id
stringlengths 30
36
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 5
878k
|
---|---|---|---|
no-problem/9908/nucl-ex9908002.html
|
ar5iv
|
text
|
# The Charge Form Factor of the Neutron from ²H⃗(𝑒⃗, 𝑒' 𝑛)𝑝
## 1 Motivation
The charge distribution of the neutron is described by the charge form factor $`G_E^n`$, which is related to the Fourier transform of the distribution and is generally expressed as a function of $`Q^2`$, the square of the four-momentum transfer. Data on $`G_E^n`$ are important for our understanding of the nucleon and are essential for the interpretation of electromagnetic multipoles of nuclei, e.g. the deuteron.
Since a practical target of free neutrons is not available, experimentalists mostly resorted to (quasi)elastic scattering of electrons from unpolarized deuterium to determine this form factor. The shape of $`G_E^n`$ as function of $`Q^2`$ is relatively well known from high precision elastic electron-deuteron scattering, but the absolute scale still contains a systematic uncertainty of about 50%. The slope of $`G_E^n`$ at $`Q^2=0`$ (GeV/c)<sup>2</sup> is known from measurements where thermal neutrons are scattered from atomic electrons .
The systematic uncertainties can be significantly reduced through the measurement of electronuclear spin observables. The scattering cross section with both longitudinal polarized electrons and a polarized target for the $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}N)`$ reaction, can be written as
$$S=S_0\left\{1+P_1^dA_d^V+P_2^dA_d^T+h(A_e+P_1^dA_{ed}^V+P_2^dA_{ed}^T)\right\},$$
(1)
where $`S_0`$ is the unpolarized cross section, $`h`$ the polarization of the electrons, and $`P_1^d`$ ($`P_2^d`$) the vector (tensor) polarization of the target. The target analyzing powers and spin-correlation parameters ($`A_i`$), depend on the orientation of the nuclear spin. The polarization direction of the deuteron is defined by the angles $`\mathrm{\Theta }_d`$ and $`\mathrm{\Phi }_d`$ in the frame where the $`z`$-axis is along the direction of the three-momentum transfer ($`𝐪`$) and the $`y`$-axis is defined by the vector product of the incoming and outgoing electron momenta. The observable $`A_{ed}^V(\mathrm{\Theta }_d=90^{},\mathrm{\Phi }_d=0^{})`$ contains an interference term, where the effect of the small charge form factor is amplified by the dominant magnetic form factor (see e.g. Refs. ). In the present paper we describe a measurement performed at NIKHEF (Amsterdam), which uses a stored polarized electron beam and a vector-polarized deuterium target, to determine $`G_E^n`$ via a measurement of $`A_{ed}^V(90^{},0^{})`$.
## 2 Experimental setup
The experiment was performed with a polarized gas target internal to the AmPS electron storage ring. An atomic beam source (ABS) was used to inject a flux of $`4.6\times 10^{16}`$ deuterium atoms/s into a cooled storage cell.
Polarized electrons were produced by photo-emission from a strained-layer semiconductor cathode (InGaAsP). After linear acceleration to 720 MeV the electrons were injected and stacked in the AmPS storage ring. Every 5 minutes the ring was refilled, after reversal of the electron polarization at the source. The polarization of the stored electrons was maintained by setting the spin tune to 0.5 with a strong solenoidal field (using the Siberian snake principle).
Scattered electrons were detected in the large-acceptance magnetic spectrometer . The electron detector was positioned at a central angle of 40, resulting in a central value of $`Q^2=0.21`$(GeV/c)<sup>2</sup>. Neutrons and protons were detected in a time-of-flight (TOF) system made of two subsequent and identical scintillator arrays. Each of the four bars in an array was preceded by two plastic scintillators used to identify and/or veto charged particles. By simultaneously detecting protons and neutrons in the same detector, one can construct asymmetry ratios for the two reaction channels $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}p)n`$ and $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}n)p`$, in this way minimizing systematic uncertainties associated with the deuteron ground-state wave function, absolute beam and target polarizations, and possible dilution by cell-wall background events.
## 3 Results
An experimental asymmetry ($`A_{exp}=\frac{N_+N_{}}{N_++N_{}}`$) can be constructed, where $`N_\pm `$ is the number of events that pass the selection criteria, with $`hP_1^d`$ either positive or negative. $`A_{exp}`$ for the $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}p)n`$-channel, integrated up to a missing momentum of 200 MeV/$`c`$, was used to determine the effective product of beam and target polarization by comparing to the predictions of the model of Arenhövel *et al.* . This advanced, non-relativistic model has shown to provide good descriptions for quasifree proton knockout from tensor-polarized deuterium. Finite acceptance effects were taken into account with a Monte Carlo code.
The spin-correlation parameter for the neutron events was obtained from the experimental asymmetry by correcting for the contribution of protons misidentified as neutrons (less than 1%, as determined from a calibration with the reaction <sup>1</sup>H($`e,e^{}p`$)), and for the product of beam and target polarization, as determined from the $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}p)n`$ channel.
Figure 2 shows the spin-correlation parameter for the $`{}_{}{}^{2}\stackrel{}{\mathrm{H}}(\stackrel{}{e},e^{}n)p`$ channel as a function of missing momentum. The data are compared to the predictions of the full model of Arenhövel *et al.*, assuming the dipole parameterization for the magnetic form factor of the neutron, folded over the detector acceptance with our Monte Carlo code for various values of $`G_E^n`$. Full model calculations are required for a reliable extraction of $`G_E^n`$. We extract $`G_E^n(Q^2=0.21(\mathrm{GeV}/c)^2)=0.066\pm 0.015\pm 0.004`$, where the first (second) error indicates the statistical (systematic) uncertainty.
In Fig. 2 we compare our experimental result to other data obtained with spin-dependent electron scattering. The figure also shows the results from Ref. . It is seen that our result favors their extraction of $`G_E^n`$ which uses the Nijmegen potential.
## 4 Conclusions
In summary, we presented the first measurement of the sideways spin-correlation parameter $`A_{ed}^V(90^{},0^{})`$ in quasifree electron-deuteron scattering from which we extract the neutron charge form factor at $`Q^2=0.21`$ (GeV/$`c`$)<sup>2</sup>. When combined with the known value and slope at $`Q^2=0`$ (GeV/$`c`$)<sup>2</sup> and the elastic electron-deuteron scattering data from Ref. , this result puts strong constraints on $`G_E^n`$ up to $`Q^2=0.7`$ (GeV/$`c`$)<sup>2</sup>.
We would like to thank the NIKHEF and Vrije Universiteit technical groups for their outstanding support and Prof. H. Arenhövel for providing the calculations. This work was supported in part by the Stichting voor Fundamenteel Onderzoek der Materie (FOM), which is financially supported by the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO), the National Science Foundation under Grants No. PHY-9504847 (Arizona State Univ.), US Department of Energy under Grant No. DE-FG02-97ER41025 (Univ. of Virginia) and the Swiss National Foundation.
|
no-problem/9908/hep-lat9908006.html
|
ar5iv
|
text
|
# Finite density [might well be easier] at finite temperature
## Acknowledgements
I would like to thank John Kogut and Don Sinclair for discussions on finite density at high temperature, Philippe de Forcrand for initial collaboration on imaginary chemical potential, Ian Barbour for kindly making his codes available, and for discussions. This work was partly supported by the NATO Collaborative Research Grant Lattice QCD at Non–zero Temperature and Chemical Potential, no. 950896, and by the TMR network Finite Temperature Phase Transitions in Particle Physics, EU contract no. ERBFMRXCT97-0122.
|
no-problem/9908/chao-dyn9908008.html
|
ar5iv
|
text
|
# Chaotic advection and relative dispersion in a convective flow
## 1 Figure captions
|
no-problem/9908/astro-ph9908058.html
|
ar5iv
|
text
|
# Polarization fluctuations due to extragalactic sources
## 1 Introduction
There are good prospects that the forthcoming space missions designed to provide high sensitivity and high resolution maps of the cosmic microwave background (CMB) will also measure the CMB polarization fluctuations (Knox 1998; Bouchet et al. 1999).
The current design of instruments for the Planck Surveyor mission (the third Medium-sized mission of ESA’s Horizon 2000 Scientific Programme) provides good sensitivity to polarization at all LFI (Low Frequency Instrument) frequencies (30, 44, 70, and 100 GHz) as well as at three HFI (High Frequency Instrument) frequencies (143, 217 and 545 GHz). The NASA’s MIDEX class mission MAP has also polarization sensitivity in all channels (30, 40 and 90 GHz).
The extraction of the very weak cosmological polarization signal requires both a great sensitivity of the instruments and a careful control of foregrounds. An analysis of the effect of Galactic polarized emissions (synchrotron and dust) on CMB measurements by the Planck and MAP missions was carried out by Bouchet et al. (1999). So far, however, the effect of extragalactic sources was not considered. On the other hand, significant linear polarization is seen in most compact, flat-spectrum radio sources which are the main contributors to small scale foreground intensity fluctuations at $`\lambda >1`$mm (Toffolatti et al. 1998) and the thermal dust emission from galaxies which dominate the counts at sub-mm wavelengths is also expected to be polarized to some extent, as the Galactic dust emission is observed to be (Hildebrand 1996).
In this paper we derive the relationship between intensity and polarization fluctuations in the case of a Poisson distribution of point sources, discuss the polarization degree of the relevant classes of extragalactic sources, and exploit recent evolutionary models to estimate the power spectrum of polarization fluctuations produced by them.
## 2 Polarization fluctuations from a Poisson distribution of polarized point sources
Following Burn (1966) we define the complex linear polarization of a source as $`P_s=\mathrm{\Pi }\mathrm{exp}(2i\chi )`$ where $`\mathrm{\Pi }`$ and $`\chi `$ are the degree and the angle of polarization, respectively. If the polarization angles of different sources within any given solid angle element $`d\mathrm{\Omega }`$ are uncorrelated, the expected value $`P_s`$ is 0 and the variance is:
$$\sigma _P^2=1/\pi _0^\pi 𝑑\chi (P_sP_s)^2=1/\pi _0^\pi 𝑑\chi \mathrm{\Pi }^2\left[\mathrm{cos}^2(2\chi )+\mathrm{sin}^2(2\chi )\right]=\mathrm{\Pi }^2.$$
(1)
Let $`N=n(S)dSd\mathrm{\Omega }`$ be the number of sources with flux $`S`$ within $`dS`$ and polarization degree $`\mathrm{\Pi }`$ in a given solid angle element $`d\mathrm{\Omega }`$. As far as the central limit theorem holds, the expected value of the linear polarization within $`d\mathrm{\Omega }`$ is also 0, with variance
$$\sigma _{P,d\mathrm{\Omega }}^2=\frac{\mathrm{\Pi }^2}{N}.$$
(2)
The fluctuation amplitude of the polarized flux $`S_P=NSP_s`$ among the different cells of the sky subtending a solid angle $`d\mathrm{\Omega }`$, due to sources with flux $`S`$, is therefore obtained integrating over the probability distribution of $`N`$, $`\rho (N)`$:
$$\sigma _{S_P,d\mathrm{\Omega }}^2=\left(S_PS_P\right)^2=S_P^2=_0^{\mathrm{}}𝑑N\rho (N)NS^2\mathrm{\Pi }^2=\mathrm{\Pi }^2S^2N(S).$$
(3)
In the case of a Poisson distribution of sources the variance is equal to the mean. Therefore, integrating the above equation over $`S`$ and over the solid angle we straightforwardly obtain $`\sigma _{I_P}=\mathrm{\Pi }\sigma _I`$, where $`\sigma _I`$ is the rms intensity fluctuation for a Poisson distribution of sources.
The assumption of an equal polarization degree for all sources is obviously unrealistic. However, it follows from the above calculations that, if the polarization degree is uncorrelated with flux, the result depends only on the mean value of $`\mathrm{\Pi }`$.
A dependence of the mean value of $`\mathrm{\Pi }`$ on $`S`$ arises, in particular, in the case of contributions from classes of sources with different polarization properties and different shapes of the $`\mathrm{log}N`$$`\mathrm{log}S`$ curves. These different populations must be dealt with separately.
## 3 Linear polarization properties of the relevant classes of sources
Studies of the polarization properties of extragalactic sources at mm wavelengths are still scanty. Table 1 summarizes the main results on radio/mm sources. For each of the main source classes (in column 1) and for each wavelength in column 2, we give the median (column 3), the minimum (column 4) and the maximum (column 5) polarization value found in literature. Column 6 lists the references from which the reported values are derived.
As shown in Table 1, the three classes of objects (BL Lacs, QSOs and bright radio galaxies) emit almost the same fraction of polarized radiation in the radio band. The major extragalactic contributors to the non-thermal polarization at sub–mm and mm wavelengths are the flat spectrum (spectral index $`\alpha 0.5`$ if $`S_\nu \nu ^\alpha `$) compact radio sources (mainly BL Lacertae objects and flat spectrum QSOs). These objects constitute about 50% of the flux limited ($`S>1`$ Jy at 5 GHz) radio catalogue compiled by Kühr et al. (1981). Stickel et al. (1991) have drawn from this catalogue a complete sample of BL Lacertae objects brighter than $`m=20`$ mag. Out of it, we have selected a complete sub–sample of 14 objects ($`RA>9^h`$ and $`\delta >0^{}`$), for which radio and/or mm polarization measurements are available.
The polarization data for objects in this sub-sample (hereinafter Stickel north) are listed in Table 2. The tabulated percentage polarization degrees were obtained averaging the measurements from long term monitoring programs. Data at 4.8 GHz, 8.1 GHz and 14.5 GHz are from Aller et al. (1985, 1999), those at 22 GHz, 31 GHz and 90 GHz are from Rudnick et al. (1985) and those at 1.1 mm are from Nartallo et al. (1998).
We have compared the average polarization percentages at the various wavelengths of sources in the Stickel north sample and in the sample by Aller et al. (1999), excluding those in common with the Stickel north sample. The agreement is very good (see Fig. 1). The mean polarization percentages (P<sub>m</sub>), the 68% confidence uncertainties ($`\mathrm{\Delta }P_m`$) obtained from the Student’s $`t`$ distribution ($`\mathrm{\Delta }P_m=t_\nu (0.32)\sigma _P(N1)^{1/2}`$, where $`N`$ is the number of sources in the bin and $`\nu =N1`$) and the dispersions, $`\sigma _P`$, for the Stickel north sample are given in Table 3.
Since for most objects in the two samples redshift information is available, we have converted the observed into the rest–frame wavelengths and computed the average polarization degrees in wavelength bins. The results for the Stickel north sample are given in Table 4 and plotted in Figure 1$`b`$, where the results for the Aller sub-sample are also reported for comparison. Table 4 gives the adopted wavelength intervals, the average $`\lambda `$ in each interval, computed as the geometric mean of the two wavelength limits, the average polarization fraction with its 68% confidence uncertainty from the Student’s $`t`$ distribution and its dispersion, and the number of available measurements in each wavelength interval.
Measurements of polarized thermal emission from dust are only available for interstellar clouds in our own Galaxy. The distribution of observed polarization degrees of dense clouds at 100$`\mu `$m shows a peak at $`2\%`$ (Hildebrand 1996). The polarization degree of emission of silicate grains in clouds opaque to visible light but optically thin in the far-IR is nearly independent of wavelength if $`\lambda a`$, $`a`$ being the grain size (Hildebrand 1988). This condition is very likely to be met at the long wavelengths of interest here.
Polarization maps of the Orion molecular cloud at $`100\mu `$m, $`350\mu `$m, $`450\mu `$m, $`1.3`$mm, and $`3.3`$mm (Schleuning 1998; Rao et al. 1998) look very similar except at clumps of higher optical depth, where the polarization increases with wavelength. In fact, the maximum polarization decreases rapidly with increasing optical depth (Hildebrand 1996); thus, the mm/sub-mm polarization may be higher than at $`100\mu `$m. On the other hand, the overall polarization of the light from a galaxy is the average of contributions from regions with different polarizing efficiencies and different orientations of the magnetic field with respect to the plane of the sky; all this works to decrease the polarization level in comparison with the mean of individual clouds.
We could not find any published polarization measurement of dust emission in external galaxies not hosting strong nuclear activity. However, the first results of the SCUBA Polarimeter include imaging polarimetry of the starburst galaxy M 82 at 850$`\mu `$m with a 15” beam size; a polarization degree of about 2% was measured (J. Greaves, private communication). An early image, available at the polarimeter Web page (http://www.jach.hawaii.edu/JCMT/scuba/scupol/
general/m82\_polmap.gif), shows very ordered polarization vectors over scales of a few hundred parsecs. This may not occur in general. We may expect that, in other galaxies, the polarization vectors from giant molecular clouds are randomly ordered and tend to cancel out.
Polarimetric measurements at $`170\mu `$m with ISOPHOT have been carried out for two galaxies (U. Klaas, private communication): NGC 1808 (P.I.: E. Krügel) and NGC 6946 (P.I.: G. Bower). The results, however, are not available yet.
## 4 Power spectrum of polarization fluctuations due to extragalactic sources
In Figures 2 and 3 the power spectrum of foreground polarization fluctuations for all the relevant Planck and MAP channels is compared with the power spectrum of CMB anisotropies and of CMB polarized components. As for the latter, we have plotted the power spectra of the combinations of Stokes parameters defined by Seljak \[1997; his eqs. (24) and (25)\] and called $`E`$ and $`B`$. The estimate of $`E`$-mode polarization fluctuations refer to a standard CDM model (scale-invariant scalar fluctuations in a $`\mathrm{\Omega }=1`$, $`\mathrm{\Lambda }=0`$ universe with $`H_0=50\text{km}\text{s}^1\text{Mpc}^1`$ and a baryon density $`\mathrm{\Omega }_b=0.05`$). The quantity $`B`$ vanishes for polarization induced by primordial scalar perturbations, and therefore provides a unique signature of tensor perturbations (Seljak 1997). The $`B`$-mode power spectrum shown in Figs. 2 and 3 refer to a tilted CDM model with power law indices $`n_s=0.9`$ and $`n_t=n_s1=0.1`$ for scalar and tensor perturbations, respectively; the other cosmological parameters keep the same values adopted for the standard CDM model. Calculations of CMB power spectra have been carried out using the CMBFAST package by Seljak & Zaldarriaga (1996).
The thick dashed lines show the $`E`$-mode of dust polarized power spectrum derived by Prunet et al. (1998), scaled to the central frequencies of Planck polarized channels using the dust emission spectrum adopted by these Authors (emission $`\nu ^2B_\nu (17.5\mathrm{K})`$, where $`B_\nu (T)`$ is the Planck function at frequency $`\nu `$ and temperature $`T`$).
The $`B`$-mode dust power spectrum turns out to be close to the $`E`$-mode one \[cf. eqs (11) and (12) of Prunet et al. (1998)\]. In fact, Seljak (1997) argued that most foregrounds should contribute on average the same amount to both modes. Therefore, for all foregrounds we consider a single polarized power spectrum, assumed to be representative of both modes.
Following Bouchet et al. (1999), we assume that, at all frequencies relevant for Planck and MAP, the synchrotron polarized emission is perfectly correlated with the total synchrotron emission (for which a power spectrum $`C_T=4.5\mathrm{}^3(\mu \mathrm{K})^2`$ at 100 GHz was adopted) and the polarization degree is 44%. An antenna temperature spectral index of 3 ($`T_{\mathrm{A},\mathrm{syn}}\nu ^3`$) has been used to extrapolate the power spectrum (in terms of brightness temperature) to the other Planck/MAP frequencies (see De Zotti et al. 1999 for references).
The polarized angular power spectra of extragalactic radiosources (thin solid lines) is estimated exploiting the model of Toffolatti et al. (1998), as updated by De Zotti & Toffolatti (1999). We have assumed that the mean polarization degree of BL Lacs in the Stickel north sample applies to all radio galaxies contributing to fluctuations in Planck’s channels; based on the results shown in Fig. 1, we have adopted a polarization degree of 5% for $`\nu 143`$GHz, of 6% at 217 GHz and of 10% at 545 GHz.
As for dusty galaxies (thin dashed lines), we have adopted a polarization degree of 2% at all frequencies and the power spectra of temperature fluctuations derived by Toffolatti et al. (1998) and, for HFI channels, also by Guiderdoni et al. (1998; their model E).
A flux cut at 1 Jy was adopted for all channels (i.e. sources brighter than 1 Jy were removed); at 30, 100 and 217 GHz we show also the angular power spectrum derived adopting a flux cut of 100 mJy for radiosources (lower thin solid lines).
In Table 5 we report the values of $`C_{\mathrm{}}`$ for temperature fluctuations due to a Poisson distribution of extragalactic radio sources and dusty galaxies for the same cases shown in Figs. 2 and 3. The corresponding values for polarization fluctuations follow immediately multiplying by $`\mathrm{\Pi }^2`$. Note that a Poisson distribution generates a simple white noise spectrum with the same power on all multipoles (Tegmark & Efstathiou 1996). The values of $`C_{\mathrm{}}`$ are given in terms of brightness temperature fluctuations and expressed in $`\mu `$K<sup>2</sup>.
As expected, polarization fluctuations due to extragalactic sources are particularly relevant at small angular scales. For multipoles $`\mathrm{}\genfrac{}{}{0pt}{}{>}{}\mathrm{\hspace{0.17em}300}`$, they in fact dominate foreground contributions at $`\nu \genfrac{}{}{0pt}{}{<}{}\mathrm{\hspace{0.17em}100}`$GHz. On these scales, at the lowest Planck frequency (30 GHz) their amplitude is, according to our estimate, close to that of CMB polarization fluctuations induced by scalar perturbations. In the “cosmological window” ($`70\genfrac{}{}{0pt}{}{<}{}\nu \genfrac{}{}{0pt}{}{<}{}\mathrm{\hspace{0.17em}200}`$GHz), however, extragalactic sources are not seriously detrimental to measurements of CMB polarization fluctuations.
The thin dotted lines in Figs. 2 and 3 show the expected power spectra of instrumental noise, for polarization measurements, averaged over the sky, for Planck’s LFI and HFI, respectively. Following Tegmark & Efstathiou (1996) we describe the noise power spectrum as $`C_{\mathrm{},\mathrm{noise}}=\sigma ^2\mathrm{FWHM}^2\mathrm{exp}(\mathrm{}^2\sigma _b^2)`$ where FWHM in expressed in radians, $`\sigma _b=\mathrm{FWHM}/2\sqrt{2\mathrm{ln}(2)}`$ and $`\sigma `$ is the rms noise for a square pixel with side FWHM.
For LFI we have adopted the sensitivities for brightness temperature measurements given by Mandolesi et al. (1998) multiplied by a factor of 2 (Mandolesi, private communication). Sensitivities of HFI channels for polarization measurements are given by Puget et al. (1998). There is a slight difference in the mission duration adopted by the two groups to derive their mean sensitivity estimates: 12 months for LFI, 14 months for HFI.
The heavier dots in Fig. 2 show the expected mean instrumental noise per resolution element for MAP’s polarization measurements at 30, 40 and 90 GHz obtained from sensitivities for brightness temperature measurements (see the MAP Web page) multiplied by a factor $`\sqrt{2}`$ (G. Hinshaw, private communication).
As shown by Figs. 2 and 3, the major hurdle in extracting the CMB polarization signal is instrumental noise. However, a simple argument shows that, at least for a limited band-power range, sufficient sensitivity can be reached. We have investigated, in particular, the potential of LFI in this respect. The expected sensitivities to polarization per resolution element, averaged over the sky, for a 12 months mission, are 6, 10, 14 and 17$`\mu `$K at 30, 44, 70, and 100 GHz, respectively; the angular resolutions (FWHM) are 33’, 23’, 14’, and 10’, respectively (Mandolesi et al. 1998). Simulations done by C. Burigana indicate sensitivities 7 times better over areas of about 25 square degrees around each of the ecliptic poles. Within these areas, by rebinning the maps at 44, 70 and 100 GHz (we leave aside the 30 GHz map which is the most contaminated by polarized foregrounds) to a $`20^{}`$ resolution and combining them, a sensitivity to polarization of about $`1\mu `$K can be achieved, allowing to image the CMB polarization induced by scalar perturbations predicted by the standard CDM model. Of course, a lower sensitivity is enough to determine the first moments of the distribution of polarization fluctuations.
## 5 Conclusions
We have shown that the polarization fluctuations due to a Poisson distribution of point sources with uniform evolutionary properties and constant polarization degree are simply equal to intensity fluctuations times the average polarization degree.
The information on the polarization degree of the classes of extragalactic sources expected to dominate in the frequency ranges relevant for the MAP and Planck missions is scanty. We have taken a conservative approach, assuming that all radio sources are as polarized as BL Lac objects and that the polarization degree of dusty galaxies is similar to that of dense clouds in our own Galaxy and of M 82, a galaxy showing a remarkably ordered magnetic field.
We find that, on small scales (multipoles $`\mathrm{}\genfrac{}{}{0pt}{}{>}{}\mathrm{\hspace{0.17em}300}`$), polarization fluctuations due to radio sources may indeed dominate foreground contributions at $`\nu \genfrac{}{}{0pt}{}{<}{}\mathrm{\hspace{0.17em}100}`$GHz. However, in the “cosmological window” ($`70\genfrac{}{}{0pt}{}{<}{}\nu \genfrac{}{}{0pt}{}{<}{}\mathrm{\hspace{0.17em}200}`$GHz), extragalactic sources are not a threat for measurements of CMB polarization fluctuations.
We have also argued that Planck/LFI can reach, in regions around the Galactic polar caps, polarization sensitivities of $`1\mu `$K, allowing to map CMB polarization fluctuations on scales $`20^{}`$. A detailed analysis of Planck and MAP capabilities for CMB polarization measurements has been carried out by Bouchet et al. (1999).
We thank J. Greaves, SCUBA Polarimeter Project Scientist, and U. Klaas, from the ISOPHOT Data Centre at MPIA Heidelberg, for useful information on polarization measurements of external galaxies with SCUBA and ISO, respectively. We also thank G. Hinshaw and N. Mandolesi for information on MAP and Planck/LFI polarization sensitivities, respectively. T his work has been done in the framework of the Planck-LFI Consortium activities; it has been supported in part by ASI, CNR and MURST.
|
no-problem/9908/cond-mat9908329.html
|
ar5iv
|
text
|
# Material failure time and the fiber bundle model with thermal noise
## Abstract
The statistical properties of failure are studied in a fiber bundle model with thermal noise. We find that in agreement with recent experiments the macroscopic failure is produced by a thermal activation of microcracks. Most importantly the effective temperature of the system is amplified by the spatial disorder (heterogeneity) of the fiber bundle.
PACS numbers: 05.70.Ln, 62.20.Mk, 05.20.-y
Material failure is a widely studied phenomenon not only for its very important technological applications but also for its fundamental statistical aspects, which are not yet very well understood. Many models have been proposed to give more insight into the statistical analysis of failure. Among the most studied ones we can mention the fuse and the non-linear spring networks which can reproduce several features of crack precursors, experimentally observed in heterogeneous materials subjected to a quasi-statically increasing stress . Specifically the power law behaviour of the acoustic emission observed in several experiments close to the failure point.
However these networks and the other related models, in their standard formulation, are unable to describe the behaviour of a material subjected to a creep-test, which consists in keeping a sample at a constant stress till it fails. Creep-tests are widely used by engineers in order to estimate the sample life time as a function of the applied stress. A modified fuse network, which takes into account the Joule effect in the fuses, has been proposed to explain the finite life time of a sample subjected to a constant stress. However also this model does not explain the recent experimental results on micro crystals, gels and heterogeneous materials. These experiments show that the life time $`\tau `$ of a sample, subjected to an imposed stress $`P`$ is well predicted by the equation
$$\tau =\tau _o\mathrm{exp}\left(\alpha \frac{\mathrm{\Gamma }^dY^{(d1)}}{KT_{eff}P^{(2d2)}}\right)$$
(1)
where $`\tau _o`$ is a constant, $`\mathrm{\Gamma }`$ the surface energy, $`Y`$ the Young modulus, $`k`$ the Boltzmann constant, $`\alpha `$ a constant which depends on the geometry, $`T_{eff}`$ is an effective temperature and $`d`$ the dimensionality of the system. The main physical hypothesis behind eq.1 is that the macroscopic failure of a material is produced by a thermal activation of micro cracks. In the original Pomeau’s theory $`T_{eff}`$ of eq.1 coincides with the thermodynamic temperature $`T`$ while experimentally $`T_{eff}>>T`$. To explain this results it has been supposed that the disorder of the material ”amplifies” the thermal noise.
The purpose of this letter is to show that adding to the spring network a noise, which plays the role of a temperature, it is possible to reproduce the behaviour of a material subjected to a creep-test. Specifically the functional dependence on the applied stress of the network life time is consistent with that observed in recent experiments, that is with eq.1. Furthermore we can prove that, for such a process, the noise is ”amplified” by the network disorder.
In order to study the influence of noise on macroscopic failure we have chosen the simplest spring network, that is the so called democratic fiber bundle model, proposed a long time ago by Pierce to study cable failure. This model, widely studied in the quasi static regime , is equivalent to $`N`$ springs in parallel subjected to a total traction force $`F`$. Specifically we have numerically studied the model by using the following rules:
* The external applied force $`F`$ produces a local stress $`f_i`$ on each fiber. $`F`$ is democratically and completely distributed in the net: $`F=_{i=1}^Nf_i`$
* The local stress $`f_i`$ on the i-th fiber produces a local deformation $`e_i`$. Being the media elastic, stress and deformation are linked by Hook’s law:
$$f_i=Ye_i$$
(2)
where $`Y`$ is the Young modulus, which is assumed to be the same for all the fibers : $`Y_i=Y`$.
* The strength of each fiber is characterized by a critical stress $`f_i^{(c)}`$: if at time $`t`$ on the i-th fiber local stress $`f_i(t)`$ is greater than critical stress $`f_i^{(c)}`$ the fiber cracks, and his local stress falls to zero at time $`t+1`$. Further, we assume that in this process some energy $`ϵ_i`$ is released proportionally to the square of local stress $`f_i`$; for sake of simplicity we will assume $`ϵ_i=f_i^2`$. Critical stress $`f_i^{(c)}`$ is a realization of a random variable that follows a normal distribution of mean $`f^{(c)}`$ and variance $`KT_d`$:
$$f_i^{(c)}=f^{(c)}+N_d(KT_d)$$
(3)
We call $`N_d`$ disorder noise.
* Each fiber is subjected to an addictive time dependent random stress $`\mathrm{\Delta }f_i(t)`$ which follows a zero mean normal distribution of variance $`KT`$:
$$\mathrm{\Delta }f_i(t)=N_T(t,KT)$$
(4)
being $`t`$ the time. We call $`N_T`$ thermal noise. We assume that $`\mathrm{\Delta }f_i(t)`$ is a white random process,which is independent in each fiber, i.e. the correlation function $`E\left[\mathrm{\Delta }f_i(t_1)\mathrm{\Delta }f_j(t_2)\right]=0`$ if $`t_1t_2`$ or $`ij`$.
The first three items are those used in the standard formulation of fiber bundle model. A new feature, which is similar to a thermal activation process, is introduced in \[A4\] to explain the dependence of failure time on a constant applied stress.
The model has the following properties. We see from \[A1\] that there exists a long-range interaction between the fibers: indeed if a number $`n(t)`$ of fibers are broken at time $`t`$, the local stress on each of the remaining fibers will be:
$$f_i(t)=\frac{F}{Nn(t)}+\mathrm{\Delta }f_i(t)$$
(5)
that is, under the specified boundary conditions, the breaking of some fibers produces an increase of local stress on the other ones. Item \[A3\] models the heterogeneity of media. The assumption that all the disorder in the model appears in the strength distribution rather than in the elastic constants may be argued by noticing that the effective elastic constant of a single fiber is essentially the average of the local elastic constant along the fiber, while the strength is determined by its weakest point . If $`KT=0`$ the model is reduced to the standard one: in this case the applied force is increased linearly $`F=A_pt`$ from zero to the critical value $`F_c`$ needed to break the whole network. The sample life time $`\tau `$ is equal to $`F_c/A_p`$ for any value of the slope $`A_p`$.
It is quite clear that when $`KT=0`$ and a constant force $`F`$ is applied the system breaks in a single avalanche only if $`F`$ is large enough otherwise it will never break. For example if $`KT_d=0`$ all the fibers are strictly equal and if there is no thermal activation ( i.e. $`KT=0`$) the system breaks in a single step when $`F=Nf^{(c)}`$.
If $`KT0`$ then the system can break also at constant imposed stress. We have numerically studied the behaviour of the model as a function of $`F`$, $`KT`$ and $`KT_d`$. In the following we assume $`N=1000`$ and $`f^{(c)}=1`$. For each set of parameter values we have repeated the numerical simulation at least ten times to estimate the scattering of the results for different realization of the noise. This corresponds to the scattering of symbols in the figures. We call event the simultaneous breaking of several fibers and event size the number $`s(t)`$ of fibers which crack. The energy $`ϵ`$ associated to an event is the sum of the energies released by the fibers which crack, that is :
$$ϵ=s(t)\left(\frac{F}{Nn(t)}\right)^2.$$
(6)
The cumulated energy $`E(t)`$ is the sum of $`ϵ`$ from 0 to t. When $`KT0`$ and $`F`$ is constant the event statistics is quite similar to that observed at $`KT=0`$ for $`F`$ increasing linearly in time . As an example of the system response to a constant force $`F=540`$ with $`KT_d=0.005`$ and $`KT=0.007`$, we plot in fig.1a the distribution $`N(ϵ)`$ of $`ϵ`$ and in fig.1b E(t) as a function of the reduced parameter $`(\tau t)/\tau `$ (notice that because of the constancy of F the only control parameter is time). In agreement with experiments we find that $`ϵ`$ is power law distributed and that the cumulated energy $`E`$ has a power law dependence on $`(\tau t)/\tau `$ for $`t\tau `$. More details on this statistical features of the model for different driving forces will be given in a longer report. Here we want to focus on failure time $`\tau `$ of the network in the case of a creep test (constant F).
We first keep $`KT_d=0`$ and and we study the evolution towards failure for various $`KT`$ and $`F`$. The results are summarized in fig.2. When $`KT0`$, we observe (see fig. 2a) that failure time $`\tau `$ as a function of $`\frac{1}{F^2}`$ follows an exponential law for any fixed value of $`KT_d`$:
$$\tau \mathrm{exp}\left[\left(\frac{F_0}{F}\right)^2\right]$$
(7)
where $`F_0`$ is a fitting parameter. Further,looking at fig. 2b we notice that the failure time $`\tau `$ depends on thermal noise $`KT`$ as follows:
$$\tau \mathrm{exp}\left(\frac{A}{KT}\right)$$
(8)
where A is a fitting parameter.
We notice that these results are similar to the prediction of Pomeau’s theory, that is eq.1 with $`d=2`$. Because of this analogy with eq.1, we are now interested in studying the dependence of failure time $`\tau `$ on disorder noise (i.e. on $`KT_d`$): to this aim, we have done simulations keeping the thermal noise variance $`KT`$ fixed to constant (not zero) values. Looking at fig.3, we observe the following facts:
* failure time $`\tau `$ decreases following a power law as the disorder noise variance $`KT_d`$ increases, that is the more the media is heterogeneous, the smaller is failure time (see fig. 3a).
* as $`KT_d`$ increases, the absolute difference between failure times $`\tau `$ for different values of thermal noise variance $`KT`$ decreases, that is failure time $`\tau `$ becomes less sensitive to the effective value of thermal noise, as shown in fig.3b.
Thus one may conclude that disorder noise amplifies the effect of thermal noise and reduces the dependence of $`\tau `$ on the temperature; these results allow us to explain the recent experimental observations. Let us recall that experiments on microcrystals , gels and macroscopic composite materials reveal that dependence on $`P`$ of the sample life time is very well fitted by eq.(1). However calculations show that thermal fluctuations are too little to activate the nucleation of microcraks in the times $`\tau `$ measured in the experiments. It had been measured that the temperature needed to have the measured life-times $`\tau `$ should be of the order of several thousands of Kelvin ($`3000K`$ for wood ). Experiments show that the life time $`\tau `$ of very heterogeneous materials , is (in the limit of experimental errors) independent of $`T`$ while the life time $`\tau `$ of quite homogeneous materials as microcristals strongly depends on $`T`$, as predicted by eq. (1). To explain these results has been supposed that disorder amplifies the thermal noise so that the nucleation time of defects becomes of the order of the measured ones and that the lifetime $`\tau `$ of the sample depend on the heterogeneity of the media. These hypothesis are now well verified by the numerical results B1 and B2 of the fiber bundle model.
As a conclusions we have shown that adding a white noise, which plays the role of temperature, to the fiber bundle model we can reproduce the recent experimental observations on the dependence of sample life time on the applied stress. Furthermore using such a simple model we can prove that the disorder of the system amplifies the thermal noise in crack nucleation process. This explains quite well why the experimentally measured $`T_{eff}`$ is more close to the thermodynamic temperature in microcrystals than in heterogeneous materials. Similar conclusions about a disordered induced high temperature in nucleation processes have been reached in other disordered systems such as foams . This is quite interesting because it seems to be a quite general properties of disordered systems where a thermal activated processes with long range correlation may be present.
We acknowledge useful discussion with Y. Pomeau. One of us (R.S.) thanks ”Le Laboratoire de Physique de l’E.N.S.L.” for the very kind hospitality during his visit in Lyon.
* on leave from ”Facoltá d’Ingegneria, Univeristá di Firenze, Italy ”, with a SOCRATES exchange program of the European Community
|
no-problem/9908/astro-ph9908067.html
|
ar5iv
|
text
|
# The complex radio spectrum of 3C 130
## 1 Introduction
Wide-angle tail radio galaxies are an interesting sub-class of the population of FRI (Fanaroff & Riley fr (1974)) objects, which typically lie at the centres of clusters and show narrow, well-collimated jets which flare abruptly into broad, diffuse plumes. Their radio power is normally intermediate between the more typical jet-dominated FRI and the ‘classical double’ FRII classes of extragalactic radio source, and an understanding of their dynamics is important to our knowledge of the relationships between these two classes and the possible evolution between them.
In an earlier paper (Hardcastle h98 (1998), hereafter Paper I) I presented new radio maps of the wide-angle tail radio galaxy 3C 130 ($`z=0.109`$). Two-frequency spectral index mapping in that paper showed flat-spectrum jets (in this paper, the term is reserved for the narrow, well-collimated features seen in the inner 50 kpc of the sources) and a hotspot (a compact, sub-kpc feature at the end of the northern jet) together with steeper-spectrum material at the edges of the plumes (the broader, more diffuse features seen between 50 and $`500`$ kpc from the nucleus). This steep-spectrum material is particularly clear in the southern plume, and is referred to here as a ‘sheath’, although I emphasise that, unlike the sheaths seen in some twin-jet FRI radio galaxies, this region has no polarization properties to distinguish it from the rest of the plume. The sheath is the only feature of the two-point radio spectrum which is not obviously consistent with a fairly simple model for the source’s dynamics, in which particles are accelerated at the base of the plumes and the spectral steepening along the plumes is a consequence of outflow of an ageing electron population. The additional data presented here show that the situation is more complicated than that simple model would imply.
B1950 co-ordinates are used throughout this paper, and spectral index $`\alpha `$ is defined in the sense $`S\nu ^\alpha `$.
## 2 Data
The X-band (8.4-GHz) and L-band (1.4-GHz) data used in this paper were described in Paper I. Data from a short (0.5 h) C-band (4.9-GHz) observation with the NRAO Very Large Array (VLA) in its C configuration, taken on 1984 Jun 11, were kindly provided by Alan Bridle. The U-band (15-GHz) images presented here are the result of a 3-hour observation with the VLA in its D configuration taken on 1999 Mar 11. All data were reduced and analysed using the aips software package.
### 2.1 Flux scales
As described in Paper I, both 3C 48 and 3C 286 were used as primary flux calibrators for the L- and X-band observations, which were taken between 1994 Nov 10 and 1995 Nov 28. Specifically, 3C 286 was used for the A-configuration X-band observations and 3C 48 for all others. The data analysis in Paper I used the older version of the SETJY aips task, which incorrectly rounded coefficients in the analytic expression for the flux densities of calibrator sources. However, the effect is very small when combined with the change between the old (1990) values of the coefficients and the more appropriate new (1995.2) values. L-band fluxes should be reduced by 0.3% and the X-band B, C and D-configuration fluxes by 2%; the A-configuration X-band flux, based on 3C 286, should be reduced by 2.5% when the effect of the partial resolution of the source is incorporated, but we use 2% for all the X-band data in what follows. The primary flux calibrator for the C-band observations was 3C 48; the fluxes derived from this are increased by 2% to take account of the variation in 3C 48’s flux density between the epoch of observation and 1995 (Perley & Taylor pt (1999)). The primary flux calibrator for the U-band observations was 3C 48, and it is assumed that 3C 48 has not varied significantly between 1995 and 1999, so that the flux densities derived from this are correct.
### 2.2 Mapping
The long, multi-array X- and L-band observations of Paper I sample the $`uv`$ plane more densely and have a much broader range of baselines than the short, single-configuration C-band observations. The U-band observations are long and so well-sampled, but still cover a narrow range of baselines. In order to try to correct for this and ensure that fluxes from maps were directly comparable I mapped the source at all four frequencies with the CLEAN-based aips task IMAGR using only baselines between 1.8 and 50 k$`\lambda `$; the $`uv`$ plane was acceptably sampled in this range at all four frequencies, though the C-band data are still sparse (Fig. 1). Reweighting the data (using different values of the ‘robustness’ parameter in IMAGR) ensured that the fitted beams were similar. The same 4.0-arcsec restoring Gaussian was then used for each map, scaling residuals appropriately. (Experiment showed that, in spite of folklore to the contrary, there was no significant difference between the results of this procedure and those of restoring clean components with the Gaussian which was the best fit to the dirty beam and then convolving to the required resolution.) Primary beam correction was applied with the aips task PBCOR. Small shifts were applied to each image so that the unresolved cores were aligned to better than 0.05 arcsec.
## 3 Results
### 3.1 Images and their fidelity
Images of 3C 130 at the four frequencies used are shown in Fig. 2.
Some apparent evidence of anomalous spectral behaviour can immediately be seen from these maps. For example, the inner jets of the source appear fainter at 5 GHz than they do at 8.4 GHz, implying an inverted spectrum; this would be very surprising in an extended source region. However, since the $`uv`$ plane coverage is sparsest at long baselines in the 5-GHz data, these apparent spectral differences may simply be a result of image infidelity. The CLEAN algorithm can be viewed as an attempt to interpolate over the missing spacings in the $`uv`$ plane, but image fidelity clearly depends on the amount of interpolation needed.
To examine the degree to which image infidelity was a problem in these datasets, I made a test model of the source (consisting of 100,000 CLEAN components from an 8-GHz map with all baselines $`<50`$ k$`\lambda `$). Using the aips task UVSUB, I replaced the real data in all four datasets with the Fourier transform of the CLEAN component model; this simulates the effect of observing an identical source with different $`uv`$ coverage. The four simulated datasets were then mapped with IMAGR. The ratios of fluxes in the resulting maps indicates the fidelity with which a particular component is likely to be reproduced. While the L- and X-band datasets produce very similar maps of the model data, the U-band sampling seems to underestimate the flux of the southern jet, and the C-band sampling gives a noticeably poorer reproduction of the original model, with regions of lower flux particularly in the jets and at the edges of the plumes.
I next made another model, consisting of 100,000 CLEAN components from a similar 1.4-GHz map. The 1.4-GHz data has denser sampling in the centre of the $`uv`$ plane. Simulating observations of this model with the $`uv`$ coverages of the four datasets shows that even the X- and U-band datasets do not perfectly reproduce structure on the largest scales; the effect is to produce spurious spectral ‘steepening’ at the very edges of the plumes, increasing the spectral index by 0.2–0.3 at the edges of detectability. The steep-spectrum sheath, however, is too strong an effect to be entirely or mostly due to this undersampling. But these results illustrate the danger of the common assumption that simply matching resolutions or longest and shortest baselines will give maps that are safe to use for spectral index analysis. All spectra at the very edges of the plumes, the U-band spectra of the jets, and the C-band data throughout the source, must therefore be treated with caution.
As an experiment, I tried making a map of spectral index between the 5-GHz data and the 8.4-GHz data remapped with the 5-GHz sampling. Assuming that the original 8.4-GHz images have ideal fidelity, we might expect this resampling to compensate fully for the poor sampling of the 5-GHz data. Although sampling the 8.4-GHz data sparsely does reduce the flux in the jets, the resulting 5–8.4-GHz spectral index is still very flat (0.1) while the spectral indices in the plumes are much more reasonable. This suggests that even identical sampling does not give completely reliable results for snapshot observations.
### 3.2 Flux densities and spectra
Flux densities of the various components of the source are tabulated in Table 1. Except for the core and hotspot flux densities, which were derived from a fit of Gaussian and zero level using the aips task JMFIT, these were measured from polygonal regions defined on the 15-GHz map using miriad. In Fig. 3 the resulting component spectra are plotted.
The spectra of the core, jets and northern plume are not unexpected. There is some slight evidence for a steeper spectrum in the N jet between 8 and 15 GHz, perhaps indicating a cutoff in the spectrum of the jet at high frequencies, but it is possible that the N jet is affected by undersampling. The S jet is certainly somewhat affected by undersampling at 15 GHz. The N plume shows a slightly concave spectrum, suggesting the presence of multiple spectral components. The hotspot in the N plume shows a flat spectrum with $`\alpha 0.5`$, consistent with a model in which it is produced by particle acceleration at the shock at the end of the N jet; its spectrum has only steepened slightly by 15 GHz.
The immediately striking result of these measurements is the steep spectral index between 8 and 15 GHz in the southern plume. We can be sure that this is not an artefact of poor sampling; the S plume lacks compact structure, and simulations show that we would not expect any flux on the scales of the observed emission to be missing from the 15-GHz maps. It appears that there is a genuine break in the spectrum between 8 and 15 GHz in the southern plume which is not present in the northern plume. As shown in Fig. 4, this effect is not limited to a single region in the plume, but is visible throughout.
The steep-spectrum ‘sheaths’ around the plumes, particularly the southern plume, which were visible in the 8-GHz data presented in Paper I, are missing in the 15-GHz images, although again simulated images show that the 15-GHz observations have sampling which should be adequate to reproduce them. This implies that the sheath regions have very steep spectra between 8.4 and 15 GHz. Using regions defined with miriad on the 1.4–8.4-GHz spectral index map, in which the sheath region is well defined, I find that $`\alpha _{8.4}^{15}2`$ for the sheaths around both north and south plumes, whereas $`\alpha _{1.4}^{8.4}1`$.
## 4 Discussion
### 4.1 Tomography and the nature of the steep-spectrum ‘sheath’
It has recently been suggested (e.g. Katz-Stone & Rudnick kr (1997)) that the jets in some FRI radio galaxies have a two-component structure, consisting of a flat-spectrum ‘core jet’ and steep-spectrum surrounding ‘sheath’. Katz-Stone et al. (krbo (1999)) show that the same picture may apply to two WAT sources from the sample of O’Donoghue et al. (o2e (1990)). The observed spectral steepening with distance from the core in FRI sources might therefore be unrelated to spectral ageing and expansion, as is frequently assumed; it might simply be a consequence of the increasing dominance of the sheath component.
To test whether such a picture is viable in 3C 130, I constructed a spectral tomography gallery as discussed by Katz-Stone & Rudnick; this involves generating a set of maps by subtracting a scaled version of the high-frequency map from the low-frequency map, so that for each pixel of the tomography map ($`I_t`$) we have
$$I_t(\alpha _t)=I_l\left(\frac{\nu _h}{\nu _l}\right)^\alpha I_h$$
where $`\alpha _t`$ is varied. Features of a given spectral index vanish on the tomography map corresponding to that spectral index; if the apparent steepening in 3C 130 is due to varying blends of a flat- and steep-spectrum component, and the steep-spectrum component is relatively smooth, the plumes should appear more uniform in a tomography map with a spectral index corresponding to that of the flat-spectrum component, as the flat-spectrum component should then have vanished, leaving only a (scaled) version of the steep-spectrum component. If there is no single, uniform flat-spectrum component, the plumes will still show structure for any value of $`\alpha _t`$.
The full gallery of tomography images is not shown, but Fig. 5 shows a representative example, made with the L and X-band maps taking $`\alpha _t=0.55`$. It will be seen that the jets and N hotspot are oversubtracted, giving rise to negative flux densities on the tomography map – this is as expected, since their spectral index is about 0.5 (Paper I). In the N plume, there is still considerable structure in this image, but the S plume has a much more uniform surface brightness after subtracting the flat-spectrum component, suggesting that a two-component model of the source is close to being adequate here. This is further illustrated in Fig. 6, which shows the results of spectral tomography on slices across the S plume; these show that, at least within 1.5 arcmin of the core, the plume can be modelled as a superposition of a flat-spectrum component with $`\alpha 0.55`$ and a broader steep-spectrum component with $`\alpha 1.2`$, with the flat-spectrum component becoming progressively fainter with distance along the plume; this is consistent with the results of Katz-Stone et al. (krbo (1999)). The spectrum of the flat-spectrum component, as estimated from the spectral index at which it disappears on tomography slices, appears to have steepened by 105 arcsec from the core; this is true even after a rough correction is applied for the effects of the undersampling of the X-band data on large spatial scales (as assessed in section 3.1).
The situation is certainly more complicated in the N plume, where there is in any case less evidence for a steep-spectrum sheath in the spectral index maps of Paper I; if a two-component model is to be viable there, it must allow for some spatial variation in the spectrum of the flat-spectrum component. But this would not be surprising, since there is much stronger evidence for ongoing particle acceleration in the N plume. I return to this point below.
If there are two spectral components, what is the origin of the steep-spectrum material? Katz-Stone & Rudnick identify several possibilities for the sheath in 3C 449. There may be a two-component jet, with the steep-spectrum material only becoming visible at a flare point; or the steep-spectrum material may have evolved from the flatter-spectrum component through ageing, adiabatic expansion, diffusion into a region of lower magnetic field or a combination of these. Without additional low-frequency observations it is impossible to say whether the injection spectral indices of the two components are the same, so we cannot rule out a two-component plume in 3C 130. But it is certainly also possible that the sheath has evolved from the flatter-spectrum component. Modelling of the synchrotron spectrum does not allow me to rule out any of the possibilities; the sheath may be substantially older than the flat-spectrum jet, or it may be of comparable age and in a weaker magnetic field, or a combination of the two. It is possible to say that the two regions cannot simultaneously be in local energy equipartition and be the same age if they have aged in the same B-field.
In any case, it is clear that the steepening of the overall spectrum of the plumes with distance from the source, as discussed in Paper I, is better modelled in terms of a two-component spectral model than in terms of spectral ageing along the jet.
### 4.2 The high-frequency spectra of the plumes
The striking difference between the high-frequency spectra of the N and S plumes (Fig. 4) is unusual in radio galaxies, particularly in a source as symmetrical at low frequencies as 3C 130. It is, of course, possible that the symmetry is illusory and that for some reason the electrons in the S plume are moving much more slowly, and therefore appear to be ageing much more rapidly, than those in the N plume. However, it seems more likely that the spectral difference is related to particle acceleration in the plumes.
In the S plume, there is no clear evidence in any single-frequency map or in the polarization maps for a compact hotspot like the one seen in the N of the source. The two-frequency spectral index maps presented in Paper I show the flat-spectrum S jet penetrating the S plume for some distance, but do not show any particularly flat-spectrum termination region; the best candidate region was in the area of maximal surface brightness at $`40`$ arcsec from the core. From those data it seemed possible that there was a hidden compact hotspot, perhaps suppressed by Doppler beaming, and that the particle-acceleration situations in the two plumes were nevertheless symmetrical. But the 15-GHz data taken together with the absence of a hotspot suggest a model in which there is currently little or no shock-related particle acceleration in the S plume, and consequently no shock-related termination of the jet. The steep 8.4–15-GHz spectrum is inconsistent with continuous injection models for the electron spectrum. If we assume for the ageing $`B`$-field the equipartition field of 0.46 nT used in Paper I, and (as in that paper) use a Jaffe & Perola (jp (1973)) aged electron spectrum then we can estimate the time for which particle acceleration must have been turned off to produce the observed spectrum of the southern plume (Table 1) from an initially power-law spectrum with $`\alpha =0.5`$, as observed in the northern hotspot; it is of order $`5\times 10^7`$ years. This is an appreciable fraction of the commonly assumed lifetime of a radio source, but it is strongly dependent on the assumed ageing $`B`$-field. (Note that, because the region of flux measurement is defined on 15-GHz maps, the plume spectrum used here is essentially that of the flat-spectrum component discussed above, and does not include a contribution from the steep-spectrum sheath.)
## 5 Speculations on source models
If there is no explicit jet termination, how does this fit in with models for WAT formation? We can clearly see a well-collimated jet entering the southern plume. By analogy with the jets in FRIIs we believe this jet to be supersonic, and numerical modelling (e.g. Norman et al. nbs (1988), Loken et al. lrb (1995)) has suggested that to make a WAT a shock should form at the end of the jet, giving rise to the characteristic flaring at the base of the plumes; in any case, we should see some evidence for a transition between the supersonic jet and the diffuse, trans-sonic plume. But there is no evidence for a jet-termination shock either in the form of a hotspot as in the northern plume, as discussed in Paper I, or in particle acceleration, as discussed above. How, then, does the southern jet terminate?
Perhaps the most attractive model is one in which the southern jet currently does not terminate, while the northern jet in 3C 130 is currently impinging on the edge of the source, causing a shock (Fig. 7); the terminations of both jets are dynamic and move about in the base of the plume, and it is only when one intersects with the plume edge that we see a shock and associated particle acceleration. Simply from the small-scale deviations from linearity seen in high-resolution maps of the jets, we know that the point at which the jet enters the plume must vary with time. In FRII radio galaxies, numerical simulations have shown (e.g. Norman n96 (1996)) that the working surface can move about as a result of turbulence in the cocoon. In WATs, bulk motions in the X-ray emitting medium on scales comparable to those of the jets may provide another source of jet buffeting.
This sort of model for 3C 130 seems incompatible with a picture in which the flaring of WAT plumes is caused by propagation across a shock in the external medium (Norman et al. nbs (1988)) or into a crosswind (Loken et al. lrb (1995)), because in these models we would expect to see hotspots on both sides at all times. The latter model is in any case hard to reconcile with the straight plumes seen in 3C 130 and in some other sources which are morphologically WATs (Paper I). Instead, it may be the case that WATs of 3C 130’s type are the natural result of the action of the external medium on a low-power FRII source.
The pressure in the lobes of an FRII is expected to fall with time (e.g. Kaiser & Alexander ka (1997), eq. 20) and if it falls below the thermal pressure of the external medium, the lobes can no longer be supported and will begin to collapse. The natural result is a breaking of the source self-similarity and a slow crushing of the cocoon, which begins with the region closest to the centre, where the thermal pressure is highest (cf. Williams w91 (1991)).<sup>1</sup><sup>1</sup>1This picture differs from the model discussed by Katz-Stone et al. (krbo (1999)), in which WATs are the remnants of FRIIs in which the jets have turned off, and in which cocoon crushing is the force that renders the source WAT-shaped. The detection of localised, compact hotspots which are clearly overpressured with respect to the surrounding emission, and must therefore be assumed to be transient features being supplied with energy by the jet, seems to rule that picture out. In many FRII sources, X-ray observations show that the thermal pressure from the external medium is greater than the minimum pressure in the radio lobes; only the very smallest sources seem to be unambiguously overpressured. Cocoon crushing is therefore a viable process. Once the radio lobes have been squeezed away from the nucleus, asymmetries in the thermal atmosphere, together with buoyancy effects, can account for the large-scale distortions in the lobes seen in many low-power FRII galaxies (Williams w91 (1991)). But, if the environment is suitable, there seems to be no reason why buoyancy cannot drive this process further in particular sources, pushing the lobes further and further away from the centre. The detailed shapes of the sources this would produce would depend on the advance speed of the front of the lobe, but if the lobes were pushed out far enough we would start to see WAT-like objects, provided that the jets continued to terminate at the end of the lobe nearest the centre. At intermediate stages we would see objects like NGC 326 (Worrall et al. wbc (1995)), which differs from a WAT only in that its tails are slightly recessed from the termination of its jets. The evolutionary sequence from FRII to WAT is represented in Fig. 8. The bending of WAT tails on large scales can still, of course, be understood in terms of bulk motions of cluster gas.
If true, this model would imply that we do not expect to see very young (small) WATs; they all evolve from FRIIs and need a certain amount of time (dependent on jet power and properties of the external medium) to do so. It is certainly the case that the WATs studied by O’Donoghue et al. (oeo (1993)) show a range of core–‘hotspot’ distances that begins around 20 kpc and is much smaller than the range of core–hotspot distances seen in classical doubles. What is not clear is whether the timescales of the processes needed to push the lobes of radio galaxies out beyond the jet termination are short enough to be active here. More detailed information on the environments of WATs will become available with the launch of Chandra and XMM; it will need to be coupled with detailed, fully three-dimensional simulations of jets in realistic atmospheres to answer all the outstanding questions on the dynamics of these sources.
###### Acknowledgements.
I am grateful to Larry Rudnick for much helpful discussion on the subjects of spectral tomography and ‘sheaths’, which prompted me to carry out the new observations described in this paper, and to Alan Bridle for allowing me to use his 5-GHz observations of 3C 130. The NRAO Very Large Array is operated by Associated Universities Inc. under contract with the National Science Foundation. This project was supported by PPARC grant GR/K98582.
|
no-problem/9908/patt-sol9908008.html
|
ar5iv
|
text
|
# Stratified spatiotemporal chaos in anisotropic reaction-diffusion systems
(Submitted: 6 April 1999; ; Revised: 28 July 1999)
## Abstract
Numerical simulations of two dimensional pattern formation in an anisotropic bistable reaction-diffusion medium reveal a new dynamical state, stratified spatiotemporal chaos, characterized by strong correlations along one of the principal axes. Equations that describe the dependence of front motion on the angle illustrate the mechanism leading to stratified chaos.
Pattern formation in nonequilibrium systems has been extensively studied in isotropic, two-dimensional media CroHo . Among the most prominent experimental examples are Rayleigh-Benard convection and the Belousov-Zhabotinsky reaction in the contexts of fluid dynamics and chemical reactions respectively. Recently, there has also been considerable interest in systems with broken rotational symmetry, like convection in liquid crystals KraPe and chemical waves in catalytic surface reactions ImbErtl . Experimental and theoretical studies of such anisotropic systems showed novel phenomena like ordered arrays of topological defects chevrons , anisotropic phase turbulence Faller , reaction-diffusion waves with sharp corners corners , and traveling wave fragments along a preferred orientation MertPRE . Anisotropy is also often present in pattern formation processes in biological media, e.g. in cardiac tissue cardiac .
In this Letter we present a new dynamical state that is possible only in anisotropic media - stratified spatiotemporal chaos. We demonstrate the phenomenon with numerical simulations of the bistable FitzHugh-Nagumo equations with anisotropic diffusion and characterize it by computing orientation dependent correlation functions. In addition, an equation for the dependence of front velocities on parameters, curvature, and the angular orientation is derived. The mechanism leading to stratified chaos is described in terms of these analytic results. It is tightly linked to the anisotropy of the system and differs from the mechanism leading to spiral chaos in isotropic bistable media LeePRE ; HaMe94 ; HaMe97 . The findings here are relevant to catalytic reactions on surfaces where anisotropy is naturally provided by crystal symmetry and in biological tissue where anisotropy comes from fiber orientation.
Many qualitative features of pattern formation in chemical and biological reaction-diffusion systems are well described by FitzHugh-Nagumo (FHN) type models for bistable media IMN89 ; HaMeNL . The specific model we choose to study is
$`{\displaystyle \frac{u}{t}}`$ $`=`$ $`ϵ^1(uu^3v)+\delta ^1^2u+{\displaystyle \frac{}{y}}\left[d\delta ^1{\displaystyle \frac{u}{y}}\right],`$
$`{\displaystyle \frac{v}{t}}`$ $`=`$ $`ua_1va_0+^2v,`$ (1)
where $`u`$ is the activator and $`v`$ the inhibitor. The parameters $`a_1`$ and $`a_0`$ are chosen so that Eqs. (1) represent a bistable medium with two stationary and uniform stable states, an “up” state, $`(u_+,v_+)`$, and a “down” state, $`(u_{},v_{})`$. Front solutions connect the two states. The number of front solutions changes, when a single front (an “Ising” front) that exists for values of $`\eta :=\sqrt{ϵ\delta }>\eta _c`$ loses stability to a pair of counter-propagating fronts (“Bloch” fronts) for $`\eta \eta _c`$. The corresponding bifurcation is referred to as the nonequilibrium Ising-Bloch (NIB) bifurcation, hereafter the “front bifurcation.” The anisotropy of the medium is reflected through the parameter $`d`$.
Figure 1 shows the formation and time evolution of stratified chaos obtained by numerical solution of Eqs. (1). The initial state is isotropic spiral chaos. As time evolves a clear orientation of up-state (grey) domains along the $`y`$ direction develops. The domains consist primarily of elongating stripe segments which either merge with other segments or shorten by emitting traveling blobs. The stratified chaos state is robust and develops from a variety of initial conditions including a single spot.
To gain insight we have followed the dynamics of $`u`$ and $`v`$ along the $`x`$ and $`y`$ axes and display them in the form of space-time plots in Fig. 2. A nearly periodic non-propagating pattern along the $`x`$ axis and irregular traveling wave phenomena along the $`y`$ axis are observed. A characteristic property of spatio-temporal chaotic patterns are correlations that decay on a length scale $`\xi `$ much smaller than the system length $`L`$. We have computed the normalized spatial two-point correlation functions, $`C_y(r)`$ and $`C_x(r)`$, for the $`u`$ field in both the $`x`$ and $`y`$ directions, where $`C_y(r)=<\mathrm{\Delta }u(x,y+r)\mathrm{\Delta }u(x,y)>/<\mathrm{\Delta }u(x,y)^2>`$, $`C_x(r)=<\mathrm{\Delta }u(x+r,y)\mathrm{\Delta }u(x,y)>/<\mathrm{\Delta }u(x,y)^2>`$, $`\mathrm{\Delta }u(x,y)=u(x,y)<u>`$, and the brackets $`<>`$ denote space and time averaging. Figure 3 shows the results of these computations. Correlations in the $`y`$ direction decay fast to zero, whereas correlations in the $`x`$ direction oscillate with constant amplitude. This observation may be used to define stratified chaos as a state that displays finite correlation length in one direction ($`x`$) and infinite correlation length in the other ($`y`$).
An important analytical tool for studying front dynamics consists of relations between the normal front velocity $`C_n`$ and other front properties like curvature. Relations of that kind have successfully been used in the study of pattern formation in isotropic media HaMe94 , and we wish to exploit this tool for anisotropic media as well.
Velocity-curvature relations are derived here for $`\lambda :=\sqrt{ϵ/\delta }1`$. The first step in this derivation is to define an orthogonal coordinate system $`(r,s)`$ that moves with the front, where $`r`$ is a coordinate normal to the front and $`s`$ is the arclength. We denote the position vector of the front by $`𝐗(s,t)=(X,Y)`$, and define it to coincide with the $`u=0`$ contour line. The unit vectors tangent and normal to the front are given by $`\widehat{𝐬}=\mathrm{cos}\theta \widehat{𝐱}+\mathrm{sin}\theta \widehat{𝐲}`$ and $`\widehat{𝐫}=\mathrm{sin}\theta \widehat{𝐱}+\mathrm{cos}\theta \widehat{𝐲}`$, respectively, where $`\theta (s,t)`$ is the angle that $`\widehat{s}`$ makes with the $`x`$ axis. A point $`𝐱=(x,y)`$ in the laboratory frame can be expressed as $`𝐱=𝐗(s,t)+r\widehat{𝐫}`$. This gives the following relations between the laboratory coordinates $`(x,y)`$ and the coordinates $`(s,r)`$ in the moving frame: $`x=X(s,t)r\mathrm{sin}\theta (s,t)`$, and $`y=Y(s,t)+r\mathrm{cos}\theta (s,t)`$ where we defined $`\widehat{𝐬}=𝐗/s`$ and $`X/s=\mathrm{cos}\theta `$, $`Y/s=\mathrm{sin}\theta `$. In terms of the moving frame coordinates the front normal velocity and curvature are given by $`C_n=\frac{r}{t}`$ and $`\kappa =\frac{\theta }{s}`$, respectively.
The second step is to express Eqs. (1) in the moving frame and use singular perturbation theory, exploiting the smallness of $`\lambda `$. We distinguish between an inner region that includes the narrow front structure, and outer regions on both sides of the front. In the inner region $`u/r𝒪(\lambda ^1)`$ and $`v/r𝒪(1)`$. In the outer regions both $`u/r`$ and $`v/r`$ are of order unity. In the inner region $`v=v_f`$ is taken to be constant. Expanding both $`u`$ and $`v_f`$ as powers series in $`\lambda `$ and using these expansions in the moving frame equations we obtain, at order $`𝒪(\lambda )`$, a solvability condition that leads to the equation
$$C_n=\frac{3}{\eta \sqrt{2}}I(\theta )v_f\frac{1+d}{\delta I(\theta )^2}\kappa ,$$
(2)
where $`I(\theta )=\sqrt{1+d\mathrm{cos}^2\theta }`$. In the outer regions to the left and to the right of the front region different approximations can be made. Here $`\frac{u}{r}\frac{v}{r}𝒪(1)`$ and to leading order all terms that contain the factor $`\lambda `$ can be neglected. The resulting equations can be solved for $`v`$ in the two outer regions. Continuity of $`v`$ and of $`\frac{v}{r}`$ at the front position $`r=0`$ yield a second relation between $`C_n`$ and $`v_f`$. Eliminating $`v_f`$ by inserting this relation into Eq. (2) gives an implicit relation between the normal velocity of the front and its curvature
$$C_n+\frac{1+d}{\delta I(\theta )^2}\kappa =\frac{3I(\theta )(C_n+\kappa )}{\eta \sqrt{2}q^2\sqrt{(C_n+\kappa )^2+4q^2}}+\frac{3I(\theta )a_0}{\eta \sqrt{2}q^2},$$
(3)
where $`q^2=a_1+1/2`$. A complete account of this derivation will be published elsewhere.
Figures 4 display solutions of Eq. (3) showing the dependence of the front velocity on front curvature and propagation direction for the parameter values of Fig. 1. In Fig. 4a $`C_n`$ vs $`\kappa `$ relations are shown for two orthogonal propagation directions. In the $`x`$ direction, $`\theta =\pi /2`$ (dashed curve), there is only one planar front solution with negative velocity (a down-state invading an up-state), indicating an Ising regime. The negative slope of the $`C_n`$ $`\kappa `$ relation implies stability to transverse perturbations. In the $`y`$ direction, $`\theta =0`$ (solid curve), there are three planar front solutions indicating a Bloch regime. The positive-velocity front (up-state invading down-state) is unstable to transverse perturbations whereas the negative-velocity front is stable. The middle branch corresponds to an unstable front.
Figure 4b shows the angular dependence of planar-front velocities. Counter-propagating (Bloch) fronts exist in a narrow sector around the $`y`$ direction ($`\theta =0`$). The other directions support only a single (Ising) front. The Ising front speed is smallest in the $`x`$ direction ($`\theta =\pi /2`$) and increases as $`\theta `$ deviates from $`\pi /2`$.
The dynamics of fronts as displayed in Fig. 1 are affected by curvature, propagation direction, and front interactions. Eq. (3) captures the effects of the first two factors but does not contain information about front interactions. The necessary information for our purpose can be summarized as follows.
The time evolution of a pair of fronts approaching one another is affected by the speed of non-interacting fronts and by the diffusion rate of the activator $`u`$ (assuming an inhibitor diffusion constant equal to unity as in Eqs. (1)). Consider a pair of fronts pertaining to up-states invading a down-state, propagating toward one another in an isotropic medium ($`d=0`$). If the distance between the fronts decreases below a critical value, $`\lambda _c𝒪(\sqrt{ϵ/\delta })`$, the two fronts collapse, leaving a uniform up-state. The accumulation of the inhibitor in the space enclosed by the fronts, however, slows their motion. If the (non-interacting) front speed is low enough, or if $`\lambda `$ is small enough, there is enough time for the inhibitor to slow the fronts down to a complete stop before they reach the critical distance $`\lambda _c`$. In that case the subsequent evolution depends on the front type. A pair of Ising fronts may either form a stationary pulse or, closer to the front bifurcation, a breathing pulse. A pair of Bloch fronts reflect and propagate away from one another HaMeNL . Thus, high front speed or fast activator diffusion ($`\delta `$ small and $`\lambda _c`$ large) lead to collapse, whereas low front speed and slow activator diffusion lead to strong front repulsion. A similar argument holds for down-state invading up-state fronts.
Returning to anisotropic media we need to know how the two factors that affect front interactions, front speed and activator diffusion, depend on the direction of propagation. The angular dependence of the front speed is already given in Fig. 4b. The angular dependence of the activator diffusion can be deduced by inspecting Eqs. (1). It is $`1/\delta `$ in the $`x`$ direction and $`(1+d)/\delta `$ in the $`y`$ direction. Since $`d>0`$ the activator diffusion constant increases as the propagation direction changes from the $`x`$ direction to the $`y`$ direction.
We can discuss now the mechanism of stratified chaos. As Fig. 4a shows, the $`x`$ direction represents a system that supports stationary or breathing planar stripe patterns. The $`y`$ direction represents a system where traveling waves prevail. The distinct characters of the medium along the two principal axes is reflected in the space-time plots of Fig. 2: nearly vertical columns in the $`x`$ direction indicate stationary or breathing motion, and diagonal stripes in the $`y`$ direction indicate traveling wave phenomena.
The irregular character of the dynamics comes from blob formation events as shown in Figure 5. The front speed and activator diffusion in the $`x`$ direction are sufficiently slow for Ising fronts to repel one another (rather than collapse). As the tip of the stripe segment grows outward and forms a bulge, propagation directions deviating from the $`x`$ direction develop. At these directions both the front speed and the activator diffusion are higher. As a result approaching fronts may collapse. This is exactly the blob pinching process in frames (c), (d) and (e) of Fig. 5. The process is periodic for extended periods of time as indicated by the vertical spot arrays in Fig. 2.
In summary, stratified chaos relies on two main elements: (i) stationary or breathing domains vs traveling wave phenomena in orthogonal directions, and (ii) an angular dependence of front interactions that leads to blob formation. Without the second element stripe segments would merge to ever longer segments until a periodic stripe pattern is formed. These elements suggest the parameter regime where stratified chaos is expected to be found. The first element implies a regime along the front bifurcation such that there is stationary or breathing motion in the direction with faster activator diffusion and traveling waves in the other (slower diffusion) direction. The width of this regime increases with the anisotropy $`d`$. The second element implies that in the direction of breathing or stationary domains the system is close to the onset of breathing motion. Deviations from this direction which move the system toward the traveling wave regime then lead to front collapse. These expectations were verified numerically. Part of the work of U. T. was supported by grant D/98/14745 of the German Academic Exchange Board (DAAD). M. B. gratefully acknowledges the hospitality of Ben-Gurion University and support of the Max-Planck-Society (MPG) by an Otto-Hahn-Fellowship. Part of this research is supported by the Department of Energy, under contract W-7405-ENG-36.
|
no-problem/9908/astro-ph9908359.html
|
ar5iv
|
text
|
# Quasi-radial modes of rotating stars in general relativity
## 1 Introduction
Radial oscillations of non-rotating relativistic stars have been studied for over thirty years. Methods for obtaining their spectra have been well established (Bardeen, Thorne & Meltzer 1966 ; see also Chapter 26 of Misner et al. 1973 and the references therein), and have been applied to several equations of state (see for example Meltzer & Thorne 1966). These works were mainly motivated by consideration of stellar stability because general relativistic effects tend to destabilize stellar models .
On the other hand, the effect of rotation on stellar oscillations is less well understood in a general relativistic context. As in the non-axisymmetric mode case, the slow rotation approximation has been the only accessible way for investigating the eigenmode behaviour of rotating stars . Recently, numerical relativistic hydrodynamic codes have been developed by several authors and some numerical simulations of rapidly rotating stars have been carried out. Stergioulas et al. and Font et al. have shown that initial small perturbations around an equilibrium star evolved to a superposition of normal mode oscillations (Note that their hydrodynamic simulation is done in the fixed background spacetime. On the other hand, Shibata et al. have solved the full system of Einstein equations to investigate the dynamical stability of rapidly rotating stars.).
Although the excitation and evolution of these modes in realistic situations should be investigated by time dependent hydrodynamic simulations, it is also important to study the mode behaviour along rotational equilibrium sequences by using linear perturbation theory.
So far we have studied a few sets of non-axisymmetric eigenmodes of rotating stars in general relativity (f-modes by Yoshida & Eriguchi 1997,1999 ; p-modes by Yoshida 1999 (unpublished; presented at the 9th Yukawa International Seminar ’BLACK HOLES AND GRAVITATIONAL WAVES \- New Eyes in the 21st Century-’) ). These results have been obtained within the Cowling approximation in which Euler perturbations of the metric coefficients have been neglected (see McDermott et al. 1983 and Finn 1988 for a definition ; see Lindblom and Splinter 1990 for the accuracy of the method when applied to non-radial modes of spherical stars). Apart from some low order modes, these results are in good agreement with those of the full perturbation theory including metric perturbations (For comparison of the eigenfrequencies for slowly rotating stars, see Yoshida & Kojima 1997 ; for comparison of the neutral points of the CFS instability, see Yoshida & Eriguchi 1999 which compare the results with the one obtained by full computation of Stergioulas & Friedman 1998 and Morsink et al. 1999)
It is therefore natural to expect that the Cowling approximation could also be successfully applied to the quasi-radial modes which are the smooth extensions of the radial modes of spherical stars to rotating stars. In the present paper, we study quasi-radial modes by using the Cowling approximation. Contrary to the expectations, our results indicate that computations with this approximation can not reproduce the relativistic instability of spherical stars. This is plausible because the instability is essentially caused by the loss of balance between gravity and the pressure gradient, and in calculations of it even the small corrections of gravity cannot be neglected. Moreover, the phase cancellation of the perturbed gravitational potentials, which may be effective in the case of non-axisymmetric modes, cannot be expected to happen for radial modes.
See the Appendix of the present paper for the comparison of two methods in the case of radial modes of non-rotating stars.
Although the validity of Cowling approximation for rotating stars is not fully assessed, we here expect that at least a qualitative picture of the eigenmode dependence on stellar rotation could be studied by this approximation.
## 2 Results
The equation of state used here is the polytropic one,
$$p=\kappa \rho ^{1+\frac{1}{N}},ϵ=\rho +Np$$
$`(1)`$
where $`\rho `$, $`ϵ`$ and $`p`$ are the rest mass density, energy density and pressure of the stellar matter, respectively. Geometrized units, $`c=G=1`$, are adopted in this paper as well as $`M_{}=1`$, following Font et al. . The constant $`N`$ is the polytropic index. The adiabatic exponent of the perturbed matter is assumed to coincide with $`1+1/N`$. The factor $`\kappa `$ is another constant.
Each equilibrium sequence is computed with $`\kappa `$ and $`N`$ fixed.
In the present study polar-like coordinates are used and the metric components are written as:
$$ds^2=e^{2\nu }dt^2+e^{2\alpha }(dr^2+r^2d\theta ^2)+e^{2\beta }r^2\mathrm{sin}^2\theta (d\varphi \omega dt)^2.$$
$`(2)`$
The rotational axis is located at $`\mathrm{sin}\theta =0`$.
The coordinates used in the actual numerical computation are surface-fitted ones $`(r^{},\theta ^{})`$ which are defined as:
$$r^{}=r/R_s(\theta ),\theta ^{}=\theta ,$$
$`(3)`$
where $`r=R_s(\theta )`$ is the form of the stellar surface in equilibrium.
The numerical method used here is basically the same as that in Yoshida & Eriguchi where non-axisymmetric modes were investigated. A minor modification is needed to obtain the axisymmetric modes. In the case of non-axisymmetric modes, the Eulerian variable $`\delta p/(ϵ+p)`$ is explicitly set to zero at the centre of the star ($`\delta p`$ is the Eulerian variation of the pressure). In the case of axisymmetric modes, however, this is not the case since the regularity of the solution requires $`(\delta p)/r`$ to be zero at the stellar centre. Therefore we simply modify the finite-difference scheme at the innermost grid points in our numerical code. Moreover to avoid the coordinate singularity on the rotation axis, points on the axis are excluded from the computational region.
Most of the results shown in this paper are computed with a resolution of 40 uniformly distributed gridpoints in the radial $`r^{}`$ direction and 10 in the angular $`\theta ^{}`$ one. The computational region is a quarter of the meridional section of stars, thus the range of the radial and the angular coordinates are $`0r^{}1`$, $`0\theta ^{}\pi /2`$.
In Figs. 1 and 2 the eigenfrequencies of the axisymmetric modes are plotted against the rotational frequency of the equilibrium model.
The model parameters are tabulated in Table 1.
The eigenfrequency and the rotational frequency are normalized by $`\sqrt{\rho _c/4\pi }`$, where $`\rho _c`$ is the central rest mass density of the models. The sequences ’$`F`$’, ’$`H_1`$’ and ’$`H_2`$’ are the fundamental, first and second overtones of the quasi-radial modes, respectively.
Some of axisymmetric f-modes and overtones of p-modes are also plotted. <sup>1</sup><sup>1</sup>1Our numerical code assumes reflection symmetry of the eigenmodes with respect to the equatorial plane of the equilibrium star. As a result, modes with odd integer $`l`$ cannot be computed. Considering the order of the even $`l`$ modes in the figure, sequences of $`{}_{}{}^{5}\text{f}`$ and $`{}_{}{}^{3}\text{p}_{1}^{}`$ may be located somewhere between the the corresponding p-modes with $`l=2,4`$. These are the continuation of the corresponding f- and p-modes with $`l=L`$ and $`m=0`$ where $`l`$ and $`m`$ are the indices of the ordinary spherical harmonics $`Y_{lm}(\theta ,\phi )`$. The label $`{}_{}{}^{L}\text{p}_{n}^{}`$ refers to a mode corresponding to the p<sub>n</sub>-mode with degree $`l=L`$ and order $`m=0`$ in the non-rotating limit. Similarly, the mode with the label $`{}_{}{}^{L}f`$ is the f-mode with $`l=L,m=0`$ in the non-rotating limit.
Generally the quasi-radial mode sequence encounters other sequences of f- or p-modes. It is seen that the so-called avoided crossing occurs on these sequences (see Fig.3). This seems to be the general relativistic extension of what has been found for the oscillations of rotating Newtonian stars. See for example Clement and Unno et al. for the Newtonian cases.
Fig.3 clearly shows the presence of the avoided crossing whereby two eigenfrequency curves approach smoothly, and then depart from each other without crossing. At the point of closest approach, the characteristics of the modes on each sequence exchange. Thus the sequence of a mode with given characteristics has a discontinuity there.
There are several discontinuities along a sequence whose number depends on the mode order as well as on the polytropic parameters of the equilibrium star. For the selected polytropic parameters, discontinuities appear in the rapidly rotating models whose rotational frequencies are over $`80\%`$ of the mass-shedding limit.
## 3 Some remarks on the analysis
### 3.1 Convergence and accuracy
Since we use a finite number of grid points, our results necessarily contain some errors. To see how much the results differ when the number of grid points is changed, we compute the fundamental quasi-radial mode sequence by using different grid numbers $`(M,N)=(40,10),(40,20),(80,10)`$ where $`M`$ is the number of grid points in $`r^{}`$-direction and $`N`$ is that in $`\theta ^{}`$-direction. By extrapolating three eigenfrequencies obtained from these grid numbers, $`\nu _{[40,10]},\nu _{[40,20]},\nu _{[80,10]}`$, we estimate the converged value of the eigenfrequency $`\nu _0`$ in the limit of infinitesimal mesh size to be:
$$\nu _0=2(\nu _{[40,20]}+\nu _{[80,10]})3\nu _{[40,10]}.$$
$`(4)`$
The ratio of $`\nu _{[M,N]}`$ to $`\nu _0`$ can be treated as a measure of convergence of the results when the grid number is changed (Fig.4). As can be seen from this figure, eigenfrequencies obtained by using $`(M,N)=(40,10)`$ mesh numbers (our standard resolution) agree with the converged values to within 3 percent.
It is noted that the relative error of $`\nu _{[80,10]}`$ is smaller than that of $`\nu _{[40,20]}`$ in the slowly rotating cases, however for rapidly rotating cases, the error of $`\nu _{[80,10]}`$ becomes larger. This is because for the rapidly rotating cases, the deformation of the stars from the spherical configuration is so large that the lack of angular resolution becomes the main source of inaccuracy.
To check the accuracy of our two dimensional (2D) code, we have compared the results obtained here with those obtained with the one dimensional (1D) code described in the Appendix (Table 4). In the 1D code we employ the standard scheme to solve the eigenvalue problem of the linear ordinary differential equation within the Cowling approximation.
As seen from this table, two results agree well to within several percent.
### 3.2 Eigenfunctions and classification of modes
Comparing with the eigenfunctions of non-axisymmetric modes, we notice that the eigenfunctions of quasi-radial modes change their shapes significantly along rotational equilibrium sequences. For example, the radial distributions of the Eulerian pressure perturbation and the radial component of the velocity perturbation change considerably near the equatorial plane of the star. The number of radial nodes of these functions increases as the stellar rotation rate increases. On the other hand, the overall radial dependence of the $`\theta `$-component of the velocity perturbation changes little as the star spins up. Therefore we can use the shape of this function as a tracer of the selected mode along the equilibrium sequence. The nomenclature of mode sequences is based upon the behaviour of the mode of a non-rotating star to which the sequence is continued smoothly.
### 3.3 Realistic neutron star models
In addition to the polytropic case presented here, we have tried to compute quasi-radial modes of realistic neutron stars by using some of the candidate zero temperature equations of state. However it was rather difficult to obtain full sequences of these modes with our method. As the star begins to rotate, the convergence to the quasi-radial modes suddenly becomes much more difficult. This may partly be due to the fact that these modes are sensitive (as compared with the non-radial modes) to the surface condition of the equilibrium star. Unfortunately, in the case of more realistic EOS, the adiabatic exponent is subject to large oscillations, becoming negative in some parts. These large variations decrease considerably the accuracy of our method which then becomes inadequate.
## Acknowledgment
We thank Luciano Rezzolla for detailed comments on the manuscript, and John Miller for his help to improve it. We also thank Kōji Uryū for useful discussions.
## Appendix A Comparison of radial modes in the full theory and in the Cowling approximation
To test the validity and accuracy of the Cowling approximation for a spherical configuration, we here present a comparison between the results obtained by the full perturbation theory and by the Cowling approximation for low order radial modes.
### A.1 Equilibrium model
The space-time of the equilibrium star is characterized by the following metric:
$$ds^2=e^{2\nu (r)}dt^2+e^{2\lambda (r)}dr^2+r^2(d\theta ^2+\mathrm{sin}^2\theta d\phi ^2).$$
$`(A1)`$
The metric coefficients, stellar pressure $`p`$ and energy density $`ϵ`$ are obtained by integrating the standard set of equations with regular boundary conditions at the stellar centre:
$$\frac{dp}{dr}=\frac{(ϵ+p)(m+4\pi pr^3)}{r(r2m)},$$
$`(A2)`$
$$\frac{dm}{dr}=4\pi r^2ϵ,$$
$`(A3)`$
$$\frac{d\nu }{dr}=\frac{m+4\pi pr^3}{r(r2m)},$$
$`(A4)`$
where $`m(r)`$ is defined from $`e^{2\lambda }=(12m/r)^1`$.
The equation of state is assumed to be polytropic, $`p=\kappa \rho ^{1+\frac{1}{N}}`$, where $`\kappa `$ and N are constants, and the rest-mass density $`\rho `$ is related to $`ϵ`$ by $`ϵ=\rho +Np`$.
### A.2 Equations for radial oscillations
For radial oscillations of spherical stars, only the Lagrangian displacement function $`\xi (r)`$ is needed to describe the physical perturbation (see Chapter 26 of Misner et al. 1973).
The equation of motion of the displacement ($`r^2e^\nu \xi \zeta `$) in the full perturbation theory is:
$$\zeta ^{\prime \prime }+A_f\zeta ^{}+B_f\zeta =0,$$
$`(A5)`$
where $`A_f`$ and $`B_f`$ are defined as:
$$A_f\frac{p^{}}{p}\frac{2}{r}+\lambda ^{}+3\nu ^{},$$
$`(A6)`$
and
$$B_f\frac{ϵ+p}{\mathrm{\Gamma }p}\left[(\nu ^{})^2+4\frac{\nu ^{}}{r}8\pi pe^{2\lambda }+\sigma ^2e^{2\lambda 2\nu }\right].$$
$`(A7)`$
The prime after a variable refers its derivative with respect to $`r`$. Here $`\sigma `$ is the frequency and the adiabatic exponent $`\mathrm{\Gamma }`$ is defined by:
$$\mathrm{\Gamma }\frac{ϵ+p}{p}\frac{\mathrm{\Delta }p}{\mathrm{\Delta }ϵ},$$
$`(A8)`$
where $`\mathrm{\Delta }`$ represents the Lagrangian perturbation. In general the adiabatic exponent need not coincide with $`1+1/N`$.
In the Cowling approximation, the equation of motion is instead ($`r^2e^\lambda \xi \eta `$):
$$\eta ^{\prime \prime }+A_c\eta ^{}+B_c\eta =0,$$
$`(A9)`$
where $`A_c`$ and $`B_c`$ are defined as:
$$A_c\frac{p^{}}{p}\frac{2}{r}\lambda ^{}+\nu ^{},$$
$`(A10)`$
and
$$B_c\frac{p^{}}{\mathrm{\Gamma }p}\left(\frac{2}{r}+\lambda ^{}\right)+\frac{ϵ+p}{\mathrm{\Gamma }p}(\nu ^{\prime \prime }+\sigma ^2e^{2\lambda 2\nu }).$$
$`(A11)`$
These equations can be solved by the matching method: i.e., we have to search for $`\sigma ^2`$ which makes the Wronskian of the solutions, obtained by integrations from the stellar centre and from the surface, vanish at some matching point inside the star.
### A.3 Boundary conditions
The boundary condition for the equation of oscillations at the centre of the star is regularity of the variables. It requires $`\zeta ,\eta r^3`$ as $`r0`$.
At the surface of the star, we impose the boundary and regularity conditions which reduce to
$$\zeta ^{}+\frac{ϵ+p}{\mathrm{\Gamma }p^{}}[(\nu ^{})^2+4\nu ^{}r^1+\sigma ^2e^{2\lambda 2\nu }]\zeta =0,$$
$`(A12)`$
or
$$\eta ^{}+\left[\frac{1}{\mathrm{\Gamma }}(2r^1+\lambda ^{})+\frac{ϵ+p}{\mathrm{\Gamma }p^{}}(\nu ^{\prime \prime }+\sigma ^2e^{2\lambda 2\nu })\right]\eta =0.$$
$`(A13)`$
### A.4 Results
Keeping the polytropic index $`N`$ fixed, we vary the compactness (mass-to-radius ratio) of the model to construct an equilibrium sequence.
In Figs. A1 and A2 we show the typical sequences of the three lowest order modes. For all of the sequences, the eigenfrequency obtained by the Cowling approximation is larger than that from the full theory: i.e., the Cowling approximation overestimates the stability of the star. Before the turning point,<sup>2</sup><sup>2</sup>2Note that with this parametrization, the turning point does not correspond to the maximum mass configuration. Thus the zeroes of the fundamental radial modes are not the turning points. the two curves are nearly parallel. As expected, the relative error of the results obtained by the Cowling approximation becomes smaller for higher overtones.
|
no-problem/9908/cond-mat9908136.html
|
ar5iv
|
text
|
# Diffusion of hydrogen in crystalline silicon
\[
## Abstract
The coefficient of diffusion of hydrogen in crystalline silicon is calculated using tight-binding molecular dynamics. Our results are in good quantitative agreement with an earlier study by Panzarini and Colombo \[Phys. Rev. Lett. 73, 1636 (1994)\]. However, while our calculations indicate that long jumps dominate over single hops at high temperatures, no abrupt change in the diffusion coefficient can be observed with decreasing temperature. The (classical) Arrhenius diffusion parameters, as a consequence, should extrapolate to low temperatures.
Submitted to Physical Review B
\]
In spite of the tremendous efforts that have been engaged in determining the rate of diffusion of hydrogen in crystalline silicon, a concensus on the “true value” has not yet been reached. Of course, the diffusion constant varies strongly with temperature — not necessarily in a perfect Arrhenius manner — making a precise determination of the diffusion parameters difficult. Other complications arise from possible collective effects (as opposed to tracer diffusion), low-temperature quantum effects, impurities and defects, etc.
Experimental estimates of the diffusion constant are numerous and vary widely, sometimes by two orders of magnitude at a given temperature, as can be appreciated from the open circles in Fig. 1(a). This is an unpleasant state of affairs, since precise knowledge of this quantity is important for both practical and fundamental reasons: Because it forms complexes with a variety of defects, hydrogen affects deeply the optical and electronic properties of semiconductors. It is usually present as a result of the fabrication process, but is often intentionally introduced in order to passivate defects. Being a light species, further, H diffuses readily, inducing additional defects along its way, thus affecting the transport properties of the material to an extent which is determined by its relative concentration. It is therefore important to understand diffusion at the atomic level in order to gain better control on the properties of semiconductors and, in view of the simplicity of the structure of the host material, it is of fundamental importance to be able to understand this prototypical system.
The diffusion coefficient can be estimated, at sufficiently high temperature, using the now well-established molecular-dynamics method; a proper model for the interatomic potentials is then needed. Empirical potentials lack the transferability and predictive power of first-principles methods. The latter, however, are subject to limitations in size and time, which makes them unpractical for the long simulations required for a proper (statistically meaningful) estimate of the rate of diffusion. In an early application of the Car-Parrinello method, Buda et al. calculated the coefficient of diffusion of H<sup>+</sup> in Si at three temperatures in the range 1200–1950 K, covering a minuscule maximum observation time of 4 ps.
The semi-empirical, semi-quantum tight-binding molecular-dynamics (TBMD) scheme, originally proposed by Khan and Broughton and Goodwin, Skinner and Pettifor (GSP), provides good accuracy at a very reasonable computational cost. Here, the attractive part of the atom-atom interactions is described quantum-mechanically using (parametrized) overlap (or hopping) integrals, while the repulsive part is fitted from known properties of the system, e.g., binding energy vs distance. The forces are then derived from the Hellman-Feynman theorem. TBMD models for Si:H were proposed by Panzarini and Colombo (PC), Boucher and DeLeo (BDL), and Kim, Lee and Lee; all three models are based on the GSP model for Si–Si interactions, and use comparable fitting schemes to take into account Si–H and H–H interactions.
The problem of hydrogen diffusion in silicon was addressed using TBMD by Panzarini and Colombo as well as by Boucher and DeLeo; both considered a single H atom in a 64-atom c-Si supercell. While extending significantly the range of temperatures that were covered by the ab initio simulations of Buda et al. (1050–2000 K for BDL, 800–1800 K for PC), the timescales covered by these MD calculations remain short (42 and 300 ps, respectively): indeed, for a diffusion constant of $`10^6`$ cm<sup>2</sup>/s — as found at about 800 K — a quick calculation indicates that the average size of the region visited by the diffusing particle over 300 ps would be about 4 Å, corresponding roughly to the second-neighbour distance in c-Si. Further, the two calculations exhibit some disagreement which might be inherent to the models or, more likely, to the statistical quality of the MD data.
In this short note, we revisit the problem using, again, TBMD (PC version), but with much longer timescales: our simulations ran during a formidable 7 nanoseconds at the lowest temperature we could decently examine — 700 K, which is 100 K below the lowest temperature looked at by PC. Our calculations generally confirm PC’s results, in particular the discrepancies with experiment observed at the lowest temperatures. However, while our calculations indicate that long jumps dominate over single hops at high temperatures, the diffusion coefficient exhibits no abrupt change with temperature. We are led to conclude that the Arrhenius diffusion parameters should extrapolate to low temperatures as far as the classical part of the motion is concerned.
As previewed earlier, Fig. 1(a) presents the results of several measurements of the diffusion constant, plotted à la Arrhenius. Also indicated are the ab initio MD data of Buda et al.; they are found to be in reasonable agreement with the high-temperature experimental points of Van Wieringen and Warmoltz, fitted to the Arrhenius law $`D(T)=D_0\mathrm{exp}(E_A/k_BT)`$, with $`D_0=9.41\times 10^3`$ cm<sup>2</sup>/s and $`E_A=0.48`$ eV. When extended to low temperatures \[dotted line in Fig. 1(a)\], one clearly sees the deviations from the Arrhenius behaviour; it should be said, however, that there is no “guarantee” that diffusion should be Arrhenius over the whole range of temperatures.
The TBMD data of PC and BDL are also plotted in Fig. 1(a), and more legibly in Fig. 1(b). The agreement with experiment is clearly excellent at high temperature — certainly within the errors that can be associated with both measurements and calculations. The data of BDL are found to be extremely well fitted by the Arrhenius law with $`D_0=6.91\times 10^3`$ cm<sup>2</sup>/s and $`E_A=0.45`$ eV all the way down to 1050 K, in striking agreement with experiment, as can be judged by the close similarity between the prefactors and energy barriers. PC, in contrast, observe significant deviations from Arrhenius already at 1200 K, a problem that can possibly be attributed to the statistical quality of their data.
In Fig. 2 we give, as an example, the calculated time-dependence of the mean-square displacement for our lowest-temperature run, viz. 700 K. The simulation at this temperature ran for a total of 7 ns and the mean-square displacement is calculated for a maximum correlation time of 1 ns in order to minimize statistical uncertainties. We also evaluated the mean-square displacement using only the first half of the run, then only the second half so as to have a feeling for the “error bar” of our estimate of the diffusion coefficient, $`D=lim_t\mathrm{}r^2(t)/6t`$. The three different calculations are indicated in Fig. 2 by full, dashed, and dotted lines, respectively. The corresponding coefficients of diffusion, which we discuss next, are presented in Fig. 1; for each temperature we investigated (700, 800, 900, 1000, 1200, and 1500 K), three data points are thus given.
As can be appreciated from Fig. 1(b), our estimates of the diffusion coefficient generally agree with the values reported by PC, but clearly with much improved statistical quality: the data points exhibit very little scatter, falling along a single, rather well-defined straight line. It is of course hazardous to assume that the diffusion process is perfectly Arrhenius and that no “exotic” diffusion mechanisms (i.e., other than single hops) are taking place. In order to “guide the eye”, however, we fitted our data to the law $`D=8.9\times 10^3\mathrm{exp}(0.58\mathrm{eV}/k_BT)`$ cm<sup>2</sup>/s; this is displayed as the dashed line in Fig. 1. Our calculations, clearly, show no evidence of a sizeable change in the diffusive behaviour as a function of temperature, as can possibly be (and was) inferred from the data of PC.
Exotic mechanisms — in the present case long jumps — are in fact present; this has already been noted by PC, who also found long jumps to be quenched in as temperature drops; this observation was based on a visual examination of the mean-square displacements. A physically more appropriate characterization of single-atom motion is provided by the self part of the van Hove correlation function (see, e.g., Ref. ), $`G_s(𝐫,t)`$, which gives the probability of finding a particle at $`𝐫`$ at time $`t`$ given that it was at the origin at $`t=0`$. Averaging out over angular space, the function of interest is $`4\pi r^2G_s(r,t)`$.
The van Hove self-correlation function for the H atom in c-Si at 700 K is displayed in Fig. 3. Here, $`4\pi r^2G_s(r,t)`$ is plotted as a function of distance for four different times. At short time (but somewhat longer than the typical vibrational period), single hops dominate the motion. As time increases, the particle is (of course) on average further away and, evidently, the diffusive motion proceeds via a sequence of single jumps — to a reasonable approximation.
Because diffusion is activated, it is more meaningful, in order to compare the modes of diffusion at different temperatures, to examine the behaviour of $`4\pi r^2G_s(r,t)`$ at fixed mean-square displacement. Thus, it is possible in this way to determine the type of motion that leads a particle a certain distance away from its original position. Here we chose (somewhat arbitrarily) a fixed mean-square displacement of 50 Å<sup>2</sup>. From the $`r^2(t)`$ curves, it is easy to determine the time at which, on average, and for each temperature, the particle will be at the required fixed position.
The results of this analysis are presented in Fig. 4. Again, here, we find that the motion consists of single hops at low temperatures, but acquires a “long-jump” character as temperature increases. At the highest temperatures, in fact, the distribution is nearly continuous, and single jumps become almost undetectable. There is some evidence, in these plots, of a discontinuous change in the manner that diffusion proceeds as a function of temperature: upon going from 800 to 900 K, nearest-neighbour jumps are found to become negligible, to the advantage of second- and further-neighbour hops. But then no sign of such a change is apparent in the diffusion coefficient (cf. Fig. 1), which perhaps lacks the sensitivity needed to reveal such details. Thus, we are led to conclude that the low-temperature ($`T<`$ about 900 K) diffusion coefficient is indeed Arrhenius-like, whereas high-temperature diffusion proceeds via a complicated sequence of jumps whose energy barriers combine into a single activation energy that is close to that for single hops.
In view of this, there are no reasons to believe that this Arrhenius law should not extrapolate to very low temperatures, if the system were classical. This is clearly not the case here and it is of course expected that diffusion will be enhanced by quantum contributions, even more so as temperature decreases. We have not included quantum corrections in the present study: this is a difficult problem and it is not clear that their role is significant at the temperatures considered here. The small discrepancies between TBMD and experimental data at high temperatures might be related to quantum effects, but are perhaps more likely due to some limitations of the model, i.e., are not significant. Likewise, Fig. 1 seems to indicate some partial “recovery” of the experimental data at very low temperature. This might be a manifestation, again, of quantum contributions. Clearly a more consistent set of experimental data is needed in order to have a proper perspective on the problem.
Aknowledgements — We are grateful to Normand Mousseau for useful suggestions and to Edward Hernandez for providing us with a TBMD code for the Goodwin-Skinner-Pettifor Si model. We also thank Luciano Colombo for useful advice on implementing the TBMD code for Si:H and for providing us an electronic version of Fig. 1 in Ref. . This work is supported by grants from the Natural Sciences and Engineering Research Council (NSERC) of Canada and the “Fonds pour la formation de chercheurs et l’aide à la recherche” (FCAR) of the Province of Québec.
|
no-problem/9908/hep-lat9908026.html
|
ar5iv
|
text
|
# Quantum gauge fixing and vortex dominance UNITU-THEP 11/99
## 1 Introduction
In recent years, evidence has been accumulated that the mechanism of quark confinement may be understood in effective theories of monopoles or $`Z_N`$ vortices -. After gauge fixing, these theories arise from projecting the full SU(N) onto a gauge theory with a reduced gauge symmetry. The observation that the reduced theories reveal the full string tension nurtures the conjecture that those degrees of freedom bear confinement. Here, we will employ center gauge fixing and center projection for the reduction of SU(2) to $`Z_2`$ gauge theory which can be understood as a theory of vortices. We define vortex dominance if two criteria are met. Firstly, the string tension is preserved by projection. Secondly, the vortices survive the continuum limit.
## 2 Center gauge fixing (standard)
Let $`U_\mu (x)`$ denote the link variable of SU(2) gauge theory and $`\mathrm{\Omega }(x)`$ a gauge transformation matrix. Maximizing the functional
$$S_{fix}[U]=\underset{\{x\}\mu }{}\frac{1}{2}\mathrm{tr}\left\{U_\mu ^\mathrm{\Omega }(x)\tau ^aU_\mu ^\mathrm{\Omega }(x)\tau ^a\right\},$$
(1)
with respect to $`\mathrm{\Omega }(x)`$ yields the gauge transformation matrices which, when applied, cast a given configuration in the center gauge. Center projection is performed by the mapping of the link variable $`U^\mathrm{\Omega }(x)`$ onto $`\pm 1`$ (see e.g. ).
In practical calculations, finding the absolute maximum of the functional $`S_{fix}`$ is a difficult task. The results which are presented in the next section were obtained by applying the algorithm presented in , which resorts to iteration with over-relaxation. Once the iteration does not change the functional $`S_{fix}`$ any more up to a given precision, we performed a random gauge transformation on the actual link configuration and repeat the center gauge fixing. The procedure is repeated up to ten times. We finally choose the configuration $`\mathrm{\Omega }(x)`$ which corresponds to the maximum value within the series of the ten fixing steps. This alleviates, but does not eliminate, the Gribov problem (see section 4).
## 3 Numerical results
Zero temperature: It turns out that the effective $`Z_2`$ gauge theory which was constructed with the ITOV-algorithm sketched above shows vortex dominance: the string tension is preserved , and the linking number of the vortex lines with a Wilson loop (vortex density) meets with the expectations from a renormalization group analysis . Also the vortex interactions scale. These observations suggest that the vortices are physical objects surviving the continuum limit .
Finite temperatures: We find that below the deconfinement temperature $`T_c`$ vortex dominance persists . In particular, the effective $`Z_2`$ gauge theory correctly reproduces the critical temperature thus indicating that the essence of the deconfinement phase transition can be captured in the vortex picture. A thorough study of the density of vortices which are linked to spatially oriented Wilson loops shows that the vortex state at high temperatures is in agreement with the predictions of dimensional reduction . On the other hand, the density of vortices linked with time-like Wilson loops only drops by a factor of three if the temperature is raised to twice $`T_c`$. Additional information is needed to understand the drop of the string tension to zero at $`T_c`$. We argue that while the vortices are organized in a huge cluster at zero temperature, they stop percolating at $`T_c`$ and the huge cluster decays into many small size clusters. If in the later case the size of the Wilson loop exceeds the average cluster size, only vortices close to the circumference of the Wilson loop contribute yielding a perimeter law. In order to substantiate this idea, we measured the probability that a link of the vortex belongs to a cluster of given size. The result is depicted in figure 1. Further details can be found in .
## 4 Gribov ambiguities
The variety of results in the previous section was obtained by implementing the center gauge (see (1)) with the help of the ITOV algorithm sketched in section 2. It was recently pointed out that this procedure does not evade the Gribov problem . In fact, it was observed that implementing the Landau gauge before applying the ITOV algorithm leads on average to a larger maximum value of $`S_{fix}`$ (1) than attained by the direct use of the ITOV algorithm. Moreover, the vortex state constructed from the gauge fixing via the Landau gauge detour does not show vortex dominance. These results cast doubt on the issue of vortex dominance in the strict maximal center gauge and call for a more accurate specification of the gauge in which the results outlined in section 3 were really obtained.
## 5 Quantum gauge fixing
For this specification, we here propose a new type of gauge fixing procedure which we will call Quantum gauge fixing (QGF). We will argue that the ITOV algorithm of section 2 already contains some of the characteristics of QGF rather than representing a numerically stable implementation of the maximal center gauge. Defining the $`SU(N)/Z_N`$ matrices
$`\zeta _{ab}(x)`$ $`:=`$ $`{\displaystyle \frac{1}{2}}\mathrm{tr}\left\{\mathrm{\Omega }(x)\tau ^a\mathrm{\Omega }^{}(x)\tau ^b\right\},`$ (2)
$`R_{ab}^\mu (x)`$ $`:=`$ $`{\displaystyle \frac{1}{2}}\mathrm{tr}\left\{U_\mu (x)\tau ^aU_\mu ^{}(x)\tau ^b\right\},`$ (3)
the gauge fixing functional $`S_{fix}`$ (1) can be cast into
$$S_{fix}=\underset{\{x\}\mu }{}\mathrm{tr}\zeta ^T(x+\mu )R^\mu (x)\zeta (x).$$
(4)
Center gauge fixing corresponds to maximizing (4) with respect to the matrices $`\zeta `$ for a given lattice configuration, i.e., $`R^\mu (x)`$. From given matrices $`\zeta (x)`$, the gauge transformation $`\mathrm{\Omega }(x)`$ can be constructed up to a center gauge transformation,
$$f_{SU}^Z:SU(N)/Z_NSU(N):\mathrm{\Omega }=f_{SU}^Z(\zeta ).$$
(5)
This ambiguity reflects the familiar fact that the maximal center gauge condition leaves a residual $`Z_N`$ gauge group unfixed.
We define the QGF as follows: by means of a functional integral over gauge transformations $`\zeta `$, we construct the matrix
$$\omega [U]:=\frac{1}{N}𝒟\zeta f_{SU}^Z(\zeta )\mathrm{exp}\left\{\beta _fS_{fix}\right\},$$
(6)
which in general is not an element of the gauge group. The gauge transformation $`\mathrm{\Omega }(x)`$ which brings a given field configuration $`U_\mu (x)`$ into the quantum gauge is then defined by the SU(N) element ”closest” to $`\omega [U]`$, i.e.,
$$\omega [U](x)\mathrm{\Omega }(x)^2\text{min},x,$$
(7)
where $`A^2:=\mathrm{tr}AA^{}`$. It can be shown that the QGF (6-7) is free of Gribov ambiguities. In the case of the center gauge fixing (4), the functional integral in (6) can be viewed as a partition function of the matrices $`\zeta `$ each interacting with its nearest neighbors via the metric $`R^\mu (x)`$. This quantum theory of matrices $`\zeta `$ is therefore interpreted as generalized spin glass. Spin glass systems are known for a complex phase structure. A numerical calculation of the expectation value (6) is therefore tedious, and we recover the Gribov problem in practical applications.
Nevertheless, QGF is helpful to put the results outlined in section 3 in the proper context. For this purpose, we consider large values of the gauge parameter $`\beta _f`$ in (6). In this case, the configuration $`\zeta (x)`$ which corresponds to a local maximum of $`S_{fix}`$ contributes to $`\omega `$ with the (semi-classical) weight
$$\mathrm{exp}\left\{\beta _fS_{fix}\left[\zeta \right]\right\}/\left[S_{fix}^{\prime \prime }\left[\zeta \right]\right]^{1/2},$$
(8)
where $`S_{fix}^{\prime \prime }`$ is the functional determinant of the second (functional) derivative of $`S_{fix}`$ with respect to $`\zeta (x)`$. Eq.(8) implies that the contributions of maxima with large curvature are suppressed. The situation is illustrated in figure 2. Since the ITOV algorithm of section 2 involves several steps of iteration which start with a random choice of $`\zeta `$, it probes the volume of the region of attraction corresponding to a particular maximum. The ITOV algorithm therefore effectively contains a similar entropy factor as the one implied by (6). We therefore argue that the results shown in section 3 refer to a gauge which is more akin (but not completely identical) to QGF than to the strict maximal center gauge.
Supported in part by DFG-RE 856/4-1, DFG-En 415/1-1.
|
no-problem/9908/cond-mat9908242.html
|
ar5iv
|
text
|
# Critical Current in the High-𝑇_𝑐 Glass model
## I Introduction
The high-$`T_c`$ glass model was introduced in 1987 to describe superconductivity in the high-$`T_c`$ cuprates (HTSC) . Originally the glass model was designed for s-wave symmetry of the superconducting wave functions . Whereas it was first argued , that the glass model is not applicable for d-wave symmetry, the experimental result of striped domains in the superconducting CuO-planes inside the high-$`T_c`$ materials only gives rise to weak disorder. Thus the high-$`T_c`$ glass model is applicable in this situation, too . It was demonstrated , that the high-$`T_c`$ glass model including the tt’–Hubbard model as a microscopic description of the striped superconducting domains is able to explain e.g. the d-wave symmetry of the superconducting phase and the pseudogap above $`T_c`$ in the density of states (DOS) . In the combined high-$`T_c`$ glass and striped Hubbard model picture the stripes in the HTSC occur at least at the same temperature, at which the pseudogap in the DOS opens, because otherwise the striped Hubbard clusters, which are responsable for these gaps, do not exist.
In this paper we will show, that within the high-$`T_c`$ glass model the maximum critical currents found in high-$`T_c`$ materials and their almost perfect isotropy in a-b-direction can be understood. Following two independent paths we calculate in the next section an upper bound for the critical current density $`j_c`$: first with a direct calculation from the high-$`T_c`$ glass model and second considering the extended Bean model . Afterwards we offer an intuitive picture for the (observed) isotropy of the critical currents and explain, why the values of $`j_c`$ usually measured in high-$`T_c`$ materials are smaller than these upper bounds. From these simple ideas follows a fabrication procedure, which could lead to a possible increase of $`j_c`$ in the high-$`T_c`$ materials.
## II Critical Currents in the High-$`T_c`$ glass model
In the high-$`T_c`$ glass model a single superconducting CuO-plane is described as an array of striped domains, figure 1, . The typical dimension of a single striped domain with constant superconducting phase inside the high-$`T_c`$ glass model can be roughly estimated following e. g. Tsuei and Doderer and Tranquada et. al. . We take for our calculation the values $`a100`$ Å, $`b10`$ Å, (e.g. $`b16`$ Å in LaSrCuO or $`b23`$ Å in YBaCuO ) and $`z10`$ Å with the directions of figure 2. This choice of $`a`$, $`b`$, and $`z`$ is not crucial for our considerations, where only their order of magnitude is important for the conclusions.
In the following calculation we show that the glass-model leads to critical current densities, which are in agreement with the experiments. The starting point is the Hamiltonian for the glass model :
$$=J\underset{i,k}{}\mathrm{cos}\left(\varphi _i\varphi _kA_{i,k}\right).$$
(1)
The phase factors $`A_{i,k}`$ are given by
$$A_{i,k}=\frac{2\pi }{\mathrm{\Phi }_0}_i^j\stackrel{}{A}d\stackrel{}{l}$$
(2)
with $`\mathrm{rot}\stackrel{}{A}=H\widehat{z}`$, where $`\widehat{z}`$ is the unity vector in z-direction, and with $`A_{i,k}=a_{i,k}H`$ we get
$$a_{i,k}=\frac{2\pi }{\mathrm{\Phi }_0}\frac{x_i+x_k}{2}\left(y_ky_i\right).$$
(3)
In equation (1) up to (3) $`\mathrm{\Phi }_0`$ is the elementary flux quantum, $`\varphi _i`$ are the phases of the superconducting wave functions, $`J`$ is the coupling energy between two clusters, $`i,k`$ is the sum over all nearest neighbors, and $`x_i`$, $`x_k`$, $`y_i`$, and $`y_k`$ are the coordinates of the center of gravity of the domains $`i`$ respectively $`k`$. Finally $`H`$ is the external magnetic field and $`\stackrel{}{A}`$ the corresponding vector potential .
In the high-$`T_c`$ glass model only weak, ”correlated” disorder was chosen in the framework of the square lattice . We should note, that this weak disorder in the high-$`T_c`$ glass model can be described as $`x_iia`$ and $`y_iib`$ with the lattice constants $`a`$ in x-direction and $`b`$ correspondingly in y-direction .
To obtain the critical current density $`j_{i,k}`$ for a connection between domain $`i`$ and $`k`$ we consider the Maxwell equation
$$\mathrm{rot}H_{i,k}=j_{i,k}.$$
(4)
Next we assume a constant increase of the internal magnetic field analogous to the assumptions of the Bean model :
$$\mathrm{rot}H_{i,k}=\frac{H_{i,k}}{\frac{x_i+x_k}{2}}.$$
(5)
The free energy $`F`$ is given as $`F=k_BT\mathrm{ln}Z`$, where the partition function $`Z=_{\{\mathrm{\Phi }\}}\mathrm{exp}\left(\beta (\mathrm{\Phi })\right)`$ with the inverse temperature $`\beta 1/(k_BT)`$ and the sum $`\{\mathrm{\Phi }\}`$ is over all possible configurations of the phases $`\mathrm{\Phi }=(\varphi _1,\mathrm{},\varphi _{L^2})`$ for a lattice with $`L\times L`$ sites ($`k_B`$ is the Boltzmann constant). The magnetization $`M`$ of a sample is given by $`M=1/VF/H`$, where $`V`$ is the volume of the sample. For the external magnetic field $`H=0`$ we obtain:
$$M=\frac{1}{V}\frac{2\pi }{\mathrm{\Phi }_0}J\underset{i,k}{}\frac{x_i+x_k}{2}\left(y_ky_i\right)\mathrm{sin}\left(\varphi _i\varphi _k\right)$$
(6)
where $`\mathrm{}`$ dennotes the thermal expectation value.
The magnetization can be expressed in terms of the internal magnetic fields $`H_{i,k}`$
$$M=\frac{1}{2L^2}\underset{i,k}{}H_{i,k}.$$
(7)
Inserting equation (6) into equation (7), using the abbreviation $`I=J2\pi /\mathrm{\Phi }_0`$ and the volume $`V=LaLbz`$ (geometries of a striped domain, figure 2) it follows
$$H_{i,k}=\frac{1}{abz}2I\frac{x_i+x_k}{2}\left(y_ky_i\right)\mathrm{sin}\left(\varphi _i\varphi _k\right).$$
(8)
Using the equations (4), (5), and (8) we get for the critical current density of the domains $`i`$ and $`k`$:
$$j_{i,k}=\frac{I}{az}\mathrm{sin}\left(\varphi _i\varphi _k\right).$$
(9)
Here we made also use of $`y_ky_ib`$, if $`i`$ and $`k`$ are nearest neighbors (n.n.) in y-direction and $`y_ky_i0`$ in the other cases.
Now we average over all $`2L^2`$ bonds between n.n. in the lattice and introduce $`j_0I/az`$, where $`az`$ is the area ”used” by a single junction between two domains. We obtain:
$$j=\frac{1}{2L^2}\underset{i,k}{}j_{i,k}=j_0\frac{1}{2L^2}\underset{i,k}{}\mathrm{sin}\left(\varphi _i\varphi _k\right).$$
(10)
The average of the sinus:
$$\overline{\mathrm{sin}\left(\varphi _i\varphi _k\right)}=\frac{1}{2L^2}\underset{i,k}{}\mathrm{sin}\left(\varphi _i\varphi _k\right)$$
(11)
can be obtained from the simulation (index <sub>sim</sub>) with
$$M_{sim}=\frac{1}{2L^2}\underset{i,k}{}\frac{x_{i,sim}+x_{k,sim}}{2}\left(y_{k,sim}y_{i,sim}\right)\mathrm{sin}\left(\varphi _i\varphi _k\right).$$
(12)
We now make following approximations: $`y_{k,sim}y_{i,sim}=1`$, if $`i`$ and $`k`$ are n.n. in y-direction, $`y_{k,sim}y_{i,sim}=0`$, if $`i`$ and $`k`$ are n.n. in x-direction, and $`x_{i,sim}+x_{k,sim}2i`$ neglecting the random placement of the domains in the glass model .
With $`i=0,1,\mathrm{},L1`$ this leads to the magnetization
$`M_{sim}`$ $`=`$ $`{\displaystyle \frac{L1}{2}}\overline{\mathrm{sin}\left(\varphi _i\varphi _k\right)}`$ (13)
With $`M_{sim}=0.5\mathrm{\Delta }M_{sim}`$ from figure 3 it follows
$$\overline{\mathrm{sin}\left(\varphi _i\varphi _k\right)}=\frac{1}{L1}\mathrm{\Delta }M_{sim}.$$
(14)
Here also as in the Bean model $`jj_c`$ leads to the critical state and the critical current density is given by:
$$j_c=j_0\frac{1}{L1}\mathrm{\Delta }M_{sim}$$
(15)
with $`j_0=2\pi J/(az\mathrm{\Phi }_0)`$.
Now we calculate the critical current density in a second way. Applying Bean’s formula in the anisotropic case is justified as we have periodic boundary conditions in y-direction leading to a very long sample in y-direction. Concerning the internal magnetic fields in the critical state we have the situation illustrated in figure 4. Therefore we have in the simulation in x-direction a sample size of double length. Thus the experimentally found magnetization $`M_{exp}`$ is the magnetization corresponding to the triangle of figure 4.
With the Bean formula of the anisotropic case :
$$M_{exp}=\frac{j_cl}{20}$$
(16)
and therefore with the length of the sample $`l=2(L1)a`$ (figure 4) we have
$$j_c=\frac{20M_{exp}}{2(L1)a}.$$
(17)
Now we calculate the value of the ”experimental” magnetization $`M_{exp}`$ from $`M_{sim}`$:
$$M_{exp}=\frac{1}{V}\frac{2\pi }{\mathrm{\Phi }_0}J\underset{i,k}{}\frac{x_i+x_k}{2}\left(y_ky_i\right)\mathrm{sin}\left(\varphi _i\varphi _k\right)$$
(18)
with $`x_i=ax_{i,sim}`$ ($`x_{i,sim}=0,1,L1`$) and $`y_i=by_{i,sim}`$ ($`y_{i,sim}=0,1,L1`$) we have
$$M_{exp}=\frac{1}{L^2abz}\frac{2\pi }{\mathrm{\Phi }_0}Jab\underset{i,k}{}\frac{x_{i,sim}+x_{j,sim}}{2}\left(y_{j,sim}y_{i,sim}\right)\mathrm{sin}\left(\varphi _i\varphi _k\right).$$
(19)
Therefore with $`I`$ from equation (8) and $`M_{sim}`$ from equation (12) and with $`\mathrm{\Delta }M_{sim}=2M_{sim}`$ (figure 3) equation (19) leads to
$$M_{exp}=\frac{1}{z}I\mathrm{\Delta }M_{sim},$$
(20)
which leads (with equation (17)) to the critical current density
$$j_c=\frac{I}{az}\frac{10}{L1}\mathrm{\Delta }M_{sim}.$$
(21)
Using the abbreviation $`j_0=I/(az)`$ and measuring $`j_c`$ in A/cm<sup>2</sup> we have finally
$$j_c=j_0\frac{1}{L1}\mathrm{\Delta }M_{sim}$$
(22)
as in our first calculation (equation (15)) of the critical current density using Bean’s assumption (equation (5)) in the high-$`T_c`$ glass model directly.
Using the numerical values $`\mathrm{\Delta }M_{sim}6`$ and $`L1=15`$ from figure 3 and the experimental values $`a=100\mathrm{\AA }=10^8\mathrm{m}`$ and $`z=10\mathrm{\AA }=10^9\mathrm{m}`$ for the stripes in the HTSC we get $`I4.210^6\mathrm{A}`$. And therefore we have for the critical current density
$`j_c`$ $`=`$ $`{\displaystyle \frac{I}{az}}\mathrm{\Delta }M_{sim}1.710^7{\displaystyle \frac{\mathrm{A}}{\mathrm{cm}^2}},`$ (23)
which is surprisingly close to the experimental values $`j_c510^7\mathrm{A}/\mathrm{cm}^2`$ at 4 K and zero field for the best films and higher than $`j_c`$ in wires ($`j_c<10^6`$ Acm<sup>-2</sup>, e. g. ). We want to note, that in figure 3 the $`\mathrm{\Delta }M_{sim}`$ was measured at $`T=0.2J`$ which corresponds to $`20`$ K for $`T_c=100`$ K.
To calculate critical current densities with equation (15) resp. (22) we made use of Bean’s assumption of a constant increase of the internal magnetic fields $`H_{i,k}`$. It is clear, that this is a rather crude approximation for the high-$`T_c`$ glass model. Furthermore for the calculation of $`j_c`$ we used the approximation of the averaged sinus (equation (11)). This gives perhaps a wrong estimate of the true critical current density of the high-$`T_c`$ glass model.
To support our line of reasoning we turn back to the glass model and estimate $`j_c`$ in the glass model directly following Ebner and Stroud . From the knowledge of the size of the domains it is possible to obtain the critical current $`J_c`$ of the glass model, which can be determined directly
$$J_c=\frac{2e}{h}J$$
(24)
with $`J=k_BT_c`$ and $`T_c=100`$ K. ($`k_B`$ is the Boltzmann constant, $`e`$ the elementary electron charge and $`h`$ the Planck constant.) Equation (24) gives an upper bound for the maximum current through a single Josephson junction in the high-$`T_c`$ glass model , which is used to estimate the coupling energy $`J`$ in the high-$`T_c`$ glass model . To determine the critical current density $`j_c`$ in a-direction we consider in a single domain (in figure 1) the area $`A`$ in b-z-direction, through which the current flows, figure 2. With $`Abz10^{18}`$ m<sup>2</sup> we obtain the critical current density
$$j_c=\frac{J_c}{A}=\frac{2e}{h}\frac{k_BT_c}{bz}=0.66710^8\frac{\mathrm{A}}{\mathrm{cm}^2}.$$
(25)
This is indeed an upper bound to $`j_c`$ in equation (23). Thus both calculations (equation (15) and (22)) of $`j_c`$ give the same formula for $`j_c`$, which is lower than the upper bound for the high-$`T_c`$ glass model in equation (25).
## III (An)isotropy of the critical current
Next we consider the critical currents in the b-direction instead of the a-direction. First we determine the magnetization (with $`I=(2\pi /\mathrm{\Phi }_0)J`$)
$$M=\frac{I}{z}\mathrm{\Delta }M_{sim}260\frac{\mathrm{A}}{\mathrm{cm}},$$
(26)
which is in good agreement with the experiments , too. Note, that $`M`$ is independent of the size of the domains in the a-b-plane. Repeating the above calculations we have for the critical current density in b-direction:
$$j_c=\stackrel{~}{j}_0\frac{1}{L1}\mathrm{\Delta }M_{sim}.$$
(27)
This is analogous to equation (15) with a different $`\stackrel{~}{j}_0=I/(bz)`$, in which $`a`$ is replaced by $`b`$. Inserting the experimental values $`ba/10=10`$ Å and $`z10`$ Å we obtain
$$j_c^b10j_c^a=1.710^8\frac{\mathrm{A}}{\mathrm{cm}^2}.$$
(28)
In b-direction the upper bound analogous to equation (25) also exists. Of course this upper bound is now also about ten times larger than in a-direction.
Thus the consideration of stripes in the high-$`T_c`$ glass model leads to a strong anisotropy of about the factor ten for the critical current densities between a- and b-direction depending on the ratio $`a/b`$ of the striped domains. But this relatively large factor was never reported in the experimental literature.
We want to note, that on the one hand the size of the stripes enters the calculation of $`j_c`$ reciprocally, but on the other hand the number of domains in one spatial direction influences $`j_c`$, too. Thus a better knowledge of the size of the domains is desirable. Additionally a finite size scaling (or the simulation of different system sizes $`L`$) for the magnetization $`\mathrm{\Delta }M_{sim}`$ of the high-$`T_c`$ glass model is necessary to calculate the critical current densities more accurately.
Now we have two features (anisotropy of $`j_c`$ in a- and b-direction and a $`j_c`$, which is in the order of magnitude of the largest experimentally found $`j_c`$ in HTSC) of the high-$`T_c`$ glass model, which do not agree with the experiment. In our opinion there are two possible mechanisms, which can lead to a small or vanishing anisotropy. The first one is an anisotropic coupling constant $`J`$ ($`J_aJ_b`$) in the high-$`T_c`$ glass model, which may be different in a- and b-direction. But this probably only reduces the anisotropy. And it is unlikely, that these anisotropic coupling constants will lead to an isotropic critical current density $`j_c`$.
We postulate therefore the existence of Weiss-type domains in the planes with the stripes, which are dominantly either pointing in a- or b-direction, figure 5. The existence of the Weiss-domains on the other hand explains the relatively lower critical current densities $`j_c`$ found in experiments for single crystals and thin films and in particular their differences.
The resulting weak links between the Weiss-domains (figure 5), which have to be assumed to be (much) ”weaker”, than the weak links of the high-$`T_c`$ glass model, restrict $`j_c`$ to lower values. Thus equation (23) and (28) are only upper bounds of $`j_c`$, too.
This possibility was first brought to our attention in private discussions with Tsuei and Doderer. In the light of the above estimated factor of $`a/b10`$ this picture is in our view one possible explanation for the experimental situation. Crystals or thin films with a predominant direction of the stripes are therefore predicted to show this anisotropy. But the main problem are the weak links between the Weiss-domains. These obviously govern the final experimental measurement of the critical current density. This picture could also explain the differences from sample to sample. While $`T_c`$ is for all samples (nearly) identical, $`j_c`$ is quite different. But in our theoretical calculations for the high-$`T_c`$ glass model both $`T_c`$ and $`j_c`$ only depend on the coupling $`J`$.
Considering this picture it is clear, that improvements of $`j_c`$ rise in general the quality of the samples. Especially in the light of the possible fabrication of electronic devices, better high-$`T_c`$ samples have to be obtained by removing the Weiss-domains and the weak links between these domains. Therefore it would be extremely fruitful to avoid the Weiss-domains or to restrict the influence of the resulting weak links. In this spirit we propose a receipt following from the fabrication of magnets. The magnets are cooled down in a strong magnetic field. For high-$`T_c`$-materials we propose a similar procedure. The cooling process should use the cooling schedule known from simulations. This schedule has been subject to extensive research in the field of (physical) optimization in particular the ”simulated annealing” method .
Procedures developed in the optimization theory may be directly transfered to the annealing of the high-$`T_c`$-materials. Furthermore the annealing has to be carried out in an electric field and/or while an electric current is flowing through the sample.
We expect from these procedures a substantial increase of the quality of the high-$`T_c`$-materials. In particular the fabrication of electronic devises as e.g. transistors should benefit greatly from these ideas. In the fabrication of wires or polycrystals additionally the weak links between the superconducting grains limit the critical current densities and should be removed, too.
## IV Summary and Conclusions
In the high-$`T_c`$ glass model the striped superconducting domains of the high-$`T_c`$ materials are the domains of constant superconducting phases . The size of the striped domains may be deduced by mainly theoretical considerations and magnetic measurements . These measurements and the corresponding theoretical framework were already known shortly after the discovery of the HTSC . Already there the size of the domains could be estimated correctly to be approximately $`10^4`$ Å<sup>2</sup> . This leads to the conclusion, that the glass behavior is not related to the early ceramic structure of the grains, that many but not all samples exhibit.
From the high-$`T_c`$ glass model the critical current density can be calculated following two different paths: first using the extended Bean model and hysteresis measurements of the high-$`T_c`$ glass model and second directly from the definition of the glass model using Bean’s assumption of the critical state model. Both approaches lead consistently to the same critical current densities $`j_c`$ and to a strong anisotropic critical current density for the a- resp. b-direction of $`j_c`$, which is much larger than the anisotropy found experimentally. But they are close to the highest measured critical current densities in thin films. This strong anisotropy follows from the underlying striped shape of the domains.
Taking the same type of approximation (following the Bean model) we obtain the same value of $`j_c`$ in the simulation as in the experiments. This is in our opinion a strong evidence in favor of the high-$`T_c`$ glass model.
Nonetheless it is puzzling, why this anisotropy in the a-b-plane was never reported for $`j_c`$. We propose two extensions of the high-$`T_c`$ glass model: first an anisotropic coupling $`J_aJ_b`$, which in our opinion can only reduce the anisotropy, and second the existence of Weiss-type domains, in which the predominant direction of the stripes is turned by 90. The latter leads to nearly isotropic critical current densities and lowers $`j_c`$ on account of the ”new” weak links between these areas with the same direction of the stripes. This explains, too, why $`T_c`$ is nearly constant between different samples whereas $`j_c`$ is quite different. Removing these Weiss-type domains should lead to higher critical current densities $`j_c`$, which may be done by simulated annealing of the HTSC (eventual in electric fields).
## V Acknowledgment
We want to thank C.C. Tsuei and T. Doderer for very helpful discussions. We would like to acknowledge illuminating discussions with K.A. Müller, H. Keller, T. Schneider, E. Stoll, and J.M. Singer. Especially we would like to thank P.C. Pattnaik and D.M. Newns for inspiring discussions and ideas. Finally we acknowledge the financial support of the UniOpt GmbH, Regensburg.
|
no-problem/9908/cond-mat9908145.html
|
ar5iv
|
text
|
# On the scaling of the (𝐻-𝑇)-phase diagram of CuGeO3.
## Abstract
The $`HT`$ phase diagram of CuGeO<sub>3</sub> has been determined, for different values of the hydrostatic pressure, utilizing optical absorption spectroscopy on the Cu<sup>2+</sup> $`dd`$ transitions. It is shown that the intensity of the related zero phonon line transition is very sensitive to the local environment of Cu<sup>2+</sup>, allowing for precise determination of all phase transitions. It is found that the phase diagrams at various pressures do not scale according to the Cross-Fischer theory. An alternative scaling is proposed.
Since the discovery of the spin-Peierls (SP) transition in the inorganic compound CuGeO<sub>3</sub> there has been an increasing experimental and theoretical interest for the physics related to low dimensional magnetism . One of the most characteristic properties of a spin-Peierls material is its magnetic field ($`H`$)-temperature ($`T`$) phase diagram (PD). For the quasi-one-dimensional Cu<sup>2+</sup> spin system CuGeO<sub>3</sub>, this PD has been explored using a variety of techniques . The investigation using thermal expansion/magnetostriction measurements has shown that CuGeO<sub>3</sub>, which crystallizes into an orthorhombic structure ($`a`$=0.481 nm, $`b`$=0.843 nm and $`c`$=0.295 nm), exhibits more drastic changes along the $`a`$ and $`b`$ axis than along the $`c`$-axis (which bears the Cu<sup>2+</sup> ions) at all lines describing the PD. This underlines the importance of the elastic energies involved in these transitions.
From a theoretical point of view the corresponding PD is believed to be universal defining in the $`HT`$ plane a high temperature uniform phase (U) and a dimerized (D) phase below $`T_{\mathrm{sp}}`$ and magnetic fields below a critical value $`H_\mathrm{c}`$ ($`T_{\mathrm{sp}}14`$ K and $`H_\mathrm{c}13`$ T for CuGeO<sub>3</sub>). Above $`H_\mathrm{c}`$ and at low temperatures the system enters an incommensurate solitonic magnetic phase (I) which evolves into a sinusoidally modulated phase for higher fields . The best established theory for spin-Peierls materials is the one due to Cross and Fisher (CF) . Apart from the universality of the PD, one of the main theoretical results is that, at low $`H`$ values the transition temperature scales with the square of the field. For CuGeO<sub>3</sub>, this scaling requires a further reduction of $`H`$ by 10% . The universal character could be eventually the consequence of approximations made in the model and since to our knowledge there is no complete experimental study of this aspect, the question deserves a careful inspection.
In order to achieve this goal one has to vary $`T_{\mathrm{sp}}`$. When accomplished by chemical substitutions , the question of the relative quality and purity of samples becomes acute. A more appropriate way is to apply, on a given sample, hydrostatic pressure which increases $`T_{\mathrm{sp}}`$ in CuGeO<sub>3</sub> . It has been argued that this increase results from an increasing frustration in the spin system under pressure . This, in turn, gives an additional interest for the study of the PD under pressure. Technically speaking the standard techniques to measure the PD are not easy to settle in a high pressure cell. We have therefore decided to adopt an optical technique which has been shown recently to demonstrate around $`\mathrm{}\omega =1.47`$ eV specific anomalies at the different phase transitions .
After a brief description of the technical aspects, we will first refine the analysis of the optical data trying to explain what is indeed measured and how it can be interpreted. The second part of this Letter presents and discusses the results obtained with this technique under high pressure.
The crystals of CuGeO<sub>3</sub> used in this study have been grown by a traveling floating zone technique and cleaved in such a way the surface of the sample contains the $`c`$ and $`b`$-axis. The absorption measurements in the visible region around 1.47 eV have been done on samples with a thickness $`d`$=2 mm using a standard optical set-up. Magnetic fields $`H`$ up to 23 T were provided by a modified resistive Bitter magnet and applied parallel to the $`a`$-axis of the sample. The high pressure measurements were made in a small piston high pressure cell filled with a mixture of ethanol-methanol, equipped with two sapphire windows to perform transmission measurements. The temperature was recorded with a cernox thermometer for which the magneto-resistance has been calibrated at different temperatures.
The optical technique consists in investigating the optical properties of CuGeO<sub>3</sub> in the low energy tail of the intra-center absorption of the Cu<sup>2+</sup>-ions. This absorption band, centered at $`\overline{E}1.63`$ eV, has been analyzed carefully . It results in a broad band with a maximum absorption coefficient $`\alpha (\overline{E})`$ ranging from 500 to 800 cm<sup>-1</sup> (depending on the polarization), and shows that in the range of 1.47 eV the remaining absorption coefficient is expected to be less than 1 cm<sup>-1</sup>. In these conditions, the appropriate way to investigate optical properties is not by measuring reflection, which becomes a complicated function of reflectivity and transmission responses, but by direct transmission measurements. If one performs such experiments, on wedged samples to avoid interference effects, the transmission $`𝒯`$ is given by the standard expression
$$𝒯(\mathrm{}\omega )=\frac{\left(1\right)^2\mathrm{exp}\left(\alpha (\mathrm{}\omega )d\right)}{1+^2\mathrm{exp}\left(2\alpha (\mathrm{}\omega )d\right)}$$
(1)
where $``$ is the reflectivity coefficient of one interface. For low absorbing media $``$ reduces to $`(n1)^2/(n+1)^2`$, where $`n`$ is the refractive index ($`2.5`$ for CuGeO<sub>3</sub> ), implying $`0.13`$. In such conditions, when measuring the ratio of the transmission coefficient with respect to that of a reference point ($`𝒯_0`$), the denominator of Eq. 1 can be approximated to 1. Furthermore, since $``$ is not expected to vary significantly neither with $`T`$ nor $`H`$ one arrives at $`𝒯/𝒯_0=\mathrm{exp}\left((\alpha \alpha _0)d\right)=\mathrm{exp}(\mathrm{\Delta }\alpha d)`$, where all optical functions are dependent on the photon energy $`\mathrm{}\omega `$. The absorption strength is strongly polarized , but all discontinuities related to the phase transitions are not dependent on this polarization . Since it is not easy to perform polarized measurements in a high pressure cell, we report here on non-polarized measurements only. The relevant physical quantity is indeed $`\mathrm{\Delta }\alpha d=\mathrm{ln}(𝒯/𝒯_0)`$. A typical variation of this quantity at fixed temperature and different values of $`H`$ is presented in Fig. 1b. One sees a clear discontinuity of the intensity, corresponding to the first-order transition at the critical field $`H_\mathrm{c}`$ as already reported . We shall see below that this value corresponds quite well to $`H_\mathrm{c}`$ values reported by other techniques.
Before detailing the main results one has to discuss the origin of the optical transition we are studying. In the picture of non-interacting ions the absorption occurs between vibronic transitions involving as the initial state an electron in the $`d`$ fundamental state and $`m=0`$ vibrational state (at low $`T`$) and as the final state the electron in the $`d^{}`$ excited state of the ion and $`m=l`$ phonons in a displaced configuration coordinate $`Q_r`$ with respect to the initial value $`Q_0`$. The system relaxes in the excited state, after absorption, towards a lower energy by emitting these $`l`$ phonons . The strength of the optical matrix element is essentially governed by the overlap of the $`m=0`$ phonon wave function and the displaced $`l`$ phonon wave function. For simple cases where only one type of phonons is involved one gets the picture sketched in Fig. 1a. Therefore one observes around 1.47 eV the so-called zero-phonon-line (ZPL) for which $`l=0`$ and the energy is $`E_{ZPL}`$. Of course due to the coupling of the different Cu<sup>2+</sup> ions, one expects the corresponding transition be broadened and this is indeed observed since the width $`\mathrm{\Gamma }`$ of the spectra is of the order of 10 meV and does not vary much with $`T`$ and $`H`$. Because this width remains much smaller than that of the main absorption band ($`0.5`$ eV) one may still apply, in a first approximation to our case, the results known for isolated centers. In this context the probability to absorb a photon in the $`l`$th vibronic state, neglecting the constant electronic part of the matrix element, is given by: $`P(l)=\mathrm{exp}(S)S^l/l!`$. $`S`$ is the number of phonons emitted after the absorption at $`\overline{E}`$ (Fig. 1a). The phonon involved in this process has been determined in Ref. 18 and corresponds very likely to the IR active phonon at $`\mathrm{}\omega _p=28.5`$ meV . With that analysis one gets $`S`$ values of about 6 which correspond to a very strong electron-phonon interaction. This gives in turn a mean absorption coefficient of the ZPL of about 0.25 cm<sup>-1</sup> which justifies the transmission technique used here. In non-polarized experiments the situation is not simple because, if there exists a nice well-defined line for the electric field of the light $`\stackrel{}{E}c`$-axis, the situation for $`\stackrel{}{E}b`$-axis is more complex and this explains the high energy shoulder of the spectra in Fig. 1b. Therefore $`S`$ has to be understood as an averaged value of the relaxation energy of the Cu<sup>2+</sup>ions in the excited state over the different relaxation routes. The total integrated quantity, $`M_0(H,T)=d\mathrm{\Delta }\alpha 𝑑E`$, has been evaluated over a fixed range of energies between 1.45 eV and 1.50 eV. If one knows $`M_0`$ at a reference point, which will be taken for each pressure at $`T=2`$ K and $`H=0`$ T, it is easy to show that the variation of $`M_0`$ mimics quite well that of $`\mathrm{\Delta }S=S(0,2K)S(H,T)`$ i.e. the variation of the relaxation energy with $`H`$ and/or $`T`$.
Fig. 2 presents characteristic variations of $`M_0(H,T)`$ for pressures $`P=1`$ bar, 0.32 GPa, and 0.74 GPa. The field dependence at fixed temperature ($`T=2`$ K , Fig. 2a) shows an abrupt discontinuity of $`M_0`$ at a critical magnetic field $`H_\mathrm{c}(P)`$. This discontinuity jump decreases upon increasing $`T`$ and disappears beyond some temperature $`T^{}(P)<T_{\mathrm{sp}}(P,H=0)`$. The existence of this discontinuity is related to the first-order phase transition from the dimerized (D) phase to the incommensurate (I) phase. Note that the discontinuity also decreases at fixed temperature when increasing $`P`$. The temperature dependence of $`M_0`$ at fixed magnetic field $`H=0`$ T (Fig. 2b) and 22 T (Fig. 2c) corresponding to fields lower and higher than $`H_\mathrm{c}`$, respectively for all values of $`P`$) shows a clear discontinuity in the slope $`dM_0/dT`$ at a critical value $`T_\mathrm{c}(H,P)`$. This discontinuity corresponds to the second-order phase transition from the D to the U phases for $`H<H_\mathrm{c}`$ and from the I to the U phases for $`H>H_\mathrm{c}`$. It is interesting to compare all the variations of $`M_0`$ observed with those of the lattice parameters of the structure. It has been shown for instance that, at $`H=0`$ T, $`db/b`$ increases by about $`8\times 10^5`$ between 2 K and 14 K whereas $`da/a`$ decreases by $`5\times 10^5`$. These variations constrain the environment of the Cu<sup>2+</sup> ions which in turn is expected to reduce the relaxation energy of these ions in the excited state. This is indeed observed since in that range of temperature $`S`$ is found to increase by a factor of about 2 corresponding to a decrease of the relaxation energy of about 57 meV.
Now that the nature of the 1.47 eV absorption feature in CuGeO<sub>3</sub> has been established, we turn to the main results reported in this Letter. Using a CF type of scaling (i.e. $`T/T_{\mathrm{sp}}`$ vs. $`g\mu _bH/T_{\mathrm{sp}}`$ where $`g=2.15`$ for $`Ha`$-axis ), the scaled PD of CuGeO<sub>3</sub> is shown in Fig. 3a for various pressures together with already published data (small squares) . Clearly one reproduces these earlier results quite well at $`P=1`$ bar. The transition temperatures and fields used in the scaling are listed in Table I. At low magnetic fields, the D–U transition temperature indeed scales with the square of the field . At higher fields, and in particular for the D-I transition, the scaling no longer holds. Already present at $`P=1`$ bar, the problem becomes more acute at higher pressures (Fig. 3a). In order to reconcile the theoretical result with the data, one now has to assume a $`T_{\mathrm{sp}}`$-dependent field scaling factor, which is not very useful if one tries to unify the PD. It seems therefore more appropriate to look for a different kind of scaling law. It is here proposed that $`H`$ does not scale with $`T_{\mathrm{sp}}`$, but rather with the singlet-triplet gap $`\mathrm{\Delta }_{\mathrm{ST}}`$ at zero field. The pressure dependence of $`\mathrm{\Delta }_{\mathrm{ST}}`$ has been determined by Nishi et al. : $`\mathrm{\Delta }_{\mathrm{ST}}(P)=\mathrm{\Delta }_{\mathrm{ST}}(0)+10.5P`$ ($`P`$ in GPa, $`\mathrm{\Delta }_{\mathrm{ST}}`$ in cm<sup>-1</sup>, $`\mathrm{\Delta }_{\mathrm{ST}}(0)=16.8`$ cm<sup>-1</sup>). The resulting scaled PD is shown in Fig. 3b. Clearly the PDs at various pressures now collapse onto each other, despite the observation that the D–I transition occurs before the full closure of the magnetic gap.
In addition to the scaling properties of D–I transition, the scaling of the D-U transition has also improved. The quadratic field dependence found for the D-U transition is:
$$\frac{T_{\mathrm{sp}}(H)}{T_{\mathrm{sp}}(0)}=\left(10.384\left[\frac{g\mu _bH}{\mathrm{\Delta }_{\mathrm{ST}}}\right]^2\right)$$
(2)
This scaling is not compatible with the CF type of scaling and appears more compatible with the mean field BCS type theory. However, CuGeO<sub>3</sub> does not obey this theory since $`\mathrm{\Delta }_{ST}/T_{sp}`$ is not constant as can be inferred from Table 1. As a consequence, one may conclude that the CF theory is not very appropriate to describe the properties of CuGeO<sub>3</sub>. There are various factors which could explain that, in particular, the presence of a next-nearest neighbor frustrating interaction or the non-adiabacity (the exchange energy is comparable to the phonon energies). It is interesting to see whether the present scaling also applies to organic spin-Peierls compounds such as (TTF)Cu(BDT) , (TTF)Au(BDT) and MEM-(TCNQ)<sub>2</sub> for which, to our knowledge, $`\mathrm{\Delta }_{ST}`$ has not been directly measured. However, if we speculate for them a value of $`\mathrm{\Delta }_{ST}/T_{sp}1.76`$ (mean field theory), the proposed scaling is surprisingly good for the D–U and even the D–I phase transitions. In particular $`g\mu _BH_c/\mathrm{\Delta }_{ST}`$ ranges between 0.77 and 0.79 for all organic compounds. The surprise comes from the fact that all compounds listed in Table 1 have very distinct properties concerning the existence of frustration, soft phonon modes or variation of $`T_{sp}`$ with pressure. This result implies that $`H_c`$ does not scale with $`\mathrm{\Delta }_{ST}`$ as already pointed by Ohta et al. .
In conclusion the study of the ZPL absorption of the intra-center Cu<sup>2+</sup> transition has allowed to determine the $`HT`$ phase diagram of CuGeO<sub>3</sub> for different pressures. It has been shown that a Cross-Fischer type of scaling of these phase diagrams is not appropriate. It is noted that a scaling which does not assume a constant ratio between $`\mathrm{\Delta }_{ST}`$ and $`T_{sp}`$ could be more effective, although it relies on two independent parameters $`\mathrm{\Delta }_{ST}`$ and $`T_{sp}`$. We have at present no strong arguments which could support these findings but we hope that this observation will motivate further theoretical research on this special class of compounds.
###### Acknowledgements.
The Grenoble High Magnetic Field Laboratory is “Laboratoire conventionné à l’UJF et l’INPG de Grenoble”. The Laboratoire de Chimie des Solides is a ”Unité de Recherche Associée CNRS n 446”. J. Z. acknowledges the partial supports from the grant ERBCHGECT 930034 of the European Commission.
|
no-problem/9908/astro-ph9908212.html
|
ar5iv
|
text
|
# ROSAT X-ray sources and exponential field decay in isolated neutron stars
## 1 Introduction
Evolution of neutron stars (NSs) can be called “magneto–rotational”, because all main astrophysical manifestations of these objects are determined by their periods and magnetic fields.
Four main regimes, as described for example in Lipunov (1992), are possible for isolated NSs: ejector, when a star represents a radio pulsar, or a dead pulsar, spinning down due to magneto–dipole radiation; propeller, when surrounding captured matter cannot penetrate through the centrifugal barrier; accretor, when matter can reach the surface, and the NSs appear as an X-ray source; and georotator, when gravitation becomes insignificant, because magnetic pressure dominates everywhere over the gravitational pull, and geo-like magnetosphere is formed.
Magnetic field decay in NSs is an uncertain subject. Many models were suggested (see, for example, Ding et al., 1993; Jahan Miri & Bhattacharya, 1994 ). The only strong observational result is that for the exponential (or nearly exponential) decay characteristic time scale, $`t_d`$, is longer than $`10^7\mathrm{yrs}`$, because no expected effects of decay are observed in radio pulsars (Lyne et al. 1998).
Field decay was used in the case of old accreting isolated NSs by Konenkov & Popov (1997) and Wang (1997) to explain properties of the source RX J0720-3125. And it seems, that if this source really is an accreting old NS and if it was born as a normal radio pulsar (with small period, $`1`$ s, and high magnetic field, $`10^{12}`$ G), its properties can be explained only if decay is working (or if this source was born with such unusual parameters). Recently the influence of the field decay in isolated NSs was investigated in Colpi et al. (1998) and Livio et al. (1998). An attempt to include field decay into population synthesis of isolated NSs was made in Popov et al. (1999).
Here we try to put some limits on the parameters of the exponential field decay, assuming, that old isolated NSs are observed as accreting X-ray sources (Walter, Wolk & Neühauser 1996; Haberl et al. 1996, 1998; Neühauser & Trümper 1999; Schwope et al. 1999), neglecting the possibility, that all of them can be highly magnetized NSs, “magnetars” (see a brief discussion on this assumption in Popov et al., 1999).
## 2 Calculations and results
The main idea of our work is to calculate the ejector time, $`t_E`$, i.e. a time interval spent by a NS on the ejector stage, for different parameters of the field decay and standard assumptions on the initial parameters of a NS, and to compare that time with the Hubble time, $`t_H`$.
For constant field $`t_E`$ monotonically depends upon NS’s velocity and ISM density:
$$t_E(\mu =const)10^9\mu _{30}^1n^{1/2}v_{10}\mathrm{yrs}.$$
(1)
If $`t_E`$ for decaying field for some sets of parameters is greater than $`t_H10^{10}\mathrm{yrs}`$ even for relatively high concentration of interstellar medium (ISM), $`1\mathrm{cm}^3`$, and small spatial velocity of a NS, $`10\mathrm{km}\mathrm{s}^1`$ (this velocity is about the sound velocity in the ISM), than for that parameters of the decay, $`t_d`$ and $`\mu _b`$, no accreting NSs would be observed, and such sets can be excluded as progenitors of old accreting isolated NSs. In addition, if we assume, that most part of NSs evolve with similar $`t_d`$ and $`\mu _b`$ and are born with some typical initial period, $`p_0`$, and initial magnetic momentum, $`\mu _0`$ (that is supported by radio pulsars observations), then such sets of parameters should be totally excluded from models of the magnetic field decay.
As the first approximation only $`t_E`$ can be taken into account, because the time interval, spent on the propeller stage, $`t_P`$, is uncertain, but usually shorter, than $`t_E`$ (see Lipunov & Popov, 1995). These authors argue, that for non-decaying magnetic field $`t_E`$ is always longer than $`t_P`$. For decaying magnetic field the situation can be different, and $`t_P`$ can be comparable to $`t_E`$. So we obtain here “upper limits”, i.e. we calculate very strong limits on parameters $`t_d`$ and $`\mu _b`$, which can become only wider if one takes into account the propeller stage.
We assume exponential field decay:
$$\mu =\mu _0e^{t/t_d},\mu >\mu _b$$
(2)
where $`\mu _0`$ is the initial magnetic momentum ($`\mu =\frac{1}{2}B_pR_{NS}^3`$, here $`B_p`$ – polar magnetic field, and $`R_{NS}`$– NS radius), $`t_d`$ – characteristic time scale of the decay, and $`\mu _b`$ – bottom magnetic momentum, which is reached in:
$$t_{cr}=t_d\mathrm{ln}\left(\frac{\mu _0}{\mu _b}\right).$$
(3)
After that moment magnetic field is assumed to be constant.
In Fig. 1 we show, as an illustration, evolutionary tracks of NSs on $`PB`$-diagram for $`v=10\mathrm{km}\mathrm{s}^1`$ and $`n=1\mathrm{cm}^3`$. Tracks start at $`t=0`$, when $`p=p_0=0.020\mathrm{s}`$ and $`\mu =mu_0=10^{30}\mathrm{G}\mathrm{cm}^3`$, and end at $`t=t_H=10^{10}`$ yrs (for $`t_d=10^{10}\mathrm{yrs}`$, $`t_d=10^9`$ yrs and for constant magnetic field) or at the moment, when $`p=p_E`$ (for $`t_d=10^8`$ yrs and $`t_d=10^7`$ yrs). The line with diamonds shows $`p=p_E(B)`$.
As far as the accretion rate from the ISM is small (even for our parameters), less than $`10^{12}\mathrm{g}\mathrm{s}^1`$, no influence of accretion onto decay was taken into account (see Urpin et al., 1996).
The ejector stage lasts until the critical ejector period, $`p_E`$, is reached:
$$p_E=11.5\mu _{30}^{1/2}n^{1/4}v_{10}^{1/2}\mathrm{s},$$
(4)
where $`v_{10}=\sqrt{v_p^2+v_s^2}/10\mathrm{km}\mathrm{s}^1`$. $`v_p`$– spatial velocity of a NS. Here the sound velocity, $`v_s`$, was taken into account, but as far as normally NSs spatial velocities are higher than $`10\mathrm{km}\mathrm{s}^1`$ and the sound velocity outside hot low density ISM regions is lower than $`10\mathrm{km}\mathrm{s}^1`$, we have $`\sqrt{v_p^2+v_s^2}v_p`$. $`n`$ is a concentration of the ISM.
Initial period should be taken to be much smaller than $`p_E`$. We used $`p_0=0`$ s. Variations of $`p_0`$, if it stays much less than $`p_E`$, have little influence on our results, i.e. in that case $`t_E`$ is determined only by $`p_E`$ and history of the decay. We calculated spin-down according to magneto–dipole formula (but other regimes are possible, see Beskin et al. 1993 for a review):
$$\frac{dp}{dt}=\frac{2}{3}\frac{4\pi ^2\mu ^2}{pIc^3},$$
(5)
where $`\mu `$ can be a function of time.
For our estimates we assumed constant velocity of NSs, $`v`$, equal to $`10\mathrm{km}\mathrm{s}^1`$ and constant ISM concentration, $`n`$, equal to $`1\mathrm{cm}^3`$. These conditions give us a lower limit on $`t_E`$, because normally velocity is significantly higher (fraction of slow velocity NSs is less than few percents, see Popov et al., 1999), and ISM density is smaller than the specified values (partly because high velocity NSs spend most of their lives in low density regions far from the Galactic plane).
After simple algebra one can obtain a formula for $`t_E`$, depending upon $`t_d`$, $`\mu _0`$, $`v`$, $`n`$ and $`\mu _b`$:
$$t_E=\{\begin{array}{cc}& t_d\mathrm{ln}\left[\frac{A}{t_d}\left(\sqrt{1+\frac{t_d^2}{A^2}}1\right)\right],t_E<t_{cr}\hfill \\ & t_{cr}+A\frac{\mu _0}{\mu _b}t_d\frac{1}{2}\left(\frac{\mu _0}{\mu _b}\right)^2\left(1e^{2t_{cr}/t_d}\right),t_E>t_{cr}\hfill \end{array}$$
(6)
where coefficient $`A`$ is determined by the formula:
$$A=\frac{3Ic}{2\mu _0\sqrt{2v\dot{M}}}10^{17}I_{45}\mu _{0_{30}}^1v_{10}^{1/2}\dot{M}_{11}^{1/2}\mathrm{s},$$
(7)
where $`\dot{M}`$ can be formally determined as a combination of intrinsic NS’s parameters, its velocity and ISM concentration using the Bondi formula even if a NS is not on the accretor stage:
$$\dot{M}10^{11}nv_{10}^3\mathrm{g}\mathrm{s}^1.$$
(8)
Results of $`t_E`$ calculations for several values of $`\mu _0`$ and $`t_d`$ are shown in Fig. 2. Horizontal regions in the left parts of all curves appear because there a NS reaches the critical period $`p=p_E`$ before it reaches bottom magnetic momentum at $`t=t_{cr}`$, and $`t_E`$ is not depended upon the rest field decay history. On the right side from the maximum on the curves a NS reaches $`\mu =\mu _b`$ with period significantly less than $`p_E(\mu _b)`$, and it takes a long time after $`t=t_{cr}`$ to reach $`p=p_E`$ (this time is longer for lower $`\mu _b`$). On the left side from the maximum, but before the horizontal region, a NS reaches $`\mu =\mu _b`$ with a period close to $`p_E`$, and it is closer if $`\mu _b`$ is lower, so very quickly after $`t=t_{cr}`$ a NS can spin down to $`p_E`$. The first point at the right corresponds to the bottom momentum equal to initial momentum, i.e. to the case without decay.
One can see from that figure, that for some combination of parameters $`t_E`$ is longer than the Hubble time. It means, that NSs never evolve further than the ejector stage.
We argue, that as far as accreting isolated NSs are observed, combinations of $`t_d`$ and $`\mu _b`$ for which no accreting isolated NS appear can be excluded. We plotted it in Figs. 3 and 4.
Filled regions represent space of the parameters where $`t_E`$ is longer than $`10^{10}\mathrm{yrs}`$, so in that region a NS never come to the accretor stage, and doesn’t appear as accreting X-ray source. With the fact of observations of accreting old isolated NSs by ROSAT this region can be called “forbidden” for selected parameters of the exponential field decay (and for specified $`\mu _0`$).
In the “forbidden” region in Fig. 3, which is plotted for $`\mu _0=10^{30}\mathrm{G}\mathrm{cm}^3`$, all NSs reach the bottom field in a Hubble time or faster, and evolution on the late stages of their lives goes on with the field equal to the bottom. The left side is determined approximately by the condition:
$$p_E(\mu _b)=p(t=t_{cr}).$$
(9)
Small difference between the line, correspondent to the condition above, and the left side of the “forbidden” region appears because a NS can slightly change its period even with the momentum $`\mu =\mu _b`$, but due to small value of the field angular momentum losses are also small.
The right side of the region is roughly determined by the value of $`\mu _b`$, with which a NS can reach the ejector stage with any $`t_d`$, i.e. this $`\mu _b`$ corresponds to the minimum value of $`\mu _0`$ with which a NS reach the ejector stage without field decay.
NSs to the right from the “forbidden” region leave the ejector stage, because their field cannot decay down to low values, and spin-down is fast enough during all their lives as ejectors, because the bottom magnetic momentum there is relatively high. To the left from the “forbidden” region the situation is different. Spin-down of NSs is very small and they leave the ejector stage not because of spin-down, but due to decreasing of $`p_E`$, which depends upon the magnetic momentum.
Dashed line in Fig. 3 shows, that for all interesting parameters a NS with $`\mu _0=10^{30}\mathrm{G}\mathrm{cm}^3`$ reach $`\mu _b`$ in less than $`10^{10}`$ yrs. Dot-dashed line shows the same for $`\mu _0=0.510^{30}\mathrm{G}\mathrm{cm}^3`$.
On Fig. 3 we also show the “forbidden” region and the line of reaching $`\mu _b`$ for $`\mu _0=0.510^{30}\mathrm{G}\mathrm{cm}^3`$.
Fig. 4 is plotted for $`\mu _0=10^{29}\mathrm{G}\mathrm{cm}^3`$. For long $`t_d`$, $`>410^9`$ yrs, a NS again is not able to leave the ejector stage. It happens because the magnetic momentum can’t decrease down to small value of $`\mu `$ (nearly $`\mu _b`$), and $`p_E`$ is not decreasing enough.
## 3 Discussion and conclusions
We tried to evaluate the region of parameters, forbidden for models of the exponential magnetic field decay in NSs using the fact of observations of old accreting isolated NSs in X-rays.
If the main fraction of NSs have nearly the same initial parameters and parameters of the decay, then the intermediate values of $`t_d`$ ($`10^710^8\mathrm{yrs}`$) in combination with the intermediate values of $`\mu _b`$ ($`10^{28}10^{29.5}\mathrm{G}\mathrm{cm}^3`$) for $`\mu _0=10^{30}\mathrm{G}\mathrm{cm}^3`$ can be excluded, because for that sets of parameters NSs spend all their lives on the ejector stage, never coming to the accretor stage.
As one can see in Fig. 2, for higher $`\mu _0`$ NSs should reach $`t_E`$ even for $`t_d<10^8`$ yrs, for smaller – the “forbidden” region should become wider. Results are depended on the initial magnetic field, $`\mu _0`$, ISM concentration, $`n`$, and NS velocity, $`v`$, so one can say, that the observed accreting isolated NSs come from, for example, the objects with high initial magnetic field, and the rest are not visible because they are in the forbidden region. To explore this idea in details population synthesis of NSs for realistic distributions of $`v`$, $`\mu _0`$ and $`n`$ is needed. But we can say immediately, that the idea of obtaining accreting old isolated NSs from initially high field objects is not very promising, because the fraction of high field NSs can not be large (basing on radio pulsars observations), and as far as the fraction of low velocity NSs is not more than several percents (Popov et al. 1999) and the volume fraction filled with relatively high density ISM is also small, accreting old isolated NSs should come from “typical” population, i.e. from NSs with $`\mu _0`$ about $`10^{30}\mathrm{G}\mathrm{cm}^3`$ or less.
Actually, limits that we obtained are even stronger than they are in nature, because we didn’t take into account, that some significant time (in the case of field decay) a NS can spend on the propeller stage (spin-down rate at this stage is very uncertain, see the list of formulae, for example, in Lipunov & Popov 1995 or Lipunov 1992). Calculations of this effect, and calculations for different models of non-exponential field decay are subjects for future work.
We cannot say anything about parameters of field decay in the case of accretion in close binaries, because there situation is completely different, and our results cannot be applied to millisecond radio pulsars or other objects in close binary systems.
We note, that there is another reason due to which very fast decay down to small values of $`\mu _b`$ also can be excluded, because it leads to huge amount of accreting isolated NSs. This situation is similar to “turning-off” the magnetic field of a NS (i.e., quenching any magnetospheric effect on the accreting matter), and for any velocity and density distributions we should expect significantly more accreting isolated NS than we have from ROSAT observations (of course, for high velocities X-ray sources will be very dim, but close NSs can be observed even for velocities $`100`$ km s<sup>-1</sup>).
So, the existence of several old isolated accreting NSs, observed by ROSAT (if it is the correct interpretation of observations), can put important limits on the models of the magnetic field decay for isolated (without influence of accretion, which can stimulate field decay) NSs, and models, from their side, should explain the fact of observations of $`10`$ accreting isolated NSs in the solar vicinity. We cannot discuss numerous details of connection between decay parameters and X-ray observations of isolated NSs without detailed calculations, we just tried to show, that this connection should be taken into account and made some illustrations of it, and future investigations in that field are wanted.
###### Acknowledgements.
We thank Monica Colpi, Denis Konenkov and Roberto Turolla for numerous comments on the text, suggestions and discussions. We also want to thank Vladimir Lipunov and Aldo Treves for advices and attention to our work. This work was supported by the RFBR (98-02-16801) and the INTAS (96-0315) grants.
|
no-problem/9908/cond-mat9908476.html
|
ar5iv
|
text
|
# Propylene Carbonate Reexamined: Mode-Coupling 𝜷 Scaling without Factorisation ?
## I Introduction
### A Motivation
The glass transition, an essentially dynamic phenomenon, can be described as the slowing down and eventual freezing of $`\alpha `$ relaxation. According to mode-coupling (MC) theory , the physical origin of this process must be sought on a picosecond scale where long-ranged transport starts to evolve from vibrational short-time dynamics.
In the long-time limit, MC theory reproduces the well-established phenomenology of $`\alpha `$ relaxation. New results are obtained for shorter times. In particular, the theory predicts a change of transport mechanism around a cross-over temperature $`T_\mathrm{c}`$, located in the moderately viscous liquid phase well above the conventional glass transition temperature $`T_g`$. On cooling towards $`T_\mathrm{c}`$, particles spend more and more time being trapped in transient cages; this process, labelled fast $`\beta `$ relaxation, is predicted to obey remarkably universal scaling.
In a couple of structural glass formers, MC predictions have been confirmed primarily by different scattering techniques . More recently, the GHz–THz dynamics also became accessible by dielectric spectroscopy . Although results are in accord with the MC scenario, in several cases the data did not fall together with dynamic susceptibilities from scattering experiments which is in conflict with the asymptotic factorisation property of MC $`\beta `$ relaxation. Recent theoretical developments suggest explanations on the basis of corrections to scaling and orientational degrees of freedom . For experimental tests, more detailed comparisons of different observables are needed.
In this context, we performed incoherent neutron and depolarised light scattering experiments on propylene carbonate (PC, 4-methyl-1,3-dioxolan-2-on, C<sub>4</sub>O<sub>2</sub>H<sub>6</sub>), a fragile glass former ($`T_g=160`$K) with low molecular weight ($`M=102.1`$) which has already been studied by various experimental techniques. A synopsis of available data shall be given in subsection I.C after collecting the essential results of MC theory in I.B.
### B The Mode-Coupling Cross-Over
In its simplest (“idealised”) formulation, MC theory describes an ergodic-to-nonergodic transition at $`T_\mathrm{c}`$. On the low-temperature side, the onset of fast $`\beta `$ relaxation leads to an anomalous decrease of the Debye-Waller or Lamb-Mößbauer factor,
$$f_q=f_q^c+h_q|\sigma |^{1/2},\sigma >0,$$
(1)
with a reduced temperature $`\sigma =(T_\mathrm{c}T)/T_\mathrm{c}`$. On the high-temperature side, the time constant of $`\alpha `$ relaxation diverges with a fractal exponent $`\gamma `$,
$$\tau |\sigma |^\gamma ,\sigma <0.$$
(2)
Such a transition has actually been observed in a colloidal suspension . In a structural glass former, the singularities (1) and (2) are smeared out because activated hopping processes restore ergodicity . Under these limitations, integral quantity like $`f_q`$ or $`\tau _\alpha `$ do not allow for decisive tests of theory. One rather has to study the full dynamics, as represented by the dynamic susceptibility $`\chi _q^{\prime \prime }(\omega )`$, or any other dynamic variable coupling to it.
Stronger predictions are made for the fast $`\beta `$ regime: around the minimum between $`\alpha `$ peak and vibrational excitations, in a temperature range close enough to, but sufficiently above $`T_\mathrm{c}`$, any susceptibility is expected to reach the same asymptotic limit
$$\chi ^{\prime \prime }(\omega )=\chi _\sigma g_\lambda (\omega /\omega _\sigma )$$
(3)
where the scaling function $`g_\lambda `$ is fully determined by one single parameter $`\lambda `$ . Amplitude and frequency scale of (3) should become singular on cooling towards $`T_\mathrm{c}`$,
$$\chi _\sigma |\sigma |^{1/2},\mathrm{and}\omega _\sigma |\sigma |^{1/2a},\sigma <0$$
(4)
where the exponent $`a`$, just as $`\gamma `$ in (2), is determined by $`\lambda `$.
### C Previous Studies of Propylene Carbonate
Table 1 summarises previous studies of fast dynamics in PC. All authors reported at least partial accord with MC predictions. However, not all MC interpretations were consistent with each other. It is an essential result of idealised MC theory that in the asymptotic $`\beta `$ regime all dynamic observables show the same spectral distribution and the same temperature dependence. Therefore, one material is characterised by just one $`\lambda `$ and one $`T_\mathrm{c}`$. Some of the $`T_\mathrm{c}`$ reported for PC are therefore incompatible with MC theory.
As with other materials, early scattering experiments concentrated on the square-root singularity of $`f_q`$. As in other materials, this singularity remains elusive: in Brillouin scattering, data fitting depends on uncontrolled approximations for the memory function . Similarly, in neutron scattering a determination of $`T_\mathrm{c}`$ from (1) works at best if the full lineshape on the $`\sigma >0`$ side is known from high-resolution spectroscopy . Therefore, the isolated results $`T_\mathrm{c}=210`$ or 270 K can be discarded from further consideration.
In comparison, the determination of $`T_\mathrm{c}`$ from viscosity or $`\alpha `$-relaxation data works better. However, available data do not allow for an independent determination of $`\lambda `$, and even for one given value $`\lambda =0.70`$, results vary between $`T_\mathrm{c}=180`$ and 196 K.
A dynamic susceptibility has been measured first by depolarised light scattering , yielding $`\lambda =0.78`$ and $`T_\mathrm{c}=187`$. These values have shown to be consistent with dielectric loss spectroscopy . However, a time-domain optical measurement of the solvation response of a solute molecule found a significantly lower $`T_\mathrm{c}`$ if the light scattering value of $`\lambda =0.78`$ was assumed .
In the following section II, we study the fast relaxation regime by neutron scattering. The $`\lambda `$ and $`T_\mathrm{c}`$ we obtain differ significantly from results of the other dynamic measurements which motivates us to remeasure some light scattering spectra (sect. III) and to reanalyse dielectric-loss and solvation-response data (sect. IV). Only then, all susceptibilities will be compared in section V.
## II Neutron Scattering
### A Experiments
Inelastic, incoherent neutron scattering has been measured on the time-of-flight spectrometers Mibémol at the Laboratoire Léon Brillouin, Saclay, and Neat at the Hahn-Meitner-Institut, Berlin.
On Mibémol, the counter-rotating choppers were operated with 10 000 rpm. With an incident neutron wavelength of $`\lambda _\mathrm{i}=8.5`$Å, we obtained a resolution (fwhm) of $`6.59.5`$GHz, depending on angle. At the Berlin reactor, the flux delivered by the undermoderated cold source decreases rather fast for wavelengths beyond 5 or 6 Å; therefore $`\lambda _\mathrm{i}=5.5`$Å was chosen. With counter-rotating choppers at 20 000 rpm, we achieved nevertheless a resolution of $`1113.5`$GHz. Total count rates were of the same order for both instruments; a precise comparison cannot be made because different sample geometries were used.
For the Mibémol experiment, we used an Al hollow cylinder with 30 mm outer diameter and a sample layer of about 0.1 mm thickness. In this container, the sample crystallised partially during a 190 K scan; at 175 K, as far as one can tell from the elastic structure factor of a predominantly incoherent scatterer, complete or nearly complete crystallisation occured within 2 hours. On the other hand, after rapidly cooling from 260 K, it was possible to perform a 2 K scan in the fully amorphous state.
On Neat, we tried to remeasure the dynamics of the supercooled state below 200 K. To this purpose, we filled the sample into nearly 200 thin capillaries (soda lime, inner diameter 0.2 mm, Hilgenberg). As in other liquids, this packaging proved highly successful; comparison to a vanadium scattering and visual inspection showed that down to 168 K no crystallisation occured. Because the detectors are located at only 250 cm from the sample, the resolution of Neat is very sensitive to flightpath differences. We therefore abstained from using a hollow cylindrical geometry; instead, we placed the capillaries in a rectangular holder with a 30 mm base length which was then mounted at $`45^{}`$ with respect to the incoming beam. For both experiments, the sample material, 1,2-propylene carbonate (99.7 %, Sigma–Aldrich), was loaded under an inert gas.
After converting the raw data to $`S(2\theta ,\omega )`$, they were binned in about 30 angular groups; a non-equidistant frequency binning was imposed by requiring statistic fluctuations to fall below a given mark. Only then was the container scattering subtracted. The absolute intensity scale was taken from the elastic scattering of the Mibémol 2 K scan; for Neat, detectors were first calibrated to vanadium before the overall scale was fit to Mibémol at 210 K.
The scattering law $`S(2\theta ,\omega )`$ depends still on $`\lambda _\mathrm{i}`$. Only after interpolation to constant wavenumbers $`q`$ one obtains spectra $`S(q,\omega )`$ which are independent of the kinematics of scattering, and only then a direct comparison between our two experiments becomes possible. This comparison is performed explicitely at 207/210 K in Fig. 1. In the quasielastic scattering range, up to some 100 GHz, the accord between Mibémol and Neat is excellent.
### B The Factorisation Property
For analysing the data, especially in the quasielastic range, it is advantageous to visualise the scattering law as a susceptibility
$$\chi _q^{\prime \prime }(\omega )=S(q,\omega )/n(\omega )$$
(5)
with the Bose factor $`n(\omega )=(\mathrm{exp}(\mathrm{}\omega /k_\mathrm{B}T)1)^1`$.
Around the susceptibility minimum, mode coupling theory predicts that the $`\beta `$ line shape is asymptotically the same for all experimental observables which couple to density fluctuations; thus a fortiori it must be the same for incoherent neutron susceptibilities measured at different wavenumbers $`q`$. We therefore expect $`\chi _q^{\prime \prime }(\omega )`$ to factorise into an $`\omega `$-independent amplitude and a $`q`$-independent spectral function,
$$\chi _q^{\prime \prime }(\omega )=h_q\chi ^{\prime \prime }(\omega ).$$
(6)
On different theoretical grounds, such a factorisation is also expected for incoherent scattering from harmonic vibrations in lowest order of mass expansion .
There are several ways to test (6) and to determine $`h_q`$: e. g., one may iteratively construct a model for $`\chi ^{\prime \prime }(\omega )`$ and fit it to the individual $`\chi _q^{\prime \prime }(\omega )`$ data sets. More simply, $`h_q`$ can be calculated from a least-squares match between neighbouring $`q`$ cuts . For the present data, we find that $`h_q`$ does not depend on the chosen procedure, nor does it vary with the frequency subrange from which it is determined.
A first surprising result is then the strictly linear wavenumber dependence of $`h_q`$ (Fig. 2a). Within MC theory, $`q`$ dependences can be calculated only if a specific microscopic structure is put in. Within harmonic theory, however, there is a clear expectation that $`h_qq^2`$. Astonishingly enough, this $`q^2`$ dependence is never seen. Instead, in at least two other molecular systems a linear $`h_qq`$ dependence is found though the temperature dependence of THz modes indicates pure harmonic behaviour . From other time-of-flight studies we suspect and a simulation confirms that this is mainly a multiple-scattering effect, Anyway, the linearity is so accurate that in our analysis we shall use (6) with $`h_q=q/q_0`$ (with an arbitrary normalisation $`q_0=1`$Å<sup>-1</sup>) instead of employing the empirical values.
Fig. 3 shows that the factorisation holds over the full experimental data range, at least from 30 to 2500 GHz, except for the onset of $`\alpha `$ relaxation at the highest temperature (which is associated with long-ranged transport and depends therefore strongly on $`q`$). This allows us to condense our huge two-dimensional data sets into $`q`$-independent functions $`\chi ^{\prime \prime }(\omega )`$ with much improved statistics; only on this basis is it possible to proceed with a quantitative analysis of noisy data accumulated in relatively short scans.
Before proceeding with the analysis we have to note that the $`\chi ^{\prime \prime }(\omega )`$ determined by this method and shown for 207/210 K in Fig. 2b reveals considerable discrepancies between Mibémol and Neat at frequencies above about 400 GHz. Although these deviations passed unnoticed on the double logarithmic scale of Fig. 1b, they are already present in the raw data. While we can exclude container scattering and dark counts as possible causes, two other frequency-dependent effects are likely to contribute: multiple scattering and an inaccurate detector efficiency correction . Multiple scattering cannot be corrected for unless a comprehensive theoretical model of $`S(q,\omega )`$ is used as input; therefore, it presents a fundamental limitation to the determination of spectral shapes by quasielastic neutron scattering.
### C Master Curves for $`𝜷`$ Relaxation
The $`q`$-independent $`\chi ^{\prime \prime }(\omega )`$ can now be used to test the scaling form (3). We use fits with a fixed $`\lambda `$ to determine $`\chi _\sigma `$ and $`\omega _\sigma `$. With these values, a master curve $`\chi ^{\prime \prime }(\omega /\omega _\sigma )/\chi _\sigma `$ is constructed, and from the master curve the scaling range is read off which is then used for improved fits to the original data. Figure 4 shows fits to the original $`\chi ^{\prime \prime }(\omega )`$ for $`\lambda =0.72`$, and Figure 5 shows master curves for three $`\lambda `$’s.
In principle, a self-consistent $`\lambda `$ can be determined in an iterative procedure from free fits to the master curves. For our $`220260`$K data, $`\lambda `$ tends towards values around 0.69 (Figure 5a), but the convergence is erratic, and the outcome depends on the subjective decision of which points to include in the master curve.
With the data on hand, a more restrictive determination of $`\lambda `$ is possible from the temperature dependence (4). The insets in Figure 5 show that a consistent linear behaviour $`\chi _{\sigma }^{}{}_{}{}^{2}\omega _{\sigma }^{}{}_{}{}^{2a}\sigma `$ is found only for $`\lambda =0.72`$ and $`T_\mathrm{c}=182`$K; the exponent $`a=0.32`$, which corresponds to $`\lambda =0.72`$, is confirmed by cross-checking $`\mathrm{ln}\chi _\sigma `$ vs. $`\mathrm{ln}\omega _\sigma `$ (inset of Figure 4).
Figure 5c demonstrates that the value $`\lambda =0.78`$ suggested by preceeding light scattering and dielectric loss measurements do not give a good description of the neutron scattering data: the master curve is of poorer quality than for $`\lambda =0.69`$ or 0.72, especially at frequencies around and below the minimum; free fits with $`g_\lambda `$ show a clear trend towards smaller values of $`\lambda `$; and the scales $`\chi _\sigma `$ and $`\omega _\sigma `$ do not consistently follow (4).
This discrepancy motivates us to remeasure some light scattering spectra around the susceptibility minimum.
## III Light Scattering
### A Previous Studies
Light scattering, just as neutron scattering, comprises coherent and incoherent contributions. Since the wavelength of visible light is much longer than molecular dimensions, coherent scattering arises only from sound modes, giving rise to discrete Brillouin lines. Incoherent scattering, on the other hand, sees local motion and yields a continuous spectrum. In a very first approximation, this spectrum can be interpreted as if it were a $`q`$ averaged neutron scattering law; microscopic models suggest that the scattering mechanism involves four-point density correlations and/or rotational motion.
Brillouin scattering yields the velocity and damping constant of sound waves. In principle , the sound dispersion through the glass transition reveals the strength of the $`\alpha `$ relaxation so that it is possible to read off the Debye-Waller factor $`f_q`$ from fits to the Brillouin lines. In practice, this led to an inconsistent estimate of $`T_\mathrm{c}`$ in PC ; for a reliable determination of $`f_q(T)`$, one should not only determine one limiting sound velocity by ultrasonic experiments, but also provide an independently determined memory function as input .
More direct information on the microscopic dynamics is obtained from the incoherent continuous spectra. In order to suppress the much stronger Brillouin lines, these spectra are preferentially gathered in depolarised (VH) back-scattering geometry. In an extensive study of PC, besides a thorough discussion of Brillouin scattering, Du et al. have also measured VH spectra over a wide frequency band, made accessible by combining a tandem interferometer with a grating monochromator. The measurements extended over a large temperature range, including $`\alpha `$ relaxation as well as hopping processes below $`T_\mathrm{c}`$, and were analysed within both the “idealised” and “extended” version of MC theory.
While these measurements gave a broad overview of the dynamics of glass-forming PC, they did not yield spectral lineshapes with the precision we need now for a quantitative comparison with neutron scattering results. In particular, the exponent parameter $`\lambda =0.78`$, obtained within idealised theory from a global fit to $`T>T_\mathrm{c}`$ data, was given with a relatively large error range of $`\pm 0.05`$. Furthermore, as in other broad band light scattering studies performed until quite recently, the tandem interferometer was used in series with an insufficient bandpass which did not fully suppress higher-order transmissions of the interferometer . For these reasons, we remeasured some depolarised spectra, concentrating on the temperature and frequency range of the asymptotic $`\beta `$ regime.
### B Experimental Set-up
Experiments were performed in Garching on a Fabry-Perot six-pass tandem interferometer. The instrument, bought from J. R. Sandercock, has been modified in several details to allow for stable operation and high contrast. The six-pass optics has been placed in a thermally isolating housing, and the scanning stage is actively temperature stabilised. The analogue-electronic stabilisation of the interferometer piezos is replaced by computer control. Entrance and exit pinholes are spatially separated from the six-pass optics, so that the most critical alignements can be done without disturbing the interferometer operation. By placing additional masks in the six-pass optics, in particular on mirror surfaces, the cross-talk between different passes could be reduced by several orders, and a straylight rejection of better than $`10^{11}`$ was achieved.
Depending on the free spectral range, the instrument is used in series with an interference filter of 0.15 or 1 GHz band width (Andover); these filters are placed in a special housing with active temperature stabilisation. Furthermore, to account for long-term drift, the instrument function is redetermined periodically by automatic white-light scans.
Although the spectrometer guaranteed excellent straylight rejection, special care was taken to prevent direct or diffuse reflections of laser light from entering the instrument. Therefore, instead of 180 back-scattering, a 169 VH geometry was chosen. From the intensity transmitted through the “ghosts” of the tandem instrument, we conclude that the straylight was about $`10^5`$ times weaker than the inelastic scattering from PC, and completely negligible compared to the detector’s dark count rate of about 2.5 sec<sup>-1</sup>. After subtraction of these dark counts, and normalisation to the corresponding white light scans, the resulting spectra showed excellent detailed balance symmetry.
The sample material was from the same source as for the neutron scattering experiment, and was vacuum sealed in a Duran cell.
### C Susceptibilities Around $`𝝎_𝝈`$
For a precise determination of the spectral lineshape, subsequent measurements over different spectral ranges were performed after stabilising the temperature over night. The most restrictive determination of the exponent parameter was possible at $`T=216`$K. As shown in Fig. 6, the matching of the three overlapping spectral ranges is excellent. Except for the leaking VV Brillouin mode, the VH susceptibility is described over more than two decades by the mode coupling asymptote $`g_\lambda `$. Fits yield $`\lambda =0.72\pm 0.01`$, which is confirmed by measurements at other temperatures as well as by a bulk of earlier experiments we had performed under less ideal experimental conditions. This result is at the margin of the error range in the literature value $`0.78\pm 0.05`$ ; the figure shows that 0.78 itself is clearly incompatible with our present data.
Secondly, for two different spectral ranges we measured temperature series. Fig. 7 shows some of the composite susceptibilities. From fits with fixed $`\lambda =0.72`$, the frequencies $`\omega _\sigma `$ were obtained. Then, the individual susceptibilities, measured around interferometer mirror spacings $`z_0=0.8`$ and 2.4 mm were fitted with fixed $`\lambda `$ and $`\omega _\sigma `$ so that we obtained two independent data sets for the amplitudes $`\chi _\sigma `$.
The temperature dependence of $`\omega _\sigma `$ is shown in Fig. 8a. Data between 190 and 230 K extrapolate to the same $`T_\mathrm{c}=182\pm 1`$ K as found from neutron scattering. Extending the fit range to higher temperatures leads to a $`T_\mathrm{c}`$ which is about 2 K lower. Again, these results are marginally compatible with $`T_\mathrm{c}=187\pm 5`$ K from the earlier light scattering study .
Measuring amplitudes in light scattering is difficult, and the prediction (4) can be verified only over a reduced temperature range. Results are visualised best in a logarithmic plot of $`\chi _\sigma `$ vs. $`TT_\mathrm{c}`$ (Fig. 8b) which suggests there are two different regions in which $`\chi _\sigma `$ is proportional to $`|\sigma |^{1/2}`$, separated by some step. Without further experiments, we must leave open whether this step comes from the sample or presents an experimental artifact, due for instance to distortions of the optical paths. We note, however, that after 30 hours and a full temperature cycle, we were able to reproduce an amplitude $`\chi _\sigma `$ within less than 1 %. Thus, irregularities in the amplitude are due not to a drift in time, but mainly to reversible effects of temperature variation.
Putting aside the amplitude problem, our results are in excellent accord with the neutron scattering data of sect. II. This motivates us to reconsider dielectric loss data which have been analysed previously with $`\lambda =0.78`$.
## IV Other Spectroscopies
### A Dielectric Loss
Dielectric spectroscopy on PC has been described recently . The measurements extended over a wide range of temperatures and many decades in frequency, and the analysis has addressed different issues which are currently debated in the context of glass transition dynamics.
Within the GHz–THz range, the dielectric susceptibility $`ϵ^{\prime \prime }(\omega )`$ passes through a minimum as suggested by MC theory. Fits with Eq. (3), approximated as a sum of two power laws , gave $`\lambda =0.78`$, and from the temperature dependence of frequency and amplitude (4) $`T_\mathrm{c}187`$K was found. Though these values had strong support in the existing literature, they differ from our neutron and light scattering results. It is therefore interesting to ask whether the dielectric loss would also allow for $`\lambda =0.72`$.
Since the available dielectric data are even noisier than the neutron susceptibilities, there is no unique way to determine the scaling range within which Eq. (3) applies; therefore, any value of $`\lambda `$ depends on the choice of the fit range. In the previous analysis , fits were applied between about 1 and 600 GHz. For temperatures around 200 K this range covered both sides of the minimum equally well. When $`\lambda =0.72`$ is imposed in an iterative master curve construction, the scaling range evades towards lower frequencies, as can be seen from the fits in Fig. 9.
For direct comparison, Fig. 10 shows master curves constructed with two different values of $`\lambda `$. In the upper curve, with $`\lambda =0.78`$, the measured susceptibilities have been rescaled with exactly the $`\omega _\sigma `$ and $`ϵ_\sigma `$ shown in Fig. 8 of Ref. . The rescaling works particularly well on the high-frequency side of the minimum and shows nicely the $`\omega ^a`$ limit which has remained elusive in so many other experimental investigations. The lower curve, constructed with $`\lambda =0.72`$, shows that the scaling property is completely lost above $`\omega _\sigma `$, whereas it has significantly improved in the low-frequency wing. Furthermore, the $`\omega ^b`$ limit of $`g_\lambda `$ describes the data nearly up to the $`\alpha `$-relaxation maximum where they almost coincide, except for the lowest two temperatures.
This unexpected observation motivates a new master curve construction which is quite common in the conventional analysis of $`\alpha `$ relaxation, but which has never before been applied to MC $`\beta `$ relaxation: in Fig. 11, the dielectric data are shown with original amplitude, rescaled only in frequency by the $`\alpha `$ maximum frequency $`\omega _{\mathrm{max}}`$. As usual for $`\alpha `$ relaxation (and in accord with the second scaling law of MC theory) the amplitude and the line shape of the $`\alpha `$ peak are temperature independent. But here, the scaling behaviour extends far down into the high-frequency wing — in fact, it extends as far as the data have been measured, except again the two lowest temperatures, 193 and 203 K.
However, a master curve which extends from the maximum up to the minimum of a dynamic susceptibility cannot simultaneously obey the first and the second scaling law of MC: we expect
$$\omega _{\mathrm{max}}/\omega _{\mathrm{min}}|\sigma |^{1/2b},$$
(7)
and, even more elementary, $`ϵ_{\mathrm{min}}|\sigma |^{1/2}`$ whereas $`ϵ_{\mathrm{max}}`$ remains constant.
Since there is no doubt about the scaling of $`\alpha `$ relaxation, we are bound to conclude that the dielectric data do not reach the asymptotic regime of fast $`\beta `$ relaxation, except in a rather small temperature range that extends at best to about 210 K. For higher temperatures, the high-frequency wing of $`\alpha `$ relaxation, though technically describable by $`g_\lambda `$, does not represent the first scaling law limit.
This conclusion is independent of any fitting, and does in particular not depend upon an imposed value of $`\lambda `$.
### B Solvation Dynamics
The solvation response of $`s`$-tetrazine has been measured in PC from 1.5 to 100 ps . A MC analysis in the time domain identified both $`\alpha `$\- and $`\beta `$-scaling regions in the dynamics . A unified analysis of both regions was consistent with MC theory for a range of $`\lambda `$$`T_\mathrm{c}`$ pairs. In the original publication, $`\lambda =0.78`$ was fixed, in accord with the light scattering analysis of Du et al. . However, this value of $`\lambda `$ yielded a cross-over temperature $`T_\mathrm{c}=176`$K substantially below the value obtained by Du et al. ($`T_\mathrm{c}=187`$K).
The solvation data have been reanalysed by the same method used in Ref. , but using $`\lambda =0.72`$. The analysis of the $`\alpha `$-scaling region is unchanged. Figure 12 shows the new beta-scaling plot and the fit to $`g_\lambda (t/t_\sigma )`$, both of which are as good as in the previous analysis. The inset shows a temperature scaling plot of $`t_\sigma `$ and $`\tau _\alpha `$. As expected, the scaling law deteriorates at temperatures far above $`T_\mathrm{c}`$. Depending on the range of temperatures fit, acceptable values of $`T_\mathrm{c}`$ lie in the range 178–182 K. Thus, the solvation response is consistent with both the $`\lambda `$ and $`T_\mathrm{c}`$ obtained in Sects. II and III by neutron and light scattering.
## V Comparison of Dynamic Observables
### A Direct Comparison of Susceptibilities
Neutron scattering, light scattering with or without proper bandpass, dielectric loss: each experiment, taken alone, seemed in full accord with the asymptotic predictions of MC theory. Taken together, the situation becomes more complicated.
In the fast $`\beta `$ regime, any observable that couples to density fluctuations is expected to tend towards the same asymptotic limit (3). For neutron scattering, this prediction takes the form of a $`q`$,$`\omega `$ factorisation, and is confirmed, in PC as in other materials, over a wide frequency range. However, it breaks down completely when depolarised light scattering or dielectric loss are included.
In Fig. 13, dynamic susceptibilities from neutron scattering, light scattering , and dielectric loss are shown for direct comparison on an arbitrary intensity scale, but in absolute frequency units. The result is in flagrant contradiction to any factorisation property: there is no one temperature for which the measured $`ϵ^{\prime \prime }(\omega )`$ and $`\chi ^{\prime \prime }(\omega )`$ fall together; in particular, their minimum frequencies differ systematically.
On a first sight, this outcome is a bad surprise, and could make us doubt whether fits of individual data sets with the asymptotic laws (3) and (4) are meaningful at all. On closer examination, the discrepancies between the three data sets can be traced back to two major differences: to the individually different temperature ranges within which the $`\beta `$ asymptote applies, and to a systematic shift in the frequency scale $`\omega _\sigma `$.
### B Temperature Range of $`𝜷`$ Relaxation
Within idealised MC theory, all dynamic observables converge towards the same scaling limit (3) and (4), characterised by just one lineshape parameter $`\lambda `$ and one frequency scale $`\omega _\sigma `$. This universality, however, is restricted to the lowest-order asymptote and does not imply a universal radius of convergence: the next-to-leading-order corrections already depend on the microscopic coupling so that different observables may reach the asymptote at different temperatures and frequencies.
Nevertheless, for neutron and light scattering, as well as for solvation response, (3) and (4) hold over rather large temperature intervals. In neutron scattering, (4) holds best between 220 and 260 K; at 210 K, the susceptibility minimum approaches the instrumental resolution function, and the signal becomes very weak: these technical limitations prevent us from following $`S(q,\omega )`$ closer to $`T_\mathrm{c}`$. For light scattering, Fig. 8 confirms (4) with reasonable precision up to nearly 1.5 $`T_\mathrm{c}`$.
On the other hand, the $`\alpha `$ master curve in Fig. 11 suggested that the range of $`\beta `$ scaling is rather small for dielectric loss. This is corroborated by Fig. 14 in which the $`\omega _\sigma `$ from fits with fixed $`\lambda =0.72`$ are compiled. For the lowest temperatures 193–213 K, the minima of $`ϵ^{\prime \prime }(\omega )`$ coincide with the light scattering data, whereas for higher temperatures the dielectric $`\omega _\sigma `$ cross over to a much steeper temperature dependence characteristic for $`\alpha `$ relaxation (actually even steeper than the MC prediction $`|\sigma |^{1/2a+1/2b}`$).
### C Frequency Scales $`𝝎_𝝈`$
As Fig. 14 shows, the $`\omega _\sigma `$ from neutron scattering fall about 35 % below the $`\omega _\sigma `$ from light scattering. The good agreement of $`\omega _\sigma `$ from neutron scattering with $`t_{\sigma }^{}{}_{}{}^{1}`$ from solvation dynamics excludes the possibility that the discrepancy between neutron and light scattering lies only in multiple scattering or other technical shortcomings of neutron scattering. Ergo, at least one of the two scattering techniques does not see the true $`\beta `$ asymptote of MC theory.
On the other hand, the observed scaling, the quality of the $`g_\lambda `$ fits, and the accord of the $`\lambda `$ and $`T_\mathrm{c}`$ call for a MC interpretation, and indicate a transient behaviour that is closely coupled to the universal $`\beta `$ asymptote.
## VI Discussion
### A Strength of $`𝜶`$ Relaxation
Although there is no universal criterion for the validity of $`\beta `$ scaling, an upper limit may be given: it is plausible that the asymptote (3) which is based on an expansion around the susceptibility minimum will no longer apply when the minimum ceases to exist. This happens when the height of the minimum, $`\chi _\sigma `$, becomes comparable to the maximum of $`\chi ^{\prime \prime }(\omega )`$ in the vibrational band. This condition also determines the highest temperature for which the proportionality (4) can hold.
On the other hand, the temperature range over which (4) holds can be related to the strength of the $`\alpha `$ peak. To explain this relation, we refer back to the $`\alpha `$ master plot in Fig. 11. Besides the dielectric $`ϵ^{\prime \prime }(\omega /\omega _{\mathrm{max}})`$, the figure also shows $`\chi _{\mathrm{VH}}^{\prime \prime }(\omega /\omega _{\mathrm{max}})`$ from depolarised light scattering. For both data sets, the same frequency scale $`\omega _{\mathrm{max}}`$ has been applied, as determined from the maximum position of the dielectric $`\alpha `$ peak. The intensity scale of the light scattering data is arbitrary and has been adjusted by a global factor so that the maximum of the microscopic excitations has about the same height as for the dielectric data which are measured in absolute units.
Compared to the high-frequency maximum, the $`\alpha `$ peak is about four times weaker in light scattering than in dielectric loss. In the high-frequency wing, at, say, $`\omega 30\omega _{\mathrm{max}}`$, the ratio is reduced to a factor of about 2. This reduction is due to different positions of the $`\alpha `$ maxima, and to different slopes of the wings.
At still higher frequencies, the susceptibilities cross over towards the minimum. For the lowest temperatures shown in the figure, the ratio of 2 between dielectric and light scattering data survives in this frequency range. Thus, for a low-temperature value of $`\sigma `$, the $`\beta `$ amplitude $`\chi _\sigma `$ for light scattering is about two times smaller than for dielectric loss.
On the other, from the condition stated above, and because we have rescaled the susceptibilities to about the same values at high frequencies, the temperature range of $`\beta `$ scaling ends for both techniques at about the same absolute value of $`\chi _\sigma `$. Using the proportionality $`\chi _\sigma |\sigma |^{1/2}`$, the dielectric data reach this limit after a 4 times smaller temperature change than the light scattering data do.
Thus, the strength of $`\alpha `$ relaxation can explain why the $`\beta `$ scaling regime extends over only 30 K in dielectric loss, compared to 90 K or more in light scattering.
### B Frequency Range of $`𝜷`$ Relaxation
One can imagine many experimental imperfections that distort spectra measured in a scattering experiment. For instance, multiple scattering could overlay $`S(q,\omega )`$ with a convolution with itself. But any such distortions would affect the wings of the susceptibility much more than the region around the minimum. In particular, convolutions lead to similar corrections as intrinsic next-to-leading-order terms. Therefore, it is not easy to explain the discrepancy observed in the $`\omega _\sigma `$.
We note, however, that the $`g_\lambda `$ fits to the neutron scattering data in Fig. 4 extend up to frequencies between 200 and 400 GHz, whereas in light scattering, the fitted $`g_\lambda `$ already deviate from the measured susceptibilities a little above 100 GHz. It is therefore quite possible that our neutron scattering analysis sees a preasymptotic transient rather than the true $`g_\lambda `$.
This would ressemble the situation in glycerol , where the $`\chi ^{\prime \prime }(\omega )`$ allowed for the construction of a master curve and followed some MC predictions, although the full $`\beta `$ asymptote was not reached. Numeric solutions of a simple two-correlator model showed how such a scenario can arise from a MC ansatz . Similar fits, with one slave correlator for each observable, also work for propylene carbonate .
### C Current Theoretical Developments
Some of the questions raised by our experiments are also addressed by recent numerical solutions of MC equations and molecular-dynamics simulations.
A MC analysis of the hard-sphere liquid as a function of wavenumber has demonstrated that corrections to scaling are of differing importance for different observables. Analytic expansions have shown that corrections to the $`\alpha `$ process are of higher order than corrections to the $`\beta `$ asymptote: this may explain why $`\alpha `$ scaling holds over a wider range than the factorisation property of $`\beta `$ relaxation .
In a molecular-dynamics simulation of a liquid made of rigid diatomic molecules , orientational correlation functions have been analysed for different angular momenta $`l=0,1,2,\mathrm{}`$ All correlators were found to fulfil the MC factorisation — with the pronounced exception of $`l=1`$ for which the position of the susceptibility minimum is shifted by a factor of 10 \[loc. cit., Fig. 8\]. The authors attribute this peculiar behaviour to $`180^{}`$ jumps, which is not necessarily the right explanation for similar behaviour in more complex fluids. A scenario with an underlying type A transition for odd $`l`$ has been proposed recently . It has been noted before that the dielectric response couples only weakly to vibrational excitations , and the peculiar strength of the $`\alpha `$ peak in the $`l=1`$ correlator has also been found in an analytic extension of MC theory to non-spherical particles .
A decomposition according to angular momenta might also explain the excellent accord between neutron scattering and solvation response found in Fig. 14, since both techniques see $`l=0`$ whereas light scattering might be dominated by $`l=2`$ .
## Acknowledgements
We thank H. Z. Cummins, M. Diehl, J. K. Krüger, H. Leyser, J. R. Sandercock, A. P. Sokolov, and J. Wiedersich for invaluable advice in setting up our Fabry-Perot spectrometer.
We are grateful to W. Petry for continuous support, and we thank him as well as M. Fuchs, W. Götze, A. P. Singh and T. Voigtmann for fruitful discussions.
We acknowledge financial aid by the Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie through Verbundprojekte 03pe4tum9 and 03lo5au28 and through contract no. 13n6917, by the Deutsche Forschungsgemeinschaft under grant no. Lo264/8–1, by the European Commission through Human Capital and Mobility Program ERB CHGECT 920001, and by the National Science Foundation under Che9809719.
The Laboratoire Léon Brillouin is a laboratoire commun CEA – CNRS.
|
no-problem/9908/hep-th9908010.html
|
ar5iv
|
text
|
# Untitled Document
hep-th/9908010 TUW-99-17
Probing the Strong Coupling Limit of Large $`N`$ SYM on Curved Backgrounds
Karl Landsteiner and Esperanza Lopez
Institut für theoretische Physik
Technische Universität Wien
Wiedner Hauptstraße 8-10
A-1040 Wien, Austria
[email protected]
[email protected]
Abstract
According to the AdS/CFT correspondence, the strong coupling limit of large $`N`$ $`𝒩=4`$ supersymmetric gauge theory at finite temperature is described by asymptotically anti de Sitter black holes. These black holes exist with planar, spherical and hyperbolic horizon geometries. We concentrate on the hyperbolic and spherical cases and probe the associated gauge theories with D3-branes and Wilson loops. The D3-brane probe reproduces the coupling of the scalars in the gauge theory to the background geometry and we find thermal stabilization in the hyperbolic case. We investigate the vacuum expectation value of Wilson loops with particular emphasis on the screening length at finite temperature. We find that the thermal phase transition of the theory on the sphere is not related to screening phenomena.
August, 1999
1. Introduction
Maldacena’s conjecture states that $`SU(N)`$ $`𝒩=4`$ supersymmetric gauge theory in four dimensions is dual to type IIB string theory on an $`AdS_5\times S^5`$ background. This conjecture is obtained by considering a collection of $`N`$ D3-branes in type IIB string theory. The effective theory describing the physics of the D3-branes is $`𝒩=4`$ $`SU(N)`$ Yang-Mills. D3-branes can also be described as classical black hole solutions in IIB supergravity. Taking $`\alpha ^{}0`$ and at the same time keeping the mass of the strings stretched between the D3-branes finite defines the near horizon limit of the supergravity solution. On the other hand one is keeping only the ground states of the open strings. This gives $`𝒩=4`$ gauge theory as the exact theory (not only the effective theory for low energies) and the $`AdS_5\times S^5`$ background as the near horizon limit in the supergravity solution.
The same arguments apply not only to extremal but also to near-extremal D3-branes . One obtains then asymptotically anti de Sitter black holes with planar horizon geometry. The euclidean continuation of these black hole solutions is conjectured to describe a thermal state of the $`𝒩=4`$ gauge theory with a temperature corresponding to the Hawking temperature of the black hole \[1,,3\]. Much work has been devoted to the study of this duality, see e.g. for a review.
The supergravity solution is however only an approximation to string theory and corresponds to the large $`N`$ and strong coupling limit in the dual gauge theory. A direct comparison of quantities calculated at weak coupling in the gauge theory with the corresponding results in the supergravity theory is therefore rather difficult. However many quantities calculated at strong coupling have been shown to be consistent at the qualitative level with general expectations from gauge theories. Among the quantities that have been studied for near-extremal D3-branes are effective potentials obtained through D3-brane probes \[5,,6\] and vacuum expectation values of Wilson loops corresponding to the action of strings ending on the conformal boundary of the AdS black hole \[7,,8,,9,,10,,11\].
Black holes in AdS space can have spherical, planar or hyperbolic horizon geometries. Although only the planar black holes can be directly related to D3-branes in a thermal state, one expects that also the spherical and hyperbolic black holes describe gauge theories. More specifically in it was conjectured that spherical black holes describe $`𝒩=4`$ gauge theories at finite temperature on $`S^3`$. One also expects that AdS black holes with hyperbolic horizon describe $`𝒩=4`$ gauge theory on the hyperbolic three-plane $`H^3`$. We will further investigate these conjectures. In section 2 we will compute the static potential for a test D3-brane in the background of spherical or hyperbolic AdS black holes. We find that at large distances the static potential reproduces the tree level coupling of the scalars to the background geometry in the gauge theory. Since in the hyperbolic case this coupling induces a negative mass term one would expect the gauge theory to be unstable. The D3-brane potential at finite temperature grows however for small distances up to a maximum value and only for large distances reproduces the negative mass term in the gauge theory. At zero temperature there is however no such stabilization and one finds only the tree level negative mass term <sup>1</sup> Similar behavior of the potential of M2-branes in the background of non-extremal D-branes with hyperbolic horizon has been found in ..
In section 3 we turn to the properties of Wilson loops. We find screening at finite temperature on the sphere and at zero temperature a potential that is very close to the form of the potential at weak coupling. We also speculate about the relation of the screening length at finite temperature and the thermal phase transition between the spherical AdS black hole and empty AdS space. In the hyperbolic case we find screening even at zero temperature.
2. Probing AdS Black Holes with D-Branes
We begin with the relevant bosonic part of the type IIB action<sup>2</sup> We take a Yang-Mills term for the five-form action with the convention of imposing self-duality at the level of the equations of motion.
$$S_{IIB}=\frac{1}{2\kappa _{10}^2}d^{10}x\sqrt{g}\left(R\frac{1}{4.5!}F_5^2\right).$$
The ten-dimensional spacetime will be the product manifold of $`S^5`$ and a five-dimensional AdS black hole. The flux of the five-form field strength through $`S^5`$ is quantized in terms of the D3-brane charge as $`F=2\kappa _{10}^2\mu _3N`$, with $`\kappa _{10}^2=64\pi ^7g_s\alpha ^4`$ and $`\mu _3^2=\frac{\pi }{\kappa _{10}^2}`$. The metric of the euclidean AdS black holes is
$$ds^2=\mathrm{}^2\left[\left(u^2+k\frac{\mu }{u^2}\right)d\tau ^2+\frac{du^2}{u^2+k\frac{\mu }{u^2}}+u^2d\mathrm{\Sigma }_{3,k}^2\right],$$
where $`\mathrm{}`$ is the radius of curvature of the asymptotic AdS geometry and $`\mu `$ is proportional to the black hole mass. The discrete parameter $`k`$ can take the values $`(1,0,1)`$ corresponding to the spherical, planar or hyperbolic cases respectively. In the following we will always consider $`k=1`$ or $`k=1`$. $`d\mathrm{\Sigma }_{3,k}^2`$ denotes then the metric on the unit 3-sphere or 3-hyperboloid
$$\begin{array}{cc}\hfill d\mathrm{\Sigma }_{3,1}^2=& d\theta ^2+\mathrm{sin}^2\theta d\mathrm{\Omega }_2^2,\hfill \\ \hfill d\mathrm{\Sigma }_{3,1}^2=& d\theta ^2+\mathrm{sinh}^2\theta d\mathrm{\Omega }_2^2.\hfill \end{array}$$
The geometry defined in (2.1) is free of conical singularities if the euclidean time coordinate $`\tau `$ is periodically identified, with period
$$\beta =\frac{2\pi u_+}{2u_+^2+k},$$
where $`u_+`$ is the horizon radius
$$u_+=\sqrt{\frac{k}{2}+\sqrt{\frac{1}{4}+\mu }}.$$
The temperature associated to (2.1) is $`T=\frac{1}{\beta }`$. Since the boundary of asymptotically AdS metrics is only defined up to conformal transformations , we have chosen (2.1) such that the boundary 3-sphere or 3-hyperboloid has radius one. Although an arbitrary radius $`R`$ can be set by redefining $`u=R\stackrel{~}{u}`$ and $`\tau =\stackrel{~}{\tau }/R`$, the thermodynamics will only depend on the ratio $`\beta _{S^1}/R`$ given by (2.1).
The black hole metrics (2.1) with $`k=1`$ only exist for $`TT_{min}=\frac{\sqrt{2}}{\pi }`$, which corresponds to $`\mu =\frac{3}{4}`$. For each temperature bigger than $`T_{min}`$ there are two solutions for $`\mu `$, one bigger than $`\frac{3}{4}`$ and one smaller. We will only consider mass parameters $`\mu \frac{3}{4}`$. Black holes with smaller values of $`\mu `$ have negative specific heat. They decay due to Hawking radiation and do not correspond to a stable thermal state in the gauge theory. There is furthermore a phase transition at $`\mu _{crit}=2`$, or equivalently
$$T_{crit}=\frac{3}{2\pi }.$$
For temperatures below $`T_{crit}`$ the black hole solution is unstable due to tunneling to empty AdS-space . The gauge theory interpretation is that of a confining/deconfining phase transition \[13,,3\]. The free energy of the black hole solutions scales like $`N^2`$, consistent with the expected deconfinement at finite temperatures in the gauge theory. For temperatures $`T<T_{crit}`$ the field theory dual is empty AdS. Its free energy scales like $`N^0`$, signalling confinement. We will discuss this phase transition in more detail in section 3.
For $`k=1`$ the horizon radius is defined even for negative mass parameters down to $`\mu =\frac{1}{4}`$. Lorentzian black holes with negative mass parameter have also an inner horizon and a timelike singularity such that their Penrose diagram is that of the Reissner-Nordstrom AdS solution . At the extremal case $`\mu =\frac{1}{4}`$ inner and outer horizon coincide. The nature of the horizon changes, it becomes a bifurcate horizon with no particular temperature associated to it. Thus the extremal solution is somewhat similar to empty AdS in the Poincaré patch and can be associated with the zero temperature ground state of $`𝒩=4`$ Yang-Mills on the hyperboloid \[16,,17\]. The fact that the groundstate is given by a solution with negative mass parameter and its implications concerning the entropy of the corresponding gauge theory have been discussed in and the field theory 1-loop calculation of the free energy has been performed recently in \[18,,19\].
Taking the metrics (2.1) as input for a Freund-Rubin like ansatz we find that the length scale $`\mathrm{}`$ is determined by the five-form flux through $`S^5`$ as $`\mathrm{}^4=4\pi g_sN\alpha ^2`$ and the potential of the five-form field strength is given by
$$A_4=\mathrm{}^4u^4d\tau d\mathrm{\Sigma }_{3,k},$$
where $`d\mathrm{\Sigma }_{3,k}`$ is the unit volume form on either $`S^3`$ or $`H^3`$. Notice that the metrics (2.1) are not derived as near-horizon limits of D3-branes on an asymptotically flat space.
$`𝒩=4`$ Yang-Mills theory in flat space has a Coulomb branch of vacua on which the gauge group is broken to a smaller group of equal rank. In D-brane language this translates into forces between parallel extremal flat D3-branes being zero. Separating D3-branes in the transverse space dimensions corresponds to turn on Higgs expectation values in the gauge theory. At finite temperature the field theory develops an effective potential which lifts the Coulomb branch and drives the system back to symmetry restauration . Accordingly, forces between near-extremal D3-branes do not cancel. A symmetry breaking pattern $`U(N+1)U(N)\times U(1)`$ is given in the gravity description by separating a single D3-brane from the rest. The effective action of such a configuration can be computed as the action of the single D3-brane in the background of the other $`N`$ branes
$$S=T_3\left(\sqrt{\widehat{g}}+A_4\right),$$
where $`T_3`$ is the D3-brane tension, $`\widehat{g}`$ the induced metric on the probe world-volume and $`A_4`$ is the background value of the type IIB four-form potential. This calculation has been done in and and qualitative agreement with the expectations from gauge theory has been found. In this section we want to extend this analysis to the spherical and hyperbolic cases. The main difference with the flat case is that the $`𝒩=4`$ scalars can get mass due to a non-minimal coupling to the boundary geometry,
$$S_{SYM}=\frac{1}{g_{YM}^2}d^4x\sqrt{g}\mathrm{Tr}\left(\frac{1}{4}F_{\mu \nu }F^{\mu \nu }+\frac{1}{2}D_\mu \mathrm{\Phi }D^\mu \mathrm{\Phi }+\frac{1}{2}\xi R\mathrm{\Phi }^2+\mathrm{}\right).$$
Therefore we expect a non-zero effective potential even at zero temperature. Since the boundary geometry is only defined as a conformal class, the $`𝒩=4`$ scalars should be conformally coupled, i.e. $`\xi =1/6`$.
We consider a test D3-brane at a distance $`u`$ from the origin. Substituting (2.1) and (2.1) in the Born Infeld action (2.1) we obtain <sup>3</sup> In \[5,,6\] the authors worked with asymptotically flat geometries describing near-extremal 3-branes solutions. In defining the probe action, they subtracted the 3-brane volume at infinity. We are directly working in asymptotically AdS spaces and the D3-brane action at infinity does not become a constant therefore no such subtraction can be done. Our expression for $`k=0`$ coincides with and up to a constant also with .
$$S=\beta T_3V_3\mathrm{}^4u^4\left(\sqrt{1+\frac{k}{u^2}\frac{\mu }{u^4}}1\right),$$
with $`V_3`$ being the volume of the unit 3-sphere or 3-hyperboloid. For distances $`u>>\frac{1}{T}`$, the potential $`V=TS`$ derived from (2.1) becomes
$$V=\frac{k}{4\pi ^2}NV_3u^2,$$
where we used $`T_3=\frac{1}{(2\pi )^3g_s}=\frac{N}{2\pi ^2\mathrm{}^4}`$ . To compare this to the gauge theory in (2.1) we introduce the ’t Hooft coupling $`\lambda =2g_{YM}^2N`$ and identify it on the the supergravity side through $`\mathrm{}^4=\lambda \alpha ^2`$. The radial coordinate $`u`$ is then related to the Higgs expectation value by $`\frac{u}{2\pi }=\frac{\mathrm{\Phi }}{\sqrt{\lambda }}`$ \[23,,5,,6\]. The curvature of a unit 3-sphere or 3-hyperboloid is $`R=6k`$. Substituting these values in (2.1) with $`\xi =1/6`$, we obtain precisely (2.1). Thus we recover at large $`u`$ precisely the tree-level coupling of the scalar fields to the background curvature in the gauge theory. At first glance it seems surprising that this term does not suffer some non-trivial renormalization since supersymmetry is broken by both the background geometry and the temperature. At large $`u`$ we are however probing the ultraviolet properties of the gauge theory. Since $`𝒩=4`$ Yang-Mills on $`S^1\times S^3`$ or $`S^1\times H^3`$ has vanishing conformal anomaly, in the large $`u`$ limit we expect to probe the bare couplings of the $`𝒩=4`$ gauge theory. This result is an example for the suggestion that large distances in the bulk correspond to small distance scales on the boundary .
It is interesting to analyze the large $`u`$ behaviour of the potential when the dual gauge theory lives on an space with non-vanishing conformal anomaly. We will consider as an example $`AdS_5`$ with $`S^4`$ boundary of radius one. The metric is
$$ds^2=\mathrm{}^2\left(\frac{du^2}{u^2+1}+u^2d\mathrm{\Omega }_4^2\right).$$
Using this metric as input for a Freund-Rubin like ansatz, we obtain its associated four-form potential
$$A_4=\mathrm{}^4\left[\left(u^3\frac{3}{2}u\right)\sqrt{1+u^2}+\frac{3}{2}\mathrm{arcsh}u\right]d\mathrm{Vol}_{S^4}.$$
Substituting these values in (2.1) and expanding for large $`u`$ we get
$$V=\frac{1}{2\pi ^2}NV_4\left(u^2\frac{3}{2}\mathrm{log}u\right),$$
where $`V_4`$ is the volume of $`S^4`$. The curvature of a four-sphere of radius one is $`12`$. The first term in (2.1) reproduces the conformal coupling of the scalars to the background curvature with the same identification between $`u`$ and $`\mathrm{\Phi }`$ as before. We associate the second term to the conformal anomaly. The trace anomaly of $`𝒩=4`$ on $`S^4`$ is $`T=\frac{3N^2}{8\pi ^2}`$ and thus the effective action picks up a logarithmic term
$$\frac{3}{8\pi ^2}N^2V_4\mathrm{log}ϵ,$$
with $`ϵ`$ playing the part of the momentum cutoff. The coefficient that multiplies the logarithm in (2.1) is precisely the leading $`O(N)`$ term in $`T_NT_{N+1}`$. The variable $`u`$ sets the energy scale in the field theory at which the gauge group is Higgsed from $`SU(N+1)`$ to $`SU(N)`$. It is therefore $`u`$ what substitutes the cutoff $`ϵ`$ in (2.1).
The identification of $`u`$ with the vacuum expectation value of the scalar fields is only correct at large $`u`$. It has been argued that in order to compare radial distances in supergravity with energy scales in the gauge theory one should use an isotropic coordinate defined by
$$\rho =\mathrm{exp}\frac{du}{u^2+k\frac{\mu }{u^2}}=\left(u^2+\frac{k}{2}+\sqrt{u^4+ku^2\mu }\right)^{\frac{1}{2}},$$
or equivalently
$$u^2=\frac{\rho ^4+\mu +1/4}{2\rho ^2}\frac{k}{2}.$$
In terms of the new coordinate $`\rho `$, the difference between the D3-brane potential at infinity and at the horizon is
$$V(\rho _+)V(\mathrm{})=\frac{V_3N}{4\pi ^2}\left(k\frac{\rho ^2}{2}|_\rho \mathrm{}+\rho _+^42k\rho _+^2\right).$$
The action of the euclidean black hole solutions (2.1) has been calculated in \[16,,17\]. Using their results and our coordinate, we find for the free energy
$$F_N=\frac{V_3N^2}{8\pi ^2}(\rho _+^42k\rho _+^2).$$
If we add one D3-brane the free energy changes by an amount $`F_{N+1}F_N`$. The leading $`O(N)`$ term of this change is precisely given by (2.1) after discarding the quadratically divergent term. The quadratic divergence does not influence the thermodynamics since it is temperature independent. For this reason we do not see this contribution in the change of the free energy.
It is convenient to introduce an “effective temperature” defined by
$$\rho _+=\pi T_{eff},$$
$$T_{eff}=\frac{T}{\sqrt{2}}\left(1+\sqrt{1\frac{2k}{\pi ^2T^2}}\right)^{\frac{1}{2}}.$$
In analogy with and we then introduce the scalar mass parameter $`M`$ by
$$M=\rho \rho _+.$$
The static D3-brane potential takes now the form
$$V(M)=\frac{\pi ^2T_{eff}^4V_3N}{4}\left[1\frac{1}{\left(1+\frac{M}{\pi T_{eff}}\right)^4}+\frac{k}{2\pi ^2T_{eff}^2}\left(\frac{3}{\left(1+\frac{M}{\pi T_{eff}}\right)^2}+\left(1+\frac{M}{\pi T_{eff}}\right)^24\right)\right],$$
where we chose to subtract a constant such that it vanishes at the horizon. The first two terms in (2.1) reproduce the potential obtained for planar horizon in \[5,,6\] by substituting $`T_{eff}`$ with $`T`$ <sup>4</sup> Note however that the potentials in \[5,,6\] have not been chosen to vanish at the horizon.. Thus the effect of the curved background on the Higgs potential is to replace $`T`$ by $`T_{eff}`$ and add an additional term proportional to the curvature, i.e. the third term in (2.1). The large temperature limit can be alternatively interpreted as the large radius limit for the boundary 3-geometries. Notice that $`T_{eff}T`$ as $`T`$ tends to infinity. For completeness we give the expansion of the potential for values of the scalar mass $`M<<T_{eff}`$
$$V=\frac{\pi ^2T_{eff}^4V_3N}{4}\underset{k=1}{\overset{\mathrm{}}{}}\left(\frac{M}{\pi T_{eff}}\right)^k\left(a_k+\frac{k}{2\pi ^2T_{eff}^2}b_k\right),$$
where the coefficients are given by
$$a_k=()^k\left(\genfrac{}{}{0pt}{}{k+3}{k}\right),b_1=4,b_2=10,b_{k>2}=3()^k(k+1).$$
Fig. 1: Static D3-brane potential in the background of spherical AdS-black holes for $`T_{min}`$, $`T_{crit}`$ and $`T>T_{crit}`$. We have also plotted the potential obtained from empty AdS in the global patch.
The definition (2.1) of $`T_{eff}`$ allows us to treat also the case of empty AdS in the global patch by taking $`T_{eff}=\frac{1}{\sqrt{2}\pi }`$. The D3-brane potential for $`k=1`$ at different temperatures as a function of the scalar mass $`M`$ is shown in fig. 1.
Fig. 2: Static D3-brane potential in the background of hyperbolic AdS-black holes (on the vertical axes we have ploted $`V/NV_3`$).
For $`k=1`$, the conformal coupling to the curvature of the hyperboloid induces a negative mass term for the $`𝒩=4`$ scalars. We have already seen in (2.1) that this tree level coupling is reproduced in the D3-brane potential. Therefore we would expect the gauge theory to be unstable. An analysis of (2.1) shows however that the static potential grows for small $`M`$ reaching a maximum at
$$M=\pi T_{eff}\left(\frac{1+\left(2\pi ^2T_{eff}^2+\sqrt{4\pi ^4T_{eff}^41}\right)^{\frac{2}{3}}}{\left(2\pi ^2T_{eff}^2+\sqrt{4\pi ^4T_{eff}^41}\right)^{\frac{1}{3}}}\right)^{\frac{1}{2}}\pi T_{eff}.$$
This formula is valid also for $`T_{eff}<\frac{1}{\sqrt{2}\pi }`$, where $`\frac{1}{\sqrt{2}\pi }`$ is the effective temperature of empty AdS. The presence of a maximum means that in spite of the negative tree level mass the gauge theory is still driven back to the symmetric phase due to thermal effects as long as the scalar vacuum expectation value is not too large. Of course these statements apply to the limit of large $`N`$ and large ’t Hooft coupling $`\lambda `$, where the supergravity approximation is valid. Notice that the high of the potential maximum is proportional to $`N`$. For finite $`N`$ and coupling the theory could still be unstable e.g. due to tunneling of D3-branes through the potential wall. It would also be interesting to see how much of this behavior is reproduced at weak coupling. For this one would have to do a one-loop calculation of the effective potential of $`𝒩=4`$ SYM on the hyperbolic background in the Higgs phase at finite temperature. A precise comparison of the gauge theory result with the results obtained here from supergravity seems problematic due to the well-known gauge dependence of the effective potential. The values of the effective potential at extrema are however gauge independent and can in principle be compared with the D3-brane potential at the maxima (2.1).
For $`T_{eff}0`$ the maximum of the potential tends to $`M=0`$. At $`T_{eff}=0`$ the potential is monotonically decreasing with $`M`$. In fact the D3-brane potential at zero temperature is precisely given by the tree level potential. This suggests that the field theory at zero temperature is unstable, a feature that, according to the AdS/CFT conjecture, should be shared by the dual supergravity solution. In a stability analysis of the extremal case has been presented. The authors found however that this geometry is stable against small perturbations. The resolution of this puzzle might go as follows. A D3-brane probe on the extremal hyperbolic black hole will be driven away from the horizon unless it is placed exactly at $`u_+`$, the unstable maximum of the potential. Since the length scale of the supergravity solution is $`\mathrm{}^4N`$, it seems that the extremal geometry is unstable against decreasing $`\mathrm{}`$ and therefore increasing its curvature. Therefore one should possibly not only consider variations in the metric but also allow for changes of the four-form potential that determines the five-form flux on the $`S^5`$.
3. Wilson Loops
We will analyze now the behavior of Wilson loops and heavy quark potentials of $`𝒩=4`$ Yang-Mills in $`S^3`$ and $`H^3`$ using the AdS/CFT correspondence. At large $`N`$ and large ’t-Hooft coupling, the expectation value of a Wilson loop $`C`$ is given in terms of the dual string theory by \[7,,8\]
$$W(C)e^{(S(C)S_0)},$$
where $`S(C)`$ is the Nambu-Goto action of the fundamental string whose world-sheet ends on the contour $`C`$ at the $`u\mathrm{}`$ boundary and minimizes the action. We will consider a contour describing the world-line of a $`q\overline{q}`$ pair of very heavy quarks. The action $`S(C)`$ is infinite due to the contribution of the world-sheet region close to the boundary. This divergence represents the self-energy of the non-dynamical quarks. In AdS language a single external quark is represented by a string extending straight along the radial coordinate into the AdS interior. It is therefore natural to regularize $`S(C)`$ by subtracting the action of two separated strings stretching down to $`u=0`$ in the case of empty AdS \[7,,8\], or to the horizon for black hole backgrounds \[9,,10\]. We have denoted this term by $`S_0`$ in (3.1).
We will restrict ourselves to the simplified case where $`q`$ and $`\overline{q}`$ have equal charge under the $`𝒩=4`$ R-symmetry group, $`SU(4)`$. This implies that the minimal area configuration will lie at a single point in the $`S^5`$ dimensions. We take $`q`$ and $`\overline{q}`$ to be static and using the symmetries of the sphere and hyperboloid we place them at $`(\theta ,\varphi ,\psi )=(0,0,0)`$ and $`(\theta ,0,0)`$ (for the sphere $`\theta (0,\pi )`$ while for the hyperboloid $`\theta (0,\mathrm{})`$).
The energy of the $`q\overline{q}`$ configuration is given by the vacuum expectation value of the Wilson loop divided by $`\beta `$. Following - we find<sup>5</sup> Recently deformations of AdS black holes with non-constant dilaton and the quark-antiquark potential in these backgrounds has been studied in .
$$E=\frac{\lambda ^{\frac{1}{2}}}{\pi z}\left[_1^{\mathrm{}}𝑑y\left(\sqrt{\frac{y^4+kz^2y^2\mu z^4}{y^4+kz^2y^21kz^2}}1\right)1+zu_+\right].$$
We have defined $`z=1/u_0`$ with $`u_0`$ being the minimal radial coordinate of the world-sheet ending on $`C`$. The mass of the black hole is related to the effective temperature defined in (2.1) by $`\mu =(\pi T_{eff})^41/4`$. As in the previous section, expression (3.1) describes both the cases of an $`S^3`$ and $`H^3`$ boundary by taking respectively $`k=1`$ or $`k=1`$. The angular distance between the quarks is given in terms of $`z`$ by
$$\theta =2z\sqrt{1+kz^2\mu z^4}_1^{\mathrm{}}\frac{dy}{\sqrt{\left(y^4+kz^2y^2\mu z^4\right)\left(y^4+kz^2y^21kz^2\right)}}.$$
The interaction potential between the pair of static quarks can be read directly from (3.1). If $`E0`$, $`V_{q\overline{q}}(\theta )=E`$. When $`E>0`$ the configuration with two separated world-sheets extending straight to the AdS interior is energetically favored. Then $`V_{q\overline{q}}=0`$ corresponding to screening of the charges \[9,,10\].
3.1. Quark-antiquark potentials on $`S^3`$.
The most interesting aspect of $`𝒩=4`$ Yang-Mills on $`S^3`$ is the existence of a phase transition at $`T=T_{crit}`$ . In the supergravity dual its counterpart is the Hawking-Page phase transition between black hole and pure AdS solutions to the Einstein’s equations . The study of quark-antiquark potentials should provide an additional piece of information about the physics involved in it.
For temperatures below $`T_{crit}`$ the supergravity dual to $`𝒩=4`$ Yang-Mills on $`S^3`$ is empty AdS with the euclidean time coordinate $`\tau `$ periodically identified with period $`\beta =1/T`$. The low temperature phase corresponds to set $`\mu =0`$, $`k=1`$ in (3.1) and (3.1), independently of $`T`$. Those expressions can be integrated to give<sup>6</sup> Quark-antiquark potentials in a Yang-Mills state associated to the uniform distribution of flat 3-branes on a disk have been studied recently in . When the $`q\overline{q}`$ pair is placed on an axes orthogonal to the disk, the resulting expressions coincide formally with (3.4). Since the Nambu-Goto action depends on products of metric elements, different metrics can produce formally equivalent expressions for the Wilson loops.
$$\begin{array}{cc}\hfill \theta =& \frac{2z}{\sqrt{(1+z^2)(2+z^2)}}\left(\mathrm{\Pi }(\kappa ^2,\kappa )K(\kappa )\right),\hfill \\ \hfill E=& \frac{\lambda ^{\frac{1}{2}}\sqrt{2+z^2}}{\pi z}\left(\kappa ^2K(\kappa )E(k)\right),\hfill \end{array}$$
where $`\kappa =1/\sqrt{2+z^2}`$, $`\kappa ^{}=\sqrt{1\kappa ^2}`$. The functions $`K(\kappa )`$, $`E(\kappa )`$ and $`\mathrm{\Pi }(\alpha ^2,\kappa )`$ denote the complete elliptic integrals of first, second and third kind <sup>7</sup> We follow the notation in .. For small $`z`$, which corresponds both to small inter-quark separations and to the large radius limit of $`S^3`$, we recover the flat limit result \[7,,8\]
$$V_{q\overline{q}}=\frac{\lambda ^{\frac{1}{2}}}{L\pi }\left(\mathrm{\Pi }(\frac{1}{2},\frac{1}{\sqrt{2}})K(\frac{1}{\sqrt{2}})\right)\left(K(\frac{1}{\sqrt{2}})2E(\frac{1}{\sqrt{2}})\right)=\frac{4\pi ^2\lambda ^{1/2}}{\mathrm{\Gamma }\left(\frac{1}{4}\right)^4L}.$$
Since we are considering 3-geometries of unit radius the inter-quark separation is just $`L=\theta `$.
Fig. 3: (a) Potential for a pair of static heavy quarks on $`S^3`$ on the large $`N`$, large ’t-Hooft coupling limit (we have set $`\lambda =1`$). (b) Normalized ratio between the strong and weak coupling limits of the potential.
The fact that the entropy of empty AdS is zero indicates that the $`N^2`$ degrees of freedom of the dual gauge theory are somehow “confined”. In it was argued that this is not dynamical confinement but just a kinematical effect: Gauss law on a compact space does not allow for net charges. In agreement with that, fig. 3 (a) shows that the strong coupling quark-antiquark potential has a Coulomb-like behavior. Kinematical confinement is also present at weak coupling since the $`𝒩=4`$ gauge theory is in a Coulomb phase at zero temperature.
We will now compare the weak and strong coupling limits of the potential more quantitatively. In the weak coupling and zero temperature limit, the potential between a pair of static charges on $`S^3`$ can be calculated explicitly. At small coupling we can neglect the non-linearity of Yang-Mills and just consider an abelian $`𝒩=4`$ gauge theory. The potential between a pair of static charges with equal R-charge is due to the interchange of gauge bosons and a linear combination of the six $`𝒩=4`$ scalars, i.e. that with R-charge aligned with the quarks R-charge
$$V_w=\frac{g_{YM}^2N}{4\pi ^2}\left((\pi \theta )\mathrm{cot}\theta +\frac{(\pi \theta )}{\mathrm{sin}\theta }\right).$$
The first term is associated to gauge boson interchange. It is obtained by solving the Laplace equation on $`S^3`$ with opposite sign delta sources at the positions of the quarks, and deriving from that the energy of the configuration. The second term in (3.1) is due to the scalar interchange. It is obtained by solving the Laplace equation with sources and a mass term since the $`𝒩=4`$ scalars couple to the curvature of the 3-sphere. Indeed we have seen in the previous section that the scalars are conformally coupled, i.e. $`m^2=1`$, as expected .
The comparison between the weak and strong coupling limits of the potential at $`T=0`$ shows important renormalization effects. As in the flat case , the dependence of the potential on the ’t-Hooft coupling changes from weak to strong coupling. On $`S^3`$, in addition, also its functional dependence on the inter-quark separation changes. The radius of the sphere introduce a new scale and even in the conformally coupled case nothing constrains the dependence of the potential on the angular separation of the quarks. This does not contradict the fact that empty AdS is an exact solution to the string equations of motion since (3.1) and (3.1) can receive $`\alpha ^{}`$ corrections due to quantum fluctuations of the string ending on the quarks world-line . In order to compare the different $`\theta `$ dependence of (3.1) and (3.1), we have plotted in fig. 3 (b) the quotient between the strong and weak coupling potentials normalized such that it tends to 1 at $`\theta 0`$, i.e. $`e=e_0V_{st}/V_w`$ with
$$e_0=\underset{\theta 0}{lim}\frac{V_w(\theta )}{V_{st}(\theta )}=\frac{\mathrm{\Gamma }(1/4)^4\lambda ^{1/2}}{2(2\pi )^3}.$$
The potential at strong coupling is proportional to $`\lambda ^{1/2}`$ instead of $`\lambda `$, which is the weak coupling result. It is interesting to notice that this reduction in the expected intensity of the potential at strong coupling decreases slightly as the angular separation of the quarks grows.
At $`TT_{crit}`$ the field theory undergoes a phase transition, detected because its supergravity dual description changes from empty AdS to an AdS Schwarzschild black hole background. The black hole solutions have an entropy $`SN^2`$. This implies that there are $`N^2`$ degrees of freedom contributing to the entropy of the dual gauge theory and suggests that the gauge theory is in a deconfined phase \[3,,13\]. Since low temperature confinement is just the statement of Gauss law on the sphere, we could think that what triggers the phase transition is screening. Let us analyze the situation at weak coupling. The $`q\overline{q}`$ potential at weak coupling is proportional to $`\frac{\mathrm{exp}(m_{el}L)}{L}`$. To lowest order we can estimate the electric screening mass to be $`m_{el}^2\lambda T^2`$ . If the screening length $`\frac{1}{m_{el}}`$ is much smaller than the size of the compact manifold the gauge theory will qualitatively behave as in flat space. Charges will be screened and we expect the entropy to scale like $`N^2`$. If however the screening length is bigger or of the order of the compact space size, the quark charges are not screened over maximal distances on the compact space. In this case physical states have to be color neutral and no quasi-free charges are possible. Thus we expect the entropy to scale like $`N^0`$. Therefore even at weak coupling there will be a phase transition in the large $`N`$ gauge theory at finite temperature on a compact space. If there is not other dominant mechanism, screening phenomena will induce a phase transition at $`T_c\frac{1}{\sqrt{\lambda }\pi }`$ for the case of a three sphere of unit radius where the maximal distance is $`\pi `$. Although at weak coupling this crude estimate gives a very high critical temperature, strong coupling effects could importantly lower it.
Fig. 4: Energy of the single string configuration (a) and quark-antiquark separation (b) as functions of the minimal radial distance that the string reaches in the black hole background. (H denotes the black hole horizon).
We derive again $`V_{q\overline{q}}`$ from (3.1) and (3.1). The high temperature phase of $`𝒩=4`$ on $`S^3`$ corresponds to $`\mu 2`$. The results parallel those obtained in the flat case . The single string configuration with boundary on the quarks positions exists only for inter-quark distances smaller than a certain threshold, $`L_m(T)`$. For each distance $`L<L_m(T)`$ there are two string configurations that extremize the Nambu-Goto action. One of them has bigger action than that of two separated strings extending straight to the black hole horizon and is therefore energetically disfavored. For separations $`L_s(T)<LL_m(T)`$, the action of the second single string configuration is also bigger than that of two separated strings (see fig. 4). Therefore $`V_{q\overline{q}}=0`$ for quark separations bigger than $`L_s(T)`$: the high temperature phase exhibits total screening.
Fig. 5: Quark-antiquark potential on $`S^3`$ for $`T>T_{crit}`$, $`T_{crit}`$ and $`T_{min}`$ ($`\lambda =1`$). In all cases the screening length is much smaller than $`\pi `$.
We observe in fig. 5 that the screening length is always much smaller than $`\pi `$. AdS Schwarzschild black hole solutions exist for $`TT_{min}`$ though they are only stable for temperatures bigger than $`T_{crit}>T_{min}`$. Black hole solutions with $`T_{crit}>TT_{min}`$ could correspond to meta-stable states in the dual gauge theory. For completeness, we have also plotted in fig. 5 the potential induced by the black hole solution at $`T_{min}`$. The screening length is still very small. Since the supergravity solutions we are treating have constant dilaton, the results obtained for the screening length between electric charges apply also to magnetic and dyonic charges. According to this result the phase transition is not triggered by screening of the electric or magnetic charges. If the degrees of freedom at strong coupling were electrically charged their charges would still be completely screened even at (and below) the critical temperature. Our results seem therefore to indicate that the relevant $`N^2`$ degrees of freedom at the phase transition do not carry electric charges. The other possible interpretation is that kinematical confinement of the charges is not the mechanism which triggers the phase transition.
3.2. Quark-antiquark potentials on $`H^3`$.
Fig. 6: Quark-antiquark potentials on $`H^3`$ at $`T=0`$, $`T_{AdS}`$ and $`T>T_{AdS}`$.
We turn now to analyze quark-antiquark potentials of $`𝒩=4`$ Yang-Mills on the 3-hyperboloid. In the large $`N`$ and large ’t-Hooft coupling limit the inter-quark potential is derived from (3.1) and (3.1) by setting $`k=1`$ and $`\mu 1/4`$; the results are plotted in fig. 6. As in the flat case, for temperatures bigger than zero there is a maximal quark separation after which the charges are completely screened. The surprising result is that even at $`T=0`$ the potential has a finite range!
For $`T>0`$ the dependence of $`V_{q\overline{q}}`$ and the inter-quark separation on the minimal radial distance of the single string configuration is as shown in fig. 4. In particular this is true for $`T=1/2\pi `$, when the background geometry is just empty AdS with a hyperbolic boundary. The situation at $`T=0`$ is however different. The action of the single string configuration is always smaller than that of two separated strings. For each inter-quark separation there is only one single string configuration. As the tip of the string gets closer to the black hole horizon, the inter-quark separation tends to the value
$$L_s=\sqrt{2}_0^{\mathrm{}}\frac{dx}{(x+1)\sqrt{x(x+2)}}=\frac{\pi }{\sqrt{2}}.$$
This relation is obtained by changing variables $`y^21=(1\frac{z^2}{2})x`$ in (3.1) with $`z<1`$ and $`\mu =0`$, and then taking the limit $`z1`$ <sup>8</sup> As for the low temperature phase on $`S^3`$, (3.1) and (3.1) with $`k=1`$ and $`\mu =1/4`$ can be integrated in terms of elliptic functions. The results coincide formally with those of when the $`q\overline{q}`$ pair lies in the plane of the disk (see footnote 6).. For separations bigger than $`L_s`$ the only possible configuration is that of two separated strings, i.e. the charges are screened. Of course since we are on a hyperbolic space, the Coulomb potential will also show exponential falloff at zero temperature in the weak coupling limit, $`V\mathrm{coth}\theta `$. The result at strong coupling deviates however from this weak coupling behavior since we find a maximum interaction distance. This indicates an additional screening of the charges at strong coupling<sup>9</sup> Screening at $`T=0`$ has been also encountered in . The geometry describing that case coincides with the extremal limit of a rotating D3-brane solution . It is interesting to notice that the hyperbolic black hole geometry at $`T=0`$ is also extremal.. The black hole background at $`T=0`$ has zero energy but non-zero entropy $`SN^2`$ . The presence of these $`N^2`$ states contributing to the entropy but not to the energy could be related with the presence of total screening.
Acknowledgments
The work of K.L. was supported by an FWF project under number P13125-TPH. The work of E.L. was supported by an FWF project under number P13126-TPH.
References
relax J. Maldacena, “The large N limit of superconformal field theories and supergravity”, Adv. Theor. Math. Phys. 2 (1998) 231, hep-th/9711200. relax Gary T. Horowitz and Simon F. Ross, “Possible Resolution of Black Hole Singularities from Large N Gauge Theory”, JHEP 9807 (1998) 014, hep-th/9803085. relax E. Witten, “Anti-de Sitter Space, Thermal Phase Transition, And Confinement In Gauge Theories”, Adv.Theor.Math.Phys. 2 (1998) 505, hep-th/9803131. relax O. Aharony, S. S. Gubser, J. Maldacena, H. Ooguri and Y. Oz, “Large N Field Theories, String Theory and Gravity”, hep-th/9905111. relax A. A. Tseytlin and S. Yankielowicz, Nucl.Phys. B541 (1999) 145, hep-th/9809032. relax E. Kiritsis and T.R. Taylor, “Thermodynamics of D-brane Probes”, hep-th/9906048; E. Kiritsis, “Supergravity, D-brane Probes and thermal super Yang-Mills: a comparison, hep-th/9906206. relax J. Maldacena, “Wilson loops in large N field theories”, Phys.Rev.Lett. 80 (1998) 4859, hep-th/9803002. relax S.-J. Rey and J. Yee, “Macroscopic Strings as Heavy Quarks of Large N Gauge Theory and Anti-de Sitter Supergravity”, hep-th/9803001. relax A. Brandhuber, N. Itzhaki, J. Sonnenschein and S. Yankielowicz, “Wilson Loops in the Large N Limit at Finite Temperature”, Phys.Lett. B434 (1998) 36, hep-th/9803137. relax S.-J. Rey, S. Theisen and J.-T. Yee, “Wilson-Polyakov Loop at Finite Temperature in Large N Gauge Theory and Anti-de Sitter Supergravity”, Nucl.Phys. B527 (1998) 171, hep-th/9803135. relax A. Brandhuber and K. Sfetsos, “Wilson loops from multicentre and rotating branes, mass gaps and phase structure in gauge theories”, hep-th/9906201. relax R. Emparan, “AdS Membranes Wrapped on Surfaces of Arbitrary Genus”, Phys.Lett. B432 (1998) 74, hep-th/9804031. relax E. Witten, “Anti De Sitter Space And Holography”, Adv.Theor.Math.Phys. 2 (1998) 253, hep-th/9802150. relax S. W. Hawking and D. Page, “Thermodynamics of Black Holes in Anti-de Sitter Space”, Comm.Math.Phys. 87 (1983) 577. relax R. B. Mann, “Topological Black Holes – Outside Looking In”, gr-qc/9709039. relax D. Birmingham, “Topological Black Holes in Anti-de Sitter Space”, Class.Quant.Grav. 16 (1999) 1197, hep-th/9808032. relax R. Emparan, C. V. Johnson and R. C. Myers, “Surface Terms as Counterterms in the AdS/CFT Correspondence”, hep-th/9903238. relax R. Emparan, “AdS/CFT Duals of Topological Black Holes and the Entropy of Zero-Energy States”, JHEP 9906 (1999) 036, hep-th/9906040. relax C. P. Burgess, N. R. Constable and R. C. Myers, “The Free Energy of N=4 Super-Yang-Mills Theory and the AdS/CFT Correspondence”, hep-th/9907188. relax P.G.O. Freund and M. A. Rubin, “Dynamics of Dimensional Reduction”, Phys.Lett.97B: (1980) 233. relax J. I. Kapusta, “Finite Temperature Field Theory”, Cambridge Monographs on Mathematical Physics (1989). relax J. Polchinski, “String Theory”, vol. 2, Cambridge University Press, 1998. relax J. Maldacena, “Probing Near Extremal Black Holes with D-branes”, hep-th/9705053; “Branes Probing Black Holes”, Nucl.Phys.Proc.Suppl. 68 (1998) 17, hep-th/9709099. relax L. Susskind and E. Witten, “The Holographic Bound in Anti-de Sitter Space”, hep-th/9805114. relax N. D. Birrell and P. C. W. Davies, “Quantum Fields in Curved Space”, Cambridge University Press, 1982. relax M. J. Duff, “Twenty Years of the Weyl Anomaly”, Class. Quant. Grav. 11 (1994) 1387, hep-th/9308075 and references therein. relax G. T. Horowitz and R. C. Myers, “The AdS/CFT Correspondence and a New Positive Energy Conjecture for General Relativity”, Phys. Rev. D59 (1999) 2605, hep-th/9808079. relax S. Nojiri and S. D. Odintsov, “Running Gauge Coupling and Quark-Antiquark Potential in Non-SUSY Gauge Theory at Finite Temperature from IIB SG/CFT correspondence”, hep-th/9906216. relax P. F. Byrd and M. D. Friedman, “Handbook of Elliptic Integrals for Engineers and Scientists”, second edition, 1971 Springer-Verlag. relax T. Banks and M. B. Green, “Non-perturbative Effects in AdS5\*S5 String Theory and d=4 SUSY Yang-Mills”, JHEP 9805 (1998) 002, hep-th/9804170. relax S. Forste, D. Ghoshal and S. Theisen, “Stringy Corrections to the Wilson Loop in N=4 Super Yang-Mills Theory”, hep-th/9903042. relax P. Kraus, F. Larsen and S. P. Trivedi, “The Coulomb Branch of Gauge Theory from Rotating Branes”, JHEP 9903 (1999) 003, hep-th/9811120. relax K. Sfetsos, “Branes for Higgs phases and exact conformal field theories”, JHEP 9901 (1999) 015, hep-th/9811167.
|
no-problem/9908/nucl-th9908005.html
|
ar5iv
|
text
|
# Describing Nuclear Matter with Effective Field Theories
## Abstract
An accurate description of nuclear matter starting from free-space nuclear forces has been an elusive goal. The complexity of the system makes approximations inevitable, so the challenge is to find a consistent truncation scheme with controlled errors. The virtues of an effective field theory approach to this problem are discussed.
Nuclear forces have been studied in depth over the past fifty years, leading to excellent phenomenological descriptions. A recent resurgence of interest in this field has been fueled by the promise of systematic results from an effective field theory (EFT) analysis . EFT techniques adapted to many-body systems may give new insight into the properties of nuclei and their connection to QCD.
The EFT lagrangian consists of long-range interactions constrained by chiral symmetry and the most general short-range interactions consistent with QCD symmetries. The coefficients of these short-range terms may eventually be derived from QCD, but at present must be fit by matching calculated and experimental observables in a momentum expansion. This effective lagrangian can then be used to systematically predict other observables, including inelastic processes.
The predictability of an EFT relies on an organizational scheme, called power counting, to determine the importance of any contribution to a calculation. This can be illustrated by using the familiar one-boson-exchange representation of the NN force as a model of the true underlying physics. One-pion exchange (OPE) is taken to be the long-range physics, as its contribution to the EFT is dictated by chiral symmetry. Exchanges of the heavier mesons will be considered short-range physics, which must be included generically in the EFT. For example, for center-of-mass momenta below the sigma mass, $`p<m_\sigma `$, the heavy-meson propagators are well approximated by momentum-dependent contact interactions,
$$\frac{1}{p^2m_\sigma ^2}=C_0+C_2p^2+C_4p^4+\mathrm{},$$
(1)
with $`C_2=C_0/m_\sigma ^2`$ and so on. Each term has an associated power of $`p/m_\sigma `$, which carries over to observables. A consistent truncation based on counting these powers leads to observables with well-defined errors . Here $`m_\sigma `$ plays the role of the EFT breakdown scale $`\mathrm{\Lambda }`$. At momenta of order $`\mathrm{\Lambda }`$, the short-distance structure is resolved, all higher-order corrections become comparable in magnitude, and consequently the EFT fails.
In the early years of research into nuclear forces, Bethe showed that low-energy scattering between two nucleons could be expressed in a form now known as the effective range expansion. For example, in the $`{}_{}{}^{1}S_{0}^{}`$ channel,
$$p\mathrm{cot}\delta =\frac{1}{a_s}+\frac{1}{2}r_ep^2+\mathrm{}.$$
(2)
A key observation is that the coefficients $`a_s`$, $`r_e`$, $`\mathrm{}`$, of this momentum expansion are independent of the details of the potential used to fit low-energy data. The scale associated with the parameters is the point at which Eq. (2) breaks down, around $`p1/r_em_\pi `$. This expansion is suggestive of an EFT with a breakdown scale around $`\mathrm{\Lambda }m_\pi `$. So we expect to reproduce Eq. (2) if we treat all interactions (including OPE) as short-ranged.
Due to the existence of nuclear bound states (such as the deuteron) near threshold, an EFT analysis for multiple nucleons requires a nonperturbative treatment. One formulation is based on the Lippmann-Schwinger equation for the $`T`$-matrix , $`\widehat{T}=\widehat{V}+\widehat{V}\widehat{G}_0\widehat{T}`$ . The most general effective potential with no long-range physics takes the form
$$\widehat{V}_{\mathrm{EFT}}=C_0+C_2\widehat{p}^2+𝒪(\widehat{p}^4/\mathrm{\Lambda }^4).$$
(3)
If the EFT phase shifts are calculated and compared with actual data, we find that the constants indeed form a hierarchy analogous to Eq. (1), but with the relevant scale being set by the pion mass $`m_\pi `$:
$$C_0\frac{4\pi }{Mm_\pi },C_2p^2\frac{4\pi }{Mm_\pi }\frac{p^2}{m_\pi ^2},\mathrm{}.$$
(4)
Enforcing this power counting order-by-order in $`p^2`$ reproduces the effective range expansion Eq. (2) and generates the solid lines in the error plot of Fig. 1. The error is dominated by the first omitted term and so behaves like a power of $`p^2`$. As more contact interactions are included in the effective potential Eq. (3), the slope of the error curves gets steeper. The point at which the linear parts of the curves converge indicates the breakdown scale $`\mathrm{\Lambda }`$, which is indeed of order $`m_\pi `$ (actually $`m_\pi /2`$).
To push beyond this scale, pions must be explicitly included in the effective lagrangian as long-range physics. There are alternative power counting schemes proposed in the literature . In Fig. 1, results are shown for a scheme in which one-pion exchange enters at leading order and pion effects are treated nonperturbatively. The range of validity of the EFT is extended to about $`\mathrm{\Lambda }=300`$ MeV in that case, as seen by the dashed lines.
A complete calculation at next-to-leading order requires irreducible two-pion exchange. Preliminary results suggest that with a careful removal of long-distance contamination from coefficients by using a modified effective range expansion , the EFT range of validity may extend as high as $`\mathrm{\Lambda }m_\rho `$ . Even then, the accuracy of the EFT results for NN scattering is not yet competitive with that of the Bonn potential . Nevertheless, the error plots manifest a systematic expansion that allows realistic error estimates for scattering and other observables, which are not available in conventional approaches.
How does this expansion and breakdown scale carry over to nuclear matter? The complications of the many-body problem can obscure the connection, but we can use a perturbative matching calculation to understand the general correspondence. To proceed, model data can be generated from an exactly solvable potential with weak coupling $`\lambda `$. The EFT contact interactions is then fit by this data up to a given order in $`\lambda `$; for example,
$$\widehat{T}_{\mathrm{EFT}}=\widehat{V}_{\mathrm{EFT}}+\widehat{V}_{\mathrm{EFT}}\widehat{G}_0\widehat{V}_{\mathrm{EFT}}+𝒪(\lambda ^3).$$
(5)
The regularization of divergences in the second term of Eq. (5) necessitates a renormalization of the first term to ensure a cutoff independent match. Solving for the exact in-medium $`T`$-matrix $`\widehat{\mathrm{\Gamma }}`$ perturbatively and comparing with the EFT result using the effective potential in Eq. (3) verifies that the match remains regulator independent to the same order as in free space,
$$\widehat{\mathrm{\Gamma }}=\widehat{\mathrm{\Gamma }}_{\mathrm{EFT}}+𝒪(\lambda ^3)+𝒪(k_F^4/\mathrm{\Lambda }^4).$$
(6)
The truncation error from the EFT expansion in nuclear matter is a power of the Fermi momentum over the free-space breakdown scale, $`k_F/\mathrm{\Lambda }`$ . This is very encouraging, since phenomenologically successful mean-field descriptions have a related expansion with $`\mathrm{\Lambda }600`$ MeV . Thus if the free-space breakdown scale is indeed as large as $`\mathrm{\Lambda }m_\rho `$, a useful EFT expansion of nuclear matter is likely.
A nonperturbative EFT calculation of nuclear matter introduces many complications compared to the free-space EFT. The analogue of scattering between two nucleons involves particles propagating off the energy-shell ($`Ep^2/M`$) and the in-medium $`T`$-matrix not only depends on the relative momentum, but also the total momentum. Also, enforcing the Pauli exclusion principle greatly complicates the momentum integrals. However, a sys-tematic EFT error analysis makes the calculation worth the effort, and the EFT approach can also provide new perspectives on old issues and practices. Some examples: i) Nuclear saturation with velocity-dependent potentials instead of a hard core; ii) Renormalization scheme independence of observables, such as the binding energy, requires the sum of ladder diagrams, providing a new justification for this approximation; iii) A self-consistent ladder calculation requires explicit three-body forces. This last feature might account for discrepancies in nuclear matter between phase-equivalent two-body potentials .
To see the necessity of three-body forces, consider three-to-three scattering with only two-body interactions, as depicted by the second Feynman diagram in Fig. 2. This could arise from the first diagram, which is a contribution to the energy of the nuclear matter ground state, by expanding the short-range interactions in contact terms and opening the hole lines. Since the EFT is only accurate at low momenta, the divergent loop integrals should be cut off at some scale $`\mathrm{\Lambda }_c`$. The intermediate states above $`\mathrm{\Lambda }_c`$ are highly virtual, and so by the uncertainty principle are well represented by a regularized series of three-body contact interactions. These interactions are needed to guarantee regularization independence of the final result and to account for contributions from the suppressed physical degrees of freedom.
Higher-body interactions are cut off in a similar fashion, implying an infinite number of constants and many-body forces would be needed to describe nuclear matter. The Pauli exclusion principle helps the situation by limiting the number of contact interactions with no derivatives to four nucleons or fewer, and the EFT power counting implies that other higher-body interactions between nucleons are suppressed. However, the nature of the power counting is still under investigation, and it is possible that three- and four-body contact terms are needed even at leading order in a nuclear matter calculation.
The inevitability of many-body forces in the EFT analysis highlights the fact that the potential is not an observable. There is no such thing as a “best” two-body potential and differences in two-body off-shell behavior in nuclear matter may be compensated by many-body counterterms . In the case of three-nucleon scattering, accounting for the three-body contact interaction leads to a compelling explanation of the Phillips line . An interesting possibility is that an extension of this analysis to nuclear matter could explain the Coester line .
While the application of effective field theory to nuclear matter is in its infancy, systematic EFT calculations could help settle long-standing issues, justify traditional expansions, or even offer new alternatives. In the long run, EFT techniques lay a foundation for ultimately connecting the physics of nuclei to QCD.
|
no-problem/9908/cond-mat9908328.html
|
ar5iv
|
text
|
# Reversible electrowetting and trapping of charge: model and experiments
## Introduction
It is possible to reduce the contact angle of a droplet on a surface by applying an electric field between the conducting liquid and a counter electrode underneath the liquid as shown in Fig. 1. This so-called electrowetting effect was observed first by Minnema in 1980 using an insulator between liquid and counter electrode and by Beni in 1981 with the liquid directly on the counter electrode. The electric field results in a distribution of charge that changes the free energy of the droplet, causing the droplet to spread and wet the surface. In systems with the liquid in direct contact with the solid electrode, the potential drops across a diffuse ionic double layer at the interface. In systems with an insulating layer of several micrometers thickness between the solid electrode and the liquid , the main voltage drop appears across the insulating layer.
In this paper, we use an insulating layer between the counter electrode and the aqueous solution to enhance the electrowetting force , achieving reversible wetting by a suitable top coating. Previously, limits have been observed for the voltage-induced reduction of contact angle: at high electric fields, the contact angle saturates . We consider the possibility that trapping of charge in or on the insulating layer affects the contact angle. We define charge to be trapped when the charge is bonded more strongly to the insulating layer than to the liquid. First, we derive the theory of electrowetting from the principle of virtual displacement. This provides a flexible method to extend Young’s equation to include the influence of an arbitrary charge distribution. We consider the case that a sheet of trapped charge is present in or on the insulating layer. Next, we present a measurement of the contact angle as a function of applied voltage and we extract the potential resulting from the trapped charge as a function of applied voltage. Finally, we suggest possible microscopic origins for trapping of charge.
## Electrowetting model
### Virtual displacement, no trapped charge.
A droplet spreads until it has reached a minimum in free energy, determined by cohesion forces in the liquid and adhesion between the liquid and the surface. In general, the energy required to create an interface is given by $`\gamma `$, the surface tension \[N/m\] or surface free energy \[J/m<sup>2</sup>\]. In case of an applied potential, a change in the electric charge distribution at the liquid/solid interface changes the free energy. We define our thermodynamic system as the droplet, the insulating layer, the metal electrode and the voltage source. Throughout the entire derivation, we assume that the system is in equilibrium at constant potential $`V`$. We focus on the change in free energy due to an infinitesimal increase in base area of the droplet on the solid surface, surrounded by vapor. When a potential $`V`$ is applied, a charge density $`\sigma _L`$ builds up in the liquid phase and induces an image charge density $`\sigma _M`$ on the metal electrode. Figure 2(a) shows the edge of a droplet and its virtual displacement. An infinitesimal increase of the base area $`\text{d}A`$ results in a contribution to the free energy from the surface energies as well as an energy contribution due to the additional charge density $`\text{d}\sigma _L`$ in the liquid and its image charge density $`\text{d}\sigma _M`$ on the metal electrode. The voltage source performs the work, $`\text{d}W_B`$. The free energy ($`F`$) of the system can be written in differential form:
$$\text{d}F=\gamma _{SL}\text{d}A\gamma _{SV}\text{d}A+\gamma _{LV}\text{d}A\mathrm{cos}\theta +\text{d}U\text{d}W_B,$$
(1)
where $`U`$ is the energy required to create the electric field between the liquid and the counter electrode. The parameters $`\gamma _{SL}`$, $`\gamma _{SV}`$ and $`\gamma _{LV}`$ are the free energies of the solid/liquid, solid/vapor and liquid/vapor interface respectively for the situation in the *absence* of an electric field. The contact angle, $`\theta `$, is the angle between the liquid/vapor interface and the solid/liquid interface at the contact line (see Fig. 1. Mechanisms for energy dissipation, which may cause contact-angle hysteresis, are not taken into account.
Let us first consider the situation in the absence of an externally applied voltage, so $`\text{d}U=\text{d}W_B=0`$. When $`\text{d}F/\text{d}A=0`$, we find the minimum in free energy, relating the surface energies to the contact angle. This equation was obtained by Young in 1805:
$$\gamma _{LV}\mathrm{cos}\theta =\gamma _{SV}\gamma _{SL}.$$
(2)
For a non-zero potential, we have to include the energy of the charge distribution. In Fig. 2(a), the droplet with charge density $`\sigma _L`$ is at constant voltage $`V`$, while the metal electrode with charge density $`\sigma _M=\sigma _L`$ is at ground potential. The electrostatic energy per unit area below the liquid is given by:
$$\frac{U}{A}=\underset{0}{\overset{d}{}}\frac{1}{2}\stackrel{}{E}\stackrel{}{D}\text{d}z,$$
(3)
where $`z`$ is the coordinate perpendicular to the surface, $`d`$ the thickness of the insulating layer, $`\stackrel{}{E}`$ the electric field and $`\stackrel{}{D}`$ the charge displacement, with $`\stackrel{}{D}=\epsilon _0\epsilon _r\stackrel{}{E}`$ . The increase of free energy due to the charge distribution in the liquid, upon an infinitesimal increase of droplet base can be written as:
$$\frac{\text{d}U}{\text{d}A}=\frac{1}{2}dED=\frac{1}{2}d\frac{V}{d}\sigma _L=\frac{1}{2}V\sigma _L.$$
(4)
The electric field originating from the liquid/vapor boundary of the droplet (the so-called fringing or stray field) makes a constant contribution to the free energy: this contribution remains unaltered when the contact line is displaced by $`\text{d}A`$. Therefore, the stray fields do not contribute to $`\text{d}U`$. The voltage source performs the work to redistribute the charge; per unit area the work is given by :
$$\frac{\text{d}W_B}{\text{d}A}=V\sigma _L.$$
(5)
Calculating the minimum of free energy, we get Young’s equation with an additional electrowetting term $`\gamma _{EW}`$, the electrowetting force per unit length due to the applied potential:
$$\gamma _{LV}\mathrm{cos}\theta =\gamma _{SV}\gamma _{SL}+\gamma _{EW},$$
(6)
with the electrowetting force:
$$\gamma _{EW}=\frac{1}{2}\frac{d}{\epsilon _0\epsilon _r}\sigma _L^2,$$
(7)
where we used $`\sigma _L=\epsilon _0\epsilon _rV/d`$ (Gauss’ law), with $`\epsilon _r`$ the dielectric constant of the insulating layer and $`\epsilon _0`$ the permittivity of vacuum. We can rewrite Eqs.(6) and (7) to get the well-known relation between $`\theta `$ and $`V`$ for electrowetting :
$$\gamma _{LV}\left[\mathrm{cos}\theta (V)\mathrm{cos}\theta _0\right]=\frac{1}{2}\frac{\epsilon _0\epsilon _r}{d}V^2,$$
(8)
where $`\theta _0`$ is the contact angle at zero volt.
### Influence of charge trapping.
When we apply a potential difference between the liquid and the metal electrode, electric forces work on the ions in the liquid and pull them toward the insulating layer. There is a possibility that charge becomes trapped in or on the insulating layer when the interaction of the ions with the solid is stronger than with the liquid. In the three-phase region, ions are trapped when the de-trapping time is large compared to the typical vibration times of the contact line. As a result of excitations of the droplet, e.g. thermal, mechanical or voltage-induced vibrations, a density of trapped charge arises on both sides of the contact line.
As yet we do not specify the precise nature of the trapped charge (e.g. electronic or ionic); we simply assume that a layer of charge with constant surface charge density $`\sigma _T`$ is trapped in the insulating layer at distance $`d_2`$ from the top of the insulator as shown in Fig. 2(b). To determine the change in electrostatic energy due to the infinitesimal base area increase \[$`\text{d}U`$ in Eq.(1)\], we have to take into account the electrostatic contribution below the liquid $`\text{d}U_L`$, as well as the contribution below the vapor phase, $`\text{d}U_V`$:
$$\text{d}U=\text{d}U_L\text{d}U_V.$$
(9)
The sign difference is due to the fact that the virtual displacement $`\text{d}A`$ increases the solid/liquid interface while the solid/vapor interface is decreased. We assume that the trapped charge is distributed uniformly at constant depth, extending sideways to the left of the contact line to a length scale of at least the insulator thickness. Then, the charge density at the liquid/vapor interface due to fringing fields at the edge of the droplet is unaltered by the virtual displacement. Therefore, we omit the electrostatic energy of the fringing field in Eq.(9).
The potential as a function of the depth in the insulator is sketched in Fig. 3(a). The solid line shows the potential in the absence of trapped charge. For the case of charge trapping, the potential beneath the liquid phase is indicated by the long dashed line while the short dashed line shows the potential beneath the vapor phase. The vertical line indicates the depth where the trapped charge is situated. Figure 3(b) shows a plot of the electrostatic fields, with $`E=V`$. It is clear that trapping of charge lowers the electric field at the liquid/solid interface and should consequently reduce the electrowetting force. The charge density of the trapped charge, $`\sigma _T`$, is at potential $`V_T^L`$ below the liquid and at $`V_T`$ below the vapor phase. On the metal electrode below the liquid, the charge density is $`\sigma _M^L=(\sigma _L+\sigma _T)`$. The charge density on the metal electrode below the vapor phase is $`\sigma _M^V=\sigma _T`$. Using the general expression for the energy density \[Eq.(3)\], we find the electrostatic energy density below the liquid phase:
$`{\displaystyle \frac{\text{d}U_L}{\text{d}A}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}d_1E_1D_1+{\displaystyle \frac{1}{2}}d_2E_2D_2`$ (10)
$`=`$ $`{\displaystyle \frac{1}{2}}d_1{\displaystyle \frac{V_T^L}{d_1}}(\sigma _T+\sigma _L)+{\displaystyle \frac{1}{2}}d_2{\displaystyle \frac{VV_T^L}{d_2}}\sigma _L`$
$`=`$ $`{\displaystyle \frac{1}{2}}V_T^L\sigma _T+{\displaystyle \frac{1}{2}}V\sigma _L.`$
The energy to create the charge distribution below the vapor phase equals:
$$\frac{\text{d}U_V}{\text{d}A}=\frac{1}{2}d_1\frac{V_T}{d_1}\sigma _T=\frac{1}{2}V_T\sigma _T.$$
(11)
The work performed by the voltage source per unit area is given by Eq.(5). Using Gauss’ law, we find the following relationships between the charge densities and the potentials:
$`\sigma _T`$ $`=`$ $`{\displaystyle \frac{\epsilon _0\epsilon _rV_T}{d_1}}`$ (12)
$`\sigma _L`$ $`=`$ $`{\displaystyle \frac{\epsilon _0\epsilon _r(VV_T)}{d}}`$ (13)
$`V_T^L`$ $`=`$ $`V_T+{\displaystyle \frac{d_1}{d}}(VV_T).`$ (14)
Using Eq.(1), Eq.(5), Eqs.(9)–(14) and $`\text{d}F/\text{d}A=0`$, we recover Eqs.(6) and (7), the modified Young equation. With Eq.(13), we find the following relation for the contact angle modulation in the presence of trapped charge:
$$\gamma _{LV}\left[\mathrm{cos}\theta (V)\mathrm{cos}\theta _0\right]=\frac{1}{2}\frac{\epsilon _0\epsilon _r}{d}(VV_T)^2.$$
(15)
The electrowetting force is proportional to the square of the applied voltage minus the voltage due to charge trapping. This causes a reduction of the electrowetting force.
## Results
We used a system as shown in Fig. 1. On a silicon substrate, a conducting aluminum layer (100 nm) and an insulating layer of parylene-N (10 $`\mu `$m, $`\epsilon _r=2.65`$, chemical vapor deposited as in Ref. ) were applied. Subsequently a thin hydrophobic AF1600 top coating was deposited by spincoating a 0.1 wt% solution of AF1600 (DuPont) in FC726 (3M) at 1000 r.p.m. during 30 s, resulting in a layer of approximately 30 nm thickness. In order to abtain a low contact-angle hysteresis, we impregnated the coating with silicon oil. The sample was left in silicon oil for several hours. Before the experiment the excess of oil was removed; this is possible because the contact angle of silicon oil on AF1600 is about 40 . The in-liquid electrode was a platinum wire. The droplet consisted of 10 $`\mu `$l of an aqueous solution, with 1.0 M, 0.1 M, 0.01 M KCl or 0.1 M K<sub>2</sub>SO<sub>4</sub>.
The capacitance was measured between the platinum electrode and the metal counter electrode as a function of applied voltage, using a 700 Hz, 5 V ac-signal which was superimposed on the dc-voltage. The capacitance gives a measure of the contact area between liquid and surface and is shown in Fig. 4(a). The potential of the liquid was increased to 500 V and subsequently decreased to 0 V; the opposite voltage polarity was measured on a different spot to avoid possible interference with previously trapped charge. The time interval between measurement points (1 s) was more than an order of magnitude longer than the time required for spreading of the droplet (experimentally verified to be about 20 ms ). We find that the droplet base increases by nearly a factor of three due to the applied voltage. The droplet recovers its original shape upon removal of the electric potential. Measurements with the solutions of different molarities and ion types resulted in identical curves.
The contact angle $`\theta `$ was derived from the measured capacitance, using the known dielectric constant and thickness of the insulating layer, the droplet volume and the droplet shape. This electrical measurement method gives the contact angle with an accuracy of 2; details are described in Ref. . Figure 4(b) shows the contact angle, derived from the capacitance measurement along with the theoretical curve according to Eq.(8). The contact-angle hysteresis is less than 2. To our knowledge, such a high degree of reversibility has not yet been reported in electrowetting experiments. We attribute the low value of contact-angle hysteresis to the penetration of the oil into nano-pores of the amorphous fluoropolymer layer , which reduces the already very low surface heterogeneity of the top coating (for water on non-impregnated AF1600, the contact-angle hysteresis is about 7). At zero volt, the value of the contact angle on the impregnated surface agrees well with the advancing contact angle on the non-impregnated surface. This indicates that the fluoropolymer determines the surface energy rather than the silicon oil.
We can distinguish two regions in the plot. In the region $`240<V<240`$ V, the measured contact angle is consistent with the theoretical contact angle of Eq.(8) within 2. According to our model \[Eq.(15)\], this means that at low voltages the ionic charge remains in the liquid without being trapped on or in the insulating layer. Within this voltage range we demonstrated reproducible droplet spreading for more than 10<sup>5</sup> times. At higher voltages, we notice contact-angle saturation at about 60and differences between the advancing and receding curves. To investigate the possibility that the saturation of contact angle is caused by trapping of charge, we applied a voltage higher than 240 V to the droplet and while maintaining the applied voltage, adsorbed the liquid into a tissue. Afterward, when blowing humid air across the sample, we observed a condensation pattern at the position and in the shape of the droplet base. During subsequent electrowetting experiments in this area, we detected a nonzero electrostatic potential, having the sign of the previously applied potential, which points at the presence of charge. Therefore, we attribute the condensation pattern to the preferential deposition of polar water molecules on the charged area of the surface. Finally, we noticed a decrease of the electrostatic surface potential after a large grounded droplet was placed on the charged coating; no decrease of the potential was observed when the large droplet was electrically floating. These experiments prove that contact-angle saturation is accompanied by charging of the insulating coating. The charge can be removed by an electrical shortcut between the metal electrode and the surface of the insulator.
We have analyzed the data of Fig. 4 with our model for voltage-induced wetting in the presence of trapped charge \[Eq.(15)\]. We ascribe the difference between the measured contact angle and the contact angle of the old model \[Eq.(8)\] to the voltage of trapped charge ($`V_T`$). Figure 5(a) shows the resulting plot for $`V_T`$. We notice a threshold voltage $`V_{th}`$ of $`240\pm 10`$ V. For $`|V|<V_{th}`$, the voltage of trapped charge equals zero and the electrowetting force $`\gamma _{EW}V^2`$. For $`|V|>V_{th}`$, $`V_TVV_{th}`$ and $`\gamma _{EW}(VV_T)^2`$. This means that above the threshold voltage, almost all charge added by the voltage source gets trapped in or on the insulating layer. Figure 4(b) shows the charge density in the liquid phase $`\sigma _L`$. For voltages below the threshold, $`\sigma _L`$ increases linearly with increasing applied voltage. Beyond the threshold voltage, $`\sigma _L`$ remains approximately constant, in line with the saturation of the contact angle.
## Discussion and Conclusions
The data of the previous section show a threshold-like saturation behavior for the electrowetting force and for the charge density in the liquid. The voltage of trapped charge shows a linear increase beyond the threshold voltage. The curves are symmetric for positive and negative potential, and independent of the ion type, ion valence and ion molarity that we have tested. Furthermore, the advancing curve is consistent with the receding curve, indicating that the trapped charge is released upon lowering of the applied voltage.
Let us now consider possible microscopic origins for trapping of charge. We defined trapped charge as charge which has a stronger interaction with the insulating layer than with the liquid. Clearly, the underlying charge bonding mechanism cannot have a chemical nature, as an expected dependence on voltage polarity, ion type, valence and molarity was not observed. Charge trapping could occur due to the attractive electrostatic force between ions in the liquid and the metal counter electrode. When the electrostatic force on the ion exceeds the force between the liquid and the ion, it moves toward the insulating layer and remains in or at the insulating layer . The ion might exchange some of its hydration shell for a bond with the surface. Although this model predicts a threshold-like trapping behavior, a dependence on the valence of the ions is expected, in disagreement with our experimental results.
At the threshold electric field, the electrowetting force is of the same order of magnitude as or larger than the surface tensions in our system (for $`V=240`$ V, $`\gamma _{EW}=68`$ mN/m). Therefore, we propose that it is possible that instabilities in the liquid/solid interface or the liquid/vapor interface occur. Small charged droplets or molecular clusters could move into nano-pores of the insulating layer and become trapped. In this concept, no dependence on molarity, ion type, valence of the ions or polarity of the applied field is expected. While this line of thought seems in agreement with the behavior of $`V_T`$ (a threshold and subsequently a slope close to one), further research is needed to determine the microscopic mechanisms of charge trapping.
In conclusion, the principle of virtual displacement provides a transparent method to calculate the influence of an arbitrary charge distribution on the contact angle. The virtual change of electric energy is calculated by integrating the energy density of the electrostatic field. Using this method, we derived a model for electrowetting that accounts for a reduction of the electrowetting force by the assumption that charge is trapped in or on the insulating layer \[Eq.(15)\].
We demonstrated reversible electrowetting using an aqueous droplet on a sample with a 10 $`\mu `$m thick parylene insulating layer and a highly fluorinated hydrophobic AF1600 top coating which was impregnated with silicon oil. The measured contact-angle hysteresis was below 2. For voltages between $`240`$ V and $`240`$ V the charge remains in the liquid and is not trapped in or on the insulating layer. At higher voltages, charge gets trapped with a threshold-like behavior, limiting the charge density that can be induced in the liquid. We observed that charge remains at areas where the droplet has receded in a high-voltage state and that discharging of the surface is possible with a zero-voltage droplet. For all solutions tested, the absolute value of the threshold voltage is independent of the polarity of the applied voltage, the ion type, ion molarity and ion valence (Cl<sup>-</sup> vs. SO$`{}_{}{}^{2}{}_{4}{}^{}`$).
From an experimental standpoint, the distribution of the trapped charge should be measured quantitatively with for instance a scanning Kelvin probe. This will clarify the dependence of the trapped charge density on the applied voltage and on the distance with respect to the contact line. The mechanisms of de-trapping of charge are interesting and need to be studied in more detail. The dependence of the threshold electric field on the insulator thickness, measured for different salts, solvents and top coatings may provide information on the mechanisms that cause the trapping of charge. Eventually, when trapping of charge in the insulator can be avoided, it may become possible to reversibly reach complete wetting of a surface by an applied electric potential.
### Acknowledgement
We thank W. Welters and A. Kemmeren for coating preparation and for valuable discussions.
## Figures
|
no-problem/9908/astro-ph9908273.html
|
ar5iv
|
text
|
# Can the Jet steepen the light curves of GRB afterglows?
## 1 Introduction
The BeppoSAX results have revolutionized our understanding of GRBs by opening a window on X-ray, optical and radio afterglows (e.g. Costa et al. 1997; Piro et al. 1998; Van Paradijs et al. 1997). The decaying power-law long-wavelength afterglows are explained as the emission from a relativistic blast wave which decelerates when sweeping up interstellar medium. The dynamical evolution of GRB fireballs and the emission features have been studied by many authors (e.g. Sari 1997; Meszaros, Rees & Wijers 1998; Wei & Lu 1998a,b; Sari, Piran & Narayan 1998), most of them considered the fireball being isotropic.
The discovery of GRB afterglow shows that GRBs are at cosmological distances. If so, the total energy for typical GRB event is about $`10^{52}`$ ergs. This year, an extraordinary event GRB990123 was detected, which was the brightest burst ever detected by BeppoSAX satellite, and is in the top $`0.3\%`$ of all bursts (Feroci et al. 1999). The detection of the redshift showed that the burst appears at $`z1.6`$, with its $`\gamma `$-ray fluence of $`5\times 10^4`$ ergs cm<sup>-2</sup>, the total energy of this source is $`1.6\times 10^{54}`$ ergs if the emission is isotropic (Andersen et al. 1999; Kulkarni et al. 1999). This energy is so large that it gives a great challenge to the popular models. For models involving stellar mass central engines it is necessary to assume that the ejecta are beamed in order to explain such a huge energy.
Now the main uncertainty of bursts’ energy is whether the bursts radiate isotropically or are beamed into a small solid angle. As shown above, the extreme large energy favors the emission being beaming. Then there is one question: how can we identify the radiation between beaming or not? Rhoads (1997, 1999) have shown that the lateral expansion of the shocked, relativistic plasma causes that at some moment the surface of the blast wave starts to increase faster than due to the cone-outflow alone, then the blast wave begins to decelerate faster than without the sideways expansion since more interstellar medium has been swept up by blast wave. Rhoads claimed that this effect will produce a sharp break in the GRB afterglow light curves. Such a break is claimed to be present in the light curves of GRB990123 and GRB990510 (Kulkarni et al. 1999; Harrison et al. 1999). Sari et al. (1999) speculate that afterglows with very steep light curves are highly beamed.
However, some other authors (Panaitescu & Meszaros 1998; Moderski et al. 1999) have performed numerical calculations of dynamical evolution of blast wave, and shown that the break of the light curve is weaker and much smoother than the one analytically predicted. Thus there are two opposite conclusions about the jet emission, the analytical treatment predicts a sharp break, while the numerical calculation shows no such sharp break.
In this paper we will first give an analytical treatment of the dynamical evolution of the jet blast wave and its emission features, and will demonstrate that the sharp break will not actually exist, we may observe the steepening of the light curve only when the jet angle is extremely small, i.e. $`\theta _0<0.1`$. We then perform a simple numerical calculation to confirm our results. In next section we discuss the dynamical evolution of blast wave, in section 3 we calculate the jet emission analytically, and finally we give some discussions and conclusions.
## 2 Dynamical evolution of the jet
Now we consider an adiabatic relativistic jet expanding in surrounding medium. For energy conservation, the evolution equation is
$$\mathrm{\Gamma }^2V=\mathrm{const}$$
(1)
where $`\mathrm{\Gamma }`$ is the bulk Lorentz factor, and $`V`$ is the jet volume, $`V=2\pi r^3(1cos\theta _j)/3r^3\theta _j^2`$ for $`\theta _j1`$, and $`\theta _j=\theta _0+\theta ^{}=\theta _0+c_st_{co}/ct`$, where $`\theta _0`$ is the initial jet opening half-angle, $`\theta ^{}`$ describes the lateral expansion, $`c_s`$ is the expanding velocity of ejecta material in its comoving frame, and $`t(t_{co})`$ is the time measured in the burster frame (comoving frame). For relativistic expanding material it is appropriate to take $`c_s`$ to be the sound speed $`c_s=c/3^{1/2}`$ (Rhoads 1997, 1999), and Rhoads (1999) has given $`t_{co}/t=2/5\mathrm{\Gamma }`$. Since the jet expands relativistically, there is the relation $`Tr/\mathrm{\Gamma }^2`$, where $`T`$ is the time measured in the observer frame, $`r=ct`$ is the radial coordinate in the burster frame. From above relations, we have
$$\mathrm{\Gamma }(1+\frac{\mathrm{\Gamma }_b}{\mathrm{\Gamma }})^{1/4}T^{3/8}$$
(2)
where $`\mathrm{\Gamma }_b=\frac{2}{5}\frac{c_s}{c}\theta _0^1`$. In this paper all quantities with the subscript ”b” denotes the point at which $`c_st_{co}=r\theta _0`$, which means that after that time the sideways expansion begins to dominate the radial divergence. Since $`Tr/\mathrm{\Gamma }^2`$, so $`\mathrm{\Gamma }_b/\mathrm{\Gamma }=(\frac{r_b}{r}\frac{T}{T_b})^{1/2}`$. For the case $`TT_b`$, it is well known that $`\mathrm{\Gamma }T^{3/8}`$, then $`r\mathrm{\Gamma }^2TT^{1/4}`$, and $`\mathrm{\Gamma }_b/\mathrm{\Gamma }=(T/T_b)^{3/8}`$. For another case $`TT_b`$, the radial coordinate $`r`$ is nearly a constant, $`rr_b`$, so $`\mathrm{\Gamma }_b/\mathrm{\Gamma }(T/T_b)^{1/2}`$. Therefore we have
$$\mathrm{\Gamma }\{\begin{array}{cc}T^{3/8}[1+(\frac{T}{T_b})^{3/8}]^{1/4},& \mathrm{if}T<T_b\\ T^{3/8}[1+(\frac{T}{T_b})^{1/2}]^{1/4},& \mathrm{if}T>T_b\end{array}$$
(3)
It is obviously that $`\mathrm{\Gamma }T^{3/8}`$ for $`TT_b`$, and $`\mathrm{\Gamma }T^{1/2}`$ for $`TT_b`$. The rapid decrease with time of Lorentz factor $`\mathrm{\Gamma }`$ is due to the fact that larger amounts of surrounding matter has been swept up by ejecta (Rhoads 1997, 1999).
In the following we calculate the value of $`T_b`$. According to the fireball model, the decelerating radius of the ejecta is
$$r_d=(\frac{E}{\pi \theta _0^2\mathrm{\Gamma }_0^2n_1m_pc^2})^{1/3}$$
(4)
where $`E`$ is the burst energy, and $`\mathrm{\Gamma }_0=E/M_0c^2`$, $`M_0`$ is the initial baryon mass. Rhoads (1999) has given $`r_b=[\frac{75\mathrm{\Gamma }_0^2\theta _0^2}{8(c_s/c)^2}]^{1/3}r_d`$. The relation between $`T`$ and $`r`$ is $`T=r/\zeta \mathrm{\Gamma }^2c`$, where the numerical value of $`\zeta `$ lies between $`3`$ and $`7`$ depending on the details of the hydrodynamical evolution and the spectrum. Sari (1997) and Waxman (1997) have shown that the typical value of $`\zeta `$ is about 4, then the time
$$T_b\frac{r_b}{4\mathrm{\Gamma }_b^2c}=70(\frac{c_s}{c/\sqrt{3}})^{8/3}(\frac{\theta _0}{0.1})^2E_{52}^{1/3}n_1^{1/3}(day)$$
(5)
We see that the break time $`T_b`$ is very large for typical parameters, which means that the transition from $`\mathrm{\Gamma }T^{3/8}`$ to $`\mathrm{\Gamma }T^{1/2}`$ is usually very slowly and smoothly.
## 3 The emission from jet
Now we calculate the emission flux from the jet. Here we adopt the formulation and notations of Mao & YI (1994). In our model the ejecta is flowing outwards relativistically (with Lorentz factor $`\mathrm{\Gamma }`$) in a cone with opening half angle $`\theta _j`$. For simplicity, we assume that the radiation is isotropic in the comoving frame of the ejecta and has no dependence on the angular positions within the cone. The radiation cone is uniquely defined by the angular spherical coordinates ($`\theta `$,$`\varphi `$) of its symmetry axis, here $`\theta `$ is the angle between the line of sight (along $`z`$-axis) and the symmetry axis, and $`\varphi `$ is the azimuthal angle. Because of cylindrical symmetry, we can assume that the symmetry axis of the cone is in the $`yz`$ plane. In order to see more clearly, let us establish an auxiliary coordinate system ($`x^{},y^{},z^{}`$) with the $`z^{}`$-axis along the symmetry axis of the cone and the $`x^{}`$ parallel the $`x`$-axis. Then the position within the cone is specified by its angular spherical coordinates $`\theta ^{}`$ and $`\varphi ^{}`$ ($`0\theta ^{}\theta _j`$, $`0\varphi ^{}2\pi `$). It can be shown that the angle $`\mathrm{\Theta }`$ between a direction ($`\theta ^{},\varphi ^{}`$) within the cone, and the line of sight satisfies $`cos\mathrm{\Theta }=cos\theta cos\theta ^{}sin\theta sin\theta ^{}sin\varphi ^{}`$. Then the observed flux is
$$F(\nu ,\theta )=_0^{2\pi }𝑑\varphi ^{}_0^{\theta _j}sin\theta ^{}𝑑\theta ^{}D^3I^{}(\nu D^1)\frac{r^2}{d^2}$$
(6)
where $`D=[\mathrm{\Gamma }(1\beta cos\mathrm{\Theta })]^1`$ is the Doppler factor, , $`\beta =(1\mathrm{\Gamma }^2)^{1/2}`$, $`\nu =D\nu ^{}`$, $`I^{}(\nu ^{})`$ is the specific intensity of synchrotron radiation at $`\nu ^{}`$, and $`d`$ is the distance of the burst source. Here the quantities with prime are measured in the comoving frame. For simplicity we have ignored the relative time delay of radiation from different parts of the cone.
For the expanding jet, we have $`r=D\mathrm{\Gamma }\beta cTD\mathrm{\Gamma }T(\beta 1)`$, $`r^{}=D\beta cTDT`$, the magnetic field strength $`B^{}\mathrm{\Gamma }`$, the peak frequency of synchrotron radiation $`\nu _m=D\nu _m^{}D\mathrm{\Gamma }^3`$, and $`I^{}(\nu _m^{})n_e^{}B^{}r^{}D\mathrm{\Gamma }^2T`$. Assuming that the emission spectrum $`I^{}(\nu ^{})\nu ^\alpha `$, then $`I^{}(\nu ^{})=I^{}(\nu _m^{})(\frac{\nu ^{}}{\nu _m^{}})^\alpha =I^{}(\nu _m^{})(\frac{\nu }{\nu _m})^\alpha D^{1+\alpha }\mathrm{\Gamma }^{2+3\alpha }T\nu ^\alpha `$. Therefore we have the flux
$$F(\nu ,\theta )\nu ^\alpha \mathrm{\Gamma }^{2(\alpha 1)}T^3g(\theta ,\mathrm{\Gamma },\alpha )$$
(7)
where
$$g(\theta ,\mathrm{\Gamma },\alpha )=_0^{2\pi }𝑑\varphi ^{}_0^{\theta _j}sin\theta ^{}𝑑\theta ^{}(1\beta cos\mathrm{\Theta })^{(6+\alpha )}$$
(8)
In general, the value of $`g`$ can only be calculated numerically. However here we consider the case $`\theta _j1`$ and $`\theta 1`$, then $`cos\mathrm{\Theta }cos\theta cos\theta ^{}`$. In this case we can calculate the value of $`g`$ analytically under certain conditions. After complicated calculation we find $`g\theta ^{2(5+\alpha )}`$ for $`\mathrm{\Gamma }^1<\theta <\theta _j`$, $`g\theta _j^2\theta ^{2(5+\alpha )2}`$ for $`\mathrm{\Gamma }^1<\theta `$ and $`\theta >\theta _j`$, $`g\mathrm{\Gamma }^{2(5+\alpha )}`$ for $`\theta <\mathrm{\Gamma }^1<\theta _j`$, and $`g\mathrm{\Gamma }^{2(5+\alpha )+2}\theta _j^2`$ for $`\theta <\mathrm{\Gamma }^1`$ and $`\theta _j<\mathrm{\Gamma }^1`$. Therefore we have the results
$$F(\nu ,\theta )\{\begin{array}{cc}\nu ^\alpha \theta ^{2(5+\alpha )}\mathrm{\Gamma }^{2(\alpha 1)}T^3,\hfill & \mathrm{for}\mathrm{\Gamma }^1<\theta <\theta _j\hfill \\ \nu ^\alpha \theta _j^2\theta ^{2(5+\alpha )2}\mathrm{\Gamma }^{2(\alpha 1)}T^3,\hfill & \mathrm{for}\mathrm{\Gamma }^1<\theta ,\theta >\theta _j\hfill \\ \nu ^\alpha \mathrm{\Gamma }^{8+4\alpha }T^3,\hfill & \mathrm{for}\theta <\mathrm{\Gamma }^1<\theta _j\hfill \\ \nu ^\alpha \theta _j^2\mathrm{\Gamma }^{10+4\alpha }T^3,\hfill & \mathrm{for}\theta <\mathrm{\Gamma }^1,\theta _j<\mathrm{\Gamma }^1\hfill \end{array}$$
(9)
For $`TT_b`$, the evolution is about $`\mathrm{\Gamma }=\mathrm{\Gamma }_0(r/r_d)^{3/2}`$. Here we define the Lorentz factor, $`\mathrm{\Gamma }_j\mathrm{\Gamma }(r_j)=\theta _j^1`$, then $`r_j=(\mathrm{\Gamma }_0\theta _j)^{2/3}r_d`$, and the corresponding timescale $`T_jr_j/4\mathrm{\Gamma }_j^2c1.3(\frac{\theta _j}{0.1})^2E_{52}^{1/3}n_1^{1/3}(\frac{\theta _j}{\theta _0})^{2/3}`$ day. Then the observed flux at fixed frequency is
$$F\{\begin{array}{cc}T^{3\alpha /2},\hfill & \mathrm{for}T<T_j\hfill \\ T^{\frac{3}{2}\alpha \frac{3}{4}},\hfill & \mathrm{for}T_j<T<T_b\hfill \\ T^{2\alpha 1},\hfill & \mathrm{for}T_b<T\hfill \end{array}$$
(10)
From above it seems that there should be two temporal index breaks in light curves. However, if we compare the values of $`T_j`$ and $`T_b`$, we will find that $`T_jT_b`$, i.e. the time interval between $`T_j`$ and $`T_b`$ is very large, the beaming break is much earlier than the break due to sideways expansion. We know that in order to see the steepening of the light curve, $`T_j`$ or $`T_b`$ must be small, so in fact we can only see one temporal break. In addition, in the Rhoads’ treatment, the effect of the sideways expansion on the $`\mathrm{\Gamma }`$ evolution was ignored when $`T<T_b`$, however in fact, there is still some sideways expansion during this phase, so the evolution of $`\mathrm{\Gamma }`$ must be affected by sideways expansion when $`\mathrm{\Gamma }\theta _0^1`$. Therefore, we expect that the evolution of $`\mathrm{\Gamma }`$ is continuous, and the transition from $`\mathrm{\Gamma }T^{3/8}`$ to $`\mathrm{\Gamma }T^{1/2}`$ is much smoother than previously claimed.
In order to test our conclusion, we make a simple numerical calculation. We assume that the blast wave evolution is adiabatic, ignore cooling of the swept-up particles. We take the following initial parameters: $`\mathrm{\Gamma }_0=300`$, the electron distribution index $`p=2.5`$. In Fig.1 we present the light curve for different initial opening angle, the solid, dotted and dashed lines represent the cases where $`\theta _0=0.1,\mathrm{\hspace{0.17em}\hspace{0.17em}0.0174},\mathrm{\hspace{0.17em}\hspace{0.17em}0.01}`$ respectively. We show that when $`\theta _0=0.1`$, the light curve is nearly not affected by sideways expansion, while when $`\theta _0=0.0174`$ and 0.01, the light curves are steepened clearly, confirming our analytic conclusion.
As for comparison, we also calculate the afterglow light curves of blast wave with no spreading. We know that, if the blast wave is highly radiative, the internal energy of the blast wave will be low, since it is converted to photons and radiated away, so the lateral expanding velocity will be very small, $`c_s<<c`$, in this case, the sideways expansion is unimportantly small, and the blast wave can be regarded as no spreading. For highly radiative evolution, it has been shown that the Lorentz factor $`\mathrm{\Gamma }T^{3/7}`$ (Wei & Lu 1998a). Taking the parameters as above, we calculate the afterglow light curves under this situation, Fig.2 gives our results, the solid, dotted and dashed lines also correspond to the cases $`\theta _0=0.1,\mathrm{\hspace{0.17em}\hspace{0.17em}0.0174},\mathrm{\hspace{0.17em}\hspace{0.17em}0.01}`$ respectively. It is obvious that, when $`T<T_c`$ ($`T_cT_0(\mathrm{\Gamma }_0\theta _0)^{7/3}`$), the light curves decay as a simple power law, and when $`T>T_c`$, the light curves deviate from the simple power law and show a steepening, and a clear break occurs at about $`T_c`$. Also we note that only when $`\theta _0<0.1`$, the steepening is obvious. Therefore we suggest that if we observe the sharp break in the GRB afterglow light curves, then it may indicate that the blast wave is highly radiative rather than adiabatic.
## 4 Discussion and conclusions
The GRB afterglows provide very good opportunity to study whether and how much the GRB ejecta are beamed. Rhoads (1997, 1999) has pointed out that the beamed outflows should diverge from the cone geometry and the sideways outflow of the shocked relativistic plasma would increase the front of the blast wave which leading to a fast deceleration. He also predicted that the afterglow light curves should have a sharp break around $`T_b`$.
However, Moderski et al. (1999) have performed numerical calculation and shown that the break of the light curve is weaker and smoother than the prediction. Here we reanalyse the dynamical evolution of the jet blast wave, calculate the emission from the jet. Our calculations show that the main reason why the results of Moderski et al. being different from that of Rhoads is that the value of $`T_b`$ is very large when taking the parameters adopted by Modersli et al. (see eq. 5). Our formular (eq. 3) indicates that the evolution of Lorentz factor $`\mathrm{\Gamma }`$ with time $`T`$ is continuous, changing the slope from -3/8 to -1/2 smoothly. In particular, if the value of $`T_b`$ is large, then the transition is much smoothly, in this case one expects that the sharp break will not exist. In order to test our analytic conclusion, we also make a simple numerical calculation, from Fig.1 it is shown that only when $`\theta _0<0.1`$, we can observe the steepening of the light curve, which is consistent with our analytic conclusion.
Our results are valid only if the remnant is still relativistic at time $`T_b`$. Since at this time the Lorentz factor $`\mathrm{\Gamma }_b\frac{2}{5}\frac{c_s}{c}\theta _0^1`$, this condition reduces to $`\theta _0<0.1`$, i.e. the jet is very narrow. If $`\theta _0>0.1`$, then before the sideways expansion is important, the remnant has already become non-relativistic. In fact, Dai and Lu (1999) have shown that even in the case of isotropic emission, the break in the light curve can appear during the transition from ultrarelativistic to non-relativistic phase in the environment of dense material.
It should be emphasized that in our calculation we have ignored the relative time delay of radiation from different parts of the cone, if this effect is considered, then the slope of the light curve should be flatter than $`T^{(2\alpha +1)}`$ (Moderski et al. 1999). However, all these calculation are based on the assumption that the material is uniformly distributed across the blast wave. In fact, it is more likely that the lateral outflow can lead to $`\theta `$ dependent structure, with the density of swept material and the bulk Lorentz factor decreasing with $`\theta `$, in this case the break in the light curve may become more prominent.
We thank the referee for several important comments that improved this paper. This work is supported by the National Natural Science Foundation (19703003 and 19773007) and the National Climbing Project on Fundamental Researches of China.
Figure Caption
Fig.1 The afterglow light curves for different initial opening angle, the solid, dotted and dashed lines represent the cases where $`\theta _0=0.1,\mathrm{\hspace{0.17em}\hspace{0.17em}0.0174}`$ and 0.01 respectively.
Fig.2 The afterglow light curves for highly radiative blast wave with no spreading. The solid, dotted and dashed lines correspond to the cases where $`\theta _0=0.1,\mathrm{\hspace{0.17em}\hspace{0.17em}0.0174}`$ and 0.01 respectively.
|
no-problem/9908/cond-mat9908189.html
|
ar5iv
|
text
|
# Yrast line for weakly interacting trapped bosons
## Abstract
We compute numerically the yrast line for harmonically trapped boson systems with a weak repulsive contact interaction, studying the transition to a vortex state as the angular momentum $`L`$ increases and approaches $`N`$, the number of bosons. The $`L=N`$ eigenstate is indeed dominated by particles with unit angular momentum, but the state has other significant components beyond the pure vortex configuration. There is a smooth crossover between low and high $`L`$ with no indication of a quantum phase transition. Most strikingly, the energy and wave function appear to be analytical functions of $`L`$ over the entire range $`2LN`$. We confirm the structure of low-$`L`$ states proposed by Mottelson, as mainly single-particle excitations with two or three units of angular momentum.
The low-lying excitations of atomic Bose Einstein condensates in harmonic traps are of considerable experimental and theoretical interest . Recently, Mottelson proposed a theory for the yrast line of weakly interacting $`N`$-boson systems , i.e. the ground states at nonvanishing angular momentum $`L`$. Physical arguments led him to assume that the yrast states are excited upon acting on the ground state $`|0`$ of vanishing angular momentum with a collective operator $`Q_\lambda =_{p=1}^Nz_p^\lambda `$ that is a sum of single-particle operators acting on the coordinates $`z_p=x_p+iy_p`$ of the $`p^{\mathrm{th}}`$ particle. For angular momenta $`LN`$ the yrast states are found to be dominated by quadrupole ($`\lambda =2`$) and octupole ($`\lambda =3`$) modes. Assuming a vortex structure of the yrast states with $`LN`$ then led to the prediction of a quantum phase transition in Fock space when passing from the low angular momentum regime $`LN`$ to the regime of high angular momenta $`LN`$. The reason for this is behavior is the approximate orthogonality of the collective states $`Q_\lambda |0`$ and the single-particle oscillator states of the vortex line in the regime $`N^{1/2}L`$. These results have been obtained for harmonically trapped bosons with a weak repulsive contact interaction. The case of an attractive interaction has been studied by Wilkin et al. . In this case the total angular momentum is carried by the center of mass motion, and there are no excitations corresponding to relative motion. This is not unexpected since internal excitations would increase the energy of the yrast state.
It is the purpose of this letter to present an independent numerical computation of the yrast line and to compare with Mottelson’s results . In particular we want to focus on the transition from low to high angular momentum yrast states. The investigation of this transition is of interest not only for the physics of Bose-Einstein condensates. Localization in Fock space is under investigation also in molecular and condensed matter physics . The numerical computation has the advantage that it does not rely on the assumptions made in the analytical calculation. However, with our numerical methods it is limited to angular momenta below about $`Ł50`$. Most interestingly, our numerical results suggest that the yrast line and the corresponding wave functions can be presented by rather simple analytical expressions.
Let us consider $`N`$ bosons in a two-dimensional harmonic trap interacting via a contact interaction<sup>*</sup><sup>*</sup>*The results obtained below extend to the three dimensional problem for $`L=L_z`$.. We are interested in the yrast line in the perturbative regime of weak interactions. Note however that experimental studies of trapped condensates are often in a regime where the interaction energy is comparable to the trapping potential, and this may introduce qualitatively different physics. We write the Hamiltonian as
$$\widehat{H}=\widehat{H}_0+\widehat{V}.$$
(1)
Here
$$\widehat{H}_0=\mathrm{}\omega \underset{j}{}j\widehat{a}_j^{}\widehat{a}_j$$
(2)
is the one-body oscillator Hamiltonian and
$$\widehat{V}=g\underset{i,j,k,l}{}V_{ijkl}\widehat{a}_i^{}\widehat{a}_j^{}\widehat{a}_k\widehat{a}_l.$$
(3)
the two-body interaction. The operators $`\widehat{a}_m`$ and $`\widehat{a}_m^{}`$ annihilate and create one boson in the single-particle oscillator state $`|m`$ with energy $`m\mathrm{}\omega `$ and angular momentum $`m\mathrm{}`$, respectively and fulfill bosonic commutation rules. The ground state energy is set to zero. Up to some irrelevant overall constant the matrix elements are given by $`V_{ijkl}=2^{kl}(k+l)!/(i!j!k!l!)^{1/2}`$ and vanish for $`i+jk+l`$. For total angular momentum $`L`$ the Fock space is spanned by states $`|\alpha |n_0,n_1,\mathrm{},n_k`$ with $`_{i=0,k}n_i=N`$, $`\widehat{a}_j^{}\widehat{a}_j|n_0,n_1,\mathrm{},n_k=n_j|n_0,n_1,\mathrm{},n_k`$ and $`_{j=0,k}jn_j=L`$. Here $`n_j`$ denotes the occupation of the $`j^{\mathrm{th}}`$ single particle state $`|j`$. For vanishing coupling $`g`$ the basis states are degenerate in energy, and the problem thus consists in diagonalizing the two-body interaction $`\widehat{V}`$ inside the Fock space basis. To set up the matrix we act with the operator (3) onto one initial basis state with angular momentum $`L`$ and onto all states created by this procedure until the the Fock space is exhausted . The resulting matrix is sparse, and the yrast state is computed using a Lanczos algorithm . We restrict ourselves to $`L50`$ corresponding to a maximal Fock space dimension of about $`d_L210^5`$.
The yrast line, i.e. the ground state energies as a function of the angular momentum may be written as $`E(L)=L\mathrm{}\omega +gϵ_L`$. Fig. 1 shows the $`L`$-dependence of the energies $`ϵ_L`$ for systems of $`N=`$ 25 and 50 bosons. The energies $`ϵ_L`$ simply decrease linearly with increasing angular momentum for $`LN`$. In fact to machine precision, the energy function is found to described by an algebraic expression,
$$ϵ_L=\frac{N(2NL2)}{2}$$
(4)
At fixed angular momentum $`L`$ and for $`LN`$ the energies $`gϵ_l`$ increase as expected with the square of the number of bosons $`N`$. Notice in the figure that there is a kink in the slope at $`N=L`$. This is a hint of condensation into a vortex state: in macroscopic superfluids, the state for $`L=N`$ would have a condensate of unit angular momentum and would be lower in energy than neighboring yrast states.
We next investigate the structure of the wave functions of yrast states. We would like to know how complex the states are and how well they can be described by single-particle operators acting on simple states. To address the question of the complexity of the states in the Fock basis, we take the wave function amplitudes $`c_\alpha ^{(L)}`$ in the Fock representation of the state
$$|L=\underset{\alpha =1}{\overset{d_L}{}}c_\alpha ^{(L)}|\alpha $$
and compute the inverse participation ratio
$$I_L\underset{\alpha =1}{\overset{d_L}{}}|c_\alpha ^{(L)}|^4.$$
The $`I_L`$ is the first nontrivial moment of the distribution of wave function intensities $`|c_\alpha ^{(L)}|^2`$. Its inverse $`1/I_L`$ measures the number of basis states $`|\alpha `$ that have significant overlap with the yrast state $`|L`$. Fig 2 shows a plot of $`1/I_L`$ and the Fock space dimension $`d_L`$ as a function of angular momentum $`L`$ for a system of $`N=50`$ bosons. The $`1/I_L`$ is seen to be much smaller than the dimensionality of the Fock space. Even where the participating is greatest, at midvalues of $`L`$, only about 30 states are active participants. A similar behavior of quantum non-ergodicity has been found previously in numerical studies . Notice that the inverse participation ratio decreases strongly as $`N=L`$ is approached. This shows that the yrast state becomes simpler, again hinting at the formation of a vortex condensate. Examining the coefficients for $`N=25`$ in detail, the largest amplitude at $`L=N`$ is in fact the vortex state, $`|\alpha =|0N0\mathrm{}0>`$, but it has less than half the probability of the complete wave function. Interestingly, our numerically obtained yrast state $`|L=N=25`$ agrees with the conjecture given by Wilkin et al. , i.e. $`|L=N=_{p=1}^N(z_pz_c)|0`$ with $`z_c=N^1_{p=1}^Nz_p`$ being the center of mass. Based on our numerical wave functions, we can generalize this conjecture. We believe that all the yrast states for $`2LN`$ are given by the formula
$$|L=\underset{p_1<p_2<\mathrm{}<p_L}{}(z_{p_1}z_c)(z_{p_2}z_c)\mathrm{}(z_{p_L}z_c)|0>$$
(5)
We have verified that this formula is correct (up to machine precision) by comparison with the numerically obtained yrast states for $`N=25`$. Since the operator acting on the ground state is translationally invariant no quanta of the center of mass motion are excited. Notice that there is a natural termination of the construction at $`L=N`$.
To further examine the structure of the yrast states $`|L`$ we show a plot of the occupation numbers $`n_j^{(L)}L|\widehat{a}_j^{}\widehat{a}_j|L`$ for $`j=0,1,2,3`$ in Fig. 3 for a system of $`N=50`$ bosons. At very low angular momenta the yrast states are dominated by single particle oscillator states with two or three units of angular momentum. This is in agreement with Mottelson’s results . However, at larger angular momentum $`L`$ the dominant fraction is carried by single particle states with one unit of angular momentum. Note that the occupation numbers $`n_j^{(L)}`$ are very small for $`j>3`$. This analysis confirms the results found for the inverse participation ratio. Note also that the observables $`n_j^{(L)}`$ are very smooth functions of $`L`$. If there were a quantum phase transition at large $`LN/2`$, we would expect to see some precursor in these observables.
In conclusion, our numerical study strongly indicates that there is no quantum phase transition to a vortex state for trapped condensates in the limit that the interaction potential is small compared to the oscillator frequency. The strongest evidence is the apparent existence of analytic expressions for the energies and the wave functions on the yrast line for $`2LN`$ One might speculate that these states are contained in a dynamical symmetry group, but we have no idea how this might come aboutWe note also that there is another symmetry group , $`SO(2,1)`$ , that produces relationships between energies of different states within a single $`L`$ subspace.. We have also examined the structure of the yrast states and the matrix elements between them, finding that the observables vary smoothly with $`L`$, for $`L`$ not too small.
We acknowledge conversations with B. Mottelson. This work was supported by the Dept. of Energy under Grant DE-FG-06-90ER40561.
|
no-problem/9908/hep-lat9908003.html
|
ar5iv
|
text
|
# Efficiencies and optimization of HMC algorithms in pure gauge theory
## Heatbath/Overrelaxation (HOR).
Each update is composed of one heatbath sweep followed by $`N_{or}`$ overrelaxation sweeps.
* Overrelaxation: microcanonical reflections in three SU(2) subgroups.
* Heatbath: Cabbibo-Marinari method with random matrices from SU(2) subgroups. SU(2) matrices are generated with the Fabricius-Haan algorithm.
A sweep consists of loops over $`x`$ and $`\mu `$. The order of these loops turned out to have a significant influence on the autocorrelation times (with the inner loop over $`x`$ being in advantage).
## Hybrid Monte Carlo (HMC).
The HMC is a member of the family of algorithms which are based on classical dynamics. To this end, one considers a Hamiltonian
$$H[P,U]=\frac{1}{2}\underset{x,\mu }{}P_{x,\mu }^2+S[U]$$
(4)
with momenta $`P_{x\mu }`$ conjugate to the link variables. In each update step, momenta are generated with a Gaussian distribution and a reversible discretized trajectory is computed. Discretization errors are corrected by an acceptance step.
Parameters which can be optimized are the trajectory length $`\tau `$ and the step size $`\delta \tau `$.
## Local Hybrid Monte Carlo (LHMC).
We also consider a local version of the HMC algorithm proposed in . Here one applies the same procedure as for the global HMC algorithm, but for one link while keeping all others fixed. This has some advantages compared to the global HMC:
* The difference in the action accumulated on a trajectory does not contain a volume factor. As a consequence, much greater step sizes sizes are possible without getting poor acceptance rates.
* The staples for a link – which are involved in the computation of the force – have to be computed only once per trajectory, so additional steps on the trajectory are cheap. This point has turned out to be of minor importance, as the optimal parameter set has $`\tau /\delta \tau `$ 2 – 3.
An advantage compared to the HOR algorithm is that $`\mathrm{\Delta }S`$ is only needed for infinitesimal $`\mathrm{\Delta }U`$. This makes no practical difference for the standard Wilson action considered here, but it is relevant for more complicated actions where terms quadratic in $`U`$ appear.
The LHMC algorithm cannot be applied to QCD with pseudofermions in a straightforward way, because the computation of the force would then require the inversion of a fermion matrix in each local step. However, for a simulation of a bermion theory with clover term, which we are preparing, it is an attractive candidate for the update of the gauge field .
We mention here that for a Wilson action, the leapfrog algorithm for generating a candidate configuration can be replaced by an exact integration of the equations of motion . The method used there makes use of the fact that in the case where subgroups of SU(3) are updated separately, the action takes the form of the energy of a ”pendulum”.
We want to represent efficiency measures in a way that allows a machine-independent comparison of algorithms. Thus, we define a measure $`S_{\text{cost}}`$ such that in order to compute $`\overline{g}^2`$ at $`1\%`$ accuracy, the equivalent in complexity of $`S_{\text{cost}}`$ computations of all staples is required.
The optimization of the three algorithms used in this study is shown in figures 1 - 5.
In the following we investigate how the different algorithms scale with $`L/a`$. The variance of the coupling between $`L/a=4`$ and $`L/a=10`$ behaves approximately like $`(L/a)^{1.4}`$. Thus, we expect a behaviour of the cost measure $`S_{\text{cost}}(L/a)^{1.4+z}`$. In figure 6 we have plotted $`S_{\text{cost}}`$ with optimized parameters for each algorithm against $`L/a`$. The dotted lines correspond to exponents $`z=1`$, $`z=2`$ resp.
From our data at $`L/a=8`$ we conclude that typical ratios in $`S_{\text{cost}}`$ for optimized parameters are 1 : 3 : 26 for HOR : LHMC : HMC. This illustrates the cost of HMC even before dynamical quarks are included. A similar performance ratio for HMC and HOR was concluded in , which recently came to our attention.
We thank DESY for allocating computer time to this project and the DFG under GK 271 for financial support.
|
no-problem/9908/cond-mat9908281.html
|
ar5iv
|
text
|
# Observation of Magnetic Fingerprints in Superconducting Au0.7In0.3 Cylinders
\[
## Abstract
Reproducible, sample-specific magnetoresistance fluctuations (magnetic fingerprints) have been observed experimentally in the low-temperature part of the superconducting transition regime of disordered superconducting Au<sub>0.7</sub>In<sub>0.3</sub> cylinders. The amplitude of the fluctuation was found to exceed that of the universal conductance fluctuation in normal metals by several orders of magnitude. The physical origin of these observations is discussed in the context of mesoscopic fluctuations of the superconducting condensation energy in disordered superconductors.
\]
In the past two decades fascinating phenomena in normal-metal mesoscopic systems have been found and, for the most part, understood . One of the most important aspects of mesoscopic physics is quantum interference over a length much larger than the atomic size. In disordered mesoscopic samples, this remarkable phenomenon is manifested in seemingly random but fully reproducible sample-specific magnetoresistance fluctuations, referred to in literature as magnetic fingerprints (MFPs) . These MFPs, which have emerged as a hallmark of mesoscopic physics, result from Aharonov-Bohm interference of electron waves. Remarkably, the amplitude of the conductance fluctuations has a universal value of the order of $`e^2/h`$, known as the universal conductance fluctuation (UCF). The physical origin of UCF lies in the energy level statistics in disordered metal, where the fluctuation in the number of energy levels within the Thouless energy is universally of the order of unity .
In the past few years, the UCF has also been examined in normal-metal samples in contact with one or more superconducting islands . Andreev reflection from the normal metal-superconductor interfaces extends phase coherence in the normal metal beyond the normal coherence length , leading to new physical phenomena . However, no significant change was found in conductance fluctuations , as anticipated theoretically .
Interesting questions arise if superconductivity is introduced in the bulk, rather than at the boundary of a normal sample. Consider a weakly disordered mesoscopic sample in which electrons become phase coherent well above the onset of superconductivity. These phase-coherent normal electrons are extremely sensitive to impurity scattering . However, when electrons form Cooper pairs, they become completely insensitive to randomness. How do electrons respond to these opposite tendencies of motion? In addition, in disordered metallic samples, energy levels fluctuate, leading to MFPs and UCF as mentioned above. What would the manifestation of the energy level fluctuation be in disordered superconductors? Experimentally, these issues have not been examined prior to the present work. In particular, no MFPs have been reported for superconductors. In this Letter, we report results obtained on superconducting Au-In cylinders, in which sample-specific magnetoresistance fluctuations, or MFPs, have been found.
Superconducting Au-In binary alloy was originally chosen for this study because its critical temperature ($`T_c`$) can be easily controlled by changing the In concentration. Au-In alloy has a rich phase diagram that includes compounds, AuIn and AuIn<sub>2</sub>, and solid solutions with varying composition ratios . For the latter, the $`T_c`$ continuously changes with In concentration . An important consequence of this is that inhomogeneity in In concentration results in spatially varying local $`T_c`$’s. Spatial fluctuation in $`T_c`$ in our samples may be related to the origin of the observed sample-specific resistance fluctuations (see below).
In the bulk form, the maximum solid solubility of In in Au is about 12% . When the In concentration exceeds this limit, a phase separation is expected to occur, with the excess In forming In-rich grains. In thick Au<sub>0.7</sub>In<sub>0.3</sub> planar films these In-rich grains can be directly observed as they form micron-size grain conglomerates (Inset (a) of Fig. 1). In thinner planar and cylindrical films such grain segregation is not found (Inset (b) of Fig. 1), probably because of the reduced mobility of atoms due to substrate effects. Nonetheless, the onset of superconductivity, marked by the initial resistance decrease, occurs at the same temperature as in thicker films, suggesting that In-rich grains are also present in these samples .
To prepare a sample, an insulating (GE 7031 varnish) filament of submicron diameter was drawn and placed across a gap in a thin glass slide. The slide was then mounted on a rotator inside an evaporation system. A cylindrical film of Au<sub>0.7</sub>In<sub>0.3</sub> was prepared by depositing 99.9999% pure Au, In, and Au sequentially in the appropriate proportion onto the rotating filament. The thickness of the films was measured with a quartz crystal thickness monitor. The length of the free-standing cylindrical film is given by the width of the gap ($``$1mm). The diameters of the cylinders were determined using scanning electron microscopy. Current and voltage leads were attached to the cylinder using Ag epoxy. The cylinders were manually aligned to be as parallel to the magnetic
field as possible. The samples were stored at room temperature for at least several days and then slowly cooled down in a dilution refrigerator. To make sure any possible residual thermal strain is relieved, the samples were kept at low temperature for several more days before any measurements were carried out. All electrical leads entering the cryostat were RF filtered. The resistance was measured in d.c. at 1$`\mu `$A.
In Fig. 1, resistance $`R`$ of two Au<sub>0.7</sub>In<sub>0.3</sub> cylindrical films, Cylinders A and B, is plotted against temperature $`T`$. The diameter, nominal thickness, and the normal-state sheet resistance of the two cylinders were respectively $`0.84\mu `$m, 350Å, and $`1.7\mathrm{\Omega }`$ for A and $`0.60\mu `$m, 300Å, and $`2.5\mathrm{\Omega }`$ for B. A wide transition regime was found for both samples, as expected for inhomogeneous films. The temperature range of the transition is consistent with the expected range of local $`T_c`$ variation, from $`0.1`$K for uniform Au<sub>0.88</sub>In<sub>0.12</sub> matrix to the maximum of $`0.6`$K ($`T_c`$ of bulk AuIn) for In-rich grains. No resistance drop was seen at 3.4K, the superconducting transition temperature of bulk In. The samples become fully superconducting around 0.1K for Cylinder A and 0.3K for Cylinder B.
In Fig. 2a, two traces of magnetoresistance (MR) scan for Cylinder A, taken at $`T=0.25`$K, deep into the superconducting transition regime, are shown. A non-periodic, asymmetric (with respect to the reversal of the magnetic field) MR pattern is seen in both traces. A comparison of the two traces shows a remarkable reproducibility of the pattern (the cross-correlation is 97%). This pattern can be seen as a reproducible resistance fluctuation, or magnetic fingerprint, in a positive, symmetric MR background expected for a superconductor. Similar MFPs have been found in most Au<sub>0.7</sub>In<sub>0.3</sub> cylinders we studied
so far. In Fig. 2b we show a set of data obtained in Cylinder B.
A small increase in temperature was seen to suppress the fluctuations surprisingly strongly (Fig. 3). For Cylinder A, at $`T=T^{}0.27`$K, still deep in the transition regime, the resistance fluctuation already disappeared completely. Magnetic field was found to have a similar effect. Above a threshold field $`H^{}(T)`$, the resistance fluctuation disappeared and the MR recovered the monotonic, symmetric behavior. It is interesting to note that the fluctuation disappeared once the resistance was above certain value, either by increasing temperature or magnetic field.
The MFPs remained essentially the same in several consecutive scans. However, after the sample was thermally cycled to around 10-15K, well above the onset of superconductivity, a different fluctuation pattern was found, as shown for example in Fig. 3 (bottom trace) and in Fig. 4 for Cylinder A. The magnitude of the zero-field resistance $`R_{H=0}`$ at a fixed temperature was also found to change randomly as a result of thermal cycling. The range of the resistance variation at $`T=0.25`$K was about 60$`\mathrm{\Omega }`$, or 10% of the normal-state resistance $`R_N`$.
Applying a high (several Tesla) magnetic field also irreversibly changed the MR, similar to thermal cycling.
A resistance maximum at zero magnetic field is clearly seen in Fig. 4. Typically, superconducting fluctuations are suppressed by an applied field, leading to a positive MR. In the conventional Little-Parks (L-P) experiment, the MR at $`H=0`$ is always a minimum . A negative MR as large as 25% of $`R_{H=0}`$ deep in the superconducting transition regime is therefore very unusual. Similar negative MR has been observed in other Au<sub>0.7</sub>In<sub>0.3</sub> cylinders, as seen for example in Fig. 2b. The negative MR was suppressed by a small temperature increase as all other features in the MR were (Fig. 3).
The data shown in Figs. 2-4 suggest that the conventional L-P resistance oscillation was too weak to be observed or even absent in Cylinders A and B, which we believe is due to the following reason. All cylinders used in the present study were measured in a free-standing configuration. Varnish undergoes a much larger thermal contraction than Au-In alloy. Therefore ”cracks” may have developed along the cylinder during the cooling down due to insufficient thermal anchoring of a free-standing sample, consistent with the AFM studies of some cylinders, which showed fine trenches parallel to the axis (Inset (b) of Fig. 1). The multiply connected part of the cylinder may be small, with widely varying local $`T_c`$’s, leading to suppression of the L-P oscillation.
Sample-specific MR could in principle result from multiple magnetic field driven transitions if the sample consisted of a collection of superconducting weak links with varying local critical field. In this picture, however, successive suppression of superconductivity of each individual weak link as the (parallel) field increases would result in monotonic, step-like features in MR, accompanied by hysteresis . Instead, MR of our samples was found to be strongly non-monotonic and non-hysteretic. Furthermore, the MR was asymmetric with respect to the magnetic field reversal, which also can not be explained
by the weak link picture. All these considerations seem to suggest that superconducting weak links, if present in our samples, do not contribute significantly to the observed sample-specific MR.
Mesoscopic conductance fluctuations in normal-metals are sensitive to impurity configurations, magnetic fields, and gate voltages . Thermal cycling to moderately high temperature can affect the impurity configuration and therefore result in a conductance change of the order of $`e^2/h`$. Similarly, thermal cycling results in a conductance change in our samples. It is possible that, due to uneven thermal contraction, thermal strains might have developed during temperature cycling. Such strains could cause structural changes in the samples and might account for the observed variation in the sample resistance. However, thermal cycling did not affect the normal state resistance, suggesting that any resulted structural changes were very small. Unlike in normal metals, conductance of our samples also changed irreversibly after they were exposed to a very high magnetic field, an issue that remains unresolved.
In mesoscopic samples of normal metals, magnetic field also modifies the sample-specific conductance, resulting in MFPs. Magnetic field of the order of the correlation field $`H_{cor}`$, corresponding to one flux quantum through the cross-section of the film, is required to change the conductance by $`e^2/h`$ . MFPs were also found in our samples, however, due to the suppression of superconductivity, the MFPs were only observed in fields up to $`H^{}`$, smaller than $`H_{cor}450`$G. As a result the most prominent fluctuation features had field scale much smaller than $`H_{cor}`$. It should be noted that conductance fluctuations on field scales much smaller than $`H_{cor}`$ have been observed in normal-metal samples , with amplitude smaller than $`e^2/h`$.
The similarities between the sample-specific conductance in our samples and in mesoscopic normal-metal systems strongly suggest that the observed features are mesoscopic in origin. However, the amplitude of these sample-specific conductance fluctuations appears to be much larger than that observed in normal samples. An order-of-magnitude estimate gives $`\mathrm{\Delta }G=\mathrm{\Delta }R_{\mathrm{}}/R_{\mathrm{}}^210^4e^2/h`$ for Cylinder A at 0.25K, where $`R_{\mathrm{}}`$ is the sheet resistance of the sample.
Theoretically, significantly enhanced sample-specific conductance fluctuations have been predicted for homogeneously disordered superconductors in the transition regime. It has been shown that under appropriate conditions, such as close to the superconductor-insulator transition or in a strong parallel magnetic field, fluctuations in superconducting condensation energy can be larger than its mean value . The physical origin of these exceedingly large fluctuations lies in the level statistics, precisely the origin of the UCF in normal metals. The fluctuation in condensation energy will in turn manifest itself in fluctuations of the local $`T_c`$ even for a homogeneously disordered superconductor . Zhou and Biagini have shown that mesoscopic fluctuations of both Aslamasov-Larkin and Maki-Thompson contributions to conductivity would lead to a sample-specific conductance fluctuation above the $`T_c`$. Because of the long-range phase coherence developing in superconductors as $`T_c`$ is approached, sample-specific conductance should be observable in arbitrarily large samples, as long as the temperature is sufficiently close to $`T_c`$. Similar to normal samples, these fluctuations are sensitive to magnetic field, impurity configuration, and gate voltage. Conductance fluctuations are greatly amplified due to the superconducting coherence resulted from Cooper pairing correlation, a spectacular example of quantum mesoscopic phenomena at a macroscopic scale.
The calculation of Zhou and Biagini has been carried out for homogeneously disordered superconductors. Therefore, strictly speaking, it is not directly applicable to our experimental system. Nonetheless, the salient features predicted by the theory are expected to be present for inhomogeneously disordered superconductors as well . Below we compare our experimental observations with these predictions. First, the predicted sample-specific conductance fluctuation was observed experimentally, and only in a narrow temperature range right above $`T_c`$, consistent with the theory. Second, the amplitude of the sample-specific conductance fluctuation close to $`T_c`$ was found to greatly exceed that of the UCF in normal samples, again consistent with the theory. Finally, negative MR, observed in our experiment, can be naturally accounted for in the same theoretical framework, as shown earlier by Spivak and Kivelson . The qualitative agreement between our experimental observations and the theory strongly suggests that the same physics as discussed above is at work in our Au<sub>0.7</sub>In<sub>0.3</sub> samples.
Before closing, we remark that the theory in reference also predicts a resistance oscillation of period $`h/4e`$, half of the L-P period. This oscillation, if present in Figs. 2-3, is masked by aperiodic features. In Fig. 4, however, reproducible MR peaks appear around $`H=`$ 0, -20, and -40G. The peak separation is very close to $`h/4e`$. Therefore, these peaks could be a signature of an $`h/4e`$ resistance oscillation. Finally, the observed asymmetry in the magnetoresistance may have a related physical origin. In normal samples, only a 4-point measuring configuration would result in asymmetic MR , because of the fundamental requirement of time-reversal symmetry. The Spivak-Kivelson theory , however, allows for time-reversal symmetry breaking in the ground state of a disordered superconductor, which may lead to asymmetric MR.
In conclusion, we have observed, for the first time, reproducible, sample-specific resistance fluctuations in disordered Au<sub>0.7</sub>In<sub>0.3</sub> cylinders. The amplitude of the fluctuation is much larger than that of the UCF in normal samples. We have argued that the physical origin of these observations lies in the mesoscopic fluctuation of superconducting condensation energy, as predicted by theory.
The authors would like to acknowledge useful discussions with S. Kivelson, B. Spivak, and F. Zhou. This work is supported by NSF under grant DMR-9702661.
|
no-problem/9908/quant-ph9908049.html
|
ar5iv
|
text
|
# Quantum Teleportation Using Quantum Non-Demolition Technique
## Abstract
We propose a new scheme and protocol for quantum teleportation of a single-mode field state, based on entanglement produced by quantum non-demolition interaction. We show that the recently attained results in QND technique allow to perform the teleportation in quantum regime. We also show that applying QND coupling to squeezed fields will significantly improve the quality of teleportation for a given degree of squeezing.
PACS numbers: 03.67.-a, 03.65.-w, 42.50.Dv
Quantum teleportation is a transport of quantum state of a system to another similar system via a classical channel . The principles of quantum mechanics demand that for realization of such a transport, the sender and the receiver must have two ancillary quantum systems, being strongly quantum-mechanically correlated (entangled). Quantum teleportation is a fundamental phenomenon of quantum world and also one of the key procedures in the rapidly growing area of quantum information . Recently several teleportation schemes have been proposed and the successful realization of some of them has been reported: teleportation of the polarization state of a single photon , of the state of a single-mode optical field , and of nuclear spin state . In this Letter we propose a new scheme for quantum teleportation of the state of single-mode optical field. In contrast to the existing schemes , , where a superposition of two squeezed fields has been used as a source of entanglement, we propose to use the entanglement produced by quantum non-demolition (QND) interaction , , realized in the last years for optical fields in various media , the best results being obtained for coupling two optical beams by parametric interaction in a $`\chi ^{(2)}`$ crystal or using cold atoms in a trap .
Our scheme is depicted in Fig. 1. The nonlinear QND interaction produces the following coupling between two bright coherent fields $`E_a`$ and $`E_b`$
$`\stackrel{~}{E}_a`$ $`=`$ $`E_a+{\displaystyle \frac{g}{2}}\left(E_b^++E_b\right),`$ (1)
$`\stackrel{~}{E}_b`$ $`=`$ $`E_b+{\displaystyle \frac{g}{2}}\left(E_a^+E_a\right),`$ (2)
where $`\stackrel{~}{E}_a`$ and $`\stackrel{~}{E}_b`$ are the outgoing fields and $`g`$ is the QND gain. It is easy to find that the quadratures $`X=E^++E`$ and $`Y=i(E^+E)`$ of both fields are transformed as
$`\begin{array}{cc}\stackrel{~}{X}_a=X_a+gX_b,\hfill & \stackrel{~}{Y}_a=Y_a,\hfill \\ \stackrel{~}{X}_b=X_b,\hfill & \stackrel{~}{Y}_b=Y_bgY_a.\hfill \end{array}`$
The quadratures $`X_b`$ and $`Y_a`$ are the so-called QND variables, which are not affected by coupling. The conjugated quadratures $`X_a`$ and $`Y_b`$ become correlated with $`X_b`$ and $`Y_a`$ respectively and for infinitely large QND gain, or for infinitely squeezed $`Y_b`$ ($`X_a`$), the measurement of $`Y_b`$ ($`X_a`$) gives a result proportional to that of measurement of $`Y_a`$ ($`X_b`$). This is the main idea of QND measurement. In our approach this measurement is not performed, but the entanglement produced by QND coupling is used for quantum teleportation, as described below.
One of the resulting from QND interaction beams ($`\stackrel{~}{E}_a`$) passes to the sender Alice, and another one ($`\stackrel{~}{E}_b`$) passes to the receiver Bob for further reconstruction of the teleported state. The protocol of quantum teleportation is as follows. Alice superimposes the field $`E_{in}`$, which is to be teleported, with the field $`\stackrel{~}{E}_a`$ at a beam splitter with the transmittance $`ϵ`$, and, using homodyne photodetection, measures the $`X`$ quadrature of the reflected (for $`E_{in}`$) field and the $`Y`$ quadrature of the transmitted field, obtaining the values
$`x`$ $`=`$ $`ϵ^{1/2}\left(X_a+gX_b\right)+(1ϵ)^{1/2}X_{in},`$ (3)
$`y`$ $`=`$ $`ϵ^{1/2}Y_{in}(1ϵ)^{1/2}Y_a,`$ (4)
where the field variables are written in the Wigner representation. The measured values are sent to Bob via a classical channel, and upon receiving them Bob prepares the bright laser beam $`E_\beta `$ by means of phase and amplitude modulation in a coherent state with the amplitude $`\beta =\mathrm{\Gamma }G(x+igy)/2`$, where $`\mathrm{\Gamma }1`$ and $`G`$ are some real gains. After that Bob superimposes the field $`E_\beta `$ with the beam $`\stackrel{~}{E}_b`$ at a beam-splitter with very low transmittance $`ϵ^{}=\mathrm{\Gamma }^21`$, so that the reflected field $`\stackrel{~}{E}_b`$ is just shifted by $`G(x+igy)/2`$, but no noise is added. The result of interference of the reflected field $`\stackrel{~}{E}_b`$ and the transmitted field $`E_\beta `$ is the output $`E_{out}`$:
$`2E_{out}`$ $`=`$ $`2\mathrm{\Gamma }^1E_\beta 2\stackrel{~}{E}_b=G(x+igy)X_biY_b+igY_a`$ (5)
$`=`$ $`Gϵ^{1/2}X_a+(Ggϵ^{1/2}1)X_b+G(1ϵ)^{1/2}X_{in}`$ (7)
$`+ig\left(1G(1ϵ)^{1/2}\right)Y_aiY_b+iGgϵ^{1/2}Y_{in}.`$
Now, if the electric gain $`G`$ and the transmittance $`ϵ`$ satisfy for a given QND gain $`g`$ the relations $`G=(1ϵ)^{1/2}`$, $`g=(1ϵ)^{1/2}ϵ^{1/2}`$, the output field reads as
$$2E_{out}=X_{in}+iY_{in}+g^1X_aiY_b,$$
(8)
that is, for $`X_a=Y_b=0`$, it reproduces the incoming field with additional noise caused by quadratures $`X_a`$ and $`Y_b`$. For infinite QND gain $`g`$ and infinitely squeezed quadrature $`Y_b`$ the teleportation is ideal. However, even for finite QND gain and not squeezed $`Y_b`$ the proposed teleportation scheme works in an essentially quantum regime.
This can be seen from comparing this scheme to classical teleportation. In the latter case Alice and Bob do not have correlated ancillary systems and all they can do in order to accomplish a true reproduction of the average values of $`X_{in}`$ and $`Y_{in}`$, is to measure this quadratures with uncertainty imposed by the Heisenberg’s principle and create a coherent field with the average values of quadratures equal to measured ones. So the protocol of classical teleportation is as follows. Alice divides the incoming field $`E_{in}`$ in two parts by a 50:50 beam splitter and measures the $`X`$ quadrature of the reflected field and the $`Y`$ quadrature of the transmitted field, obtaining the values
$`x`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}X_{in}+{\displaystyle \frac{1}{\sqrt{2}}}X_v,`$ (9)
$`y`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}Y_{in}{\displaystyle \frac{1}{\sqrt{2}}}Y_v,`$ (10)
where $`E_v=\left(X_v+iY_v\right)/2`$ is the vacuum field at the unused port of the beam-splitter. The measured values are send to Bob who creates a coherent state
$`2E_{out}`$ $`=`$ $`\sqrt{2}(x+iy)+2E_w`$ (11)
$`=`$ $`X_{in}+iY_{in}+X_viY_v+X_w+iY_w,`$ (12)
where $`E_w=\left(X_w+iY_w\right)/2`$ is the vacuum fluctuation inevitably added at the reconstruction stage (e.g., as in the previous protocol, Bob can superimpose a strong coherent field $`2E_\beta =\mathrm{\Gamma }\sqrt{2}(x+iy)`$ with vacuum field $`E_w`$ at a low transmittance $`ϵ^{}=\mathrm{\Gamma }^2`$ beam splitter and use the reflected (for $`E_w`$) field as output). Now, to compare quantum (Eq. (8)) and classical (Eq. (11)) teleportation, we must calculate a quantitative measure of teleportation success, connected with some verification procedure. Let us consider the verification scheme shown in Fig. 2, similar to that of Ref. . The verifier Victor, independent on Alice and Bob, prepares an ensemble of states of field $`E_0`$. Victor can directly measure the average value and the variance of the field quadrature $`X_0`$. Or he can shift the phase of the field by random phase $`\phi (t)`$ and give it to Alice who teleports the state to Bob. After receiving the state $`E_{out}`$ from Bob, Victor again shifts the phase, this time by $`\phi (t)`$ and measures the $`X`$-quadrature of the obtained field $`E_T`$. If the teleportation is ideal, the average value and variance of $`X_T`$ will be the same as for $`X_0`$. The proposed in this Letter teleportation scheme gives the following expression for $`X_T`$:
$$X_T=X_0+g^1X_a\mathrm{cos}\phi Y_b\mathrm{sin}\phi .$$
(13)
So for $`X_a=Y_b=0`$ the average value of $`X_T`$ is equal to that of $`X_0`$. The variance of $`X_T`$ is given by
$`\left(\delta X_T\right)^2`$ $`=`$ $`\left(\delta X_0\right)^2+g^2\left(\delta X_a\right)^2\overline{\mathrm{cos}^2\phi }`$ (15)
$`+\left(\delta Y_b\right)^2\overline{\mathrm{sin}^2\phi },`$
where $`\delta X=XX`$ and the bar denotes averaging over the random phase shift. If the phase is distributed homogeneously in $`[0,2\pi ]`$, then $`\overline{\mathrm{cos}^2\phi }=\overline{\mathrm{sin}^2\phi }=\frac{1}{2}`$, and
$`N_{add}`$ $``$ $`\left(\delta X_T\right)^2\left(\delta X_0\right)^2`$ (16)
$`=`$ $`{\displaystyle \frac{1}{2g^2}}\left(\delta X_a\right)^2+{\displaystyle \frac{1}{2}}\left(\delta Y_b\right)^2,`$ (17)
where $`N_{add}`$ is the noise added to the teleported light. We use the added noise as a measure of teleportation success instead of commonly used fidelity $`F=\psi _{in}|\rho _{out}|\psi _{in}`$ (where $`|\psi _{in}`$ is the incoming state and $`\rho _{out}`$ is the density matrix of the output field), for two following reasons. Firstly, fidelity generally depends on the incoming state, while for the teleportation of field state, which is a linear in field transformation, the output field is always the incoming one plus some noise; so it is natural to describe the teleportation by this added noise, independent on the incoming state. Secondly, the output state can differ from the input one in two ways: in averages (displacement) and in variances of quadratures (distortion). Fidelity (as well as the parameters introduced in Ref. ) takes into account both these effects. However, classical teleportation always can be arranged so that the average values (of quadratures) are transported correctly. So only a measure of state distortion is necessary for characterizing the quantum nature of teleportation, and the added noise just provides such a measure.
It is easy to find from Eq. (11) that for classical teleportation
$`N_{add}^{cl}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\delta X_v\right)^2+{\displaystyle \frac{1}{2}}\left(\delta Y_v\right)^2+{\displaystyle \frac{1}{2}}\left(\delta X_w\right)^2`$
$`+{\displaystyle \frac{1}{2}}\left(\delta Y_w\right)^2`$ $``$ $`2`$
for any states of fields $`E_v`$ and $`E_w`$, which can be seen from the inequality $`a^2+b^22ab`$ and the Heisenberg’s uncertainty principle. So, to reach the quantum regime of teleportation, one must reach added noise $`0N_{add}<2`$, under condition that $`X_T=X_0`$ for any phase $`\phi `$. In the proposed scheme it can be obtained by using coherent fields $`E_a`$ and $`E_b`$ (with imaginary and real amplitudes respectively) and QND gain $`g>1/\sqrt{3}`$.
The entanglement produced by a QND device (or its quantum state preparation regime) is usually characterized by conditional variance of signal given a measured value of meter, $`V_c`$ . For transformation described by Eqs. (1), (2) and coherent input fields $`E_a`$ and $`E_b`$ this variance is given by
$$V_c\left(\delta \stackrel{~}{X}_b\right)^2\frac{|\delta \stackrel{~}{X}_a\delta \stackrel{~}{X}_b|^2}{\left(\delta \stackrel{~}{X}_a\right)^2}=\frac{1}{1+g^2}$$
(18)
and the condition $`g>1/\sqrt{3}`$ corresponds to the condition $`V_c<\frac{3}{4}`$. In a recent QND experiment using cold atoms in a trap as a nonlinear medium the value $`V_c=0.45`$ has been measured ($`V_c=0.37`$ with corrections for optical losses). In the same time, our requirements for field phases are exactly that needed for QND interaction based on cross-phase modulation, implemented in this experiment. For the QND schemes based on $`\chi ^{(2)}`$ crystals, a value $`V_c=0.650.7`$ has been achieved , without restrictions on the phases of interacting beams. These results speak about a very good possibility to reach the quantum regime of teleportation using the existing experimental technique and the protocol presented in this Letter.
Let us briefly compare our scheme to that using squeezed states , . In classical teleportation the output field is a sum of the input one and four additional noises (quantum duties), as shown by Eq. (11), and the noise added to a random quadrature is always more than 2. Quantum teleportation can be realized by using correlated ancillary fields $`E_v`$ and $`E_w`$, produced by mixing two squeezed fields $`E_1`$ and $`E_2`$ at a 50:50 beam-splitter, so that in Eq. (11) $`X_v+X_w=\sqrt{2}X_1`$ and $`Y_vY_w=\sqrt{2}Y_2`$, and the added noise is given by
$$N_{add}^{sq}=V_1+V_2,$$
(19)
where $`V_1`$ and $`V_2`$ are variances of quadratures $`X_1`$ and $`Y_2`$ respectively. When the quadratures $`X_1`$ and $`Y_2`$ are squeezed, the added noise becomes less than the classical value of 2. In our approach the ancillary fields are correlated via a QND interaction. Two QND variables ($`X_b`$ and $`Y_a`$) are automatically cancelled in the output field due to QND correlations. One of the remaining noises ($`X_a`$) can be made as small as desired by increasing the QND gain and/or by squeezing this quadrature. The noise of $`Y_b`$ can be reduced by squeezing this quadrature only. The added noise in this case reads as ($`V_c<1`$)
$$N_{add}^{QND}=\frac{1}{2}\frac{V_c}{1V_c}V_a+\frac{1}{2}V_b,$$
(20)
where $`V_a`$ and $`V_b`$ are variances of quadratures $`X_a`$ and $`Y_b`$ respectively. Comparing Eq. (20) with Eq. (19) we see that with decreasing conditional variance the QND coupling decreases significantly the requirements for the degree of squeezing of ancillary beams (so that for $`V_c<\frac{3}{4}`$ they can be not squeezed at all).
It should be noted that a principal possibility to use QND coupling for quantum teleportation has been pointed in Ref. . However this paper considers only QND gain $`g=1`$ and two ideally squeezed ancillary beams, in which case no benefits from QND coupling can be seen, compared to mixing these beams at a beam-splitter.
In summary, we have described a protocol for quantum teleportation of a single mode optical field using QND coupling as a source of entanglement of ancillary laser beams. In our scheme the QND device is used in a quantum state preparation regime, and the recent achievements for realizing this regime satisfy very well the requirements for quantum teleportation. From the other side, the existing schemes for quantum teleportation of single-mode field state are limited by the restricted degree of squeezing, available at present, and using QND coupling for squeezed ancillary beams can provide further improvement of quantum teleportation technique.
This work was supported by INTAS, grant # 96-0167.
|
no-problem/9908/math-ph9908023.html
|
ar5iv
|
text
|
# Singularity theory study of overdetermination in models for L–H transitions
## The SH model:
This may be expressed as the dimensionless bifurcation equation
$$\begin{array}{ccc}\hfill g(u,q,d_a)& =& \left(qd_au^21\right)\left(q+um\left(u\right)\right),\hfill \\ \hfill m(u)& =& u^p(b+au^{1p}),\hfill \end{array}$$
(1)
obtained by eliminating in the steady state the two other dynamical variables $`f`$ and $`k`$ in favor of $`u`$ the potential energy of the pressure gradient. The control parameter $`q`$ is the power input, $`d_a`$ is the reciprocal of the anomalous diffusivity, and $`m(u)`$ is the anomalous viscosity. In Sugama and Horton’s numerical work $`d_a`$ was set to 1, $`p`$ was given values of $`3/2`$ (case A) and $`1`$ (case B), and $`a`$ and $`b`$ were given as positive numerical factors. (Note: The dynamical equations also contain a parameter $`c`$ which cancels from Eq. (1).) Figure 1 shows the bifurcation diagrams for case A and case B. (In all diagrams stable solutions are indicated by solid lines, unstable solutions by dashed lines, and branches of limit cycles by dotted lines marking the maximum and minimum amplitude.) It was assumed that the transition from the lower stable solution branch (L-mode) to the upper stable branch (H-mode) must occur at the singular point $`A`$, where the steady-state shear flow kinetic energy $`f=\left(u^2d_aq\right)/cu`$ becomes unphysical. The transition is discontinuous for case A and continuous for case B. The H-mode branch becomes unstable at a Hopf bifurcation to stable limit cycles, identified as ELMs. The SH model thus predicts hysteresis of the L–H transition and oscillating and quiescent H-modes, which accords with recent experimental observations . However, the derivative discontinuity at $`A`$ is problematic. For case A the transition was described as first-order, but it occurs at what appears to be a highly degenerate point. For case B the transition was described as second-order. It should also be noted that the singular point $`A`$ is persistent to variations in $`d_a`$, $`a`$, and $`b`$. For these reasons we suspect that there may not be enough independent parameters in the model. Solution of the recognition problem, step (2), indicates that the model may be overdetermined as a bifurcation problem.
Proposition 1.—Equation (1) with $`d_a=d_{a0}`$, $`p<1`$ is a germ that is strongly equivalent to the normal form
$$h(x,\lambda )=x^3+\lambda x.$$
(2)
(The term “germ” is explained as follows: two functions $`g_1(x,\lambda )`$ and $`g_2(x,\lambda )`$ are equal as germs if they coincide on some neighborhood of a fixed point $`x_0,\lambda _0`$.) Proof.—We apply the following theorem, adapted from : Theorem.—A germ $`g(x,\lambda )`$ is strongly equivalent to the normal form $`h(x,\lambda )=\epsilon x^3+\delta \lambda x`$ if and only if, at the fixed point $`(x_0,\lambda _0)`$,
$`g=g_x`$ $`=`$ $`g_{xx}=g_\lambda =0,g_{xxx}0,g_{\lambda x}0`$ (3)
where $`\epsilon =\text{sgn}g_{xxx}`$, $`\delta =\text{sgn}g_{\lambda x}`$. In Eq. (1) we identify the state variable $`ux`$ and the distinguished parameter $`q\lambda `$ and evaluate the defining and non-degeneracy conditions (3) at the point $`A=(u_0,q_0)`$. We find that $`g=g_u=g_q=0`$, $`g_{uu}=4a\left(1+p\right)4\left(1+p\right)/d_a=0`$ for $`d_a=d_{a0}=\left(1+p\right)/a\left(1+p\right)`$, and $`g_{uuu}=12\left(1+p\right)/u_0d_{a0}`$, $`g_{uq}=2/u_0`$. Equation (2) for the normal form is inferred. It is the prototypic pitchfork , a codimension 2 bifurcation which requires two auxiliary parameters for an unfolding that contains, to qualitative equivalence, all possible perturbations of $`g`$. We see that the defining conditions for the point $`A`$ yield a system of four algebraic equations in what is effectively only two variables — $`u`$ and $`q`$. To resolve the overdetermination we propose a universal unfolding of Eq. (1).
Proposition 2.—The bifurcation function
$$G(u,q,d_a,\alpha )=g(u,q,d_a)+\alpha $$
(4)
is a universal unfolding of the germ (1) for $`p<1`$. It is equivalent to the prototypic universal unfolding of the pitchfork $`G(x,\lambda ,\alpha ,\beta )=x^3+\beta x^2+\lambda x+\alpha ,`$ where $`d_a=d_{a0}\pm \beta `$. The proof is not presented here; instead we focus on the qualitative consequences. (The physical interpretation of the unfolding parameter $`\alpha `$ is discussed below.) Specifically, Eq. (4) encapsulates the generic behavior of the SH system. The four qualitatively distinct bifurcation diagrams are shown in Fig. 2(a)–(d), of which (a) and (b) are physically relevant because $`\alpha <0`$ leads to dynamical violation of the condition $`f0`$. In (a) the L–H and H–L transitions occur at non-degenerate limit points. No marked transition to H-mode occurs at all in (b). Now it can be seen why the unperturbed bifurcation set, Fig. 2(e), and the partially perturbed bifurcation set, Fig. 1(a), cannot predict the results of experiments. The singularity that exists in these sets (point $`A`$) is not even present when $`\alpha `$ is nonzero. We also see that changes in the auxiliary parameters around the critical values can lead to incomparably different bifurcation behavior.
What of case B? Proposition 3.—Equation (1) with $`p=1`$ is a germ that is strongly equivalent to the normal form $`h(x,\lambda )=x^2+\lambda ^2,`$ a codimension 1 bifurcation known as the transcritical bifurcation. Proof.—We apply the following theorem from : Theorem 2.—A germ $`g(x,\lambda )`$ is strongly equivalent to the normal form $`h(x,\lambda )=\epsilon \left(x^2\lambda ^2\right)`$ if and only if, at the fixed point $`(x_0,\lambda _0)`$,
$$g=g_x=g_\lambda =0,g_{xx}0,\mathrm{d}et\left(\begin{array}{cc}g_{xx}& g_{\lambda x}\\ g_{\lambda x}& g_{\lambda \lambda }\end{array}\right)<0$$
(5)
where $`\epsilon =\mathrm{s}gng_{xx}`$. These conditions in Eq. (1) yield $`g=g_u=g_q=0`$, $`g_{uu}=8a`$, $`detd^2g=4\left(ad_a1\right)^2/u^2`$. Equation (4) in this special, fragile case is a one-parameter universal unfolding, indifferent to the value of $`d_a`$. It yields two qualitatively distinct bifurcation diagrams, shown in Fig. 3. Note that the bifurcation structure here excludes the possibility of hysteresis.
## The LDGC model:
The steady states are summarized in the bifurcation diagram of Fig. 4, where the control parameter $`\varphi `$ is the particle flux and $`p`$ is the pressure gradient. The lower stable branch is identified as L-mode. At $`A`$ the transition to the intermediate stable branch $`AB`$, identified as H-mode, is described by Lebedev et al as analogous to a second-order phase transition. At $`B`$ the system moves onto the $`p=1`$ branch in another continuous transition, but is said to remain in H-mode. The first Hopf bifurcation initiates a branch of unstable limit cycles and the second terminates a branch of stable limit cycles, identified as ELMs. The point $`C`$ is the intersection of the $`p=1`$ branch and the unstable $`AC`$ branch. Near $`B`$ and $`C`$ the bifurcation equations may be written, respectively, as
$`g_B(p,\varphi )`$ $`=`$ $`\gamma \left(\varphi \stackrel{~}{d}\mu p\right)\left(p1\right)/p\left(\stackrel{~}{d}\stackrel{~}{d}_m\right),\text{and}`$ (6)
$`g_C(p,\varphi )`$ $`=`$ $`\gamma \left(p^2\stackrel{~}{d}\varphi \right)\left(p1\right)/p\stackrel{~}{d}_m.`$ (7)
As before, we use the singularity theory analysis to focus on qualitative structure. Using theorem 2 we find that at points $`B`$ and $`C`$ there is a transcritical bifurcation, which requires the single auxiliary parameter $`\alpha ^{}`$ for a universal unfolding. The two qualitatively distinct bifurcation diagrams are shown in Fig. 5. In (a) a branch of stable limit cycles connects the two stable stationary branches. In (b) the branches of stable stationary solutions are unconnected. The structure of the limit cycle branch implies that (on a phase plane) a stable orbit is surrounded by an unstable orbit. The point $`A`$ in Fig. 4 clearly is not unfolded by the one-parameter perturbation. Somewhat surprisingly, it is the limit point of the branch of the branch $`CAB`$, which actually coincides along $`AB`$ with the continuous branch $`0AB`$. (This result is detailed elsewhere.) A limit point is its own universal unfolding, i.e., persistent to small perturbations.
## In summary:
(1) The SH model in general is a codimension 2 bifurcation problem, containing a pitchfork, that requires two unfolding parameters for a universal unfolding and hence complete determination. The critical values of the unfolding parameters $`\alpha `$ and $`d_a`$ are respectively $`0`$ and $`\left(1+p\right)/a\left(1+p\right)`$, $`p<1`$. (2) The LDGC model is a codimension 1 bifurcation problem, containing two transcritical bifurcations. A universal unfolding is provided by a single auxiliary parameter $`\alpha ^{}`$. The two models are therefore structurally dissimilar in general form. However, the fact that they describe the same phenomena suggests that the LDGC model may be a partially collapsed codimension 2 system, and in a forthcoming work we show that this is indeed the case. A fortiori we can also say that second-order phase transitions in these systems, if they exist, could only be observed on variation of at least two parameters simultaneously. In many bifurcation problems pitchforks occur in the presence of $`𝐙_2`$ equivariance in the governing equations for the system, that manifests as a physically invariant property. (A function $`\varphi (x)`$ has $`𝐙_2`$ symmetry if $`\varphi (x)=\varphi (x)`$.) A symmetry arises in the dynamical equations for the SH model because the shear of the poloidal flow $`v^{}`$ is invariant under the transformation $`v^{}v^{}`$. The unfolding parameter $`\alpha `$ can therefore be interpreted as a symmetry-breaking term, representing an intrinsic energy (or angular momentum) generation rate that occurs even in a pressure gradient of zero. The $`𝐙_2`$ invariance of the flow shear is not evident in the bifurcation structure of the LDCG model, and $`\alpha ^{}`$ represents a perturbation of the MHD turbulence level.
Other models for L–H transitions that have multiple solutions include those where the flow shear is due to ion-orbit losses on the plasma edge or magnetic field ripple induced particle flux in the core . We feel that singularity theory could play an important role in developing and unifying these models and elucidating the physics of L–H transitions. There is a reasonable expectation that different models, if they appeal to the same general physics, should belong to the same qualitative universality class even though they may differ quantitatively. A wider question is whether the dynamics of infinite-dimensional systems can be approximated by low-dimensional systems such as these. The practical advantages are obvious, and developments in inertial manifold theory have shown that the long-time-scale behavior of infinite-dimensional dissipative systems can occur in a defined finite-dimensional subspace.
Acknowledgment: we are grateful to Professor Jeffrey Harris for helpful discussions about this work.
|
no-problem/9908/cond-mat9908363.html
|
ar5iv
|
text
|
# Theory of PbTiO3, BaTiO3, and SrTiO3 Surfaces
## I Introduction
The surfaces of insulating cubic perovskite materials such as PbTiO<sub>3</sub>, BaTiO<sub>3</sub>, and SrTiO<sub>3</sub> are of interest from several points of view. First, some of these materials (notably SrTiO<sub>3</sub>) are very widely used as substrates for growth of other oxide materials (e.g., layered high-$`T_c`$ superconductors and “colossal magnetoresistance” materials). Second, this class of materials is of enormous importance for actual and potential applications that make use of their unusual piezoelectric, ferroelectric, and dielectric properties (e.g., for piezoelectric transducers, non-volatile memories, and wireless communications applications, respectively). Many of these applications are increasingly oriented towards thin-film geometries, where surface properties are of growing importance. Third, the bulk materials display a variety of structural phase transitions; the ferroelectric (FE) structural phases are of special interest, but antiferroelectric (AFE) or antiferrodistortive (AFD) transitions can also take place. It is then of considerable fundamental interest to consider how these structural distortions couple to the surface, e.g., whether the presence of the surface acts to enhance or suppress the structural distortion. The ferroelectric properties are well known to degrade in thin-film and particulate geometries, and it is very important to understand whether such behavior is intrinsic to the presence of a surface, or whether it arises from extrinsic factors such as compositional non-uniformities or structural defects in the surface region. Finally, the cubic perovskites can serve as model systems for the study of transition-metal oxide surfaces more generally.
In the last decade, there has been a surge of activity in the application of first-principles computational methods based on density-functional theory (DFT) to the study of the bulk properties, and especially the ferroelectric transitions, in bulk perovskite oxides. (For a recent review, see Ref. or .) The importance of these methods was recently underlined by the award of the Nobel Prize in Chemistry to Walter Kohn, the primary originator of DFT. In the materials theory community, these methods have been widely used for two decades to predict properties of semiconductors and simple metals. However, recent advances in computational algorithms and computer power now allow these methods to be applied to more complex materials (e.g., perovskites) and more complex geometries (e.g., defects and surfaces). In particular, pioneering studies of BaTiO<sub>3</sub> and SrTiO<sub>3</sub> surfaces have recently appeared.
Experimental investigations of the surface structure of cubic perovskites have not been very extensive. Such studies are hindered by the difficulties of preparing clean and defect-free surfaces, and of overcoming charging effects associated with many experimental probes. Even for SrTiO<sub>3</sub>, the best-studied of these surfaces, there is a disappointing level of agreement among experimental results and between experiment and theory. We are not aware of comparable studies of BaTiO<sub>3</sub> and PbTiO<sub>3</sub> surfaces.
The purpose of the present contribution is to present new theoretical work on the structural properties of the PbTiO<sub>3</sub> (001) surface, and to compare and contrast these results with the previous work of our group on BaTiO<sub>3</sub> and SrTiO<sub>3</sub> surfaces. As regards bulk properties, lead-based compounds such as PbTiO<sub>3</sub> and PbZrO<sub>3</sub> are known to behave quite differently from alkaline-earth based perovskites such as BaTiO<sub>3</sub> and SrTiO<sub>3</sub>. Previous theoretical work has shown that the FE distortion is typically larger and that Pb atoms participate much more strongly in (and sometimes even dominate) the FE distortion, compared with non-Pb perovskites. Moreover, the Pb-based compounds are generally more susceptible to more complex AFD and AFE instabilities involving tilting of the oxygen octahedra, and the ground-state structures often involve the formation of some quite short Pb–O bonds. All of these effects point to a strong and active involvement of the Pb atoms in the bonding, most naturally interpreted in terms of the formation of partially covalent Pb–O bonds with the closest oxygen neighbors. Finally, a focus on Pb-based materials is motivated by the fact that these are the leading candidates for many practical piezoelectric and switching applications, especially in the form of solid solutions such as PZT (PbZr<sub>x</sub>Ti<sub>1-x</sub>O<sub>3</sub>), PMN (PbMg<sub>1/3</sub>Nb<sub>2/3</sub>O<sub>3</sub>), and PZN (PbZn<sub>1/3</sub>Nb<sub>2/3</sub>O<sub>3</sub>).
The manuscript is organized as follows. Section II contains a brief account of the technical details of the work, including the theoretical methods used, the slab geometries studied, and the formulation of the surface energy. In Sec. III we present the computed structural relaxations of the PbTiO<sub>3</sub> surfaces, and compare these to the previous results on BaTiO<sub>3</sub> and SrTiO<sub>3</sub> surfaces. Additionally, we discuss the surface energetics (surface energies and surface relaxation energies), and point out some characteristic differences in the surface electronic structure of the three compounds. Finally, the paper ends with a summary in Sec. IV.
## II Preliminaries
### A Theoretical Methods
We carried out self-consistent plane-wave pseudopotential calculations within Kohn-Sham density-functional theory using a conjugate-gradient technique. Exchange and correlation were treated using the Ceperley-Alder form. Vanderbilt ultrasoft pseudopotentials were employed, with semicore Pb $`5d`$, Ba $`5s`$ and $`5p`$, Sr $`4s`$ and $`4p`$, and Ti $`3s`$ and $`3p`$ orbitals included as valence states. A plane-wave cutoff of 25 Ry has been used throughout. Relaxations of atomic coordinates are iterated until the forces are less than 0.01 eV/Å. Justification of the convergence and accuracy of this approach can be found in the previously published work.
### B Surface and Slab Geometries
In this work we consider only II-IV cubic perovskites, i.e., ABO<sub>3</sub> perovskites in which atoms A and B are divalent and tetravalent, respectively. In this case, two non-polar (001) surface terminations are possible: the AO–terminated surface, and the BO<sub>2</sub>–terminated surface.
We have studied both types of surface termination for all three materials (PbTiO<sub>3</sub>, BaTiO<sub>3</sub>, and SrTiO<sub>3</sub>) using a repeated slab geometry. The slabs are symmetrically terminated and typically contain seven layers (17 or 18 atoms), as illustrated in Fig. 1. The vacuum region was chosen to be two lattice constants thick. The calculations were done using a (4,4,2) Monkhorst-Pack mesh, corresponding to three or four k-points in the irreducible Brillouin zone for cubic and tetragonal surfaces respectively. The convergence of the calculations has been very carefully checked for PbTiO<sub>3</sub> by repeating some of the calculations with asymmetrically terminated eight-layer slabs and symmetrically terminated nine-layer slabs. Additionally, we have enlarged the vacuum region to a thickness of three lattice constants, and we have checked the convergence of the Brillouin zone integration by going to a (6,6,2) k-point mesh. In all cases, the results for the structural properties of the surfaces given in the Tables I to V change by less than 0.2%.
For all three materials, we first computed the relaxations for the “cubic” surface, i.e., for the case where there is no symmetry lowering relative to a slab of ideal cubic material. In this case we preserved $`M_x`$, $`M_y`$, and $`M_z`$ mirror symmetries relative to the center of the slab, and set the lattice constants in the $`\widehat{x}`$ and $`\widehat{y}`$ directions equal to those computed theoretically for the corresponding bulk material (3.89 Å, 3.95 Å, and 3.86 Å for PbTiO<sub>3</sub>, BaTiO<sub>3</sub>, and SrTiO<sub>3</sub>, respectively). The symmetry-allowed displacements of the atoms in the $`z`$ (surface-normal) direction were then fully relaxed.
Each of the three materials studied displays a different sequence of structural phase transitions from the cubic paraelectric phase as the temperature is lowered. PbTiO<sub>3</sub> undergoes a single transition into a tetragonal ferroelectric (FE) phase at 763 K and then remains in this structure down to zero temperature. BaTiO<sub>3</sub> displays a series of three transitions to tetragonal, orthorhombic, and rhombohedral FE phases at 403 K, 278 K, and 183 K, respectively. SrTiO<sub>3</sub> remains cubic down to 105 K, at which point it undergoes an antiferrodistortive transition involving rotation of the oxygen octahedra and doubling of the unit cell. The material nearly goes ferroelectric at about $`T=30`$ K, but is evidently prevented from doing so by quantum zero-point fluctuations.
Because we are primarily interested in the room-temperature structures of these materials and their surfaces, we have chosen to focus on the tetragonal FE phases of PbTiO<sub>3</sub> and BaTiO<sub>3</sub> for our surface studies. We consider only the case of the tetragonal $`c`$ axis (i.e., polarization) lying parallel to the surface, since polarization normal to the surface is strongly suppressed by the depolarization fields that would arise from the accumulated charge at the surfaces. We take the tetragonal axis to lie along $`\widehat{x}`$, and relax the $`M_x`$ symmetry while retaining the $`M_y`$ and $`M_z`$ symmetries with respect to the center of the slab. For PbTiO<sub>3</sub>, which is tetragonal at $`T=0`$, this will indeed be the ground-state structure of the slab. For BaTiO<sub>3</sub>, on the other hand, the $`M_y`$ symmetry is artificially imposed so that the theoretical $`T=0`$ calculation will mimic the experimental room-temperature surface structure. In both cases, the slab lattice constants in the $`\widehat{x}`$ and $`\widehat{y}`$ directions were set equal to the corresponding theoretical equilibrium lattice constants computed for the bulk tetragonal phase: $`c`$=4.04 Å and $`a`$=3.86 Å for PbTiO<sub>3</sub>, and $`c`$=3.99 Å and $`a`$=3.94 Å for BaTiO<sub>3</sub>.
### C Surface Energies
A comparison of the relative stability of the AO and TiO<sub>2</sub> surface terminations is problematic because the corresponding surface slabs contain different numbers of AO and TiO<sub>2</sub> formula subunits. We treat this problem by introducing chemical potentials $`\mu _{\mathrm{AO}}`$ and $`\mu _{\mathrm{TiO}_2}`$ for these subunits, defined in such a way that $`\mu _{\mathrm{AO}}=0`$ and $`\mu _{\mathrm{TiO}_2}=0`$ correspond to thermal equilibrium with bulk crystalline AO and TiO<sub>2</sub>, respectively. We have computed the cohesive energies $`E_{\mathrm{AO}}`$ and $`E_{\mathrm{TiO}_2}`$ of crystalline PbO, BaO, SrO, and TiO<sub>2</sub> using the same first-principles pseudopotential method in order to provide these reference values. The grand potential for a given surface structure can then be computed as
$$F_{\mathrm{surf}}=\frac{1}{2}[E_{\mathrm{slab}}N_{\mathrm{TiO}_2}(E_{\mathrm{TiO}_2}+\mu _{\mathrm{TiO}_2})N_{\mathrm{AO}}(E_{\mathrm{AO}}+\mu _{\mathrm{AO}})],$$
(1)
where $`N`$ is the number of formula subunits contained in the slab, and the factor of $`1/2`$ accounts for the fact that each slab contains two surfaces. Assuming that the surface of the ATiO<sub>3</sub> is in equilibrium with its own bulk, it follows that
$$\mu _{\mathrm{AO}}+\mu _{\mathrm{TiO}_2}=E_\mathrm{f},$$
(2)
where $`E_\mathrm{f}`$ is the heat of formation of bulk ATiO<sub>3</sub> from bulk AO and bulk TiO<sub>2</sub>. The two chemical potentials are thus not independent, and we choose to treat $`\mu _{\mathrm{TiO}_2}`$ as the independent variable when presenting our results. Accordingly, $`\mu _{\mathrm{TiO}_2}`$ is allowed to vary over the range
$$E_\mathrm{f}\mu _{\mathrm{TiO}_2}0,$$
(3)
the lower and upper limit corresponding to the precipitation of particulates of AO and TiO<sub>2</sub> on the surface, respectively.
## III Results and Discussions
### A Structural relaxations
We begin by presenting our new results on the structural properties of the PbTiO<sub>3</sub> (001) surfaces. The equilibrium atomic positions for both surface terminations in the two phases were obtained by starting from the ideal structures of the surfaces and then relaxing the atomic positions while preserving the symmetries described in section II B. The results for the fully relaxed geometries are summarized in Table I and II. By symmetry, there are no forces along $`\widehat{x}`$ and $`\widehat{y}`$ for the cubic surface, and no forces along $`\widehat{y}`$ for the tetragonal surface.
Tables I and II show for both surfaces a substantial inward contraction towards the bulk for the uppermost surface layers, whereas for the second layers we find an outward relaxation of the atoms relative to the positions of the atoms on the ideal surface. Generally, the metal and the oxygen atoms move in the same direction, but the relaxations of the metal atoms are much larger, leading to a rumpling of the layers. The single exception is the surface layer of the tetragonal TiO<sub>2</sub>–terminated surface, where one of the two oxygen atoms moves in the opposite direction to the metal atom. Therefore we can see here a significant asymmetry between the O atoms with respect to their positions perpendicular to the surface. This asymmetry between the oxygen atoms in the topmost surface layer of the tetragonal TiO<sub>2</sub>–terminated surface was also found for BaTiO<sub>3</sub> but with a much smaller amplitude. As expected, we find the largest relaxations for the surface–layer atoms, but the displacement of the Pb atom in the second layer of the TiO<sub>2</sub>–terminated surface is of the same magnitude.
In order to compare these results with previous calculations for SrTiO<sub>3</sub> and BaTiO<sub>3</sub>, we have calculated the changes in the interlayer distances $`\mathrm{\Delta }d_{ij}`$ and the amplitudes of the rumpling $`\eta _i`$ of the layers in the surface slabs for all three perovskites. The results for both surface terminations and the different phases are given in the Tables III and IV. We denote the change in the $`z`$ position of a
metal atom relative to the ideal unrelaxed structure as $`\delta _z(\mathrm{M})`$, and $`\delta _z(\mathrm{O})`$ is the same for the oxygen atom in the same layer (defined as $`[\delta _z(\mathrm{O}_\mathrm{I})+\delta _z(\mathrm{O}_{\mathrm{II}})]/2`$ for a TiO<sub>2</sub> layer). We then define the change of the interlayer distance $`\mathrm{\Delta }d_{ij}`$ as the difference between the averaged atomic displacements $`[\delta _z(\mathrm{M})+\delta _z(\mathrm{O})]/2`$ of layer $`i`$ and $`j`$, and the rumpling $`\eta _i`$ is defined as the amplitude of these displacements $`|\delta _z(\mathrm{M})\delta _z(\mathrm{O})|`$. From Tables III and IV we can see that, for all three perovskites and for both terminations, the surfaces display a similar oscillating relaxation pattern with a reduction of the interlayer distance $`d_{12}`$, an expansion of $`d_{23}`$ and again a reduction for $`d_{34}`$. However, compared to BaTiO<sub>3</sub> and SrTiO<sub>3</sub>, the amplitudes of the relaxations in PbTiO<sub>3</sub> are significantly increased.
The second interesting feature of Tables III and IV is that for BaTiO<sub>3</sub>, there is almost no difference in the relaxations of the surface layers between the cubic and the tetragonal phase. The same is true for the TiO<sub>2</sub>–terminated surface of PbTiO<sub>3</sub>. For the PbO–terminated surface, in contrast, the changes in the interlayer distances and the layer rumplings are strongly reduced in the tetragonal phase. We will come back to this point at the end of the next subsection.
### B Influence of the surface upon ferroelectricity
We turn now to the question of whether the presence of the surface has a strong effect upon the near–surface ferroelectricity. To analyze whether the ferroelectric order is enhanced or suppressed near the surface, we introduce average ferroelectric distortions $`\delta _{\mathrm{FE}}`$ for each layer of the surface slabs:
$$\begin{array}{cccc}\hfill \delta _{\mathrm{FE}}& =& |\delta _x(\mathrm{A})\delta _x(\mathrm{O}_{\mathrm{III}})|\hfill & \text{for AO planes and}\hfill \\ \hfill \delta _{\mathrm{FE}}& =& |\delta _x(\mathrm{Ti})[\delta _x(\mathrm{O}_\mathrm{I})+\delta _x(\mathrm{O}_{\mathrm{II}})]/2|\hfill & \text{for TiO}\text{2}\text{ planes.}\hfill \end{array}$$
(4)
The calculated values of $`\delta _{\mathrm{FE}}`$ for BaTiO<sub>3</sub> and PbTiO<sub>3</sub> are given in Table V; the last row of the table gives the bulk values for reference.
For the PbO–terminated surface of PbTiO<sub>3</sub>, one can see a clear increase in the average ferroelectric distortions $`\delta _{\mathrm{FE}}`$ when going from the bulk values to the surface layer. On the other hand, for the TiO<sub>2</sub>–terminated surface, the average distortions are slightly decreased at the surface. Surprisingly, this is just the opposite of what one observes for BaTiO<sub>3</sub>, where one sees a reduction of the ferroelectric distortions for the BaO–terminated surface and a moderate enhancement for the TiO<sub>2</sub>–terminated surface. (Of course, the distortions are also much smaller for BaTiO<sub>3</sub> surfaces, as they are in the bulk, compared to PbTiO<sub>3</sub>.) These results tend to confirm that Pb is a much more active constituent in PbTiO<sub>3</sub> than is Ba in BaTiO<sub>3</sub>, presumably because of the partially covalent nature of the Pb–O bonds as discussed in Sec. I.
In any case, the present results again confirm that the presence of the surface does not lead to any drastic suppression of the ferroelectric order near the surface, supporting the view that extrinsic effects must be responsible for degradation of ferroelectricity in thin-film geometries.
Finally, we note that there are interesting signs of interplay between the relaxations parallel and perpendicular to the surface for PbTiO<sub>3</sub>. In particular, the relaxations perpendicular to the surface are substantially reduced (by a factor of $``$3) on the PbO-terminated surface when going from the cubic to the tetragonal case. This can be rationalized as follows. Because of the partial covalency of the Pb–O bonds, there is a tendency to reduce the Pb–O bond length (this length is 2.75, 2.51, and 2.30 Å in cubic PbTiO<sub>3</sub>, tetragonal PbTiO<sub>3</sub>, and PbO, respectively). For the cubic surface, by symmetry, the only possibility to shorten this bond length is by a strong movement of the Pb atom towards the bulk and a strong movement upwards of the O atoms in the second layer. This leads to the strong rumpling and the decrease of $`d_{12}`$. But in the tetragonal phase there is also the possibility to enlarge the ferroelectric distortion in order to shorten the Pb–O bond length. Evidently, the enlargement of the ferroelectric distortion is preferred to the relaxation perpendicular to the surface.
### C Surface energies
In this section we discuss the surface energetics of the three perovskite compounds. In order to compare the relative stability of the AO– and TiO<sub>2</sub>–terminated surfaces, we have calculated the grand thermodynamic potential $`F_{\mathrm{surf}}`$ (as introduced in Sec. II C) for the different surfaces as a function of the chemical potential $`\mu _{\mathrm{TiO}_2}`$. The results for the tetragonal surfaces of BaTiO<sub>3</sub> and PbTiO<sub>3</sub> are shown in Fig. 3. The graphs of the grand thermodynamic potentials for the SrTiO<sub>3</sub> surfaces are very similar to those of BaTiO<sub>3</sub> and are therefore not shown separately.
Figure 3 shows a very different behavior for the BaTiO<sub>3</sub> and PbTiO<sub>3</sub> surfaces. First of all, the formation energy $`E_\mathrm{f}`$ of PbTiO<sub>3</sub> (when formed from bulk PbO and TiO<sub>2</sub>) is 0.36 eV, much lower than the formation energies of SrTiO<sub>3</sub> and BaTiO<sub>3</sub> which are about 3.2 eV. This leads to a much smaller range for the chemical potential $`\mu _{\mathrm{TiO}_2}`$ for which PbTiO<sub>3</sub> surfaces can grow in thermodynamic equilibrium. Second, for BaTiO<sub>3</sub> the two different surfaces have a comparable range of thermodynamic stability, indicating that either BaO–terminated surfaces or TiO<sub>2</sub>–terminated surfaces could be formed depending on whether growth occurs in Ba–rich or Ti–rich conditions. In contrast, for PbTiO<sub>3</sub> only the PbO–terminated surface can be obtained in thermodynamic equilibrium.
To get a quantity describing the surface energetics that is independent of the chemical potential $`\mu _{\mathrm{TiO}_2}`$ and therefore allows a more direct comparison of the three compounds, we define the average surface energy per surface unit cell
$$E_{\mathrm{surf}}=\frac{1}{4}\left(E_{\mathrm{slab}}^{\mathrm{AO}}+E_{\mathrm{slab}}^{\mathrm{TiO}_2}7E_{\mathrm{bulk}}\right),$$
(5)
which is equal to the average of the grand thermodynamic potential $`F_{\mathrm{surf}}`$ for the two kinds of surfaces. Again, the results for $`E_{\mathrm{surf}}`$ shown in Table VI are very similar for SrTiO<sub>3</sub> and BaTiO<sub>3</sub>, whereas the value for PbTiO<sub>3</sub> is significantly lower.
Finally we have computed the average relaxation energy $`E_{\mathrm{relax}}`$ of the three perovskite compounds. $`E_{\mathrm{relax}}`$ is defined as the difference between the average surface energy $`E_{\mathrm{surf}}`$ of the ideal surface without relaxation of the atoms, and the fully relaxed surfaces. The largest and smallest value for $`E_{\mathrm{relax}}`$ (see Table VI) were found for PbTiO<sub>3</sub> and BaTiO<sub>3</sub>, respectively, which is in agreement with the observation that the atomic relaxations are largest in PbTiO<sub>3</sub> and smallest in BaTiO<sub>3</sub>.
For all three compounds the average relaxation energy $`E_{\mathrm{relax}}`$ is many times larger than a typical bulk ferroelectric well depth, which is approximately 0.03 eV for BaTiO<sub>3</sub> and 0.05 eV for PbTiO<sub>3</sub>. This would indicate that the surface is capable of acting as a strong perturbation on the ferroelectric order. As we have shown in Sec. III B, this is not the case for BaTiO<sub>3</sub> and PbTiO<sub>3</sub>. One reason why the ferroelectric order is not as strongly affected by the surface as one might have thought has been pointed out in Ref. : the soft phonon eigenmode, which is responsible for the ferroelectric distortion, is only one of three zone center modes having the same symmetry. By looking at how strongly the surface relaxations are related to each of these zone center modes it has turned out that the distortions induced by the presence of the surface are to a large extent of non–ferroelectric character.
### D Surface band structure
For all three perovskite compounds we have carried out LDA calculations of the bulk and the surface electronic structure for our various surface slabs. It is well known that the LDA is quantitatively unreliable regarding excitation properties such as band gaps. Since we are in the following only looking at differences between band structures, we think that our conclusions drawn from the LDA results are nevertheless qualitatively correct.
As has already been shown in Ref. , the bulk band structures of SrTiO<sub>3</sub> and BaTiO<sub>3</sub> are very similar, whereas PbTiO<sub>3</sub> shows some significant differences. In SrTiO<sub>3</sub> and BaTiO<sub>3</sub> the upper edge of the valence band is very flat throughout the Brillouin zone. On the other hand, in PbTiO<sub>3</sub> the shallow $`6s`$ semicore states of the Pb atoms hybridize with the $`2p`$ states of the O atoms, leading to a lifting of the upper valence band states near the X point of the Brillouin zone.
This fact is responsible for a different behavior of the PbTiO<sub>3</sub> surface band structure compared to SrTiO<sub>3</sub> and BaTiO<sub>3</sub>. If we look at the calculated band gaps in Table VII, we see that for TiO<sub>2</sub>–terminated surfaces the band gap is significantly reduced for SrTiO<sub>3</sub> and BaTiO<sub>3</sub>, whereas for PbTiO<sub>3</sub> the band gap is almost unchanged. The reduction of the band gap in SrTiO<sub>3</sub> and BaTiO<sub>3</sub> is mainly due to an upward intrusion of the upper valence band states near the M point into the lower part of the band gap (as pointed out in Ref. , this is caused by the suppression of the hybridization of certain O $`2p`$ and Ti $`3d`$ orbitals in the surface layer). In PbTiO<sub>3</sub> we find the same upward movement of the upper valence band states near the M point, but these states stay just below the highest valence states at the X point, and so the band gap is almost unchanged.
On the other hand, for the AO–terminated surfaces we see no reduction of the band gap for any of the three perovskite compounds. Even here, however, there is a subtle difference between PbTiO<sub>3</sub> and the other materials, this time concerning the conduction band edge. According to our calculations, the Pb $`6p`$ states overlap the Ti $`3d`$ states to some degree in bulk PbTiO<sub>3</sub>, and this effect is accentuated at the $`\mathrm{\Gamma }`$ point of the surface Brillouin zone on the Pb–O terminated surface, where the lowest Pb $`6p`$ state falls just below the lowest Ti $`3d`$ state. We thus suggest that the conduction band minimum may actually have Pb $`6p`$ character at this surface, although the effect is too small to affect the band gaps in Table VII substantially. This might be an interesting target of investigation for future spectroscopic experimental studies.
## IV Summary
In summary, we have calculated structural and electronic properties of PbTiO<sub>3</sub> (001) surfaces using a first-principles density-functional approach. The results are compared and contrasted with corresponding previous calculations on BaTiO<sub>3</sub> and SrTiO<sub>3</sub> surfaces. We observe qualitatively different behavior of the PbTiO<sub>3</sub> surfaces in several respects. First, within the narrow range of PbO and TiO<sub>2</sub> chemical potentials permitted by bulk thermodynamics, we find that the TiO<sub>2</sub>-terminated surface is never thermodynamically stable. Thus, the PbO-terminated surface is expected to be the one observed experimentally. Second, the interaction between the ferroelectric distortion and the presence of the surface is quite different for PbTiO<sub>3</sub>, compared to BaTiO<sub>3</sub>. In particular, the ferroelectricity is strongly enhanced at the AO-terminated surface and suppressed at the TiO<sub>2</sub>-terminated surface, just the opposite of the behavior found for BaTiO<sub>3</sub>. Moreover, the ferroelectric distortion at the surface allows for a drastic reduction of the rumpling of the surface layer on the PbO-terminated surface, an effect which is not seen on the BaO-terminated of BaTiO<sub>3</sub>. Third, the surface electronic band structure is qualitatively modified in the case of PbTiO<sub>3</sub> by the presence of Pb $`6s`$ and $`6p`$ states in the upper valence and lower conduction regions.
###### Acknowledgements.
This work was supported by the ONR grant N00014-97-1-0048.
|
no-problem/9908/hep-th9908194.html
|
ar5iv
|
text
|
# Cosmic String Helicity: Constraints on Loop Configurations, and the Quantization of Baryon Number
## I Introduction
Cosmic strings are a special case of the vorticity phenomenon, which appears in many forms - in the flow of rivers and streams, in smoke rings, in hurricanes and tornadoes, etc. A special case of vorticity are the quantum vortices which include vortices in superfluids, flux tubes in superconductors and cosmic strings. Global cosmic strings are analogues to Vortices in superfluids while local cosmic strings are analogues to flux tubes in superconductors . All these vortices are described by a complex wave function or scalar field with a multivalued phase function $`\theta `$, such that
$$\theta d𝐥=2\pi nn=\pm 1,\pm 2,\mathrm{}$$
(1)
over a closed contour surrounding the string. As a result different physical quantities get quantized: the circulation of a superfluid vortex, the magnetic flux in a superconductor, the magnetic flux in a local cosmic string.
Vorticity was first investigated by Helmholz and Kelvin who set down the foundations for the study of vorticity in hydrodynamics. In this work we try to implement concepts born in the field of classical hydrodynamics to cosmic strings. As we shall see the interaction between these two fields can be very rewarding. Turbulent fluid is regarded as full of vortex filaments . Quantum turbulence in superfluids has likewise been defined as a superfluid state featuring a tangle of quantum vortex filaments . This paper is concerned with helicity of a tangle of cosmic strings.
The concept of fluid helicity was introduced by Moffatt as a useful measure of the degree of linkage of ordinary fluid vortice loops. It is defined as the volume integral of the scalar product of a velocity field $`𝐯`$ and its curl, the vorticity $`\xi `$. Following this, Moffatt and others considered helicity integrals in general, i.e. the space integral over the scalar product of a vector field and its curl
$$H=𝐕\times 𝐕d^3r,$$
(2)
and found them to have a topological interpretation as the linking of the field lines of the divergence free field $`\times 𝐕`$. Thus the conservation of helicity can be explained for systems in which the field lines may not cross each other, such as vortex lines in an inviscid fluid and the magnetic field lines in a perfectly conducting fluid . For these flows the field lines are said to be frozen into the flow, and the helicity is a topological invariant mathematically related to the Hopf invariant classifying the nontrivial homotopy classes of maps from $`S^3`$ to $`S^2`$.<sup>*</sup><sup>*</sup>*The existence of a relation between helicity integrals and the topological Hopf invariant has been mentioned by . However, the exact relation had been derived by Arnold . The Hopf invariant has an elementary geometric interpretation as the linking number of the preimages of the points of the target space $`S^2`$, which are isomorphic to circles in $`S^3`$. The Hopf invariant is equivalent to the helicity integral of a vector field who’s field lines lie tangent to the preimage circles. In this sense the Hopf invariant is a helicity integral term, but the inverse is usually not the case.
Superfluids and cosmic strings are both described by field theories of a complex field $`\psi =\rho e^{ı\theta }`$ which obeys a nonlinear field equation , e.g. the nonlinear Ginzburg-Pitaevskii equation or the Higgs equation. Cosmic strings and superfluid vortices both occur in the presence of spontaneous symmetry breaking . A vortex filament or cosmic string is a configuration of $`\psi `$ which approaches asymptotically the broken symmetry vacuum, and is characterized by a phase $`\theta `$ which changes by $`2\pi N`$ ($`N𝐙`$) when one goes once around the filament or string axis. $`N`$, is termed the winding number of the string. Single valuedness of the field $`\psi `$ on the filament axis requires that $`\rho =0`$ along it. The curves along which $`\rho =0`$ define the position of the strings. A vortex filament or cosmic string must either be infinite in both directions, close on itself (cosmic string loop, vortex ring), or terminate on a boundary of the system. Otherwise the behavior of the phase just beyond the filament’s free ends would be ambiguous. We assume throughout that all strings are closed un-knotted loops confined to a finite region.
The analogy between vortices in superfluids and global cosmic strings has inspired the construction of a helicity for global cosmic strings. Bekenstein defined a helicity for global strings, and Vachaspati and Field , Sato and Yahikozawa and Sato did this for local strings. These works tried to relate helicity for cosmic strings with the topological and geometrical structure of these strings (such as linking, knotting, writhing and twisting). In all these works bizarre physical conclusions were drawn. Bekenstein concluded that an isolated unknotted loop of global string is restricted to a plane i.e., only planar configurations may exist. This is strange since dynamically there seems to be no constraint on a loop accumulating a distortion continuously in an arbitrary direction. Furthermore simulations of the formation and evolution of cosmic strings show the formation of loops which are not planar . Bekenstein also claimed that the contortion of a single knot is quantized in integers. This is also strange because it says a single knot may not wiggle freely about, but is frozen in a configuration of integer contortion.
For local cosmic strings there is a relation between the electroweak magnetic helicity of a tangle of strings and the violation of baryon number conservation via what is known as the chiral anomaly . Vachaspati and Field, Sato and Yahikozawa and Sato found the helicity to change continuously with the shape of the strings, which implies that baryon number changes continuously too. This clashes with the expectation that baryon number is quantized in integers.
We show here that the topological interpretation of helicity as the linking of field lines themselves is not fully incorporated in these works, and hence an important term was left out, which accounts for all the mentioned unacceptable conclusions. Helicity arises from internal structure within a string, determined by the twist and writhe of the string, in addition to the external relations between strings, i.e. linking and knotting. The missing term arises from the internal structure of field lines within the strings which we name internal helicity. It was previously believed that since the strings length is much greater than its width, the internal structure of the string becomes unimportant and physical quantities of interest can be averaged over the string cross-section. However, as we will see this is not always the case.
In order to clarify the contribution of internal helicity, the relation of helicity to the topological structure of field lines will be further developed here. Field lines of vector fields which are divergence-free do not have endpoints. This property allows us to examine such field structures in terms of the topology of closed curves. The link between helicity and topological invariants of curves was first conjectured by Moffatt in 1981 and then developed by Berger and Field in 1984 . In 1992 Moffatt and Ricca managed to derive the extra term, arising from the internal structure of magnetic flux tubes, directly from the field equations of motion for a magnetohydrodynamic fluid. This is not possible when calculating the helicity of cosmic strings since their equations are second order non-linear equations, and therefore the calculation is too complicated. However, using formulas describing curves, developed by mathematical biologists investigating the structure of DNA , the internal helicity may be calculated. By adding it to the external helicity we correct the mentioned results.
In section II we review the concept of helicity of a solenoidal (divergence free) vector field, and dwell on the relation of the helicity to the linking of the field lines of the solenoidal field. In section III we review the construction and calculation of the helicity for a tangle of global cosmic string loops, as a function of different geometrical and topological properties of the strings . Next we correct the previously mentioned conclusions of Bekenstein by means of the internal helicity. In section IV we deduce new constraints on the permitted linkages of string loops. We find that the linking possibilities of string loops depends on their winding numbers. In section V we review how the electroweak helicity of local cosmic strings in the Weinberg-Salam model is related to the baryon number, and clarify how baryon number conservation is violated. The baryon number is found to be quantized, in contrast to previous results. Again, we show how internal helicity accounts for the new results.
## II Helicity as Topological Invariant
The term helicity is used in particle physics for the scalar product of the momentum and spin of a particle. Moffatt adopted this term to describe the volume integral over the scalar product of a velocity field $`𝐯`$ and its curl, the vorticity $`\xi `$
$$H=𝐯\xi 𝑑V$$
(3)
which he named the helicity of the flow. This idea stemmed from the essentially kinematical interpretation of the quantity $`𝐯\xi `$. The streamlines of the flow passing near any representative point $`0`$ in a small volume element $`dV`$ are (locally) helices about the streamline through $`0`$, and the contribution
$$𝐯\xi dV𝐯_\mathrm{𝟎}\xi _\mathrm{𝟎}dV$$
(4)
to $`H`$ from $`dV`$ is positive or negative according as the screw of these helices is right-handed or left-handed.
More generally, the concept of helicity describes the volume integral over the scalar product of any general vector and its curl. In particular, the magnetic helicity of magnetohydrodynamic flow
$$H=𝐀𝐁𝑑V$$
(5)
had already been shown by Woltjer to be conserved for a perfectly conducting fluid. Later Moffatt showed that the fluid helicity of an ideal barotropic fluid is conserved as well.This is true for any inviscid flow which conserves circulation, including irrotational flow. Both these cases share the property that the divergence free field lines (vortex lines or magnetic field lines) are frozen into the fluid, and that in consequence knots and linkages of the field lines are inevitably conserved. It was through struggling to understand the physical meaning of this result that Moffatt was led to the topological interpretation of helicity in terms of links and knots in convected vector fields generally, and to the characterization of helicity as a topological invariant for these cases. Moreau observed that conservation of fluid helicity can be deduced from Noether’s theorem. Helicity thus appears to have status comparable to Energy, momentum and angular momentum in this sense. As we shall see, unlike these quantities helicity also has a strong topological character. To demonstrate this we review the calculation of helicity for the simple case of two linked flux tubes and show that the helicity counts the Gauss linking number of the tubes .
Consider two flux tubes linked once (Fig. 1), carrying fluxes $`\varphi _1`$ and $`\varphi _2`$ respectively, of the vector field $`𝐁=\times 𝐀`$ (the flux is zero everywhere outside the filaments).
Within each filament, the $`𝐁`$-lines are unlinked curves which close on themselves after just one passage round the filament. The helicity $`H`$ of the two flux tubes has the same form as the pseudo-scalar quantity described in (5).It is straightforward that $`H`$ does not depend on the gauge of $`𝐀`$; for if $`𝐀`$ is replaced by $`𝐀+\psi `$, then $`H`$ is unchanged.
Adopting the coulomb gauge for $`𝐀`$ (i.e. $`𝐀=0`$), and imposing the further condition $`𝐀=O(|r|^3)`$ as $`|r|\mathrm{}`$, $`𝐀`$ is given by the Biot-Savart law:
$$𝐀(r)=\frac{1}{4\pi }\frac{𝐁(r^{})\times (𝐫𝐫^{})}{|𝐫𝐫^{}|^3}𝑑V^{},$$
(6)
so that from (5),
$$H=\frac{1}{4\pi }\frac{𝐁(r)\times 𝐁(r^{})(𝐫𝐫^{})}{|𝐫𝐫^{}|^3}𝑑V𝑑V^{}.$$
(7)
Each tube may be built of many infinitely small flux tubes carrying fluxes $`\delta \varphi _1`$ and $`\delta \varphi _2`$. Replacing $`𝐁dV`$ by $`\delta \varphi d𝐥`$: allowing for the fact that each flux filament is integrated over twice once as $`r`$ and once as $`r^{}`$, we find that $`\delta H=2_{12}\delta \varphi _1\delta \varphi _2`$ with
$$_{12}=\frac{1}{4\pi }_{C_1}_{C_2}\frac{(d𝐥\times d𝐥^{})(𝐫𝐫^{})}{|𝐫𝐫^{}|^3}$$
(8)
which is the Gauss formula for the linking of two curves. The sign of $`_{12}`$ depends on the relative orientations of the field in the two filaments (according to the right hand rule).<sup>§</sup><sup>§</sup>§In this derivation, it is essential that each flux tube should by itself have zero helicity and this is ensured by the above assumption that the $`𝐁`$-lines within either tube on its own are unlinked closed curves.
It is simple to obtain the total helicity of the two flux tubes. Each pair of filaments (one from each tube) make a contribution $`2_{12}\delta \varphi _1\delta \varphi _2`$ to the total helicity, so that summing over all such pairs, this is now given by
$$H=2_{12}\varphi _1\varphi _2.$$
(9)
Helicity has been shown to be an invariant for flows in which the divergence free field lines are frozen into the flow (Woltjer for a perfectly conducting fluid and Moffatt for ideal barotropic flow). Combining this knowledge with (9), the interpretation of the invariant in terms of conservation of linkage of the (vortex/magnetic) field lines which are frozen in the fluid is immediate.
## III Helicity of Global Cosmic Strings
States of a Higgs field containing a tangle of cosmic strings resemble a tangle of quantum vortex filaments which arise in quantum turbulent flow in superfluids. They are both described by a complex wave function $`\psi =\rho e^{ı\theta }`$ and a winding number $`N`$. The phase $`\theta `$ changes by $`2\pi N`$ when one goes once around the filament or string axes. The curve $`\rho =0`$ defines the position of the strings and the passage from $`\rho =0`$ to the vacuum value of $`\rho `$ takes place over a certain lengthscale denoted by $`r_0`$. This lengthscale is very small in comparison with the strings lengths and distances between the strings, i.e. the strings are very thin.
Motivated by this resemblance Bekenstein designed a helicity for cosmic strings in analogy with fluid helicity of a superfluid. We first review the original calculation of the helicity of a tangle of strings, and then correct it for the left-out contribution of internal helicity.
### A Constructing the helicity of global cosmic strings
In a superfluid the velocity is proportional to the gradient of the phase of the wave function; Bekenstein defines the vector
$$𝐯=\mathrm{\Theta }(\rho ^2)\theta ,$$
(10)
where $`\mathrm{\Theta }(\rho ^2)`$ is taken to be a function which rises rapidly from its value at a string axis $`\mathrm{\Theta }(0)=0`$ to near its asymptotic value, $`\mathrm{\Theta }(\rho _0^2)=1`$, over an interval of $`\rho ^2`$ which corresponds to a distance of order $`r_0`$. $`𝐯`$ is a kind of regularised velocity in that the singularity of $`\theta `$ on the axis is defused by the $`\mathrm{\Theta }`$ function (far from the axis $`𝐯`$ will drop inversely with the distance from it as befits a vortex velocity field). We have
$$\times 𝐯=\mathrm{\Delta }(\rho ^2)\rho ^2\times \theta ,$$
(11)
where $`\mathrm{\Delta }(\rho ^2)`$ is the derivative of $`\mathrm{\Theta }(\rho ^2)`$ with respect to $`\rho ^2`$. $`\mathrm{\Delta }(\rho ^2)`$ has a rather sharp peak at a distance $`r_0`$ from a string axis and the integral of $`\mathrm{\Delta }`$ over all $`\rho ^2`$ is unity.
The helicity of the field $`𝐯`$ is defined as
$$h=𝐯\times 𝐯d^3r.$$
(12)
Unlike ordinary fluid helicity, this one vanishes.A helicity which vanishes inherently cannot be used to study the dynamics of a system; only results that relate to static configurations can be obtained. There is no physical meaning to the conservation of a helicity of this kind. If one is interested in drawing conclusions on the dynamics of a network of strings, a helicity which does not vanish must be constructed, and then conditions on its conservation must be found. This is clear since the vectors $`𝐯`$ and $`\times 𝐯`$ are perpendicular by construction. Nevertheless, string loops (and vortex rings) can link, and knot, and their linkage is correlated with other topological properties of the strings to be introduced later.
Throughout it is assumed that linked and knotted loops can occur over some periods with no intersections. Situations where there are intersecting strings are viewed as representing singular moments, not generic ones. (Simulations of the development of superfluid vortices and global cosmic strings confirm this assumption). Bekenstein develops from the zero helicity a quantitative relation between the linkage and other properties of the loops.
### B Evaluation of global cosmic strings helicity
In order to evaluate the helicity explicitly for a tangle of cosmic string loops, it is convenient to rewrite (12) as a two-point functional of $`𝐯`$ (This is valid if the loops are confined to a finite region )
$$h=H[𝐯]=d^3rd^3r^{}\frac{(\times 𝐯)\times (\times 𝐯)^{}(𝐫𝐫^{})}{4\pi |𝐫𝐫^{}|^3}.$$
(13)
We follow Bekenstein in the following discussion of the reduction of $`H[𝐯]`$, and the conclusions that may be drawn from it. For field configurations of size large compared to $`r_0`$, and where all the string axis are always well apart on scale $`r_0`$, $`H`$ receives a contribution only when each of $`𝐫`$ and $`𝐫^{}`$ lies very close to a string axis. This stems from the fact that the $`\mathrm{\Delta }`$ function in (11) confines the integrals to very near the axis. Hence, Bekenstein assumes $`H`$ is equivalent to a double line integral taken along the string axis. In addition the cross section of each loop is very little distorted by the presence of the other loops and nearby portions of itself and therefore is assumed to possess a line symmetry. Under these assumptions and a few manipulations Bekenstein concludes that
$$d^3r\times 𝐯=2\pi |N|d𝐬$$
(14)
for the contribution of a little piece of string, where $`d𝐬`$ here is the vectorial line element along the piece of string considered.
Inserting (14) into (13) we obtain a double contribution to helicity from each pair of loops, because $`d𝐬`$ may progress through one loop (call it $`\mathrm{}_i`$) while $`d𝐬^{}`$ does so through the other ($`\mathrm{}_j`$) and vice-versa. In addition we obtain a contribution from each string loop alone so that
$$H[𝐯]=\frac{1}{2}\underset{ij}{}2_{ij}(2\pi |N_i|)(2\pi |N_j|)+\underset{i}{}𝒲_i(2\pi |N_i|)^2.$$
(15)
where
$$_{ij}=\frac{1}{4\pi }_\mathrm{}_i_\mathrm{}_j\frac{d𝐬\times d𝐬^{}(𝐫𝐫^{})}{|𝐫𝐫^{}|^3}$$
(16)
is the Gauss linking number (8) of the two loops $`\mathrm{}_i`$ and $`\mathrm{}_j`$ and
$$𝒲_i=\frac{1}{4\pi }_\mathrm{}_i_\mathrm{}_i\frac{d𝐬\times d𝐬^{}(𝐫𝐫^{})}{|𝐫𝐫^{}|^3}$$
(17)
is the writhing number of loop $`i`$ (Fuller ).Equation (15) differs from the original one in by a factor of $`4\pi `$ which was missed in the previous calculation. While $`_{ij}`$ is a topological invariant of the linking of the string loops and must be an integer, $`𝒲_i`$ is not a topological invariant. It changes with the shape of the loop and can obtain any value. $`𝒲_i`$ measures the deviation of a loop from its plane; the writhe of a planar loop is zero.
When all loops are confined to a finite region, $`H=h=0`$, and (15) gives
$$0=(2\pi )^2\underset{ij}{}_{ij}|N_i||N_j|+(2\pi )^2\underset{i}{}𝒲_iN_i^2.$$
(18)
We thus obtain a constraint connecting the linking of a collection of loops with their geometrical shapes. Particularly simple is the case of an isolated loop. Then $`_{ij}=0`$. For such a loop we must have $`𝒲_i=0`$ too. According to this result an isolated un-knotted loop must be a planar loop at any instant . This is hard to accept since dynamically there seems to be no constraint on a loop accumulating a distortion continuously in an arbitrary direction. Furthermore simulations of the formation and evolution of cosmic strings show the formation of loops which are not planar! Similarly, according to (18) two loops linked once may never be planar since the contribution of their writhe to the helicity must cancel out the contribution from their linkage. This restriction is again hard to accept.
To obtain another constraint Bekenstein introduces the relation
$$𝒦=𝒲+𝒞.$$
(19)
$`𝒦`$ measures the knotting of the curve and can jump by increments of $`\pm 2`$, and $`𝒞`$ is the contortion of the loop defined by $`𝒞=(2\pi )^1\tau 𝑑s`$ which integrates the torsion over the whole loop.<sup>\**</sup><sup>\**</sup>\**The axis of a string can be treated as a curve in differential geometry. A curve $`r=X(s)`$ ($`s`$ is the arc length) has $`T=\dot{X}(s)`$ as unit tangent vector, has a unit normal $`N`$, and a unit binormal $`B=T\times N`$. They are connected by the Frenet-Serret equations $`{\displaystyle \frac{d}{ds}}\left(\begin{array}{c}T\\ N\\ B\end{array}\right)=(\kappa B\tau T)\times \left(\begin{array}{c}T\\ N\\ B\end{array}\right),`$where $`\kappa `$, a positive scalar, is called the curvature of the curve, and $`\tau `$, which may take on either sign, is called the torsion. Inserting $`𝒲`$ from (19) into (18) gives the following constraint connecting the contortion, knotting and linking of a collection of loops:
$$\underset{i}{}𝒞_iN_i^2=\underset{ij}{}_{ij}|N_iN_j|+\underset{i}{}𝒦_iN_i^2.$$
(20)
This relation says that the sum of contortions of a collection of unit winding number loops is quantized in integers . This implies that the contortion of any single knot is quantized in integers: the knot cannot wiggle freely about, but is frozen in a configuration of integer contortion. This is a very strong constraint on the dynamics of such a configuration, and together with our previous observations, makes us suspect that the calculation has missed a term.
### C Internal helicity
The first clue of the missing term was presented by Moffatt , who notes that close inspection of an expression of the form
$`H_\mathrm{}_i={\displaystyle _\mathrm{}_i}d^3r{\displaystyle _\mathrm{}_i}d^3r^{}{\displaystyle \frac{(\times 𝐯)\times (\times 𝐯)^{}(𝐫𝐫^{})}{4\pi |𝐫𝐫^{}|^3}},`$ (21)
integrated over a single flux loop $`\mathrm{}_i`$ (of an arbitrary divergence-free vector field), shows that the helicity in this limit has an additional contribution from pairs of points $`𝐫`$, $`𝐫^{}`$ separated by a distance comparable to the cross-sectional span of the loop. In addition to the writhe $`𝒲`$ (17), which we introduced earlier as the sole contribution to helicity from a single loop, Moffatt conjectures that the close points contribute an extra term namely
$$H_\mathrm{}_i=𝒞_\mathrm{}_i\varphi ^2$$
(22)
where $`𝒞_\mathrm{}_i`$ is the contortion of loop $`i`$ defined earlier, and $`\varphi `$ is the flux of the loop. Actually it is easy to see that Bekenstein’s calculation in Sec. III B ignores the thickness of the string, i.e. the contribution from close pairs of points $`𝐫`$ and $`𝐫^{}`$. Equation (14) reduces a volume integral to a line integral along the core of the string and thus treats the string as though it where infinitely thin. Moffatt speculates that the additional contribution, from close pairs of points $`𝐫`$ and $`𝐫^{}`$, is related to the contortion so that the helicity will be a topological invariant of the loop, i.e., $`H_\mathrm{}_i=(𝒲_\mathrm{}_i+𝒞_\mathrm{}_i)\varphi ^2=𝒦_\mathrm{}_i\varphi ^2`$ by (19). But, thinking this over we realize that the contortion is also just a property of the flux tube’s (string’s) axis and not of its internal structure hence we should expect a different term to appear.
Later Moffatt explicitly calculated the helicity of a magnetohydrodynamic system of flux tubes to find that his earlier speculation was not exact, and that $`𝒦`$ is not a topological invariant after all since its value jumps discontinuously at inflexion points (points of zero curvature). In this later calculation Moffatt manipulates the magnetohydrodynamic equation of motion of a configuration of field lines moving with the flow. This calculation is much too complicated to carry over to the case of cosmic strings since the field equations for these are non-linear. Hence we will follow in the footsteps of Berger and Field who also calculate the internal helicity due to the internal structure of a flux tube, but we will present a slightly different calculation, and a slightly different result. Berger and Field’s result was that $`H=T\varphi ^2`$ where $`T`$ is any real number, while we obtain the same result as Moffatt with $`H=n\varphi ^2`$ and $`n`$, an integer, is the linking number of the field lines in the flux tube.
We first calculate exactly the contribution to the helicity of a single vortex loop $`\mathrm{}`$. First we consider the inner structure of the string to consist of nested toroidal vorticity surfaces (a peace of the loop is shown in Fig. 2). Then we may build up the string by increments of vorticity flux $`d\varphi `$ composed of many thin flux tubes $`\delta \varphi `$ that wind around the inner flux tube $`\varphi `$ $`n`$ times.
The increment in helicity for each such flux tube $`\delta \varphi `$ is $`2n\varphi \delta \varphi `$ where $`n`$ is the Gauss linking number of element $`\delta \varphi `$ with $`\varphi `$ (the inner flux tube). The increment in helicity for the whole of $`d\varphi `$ is simply $`2n\varphi d\varphi `$. (The flux $`d\varphi `$ simply counts the number of field lines intersecting the increment in the cross section of the string, and hence the number of field lines wrapped around the flux tube in the center.) We thus have according to (13)
$`H_{\mathrm{}}`$ $`=`$ $`{\displaystyle _{\mathrm{}}}d^3r{\displaystyle _{\mathrm{}}}d^3r^{}{\displaystyle \frac{(\times 𝐯)\times (\times 𝐯)^{}(𝐫𝐫^{})}{4\pi |𝐫𝐫^{}|^3}}`$ (23)
$`=`$ $`{\displaystyle _0^\varphi }𝑑\varphi \varphi {\displaystyle \frac{(d𝐥\times d𝐥^{})(𝐫𝐫^{})}{4\pi |𝐫𝐫^{}|^3}}={\displaystyle _0^\varphi }𝑑\varphi \varphi \mathrm{\hspace{0.17em}2}n`$ (24)
Where $`d𝐥`$ is along the axis of the string and $`d𝐥^{}`$ is along one of the field lines in $`\delta \varphi `$ wrapping $`\varphi `$, and vice-versa. Finally, integrating over all the nested flux tubes
$$H_{\mathrm{}}=\frac{\varphi ^2}{2}2n=n(2\pi N)^2,$$
(25)
where $`N`$ is the winding number of the cosmic string (the flux $`\varphi `$ is easily calculated using the definition of $`𝐯`$, Stoke’s theorem and (1)).
Throughout this calculation we have assumed that $`\delta \varphi `$ and $`\varphi `$ have the same linking number as a single field line in $`\delta \varphi `$ with the axis of the string. We have also assumed that all field lines have the same linking number with the axis of the string (and with each other). This is plausible for a continuous field. As we see, the internal helicity is due to the linking of the $`\times 𝐯`$ field lines inside the string. This linking occurs when the field $`\times 𝐯`$ is twisted: if the field lines do not link the internal helicity is zero.
We would now like to express the linking number of the field lines $`n`$ as a sum of $`𝒲`$ and some other term or terms, so that we may reconcile this calculation with Bekenstein’s result (15). The linking number $`n`$ of $`\delta \varphi `$ with $`\varphi `$ is equal to the linking number of a single field line in $`\delta \varphi `$ with the axis of the string. Since these are two closed curves at a very short distance apart, the axis of the string can be treated as a curve in differential geometry, forming the edge of a ribbon whose other edge is traced by the field line.
This type of problem appears first to have been addressed by Călugăreanu , who considered two neighboring closed curves $`C`$, $`C^{}`$ forming the boundaries of a (possibly knotted) ribbon of small spanwise width $`ϵ`$, and showed that for sufficiently small $`ϵ`$ the linking number $`n`$ of $`C`$ and $`C^{}`$ can be expressed in the form
$$n=𝒲+𝒞+𝒩,$$
(26)
where $`𝒲`$ and $`𝒞`$ are respectively the writhe and the contortion of $`C`$, and $`𝒩`$ is an integer representing the number of rotations of the unit spanwise vector $`U`$ on the ribbon relative to the Frenet pair (the unit normal and unit binormal<sup>††</sup><sup>††</sup>††See footnote \**III B.) in one passage round $`C`$. (the contortion $`𝒞`$ actually measures the rotation round the curve of the Frenet pair itself.)
The Frenet pair has discontinuous behavior in going through a point of inflexion (zero curvature). As a result $`𝒞`$ and $`𝒩`$ are not well defined if $`C`$ has inflexion points. If $`C`$ is continuously deformed through an inflexional configuration, then $`𝒞`$ is discontinuous by $`\pm 1`$, but $`𝒩`$ is simultaneously discontinuous by an equal and opposite amount $`1`$, so that the sum $`𝒞+𝒩`$ does vary continuously through inflexion points . This sum is known as the twist of the ribbon
$$𝒯=𝒞+𝒩.$$
(27)
Thus, as we remarked earlier, $`𝒦=𝒲+𝒞`$ is not a topological invariant while $`n`$ (26) is.
With the help of (26) and (27) we are now ready to complete the calculation of the helicity for a single (un- knotted) cosmic string loop $`\mathrm{}`$. We take $`C`$ above to be the string’s axis and $`C^{}`$ to be along one of the field lines twisted around the axis. Then for the $`i`$-th loop
$$H_\mathrm{}_i=n_i\varphi ^2=𝒲_i(2\pi N_i)^2+𝒯_i(2\pi N_i)^2.$$
(28)
Our final expression for the helicity of a tangle of cosmic string loops is therefore
$$H[𝐯]=\underset{ij}{}_{ij}(2\pi |N_i|)(2\pi |N_j|)+\underset{i}{}𝒲_i(2\pi N_i)^2+𝒯_i(2\pi N_i)^2.$$
(29)
Contrast this (correct) result with (15). The twist $`𝒯`$ is the contribution from the close pairs of points $`𝐫`$ and $`𝐫^{}`$ inside the string; it is not a property of the axes of the string alone, but rather of the internal structure of the field lines inside the string core.
## IV New Constraints on String Configurations
In the previous sections we learned that helicity measures the linkage of the field lines of $`\times 𝐯`$, and that the total helicity of global cosmic strings is zero whatever their configuration may be. These properties must put some constraint on the possible configurations of the strings. In this section we show that the knotting and external linkage of a cosmic string determines the internal structure of the field lines inside the string, and thus external and internal helicity cancel each other out so that the total helicity is zero. Since the field lines in a string must close on themselves, and their twist depends on external knotting and linking, we will get constraints on the possible configurations of string loops. These constraints are weaker than those proposed by Bekenstein, and are free from the latter’s paradoxical aspects.
### A Determining internal linking from winding numbers of a string loop
Let us start by examining $`\theta `$ (the gradient of the phase) of the most simple configuration, a single loop. As we know the phase varies when one goes once around the string by $`2\pi N`$, where $`N`$ is the winding number of the string.
What happens if the phase varies along the string too? Because of periodicity this is only possible if
$$\theta _sd𝐬=2\pi N^{}$$
(30)
where $`d𝐬`$ is along the string axis. But in this case there will be another topological requirement. Because the phase varies by $`2\pi N^{}`$ around the perimeter of the loop, for every surface spanning the loop there must be at least one place where the amplitude $`\rho =0`$ so that the field $`\psi `$ will be single-valued. This is actually the requirement that the loop be threaded by another $`N^{}`$ unit winding-strings or a single string with winding number $`N^{}`$. Hence an isolated loop must have $`N^{}=0`$; for a totally symmetric loop $`\theta `$ will have only one component, in the direction around the string.<sup>‡‡</sup><sup>‡‡</sup>‡‡To be more accurate, the average and not the value at every point, of the component of $`\theta `$ along the loop must be zero for an isolated loop. However, any configuration where the average of the component of $`\theta `$ along the loop is zero can be continuously deformed into a configuration where the component of $`\theta `$ along the loop equals zero at every point. Hence, as we shall see the linking of the $`\times 𝐯`$ field lines is not affected.
How does this affect the configuration of the field lines of $`\times 𝐯`$ ? We recall that by definition (10), $`𝐯`$ is parallel to $`\theta `$ and $`\times 𝐯`$ (11) is perpendicular to $`𝐯`$. Hence if $`𝐯`$ has a component only around the string, $`\times 𝐯`$ will have a component only along the string, and its field lines will not be twisted or linked. Since linkage of the field lines is a topological property, it will not change for continuous deformation of the loop; therefore the field lines of any isolated loop with arbitrary shape will be unlinked. The immediate result is that the total helicity of any isolated loop with arbitrary shape is zero. This corrects Bekenstein’s conclusion (18) that an isolated loop must be a planar loop at any instant so that the helicity will be zero.
We next investigate a configuration of two linked loops $`\mathrm{}_1`$ and $`\mathrm{}_2`$ with winding numbers $`N_1`$ and $`N_2`$ respectively. We now have a component of $`\theta `$ in the direction along the strings too. The phase will vary around the perimeter of loop $`\mathrm{}_1`$ by $`2\pi N_2`$ and around the perimeter of $`\mathrm{}_2`$ by $`2\pi N_1`$ (Fig. 5). Now that $`\theta `$ has components along the string and around it, we must introduce a new concept, the chirality (handedness) of a string loop. A loop is right-handed (and has positive chirality) when the component of $`\theta `$ along the string is in the direction of the thumb of the right hand, whose fingers point along the component of $`\theta `$ around the string. If the component of $`\theta `$ along the string is in an opposite direction to the thumb, the loop is left-handed (and the chirality is negative).
Along with the definition of chirality comes the definition of the sign of the directions along and around the string. We define the positive direction around the string to coincide with the direction of $`\theta `$ around it, and the positive direction along the string to coincide with the direction of the thumb of the right hand whose fingers follow the direction of $`\theta `$ around the string axis. In a right-handed (left-handed) string loop whose phase changes around the string by $`2\pi N`$ and along the string by $`2\pi N^{}`$, $`N`$ and $`N^{}`$ will have equal (opposite) signs. It is easy to see that two strings may link only if they have the same chirality; otherwise there would be a discontinuity in the phase.
What will be the structure of the $`\times 𝐯`$ field lines in such a configuration? Now that $`\theta `$ has components in two directions, $`\times 𝐯`$ will also have at least two components, one along the string and one around it. The field lines will be twisted and linked (Fig. 6).
In such a case we expect to have two contributions to helicity, a contribution from the external linkage of the string loops, and a contribution from the internal linkage of the field lines inside the strings. These two contributions must cancel each other so that this configuration may exist. The external helicity for such a configuration with $`_{12}=_{21}=\pm 1`$ and winding numbers $`N_1`$ and $`N_2`$ is simply
$$H_{ext}=_{12}(2\pi N_1)(2\pi N_2)+_{21}(2\pi N_2)(2\pi N_1)=\pm 2(2\pi N_1)(2\pi N_2),$$
(31)
where the sign of $`_{12}`$ is determined by the relative direction of $`\times 𝐯`$ along the perimeter of both loops. We will show below that $`_{12}`$ is positive for right-handed configurations and negative for left-handed ones.<sup>\**</sup><sup>\**</sup>\**Since the sign of the winding numbers ($`N_1,N_2`$) is always positive by our current definition, there is no need to take the absolute of the winding numbers as in (29).
But computing the internal helicity
$$H_{int}=n_1(2\pi N_1)^2+n_2(2\pi N_2)^2$$
(32)
is not straightforward. The internal helicity of magnetic flux tubes in a magnetohydrodynamic fluid or of vortices in an ordinary fluid is arbitrary since there is no restriction on the linking of the field lines inside the flux tube. In those systems the internal structure of the field lines is not determined by the topology of the flux tubes or vortex filaments and, therefore, the total helicity cannot be determined exactly from their configuration. By contrast we shall now show that the internal helicity of cosmic strings is solely determined by the winding numbers of the strings and their topological configuration, so that the total helicity for a configuration of linked vortices can be expressed as a function of their linkage and winding numbers alone.
We now attempt to derive the internal linking number $`n`$, that measures the linkage of the field lines inside a cosmic string loop. The loop is characterized by winding numbers $`N`$ and $`N^{}`$ around and along the perimeter of the loop respectively. Any (un-knotted) cosmic string loop may be deformed continuously into a planar circle loop, and the twist of the field lines inside it uniformly distributed, without changing the linking of the field lines. This is possible since the linking is a topological property of the field lines, and does not change by continuous deformations. It would further simplify our calculation if we could perform it in cylindrical symmetric coordinates, by virtually cutting the loop and standing it upright, such that it acquires cylindrical symmetry (Fig. 7). We may convince ourselves that this is possible since very close to the axis of the string the field lines have cylindrical symmetry around the string, and for reasons of continuity the linkage of the field lines cannot change as we move away from the string.
The linking number $`n`$ is simply the number of times $`\times 𝐯`$ is wrapped around the $`z`$ axis (which coincides with the axis of the string). This must be an integer for the string to form a loop. The vectors $`𝐯`$ and $`\times 𝐯`$ are computed in cylindrical symmetric coordinates $`(r,\phi ,z)`$ as follows: First
$$𝐯=f(r)\theta ,$$
(33)
where $`\mathrm{\Theta }(\rho ^2)`$ in (10) is taken to be $`f(r)`$, a function of $`r`$ only. The gradient of $`\theta `$ must obey
$$\theta _\phi rd\phi =2\pi Nand_0^L\theta _zdz=2\pi N^{},$$
(34)
where $`L`$ is the length of the string. Hence for uniformly distributed twist
$$\theta =\frac{N}{r}\widehat{\phi }+\frac{2\pi N^{}}{L}\widehat{z}.$$
(35)
The calculation of $`\times 𝐯`$ gives
$$\times 𝐯=f^{}(r)\left(N\widehat{z}\frac{2\pi r}{L}N^{}\widehat{\phi }\right).$$
(36)
We are interested in the number of times each field line encircles the $`z`$ axis. Often we associate flow lines with vector fields by imagining how a test particle would move if the vector field were its velocity field. Let us assume the field line traces out the trajectory of a point particle traveling with constant velocity during a period of time $`t`$. The velocity components of the particle are equal to the components of $`\times 𝐯`$. The following simple equations must apply so that the field lines may reconnect when the string forms a loop
$$(\times 𝐯)_zt=Land(\times 𝐯)_\phi t=n2\pi r.$$
(37)
By eliminating $`t`$ between the two equations and using $`\times 𝐯`$ from (36) we obtain the result
$$n=N^{}/N.$$
(38)
Thus the internal linking number $`n`$ depends exclusively on the winding numbers of the loop.
To evaluate the total helicity for two linked un-knotted cosmic string loops, we first determine the sign of the Gauss linking number $`_{12}`$ for left-handed and right-handed configurations, which depends on the direction of $`\times 𝐯`$ along the perimeter of both loops. By our definition above, the winding numbers around the axis of the string are always positive, since we have defined the positive direction around the string to coincide with the direction of $`\theta `$ around the axis of the string. Hence from (36) the component of $`\times 𝐯`$ along the axis is always in the positive direction i.e. along the thumb of the right hand whose fingers follow the positive direction around the axis of the string. Fig. 8 makes it clear that a right-handed configuration will have $`_{12}=+1`$ and the left-handed configuration will have $`_{12}=1`$.
The sign of the internal linking number $`n`$ also depends on whether the configuration is left-handed or right-handed. As we explained earlier, in a right-handed configuration $`N`$ and $`N^{}`$ have equal signs while in a left-handed configuration, they have opposite signs. Therefore, by (38) $`n`$ is negative for a right-handed pair and positive for a left-handed one. The signs of the external and internal linkages are opposite.
We are now able to write down the total helicity for a pair of linked strings. For two right-handed linked strings with winding numbers $`N_1`$ and $`N_2`$, we have $`N_1^{}=N_2`$ and $`N_2^{}=N_1`$ (since the axis of each loop encircles the other loop) and by (31) and (32) the total helicity is
$$H=2(2\pi N_1)(2\pi N_2)\frac{N_2}{N_1}(2\pi N_1)^2\frac{N_1}{N_2}(2\pi N_2)^2=0.$$
(39)
For two left-handed linked loops the total helicity is
$$H=2(2\pi N_1)(2\pi N_2)+\frac{N_2}{N_1}(2\pi N_1)^2+\frac{N_1}{N_2}(2\pi N_2)^2=0.$$
(40)
The external and internal helicities thus cancel each other out without constraining the shapes of the loops.
We have thus removed the constraint on the geometrical shape of a single loop, a pair of linked loops or any other configuration. The contributions to helicity are purely topological. But (38) introduces a new topological constraint. The internal Gauss linking number $`n`$ of the field lines in the string must be an integer. Hence the winding number $`N^{}`$ along the perimeter of the string is not free to take on just any value, i.e., a string loop with winding number $`N`$ cannot be linked arbitrarily with other loops. For the right handed pair of linked loops $`\mathrm{}_1`$ and $`\mathrm{}_2`$, for example, the internal linking number of $`\mathrm{}_1`$ is $`n_1=N_2/N_1`$ and the internal linking number of $`\mathrm{}_2`$ is $`n_2=N_1/N_2`$. Therefore we must have $`N_1=N_2`$ so that both $`n_1`$ and $`n_2`$ be integers. This means that only string loops with the same winding number may link once. In the next section we generalize this result to a generic configuration of linked un-knotted loops.
### B Generalized topological constraint on configurations of un-knotted linked loops
Examining a couple more examples will help us formulate a more general result.
Fig. 9 depicts a right-handed cosmic string loop $`\mathrm{}_1`$ with winding number $`N_1`$ linked twice with a cosmic string loop $`\mathrm{}_2`$ with winding number $`N_2`$ such that their Gauss linking number is $`_{12}=2`$. As explained earlier the change in phase along the perimeter of a loop is determined by the winding numbers of the strings threading it. Since $`\mathrm{}_2`$ links $`\mathrm{}_1`$ twice the change in phase along $`\mathrm{}_1`$ is $`2\pi 2N_2`$ so that $`N_1^{}=2N_2`$; correspondingly the phase along $`\mathrm{}_2`$ changes by $`2\pi 2N_1`$ so that $`N_2^{}=2N_1`$. Hence the internal linking number of $`\mathrm{}_1`$ is $`n_1=2N_2/N_1`$ and that of $`\mathrm{}_2`$ is $`n_2=2N_1/N_2`$. Thus the values of $`N_2`$ must be restricted to the integers that divide $`2N_1`$ and are greater than or equal to $`N1/2`$. For a few examples see Table I, only such loops may link twice. The total helicity for this configuration is
$$H=22(2\pi N_1)(2\pi N_2)\frac{2N_2}{N_1}(2\pi N_1)^2\frac{2N_1}{N_2}(2\pi N_2)^2=0.$$
(41)
Our second example in Fig. 10 depicts a left-handed string loop with winding number $`N_1`$ linked once with two other left-handed string loops, one with winding number $`N_2`$ and the other with winding number $`N_3`$.
The external Gauss linking numbers are $`_{12}=1`$ and $`_{13}=1`$ and $`_{23}=0`$. The change in phase around the perimeter of the strings, and the internal linking numbers are determined by the same arguments as above, such that $`n_1=(N_2+N_3)/N_1`$, $`n_2=N_1/N_2`$ and $`n_3=N_1/N_3`$. The total helicity will thus be
$`H`$ $`=`$ $`2(1)(2\pi N_1)(2\pi N_2)+2(1)(2\pi N_1)(2\pi N_3)`$ (42)
$`+`$ $`{\displaystyle \frac{(N_2+N_3)}{N_1}}(2\pi N_1)^2+{\displaystyle \frac{N_1}{N_2}}(2\pi N_2)^2+{\displaystyle \frac{N_1}{N_3}}(2\pi N_3)^2=0`$ (43)
as we would expect.
By generalizing from these examples we see that the internal linking number $`n_i`$ for a loop $`\mathrm{}_i`$ linked with an arbitrary number of other loops is
$$n_i=\frac{p_{ji}|_{ij}|N_j}{N_i},$$
(44)
where $`p`$ is equal to +1 for right handed loops and -1 for left handed ones. This equation is also a constraint on the winding numbers and the Gauss linking numbers of the loops linking $`\mathrm{}_i`$, since $`n_i`$ must come out an integer.
The total helicity of loops linked in an arbitrary manner thus vanishes:
$`H`$ $`=`$ $`4\pi ^2\left(p{\displaystyle \underset{i,ji}{}}|_{ij}|N_iN_j+{\displaystyle \underset{i}{}}n_iN_i^2\right)`$ (45)
$`=`$ $`4\pi ^2\left(p{\displaystyle \underset{i,ji}{}}|_{ij}|N_iN_j{\displaystyle \underset{i}{}}{\displaystyle \frac{p_{ji}|_{ij}|N_j}{N_i}}N_i^2\right)`$ (46)
$`=`$ $`4\pi ^2\left(p{\displaystyle \underset{i,ji}{}}|_{ij}|N_iN_jp{\displaystyle \underset{i}{}}{\displaystyle \underset{ij}{}}|_{ij}|N_jN_i\right)=0.`$ (47)
## V Electroweak Helicity of Local Cosmic Strings and Baryogenesis
Local cosmic strings are solutions to Lagrangians invariant under local gauge transformations. The simplest model for a local cosmic string is the Abelian Higgs model; it has $`U(1)`$ local symmetry. Nielsen and Olesen found straight string (cylindrically symmetric) solutions to this model characterized by a magnetic field confined to the string’s core whose flux is quantized in proportion to the winding number $`N`$ of the string. Another important property of these strings is that they have finite energy per unit length (in contrast to the global strings) since the gauge field $`A_\mu `$ can cancel out the gradient of the phase at spatial infinity.
The Nielsen-Olesen string is actually a two dimensional vortex solution with energy density localized around a point in space, which can trivially be embedded into the extra dimension to produce an infinite string-like object whose energy density is localized along a line. By embedding the vortex along a closed curve, rather than a straight line, the ends of the string can be joined to produce a closed string of finite length. Thus we may have a configuration that is a tangle of such local cosmic strings which are both knotted and linked with each other. Since these strings have quantized magnetic flux tubes in their cores, we actually have a tangle of knotted and linked magnetic flux tubes, and each configuration may be characterized by a magnetic helicity quantifying the linking of the cosmic strings, and the internal linkage of the magnetic field lines inside each string.
In this section we refer to the result that the change in the electroweak magnetic helicity of a tangle of local cosmic strings, in the standard model, is related to baryon number violation. Previous works indicated a continuous change in baryon number. In contradiction to these works, We show here, by using our knowledge of the topological interpretation of helicity as the linking of field lines that, baryon number on local strings is quantized in this model.
### A The term $`I=d^4xF_{\mu \nu }\stackrel{~}{F}^{\mu \nu }`$ for Nielsen-Olesen strings
Sato and Yahikozawa scrutinize the geometrical and topological properties of the Abelian Higgs model with vortex strings by calculating the topological term $`I`$
$$Id^4xF_{\mu \nu }\stackrel{~}{F}^{\mu \nu },$$
(48)
which originates from the chiral anomaly, and is the CP violating $`\theta `$ term in the action.<sup>\*†</sup><sup>\*†</sup>\*†See Weinberg ch. 22-23, Peskin & Schroeder ch. 19. The definition of the dual is the following: $`\stackrel{~}{F}^{\mu \nu }=\frac{1}{2}ϵ^{\mu \nu \rho \sigma }F^{\rho \sigma }`$. $`I`$ is evaluated for cases where all vortex strings are closed, with winding number $`N=1`$, and the number of the vortex strings is conserved i.e. vortex string reconnection does not occur. Using a topological formulation which they developed, and via a complicated calculation, Sato and Yahikozawa show that
$$I=\frac{8\pi ^2}{e^2}\left[\underset{ij=1}{\overset{n}{}}_{ij}+\underset{i=1}{\overset{n}{}}𝒲_i\right]_{t=\mathrm{}}^{t=+\mathrm{}}.$$
(49)
The first term is the linking number between each pair of strings and comes from the mutual interactions between different vortex strings. The second term is the writhing number; it comes from the self-interactions of each vortex string, and can vary with the shape of the string.
It is plain that the value of $`I`$ changes continuously with the shape of the strings as time elapses. Equation (49) strongly resembles Bekensteins result (15) for the helicity of global cosmic strings. We shall now show that $`I`$ is just the total change in time of the magnetic helicity of the Abelian Higgs model with local vortex strings, and hence equal to the change in the external and internal linking numbers of the magnetic field lines each weighed by the square of the quantized flux in the corresponding vortex.
By construction $`F_{\mu \nu }\stackrel{~}{F}^{\mu \nu }=4𝐄𝐁`$, so we may write
$$I=4d^4x𝐄𝐁.$$
(50)
Using Maxwell’s equations
$`𝐄={\displaystyle \frac{𝐀}{t}}\varphi ,𝐁=\times 𝐀`$and
$`{\displaystyle \frac{𝐁}{t}}=\times 𝐄,`$and integration by parts,
$`{\displaystyle d^4x𝐄𝐁}`$ $`=`$ $`{\displaystyle d^3x𝐀𝐁}|_{t=\mathrm{}}^{t=+\mathrm{}}+{\displaystyle d^4x𝐀\frac{𝐁}{t}}`$ (51)
$`=`$ $`{\displaystyle d^3x𝐀𝐁}|_{t=\mathrm{}}^{t=+\mathrm{}}{\displaystyle d^4x𝐄\times 𝐀}`$ (52)
where we have used the relation $`(𝐄\times 𝐀)=𝐀(\times 𝐄)𝐄(\times 𝐀)`$and the assumption that all fields die off at spatial infinity. Our final result is thus
$$I=2d^3x𝐀𝐁|_{t=\mathrm{}}^{t=+\mathrm{}}=2\mathrm{\Delta }H$$
(53)
where $`H`$ stands for the magnetic helicity.
As we already know, the magnetic helicity may be written as the sum of the external linkages of the flux tubes and the internal linkage of the field lines in each flux tube, times the square of the flux (29)
$$H=\underset{ij}{}_{ij}(\frac{2\pi }{e})^2+\underset{i}{}𝒲_i(\frac{2\pi }{e})^2+\underset{i}{}𝒯_i(\frac{2\pi }{e})^2.$$
(54)
The two terms $`𝒲_i+𝒯_i`$ add to $`n_i`$, the internal linking number (28), and the magnetic flux in a local string of unit winding number in this model is $`2\pi /e`$.<sup>\*‡</sup><sup>\*‡</sup>\*‡Finite energy of the string requires $`D_\mu \psi =(_\mu ıeA_\mu )\psi 0`$ as $`r\mathrm{}`$. Hence, $`_\mu \theta =eA_\mu `$ and via Stoke’s theorem and (1) the magnetic flux is $`2\pi /e`$. Therefore, according to (53) and (54) the term $`I`$ is simply equal to
$$I=\frac{8\pi ^2}{e^2}\left[\underset{ij}{}_{ij}+\underset{i}{}𝒲_i+\underset{i}{}𝒯_i\right]_{t=\mathrm{}}^{t=+\mathrm{}}.$$
(55)
Comparing this with (49) we see that Sato and Yahikozawa have missed out the twist term. This resulted from neglecting the internal structure of the vortex strings in their calculation. Our result is crucially different from theirs: according to their result the term $`I`$ may have a continuous range of values, while according to our result $`I`$ is quantized because $`\mathrm{\Delta }H`$ is quantized by the internal and external linking numbers of the magnetic flux tubes threading the strings cores. This fact has an important physical consequence, as we shall see presently.
### B The chiral anomaly and baryogenesis
The origin of the excess of matter over antimatter in our universe remains one of the fundamental problems. A widely discussed model is baryogenesis in the course of the broken symmetry electroweak transition, where the baryonic asymmetry is induced by the quantum chiral Adler-Bell-Jackiw anomaly . In models of fermions coupled to gauge fields, certain current-conservation laws are violated by this anomaly. The term $`F_{\mu \nu }\stackrel{~}{F}^{\mu \nu }`$ calculated on the background of the Nielsen-Olesen strings has been shown to cause violation of baryon number conservation via this anomaly.
The anomaly arises when a classical symmetry of the Lagrangian does not survive the process of quantization and regularization: the symmetry of the Lagrangian is not inherited by the effective action. An axial current that is conserved at the level of the classical equations of motion can thus acquire a nonzero divergence through one-loop diagrams that couple this current to a pair of gauge boson fields. The Feynman diagram that contains this anomalous contribution is a triangle diagram with the axial current and the two gauge currents at its vertices.A good account of this subject can be found in Peskin . Left and right handed fermions contribute terms with opposite signs to the anomaly; however, in chiral theories in which the gauge bosons do not couple equally to right- and left-handed species, the sum of these terms can give a non-zero contribution. In theories such as QED or QCD in which the gauge bosons couple equally to right- and left-handed fermions, the anomalies automatically cancel.
Within the standard model of weak interactions (the Glashow-Weinberg-Salam $`SU(2)\times U(1)`$ invariant theory), the requirement from experiment that weak interaction currents are left-handed forces us to choose a chiral gauge coupling. The coupling of the $`W`$ boson to quarks and leptons can be derived by assigning the left-handed components of quarks and leptons to doublets of an $`SU(2)`$ gauge symmetry, and then identifying the $`W`$ bosons as gauge fields that couple to this $`SU(2)`$ group, while making the right-handed fermions singlets under this group. The restriction of the symmetry to left-handed fields leads to the helicity structure of the weak interactions effective Lagrangian, and thus causes the chiral anomaly of the baryon current to arise.All the possible gauge anomalies of weak interaction theory must vanish for the Glashow-Wienberg-Salam theory to be consistent. It turns out that the leptons and quarks exactly cancel each others anomalies. In fact, the charge assignments of the quarks and leptons in the Standard Model are precisely the ones that cancel the anomaly. The baryon number current in this model is $`j_B^\mu =\overline{Q}\gamma ^\mu Q`$ and the anomalous baryonic current conservation equation is
$$_\mu j_B^\mu =\frac{N_F}{32\pi ^2}(g^2W_{\mu \nu }^a\stackrel{~}{W}^{a\mu \nu }+g^2Y_{\mu \nu }\stackrel{~}{Y}^{\mu \nu }),$$
(56)
where $`W_{\mu \nu }^a`$ and $`Y_{\mu \nu }`$ are the field strengths for the three $`SU(2)`$ and one $`U(1)`$ gauge fields $`W_\mu ^a`$ ($`a=1,2,3`$) and $`Y_\mu `$ respectively, and $`g`$ and $`g^{}`$ are their charges. $`N_F`$ is the number of quark families.
The baryon number current can be integrated to yield the change in the baryon number between two different times, if we assume that the baryonic flux through the surface of the three-volume of interest vanishes:
$`Q_B`$ $`=`$ $`{\displaystyle d^3xj^0}`$ (57)
$`\mathrm{\Delta }Q_B`$ $`=`$ $`{\displaystyle 𝑑t_0Q_B}={\displaystyle d^4x_0j_B^0}={\displaystyle d^4x_\mu j_B^\mu }.`$ (58)
Equations (56) and (58) relate the change in the baryon number to terms of the form $`d^4xF_{\mu \nu }\stackrel{~}{F}^{\mu \nu }`$ which we calculated in the previous section for an abelian model on the background of Nielsen-Olesen vortices. This term was found to be related to the change in the sum of the linking of the strings from their initial to final configurations (55). This leads us to an exciting idea : if chiral fermions are coupled to electroweak vortices, baryon number may be violated because of the anomaly as vortices link and de-link. For this to happen there must be a stable string solution in this model, and the change in the baryon number must be related to the helicity of an electroweak field whose flux is confined to the string.
It is generally believed that the standard electroweak model is free from topological defects. The reason is that the first homotopy group of
$`M=G/H=\left(SU(2)_L\times U(1)\right)/U(1)_{em}`$can be shown to be trivial i.e. $`\mathrm{\Pi }_1(M)=0`$ . However, this does not mean that the model is free from non-topological defects. It has been shown by Manton that the configuration space of classical bosonic Weinberg-Salam theory has a non-contractible loop. Hence an unstable, static, finite-energy solution of the field equations can exist. Vachaspati showed that exact vortex solutions exist, which are stable to small perturbations for large values of the Weinberg angle $`\theta _W`$ and small values of the Higgs boson mass.<sup>\*∥</sup><sup>\*∥</sup>\*∥String like configurations in the Weinberg-Salam theory had been discussed earlier by Nambu in 1977 . This expectation has been confirmed by the numerical results of James, Perivolaropoulos and Vachaspati .<sup>\***</sup><sup>\***</sup>\***Somewhat disappointingly, they also found that the strings are unstable for realistic parameter values in the electroweak model. It is possible, however, that there are extensions of the Standard Model for which the string solutions are stable.
Defining
$`Z_\mu `$ $`=`$ $`\mathrm{cos}\theta _WW_\mu ^3\mathrm{sin}\theta _WY_\mu `$ (59)
$`A_\mu `$ $`=`$ $`\mathrm{sin}\theta _WW_\mu ^3+\mathrm{cos}\theta _WY_\mu ,`$ (60)
where $`\mathrm{tan}\theta _W=g^{}/g`$, the static vortex solution that extremizes the energy functional was found to be:
$$\varphi =f_{NO}(r)e^{im\theta }\left(\begin{array}{c}0\\ 1\end{array}\right),Z_\mu =A_{\mu }^{}{}_{NO}{}^{},$$
(61)
and
$$A_\mu =0=W_\mu ^a(a=1,2).$$
(62)
The subscript $`NO`$ on the functions $`f`$ and $`A_\mu `$ means that they are identical to the corresponding functions found by Nielsen and Olesen for the usual Abelian-Higgs string. In other words, the Weinberg-Salam theory has a vortex solution which is simply the $`NO`$ vortex of the Abelian Higgs model embedded in the non-Abelian theory. Substituting this solution into (56) we have
$$_\mu j_B^\mu =\frac{N_F\alpha ^2}{32\pi ^2}\mathrm{cos}2\theta _WZ_{\mu \nu }\stackrel{~}{Z}^{\mu \nu }$$
(63)
where $`\alpha =g^{}\mathrm{sin}\theta _W+g\mathrm{cos}\theta _W`$ and $`\alpha ^2=g^2+g^2`$.
The change in the baryon number is now
$$\mathrm{\Delta }Q_B=\frac{N_F\alpha ^2}{16\pi ^2}q\mathrm{\Delta }d^3x𝐙\times 𝐙$$
(64)
via equations (58), (63) and (53) and with $`q=\mathrm{cos}2\theta _W`$ playing the role of baryon number charge. We see that the change in the baryon number is proportional to the change in the $`𝐙`$ field’s helicity. For a configuration of tangled strings with arbitrary winding numbers $`N_i`$
$$\mathrm{\Delta }Q_B=\frac{N_F\alpha ^2}{16\pi ^2}q\mathrm{\Delta }\left(\underset{ij}{}_{ij}N_iN_j+\underset{i}{}n_iN_i^2\right)\left(\frac{4\pi }{\alpha }\right)^2$$
(65)
where the flux of each vortex is $`4\pi N_i/\alpha `$.<sup>\*††</sup><sup>\*††</sup>\*††The covariant derivative for this model is $`D_\mu =_\mu +ı{\displaystyle \frac{\alpha }{2}}Z_\mu `$and hence the factor of $`2`$ for the flux. The final result for a configuration of strings with unit winding number is
$$\mathrm{\Delta }Q_B=N_Fq\mathrm{\Delta }\left(\underset{ij}{}_{ij}+\underset{i}{}n_i\right)$$
(66)
and baryon number is quantized.
## VI Summary
Earlier works failed to fully implement the topological interpretation of helicity as the measure of the linkage of field lines of the divergence free field. Since field lines may be twisted and linked inside a vortex core, the inner structure of the field inside the core may not be ignored when calculating helicity terms. Neglecting internal helicity had led to peculiar results: unacceptable constraints on string configurations and continuous violation of baryon number. Once we add the contribution from the internal structure of the strings, we find new constraints on the configurations of linked global cosmic strings, which are physically more pleasing, and we also find that, the baryon number is quantized as we would expect.
Our work is another demonstration of the fact that helicity counts the linkage of the divergence free field lines. We explain how there is no contradiction between the fact that helicity counts the linkage of field lines, and that the total helicity of any configuration of knotted linked loops always manages to vanish. The vector $`𝐯`$ (10) constructed for the purpose of obtaining a helicity term for global cosmic strings is proportional to $`\theta `$. Hence it is perpendicular to $`\times 𝐯`$ (11) and the helicity must vanish. On the other hand the helicity counts the linkage of field lines of $`\times 𝐯`$, and these lines must have external and internal linkages for linked configurations of cosmic string loops. What we have discovered is that the field lines arrange themselves in such a manner that the external and internal helicities exactly cancel each other. The field lines inside a string may not link in an arbitrary manner, their linkage is constrained by the topological configuration of the string and by its linking relations to other strings. Further, we have discovered a constraint on the permitted configurations of un-knotted linked loops, in the form of (44) which must produce integer values. This constraint results from the combination of the topological character of helicity integrals with the topological nature of cosmic strings as topological defects.
In this work we confined ourselves to the problem of un-knotted linked loops. The problem of knotted configurations is far more complicated since it is hard to separate the external contribution to helicity (depending on the topology of the knot) from the internal contribution depending on the twist of the field lines inside the knot. Moffatt proposed a method for calculating the external helicity, but it turns out that for some knotted configurations this method is ambiguous (it may be that it works only for chiral knots, i.e. knots that are not isotopic to their mirror images). Hence, we leave the treatment of knots to later work.
For local cosmic strings in the Standard Model, the change in the helicity of the electroweak magnetic field confined to the strings core is related to the change in baryon number over a period of time (64). Thus, baryon number conservation is violated as cosmic strings link and de-link. The idea that violation of baryon number may be related to the change in the linking of electroweak strings has been developed by Cohen and Manohar , Vachaspati and Field and Garriga and Vachapati . However, they neglected the contribution of the writhe of the strings to helicity. Sato adds the writhe term, but misses out the twist term. This led to his conclusion that baryon number conservation is violated as vortices change their shape, which implies that baryon number may change continuously and, therefore, is not quantized. Only by including both the writhe and the twist terms, which together give the internal linkage of the field lines, do we recover a quantized baryon number!
Charge quantization has previously been related to flux quantization. It was suggested by Dirac in 1931 to explain the quantization of the electric charge. This idea was developed further by Jehle . In general, quantization of charges is not a direct consequence of standard (non-GUT) field theories. Particularly, the quantization of electric charge has no widely accepted explanation within the Standard Model. During the last decade it has been suggested that the electric charges can be heavily constrained within the framework of the Standard Model. This is achieved partly by constraints related to the classical structure of the theory (such as the requirement that the Lagrangian be gauge-invariant), and from the cancelation of anomalies, at the quantum level. The novel idea in this work is that quantization of the baryon number may arise from the topological nature of helicity integrals calculated on the background of local cosmic strings and from the flux quantization in the strings.
## Acknowledgments
I thank Prof. Jacob. D. Bekenstein for his guidance during the course of this work and for many discussions. I am also grateful to Dr. L.Sriramkumar for many stimulating discussions. This research is supported by a grant from the Israel Science Foundation, established by the Israel Academy of Sciences and Humanities.
|
no-problem/9908/cond-mat9908421.html
|
ar5iv
|
text
|
# Untitled Document
Title: An integrable model for the integer quantum Hall transition I: The vertex model Author: R.M. Gade Comments: Paper withdrawn by the autor. A revised analysis of the model will appear.
|
no-problem/9908/patt-sol9908004.html
|
ar5iv
|
text
|
# Bilateral symmetry breaking in a nonlinear Fabry-Perot cavity exhibiting optical tristability
## Abstract
We show the existence of a region in the parameter space that defines the field dynamics in a Fabry-Perot cylindrical cavity, where three output stable stationary states of the light are possible for a given localized incident field. Two of these states do not preserve the bilateral (i.e. left-right) symmetry of the entire system. These broken-symmetry states are the high-transmission nonlinear modes of the system. We also discuss how to excite these states.
Symmetries are at the backbone of physical theories. In Hamiltonian systems, symmetries are related to conserved quantities through Noether’s theorem, and they provide us with a powerful tool to understand Nature. The left-right symmetry of the fundamental laws of Nature (apart from the nonconservation of parity for the weak interaction) is a well-known fact . In particular, the electromagnetic interaction possesses this symmetry at the fundamental level. However, there is no reason why an invariance of the evolution equations should be an invariance of the stationary states of the system . In fact, most of the time the state of really big systems does not have the symmetry of the laws which govern it . Here in a simple symmetrical dynamical system with a symmetric localized driving source, we have found an example of the dynamical breaking of a discrete fundamental symmetry of the system, namely, its left-right, or bilateral, symmetry.
Shortly after the proposal and demonstration of nondissipative optical bistability using Fabry-Perot cavities , the effect on the bistable behaviour of the transverse field amplitude profile of the fundamental mode of the cavity was taken into account . Nonlinearity and diffraction can excite several transverse modes of the cavity, and this will considerably modify the dynamics of the light . For high finesse cavities, when only one longitudinal mode of the cavity is excited, the dynamics can be described with a single scalar wave equation , which facilitates numerical analysis and physical interpretation. In this paper, we restrict ourselves to this regime. Generally speaking, a great variety of light dynamics can be observed . Concerning bistability, it has been shown that diffraction gives rise to transverse instability in one of the two branches of the plane wave bistable regime . This raises the question if bistability can exist at all when important transverse effects are taken into account. Some families of stationary solutions have been found , but their role in the dynamics of the light has not been discussed. Here we consider new families of stationary solutions with broken symmetry which are relevant to the dynamics of the light that we numerically observed, and we discuss their stability and excitation.
We consider a cylindrical Fabry-Perot cavity filled with a self-focusing Kerr nonlinear medium, $`n=n_0+n_2|E|^2`$, where $`n`$ is the refractive index of the medium, $`n_0`$ is the low-power refractive index, $`n_2>0`$ is the nonlinear coefficient, and $`E`$ is the electric field inside the cavity (see Fig. 1). The cavity is driven by a gaussian beam $`E_i`$ with beam widths in the transverse $`x`$ and $`y`$ directions given by $`w_x`$ and $`w_y`$ ($`1/e^2`$ half width intensity). The concave mirror in the $`y`$-direction has a radius of curvature $`R`$, and the width $`w_y`$ is chosen so that the incident beam matches the linear Hermite-Gaussian mode of the cavity in the $`y`$-direction. Both mirrors have very high reflectivity $``$, and the incident beam is assumed to have a Rayleigh range in the $`x`$ direction $`z_x=\pi w_x^2n_0/\lambda _0`$ much larger than the cavity length $`L`$. $`\lambda _0`$ is the light wavelength in vacuum. Under these conditions, only a single longitudinal mode of the cavity is excited and the electric field $`E`$ has the structure $`E(x,y,z,t)\mathrm{\Psi }(x,t)\mathrm{exp}(y^2/w_y^2)\mathrm{sin}(k_cz)\mathrm{exp}(i\omega t)+c.c.`$, where $`\mathrm{\Psi }`$ is the normalized complex field envelope inside the cavity, $`\omega `$ is the frequency of the input beam $`E_i`$, $`t`$ is the time and $`k_c`$ is the resonance wavenumber of the linear cavity. Only one transverse mode in the $`y`$ dimension is selected out by the use of the cylindrical cavity. The field transmitted by the system is proportional to $`\mathrm{\Psi }`$ , which obeys the driven Ginzburg-Landau equation
$$\frac{\mathrm{\Psi }}{\tau }=\frac{i}{2}\frac{^2\mathrm{\Psi }}{\xi ^2}i\theta \mathrm{\Psi }+i|\mathrm{\Psi }|^2\mathrm{\Psi }+\mathrm{\Gamma }(\mathrm{\Psi }_d\mathrm{\Psi })$$
(1)
where $`\mathrm{\Psi }_d`$ is the driving field, which is proportional to the incident field, $`\xi =x/w_x`$ is the normalized $`x`$ coordinate and $`\tau =t/t_0`$ is the normalized time. The normalizing factor $`t_0`$ is given by $`t_0=2n_0z_x/c`$, where $`c`$ is the velocity of light in vacuum. The normalized cavity detuning $`\theta `$ is given by $`\theta =2\mathrm{\Delta }kz_x`$, where $`\mathrm{\Delta }k=k_ck`$ and the wavenumber $`k`$ is $`k=n_0\omega /c`$. The amplitude decay rate $`\mathrm{\Gamma }`$ is $`\mathrm{\Gamma }=(𝒯+\alpha L)z_x/L`$, where $`𝒯=1`$ is the transmissivity of each mirror, and $`\alpha `$ is the loss coefficient of the material filling the cavity. The normalized amplitude $`q`$ of the driving field ($`q\mathrm{\Psi }_d(0)`$), the normalized cavity detuning $`\theta `$, and the amplitude decay rate $`\mathrm{\Gamma }`$ define the parameter space that determines the dynamics of the light in the Fabry-Perot etalon.
Let us consider a laser light that excites $`{}_{}{}^{85}\mathrm{Rb}`$ atoms in their D<sub>2</sub> transition near $`\lambda _0=780`$ nm to provide a resonantly-enhanced optical nonlinearity. The mirror reflectivity is $``$=0.995 and the cavity length $`L=2`$ mm. The intracavity loss is $`\alpha 0.1`$ dB/cm and the linear refractive index is taken to be $`n_01`$. A beam width $`w_x100\mu `$m corresponds to $`\mathrm{\Gamma }0.15`$, and $`\theta 0.4`$ corresponds to a cavity frequency detuning $`\mathrm{\Delta }f240`$ MHz. The Rayleigh range in the $`x`$ direction is $`z_x40`$ mm. The mode beam waist in the $`y`$ direction is $`w_y=\sqrt{g^2\lambda L/n_0\pi (1g)^2}`$, where $`g=1L/R`$. For $`g=0.6`$, we obtain $`w_y28\mu `$m, that can be achieved through an appropriate cylindrical mode-matching lens. The maximum nonlinear refractive index change $`\mathrm{\Delta }n`$ is given by $`\mathrm{\Delta }n=\lambda |\mathrm{\Psi }|^2/4\pi z_xA`$, where $`A=3/2\sqrt{8}`$ is a mode overlap factor given by the field profile in the $`y`$ and $`z`$ directions. For the parameters considered here, $`\mathrm{\Psi }1`$ corresponds to $`\mathrm{\Delta }n3\times 10^6`$, which has been observed in a cylindrical Fabry-Perot experiment in rubidium vapor .
The time evolution of the field $`\mathrm{\Psi }`$, as described by Eq. (1), results from the interplay between different processes: transverse effects arising from the diffraction (second derivative) term in the wave equation, self-focusing due to the nonlinearity, energy input due to the driving field and energy loss due to the finite transmissivity of both mirrors. The actual values of the parameters $`\{\mathrm{\Gamma },\theta ,q\}`$, will determine the influence on light dynamics of every one of the above mentioned processes. For large beam width, we have $`\mathrm{\Gamma }`$,$`\theta 1`$, so transverse effects are expected to be negligible in this limit. For small beam width, but large enough in order to be in the regime of validity of Eq. (1), the different competing processes depicted above are comparable. This is the regime we are considering in this paper. We will see that there is a region in parameter space where three possible output states with different broken and unbroken symmetries can be excited, and under appropriate circumstances, all stable output states can be excited. In all our considerations we will restrict ourselves to the case $`\theta >0`$, corresponding to a red-detuning of the laser from the cavity resonance. Finally, we point out that the total energy inside the cavity and the transmitted beam power are both proportional to $`I=_{\mathrm{}}^{\mathrm{}}|\mathrm{\Psi }|^2𝑑\xi `$, and this varies at a rate
$$\frac{dI}{dt}=2\mathrm{\Gamma }_{\mathrm{}}^{\mathrm{}}|\mathrm{\Psi }|^2𝑑\xi +_{\mathrm{}}^{\mathrm{}}[\mathrm{\Psi }^{}\mathrm{\Psi }_d+\mathrm{\Psi }\mathrm{\Psi }_d^{}]𝑑\xi $$
(2)
Thus Eq. (1) corresponds to a non-conservative system.
Let us first review the main results for the stability of plane wave solutions against transverse perturbations . In the plane wave limit, when the second derivative of the field in Eq. (1) is neglected, the curve $`|\mathrm{\Psi }_0|\mathrm{\Psi }_d`$, where $`\mathrm{\Psi }_0`$ are the stationary plane wave solutions, is single-valued for $`\theta <\sqrt{3}\mathrm{\Gamma }`$, whereas for $`\theta >\sqrt{3}\mathrm{\Gamma }`$, is S-shaped and can lead to bistability. Stationary plane-wave solutions are unstable against transverse perturbation when both conditions $`|\mathrm{\Psi }_0|^2>\mathrm{\Gamma }`$ and $`|\mathrm{\Psi }_0|^2>\theta /2`$ are fulfilled. Since the upper branch of the S-shaped curve $`|\mathrm{\Psi }_0|\mathrm{\Psi }_d`$ always begins at $`|\mathrm{\Psi }_0|^2>2\theta /3`$, it turns out that the upper branch is always unstable against transverse perturbation. This raises the question whether or not bistability exists at all when one takes into account the transverse structure of the beams.
For that purpose, we first look for stationary solutions of Eq. (1). We set $`\mathrm{\Psi }/\tau =0`$, and solve the corresponding ordinary differential equation with a Newton-Raphson scheme . Note that contrary to travelling wave configurations, where stationary solutions are related to a nonlinear wavenumber shift , in the cavity configuration this is not the case, since the presence of the driving field makes Eq. (1) no longer invariant under a global phase change. We do not explore all the variety of stationary solutions that Eq. (1) can support. Instead, we restrict ourselves to the stationary solutions that will play a role in our discussion of bistability. We plot in Fig. 2 families of stationary solutions corresponding to $`\mathrm{\Gamma }=0.15`$ and $`\theta =0.4`$. Stationary solutions plotted in Fig. 2 can be divided into two families. The curve labeled $`S`$ corresponds to symmetric solutions $`\mathrm{\Psi }_0(\xi )`$, such that $`\mathrm{\Psi }_0(\xi )=\mathrm{\Psi }_0(\xi )`$. The curve labeled $`B`$ corresponds to to a pair of broken-symmetry solutions, $`\mathrm{\Psi }_0(\xi )`$ and $`\mathrm{\Psi }_0(\xi )`$. The curve $`S`$ has three branches defined by the sign of the slope of the curve in the figure, and we will refer to them as lower, middle, and upper branches. Figure 3 shows the amplitude profiles of two stationary solutions, one with broken symmetry and the other with unbroken symmetry. Note the breaking of the left-right symmetry of the output field relative to that of the drive, for the broken-symmetry state.
To investigate the stability of the stationary solutions plotted in Fig. 2, we solve Eq. (1) with a split-step Fourier Transform algorithm . We take as input some selected slightly perturbed stationary solution $`\mathrm{\Psi }(\xi ,\tau =0)=\mathrm{\Psi }_0(\xi )\left[1+v(\xi )\right]`$, where $`v(\xi )`$ is a complex gaussian random variable for every value of the transverse coordinate with mean $`\mu =0`$ and typical deviation $`\sigma =0.01`$, for both the real and imaginary parts. The perturbative noise will seed any instability present in the system. Since some of the stationary solutions have broken symmetry, in order to monitor the time evolution of the field, in addition to the peak amplitude, we will also follow the centroid $`\overline{\xi }`$, defined as
$$\overline{\xi }(\tau )=\frac{_{\mathrm{}}^{\mathrm{}}\xi |\mathrm{\Psi }(\xi ,\tau )|^2}{_{\mathrm{}}^{\mathrm{}}|\mathrm{\Psi }(\xi ,\tau )|^2}.$$
(3)
For a symmetric beam, $`\overline{\xi }(\tau )=0`$. After extensive numerical solutions, we find stable stationary solutions in the lower branch of curve $`S`$, and in curve $`B`$ below $`q1.48`$. This means that the outcome of the simulations are the corresponding stationary solutions $`\mathrm{\Psi }_0`$ in each case. Stationary solutions corresponding to all other parts of curves $`S`$ and $`B`$ have been found to be unstable. In some cases, the stability analysis leads to one or the other of the two corresponding stable stationary solutions for a given amplitude of the driving field, which one depending on the noise. In other cases, the output is a time-varying pattern that depends on the particular value of the peak amplitude of the driving field. In most cases, when the stationary solution of the lower branch of curve $`S`$ is not excited, the output of simulations always show that the beam does not preserve the bilateral symmetry of the driving field. For values of the peak amplitude of the driving field $`1.37<q<1.48`$, there are three stable stationary solutions. The one corresponding to the lower branch of curve $`S`$, possesses an unbroken symmetry, and the transmissivity, defined as the ratio of the output beam power to the incident beam power, is low ($`1\%`$). The pair of solutions corresponding to curve $`B`$, one displaced to the left, the other to the right of the drive beam by equal amounts, possesses broken symmetry. The transmissivity of this pair of solutions is remarkably high ($`75\%`$). Thus the nonlinear Fabry-Perot cavity can be viewed as a nonlinear transmission device which selects out the broken-symmetry solutions as the high-transmission modes of the system.
Numerical simulations of Eq. (1) show that the symmetric state in the lower branch of curve $`S`$ can be excited after switching on the driving field with the appropriate amplitude of the driving field. In order to excite the broken-symmetry states, a route to bistability has to be properly devised. Which one of the two broken-symmetry states is actually excited in the numerical simulations shown in Figs. 4 and 5, depends on round-off noise, and the method of excitation. One method is to tune the amplitude of the driving field in a step-like way to a value $`q>1.98`$, so that a time-varying intermediate output state associated with curve $`B`$ is excited. Then the amplitude of the driving field $`q`$ is reset to the value where the broken-symmetry state is stable. We plot an example of such route to tristability in Fig. 4. Note the delay on the onset of the spontaneous symmetry breaking from the time the drive is switched down. The actual excitation of the broken-symmetry state depends on the characteristics of the intermediate unstable state that is excited, and on the duration of the round-trip time around the hysteresis loop. Another way to excite the stable state in curve $`B`$, it is to switch on the driving field and then to slowly ramp it down. Figure 5 show the actual excitation of both the symmetric and broken-symmetry states by this method.
We anticipate that these results, which we have obtained at the nonlinear classical field level, should be related to the scattering resonances of the underlying quantum field theory, in view of the correspondence principle.
This work has been supported by the Spanish Government under contract PB95-0768. Juan P. Torres is grateful to the Spanish Government for funding of his sabbatical leave through the Secretaría de Estado de Universidades, Investigación y Desarrollo. Jack Boyce and Raymond Chiao acknowledge support of the ONR and the NSF. We would also like to thank Eric Bolda, John Garrison, Morgan W. Mitchell, and Ewan Wright for very helpful discussions.
|
no-problem/9908/math9908075.html
|
ar5iv
|
text
|
# Topological characterization of torus groups
## 1. Introduction
The problem of topological characterization of torus groups (i.e., powers of the circle group 𝕋) has been of considerable importance in the theory of compact Abelian groups. The following statement collects most of the known information.
###### Theorem A.
Let $`G`$ be a connected compact Abelian group. Then the following conditions are equivalent:
* $`G`$ is arcwise connected.
* $`G`$ is locally arcwise connected.
* The map $`\mathrm{exp}_G:L(G)G`$ is surjective.
* The map $`\mathrm{exp}_G:L(G)G`$ is open.
* $`\mathrm{Ext}(\widehat{G},\text{})=0`$, i.e., the character group $`\widehat{G}`$ of $`G`$ is a Whitehead group.
* $`\mathrm{Ext}(\widehat{G},\text{})=0`$ and each pure subgroup of $`\widehat{G}`$ of finite rank splits.
In addition, if $`G`$ is metrizable, then the above conditions are equivalent to each of the following:
* $`\widehat{G}`$ is free.
* $`\widehat{G}`$ is $`\mathrm{}_1`$-free, i.e., every countable subgroup of $`\widehat{G}`$ is free.
* $`\widehat{G}`$ is projective.
* $`G`$ is injective (in the category of compact Abelian groups).
* $`G`$ is locally connected.
* $`G`$ is a torus.
In the nonmetrizable case situation is quite different. For instance, the equivalence $`(8)(11)`$ remains true, whereas $`(11)(12)`$ does not. The equivalence $`(5)(7)`$ is undecidable in ZFC . In topological terms this means that:
* There are (necessarily nonmetrizable) connected and locally connected compact Abelian groups which are not tori.
* It is undecidable within ZFC that the arcwise connectedness of a nonmetrizable compact Abelian group forces the group to be isomorphic to a torus.
Below we show (Theorem E) that a simple topological property gives a complete characterization of torus groups. Statements of this sort are not unusual. Here are some examples (see also ).
###### Theorem B (Anderson-Kadec , $`\tau =\omega `$; $`\tau >\omega `$).
The following conditions are equivalent for an infinite-dimensional locally convex linear topological space $`E`$ of weight $`\tau \omega `$:
* $`E`$ is topologically equivalent to $`R^\tau `$ (the $`\tau `$-th power of the real line).
* $`E`$ is an $`AE(0)`$-space.
###### Theorem C (Pontryagin-Haydon, , ).
Every compact group is an $`AE(0)`$-space.
###### Theorem D ().
The following conditions are equivalent for a zero-dimensional topological group $`G`$:
* $`G`$ is topologically equivalent to the product $`(\text{}_2)^\tau \times \text{}^\kappa `$.
* $`G`$ is an $`AE(0)`$-space.
Our goal in these notes is to prove the following statement.
###### Theorem E.
The following conditions are equivalent for a compact Abelian group $`G`$:
* $`G`$ is a torus group (both in topological and algebraic senses).
* $`G`$ is an $`AE(1)`$-compactum.
* The map $`\mathrm{exp}_G:L(G)G`$ is $`0`$-soft.
There are various equivalent versions of definitions of the concepts of $`AE(n)`$-spaces and $`n`$-soft maps. In this particular situation categorical terminology seems to be more appropriate. Suppose that $`𝓑`$ is a full subcategory of a category $`𝓐`$. An object $`a\mathrm{Ob}(𝓐)`$ is called $`𝓑`$-injective, if for any monomorphism $`i\mathrm{Mor}_𝓑(b,c)`$ of the category $`𝓑`$ and for any morphism $`f\mathrm{Mor}_𝓐(b,a)`$ there exists a morphism $`\stackrel{~}{f}\mathrm{Mor}_𝓐(c,a)`$ such that $`f=\stackrel{~}{f}i`$.
Let now $`𝓣𝓨𝓒𝓗`$ denote the category of Tychonov spaces and their continuous maps and $`𝓣𝓨𝓒𝓗_n`$ be the full category consisting of at most $`n`$-dimensional (in the sense of $`dim`$) spaces. Then $`AE(n)`$-spaces are precisely $`𝓣𝓨𝓒𝓗_n`$-injective objects of $`𝓣𝓨𝓒𝓗`$. $`n`$-soft maps can be defined similarly: these are precisely $`𝑴𝒐𝒓(𝓣𝓨𝓒𝓗_n)`$-injective objects of the category $`𝑴𝒐𝒓(𝓣𝓨𝓒𝓗)`$. The class of metrizable $`AE(0)`$-spaces coincides with the class of separable completely metrizable spaces and in positive dimensions the strictly decreasing sequences
$$AE(1)AE(2)\mathrm{}AE(n)AE(n+1)\mathrm{}$$
and
$$LC^0C^0LC^1C^1\mathrm{}\mathrm{}LC^{n1}C^{n1}LC^nC^n\mathrm{}$$
are identical (i.e., $`AE(n)=LC^{n1}C^{n1}`$) in the presence of metrizability and differ (i.e. $`AE(n)LC^{n1}C^{n1}`$) in general. Important examples of nonmetrizable $`AE(0)`$-spaces are uncountable powers of the real line. This fact is exploited below when considering the above mentioned space $`L(G)\text{}^{\mathrm{rank}(\widehat{G})}`$. Every $`0`$-soft map is open and surjective and for maps between completely metrizable separable spaces the converse is also true. This fact is also exploited when considering the map $`\mathrm{exp}_G:L(G)G`$.
Since, by duality, any statement in the category of compact Abelian groups generates an equivalent statement in the category of Abelian groups (see Theorem A), it might be worth examining Theorem E from this point of view. By the equivalence $`(7)(12)`$ in Theorem A, the statement corresponding to Theorem E in the category of Abelian groups must provide a characterization of free groups. Since $`dimG=\mathrm{rank}(\widehat{G})`$ for a finite dimensional compact Abelian group $`G`$ and since the concept of $`AE(1)`$ is defined in terms $`1`$-dimensional testing compacta, this would lead us to the concept of $`1`$-projectivness (restrict, in the standard definition of projectivness, domains of testing epimorphisms by groups of rank $`1`$) for Abelian groups and to a hope of subsequent characterization of free Abelian groups in these terms. The so obtained version of $`AE(1)`$ will differ from the original topological concept of $`AE(1)`$ and will not be a replacement of the latter in Theorem E. Reason is simple - every $`1`$-dimensional connected compact (Abelian) group is metrizable and, consequently, the changed version of $`AE(1)`$ would essentially coincide with the arcwise connectedness. This shows that Theorem E provides a purely topological characterization of tori and leaves the problem of characterizing free uncountable Abelian groups until Section 4.
## 2. Preliminaries
All spaces below are Tychonov (i.e., completely regular and Hausdorff) and maps are continuous. , and $`\text{𝕋}=\text{}/\text{}`$ stand for the additive group of real numbers, the integers and the circle group respectively. We assume the familiarity with the general theory of $`ANE(n)`$-spaces and $`n`$-soft maps as well as with the spectral machinery based on the Ščepin’s Spectral Theorem (a comprehensive discussion of these topics including a noncompact version of the Spectral Theorem can be found in ). The space $`L(G)`$, defined by letting $`L(G)\stackrel{\mathrm{def}}{=}\mathrm{Hom}(\text{},G)`$, is a real topological linear space (with the topology of uniform convergence on compact sets) algebraically and topologically isomorphic to $`\text{}^{\mathrm{rank}(\widehat{G})}`$. The continuous homomorphism $`\mathrm{exp}_G:L(G)G`$ is defined naturally: $`\mathrm{exp}_G(f)=f(1)`$, $`fL(G)`$. If $`p:GT`$ is a continuous homomorphism of compact Abelian groups, then the map $`L(p):L(G)L(T)`$, defined by letting $`L(p)(f)=pf`$, $`fL(G)`$, is a continuous homomorphism.
As was noted above $`X`$ is an $`AE(1)`$-compactum if any map $`g_0:Z_0X`$, defined on a closed subspace $`Z_0`$ of an at most $`1`$-dimensional compactum $`Z`$, admits an extension $`g:ZX`$. Note that the arcwise connectedness of $`X`$ is obtained by restricting the previous definition to the one particular case: $`Z_0=\{0,1\}`$ and $`Z=[0,1]`$. $`1`$-soft maps are defined similarly. A map $`f:XY`$ between compacta is $`1`$-soft if for any maps $`g_0:Z_0X`$ and $`h:ZY`$ such that $`fg_0=h`$, where $`Z_0`$ is a closed subspace of an at most $`1`$-dimensional compactum $`Z`$, there exists a map $`g:ZX`$ such that $`g_0=g|Z_0`$ and $`h=fg`$.
When a well-ordered inverse spectrum $`𝒮_G=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$ is used to investigate properties of its limit, a crucial information is contained not only in spaces $`G_\alpha `$ but in short projections $`p_\alpha ^{\alpha +1}:G_{\alpha +1}G_\alpha `$ as well. In other words, in order to obtain information about the limit object one needs information about morphisms of the spectrum representing the given object. This shows how Lemma 3.1 is related to Proposition 3.2.
If two well-ordered spectra $`𝒮_G=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$ and $`𝒮_T=\{T_\alpha ,q_\alpha ^{\alpha +1},\tau \}`$ are given and we wish to investigate properties of a map $`f:lim𝒮_Glim𝒮_T`$, even an ability to obtain a complete information hidden in spaces and short projections of these spectra will not suffice. This is not surprising at all since spectra $`𝒮_G`$ and $`𝒮_T`$, while describing domain and range of $`f`$, have nothing to do with $`f`$ itself. After accepting this conclusion one would naturally make the next step which would basically consist of representing the map $`f`$ itself as the limit of “level maps”. In other words we need to find a morphism
$$\{f_\alpha :G_\alpha T_\alpha ,\tau \}:𝒮_G𝒮_T$$
limit of which is $`f`$. Here is the corresponding diagram
$$\begin{array}{ccccccccccc}lim𝒮_G& & \mathrm{}& & G_{\alpha +1}& \stackrel{p_\alpha ^{\alpha +1}}{}& G_\alpha & & \mathrm{}& & G_0\\ f& & & & f_{\alpha +1}& & f_\alpha & & & & f_0& & \\ lim𝒮_T& & \mathrm{}& & T_{\alpha +1}& \stackrel{q_\alpha ^{\alpha +1}}{}& T_\alpha & & \mathrm{}& & T_0\end{array}$$
Diagram 1
Representation of $`f`$ as the limit of such a morphism is not always possible. Trivial counterexamples can be found for maps between metrizable compacta. But according to “the effect of uncountability” discovered by Ščepin, any map between nonmetrizable compacta admits such a representation (this is basically what the Spectral Theorem states). Having now information about $`f_\alpha `$’s available in principle we find ourselves in a situation described above. This information will not suffice to restore properties of $`f`$. This is information about objects and information needed is, as we have already seen, hidden in morphisms. What are objects and what are morphisms in this case? Clearly $`f_\alpha `$’s are objects (representing the given object $`f`$) and morphisms are nothing else but commutative square diagrams
$$\begin{array}{ccc}G_{\alpha +1}& \stackrel{f_{\alpha +1}}{}& T_{\alpha +1}\\ p_\alpha ^{\alpha +1}& & q_\alpha ^{\alpha +1}& & \\ G_\alpha & \stackrel{f_\alpha }{}& T_\alpha .\end{array}$$
Diagram 2
The rest is a matter of technicalities. How one would investigate properties of such square diagrams? A possible approach is through their characteristic maps. In the above notations this is the diagonal product $`\chi :p_\alpha ^{\alpha +1}\mathrm{}f_{\alpha +1}:G_{\alpha +1}K`$, where $`K`$ is the pullback
$$K=\{(g,t)G_\alpha \times T_{\alpha +1}:f_\alpha (g)=q_\alpha ^{\alpha +1}(t)\}$$
of maps $`f_\alpha `$ and $`q_\alpha ^{\alpha +1}`$. Here is the corresponding diagram
Diagram 3
Obviously $`\chi `$ measures how close the above diagram is to the pullback square (or the Cartesian square in the terminology of ).
It is easy to see that the surjectivness of all vertical arrows in Diagram 1 does not imply surjectivness of $`f`$. Explanation is simple: these are requirements about $`f_\alpha `$’s, i.e. objects, but not about morphisms. Having this in mind, it is clear that requesting additionally surjectivness of characteristic maps of all square diagrams in Diagram 1 (thus imposing conditions on morphisms), one would now expect that the limit would also be surjective. This is indeed true. Moreover it can be shown that this is the only right way to proceed: if one starts with surjective $`f`$, then “almost all” square diagrams formed by any morphism representing $`f`$ would have surjective characteristic maps. Same applies to the openness of $`f`$. Requesting openness of not only $`f_\alpha `$’s but of all characteristic maps of all square diagrams we arrive to the $`0`$-softness of the limit map $`f`$. Conversely, $`0`$-soft maps are precisely those which can be obtained in this way. Noting that there are open but not $`0`$-soft maps, we see where the difference between Theorems A and E comes from. This also shows how Lemma 3.3 is related to Proposition 3.4.
Definitions of few remaining concepts are discussed below. For the readers convenience we mainly cite where a complete discussion, original sources and omitted details can be found.
## 3. Characterization of torus groups
The following simple and well-known fact plays an important role below. A standard proof is based on Pontryagin-van Kampen duality (I thank K. H. Hofmann for pointing this to me out). Here we present its topological translation.
###### Lemma 3.1.
Let $`p:GT`$ be a continuous surjective homomorphism between compact Abelian groups. If $`\mathrm{ker}(p)`$ is a metrizable $`AE(1)`$-compactum, then $`p`$ is (in both topological and algebraic senses) a trivial fibration.
###### Proof.
Since the Abelian group $`\mathrm{ker}(p)`$ is a metrizable $`AE(1)`$-compactum, it is a torus group (the equivalence $`(11)(12)`$ of Theorem A). Since torus groups are injective objects in the category of compact Abelian groups (the equivalence $`(10)(12)`$ of Theorem A), there exists a continuous retraction $`r:G\mathrm{ker}(p)`$ which is a homomorphism. Obviously, there is an isomorphism $`h:G\mathrm{ker}(r)\times \mathrm{ker}(p)`$ (here is the explicit formula: $`h(g)=(gr(g),r(g))`$, $`gG`$). Now observe that $`q=p|\mathrm{ker}(r):\mathrm{ker}(r)T`$ is an isomorphism. Consequently, the following diagram
$$\begin{array}{ccc}G& \stackrel{h}{}& \mathrm{ker}(r)\times \mathrm{ker}(p)\\ p& & \mathrm{pr}_1& & \\ T& \stackrel{q^1}{}& \mathrm{ker}(r)\end{array}$$
commutes. This proves the lemma. ∎
The following statement provides a characterization of $`AE(1)`$-compact Abelian groups in terms of well-ordered inverse spectra. It should be especially emphasized that characterization of these groups in terms of $`\omega `$-spectra is an immediate consequence of \[3, Proposition 6.3.5 and Lemma 8.2.1\] and Ščepin’s Spectral Theorem for compacta \[3, Theorem 1.3.4\]. But for establishing our results we need well-ordered spectra - they allow us to proceed by induction. On the other hand for well ordered spectra we are forced to reprove already known topological versions in the presence of group structure. The latter only slightly simplifies a typical proof (see ) although the outcome is much stronger.
###### Proposition 3.2.
Let $`G`$ be a compact Abelian group of weight $`\tau >\omega `$. Then the following conditions are equivalent:
* $`G`$ is the torus $`\text{𝕋}^\tau `$.
* $`G`$ is an $`AE(1)`$-compactum.
* There exists a well-ordered inverse spectrum $`𝒮_G=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$ satisfying the following properties:
1. $`G_\alpha `$ is a compact Abelian group and $`p_{\alpha +1}:G_{\alpha +1}G_\alpha `$ is a continuous homomorphism, $`\alpha <\tau `$.
2. If $`\beta <\tau `$ is a limit ordinal, then the diagonal product
$$\mathrm{}\{p_\alpha ^\beta :\alpha <\beta \}:G_\beta lim\{G_\alpha ,p_\alpha ^{\alpha +1},\alpha <\beta \}$$
is an isomorphism.
3. $`G`$ is isomorphic to $`lim𝒮`$.
4. The short projection $`p_\alpha ^{\alpha +1}:G_{\alpha +1}G_\alpha `$ is isomorphic to the trivial fibration $`G_\alpha \times \mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)G_\alpha `$ the fiber $`\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)`$ of which is a metrizable torus, $`\alpha <\tau `$.
5. $`G_0`$ is a metrizable torus.
###### Proof.
$`(a)(b)`$. Note that $`\text{𝕋}^\tau `$ is an $`AE(1)`$-compactum.
$`(c)(a)`$. Apply a straightforward transfinite induction to conclude that $`G`$ is topologically and algebraically equivalent to the product $`G_0\times \{\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right):\alpha <\tau \}`$. Properties 4 and 5 show that such a product is isomorphic to the torus group $`\text{𝕋}^\tau `$.
$`(b)(c)`$. We may assume that $`G`$ is a closed subgroup of a product $`\{X_a:aA\}`$, $`|A|=\tau `$, of compact metrizable Abelian groups $`X_a`$, $`aA`$, each of which is connected and locally connected (clearly each of this $`X_a`$’s can be chosen to be the circle group 𝕋 but in order to avoid a confusion we deliberately use a different notation). By \[6, Theorem 4.1\], there exists an open $`1`$-invertible map $`f:Y\{X_a:aA\}`$, where $`Y`$ is a $`1`$-dimensional compactum of weight $`\tau `$. Consider the inverse image $`f^1(G)Y`$ of $`G`$ and the map $`f|f^1:f^1(G)G`$. Since $`dimY=1`$ and since $`G`$, according to (b), is an $`AE(1)`$-compactum, there exists a map $`g:YG`$ such that $`g|f^1(G)=f|f^1(G)`$.
Next let us denote by
$$\pi _B:\{X_a:aA\}\{X_a:aB\}$$
and
$$\pi _C^B:\{X_a:aB\}\{X_a:aC\}$$
the natural projections onto the corresponding subproducts ($`CBA`$). We call a subset $`BA`$ admissible (compare with the proof of \[3, Theorem 6.3.1\]) if the following equality
$$\pi _B(g(f^1(x)))=\pi _B(x)$$
is true for each point $`x\pi _B^1\left(\pi _B(G)\right)`$. We need the following properties of admissible sets.
Claim 1. The union of arbitrary collection of admissible sets is admissible.
Indeed let $`\{B_t:tT\}`$ be a collection of admissible sets and $`B=\{B_t:tT\}`$. Let $`x\pi _B^1\left(\pi _B(G)\right)`$. Clearly $`x\pi _{B_t}^1\left(\pi _{B_t}(G)\right)`$ for each $`tT`$ and consequently
$$\pi _{B_t}(g(f^1(x)))=\pi _{B_t}(x)\text{for each}tT.$$
Obviously, $`\pi _B(x)\pi _B(g(f^1(x)))`$ and it therefore suffices to show that the set $`\pi _B(g(f^1(x)))`$ contains only one point. Assuming that there is a point $`y\pi _B(g(f^1(x)))`$ such that $`y\pi _B(x)`$ we conclude (having in mind that $`B=\{B_t:tT\}`$) that there must be an index $`tT`$ such that $`\pi _{B_t}^B(y)\pi _{B_t}^B\left(\pi _B(x)\right)`$. But this is impossible
$$\pi _{B_t}^B(y)\pi _{B_t}^B\left(\pi _B(g(f^1(x)))\right)=\pi _{B_t}(g(f^1(x)))=\pi _{B_t}(x)=\pi _{B_t}^B\left(\pi _B(x)\right).$$
Claim 2. If $`BA`$ is admissible, then the restriction $`\pi _B|G:G\pi _B(G)`$ is $`1`$-soft.
Let $`\phi :Z\pi _B(G)`$ and $`\phi _0:Z_0G`$ be two maps defined on a $`1`$-dimensional compactum $`Z`$ and its closed subset $`Z_0`$ respectively. Assume that $`\pi _B\phi _0=\phi |Z_0`$. We wish to construct a map $`\varphi :ZG`$ such that $`\varphi |Z_0=\phi _0`$ and $`\pi _B\varphi =\phi `$, i.e. $`\varphi `$ makes the diagram
commutative. Since, according to our choice, all $`X_a`$’s are $`AE(1)`$-compacta, so is the product $`\{X_a:aAB\}`$. This implies the $`1`$-softness of the projection $`\pi _B`$ and hence of its restriction
$$\pi _B|\pi _B^1\left(\pi _B(G)\right):\pi _B^1\left(\pi _B(G)\right)\pi _B(G).$$
Then there exists a map $`\varphi ^{\prime \prime }:Z\pi _B^1\left(\pi _B(G)\right)`$ such that $`\varphi ^{\prime \prime }|Z_0=\phi _0`$ and $`\pi _B\varphi ^{\prime \prime }=\phi `$. Since $`f`$ is $`1`$-invertible (and $`dimZ=1`$), there exists a map $`\varphi ^{}:ZY`$ such that $`f\varphi ^{}=\varphi ^{\prime \prime }`$. Now let $`\varphi =g\varphi ^{}`$. Since $`g|f^1(G)=f|f^1(G)`$, we have $`\phi _0=\varphi |Z_0`$. Finally observe that the admissibility of $`B`$ implies $`\phi =\pi _B\varphi `$ as required.
Claim 3. For each countable subset $`CA`$ there exists a countable admissible subset $`BA`$ such that $`CB`$.
Since $`w(Y)=\tau `$, it follows (consult ) that $`Y`$ can be represented as the limit space of an $`\omega `$-spectrum $`𝒮_Y=\{Y_B,q_C^B,\mathrm{exp}_\omega A\}`$ consisting of metrizable compacta $`Y_B`$, $`B\mathrm{exp}_\omega A`$, and continuous surjections $`q_C^B:Y_BY_C`$, $`CB`$, $`C,B\mathrm{exp}_\omega A`$. Consider also the standard $`\omega `$-spectrum $`𝒮_X=\{\{X_a:aB\},\pi _C^B,\mathrm{exp}_\omega A\}`$ consisting of countable subproducts of the product $`\{X_a:aA\}`$ and corresponding natural projections. Obviously the full product coincides with the limit of $`𝒮_X`$. One more $`\omega `$-spectrum arises naturally. This is the spectrum $`𝒮_G=\{\pi _B(G),\pi _C^B|\pi _B(G),\mathrm{exp}_\omega A\}`$ the limit of which coincides with $`G`$.
Consider the map $`f:lim𝒮_Ylim𝒮_X`$. Applying the Spectral Theorem (\[3, Theorem 1..3.4\]) there is a cofinal and $`\omega `$-closed subset $`𝒯_1`$ of $`\mathrm{exp}_\omega A`$ such that for each $`B𝒯_f`$ there is a map $`f_B:Y_B\{X_a:aB\}`$ such that $`f_Bq_B=\pi _Bf`$. Moreover, these maps form a morphism
$$\{f_B;B𝒯_f\}:𝒮_Y𝒮_X$$
limit of which coincides with $`f`$. Since $`f`$ is open (and closed), we may assume without loss of generality (considering a smaller cofinal and $`\omega `$-subset of $`𝒯_f`$ if necessary) that the above indicated morphism is bicommutative. This simply means that $`q_Bf^1(K)=f_B^1\left(\pi _B(K)\right)`$ for any $`B𝒯_f`$ and any closed subset $`K`$ of the product $`\{X_a:aA\}`$.
Similarly applying the Spectral Theorem to the map $`g:lim𝒮_Ylim𝒮_G`$ we obtain a cofinal and $`\omega `$-closed subset $`𝒯_g`$ of $`\mathrm{exp}_\omega A`$ and the associated to it morphism
$$\{g_B:Y_B\pi _B(G);B𝒯_g\}:𝒮_Y𝒮_G$$
limit of which coincides with the map $`g`$.
By \[3, Proposition 1.1.27\], the intersection $`𝒯=𝒯_f𝒯_g`$ is still a cofinal and $`\omega `$-closed subset of $`\mathrm{exp}_\omega A`$. It therefore suffices to show that each $`B𝒯`$ is an admissible subset of $`A`$. Consider a point $`x\pi _B^1\left(\pi _B(G)\right)`$. First observe that bicommutativity of the morphism associated with $`𝒯_f`$ implies that $`q_B(f^1(x))=f_B^1(\pi _B(x))`$. Since the maps $`f_B`$ and $`g_B`$ coincide on $`f_B^1(\pi _B(G))`$ we have
$$\begin{array}{c}\pi _B(g(f^1(x)))=g_B(q_B(f^1(x)))=g_B(f_B^1(\pi _B(x)))=\hfill \\ \hfill f_B(f_B^1(\pi _B(x)))=\pi _B(x)\end{array}$$
as required.
Claim 4. If $`C`$ and $`B`$ are admissible subsets of $`A`$ and $`CB`$, then the map $`\pi _C^B|\pi _B(G):\pi _B(G)\pi _C(G)`$ is $`1`$-soft.
This property follows from Claim 2 and \[3, Lemma 6.1.15\].
After having all the needed properties of admissible subsets established we proceed as follows. Since $`|A|=\tau `$ we can write $`A=\{a_\alpha :\alpha <\tau \}`$. By Claim 3, each $`a_\alpha A`$ is contained in a countable admissible subset $`B_\alpha A`$. Let $`A_\alpha =\{B_\beta :\beta \alpha \}`$. We use these sets to define a transfinite inverse spectrum $`𝒮=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$ as follows. Let $`G_\alpha =\pi _{A_\alpha }(G)`$ and $`p_\alpha ^{\alpha +1}=\pi _{A_\alpha }^{A_{\alpha +1}}|G_{\alpha +1}`$ for each $`\alpha <\tau `$. Properties 1-3 of the spectrum $`𝒮_G`$ are satisfied by construction. Since, $`G_0`$ is a metrizable $`AE(1)`$-compactum, it follows from the equivalence $`(11)(12)`$ of Theorem A, that $`G_0`$ is a metrizable torus. Finally, property 4 is a consequence of Claim 4 and Lemma 3.1. Proof is completed. ∎
Next we investigate the map $`\mathrm{exp}_G:L(G)G`$.
###### Lemma 3.3.
Let $`p:GT`$ be a continuous homomorphism of compact Abelian groups and $`\mathrm{ker}(p)`$ is a metrizable $`AE(1)`$-compactum. Then the diagram
$$\begin{array}{ccc}L(G)& \stackrel{\mathrm{exp}_G}{}& G\\ L\left(p\right)& & p& & \\ L(T)& \stackrel{\mathrm{exp}_T}{}& T\end{array}$$
is $`0`$-soft.
###### Proof.
By Lemma 3.1, we may without loss of generality assume that $`G`$ is the product $`T\times \mathrm{ker}(p)`$ and the homomorphism $`p`$ coincides with the projection $`\pi _T:T\times \mathrm{ker}(p)T`$. Now consider the following commutative diagram (compare it with Diagram 3):
in which $`\chi =L(p)\mathrm{}\mathrm{exp}_G`$ is the characteristic map of the previous diagram \[3, Definition 6.2.1\]. We therefore need to show the $`0`$-softness of $`\chi `$. A straightforward verification shows that in this situation $`L(G)`$ is isomorphic to the product $`L(T)\times L(\mathrm{ker}(p))`$ and the homomorphism $`\chi `$ coincides with the product $`\mathrm{id}\times \mathrm{exp}_{\mathrm{ker}(p)}`$. Observe also that with these identifications $`L(p)`$ becomes isomorphic to the projection $`L(T)\times L(\mathrm{ker}(p))L(T)`$. Since $`\mathrm{ker}(p)`$ is a metrizable $`AE(1)`$-compactum, the map $`\mathrm{exp}_{\mathrm{ker}(p)}`$ is an open surjection between completely metrizable separable spaces (equivalence $`(3)(4)(11)`$). By \[3, Corollary 6.1.27\], $`\mathrm{exp}_G`$ is $`0`$-soft. This implies $`0`$-softness of $`\chi `$. ∎
###### Proposition 3.4.
Let $`G`$ be a compact connected Abelian group. Then the following conditions are equivalent:
* $`G`$ is an $`AE(1)`$-compactum.
* The map $`\mathrm{exp}_G:L(G)G`$ is $`0`$-soft.
###### Proof.
Since the statement is known to be true in the metrizable case we assume that $`w(G)=\tau >\omega `$.
$`(a)(b)`$. Represent $`G`$ as the limit of a well-ordered inverse spectrum $`𝒮=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$ with the properties listed in Proposition 3.2(c). The space $`L(G)`$ is then the limit of the associated spectrum $`L(𝒮)=\{L(G_\alpha ),L\left(p_\alpha ^{\alpha +1}\right),\tau \}`$ and the map $`\mathrm{exp}_G:L(G)G`$ is the limit of the morphism
$$\{\mathrm{exp}_{G_\alpha }:L(G_\alpha )G_\alpha ,\tau \}:L(𝒮)𝒮.$$
According to Lemma 3.3 this morphism is $`0`$-soft. $`0`$-softness of the limit map $`\mathrm{exp}_G`$ follows now from \[3, Theorem 6.3.1\] (applied with $`n=0`$).
$`(b)(a)`$. The proof of this part follows the proof of \[3, Theorem 6.3.1\]. Only two things have to be taken into account. The first has already been mentioned: the space $`L(G)`$ is isomorphic to $`\text{}^{\mathrm{rank}(\widehat{G})}`$ and hence is an $`AE(0)`$-space in the sense of \[3, Definition 6.1.12\]. The second: $`G`$ being a compact group is also an $`AE(0)`$-compactum (Theorem C). Following the indicated proof and making slight adjustments (as in the proof of Proposition 3.2) we arrive to the following situation. There exists a well-ordered spectrum $`𝒮_G=\{G_\alpha ,p_\alpha ^{\alpha +1},\tau \}`$, satisfying properties 1-3 and 5 of Proposition 3.2(c) and a part of property 4 which guarantees that the fiber $`\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)`$ is metrizable. Also there exists a morphism (we keep notation of the first part of the proof)
$$\{\mathrm{exp}_{G_\alpha }:L(G_\alpha )G_\alpha ,\tau \}:L(𝒮)𝒮$$
between this spectra which is $`0`$-soft in the sense that characteristic maps of all the square diagrams
$$\begin{array}{ccc}L(G_{\alpha +1})& \stackrel{\mathrm{exp}_{G_{\alpha +1}}}{}& G_{\alpha +1}\\ L\left(p_\alpha ^{\alpha +1}\right)& & p_\alpha ^{\alpha +1}& & \\ L(G_\alpha )& \stackrel{\mathrm{exp}_{G_\alpha }}{}& G_\alpha \end{array}$$
are $`0`$-soft. In addition, it follows from the above remarks that typical projection $`L(p_\alpha ^{\alpha +1})`$ is isomorphic to the projection $`\mathrm{pr}_1:L(G_\alpha )\times \text{}^\omega L(G_\alpha )`$. Since the characteristic map of the above diagram is $`0`$-soft, it is surjective \[3, Lemma 6.1.13\]. This means that $`\left(p_\alpha ^{\alpha +1}\right)^1(\mathrm{exp}_{G_\alpha }(x))=\mathrm{exp}_{G_{\alpha +1}}\left(L^1(p_\alpha ^{\alpha +1})(x)\right)`$ for any point $`xL(G_\alpha )`$. Apply the latter to the point $`\mathrm{𝟎}L(G_\alpha )`$. We have $`\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)=\mathrm{exp}_{G_{\alpha +1}}\left(L^1(p_\alpha ^{\alpha +1})(\mathrm{𝟎})\right)`$. Since, as was noted, the map $`L(p_\alpha ^{\alpha +1})`$ is isomorphic to the projection $`\mathrm{pr}_1`$, it follows that $`L^1(p_\alpha ^{\alpha +1})(\mathrm{𝟎})`$ is isomorphic to $`\text{}^\omega `$. Consequently a metrizable compact Abelian group $`\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)`$ being a surjective image of $`\text{}^\omega `$ (through the map $`\mathrm{exp}_{G_{\alpha +1}}`$) is arcwise connected. Equivalence $`(1)(12)`$ applied to $`\mathrm{ker}\left(p_\alpha ^{\alpha +1}\right)`$ shows that the latter is a metrizable torus. This finishes verification of property 4 of Proposition 3.2(c). Since $`G_0`$ is obviously a metrizable torus, it follows (Proposition 3.2) that $`G`$ is an $`AE(1)`$-compactum. Proof is completed. ∎
Theorem E from the introduction is a consequence of Propositions 3.2 and 3.4.
## 4. Free Abelian Groups
Discussion of spectral analysis in Section 2 indicates that although the Spectral Theorem has originally been designed for topological category, its explicit categorical nature allows us to obtain corresponding counterparts in other categories (such as category of compact (Abelian) groups, Shape category, category of spaces admitting effective actions of compact groups etc. ) as well. Here we consider the category of Abelian groups.
Let $`\tau \omega `$ be a given cardinal number. Every Abelian group $`G`$ can be represented as the union $`\{G_\alpha :\alpha A\}`$ of its subgroups $`G_\alpha `$, $`\alpha A`$, of cardinality $`|G_\alpha |\tau `$, where the partially ordered and directed indexing set $`A`$ is $`\tau `$-complete (this basically means that $`A`$ is not only directed but also $`\tau `$-directed - contains elements majorating all elements of any subset of cardinality $`\tau `$; see for details). Basic example of such an indexing set $`A`$ is the one ($`\mathrm{exp}_\tau G`$) corresponding to the representation of $`G`$ as the union of all subgroups of cardinality $`\tau `$. Other $`\tau `$-complete indexing sets would be (at least in all the situations we deal with) cofinal and $`\tau `$-closed (contain supremums of subsets of cardinality $`\tau `$) subsets of $`\mathrm{exp}_\tau G`$.
Let us say that the above representation is $`\tau `$-smooth<sup>1</sup><sup>1</sup>1This terminology is suggested by . if, in addition, $`G_\beta =\{G_\alpha :\alpha B\}`$, whenever $`B`$ is a chain in $`A`$, $`|B|\tau `$ and $`\beta =supB`$. Observe that $`\tau `$-smooth decompositions (i.e., direct $`\tau `$-spectra) are in one to one correspondence with $`\tau `$-spectra of compact Abelian groups.
###### Theorem 4.1 ($`\tau `$-Spectral Theorem).
Let $`f:GL`$ be a homomorphism between Abelian groups. Suppose also that $`\{G_\alpha :\alpha A\}`$ and $`\{L_\alpha :\alpha A\}`$ are any two $`\tau `$-smooth decompositions of $`G`$ and $`L`$ with the same indexing set $`A`$. Then there exists a cofinal and $`\tau `$-closed subset $`B`$ of $`A`$ such that $`f(G_\alpha )L_\alpha `$ for each $`\alpha B`$.
###### Proof.
Apply the Spectral Theorem for $`\tau `$-spectra of compact Abelian groups (see \[3, Lemma 8.2.1\] where the non Abelian and non compact version is proved, consult also ), note that $`\tau `$-smooth decompositions are nothing else but $`\tau `$-spectra and apply duality. ∎
There is no difficulty in defining the concept corresponding to well-ordered spectra of compact Abelian groups. These are decompositions of Abelian groups into increasing well-ordered systems of subgroups which are called smooth , i.e., $`\{G_\alpha :\alpha <\tau \}`$ is a smooth decomposition of $`G`$ if $`G_\alpha G_{\alpha +1}`$, $`\alpha <\tau `$, and if $`G_\beta =\{G_\alpha :\alpha <\beta \}`$ for any limit ordinal $`\beta <\tau `$.
###### Theorem 4.2 (Eklof,).
Let $`\{G_\alpha :\alpha <\tau \}`$ be a smooth decomposition of an Abelian group $`G`$ such that $`G_0`$ is free and $`G_{\alpha +1}/G_\alpha `$ is free for every $`\alpha <\tau `$. Then $`G`$ is free.
The above statement in essence gives one half of characterization of free Abelian groups. This half corresponds to the implication $`(c)(a)`$ of Proposition 3.2. Since, as we have seen, the other implication $`(a)(c)`$ of Proposition 3.2 is also true, we obtain the following result.
###### Theorem 4.3.
An (uncountable) Abelian group $`G`$, $`|G|=\tau `$, is free if and only if it admits a smooth decomposition $`\{G_\alpha :\alpha <\tau \}`$ such that $`G_0`$ is countable and free and $`G_{\alpha +1}/G_\alpha `$ is countable and free for every $`\alpha <\tau `$.
From methodological point of view it is perhaps of some interest to note that the well known \[11, Theorem 4,p.45\] statement “a subgroup of a free Abelian group is free” has in fact a countable nature.
###### Corollary 4.4.
Statements (a) and (b) are equivalent:
* A subgroup of a countable free Abelian group is free.
* A subgroup of free Abelian group is free.
###### Proof.
Let $`L`$ be a subgroup of an Abelian group $`G`$. If $`G`$ is free it has a smooth decomposition $`\{G_\alpha :\alpha <\tau \}`$ with the properties indicated in Theorem 4.3. Then $`\{L_\alpha =LG_\alpha :\alpha <\tau \}`$ is a smooth decomposition of $`L`$. Since $`L_0G_0`$ and $`L_{\alpha +1}/L_\alpha G_{\alpha +1}/G_\alpha `$, $`\alpha <\tau `$, we conclude (by (a)) that $`\{L_\alpha :\alpha <\tau \}`$ satisfies conditions of Theorem 4.3 and hence $`L`$ is free. ∎
###### Corollary 4.5.
A continuous and homomorphic image of a torus is a torus.
|
no-problem/9908/nucl-th9908045.html
|
ar5iv
|
text
|
# The quark-photon vertex and meson electromagnetic form factors
## 1 DYSON–SCHWINGER EQUATIONS
The Dyson–Schwinger equations \[DSEs\] form a useful tool for nonperturbative QCD modeling of hadrons and their interactions. They have been successfully applied to calculate properties of light vector and pseudoscalar mesons , as described elsewhere in these proceedings . The dressed-quark propagator, as obtained from its DSE, together with the Bethe–Salpeter amplitude as obtained from the Bethe–Salpeter equation \[BSE\] for $`q\overline{q}`$ bound states, form essential ingredients for calculations of meson couplings and form factors . To describe electromagnetic interactions of hadrons, we also need the nonperturbatively dressed quark-photon vertex. Here, we use a solution of the DSE for the quark-photon vertex under the same truncation as used in Refs. for light mesons.
The DSE for the renormalized dressed-quark propagator in Euclidean space is
$`S(p)^1`$ $`=`$ $`Z_2i\gamma p+Z_4m(\mu )+Z_1{\displaystyle \frac{d^4q}{(2\pi )^4}g^2D_{\mu \nu }(pq)\frac{\lambda ^a}{2}\gamma _\mu S(q)\mathrm{\Gamma }_\nu ^a(q,p)},`$ (1)
where $`D_{\mu \nu }(k)`$ is the dressed-gluon propagator and $`\mathrm{\Gamma }_\nu ^a(q;p)`$ the dressed-quark-gluon vertex. The solution of Eq. (1) has the form $`S(p)^1=i\gamma pA(p^2)+B(p^2)`$ and is renormalized at spacelike $`\mu ^2`$ according to $`A(\mu ^2)=1`$ and $`B(\mu ^2)=m(\mu )`$ with $`m(\mu )`$ the current quark mass.
The DSE for the quark-photon vertex $`\stackrel{~}{\mathrm{\Gamma }}_\mu (p;Q)=\widehat{Q}\mathrm{\Gamma }_\mu (p;Q)`$ is the inhomogeneous BSE
$$\stackrel{~}{\mathrm{\Gamma }}_\mu (p;Q)=Z_2\widehat{Q}\gamma _\mu +^\mathrm{\Lambda }\frac{d^4q}{(2\pi )^4}K(p,q;Q)S(q+Q/2)\stackrel{~}{\mathrm{\Gamma }}_\mu (q;Q)S(qQ/2),$$
(2)
where $`\widehat{Q}`$ is the charge operator. The kernel $`K`$ is the renormalized, amputated $`\overline{q}q`$ scattering kernel that is irreducible with respect to a pair of $`\overline{q}q`$ lines. Solutions of the homogeneous version of Eq. (2) define the vector meson bound states at $`Q^2=m^2`$. It follows that $`\stackrel{~}{\mathrm{\Gamma }}_\mu (p;Q)`$ has poles at those locations.
We use a ladder truncation for the BSE in conjunction with a rainbow truncation $`\mathrm{\Gamma }_\nu ^a(q,p)\gamma _\nu \lambda ^a/2`$ for the quark DSE. Both the vector Ward–Takahashi identity \[WTI\] for the quark-photon vertex and the axial-vector WTI are preserved in this truncation. This ensures both current conservation and the existence of massless pseudoscalar mesons if chiral symmetry is broken dynamically: pions are Goldstone bosons .
The details of the model can be found in Refs. . It leads to chiral symmetry breaking and confinement; furthermore, at large momenta, our effective interaction reduces to the perturbative running coupling and thus preserves the one-loop renormalization group behavior of QCD and reproduces perturbative results in the ultraviolet region. The model gives a good description of the $`\pi `$, $`\rho `$, $`K`$, $`K^{}`$ and $`\varphi `$ masses and decay constants .
## 2 THE DRESSED QUARK-PHOTON VERTEX
The general form of the quark-photon vertex $`\mathrm{\Gamma }_\mu (q;Q)`$ can be decomposed into twelve independent Lorentz covariants. Four of these covariants, representing the longitudinal components, are uniquely determined by the vector WTI
$$iQ_\mu \mathrm{\Gamma }_\mu (p;Q)=S^1(p+Q/2)S^1(pQ/2).$$
(3)
The eight transverse components of $`\mathrm{\Gamma }_\mu (p;Q)`$ are not constrained by the WTI, except at $`Q=0`$, where the WTI reduces to $`i\mathrm{\Gamma }_\mu (p;0)=`$ $`S(p)^1/p_\mu `$.
It is obvious from Eq. (3) that the bare vertex $`\gamma _\mu `$ is a bad approximation if the quark self-energy is momentum dependent as is realistically the case. For QCD modeling of electromagnetic coupling to hadrons, it has been common practice to avoid a numerical study of the quark-photon vertex, and use the so-called Ball–Chiu \[BC\] Ansatz , which expresses the vertex in terms of the dressed-quark propagator functions $`A`$ and $`B`$
$$\mathrm{\Gamma }_\mu ^{BC}(p;Q)=\gamma _\mu \frac{A(p_+^2)+A(p_{}^2)}{2}+2(\gamma p)p_\mu \frac{A(p_+^2)A(p_{}^2)}{p_+^2p_{}^2}2ip_\mu \frac{B(p_+^2)B(p_{}^2)}{p_+^2p_{}^2},$$
(4)
where $`p_\pm =p\pm Q/2`$. This satisfies the WTI, Eq.(3), transforms correctly under CPT, and has the correct perturbative limit $`\gamma _\mu `$ in the extreme ultraviolet. The longitudinal components of $`\mathrm{\Gamma }_\mu ^{BC}`$ are exact, but the transverse components are correct only at $`Q=0`$. In particular, $`\mathrm{\Gamma }_\mu ^{BC}`$ does not have the vector meson poles. This should be of little concern for form factors at large spacelike $`Q^2`$; however, for $`Q^20`$ the situation is less clear.
Our numerical solution of Eq. (2) shows clearly the vector meson pole in all eight transverse amplitudes. The solution for the four longitudinal amplitudes agrees perfectly with the BC Ansatz, as required by the WTI. Our transverse solution agrees with the BC Ansatz only at spacelike asymptotic momenta. At low $`Q^2`$ it departs significantly from this Ansatz; although there is necessarily agreement at the point $`Q=0`$, the $`Q`$-dependence of the DSE solution is much larger than that of the BC Ansatz. Near the $`\rho `$ pole, the quark-photon vertex behaves like
$$\stackrel{~}{\mathrm{\Gamma }}_\mu (p;Q)\frac{\mathrm{\Gamma }_\mu ^\rho (p;Q)m_\rho ^2/g_\rho }{Q^2+m_\rho ^2},$$
(5)
where the $`\rho \gamma `$ coupling strength $`m_\rho ^2/g_\rho `$ associated with the $`\rho e^+e^{}`$ decay is well reproduced by the present model . There is no unique decomposition of the vertex into resonant and non-resonant terms away from the pole, but over a limited interval near $`Q^20`$ the difference between the DSE solution and the BC Ansatz can be approximated by Eq. (5) with $`m_\rho ^2Q^2`$ in the numerator, which one can call the tail of the $`\rho `$ resonance. However, the DSE solution for the vertex is the appropriate representation containing both the resonant and non-resonant parts of the vertex.
## 3 MESON FORM FACTORS
In the impulse approximation, both the pion charge form factor and the $`\gamma ^{}\pi \gamma `$ transition form factor are described by a triangle diagram, with one or two pion Bethe–Salpeter amplitudes, one or two quark-photon vertices, and three dressed-quark propagators. We obtain these elements from the appropriate BSE and DSE, within the same model. We compare the form factor results using: 1) a bare vertex; 2) the BC Ansatz; 3) the DSE solution for the quark-photon vertex. Note that the WTI ensures electromagnetic current conservation in both 2) and 3), but not in approximation 1), which violates the WTI.
The impulse approximation for the pion form factor gives
$$F_\pi (Q^2)P_\nu =N_c\frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[\mathrm{\Gamma }_\pi (k_+;P_+)S(q_+)i\mathrm{\Gamma }_\nu (q_+;Q)S(q_{++})\mathrm{\Gamma }_\pi (k_{};P_{})S(q_{})\right],$$
(6)
where $`Q`$ is the photon momentum, $`P_\pm =P\pm Q/2`$, $`q_\pm =q\pm P/2`$, $`q_{+\pm }=q_+\pm Q/2`$ and $`k_\pm =q\pm Q/4`$. It is evident that the $`Q^2`$ dependence of $`F_\pi `$ comes from both the quark substructure of the pion and the $`Q`$-dependence of the quark-photon vertex. Due to Eq. (5), $`F_\pi (Q^2)`$ will exhibit a resonance peak at timelike momenta $`Q^2`$ near $`m_\rho ^2`$.
A long-standing issue in hadronic physics is the question of the extent to which $`F_\pi (Q^2)`$ at low spacelike $`Q^2`$ can still be described by the $`\rho `$ resonance mechanism. This is an essential element of vector meson dominance (VMD), which leads to
$$F_\pi ^{VMD}(Q^2)=1\frac{g_{\rho \pi \pi }Q^2}{g_\rho (Q^2+m_\rho ^2)}.$$
(7)
The first term arises from the non-resonant photon coupling to a point pion; the only $`Q^2`$ dependence in VMD comes from the resonant mechanism with the produced $`\rho `$ having a point coupling to the pion. The pion charge radius $`r_\pi ^2=`$ $`6F_\pi ^{}(0)`$ thus becomes $`6g_{\rho \pi \pi }/(m_\rho ^2g_\rho )0.48\mathrm{fm}^2`$ which compares favorably with the experimental value $`0.44\mathrm{fm}^2`$.
In Fig. 1(a) we show our results from Eq. (6). Clearly a bare vertex is incorrect: current conservation, which ensures $`F_\pi (0)=1`$, is violated. Use of the BC Ansatz conserves the current; but the resulting $`r_\pi ^2=0.18\mathrm{fm}^2`$ is less than half the experimental value and the curve misses the data completely. The DSE solution for the vertex agrees very well with the data and produces $`r_\pi ^2=0.45\mathrm{fm}^2`$, without fine tuning the model parameters: the parameters are completely fixed in Refs. . This indicates that as much as half of $`r_\pi ^2`$ can be attributed to a reasonable extrapolation of the $`\rho `$ resonance mechanism. On the other hand, the strict VMD picture is too simple; at least 40% of $`r_\pi ^2`$ arises from the non-resonant photon coupling to the quark substructure of the pion.
The impulse approximation for the $`\gamma ^{}\pi \gamma `$ vertex with $`\gamma ^{}`$ momentum $`Q`$ is
$`\mathrm{\Lambda }_{\mu \nu }(P,Q)=i{\displaystyle \frac{\alpha }{\pi f_\pi }}ϵ_{\mu \nu \alpha \beta }P_\alpha Q_\beta g_{\pi \gamma \gamma }F(Q^2)`$
$`={\displaystyle \frac{N_c}{3}}{\displaystyle \frac{d^4k}{(2\pi )^4}\mathrm{Tr}\left[S(q^{})i\mathrm{\Gamma }_\nu (k^{};Q)S(q^{\prime \prime })i\mathrm{\Gamma }_\mu (k^{\prime \prime };PQ)S(q^{\prime \prime \prime })\mathrm{\Gamma }_\pi (k;P)\right]}.`$
where the momenta follow from momentum conservation. In the chiral limit, the value at $`Q^2=0`$, corresponding to the decay $`\pi ^0\gamma \gamma `$, is given by the axial anomaly and its value $`g_{\pi \gamma \gamma }^0=1/2`$ is a direct consequence of only gauge invariance and chiral symmetry; this value corresponds well with the experimental width of $`7.7\mathrm{eV}`$. In Fig. 1(b) we show our results, normalized to the experimental $`g_{\pi \gamma \gamma }`$. A bare vertex does not reproduce the anomaly since it violates WTIs. Both the BC Ansatz and the DSE vertex solution reproduce the anomaly value, but the BC Ansatz overestimates the form factor at small but nonzero spacelike momenta and gives an interaction radius $`r^2=0.13\mathrm{fm}^2`$, compared to the experimental value $`r^20.42\mathrm{fm}^2`$ . The vertex DSE solution gives results remarkably close to the data and $`r^2=0.40\mathrm{fm}^2`$, again indicating that the BC Ansatz underestimates the $`Q^2`$ dependence of the form factor, related to the absence of a $`\rho `$ resonance.
|
no-problem/9908/astro-ph9908077.html
|
ar5iv
|
text
|
# Markarian 421’s Unusual Satellite Galaxy
## 1 Introduction
Markarian 421 is a giant elliptical galaxy that contains one of the nearest BL Lac objects, at a redshift of 0.0308 (Ulrich et al. 1975). This object is among the most intensively studied of all active galactic nuclei (AGN). MKN 421 is a strong cm-wavelength radio source that has shown reported superluminal motion in its compact radio jet (Zhang and Baath 1990), although recent measurements (Piner et al. 1999) do not confirm this. It is optically highly variable in both intensity and polarization. It is seen in X-rays (Comastri et al. 1997), GeV gamma-rays (Mukherjee et al. 1997), and up to multi-TeV energies (Krennrich et al. 1997). Episodes of rapid variability have been seen repeatedly at many wavelengths (Tosti et al. 1998; Gaidos et al. 1996) strengthening the evidence for the presence of a compact object as the source of the nuclear activity.
The host galaxy of MKN 421 has been the subject of several spectroscopic and photometric studies. The first such study (Ulrich 1975) established the redshift based on weak stellar absorption lines, and also noted that a nearby galaxy 14 arcsec to the ENE had a similar redshift (z=0.0316), indicating that it was probably physically related, although if the velocity difference were due to the Hubble flow, the distance could be a few Mpc or more. The companion galaxy was classified as a normal elliptical (Hickson et al. 1984). Further work by Ulrich (1978) showed that MKN 421 was the brightest member of a group of 5–7 galaxies of similar redshift spread over a region of sky of order 10 arcmin in radius. The presence of this group increases the likelihood that the companion’s proximity to MKN 421 is physical rather than a random alignment.
There is mounting evidence that AGN phenomena appear to be associated with galaxy mergers or encounters (cf. Shlosman, Begelman, and Frank 1990; Hernquist & Mihos 1995). In the case of BL Lac objects, a significant number have been found in the last decade or so to be associated with close companions or groups of nearby galaxies (cf. Falomo 1996; Heidt 1999), although MKN 421 has apparently been overlooked in this regard. Intrigued by the proximity of these two galaxies, we have analyzed archival HST imagery of the system, and performed Hale 5 m long-slit spectroscopic observations aimed at clarifying this association.
Our goal was to understand the nature of the nearby galaxy in relation to MKN 421, and to investigate the properties of the companion galaxy itself. If the galaxy is as close to MKN 421 as its projected distance ($`10`$ Kpc) suggests, it is deep within the gravitational potential well of MKN 421, and is probably sweeping through its stellar halo. The conditions under which such an encounter can take place are of general interest in the understanding of galaxy evolution. Our results will show that the companion galaxy contains a Seyfert-like nucleus, and is likely to be tidally interacting with MKN 421. Although the evidence is circumstantial, this association does appear to lend weight to the suggestions that galaxy encounters are an important factor in AGN evolution, and that close companions are associated in some way with the BL Lac phenomenon.
In the following section we summarize the the imagery and photometry which indicate nuclear activity in the companion galaxy. Section 3 summarizes the spectroscopic results, from which we derive a velocity profile which indicates a system that is tidally bound. Section 4 presents some further analysis of the spectroscopic results, including a derived lower limit of the bulge mass and mass–to–light ratio of MKN 421 under some plausible assumptions. Section 5 summarizes and concludes the article.
## 2 Observations
A complete log of all observations presented here is shown in Table 1.
In the work of Ulrich (1978), the system $`14\mathrm{}`$ to the ENE of MKN 421 was denoted as galaxy number 5 of the group of galaxies associated with MKN 421. Although the galaxy had been noticed prior to this work, it appears that Ulrich was the first to provide a designation for it. Thus we will refer to it in this present work as MKN 421-5.
We note that users of NASA’s Extragalactic Database (NED) may find a reference to this object as RX J1104.4+3812:\[BEV98\] 014. However, since this reference is only used to distinguish objects within the ROSAT error circle for MKN 421, and the object is previously known, we prefer the designation given by Ulrich (1978).
### 2.1 HST Imagery and Photometry
#### 2.1.1 HST images
MKN 421 was observed with the HST wide field/planetary camera (WF/PC2) on 1997 March 5, using the F702W filter. Five exposures, one of duration 2 s, two of duration 30 s, and two of duration 120 s were made <sup>1</sup><sup>1</sup>1The HST observations used here are available as part of the Space Telescope Science Institute public archive, and were made originally as a result of a proposal by C. M. Urry.. The images were prepared by standard techniques described in Holtzman et al. (1995), and the moderate cosmic ray contamination of the frames was repaired by hand using linear interpolation. In the 2 s exposure, the MKN 421 nucleus is not saturated, but the companion galaxy is underexposed. The remaining two images overexpose the AGN, but for both MKN 421’s host galaxy and the companion galaxy the exposures provide good signal- to-noise ratios over the sky background. The pixel scale in these images is 0.0455 arcsec per pixel, and the resolution of HST at the mean filter wavelength (690 nm) is $`0.080`$ arcsec; thus the images are sampled just slightly under the Nyquist frequency.
Secondary corrections, including those associated with the known gradient in charge transfer efficiency and the pixel area variation across the frame (Holtzman et al. 1995), are not corrected for in the displayed images, but we have applied corrections for these effects in the relative photometry presented in the next section.
MKN 421 was again observed with WF/PC2 on 1999 May 24 with both the F555W and F814W filters in snapshot mode, with 300s exposures in each case.<sup>2</sup><sup>2</sup>2These HST observations are also available as part of the Space Telescope Science Institute public archive, and were made originally as a result of a proposal by R. Fanti. Since these were single frame integrations, the cosmic ray removal for presentation purposes is more difficult, and the images do not add significantly to the structure seen in the F702W image. Thus we do not present these as part of the imagery here, although we do include results of the photometry performed on these additional images in the following section.
Figure 1 shows a slightly smoothed grayscale of the summed F702W image, with a logarithmic stretch and quantized levels chosen to show the details of the host galaxy of MKN 421 and MKN 421-5. Several features of MKN 421-5 are evident even from this image. First, its structure is not a simple elliptical. A suggestion of spiral arms is evident, and possible evidence of barlike structure appears at the outer edges of the galaxy. Second, the nucleus of the companion is clearly brightened relative to the galactic bulge, as is shown in the inset frame. Third, there is no evidence for any obvious dust lanes or similar absorption features in either the outer or bulge regions of the galaxy.
Figures 23 illustrate these features. In Figure 2 we show more quantitative evidence for the presence of spiral structure, since the structure is difficult to reproduce adequately in paper copies of the images themselves. Fig. 2(a) is a contour plot centered on the companion galaxy showing structure which is suggestive of spiral arms. In Fig. 2(b) we show centroids of the surface brightness distributions in slices along the galaxy major axis (dashed line) in Fig. 2(a). The displacement of the brightness centroids clearly shows the presence of spiral structure. (The amplitude of the centroid displacement does not directly track the location of the arms, since it depends on the brightness of the arm structure relative to the surrounding disk.)
In Figure 3 we show a profile of relative surface photometry using the F702W filter along the major axis of the companion galaxy. We have averaged the surface brightness in opposite $`15^{}`$ sectors, and have performed one-dimensional fits to 3 components that are apparent in the image: an unresolved core; a bulge component, here represented by a Gaussian; and an exponential disk. We have not truncated the disk at an inner radius here; our goal is primarily to show that the data are not consistent with a single power-law or other one-parameter model for radial brightness, but are reasonably well-modeled by the standard parameters of a disk system.
#### 2.1.2 HST photometry
We estimated photometric parameters from the HST images using the prescription given by Holtzman et al. 1995, including all of the relevant corrections. We have also transformed the resulting magnitudes to Cousins VRI magnitudes using the transformations in Holtzman et al., and we include also the transformed colors although we note that the $`V`$ and $`I`$ measurements were not contemporaneous with the $`R`$ measurements. In Table 2 we summarize the results for the nuclear magnitudes of the two galaxies, and the annular bulge magnitude of MKN 421-5. We have also converted the measured magnitudes to absolute magnitudes using $`h=0.65`$, for use in later discussion. For MKN 421, we show only an estimate of the $`R`$ magnitude of the nucleus for comparison with MKN 421-5; in the 1999 May observations in the F555W and F814W filters there were no exposures short enough to avoid saturation of the nucleus.
Based on the expected contributions of the different components indicated by Fig. 3, The nuclear magnitude may contain a significant bulge contribution, of order 30-40%. In spite of this, it still appears that the bright nucleus is too compact to be a typical star forming region. As we discuss in a later section, it is likely to be either a compact nuclear star cluster or more probably an active nucleus. And although there is only a marginal difference in color between the nucleus and the annular bulge region, it does favor a rise of the nucleus relative to the bulge in the near–infrared, which is consistent with Seyfert–nucleus behavior.
The nucleus of MKN 421 itself is highly variable in the visual bands. In fact, the HST observations in early 1997 occurred shortly after a large optical outburst (Tosti et al. 1998) during which the R-band magnitude peaked at brighter than 12. During March 1997, Tosti et al. estimated $`R=12.4`$ from ground-based photometry. The HST observations, taken approximately 2 months later, show a decrease of $`0.34`$ mag.
## 3 Optical Spectroscopy
Low resolution optical spectra of MKN 421 and its companion were obtained with the Double Spectrograph on the 5m Palomar<sup>3</sup><sup>3</sup>3These observations at the Palomar Observatory were made as part of a continuing cooperative agreement between Cornell University and the California Institute of Technology. telescope during the night of 1999 February 19. The long slit (2′) with a 2″ aperture was centered on the companion galaxy and two 600 sec exposures were obtained. The slit was positioned at an angle of 53°, and passed through both the companion galaxy and the nucleus of MKN 421. A 5500 Å dichroic was used to split the light to the two sides (blue and red), providing nearly complete spectral coverage from 3600–7600 Å. The blue spectra were acquired with the 600 lines/mm diffraction grating (blazed at 4000 Å). The red spectra were acquired with the 316 lines/mm diffraction grating (blazed at 7500 Å). Thinned 1024$`\times `$1024 Tek CCDs, with read noises of 8.6 e<sup>-</sup> (blue) and 7.5 e<sup>-</sup> (red), were used on the two sides of the spectrograph. Both CCDs had a gain of 2. e<sup>-</sup>/(digital unit). The effective spectral resolution of the blue camera was 5.0 Å (1.72 Å/pix); the effective spectral resolution of the red camera was 7.9 Å (2.47 Å/pix). The spatial scale of the long slit was 0.62 ″/pix for the blue camera and 0.48 ″/pix for the red camera.
The spectra were reduced and analyzed with the IRAF<sup>4</sup><sup>4</sup>4IRAF is distributed by the National Optical Astronomy Observatories. package. The spectral reduction included bias subtraction, scattered light corrections, and flat fielding with both twilight and dome flats. The 2–dimensional images were rectified based on arc lamp observations (Fe and Ar for the blue side; He, Ne, and Ar for the red side) and the trace of stars at different positions along the slit. The sky background was removed from the 2-dimensional images by fitting a low order polynomial along each row of the spectra. One dimensional spectra of MKN 421 and its companion were extracted from the rectified images using a 1.5″ extraction region (the seeing disk at the time of the observations) centered on the peak emission of each system. In addition, a galaxy spectrum for MKN 421 was obtained by averaging together two 4″ regions offset from the nucleus by $`\pm `$4″. While the night was non–photometric, observations of standard stars from the list of Oke (1990) provided calibration for the fluxes which are estimated to be accurate to $`10`$%. The flux–calibrated spectra are presented in Figure 4, in units of erg cm<sup>-2</sup> s<sup>-1</sup> Å<sup>-1</sup>. We estimate an overall uncertainty of 20% in the flux calibration due to variations in the transparency during our observations.
### 3.1 Spectrum of MKN 421
As seen in Figure 4a, the nucleus of MKN 421 is dominated by nonthermal emission with very weak absorption features. This new spectrum is similar to other observations of the nucleus of MKN 421 (e.g., Marchã et al. 1996), with the exception that \[OI\], \[NII\], and H$`\alpha `$ emission lines have been detected. While measurement of relative fluxes for the narrow emission features is complicated by the presence of broad H$`\alpha `$ emission (Figure 5), and by the contamination of H$`\alpha `$ absorption from the underlying stellar population (see Figure 4b), the \[NII\] lines are significantly stronger than the narrow H$`\alpha `$ feature. Large \[NII\]/H$`\alpha `$ ratios are not uncommon in AGN, however (e.g., Veilleux & Osterbrock 1987). Both the broad and narrow emission lines appear to be associated only with the nucleus of MKN 421.
In Fig. 5 we plot a fitted continuum model for MKN 421 along with the measured spectrum, in the region near the detected emission lines. Although the continuum model does not account for all of the structure in the spectrum near the emission lines, it does qualitatively indicate the instrinsic width of the broad-line emission, which has a full-width-at-half-maximum (FWHM) of order 80 Å with detectable emission that extends out to nearly twice this value. The implied velocity dispersion is $`5000`$ km s<sup>-1</sup>. We discuss some of the implications of this in a later section.
Recently, Morganti et al. (1992) obtained narrow band images roughly centered on the H$`\alpha `$ and \[NII\] lines and reported a total H$`\alpha +`$\[NII\] flux of 1.6 $`\times `$ 10<sup>-14</sup> erg s<sup>-1</sup> cm<sup>-2</sup> for MKN 421. The present observations are the first spectroscopic detection of emission lines associated with the nucleus of MKN 421, confirming this result. The total integrated flux in the broad and narrow emission lines is 1.7 $`\times `$ 10<sup>-14</sup> erg s<sup>-1</sup> cm<sup>-2</sup> (EW $``$ 2.2 Å). We estimate an overall uncertainty in this value of $`20`$%.
This apparent agreement with Morganti et al. is somewhat misleading, since our estimate includes flux spread over more than 100 Å, while the Morganti et al. estimate was based on a 50 Å FWHM filter which was apparently offset with respect to the centroid of the emission. Thus we actually have detected a significantly lower total flux level than Morganti et al., although we cannot quantify the difference without a more accurate knowledge of the filters used by Morganti et al. Despite the new detection of emission lines, MKN 421 still falls under the general class of BL Lac objects: the 4000 Å break has a low contrast in the nuclear regions and the derived equivalent widths for the emission lines are significantly less than 5 Å.
The sudden appearance of emission features in previously featureless spectra of BL Lac–type objects is becoming quite common. Just a few years ago, such features were detected in the prototypical BL Lac, BL Lac itself, for the first time (Vermeulen et al. 1995; Corbett et al. 1996). Similar events were discovered in OJ 287 (Sitko & Junkkarinen 1985) and PKS 0521–365 (Ulrich 1981; Scarpa, Falomo, & Pian 1995) as well. The rise and decline of emission–line features in BL Lac–type objects poses an interesting puzzle, and monitoring of these lines should be done as often as possible to determine if their intensity is correlated to that of other bands.
### 3.2 Spectrum of the Host Galaxy
A spectrum of the underlying host galaxy was obtained by averaging spectra on either side of the nucleus (Figure 4b). As found in other spectroscopic observations of the host galaxy (e.g., Ulrich et al. 1975; Ulrich 1978), the spectrum is dominated by absorption features. In addition to the usual strong absorption features in the blue, the Na I $`\lambda \lambda `$5890,5896 doublet is remarkably strong throughout the galaxy. We note that the presence of an apparent broad emission bump in the region of the spectrum surrounding the H$`\alpha `$ absorption line is apparently due to scattered emission from the extremely bright active nucleus of MKN 421.
### 3.3 Spectrum of MKN 421-5
Similar to the host galaxy of MKN 421, the spectrum of the companion galaxy is dominated by absorption lines (Figure 4c). However, the presence of strong Balmer absorption features indicates that this system is not dominated only by an old stellar population; rather, it must have had recent star formation activity within the last Gyr or so.
The plethora of absorption lines in the optical spectrum permit an accurate redshift determination for this system: 9380 $`\pm `$ 50 km s<sup>-1</sup>. Ulrich (1978) was one of the first to point out that MKN 421 is part of a large group of galaxies; we now know that radio galaxies tend to form in groups (Zirbel 1997). The close proximity (spatially and in velocity space) of MKN 421 and MKN 421-5 suggests that these two systems may have had significant tidal interactions in the past and may explain the unusual stellar population of the companion.
### 3.4 Spatially resolved Na absorption
Both MKN 421 and its companion galaxy have high signal–to–noise Na I absorption features which can be used to trace their gas kinematics. A position–wavelength diagram for the two systems is shown in Figure 6. The Na line can be traced almost continuously from MKN 421 to MKN 421-5. The mean velocity as a function of position along the slit for several of the absorption lines is shown in Figure 7. The errors in each individual measurement are largely due to the relatively low signal–to–noise ratio and the somewhat coarse spectral resolution of the observations; nonetheless, the two systems are clearly offset in velocity, with a sense of velocity continuity between the two galaxies.
Fig. 7 also displays the results of unconstrained least-squares fits for the velocity gradients across each of the two systems. The results of these fits are shown in Table 3. The velocities are not corrected for earth orbital motion; however, the correction for the hour angle of our observation is negligible compared to the errors in the estimates in Table 3.
### 3.5 On limits to radio emission from MKN 421-5
A number of VLA studies of MKN 421 and its immediate vicinity were conducted soon after it was recognized to be a BL Lac object (Ulvestad et al. 1984; 1983); however, the resolution of these studies was not adequate to show the companion due in part to the high dynamic range required (MKN 421 is $`0.57`$ Jy at 20cm). More recent 20cm observations (Laurent-Muhleisen et al. 1993) with higher resolution show complex extended emission in the vicinity of MKN 421, but probably do not have enough dynamic range to detect the companion galaxy nucleus, if its brightness is comparable to the average of its class.
Although there are conspicuous examples of radio–loud Seyferts (primarily Seyfert 1) more typical Seyfert nuclei of both type 1 and 2 emit total fluxes of order $`10^{21}`$ W Hz<sup>-1</sup> at cm wavelengths (cf. Ulvestad and Wilson 1989 and references therein). At the distance of the companion, this luminosity would produce about 0.5 mJy of 3.5 cm flux density, and a factor of 2-3 more than this at 20 cm for a typical Seyfert spectral index of $`0.5`$ to $`0.7`$. Inspection of a number of VLA maps at various resolutions shows no source of significant strength at the companion galaxy position. However, the complex halo of emission around MKN 421 precludes identifying a source at a level below 1 mJy. Further radio observations should be made to attempt to identify any compact source associated with the companion galaxy nucleus.
## 4 Discussion
### 4.1 The nature of the companion
The absolute magnitude of the nucleus of MKN 421-5 was given in Table 2 as $`M_R=16.6`$. If we assume this magnitude contains a 40% background contribution from the galaxy of MKN 421-5, the nucleus still has $`M_R=16.0`$, and a corresponding $`M_V15.5`$. It is thus apparently as bright or brighter than any of the compact nuclear clusters described from HST observations by Phillips et al. (1996) and Carollo et al. (1997) which fell mostly in the range of $`M_V=12`$ to $`14`$. A recent detailed HST study of four nearby spirals with compact nuclei (Matthews et al. 1999) gave $`M_B`$ values that range from -8.5 to -10.4, and corresponding $`M_I`$ values of -9.9 to -11.4.
Compact star cluster nuclei are not uncommon among late-type spiral galaxies, although they are often difficult to detect in ground-based images. However, based on the photometric results presented above, the nucleus of MKN 421-5 must be among the most luminous in its class if it comprises an overluminous nuclear cluster.
The alternative is that we are viewing a Seyfert nucleus near the low end of the Seyfert luminosity range. In fact, recent studies have shown that the Seyfert luminosity function extends well below the luminosity of MKN 421-5 (Ho et al. 1996, Ho et al. 1997). It is also worth noting that Malkan et al. (1998), in an HST survey targeting known or potential Seyfert galaxies, found that the presence of a bright, unresolved nucleus in HST images was a nearly perfect indicator of Seyfert activity in a survey of several hundred objects. Thus the lack of emission lines from the nucleus of the companion galaxy is puzzling, given the other indications that the source is non-thermal in nature. However, given the mediocre seeing conditions during the spectroscopic observations, and the extreme compactness of the core, our constraints on emission lines are not yet very strong.
The HST results indicate that we are viewing the companion at a substantial inclination relative to its spiral axis. This could contribute to the lack of observed emission lines due to obscuration. If the galaxy is a Seyfert 2, the narrow line emission may in fact be sufficiently diluted by the galactic bulge that it will be difficult to observe from the ground. From the HST image, the nuclear source is probably unresolved; even if we assume its diameter to be 0.1”, the dilution factor under our seeing conditions (1.2”) is at least a factor of 10, after accounting for the brightness of the nucleus relative to its surrounding bulge. Clearly, interpretation of any ground-based observations of this and other similar objects must be tempered with caution because of the danger of these dilution effects.
### 4.2 Implications of the velocity measurements
#### 4.2.1 Mass of MKN 421 bulge
As noted above, the shape of the velocity curve is consistent with the behavior expected from tidal interaction. This, combined with the proximity of MKN 421-5 to MKN 421, and the presence of a group of galaxies of similar redshift, leads to the conclusion that MKN 421-5 is undergoing an orbital encounter with MKN 421. Also, since MKN 421-5 is much closer to MKN 421 than any other galaxy in the group, we may treat the pair as a binary system for purpose of estimating the dynamical masses involved.
The statistics of binary galaxies have been treated in detail by Noerdlinger (1975) who showed that the assumption of circular orbits for a binary pair is a conservative one with regard to mass estimation. In fact, the circular orbit approximation is valid up to eccentricities of $`ϵ0.4`$, and underestimates the mass by only a factor of 3 in the limit of a parabolic encounter ($`ϵ=1`$).
Thus we estimate a lower limit to MKN 421’s mass $`M`$ within the companion’s present projected orbital radius $`R_p`$:
$$M(R_p)\frac{U_r^2R_p}{G}$$
(1)
where $`U_r`$ is the companion’s radial velocity and $`G`$ is the gravitational constant.
For $`U_r=496`$ km s<sup>-1</sup> and $`R_p=10`$ Kpc in the comoving frame, the lower limit of the mass is
$$M(R_p)5.9\times 10^{11}M_{\mathrm{}}.$$
(2)
For moderate orbital eccentricity ($`ϵ0.4`$), Noerdlinger (1975) has shown that the most probable values of the total velocity $`U_T`$ and true separation $`R_T`$ are likely to be 20-30% higher than the observed values. Thus the most probable value for the bulge mass is about twice the value above, of order $`10^{12}M_{\mathrm{}}`$. Typical giant elliptical rotation curves increase out to radii of order 10 kpc, then flatten and continue flat out to 50 kpc or more. Since a flat rotation curve implies a linear increase of the dark halo mass with radius, we expect that only of order 10–20% of the total mass of the MKN 421 system is likely to reside within $`R_p`$, the total mass of the galaxy probably approaches $`10^{13}M_{\mathrm{}}`$, among the highest masses of giant elliptical galaxies. Further measurements of the virial velocities of other members of MKN 421’s group should be able to confirm this.
#### 4.2.2 Mass–to–light ratio
The CCD photometry of MKN 421 reported by Kikuchi et al. (1987) gives the absolute V magnitude of the MKN 421 host galaxy within a 13″ radius of M<sub>V</sub>(13″) = –21.51 $`\pm `$ 0.03, implying a luminosity of 3.5 $`\times `$ 10<sup>10</sup> L. Note that the emission from other bands, such as radio continuum and X–ray, is almost entirely associated with the AGN rather than the galaxy. Thus, the implied minimum mass–to–light ratio is M/L $``$ 17, indicating a substantial dark matter contribution to the total mass of MKN 421 within $`R_p`$.
This calculation ignores the possible contribution of atomic or molecular gas to the total mass, but no HI has been detected in this system (van Gorkom et al. 1989). IRAS measurements of far-infrared fluxes of MKN 421 (van Gorkom et al.) have been interpreted as due to the presence of a total dust mass of order $`5\times 10^7`$ M. If a gas component were present corresponding to this dust level, one might expect up to $`10^9`$ M of gas in the total galaxy; only a fraction of this would be within $`R_p`$. Thus we expect our derived M/L ratio to be robust.
A number of statistical estimates of the mass–to–light ratios of binary galaxies have yielded values with ranges of 12–32 for pure spiral pairs (Schweizer 1987; Honma 1999) and 22–60 for elliptical pairs (Schweizer 1987). These estimates use galaxies with mean separations of order 60–100 Kpc, and thus include a much greater mass fraction of the whole galaxy than our estimate. As discussed above, typical giant elliptical rotation curves lead one to expect that the total M/L for MKN 421 may be 5-10 times higher than what is measured for this close pair.
#### 4.2.3 Estimates of MKN 421 central black hole mass
Wandel (1999), Wandel et al. (1999), Laor (1998) and others have noted the correlation of AGN host galaxy bulge mass $`M_{bulge}`$ to the mass of a central black hole $`M_{BH}`$ in cases where a reliable virial mass or reverberation mapping mass could be estimated. Wandel (1999) finds a relation of $`M_{BH}3\times 10^4M_{bulge}`$ in a sample composed primarily of low-luminosity AGN for which the statistics are reasonably good. Magorrian et al. (1998), in a comprehensive study of kinematics of normal galaxies, found a relation $`M_{BH}6\times 10^3M_{bulge}`$. If we use these values to bound the likely value of $`M_{BH}`$ for MKN 421, and assuming that our estimate of the lower limit to the mass within $`R_p`$ is dominated by the bulge mass of MKN 421, we expect a central black hole mass in the range $`1.8\times 10^8M_{BH}3.6\times 10^9`$ M.
We can also use the parametric relations given by Laor (1998) and Wandel et al. (1999) to make an independent estimate of the size of the broad-line region (BLR) based on the AGN bolometric luminosity. Wandel et al. give a black hole mass estimate based on the size of the BLR and $`\mathrm{\Delta }v_{FWHM}`$, the FWHM of the measured broad-line velocity dispersion:
$$M_{BH}1.45\times 10^5M_{\mathrm{}}\left(\frac{c\tau _{BLR}}{1\mathrm{light}\mathrm{day}}\right)\left(\frac{\mathrm{\Delta }v_{FWHM}}{1000\mathrm{km}\mathrm{s}^1}\right)^2.$$
(3)
The size of the BLR can be estimated from (Kaspi et al. 1997; Wandel et al. 1999):
$$r_{BLR}=15L_{44}^{1/2}\mathrm{light}\mathrm{days}.$$
(4)
where $`L_{44}`$ is the bolometric liminosity For MKN 421, estimates of the bolometric luminosity over the $`0.1`$ to $`1\mu `$m band vary considerably, and it is unclear whether the quiescent or high-state luminosity is more appropriate. Using a bolometric luminosity derived from our HST magnitudes and the conversion given in Laor (1998), we find $`L_{44}40`$, which gives $`r_{BLR}=90`$ light days. Assuming that $`H\alpha `$ is dominant in forming the broad line emission, then we use $`\mathrm{\Delta }v_{FWHM}3300`$ km s<sup>-1</sup> from our spectrum above. The implied black hole mass is then $`M_{BH}1.4\times 10^8M_{\mathrm{}}`$ which is of of the same order of magnitude as the lower limit estimate from the bulge mass above.
A black hole mass of $`10^9M_{\mathrm{}}`$ is of order 0.25% of the total dynamical mass out to 10 Kpc. Thus it is not in itself enough to affect the large scale dynamics between the two galaxies, although it would certainly strongly influence the inner bulge regions. It is notable, however, that the center of mass of the system will likely be significantly displaced from the position of the black hole, perhaps by a few hundred pc or more, depending on the companion mass and the dark matter distribution. Careful measurements of the virial motion of the stars near the center of MKN 421 may be able to detect such an offset, which would provide independent confirmation of the proximity of MKN 421-5 to MKN 421.
#### 4.2.4 The dynamical state of the system
The smooth rise in the rotation curve of MKN 421 out to $`R_p`$ is consistent with expectations for a post-encounter system, in which any major velocity distortions produced by the encounter have relaxed and bulk tidal motions now dominate. However we can not rule out the possibility that this is the companion’s first approach toward MKN 421. It is also possible that there are distortions in the velocity curves that are unresolved at our present sensitivity. Since we cannot unambiguously infer the companion’s true velocity from radial velocity measurements alone, we cannot prove that the system is bound, and we can only base our estimates of the tidal effects on the global properties we observe. Despite these caveats, there are a number of different lines of evidence that point to a bound and tidally relaxed system.
An obvious, though not compelling, piece of evidence is the fact that we observe the companion galaxy this close to MKN 421 at all, at the relatively late epoch of the system. The probability that we are observing a first encounter between two high-ranking cluster galaxies cannot be estimated based on this single case alone. However, based on the galaxy count of Ulrich (1978) and on our own Palomar images over a 12 arcmin field around MKN 421, there are potentially 7 to 10 galaxies of comparable brightness to MKN 421-5, in a volume of order $`6\times 10^{16}`$pc<sup>3</sup>, of order $`10^4`$ times the volume containing MKN 421 and its companion. Since MKN 421 will probably tend to significantly perturb the local potential, the chance of near encounters with it are not determined simply by chance; however, we estimate that the chance observation of such an encounter is of order 1% or less for any snapshot of similar clusters.
If the system is not in the initial phase of a first encounter, it follows that (a) it is likely to be bound (due to the mass dominance of MKN 421 and the narrow range of phase space required for MKN 421-5 to have the required escape velocity); and (b) that it is tidally relaxed, since the relaxation time for the system as observed is in the range of 100 Myr, relatively short compared to the time of cluster formation.
The evidence for recent star-formation in the companion galaxy is also relevant to discussion of the dynamics of this pair. Recent simulations of hierarchical galaxy encounters (cf. Bekki 1999 & references therein) have shown that star forming activity in the satellite galaxy is a likely result of such an interaction. A number of recent HST studies have found a significant number of companions near QSO host galaxies (Bahcall et al. 1995; Disney et al. 1995) which appear to be undergoing some type of interaction with the QSO host. Canalizo and Stockton (1997) found that at least one companion galaxy in such a system (PG 1700+518) showed evidence of a relatively young stellar population.
What is perhaps surprising in the case of MKN 421 and MKN 421-5 is that, although simulations such as those of Bekki (1999) indicate that the companion galaxy is likely to evolve to a dwarf elliptical or irregular as a result of the encounter, MKN 421-5 appears to in fact be a relatively normal spiral, without any apparent major disruption or irregularity. This system thus presents a challenge to models which cannot account for preservation of such structure even in what was probably a strong tidal encounter between these two systems. It is notable that MKN 421-5, as a probable Seyfert, is likely to have a massive compact object in its nucleus. Based on arguments similar to the previous section and our estimates of the bulge luminosity of MKN 421-5, the expected mass is in the range of $`25\times 10^6M_{\mathrm{}}`$. The presence of this large central mass may in fact have important dynamical consequences in stabilizing a satellite galaxy in such an encounter.
#### 4.2.5 A schematic geometry for the system
To postulate a possible three-dimensional geometry for this pair of galaxies, we combine a number of related pieces of evidence that bear on the geometry, including our measured velocity curves, the apparent inclination of the companion spiral, the major and minor axes of MKN 421, and parameters associated with the central mass of MKN 421, although the latter are included mainly for comparison rather than as constraints on the global geometry.
The position angle of the center of the companion relative to MKN 421 is $`53^{}`$, and we estimate that the projected major axis of the spiral lies at a position angle of $`65^{}`$. Thus the projected rotation axis (perpendicular to the major axis of the spiral) of MKN 421-5 is at a position angle of $`335^{}`$. The inclination of the spiral is more difficult to estimate, since the shape may be confused by an inner bar feature not easily distinguished in the images. However, it appears that the spiral inclination is in the range of $`45^{}`$ to $`60^{}`$ to the line-of-sight.
If we assume that the disk of the spiral and the orbital plane are roughly coincident (this is reasonable if the systems are tidally relaxed), then we can make a first order schematic estimate of the overall geometry of the system, if the two galaxies are bound in an orbit of only moderate eccentricity. This schematic view of the system is depicted graphically in Figure 8. We have not attempted to quantify this geometry further; we present it simply as a qualitative baseline which can be refined (or revised) by future observations.
Recent VLBI imagery and estimates of the jet doppler factors from both VLBI and gamma-ray observations (Piner et al. 1999; Gaidos et al. 1996) indicate conclusively that the jet angle must be $`5^{}`$ with respect to the observation vector. The projected position angle of the inner jet from VLBI observations is $`320^{}`$. We have plotted this position angle in Fig. 8 for comparison. For the case of a small angle with respect to the line-of-sight, the approximate coincidence of the projected position angle with that of the spiral axis and the possible orbital plane is presumably an accident. However, if the jet is actually physically aligned with the other angular momentum vectors in the system, then a much larger angle for it with respect to the line-of-sight is favored.
An outstanding question in MKN 421 and in BL Lac objects in general is how well-aligned the jet axis is with the line-of-sight to the AGN. If the results based on estimates of the doppler factors are correct, and the jet axis is inclined by $`5^{}`$, then we are led to conclude that either the orbital plane of the two galaxies are not well-aligned with the perpendicular to the jet axis, or plane of the spiral companion is not well-aligned with the overall orbital plane. In either case the residual torques may have observable consequences, either in precession effects or tidal velocity distortions of the system. More complete velocity mapping may provide information on the latter consequence. As to the former effect, further analysis is necessary to estimate whether there might be observable effects, since the precession time scale would be expected to be long compared to the orbital time scale.
#### 4.2.6 Companion rotation: Prograde or Retrograde?
We conclude this section by briefly noting the sense of the companion rotation. The rotation curve shown in Fig. 7 shows that the eastern end of MKN 421-5 is redshifted relative to the western end which is closest to MKN 421. This indicates a rotation which is clockwise on the sky, and corresponds to a prograde rotation relative to the projected rotation of the bulge of MKN 421.
This feature of the system may be important to the stability of the companion. Future simulations should look in more detail at the effects of retrograde vs. prograde rotation in a tidal encounter.
## 5 Conclusions
We have gathered evidence that supports the following conclusions:
1. MKN 421’s nearest companion galaxy MKN 421-5 appears to be an early-type spiral, rather than elliptical. Spectral evidence suggests moderately recent star-formation activity in the companion, a feature which is known to accompany galaxy interactions.
2. MKN 421-5 contains a Seyfert-like nucleus, but without detectable emission lines in ground–based spectroscopy. We find that its luminosity is probably too high for a compact nuclear star cluster and conclude it is most likely to be an AGN, in spite of the lack of spectroscopic evidence.
3. We confirm the published radial velocity for MKN 421-5. We find that MKN 421-5 is very likely to be bound to MKN 421, and its orbital velocity is consistent with the trend in the bulk rotational velocity of MKN 421, suggesting tidal interaction.
4. We report the first spectroscopic observation of emission lines from MKN 421’s nucleus, notably H$`\alpha `$ and NII. Both broad and narrow components of H$`\alpha `$ are present, with the broad line emission showing a velocity dispersion of several thousand km s<sup>-1</sup>, typical of QSO emission lines. Spectroscopic monitoring of these lines should continue.
5. The observed orbital velocity of the companion MKN 421-5 provides a lower limit on the mass of MKN 421’s bulge of $`5.9\times 10^{11}`$ solar masses, with an estimated bulge mass–to–light ratio of $`17`$.
We thank R. Linfield and the staff of Palomar Observatory for their invaluable help with the observations, and Glenn Piner for his helpful comments on the manuscript. This work was performed at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under a cooperative agreement by Associated Universities Inc. This research is based in part on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS 5-26555.
|
no-problem/9908/hep-ex9908033.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The production of final state hadrons from primary hard partons, e.g. the quark and antiquark in $`e^+e^{}Z^0q\overline{q}`$, is currently believed to proceed in three stages. The first stage involves the radiation of gluons from the primary quark and antiquark, which in turn radiate gluons or split into $`q\overline{q}`$ pairs until their virtuality approaches the hadron mass scale. Such a “parton shower” is calculable in perturbative QCD, for example in the Modified Leading Logarithm Approximation (MLLA) .
The second stage, in which these partons turn into “primary” hadrons, is not understood quantitatively, although several hadronization models exist. A simple model is the ansatz of Local Parton-Hadron Duality (LPHD) , which hypothesizes that distributions of kinematic quantities for a given hadron species are directly proportional to the parton distributions at some appropriate parton virtuality. This allows the prediction via MLLA QCD of the shapes of differential cross sections for primary hadrons, and of, for example, the energy- and mass-dependences of the peak of the distribution of $`\xi =\mathrm{ln}(x_p)`$, where $`x_p=2p/E_{cm}`$, $`p`$ is the hadron momentum and $`E_{cm}`$ is the $`e^+e^{}`$ center-of-mass energy.
The third stage, in which unstable primary hadrons decay into final state hadrons, complicates the interpretation of inclusive measurements. It is desirable to remove the effects of these decays when comparing with the predictions of QCD$`+`$LPHD. Additional complications arise in jets initiated by heavy ($`c`$ or $`b`$) quarks in which the leading heavy hadrons carry a large fraction of the beam energy, restricting that available to other primary particles, and then decay into a number of secondary particles. It is thus also desirable to restrict measurements to events with light primary flavors.
A particularly interesting aspect of jet fragmentation is the question of what happens to the primary quark or antiquark that initiated the jet. Many fragmentation models assume that the initial quark is “contained” as a valence constituent of a particular hadron, and that this “leading” hadron has on average a higher momentum than the other particles in the jet. This phenomenon has not been studied precisely for high-energy light-flavor jets, since it is difficult to identify the sign and flavor of the initial $`q`$/$`\overline{q}`$ on a jet-by-jet basis. The quantification of leading particle effects could lead to ways to identify the primary flavor of arbitrary samples of jets, enabling a number of new measurements in $`e^+e^{}`$, as well as in $`e`$p and p$`\overline{\mathrm{p}}`$, collisions.
In this paper we present an update of our analysis of $`\pi ^\pm `$, $`K^\pm `$, and p/$`\overline{\mathrm{p}}`$ production in hadronic $`Z^0`$ decays collected by the SLC Large Detector (SLD). The analysis is based upon the full sample of 550,000 hadronic events obtained in runs of the SLAC Linear Collider (SLC) between 1993 and 1998. We measure differential cross sections in an inclusive sample of hadronic events of all flavors, and also in high-purity samples of light- ($`Z^0u\overline{u},d\overline{d},s\overline{s}`$) and $`b`$-flavor ($`Z^0b\overline{b}`$) events. From these three samples we extract corrected differential cross sections in light- and $`b`$-, as well as $`c`$-flavor ($`Z^0c\overline{c}`$) events. The unfolded differential cross sections for the light-flavor events are free from effects of heavy quark production and decay, and as such provide a more appropriate sample for comparison with QCD predictions, which generally assume massless quarks, although the influence of decay products of other unstable primary hadrons remains. We use these measurements to test the predictions of various fragmentation models.
We also select samples of quark and antiquark jets from our light-flavor event sample, using the large forward-backward production asymmetry in polar angle inherent in collisions of highly polarized electrons with positrons. The differential cross sections are measured separately for hadrons and antihadrons in light-quark jets, and the observed differences are interpreted in terms of leading particle effects. These measurements provide precise, unique tests of fragmentation models.
## 2 The SLD and Hadronic Event Selection
A general description of the SLD can be found elsewhere . The trigger and initial selection criteria for hadronic $`Z^0`$ decays are described in Ref. . This analysis used charged tracks measured in the Central Drift Chamber (CDC) and Vertex Detector (VXD) , and identified using the Cherenkov Ring Imaging Detector (CRID) . Momentum measurement is provided by a uniform axial magnetic field of 0.6T. The CDC and VXD give a momentum resolution of $`\sigma _p_{}/p_{}`$ = $`0.010.0026p_{}`$, where $`p_{}`$ is the track momentum transverse to the beam axis in GeV/$`c`$. One quarter of the data were taken with the original vertex detector (VXD2), and the rest with the upgraded detector (VXD3). In the plane normal to the beamline the centroid of the micron-sized SLC IP was reconstructed from tracks in sets of approximately thirty sequential hadronic $`Z^0`$ decays to a precision of $`\sigma _{IP}7`$ $`\mu `$m for the VXD2 data and $``$3 $`\mu `$m for the VXD3 data. Including the uncertainty on the IP position, the resolution on the charged track impact parameter ($`\delta `$) projected in the plane perpendicular to the beamline is $`\sigma _\delta =`$11$``$70/$`(p\mathrm{sin}^{3/2}\theta )`$ $`\mu `$m for VXD2 and $`\sigma _\delta =`$8$``$29/$`(p\mathrm{sin}^{3/2}\theta )`$ $`\mu `$m for VXD3, where $`\theta `$ is the track polar angle with respect to the beamline. The CRID comprises two radiator systems that allow the identification of charged pions with high efficiency and purity in the momentum range 0.3–35 GeV/c, charged kaons in the ranges 0.75–6 GeV/c and 9–35 GeV/c, and protons in the ranges 0.75–6 GeV/c and 10–46 GeV/c . The event thrust axis was calculated using energy clusters measured in the Liquid Argon Calorimeter .
A set of cuts was applied to the data to select well-measured tracks and events well contained within the detector acceptance. Charged tracks were required to have a distance of closest approach transverse to the beam axis within 5 cm, and within 10 cm along the axis from the measured IP, as well as $`|\mathrm{cos}\theta |<0.80`$, and $`p_{}>0.15`$ GeV/c. Events were required to have a minimum of seven such tracks, a thrust axis polar angle w.r.t. the beamline, $`\theta _T`$, within $`|\mathrm{cos}\theta _T|<0.71`$, and a charged visible energy $`E_{vis}`$ of at least 20 GeV, which was calculated from the selected tracks assigned the charged pion mass. The efficiency for selecting a well-contained $`Z^0q\overline{q}(g)`$ event was estimated to be above 96% independent of quark flavor. The VXD, CDC and CRID were required to be operational, resulting in a selected sample of roughly 303,000 events, with an estimated non-hadronic background contribution of $`0.10\pm 0.05\%`$ dominated by $`Z^0\tau ^+\tau ^{}`$ events.
Samples of events enriched in light and $`b`$ primary flavors were selected based on charged track impact parameters $`\delta `$ with respect to the IP in the plane transverse to the beam . For each event we define $`n_{sig}`$ as the number of tracks with impact parameter greater than three times its estimated error, $`\delta >3\sigma _\delta `$. Events with $`n_{sig}=0`$ were assigned to the light flavor sample and those with $`n_{sig}4`$ were assigned to the $`b`$ sample; the remaining events were classified as a $`c`$ sample. The light, $`c`$ and $`b`$ samples comprised 176,000, 88,000 and 38,000 events, respectively; selection efficiencies and sample purities were estimated from our Monte Carlo simulation and are listed in table 1.
Separate samples of hemispheres enriched in light-quark and light-antiquark jets were selected from the light-tagged event sample by exploiting the large electroweak forward-backward production asymmetry wrt the beam direction. The event thrust axis was used to approximate the initial $`q\overline{q}`$ axis and was signed such that its $`z`$-component was positive, $`\widehat{t}_z>0`$. Events in the central region of the detector, where the production asymmetry is small, were removed by the requirement $`|\widehat{t}_z|>0.2`$, leaving 125,000 events. The quark-tagged hemisphere in events with left-(right-)handed electron beam was defined to comprise the set of tracks with positive (negative) momentum projection along the signed thrust axis. The remaining tracks in each event were defined to be in the antiquark-tagged hemisphere. The sign and magnitude of the electron beam polarization were measured for every event. For the selected event sample, the average magnitude of the polarization was 0.73. Using this value and assuming Standard Model couplings at tree-level, the purity of the quark-tagged sample is 0.73.
For the purpose of estimating the efficiency and purity of the event flavor tagging and the particle identification, we made use of a detailed Monte Carlo (MC) simulation of the detector. The JETSET 7.4 event generator was used, with parameter values tuned to hadronic $`e^+e^{}`$ annihilation data , combined with a simulation of $`B`$-hadron decays tuned to $`\mathrm{{\rm Y}}(4S)`$ data and a simulation of the SLD based on GEANT 3.21 . Inclusive distributions of single-particle and event-topology observables in hadronic events were found to be well described by the simulation .
## 3 Measurement of the Charged Hadron Fractions
Charged tracks were identified as pions, kaons or protons, in the CRID using a likelihood technique . Information from the liquid (gas) radiator only was used for tracks with $`p<2.5`$ ($`p>7.5`$) GeV/c; in the overlap region, $`2.5<p<7.5`$ GeV/c, liquid and gas information was combined. Additional track selection cuts were applied to remove tracks that scattered through large angles before exiting the CRID and to ensure that the CRID performance was well-modelled by the simulation. Tracks were required to have at least 40 CDC hits, at least one of which was in the outermost superlayer, to extrapolate through an active region of the appropriate radiator(s), and to have at least 80 (100)% of their expected liquid (gas) ring contained within a sensitive region of the CRID TPCs. The latter requirement included rejection of tracks with $`p>2.5`$ GeV/c for which there was a saturated CRID hit (from passage of miminum-ionizing particles) within a 5 cm radius (twice the maximum ring radius) of the expected gas ring center. Tracks with $`p<7.5`$ GeV/c were required to have a saturated hit within 1 cm of the extrapolated track, and tracks with $`p>2.5`$ GeV/c were required to have either such a saturated hit or the presence of at least four hits consistent with a liquid ring. These cuts accepted 47, 28 and 49% of tracks within the barrel acceptance in the momentum ranges $`p<2.5`$, $`2.5<p<7.5`$ and $`p>7.5`$ GeV/c, respectively. For momenta below 2 GeV/c, only negatively charged tracks were used to reduce the background from protons produced in interactions with the detector material. For momenta below 2.5 GeV/c, only the VXD2 data were used (due to time constraints), and the results in this region are identical to our published results .
For tracks with $`p<2.5`$ ($`p>2.5`$) GeV/c, we define a particle to be identified as type $`j`$, where $`j=\pi ,K`$,p, if $`_j`$ exceeds both of the other log-likelihoods by at least 5 (3) units. Efficiencies for identifying selected particles of true type $`i`$ as type $`j`$ were determined where possible from the data, using tracks from tagged $`K_s^0`$, $`\tau ^\pm `$ and $`\mathrm{\Lambda }^0`$ decays, as described in . An example is shown in fig. 1. A detailed Monte Carlo (MC) simulation of the detector was then used to make small corrections to these measurements, and to derive the remaining efficiencies from those measured. These efficiencies are parametrized in terms of continuous functions in each of the three momentum ranges, and are shown in fig. 2, in which the pairs of lines represent our estimated efficiencies plus and minus their systematic uncertainties. For the diagonal entries, these uncertainties correspond to statistical errors on the parameters fitted from the data, and are completely positively correlated across each of the three momentum regions. For the off-diagonal terms, representing misidentification rates, a more conservative 25% relative error was assigned at all points to account for the limited experimental constraints on the momentum dependence. These errors are also strongly positively correlated among momenta. The diagonal elements peak near or above 0.9 and the pion coverage is continuous from 0.5 GeV/c up to approximately 35 GeV/c. There is a gap in the kaon-proton separation between 7 and 10 GeV/c due to limited resolution of the liquid system and the fact that both particles are below Cherenkov threshold in the gas system. The proton coverage extends to the beam momentum. Misidentification rates are typically less than 0.03, with peak values of up to 0.07.
In each momentum bin we measured the fractions of the selected tracks that were identified as $`\pi `$, $`K`$ and p. The observed fractions were related to the true production fractions by an efficiency matrix, composed of the values in fig. 2 for that bin. This matrix was inverted and used to unfold our observed identified particle rates. This analysis procedure does not require that the sum of the charged particle fractions be unity; instead the sum was used as a consistency check and was found to be within statistical errors of unity for all momenta. In some momentum regions we cannot distinguish two of the three species, so the procedure was reduced to a 2$`\times `$2 matrix analysis and we present only the fraction of the identified species, i.e. protons above 35 GeV/c and pions between 6 and 9.5 GeV/c.
Electrons and muons were not distinguished from pions in this analysis; this background was estimated from the simulation to be about 5% in the inclusive flavor sample, predominantly from $`c`$\- and $`b`$-flavor events. The flavor-inclusive fractions were corrected using the simulation for the lepton backgrounds, as well as for the effects of beam-related backgrounds, particles interacting in the detector material, and particles with large flight distance, such that the conventional definition of a final-state charged hadron is recovered, namely charged pions, kaons or protons that are either from the primary interaction or decay products of particles with lifetime less than 3$`\times 10^{10}`$s.
The measured charged particle fractions for inclusive hadronic $`Z^0`$ decays are shown in fig. 3. The errors on the points below 15 GeV/c are dominated by the systematic uncertainties on the identification efficiencies and are strongly positively correlated across the entire momentum range. For $`p>15`$ GeV/c the errors have roughly equal statistical and systematic contributions, and the systematic errors are positively correlated and increase in magnitude with momentum.
Pions are seen to dominate the charged hadron production at low momentum, and to decline steadily in fraction as momentum increases. The kaon fraction rises steadily to about one-third at high momentum. The proton fraction rises to a maximum of about one-tenth at about 10 GeV/c, then declines slowly. At high $`x_p`$, the pion and kaon fractions appear to be converging. This convergence could indicate reduced strangeness suppression at high momentum, or that production is becoming dominated by leading particles, such that kaons from $`s\overline{s}`$ events are as common as pions from $`u\overline{u}`$ and $`d\overline{d}`$ events.
Where the momentum coverage overlaps, these measured fractions were found to be in agreement with our previous results and with other measurements at the $`Z^0`$ . Measurements based on ring imaging and those based on ionization energy loss rates cover complementary momentum ranges and can be combined to provide continuous coverage over the range $`0.2<p<35`$ GeV/c.
In fig. 4 we compare our measured charged hadron fractions with the predictions of the JETSET 7.4 , UCLA and HERWIG 5.8 fragmentation models, using default parameters. The momentum dependence of each fraction is reproduced qualitatively by all three models. The HERWIG and UCLA predictions for the pion fraction are high at intermediate $`x_p`$; the three model predictions differ widely at very high $`x_p`$, but the statistics of the data are not sufficient to distinguish between them. All three predictions for the kaon fraction are too low (high) at small (large) $`x_p`$. The JETSET prediction for the proton fraction is too high at all $`x_p`$; those of HERWIG and UCLA show structure in the proton fraction at large $`x_p`$ that is inconsistent with the data.
## 4 Flavor-Dependent Analysis
The analysis was repeated separately on the high-purity light- and $`b`$-tagged event samples described in section 2, and on the remaining sample of events satisfying neither tag requirement, which we denote $`c`$-tagged. In each momentum bin the measured differential cross sections $`r_j^{meas}`$ of each hadron species for these three samples, $`j=`$light-tag, $`c`$-tag, $`b`$-tag, were unfolded by inverting the relations:
$$r_j^{meas}=\frac{\mathrm{\Sigma }_ib_{ij}ϵ_{ij}R_ir_i^{true}}{\mathrm{\Sigma }_iϵ_{ij}R_i}$$
(1)
to yield true differential cross sections $`r_i^{true}`$ in events of the three flavor types, $`i=`$1, 2, 3, corresponding to $`Z^0u\overline{u},d\overline{d},s\overline{s}`$, $`Z^0c\overline{c}`$ and $`Z^0b\overline{b}`$. Here, $`R_i`$ is the fraction of hadronic $`Z^0`$ decays of flavor type $`i`$, taken from , $`ϵ_{ij}`$ is the event tagging efficiency matrix, estimated from the simulation and listed in table 1, and $`b_{ij}`$ represents the momentum-dependent bias of tag $`j`$ toward selecting events of flavor $`i`$ that contain hadrons of the type in question. The diagonal bias values are within a few percent of unity, reflecting a small multiplicity dependence of the flavor tags. The off-diagonal bias values are larger, but these have little effect on the unfolded results.
In fig. 5 we compare our measured charged hadron fractions in light-flavor flavor events with the predictions of the three fragmentation models. Qualitatively there is little difference between these data and those for the inclusive sample (fig. 3), however these are more relevant for comparison with QCD predictions based on the assumption of massless primary quark production, as well as for determining parameters in fragmentation models. We observe the same general differences between the predictions of the three fragmentation models and the data as were seen above in the flavor-inclusive sample. This indicates that these deficiencies are in the fragmentation simulation and not simply in the modelling of heavy hadron production and decay.
In fig. 6 we show the ratios of production in $`b`$\- to light-flavor and $`c`$\- to light-flavor events for the three species. The systematic errors on the particle identification largely cancel in these ratios, and the resulting errors are predominantly statistical. There is greater production of charged pions in $`b`$-flavor events at low momentum, with an approximately constant ratio for $`0.02<x_p<0.07`$. The production charged kaons is approximately equal in the two samples at $`x_p=0.02`$, but the relative production in $`b`$-flavor events then increases with $`x_p`$, peaking at $`x_p0.07`$. There is approximately equal production of protons in $`b`$-flavor and light-flavor events below $`x_p=0.15`$. For $`x_p>0.1`$, production of all these particle species falls faster with increasing momentum in $`b`$-flavor events. These features are consistent with expectations based on the known properties of $`Z^0b\overline{b}`$ events, namely that a large fraction of the event energy is carried by the leading $`B`$\- and $`\overline{B}`$-hadrons, which decay into a large number of lighter particles. Also shown in fig. 6 are the predictions of the three fragmentation models, which reproduce these features qualitatively, although HERWIG overestimates the pion and kaon ratios by a large factor at low $`x_p`$.
There is higher kaon production in $`c`$-flavor events than in light-flavor events at $`x_p0.1`$, reflecting the tendency of $`c`$-jets to produce a fairly hard charmed hadron whose decay products include a kaon carrying a large fraction of its momentum. There are fewer additional charged pions produced in $`D`$ decays than in $`B`$ decays, so that pion production is only slightly higher in $`c`$-flavor events at very small $`x_p`$. The pion $`c`$:light ratio starts to cut off at a larger value, $`x_p0.3`$, than the corresponding $`b`$:light ratio, attributable to the lower average decay multiplicity and softer fragmentation function of $`D`$ hadrons, and the kaon and proton ratios are consistent with this cutoff point. Again, all three fragmentation models reproduce the data qualitatively, although HERWIG overestimates the pion ratio at small $`x_p`$, as it did in the $`b`$:light case, and underestimates the proton ratio is large $`x_p`$.
## 5 Leading Particle Effects
We extended these studies to look for differences between particle and antiparticle production in quark (rather than antiquark) jets, in order to address the question of whether e.g. a primary $`u`$-initiated jet contains more particles that contain a valence $`u`$-quark (e.g. $`\pi ^+`$, $`K^+`$, p) than particles that do not (e.g. $`\pi ^{}`$, $`K^{}`$, $`\overline{\mathrm{p}}`$). To this end we used the light quark- and antiquark-tagged hemispheres described in section 2.
We measured the production rates per light quark jet
$`R_h^q`$ $`=`$ $`{\displaystyle \frac{1}{2N_{evts}}}{\displaystyle \frac{d}{dx_p}}\left[N(qh)+N(\overline{q}\overline{h})\right],`$ (2)
$`R_{\overline{h}}^q`$ $`=`$ $`{\displaystyle \frac{1}{2N_{evts}}}{\displaystyle \frac{d}{dx_p}}\left[N(q\overline{h})+N(\overline{q}h)\right],`$ (3)
where: $`q`$ and $`\overline{q}`$ represent light-flavor quark and antiquark jets respectively; $`N_{evts}`$ is the total number of events in the sample; $`h`$ represents any of the identified hadrons $`\pi ^{}`$, $`K^{}`$, and p, and $`\overline{h}`$ indicates the corresponding antiparticle. Then, for example, $`N(qh)`$ is the number of hadrons of type $`h`$ in light quark jets.
The charged hadron fractions analysis was repeated separately on the positively and negatively charged tracks in each of the quark- and antiquark-tagged samples. Results for the positively charged tracks in the quark-tagged sample and the negatively-charged tracks in the antiquark-tagged sample were consistent, so these two samples were combined and labelled as positively charged hadrons from light quark jets, yielding measured values of $`R_{\pi ^+}^q`$, $`R_{K^+}^q`$, and $`R_\mathrm{p}^q`$ in the tagged samples. The same procedure applied to the remaining tracks yielded $`R_\pi ^{}^q`$, $`R_K^{}^q`$, and $`R_{\overline{\mathrm{p}}}^q`$.
It is essential to understand the contributions to these rates from heavy-flavor events, which are typically large in the momentum range we cover and show substantial differences between hadron and antihadron due to decay products of the heavy hadrons. This motivated our use of light-tagged events, and the residual heavy flavor contributions were estimated from the simulation to be typically 15% of the observed hadrons. This estimate was applied as a correction, yielding differential cross sections per light-quark-tagged jet. The effect of this correction on the results was negligible compared with the statistical errors.
For each hadron type, differential cross sections in light quark jets were then extracted by correcting for the light-tag bias and unfolding for the effective quark (vs. antiquark) purity. The purity was estimated from the simulation to be 0.72, which is slightly lower than the value of 0.73 noted in section 2, reflecting the cutoff in acceptance of the barrel CRID at $`|\mathrm{cos}\theta |=0.68`$.
The measured differential cross sections per light quark jet are shown in fig 7. The errors shown are are the sum in quadrature of statistical errors and those systematic errors arising from uncertainties in the heavy-flavor background correction and the effective quark purity; the statistical errors dominate this total. Systematic errors common to hadron and antihadron, such as those due to their identification efficiencies, are not included,
In all cases the hadron and antihadron differential cross sections are consistent at low $`x_p`$. For charged pions there are small differences at high $`x_p`$, and for the other particles there are substantial differences, all of which appear to increase with increasing $`x_p`$. It is convenient to show these data in the form of the difference between hadron and antihadron differential cross sections normalized by the sum:
$$D_h=\frac{R_h^qR_{\overline{h}}^q}{R_h^q+R_{\overline{h}}^q},$$
(4)
The common systematic errors cancel explicitly in this variable. Results are shown in fig 8, along with our previous similar results for the strange vector meson $`K^0`$ and the $`\mathrm{\Lambda }^0`$ hyperon. A value of zero corresponds to equal production of hadron and antihadron, and the data are consistent with zero at low $`x_p`$. A value of $`+1`$ (–1) corresponds to complete dominance of (anti)hadrons $`h`$.
The baryon results are most straightforward to interpret. Since baryons contain valence quarks and not antiquarks, the excess of baryons over antibaryons in light quarks jets provides clear evidence for the production of leading baryons at high scaled momentum. The data suggest that the effect increases with $`x_p`$.
The interpretation for the mesons is more complicated, since they contain one valence quark along with one antiquark. All down-type quarks are produced equally and with the same SM forward-backward asymmetry in $`Z^0`$ decays, so that if a leading neutral particle such as $`K^0`$ ($`d\overline{s}`$) were produced equally in $`d`$ and $`\overline{s}`$ jets then one would observe $`D_{\overline{K}^0}=0`$. In the case of charged mesons such as $`\pi ^{}`$ ($`d\overline{u}`$), the different production rates and forward-backward asymmetries of up- and down-type quarks cause a nonzero dilution of leading particle effects. At the $`Z^0`$, equal leading pion production in $`u`$\- and $`d`$-jets would lead to a dilution factor of 0.27.
Our measured $`D_\pi ^{}`$ are consistently above zero at high $`x_p`$, and consistently below 0.27$`D_\mathrm{p}`$, although statistically consistent at each point with both. This suggests that leading primary pions are produced, but indicates that nonleading production of pions must be relatively large. This could be due to a very soft leading pion momentum distribution and/or a large “background” contribution from decays of $`\rho ^0`$, $`K^{}`$, etc. Our measured $`D_K^{}`$ are well above both zero and 0.27$`D_\mathrm{p}`$ for $`x_p>0.2`$. This indicates both substantial production of leading $`K^\pm `$ mesons at high momentum, and a depletion of leading kaon production in $`u\overline{u}`$ and $`d\overline{d}`$ events relative to $`s\overline{s}`$ events.
Assuming these high-momentum kaons to be directly produced in the fragmentation process, this amounts to a direct observation of a suppression of $`s\overline{s}`$ production from the vacuum with respect to $`u\overline{u}`$ or $`d\overline{d}`$ production. Assuming all $`K^\pm `$ in the range $`x_p>0.5`$ to be leading, we calculate $`\gamma _s=0.26\pm 0.06`$, consistent with values derived from inclusive measurements of the relative production rates of strange and non-strange, pseudoscalar and vector mesons.
Also shown in fig. 8 are the predictions of the three Fragmentation models. All three are consistent with the meson data and with the $`\mathrm{\Lambda }^0`$ data. The JETSET model is also consistent with the proton data, however the other two models predict a saturated value of $`D_\mathrm{p}`$ for $`x_p>0.4`$ that is inconsistent with the data.
## 6 Summary and Conclusions
Using the SLD Cherenkov Ring Imaging Detector we have made preliminary measurements of charged pion, kaon and proton production over most of the momentum range in hadronic $`Z^0`$ decays. We find the predictions of the JETSET, UCLA and HERWIG fragmentation models to be in qualitative agreement with our data. These results are in agreement with those from previous experiments.
By isolating high-purity light- and $`b`$-flavor samples, we have measured hadron production in light-flavor events, as well as in $`c`$\- and $`b`$-flavor events. We find substantial differences in particle production between light- and heavy-flavor events, with the latter producing more mesons overall, but far fewer at high momentum. These qualitative features are expected given the hard fragmentation and high average decay multiplicity of heavy hadrons. The light-flavor sample is more suitable for testing predictions of QCD that assume massless quarks, as well as for testing fragmentation models. We find differences between fragmentation model predictions and our data similar to those found in the inclusive sample, indicating that the deficiencies lie in the simulation of fragmentation rather than in that of heavy hadron production and decay.
By isolating high-purity light-quark and light-antiquark samples, we have made the first comparison of hadron and antihadron production in light-quark jets in $`e^+e^{}`$ annihilation. We observed an excess of p over $`\overline{\mathrm{p}}`$, which appears to increase with momentum, and provides direct evidence for the “leading particle” hypothesis that high momentum protons are more likely to contain the primary quark. We also observed a large excess of high momentum $`K^{}`$ over $`K^+`$ indicating that a high momentum kaon is likely to contain a primary quark or antiquark from the $`Z^0`$ decay, and that leading kaons are produced predominantly in $`s\overline{s}`$ events rather than $`d\overline{d}`$ or $`u\overline{u}`$ events. We observe only a small excess of $`\pi ^{}`$ over $`\pi ^+`$ at high momentum, due in part to the cancellation of the signal from $`u\overline{u}`$ and $`d\overline{d}`$ events, but also suggesting a large nonleading pion fraction even in this momentum region.
## Acknowledgements
We thank the personnel of the SLAC accelerator department and the technical staffs of our collaborating institutions for their outstanding efforts on our behalf.
This work was supported by Department of Energy contracts: DE-FG02-91ER40676 (BU), DE-FG03-91ER40618 (UCSB), DE-FG03-92ER40689 (UCSC), DE-FG03-93ER40788 (CSU), DE-FG02-91ER40672 (Colorado), DE-FG02-91ER40677 (Illinois), DE-AC03-76SF00098 (LBL), DE-FG02-92ER40715 (Massachusetts), DE-FC02-94ER40818 (MIT), DE-FG03-96ER40969 (Oregon), DE-AC03-76SF00515 (SLAC), DE-FG05-91ER40627 (Tennessee), DE-FG02-95ER40896 (Wisconsin), DE-FG02-92ER40704 (Yale); National Science Foundation grants: PHY-91-13428 (UCSC), PHY-89-21320 (Columbia), PHY-92-04239 (Cincinnati), PHY-95-10439 (Rutgers), PHY-88-19316 (Vanderbilt), PHY-92-03212 (Washington); The UK Particle Physics and Astronomy Research Council (Brunel, Oxford and RAL); The Istituto Nazionale di Fisica Nucleare of Italy (Bologna, Ferrara, Frascati, Pisa, Padova, Perugia); The Japan-US Cooperative Research Project on High Energy Physics (Nagoya, Tohoku); The Korea Science and Engineering Foundation (Soongsil).
## <sup>∗∗</sup>List of Authors
Kenji Abe,<sup>(21)</sup> Koya Abe,<sup>(33)</sup> T. Abe,<sup>(29)</sup> I.Adam,<sup>(29)</sup> T. Akagi,<sup>(29)</sup> N. J. Allen,<sup>(5)</sup> W.W. Ash,<sup>(29)</sup> D. Aston,<sup>(29)</sup> K.G. Baird,<sup>(17)</sup> C. Baltay,<sup>(40)</sup> H.R. Band,<sup>(39)</sup> M.B. Barakat,<sup>(16)</sup> O. Bardon,<sup>(19)</sup> T.L. Barklow,<sup>(29)</sup> G. L. Bashindzhagyan,<sup>(20)</sup> J.M. Bauer,<sup>(18)</sup> G. Bellodi,<sup>(23)</sup> R. Ben-David,<sup>(40)</sup> A.C. Benvenuti,<sup>(3)</sup> G.M. Bilei,<sup>(25)</sup> D. Bisello,<sup>(24)</sup> G. Blaylock,<sup>(17)</sup> J.R. Bogart,<sup>(29)</sup> G.R. Bower,<sup>(29)</sup> J. E. Brau,<sup>(22)</sup> M. Breidenbach,<sup>(29)</sup> W.M. Bugg,<sup>(32)</sup> D. Burke,<sup>(29)</sup> T.H. Burnett,<sup>(38)</sup> P.N. Burrows,<sup>(23)</sup> A. Calcaterra,<sup>(12)</sup> D. Calloway,<sup>(29)</sup> B. Camanzi,<sup>(11)</sup> M. Carpinelli,<sup>(26)</sup> R. Cassell,<sup>(29)</sup> R. Castaldi,<sup>(26)</sup> A. Castro,<sup>(24)</sup> M. Cavalli-Sforza,<sup>(35)</sup> A. Chou,<sup>(29)</sup> E. Church,<sup>(38)</sup> H.O. Cohn,<sup>(32)</sup> J.A. Coller,<sup>(6)</sup> M.R. Convery,<sup>(29)</sup> V. Cook,<sup>(38)</sup> R. Cotton,<sup>(5)</sup> R.F. Cowan,<sup>(19)</sup> D.G. Coyne,<sup>(35)</sup> G. Crawford,<sup>(29)</sup> C.J.S. Damerell,<sup>(27)</sup> M. N. Danielson,<sup>(8)</sup> M. Daoudi,<sup>(29)</sup> N. de Groot,<sup>(4)</sup> R. Dell’Orso,<sup>(25)</sup> P.J. Dervan,<sup>(5)</sup> R. de Sangro,<sup>(12)</sup> M. Dima,<sup>(10)</sup> A. D’Oliveira,<sup>(7)</sup> D.N. Dong,<sup>(19)</sup> M. Doser,<sup>(29)</sup> R. Dubois,<sup>(29)</sup> B.I. Eisenstein,<sup>(13)</sup> V. Eschenburg,<sup>(18)</sup> E. Etzion,<sup>(39)</sup> S. Fahey,<sup>(8)</sup> D. Falciai,<sup>(12)</sup> C. Fan,<sup>(8)</sup> J.P. Fernandez,<sup>(35)</sup> M.J. Fero,<sup>(19)</sup> K.Flood,<sup>(17)</sup> R. Frey,<sup>(22)</sup> J. Gifford,<sup>(36)</sup> T. Gillman,<sup>(27)</sup> G. Gladding,<sup>(13)</sup> S. Gonzalez,<sup>(19)</sup> E. R. Goodman,<sup>(8)</sup> E.L. Hart,<sup>(32)</sup> J.L. Harton,<sup>(10)</sup> A. Hasan,<sup>(5)</sup> K. Hasuko,<sup>(33)</sup> S. J. Hedges,<sup>(6)</sup> S.S. Hertzbach,<sup>(17)</sup> M.D. Hildreth,<sup>(29)</sup> J. Huber,<sup>(22)</sup> M.E. Huffer,<sup>(29)</sup> E.W. Hughes,<sup>(29)</sup> X.Huynh,<sup>(29)</sup> H. Hwang,<sup>(22)</sup> M. Iwasaki,<sup>(22)</sup> D. J. Jackson,<sup>(27)</sup> P. Jacques,<sup>(28)</sup> J.A. Jaros,<sup>(29)</sup> Z.Y. Jiang,<sup>(29)</sup> A.S. Johnson,<sup>(29)</sup> J.R. Johnson,<sup>(39)</sup> R.A. Johnson,<sup>(7)</sup> T. Junk,<sup>(29)</sup> R. Kajikawa,<sup>(21)</sup> M. Kalelkar,<sup>(28)</sup> Y. Kamyshkov,<sup>(32)</sup> H.J. Kang,<sup>(28)</sup> I. Karliner,<sup>(13)</sup> H. Kawahara,<sup>(29)</sup> Y. D. Kim,<sup>(30)</sup> M.E. King,<sup>(29)</sup> R. King,<sup>(29)</sup> R.R. Kofler,<sup>(17)</sup> N.M. Krishna,<sup>(8)</sup> R.S. Kroeger,<sup>(18)</sup> M. Langston,<sup>(22)</sup> A. Lath,<sup>(19)</sup> D.W.G. Leith,<sup>(29)</sup> V. Lia,<sup>(19)</sup> C.Lin,<sup>(17)</sup> M.X. Liu,<sup>(40)</sup> X. Liu,<sup>(35)</sup> M. Loreti,<sup>(24)</sup> A. Lu,<sup>(34)</sup> H.L. Lynch,<sup>(29)</sup> J. Ma,<sup>(38)</sup> G. Mancinelli,<sup>(28)</sup> S. Manly,<sup>(40)</sup> G. Mantovani,<sup>(25)</sup> T.W. Markiewicz,<sup>(29)</sup> T. Maruyama,<sup>(29)</sup> H. Masuda,<sup>(29)</sup> E. Mazzucato,<sup>(11)</sup> A.K. McKemey,<sup>(5)</sup> B.T. Meadows,<sup>(7)</sup> G. Menegatti,<sup>(11)</sup> R. Messner,<sup>(29)</sup> P.M. Mockett,<sup>(38)</sup> K.C. Moffeit,<sup>(29)</sup> T.B. Moore,<sup>(40)</sup> M.Morii,<sup>(29)</sup> D. Muller,<sup>(29)</sup> V.Murzin,<sup>(20)</sup> T. Nagamine,<sup>(33)</sup> S. Narita,<sup>(33)</sup> U. Nauenberg,<sup>(8)</sup> H. Neal,<sup>(29)</sup> M. Nussbaum,<sup>(7)</sup> N.Oishi,<sup>(21)</sup> D. Onoprienko,<sup>(32)</sup> L.S. Osborne,<sup>(19)</sup> R.S. Panvini,<sup>(37)</sup> C. H. Park,<sup>(31)</sup> T.J. Pavel,<sup>(29)</sup> I. Peruzzi,<sup>(12)</sup> M. Piccolo,<sup>(12)</sup> L. Piemontese,<sup>(11)</sup> K.T. Pitts,<sup>(22)</sup> R.J. Plano,<sup>(28)</sup> R. Prepost,<sup>(39)</sup> C.Y. Prescott,<sup>(29)</sup> G.D. Punkar,<sup>(29)</sup> J. Quigley,<sup>(19)</sup> B.N. Ratcliff,<sup>(29)</sup> T.W. Reeves,<sup>(37)</sup> J. Reidy,<sup>(18)</sup> P.L. Reinertsen,<sup>(35)</sup> P.E. Rensing,<sup>(29)</sup> L.S. Rochester,<sup>(29)</sup> P.C. Rowson,<sup>(9)</sup> J.J. Russell,<sup>(29)</sup> O.H. Saxton,<sup>(29)</sup> T. Schalk,<sup>(35)</sup> R.H. Schindler,<sup>(29)</sup> B.A. Schumm,<sup>(35)</sup> J. Schwiening,<sup>(29)</sup> S. Sen,<sup>(40)</sup> V.V. Serbo,<sup>(29)</sup> M.H. Shaevitz,<sup>(9)</sup> J.T. Shank,<sup>(6)</sup> G. Shapiro,<sup>(15)</sup> D.J. Sherden,<sup>(29)</sup> K. D. Shmakov,<sup>(32)</sup> C. Simopoulos,<sup>(29)</sup> N.B. Sinev,<sup>(22)</sup> S.R. Smith,<sup>(29)</sup> M. B. Smy,<sup>(10)</sup> J.A. Snyder,<sup>(40)</sup> H. Staengle,<sup>(10)</sup> A. Stahl,<sup>(29)</sup> P. Stamer,<sup>(28)</sup> H. Steiner,<sup>(15)</sup> R. Steiner,<sup>(1)</sup> M.G. Strauss,<sup>(17)</sup> D. Su,<sup>(29)</sup> F. Suekane,<sup>(33)</sup> A. Sugiyama,<sup>(21)</sup> S. Suzuki,<sup>(21)</sup> M. Swartz,<sup>(14)</sup> A. Szumilo,<sup>(38)</sup> T. Takahashi,<sup>(29)</sup> F.E. Taylor,<sup>(19)</sup> J. Thom,<sup>(29)</sup> E. Torrence,<sup>(19)</sup> N. K. Toumbas,<sup>(29)</sup> T. Usher,<sup>(29)</sup> C. Vannini,<sup>(26)</sup> J. Va’vra,<sup>(29)</sup> E. Vella,<sup>(29)</sup> J.P. Venuti,<sup>(37)</sup> R. Verdier,<sup>(19)</sup> P.G. Verdini,<sup>(26)</sup> D. L. Wagner,<sup>(8)</sup> S.R. Wagner,<sup>(29)</sup> A.P. Waite,<sup>(29)</sup> S. Walston,<sup>(22)</sup> J.Wang,<sup>(29)</sup> S.J. Watts,<sup>(5)</sup> A.W. Weidemann,<sup>(32)</sup> E. R. Weiss,<sup>(38)</sup> J.S. Whitaker,<sup>(6)</sup> S.L. White,<sup>(32)</sup> F.J. Wickens,<sup>(27)</sup> B. Williams,<sup>(8)</sup> D.C. Williams,<sup>(19)</sup> S.H. Williams,<sup>(29)</sup> S. Willocq,<sup>(17)</sup> R.J. Wilson,<sup>(10)</sup> W.J. Wisniewski,<sup>(29)</sup> J. L. Wittlin,<sup>(17)</sup> M. Woods,<sup>(29)</sup> G.B. Word,<sup>(37)</sup> T.R. Wright,<sup>(39)</sup> J. Wyss,<sup>(24)</sup> R.K. Yamamoto,<sup>(19)</sup> J.M. Yamartino,<sup>(19)</sup> X. Yang,<sup>(22)</sup> J. Yashima,<sup>(33)</sup> S.J. Yellin,<sup>(34)</sup> C.C. Young,<sup>(29)</sup> H. Yuta,<sup>(2)</sup> G. Zapalac,<sup>(39)</sup> R.W. Zdarko,<sup>(29)</sup> J. Zhou.<sup>(22)</sup>
<sup>(1)</sup>Adelphi University, Garden City, New York 11530, <sup>(2)</sup>Aomori University, Aomori , 030 Japan, <sup>(3)</sup>INFN Sezione di Bologna, I-40126, Bologna Italy, <sup>(4)</sup>University of Bristol, Bristol, U.K., <sup>(5)</sup>Brunel University, Uxbridge, Middlesex, UB8 3PH United Kingdom, <sup>(6)</sup>Boston University, Boston, Massachusetts 02215, <sup>(7)</sup>University of Cincinnati, Cincinnati, Ohio 45221, <sup>(8)</sup>University of Colorado, Boulder, Colorado 80309, <sup>(9)</sup>Columbia University, New York, New York 10533, <sup>(10)</sup>Colorado State University, Ft. Collins, Colorado 80523, <sup>(11)</sup>INFN Sezione di Ferrara and Universita di Ferrara, I-44100 Ferrara, Italy, <sup>(12)</sup>INFN Lab. Nazionali di Frascati, I-00044 Frascati, Italy, <sup>(13)</sup>University of Illinois, Urbana, Illinois 61801, <sup>(14)</sup>Johns Hopkins University, Baltimore, MD 21218-2686, <sup>(15)</sup>Lawrence Berkeley Laboratory, University of California, Berkeley, California 94720, <sup>(16)</sup>Louisiana Technical University - Ruston,LA 71272, <sup>(17)</sup>University of Massachusetts, Amherst, Massachusetts 01003, <sup>(18)</sup>University of Mississippi, University, Mississippi 38677, <sup>(19)</sup>Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, <sup>(20)</sup>Institute of Nuclear Physics, Moscow State University, 119899, Moscow Russia, <sup>(21)</sup>Nagoya University, Chikusa-ku, Nagoya 464 Japan, <sup>(22)</sup>University of Oregon, Eugene, Oregon 97403, <sup>(23)</sup>Oxford University, Oxford, OX1 3RH, United Kingdom, <sup>(24)</sup>INFN Sezione di Padova and Universita di Padova I-35100, Padova, Italy, <sup>(25)</sup>INFN Sezione di Perugia and Universita di Perugia, I-06100 Perugia, Italy, <sup>(26)</sup>INFN Sezione di Pisa and Universita di Pisa, I-56010 Pisa, Italy, <sup>(27)</sup>Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX United Kingdom, <sup>(28)</sup>Rutgers University, Piscataway, New Jersey 08855, <sup>(29)</sup>Stanford Linear Accelerator Center, Stanford University, Stanford, California 94309, <sup>(30)</sup>Sogang University, Seoul, Korea, <sup>(31)</sup>Soongsil University, Seoul, Korea 156-743, <sup>(32)</sup>University of Tennessee, Knoxville, Tennessee 37996, <sup>(33)</sup>Tohoku University, Sendai 980, Japan, <sup>(34)</sup>University of California at Santa Barbara, Santa Barbara, California 93106, <sup>(35)</sup>University of California at Santa Cruz, Santa Cruz, California 95064, <sup>(36)</sup>University of Victoria, Victoria, B.C., Canada, V8W 3P6, <sup>(37)</sup>Vanderbilt University, Nashville,Tennessee 37235, <sup>(38)</sup>University of Washington, Seattle, Washington 98105, <sup>(39)</sup>University of Wisconsin, Madison,Wisconsin 53706, <sup>(40)</sup>Yale University, New Haven, Connecticut 06511.
|
no-problem/9908/astro-ph9908047.html
|
ar5iv
|
text
|
# Domain wall dominated universes
## Abstract
We consider a cosmogony with a dark matter component consisting of a network of frustrated domain walls. Such a network provides a solid dark matter component with $`p=(2/3)\rho `$ that remains unclustered on small scales and with $`\mathrm{\Omega }_{\mathrm{dw}}0.7`$ can reconcile a spatially flat universe with the many observations indicating $`\mathrm{\Omega }_\mathrm{m}0.3.`$ Because of its large negative pressure, this component can explain the recent observations indicating an accelerating universe without recourse to a non-vanishing cosmological constant. We explore the viability of this proposal and prospects for distinguishing it from other kinds of proposed dark matter with significant negative pressure.
Recent observations of apparent luminosities of Type Ia supernovae (SNIa) at moderate redshift $`(z0.6)`$ suggest that the expansion of the universe is now accelerating, indicating a form of matter with a significant negative pressure . From a theoretical standpoint, these observations are remarkable because most conceivable contributions to the stress-energy entail a positive (radiation) or vanishing (matter) pressure, rather than a large negative pressure.
Two proposals to explain these observations are a non-vanishing cosmological constant $`(\mathrm{\Lambda })`$ or a very slowly rolling scalar field, often dubbed quintessence. Both proposals, however, are plagued with formidable fine tuning problems. The fine tuning problems of $`\mathrm{\Lambda }0`$ are well known . For the alternative of a slowly rolling scalar field extraordinarily flat potentials are required, so that the field is unable to roll to the true minimum by the present day. If the necessary flatness of such a potential is characterized by a mass scale $`m`$, one requires $`m10^{33}\mathrm{eV}`$.
In this letter we suggest another form of dark matter with significant negative pressure—a solid dark matter (SDM) component with the properties of a relativistic solid . The term solid here denotes a substance with a harmonic, non-dissipative resistance to pure shear (volume preserving) deformations. A perfect fluid with negative pressure ($`w=p/\rho <0`$) would have an imaginary sound speed, indicating instabilities most severe on the smallest scales. For a solid, however, with a sufficiently large shear modulus these instabilities are removed. Since the sound speed of perturbations of the solid should comprise a substantial fraction of the speed of light, its Jeans length today is comparable to the size of the current horizon, and hence an SDM component would remain unclustered on the smaller scales over which $`\mathrm{\Omega }_\mathrm{m}`$ is measured and thus evade detection.
To compute the effect of an exotic dark matter component on the evolution of cosmological perturbations, more is required than just knowing how $`w`$ evolves throughout cosmic history . As we will see, it is possible to construct dark matter components for which $`w`$ agrees at all times, but which respond differently to cosmological perturbations. Because of the non-vanishing shear modulus in the SDM case, large anisotropic stresses are generated, whereas in models with a $`\mathrm{\Lambda }0`$ or a slowly rolling scalar field these stresses vanish. Furthermore, long-wavelength gravity waves entering the horizon at late times acquire an effective mass because of the energetic cost associated with pure shear deformations of spacetime. In a previous paper two of us (MB and DNS) developed a formalism for computing the evolution of cosmological perturbations in the presence of a SDM component and computed the cosmic microwave background (CMB) anisotropy on large scales. Here we discuss the viability of such scenarios for creating cosmic structure and the prospects for distinguishing SDM from other types of dark matter with significant negative pressure using future measurements of the CMB. A subsequent paper will elucidate in more detail the differences between these scenarios and other models, such as quintessence and $`\mathrm{\Lambda }0.`$
A possible microphysical origin for SDM is a frustrated network of topological defects, either cosmic strings or domain walls (see ref. for a review of cosmic defects). The standard picture for the formation and evolution of topological defects is that of a random network of defects produced in a cosmological phase transition which then evolves toward a self-similar scaling regime with the number of defects within a Hubble volume approaching a fixed number . However, such behavior relies on the ability of the defects to untangle and lose energy essentially as fast as allowed by causality, and in more complicated models this need not occur. In particular, in theories with several species of defects or with non-Abelian symmetries, topological obstructions to such untangling may arise and defect-dominated evolution can occur. The basic features of such evolution would be that after the decay of an initial transient the number of defects per co-moving volume approaches a constant. For the simple domain walls of most interest here this implies an equation of state with $`w=2/3`$ ($`\rho _{\mathrm{dw}}a^1`$) and for strings $`w=1/3`$ ($`\rho _{\mathrm{str}}a^2`$), although other values of $`w`$ may be possible as well. Simulations of non-Abelian cosmic strings indicate behavior leading to a string-dominated universe and simulations of domain walls find that the density of walls in a model with several vacua falls less rapidly than would be required by scaling. In axion models, where an additional symmetry breaking $`U(1)Z_N`$ allows for $`N`$ domain walls to meet at a string, scaling behavior has been observed . The precise criterion for when defect domination takes place is not clear at present, although it seems clear that such models can exist. This is currently under investigation.
The symmetry breaking scale $`\eta `$ required for the formation of a network of frustrated domain walls with $`\mathrm{\Omega }_{\mathrm{dw}}1`$ today is around $`\eta 100\mathrm{KeV}`$, assuming a phase transition at $`T\eta `$, with initially one wall per horizon volume and a network immediately settling down to an equilibrium configuration which is subsequently swept along by the Hubble flow. This estimate, however, is subject to considerable uncertainties in either direction. For example, a long transient before settling down to an equilibrium configuration could considerably raise $`\eta `$. The mean separation between walls is approximately 30 parsecs, much smaller than the scale on which the network clusters in response to gravitational perturbations. It is therefore justified to treat the network as a continuum solid when studying its response to cosmological perturbations.
In a recent article it was suggested that a quintessence model with a flat geometry $`w2/3`$, $`\mathrm{\Omega }_\mathrm{c}=0.25`$, $`\mathrm{\Omega }_\mathrm{b}=0.05`$, $`h=0.65`$, $`n=1`$ gave a adequate fit to data probing a range of epochs and scales, where $`\mathrm{\Omega }_\mathrm{X}`$ is the fractional density relative to the critical density in particle species X (X=c is CDM, X=b is baryons), the Hubble constant is $`H_0=100h\mathrm{km}\mathrm{sec}^1\mathrm{Mpc}^1`$ and $`n`$ is the spectral index of the initial density perturbations. One might, therefore, wonder whether a domain wall dominated model with $`\mathrm{\Omega }_{\mathrm{dw}}0.7`$ might fit the data equally well and, in fact, a number of the calculations presented in ref. apply equally well to the case of a SDM model. However, important differences in how perturbations evolve once the SDM component comes to dominate must be considered for an accurate comparison to the data. Due to the apparently good fit of the $`w=2/3`$ case, most of the discussion focuses on the domain wall dominated case, but we also discuss the string dominated case and other values of $`w`$. Except where expressly stated, all our results use the parameter choices listed above.
We have included the evolution of a SDM component and its effects on the other perturbation variables into the standard Einstein-Boltzmann solver CMBFAST. As well as specifying $`w`$, this requires the introduction of a another parameter $`c_s`$ which is the sound speed of scalar perturbations in the solid, related to the the vector sound speed, $`c_v`$, by $`c_s^2=4c_v^2/3+w`$. For a given $`w`$, the evolution of the Newtonian potentials $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$ at late times differs as one varies $`c_s`$, and also differs from quintessence models. This leads to distinct integrated Sachs-Wolfe (ISW) contributions to the CMB anisotropies. At recombination, when the density of the SDM component is negligible (for $`w<1/3`$), these differences are suppressed and the small-angle anisotropies are identical for the same initial fluctuations. For this reason, when the results of our computations are plotted in Fig. 1, they are normalized at an angular scale corresponding to $`\mathrm{}=500`$ and spectra for different values of $`c_s`$ are then identical for $`\mathrm{}>100`$. Models with different $`w`$, however, have different peak positions with all the other cosmological parameters fixed, since the angular diameter distance $`\mathrm{}_D=k_\mathrm{p}\eta _0`$ depends on $`w.`$ The wavenumber $`k_\mathrm{p}`$ corresponds to the first acoustic peak and any change in $`\eta _0`$ due to a variation in $`w`$ can be offset by modifying $`k_\mathrm{p}`$, as discussed in ref. . In Fig. 1, the angular scale has been rescaled so that all the models have the angular diameter distance of the model with $`w=1`$. Distinguishing among the range of models considered will require accurate measurements of the CMB anisotropies on large scales.
Table 1 examines the significance of these differences by comparing flat models with identical cosmological parameters, but with differing properties for the component with significant negative pressure rescaled and normalized as described above. An estimate of the significance may be obtained by ignoring observational uncertainties and incomplete sky coverage because of the galaxy and considering only the effect of cosmic variance, arising from the fact that for each $`\mathrm{}`$ one is able to observe only $`(2\mathrm{}+1)`$ realizations of a Gaussian random process. Assuming that a model A is correct, the expectation value of the natural logarithm of the relative likelihood of model A relative to a model B is
$$\mathrm{ln}\left(\frac{P(\{a_\mathrm{}m\}|A)}{P(\{a_\mathrm{}m\}|B)}\right)_A=\frac{1}{2}\underset{\mathrm{}}{}(2\mathrm{}+1)\left[1\frac{C_{\mathrm{}}^{(A)}}{C_{\mathrm{}}^{(B)}}+\mathrm{ln}\left(\frac{C_{\mathrm{}}^{(A)}}{C_{\mathrm{}}^{(B)}}\right)\right].$$
(1)
For a domain wall dominated universe $`(w=2/3)`$ we find it virtually impossible to distinguish between differing sound speeds or between SDM and quintessence, while for $`w=1/3`$ we observe substantial differences between differing sound speeds, and between SDM and quintessence. Moreover, all the models are distinguishable from $`\mathrm{\Lambda }`$ model with the same value of $`\mathrm{\Omega }_\mathrm{m}`$. The relative likelihood of different models increases as $`\mathrm{\Omega }_\mathrm{m}`$ is decreased .
An important test for any cosmological model is whether it can create the large-scale structure (LSS) seen today when normalized to the the CMB anisotropies observed by COBE. The simplest version of this is to compare the COBE normalized value of $`\sigma _8`$, the variance of the density field in spheres of radius $`8h^1\mathrm{Mpc},`$ with that obtained from the observations of X-ray clusters . Models with $`w=1`$ pass this test, but two effects can cause models with $`w>1`$ to predict a lower $`\sigma _8`$ when normalized to COBE: a large integrated Sachs-Wolfe component reduces the predicted value of $`\sigma _8`$, and for larger $`w`$ the growth of perturbations becomes stunted earlier because the SDM (or quintessence) component begins to dominate the universe earlier. To investigate the viability of these models we compute the COBE normalized $`\sigma _8`$ for $`w=2/3`$ and a range of values in the $`\mathrm{\Omega }_\mathrm{m}h`$ plane, which are then compared to the recently computed observational value $`\sigma _8\mathrm{\Omega }_\mathrm{m}^\gamma =(0.50.1\mathrm{\Theta })\pm 0.1`$, where $`\gamma =0.210.22w+0.33\mathrm{\Omega }_\mathrm{m}+0.25\mathrm{\Theta }`$ and $`\mathrm{\Theta }=(n1)+(h0.65)`$, valid in the range of parameters considered. Fig. 2 shows the regions of $`\mathrm{\Omega }_\mathrm{m}h`$ plane which pass this test for spectral indices $`n=1.0`$ and $`n=1.1.`$ For $`n=1.0,`$ the preferred values for quintessence models $`h=0.65`$ and $`\mathrm{\Omega }_\mathrm{m}=0.3`$ are marginally incompatible with the quoted values for $`\sigma _8`$, but this situation is easily rectified by increasing $`h`$. Furthermore, these parameters are compatible with $`n=1.1`$, although the viability of scenarios with $`w=1/3`$ would require even larger values of $`n`$. The constraint from the upper bound on $`\sigma _8`$ is weaker because of the possibility of including a tensor contribution to the CMB anisotropies .
We have shown that the SDM proposal, and in particular a domain wall dominated universe, is compatible with LSS via the COBE normalization/$`\sigma _8`$ test and that the inclusion of a SDM component, as opposed to quintessence, can have some potentially interesting effects on the CMB anisotropies. If the SNIa results are confirmed, an important question will be how to distinguish between SDM and quintessence models. Some possibilities include cross-correlating the CMB with LSS using, for example the X-ray background , or the gravitational lensing of the CMB . Observations of gravitational lensing can measure the evolution of the power spectrum with redshift , and in particular the proposed Dark Matter Telescope would measure the growth of perturbations with high accuracy.
We would like to thank Brandon Carter, Neil Turok, and Alexander Vilenkin for useful discussions, and Uros Seljak and Matias Zaldariagga for the use of CMBFAST. RAB was funded by Trinity College, MB was supported by PPARC, and DNS was supported in part by MAP.
|
no-problem/9908/astro-ph9908322.html
|
ar5iv
|
text
|
# Infrared emission from interstellar dust cloud with two embedded sources: IRAS 19181+1349
## 1 Introduction
Galactic star forming regions mostly comprise of Young Stellar Objects (YSOs) / protostars still buried inside / in the vicinity of the parent interstellar cloud from which they are formed. Hence the study of YSOs leads to understanding of the interstellar medium in the close neighborhood of the starbirth. Early evolution of star forming regions is obviously more important from the point of understanding the star formation process itself. Typically the photons from the embedded energy source, protostar / ZAMS star, get reprocessed by the dust component of the interstellar medium in the immediate neighborhood. The dust grains absorb / scatter the incident radiation, depending on their dielectric properties. The dust grains acquire an equilibrium temperature based on the local radiation field which depends on the distance from the energy source and secondary radiation from other grains. The reradiation from the grains in the outer regions, is what is observable. Hence, the emerging observable spectrum can in principle be connected to the spectrum of the embedded protostar / ZAMS star through detailed radiation transfer provided some details about the geometry are known.
In order to study the earliest stage of star formation, a large amount of observational effort is directed towards far infrared / sub-mm observations of prospective young star forming regions. Many Galactic star forming regions have been mapped with near diffraction limited angular resolutions, using Kuiper Airborne Observatory, Infrared Space Observatory (ISO), James Clerk Maxwell Telescope etc. in recent times. In many cases, the maps of continuum emission resolve several closeby individual intensity peaks as evident from the morphology of their isophots.
The present study is a step in the direction of extracting maximum possible information about the geometrical and physical details of the source by comparing radiative transfer models with observations in cases where two nearby sources are resolved. Here, the “nearby” implies the interference in energetics of each of the two resolved source by the other one. The heating of dust in a cloud with multiple sources has been studied by Rouan (1979) under certain simplification. However, the problem of more than one embedded source has rarely been addressed quantitatively. The existing observational data showing clear evidence of resolved multiple embedded sources justify the need to explore geometries dealing with more than one embedded source. Ghosh & Tandon (1985) attempted to study the case of two embedded sources with many simplifying assumptions. They neglected some basic phenomena like scattering which limited its applicability to $`\lambda 50`$ $`\mu `$m only. The present work is an extension of this earlier attempt by including the effect of isotropic scattering.
In section 2 the problem has been formulated and the radiation transfer scheme is described. In section 3, we model observations of IRAS 19181+1349 which shows evidence of having two embedded sources. The results of our modelling are then discussed.
## 2 Model Formulation
The primary aim of the present model is to reproduce the infrared emission from a star forming cloud with two embedded sources, keeping the computational complexities at a minimum level.
The interstellar cloud is assumed to be of cylindrical shape. As a starting point, an uniform density of the cloud has been assumed. The line joining the two embedded ZAMS stellar / protostellar energy sources defines the symmetry axis of the problem. Around each of the sources there will be a dust free cavity (Fig. 1). The existence of such a cavity is widely accepted due to evaporation of the dust grains in the intense radiation field. In addition, radiation pressure effects on the dust grains may also play a role in deciding the cavity size. The radiation transfer is carried out through the dust component alone. Dust grains with a continuous size distribution have been considered, and their composition is a parameter of modelling. Three types of grains, viz., Graphite, Astronomical Silicate and Silicon Carbide have been invoked since their existence is generally accepted. All properties of the dust grains, viz., absorption and scattering coefficients as a function of wavelength, for various sizes and all three types of grains, have been taken from Laor & Draine (1993). The size distribution of all the three types of grains has been assumed to be power law ($`n(a)daa^{3.5}da`$) as per Mathis, Rumpl & Nordsieck (1977). The wavelength grid used 89 points covering from the Lyman continuum limit to the millimeter wavelengths.
The geometrical parameters relevant to the model (Fig. 1) include the radius of the cylinder (R<sub>cyl</sub>), radii of dust free cavities near the two sources ( R<sub>c1</sub>, R<sub>c2</sub>), and distance between two exciting sources (D). Other physical parameters are the composition of the dust (relative abundances of three types of dust components) and the dust density expressed in optical depth at 100 $`\mu `$m ($`\tau _{100}`$). The optical depth at any other wavelength is uniquely connected to $`\tau _{100}`$ via the dust properties and composition assumed in the perticular model. The cylindrical interstellar cloud is divided into $`n_z`$ identical discs. Each of these discs are further divided into $`n_r`$ annular rings. The $`n_z`$ is chosen such that each disc is optically thin even at UV, along the z-axis. Along the radial direction (total number of grid points being $`n_r`$), a two stage grid has been employed which is initially nearly logarithmically spaced (near the symmetry axis) and linearly spaced in the outer regions of the cylinder. This scheme of radial grid has been arrived at by keeping the optical depth related inaccuracies under check, for the entire wavelength region considered for the radiation transfer. Both the near logarithmic and the linear grids are matched by ensuring that radial cell size, $`\delta r(n_r)`$, is a smooth function of $`n_r`$. For modelling attempts of IRAS 19181+1349 a grid of 600 points in axial direction and of 25 points in radial direction were employed.
The relevant calculations are represented in equations 1-7. For clarity the frequency suffix has been dropped from all the terms, though calculations are performed for each frequency grid point. The quantities in angled brackets are averages over the dust size distribution. The code is simplified and optimized in many ways in view of the memory requirements and speed. Initially, factors totally dependent on the geometry of the problem and which do not change in each iteration are calculated. These include the optical depth terms and the geometric integrals involved in computations of radiation received by an unit volume element of the ring $`i`$ from the ring $`j`$ (equation 1). The geometric symmetry is such that such terms only depend on the axial separation between the two rings and their respective radii. The total flux absorbed and scattered by unit volume in each ring due to radiation from other rings is then calculated (equation 2). Also, the radiation received by an unit volume element of the ring $`i`$ from embedded sources depends only on the geometry and the optical depth per unit length of the particular model and hence is fixed over iterations (equation 3). The equilibrium temperature of the dust grains (on the median circle) for any particular annular ring is calculated using an iterative scheme by equating the power radiated by the dust (equation 6) to the power absorbed (equation 7). The latter is contributed by the embedded exciting sources (attenuated by the line of sight dust) as well as secondary emission and scattering from dust grains in all other annular rings (equation 4). This simplifies the calculations leading to a set of coupled equations with only two parameters per ring, temperature ( $`T_i`$ ) and $`F^i`$ changing from iteration to iteration thus greatly reducing the memory requirements .
$$F_r^{i,j}=\underset{\theta }{}\frac{e^{<\pi a^2Q_{ext}>n_dd_{i,j,\theta }}}{d_{i,j,\theta }^2}[<\pi a^2Q_{abs}>n_dB(T_j)+\frac{<\pi a^2Q_{scat}>F^j}{4\pi <\pi a^2Q_{ext}>}]\mathrm{\Delta }r_j\mathrm{\Delta }z\mathrm{\Delta }\theta r_j$$
(1)
$$F_r^i=\underset{j=1,ji}{\overset{n_r\times n_z}{}}<\pi a^2Q_{ext}>n_dF_r^{i,j}$$
(2)
$$F_s^i=\underset{s=1}{\overset{2}{}}<\pi a^2Q_{ext}>\frac{\pi R_s^2F_s}{r_{i,s}^2}e^{<\pi a^2Q_{ext}>n_d(r_{i,s}r_{c,s})}$$
(3)
$$F^i=F_r^i+F_s^i$$
(4)
$$F_e^i=<\pi a^2Q_{abs}>4\pi B(T_i)$$
(5)
$$P_{Emitted}=F_e^i𝑑\nu $$
(6)
$$P_{Absorbed}=\frac{<\pi a^2Q_{abs}>}{<\pi a^2Q_{ext}>}(F_r^i+F_s^i)𝑑\nu $$
(7)
where
$`d^2`$ $`(z_iz_j)^2+(r_i)^2+(r_j)^22r_ir_j\mathrm{cos}(\theta )`$
$`\theta `$ is azimuthal angular difference between two volume units under consideration.
$`r_{i,s}^2`$ $`(z_iz_s)^2+r_i^2`$
$`r_{c,s}^2`$ $``$ dust cavity radius for source $`s`$.
$`Q_{ext}`$ $`Q_{abs}+Q_{scat}`$
$`n_d`$ $``$ number density of dust grains.
$`T_i`$ $``$ Temperature of dust in ring $`i`$.
$`F_s`$ $``$ surface flux spectrum for source $`s`$.
$`F_r^i`$ $``$ Flux absorbed and scattered by unit volume of ring $`i`$, due to other rings.
$`F_s^i`$ $``$ Flux absorbed and scattered by unit volume of ring $`i`$, due to sources.
$`F^i`$ $``$ Total flux absorbed and scattered by unit volume of ring $`i`$.
$`B(T)`$ $``$ Planck function.
The dust temperature for each annular ring, leading to the temperature distribution throughout the cloud, is determined iteratively. Initially (the very first iteration), only the two embedded sources power the heating of the grains. From the second iteration onwards the effects of secondary heating and scattering are taken into account. In this iterative procedure the temperature of each annular ring is gradually updated in each iteration satisfying the condition $`P_{Emitted}=P_{Absorbed}`$. The iterations are continued till the fractional changes in absorbed power for each annular ring, between successive iterations, reduces below the convergence criteria.
The emergent intensity distribution, as seen by a distant observer is predicted by integrating the emitted and scattered radiation along relevant lines of sight and taking account of extinction due to the line of sight optical depth. This spatial intensity distribution at any selected wavelengths is convolved with the relevant instrumental beam profile (PSF) for direct comparison with observations.
Before applying the scheme developed above to any astrophysical source it is necessary to verify its reliability and quantify its accuracy. For this, we simulate the case of a single embedded source by “dimming” one of the two sources thereby keeping our original code intact during the test runs. This simulates a single exciting source embedded on the symmetry axis of the uniform density cylindrically shaped cloud. The size of the cylindrical cloud has been chosen such that the results from other codes using spherically symmetric geometry could be compared effectively. We have used the well established code CSDUST3 (Egan, Leung & Spagna 1988) for such a comparison.
The radius of the cylindrical cloud used for comparing our code with the spherically symmetric code was identical to that of the “equivalent” spherical cloud. The length of the cylinder (along z-axis) is twice this radius. Figure 2 is a schematic of the cylindrical as well as the spherical models.
The assumed radius of the cylinder, as well as the sphere, was 2 pc. The radius of the dust free cavity near the source was adopted to be .01 pc. In order to make the comparison possible and effective, all model parameters were made identical for both the codes. These include : dust size distribution; dust number density (and hence total optical depth along the line of sight between the embedded source and the distant observer). A typical dust composition has been assumed consisting of 50 % Graphite and 50 % Astronomical Silicate. The embedded energy source was assumed to be a single ZAMS O5 star with luminosity $`9.0\times 10^5L_{}`$. Some parameters were explored (e.g. optical depth) to study regions of validity with specified accuracy, by changing them identically for both the schemes. A large range in the optical depth $`\tau _{100}`$, viz., 1.0 x $`10^3`$ to 0.1 was covered during the test runs.
The comparison of the emergent spectral energy distributions (SEDs) from both the schemes for various optical depths are shown in Fig. 3. It may be seen that the SEDs match quantitatively over a wide range of optical depths from mid infrared to millimeter wavelengths. There are some differences at near infrared wavelengths; this is to be expected from the differences in the cell sizes and the geometry. However, since the main motive of the present study is to interpret measurements in the wavebands beyond the mid–IR, our code may be considered to be satisfactory. From this comparative study we conclude that our code is accurate upto an optical depth corresponding to $`\tau _{100}0.06`$ ($`\tau _{1\mu m}6`$), for the present choice of grid points. This limit already corresponds to much denser clouds than generally found in the Galactic star forming regions.
## 3 IRAS 19181+1349
The Galactic star forming region IRAS 19181+1349 is a IRAS Point Source Catalog (IRAS PSC) source associated with the radio source G48.60+00. The presence of a radio continuum suggests ongoing high mass star formation in the region. This source has been resolved in two components in the 210 $`\mu `$m map (Fig. 4b) generated from the observations using the TIFR 1–meter balloon borne telescope (Karnik et al, 1999). Although a single IRAS PSC source is associated with this star forming region, the HIRES processing of the IRAS survey data has led to the resolution of these two sources in the 12 & 25 $`\mu `$m band maps. The map at 12 $`\mu `$m is shown in Fig. 4a. We denote FIR peak with higher R.A. as S1 and that of lower R.A. S2, respectively, hereafter. Zoonematkermani et al. (1990) have reported four compact radio sources 48.603+0.026,48.609+0.027,48.606+0.023 and 48.592+0.044 in this region. First three of these closely match S1 (positionally) while the fourth one matches S2. Kurtz, Churchwell & Wood (1994) have also reported three ultracompact HII regions which are positionally close to S1, indicating ongoing high mass star formation in a very early evolutionary phase. This source has also been observed using ISOCAM instrument onboard Infrared Space Observatory (ISO) using seven filters centred at four PAH features and 3 continuua (3.30, 3.72, 6.00, 6.75, 7.75, 9.63, 11.37 $`\mu `$m; Verma et al 1999). The ISOCAM images at all the bands show two extended prominent complexes with multiple peaks. Hence from radio as well as mid to far infrared observations the nature of IRAS 19181+1349 is established to be of double embedded source/cluster type.
This seems to be an ideal astrophysical example of interstellar cloud with two embedded sources. We attempt to model this source with our scheme to extract important physical parameters about the embedded energy sources as well as the intervening interstellar medium in IRAS 19181+1349. Next we describe the results obtained from such a study.
The source IRAS 19181+1349 was considered as a cylindrical dust cloud with two protostellar /ZAMS stellar sources embedded along the axis of the cylinder. The size of the cloud and the dust density were used as free parameters. The sum of the luminosities of the two embedded sources is determined by integrating the observed spectral energy distribution (SED). The observations used for this integration includes the four IRAS bands and the two TIFR bands. This total luminosity is treated as an observational constraint in the modelling. Further observational constraints include : the shape of the SED which was obtained from HIRES–IRAS, TIFR and ISO observations, and the structural morphology of IRAS 19181+1349 as reflected by the isophote contours of the high angular resolution maps at 12 & 210 $`\mu `$m. As our code does not deal with the physics of PAH emission at present, only the data in ISOCAM filters sampling the continuua have been used to constrain the model.
Two completely independent approaches have been followed in modelling IRAS 19181+1349 using our scheme. In each approach, all the model parameters are floated to obtain the best fit model. The parameters fine tuned to achieve the best fit model are : luminosities of individual embedded sources; geometrical size details of the cylindrical cloud (including the size of the cavity); and the dust density / optical depth. The main aim of the first approach is to optimize the fit to the observed SED (hereafter $`M_{SED}`$). The second approach optimizes the one dimensional radial intensity profiles at selected wavelengths covering mid to far infrared bands (hereafter $`M_{RC}`$). The radial profiles are taken along geometrically interesting axes, viz., along the line joining the two embedded sources and along lines perpendicular to the cylinder axis and passing through either of the two sources (see Fig. 4). Whereas the first approach gives precedence to the overall energetics the latter gives more importance to the structural details in the isophotes, particularly close to the embedded sources and the region between them. The actual reality may lie somewhere in between these two approaches.
The axes where radial cuts have been taken are also displayed on the 12 & 210 $`\mu `$m maps in Fig. 4. The fits to the observed SED for both $`M_{SED}`$ as well as $`M_{RC}`$ are presented in Fig. 5. Comparison of the radial cuts at 12 & 210 $`\mu `$m along the three axes between the observed maps and the best fit $`M_{RC}`$ model are shown in Fig. 6.
The $`M_{SED}`$ approach fits the observed SED very well right through near–IR to sub-mm, though there is some discrepancy in the far–IR (Fig. 5). On the other hand, fit from the $`M_{RC}`$ approach is resonable for $`\lambda `$ 25 $`\mu `$m only. The radial cuts at 210 $`\mu `$m fit the observed data very well. However, at 12 $`\mu `$m although the model predictions qualitatively agree with the data, there are discrepancies viz., the extent of emission along vertical axis and the relative contrast of the minima between the two sources. From the above it appears that the $`M_{RC}`$ approach is very sensitive to the exact distribution of sources, specially at shorter wavelengths (which traces much hotter dust and hence physically closer to the exciting source). One possibility for these discrepancies is that our model is too simplified than the actual reality in IRAS 19181+1349. For example there may be a distribution of sources along the line joining S1 and S2 and away from it, but with more concentration near S1 and S2. This way the discrepancy for $`M_{RC}`$ in the mid–IR part of SED will also be explained. Assuming a constant dust density in the cloud may also be a major reason for the discrepancy.
The results of modelling IRAS 19181+1349 in the form of best fit model parameters as found under both the approaches ($`M_{SED}`$ and $`M_{RC}`$) are presented in Table 1. The value of $`\tau _{100}`$ found from both the approaches are well within the range of validity of the code, i.e. much less than 0.06. The luminosities of both the sources are similar and the dust is found to be predominantly (80%) Silicate with no Silicon Carbide, the rest (20%) being Graphite. It is remarkable that almost all the important parameters are identical as found from both the approaches. The only difference is seen at the physical sizes of the cavities. The remarkable similar physical parameters obtained from both the approaches gives good confidence on the derived astrophysical parameters for the Galactic star forming region IRAS 19181+1349.
This scheme will be useful to model other similar double sources. A huge database from ISO (SEDs as well as images leading to radial cuts, covering 2.5 to 200 $`\mu m`$ of the spectrum) will become available soon for such modelling.
## 4 Summary
High angular resolution mid / far infrared maps of many Galactic star forming regions show evidence for multiple embedded energy sources in the corresponding interstellar clouds. With the aim of studying such star forming regions with two embedded sources, a scheme has been developed to carry out radiative transfer calculations in a uniform density dust cloud of cylindrical geometry, and which includes the effect of isotropic scattering in addition to the absorption and emission processes. In addition to the luminosities of the two embedded energy sources, the cylindrical cloud size, separation between the two sources, dust density and composition are the parameters for the modelling. The accuracy of our scheme has been tested by comparing the results with a well established 1-D code.
An attempt was made to model the Galactic star forming region associated with IRAS 19181+1349 which shows double peaks in the mid and far infrared maps using the scheme described here. Two independent approaches were employed to find the best fit models for IRAS 19181+1349. Whereas the first ($`M_{SED}`$) approach aims to fit the observed SED best; the latter ($`M_{RC}`$) aims to fit the radial intensity profiles (along a few important axes) at mid to far infrared wavebands. Interestingly most of the crucial model parameters like luminosities, effective temperatures, dust composition, optical depth etc turn out to be identical under both the approaches.
Acknowledgments
It is a pleasure to thank the members of the Infrared Astronomy Group of T.I.F.R. for their encouragement.
|
no-problem/9908/hep-ph9908397.html
|
ar5iv
|
text
|
# Neutralino–Nucleus Elastic Scattering in the MSSM with Explicit CP Violation
## I Introduction
Supersymmetry (SUSY) is one of the most promising theoretical frameworks for a successful unification of gravity with all other fundamental forces and is the most appealing perturbative solution to the gauge hierarchy problem of the standard model (SM). However, SUSY is not an exact symmetry of nature. The minimal supersymmetric SM (MSSM) , the minimal SUSY realization, must break SUSY softly in order to accomplish agreement with experimental observations and its breaking scale should not be much larger than a few TeV in order to retain naturalness. In general, the SUSY breakdown introduces a large number of unknown parameters, many of which can be complex . CP–violating phases associated with sfermions of the first and, to a lesser extent, second generations are severely constrained by bounds on the electric dipole moments of the electron, neutron and muon. The present experimental upper bounds on the neutron EDM $`d_n`$ and electron EDM $`d_e`$ are very tight : $`|d_n|<1.12\times 10^{25}e\mathrm{cm}`$ and $`|d_e|<0.5\times 10^{26}e\mathrm{cm}`$ at the 2–$`\sigma `$ level. As a result of the CP crises, some fine–tuning mechanisms are necessary in generic supersymmetric theories to avoid these problems. There have been several phenomenologically attractive solutions to evade these constraints without suppressing the CP–violating phases. One option is to make the first two generations of scalar fermions rather heavy so that one–loop EDM constraints are automatically evaded. As a matter of fact one can consider so–called effective SUSY models which seem to combine all healthy features of both the MSSM and technicolor theories. The main virtue of the effective SUSY model is that any non–SM source of CP violation and FCNC involving the first two generations is suppressed by allowing their respective soft–SUSY–breaking masses to be as high as 20 TeV, whereas third generation scalar quarks and leptons may naturally be light well below the TeV scale. Another possibility is to arrange for partial cancellations among various contributions to the electron and neutron EDM’s .
Following the suggestions that the CP–violating phases do not have to be always suppressed, many important works on the effects due to the CP–violating phases in the MSSM have been already reported; the effects are very significant in extracting the parameters in the SUSY Lagrangian from experimental data , estimating dark matter densities and scattering cross sections and Higgs boson mass limits , CP violation in the $`B`$ and $`K`$ systems , and so on. In particular, it has been found that the Higgs–sector CP violation induced via loop corrections of soft CP–violating Yukawa interactions may drastically modify the couplings of the light neutral Higgs boson to the gauge bosons. As a result, the production cross sections as well as the decay branching ratios of neutral Higgs bosons changes so significantly. One of the crucial modifications is that the current experimental lower bound on the lightest Higgs boson mass may be dramatically relaxed up to a 60–GeV level in the presence of large CP violation in the Higgs sector of the MSSM. Moreover, the explicit CP violation in the MSSM Higgs sector radiatively induces a finite unremovable misalignment between two Higgs doublets. This additional phase can be as large as the original CP phases in certain portions of the MSSM parameter space and affect the chargino and neutralino systems. Therefore, it would be very important to re–visit all the relevant phenomena by including the CP–violating induced phase while avoiding the severe constraints from the neutron and electron EDMs. In particular, one of the most phenomenologically interesting subjects is to investigate the possibility of detecting lightest neutralinos when a large portion of the dark matter in the Universe is composed of the lightest and (almost) stable lightest supersymmetric particle (LSP).
The observations of the dynamics of galaxies and clusters of galaxies , and the constraints on the baryon density from big bang nucleosynthesis requires the existence of a considerable amount of non-baryonic dark matter. Therefore, it is almost universally accepted that most of the mass in the Universe and most of the mass in the Galactic halo is dark and the dark matter consists of some new, as yet undiscovered, weakly–interacting massive particle (WIMP). Of the many WIMP candidates, one of the best motivated and the most theoretically developed is the lightest neutralino, the lightest supersymmetric particle (LSP) in most SUSY theories. In this light, there have been an intensive investigation for its detection and identification . In the present work, we re–investigate the LSP–nucleus elastic scattering process in the MSSM framework with R-parity and with CP–violating complex parameters. There have been already several works on the effects of the phase $`\mathrm{\Phi }_\mu `$ of the higgsino mass parameter $`\mu `$ on the neutralino–nucleus elastic scattering as well as the neutralino relic density . So, referring to those works for the effects from the phase $`\mathrm{\Phi }_\mu `$, we will mainly concentrate on the impact of the scalar–pseudoscalar mixing and the induced phase between two Higgs doublets in the MSSM Higgs sector stemming from the radiative corrections due to the CP–violating scalar top and bottom sectors in the MSSM on the neutralino–nucleus elastic scattering. For a more concrete, quantitative investigation, our analysis through the paper is based on a specific scenario with the following assumptions:
* The first and second generation sfermions are very heavy so that they are decoupled from the theory. In this case, there are no constraints on the CP–violating phases from the neutron and electron EDMs. On the other hand, the annihilation of the neutralinos into tau pairs through the exchange for relatively light scalar tau leptons guarantees that the cosmological constraints on the dark matter densities be satisfied.
* The explicit CP violation in the Higgs sector through the CP–violating radiative corrections from the scalar top and bottom sectors is included.
* Simultaneously, the effects of the induced CP–violating phase between two Higgs doublets on the chargino and neutralino systems are explicitly included.
* It is necessary to avoid the possible constraints from the so–called Barr-Zee–type diagrams to the electron and neutron EDMs as well as from the null results of the Higgs boson searches at LEP . We take two values 3 and 30 for $`\mathrm{tan}\beta `$, the ratio of the vacuum expectation values of two neutral Higgs fields in the present analysis.
Certainly, the major issue concerning the supersymmetric dark matter is its detection and identification. Indeed, there are a multitude of ongoing experiments involved in the direct and indirect detection of dark matter, many with a specific emphasis on searching for supersymmetric dark matter . The event rates for either direct or indirect detection depend crucially on the LSP-nucleon, or LSP-nucleus, cross-section. Because the neutralinos have Majorana mass terms, their interactions with matter are generally spin dependent, coming from an effective interaction term of the form $`(\overline{\chi }\gamma ^\mu \gamma _5\chi )(\overline{f}\gamma _\mu \gamma _5f)`$. In the regions of the MSSM parameter space where the LSP is a mixture of both gaugino and Higgsino components, there is also an important contribution to the scattering cross-section due to a term in the interaction Lagrangian of the form $`(\overline{\chi }\chi )(\overline{f}f)`$ which is spin independent. These terms are particularly important for scattering off of large nuclei, where coherent nucleon scattering effects can quickly come to dominate all others. The scalar–pseudoscalar mixing and the modified couplings of the neutral Higgs bosons to fermions and neutralinos affect the spin–independent part while the CP–violating induced phase affects both the spin–dependent and the spin–independent parts through its modifications of the structure of the neutralino mass matrix. In this light it is worthwhile to make a systematic investigation of the effects of the CP–violating phases on the neutralino–nucleus scattering cross section, which is the goal of the present work.
The organization of the present paper is as follows. In Section II, a brief review on the explicit CP violation in the Higgs sector is given following the work by Pilaftsis and Wager . Section III is devoted to a detailed analysis of various effects of the CP–violating induced phase between two Higgs doublets on the chargino and neutralino sectors. Then we give the fully analytic expressions for the spin–dependent and spin–independent neutralino–nucleus scattering cross sections in Section IV and give a detailed numerical analysis of the dependence of the cross sections on the CP–violating phases as well as real SUSY parameters such as $`\mathrm{tan}\beta `$, the size of the higgsino mass parameter $`|\mu |`$ and the trilinear terms $`|A_{t,b}|`$. Finally, we summarize our findings and conclude in Section V.
## II CP Violation in the MSSM Higgs sector
### A CP–violating radiative corrections
The MSSM introduces several new parameters in the theory that are absent from the SM and could, in principle, possess many CP–violating phases. Specifically, the new CP phases may come from the following parameters: (i) the higgsino mass parameter $`\mu `$, which involves the bilinear mixing of the two Higgs chiral superfields in the superpotential; (ii) the soft SUSY–breaking gaugino masses $`M_a`$ ($`a=1,2,3`$), where the index $`a`$ stands for the gauge groups U(1)<sub>Y</sub>, SU(2)<sub>L</sub> and SU(3)<sub>C</sub>, respectively; (iii) the soft bilinear Higgs mixing masses $`m_{12}^2`$, which is sometimes denoted as $`B\mu `$ in the literature; (iv) the soft trilinear Yukawa couplings $`A_f`$ of the Higgs particles to scalar fermions; and (v) the flavor mixing elements of the sfermions mass matrices. If the universality condition is imposed on all gaugino masses at the unification scale $`M_X`$, the gaugino masses $`M_a`$ have a common phase, and if the diagonal boundary conditions are added to the universality condition for the sfermion mass matrices at the GUT scale, the flavor mixing elements of the sfermions mass matrices vanish and the different trilinear couplings $`A_f`$ are all equal, i.e. $`A_f=A`$.
The conformal–invariant part of the MSSM Lagrangian has two global U(1) symmetries; the U(1)<sub>Q</sub> Peccei–Quinn symmetry and the U(1)<sub>R</sub> symmetry acting on the Grassmann–valued coordinates. As a consequence, not all CP–violating phases of the four complex parameters $`\{\mu ,m_{12}^2,M_a,A\}`$ turn out to be physical, i.e. two phases may be removed by redefining the fields accordingly . Employing the two global symmetries, one of the Higgs doublets and the gaugino fields can be rephased such that $`M_a`$ and $`m_{12}^2`$ become real. In this case, arg($`\mu `$) and arg($`A`$) are the only physical CP–violating phases in the low–energy MSSM supplemented by universal boundary conditions at the GUT scale. Denoting the scalar components of the Higgs doublets $`H_1`$ and $`H_2`$ by $`H_1=i\tau _2\mathrm{\Phi }_1^{}`$ ($`\tau _2`$ is the usual Pauli matrix) and $`H_2=\mathrm{\Phi }_2`$, the most general CP–violating Higgs potential of the MSSM can be conveniently described by the effective Lagrangian
$`_V`$ $`=`$ $`\mu _1^2(\mathrm{\Phi }_1^{}\mathrm{\Phi }_1)+\mu _2^2(\mathrm{\Phi }_2^{}\mathrm{\Phi }_2)+m_{12}^2(\mathrm{\Phi }_1^{}\mathrm{\Phi }_2)+m_{12}^2(\mathrm{\Phi }_2^{}\mathrm{\Phi }_1)`$ (4)
$`+\lambda _1(\mathrm{\Phi }_1^{}\mathrm{\Phi }_1)^2+\lambda _2(\mathrm{\Phi }_2^{}\mathrm{\Phi }_2)^2+\lambda _3(\mathrm{\Phi }_1^{}\mathrm{\Phi }_1)(\mathrm{\Phi }_2^{}\mathrm{\Phi }_2)+\lambda _4(\mathrm{\Phi }_1^{}\mathrm{\Phi }_2)(\mathrm{\Phi }_2^{}\mathrm{\Phi }_1)`$
$`+\lambda _5(\mathrm{\Phi }_1^{}\mathrm{\Phi }_2)^2+\lambda _5^{}(\mathrm{\Phi }_2^{}\mathrm{\Phi }_1)^2+\lambda _6(\mathrm{\Phi }_1^{}\mathrm{\Phi }_1)(\mathrm{\Phi }_1^{}\mathrm{\Phi }_2)+\lambda _6^{}(\mathrm{\Phi }_1^{}\mathrm{\Phi }_1)(\mathrm{\Phi }_2^{}\mathrm{\Phi }_1)`$
$`+\lambda _7(\mathrm{\Phi }_2^{}\mathrm{\Phi }_2)(\mathrm{\Phi }_1^{}\mathrm{\Phi }_2)+\lambda _7^{}(\mathrm{\Phi }_2^{}\mathrm{\Phi }_2)(\mathrm{\Phi }_2^{}\mathrm{\Phi }_1).`$
In the Born approximation, the quartic couplings $`\lambda _{1,2,3,4}`$ are solely determined by the gauge couplings and $`\lambda _{5,6,7}`$ are zero. However, beyond the Born approximation, the quartic couplings $`\lambda _{5,6,7}`$ receive significant radiative corrections from trilinear Yukawa couplings of the Higgs fields to scalar–top and scalar–bottom quarks. These parameters are in general complex and so lead to CP violation in the Higgs sector through radiative corrections. The explicit form of the couplings with radiative corrections can be found in Refs. .
It is necessary to determine the ground state of the Higgs potential to obtain physical Higgs states and their self–interactions. To this end we introduce the linear decompositions of the Higgs fields
$`\mathrm{\Phi }_1=\left(\begin{array}{cc}\varphi _1^+& \\ \frac{1}{\sqrt{2}}(v_1+\varphi _1+ia_1)& \end{array}\right),\mathrm{\Phi }_2=\mathrm{e}^{i\xi }\left(\begin{array}{cc}\varphi _2^+& \\ \frac{1}{\sqrt{2}}(v_2+\varphi _2+ia_2)& \end{array}\right),`$ (9)
with $`v_1`$ and $`v_2`$ the moduli of the vacuum expectation values (VEVs) of the Higgs doublets and $`\xi `$ their CP–violating induced relative phase. These VEVs and the relative phase can be determined by the minimization conditions on $`_V`$, which can be efficiently performed by the so–called tadpole renormalization techniques . It is always guaranteed that one combination of the CP–odd Higgs fields $`a_1`$ and $`a_2`$ ($`G^0=\mathrm{cos}\beta a_1\mathrm{sin}\beta a_2`$) defines a flat direction in the Higgs potential and so it is absorbed as the longitudinal component of the $`Z`$ boson. \[Here, $`\mathrm{sin}\beta =v_2/\sqrt{v_1^2+v_2^2}`$ and $`\mathrm{cos}\beta =v_1/\sqrt{v_1^2+v_2^2}`$.\] As a result, there exist one charged Higgs state and three neutral Higgs states that are mixed in the presence of CP violation in the Higgs sector. Denoting the remaining CP–odd state $`a=\mathrm{sin}\beta a_1+\mathrm{cos}\beta a_2`$, the $`3\times 3`$ neutral Higgs–boson mass matrix describing the mixing between CP–even and CP–odd fields can be decomposed into four parts in the weak basis $`(a,\varphi _1,\varphi _2)`$ :
$`_H^2=\left(\begin{array}{cc}_P^2& _{PS}^2\\ _{SP}^2& _S^2\end{array}\right),`$ (12)
where $`_P^2`$ and $`_S^2`$ describe the CP–preserving transitions $`aa`$ and $`(\varphi _1,\varphi _2)(\varphi _1,\varphi _2)`$, respectively, and $`_{PS}^2=(_{SP}^2)^T`$ contains the CP–violating mixings $`a(\varphi _1,\varphi _2)`$. The analytic form of the sub–matrices can be found in Ref. .
On the other hand, the charged Higgs-boson mass $`m_{H^\pm }`$ is related to the pseudoscalar mass term $`m_a`$ as
$`m_a^2=m_{H^\pm }^2{\displaystyle \frac{1}{2}}\lambda _4v^2+(\lambda _5\mathrm{e}^{2i\xi })v^2.`$ (13)
Taking this very last relation between $`m_{H^\pm }`$ and $`m_a`$ into account, we can express the neutral Higgs–boson masses as functions of $`m_{H^\pm }`$, $`\mu `$, $`A`$, a common SUSY scale $`M_{\mathrm{SUSY}}`$, $`\mathrm{tan}\beta `$ and the physical phase $`\xi `$, which cannot be rotated away in the presence of the chargino and neutralino contributions. Clearly, the CP–even and CP–odd states mix unless all of the imaginary parts of the parameters $`\lambda _{5,6,7}`$ vanish. Since the Higgs—boson mass matrix $`_H^2`$ describing the scalar–pseudoscalar mixing is symmetric, we can diagonalize it by means of an orthogonal rotation $`O`$; $`O^T_H^2O=\mathrm{diag}(m_{H_3}^2,m_{H_2}^2,m_{H_1}^2)`$ with the ordering of masses $`m_{H_1}m_{H_2}m_{H_3}`$. The neutral Higgs–boson mixing affects the couplings of the Higgs fields to fermions, gauge bosons, and Higgs fields themselves as shown in the following.
The CP–violating effects due to radiative corrections to the Higgs potential is characterized by a dimensionless parameter $`\eta _{{}_{C}{}^{}P}`$
$`\eta _{_{CP}}={\displaystyle \frac{m_{t,b}^4}{v^4}}\left({\displaystyle \frac{|\mu ||A|}{32\pi ^2M_{\mathrm{SUSY}}^2}}\right)\mathrm{sin}\mathrm{\Phi }_{_{CP}},`$ (14)
where $`\mathrm{\Phi }_{_{CP}}=\mathrm{arg}(A\mu )+\xi `$, i.e. the sum of three CP–violating phases. So, for $`|\mu |`$ and/or $`|A|`$ values larger than the SUSY–breaking scale $`M_{\mathrm{SUSY}}`$, the CP–violating effects can be significant.
### B An induced CP–violating phase
As shown in the previous section, the first derivatives of the Higgs potention with respect to the neutral fields $`\{\varphi _1,\varphi _2,a\}`$ do not vanish any longer; hence, one has to redefine the Higgs doublet fields with a relative phase $`\xi `$. This CP–violating induced phase can be obtained analytically by combining the two relations for the minimization. First of all, we make use of the fact that a U(1)<sub>PQ</sub> rotation allows us to take $`m_{12}^2`$ to be real and for a notational convenience define $`\stackrel{~}{\lambda }_6`$, $`\delta `$, and $`\delta ^{}`$;
$`\stackrel{~}{\lambda }_6=\lambda _6c_\beta ^2+\lambda _7s_\beta ^2,`$ (15)
$`\delta =\left({\displaystyle \frac{m_{H^\pm }^2}{v^2}}{\displaystyle \frac{\lambda _4}{2}}+\lambda _5\right)\mathrm{sin}2\beta ,\delta ^{}=\left({\displaystyle \frac{m_{H^\pm }^2}{v^2}}{\displaystyle \frac{\lambda _4}{2}}\lambda _5\right)\mathrm{sin}2\beta .`$ (16)
Then, the CP–violating induced phase $`\xi `$ is determined by the relations;
$`\mathrm{sin}\xi ={\displaystyle \frac{1}{|\delta |^2}}\left\{(\delta )(\stackrel{~}{\lambda }_6)(\delta )\sqrt{|\delta |^2^2(\stackrel{~}{\lambda }_6)}\right\},`$ (17)
$`\mathrm{cos}\xi =+{\displaystyle \frac{1}{|\delta |^2}}\left\{(\delta )(\stackrel{~}{\lambda }_6)+(\delta )\sqrt{|\delta |^2^2(\stackrel{~}{\lambda }_6)}\right\},`$ (18)
and the soft–breaking positive bilinear mass squared $`m_{12}^2`$ is given by
$`m_{12}^2={\displaystyle \frac{v^2}{2|\delta |^2}}\left\{(\delta \delta ^{})(\stackrel{~}{\lambda }_6)+(\delta \delta ^{})\sqrt{|\delta |^2^2(\stackrel{~}{\lambda }_6)}|\delta |^2(\stackrel{~}{\lambda }_6)\right\}.`$ (19)
We note that the induced phase vanishes if $`\mathrm{arg}(A\mu )`$, the sum of the phases $`\mathrm{\Phi }_\mu `$ and $`\mathrm{\Phi }_A`$, vanishes, even if each of the CP–violating phases might not have to vanish. On the other hand, the size of $`\delta `$ or $`\delta ^{}`$ is proportional to the pseudoscalar mass $`m_a`$ to a very good approximation so that if it becomes large, i.e, decoupled, the induced phase $`\xi `$ is diminished. Since the size of the induced phase is also inversely proportional to $`\mathrm{sin}2\beta `$, the phase grows with increasing $`\mathrm{tan}\beta `$. In general, this induced phase will remain as a non–trivial physical phase and lead to a modification in the chargino and neutralino mass matrices.
Analytically, the induced phase $`\xi `$ palys a role of rotating the vacuum expectation value $`v_2`$ into $`v_2\mathrm{e}^{i\xi }`$. So, the chargino mass matrix is given in the $`(\stackrel{~}{W}^{},\stackrel{~}{H}^{})`$ basis by
$`_C=\left(\begin{array}{cc}M_2& \sqrt{2}m_Wc_\beta \\ \sqrt{2}m_Ws_\beta \mathrm{e}^{i\xi }& |\mu |\mathrm{e}^{i\mathrm{\Phi }_\mu }\end{array}\right),`$ (23)
which is built up by the fundamental SUSY parameters; the SU(2) gaugino mass $`M_2`$, the higgsino mass parameter $`|\mu |`$, its phase $`\mathrm{\Phi }_\mu `$ and the ratio $`\mathrm{tan}\beta `$ of the vacuum expectation values of the two neutral Higgs doublet fields<sup>*</sup><sup>*</sup>*We note that the vacuum expectation value of the Higgs doublet $`H_1`$ has an opposite sign to the conventional one in the literature. This is the reason why there appears a negative sign in the $`(12)`$ component of the chargino mass matrix. The gaugino mass $`M_2`$ was made real already by appropriate field redefinitions. Since the chargino mass matrix $`_C`$ is not symmetric, two different unitary matrices acting on the left– and right–chiral $`(\stackrel{~}{W}^{},\stackrel{~}{H}^{})`$ states are needed to diagonalize the matrix:
$`U_{L,R}\left(\begin{array}{c}\stackrel{~}{W}^{}\\ \stackrel{~}{H}^{}\end{array}\right)=\left(\begin{array}{c}\stackrel{~}{\chi }_1^{}\\ \stackrel{~}{\chi }_2^{}\end{array}\right),`$ (28)
and the mass eigenvalues are given by
$`m_{\stackrel{~}{\chi }_{1,2}^\pm }^2={\displaystyle \frac{1}{2}}\left[M_2^2+|\mu |^2+2m_W^2\mathrm{\Delta }\right],`$ (29)
with $`\mathrm{\Delta }`$ involving the phases $`\mathrm{\Phi }_\mu `$ and the induced phase $`\xi `$:
$`\mathrm{\Delta }=\left\{(M_2^2|\mu |^2)^2+4m_W^4\mathrm{cos}^22\beta +4m_W^2(M_2^2+|\mu |^2M_2|\mu |\mathrm{sin}2\beta \mathrm{cos}(\mathrm{\Phi }_\mu \xi ))\right\}^{1/2}.`$ (30)
As a matter of fact, an additional field redefinition enables one to find that every CP–violating phenomenon due to the chargino mixing is dependent on the difference $`\mathrm{\Phi }_\mu \xi `$ of two phases $`\mathrm{\Phi }_\mu `$ and $`\xi `$. Keeping in mind this point, one can infer the following aspects:
* Even if $`\mathrm{\Phi }_\mu `$ vanishes there is still a source of CP violation due to the presence of the induced phase $`\xi `$ stemming from the CP–violating phases in the scalar top and bottom sectors.
* If both $`\mathrm{\Phi }_\mu `$ and $`\mathrm{\Phi }_A`$ vanish as in the CP–invariant theory, then the induced phase $`\xi `$ vanishes, leaving no source of CP violation.
These effects may be investigated through the chargino pair production as soon as the collider energies become high enough to go over their production thresholds. However, this investigation is beyond the regime of the present work and so, it will be not touched upon here.
Similarly, the neutralino mass matrix describing neutralino mixing is modified by the introduction of the CP–violating induced phase. Following the same prescription for the chargino mass matrix yields the neutralino mass matrix in the $`(\stackrel{~}{B},\stackrel{~}{W}^3,\stackrel{~}{H}_{}^{0}{}_{1}{}^{},\stackrel{~}{H}_{}^{0}{}_{2}{}^{})`$ basis:
$`_N=\left(\begin{array}{cccc}|M_1|\mathrm{e}^{i\mathrm{\Phi }_1}& 0& m_Zs_Wc_\beta & m_Zs_Ws_\beta \mathrm{e}^{i\xi }\\ 0& M_2& m_Zc_Wc_\beta & m_Zc_Ws_\beta \mathrm{e}^{i\xi }\\ m_Zs_Wc_\beta & m_Zc_Wc_\beta & 0& |\mu |\mathrm{e}^{i\mathrm{\Phi }_\mu }\\ m_Zs_Ws_\beta \mathrm{e}^{i\xi }& m_Zs_Ws_\beta \mathrm{e}^{i\xi }& |\mu |\mathrm{e}^{i\mathrm{\Phi }_\mu }& 0\end{array}\right).`$ (35)
The neutralino mass matrix $`_N`$ is a complex but symmetric matrix so that it can be diagonalized by just one unitary matrix $`N`$ such that $`N^{}_NN^{}=\mathrm{diag}(m_{\stackrel{~}{\chi }_1^0},m_{\stackrel{~}{\chi }_2^0},m_{\stackrel{~}{\chi }_3^0},m_{\stackrel{~}{\chi }_4^0})`$ with the increasing ordering in masses. A simple orthogonality transformation of a phase matrix enables one to confirm that any physical observable related with the neutralino mixing depends only on the phase $`\mathrm{\Phi }_1`$ and the combination $`\mathrm{\Phi }_\mu \xi `$ of the phases $`\mathrm{\Phi }_\mu `$ and $`\xi `$ as in the chargino system. So, it is clear that except for the phase $`\mathrm{\Phi }_1`$ the neutralino system exhibits a similar dependence on the phase $`\mathrm{\Phi }_\mu `$ and the induced phase $`\xi `$. In the present work, we take the assumption of gaugino mass unification for which the phase $`\mathrm{\Phi }_1`$ should be zero at least up to one–loop level against the renormalization group running from the unification scale to the electroweak scale. In this scenario, there exists only one CP–violating rephasing–invariant phase $`\mathrm{\Phi }_\mu \xi `$ in the chargino and neutralino mass sectors.
## III Neutralino–nucleus elastic scattering
### A Four–Fermi effective Lagrangian
In this section we investigate the importance of the CP-violating phases on the elastic scattering cross-sections of neutralinos on nuclei. To this end, we calculate the four-Fermi effective $`\chi `$-quark interaction Lagrangian with the inclusion of the CP violating phases $`\mathrm{\Phi }_\mu `$, $`\mathrm{\Phi }_1`$, and $`\xi `$ for the standard spin–dependent and spin–independent neutralino–nucleus interactions. There are two types of diagrams contributing to the elastic scattering; $`Z`$-exchange diagram and neutral Higgs boson exchange diagrams as shown in Fig. 1. In order to derive the analytic expression for the scattering cross sections, it is necessary to determine the interactions of the $`Z`$ and Higgs bosons to neutralinos and fermions.
Firstly, the interactions of the neutral Higgs fields with SM fermions are described by the Lagrangian
$`_{H\overline{f}f}=H_{4\alpha }\left\{{\displaystyle \frac{gm_d}{2m_Ws_\beta }}\overline{d}\left[O_{2\alpha }is_\beta O_{1\alpha }\gamma _5\right]d+{\displaystyle \frac{gm_u}{2m_Wc_\beta }}\overline{u}\left[O_{3\alpha }ic_\beta O_{1\alpha }\gamma _5\right]u\right\}.`$ (36)
Here, $`u`$ denotes one of up–type fermions and $`d`$ one of down–type fermions. Obviously, the Higgs–fermion–fermion couplings are significant for the third–generation fermions, $`t`$, $`b`$ and $`\tau `$ because of their relatively large Yukawa couplings. On the contrary, because any ordinary nucleus is mainly composed of the first (and second) generation fermions, the couplings are very small. Nevertheless, these contributions become important for a nucleus with a large atomic/mass number, because they can contribute to the spin–independent cross section in a coherent manner. We note in passing that the effect of CP–violating Higgs mixing is to induce a simultaneous coupling of $`H_i`$ ($`i=1,2,3`$) to CP–even and CP–odd fermion bilinears $`\overline{f}f`$ and $`\overline{f}i\gamma _5f`$ . This can lead to a sizable phenomenon of CP violation in the Higgs decays into polarized top-quark or tau-lepton pairs .
Secondly, the interactions of the Higgs bosons to neutralinos involve the neutralino mixing and the Higgs boson mixing simultaneously. As a result, the expression can become lengthy. So, for a notational convenience, we introduce the expression $`G_\alpha `$ defined in terms of the induced phase $`\xi `$, the neutralino diagonalization matrix $`N`$, and the Higgs diagonalization matrix $`O`$ as follows;
$`G_\alpha =(N_{12}t_WN_{11})\left[i(N_{13}s_\beta +N_{14}c_\beta \mathrm{e}^{i\xi })O_{1\alpha }+N_{13}O_{2\alpha }+N_{14}\mathrm{e}^{i\xi }O_{3\alpha }\right].`$ (37)
Then the interaction Lagrangian for the couplings of the neutral Higgs bosons to a lightest neutralino–pair is cast into a very simple form
$`_{H\chi \chi }={\displaystyle \frac{g}{2}}{\displaystyle \underset{\alpha =1}{\overset{3}{}}}\overline{\chi }\left[(G_\alpha )+i(G_\alpha )\gamma _5\right]\chi H_{4\alpha }.`$ (38)
Here and from now on, we use a simplified notation $`\chi `$ instead of the conventional notation $`\stackrel{~}{\chi }_1^0`$ to denote the lightest neutralino state. We note that a sizable coupling is expected when the LSP has the significant compositions of both the gaugino states and the higgsino states.
Thirdly, since both $`\stackrel{~}{W}^0`$ and $`\stackrel{~}{B}`$ have $`T_3=Q=0`$, these neutral gaugino states do not couple to the $`Z`$ boson. Therefore, the neutral current coupling of the $`Z`$ boson to neutralinos occurs only from the higgsino component of neutralinos. The interaction Lagrangian is given by
$`_{Z\chi \chi }={\displaystyle \frac{g}{4c_W}}\left[|N_{13}|^2|N_{14}|^2\right]\overline{\chi }\gamma ^\mu \gamma _5\chi Z_\mu ,`$ (39)
Note that the $`Z`$ boson couples to only an axial–vector current, reflecting the Majorana property of neutralinos. One the other hand, the coupling of the $`Z`$ boson to fermions are not changed even in the presence of new CP–violating phases.
Since the recoil momenta of the nucleus in the elastic scattering of the neutralinos with fixed nuclei are very small compared to the masses of the exchanged $`Z`$ and neutral Higgs bosons, it is appropriate to have an effective four–Fermi Lagrangian by taking the momentum transfer to be zero. The general form of the four-Fermi effective Lagrangian can be written as
$``$ $`=`$ $`\alpha _{1f}\left(\overline{\chi }\gamma ^\mu \gamma _5\chi \right)\left(\overline{f}\gamma _\mu f\right)+\alpha _{2f}\left(\overline{\chi }\gamma ^\mu \gamma _5\chi \right)\left(\overline{f}\gamma _\mu \gamma _5f\right)+\alpha _{3f}\left(\overline{\chi }\chi \right)\left(\overline{f}f\right)`$ (41)
$`+\alpha _{4f}\left(\overline{\chi }\gamma _5\chi \right)\left(\overline{f}f\right)+\alpha _{5f}\left(\overline{\chi }\chi \right)\left(\overline{f}\gamma _5f\right)+\alpha _{6f}\left(\overline{\chi }\gamma _5\chi \right)\left(\overline{f}f\right).`$
The effective Lagrangian should be summed over fermions and the coefficients $`\alpha _{if}`$ ($`i=1`$ to 6) based on the effective Lagrangian can be obtained by evaluating two Feynman diagramsi in Fig. 1 as
$`\alpha _{1f}=+{\displaystyle \frac{G_F}{\sqrt{2}}}\left[|N_{13}|^2|N_{14}|^2\right](T_3^f2Q_fs_W^2),`$ (42)
$`\alpha _{2f}={\displaystyle \frac{G_F}{\sqrt{2}}}\left[|N_{13}|^2|N_{14}|^2\right]T_3^f,`$ (43)
$`\alpha _{3f}={\displaystyle \frac{gY_f}{2\sqrt{2}}}{\displaystyle \underset{\alpha =1}{\overset{3}{}}}{\displaystyle \frac{(G_\alpha )}{m_{H_{4\alpha }}^2}}\{\begin{array}{cc}O_{3\alpha }& \mathrm{for}u\hfill \\ O_{2\alpha }& \mathrm{for}d\hfill \end{array},`$ (46)
$`\alpha _{4f}={\displaystyle \frac{gY_f}{2\sqrt{2}}}{\displaystyle \underset{\alpha =1}{\overset{3}{}}}{\displaystyle \frac{(G_\alpha )}{m_{H_{4\alpha }}^2}}O_{1\alpha }\{\begin{array}{cc}c_\beta & \mathrm{for}u\hfill \\ s_\beta & \mathrm{for}d\hfill \end{array},`$ (49)
$`\alpha _{5f}=+{\displaystyle \frac{igY_f}{2\sqrt{2}}}{\displaystyle \underset{\alpha =1}{\overset{3}{}}}{\displaystyle \frac{(G_\alpha )}{m_{H_{4\alpha }}^2}}O_{1\alpha }\{\begin{array}{cc}c_\beta & \mathrm{for}u\hfill \\ s_\beta & \mathrm{for}d\hfill \end{array},`$ (52)
$`\alpha _{6f}={\displaystyle \frac{igY_f}{2\sqrt{2}}}{\displaystyle \underset{\alpha =1}{\overset{3}{}}}{\displaystyle \frac{(G_\alpha )}{m_{H_{4\alpha }}^2}}\{\begin{array}{cc}O_{3\alpha }& \mathrm{for}u\hfill \\ O_{2\alpha }& \mathrm{for}d\hfill \end{array}.`$ (55)
In these expressions, $`G_F`$ is the Fermi constant, $`Y_f`$ the Yukawa coupling of the fermion $`f`$, which is $`gm_u/(\sqrt{2}m_Ws_\beta )`$ for $`u`$–type fermions and $`gm_d/(\sqrt{2}m_Wc_\beta )`$ for $`d`$–type fermions, and $`T_3^f`$ the third component of the isospin of the fermion $`f`$. In the limit of vanishing CP-violating phases, these expressions agree with those in and .
Among the six independent terms in the effective Lagrangian, only the terms with the coefficients $`\alpha _{2f}`$ and $`\alpha _{3f}`$ survive in the vanishing momentum–transfer limit. The coefficient $`\alpha _{2f}`$ contains contributions from the $`Z`$ boson exchange while the coefficient $`\alpha _{3f}`$ has contributions from the neutral Higgs–boson exchanges. The spin–dependent contribution from the $`\alpha _{2f}`$ terms contains are not suppressed by the fermion mass and can be large over much of the parameter space. In contrast, the spin–independent contribution from the $`\alpha _{3f}`$ terms is always proportional to fermion masses and relies on the LSP being a well–balanced mixture of gaugino and higgsino states and the size of the scalar–pseudoscalar mixing. It might be naively expected that the small first and second generation fermion masses will give rise to a very small spin–independent cross section. However, the spin–independent cross-section can be enhanced by the effects of coherent scattering in a nucleus and can dominate over the spin–dependent cross-section for heavy nuclei. In the following subject, we present the analytic form of both the spin–dependent and spin–independent elastic cross sections and investigate the physical parameters determining them.
### B Elastic cross sections: spin–dependent versus spin–independent
The elastic scattering cross sections based on $`\alpha _{2,3f}`$ have been conveniently expressed in . The spin–dependent cross-section can be written as
$`\sigma __S={\displaystyle \frac{32}{\pi }}G_F^2m_r^2\mathrm{\Lambda }^2J(J+1),`$ (56)
where $`m_r=m_fm_N/(m_f+m_N)`$ is the reduced neutralino-nucleus mass, $`J`$ is the spin of the nucleus, the value of which is $`4.5`$ for <sup>73</sup>Ge and $`0.5`$ for <sup>19</sup>F, respectively, and the quantity $`\mathrm{\Lambda }`$ is given by
$`\mathrm{\Lambda }={\displaystyle \frac{1}{J}}\left(a_pS_p+a_nS_n\right),`$ (57)
with the coefficients $`a_p`$ and $`a_n`$:
$`a_p={\displaystyle \underset{f=u,d,s}{}}{\displaystyle \frac{\alpha _{2f}}{\sqrt{2}G_F}}\mathrm{\Delta }_f^{(p)},a_n={\displaystyle \underset{f=u,d,s}{}}{\displaystyle \frac{\alpha _{2f}}{\sqrt{2}G_F}}\mathrm{\Delta }_f^{(n)}.`$ (58)
The factors $`\mathrm{\Delta }_f^{(p,n)}`$ depend on the spin content of the nucleus and the values of the factors are taken to be
$`\mathrm{\Delta }_u^{(p)}=+0.77,\mathrm{\Delta }_d^{(p)}=0.38,\mathrm{\Delta }_s^{(p)}=0.09,`$ (59)
$`\mathrm{\Delta }_u^{(n)}=0.38,\mathrm{\Delta }_d^{(n)}=+0.77,\mathrm{\Delta }_s^{(n)}=0.09,`$ (60)
in our analysis. The expectation values $`S_{p,n}`$ are the averaged values of the spin content in the nucleus and therefore are dependent on each target nucleus. We will display results for scattering off of a <sup>73</sup>Ge target and a <sup>19</sup>F for which in the shell model
$`S_p_{\mathrm{Ge}}=0.011,S_n_{\mathrm{Ge}}=0.491,`$ (61)
$`S_p_\mathrm{F}=0.415,S_n_\mathrm{F}=0.047.`$ (62)
For a more detailed information on the these quantities, we refer to the review paper by Jungman, Kamionkowski and Griest .
On the other hand, the spin–independent cross section is written as
$`\sigma __I={\displaystyle \frac{4m_r^2}{\pi }}\left[Zf_p+(AZ)f_n\right]^2,`$ (63)
where $`Z`$ and $`A`$ are the atomic number and the mass number of the nucleus, respectively, and the coefficients $`f_p`$ are given by
$`{\displaystyle \frac{f_p}{m_p}}={\displaystyle \underset{q=u,d,s}{}}f_{Tq}^{(p)}{\displaystyle \frac{\alpha _{3}^{}{}_{q}{}^{}}{m_q}}+{\displaystyle \frac{2}{27}}f_{TG}^{(p)}{\displaystyle \underset{q=c,b,t}{}}{\displaystyle \frac{\alpha _{3}^{}{}_{q}{}^{}}{m_q}},`$ (64)
and $`f_n`$ is given by an expression similar to that for $`f_n`$. The parameters $`f_{Tq}^{(p)}`$ are defined by $`p|m_q\overline{q}q|p=m_pf_{Tq}^{(p)}`$, while $`f_{TG}=1(f_{Tu}+f_{Td}+f_{Ts})`$ . For our numerical analysis, we adopt
$`f_{Tu}^{(p)}=0.019,f_{Td}^{(p)}=0.041,f_{Ts}^{(p)}=0.140,`$ (65)
$`f_{Tu}^{(n)}=0.023,f_{Td}^{(n)}=0.034,f_{Ts}^{(n)}=0.140.`$ (66)
There exist additional contributions due to one–loop Higgs couplings to gluons and so-called twist-2 operators; however the change from a more careful treatment of loop effects for heavy quarks and the inclusion of twist-2 operators is expected to be numerically small .
## IV LSP–nucleus elastic scattering cross sections
### A Independent SUSY parameters
The elastic scattering cross sections depends on a large number of SUSY parameters; more than ten parameters. So, in order to make a realistic analysis, it will be necessary to make a few assumptions which are reasonable for physics point of view.
Firstly, we note that the neutral Higgs mass spectrum depends on the chargino mass $`m_{H^\pm }`$, a SUSY breaking scale $`M_{\mathrm{SUSY}}`$, two trilinear terms $`|A_{t,b}|`$ and their phases $`\mathrm{\Phi }_{A_t,A_b}`$ as well as $`\mathrm{tan}\beta `$, the higgsino mass parameter $`|\mu |`$ and its phase $`\mathrm{\Phi }_\mu `$. Let us assume a universal trilinear parameter in our analysis:
$`|A_t|=|A_b||A|,\mathrm{\Phi }_{A_t}=\mathrm{\Phi }_{A_b}\mathrm{\Phi }_A,`$ (67)
This assumption will not forbid us from finding out the general trend of the Higgs masses and their couplings to fermions and neutralinos. The size of the Higgs boson mixing is determined by the dimensionless parameter $`\eta _{_{CP}}`$. We take in our analysis
$`M_{\mathrm{SUSY}}=0.5\mathrm{TeV},|A|=1.5\mathrm{TeV}.`$ (68)
The charged Higgs mass plays a crucial role in determining the contribution of the spin–independent cross section so that it will be treated as a free parameter along with the parameters $`\{\mathrm{tan}\beta ,|\mu |,\mathrm{\Phi }_\mu \}`$.
Secondly, the neutralino masses and mixing are determined by the SU(2) gaugino mass parameter $`M_2`$, the U(1) gaugino mass $`|M_1|`$ and its phase $`\mathrm{\Phi }_1`$ as well as the parameters $`\{\mathrm{tan}\beta ,|\mu |,\mathrm{\Phi }_\mu \}`$. It will be reasonable to take the gaugino mass unification condition between two gaugino masses so that at least up to the one–loop level one have
$`|M_1|={\displaystyle \frac{5}{3}}t_W^2M_20.5M_2,\mathrm{\Phi }_1=0,`$ (69)
where $`t_W=\mathrm{tan}\theta _W`$. In this case, the lightest neutralino will be Bino–like for $`|\mu |M_2`$ and it will be higgsino–like for $`|\mu |M_1`$. Note that the higgsino parameter $`|\mu |`$ and its phase $`\mathrm{\Phi }_\mu `$ affect both the neutralino mixing and neutral Higgs mixing. For a large Higgs mixing, a large $`|\mu |`$ is preferred. On the contrary, a large neutralino mixing requires a relatively small value of $`|\mu |`$. In this light, $`|\mu |`$ is a crucial SUSY parameter in determining the relative importance of the spin–dependent and spin–independent contributions. We will take $`|\mu |`$ and $`\mathrm{\Phi }_\mu `$ as free parameters.
Thirdly, the spin–independent cross section strongly depends on fermion masses. For the fermion masses, we will use their maximum values coded by the Particle Data Group :
$`m_u=5.0\mathrm{MeV},m_d=9.0\mathrm{MeV},m_s=170\mathrm{MeV},`$ (70)
$`m_c=1.4\mathrm{GeV},m_b=4.4\mathrm{GeV},m_t=174\mathrm{GeV}.`$ (71)
It is certain that there still exist large uncertainties in the values of the fermion masses. Nevertheless, our numerical analysis will be qualitatively reasonable and even quantitatively meaningful with some improved determinations of fermion masses.
Consequently, the independent SUSY parameters which we manipulate in our numerical estimates of the LSP–nucleus scattering cross sections are
$`\mathrm{tan}\beta ,M_2,|\mu |,m_{H^\pm },\mathrm{\Phi }_\mu ,\mathrm{\Phi }_A.`$ (72)
We set $`\mathrm{\Phi }_\mu =0`$ and take two values 3 and 30 for $`\mathrm{tan}\beta `$ which will satisfy the constraints from the Higgs search experiments at LEP and the 2–loop Barr–Zee–type electron and neutron EDMs.
### B Numerical results
We are now ready to show the importance of the CP–violating phases in the LSP detection through the LSP–nucleus elastic scattering. For a systematic analysis, it is useful to understand the dependence of the CP–violating induced phase on the parameters $`|\mu |`$, $`\mathrm{\Phi }_\mu `$ and $`\mathrm{\Phi }_A`$. Incidentally, the induced phase depends only on one combination of two phases $`\mathrm{\Phi }_\mu +\mathrm{\Phi }_A`$. So, for this analysis, we simply introduce $`\mathrm{\Phi }=\mathrm{\Phi }_\mu +\mathrm{\Phi }_A`$ and set the phase to be $`\pi /2`$, which will give (almost) the maximal absolute value of $`\mathrm{sin}\xi `$. In this case, the sine of the induced phase is determined by the higgsino mass parameter $`|\mu |`$ and the charged Higgs mass for which we take into account two values; 250 GeV and 500 GeV. Also, $`\mathrm{sin}\xi `$ in this limit is is given by a simple analytical expression
$`\mathrm{sin}\xi ={\displaystyle \frac{|\stackrel{~}{\lambda }_6|}{m_a^2\mathrm{sin}2\beta }}.`$ (73)
Therefore, $`\mathrm{sin}\xi `$ is always negative and it is expected that the existence of $`\mathrm{sin}2\beta `$ in the denominator forces its absolute value to increase with increasing $`\mathrm{tan}\beta `$.
Figure 1 shows $`|\mathrm{sin}\xi |`$ as a function of the higgsino mass parameter $`|\mu |`$ by taking four different sets of $`\{\mathrm{tan}\beta ,m_{H^\pm }\}`$; $`\{3,250\mathrm{GeV}\}`$ (solid line), $`\{30,250\mathrm{GeV}\}`$ (dashed line), $`\{3,500\mathrm{GeV}\}`$ (dot–dashed line), and $`\{30,500\mathrm{GeV}\}`$ (dotted line). The values of the other SUSY parameters are given in the previous section. It is clear that $`|\mathrm{sin}\xi |`$ increases with increasing $`\mathrm{tan}\beta `$ but decreases with increasing the charged Higgs mass $`m_{H^\pm }`$. It depends very strongly on the higgsino mass parameter $`|\mu |`$; for a small $`\mathrm{tan}\beta `$ the absolute value of $`\mathrm{sin}\xi `$ can be at most as large as 0.1 for a large value of $`|\mu |`$ and for a relatively small value of $`m_{H^\pm }`$, but it can be as large as unity for a large value of $`|\mu |`$ and a small value of $`m_{H^\pm }`$. This is due to the fact that for a small value of $`\mathrm{tan}\beta `$ the scalar top quarks contribute (almost) exclusively, but for a large value of $`\mathrm{tan}\beta `$ the scalar bottom quarks as well as the scalar top quarks contribute to the CP–violating induced phase. In this light, one may expect that the CP–violating induced mass $`\xi `$ can affect the chargino and neutralino sectors significantly for a certain regime of the SUSY parameter space. However, we note that the relative size of the induced phase contribution to the chargino or neutralino mass spectrum is at most as large as $`(m_W/|\mu |)^2\mathrm{sin}2\beta `$ for $`|\mu |`$ much larger than $`M_2`$ so that the enhancement effect in the induced phase by a large $`|\mu |`$ and a large $`\mathrm{tan}\beta `$ is washed out through the strong suppression by the same parameters. Therefore, in most cases there are no significant modifications in the chargino and neutralino mass spectrums.
To begin with, we estimate the spin–dependent cross section. For this case we consider the scattering of neutralinos on fluorine $`{}_{}{}^{19}F`$, for which the spin–dependent contribution typically dominates by a factor of 20 . As shown in Section 2B, the crucial parameters determining the spin–dependent cross section are the coefficients $`\alpha _{2f}`$ and on the whole by the coefficients $`a_p`$ and $`a_n`$. Because there is simply one $`Z`$–exchange contribution to the spin–independent process, all the $`\alpha _{2f}`$ coefficients are proportional to the coupling of the $`Z`$ boson to neutralinos except for their relative factors; $`|N_{13}|^2|N_{14}|^2`$. We find by a comprehensive numerical scan on the relevant SUSY parameters that the CP–violating induced phase hardly changes the values of $`a_p`$ and $`a_n`$. This property is in a sharp contrast with the aspect that the coefficients and the spin–dependent cross section are very sensitive to $`\mathrm{\Phi }_\mu `$, the phase of the higgsino. For the details of the significant dependence on the phase $`\mathrm{\Phi }_\mu `$, we refer to the work by Falk, Ferstl and Olive .
On the other hand, the spin–independent cross section is dominant in much of the SUSY parameter space for scattering off heavy nuclei so that we consider the scattering of neutralinos on <sup>73</sup>Ge. The cross section is determined by the coefficients $`\alpha _{3f}`$ which involve the contributions from the neutral Higgs exchanges and the scalar part of their couplings to neutralinos and fermions. Because the explicit CP violation through radiative corrections to the Higgs sector can lead to a large mixing among scalar Higgs bosons and pseudoscalar Higgs boson, it is naturally expected to exhibit a rather strong dependence of the cross section on the phases such as the phase $`\mathrm{\Phi }_A`$ and the induced phase $`\xi `$, even if the phase $`\mathrm{\Phi }_\mu `$ is taken to be zero in favor of the naturalness conditions at the GUT scale. Nevertheless, the structure of the coupling $`G_\alpha `$ requires the lightest neutralino state to be composed of both gauginos and higgsinos with significant compositions. This means that a large cross section can be obtained when the gaugino mass $`M_2`$ and the higgsino mass parameter $`|\mu |`$ are comparable in size.
Keeping in mind these aspects, we take for our numerical analysis two values of $`\mathrm{tan}\beta `$ (3 and 30), set $`m_{H^\pm }=250\mathrm{GeV}`$ and use the same values for the other SUSY parameters as those given in the previous analyses. Figure 3 shows $`f_p`$The dependence of $`f_n`$ on the phase $`\mathrm{\Phi }`$ and the parameter $`|\mu |`$ is very similar to that of $`|f_p|`$ and furthermore we find that two quantities have a very similar value for the whole scanned space of the SUSY parameters. for five values of the higgsino mass parameter $`|\mu |`$; 200 GeV (solid line), 400 GeV (long dashed line), 600 GeV (dot–dashed line), 800 GeV (dotted line) and 1000 GeV (dashed line). The value of $`\mathrm{tan}\beta `$ is taken to be 3 in the left frame and 30 in the right frame, respectively. We note several interesting aspects from two figures:
* The magnitude of $`f_p`$ (as well as $`f_n`$) is very sensitive to the value of $`|\mu |`$. In most cases except for the region of large values of $`\mathrm{\Phi }`$, it decreases with increasing $`|\mu |`$. The main reason for the suppression is that the LSP is gaugino–dominated for a large value $`|\mu |`$, while an intermediate state of the LSP is required to have a sizable $`f_p`$.
* Comparing the small and large $`\mathrm{tan}\beta `$ cases, one can find that the magnitude of $`f_p`$ increases with increasing $`\mathrm{tan}\beta `$.
* The relative sensitivity of $`f_p`$ to the phase $`\mathrm{\Phi }`$ is very much enhanced for a large value of $`|\mu |`$, in particular for a small value of $`\mathrm{tan}\beta `$. In this case, $`f_p`$ increases by one order of magnitude as $`\mathrm{\Phi }`$ changes from zero to 180 degrees. As a result, one can expect to have a large LSP detection rate even for a large $`|\mu |`$.
* On the contrary, for a large $`\mathrm{tan}\beta `$ and a large $`|\mu |`$, the value of $`f_p`$ tends to be suppressed for non–trivial values around $`\mathrm{\Phi }=90^0`$.
On the whole, it is clear that $`f_p`$ is very sensitive to $`|\mu |`$ and $`\mathrm{tan}\beta `$, and it becomes sensitive to the phase $`\mathrm{\Phi }`$, equivalently, $`\mathrm{\Phi }_A`$ in the present work, for a large $`|\mu |`$. However, the behavior of $`f_p`$ with respect to the phase $`\mathrm{\Phi }`$ strongly depends on the value of $`\mathrm{tan}\beta `$.
From the discussion in the previous paragraph, it is expected that the spin–independent cross section also will show the same behavior with respect to the SUSY parameters and CP–violating phases. With the fact that $`f_p`$ and $`f_n`$ are (almost) similar to each other in size, one can see that the spin–independent cross section is proportional to the square of $`f_p`$ so that the dependence of the cross section is enhanced. Figure 4 shows the dependence of the spin–independent cross section $`\sigma _I(\chi {}_{}{}^{73}\mathrm{Ge}\chi {}_{}{}^{73}\mathrm{Ge})`$ on the phase $`\mathrm{\Phi }`$ with the other SUSY parameters as those in Figure 3. As expected, the behavior of the cross section is similar to that of $`f_p`$. For $`\mathrm{tan}\beta =3`$ and $`|\mu |=1000`$ GeV, the cross section changes by one order of magnitude, depending on the phase, while the cross section tends to be suppressed for non–trivial phases in the case of a large $`\mathrm{tan}\beta `$.
## V Summary and Conclusions
In this paper, we have re–visited the neutralino–nucleus elastic scattering process portion of dark matter in the Universe under a specific SUSY scenario (which has been recently suggested to avoid the severe EDM constraints); in the scenario, the first and second generation sfermions are so heavy that they are decoupled from the low–energy supersymmetric theories, while the third generation sfermions are required to be relatively light not to spoil naturalness. In this case, the complex parameters, in particular, the phase of the trilinear terms and the phase of the higgsino mass parameter, of the third–generation stop and sbottom sectors can lead to a significant mixing between scalar neutral Higgs bosons and a pseudoscalar neutral Higgs boson. As a result, there appears a CP–violating induced phase due to the misalignment between two Higgs doublet fields. We have focussed on the impact of the scalar–pseudoscalar mixing and the induced phase on the LSP–nucleus elastic scattering process.
For the sake of our numerical analysis, we have assumed a universal trilinear parameters $`|A|`$ and $`\mathrm{\Phi }_A`$ and set the phase of the higgsino mass parameter $`\mathrm{\Phi }_\mu `$ to be zero, and then we have varied the charged Higgs boson mass, $`\mathrm{tan}\beta `$, $`|\mu |`$ and the phase $`\mathrm{\Phi }_A`$ while keeping $`|A|=2`$ TeV and $`M_{\mathrm{SUSY}}=0.5`$ TeV, the SUSY breaking scale. To summarize, we have several interesting aspects related with the induced phase and the detection of the neutralino dark matter:
* The induced phase $`\xi `$ itself can be large if $`\mathrm{tan}\beta `$ is large, and $`|\mu |`$ is comparable to $`M_{\mathrm{SUSY}}`$ or larger. However, we have found that in this case its contribution to chargino and neutralino masses is strongly suppressed because the contribution is determined by the combination $`(m_W/|\mu |)^2\mathrm{sin}2\beta `$.
* The coupling of the $`Z`$ boson to neutralinos is affected only through the CP–violating induced phase from the third generation stop and sbottom sectors. We have found that because of the mutually destructive properties mentioned before the CP–violating induced phase hardly changes the spin–dependent cross section.
* The spin–independent part, to which the main contribution comes from the neutral Higgs boson exchanges can undergo a big change with respect to the phase $`\mathrm{\Phi }_A`$ and the induced phase $`\xi `$ as well. The Higgs mass spectrum and the couplings of the neutral Higgs bosons to neutralinos and fermions vary very significantly with respect to the phases as well as the other SUSY parameters. The dimensionful parameter $`f_p`$ as well as $`f_n`$ dictating the size of the spin–independent cross section increases with decreasing $`|\mu |`$ and increasing $`\mathrm{tan}\beta `$, while the sensitivity of the parameter to the phase $`\mathrm{\Phi }`$. For a small value of $`\mathrm{tan}\beta =3`$, the sensitivity of $`f_p`$ and the spin–independent cross section to the phase $`\mathrm{\Phi }`$ can be huge for a large value of $`|\mu |`$. On the contrary, the spin–independent cross section is suppressed for non–trivial values of the phase $`\mathrm{\Phi }`$ for a large value of $`\mathrm{tan}\beta `$.
Even though we have not presented here, we have confirmed the result by Falk, Ferstl and Olive that the spin–dependent and spin–independent cross sections are strongly dependent on the phase $`\mathrm{\Phi }_\mu `$. In some cases, for a broad range of non–zero $`\mathrm{\Phi }_\mu `$, there are cancellations in the cross sections which reduce both the spin–dependent and spin–independent cross sections by more than an order of magnitude. In other cases, there may be enhancements as one varies $`\mathrm{\Phi }_\mu `$.
In general there are several CP–violating phases which can affect many important physics phenomena. In addition to the supersymmetric CP–violating phase $`\mathrm{\Phi }_\mu `$, we have found that even in the scenario of the gaugino mass unification and the decoupling of the first and second sfermion states, the CP–violating scalar–pseudoscalar neutral Higgs boson mixing and the CP–violating induced phase between two Higgs doublets can affect the spin–independent part of the neutralino–nucleus elastic scattering cross section significantly, depending on the values of $`\mathrm{tan}\beta `$, the charged Higgs boson mass, the SUSY breaking scale, and so on. Therefore, we can conclude that in the present situation of no SUSY signatures it is important to clearly understand the impact of all the CP–violating phases, (which are not constrained by low–energy measurements such as the electron and neutron EDMs), on the neutralino–nucleus elastic scattering, one of the most promising dark matter detection mechanism.
### Acknowledgments
The author is grateful to Manuel Drees and Toby Falk for helpful discussions and thanks Francis Halzen and the Physics Department, University of Wisconsin–Madison where the work was initiated. The author wishes to acknowledge the financial support of 1997–sughak program of Korea Research Foundation.
Note added
While finalizing our paper, we became aware of one very recent work by Gondolo and Freese which treats some of the features we have been studying here. In the paper, they have mainly concentrated on the scalar–pseudoscalar mixing and have provided a global parameter scan in estimating the elastic scattering cross section. Instead, our work have studied the effect of the CP–violating induced phase between two Higgs doublets as well as of the scalar–pseudoscalar mixing.
|
no-problem/9908/cond-mat9908319.html
|
ar5iv
|
text
|
# 4f–spin dynamics in 𝐋𝐚_{𝟐-𝐱-𝐲}𝐒𝐫_𝐱𝐍𝐝_𝐲𝐂𝐮𝐎_𝟒
## I Introduction
In the high $`\mathrm{T}_\mathrm{c}`$ superconductors a close interplay between superconductivity and magnetism exists. The parent compounds of the cuprates are antiferromagnetic insulators. Doping with holes or electrons destroys the long range order but antiferromagnetic correlations persist even in the superconducting region. Inelastic incommensurate magnetic peaks in superconducting $`\mathrm{La}_{2\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{CuO}_4`$ (LSCO) show that superconductivity and magnetic correlations coexist. This phenomenon has regained attraction since in 1995 elastic peaks at the same incommensurate positions were found in $`\mathrm{La}_{1.48}\mathrm{Sr}_{0.12}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$. These elastic peaks are interpreted in terms of hole–rich and spin–rich domains in the $`\mathrm{CuO}_2`$–layers, i.e. the well known stripe picture. In LSCO (T–phase) doping with Nd induces a further low temperature phase transition for $`\mathrm{Nd}0.18`$. For x = 0 there is a transition from the low temperature orthorhombic LTO to the less orthorhombic Pccn phase whereas for $`\mathrm{x}\genfrac{}{}{0pt}{}{_>}{^{}}0.1`$ the transition is to the tetragonal LTT phase. In the latter superconductivity is strongly suppressed for certain Sr concentrations. In the LTT phase the tilt of the $`\mathrm{CuO}_6`$–octahedra can serve as pinning potential for the dynamical stripe correlations. Hence, the inelastic peaks become elastic indicating a formation of static anti phase antiferromagnetic domains in the $`\mathrm{CuO}_2`$–planes which are separated by quasi one–dimensional stripes containing the doped charge carriers. Recently, inelastic incommensurate peaks have also been observed in superconducting $`\mathrm{YBa}_2\mathrm{Cu}_3\mathrm{O}_{6.6}`$ giving rise to the question whether stripes are a general feature of cuprate based high $`\mathrm{T}_\mathrm{c}`$ superconductors.
In addition to the investigation of the Cu subsystem, the dynamics of the Nd spins in $`\mathrm{Nd}_{2\mathrm{x}}\mathrm{Ce}_\mathrm{x}\mathrm{CuO}_4`$ (NCCO, $`\mathrm{T}^{^{}}`$–phase) has been examined extensively by several groups. Henggeler et al. performed neutron scattering experiments to examine the spin excitation spectrum at low temperatures. The same group explained the heavy fermion like large $`\gamma `$ coefficient in low temperature specific heat measurements of NCCO by the shift of spectral weight of the Nd modes to lower energies with increasing number of charge carriers. Specific heat measurements show a Schottky anomaly in the parent compound $`\mathrm{Nd}_2\mathrm{CuO}_4`$. This is explained by the presence of Nd–Cu interactions being responsible for the splitting of the $`\mathrm{Nd}^{3+}`$ ground state Kramers doublet as e.g. observed in Raman and neutron scattering experiments. At higher temperatures in neutron scattering experiments on powder samples of NCCO and on a single crystal of $`\mathrm{Nd}_2\mathrm{CuO}_4`$ a quasielastic (QE) Lorentzian is observed with a line width that increases almost linearly with increasing temperature. At lower temperatures the line shape turns into a Gaussian with an almost constant line width. Similar features have been presented for $`\mathrm{NdBa}_2\mathrm{Cu}_3\mathrm{O}_{7\delta }`$. In this compound the QE Gaussian and the Lorentzian coexist.
In the present work we report on inelastic magnetic neutron scattering experiments on $`\mathrm{La}_{2\mathrm{x}\mathrm{y}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ at various temperatures. We have investigated the 4f magnetic response for samples with Sr–concentrations $`0\mathrm{x}0.2`$ and Nd–concentrations $`0.1\mathrm{y}0.6`$ at low energies (typically $`1\mathrm{meV}`$) in order to obtain information about the Cu magnetism in the $`\mathrm{CuO}_2`$–layers via the Nd–Cu interaction. The paper is organized as follows : the next chapter describes the experimental technique. The presentation of our results and their discussion follows in chapter III which is divided into two parts as justified by our experimental findings. In section IV we will give a brief summary.
## II Experimental
We performed temperature dependent studies on $`\mathrm{La}_{1.9}\mathrm{Nd}_{0.1}\mathrm{CuO}_4`$, $`\mathrm{La}_{1.7\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ (x = 0, 0.12, 0.15, 0.2) and $`\mathrm{La}_{1.4\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ (x = 0.1, 0.12, 0.15, 0.18, 0.2) using the time–of–flight (TOF) spectrometers V3 NEAT (HMI Berlin), G6.2 MIBEMOL (LLB Saclay) and IN5 (ILL Grenoble). All spectrometers are located at cold neutron beam lines and use chopper systems for monochromatization which give very sharp and clean resolution functions. The energy chosen for the incident neutrons ranges from $`\mathrm{E}_\mathrm{i}=3.15\mathrm{meV}`$ ($`5.1\mathrm{\AA }`$) to $`\mathrm{E}_\mathrm{i}=1.28\mathrm{meV}`$ ($`8\mathrm{\AA }`$) resulting in energy resolutions between $`\mathrm{\Delta }\mathrm{E}50\mu \mathrm{eV}`$ and $`\mathrm{\Delta }\mathrm{E}20\mu \mathrm{eV}`$ (HWHM), respectively. Additionally, $`\mathrm{La}_{1.45}\mathrm{Sr}_{0.15}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$ was measured at the IN6 (ILL, Grenoble) with $`\mathrm{E}_\mathrm{i}=3.15\mathrm{meV}`$ ($`\mathrm{\Delta }\mathrm{E}45\mu \mathrm{eV}`$). This spectrometer uses Bragg diffraction on single crystals to monochromize the neutron beam. In all cases a Vanadium standard for calibration and an empty can measurement of the Al flat plate for background correction were carried out. For the experiments we used well characterized powder samples of typically m = 20 g and a standard cryostat for cooling. Details of the data analysis are described elsewhere. Since we could not find any Q–dependence of the magnetic signals $`\mathrm{S}(\mathrm{Q},\omega )`$ is averaged over a broad Q–window for all spectra to obtain a better statistics.
## III Results and Discussion
### A 4f spin relaxation due to coupling of phonons and crystal field excitations
In all samples a quasielastic line of Lorentzian shape is observed at high temperatures. This is illustrated in Fig. 1 showing spectra of $`\mathrm{La}_{1.25}\mathrm{Sr}_{0.15}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ for two different temperatures. The line width $`\mathrm{\Gamma }/2(\mathrm{T})`$ of the Lorentzian decreases with decreasing temperature and shows the same temperature dependence in all compounds within experimental error. In Fig. 2 the temperature dependence of the QE line width for samples with different Sr– and Nd–concentrations is plotted. Above about 100 K the line width increases almost linearly with increasing temperature. Below this temperature the slope is drastically reduced although the line width decreases furthermore on lowering the temperature. The residual line width for $`\mathrm{T}0\mathrm{K}`$ is below the resolution limit even in the experiments when an energy resolution of $`20\mu \mathrm{eV}`$ (HWHM) was chosen. Generally, it is hard to detect the QE Lorentzian for $`\mathrm{T}\genfrac{}{}{0pt}{}{_<}{^{}}\mathrm{\hspace{0.17em}20}\mathrm{K}`$. However, when we do not include a QE Lorentzian in our fitting procedure we observe an increase of the elastic intensity in this temperature region. Since the coherent and incoherent elastic scattering intensity should be almost temperature independent for each sample we can conclude that this additional intensity originates from magnetic scattering. If the fit for the QE Lorentzian yielded a value smaller than half of the resolution (HWHM) we fixed the line width at $`\mathrm{\Gamma }/2=0`$ for this temperature.
The QE line width is related to the spin fluctuation frequency via $`\mathrm{\Gamma }/2=\mathrm{}/\tau `$. Thus the decrease of the line width with decreasing temperature is a direct evidence for the lowering of the 4f–spin fluctuation frequency.
The observation of a magnetic QE line in high $`\mathrm{T}_\mathrm{c}`$ superconductors and related materials has been a subject for discussion for over a decade. During this time two attempts to describe the 4f spin relaxation, i.e. the temperature dependence of the QE line width, were suggested :
* via the interaction of 4f spins with conduction electron spins
* by exchange interaction between 4f spins themselves
In the first case the line width should increase linearly with temperature as $`\mathrm{\Gamma }/2\left[\mathrm{N}(ϵ_\mathrm{F})\mathrm{J}_{\mathrm{ex}}\right]^2\mathrm{T}`$ (Korringa law) with $`\mathrm{N}(ϵ_\mathrm{F})`$ the density of electron states at the Fermi energy and $`\mathrm{J}_{\mathrm{ex}}`$ the exchange integral between 4f– and conduction electron spins. Such a temperature dependence was often found in intermetallic systems. In the second case a power law is expected.
From our data we can exclude both cases, since we observe the same temperature dependence of $`\mathrm{\Gamma }/2(\mathrm{T})`$ in samples with ($`\mathrm{x}>0`$) and without charge carriers ($`\mathrm{x}=0`$) which rules out scenario (i). Again, the same temperature dependence of $`\mathrm{\Gamma }/2(\mathrm{T})`$ is obtained for samples with high Nd content (y = 0.6) and low Nd content (y = 0.1). For a small Nd concentration the Nd–Nd exchange interaction should be weak, which rules out scenario (ii). After a detailed analysis of our data we found that the relaxation of the 4f–spins can be well described with a two–phonon Orbach process assuming a coupling between crystal field (CF) excitations and phonons. The Orbach process is sketched in Fig. 3. In this relaxation process a Nd ion which is in state $`|\mathrm{b}>`$ can absorb a phonon and hence is excited to state $`|\mathrm{c}>`$ (or $`|\mathrm{d}>`$). From this state it can fall down to state $`|\mathrm{a}>`$ by emitting a second phonon. The temperature dependence of the QE line width then follows
$$\mathrm{\Gamma }/2(\mathrm{T})=\mathrm{\Gamma }_0/2+\mathrm{c}\mathrm{\Delta }_{\mathrm{CF}}^3/(\mathrm{e}^{\mathrm{\Delta }_{\mathrm{CF}}/\mathrm{T}}1)$$
(1)
where $`\mathrm{\Delta }_{\mathrm{CF}}`$ is the energy of an excited crystal field state and c is a factor which among others considers the coupling of the CF ground state with the excited CF state. In equation (1) only for $`\mathrm{T}\mathrm{\Delta }_{\mathrm{CF}}`$ the line width is proportional to the temperature. We fit this function to the experimental data of many different samples and obtained values around 200 K for $`\mathrm{\Delta }_{\mathrm{CF}}`$. Since the energy of the excited CF state should not differ much between our samples, we took a mean value of $`\mathrm{\Delta }_{\mathrm{CF}}=200\mathrm{K}`$ in the following. Unfortunately, the crystal field scheme for Nd doped LSCO has yet not been evaluated. However, $`\mathrm{\Delta }_{\mathrm{CF}}200\mathrm{K}`$ coincides roughly with the energy of the first excited state ($`\mathrm{\Delta }_{\mathrm{CF}}=173\mathrm{K}`$) in $`\mathrm{Nd}_2\mathrm{CuO}_4`$.
A fit with equation (1) reveals $`\mathrm{c}=1.5410^7\mathrm{meV}/\mathrm{K}^3`$ and $`\mathrm{c}=1.9310^7\mathrm{meV}/\mathrm{K}^3`$ for $`\mathrm{La}_{1.25}\mathrm{Sr}_{0.15}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ and $`\mathrm{La}_{1.5}\mathrm{Sr}_{0.2}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$, respectively. A residual line width of $`\mathrm{\Gamma }_0/2=10\mu \mathrm{eV}`$ has also been taken into account but does not influence the other parameters markedly. The values for c are in the expected range and are in good agreement with the reported value for LSCO probed with $`\mathrm{Er}^{3+}`$ spins.
We now turn to a discussion of the relevance of the Orbach relaxation process in other systems. A comparison of the QE line widths in $`\mathrm{La}_{2\mathrm{x}\mathrm{y}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ with the reported values in $`\mathrm{Nd}_2\mathrm{CuO}_4`$ and $`\mathrm{NdBa}_2\mathrm{Cu}_3\mathrm{O}_{7\delta }`$ shows that the absolute values of $`\mathrm{\Gamma }/2`$ are of the same order of magnitude in all compounds. For $`\mathrm{Nd}_2\mathrm{CuO}_4`$ only a few data points of the QE line width exist. The agreement between the data of Casalta et al. for a single crystal and Loewenhaupt et al. for a powder sample is rather poor. A fit with the above function yields $`\mathrm{\Gamma }_0/2=0.24\mathrm{meV}`$, $`\mathrm{c}=1.4210^7\mathrm{meV}/\mathrm{K}^3`$ and $`\mathrm{\Gamma }_0/2=0.17\mathrm{meV}`$, $`\mathrm{c}=8.510^8\mathrm{meV}/\mathrm{K}^3`$ for the data of Loewenhaupt et al. and Casalta et al., respectively ($`\mathrm{\Delta }_{\mathrm{CF}}`$ was fixed at 175 K). Note that in both cases a large residual line width is obtained from the fit. This broad Lorentzian line is probably masked by the broad Gaussian line. A large value of $`\mathrm{\Gamma }_0/2`$ was directly observed in $`\mathrm{NdBa}_2\mathrm{Cu}_3\mathrm{O}_{7\delta }`$. A fit of the data on $`\mathrm{NdBa}_2\mathrm{Cu}_3\mathrm{O}_6`$ with eq. (1) reveals $`\mathrm{c}=3.3510^8\mathrm{meV}/\mathrm{K}^3`$ ($`\mathrm{\Gamma }_0/2`$ and $`\mathrm{\Delta }_{\mathrm{CF}}`$ fixed at 0.235 meV and 410 K, respectively). Two things are worth mentioning. Firstly, the choice of 410 K for $`\mathrm{\Delta }_{\mathrm{CF}}`$ has the consequence that the increase of the line width is suppressed up to higher temperatures. This is indeed observed in $`\mathrm{NdBa}_2\mathrm{Cu}_3\mathrm{O}_{7\delta }`$. Secondly, c is roughly an order of magnitude smaller than in $`\mathrm{La}_{2\mathrm{x}\mathrm{y}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$. This is in agreement with the observations of Shimizu et al. for an $`\mathrm{Er}^{3+}`$ spin probe in LSCO and YBCO. The broad residual line width in the concentrated systems might be caused by a strongly enhanced Nd–Nd interaction.
A QE Lorentzian was also observed in $`\mathrm{Pb}_2\mathrm{Sr}_2\mathrm{TbCu}_3\mathrm{O}_8`$ and $`\mathrm{Pb}_2\mathrm{Sr}_2\mathrm{Tb}_{0.5}\mathrm{Ca}_{0.5}\mathrm{Cu}_3\mathrm{O}_8`$. In both compounds Tb has a quasi–doublet ground state, i.e. two singlets separated by only a few $`\mu \mathrm{eV}`$. It was found that in $`\mathrm{Pb}_2\mathrm{Sr}_2\mathrm{Tb}_{0.5}\mathrm{Ca}_{0.5}\mathrm{Cu}_3\mathrm{O}_8`$ the temperature dependence of $`\mathrm{\Gamma }(\mathrm{T})`$ obeys a power law $`\mathrm{t}^\nu `$ with $`\nu =2.8[\mathrm{t}=(\mathrm{T}\mathrm{T}_\mathrm{N})/\mathrm{T}_\mathrm{N}]`$. Conclusively the authors claimed that the rare–earth exchange interaction might be the dominant process for the 4f–spin relaxation. The same group obtained similar results for $`\mathrm{Y}_{0.9}\mathrm{Tb}_{0.1}\mathrm{Ba}_2\mathrm{Cu}_3\mathrm{O}_7`$. Since the concentration of Tb in this compound is very low the above interpretation of a strong rare–earth exchange interaction seems rather unlikely. In contrast, a very recent reanalysis of the data showed that the temperature dependence of the QE line width also follows the two–phonon Orbach process. Taking all these facts into account it seems reasonable to assume that the 4f spin relaxation in high $`\mathrm{T}_\mathrm{c}`$ superconductors and related materials is caused by CF transitions assisted by phonons and not by interaction with conduction electrons. Moreover, the interpretation of the deviation from a linear temperature dependence of $`\mathrm{\Gamma }/2(\mathrm{T})`$ as opening of a gap has to be reexamined.
To conclude this section we want to mention that a coupling between these two elementary excitations was already reported in Raman scattering studies on several high $`\mathrm{T}_\mathrm{c}`$ superconductors and related compounds. Furthermore, ESR data have been discussed in terms of an Orbach process. These data are consistent with our interpretation of the temperature dependence of the QE line width (see also our previous work ).
### B Magnetic response due to Nd–Cu interaction
In contrast to the above described behavior the 4f magnetic response at low temperatures correlates with the electronic properties of the $`\mathrm{CuO}_2`$–layers. Depending on the dopant concentration we find different features of the magnetic response which we will now discuss in detail.
x = 0
In insulating $`\mathrm{La}_{2\mathrm{y}}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ with y = 0.1, 0.3 the 4f magnetic response changes from a QE Lorentzian to an inelastic (INE) excitation below about 80 K (see Fig. 4). This INE excitation clearly indicates the splitting of the $`\mathrm{Nd}^{3+}`$ ground state Kramers doublet due to the internal exchange field of ordered Cu moments and shows the strong interaction between the Cu– and Nd–subsystems. Although the Cu ordering temperature is much higher, this excitation becomes first detectable below 80 K since at higher temperature the line width is larger than the observed energy splitting. We mention that this is in contrast to neutron scattering results on $`\mathrm{Nd}_2\mathrm{CuO}_4`$ where the magnetic signal remains QE down to 5 K possibly due to Nd–Nd interactions. For the sample with smaller Nd content, namely $`\mathrm{La}_{1.9}\mathrm{Nd}_{0.1}\mathrm{CuO}_4`$, we find a similar behavior as for y = 0.3. In contrast to y = 0.3 in this sample only a minority fraction of about 20 % changes to the Pccn phase.
The temperature dependence of the energy splitting for both compounds is plotted in Fig. 5. An increase of the splitting, i.e. of the internal exchange field at the Nd site, is clearly visible. Our findings agree roughly with the data of Chou et al. who measured the internal field at the La site in $`\mathrm{La}_2\mathrm{CuO}_4`$ with $`{}_{}{}^{139}\mathrm{La}`$ NQR. According to their results the temperature dependence of the internal field can be described with a power law $`(1\mathrm{T}/\mathrm{T}_\mathrm{N})^\beta `$ ($`\mathrm{T}_\mathrm{N}=250\mathrm{K}`$ and $`\beta =0.41`$). This means an increase of about 17% from 80 K to 3 K and agrees roughly with our results in $`\mathrm{La}_{1.9}\mathrm{Nd}_{0.1}\mathrm{CuO}_4`$. It is obvious that the exchange field at the Nd site is influenced by both, the staggered magnetization and the direction of the Cu spins. In $`\mathrm{La}_{1.7}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ the structural transition from $`\mathrm{LTO}\mathrm{Pccn}`$ is accompanied by a Cu spin reorientation. Thus, the enhanced splitting in y = 0.3 compared to y = 0.1 for $`\mathrm{T}0\mathrm{K}`$ might be due to the difference in the direction of the Cu spins. Finally, we want to mention that the value of $`\mathrm{\Delta }\mathrm{E}`$ coincides with that derived from the Schottky anomaly found in low temperature specific heat measurements.
x = 0.12
Static ordering of charges and spins was first reported in $`\mathrm{La}_{1.48}\mathrm{Sr}_{0.12}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$. Similar results are obtained for x = 0.15 where magnetic order has also been observed with $`\mu ^+\mathrm{SR}`$–experiments and Mössbauer experiments.
We performed measurements on several compounds related to this composition and found similar properties at low temperatures. As a representative sample we chose $`\mathrm{La}_{1.28}\mathrm{Sr}_{0.12}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ to explain the general features that are observed when the type of ordering changes from the well known spin structure for x = 0 into an antiferromagnetic stripe pattern of ordered spins and charges in the $`\mathrm{CuO}_2`$–layers. Above the antiferromagnetic ordering temperature ($`\mathrm{T}_\mathrm{N}30\mathrm{K}`$) a single QE Lorentzian line is found as discussed in section A. When the temperature is further lowered below $`\mathrm{T}_\mathrm{N}`$ the line width does not become smaller as expected from the two–phonon Orbach process. Instead, a broad magnetic response of almost constant width centered around the elastic peak is visible (see Fig. 6). The analysis of our results shows that in addition to the Lorentzian a Gaussian line is necessary to accurately describe the data. The width of this Gaussian is almost temperature independent ($`\mathrm{\Gamma }_{\mathrm{Gaussian}}/20.13\mathrm{meV}`$, see Fig. 7).
The observation of a broad QE Gaussian line instead of a well resolved INE excitation as for x = 0 infers a distribution of different energy splittings on different Nd sites. This is a direct hint on spatial inhomogeneities in the $`\mathrm{CuO}_2`$–planes and is probably caused by the formation of stripes. In this picture the Kramers doublet of a $`\mathrm{Nd}^{3+}`$ ion, which is located ’near’ a charge stripe and thus is not split up, contributes to the Lorentzian signal which is still observable (at least the splitting must be small compared to the width of this Lorentzian). The comparison of the width of the Gaussian line in $`\mathrm{x}=0.12`$ with the energy excitation of the inelastic line (x = 0) reveals a reduced (average) splitting, which is related to a reduced zero temperature staggered magnetization in the $`\mathrm{CuO}_2`$–planes. This finding coincides with the $`\mu ^+\mathrm{SR}`$–experiments on $`\mathrm{La}_{1.85\mathrm{y}}\mathrm{Sr}_{0.15}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ where a decrease of the muon spin rotation frequency compared to $`\mathrm{La}_{1.7}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ was found. This was interpreted as a decrease of the average magnetic field at the muon site in Sr doped compounds.
In Fig. 8 the intensities of the Lorentzian and Gaussian signals are plotted versus temperature. Above 50 K the intensity of the single Lorentzian line raises with decreasing temperature due to an increase in the thermal occupation of the ground state (not shown). Below this temperature firstly the intensity of the Lorentzian decreases linearly for $`\mathrm{T}\genfrac{}{}{0pt}{}{_>}{^{}}\mathrm{\hspace{0.17em}10}\mathrm{K}`$ whereas the intensity of the Gaussian raises. This can be interpreted as a decrease in the number of paramagnetic Nd ions (i.e. with a splitting smaller than the Lorentzian line width). This is expected since the Lorentzian line width decreases with decreasing temperature and therefore the number of Nd ions for which the above condition holds reduces. Secondly, between 10 K and 3 K, a plateau–like level is reached. For $`\mathrm{T}\genfrac{}{}{0pt}{}{_<}{^{}}\mathrm{\hspace{0.17em}10}\mathrm{K}`$ the Lorentzian line width is well below the resolution limit and so small that we cannot distinguish between a Lorentzian line and the elastic line. Finally, below 3 K a strong increase of the Lorentzian intensity (which we cannot distinguish from an increase of the elastic intensity) is observed. At a similar temperature a pronounced increase of the magnetic intensity due to ordering of the Nd moments was reported in $`\mathrm{La}_{1.48}\mathrm{Sr}_{0.12}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$.
Sr dependence of the magnetic response
Besides the above mentioned two Sr–concentrations x = 0 (with y = 0.1, 0.3) and x = 0.12 (with y = 0.6) experiments with varying Sr compositions were carried out on $`\mathrm{La}_{1.7\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ with x = 0.12, 0.15, 0.2 and $`\mathrm{La}_{1.4\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ with x = 0.1, 0.15, 0.18, 0.2. For clarity the data of the series with y = 0.3 are not shown since the observations in these compounds are similar to the findings in the related samples with y = 0.6.
To study the Sr dependence of the magnetic signal at low temperatures in detail we extended our measurements to samples with a Nd content of y = 0.6 with $`0.1\mathrm{x}0.2`$. In contrast to $`\mathrm{La}_{1.7\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ all of these samples lie in the region of the phase diagram where superconductivity is strongly suppressed and thus a broad magnetic response at low temperatures is expected. Such a response was found in all samples except for $`\mathrm{La}_{1.2}\mathrm{Sr}_{0.2}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$. The line widths of the QE Gaussians are plotted in Fig. 9 (these data points are fits under the assumption that $`\mathrm{\Gamma }_{\mathrm{Gaussian}}/2`$ is temperature independent for each compound). The decrease of the line width with increasing Sr content is related to a reduction of the average staggered magnetization in the $`\mathrm{CuO}_2`$–planes. This observation is consistent with the findings in $`\mathrm{La}_{1.6\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$ with x = 0.12, 0.15 and 0.2.
Unfortunately the spectra of $`\mathrm{La}_{1.3}\mathrm{Sr}_{0.1}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ showed strongly enhanced background probably due the diffusion of air into the neutron flight path. Nevertheless, the analysis showed that a QE Gaussian is not consistent with the data in the whole temperature range 1.8 K – 30 K although it is accurate enough at certain temperatures. A better agreement is obtained when we use an INE Gaussian line. This might reflect a mixture of both types of signals which we observe in $`0<\mathrm{x}=0.1<0.12,\mathrm{\hspace{0.17em}0.15},\mathrm{\hspace{0.17em}0.18}`$, i.e. an INE excitation and QE Gaussian, respectively. Furthermore we note that this compound is closest to the Pccn phase.
We could not detect any QE broadening in $`\mathrm{La}_{1.8\mathrm{y}}\mathrm{Sr}_{0.2}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ with y = 0.3 and 0.6 at lowest temperature (see Fig. 10). For y = 0.3 this behavior is expected since this compound is a bulk superconductor below $`\mathrm{T}_\mathrm{c}25\mathrm{K}`$ and hence no magnetic order in the $`\mathrm{CuO}_2`$–planes is expected. Indeed, in a superconductor with even higher Nd–concentration ($`\mathrm{La}_{1.4}\mathrm{Sr}_{0.2}\mathrm{Nd}_{0.4}\mathrm{CuO}_4`$) Nachumi et al. could not find any hints for magnetic order in a recent $`\mu \mathrm{SR}`$–experiment. In contrast to this, the absence of a QE broadening is surprising in the compound with y = 0.6 because superconductivity is strongly suppressed and hence a broad magnetic response is expected. Our sample has also been studied in a recent $`\mu \mathrm{SR}^+`$–experiment which shows magnetic order below about 15 K. This finding seems to contradict our neutron scattering results. There might be two reasons why a broadening is not observable in the present neutron scattering experiment : (i) the splitting of the Nd Kramers ground state is too small or (ii) the intensity of the QE Gaussian is too low. From the energy resolution chosen in our experiment we can conclude that a possible average splitting of the Nd ground state must be below $`20\mu \mathrm{eV}`$ if case (i) were true. This contradicts the value which is obtained by extrapolating the Gaussian widths of Fig. 9 to x = 0.2. Indeed, scenario (ii) is more likely since we observe a drastic drop of the intensity of the QE Gaussian in y = 0.18 compared to y = 0.15 at lowest temperature. At present we have no interpretation for this drop of the intensity. Further combined neutron and $`\mu \mathrm{SR}`$ studies are necessary in order to investigate and understand this pronounced concentration dependence of the intensity.
## IV Summary
To summarize, we have presented inelastic magnetic neutron scattering experiments on Nd doped $`\mathrm{La}_{2\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{CuO}_4`$. In all samples at higher temperatures a quasielastic line of Lorentzian shape is observed with a line width which decreases with decreasing temperature. The temperature dependence of this width, i.e. the relaxation of the Nd 4f–moments is dominated by the Orbach relaxation process via the coupling of phonons and CF excitations and does not depend on the charge carrier concentration in the $`\mathrm{CuO}_2`$–planes. The low temperature behavior of the magnetic response clearly correlates with the electronic properties of the $`\mathrm{CuO}_2`$–layers. In the undoped samples (x = 0) below about 80 K an inelastic excitation occurs which shows the splitting of the $`\mathrm{Nd}^{3+}`$ Kramers doublet ground state due to the Cu exchange field at the Nd site. In $`\mathrm{La}_{1.7\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.3}\mathrm{CuO}_4`$ with x = 0.12, 0.15 and $`\mathrm{La}_{1.4\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{Nd}_{0.6}\mathrm{CuO}_4`$ with x = 0.1, 0.12, 0.15, 0.18 superconductivity is strongly suppressed. In all these compounds we observe an additional quasielastic Gaussian below about 30 K. The width of this Gaussian is almost temperature independent and decreases with increasing Sr concentration. The observation of a Gaussian line infers a distribution of various Cu exchange fields on different Nd sites and is interpreted in terms of the stripe model. In $`\mathrm{La}_{1.8\mathrm{y}}\mathrm{Sr}_{0.2}\mathrm{Nd}_\mathrm{y}\mathrm{CuO}_4`$ (y = 0.3, 0.6) no indication for a Nd–Cu interaction has been found, i.e. a single quasielastic Lorentzian is observed.
## V Acknowledgment
We gratefully acknowledge useful discussions with V. Kataev and M.M. Abd–Elmeguid. Our work was supported by the ’BMBF’ under contract number 03-HO4KOE.
|
no-problem/9908/cond-mat9908087.html
|
ar5iv
|
text
|
# SPONTANEOUS VS. PIEZOELECTRIC POLARIZATION IN III-V NITRIDES: CONCEPTUAL ASPECTS AND PRACTICAL CONSEQUENCES
## I INTRODUCTION
Macroscopic polarization in low-symmetry crystals is well known in the ferroelectrics community. Outside that community, it was so far held for a curiosity devoid of practical, let alone technological, relevance. In particular, it was known that a large spontaneous polarization (henceforth SP), at most an order of magnitude smaller than the giant values of ferroelectric perovskites, was present in some wurtzite semiconductors, but due to their minor technological relevance compared to zincblende III-V arsenides or phosphides, this issue was largely overlooked outside a limited circle of specialists.
Such state of affairs has changed after first-principles calculations provided evidence of the considerable SP and larger-then-usual piezoelectric coupling constants of wurtzite III-V nitrides. Because of the huge importance of the latter as materials for optoelectronic devices in the blue–UV range , the issue of spontaneous polarization fields was brought to a larger community.
The obvious relevance of polarization in nitrides is due to the fact that it induces large ($``$MV/cm) built-in electrostatic fields in layered nitride nanostructures, as first shown in Ref. . These fields affect dramatically the optical and electrical properties of those structures \[5-14\]. Quite a flurry of activity has recently revealed a host of new and unexpected effects: to name but some, Stark-like red shifts in recombination and absorption energies for increasing quantum well widths , concurrent suppression of oscillator strength , ensuing anomalies in recombination dynamics , and their interplay with charge injection driving a recovery of oscillator strength and a blue back-shift of transition energies, “self-doping” effects at HEMT heterointerfaces , have been demonstrated theoretically and experimentally.
In view of these exciting developments, it is likely that most of the potentially observable polarization-related effects in nitrides are still to be dreamt about. It is safe to state, in any event, that conceptual clarity on the determination of the electric fields produced by polarization, and in particular by SP, is a must if we are to understand and exploit these effects in practical applications. To mention just a pair of issues, the built-in fields depend on boundary conditions and on device design (number of quantum wells, thickness, composition, contacts, etc.); also, the field values and sign patterns are not those expected if only piezoelectricity is considered. This work therefore aims at clarifying some of these subtle questions, focusing on three inter-related systems, i.e. large samples, multi-quantum-wells (MQWs), and isolated quantum-wells (QWs).
## II SPONTANEOUS VS. PIEZOELECTRIC
Spontaneous polarization and piezoelectric constants can be reliably determined via simple electronic structure calculations based on the Berry geometric quantum phase concept. We start by discussing our recent result thereon , which are summarized in Figs. 1 and 2. The values calculated ab initio are for III-N binaries (AlN, GaN, InN) only; since direct calculations for the alloys are not yet available, an estimate of the SP $`𝐏^{(\mathrm{sp})}`$ in Al<sub>x</sub>In<sub>y</sub>Ga<sub>(1-x-y)</sub>N alloys was obtained by Vegard interpolation:
$`𝐏^{(\mathrm{sp})}(\mathrm{Al}_\mathrm{x}\mathrm{In}_\mathrm{y}\mathrm{Ga}_{(1\mathrm{x}\mathrm{y})}\mathrm{N})`$ $`=`$ $`x𝐏_{\mathrm{AlN}}^{(\mathrm{sp})}+y𝐏_{\mathrm{InN}}^{(\mathrm{sp})}`$ (1)
$`+`$ $`(1xy)𝐏_{\mathrm{GaN}}^{(\mathrm{sp})},`$ (2)
similar relations holding for the piezoelectric constants. Thus, the triangle in Fig. 1 borders the (theoretical) SP values achievable in a general III-N alloy as a function of in-plane lattice constant. The triangles in Fig. 2 border the accessible values of the two piezoconstants relevant to $`a`$-plane epitaxial strains. The large range of accessible polarizations and lattice constants is due to the concurrent large lattice mismatch between GaN and InN on one side, and the very large SP in AlN on the other.
Three main points are evident from the data reported. First, the piezoelectric constants are an order of magnitude larger than in other III-V’s, and reversed in sign. Second, for basal-plane strains typical of epitaxial wurtzite nitride multilayers (2-5%), SP is comparable to, or larger than piezoelectric polarization. Therefore, SP is all but negligible in any typical III-V nitride nanostructures. Third, GaN has nearly the same SP as InN, but a large lattice mismatch to it; on the other hand, AlN has a SP about three times larger than GaN, but a much smaller mismatch. This implies that InGaN/GaN structures will tend to be mostly influenced by piezoelectricity, whereas in AlGaN/GaN systems, SP effects will be dominant. Mixed regimes can also occur, which we will come back to in Sec. V.
## III MASSIVE SAMPLES
By elementary electrostatics, polarization and electrostatic field are related by
$`𝐄=(𝐃𝐏)/\epsilon `$ (3)
where $`𝐃`$ is the displacement field, $`𝐏`$ the macroscopic polarization, and $`\epsilon `$ the static dielectric constant. The latter have been calculated to be 10.31, 10.28, and 14.61 for AlN, GaN, and InN respectively; the value for generic alloys are estimated by Vegard interpolation. The polarization in Eq. 3 is known as transverse, and it is the sum of the spontaneous and piezoelectric polarization contributions: $`𝐏=𝐏^{(\mathrm{sp})}+𝐏^{(\mathrm{pz})}`$. The displacement field is determined by the free-charge distribution in the material:
$`𝐃=e(pn),`$ (4)
with $`e`$ the electron charge, and $`p`$ and $`n`$ the hole and electron densities, respectively. The evaluation of the electric field $`𝐄`$ inside the semiconductor requires in general a self-consistent solution of Eq. 3 and 4 as outlined in Refs. and . However, many cases can be discussed without explicit calculations, an useful exercise especially in the low free-carrier–density limit.
An interesting such case is the evaluation of polarization effects in large, massive samples. In effect, misunderstanding the role of the displacement field in massive samples leads to predicting wrong electrostatic fields in a MQWs system. At the outset, let it be pointed out that a strong polarization field in a III-V nitride massive sample does not imply the existence of a macroscopic electrostatic field over the whole sample. Qualitatively, since such an electric field would produce (at non-zero temperature) a persistent current due to thermally generated intrinsic carriers, and because this current cannot be sustained indefinitely, the electric field inside a massive sample must be zero. One deduces thereforth that the displacement field in the sample is
$`𝐃=𝐏^{(\mathrm{sp})}.`$ (5)
For a neutral (albeit polarized) sample, displacement field conservation across the surface would imply an electric field $`𝐄=𝐏^{(\mathrm{sp})}`$ in the vacuum off the sample surface. Of course this is unrealistic, as outside the sample $`𝐏`$, $`𝐄`$, and $`𝐃`$ must all vanish. The SP discontinuity across the surface must be counterbalanced by an equal change in the displacement field. Thus, near the surface of our sample, there exists a region where the free-carriers distribution satisfies
$`{\displaystyle _{\mathrm{}}^0}\left(e(pn)𝐏^{(\mathrm{sp})}\right)𝑑z=0,`$ (6)
with $`\mathrm{}`$ the thickness of the surface region, which may range easily in the hundreds of Å. In this region the free carriers will constitute an approximately two-dimensional electron gas (2DEG) whose areal density equals the SP. The electric field in this region depends strongly on material parameters and on surface states. It can be concluded however that at distances in excess of $`\mathrm{}`$ from the surface (or, in the case of devices, the interface with contacts, buffers, caps, etc.), the electric field will be zero, and the polarization will equal $`𝐏^{(\mathrm{sp})}`$, as will the displacement field. This is the starting point to discuss MQWs.
## IV MULTI-QUANTUM-WELLS
As pointed out above (especially in a semi-insulating sample) it can be safely assumed that the displacement field in a large sample is uniform all over the crystal, and equal to $`𝐏^{(\mathrm{sp})}`$, and $`𝐄=0`$. If a sequence of QWs, of fixed composition, is now inserted in the (otherwise homogeneous) sample, the electric field inside the generic $`j`$-th layer (either QW or barrier) is given by:
$`𝐄_j=(𝐃𝐏_j)/\epsilon _j`$ (7)
with $`𝐏_j`$ and $`\epsilon _j`$ the total polarization and dielectric constant in layer $`j`$.
If the displacement field is constant, i.e. if Eq. 5 holds, the electric field in the $`j`$-th layer is
$`𝐄_j=(𝐏^{(\mathrm{sp})}𝐏_j)/\epsilon _j,`$ (8)
where $`𝐏^{(\mathrm{sp})}`$ is the SP in the massive sample material, which now functions locally as barrier material. The field is then zero in the barriers, which are made up of unstrained material with spontaneous polarization $`𝐏^{(\mathrm{sp})}`$, while each QW in the series is subject to the same electric field $`𝐄_j`$. A potential drop given by
$`\mathrm{\Delta }V_j=|𝐄_j|l_j`$ (9)
will thus occur in the generic $`j`$-th well of thickness $`l_j`$. Now Eq. 5 holds only subject to the following four constraints:
i) the Debye-Hückel screening length is at least larger than, say, the MQW region thickness;
ii) the QWs are far enough from any interface with the outer world that the surface effects mentioned in the previous Section are negligible;
iii) none of the $`\mathrm{\Delta }V_i`$, nor their sum, must exceed the band gap energy of the well material:
$`{\displaystyle \underset{k}{}}l_k𝐄_k<E_{\mathrm{gap}};`$ (10)
iv) the well-barrier interfaces are free of interface states.
These requirements are typically met in nitride MQWs, except for condition (iii), which breaks down if the QWs form a sufficiently thick superlattice. Thereby, a self-consistent calculation explicitly accounting for free carriers must be performed. To approximately enforce condition $`(iii)`$ avoiding cumbersome calculations, it is convenient to approximate Eq. 10 applying periodic boundary conditions,
$`{\displaystyle \underset{k}{}}l_k𝐄_k=0,`$ (11)
where the sum runs over all layers in the MQW, in particular including the barrier layers. Thereby the displacement field is determined subject to the condition of zero average electric field in the MQW region. The maximal error in the predicted fields due to using Eq. 11 instead of Eq. 10 is of the order of $`E_{\mathrm{gap}}/d`$, with $`d`$ the overall thickness of the superlattice; hence this approximation becomes exact in the limit of very thick superlattices. The error may be of course smaller depending on the specific conditions in the device (doping, contacts, etc.).
Substituting Eq. 7 in Eq. 11, the displacement field is then obtained:
$`𝐃={\displaystyle \frac{_kl_k𝐏_k/\epsilon _k}{_kl_k/\epsilon _k}}.`$ (12)
Plugging this back into Eq. 7 gives a simple expression for the field in the generic well or barrier of the MQW, namely
$`𝐄_j={\displaystyle \frac{_kl_k𝐏_k/\epsilon _k𝐏_j_kl_k/\epsilon _k}{\epsilon _j_kl_k/\epsilon _k}},`$ (13)
with sums running on all layers (including the $`j`$-th). This is a general expression for any thickness of wells and barriers in a generic superlattice or MQW, which reduces to the formulas in Ref. assuming a single type of well and barrier.
It is to be noted again that in this case the fields are non-zero both in the wells and the barriers. Typically, the sign of the barrier field will be opposite to that of the well, leading to an effectively triangular confinement potential, which favors transfer of confined electrons into the barrier layer on one side of the well, which may lead to anomalous effects in excitonic transitions .
In practice, if the conditions discussed above are satisfied for a specific system, the properties of the latter can be calculated by simply solving a Schrödinger equation for the compositional potential of the superlattice with the built-in fields as given by Eq. 13. Of course, this is emphatically not the case in general, because of e.g. doping, excitation, boundary conditions, etc. A self-consistent solution as outlined e.g. in Refs. and will be needed.
## V ISOLATED QUANTUM WELLS
Consider now a single QW of material A; let the latter have dielectric constant $`\epsilon _\mathrm{A}`$ and spontaneous polarization $`𝐏_\mathrm{A}^{(\mathrm{sp})}`$, and the QW made thereof be strained so as to be subject to a piezoelectric polarization $`𝐏^{(\mathrm{pz})}`$. The total polarization in the QW is then $`𝐏_\mathrm{A}^{(\mathrm{sp})}+𝐏^{(\mathrm{pz})}`$. The well is embedded in an extended, insulating, and unstrained material with spontaneous polarization $`𝐏^{(\mathrm{sp})}`$. According to Eq. 7, the field in the QW is
$`𝐄`$ $`=`$ $`[𝐏^{(\mathrm{sp})}𝐏_\mathrm{A}^{(\mathrm{sp})}]/\epsilon _\mathrm{A}𝐏^{(\mathrm{pz})}/\epsilon _\mathrm{A}`$ (14)
$`=`$ $`𝐄_\mathrm{A}^{(\mathrm{sp})}+𝐄_\mathrm{A}^{(\mathrm{pz})}.`$ (15)
At variance with conventional III-V’s (having no SP), an additional term $`𝐄_\mathrm{A}^{(\mathrm{sp})}`$ appears, due solely to the difference in the SP between the QW active layer and the barrier material . This additional term has far-reaching consequences, and can be used to tune the value of the electric field inside the QW. In Fig. 3 we depict the value of the electric field $`𝐄`$ and of its purely piezoelectric component $`𝐄^{(\mathrm{pz})}`$ vs in-plane strain for an AlInGaN QW embedded in GaN. Without SP, the values of the electric field inside the QW would fall into the hatched region delimited by dashed lines. The inclusion of SP gives rise to a different, wider accessible region, especially for small strains. The region in question is white and delimited by a continuous line. On the negative strain side of Fig. 3, this region barely touches the upper side of the piezoelectric region, and shrinks becoming a line for large strains. The various region boundaries are curved due to the quadratic dependence of the piezoelectric component on alloy composition (through \[piezoconstants\]$`\times `$\[strain\] terms).
Several features are worth a mention. First, the two regions do not overlap, i.e. for any composition the total field differs from its pure piezoelectric component (this of course is due to the fact that $`𝐄^{(\mathrm{sp})}`$ never vanishes). This difference is rather dramatic for positive strains, i.e. in Al-rich alloy wells. A noticeable exception to this general behavior is that of GaN/InGaN QWs. These correspond to the region where the total and piezo field regions in Fig. 3 (almost) touch each other: in that case indeed, as InN has nearly the same SP as GaN, the SP difference between barrier and active layer, and hence $`𝐄^{(\mathrm{sp})}`$, is negligible with respect to the piezoelectric term. That is why early investigations on GaN/InGaN MQWs yielded results in good agreement with theory even though SP was neglected. Second, without SP the electric field is zero at zero strain (the typical situation in zincblende III-Vs). This is no more the case in III-V nitrides, where SP is an additional degree of freedom potentially producing fields up to 4 MV/cm at zero strain. This requires growing AlInGaN with appropriate compositions, which appears to be difficult because of thermodynamic solubility constraints; however, fields of several hundreds of KV/cm can be achieved already at minute Al and In concentrations, which may perhaps become accessible in the future. Third, last but not least, SP provides a handle for reducing or even, in principle, zeroing the field in strained QWs . Specifically, this occurs on the zero field line in Fig. 3. This can be a breakthrough for the many applications needing a good confinement, but at the same time no built-in field in the active region. The same considerations on growth hold as in the previous point.
## VI Summary and acknowledgements
In conclusion, SP can be considered as a degree of freedom to tune the value of the polarization-induced electric fields in QW systems. It was shown, for a GaN/AlInGaN QW, that fields up to 4MV/cm can be obtained in absence of strain, and that, conversely, a vanishing field can also be obtained, despite lattice mismatch, for judicious choices of composition and strain.
Support from the PAISS program of INFM is acknowledged. VF thanks the Alexander von Humboldt-Stiftung for supporting his stay at the Walter Schottky Institut.
|
no-problem/9908/cond-mat9908135.html
|
ar5iv
|
text
|
# Non-Kondo zero bias anomaly in electronic transport through an ultra-small Si quantum dot
\[
## Abstract
We have studied low-temperature single electron transport through ultra-small Si quantum dots. We find that at low temperatures Coulomb blockade is partially lifted at certain gate voltages. Furthermore, we observed an enhancement of differential conductance at zero bias. The magnetic field dependence of this zero bias anomaly is very different from the one reported in GaAs quantum dots, inconsistent with predictions for the Kondo effect.
preprint: Submitted to Phys. Rev. B, Rapid Communications
\]
Quantum dots (QDs) formed in GaAs/AlGaAs heterostructures have been used as model systems to study transport in the Coulomb blockade regime. Following advances in nanolithography QD size was reduced down to the level where quantum effects start to play a role. As the size become smaller collective phenomena, such as the Kondo effect, were recently reported. To further reduce the QD size for exploring new phenomena, one must abandon the conventional approach of using field–induced barriers. Recently, Si dots with confinement provided by the sharp Si-SiO<sub>2</sub> interface have been realized. These dots can be fabricated so small that the single-electron transistor (SET) can operate at room temperature. Although an increase of the operating temperature of SETs was the primary driving force behind the development of the Si quantum dot technology, there were some limited studies of electron transport at low temperatures, which provided information about energy spectrum in these structures and probed first-order quantum corrections to the conductivity in the Coulomb blockade regime.
In this paper we report low temperature electron transport in ultra–small Si quantum dots. We found that at certain range of gate voltages Coulomb blockade is lifted at $`T<1`$ K and differential conductance at zero source-drain bias increases as the temperature is lowered. Although it is appealing to attribute the enhancement to the Kondo effect, we find that the magnetic field dependence of this zero bias anomaly is inconsistent with such an interpretation.
We have investigated transport in quantum dot samples which are metal–oxide–semiconductor field effect transistors (MOSFETs) with a Si dot connected to the source and drain leads through tunneling barriers. The dot is surrounded by 40-50 nm of SiO<sub>2</sub> and wrapped by a poly-Si gate (fabrication details can be found in Ref. ). The gate is also extended over the tunneling barriers and parts of the leads, adjacent to the dot. Outside the gate, the source and drain are $`n`$-type. An inversion layer is formed at the Si-SiO<sub>2</sub> interface by applying a voltage to the poly-Si gate. Unlike GaAs dots, there are no separate gates to control the coupling between the dot and the source/drain. In fact, the coupling is a function of the applied gate voltage $`V_g`$. We studied more than 30 samples which show Coulomb blockade above 10 K. However, at low temperatures ($`T<4`$ K) and low source–drain bias ($`V_b<100`$ $`\mu `$V) the conducting channel under the gate breaks apart and the samples have electrical characteristics of multiply–connected dots. Wide sweeps of the gate voltage are accompanied by sudden switching, which could be due to charging/discharging of some traps in the oxide. If we restrict the sweeps to $`<1`$ V, we can obtain reproducible results for several days.
In Fig. 1 the differential conductance $`G`$ is plotted as a function of $`V_g`$ at a source-drain bias $`V_b=0`$ for six different temperatures from one of the samples. From the device geometry the dot–gate capacitance is estimated to be 1-2 aF. We attribute large peaks at $`V_g=3.46`$, 3.54, 3.67 and 3.73 V to the main lithographically defined quantum dot. From the analysis of $`G`$ vs. $`V_b`$ and $`V_g`$ data we estimate gate voltage – to – single particle energy conversion coefficient $`\alpha 8.5`$ mV/meV.
The $`G`$ in the valleys between most of the peaks is thermally activated and vanishes rapidly at low $`T`$. However, in some valleys (for example between peaks at 3.54 and 3.67 V) the $`G`$ is almost $`T`$-independent. Remarkably, in such valleys the $`G`$ vs. $`V_b`$ data reveals a maximum close to the zero bias in the entire $`V_g`$ range of the valley. This is in striking contrast to the broad minimum around $`V_b=0`$ observed in the neighboring Coulomb blockade valleys. In Fig. 2a we plot a representative $`G`$ vs. $`V_b`$ curve measured at $`V_g=3.57`$ V. The peak at $`V_b=0.08`$ mV has a weak dependence on the $`V_g`$: it shifts from $`V_b=0.06`$ mV at $`V_g=3.55`$ V to $`V_b=0.21`$ mV at $`V_g=3.66`$ V. Similar results which show slightly off–zero bias peaks that shift as a function of the $`V_g`$ have been reported in GaAs quantum dots. In the case of the Kondo regime, the shift of the peak from $`V_b=0`$ can be qualitatively explained by the energy– and $`V_g`$– dependent coupling of the dot to the leads. The maximum at $`V_b=0.08`$ mV vanishes at $`T>1`$ K and becomes a broad minimum.
This zero bias anomaly is sensitive to an external magnetic field $`B`$. At $`B>2`$ T, applied parallel to the conducting channel, the enhanced conductivity is suppressed and Coulomb blockade is restored in the entire range $`3.55`$ V $`<V_g<3.66`$ V. At $`B>2`$ T the conductance shows a broad minima near $`V_b=0`$. In Fig. 2b we plotted $`G`$ vs. $`V_b`$ at $`V_g=3.57`$ for different $`B`$. There is no apparent dependence of the peak position on $`B`$, while the peak magnitude decreases as $`B`$ is increased. The peak is completely suppressed by $`B2`$ T.
There are striking differences between the Kondo effect reported in GaAs dots and the zero bias anomaly in our data. One of the signatures of the Kondo effect is that at $`B>0`$ the zero bias peak in $`G`$ is split into two peaks separated by twice the Zeeman energy, $`\mathrm{\Delta }V_b=2E_Z/e`$. Such a splitting was reported in GaAs quantum dots as well as in metallic grains. $`E_Z=g\mu _BB=0.12`$ meV at $`B=1`$ T (assuming $`g=2`$ in Si) and the splitting is expected to be $`\mathrm{\Delta }V_b=0.23`$ mV (indicated by a bar in Fig. 2b). The width of the zero bias peak in our data is $`0.15`$ mV at $`B=1`$ T, half the expected Kondo splitting. However, we have seen no splitting of the zero bias maximum as a function of $`B`$ in our data up to $`B=2`$ T, the highest field at which the maximum is still observed. Also, the position of this maximum is not effected by magnetic field.
An underlying physics for the Kondo effect requires the highest occupied level in the dot to be at least doubly degenerate. Adding an extra electron to the dot costs just a bare charging energy $`U_c=e^2/2C`$, where $`C`$ is the total capacitance from the dot to the gate and leads. Adding a second electron should cost $`U_c+\mathrm{\Delta }E`$, where $`\mathrm{\Delta }E`$ is due to the size quantization in the dot (or it can be the same $`U_c`$ if the level is more than two–fold degenerate). Thus, the Kondo effect is expected to be observed in the narrower valley between two adjacent charge–degenerate peaks which are separated by $`\mathrm{\Delta }V_g=U_c/\alpha `$, while neighboring valleys are expected to be wider with gate voltage separation of $`\mathrm{\Delta }V_g=(U_c+\mathrm{\Delta }E)/\alpha `$. However, we observed zero bias anomaly in the widest valley with $`\mathrm{\Delta }V_g=170`$ mV, while the neighboring valleys with widths 80 and 60 mV have no zero bias anomalies at $`B=0`$. There is also a characteristic shift of the charge–degeneracy peaks as a function of temperature in the Kondo regime. As temperature decreases, off-resonant conductance is enhanced which results in the shift of the charge–degeneracy peaks toward each other. Instead, we observed that the peak at $`V_g=3.54`$ V shifts to the lower gate voltages as the temperature is decreased, while the position of the peak at $`V_g=3.67`$ V is almost temperature independent.
Dependence of the conductance of the zero bias peak $`G^P`$ on $`T`$, $`B`$ and $`V_b`$ is shown in Fig. 3. The temperature range $`0.3<T<1`$ K, where zero bias anomaly is observed, is not sufficient to extract the functional dependence $`G^P(T)`$ with certainty, although it is close to being logarithmic. The zero bias peak is superimposed on a parabolic $`V_b`$–dependent background, thus we cannot unambiguously conclude what is the functional dependence of $`G^P`$ on the bias voltage. In contrast, $`G^P`$ is a strong function of the magnetic field. As shown in the inset in Fig. 3, magnetic field exponentially suppresses the conductance by more than an order of magnitude. Note, that one expects a weak logarithmic suppression of $`G`$ by magnetic field at $`V_b=0`$ in the Kondo regime.
Another striking result is that in some Coulomb blockade valleys the zero bias anomaly appear only at non-zero magnetic field. These valleys may group around the valley where the zero bias anomaly is observed at $`B=0`$. For example, at $`B=0`$ we observe a peak in $`G`$ at $`V_b0`$ in the valley 3.54 V $`<V_g<`$ 3.67 V (Fig. 2), while there are broad minima at $`V_b=0`$ in the neighboring valleys (3.46 V $`<V_g<`$ 3.54 V and 3.67 V $`<V_g<`$ 3.72 V). The zero bias anomaly peak at 3.55 V $`<V_g<`$ 3.66 V is destroyed by $`B2`$ T and $`G`$ has a broad minimum centered at $`V_b=0`$ at higher magnetic field. However, at $`B=3`$ T $`G`$ has a maximum in the valley 3.46 V $`<V_g<`$ 3.54 V. In Fig. 4a the $`V_b`$–dependence of $`G`$ is shown in the center of that valley at $`V_g=3.49`$ V. While there is a minimum around $`V_b=0`$ at $`B<2`$ T and $`B>4`$ T, there is a pronounced peak at $`V_b=0.1`$ mV at $`B=3`$ T. In the neighboring valley (at $`V_g=3.707`$ V) we observed a peak at $`V_b=0.3`$ mV at $`B=5`$ T. At yet higher $`B=9`$ T there is a maximum at $`V_b=0.1`$ mV at $`V_g=3.78`$ V, as shown in Fig. 4b. These maxima are observed over a limited $`B`$ range of $`\mathrm{\Delta }B1`$ T.
In some Coulomb blockade valleys zero bias anomaly is observed at $`B=0`$, although the strongest zero bias peak is found at $`B>0`$. As shown in Fig. 4c, at $`V_g=4.6`$ V the strongest zero bias peak is at $`B=0.6`$ T; the peak becomes a broad minimum at $`B>2.5`$ T. There are no zero bias anomalies developed in the adjacent valleys in the experimental range of $`0<B<10`$ T.
To summarize our findings, we observed a suppression of the Coulomb blockade and an enhancement of the differential conductance at low temperatures at certain gate voltages. This anomaly is destroyed by i) raising the temperature, ii) increasing the bias, or iii) applying a magnetic field. Unlike in the Kondo effect, the zero bias anomaly in our experiment is not split by the magnetic field but, instead, the magnetic field suppresses it exponentially. Also, at certain gate voltages we observed a zero bias anomaly at $`B>0`$. Our data cannot be understood within the framework of the theories of the Kondo effect.
Authors gratefully acknowledge discussions with Ned Wingreen. The work was supported by ARO, ONR and DARPA.
|
no-problem/9908/hep-th9908039.html
|
ar5iv
|
text
|
# A Remark on Witten Effect for QCD Monopoles in Matrix Quantum Mechanics
## Abstract
In a recent work (hep-th/9905198) we argued that a certain matrix quantum mechanics may describe ’t Hooft’s monopoles which emerge in QCD when the theory is projected to its maximal Abelian subgroup. In this note we find further evidence which supports this interpretation. We study the theory with a non-zero theta-term. In this case, ’t Hooft’s QCD monopoles become dyons since they acquire electric charges due to the Witten effect. We calculate a potential between a dyon and an anti-dyon in the matrix quantum mechanics, and find that the attractive force between them grows as the theta angle increases.
preprint: NYU-TH-99/07/03, HUTP-99/A040, NUB 3204
’t Hooft has shown that a new and rich structure emerges in QCD when certain unusual gauge fixing conditions are imposed on the theory. This class of unitary gauges projects QCD onto its maximal Abelian subgroup, i.e., breaks $`SU(N)`$ down to $`U(1)^{N1}`$. In addition to massless and massive gauge fields, which are present after this symmetry breaking, some point-like monopoles emerge in the theory . These monopoles appear in a somewhat unusual way, as singularities of a chosen gauge fixing condition. Given the importance of these monopoles in various analytic and lattice studies of QCD (see and references therein), it is desirable to have some gauge independent description of the dynamics of these objects. Recently, we addressed this issue in by considering pure QCD on a spatial three-torus. The T-dual form of pure QCD on a spatial torus can be interpreted as a certain matrix quantum mechanics. In some arguments were presented which lead us to conjecture that this quantum mechanics describes the dynamics of ’t Hooft’s monopoles. The aim of this note is to seek for further evidence in favor of this conjecture. Below we consider pure QCD with the theta term. It is known that ’t Hooft’s monopoles acquire electric charges due to the Witten effect once non-zero theta angle is introduced . If the identification of the excitations of the matrix quantum mechanics with ’t Hooft’s monopoles is correct, then the Witten effect should also be seen within the matrix model. In other words, if one calculates the interaction force between the point-like objects of the matrix model, then this force should depend on the theta angle. In fact, the attractive force between a monopole and an antimonopole should be greater then it is for a zero theta angle. In what follows we will show that this is indeed the case: An interacting monopole-antimonopole pair becomes a dyon-anti-dyon pair once the theta angle is switched on, and the attractive force between them grows as $`\theta `$ increases.
Consider four-dimensional pure $`SU(N)`$ QCD in the presence of the $`\theta `$ angle. The corresponding Lagrangian density reads:
$$_{\mathrm{YM}}=\frac{1}{4g_{\mathrm{YM}}^2}G_{\mu \nu }^aG^{a\mu \nu }+\frac{\theta }{32\pi ^2}G_{\mu \nu }^a\stackrel{~}{G}^{a\mu \nu }=\frac{1}{4}\mathrm{Re}\left(\tau \left[G_{\mu \nu }^aG^{a\mu \nu }+iG_{\mu \nu }^a\stackrel{~}{G}^{a\mu \nu }\right]\right).$$
(1)
Here $`\stackrel{~}{G}^{a\mu \nu }\frac{1}{2}ϵ^{\mu \nu \lambda \rho }G_{\lambda \rho }^a`$, and $`\tau =\tau _1+i\tau _21/g_{\mathrm{YM}}^2+i\theta /8\pi ^2`$. To avoid complications with Gribov copies, in the following we will be working in the $`A_0=0`$ gauge.
Let $`\zeta 1/\mathrm{\Lambda }_{\mathrm{YM}}`$ be the effective correlation length of the theory, where $`\mathrm{\Lambda }_{\mathrm{YM}}`$ is the dynamically generated QCD scale. Let us compactify the theory on a rectangular three-torus $`T_LS^1\times S^1\times S^1`$ with the radii of all three circles equal $`L\zeta `$. This corresponds to the strong coupling regime of the theory (see discussions in ). We can rewrite pure QCD compactified on $`T_L`$ as a matrix quantum mechanics compactified on a dual three-torus $`T_R\stackrel{~}{S}^1\times \stackrel{~}{S}^1\times \stackrel{~}{S}^1`$ with the radii of all three circles equal $`R\alpha ^{}/L`$, where the parameter $`\alpha ^{}`$ is defined via the QCD scale $`\mathrm{\Lambda }_{\mathrm{YM}}`$ as follows: $`\alpha ^{}\zeta ^2=1/\mathrm{\Lambda }_{\mathrm{YM}}^2`$. In this T-dual formulation of the theory the dynamical variables are time-dependent matrices $`\mathrm{\Phi }_i(t)`$, $`i=1,2,3`$, transforming in the adjoint representation of (global) $`SU(N)`$. In addition to the color indices, for a given value of the index $`i=1,2,3`$, the matrices $`\mathrm{\Phi }_i`$ also carry indices corresponding to the winding modes. In general, $`\mathrm{\Phi }`$’s give a matrix representation of a covariant derivative on a torus . In the following we will suppress for simplicity these winding indices. The corresponding Lagrangian of the matrix quantum mechanics (with the appropriate normalization for $`\mathrm{\Phi }_i`$) is given by:
$$=\frac{1}{2g\sqrt{\alpha ^{}}}\mathrm{Tr}\left(\dot{\mathrm{\Phi }}_i^2+\frac{1}{2(2\pi \alpha ^{})^2}[\mathrm{\Phi }_i,\mathrm{\Phi }_j]^2\frac{i\lambda }{2\pi \alpha ^{}}ϵ_{ijk}[\mathrm{\Phi }_i,\mathrm{\Phi }_j]\dot{\mathrm{\Phi }}_k\right),$$
(2)
where $`\dot{\mathrm{\Phi }}_i`$ denotes the time derivative of $`\mathrm{\Phi }_i`$, and the traces over color and winding indices (with appropriate normalizations) are implicit. This Lagrangian should be amended by a corresponding constraint equation (a counterpart of the Gauss’s law) to describe pure QCD in a T-dual picture . The new coupling constant $`g`$ is defined as follows:
$$g(R/L)^{3/2}g_{\mathrm{YM}}^2/4\pi .$$
(3)
Also, the last term in (2) is due to the $`\theta `$-term in (1), and the corresponding coupling $`\lambda `$ is given by:
$$\lambda \tau _2/\tau _1=\theta g_{\mathrm{YM}}^2/8\pi ^2.$$
(4)
For $`\theta =0`$ (2) reduces to the usual bosonic matrix quantum mechanics Lagrangian .
Following the Lagrangian (2) describes the dynamics of ’t Hooft’s QCD monopoles. Let us notice that in the limit $`L\zeta `$, that is, $`R\zeta `$, which we are interested in, the matrix quantum mechanics (2) is as complicated a theory as strongly coupled pure QCD, the reason being that it contains light winding modes (which map to the Kaluza-Klein modes in the T-dual QCD description) whose masses scale as $`R/\alpha ^{}=1/L`$ . However, certain aspects of pure QCD in a large volume (which is a strongly coupled theory) might be more transparent in the matrix quantum mechanics approach. In particular, the monopole mass in the theory is given by $`M=1/2g\sqrt{\alpha ^{}}`$, and in the regime we are discussing $`M\mathrm{\Lambda }_{\mathrm{YM}}`$ . Thus, it is reasonable to consider interactions between monopoles when they are moving very slowly (or, are almost at rest). At such a low energies the light winding modes are not yet excited in the model. Thus, we can neglect the contributions of these modes in the calculation. On the matrix model side interactions between monopoles (a monopole-antimonopole pair) are described by off-diagonal elements in $`\mathrm{\Phi }_i`$. Thus, following , consider the $`U(2)`$ case where we have two monopoles with opposite magnetic charges. Let us make the standard decomposition of the $`\mathrm{\Phi }`$ field into it’s classical and quantum parts:
$$\mathrm{\Phi }_i=\mathrm{\Phi }_i^{\mathrm{cl}}+\delta \mathrm{\Phi }_i.$$
(5)
Here we choose the classical solution as follows:
$`\mathrm{\Phi }_1^{\mathrm{cl}}={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}r& \text{0}\\ \text{0}& r\end{array}\right),\mathrm{\Phi }_2^{\mathrm{cl}}=0,\mathrm{\Phi }_3^{\mathrm{cl}}=0.`$ (8)
That is, the two monopoles are at a distance $`r`$ apart from each other in one of the spatial directions. In order to find an effective potential between them one performs integration with respect to off-diagonal fluctuations. In the absence of the $`\theta `$-angle the effective potential between the monopoles is given by (see and references therein):
$$V_{\mathrm{eff}}^{(\theta =0)}\frac{r}{\alpha ^{}}.$$
(9)
In fact, the easiest way to deduce the effective potential in the presence of the $`\theta `$-angle is to rewrite the Lagrangian (2) as follows:
$$=\frac{1}{2g\sqrt{\alpha ^{}}}\mathrm{Tr}\left(\left(\dot{\mathrm{\Phi }}_i\frac{i\lambda }{2(2\pi \alpha ^{})}ϵ_{ijk}[\mathrm{\Phi }_i,\mathrm{\Phi }_j]\right)^2+\frac{1+\lambda ^2}{2(2\pi \alpha ^{})^2}[\mathrm{\Phi }_i,\mathrm{\Phi }_j]^2\right).$$
(10)
Note that the difference between the Lagrangians with $`\lambda =0`$ and $`\lambda 0`$ is in the redefinition of the conjugate momentum and rescaling $`\alpha ^{}\alpha ^{}/\sqrt{1+\lambda ^2}`$ in the term containing the commutator $`[\mathrm{\Phi }_i,\mathrm{\Phi }_j]^2`$ (which is responsible for interactions between monopoles) This is, however, not true for the Hamiltonian of the theory which will contain a term linear in theta along with the quadratic term arising in front of the commutator $`[\mathrm{\Phi }_i,\mathrm{\Phi }_j]^2`$.. Performing explicitly integration of the off-diagonal excitations in (10), and calculating the corresponding functional determinant, one finds the effective potential in the presence of the $`\theta `$-angle
$$V_{\mathrm{eff}}^{(\theta 0)}=\sqrt{1+\lambda ^2}V_{\mathrm{eff}}^{(\theta =0)}.$$
(11)
We see that, as in the case without the $`\theta `$-angle, there is a linearly rising potential between a monopole and an anti-monopole. Thus, there is a string stretched between them, and the string tension $`T_s`$ has a non-trivial $`\theta `$-dependence:
$$T_s\sqrt{1+\lambda ^2}\mathrm{\Lambda }_{\mathrm{YM}}^2=\sqrt{1+\left(\theta g_{\mathrm{YM}}^2/8\pi ^2\right)^2}\mathrm{\Lambda }_{\mathrm{YM}}^2.$$
(12)
As a result, the attraction force between the pair increases when the theta angle is switched on. This corresponds to the fact that monopoles acquire electric charges and become dyons. This is consistent with the fact that we expect Witten’s effect to take place for magnetic monopoles - a monopole with a magnetic charge $`h`$ becomes a dyon with the electric charge
$`e=\left({\displaystyle \frac{\theta g_{\mathrm{YM}}^2}{16\pi ^2}}\right)h`$ (13)
in the presence of the $`\theta `$ angle . Thus, the fact that the interaction force derived from the above matrix quantum mechanics depends non-trivially on the $`\theta `$-angle gives additional evidence that ’t Hooft’s QCD monopoles might indeed be described by the former. Dyons in this case can (very roughly) be thought of as complicated bound states of a monopole and off-diagonal gluons.
Let us also point out that the string tension $`T_s`$ is invariant under the S-duality transformation $`\tau 1/\tau `$. On the other hand, at first it might appear strange that it is not invariant under the shift $`\theta \theta +2\pi `$. This is, however, expected as the electric charge $`e`$ given by (13) is not invariant under such shifts either; the charge (13) gets shifted by a fundamental unit of the “electric charge” (which in our notations is $`g_{\mathrm{YM}}^2h/8\pi `$), i.e., $`ee+g_{\mathrm{YM}}^2h/8\pi `$. This can be interpreted as follows: ’t Hooft’s dyons at $`\theta `$ can be viewed as a bound state of the corresponding dyon at $`\theta 2\pi `$ and a gluon .
The work of G.G. was supported by the grant NSF PHY-94-23002. The work of Z.K. was supported in part by the grant NSF PHY-96-02074, and the DOE 1994 OJI award. Z.K. would also like to thank Albert and Ribena Yu for financial support.
|
no-problem/9908/astro-ph9908358.html
|
ar5iv
|
text
|
# PSCz-1.2 Jy Comparison: A Spherical Harmonics Approach
## 1 Introduction
Nusser & Davis (1994) show that in linear gravitational instability (GI) theory the peculiar velocity field in redshift space is irrotational and thus can be expressed in terms of a potential: $`\stackrel{}{v}=\mathrm{\Phi }(\stackrel{}{s})`$. The angular dependencies of the potential velocity field and the galaxy overdensity field \[both measured in redshift space and expanded in spherical harmonics, $`\mathrm{\Phi }_{lm}(s)`$ and $`\widehat{\delta }_{lm}(s)`$, respectively\] are related by a modified Poisson equation:
$$\frac{1}{s^2}\frac{d}{ds}\left(s^2\frac{d\mathrm{\Phi }_{lm}}{ds}\right)\frac{1}{1+\beta }\frac{l(l+1)\mathrm{\Phi }_{lm}}{s^2}=\frac{\beta }{1+\beta }\left(\widehat{\delta }_{lm}\frac{1}{s}\frac{d\mathrm{ln}\varphi }{d\mathrm{ln}s}\frac{d\mathrm{\Phi }_{lm}}{ds}\right),$$
(1)
where $`\varphi (s)`$ is the selection function. To solve this differential equation we first compute the density field on an angular grid using cells of equal solid angle and 52 bins in redshift out to $`s=\mathrm{18\hspace{0.17em}000}`$ kms<sup>-1</sup>.
$$1+\widehat{\delta }_j(\stackrel{}{s}_n)=\frac{1}{(2\pi )^{3/2}\sigma _{1.2n}^3}\underset{i}{\overset{N_j}{}}\frac{1}{\varphi (s_i)}\mathrm{exp}\left[\frac{(\stackrel{}{s}_n\stackrel{}{s}_i)^2}{2\sigma _{1.2n}^2}\right]$$
(2)
where the sum is over all the galaxies within the catalogue $`j`$, $`N_j`$. The Gaussian smoothing width for the cell $`n`$ at redshift $`s_n`$, $`\sigma _{1.2n}`$, is given by $`\sigma _{1.2n}=[\overline{n}_{1.2}\varphi _{1.2}(s_n)]^{1/3}`$ (or 100 km s$`^1`$when such a length is smaller than this), where $`\overline{n}_{1.2}`$ and $`\varphi _{1.2}`$ are the 1.2 Jy mean number density and selection function, respectively.
## 2 Datasets
PSCz is a new redshift survey which resulted from a collaborative effort involving several British institutions (Durham, Oxford, London, Edinburgh and Cambridge). The catalogue contains some 15500 IRAS PSC (Point Source Catalogue) galaxies with a 60 $`\mu `$m flux larger than 0.6-Jy. This survey covers 84$`\%`$ of the sky. A more detailed description of the catalogue is given by Saunders et al. 1999 (these proceedings). The 1.2-Jy catalogue (Fisher et al. 1995) contains 5321 IRAS PSC galaxies with a 60 $`\mu m`$ limit of 1.2-Jy which covers 87.6$`\%`$ of the sky.
## 3 Results, Discussion and Conclusions
We compare the line of sight peculiar velocity, $`u(\stackrel{}{s})[\stackrel{}{v}(\stackrel{}{v})\stackrel{}{v}_{LG}]\widehat{s}`$, and density perturbations, $`\widehat{\delta }(s)`$, fields inferred from the IRAS PSCz and 1.2-Jy redshift surveys, in redshift-space. To compute the spherical harmonics coefficients up to $`l_{max}=16`$ we apply the algorithm introduced by Nusser & Davis (1994) setting $`\beta =0.6`$.
In Fig. 1 we display the monopole of the velocity field, $`u_{00}`$ (two top panels) and the monopole of the density $`\widehat{\delta }_{00}`$ (bottom panel). In the top-left panel, the estimate of the velocity monopole in the 1.2-Jy (dashed line) is systematically larger than the PSCz one (continuous thick line) within a redshift of $`s8000`$ km s<sup>-1</sup>. The density monopole of the two surveys show a similar qualitative radial dependency (bottom panel). In Fig. 2 we show the three velocity dipole terms of the two surveys along with the total amplitude (bottom left panel). The various dipole moments exhibit good agreement, except $`u_{11}`$ (top left panel) for which there is a discrepancy of 100 km$`s^1`$ in the redshift range $`15007500`$ kms<sup>-1</sup>. Good agreement is also found when comparing higher order multipoles (Teodoro et al. 1999). Note that when all the multipoles are considered (i.e. when performing a full v-v comparison) the PSCz and 1.2-Jy gravity fields look fully consistent (Branchini et al. 1999).
In all plots, shaded and hatched regions represent 1-$`\sigma `$ shot-noise uncertainties computed from 20 bootstrap realizations of PSCz and 1.2 Jy surveys, respectively. Where does the discrepancy between the monopoles of the PSCz and 1.2 Jy surveys come from? As shown in the plots, the difference is larger than that expected from the shot-noise. Tadros et al. (1999) have suggested that the PSCz catalogue may be incomplete for fluxes $`0.7`$ Jy. If true, then we would expect that the velocity monopole for the PSCz with a flux cut at 0.7 Jy (PSCz<sub>0.7</sub>) would be in good agreement with the 1.2-Jy survey. The dotted line ($`u_{00,\text{PSCz}_{0.7}}`$) in the top panel of Fig. 1 shows however that this is not case. It is only in cutting the PSCz catalogue at a flux level of 1.2 Jy that, as expected, the discrepancy disappears, provided that we use the same mask for both catalogues. This is clearly seen in the top right panel of Fig. 1 in which the thin and dot-dashed lines indicate 1.2-Jy<sub>PSCz</sub>(1.2 Jy with the same mask as PSCz) and PSCz<sub>1.2-Jy</sub> (PSCz with a 1.2 Jy flux limit) velocity monopoles, respectively.
A more detailed comparison between the two catalogues is in progress in which we assess the importance of systematic errors by using a suite of 1.2 JY and PSCz mock catalogues drawn from N-body simulations.
###### Acknowledgements.
We thank Marc Davis for providing the original version of the reconstruction code and Adi Nusser for many dicussions. LT has been supported by the grants PRAXIS XXI/BPD/16354/98 and PRAXIS/C/FIS/ 13196/1998.
|
no-problem/9908/hep-th9908090.html
|
ar5iv
|
text
|
# Tiling the plane without supersymmetry
\[
## Abstract
We present a way of tiling the plane with a regular hexagonal network of defects. The network is stable and follows in consequence of the three-junctions that appear in a model of two real scalar fields that presents $`Z_3`$ symmetry. The $`Z_3`$ symmetry is effective in both the vacuum and defect sectors, and no supersymmetry is required to build the network.
\]
Domain walls appear in diverse branches of physics, envolving energy scales as different as the ones for instance in magnetic materials and in cosmology . They live in three spatial dimensions as bidimensional objects that arise in systems with at least two isolated degenerate minima. In field theory they appear in the $`(3,1)`$ dimensional space-time, and this may happen in supersymmetric theories, although supersymmetry plays no fundamental role for the presence of domain walls.
Very recently, in a paper by Gibbons and Townsend , and also in Refs. , one investigates the presence of domain walls and their possible intersections in a Wess-Zumino model, with a polynomial superpotential. In the supersymmetric theory, one can classify the classical solutions as BPS and non-BPS states, according to the work of Bogomol’nyi, and of Prasad and Sommerfield . The BPS states are stable, and are expected to play some role in investigating duality in supersymmetric models. We recall that no BPS state can be annihilated under continuum variation of the parameters that define the supersymmetric theory.
In Ref. one investigates models of coupled real scalar fields in bidimensional space-time. These investigations provide a concrete way of finding BPS states and suggest other studies, in particular on the subject of defects inside defects – see Ref. . Most of the models investigated in can be seen as real bosonic portions of supersymmetric theories. In supersymmetric models the presence of discrete symmetry may produce BPS and non-BPS defects. The BPS states lie in shorter multiplets, and preserve the supersymmetry only partially . There are BPS states that preserve $`1/2`$ of the supersymmetry, but the possibility of BPS states preserving $`1/4`$ supersymmetry is subtler, and is shown to appear as junctions of domain walls in the recent papers .
In the present work we start dealing with the bosonic portions of supersymmetric theories. We do this guided by the discrete $`Z_3`$ symmetry, with the aim of describing the presence of three-junctions and the network of defects that it can generate. We first point out that supersymmetry introduces restrictions that may lead to instability of the junction, or at least of the network that it could generate. We then examine another model, and show that all the difficulties found in the supersymmetric context are circumvented by just giving up supersymmetry.
The subject of this work may be of interest to several different branches of physics, in particular in applications concerning the entrapment of networks of defects inside domain walls. This possibility can be implemented with three scalar fields, in models engendering the $`Z_2\times Z_3`$ symmetry, following the lines of Refs. . Other applications may include strong interactions, if we recall that the $`Z_3`$ group is the center of the $`SU(3)`$ group – see for instance Refs. . Also, there are applications to systems of condensed matter, in particular on issues concerning pattern formation , as for instance in the case of the thermal convection studied in Ref. .
We start describing the two real scalar fields $`\varphi `$ and $`\chi `$ in bidimensional space-time with the Lagrangian density
$$=\frac{1}{2}_\alpha \varphi ^\alpha \varphi +\frac{1}{2}_\alpha \chi ^\alpha \chi V$$
(1)
Here $`V=V(\varphi ,\chi )`$ is the potential. In the supersymmetric case it has the general form
$$V(\varphi ,\chi )=\frac{1}{2}W_\varphi ^2+\frac{1}{2}W_\chi ^2$$
(2)
where $`W=W(\varphi ,\chi )`$ is the superpotential.
The superpotential allows introducing several properties, as shown in Refs. . For instance, for static fields the equations of motion are solved by first-order differential equations $`d\varphi /dx=W_\varphi `$ and $`d\chi /dx=W_\chi `$. The energy of solutions of the first-order equations are given by $`E_B^{ij}=|\mathrm{\Delta }W_{ij}|`$, with $`\mathrm{\Delta }W_{ij}=W_iW_j`$ and $`W_i=W(\varphi _i,\chi _i)`$, where $`(\varphi _i,\chi _i)`$ represents the i-th vacuum state of the model. This is the Bogomol’nyi bound, and the corresponding solutions are BPS solutions. The BPS solutions are linearly or classically stable.
We guide ourselves toward the topological solutions by introducing the topological current
$$J^\alpha =\epsilon ^{\alpha \beta }_\beta \left(\genfrac{}{}{0pt}{}{\varphi }{\chi }\right)$$
(3)
It obeys $`_\alpha J^\alpha =0`$, and it is also a vector in the $`(\varphi ,\chi )`$ plane. For static configurations we have $`J_\alpha ^tJ^\alpha =\rho ^t\rho `$, where $`\rho =\rho (\varphi ,\chi )`$ is the charge density. This charge density allows writing $`\rho ^t\rho `$ as twice the kinetic energy density of the topological solution, and this can be used to infer stability of junctions – see below.
Let us now consider a specific model, defined by the superpotential
$`W(\varphi ,\chi )`$ $`=`$ $`\lambda \varphi ^33\lambda \varphi +\lambda \varphi \left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^2`$ (5)
$`+\lambda \sqrt{{\displaystyle \frac{4}{27}}}\left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^3\lambda \left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^2`$
The first-order differential equations are given by
$`{\displaystyle \frac{d\varphi }{dx}}`$ $`=`$ $`3\lambda (\varphi ^21)+\lambda \left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^2`$ (6)
$`{\displaystyle \frac{d\chi }{dx}}`$ $`=`$ $`2\lambda \left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)\left[\varphi +\sqrt{{\displaystyle \frac{1}{3}}}\left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)1\right]`$ (7)
The model is described by the fourth-order polynomial potential
$`V_4(\varphi ,\chi )`$ $`=`$ $`{\displaystyle \frac{9}{2}}\lambda ^2(\varphi ^21)^2+{\displaystyle \frac{1}{2}}\lambda ^2\left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^4`$ (10)
$`+3\lambda ^2(\varphi ^21)\left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^2`$
$`+2\lambda ^2\left(\chi +\sqrt{{\displaystyle \frac{1}{3}}}\right)^2\left[\varphi +\sqrt{{\displaystyle \frac{1}{3}}}\chi {\displaystyle \frac{2}{3}}\right]^2`$
Thus, it behaves standardly in one, two, and three spatial dimensions.
The above potential presents the three vacuum states $`(0,\sqrt{4/3}),(1,\sqrt{1/3})`$, and $`(1,\sqrt{1/3})`$. These minima form an equilateral triangle invariant under the $`Z_3`$ rotations, with side $`l=2`$. The values of the superpotential at the minima are $`W_1=\lambda ,W_2=2\lambda `$ and $`W_3=2\lambda `$. The energies of the BPS states are $`|\lambda |,\mathrm{\hspace{0.17em}3}|\lambda |,`$ and $`4|\lambda |`$. All the three sectors are BPS sectors, but they do not present the $`Z_3`$ symmetry that connects the vacuum states. The highest energy is associated to the sector connecting the second and third vacuum states. This is the only sector where we can find explicit solutions. They are given by $`\varphi (x)=\mathrm{tanh}(3\lambda x)`$ and $`\chi (x)=\sqrt{1/3}`$. This is a BPS state, representing an orbit in the $`(\varphi ,\chi )`$ plane. The orbit is a straight-line segment that connects the corresponding vacuum states. The orbits connecting the other vacua cannot be straight-line segments. They cannot be obtained by rotating the $`(\varphi ,\chi )`$ plane according to the $`Z_3`$ symmetry, and so the defect sectors do not present the $`Z_3`$ symmetry that connects the vacuum states. This fact also appears when one identifies the tensions of the BPS defects. They are given by $`t_1=|\lambda |`$, $`t_2=3|\lambda |`$, and $`t_3=4|\lambda |`$. They are different and do not obey the $`Z_3`$ symmetry. They are such that $`t_3=t_1+t_2`$, and do not strictly obey the triangle inequality one needs to ensure stability of the three-junction that appears in this model.
To circumvent instability of the three-junction we now follow Refs. . We make contact with these works after considering superpotentials that satisfy $`W_{\varphi \varphi }+W_{\chi \chi }=0`$. In this case, for harmonic superpotentials one adds to the two first-order equations $`d\varphi /dx=W_\varphi `$ and $`d\chi /dx=W_\chi `$ the two new first-order equations: $`d\varphi /dx=W_\chi `$ and $`d\chi /dx=W_\varphi `$. Solutions to these equations also minimize the energy and solve the equations of motion. This allows introducing $`\stackrel{~}{W}(\varphi ,\chi )`$ such that $`\stackrel{~}{W}_\varphi =W_\chi `$ and $`\stackrel{~}{W}_\chi =W_\varphi `$. We use $`W`$ and $`\stackrel{~}{W}`$ to introduce the complex superpotential, $`𝒲=W+i\stackrel{~}{W}`$. We write the complex superpotential in terms of the complex field $`\varphi +i\chi `$, and this is the way one gets from the investigations of Refs. to the recent possibility of describing three-junctions preserving $`1/4`$ supersymmetry. However, junctions require the presence of at least three minima, and this is only achieved when the superpotential is of at least the fourth-order power in the complex field. This means that the model behaves standardly only in one and two spatial dimensions. In this case one can show explicitly that the three-junction is stable and breaks $`1/4`$ supersymmetry, although supersymmetry itself does not allow the presence of a stable network of defects . Owing to the fact that each adjacent junction in the network has opposite winding number, any adjacent vacua should be connected with defect solutions also having opposite winding numbers along the same orbit . Since we have to use different conjugate Bogomol’nyi equations to take into account these winding numbers, the network cleary cannot be BPS and then can decay.
We then give up supersymmetry, turning our attention to polynomial potentials that engenders the $`Z_3`$ symmetry, and that supports stable three-junctions that generate a regular hexagonal network of defects. Interestingly, we have found a fourth-order polynomial potential that do the job. It is given by
$`V(\varphi ,\chi )`$ $`=`$ $`\lambda ^2\varphi ^2\left(\varphi ^2{\displaystyle \frac{9}{4}}\right)+\lambda ^2\chi ^2\left(\chi ^2{\displaystyle \frac{9}{4}}\right)`$ (12)
$`+\mathrm{\hspace{0.17em}2}\lambda ^2\varphi ^2\chi ^2\lambda ^2\varphi (\varphi ^23\chi ^2)+{\displaystyle \frac{27}{8}}\lambda ^2`$
This potential was introduced in Ref. . The equations of motion for static configurations are
$`{\displaystyle \frac{d^2\varphi }{dx^2}}`$ $`=`$ $`\lambda ^2\varphi \left(4\varphi ^2+4\chi ^23\varphi {\displaystyle \frac{9}{2}}\right)+3\lambda ^2\chi ^2`$ (13)
$`{\displaystyle \frac{d^2\chi }{dx^2}}`$ $`=`$ $`\lambda ^2\chi \left(4\varphi ^2+4\chi ^2+6\varphi {\displaystyle \frac{9}{2}}\right)`$ (14)
The potential has three degenerate minima, at the points $`v_1=(3/2)(1,0)`$ and $`v_{2,3}=(3/4)(1,\pm \sqrt{3})`$. These minima form an equilateral triangle, invariant under the $`Z_3`$ symmetry. The distance between the minima is $`(3/2)\sqrt{3}`$.
We can obtain the topological solutions explicitly. The easiest way to do this follows by first examining the sector that connects the vacua $`v_2`$ and $`v_3`$. This is so because in this case we set $`\varphi =3/4`$, searching for a strainght-line segment in the $`(\varphi ,\chi )`$ plane. This is compatible with the Eq. (13), and reduces the other Eq. (14) to the form
$$\frac{d^2\chi }{dx^2}=\lambda ^2\left(4\chi ^3\frac{27}{4}\chi \right)$$
(15)
This implies that the orbit connecting the vacua $`v_2`$ and $`v_3`$ is a straight line. It is such that, along the orbit the $`\chi `$ field feels the potential $`\lambda ^2[\chi ^2(27/16)]^2`$. This shows that the model reduces to a model of a single field, and the solution satisfies the first-order equation
$$\frac{d\chi }{dx}=\sqrt{2}\lambda \left(\chi ^2\frac{27}{16}\right)$$
(16)
The solution is
$$\chi (x)=\frac{3}{4}\sqrt{3}\mathrm{tanh}\left(\sqrt{\frac{27}{8}}\lambda x\right)$$
(17)
The other solutions can be obtained by rotations obeying the $`Z_3`$ symmetry of the model.
The full set of solutions of the equations of motion are collected below. In the sector connecting the minima $`v_2`$ and $`v_3`$ they are
$`\varphi _{(2,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{4}}`$ (18)
$`\chi _{(2,3)}^{(\pm )}`$ $`=`$ $`\pm {\displaystyle \frac{3}{4}}\sqrt{3}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)`$ (19)
In the sector connecting the minima $`v_1`$ and $`v_2`$ they are
$`\varphi _{(1,2)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\pm {\displaystyle \frac{9}{8}}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)`$ (20)
$`\chi _{(1,2)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\sqrt{3}{\displaystyle \frac{3}{8}}\sqrt{3}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)`$ (21)
In the sector connecting the minima $`v_1`$ and $`v_3`$ they are
$`\varphi _{(1,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}{\displaystyle \frac{9}{8}}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)`$ (22)
$`\chi _{(1,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\sqrt{3}{\displaystyle \frac{3}{8}}\sqrt{3}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)`$ (23)
The label $`(\pm )`$ is used to identify kink and antikink. All the solutions have the same energy, $`(9/4)\sqrt{27/8}|\lambda |`$.
We examine how the bosonic fields behave in the background of the classical solutions. We do this by considering fluctuations around the static solutions $`\varphi (x)`$ and $`\chi (x)`$. We use the equations of motion to see that the fluctuations depend on the potential
$$𝐔(x)=\left(\genfrac{}{}{0pt}{}{V_{\varphi \varphi }V_{\varphi \chi }}{V_{\chi \varphi }V_{\chi \chi }}\right)$$
(24)
Evidently, after obtaining the derivatives we substitute the fields by their classical static values $`\varphi (x)`$ and $`\chi (x)`$. The model under consideration is defined by the potential (12). In this case we use (18) and (19) to obtain two decoupled equations for the fluctuations. The potentials of the corresponding Schrödinger-like equations are
$`U_{11}(x)`$ $`=`$ $`{\displaystyle \frac{27}{8}}\lambda ^2\left[42\mathrm{sech}^2\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)\right]`$ (25)
$`U_{22}(x)`$ $`=`$ $`{\displaystyle \frac{27}{8}}\lambda ^2\left[46\mathrm{sech}^2\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda x\right)\right]`$ (26)
The eigenvalues can be obtained explicitly: in the $`\chi `$ direction we get $`w_0^\chi =0`$ and $`w_1^\chi =(9/2)\sqrt{\lambda ^2/2}`$, and in the $`\varphi `$ direction we have $`w_0^\varphi =(9/2)\sqrt{\lambda ^2/2}`$. This shows that the pair $`(\text{18})`$ and $`(\text{19})`$ is stable, and by symmetry we get that all the three topological solutions are stable solutions.
The classical solutions present the nice property of having energy evenly distributed in their kinetic (k) and potential (p) portions. In terms of energy density they are
$$\mathrm{k}(x)=\mathrm{p}(x)=\frac{1}{4}\left(\frac{27}{8}\right)^2\lambda ^2\mathrm{sech}^4\left(\sqrt{\frac{27}{8}}\lambda x\right)$$
(27)
To understand this feature we recall the calculation done explicitly in the sector with $`\varphi =3/4`$, constant. There the model is shown to reduce to a model of a single field, a model that supports BPS solutions. Within this context, the above solutions are very much like the non-BPS solutions that appear in supersymmetric systems . We use this property and the topological current (3) to obtain $`\rho ^t\rho =\epsilon `$, where $`\epsilon (x)=\mathrm{k}(x)+\mathrm{p}(x)`$ is the (total) energy density of the solution. We use this result and the notation $`ij`$, to identify the sector connecting the vacua $`(\varphi _i,\chi _i)`$ and $`(\varphi _j,\chi _j)`$, to show that for any two different sectors $`ij`$ and $`jk`$, $`i,j,k=1,2,3`$ we get that
$$(\rho _{ij}+\rho _{jk})^t(\rho _{ij}+\rho _{jk})<\rho _{ij}^t\rho _{ij}+\rho _{jk}^t\rho _{jk}$$
(28)
This condition shows that the three-junction is a process of fusion of defects that occurs exothermically, providing stability of junctions in the present model. This result is more general than the one in Ref. , which appears within the context of supersymmetry. Evidently, our result also works for BPS and non-BPS solutions that appears in supersymmetric models, with the property of having energy evenly distributed in their kinetic and potential portions .
We notice that the orbits corresponding to the stable defect solutions form an equilateral triangle in the $`(\varphi ,\chi )`$ plane. This is so because the solutions are straight-line segments joining the three vacuum states in configuration space. They are degenerate in energy, and this allows associating to each defect the same tension
$$t=\frac{9}{4}\sqrt{\frac{27}{8}}|\lambda |$$
(29)
This makes $`t_{ij}<t_{jk}+t_{ki},i,j,k=1,2,3`$, and now the inequality is strictly valid in this case, stabilizing the three-junction that appears in this model when one enlarges the space-time to three spatial dimensions.
We consider the possibility of junctions in the plane, which may give rise to a planar network of defects. We work in the $`(2,1)`$ space-time, in the plane $`(x,y)`$. We identify the plane $`(x,y)`$ with the space of configurations, the plane $`(\varphi ,\chi )`$. We illustrate this situation by considering, for instance, the solutions we have already obtained. They are collected in Eqs. (18)-(23) in (1,1) dimensions. In the planar case they change to
$`\varphi _{(2,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{4}}`$ (30)
$`\chi _{(2,3)}^{(\pm )}`$ $`=`$ $`\pm {\displaystyle \frac{3}{4}}\sqrt{3}\mathrm{tanh}\left(\sqrt{{\displaystyle \frac{27}{8}}}\lambda y\right)`$ (31)
and
$`\varphi _{(1,2)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}{\displaystyle \frac{9}{8}}\mathrm{tanh}\left({\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{27}{8}}}\lambda (y+\sqrt{3}x)\right)`$ (32)
$`\chi _{(1,2)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\sqrt{3}\pm {\displaystyle \frac{3}{8}}\sqrt{3}\mathrm{tanh}\left({\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{27}{8}}}\lambda (y+\sqrt{3}x)\right)`$ (33)
and
$`\varphi _{(1,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\pm {\displaystyle \frac{9}{8}}\mathrm{tanh}\left({\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{27}{8}}}\lambda (y\sqrt{3}x)\right)`$ (34)
$`\chi _{(1,3)}^{(\pm )}`$ $`=`$ $`{\displaystyle \frac{3}{8}}\sqrt{3}\pm {\displaystyle \frac{3}{8}}\sqrt{3}\mathrm{tanh}\left({\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{27}{8}}}\lambda (y\sqrt{3}x)\right)`$ (35)
These planar defects are domain walls, and can be used to represent the three-junction in the limit of thin walls.
The three-junction that appears in this $`Z_3`$-symmetric model allows building a network of defects, precisely in the form of a regular hexagonal network, as depicted in FIG. 1 in the thin wall approximation. In this network the tension associated to the defect is the tipical value of the energy in this tiling of the plane with a regular hexagonal network, which seems to be the most efficient way of tiling the plane. As we have shown, our model behaves standardly in $`(3,1)`$ dimensions. It supports stable three-junctions that generate a stable regular hexagonal network of defects.
We thank R.F. Ribeiro for discussions, and CAPES, CNPq, and PRONEX for partial support.
|
no-problem/9908/physics9908031.html
|
ar5iv
|
text
|
# Guiding neutral atoms around curves with lithographically patterned current-carrying wires
## Abstract
Laser-cooled neutral atoms from a low-velocity atomic source are guided via a magnetic field generated between two parallel wires on a glass substrate. The atoms bend around three curves, each with a 15-cm radius of curvature, while traveling along a 10-cm-long track. A maximum flux of $`210^6\mathrm{atoms}/\mathrm{sec}`$ is achieved with a current density of $`310^4\mathrm{A}/\mathrm{cm}^2`$ in the $`100\times 100`$-$`\mu \mathrm{m}`$-cross-section wires. The kinetic energy of the guided atoms in one transverse dimension is measured to be 42 $`\mu \mathrm{K}`$.
Just as optical waveguides play a central role in many aspects of modern optics, from communications to integrated optics, atom waveguides are likely to be an enabling technology for future atom-optics-based science. In particular, well-characterized atom waveguides may make possible inertial and rotation measurements of exquisite sensitivity via large-enclosed-area atom interferometers . One of the first guides for atoms was based on optical forces, where hollow glass fibers guide light, and the light in turn guides atoms .
Atom guiding using magnetic forces from current-carrying wires has been demonstrated more recently . From the point-of-view of using atom-guides to pursue precision metrology goals much benefit can be derived from patterning the waveguides on a rigid substrate. First, beamsplitters can be precisely and reproducibly fabricated. Second, the enclosed area of an interferometer can be precisely controlled. Third, the use of well-established lithographic techniques means that progress on individual optical elements (a beamsplitter, or a monochromator, for instance) can be rapidly extended to multi-component experiments. Mirrors based on micro-patterned wires have already been introduced . We report magnetic guiding by a pair of parallel wires produced on a glass substrate by photolithography and subsequent electro-plating. Intricate two-dimensional guiding structures are easily produced by this manufacturing technique; in the present case, it allows us to demonstrate guiding around curves in a 10 cm long guide .
We guide weak-field-seeking atoms along a one-dimensional magnetic-field minimum. Our magnetic field is produced by two parallel wires with equal currents in the same direction. The track consists of two $`100\times 100`$ $`\mu \mathrm{m}`$ wires spaced 200 microns from center to center, providing a 100 $`\mu \mathrm{m}`$ space between the wires. The resulting magnetic field is zero at the center between the wires and increases linearly outward. A small longitudinal field is applied to prevent the field magnitude from vanishing at the track center. The maximum transverse guiding potential increases linearly with applied current. The transverse magnetic-field gradient around the center is proportional to the track current and inversely proportional to the spacing between the wires. The wire spacing, applied wire current, and the longitudinal velocity of the guided atoms determine the minimum radius of curvature around which the atoms can be bent.
Our experimental apparatus consists of two chambers connected by a 2-inch-diameter steel tube that holds the substrate of the wire guide as shown in figure 1. The chambers are evacuated by separate ion pumps that maintain the pressure in each chamber typically at $`10^9`$ Torr. The source chamber provides a beam of laser-cooled atoms, and the detection chamber houses a hot wire and channeltron electron multiplier to measure the atom flux.
A modified magneto-optical trap (MOT) in the source chamber produces the beam of laser-cooled atoms . A diode laser in a master-oscillator power-amplifier configuration (MOPA) provides 350 mW of single-frequency light tuned near the $`5S_{1/2}(F=2)5P_{3/2}(F^{}=3)`$ transition in rubidium for trapping and cooling in the MOT. This light is divided into three beams, which are directed into the chamber along orthogonal axes, and retro-reflected to supply cooling along all directions. A 30-mW external-cavity diode laser supplies light tuned to the $`5S_{1/2}(F=1)5P_{3/2}(F^{}=2)`$ transition to repump atoms that fall into the $`F=1`$ ground state back into the cycling transition. A 500-$`\mu \mathrm{m}`$ hole is drilled in the center of one of the retro-reflecting mirrors, and this mirror is placed inside the vacuum chamber. Thus, one of the six confining laser beams has a dark region in the center of its cross-section. The radiation-pressure imbalance for atoms in the MOT that enter into the shadow of the hole accelerates those atoms toward and then through the hole in the mirror. The resulting atomic beam is referred to as a low-velocity intense source (LVIS) . Our observations show that $`90\%`$ of the LVIS flux atoms are optically pumped into the $`F=1`$ ground state by the MOT light. We observe that roughly $`50\%`$ are in the $`m_F=0`$ state and the rest of the atoms are roughly equally divided between the two $`m_F=\pm 1`$ sublevels. Therefore, only $`25\%`$ of LVIS atoms are in the correct state to be guided. Typically, we measure an overall LVIS flux of $`510^8`$ atoms/sec and a beam brightness of $`510^{12}\mathrm{atoms}/\mathrm{sr}\mathrm{sec}`$. We estimate the transverse-velocity distribution entering our guide to be about $`v_t=5.0\pm 2.0`$ cm/sec. A time-of-flight measurement found the longitudinal velocity of LVIS to be $`v_l=10.1\pm 2.0`$ m/sec .
Our tracks were manufactured by Metrigraphics using photolithography and electro-plating techniques. A layer of photoresist is applied on top of a 3-$`\mu \mathrm{m}`$-thick layer of copper deposited onto a $`10\times 10`$ cm glass substrate (Fig. 2). This photoresist is then exposed through a mask and removed where the tracks will be grown. Using an electro-plating technique, the tracks are grown through the gaps in the photoresist to a height of 100 microns. The excess photoresist and 3-$`\mu \mathrm{m}`$ copper layer are removed, leaving behind a track structure. The final track structure extends $`10`$ cm. We solder current-feeding wires to connection pads on the glass substrate.
The tracks are aligned with the mirror hole before the vacuum chambers are assembled. We define the track axis as the line joining the beginning and the end of the track (Fig. 2). The beginning of the track sits 1 mm behind the mirror hole and its axis is aligned parallel to the mirror axis (Fig. 1). The magnetic guide starts with a 1.2 cm straight region followed by a 1.9-cm-long curve to the right with a 15 cm radius of curvature (Fig. 2). The curvature is then reversed for a 3.8-cm-long left curve. This region is followed by a 1.9-cm-long right curve, completing the three alternating curves and leading back to the track axis. The three curves lead atoms around a bend, diverting their trajectory by 2 mm transverse to the track axis. After the bend, the guide confines atoms for another 1.2 cm to a straight trajectory before they exit the magnetic guide and travel 7 cm through free space to the detector hot wire. We place a glass barrier halfway along the track axis (Fig. 2) to block out direct LVIS flux. Guided atoms are led around this barrier by the magnetic guide and can be detected downstream. For our guiding experiment we run 35-msec-long current pulses of up to 4.5 A through the two wires. We choose short current pulses to prevent the substrate from overheating, allowing us to run larger guiding currents than continuous currents would allow.
After exiting the guide, atoms are ionized by the hot wire and the subsequent ions are then detected by the channeltron. The 70-$`\mu \mathrm{m}`$-diameter hot wire placed $`7`$ cm from the output of the magnetic guide intercepts a small fraction of the diverging atomic beam. We determine the total flux from our guide by integrating over the atom-beam profile. At a current of 3.0 A, we guide up to $`210^6`$ atoms/sec.
Figure 3 shows the atom-flux dependence on the track wire current. For low currents ($`<`$0.7 A), the guiding potential should be sufficient to confine the initial transverse-velocity distribution, but it does not provide sufficient force to bend the atoms around the curve; hence no flux is observed at the detector (Fig. 3). Above the estimated centripetal track current threshold of 0.7 A, atoms with low longitudinal velocity are guided. At currents above 2.3 A, the flux saturates as we have sufficient magnetic gradient to guide all longitudinal velocities around the bend.
In our design, the magnetic-field minimum is 50 $`\mu \mathrm{m}`$ above the substrate. By lowering the position of the magnetic-field minimum, and thus the guiding center of the track, to near or below the substrate, we should be able to guide the atoms into the glass. We do this by applying a bias field transverse to the track and parallel to the substrate with a pair of Helmholtz coils. Depending on the polarity of our Helmholtz coils, the minimum is either raised or lowered. This variable bias field is set up at the final 1-2 cm stretch of our guide, after the atoms have negotiated the bend. We observe that for one polarity of the coils the guided flux changes by $`10\%`$, but as the bias field polarity is reversed, the flux is completely eliminated (Fig. 4). In the first case, the bias field lifts the field minimum out of the tracks, but atoms still make it to the hot wire, although they are shifted vertically. In the second case, we lower the field minimum and atoms are guided into the substrate surface. Once atoms touch the substrate, they stick and can no longer be guided. As expected, larger track currents require larger bias fields to push atoms into the substrate (Fig. 4). The bias field required to entirely squelch the guided-atom flux is directly determined by the track current. The three track currents of 1.25, 2.0 and 3.0 A should correspond to squelching bias fields of 21, 33, and 49 Gauss, respectively. Our bias-field measurement shows that for the above track currents 21, 34, and 43 Gauss are necessary to cut off the flux, which is in good agreement with the estimates .
We measure the guided atoms’ transverse velocity profile by translating the hot wire to map out the spatial extent of the atom beam as it diverges from the exit of the track (Fig. 5). We calculate that the atoms’ emergence from the confining fields of the tracks is almost completely non-adiabatic—the transverse kinetic energy of the emerging beam should thus be a faithful reflection of the transverse kinetic energy in the guide. The virial theorem tells that the mean potential energy in the linear confining field should be twice the mean kinetic energy. The total mean energy per transverse dimension of the atoms in the guide should thus be three times the observed kinetic energy of the emerging atoms. At a track current of 2 A, we can determine that the guided atoms have $`3\times 42\mu \mathrm{K}=126\mu \mathrm{K}`$ total mean energy per transverse dimension (Fig. 5). This is smaller than the lowest point on the rim of the confining potential, which is 1.1 mK.
In the discussion above we apply a bias field parallel to the substrate and perpendicular to the track. Similarly, we can apply a transverse bias field in the vertical direction, perpendicular to the substrate. In this configuration the guided atom beam is moved close to one of the wires and the highest-energy component of the beam is ”skimmed off”. Figure 5 shows a case in which the mean energy in the corresponding direction has been reduced by a factor of 3.1 using a 43 Gauss bias field.
Due to the non-adiabatic loading into our tracks, atoms entering the center portion of the guide are more likely to be guided. The effective aperture through which atoms can enter the guide is consequently smaller than the physical wire separation. Comparing the estimated source flux to the number of atoms detected after the guide, we estimate that $`25\%`$ of the $`m_F=1`$ atoms that hit the $`100\times 100`$ $`\mu \mathrm{m}`$ opening are detected.
In summary, we have guided a beam of laser-cooled atoms between two $`100\times 100`$ $`\mu \mathrm{m}`$ wires. We are able to bend the atoms’ trajectory around 15-cm-radius curves. The guiding-current threshold agrees with our theoretical prediction to within our uncertainty from the atoms’ longitudinal velocity. We demonstrate that moderate current densities give guiding potentials of several milli Kelvin. No external cooling for the wires was necessary. Our magnetic guide design has great promise for applications in atom interferometers due to its versatility and simplicity. We tested the tracks with current pulses up to 8 A before substrate heating became a problem. This test indicates that we can reduce the radius of curvature by a factor of 4 to achieve a 3-4 cm track radius. In future experiments with larger currents, we hope to guide the atoms around a full 360 degree bend, which is an important step for a possible future Sagnac interferometer using magnetic-field confinement. Photolithographic technology will provide a reproducible method for producing an atom beamsplitter in future atom-interferometer applications.
The authors would like to thank Carl Wieman and Eric Abraham for helpful discussions. This work was made possible by funding from the Office of Naval Research (Grant No. N00014-94-1-0375) and the National Science Foundation (Grant No. Phy-95-12150).
|
no-problem/9908/physics9908036.html
|
ar5iv
|
text
|
# Golden Section and the Art of Painting
## 1 Introduction
In mathematics, the Golden Section is a geometric proportion created by a point C on a segment of line AB when AC/AB=CB/AC, as shown in Fig. 1.
This ratio has the value $`\mathrm{\Phi }`$=0.618…
Since the times of antiquity many philosophers, artists and mathematicians have been preoccupied by the Golden Section, which the writers of the Renaissance have called the ”Divine Proportion”. The mathematician Lucas Pacioli has characterized the Golden Section as aesthetically satisfying and wrote on this theme the treaty ”Divina Proportione”. It is largely accepted that a rectangle having the sides in this ratio has special aesthetic qualities <sup>1</sup>. Moreover the Golden Section has been used as an ideal proportion on which the pattern of lines and shapes in the composition of a painting should be based. Taking this idea as a point of departure in this paper was done a statistical study on a series of paintings, belonging to various authors and from different periods to see how the Golden Section is applied in painting. The ratio between the sides of the paintings chosen by these painters is regarded as the most appropriate and beautiful proportion. Let’s remark that 1/$`Phi`$=1.618 is also related to the golden section by the relation $`1/\mathrm{\Phi }=\mathrm{\Phi }+1`$.
## 2 Statistical study of paintings
It was done a statistical study on 565 works of art of different great painters: Bellini, Caravaggio, Cesanne<sup>2</sup>, Goya, van Gogh, Delacroix, Pallady (Romanian painter), Rembrandt, Toulouse-Lautrec. It was calculated the ratio of the 2 dimensions of painting: the longer part to the shorter part of the painted rectangle. In Table 1 are given the average values and the errors of the average for the ratio of the sides for various paintings of the painters studied. The paintings considered in this statistics have been selected from the specified references, where the sides of the paintings have been indicated.
Assuming that all the painters under discussion enter in a statistics with equal weights, in Fig. 2a is shown the total distribution, for the number of paintings N= 565. The average value obtained for the ratio of the sides is
1.34 $`\pm `$ 0.12.
This value, determined experimentally, is the result of the intuitive choice of great creators of art and is significsantly different from the value of the Golden Section $`1/\mathrm{\Phi }`$=1.618, which is a theoretical ratio, obtained from an abstract, mathematical theory, which supposedly ought to impress on a painting a supreme harmony.
In Fig. 2b is illustrated the ratio L/l=1.34, lying on the maximum of the distribution, by the painting of Toulouse-Lautrec, ”La Goulue entering at Moulin Rouge”, dated 1891-1892, Museum of Modern Art from New York, having the dimensions 79.4 x 59 cm, and alongside this painting it is drawn a rectangle whose sides are in the Golden Section.
References
1. Peter B. Norton, Josph J. Esposito, The New Encyclopaedia Britannica,
15<sup>th</sup> Edition, 1995 2. Nicolas Pioch, WebMuseum Data Base
|
no-problem/9908/cond-mat9908368.html
|
ar5iv
|
text
|
# Low Temperature AC Conductivity of Bi2Sr2CaCu2O8+δ
## Abstract
We report measurements of anamolously large dissipative conductivities, $`\sigma _1`$, in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> at low temperatures. We have measured the complex conductivity of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> thin films at 100-600 GHz as a function of doping from the underdoped to the overdoped state. At low temperatures there exists a residual $`\sigma _1`$ which scales with the $`T=0`$ superfluid density as the doping is varied. This residual $`\sigma _1`$ is larger than the possible contribution to $`\sigma _1`$ from a thermal population of quasiparticles (QP) at the d-wave gap nodes.
preprint: HEP/123-qed
The conductivity of a superconductor combines two components: the superfluid and the quasiparticle (QP) excitations. As one cools a superconductor from $`T_C`$ towards $`T=0`$, the fraction of the charge carriers excited out of the superfluid decreases towards zero. Therefore, at low temperatures the finite frequency conduction is increasingly dominated by the response of the superfluid rather that the dissipative QP conductivity. This is exemplified in YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (YBCO) by the low temperature behavior of $`\sigma _1`$. Hosseini, et al. find $`\sigma _1(T)`$ to exhibit a broad peak below $`T_C`$ and approach zero $``$ linearly in the $`T=0`$ limit. The peak in $`\sigma _1(T)`$ is thought to result from a competition between the increasing QP lifetime, $`\tau _{QP}`$, and the decreasing QP density, $`n_{QP}`$, as $`T0`$.
In the cuprate Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> (BSCCO), however, very different low temperature behavior has been observed. $`\sigma _1`$ of optimally doped BSCCO single crystals at 35 GHz approaches a constant value which is $`10`$ times $`\sigma _{1Normal}`$ at the lowest temperatures measured ($``$ 5K). Using Time-Domain-THz-Spectroscopy we extended measurements of the complex conductivity, $`\sigma =\sigma _1+i\sigma _2`$, in BSCCO to higher frequencies and different carrier concentrations. We studied a set of BSCCO thin films varying from under ($`T_C=33K`$) to over-doped($`T_C=74K`$). Figure 1 is a semi-log plot of $`\sigma _1`$(140 GHz) versus $`T`$. $`\sigma _1`$ is seen to have two components: a peak centered near $`T_C`$ superimposed on a broad background. This peak, less prominent as doping increases, results from thermally generated phase fluctuations. The background $`\sigma _1`$ is seen, at all dopings, to increase to a value well above $`\sigma _{1Normal}`$ and does not decrease to zero as $`T0`$.
In the optimal and overdoped samples the superfluid density, $`\rho _S`$, is seen to decrease linearly with $`T`$. If we ascribe this low $`T`$ decrease in superfluid density $`\mathrm{\Delta }\rho _ST`$ with a $`n_{QP}=n_{Total}(T/T_C)`$ , then the fraction of $`\sigma _1`$ which can be attributed to the QP, $`\sigma _{1QP}`$, must tend to zero as $`T0`$. We illustrate this using the Drude form for the QP conductivity, $`\sigma _{1QP}=n_{QP}\tau _{QP}/(1+(\omega \tau _{QP})^2)`$, with $`n_{QP}=n_{Total}(T/T_C)`$. Using our value of $`\sigma _{1Normal}`$ and the measured normal state value of $`\tau _{QP}`$ we can determine $`n_{Total}`$. Following the recent ARPES measurements by Valla et al. we assume the $`T`$ dependence of $`\tau _{QP}`$ to be $`1/\tau _{QP}(T)T+T_0`$ .
Figure 2 shows, for the $`T_C=85K`$ sample at 200 GHz, $`\sigma _{1QP}`$ from this model, $`\sigma _{1Total}`$, and the residual conductivity not due to the QP, $`\sigma _{1Res}\sigma _{1Total}\sigma _{1QP}`$. $`\sigma _{1QP}`$ accounts for all of $`\sigma _1`$ down to 100 K, the temperature identified in previous work with the onset of identifiably non-zero bare $`\rho _S`$ . $`\sigma _{1Res}`$ appears at 100 K, exhibits the thermal phase fluctuation peak at 80 K and then rises linearly in $`T`$ over the rest of the temperature range down to zero.
Finally, we find $`\sigma _{1Res}`$ to be proportional to $`\rho _S(T=0)`$ over our range of doping from the overdoped to the $`T_C=50K`$ underdoped sample. The width of $`\sigma _{1Res}(\omega )`$ in frequency seems to be $``$ 200 GHz for all of the samples studied.
If, instead of the Valla et al. $`\tau _{QP}(T)`$, we use a more conventional form $`\tau _{QP}(T)T^\alpha +T_0`$, $`\sigma _{1QP}`$ develops a peak below $`T_C`$ and $`\sigma _{1Res}(T)`$ is no longer linear in T from 80K to 5K. However, $`\tau _{QP}`$ still approaches a small constant as $`T0`$ and the other result is unchanged: $`\sigma _{1QP}`$ is still a small fraction of $`\sigma _{1Total}`$ at low temperatures.
Since $`\sigma _1`$ is too large to be ascribed entirely to the QP regardless of the choice for $`\tau _{QP}`$, and $`\sigma _{1Res}\rho _S(T=0)`$ as one might expect for a collective excitation in the superfluid, we speculate that this low $`\omega `$, low $`T`$ peak in $`\sigma _1`$ could signal observable quantum fluctuations in the BSCCO system.
###### Acknowledgements.
This work was supported by NSF, DOE and ONR
|
no-problem/9908/hep-ph9908466.html
|
ar5iv
|
text
|
# Single target-spin asymmetries in semi-inclusive pion electroproduction on longitudinally polarized protonstalk presented by K. Oganessyan at the Workshop on the structure ot the Nucleon (N99), Frascati, June 7-9, 1999.
## Abstract
We evaluate the single target-spin sin$`\varphi _h`$ and sin$`\mathrm{\hspace{0.17em}2}\varphi _h`$ azimuthal asymmetries in the semi-inclusive deep inelastic lepton scattering off longitudinally polarized proton target under HERMES kinematic conditions. A good agreement with the HERMES data can be achieved using only the twist-2 distribution and fragmentation functions.
Significant single-spin asymmetries have been observed in experiments with transversely polarized proton and anti-proton beams . Recently new experimental results on azimuthal asymmetries became available. Specifically, the first measurements of single target-spin azimuthal asymmetries of pion production in semi-inclusive deep inelastic scattering (SIDIS) of leptons off a longitudinally polarized target at HERMES and off a transversely polarized target at SMC , and the observation of the azimuthal correlations for particles produced from opposite jets in $`Z`$ decay at DELPHI .
In this note we present estimates of the single spin azimuthal asymmetry in the SIDIS on a longitudinally polarized nucleon target for the HERMES kinematic conditions. Our approach is based on the parton model description of polarized SIDIS . The cross-section contains the $`(1/Q)^0`$-order terms coming from leading dynamical twist-two distribution and fragmentation functions (DF’s and FF’s) as well as $`(1/Q)`$-order kinematic twist-three terms arising due to the intrinsic transverse momentum of the quark in the nucleon. We will neglect the $`(1/Q)`$-order contributions of the higher twist DF’s and FF’s obtained in . Thus, our approach is similar to that of in describing the cos$`\varphi _h`$ asymmetry in unpolarized SIDIS.
Let $`k_1`$ ($`k_2`$) be the initial (final) momentum of the incoming (outgoing) charged lepton, $`Q^2=q^2`$, $`q=k_1k_2`$ – the momentum of the virtual photon, $`P`$ and $`P_h`$ ($`M`$ and $`M_h`$) – the target and final hadron momentum (mass), $`x=q^2/2(Pq)`$, $`y=(Pq)/(Pk_1)`$, $`z=(PP_h)/(Pq)`$, $`P_{hT}`$ ($`k_{1T}`$) – the hadron (lepton) transverse with respect to virtual photon momentum direction and $`\varphi _h`$ – the azimuthal angle between $`P_{hT}`$ and $`k_{1T}`$ around the virtual photon direction. Note that the azimuthal angle of the transverse (with respect to the virtual photon) component of the target polarization, $`\varphi _S`$, is equal to 0 ($`\pi `$) for the target polarized parallel (antiparallel) to the beam (Fig. 1).
We use the approach developed in and consider the cross-section integrated with different weights depending on the final hadron transverse momenta $`w_i(P_{hT})`$ <sup>1</sup><sup>1</sup>1More details can be found in .:
$$\mathrm{\Sigma }_i=\frac{Q^2y}{2\pi \alpha ^2}d^2P_{hT}w_i(P_{hT})𝑑\sigma ,$$
(1)
with $`w_1(P_{hT})=1`$, $`w_2(P_{hT})=|P_{hT}|\mathrm{sin}\varphi _h/M_h`$ and $`w_3(P_{hT})=|P_{hT}|^2\mathrm{sin}2\varphi _h/2MM_h`$. Considering only the twist-two contributions, we have:
$$\mathrm{\Sigma }_1=(1+(1y)^2)f_1(x)D_1(z),$$
(2)
where $`f_1(x)`$ and $`D_1(z)`$ are the usual unpolarized DF’s and FF’s. Moreover
$$\mathrm{\Sigma }_2=\mathrm{\Sigma }_{2L}+\mathrm{\Sigma }_{2T},$$
(3)
where
$$\mathrm{\Sigma }_{2L}=8S_L\frac{M}{Q}(2y)\sqrt{1y}zh_{1L}^{(1)}(x)H_1^{(1)}(z)$$
(4)
is the $`(1/Q)`$-order contribution from twist-two DF $`h_{1L}^{(1)}(x)`$ and FF $`H_1^{(1)}(z)`$ arising due to intrinsic transverse momentum and
$$\mathrm{\Sigma }_{2T}=2S_{Tx}(1y)zh_1(x)H_1^{(1)}(z)$$
(5)
is arising due to the small ($`(1/Q)`$) transverse component of the target polarization ($`S_{Tx}`$) . Finally
$$\mathrm{\Sigma }_3=8S_L(1y)z^2h_{1L}^{(1)}(x)H_1^{(1)}(z).$$
The weighted cross sections involve the $`p_T^2`$ ($`k_T^2`$) moment of the DF’s (FF’s), defined as
$$h_{1L}^{(1)}(x)d^2p_T\left(\frac{p_T^2}{2M^2}\right)h_{1L}^{}(x,p_T^2),$$
(6)
$$H_1^{(1)}(z)z^2d^2k_T\left(\frac{k_T^2}{2M_h^2}\right)H_1^{}(z,z^2k_T^2).$$
(7)
We note that $`h_{1L}^{}(x)`$ and $`h_1(x)`$ describe the quark transverse spin distribution in the longitudinally and transversely polarized nucleon respectively, while $`H_1^{}(z)`$ describes the analyzing power of transversely polarized quark fragmentation (Collins effect) .
The single target-spin asymmetries for SIDIS on a longitudinally polarized target are defined as
$$\frac{|P_{hT}|}{M_h}\mathrm{sin}\varphi _h\frac{d^2P_{hT}\frac{|P_{hT}|}{M_h}\mathrm{sin}\varphi _h\left(d\sigma ^+d\sigma ^{}\right)}{d^2P_{hT}\left(d\sigma ^++d\sigma ^{}\right)},$$
(8)
$$\frac{|P_{hT}|^2}{MM_h}\mathrm{sin}2\varphi _h\frac{d^2P_{hT}\frac{|P_{hT}|^2}{MM_h}\mathrm{sin}2\varphi _h\left(d\sigma ^+d\sigma ^{}\right)}{d^2P_{hT}\left(d\sigma ^++d\sigma ^{}\right)},$$
(9)
where $`+()`$ denotes positive (negative) longitudinal polarization of the target. Using $`\mathrm{\Sigma }_{1,2,3}`$ one can see that for both polarized and unpolarized lepton these asymmetries are given by
$$\frac{|P_{hT}|}{M_h}\mathrm{sin}\varphi _h(x,y,z)=\frac{\mathrm{\Sigma }_2(x,y,z)}{\mathrm{\Sigma }_1(x,y,z)}$$
(10)
$$\frac{|P_{hT}|^2}{MM_h}\mathrm{sin}2\varphi _h(x,y,z)=\frac{\mathrm{\Sigma }_3(x,y,z)}{\mathrm{\Sigma }_1(x,y,z)}.$$
(11)
We use the non-relativistic approximation $`h_1(x)=g_1(x)`$, the upper limit from Soffer’s inequality $`h_1(x)=(f_1(x)+g_1(x))/2`$, and the relation between $`h_{1L}^{(1)}(x)`$ and $`h_1(x)`$ obtained by neglecting the interaction dependent twist-three part of the DF and the term proportional to the current quark’s mass:
$$h_{1L}^{(1)}(x)=x^2_x^1𝑑y\frac{h_1(y)}{y^2}.$$
(12)
We took the parameterisations of DF’s $`f_1(x)`$ and $`g_1(x)`$ from Ref. .
To calculate the T-odd FF $`H_1^{(1)}(z)`$ we adopt the Collins parameterisation for the analyzing power of transversely polarized quark fragmentation
$$A_C(z,k_T)\frac{|k_T|}{M_h}\frac{H_1^{}(z,k_T^2)}{D_1(z,k_T^2)}=\frac{M_C|k_T|}{M_C^2+k_T^2}$$
(13)
and assume a Gaussian parameterisation of the unpolarized FF with $`z^2k_T^2=b^2`$ (in the numerical calculations we use $`b=0.5`$ GeV ). For $`D_1^{\pi ^\pm }(z)`$ we use the parameterisation from Ref. .
The $`A_{UL}^{\mathrm{sin}\varphi _h}(x)`$ asymmetry for $`\pi ^\pm `$ production on the proton target is obtained from the defined asymmetry (Eq.(10)) by the relation $`A_{UL}^{\mathrm{sin}\varphi _h}\frac{2M_h}{P_{hT}}\frac{|P_{hT}|}{M_h}\mathrm{sin}\varphi _h`$ and is presented in Fig. 2 in comparison with preliminary HERMES data . The data corresponds to $`Q^21`$ GeV<sup>2</sup>, $`E_\pi 4`$ GeV, and the ranges $`0.2z0.7`$, $`0.2y0.8`$. The theoretical curves are calculated by integrating over the same ranges with $`P_{hT}=0.52`$ GeV, $`P_{hT}^2=0.35`$ GeV<sup>2</sup>. These average values of $`P_{hT}`$, $`P_{hT}^2`$ are obtained in mentioned kinematics assuming a Gaussian parameterisation of DF’s and FF’s with $`a=0.7`$ GeV ($`p_T^2=a^2`$. From Fig. 2 one can see that a good agreement with HERMES data can be achieved by varying $`h_1(x)`$ and $`M_C`$. Note that the main effect comes from the $`\mathrm{\Sigma }_{2L}`$ term, the contribution of $`\mathrm{\Sigma }_{2T}`$ is about $`20÷25\%`$.
We calculate the $`sin2\varphi _h`$-weighted asymmetry in the same manner as well and show that the amplitude of the $`sin2\varphi _h`$ modulation is about a factor of 2-3 smaller than that of the $`sin\varphi _h`$ modulation (see Fig. 3) in the HERMES kinematics. Note that the ratio of these asymmetries is almost independent of the choice of $`h_1(x)`$ and $`M_C`$.
In conclusion, the sin$`\varphi _h`$ and sin$`\mathrm{\hspace{0.17em}2}\varphi _h`$ single target-spin asymmetries of SIDIS off longitudinally polarized protons related to the time reversal odd FF was investigated. It was shown that the main $`(1/Q)`$-order contribution to the spin asymmetry arises from intrinsic $`k_T`$ effects similar to the $`cos\varphi _h`$ asymmetry in unpolarized SIDIS. A good agreement with the HERMES data can be achieved using only the twist-2 DF’s and FF’s. The $`(1/Q)^0`$-order $`sin2\varphi _h`$ asymmetry, in contrast to the naive expectations, is suppressed comparing to the $`(1/Q)`$-order $`sin\varphi _h`$ asymmetry at HERMES kinematics.
The authors would like to thank D. Boer, R. Jakob, and P. Mulders for useful discussions. The work of (K.O) and (H.A) was in part supported by the INTAS contributions (contract number 93-1827) from the European Union.
|
no-problem/9908/cond-mat9908222.html
|
ar5iv
|
text
|
# A study of supercooling of the disordered vortex phase via minor hysteresis loops in 2H-NbSe2 G. Ravikumar1,∗, P. K. Mishra1, V. C. Sahni1, S. S. Banerjee2, A. K. Grover2, S. Ramakrishnan2, P. L. Gammel3, D. J. Bishop3, E. Bucher3, M. J. Higgins4 and S. Bhattacharya2,4,∗ 1TPPED, Bhabha Atomic Research Centre, Mumbai-400085, India 2 Dept. of Condensed Matter Physics and Materials Science, Tata Institute of Fundamental Research, Mumbai-400005, India 3Bell Laboratories, Lucent Technologies, Murray Hill, NJ 07974 4 NEC Research Institute, 4 Independence Way, Princeton, NJ 08540 We report on the observation of novel features in the minor hysteresis loops in a clean crystal of 𝑁𝑏𝑆𝑒₂ which displays a peak effect. The observed behavior can be explained in terms of a supercooling of the disordered vortex phase while cooling the superconductor in a field. Also, the extent of spatial order in a flux line lattice formed in ascending fields is different from (and larger than) that in the descending fields below the peak position of the peak effect; this is attributed to unequal degree of annealing of the state induced by a change of field in the two cases. PACS numbers :64.70 Dv, 74.25 Ha 74.60 Ge, 74.60 Jg
## I INTRODUCTION
A variety of recent transport, magnetic and structural studies in weakly pinned superconductor 2H-NbSe<sub>2</sub> support the view that the peak effect (PE) phenomenon in the critical current density ($`J_c`$) marks an amorphisation of the flux line lattice (FLL) \[2-10\]. In a varying field experiment, for example, , $`J_c(H)`$ begins to increase from an initial low value at the onset field $`H_{pl}^+`$, reaches a maximum at the peak field $`H_p`$ and eventually collapses below the measurable limit above the irreversibility field $`H_{irr}`$ ( $`<`$ $`H_{c2}`$). The vortex matter is thought to undergo a transformation from a state with a high spatial order for $`H`$ $`<`$ $`H_{pl}^+`$ to a highly disordered state for $`H`$ $`>`$ $`H_p`$. This interpretation is usually rationalized within the Larkin-Ovchinnikov (LO) collective pinning formalism , where the Larkin volume $`V_c`$ ($``$ $`R_c^2`$ $`L_c`$, where $`R_c`$ and $`L_c`$ are the radial and longitudinal correlation lengths, respectively) is a measure of the spatial extent of order in FLL and the critical current density is determined through the relation, $`\mu _0HJ_c(H)`$ $``$ $`(n_p<f^2>/V_c)^{1/2}`$, where $`n_p`$ and $`f`$ are pin density and the elementary pinning force parameter, respectively. $`J_c`$ measurements therefore can reveal the extent of spatial order in the FLL. Magnetization hysteresis measurements are a convenient tool for estimating $`J_c(H)`$ and thus for detecting the occurrence of phase transformations and associated changes in the vortex correlations across the PE regime.
Recently, the PE phenomenon has received a great deal of attention due to a characteristically rich phenomenology that accompanies it. Prominent among them is the marked history dependence in the critical current. In the LO scenario, this translates into a history dependence of the Larkin volume $`V_c`$, since quantities such as $`f`$ and $`n_p`$ cannot be history dependent. The history dependence of the correlations implies strong metastability and is thus a hallmark of disorder in condensed matter systems. Recent theoretical studies emphasizing the role of quenched random disorder in producing novel disordered (glassy) phases (see, for instance, Refs. 12-14) further illustrate the need to understand and ultimately unravel the complex effects of disorder readily observed in the PE regime. Ravikumar et al have experimentally demonstrated via dc magnetization technique the presence of a highly disordered vortex state when a sample is field cooled (FC) in $`H`$ $`<`$ $`H_p`$. In crystals of $`2H`$-$`NbSe_2`$ and $`CeRu_2`$ with comparable levels of effective pinning, they had shown that the critical current density in the FC state is larger than that in the zero field cooled (ZFC) state for $`H`$ $`<`$ $`H_p`$ , i.e., $`J_c^{FC}(H)`$ $`>`$ $`J_c^{ZFC}(H)`$. They had also shown that the disordered FC state could be annealed to the ordered ZFC like state when the sample was subjected to a small change in magnetic field. Independently, a different type of history effects has been reported in a wide variety of polycrystalline and single crystal samples of pure and doped $`CeRu_2`$ showing PE phenomenon. In these cases, the minor magnetization curves, starting at a field $`H`$ on the forward branch of the hysteresis loop (such that $`H_{pl}^+`$ $`<`$ $`H`$ $`<`$ $`H_p`$), were reported to saturate without merging with the reverse leg of the hysteresis loop. The minor curves starting on the reverse branch, on the other hand, merge with the forward branch of the hysteresis loop. Roy et al attribute the observed behavior to an anomalous nature of PE phenomenon in the mixed valent superconducting system $`CeRu_2`$, in contrast with, and as distinct from, the conventional PE in $`2H`$-$`NbSe_2`$ and most other weakly pinned low $`T_c`$ superconductors . They propose that this novel behavior of the minor loops reflects a thermodynamic evidence for a first order transformation to a new phase caused by the positional dependence of the order parameter in superconductors with large normal state paramagnetism, like, heavy fermion superconductors, mixed valent rare earth systems, etc. . This is in contrast to the explanation based on metastability effects in the vortex matter caused by quenched disorder.
Very recently, a similar effect with the minor hysteresis curves has been observed in a cuprate superconductor as well . The microscopics in the cuprate system appears to bear little resemblance to that in the mixed valent systems such as CeRu<sub>2</sub>. We propose that the above sets of experiments exemplify the ubiquitous nature of thermomagnetic history dependence of $`J_c`$ in superconductors in general and thus require an explanation that is independent of the microscopics relevant for superconductivity in the diverse systems in which these effects are present.
In this paper, through detailed measurements of minor magnetisation curves on a clean single crystal of $`2H`$-$`NbSe_2`$ (belonging to the category of most weakly pinned samples of type II superconductors ), we present an understanding, based on the LO collective pinning description for the observed path dependence in the critical currents and vortex correlations. We invoke the notion of a disorder-assisted supercooling of a metastable disordered phase, that is otherwise thermodynamically stable only above $`H_p`$. We further propose that changes in the applied field act as a driving force that helps to anneal the system, often partially. As a result, systems with different field histories are often in metastable states with different degrees of annealing and thus with different values of vortex correlations, leading to the different critical currents, as observed experimentally. In the present work, three types of isothermal minor hysteresis loops were studied : (I) Decreasing field after cooling the sample in a field (FC-REV), (II) and (III) : decreasing /increasing field from a given point on the forward/reverse leg of the envelope hysteresis curve. Fig.1 shows a schematic view of all these loops for the case where $`J_c(H,T)`$ is single-valued, independent of the field/temperature history. In all such cases, the minor curves merge into the envelope curve while always remaining within it, without any overshoot effects. We show below that violations of this standard scenario provide an understanding of the aforementioned anomalous behavior.
## II EXPERIMENTAL
We carried out dc magnetization measurements using a Quantum Design (QD) SQUID magnetometer (Model MPMS5) on a $`2H`$-$`NbSe_2`$ single crystal with $`T_c`$ $``$ 7.25 K. The crystal was mounted with field applied parallel to its c-axis. The thermomagnetic history dependent measurements have been performed at 6.95 K, where a well recognized peak effect, manifesting as a sharp increase in the magnetization hysteresis, is observed at a relatively low field of about 1000 Oe. Occurrence of PE at such low fields, where the flux line lattice constant $`a_0`$ (= 1600 A) is of the same order as the range of interaction ( i.e., the penetration depth $`\lambda _{ab}`$) at 6.95K in this system , confirms that the sample has weak quenched disorder . In the field range 1$``$2 kOe, the inhomogeneity experienced by the sample in a 2 cm full-scan in a QD SQUID magnetometer is of the order of 0.1 Oe. This value is much smaller than the threshold field $`H_{II}`$ required to change the sign of the induced shielding currents throughout the sample. $`H_{II}`$ value is estimated by measuring the minor magnetisation curves for fields above $`H_p`$ and was found to be about 10 Oe. Thus, the field inhomogeneity in a full scan of 2 cm does not introduce any error in the measured magnetization values, and we have recorded all the present data with a 2 cm scan length instead of half scan technique utilized earlier by Ravikumar et al , in their study of $`CeRu_2`$ and more strongly pinned sample of 2H-NbSe<sub>2</sub>.
## III RESULTS AND DISCUSSION
In Fig.2, we show the magnetization hysteresis loop in the present 2H-NbSe<sub>2</sub> crystal at 6.95 K, indicating the onset field of PE on the ascending field cycle ($`H_{pl}^+`$ $``$ 800 Oe), the peak field $`H_p`$ ($``$ 1000 Oe) and the irreversibility field $`H_{irr}`$ ($``$ 1250 Oe). Within the LO collective pinning description, the Larkin volume $`V_c`$ begins to shrink at $`H_{pl}^+`$ and FLL reaches nearly amorphous (though pinned) state at $`H_p`$ . Note that the onset of PE is much sharper (see inset(ii)) on the ascending field cycle than on the descending field cycle. This immediately shows that at a given field value in the PE region, $`J_c`$ and thus $`V_c`$ in the vortex state are not the same on the ascending and descending field cycles in the PE regime. In what follows, we examine this non-uniqueness of the vortex correlations in greater detail.
In Fig.3(a), we show the magnetization curves measured for FC-REV case, i.e., magnetization of a field cooled sample measured in decreasing magnetic fields. The field value in which the sample was cooled was varied across the PE region. Note that in Fig.3(a), the FC-REV curve originating at $`H`$ $`>`$ $`H_p`$ merges with the reverse magnetization envelope curve in accordance with the critical state model. On the other hand, FC-REV curves, originating from a field $`H`$ $`<`$ $`H_p`$, overshoot the reverse magnetization envelope curve in a clear departure from the conventional behavior shown in Fig.1. The difference between the highest magnetisation value recorded on the FC-REV curve and the notional equilibrium magnetisation value could be taken as a measure of the critical current density in the FC state . Thus, the minor curves in the FC-REV case produce higher $`J_c`$ values than those in the conventional descending part of the envelope curve.
This anomalous behavior can be explained by assuming a supercooling of the disordered phase and the annealing effect (i.e., increase in vortex lattice correlations or growth of Larkin domain volume $`V_c`$) due to a subsequent field change. In the FC state, the FLL traverses through a pinned amorphous state as it is cooled down across the $`H_p(T)`$ line (see the phase diagram drawn in the inset (i) of Fig.2). The highly disordered vortex state that is stable above PE curve, with a large density of defects/dislocations , is then effectively supercooled when the sample is cooled to a given T in a field less than H<sub>p</sub>(T). Vortex state obtained by cooling the sample in a field H is more disordered than that at the same field value on the descending branch of the envelope loop. In the latter case, the process of lowering of field (below $`H_p`$) induces partial annealing and produces a vortex state, which is more correlated and thus has a smaller critical current than that in the field cooled state. Furthermore, annealing induced by the field change as mentioned above is also clearly seen in the FC-REV magnetization values moving towards the reverse envelope curve (see inset of Fig.3(a)).
The FC-REV curves initiated from different fields below $`H_p`$ form a family of curves. FC-REV curve originating from $`H`$ $`=`$ $`H_p`$ essentially retraces the reverse magnetisation envelope curve below $`H_p`$. As stated earlier, the vortex state at any field value below $`H_p`$ on the reverse magnetization curve is the result of a gradual healing of the disordered state existing at $`H=H_p`$. Thus, the reverse magnetization envelope curve (for $`H`$ $`<`$ $`H_p`$), in principle need not, and in practice will not, be a mirror reflection of the forward magnetization envelope curve as the specific kinetics of the annealing processes are different. That the process of healing of FLL dislocations could continue down to a field well below $`H_{pl}^+`$ (during the descending cycle) is a clear indicator of this difference. The data shows that for a given field $`H(<H_p)`$, the Larkin domains on the reverse magnetization envelope curve are smaller than those on the forward magnetization envelope curve. In other words, $`J_c`$ values on the descending field cycle are larger than those on the ascending field cycle. This difference is likely to be due to the difference in the starting configurations on the ascending and descending branches. For the latter, one starts from a much more disordered state; thus the residual disorder after a comparable level of (incomplete) annealing is more than that in the former.
Using the scenario described above, we now examine the minor magnetisation curves of the type II and III (cf. Fig.1) in the PE region of 2H$``$NbSe<sub>2</sub> and compare and contrast them with anomalous behavior of minor magnetisation curves in the PE region of $`CeRu_2`$ . In Fig.3(b), we first show the minor magnetization curves of type II, initiating from different fields lying on the forward magnetization curve. The minor magnetization curves initiated from $`H>H_p`$ and $`H<H_{pl}^{}`$ merge with the reverse magnetization curve within a field change of about 10 Oe. The threshold field $`H_{II}`$ is thus estimated from these curves to be of the order of 10 Oe. However, for $`H_{pl}^{}<H<H_p`$, the minor magnetization curves do not merge into the reverse envelope curve, because $`J_c(H)`$ values on the ascending field cycle are smaller than those on the descending field cycle, as asserted earlier. The significance of $`H_{pl}^{}`$ is that the disorder present at $`H=H_p`$ is maximally annealed at this field value on the descending cycle.
In Fig. 3(c), we show the measured minor magnetization curves of the type III, i.e., by increasing the field from different points on the reverse magnetization envelope curve. These minor curves now overshoot the forward envelope magnetisation curve when the field is increased by about 10 Oe. The annealing due to the field increase (of about 10 Oe) is inadequate to produce a comparable level of lattice order existing at the corresponding fields on the forward envelope loop. When the field is further increased, the residual disorder gets annealed. Thus the minor magnetisation curves eventually merge into the more ordered forward envelope curve. However, Roy et al reported that in $`CeRu_2`$ the minor curves initiated from the reverse magnetisation envelope curve readily merge with the forward magnetisation curve, in contrast with our data. Whether this is a significant difference, or merely a trivial one, in the comparative levels of annealing in the two instances, is unclear .
The critical current densities, $`J_c^{for}`$ and $`J_c^{rev}`$ on the ascending and descending field branches can now be estimated from the maximum width of these two sets of minor magnetisation curves. We collate, in Fig.4, the relative $`J_c(H)`$ values corresponding to three thermomagnetic histories of the sample, viz., the FC state and the states along the forward and the reverse legs of the envelope hysteresis loop. The three sets of $`J_c(H)`$ values have been estimated by taking the notional half width of the magnetization hysteresis at a given $`H`$ to be proportional to the corresponding critical current density. The data in Fig.4 can be summarized by the inequality,
$$J_c^{FC}(H)>J_c^{rev}(H)>J_c^{for}(H),$$
which is consistent with an early report of transport data in a Nb crystal by Steingart et al . In the framework mentioned above, this inequality corresponds to the least annealed state in the FC mode and the most annealed state on the forward curve, while the $`J_c^{rev}`$ is intermediate between the two.
## IV Conclusion
We propose that the PE phenomenon marks a true thermodynamic phase transformation between an ordered solid that is stable below $`H_{pl}^+`$ (on forward envelope loop) and a fully disordered vortex state that is stable above $`H_p`$. Further, this transformation is first order in character. Thus, it is possible to supercool or superheat one phase into the regime of stability of the other phase. For the PE regime, when the free energies of the ordered and disordered phases are not significantly different, the metastability effects are expected to be prominent. Moreover, thermal fluctuations are inadequate, at least below $`H_p`$, for the system to fully explore the phase space, which helps the metastability of the supercooled phase, aided by the fact that both phases have finite pinning. Substantial driving forces experienced when the field is changed allow the system to anneal (or fracture, as the case may be) towards the stable state. When the levels of annealing are incomplete and, as is often the case, unequal due to the previous thermomagnetic history, the system will typically exhibit different levels of correlations and thus different critical currents, as expected within the LO mechanism. The history dependence of critical current should thus be a generic occurrence in this regime for comparable levels of disorder. We emphasize that the proposed explanation of the various anomalous history dependent magnetization hysteresis data shown here and elsewhere are in terms of an order-disorder transformation and disorder-aided supercooling. This explanation is , prima facie, independent of the specifics of microscopic considerations consistent with the ubiquitous nature of the phenomena under consideration. It is tempting to suggest that the so-called anomalous PE in $`CeRu_2`$ may also find an explanation within the scenario described above; whether this is indeed the case remains to be concluded.
[email protected] or [email protected]. nec.com
REFERENCES
1. M. Tinkham, Introduction to Superconductivity, second edition, McGraw-Hill International Editions (1996), U.S.A., Chapter 9.
2. S. Bhattacharya and M. J. Higgins, Phys. Rev. Lett. 70, 2617 (1993); Phys. Rev. B 52, 64 (1995); M. J. Higgins and S. Bhattacharya, Physica C 257, 232 (1996) and references therein.
3. K. Ghosh, S. Ramakrishnan, A. K. Grover, G. I. Menon, G. Chandra, T. V. Chandrasekhar Rao, G. Ravikumar, P. K. Mishra, V. C. Sahni, C. V. Tomy, G. Balakrishnan, D. Mck Paul, S. Bhattacharya, Phys. Rev. Lett. 76, 4600 (1996).
4. L. A. Angurel, F. Amin, M. Polichetti, J. Aarts, P. H. Kes , Phys. Rev. B 56, 3425 (1997).
5. C. Tang, X. S. Ling, S. Bhattacharya and P. M. Chaikin, Europhys. Lett. 35 , 597 (1996).
6. W. Henderson, E. Y. Andrei, M. J. Higgins and S. Bhattacharya, Phys. Rev. Lett. 77, 2077 (1996); 80, 381 (1998).
7. S. S. Banerjee et al, Physica C 308, 25 (1998), Phys. Rev. B 58, 995 (1998); 59, 6043 (1999) and references therein.
8. T. V. Chandrasekhar Rao et al, Physica C 299, 267 (1998); BARC (India) preprint 1998.
9. G. Ravikumar et al, Phys. Rev. B 57, R11069 (1998).
10. P. L. Gammel, U. Yaron, A. P. Ramirez, D. J. Bishop, A. M. Chang, R. Ruel, L. N. Pffeifer, E. Bucher, G. D’ Anna, D. A. Huse, K. Mortensen, M. R. Eskildsen and P. H. Kes, Phys. Rev. Lett. 80, 833 (1998).
11. A. I. Larkin and Y. N. Ovchinnikov, Sov. Phys. JETP 38, 854 (1974); J. Low Temp Phys. 34, 409 (1979); A. I. Larkin, Sov. Phys. JETP 31, 784 (1970).
12. T. Giamarchi and P. Le Doussal, Phys. Rev. Lett. 72, 1530 (1994).
13. T. Giamarchi and P. Le Doussal, Phys. Rev. B 52, 1242 (1995); 55, 6577 (1997).
14. M. J. P. Gingras and D. A. Huse, Phys. Rev. B 53, 15193 (1996).
15. S. B. Roy and P. Chaddah, Physica C 273, 120 (1996); J. Phys: Condens. Matter 9, L625 (1997); S. B. Roy, P. Chaddah and S. Chaudhary, ibid 10, 4885 (1998).
16. S. Chaudhary, S. B. Roy, P. Chaddah and L. F. Cohen, Proceedings of the 41st annual DAE Solid State Physics Symposium, Universities Press, Hyderabad (India), 41, 367 (1998); P. Chaddah and S. B. Roy, Bull. Mater. Sci.,22, 275 (1999).
17. C. V. Tomy, G. Balakrishnan and D. Mck Paul, Physica C 280, 1 (1997); Phys. Rev. B 56, 8346 (1997) and references therein.
18. M. Tachiki, S. Takahashi, P. Gegenwart, M. Weiden, M. Lang, C. Geibel, F. Steglich, R. Modler, C. Paulsen, Y. Onuki, Z. Phys. B 100, 369(1996).
19. S. Kokkaliaris, P. A. J. de Groot, S. N. Gordeev, A. A. Zhukov, R. Gagnon and L. Taillefer, Phys. Rev. Lett. 82, 5116 (1999).
20. S. S. Banerjee et al, Physica B 237-238, 315 (1997).
21. G. Ravikumar et al, Physica C 298, 122 (1998) and references therein.
22. P. Chaddah, K. V. Bhagwat and G. Ravikumar, Physica C 159, 570 (1989).
23. A. K. Grover, in Studies of High Temperature Superconductors, Vol. 14, edited by A. V. Narlikar, Nova Science Publishers, Inc., Comack, NY, USA, 1995, pp.185-244.
24. P. Chaddah, ibid., pp 245-273.
25. S. S. Banerjee et al, Appl. Phys. Lett. 74, 126 (1999).
26. Inspection of Fig.3 in Ref.15 (J. Phys : Condens. Matter 10, 4885(1998)) suggests that the minor loop initiated from the reverse leg indeed overshoots the forward envelope curve even in CeRu<sub>2</sub>.
27. M. Steingart, A. G. Putz and E. J. Kramer, J. Appl. Phys. 44, 5580 (1973).
FIGURE CAPTIONS
Fig. 1. Schematic behavior of minor magnetisation curves initiated from field cooled (I) magnetisation value (M<sub>FC</sub>) and those from forward (II) and reverse (III) legs of the envelope hysteresis loop within the framework of critical state model, i.e., assuming J<sub>c</sub>(H) is uniquely defined by H. The field interval H<sub>II</sub>, corresponding to a threshold field change required to change the sign of shielding currents throughout the sample, is also indicated.
Fig.2. Magnetization hysteresis of a $`NbSe_2`$ crystal recorded at 6.95K for $`Hc`$. The inset (i) schematically shows three different paths, viz., the zero field cooled (ZFC), field cooled (FC) and descending fields from above $`H_{c2}`$. The PE line H<sub>p</sub>(T) and the upper critical field line H<sub>c2</sub>(T) have been determined from the temperature dependent in-phase ac susceptibility data, as in Ref.3. Note that in the FC mode, the sample would cross the $`H_p(T)`$ line at different point each time, while reaching a given $`(H,T)`$ value. Inset (ii) shows an enlarged view of the $`M`$-$`H`$ loop in the PE region, indicating the onset field $`H_{pl}^+`$, the peak field $`H_p`$ and the irreversibility field $`H_{irr}`$.
Figs.3(a) to 3(c). Minor magnetisation curves in the given $`NbSe_2`$ crystal at 6.95 K measured along three paths, as schematically sketched in Fig.1. The inset in Fig.3(a) shows the merger of FC-REV curves initiated from M<sub>FC</sub>(H) (where $`H`$ $`=`$ 0.4 kOe and 0.6 kOe) into the reverse envelope curve. Note that the minor curves in Fig.3(b) do not readily overlap with reverse envelope curve, whereas those in Fig.3(c) cut across the forward envelope curve.
Fig.4. Field dependence of J<sub>c</sub> for $`Hc`$at 6.95K in $`NbSe_2`$ for three different histories as indicated.
|
no-problem/9908/cond-mat9908392.html
|
ar5iv
|
text
|
# Electrical Conductance of Molecular Wires
## 1 Introduction
A molecular wire (MW)in its simplest definition consists of a molecule connected between two reservoirs of electrons. The molecular orbitals of the molecule when they couple to the leads provide favourable pathways for electrons. Such a system was suggested in the early ’70’s by Aviram and Ratner to have the ability to rectify current. Experimental research on MW has increased over the past few years looking into the possibility of rectification and other phenomena. There has also been an increase in the theoretical modeling of MW systems. For a comprehensive overreview of the current status of the molecular electronics field see .
Theoretical studies of the electronic conductance of a MW bring together different methods from chemistry and physics. Quantum chemistry is used to model the energetics of the molecule. It is also incorporated into the study of the coupling between the molecule and the metallic reservoirs. Once these issues have been addressed it is possible to proceed to the electron transport problem. Currently, Landauer theory is used which relates the conductance to the electron transmission probability.
A molecule of current experimental interest as a MW is 1,4 benzene-dithiolate (BDT). It consists of a benzene molecule with two sulfur atoms attached, one on either end of the benzene ring. The sulfurs bond effectively to the gold nanocontacts and the conjugated $`\pi `$ ring provides delocalized electrons which are beneficial for transport. Two major unknowns of the experimental system are the geometry of the gold contacts and the nature of the bond between the molecule and these contacts. This paper attempts to highlight these important issues by showing how the differential conductance varies with bond strength.
For mesoscopic systems with discrete energy levels (such as MW) connected to continuum reservoirs, the transmission probability displays resonance peaks. Another potentially important transport phenomenon that has been predicted is the appearance of antiresonances. These occur when the transmission probability is zero and correspond to the incident electrons being perfectly reflected by the molecule. We derive a simple condition controlling where the antiresonances occur in the transmission spectrum. We apply our formula to the case of a MW consisting of an “active” molecular segment connected to two metal contacts by a pair of finite $`\pi `$ conjugated chains. In this calculation we show how an antiresonance can be generated near the Fermi energy of the metallic leads. The antiresonance is characterized by a drop in conductance. We find that for this calculation our analytic theory of antiresonances has predictive power.
In Sec. II, we describe the scattering and transport theory used in this work. The first subsection deals with Landauer theory. The second subsection outlines a method for evaluating the transmission probability for MW systems. The last subsection outlines our derivation of the antiresonance condition. In Sec. III, we present some calculations on BDT for different lead geometries and different binding strengths. Sec. IV describes a numerical calculation for a system displaying antiresonances which are predicted by the formula in Sec. II. Finally we conclude with some remarks in Sec. V.
## 2 Transport Theory
### 2.1 Landauer Formula
We consider the transport of electrons through a molecular system by modeling it as a one electron elastic scattering problem. The molecule acts as a defect between two metallic reservoirs of electrons. An electron incident from the source lead with an energy $`E`$, has a transmission probability $`T(E)`$ to scatter through the molecule into the drain lead. A schematic of the model system is shown in Fig. 1. By determining the transmission probability for a range of energies around the Fermi energy $`ϵ_F`$ of the source lead, the finite temperature, finite voltage, Landauer formula can be used to calculate the transmitted current $`I`$ as a function of the bias voltage, $`V`$, applied between the source (left lead) and drain (right lead)
$$I(V)=\frac{2e}{h}_{\mathrm{}}^{\mathrm{}}𝑑ET(E)\left(\frac{1}{\mathrm{exp}[(E\mu _s)/kT]+1}\frac{1}{\mathrm{exp}[(E\mu _d)/kT]+1}\right)$$
(1)
The two electro-chemical potentials $`\mu _s`$ and $`\mu _d`$, refer to the source and drain, respectively. They are defined to be, $`\mu _s=ϵ_F+eV/2`$ and $`\mu _d=ϵ_FeV/2`$. The differential conductance is then given by the derivative of the current with respect to voltage.
### 2.2 t Matrix Method
We find the transmission probability $`T(E)`$ used in the Landauer formula, Eq.(1), by solving the Schroedinger equation directly for the scattered wavefunction of the electron. The electron is initially propagating in a Bloch wave in one of the modes of the source lead. The molecule will reflect some of this wave back into the various modes of the source lead. The molecule is represented by a discrete set of molecular orbitals (MO’s) through which the electron can tunnel. Hence, some of the wave will be transmitted through the molecule and into the modes of the drain lead. By finding the scattered wavefunction it is then possible to determine how much was transmitted, yielding $`T(E)`$.
We start with Schroedinger’s equation, $`H|\mathrm{\Psi }^\alpha =E|\mathrm{\Psi }^\alpha `$, where $`H`$ is the Hamiltonian for the entire MW system consisting of the leads and the molecule. $`|\mathrm{\Psi }^\alpha `$ is the wavefunction of the electron propagating initially in the $`\alpha ^{th}`$ mode of the left lead with energy $`E`$. It is expressed in terms of the transmission and reflection coefficients, $`t_{\alpha ,\alpha ^{}}`$ and $`r_{\alpha ,\alpha ^{}}`$ and has different forms on the left lead (L), molecule (M), and right lead (R). The total wavefunction is a sum of these three, $`|\mathrm{\Psi }^\alpha =|\mathrm{\Psi }_L^\alpha +|\mathrm{\Psi }_M^\alpha +|\mathrm{\Psi }_R^\alpha `$, where
$`|\mathrm{\Psi }_L^\alpha `$ $`=`$ $`|\mathrm{\Phi }_+^\alpha +{\displaystyle \underset{\alpha ^{}}{}}r_{\alpha ^{},\alpha }|\mathrm{\Phi }_{}^\alpha ^{}`$ (2)
$`|\mathrm{\Psi }_M^\alpha `$ $`=`$ $`{\displaystyle \underset{j}{}}c_j|\varphi _j`$ (3)
$`|\mathrm{\Psi }_R^\alpha `$ $`=`$ $`{\displaystyle \underset{\alpha ^{}}{}}t_{\alpha ,\alpha ^{}}|\mathrm{\Phi }_+^\alpha ^{}`$ (4)
In the above, $`|\mathrm{\Phi }_\pm ^\alpha `$ are forward/backward propagating Bloch waves in the $`\alpha ^{th}`$ mode, and $`|\varphi _j`$ is the $`j^{th}`$ MO on the molecule.
We then consider our Hamiltonian in the tight-binding (or Hückel) approximation and use Schroedinger’s equation to arrive at a system of linear equations in the unknown quantities, $`r_{\alpha ^{},\alpha }`$, $`c_j`$ and $`t_{\alpha ^{},\alpha }`$. We solve numerically for the reflection and transmission coefficients for each rightward propagating mode $`\alpha `$ at energy $`E`$ in the left lead. The total transmission is then given by
$$T(E)=\underset{\alpha ϵL}{}\underset{\alpha ^{}ϵR}{}\frac{v^\alpha ^{}}{v^\alpha }|t_{\alpha ^{},\alpha }|^2$$
(5)
where $`v^\alpha `$ is the velocity of the electron in the $`\alpha ^{th}`$ rightward propagating mode. The sum over $`\alpha ^{}`$ in the expression for $`T`$ is over the rightward propagating modes at energy $`E`$ in the right lead. With $`T(E)`$ determined the current and differential conductance can then be calculated using Eq. (1).
### 2.3 Antiresonance Condition
The above section outlined a method for numerically evaluating the transmission probability for a MW system. We now focus on an interesting physical feature of the transmission probability, namely the occurrence of antiresonances. These are characterized by the perfect reflection of electrons incident on the molecule at a certain energy $`E`$. The basis of our analysis in this section will be the Lippmann-Schwinger (LS) equation applied to the scattering problem in a highly idealized model. It will lead to an analytic formula for $`T(E)`$.
Before we proceed further we wish to highlight a problem that arises in most quantum chemistry applications. In studying quantum systems, such as MW, which are composed out of atomic building blocks, it is customary to solve the problem by expressing the electron’s wavefunction in terms of the atomic orbitals on the various atoms. A problem that arises is that these orbitals are not usually orthogonal to each other. Including this nonorthogonality complicates the solution and so it is often neglected. In the analysis that follows we have utilized a transformation which removes the nonorthogonality and allows for a straightforward solution. This transformation leads to an energy dependent Hamiltonian which will play an important role in determining the presence of antiresonances.
The LS equation we consider has the form
$$|\mathrm{\Psi }^{}=|\mathrm{\Phi }^{}+G^{}(E)W^E|\mathrm{\Psi }^{}.$$
(6)
Here $`|\mathrm{\Psi }^{}`$ is the scattered electron wavefunction of the transformed Hamiltonian $`H^E`$. $`W^E`$ couples the molecule to the adjacent lead sites. $`|\mathrm{\Phi }^{}`$ is the initial electron state which is a propagating Bloch wave that is confined to the left lead. $`G^{}(E)=(EH_0^E)^1`$ is the Green’s function of the decoupled system.
We consider the LS equation in the tight binding approximation and solve for $`|\mathrm{\Psi }^{}`$. This gives us the value for $`\mathrm{\Psi }_1`$ which is the value of the wavefunction on the first atomic site on the right lead. The transmission probability, $`T`$ is simply $`|\mathrm{\Psi }_1|^2`$. The result is
$$\mathrm{\Psi }_1=\frac{P\mathrm{\Phi }_1^{}}{[(1Q)(1R)PS]}$$
(7)
where
$`P`$ $`=`$ $`G_{1,1}^{}{\displaystyle \underset{j}{}}W_{1,j}^EG_j^{}W_{j,1}^E`$
$`Q`$ $`=`$ $`G_{1,1}^{}{\displaystyle \underset{j}{}}(W_{1,j}^E)^2G_j^{}`$
$`R`$ $`=`$ $`G_{1,1}^{}{\displaystyle \underset{j}{}}(W_{1,j}^E)^2G_j^{}`$
$`S`$ $`=`$ $`G_{1,1}^{}{\displaystyle \underset{j}{}}W_{1,j}^EG_j^{}W_{j,1}^E`$
The sum over $`j`$ is over only the MO’s. In the above, $`W_{1,j}^E=H_{1,j}ES_{1,j}`$ is the energy-dependent hopping element of $`H^E`$ between the first lead site and the $`j^{th}`$ MO in terms of the hopping element of the original Hamiltonian $`H`$ and the overlap $`S`$ in the non-orthogonal basis. The Green’s function on the molecule is expanded in terms of its molecular eigenstates and this gives $`G_j^{}=1/(Eϵ_j)`$ for the $`j^{th}`$ MO with energy $`ϵ_j`$. $`G_{1,1}^{}`$ is the diagonal matrix element of the Green’s function $`G^{}(E)`$ at the end site of the isolated lead.
Antiresonances of the MW occur where the transmission $`T`$ is equal to zero. These occur at Fermi energies $`E`$ that are the roots of Eq. (7), namely
$$\underset{j}{}\frac{(H_{1,j}ES_{1,j})(H_{j,1}ES_{j,1})}{Eϵ_j}=0.$$
(8)
Antiresonances can arise due to an interference between molecular states that may differ in energy, as is seen directly from Eq. (8). An electron hops from the left lead site adjacent to the molecule onto each of the MO’s with a weight $`W_{j,1}^E`$. It then propagates through each of the different orbitals and hops onto the right lead with a weight $`W_{1,j}^E.`$ These processes interfere with each other and where they cancel (8) is satisfied and an antiresonance occurs.
Antiresonances can also arise solely from the nonorthogonality of atomic orbitals. This occurs when only a single MO $`k`$ couples appreciably to the leads. Eq. (8) then becomes $`(H_{k,1}ES_{k,1})(H_{1,k}ES_{1,k})=0`$. This equation gives two antiresonances. The non-orthogonality of two orbitals can actually stop electron hopping between these orbitals, which blocks electron transport and creates an antiresonance. This is a counterintuitive effect since one would normally expect orbital overlap to aid electron transfer between the orbitals rather than hinder it. Thus the inclusion of orbital overlap leads to non-trivial physical consequences in the tight binding treatment of MW.
## 3 Conductance of BDT
For our first application, we apply our t matrix formalism to study the conductance of 1,4 benzene-dithiolate, a MW of recent experimental interest. The molecule consists of a benzene molecule with the hydrogen atoms at the 1 and 4 positions replaced with sulphur atoms. The sulphur atoms act like alligator clips when they bond to the gold leads. We calculate the conductance of the MW geometries shown in Fig. 2. The gold leads are oriented in the (111) direction. Attached to the molecule are gold clusters which form the tips for the lead. Experimentally it has been found that sulphur atoms preferentially bind over the hollow sites formed on gold surfaces and so for our simulation the BDT molecule is bonded over the hollow site on each tip.
The strength of the bond between the molecule and the gold surface plays an important role in determining the transmission characteristics of the BDT MW. Isolated BDT has a discrete set of MO’s with the highest occupied molecular orbital (HOMO) calculated to be around -10.5 eV and the lowest unoccupied molecular orbital (LUMO) found to be around -8.2 eV. These levels when bonded with the leads become part of the continuum of energy states that exist within the metallic reservoirs. For strong bonding they can become significantly altered as their chemical nature becomes mixed with the surface states of the gold tips. For weaker bonding the MO’s retain the character of the isolated molecule.
We consider strong binding to the (111) leads first. The transmission diagram is shown in Fig. 3a. There is strong transmission in the energy regions where the gold tip states have mixed with the molecular states. This occurs most prominently around -11.5 eV, where there exist resonances that can be connected with the HOMO states of the molecule. The HOMO and states around it in the isolated molecule have been mixed and lowered in energy due to the bonding to the lead. The other region of significant transmission is at around -8 eV, which is due to states connected with the LUMO of the BDT. The region in between has resonances that arise from those states that are complex admixtures of gold tip and molecule levels. The differential conductance was calculated with a Fermi level chosen at -10 eV which lies in the HOMO-LUMO gap. The molecule seems to be very conductive when attached strongly to the (111) oriented wide leads.
The magnitude of the conductance found in the above calculation exceeds that found experimentally. We now consider the case of weak binding to see what effect it will have on the transmission. One lead is allowed to bond strongly to the molecule, while the other lead is pulled away from the molecule so as to be at non-bonding distances. The transmission and conductance diagrams for the weak bonding case are shown in Fig. 3b. It can be seen that the transmission has gone down significantly in the region between -10.5 and -8 eV. There are also now strong resonances that correspond almost exactly with the isolated molecular levels. The weaker coupling of the molecule to the contacts has lowered the transmission and the conductance. Because there are order of magnitude fluctuations in $`T(E)`$ the conductance is now quite sensitive to the selection of Fermi energy. Different conductance curves can be obtained if the Fermi energy is chosen to occur on or off resonance.
## 4 Antiresonance Calculations
We now look at a MW system that displays antiresonances that are predicted by the formula derived in Sec. II. Eq. (8) was derived for an idealized MW system where there was only one incident electron mode on the molecule. Thus in designing the MW system for the numerical study we tried to approximate the ideal system as closely as possible. The MW system we suggest consists of left and right $`\pi `$ conjugated chain molecules attached to what we will call the “active” molecule. The purpose of these conjugated chains is to act as filters to the many modes that will be incident from the metallic leads. For appropriate energies they will restrict the propagating electron mode to be only $`\pi `$ like. This $`\pi `$ backbone will only interact with the $`\pi `$ orbitals of the molecule if they are bonded in an appropriated fashion.
An atomic diagram of our system is shown in Fig. 4. The leads are oriented in the (111) direction. The chain molecules are bonded to clusters of gold atoms that form the tips of the leads. The carbon atom nearest to the gold tip binds over the hollow site of the tip. The chains each have eight CH groups, which for the energies considered will only admit a $`\pi `$ like state to propagate along them. This gives rise to the filtering process mentioned above. The active molecule is chosen to have two $`\sigma `$ states and two $`\pi `$ states. The chain’s $`\pi `$ like orbitals will only couple to the two $`\pi `$ states of the active molecule.
The Fermi energy for our gold leads is around -10 eV which lies within the $`\pi `$ band. We would like an antiresonance to occur somewhere near this energy. The parameters entering Eq. 8 (i.e. the molecular orbital energies and their overlaps with the chains) are chosen so that one of the roots of the equation yields a value near -10 eV. The numerically calculated electron transmission probability for this model is shown in Fig. 5a; an antiresonance is seen in the plot at the predicted energy of -10.08 eV. The differential conductance at room temperature was calculated using Eq. (1). Two different conductance calculations are shown in Fig. 5b. The solid curve corresponds to a choice of Fermi energy of -10.2 eV. Because it lies to the left of the antiresonance in a region of strong transmission, the conductance is strong at 0 V. It then drops at around 0.2 V when the antiresonance is crossed. The dashed curve was calculated using a Fermi energy of -10.0 eV. It starts in a region of lower transmission and thus the antiresonance suppresses the increase in current. After 0.2 V the large transmission to the left of the antiresonance is sampled and the current rises sharply. So in both cases the antiresonance has lowered the conductance. It is conceivable to think of utilizing more antiresonances in a narrow energy range to create a more pronounced conductance drop.
## 5 Conclusions
We have presented a theoretical study of the electrical conductance properties of molecular wires in the context of one-electron theory. A numerical method has been developed for the study of transport in molecular wires which solves for the transmission coefficients using Schroedinger’s equation. It allows for the study of multimode leads attached to a molecule. We then proceeded to present a simple analytically solvable model which highlighted the interesting phenomena of antiresonances. A formula was derived that predicts Fermi energies for which a molecule with a given set of molecular energy levels should display antiresonances. It predicts two mechanisms by which antiresonances arise: one due to interference between the molecular orbitals and the other due to a cancellation of the effective hopping parameter.
As an application of our numerical method studied the conductance of BDT. We examined the role of coupling by considering both the strong and weak regimes. For strong coupling it was found that the MW has regions of strong transmission. These regions occur at energies which differ from the isolated molecule’s energy levels because of state hybridization with the surface states of the gold tips. The conductance found for the strong coupling case was orders of magnitude greater than that found experimentally, although qualitatively it shared common features. In the weak coupling study the transmission displayed resonances at energies corresponding to those of the isolated molecule. The magnitude was also significantly down. This resulted in a conductance curve that was of the magnitude found in the experiment. Future work will need to focus on the electrostatic problem of the molecule within an applied electric field and the consequences of this in the context of Landauer theory, many-electron and possible polaronic effects within the MW.
Using our analytic formula for antiresonances, we were able to predict the occurrence of an antiresonance within a more sophisticated numerical study utilizing a molecule attached bridging a metal wire break junction. The model should also be of interest for future MW work since it introduces the idea of “filter” molecules. Our numerical studies showed that $`\pi `$ conjugated chain molecules act as effective mode filters to electrons incident from the metallic leads. The filter chains in our model reduced the number of propagating modes down to one, which was then coupled to the “active” molecule. Our formula was derived on the assumption of only a single propagating mode, and for the “active” molecule considered it was able to successfully predict the energy at which the antiresonance occurred. The antiresonance was charaterized by a drop in the differential conductance.
|
no-problem/9908/astro-ph9908208.html
|
ar5iv
|
text
|
# What is Beta?
## 1 Analysis
We estimate $`\beta `$ from a variety of simulated galaxy distributions, created by applying simple local and deterministic biasing algorithms to the mass distributions obtained from cosmological N-body simulations.
The cosmological models that we use are: (1) $`\mathrm{\Omega }=1.0`$, CDM model with a tilted power spectrum designed to simultaneously satisfy both COBE and cluster normalization constraints. (2) $`\mathrm{\Omega }=0.4`$, CDM, cluster normalized model. (3) $`\mathrm{\Omega }=0.2`$, CDM, cluster normalized model. Cluster normalization requires that $`\sigma _{8,m}\mathrm{\Omega }^{0.6}=0.55`$ (White, Efstathiou & Frenk 1993).
The biasing algorithms that we use are: (1) Semi-analytic: An empirical bias prescription derived by Narayanan et al. (1999, in prep.) which characterizes the relation between the galaxy and mass density fields in the semi-analytic galaxy formation models of Benson et al. (1999). (2) Sqrt-Exp. (Square-root Exponential): A biasing prescription in which $`(1+\delta _g)\sqrt{(1+\delta _m)}e^{\alpha (1+\delta _m)}`$. (3) Power-law: A biasing prescription in which $`(1+\delta _g)(1+\delta _m)^\alpha `$. Here, $`(1+\delta _g)`$ and $`(1+\delta _m)`$ are the galaxy and mass over-densities, respectively.
All the bias models are designed to yield galaxy distributions with an rms fluctuation, in top-hat spheres of radius $`12h^1`$Mpc, of $`\sigma _{12}0.7`$. The values of $`\mathrm{\Omega }^{0.6}/b_\sigma (12)`$ for the galaxy distributions are 0.6, 0.58, and 0.56 for $`\mathrm{\Omega }=1.0`$, 0.4, and 0.2, respectively. All the results have been averaged over four independent realizations of the cosmological models. The simulation volumes are periodic cubes, with a box size of $`400h^1`$Mpc.
The first method that we use to estimate $`\beta `$ involves the anisotropy of the redshift-space power spectrum. Kaiser (1987) and Cole, Fisher & Weinberg (1994) showed that in linear theory, and under the plane-parallel approximation, the multipole moments of the redshift-space power spectrum may be used to estimate $`\beta `$. Here we employ two different $`\beta `$ estimators: $`P^S(k)/P^R(k)`$, the ratio of the angle-averaged redshift-space power spectrum (monopole) to the real-space power spectrum, and $`P_2(k)/P_0(k)`$, the ratio of the quadrupole and monopole moments of the redshift-space power spectrum. These ratios are, in principle, measurable, and they are related to $`\beta `$ as follows: $`\frac{P^S(k)}{P^R(k)}=1+\frac{2}{3}\beta +\frac{1}{5}\beta ^2`$, $`\frac{P_2(k)}{P_0(k)}=(\frac{4}{3}\beta +\frac{4}{7}\beta ^2)/(1+\frac{2}{3}\beta +\frac{1}{5}\beta ^2)`$. These ratios yield an estimate of $`\beta `$ at each wavenumber $`k`$, and the scale dependence of this estimate is caused mainly by non-linearity in the peculiar velocity field. It is possible to obtain a global estimate of $`\beta `$ by modeling this non-linearity. We use two such models in our analysis: (1) Exponential velocity distribution model (Cole, Fisher & Weinberg 1995), where we assume that the galaxies have uncorrelated small scale peculiar velocities which are drawn from an exponential distribution. We use this model in our analysis of $`P^S(k)/P^R(k)`$ (the results of which are shown in Fig. 1). (2) Empirical model (Hatton & Cole 1999), where the scale dependence of $`P_2(k)/P_0(k)`$ is found empirically by examining a large number of N-body simulations spanning a broad range of parameter space.
The second method that we use to estimate $`\beta `$ involves a direct comparison between the mass and galaxy density fields. The POTENT method (Bertschinger & Dekel 1989) reconstructs the full three dimensional peculiar velocity field $`𝐯`$ from the observed radial peculiar velocity field, allowing a measurement of $`\beta `$ directly from the linear theory relation $`𝐯=\beta H\delta _g`$. We estimate $`\beta `$ using an idealized POTENT method, in which we have perfectly reconstructed the three dimensional peculiar velocity field. We obtain the value of $`\beta `$ by fitting a line of the form $`y=ax+c`$ to the $`(𝐯)`$ vs. $`\delta _g`$ relation. We perform this fit in the region: $`0.5<\delta _g<0.5`$. By repeating this procedure for different smoothing of the velocity and density fields, we can measure $`\beta `$ as a function of scale (see Fig. 1).
## 2 Conclusions
Figure 1 compares $`\beta `$ estimates from the two methods described previously to the function $`\beta _\sigma (R)=\mathrm{\Omega }^{0.6}/b_\sigma `$. Here, $`b_\sigma (R)`$ is the bias function defined as $`b_\sigma (R)=\sigma _g(R)/\sigma _m(R)`$, where $`\sigma _g(R)`$ and $`\sigma _m(R)`$ are the rms fluctuations of the galaxy and mass density fields, smoothed with a Gaussian filter of radius $`R`$, of the galaxy and mass distributions, respectively. In our full analysis, we also compare our $`\beta `$ estimates to the function $`\beta _P(k)=\mathrm{\Omega }^{0.6}/b_P(k)`$. Here $`b_P(k)`$ is the bias function defined as $`b_P(k)=\sqrt{P_g(k)/P_m(k)}`$, where $`P_g(k)`$ and $`P_m(k)`$ are the power spectra of the galaxy and mass distributions, respectively. In the case of a simple, linear bias model, $`\delta _g=b\delta _m`$, all these definitions of $`b`$ are equivalent, i.e., $`b_\sigma =b_P=b`$.
In most cases, we find that the bias factor $`b=\mathrm{\Omega }^{0.6}/\beta `$ derived from redshift-space anisotropy or POTENT is similar to the large-scale value of $`b_P(k)`$ or $`b_\sigma (R)`$ defined by rms fluctuation amplitudes. Discrepancies at the 10-20% level may reflect non-linear dynamics as much as non-linearity of bias, since they also occur for unbiased models. In greater detail: (1) Measurements of $`\beta `$ derived by fitting the Cole et al. (1995) exponential velocity distribution model to $`P^S(k)/P^R(k)`$ underestimate the large-scale values of $`\beta _P`$ by about 10-20% for all cosmological models and biasing prescriptions. (2) Measurements of $`\beta `$ derived by fitting the Hatton & Cole (1999) empirical model to $`P_2(k)/P_0(k)`$ accurately reproduce the large-scale values of $`\beta _P`$ for all cosmological models and biasing prescriptions. (3) Measurements of $`\beta (R)`$ derived from a POTENT-like comparison of $`(𝐯)`$, the negative divergence of the peculiar velocity field, to $`\delta _g`$, the galaxy density field, underestimate the large-scale values of $`\beta _\sigma `$ by about 10-20% for all cosmological models and biasing prescriptions, with the exception of the Sqrt-Exp. biasing prescription. The Sqrt-Exp. model gives erratic results because the non-linearity of the bias model makes the estimated $`\beta `$ sensitive to the range of $`\delta _g`$ used in the linear fit. (4) Estimates of $`\beta `$ from the anisotropy of the redshift-space power spectrum approximately agree with those from the POTENT method on large scales, except in the case of the Sqrt-Exp. model.
|
no-problem/9908/cond-mat9908461.html
|
ar5iv
|
text
|
# Transition Matrix Monte Carlo
## 1 Introduction
This paper presents both an approach to analyzing the information contained in configurations generated by general Monte Carlo simulations, and a closely related method of simulation that provides great flexibility and surprising efficiency. The approach uses information contained in the configurations about the set of possible changes on the next Monte Carlo step, which we encode in a “Transition Matrix.”<sup>-</sup> This information is complementary to a histogram analysis<sup>,</sup>. Indeed, it was originally developed as an improvement to the histogram approach. The discovery that the transition matrix, by itself, had definite advantages over the older methods has opened up the development of several new techniques.
Combining the Transition Matrix with an N-fold way simulation technique produces an extremely flexible and efficient approach to rather general Monte Carlo simulations. Because the same information is needed for the simulation and the analysis, the two methods work extremely efficiently together. In particular, the fact that this information is updated and used for every step of the N-fold way simulation enables contributions to the transition matrix to be made at every MC step, instead of after every sweep, as in cluster simulations.
Since no use is made of cluster methods that could be limited to non-frustrated systems, the techniques described in this paper are extremely general. They can be used directly with any system that has equally spaced energy levels, and, as is the case for histograms, can be generalized to systems with continuous symmetry with the use of binning.
In the following sections, we will define the transition matrix, describe how thermodynamic information is extracted from it, describe the N-fold way (as it is used with the transition matrix), describe the creation of generalized ensembles, and discuss the advantages of using these methods as part of parallel simulations similar to Replica Monte Carlo or Parallel Simulated Tempering.
## 2 The Transition Matrix
We will use the two-dimensional Ising model to demonstrate the transition matrix approach, although generalizations to higher dimensions and more complicated models is trivial. In particular, the generalization to a spin glass is obvious. The Ising model is given by the Hamiltonian
$$H=J\underset{i,j}{}\sigma _i\sigma _j,$$
(1)
where $`\sigma _i`$ takes on the values $`\pm 1`$.
Although transition matrices can be defined for any kind of Monte Carlo simulation that depends only of the energies of the initial and final states, the primary transition matrix is defined in terms of the standard Monte Carlo dynamics at infinite temperature. For this dynamics, each spin is chosen with equal probability and “flipped” with probability one.
For every spin in a configuration, it is easy to calculate the change in energy of the configuration when that spin is reversed by simply counting the number of neighbors with the same sign. Denoting the number of sites that will produce an energy change $`\delta E`$ by $`n_{\delta E}`$, we define the “infinite temperature transition matrix” by
$$T_{E,\delta E}n_{\delta E}_E/N,$$
(2)
where $`N`$ is the total number of sites and the average is taken over all configurations at energy $`E`$.
The largest eigenvalue of this matrix is unity, and the corresponding eigenvector is the density of states, $`W(E)`$.
$$\underset{\delta E}{}W(E\delta E)T_{E\delta E,\delta E}=W(E)$$
(3)
The elements of the transition matrix must satisfy the condition of detailed balance,
$$W(E)T_{E,\delta E}=W(E+\delta E)T_{E+\delta E,\delta E},$$
(4)
which requires the “TTT-identity”,
$$T_{E,\mathrm{\Delta }}T_{E+\mathrm{\Delta },\mathrm{\Delta }}T_{E+2\mathrm{\Delta },2\mathrm{\Delta }}=T_{E,2\mathrm{\Delta }}T_{E+2\mathrm{\Delta },\mathrm{\Delta }}T_{E+\mathrm{\Delta },\mathrm{\Delta }},$$
(5)
where $`\mathrm{\Delta }`$ is the smallest allowable energy change.
For Potts models, the Ising model in three-dimensions, or other models with a greater number of possible energy changes, there are additional TTT-identities of the form,
$$T_{E,(m1)\mathrm{\Delta }}T_{E+(m1)\mathrm{\Delta },\mathrm{\Delta }}T_{E+m\mathrm{\Delta },m\mathrm{\Delta }}=T_{E,m\mathrm{\Delta }}T_{E+m\mathrm{\Delta },\mathrm{\Delta }}T_{E+(m1)\mathrm{\Delta },(m1)\mathrm{\Delta }}.$$
(6)
The situation is only slightly more complicated if other quantities, like the magnetization or second-neighbor interactions, are introduced. In each case, the introduction of an additional energy change creates two new elements at each energy ($`+\delta E`$ and $`\delta E`$), along with one new identity (possibly involving four matrix elements), leaving one new independent variable.
Transition matrices corresponding to general Monte Carlo dynamics, $`\stackrel{ˇ}{T}_{E,\delta E}`$, can be constructed from the acceptance probabilities,
$$\stackrel{ˇ}{T}_{E,\delta E}=\{\begin{array}{cc}T_{E,\delta E}\times a_{E,\delta E}\hfill & \mathrm{for}\delta E0\hfill \\ T_{E,\delta E}\times a_{E,\delta E}+\mathrm{\Sigma }_{\delta E0}T_{E,\delta E}\times \left(1a_{E,\delta E}\right)\hfill & \mathrm{for}\delta E=0\hfill \end{array}.$$
(7)
The corresponding probability distribution can then be constructed from the leading eigenvector of this matrix.
For example, the standard Metropolis acceptance probabilities are given by
$$a_{E,\delta E}=\mathrm{min}[1,\mathrm{exp}\left(\delta E/kT\right)],$$
(8)
and the leading eigenvector of the corresponding transition matrix is the usual canonical probability distribution.
## 3 The N-Fold Way
Instead of choosing spins and determining whether to accept each move, it is possible to use the set of numbers, $`\left\{n_{\delta E}\right\}`$, to determine in advance what the probability of picking a spin that will change the energy by $`\delta E`$ and calculating its probability of acceptance. A class of spins with $`\delta E`$ is then picked with probability
$$\left(\frac{n_{\delta E}}{N}\right)\left(\frac{a_{E,\delta E}}{A}\right),$$
(9)
where
$$A=\underset{\delta E}{}\left(\frac{n_{\delta E}}{N}\right)a_{E,\delta E}.$$
(10)
A spin from that class is chosen and flipped. The contributions to the transition matrix (and any other quantity being computed) are weighted with a factor of $`\left(1/A\right)`$.
Since the set of numbers $`\left\{n_{\delta E}\right\}`$ are updated at every step for both the simulation and the transition matrix, little extra computer time is needed to record contributions to the transition matrix at every Monte Carlo step.
It is clear that any acceptance rates may be used in defining an ensemble for Monte Carlo simulations, as long as detailed balance is satisfied. Because of the reweighting of the contributions of each configuration, it is not even necessary for the acceptance rates to be normalized.
## 4 Generalized Ensembles
To create an ensemble with a desired probability distribution, $`P(E)`$, a necessary condition is
$$P(E)\times T_{E,\delta E}\times a_{E,\delta E}=P(E+\delta E)\times T_{E+\delta E,\delta E}\times a_{E+\delta E,\delta E}.$$
(11)
For example, the multicanonical ensemble is characterized by $`P(E)=P(E+\delta E)`$, so that the condition on the acceptance rates is
$$\frac{a_{E,\delta E}}{a_{E+\delta E,\delta E}}=\frac{T_{E+\delta E,\delta E}}{T_{E,\delta E}}.$$
(12)
Eqn. 4 relates this to the usual condition that the ratios of acceptance rates is equal to the ratio of the densities of state at the two energies.
Eqn. 12 is not sufficient to determine the acceptance rates uniquely. In addition to the usual acceptance rates given by the minimum of the ratio in Eqn. 12 or unity, both
$$a_{E,\delta E}=T_{E+\delta E,\delta E}$$
(13)
and
$$a_{E,\delta E}=1/T_{E,\delta E}$$
(14)
are valid. Many other options exist.
## 5 The “Equal-Hit” Ensemble
Although the multicanonical ensemble was designed to visit every energy level with equal probability, that is not really optimal when using the N-fold way. For low energies, the acceptance ratio is small and the N-fold way achieves a uniform $`P(E)`$ by visiting an energy level very few times, but weighting it with a large factor $`A^1_{E,Nf}`$. To scan all energy levels equally, it would be more appropriate to specify that the number of visits to each energy level, $`H(E)`$, is uniform. Note that the average of $`A^1`$ over the visits to energy level, $`E`$, with the N-fold way is not the same as the microcanonical average over configurations with that energy. Denoting the microcanonical average of a quantity, $`B`$, by, $`B_E`$, we have
$$B_E=\frac{BA^1_{E,Nf}}{A^1_{E,Nf}}.$$
If we take $`B=A`$, this gives
$$A^1_{E,Nf}=\frac{1}{A_E}.$$
Since the probability, $`P(E)`$, is given by the product of the number of times an N-fold way move ends at the energy $`E`$ times the average inverse acceptance ratio, $`A^1_{E,Nf}`$, the condition for an equal-hit ensemble with $`H(E)=H(E+\delta E)`$ is
$$A^1_{E,Nf}\times T_{E,\delta E}\times a_{E,\delta E}=A^1_{E+\delta E}\times T_{E+\delta E,\delta E}\times a_{E+\delta E,\delta E}.$$
(15)
This gives the condition on the acceptance rates.
$$\frac{a_{E,\delta E}}{a_{E+\delta E,\delta E}}=\frac{A^1_{E+\delta E,Nf}\times T_{E+\delta E,\delta E}}{A^1_{E,Nf}\times T_{E,\delta E}}$$
(16)
This condition can also be fulfilled in many ways, including
$$a_{E,\delta E}=A^1_{E+\delta E,Nf}\times T_{E+\delta E,\delta E}$$
(17)
and
$$a_{E,\delta E}=A^1_{E+\delta E,Nf}/T_{E,\delta E}.$$
(18)
## 6 Equilibration
At the beginning of a Monte Carlo simulation, there is usually little information about either the density of states or the transition matrix. For a canonical simulation at a given temperature, this does not cause a problem. However, for either a multicanonical simulation or an equal-hit simulation, such information is necessary. However, it turns out that the problem of dealing with the lack of information is greatly simplified by the transition matrices.
The primary approach consists of three stages.
For the first stage, we start with a random configuration and use one of the acceptance rates if it is known, otherwise we arbitrarily set it equal to unity. As the simulation in this stage progresses, we have a changing random walk that does not satisfy detailed balance and is biased towards states that have not yet been visited.
For the second stage, we take the approximate (and biased) transition matrix from the first stage, and impose the TTT-identities. In the second stage of the simulation, we use this approximate transition matrix to define the acceptance rates. Although the acceptance rates are not those of the ensemble that we really want to simulate, they do satisfy detailed balance by their construction, so that the transition matrix calculated in the second stage is not biased.
For the third stage, we use the estimate of the transition matrix (after imposing the TTT-identities) from the second stage to determine the acceptance rates. This algorithm represents a good approximation to the ensemble we wish to simulate, as well as satisfying detailed balance. The third stage is the main part of the simulation; it is where most of the computer time is used.
A modification of the third stage would be to continually update the transition matrix as the simulation progresses. This has the advantage that the simulation provides an increasingly accurate representation of the desired ensemble. The obvious disadvantage is that, strictly speaking, the algorithm no longer satisfies detailed balance. However, we have carried out some very long simulations on small systems with this modification, and found that the errors decrease with the square root of the length of the simulation, as expected. Note that the effect of any bias from the early part of the simulation is expected to decrease directly as the length of the simulation, rendering it unimportant. Therefore, we believe that the small violation of detailed balance is harmless, and will improve the efficiency of the algorithm.
The two-stage equilibration before the main simulation in the third stage requires relatively little computer time, so that this approach turns out to be more efficient than the standard multicanonical procedure, as well as being much simpler.
## 7 Parallel Simulations
A great advantage of a multicanonical simulation, which is retained by the equal-hit simulation, is that information over the entire range of energies can be obtained. This means that a single simulation provides the thermodynamic behavior at all temperatures – even negative temperatures (also known as the antiferromagnetic version of the model).
However, these advantages come with a price. The relaxation time to equilibrate a random walk is proportional to the square of its range. Since the total energy range is proportional to the number of particles, $`N`$, the relaxation time in units of MC steps is proportional to $`N^2`$. This does, indeed, turn out to be the case for the simulations described above.
We can greatly reduce this disadvantage by introducing parallel simulations. If we consider a number, $`l`$, of parallel simulations on independent replicas of the system, in which each replica is restricted to a range proportional to $`N/l`$, the relaxation time for each replica is only proportional to $`\left(N/l\right)^2`$. Taking the increased number of simulations into account, we are left with a relaxation time in units of MC steps proportional to $`N^2/l`$. The relaxation time in units of MC steps/site is proportional to $`N/l`$.
Note that this improvement occurs even with a serial machine. Implementation on a parallel machine is trivial.
We can further improve the algorithm by exchanging replicas, as introduced in the Replica Monte Carlo method<sup>-</sup> and later rediscovered as parallel simulated tempering.
Break up the total energy range into pieces with equal width. Beginning with the lowest energies, restrict the first replica to the first two pieces of the energy spectrum. The second replica is restricted to the second and third pieces, and so on. Replica $`n`$ and replica $`n+1`$ therefore overlap for half of their range of energies. After simulating for a few relaxation times (proportional to $`N^2/l`$), all neighboring replicas that are in the overlapping half of their ranges are exchanged. Since we are either using a multicanonical simulation with single spin flips, or an equal-hit ensemble with the N-fold way, the acceptance probability for such exchanges is unity.
This combination of replica exchange with an equal-hit (or multicanonical) simulation provides improvements over both. Although this approach has been discussed in terms of transition matrix Monte Carlo, its use is much broader. For example, it could easily be applied to multicanonical simulations of biological molecules, as we intend to do in future work.
|
no-problem/9908/astro-ph9908343.html
|
ar5iv
|
text
|
# A robust method for measuring the Hubble parameter
## 1 Introduction
Despite the recent emergence of a broad consensus in estimates of $`H_0`$, the issue of observational selection effects remains an important one for studies of the distance scale and peculiar velocity field. Current recipes for eliminating Malmquist bias make parametric model assumptions about the distribution function of the distance indicator and the selection effects, and the spatial distribution of the observed galaxies. There is clearly an advantage in developing more robust techniques in which only minimal model assumptions are required. In this paper we present such a robust method.
## 2 Method and Application Using Galaxy Diameters
Our technique is based on the $`C^{}`$ method of Lynden-Bell (1971), and provides an estimate of the cumulative distribution function (CDF) of galaxy diameter independent of any model assumption about its parametric form. Moreover, the method may be applied to data of arbitrary spatial distribution and thus requires no correction for Malmquist bias. The method is, however, applicable only to samples which are strictly complete to a given apparent magnitude or diameter. We have developed an objective test of the validity of this assumption, based on the approach of Efron & Petrosian (1992). (See Rauzy & Hendry in prep. for more details).
We applied the method to reconstruct the CDF of linear isophotal diameter, $`D`$, from a sample of 4005 galaxies – complete to an angular diameter limit of $`D1.5^{}`$ – from the LEDA database (Paturel et al. 1997). We carried out our analysis using the variables, $`m`$ and $`M`$, analogous to apparent magnitude and absolute magnitude, given by
$$m=205\mathrm{log}_{10}D$$
(1)
$$M=mZ=205\mathrm{log}_{10}D5\mathrm{log}_{10}\frac{cz}{H_0}25$$
(2)
We compared the CDF of $`M`$ from the LEDA galaxies with that obtained from a set of 14 local calibrators, with HST Cepheid distances, from Theureau et al. (1997). We then varied $`H_0`$ in eq. (2) and, for each value of $`H_0`$, determined the Kolmogorov-Smirnoff (KS) distance between the two CDFs as a function of $`M`$. We took as our ‘best-fit’ estimate of $`H_0`$ the value which gave the minimum KS distance – obtaining $`H_0=42`$ kms<sup>-1</sup> Mpc<sup>-1</sup>. This value agrees with Sandage (1993a,b), who used M31 and M101 as standard rulers, and is consistent with the analysis of Goodwin, Gribbin & Hendry (1997), who obtained $`H_0=52\pm 6\pm 8`$ kms<sup>-1</sup> Mpc<sup>-1</sup> using galaxy linear diameters and a similar calibrating sample. However, their second uncertainty was an estimate of the difference in the mean intrinsic diameter of the local calibrating galaxies compared with the distant sample (even after correction for Malmquist bias) – a difference which might be systematically negative given the strategy of the HST Key Project to observe ‘Grand Design’ spirals (Kennicutt et al. 1995). We find evidence for a similar negative bias in our calibrating sample: galaxies of small diameter are relatively under-represented in the CDF of the local calibrators. This would lead to a systematic underestimate in the value of $`H_0`$.
## 3 Including Tully-Fisher information
We can improve our analysis by introducing galaxy rotation velocity to reduce the dispersion of the distance indicator. This is analogous to the conventional Tully-Fisher relation but – crucially – retains completely the robustness of our previous analysis. In a similar manner to the ‘Sosie’ method (c.f. Paturel et al. 1998), we select, for each local calibrator, the subset of galaxies from the distant sample with similar log rotation velocity and morphological type. For each subset we then reconstruct the diameter function, assuming initially a fiducial value of $`H_0`$. Finally we determine the value of $`H_0`$ required to match the median of the CDF to the observed diameter for that calibrator.
We applied this technique to the KLUN sample of spiral galaxies (c.f. Theureau et al. 1997). Fig. 1 shows the CDFs reconstructed from the subsets corresponding to each local calibrator. The median value of the reconstructed distribution is indicated on 9 of the panels. The remaining 3 panels correspond to the 4 local calibrators with the smallest rotation velocities: N300, N598, N925 and N4496A. We exclude these calibrators since it is not clear from Fig. 1 whether their corresponding reconstructed CDFs are completely sampled – due to the presence of the lower diameter limit. Matching the median values from the 9 remaining panels to the linear diameters of the 10 remaining calibrators, and taking the logarithmic mean of these individual values gives
$$H_0=66\pm 6\mathrm{kms}^1\mathrm{Mpc}^1$$
(3)
## 4 Discussion
When we incorporate information on the rotation velocities of the local calibrators our results are in excellent agreement with recent determinations of $`H_0`$ from the conventional Tully-Fisher relation (c.f. Giovanelli et al 1997). Note, however, that our analysis is completely free of assumptions about the form of the galaxy diameter and luminosity function and the conditional distribution function of diameter at a given value of $`\mathrm{log}V_m`$. We do not, for example, require to assume that this conditional distribution is Gaussian, nor indeed even that it has zero mean or constant dispersion – as is often assumed in calibrating the Tully-Fisher relation. In particular, therefore, we do not require that the Tully-Fisher relation is a straight line, nor that the distribution of residuals is symmetrical. Our method would remain applicable if, for example, the distribution of Tully-Fisher residuals displayed a long ‘tail’ for galaxies of small rotation velocity. Our method is also completely independent of Malmquist bias corrections, so that our results are unaffected by the precise ‘recipe’ adopted to correct for Malmquist bias.
In summary, the robustness and assumption-free nature of this method has important ramifications for any current debate on the cosmic distance scale. There appears to be little possibility of any remaining systematic error in $`H_0`$ which is introduced at the stage of linking the primary and secondary distance scales. Improving the calibration of the Cepheid distance scale, via the lowest rungs of the Cosmic Distance Ladder, must now be the main priority for the Distance Scale community.
|
no-problem/9908/physics9908062.html
|
ar5iv
|
text
|
# References
Microscopic discontinuity of fluids
Dept. of Physics, Beijing University of Aeronautics
and Astronautics, Beijing 100083, PRC
C.Y. Chen, Email: [email protected]
Abstract: We reveal that realistic fluids generate microscopic-level discontinuity constantly and the discontinuity spreads out with motion of particles rather rapidly and widely. These things cannot be treated by the standard kinetic equations, and thus the existing fluid theories, macroscopic ones and microscopic ones, need to be revised considerably.
It has been unanimously believed that the following standard framework Continuity equations and Liouville’s theorem $``$ BBGKY hierarchy equations $``$ Kinetic equations $``$ Fluid equations $``$ Equilibrium and nonequilibrium phenomena
gives sound microscopic and macroscopic descriptions of all classical fluids observed in the physical reality.
Nevertheless, discoveries in various fields, such as those related to turbulence, chaos and dissipative structures, constantly revealed fluid phenomena that appear to be inconsistent with our existing theoretical knowledge. In the studies of statistical behavior of charged particles, we ourselves were in trouble to construct intuition out of what was obtained from the kinetic equations. The situation motivated us to make a lot of investigation on related fundamental subjects and finally resulted in our critical attitude toward the existing kinetic theory.
In the usual textbook elaboration of the kinetic theory, all the particles are at first assumed to make motion in the way as if they constitute a continuous medium and the mediumlike behavior is cast into a differential operator similar to that in fluid mechanics; an integral operator, as a correction term, is then attached, in which particles interact with each other as if they are ordinary colliding particles. In our view, the formalism, though seems to give a comprehensive description of the two contradictory aspects, medium aspect and particle aspect, runs into conceptual difficulties. According to the theory, if the number of particles becomes more and more (the gas becomes denser and denser), then the collisional term becomes overwhelming; otherwise, the medium aspect becomes more important. (In some textbook, it is explicitly stated that if the collisions between particles are not significant the Boltzmann equation is reduced to the collisionless Boltzmann equation, called sometimes the Vlasov equation.) However, this picture is not in harmony with the simple physical intuition which says that a “fluid” consisting of a very small number of particles behaves as a collection of particles only. The particles in such a “fluid” move like individual particles and collide with each other like individual particles; no mediumlike character can be observed.
In this letter, we will investigate the issue thoroughly. It will be shown that a realistic Boltzmann gas (by “Boltzmann gas” we mean a gas in which interforces between particles are short-ranged) is capable of generating several kinds of discontinuous distribution functions and the dynamical behavior of such distribution functions is far from mediumlike and cannot be described by the existing kinetic theory.
We start our discussion by recalling the standard kinetic equations briefly. These equations can be written in the following unifying form
$$\left(\frac{}{t}+𝐯\frac{}{𝐫}+\frac{𝐅}{m}\frac{}{𝐯}\right)f=\left(\frac{\delta f}{\delta t}\right)_{\mathrm{collision}},$$
(1)
where the right side represents an integral collisional operator, which takes different forms for different kinetic equations. For purposes of this paper, we confine ourselves to a gas in which collisions are not significant. Under this understanding, Eq. (1) becomes
$$\left(\frac{}{t}+𝐯\frac{}{𝐫}+\frac{𝐅}{m}\frac{}{𝐯}\right)f=0.$$
(2)
Moving together with one particle in the fluid, we simply find that the equation can be expressed as a convective derivative
$$\left(\frac{df}{dt}\right)_{\mathrm{path}}=0.$$
(3)
In words, this is to say that the distribution function keeps invariant along any particle path in the six-dimensional phase space ($`\mu `$-space).
The invariance expressed by (3) is quite amazing in the sense that the distribution function can have such nice and regular behavior while each particle in the fluid seems free to make any kinds of motions, including some “irregular ones”. This is by no means possible and seeking for hidden requirements is kind of necessary. A careful inspection leads us to the following three requirements. (i) Throughout the dynamical process, the distribution function must be continuous both in terms of position and in terms of velocity. (ii) The force term $`𝐅`$ in the equation must be free from dissipation. (iii) No stochastic forces (collision-type) get involved, namely the force $`𝐅`$ must be sufficiently smooth in space. In what follows, we are mainly concerned with the first requirement and it will be shown that a realistic fluid can constantly generate discontinuity at the miroscopic level and the discontinuity spread out rather rapidly and widely.
Firstly, consider a piece of boundary shown in Fig. 1. On the right side of it, there are particles belonging to an ordinary gas initially. Suppose that we are able to observe these particles continuously. It is easy to find that the distribution function at every point on the left side of the boundary involves discontinuity in terms of velocity angle.
Then, let’s look at the particles leaking through a small hole of a gas container (free-expansion gas), shown in Fig. 2. We can find that the distribution function related to all these particles is continuous in the position space but not in the velocity space.
Actually, there are more sources of discontinuity. To get a complete knowledge about this, it is instructive to study another kind of interaction between particles and boundaries, collisions between particles and boundaries. Such collisions are usually assumed to be kind of deterministic and the assumption is adopted by many college-level textbooks, in which molecules are considered as having hard-spheres, walls as perfectly smooth and collisions as classically elastic, shown in Fig. 3. According to this elastic model, the distribution function must keep invariant after the collisions, which partly explains why boundary problems have not received much attention in the standard kinetic theory.
It is easy to see that the model outlined above is an ideal one serving for educational purposes. Actual observations in fluid mechanics tell us a different story: the fluid particles immediately next to a solid surface remain almost stationary with respect to the surface (apparently in the average sense), shown in Fig. 4. This is called the no-slip condition by scientists in fluid mechanics.
To conform to the no-slip condition, a different model, called the statistical model, should be adopted. In this model, particles leave, after collisions, the solid surface in the way as if they are emitted from there, and their velocities will distribute according to a statistical law irrespective of how they come to there. (Though more sophisticated models may be better off in the academic sense, we take this simple one as our “zeroth-order approximation” for relative simplicity.)
Taking this statistical model implies that we are ready not only to admit that interaction between particles and boundaries is of a dissipative nature and stochastic nature but also to admit that the distribution functions of emitted particles are discontinuous: if the area of the surface is small enough the situation is like the gas leaking out of a container in Fig. 2; if the surface is of a finite-size, the distribution function will be roughly the same as that created by the finite boundary in Fig. 1.
Examples aforementioned have shown that realistic boundaries can indeed generate, by blocking particles or by emitting particles, certain kinds of discontinuous distribution functions. A conception which may come to one’s mind immediately is that by introducing the $`\delta `$-function and/or the step function these discontinuous distribution functions may be treatable as if they are continuous ones. Such conceptions got a lot of successes in other physical fields, why not in this field? We will see, however, that the position-velocity phase space, in which distribution functions are defined, possesses special features and one of the features is that the position $`𝐫`$ and the velocity $`𝐯`$ are assumed to be completely independent of each other. (Note that we have $`𝐯=\dot{𝐫}`$ for each particle in the fluid.) It is because of this assumption that the existing kinetic theory is inapplicable in the situations even if the $`\delta `$-function and step-function are properly introduced.
We investigate the following distribution function
$$f=\frac{n_0}{r^2}\delta (𝐯v_0\frac{𝐫}{r}),$$
(4)
where $`n_0`$ and $`v_0`$ represent two constants, $`𝐫`$ the position vector and $`r=|𝐫|`$. By referring to Fig. 2, we can recognize that the distribution function characterizes a kind of free expansion gas. The time development of these particles can be expressed by
$$\begin{array}{c}x^{}=x+v_xt\hfill \\ y^{}=y+v_yt\hfill \\ z^{}=z+v_zt\hfill \end{array}$$
(5)
and
$$v_x^{}=v_x,v_y^{}=v_y,v_z^{}=v_z.$$
(6)
By a direct calculation, we know that along a particle trajectory
$$\frac{f(t,𝐫,𝐯)}{f(t^{},𝐫^{},𝐯^{})}=\frac{r_{}^{}{}_{}{}^{2}}{r^2}.$$
(7)
Actually, Fig. 2 can directly tell us that, in contrast with (3), the distribution function is not invariant along the particle trajectories.
The difficulty becomes even more striking if we substitute the distribution function (4) into the kinetic equation (2). The following two partial derivatives
$$\frac{[\delta (𝐯v_0𝐫/r)]}{𝐫}|_𝐯=\mathrm{?}$$
(8)
and
$$\frac{[\delta (𝐯v_0𝐫/r)]}{𝐯}|_𝐫=\mathrm{?}$$
(9)
are rather puzzling. It is obvious that the troublesome situation arises not because we do not know how to differentiate the $`\delta `$-function but because we cannot determine whether or not $`𝐫`$ and $`𝐯`$ correlate.
Up to now, we have seen that such discontinuous distribution functions pose difficulty in the two related respects: (i) The discontinuity spreads throughout almost the entire space instead of “staying” on certain boundaries. (ii) Since the discontinuity spreads the distribution function must contain path information of particles; this is in conflict with the partial-differential equations where $`𝐫`$ and $`𝐯`$ are considered independent of each other. The latter respect, called the paradox of $`𝐫`$-$`𝐯`$ correlation in this letter, holds its significance rather generally.
Before finishing this letter, it is important and relevant to look at whether or not other related theories, such as the continuity equations and Liouville’s theorem, hold in the presence of such discontinuity. (In fact, by referring to the paradox of $`𝐫`$-$`𝐯`$ correlation, readers themselve can come up with the answer to the question without many studies.) Adopting that the force term in (2) is independent of velocity, we find that the collisionless Boltzmann equation (2) is identical to the continuity equation in $`\mu `$-space
$$\frac{f}{t}+\frac{(𝐯f)}{𝐫}+\frac{(\dot{𝐯}f)}{𝐯}=0.$$
(10)
It is then obvious that the continuity equation in $`\mu `$-space does not generally hold for discontinuous distribution functions.
In a similar way, it can be shown that the continuity equation in the grand phase space ($`\mathrm{\Gamma }`$-space), which can be regarded as a natural extension of the continuity equation in $`\mu `$-space, is not valid for systems having discontinuous distribution functions.
Finally, we study a more fundamental issue: Is Liouville’s theorem applicable in statistical mechanics? To get a definite answer to the question and, at the same time, to gain insight into the important subject, it is necessary to examine derivations of the theorem in detail.
In textbooks, one of the derivations is based on the continuity equation in $`\mathrm{\Gamma }`$-space and will be discussed no more. Another derivation, which invokes the Hamiltonian formalism and the Jacobian approach, is worth our special attention.
Suppose that the Hamiltonian of a system with $`2N`$ degrees of freedom is in the form
$$H=\underset{N}{}\frac{p_i^2}{2m}+\mathrm{\Phi }(q_1,q_2,\mathrm{},t).$$
(11)
We have explicitly assumed that the potential term $`\mathrm{\Phi }`$ only contains the position coordinates $`q_1,q_2,\mathrm{}`$ and implicitly assumed that the potential term $`\mathrm{\Phi }`$ is differentiable in terms of these coordinates. Note that these two things have defined the syatem as a pure mechanical one with no dissipative and stochastic forces involved. Suppose that $`d\mathrm{\Gamma }`$ represents a phase volume element in $`\mathrm{\Gamma }`$-space. One version of Liouville’s theorem states that the value of $`d\mathrm{\Gamma }`$ is motion-invariant.
To prove it, we adopt that the time-development of the system is a canonical transformation
$$p(t_0),q(t_0)P(t),Q(t).$$
(12)
In the grand space we have, generally,
$$dqdp=DdQdP,$$
(13)
where
$$D=\frac{(Q,P)}{(q,p)}$$
(14)
is the Jacobian of the transformation. By applying the properties of the canonical transformation, we get (with the details omitted)
$$D=1.$$
(15)
Equation (13) can then be written in the form
$$d\mathrm{\Gamma }_0(t_0)=d\mathrm{\Gamma }(t).$$
(16)
Eq. (16) is regarded as Liouville’s theorem in classical mechanics. To have Liouville’s theorem in statistical mechanics, which states that the grand density $`\rho `$ conserves in statistical processes, several requirements have to get involved. By no coincidence, the requirements are almost exactly the same as those we listed for the invariance of distribution functions: In the entire statistical process the grand density $`\rho `$ has to keep continuous and the external forces, which are associated with $`\mathrm{\Phi }`$ in (11), must be completely free from the dissipative nature and stochastic nature. As has been shown, these requirements cannot be generally fulfilled for realistic statistical systems. All these tell us that Liouville’s theorem should be regarded as a theory in classical mechanics only.
In summary, this letter has revealed that seemingly simple fluids may actually be beyond the scope of the existing fluid mechanics and beyond the scope of the existing kinetic theory.
Discussion with Professor Keying Guan is gratefully acknowledged. His mathematical viewpoint on turbulence is a stimulating factor of this letter. This work is partly supported by the fund provided by Education Ministry, PRC.
|
no-problem/9908/nucl-th9908043.html
|
ar5iv
|
text
|
# Dibaryons with Strangeness: their Weak Nonleptonic Decay using SU(3) Symmetry and how to find them in Relativistic Heavy-Ion Collisions
## Abstract
Weak SU(3) symmetry is successfully applied to the weak hadronic decay amplitudes of octet hyperons. Weak nonmesonic and mesonic decays of various dibaryons with strangeness, their dominant decay modes, and lifetimes are calculated. Production estimates for BNL’s Relativistic Heavy-Ion Collider are presented employing wave function coalescence. Signals for detecting strange dibaryon states in heavy-ion collisions and revealing information about the unknown hyperon-hyperon interactions are outlined.
Relativistic heavy-ion collisions provide a prolific source of strangeness: dozens of hyperons and kaons are produced in central collisions at BNL’s AGS and at CERN SPS (see e.g. ). This opens the exciting perspective of forming composites with multiple units of strangeness hitherto unachievable with conventional methods.
Exotic forms of deeply bound objects with strangeness have been proposed by Bodmer as collapsed states of matter, either consisting of baryons or quarks. A six-quark bag state, the H dibaryon, was predicted by Jaffe . Other bound dibaryon states with strangeness were proposed using quark potentials or the Skyrme model . On the hadronic side, hypernuclei are known to exist already for a long time. The double $`\mathrm{\Lambda }`$ hypernuclear events reported so far are closely related to the H dibaryon . Metastable exotic multihypernuclear objects (MEMOs) as well as purely hyperonic systems of $`\mathrm{\Lambda }`$’s and $`\mathrm{\Xi }`$’s were introduced in as the hadronic counterparts to multistrange quark bags (strangelets) . Most recently, the Nijmegen soft-core potential was extended to the full baryon octet and bound states of $`\mathrm{\Sigma }\mathrm{\Sigma }`$, $`\mathrm{\Sigma }\mathrm{\Xi }`$, and $`\mathrm{\Xi }\mathrm{\Xi }`$ dibaryons were predicted .
One major uncertainty for the detection of such speculative states is their (meta)stability. MEMOs, for example, consists of nucleons, $`\mathrm{\Lambda }`$’s, and $`\mathrm{\Xi }`$ and are metastable by virtue of Pauli-blocking effects. Only two investigations about the weak decay of dibaryons exist so far: In , the H dibaryon was found to decay dominantly by H$`\mathrm{\Sigma }^{}+p`$ for moderate binding energies. The $`(\mathrm{\Lambda }\mathrm{\Lambda })_b`$, which has exactly the same quantum numbers as the H dibaryon, was studied in . Here, the main nonmesonic channel was found to be $`(\mathrm{\Lambda }\mathrm{\Lambda })_b\mathrm{\Lambda }+n`$.
In the following, we will revive an ’old’ approach to calculate weak decay channels and lifetimes of various strange dibaryons using SU(3) symmetric contact interactions. Finally, we present production estimates for RHIC combining transport simulations using Relativistic Quantum Molecular Dynamics, which is widely used for simulations of relativistic heavy ion collisions, with wave function coalescence.
The weak decays of the octet hyperons ($`\mathrm{\Lambda }`$, $`\mathrm{\Sigma }`$, and $`\mathrm{\Xi }`$) can be described by an effective SU(3) symmetric interaction with a parity-violating ($`A`$) and a parity conserving ($`B`$) amplitude . The weak operator is assumed to be proportional to the Gell-Mann matrix $`\lambda _6`$ which ensures hypercharge violation $`|\mathrm{\Delta }Y|=1`$, the $`\mathrm{\Delta }I=1/2`$ rule and the Lee-Sugawara relation for the $`A`$ amplitudes. There are three $`𝒞𝒫`$ invariant terms for the $`A`$ amplitude. One contributes to $`\mathrm{\Sigma }^+n+\pi ^+`$ and can be ignored. The two remaining parameters can be well fitted to the experimental data (see below).
The problem is to describe correctly the $`B`$ amplitudes which defy a consistent explanation. Traditionally, one uses the pole model which in its basic version is not able to describe the experimental measured amplitudes . Various solutions have been proposed to remedy the situation like including the vector meson pole or hyperon resonances . On the other hand, as pointed out in , there is no serious consideration about a contact interaction for the $`B`$ amplitudes in the literature.
General SU(3) symmetry and $`𝒞𝒫`$ invariance results in five independent terms for the $`B`$ amplitudes . We find that one term gives the wrong sign either to the $`B`$ amplitudes for the $`\mathrm{\Lambda }`$ or for the $`\mathrm{\Xi }`$’s. Hence, it must be small compared to the others. Another term gives a contribution to $`\mathrm{\Sigma }^{}n+\pi ^{}`$ and can be neglected. Only three terms remain with coupling constants to be adjusted to the seven measured $`B`$ amplitudes.
The corresponding Lagrangian for both amplitudes reads
$``$ $`=`$ $`D\mathrm{Tr}\overline{B}B[P,\lambda _6]+F\mathrm{Tr}\overline{B}[P,\lambda _6]B+G\mathrm{Tr}\overline{B}P\gamma _5B\lambda _6`$ (2)
$`+H\mathrm{Tr}\overline{B}\lambda _6\gamma _5BP+J\mathrm{Tr}\overline{B}\{P,\lambda _6\}\gamma _5B`$
$`B`$ stands for the baryon octet and $`P`$ for the pseudoscalar nonet. The choice $`D=4.72`$ and $`F=1.62`$ for the $`A`$ amplitudes and $`G=40.0`$, $`H=47.8`$, and $`J=7.1`$ for the $`B`$ amplitudes in units of $`10^7`$ gives a good agreement with the experimental data as shown in Table I. We point out that the $`B`$ amplitudes do not follow a Lee-Sugawara relation .
Using this model for the weak hyperon decay, one can calculate the weak mesonic and nonmesonic decay of strange dibaryons using a Hulthen-like wave function . The meson exchange model for the weak nonmesonic decay of hypernuclei has been proven to be quite successful . We include pion and kaon exchange in our model for the nonmesonic decay as they are the dominant contributions. Effects from short-range contributions like vector meson exchange and direct quark-quark contributions have been found to be less important. We find that the p-wave contributions originating from the $`B`$ amplitudes, the kaon exchange terms and the interference terms are particularly important for the nonmesonic decay channels. Hence, a consistent scheme of both amplitudes turns out to be a crucial ingredience. Clearly, a more fundamental approach is desirable but is at present not at hand before we understand strong interactions at the confinement scale.
For a detection in heavy-ion experiments we are mainly interested in candidates whose final decay products are charged:
$`(\mathrm{\Sigma }^+p)_b`$ $``$ $`p+p`$ (4)
$`(\mathrm{\Xi }^0p)_b`$ $``$ $`p+\mathrm{\Lambda }`$ (5)
$`(\mathrm{\Xi }^0\mathrm{\Lambda })_b`$ $``$ $`p+\mathrm{\Xi }^{}\text{ or }\mathrm{\Lambda }+\mathrm{\Lambda }`$ (6)
$`(\mathrm{\Xi }^0\mathrm{\Xi }^{})_b`$ $``$ $`\mathrm{\Xi }^{}+\mathrm{\Lambda }.`$ (7)
We find that the decay lengths for all of the above strange dibaryons is between $`c\tau 15`$ cm. Fig. 1 shows the calculated branching ratios as a function of the binding energy.
(a): There is only one nonmesonic decay channel for $`(\mathrm{\Sigma }p)_bp+p`$ which we find to be dominant above 5 MeV binding energy. The dibaryon should show up in the invariant $`pp`$ mass spectrum after background subtraction from event-mixing at $`M=2.128\text{ GeV }ϵ`$ where $`ϵ`$ is the binding energy. With this method the weak decay of the lightest hypernucleus $`{}_{\mathrm{\Lambda }}{}^{}{}_{}{}^{3}`$H$`^3`$He$`+\pi ^{}`$ has been detected in heavy-ion collisions by the E864 collaboration .
(b): For the $`(\mathrm{\Xi }^0p)_b`$ bound state only one mesonic but three different nonmesonic channels contribute. The dominant nonmesonic decay turns out to be $`(\mathrm{\Xi }^0p)_b\mathrm{\Lambda }+p`$ already for a binding energy of 2 MeV or more. The decay itself resembles the one for the weak decay of the $`\mathrm{\Xi }^{}`$ or $`\mathrm{\Omega }^{}`$, which have already been detected by several experiments (see contributions in ). Instead of an outgoing $`\pi ^{}`$ or $`K^{}`$ there is a proton leaving the first weak vertex.
(c): The dibaryon $`(\mathrm{\Xi }^0\mathrm{\Lambda })_b`$ decays to $`\mathrm{\Xi }^{}+p`$ and, with a small fraction, to two $`\mathrm{\Lambda }`$’s. Therefore, it can be seen in $`\mathrm{\Xi }^{}p`$ or $`\mathrm{\Lambda }\mathrm{\Lambda }`$ invariant mass plots. One has indeed seen two-Lambda events at the AGS by experiment E896 and experiment WA97 at the SPS has already published two-Lambda correlation functions . There are plans to study the correlation of two $`\mathrm{\Lambda }`$’s on an event-by-event basis at the STAR detector at BNL’s RHIC .
(d): The $`(\mathrm{\Xi }^0\mathrm{\Xi }^{})_b`$ dibaryon has been predicted to be bound and its decay to $`\mathrm{\Xi }^{}+\mathrm{\Lambda }`$ has a branching ratio of a few percent.
The other bound candidates predicted by the Nijmegen model involve weak decays with $`\mathrm{\Sigma }`$ hyperons in the final state. If one can measure neutrons, one is sensitive to all proposed states:
$`(\mathrm{\Sigma }^{}\mathrm{\Sigma }^{})_b`$ $``$ $`\mathrm{\Sigma }^{}+n+\pi ^{}`$ (9)
$`(\mathrm{\Sigma }^+\mathrm{\Sigma }^+)_b`$ $``$ $`\mathrm{\Sigma }^++p`$ (10)
$`(\mathrm{\Xi }^0\mathrm{\Sigma }^+)_b`$ $``$ $`\mathrm{\Sigma }^++\mathrm{\Lambda }`$ (11)
$`(\mathrm{\Xi }^{}\mathrm{\Sigma }^{})_b`$ $``$ $`\mathrm{\Sigma }^{}+\mathrm{\Sigma }^{}`$ (12)
$`(\mathrm{\Xi }^0\mathrm{\Xi }^0)_b`$ $``$ $`\mathrm{\Sigma }^++\mathrm{\Xi }^{}`$ (13)
$`(\mathrm{\Xi }^{}\mathrm{\Xi }^{})_b`$ $``$ $`\mathrm{\Sigma }^{}+\mathrm{\Xi }^{}.`$ (14)
In addition, one can see the nonmesonic decay involving a direct neutron in the final state, like $`(\mathrm{\Lambda }\mathrm{\Lambda })_b\mathrm{\Lambda }+n`$ and $`(\mathrm{\Xi }^{}\mathrm{\Lambda })_b\mathrm{\Xi }^{}+n`$. Thus, a possible $`\mathrm{\Lambda }n`$ or $`\mathrm{\Xi }^{}n`$ invariant mass distribution might reveal important information about the unknown hyperon-hyperon interactions hitherto unaccessible by experiment. We find that the dominant nonmesonic decay for $`(\mathrm{\Lambda }\mathrm{\Lambda })_b`$ is the same as for the H dibaryon, i.e. $`(\mathrm{\Lambda }\mathrm{\Lambda })_b\mathrm{\Sigma }^{}+p`$. This means that the two dibaryons are indistinguishable experimentally. Note, that the nonmesonic decay of the $`(\mathrm{\Lambda }\mathrm{\Lambda })_b`$ always involves a neutral particle in the final state. Searches for the H dibaryon in heavy-ion collisions are indeed sensitive for a weak decay with a $`\mathrm{\Sigma }^{}`$ in the final state and may be utilized to look for other exotic candidates. Especially the weak decay (10) looks very similar to the weak decay of the H dibaryon one is already looking for, but with the opposite sign for the $`\mathrm{\Sigma }`$ hyperon.
Let us now focus on formation probabilities for strange baryon clusters $`\mathrm{\Lambda }\mathrm{\Lambda }`$, $`p\mathrm{\Sigma }^+`$, $`p\mathrm{\Xi }^0`$, $`\mathrm{\Xi }^0\mathrm{\Lambda }`$ and $`\mathrm{\Xi }^0\mathrm{\Xi }^{}`$. The coalescence model provides estimates by simple phase-space arguments. Momentum coalescence has been successful in describing data at low energies (see e.g. ). At relativistic bombarding energies, however, expansion of the source and collective flow have been shown to strongly modify the production rates . Therefore, we will combine source distributions for baryons borrowed from microscopic transport calculations with a coalescence prescription in phase space as detailed in . This procedure has been successful to describe deuteron yields, momentum distributions and is in accord with studies of proton-deuteron correlations . Assuming uncorrelated emission the formation rate can be expressed as
$`{\displaystyle \frac{dN}{d\stackrel{}{P}}}`$ $`=`$ $`g{\displaystyle f_\mathrm{A}(\stackrel{}{x}_1,\stackrel{}{p}_1)f_\mathrm{B}(\stackrel{}{x}_2,\stackrel{}{p}_2)\rho _{\mathrm{AB}}(\mathrm{\Delta }\stackrel{}{x},\mathrm{\Delta }\stackrel{}{p})}`$ (16)
$`\delta (\stackrel{}{P}\stackrel{}{p}_1\stackrel{}{p}_2)d^3x_1d^3x_2d^3p_1d^3p_2`$
where $`\mathrm{\Delta }\stackrel{}{x}=\stackrel{}{x}_1\stackrel{}{x}_2`$ and $`\mathrm{\Delta }\stackrel{}{p}=(\stackrel{}{p}_1\stackrel{}{p}_2)/2`$ are given in the respective 2-body cms (i.e. $`\stackrel{}{P}0`$). One has to multiply the rate with a symmetry factor of 1/2, if the outgoing particles are identical. For the wave function we assume a Hulthen-shape as for the calculations of weak decay properties $`\mathrm{\Psi }(r)=c/r\left(\mathrm{e}^{\kappa r}\mathrm{e}^{\alpha \kappa r}\right)`$. The statistical prefactors $`g`$ account for the lack of information about 2-body correlations with respect to internal degrees of freedom. It includes spin average and the projection on one particular final isospin state of the dibaryon. All strange dibaryons are assumed to be formed in spin-singlet states. The reduction to the correct ’number’ of possible quantum states depends crucially on the assumption of uncorrelated emission and the nature of the bound state. Since the multiparticle correlations during the break up are not well known we consider the g-values as estimates which need further guidance – and insight. The predictions for strange dibaryons are depicted in Fig. 2. Variations in the wave function parameters $`E_b=120`$ MeV ($`\kappa =\sqrt{2\mu E_b}`$) and $`\alpha =26`$ lead only to minor changes in the final result ($`\pm 20\%`$). Therefore, we have chosen to present calculations for the two most extreme parameter-sets ($`E_b5`$ MeV, $`\alpha =2`$) and ($`E_b1`$ MeV, $`\alpha =2`$). The formation of $`\mathrm{\Lambda }\mathrm{\Lambda }`$-states and deuterons (see also ) is diminished by the volume expansion close to midrapidity. For nucleon-hyperon bound states the rapidity shift towards projectile and target is somewhat stronger due to enhanced nuclear freeze-out densities at forward/backward rapidities. Note, that strange dibaryons produced at these rapidities have substantially longer decay lengths which opens the possibility of detecting them at small forward or backward angles. The B-parameter $`B_{\mathrm{AB}}\frac{1}{g_{\mathrm{AB}}}N(\mathrm{A},\mathrm{B})/N(\mathrm{A})N(\mathrm{B})`$ measuring the production rates of dibaryons increases by a factor of two to three comparing $`\mathrm{\Lambda }\mathrm{\Lambda }`$ and $`\mathrm{\Xi }\mathrm{\Xi }`$-states. This enhancement is compatible with an ’early’ freeze-out scenario for multiple strange baryons as argued in : Clusters with high strangeness might be formed more likely, as they decouple earlier from the collisions zone.
There are several searches in heavy-ion collisions for the H dibaryon and for long-lived strangelets with high sensitivities. Hypernuclei have been detected most recently in heavy-ion reactions at the AGS by the E864 collaboration . The dibaryon states studied here are short-lived. They can in principle be detected in present and future experiments by the following means:
1) Experiments with a time-projection chamber can track for unique exotic decays like a charged particle decaying to two charged particles or tracks forming a vertex a few cm outside the target.
2) Experiments sensitive to hyperons can look for peaks in the invariant mass spectrum of $`pp`$, $`p\mathrm{\Lambda }`$, $`\mathrm{\Lambda }\mathrm{\Lambda }`$, $`p\mathrm{\Xi }^{}`$, and $`\mathrm{\Lambda }\mathrm{\Xi }^{}`$ by background subtraction using event mixing.
3) Resonances (unbound states) can be seen in the correlation function of $`\mathrm{\Lambda }\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }\mathrm{\Xi }^{}`$. Two-particle interferometry is a powerful tool to extract information about their (unknown) strong interaction potential as the correlation function depends sensitively on final-state interactions . The Coulomb potential does not mask the strong interactions at low momenta as pointed out in for $`\mathrm{\Lambda }p`$ so that information about the presently unknown hyperon-hyperon forces can be extracted as shown in ).
The STAR experiment at the BNL’s Collider RHIC is able to detect short-lived candidates as well as exotic resonances . One $`(\mathrm{\Lambda }\mathrm{\Lambda })`$ resonance can be seen out of 100 uncorrelated $`\mathrm{\Lambda }`$’s . For the production rates given in Fig. 2 and $`10^6`$ central events, even the bound $`(\mathrm{\Xi }^0\mathrm{\Xi }^{})_b`$ dibaryon can be seen by backtracking for a reconstruction efficiency of only .2 % or better which is indeed feasible for lifetimes around $`10^{10}`$ s .
In this paper we have calculated production rates of strange dibaryons via the coalescence mechanism of independently produced baryons. Finally, we want to point out that another mechanism for their formation might be possible. Via the separation and distillation process, a hot quark-gluon plasma gets enriched with strangeness leading to strangelet creation. If strangelets are unstable they can form a doorway state by decaying to strange dibaryons and increase the production rates.
We are pleased to acknowledge helpful discussions with Ken Barish, Hank Crawford, Carsten Greiner, Huan Huang, Sonja Kabana, Richard Majka, Jamie Nagle, Grazyna Odyniec, Klaus Pretzl, Gulshan Rai, Jack Sandweiss, and Horst Stöcker.
|
no-problem/9908/cond-mat9908409.html
|
ar5iv
|
text
|
# UFRJ–IF–FPC/99 Vortex Reconnection as the Dissipative Scattering of Dipoles
\[
## Abstract
We propose a phenomenological model of vortex tube reconnection at high Reynolds numbers. The basic picture is that squeezed vortex lines, formed by stretching in the region of closest approach between filaments, interact like dipoles (monopole-antimonopole pairs) of a confining electrostatic theory. The probability of dipole creation is found from a canonical ensemble spanned by foldings of the vortex tubes, with temperature parameter estimated from the typical energy variation taking place in the reconnection process. Vortex line reshuffling by viscous diffusion is described in terms of directional transitions of the dipoles. The model is used to fit with reasonable accuracy experimental data established long ago on the symmetric collision of vortex rings. We also study along similar lines the asymmetric case, related to the reconnection of non-parallel vortex tubes.
\]
There is growing evidence that the dynamics of vortex tubes is a necessary ingredient for a deeper understanding of turbulence. This view, initially suggested by images of the vorticity field produced through direct numerical simulations , received strong support from the recent accurate determination of scaling exponents for the velocity structure functions, within a phenomenological theory which places filamentary configurations on a fundamental status . However, the present knowledge on vortex dynamics is still far from being complete, so that even in simplified situations, as in the collision of vortex rings, a formal theory remains to be developed. This difficulty is related in part to the absence of comprehensive phenomenological descriptions that could provide a starting point for more elaborate discussions.
We will focus our attention on the problem of vortex reconnection. Its importance – considered more as an expectation in the long run – relies on the idea that the global structure of turbulent flows may depend on topology changing processes, like the intertwining of closed vortex tubes. Previous theoretical studies on vortex reconnection basically correspond to the case of low and moderate Reynolds numbers. These attempts may be regarded as the counterpart to interesting experimental observations reported by different groups on the collision of vortex rings . On the other side, at higher Reynolds numbers, relevant effects come into play, as stretching and the possible existence of singularities . Another important element to be considered at high Reynolds numbers is that vortex tube evolution is hardly reproducible, due to sensitivity to initial conditions, and one has to resort, therefore, to statistical methods. While it could seem there is no hope of an analytical treatment, since there are no standard techniques to find statistical measures in unstable dynamical systems, we show in this note that the well-known Boltzman distribution appears as a natural candidate from which plausible consequences may be derived.
In an experiment performed about 25 years ago, Fohl and Turner , studied the symmetric collision of two identical vortex rings in water at high Reynolds number ($`Re4000`$), approaching each other with variable angle $`2\theta `$. They found that the fusion of colliding vortex rings, both with radius $`r`$ and velocities $`(0,v\mathrm{sin}\theta ,v\mathrm{cos}\theta )`$ and $`(0,v\mathrm{sin}\theta ,v\mathrm{cos}\theta )`$, into a single ring with velocity $`(0,0,v/2)`$ and radius $`2r`$ allways occurs in a first stage. The fused ring exhibits amplitude oscillations , so that a second stage characterized by a splitting reconnection takes place with probability $`p=p(\theta )`$. The two rings created after the second reconnection move in a plane perpendicular to the initial collision plane. An important feature in the experiment is the existence of a critical angle. For $`\theta >\theta _c16^{}`$, it holds $`p(\theta )=1`$, where as $`p(\theta )0`$ as $`\theta 0`$. An explanation of $`\theta _c`$ was given by Fohl and Turner, based on the fact that the modes of amplitude oscillations, which describe perturbations around a vortex ring of radius $`r`$ and velocity $`v`$, have wavelength $`\lambda _n=2\pi r/n`$ and period
$$T_n(r,v)=\frac{2\pi r}{n(n^21)^{\frac{1}{2}}v},$$
(1)
with $`n2`$. In a collision defined by the angle $`2\theta `$, the projection of the ring’s velocity on the direction transverse to the symmetry plane is $`v\mathrm{sin}\theta `$. One may expect the amplitude of oscillations in the fused ring, which depends on the collision angle, to be $`2r`$ when
$$v\mathrm{sin}\theta =\frac{2r}{T_2(2r,v/2)}.$$
(2)
In this case, diametrically opposite parts of the fused ring will touch each other, producing reconnection. Eq. (2) may be readily solved, yielding $`\theta _c=\mathrm{arcsin}(\sqrt{3}/2\pi )16^{}`$.
There are important questions not answered on this problem. We would like to understand in a more detailed way the form of $`p(\theta )`$, taking into account its behavior at small angles. As we show below, this may be achieved in an effective model where an important role is played by the dynamics of interacting dipoles.
To start, imagine two vortex tubes, locally antiparallel in some neighbourhood $`\mathrm{\Omega }`$, both carrying the same flux and having identical circular cross sections. A description of the physical mechanism underlying reconnection was put forward by Saffman . In his model, the strain field shrinks $`\mathrm{\Omega }`$ in the plane perpendicular to the vortex tubes, so that viscous annihilation of vortex lines occurs, reducing the circulation in $`\mathrm{\Omega }`$. Therefore, a pressure gradient along the vortex’s cores develops, which increases strain and then viscous diffusion, enhancing reconnection in a self-induced way. The equations assumed to describe these steps give reasonable answers, in spite of some disagreements with real and numerical experiments . Saffman’s model is in fact devoted to the situation of two antiparallel vortex tubes interacting at close enough distance. The model works better in the case of strong viscous diffusion (low Reynolds numbers), where vorticity amplification is not very large.
We suggest here a scenario of reconnection at high Reynolds numbers, when strecthing effects become relevant, which probably does not disagree with Saffman’s model, since its application will be related to a different range of physical parameters. The picture we will consider is that in the collision of vortex tubes a system of “dipoles”, emerges after some stretching in a process characterized by very small energy lost. Reconnection is finished with right-angle transitions and subsequent collapse of dipoles, through vortex line reshuffling by viscous diffusion. The main events are depicted in Fig. 1. It is important to keep in mind that dipoles are just effective structures which represent stretched vortex tube segments<sup>*</sup><sup>*</sup>*We mean, throughout this work, an analogy with the electric dipole definition.. In a more rigorous approach the reconnection problem should be addressed in terms of vortex sheets rather than quasi one-dimensional supports of vorticity, since numerical simulations show that vortex tubes flatten in the reconnection region due to stretching. Dipoles have to be regarded only as an useful approximation, from which it is possible to get phenomenological results in the simplest computational way.
We are interested in studying the energy of a flow given by two long vortex tubes, both carrying vorticity flux $`\varphi `$, with stretched segments of length $`\delta `$ and separated by the distance $`d`$, as shown in Fig. 1. Vorticity may be decomposed as $`\stackrel{}{\omega }=\stackrel{}{\omega }^{(1)}+\stackrel{}{\omega }^{(2)}`$, where $`\stackrel{}{\omega }^{(1)}`$ is the field locally amplified by stretching, and $`\stackrel{}{\omega }^{(2)}`$ is the field at other places along the tubes. The energy $`E=E(\delta ,d)`$ may be written as
$$E=E_1+E_2+E_{12},$$
(3)
where, introducing the notation
$$\stackrel{}{\omega },\stackrel{}{\eta }\frac{1}{8\pi }d^3\stackrel{}{x}d^3\stackrel{}{x}^{}\frac{1}{|\stackrel{}{x}\stackrel{}{x}^{}|}\omega _i(\stackrel{}{x})\eta _i(\stackrel{}{x}^{}),$$
(4)
we have
$`E_1=\stackrel{}{\omega }^{(1)},\stackrel{}{\omega }^{(1)},E_2=\stackrel{}{\omega }^{(2)},\stackrel{}{\omega }^{(2)},`$ (5)
$`E_{12}=2\stackrel{}{\omega }^{(1)},\stackrel{}{\omega }^{(2)}.`$ (6)
To simplify expressions, the fator $`1/8\pi `$ in (4) will be supressed henceforth, corresponding to the replacement $`\varphi 2\sqrt{2\pi }\varphi `$. Taking the stretched segments to be identical, both with circular cross section of radius $`ϵ\delta `$, it follows that
$$\frac{E_1}{4\varphi ^2\delta }=\mathrm{ln}(\frac{2\delta }{ϵ})\text{arcsinh}(\frac{\delta }{d})1+[(\frac{d}{\delta })^2+1]^{\frac{1}{2}}\frac{d}{\delta }.$$
(7)
In the computation of $`E_1`$, the kernel in (4) is regularized by means of
$$\frac{1}{|\stackrel{}{x}\stackrel{}{x}^{}|}\frac{1}{[(\stackrel{}{x}\stackrel{}{x}^{})^2+ϵ^2]^{\frac{1}{2}}}.$$
(8)
The contribution to the energy which comes from the interaction between the stretched and non-stretched parts is
$`E_{12}`$ $`=\stackrel{}{\omega }^{(1L)},\stackrel{}{\omega }^{(2L)}+\stackrel{}{\omega }^{(1R)},\stackrel{}{\omega }^{(2R)}`$ (10)
$`+\stackrel{}{\omega }^{(1L)},\stackrel{}{\omega }^{(2R)}+\stackrel{}{\omega }^{(1R)},\stackrel{}{\omega }^{(2L)},`$
with the superscripts L and R denoting the left and right vortex tubes. The effect of screening between the two vortex tubes is evaluated from the following estimative
$$\frac{E_{12}}{4\varphi ^2\delta }\underset{n=1}{\overset{\mathrm{}}{}}\left[\frac{1}{n}\frac{\delta }{[n^2\delta ^2+d^2]^{\frac{1}{2}}}\right].$$
(11)
The above expression is derived from a discretized version of the integral (4) in terms of vortex tube segments of length $`\delta `$, as discussed in ref. , where we additionally used reflection symmetry (invariance of the vorticity field under the interchange L $``$ R. In the reconnection experiments, $`\delta `$ typically fluctuates around some mean value $`\overline{\delta }>d`$. This inequality implies, with (7) and (11), that for $`ϵ/\overline{\delta }1`$ the interaction energy $`\overline{E}_{12}`$ may be neglected, when compared to $`\overline{E}_1`$, and, therefore, $`EE_1+E_2`$ (as an example, consider the choice $`ϵ/\overline{\delta }10^2`$, which gives $`E_{12}(\overline{\delta },d)/E_1(\overline{\delta },d)<0.1`$). We conclude that the positive energy variation due to stretching is local, being compensated by reduction of energy through foldings of the vortex tubes .
There is a close relationship between the stretched vortex tube segments and dipoles of a confining electrostatic theory. Let $`\stackrel{}{p}_1(\stackrel{}{x})`$ and $`\stackrel{}{p}_2(\stackrel{}{x})`$ be the dipole moments of two charge distributions, respectively defined in compact regions $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$. The interaction energy associated to a linear confining potential is
$$E_{int}d^3\stackrel{}{x}d^3\stackrel{}{x}^{}|\stackrel{}{x}\stackrel{}{x}^{}|\stackrel{}{}\stackrel{}{p}_1(\stackrel{}{x})\stackrel{}{}\stackrel{}{p}_2(\stackrel{}{x}^{}).$$
(12)
Through integration by parts we get
$$E_{int}\stackrel{}{p}_1,\stackrel{}{p}_2(\stackrel{}{p}_1(\stackrel{}{x})\stackrel{}{r})(\stackrel{}{p}_2(\stackrel{}{x}^{})\stackrel{}{r})|\stackrel{}{r}|^3,$$
(13)
where $`\stackrel{}{r}=\stackrel{}{x}\stackrel{}{x}^{}`$. Taking $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$ as oriented segments, parallel to the vorticity field and parametrized by $`(0,0,s)`$ and $`(d,0,\delta s)`$, with $`0s\delta `$ and $`\delta d`$, the scalar products in the second term in eq. (13) may be neglected. In this situation the result may be identified with the interaction energy associated to vorticity fields given by $`\stackrel{}{p}_1`$ and $`\stackrel{}{p}_2`$.
An alternative and straightforward way to obtain the connection with confining electrostatics is to represent the dipoles as “monopole-antimonopole” pairs. This is done by replacing the stretched vortex tube segments by monopoles at positions $`(0,0,\delta )`$ and $`(d,0,0)`$ and antimonopoles at $`(0,0,0)`$ and $`(d,0,\delta )`$. These points are just the boundaries of $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$. The monopole (antimonopole) is the source (sink) of a radially symmetric vorticity field, which is not solenoidal – and hence deprived of direct physical meaning. The field of a monopole-antimonopole pair is given as $`\stackrel{}{\omega }=\stackrel{}{\omega }^{(+)}+\stackrel{}{\omega }^{()}`$, where
$$\omega _i^{(\pm )}(\stackrel{}{x})=\frac{\varphi }{4\pi }_i\frac{1}{|\stackrel{}{x}\stackrel{}{x}_\pm |}.$$
(14)
Above, $`\stackrel{}{x}_\pm `$ gives the position of the monopole with charge $`\pm \varphi `$. To compute the energy of a system of monopole-antimonopole pairs, it is necessary to regularize infrared divergencies. Defining the flow inside a large sphere of radius $`\mathrm{\Lambda }\mathrm{}`$, we will have
$$\stackrel{}{\omega }^{(p)},\stackrel{}{\omega }^{(q)}=2\eta (p,q)\varphi ^2(\mathrm{\Lambda }|\stackrel{}{x}_p\stackrel{}{x}_q|),$$
(15)
where $`p,q=\pm `$, and $`\eta (p,q)=1`$ for $`p=q`$, otherwise $`\eta (p,q)=1`$. The energy of the interacting monopole-antimonopole pairs considered here is, thenNote that the infrared divergencies cancel for a neutral system of monopoles.,
$$E=4\varphi ^2[\delta +d(d^2+\delta ^2)^{\frac{1}{2}}].$$
(16)
When $`\delta d`$, the above expression differs from eq. (7), only by self-energy terms, which are independent of $`d`$ (for $`d/\delta >1`$ the agreement is better than $`90\%`$). Since the interaction potential is linear, the force between monopoles has constant strength and is directed along the line joining the particles. This suggests, in the description of stretched vortex tube segments as monopole-antimonopole pairs, that whenever $`\delta >d`$, reconnection takes place. This is in fact the same condition that follows from $`E_1(\delta ,d)>E_1(d,\delta )`$, using the more precise result, eq. (7). These are the energies for the configurations shown in Fig. 1. Furthermore, the energy dissipated in the reconnection process is $`\mathrm{\Delta }E=E_1(\delta ,d)E_1(d,\delta )`$.
To model the symmetric collision of vortex rings, it is necessary to know how $`d`$ depends on the collision angle. This may be obtained by replacing the numerator $`2r`$ in eq. (2) by $`2rd/(2c_1)`$, which we propose to be the oscillation amplitude at angle $`\theta `$, where $`c_1`$ is a phenomenological parameter. We find $`d(\theta )=4c_1r(12\pi \mathrm{sin}\theta /\sqrt{3})`$. As only small angles are involved ($`0\theta \theta _c16^{}`$), this expression may be effectively thought as a linear interpolation between the maximum value $`d4c_1r`$, which occurs at $`\theta =0^{}`$, and the minimum value $`d=0`$, at $`\theta =\theta _c`$. Another important parameter is the mean length of the stretched vortex tube segments. We define it as $`\overline{\delta }=c_2r`$. It is convenient to use dimensionless units where $`\overline{\delta }=1`$ and
$$d(\theta )=4\frac{c_1}{c_2}(1\frac{2\pi }{\sqrt{3}}\mathrm{sin}\theta ).$$
(17)
Due to the additive property of energies, $`E=E_1+E_2`$, we may interpret the random behavior of $`\delta `$ as being related to fluctuations of $`E_1`$ in a canonical ensemble. The elements of the canonical ensemble correspond to folded configurations of the non-stretched parts of the vortex tubes, which act like a reservoir exchanging energy with the stretched region. The probability density to have stretching length $`0\delta \mathrm{}`$ is
$$\rho (\delta ,d)=Z^1\mathrm{exp}[\beta E_1(\delta ,d)],$$
(18)
where $`\beta `$ is the “inverse temperature” and
$$Z=_0^{\mathrm{}}𝑑x\mathrm{exp}[\beta E_1(x,d)].$$
(19)
is the partition function. The definition $`\overline{\delta }=1`$ is used to find the temperature, $`\beta ^1E_1(1.0,d)`$, which is the energy necessary for the creation of stretched segments of length $`\overline{\delta }`$. The probability to have reconnection is, thus,
$$p(\theta )=_{d(\theta )}^{\mathrm{}}𝑑x\rho (x,d(\theta )).$$
(20)
It is not our purpose to derive the above statistical mechanical correspondence from first principles; we take it as a hypothesis. The main problem would be to show that energy fluctuations take place in a time scale much larger than the one of reconnection ($`ϵ^2/\varphi `$, see ). On the other hand, the relaxation time for the vortex system to reach thermal equilibrium has to be much smaller than the time spent for the whole process ($`r^2/\varphi `$), lasting from the fusion of the vortex rings up to the instant of split. Amplitude oscillations may be an important aspect of such an analysis: at large $`n`$, the wave velocity is $`\lambda _n/T_nnv`$, according to eq. (1). Therefore, “thermal equilibrium” would be assured by the fast propagation of perturbations along the vortex tubes. Actually, this equilibrium is not stable, due to the attraction between dipoles. It is worth noting that a more complete study of the right-angle transition of dipoles would probably deal with methods of non-equilibrium statistical mechanics.
We need to set the value of the dimensionless phenomenological parameter $`c_1/c_2`$. This is done in principle by searching for the best agreement with experimental data, but we do not expect this parameter to significantly depart from unit. The reason is that $`\overline{\delta }r`$, as indicated in numerical and real experiments, and $`d(0^{})4r`$, if one assumes that the amplitude of oscillations in the fused ring vanishes when $`\theta =0^{}`$. We note that $`d(0^{})/\overline{\delta }=4c_1/c_2`$ is a quantity suitable of experimental determination. The resulting $`p(\theta )`$ is shown in Fig. 2, with $`c_1/c_2=1.0`$ and $`ϵ=0.01`$. As a matter of fact, the form of $`p(\theta )`$ is not altered in an important way for $`ϵ<0.1`$. In the limit $`ϵ/\overline{\delta }0`$, we may solve eq. (20) exactly, to get that the probability of reconnection decreases exponentially with the distance between the vortex tubes, that is, $`p(\theta )=\mathrm{exp}[d(\theta )/\overline{\delta }]`$.
We may proceed through similar computations to study the asymmetric scattering of vortex rings, aiming at predictions that could be tested in future experiments.
In the symmetric case considered so far, let $`A`$ be the point where the colliding vortex rings first touch. $`A`$ is diametrically opposite to points $`B`$ and $`C`$, which are placed in different rings. The angle at the vertex $`A`$, defined by the segments $`AB`$ and $`AC`$ is $`180^{}2\theta `$. We study the asymmetric collision obtained from the following initial configuration: while the ring which contains $`C`$ is fixed, the other ring is rotated around the axis $`AB`$ by the angle $`\alpha `$. This is precisely the angle between the vortex tubes at the point of contact.
We want to find now the probability $`p(\theta ,\alpha )`$ for the splitting reconnection. As the fused ring evolves, points $`B`$ and $`C`$ move toward each other with relative velocity
$$2v\mathrm{sin}\theta \mathrm{cos}^2(\alpha /2).$$
(21)
This amounts to replacing $`\mathrm{sin}\theta `$ which appears in eq. (17) by $`\mathrm{sin}\theta \mathrm{cos}^2(\alpha /2)`$, to define $`d(\theta ,\alpha )`$, the distance between the non-parallel dipoles. Assuming that the second reconnection occurs close to points $`B`$ and $`C`$, with stretched vortex tubes keeping the initial relative angle $`\alpha `$, all we need to know is the energy of such a configuration. Neglecting the interaction terms in the expression for the energy, which depend on the distance between dipoles, the condition for reconnection is then $`E_1(\delta )>E_1(\mathrm{}(\theta ,\alpha ))`$, where
$$\mathrm{}(\theta ,\alpha )=[d^2(\theta ,\alpha )+\delta ^2\mathrm{sin}^2(\alpha /2)]^{\frac{1}{2}}.$$
(22)
Therefore, reconnection occurs when $`\delta >\mathrm{}(\theta ,\alpha )`$, or, equivalently, $`\delta >d(\theta ,\alpha )/\mathrm{cos}(\alpha /2)`$. We just have to replace the lower bound $`d(\theta )`$ in the integral (20) by $`d(\theta ,\alpha )/\mathrm{cos}(\alpha /2)`$. We carried out computations using the same set of parameters $`\beta `$ and $`c_1/c_2`$ for the former case ($`\alpha =0`$). It is possible that deviations grow as $`\alpha `$ gets larger, where the dipole model may loose its applicability. However, there is some indication that reconnection is supressed at large $`\alpha `$ , which is also verified through an explicit computation of $`p(\theta ,\alpha )`$. In Fig. 3 $`p(\theta ,\alpha )`$ is shown with $`0\alpha 180^{}`$, for $`\theta =15^{}`$ and $`\theta =17^{}`$. The latter situation is perhaps the more interesting, because of the plateau given by $`p(\theta ,\alpha )=1`$ at small values of $`\alpha `$.
To summarize, we studied the problem of vortex reconnection at high Reynolds numbers, where stretching effects become important. A simple correspondence with confining eletrostatics and statistical mechanics allowed us to investigate the way how “dipoles”, i.e., stretched vortex tube segments, behave in the process of reconnection. The initial dipole configuration evolves, thanks to diffusion, towards a state which links the interacting vortex tubes. The measurements of Fohl and Turner for the probability of a splitting reconnection after the initial merger of vortex rings are successfully accounted by the present model. We also defined some predictions concerning the case of asymmetric collisions, to be compared with possible future experimental observations.
Acknowledgments
This work was partially supported by CNPq.
FIGURE CAPTIONS
Fig. 1
Sketch of the configuration of the vortex tubes in the region of closest approximation. An attractive force betwen the oppositely oriented stretched segments (“dipoles”) is followed by vorticity cancelation (not represented in the picture) and then by the right-angle transition of the dipoles.
Fig. 2
The probability $`p(\theta )`$ for the occurrence of a splitting reconnection. Dots represent the values measured by Fohl and Turner. The continuous line is the prediction of the dipole model obtained with $`\beta =1/E_1(1.0,d)`$, $`c_1/c_2=1.0`$, and $`ϵ=0.01`$.
Fig. 3
The probability $`p(\theta ,\alpha )`$ for the ocurrence of a splitting reconnection in the asymmetric case, where $`\theta `$ is fixed and $`0\alpha 180^{}`$. a) $`\theta =15^{}`$. b) $`\theta =17^{}`$.
|
no-problem/9908/cond-mat9908100.html
|
ar5iv
|
text
|
# Phase transition from a 𝑑_{𝑥²-𝑦²} to 𝑑_{𝑥²-𝑦²}+𝑑_{𝑥𝑦} superconductor
\[
## Abstract
We study the phase transition from a $`d_{x^2y^2}`$ to $`d_{x^2y^2}+d_{xy}`$ superconductor using the tight-binding model of two-dimensional cuprates. As the temperature is lowered past the critical temperature $`T_c`$, first a $`d_{x^2y^2}`$ superconducting phase is created. With further reduction of temperature, the $`d_{x^2y^2}+d_{xy}`$ phase is created at temperature $`T=T_{c1}`$. We study the temperature dependencies of the order parameter, specific heat and spin susceptibility in these mixed-angular-momentum states on square lattice and on a lattice with orthorhombic distortion. The above-mentioned phase transitions are identified by two jumps in specific heat at $`T_c`$ and $`T_{c1}`$.
PACS number(s): 74.20.Fg, 74.62.-c, 74.25.Bt
\]
Inspite of many theoretical and experimental studies on high-$`T_c`$ cuprates the exact symmetry of the order parameter is still a subject of active research . However, there is evidence that the cuprates have singlet $`d`$-wave Cooper pairs and the order parameter has $`d_{x^2y^2}`$ symmetry in two dimensions . Recent measurements of penetration depth and superconducting specific heat at different temperatures $`T`$ and related theoretical analyses also support this. However, several phase-sensitive measurements of the order parameter of the cuprates indicate a significant mixing of a distinct angular momentum component with a predominant $`d_{x^2y^2}`$ state at temperatures below a second critical temperature $`T_{c1}`$. For temperatures between $`T_{c1}`$ and $`T_c`$ only the $`d_{x^2y^2}`$ state survives. Below $`T_{c1}`$ the order parameter can have a mixed-symmetry state of type $`d_{x^2y^2}+\mathrm{exp}(i\theta )\chi `$, where $`\chi `$ represents a state of different symmetry. The most probable possibilities for $`\chi `$ are the $`s`$ or $`d_{xy}`$ wave. The possibility of a mixed $`(sd)`$-wave symmetry was first suggested theoretically by Ruckenstein et al. and Kotliar .
There are experimental evidences based on Josephson supercurrent for tunneling between a conventional $`s`$-wave superconductor (Pb) and twinned or untwinned single crystals of YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (YBCO) that YBCO has mixed $`d_{x^2y^2}\pm s`$ or $`d_{x^2y^2}\pm is`$ symmetry at lower temperatures. Recently, the existence of these mixed-symmetry states has been explored to explain the nuclear magnetic resonance data in the superconductor YBCO and the Josephson critical current observed in YBCO-SNS and YBCO-Pb junctions .
Kouznetsov et al. performed some c-axis Josephson tunneling experiments by depositing conventional superconductor (Pb) across a single twin boundary of a YBCO crystal. By measuring the critical current as a function of the angle and magnitude of a magnetic field applied in the plane of the junction they also found the evidence of a mixed-symmetry order parameter in YBCO involving $`d_{x^2y^2}`$ and $`s`$ waves. By measuring the microwave complex conductivity in the superconducting state of high quality YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-δ</sub> single crystals at 10 GHz using a high-Q Nb cavity Sridhar $`etal`$ also suggested the existence of a multicomponent superconducting order parameter in YBCO .
A similar conclusion of the existence of mixed-symmetry states may also be obtained based on the results of angle-resolved photoemission spectroscopy experiment by Ma et al. in which a temperature dependent gap anisotropy in oxygen-annealed Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+x</sub> was detected . The measured gaps along directions $`\mathrm{\Gamma }M`$ and $`\mathrm{\Gamma }X`$ are non-zero at low temperatures and their ratio was strongly temperature dependent. Using Ginzburg-Landau theory, Betouras and Joynt demonstrated that one way of explaining this behavior is to employ a mixed-symmetry state of the $`d_{x^2y^2}+s`$-wave type. They also conclude that the actual symmetry of the order parameter should vary substantially from one compound to another and for different levels of doping. This also suggests the possible appearance of a $`d_{x^2y^2}+d_{xy}`$ state under favorable conditions.
More recently, Krishana et al. reported a phase transition in the superconductor Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> induced by a magnetic field from a study of the thermal conductivity as a function of temperature and applied field. Laughlin provided a theoretical explanation of the observation by Krishana et al. that for weak magnetic field a time-reversal symmetry breaking state of mixed symmetry is induced in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub>. From a study of vortex in a $`d`$-wave superconductor using a self-consistent Bogoliubov-de Gennes formalism, Franz and Teśanović also predicted the possibility of a superconducting state of mixed symmetry. This mixed-symmetry state is likely to be a minor $`s`$ or $`d_{xy}`$ component superposed on a $`d_{x^2y^2}`$ state for $`T<T_{c1}`$.
From different experimental observations it is now generally accepted that a time-reversal symmetry breaking state of type $`d_{x^2y^2}+i\chi `$ is possible in the presence of an external field or magnetic impurity. This mixed-symmetry state is observed close to these impurities, surfaces/twin boundaries in the ab-plane or vortices. The nature of the mixed state varies from compound to compound. There are physical reasons for the appearance of these states. Either spin-orbit coupling with magnetic impurities or Andreev reflected bound states which create internal currents at the boundaries is responsible for these states . However, orthorhombicity plays a crucial role in the generation of time-reversal symmetric mixed states. For example, it is established from a Ginzburg-Landau functional analysis that the orthorhombicity has a consequence in the development of a $`d+s`$ state instead of a time-reversal symmetry broken one. Moreover, from a theoretical point of view, time-reversal symmetric states of type $`d_{x^2y^2}+\chi `$ are expected to be allowed depending on the orthorhombic distortion.
There have been some studies on the phase transition to a $`d_{x^2y^2}+\mathrm{exp}(i\theta )\chi `$ phase from a $`d_{x^2y^2}`$ phase with $`\theta =\pi /2`$ and $`\chi =d_{xy}`$ or a $`s`$ state. From theoretical considerations we find that there are two possibilities for the phase $`\theta `$: 0 or $`\pi /2`$. For $`\theta =0`$, we find numerically that there is no stable $`d_{x^2y^2}+s`$ phase. Here we study the phase transition to a $`d_{x^2y^2}+d_{xy}`$ phase from a $`d_{x^2y^2}`$ phase below $`T_{c1}`$. In particular we study the temperature dependencies of the order parameter, specific heat, and spin susceptibility in the mixed-symmetry state.
There is no suitable microscopic theory for high-$`T_c`$ superconductors and there is controversy about a proper description of the normal state and the pairing mechanism for such materials . In the absence of a microscopic theory, a phenomenological tight-binding model in two dimensions with the proper lattice symmetry will be used . This model has been successful in describing many properties of high-$`T_c`$ materials.
We study the temperature dependencies of specific heat and susceptibility of a $`d_{x^2y^2}+d_{xy}`$-wave superconductor with a weaker $`d_{xy}`$ wave both on square lattice and on a lattice with orthorhombic distortion. The order parameter of a $`d_{x^2y^2}+d_{xy}`$-wave superconductor has nodes on the Fermi surface and changes sign across it, and consequently, its superconducting observables also exhibit power-law dependencies on temperature. On the other hand, the order parameters for the mixed $`d_{x^2y^2}+is`$ and $`d_{x^2y^2}+id_{xy}`$-wave states do not have a node on the Fermi surface and the corresponding observables have a exponential dependencies on temperature. In the present study on $`d_{x^2y^2}+d_{xy}`$-wave states the specific heat exhibits two jumps at $`T=T_{c1}`$ and $`T=T_c`$, which clearly exhibits the phase transition at $`T_{c1}`$.
In the present two-dimensional tight binding model the effective interaction $`V_{\mathrm{𝐤𝐪}}`$ for transition from a momentum $`𝐪`$ to $`𝐤`$ is taken to be separable, and is expanded in terms of some general basis functions $`\eta _{i𝐤}`$, labelled by the index $`i`$, as $`V_{\mathrm{𝐤𝐪}}=_iV_i\eta _{i𝐤}\eta _{i𝐪}`$ . The functions $`\eta _{i𝐤}`$ are associated with a one dimensional irreducible representation of the point group of square lattice $`C_{4v}`$ and are appropriate generalizations of the circular harmonics incorporating the proper lattice symmetry. The effective interaction after including the two appropriate basis functions for singlet pairing is taken as
$$V_{\mathrm{𝐤𝐪}}=V_1\eta _{1𝐤}\eta _{1𝐪}V_2\eta _{2𝐤}\eta _{2𝐪},$$
(1)
where $`\eta _{1𝐪}(\mathrm{cos}q_x\beta \mathrm{cos}q_y)`$ corresponds to $`d_{x^2y^2}`$ symmetry, $`\eta _{2𝐪}\mathrm{sin}q_x\mathrm{sin}q_y`$ corresponds to $`d_{xy}`$ symmetry, and where $`\beta =1`$ corresponds to a square lattice, and $`\beta 1`$ represents orthorhombic distortion. In this case the quasiparticle dispersion relation is given by $`ϵ_𝐤=2t[\mathrm{cos}k_x+\beta \mathrm{cos}k_y\gamma \mathrm{cos}k_x\mathrm{cos}k_y],`$ where $`t`$ and $`\beta t`$ are the nearest-neighbour hopping integrals along the in-plane $`a`$ and $`b`$ axes, respectively, and $`\gamma t/2`$ is the second-nearest-neighbour hopping integral. The energy $`ϵ_𝐤`$ is measured with respect to the Fermi surface.
At a finite $`T`$, one has the following BCS equation
$`\mathrm{\Delta }_𝐤`$ $`=`$ $`{\displaystyle \underset{𝐪}{}}V_{\mathrm{𝐤𝐪}}{\displaystyle \frac{\mathrm{\Delta }_𝐪}{2E_𝐪}}\mathrm{tanh}{\displaystyle \frac{E_𝐪}{2k_BT}}`$ (2)
with $`E_𝐪=[(ϵ_𝐪E_F)^2+|\mathrm{\Delta }_𝐪|^2]^{1/2},`$ where $`E_F`$ is the Fermi temperature and $`k_B`$ the Boltzmann’s constant. The order parameter has the following anisotropic form:
$$\mathrm{\Delta }_𝐪\mathrm{\Delta }_1\eta _{1𝐪}+C\mathrm{\Delta }_2\eta _{2𝐪},$$
(3)
where $`C`$ is a complex number of unit modulas $`|C|^2=1`$. If we substitute Eqs. (1) and (3) into the BCS equation (2), one can separate the resultant equation in its real and imaginary parts. The resultant equations only have solution for real $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$, when the complex parameter $`C`$ is either purely real or purely imaginary. The solution for purely imaginary $`C`$, e.g., $`C=i`$ have been extensively studied in relation to mixed $`d_{x^2y^2}+is`$ and $`d_{x^2y^2}+id_{xy}`$ states . Here we consider the solution for $`C=1`$, for $`d_{x^2y^2}+d_{xy}`$ state. Using the form (3) of $`\mathrm{\Delta }_𝐪`$ with $`C=1`$ and potential (1), Eq. (2) becomes the following coupled set of BCS equations
$`\mathrm{\Delta }_1=V_1{\displaystyle \underset{𝐪}{}}{\displaystyle \frac{\eta _{1𝐪}[\mathrm{\Delta }_1\eta _{1𝐪}+\mathrm{\Delta }_2\eta _{2𝐪}]}{2E_𝐪}}\mathrm{tanh}{\displaystyle \frac{E_𝐪}{2k_BT}}`$ (4)
$`\mathrm{\Delta }_2=V_2{\displaystyle \underset{𝐪}{}}{\displaystyle \frac{\eta _{2𝐪}[\mathrm{\Delta }_1\eta _{1𝐪}+\mathrm{\Delta }_2\eta _{2𝐪}]}{2E_𝐪}}\mathrm{tanh}{\displaystyle \frac{E_𝐪}{2k_BT}},`$ (5)
where both the interactions $`V_1`$ and $`V_2`$ are assumed to be energy-independent constants for $`|ϵ_𝐪E_F|<k_BT_D`$ and zero for $`|ϵ_𝐪E_F|>k_BT_D`$, where $`k_BT_D`$ is a purely mathematical cutoff introduced to eliminate the ultraviolet divergence in the BCS equation and should be compared with the physically motivated Debye cutoff in the case of the conventional superconductors.
We solved the coupled set of equations (4) and (5) numerically and calculated the gaps $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$ at various temperatures for $`T<T_c`$. We have performed calculations (1) on a perfect square lattice and (2) in the presence of an orthorhombic distortion with cut off $`k_BT_D=0.02586`$ eV ($`T_D=300`$ K) in both cases. The parameters for these two cases are the following: (1) Square lattice $``$ (a) $`t=0.2586`$ eV, $`\beta =1`$, $`\gamma =0`$, $`V_2=8.5t`$, and $`V_1=0.73t`$, $`T_c=71`$ K, $`T_{c1}`$ = 28 K; (b) $`t=0.2586`$ eV, $`\beta =1`$, $`\gamma =0`$, $`V_2=9.0t`$, and $`V_1=0.73t`$, $`T_c=71`$ K, $`T_{c1}`$ = 55 K; (2) Orthorombic distortion $``$ (a) $`t=0.2586`$ eV, $`\beta =0.95`$, and $`\gamma =0`$, $`V_2=8.35t`$, and $`V_1=0.97t`$, $`T_c`$ = 70 K, $`T_{c1}`$ = 30 K; (b) $`t=0.2586`$ eV, $`\beta =0.95`$, and $`\gamma =0`$, $`V_2=8.7t`$, and $`V_1=0.97t`$, $`T_c`$ = 70 K, $`T_{c1}`$ = 50 K. For a very weak $`d_{x^2y^2}`$-wave ($`d_{xy}`$-wave) coupling the only possible solution corresponds to $`\mathrm{\Delta }_1=0`$ ($`\mathrm{\Delta }_2=0`$). We have studied the solution only when a coupling is allowed between Eqs. (4) and (5).
Fig. 1. The order parameters $`\mathrm{\Delta }_1`$, $`\mathrm{\Delta }_2`$ in Kelvin at different temperatures for models 1(a) (full line) and 1(b) (dashed line) on square lattice described in text.
Fig. 2. The same as in Fig. 1 at different temperatures for models 2(a) (full line) and 2(b) (dashed line) in the presence of orthorhombic distortion described in text.
In Figs. 1 and 2 we plot the temperature dependencies of different $`\mathrm{\Delta }`$’s for the following two sets of $`d_{x^2y^2}+d_{xy}`$ wave corresponding to models 1 and 2, respectively. In all cases, with the lowering of temperature passed $`T_c`$, the parameter $`\mathrm{\Delta }_1`$ increases up to $`T=T_{c1}`$. As $`T`$ is lowered further, $`\mathrm{\Delta }_2`$ becomes nonzero at $`T=T_{c1}`$ and begins to increase. As temperature is lowered, both $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$ first increase and then attain a constant value at zero temperature.
The different superconducting and normal specific heats are plotted in Figs. 3 and 4 for square lattice \[models 1(a) and 1(b)\] and orthorhombic distortion \[models 2(a) and 2(b)\], respectively. In both cases the specific heat exhibits two jumps $``$ one at $`T_c`$ and another at $`T_{c1}`$. From Figs. 1 and 2 we see that the temperature
Fig. 3. Specific heat ratio $`C(T)/C_n(T_c)`$ versus $`T/T_c`$ for models 1(a) and 1(b) on square lattice: 1(a) (full line), 1(b) (dashed line), $`d_{xy}`$ (dotted line), normal (dashed-dotted line).
Fig. 4. Specific heat ratio $`C(T)/C_n(T_c)`$ versus $`T/T_c`$ for models 2(a) and 2(b) in the presence of orthorhombic distortion: 2(a) (full line), 2(b) (dashed line), $`d_{xy}`$ (dotted line), normal (dashed-dotted line).
derivative of $`|\mathrm{\Delta }_𝐪|^2`$ has discontinuities at $`T_c`$ and $`T_{c1}`$ due to the vanishing of $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$, respectively, responsible for the two jumps in specific heat (see, definition in Ref. ). For a pure $`d_{x^2y^2}`$ wave we find that the specific heat exhibits a power-law dependence on temperature. However, the exponent of this dependence varies with temperature. For small $`T`$ the exponent is approximately 2.5, and for large $`T`$ ($`TT_c`$) it is nearly 2. For the mixed $`d_{x^2y^2}+d_{xy}`$-wave model, for $`T_c>T>T_{c1}`$ the specific heat exhibits $`d`$-wave power-law behavior. For $`d`$-wave models $`C_s(T_c)/C_n(T_c)`$ is a function of $`T_c`$ and $`\beta `$. In Figs. 2 and 3 this ratio for the $`d_{x^2y^2}`$-wave case, for $`T_c`$ = 70 K, is approximately 3 (2.44) for $`\beta =`$ 1 (0.95). For the $`d_{xy}`$-wave case, for $`T_c`$ = 70 K, this ratio is approximately 1.81 (1.9) for $`\beta =`$ 1 (0.95). In a continuum calculation this ratio was 2 in the absence of a van Hove singularity .
Fig. 5. Susceptibility ratio $`\chi (T)/\chi (T_c)`$ versus $`T/T_c`$ for square lattice: pure $`d_{x^2y^2}`$ wave (solid line), pure $`d_{xy}`$ wave (dashed line), model 1(a) (dotted line), and model 1(b) (dashed-double-dotted line).
Fig. 6. Susceptibility ratio $`\chi (T)/\chi (T_c)`$ versus $`T/T_c`$ in the presence of orthorhombic distortion: pure $`d_{x^2y^2}`$ wave (solid line), pure $`d_{xy}`$ wave (dashed line), model 1(a) (dotted line), and model 1(b) (dashed-double-dotted line).
Next we exhibit the temperature dependence of spin susceptibility (defined in Ref. ) in Figs. 5 and 6 where we also plot the results for pure $`d_{x^2y^2}`$ and $`d_{xy}`$ waves for comparison. In Figs. 5 and 6, we show the results for models 1 and 2 on square lattice and with orthorhombic distortion, respectively. For pure $`d_{x^2y^2}`$ and $`d_{xy}`$ waves we obtain power-law dependencies on temperature. The exponent for this power-law scaling was independent of critical temperature $`T_c`$ but varied from a square lattice to that with an orthorhombic distortion. For $`d_{x^2y^2}`$ wave, the exponent for square lattice (orthorhombic distortion, $`\beta `$ = 0.95) is 2.6 (2.4). For $`d_{xy}`$ wave, the exponent for square lattice (orthorhombic distortion, $`\beta `$ = 0.95) is 1.1 (1.6). For the mixed $`d_{x^2y^2}+d_{xy}`$ wave these exponents are nearly identical to the pure $`d_{x^2y^2}`$ wave case. Hence, by studying the temperature dependency of spin susceptibility, it will be impossible to detect the phase transition at $`T=T_{c1}`$ from a $`d_{x^2y^2}`$ wave to a $`d_{x^2y^2}+d_{xy}`$ wave, at least within the present tight-binding model.
In conclusion, we have studied the $`d_{x^2y^2}+d_{xy}`$-wave superconductivity employing the two-dimensional tight binding BCS model on square lattice and also on a lattice with orthorhombic distortion and confirmed a second second-order phase transition at $`T=T_{c1}`$ in the presence of a weaker $`d_{xy}`$ wave. This phase transition is marked by a jump in the specific heat at $`T=T_{c1}`$. We have kept the $`s`$\- and $`d`$-wave couplings in such a domain that a coupled $`d_{x^2y^2}+d_{xy}`$-wave solution is allowed. The $`d_{x^2y^2}+d_{xy}`$-wave state is similar to a $`d_{x^2y^2}`$-wave-type state with nodes on the Fermi surface in the order parameter. Consequently, we find power-law temperature dependencies of specific heat and spin susceptibility in the $`d_{x^2y^2}+d_{xy}`$ wave. The exponents of these power laws for the mixed $`d_{x^2y^2}+d_{xy}`$ wave are very close to those for the pure $`d_{x^2y^2}`$ wave.
The work was supported by the CNPq and FAPESP.
|
no-problem/9908/cond-mat9908415.html
|
ar5iv
|
text
|
# Dynamics of simulated water under pressure
## I Introduction
The “slow dynamics” and glass transition of both simple and molecular liquids has been a topic of significant interest in recent years. The initial slowing down of liquids at temperatures down to $`T_c1.2T_g`$, where relaxation times approach $`1`$ ns, has been well described by the mode-coupling theory (MCT) . MCT has been successfully applied to a wide variety of real and model systems , including hard spheres , Ni<sub>80</sub>P<sub>20</sub> , SiO<sub>2</sub> , and polymer melts . However, there has not been an extensive test of the validity of the MCT predictions for a model system over a wide range of pressures and along different thermodynamic paths.
At low pressure, it was shown previously that the power-law behavior of dynamic properties in the SPC/E model can be explained using MCT . Furthermore, the possible relationship between the experimentally-observed power-law behavior and the predictions of MCT has been discussed . The experimentally-observed locus of apparent power-law singularities of dynamic and thermodynamic properties \[Fig. 1\] is of particular interest , and has catalysed the development of three scenarios to explain the anomalous properties of water: (i) the existence of a spinodal bounding the stability of the liquid in the superheated, stretched, and supercooled states ; (ii) the existence of a liquid-liquid phase transition line separating two liquid phases differing in density ; (iii) a singularity-free scenario in which the thermodynamic anomalies are related to the presence of low-density and low-entropy structural heterogeneities . The predictions of MCT are of interest since MCT might account for the apparent power-law behavior of dynamic properties on cooling, thereby removing the need for a thermodynamic explanation of the dynamic properties of water.
In this article we focus on two related issues: (i) the possibility of using MCT to explain the slow dynamics of water under pressure, and (ii) a test of the validity of the MCT predictions over an extremely wide pressure range in a system with dramatic structural changes. We find that MCT provides a good account of the slow dynamics of the SPC/E model for water at all pressures, with the structure evolving continuously from an open tetrahedral network to a densely packed fluid, similar to a Lennard-Jones type liquid. By examining the wave-vector dependence of collective dynamics, we are able to discover how these structural changes are reflected in the dynamic behavior of the liquid. We are also able to test the validity of the relationship predicted by MCT for the diffusivity exponent $`\gamma `$ and the von Schweidler exponent $`b`$ over a wide range of values $`\gamma `$ and $`b`$ \[Fig. 2\]. Our results support the predicted relationship of these exponents. A brief report of a subset of the present results for the SPC/E potential has recently appeared . The dynamic properties of the ST2 potential at one pressure, and of the TIP4P potential at several pressures, have also recently been discussed.
## II Mode Coupling Theory
We will focus our discussion on the idealized form of MCT, originally formulated to describe spherically-symmetric potentials. Recent extensions have been made to account for the rotational motion present in non-spherical molecular systems , such as water. The idealized version of MCT has been shown to provide a good account for the center-of-mass motion for the SPC/E model . We provide only a brief account of the MCT predictions relevant to the results of this article, and we refer the reader to extensive reviews for more information .
MCT assumes that localization, or “caging” of molecules due to the slow rearrangement of neighboring molecules, is the source of the dramatic increase of relaxation times on cooling, leading to a strong coupling between single particle motion and the density fluctuations of the liquid. Indeed, according to MCT, the static density fluctuations, measured by the structure factor $`S(q)`$, entirely determine the long time dynamic behavior. MCT accounts for the loss of correlation by the interaction of density mode fluctuations, ignoring other possible mechanisms for relaxation. MCT predicts the asymptotic power-law divergence of correlation times, and power-law vanishing of the diffusion constant
$$DD_0(T/T_c1)^\gamma $$
(1)
at a critical temperature $`T_c=T_c(P)`$, where we refer to $`\gamma =\gamma (P)`$ as the diffusivity exponent. In real systems, “freezing” of the system dynamics is avoided at $`T_c`$, as
relaxation mechanisms not accounted for by MCT become significant. However, $`T_c`$ can still be interpreted as a “crossover temperature” where the dynamics change from being dominated by density fluctuations to being controlled by “activated” processes. Some recent work has also demonstrated the significance of $`T_c`$ as a crossover temperature where relaxation occurs primarily through basin hopping , in the energy landscape view of liquid dynamics .
MCT predicts that the Fourier transform of the density-density correlation function or intermediate scattering function
$$F(q,t)\frac{1}{S(q)}\underset{j,k=1}{\overset{N}{}}e^{i𝐪[𝐫_k(t)𝐫_j(0)]}$$
(2)
decays via a two-step process. In the first relaxation step, $`F(q,t)`$ approaches a plateau value $`F_{\text{plateau}}(q)`$ which is described, to leading order in time, by a power law with exponent $`a=a(P)`$,
$$F(q,t)F_{\text{plateau}}(q)t^a,$$
(3)
At larger times, $`F(q,t)`$ decreases from $`F_{\text{plateau}}(q)`$ and MCT predicts the decay obeys the von Schweidler power law to leading order in time
$$F_{\text{plateau}}(q)F(q,t)t^b,$$
(4)
where $`b=b(P)`$ is known as the von Schweidler exponent. The region of validity of Eqs. (3) and (4) can be quite limited.
The slow relaxation of $`F(q,t)`$ has a characteristic relaxation time $`\tau `$ that is also predicted to have asymptotic power law dependence on temperature,
$$\tau \tau _0(T/T_c1)^\gamma $$
(5)
with the same value of the exponent $`\gamma `$ as for the diffusion constant. Hence, Eqs. (1) and (5) predict that the product $`D\tau `$ is not singular as $`TT_c`$, hence we take the product to be constant over the range that Eqs. (1) and (5) are valid (neglecting corrections to scaling).
MCT predicts that the scaling exponents $`a`$, $`b`$, and $`\gamma `$ are not independent; $`a`$ and $`b`$ are related by the exponent parameter $`\lambda `$ using the relationship
$$\lambda =\frac{[\mathrm{\Gamma }(1a)]^2}{\mathrm{\Gamma }(12a)}=\frac{[\mathrm{\Gamma }(1b)]^2}{\mathrm{\Gamma }(1+2b)}$$
(6)
where $`\mathrm{\Gamma }(x)`$ is the gamma function. MCT also relates $`\gamma `$ to $`a`$ and $`b`$ via
$$\gamma =\frac{1}{2a}+\frac{1}{2b}.$$
(7)
Because of Eqs. (6) and (7), only one exponent value is needed to determine all others, so calculation of two exponents determines if the dynamics of a system are consistent with the predictions of MCT. Furthermore, these exponents are expected to depend on the path along which $`T_c`$ is approached.
After $`F(q,t)`$ departs from the plateau, $`F(q,t)`$ is well-described by a Kohlrausch-Williams-Watts stretched exponential
$$F(q,t)=A(q)\mathrm{exp}\left[\left(\frac{t}{\tau (q)}\right)^{\beta (q)}\right],$$
(8)
where $`\tau (q)`$ is the relevant relaxation time. Moreover, it has been shown that the exponent $`\beta =\beta (q)`$ is related to the von Schweidler exponent
$$\underset{q\mathrm{}}{lim}\beta (q)=b.$$
(9)
This relation facilitates evaluation of $`b`$, since the region of validity of Eq. (4) is difficult to identify in practice.
## III Simulations
We perform MD simulations of 216 water molecules interacting via the SPC/E pair potential . The SPC/E model treats water as a rigid molecule consisting of three point charges located at the atomic centers of the oxygen and hydrogen, which have an OH distance of 1.0 Å and HOH angle of 109.47, the tetrahedral angle. Each hydrogen has charge $`q_H=0.4238e`$, where $`e`$ is the fundamental unit of charge, and the oxygen has charge $`q_O=2q_H`$. In addition, the oxygen atoms of separate molecules interact via a Lennard-Jones potential with parameters $`\sigma =3.166`$ Å and $`ϵ=0.6502`$ kJ/mol.
Our simulation results are summarized in Table I. For $`T300`$ K, we simulate two independent systems to improve statistics, as the long relaxation time makes time averaging more difficult. We equilibrate all simulated state points to constant $`T`$ and $`\rho `$ by monitoring the pressure and internal energy. We control the temperature using the Berendsen method of rescaling the velocities , while the reaction field technique with a cutoff of 0.79 nm accounts for the long-range Coulombic interactions. The equations of motion evolve using the SHAKE algorithm with a time step of 1 fs, except at $`T=190`$, where a time step of 2 fs is used due to the extremely slow motion of the molecules. Equilibration times at high temperatures are relatively small. At low $`T`$, extremely long equilibration times are needed. The structural and thermodynamics properties may be obtained after relatively short equilibration times. However, dynamic properties show significant aging effects (i.e. dependence of measured properties on the chosen starting time) if great care is not taken in equilibration.
For production runs, it is desirable to make measurements in the isoenergetic/isochoric ensemble (NVE). However, a small energy drift is unavoidable for the long runs presented here, so we again employ the heat bath of Berendsen, using a relaxation time of 200ps . The large relaxation time prevents an energy drift but achieves results that are very close to those that would be found if it were possible to perform a simulation in the NVE ensemble.
Since we perform long runs for many state points, we store the molecular trajectories $`\{𝐫_i,𝐩_i\}`$ at logarithmic intervals to avoid storage problems that linear sampling presents. Specifically, we sample configurations at times growing in powers of $`2`$ up a maximum time $`t_{\text{max}}`$. We begin a new sampling cycle each time $`t_{\text{max}}`$ (relative to the cycle starting time) is reached. This sampling method allows for calculation of dynamic properties on time scales spanning eight orders of magnitude (from 1 fs to 100 ns) using a relatively small amount of disk space. Still, more then 2 GB of storage was required for storing configurations at $`T=210K`$. Our simulations have a speed of approximately 200 $`\mu `$s per update per molecule on a MIPS R10000 processor, representing a total calculation of approximately 8.4 years of CPU time, including the systems of 1728 molecules discussed in the appendix. For the larger systems, we utilize a parallelized version of our simulation code on eight processors to improve performance.
## IV Static Structure Factor
We first summarize the structural properties of our simulations in order to better understand the relationship between the changes in structure with the changes in dynamic behavior, which we will detail in Sec. VIII. Other studies have considered the structural and thermodynamic properties of SPC/E in a large region of the $`(P.T)`$ plane , so the present discussion is brief.
The MCT theory requires as input the static density-density correlation functions. In the case of water, the structure of the system is very sensitive to the value of the external control parameter $`(P,T)`$. Hence, for all state points simulated, we calculate the oxygen-oxygen partial structure factor
$$S(q)\frac{1}{N}\left|\underset{j=1}{\overset{N}{}}e^{i𝐪𝐫_j}\right|^2.$$
(10)
Several studies have carefully calculated the structure of simulated water, and found surprisingly good agreement with experiments . At $`T=210`$ K, we show the structural changes from low to high density \[Fig. 3(a)\]. The structure at low density/pressure is similar to that observed for low-density amorphous (LDA) solid water, consisting of an open tetrahedral network. At high density/pressure, water is very similar to high density amorphous (HDA) solid water, where core-repulsion dominates, similar to simple liquids under pressure.
We show the evolution of $`S(q)`$ as a function of $`T`$ along the $`\rho =1.0`$ g/cm<sup>3</sup> isochore in Fig. 3(b). We note that in the temperature range from 190 K to 300 K, where the dynamics show the most dramatic change in behavior, $`S(q)`$ shows only small changes in the first two peaks. Also, the location of the first maximum $`q_0`$ in $`S(q)`$, the wave vector at which $`F(q,t)`$ typically shows the slowest relaxation, does not appear to change significantly. All other densities and temperatures show a relative smooth interpolation of Figs. 3(a) and (b).
## V Mean-Squared Displacement and Diffusion
The mean-squared displacement $`r^2(t)|𝐫(t)𝐫(0)|^2`$ is shown in Fig. 4. All the curves show $`t^2`$ dependence at small time, as expected in the “ballistic” regime. For low T (e.g. $`T=210`$ K \[Fig. 4(a)\]), $`r^2(t)`$ shows relatively flat behavior over 3-4 decades in time. This is the “cage” region, in which a molecule is trapped by its neighbors and cannot diffuse, and is only vibrating within its “cage.” At low $`P`$, the cage consists of hydrogen-bonded neighbors in a tetrahedral configuration. This cage is relatively strong, compared to simple liquids, because of the H bonds. The size of the cage may be estimated by the value $`r^2(t)`$ at the plateau, as shown in the inset of \[Fig. 4(a)\]. Surprisingly, the size of the cage is not monotonic with density, and has a maximum at $`\rho 1.1`$ g/cm<sup>3</sup>. We shall see that this corresponds roughly to the $`\rho `$ at which $`D`$ also has a maximum. We observe a small bump in $`r^2(t)`$ at $`t0.35`$ ps, as observed in Ref. . A system size study indicates that this may be attributed to finite size effects .
For long times, all the curves show linear $`t`$ dependence, indicating that our simulations are in the diffusive regime. We extract the diffusion constant $`D`$ using the asymptotic relation $`r^2(t)=6Dt`$. We plot the density dependence of $`D`$ in Fig. 5 and find that the SPC/E potential, like water, shows an anomalous increase in $`D`$ on increasing density. We also point out the feature that $`D`$ shows a slight increase at very low density; namely, at $`\rho =0.90`$ g/cm<sup>3</sup> and $`T=210`$ K. This can be attributed to the fact that the liquid is extremely stretched at this density, causing an increase in the defects of the bond network, and thus increased diffusivity. We show $`D`$ as a function of pressure along several isotherms to compare with experimental measurements \[Fig. 6. The anomalous increase in $`D`$ is qualitatively reproduced by our calculations for the SPC/E model, but the quantitative increase of $`D`$ is significantly larger than that
observed experimentally. This discrepancy may arise from the fact that the SPC/E potential is under-structured relative to water , so applying pressure allows for more bond breaking and thus greater diffusivity than observed experimentally. We also find that the pressure where $`D`$ begins to decrease with pressure — normal behavior for a liquid — is larger than that observed experimentally . This comparison of $`D`$ with experiment leads us to expect that while the qualitative dynamic features we observe in the SPC/E potential may aid in the understanding of the dynamics of water under pressure, they will likely not be quantitatively accurate.
We estimate $`D`$ along the isobars $`P=80`$ MPa, 0 MPa, 100 MPa, 200 MPa, 300 MPa, and 400 MPa from the isochoric data. We confirm that along the -80 MPa isobar, our estimates agree with the $`80`$ MPa calculations of Ref. , which employs the same truncation of the potential used here (see Sec. III). Along the 0 MPa
isobar, our estimates of $`D`$ are smaller than those calculated for SPC/E in Ref. , perhaps because Ref. chooses a different truncation of the electrostatic terms—highlighting the extreme sensitivity of the dynamics to changes in the potential.
We fit $`D`$ by the power law of Eq. (1) along both isochores and the estimated isobars for $`T300`$ K \[Figs. 7 and 8\]. The values of the two fit parameters $`T_c`$ and $`\gamma `$ are given in Table II . We also include the data from Ref. along the $`\rho =1.0`$ g/cm<sup>3</sup> isochore and from Ref. along the $`P=80`$ MPa isobar to improve the quality of the fits. At $`\rho =1.40`$ g/cm<sup>3</sup>, we exclude $`T=210`$ K when fitting $`D`$ and obtain $`T_c=209.3`$, since we expect the power law of Eq. (1) to fail for $`TT_c+5`$ K, because activated processes – such as “hopping” not accounted for in the idealized MCT – become significant and aid diffusion. To demonstrate the presence of hopping at $`\rho =1.40`$ and $`T=210`$ K, we plot the “self” part of the van Hove correlation function $`G_s(r,t)`$, which measures the distribution of particle displacements $`r`$ at time $`t`$, for several densities at $`T=210`$ K \[Fig. 9\]. For these densities where a power law adequately describes $`D`$, there is a single peak. At $`\rho =1.40`$ g/cm<sup>3</sup>, we see a “shoulder” in $`G_s(r,t)`$ at $`r0.2`$ nm in addition to a well-defined peak at $`r0.05`$ nm, indicating that particle hopping is significant.
Along the $`\rho =1.00`$ g/cm<sup>3</sup> we have simulated to significantly lower $`T`$, allowing us to study the temperature dependence of $`D`$ for $`TT_c`$. Fig. 10 shows that the lowest temperatures are consistent with the Arrhenius form
$$D=D_{\mathrm{}}\mathrm{exp}(E/k_BT).$$
(11)
Arrhenius temperature dependence of $`D`$ is not surprising in this region, since for $`T<T_c`$ it is expected that the energy barriers the system must overcome to
rearrange exceed the thermal energy . Hence the motion of the system is dominated by activated jumps over the energy barriers, as described by Goldstein . We obtain an activation energy of $`E65`$ kJ/mol and extrapolate a glass transition temperature $`T_g125`$ K , surprisingly close to the experimental value of 136 K. The extrapolated value of $`T_g`$ is similar to that estimated in Ref. which studied hydrogen bond dynamics. Moreover, our results are consistent with a crossover from “fragile” behavior (the behavior described by MCT) for $`TT_c`$, to “strong” behavior (Arrhenius behavior with
$`Ek_BT_g/2540`$ kJ/mol for our estimate of $`T_g`$) for $`TT_c`$. The possibility of a “fragile-to-strong” crossover in water has been discussed recently based on experimental findings , but lower temperatures are required to test this possibility in the SPC/E model.
## VI Isochrones of $`D`$ and the Locus of $`T_c(P)`$
To construct isochrones of $`D`$ (lines of constant $`D`$), we first estimate $`T(D)`$ at values of $`D=10^5`$cm<sup>2</sup>/s, $`10^{5.5}`$ cm<sup>2</sup>/s, $`10^6`$cm<sup>2</sup>/s, and $`10^7`$cm<sup>2</sup>/s, using the fits of Figs. 7 and 8. Along the isobaric paths, we know already $`P`$ for these points, and along isochores we may estimate the value of $`P`$ using the results presented in Table I. We plot the isochrones in Fig. 11.
We also show the the loci of $`T_c(P)`$ in Fig. 11(a), obtained from the fits in the previous section. We know $`P`$ at $`T_c`$ along the isobaric paths, and we estimate the $`P`$ at $`T_c`$ along isochores by extrapolating $`P`$ in Table I to $`T_c`$.
Using the experimental diffusion data of Ref. , we also construct the behavior of the experimental isochrones following the same technique \[Fig. 11(b)\]. The shape of the locus of $`T_c(P)`$ compares well with that observed experimentally , and changes slope at roughly the same pressure \[Fig. 11\]. Therefore, an explanation of the SPC/E dynamics using the MCT would support using the MCT framework as an interpretation of the experimentally found locus of $`T_c(P)`$. We find, however, that $`\gamma `$ decreases with $`P`$ for the SPC/E model, while $`\gamma `$ increases with $`P`$ \[Fig. 12\]. This disagreement underscores the need to improve the dynamic properties of water models, most of which already provide an adequate account of static properties .
## VII Intermediate Scattering Function
We plot the intermediate scattering function $`F(q_0,t)`$ in Fig. 13(a) for all $`T`$ along the $`\rho =1.00`$ g/cm<sup>3</sup> isochore, where $`q_0=18.55`$ nm<sup>-1</sup>, the approximate value of the first peak of $`S(q)`$ where the relaxation of $`F(q,t)`$ is slowest. We define the relaxation time $`\tau `$ by $`F(t=\tau )=e^1`$. We show $`\tau `$ along isotherms in Fig. 5(b), from which it is obvious that $`\tau `$ has very similar behavior to $`D^1`$. Indeed, MCT predicts that the product $`D\tau `$ is constant along isochores, which we test in Fig. 5(c). We find that $`D\tau `$ increases slightly on cooling, but remains relatively constant along each isochore. The weak residual
$`T`$-dependence in $`D\tau `$ should be subjected to a deeper scrutiny to find out if it is related to a $`q`$vector dependent correction to scaling (since $`D`$ is a $`q=0`$ quantity) or to the progressive breakdown of the validity of the ideal MCT on approaching $`T_c`$.
The study of the time dependence of $`F(q_0,t)`$ allows us to test the predicted relation between the exponents $`b`$ and $`\gamma `$ (see Eqs. (6) and (7)). Since the value of $`b`$ is completely determined by the value of $`\gamma `$ , calculation of these exponents for SPC/E determines if MCT is consistent with our results. The range of validity of the van Schweidler power law \[Eq. (4)\] is strongly $`q`$-dependent , making unambiguous calculation of $`b`$ difficult.
Fortunately, according to MCT , at large $`q`$-vectors,
the stretching exponent $`\beta (q)`$, which characterizes the the long-time behavior of $`F(q,t)`$ (see Eq. (8)), is controlled by the same exponent $`b`$ at large $`q`$. Fits of $`F(q_0,t)`$ according to Eq. (8) are shown for many $`q`$ values at $`T=210`$ K and $`\rho =1.00`$ g/cm<sup>3</sup>. The same fit quality is observed for all other low $`T`$ state points. The $`q`$-dependence of $`\beta (q)`$ for $`\rho 1.30`$ g/cm<sup>3</sup> and $`T=210`$ K is shown in Fig. 14 for $`F(q,t)`$. In addition, we show the expected value of $`b`$ according to MCT, using the values of $`\gamma `$ extrapolated from Fig. 12. The large-$`q`$ limit of $`\beta `$ appears to approach the value predicted by MCT. Hence we conclude that the dynamic behavior of the SPC/E potential in the pressure range we study is consistent with slowing down as described by MCT \[Fig. 2\]. We also checked that the values of $`b`$ calculated from Eq. (9) are consistent with the von Schweidler power law Eq. (4), but that corrections to scaling in $`t^{2b}`$ are relevant at several $`q`$ vectors, as discussed in Ref. .
## VIII Relationship of Structure to Dynamics
The results shown in Fig. 14, and the observed power-law dependence of diffusivity, suggest that MCT is able to predict the dynamical behavior of SPC/E water in a wide range of $`P`$ and $`T`$. As discussed above, the structure of the liquid changes significantly under increased pressure. To highlight the effect of structural changes on dynamic properties, we consider an approximately isochronic path – along which $`D`$ remains nearly constant – such that the changes in dynamic properties we observe on increasing $`P`$ are confined to their $`q`$-vector dependence. We select 5 state points with $`D=(0.30\pm 0.09)\times 10^{}6`$ cm<sup>2</sup>/s: (i) $`T=220`$ K, $`\rho =1.00`$ g/cm<sup>3</sup>, (ii) $`T=210`$ K, $`\rho =1.05`$ g/cm<sup>3</sup>, (iii) $`T=210`$ K, $`\rho =1.10`$ g/cm<sup>3</sup>, (i) $`T=210`$ K, $`\rho =1.20`$ g/cm<sup>3</sup>, (iv) $`T=220`$ K, $`\rho =1.30`$ g/cm<sup>3</sup>, and (v) $`T=240`$ K, $`\rho =1.40`$ g/cm<sup>3</sup>. We show in Fig. 15 the $`q`$-dependence of the $`\alpha `$-relaxation time $`\tau (q)`$ extracted from the fit of $`F(q,t)`$ to the stretched exponential of Eq. (8). For all state points, the $`q`$-dependence of $`\tau `$ follows the $`q`$-dependence of $`S(q)`$, as commonly observed in supercooled liquids and in solutions of the full $`q`$-vector dependent mode coupling equations. We also note that $`\tau (q)`$ is well described by the relation
$$\tau (q)S(q)/q^2$$
(12)
(the de Gennes narrowing relation), as shown in the same figure. The MCT prediction for the $`q`$-dependence of $`\tau `$ is often very close to the relation (12).
## IX Discussion
We have presented extensive simulations that provide evidence for interpreting the dynamics of the SPC/E potential in the framework of MCT. Our calculations also provide a necessary test of the relation predicted between the diffusivity exponent $`\gamma `$ and the von Schweidler exponent $`b`$ for a wide range of values $`\gamma `$ and $`b`$. Our results support interpretation of the experimental locus $`T_c(P)`$ as the locus of MCT transitions.
We found that on increasing pressure, the values of the exponents become closer to those for hard-sphere ($`\gamma =2.58`$ and $`b=0.545`$) and Lennard-Jones ($`\gamma =2.37`$ and $`b=0.617`$) systems , thereby confirming that the hydrogen-bond network is destroyed under pressure and that the water dynamics become closer to that of normal liquids, where core repulsion dominates. A significant result of our analysis is the demonstration that MCT is able to rationalize the dynamic behavior of the SPC/E model of water at all pressures. In doing so, MCT encompasses both the behavior at low pressures, where the mobility is essentially controlled by the presence of strong energetic cages of hydrogen bonds, and at high pressures, where the dynamics are dominated by excluded volume effects. We also showed how these structural changes are reflected in the $`q`$-dependence of $`F(q,t)`$.
Our results underscore the need to improve the dynamic properties of potentials for realistic simulations of
water and other other materials. Of the many potentials available for studying water, only the SPC/E potential is known to display the power-law dependence of dynamic properties, but even SPC/E fails to reproduce the power law quantitatively. A recent study of the ST2 potential found that the $`T`$-dependence of $`D`$ is consistent with an Arrhenius $`T`$-dependence for $`T300`$ K, crossing over to a another region of Arrhenius behavior for $`T275`$ K , in contrast to the non-Arrhenius behavior observed in real water and to our interpretation based on MCT for $`TT_c`$. The presence of a low-$`T`$ Arrhenius regime in the ST2 potential might be due to activated processes, that are expected to dominate the dynamics of fragile liquids below $`T_c`$, as we observed for the SPC/E potential. Hence the ST2 potential may provide an excellent opportunity to study these activated processes on a smaller time scale than is typically observed for most fragile liquids.
Finally, we stress that a full comparison between theory and simulation data requires a complete solution of the recently proposed molecular-MCT (the extension of MCT to molecules of arbitrary shape) . A detailed solution of the complicated molecular-MCT equations in such large region of $`T`$ and $`P`$ values would requires computational effort beyond the present possibilities, but a detailed comparison between molecular-MCT and MD data for one selected isobar is underway .
## X Acknowledgments
We thank C.A. Angell, A. Geiger, E. La Nave, A. Rinaldi, S. Sastry, A. Scala and R.J. Speedy for enlightening discussions and comments on the manuscript. We especially thank S. Harrington for his contributions to the early stages of this work. We thank the Boston University Center for Computational Science for access to the 192 processor SGI/Cray Origin supercomputer. F.S. is supported in part by MURST (PRIN 98). The Center for Polymer Studies is supported by the NSF Grant No. CH9728854.
## A Finite-Size Effects
Some recent work indicates that significant finite-size effects can affect results at temperatures close to the MCT $`T_c`$. Most of our systems are farther than 5% from $`T_c`$ (i.e., $`T/T_c10.05`$) so that the relatively small size of our system should not affect our results. We simulated two independent systems of 1728 molecules at both $`T=200`$ K and $`190K`$ and $`\rho =1.00`$ g/cm<sup>3</sup> to check if significant finite size effects appear at low $`T`$. The results are shown in Table. III. We observe no significant deviations from the system of 216 molecules at $`T=200`$ K. Hence we believe that no strong finite-size effects are present in the 216 molecule system for $`T200`$ K. However at $`T=190`$ K, the potential energy of the larger system appears to be significantly smaller than that in the smaller system. We lack adequate computer resources to make a reliable estimate the diffusivity in the larger system, but simulations are continuing in order to check the possible finite-size effects at this temperature.
|
no-problem/9908/hep-ph9908262.html
|
ar5iv
|
text
|
# References
TIFR/TH/99-37
July, 1999
A Minimal See-Saw Model for Hierarchical Neutrino Masses with Large Mixing<sup>1</sup><sup>1</sup>1To appear in the proceedings of the 6th San Miniato Topical Seminar on ‘Neutrino and Astroparticle Physics’, Nuclear Physics B (Proc. Suppl.)
D.P. Roy
Tata Institute of Fundamental Research, Mumbai 400 005, India
## Abstract
The atmospheric and solar neutrino oscillation data suggest hierarchical neutrino masses with at least one large mixing. The simplest see-saw models for reconciling the two features are $`U(1)`$ extensions of the SM with flavour dependent gauge charges. I discuss a minimal model of this type containing two heavy right-handed neutrinos, which have normal Dirac couplings to $`\nu _\mu `$ and $`\nu _\tau `$ but suppressed ones to $`\nu _e`$. It can naturally account for the large (small) mixing solutions to the atmospheric (solar) neutrino oscillation data.
The recent Superkamiokande data has provided convincing evidence for atmospheric neutrino oscillation and confirmed earlier results of solar neutrino oscillation . The atmospheric neutrino data seems to imply a large mixing between $`\nu _\mu `$ and $`\nu _\tau `$, $`\mathrm{sin}^22\theta _{\mu \tau }>0.86`$, along with $`\mathrm{\Delta }M^2=(1.56)\times 10^3`$ eV<sup>2</sup> at 90% CL. They correspond to
$$\theta _{\mu \tau }=45\pm 11^{}$$
(1)
and
$$\mathrm{\Delta }M0.06eV,$$
(2)
the latter representing the central value of $`\mathrm{\Delta }M`$ for hierarchical masses and an upper limit on this quantity for degenerate ones. By far the simplest explanation of the solar neutrino oscillation data is provided by the small mixing angle MSW solution although one can get equally good descriptions in terms of the large mixing angle MSW or vacuum oscillation solutions. The SMA solution corresponds to a small mixing of $`\nu _e`$ with one of the above states, $`\mathrm{sin}^22\theta _e=10^310^2`$, along with a small $`\mathrm{\Delta }m^2=(0.51)\times 10^5`$ eV<sup>2</sup>. They correspond to
$$\mathrm{sin}\theta _e=(15)\times 10^2$$
(3)
and
$$\mathrm{\Delta }m0.003eV.$$
(4)
By itself the atmospheric neutrino oscillation result of Eqs. (1,2) could be naturally explained in terms of a nearly degenerate pair of $`\nu _\mu `$ and $`\nu _\tau `$. Indeed a pseudo-Dirac mass matrix for this pair would lead to degenerate masses and maximal mixing on diagonalisation, i.e.
$$\left(\begin{array}{cc}0& M\\ M& 0\end{array}\right)\left(\begin{array}{cc}M& 0\\ 0& M\end{array}\right),\theta =45^{}.$$
(5)
But explaining the solar neutrino oscillation result of Eqs. (3,4) would then imply an even finer level of degeneracy between $`\nu _e`$ and one of this pair, which is totally ad-hoc. Therefore it is generally considered more natural to interpret them as hierarchical states, i.e.,
$`m_1\mathrm{\Delta }M`$ $``$ $`0.06eV,`$
$`m_2\mathrm{\Delta }m`$ $``$ $`0.003eV,`$
$`m_3m_2`$ $``$ $`0,`$ (6)
where the first two states are large admixtures of $`\nu _\mu `$ and $`\nu _\tau `$ and the third one is dominantly $`\nu _e`$. Indeed much of the recent literature on neutrino physics is focussed on theoretical models, mainly in the see-saw frame work, which can naturally reconcile such hierarchical masses with large mixing . Note that the mass of the 3rd state can be exactly zero as far as the atmospheric and solar neutrino oscillation data are concerned. Thus a minimal see-saw model for explaining these oscillations requires two right-handed neutrinos with normal Dirac couplings to $`\nu _\mu `$ and $`\nu _\tau `$, but suppressed ones to $`\nu _e`$.
It may be noted here that the standard see-saw model represents a $`U(1)`$ extension of the standard model (SM) gauge group into
$$SU(3)_C\times SU(2)\times U(1)_Y\times U(1)_Y^{}$$
(7)
with the gauge charge
$$Y^{}=BL=B(L_e+L_\mu +L_\tau ).$$
(8)
Then the requirement of anomaly cancellation implies the existence of three right-handed singlet neutrinos ($`N_i`$) with $`Y^{}=1`$ to match the three left-handed neutrinos ($`\nu _{e,\mu ,\tau }`$) carrying this gauge charge. Cancellation of the axial parts of the $`Y^{}`$ current between the left and right handed fermions ensures purely vector coupling for $`Y^{}`$, which in turn ensures that the model is anomaly free . The flavour independence of $`Y^{}`$ implies however that the singlet neutrinos have normal Dirac couplings to all the left-handed doublets $`\nu _{e,\mu ,\tau }`$ along with the SM Higgs doublet $`\varphi `$ instead of preferential couplings to $`\nu _{\mu ,\tau }`$ as suggested by data. In order to accomplish the latter one has to invoke a horizontal symmetry with flavour dependent charges . In other words one first takes a flavour blind step beyond the SM and then applies correctives via additional symmetry groups with flavour dependent charges. Let us consider instead a one-step process, where the desired flavour depence is incorporated into the gauge charge $`Y^{}`$ of the $`U(1)`$ extension of SM (Eq. 7). While such flavour dependent $`U(1)`$ extensions of the SM gauge group are hard to embed in the familiar GUTs they can arise naturally from string theories .
We have studied two such $`U(1)`$ extensions of the SM , corresponding to the gauge charges
$$Y^{}=B3L_e$$
(9)
and
$$Y^{}=B\frac{3}{2}(L_\mu +L_\tau ),$$
(10)
in the context of the atmospheric and solar neutrino oscillations. I shall concentrate on the simpler of the two models , corresponding to the gauge charge (10). Indeed it seems to represent a minimal see-saw model for explaining these neutrino oscillation data. In this case the anomaly cancellation requirement implies the existence of two right-handed singlet neutrinos $`(N_{1,2})`$ with $`Y^{}=\frac{3}{2}`$ to match the two left-handed neutrinos $`(\nu _{\mu ,\tau })`$ carrying this gauge charge.
The minimal Higgs sector of this model consists of
$$\left(\begin{array}{c}\varphi ^+\\ \varphi _0\end{array}\right)_{Y^{}=0}\&\chi _{Y^{}=3}^0,$$
(11)
i.e. the SM Higgs doublet along with a singlet carrying non-zero $`Y^{}`$ charge. The $`Y^{}`$ symmetry is spontaneously broken via the vacuum expectation value of $`\chi `$, $`<\chi >`$, at a high mass scale. The coupling of this $`\chi `$ to $`\overline{N}_1^CN_1`$ and $`\overline{N}_2^CN_2`$ gives them large Majorana masses $`<\chi >`$. Moreover the coupling of $`\varphi `$ to $`\overline{\nu }_\mu N_{1,2}`$ and $`\overline{\nu }_\tau N_{1,2}`$ gives them Dirac masses $`<\varphi >`$, while there is no such coupling to $`\nu _e`$. Thus the see-saw mechanism would generate two non-zero mass states, which are large admixtures of $`\nu _\mu `$ and $`\nu _\tau `$, while $`\nu _e`$ remains massless.
One can generate a small mixing of $`\nu _e`$ with the non-zero mass states, as required by the SMA solution (3) to the solar neutrino oscillation, by expanding the Higgs sector. For this purpose we add another doublet and a singlet with
$$\left(\begin{array}{c}\eta ^+\\ \eta ^0\end{array}\right)_{Y^{}=3/2}\&\zeta _{Y^{}=3/2}^0.$$
(12)
The coupling of the doublet $`\eta `$ to $`\overline{\nu }_eN_{1,2}`$ generates Dirac mass terms $`<\eta >`$. The singlet $`\zeta ^0`$ does not couple to fermions; but it is required to avoid an unwanted Goldstone boson. The latter comes about because there are 3 global $`U(1)`$ symmetries, corresponding to rotating the phases of $`\varphi `$, $`\eta `$ and $`\chi ^0`$ independently in the Higgs potential, while only 2 local $`U(1)`$ symmetries are spontaneously broken. The addition of the singlet $`\zeta ^0`$ introduces two more terms in the Higgs potential, $`\eta ^+\varphi \zeta ^0`$ and $`\chi ^0\zeta ^0\zeta ^0`$, so that the phases can no longer be rotated independently. While the $`\zeta ^0`$ is expected to acquire a large vev at the $`U(1)_Y^{}`$ symmetry breaking scale, the doublet $`\eta `$ must have a positive mass squared term in order to avoid $`SU(2)`$ breaking at this scale. Nonetheless it can acquire a small but non-zero vev at the $`SU(2)`$ symmetry breaking scale, which can be estimated from the relevant part of the potential
$$m_\eta ^2\eta ^{}\eta +\lambda (\eta ^{}\eta )(\chi ^{}\chi )+\lambda ^{}(\eta ^{}\eta )(\zeta ^{}\zeta )\mu \eta ^{}\varphi \zeta .$$
(13)
Although we start with a positive $`m_\eta ^2`$ term, after minimization of the potential with respect to $`\eta `$ we see that this field has acquired a small vev,
$$<\eta >=\mu <\varphi ><\zeta >/2M_\eta ^2,$$
(14)
where $`M_\eta ^2=m_\eta ^2+\lambda <\chi >^2+\lambda ^{}<\zeta >^2`$ represents the physical mass of $`\eta `$. The size of the soft term is bounded by the $`Y^{}`$ symmetry breaking scale, i.e. $`\mu <\zeta >`$. Thus with a choice of $`M_\eta 5<\zeta >`$, we get
$$<\eta >/<\varphi >1/50,$$
(15)
which will account for the small mixing angle of $`\nu _e`$ (3).
Let us write down the $`5\times 5`$ neutrino mass-matrix in this model. We shall be working in the basis where the charged lepton mass matrix, arising from their couplings to the SM Higgs boson $`\varphi `$, is diagonal. This defines the flavour basis of the doublet neutrinos. Since the two singlet neutrinos do not couple to the charged leptons, their Majorana mass matrix can be independently diagonalised in this basis. While the overall size of their masses will be at the $`Y^{}`$ symmetry breaking scale, it is reasonable to assume a modest hierarchy between them,
$$M_1/M_21/20,$$
(16)
in analogy with those observed in the quark and the charged lepton sectors. This will account for the desired mass ratio for the doublet neutrinos (6). Thus we have the following $`5\times 5`$ mass matrix $``$ in the basis $`(\nu _e,\nu _\mu ,\nu _\tau ,N_1^C,N_2^C)`$:
$$\left(\begin{array}{ccccc}0& 0& 0& f_e^1<\eta >& f_e^2<\eta >\\ 0& 0& 0& f_\mu ^1<\varphi >& f_\mu ^2<\varphi >\\ 0& 0& 0& f_\tau ^1<\varphi >& f_\tau ^2<\varphi >\\ f_e^1<\eta >& f_\mu ^1<\varphi >& f_\tau ^1<\varphi >& M_1& 0\\ f_e^2<\eta >& f_\mu ^2<\varphi >& f_\tau ^2<\varphi >& 0& M_2\end{array}\right),$$
(17)
where the $`f_{e,\mu \tau }^{1,2}`$ are the Higgs Yukawa couplings. We shall assume these couplings to be of similar order of magnitude, i.e. the elements of a mass-matrix arising from the same Higgs vev are expected to be of similar size. There is of course no conflict between such democratic mass-matrix elements and the hierarchical mass eigen-values assumed above (16). In fact they are closely related - the former implies large cancellation in the determinant as required by the latter.
The resulting $`3\times 3`$ mass-matrix for the doublet neutrinos is given by the see-saw formula in this basis,
$$m_{ij}=\frac{D_{1i}D_{1j}}{M_1}+\frac{D_{2i}D_{2j}}{M_2},$$
(18)
where $`D`$ is the $`2\times 3`$ Dirac mass matrix at the bottom left of (17). One can then calculate the corresponding mass eigen-values $`m_{1,2,3}`$ and mixing-angles by diagonalising this matrix . Alternatively we can read off the approximate magnitudes of these quantities directly from the mass matrix (17), i.e.
$`\mathrm{tan}\theta _{\mu \tau }`$ $``$ $`_{42}/_{43}f_\mu ^1/f_\tau ^11,`$
$`\mathrm{sin}\theta _e`$ $``$ $`_{51}/_{52}<\eta >/<\varphi >{\displaystyle \frac{1}{50}},`$
$`m_2/m_1`$ $``$ $`M_1/M_21/20.`$ (19)
They are clearly in good agreement with the corresponding experimental quantities of Eqs. (1), (3) and (6). Note that in this model the $`\nu _e`$ mixing with the higher mass $`(m_1)`$ eigen-state is also expected to be of similar size as above, i.e.,
$$_{41}/_{42}<\eta >/<\varphi >1/50.$$
(20)
This prediction is well within the present experimental limit on this quantity $`(0.2)`$ from CHOOZ data ; but can be tested by future long base line experiments.
Finally, the scale of the $`Y^{}`$ symmetry breaking can be estimated from the larger Majorana mass $`M_2`$, i.e.
$$M_2f^2<\varphi >^2/m_2f^210^{16}\mathrm{GeV}10^{1216}\mathrm{GeV}.$$
(21)
The lower limit corresponds to $`f10^2`$ as in the case of $`\tau `$ Yukawa coupling, while the upper limit corresponds to $`f1`$ as in the case of top. Thus the observed scale of neutrino masses (6) can be explained if one assumes the $`Y^{}`$ symmetry breaking scale to be in the range of $`10^{12}10^{16}`$ GeV.
One can get a more exact derivation of the masses and mixing angles via the $`3\times 3`$ mass-matrix of the doublet neutrinos (18), i.e.
$$\left(\begin{array}{ccc}c_1^2+c_2^2& c_1a_1+c_2a_2& c_1b_1+c_2b_2\\ c_1a_1+c_2a_2& a_1^2+a_2^2& a_1b_1+a_2b_2\\ c_1b_1+c_2b_2& a_1b_1+a_2b_2& b_1^2+b_2^2\end{array}\right),$$
(22)
where
$$a_{1,2}=\frac{f_\mu ^{1,2}<\varphi >}{\sqrt{M_{1,2}}},b_{1,2}=\frac{f_\tau ^{1,2}<\varphi >}{\sqrt{M_{1,2}}},c_{1,2}=\frac{f_e^{1,2}<\eta >}{\sqrt{M_{1,2}}}.$$
(23)
Note that the assumed hierarchies of (15) and (16) imply
$$a_1,b_1a_2,b_2,c_1c_2.$$
(24)
The determinant of (22) vanishes identically, ensuring that one of the mass eigenvalues is zero. The other two are
$$m_1a_1^2+b_1^2,m_2\frac{(a_1b_2a_2b)^2}{a_1^2+b_1^2}.$$
(25)
The corresponding mixing matrix $`U`$ between the flavour and the mass eigenstates is
$$\left(\begin{array}{c}\nu _e\\ \nu _\mu \\ \nu _\tau \end{array}\right)=\left(\begin{array}{ccc}1& \frac{c_2\sqrt{a_1^2+b_1^2}}{a_1b_2b_1a_2}& \frac{c_1}{\sqrt{a_1^2+b_1^2}}\\ \frac{b_1c_2c_1b_2}{a_1b_2b_1a_2}& \frac{b_1}{\sqrt{a_1^2+b_1^2}}& \frac{a_1}{\sqrt{a_1^2+b_1^2}}\\ \frac{c_1a_2a_1c_2}{a_1b_2b_1a_2}& \frac{a_1}{\sqrt{a_1^2+b_1^2}}& \frac{b_1}{\sqrt{a_1^2+b_1^2}}\end{array}\right)\left(\begin{array}{c}\nu _3\\ \nu _2\\ \nu _1\end{array}\right).$$
(26)
One can easily check that the Eqs. (23-26) lead to the masses and mixing angles of Eqs. (19-21).
|
no-problem/9908/cond-mat9908128.html
|
ar5iv
|
text
|
# Magnetic anisotropy, first-order-like metamagnetic transitions and large negative magnetoresistance in the single crystal of Gd2PdSi3
## Abstract
Electrical resistivity ($`\rho `$), magnetoresistance (MR), magnetization, thermopower and Hall effect measurements on the single crystal Gd<sub>2</sub>PdSi<sub>3</sub>, crystallizing in an AlB<sub>2</sub>-derived hexagonal structure are reported. The well-defined minimum in $`\rho `$ at a temperature above Néel temperature (T<sub>N</sub>= 21 K) and large negative MR below $``$ 3T<sub>N</sub>, reported earlier for the polycrystals, are reproducible even in single crystals. Such features are generally uncharacteristic of Gd alloys. In addition, we also found interesting features in other data, e.g., two-step first-order-like metamagnetic transitions for the magnetic field along direction. The alloy exhibits anisotropy in all these properties, though Gd is a S-state ion.
The observation of large negative magnetoresistance (MR) above respective magnetic transition temperatures in some polycrystalline Gd (also Tb, Dy) alloys is of considerable interest. Among the Gd alloys, we have studied the transport properties of the compound, Gd<sub>2</sub>PdSi<sub>3</sub>, crystallizing in an AlB<sub>2</sub>-type structure, which has been found to show unusual nature. While this compound orders antiferromagnetically at (T<sub>N</sub>=) 21 K, there is unexpectedly a distinct minimum in the temperature dependent electrical resistivity ($`\rho `$) at about 45 K. This minimum disappears by the application of a magnetic field (H), thereby resulting in large MR in the vicinity of T<sub>N</sub> \[Ref. 3\]. These properties are also characteristic of Ce/U-based Kondo lattices, but uncharacteristic of Gd systems, considering that the Gd-4f orbital is so deeply localised that it cannot exhibit the Kondo effect. Though magnetic-polaronic effect (even in metallic environments) has been proposed in references 1-5 as one of possible mechanisms behind this large MR, its origin is not clear yet. It, however, appears that short-range correlation as a magnetic precursor effect may be the primary ingredient for the origin of the resistivity minimum above T<sub>N</sub> and negative MR. The importance of such findings is obvious from similar recent reports from other groups and among these the observation of $`\rho `$ minimum and resultant colossal magnetoresistance (CMR) in a pyrochlore-based oxide, Tl<sub>2</sub>Mn<sub>2</sub>O<sub>7</sub> \[Ref. 7,8\], has attracted recent attention. In view of the importance of the observations on polycystals of this Gd compound, we considered it important to confirm the findings on the single crystals. With this primary motivation, we have investigated $`\rho `$, MR, thermopower (S), Hall-effect and magnetization behavior on single crystals of Gd<sub>2</sub>PdSi<sub>3</sub>, and these results are presented in this article.
Single crystals of Gd<sub>2</sub>PdSi<sub>3</sub> have been prepared by the Czochralsky pulling method using a tetra-arc furnace in an argon atmosphere. The single-crystalline nature has been confirmed using back scattering x-ray technique. The $`\rho `$, MR and Hall effect (employing a magnetic field of 15 kOe) measurements have been performed by a conventional DC four-probe method down to 1.2 K; the MR and Hall effect measurements have also been performed as a function of H at 4.2 K. The magnetic measurements have been carried out with a Quantum Design Superconducting Quantum Interference Device. The thermopower data were taken by the differential method using Au-Fe (0.07%)-chromel thermocouples.
Fig. 1a shows the temperature dependence (1.2-300 K) of $`\rho `$ for the sample with the current j//\[$`10\overline{1}0`$\] and j// in zero field. In Fig. 1b, the low temperature data, normalised to the 300 K value, in the absence of a magnetic field as well as in the presence of 50 kOe (in the longitudinal geometry, H//j) are shown. In zero field, the $`\rho `$(T) gradually decreases with decreasing temperature like in ordinary metals, however, only down to about 45 K below which there is an upturn. There is a kink at about 21 K for both directions, marking the onset of magnetic ordering. The $`\rho `$, however, does not drop sharply at T<sub>N</sub> expected due to the loss of spin-disorder contribution, but exhibits a tendency to flatten or a fall slowly with decreasing temperature. Presumably, the magnetic structure could be a complex one, resulting in the formation of superzone-zone boundary gaps in some portions of the Fermi surface. The application of a magnetic field, say H = 50 kOe, in the geometry discussed above, however depresses the $`\rho `$ minimum restoring metallic behavior in the entire temperature range of investigation. This naturally means that there is a large negative magnetoresistance, MR = $`\mathrm{\Delta }\rho /\rho =[\rho (H)\rho (0)]/\rho (0)`$, at low temperatures, the magnitude of which increases with decreasing temperature. A large negative value close to -30% could be seen for moderate fields (15 kOe) at 4.2 K (Fig. 2a), an indication of giant magnetoresistance. Thus, all these features observed in polycrystals, are reproducible in single crystals as well. It is obvious (Fig. 1a) that the absolute values of $`\rho `$ are relatively higher for j//\[$`10\overline{1}0`$\] than that for j//. It is to be noted that, though these values still fall in the metallic range, the temperature dependence is rather weak, for instance, $`\rho `$(4.2K)/$`\rho `$(300K) is not less than 0.75 (in zero field), in sharp contrast to a value of about 0.2 even for polycrystalline Lu<sub>2</sub>PdSi<sub>3</sub> (Ref. 11). It is not clear whether this fact is associated with some kind of disorder effect on $`\rho `$ in magnetic sample compared to that in nonmagnetic Lu<sub>2</sub>PdSi<sub>3</sub> or with an intrinsic mechanism responsible for the $`\rho `$ minimum.
We have also measured MR as a function of H at 4.2 K with H varying from -15 kOe to 15 kOe, both in the longitudinal and transverse geometries for H// and H//\[$`10\overline{1}0`$\]. For H//, the transverse MR (j//\[$`10\overline{1}0`$\]) is positive with a small magnitude, while the longitudinal MR (j//) is negative (see Fig. 2a). The contribution from the anisotropic MR due to spin-orbit coupling is negligible for the Gd ion. The cyclotron contribution to the resistivity is also small. Possible contribution resulting from the reduction of magnetic scattering due to the metamagnetic transitions should be the same for both geometries, since the field directions are the same. Therefore, the anisotropy in MR reflects the anisotropy of the Fermi surface for the two current directions. We believe that the conductivity parallel to is favored by the disappearance of the magnetic superzone gaps in some portions of the Fermi surface, resulting in a decrease of $`\rho `$ in the longitudinal MR geometry for H//. However, it is interesting to note that, for H//\[$`10\overline{1}0`$\], one sees negative MR for both geometries (j//\[$`10\overline{1}0`$\] and j//). A noteworthy finding is that, for H//, there are sharp changes in MR when measured as a function of H as indicated by arrows in Fig. 2a, but occurring at different fields for the two geometries due to the difference in the demagnetizing fields. There is a small hysteresis at the region around the sharp changes and we see similar behavior even in the magnetization data (see below); such sharp variations are absent for H//\[$`10\overline{1}0`$\]. All these results bring out anisotropic nature of MR.
Fig. 3a shows the temperature dependence of magnetic susceptibility ($`\chi `$), measured in the presence of a field of 1 kOe for both H// and H//\[$`10\overline{1}0`$\]. There is a well-defined peak in $`\chi `$ at 21 K confirming the antiferromagnetic nature of the magnetic transition; below 21 K, however, there is only a small difference in the values for these two geometries. The paramagnetic Curie-temperature turns out to be the same for both geometries, with the same magnitude as that of T<sub>N</sub>, however, with a positive sign suggesting the existence of strong ferromagneticcorrelations. There is no difference between
field-cooled (FC) and zero-field-coooled (ZFC) $`\chi `$ values below 21 K, unlike the situation in polycrystals. This suggests that such difference in FC and ZFC data in polycrystals is not intrinsic to this material. This fact supports our earlier conclusion that this alloy is not a spin-glass, unlike U<sub>2</sub>PdSi<sub>3</sub> (Ref. 12).
The isothermal magnetization (M) behavior at 2 K is shown in Fig. 3b, both for increasing and decreasing fields. For the field along , there are two step-like metamagnetic transitions, one around 3 kOe and the other around 9 kOe. Apparently, there is a small hysteresis around these transitions, indicating first-order nature of the transitions. The inset of figure 3b shows the metamagnetic transition fields H<sub>m1</sub> and H<sub>m2</sub> (the magnetic fields corresponding to the highest dM/dH at the low-field and high-field transitions, respectively) versus temperature; both H<sub>m1</sub> and H<sub>m2</sub> decrease with increasing temperature. M vs H for H//\[$`10\overline{1}0`$\] also shows a faint meta-magnetic anomaly (Fig. 3c), however with M varying relatively smoothly with H, unlike the situation for H//; the inset shows the characteristic magnetic field, H<sub>m</sub> (estimated in the same way as H<sub>m1</sub> and H<sub>m2</sub>). The results establish the existence of anisotropy in the isothermal magnetization. The observation of metamagnetic transitions are consistent with the anomalies in MR, discussed above.
Fig. 4 shows the temperature dependence of thermopower. The absolute value is large at 300 K as in the case of Lu<sub>2</sub>PdSi<sub>3</sub> (Ref. 3). Therefore, the large S might arise from 4d band of Pd as in the case of 3d band of Co in YCo<sub>2</sub>. There is no anomaly, however, at T<sub>N</sub>. S decreases with decreasing temperature and the features are qualitatively the same as those observed in the non-magnetic Lu<sub>2</sub>PdSi<sub>3</sub>. Though the overall S behavior mimics the one in polycrystals, there is a distinct anisotropy in the values when measured along different directions, that is, the value of S depends on whether the temperature gradient, $`\mathrm{\Delta }`$T, is parallel to \[$`10\overline{1}0`$\] or . The absence of any peak-like behavior expected for Kondo system indicates the absence of Kondo effect in Gd<sub>2</sub>PdSi<sub>3</sub>.
The temperature dependence of Hall coefficient (R<sub>H</sub>), shown in Fig. 5a, also reflects anisotropic nature of this material. The R<sub>H</sub> shows large temperature dependence, in contrast to the temperature independent behavior in Lu<sub>2</sub>PdSi<sub>3</sub> (Ref. 3), with a negative peak for both geometries, H// and H//\[$`10\overline{1}0`$\], in the vicinity of T<sub>N</sub>, however at slightly different temperatures (the reason for which is not clear). Clearly there is a dominant 4f contribution in the Gd case. The Hall effect of magnetic metals like those of Gd is generally a sum of two terms \- an ordinary Hall effect (R<sub>0</sub>) due to Lorentz force and an anomalous part arising from magnetic scatterring (skew scatterring). Thus, in the paramagnetic state, R<sub>H</sub>= R<sub>0</sub> \+ A$`\rho `$$`\chi `$, where A is a constant. Using this relation, R<sub>0</sub> is estimated by plotting R<sub>H</sub> versus $`\rho `$$`\chi `$ (Fig. 5b). From Fig. 5b, it is obvious that the plot is linear for H//\[$`10\overline{1}0`$\] in the paramagnetic state with a value of R<sub>0</sub>= 0.92$`\times `$10<sup>-10</sup> m<sup>3</sup>/coul. However, for H//, there is a distinct change in the magnitude as well as in the sign around 110 K as if there is a change in the sign of the carrier; (1.6$`\times `$10<sup>-10</sup> m<sup>3</sup>/coul and -1.3$`\times `$10<sup>-10</sup> m<sup>3</sup>/coul for T$`>`$ 110 K and T$`<`$110 K, respectively). Below 17 K, the data however deviate from the high temperature linear variation as the state is no longer paramagnetic.
Fig. 6a shows the field dependence of Hall resistivity ($`\rho `$<sub>H</sub>) for H//. $`\rho `$<sub>H</sub> shows distinct anomaly across the two metamagnetic transition fields and traces the hysteresis. Again considering the concept of anomalous Hall effect discussed in the above paragraph, the field dependence of the corresponding $`\rho `$M is ploted in Fig. 6b. Theoretically, when anomalous Hall effect is dominant, $`\rho `$<sub>H</sub> should vary linearly with $`\rho `$M. In other words, $`\rho `$<sub>H</sub> and $`\rho `$M should vary in the same way with the corresponding applied fields. But in the present case, the field dependence of $`\rho `$<sub>H</sub> completely differ from that of $`\rho `$M, particularly around H<sub>m2</sub>. This fact strongly indicates the modification of the Fermi surface across the metamagnetic anomaly.
Summarising, we have explored anisotropy in the transport and magnetic properties on the single crystal Gd<sub>2</sub>PdSi<sub>3</sub>. Possibly the anysotropic exchange interaction due to crystalline anisotropy and the anisotropy in the Fermi surface are responsible for the observed anysotropy. Interestingly, there are magnetic field induced first-order-like magnetic transitions in the magnetically ordered state, resulting in large MR and consequent Fermi surface modification. Recently, first-order transition has been reported in another Gd-alloy, Gd<sub>5</sub>(Si,Ge)<sub>4</sub> (Ref. 15), which has been found to be a simultaneous crystallograpic and magnetic transition, and in view of this it is of interest to explore whether there is any structural transition with the application of H in our case as well. In short, the single crystal of Gd<sub>2</sub>PdSi<sub>3</sub> exhibits interesting features. Above all, there is a well-defined $`\rho `$ minimum above T<sub>N</sub>, the origin of which is still not completely clear; the negative MR persists till about 3T<sub>N</sub> even in single crystals with the magnitude gradually increasing with decreasing temperature towards T<sub>N</sub>. The results overall establish that this compound is a novel magnetic material.
This work has been partially supported by a grant-in-Aid for Scientific Research from the Minstry of Education, Science and Culture of Japan
|
no-problem/9908/hep-ph9908368.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Since the start of the HERA machine, the A RIADNE program has generally been regarded the event generator giving the best overall description of the measured DIS hadronic final states. For most observables it gives a better reproduction of experimental results than more conventional generators such as HERWIG and L EPTO which implements initial- and final-state parton showers based on DGLAP evolution . This has in particular been true in the region of small $`x`$ and $`Q^2`$, and it has often been taken as an indication that DGLAP evolution is inappropriate for this region. The fact that A RIADNE has the feature of non-ordering of transverse momenta in common with the BFKL evolution , has also been taken as an indication that the latter is a more correct description of small-$`x`$ evolution
As the experimental statistics increased also for events at high $`Q^2`$, where we assume that our description of the underlying physics is on more solid theoretical grounds, it became apparent that the A RIADNE program was unable to describe this data . Even in the so-called Current Breit Hemisphere (CBH), where the events are expected to look much like half an event in an e<sup>+</sup>e<sup>-</sup> collision at the same $`Q^2`$ , the A RIADNE program gave far too little transverse momentum, despite the fact that the same program gives a very good description of e<sup>+</sup>e<sup>-</sup> data e.g. from LEP.
In this paper I will comment on the reason of this poor description of data at high $`Q^2`$, and present a modification of the underlying Colour-Dipole Model (CDM) for DIS . The resulting improvement in the reproduction of measured data is presented elsewhere in these proceedings.
## 2 The problem
In the CDM, it is assumed that all gluon radiation in a DIS event can be described in terms of dipole radiation between the struck quark and the proton remnant, much in the same way as the radiation in a hadronic e<sup>+</sup>e<sup>-</sup> collision can be described as dipole radiation from the initial quark and anti-quark. The main difference is that, while in e<sup>+</sup>e<sup>-</sup> the initial $`q\overline{q}`$-pair is essentially point-like, the proton remnant in a DIS event is an extended object.
Looking at the phase space available for gluon radiation in a DIS event (conveniently described as an approximately triangular area in the plane of rapidity and logarithm of the transverse momentum of the emitted gluon in figure 1), it is ultimately limited by the momenta (given here in light-cone components) of the incoming virtual photon $`(Q_+,xP_{},\stackrel{}{0})`$ and proton $`(0,P_{},\stackrel{}{0})`$. In the emission of a gluon with momentum $`(q_+,q_{},\stackrel{}{q}_{})`$, the phase space restriction is given by $`q_+=q_{}e^y<Q_+`$ and $`q_{}=q_{}e^y<P_{}`$. But since the proton remnant is an extended object, say with transverse extension $`1/\mu 1`$ fm, it is reasonable to assume that the gluon only can access a fraction of the negative light-cone momentum of the remnant. Just as radiation of short wavelengths from an extended antenna is suppressed, we get an extra phase space restriction from the condition that the gluon with a transverse wavelength $`1/q_{}`$ can only resolve a fraction $`(\mu /q_{})^\alpha `$ of the remnants momentum (where $`\alpha `$ is the dimensionality of the remnant)
$$q_{}e^y<(\mu /q_{})^\alpha P_{},$$
(1)
corresponding to the dotted line in figure 1.
For small values of $`Q^2`$ the struck quark can no longer be considered point like, and one can argue that emissions of gluons with $`q_{}>Q`$ should also be suppressed, since the gluon only can resolve a fraction $`(Q/q_{})^\alpha ^{}`$ of the positive light-cone component of the virtual photon. This gives an extra phase-space restriction corresponding to the dashed line in figure 1. In this way the concept of a resolved virtual photon is present in A RIADNE .
At high $`Q^2`$, we expect the radiation in the CBH, which corresponds to the shaded area in figure 2, to look very much like half an e<sup>+</sup>e<sup>-</sup> event. But the phase space restriction due to the extension of the proton remnant actually cuts away part of this region. This is most likely the reason why A RIADNE has given too little radiation in high-$`Q^2`$ events.
## 3 The solution
To solve this deficiency in the model we note that in the quark-parton model we have a collision between a virtual photon and an incoming quark with momentum $`(0,xP_{},\stackrel{}{0})`$. So in some sense a fraction x of the incoming protons momentum has already been localized, and this fraction should be easily available for gluon emission. We can then rewrite the condition in eq. (1)
$$q_{}e^y<\mathrm{max}(x,(\mu /q_{})^\alpha )P_{},$$
(2)
and the restriction on the phase space would instead correspond to the dashed line in figure 2.
It could, of course, be argued that the whole of the negative light-cone momentum $`xP_{}`$ of the incoming quark is needed to put the quark on-shell, due to the virtuality of the incoming photon with a negative light-cone momentum of $`xP_{}`$. And that to radiate an extra gluon, more negative light-cone momenta has to be accessed from the proton remnant. But we should remember that the picture of the proton remnant, where the momentum is spread out evenly over its transverse extension, certainly is an oversimplification. It is not unnatural to assume that in the vicinity where the struck quark was localized, the momentum is more concentrated. In any case it is clear that the restriction in eq. (1) is too hard, and seems like eq. (2) is a reasonable modification. Indeed, as reported in , it seems that this modification is approximately what is needed to make CDM reproduce measured data at high $`Q^2`$.
## 4 Conclusions
The modification of the colour dipole model for DIS presented here does fix the most serious problems with the reproduction of data at high $`Q^2`$ for A RIADNE . All problems are, however, not completely solved. Elsewhere in these proceedings , it has been proven to be difficult to find a parameter set which at the same time can describe general event shapes and jet observables. Jet shapes seem to be particularly difficult to describe with A RIADNE . Whether this can be solved with small adjustments or if it is due to an inherent flaw in the model remains to be seen.
|
no-problem/9908/astro-ph9908227.html
|
ar5iv
|
text
|
# Contact Discontinuities in Models of Contact Binaries Undergoing Thermal Relaxation Oscillations
## 1. Introduction
W Ursae Majoris (W UMa) binary stars were first thought to be a particular binary population due to their abnormal mass-radius relationship, namely, the so-called Kuiper’s paradox, $`R_2/R_1=\left(M_2/M_1\right)^{0.46}`$ (Kuiper 1941). These particular binaries appear to consist of two main-sequence stars that possess photospheres exhibiting the almost same effective temperatures for the two components despite the fact that typical mass ratio in a system is 0.5. It was originally proposed that a common convective envelope may be formed due to dynamic equilibrium (Osaki 1965), and mass and energy transfers would take place in CCE in order to interpret the Kuiper’s paradox (Lucy 1968) although the specific mechanism of energy transfer for the circulation has not been fully understood (Robertson 1980, Sinjab, Robertson & Smith 1990, Tassoul 1992). It is now firmly believed that W UMa stars are contact binaries in which both components are full of their Roche lobe, showing strong interactions (Mochnacki 1981). Lucy’s iso-entropy model (1968) as a zero-order model with thermal equilibrium, however, cannot explain the color – period diagram by Eggen (1967) which leads to the establishments of two parallel first-order theories of thermal relaxation oscillation (TRO) and contact discontinuity (DSC). TRO model was advanced by Lucy (1976), Flannery (1976), and Robertson & Eggleton (1977), who suggest that contact system can not reach thermal equilibrium at dynamical equilibrium configuration and may thus undergo thermal relaxation oscillation. DSC theory was proposed by Shu and his collaborators \[Shu, Lubow & Anderson 1976, 1979, Lubow & Shu 1977, also see Biermann & Thomas (1972) and Vilhu (1973) for some earlier elements of DSC model\] who hypothesize contact binary can attain thermal and dynamical equilibrium but there is a temperature inversion layer in the secondary. With great attempts the so-called Kuiper’s paradox and period - colour diagram may be resolved by the two different hypotheses independently. However, both theories have some difficulties to explain observations, such as the so-called W-phenomenon, i.e., the less star is hotter than the massive component (Binnendijik 1970) (see a concise summary of observations and theory by Smith 1984). Especially there are great debases between the two in nature in their simplest version (Kähler 1989). Observational studies continue and the theoretical controversy still remains (Ruciński 1997). These imply that the first-order theories of contact binary (i.e. TRO and DSC) should be improved.
The intensive disputes by the two contending schools (Lucy & Wilson 1979, Shu, Lubow & Anderson 1979) lead to an intriguing suggestion by Shu, Lubow & Anderson (1979), Shu (1980) (from theoretical viewpoint) and Wang (1995) (from the analysis of observational data) that TRO theory needs the contact discontinuity in some phases. Some observations seem to support TRO theory (Lucy & Wilson 1979, Hilditch et al 1989, Samec et al 1998). Although some criticisms about DSC model exist (Shu, Lubow & Anderson 1980), this theory is still attractive because it is successful in many aspects (Smith 1984). The contact discontinuity may be ironed out within the thermal timescale (Webbink 1977, Hazlehurst & Resfdal 1978, Papaloizou & Pringle 1979, Smith, Robertson & Smith 1980) in a steady state, however, the existence of time-dependent contact discontinuity can not be excluded because it does not violate the second law of thermodynamics (Papaloizou & Pringle 1979). However a detailed analysis is needed for this. It is highly desired to reconcile the two theories not only for removing the discrepancies but also for explaining more detailed observations (Shu 1980).
The difficulties of pure TRO and DSC models in explanation of W-phenomenon motivate us to explore the possibility to develop a second-order theory. The interaction between the secondary and the common convective envelope is thought as an important role in the W-phenomenon, Wang (1994) find that the W-phenomenon can be explained by the released gravitational energy of the secondary through its contraction corresponding to the TRO contracting phase in W-type contact systems. This is encouraging and leads to the suggestions by Wang (1995) from a sample with 32 contact systems that the A-type systems may undergo thermal relaxation oscillation with contact discontinuity whereas the contraction of secondary in W-type systems irons out this contact discontinuity.
The over-riding virtue of a contact discontinuity is that it gives a clear mechanism for making the secondary physically larger, and the primary physically smaller, than their main-sequence single-star counterparts, as is needed to satisfy the Roche-lobe filling requirements of the Kuiper paradox (Lubow & Shu 1977). On the other hand, if the system cannot be maintained in steady state by heat-carrying flows, then the capping of the radiative heat flow from the secondary by the hotter overlying common envelope should lead to an expansion of the secondary, with a resulting transfer of mass from the secondary to the primary. Such Roche-lobe overflow, from a less massive star to a more massive star to a more massive one, is known to occur slowly, so the ultimate breaking of contact caused by the expansion of the binary orbit takes a relatively long time. Once contact has been broken, however, the common envelope disappears; the secondary is no longer capped; and it will begin to shrink toward its normal single-star size. Conversely, because the larger area of the common envelope is no longer available to carry away much of the primary’s interior luminosity, the primary can no longer be maintained at its suppressed contact size, and it will begin to overflow its Roche lobe. The transfer of mass from a more massive star to a less massive one is known to be unstable (e.g. Paczyński 1971), and the rapid shrinkage of the binary orbit causes the system to come into contact again. The re-establishment of the common envelope and the capping of the secondary results in its refilling its Roche lobe. Thus, would the DSC hypothesis provide the physical mechanism for the TRO hypothesis, together with a justification why the duty cycle is long for the contact phase and short for the semidetached phase, as is required by the observational statistics. The rest of this paper attempts to establish the above ideas on a more quantitative basis.
This paper is organized as following: the instability of Roche lobe and its operation in contact system are found in Sec.2; the surviving condition of DSC layer is derived from the thermodynamics in Sec. 3; and the conclusions are remarked in last sectionr.
## 2. The Roche lobe instability and DSC model
It is generally believed that the two components of W UMa stars share an optically thick common convective envelope due to the dynamics equilibrium (Osaki 1965, Mochnacki 1981). The redistribution of the total luminosities (the plus of luminosities of each star), which takes place in CCE, deals with comprehensive fluid processes (Lucy 1968). Why and how to redistribute the luminosties is the main task to theoreticians. The debate of the exsiting theories of contact binaries had been attracted much consideration between 1970s and 1980s (Lucy & Wilson 1979, Shu 1980, Shu, Lubow & Anderson 1981). Shu (1980) clearly stated that the two superficially distinct theories are complementary with the crucial theoretical issue to be resolved being the secular stability of temperature inversion layer from his thought-provoking analysis. Here we argue that DSC layer is a natural results of TRO theory via the mechanism of Roche lobe instability, showing the presence of DSC layer during the expansion TRO phase.
In the following discussions we assume that the total mass and angular momentum are conserved, neglecting the spin angular momentum of two components. These assumptions are basic and the same in TRO theory, but they are unnecessary in the DSC theory. In principle, the two assumptions put more strong constraints on the theoretical model. In the conserved systems there are mainly two other parameters: mass ratio $`q`$, and mass ratio changing rate due to mass exchange $`\dot{q}`$, to determine the structure of the contact binaries in TRO theory. The most serious shortcoming of TRO (mentioned in the previous section) is a strong indicator that we should relax some of assumptions in TRO model. One possible way to remove this shortcoming is to supplement the interaction between CCE and the component. This inclusion may reconcile the two contending schools each other (Wang 1995).
We first show that the instability of mass exchange may prevent from the mass in the secondary being pushed into the primary through the inner Lagrangian point $`L_1`$ due to the lid effects of CCE placed on the secondary (Shu, Lubow & Anderson 1976). For a contact system with total angular momentum $`𝒥`$ in a circular orbit and total mass $`=M_1+M_2`$, the separation between components reads
$$𝒟=\left(\frac{𝒥^2}{G^3}\right)\frac{(1+q)^4}{q^2},$$
(1)
where $`q=M_2/M_1`$ (for the convenience we take $`q1`$), and $`G`$ is the gravitation constant. The Roche lobe radius $`R_\mathrm{L}`$ of the secondary approximates for all mass ratio (Eggleton 1983)
$$r_\mathrm{L}=\frac{R_\mathrm{L}}{𝒟}=\frac{0.49}{0.6+q^{\frac{2}{3}}\mathrm{ln}(1+q^{\frac{1}{3}})}.$$
(2)
The Roche lobe of the primary will be obtained when we replace $`q`$ by $`1/q`$. It is important to note that the Roche lobe is changing due to the mass transfer between the two components. The variation rate of the Roche lobe due to mass transfer between the two components reads
$$\frac{d\mathrm{ln}R_\mathrm{L}}{dq}=\frac{2r_\mathrm{L}}{3q^{\frac{1}{3}}}\left[\frac{1}{1+q^{\frac{1}{3}}}2\mathrm{ln}(1+q^{\frac{1}{3}})\right]+\frac{2(q1)}{q(1+q)},$$
(3)
and then we have the timescale for this change with the helps of $`d\mathrm{ln}R_\mathrm{L}/dt=(d\mathrm{ln}R_\mathrm{L}/dq)(dq/dt)`$
$$t_{\mathrm{R}_\mathrm{L}}=\left(\frac{d\mathrm{ln}R_\mathrm{L}}{dt}\right)^1=f(q)t_\mathrm{M},$$
(4)
where $`t_\mathrm{M}`$ is the timescale of mass transfer defined as
$$t_\mathrm{M}=\frac{M_1}{\dot{M}_1},$$
(5)
here the parameter $`\dot{M}_1`$ is the rate of mass transfer, and the function $`f(q)`$ is
$$f(q)=\left\{\frac{2(1q)}{q}\frac{2r_\mathrm{L}}{3q^{\frac{1}{3}}}\left[\frac{1}{1+q^{\frac{1}{3}}}2\mathrm{ln}(1+q^{\frac{1}{3}})\right](1+q)\right\}^1.$$
(6)
The function $`f(q)`$ represents the ratio of the two timescales. We have calculated the function $`f(q)`$ in Figure 1, showing its value for the range of $`q`$ from 0.0 to 1.0. If $`f(q)>0`$ then Roche lobe will expand with increases of $`q`$ or shrink with decreases of $`q`$. If $`f(q)<0`$ then Roche lobe will shrink with the increases of $`q`$ or expand with decreases of $`q`$. It is very important to address that if $`|f(q)|<1`$ then the expansion or shrinkage will be rapid than the process of mass transfer in the contact system known from equation (4). From Figure 1 the Roche lobe of the secondary expands with the increases of mass ratio. The expansion timescale is shorter than that of gaining mass from the primary until the mass ratio $`q>0.8`$, indicating that the secondary is capable of swallowing more mass when $`q<0.8`$. In this range of mass ratio the mass gaining is unstable. The Roche lobe of the primary will shrink due to the mass exchange. There also is an instability of mass exchange in the range of $`q<0.35`$. This implies that the timescale of Roche lobe shrinkage due to mass transfer is shorter than that of mass transfer. This means that Roche lobe shrinks more rapid than the mass loss. This mass exchange instability plays important role in the structure of contact binaries. The maximum of mass ratio for the instability of the primary is 0.35. It is believed that the mass transfer will be more efficient when $`q<0.35`$ with the presence of the instability of mass exchange. Thus it is expected that the mean mass ratio of A-type systems will be less than that of W-types. This is consistent with the observations. It should be noted that here we do not specify the mechanism for the energy and mass transfer. Of course the direction of mass exchange between the two components is determined by the relative potential of the star surface. Here we are trying to show the instability of mass exchange, namely, represented by $`|f(q)|<1`$.
In the original DSC version the rising convective elements interior to the Roche lobe of the secondary cannot penetrate into the common convective envelope because the resulting buoyancy deficit opposes such penetration. SLA76 argued that there is a mass flow pushed by the slight excess of pressure due to a slight heating of the interior of the secondary under constant volume. It is very important to note that this mass flow process is based on the assumption with constant volume of the secondary. The mass flow was estimated by SLA79, however, their estimation follows up another assumption that all the energy transferred from the primary to the secondary radiates again from the secondary, neglecting the interaction between CCE and the secondary. Now we can work out a condition inhibits the returning of mass flow to the primary through the inner Lagrangian point $`L_1`$. If there is no excess pressure, then the mass flow stops. This is equivalent to $`d\overline{\rho }/dt0`$ if the contact discontinuity being lower than the temperature of CCE survives, namely, $`T=`$constant (we assume the gas is ideal), where $`\overline{\rho }`$ is the mean mass density within the Roche lobe of the secondary defined as $`\overline{\rho }M_2/R_\mathrm{L}^3`$. We thus have
$$\frac{d\mathrm{ln}\overline{\rho }}{dt}\frac{d}{dt}\left[\mathrm{ln}\left(\frac{q}{1+q}\right)\right]\frac{3}{t_{R_\mathrm{L}}},$$
(7)
then the condition no returning of mass flow reads
$$t_{\mathrm{R}_\mathrm{L}}3qt_\mathrm{M}.$$
(8)
We draw the line $`t_{\mathrm{R}_\mathrm{L}}=3qt_\mathrm{M}`$ in Figure 1. It is obvious that all the value of $`f(q)`$ is always less than $`3q`$. This means that all the cases satisfy the condition that no mass flow returns to the primary even beyond the mass exchange instability. Therefore the assumption of constant volume of the secondary should be relaxed in the advanced DSC model. The contact discontinuity is time-dependent from this viewpoint at least, coinciding with that DSC layer could be maintained in a time-dependent model (Papaloizou & Pringle 1979).
## 3. Thermodynamics of DSC layer
By defining the thermal timescale as $`t_{\mathrm{Th}}=_{M\delta m}^M(4\pi r^2\rho v_c)^1𝑑m`$, Webbink(1977) first showed that the thermal diffusion time scale in the common envelope is typically of the same order as the dynamical timescale (is roughly of one orbital period). This makes the contact discontinuity disappear within one orbital period. We call the thermal diffusive process as interaction $`ϵ`$. Papaloizou & Pringle (1979) show the steady contact discontinuity violates the second law of thermodynamics, but the time-dependent contact discontinuity may exist. However in the time-dependent model it is the interaction $`ϵ`$ that keeps the contact discontinuity in contact system undergoing thermal relaxation oscillation. The controversy of inner structure may be removed by this kind of interaction (Wang 1995). With the helps of the conservation of mass and momentum we can rewrite the energy equation beyond the energy generation region as
$$\rho \frac{}{t}(\mathrm{\Psi }+Ts)=\rho T\stackrel{}{v}s\stackrel{}{F}+ϵ,$$
(9)
for the inviscid fluid (e.g. Webbink 1977, and 1992 ApJ, 396, p378 for the erratum), where $`t`$ is time; $`\rho `$, the density; $`T`$, the temperature; $`\stackrel{}{v}`$, the velocity; $`s`$, the specific entropy; $`\stackrel{}{F}`$, the energy flux radiated from the star; $`ϵ`$, the energy density absorbed by the secondary in the unit time due to the interaction with CCE; and $`\mathrm{\Psi }`$, the gravitational energy per unit mass. Following the assumption by Shu, Lubow & Anderson (1980) that the specific entropy $`s`$ can be decomposed in terms of a barotropic and a baroclinic one as $`s=s_0(\mathrm{\Psi }_\mathrm{D})+s_1(\stackrel{}{x},t)`$, we integrate the above equation over the volume enclosed by the equipotential surfaces $`C`$ and $`D`$, and obtain
$`{\displaystyle \frac{dS}{dt}}`$ $`=`$ $`s_0(\mathrm{\Psi }_D){\displaystyle \frac{d(\mathrm{\Delta }M)}{dt}}{\displaystyle _{\mathrm{CD}}\frac{\rho }{T}\frac{\mathrm{\Psi }}{t}𝑑V}+{\displaystyle \frac{ϵ}{T}}\mathrm{\Delta }V`$ (10)
$`{\displaystyle _{\mathrm{CD}}}\rho s_1(\stackrel{}{x},t)\stackrel{}{v}\stackrel{}{n}𝑑A{\displaystyle _{\mathrm{CD}}\frac{1}{T}\stackrel{}{F}𝑑V},`$
where $`S=\rho s𝑑V`$, $`\mathrm{\Delta }M=\rho 𝑑V`$, $`\mathrm{\Delta }V=𝑑V`$ is the volume enclosed by the two surfaces $`C`$ and $`D`$, $`dA`$ is the area of the surface of contact discontinuity, and $`\stackrel{}{n}`$ is its normal vector. For the time-dependent case we assume that the last two terms offset approximately as in steady case (Shu, Lubow & Anderson 1980), thus we have more physically concise form of equation (10)
$$\frac{dS}{dt}=s_0(\mathrm{\Psi }_D)\frac{d(\mathrm{\Delta }M)}{dt}_{\mathrm{CD}}\frac{\rho }{T}\frac{\mathrm{\Psi }}{t}𝑑V+\frac{ϵ}{T}\mathrm{\Delta }V,$$
(11)
The first term of right hand in equation (11) represents the entropy increases due to mass exchange between CCE and the secondary, the last term does the same meanings but due to energy exchange, the second term does the entropy decreases of entropy due to the Roche lobe expansion. The enclosed volume is an open system undergoing mass and energy exchanges with its surroundings rather than an isolated volume. This equation also tells us the resulting expansion due to the interaction $`ϵ`$ if the contact discontinuity layer survives: 1) if there is only exchanges of energy by thermal diffusion, namely, $`\mathrm{\Delta }M=`$const, we have
$$_{\mathrm{CD}}\rho \frac{\mathrm{\Psi }}{t}𝑑Vϵ\mathrm{\Delta }V.$$
(12)
This clearly states that the surviving of contact discontinuity must be provided by the expansion. Detailed calculation should be done in the future. 2) There is a mass exchange between CCE and the secondary accompanying the energy interaction, i.e., $`d(\mathrm{\Delta }M)/dt>0`$, the expansion is at least
$$_{\mathrm{CD}}\rho \frac{\mathrm{\Psi }}{t}𝑑Vϵ\mathrm{\Delta }V+s_0(\mathrm{\Psi }_D)T\frac{d(\mathrm{\Delta }M)}{dt}.$$
(13)
This equation predicts the secular change of orbital period due to the shift of mass ratio. 3) If the secondary keeps constant volume as originally suggested by Shu, Lubow & Anderson (1976), the term $`d\mathrm{\Psi }/dt=0`$, we always have $`dS/dt0`$ which means the discontinuity will be ironed out with the thermal diffusive timescale. The only possible way to relax this condition is the inclusion of the changes of Roche lobe. This way will permits us unifying the two contending hypotheses. According the simplest version of star structure, equations (12) and (13) will provide the expansion velocity
$$v_{\mathrm{int}}R_2\left(\frac{ϵ}{\overline{ϵ}}\right)\left(\frac{\mathrm{\Delta }V}{V_2}\right)+Ts_0(\mathrm{\Psi }_\mathrm{D})g_2^1\frac{d(\mathrm{ln}\mathrm{\Delta }M)}{dt},$$
(14)
where $`\overline{ϵ}=(GM_2\mathrm{\Delta }M/R_2)/V_2`$ is the mean density of the gravitational energy between the secondary and CCE, $`g_2=GM_2/R_2^2`$ the gravitational acceleration, $`V_2`$ is the volume of the secondary. One should note that in the above estimation we neglect the down directed propagation of energy to the interior with dynamical timescale. Therefore the present estimation is somewhat higher than that of the actual. Both of the mass exchange and energy interaction result in the expansion of the secondary, we thus have the minimum velocity
$$V=\mathrm{max}(v_{\mathrm{int}},v_{\mathrm{R}_\mathrm{L}}),$$
(15)
where $`v_{\mathrm{R}_\mathrm{L}}=dR_\mathrm{L}/dt`$ is the expanding velocity of the Roche lobe. This is the condition maintaining contact discontinuity. From the viewpoint of total energy (by the nuclear) conservation, the exhausted energy (i.e. $`ϵ`$) to expand the secondary lowers the re-radiating efficiency of the transferred energy from the primary. The lower temperature of the secondary than the primary may be an indicator of the presence of contact discontinuity.
## 4. Conclusions and Discussions
Introducing a discontinuity of temperature by Shu and his collaborators (1976, 1979) the thermal instability of binary (Lucy 1976, Flannery 1976) can be suppressed, but its maintenance of DSC layer opens. In this paper we try to construct the physical scenario of time-dependent model of contact binary. Only two assumptions that total mass and angular momentum of the contact system are conserved are employed in this paper. It is found that the mass exchange results in the instability of Roche lobe in some ranges of mass ratio. We show that this instability always satisfies the condition that keeps the mean density of the secondary $`d\mathrm{ln}\overline{\rho }/dt0`$. Therefore it ensures that no mass flow returns to the primary through the inner Lagrangain point $`L_1`$. The second-order theory predicts that the contact binary may be in oscillations take place about a state with a contact discontinuity. The temperature differences of DSC layer across the interface is determined by the expansion velocity.
The existing TRO and DSC theories (Lucy 1976, Flannery 1976, Robertson & Eggleton 1977; Shu, Lubow & Anderson 1976, 1979) neglect the effects of interaction $`ϵ`$ between CCE and the secondary. Here we argue that it is the contact discontinuity layer that results in the interaction between CCE and the secondary and in the meanwhile it is the interaction that maintains the contact discontinuity. We find this interaction $`ϵ`$ can result in some interesting issues. First it is the reason why the temperature of the secondary in A-types is lower than that of the primary. Second the maintenance of contact discontinuity needs faster mass transfer which breaks down the deep contact. Thus the shortcoming of TRO theory will be removed. It is highly desired that the unification of TRO theory and DSC model should be calculated in order to discover the nature of the contact binaries.
In the present work we do not specify the mechanism of thermal diffusion process. Although we have not performed the time-dependent model unifying TRO and DSC hypotheses, this time-dependent unified theory might give some predictions. First the secular behavior of period change of A-type systems is violent than that in W-type system in order to survive the existence of contact discontinuity. Second, the maintenance of contact discontinuity may lead to the radial oscillation of the secondary with period from a few to several ten minus. The interaction between CCE and the secondary drives such a oscillation similar to the $`\kappa `$-mechanism working in other types of stars. It is thus expected to find the light variation during the primary eclipse as another probe of contact discontinuity in A-type systems.
The author would like to express his honest thanks to the referee, Professor Frank H. Shu, for his input of physical insight to the title and the fourth paragraph of the first section, enhancing the scientific clarity of this paper. I appreciate the useful discussions with Drs. Zhanwen Han and Fangjun Lu. This project is supported by Climbing Plan of The Ministry of Science and Technology of China and the Natural Science Foundation of China under Grant No. 19800302.
|
no-problem/9908/cond-mat9908422.html
|
ar5iv
|
text
|
# On the problem of an electron scattering in an arbitrary one-dimensional potential field
## Abstract
Recurrent representations for an electron transmission and reflection amplitudes for a one-dimensional chain are obtained. The linear differential equations for scattering amplitudes of an arbitrary potential are found.
In this work the electron scattering problem is considered, when the potential energy has the form .
$$V(x)=\underset{n=1}{\overset{N}{}}V_n(xx_n)$$
(1)
Here $`V(xx_n)`$ is an individual potential, which has located near the $`x_n`$point. It is suggested, that individual potentials $`V(xx_n)`$ do not have general points.
There have been many studies to investigate electronic properties of one-dimensional quaziperiodical and disordered systems. It is well known, that the problem of determination of transmission amplitude (TA) and reflection amplitude (RA) in the field of (1) type is reduced to the problem of $`N`$ matrixes product calculation ;
$$\left(\begin{array}{c}T_N\\ 0\end{array}\right)=\underset{N}{\overset{n=1}{}}\left(\begin{array}{cc}1/t_n^{}& r_n^{}/t_n^{}\\ r_n/t_n& 1/t_n\end{array}\right)\left(\begin{array}{c}1\\ R_N\end{array}\right)$$
(2)
where $`T_N`$, $`R_N`$are TA and RA of potential $`V(x)`$, $`t_n`$ and $`r_n`$are TA and RA of the individual potential $`V(xx_n)`$ of the system (1).
The problem of calculation of the $`N`$ two-order matrixes product (2), in general, is equivalent to the problem of solution of some linear equations . So, the problem is represented as
$$D_N=r_N/t_NP_{N1}+1/t_ND_{N1}N1$$
(3)
$$P_N=r_N^{}/t_N^{}D_{N1}+1/t_N^{}P_{N1}.N1$$
(4)
Here $`D_N=1/T_N`$, $`P_N=R_N^{}/T_N^{}`$ and $`D_{Nn}=1/T_{Nn}`$, $`P_{Nn}=R_{Nn}^{}/T_{Nn}^{}`$, where $`T_{Nn}`$ and $`R_{Nn}`$are TA and RA of the first potentials of the field (1). Note, that in (3, 4) $`N`$ is a variable.
Let us now introduce (3, 4) in more confortable form. Excluding $`P_{N1}`$ from the first equation and $`D_{N1}`$ from the second one, we will get two equations for $`D_N`$ and $`P_N`$ ;
$$D_N=A_ND_{N1}B_ND_{N2}N2$$
(5)
$$P_N=A_N^{}P_{N1}B_N^{}P_{N2},N2$$
(6)
where $`A_N=1/t_N+B_N/t_{N1}^{}`$ and $`B_N=r_Nt_{N1}/r_{N1}t_N`$.
So the problem of determination of $`T_N`$and $`R_N`$ is reduced to the problem of solution of the equations (5, 6). In order to have the simple solution of (5, 6), it is necessary to give initial conditions for equations (5) and (6). They are
$$D_1=t_1^1,D_0=1,P_1=r_1^{}/t_1^{},P_0=0.$$
(7)
Note, that representations (5, 6) take place for arbitrary potentials of the type (1). So, if we take $`t_N=1+i\alpha _N/2k`$ and $`r_N=i(\alpha _N/2k)\mathrm{exp}2ikx_N`$, then (5) will coincide with the result of , received by the ”determinant” method for $`\delta `$\- potentials. In a case of layered system from homogeneous mediums, (5, 6) yields to the result of .
It is interesting to apply the obtained result (5, 6) to a simple case, when potentials $`V(xx_n)`$ are located periodically and have the same form. Then $`t_n`$and $`r_n`$ $`\left(n=1,2,\mathrm{},N\right)`$ can be represented as $`t_n=t_1`$ and $`r_n=r_1\mathrm{exp}2ik(x_nx_1)`$ . In this case, according to (5, 6), it is clear that coefficients $`A_N`$ and B<sub>N</sub>become independent from $`N`$ and the equations (5, 6) are easily solved .
Taking into account conditions (7) and denoting the systems period by $`a`$, for $`D_N`$ and $`P_N`$ the following expressions are obtained
$$D_N=\mathrm{exp}iNka\left\{\mathrm{cos}N\beta +iIm(t_1^1\mathrm{exp}(ika))\mathrm{sin}N\beta /\mathrm{sin}\beta \right\},$$
(8)
$$P_N=r_1^{}/t_1^{}\mathrm{exp}i(N1)ka\mathrm{sin}N\beta /\mathrm{sin}\beta .$$
Here $`k^2=E(\mathrm{}=2m_0=1)`$ is the electron energy. The electron energy spectrum is given by $`\mathrm{cos}\beta =Re(t_1^1\mathrm{exp}(ika))`$ , which was obtained in .
Let us now considered one important form of the potential $`V(x)`$ (1), when $`V(xx_n)`$ the are rectangular potentials with arbitrary width $`2d_n`$ and magnitudes of potentials $`V_n`$. Then $`t_n`$and $`r_n`$ are given by well-known formulas ;
$$t_n^1=\mathrm{exp}i2kd_n\left\{\mathrm{cos}2kd_ni\frac{k_n^2+k^2}{2k_nk}\mathrm{sin}2kd_n\right\},$$
(9)
$$r_n/t_n=i\mathrm{exp}i2kx_n\frac{k_n^2k^2}{2k_nk}\mathrm{sin}2kd_n.$$
(10)
where $`k_n=\sqrt{EV_n}`$, $`x_n`$ is a coordinate of the middle point of the potential.
This model (9, 10) can be used to obtain differential equations for $`D(x)=1/T(x)`$ and $`P(x)=R^{}(x)/T^{}(x)`$ quantities, where $`T(x)`$ and $`R(x)`$ are TA and RA of the potential
$$V(x)=U(y)\theta (xy)$$
(11)
where $`U(y)`$ is an arbitrary function limited within some interval $`ayb`$ and is equal to zero outside of that.
Indeed, approximating $`V(x)`$ by rectangular potentials, it is possible with the help of representation (5, 6) and corresponding limit transition to get the sought equation. If we denote in (5, 6) $`D_{N1}=D(x)`$, then $`D_N=D(x+\mathrm{\Delta }x)`$ will correspond to the potential $`V(x+\mathrm{\Delta }x)`$. Respectively, $`D_{N2}=D(x\mathrm{\Delta }x)`$ will correspond to the potential $`V(x\mathrm{\Delta }x)`$. Presenting in the expression (9, 10,) $`V_N=V(x+\mathrm{\Delta }x),V_{N1}=V(x),2d_n=\mathrm{\Delta }x`$ and expanding (5, 6) in a series by infinitely small quantity $`\mathrm{\Delta }x`$, we get
$$\frac{1+B}{2}\frac{d^2D}{dx^2}(\mathrm{\Delta }x)^2+(1B)\frac{dD}{dx}\mathrm{\Delta }x(AB1)D+\theta \left((\mathrm{\Delta }x)^3\right)=0,$$
(12)
where
$$B1=\left(2ik+\frac{1}{V(x)}\frac{dV(x)}{dx}\right)\mathrm{\Delta }x+\theta \left((\mathrm{\Delta }x)^2\right),$$
$$AB1=V(x)(\mathrm{\Delta }x)^2+\theta \left((\mathrm{\Delta }x)^3\right).$$
(13)
Inserting (13) in to (12) and taking $`\mathrm{\Delta }x0`$, we find the desired equation for $`D(x)`$:
$$\frac{d^2D}{dx^2}\left(2ik+\frac{1}{V(x)}\frac{dV(x)}{dx}\right)\frac{dD}{dx}V(x)D=0$$
(14)
To find TA and RA of the potential $`V(x)`$ it is necessary to solve the equation (14) with the following initial conditions:
$$D(a)=1,dD/dx_{x=a}=iV(a)/2k$$
(15)
The value of the function $`D(x)`$ in the point $`x`$ determines TA of the potential $`V(x)`$. According to the equations (5, 6), the quantity $`P^{}(x)`$ satisfies the obtained equation (14) too. The initial conditions for $`P^{}(x)`$ are the following
$$P^{}(a)=0,dP^{}/dx_{x=a}=iV(a)/2k\mathrm{exp}i2ka$$
(16)
Let us now consider the obtained result (14) for a case of a simple rectangular potential wall with width $`L=ba`$ and a constant potential $`V(x)=V_n`$. Then, it is easy to see, that the solution of the problem (14)-(15) is
$$D(b)=\mathrm{exp}ikL\left\{\mathrm{cos}k_nLi\frac{k_n^2+k^2}{2k_nk}\mathrm{sin}k_nL\right\},$$
(17)
which, as was expected, agrees with (9) for $`L=2d_n`$.
In conclusion, it is interesting to note, that taking $`\mathrm{exp}ikxD(x)`$ as $`F_1(x)`$ and $`\mathrm{exp}ikxP(x)`$ as $`F_2(x)`$, from equation (14), we obtain the Schrödinger equation for function $`F_1(x)F_2(x)=L(x)`$;
$$\frac{d^2L}{dx^2}+\left(EV(x)\right)L=0$$
(18)
while the $`F_1(x)+F_2(x)=\frac{i}{k}\frac{dL(x)}{dx}`$.
Thus, the problem of determination of TA and RA is reduced to the Cauchy problem for Shrodinger equation (18), with initial condition (7).
|
no-problem/9908/hep-ph9908396.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Primordial Big Bang Nucleosynthesis (BBN) is without doubt one of the biggest successes of early universe cosmology. Not only does it provide a stringent test of the Big Bang model, predicting the light element abundences as a function of only a single parameter, $`\eta =n_b/n_\gamma `$, the cosmological baryon to photon ratio, but it also supplies important constraints on particle physics, the most well-known example being the determination of the number of light neutrino species. Given the consistency between the primordial abundance of light elements (inferred from observation extrapolated back to the primordial values) and theoretical calculations, BBN does not leave much room for extra particles which otherwise could have existed in the early universe.
Many extensions of physics beyond the Standard Model (SM), however, introduce additional relativistic degrees of freedom at the epoch of BBN. A small selection of such new light degrees of freedom include: one or more sterile neutrinos which might be required by the neutrino oscillation data; a light gravitino as a consequence of a low fundamental scale of supersymmetry breaking as in the gauge-mediated scenarios; a hadronic axion in the hot dark matter window, and many other examples. If such new light degrees of freedom exist, the expansion rate at the BBN epoch is faster, resulting in an earlier freeze-out of neutrons and hence a larger number of them, therefore overproducing <sup>4</sup>He. Taking the BBN constraint seriously, it is then necessary to modify standard BBN, and the simplest and most elegant possibility is a large lepton asymmetryOne other possibility discussed in the literature is that of a late-decaying $`\nu _\tau `$., a possibility which is not ruled out by current observational limits . Specifically, a large positive asymmetry in the electron number implies an excess in the number of electron neutrinos over that of electron anti-neutrinos, thereby shifting the chemical equilibrium between protons and neutrons towards protons. This results in a smaller number density of neutrons after the freeze-out and hence in a reduced <sup>4</sup>He abundance. This effect can therefore compensate the effect of the larger expansion rate due to the additional particle species.
Moreover in recent years, with the advent of new and refined data on the relative abundances of the light elements, there may be appearing a slight but significant discrepancy between the data and the theoretical predictions. In particular, if the recent low measurements of the primordial Deuterium abundance are correct and the <sup>4</sup>He abundance is as low as reported in , then some modification of the standard BBN scenario seems to be required independent of the conjectured existence of new light degrees of freedom. The most promising such modification is again the assumption of a positive chemical potential for electron neutrinos which reduces the final <sup>4</sup>He abundance closer to the reported value. It is noteworthy that the preferred sign of the electron asymmetry is the same for both purposes: to compensate the effect of additional particle species and to bring the BBN prediction closer to observations. Of course, given the uncertainties in the data, it is not clear if this is really required by primordial nucleosynthesis. It is, however, useful to explore such modifications of the standard Big-Bang scenario to see if they are either disfavored by other data, or serve some further, unexpected, purpose.
On the other hand, there is an apparent contradiction of an assumption of large lepton asymmetry with the very small observed baryon asymmetry. This arises from the presence of sphaleron mediated transitions at temperatures of the weak scale and above which tend to quickly equilibriate the lepton and baryon asymmetries, resulting in far too large a baryon asymmetry today. There are three logical possibilities for how a large $`\nu _e`$ lepton asymmetry can be compatible with the small baryon asymmetry: (1) Sphalerons were never in equilibrium, (2) The lepton asymmetry is generated after the electro-weak phase transition but before BBN, and, (3) The total lepton asymmetry across all three generations is zero.
In this letter we focus on the third possibility – in particular the case where $`L_e=L_\mu 0`$ and $`L_\tau =0`$ – and show, in Section 3, that it has a very pretty and unexpected consequence – the natural generation within the Standard Model of a small baryon asymmetry of the observed size, at least within a factor of two, and with the correct sign! This numerical coincidence is quite remarkable, especially given the simplicity and naturalness of the baryon asymmetry generation mechanism. The fundamental reason for the small baryon asymmetry in this case, $`L_e=L_\mu 0`$, is quite simple; it is just a consequence of the small muon Yukawa coupling. As we show in Section 4, if one goes to the minimal supersymmetric standard model (MSSM) then even the factor of two discrepancy between the predicted and observed baryon asymmetry disappears for large $`\mathrm{tan}(\beta )`$. (Section 2 contains a more extensive discussion of the reasons for considering a Lepton asymmetry, together with its possible size and sign.)
## 2 BBN with Large Lepton Number
We will now argue in detail that it is useful to explore the possibility that there may be a slight modification of standard BBN, and that such modifications are certainly not disallowed and are possibly even favored by the light element abundances.
Many particle physics models beyond the SM introduce additional particle species which could be relativistic and thermal at the BBN epoch. Probably the most discussed such example is a sterile neutrino (or many of them, especially in the context of neutrinos from large extra dimensions ). If one takes all existent hints for neutrino oscillations seriously, namely the atmospheric neutrino oscillations, solar neutrino deficit and the results from the LSND experiment, the data cannot be accommodated by neutrino oscillations between the three known species: $`\nu _e`$, $`\nu _\mu `$, $`\nu _\tau `$. The reason is simple. The three hints for oscillations listed above require different values of the mass-squared differences $`\mathrm{\Delta }m^2`$, and with three neutrinos only, the sum of $`\mathrm{\Delta }m^2`$ should vanish. The only known way to explain the data fully by neutrino oscillations is by introducing an additional “sterile neutrino” $`\nu _s`$, thereby allowing yet another mass-squared difference to account for three oscillation modes. However, neutrino oscillations should have occurred in the early universe as well, thus producing sterile neutrino states. In order not to overproduce <sup>4</sup>He due to the additional sterile neutrino energy density, the quoted bounds are
$`\mathrm{\Delta }m^2\mathrm{sin}^42\theta \begin{array}{c}<\hfill \\ \hfill \end{array}5\times 10^6\mathrm{eV}^2,\nu =\nu _e,`$ (9)
$`\mathrm{\Delta }m^2\mathrm{sin}^42\theta \begin{array}{c}<\hfill \\ \hfill \end{array}3\times 10^6\mathrm{eV}^2,\nu =\nu _{\mu ,\tau }.`$ (18)
These constraints, taken literally, imply that sterile neutrinos cannot be responsible for atmospheric neutrino oscillations or the large angle MSW solution to the solar neutrino problem. The existence of a sterile neutrino exceeding the above bounds would increase the effective number of neutrinos at BBN by one: $`\mathrm{\Delta }N_\nu =1`$.
In supersymmetric theories, a light gravitino $`\stackrel{~}{G}`$ may be present at the BBN epoch as well. According to the estimate in Ref. , the gravitinos remain thermal down to the BBN epoch if
$$m_{3/2}\begin{array}{c}<\hfill \\ \hfill \end{array}10^{13}\mathrm{GeV}\left(\frac{m_{\stackrel{~}{l}}}{100\mathrm{G}\mathrm{e}\mathrm{V}}\right),$$
(19)
due to the process $`l^+l^{}\stackrel{~}{G}\stackrel{~}{G}`$. This roughly corresponds to a primordial supersymmetry breaking scale below a TeV. Such a low scale is not expected in the conventional hidden sector models or gauge mediation, but can occur in models where the supersymmetric standard model is directly involved in the mechanism of dynamical supersymmetry breaking (see, e.g., the model in Ref. ). Because the produced gravitino states are dominantly helicity $`\pm 1/2`$ (the would-be Nambu-Goldstino state), they increase the effective number of neutrinos by $`\mathrm{\Delta }N_\nu =1`$.
Invisible axions are another candidate particle that could be present at the BBN epoch. Despite strong constraints from astrophysics, a hadronic (KSVZ) axion in the mass range 3–20 eV is allowed as long as its coupling to the photon is accidentally suppressed . This is an interesting window for a Hot Dark Matter component of the universe which some recent analyzes of large scale structure prefer (however, for conflicting views, see ). The axion in this mass window would contribute to the energy density as an equivalent of $`\mathrm{\Delta }N_\nu =0.4`$–0.5 and is marginal from the BBN point of view.
Yet another example of an exotic particle which might be in thermal contact during BBN is represented by the majoron, the Goldstone boson associated to the spontaneous breakdown of lepton number. Majorons stay in thermal equilibrium as long as $`\tau `$-neutrinos, and provide a contribution to $`\mathrm{\Delta }N_\nu `$ of about 0.6 .
Given these important constraints from BBN on particle physics models, it is important to ask how rigid the constraint actually is. In this regard it is interesting to note that the BBN itself may require some modifications.
Specifically, if one takes the low Deuterium measurement and the reported statistical average of the <sup>4</sup>He abundance extrapolated to zero metalicity , they cannot be reconciled with detailed BBN calculations by choosing an appropriate value of $`\eta `$, the baryon to photon ratio. Of course, it is not yet established that these measurements are reliable. For instance, one should take seriously the conflicting measurement of the Deuterium abundance based on the same technique which returns a high value , even though it has been challenged on the basis of a possible overlap with a foreground cloud and less systematic checks than the low abundance observation. (It is interesting to note that by including turbulence effects in the extraction of the D/He ratio , all the data is cosistent with a low value of D/He $`3.55.2\times 10^5`$.) The “best” determination of the <sup>4</sup>He abundance has also been challenged by a re-analysis of the more-or-less the same data set . Nevertheless there is motivation for considering modifications to BBN which can reconcile the “best” determinations of element abundances. Most certainly, such a modification is allowed by current data. (For a recent review see Ref. .)
It is noteworthy that both the presence of additional relativistic degrees of freedom and the apparent inconsistency between the D and <sup>4</sup>He abundances prefer a mechanism to reduce the effective number of neutrinos $`N_\nu `$. Two such possibilities have been proposed in the literature:
1. A late-decaying $`\nu _\tau `$ with a mass of $`m_{\nu _\tau }10`$ MeV and a lifetime of $`\tau 10^2`$$`1`$ sec .
2. A large chemical potential for $`\nu _e`$ .
The former proposal is interesting from the collider physics point of view because it is testable in the forthcoming $`B`$-factory experiments .
In this letter we focus on the second possibility. Here the idea is that the presence of a large chemical potential for $`\nu _e`$ makes the $`\nu _e`$ number density larger than the thermal number density without chemical potential, which in turn changes the chemical equilibrium of the reaction $`\nu _ene^{}p`$ etc. The presence of a positive chemical potential for $`\nu _e`$ shifts the equilibrium towards the right-hand side, which reduces the neutron number density at the freeze-out. Therefore the <sup>4</sup>He abundance is reduced for a given value of $`\eta `$. Since the $`D`$ abundance prefers a relatively large value of $`\eta `$, which prefers a large <sup>4</sup>He abundance, the reduced prediction for the <sup>4</sup>He abundance would allow additional relativistic degrees of freedom present at the BBN epoch or reconciles the apparent conflict between the observations and the calculations.In the case of neutrino oscillations to a sterile neutrino, the interplay between the neutrino oscillations and thermalization can be quite complicated . However, a large primordial lepton asymmetry which exists from the pre-BBN era does persist and can allow the sterile neutrinos. This differs from the situations discussed in where the lepton asymmetry was assumed to vanish primordially (i.e., before the BBN era).
The electron-neutrino chemical potential affects the neutron-to-proton ratio at the freeze-out as
$$\left(\frac{n}{p}\right)_{\xi _{\nu _e}0}=\left(\frac{n}{p}\right)_{\xi _{\nu _e}=0}e^{\xi _{\nu _e}},$$
(20)
where $`\xi _{\nu _e}=\mu _{\nu _e}/T`$ at the freeze-out temperature. The effect of the extra degrees of freedom on <sup>4</sup>He abundance is given by an analytic fit : $`\mathrm{\Delta }Y_P=0.0075\mathrm{\Delta }g_{}=0.013\mathrm{\Delta }N_\nu `$. Therefore, an approximate dependence of $`Y_P`$ on the extra degrees of freedom and the chemical potential is given by
$$Y_P=\left(0.225+0.025\mathrm{log}_{10}\left(\frac{\eta }{10^{10}}\right)+0.013\mathrm{\Delta }N_\nu \right)e^{\xi _{\nu _e}}$$
(21)
for $`\tau _{1/2}(n)=10.24`$ minutes. The low D measurement requires $`\eta 5\times 10^{10}`$ and hence $`Y_P0.242`$ which is beyond the quoted $`Y_P=0.234\pm 0.002\pm 0.005`$ (see, however, a conflicting number $`Y_P=0.244\pm 0.002\pm 0.005`$ ). This would require $`\xi _{\nu _e}0.0034`$. This approximate discussion also tells us that an additional degrees of freedom with $`\mathrm{\Delta }N_\nu =1`$ can be compensated by $`\xi _{\nu _e}=0.056`$.
The size of the chemical potential favored to reconcile the observations and the BBN calculations of the light element abundances were studied by intensive numerical analysis in Ref. . The result is $`\xi _{\nu _e}=(4.3\pm 4.0)\times 10^2`$ at 95% CL, quite close to the rough estimate given above. From this the electron-number per photon ratio is given by
$$\frac{n_{\nu _e}n_{\overline{\nu }_e}}{n_\gamma }=\frac{\pi ^3}{12\zeta (3)}\left(\frac{T_{\nu _e}}{T_\gamma }\right)^3\left(\frac{\xi _{\nu _e}}{\pi }\right)+𝒪(\xi ^3).$$
(22)
Since $`T_{\nu _e}=T_\gamma `$ in the relevant temperature regime, and the total entropy density is $`s=\frac{43}{4}\frac{2\pi ^2}{45}T^3`$ (from photons, electrons, positrons and three neutrinos), we find the “preferred” electron-number to entropy ratio $`L_e`$ to be
$$L_e^{\mathrm{NUC}}=\frac{15}{43}\frac{\xi _{\nu _e}}{\pi ^2}=(1.52\pm 1.41)\times 10^3.$$
(23)
For the purpose of allowing an extra relativistic degree of freedom at the epoch of BBN, we would also require an additional contribution the electron-number to entropy ratio of this same magnitude and sign. Thus we take
$$L_e^{}2L_e^{\mathrm{NUC}}=(3.04\pm 2.82)\times 10^3$$
(24)
as the favored value of the lepton asymmetry both by compensating an additional relativistic degree of freedom at the BBN epoch and by reconciling the discrepancy between the theory and observation in the BBN itself.
## 3 Small Baryon Number from Large Lepton Number
The most uncomfortable aspect of a large chemical potential for $`\nu _e`$ is the consistency with the small observed baryon asymmetry. An almost universal theoretical prejudice is that the baryon asymmetry is a consequence of non-trivial dynamics in the Early Universe, with the three Sakharov conditions being met: (1) the existence of a baryon-number violating interaction, (2) departure from thermal equilibrium, and (3) CP-violation. If there were also a chemical potential for $`\nu _e`$, or in other words, an asymmetry in the electron number, it should also be a consequence of similar dynamics in the Early Universe. It then appears unnatural that the lepton asymmetry is many orders of magnitude larger than the baryon asymmetry if they are generated by similar mechanisms.
The uncomfortableness mentioned above becomes a conflict in the view of the following consideration. Given the difficulty in generating a large enough baryon asymmetry purely from the electroweak phase transition, the much larger preferred size of the lepton asymmetry from the BBN, Eqn. (23), is highly unlikely to be a consequence of physics at or below the electroweak scale. However, above the electroweak phase transition, neither baryon- or lepton-number is conserved, but only $`BL`$ because of sphaleron mediated transitions and the electroweak $`B`$ and $`L`$ anomalies . Furthermore, the chemical equilibrium induced by sphaleron transitions enforces the baryon- and lepton-asymmetries to be of the same orders of magnitude.
There are three logical possibilities to avoid this conflict:
1. The large lepton asymmetry is generated below the electroweak scale.
2. The sphaleron transition was never in equilibrium below the temperature at which the lepton asymmetry was generated.
3. The total lepton asymmetry vanishes, while the individual lepton-flavor asymmetries do not.
We already argued that the first possibility is unlikely, even though it is logically possible. The second possibility arises if the large lepton number asymmetry causes a Bose condensate of electroweak-doublet scalar fields . In the Standard Model the preferred value of the lepton asymmetry from nucleosynthesis considerations is below the critical value at which the Higgs doublet acquires a large expectation value and thus at temperatures above the electroweak scale the sphaleron transition is still in equilibrium. The same is true in the case of the MSSM as recently shown in Ref. . Note, in particular, that if the squark and slepton masses are heavier than the electroweak phase transition temperature of 100–200 GeV, they are irrelevant to this discussion and the situation is the same as in the SM and hence the sphaleron transitions are active. Moreover, even if one manages to keep sphaleron transitions out of equilibrium, it still does not resolve the question why the lepton asymmetry is so much larger than the baryon asymmetry. From these considerations, we find the third possibility to be the most interesting one, which has not been discussed in the literature so far.
The baryon and the lepton asymmetries are determined by the $`BL`$ asymmetry via sphaleron-induced chemical equilibrium. For the Standard Model :
$$B=\frac{8N_G+4N_H}{22N_G+13N_H}(BL),$$
(25)
where $`N_G=3`$ is the number of generations and $`N_H`$ is the number of Higgs doublets (1 in the SM). In the presence of the supersymmetric particles, the formula is slightly modified . Therefore, if the total lepton asymmetry vanishes, the total baryon number also vanishes. This way, one can obtain a vanishing baryon asymmetry even in the presence of individual flavor-dependent lepton asymmetries.
The above formula is usually assumed to hold above the electroweak phase transition temperature, while it requires modification after the phase transition because of finite mass effects. However, even above the phase transition temperature, the effects of thermal masses need to be considered. Such effects are small and usually ignored, but they cannot be ignored in the presence of the large individual lepton numbers of interest in this letter.
The final resulting baryon asymmetry depends on when the sphaleron transition freezes out, which in turn depends on whether the electroweak phase transition is strongly first-order or not . Given the experimental lower bound on the Higgs mass of about 95 GeV together with the results of current large-scale numerical lattice simulations and analytic arguments , the phase transition in the Standard Model is certainly not a strongly first order transition, while in the case of the MSSM a weakly first-order transition or smoother is favored over much of the parameter space. In the case that the phase transition is second order, or if the sphalerons are still active after a first order phase transition (i.e., a weakly first-order transition with $`\varphi (T)/T1`$, being $`\varphi `$ the vacuum expectation value of the Higgs field), there are two contributions to the resulting baryon asymmetry. These flavor-dependent effects both arise from the interaction of electrons and muons with the Higgs boson via their Yukawa couplings. (The two effects correspond to the interactions with condensed and real Higgs bosons respectively.) The total flavor-dependent effect was estimated in Ref. , and in the case of vanishing total lepton asymmetry $`L_e+L_\mu =0`$, we find
$$B=A\frac{6}{13\pi ^2}\frac{\overline{m}_\mu ^2(T)}{T^2}L_e^{},$$
(26)
where $`A1`$ and
$$\frac{\overline{m}_\mu ^2(T)}{T^2}=\frac{1}{6}f_\mu ^2+\frac{1}{3}f_\mu ^2\left(\frac{v(T)}{T}\right)^2\frac{1}{2}f_\mu ^2=1.8\times 10^7.$$
(27)
The resulting baryon-to-photon ratio in this case is
$$\eta =(1.8\pm 1.68)\times 10^{10}$$
(28)
This should be compared to the preferred value from BBN, e.g. $`\eta =(4.0_{0.9}^{+1.5})\times 10^{10}`$. Thus we find agreement with the required value at the upper edge of the 95% CL region!.
Notice that, if the electroweak phase transition is strongly first order with $`\varphi (T)/T`$ larger than unity after the transition, the sphaleron processes are frozen-out and absent after the transition. In this case the chemical equilibrium before the transition determines the baryon asymmetry. The only flavor-dependent effects before the transition are the Yukawa interactions of electrons and muons with the uncondensed Higgs boson. In the case of vanishing total lepton asymmetry $`L_e+L_\mu =0`$, we now find
$$\frac{\overline{m}_\mu ^2(T)}{T^2}=\frac{1}{6}f_\mu ^2=\frac{\pi \alpha _W}{3}\frac{m_\mu ^2(0)}{m_W^2}=6.0\times 10^8.$$
(29)
This translates into baryon and lepton to entropy ratios of
$$B=\frac{6}{13\pi ^2}(6.0\times 10^8)L_e^{}=(8.6\pm 8.0)\times 10^{12}.$$
(30)
This corresponds to a current baryon-to-photon ratio of $`\eta =(6.0\pm 5.6)\times 10^{11}`$, which is off by more than a factor of three.
## 4 Model Building
We have seen that the preferred value of $`L_e`$ from Eqn. (23), $`L_e^{}=2L_e^{\mathrm{NUC}}`$, together with the relation $`L_e=L_\mu `$ gives the correct order of magnitude and sign for the baryon asymmetry. We find this a remarkable coincidence.
Suppose however one takes the preferred value of the lepton asymmetry to be $`L_e^{\mathrm{NUC}}`$, i.e. let us not allow any room for extra degrees of freedom during nucleosynthesis. Then from Eqn. (26) the baryon asymmetry turns out to be correct except for a factor of two or so. A natural question then is if there are corrections that can fix this factor-of-two discrepancy so that the generation of the observed small baryon asymmetry from the magnitude of the lepton asymmetry currently preferred from the BBN is a realistic possibility. We find that there are many ways to achieve this. Another natural question is if there is an appropriate leptogenesis mechanism which can create a large lepton asymmetry with $`L_e=L_\mu `$ in a simple way.
The simplest possibility to enhance the baryon asymmetry is to consider the MSSM where all sleptons and squarks are heavier than the electroweak phase transition temperature while the entire Higgs sector, $`h^0`$, $`H^0`$, $`A^0`$ and $`H^\pm `$ is light. In the approximation where one ignores their masses, the lepton doublets interact only with the $`H_d`$ doublet with the Yukawa coupling $`f_l/\mathrm{cos}\beta `$. Here $`\mathrm{tan}\beta H_u/H_d`$ is the vacuum angle. In this limit, the only change from the case of the SM is to replace the Yukawa couplings $`f_l`$ by $`f_l/\mathrm{cos}\beta `$, which enhances the plasma mass effects. The net result is an enhanced baryon asymmetry which brings the predicted value into the required range for a moderate value of $`\mathrm{tan}\beta `$. If the masses of the Higgs bosons cannot be neglected, the enhancement effect is reduced. But it is clear that a realistic value of the baryon asymmetry can be easily achieved.
There are many other possible enhancement mechanisms of the baryon asymmetry. For instance, light higgsinos and sleptons also contribute to the plasma mass of the lepton doublets. These effects are enhanced by $`1/\mathrm{cos}^2\beta `$ but Boltzmann-suppressed by their masses $`e^{m/T}`$. For suitable values of $`\mathrm{tan}\beta `$ and the slepton and higgsino masses estimates indicate that the required factor of two is generated. (A detailed quantitative analysis involves the generalization of the formulae in to include the individual lepton asymmetries.) It is therefore clear that there are quite simple extensions of the SM which fairly naturally provide the required factor of two.
We now turn to the question of whether it is possible to generate a large lepton asymmetry with $`L_e=L_\mu `$ in a natural and elegant fashion. Such leptogenesis with $`L_e=L_\mu `$ can be achieved naturally by utilizing the Affleck–Dine mechanism .<sup>§</sup><sup>§</sup>§It was discussed recently also in that one can generate a large lepton asymmetry by the Affleck–Dine mechanism. The author however required an even larger asymmetry than what we discuss to keep the electromagnetism as well as sphalerons out of equilibrium to solve the monopole problem and avoid the overproduction of baryon asymmetry . Therefore the aim of the paper is orthogonal to ours. This requires the operator
$$d^4\theta (m_{3/2}\theta ^2)(m_{3/2}\overline{\theta }^2)\frac{L_e^{}L_\mu H_u^{}H_u}{M_X^2},$$
(31)
where the supersymmetry-breaking spurions are inserted. This operator preserves the total lepton number, while breaking $`L_e`$ and $`L_\mu `$ individually. The energy scale of this operator, $`M_X`$, can be, for example, the (reduced) Planck scale $`M_{}=2\times 10^{18}\mathrm{GeV}`$. The $`D`$-flat direction $`|L_e|^2+|L_\mu |^2=|H_u|^2`$ is lifted by this operator and the field acquires a large “angular momentum” in the internal space. This corresponds to the generation of individual lepton numbers satisfying $`L_e=L_\mu `$. This leads to an estimate of the lepton number,
$$L_e=L_\mu \frac{\varphi _0^4T_{RH}}{m_{3/2}M_X^2M_{}^2},$$
(32)
where $`T_{RH}`$ is the reheating temperature of primordial inflation, $`m_{3/2}`$ is the typical mass of the sleptons and Higgs bosons, and $`\varphi _0`$ is the initial amplitude of the slepton expectation values. Even taking account of the constraint from gravitino overproduction $`T_{RH}\begin{array}{c}<\hfill \\ \hfill \end{array}10^9`$ GeVThis bound is obtained considering the thermal production of gravitinos. However, it has been recently pointed out that this mechanism of productio is overcome by the non-thermal generation of gravitinos . and $`m_{3/2}1`$ TeV, the initial value of the amplitude can be relatively small $`\varphi _0\begin{array}{c}>\hfill \\ \hfill \end{array}(10^3L_eM_XM_{})^{1/2}`$. Taking $`M_XM_{}`$, Eqn. (32) shows that $`\varphi _010^{15}`$ GeV is sufficient to generate the large lepton asymmetry that we require. Note that the detailed mechanism for generating a large initial amplitude, $`\varphi _0`$, is model-dependent; it could be a negative mass-squared during the inflationary epoch or quantum effects .Since the total lepton number is preferably conserved within our scenario, the neutrino masses should be Dirac rather than Majorana. The atmospheric neutrino oscillation prefers a small Yukawa coupling of order $`h_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}10^{12}`$. Even though this Yukawa coupling lifts our flat direction, a negative mass squared of, for instance, $`H_{inf}^2`$ during inflation, generates an initial amplitude of $`\varphi _0H_{inf}/h_\nu `$ which is well beyond what we need given the typical value of the Hubble constant during inflation $`H_{inf}10^{11}`$$`10^{13}`$ GeV. Such a small Yukawa coupling could be a natural consequence of a flavor symmetry .
One final concern is if this scenario is consistent with the reported atmospheric neutrino oscillation: If the generated asymmetry $`L_\mu `$ is converted partially to an asymmetry in $`L_3`$, it could then generate too large a baryon asymmetry because of the large tau Yukawa coupling $`f_\tau `$. This fortunately does not happen. By the time of the electroweak phase transition, the probability for neutrino oscillation is suppressed by $`\mathrm{sin}^2(\mathrm{\Delta }m^2t/4E_\nu )`$, where $`tM_{}/m_W^2`$ and $`E_\nu m_W`$. Substituting the relevant $`\mathrm{\Delta }m^2`$ into this expression then shows that the oscillation to $`L_3`$ is negligible.
## 5 Conclusion
Over the years, many mechanisms for the generation of the tiny observed baryon asymmetry have been proposed and we have very little idea which if any is the correct one. Furthermore, many of the proposed baryogenesis mechanisms are not able to predict the resulting baryon asymmetry to better than an order of magnitude (sometimes many). On the other hand, so far there is no observational evidence excluding the possibility that the lepton asymmetry in the Universe is almost as large as the present entropy density. On the contrary, the current measurements of the light element abundances may prefer such an asymmetry to reconcile BBN theory with observations. In this paper, we have made a simple observation which seems quite surprising to us: If the asymmetries in electron number and muon number are equal and opposite and of the size indicated by nucleosynthesis considerations, a baryon asymmetry of the observed size is naturally generated within the Standard Model itself due to the small but non-zero muon Yukawa coupling. This might just be a coincidence, but it is quite an intriguing one!
## Acknowledgements
AR thanks M. Shaposhnikov for useful conversations. JMR and HM thank the Aspen Center for Physics where a part of this work was done. HM also thanks the Institute for Nuclear Theory at the University of Washington for its hospitality and the Department of Energy for partial support during the completion of this work and Jose Valle and Raymond Volkas for useful discussions. HM was supported in part by the U.S. Department of Energy under Contracts DE-AC03-76SF00098, in part by the National Science Foundation under grant PHY-95-14797. Both JMR and HM were supported by the Alfred P. Sloan Foundation.
|
no-problem/9908/astro-ph9908209.html
|
ar5iv
|
text
|
# 1 Introduction:
## 1 Introduction:
For many years the X-ray emission of shell-type (as opposed to plerionic) supernova remnants was modeled in terms of only the emission of hot thermal plasmas. However, it has recently become clear that at least some young shell-type supernova remnants (SN 1006, Koyama et al. 1995, Allen et al. 1999; Cas A, Allen et al. 1997; G347.3$``$0.5, Koyama et al. 1997; IC 443, Keohane et al. 1997) produce non-thermal emission as well. At least for SN 1006 and Cas A, the only plausible description of the non-thermal X-ray emission is synchrotron emission by 10–100 TeV electrons (Allen et al. 1999; Allen et al. 1997). The non-thermal X-ray spectra of these two remnants are qualitatively consistent with simple models of the radio-to-X-ray synchrotron spectra, but are inconsistent with the predicted shapes and flux levels of other processes such as bremsstrahlung emission or inverse Compton scattering of the cosmic microwave background radiation (fig. 1). Furthermore, the high-energy (i.e. non-thermal–dominated) X-ray images differ substantially from lower-energy (thermal-dominated) X-ray images (Willingale et al. 1996; Holt et al. 1994; Vink et al. 1999), but are similar to the radio synchrotron images of the remnants (Reynolds and Gilmore 1986; Anderson and Rudnick 1995). These two clues provide very-strong support for the idea that the non-thermal X-ray emission of SN 1006 and Cas A is synchrotron radiation. The presence of very-high–energy electrons in SN 1006 is corroborated by the detection of TeV gamma-ray emission from this remnant (Tanimori et al. 1998)
These results have very important implications for the study of Galactic cosmic-ray acceleration. Galactic cosmic rays, up to an energy of about 3000 TeV (the “knee”), are thought to be predominantly accelerated in the shocks of supernova remnants. Since the shock-acceleration process of supernova remnants depends on the magnetic rigidity of the particles, it may be the case that the cosmic-ray particles at 3000 TeV are principally iron and that protons (and electrons) are only accelerated to energies of about 100 TeV (Lagage and Cesarsky 1983). However, it has been difficult to confirm that supernova remnants accelerate particles to such high energies. Although previous radio and gamma-ray observations of remnants reveal evidence of non-thermal particles in many supernova remnants, these results are limited to particle energies ($`<`$ $``$ 10 GeV) that are well below 100 TeV (except for the TeV gamma-ray results of Tanimori et al. 1998). The non-thermal X-ray data of SN 1006 provided the first evidence that supernova remnants accelerate particles to very-high energies (Koyama et al. 1995; Reynolds 1996). Subsequent observations reveal that such X-ray emission is not unique to SN 1006 (Allen et al. 1997; Koyama et al. 1997; Keohane et al. 1997). Simple models of the radio-to-X-ray synchrotron spectra of supernova remnants (Reynolds 1998) have been used to show that the electron spectra of SN 1006 and Cas A have exponential cut offs with e-folding energies $`ϵ20`$ and 5 TeV, respectively. Therefore, the study of the non-thermal X-ray emission of young shell-type supernova remnants provides a powerful means of studying the acceleration of the highest-energy cosmic-ray electrons in situ.
## 2 Data and Analyses:
Several supernova remnants have been observed using the Proportional Counter Array (PCA) on the Rossi X-Ray Timing Explorer satellite. The PCA is a spectrophotometer comprised of an array of five co-aligned proportional counter units that are mechanically collimated to have a field-of-view of $`1\text{}`$ FWHM (Jahoda et al. 1996). We have analyzed PCA data for the five, young, shell-type remnants Cas A, Kepler, Tycho, SN 1006, and RCW 86. These data were screened to exclude the time intervals during which (1) one or more of the five proportional counter units was off, (2) the elevation of the source above the limb of the Earth $`<10^{}`$, (3) the PCA background model is not well behaved, and (4) the nominal pointing direction of the PCA $`>0.^{}02`$ from the specified direction of the source. After applying these selection criteria, 169, 44, 81, 18, and 32 ks of data were used to construct the X-ray spectra of Cas A, Kepler, Tycho, SN 1006, and RCW 86, respectively.
The results of the spectral analyses are shown in figure 2. The spectral data below 10 keV for Cas A, Kepler, Tycho, and RCW 86 and below 7 keV for SN 1006 are excluded to insure that the spectra of the remnants are dominated by non-thermal emission.
## 3 Discussion and Conclusion:
At energies above 10 keV, non-thermal X-ray emission dominates the spectra of both SN 1006 and Cas A. Both of these spectra can be described by power laws with photon indices of $`\mathrm{\Gamma }=3.0\pm 0.2`$ (Allen et al. 1997; Allen et al. 1999). As shown in figure 2, similar results are obtained for the remnants Kepler ($`\mathrm{\Gamma }=3.0\pm 0.2`$), Tycho ($`\mathrm{\Gamma }=3.2\pm 0.1`$), and RCW 86 ($`\mathrm{\Gamma }=3.3\pm 0.2`$). Therefore, the high-energy non-thermal X-ray spectra of the five remnants may be produced by a common emission mechanism. Since the high-energy non-thermal X-ray spectra of SN 1006 and Cas A are produced by synchrotron radiation from 10–100 TeV electrons, the results shown in figure 2 support the conclusion that all young shell-type supernova remnants accelerate electrons to very–high-energies. If more detailed analyses of the X-ray emission of Kepler, Tycho, RCW 86, and other young shell-type supernova remnants confirm this conclusion, the results will have very important implications for the origins of Galactic cosmic rays.
The energy at which the radio-to-X-ray synchrotron spectra roll over (see fig. 1) can be used to determine the exponential cut-off energy of the electron spectra of the remnants. For example, if the magnetic field strength of SN 1006 $`10`$ $`\mu `$G (Tanimori et al. 1998) and the energy of the roll off in the synchrotron spectrum $`100`$ eV (fig. 1), the e-folding energy of the cut off in the electron spectrum of SN 1006 $`ϵ20`$ TeV. Estimates of the e-folding energies of the cut offs in the electron spectra of the other remnants yield similar results ($`ϵ5`$, 20, 10, and 10 TeV for Cas A, Kepler, Tycho, and RCW 86, respectively). Since these results are sensitive to the assumed shape of the electron spectrum and the strength of the magnetic field, they should be regarded as order-of-magnitude estimates. Nevertheless, it is interesting that all five of the estimates are below 100 TeV. If the estimates are accurate, the results may indicate that the cosmic-ray electrons in the five remnants have not yet reached their maximum energy, that these remnants do not accelerate cosmic-ray electrons to energies above $`10`$ TeV, or that the maximum energy of the electrons (but not the nuclei) is regulated by radiative losses. More work is needed to differentiate between these possibilities.
Two other clues support the idea that Galactic cosmic rays are accelerated in the shocks of supernova remnants. One clue is the typical electron spectral index of the young remnants of figure 2. A review of radio spectral indices of the five remnants ($`\alpha =0.77`$, 0.64, 0.61, 0.57, and 0.6 for Cas A, Kepler, Tycho, SN 1006, and RCW 86, respectively, Green 1998) reveals that the typical differential spectral index of the electrons producing the radio spectra is about $`\mathrm{\Gamma }=2.2`$ ($`=2\alpha +1`$). This index is consistent with the index expected for cosmic-ray accelerators because the observed differential spectral index of the cosmic-ray protons observed at Earth ($`\mathrm{\Gamma }=2.8`$, Asakimori et al. 1998) less the inferred spectral steepening due to an energy-dependent escape of the cosmic rays from the Galaxy ($`\mathrm{\Delta }\mathrm{\Gamma }=0.6`$, Swordy et al. 1990) is about 2.2.
The second clue also stems from an analysis of the radio spectral data. The radio data of a remnant may be used to estimate the total energy of the cosmic-ray particles in the remnant. Such estimates depend on some assumptions about the strengths of the magnetic fields and about the cosmic-ray electron and proton spectra of the remnants. For simplicity, we assume (1) that the magnetic fields of the remnants are $`B=10^3`$, $`10^2`$, $`10^2`$, $`10`$, and $`10^2`$, for Cas A, Kepler, Tycho, SN 1006, and RCW 86, respectively, (2) that the cosmic-ray electron and proton spectra are described by $`dN_{\mathrm{e},\mathrm{p}}/dE(E+m_{\mathrm{e},\mathrm{p}}c^2)(E^2+2m_{\mathrm{e},\mathrm{p}}c^2E)^{(\mathrm{\Gamma }+1)/2}e^{E/ϵ}`$ (Bell 1978), (3) that the high-energy spectral indices ($`\mathrm{\Gamma }`$ ($`=2\alpha +1`$) of the electrons and protons are the same, (4) that the e-folding energy of the exponential cut offs $`ϵ=10`$ TeV, and (5) that non-thermal electrons outnumber non-thermal protons by a factor of 1.2. The last assumption follows from the assumptions that the relative elemental abundances of the cosmic rays are comparable to the relative elemental abundances of the solar system and that all hydrogen and helium nuclei are fully ionized. This set of assumptions naturally leads to the result that cosmic-ray protons outnumber cosmic-ray electrons by a factor $`100`$ at 1 GeV, as is observed at Earth. Using these assumptions, we find that, at the present, the total energies of the cosmic-ray particles in the remnants $`U_{\mathrm{cr}}5`$, 2, 1, 2, and $`1\times 10^{49}`$ erg for Cas A, Kepler, Tycho, SN 1006, and RCW 86, respectively. These order-of-magnitude estimates are comparable to estimates of the total energy that is needed, on average, per remnant, over their lifetimes ($`3`$$`10\times 10^{49}`$ erg, Blandford and Eichler 1987; Lingenfelter 1992).
In summary, analyses of the high-energy X-ray spectra of five young supernova remnants suggest that the remnants have similar non-thermal spectral properties at energies $`>\text{ }10`$ keV. These spectra are qualitatively consistent with models of the radio-to-X-ray synchrotron spectra of supernova remnants. If this emission is produced by synchrotron radiation, the radio-to-X-ray synchrotron spectra imply that the electron spectra have differential spectral indices of about 2.2 and exponential cut offs at energies $`10`$ TeV. If the remnants also accelerate cosmic-ray nuclei, the total energies of the cosmic rays in the remnants are estimated to be $`1`$$`5\times 10^{49}`$ erg. These results appear to be consistent with the idea that Galactic cosmic-ray electrons (and, presumably, nuclei) are predominantly accelerated in the shocks of supernova remnants.
References
Allen, G. E. et al. 1997, ApJ, 487, L97
Allen, G. E. et al. 1999, In preparation
Anderson, M. C., & Rudnick, L. 1995, ApJ,
441, 307
Asakimori, K. et al. 1998, ApJ, 502, 278
Bell, A. R. 1978, MNRAS, 182, 443
Blandford, R. & Eichler, D. 1987, Phys.
Rep., 154, 1
Holt, S. S., et al. 1994, PASJ, 46, L151
Jahoda, K., et al. 1996, in EUV, X-ray and
Gamma-ray Instrumentation for Space
Astronomy VII, ed. Siegmund, O. H.
W, & Grummin, M. A. Proc. SPIE,
2808, 59
Keohane, J. W. et al. 1997, ApJ 484, 350
Koyama, K. et al. 1995, Nature, 378, 255
Koyama, K. et al. 1997, PASJ, 49, L7
Lagage, P. O., & Cesarsky, C. J. 1983, A&A,
125, 249
Lingenfelter, R. E. 1992, In The Astronomy
and Astrophysics Encyclopedia, Edited
by Maran, S. 139, Van Nostrand Rein-
hold Publishers
Reynolds, S. P. 1996, ApJ, 459, L13
Reynolds, S. P. 1998, ApJ, 493, 375
Reynolds, S. P., & Gilmore, D. M. 1986, AJ,
92, 1138
Swordy, S. P. et al. 1990, ApJ, 349, 625
Tanimori, T. et al. 1998, ApJ, 497, L25
The, L.-S., et al. 1996, A&ApS, 120, 357
Vink, J. et al. 1999, A&A, In press
Willingale, R., et al. 1996, MNRAS, 278, 749
|
no-problem/9908/hep-th9908192.html
|
ar5iv
|
text
|
# Untitled Document
hep-th/9908192 TIFR/TH/99-44
Killing Spinors and Supersymmetric AdS Orbifolds
Bahniman Ghosh<sup>1</sup> E-mail: [email protected] and Sunil Mukhi<sup>2</sup> E-mail: [email protected]
Tata Institute of Fundamental Research,
Homi Bhabha Rd, Mumbai 400 005, India
ABSTRACT
We examine the behaviour of Killing spinors on $`AdS_5`$ under various discrete symmetries of the spacetime. In this way we discover a number of supersymmetric orbifolds, reproducing the known ones and adding a few novel ones to the list. These orbifolds break the $`SO(4,2)`$ invariance of $`AdS_5`$ down to subgroups. We also make some comments on the non-compact Stiefel manifold $`W_{4,2}`$.
August 1999
1. Introduction
Supersymmetric compactifications of type IIB string theory on spacetimes of the form $`AdS_5\times X_5`$ have yielded a number of interesting results about superconformal gauge theory in four dimensions following the discovery\[1\] of the AdS/CFT correspondence. Here the compact manifold $`X_5`$ can be the sphere $`S^5`$, or one of many possible “non-spherical horizons” including spherical orbifolds and other Einstein spaces like the conifold base.
The presence of an $`AdS_5`$ factor guarantees conformal invariance of the dual field theory, while varying the $`X_5`$ affects the spectrum of the field theory and in particular the total number of supersymmetries. Recently a number of situations have been discussed where instead one keeps $`X_5`$ fixed (for example, to be $`S^5`$) and chooses different noncompact Einstein spaces in lieu of $`AdS_5`$. Some examples of this are the five-dimensional Stiefel manifold $`W_{4,2}`$\[2\], spaces with a nontrivial $`H=dB`$\[3,,4\] and orbifolds of $`AdS_p`$\[5,,6,,7,,8\]. The physical interpretation of all these spaces is not completely clear at present – for example, some of the spaces discussed in \[2,,3,,4\] have singular behaviour at infinity leading to boundaries of dimension less than 4.
Orbifolds of $`S^5`$\[9\] are interesting because they allow us to design spacetimes that are dual to a wide class of conformally invariant, supersymmetric field theories in 4 dimensions. Some orbifolds of $`AdS_5`$ have been interpreted as topological black holes\[5\] generalizing the famous BTZ black hole in 3 dimensions\[10\], while other orbifolds represent cosmological solutions\[6\]<sup>1</sup> Much of the previous work on $`AdS`$ orbifolds deals with non-supersymmetric cases, and hence our discussion below will not be closely related to it.. A particular supersymmetric $`AdS_5`$ orbifold was discussed in Ref.\[8\] where it was proposed to be dual to a 3-brane field theory with a $`pp`$-wave propagating on it. Hence in this example one finds a type IIB supergravity background that is dual to a 3-brane worldvolume theory, not in its ground state but in a BPS excited state. This is an intriguing direction in which to generalize the AdS/CFT correspondence.
The purpose of this note is to examine conditions under which orbifolds of $`AdS_5`$ (with or without fixed points) preserve some supersymmetry. The analogous conditions for $`S^5`$ have been analyzed in some depth in Refs.\[11,,12\]. One key result that helped in that classification was a theorem relating Killing spinors on Einstein 5-manifolds to parallel spinors on a 6d cone above them. However, one could also reproduce many of those results by directly studying the transformation properties of Killing spinors on $`S^5`$ under the orbifolding action.
In what follows, we construct Killing spinors on $`AdS_5`$ in three different coordinate systems and examine their behaviour under various possible orbifolding actions. This enables us to construct a number of supersymmetric orbifolds, including some known ones and some that are apparently new. As we will see, different orbifolds can be conveniently studied in different coordinates systems on $`AdS_5`$. A complete classification of orbifolds of $`AdS_5`$, on the lines of Refs.\[11,,12\], would be interesting to attempt. This would perhaps follow if one could prove a theorem relating Killing spinors on a (non-compact) 5-dimensional Einstein space to spinors on a “cone” over it with two timelike directions.
We also compute Killing spinors for $`W_{4,2}`$ and discuss some supersymmetric orbifolds of this space. The physical meaning of this space and its relevance to the AdS/CFT correspondence are not very clear, and the same holds for the $`AdS`$ orbifolds we consider except in a few cases. We leave the detailed analysis of this question, along with the study of the global structure of these orbifold spacetimes, for the future.
Like the cases discussed in Refs.\[8,,2\], the orbifolds discussed here break the $`SO(4,2)`$ invariance of $`AdS_5`$ down to subgroups, while preserving the $`S^5`$ factor and hence the $`SO(6)`$ symmetry associated to R-symmetry of the boundary CFT. One can of course combine the orbifolds discussed here with the ones proposed in Ref.\[9\] to get compactifications with still lower symmetry and supersymmetry.
2. Killing Spinors on $`AdS^5`$
$`AdS^5`$ spacetime can be described as a hyperboloid in a 6-dimensional spacetime with 2 timelike directions. Labelling the coordinates of this ambient spacetime as $`X_1,X_0,X_1,\mathrm{}X_4`$, the metric is
$$ds^2=(dX_1)^2(dX_0)^2+(dX_1)^2+\mathrm{}+(dX_4)^2$$
and the equation of the hyperboloid is:
$$1=(X_1)^2(X_0)^2+(X_1)^2+(X_2)^2+(X_3)^2+(X_4)^2$$
The metric on $`AdS^5`$ is the one induced from the ambient space.
We will find it convenient to work in three different sets of coordinates.
2.1. Light-Cone Type Coordinates
These consist of two pairs of lightlike coordinates and one complex coordinate. It is defined by
$$z_1^\pm =X_0\pm X_1,z_2^\pm =X_2\pm X_1,w=(X_3+iX_4)$$
and the hyperboloid is
$$1=z_1^+z_1^{}+z_2^+z_2^{}+w\overline{w}$$
For this set of coordinates, it is convenient to choose an explicit basis for the Gamma-matrices as:
$$\mathrm{\Gamma }_1=\sigma _21,\mathrm{\Gamma }_2=i\sigma _3\sigma _1,\mathrm{\Gamma }_3=\sigma _3\sigma _2,\mathrm{\Gamma }_4=\sigma _3\sigma _3,\mathrm{\Gamma }_5=\sigma _11$$
Note that $`(\mathrm{\Gamma }_2)^2=1`$, while the other $`\mathrm{\Gamma }`$-matrices square to $`+1`$.
From the light-cone type coordinates we go to a set of five independent coordinates $`\theta _1,\theta _2,\alpha ,\beta ,\delta `$ where $`0\theta _2\pi `$, $`0\beta 2\pi `$ and $`\alpha `$,$`\delta `$ and $`\theta _1`$ are non-compact. These coordinates are defined by:
$$\begin{array}{cc}\hfill z_1^\pm & =\mathrm{cosh}\frac{\theta _1}{2}e^{\pm \delta }\hfill \\ \hfill z_2^\pm & =\mathrm{sinh}\frac{\theta _1}{2}\mathrm{cos}\frac{\theta _2}{2}e^{\pm \alpha }\hfill \\ \hfill w& =\mathrm{sinh}\frac{\theta _1}{2}\mathrm{sin}\frac{\theta _2}{2}e^{i\beta }\hfill \end{array}$$
By abuse of notation we will refer to these also as light-cone type coordinates, though they are actually an angular parametrization of those coordinates which solves the hyperboloid constraint. The metric on $`AdS^5`$ in these coordinates is:
$$ds^2=\mathrm{sinh}^2\frac{\theta _1}{2}\mathrm{cos}^2\frac{\theta _2}{2}d\alpha ^2+\mathrm{cosh}^2\frac{\theta _1}{2}d\delta ^2+\frac{1}{4}d\theta _1^2+\mathrm{sinh}^2\frac{\theta _1}{2}(\frac{1}{4}d\theta _2^2+\mathrm{sin}^2\frac{\theta _2}{2}d\beta ^2)$$
For fixed $`\theta _2`$ and $`\beta `$, the metric is proportional to that of $`AdS^3`$.
For this case, we have the vielbeins:
$$\begin{array}{cc}& e^{\underset{¯}{1}}=\frac{1}{2}\mathrm{sinh}\frac{\theta _1}{2}d\theta _2e^{\underset{¯}{2}}=\mathrm{sinh}\frac{\theta _1}{2}\mathrm{cos}\frac{\theta _2}{2}d\alpha \hfill \\ & e^{\underset{¯}{3}}=\mathrm{cosh}\frac{\theta _1}{2}d\delta e^{\underset{¯}{4}}=\frac{1}{2}d\theta _1e^{\underset{¯}{5}}=\mathrm{sinh}\frac{\theta _1}{2}\mathrm{sin}\frac{\theta _2}{2}d\beta \hfill \end{array}$$
and the spin connections:
$$\begin{array}{cc}& \omega ^{\underset{¯}{1}\underset{¯}{2}}=\mathrm{sin}\frac{\theta _2}{2}d\alpha \omega ^{\underset{¯}{1}\underset{¯}{4}}=\frac{1}{2}\mathrm{cosh}\frac{\theta _1}{2}d\theta _2\omega ^{\underset{¯}{1}\underset{¯}{5}}=\mathrm{cos}\frac{\theta _2}{2}d\beta \hfill \\ & \omega ^{\underset{¯}{2}\underset{¯}{4}}=\mathrm{cos}\frac{\theta _2}{2}\mathrm{cosh}\frac{\theta _1}{2}d\alpha \omega ^{\underset{¯}{3}\underset{¯}{4}}=\mathrm{sinh}\frac{\theta _1}{2}d\delta \omega ^{\underset{¯}{4}\underset{¯}{5}}=\mathrm{sin}\frac{\theta _2}{2}\mathrm{cosh}\frac{\theta _1}{2}d\beta \hfill \end{array}$$
Note that the tangent-space metric has $`\eta _{\underset{¯}{2}\underset{¯}{2}}=1`$ while the other components are $`+1`$.
We are interested in studying the Killing spinors on $`AdS^5`$ in this coordinate basis. The relevant equation is:
$$(_\mu +\frac{1}{4}\omega _\mu ^{\underset{¯}{a}\underset{¯}{b}}\mathrm{\Gamma }_{\underset{¯}{a}\underset{¯}{b}}\frac{1}{2}e_\mu ^{\underset{¯}{a}}\mathrm{\Gamma }_{\underset{¯}{a}})ϵ=0$$
It is fairly straightforward to compute the solutions to this equation, which are given by:
$$ϵ=e^{\frac{1}{4}\mathrm{\Gamma }_4\theta _1}e^{\frac{1}{4}\mathrm{\Gamma }_{14}\theta _2}e^{\frac{1}{2}\mathrm{\Gamma }_{24}\alpha }e^{\frac{1}{2}\mathrm{\Gamma }_3\delta }e^{\frac{1}{2}\mathrm{\Gamma }_{15}\beta }ϵ_0$$
where $`ϵ_0`$ is an arbitrary constant spinor.
2.2. Complex Coordinates
Another coordinate system will turn out to be useful to investigate a different class of orbifolds. These will be called complex coordinates – they are actually complex coordinates of the ambient 6-dimensional spacetime, a complex time and two complex space dimensions. Thus we define:
$$u=X_1+iX_0,v=X_1+iX_2,w=(X_3+iX_4)$$
in terms of which the hyperboloid is
$$1=u\overline{u}+v\overline{v}+w\overline{w}$$
The coordinate $`w`$ is the same as was used for the light-cone type coordinates. This time it is convenient to go to five independent coordinates $`\theta _1,\theta _2,\alpha ^{},\beta ,\delta ^{}`$ where $`0\theta _2\pi `$, $`0\beta ,\alpha ^{},\delta ^{}2\pi `$ and $`\theta _1`$ is non-compact. These coordinates are defined by:
$$\begin{array}{cc}\hfill u& =\mathrm{cosh}\frac{\theta _1}{2}e^{i\delta ^{}}\hfill \\ \hfill v& =\mathrm{sinh}\frac{\theta _1}{2}\mathrm{cos}\frac{\theta _2}{2}e^{i\alpha ^{}}\hfill \\ \hfill w& =\mathrm{sinh}\frac{\theta _1}{2}\mathrm{sin}\frac{\theta _2}{2}e^{i\beta }\hfill \end{array}$$
Again, although these angles parametrize the complex coordinates in a way which solves the hyperboloid constraint, we will refer to the angles themselves as complex coordinates.
Note that these coordinates can be obtained from the previous ones by the formal replacement $`\alpha =i\alpha ^{}`$ and $`\delta =i\delta ^{}`$, which interchanges one space with one time direction. This replacement in Eq.(2.1) also gives us the metric in these coordinates. It is evident that the vielbeins are formally the same as before, though the tangent-space metric now has $`\eta _{\underset{¯}{3}\underset{¯}{3}}=1`$. Careful inspection shows that the spin connections also turn out to be exactly the same as in Eq.(2.1), as the sign changes introduced by the interchange of a space and a time direction eventually cancel out. The change of tangent space metric necessitates a slightly different basis of $`\mathrm{\Gamma }`$-matrices. We multiply $`\mathrm{\Gamma }_2`$ of Eq.(2.1) by $`i`$ and $`\mathrm{\Gamma }_3`$ by $`i`$. Hence the new set of $`\mathrm{\Gamma }`$-matrices (which we label $`\mathrm{\Gamma }^{}`$ to avoid confusion with the previous set) becomes:
$$\mathrm{\Gamma }_1^{}=\sigma _21,\mathrm{\Gamma }_2^{}=\sigma _3\sigma _1,\mathrm{\Gamma }_3^{}=i\sigma _3\sigma _2,\mathrm{\Gamma }_4^{}=\sigma _3\sigma _3,\mathrm{\Gamma }_5^{}=\sigma _11$$
This time, $`(\mathrm{\Gamma }_3^{})^2=1`$ while the others square to $`+1`$. The advantage of this choice is that we find (formally) the same Killing spinor as in Eq.(2.1), but now with the $`\mathrm{\Gamma }^{}`$-matrices:
$$ϵ=e^{\frac{1}{4}\mathrm{\Gamma }_4^{}\theta _1}e^{\frac{1}{4}\mathrm{\Gamma }_{14}^{}\theta _2}e^{\frac{1}{2}\mathrm{\Gamma }_{24}^{}\alpha }e^{\frac{1}{2}\mathrm{\Gamma }_3^{}\delta }e^{\frac{1}{2}\mathrm{\Gamma }_{15}^{}\beta }ϵ_0$$
2.3. Horospherical Coordinates
Let us finally recall\[13\] the Killing spinors in horospherical coordinates, which consist of five independent real coordinates, $`r,x^1,x^2,x^3,x^4`$ in terms of which the metric is:
$$ds^2=(dr)^2+e^{2r}((dx^1)^2+(dx^2)^2+(dx^3)^2+(dx^4)^2)$$
Choosing the $`\mathrm{\Gamma }`$ matrices as
$$\mathrm{\Gamma }_1=i\sigma _3\sigma _2,\mathrm{\Gamma }_2=\sigma _3\sigma _1,\mathrm{\Gamma }_r=\sigma _3\sigma _3,\mathrm{\Gamma }_3=\sigma _11,\mathrm{\Gamma }_4=\sigma _21$$
the Killing spinor is found to be\[13\]:
$$\begin{array}{cc}\hfill ϵ& =e^{\frac{1}{2}r\mathrm{\Gamma }_r}(1+\frac{1}{2}x^\alpha \mathrm{\Gamma }_\alpha (1\mathrm{\Gamma }_r))ϵ_0\hfill \\ & =\left(\begin{array}{c}e^{\frac{r}{2}}(ϵ_0^{(1)}+x^+ϵ_0^{(2)}+(x^3ix_4)ϵ_0^{(3)})\\ \\ e^{\frac{r}{2}}ϵ_0^{(2)}\\ \\ e^{\frac{r}{2}}ϵ_0^{(3)}\\ \\ e^{\frac{r}{2}}(ϵ_0^{(4)}+x^{}ϵ_0^{(3)}+(x^3+ix^4)ϵ_0^{(2)})\end{array}\right)\hfill \end{array}$$
where $`\alpha =1,2,3,4`$, $`x^+=x^1+x^2`$, $`x^{}=x^1x^2`$ and
$$ϵ_0=\left(\begin{array}{c}ϵ_0^{(1)}\\ ϵ_0^{(2)}\\ ϵ_0^{(3)}\\ ϵ_0^{(4)}\end{array}\right)$$
The transformation between these horospherical coordinates and the light-cone type coordinates $`z_1^\pm ,z_2^\pm ,w`$ is: is
$$e^r=z_2^+,x^+=\frac{z_1^+}{z_2^+},x^{}=\frac{z_1^{}}{z_2^+},x^3=\frac{(w+\overline{w})}{2z_2^+},x^4=\frac{(w\overline{w})}{2iz_2^+},$$
3. Orbifolds of $`AdS_5`$
3.1. Half-supersymmetric Orbifolds
Now we will examine the orbifolding actions which are natural in the various coordinates. In the light-cone type coordinates, one natural action follows from the following transformation:\[8\]:
$$z_1^\pm e^{\pm 2\pi /k}z_1^\pm ,z_2^\pm e^{\pm 2\pi /k}z_2^\pm $$
which can be expressed as a simple translation:
$$\delta \delta +\frac{2\pi }{k},\alpha \alpha +\frac{2\pi }{k}$$
From Eq.(2.1), this clearly has no fixed points, since the hyperboloid does not pass through $`z_1^\pm =z_2^\pm =0`$. In order for the Killing spinor $`ϵ`$ to be invariant under the above transformation we must require that the constant spinor $`ϵ_0`$ satisfy
$$e^{\frac{\pi }{k}(\mathrm{\Gamma }_{24}+\mathrm{\Gamma }_3)}ϵ_0=ϵ_0$$
which means the matrix $`(\mathrm{\Gamma }_{24}+\mathrm{\Gamma }_3)`$ must annihilate $`ϵ_0`$. In the basis chosen in Eq.(2.1), we have:
$$\mathrm{\Gamma }_{24}+\mathrm{\Gamma }_3=2\left(\begin{array}{cc}0& 0\\ 0& \sigma _2\end{array}\right)$$
and hence the Killing spinors that are preserved, in this basis, are the ones for which
$$ϵ_0=\left(\begin{array}{c}ϵ_0^{(1)}\\ ϵ_0^{(2)}\\ 0\\ 0\end{array}\right)$$
This in particular gives a direct proof that the orbifold discussed in Ref.\[8\] preserves half the supersymmetries.
The above orbifold action is generated by an $`SU(1,1)`$ matrix in the full isometry group $`SO(4,2)`$ of $`AdS_5`$, hence the surviving symmetry group is the commutant of $`SU(1,1)`$ in $`SO(4,2)`$ which is $`SU(1,1)\times U(1)`$. This is analogous to the fact that the simplest half-supersymmetric orbifold of $`S^5`$ (corresponding to D3-branes at an ALE singularity) has an R-symmetry group $`SU(2)\times U(1)`$.
Turning now to the complex coordinates, it is natural to consider orbifold actions of the type
$$u\gamma ^du,v\gamma ^av,w\gamma ^bw$$
where $`\gamma =\mathrm{exp}(2\pi i/k)`$ and $`a,b,d`$ are some integers. These are quite analogous to corresponding orbifolds of $`S^5`$. The result is also analogous: the orbifolding action above leaves the Killing spinor invariant if
$$\left(a\mathrm{\Gamma }_{24}^{}+d\mathrm{\Gamma }_3^{}+b\mathrm{\Gamma }_{15}^{}\right)ϵ_0=0$$
The above matrix has eigenvalues $`(a+bd)`$, $`(a+b+d)`$, $`(ab+d)`$ and $`(abd)`$. If one of $`a,b,d`$ is zero then we have two vanishing eigenvalues and $`\frac{1}{2}`$-supersymmetry.
Note, however, that because of the signature of the spacetime, all the $`\frac{1}{2}`$-supersymmetric cases are not equivalent. The case with $`d=0`$ has a circle of fixed points $`u\overline{u}=1`$, while the cases with $`a=0`$ or $`b=0`$ have no fixed points and are equivalent to each other.
In the case $`d=0`$, the orbifold generator lies in an $`SU(2)`$ subgroup of $`SO(4)SO(4,2)`$ hence the symmetry of the quotient space is $`SU(2)\times U(1)`$. On the other hand, for $`a=0`$ or $`b=0`$ the orbifold is generated by an element in an $`SU(1,1)`$ subgroup of $`SO(2,2)SO(4,2)`$ and the surviving symmetry is $`SU(1,1)\times U(1)`$.
Next, it is useful to examine the orbifold described in Eq.(3.1), in horospherical coordinates. The action becomes:
$$rr+a,x^+x^+,x^{}e^{2a}x^{},x^3e^ax^3,x^4e^ax^4,$$
Hence the Killing spinor transforms as
$$ϵe^{\frac{1}{2}a(1\sigma _3)}\left(\begin{array}{c}e^{\frac{(r+a)}{2}}(ϵ_0^{(1)}+x^+ϵ_0^{(2)}+e^a(x^3ix^4)ϵ_0^{(3)})\\ \\ e^{\frac{(r+a)}{2}}ϵ_0^{(2)}\\ \\ e^{\frac{(r+a)}{2}}ϵ_0^{(3)}\\ \\ e^{\frac{(r+a)}{2}}(ϵ_0^{(4)}+e^{2a}x^{}ϵ_0^{(3)}+e^a(x^3+ix^4)ϵ_0^{(2)})\end{array}\right)$$
Thus the Killing spinors that are preserved by this orbifold, in this basis, are the ones for which
$$ϵ_0=\left(\begin{array}{c}ϵ_0^{(1)}\\ ϵ_0^{(2)}\\ 0\\ 0\end{array}\right)$$
Another apparently trivial kind of $`\frac{1}{2}`$-supersymmetric orbifold is apparent from Eqn.(2.1). Suppose we choose $`ϵ_0^{(2)}=ϵ_0^{(3)}=0`$. Then the Killing spinor becomes independent of $`x^\pm ,x^3,x^4`$. As a result, periodic identifications in these coordinates preserve the Killing spinor. This is essentially what was noted in Ref., and corresponds to the fact that the identification of these coordinates breaks conformal invariance by introducing a scale, hence the conformal part of the superconformal invariance goes away. Thus such orbifolds preserve half the supersymmetries. (One can further deform the space in the $`x^3,x^4`$ directions and add a nontrivial $`B`$-field, preserving the remaining supersymmetry, as was done in Ref.\[3,,4\]. In this case one does not expect the deformed manifold to have a Killing spinor, since the field strength $`dB`$ also contributes to the supersymmetry variation.)
3.2. One-fourth Supersymmetry
We have already considered orbifold actions, in complex coordinates, of the general type
$$u\gamma ^du,v\gamma ^av,w\gamma ^bw$$
where $`\gamma =\mathrm{exp}(2\pi i/k)`$ and $`a,b,d`$ are some integers. We saw that the relevant matrix acting on Killing spinors has eigenvalues $`(a+bd)`$, $`(a+b+d)`$, $`(ab+d)`$ and $`(abd)`$. Hence if all of $`a,b,d`$ are nonzero then at most one of these eigenvalues can be zero and in that case we have a $`\frac{1}{4}`$-supersymmetric orbifold. For the $`\frac{1}{4}`$ supersymmetric cases we have no fixed points for the orbifold action.
The orbifold generator lies in an $`SU(2,1)`$ subgroup of $`SO(4,2)`$, hence it preserves only a $`U(1)`$ symmetry. This is the analogue of the $`U(1)`$ symmetry preserved by $`\frac{1}{4}`$-supersymmetric orbifolds of $`S^5`$, which is realized as a $`U(1)`$ R-symmetry in the boundary theory.
Another class of $`\frac{1}{4}`$-supersymmetric $`AdS_5`$ orbifolds comes from quotienting by a pair of transformations each of which preserves half the supersymmetry. If $`(k,k^{})`$ are co-prime there are two inequivalent cases:
$$\begin{array}{cc}\hfill u\gamma u,& v\gamma ^1v\hfill \\ \hfill v\gamma ^{}v,& w\gamma ^1w\hfill \end{array}$$
and
$$\begin{array}{cc}\hfill u\gamma u,& v\gamma ^1v\hfill \\ \hfill u\gamma ^{}u,& w\gamma ^1w\hfill \end{array}$$
where $`\gamma ^k=(\gamma ^{})^k^{}=1`$.
Another interesting $`\frac{1}{4}`$-supersymmetric orbifold arises by combining the periodic identification in $`x^2,x^3,x^4`$ in the horospherical coordinates, with the orbifolding action of Eq.(3.1). Here the constant Killing spinor satisfies $`ϵ_0^{(2)}=ϵ_0^{(3)}=ϵ_0^{(4)}=0`$.
We have encountered a number of supersymmetric orbifolds, but it turns out that each of them is natural in a certain coordinate system and not so easy to describe in another. Thus, it becomes hard to combine the different actions discussed in the previous section and find more general orbifolds. For example, one of the simplest orbifolds described in complex coordinates in Eq.(3.1) arises by choosing $`k=2,d=a=1,b=0`$. This just corresponds to the reflection $`uu`$, $`vv`$. In terms of the light cone type coordinates this means $`z_i^\pm z_i^\pm `$, which cannot be carried out using the independent coordinates defined in Eq.(2.1), which only cover the region $`z_1^\pm >0`$.
In contrast, the orbifold corresponding to $`k=2,a=b=1,d=0`$ can be realized in the light cone type coordinates. In this case we have $`vv`$, $`ww`$. This corresponds to the action
$$z_1^+z_1^{},z_2^+z_2^{},ww$$
which in terms of the independent coordinates in Eq.(2.1) is just
$$\delta \delta ,\alpha \alpha ,\theta _1\theta _1$$
This acts on the Killing spinor in Eq.(2.1) as follows. In our basis, this Killing spinor is explicitly given by
$$ϵ=\left(\begin{array}{c}\begin{array}{c}\mathrm{cos}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha \delta }{2}ϵ_0^{(1)}+i\mathrm{sinh}\frac{\alpha \delta }{2}ϵ_0^{(2)}\right\}e^{i\frac{\beta }{2}}\\ i\mathrm{sin}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha +\delta }{2}ϵ_0^{(3)}+i\mathrm{sinh}\frac{\alpha +\delta }{2}ϵ_0^{(4)}\right\}e^{i\frac{\beta }{2}}\end{array}\\ \\ \begin{array}{c}\mathrm{cos}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha \delta }{2}ϵ_0^{(2)}i\mathrm{sinh}\frac{\alpha \delta }{2}ϵ_0^{(1)}\right\}e^{i\frac{\beta }{2}}\\ +i\mathrm{sin}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha +\delta }{2}ϵ_0^{(4)}i\mathrm{sinh}\frac{\alpha +\delta }{2}ϵ_0^{(3)}\right\}e^{i\frac{\beta }{2}}\end{array}\\ \\ \begin{array}{c}\mathrm{cos}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha +\delta }{2}ϵ_0^{(3)}+i\mathrm{sinh}\frac{\alpha +\delta }{2}ϵ_0^{(4)}\right\}e^{i\frac{\beta }{2}}\\ i\mathrm{sin}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha \delta }{2}ϵ_0^{(1)}+i\mathrm{sinh}\frac{\alpha \delta }{2}ϵ_0^{(2)}\right\}e^{i\frac{\beta }{2}}\end{array}\\ \\ \begin{array}{c}\mathrm{cos}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha +\delta }{2}ϵ_0^{(4)}i\mathrm{sinh}\frac{\alpha +\delta }{2}ϵ_0^{(3)}\right\}e^{i\frac{\beta }{2}}\\ +i\mathrm{sin}\frac{\theta _2}{4}e^{\frac{\theta _1}{4}}\left\{\mathrm{cosh}\frac{\alpha \delta }{2}ϵ_0^{(2)}i\mathrm{sinh}\frac{\alpha \delta }{2}ϵ_0^{(1)}\right\}e^{i\frac{\beta }{2}}\end{array}\end{array}\right)$$
Now one finds that, setting $`ϵ_0^{(1)}=ϵ_0^{(2)}`$ and $`ϵ_0^{(3)}=ϵ_0^{(4)}`$, the expression for $`ϵ`$ above goes over to $`ϵ^{}`$ satisfying
$$ϵ^{}=\left(\begin{array}{cc}\sigma _1& 0\\ 0& \sigma _1\end{array}\right)ϵ=(\sigma _3\sigma _1)ϵ$$
This coincides with the Lorentz transformation of $`ϵ`$ under the action in Eq.(3.1)<sup>2</sup> The action in Eq.(3.1) inverts the sign of the vielbeins $`e^{\underset{¯}{1}},e^{\underset{¯}{3}},e^{\underset{¯}{4}},e^{\underset{¯}{5}}`$, hence it must be represented on spinors by a matrix $`P`$ which anticommutes with $`\mathrm{\Gamma }_{\underset{¯}{1}},\mathrm{\Gamma }_{\underset{¯}{3}},\mathrm{\Gamma }_{\underset{¯}{4}},\mathrm{\Gamma }_{\underset{¯}{5}}`$. Thus $`P`$ is proportional to $`\mathrm{\Gamma }_2=i\sigma _3\sigma _1`$. Since $`P^2=1`$, it follows that $`P=\sigma _3\sigma _1`$.
Thus we see (as expected from the analysis of the same orbifold in complex coordinates) that this orbifold preserves half the supersymmetries. But now that we know the preserved Killing spinors in the basis appropriate to light-cone coordinates, we can combine this orbifold with the orbifolding along lightlike directions defined in Eq.(3.1) and determine if there is any residual supersymmetry. Indeed we have seen that Eq.(3.1) preserves Killing spinors of the form Eq.(2.1) where the constant spinor $`ϵ_0`$ has the form:
$$ϵ_0=\left(\begin{array}{c}ϵ_0^{(1)}\\ ϵ_0^{(2)}\\ 0\\ 0\end{array}\right)$$
while Eq.(3.1) preserves Killing spinors where $`ϵ_0`$ has the form
$$ϵ_0=\left(\begin{array}{c}ϵ_0^{(1)}\\ ϵ_0^{(1)}\\ ϵ_0^{(3)}\\ ϵ_0^{(3)}\end{array}\right)$$
Choosing now $`ϵ_0^{(1)}=ϵ_0^{(2)}`$ in the first of these equations and $`ϵ_0^{(3)}=0`$ in the second, we find that both the transformations together preserve $`\frac{1}{4}`$ of the supersymmetry, namely the Killing spinor for which
$$ϵ_0=\left(\begin{array}{c}1\\ 1\\ 0\\ 0\end{array}\right)$$
This is a $`Z\times Z_2`$ orbifold of $`AdS_5`$. It is much more difficult, if not impossible, to express the $`Z_k`$ orbifold given in complex coordinates by $`v\gamma v,w\gamma ^1w,\gamma ^k=1`$ in terms of light-cone coordinates. Luckily it is not necessary to do this. These orbifolds preserve the same Killing spinors for all $`k`$. Since we have shown that for $`k=2`$ there is a $`\frac{1}{4}`$-supersymmetric $`Z\times Z_2`$ orbifold obtained by combining with the action Eq.(3.1), it follows that there is also a $`Z\times Z_k`$ orbifold with $`\frac{1}{4}`$-supersymmetry for all $`k`$.
4. The Stiefel Manifold $`W_{4,2}`$
Although not directly related to orbifolds of $`AdS_5`$, the noncompact Stiefel manifold $`W_{4,2}`$\[2\] is an interesting $`\frac{1}{4}`$-supersymmetric coset spacetime. We will speculate later on its possible relation with $`AdS_5`$ orbifolds. In the present section we will compute its single Killing spinor and comment on orbifolds of this spacetime.
This case corresponds to an analytic continuation of the compact manifold variously denoted $`T_{1,1}`$ or $`V_{4,2}`$, which is the base of the conifold geometry and hence appears in the study of D3-branes at conifolds\[14\]. It can be thought of as the coset space $`(AdS_3\times AdS_3)/U(1)`$, and also as a (timelike) $`U(1)`$ fibration above Euclidean $`AdS_2\times AdS_2`$. The natural coordinates for $`W_{4,2}`$ are $`(\theta _1,\varphi _1,\theta _2,\varphi _2,\psi )`$ where $`0\psi <4\pi `$, $`0\varphi _i<2\pi `$ while $`\theta _i`$ are noncompact. In terms of these, the metric of $`W_{4,2}`$ is
$$\begin{array}{cc}\hfill ds^2=& \frac{1}{9}(d\psi +\mathrm{cosh}\theta _1d\varphi _1+\mathrm{cosh}\theta _2d\varphi _2)^2\hfill \\ & +\frac{1}{6}(d\theta _1^2+\mathrm{sinh}^2\theta _1d\varphi _1^2)+\frac{1}{6}(d\theta _2^2+\mathrm{sinh}^2\theta _2d\varphi _2^2)\hfill \end{array}$$
which explicitly exhibits the $`U(1)`$ fibration, with $`\psi `$ being the fibre coordinate.
This spacetime breaks the maximal $`SO(4,2)`$ isometry to $`SO(2,2)\times SO(2)`$, much as its compact version breaks the maximal $`SO(6)`$ isometry of $`S^5`$ to $`SO(4)\times SO(2)`$. From the metric one can read off the vielbeins:
$$\begin{array}{cc}\hfill e^{\underset{¯}{1}}& =\frac{1}{\sqrt{6}}d\theta _1e^{\underset{¯}{2}}=\frac{1}{\sqrt{6}}\mathrm{sinh}\theta _1d\varphi _1\hfill \\ \hfill e^{\underset{¯}{3}}& =\frac{1}{\sqrt{6}}d\theta _2e^{\underset{¯}{4}}=\frac{1}{\sqrt{6}}\mathrm{sinh}\theta _2d\varphi _2\hfill \\ \hfill e^{\underset{¯}{5}}& =\frac{1}{3}(d\psi +\mathrm{cosh}\theta _1d\varphi _1+\mathrm{cosh}\theta _2d\varphi _2)\hfill \end{array}$$
and compute the spin connections:
$$\begin{array}{cc}\hfill \omega ^{\underset{¯}{1}\underset{¯}{2}}& =\frac{2}{3}\mathrm{cosh}\theta _1d\varphi _1+\frac{1}{3}d\psi +\frac{1}{3}\mathrm{cosh}\theta _2d\varphi _2\hfill \\ \hfill \omega ^{\underset{¯}{1}\underset{¯}{3}}& =\omega ^{\underset{¯}{1}\underset{¯}{4}}=\omega ^{\underset{¯}{2}\underset{¯}{3}}=\omega ^{\underset{¯}{2}\underset{¯}{4}}=0\hfill \\ \hfill \omega ^{\underset{¯}{3}\underset{¯}{4}}& =\frac{2}{3}\mathrm{cosh}\theta _2d\varphi _2+\frac{1}{3}d\psi +\frac{1}{3}\mathrm{cosh}\theta _1d\varphi _1\hfill \\ \hfill \omega ^{\underset{¯}{1}\underset{¯}{5}}& =\frac{1}{\sqrt{6}}\mathrm{sinh}\theta _1d\varphi _1\omega ^{\underset{¯}{2}\underset{¯}{5}}=\frac{1}{\sqrt{6}}d\theta _1\hfill \\ \hfill \omega ^{\underset{¯}{3}\underset{¯}{5}}& =\frac{1}{\sqrt{6}}\mathrm{sinh}\theta _2d\varphi _2\omega ^{\underset{¯}{4}\underset{¯}{5}}=\frac{1}{\sqrt{6}}d\theta _2\hfill \end{array}$$
A convenient basis for the $`\mathrm{\Gamma }`$-matrices this time is
$$\mathrm{\Gamma }_1=\sigma _11,\mathrm{\Gamma }_2=\sigma _21,\mathrm{\Gamma }_3=\sigma _3\sigma _1,\mathrm{\Gamma }_4=\sigma _3\sigma _2,\mathrm{\Gamma }_5=i\sigma _3\sigma _3$$
In this basis, the Killing spinor equations become
$$\frac{}{\theta _1}\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)=\left(\begin{array}{cccc}0& 0& \frac{1}{\sqrt{6}}& 0\\ 0& 0& 0& 0\\ \frac{1}{\sqrt{6}}& 0& 0& 0\\ 0& 0& 0& 0\end{array}\right)\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)$$
$$\frac{}{\varphi _1}\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)=\left(\begin{array}{cccc}\frac{i}{3}\mathrm{cosh}\theta _1& 0& \frac{i}{\sqrt{6}}\mathrm{sinh}\theta _1& 0\\ 0& \frac{i}{3}\mathrm{cosh}\theta _1& 0& 0\\ \frac{i}{\sqrt{6}}\mathrm{sinh}\theta _1& 0& \frac{2i}{3}\mathrm{cosh}\theta _1& 0\\ 0& 0& 0& 0\end{array}\right)\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)$$
$$\frac{}{\theta _2}\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)=\left(\begin{array}{cccc}0& \frac{1}{\sqrt{6}}& 0& 0\\ \frac{1}{\sqrt{6}}& 0& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 0& 0\end{array}\right)\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)$$
$$\frac{}{\varphi _2}\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)=\left(\begin{array}{cccc}\frac{i}{3}\mathrm{cosh}\theta _2& \frac{i}{\sqrt{6}}\mathrm{sinh}\theta _2& 0& 0\\ \frac{i}{\sqrt{6}}\mathrm{sinh}\theta _2& \frac{2i}{3}\mathrm{cosh}\theta _2& 0& 0\\ 0& 0& \frac{i}{3}\mathrm{cosh}\theta _2& 0\\ 0& 0& 0& 0\end{array}\right)\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)$$
$$\frac{}{\psi }\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)=\frac{i}{6}\left(\begin{array}{cccc}1& 0& 0& 0\\ 0& 1& 0& 0\\ 0& 0& 1& 0\\ 0& 0& 0& 3\end{array}\right)\left(\begin{array}{c}ϵ^{(1)}\\ ϵ^{(2)}\\ ϵ^{(3)}\\ ϵ^{(4)}\end{array}\right)$$
A remarkable simplification takes place on observing that a spinor with $`ϵ^{(1)}=ϵ^{(2)}=ϵ^{(3)}=0`$ automatically satisfies the first four equations. Inserting this form in the last equation, we find
$$\frac{}{\psi }ϵ^{(4)}=\frac{i}{2}ϵ^{(4)}$$
from which the corresponding Killing spinor is
$$ϵ=e^{\frac{i}{2}\psi }\left(\begin{array}{c}0\\ 0\\ 0\\ 1\end{array}\right)$$
It is easy to see that there are no other solutions to the coupled set of equations. For each of the components $`ϵ^{(1)},ϵ^{(2)},ϵ^{(3)}`$, there is an irrational factor $`\sqrt{6}`$ in one or other equation, which implies that we can never get a solution that is single-valued in the angles $`\varphi _1,\varphi _2`$.
Thus, as expected, there is a single Killing spinor for this manifold, proving explicitly that it has $`\frac{1}{4}`$-supersymmetry. The Killing spinor has the very simple form given in Eq.(4.1), and is quite similar to a solution obtained in Ref.\[15\] in the context of 7-dimensional Einstein spaces.
Because this Killing spinor depends only on $`\psi `$, any orbifold of $`W_{4,2}`$ by an action under which spinors transform trivially (such as a translation in the angles) will preserve it. One example is provided by the $`Z_k\times Z_k^{}`$ transformation:
$$\theta _1\theta _1+\frac{2\pi }{k},\theta _2\theta _2+\frac{2\pi }{k^{}}$$
This introduces conical singularities into the Euclidean $`AdS_2`$ factors that make up the base of $`W_{4,2}`$, but as usual one expects that string propagation on this space is smooth.
5. Discussion
We have constructed Kiling spinors in various coordinate systems and thereby discovered a number of supersymmetric orbifolds of $`AdS_5`$. One should attempt to understand the global properties and causal structure of such spacetimes. Some of them are already known to be topological black holes.
One interesting proposal emerges from our discussion. There is an $`AdS_5/Z_2`$ orbifold with a circle of fixed points, very similar to the $`S^5/Z_2`$ orbifold obtained by placing D3-branes at a $`Z_2`$ ALE singularity. Both are cases of $`\frac{1}{2}`$-supersymmetry. Now for the latter, it is known\[14\] that blowing up the circle of fixed points is a relevant deformation which causes the theory to flow to the $`\frac{1}{4}`$-supersymmetric theory obtained by replacing $`S^5/Z_2`$ with the Stiefel manifold $`V_{4,2}`$. The corresponding conformal theories flow from $`N=2`$ to $`N=1`$ and can be obtained very simply by rotating branes in a brane construction\[16,,17\]. One could perhaps expect an analogous blowup of our $`AdS_5/Z_2`$ orbifold to lead to the non-compact Stiefel manifold $`W_{4,2}`$ discussed in Ref.\[2\]. The physics of this would be quite different from the compact case and possibly very interesting<sup>3</sup> This proposal arose in discussions with Debashis Ghoshal..
In comparing our results with those of Ref., we find that we have reproduced the $`Z`$-orbifold discussed there and explicitly shown that it is $`\frac{1}{2}`$-supersymmetric, which is important for the identification proposed with $`pp`$-waves on a brane. However, it is not clear if we have found the $`Z\times Z`$ orbifold that also finds brief mention in their work. This is supposed to be dual to a 3-brane with a pp-wave together with a D-instanton, and one would expect it to be $`\frac{1}{4}`$-supersymmetric. We have found two $`\frac{1}{4}`$-supersymmetric orbifolds that include the action in Eq.(3.1). One is found by adjoining the $`Z`$ action which compactifies one or more of the spatial coordinates in the brane, as discussed below Eq.(3.1). The other is obtained by adjoining a $`Z_k`$ action, as discussed after Eq.(3.1). In neither of these cases does the second orbifold group appear symmetrically with the first, while the authors of Ref.\[8\] seem to suggest that such an example should exist. We hope to return to this point in the future, along with a study of the physical interpretation, in terms of the brane worldvolume theory, for the various orbifolds we have constructed here.
The compactification of type IIB string theory on $`AdS_5\times S^5`$ possesses a remarkable “symmetry” between the two factors. Geometrically, the only difference is that while $`S^5`$ solves a quadratic equation in $`R^6`$, $`AdS_5`$ solves a quadratic equation in $`R^{4,2}`$. This “symmetry” is not reflected in the emergence of this background as the near-horizon geometry of D3-branes, where $`R^6`$ has a physical interpretation as the flat space transverse to the 3-branes, but $`R^{4,2}`$ does not appear. One may speculate that such a symmetry may be visible or partly visible in F-theory, which can sometimes be given a 12-dimensional interpretation. This philosophy has been partially explored in Ref.\[8\] using earlier observations in Ref.\[18\]. It would be interesting to find out whether such a symmetry can be exploited to systematically classify the supersymmetric orbifolds of $`AdS_5`$.
Acknowledgements:
We are grateful to Sumit Das, Justin David, Debashis Ghoshal and Nemani V. Suryanarayana for helpful comments.
References
relax J. Maldacena, “The Large N Limit of Superconformal Field Theories and Supergravity”, Adv. Theor. Math. Phys. 2 (1998) 231; hep-th/9711200. relax R. Britto-Pacumio, A. Strominger and A. Volovich, “Holography for Coset Spaces”, hep-th/9905211. relax A. Hashimoto and N. Itzhaki, “Non-commutative Yang-Mills and the AdS/CFT Correspondence”, hep-th/9907166. relax J. Maldacena and J. Russo, “Large N Limit of Non-commutative Gauge Theories”, hep-th/9908134. relax R.B. Mann, “Topological Black Holes: Outside Looking In”, gr-qc/9709039; M. Bañados, A. Gomberoff and C. Martinez, “Anti-deSitter Space and Black Holes”, Class. Quant. Grav. 15 (1998) 3575; hep-th/9805087. relax G. Horowitz and D. Marolf, “A New Approach to String Cosmology”, hep-th/9805207. relax Y.-H. Gao, “AdS/CFT Correspondence and Quotient Space Geometry”, hep-th/9908134. relax K. Behrndt and D. Lüst, “Branes, Waves and AdS Orbifolds”, hep-th/9905180. relax S. Kachru and E. Silverstein, “4-D Conformal Theories and Strings on Orbifolds”, Phys. Rev. Lett. 80 (1998) 4855; hep-th/9802183. relax M. Bañados, C. Teitelboim and J. Zanelli, “The Black Hole in Three-dimensional Space-time”, Phys.Rev.Lett. 69 (1992) 1849; hep-th/9204099; M. Bañados, M. Henneaux, C. Teitelboim and J. Zanelli, “Geometry of the $`2+1`$ Black Hole”, Phys. Rev. D48 (1993) 1506; gr-qc/9302012. relax B.S. Acharya, J. Figueroa O’Farrill, C. Hull and B. Spence, “Branes at Conical Singularities and Holography”, Adv. Theor. Math. Phys. 2 (1999) 1249; hep-th/9808014. relax D. Morrison and R. Plesser, “Non-Spherical Horizons-I”, Adv. Theor. Math. Phys. 3 (1999) 1; hep-th/9810201. relax H. Lü, C.N. Pope and P.K. Townsend, “Domain Walls from Anti-deSitter Space-time”, Phys. Lett. B391 (1997) 39; hep-th/9607164. relax I. Klebanov and E. Witten, “Superconformal Field Theory on Threebranes at a Calabi- Yau Singularity”, Nucl.Phys. B536 (1998) 199; hep-th/9807080. relax C.N. Pope and N.P. Warner, “Two New Classes of Compactifications of $`d=11`$ Supergravity”, Class. Quant. Grav 2 (1985) L1. relax A. Uranga, “Brane configurations for Branes at Conifolds”, JHEP 01 (1999) 022; hep-th/9811004. relax K. Dasgupta and S. Mukhi, “Brane Constructions, Conifolds and M-Theory”, Nucl. Phys. B551 (1999) 204; hep-th/9811139. relax A.A. Tseytlin, “Type IIB Instanton as a Wave in Twelve Dimensions”, Phys. Rev. Lett. 78 (1997) 1864; hep-th/9612164.
|
no-problem/9908/cond-mat9908486.html
|
ar5iv
|
text
|
# Probing Pairing Symmetry in Sr2RuO4
## Abstract
We study the spin dynamics in a p-wave superconductor at the nesting vector associated with $`\alpha `$ and $`\beta `$ bands in Sr<sub>2</sub>RuO<sub>4</sub>. We find a collective mode at the nesting vector in the superconducting phase identified as the odd-parity pairing state which breaks time reversal symmetry. This mode in the spin channel only exists in the p-wave superconductor, not in s- or d-wave superconductors. We propose that probing this mode would clarify the pairing symmetry in Sr<sub>2</sub>RuO<sub>4</sub>. The possibility of second superconducting phase transition is also discussed.
The nature of superconductivity discovered in Sr<sub>2</sub>RuO<sub>4</sub> has been the subject of intense theoretical and experimental activity. Although Sr<sub>2</sub>RuO<sub>4</sub> has the same layered perovskite structure as La<sub>2</sub>CuO<sub>4</sub>, the prototype of the cuprates, the behavior is remarkably different. At present, not much is known about any possible relation to the cuprates.
While it is clear that electron correlation effects are important in Sr<sub>2</sub>RuO<sub>4</sub>, the normal state is characterized as essentially a Fermi liquid below 50 K. The resistivities in all directions show $`T^2`$ behavior for $`T50K`$. The effective mass is about $`34m_{electron}`$ and the susceptibility is also about $`34\chi _0`$ where $`\chi _0`$ is the pauli spin susceptibility. In contrast to the conventional normal state (below 50 K), there are considerable evidences that the superconducting state (below about 1 K) is unconventional. The nuclear quadrupole resonance(NQR) does not show the Hebel-Slicheter peak . The transition temperature is very sensitive to non-magnetic impurities. The <sup>17</sup>O NMR knight shift experiment shows that the spin susceptibility has no change across $`T_c`$ but stays just the same as in the normal state for the magnetic field parallel to the RuO<sub>2</sub> plane.
Shortly after the discovery of the superconductivity in Sr<sub>2</sub>RuO<sub>4</sub>, it was proposed that the odd-pairity(spin triplet) Cooper pairs are formed in the superconducting state in analogy with <sup>3</sup>He. Further evidence favoring spin triplet pairing is the observation of the ferromagnetic metallic state in SrRuO<sub>3</sub> which is the three dimensional analogue of the layered Sr<sub>2</sub>RuO<sub>4</sub>. Since a weak coupling analysis of the spin triplet state implies nodeless gap, it is puzzling that the specific heat and NQR measurements show a large residual density of states (DOS), $`5060\%`$ of DOS of the normal state, in the superconducting phase. As a consequence, nonunitary superconducting state like <sup>3</sup>He A<sub>1</sub> phase, has been proposed. However, recent specific heat measurement shows that the residual DOS is about $`25\%`$ of the normal DOS, which indicates that the nonunitary state may not be stabilized.
An alternative explanation, so called, orbital dependent superconductivity was proposed. Since four $`4d`$ electrons in Ru<sup>4+</sup> partially fill the $`t_{2g}`$ band, the relevant orbitals are $`d_{xy}`$, $`d_{xz}`$, and $`d_{yz}`$ which determine the electronic properties. Using the quasi-two dimensional nature of the electronic dispersion, they show that there are two superconducting order parameters for two different classes of the orbitals. The gap of one class of bands is substantially smaller than that of other class of bands. The presence of gapless excitations for temperatures greater than the smaller gap would account for a residual DOS. The recent analysis of London penetration depth and coherence length led to the evidence for orbital dependent superconductivity identifying $`d_{xy}`$ as the orbital relevant for superconductivity. The possibility of the second superconducting phase transition was also discussed when the pairing symmetries are different for different classes of the bands.
Sigrist et al proposed the following order parameter which is claimed to be compatible with all the present experimental data.
$$𝐝=\widehat{d}(k_1\pm ik_2),$$
(1)
where $`\widehat{d}`$ is parallel to the $`\widehat{c}`$ axis and the gap is described as the tensor represented by $`𝐝`$ in the following way.
$$\widehat{\mathrm{\Delta }}(𝐤)=𝝈𝐝i\sigma _2,$$
(2)
where $`𝝈`$ is the Pauli matrix. Here $`\widehat{d}`$ is the spin vector whose direction is perpendicular to the direction of the spin associated with the condensed pair. Notice that the direction of the order parameter vector is frozen along the $`\widehat{c}`$ direction due to the crystal field and there is a full gap on the whole Fermi surface.
The details of the Fermi surface have been observed by quantum oscillations . The Fermi surface consists of three nearly-cylindrical sheets, which is consistent to the electronic band calculations. Three Fermi sheets are labeled by $`\alpha `$, $`\beta `$, and $`\gamma `$. While the $`\gamma `$ sheet of the Fermi surface can be attributed solely to the $`d_{xy}`$ Wannier function, the $`\alpha `$ and $`\beta `$ sheets are due to the hybridization of the $`d_{xz}`$ and $`d_{yz}`$ Wannier functions. Combining the orbital dependent superconductivity and experimental observation, the gap associated with $`\gamma `$ band is larger than that of $`\alpha `$ and $`\beta `$ bands. Therefore the $`\gamma `$ band, which is essentially quasi-isotropic two dimensional, is responsible for the existing superconductivity. On the other hand, $`\alpha `$ and $`\beta `$ sheets are quasi-one dimensional which can be visualized as a set of parallel planes separated by $`Q=2\pi /3`$ running both in $`k_x`$ and $`k_y`$ directions. Therefore, it is natural to expect a sizable nesting effects at the wave vector $`𝐐=(2\pi /3,2\pi /3)`$ originated from $`\alpha `$ and $`\beta `$ bands. In the normal state, one can see that there should be a collective mode in the spin channel due to the nesting, and has been shown by the numerical calculation on the static susceptibility. The neutron scattering experiment also shows a peak at the wave vector, $`(0.6\pi ,0.6\pi ,0)`$ close to the nesting vector, with energy transfer $`6.2meV`$. Mazin and Singh discussed the possibility of a competition between p-wave and d-wave superconductivity in Sr<sub>2</sub>RuO<sub>4</sub>. The experimental result also casts some doubt on the predominant role of ferromagnetic spin fluctuations in the mechanism of spin triplet superconductivity. Although it is generally accepted that the pairing symmetry in Sr<sub>2</sub>RuO<sub>4</sub> has the odd pairity, a direct theoretical prediction is still necessary to determine the pairing symmetry among the possible order parameters which have the odd pairity.
In this paper, we propose a way to probe the pairing symmetry in Sr<sub>2</sub>RuO<sub>4</sub>. We calculate the spin-spin correlation function at the nesting vector, $`𝐐=(2\pi /3,2\pi /3)`$, using the Green function method. It is important to include the coupling between the spin density and the vectorial order parameter fluctuation which is the unique property of p-wave superconductor. We find a collective mode in the spin channel in the superconducting state only in p-wave superconductor with the pairing symmetry which breaks time reversal symmetry. Since the position of the resonant peak is just below $`2\mathrm{\Delta }`$, this will also determine the size of the smaller gap related to $`\alpha `$ and $`\beta `$ bands which have the nesting. On the other hand, no observation of the mode will indicate that the pairing symmetry associated with $`\alpha `$ and $`\beta `$ bands is different from that with $`\gamma `$ band, assuming that the pairing symmetry in $`\gamma `$ band, which does not have any nesting effect on the Fermi surface, is the proposed one as Eq. (1). Therefore there must be a second superconducting phase transition at a rather low temperature.
Using the Nambu’s representation, the Green function can be written as
$$G^1(\omega _n,𝐤)=i\omega _n\xi _𝐤\rho _3\sigma _3\mathrm{\Delta }\rho _1𝝈\widehat{d}i\sigma _2,$$
(3)
where $`𝝆`$ and $`𝝈`$ are Pauli matrices which operate in the particle-hole and spin spaces, respectively. Here $`\xi _𝐤=k^2/(2m)\mu `$, where $`\mu `$ is the chemical potential. In the superconducting state, the bare susceptibility which represents the spin flip procedure can be written by using the Green functions.
$$\chi ^{00}(\omega _\nu ,𝐪)=T\underset{n}{}\underset{𝐤}{}Tr[G(\omega _n,𝐤)\alpha _+G(\omega _n+\omega _\nu ,𝐤+𝐪)\alpha _{}],$$
(4)
where $`\omega _n`$ is the Matubara frequency and the spin vertex $`𝜶`$ is given by
$$𝜶=\frac{1+\rho _3}{2}𝝈+\frac{1\rho _3}{2}\sigma _2𝝈\sigma _2,$$
(5)
and $`\alpha _\pm =\alpha _1\pm i\alpha _2`$.
Since the gap order parameter fluctuation couples to the spin density, the susceptibility renormalized by order parameter fluctuations consists of two parts.
$$\chi ^0(\omega ,𝐪)=\chi ^{00}(\omega ,𝐪)\frac{V(\omega ,𝐪)g\overline{V}(\omega ,𝐪)}{1g\mathrm{\Pi }(\omega ,𝐪)},$$
(6)
where the $`g`$ is the interaction strength and responsible for the superconductivity. Here $`V(\omega ,𝐪)`$ and $`\mathrm{\Pi }(\omega ,𝐪)`$ can be computed as follows.
$`V(\omega _\nu ,𝐪)`$ $`=`$ $`T{\displaystyle \underset{n}{}}{\displaystyle \underset{𝐤}{}}Tr[G(\omega _n,𝐤)\alpha _+G(\omega _n+\omega _\nu ,𝐤+𝐪)(\alpha _{}\rho _1\sigma _1)],`$ (7)
$`\mathrm{\Pi }(\omega _\nu ,𝐪)`$ $`=`$ $`T{\displaystyle \underset{n}{}}{\displaystyle \underset{𝐤}{}}Tr[G(\omega _n,𝐤)(\alpha _+\rho _1\sigma _1)G(\omega _n+\omega _\nu ,𝐤+𝐪)(\alpha _{}\rho _1\sigma _1)].`$ (8)
Using $`\xi _𝐤=\xi _{𝐤+𝐐}`$ for the nesting vector $`𝐐`$, we found the following results at $`T=0`$.
$`\mathrm{Re}\chi ^{00}(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}g^1N_0\frac{|\omega |\mathrm{arcsin}|\omega |/2\mathrm{\Delta }}{2\sqrt{|\omega ^24\mathrm{\Delta }^2|}}\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ g^1N_0\frac{|\omega |\mathrm{ln}\left(\frac{4\mathrm{\Delta }^2}{\omega ^24\mathrm{\Delta }^2}\right)}{2\sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (9)
$`\mathrm{Im}\chi ^0(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}0\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ N_0\frac{\pi \omega }{2\sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (10)
$`\mathrm{Re}V(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}N_0\frac{\mathrm{\Delta }sgn(\omega )\mathrm{arcsin}|\omega |/2\mathrm{\Delta }}{\sqrt{|\omega ^24\mathrm{\Delta }^2|}}\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ N_0\frac{\mathrm{\Delta }sgn(\omega )\mathrm{ln}\left(\frac{4\mathrm{\Delta }^2}{\omega ^24\mathrm{\Delta }^2}\right)}{\sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (11)
$`\mathrm{Im}V(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}0\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ N_0\frac{\pi \mathrm{\Delta }}{\sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (12)
$`\mathrm{Re}\mathrm{\Pi }(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}N_0\frac{2\mathrm{\Delta }^2\mathrm{arcsin}|\omega |/2\mathrm{\Delta }}{|\omega |\sqrt{|\omega ^24\mathrm{\Delta }^2|}}\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ N_0\frac{2\mathrm{\Delta }^2\mathrm{ln}\left(\frac{4\mathrm{\Delta }^2}{\omega ^24\mathrm{\Delta }^2}\right)}{|\omega |\sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (13)
$`\mathrm{Im}\mathrm{\Pi }(\omega ,𝐐)`$ $`=`$ $`\{\begin{array}{cc}0\hfill & |\omega |<2\mathrm{\Delta }\hfill \\ N_0\frac{2\pi \mathrm{\Delta }^2}{\omega \sqrt{\omega ^24\mathrm{\Delta }^2}}\hfill & |\omega |>2\mathrm{\Delta }\text{,}\hfill \end{array}`$ (14)
where $`N_0`$ is the DOS at the Fermi level. Using the above result, the renormalized susceptibility is obtained as follows for $`|\omega |<2\mathrm{\Delta }`$,
$`\mathrm{Re}\chi ^0(\omega ,𝐐)`$ $`=`$ $`g^1{\displaystyle \frac{N_0}{2}}{\displaystyle \frac{|\omega |\mathrm{arcsin}\frac{|\omega |}{2\mathrm{\Delta }}}{\sqrt{4\mathrm{\Delta }^2\omega ^2}}}+{\displaystyle \frac{N_0^2\mathrm{\Delta }^2|\omega |(\mathrm{arcsin}\frac{|\omega |}{2\mathrm{\Delta }})^2}{g^1|\omega |(4\mathrm{\Delta }^2\omega ^2)2N_0\mathrm{\Delta }^2\mathrm{arcsin}\frac{|\omega |}{2\mathrm{\Delta }}\sqrt{4\mathrm{\Delta }^2\omega ^2}}},`$ (15)
$`\mathrm{Im}\chi ^0(\omega ,𝐐)`$ $`=`$ $`0.`$ (16)
Including the effects of the exchange interaction within the random phase approximation(RPA), the full dynamical spin susceptibility is expressed as
$$\chi (\omega ,𝐪)=\frac{\chi ^0(\omega ,𝐪)}{1I_𝐪\chi ^0(\omega ,𝐪)},$$
(17)
where the exchange interaction $`I_𝐐I`$ for $`𝐐`$. Since $`\mathrm{Im}\chi ^0=0`$ and $`\mathrm{Re}\chi ^0`$ diverges as $`\omega `$ approaches to $`2\mathrm{\Delta }`$, there exists a collective mode when $`I<g`$. The position of the mode is at
$$\omega =2\mathrm{\Delta }\frac{\pi ^2}{4}\frac{g^2I^2}{(gI)^2}\mathrm{\Delta }N_0^2.$$
(18)
Here we have assumed that the $`\mathrm{arcsin}|\omega |/2\mathrm{\Delta }\pi /2`$ consistent with the result, and $`g,I<1/N_0`$. Notice that the position of the mode is very close to $`2\mathrm{\Delta }`$. The intensity of the peak is
$$\frac{\pi ^2}{2}\frac{g^3I}{(gI)^3}\mathrm{\Delta }N_0^2.$$
(19)
If the coupling between the spin density and the order parameter fluctuation, $`g`$, is rather large, then there are two solutions which satisfy the condition, Re $`\chi _0(\omega )=1/I`$, for $`g<I`$. However, the separation between two modes is $`\frac{\pi ^2g^3}{4(gI)^2}\sqrt{g8Ig+8I^2}\mathrm{\Delta }N_0^2`$ which is very tiny so that we expect to observe only one mode at
$$\omega =2\mathrm{\Delta }\frac{\pi ^2g^2(g2I)^2}{8(gI)^2}\mathrm{\Delta }N_0^2.$$
(20)
Now, let us investigate the case of the spin singlet superconductors, such as s- or d-wave superconductor. In the case of the spin singlet superconductor, the bare spin-spin correlation function can be obtained through the following expression.
$$\chi ^{00}(\omega ,𝐐)=\underset{𝐤}{}\left(1\frac{\xi _𝐤\xi _{𝐤+𝐐}+\mathrm{\Delta }_𝐤\mathrm{\Delta }_{𝐤+𝐐}}{E_𝐤E_{𝐤+𝐐}}\right)\left(\frac{1}{\omega E_𝐤E_{𝐤+𝐐}}\frac{1}{\omega +E_𝐤+E_{𝐤+𝐐}}\right).$$
(21)
For s-wave superconductor, i.e., $`\mathrm{\Delta }_𝐤=\mathrm{\Delta }_{𝐤+𝐐}=\mathrm{\Delta }`$, one can obtain the following results for $`|\omega |<2\mathrm{\Delta }`$.
$`\mathrm{Re}\chi ^0(\omega ,𝐐)`$ $`=`$ $`N_0\mathrm{ln}({\displaystyle \frac{\sqrt{|\omega ^24\mathrm{\Delta }^2|}}{\mathrm{\Delta }}}+{\displaystyle \frac{\sqrt{|\omega ^23\mathrm{\Delta }^2|}}{\mathrm{\Delta }}})+\mathrm{ln}(C),`$ (22)
$`\mathrm{Im}\chi ^0(\omega ,𝐐)`$ $`=`$ $`0,`$ (23)
where $`C`$ is a constant. This implies that one needs enomously large interaction $`I`$ to get the collective mode, i.e., $`I1/N_0`$, which is practically impossible.
Let us study the possibility of having the resonance peak in the d-wave superconductor at the nesting vector $`𝐐=(2\pi /3,2\pi /3)`$. Assuming that the superconducting phase is described by the conventional BCS superconductor with d-wave pairing symmetry, $`\mathrm{\Delta }(𝐤)=\frac{\mathrm{\Delta }}{2}[cos(k_x)cos(k_y)]`$, we use the same expression as Eq. (21) for the bare spin-spin correlation function. Due to the coherence factor, there is a collective mode at $`𝐐=(\pi ,\pi )`$ even without the nesting in the electronic dispersion. However, in the case of $`𝐐=(2\pi /3,2\pi /3)`$, we do not have simple relation as that for $`𝐐=(\pi ,\pi )`$. In fact, $`\mathrm{\Delta }_𝐤`$ is equal to $`\mathrm{\Delta }_{𝐤+𝐐}`$ for the line from $`𝐤=(2\pi /3,0)`$ to $`(0,2\pi /3)`$, which makes the coherence factor $`O(1)`$, but $`\mathrm{\Delta }_𝐤`$ and $`\mathrm{\Delta }_{𝐤+𝐐}`$ have the linear dispersion in $`(k_x,k_y)`$ so it does not produce any singularity in the spin susceptibility. If the momentum lies near the node, then we have the same dispersion relation for the $`𝐐=(\pi ,\pi )`$ and it was found that there is no singularity in spin channel if $`𝐤`$ and $`𝐤+𝐐`$ are near the nodes. Therefore we do not expect any collective mode in either s- or d-wave superconductor at the wave vector $`𝐐=(2\pi /3,2\pi /3)`$.
The possible Cooper pairing states were classified according to the irreducible representation of the tetragonal point group D<sub>4h</sub> which include four one-dimensional and one two-dimensional representations for both even and odd parity. Assuming that the order parameter associated with $`\alpha `$($`\beta `$) band has the odd parity with a different pairing symmetry from Eq. (1), we also study the existence of the resonance peak with the following order parameters classified as the odd-parity pairing.
$$𝐝=\{\begin{array}{cc}\widehat{x}k_1+\widehat{y}k_2\hfill & \\ \widehat{x}k_1\widehat{y}k_2\hfill & \\ \widehat{x}k_2+\widehat{y}k_1\hfill & \\ \widehat{x}k_2\widehat{y}k_1.\hfill & \end{array}$$
(24)
Due to the coherence factor, we found that the spin-spin correlation function behaves as in Eq. 23. However, once the fluctuation effect gets strong, one might get a collective mode with the other order parameters as in Eq. 24. Since the coupling between the spin density and the fluctuation of the order parameter should not destabilize the ground state, one needs careful investigation of the values such as $`\mathrm{\Delta }N_0`$, $`gN_0`$ and $`I_𝐐N_0`$ which would determine the possibility of the resonance peak due to the fluctuation effect. . Therefore, we conclude that the collective mode at the nesting vector only exits with the proposed order parameter as in Eq. (1) unless there is unusually strong coupling between the spin density and the order parameter fluctuation. Notice that only the order parameter which breaks the time reversal symmetry shows the resonance peak at the nesting vector.
In conclusion, we studied the spin dynamics in p-wave superconductor at the nesting vector associated with $`\alpha `$ and $`\beta `$ bands in Sr<sub>2</sub>RuO<sub>4</sub>. We found that there is a collective mode at the frequency just below $`2\mathrm{\Delta }`$ where $`\mathrm{\Delta }`$ is a smaller gap according to the orbital dependent superconductivity. We show that this mode exists only in p-wave superconductor, not in s- or d-wave superconductor. We also presented that the other odd pairing states do not produce the collective mode unless there is unusually strong coupling between the spin density and the order parameter fluctuation. Therefore we suggest that probing this mode will determine the pairing symmetry in Sr<sub>2</sub>RuO<sub>4</sub> which breaks time reversal symmetry assuming that all bands favor the same pairing symmetry. This will clarify the controversial situation about the possibility of having s- or d-wave superconductor in Sr<sub>2</sub>RuO<sub>4</sub>. Moreover, no observation of this mode will indicate that the order parameter associated with $`\alpha (\beta )`$ band is different from the proposed order parameter as in Eq. (1). This implies that there must be a second superconducting phase transition at rather low temperature. The observation of the strong pinning of the vortex about $`50mK`$ might be an indication of second superconducting phase transition, although we believe that second superconducting phase is described as one of the possible p-wave pairing states.
###### Acknowledgements.
I thank S. Chakravarty, Yong Baek Kim, Kazumi Maki, and especially M. Sigrist for helpful discussions. This work was conducted under the auspices of the Department of Energy, supported (in part) by funds provided by the University of California for the conduct of discretionary research by Los Alamos National Laboratory.
|
no-problem/9908/hep-ph9908273.html
|
ar5iv
|
text
|
# 𝜂'→𝛾𝛾 and the topological susceptibility
## 1 The Result
Radiative decays of the $`\eta ^{}`$ are currently attracting renewed interest. (See for references.) In this talk we summarise the special features that arise due to the gluonic contribution to the $`U_A(1)`$ anomaly when PCAC methods (including chiral Lagrangians) are used in the flavour singlet channel. In , we presented an analysis of $`\eta ^{}\gamma \gamma `$ decay in the chiral limit of QCD, taking into account the gluonic anomaly and the associated anomalous scaling implied by the renormalisation group. Here, we summarise the results of a recent new analysis extending this to QCD with massive quarks, incorporating $`\eta \eta ^{}`$ mixing. In particular, we show how a combination of the radiative decay formula and a generalisation of the Witten–Veneziano mass formula for the $`\eta ^{}`$ may be used to measure the gluon topological susceptibility $`\chi (0)`$ in full QCD with massive quarks.
Our main result is summarised in the formulae:
$$f^{a\alpha }g_{\eta ^\alpha \gamma \gamma }+2n_fAg_{G\gamma \gamma }\delta _{a0}=a_{\mathrm{em}}^a\frac{\alpha }{\pi }$$
(1)
which describes the radiative decays, and
$`f^{a\alpha }(m^2)_{\alpha \beta }f^{T\beta b}=(2n_f)^2A\delta _{a0}\delta _{b0}`$ (2)
$`2d_{abc}\mathrm{tr}T^c\left(\begin{array}{ccc}m_u\overline{u}u& 0& 0\\ 0& m_d\overline{d}d& 0\\ 0& 0& m_s\overline{s}s\end{array}\right)`$ (3)
which defines the decay constants appearing in (1) through a modification of Dashen’s formula to include the gluon contribution to the $`U_A(1)`$ anomaly.
In these formulae, $`\eta ^\alpha `$ denotes the neutral pseudoscalars $`\pi ^0,\eta ,\eta ^{}`$. The (diagonal) mass matrix is $`(m^2)_{\alpha \beta }`$ and $`g_{\eta ^\alpha \gamma \gamma }`$ is the appropriate coupling, defined as usual from the decay amplitude by $`\gamma \gamma |\eta ^\alpha =ig_{\eta ^\alpha \gamma \gamma }ϵ_{\lambda \rho \alpha \beta }p_1^\alpha p_2^\beta ϵ^\lambda (p_1)ϵ^\rho (p_2)`$. The constant $`a_{\mathrm{em}}^a`$ is the coefficient of the electromagnetic contribution to the axial current anomaly:
$$^\mu J_{\mu 5}^a=d_{acb}m^c\varphi _5^b+2n_f\delta _{a0}Q+a_{\mathrm{em}}^a\frac{\alpha }{8\pi }F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }$$
(5)
Here, $`J_{\mu 5}^a`$ is the axial current, $`\varphi _5^a=\overline{q}\gamma _5T^aq`$ is the quark pseudoscalar and $`Q=\frac{\alpha _s}{8\pi }\mathrm{tr}G^{\mu \nu }\stackrel{~}{G}_{\mu \nu }`$ is the gluon topological charge. $`m^a`$ are the quark masses (see eq(11)). $`a=0,3,8`$ is the flavour index, $`T^{3,8}`$ are $`SU(3)`$ generators and $`T^0=\mathrm{𝟏}`$. The $`d`$-symbols are defined by $`\{T^a,T^b\}=d_{abc}T^c`$.
The decay constants $`f^{a\alpha }`$ in (1) are defined by the relation (2). In general they are not the couplings of the pseudoscalar mesons to the axial current. In the flavour singlet sector, such a definition would give a RG non-invariant decay constant which would not coincide with the quantities arising in the correct decay formula (1). In contrast, all the quantities in the formulae (1),(2) are separately RG invariant,. The proof is not immediately obvious, and depends on the RGEs for the various Green functions and vertices defining the terms in (1),(2) being evaluated on-shell or at zero-momentum.
In practice, since flavour $`SU(2)`$ symmetry is almost exact, the relations for $`\pi ^0`$ decouple and are simply the standard ones with $`f^{3\pi }`$ identified as $`f_\pi `$ (see eqs(24),(27)). In the octet-singlet sector, however, there is mixing and the decay constants form a $`2\times 2`$ matrix:
$$f^{a\alpha }=\left(\begin{array}{cc}f^{0\eta ^{}}& f^{0\eta }\\ f^{8\eta ^{}}& f^{8\eta }\end{array}\right)$$
(6)
The four components are independent. In particular, for broken $`SU(3)`$, there is no reason to express $`f^{a\alpha }`$ as a diagonal matrix times an orthogonal $`\eta \eta ^{}`$ mixing matrix, which would give just three parameters. Several convenient parametrisations may be made, e.g. involving two constants and two mixing angles, but this does not seem to reflect any special dynamics.
The novelty of our results of course lies in the extra terms arising in (1) and (2) due to the gluonic contribution to the $`U_A(1)`$ anomaly. The coefficient $`A`$ is the non-perturbative number which specifies the topological susceptibility in full QCD with massive dynamical quarks. The topological susceptibility is defined as
$$\chi (0)=d^4xi0|TQ(x)Q(0)|0$$
(7)
The anomalous chiral Ward identities determine its dependence on the quark masses and condensates up to an undetermined parameter, viz.
$$\chi (0)=A\left(1A\underset{q}{}\frac{1}{m_q\overline{q}q}\right)^1$$
(8)
Notice how this satisfies the well-known result that $`\chi (0)`$ vanishes if any quark mass is set to zero.
The modified flavour singlet Dashen formula is in fact a generalisation of the Witten–Veneziano mass formula for the $`\eta ^{}`$. Here, however, we do not impose the leading order in $`1/N_c`$ approximation that produces the Witten–Veneziano formula. Recall that this states
$$m_\eta ^{}^2+m_\eta ^22m_K^2=\frac{6}{f_\pi ^2}\chi (0)|_{\mathrm{YM}}$$
(9)
To recover (7) from our result (see the first of eqs(9)) the condensate $`m_s\overline{s}s`$ is replaced by the term proportional to $`f_\pi ^2m_K^2`$ using a standard Dashen equation, and the singlet decay constants are set to $`\sqrt{2n_f}f_\pi `$. The identification of the large $`N_c`$ limit of the coefficient $`A`$ with the non-zero topological susceptibility of pure Yang-Mills theory follows from large $`N_c`$ counting rules and is explained in ref.
The final element in (1) is the extra ‘coupling’ $`g_{G\gamma \gamma }`$ in the flavour singlet decay formula, which arises because even in the chiral limit the $`\eta ^{}`$ is not a Goldstone boson because of the gluonic $`U_A(1)`$ anomaly. A priori, this is not a physical coupling, although (suitably normalised) it could be modelled as the coupling of the lightest predominantly glueball state mixing with $`\eta ^{}`$. However, this interpretation would probably stretch the basic dynamical assumptions underlying (1) too far, and is not necessary either in deriving or interpreting the formula. In fact, the $`g_{G\gamma \gamma }`$ term arises simply because in addition to the electromagnetic anomaly the divergence of the axial current contains both the quark pseudoscalar $`\varphi _5^a`$ and the gluonic anomaly $`Q`$. Diagonalising the propagator matrix for these operators isolates the $`\eta `$ and $`\eta ^{}`$ poles, whose couplings to $`\gamma \gamma `$ give the usual terms $`g_{\eta \gamma \gamma }`$ and $`g_{\eta ^{}\gamma \gamma }`$. However, the remaining operator (which we call $`G`$) also couples to $`\gamma \gamma `$ and therefore also contributes to the decay formula, whether or not we assume that its propagator is dominated by a ‘glueball’ pole.
Of course, the presence of the coupling $`g_{G\gamma \gamma }`$ in (1) appears to remove any predictivity from the $`\eta ^{}\gamma \gamma `$ decay formula. In a strict sense this is true, but we shall argue that it may nevertheless be a good dynamical approximation to assume $`g_{G\gamma \gamma }`$ is small compared to $`g_{\eta ^{}\gamma \gamma }`$. In this case, we can combine eqs (1) and (2) to give a measurement of the non-perturbative coefficient $`A`$ in $`\chi (0)`$. To see this, assume $`m_u=m_d0`$. The terms in (1), (2) involving the pion then decouple leaving the following five equations, in which we assume the physical quantities $`m_\eta ,m_\eta ^{},g_{\eta \gamma \gamma }`$ and $`g_{\eta ^{}\gamma \gamma }`$ are all known and we neglect $`g_{G\gamma \gamma }`$. The decay equations are:
$`f^{0\eta ^{}}g_{\eta ^{}\gamma \gamma }+f^{0\eta }g_{\eta \gamma \gamma }+6Ag_{G\gamma \gamma }=a_{\mathrm{em}}^0{\displaystyle \frac{\alpha }{\pi }}`$ (10)
$`f^{8\eta }g_{\eta \gamma \gamma }+f^{8\eta ^{}}g_{\eta ^{}\gamma \gamma }=a_{\mathrm{em}}^8{\displaystyle \frac{\alpha }{\pi }}`$ (11)
where $`a_{\mathrm{em}}^0=\frac{4}{3}N_c`$ and $`a_{\mathrm{em}}^8=\frac{1}{3\sqrt{3}}N_c`$, and the Dashen equations are:
$`\left(f^{0\eta ^{}}\right)^2m_\eta ^{}^2+\left(f^{0\eta }\right)^2m_\eta ^2=4m_s\overline{s}s+36A`$ (13)
$`f^{0\eta ^{}}f^{8\eta ^{}}m_\eta ^{}^2+f^{0\eta }f^{8\eta }m_\eta ^2={\displaystyle \frac{4}{\sqrt{3}}}m_s\overline{s}s`$ (14)
$`\left(f^{8\eta }\right)^2m_\eta ^2+\left(f^{8\eta ^{}}\right)^2m_\eta ^{}^2={\displaystyle \frac{4}{3}}m_s\overline{s}s`$ (15)
Clearly, the two purely octet formulae can be used to find $`f^{8\eta }`$ and $`f^{8\eta ^{}}`$ if both $`g_{\eta \gamma \gamma }`$ and $`g_{\eta ^{}\gamma \gamma }`$ are known. The off-diagonal Dashen formula then expresses $`f^{0\eta }`$ in terms of $`f^{0\eta ^{}}`$. This leaves the two purely singlet formulae involving the still-undetermined decay constant $`f^{0\eta ^{}}`$, the topological susceptibility coefficient $`A`$, and the coupling $`g_{G\gamma \gamma }`$. The advertised result follows immediately. If we neglect $`g_{G\gamma \gamma }`$, we can find $`f^{0\eta ^{}}`$ from the singlet decay formula and thus determine $`A`$ from the remaining flavour singlet Dashen formula. This is the generalisation of the Witten–Veneziano formula.
Without neglecting $`g_{G\gamma \gamma }`$, the five equations give a self-consistent description of the radiative decays, but are non-predictive. It is therefore important to analyse more carefully whether it is really legitimate to neglect $`g_{G\gamma \gamma }.`$ The argument is based on the fact that $`g_{G\gamma \gamma }`$ is both OZI suppressed and renormalisation group (RG) invariant. Since violations of the OZI rule are associated with the $`U_A(1)`$ anomaly, it is a plausible conjecture that we can identify OZI-violating quantities by their dependence on the anomalous dimension associated with the non-trivial renormalisation of $`J_{\mu 5}^0`$ due to the anomaly. In this way, RG non-invariance can be used as a flag to indicate those quantities expected to show large OZI violations. If this conjecture is correct, then we would expect the OZI rule to be reasonably good for the RG invariant $`g_{G\gamma \gamma }`$, which would therefore be suppressed relative to $`g_{\eta ^{}\gamma \gamma }`$. (An important exception is of course the $`\eta ^{}`$ mass itself, which although obviously RG invariant is not zero in the chiral limit as it would be in the OZI limit of QCD.) Notice that this conjecture has been applied already with some success to the ‘proton spin’ problem in polarised deep inelastic scattering. A related $`1/N_c`$ discussion is given in .
## 2 The Proof
Consider first QCD by itself without the coupling to electromagnetism. The axial anomaly is
$$^\mu J_{\mu 5}^a=M_{ab}\varphi _5^a+2n_fQ\delta _{a0}$$
(17)
The notation is defined in ref. The quark mass matrix is written as $`m^aT^a`$, so
$$\left(\begin{array}{ccc}m_u& 0& 0\\ 0& m_d& 0\\ 0& 0& m_s\end{array}\right)=m^0\mathrm{𝟏}+m^3T^3+m^8T^8$$
(18)
The condensates are written as
$$\left(\begin{array}{ccc}\overline{u}u& 0& 0\\ 0& \overline{d}d& 0\\ 0& 0& \overline{s}s\end{array}\right)=\frac{1}{3}\varphi ^0\mathrm{𝟏}+2\varphi ^3T^3+2\varphi ^8T^8$$
(19)
where $`\varphi ^c`$ is the VEV $`\overline{q}T^cq`$. Then
$$M_{ab}=d_{acb}m^c,\mathrm{\Phi }_{ab}=d_{abc}\varphi ^c$$
(20)
The anomalous chiral Ward identities, at zero momentum, for the two-point Green functions of these operators are
$`2n_fQQ\delta _{a0}+M_{ac}\varphi _5^cQ=0`$ (21)
$`2n_fQ\varphi _5^b\delta _{a0}+M_{ac}\varphi _5^c\varphi _5^b+\mathrm{\Phi }_{ab}=0`$ (22)
which imply
$$M_{ac}M_{bd}\varphi _5^c\varphi _5^d=(M\mathrm{\Phi })_{ab}+(2n_f)^2QQ\delta _{a0}\delta _{b0}$$
(24)
We also need the result for the general form of the topological susceptibility (see eq(6)):
$$\chi (0)QQ=\frac{A}{1(2n_f)^2A(M\mathrm{\Phi })_{00}^1}$$
(25)
Although the pseudoscalar operators $`\varphi _5^a`$ and $`Q`$ indeed couple to the physical states $`\eta ^\alpha =\eta ^{},\eta ,\pi ^0`$, it is more convenient to redefine linear combinations such that the resulting propagator matrix is diagonal and properly normalised. So we define operators $`\eta ^\alpha `$ and $`G`$ such that
$$\left(\begin{array}{cc}QQ& Q\varphi _5^b\\ \varphi _5^aQ& \varphi _5^a\varphi _5^b\end{array}\right)\left(\begin{array}{cc}GG& 0\\ 0& \eta ^\alpha \eta ^\beta \end{array}\right)$$
(26)
This is achieved by
$`G`$ $`=QQ\varphi _5^a(\varphi _5\varphi _5)_{ab}^1\varphi _5^b`$ (27)
$`=Q+2n_fA\mathrm{\Phi }_{0b}^1\varphi _5^b`$ (28)
and
$$\eta ^\alpha =f^{T\alpha a}\mathrm{\Phi }_{ab}^1\varphi _5^b$$
(30)
With this choice, the $`GG`$ propagator is
$$GG=A$$
(31)
and we impose the normalisation
$$\eta ^\alpha \eta ^\beta =\frac{1}{k^2m_{\eta ^\alpha }^2}\delta ^{\alpha \beta }$$
(32)
This implies that the constants $`f^{a\alpha }`$ in (19), which are simply the decay constants, must satisfy the (Dashen) identity
$`f^{a\alpha }m_{\alpha \beta }^2f^{T\beta b}`$ $`=\mathrm{\Phi }_{ac}(\varphi _5\varphi _5)_{cd}^1\mathrm{\Phi }_{db}`$ (33)
$`=(M\mathrm{\Phi })_{ab}+(2n_f)^2A\delta _{a0}\delta _{b0}`$ (34)
The last line follows from the Ward identities (15) and (16) . In terms of these new operators, the anomaly equation (10) is:
$$^\mu J_{\mu 5}^a=f^{a\alpha }m_{\alpha \beta }^2\eta ^\beta +2n_fG\delta _{a0}$$
(36)
Now recall how conventional PCAC is applied to the calculation of $`\pi ^0\gamma \gamma `$. The pion decay constant is defined as the coupling of the pion to the axial current
$$0|J_{\mu 5}^3|\pi =ik_\mu f_\pi 0|^\mu J_{\mu 5}^3|\pi =f_\pi m_\pi ^2$$
(37)
and satisfies the usual Dashen formula The next step is to define a ‘phenomenological pion field’ $`\pi `$ by
$$^\mu J_{\mu 5}^3f_\pi m_\pi ^2\pi $$
(38)
To include electromagnetism, the full anomaly equation is extended as in (3) to include the $`F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }`$ contribution. Using (24) we have
$`ik^\mu \gamma \gamma |J_{\mu 5}^3|0`$ (39)
$`=f_\pi m_\pi ^2\gamma \gamma |\pi |0+a_{\mathrm{em}}^a{\displaystyle \frac{\alpha }{8\pi }}\gamma \gamma |F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }|0`$ (40)
$`=f_\pi m_\pi ^2\pi \pi \gamma \gamma |\pi +a_{\mathrm{em}}^a{\displaystyle \frac{\alpha }{8\pi }}\gamma \gamma |F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }|0`$ (41)
where $`\pi \pi `$ is the pion propagator $`1/(k^2m_\pi ^2)`$. At zero momentum, the l.h.s. vanishes because of the explicit $`k_\mu `$ factor and the absence of massless poles. We therefore find,
$$f_\pi g_{\pi \gamma \gamma }=a_{\mathrm{em}}^3\frac{\alpha }{\pi }$$
(43)
In the full theory including the flavour singlet sector and the gluonic anomaly, we find a similar result. The ‘phenomenological fields’ are defined by (23) where the decay constants satisfy the generalised Dashen formula (22) . Notice, however, that they are not simply related to the couplings to the axial current as in (24) for the flavour non-singlet. We therefore find:
$`ik^\mu \gamma \gamma |J_{\mu 5}^a|0=f^{a\alpha }m_{\alpha \beta }^2\gamma \gamma |\eta ^\beta |0`$ (44)
$`+2n_f\gamma \gamma |G|0\delta _{a0}+a_{\mathrm{em}}^a{\displaystyle \frac{\alpha }{8\pi }}\gamma \gamma |F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }|0`$ (45)
$`=f^{a\alpha }m_{\alpha \beta }^2\eta ^\beta \eta ^\gamma \gamma \gamma |\eta ^\gamma `$ (46)
$`+2n_fGG\gamma \gamma |G\delta _{a0}+a_{\mathrm{em}}^a{\displaystyle \frac{\alpha }{8\pi }}\gamma \gamma |F^{\mu \nu }\stackrel{~}{F}_{\mu \nu }|0`$ (47)
using the fact that the propagators are diagonal in the basis $`\eta ^\alpha ,G`$. Using the explicit expressions (20),(21) for the Green functions, we find in this case:
$$f^{a\alpha }g_{\eta ^a\gamma \gamma }+2n_fAg_{G\gamma \gamma }\delta _{a0}=a_{\mathrm{em}}^a\frac{\alpha }{\pi }$$
(49)
where the extra coupling $`g_{G\gamma \gamma }`$ is defined through (28). This completes the derivation. It is evidently a straightforward generalisation of conventional PCAC with the necessary modification of the usual formulae to take account of the extra gluonic contribution to the axial anomaly in the flavour singlet channel, the key point being the identification of the operators $`\eta ^\alpha `$ and $`G`$ in (23).
Finally, notice that these methods may equally be applied to other decays involving the $`\eta ^{}`$ such as $`\eta ^{}V\gamma `$ where $`V`$ is a light $`1^{}`$ meson, $`\eta ^{}\pi \pi \gamma `$, $`\psi \eta ^{}\gamma `$, etc.
## 3 Acknowledgements
I would like to thank S. Narison and G. Veneziano for helpful discussions. This work was supported by the PPARC grant GR/L56374 and by the EC TMR grant FMRX-CT96-0008.
|
no-problem/9908/quant-ph9908021.html
|
ar5iv
|
text
|
# Quantum gates by coupled quantum dots and measurement procedure in Si MOSFET
## I Introduction
Since Shor’s factorization program was proposed, many studies have been carried out in order to realize the quantum computer . Nakamura succeeded in the control of the macroscopic quantum state as the solid devices of Josephson junctions. Recently, we have proposed the quantum computer in the asymmetric coupled dot in Si nanocrystals. The advantage of using the coupled dots attached by the gate electrode is that a quantum state can be controlled by the gate voltage in the solid circuits. Moreover, when the coupled Si quantum dots are embedded in the gate insulator of the MOSFET, the quantum state in the two coupled dots (qubits) is expected to be detected by the channel current. In the previous paper, we investigated the case where the sizes of Si nanocystals were small such that the energy-levels in the quantum dots were discrete and the two-state system was constituted by the lowest energy-levels of the two quantum dots. Here, in this paper, we consider the case where the sizes of the quantum dots are larger (of the order of a few tens of nm) and the energy-levels of the individual quantum dots are almost continuous. We also assume that the charging energy of each tunneling junction is large enough that the Coulomb blockade effects can be seen. This was realized in the quantum gate of the Josephson junction by Shnirman and Averin. The superconducting state has the advantage from the viewpoint of the decoherence that the qubits are connected by the coherent circuits. On the other hand, we use the normal state of the capacitively coupled quantum dots, because there exist some predictions that the decoherence time in the normal quantum dot array in semiconductors is not expected to be so short. First, in this paper, we show that the capacitively coupled dots can be reduced to a two-state system of interacting two qubits. Next, we show the detecting mechanism of the quantum state by the channel current based on the conventional MOSFET model. Asymmetry of the coupled dots is not assumed and we set $`e`$=1.
## II Capacitively coupled quantum dots as two qubits
The configuration of the quantum dots and capacitances are shown in Fig.1. We assume that the coupling between qubits is weaker than that in a qubit(The full formation of the electrostatic energy is shown in the Appendix), and we take
$$C_1=C_2=\sqrt{2}C_3=\sqrt{2}C_4C_5(=C_8),C_6(=C_9),C_7(=C_{10}).$$
(1)
By this assumption, we can expand the charging energy as a function of $`n_aN_\mathrm{A}N_\mathrm{B}`$ and $`n_bN_\mathrm{C}N_\mathrm{D}`$ ($`N=N_\mathrm{A}+N_\mathrm{B}=N_\mathrm{C}+N_\mathrm{D}`$), by a small $`C_3`$, and obtain the electrostatic energy:
$`U(n_a,n_b)`$ $`=`$ $`E_c\left[n_a+{\displaystyle \frac{\eta }{2E_c}}n_b+{\displaystyle \frac{C_7C_5}{C_5+C_7}}N{\displaystyle \frac{2C_5C_7}{C_5+C_7}}V_a\right]^2`$ (2)
$`+`$ $`E_c\left[n_b+{\displaystyle \frac{C_7C_5}{C_5+C_7}}N{\displaystyle \frac{2C_5C_7}{C_5+C_7}}V_b\right]^2`$ (3)
where $`E_c(C_5+C_7)/(8(C_5C_6+C_6C_7+C_7C_5))`$ and
$$\eta =\frac{\sqrt{2}(C_5^2+C_7^2\sqrt{2}C_5C_7)}{4(C_5C_6+C_6C_7+C_7C_5)^2}C_3$$
(4)
If we consider the gate voltage region where the electrostatic energy of $`n_j=0(j=a,b)`$ state crosses that of $`n_j=1`$ state described by Shnirman and Averin, the electrostatic energy and the tunneling amplitude, $`\mathrm{\Omega }_j`$, constitute the Hamiltonian of the two-state system as
$$H=\underset{j=a,b}{}(ϵ_j\sigma _{zj}+\mathrm{\Omega }_j\sigma _{xj})(\eta /4)\sigma _{za}\sigma _{zb}$$
(5)
where $`ϵ_a=E_c(1/2+[(C_7C_5)N2C_5C_7V_a]/(C_5+C_7))`$ and $`ϵ_b=E_c(1/2+[(C_7C_5)N2C_5C_7V_b]/(C_5+C_7))`$. $`\sigma _x`$ and $`\sigma _z`$ are Pauli matrices. When the capacitances are approximated as $`C_i=2\pi ϵ_{\mathrm{ox}}r^2/(d_i+(ϵ_{\mathrm{ox}}/ϵ_{\mathrm{Si}})r)`$ where $`ϵ_{\mathrm{ox}}=4`$, $`ϵ_{\mathrm{Si}}=12`$, $`d_i`$ is the size of the capacitance and $`r`$ is the radius of each quantum dot, then $`V_a,V_b`$ is of the order of tens of meV.
The mechanism of the controlled NOT operation is similar to that of the Josephson junction. Whether $`n_b=0`$ or $`n_b=1`$, the level-crossing gate voltage, $`V_a`$, shifts(see Eq.(3) and the controlled NOT operation is realized.
The dynamical motion of the excess charge in the two-state system can be easily considered by solving the time-dependent Schrödinger equation, and the excess charge shows the oscillating behavior depending on the energy of the two states. The time period of the oscillation, $`\tau _\delta `$, can be simply approximated as $`\tau _\delta \mathrm{}/\sqrt{\mathrm{\Omega }^2+(ϵ_aϵ_b)/4}`$.
## III Dissipation by environmental phonons
The polarized charged state of a qubit (coupled dots) behaves as the dipole moment under the electric field generated by the gate electrode. By the coupling of the dipole with the electric field, the two-state system can be described by the Bloch equation and the quantum calculation is realized similar to the NMR quantum computer. It is well known that the one of the attractive characteristics of the NMR quantum computer is its long decoherence time. Here, we comment on the decoherence time of the semiconductor dot array. The decoherence time in semiconductor dot array has been considered to be short. This is the reason why Shnirman and Averin used the Josephson effects in quantum gates. The decoherence in this case is considered mainly to originate from the phonon environments, where the interaction between the two-state system and the phonon bath is largely given by a deformation potential. The estimated decoherence time is not so short and of the order of $`10^7`$ secfrom the analysis based on the model of Leggett. This relatively long decoherence time will be related to the ’phonon bottleneck’ derived by Zanardi. We will have to include the effects of the higher excited energy-levels and temperature for more detailed estimation.
## IV Measurement mechanism in MOSFET
The qubit which changes the charge distribution can be detected by the MOSFET structure. MOSFET structure is considered to be a more efficient detecting devise in semiconductor quantum dots compared to the SET structure. This is due to the change in the charge distribution in the qubit being detected by the capacitance effects similar to those of the quantum point contact. In this section, we show the detailed detecting mechanism of the MOSFET based on the long-channel MOSFET model in the case of two qubits. The qubit system in the MOSFET proposed here can be seen as series of single coupled-dot MOSFETs (Fig.2). When bias, $`V_D`$, is applied between the source and drain, the depletion region expands from the source and drain such that the width of the depletion region increases toward the drain. Thus, the channel current differs depending on the positions of the qubits which change the charge distribution. The channel current between the $`i`$-th qubit and $`(i1)`$-th qubit is given for a small $`V_D`$ region as,
$$I_D^{(i)}\beta _0\left([V_G^{(i)}V_{\mathrm{th}}^{(i)}](V_iV_{i1})\frac{1}{2}\alpha (V_i^2V_{i1}^2)\right),$$
(6)
where $`\beta _0Z\mu _0C_0/L_i`$ ($`Z`$ is the channel width, $`\mu _0`$ is the mobility, $`L_i`$ is the channel length of $`i`$-th qubit where we set $`L_1=L_2=\mathrm{}=L_\mathrm{N}`$, and $`C_0`$ is the capacitance of the SiO<sub>2</sub>) and $`\alpha 1+\frac{1}{4\phi _\mathrm{B}}\frac{Q_B}{C_0}`$ where $`Q_\mathrm{B}`$ is the charge within the surface depletion region. $`V_\mathrm{G}`$ is the gate voltage, and the threshold voltage, $`V_{\mathrm{th}}^{(i)}`$, is given by $`V_{\mathrm{th}}^{(i)}=V_{\mathrm{th}}+\mathrm{\Delta }V_{\mathrm{th}}^{(i)}`$ where $`V_{\mathrm{th}}V_{\mathrm{FB}}+2\phi _\mathrm{B}+\frac{Q_B}{C_0}`$ ($`V_{\mathrm{FB}}`$ is a flat band voltage, $`\phi _\mathrm{B}`$ is the potential difference between the Fermi level and the intrinsic Fermi level of substrate), and the shift by the change of the charge distribution, $`\mathrm{\Delta }V_{\mathrm{th}}^{(i)}`$ in the $`i`$-th qubit. Then, we have the following conditions:
$$V_N=V_{DS}\mathrm{and}I_D^{(1)}=I_D^{(2)}=\mathrm{}=I_D^{(N)}$$
(7)
In the case of two qubits, with $`V_{Gi}=V_G^{(i)}V_{\mathrm{th}}^{(i)}`$ ($`V_{Gi}V_D`$ is assumed),
$$I_D=\frac{\beta _0}{(V_{G1}+V_{G2})}(V_{G1}V_{G2}V_D\frac{\alpha V_{G1}}{2}V_D^2)$$
(8)
Thus, whether ($`V_{G1}=V_g\mathrm{\Delta }V_{\mathrm{th}}`$ and $`V_{G2}=V_g`$ ) or ($`V_{G1}=V_g`$ and $`V_{G2}=V_g\mathrm{\Delta }V_{\mathrm{th}}`$ ), the difference of the corresponding currents, $`\mathrm{\Delta }I_D^{(12)}`$ is given as
$$\mathrm{\Delta }I_D^{(12)}\frac{\beta _0\alpha }{2(2V_g\mathrm{\Delta }V_{\mathrm{th}})}\mathrm{\Delta }V_{\mathrm{th}}V_D^2.$$
(9)
This difference can be observed in the nonlinear $`I_D`$-$`V_D`$ region and, in the pure linear region where the terms which include $`\alpha `$ disappear, the changed qubits cannot be distinguished.
## V Conclusions
We have investigated the quantum gates of the capacitively coupled quantum dot array in the MOSFET structure, and have derived the two-state Hamiltonian by the capacitances of the quantum dots. We also have shown the detecting mechanism of the MOSFET structure by analyzing the channel current based on the conventional MOSFET model.
## ACKNOWLEDGMENTS
The author is grateful to N. Gemma, S. Fujita, K. Ichimura, and J. Koga for fruitful discussions.
## A The electrostatic energy of the two qubits
In this section we show the electrostatic energy of the two qubits in terms of the capacitances of dots and the gate electrodes (Fig. 1). We assume $`C_1=C_2=\sqrt{2}C_3=\sqrt{2}C_4`$ , $`C_8=C_5`$, $`C_9=C_6`$ and $`C_{10}=C_7`$. The total electrostatic energy of the two qubits is given by
$$U=\underset{i=1,\mathrm{},10}{}\frac{q_i^2}{2C_i}q_7V_aq_{10}V_b,$$
(A1)
where $`q_i`$ shows the charge of the $`i`$-th capacitance and we have the relation between $`q_i(i=1,2\mathrm{}10)`$ and the total charge of the fore dots, $`N_\mathrm{A}`$, …,$`N_\mathrm{D}`$ as
$`N_\mathrm{A}`$ $`=`$ $`q_1+q_3+q_5+q_6,N_\mathrm{B}=q_2+q_4q_5+q_7,`$ (A2)
$`N_\mathrm{C}`$ $`=`$ $`q_1q_4+q_8+q_9,N_\mathrm{D}=q_2q_3q_9+q_{10}.`$ (A3)
By minimizing the energy Eq.(A1) at the fixed values of $`V_a`$, $`V_b`$ and $`N_\mathrm{A}\mathrm{}N_\mathrm{D}`$, we have
$`U(n_a,n_b)={\displaystyle \frac{1}{16}}({\displaystyle \frac{1}{C_a}}+{\displaystyle \frac{1}{C_b}})n_a^2+{\displaystyle \frac{1}{4}}\{{\displaystyle \frac{C_7C_5}{C_AC_B(C_3+C_6)^2}}N{\displaystyle \frac{C_7C_5}{C_AC_B(C_3+C_6)^2}}V^++{\displaystyle \frac{C_6C_3C_C}{C_CC_D(C_6C_3)^2}}V^{}\}n_a`$ (A4)
$`+`$ $`{\displaystyle \frac{1}{16}}({\displaystyle \frac{1}{C_a}}+{\displaystyle \frac{1}{C_b}})n_b^2+{\displaystyle \frac{1}{4}}\left\{{\displaystyle \frac{C_7C_5}{C_AC_B(C_3+C_6)^2}}N{\displaystyle \frac{C_7C_5}{C_AC_B(C_3+C_6)^2}}V^+{\displaystyle \frac{C_6C_3C_C}{C_CC_D(C_6C_3)^2}}V^{}\right\}n_b`$ (A5)
$`+`$ $`{\displaystyle \frac{1}{8}}({\displaystyle \frac{1}{C_a}}{\displaystyle \frac{1}{C_b}})n_an_b+{\displaystyle \frac{1}{4C_A}}N^2+{\displaystyle \frac{C_A}{4(C_AC_B(C_3+C_6)^2)}}\left[{\displaystyle \frac{C_3+C_6+C_A}{C_A}}N+C_7V^+\right]^2+{\displaystyle \frac{C_C}{4(C_CC_D(C_6C_3)^2)}}(C_7V^+)^2`$ (A6)
where
$`C_A`$ $`=`$ $`C_3+C_5+C_6,C_B=C_3+C_6+C_7,`$ (A7)
$`C_C`$ $`=`$ $`\varrho C_3+C_5+C_6,C_D=\varrho C_3+C_6+C_7,`$ (A8)
and $`V^\pm =V_a\pm V_b`$, $`1/C_a=(C_5+C_7)/((C_5+C_7)(C_3+C_6)+C_5C_7)`$, $`1/C_b=[C_5+C_7+2(\varrho +1)C_3]/[(C_5+C_6+\eta C_3)(C_6+C_7+\varrho C_3)(C_6C_3)^2]`$ and $`\varrho =2\sqrt{2}+1`$.
|
no-problem/9908/hep-ex9908035.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Correlations between particles produced in hadronic jets can be used to probe details of the jet fragmentation process. In $`e^+e^{}`$ annihilations into hadrons, the total charge, strangeness, baryon number, etc. of the final state particles in each event must be zero, and it is interesting to ask whether the conservation of such quantum numbers is local, or is longer-range in character. For example, in the case of strangeness, one can ask whether a strange particle and a corresponding antistrange particle tend to be produced “close” to each other within the event, whether strange and antistrange particles are associated with the initial $`\overline{s}`$ and $`s`$ quarks, respectively, in $`s\overline{s}`$ events (or with the initial $`u`$ and $`\overline{u}`$ in $`u\overline{u}`$ events, etc.), or whether strange and antistrange particles are distributed randomly throughout the event. Similar questions can be posed for other relevant quantum numbers. Previous studies of differences in rapidity between associated identified hadrons have shown that the conservation of charge, strangeness and baryon number is predominantly local. Most fragmentation models implicitly implement this feature, and the form and range of such short-range correlations provide useful tests of these models. Short-range correlations can also arise from the decays of heavier hadrons, for example the decay $`\rho ^0\pi ^+\pi ^{}`$ will produce opposite-charge pion pairs with a characteristic degree of locality.
Long-range correlations between particles of opposite charge and strangeness in opposite jets of an event have also been observed . These can be understood in terms of leading particle production whereby the higher-rapidity tracks in each jet tend to carry the quantum numbers of the initial quark or antiquark. In the case of $`e^+e^{}s\overline{s}`$ events, the $`s`$ and $`\overline{s}`$ quarks may hadronize, for example, into a high momentum $`K^{}`$ and $`K^+`$, respectively, and there need be no other strange particles in the event. In $`u\overline{u}`$ and $`d\overline{d}`$ events, however, the locality of quantum number conservation implies that a high-momentum strange-antistrange pair must be produced in each jet, which will dilute any long-range correlation. Nevertheless, improved measurements of long-range correlations may provide better understanding of leading particle production.
The rapidity of a particle is typically defined with an arbitrary sign. If a sign could be given to each measured rapidity such that, for example, positive (negative) rapidity corresponds to the initial quark (antiquark) direction, then one might probe more deeply into both leading and nonleading particle production. One could measure, for example, the extent to which a leading particle has higher rapidity than its associated antiparticle, and the extent to which low-momentum particles in jets remember the initial quark/antiquark direction.
In this paper we present a study of correlations in rapidity between identified charged pions, kaons and protons based on about 550,000 hadronic $`Z^0`$ decays recorded by the SLD experiment at the SLAC Linear Collider. Clean samples of identified particles were obtained using the Cherenkov Ring Imaging Detector. The hadronic event sample was divided into samples enriched in light-flavor ($`e^+e^{}u\overline{u}`$, $`d\overline{d}`$ or $`s\overline{s}`$), $`c\overline{c}`$ and $`b\overline{b}`$ events in order to study the effects of the decays of the leading bottom and charmed hadrons. In section 4 we present a study of short-range correlations between all pair-combinations of these hadron species as a function of hadron momentum. We quantify the range of each correlation, and compare with the predictions of the JETSET fragmentation model. In section 5 we search for long-range correlations between these species. In section 6 we exploit the SLC electron beam polarization to identify the quark (vs. antiquark) hemisphere in each event and sign the rapidities such that particles in the quark-tagged hemisphere have positive rapidity. The signed rapidity distributions themselves provide new information on leading particle production, and ordered rapidity differences between particle pairs allow new and unique probes of the fragmentation process.
## 2 Apparatus and Hadronic Event Selection
A general description of the SLD can be found elsewhere . The trigger and initial selection criteria for hadronic $`Z^0`$ decays are described in Ref. . This analysis used charged tracks measured in the Central Drift Chamber (CDC) and Vertex Detector (VXD) , and identified using the Cherenkov Ring Imaging Detector (CRID) . Momentum measurement is provided by a uniform axial magnetic field of 0.6T. The CDC and VXD give a momentum resolution of $`\sigma _p_{}/p_{}`$ = $`0.010.0026p_{}`$, where $`p_{}`$ is the track momentum transverse to the beam axis in GeV/$`c`$. One quarter of the data were taken with the original vertex detector (VXD2), and the rest with the upgraded detector (VXD3). In the plane normal to the beamline the centroid of the micron-sized SLC IP was reconstructed from tracks in sets of approximately thirty sequential hadronic $`Z^0`$ decays to a precision of $`\sigma _{IP}7`$ $`\mu `$m for the VXD2 data and $``$3 $`\mu `$m for the VXD3 data. Including the uncertainty on the IP position, the resolution on the charged track impact parameter ($`\delta `$) projected in the plane perpendicular to the beamline is $`\sigma _\delta =`$11$``$70/$`(p\mathrm{sin}^{3/2}\theta )`$ $`\mu `$m for VXD2 and $`\sigma _\delta =`$8$``$29/$`(p\mathrm{sin}^{3/2}\theta )`$ $`\mu `$m for VXD3, where $`\theta `$ is the track polar angle with respect to the beamline. The CRID comprises two radiator systems that allow the identification of charged pions with high efficiency and purity in the momentum range 0.3–35 GeV/c, charged kaons in the ranges 0.75–6 GeV/c and 9–35 GeV/c, and protons in the ranges 0.75–6 GeV/c and 10–46 GeV/c . The event thrust axis was calculated using energy clusters measured in the Liquid Argon Calorimeter .
A set of cuts was applied to the data to select well-measured tracks and events well contained within the detector acceptance. Charged tracks were required to have a distance of closest approach transverse to the beam axis within 5 cm, and within 10 cm along the axis from the measured IP, as well as $`|\mathrm{cos}\theta |<0.80`$, and $`p_{}>0.15`$ GeV/c. Events were required to have a minimum of seven such tracks, a thrust axis polar angle w.r.t. the beamline, $`\theta _T`$, within $`|\mathrm{cos}\theta _T|<0.71`$, and a charged visible energy $`E_{vis}`$ of at least 20 GeV, which was calculated from the selected tracks assigned the charged pion mass. The efficiency for selecting a well-contained $`Z^0q\overline{q}(g)`$ event was estimated to be above 96% independent of quark flavor. The VXD, CDC and CRID were required to be operational, resulting in a selected sample of roughly 285,000 events, with an estimated non-hadronic background contribution of $`0.10\pm 0.05\%`$ dominated by $`Z^0\tau ^+\tau ^{}`$ events.
Samples of events enriched in light and $`b`$ primary flavors were selected based on charged track impact parameters $`\delta `$ with respect to the IP in the plane transverse to the beam . For each event we define $`n_{sig}`$ as the number of tracks with impact parameter greater than three times its estimated error, $`\delta >3\sigma _\delta `$. Events with $`n_{sig}=0`$ were assigned to the light-flavor sample and those with $`n_{sig}4`$ were assigned to the $`b`$-flavor sample; the remaining events were classified as a $`c`$-flavor sample. The light-, $`c`$\- and $`b`$-tagged samples comprised 166,000, 82,000 and 37,000 events, respectively; selection efficiencies and sample purities were estimated from our Monte Carlo simulation and are listed in table 1.
Separate samples of hemispheres enriched in quark and antiquark jets were selected by exploiting the large electroweak forward-backward production asymmetry wrt the beam direction. The event thrust axis was used to approximate the initial $`q\overline{q}`$ axis and was signed such that its $`z`$-component was positive, $`\widehat{t}_z>0`$. Events in the central region of the detector, where the production asymmetry is small, were removed by the requirement $`|\widehat{t}_z|>0.2`$, leaving 235,000 events. The quark-tagged hemisphere in events with left-(right-)handed electron beam was defined to comprise the set of tracks with positive (negative) momentum projection along the signed thrust axis. The remaining tracks in each event were defined to be in the antiquark-tagged hemisphere. The sign and magnitude of the electron beam polarization were measured for every event. For the selected event sample, the average magnitude of the polarization was 0.73. Using this value and assuming Standard Model couplings at tree-level, the purity of the quark-tagged sample is 0.73.
For the purpose of estimating the efficiency and purity of the event flavor tagging and the particle identification, we made use of a detailed Monte Carlo (MC) simulation of the detector. The JETSET 7.4 event generator was used, with parameter values tuned to hadronic $`e^+e^{}`$ annihilation data , combined with a simulation of $`B`$-hadron decays tuned to $`\mathrm{{\rm Y}}(4S)`$ data and a simulation of the SLD based on GEANT 3.21 . Inclusive distributions of single-particle and event-topology observables in hadronic events were found to be well described by the simulation .
## 3 Identified Particle Selection
The identification of charged tracks as pions, kaons or protons using the CRID is described in detail in . For this analysis we used a relatively loose identification algorithm, since the presence of misidentified hadrons at the 10% level has little effect on the measured correlations. Tracks with poor CRID information or that were likely to have scattered or interacted before exiting the CRID were removed by requiring each track to have at least 40 CDC hits, at least one of which was at a radius of at least 92 cm, to extrapolate through an active region of the appropriate CRID radiator and through a live CRID TPC, and, in the case of the gas radiator, to have fewer than four saturated hits within the volume in which the gas ring is expected. Approximately 85% of the tracks within the CRID acceptance satisfied these cuts.
Tracks identified in the calorimeters as electrons or muons were rejected, and for the remaining tracks log-likelihoods were calculated for each of the three charged hadron hypotheses $`i=\pi `$, $`K`$ and p, and for each of the liquid and gas radiators. A track was tagged as a hadron of species $`i`$ by the liquid (gas) system if the log-likelihood for hypothesis $`i`$ exceeded both of the other log-likelihoods by at least 5 (3) units. In addition, for those tracks with good information from both the liquid and gas systems, the liquid and gas log-likelihoods were added together, and a track was tagged by the combined system if the log-likelihood for hypothesis $`i`$ exceeded both of the others by at least 3 units. A track was identified as an $`i`$-hadron if it was tagged as type $`i`$ by any of the liquid, gas or combined systems, and it was not tagged as any other type by any other system. The efficiencies for identifying accepted tracks are similar to those given in : for charged pions there is roughly constant efficiency of about 80% in the momentum range 0.5–25 GeV/c; charged kaons and protons have similar efficiency except for a dip in the range 5–10 GeV/c. The simulation was found to provide a good description of the momentum distributions of the identified hadrons.
For each identified track the rapidity $`y=0.5\mathrm{ln}((E+p_{})/(Ep_{}))`$ was calculated using the measured momentum and its projection $`p_{}`$ along the thrust axis, and the appropriate hadron mass. The distributions of rapidity are shown in fig. 1 for each identified hadron species, along with prediction of the simulation. Note that the overall sign of the thrust axis vector, and therefore the sign of the rapidity, is arbitrary. These distributions are not flat in the central region, but show structure due to the momentum dependence of the CRID identification efficiency (see ). The simulation provides a good qualitative description of the rapidity distributions; the dependence of the identification efficiency on momentum has only modest effects on the correlation studies, as discussed below.
The absolute value of the difference between the rapidities of each pair of identified particles was calculated, and the distribution of this quantity is shown in fig. 2 for each of the six possible pairs of hadron species. In each case the distribution for those pairs with opposite charge is shown as the solid histogram and that for pairs with the same charge is shown as the dashed histogram.
## 4 Short-Range Correlations
For every type of hadron pair in fig. 2 there is an excess of opposite-charge pairs over same-charge pairs at small values of the absolute rapidity difference $`|\mathrm{\Delta }y|=|y_1y_2|`$. We expect more opposite-charge than same-charge track pairs, as well as more $`K\overline{K}`$ than $`KK`$/$`\overline{K}\overline{K}`$ pairs and more p$`\overline{\mathrm{p}}`$ than pp/$`\overline{\mathrm{p}}\overline{\mathrm{p}}`$ pairs, due to conservation of electric charge, strangeness and baryon number, respectively. In the case of $`KK`$ (pp) pairs the size of the excess is sensitive to the relative fraction of strange-antistrange (baryon-antibaryon) pairs that are both charged rather than neutral; if e.g. p-$`\overline{\mathrm{n}}`$ were the only type of baryon-antibaryon pair produced, there would be no such excess. The fact that the excess of opposite-charge $`KK`$ and pp pairs peaks at low values of $`|\mathrm{\Delta }y|`$ indicates that conservation of strangeness and baryon number, respectively, is local, as has been observed previously .
In order to study these short-range correlations in more detail, we assumed that the tracks in each same-charge pair are unassociated, and subtracted their $`|\mathrm{\Delta }y|`$ distributions from those of the respective opposite-charge pairs. These differences are shown for the low $`|\mathrm{\Delta }y|`$ region in fig. 3, and are seen to differ significantly from each other in form. Monte Carlo studies indicate that these differences are not due to acceptance, momentum dependence of the particle identification efficiency or to background from misidentified particles. Also shown in fig. 3 are the predictions of the simulation for this difference. In all cases, the simulation gives a reasonable description of the form of the difference, although there are small differences in the details of the form for $`\pi \pi `$, $`\pi `$p, $`KK`$ and $`\pi `$p pairs. The predictions for the amplitudes have $``$5% normalization uncertainties (not shown) due to uncertainties in the particle identification efficiencies. The predicted amplitudes are thus consistent with the data, except that for the $`K`$p correlation, which is low by about 40%.
Pairs of tracks from the decays of resonances (e.g. $`\rho ^0\pi ^+\pi ^{}`$) contribute to varying degrees for all pair combinations; each decay mode gives a $`|\mathrm{\Delta }y|`$ difference distribution with a characteristic form, and collectively they have a substantial influence on the overall form of the distributions. Since the resonance production rates are adjustable in the simulation, we suspect that it is able to provide an adequate description of the data.
Pairs in which one or both particles are misidentified are an important source of background, and lead to a convergence in the forms of the distributions for different pair combinations. The simulated differences for those pairs in which at least one track is misidentified are also shown in fig. 3, and are seen to contribute a sizeable fraction of the observed differences in the cases of unlike particles ($`\pi K`$, $`\pi `$p and $`K`$p). There are normalization uncertainties on these predictions of $``$20% due to uncertainties in the misidentification rates. Excellent particle identification is required to demonstrate, as seen in the figure, that there are excesses at short range for true pairs of these three combinations, indicating local conservation of electric charge between these different particle species and suggesting that there is charge ordering among hadrons of all species in the fragmentation process. The prediction of the simulation for the observed $`K`$p correlation is dominated by misidentification; the prediction for the true correlation is therefore low by a factor of $``$3.
We have studied these short-range correlations in 6 bins of the momentum of the heavier (higher momentum) particle in the case of $`\pi K`$, $`\pi `$p and $`K`$p ($`\pi \pi `$, $`KK`$ and pp) pairs. We find the quality of the simulated description of the data to be independent of the momentum bin. In order to quantify the range of each correlation, we fitted an ad hoc function, the sum of two Gaussians, to each difference in each bin over the range $`0<|\mathrm{\Delta }y|<3`$ units. The centers of both Gaussians were fixed to zero, and the amplitude (width) of the wider Gaussian was fixed to 0.4 (2.2) times that of the narrower Gaussian, leaving the amplitude and width of the narrower Gaussian as free parameters. We used the width as a measure of the range of the correlation. This function provided a reasonable qualitative description of both the data and simulation for all pair combinations in all momentum bins.
The fitted widths are shown in fig. 4 as a function of momentum for each pair combination. If the fragmentation process is scale invariant, then we expect the range to be independent of momentum, except for biases introduced by the bin edges and the particle identification. In the case of $`\pi \pi `$, the range is constant within ten percent except for the highest momentum bin. Significant momentum dependence is observed for $`\pi `$p, $`KK`$, $`K`$p and pp pairs, however this dependence is reproduced by the simulation, with the possible exception of differences at low momentum for $`\pi `$p and $`K`$p pairs, and at high momentum for $`KK`$ pairs. In the simulation, there is a $``$10% momentum dependence due to binning and particle decays. The large slopes and point-to-point structure are due to the momentum dependence of the particle identification efficiencies. Thus the hypothesis of momentum-independence of the range of the short-range rapidity correlation for a given pair combination is consistent with the data, within the context of the JETSET model.
The qualitative results presented above were found to be independent of the primary flavor of the event, as were the quality of the predictions of the simulation and the feature of scale independence. This is expected, as the only difference at short range should be due to the decay properties of the leading $`B`$ and $`D`$ hadrons; we do observe some small differences between the flavor-tagged samples in the measured ranges of the correlations, which are reproduced by the simulation.
## 5 Long-Range Correlations
We next searched for long-range correlations between all pair combinations. In fig. 2, a difference between opposite-charge and same-charge pairs at high $`|\mathrm{\Delta }y|`$ is visible only in the case of $`KK`$, and even here the background from uncorrelated pairs is dominant. Since we expect long-range correlations from leading particle production to be relatively more important at high momentum, we have studied these differences for pairs in which both tracks have momentum $`p>9`$ GeV/c. The corresponding $`|\mathrm{\Delta }y|`$ distributions for each of the six pair combinations are shown in fig. 5. For these high-momentum pairs, there is a separation between pairs in the same jet ($`|\mathrm{\Delta }y|<2.5`$) and those in opposite jets ($`|\mathrm{\Delta }y|>4`$). At short range there is a large excess of opposite-charge pairs over same-charge pairs for all pair combinations, confirming that locality holds even at the highest momenta. At long range, there are clear correlations for $`\pi \pi `$ and $`KK`$ pairs, as well as significant correlations for $`\pi `$p, $`K`$p and pp pairs.
These long-range correlations are found to be strongly flavor-dependent. Figure 6 shows the differences between the opposite-charge and same-charge pairs in the high $`|\mathrm{\Delta }y|`$ region for each of the flavor-tagged samples. The $`b\overline{b}`$ events are seen to contribute very little to the difference for any pair combination, primarily because there are relatively few tracks with such high momentum in these events. Light flavor and $`c\overline{c}`$ events contribute roughly equally to the observed correlations for $`\pi \pi `$, $`\pi `$p and pp pairs; light-flavor events dominate the $`K`$p correlation. In the case of $`\pi `$K pairs, there is a strong anticorrelation in $`c\overline{c}`$ events along with a correlation in light-flavor events, both of which were invisible in the flavor-inclusive sample.
The integrated differences for $`|\mathrm{\Delta }y|>4`$ are given for each flavor-tagged sample in table 2. The predictions of the simulation are also given and are generally consistent with the data. The simulation also describes the rapidity dependence in fig. 5 within the experimental uncertainties.
## 6 Signed Rapidities and Correlations
We next tagged the quark (vs. antiquark) direction in each hadronic event using the electron beam polarization for that event, exploiting the large forward-backward quark production asymmetry in $`Z^0`$ decays. If the beam was left-(right-)handed then the thrust axis was signed such that $`\mathrm{cos}\theta _T`$ was positive (negative). Events with $`|\mathrm{cos}\theta _T|<0.15`$ were removed, as the production asymmetry is small in this region. The probability to tag the quark direction correctly in these events was 73%.
The rapidity of a particle with respect to the signed thrust axis is naturally signed such that positive rapidity corresponds to the hemisphere in the tagged direction of the initial quark, and negative rapidity corresponds to the tagged antiquark hemisphere. The signed rapidity distributions for identified $`K^+`$ and $`K^{}`$ are shown in fig. 7. There is a clear difference between the two, with more $`K^{}`$ than $`K^+`$ in the quark hemisphere, as expected due to leading $`K^{}`$ produced in $`s`$ quark jets . The difference between the two distributions is also shown in the figure and is compared with the prediction of the simulation, which is consistent with the data.
For pairs of identified particles, one can define an ordered rapidity difference. For particle-antiparticle pairs, we define $`\mathrm{\Delta }y^+=y_+y_{}`$ as the difference between the signed rapidities of the positively charged particle and the negatively charged particle. In fig. 8 we show the distribution of $`\mathrm{\Delta }y^+`$ for $`\pi ^+\pi ^{}`$, $`K^+K^{}`$ and p$`\overline{\mathrm{p}}`$ pairs. Asymmetries in these distributions are indications of ordering along the event axis, and the differences between the positive and negative sides of these distributions are also shown. The negative difference at high $`|\mathrm{\Delta }y^+|`$ for the $`K^+K^{}`$ pairs can be attributed to the fact that leading kaons are produced predominantly in $`s\overline{s}`$ events. Similar but smaller effects for expected for $`\pi ^+\pi ^{}`$ and p$`\overline{\mathrm{p}}`$ pairs, and the positive difference in the latter at $`|\mathrm{\Delta }y|4`$ may be attributed to this effect. For $`\pi ^+\pi ^{}`$ pairs we observe a large positive difference at high $`|\mathrm{\Delta }y^+|`$ rather than the expected small negative difference, which is due entirely to $`c\overline{c}`$ events (see below). The predictions of the simulation are also shown and are consistent with the data at high $`|\mathrm{\Delta }y^+|`$.
The positive difference in the lowest $`|\mathrm{\Delta }y^+|`$ bins for the p$`\overline{\mathrm{p}}`$ pairs indicates that the baryon in an associated baryon-antibaryon pair follows the quark direction more closely than the antibaryon. This could be due to leading baryon production and/or to baryon number ordering along the entire fragmentation chain. We find a significant effect in all six of our momentum bins, and that the bulk of the observed difference occurs at low momentum (see fig. 9). We therefore conclude that both of these effects contribute; this is the first direct observation of baryon number ordering along the entire chain. Figure 9 also shows that the effect at long range noted above is confined to high-momentum pairs, as expected if due to leading baryon production. The prediction of the simulation is low by a factor of two at low $`|\mathrm{\Delta }y^+|`$ in both fig. 8 and in the low $`p_{max}`$ side of fig. 9; it is consistent with the high $`p_{max}`$ data.
In fig. 10 we show the difference for $`K^+K^{}`$ pairs in two bins of $`p_{max}`$ for the light- and $`c`$-tagged samples separately. For large $`p_{max}`$ one can see differences in amplitude and in $`|\mathrm{\Delta }y|`$ between the contributions from these two flavor-tagged samples. At low $`p_{max}`$ there is a positive difference of three standard deviations at low $`|\mathrm{\Delta }y|`$ for the light-flavor sample, a possible indication of strangeness ordering along the quark-antiquark axis for associated $`K^+K^{}`$ pairs, similar to the baryon number ordering observed for p$`\overline{\mathrm{p}}`$ pairs. Such an effect is expected to be diluted at high momentum by an effect of the opposite sign due to triplets of high-momentum kaons produced in $`s\overline{s}`$ events. We do not observe such a signal in the heavy flavor events, possibly due to limited statistics, dilution due to heavy hadron decays that include a $`K^+`$ and a $`K^{}`$, and/or dilution due to the negative difference from opposite hemisphere pairs, that populate a lower $`|\mathrm{\Delta }y|`$ region in heavy-flavor events than in light-flavor events.
In fig. 11 we show the difference for $`\pi ^+\pi ^{}`$ pairs in two bins of $`p_{max}`$ for the light-flavor and $`c\overline{c}`$ samples separately. For large $`p_{max}`$ we observe a negative difference at high $`|\mathrm{\Delta }y|`$ in the light-flavor sample, as expected from leading pion production. In the $`c\overline{c}`$ events, there are positive differences in both bins of $`p_{max}`$ at high $`|\mathrm{\Delta }y^+|`$ and also in the low $`p_{max}`$ bin at low $`|\mathrm{\Delta }y^+|`$, all of which can be attributed to pairs involving a $`\pi ^+`$ from the $`D`$ meson decay and a $`\pi ^{}`$ from the $`\overline{D}`$ meson decay. The predictions of the simulation are consistent with these data.
For unlike particles, both opposite-charge and same-charge pairs may be of interest. We define the ordered difference as the rapidity of the heavier particle minus that of the lighter particle multiplied by the charge of the heavier particle. That is, $`\mathrm{\Delta }y^+=y_{K^+}y_\pi ^{}`$ or $`y_\pi ^{}y_{K^+}`$, $`\mathrm{\Delta }y^{++}=y_{K^+}y_{\pi ^+}`$, $`\mathrm{\Delta }y^{}=y_\pi ^{}y_K^{}`$, etc. In fig. 12 we show the distributions of $`\mathrm{\Delta }y^+`$ for opposite-charge pairs of each of the three combinations of unlike particles, as well as the sum of the $`\mathrm{\Delta }y^{++}`$ and $`\mathrm{\Delta }y^{}`$ distributions for the corresponding same-charge pairs. A significant negative asymmetry is observed for $`\pi K`$ pairs of both opposite- and same-charge at all $`\mathrm{\Delta }y`$, which may simply be due the combination of leading kaons and randomly selected pions. A similar effect in $`K`$p pairs can be attributed to leading kaons combined with randomly chosen protons. In this case the asymmetry is negative for opposite-charge pairs and positive for same-charge pairs, as expected given our sign convention. In the case of $`\pi `$p pairs there is a small positive asymmetry for both opposite- and same-charge pairs. The asymmetries predicted by the simulation are also shown; they are consistent with the p$`K^+`$ data, but overestimate the magnitudes of the effects for p$`K^{}`$, $`K^+\pi ^{}`$ and $`K^+\pi ^+`$ pairs slightly. The latter two differences are expected in light of the absence of a $`\pi K`$ correlation in the simulation of light-flavor events noted above. A small negative asymmetry is predicted for $`\pi `$p pairs, which is inconsistent with the observed positive asymmetry.
## 7 Conclusion
We have presented a preliminary study of rapidity differences between pairs of identified charged pions, kaons and protons in light-flavor hadronic $`Z^0`$ decays. The SLD Cherenkov Ring Imaging Detector was used to select clean samples of identified charged hadrons, and the Vertex Detector was used to investigate the flavor dependence of the results.
We observe excesses of opposite-charge over same-charge pairs for all pair combinations at low values of the absolute rapidity difference, indicating that there is a high degree of local conservation of baryon number, strangeness and electric charge in the fragmentation process. The predictions of the JETSET fragmentation model are found to provide a qualitative description of the data, although they fail to describe the forms of some of the correlations in detail. The range of these short-range correlations was studied as a function of momentum; the variations found were reproduced by the simulation, verifying within the context of the JETSET model that the correlations are scale invariant.
We observe a large excess of high momentum $`K^+K^{}`$ pairs over same-charge kaon pairs at large values of the absolute rapidity difference, and the effect is larger for higher momenta, as expected from leading kaon production in $`s\overline{s}`$ events. Weaker correlations are observed for protons, indicating that events with a leading baryon in one jet, a leading antibaryon in the other and no additional baryons do not contribute significantly to baryon production in $`e^+e^{}`$ annihilations. Significant long-range correlations are observed between opposite-charge pairs of all combinations in light-flavor events; in $`c\overline{c}`$ events, long-range $`\pi \pi `$, $`\pi `$p and $`KK`$ correlations are observed, along with a strong $`\pi K`$ anticorrelation. The simulation provides a good description of the data in general, but does not predict the long-range $`\pi K`$ correlation in light-flavor events.
We have studied distributions of rapidity signed so that positive rapidity corresponds to the quark (rather than antiquark) direction. Differences between signed rapidity distributions for positive and negative hadrons of all three species are observed, giving further evidence for leading production of charged pions kaons and protons. The distribution of the difference between the signed rapidities of $`K^+`$ and $`K^{}`$ shows a large asymmetry at large values of the absolute rapidity difference, a direct indication that the long-range correlated $`KK`$ pairs are dominated by $`s\overline{s}`$ events. A similar but smaller difference for $`\pi ^+\pi ^{}`$ pairs indicates roughly equal production of leading pions in $`u\overline{u}`$ and $`d\overline{d}`$ events. There is a large asymmetry at small rapidity difference for p$`\overline{\mathrm{p}}`$ pairs, a clear indication of ordering of baryons along the event axis. A similar effect is observed for $`K^+K^{}`$ pairs at low momentum in light-flavor events.
## Acknowledgements
We thank the personnel of the SLAC accelerator department and the technical staffs of our collaborating institutions for their outstanding efforts on our behalf.
Work supported by Department of Energy contracts: DE-FG02-91ER40676 (BU), DE-FG03-91ER40618 (UCSB), DE-FG03-92ER40689 (UCSC), DE-FG03-93ER40788 (CSU), DE-FG02-91ER40672 (Colorado), DE-FG02-91ER40677 (Illinois), DE-AC03-76SF00098 (LBL), DE-FG02-92ER40715 (Massachusetts), DE-FC02-94ER40818 (MIT), DE-FG03-96ER40969 (Oregon), DE-AC03-76SF00515 (SLAC), DE-FG05-91ER40627 (Tennessee), DE-FG02-95ER40896 (Wisconsin), DE-FG02-92ER40704 (Yale); National Science Foundation grants: PHY-91-13428 (UCSC), PHY-89-21320 (Columbia), PHY-92-04239 (Cincinnati), PHY-95-10439 (Rutgers), PHY-88-19316 (Vanderbilt), PHY-92-03212 (Washington); The UK Particle Physics and Astronomy Research Council (Brunel, Oxford and RAL); The Istituto Nazionale di Fisica Nucleare of Italy (Bologna, Ferrara, Frascati, Pisa, Padova, Perugia); The Japan-US Cooperative Research Project on High Energy Physics (Nagoya, Tohoku); The Korea Research Foundation (Soongsil, 1997).
## <sup>∗∗</sup>List of Authors
Kenji Abe,<sup>(21)</sup> Koya Abe,<sup>(33)</sup> T. Abe,<sup>(29)</sup> I.Adam,<sup>(29)</sup> T. Akagi,<sup>(29)</sup> N. J. Allen,<sup>(5)</sup> W.W. Ash,<sup>(29)</sup> D. Aston,<sup>(29)</sup> K.G. Baird,<sup>(17)</sup> C. Baltay,<sup>(40)</sup> H.R. Band,<sup>(39)</sup> M.B. Barakat,<sup>(16)</sup> O. Bardon,<sup>(19)</sup> T.L. Barklow,<sup>(29)</sup> G. L. Bashindzhagyan,<sup>(20)</sup> J.M. Bauer,<sup>(18)</sup> G. Bellodi,<sup>(23)</sup> R. Ben-David,<sup>(40)</sup> A.C. Benvenuti,<sup>(3)</sup> G.M. Bilei,<sup>(25)</sup> D. Bisello,<sup>(24)</sup> G. Blaylock,<sup>(17)</sup> J.R. Bogart,<sup>(29)</sup> G.R. Bower,<sup>(29)</sup> J. E. Brau,<sup>(22)</sup> M. Breidenbach,<sup>(29)</sup> W.M. Bugg,<sup>(32)</sup> D. Burke,<sup>(29)</sup> T.H. Burnett,<sup>(38)</sup> P.N. Burrows,<sup>(23)</sup> A. Calcaterra,<sup>(12)</sup> D. Calloway,<sup>(29)</sup> B. Camanzi,<sup>(11)</sup> M. Carpinelli,<sup>(26)</sup> R. Cassell,<sup>(29)</sup> R. Castaldi,<sup>(26)</sup> A. Castro,<sup>(24)</sup> M. Cavalli-Sforza,<sup>(35)</sup> A. Chou,<sup>(29)</sup> E. Church,<sup>(38)</sup> H.O. Cohn,<sup>(32)</sup> J.A. Coller,<sup>(6)</sup> M.R. Convery,<sup>(29)</sup> V. Cook,<sup>(38)</sup> R. Cotton,<sup>(5)</sup> R.F. Cowan,<sup>(19)</sup> D.G. Coyne,<sup>(35)</sup> G. Crawford,<sup>(29)</sup> C.J.S. Damerell,<sup>(27)</sup> M. N. Danielson,<sup>(8)</sup> M. Daoudi,<sup>(29)</sup> N. de Groot,<sup>(4)</sup> R. Dell’Orso,<sup>(25)</sup> P.J. Dervan,<sup>(5)</sup> R. de Sangro,<sup>(12)</sup> M. Dima,<sup>(10)</sup> A. D’Oliveira,<sup>(7)</sup> D.N. Dong,<sup>(19)</sup> M. Doser,<sup>(29)</sup> R. Dubois,<sup>(29)</sup> B.I. Eisenstein,<sup>(13)</sup> V. Eschenburg,<sup>(18)</sup> E. Etzion,<sup>(39)</sup> S. Fahey,<sup>(8)</sup> D. Falciai,<sup>(12)</sup> C. Fan,<sup>(8)</sup> J.P. Fernandez,<sup>(35)</sup> M.J. Fero,<sup>(19)</sup> K.Flood,<sup>(17)</sup> R. Frey,<sup>(22)</sup> J. Gifford,<sup>(36)</sup> T. Gillman,<sup>(27)</sup> G. Gladding,<sup>(13)</sup> S. Gonzalez,<sup>(19)</sup> E. R. Goodman,<sup>(8)</sup> E.L. Hart,<sup>(32)</sup> J.L. Harton,<sup>(10)</sup> A. Hasan,<sup>(5)</sup> K. Hasuko,<sup>(33)</sup> S. J. Hedges,<sup>(6)</sup> S.S. Hertzbach,<sup>(17)</sup> M.D. Hildreth,<sup>(29)</sup> J. Huber,<sup>(22)</sup> M.E. Huffer,<sup>(29)</sup> E.W. Hughes,<sup>(29)</sup> X.Huynh,<sup>(29)</sup> H. Hwang,<sup>(22)</sup> M. Iwasaki,<sup>(22)</sup> D. J. Jackson,<sup>(27)</sup> P. Jacques,<sup>(28)</sup> J.A. Jaros,<sup>(29)</sup> Z.Y. Jiang,<sup>(29)</sup> A.S. Johnson,<sup>(29)</sup> J.R. Johnson,<sup>(39)</sup> R.A. Johnson,<sup>(7)</sup> T. Junk,<sup>(29)</sup> R. Kajikawa,<sup>(21)</sup> M. Kalelkar,<sup>(28)</sup> Y. Kamyshkov,<sup>(32)</sup> H.J. Kang,<sup>(28)</sup> I. Karliner,<sup>(13)</sup> H. Kawahara,<sup>(29)</sup> Y. D. Kim,<sup>(30)</sup> M.E. King,<sup>(29)</sup> R. King,<sup>(29)</sup> R.R. Kofler,<sup>(17)</sup> N.M. Krishna,<sup>(8)</sup> R.S. Kroeger,<sup>(18)</sup> M. Langston,<sup>(22)</sup> A. Lath,<sup>(19)</sup> D.W.G. Leith,<sup>(29)</sup> V. Lia,<sup>(19)</sup> C.Lin,<sup>(17)</sup> M.X. Liu,<sup>(40)</sup> X. Liu,<sup>(35)</sup> M. Loreti,<sup>(24)</sup> A. Lu,<sup>(34)</sup> H.L. Lynch,<sup>(29)</sup> J. Ma,<sup>(38)</sup> G. Mancinelli,<sup>(28)</sup> S. Manly,<sup>(40)</sup> G. Mantovani,<sup>(25)</sup> T.W. Markiewicz,<sup>(29)</sup> T. Maruyama,<sup>(29)</sup> H. Masuda,<sup>(29)</sup> E. Mazzucato,<sup>(11)</sup> A.K. McKemey,<sup>(5)</sup> B.T. Meadows,<sup>(7)</sup> G. Menegatti,<sup>(11)</sup> R. Messner,<sup>(29)</sup> P.M. Mockett,<sup>(38)</sup> K.C. Moffeit,<sup>(29)</sup> T.B. Moore,<sup>(40)</sup> M.Morii,<sup>(29)</sup> D. Muller,<sup>(29)</sup> V.Murzin,<sup>(20)</sup> T. Nagamine,<sup>(33)</sup> S. Narita,<sup>(33)</sup> U. Nauenberg,<sup>(8)</sup> H. Neal,<sup>(29)</sup> M. Nussbaum,<sup>(7)</sup> N.Oishi,<sup>(21)</sup> D. Onoprienko,<sup>(32)</sup> L.S. Osborne,<sup>(19)</sup> R.S. Panvini,<sup>(37)</sup> C. H. Park,<sup>(31)</sup> T.J. Pavel,<sup>(29)</sup> I. Peruzzi,<sup>(12)</sup> M. Piccolo,<sup>(12)</sup> L. Piemontese,<sup>(11)</sup> K.T. Pitts,<sup>(22)</sup> R.J. Plano,<sup>(28)</sup> R. Prepost,<sup>(39)</sup> C.Y. Prescott,<sup>(29)</sup> G.D. Punkar,<sup>(29)</sup> J. Quigley,<sup>(19)</sup> B.N. Ratcliff,<sup>(29)</sup> T.W. Reeves,<sup>(37)</sup> J. Reidy,<sup>(18)</sup> P.L. Reinertsen,<sup>(35)</sup> P.E. Rensing,<sup>(29)</sup> L.S. Rochester,<sup>(29)</sup> P.C. Rowson,<sup>(9)</sup> J.J. Russell,<sup>(29)</sup> O.H. Saxton,<sup>(29)</sup> T. Schalk,<sup>(35)</sup> R.H. Schindler,<sup>(29)</sup> B.A. Schumm,<sup>(35)</sup> J. Schwiening,<sup>(29)</sup> S. Sen,<sup>(40)</sup> V.V. Serbo,<sup>(29)</sup> M.H. Shaevitz,<sup>(9)</sup> J.T. Shank,<sup>(6)</sup> G. Shapiro,<sup>(15)</sup> D.J. Sherden,<sup>(29)</sup> K. D. Shmakov,<sup>(32)</sup> C. Simopoulos,<sup>(29)</sup> N.B. Sinev,<sup>(22)</sup> S.R. Smith,<sup>(29)</sup> M. B. Smy,<sup>(10)</sup> J.A. Snyder,<sup>(40)</sup> H. Staengle,<sup>(10)</sup> A. Stahl,<sup>(29)</sup> P. Stamer,<sup>(28)</sup> H. Steiner,<sup>(15)</sup> R. Steiner,<sup>(1)</sup> M.G. Strauss,<sup>(17)</sup> D. Su,<sup>(29)</sup> F. Suekane,<sup>(33)</sup> A. Sugiyama,<sup>(21)</sup> S. Suzuki,<sup>(21)</sup> M. Swartz,<sup>(14)</sup> A. Szumilo,<sup>(38)</sup> T. Takahashi,<sup>(29)</sup> F.E. Taylor,<sup>(19)</sup> J. Thom,<sup>(29)</sup> E. Torrence,<sup>(19)</sup> N. K. Toumbas,<sup>(29)</sup> T. Usher,<sup>(29)</sup> C. Vannini,<sup>(26)</sup> J. Va’vra,<sup>(29)</sup> E. Vella,<sup>(29)</sup> J.P. Venuti,<sup>(37)</sup> R. Verdier,<sup>(19)</sup> P.G. Verdini,<sup>(26)</sup> D. L. Wagner,<sup>(8)</sup> S.R. Wagner,<sup>(29)</sup> A.P. Waite,<sup>(29)</sup> S. Walston,<sup>(22)</sup> J.Wang,<sup>(29)</sup> S.J. Watts,<sup>(5)</sup> A.W. Weidemann,<sup>(32)</sup> E. R. Weiss,<sup>(38)</sup> J.S. Whitaker,<sup>(6)</sup> S.L. White,<sup>(32)</sup> F.J. Wickens,<sup>(27)</sup> B. Williams,<sup>(8)</sup> D.C. Williams,<sup>(19)</sup> S.H. Williams,<sup>(29)</sup> S. Willocq,<sup>(17)</sup> R.J. Wilson,<sup>(10)</sup> W.J. Wisniewski,<sup>(29)</sup> J. L. Wittlin,<sup>(17)</sup> M. Woods,<sup>(29)</sup> G.B. Word,<sup>(37)</sup> T.R. Wright,<sup>(39)</sup> J. Wyss,<sup>(24)</sup> R.K. Yamamoto,<sup>(19)</sup> J.M. Yamartino,<sup>(19)</sup> X. Yang,<sup>(22)</sup> J. Yashima,<sup>(33)</sup> S.J. Yellin,<sup>(34)</sup> C.C. Young,<sup>(29)</sup> H. Yuta,<sup>(2)</sup> G. Zapalac,<sup>(39)</sup> R.W. Zdarko,<sup>(29)</sup> J. Zhou.<sup>(22)</sup>
<sup>(1)</sup>Adelphi University, Garden City, New York 11530, <sup>(2)</sup>Aomori University, Aomori , 030 Japan, <sup>(3)</sup>INFN Sezione di Bologna, I-40126, Bologna Italy, <sup>(4)</sup>University of Bristol, Bristol, U.K., <sup>(5)</sup>Brunel University, Uxbridge, Middlesex, UB8 3PH United Kingdom, <sup>(6)</sup>Boston University, Boston, Massachusetts 02215, <sup>(7)</sup>University of Cincinnati, Cincinnati, Ohio 45221, <sup>(8)</sup>University of Colorado, Boulder, Colorado 80309, <sup>(9)</sup>Columbia University, New York, New York 10533, <sup>(10)</sup>Colorado State University, Ft. Collins, Colorado 80523, <sup>(11)</sup>INFN Sezione di Ferrara and Universita di Ferrara, I-44100 Ferrara, Italy, <sup>(12)</sup>INFN Lab. Nazionali di Frascati, I-00044 Frascati, Italy, <sup>(13)</sup>University of Illinois, Urbana, Illinois 61801, <sup>(14)</sup>Johns Hopkins University, Baltimore, MD 21218-2686, <sup>(15)</sup>Lawrence Berkeley Laboratory, University of California, Berkeley, California 94720, <sup>(16)</sup>Louisiana Technical University - Ruston,LA 71272, <sup>(17)</sup>University of Massachusetts, Amherst, Massachusetts 01003, <sup>(18)</sup>University of Mississippi, University, Mississippi 38677, <sup>(19)</sup>Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, <sup>(20)</sup>Institute of Nuclear Physics, Moscow State University, 119899, Moscow Russia, <sup>(21)</sup>Nagoya University, Chikusa-ku, Nagoya 464 Japan, <sup>(22)</sup>University of Oregon, Eugene, Oregon 97403, <sup>(23)</sup>Oxford University, Oxford, OX1 3RH, United Kingdom, <sup>(24)</sup>INFN Sezione di Padova and Universita di Padova I-35100, Padova, Italy, <sup>(25)</sup>INFN Sezione di Perugia and Universita di Perugia, I-06100 Perugia, Italy, <sup>(26)</sup>INFN Sezione di Pisa and Universita di Pisa, I-56010 Pisa, Italy, <sup>(27)</sup>Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX United Kingdom, <sup>(28)</sup>Rutgers University, Piscataway, New Jersey 08855, <sup>(29)</sup>Stanford Linear Accelerator Center, Stanford University, Stanford, California 94309, <sup>(30)</sup>Sogang University, Seoul, Korea, <sup>(31)</sup>Soongsil University, Seoul, Korea 156-743, <sup>(32)</sup>University of Tennessee, Knoxville, Tennessee 37996, <sup>(33)</sup>Tohoku University, Sendai 980, Japan, <sup>(34)</sup>University of California at Santa Barbara, Santa Barbara, California 93106, <sup>(35)</sup>University of California at Santa Cruz, Santa Cruz, California 95064, <sup>(36)</sup>University of Victoria, Victoria, B.C., Canada, V8W 3P6, <sup>(37)</sup>Vanderbilt University, Nashville,Tennessee 37235, <sup>(38)</sup>University of Washington, Seattle, Washington 98105, <sup>(39)</sup>University of Wisconsin, Madison,Wisconsin 53706, <sup>(40)</sup>Yale University, New Haven, Connecticut 06511.
|
no-problem/9908/astro-ph9908274.html
|
ar5iv
|
text
|
# On the nature of the compact star in 4U 1728-34
## 1 Introduction
With the Rossi X-ray Timing Explorer kHz QPOs have been discovered in about 20 neutron star LMXBs (see van der Klis vk98 (1998) for a review). In many cases, two simultaneous kHz peaks ($`2001200`$ Hz) are observed in the power spectra of the X-ray count rate variations, with the separation frequency $`\mathrm{\Delta }\nu `$ roughly constant (e.g., Strohmayer et al. s96 (1996); Ford et al. f97 (1997); Wijnands et al. w97 (1997)). Sometimes a third kHz peak is detected in a few atoll sources during type I X-ray bursts at a frequency $`\nu _\mathrm{b}`$ equal to the separation frequency $`\mathrm{\Delta }\nu `$ of the two peaks or twice that (see Strohmayer, Swank, & Zhang ssz98 (1998) for a review). These two results suggest that a beat-frequency mechanism is at work, with the upper kHz peak at the Keplerian orbital frequency at the inner edge of the accretion disk around the neutron star, the third peak at the neutron star spin frequency (or twice that), and the lower kHz peak at the difference frequency between them (Strohmayer et al. s96 (1996); Miller, Lamb, & Psaltis mlp98 (1998)).
Within the beat-frequency model, the burst oscillations $`\nu _\mathrm{b}`$ and the kHz peak separations $`\mathrm{\Delta }\nu `$ are both believed to be close to the neutron star spin frequency and thus $`\mathrm{\Delta }\nu `$ should remain constant. However, recent observations have challenged this interpretation: in Sco X-1 (van der Klis et al. vk97 (1997)), 4U 1608-52 (Méndez et al. m98 (1998)), and 4U 1735-44 (Ford et al. f98 (1998)), $`\mathrm{\Delta }\nu `$ slowly decreases as the frequencies of both QPOs increase. Moreover, in 4U 1636-53 the frequeny $`\nu _\mathrm{b}`$ of the burst oscillations (or half this value) does not match the frequency difference $`\mathrm{\Delta }\nu `$ between the kHz QPOs (Méndez, van der Klis, & van Paradijs mkp98 (1998)). In the atoll source 4U 1728-34, Méndez & van der Klis (mk99 (1999)) find that $`\mathrm{\Delta }\nu `$ is always significantly smaller than $`\nu _\mathrm{b}`$, even at the lowest inferred mass accretion rate, when $`\mathrm{\Delta }\nu `$ seems to reach its maximum value, and that $`\mathrm{\Delta }\nu `$ decreases significantly as the frequency of the lower kHz QPO increases.
A different approach was recently invoked by Osherovich & Titarchuk (1999a ) and Titarchuk & Osherovich (to99 (1999)), who proposed a unified classification of kHz QPOs and the related low frequency phenomena. In this model, kHz QPOs are modeled as Keplerian oscillations under the influence of the Coriolis force in a rotating frame of reference (magnetosphere). The frequencies $`\nu _\mathrm{u}`$ of the upper kHz QPO branch hold a hybrid frequency relation with the Keplerian frequencies $`\nu _\mathrm{k}`$ (referred to the lower kHz QPO branch): $`\nu _\mathrm{u}^2=\nu _\mathrm{k}^2+(2\nu _\mathrm{m})^2`$, where $`\nu _\mathrm{m}`$ is the rotational frequency of the star’s magnetosphere. For Sco X-1, the QPOs with frequencies $`45`$ and 90 Hz are interpreted as the 1st and 2nd harmonics of the lower branch of the Keplerian oscillations (Osherovich & Titarchuk 1999a ). The same interpretation is applied to the $`35`$ Hz QPOs in the atoll source 4U 1702-42 (Osherovich & Titarchuk 1999b ). The observed low Lorentzian frequency in 4U 1728-34 is suggested to be associated with radial oscillations in the boundary layer of the disk, whereas the observed break frequency is determined by the characteristic diffusion time of the inward motion of the matter in the accretion flow (Titarchuk & Osherovich to99 (1999)). Predictions of this model regarding relations between the QPO frequencies mentioned above compare favorably with recent observations of Sco X-1, 4U 1608-52, 4U 1702-429, and 4U 1728-34.
The discovery of kHz QPO features allows to probe not only the interaction between the accretion disk and the stellar magnetosphere, but also the structure of the compact stars involved. In this Letter, based on the work of Osherovich & Titarchuk (1999a ) and Titarchuk & Osherovich (to99 (1999)), we study the mass - radius ($`MR`$) relation of the atoll X-ray source 4U 1728-34, and show that it is possibly a strange star.
## 2 The mass - radius relation for 4U 1728-34
Titarchuk & Osherovich (to99 (1999)) suggest a model for the radial oscillations and diffusion in the viscous boundary layer of the accretion disk in LMXBs. Their dimensional analysis has identified the corresponding frequencies which are consistent with the low Lorentzian and break frequencies for 4U 1728-34 observed by Ford & van der Klis (fk98 (1998)), the predicted values for the break frequency related to the diffusion in the boundary layer are also consistent with the observed ones for the same source. The presence of the break frequency and the correlated Lorentzian frequency suggests the introduction of a new scale in the phenomenon. One attractive feature of the model is the introduction of such a scale in the model through the Reynolds number. The best fit for the observed data was obtained when
$$a_\mathrm{k}=(M/M_{})(R_0/3R_\mathrm{s})^{3/2}(\nu /364\mathrm{Hz})=1.03,$$
(1)
where $`M`$ is the stellar mass, $`R_0`$ is the inner edge of the accretion disk, $`R_\mathrm{s}=2GM/c^2`$ is the Schwarzschild radius, and $`\nu `$ is the spin frequency of the star. Given the 364 Hz spin frequency of 4U 1728-34 (Strohmayer et al. s96 (1996)), from Eq. (1) we derive the inner disk radius
$$R_09(a_\mathrm{k}/1.03)^{2/3}m^{1/3}\mathrm{km},$$
(2)
where $`m=M/M_{}`$. Since the innermost radius of the disk must be larger than the radius of the star itself, this leads to a mass-dependent upper bound on the stellar radius,
$$R\stackrel{<}{}9(a_\mathrm{k}/1.03)^{2/3}m^{1/3}\mathrm{km},$$
(3)
which is plotted in dot-dashed curve in Fig. 1. The dotted curve in Fig. 1 represents the Schwarzschild radius, which presents the lower limit of $`R`$.
Figure 1 compares Eq. (3) with the theoretical $`MR`$ relations (solid curves) for nonrotating neutron stars given by six recent realistic models for the equation of state (EOS) of dense matter. In models UU (Wiringa, Fiks, & Fabrocini wff88 (1988)), BBB1 and BBB2 (Baldo, Bombaci, & Burgio bbb97 (1997)), the neutron star core is assumed to be composed by an uncharged mixture of neutrons, protons, electrons and muons in equilibrium with respect to the weak interaction ($`\beta `$-stable nuclear matter). These models are based on microscopic calculations of asymmetric nuclear matter by use of realistic nuclear forces which fit experimental nucleon-nucleon scattering data, and deuteron properties. In model Hyp (Prakash et al. p97 (1997)), hyperons are considered in addition to nucleons as hadronic constituents of the neutron star core. The curve labeled BPAL12 is a very soft EOS for $`\beta `$-stable nuclear matter (Prakash et al. p97 (1997)), giving a lower limiting mass and a smaller radius with respect to a stiff EOS. Finally, we consider the possibility that neutron stars may possess a core with a Bose - Einstein condensate of negative kaons (e.g. Glendenning & Schaffner-Bielich gs99 (1999)). The main physical effect of the onset of K<sup>-</sup> condensation is a softening of the EOS with a consequent lowering of the neutron star maximum mass and possibly of the radius. This is shown with the curve labeled K<sup>-</sup>. It is clearly seen that only the upper part of EOSs UU, BBB1, BBB2, and BPAL12 are compatible with Eq. (3).
Another constraint on the mass and radius of 4U 1728-34 results from the requirement that the inner radius $`R_0`$ of the disk must be larger than the last stable circular orbit $`R_{\mathrm{ms}}`$ around the star, which is given by (Bardeen, Press, & Teukolsky bpt72 (1972)),
$$R_{\mathrm{ms}}=R_\mathrm{h}(3+Z_2[(3Z_1)(3+Z_1+2Z_2)]^{1/2}),$$
(4)
where
$$Z_1=1+[1(a/R_\mathrm{h})^2]^{1/3}[(1+a/R_\mathrm{h})^{1/3}+(1a/R_\mathrm{h})^{1/3}],$$
(5)
$$Z_2=[3(a/R_\mathrm{h})^2+Z_1^2]^{1/2},$$
(6)
with $`R_\mathrm{h}=GM/c^2`$ and $`a=I\mathrm{\Omega }/Mc`$ (where $`I`$ is the moment of inertia and $`\mathrm{\Omega }=2\pi \nu `$ is the angular velocity of the star, respectively). For a given mass $`M`$, from Eq. (2) we can find $`R_0`$. From the inequality $`R_0`$$`\stackrel{>}{}`$$`R_{\mathrm{ms}}`$ and Eqs. (4-6), one can calculate the lower limit of $`a`$, and from there the lower limit of the moment of inertia $`I`$. To account for the inflation of the radius caused by rotation, we adopt the formula
$$I=0.21MR^2(1R_\mathrm{s}/R)^1,$$
(7)
proposed by Burderi et al. (b99 (1999)), to obtain the lower limit on the stellar radius $`R`$ <sup>1</sup><sup>1</sup>1There are actually two solutions of Eq. (7) with given $`I`$ and $`M`$, one serves as the upper limit of $`R`$ and the other the lower limit of $`R`$. But the former solution gives an unphysically small radius, and is rejected.. The results are shown in the dashed curve in Fig. 1. Combining the constraints $`R\stackrel{<}{}R_0`$ and $`R_0`$$`\stackrel{>}{}`$$`R_{\mathrm{ms}}`$ reveals that the allowed distribution of the mass and radius of 4U 1728-34 must lie between the dot-dashed curve and the dashed curve. None of the neutron star EOSs extends to this region. Thus it seems to rule out the possibility that 4U 1728-34 is a neutron star! Including rotational effects will shift the theoretical neutron star $`MR`$ curves to up-right in Fig. 1, making the contrast between the theoretical neutron star models and the derived mass and radius range of 4U 1728-34 even worse.
The difficulty of the neutron star model for 4U 1728-34 suggests that it may belong to another type of compact objects, that is, strange stars, which are entirely made of deconfined u,d,s quark matter (strange matter). The possible existence of strange stars is a direct consequence of the conjecture (Bodmer b71 (1971); Witten w84 (1984)) that strange matter may be the absolute ground state of strongly interacting matter. Detailed studies have shown that the existence of strange matter is allowable within uncertainties inherent in a strong interaction calculation (Farhi & Jaffe fj84 (1984)); thus strange stars may exist in the universe.
Most of the previous calculations (e.g. Alcock, Farhi, & Olinto afo86 (1986); Haensel, Zdunik, & Schaeffer hzs86 (1996)) of strange star properties used an EOS for strange matter based on the phenomenological nucleonic bag model, in which the basic features of quantum chromodynamics, such as quark confinement and asymptotic freedom are postulated from the beginning, though the deconfinement of quarks at high density is not obvious in the bag model. Recently, Dey et al. (d98 (1998)) derived an EOS for strange matter, which has asymptotic freedom built in, shows confinement at zero baryon density, deconfinement at high density, and gives a stable configuration for chargeless, $`\beta `$-stable strange matter. In this model the quark interaction is described by an interquark vector potential originating from gluon exchange, and by a density dependent scalar potential which restores the chiral symmetry at high density. Using the same model (but different values of the parameters with respect to those employed in Dey et al. d98 (1998)) we calculated the $`MR`$ relations, which are also shown in solid curves labeled ss1 and ss2 in Fig. 1, corresponding to strange stars with maximum masses of $`1.44M_{}`$ and $`1.32M_{}`$ and radii of 7.07 km and 6.53 km, respectively. It is seen that the region confined by the dashed and dot-dashed curves for 4U 1728-34 in Fig. 1 is in remarkable accord with the strange star models with masses $`\stackrel{<}{}1.1M_{}`$. Figure 1 clearly demonstrates that a strange star model is more compatible with 4U 1728-34 than a neutron star one. This seems to be important since the model of Dey et al. (d98 (1998)) employs quark masses which decrease with increasing density, a manifestation of chiral symmetry restoration (CSR), and the parametrization of CSR can be deduced from the mass and radius of the strange star fitted.
The lower limit on the moment of inertia deduced above does not depend on the EOS. In particular Eq. (7) is an extrapolation from neutron stars and it may be doubted whether it is valid for strange stars. We have checked that the standard calculated moment of inertia of the strange star of mass $`M1.1M_{}`$ agrees with Eq. (7) with an error of about $`20\%`$, and the numerical factor 0.21 can change to 0.28 (ss1) or 0.26 (ss2) depending on the model parameters. The moment of inertia calculation of the strange star with variable density layers may not in fact satisfy the assumptions of the standard calculation. It is satisfying to find that an EOS-independent estimate can already reproduce the inertia to a reasonable accuracy.
## 3 Discussion and conclusion
We have shown that strange stars are more consistent with the properties of 4U 1728-34 compared to neutron stars, if the model of Osherovich & Titarchuk (1999a ) and Titarchuk & Osherovich (to99 (1999)) correctly interprets the high- and low-frequency QPO phenomena in 4U 1728-34. Since our analysis does not involve any unknown physical factors such as the magnetic field strength and topology, radiation processes, and the mass accretion rate, it seems to provide the first definite evidence for the existence of strange stars. It remains to see whether other compact stars in LMXBs would be identified as strange stars from observations of their QPO phenomena. If so, there will be very deep consequences for both physics of strong interactions and astrophysics.
Of course our conclusion is subject to the uncertainty in the parameter $`a_\mathrm{k}`$, which is determined by the best fitting of the observed data. Increase or decrease in $`a_\mathrm{k}`$ will enlarge or reduce the allowed region of the mass and radius of 4U 1728-34 in Fig. 1. For example, if $`a_\mathrm{k}=2`$, nearly all neutron star EOSs are compatible with 4U 1728-34. However, according to Titarchuk & Osherovich (to99 (1999)), the dependence of $`\chi ^2`$ on $`a_\mathrm{k}`$ is quite strong: $`\chi ^2=3802473076a_\mathrm{k}+35732a_\mathrm{k}^235732[(a_\mathrm{k}1.03)^2+0.019]`$. A $`15\%`$ variation of $`a_\mathrm{k}`$ from the preferred value 1.03 will double the value of $`\chi ^2`$, greatly reducing the confidence level of the fit. Thus we expect that the possible change in $`a_\mathrm{k}`$ and in the derived $`MR`$ relation is rather small.
If 4U 1728-34 is a strange star, Fig. 1 reveals that it possesses a mass $`\stackrel{<}{}1.1M_{}`$, which is quite small compared to the $`1.4M_{}`$ masses of standard neutron stars like PSR 1913+16, though still barely consistent with the estimated masses of quite a few radio pulsars within $`68\%`$ confidence limit (see Thorsett & Chakrabarty tc99 (1999)). Since the mass accretion rates in LMXB atoll sources are lower than the Eddington limit accretion rate by $`12`$ orders of magnitude, it is conventionally thought that in such systems almost all of the matter transferred from the companion stars should be accreted by the compact stars. The light mass of 4U 1728-34, however, seems to support the conclusion of Thorsett & Chakrabarty (tc99 (1999)) that there is no evidence for extensive mass accretion occurred in LMXBs, implying that even the simplest evolution of LMXBs is far from being fully understood. However, formation of strange stars may be different from that of neutron stars (i.e. type II/Ib supernova explosions) concerning the glitching phenomena observed in some radio pulsars (e.g. Alpar a87 (1987)), suggesting that some exotic formation channels may be at work. Accreting massive neutron stars in binary systems are not favorable progenitor candidates of strange stars, at least in the case of 4U 1728-34, since it requires too large mass ($`\stackrel{>}{}0.7M_{}`$) ejected during the conversion as the neutron star masses exceed $`\stackrel{>}{}1.8M_{}`$. Accretion-induced collapse of white dwarfs remains an open possibility of strange star formation, and could be responsible for the light mass of 4U 1728-34. The detailed investigation of this subject is beyond the scope of this paper and wil be discussed elsewhere.
We are grateful to Lev Titarchuk for helpful and prompt e-mails. X. L. was supported by National Natural Science Foundation of China. S. R., M. D. and J. D. were supported in part by DST grant no. SP/S2/K18/96, Govt. of India.
|
no-problem/9908/cond-mat9908046.html
|
ar5iv
|
text
|
# Electron spin resonance study of Na1-xLixV2O5
## I Introduction
Since 1996, when Isobe and Ueda first reported the observation of an exponential decrease of the susceptibility in NaV<sub>2</sub>O<sub>5</sub> below 34 K, this material has been subject of intense investigation. The transition was first considered to be a spin-Peierls transition similar to that observed in CuGeO<sub>3</sub> . This assumption was based on an early determination of the structure by Carpy et al. , who proposed alternating chains of V<sup>4+</sup> (spin 1/2) and nonmagnetic V<sup>5+</sup>. This picture was able to explain the physical properties above the transition, like the susceptibility that closely follows that of a one-dimensional spin 1/2 Heisenberg antiferromagnet as calculated by Bonner and Fisher or more recently by Eggert et al. . It could not explain most of the experimental findings connected with the transition itself nor the low-temperature state: the ratio of the energy gap $`\mathrm{\Delta }(0)`$ to the transition temperature was found to be much larger than the expected mean-field value of $`2\mathrm{\Delta }/k_\mathrm{B}T_{\mathrm{SP}}=3.53`$ ; the entropy of the jump in the specific heat is also much higher than expected ; and in thermal-expansion measurements two transitions close to each other were observed . In the low-temperature phase satellite reflections were reported in X-ray measurements corresponding to a doubling of the unit cell in $`a`$ and $`b`$ and a quadrupling in $`c`$ direction .
However, recent structural investigations have shown that instead of the originally proposed non-centrosymmetric space group $`P2_1mn`$, the structure of NaV<sub>2</sub>O<sub>5</sub> at room temperature has to be described by the centrosymmetric space group $`Pmmn`$. In this structure only one kind of vanadium sites exists with an average vanadium valence of V<sup>+4.5</sup>. NaV<sub>2</sub>O<sub>5</sub> can therefore be regarded as a quarter-filled ladder system with one electron per rung. This excludes the possibility of a simple spin-Peierls transition in this material. The occurrence of a charge-ordering transition followed by a dimerization is discussed . Different types of low temperature structures were proposed. Whereas theoretical models mainly discuss an inline or a zig-zag ordering, a recent determination of the low-temperature structure suggests a separation into modulated and unmodulated vanadium ladders .
The first ESR measurements of NaV<sub>2</sub>O<sub>5</sub> were carried out in 1986 by Ogawa et al.. Due to a large Curie contribution in the susceptibility they did not observe the characteristic decrease below $`34`$K. The discovery of the transition by Isobe and Ueda stimulated many other ESR studies in this compound . In this article we present electron-spin resonance (ESR) results of single crystalline Na<sub>1-x</sub>Li<sub>x</sub>V<sub>2</sub>O<sub>5</sub> for $`x=0,0.15\%,0.3\%,0.5\%,0.7\%`$, $`0.9\%`$, and $`1.3\%`$ in the temperature range 4.2 K – 700 K. We discuss the ESR linewidth and the signal intensity that is directly proportional to the spin susceptibility. Assuming a mean-field like dependence of the energy gap $`\mathrm{\Delta }(T)`$ that opens below the transition, we determine the value of the energy gap at zero temperature and the transition temperature as a function of the lithium concentration.
## II Sample preparation and experiment
The samples were small single crystals, prepared from a NaVO<sub>3</sub> flux . In a first step a mixture of Na<sub>2</sub>CO<sub>3</sub> and V<sub>2</sub>O<sub>5</sub> is heated up to 550 C in air to form NaVO<sub>3</sub>. In a second step the NaVO<sub>3</sub> is mixed with VO<sub>2</sub> in the ratio of 8:1 and then heated up to 800 C in an evacuated quartz tube and cooled down at a rate of 1 K per hour. The excess NaVO<sub>3</sub> was dissolved in water. The doped Samples were produced by substituting in the first step Na<sub>2</sub>CO<sub>3</sub> by Li<sub>2</sub>CO<sub>3</sub>. However, due to a low distribution coefficient during the flux growth process, the real amount of Li in the sample is much lower. The real cation composition was determined in two doped samples using inductive coupled plasma for the V content and atomic absorption spectroscopy for the Li and Na content (see table I). The result shows that the real Li content is a factor of 7.5 lower than the nominal one. For the other samples the Li concentration was scaled accordingly, as given in table I. All the samples were investigated using X-ray powder diffraction. Only at high Li-content, a small decrease of the c lattice parameter was observed.
The ESR measurements were performed using a Bruker Elexsys 500 CW spectrometer at X-band frequency (9.48 GHz). In the temperature range 4.2 – 300 K a continuous flow He-cryostat (Oxford Instruments) and between 300 K and 700 K a nitrogen cryostat (Bruker) was used. The samples were orientated in a way that the applied external field was always perpendicular to the crystallographic $`b`$ axis and could be rotated about this axis. All measurements were made at the orientation with the narrowest resonance line, i.e. the external field $`H`$ being parallel to the $`a`$ axis.
## III Electron-spin resonance
NaV<sub>2</sub>O<sub>5</sub> shows one single lorentzian-shaped resonance line with an anisotropic $`g`$ value between 1.976 ($`H`$ parallel $`a`$ axis) to 1.936 ($`H`$ parallel $`c`$ axis) . At high temperatures the linewidth of this resonance decreases monotonically with decreasing temperature and is independent from lithium doping as shown in the inset of figure 1 for the undoped and the 0.7 % lithium doped sample. Below 34 K the linewidth increases again. This increase was found to be rather strongly suppressed by doping (figure 1). While the linewidth in the undoped sample increases by a factor of 4 from 34 K down to 15 K, for 1.3 % lithium content the increase is only about 40 %. This clearly indicates that the increase of the linewidth below 34 K is directly connected to the transition, which is suppressed upon lithium doping as will be shown below. In the whole temperature range the ESR signal is strongly exchange narrowed and no hyperfine structure due to the <sup>51</sup>V-spin ($`I=7/2`$) is observed .
We therefore propose that the broadening of the linewidth below the transition appears because the exchange narrowing becomes less effective, probably due charge localisation.
A similar overall temperature dependence of the linewidth is observed in CuGeO<sub>3</sub> . Yamada et al. qualitatively explained the high-temperature behavior in both CuGeO<sub>3</sub> and NaV<sub>2</sub>O<sub>5</sub> by identifying the anisotropic Dzyaloshinsky-Moriya exchange interaction $`H_{\mathrm{DM}}`$ as the dominating interaction responsible for the linebroadening . The Dzyaloshinsky-Moriya interaction is given by $`_id_{ii+1}(𝐒_i\times 𝐒_{i+1})`$ for neighboring spins $`𝐒`$, were $`d_{ii+1}`$ can be estimated as $`d_{ii+1}(\mathrm{\Delta }g/g)|J|`$ . We found that both $`g`$ value and exchange coupling constant $`J`$ (that can be determined from the spin susceptibility, see fig. 2a) remain nearly unaffected by doping. This is consistent with the fact that no concentration dependence of the linewidth was detected at high temperatures.
We also determined the spin susceptibility of Na<sub>1-x</sub>Li<sub>x</sub>V<sub>2</sub>O<sub>5</sub> from the intensity of the ESR signal. Since it is difficult to determine the absolute values of the susceptibility by ESR, only relative values are given, the curves being scaled to one at 300 K. An estimation of the absolute intensity is consistent with one vanadium per formular unit contributing to the signal. As mentioned before, the spin susceptibility above the transition is nearly insensitive to lithium doping. In figure 2a the undoped sample is compared with the 0.7 % lithium doped sample. For $`T>200`$ K both curves nicely agree with the theoretical fit using the dependence calculated by Bonner and Fisher or Eggert et al. with $`J=578`$ K. Both calculations give the same results above $`T=0.3J175`$ K. Below this temperature the more exact calculation of Eggert et al. shows an even more pronounced disagreement with the data. The reason for this deviation is not totally resolved. It could be due to a dimensional crossover as was suggested from X-ray investigations (Ravy et al. predict a deviation from the Bonner-Fisher theory up to temperatures much higher than 90 K) or due to the existence of structural fluctuations.
Figure 2b displays the spin susceptibility below 60 K for different lithium concentrations. The transition shifts to lower temperatures and the decrease of the susceptibility becomes less pronounced with increasing lithium content. We also observe a Curie like increase at lowest temperatures that increases with doping. In the sample Na<sub>1-x</sub>Li<sub>x</sub>V<sub>2</sub>O<sub>5</sub> with $`x=1.3\%`$ the transition is no longer visible (see figure 2b).
To analyse the data, a Curie law was fitted to the data points below 10 K and subtracted. The curves were then analyzed using a mean-field like temperature dependence of the energy gap and $`\chi (T)\mathrm{exp}(2\mathrm{\Delta }/\mathrm{k}_\mathrm{B}\mathrm{T})`$. For the temperature dependence of the energy gap $`\mathrm{\Delta }(T)`$ the exact mean-field values were taken; $`\mathrm{\Delta }(0)`$, and $`T_{\mathrm{SP}}`$ being the only fitting parameters. In this case it is preferable to use this method rather than fitting with the theory of Bulaevskii because the uncertainty at low temperatures caused by the Curie contribution strongly influences the determination of the energy gap $`\mathrm{\Delta }(0)`$. Examples of the fitting procedure for different $`x`$ are given in figure 3. In the samples with $`x5\%`$ perfect agreement of the data and the fitting curves is found. The transition is broadened with increasing lithium content thus causing an increasing uncertainty for the high doped samples $`x=0.9\%`$ and $`x=1.3\%`$. While a determination of $`\mathrm{\Delta }(0)`$, and $`T_{\mathrm{SP}}`$ is still possible in the $`x=0.9\%`$ lithium doped sample, in the 1.3% doped sample no clear choice of $`\mathrm{\Delta }(0)`$ and $`T_{\mathrm{SP}}`$ could be made, because the phase transition is strongly broadened in temperature and it is not clear how to determine the Curie contribution exactly (if the data are treated like those of the other samples assuming that a low temperature only the Curie contribution exists this contribution is probably overestimated leading to a seemingly linear decrease of the susceptibility as shown in figure 3).
The results for the transition temperature $`T_{\mathrm{SP}}`$ and the energy gap $`\mathrm{\Delta }(0)`$ are displayed in figure 4. The transition temperature is seems to follow a $`T_{\mathrm{SP}}abx^2`$ function (dashed line). The energy gap $`\mathrm{\Delta }(0)`$ varies linearly with the lithium content. However, since the errors in the determination of the lithium content have to be taken into account, further investigation is necessary to confirm the exact dependencies. For both cases the value of the assumed functions differs from zero (i.e. no transition occurs) at $`x=1.3\%`$ lithium. This suggests that even in the case of 1.3 % lithium doping the transition is not completely suppressed. Another interesting result is that the ratio $`2\mathrm{\Delta }/k_\mathrm{B}T_{\mathrm{SP}}`$ decreases from the strong coupling value of 5–6 in undoped NaV<sub>2</sub>O<sub>5</sub> to values close to the mean field result of 3.53, i.e. 3.7–4 in the samples with $`x=0.5\%`$ and $`x=0.7\%`$.
## IV Conclusions
In conclusion we have presented ESR results on Na<sub>1-x</sub>Li<sub>x</sub>V<sub>2</sub>O<sub>5</sub> for $`0x1.3\%`$. The linewidth and the spin susceptibility above the transition were found to be nearly independent from the lithium concentration. At low temperatures the increase of the linewidth is suppressed with growing lithium content. The spin susceptibility was analyzed using a mean-field like model to extract the transition temperatures and the $`T=0`$ value of the energy gap with respect to the doping. It was found that the transition temperature and the energy gap decrease monotonically on increasing Li concentration, suggesting a square dependence of the transition temperature and a linear decrease of the energy gap. Considering these dependencies it is highly probable that even in the highest doped sample a transition still persists.
Although there is no theoretical prediction for the suppression of the transition upon doping in NaV<sub>2</sub>O<sub>5</sub>, one can speculate about the relevant physical properties. The lattice parameters (table 1) show only a slight doping dependence. It is consequently very improbable that the suppression of the transition can be explained with the change of the lattice. In a normal spin-Peierls system the transition depends on the spin-phonon coupling $`g`$ and the phonon frequency $`\omega `$ . The transition temperature should be in the order of $`g/\omega ^2`$. Substitution of the lighter lithium ions for sodium is expected to increase the phonon frequency $`\omega `$ thus reducing the transition temperature. This scenario could explain the monotonic decrease of the transition temperature upon doping. In this context a direct observation of the phonon frequencies in lithium-doped samples would be very interesting.
In Na<sub>1-x</sub>Li<sub>x</sub>V<sub>2</sub>O<sub>5</sub> the lithium ions are located on the off-chain sodium positions. In contrast to CuGeO<sub>3</sub> doped off-chain with silicon , where antiferromagnetic order appears for concentrations as low as 0.5%, no signs of magnetic order were found . In CuGe<sub>1-x</sub>Si<sub>x</sub>O<sub>3</sub> the spin-Peierls transition decreases linearly as $`T_{\mathrm{SP}}(x)abx`$ . While in CuGeO<sub>3</sub> off-chain substitutions (like Si ) and in-chain substitutions (like Zn or Mg ) have been extensively studied, in NaV<sub>2</sub>O<sub>5</sub> many interesting work in this field remains to be done.
We gratefully acknowledge helpful discussion with A. Kampf. This work was partly supported by BMFT under contract no. 13N6917/0 and DFG under contract no. 20 264/10-1.
|
no-problem/9908/astro-ph9908183.html
|
ar5iv
|
text
|
# Evidence for Large-Scale Structure at 𝑧≈2.4 From Lyman 𝛼 Imaging
## 1 Introduction
As our ability to measure galaxy evolution has grown, so has the possibility of observing how galaxy clustering has evolved. On various linear scales, this is relevant to the merger rate of both galaxies (and perhaps their smaller progenitors) and groups, and to cosmological parameters. As reviewed recently by Cen (1998), the growth of representative cluster masses depends strongly on $`\mathrm{\Omega }_0`$. This is due not to a direct relation between $`\mathrm{\Omega }_0`$ and structure growth per se, but more to the fact that the calculations must be normalized to match the present-epoch mass spectrum, which introduces a coupling between the amplitude $`\sigma _8`$ of the power spectrum and $`\mathrm{\Omega }_0`$ for viable models.
Several studies have indeed shown evidence for cluster-scale structures at redshifts $`z>2`$. Our HST imaging (Pascarelle et al. 1996b, herefter P96b; Pascarelle et al. 1998, hereafter P98), using a combination of broadband and medium-band filters to isolate Lyman $`\alpha `$ emission in the relevant redshift range, showed that the $`z=2.4`$ radio galaxy 53W002 is part of a rich assemblage of Lyman $`\alpha `$ emitters. Most of these are compact (effective radius $`r_e0.1`$” or 0.8 kpc), and are powered by star formation rather than by classical active nuclei. In P98, we showed that the surface density of such objects varies between different random lines of sight by approximately a factor of 4, with the 53W002 field being the richest we have observed thus far. A somewhat different grouping at similar redshift ($`z=2.38`$) was identified by Francis et al. (1996, 1997), who found four Lyman $`\alpha `$ emitters very close to the redshifts of Lyman $`\alpha `$ absorbers seen against two background QSOs. These emitters are seen over a projected span of 0.63 Mpc, and are much redder than the objects found by P96b. In an analogous way, Malkan et al. (1995, 1996) used narrow-band near-infrared imagery to find three H$`\alpha `$–emitting objects at $`z=2.50`$ in the foreground of the QSO SBS 0953+545 at $`z=2.58`$, closely matching the redshifts of metal-line absorption systems seen in the QSO spectrum. Starting from a sample of Lyman-break galaxies, Steidel et al. (1998, also Steidel 1999) have found a concentration of galaxies at $`z=3.090\pm 0.015`$ spanning about $`4\times 8`$ Mpc. And at even higher redshift, Hu & McMahon (1996) report spectroscopically-confirmed Lyman $`\alpha `$ companions to the $`z=4.55`$ QSO BR2237–0607.
These results show that it is now possible to trace developing clusters, and other large-scale structure, at high redshift. The new generation of wide-field imagers has enabled survey strategies that can tell how common, how extensive, and of what amplitude structures are in the galaxy distribution at various redshifts. As a first step in this direction, we present here a Lyman $`\alpha `$ survey of large fields around the regions we have searched with HST, to place the object counts from those fields in a larger context, and in particular to probe the spatial extent and bright end of the luminosity function of the cluster which includes 53W002.
In evaluating size and luminosity, we use $`\mathrm{H}_0=80`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, $`q_0=1/2`$, which gives an angular scale of 128” per Mpc. Scaling for other values, linear sizes scale directly with $`\mathrm{H}_0`$ and luminosities as $`\mathrm{H}_{0}^{}{}_{}{}^{2}`$. For other values of $`q_0`$, as a shortcut, we note that linear sizes (luminosities) quoted here would be multiplied by 2.9 (8.2) for $`q_0=0.1`$ and by 0.49 (0.24) for $`q_0=1`$.
## 2 Observations
We observed several fields around the radio galaxy 53W002 at $`z=2.39`$ (Windhorst et al. 1991), which had been shown from imaging in redshifted Lyman $`\alpha `$ to be part of a structure containing additional AGN and star-forming galaxies (P96ab, P98). This field therefore offered a unique opportunity to probe a known structure at significant redshift. We observed two further WFPC2 fields using the same filter set, as part of a parallel survey for additional objects in the window around $`z=2.4`$ (P98). For comparison, we also observed a large region adjacent to the 53W002 area using a filter tuned for Lyman $`\alpha `$ emission at $`z2.55`$.
For the current wide-field extension of the HST medium-band survey, we used the PFCCD imager, with $`2048^2`$ Tektronix CCD, for observing runs in the 1997 and 1998 summer seasons on the 4m Mayall telescope of Kitt Peak National Observatory. At the time, this system had significantly better throughput at 4100–4300 Å than the wider-field Mosaic system. Each exposure covered a region 14.3′ on a side with 0.420″ pixels. We isolated Lyman $`\alpha `$ in the redshift ranges $`z=2.322.45`$ and $`z=2.492.61`$ with intermediate-band filters, the first of which was intended as a clone of the WFPC2 F410M filter, manufactured by Custom Scientific, Inc., to the same specifications as the HST filter set. We refer to these as F413M and F433M to avoid confusion with the WFPC2 F410M filter; WFPC2 has no close counterpart to F433M. These filters have FWHM=150 Å and peak transmission at 4150 and 4330 Å respectively, as measured in a parallel beam; the peak transmission moves blueward by $`12`$ Å and the FWHM increases by about $`19`$ Å in the $`f/2.7`$ prime-focus beam of the Mayall telescope (Marcus 1998). In addition to the medium-band Lyman $`\alpha `$ filters, we also observed each field in $`B`$, for continuum magnitudes, and $`V`$ to account for color terms in the continuum subtraction (as in P96b). As it happened, we were able to observe three contiguous fields just to the northeast of 53W002 in F433M. No Lyman $`\alpha `$ emission candidates in the overlapping region were common to both filters, except for the brighter of the new QSOs discussed later, whose Lyman $`\alpha `$ emission is so strong that it was detected even in the extreme wings of the redder filter’s passband.
Total exposure times for each region and filter are listed in Table 1, with the area of full exposure extending 420″ in each coordinate from the listed position. Individual exposures were 30 minutes for the medium passbands, and 10-20 minutes in the broad bands, with dither motions of 20-30 arcseconds between successive exposures to suppress residual flat-field and cosmetic effects in the stacked combination images. The various field pointings were sometimes shifted from an exact rectangular pattern to avoid stray light from bright stars within a region extending about 5′ outward from the CCD edge. The image stacks show image FWHM in the range 1.2–1.6″ . The number of objects detected in each field depends on both seeing and total exposure times, and the number shown in Table 1 reflects detections in both $`B`$ and medium-band filters. The $`B`$ limiting magnitude is given for each field using a $`3\sigma `$ threshold.
The broadband data were converted to standard Johnson $`BV`$ magnitudes via secondary standard stars in M92, NGC 7006, and NGC 4147 (Christian et al. 1985, Odewahn et al. 1992). The photometric zero points were consistent to 0.02 magnitude or better from night to night. Both $`B`$ and $`V`$ magnitudes show color terms at the 0.03-magnitude level per unit change in $`(BV)`$. A more important issue is that of the color correction in continuum subtraction, as outlined below.
The 1997 data suffered from an additive ghost image of the telescope pupil occupying much of the field, produced by internal reflections in the optical corrector. This ghost image was not present in the 1998 data, since the dewar had been offset from the optical axis to avoid the problem. As an additive artifact, the ghost image could be isolated by comparing medium- and broad-band sky flats, then removed by subtracting scaled versions to eliminate the ghosting in our stacked images as completely as possible. Many of the images suffered from spatially variable background structure in the “blank sky” regions due to scattered starlight from stars both within and outside the field of view, sometimes modulated by passing cirrus clouds, which we subtracted using a $`101\times 101`$-pixel (42”) median filter, clipped around the brightest galaxies which would otherwise be partially subtracted. This allowed higher quality in the final average images, since pixels would not be artificially flagged for rejection because of a temporarily high background. This leaves spurious residual dark halos around bright stars, but since these are additive, local background subtraction will still give accurate photometry quite close to such stars.
We identified emission- and absorption-line candidates starting with object lists and photometry generated using version 1.0a of SExtractor (Bertin & Arnouts 1996), using visual inspection to reject putative detections which were compromised by bright stars or artifacts near the edges of individual exposures comprising the stacked mosaics. The detection parameters were: object detection threshold 2.5$`\sigma `$ above background over 5 contiguous pixels, and a deblending parameter 0.005 (which turned out to be essentially irrelevant at this level of crowding). Table 1 includes numbers of objects in each field appearing in the matched $`B,V,m_{413}/m_{430}`$ catalogs. Detections in all three bands were required to deal with color terms in the continuum-to-line comparison. The relative exposure depths suggest that we should not be missing comparable objects due to color effects (though the possibility of extreme colors, such as very red objects, still exists). This multiband matching requirement means that the listed detection totals do not simply reflect the relative exposure times. Coordinates were measured by fitting a celestial coordinate system to stars from the HST Guide-Star Catalog (GSC) on each frame; the formal accuracy is 0.25” rms, borne out by recovering positions of individual GSC stars.
The threshold for emission-line detection is not completely straightforward, since each object’s detectability depends both on the line flux and its equivalent width. Our primary criterion was for equivalent width incorporating individual error estimates, with a secondary list using formal significance of line emission as a basis for selection. Since the F413M medium-band filter sits on the blue edge of the $`B`$ passband, there is a color term accounting for the continuum slope between $`B`$ and 4130 Å (a similar but smaller term exists for the F433M filter). We follow W91 and P96B in using the traced filter properties to compute the locus of featureless power-law spectra (a reasonable approximation for galaxies in the emitted ultraviolet) in the $`(m_{413}B)(BV)`$ plane, as shown in Fig. 1 for the 53W002 field. This locus is well approximated by the line
$$(F413B)=0.32(BV)0.08,$$
which is a good fit to the observed distribution of “field” objects in our data (the numerical constants become 0.10 for the slope and 0.0 for the intercept in the case of the F433M filter). Our primary sample of emission-line candidates consists of objects which fall more than $`4\sigma `$ below this relation where $`\sigma `$ applies to the scatter of points on the emission side of the distribution’s ridge line (Fig. 1), which puts our threshold at 0.6 magnitude in F413M/F433M excess (observed equivalent width about 110 Å , corresponding to an emitted equivalent width 30-32 Å at $`z=2.42.6`$). These candidates are listed in Tables 2 and 3 for the two filters, and enlargements of the intermediate-band and $`B`$ images are shown in Figs. 2 and 3. Here, the tabulated $`(m_{413}B)`$ and $`(m_{430}B)`$ have been corrected for first-order color terms as described below; negative values indicate an excess in the narrower passband. The listed equivalent widths are in the observed frame; the emitted value will be smaller by $`(1+z)3.43.6`$. The Lyman $`\alpha `$ EW and flux for the three objects previously reported by P96b – their object numbers 18 and 19 plus 53W002 itself – are somewhat uncertain because each has a resolved Lyman $`\alpha `$ emission region (see section 6 below), which produces somewhat different values depending on how the flux is extracted. There is evidence that object 19 is itself variable as well (P96a).
The reliability of this sample is supported by the fact that all the members observed spectroscopically (in the 53W002 field) are indeed active nuclei at $`z=2.39`$. To assess whether there is an additional population of detections with lower equivalent width but comparable statistical reliability, we also considered object selection by significance in the deep 53W002 F413M data, defined as
$$S=[(F413B)0.32(BV)+0.08]/\sigma $$
(where $`\sigma `$ here is the statistical error in the F413-B color) with the additional requirements of $`B>23.5`$ and computed equivalent width $``$ 90 Å to avoid spurious detections of bright objects where the formal errors are much smaller than the scatter introduced by spectral features in stars and lower-redshift galaxies. All of the detections in the primary list have significance $`>5\sigma `$ by this criterion. Within the range of $`B`$ and $`S`$ that contains all the primary detections, the 53W002 field includes an additional five candidates, thus potentially augmenting the total number by about one third (which occur all over the field, unlike the clustered equivalent-width candidates). These are listed at the bottom of Table 2, but since this technique doesn’t generate any additional objects with line flux significantly above the threshold of the original list (even while relaxing the possible error bounds), we concentrate on the equivalent-width defined list. This list includes some objects with line fluxes as low as $`3.7\times 10^{17}`$ erg cm<sup>-2</sup> s<sup>-1</sup>, but a characteristic flux limit for approximate comparison with other results would be close to $`5\times 10^{17}`$ for the 53W002 field in the F410M filter. Corresponding values for the other fields at are $`1.4\times 10^{16}`$ for HU Aqr and $`1.0\times 10^{16}`$ for NGC 6251 and the three fields observed with the F430M filter.
Contamination of the Lyman $`\alpha `$ sample by objects with \[O II\] $`\lambda 3727`$ emission at $`z0.11`$ (F413M) or $`z=0.15`$ (F433M) should not be important, for the following reasons. The Keck spectroscopy of several of our candidates, plus objects from the HST lists, by Armus et al. (1999), shows that all the emission-line objects either have multiple lines at $`z2.4`$ or a single line with equivalent width plus continuum shape inconsistent with \[O II\] as judged from nearby objects. Finally, these objects are all smaller than 1″ in effective radius (that is, the blue continuum is either unresolved or almost so from the ground) and fainter than $`B=23`$, which would translate simultaneously into linear extent of less than 4 kpc and absolute magnitude fainter than $`M_B=15`$ for objects at redshift low enough to have \[O II\] emission in our passbands.
These data are not deep enough to recover the star-forming objects seen by Lyman $`\alpha `$ emission in the WFPC2 data of P96b and P98, even in the deepest Kitt Peak F413 exposure on the 53W002 field. The brightest such candidate in the HU Aqr WFPC2 field from P98 also falls slightly below our equivalent-width limit. These known $`z=2.4`$ objects have emission-line intensities which would correspond to Kitt Peak detection levels typically $`0.5\sigma `$, which is confirmed by comparison of our ground-based detections. Therefore we are tracing structures using objects which are, as far as we can tell from the spectroscopically identified subset, fairly luminous active nuclei. For 53W002 and its two immediate neighbors, HST imagery shows these to be accompanied by dimmer star-forming objects. Whether these are smoothly distributed throughout the clumping or form smaller structures in the regions traced by AGN is an important question for further work.
The two strongest-emission new candidates in the 53W002 field, bright enough to appear on blink inspection of the data during the observing run, were observed spectroscopically using the Ritchey-Chretien spectrograph and T2KB CCD at the Mayall telescope a week after the 1997 imaging observations. Grating KPC-10A (316 lines/mm, first-order blaze at 4000 Å ) gave 2.77 Å pixels with usable sensitivity over the 3700–8000 Å region, and resolution typically 2.1 pixels = 5.8 Å FWHM. Each object was acquired by offset from bright stars, as measured on the CCD frames. Mediocre seeing mandated a relatively wide 1.7″ slit opening, and even so the seeing was variable enough that only single 60-minute exposures of high quality were obtained for each object. While conditions were not photometric, the data could be placed on a relative flux scale using observations of the hot standard star PG 1708+602 (and the objects’ broadband magnitudes are accurately known from the multiband imagery).
Both new candidates observed were found to be QSOs, as shown in Fig. 4. Redshifts were measured by taking centroids of the bright emission lines, by Gaussian fits to the profile peaks, and by cross-correlation with the mean QSO spectrum assembled by Francis et al. (1991), adopting a mean of these redshift measures and using the differences as a measure of the error. Their spectroscopic properties are given in Table 4, with redshifts measured from Lyman $`\alpha `$ and C IV individually and from cross-correlation. Both have redshifts within $`\mathrm{\Delta }z=0.005`$ of the other objects in the 53W002 grouping (P96ab, Armus et al. 1999). We also include emitted-frame Lyman $`\alpha `$ widths, to make the point that these are fairly narrow-lined QSOs.
As this paper was in revision, we received word that Pascarelle, Yahil, & Puetter (1999) have confirmed candidate 6 from Table 2 with a redshift close to $`z=2.38`$ based on Keck spectroscopy, giving a total of six confirmations of the imaging candidates.
## 3 Emission-Line Detections and Field-to-Field Variations
Our most striking result is seen from Tables 2 and 3, and Fig. 5, where we show the distribution of candidate Lyman $`\alpha `$ emitters in part of the 53W002 F413M field. There is an extensive grouping of emission-line candidates including 53W002 which has no counterpart in the other fields observed at either $`z=2.4`$ or $`z=2.55`$. Fourteen objects in the 53W002 field passed our equivalent-width criterion for line emission, while only one at the edge of the HU Aqr field did, and none in the NGC 6251 field. Among the F413M fields, with differing exposure times, the field-to-field ratios to a highest common limiting line flux would be 4:1:0 (the first value rising to 6 for the intermediate limit appropriate to the empty NGC 6251 field). There are 6 candidates in the $`z=2.55`$ range sampled by the F433M filter, over almost three times the solid angle of the 53W002 field (and twelve times the solid angle encompassing the candidate emitters in that field). This shows that there are significant structures in place at $`z=2.4`$, but not necessarily over a large fraction of the sky; a simple estimate based on these data alone is that such assemblages cover less than 0.04 of the sky in a redshift range $`\mathrm{\Delta }z=0.15`$, with the limit arising from the fact that the 53W002 field was observed precisely because we already knew that some additional objects were present. The amplitude we find from field to field is even greater than that found by P98 from fainter HST detections at the center of each of these fields, which might indicate that luminous AGN are more clumped than fainter objects. This could mean, for example, that the more massive objects (thus more likely to host AGN) start life more strongly biased toward initial mass peaks.
We can assess the statistical significance of the grouping around 53W002 in several ways. First, we address the reality of the clumping seen within the 53W002 field at $`z=2.4`$. Most simply, the probability of $`n`$ objects falling within a single region covering a fraction $`f`$ of the solid angle surveyed will be $`p=f^{(n1)}`$, where the location of the region is not otherwise specified. In this case, using the superscribed circle about the 14 candidates for the region size and the area of full exposure in the stacked F410M as the overall field surveyed, $`f=0.38`$ and $`p=4\times 10^6`$. A Monte Carlo simulation indicates a somewhat higher (though still small) probability of having 14 points drawn from a uniform random distribution fall within a circle of this size, $`p=0.0014`$. Finally, we used the two-dimensional version of the Kolmogorov-Smirnov test as proposed by Peacock (1983) and examined in detail by Fasano & Francheschini (1987), employing the routines presented by Press et al. (1992). This test gives a significance level of 95% for the clustering within the 53W002 field. The critical values for the 2-dimensional test are slightly dependent on the distribution of sample points, but these points are not strongly correlated (Pearson $`r=0.40`$) and the effect would only act at the $`\pm 1`$% level in this regime. This most conservative of the tests still shows the grouping with high significance.
To test the sigificance of variations in the number of objects from field to field, we use combinatorics to ask how likely it is that a uniform distribution sampled with our total number of detections (to a common flux limit) would be so strongly weighted toward a specified field (since we already knew from previous data that there is an excess around 53W002). As noted above, to the highest of the three limiting fluxes, there are four objects in the 53W002 field, one in the HU Aqr field, and none in the NGC 6251 parallel field. Of the 207 ways to distribute 5 objects among three bins, 11 are at least this strongly weighted to the specified one, yielding a probability of $`11/207=0.053`$ of achieving this result by chance. Thus the significance of the higher number of objects in the 53W002 field is 95%, even without taking into account their concentration within the observed area in this field. We note that there is somewhat weaker evidence for clumping of the detections in the F430M filter ($`z2.55`$) in the 53W002 NE field; these objects all have derived line luminosities in the QSO range.
## 4 Lyman-$`\alpha `$ absorption candidates
Many of the brightest Lyman-break galaxies observed by Steidel et al. (1996, 1998) show net absorption at Lyman $`\alpha `$, indeed sometimes with no significant emission. One of the factors contributing to the strength or weakness of the emission may be metallicity, through the enhanced formation of grains which can absorb resonantly scattered Lyman $`\alpha `$ photons (Bonilha et al. 1979), though observations of Lyman $`\alpha `$ in starbursts of different metallicity show that there must be more to the story than this single parameter (Giavalisco et al. 1996, Lequeux et al. 1995, Thuan & Izotov 1997). Using imaging techniques, we are sensitive only to net emission, while some of the line emission may be cancelled by the line in absorption from stellar atmospheres and H I in the galaxy. Indeed, about 1/3 of the Lyman-break galaxies observed by Steidel et al. (1998) show net absorption at Lyman $`\alpha `$. Therefore, we consider here the possibility of detecting objects with strong absorption at Lyman $`\alpha `$.
As a guide to the strength of Lyman $`\alpha `$ absorption expected from star-forming galaxies, we use the HST GHRS spectrum of the bright knot in NGC 4214 obtained by Leitherer et al. (1996). After excising the narrow emission component, this spectrum shows an absorption line of equivalent width $`28`$ Å , which would be an observed value of EW=95 Å at $`z=2.4`$. This is just within our detection threshold for objects to $`B=25`$ in the 53W002 field, so that we can extract absorption candidates in the same way as the emission candidates. Since even luminous star-forming galaxies are unlikely to exceed QSO luminosities, we restrict the selection to objects in the range $`24<B<25`$, thereby avoiding much of the potential confusion from foreground stars and galaxies which have an absorption edge between the continuum and narrowband filters, and using the equivalent-width criterion to screen out stars with strong H$`\delta `$ absorption. This is a particular issue for white dwarfs, which would also be distinguished by broadband colors much bluer than expected for any high-redshift galaxies. Accordingly, we restrict the candidate absorption objects to the range $`0.15<(BV)<1`$ and require significance of the absorption to exceed $`4\sigma `$. This color range is wider than we observe for the star-forming emitters in the WFPC2 data (P96b). These criteria leave 4 candidates in the 53W002 field (Table 5, Fig. 6). Three of these are in same spatial region as the emission candidates, but our ability to select these objects in a more precise way is limited by the fact that the scatter in the $`(m_{413}B),(BV)`$ diagram is asymmetric and larger to the absorption side, largely due to the natural signal-to-noise limitations at faint levels, so that there are many more interlopers at a given equivalent width for absorption than for emission.
## 5 The 53W002 “Cluster” at $`z=2.39`$
These results strengthen the evidence for some sort of clustering at early cosmic times. We consider here what kind of assemblage we see in the 53W002 field, and how it might relate to the clustering we see today. This entails measures of its size, population, and dynamical state.
### 5.1 Cluster “Size” and Radial Distribution
The virial radius $`R_v=\frac{1}{n}[\mathrm{\Sigma }_{j<i}(\frac{1}{|𝐫_i𝐫_j|})]^1`$ of this assemblage of 14 objects is 157” or 1.2 Mpc in proper coordinates, which would correspond to 1.9 Mpc (a factor $`\pi /2`$ larger) in three dimensions for a typical projection geometry. The radial distribution is so extended that fewer than half the candidates (four) lie within this projected radius of the centroid.
For a distribution this sparsely sampled, whose centroid is not well determined by a strong central concentration, it may be more enlightening to consider the fraction of objects encompassed by circumscribed (projected) circles than by such a specific physical measure as the virial radius. All fourteen candidates are contained within a radius of 327”, with 2/3 (10) contained within r=218” and half (seven) inside r=164”.
To examine whether the distribution of these objects looks like contemporary relaxed systems, we consider how well a King-law profile with any core radius can be fitted to our observations. Because the center is ill-defined from such a sparse sample, we use as a statistic for comparison the cumulative number of objects within an encircled radius, whose center can drift to accomodate the maximum number within a given radius. From $`10^4`$ Monte Carlo samples of 14 objects each drawn from a King profile (in number density), we generated the bounds containing various fractions of the trials and identify these with confidence intervals. Since the $`z=2.4`$ structure is traced by active nuclei at B$``$ 24.5 – which may occur in rather low-luminosity galaxies this close to the peak redshift for QSO number density – we use a model for number density rather than incorporating some level of mass segregation to represent the luminosity density in a typical rich cluster. The core radius is also determined by fitting the radial scale of the distributions so that the true significance of each band may be slightly greater. The core radius was left as an adjustable parameter to be determined by the best fit to the observed cumulative distribution. This exercise should tell whether the observed two-dimensional distribution is likely to be drawn from one like the relaxed profiles of nearby (rich) clusters, and if so what its radial scale (the core radius for a King model) is. As shown in Fig. 7, the 53W002 association is less centrally condensed than a King model of any core radius. Specifically, if either the inner or outer four points in the number-radius relation are used to anchor the data to the Monte Carlo predictions, some points fall outside the 90% band (and if the inner points are fit, outside the 97% band). The difference is such that the inner points imply a core radius of 76” (0.6 Mpc), while the outer ones a core radius of 42” (0.3 Mpc). At this 90-97% confidence level, we can reject a relaxed King distribution for these objects.
If this grouping is not yet relaxed, we are left with an ambiguity in intepreting its linear scale – would it evolve more nearly in comoving or proper coordinates? If it has yet to turn around from the Hubble expansion, it will grow for some time in proper coordinates but shrink (slightly) in comoving ones until it turns around. On the other hand, it may have already turned around but not yet have virialized, in which case the proper-coordinate linear scale will remain nearly constant. Other reports of high-redshift structures have made differing assumptions on this matter – note that Steidel et al. (1998) quote a comoving extent for their structure at $`z=3.1`$, while some other workers use proper length. For the 53W002 structure, some intermediate case would be most appropriate, still allowing a wide range of current length scales for comparison with the present epoch.
We can address the correlation function $`w(\theta )`$ as measured in the 53W002 field from these data. While it is clearly a high-amplitude structure, this measurement might furnish useful information on length scales as well as just how large an amplitude could be reached by $`z=2.4`$. Edge effects were assessed by Monte Carlo trials, employing the expression from Landy & Szalay (1993) as used by Neuschaefer & Windhorst (1995). At $`\theta =30`$″ (200 kpc), $`w(\theta )=3.2`$, and we see positive correlation out to a radius $`r=5.5`$′ (2.4 Mpc) above a threshold value $`w1`$. Following the treatment by Neuschafer & Windhorst (1995), and using their sample to $`g=25`$, thus result does indicate that the region around 53W002 (sliced in both angle and redshift) is more strongly clustered than the field by a factor $`3`$, since the “field” objects have an amplitude of only about 0.03 covering a redshift range roughly $`z=0.52`$.
### 5.2 Galaxy and AGN Content
The objects from our emission candidate list which have been spectroscopically confirmed are all obvious AGN, with a mix of broad- and narrow-lined cases. This is not surprising given the flux constraints on spectroscopy with 4-m class telescopes, but it is already an unusual AGN population for a single group. Their absolute magnitudes are in the range associated with, for example, low-redshift PG quasars ($`M_B=21.4`$ to $`22.4`$ for our adopted cosmology, following Weedman 1986 in dealing with spectral slope). The brightest objects known from the WFPC2 field not to be such AGN have $`M_B20.5`$, so it remains unclear what the fainter KPNO detections at $`B=2425`$ represent. Certainly these would not be unusual luminosities for additional AGN, but this is a regime in which star-forming objects are not unreasonable either.
One hint might come from image structure. If AGN are weaker for the fainter objects, they might appear more clearly resolved since the core is less dominant. We compared image FWHM values (from the SExtractor tables and as computed by the IRAF imexamine task) for the emission candidates with those of bright, unsaturated stellar images nearby in the $`B`$ frame. Except for the extended structures around 53W002 and object 18, which have substantial line-emission components, all the candidates are unresolved. This fits with the sizes of the objects in the WFPC2 field, whose typical half-light radii are 0.10″ , but doesn’t furnish any further constraints on whether these new objects are more likely to be AGN or bright star-forming systems.
Even for only passive evolution of the star-forming objects, it is significant that we see none brighter than $`M_B=21`$, consistent with the HST results but now covering a much larger region. A typical $`L^{}`$ galaxy would have $`M_B=23`$ at $`z=2.4`$ unless either active evolution continued to substantially lower redshift, or merging of these small objects continued to form today’s luminous galaxies. More luminous galaxies could hide by lacking Lyman $`\alpha `$ emission, perhaps if they are more metal-rich and hence can suppress emission in this line, or if their star formation rate dropped quickly at early times. Near-IR line surveys could test the first possibility.
### 5.3 Velocity Dispersion
The five spectroscopically confirmed members of the 53W002 grouping have a velocity dispersion $`\sigma _z=0.0060`$, translating into $`\sigma _v=532`$ km s<sup>-1</sup> in the objects’ frame. Adding the two additional faint emitters from P98 drops this value to 467 km s<sup>-1</sup>. While one should respect the errors in estimating the velocity dispersion from such small samples, it does seem clear that we are not dealing with the dynamics of a rich virialized cluster with $`\sigma _v=1000`$ km s<sup>-1</sup>.
Since many of the previously reported HST objects around 53W002 are apparently star-forming complexes, with very narrow Lyman $`\alpha `$, there is the possibility of introducing a systematic offset in comparison with redshift measurements of broad-lined AGN using the same line. This is less of an issue with the three narrow-lined (“type 2”) AGN previously reported in this region (P96a,b). Comparison of the strong UV lines (Lyman $`\alpha `$, C IV, C III\]) with lower-ionization species or with narrow emitted-optical lines expected to arise far from the core (especially \[O II\] $`\lambda 3727`$) have shown that substantial differences in central velocity can exist. The shifts can exceed 1000 km s<sup>-1</sup> for radio-loud objects, but have a mean close to zero for radio-quiet QSOs (Espey et al. 1989, Marziani et al. 1996). In addition to being radio-quiet (Richards et al. 1999), the two new QSOs have rather narrow lines compared to many of the ones studied for velocity shifts (and compared to the Francis et al. 1991 composite), so the shifts may not be as large. In fact, their close match to the redshifts of other objects in the field would be a remarkable coincidence if systematic shifts of more than a few hundred km s<sup>-1</sup> are present, but the possibility remains that the actual velocity range of these objects is larger than the value we measure from Lyman $`\alpha `$ and C IV alone.
Furthermore, since the radial distribution suggests that the structure is not virialized and may still be coupled to the Hubble flow, we consider the limiting case in which the velocity range represents the Hubble flow across the depth of the structure rather than internal motions driven by gravity. For $`q_0=1/2`$ (or its $`\mathrm{\Omega }+\mathrm{\Lambda }`$ counterpart), the Hubble parameter $`H`$ would have been greater at $`z=2.4`$ than today’s $`H_0`$ by a factor about 6 (scaling inversely with cosmic time for this cosmology), so that the relevant expansion rate would have been in the range 300–600 km s<sup>-1</sup> Mpc<sup>-1</sup> for $`\mathrm{H}_0=`$50–100 km s<sup>-1</sup> Mpc<sup>-1</sup>. For a characteristic line-of-sight depth of 1.5 Mpc, comparable to the observed transverse extent containing most of the members, this implies a “velocity dispersion” of 450-900 km s<sup>-1</sup> even for a completely unbound assemblage. Since the positional data show clearly that the grouping has decoupled from the Hubble flow to the extent of showing a density contrast of at least a factor 4, we interpret this comparison as showing that this group is still turning around from the Hubble expansion, so that the velocity data do not necessarily allow us to measure its mass (or anything else about the detailed dynamics).
## 6 Lyman $`\alpha `$ Haloes of Constituent AGN
Three of the bright AGN in this field show extended Lyman $`\alpha `$ structures in WFPC2 data (P96b, P98). These are either linear or roughly biconical, fitting with a general paradigm of ionizing radiation directed mostly along the poles of some disklike structure. The KPNO data have better sensitivity to large regions of low surface brightness than does HST, and reveal new aspects of the extended line emission. For 53W002 and object 19 (in the P96b nomenclature), this is an extension of the Lyman $`\alpha `$ structure seen in WFPC2 images (Windhorst, Keel, & Pascarelle 1998), but for object 18, this resolved structure is not only much larger (extending more than 5″ from the core) than the ionization or scattering cone inferred from WFPC2 data, but it is most extended in a different direction. The inner parts of these structures are detected as well in H$`\alpha `$ using IRTF narrowband imagery and in \[O III\] using NICMOS multiband and grism data (Keel et al. 1999).
We can examine the structure of the Lyman $`\alpha `$ images by comparing both the $`B`$ continuum and emission-line images of each candidate emitter to stellar profiles from the same region of each image. This gives some insurance against minor PSF changes across the field, and avoids problems due to somewhat different PSF widths between the $`B`$ and F413M images. We consider extended emission to be detected when there is some scaling between broad– and medium–band images for which the difference is flat across the core and shows flux more extensive than the PSF. Requiring a flat central profile is conservative, to minimize the possibility of false detections at the expense of underestimating the flux in the spatially extended component. Several of the brightest emission-line objects show Lyman $`\alpha `$ emission more extended than their continuum structures.
This analysis suggests that both scattering and local recombination play roles in these emission-line halos. The three objects with extended emission-line regions illustrate this:
Object 18 (P96b): The PSF subtraction shows that more than 75% of the Lyman $`\alpha `$ flux from object 18 comes from outside the core, and recovers the gross features of the WFPC2 image. Similar results come from analysis of the $`B`$ image, while the relative count rates indicate that most of the $`B`$ light is in fact Lyman $`\alpha `$. Spectroscopy by Armus et al. (1999) shows that the extended cloud has almost no continuum component, consistent with these results. This accounts for the very blue color of the extended structure ($`(BV)=0.4`$), since there are no strong emission lines in the $`V`$ band.
53W002: For 53W002 and object 19, about half the line flux is spatially resolved, in accord with the HST PC data of Windhorst et al. (1998) and the ground-based Lyman $`\alpha `$ imaging from Windhorst et al. (1991). As noted earlier, the emission-line structure is approximately along the orientation of the 1″ radio double source, but much larger.
Object 19 (P96b): As in 53W002, about half the line flux is resolved, in accord with the HST data as well. In this case, the extended emission is all in Lyman $`\alpha `$ to our detection threshold; less than 10% of the $`B`$-band flux comes from outside the core.
These extended structures are illustrated in Fig. 8, comparing the medium-band image, the PSF-subtracted version, and HST imagery of the brightest regions. The large-scale line emission is well aligned with the small-scale emission observed with HST, which is well shown in the color figure of Windhorst et al. (1998) including scattered continuum components. For object 18, the KPNO data reveal that the inner emission region is identical with the two major components seen with HST, but much more extensive and amorphous material appears at this deeper surface-brightness threshold.
For the two newly detected QSOs, any such resolved line-emitting region must have less than 10% of the total Lyman $`\alpha `$ flux (and as low as 5% for the brighter QSO 2). These values apply to structures that are extended on the scale resolved by the PFCCD images; as a guide, the image size in the final F413M stack has 1.2″ FWHM.
Lyman $`\alpha `$ emission by itself is difficult to interpret, since we lack useful density indicators and its radiative transfer is sensitive to the velocity field and dust content. At a minimum, if mechanical energy input isn’t important in the extended nebulae, the number of Lyman $`\alpha `$ photons can give a lower limit to the number of ionizing photons reaching the gas, provided only that the situation is in a steady state. In turn, this can tell us whether the radiation field must be anisotropic to account for the structures we see - that is, whether we are correct in referring to some of these structures as ionization “cones”. The continua on our line of sight are measured from about 1100-2000 Å in the emitted frame, so that we should be able to do a reasonable extrapolation to the Lyman limit and estimate the expected number of ionizing photons in the isotropic case. For the simple case of a photoionized cloud occupying solid angle $`\mathrm{\Omega }`$ as seen from the central source, if we see the same ionizing continuum as the cloud does, the extrapolated continuum and observed Lyman $`\alpha `$ emission should satisfy
$$n_{LyC}\frac{\mathrm{\Omega }}{4\pi }\frac{n_{Ly\alpha }}{f_\alpha }$$
where $`n_{LyC}`$ is the number of Lyman continuum photons per second extrapolated from the observed continuum, $`n_{Ly\alpha }`$ is the observed number of Lyman $`\alpha `$ photons per second, and $`f_\alpha `$ is the fraction of recombinations whose cascade includes Lyman $`\alpha `$ (0.64 for case A, following the tabulations in Osterbrock 1989). The luminosity distance has cancelled on both sides, though we still need to make a plausible assumption about the clouds’ geometry to assign a subtended $`\mathrm{\Omega }`$. The equality holds for an ionization-bounded nebula which is optically thin to all the Lyman lines, in the sense that violating these conditions increases the continuum/line ratio and therefore makes the observed continuum more sufficient to power the extended line region.
Applying this test to the three resolved Lyman $`\alpha `$ regions shows that at least object 18, with its very extensive line emission, has an ionization source that we don’t see. Extrapolating the observed continuum at its flat level in flux falls short of creating the observed Lyman $`\alpha `$ emission by at least a factor 2, suggesting either a bump in the ionizing spectrum or anisotropic radiation. The $`BK`$ continuum shape is not unusually red, in fact quite normal for narrow-line AGN and almost identical to the other two objects with extended line emission, so that anisotropic illumination makes sense if it is not caused by material that would redden the observed continuum. A similar issue appears for many type 2 Seyfert nuclei, with a Lyman-continuum deficit implied by the observed continuum and line intensities, and blue UV continuum slopes. This has been variously attributed to scattering or reflection of radiation from a small continuum region (as in Antonucci, Hurt, & Miller 1994), and surrounding star formation (Colina et al. 1997), with recent results suggesting that the nucleus itself may not be an important contributor to the UV flux in narrow-line objects. The geometry of the Lyman $`\alpha `$ cloud near object 18 offers little help; while the inner parts, as detected with HST (P96a, Windhorst et al. 1998) resemble an ionization cone, the outer regions are extended at $`90^{}`$ in projection to this axis. Of course, additional energy sources might be considered, such as the radio jet interactions proposed for powerful radio galaxies. However, of these three objects, only 53W002 itself has significant resolved radio emission; the other two are both substantially weaker and unresolved by the VLA at the 1″ level (Richards et al. 1999). The flux data alone do not require anisotropic radiation for 53W002 and object 19, though the emission-line structure at least suggests an anisotropic gas distribution, and it is suspicious that the Lyman $`\alpha `$ structure in 53W002 aligns with the smaller double radio source (Windhorst, Keel, & Pascarelle 1998).
Independent of the ionization mechanism, such a rich collection of large clouds around the brightest illuminating sources raises the question of whether the extended gas belongs exclusively to the AGN hosts or exists more widely throughout this cluster, where we cannot observe it so easily. Detection of these structures in the continuum at a level above the weak free-free emission accompanying recombination would imply the presence of dust, a tracer of the level of star formation early in the galaxies’ history. Furthermore, if the nuclei are more often obscured early in cosmic time, we might expect to see “disembodied” Lyman $`\alpha `$ clouds in upcoming deep surveys.
## 7 The Evolution of Structure: Forward to the Past
We have reported a Lyman $`\alpha `$ survey aimed at tracing structures in the range $`z=2.32.6`$, finding a clumping or clustering in one field, represented by 14 luminous objects spanning about 3 Mpc. This adds to the existing evidence for structure in place, if not necessarily well developed, at cosmologically early epochs.
What does the 53W002 structure turn into? Based on its extent as found here, we can ask how many members might exist to the HST detection threshold. The existing WFPC2 data cover only a single 5.7-arcminute<sup>2</sup> area, while we find candidate members spread over an area of about 93 arcminutes<sup>2</sup>. If the WFPC2 field is representative, there would be 16 times as many faint star-forming members as we’ve detected to date. With 8 objects in the HST field now spectroscopically confirmed as members (Armus et al. 1999), that means the total membership would surpass 120 if we’re seeing a smooth distribution. Alternatively, if the star-forming objects are in clumps traced by the AGN that we detect from KPNO, there would still be $`70`$ in this structure. These numbers are lower estimates, since objects undoubtedly occur below our detection thresholds. These two cases represent rather different proposed histories for cluster and group formation – in one case, that clumps of objects will merge into today’s galaxies, and in the other, that individual objects we see at $`z=2.4`$ will either passively evolve as they begin to exhaust their gas or continue to acquire infalling material, with merging of initially separate galaxies a less important process.
Our velocity information is largely confined to the original WFPC2 field, with the addition of the two newly-identified QSOs. Thus it is not very clear how the small velocity dispersion of these objects (of order 385 km s<sup>-1</sup>, including redshifts of members from Armus et al. 1999) should be interpreted for the whole structure. Furthermore, the extended spatial distribution suggests that the structure has not yet relaxed, and may not yet be fully decoupled from the Hubble flow. Recent simulations of galaxy formation from a clumpy medium by Haehnelt, Steinmetz, & Rauch (1998) indicate that line-of-sight velocity measurements not only have a factor 2 dispersion as seen from various directions, but underestimate the relaxed virial velocities by $`60`$%. These considerations all make a virial mass estimate very uncertain, and likely a lower limit. It may be more realistic to consider the velocity dispersion as applying to the megaparsec-scale clumping including 53W002 itself and the AGN in objects 18 and 19. The velocity range we see is comparable to the dispersion expected purely from the Hubble flow on an assemblage 3 Mpc deep at this epoch, so we may well be seeing the group near the time of turnaround from cosmological expansion, in which case the velocity dispersion tells very little about the internal dynamics.
These questions suggest several potentially fruitful lines for further work. Most notably, we need to know more about the content of this structure, especially for fainter objects both with and without strong Lyman $`\alpha `$ emission. Multiband imagery sufficient to derive photometric redshifts and narrowband near-infrared measurements tailored to find emission from \[O II\], \[O III\], or H$`\alpha `$ can help fill out our census of members. A more accurate accounting of how common such structures are in the early Universe will require wider-field multiband surveys, preferably with fine enough wavelength bands to both pick out line emitters and resolve multiple line-of-sight sheets or clusters. Eventually, dynamical studies should tell us how these early assemblages become the rich structural spectrum seen in today’s Universe.
We are grateful to Richard Green for approving, and the KPNO staff for implementing, a scheduling switch between the imaging and spectroscopic observing runs which allowed us to confirm the new QSO candidates. We acknowledge C. Leitherer and colleagues for making their starburst spectral templates available via WWW. Portions of this work were supported by NSF grant AST-9802963 and HST STScI grants GO-5985.0\*.96A and AR-8388.0\*.98A. We thank Paul Francis, the referee, for goading us into more quantitave probability assessments than we had originally incorporated, as well as some interesting suggestions on cosmology versus dynamics in the cluster redshift distribution.
| Table 1 | | | | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- |
| Summary of Fields Observed | | | | | | | |
| | | Center (2000) | | Exposure (min) | | Matched | |
| Field | Filter | $`\alpha `$ | $`\delta `$ | F413/430M | B | Objects | B(lim) |
| 53W002 | 413M | 17 14 10.3 | +50 16 07 | 480 | 100 | 3161 | 26.5 |
| HU Aqr par | 413M | 21 07 27.7 | -05 23 07 | 150 | 60 | 928 | 25.2 |
| NGC 6251 par | 413M | 16 36 37.0 | +82 34 10 | 180 | 75 | 930 | 25.5 |
| 53W002 E | 430M | 17 15 22.4 | +50 17 23 | 240 | 30 | 1588 | 25.0 |
| 53W002 N | 430M | 17 14 01.6 | +50 28 42 | 240 | 30 | 2215 | 25.5 |
| 53W002 NE | 430M | 17 15 06.6 | +50 28 23 | 240 | 40 | 2167 | 25.5 |
| Table 2 | | | | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- |
| Candidate Lyman $`\alpha `$ emitters from F413M, $`z2.4`$ | | | | | | | |
| No. | $`\alpha `$(2000) | $`\delta `$(2000) | $`B`$ | $`(m_{4130}B)`$ | Obs. EW | F(Ly $`\alpha `$) | Notes |
| | | | (mag) | (mag) | EW | (cgs) | |
| 53W002 field: primary equivalent-width sample | | | | | | | |
| 1 | 17:14:12.01 | +50:16:02.3 | 23.41 | $`1.29\pm 0.04`$ | 342 | 1.1(-15) | P96b object 18 |
| 2 | 17:14:44.72 | +50:17:44.9 | 23.31 | $`1.24\pm 0.03`$ | 320 | 1.1(-15) | New QSO |
| 3 | 17:14:11.30 | +50:16:09.4 | 23.91 | $`1.01\pm 0.04`$ | 230 | 4.5(-16) | P96b object 19 |
| 4 | 17:14:12.39 | +50:18:18.0 | 24.37 | $`1.00\pm 0.04`$ | 226 | 2.9(-16) | New QSO |
| 5 | 17:14:14.70 | +50:15:29.7 | 24.05 | $`0.80\pm 0.05`$ | 164 | 2.8(-16) | 53W002 |
| 6 | 17:14:39.82 | +50:21:52.3 | 25.46 | $`0.85\pm 0.09`$ | 178 | 8.4(-17) | |
| 7 | 17:14:32.80 | +50:15:50.7 | 26.19 | $`0.78\pm 0.15`$ | 158 | 3.8(-17) | |
| 8 | 17:14:31.66 | +50:19:06.8 | 25.54 | $`0.66\pm 0.11`$ | 125 | 5.5(-17) | |
| 9 | 17:14:24.76 | +50:20:45.7 | 25.36 | $`0.62\pm 0.09`$ | 115 | 5.9(-17) | |
| 10 | 17:14:42.15 | +50:16:51.8 | 24.34 | $`0.73\pm 0.08`$ | 144 | 1.9(-16) | |
| 11 | 17:14:39.20 | +50:21:32.8 | 25.66 | $`0.83\pm 0.13`$ | 172 | 6.7(-17) | |
| 12 | 17:14:53.75 | +50:22:31.8 | 25.87 | $`0.62\pm 0.15`$ | 116 | 3.7(-17) | |
| 13 | 17:14:42.48 | +50:16:01.6 | 25.09 | $`0.67\pm 0.09`$ | 128 | 8.5(-17) | |
| 14 | 17:14:30.77 | +50:12:09.1 | 26.21 | $`0.82\pm 0.16`$ | 169 | 4.0(-17) | |
| 53W002 field: additional significance-selected candidates | | | | | | | |
| 15 | 17:13:49.35 | +50:14:29.0 | 25.05 | $`0.52\pm 0.07`$ | 92 | 6.3(-17) | |
| 16 | 17:14:04.15 | +50:18:14.6 | 25.24 | $`0.53\pm 0.09`$ | 94 | 5.4(-17) | |
| 17 | 17:14:30.46 | +50:13:13.7 | 25.58 | $`0.58\pm 0.10`$ | 106 | 4.4(-17) | |
| 18 | 17:13:42.86 | +50:21:20.4 | 25.35 | $`0.57\pm 0.11`$ | 104 | 5.4(-17) | |
| 19 | 17:14:17.55 | +50:10:13.6 | 25.50 | $`0.55\pm 0.12`$ | 98 | 4.4(-17) | |
| HU Aqr field | | | | | | | |
| 20 | 21:07:57.26 | -05:25:41.1 | 22.25 | $`1.04\pm 0.03`$ | 241 | 2.2(-15) | |
| Table 3 | | | | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- |
| Candidate Lyman $`\alpha `$ emitters from F433M, $`z2.6`$ | | | | | | | |
| No. | $`\alpha `$(2000) | $`\delta `$(2000) | $`B`$ | $`(m_{4300}B)`$ | Obs. EW | F(Ly $`\alpha `$) | Notes |
| | | | (mag) | (mag) | EW | (cgs) | |
| 53W002 N field: | | | | | | | |
| 21 | 17:13:17.36 | +50:27:12.8 | 24.55 | $`1.17\pm 0.15`$ | 291 | 3.2(-16) | |
| 53W002 E field: | | | | | | | |
| 22 | 17:15:23.23 | +50:19:35.2 | 22.38 | $`1.17\pm 0.05`$ | 155 | 1.2(-15) | |
| 53W002 NE field: | | | | | | | |
| 23 | 17:14:51.62 | +50:23:13.4 | 24.54 | $`0.87\pm 0.16`$ | 184 | 2.0(-16) | |
| 24 | 17:15:33.80 | +50:28:49.6 | 24.56 | $`0.99\pm 0.15`$ | 223 | 2.4(-16) | |
| 25 | 17:15:28.13 | +50:23:46.2 | 24.57 | $`1.02\pm 0.16`$ | 234 | 2.5(-16) | |
| 26 | 17:15:32.92 | +50:30:52.1 | 24.25 | $`1.06\pm 0.11`$ | 248 | 3.5(-16) | |
| Table 4 | | | | | |
| --- | --- | --- | --- | --- | --- |
| Lyman $`\alpha `$ Absorption Candidates ($`z2.4`$) in the 53W002 Field | | | | | |
| No. | $`\alpha _{2000}`$ | $`\delta _{2000}`$ | $`B`$ | $`(m_{413}B)`$ | $`(BV)`$ |
| 27 | 17:13:44.51 | +50:21:08.9 | $`24.01\pm 0.04`$ | $`1.94\pm 0.19`$ | $`0.18\pm 0.10`$ |
| 28 | 17:14:05.79 | +50:19:17.8 | $`24.81\pm 0.05`$ | $`1.46\pm 0.18`$ | $`0.24\pm 0.14`$ |
| 29 | 17:14:23.37 | +50:23:09.2 | $`24.25\pm 0.04`$ | $`1.56\pm 0.14`$ | $`0.57\pm 0.07`$ |
| 30 | 17:14:20.42 | +50:10:51.2 | $`24.29\pm 0.03`$ | $`1.56\pm 0.14`$ | $`0.86\pm 0.05`$ |
| Table 5 | | | | | |
| --- | --- | --- | --- | --- | --- |
| New QSO redshift measurements | | | | | |
| Object | Ly $`\alpha `$ FWHM (Å ) | Ly $`\alpha `$ | C IV | Cross-correlation | Mean |
| | | (centroids) | | | |
| 2 (171444.72+501744.9) | 11 | 2.384 | 2.377 | 2.382 | $`2.381\pm 0.002`$ |
| 4 (171412.39+501818.0) | 17 | 2.391 | 2.400 | 2.388 | $`2.393\pm 0.006`$ |
|
no-problem/9908/cond-mat9908217.html
|
ar5iv
|
text
|
# Structural instability associated with the tilting of CuO6 octahedra in La2-xSrxCuO4
## I Introduction
Number of experimental results accumulated for a decade have shown that spin fluctuations, appearing as incommensurate peaks around $`(\pi ,\pi )`$, play a significant role in the high-$`T_\mathrm{c}`$ superconductivity. In particular, a recent neutron-scattering work on high-quality La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> (LSCO) single crystals by Yamada et al. has clearly demonstrated that there exists a linear relationship between the incommensurability and $`T_\mathrm{c}`$ in a wide hole concentration range. On the other hand, a primary contribution of phonon to high-$`T_\mathrm{c}`$ superconductivity has not been seriously considered because electrical resistivity measurements suggest that the electron-phonon (e-ph) interaction is not as large as that of BCS superconductors such as A15-type compounds.
It is known that the structural properties of LSCO is closely related to the tilting pattern of CuO<sub>6</sub> octahedra. Birgeneau et al. and Böni et al. clarified that the tetragonal X-point phonon mode, associated with tilting of CuO<sub>6</sub> octahedra along the \[$`\mathrm{1\; 1\; 0}`$\] or \[$`\overline{1}\mathrm{1\; 0}`$\] tetragonal axis, goes soft at the high-temperature tetragonal (HTT) to the low-temperature orthorhombic (LTO) structural phase transition. Thurston et al. studied temperature dependence of the $`\mathrm{\Gamma }`$\- and Z-point modes in the LTO phase, which are split from the degenerated X-point mode in the HTT phase. They have shown that the Z-point mode first hardens below the HTT-LTO transition, then softens again with further decreasing temperature, indicating incipient structural transition from LTO to the low-temperature tetragonal (LTT) phase. The structural phase transitions and soft phonons in this system are summarized in Appendix A.
Recently, Lee et al. have discovered that the softening of Z-point phonon breaks at $`T_\mathrm{c}`$ in the optimally doped La<sub>1.85</sub>Sr<sub>0.15</sub>CuO<sub>4</sub> sample. This new finding, suggesting a competition between the LTO-LTT transition and the appearance of superconductivity, is contrary to the conventional view that the e-ph interaction in LSCO is weak thus not essential. Therefore it is now increasingly important to revisit structural properties of LSCO and elucidate their relation to the high-$`T_\mathrm{c}`$ superconductivity. In the present study, we have carried out comprehensive inelastic neutron-scattering measurements of the Z-point phonon using $`x=0.10`$, 0.12 and 0.18 single crystals. Sample characterizations and neutron-scattering measurements are described in Sec. II. To study the phonon softening in detail, we utilized a new technique for optimizing resolution function of neutron scattering, which is described in Appendix B. In the course of studying the soft phonons, we discovered incommensurate structural diffuse signals around $`(0k\pm \delta l)_{\mathrm{LTO}}`$ $`(k=odd,l=even)`$ in the HTT phase for both $`x=0.12`$ and 0.18. The newly found diffuse signals as well as the soft phonons are reported in Sec. III, and discussed in Sec. IV. @
## II Experimental Details
In order to grow a LSCO single crystal which is large, highly crystalline and homogeneous in the Sr distribution, Traveling-Solvent-Floating-Zone (TSFZ) method has been utilized and improved over past few years. In the present study, we have grown $`x=0.10`$, 0.12 and 0.18 single crystals by using the TSFZ method; all the crystals have a volume of about 1 cm<sup>3</sup> with the mosaicness of 0.4 full-width at half-maximum (FWHM). To make these crystals stoichiometric, i.e., oxygen deficiency free, as-grown samples were annealed under pure oxygen gas flow at 900 C for 50 hours, then cooled down to 500 C with a cooling rate of 10 C per hour and kept there for 50 hours, and finally cooled down slowly to room temperature.
Bulk magnetic susceptibilities were measured with a SQUID magnetometer. All the samples show superconductivity with shielding signal $`100`$ % and the onset of $`T_\mathrm{c}`$’s were determined to be 28.5 K, 31.5 K and 36.5 K for $`x=0.10`$, 0.12 and 0.18, respectively. The transition width defined as the temperature range between 5 % and 95 % of the maximum shielding signal is about 3 K for all the samples, which sharpness suggests that doped Sr are homogeneously distributed. The c-axis lengths of all the samples, measured at room temperature by using X-ray powder diffractometer, are in good agreement with previous report. The HTT-to-LTO transition temperatures, defined as $`T_{\mathrm{s1}}`$ in this paper, were determined to be 240 K and 125 K for $`x=0.12`$ and 0.18 by neutron diffraction.
Neutron scattering experiments were performed on the triple-axis spectrometer TOPAN installed at the JRR-3M research reactor in the Japan Atomic Energy Research Institute (JAERI). The final energy of neutrons was fixed at $`E_\mathrm{f}=14.1`$ meV. The $`(\mathrm{0\; 0\; 2})`$ Bragg reflection of Pyrolytic Graphite (PG) was used in order to monochromatize and analyze neutron beam. A typical horizontal collimation is Blank-60-S-60-Blank. PG filter was inserted into scattered beam to reduce higher-order contaminations. All the single crystals have twined domains in the LTO phase. Therefore, we should consider that all the measurements were performed in the superposed $`(h0l)_{\mathrm{LTO}}`$ / $`(0kl)_{\mathrm{LTO}}`$ zone. Throughout this paper, the reciprocal space is described by using reciprocal lattice unit (r.l.u.) in the LTO (Bmab) coordinate system. A crystal was put into an Aluminum container filled with He gas then attached to the cold finger of a closed-cycle <sup>4</sup>He cryostat. Temperature was measured by using Si-diode thermometers attached to the cold finger and the bottom part of the Al container.
## III Results
### A Soft phonons
The Z-point phonons on $`x=0.10`$, 0.12 and 0.18 around $`𝐐_0=(\mathrm{3\; 0\; 2})`$ were studied as a function of temperature. Constant-$`Q`$ scans were carried out at $`𝐐^{}=𝐐_0+𝐪`$ where the instrumental resolution matches most effectively with the dispersion of phonon (focusing). The procedure we employed to determine the most effective focusing point $`𝐐^{}`$ is described in Appendix B. The phonon spectra for $`x=0.10`$ and 0.12 were measured at $`𝐐^{}=(\mathrm{3.09\; 0\; 1.88})`$, while for $`x=0.18`$ the scan was carried out at $`𝐐^{}=(\mathrm{3.10\; 0\; 1.96})`$. Well-defined phonon spectra were obtained for all the samples, indicating that the resolution ellipsoid at $`𝐐^{}`$ sufficiently focuses with the slope of dispersion. To quantitatively analyze the phonon spectra, data were fitted with the following scattering function $`S(Q,\omega )`$ convoluted with a proper instrumental resolution:
$`S(Q,\omega )`$ $`=`$ $`{\displaystyle \frac{\omega }{1\mathrm{exp}(\omega /k_BT)}}\left\{{\displaystyle \frac{1}{\pi }}{\displaystyle \frac{A}{\mathrm{\Gamma }_{\mathrm{ph}}/2}}{\displaystyle \frac{1}{1+\left(\frac{\omega \omega _{\mathrm{ph}}}{\mathrm{\Gamma }_{\mathrm{ph}}/2}\right)^2}}\right\},`$
$`A`$ $`=`$ $`\chi (Q)={\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{\chi (Q,\omega )}{\omega }}𝑑\omega ,`$
where $`\omega _{\mathrm{ph}}`$ and $`\mathrm{\Gamma }_{\mathrm{ph}}`$ are the phonon energy and the intrinsic line-width at $`𝐐^{}`$, respectively. Fitting thus carried out quite well reproduce all the observed data.
Temperature dependences of the soft phonon energy $`\omega _{\mathrm{ph}}`$ for $`x=0.10`$, 0.12 and 0.18 are shown in Fig. 1. All the data exhibit the softening of phonons with decreasing temperature toward $`T_\mathrm{c}`$, which confirms that the phonons at $`𝐐^{}`$ indeed reflect the behaviors of Z-point phonons. As a reference, the Z-point ($`𝐐_0=(\mathrm{1\; 0\; 4})`$) phonon for $`x=0.15`$ reported by Lee et al. is also shown in Fig. 1(c). For $`x=0.18`$, the softening of phonon breaks with the appearance of superconductivity as seen in $`x=0.15`$. In contrast, the softening continues even below $`T_\mathrm{c}`$ for $`x=0.10`$ and 0.12. These results indicate that the structural instability for the LTT phase persists even in the superconducting phase for $`x=0.10`$ and 0.12.
The intrinsic line-widths of the soft phonons $`\mathrm{\Gamma }_{\mathrm{ph}}`$ for $`x=0.10`$, 0.12 and 0.15 are shown as a function of temperature in Figs. 2(a)-(c). The data for $`x=0.15`$ is also after Lee et al.. Down to 40 K, the line-widths for all the samples decrease with decreasing temperature, corresponding to increase of phonon life-time. However, this narrowing suddenly stops, i. e. breaks, around 40 K ($`T_d`$) and the line-widths become constant at lower temperatures. On the other hand, in La<sub>2</sub>CuO<sub>4</sub> studied by Lee et al., the narrowing continues down to the lowest temperature, indicating that the break of the line-width narrowing is a characteristic phenomenon of
the Sr-doped samples. As for $`x=0.18`$, it is difficult to estimate the accurate value of $`\mathrm{\Gamma }_{\mathrm{ph}}`$ because the two phonons, i.e., $`\mathrm{\Gamma }`$ and Z-point modes, were so close that the phonon spectra were superposed with each other.
### B Incommensurate diffuse peak in the HTT phase
Upon cooling, the $`(0kl)`$ $`(k=odd,l=even)`$ superlattice reflections emerge at the HTT-LTO transition temperature $`T_{\mathrm{s1}}`$. As briefly described in Sec. II, $`T_{\mathrm{s1}}`$ values of $`x=0.12`$ and $`x=0.18`$ were determined to be 240 K and 125 K from the temperature dependence of the
$`(\mathrm{0\; 3\; 2})`$ peak intensity. Note that we inserted a PG filter between sample and analyzer as well as between monochromator and sample, which almost completely eliminates higher-order contaminations. In the course of studying the soft-mode phonons and superlattice reflections, we found that there remains a weak diffuse peak centered at $`(\mathrm{0\; 3\; 2})`$ and $`\omega =0`$ even above $`T_{\mathrm{s1}}`$. Since this peak is associated with the softening of X-point phonon and diverges at $`T_{\mathrm{s1}}`$, we have noticed that this is a so-called “central peak” as studied in SrTiO<sub>3</sub> and many other systems.
At higher temperature, we found that this central peak starts splitting into two incommensurate components at $`(\mathrm{0\; 3}\pm \delta 2)`$ or more generally $`(0k\pm \delta l)`$ $`(k=odd,l=even)`$. This phenomenon was first reported by Shirane et al. in their preliminary experiment for $`x=0.15`$. As shown in Fig. 3(a), the incommensurate “central” peaks are clearly seen in $`x=0.12`$ at 315 K, temperature much higher than $`T_{\mathrm{s1}}`$. Similar incommensurate peaks were also observed for $`x=0.18`$. The peak profiles for both samples were well fitted with a double Lorentzian. The sharp peak in Fig. 3(a) shows the $`(\mathrm{0\; 3\; 2})`$ superlattice peak
at $`T=13.5`$ K, of which the line-width corresponds to the instrumental resolution at $`(\mathrm{0\; 3\; 2})`$. By fitting with a double Lorentzian convoluted with the resolution, the intrinsic line-widths of central peaks were estimated to be $`0.06`$ Å<sup>-1</sup> for both the samples and for all the temperature range above $`T_{\mathrm{s1}}`$, indicating short range in-plane correlations ($`\xi 17`$ Å).
The temperature dependence of the splitting $`\delta `$ for $`x=0.12`$ and 0.18 were also measured. As shown in Fig. 3(b), the $`\delta `$ value of $`x=0.18`$ (open diamonds) increases with increasing temperature and is saturated around $`\delta 0.12`$ (r.l.u.). Upon cooling, on the other hand, $`\delta `$ approaches zero at $`T_{\mathrm{s1}}`$, suggesting that the central peak indeed starts splitting just at $`T_{\mathrm{s1}}`$. It is remarkable that the $`\delta `$ values for both $`x=0.12`$ (closed circles) and 0.18 can be scaled using each $`T_{\mathrm{s1}}`$ value. Note that the saturated value $`\delta 0.12`$ is very close to the incommensurability $`ϵ`$ of elastic and inelastic magnetic scattering peaks seen in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> at $`(\pi (1\pm 2ϵ),\pi )`$ and $`(\pi ,\pi (1\pm 2ϵ))`$. In the $`x=0.25`$ sample, which has no HTT to LTO structural phase transition, the central peak was also searched around $`(\mathrm{0\; 3\; 2})`$ at room temperature. However, no signature was found.
The incommensurate central peak was also measured around $`(\mathrm{0\; 1\; 4})`$ at 315 K for $`x=0.12`$. The intrinsic line-width is almost same as that of $`(\mathrm{0\; 3\; 2})`$. The width of the central peak along the $`l`$ direction, which corresponds to the out-of-plane correlation, was measured at $`(\mathrm{0\; 1\; 4})`$. The peak-width along l is almost identical to that along k, indicating short-range isotropic correlations for the in-plane and out-of-plane directions. The intensity ratio of the $`(\mathrm{0\; 3\; 2})`$ and $`(\mathrm{0\; 1\; 4})`$ central peaks is 2.9, which is close to the ratio of $`Q^2`$ values for these two peaks, 2.7, indicating that the central peak originates from atomic displacement. The intensity ratio between the central peak and the fundamental Bragg peak is about 10<sup>-3</sup> in the present neutron-scattering measurement. We found a similar diffuse peak in X-ray diffraction measurements. However, this diffuse peak intensity in X-ray diffraction is 10<sup>-5</sup> times weaker than that of the fundamental reflection. This difference between neutron and X-ray measurements suggests that the central peaks are mainly contributed from the displacement of oxygen atoms.
## IV Discussion
### A Soft phonons
The present study, combined with the work by Lee et al. , has established that the softening of Z-point phonon suggesting incipient LTO-LTT transition breaks at $`T_\mathrm{c}`$ for optimally and overdoped LSCO and that the softening persists even below $`T_\mathrm{c}`$ in the underdoped region. This breaking found in optimally and overly doped LSCO is consistent with previous results obtained by other techniques suggesting competition between the structural phase transition and the superconductivity in LSCO and related compounds. Such a competition is also seen in other systems such as the A15 superconductors. On the contrary, it appears that the behaviors of soft phonons in the underdoped region are qualitatively different from those in the overdoped region. Note also that the persistence of phonon softening below $`T_\mathrm{c}`$ is not a characteristic feature for the 1/8-doping. The change in behaviors of phonon softening across the optimum hole concentration might be related to the incommensurate elastic magnetic peaks in the underdoped region and/or the energy-gap in the spin fluctuations near the optimally doped region. It is intriguing to relate the appearance of elastic magnetic signals to the LTO-to-LTT instability persisting below $`T_\mathrm{c}`$ in underdoped LSCO. It is clear, however, that even more systematic investigation is required to elucidate the origin of soft-phonon breaking and its influence to other properties.
Another remarkable feature is that the narrowing of soft-phonon line-width breaks below $`T_\mathrm{c}`$, which is confirmed in underdoped and optimally doped LSCO. This phenomenon is contrary to that observed in BCS superconductors, where e-ph scattering disappears because of the BCS gap opening. It is thus suggested that the intrinsic phonon line-width is not owing to e-ph interactions, though the origin of the breaking of soft-phonon narrowing is still not clear.
### B Incommensurate diffuse peaks
Residual week diffuse peak observed above $`T_{\mathrm{s1}}`$ at the LTO superlattice position is very similar to the central peak associated with R-point phonon softening in SrTiO<sub>3</sub>. Interestingly, the central peak in LSCO splits into incommensurate peaks and the incommensurability $`\delta `$ saturates around $`\delta 0.12`$ r.l.u., which value is close to the splitting of incommensurate magnetic signals. This implies that the incipient lattice modulation starts appearing at very high temperature. It has been reported that the averaged (long-ranged) structure is different from local (short-ranged) structures in the LSCO cuprates. Quantitative comparison between the experimental results and structural models including local distortions is required. Further neutron scattering along this line is planned in the near future.
## Acknowledgments
We thank Y. Endoh, N. Toyota, R. J. Birgeneau, T. Imai, and H. Fukuyama, for several stimulating discussions. We also acknowledge M. Onodera and K. Nemoto for their technical supports on neutron scattering experiments. This work was supported in part by a Grant-In-Aid for Scientific Research from the Japanese Ministry of Education, Science, Sports and Culture, and by a Grant for the Promotion of Science from the Science and Technology Agency and also supported by CREST and the US-Japan cooperative research program on Neutron Scattering. Work at Brookhaven National Laboratory was carried out under contact No. DE-AC02-98CH10886, Division of Material Science, U. S. Department of Energy.
## A Structural phase transitions and soft phonons in LSCO
As shown in Fig. 4(a), all the structural phases of LSCO can be characterized by two order parameters $`Q_1`$ and $`Q_2`$, which correspond to the rotation of the CuO<sub>6</sub> octahedron along \[$`\mathrm{1\; 0\; 0}`$\] and \[$`\mathrm{0\; 1\; 0}`$\] in Bmab notation. In the high-temperature tetragonal (HTT) phase, no static tilting exists and the structure has a tetragonal symmetry with the space group of I4/mmm; $`|Q_1|=|Q_2|=0`$. With decreasing temperature, coherent tilting which rotation axis is parallel to \[$`\mathrm{1\; 0\; 0}`$\] develops below $`T_{\mathrm{s1}}`$ and it leads to the low-temperature orthorhombic phase (LTO) with the space group Bmab; $`|Q_1|0,|Q_2|=0`$. $`T_{\mathrm{s1}}`$ changes from 520 K for $`x=0.00`$ to 0 K for $`x>0.21`$. In some doped LaCuO<sub>4</sub> such as La<sub>2-x</sub>Ba<sub>x</sub>CuO<sub>4</sub>, with further lowering temperature, a low-temperature tetragonal (LTT) phase appears with the space group P4<sub>2</sub>/ncm; $`|Q_1|=|Q_2|0`$. An intermediate phase
between LTO and LTT, which is written in the Pccn orthorhombic symmetry, is also known; $`|Q_{1,2}|0,|Q_1||Q_2|`$.
The HTT-to-LTO structural phase transition in LSCO is driven by softening of the zone-boundary optical phonon at X-point $`𝐪=\frac{1}{2}(1\pm \mathrm{1\; 0})`$. The temperature dependence of the soft phonon energy is shown schematically in Fig. 4(b), which is based on the Landau free energy expanding to eighth order for {$`Q_1`$, $`Q_2`$}. At $`T_{\mathrm{s1}}`$ the doubly degenerate X-point phonons freeze, resulting in the HTT-to-LTO phase transition. In the LTO phase, due to the crystal symmetry, the degenerate phonon branches split into $`\mathrm{\Gamma }`$\- and Z-point phonon of which amplitudes correspond to $`Q_1`$ and $`Q_2`$, respectively. The $`\mathrm{\Gamma }`$-point phonon hardens with decreasing temperature suggesting the stability of coherent $`|Q_1|`$ tilting, while the Z-point phonon softens at lower temperature, indicating the instability toward the LTT phase. If the Z-point phonon completely freezes, a further transition from the LTO to Pccn or LTT phase occurs. As for LSCO, no Pccn or LTT phase exists but the softening of the Z-point phonon is observed, indicating incipient transition to these phases. Note that since the present study cannot distinguish a qualitative difference between the Pccn and LTT, we define a structural phase below LTO as the LTT phase in this paper for convenience.
## B Experimental techniques in phonon measurements
For the present study, it is essential to precisely measure the frequency and intrinsic width of phonon. As shown in the inset of Fig. 5(a), the instrumental resolution of neutron triple-axis spectrometer has a slope in the $`Q\omega `$ space. Therefore, there exist focusing and defocusing side for the phonon dispersion. Instead of performing a constant-$`Q`$ scan at the zone-boundary $`X`$-point $`𝐐_0`$, we looked for the most effective focusing point in the vicinity of $`Q_0`$. We made constant-$`Q`$ scans at several different position $`𝐐^{}`$ on the arc $`𝐐^{}=𝐐_0+𝐪`$ as shown in Fig. 5(a), making a natural assumption that the phonon dispersion is isotropic near $`𝐐_\mathrm{𝟎}`$. Figure 5(b) shows the $`\theta `$ dependence of the line-width of the phonon spectrum for $`x=0.12`$ at $`T=315`$ K, where $`\theta `$ is defined as shown in Fig. 5(a). It is clearly seen that the line-width has a minimum at $`\theta 37^{}`$. We have thus chosen $`𝐐^{}=(\mathrm{3.09\; 0\; 1.88})`$ for $`x=0.10`$ and 0.12. The well-defined phonon spectrum for $`x=0.12`$ obtained at $`𝐐^{}`$ is shown in the inset of Fig. 5(b). As for $`x=0.18`$, $`𝐐^{}`$ was determined to be $`(\mathrm{3.10\; 0\; 1.96})`$ by using the same procedure.
Note that this method is based upon the assumption that, in the formula of phonon dispersion $`\omega _{\mathrm{ph}}^2(q)=\omega ^2(Q_0)+Aq^2`$, the coefficient $`A`$ is weakly temperature dependent. This assumption was justified by the work done by Birgeneau et al..
<sup>1</sup><sup>1</sup>footnotetext: Present address: Research Institute for Scientific Measurements, Tohoku University, Aoba-ku, Sendai 980-8577, Japan.
|
no-problem/9908/nucl-ex9908010.html
|
ar5iv
|
text
|
# Trace elements analysis of aerosol samples from some Romanian urban zones
## 1 Introduction
In this paper we have analyzed aerosols deposits on filters from ten Romanian towns: Pitesti, Giurgiu, Resita, Ramnicu-Valcea, Baia-Mare, Craiova, Timisoara, Calarasi, Braila and Arad with different kinds and levels of industrial development by method of particle-induced X-ray emission (PIXE). This method is based on the fact that the bombardment of the sample charged particles causes the ionisation of the atomic inner shells followed by a subsequent of the characteristic X-rays. When the X-rays spectrum is detected by a high resolution detector, the well-known Z-dependence of the X-rays energies, as well as the intensities of the individual X-rays line, allow a straight forward determination of elements present in the target. The properties of PIXE can summarized as: high sensitivity in small samples, high speed, surface analysis, genuinely multielemental and quantitative, partialy nondestructive, possible to combine simultaneously with other ion-beam techniques and microprobes.
The use of protons or alpha particles for the production of inner-shell vacancies combines a high ionization cross section with low X-ray background. The background in the region of low-Z elements is determined by the bremsstrahlung from secondary electrons while at higher X-ray energies the background is normaly determined by $`\gamma `$-rays produced in the target and the Compton electrons scattered in the crystal of the detector. The selection of various X-rays absorbers can improves the sensitivity over the whole elemental range. Although the ionization cross sections also increase for high elements with increasing particle energy up to rather high energies, the variation in the background radiation leads to the lowest general detection limits being obtained for 1.5-3.5 MeV protons. While the absolute detection limits in thick samples of low Z-elements are normaly in the interval from 0.1 to 10 $`\mu `$g/g. The advantages and disadvantages of the method as long in use as the PIXE analysis are well known and documented in several reviews, articles and textbooks .
In the present work we report the first PIXE analysis of atmospheric aerosols deposits filters from ten Romanian towns with different kinds and levels of industrialization.
## 2 Experimental method
The experimental set-up was described previously . The irradiation chamber has a 0.2 mm Be window for X-rays. The target was oriented at the 45 angle with respect to both the beam and detector direction. The beam passes through a collimator ($`\mathrm{\Phi }`$=2mm) before reaching the target. For analysis we used proton beams of 3 MeV energy supplied by the FN tandem accelerator from the Institute for Physics and Nuclear Engineering - Bucharest. The beam current was kept below 10 nA to maintain a count rate of about 250 counts/s, which implies negligible dead-time and pile-up corrections. X-rays were detected with a HPGe (100 mm<sup>2</sup>mm) detector with 160 eV energy resolution at 5.9 keV. Sample targets to be annualized were collected by the Institute of Hydrology and Waters of Bucharest and prepared in the following manner: aerosol particles were collected on cellulose fiber filter (Whatman 41). The flow rate was 15 to 20 liters per minute. Air volumes were measured with calibrated gasmeters with a precision of about 5% (to our regret half of samples have an unknown air volumes). The absolute concentrations of observed elements in aerosol samples were determined by advantage of the internal standard . This calibration method implies doping the sample with a known amount of the standard element and relating the unknown concentrations to those of the standard element. We choose Yttrium as the calibration standard because it is very rare element in the environment items. The intense peaks of Yttrium in the X-spectrum could obscure the peaks of some elements possible existing in the samples: L and K lines of Yttrium overlap SK<sub>α</sub> and (Rb and Sr)K lines respectively. Therefore we have analyzed targets without Yttrium too and we have not observed any new elements. A sample of Yttrium on Whatman 41 filter was measured too. Weak impurities of Ca, Fe and Zn were found. Concentrations of elements present in the aerosol samples were corrected for these impurities of the filter.
## 3 Results and discussions
A typical PIXE spectrum of an air filter sample is shown in Fig. 1. We have identified 15 elements: S, K, Ca, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, As, Hg and Pb. The measured elemental concentrations are given in $`\mu `$g per m<sup>3</sup> air in Table 1 for five samples with the known processed air volumes, and with respect to the concentration of the Ca for all analyzed samples in Table 2.
The great number of identified elements in the samples is similar to these of big cities as Livermore (USA) and Munich (Germany) . It is worth to mention the absence of the Cadmium element in our studied samples.
One could remark from the Table 1 the air filters from the town Pitesti, for the highest concentrations of the following elements: Ti: 9.63 $`\mu `$g/m3, V: 1.20 $`\mu `$g/m3, Ni: 0.59 $`\mu `$g/m3, S: 7.6 $`\mu `$g/m3
the town Resita for: Cr: 1.24 $`\mu `$g/m3, Mn: 10.04 $`\mu `$g/m3, Fe:253 $`\mu `$g/m3,
Baia-Mare for Cu: 36 $`\mu `$g/m3, Zn: 13.5 $`\mu `$g/m3, As: 5.09 $`\mu `$g/m3, Pb:12.1 $`\mu `$g/m3.
From the ratios of concentrations shown in the Table 2 we could make a comparison between the analyzed filters from all the towns considered here, from the point of view of the pollutant elements: the town Craiova is put in evidence by its high ratios of concentrations : Ti/Ca: 0.505, Cr/Ca: 0.035, Fe/Ca: 5.38, Co: 0.015, Zn/Ca: 0.036, As: 0.005 and Pb: 0.043. Calarasi has the highest ratio of Mn/Ca: 0.078 and the filter from Braila is put in evidence by the presence of Mercury, Hg/Ca: 0.003. Certainly the level of pollution of a region can not be determined by a single filter and it is need of a good statistics to draw conclusions. This work want rather to demonstrate that the PIXE method is a suitable tool in the analysis of air filters in the pollution studies and and also the results of this analysis draw the attention on the presence of the pollutants elements from the atmosphere of towns discussed in this paper.
Table 1. Concentrations in air filters, $`\mu `$g/m<sup>3</sup>, towns in Romania, by PIXE, with known processed air volumes
| Element | Pitesti | Giurgiu | Resita | Ramnicu-Valcea | Baia-Mare |
| --- | --- | --- | --- | --- | --- |
| S | 7.58 | - | - | - | - |
| K | 38.2 | 28.7 | 58.3 | - | 21.29 |
| Ca | 67.500 | 105.6 | 475 | 63.9 | 406.2 |
| Ti | 9.63 | 5.33 | 9.096 | 3.98 | - |
| V | 1.200 | - | - | 0.784 | - |
| Cr | 0.157 | 0.733 | 1.243 | 0.432 | - |
| Mn | 1.8 | 1.49 | 10.04 | 11.084 | 2.44 |
| Fe | 73.5 | 51.60 | 253 | 41.1 | 30.41 |
| Ni | 0.599 | - | - | 0.410 | - |
| Cu | 0.238 | - | 0.344 | 1.31 | 36.03 |
| Zn | 0.802 | - | 9.14 | 1.45 | 13.45 |
| As | - | - | - | 0.284 | 5.097 |
| Pb | - | - | - | 0.369 | 12.06 |
Table 2. Ratios of concentrations with respect to the concentration of the Ca, air filters from industrialzed towns in Romania, by PIXE
| Element | Pitesti | Giurgiu | Resita | Ramnicu-Valcea | Baia-Mare | Craiova | Timisoara | Calarasi | Braila | Arad |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| S | 0.112 | - | - | - | - | - | 0.02 | - | 0.012 | 0.047 |
| K | 0.57 | 0.272 | 0.122 | - | 0.052 | 2.51 | 0.2 | 0.26 | 0.174 | 0.36 |
| Ca | 1 | 1 | 1 | 1 | 1 | 1 | 1 | 1 | 1 | 1 |
| Ti | 0.14 | 0.05 | 0.019 | 0.063 | - | 0.505 | 0.15 | 0.66 | 0.017 | 0.046 |
| V | 0.018 | - | - | 0.012 | - | - | - | - | - | - |
| Cr | 0.002 | 0.007 | 0.003 | 0.007 | - | 0.035 | - | 0.12 | 0.006 | 0.001 |
| Mn | 0.027 | 0.014 | 0.021 | 0.017 | 0.006 | 0.05 | 0.007 | 0.078 | 0.024 | 0.017 |
| Fe | 1.09 | 0.49 | 0.532 | 0.647 | 0.075 | 5.38 | 0.52 | 4.32 | 0.28 | 0.58 |
| Co | - | - | - | - | - | 0.015 | - | - | - | - |
| Ni | 0.009 | - | - | 0.006 | - | - | - | - | - | - |
| Cu | 0.004 | - | 0.001 | 0.021 | 0.089 | - | 0.001 | 0.008 | 0.002 | 0.03 |
| Zn | 0.012 | - | 0.019 | 0.023 | 0.033 | 0.036 | 0.002 | 0.015 | 0.004 | 0.011 |
| As | - | - | - | 0.004 | 0.004 | 0.005 | - | - | - | 0.001 |
| Hg | - | - | - | - | - | - | 0.001 | - | 0.003 | 0.001 |
| Pb | - | - | - | 0.006 | 0.03 | 0.043 | - | 0.005 | 0.002 | 0.0003 |
|
no-problem/9908/astro-ph9908260.html
|
ar5iv
|
text
|
# Research Note Time-resolved spectroscopy of the peculiar H𝛼 variable Be star HD 76534
## 1 Introduction
In a previous paper we reported on the discovery of an H$`\alpha `$ burst in the B emission line star HD 76534 (Oudmaijer & Drew, 1997). During an observing run in 1995, two days after a strong H$`\alpha `$ emission line was observed, the star was re-observed, and found to have only a photospheric absorption line. Spectacularly, two hours later, the line was again in emission, still increasing its strength with respect to the continuum.
Especially marked spectral variations have been detected in a few other Be stars, notably $`\mu `$ Cen (Peters, 1986; Hanuschik et al. 1993, Rivinius et al. 1998) and $`\lambda `$ Eri (Smith, Peters & Grady 1991, Smith et al. 1997a+b). These two stars have long periods with relatively stable H$`\alpha `$ emission, and sometimes undergo an H$`\alpha `$ outburst where the emission grows to a maximum within days, displaying rapid variations of the violet and red peaks of the H$`\alpha `$ line. The emission then fades on longer timescales.
In contrast to both $`\mu `$ Cen and $`\lambda `$ Eri, the observed timescale of the H$`\alpha `$ outburst of HD 76534 was an order of magnitude shorter, while the line profile did not show any V/R variability. Instead, the line that was present only two hours after the absorption was observed, was similar in profile and V/R ratio to existing high resolution H$`\alpha `$ spectra of the object, and did not betray any signs of on-going formation of recently ejected disk material.
Two hypotheses were put forward by us to explain the rapid variations. In analogy with the hypothesis for $`\mu `$ Cen, a sudden burst of mass loss was first considered (see Hanuschik et al. 1993, Rivinius et al. 1998), but based on the above arguments it was discarded in favour of the idea that a stable rotating Keplerian disk was already present around the star, but that a lack of ionizing photons from, for example, the stellar photosphere failed to produce sufficient ionizations and subsequent recombinations to push the H$`\alpha `$ line into emission. From simple considerations, it was found that a slight change in ionizing flux can indeed ionize an existing stable neutral disk, and result in detectable H$`\alpha `$ emission.
It is not clear however where the change in ionizing radiation should come from. By analogy with the EUV variations of the $`\beta `$ Cep star $`\beta `$ CMa (Cassinelli et al. 1996), it is possible that stellar pulsations are responsible for this behaviour. $`\beta `$ CMa shows relatively large (30%) variations in its Lyman continuum, which are not as readily visible in the optical (Cassinelli et al. 1996). A similar effect could be happening in the case of HD 76534; at the times when the Lyman continuum is at minimum, no H$`\alpha `$ emission is visible, while at maximum the line will develop. A critical test of the stellar pulsation hypothesis would be to monitor the star for several hours up to several days to investigate whether any periodicity would be present in the H$`\alpha `$ emission of the object.
On the other hand, Smith et al. (1997b) reaches a similar conclusion to explain, amongst other phenomena, the H$`\alpha `$ variations in $`\lambda `$ Eri. A source of extra Lyman continuum photons could be responsible for extra ionizations and recombinations in the circumstellar material. Smith et al. (1997b) find that this can be explained by the occurrence of heated slabs, possibly related to magnetic activity, close to the stellar photosphere. This activity does not appear regularly, so no apparent periodicity in the H$`\alpha `$ line, certainly not on timescales of hours, would be expected.
Previous to Oudmaijer & Drew (1997), observations of the H$`\alpha `$ line of HD 76534 were reported only twice, by Thé et al. (1985) and Praderie et al. (1991). Their spectra were taken one year apart, and showed ‘indistinguishable’ (cf. Praderie et al. 1991) line profiles. We measure an equivalent width (EW) from Thé et al. of –7 $`\mathrm{\AA }`$. The exposure times used by these authors were 2 and 2.5 hours respectively, so that any shorter term variations in either the line profile or EW would have been washed out.
Since our previous paper, several new datasets on HD 76534 have been published. Recently, Corcoran & Ray (1998), report a measurement of the EW of the H$`\alpha `$ of –14.3 $`\mathrm{\AA }`$, on a spectrum that was obtained in the last week of 1991 (their only Southern observing run) and Reipurth, Pedrosa & Lago (1996) show a spectrum, obtained in February 1993, with an EW of –10 $`\mathrm{\AA }`$. These two papers do not mention the exposure times. Oudmaijer & Drew (1999) report on observations taken in December 1995 and December 1996 and found an EW of –6 $`\mathrm{\AA }`$ and –4 $`\mathrm{\AA }`$ respectively (we note a typographical error in their Table 2, the data for 30 December 1996 and December 1995 should be interchanged). It is clear that the H$`\alpha `$ line of HD 76534 is variable on long timescales.
The main aim of this paper is to present time-resolved observations of the object at high resolution as a first step in determining whether H$`\alpha `$ is prone to any variations on timescales of minutes. The observations thus serve as a check whether the long integration times by Praderie et al. (1991) and Thé et al. (1985) would indeed have washed out any short-term variations.
This Research Note is organized as follows: In Sec. 2 we present the time-resolved observations of the object obtained with the UCLES spectrograph on the AAT. In Sec. 3 we will search for variations in the data and revisit the multi-epoch V band photometry obtained by Hipparcos. We will conclude in Sec. 4.
## 2 The Observations
During the night of 7 June 1996 (UT), HD 76534 was observed in service time, employing the UCLES spectrograph mounted on the 3.9 m Anglo-Australian Telescope. The observational set-up included the 31 grooves/mm echelle and a 1024$`\times `$1024 Tektronix CCD detector. The resulting spectrum contains 44 orders covering the spectral coverage of 4650 to 7240 Å, including both H$`\alpha `$ and H$`\beta `$. The observing strategy was fairly straightforward, we aimed to obtain spectra every few minutes, and employed different exposure times in order to compromise between time-resolution and signal-to-noise. Within a time span of two hours, we obtained 30 spectra with signal-to-noise ratios (SNR) in the H$`\alpha `$ setting ranging from $``$ 60 in the shortest (60s) exposures to $``$ 140 in the longest (240s) exposures. The observing conditions were not ideal, so the total count-rates and SNR are slightly variable.
Data reduction was performed in iraf (Tody, 1993), and included the procedures of bias-subtraction, flatfielding and wavelength calibration. An observation of a Th-Ar arc lamp was used to provide the wavelength calibration of each spectrum. The resulting spectral resolution was determined to be $``$ 10 km s<sup>-1</sup> from Gaussian fits through telluric absorption lines.
In total 30 spectra with different exposure times were taken. The summed total spectrum, with a total exposure time of 4500s has a SNR of $``$200 in the H$`\beta `$ order and $``$300 in the H$`\alpha `$ order. The log of the observations and the SNR in the H$`\alpha `$ order are provided in Table 1.
## 3 Results
### 3.1 Short description of the spectrum
Apart from the strong H$`\alpha `$ emission (EW = –8 $`\mathrm{\AA }`$) and the filled-in H$`\beta `$ absorption, no apparent emission lines are found. Strong Hei lines at 4921, 5047, 5876 and 6678 $`\mathrm{\AA }`$ show absorption and allow for a determination of the stellar velocity, which is found to be 14 $`\pm `$ 3 km s<sup>-1</sup> (heliocentric). This is consistent with other determinations (see Oudmaijer & Drew 1997). Within the observational error-bars, we do not find evidence for radial velocity variations.
The continuum corrected spectra around the H$`\alpha `$ and H$`\beta `$ lines are shown in Fig. 1. H$`\alpha `$ is a strong doubly-peaked emission line, while the doubly-peaked H$`\beta `$ emission hardly reaches above the local continuum. The peak separation in the H$`\beta `$ line is 242 $`\pm `$ 5 km s<sup>-1</sup> which is almost twice as large as observed in the H$`\alpha `$ line (143 $`\pm `$ 5 km s<sup>-1</sup>). The larger peak separation in H$`\beta `$ is consistent with the fact that both lines are formed in a rotating Keplerian type disk. If both lines, or H$`\alpha `$ alone, are optically thick, the H$`\alpha `$ forming region is larger than the H$`\beta `$ forming region, and will thus trace lower rotational velocities.
### 3.2 Variability on short time scales
#### 3.2.1 Method
Visual inspection of the 30 individual spectra did not reveal any obvious variability in the H$`\alpha `$ line, so instead, we adopt the simple statistical formalism presented by Henrichs et al. (1994), and already exploited by Oudmaijer & Bakker (1994) for a similar experiment for the post-AGB star HD 56126. This method was devised by Henrichs et al. to spot the regions of interest in their multi-epoch data on hot stars. The variability can be expressed in a temporal variance spectrum (TVS):
$$(TVS)_\lambda \frac{1}{N1}\underset{i=1}{\overset{N}{}}\left(\frac{F_i(\lambda )F_{av}(\lambda )}{\sigma _i(\lambda )}\right)^2$$
(1)
where $`N`$ is the number of spectra, $`F_{av}(\lambda )`$ represents the constructed average spectrum, $`F_i(\lambda )`$ the individual spectra, and $`\sigma _i(\lambda )`$ = $`F_i(\lambda )`$/(SNR) of each individual pixel of the spectra.
Then, the so-called temporal sigma spectrum, TSS = $`\sqrt{TVS}`$), is calculated. This quantity represents approximately ($`\sigma _{obs}/\sigma _{av}`$). $`\sigma _{obs}`$ traces the standard deviation of the variations of the individual spectra with respect to the average spectrum, while $`\sigma _{av}`$ represents the standard deviation of the average spectrum. If no significant variations are present in the spectra, the ratio of these two numbers, $`\sigma _{obs}/\sigma _{av}`$, will be close to one, deviations are directly represented in units of the noise level, that is to say a peak ‘Temporal Sigma Spectrum’ of three corresponds to a variability at a 3$`\sigma `$ level.
During the extraction process, IRAF provides a SNR spectrum based on the photon-statistics of the data. This is very convenient, as the SNR changes strongly over a given order because the blaze of the spectrograph results in lower count-rates and thus lower SNR at its edges, while, of course, the countrates and SNR also change across absorption and emission lines compared to the local continuum. As a check, we measured the SNR in several wavelength intervals. The IRAF-extracted SNR were scaled up by 40% to bring the measured and the IRAF SNR in agreement. In the remainder of this exercise, we will use these SNR spectra as input for Eq. 1.
The average spectrum was constructed by summing all individual spectra. After this, the spectra were continuum rectified as input for Eq. 1. In the case of the H$`\alpha `$ line, the wavelength region used for the fit were the blue and red continua beyond 13 $`\mathrm{\AA }`$ from the center of the line. The $`\sigma (\lambda )`$ spectrum was computed by dividing the individual continuum rectified spectra by their respective SNR spectra.
Fig 2 shows an example of the usefulness of the method. In the middle panel the total spectrum in the order around the telluric absorption bands at $``$ 6870 $`\mathrm{\AA }`$ are shown, the lower panel shows the derived TSS. The continuum shows a trend from TSS $``$ 0.9 to $``$ 1.1, indicating a slight variation in the continuum of the extracted spectrum. The telluric lines vary however at the 2-4$`\sigma `$ level. This is due to the changing airmass during the observations: as the airmass increases from 1.30 (zenith distance 39<sup>o</sup>) to 2.0 (59<sup>o</sup>) the telluric absorption becomes stronger. This is visible in the upper panel which shows an overplot of the first and last spectrum taken. The changes, which are only a few percent of the continuum level, are real.
The fact that a slight variability is traced in the continuum illustrates a very important caveat of the method. The TSS only depends on photon-statistics, and is insensitive to any systematic errors that may be present. In particular, a variable response curve (‘blaze’) of the echelle, can show up as variability, while in reality such variations are purely systematic rather than intrinsic. In the case of Fig 2 this is not so important, as the entire order can be used for the continuum rectification, effectively removing this effect. In the case of H$`\alpha `$ however, a large part of the spectrum can not be used for the continuum rectification as, of course, it is covered by the H$`\alpha `$ line itself.
#### 3.2.2 The H$`\alpha `$ line
Fig. 3 shows the resulting TSS for the H$`\alpha `$ order. The telluric lines show variability at the 2.5$`\sigma `$ level. This is smaller than the variability observed for telluric lines in Fig. 2, but can be explained by the fact that these lines and their changes are weaker. It is nevertheless an important check to note that the method also works in the H$`\alpha `$ order.
H$`\alpha `$ itself hardly shows any variability. In fact, the most significant variability is due to the telluric absorption in the central minimum of the line. The low $`<`$1.5$`\sigma `$ variability observed across the line is statistically not significant. Nevertheless, we investigated the possible cause of these marginal variations. This was done bearing in mind the fact that a heavily rebinned spectrum has a much larger SNR, and thus any variations should show up with greater significance.
Unfortunately, the echelle orders’ wavelength ranges are rather limited (about 67 $`\mathrm{\AA }`$ or 3000 km s<sup>-1</sup> in the H$`\alpha `$ order) compared to the extent of the line itself (full-width at zero intensity $``$ 1000 km s<sup>-1</sup>). It is thus possible that systematic variations in the continuum interpolated underneath H$`\alpha `$ may be mis-interpreted as revealing intrinsic variations in the line. Indeed, by dividing all individual spectra around H$`\alpha `$ by the same (rescaled) continuum fit, it became clear that the curvature of the spectra varies in time on a level less than a few %, having biggest impact on the red end of the spectrum. This is probably related to a well-known varying blaze due to the de-rotator optics in UCLES.
In order to check whether the line may be intrinsically variable, we performed some tests. The main reasoning behind these tests is that if the response curve of the echelle is variable in time, the adjacent (line-free) orders should show a similar variability. We therefore investigated the two orders next to the H$`\alpha `$ order in the echellogram, and continuum rectified these using the same pixel range as the H$`\alpha `$ order, i.e. not using the $``$ 25 $`\mathrm{\AA }`$ around the center of H$`\alpha `$, and looked for evidence of variability.
We measured the EW of a fiducial line over the same pixel-range in these orders (corresponding to 26 $`\mathrm{\AA }`$) as H$`\alpha `$. The measured EW in both orders is close to 0 $`\mathrm{\AA }`$, but has a scatter of 0.12 $`\mathrm{\AA }`$. This is to be compared with the scatter in the EW of the H$`\alpha `$ line of 0.21 $`\mathrm{\AA }`$. Based on the variations of the EW of the fiducial lines in the continuum of the adjacent echelle orders and the mean height of H$`\alpha `$ line over the measured interval, we would expect a scatter of 0.16 $`\mathrm{\AA }`$ in the EW of H$`\alpha `$. The scatter of the EW of H$`\alpha `$ is slightly larger than this, corresponding to variability at a 1.3 $`\sigma `$ level.
Having established that the total H$`\alpha `$ EW is hardly variable, the question is whether this is because the total line is not variable at all, or whether the line-profile changes in such a way that the total line-flux is nearly constant. Checks on the individual spectra show that the small changes in the line-profile are in phase with each other, and more importantly, in phase with changes in the red continuum flux. This indicates that the line-profile as such does not vary, while it traces the changes in the continuum level. Hence the insignificant variability in the EW is not due to changes in the line-profile.
Based on the facts that the EW of the H$`\alpha `$ line changes at a similar amplitude as the EW in the same pixel-range of the adjacent orders, and that the ‘changes’ traced by the TSS spectrum are in phase with the red continuum, we conclude that during the two hours of these observations, no significant variability was present in the H$`\alpha `$ line of HD 76534.
### 3.3 Hipparcos photometry re-visited
The Hipparcos satellite observed HD 76534 125 times between 1989 and 1993 photometrically in a passband similar to the V band (ESA 1997). These observations were reported on in the paper on Herbig Ae/Be stars by van den Ancker, de Winter & Tjin A Djie (1998). These authors mentioned that the star is probably photometrically variable, which they based on the fact that the variance of the individual photometric points is larger than the observational errors. No lightcurves were provided however.
Here we look at the data provided by the Hipparcos catalogue into more detail. The photometry is plotted as function of Julian Date in Fig. 4. In the first year of the mission, HD 76534 was constant within the errorbars until the object brightened by about 0.1 magnitude, reaching a maximum around May 1991. The period of brightening and fading lasted about 100 days. Afterwards the object ‘flickers’ around a mean value close to what was measured before the maximum.
Could this rise in brightness be associated with the spectral behaviour of the object? Mennickent et al. (1998) published 11 year long photometric monitoring of $`\lambda `$ Eri, and found several similar changes in the Strömgren photometry of the object. A period search revealed that rises in brightness of $``$0.1 mag. occurred with a period of 486 days, while the rising and fading of the object lasted about 100 days. From the colour changes, they found that the brightness maxima correspond to slight increases in effective temperature of the star. Although the overlap between spectroscopic and photometric data is not very complete, Mennickent et al. find a rough correlation between the jumps in brightness and periods of H$`\alpha `$ emission in the star.
## 4 Concluding remarks
In this paper we have investigated whether the H$`\alpha `$ flaring star HD 76534 is variable on short timescales of order minutes and hours. We find no evidence for statistically significant variations. One of our original suggestions to explain the H$`\alpha `$ outburst in 1995 was to invoke a pulsation-type behaviour of the star. If this would have been the case, we might have seen at least some variability on short time scales, but these are not seen.
Including the current data, H$`\alpha `$ measurements for this object have now been reported 9 times in the astronomical literature (see the Introduction), and all but one show the line in emission. Considering this, one would conclude that the observed collapse and the subsequent rapid recovery of the H$`\alpha `$ emission is only a sporadic event. If we were to associate the brightening of the object in the Hipparcos photometric data to a strong emission variability, it may be that the relevant timescale is the hundreds of days between events. A monitoring programme sampling a range of timescales is needed to clear this up.
In terms of variability, HD 76534 can be regarded as belonging to the same class as $`\mu `$ Cen and $`\lambda `$ Eri. But there are some striking differences. In the first place, the H$`\alpha `$ variations in 1995 were the strongest and fastest ones observed to date in a Be star. Secondly, whereas $`\mu `$ Cen and $`\lambda `$ Eri are most often observed in a quiescent state, interrupted by bursts of emission that gradually declines, HD 76534 is most often observed with H$`\alpha `$ emission, interrupted by absorption, suggesting a different process governing the H$`\alpha `$ behaviour. Indeed, the H$`\alpha `$ line that was observed only two hours after it had been in absorption showed a smooth doubly-peaked emission profile, similar to what has been observed now, suggesting that a neutral disk was present before the source of ionizing photons increased in strength contrary to the common idea that a disk has been built up in the case of $`\mu `$ Cen and $`\lambda `$ Eri.
##### Acknowledgments:
We thank Myron Smith for interesting and inspiring discussions. We thank the staff at the Anglo-Australian Telescope for their expert advice and support. RDO is funded by the Particle Physics and Astronomy Research Council of the United Kingdom. The data analysis facilities are provided by the Starlink Project, which is run by CCLRC on behalf of PPARC. This research is partly based on results from the ESA Hipparcos astrometric satellite.
|
no-problem/9908/astro-ph9908241.html
|
ar5iv
|
text
|
# Extinction and polarization of radiation by absorbing spheroids: shape/size effects and benchmark results
## 1 Introduction
In many scientific and engineering applications prolate and oblate spheroids are appropriate models for real particles. The scattering of electromagnetic radiation by spheroids is usually calculated by the separation of variables method ay ; vf (SVM) and the T-matrix method bh (TMM). Although both approaches are well-suited for numerical computations, it is often difficult to give reliable error estimates. For non-absorbing particles, there always exists an internal check as the absorption cross-sections (and efficiency factors) must be equal to zero. Obviously, for particles consisting of absorbing materials like silicates, graphite or metals, this test does not work. Despite the fact that rather elaborated methods for the calculation of the optical properties of highly absorbing non-spherical particles can be found in literature (see, for example, Michel et al. mich ), these methods have never been used for detailed numerical studies.
In this paper, we consider the light scattering by homogeneous spheroids using the most popular approaches – SVM and TMM. The extinction and polarization cross-sections for prolate and oblate spheroids of various aspect ratios $`a/b`$ are calculated for the refractive indices $`m`$ typical of soot, graphite, iron, silicate, and water ice and compared with the results for the particles of the same volume. A part of the results is presented in tabular form serving as benchmarks for forthcoming calculations. We also describe a combined approach to calculate the light scattering by axisymmetric particles. The method should allow to treat easily (very) elongated and flattened absorbing particles.
## 2 Description of methods
### 2.1 General definitions
A spheroid (ellipsoid of revolution) is obtained by the rotation of an ellipse around its major axis (prolate spheroid) or its minor axis (oblate spheroid). The ratio of the major semiaxis $`a`$ to the minor semiaxis $`b`$ (i.e. the aspect ratio $`a/b`$) characterizes the particle shape which may vary from a nearly spherical one ($`a/b1`$) to a needle or a disk ($`a/b1`$).
We assume an incident plane wave having the wavelength $`\lambda `$. Let $`\alpha `$ denote the angle between the propagation direction and the rotation axis of the spheroid ($`0^{}\alpha 90^{}`$).
For axial propagation ($`\alpha =0^{}`$), there is no polarization of transmitted radiation due to symmetry. If $`\alpha 0^{}`$, two cases of polarization of the incident radiation have to be considered: the electric vector $`\stackrel{}{E}`$ is parallel (TM mode) or perpendicular (TE mode) to the plane defined by the spheroid’s rotation axis and the wave propagation vector.
The size parameter is given by
$$x_\mathrm{V}=\frac{2\pi r_\mathrm{V}}{\lambda },$$
(1)
where $`r_\mathrm{V}`$ is the radius of the sphere whose volume is equal to that of the spheroid. The radius $`r_\mathrm{V}`$ is defined as
$$r_\mathrm{V}^3=ab^2\text{for prolate spheroids}$$
(2)
and
$$r_\mathrm{V}^3=a^2b\text{for oblate spheroids.}$$
(3)
One usually calculates the efficiency factors $`Q=C/G`$ which are the ratio of the corresponding cross-sections $`C`$ to the geometrical cross-section $`G`$ of the spheroid (the area of the particle’s shadow)
$$G(\alpha )=\pi b\left(a^2\mathrm{sin}^2\alpha +b^2\mathrm{cos}^2\alpha \right)^{1/2}\text{for prolate spheroids}$$
(4)
and
$$G(\alpha )=\pi a\left(a^2\mathrm{cos}^2\alpha +b^2\mathrm{sin}^2\alpha \right)^{1/2}\text{for oblate spheroids.}$$
(5)
In order to compare the optical properties of the particles of different shapes it is convenient to consider the ratios of the cross-sections for spheroids to the geometrical cross-sections of the equal volume spheres, $`C/\pi r_\mathrm{V}^2`$. They can be found as
$$\frac{C}{\pi r_\mathrm{V}^2}=\frac{[(a/b)^2\mathrm{sin}^2\alpha +\mathrm{cos}^2\alpha ]^{1/2}}{(a/b)^{2/3}}Q$$
(6)
for a prolate spheroid and
$$\frac{C}{\pi r_\mathrm{V}^2}=\frac{[(a/b)^2\mathrm{cos}^2\alpha +\mathrm{sin}^2\alpha ]^{1/2}}{(a/b)^{1/3}}Q$$
(7)
for an oblate spheroid.
The albedo of a particle can be calculated from the extinction and scattering cross-sections
$$\mathrm{\Lambda }=\frac{C_{sca}}{C_{ext}}.$$
(8)
The dichroic polarization efficiency is defined by the extinction cross-sections for TM and TE modes
$$\frac{P}{\tau }=\pm \frac{C_{ext}^{\mathrm{TM}}C_{ext}^{\mathrm{TE}}}{C_{ext}^{\mathrm{TM}}+C_{ext}^{\mathrm{TE}}}100\%,$$
(9)
where the upper (lower) sign is related to prolate (oblate) spheroids. This ratio describes the efficiency of the polarization of light transmitted through the uniform slab consisting of non-rotating particles of the same orientation.
The optical properties of spheroidal particles can be determined by various methods of light scattering theory. Most frequently, the separation of variables method obtained by Asano & Yamamoto ay and Farafonov vf ; v and the T-matrix method bh ; mtm are used. The calculations can be also performed with the discrete dipole approximation (DDA) which can be applied to arbitrarily inhomogeneous and irregular particles, but requires large computation time and is less accurate than the other methods (see discussion in the papers hov ; vis ).
### 2.2 Asano & Yamamoto’s solution for spheroids (SVM1)
Historically it was the first approach to the light scattering problem for spheroids with a complex refractive index. The method is based on the solution to the Helmholtz equation in the spheroidal coordinate system.
In their pioneering paper, Asano & Yamamoto ay used the Debye potentials to describe the electromagnetic fields, which is similar to the Mie solution for spheres. The scattering coefficients then are bound in the infinite systems of the linear algebraic equations and can be found by solving the truncated systems.
In our comparison, we made the calculations with the numerical code of Peter Martin (see rm ; km ).
### 2.3 Farafonov’s solution for spheroids (SVM2)
The principal distinction of this solution (see vf ; vf851 ) from the previous one is the special basis for the representation of the electromagnetic fields – a combination of the Debye and Hertz potentials (i.e. the potentials introduced to solve the light scattering problem for spheres and infinitely long cylinders, respectively).
The approach has an incontestable advantage for strongly elongated or flattened particles. In this paper, the most recent version of our numerical code has been used. It involves different methods to calculate the spheroidal functions with the automatic choice of the most appropriate method among them.
### 2.4 Solution for spheroids by the T-matrix method (TMM1 and TMM2)
Besides SVM, another very popular approach to solve the light scattering problem for spheroids is the T-matrix method. It is mainly applied to the axisymmetric particles: finite cylinders, spheroids, Chebyshev particles.
The main idea of the method is the expansion of the incident, internal and scattered radiation fields in terms of the vector spherical harmonic functions. Because of the linearity of the Maxwell equations, the relations between the expansion coefficients of the fields are linear and given by two matrices. Thus, solution consists of calculations of the matrix elements being the surface integrals and inversion of a matrix. The advantages of the approach are simple compact codes and their high speed. Another advantage is also the ability to treat more the complex case of elastic waves, the potential applicability to particles of complex shape, and the possible analytical averaging over the particle orientations within the expansion coefficients which can greatly enlarge the field of the TMM applications.
We used the T-matrix codes of Barber and Hill bh (TMM1) and Mishchenko (TMM2; see mtm and references therein).
### 2.5 A comparison of methods
We do not intend here to compare in detail the computational efficiency of different methods. The goal of this subsection is only to illustrate it on one example. A crude impression of the range of applicability of the methods may be obtained from Fig. 1 where the normalized cross-sections are plotted for oblate spheroids with $`m=1.7+0.7i`$ and $`a/b=4`$. In all four cases double precision was used.
It should be noted that the advantage of the T-matrix codes appears mainly for particles more spherical than presented in the Fig. 1. On the other hand, for larger aspect ratios, the SVM2 code becomes unrivalled.
### 2.6 A new method for axisymmetric particles
Each of the methods (SVM and TMM) has its own problems. We suggest the following approach that could solve some of them.
The incident, scattered and internal fields are divided in two parts: axisymmetric and non-axisymmetric ones
$$\stackrel{}{E}=\stackrel{}{E}_1+\stackrel{}{E}_2,\stackrel{}{H}=\stackrel{}{H}_1+\stackrel{}{H}_2.$$
(10)
It is possible to do so that the axisymmetric part does not depend on the azimuthal angle ($`\phi `$) and the averaging of the non-axisymmetric part over all the azimuthal angle values gives zero vf . Then the scattering problem can be solved for each of the parts separately.
The scattering problem is formulated in the integral form like in the TMM (e.g., bh ). The Abraham’s potentials
$$p=E_{1\phi }\mathrm{cos}\phi ,q=H_{1\phi }\mathrm{cos}\phi $$
(11)
are introduced for the axisymmetric part. A superposition of the Debye and Hertz potentials is used for the non-axisymmetric part:
$$\stackrel{}{E}_2=\stackrel{}{}\times (U\stackrel{}{e}_z+V\stackrel{}{r}),\stackrel{}{H}_2=\frac{1}{i\mu k_0}\stackrel{}{}\times \stackrel{}{}\times (U\stackrel{}{e}_z+V\stackrel{}{r})$$
(12)
for TE mode and
$$\stackrel{}{E}_2=\frac{1}{i\epsilon k_0}\stackrel{}{}\times \stackrel{}{}\times (U\stackrel{}{e}_z+V\stackrel{}{r}),\stackrel{}{H}_2=\stackrel{}{}\times (U\stackrel{}{e}_z+V\stackrel{}{r})$$
(13)
for TM mode. Here $`\epsilon `$ is the complex permittivity, $`\mu `$ the magnetic permeability, $`k_0`$ the wavevector in vacuum, $`\stackrel{}{r}`$ the radius-vector, $`\stackrel{}{e}_z`$ the unit vector, and $`i`$ the complex unity.
The scalar potentials are expanded in terms of the spherical wave functions. The coefficients of the expansions are determined from solution to the algebraic systems similar to those obtained in the TMM (e.g., bh ).
Thus, the new approach should combine the strong aspects of the basic methods: the simple solution scheme typical of the TMM and the ability to treat particles whose shape can be very elongated or flattened as in the SVM2. The theoretical description of this approach is given in f the development of a numerical code realizing it is in progress fih .
## 3 Numerical results
Here we present some results illustrating the behaviour of the optical properties of spheroidal particles for the case of transmitted radiation. Because this is the first consideration of optical properties of highly absorbing non-spherical particles beyond the Rayleigh limit, we try to demonstrate the features of such particles in extinction, scattering and polarization.
Note that in the standard case of light scattering by spheres (Mie theory) only the refractive index and the size parameter of particles may be varied. For the homogeneous spheroids, one can also study the effects of the particle shape (the aspect ratio $`a/b`$) and orientation (the angle $`\alpha `$) as well as that of the polarization state of the incident radiation (TM and TE modes) for both elongated (prolate spheroids) and flattened (oblate spheroids) particles.
We consider spheroids with aspect ratios $`a/b=1.1,2,4,`$ and 10 in a fixed orientation. The first $`a/b`$ value illustrates the influence of small deviations from the spherical shape. All the calculations presented in this Section were performed using our new SVM2 code.
### 3.1 Extinction
Figures 2 and 3 show the normalized extinction cross-sections for non-absorbing and absorbing spheroids with different aspect ratios. The similar scale of the $`y`$axes allows to see clearly the shape effects for the particles of equal volume.
Note that large oscillations and the ripple-like structure disappear for absorbing spheroids – the curves become very smooth. These features depend mainly on the value of the imaginary part of the refractive index. The curves $`C_{\mathrm{ext}}/\pi r_V^2(x)`$ for $`\alpha =0^{}`$ and $`\alpha =90^{}`$ differ more and more when the ratio $`a/b`$ grows (Fig. 3). For absorbing spheroids, the position of the first (main) maximum slightly shifts to larger values of $`x_V`$ with increasing $`a/b`$, while for non-absorbing particles, we have the opposite case.
Now let us consider the particle of a fixed volume and begin to change its shape from a sphere to a needle or disk. The extinction by the particle is shown in Fig. 4 for three values of the refractive index. As follows from this Figure, in almost all cases the extinction cross-sections seem to reach their asymptotic values already at $`a/b10`$. It is interesting that the extinction by non-spherical transparent particles is usually smaller than that by spheres. The absorbing spheroids normally remove from the incident beam 1.5 – 2 times more energy than spheres of the same volume. Though the behaviour of $`C_{\mathrm{ext}}`$ is refractive index/size dependent, the conclusions we made are rather general for absorbing spheroids. From Figs. 3, 4 one can also see that even after averaging over particle’s orientations we should get a stronger extinction for particles with larger aspect ratios $`a/b`$.
### 3.2 Scattering and albedo
If the imaginary part of the refractive index is not zero, we can study independently the scattering and absorbing properties of particles. In the case presented as an example in Fig. 3, the behaviour of extinction is mainly determined by that of absorption.
The integral scattering properties of particles are characterized by their albedo (see Eq. (8)). This quantity depends on the particle size and, in general, on the particle shape. As far as we know the shape effects have not yet been discussed in the literature. Figure 5 shows the size dependence of the albedo for spheroids with $`m=1.7+0.7i`$. The calculations were made for the particles with four aspect ratios $`a/b=1.1,2,4,10`$ and for two orientations $`\alpha =0^{}`$ and $`90^{}`$ (we adopt that the incident radiation is non-polarized).
It can be clearly seen that the albedo for non-spherical particles whose sizes are larger than a critical value ($`x_V23`$) does not deviate more than $`20\%`$ from that of spheres. The largest deviations in Fig. 5 occur for the aspect ratio $`a/b=10`$ and $`\alpha =0^{}`$ (prolate spheroids) and $`\alpha =90^{}`$ (oblate spheroids). Note that for arbitrarily oriented particles (3D-alignment), the distinction between the albedo of spheres and spheroids should be smaller (see Fig. 11 in mtm where the albedo for particles with $`m=1.53+0.008i`$ and $`a/b=2`$ is plotted).
Our calculations made for particles with different absorption show that the distinction of the albedo for spheres and spheroidal particles remains rather small (within $`20\%`$) if the ratio of the imaginary part of the refractive index to its real part $`k/n\begin{array}{c}>\\ \end{array}0.20.3`$, which corresponds to $`k0.30.4`$ if $`n=1.7`$. Thus, we can conclude that the albedo of spheroidal particles from soot, graphite, iron very slightly depends on the particle shape.
### 3.3 Polarization
If a volume contains aligned non-spherical particles, the initially non-polarized incident radiation will be partially polarized after passing the volume. The polarization degree can be used to estimate the alignment degree of the particles. The simplest and at the same time extreme case of particles’ alignment is the perfect alignment of non-rotating particles (picked fence orientation). The maximum polarization usually occurs when the major axis of the particles is perpendicular to the direction of the incident radiation ($`\alpha =90^{}`$).
The previous calculations of the polarization (see vf ; rm ; km ) did not deal with large, highly absorbing particles. However, such particles exist in nature and can be aligned under terrestrial and extraterrestrial conditions.
The behaviour of the polarization efficiency $`P/\tau `$ (see Eq. (9)) for non-absorbing and absorbing spheroids is shown in Fig. 6 for the case when a maximum polarization is expected (picked fence alignment, $`\alpha =90^{}`$). It is clearly seen from this Figure that the relatively large particles produce no polarization—independent of their shape. For absorbing particles, it occurs at smaller $`x_V`$ values than for non-absorbing particles. Note also that the polarization reversal takes place for disk-like particles. This effect seems to depend on the imaginary part of the refractive index as well.
### 3.4 Benchmarks
The results of computations for spheroids were previously published in the tabular form only in the papers of Voshchinnikov & Farafonov vf851 ; vf852 ; vf86 (efficiency factors for spheroids with the refractive index $`m=1.2`$), Kuik et al. khh and Hovenier et al. hov (the scattering matrices for spheroids with $`m=1.5+ki,k<0.01`$). Extensive calculations of the elements of the scattering matrices for randomly oriented slightly absorbing spheroids with $`a/b2`$ were performed by Mishchenko and others (see mtm and references therein).
In Tables 1 – 4, we present the normalized extinction and scattering cross-sections for prolate and oblate spheroids with the refractive indices $`m=2.5+0.0i,1.7+0.7i,2.5+1.5i`$ and $`3.0+4.0i`$. The results — all 7 (or in a few cases 6) significant digits — were obtained with at least two of four codes used (see Sects. 2.2–2.4).
Editor: Tables 1 - 4 are put here!!!
The numbers printed by italics in Table 1 were calculated using the SVM2 code only. As the extinction and scattering cross-sections were the same (with 8 and more digits), the values presented in the Table are expected to be correct.
## 4 Conclusions
On the basis of the solutions to the light scattering problem by the separation of variables method and the T-matrix method we have considered the optical properties of absorbing spheroidal particles of different aspect ratios. Distinct approaches allowed to get the reliable results which may be used as benchmarks.
The following conclusions can be made from the study:
1. The extinction cross-sections for highly absorbing spheroids are typically 1.5 – 2 times larger than those for spheres of the same volume.
2. The albedo of non-spherical particles exhibits only a weak dependence on the particle shape and is determined mainly by the imaginary part of the refractive index.
3. For particles larger than a certain minimum size, the spheroidal particles do not polarize the transmitted radiation independent of their shape.
Acknowledgements
This work was supported by a grant from the Volkswagen Foundation.
|
no-problem/9908/cond-mat9908167.html
|
ar5iv
|
text
|
# Exact Solutions of a Model for Granular Avalanches
## 1 Introduction and Model
The study of avalanches and surface flows in granular materials has attracted much attention recently, both from a theoretical and an experimental point of view . A simple model, thought to capture some of the essential phenomena, has been proposed in . It is based on the assumption that a strict separation between rolling grains and static grains can be made. Coupled dynamical equations for these two species, based on phenomenological arguments, can then be written. Calling $`R`$ the local density of rolling grains and $`h`$ the height of static grains, the simplest form of the bcre equations read:
$`H_t`$ $`=`$ $`\gamma RH_x,`$ (1)
$`R_t`$ $`=`$ $`R_x+\gamma RH_x,`$ (2)
where $`H`$ is the height of static grains, counted from the repose slope of angle $`\theta _r`$: $`h(t,x)=H(t,x)+x\mathrm{tan}(\theta _r)`$ (the heap is sloping upwards from left to right). In the above equations, the units of lengths and time are chosen such that the (downhill) velocity of grains is $`v=1`$, while $`H`$ and $`R`$ are counted in units of the grains diameter. The term $`\gamma RH_x`$ describes the conversion of static grains into rolling grains if $`H_x>0`$, or vice versa if $`H_x<0`$. $`\gamma `$ is a grain collision frequency, typically of the order of $`100`$ Hz.
Many important phenomenon are left out from the above description, and can be included by adding more terms. For example, diffusion terms (such as $`D_1R_{xx}`$ or $`D_2H_{xx}`$, describing, e.g. non local dislodgement effects) will generically be present, and qualitatively change the structure of the solutions . Another aspect not described by the linear form of the conversion term above is the expected saturation of rolling grains with time, rather than the exponential growth predicted by Eq. (2) for a constant positive slope $`H_x`$. Non linear saturation terms, as well as a dependence of the velocity of the rolling grains on $`R`$, are thus expected in general, and can lead to important differences with the above equations .
Recently, these equations has been studied by Mahadevan and Pomeau (mp) . They found a conservation law, which relates the solutions $`R(t,x)`$ and $`H(t,x)`$ in a frame moving with the velocity of the grains. From this law, they concluded that the bcre equations have characteristics that are straight lines, along which both $`R(t,x)`$ and $`H(t,x)`$ are constant. Independently of the initial profile $`H_0(x)`$, they found that a shock forms at time $`t_s=1/(\gamma R_{0,\mathrm{max}}^{})`$ with $`R_{0,\mathrm{max}}^{}`$ is the maximum (in absolute value) of the initial gradient of rolling grains. Whereas our exact solution fulfills the same conservation law, our results for the characteristics and the shock time disagree with the results of mp. As we will discuss below, the reason for this disagreement is their implicit assumption of a very restrictive relation between the initial profiles $`R_0(x)`$ and $`H_0(x)`$.
## 2 Characteristic coordinates
The general basis of the method we used to solve Eqs. (1,2) consists in a replacement of the original equations by an equivalent system of four partial differential equations for the functions $`t`$, $`x`$, $`R`$ and $`H`$, but now considered as functions of new coordinates $`\mu `$ and $`\nu `$, which will be defined below.<sup>1</sup><sup>1</sup>1The theory used here is actually more general and can be used in the presence of non-linear saturation terms or for ripple models . These new equations will be particularly simple inasmuch as each equation has derivatives with respect to either $`\mu `$ or $`\nu `$, though the mapping between the coordinates $`(t,x)`$ and $`(\mu ,\nu )`$ will be in general complicated. To define the characteristic coordinates $`(\mu ,\nu )`$, we have to specify first the characteristic curves of the system (1,2). For practical reasons, we introduce new functions $`u(t,x)=1R(t,x)/\alpha `$ and $`v(t,x)=(\alpha +xR(t,x)H(t,x))/\alpha `$ instead of $`H(t,x)`$ and $`R(t,x)`$. For this new functions the differential expressions become
$`L_1[u,v]`$ $`=`$ $`u_t\gamma \alpha (1u)u_x+v_t+\gamma \alpha (1u)v_x\gamma (1u)=0,`$ (3)
$`L_2[u,v]`$ $`=`$ $`u_t+[1+\gamma \alpha (1u)]u_x\gamma \alpha (1u)v_x+\gamma (1u)=0.`$ (4)
Both operators $`L_1`$ and $`L_2`$ contain linear combinations of the type $`au_t+bu_x`$ of the derivatives of $`u`$ (and the same holds for $`v`$). This combination means that $`u`$ is differentiated in the direction given by the ratio $`t/x=a/b`$. Since the coefficients $`a`$ and $`b`$ differ for $`u`$ and $`v`$ and also for $`L_1`$ and $`L_2`$, the functions $`u`$ and $`v`$ are differentiated in each of the operators in different directions in the $`(t,x)`$ plane. Notice that the directions depend also on $`u`$ itself, and therefore on the solution under consideration, which is a typical feature of non-linear systems. As noted above, our goal is to find equivalent differential equations of which each contains derivatives in only one (local) direction corresponding to one of the new coordinates $`\mu `$ and $`\nu `$. Therefore we take a linear combination $`L=\lambda _1L_1+\lambda _2L_2`$ of the operators in Eqs. (3,4) such that the derivatives of $`u`$ and $`v`$ in $`L`$ combine to derivatives in the same direction, which is called a characteristic direction. Moreover we assume that these local directions change smoothly as functions of $`t`$ and $`x`$, and are given by the tangential vectors $`(t_\sigma (\sigma ),x_\sigma (\sigma ))`$ of a smooth path $`(t(\sigma ),x(\sigma ))`$ with $`\sigma `$ as parameter. Considering the functions $`u`$ and $`v`$ along this path, they depend only on $`\sigma `$ and we have, e.g., $`u_\sigma =u_tt_\sigma +u_xx_\sigma `$. Using these conditions, we obtain four homogeneous linear equations for the coefficients $`\lambda _1`$ and $`\lambda _2`$ with coefficients depending on $`t`$, $`x`$, $`u`$, $`v`$ and their derivatives with respect to $`\sigma `$. For non-trivial solutions all possible determinants of the matrix of these coefficients have to vanish, leading to three independent equations or characteristic relations (cr). The first one can be written as a quadratic equation for the local direction $`\zeta =x_\sigma /t_\sigma `$ of differentiation, the solution of which are: $`\zeta _+=1`$ and $`\zeta _{}=\gamma \alpha (1u)`$. Now, for a fixed solution $`u`$, the equations $`dx/dt=\zeta _+`$ and $`dx/dt=\zeta _{}`$ are ordinary differential equations, which define two families of paths with the starting position $`x_0`$ at $`t=0`$ as parameter. These families of paths are the characteristics $`C_+`$ and $`C_{}`$ of the system (1,2). From a physical point of view, they are simply the paths along which $`R(x,t)`$ ($`\zeta _+`$) and $`H(t,x)`$ ($`\zeta _{}`$) evolves with time.
The new curved coordinate frame $`(\mu ,\nu )`$ is now defined such that the two one-parametric families of characteristics are mapped by the coordinate transformation on an usual Cartesian coordinate frame in the $`(\mu ,\nu )`$-plane, i.e., along the characteristics the coordinate functions $`\mu (t,x)`$ and $`\nu (t,x)`$, respectively, are constant. Here we have chosen to map the line $`t=0`$ on the line given by $`\mu =\nu `$. In terms of the new coordinates we find
$$x_\nu +t_\nu =0,x_\mu \gamma \alpha (1u)t_\mu =0.$$
(5)
Now we make use of another cr, which evaluated along $`C_+`$ and $`C_{}`$ by identifying $`\sigma `$ with $`\nu `$ and $`\mu `$, respectively, yields the conditions
$$u_\nu +\gamma \alpha (1u)v_\nu +\gamma (1u)t_\nu =0,u_\mu v_\mu +\gamma (1u)t_\mu =0.$$
(6)
These equations together with the Eqs. (5) form the desired set of four equations mentioned before. Every solution of this new system satisfies the original Eqs. (1,2), since the Jacobian $`t_\nu x_\mu t_\mu x_\nu 1+\gamma R(\mu ,\nu )`$ of the coordinate map does not vanish due to $`\gamma R(\mu ,\nu )>0`$.
## 3 General solution
Before we can construct a solution to the equivalent system (5,6), we have to specify initial data along the line $`\mu =\nu `$ corresponding to $`t=0`$. We choose an general profile $`H_0(x)`$, perturbed at $`t=0`$ by a uniform ‘rain’ of rolling grains: $`R_0(x)=\alpha `$. In terms of the new coordinates, the initial conditions become $`t_0(\mu )=0`$, $`x_0(\mu )=\mu `$, $`u_0(\mu )=0`$, $`v_0(\mu )=(\mu +H_0(\mu ))/\alpha `$. By introducing the function $`\mathrm{\Delta }(\mu ,\nu )=1\gamma \alpha (1u(\mu ,\nu ))`$, one can show that the problem of solving the system given by Eqs. (5,6) can be reduced to the task of finding a solution to the equation $`\mathrm{\Delta }_\nu =\gamma H_0^{}(\nu )(1+1/\mathrm{\Delta })`$, with initial condition $`\mathrm{\Delta }(\mu ,\mu )=1\gamma \alpha `$. The solution of this equation can be simply expressed in terms of the so-called Lambert function $`W`$ :
$$\mathrm{\Delta }(\mu ,\nu )=1W\left\{\alpha \gamma \mathrm{exp}[\alpha \gamma +\gamma (H_0(\mu )H_0(\nu ))]\right\}.$$
(7)
With this solution at hand, the solution to the system (5,6) is determined by
$`t(\mu ,\nu )={\displaystyle _\nu ^\mu }{\displaystyle \frac{ds}{\mathrm{\Delta }(s,\nu )}}=\mu \nu +{\displaystyle _\nu ^\mu }{\displaystyle \frac{\mathrm{\Delta }_\mu (s,\nu )ds}{\gamma H_0^{}(s)}}`$ , $`x(\mu ,\nu )=\mu t(\mu ,\nu )`$
$`R(\mu ,\nu )={\displaystyle \frac{1+\mathrm{\Delta }(\mu ,\nu )}{\gamma }}`$ , $`H(\mu ,\nu )=H_0(\nu ),`$ (8)
where we have expressed already the original fields $`R(\mu ,\nu )`$ and $`H(\mu ,\nu )`$ in terms of the functions $`u`$ and $`v`$. To get the fields as functions of $`t`$ and $`x`$, one has to invert the coordinate map. This can be done by using $`\mu (t,x)=tx`$ and integrating the equation for $`t(\mu ,\nu )`$ to obtain also $`\nu (t,x)`$. As announced before, the height profile $`H(t,x)=H_0(\nu (t,x))`$ turns out to be constant along the characteristics $`C_{}`$.
## 4 Generic shape for $`H(t,x)`$
In the following we will consider a situation which is generic for sandpile surfaces. Suppose that one starts with a sandpile profile, which consists of two regions with constant but different slopes matching with a kink at $`x=0`$, and again with a constant amount of rolling grains. The slopes may be either larger or smaller than the angle of repose $`\theta _r`$. If we denote the slope to the right (left) by $`\theta _r+\theta _>`$ ($`\theta _r+\theta _<`$), we have $`H_0(x)=\theta _>x`$ for $`x>0`$ and $`H_0(x)=\theta _<x`$ for $`x<0`$. In the case of a piecewise constant $`H_0^{}(x)`$ one can integrate the equation for $`t(\mu ,\nu )`$ easily as can be seen from Eq. (8). The structure of Eq. (7) suggests to distinguish between three regions given by $`\mu >0`$,$`\nu <0`$ (I), $`\mu ,\nu <0`$ (II) and $`\mu <0,\nu >0`$ (III). <sup>2</sup><sup>2</sup>2The region where $`\mu ,\nu >0`$ turns out to be mapped on the half space with $`t<0`$ and is therefore not of physical interest. In regions I and III one can find the explicit expression $`\nu (t,x)=x+\frac{\alpha }{\theta }(1e^{\gamma \theta t})`$ with $`\theta =\theta _<`$ (I) or $`\theta =\theta _>`$ (III), i.e., the characteristics $`C_{}`$ are in these regions simple exponential curves. As a consequence, no shocks can appear in these two regions and the corresponding solutions are particularly simple:
$$R_{I(III)}(t,x)=\alpha e^{\gamma \theta _{<(>)}t},H_{I(III)}(t,x)=H_0(x)+\alpha R_{I(III)}(t,x).$$
(9)
The boundaries of the regions I and III in real space $`(t,x)`$ are given by the conditions $`x<t`$ and $`x>x_1(t)=\frac{\alpha }{\theta _>}(e^{\gamma \theta _>t}1)`$ corresponding to the $`\mu =0`$ and $`\nu =0`$ characteristics, respectively, see Fig. 1. The boundary for region I has an obvious physical meaning: The information that there is a kink at $`x=0`$ can only propagate to the left with the velocity of the moving grains, which is $`1`$ in our rescaled units. Moreover, it is important to note that the ‘uphill’ velocity with which the kink moves is only equal to $`\gamma \alpha `$ at small times, before growing exponentially. As discussed in the introduction, this growth eventually saturates, as does the value $`R`$, or else the characteristic $`C_{}`$ quickly reaches the edge of the pile.
The range of $`x`$ in between the above two regions corresponds to the intermediate region II. Within this range one can obtain only an implicit solution for the coordinate map $`\nu (t,x)`$. It reads
$$\nu =x+\frac{1}{\gamma }\left[\frac{\mathrm{\Delta }(xt,\nu )\mathrm{\Delta }(0,\nu )}{\theta _>}+\frac{\mathrm{\Delta }(0,\nu )+1+\alpha \gamma }{\theta _<}\right],$$
(10)
where $`\mathrm{\Delta }(\mu ,\nu )=1W\left\{\alpha \gamma \mathrm{exp}[\alpha \gamma \gamma (\theta _>\mu +\theta _<\nu )]\right\}`$ as follows from Eq. (7). The shape of $`R(t,x)`$ and $`H(t,x)`$ can be obtained directly from the last two equations of (8). In general, Eq. (10) has to be solved numerically although several results can be obtained in an analytic way. It turns out that the solutions of Eq. (10) fall into two qualitatively different classes, according to the values of $`\beta =\theta _>/\theta _<`$ and $`\theta _<`$: for $`\beta >1\alpha \gamma `$ or $`\theta _<<0`$, both $`R(t,x)`$ and $`H(t,x)`$ remain continuous for all times, while for $`\beta <1\alpha \gamma `$ and $`\theta _<>0`$, the solutions develop a discontinuity in $`R(t,x)`$ and $`H(t,x)`$ beyond a finite shock time $`t_s`$. This must be contrasted with mp, since in the present case $`R_0(x)=\alpha `$, they predict that shocks are absent for all times.
## 5 Examples
The characteristics resulting from numerical solutions of Eq. (10) have been plotted in Fig. 1. The left part of this figure has been obtained for $`\theta _>>0`$ and $`\theta _<<0`$, corresponding to $`\beta <0`$. In this case, the characteristics are more and more ‘diluted’ as time increases, and therefore never cross – no shock. In the limit of large times, the argument of the Lambert $`W`$ function becomes very large. Using the first two terms of the asymptotic expansion of $`W`$ we get $`\nu (t,x)=[\beta t+\mathrm{ln}(x+\frac{\beta t}{\beta 1})/(\gamma \theta _<)]/(\beta 1)`$. The corresponding expression for $`R(t,x)`$ and $`H(t,x)`$ can be obtained from Eq. (8). A particularly interesting quantity to look at is the local slope at, say, $`x=0`$. In this limit the slope is negative and decays with time as $`H_x(t,x=0)=1/(\gamma \beta t)`$.<sup>3</sup><sup>3</sup>3Note that this $`t^1`$ relaxation of the slope has also been obtained in within a very different model. It means that the ‘true’ slope $`h_x`$ actually relaxes to the angle of repose $`\theta _r`$ for very large time. If $`L`$ is the size of experimental system, then $`C_{}`$ reaches the boundary of the system at a time $`t^{}`$ such that $`L\frac{\alpha }{\theta _>}e^{\gamma \theta _>t^{}}`$. One should therefore measure a final slope $`h_x\theta _r+\theta _</\mathrm{ln}(\theta _>L/\alpha )`$ smaller than the repose angle. This result is consistent with the qualitative discussion of Boutreux and de Gennes for a similar situation .
Another experimentally important quantity is the velocity $`v_R`$ of the “active” region. Following , this region can be defined by the condition $`R(t,x)>R_{\mathrm{min}}`$, where $`R_{\mathrm{min}}`$ is a small threshold. $`v_R`$ is then given by the slope of the curves of constant $`R(t,x)`$, which tends to a constant in the large $`t`$ limit as can be seen in Fig. 1(a). The asymptotic analysis yields $`v_R=\beta /(1\beta )`$. Since $`\beta <0`$, $`1<v_R<0`$, and the avalanche proceeds downhill, but slower than the grains themselves. This is an effect of the non-linear term in the bcre equations since the linearized theory yields $`v_R=1`$ .
The situation where $`\theta _><0`$ and $`\theta _<>0`$ is qualitatively different. In this case, the characteristics cross at some finite time: a shock occurs – see Fig. 1(b). A crossing point of two characteristics means indeed that at this point, two different values of $`R`$ (or $`H`$) are possible and these functions then become discontinuous. Strictly speaking, the Eqs. (1,2) are no longer valid, and the diffusion terms left of from the analysis become important to smooth out this discontinuity. In Fig. 2, we plotted snapshots of the $`h`$ and $`R`$ profiles at different times, for both situations (with and without the occurrence of a shock).
One can calculate the time $`t_s`$ and location $`x_s`$ at which the shock occurs. For that purpose, let us introduce the envelop of the characteristic curves $`x(t,\nu )`$, where $`\nu `$ is a label. The envelop can be represented in a parametric way as $`(t_e(\nu ),x_e(\nu ))`$. It has the property that for each of its points exists a characteristic, which touches it tangentially. It has then to fulfill the conditions $`x(t_e(\nu ),\nu )=x_e(\nu )`$, $`x_\nu (t_e(\nu ),\nu )=0`$. After some calculations, one can find the explicit expression for the envelop,
$`x_e(\nu )`$ $`=`$ $`\nu {\displaystyle \frac{1}{\gamma \theta _<}}\left[1+\alpha \gamma +\mathrm{\Delta }(0,\nu )\left(1+{\displaystyle \frac{1}{1\beta }}\right)\right]`$ (11)
$`t_e(\nu )`$ $`=`$ $`x_e(\nu )+{\displaystyle \frac{\nu }{\beta }}{\displaystyle \frac{1}{\gamma \theta _>}}\left[\alpha \gamma +\mathrm{ln}(\alpha \gamma )+1+{\displaystyle \frac{\mathrm{\Delta }(0,\nu )}{1\beta }}\mathrm{ln}\left(1{\displaystyle \frac{\mathrm{\Delta }(0,\nu )}{1\beta }}\right)\right].`$ (12)
This envelop has two branches, separated by a kink, see Fig. 1(b), given by $`\nu =\nu _s=[\alpha \gamma +\mathrm{ln}(\alpha \gamma )1+\beta \mathrm{ln}\left(1\beta \right)]/(\gamma \theta _<)`$. Whereas the upper branch is parameterized by $`\mathrm{}<\nu <\nu _s`$, the lower one corresponds to $`\nu _s<\nu <\nu _c=[\beta +\alpha \gamma +\mathrm{ln}(\alpha \gamma /\beta )]/(\gamma \theta _<)`$. The resulting shock coordinates are
$$x_s=\frac{1}{\gamma \theta _<}\left[\mathrm{ln}(\alpha \gamma )\mathrm{ln}(1\beta )+1+\frac{1}{1\beta }\right],t_s=\frac{1}{\gamma \theta _<}\left[\left(1\frac{2}{\beta }\right)\mathrm{ln}\left(1\beta \right)\mathrm{ln}(\alpha \gamma )\right].$$
(13)
The condition that ($`t_s,x_s`$) has to be located inside region II leads to the boundary between the classes with and without shock as mentioned above. At the shock position, the amount of moving grains is universal (independent of the initial value $`\alpha `$), and given by $`R_s=1/(\gamma (1\beta ))`$, while $`H_s=\theta _<\nu _s`$. Since typically $`v\gamma d`$ with $`d`$ the grain diameter, we have in our rescaled units $`\gamma 1`$ showing that due to $`R_s\stackrel{<}{}1`$ non linear saturation terms can be neglected at the shock if $`\beta \stackrel{<}{}1`$. The lower branch of the envelop saturates for large $`t`$ exponentially fast with a characteristic time $`1/(\gamma \theta _>)`$ at $`x_{\mathrm{}}=[1+\mathrm{ln}(\alpha \gamma /\beta )]/(\gamma \theta _<)`$, which is always larger than $`x_s`$. This means that the shock stops propagating upwards. A large time expansion in the shock free range $`t<x<x_{\mathrm{}}`$ gives, taking the two leading terms of $`W`$, $`\nu (t,x)=(\alpha /\theta _<)\mathrm{exp}[\gamma \theta _<t(\theta _</\alpha )xe^{\gamma \theta _<t}]`$. Thus the slope is non monotonous within this range: after increasing for small times it relaxes again to the initial value $`\theta _<`$ as $`H_x(t,x)=\theta _<\mathrm{exp}[(\theta _</\alpha )xe^{\gamma \theta _<t}]`$.
## 6 Discussion
Let us summarize the major results of this paper, which could be explored experimentally. Starting from an initial profile made up of two different slopes, we find that shocks can occur after a finite time, depending on the value of the two slopes and the initial density of rolling grains. When shocks are absent, we find that the evolution surface profile is characterized by different velocities: the kink moves upwards with a velocity of the order of $`\alpha \gamma `$ for early times, while the edge of the “active” region moves downwards at a velocity which only depends on the initial slopes, and is smaller than the velocity of the grains. The final slope is shown to be the angle of repose; however, for finite size systems, one expects the final slope to be smaller by an amount which varies as $`1/\mathrm{ln}L`$. When a shock appears, we predict the time and position of this shock, as well as the density of rolling grains there, which takes a universal value. The shock is found to stop progressing upwards.
Our results are in disagreement with those of mp. For the situation considered here, they predict that the initial profile is rigidly shifted along straight characteristics. Therefore, for example, the final slope would be given by $`H_x(t,x=0)=\theta _<`$, which is completely different from our prediction of a decaying slope. The reason for this discrepancy comes from their implicit assumption that $`R_0(x)+H_0(x)+\mathrm{ln}(R_0(x))/\gamma =\mathrm{const}.`$, which does not hold in the cases considered here.
The method presented here can be extended to more general situations. For example, each profile $`H_0(x)`$ can be approximated by a piecewise linear function. Therefore, our analysis can be used to obtain analytical results for more complicated situations as, e.g., bumps or sinusoidal shapes. Another interesting situation is the case where $`R_0(x)`$ is localized in space. Applications of this method to the problem of ripple formation are under way.
Two important physical phenomena have been neglected: diffusion terms, which are expected to be important in the presence of shocks or in the case of a localized initial $`R_0(x)`$ (see ), and non-linear effects, which lead to a saturation of the static/rolling grains conversion term. A simple way to account for the latter effect is to replace the characteristics by straight lines of velocity $`\gamma R_{\mathrm{}}`$ as soon as $`R=R_{\mathrm{}}`$. The influence of a dependence of the velocity of grains on their density would also be worth investigating .
\***
This research was partly supported by the Deutsche Forschungsgemeinschaft (DFG) under grant EM70/1-1.
|
no-problem/9908/astro-ph9908137.html
|
ar5iv
|
text
|
# HIGH VELOCITY LINE EMISSION IN THE NLR OF NGC 4151Based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by AURA Inc under NASA contract NAS5-26555
## 1 Introduction
The Seyfert galaxy NGC 4151 is close enough and well oriented to enable a detailed study of the associated cloud structure around the nuclear engine, with the ultimate goal of understanding the Seyfert mechanism. HST instrumentation has advanced these studies with its high spatial imaging (e.g. Evans et al 1993, Boksenberg et al 1995, Winge et al 1997). The power of HST was greatly extended by the installation of the Space Telescope Imaging Spectrograph (STIS), with both slitless and longslit capability. The structure and kinematics of the narrow line region of NGC 4151 have been studied recently using slitless spectra obtained with STIS (Hutchings et al , 1998, Kaiser et al , 1999). More detailed investigation along subsequently defined long slit locations have been done by Nelson el al (1999) and Kraemer et al (1999). This work revealed a much fuller picture of the kinematics and ionization structure of the ensemble of emission line clouds that form a ”biconical” region on both sides of the nucleus and roughly along the axis of the extended radio structure on either side of the nucleus. The majority of the clouds appear to be outflowing from the nucleus at velocities $`<`$400km. s<sup>-1</sup> , approaching us to the west side and receding on the eastern side. These clouds appear to be part of a general outflow away from the nucleus, but they are not obviously connected with the radio jet outflow. Beyond the inner 5 arcsec, the emission line clouds have lower velocities and appear to be participating in the general galaxy rotation rather than under any influence of the nuclear processes. They do appear to be illuminated and ionized by the beamed nuclear radiation.
In the course of the work on the bright clouds, some weaker and highly shifted emission components were seen in the slitless spectra. Kaiser et al (1999) identified 6 of these with velocities ranging from +846 to -1716 km s<sup>-1</sup>. The location of the material with these high velicites was established by use of two different dispersion directions in the slitless data, as described by Kaiser et al . These locations did not generally correspond with any bright region seen in the undispersed narrow band images, but lay within the ensemble of such clouds, which have much larger flux than the high velocity material, as measured from the dispersed images.
The high velocity material did not fit the simple biconical outflow scenario that describes the bright clouds, and Kaiser et al noted that there seem to be more than one kinematic components superposed in the region. In addition, there is a bimodal distribution of velocity dispersions among the bright clouds that is not related to the radial velocity, although the high velocity clouds all have high velocity dispersion.
In order to study the high velocity material in more detail, and to distinguish it from the bright clouds, we obtained a number of narrow band filter images with STIS and WFPC2. These have bandpasses near to the wavelengths of several strong emission lines, and comparison between them was intended to indicate the location of emission material in various velocity ranges. In all cases, these ranges are more than 400 km s<sup>-1</sup> from that of the nucleus, and thus help isolate the high velocity gas.
We report here on these imaging data, and use them in conjunction with the slitless spectra, and two long-slit STIS spectra through the region, to discuss the high velocity gas in the nuclear region of NGC 4151.
## 2 Image data
Table 1 shows the observational log. The redshift of NGC 4151 is close to 1000 km s<sup>-1</sup>, and the high velocities range over more than 2000 km s<sup>-1</sup> about that, so that the bandpasses of the narrow-band filters do not in general include all the velocities present in the region. Figure 1 illustrates the velocity ranges that are covered for the lines in the passbands, as defined by the FWHM of their transmission functions. In several cases there are two or more lines within the passband, so that different velocities are sampled for the different lines. These are the \[O II\] 3727A doublet and the \[S II\] 6717-31A doublet, and the \[N II\] plus H$`\alpha `$ triplet at $``$6548-84A. We also obtained continuum images and used an archival F502N image from WFPC2.
Figure 2 shows some images and differences of interest. The LRF image of \[O III\] covers all known velocities from the slitless data, and thus is the full representation of the \[O III\] flux. The STIS \[O III\] image covers velocities between -1200 and -860 km s<sup>-1</sup> with respect to the nucleus. It shows almost no resemblance to the LRF image, and some of the bright knots are in areas of low flux in the total intensity image. Thus, the high velocity material in this velocity range is much fainter than the main bright clouds. It also is seen on both sides of the nucleus and outside the main biconical emission region. The difference between the F502N and LRF image shows regions of high positive velocity. These too do not correspond with the main bright knots, and lie on the E side of the nucleus.
The other difference images in Figure 2 are more complex as they compare different lines and combinations of lines. They do however indicate the same regions of high positive and negative velocity, although the contrast is lower because of including continuum or some flux from the low velocity lines.
## 3 Spectroscopic data
The slitless spectra have been described fully in the Hutchings et al and Kaiser et al references. In Figure 3 we show one of the \[O III\] 5007A line, intended to show the high velovity material clearly. The locations of the high velocity material in the total \[O III\] image are ahown in Figure 4. The long slit data are described in detail by Nelson et al (1999). For the purpose of this paper we show the orientation of the two 0.1” wide slits in Figure 5.
From the slitless data we can locate the high velocity regions labelled with letters in Figures 4 and 5. These positions depend on identifying the same feature in both dispersed images. In crowded regions there is occasionally some ambiguity between regions that lie along the same dispersion direction. The image data have been valuable in showing up regions where some high velocity material lies, and provide significant overlap with the slitless spectra.
The two long slit spectra do not sample the whole region, but they do reveal a few high velocity regions that correspond with (parts of) H, D, N(=26), and M, as well as verifying the bright knot velocities reported in the slitless papers (see Figure 5). Table 2 shows the velocities from all our spectroscopic data. These differ from values given for some of them by Kaiser et al (1999), mainly because of more detailed attention to these weak features. The differences are in most cases within the uncertainties, and in most cases the large line width (often blended with other stronger features) make the values reliable only within $``$10% of their given value.
The long slit data do not show any signs of E, F, B, C, I, or J, although their estimated positions lie close to the slits. The positions for H and D also lie near the slit edges, but so do G and A, which have similar velocities. It is thus possible that there is extended high velocity emission lying between G-H and A-D. Since there is no doubt of the reality of the high velocity material at E, F, B, C, I, and J in the slitless images, we conclude that either the regions are compact, or faintness and uncertainty in their locations explains why they are not seen in the long slit data.
The long slit data do indicate material with high positive velocity (+400 km s<sup>-1</sup>) near bright cloud 19 of Hutchings et al (1998), and some material at $``$600 km s<sup>-1</sup> inside region M. The STIS \[O III\] image also indicates regions of high negative velocity marked with small squares in Figure 4. These are not seen in the slitless data as these velocity shifts place the dispersed line emission in the region dominated by the bright emission clouds. Thus, it is clear that we do not have the complete picture of high velocity emissions, and that a full set of long slit spectra, possibly with different orientations, would be required to get that.
However, with the exception of bright knot 26, the high velocity line emission has low flux and comes from locations superposed on the whole inner bright emission cloud region. Region O also lies at the end of a plume of material seen extending out of the main emission cloud region, although very close to the nucleus.
Winge et al (1999) report velocities from several long slit spectra obtained with the FOC. These have lower resolution and S/N and greater distortions than the STIS spectra. However, their results also show up high velocity material that stands apart from the lower velocities found in the main cloud ensemble. Their results detect the material we designate N and G-H in their nuclear spectrum; C and G-H in their 0.23” NW offset; D and the knots seen in the STIS \[O III\] image in their 0.41” and 0.46” offsets to the NW. Thus their results are consistent with those reported here.
## 4 Ionisation structure
With the complete wavelength coverage of the \[O II\], \[O III\], and H$`\beta `$ line velocities in our images, it is possible to construct ratio images that show the variation of O ionisation and \[O III\]/H$`\beta `$ in the emission line region. Figure 6 shows these ratios. The \[O III\]/\[O II\] ratio was not corrected for the different continuum and nuclear scattering in the two images, but these appear to be small as there is no systematic gradient away from the nucleus. Both ratio images were tested empirically for sensitivity to alignment, sampling, and rotation and the morphological features that are seen are robust to any expected uncertainties.
We first note that the ionisation of O is higher along radial locations on both sides of the nucleus. The rays are longer to the W, which is the approaching side for the general outflow. This suggests these are paths where the nuclear radiation is less obscured, and ionise the gas more strongly. These paths are seen more clearly in the approaching cone. Making a detailed comparison with the high velocity gas locations, we find that many of them lie in the regions of higher ionisation. Some (e.g. G and H, B, C, I, J lie around the edges of a high ionisation region. A few (E, F, and A) lie in regions of low or average ionisation. Region O lies in the part of the diagram affected by the telescope diffraction spikes, so we cannot comment on it. Generally there is an association of high velocity with higher ionisation. There are, however regions of high ionisation (mainly to the E) with no known high velocity gas. In view of the higher signal from the bright clouds and the fact that not all high velocity regions appear to lie in high ionisation places, we need to confirm these conclusions with systematic slit spectroscopy.
Turning to the \[O III\]/H$`\beta `$ ratio, there is good signal over a smaller region, but the main radial features are still present. This presumably is a result of the same cause: where O is more highly ionised, the H is weaker because it is ionised. Kaiser et al have discussed the quantitative \[O III\]/H$`\beta `$ values for many bright clouds based on dispersed images. The present images will superpose the high velocity fluxes on the bright clouds, and the latter presumably dominate. In detail, the high velocity gas locations lie in or close to regions of high \[O III\]/H$`\beta `$, even more markedly than the O line ratio. Once again, we have a few regions of high ratio that do not lie near any known high velocity gas, and the same cabest apply to any general conclusions about the high velocity gas.
Thus, the high velocity gas is weak in H$`\beta `$ and \[O II\], and appears to occur in highly ionised material. In the bright clouds Kaiser et al find a decrease in \[O III\]/H$`\beta `$ with increasing nulcear distance. The high velocity gas has higher \[O III\]/H$`\beta `$: it is not known how high but the non-detection in \[O II\] and H$`\beta `$ implies numbers larger than 15. The measured bright cloud line ratios from Kaiser et al do not show up as features in the ratio image, so that there may be gradients within large clouds, or confusion from overlap. Table 2 shows nuclear distances for the high velocity regions, and there is no correlation of line ratio with this: the regions all lie within the high ionisation inner region of Figure 5.
## 5 Radio structure
The radio structure of NGC4151 lies along a line with the brightest feature at the nucleus. There are several bright knots along the structure, on both sides of the nucleus, but with more flux and structure on the E (approaching) side. As noted in Kaiser et al , there is little detailed correspondence between the bright clouds and the radio structure, although the radio axis lies along the overall inner emission line region. In addition, the innermost radio structure shows a bend away from the large-scale axis by some 50<sup>o</sup> (to the N), but on a spatial scale unresolved by HST.
Figure 4 shows the radio structure ridge line and the locations of the main bright features at 6cm. The brightest knot C3 lies close to the brightest emission line cloud #20 from Hutchings et al . Several high velocity gas locations lie close to the radio ridge line (E, F, D, A), while others occur near bright radio knots or bends in the ridge line (G, C, I). About half the high velocity locations do not have any apparent relationship to the radio structure. There is no correlation between the magnitude of the gas velocity and the projected distance from the radio jet. There is also no high velocity material in the direction of the innermost (VLBI) radio jet orientation.
## 6 Discussion
Table 2 summarizes what is known about the high velocity material. The high velocity gas appears to be a different kinematic component of the nuclear environment from the bright biconical clouds. The high velocity gas has higher ionisation, probably due to less obscured illumination by the nucleus. The material has much lower flux (by more than an order of magnitude in \[O III\] flux, and even more in other lines) than the bright clouds, with the single exception of cloud 26 whose flux is typical of the bright clouds. The high velocity material is seen with velocities of approach and recession on both sides of the nucleus, quite distinct from the clear separation of approach and recession across the nucleus, in the bright clouds.
The location of the known high velocity material (which is probably not complete) is along ridges of higher ionisation. The ridges are more marked and longer on the W side, where the bright clouds are approaching (i.e. the near side of the bicone). The central ridge lies along the line of the radio jet axis, and the radio ridge line appears to wind around regions of highest ionisation (Figure 4). This may indicate that the jet is deflected by higher density in these regions, or that it adds to the ionisation by its interaction with gas clouds, as proposed in a more general sense by Winge et al (1999).
12 of the 17 known regions of high velocity material lie in the W (approaching) cone. This implies that there is less overall obscuration on this side, considering that the flux from all such regions is low. There is no flux difference across the nucleus, in the regions that are seen. The high velocity material does appear to be distinct from the bicone outflow inferred from the bright clouds, but its overall velocity and visibility is consistent with that picture.
All of the high velocity material has a high velocity dispersion, by the criterion of 130 km s<sup>-1</sup> set by Kaiser et al . Some of the velocity dispersions (as noted in the table) are much higher, but also not well determined because of blending in the slitless images. However, there are also high velocity dispersion bright clouds with low radial velocities, as shown by Kaiser et al . There are some 10 bright clouds that have velocity dispersions comparable with the low flux high velocity regions. High velocity gas is found only within about 1 arcsec of the nucleus, and all high velocity dispersion material lies within about 2 arcsec of the nucleus.
It is tempting to suppose that the high velocity material is accelerated radially away from the nucleus with a velocity gradient that is also radial. That would predict that there is a correlation between radial velocity and velocity dispersion due to projection effects, and we do not in fact see clear evidence of this. Also, it would not explain why there are bright low velocity, high velocity dispersion clouds. Thus, it appears that this effect is not a strong one, and high velocity dispersion is found in both populations - bright outflow clouds and the high velocity gas.
The high velocity gas is generally lower in density and appears in places that may be directly illuminated by nuclear radiation or shocked by nuclear material. The concidence with radio features may indicate that the jet also moves along one line of direct illumination (travel) from the nucleus, rather than direct acceleration of the high velocity gas by the jet.
The fact that velocities of approach and recession are not connected with position E or W of the nucleus suggests that the whatever drives the high velocity material may push it aside from the general radial motion, so that it may appear with radial velocity counter to the general flow, by projection. It may be that there is high velocity material projected with low radial velocity that is not recognised against the backdrop of lower velocity motions of the bright clouds. We also note that the high velocity material may be consistent with the ballistic ejection model proposed by Smith (1994), as it is faint, has large velocity dispersion, and may be highly ionised.
Finally, we note that the high velocities are similar to those seen strongly in absorption in C IV, against the nucleus. However, those absorbers have very low velocity dispersions (Weymann et al 1997), with one exception. The exception is a broad transient absorber seen at high approach velocity in one GHRS spectrum. The location of the high velocity emitting gas suggests a low covering factor against the nucleus from our line of sight. The relationship of the narrow high velocity C IV absorbers to the optical high velocity emitters is not clear.
References
Boksenberg A., et al , 1995, ApJ, 440, 151
Evans I.N., Tsvetanov Z., Kriss G.A., Ford H.C., Caganoff S., Koratkar A.P., ApJ, 417, 82
Hutchings J.B., et al 1998, ApJ, 492, L115, 1998
Kaiser M.E., et al 1999, ApJ (in press: Astro-ph 9906283)
Kraemer S.B., Crenshaw D.M., Hutchings J.B., Gull T.R., Kaiser M.E., Nelson C.H., Weistrop D., 1999, preprint
Nelson C.H., Weistrop D., Hutchings J.B., Crenshaw D.M., Gull T.R., Kaiser M.E., Kraemer S.B., Lindler D., 1999, preprint
Pedlar A., et al , 1993, MNRAS, 263, 471
Smith M.D., 1994, MNRAS, 209, 913
Weymann R.J., Morris S.L., Gray M.E., Hutchings J.B., 1997, ApJ, 483, 717
Winge C., Axon D.J., Macchetto F.D., Capetti A., 1997, ApJ, 487, L121
Winge C., Axon D.J., Macchetto F.D., Capetti A., Marconi A., 1999, Astro-ph 9901415
Captions to Figures
1. Ranges of radial velocity with respect to NGC 4151 nucleus covered by narrow-band filters, for emission lines that lie within their passbands. The filters are shown and exposures are listed in Table 1. The dots in the bottom row indicate the approximate velocities of the known high velocity emission lines in NGC 4151. The range between -400 and 400 is occupied by almost all the bright emission line clouds.
2. Images and difference images of the central $``$5 arcsec of NGC 4151, as labelled. The orientation is the same for all and North is indicated in Figure 4. The central emission line region shown is approximately 1 x 4 arcsec. The bright clouds are approaching on the lower side and receeding on the upper side of the nucleus. The material visible has velocity ranges as shown in Figure 1. Individual images are discussed in the text.
3. A slitless spectral image of \[O III\] line showing dispersed locations of high velocity material. The lower panel is the unprocessed image with the nuclear continuum removed, and stretch to show the faint high velocity components. The upper lanel is an unsharp mask that shows some of the high velocity features more clearly.
4. Positions of high velocity material identified on the total \[O III\] image. Table 2 gives details of the individual positions as labelled. The key shows how the positions are derived. The jagged line sketches the ridge of the radio structure, with its bright knots identified by X and numbered following Pedlar et al 1993. The arrow indicates North.
5. The positions of our long slits shown with respect to the lettered high velocity material from Figure 3. The slit widths are somewhat oversized and relative positions of the high velocity gas are uncertain by up to 0.1”.
6. Matched \[O III\] and ratio images of central NGC 4151. Note the radially extended regions of higher ionisation, particularly to the W (lower right). The images are discussed in the text.
|
no-problem/9908/physics9908053.html
|
ar5iv
|
text
|
# Untitled Document
Oscar’s Physics Phaire
A Collection of Problems
You will find here a number of (mostly) elementary physics problems dealing mainly with uniform motion kinematics. In preparing this collection I have tried to create original situations that could help bring motivation to an introductory course. You are welcome to suggest improvements and … provide solutions!
Oscar Bolina
Department of Mathematics
University of California, Davis
Davis, CA 95616-8633
E-mail: [email protected]
The Four Spiders
Four spiders move in straight lines away from a common origin on a plane in such a way that at any time they are situated on the corners of a rectangle. If three spiders move at rates of 2, 3 and 4 cm/s, find out the possible values of the speed of the fourth spider.
Dangerous Nanny
On his everyday commute from home to work at 50 km/h, a man always met his nanny, heading to his house, halfway on his journey to work. One day the man left home 5 minutes later and had to drive at 70 km/h in order to get to work on time. On this day, he met his nanny 9 km from his house. What is the speed the nanny used to maintain on her daily ride?
A Sliding Problem
A bead slides with constant speed v along lines perpendicular to the sides of a cone of side l and base b. How long does it take for the bead to reach the vertex V?
Cubic River
A boat that can move at speed v in still water crosses a river of width l that flows with speed w. Prove that the shortest time the boat takes to complete its trip is $`l/v`$. Show that the boat reaches the opposite bank at a distance $`x=wl/v`$ (above or below?) from its starting position. Verify that this value for x is a root of the equation
$$v^2x^6+l^2(2v^2w^2)x^4+l^4(v^22w^2)x^2w^2l^6=0.$$
Solve this (cubic) equation for other roots and interpret your result.
Bugs’ Lives
A bug A that was confined to live on the rim of a circle of radius R realized that a tasty bug B it preyed on used to sneak into its territory always with the same constant velocity v along a same straight line located at a distance r from the enter of the circle. Bug A would like to figure out the minimum velocity it should start moving as soon as bug B showed up at p so that, no matter where it was on the circumference of the circle, it would always arrive at q before bug B, and, hopefully, have a good snack. Could you help?
The Closer the Faster
Two particles describe a rectilinear motion in a plane in such a way that the component of their relative velocity along the straight line joining them has a constant magnitude u. Prove that their relative velocity as a function of the distance d between them is given by
$$V=\frac{u}{\sqrt{1(d_{min}/d)^2}}$$
(1)
where $`d_{min}`$ is the minimum distance between the particles
Square Dance
Four ants, initially located at the corners of a square of side d, start to move at the same time, in the same sense and with constant speeds, along the sides of the square. How long does it take for the ants to be, for the first time, all moving on the same side of the square? Are there conditions on the speeds for this to be possible?
Probabilistic Kinematics
Particles are created at random on a unit line AB and move towards B with speed v. What is the probable elapsed time between creation and detection at B?
Storm Ahead
Forty minutes into its straight line flight of 1,800 miles to Clear Skies City at 360 miles an hour, an airplane is warned against a storm ahead and ordered to take an alternative route and increase its speed $`30\%`$. Determine how far from its routine route can the plane get without delaying itself.
Of Life and Death
A photon, created at P with speed c, reflects twice on the walls of a square box before being absorbed at the origin O. How long does the photon survive?
Squirrels Are Not Pets
A very fearful squirrel is careful enough never to go farther than a distance R from its burrow. One day the squirrel leaves its burrow, gets a cone near a pine tree, gets a nut near an oak tree and returns to the burrow leaving behind a path in the shape of a triangle having maximum area. If the squirrel can run with speed v, how much time does it spend on the round-trip?
Fast and First
Two particles, initially at a distance r apart, move for a time T with speeds v and w along straight lines towards the same point. Determine all the points that the faster particle reaches before the slower one.
Askance and …
Two particles that move in a plane along intersecting straight lines with constant speeds $`v`$ and $`w`$ have the component of these velocities along the line joining the particles always in the ratio 1:k. Prove that angle $`\beta `$ between the directions of motion of the two particles is given by
$$\mathrm{cos}\beta =\frac{v^2kw^2}{(1k)vw}$$
… Oblique
The components of the acceleration of a particle along two direction in the plane of the acceleration vector that make an angle $`\xi `$ with each other are $`a_1=ap`$ and $`a_2=aq`$, where $`a`$ is the magnitude of the acceleration vector. Show that
$$\mathrm{sin}\xi =\frac{1}{2pq}\sqrt{2(p^2+q^2)(p^2q^2)^21}.$$
Are there restrictions on the values of $`p,q`$?
Under Arrest
When the police arrived at a bank responding to a robbery, the robbers had already fled in two cars speeding at 80 km/h and 90 km/h. The police pursue the cars, arresting the slower car at 20 km from the bank and the faster car at 30 km from the bank. Determine the speed of the police car in the pursuit.
Math Illusion
A particle moves with speed v on concentric triangular lines having lengths $`p\%`$ shorter than the previous one. If the longest line has unit length, how long does it take for the particle to reach the center?
Downstreaming
A boat capable of developing speed v in still water crosses a river of width l and reaches the opposite shore at a distance x below its starting position. Determine the speed of the current that minimizes the time of travel.
Here and There
A particle moves in the plane defined by two straight lines that meet at an angle $`\psi `$. If the distances of the particle to both lines are always in the ratio $`1:r`$, show that the components of its velocity along the two lines are in the ratio
$$\frac{r+\mathrm{cos}\xi }{1+r\mathrm{cos}\xi }$$
Work Out
You are willing to burn some calories (without much effort) and decide to run from some point A to B, C and back to A, all on the side of an equilateral triangle of length l. Before starting, find out which initial position x minimizes your route.
A Train Breakdown
The trains that serve two stations 60 km apart leave at regular intervals of 20 min. One day, one of the trains experiences mechanical problems after traveling 40 km and has to complete the rest of the trip at half its usual speed. As a consequence, this train arrives at the other station just 8 min before the next train. Determine the usual speed of the trains.
Animal Procession
A hen, a pig, and a dog leave a barn at equal time intervals and with speeds $`p,q,r`$ and arrive at the farmhouse also at equal time intervals but in inverse order. Show that
$$\frac{1}{p}+\frac{1}{r}=\frac{2}{q}$$
Bouncing Molecule
An air molecule on the corner of a container makes its way out of the container by progressively bouncing off its walls until it escapes through the opening. Determine the initial directions of the molecule’s velocity vector that allow an escape in the case the molecule experiences no loss of speed. How would you change your answer if the molecule started off with speed $`v_0`$ and lost one third of it after each strike on the wall? What is the minimum possible value of $`v_0`$ in this case?
Back and Forth
A particle moves back and forth with speed v on a straight line OA of length p, while a second particle moves back and forth with speed w on a straight line OB of length q. If both particles start from O at the same time, how long does it take for them to cross, simultaneous and for the first time, a circle of radius r and center O in the plane AOB when a. both particles move toward O? b. one particle (pick one) moves toward O, the other moves away from it? c. both particles move away from O?
|
no-problem/9908/hep-ph9908519.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
This note includes the most recent HERA results on exclusive vector meson (VM) production: $`e+pe+VM+p`$. The H1 and ZEUS experiments studied the production of $`\rho `$ , $`\omega `$ , $`\varphi `$ , $`J/\psi `$ and $`\mathrm{{\rm Y}}`$ mesons, in a $`Q^2`$ domain ranging from photoproduction ($`Q^2`$ $``$ 0) to $`Q^2`$ = 60 $`\mathrm{GeV}^2`$ (where $`Q^2`$ is the virtuality of the photon exchanged in the interaction).
The note is divided in two parts. The first part presents the $`Q^2`$ and $`W`$ dependences of the total, transverse and longitudinal cross sections, for the various vector mesons. The second part is concerned only with $`\rho `$ meson production, and presents measurements of the helicity amplitude ratios $`|T_{ij}|/|T_{00}|`$.
## 2 The cross section $`𝝈\mathbf{(}𝜸^{\mathbf{(}\mathbf{}\mathbf{)}}𝒑\mathbf{}𝑽𝑴𝒑\mathbf{)}`$
### 2.1 Scale of the interaction
Recent measurements of $`𝝆`$ meson electroproduction at $`𝑸^\mathrm{𝟐}`$ $`\text{ }\mathbf{}\mathbf{>}`$ 10 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and of $`𝑱\mathbf{/}𝝍`$ meson photo– and electroproduction indicate a strong energy dependence of the $`𝜸^{\mathbf{(}\mathbf{}\mathbf{)}}𝒑\mathbf{}𝑽𝑴𝒑`$ cross sections (“hard” behaviour): $`𝝈\mathbf{}𝑾^𝜹`$ with $`𝜹`$ $`\mathbf{}`$ 0.8, where $`𝑾`$ is the energy in the photon–proton centre of mass. This behaviour indicates that the mass of the $`𝒄`$ quark or a high $`𝑸^\mathrm{𝟐}`$ value provides a “scale” $`𝑲^\mathrm{𝟐}`$ in the interaction. In this section, results are presented for all VM production in a global way as a function of the scale $`𝑲^\mathrm{𝟐}\mathbf{=}\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}\mathbf{/}\mathrm{𝟒}`$, where $`𝑴_𝑽`$ is the VM mass. This approach is inspired e.g. by the discussions in reference .
A compilation of the HERA measurement of the $`𝜸^{\mathbf{(}\mathbf{}\mathbf{)}}𝒑\mathbf{}𝑽𝑴𝒑`$ cross sections is presented in Fig. 1 as a function of $`\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}\mathbf{/}\mathrm{𝟒}`$. The cross sections were scaled by SU(4) factors, according to the quark content of the VM : $`\mathbf{}`$9/1 for the $`𝝎`$, $`\mathbf{}`$9/2 for the $`\mathit{\varphi }`$, $`\mathbf{}`$9/8 for the $`𝑱\mathbf{/}𝝍`$ and $`\mathbf{}`$9/2 for the $`𝚼`$ mesons. In Fig. 1 and in all following plots, the errors on the data points represent the full errors (including the statistical, systematic and normalisation errors, added in quadrature).
The cross sections are measured at $`𝑾`$ = 75 GeV, or are moved to this value according to the parametrisation $`𝝈\mathbf{}𝑾^𝜹`$, using the $`𝜹`$ value measured by the relevant experiment. The ZEUS $`𝝆`$ and $`\mathit{\varphi }`$ cross sections were corrected for the $`\mathbf{|}𝒕\mathbf{|}`$ cut ($`\mathbf{|}𝒕\mathbf{|}`$ $`\mathbf{<}`$ 0.5 or 0.6 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$), following an exponentially falling distribution d$`𝝈\mathbf{/}𝒅𝒕\mathbf{}𝒆^{𝒃𝒕}`$, with a $`𝑸^\mathrm{𝟐}`$ dependent $`𝒃`$ parameter according to measurements (this correction is $`\text{ }\mathbf{}\mathbf{<}`$ 7 %).
Within experimental errors, the total cross sections for VM production, including the SU(4) normalisation factors, appear to lay on a universal curve when plotted as a function of the scale $`𝑲^\mathrm{𝟐}\mathbf{=}\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}\mathbf{/}\mathrm{𝟒}`$, except possibly for the $`𝚼`$ photoproduction <sup>2</sup><sup>2</sup>2The cross sections $`\sigma (\gamma p\mathrm{{\rm Y}}(1\mathrm{S})p)`$ measured by H1 and ZEUS at $`W`$ = 160 and 120 GeV respectively, were moved to the value $`W`$ = 75 GeV using the parametrisation $`\sigma W^\delta `$, with $`\delta `$ = 1.7. This high value of the parameter $`\delta `$ comes from the prediction of . Note that if the value $`\delta `$ = 0.8 is used (a value measured in case of $`J/\psi `$ photoproduction), the cross sections are higher by a factor 1.5 for ZEUS and 2.0 for H1..
A fit performed on the H1 and ZEUS $`𝝆`$ data using the parametrisation $`𝝈\mathbf{=}𝒂\mathbf{(}𝑲^\mathrm{𝟐}\mathbf{+}𝒃^\mathrm{𝟐}\mathbf{)}^𝒄`$, with $`𝒃^\mathrm{𝟐}`$ = 0.11 $`\mathbf{\pm }`$ 0.03 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and $`𝒄`$ = – 2.37 $`\mathbf{\pm }`$ 0.10 ($`𝝌^\mathrm{𝟐}\mathbf{/}𝒏𝒅𝒇`$ = 0.67) is shown as the full curve in Fig. 1. The ratio of the $`𝝎`$, $`\mathit{\varphi }`$ and $`𝑱\mathbf{/}𝝍`$ cross sections to this parametrisation is presented in the lower plot of Fig. 1.
It is interesting to recognise that the universal $`\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}\mathbf{/}\mathrm{𝟒}`$ dependence is for the total cross section measurements.
### 2.2 Ratio of the longitudinal to transverse cross sections
The measured ratio of longitudinal to transverse cross sections $`𝑹`$ = $`𝝈_𝑳\mathbf{/}𝝈_𝑻`$ at $`𝑸^\mathrm{𝟐}`$ = 6 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ is approximately 2.5 for $`𝝆`$ meson production but approximately 0.4 for $`𝑱\mathbf{/}𝝍`$ meson production. However, the ratio $`𝑹`$ presents a similar dependence for the $`𝝆`$, $`\mathit{\varphi }`$ and $`𝑱\mathbf{/}𝝍`$ meson production when plotted as a function of $`𝑸^\mathrm{𝟐}\mathbf{/}𝑴_𝑽^\mathrm{𝟐}`$ (see Fig. 2). All data are well described by a common empirical parametrisation: $`𝑹\mathbf{=}𝒂\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{/}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}^𝒃\mathbf{(}\mathrm{𝐥𝐧}𝑸^\mathrm{𝟐}\mathbf{/}𝑴_𝑽^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟎}\mathbf{)}^𝒄`$ (curve in Fig. 2).
It is observed that $`𝑹`$ rises steeply at small $`𝑸^\mathrm{𝟐}`$, with a weaker dependence at large $`𝑸^\mathrm{𝟐}`$ values. This behaviour is consistent with the fact that the $`𝑸^\mathrm{𝟐}\mathbf{/}𝑴_𝑽^\mathrm{𝟐}`$ dependence expected at leading order is modified by higher order corrections. The data in Fig. 2 suggest that, within the present statistical precision, this modification preserves the ratio $`𝑸^\mathrm{𝟐}\mathbf{/}𝑴_𝑽^\mathrm{𝟐}`$ as the relevant variable for $`𝝈_𝑳\mathbf{/}𝝈_𝑻`$.
### 2.3 The transverse and longitudinal cross sections
Figs. 3 (a) and (b) present the transverse $`𝝈_𝑻`$ and the longitudinal $`𝝈_𝑳`$ cross sections separately, as a function of $`\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}\mathbf{/}\mathrm{𝟒}`$. The parametrisation for $`𝑹\mathbf{=}𝝈_𝑳\mathbf{/}𝝈_𝑻`$, described above, was used to separate the longitudinal and the transverse parts of the cross sections for the different vector mesons:
$$𝝈\mathbf{=}𝝈_𝑻\mathbf{+}𝜺𝝈_𝑳\mathbf{=}𝝈_𝑻\mathbf{(}\mathrm{𝟏}\mathbf{+}𝜺𝑹\mathbf{)}\mathbf{,}$$
where the polarisation parameter $`\mathbf{}𝜺\mathbf{}`$ = 0.996 at HERA.
The transverse cross sections are well described by a simple power law dependence: $`𝝈_𝑻\mathbf{}\mathbf{(}𝑸^\mathrm{𝟐}\mathbf{+}𝑴_𝑽^\mathrm{𝟐}\mathbf{)}^𝒏`$, with $`𝒏`$ = – 2.47 $`\mathbf{\pm }`$ 0.03 for $`𝝆`$, $`𝒏`$ = – 2.4 $`\mathbf{\pm }`$ 0.1 for $`𝝎`$, $`𝒏`$ = – 2.8 $`\mathbf{\pm }`$ 0.1 for $`\mathit{\varphi }`$, and $`𝒏`$ = – 3.1 $`\mathbf{\pm }`$ 0.2 for $`𝑱\mathbf{/}𝝍`$ meson production.
In view of these values of the parameter $`𝒏`$ ($`𝒏\mathbf{}\mathrm{𝟐}`$), the parametrisation proposed in the GVDM approach of Schildknecht, Schuler and Surrow does not provide good fits <sup>3</sup><sup>3</sup>3If the transverse and the longitudinal cross sections are fitted using the Schildknecht, Schuler and Surrow parametrisation, one obtains for the square of the transverse masses 0.29 $`\pm `$ 0.01 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 3.6), 0.39 $`\pm `$ 0.03 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 5.7) and 4.1 $`\pm `$ 0.4 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 1.9) for the $`\rho `$, $`\varphi `$ and $`J/\psi `$ mesons respectively. The results for the square of the longitudinal masses are 0.38 $`\pm `$ 0.02 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 3.1), 0.53 $`\pm `$ 0.05 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 2.2) and 5.4 $`\pm `$ 0.7 $`\mathrm{GeV}^2`$ ($`\chi ^2/ndf`$ = 0.8) for the $`\rho `$, $`\varphi `$ and $`J/\psi `$ mesons respectively..
### 2.4 Energy dependence
The energy dependence of the cross section for VM production at HERA can be parametrised as $`𝝈\mathbf{}𝑾^𝜹`$. In a Regge context, the parameter $`𝜹`$ can be related to the exchanged trajectory $`𝜶\mathbf{(}𝒕\mathbf{)}`$, which is assumed to take a linear form $`𝜶\mathbf{(}𝒕\mathbf{)}\mathbf{=}𝜶\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{+}𝜶^{\mathbf{}}𝒕`$, where $`𝒕`$ is the square of the four–momentum exchanged at the proton vertex. The value $`𝜶^{\mathbf{}}\mathbf{=}\mathbf{0.25}\mathrm{𝐆𝐞𝐕}^\mathbf{}\mathrm{𝟐}`$ is assumed for the $`𝝆`$ and $`\mathit{\varphi }`$ mesons, as measured in hadron$`\mathbf{}`$hadron interactions . For the $`𝑱\mathbf{/}𝝍`$ meson, the value $`𝜶^{\mathbf{}}\mathbf{=}\mathbf{0.05}`$ is used .
The values obtained for the parameter ($`𝜶\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{}\mathrm{𝟏}`$) for the $`𝝆`$, $`\mathit{\varphi }`$ and $`𝑱\mathbf{/}𝝍`$ meson production are presented in Fig. 4 as a function of the scale $`𝑲^\mathrm{𝟐}`$ = ($`𝑸^\mathrm{𝟐}`$ \+ $`𝑴_𝑽^\mathrm{𝟐}`$)/4. The error bars on the data represent the full errors. For the $`𝝆`$ meson production, the sensitivity to the choice of $`𝜶^{\mathbf{}}`$ is shown by the outer bars, which contain the variation due to the assumption $`𝜶^{\mathbf{}}\mathbf{=}\mathrm{𝟎}`$ (i.e. no shrinkage) added in quadrature. The points are compared to the values of the parameter $`𝝀`$ obtained from fits to the $`𝑾`$ dependence of inclusive $`𝑭_\mathrm{𝟐}`$ measurements ($`𝝈_{𝒕𝒐𝒕}\mathbf{}𝑭_\mathrm{𝟐}\mathbf{(}𝒙\mathbf{,}𝑸^\mathrm{𝟐}\mathbf{)}\mathbf{}\mathbf{(}\mathrm{𝟏}\mathbf{/}𝒙\mathbf{)}^𝝀\mathbf{}𝑾^{\mathrm{𝟐}𝝀}`$), plotted as a function of the scale $`𝑲^\mathrm{𝟐}\mathbf{=}𝑸^\mathrm{𝟐}`$ (see also ).
Within experimental errors, a common rise is observed for the ($`𝜶\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{}\mathrm{𝟏}`$) and the $`𝝀`$ parameters when plotted as a function of the relevant scale $`𝑲^\mathrm{𝟐}`$. This is at variance with the values $`\mathbf{1.08}\mathbf{}\mathbf{1.10}`$ obtained from fits to the total and elastic hadron–hadron cross sections .
## 3 Helicity amplitudes
This section presents the H1 and ZEUS elastic $`𝝆`$ meson production results on the study of angular distributions of the $`𝝆`$ meson production and decay , which provides information on the photon and $`𝝆`$ polarisation.
Following the formalism of , the normalised angular decay distribution is a function of 15 combinations of spin density matrix elements, $`𝒓_{𝒊𝒋}^{𝜶𝜷}`$. Each of these $`𝒓_{𝒊𝒋}^{𝜶𝜷}`$ is a sum of bilinear combinations of the helicity amplitudes $`𝑻_{𝝀_𝝆𝝀_𝜸}`$. One can thus invert the system and extract the helicity amplitudes from the measurement of the 15 elements. The motivations for extracting the helicity amplitudes from the 15 matrix elements are the following:
1. fundamental quantities are computed, on which $`𝒔`$-channel helicity conservation <sup>4</sup><sup>4</sup>4 In case of $`s`$-channel helicity conservation (SCHC), the helicity of the vector meson is equal to that of the photon when the spin quantisation axis is taken along the direction of the meson momentum is the $`\gamma ^{}p`$ centre of mass system. In that case, the non-flip helicity amplitudes $`T_{\lambda _\rho \lambda _\gamma }=T_{00}`$ and $`T_{11}`$ have non-zero values, the single flip ($`T_{10}`$, $`T_{01}`$, $`T_{01}`$, $`T_{10}`$) and double flip amplitudes ($`T_{11}`$, $`T_{11}`$) being zero. (SCHC) hypothesis can be checked directly,
2. measurements of the real and imaginary parts of the helicity amplitudes can be obtained.
Only 15 equations are available for 18 unknowns (9 complex helicity amplitudes $`𝑻_{𝝀_𝝆𝝀_𝜸}\mathbf{=}\mathbf{|}𝑻_{𝝀_𝝆𝝀_𝜸}\mathbf{|}𝒆^{𝒊𝝋_{𝝀_𝝆𝝀_𝜸}}`$). However, both H1 and ZEUS results are compatible with natural parity exchange <sup>5</sup><sup>5</sup>5 Natural parity exchange (NPE) is defined by the following relations between the helicity amplitudes: $`T_{\lambda _\rho \lambda _\gamma }=(1)^{\lambda _\rho \lambda _\gamma }T_{\lambda _\rho \lambda _\gamma }`$. (NPE). Under this hypothesis, 5 helicity amplitudes remain independent (10 unknown): $`𝑻_{\mathrm{𝟎𝟎}}\mathbf{,}𝑻_{\mathrm{𝟏𝟏}}\mathbf{,}𝑻_{\mathrm{𝟎𝟏}}`$, $`𝑻_{\mathrm{𝟏𝟎}}`$ and $`𝑻_{\mathrm{𝟏}\mathbf{}\mathrm{𝟏}}`$.
### 3.1 The ratios $`\mathbf{|}𝑻_{𝒊𝒋}\mathbf{|}\mathbf{/}\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$
Minimum $`𝝌^\mathrm{𝟐}`$ fits were performed to measurements of the combinations of the matrix element to extract the 5 independent complex helicity amplitudes. The normalisation $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ = 1 and $`𝝋_{\mathrm{𝟎𝟎}}`$ = 0 is used. In the full fit, the parameters $`\mathbf{|}𝑻_{\mathrm{𝟏}\mathbf{}\mathrm{𝟏}}\mathbf{|}`$, $`𝝋_{\mathrm{𝟏𝟎}}`$, $`𝝋_{\mathrm{𝟎𝟏}}`$ and $`𝝋_{\mathrm{𝟏}\mathbf{}\mathrm{𝟏}}`$ were found compatible with zero within large errors. These parameters are then put to zero, and the remaining free parameters of the fit are $`\mathbf{|}𝑻_{\mathrm{𝟏𝟏}}\mathbf{|}`$, $`\mathbf{|}𝑻_{\mathrm{𝟎𝟏}}\mathbf{|}`$, $`\mathbf{|}𝑻_{\mathrm{𝟏𝟎}}\mathbf{|}`$ and $`𝝋_{\mathrm{𝟏𝟏}}`$.
The measurement of the ratios $`\mathbf{|}𝑻_{\mathrm{𝟏𝟏}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$, $`\mathbf{|}𝑻_{\mathrm{𝟎𝟏}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ and $`\mathbf{|}𝑻_{\mathrm{𝟏𝟎}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ as a function of $`𝑸^\mathrm{𝟐}`$ is presented in Figs. 5 (a), (b) and (c) respectively (see also table 1). The H1 data cover the kinematic range 2.5 $`\mathbf{<}`$ $`𝑸^\mathrm{𝟐}`$ $`\mathbf{<}`$ 60 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$, 30 $`\mathbf{<}`$ W $`\mathbf{<}`$ 140 GeV ($`\mathbf{}𝑾\mathbf{}`$ = 75 GeV) and $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}`$ = 0.138 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$. For the ZEUS data, the kinematic domain is 0.25 $`\mathbf{<}`$ $`𝑸^\mathrm{𝟐}`$ $`\mathbf{<}`$ 0.85 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$, 20 $`\mathbf{<}`$ W $`\mathbf{<}`$ 90 GeV ($`\mathbf{}𝑾\mathbf{}`$ = 45 GeV) and $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}`$ = 0.14 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ for the low $`𝑸^\mathrm{𝟐}`$ data and 3 $`\mathbf{<}`$ $`𝑸^\mathrm{𝟐}`$ $`\mathbf{<}`$ 30 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$, 40 $`\mathbf{<}`$ W $`\mathbf{<}`$ 120 GeV ($`\mathbf{}𝑾\mathbf{}`$ = 73 GeV) and $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}`$ = 0.17 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ for the high $`𝑸^\mathrm{𝟐}`$ data.
The ratio $`\mathbf{|}𝑻_{\mathrm{𝟏𝟏}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ decreases when $`𝑸^\mathrm{𝟐}`$ increases, indicating that at high $`𝑸^\mathrm{𝟐}`$ longitudinal polarisation dominates. The ratio $`\mathbf{|}𝑻_{\mathrm{𝟎𝟏}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ is observed to be around 8 % (14.0 $`𝝈`$ from zero, 5 data points), and the ratio $`\mathbf{|}𝑻_{\mathrm{𝟏𝟎}}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ around 3 – 4 % (4.5 $`𝝈`$ from zero), which indicate a small but significant violation of SCHC for $`𝝆`$ meson production at HERA.
The predictions of three models based on perturbative QCD are compared to the HERA measurements in Fig. 5. The dashed lines represent the predictions of the Royen and Cudell model , computed for $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}\mathbf{=}\mathbf{0.135}`$ $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$. The dotted lines are the predictions of the Nikolaev and Akushevich model , for $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}\mathbf{=}\mathbf{0.13}`$ $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and $`\mathbf{}𝑾\mathbf{}`$ = 75 GeV. The dash-dotted and full lines present the predictions of the Ivanov and Kirschner model , computed for $`\mathbf{}\mathbf{|}𝒕\mathbf{|}\mathbf{}\mathbf{=}\mathbf{0.13}`$ $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and $`\mathbf{}𝑾\mathbf{}`$ = 75 GeV, using respectively the GRV94HO and the CTEQ4lQ parametrisations for the gluon density in the proton. These three models predict a violation of SCHC, the amplitudes $`𝑻_{\mathrm{𝟎𝟏}}`$ and $`𝑻_{\mathrm{𝟏𝟎}}`$ being different from zero, in agreement with the data.
The $`𝑾`$ dependence of the ratios $`\mathbf{|}𝑻_{𝒊𝒋}\mathbf{|}`$ / $`\mathbf{|}𝑻_{\mathrm{𝟎𝟎}}\mathbf{|}`$ is presented in Figs. 6 (a), (b) and (c) for $`\mathbf{}𝑸^\mathrm{𝟐}\mathbf{}`$ = 4.8 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$, together with predictions of the Nikolaev and Akushevich model (dotted lines) and the Ivanov and Kirschner model (dash-dotted and full lines) (the Royen and Cudell model gives no prediction for the $`𝑾`$ dependence).
### 3.2 The phase $`𝝋_{\mathrm{𝟏𝟏}}\mathbf{}𝝋_{\mathrm{𝟎𝟎}}`$
Figs. 7 (a) and (b) present results for the phase difference $`𝝋_{\mathrm{𝟏𝟏}}\mathbf{}𝝋_{\mathrm{𝟎𝟎}}`$, as a function of $`𝑸^\mathrm{𝟐}`$ and $`𝑾`$ respectively (see also table 1). The phase difference is clearly different from zero, of the order of $`\mathit{\varphi }`$ = 25 degrees. This value is in agreement with the H1 measurement of $`\mathrm{𝐜𝐨𝐬}\mathit{\varphi }\mathbf{=}\mathbf{0.925}\mathbf{\pm }\mathbf{0.022}`$ $`{}_{\mathbf{}\mathbf{0.022}}{}^{}{}_{}{}^{\mathbf{+}\mathbf{0.011}}`$ obtained from the angular distribution when supposing SCHC and NPE. The dotted lines on Figs. 7 (a) and (b) represent the predictions of the Nikolaev and Akushevich model <sup>6</sup><sup>6</sup>6 The models of Royen and Cudell and of Ivanov and Kirschner suppose that the amplitudes are purely imaginary..
### 3.3 Fixed target experiments
The same procedure was used to extract the helicity amplitudes for the fixed target experiments for which the full set of matrix elements have been measured. The data in references and correspond to $`𝑸^\mathrm{𝟐}`$ values between 0.4 and 1.1 $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and $`𝑾`$ values between 2.0 and 3.1 GeV, the CHIO experiment is at higher energy ($`\mathrm{𝟎}\mathbf{<}𝑸^\mathrm{𝟐}\mathbf{<}\mathrm{𝟑}`$ $`\mathrm{𝐆𝐞𝐕}^\mathrm{𝟐}`$ and $`\mathbf{12.5}\mathbf{<}𝑾\mathbf{<}\mathrm{𝟏𝟔}`$ GeV). The fitted helicity amplitudes for are affected by large errors. Fits to the data in reference have a very bad $`𝝌^\mathrm{𝟐}`$. Fits to the data in reference and to the CHIO data give helicity amplitude ratios in agreement, within the experimental errors, with the Royen and Cudell predictions. The experimental errors on the helicity ratios for these two fixed target experiments are typically a factor five greater than the errors on the HERA measurements.
## Acknowledgements
I am grateful to I. Akushevich, D. Ivanov, P. Marage, N. Nikolaev, I. Royen for numerous interesting discussions.
|
no-problem/9908/hep-th9908069.html
|
ar5iv
|
text
|
# 1 2+1-dimensional periodic potential V(A₀), eq. ().
NORDITA-1999/47 HE
Vortex Solution in 2+1 Dimensional Pure Yang–Mills
Theory at High Temperatures
Dmitri Diakonov
NORDITA, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark
and
Petersburg Nuclear Physics Institute, Gatchina, St.Petersburg 188350, Russia
E-mail: [email protected]
## Abstract
At high temperatures the $`A_0`$ component of the Yang–Mills field plays the role of the Higgs field, and the 1-loop potential $`V(A_0)`$ plays the role of the Higgs potential. We find a new stable vortex solution of the Abrikosov–Nielsen–Olesen type, and discuss its properties and possible implications.
Recently there has been renewed interest in the quantized $`Z(N_c)`$ vortices as possible candidates for the confinement mechanism . If vortices are physical objects (and not lattice or gauge artifacts) playing a role in the dynamics of the vacuum fluctuations they have to be found as stable solutions of the effective action obtained from integrating out high frequencies of the field. In the zero-temperature case the one-loop effective action for vortices has been introduced in our previous work . The zero-derivative term of the effective action, i.e. the ‘potential energy’ of the vortex has been found in that paper: it does not lead to any stable solution. This is probably not surprising since all derivatives of the effective action should be summed up before one gets to a definite conclusion on the existence of vortex solutions.
In this letter we address a related but simpler question on whether there are stable vortex solutions in the case of nonzero temperatures. Moreover, we restrict ourselves to a simple and more ‘pure’ case of $`2+1`$ dimensions. If the temperature $`T`$ is high enough vortices can be viewed as cylinders pointing in the ‘time’ direction, with a nontrivial profile in the transverse ‘spatial’ plane. We shall show that the Abrikosov–Nielsen–Olesen-type vortex solution indeed exists in this case, and we shall present its profile and energy per unit length, as functions of temperature.
The $`2+1`$-dimensional YM theory has a coupling constant $`g_3^2`$ of the dimension of mass; high temperatures means $`Tg_3^2`$. One can always choose a gauge with the ‘time’ component of the YM field $`A_0^a(x)`$ independent of ‘time’. In this gauge, one can integrate out the high-momenta components of the fields to obtain the low-momenta effective action for time-independent fields $`A_\mu ^a(x),\mu =0,1,2`$. The separation scale between high and low momenta is given, naturally, by the temperature: high momenta means $`p>2\pi T`$. The effective action is obtained by integrating over nonzero Matsubara frequencies, $`\omega _n=2\pi nT,n>0`$, and over zero ($`n=0`$) Matsubara frequency but large spatial momenta. It can be expanded in powers of the spatial derivatives of the $`A_\mu ^a(x)`$ fields, divided by appropriate powers of $`T`$. The zero-derivative term is the ‘potential energy’ $`V(A_0)`$. Since the spatial size of the vortex solution is expected to be larger than $`1/T`$ one can neglect all terms of the derivative expansion except the first, zero-derivative term $`V(A_0)`$. The latter can be written as
$`V(A_0)`$ $`=`$ $`{\displaystyle \frac{2T^3}{\pi }}\mathrm{sin}^2\pi \nu {\displaystyle _0^{\mathrm{}}}𝑑p{\displaystyle \frac{p^2\mathrm{cosh}p}{\mathrm{sinh}p(\mathrm{sinh}^2p+\mathrm{sin}^2\pi \nu )}}=\pi T^3\nu ^2\left(\mathrm{ln}{\displaystyle \frac{1}{\nu ^2}}+2.67575\right)+O(\nu ^4)`$ (1)
$`=`$ $`{\displaystyle \frac{T}{4\pi }}(A_0^a)^2\mathrm{ln}{\displaystyle \frac{\mathrm{const}T^2}{(A_0^a)^2}}+O(A_0^4),\nu ={\displaystyle \frac{\sqrt{A_0^aA_0^a}}{2\pi T}}.`$
It is a periodic function of the dimensionless variable $`\nu `$ with unit period, depicted in Fig.1. Notice that the potential energy (1) is nonanalytic at the minima $`\nu =integer`$. The fact that the second derivative of $`V(A_0)`$ has a logarithmic singularity at $`\nu =integer`$ is related to the infrared divergency of the Debye mass in $`2+1`$ dimensions , and will have important consequences for the vortex solution.
Adding $`V(A_0)`$ to the classical YM action, $`F_{\mu \nu }^aF_{\mu \nu }^a/4g_3^2`$, we get the energy functional in the spatial transverse plane:
$$E_{}=d^2x\left\{\frac{1}{2g_3^2}\left[ϵ_{ij}\left(_iA_j^a+\frac{1}{2}f^{abc}A_i^bA_j^c\right)\right]^2+\frac{1}{2g_3^2}\left(D_i^{ab}A_0^b\right)^2+V(A_0^a)\right\}$$
(2)
where $`i,j=1,2`$ denote spatial components, $`D_i`$ is the covariant derivative, $`D_i^{ab}=_i\delta ^{ab}+f^{acb}A_i^c`$.
Eq. (2) presents a 2-dimensional YM $`+`$ Higgs system where $`A_0^a`$ plays the role of the adjoint Higgs field, with a peculiar form of the potential energy (1).
We choose the following vortex-type Ansatz for the fields, restricting ourselves to the $`SU(2)`$ colour group for simplicity:
$`A_i^a`$ $`=`$ $`\delta ^{a3}ϵ_{ij}n_j{\displaystyle \frac{\mu (\rho )}{\rho }},\rho =\sqrt{x_1^2+x_2^2},n_j={\displaystyle \frac{x_j}{\rho }},`$ (3)
$`A_0^a`$ $`=`$ $`\delta ^{ai}ϵ_{ij}n_j\nu (\rho )\mathrm{\hspace{0.17em}2}\pi T,`$ (4)
where $`\mu (\rho )`$ and $`\nu (\rho )`$ are trial profile functions of the distance $`\rho `$ from the vortex center. The Ansatz (3) corresponds to the radial component $`A_\rho ^a=0`$ and the azimuthal component $`A_\varphi ^a(\rho )=\delta ^{a3}\mu (\rho )/\rho `$. The closed Wilson loop in the representation labelled by spin $`J`$, circling around the center of the vortex in the transverse plane at distance $`\rho `$, is
$$W_J(\rho )=\frac{1}{2J+1}\text{Tr}\mathrm{P}\mathrm{exp}i𝑑\varphi \rho A_\varphi ^aT^a=\frac{1}{2J+1}\frac{\mathrm{sin}[(2J+1)\pi \mu (\rho )]}{\mathrm{sin}[\pi \mu (\rho )]},\mu (\rho )=\sqrt{A_\varphi ^aA_\varphi ^a}\rho ,$$
(5)
where $`\mu (\rho )`$ is the magnetic field flux along the vortex inside a tube of radius $`\rho `$. If $`\mu (\rho )1`$ at large $`\rho `$ the Wilson loop $`W_J(\rho )(1)^{2J}`$. In particular, in the fundamental representation one has $`W_{1/2}1`$. This is the definition of the quantized $`Z(2)`$ vortex. Our solution will be precisely of this type.
Let us introduce the dimensionless radius
$$x=\rho g_3^{1/2}T^{3/4}$$
(6)
and the dimensionless parameter
$$\alpha =\frac{g_3}{\sqrt{T}(2\pi )^2}.$$
(7)
At high temperatures $`\alpha 1`$. In terms of these quantities the action functional for the vortex of length $`1/T`$ becomes
$$S_{\mathrm{vortex}}=\frac{E_{}}{T}=\frac{1}{2\pi \alpha ^2}_0^{\mathrm{}}𝑑xx\left[\frac{1}{2}\nu ^{\mathrm{\hspace{0.17em}2}}+\frac{1}{2x^2}\nu ^2(1\mu )^2+\alpha \frac{1}{2x^2}\mu ^{\mathrm{\hspace{0.17em}2}}+\alpha v(\nu )\right],$$
(8)
where we have introduced the dimensionless ‘Higgs potential’
$$v(\nu )=\frac{2}{\pi }\mathrm{sin}^2\pi \nu _0^{\mathrm{}}𝑑p\frac{p^2\mathrm{cosh}p}{\mathrm{sinh}p(\mathrm{sinh}^2p+\mathrm{sin}^2\pi \nu )}=\{\begin{array}{cc}\pi \nu ^2\mathrm{ln}\frac{1}{\nu ^2}+\mathrm{}& \mathrm{at}\nu 0,\\ \pi (1\nu )^2\mathrm{ln}\frac{1}{(1\nu )^2}+\mathrm{}& \mathrm{at}\nu 1.\end{array}$$
(9)
The Euler–Lagrange equations of motion for the profile functions $`\mu ,\nu (\rho )`$ are:
$`\alpha {\displaystyle \frac{d}{dx}}\left({\displaystyle \frac{1}{x}}{\displaystyle \frac{d\mu }{dx}}\right)`$ $`=`$ $`{\displaystyle \frac{1}{x}}\nu ^2(1\mu ),`$ (10)
$`{\displaystyle \frac{d}{dx}}\left(x{\displaystyle \frac{d\nu }{dx}}\right)`$ $`=`$ $`{\displaystyle \frac{1}{x}}\nu (1\mu )^2+\alpha x{\displaystyle \frac{dv}{d\nu }},`$ (11)
$$\frac{dv}{d\nu }=2\mathrm{sin}2\pi \nu _0^{\mathrm{}}\frac{dpp}{\mathrm{sinh}^2p+\mathrm{sin}^2\pi \nu }=\{\begin{array}{cc}2\pi \nu \left(\mathrm{ln}\frac{1}{\nu ^2}+1.67575\right)& \mathrm{at}\nu 0,\\ 2\pi (1\nu )\left(\mathrm{ln}\frac{1}{(1\nu )^2}+1.67575\right)& \mathrm{at}\nu 1.\end{array}$$
(12)
We look for the solution with the following boundary conditions:
$$\mu (0)=\nu (0)=0,\mu (\mathrm{})=\nu (\mathrm{})=1,$$
(13)
corresponding to the quantized $`Z(2)`$ vortex, with the ‘Higgs field’ $`A_0`$ going from the trivial minimum at $`\rho =0`$ to a non-trivial one at $`\rho \mathrm{}`$.
The behaviour of the profile functions near the origin and at infinity can be found analytically. At small $`x`$ we get:
$`\mu (x)`$ $`=`$ $`c_1x^2+c_2x^4+\mathrm{},`$
$`\nu (x)`$ $`=`$ $`d_1x+d_2x^3\mathrm{ln}x+d_3x^3+\mathrm{}.`$ (14)
The coefficients $`c_1,d_1`$ are arbitrary but the higher coefficients are determined from eqs. (10, 11):
$$c_2=\frac{d_1^2}{8\alpha },d_2=\frac{\alpha \pi d_1}{2},d_3=\frac{d_1}{8}\left[2c_1+\alpha \pi (3+2h4\mathrm{ln}d_1)\right],h=1.67575\mathrm{}$$
(15)
At large values of $`x`$ the analytical solution is:
$`\mu (x)`$ $``$ $`1e_1\sqrt{x}\mathrm{exp}\left({\displaystyle \frac{x}{\sqrt{\alpha }}}\right),`$ (16)
$`\nu (x)`$ $``$ $`1f_1\mathrm{exp}\left(\pi \alpha x^2\right),`$ (17)
where the constants $`e_1,f_1`$ are not determined by the equations. Notice that the ‘Higgs field’ $`\nu (x)`$ approaches its asymptotic value at infinity not as an exponent but as a gaussian. This is a consequence of the logarithmic divergence of the Debye mass in $`2+1`$ dimensions. Being rewritten in original notations eq. (17) reads
$$\left|A_0^a(\rho )\right|2\pi T\left[1f_1\mathrm{exp}\left(\frac{g_3^2T}{4\pi }\rho ^2\right)\right],\mathrm{at}\rho \mathrm{}.$$
(18)
We notice that the characteristic scale for the variation of $`A_0(\rho )`$, namely $`1/g_3\sqrt{T}`$, is, at high temperatures, much larger than $`1/T`$. It justifies neglecting derivative terms of $`A_0`$ in the effective action and leaving only the zero-derivative term $`V(A_0)`$, as in eq. (2).
The coefficients $`c_1,d_1,e_1,f_1`$ are determined by solving eqs. (10, 11) starting from the expansion (14) towards larger values of $`x`$ and starting from the asymptotic form (16,17) towards smaller values of $`x`$, and matching the functions and their derivatives at some intermediate point $`x1`$; this is done numerically. The resulting profile functions $`\mu (x)`$ and $`\nu (x)`$ are shown in Fig.2, for two values of $`\alpha =0.15`$ and 0.5.
Integrating the profile functions over the whole range of $`x`$ we get the action of the vortex $`S_{\mathrm{vortex}}=E_{}/T`$ plotted in Fig.3, as function of temperature $`T`$. At $`Tg_3^2`$ the action is large, so that the vortices are exponentially suppressed. This is a theoretically ‘clean’ case: as explained above, the use of the energy functional (2) for finding the vortex solution is justified. The vortices are in fact short and broad cylinders oriented in the ‘time’ direction. Indeed, the length of the cylinders is $`1/T`$ while the characteristic radius where the ‘Higgs field’ $`A_0`$ reaches its asymptotic value $`2\pi T`$ is of the order of $`1/g_3\sqrt{T}1/T`$, see eq. (17). It should be mentioned, however, that the magnetic flux $`\mu (\rho )`$ reaches its asymptotic value of 1 at a smaller radius $`\rho 1/T`$, see eq. (16).
To build the $`2+1`$-dimensional vacuum at high temperatures out of the vortices, one needs first of all to gauge rotate the ‘Higgs field’ $`A_0^a`$ (4) to, say, the third direction in colour space: one cannot add up vortices with different colour orientations of $`A_0`$ at infinity. This is achieved with a help of a $`\varphi `$-dependent gauge transformation
$$A_\mu ^{}=U^{}(\varphi )A_\mu U(\varphi )+i\delta _{\mu \varphi }\frac{1}{\rho }U^{}(\varphi )\frac{d}{d\varphi }U(\varphi ),U(\varphi )=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}e^{i\varphi /2}& ie^{i\varphi /2}\\ e^{i\varphi /2}& ie^{i\varphi /2}\end{array}\right).$$
(19)
Under this gauge transformation the $`A_0`$ component of the YM field becomes proportional to $`\tau _3`$, $`A_0^a=\delta ^{a3}\mathrm{\hspace{0.17em}2}\pi T\nu (\rho )`$. The gauge transformation (19) is, however, discontinuous at $`\varphi =2\pi `$, therefore the azimuthal component $`A_\varphi ^{}`$ will now have a ‘Dirac-surface’ singularity at $`\varphi =0`$:
$$A_\varphi ^a=\delta ^{a2}\frac{1}{\rho }\left[1\mu (\rho )2\pi \delta (\varphi )\right].$$
(20)
The Wilson loop (5) is, naturally, preserved by this gauge transformation: it remains $`(1)^{2J}`$ at $`\rho \mathrm{}`$ (for the representation labelled by spin $`J`$).
Having oriented the ‘Higgs expectation value’ $`A_0(\mathrm{})`$ in one colour direction it is now possible to add up many vortices, each of them necessarily carrying a singular Dirac surface. Apart from the factor $`\mathrm{exp}(S_{\mathrm{vortex}})`$ the statistical weight of a vortex is determined by the fluctuation determinant, in particular by the zero modes of the solution. We expect three zero modes here, two of which are associated with the spatial position of the vortex center $`z_i`$, and one pure gauge cyclic mode associated with shifts in the ‘time’ direction. (The latter can be revealed if one uses a gauge with $`A_0`$ explicitly dependent on time.)
Denoting the (uncalculated) prefactor arising from the fluctuation determinant and from zero modes by $`\kappa (T,g_3)`$ we can write the vortex partition function as that of a gas,
$$𝒵_{\mathrm{vortex}}=\underset{N}{}\frac{1}{N!}\left(d^2z\kappa \mathrm{exp}(S_{\mathrm{vortex}})\right)^N,$$
(21)
implying that at large $`T`$ vortices are dilute and thus neglecting their interactions. It follows immediatelly from eq. (21) that the spatial density of vortices (i.e. number per unit area) is
$$n=\kappa \mathrm{exp}(S_{\mathrm{vortex}}).$$
(22)
At large $`T`$ the numbers of noninteracting vortices inside any large area are Poisson-distributed. Therefore, the average of large Wilson loops can be calculated as
$$W_J=\underset{N}{}\left[(1)^{2J}\right]^N\frac{(n\mathrm{Area})^N}{N!}e^{n\mathrm{Area}}=\{\begin{array}{cc}e^{2n\mathrm{Area}}& \mathrm{at}J=\mathrm{half}\mathrm{integer},\\ 1& \mathrm{at}J=\mathrm{integer},\end{array}$$
(23)
so that the string tension for half-integer representations is twice the vortex density , and zero for integer represenations.
It should be recalled, however, that at $`T\mathrm{}`$ the $`2+1`$ dimensional system reduces to two dimensions with a coupling constant $`g_2^2=g_3^2T`$, and in two dimensions there is a trivial perturbative confinement with a nonzero string tension in any representation,
$$\sigma _2=g_2^2J(J+1).$$
(24)
The above vortex-induced string tension should be thus regarded as an exponentially small addition to eq. (24): it leads to a deviation from the ‘Casimir scaling’ of eq. (24). It would be instructive to study the high-temperature $`2+1`$ dimensional YM theory in lattice simulations, if only because one can learn to identify physical vortices by comparing objects found from maximal center gauge fixing with the continuum profiles presented here. Such experience might be useful in higher dimensions.
As one lowers T the vortex action decreases; at $`Tg_3^2`$ the action becomes of the order of unity, therefore, the vortices are not suppressed anymore. Unfortunately, the quantitative theory fails at this point: first, because the effective action (2) is not accurate anymore, second, because the vortices become long and start to bend, third, because one can hardly neglect interactions between dense vortices. Much more serious efforts are needed to describe vortices in this case (if they exist). Nevertheless, to our mind it is useful to know that at least in the limit of high temperatures vortices do exist in a pure Yang–Mills theory, and their basic properties are established.
Useful discussions with Jeff Greensite, Victor Petrov and Gerard ’t Hooft are gratefully acknowledged.
|
no-problem/9908/astro-ph9908290.html
|
ar5iv
|
text
|
# BLACK HOLE – NEUTRON STAR MERGERS AS CENTRAL ENGINES OF GAMMA-RAY BURSTS
## 1. INTRODUCTION
BH/NS and NS/NS mergers are discussed as promising candidates for the origin of gamma-ray bursts (GRBs) (e.g., Blinnikov et al. 1984; Eichler et al. 1989; Paczyński 1991; Narayan, Piran, & Shemi 1992; Mészáros 1999; Fryer, Woosley, & Hartmann 1999; Bethe & Brown 1998, 1999), at least for the subclass of less complex and less energetic short and hard bursts (Mao, Narayan, & Piran 1994) with durations of fractions of a second (Popham, Woosley, & Fryer 1999; Ruffert & Janka 1999). Optical counterparts and afterglows of this subclass have not yet been observed. Due to the presence of a region of very low baryon density above the poles of the black hole, BH/NS mergers are considered as more favorable sources than NS/NS mergers (e.g., Portegies Zwart 1998; Brown et al. 1999).
Previous Newtonian SPH simulations of BH/NS mergers using a polytropic equation of state indicate that the neutron star may slowly lose gas in many mass transfer cycles (Kluźniak & Lee 1998; Lee & Kluźniak 1998,1999). Whether dynamical instability sets in at a minimum separation (Rasio & Shapiro 1994; Lai, Rasio, & Shapiro 1994) or whether stable Roche lobe overflow takes place, however, can depend on the neutron star to black hole mass ratio (Bildsten & Cutler 1992) and the properties of the nuclear equation of state, expressed by the adiabatic index (Uryū & Eriguchi 1999).
In this Letter we report about the first Newtonian BH/NS merger simulations (Eberl 1998) which were done with a realistic nuclear equation of state (Lattimer & Swesty 1991) and which therefore yield information about the thermodynamic evolution and the neutrino emission. They allow one to compare the strength of the gravitational wave (GW) emission relative to NS/NS mergers and to investigate neutrino-antineutrino ($`\nu \overline{\nu }`$) annihilation as potential source of energy for GRBs.
## 2. NUMERICAL METHODS
The three-dimensional hydrodynamic simulations were performed with a Eulerian PPM code using four levels of nested cartesian grids which ensure a good resolution near the center of mass and a large computational volume simultaneously. Each grid had 64<sup>3</sup> zones, the size of the smallest zone was 0.64 or 0.78 km in case of NS/NS and 1.25 or 1.5 km for BH/NS mergers. The zone sizes of the next coarser grid levels were doubled to cover a volume of 328 or 400 km side length for NS/NS and 640 or 768 km for BH/NS simulations. GW emission and its backreaction on the hydrodynamics were taken into account by the method of Blanchet, Damour, & Schäfer (1990) (see Ruffert, Janka, & Schäfer 1996). The neutrino emission and corresponding energy and lepton number changes of the matter were calculated with an elaborate neutrino leakage scheme (Ruffert, Janka, & Schäfer 1996), and $`\nu \overline{\nu }`$ annihilation around the merger was evaluated in a post-processing step (Ruffert et al. 1997).
## 3. SIMULATIONS
Table 1 gives a list of computed NS/NS and BH/NS merger models. Besides the baryonic mass of the neutron star and the mass of the black hole, the spins of the neutron stars were varied. “Solid” means synchronously rotating stars, “none” irrotational cases and “anti” counter-rotation, i.e., spin vectors opposite to the direction of the orbital angular momentum. The cool neutron stars have a radius of about 15 km (Ruffert, Janka, & Schäfer 1996) and the runs were started with a center-to-center distance of 42–46 km for NS/NS and with 47 km in case of BH/NS for $`M_{\mathrm{BH}}=2.5M_{}`$, 57 km for $`M_{\mathrm{BH}}=5M_{}`$ and 72 km for $`M_{\mathrm{BH}}=10M_{}`$. The simulations were stopped at a time $`t_{\mathrm{sim}}`$ between 10 ms and $`20`$ms. The black hole was treated as a point mass at the center of a sphere with radius $`R_\mathrm{s}=2GM_{\mathrm{BH}}/c^2`$ which gas could enter unhindered. Its mass and momentum were updated along with the accretion of matter. Model TN10, which is added for comparison, is a continuation of the NS/NS merger model B64 where at time $`t_{\mathrm{sim}}=10`$ms the formation of a black hole was assumed and the accretion was followed for another 5 ms until a steady state was reached (Ruffert & Janka 1999).
## 4. RESULTS
### 4.1. Evolution of BH/NS mergers
Due to the emission of GWs the orbital separation decreases. During its first approach, the neutron star transfers matter to the black hole at huge rates of several 100 up to $`1000M_{}`$/s. Within 2–3 ms it loses 50–75% of its initial mass. In case of the 2.5$`M_{}`$ black hole the evolution is catastrophic and the neutron star is immediately disrupted (Lattimer & Schramm 1974). A mass of 0.2–0.3$`M_{}`$ remains in a thick disk around the black hole ($`M_\mathrm{d}`$ in Table 2). In contrast, the orbital distance increases again for $`M_{\mathrm{BH}}=5M_{}`$ and $`M_{\mathrm{BH}}=10M_{}`$ and a significantly less massive neutron star begins a second approach. Again, the black hole swallows gas at rates of more than $`100M_{}`$/s. Even a third cycle is possible (Fig. 1). Finally, at a distance $`d_{\mathrm{ns}}`$ and time $`t_{\mathrm{ns}}`$ the neutron star with a mass of $`M_{\mathrm{ns}}^{\mathrm{min}}`$ is destroyed and most of its mass ends up in an accretion disk (Table 2). (In case of NS/NS mergers $`t_{\mathrm{ns}}`$ means the time when the two density maxima of the stars are one stellar radius, i.e., $`d_{\mathrm{ns}}=15`$km, apart).
The increase of the orbital separation is connected with a strong rise of the specific (orbital) angular momentum of the gas (Fig. 1). Partly this is due to the fact that the black hole can capture gas with low specific angular momentum first, but mainly because only a fraction of the orbital angular momentum of the accreted gas is fed into spinning up the black hole. This fraction, which is lost for the orbital motion, is proportional to the quantity $`\alpha `$ in Fig. 2. Figure 2 is based on the parameterized analysis of non-conservative mass-transfer by Podsiadlowski, Joss, & Hsu (1992) (see also Fryer et al. 1999) assuming that mass ejection from the system is negligible. It shows that disregarding GW emission, the orbital separation can increase for small initial black hole mass only after the neutron star has lost much mass, while for larger initial $`M_{\mathrm{BH}}`$ and smaller $`\alpha `$ orbital widening is easier. Without GWs the separation increases when $`\alpha <(M_{\mathrm{BH}}M_{\mathrm{NS}})/(M_{\mathrm{BH}}+M_{\mathrm{NS}})`$. Including angular momentum loss by GWs in the point-mass approximation and using the mass-loss rates from the hydrodynamic models (dashed lines in Fig. 2) yields a qualitative understanding of the behavior visible in Fig. 1 and suggests that $`\alpha `$ is between 0.2 and 0.5.
During the merging a gas mass $`\mathrm{\Delta }M_{\mathrm{ej}}`$ of $`10^4M_{}`$ (in case of counter-rotation and $`M_{\mathrm{BH}}=2.5M_{}`$) to $`0.1M_{}`$ (corotation and $`M_{\mathrm{BH}}=10M_{}`$) is dynamically ejected (Table 2). In the latter case the associated angular momentum loss is about 7%, in all other cases it is less than 5% of the total initial angular momentum of the system. Another fraction of up to 24% of the initial angular momentum is carried away by GWs. In Table 2 the rotation parameter $`a=Jc/(GM^2)`$ is given for the initial state of the binary system ($`a_\mathrm{i}`$) and at the end of the simulation ($`a_\mathrm{f}`$) for the remnant of NS/NS mergers or for the black hole in BH/NS systems, respectively, provided the black hole did not have any initial spin. When the whole disk mass $`M_\mathrm{d}`$ has been swallowed by the Kerr black hole, a final value $`a_{\mathrm{BH}}^{\mathrm{}}`$ (Table 2) will be reached in case of the accretion of a corotating, thin disk with maximum radiation efficiency.
The phase of largest mass flow rate to the black hole (between 2 and 5 ms after the start of the simulations) is connected with a maximum of the GW luminosity $`L_{\mathrm{GW}}`$ which reaches up to $`7\times 10^{55}`$erg/s (Table 1). The peak values of $`L_{\mathrm{GW}}`$ and the wave amplitude $`rh`$ (for distance $`r`$ from the source) increase with the black hole mass. The total energy $`E_{\mathrm{GW}}`$ radiated in GWs can be as much as $`0.1M_{}c^2`$ for $`M_{\mathrm{BH}}=10M_{}`$.
### 4.2. Neutrino Emission and GRBs
Compressional heating, shear due to numerical viscosity, and dissipation in shocks heat the gas during accretion to maximum temperatures $`kT^{\mathrm{max}}`$ of several 10 MeV. Average temperatures are between 5 and 20 MeV, the higher values for the less massive and more compact black holes. At these temperatures and at densities of $`10^{10}`$$`10^{12}`$g/cm<sup>3</sup> in the accretion flow, electrons are non-degenerate and positrons abundant. Electron neutrinos and antineutrinos are therefore copiously created via reactions $`p+e^{}n+\nu _e`$ and $`n+e^+p+\overline{\nu }_e`$ and dominate the neutrino energy loss from the accreted matter. Dense and hot neutron matter is not completely transparent to neutrinos. By taking into account the finite diffusion time, the neutrino trapping scheme limits the loss of energy and lepton number.
In Table 1 maximum and average values of the luminosities ($`L_{\nu _i}^{\mathrm{max}}`$ and $`L_{\nu _i}^{\mathrm{av}}`$, respectively, the latter in brackets) in the simulated time intervals are listed for $`\nu _e`$ and $`\overline{\nu }_e`$ and for the sum of all heavy-lepton neutrinos. The latter are denoted by $`\nu _x\nu _\mu ,\overline{\nu }_\mu ,\nu _\tau ,\overline{\nu }_\tau `$ and are mainly produced by $`e^+e^{}`$ annihilation. The total neutrino luminosities $`L_\nu (t)`$ (Fig. 3) fluctuate strongly with the varying mass transfer rate to the black hole during the cycles of orbital decay and widening (compare with Fig. 1). The total energy $`E_\nu `$ radiated in neutrinos in 10–20 ms is typically several $`10^{51}`$erg. Time averages of the mean energies $`ϵ`$ of the emitted neutrinos are $`15`$MeV for $`\nu _e`$, 20 MeV for $`\overline{\nu }_e`$, and 30 MeV for $`\nu _x`$. Luminosities as well as mean energies, in particular for smaller black holes, are significantly higher than in case of NS/NS mergers.
At the end of the simulations, several of the BH/NS models have reached a steady state, characterized by only a slow growth of the black hole mass with a nearly constant accretion rate. Corresponding rates $`\dot{M}_\mathrm{d}`$ are given in Table 2 and are several $`M_{}`$/s. From these we estimate torus life times $`t_{\mathrm{acc}}=M_\mathrm{d}/\dot{M}_\mathrm{d}`$ of 50–150 ms. Values with $`>`$ and $`<`$ signs indicate cases where the evolution and emission are still strongly time-dependent at $`t_{\mathrm{sim}}`$. In these cases the accretion torus around the black hole has also not yet developed axial symmetry. In all other cases the effective disk viscosity parameter $`\alpha _{\mathrm{eff}}v_\mathrm{r}/v_{\mathrm{Kepler}}3\sqrt{6}R_\mathrm{s}/(t_{\mathrm{acc}}c)`$, evaluated at a representative disk radius of $`3R_\mathrm{s}=6GM_{\mathrm{BH}}/c^2`$, has the same value, 4–$`5\times 10^3`$. This value is associated with the numerical viscosity of the hydro code (which solves the Euler equations) for the chosen resolution. The further disk evolution is driven by the angular momentum transport mediated by viscous shear forces, which determines the accretion rate. The physical value of the disk viscosity is unknown. The numerical viscosity of our code, however, is in the range where the viscous energy dissipation and the energy emission by neutrinos should be roughly equal, i.e., where the conversion efficiency $`q_\nu =L_\nu /(\dot{M}_\mathrm{d}c^2)`$ of rest-mass energy to neutrinos is nearly maximal (see Ruffert et al. 1997; Ruffert & Janka 1999).
Assuming that the average neutrino luminosity $`L_\nu `$ at $`t_{\mathrm{sim}}`$ is representative for the subsequent accretion phase, we obtain for $`q_\nu `$ numbers between 4 and 6% and total energies $`E_\nu L_\nu t_{\mathrm{acc}}`$ around $`3\times 10^{52}`$erg (Table 2). Annihilation of neutrino pairs, $`\nu \overline{\nu }e^+e^{}`$, deposits energy at rates up to $`\dot{E}_{\nu \overline{\nu }}2\times 10^{52}`$erg/s in the vicinity of the black hole (Fig. 4). This corresponds to total energies $`E_{\nu \overline{\nu }}\dot{E}_{\nu \overline{\nu }}t_{\mathrm{acc}}`$ as high as $`10^{51}`$erg and annihilation efficiencies $`q_{\nu \overline{\nu }}=\dot{E}_{\nu \overline{\nu }}/L_\nu `$ of 1–3%. These estimates should not change much if the different effects of general relativity on $`\nu \overline{\nu }`$ annihilation are taken into account in combination (Ruffert & Janka 1999; Asano & Fukuyama 1999), but general relativistic simulations of the merging are very important. More energy could be pumped into the $`e^\pm \gamma `$ fireball when the black hole rotates rapidly (Popham, Woosley, & Fryer 1999) or if magnetic fields are able to tap the rotational energy of the accretion torus and of the black hole with higher efficiency than $`\nu \overline{\nu }`$ annihilation does (Blandford & Znajek 1977). This seems to be necessary for the long and very energetic GRBs (Mészáros, Rees, & Wijers 1999; Brown et al. 1999; Lee, Wijers, & Brown 1999).
HTJ was supported by DFG grant SFB 375 für Astro-Teilchenphysik, MR by a PPARC Advanced Fellowship, and CLF by NASA (NAG5-8128) and the US DOE ASCI Program (W-7405-ENG-48).
|
no-problem/9908/hep-th9908031.html
|
ar5iv
|
text
|
# Chaotic Quantization of Classical Gauge Fields
## 1 Introduction
Although much progress has been made in recent years, the question, how gravitation and quantum mechanics should be combined into one consistent unified theory of fundamental interactions, is still open. Superstring theory , which describes four-dimensional space-time as the low-energy limit of a ten- or eleven-dimensional theory (“M-theory” ), may provide the correct answer, but the precise form and content of the theory is not yet entirely clear. It is therefore legitimate to raise the question whether the fundamental description of nature at the Planck scale is really quantum mechanical, or whether the underlying theory could be a classical extension of general relativity. This questions was initially raised by ’t Hooft, who has argued that quantum mechanics can logically arise as low-energy limit of a microscopically deterministic, but macroscopically dissipative theory .
It is the goal of this manuscript to present an explicit example that shows how (Euclidean) quantum field theory can emerge in the infrared limit of a higher-dimensional, classical field theory. It is well known that relativistic quantum field theory in $`(3+1)`$-dimensional Minkowski space can be obtained by analytic continuation (“Wick rotation”) of the analogous statistical field theory defined on a four-dimensional Euclidean space. In fact, this concept provides the only known mathematically rigorous definition of interacting quantum field theories. Physical observables, such as vacuum expectation value of self-adjoint operators, can be reliably calculated in the Euclidean path integral formulation of the quantum field theory. This method has been extensively used to obtain numerical solutions of many relativistic quantum field theories.
We here show that in some cases, specifically for non-Abelian gauge fields, the functional integral of the three-dimensional Euclidean quantum field theory arises naturally as the long-distance limit of the corresponding classical gauge theory defined in $`(3+1)`$-dimensional Minkowski space. Because of the general nature of the mechanism underlying this transformation, for which we have coined the term chaotic quantization, it is expected to work equally well in other dimensions. For example, the four-dimensional Euclidean quantum gauge theory arises as the infrared limit of the $`(4+1)`$-dimensional classical gauge theory. We emphasize that the dimensional reduction is not caused by compactification; the classical field theory does not exhibit periodicity either in real or imaginary time.
The mechanism discussed below can be viewed as a physical analogue of the well-known technique of stochastic quantization . In both cases, the quantum fluctuations arise from the stochastic noise in a higher-dimensional theory. However, chaotic quantization differs from stochastic quantization in two essential aspects: It only applies to certain field theories, including non-Abelian gauge theories, and it allows us to calculate Planck’s constant $`\mathrm{}`$ in terms of fundamental physical quantities of the underlying higher-dimensional classical field theory. Accordingly, chaotic quantization provides a physical mechanism generating the quantum mechanics of fields and particles, while stochastic quantization is generally regarded as a convenient calculational technique, but not as a physical principle.
## 2 Chaoticity of Classical Yang-Mills Fields
The chaotic nature of classical non-Abelian gauge theories was first recognized twenty years ago . Over the past decade, extensive numerical solutions of spatially varying classical non-Abelian gauge fields on the lattice have revealed that the gauge field has positive Lyapunov exponents that grow linearly with the energy density of the field configuration and remain well-defined in the limit of small lattice spacing $`a`$ or weak-coupling . More recently, numerical studies have shown that the $`(3+1)`$-dimensional classical non-Abelian lattice gauge theory exhibits global hyperbolicity. This conclusion is based on calculations of the complete spectrum of Lyapunov exponents and on the long-time statistical properties of the Kolmogorov-Sinai (KS) entropy of the classical SU(2) gauge theory .
These results imply that correlation functions of physical observables decay rapidly, and that long-time averages of observables for a single initial gauge field configuration are identical to their microcanonical phase-space average, up to Gaussian fluctuations which vanish in the long-time limit as $`t_s^{1/2}`$, where $`t_s`$ is the observation time. Since the relative fluctuations of extensive quantities scale as $`L^{3/2}`$, the microcanonical (fixed-energy) average can be safely replaced by the canonical average when the spatial volume probed by the observable becomes large. In the following we discuss the hierarchy of time and length scales on which this transformation occurs.
Accoridng to the cited results, the classical non-Abelian gauge field self-thermalizes on a finite time scale $`\tau _{\mathrm{eq}}`$ given by the ratio of the equilibrium entropy and the KS-entropy, which determines the growth rate of the course-grained entropy:
$$\tau _{\mathrm{eq}}=S_{\mathrm{eq}}/h_{\mathrm{KS}}.$$
(1)
At weak coupling, the KS-entropy for the $`(3+1)`$-dimensional SU(2) gauge theory scales as
$$h_{\mathrm{KS}}g^2Eg^2T(L/a)^3,$$
(2)
where $`E`$ is the total energy of the field configuration and $`T`$ is the related temperature defined by
$$E=T^2Z/T.$$
(3)
Here $`Z(T)`$ is the partition function of the classical gauge field regularized by the lattice cut-off $`a`$. The equilibrium entropy of the lattice is independent of the energy and proportional to the number of degrees of freedom of the lattice: $`S_{\mathrm{eq}}(L/a)^3`$. The time scale for self-equilibration is thus given by<sup>1</sup><sup>1</sup>1In the convention adopted here, $`g`$ is the coupling strength of the classical Yang-Mills theory with dimension (energy$`\times `$length)<sup>-1/2</sup>. This choice ensures the proper dimensionality of the Yang-Mills action.
$$\tau _{\mathrm{eq}}(g^2Ea^3/L^3)^1(g^2T)^1.$$
(4)
When one is interested only in long-term averages of observables, it is thus sufficient to consider the thermal classical gauge theory on a three-dimensional spatial lattice. Note that, although we are interested only in the long-distance properties of the quasi-thermalized field, we have to define the classical gauge field on a lattice rather than in the continuum. Due to the nonlinear interactions most of the energy contained in the initial field configuration ultimately cascades into modes with wave lengths near the ultraviolet cutoff $`a`$, and the limit of vanishing lattice spacing $`a`$ is not well defined. Without the lattice cutoff, we would be unable to replace the microcanonical average by a thermal average, because the cascade toward the ultraviolet would not end and no stationary limit would exist. We shall see later that the lattice cutoff $`a`$ also takes on an important physical role, as it enters into the definition of Planck’s constant $`\mathrm{}`$.
## 3 Chaoticity of Classical Yang-Mills Theory
The long-distance dynamics of non-Abelian gauge theories at finite temperature $`T`$ is known to reduce to the dynamics of the static chromomagnetic sector of the gauge field . The scale beyond which this dimensional reduction is valid is given by the magnetic length scale $`d_{\mathrm{mag}}(g^2T)^1`$, where $`g`$ is the classical gauge coupling. The magnetic length scale is the same for the classical and the quantized gauge theories at finite temperature. $`d_{\mathrm{mag}}`$ is independent of the lattice cutoff $`a`$. It is well recognized that the static magnetic sector in the thermal quantum field theory is essentially classical in nature and depends on $`\mathrm{}`$ only via the scale of the thermal effective gauge coupling $`g(T)`$.
It is worth noting that in spite of many similarities between the thermal classical field theory and the thermal quantum field theory, there are major differences. The ultraviolet properties of the quantum field theory at finite temperature are controlled by the thermal length scale $`d_{\mathrm{th}}\mathrm{}/T`$, which is a basically quantum mechanical concept.<sup>2</sup><sup>2</sup>2The need for this length scale in the derivation of the Stefan-Boltzmann radiation law motivated Planck in 1900 to postulate the existence of the quantum of action $`\mathrm{}`$. In the thermal classical field theory, the lattice spacing $`a`$ serves as ultraviolet regulator. Similarly, the electric screening length of the thermal quantum field theory is $`d_{\mathrm{el}}\sqrt{\mathrm{}}/gT`$. In the thermal classical field theory electric fields are screened on the length scale $`d_{\mathrm{el}}\sqrt{a/g^2T}`$. Only the magnetic length scales are equal for classical and quantal gauge fields. The inverse electric screening length is proportional to the plasma frequency $`\omega _{\mathrm{pl}}`$ governing propagating long-wavelength modes. The damping of these plasma modes is of the order $`\gamma _{\mathrm{pl}}d_{\mathrm{mag}}^1`$ , rendering the dynamics of the thermal gauge field purely dissipative and noisy on distances larger than $`d_{\mathrm{mag}}`$.
The dynamic properties of thermal non-Abelian gauge fields at such long distances have been studied in much detail . It is now understood that the real-time dynamics of the gauge field at such scales can be described, at leading order, by a Langevin equation
$$\sigma \frac{A}{t}=D\times B+\xi ,$$
(5)
where $`D`$ is the gauge covariant spatial derivative, $`B=D\times A`$ is the magnetic field strength, and $`\xi `$ denotes Gaussian distributed (white) noise with zero mean and variance
$$\xi _i(x,t)\xi _j(x^{},t^{})=2\sigma T\delta _{ij}\delta ^3(xx^{})\delta (tt^{}).$$
(6)
Here $`\sigma `$ denotes the color conductivity of the thermal gauge field which is determined by the ratio $`\omega _{\mathrm{pl}}^2/\gamma _{\mathrm{th}}`$ of the plasma frequency $`\omega _{\mathrm{pl}}`$ and the damping rate $`\gamma _{\mathrm{th}}`$ of a thermal gauge field excitation.
At leading logarithmic order in the quantum field theory, the color conductivity satisfies
$$\sigma ^1\frac{\mathrm{}}{T}\mathrm{ln}[d_{\mathrm{mag}}/d_{\mathrm{el}}].$$
(7)
The derivation of the Langevin equation (5) for the classical thermal gauge theory proceeds completely in parallel to that for the quantum field theory, except that the plasma frequency $`\omega _{\mathrm{pl}}^2g^2T/a`$, and the ratio between the magnetic and electric length scales depends on the combination $`g(Ta)^{1/2}`$ instead of $`g\mathrm{}^{1/2}`$. The separation of length scales requires $`g^2Ta1`$. The color conductivity then scales as
$$\sigma ^1a\mathrm{ln}[d_{\mathrm{mag}}/d_{\mathrm{el}}].$$
(8)
This relation implies that the color conductivity is an ultraviolet sensitive quantity, which depends on the lattice cutoff. The various relations derived in this section are summarized in Table 1, where the results for the thermal quantum field theory are listed in parallel with those of the classical gauge theory.
## 4 Dimensional Reduction and Quantization
We will now show that an observer restricted to long distances in three-dimensional Euclidean space would interpret the dynamics of the classical gauge field as that of a quantum field in its vacuum state with the Planck constant
$$\mathrm{}_3=aT.$$
(9)
The starting point of our argument is the well-established fact that the random Gaussian process defined by the Langevin equation (5) generates three-dimensional field configurations with a probability distribution $`P[A]`$ determined by the Fokker-Planck equation
$$\sigma \frac{}{t}P[A]=d^3x\frac{\delta }{\delta A}\left(T\frac{\delta P}{\delta A}+\frac{\delta W}{\delta A}P[A]\right).$$
(10)
Here $`W[A]`$ denotes the magnetic energy functional
$$W[A]=d^3x\frac{1}{2}B(x)^2.$$
(11)
Any non-static excitations of the magnetic sector of the gauge field, i.e. magnetic fields $`B(k)`$ not satisfying $`k\times B=0`$, die away rapidly on a time scale of order $`\sigma /k^2`$, where $`k`$ denotes the wave vector of the field excitation. For observers sensitive only to distances much larger than $`d_{\mathrm{mag}}`$ and times much longer than $`\sigma d_{\mathrm{mag}}^2`$, measurements of the magnetic field yield averages with the equilibrium weight given by the stationary solution of the Fokker-Planck equation (10):
$$P_0[A]=e^{W[A]/T}.$$
(12)
The three-vector $`B_iϵ_{ijk}F^{jk}`$ incorporates all components of the field strength tensor $`f^{jk}`$ in three dimensions. $`W/T`$ is now identified as the three-dimensional action $`S_3`$ measured in units of Planck’s constant $`\mathrm{}_3`$:
$$W/T=S_3/\mathrm{}_3,$$
(13)
where
$$S_3[A]=\frac{1}{4}𝑑x_3d^2xf^{ik}f_{ik}.$$
(14)
Dimensional reasons require a rescaling of the gauge field strength with the fundamental length scale
$$f^{ik}=\sqrt{a}F^{ik}.$$
(15)
That the lattice spacing $`a`$ is the proper rescaling parameter is seen by noting that the lattice versions of the two-dimensional and the three-dimensional integrals
$$d^nxa^n\underset{x}{}$$
(16)
differ by a factor $`a`$. Together with the relations (11), (13), and (14) this reasoning demonstrates that $`\mathrm{}_3=aT`$, as stated at the beginning of this section. The rescaling of the gauge field also fixes the three-dimensional coupling constant
$$g_3^2=\frac{g^2}{a}=\frac{g^2T}{\mathrm{}_3},$$
(17)
so that
$$\sqrt{a}(A+gA\times A)=(A_{(3)}+g_3A_{(3)}\times A_{(3)}).$$
(18)
According to (10), an observer confined to the measurement of long-time and long-distance averages of microscopic observables associated with the classical gauge field measures the same values as would an observer “living” in the three-dimensional Euclidean world in the presence of a quantized gauge field in its vacuum state. It is important to note that this correspondence is not induced by a compactification of the Minkowskian time coordinate. There is no true thermal bath of gauge fields in the original Minkowski space theory, and the quasi-thermal solution of the $`(3+1)`$-dimensional classical field theory does not satisfy periodic boundary conditions in imaginary time.
The effective dimensional reduction found here is not caused by the discreteness of the excitations with respect to the time-like dimension, but by the dissipative nature of the (3+1)-dimensional dynamics. Magnetic field configurations satisfying $`D\times B=0`$ can be thought of as low-dimensional attractors of the dissipative motion, and the chaotic dynamical fluctuations of the gauge field around the attractor can be consistently interpreted as quantum fluctuations of a vacuum gauge field in 3-dimensional Euclidean space.
We thus see that the mechanism of dimensional reduction discussed above is distinct from the mechanism that operates in thermal quantum field theories. In fact, the dimensional reduction by chaotic fluctuations and dissipation does not occur in scalar field theories, because – even in cases that exhibit chaos, such as two quarticly coupled scalar fields – there is no dynamical sector that survives after long-time averaging. Quasi-thermal fluctuations generate a dynamical “mass” for the scalar field(s) and thus eliminate any arbitraily slow field modes.<sup>3</sup><sup>3</sup>3An exception may be the case where the excitation energy of the scalar field is just right to put the quasi-thermal field at the critical temperature of a second-order phase transition, where arbitraily slow modes exist as fluctuations of the order parameter. In the case of gauge fields, the transverse magnetic sector is protected by the gauge symmetry, and it is this sector which survives the time average, without any need for fine-tuning of the microscopic theory.
## 5 General Considerations
It is worthwhile to review the essential ingredients of chaotic quantization. First, the underlying classical field theory must contain strongly coupled massless degrees of freedom. Such theories are generally chaotic at the classical level . When observations are restricted to the infrared degrees of freedom, this corresponds to a coarse graining of the dynamical system, leading to strongly dissipative long-distance dynamics. The coupling to the short-distance modes generates uncorrelated noise, and the coarse-grained system obeys a dissipation-fluctuation theorem .
Second, the classical field theory at finite temperature must have degrees of freedom that remain unscreened. This condition generally requires the presence of a symmetry, such as gauge invariance. It is a reasonable expectation that such symmetries occur in any unified theory containing general relativity. The requirement also provides a simple and consistent explanation for the empirical fact that all fundamental interactions are described by gauge fields.
Our example for the chaotic quantization of a three-dimensional gauge theory in Euclidean space raises a number of questions:
1. Does the principle of chaotic quantization generalize to higher dimensions, in particular, to quantization in four dimensions?
2. Can the method be extended to describe field quantization in Minkowski space?
3. What type of deviations from the standard quantum field theory are caused by the existence of a microscopic classical dynamics?
The first question is most easily answered. As long as globally hyperbolic classical field theories can be identified in higher dimensions, our proposed mechanism should apply. Although we do not know of any systematic study of dicretized field theories in higher dimensions, a plausibility argument can be made that Yang-Mills fields exhibit chaos in (4+1) dimensions. For this purpose, we consider the infrared limit of a spatially constant gauge potential, as studied in refs. . For the SU($`N`$) gauge field in $`(D+1)`$ dimensions in the $`A_0=0`$ gauge, there are $`3(N^21)`$ interacting components of the vector potential and $`3(N^21)`$ canonically conjugate momenta (the components of the electric field) that depend only on the time coordinate. The remaining gauge transformations and Gauss’ law allow to eliminate $`(N^21)`$ degrees of freedom from each. Next, rotational invariance in $`D`$ dimensions permits to reduce the number of dynamical degrees of freedom by twice the number of generators of the group SO($`D`$), i.e. by $`D(D1)`$. This leaves a $`(D1)(2N^22D)`$-dimensional phase space of the dynamical degrees of freedom and their conjugate momenta. For the dynamics to be chaotic, this number must be at least three. For the simplest gauge group SU(2), this condition permits infrared chaos in $`2D5`$ dimensions, including the interesting case $`D=4`$. Higher gauge groups are needed to extend the chaotic quantization scheme to gauge fields in $`D>5`$ dimensions. Of course, this reasoning does not prove full chaoticity of the Yang-Mills field in these higher dimensions, it just indicates the possibility. Numerical studies will be required to establish the presence of strong chaos in these classical field theories.
The second question is more difficult to address. A formal answer would be that the Minkowski-space quantum field theory can (and even must) be obtained by analytic continuation from the Euclidean field theory. Any observable in the Minkowski-space theory that can be expressed as a vacuum expectation value of field operators can be obtained in this manner. If this argument appears somewhat unphysical, one might consider a completely different approach, beginning with a chaotic classical field theory defined in (3+2) dimensions. Field theories defined in spaces with two time-like dimensions were first proposed by Dirac in the context of conformal field theory and have recently been considered as generalizations of superstring theory . In that case, the reduction to one time dimension is achieved by gauge fixing. In the present case, the physical time dimension may be defined as the coordinate orthogonal to the total 5-momentum vector $`P^\mu `$ of the initial field configuration.<sup>4</sup><sup>4</sup>4In the four-dimensional case, the total 4-momentum vector $`P^\mu `$, which is assumed to be time-like, defines the 4-velocity vector $`u^\mu `$ of the thermal rest frame via the relation $`P^\mu =E(T)u^\mu `$. The three-dimensional Euclidean quantum field theory lives in the hypersurface orthogonal to $`u^\mu `$.
In the presence of two time-like dimensions, “energy” becomes a two-component vector $`\stackrel{}{E}`$ that is a part of the $`(D+2)`$-dimensional energy-momentum vector. If we select an initial field configuration with energy $`E_0\stackrel{}{n}`$, where $`\stackrel{}{n}`$ is a two-dimensional unit vector, this choice defines a preferred time-like direction $`\stackrel{}{n}`$ in which the field thermalizes. Conservation of the energy-momentum vector ensures that the total energy component orthogonal to $`\stackrel{}{n}`$ always remains zero. In this sense, the choice of an initial field configuration corresponds to a spontaneous breaking of the global SO($`D`$,2) symmetry down to a global SO($`D`$,1) symmetry. Whether this leads to an effective quantum field theory in $`(D+1)`$ dimensional Minkowski space, remains to be investigated.
Finally, it is interesting to ask which deviations from the quantum field theory could be detected by a “slow” observer by means of very precise measurements. Clearly, an observer able to resolve the dynamics on the thermal or electric length scales of the underlying classical field theory would observe deviations from the dimensionally reduced vacuum field theory. For a space-time volume of linear dimension $`L`$, the amplitude of the fluctuations is of the order $`(g^2TL)^2`$. If, as one might suspect, the relevant microscopic length scale is of the order of the Planck scale, $`g^2TM_P`$, quantities sensitive to the fluctuations around the infrared dynamics are suppressed by $`(M_PL)^2`$. For presently accessible length scales, the suppression factor is at least $`10^{34}`$, and even smaller in low-energy precision tests of quantum mechanics. However, in principle, tests of Bell’s inequality in systems prepared with strong correlations on short time and distance scales can be used to establish at least an upper bound for the scale at which the transition from the classical dissipative dynamics to the quantum dynamics occurs.
It is a natural question to ask whether the mechanism of chaotic quantization outlined above corresponds to a hidden parameter theory of quantum mechanics. The answer is obviously positive, as the microscopic state of the system in a higher dimension is always precisely and deterministically defined. However, it is important to realize that the impossibility of hidden parameter descriptions of quantum mechanics is restricted to local theories, while our proposed mechanism operates in a higher dimensional space. A local dynamical theory in more dimensions generates fundamentally non-local effects in the lower-dimensional space. One can speculate that the time-scale associated with dimensional reduction, $`(g^2T)^1`$, is the time for the collapse of the wave function. Our analysis predicts that this is the time required to average over the noise generated by the classical dynamics and to establish the stationary distribution of the Fokker-Planck equation (10).
## 6 Summary and Conclusions
Let us summarize. We have shown that a homogeneously excited, classical field theory in four dimensions can generate a three-dimensional Euclidean quantum field theory. Whereas the classical theory appears thermal for a four-dimensional observer, the three-dimensional observer experiences quantum fields at zero temperature. The essential feature facilitating this transformation is that the underlying deterministic theory contains a mechanism for information loss , here realized through its chaotic dynamics.
We can go further and speculate that the randomness caused by this intrinsic chaoticity of the underlying theory could generally lead to a reduction of the effective space-time dimensionality of the theory. In our example, the dimensional reduction is an effect of the quasi-thermal fluctuations. Another related, well-known phenomenon is the dimensional reduction of spin systems in arbitrarily weak, random magnetic fields , which finds its explanation in the hidden supersymmetry of the system .
We note that symmetries and physical laws may arise naturally from some essentially random dynamics, rather than being postulated from the beginning . The goal of the program formulated in our example, is more restricted: microscopic randomness is utilized as a foundation for “large-scale” physics that is described by quantum mechanics.
It is not clear whether the mechanism presented here for non-Abelian gauge fields is also at work in general relativity. Examples of chaotic behavior have been identified in the dynamics of classical gravitational fields . It has been found that the chaotic nature of the solutions may depend on the number of dimensions. A famous case is the evolution toward the singularity in the Bianchi type IX geometry, where the chaotic oscillatory approach changes to a monotonic approach in more than 10 dimensions .
It has not been demonstrated that chaoticity is a general property of solutions of Einstein’s equations. This may not even be required, because an entirely different mechanism of information loss may be at work in general relativity than in Yang-Mills theory. Indeed, ’t Hooft has speculated that black hole formation may be the mechanism operating in the case of gravity . Finally, our example does not contain fermion fields. It would be interesting to extend our study to supersymmetric theories, some of which have been shown to exhibit chaos in the infrared limit .
Acknowledgments: This work was supported in part by a grant from the U.S. Department of Energy (DE-FG02-96ER40495), by the American-Hungarian Joint Fund TÉT (JFNo. 649) and by the Hungarian National Science Fund OTKA (T 019700). One of us (B.M.) acknowledges the support by a U. S. Senior Scientist Award from the Alexander von Humboldt Foundation.
## Appendix
Here we present a qualitative analysis of the various length and time scales of a thermal Yang-Mills field in $`D=(d+1)`$ space-time dimensions, for $`d2`$. The results are accurate up to logarithmic corrections that arise for various quantities in some dimensions. We decompose the field into Fourier components (suppressing vector and color indices):
$$A(x,t)=V^{1/2}d^dka(k)e^{ikxi\omega _kt}.$$
(19)
Equipartition of the energy over the classical modes then implies that
$$|a(k)|^2=T\omega _k^2.$$
(20)
The presence of an “external” static color potential $`A^0`$ induces a polarization density
$$\rho _{\mathrm{pol}}g^2V^1d^dxA(x,t)^2A^0g^2A^0d^dk|a(k)|^2g^2TA^0d^dk\omega _k^2.$$
(21)
With the ultraviolet lattice cut-off $`ka^1`$, one obtains:
$$d_{\mathrm{el}}^2=\rho _{\mathrm{pol}}/A_0g^2Ta^{2d}.$$
(22)
The field theory can be considered to in the weak coupling regime, when the electric screening length is much larger than the lattice constant $`a`$. This amounts to the condition
$$\stackrel{~}{g}^2g^2Ta^{4d}=(a/d_{\mathrm{el}})^21,$$
(23)
which defines the effective weak coupling parameter $`\stackrel{~}{g}`$ of the $`(d+1)`$-dimensional classical Yang-Mills theory. With the help of this parameter, the electric screening length can be expressed simply as $`d_{\mathrm{el}}=a/\stackrel{~}{g}`$.
The coupling constant of the dimensionally reduced quantum field theory in $`D1=d`$ dimensional Euclidean space is given by
$$g_{D1}^2\mathrm{}=g^2a^1(Ta)=g^2T,$$
(24)
independent of the number $`d`$ of space dimensions.
The transport coefficient describing color conductivity $`\sigma `$ is obtained by considering the polarization current induced by a constant electric field $`E`$. Schematically, it is given by
$$j_{\mathrm{pol}}=g^2V^1d^dx𝑑tA(x,t)^2Eg^2Ed^dk|a(k)|^2\gamma (k)^1,$$
(25)
where $`\gamma (k)`$ is the damping rate of a thermal field mode. This damping rate can be calculated in the standard way using the formula $`\gamma =\sigma _{\mathrm{coll}}n_{\mathrm{th}}`$, where $`\sigma _{\mathrm{coll}}`$ denotes the “cross section” for a thermal excitation, and $`n_{\mathrm{th}}`$ stands for the density of hard thermal excitations. In $`d`$ spatial dimensions one finds
$$\sigma _{\mathrm{coll}}g^4Tad^{d1}q(q^2+d_{\mathrm{el}}^2)^2g^4Tad_{\mathrm{el}}^{5d}.$$
(26)
The classical formula contains an additional factor $`(Ta)`$ describing the enhancement due to the classical occupation of thermal modes of the gauge field. The density of thermal excitations is
$$n_{\mathrm{th}}d^dk|a(k)|^2\omega _kTa^{1d}.$$
(27)
Since $`\gamma `$ does not depend on $`k`$ in this approximation, we can pull it out of the integral over $`k`$ in (25) to obtain:
$$j_{\mathrm{pol}}g^2E\gamma ^1d^dk|a(k)|^2g^2E\gamma ^1Ta^{2d}.$$
(28)
Combining the expressions (2628) we final get the desired expression for the color conductivity:
$$\sigma =j_{\mathrm{pol}}/Eg^2a/\sigma _{\mathrm{coll}}d_{\mathrm{el}}^{d5}(g^2T)^1.$$
(29)
Specifically, in $`d=4`$ spatial dimensions, the result is
$$\sigma a(g^2T)^{3/2}.$$
(30)
All results are summarized, and compared to the results obtained in the $`(d+1)`$-dimensional thermal quantum field theory, in Table 2.
|
no-problem/9908/hep-ph9908408.html
|
ar5iv
|
text
|
# References
LU TP 99–23
hep-ph/9908408
August 1999
QCD Interconnection Effects<sup>1</sup><sup>1</sup>1To appear in the Proceedings of the International Workshop on Linear Colliders, Sitges (Barcelona), Spain, April 28 – May 5, 1999
Torbjörn Sjöstrand<sup>2</sup><sup>2</sup>[email protected]
Department of Theoretical Physics,
Lund University, Lund, Sweden
and
Valery A. Khoze<sup>3</sup><sup>3</sup>[email protected]
INFN – Laboratori Nazionali di Frascati,
P.O. Box 13, I-00044 Frascati (Roma), Italy
Abstract
Heavy objects like the $`W`$, $`Z`$ and $`t`$ are short-lived compared with typical hadronization times. When pairs of such particles are produced, the subsequent hadronic decay systems may therefore become interconnected. We study such potential effects at Linear Collider energies.
The widths of the $`W`$, $`Z`$ and $`t`$ are all of the order of 2 GeV. A Standard Model Higgs with a mass above 200 GeV, as well as many supersymmetric and other Beyond the Standard Model particles would also have widths in the (multi-)GeV range. Not far from threshold, the typical decay times $`\tau =1/\mathrm{\Gamma }0.1\mathrm{fm}\tau _{\mathrm{had}}1\mathrm{fm}`$. Thus hadronic decay systems overlap, between pairs of resonances ($`W^+W^{}`$, $`Z^0Z^0`$, $`t\overline{t}`$, $`Z^0H^0`$, …), so that the final state may not be just the sum of two independent decays. Pragmatically, one may distinguish three main eras for such interconnection:
* Perturbative: this is suppressed for gluon energies $`\omega >\mathrm{\Gamma }`$ by propagator/timescale effects; thus only soft gluons may contribute appreciably.
* Nonperturbative in the hadroformation process: normally modelled by a colour rearrangement between the partons produced in the two resonance decays and in the subsequent parton showers.
* Nonperturbative in the purely hadronic phase: best exemplified by Bose–Einstein effects.
The above topics are deeply related to the unsolved problems of strong interactions: confinement dynamics, $`1/N_\mathrm{C}^2`$ effects, quantum mechanical interferences, etc. Thus they offer an opportunity to study the dynamics of unstable particles, and new ways to probe confinement dynamics in space and time , but they also risk to limit or even spoil precision measurements .
So far, studies have mainly been performed in the context of $`W`$ mass measurements at LEP2. Perturbative effects are not likely to give any significant contribution to the systematic error, $`\delta m_W\text{ }\stackrel{<}{}\text{ }5`$ MeV . Colour rearrangement is not understood from first principles, but many models have been proposed to model effects , and a conservative estimate gives $`\delta m_W\text{ }\stackrel{<}{}\text{ }40`$ MeV. For Bose–Einstein again there is a wide spread in models, and an even wider one in results, with about the same potential systematic error as above . The total QCD interconnection error is thus below $`m_\pi `$ in absolute terms and 0.1% in relative ones, a small number that becomes of interest only because we aim for high accuracy.
More could be said if some experimental evidence existed, but a problem is that also other manifestations of the interconnection phenomena are likely to be small in magnitude. For instance, near threshold it is expected that colour rearrangement will deplete the rate of low-momentum particle production . Even with full LEP2 statistics, we are only speaking of a few sigma effects, however. Bose-Einstein appear more promising to diagnose, but so far experimental results are contradictory .
One area where a linear collider could contribute would be by allowing a much increased statistics in the LEP2 energy region. A 100 fb<sup>-1</sup> $`W^+W^{}`$ threshold scan would give a $`6`$ MeV accuracy on the $`W`$ mass , with negligible interconnection uncertainty. This would shift the emphasis from $`m_W`$ to the understanding of the physics of hadronic cross-talk. A high-statistics run, e.g. 50 fb<sup>-1</sup> at 175 GeV, would give a comfortable signal for the low-momentum depletion mentioned above, and also allow a set of other tests . Above the $`Z^0Z^0`$ threshold, the single-$`Z^0`$ data will provide a unique $`Z^0Z^0`$ no-reconnection reference.
Thus, high-luminosity, LEP2-energy LC (Linear Collider) runs would be excellent to establish a signal. To explore the character of effects, however, a knowledge of the energy dependence could give further leverage.
In QED, the interconnection rate dampens with increasing energy roughly like $`(1\beta )^2`$, with $`\beta `$ the velocity of each $`W`$ in the CM frame . By contrast, the nonperturbative QCD models we studied show an interconnection rate dropping more like $`(1\beta )`$ over the LC energy region (with the possibility of a steeper behaviour in the truly asymptotic region). If only the central region of $`W`$ masses is studied, also the mass shift dampens significantly with energy. However, if also the wings of the mass distribution are included (a difficult experimental proposition, but possible in our toy studies), the average and width of the mass shift distribution do not die out. Thus, with increasing energy, the hadronic cross-talk occurs in fewer events, but the effect in these few is more dramatic.
The depletion of particle production at low momenta, close to threshold, turns into an enhancement at higher energies . However, in the inclusive $`W^+W^{}`$ event sample, this and other signals appear too small for reliable detection. One may instead turn to exclusive signals, such as events with many particles at low momenta, or at central rapidities, or at large angles with respect to the event axis. Unfortunately, even after such a cut, fluctuations in no-reconnection events as well as ordinary QCD four-jet events (mainly $`q\overline{q}gg`$) give event rates that overwhelm the expected signal. It could still be possible to observe an excess, but not to identify reconnections on an event-by-event basis. The possibility of some clever combination of several signals still remains open, however.
Since the $`Z^0`$ mass and properties are well-known, $`Z^0Z^0`$ events provide an excellent hunting ground for interconnection. Relative to $`W^+W^{}`$ events, the set of production Feynman graphs and the relative mixture of vector and axial couplings is different, however, and this leads to non-negligible differences in angular distributions. Furthermore, the higher $`Z^0`$ mass means that a $`Z^0`$ is slower than a $`W^\pm `$ at fixed energy, and the larger $`Z^0`$ width also brings the decay vertices closer. Taken together, at 500 GeV, the reconnection rate in $`Z^0Z^0`$ hadronic events is likely to be about twice as large as in $`W^+W^{}`$ events, while the cross section is lower by a factor of six. Thus $`Z^0Z^0`$ events are interesting in their own right, but comparisons with $`W^+W^{}`$ events will be nontrivial.
As noted above, the Bose–Einstein interplay between the hadronic decay systems of a pair of heavy objects is at least as poorly understood as is colour reconnection, and less well studied for higher energies. In some models , the theoretical mass shift increases with energy, when the separation of the $`W`$ decay vertices is not included. With this separation taken into account, the theoretical shift levels out at around 200 MeV. How this maps onto experimental observables remains to be studied, but experience from LEP2 energies indicates that the mass shift is significantly reduced, and may even switch sign.
The $`t\overline{t}`$ system is different from the $`W^+W^{}`$ and $`Z^0Z^0`$ ones in that the $`t`$ and $`\overline{t}`$ always are colour connected. Thus, even when both tops decay semileptonically, $`tbW^+b\mathrm{}^+\nu _{\mathrm{}}`$, the system contains nontrivial interconnection effects. For instance, the total hadronic multiplicity, and especially the multiplicity at low momenta, depends on the opening angle between the $`b`$ and $`\overline{b}`$ jets: the smaller the angle, the lower the multiplicity . On the perturbative level, this can be understood as arising from a dominance of emission from the $`b\overline{b}`$ colour dipole at small gluon energies , on the nonperturbative one, as a consequence of the string effect .
Uncertainties in the modelling of these phenomena imply a systematic error on the top mass of the order of 30 MeV already in the semileptonic top decays. When hadronic $`W`$ decays are included, the possibilities of interconnection multiply. This kind of configurations have not yet been studied, but realistically we may expect uncertainties in the range around 100 MeV.
In summary, LEP2 may clarify the Bose–Einstein situation and provide some hadronic cross-talk hints. A high-luminosity LEP2-energy LC run would be the best way to establish colour rearrangement, however. Both colour rearrangement and BE effects (may) remain significant over the full LC energy range: while the fraction of the (appreciably) affected events goes down with energy, the effect per such event comes up. If the objective is to do electroweak precision tests, it appears feasible to reduce the $`WW/ZZ`$ “interconnection noise” to harmless levels at high energies, by simple proper cuts. It should also be possible, but not easy, to dig out a colour rearrangement signal at high energies, with some suitably optimized cuts that yet remain to be defined. The $`Z^0Z^0`$ events should display about twice as large interconnection effects as $`W^+W^{}`$ ones, but cross sections are reduced even more. The availability of a single-$`Z^0`$ calibration still makes $`Z^0Z^0`$ events of unique interest. While detailed studies remain to be carried out, it appears that the direct reconstruction of the top mass could be uncertain by maybe 100 MeV. Finally, in all of the studies so far, it has turned out to be very difficult to find a clean handle that would help to distinguish between the different models proposed, both in the reconnection and Bose–Einstein areas. Much work thus remains for the future.
A copy of the transparencies of this talk, including all the figures not shown here (for space reasons) may be found on
http://www.thep.lu.se/$``$torbjorn/talks/sitges99ww.ps.
A longer writeup is in preparation.
|
no-problem/9908/astro-ph9908129.html
|
ar5iv
|
text
|
# On the Formation of Boxy and Disky Elliptical Galaxies
## 1 Introduction
Elliptical galaxies have long been considered as old, coeval systems, consisting of a dynamically relaxed, spheroidal stellar population with a universal $`r^{1/4}`$ surface brightness profile. More detailed observations have however shown that these systems can be subdivided into two classes with distinct kinematical and orbital properties. Low-luminosity ellipticals are isotropic and rotationally supported, with small minor axis rotation and disky deviations of their isophotal shapes from perfect ellipses (Bender (1988); Bender, Döbereiner & Möllenhoff (1988), hereafter BDM; Kormendy & Bender (1996) and references therein). High-luminosity ellipticals, on the other hand, are anisotropic, slow rotators with large minor axis rotation, boxy isophotes and, occasionally, with kinematically distinct cores. Bender el al. (1989) demonstrated that both groups have different radio and X-ray luminosities and recent high-resolution observations exhibit that disky ellipticals have steep power-law cores in contrast to boxy ellipticals with flat cores and central density cusps (Lauer et al. (1995), Faber et al. (1997)).
On the theoretical side, Toomre & Toomre (1972) proposed that early type galaxies originate from major mergers of disk galaxies. This ”merger hypothesis” has been investigated in great details by numerous authors using numerical simulations (see Barnes & Hernquist (1992) for a review). The first fully self-consistent merger models of two equal-mass rotationally supported stellar disks, embedded in dark matter halos were performed by Barnes (1988) and Hernquist (1992). They found that mergers indeed lead to a slowly rotating, pressure supported, anisotropic spheroidal system. In the inner regions, the remnants were however too diffuse, leading to strong deviations from the observed de Vaucouleurs profiles which requires an inner $`r^1`$ density profile. This result can be explained by phase space limitations (Carlberg (1986)). Subsequent investigations by Hernquist et al. (1993) showed that mergers of progenitors with massive bulge components could resolve this problem, leading to ellipticals with small core radii and surface brightness profiles that are in excellent agreement with observations. Hernquist (1993b) and subsequently Heyl, Hernquist & Spergel (1994) and Steinmetz & Buchner (1995) noted departures from pure ellipses in their equal-mass merger remnants. The same remnant when viewed from different orientations appeared either boxy or disky. This seems to be in contradiction with the observations. As disky and boxy ellipticals have different radio and X-ray properties that should not depend on viewing angle their isophotal shapes cannot change as a result of projection effects. In agreement with observations of boxy ellipticals, Barnes(1992) and Heyl et al. (1996) also found misalignments between the spin and minor axis in major merger remnants which however seemed to be larger than observed (Franx Illingworth & De Zeeuw (1991)).
It has been argued by Kormendy & Bender (1996) and Faber et al. (1997) that gaseous mergers lead to distinct inner gaseous disks in the merger remnants which subsequently turn into stars, generating disky isophotes. In contrast, boxy ellipticals would form from purely dissipationless mergers. This idea has been addressed in details by Bekki & Shioya (1997) and Bekki (1998). Bekki & Shioya (1997) simulated mergers including gaseous dissipation and star formation. They found that the rapidity of gas consumption affects the isophotal shapes. Secular star formation however leads to final density profiles which deviate significantly from the observed r<sup>1/4</sup>-profiles in radial regimes where all ellipticals show almost perfect de Vaucouleurs laws (Burkert 1993). These calculations and models by Mihos & Hernquist (1996) demonstrate that the effect of gas and star formation changes the structure of merger remnants as such a dissipative component would most likely lead to strong deviations from the r<sup>1/4</sup>-profiles which seems to be a result of dissipationless, violent relaxation processes. Nevertheless the observations of metal enhanced, decoupled and rapidly spinning disk-like cores (Bender & Surma (1992); Davies Sadler & Peletier (1993); Bender & Davies (1996)) shows that even in boxy ellipticals gas must have present. Numerical simulations show, that these features would result naturally from gas infall during the merger process (Barnes & Hernquist (1996); Mihos & Hernquist (1996)). The influence of gas on the global structure of elliptical galaxies is not well understood as it is sensitive to uncertain details about star formation (Barnes & Hernquist (1996)).
Recently Barnes (1998) proposed a scenario of the origin of rapidly rotating ellipticals that does not require gaseous dissipation. He showed that such systems would result from the merger of a large disk galaxy with a smaller companion. In addition, the edge on view shows a disky morphology. Taking this model into account, boxy and disky ellipticals should result from equal- and unequal-mass mergers, respectively.
In this letter the Barnes hypothesis is investigated in greater details. We present the results of two merger simulations, a 1:1 merger and a 3:1 merger. The definition of the global orbital and kinematical properties depends strongly on the method used. We therefore apply the same data reduction method to our model galaxies which is used to derive the global parameters of observed ellipticals and with which we compare our results. The equal-mass merger indeed leads to a boxy elliptical whereas the unequal-mass merger forms a disky elliptical, with detailed kinematical properties that are in perfect agreement with the observations.
## 2 The merger models
The spiral galaxies are constructed in dynamical equilibrium using the method described by Hernquist (1993a). We use the following system of units: gravitational constant G=1, exponential scale length of the larger disk h=1 and mass of the larger disk $`M_d=1`$. Each galaxy consists of an exponential disk, a spherical, non-rotating bulge with mass $`M_b=1/3`$, a Hernquist density profile (Hernquist (1990)) and a scale length $`r_b=0.2h`$ and a spherical pseudo-isothermal halo with a mass $`M_d=5.8`$, cut-off radius $`r_c=10h`$ and core radius $`\gamma =1h`$.
The N-body simulations were performed using a direct summation code with the special purpose hardware GRAPE (Sugimoto et al. (1990)). The 1:1 merger was calculated adopting in total 400000 particles with each galaxy consisting of 20000 bulge particles, 60000 disk particles and 120000 halo particles. For the 3:1 merger the parameters of the more massive galaxy were as described above. The low-mass galaxy contained 1/3 the mass and number of particles in each component, with a disk scale length of $`h=\sqrt{1/3}`$, as expected from the Tully-Fisher relation. For the gravitational softening we used a value of $`ϵ=0.07`$. In agreement with Walker et al. (1996) we noticed a growing bar mode in the disk for test cases where we evolved galaxies in isolation. This effect is however reduced considerably with respect to previous calculations, due to our choice of twice as many halo particles than disk particles.
For both mergers, the galaxies approach each other on nearly parabolic orbits with an initial separation of 30 length units and a pericenter distance of 2 length units. The inclinations of the two disks relative to the orbit plane are $`t_1=30^{}`$ and $`t_2=30^{}`$ with arguments of pericenter of $`\omega _1=30^{}`$ and $`\omega _2=30^{}`$. These values are most likely for random encounters. In both simulations the merger remnants were allowed to settle into equilibrium for approximately 10 dynamical times after the merger was complete. Then their equilibrium state was analysed.
## 3 Analysis of the equilibrium states
In order to compare the simulation with observations we follow as closely as possible the analysis used by BDM. An artificial image of the remnant is created by binning the central 10 length units into $`128\times 128`$ pixels. This picture is smoothed with a Gaussian filter of standard deviation 1.5 pixels. The isophotes and their deviations from perfect ellipses are then determined using a reduction package kindly provided by Ralf Bender.
Figure 1 shows the radial distribution of $`a_4`$ along the major axis for 200 random projections. There is a clear trend for the 1:1 merger to have boxy (negative $`a_4`$) isophotes while the 3:1 merger has disky deviations (positive $`a_4`$) inside one half-mass radius. Following the definition of BDM for the global properties of observed elliptical galaxies, we determine for every projection $`a_4`$ as the mean value between $`0.25r_e`$ and $`1.0r_e`$ , with $`r_e`$ being the spherical half light radius. In case of a strong peak in the $`a_4`$-distribution with an absolute value that is larger than the absolute mean value, we choose the peak value. In Figure 2 representative isophotal contours of the simulation of a boxy (1:1 merger) and a disky (3:1 merger) merger remnant are shown.
The characteristic ellipticity $`ϵ`$ is defined as the isophotal ellipticity at $`1.5r_e`$. The central velocity dispersion $`\sigma `$ is determined as the average projected velocity dispersion of the stars inside a projected galactocentric distance of $`0.2r_e`$. Finally, we define the characteristic rotational velocity along the major and the minor axis as the projected rotational velocity determined around $`1.5r_e`$ and $`0.5r_e`$, respectively.
Figure 3 shows the characteristic isophotal and kinematical properties of the 1:1 merger (filled circles) and of the 3:1 merger (open circles) for 200 random projections. One can clearly see that the 1:1 merger produces remnants that are anisotropic and boxy with large minor axis rotation whereas the 3:1 merger forms an isotropic and disky elliptical with small minor axis rotation. These results are in excellent agreement with the observations of BDM.
## 4 Discussion and Conclusions
Our calculations demonstrate that the observed dichotomy between boxy and disky ellipticals could originate from variations in the mass ratios of the merger components. In general, the isophotal shapes change with radius. An analysis applying methods and definitions similar to those used for observed ellipticals does however lead to a clear separation of properties between the models of equal and unequal-mass mergers in agreement with the observed structure of boxy and disky ellipticals, respectively. A trend for equal-mass mergers to form preferentially boxy ellipticals has already been noted before by Steinmetz & Buchner (1995), who did however neglect the bulge component. We find that a 3:1 merger is still efficient enough in order to disrupt the disks, leading to spheroidal galaxies with de Vaucouleurs profiles. Our analysis also shows that these systems are rotationally supported with disky isophotes in the region inside one effective radius.
Projection effects do not change the fundamental difference between equal and unequal-mass merger remnants. They do however lead to a large spread in the global parameters. This is in very good agreement with the observed parameter distribution (BDM; Bender et al. (1989)).
In contradiction to the common believe that disky E/S0 galaxies are formed involving dissipative processes like star formation (Bekki & Shioya (1997)) we conclude that pure stellar mergers can in principle explain the observed dichotomy between disky and boxy ellipticals. This result is supported by additional merger simulations with varying mass ratios and orbital parameters which will be discussed in more details in a subsequent paper.
Observations show that disky ellipticals have on average lower luminosities than boxy ellipticals (BDM). Our results would indicate that low-mass elliptical galaxies preferentially formed by unequal-mass mergers of disk-galaxies whereas equal-mass mergers dominated the formation of high-mass ellipticals. This result is puzzeling as there does not exist a convincing argument for why low mass ellipticals should have suffered mainly minor mergers while high mass ones should have evolved mainly through major mergers.
Since dissipative features are observed in all types of elliptical galaxies in can not generally be ruled out, that the gas dynamics and star formation have played an important role for the formation of boxy and disky ellipticals.
We thank Ralf Bender, Hans-Walter Rix and Ralf Klessen for helpful disussions.
|
no-problem/9908/nucl-ex9908016.html
|
ar5iv
|
text
|
# An electronic clock for correlated noise corrections.
## 1 Introduction
Despite dedicated efforts to insure otherwise, a voltage ripple with the same frequency as the main AC power line (60 Hz in the U.S.) but a complicated shape and amplitude may exist on the electrical ground used in an experiment. The cause of the ripple is the existence somewhere in the experiment of at least one electrical path between the experimental ground and other grounds. These “dirty” grounds may be at different mean voltages relative to the experimental ground and may be correlated with the AC power with different phases, line shapes and amplitudes. Generally, experimenters use oscilloscopes and voltmeters to locate unwanted electrical connection(s) in the experiment that tie the experimental ground to dirty grounds. The most egregious paths can generally be located and removed in this manner, and the grounds may hence appear very clean when only the experiment itself is fully powered up. During an actual run, however, the experimental ground may be compromised in ways beyond the control of the experimenters. The contaminating voltage ripple is a noise current which may differ for different detectors in the experiment and for different elements of any given detector, and may be weakly time dependent over periods of hours. The digitization in an Analog-Digital Converter (ADC) of the total charge of electrical pulses from active detector elements is then made versus a baseline ground that itself carries current. This smears ADC data, possibly significantly, and makes less significant all quantities inferred from it. The custom circuit developed to address this problem for Experiment 896 at the BNL-AGS, and its performance for correlated noise corrections, is described here.
Detector channels involving PhotoMultiplier Tubes (PMTs) are particularly sensitive to such ground contaminations, as PMTs are nearly perfect amplifiers. Depending on certain factors such as the widths of ADC gates that are used, even a few millivolts of correlated noise can significantly worsen the ADC resolution both on the pedestal and for hits. The Time of Flight (TOF) system in BNL-AGS Experiment 896 is composed of $``$190 Bicron BC404 plastic scintillator slats of various dimensions each read out by two PMTs (Hamamatsu R2076 or 1398 depending on the slat). The signals from the E896 TOF System have a width of $``$10 ns FWHM and peak at about -300 mV for single hits of relativistic charge $`Z`$$`=`$1 particles. This implies that the integrated charge in a typical PMT pulse for a $`Z`$$`=`$1 hit is on the order of 45 pC into 50 $`\mathrm{\Omega }`$. For this system, the ADC gate width used is 200 ns wide. If the ripple in the ground at the time of a particle hit is 2 mV, the integral within the same 200 ns wide gate is $``$8 pC into 50 $`\mathrm{\Omega }`$, which is $``$10% of the charge measured for hits and time dependent. Depending on how ADC information from a particular detector is used in subsequent analyses, correlated noise may thus have far reaching implications. The resolution on the charge $`|Z|`$/e of particles striking (thin) E896 TOF slats, inferred from the TOF ADC values, can be worsened. Corrections to the TOF timing information based on the ADC information, i.e. “slewing” corrections, would be degraded, worsening the TOF timing performance. There are other detectors in E896 that include the PMT read-out of scintillation or Ĉerenkov light. These detectors include the Multi-Functional Neutron Spectrometer (MUFFINS), the Beam Counters (BCs), the Exit Charge Detector (ECD), and the MuLTiplicity detector (MLT). Correlated noise degrades $`|Z|`$/e measurements and the quality of the slewing corrections for the MUFFINs and BC data, and degrades inferences on the event centrality based on the MLT and ECD data.
This paper is organized as follows. Section 2 describes the apparent symptoms of a correlated noise problem, using as an example those in experimental data collected by the E896 Collaboration during the Spring 1998 run of 11.5 GeV/c/N <sup>197</sup>Au beams at the BNL-AGS. As alluded to above, the E896 experiment is composed of many different subsystems arranged in a mechanically and electrically complicated configuration. There are two large spectroscopic magnets in E896, and many more beam line magnets nearby. As will be shown, correlated noise is rampant in E896 during full beam-on running conditions, although the experimental grounds are clean otherwise. The general solution chosen to combat this problem offline involves custom electronics which are described Section 3. Sections 4 and 5 describe the results obtained when this information was inserted into the E896 data stream during full beam-on running conditions. The summary and conclusions are presented in Section 6.
## 2 The Symptom
In general, an ADC pedestal for a PMT-equipped detector channel has contributions from the intrinsic offsets of the ADC itself, and the PMT’s dark current. For the Hamamatsu R2076 and 1398 PMTs and the LeCroy 1885F ADCs used in the E896 TOF System, these contributions typically result in a pedestal variance on the order of 3-5 ADC channels. A correlation between the pedestals of different detector channels is the unambiguous signature of correlated noise, also known as “common mode noise,” or “ground loops.”
All of the $``$600 ADC pedestals in five different detector systems in E896 are, in some way, highly correlated during full running conditions. An example is shown in Fig. 1 for two PMTs attached to two different E896 TOF slats. The upper right and lower left frames in Fig. 1 indicate that the ADC pedestals for these two channels have variances on the order of 40 channels, which is an order of magnitude larger than would be expected in the absence of correlated noise.
To recover the needed ADC resolution, one must measure the contribution to the ADC values arising solely from the correlated noise component in each experimental event. One method is to add one or more “blackened PMTs” (bPMTs) to the detector, and digitizing these in ADCs exactly as if these bPMTs were attached to active elements. Such blackened PMTs are optically isolated by definition, but as they are mechanically and electrically part of the detector, the ADC values measured are, up to (time-independent) offsets, exactly the correlated noise contribution event by event.
If the correlation between the ADC pedestal from an active channel and that from a bPMT channel is sufficiently strong and single valued (i.e. something like the left frame in Figure 2, the ADC values in active channels can be corrected with some efficiency for the correlated noise contribution using the value of the bPMT in each experiment event. One makes two passes through the experimental data. First one records the correlation between the ADC values for the active channel versus the ADC values from a bPMT, in only those events when the active channel is known not to be struck by a particle. Once this dependence versus one (or more) bPMTs is known for each active channel, in the second pass through the data the bPMT correction can be applied to the active channels in all events, independent of whether the active channel was struck by a particle or not.
However, there is a crucial shortcoming of such bPMT-based approaches. A complete offline correction is possible only if this correlation is indeed strong and single-valued. That is, the blackened PMT must sit on the same local ground as the PMT to be corrected. In electrically simple experiments, i.e. if there is only one or two variants of the (dirty) electrical ground seen by different detector elements, a handful of blackened PMTs could provide all of the information needed for a sufficient correction for correlated noise. In more complicated experiments, i.e. anything on a heavy-ion beam-line, it is possible that innumerable different local grounds exist with different shapes, amplitudes, and phases. An example of a detector channel that cannot efficiently be corrected using a particular bPMT is seen in the right frame of Figure 2.
Shown in Figure 3 is the correlation of ADC pedestals from a number of detectors in E896 versus a particular time recorded for each experimental event and inserted into the main data stream. This time is obtained from a clock, described below, that resets at 60 Hz in active synchronization with the experimental AC power. These plots can be considered oscilloscope traces for ADC pedestals where the horizontal axis is 16.$`\overline{6}`$ ms in total, and the “trigger” is by definition always at exactly the same phase relative to the experimental AC power line. Clear signatures of correlated noise are in fact seen in all PMT-equipped channels in this experiment during full beam-on running conditions. It is also clear that there are many variants of the (dirty) ground local to the different subdetectors in E896, and indeed differing local grounds also exist between different channels in a given subdetector.
If one wanted to combat the noise problem in E896 with blackened PMTs, one would need many of them, and the time and tenacity to find all of the different local grounds at work in the experiment. A more universal and technically simpler correction is possible given the availability of a single number for each experimental event. This number is the time the event occurred as measured within 16.$`\overline{6}`$ ms intervals that are actively synchronized with the experimental AC power, i.e. the time defining the horizontal axis of Fig. 3. The custom electronics that we developed that allows this number to be trivially inserted into any experimental data stream for each experimental event is described in the next Section.
## 3 The Ramp
Given the time that each event occurs, measured relative to 16.$`\overline{6}`$ ms intervals that are actively synchronized with the experimental AC power, one can correct completely for correlated noise offline. In experiments in the U.S., the line frequency is 60 Hz. One thus can use an AC line-synchronized 60 Hz pulser and a latching scaler to provide such clock information to a data acquisition system. We describe here another approach. The present approach was developed primarily because it is more suitable for integration into the E896 data stream than the pulser+latching scaler approach, although the circuit we built is also smaller, considerably cheaper, and more generally portable. The present device can be made part of any experiment that has one spare AC power socket and one spare ADC channel.
The circuit generates a precision negative-voltage sawtooth waveform which resets from V<sub>max</sub> to V<sub>min</sub> in active synchronization with the experimental AC power line. The output is sent to a spare channel of the ADCs used to digitize the detector signals. When an event occurs, the ADC channel connected to the ramp is thus gated at the same absolute time as the detector channel ADCs are gated. Experimental gates are typically 100-200 ns wide, which is a factor of 10<sup>5</sup> shorter than the 16.$`\overline{6}`$ ms period of the AC line. This implies the gating of the ADC channel connected to the ramp digitizes a very thin vertical slice, so the pulse area measured by the ramp ADC is thus effectively just a measurement of the instantaneous value of the ramp voltage at the time the experimental event occurs.
By construction, the ramp voltage so measured is related linearly to the time within actively AC line-synchronized 16.$`\overline{6}`$ ms intervals. The depiction of an ADC pedestal, from detectors anywhere in the experiment, versus the (single) ramp ADC value in the same event thus measures exactly, event by event, the time dependence of the dirty experimental ground local to each detector channel. This allows straightforward and complete offline corrections (examples described below), no matter how many different local grounds exist in the experiment. As there is by definition a ramp value for every experimental event, the correction (described in the next section) is $``$100% efficient in practice.
The circuit that was built in early 1998 and used in the E896 1998 Au run is shown schematically in Figure 4. The circuit is connected to the experimental AC power using a US-standard three prong plug. The AC voltage is shown, attenuated by 100 dB, as trace (a) in this figure. The neutral and phase lines from the AC plug are sent to a power supply, which provides the low voltages used by the circuit, and to a line conditioner. The line conditioner searches for the time at which the experimental AC power crosses zero Volts from below, and generates a bipolar square wave in phase with this time, and hence with the AC power cycle, as seen in trace (b). The bipolar square wave is sent to a “one-shot,” which controls a switch. The leading edges of the one-shot output, shown in trace (c), open the switch and the falling edges close the switch.
A voltage reference and a constant current supply continuously add charge to a 100 nF capacitor (trace d), and by a related amount to the rest of the circuit. This produces the precision voltage ramp, shown in trace (e). When the next 60 Hz cycle begins, i.e. on a leading edge of trace (b), the one shot closes the switch and discharges the capacitor. When the circuit senses the capacitor is fully discharged, which is approximately 3 $`\mu `$s later, the one-shot reopens the switch and the output (trace e) returns to V<sub>min</sub>. Current then begins flowing into the capacitor again, charging it and producing an output voltage that again becomes more negative in precise linearity with the time within the present AC cycle.
The synchronization is active, occurring independently for each cycle of the AC power. Thus, in principle there is nothing in the present circuit that makes it specific to 60Hz line frequencies. The device should work without modification for others, such as are used in experiments abroad.
The discharging of the capacitor implies a “dead time” of the circuit of $``$3 $`\mu `$s/16.$`\overline{6}`$ ms , or about 0.02%. The current out of the constant current supply is generated against a 5.000V voltage reference, which holds the voltage on the capacitor precise to one part in a few thousand. This implies that the circuit provides the time within AC line-synchronized 16.$`\overline{6}`$ ms periods with a resolution of a few microseconds. Such a time resolution is well better than necessary to correct for typical correlated noise levels - typically one needs sufficient statistical certainty in only 50-100 bins of the 60Hz time in order to perform a complete correction (as shown below).
The values of V<sub>min</sub> and V<sub>max</sub> are adjustable via externally mounted potentiometers. This allows the device to be trivially implemented in other experiments, which may use ADC gates of different widths or ADC modules with different full-scale ranges or charge conversion factors. The present circuit uses buffers to produce five copies of the ramp signal, which were each sent to a spare channel in five different TOF ADC modules in the E896 TOF system. In the end only one output channel was actually needed, as the circuit proved to be very stable and its data easy to understand.
The circuit is packaged in a small metal box with external signal connectors and the pots for the V<sub>min</sub> and V<sub>max</sub> control, which is shown in Fig. 5. The cost was about $100 in total.
## 4 The Performance
The event-inclusive distribution of ADC values for the ADC channel connected to the present ramp circuit should, except for spill-structure or other known event trigger time correlations, be essentially flat. Such a distribution for one data-taking run from the E896 Au98 run is shown in Figure 6. The raw ramp ADC distribution is shown in the solid histogram. The dashed histogram depicts the same ramp ADC distribution following the subtraction of the two intrinsic pedestals in the (dual range) LRS 1885F ADCs. This subtraction was done following a measurement of these two intrinsic pedestals for each ADC channel using calibration logic internal to the ADC and a Tcl/Tk script.
Based on the dashed histogram in Figure 6, the times $`t_0`$ at which new 60 Hz cycles begin (and the ramp resets to the voltage $`V_{min}`$) are identified as those events with a pedestal-subtracted ramp ADC value of 0. The time 16.$`\overline{6}`$ ms later is identified as an pedestal-subtracted ramp ADC value of $``$15,000. Thus the event time, $`T_{60Hz}`$, in milliseconds relative to the AC cycle is then computed from the intrinsic pedestal-subtracted ramp ADC values in each event, $`ADC_{ramp}`$, simply from $`T_{60Hz}`$ $`=`$ (16.67 ms)\*$`ADC_{ramp}`$/$`15000`$. This is how the horizontal axis of Figures 3 and 7 was obtained. It should be noted that, as the digitization of the ramp voltage is performed entirely in the counting house, and no PMTs are involved, there is no correlated noise contribution in the ramp ADC values themselves. This allows the ramp ADC values to be applied directly for the noise correction of all of the PMT-equipped detector channels in the experiment.
During the 1998 Au run, hundreds of data runs were recorded to tape in E896, and each was approximately 20 minutes long ($``$200k events/data run). This is well matched to the time scale of roughly hours over which the correlated noise line shapes for a given detector channel slowly drift. Thus, the recording the pedestal profiles for each of the $``$600 ADC pedestals in E896 once per data run allows sufficiently accurate corrections.
A correction based on the present device involves two passes through the experimental data. In the first, the correlated noise “profiles” versus the ramp ADC are collected and saved for each PMT in the experiment. Shown in Figure 7 are such profiles, i.e. the ordinates are average values of an ADC pedestal for all events in a particular 60Hz time bin (abscissa), for the same sampling of E896 PMTs across different detector systems that were shown in Fig. 3.
In any one of these frames, the average ADC values are the same at the extreme values, indicating continuity and the appropriateness of the present 60 Hz clock-based corrections. It is also interesting to notice in Figure 7 (or Figure 3) that frequency modes that are some multiple of 60 Hz are clearly also present. An analysis of the frequency spectra obtained from Fourier transforms of these line-shapes for each of the $``$380 PMTs in the E896 TOF system will be presented in the next section.
In all subsequent passes through the experimental data, the value of the ramp ADC in each experimental event is then used to look up $``$600 values from the $``$600 profile histograms recorded in the previous pass. For each detector channel in this event the appropriate 60Hz-time-dependent pedestal is subtracted from the detector channel’s ADC value, whether or not there is a “hit” in this particular detector channel in this event. The ADC pedestal is then, in software, completely restored to the much-tighter distribution expected in the absence of the correlated noise. The resolution for “hits” may also be improved.
Examples of the dramatic improvement to correlated noise problems in these data is shown in Figures 8 and 9. In either figure, the upper left frame is, for reference, a correlation between a TOF ADC pedestal and a TOF bPMT pedestal. This TOF pedestal is plotted versus the (dual pedestal subtracted) ramp ADC value, i.e. the time in AC line synchronized 16.$`\overline{6}`$ ms intervals, in the lower left frame. The comparison of the two corrections, one based on this bPMT ADC and the other based on the ramp ADC, is shown in the right frame. The raw pedestal is shown as the solid histogram, the pedestal following the bPMT correction is the dotted histogram, while the pedestal following the ramp ADC correction is the dashed histogram.
One observes in the right frame of Figure 8 that the bPMT correction leads to only a partial correction, while in the right frame of Figure 9, the bPMT provides no significant correction at all. This is due to the poor correlations seen in the upper left frames. However, as seen in both of the lower left frames, a far stronger correlation exists between the TOF pedestals and the ramp ADC values, allowing an efficient and complete correction. According to the two right frames of these Figures, the ramp correction restores the ADC pedestals to perfect Gaussian’s with a variance of 3-5 channels - as would be expected in the absence of correlated noise.
## 5 Spectral Analysis
That the time dependence of the value of any E896 ADC pedestal is manifestly periodic at 60 Hz is apparent from the tight correlation, and the continuity at the endpoints, of the profiles shown in Figures 3 and 7. This makes no statement on higher frequency modes though. As the correlated noise is absent when only the experiment itself is fully powered, but is present during full beam-on running, one suspects a contributor to correlated noise at the AGS is the massive power supplies connected to the numerous beam-line magnets. These regulate at 720 Hz in multiples of 60, so one might expect that frequency modes that are multiples of 60 Hz exist in the line-synchronized time dependence of the ADC pedestals. To investigate this possibility, the measured values of the ADC pedestals versus the time from our 60 Hz clock are Fourier analyzed to extract frequency spectra.
For each of the $``$384 channels of the TOF system, the ADC pedestal versus ramp ADC profile, e.g. Figure 7, was measured using the data from a single data run. For each profile, an array of dimension 2<sup>11</sup> was filled with 60 copies of this profile placed end to end in this array. This array thus contains an average PMT pedestal as a function of the time in 2<sup>11</sup> bins that spans a total period of 1 second. The Fourier transform of this array then contains the weights for frequencies measured in Hertz.
A typical frequency spectrum is shown in Figure 10. The weight for the zero-frequency mode, which is related to the (time independent) mean value of the pedestal, is suppressed in this histogram. The dominant frequency modes are identified as the bins in this histogram with the largest absolute values of the Fourier coefficient in this bin. The statistical nature of the mean ADC values versus the 60Hz time used as input to the Fourier transforms results in a finite weight for all possible frequency modes. For the TOF channel shown in this Figure, the dominant mode is 60 Hz, although there is significant weight in modes at 300 Hz and 420 Hz.
For each of the 384 channels of the E896 TOF system and in one run, the first and second largest Fourier coefficients were located for each channel using this procedure. The histogram of the number of TOF detector channels having a particular value of the frequency with the largest(second largest) Fourier coefficient is shown in the left(right) frame of Figure 11. Roughly half of the channels of the E896 TOF system have pedestals that depend on time with the dominant frequency mode being exactly 60 Hz. However, the other half of the TOF channels have ADC pedestals that depend on time dominantly at frequencies that are 3, 5, or 7 times 60 Hz. Even multiples of any significance were not seen in the dominant frequency mode (left frame), although there are channels for which even multiples contribute to the sub-dominant mode (right frame).
It is interesting to note that these is a certain amount of correlation between the dominant frequency mode of the time dependence of the ADC pedestal for a given channel and the location of this channel in the apparatus. In principle this information could be used to provide more clues as to the exact sources of the correlated noise. In practice, however, this is not necessary, as the offline 60 Hz clock-based corrections completely solve the problem anyway.
## 6 Summary
A small, highly portable, and inexpensive custom circuit was developed to provide the information needed for an efficient and complete offline correction for correlated noise in all of the $``$600 PMT-equipped detector channels in BNL-AGS Experiment 896. The circuit generates a precision voltage ramp that resets in active synchronization with the AC power, and this ramp voltage is simply digitized in a spare ADC channel to provide line synchronized clock information for every experimental event. The variances of the experimental ADC pedestals following the “ramp ADC” correction were reduced from, in some cases, tens of channels to the values of 3-5 channels expected in the absence of the correlated noise. Experimentally, the correlated noise profiles for each and every PMT was fundamentally periodic and repeating at 60Hz, yet a considerable number of channels indicate predominant frequencies of odd multiples of 60 Hz, primarily 180 Hz and 300 Hz.
We thank the BNL-AGS E896 Collaboration for the use of preliminary ADC pedestal data from the Au98 run that was discussed above. We gratefully acknowledge funding from the US Department of Energy under Grant No. DE-FG03-96ER40772, as well as helpful comments from H. Takai, S. Costa, C. Nociforo, G.S. Mutchler and E.D. Platner.
|
no-problem/9908/hep-ph9908272.html
|
ar5iv
|
text
|
# “Secret” neutrino interactions11footnote 1 Neutrino Mixing (Torino, March 1999) In honour of Samoil Bilenky’s 70-th Birthday
## 1 “Secret” neutrino interactions?
In the standard model (SM) the interaction between neutrinos is given by the exchange of the $`Z`$-boson and the effective Hamiltonian of the neutrino-neutrino interaction has the form
$$_{SM}^{\nu \nu }=\frac{G_F}{\sqrt{2}}\underset{\mathrm{},\mathrm{}^{}=e,\mu ,\tau }{}(\overline{\nu }_\mathrm{}L\gamma _\alpha \nu _\mathrm{}L)(\overline{\nu }_\mathrm{}^{}L\gamma ^\alpha \nu _\mathrm{}^{}L),$$
(1)
where $`G_F`$ is the Fermi constant. However, it is very difficult to perform direct experimental tests on this interaction.
Many years ago the question was raised whether an additional non-standard four-neutrino interaction exists:
$$_I^{\nu \nu }=F(\overline{\nu }\gamma _\alpha \nu )(\overline{\nu }\gamma ^\alpha \nu )$$
(2)
Such an effective interaction could arise, for instance, from the exchange of a strongly interacting heavy vector boson, $`V_\mu `$, coupled to neutrinos only
$$=g_V(\overline{\nu }_i\gamma _\mu \nu _i)V^\mu ,$$
(3)
when it is considered at energy scales much lower than the vector boson mass, $`m_V`$, e.i. $`q^2m_V^2`$. In this case we have the relation $`F=g_V^2/m_V^2`$.
Interactions mediated by scalars can also be written in the form of $`_I^{\nu \nu }`$ after a Fierz transformation. The flavour structure could be, however, more general.
Obviously, the possible effect of the exchange of a light particle coupled to neutrinos can not be approximated correctly by the contact four-neutrino interaction of the form (2) at $`q^2m_V^2`$. However, in this case the non-standard particles could be produced directly and such models are in general strongly constrained. This is what happens for a class of popular models in which the non-standard interaction among neutrinos is due to the exchange of a massless majoron. In this case very strong bounds on neutrino-majoron coupling constants $`g_{\mathrm{}\mathrm{}}`$ follow from searches for massive neutrinos and neutral particles in $`K\mathrm{}+\mathrm{}`$. One obtains $`g_{ee}^2<1.810^4`$, $`g_{\mu \mu }^2<2.410^4`$. Consequently majoron bremsstrahlung from neutrinos can give only a small contribution to the invisible $`Z`$-width. Majorons with non-vanishing hypercharge could potentially give a large contribution to the invisible $`Z`$-width because the $`Z`$-boson can decay directly to scalars. Therefore, models of this type have already been excluded by LEP data (in the case of the triplet majoron with hypercharge one this contribution is equivalent to the existence of two additional neutrinos and in the case of the doublet majorons a contribution, equivalent to half the contribution of an additional neutrino, arises). Singlet majorons or non-singlet majorons without hypercharge cannot be excluded by LEP data.
We review in this paper, from a historical point of view, the different bounds set on the exotic four-neutrino contact interaction. In section 2 we review the old low-energy bounds. In section 3 we take a look to the limits obtained from the supernova SN1987A. Sections 4 and 6 are devoted to the limits obtained from the invisible decay width of the $`Z`$ gauge boson measured at LEP: section 4 by using the four-neutrino decay at tree level and section 6 by using the one loop contribution of the SNI to the two-neutrino decay. In section 5 we briefly discuss the possibility of gaining some information on the SNI by modification of the lepton spectra in the $`W`$-boson decay due to the process $`W^+\mathrm{}^+\nu _{\mathrm{}}\nu \overline{\nu }`$. In section 7 we present the stringent bounds obtained from nucleosynthesis and finally in section 8 we present our conclusions.
## 2 Low-energy bounds
In the pioneering paper the contributions of $`_I^{\nu \nu }`$ to different low-energy processes were analysed and different bounds were set.
The SNI, contributes to the decays $`\pi ^+e^+\nu _e\overline{\nu }\nu `$ and $`K^+l^+\nu _l\overline{\nu }\nu (l=e,\mu )`$, and could modify the inclusive lepton energy spectra in $`K^+`$ and $`\pi ^+`$ decays, which are dominated by standard decays, $`K^+(\pi ^+)l^+\nu _l`$. From an analysis of these spectra the following bounds on the coupling $`F`$ were obtained
$$|F|10^7G_F,|F|2\times 10^6G_F,$$
(4)
where $`G_F`$ denotes the weak Fermi constant.
Similar bounds were found from the absence of leptons with “wrong” charge in the process $`\nu _\mu +N\mu ^++\nu _\mu +\nu _\mu +X`$.
Later on these bounds were improved in a special experiment (for more recent experiments see also ) searching for the decay $`K^+\mu ^+\nu _\mu \overline{\nu }\nu `$. From the negative result of this experiment the following limit was set:
$$F1.7\times 10^5G_F.$$
(5)
It seemed at that time that four-neutrino interactions could be much stronger than standard model neutral current interactions.
The reason why bounds on the non-standard neutrino interaction coming from low-energy experiments are so loose is evident. The SNI contributes only to the decays with four particles in the final state, and such processes are strongly suppressed by phase space compared with the standard leptonic $`\pi `$ and $`K`$ decays.
## 3 Supernova bounds
The detection of (anti)neutrinos from SN1987A stimulated again the interest on SNI. Using that data some new limits were set.
In particular, the detection of neutrinos from SN1987A requires that the value of the mean free path of neutrinos through the cosmic background particles (CBP) is comparable or greater than the distance to the supernova.
Stable neutrinos should be present today as CBP, therefore, a four-neutrino interaction will contribute to the mean free path of supernova neutrinos and a bound can be set.
If neutrinos have an interaction with neutrinos mediated by heavy vector bosons with mass $`M`$ and coupling $`g`$, a bound on $`g/M`$ was obtained
$$\frac{g}{M}<\frac{12}{MeV},$$
(6)
which can be translated into the following bound on the constant $`F`$:
$$F<10^{13}G_F.$$
(7)
This bound is much worse than the obtained low-energy bounds.
On the other hand, from the estimate of the diffusion time of neutrinos in the supernova and its comparison with the duration of the detected neutrino pulse, an upper bound on neutrino-neutrino cross section $`\sigma _{\nu \nu }`$ can be obtained (see also )
$$\sigma _{\nu \nu }<10^{35}cm^2.$$
(8)
If this cross section arises from a strong four-neutrino interaction one can obtain the following estimate
$$F<10^3G_F,$$
(9)
which is two orders of magnitude better than the best of the low-energy bounds but still allows for a rather strong SNI.
## 4 The decay $`Z\nu \overline{\nu }\nu \overline{\nu }`$
The decays $`\pi (K)\mathrm{}\nu _{\mathrm{}}\nu \overline{\nu }`$ with four particles in the final state are strongly suppressed by phase space in comparison with the usual two-body lepton decays. Decays of much heavier particles, such as gauge bosons, will provide a much larger phase space for multi-neutrino production.
If strong four-neutrino interactions exist, four-neutrino decays
$$Z\nu \overline{\nu }\nu \overline{\nu }$$
(10)
will also contribute to the invisible width of the $`Z`$ gauge boson and therefore the strength of such interaction can be constrained from the precise LEP measurement of $`\mathrm{\Gamma }_{invis}`$.
To be definite we take the Hamiltonian of $`\nu \nu `$ interactions with a general $`V,A`$ form
$$_I^{\nu \nu }=F\underset{\mathrm{},\mathrm{}^{}=e,\mu ,\tau }{}(\overline{\nu }_{\mathrm{}}O_\mathrm{}\alpha \nu _{\mathrm{}})(\overline{\nu }_{\mathrm{}^{}}O_{\mathrm{}^{}}^\alpha \nu _{\mathrm{}^{}}).$$
(11)
Here
$$O_{\mathrm{}}^\alpha =a_{\mathrm{}}\gamma ^\alpha P_L+b_{\mathrm{}}\gamma ^\alpha P_R,$$
(12)
$`P_L`$ and $`P_R`$ are the left and right chirality projectors $`P_L=\frac{1}{2}(1\gamma _5)`$, $`P_R=\frac{1}{2}(1+\gamma _5)`$ and $`F,a_{\mathrm{}},b_{\mathrm{}}`$ are real parameters.
For the total probability of the decay of $`Z`$-bosons into two neutrino pairs (identical and non-identical) we have found the following expression
$$\mathrm{\Gamma }(Z\nu \overline{\nu }\nu \overline{\nu })=\frac{G_F}{\sqrt{2}}m_Z^7F^2\frac{1}{1024\pi ^5}\underset{\mathrm{},\mathrm{}^{}}{}\left\{(a_{\mathrm{}}^2a_{\mathrm{}^{}}^2+a_{\mathrm{}}^4\delta _{\mathrm{}\mathrm{}^{}})C_1+a_{\mathrm{}}^2b_{\mathrm{}^{}}^2C_2\right\},$$
(13)
where
$$C_1=\frac{1651}{486}+\frac{28}{81}\pi ^2,C_2=\frac{259}{486}\frac{4}{81}\pi ^2$$
(14)
The summation runs over the three generations $`\mathrm{},\mathrm{}^{}=e,\mu ,\tau `$.
The decays $`Z\nu \overline{\nu }\nu \overline{\nu }`$ are not sensitive to pure right-handed $`\nu \nu `$ interactions. This is also true for any process involving neutrinos produced through the standard interaction.
Assuming $`e\mu \tau `$ universality in the non-standard $`\nu \nu `$ interaction ($`a_{\mathrm{}}=a,b_{\mathrm{}}=b,\mathrm{}=e,\mu ,\tau `$), one can rewrite the expression for the decay width of $`Z\nu \overline{\nu }\nu \overline{\nu }`$ in the following form :
$$\mathrm{\Gamma }(Z\nu \overline{\nu }\nu \overline{\nu })=\frac{G_F}{\sqrt{2}}m_Z^7\frac{1}{1024\pi ^5}\overline{F}^2(12C_1+9\overline{b}^2C_2).$$
(15)
Here $`\overline{F}^2F^2a^4`$ and the parameter $`\overline{b}^2b^2/a^2`$ characterises the relative contribution of right-handed currents into the $`\nu \nu `$ interaction.
Assuming that only the standard decays $`Z\nu _{\mathrm{}}\overline{\nu }_{\mathrm{}}(\mathrm{}=e,\mu ,\tau )`$ and the decays $`Z\nu \overline{\nu }\nu \overline{\nu }`$ contribute to the invisible width of the $`Z`$-boson $`\mathrm{\Gamma }_{invis}`$ we can obtain a bound on $`\overline{F}`$.
$$\mathrm{\Gamma }_{invis}=3\mathrm{\Gamma }(Z\nu _{\mathrm{}}\overline{\nu }_{\mathrm{}})^{SM}+\mathrm{\Delta }\mathrm{\Gamma }_{invis}.$$
(16)
In our case,
$$\mathrm{\Delta }\mathrm{\Gamma }_{invis}=\mathrm{\Gamma }(Z\nu \overline{\nu }\nu \overline{\nu })$$
(17)
On the other hand this quantity can also be expressed as
$$\mathrm{\Delta }\mathrm{\Gamma }_{invis}=\mathrm{\Gamma }_{invis}3\left(\frac{\mathrm{\Gamma }_{\overline{\nu }\nu }}{\mathrm{\Gamma }_{\overline{l}l}}\right)^{SM}\mathrm{\Gamma }_{\overline{l}l}.$$
(18)
From LEP measurements we have
$$\mathrm{\Gamma }_{invis}=500.1\pm 1.8\mathrm{MeV},\mathrm{\Gamma }_{\overline{l}l}=83.91\pm 0.1\mathrm{MeV},$$
(19)
then, using the ratio of the neutrino and charged leptons partial widths calculated within the SM
$$\left(\frac{\mathrm{\Gamma }_{\overline{\nu }\nu }}{\mathrm{\Gamma }_{\overline{l}l}}\right)^{SM}=1.991\pm 0.001.$$
(20)
we obtain from eq. (18)
$$\mathrm{\Delta }\mathrm{\Gamma }_{invis}1.1\pm 1.9\mathrm{MeV}.$$
(21)
Therefore, from eq. (13), eq. (17) and eq. (21) we obtain for $`\overline{b}^2=1`$ (pure vector or pure axial $`\nu `$$`\nu `$ interaction)
$$\overline{F}<90G_F,$$
(22)
while for pure $`VA`$ couplings ( $`\overline{b}^2=0`$) we obtain
$$\overline{F}<160G_F.$$
(23)
The upper bound on this constant is much lower than earlier existing particle physics bounds and one order of magnitude lower than the estimate obtained from the supernova neutrino diffusion time <sup>2</sup><sup>2</sup>2Note also the important improvement with respect the results obtained in which is due to the updated values of LEP results we use here..
## 5 $`W^+\mathrm{}^+\nu \nu \overline{\nu }`$
Four-neutrino interactions will also give rise to the decay
$$W^+\mathrm{}^+\nu _{\mathrm{}}\nu \overline{\nu }$$
(24)
Using our effective Hamiltonian for the SNI we get the following lepton spectrum in the rest frame of the $`W`$.
$$\frac{d\mathrm{\Gamma }}{dE}=\frac{1}{9}\frac{1}{(2\pi )^5}\frac{G_F}{\sqrt{2}}F^2m_W^2a_{\mathrm{}}^2\left[a_{\mathrm{}}^2+\underset{\mathrm{}^{}}{}(a_{\mathrm{}^{}}^2+b_{\mathrm{}^{}}^2)\right]$$
$$\times \sqrt{E^2m_{\mathrm{}}^2}(E_0E)(3m_WE2E^2m_{\mathrm{}}^2),$$
(25)
where E is the total energy of the charged lepton, $`m_W`$ and $`m_{\mathrm{}}`$ are the masses of the $`W`$-boson and the lepton and $`E_0=(m_W^2+m_{\mathrm{}}^2)/2m_W`$ is the maximum energy of the lepton.
The search for decays of the $`W`$-boson with a single lepton with energy less than $`E_0`$ could give additional information about $`\nu \nu `$ interactions. This analysis could be done using present LEPII data.
## 6 “Secret” neutrino interactions at one loop
All previous bounds are extracted from processes in which the new interaction is the only relevant one and, therefore, observables depend quadratically on the coupling $`F`$. Obviously, if the new interaction enters in loop corrections to a SM process, modifications come through its interference with the SM amplitude and, then, the deviations from the SM predictions will depend linearly on the coupling $`F`$.
For example, $`\nu `$$`\nu `$ interactions will contribute to the decay $`Z\overline{\nu }\nu `$ at the one-loop level and consequently to the invisible width of the $`Z`$-boson.
It is very simple to estimate the order of magnitude of the corresponding contribution of the non-standard interactions at one-loop:
$$\frac{\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu }}{\mathrm{\Gamma }_{\overline{\nu }\nu }}\frac{FM_Z^2}{(4\pi )^2}.$$
(26)
As the invisible width of the $`Z`$-boson is now measured with an accuracy better than 1%, one finds the following bound on the non-standard coupling $`F`$:
$$F(110)G_F.$$
(27)
Therefore, one expects stronger bounds of $`F`$ coming from the one-loop analysis than those which follow from its contribution to the invisible width of the $`Z`$-boson at tree-level.
Although four-fermion interactions are not renormalizable in the “text-book-sense”, still one can obtain some information on their couplings by using them at the one-loop level. This is done by considering additional dimension-six operator(s) which mix with the four-fermion operator under the renormalization group and serve as counterterms to cancel the divergences arising from the use of the four-fermion operator in the loop. The price to be paid is the introduction of more unknown parameters in the analysis which depend on the details of the full theory giving rise to the effective theory. However, if the scale of new physics and the EW scale are well separated, the dominant contributions are the logarithmic terms coming from the running between the two scales. These contributions are quite model independent and can be unambiguously computed in the effective theory.
The partial decay width of the $`Z`$-boson into two neutrinos can be written in the following form:
$$\mathrm{\Gamma }(Z\overline{\nu }\nu )=\mathrm{\Gamma }^{SM}(Z\overline{\nu }\nu )+\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu },$$
(28)
where $`\mathrm{\Gamma }^{SM}(Z\overline{\nu }\nu )`$ is the SM contribution and $`\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu }`$ contains the effects of the non-standard operators.
At lowest order these effects come from the interference of the non-standard amplitude with the SM amplitude and we have
$$\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu }=\mathrm{\Gamma }^{SM}(Z\overline{\nu }\nu )2Re\left\{g_L(M_Z^2)\right\},$$
(29)
where
$$Re\left\{g_L(q^2)\right\}=G_Fq^2\left(c_2+c_1\kappa +c_1\gamma \mathrm{log}(M_Z^2/|q^2|)\right),$$
(30)
gives the vertex $`Z\nu \overline{\nu }`$ induced by the four-fermion interaction at one loop. The constants $`\gamma `$ and $`\kappa `$ and $`c_1`$ are
$$\gamma =\frac{1}{3\pi ^2},\kappa =\frac{17}{36\pi ^2},c_1=\frac{Fa^2}{G_F}=\frac{\overline{F}}{G_F},$$
(31)
and $`c_2`$ is just the finite part of the counterterm needed to absorb the divergences encountered in the loop calculation (see fig. 2c).
Because the standard $`Z\overline{\nu }\nu `$ coupling only involves left-handed neutrinos, and because the lowest order effect of the non-standard interaction comes via its interference with the standard coupling, only interactions of left-handed neutrinos contribute.
Assuming that there are three generations of neutrinos, the non-standard contribution to the invisible width of the $`Z`$-boson is now
$$\mathrm{\Delta }\mathrm{\Gamma }_{invis}=3\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu },$$
(32)
where $`\mathrm{\Delta }\mathrm{\Gamma }_{\overline{\nu }\nu }`$ is given above. Using the limits on $`\mathrm{\Delta }\mathrm{\Gamma }_{invis}`$ obtained in section 4 we get
$$0.03c_2+c_1\kappa 0.007.$$
(33)
If there are no unnatural cancellations between the couplings each of the couplings can be bounded independently of the others and we obtain:
$$\left|c_1\right|=\frac{\overline{F}}{G_F}0.6,\left|c_2\right|0.039,$$
(34)
However, even if there are cancellations at this particular scale ($`q^2=m_Z^2`$) there will not be cancellations at other scales, because the logarithmic dependence of $`g_L(q^2)`$ on $`q^2`$. Therefore, it will still be possible to get some interesting bounds on the coupling $`\overline{F}`$ if additional data obtained at different scales are used (for instance DIS experiments at high energy $`(q^21001000\mathrm{GeV}^2)`$.
Thus, from the analysis of LEP data we can say that contact four-fermion neutrino interactions, involving only left-handed neutrinos, cannot be larger than standard neutral current interactions.
## 7 Primordial Nucleosynthesis
Data on primordial nucleosynthesis offers a limit on the number of the massless degrees of freedom contributing to the early universe expansion for temperatures $`T1`$ MeV.
Three right-handed neutrinos in equilibrium with their left-handed partners at these temperatures are completely excluded. In fact, the three right-handed neutrinos should have decoupled at $`T200`$ MeV.
Four-neutrino interactions of the type considered involving both, left-handed and right-handed neutrinos, could keep right-handed neutrinos in thermal equilibrium through the reactions
$$\overline{\nu }_{Li}+\nu _{Ri}\overline{\nu }_{Rj}+\nu _{Lj}$$
(35)
Requiring that there is decoupling at a temperature $`T`$, that is, enforcing that the interaction rate $`\mathrm{\Gamma }`$ is smaller that the expansion rate of the universe $`H`$, one obtains a rather stringent limit on four-neutrino couplings:
$$F_V<3\times 10^3G_F$$
(36)
for pure vector interactions.
If four-fermion interactions involve only left-handed (or only right-handed) neutrinos the limit does not apply at all.
Other constraints can be obtained from ultra-high energy AGN neutrinos. However they are only relevant for relatively light mediators ($`m_V<0.5`$ GeV).
## 8 Conclusions
We have reviewed, from a historical point, the information obtained on the possible existence of a strong four-fermion contact neutrino interaction.
Bounds on four-neutrino interactions coming from $`K^+`$ and $`\pi ^+`$ decays are very soft and still allow for interactions much stronger than standard model interactions.
From neutrino diffusion time in the supernova one can set better bounds but “Secret” interactions could still be large.
The invisible decay width of the $`Z`$ gauge boson constrains strongly four-neutrino interactions, if they involve left-handed neutrinos. It gives contributions to four-neutrino decay at tree level and to two neutrino decay at one-loop. Using present data we find that SNI interactions involving left-handed neutrinos cannot be larger than neutral current standard model interactions.
“secret” interactions involving both, left-handed and right-handed neutrinos are severely constrained by primordial nucleosynthesis. Their strength must be at least two orders of magnitude smaller than in the standard model, although, these bounds do not apply to pure left-handed or pure right-handed couplings.
Taking all information together we find that there is no room for strong four-fermion contact neutrino interactions unless they involve only right-handed neutrinos.
## Acknowledgements
We would like to acknowledge especially Samoil Bilenky for an enjoyable collaboration and for introducing us to the subject discussed in this paper. Many of the results presented here belong to Samoil. A.S. also likes to thank Wanda Alberico for giving him the opportunity to join Samoil and his friends in this special occasion. Finally A.S. is also indebted to the CERN Theory division, where this work has been written in part, for its hospitality. This work has been funded, in part, by CICYT under the Grant AEN-96-1718, by DGESIC under the Grant PB97-1261 and by the DGEUI of the Generalitat Valenciana under the Grant GV98-01-80.
|
no-problem/9908/cond-mat9908440.html
|
ar5iv
|
text
|
# Berry phase and persistent current in disordered mesoscopic rings
## Abstract
A novel quantum interference effect in disordered quasi-one-dimensional rings in the inhomogeneous magnetic field is reported. We calculate the canonical disorder averaged persistent current using the diagrammatic perturbation theory. It is shown that within the adiabatic regime the average current oscillates as a function of the geometric flux which is related to the Berry phase and the period becomes half the value of the case of a single one-dimensional ring. We also discuss the magnetic dephasing effect on the averaged current.
Since the discovery of the Berry phase, there has been much interest in the study of topological effects in the fields of quantum mechanics and condensed matter physics. The typical example to illustrate the Berry phase is the Aharonov-Bohm (AB) effect in the mesoscopic ring, where the relative phase would accumulate on the wave function of a charged particle due to the presence of a electromagnetic gauge potential. Similarly, when a quantum spin follows adiabatically a magnetic field that rotates slowly in time, the spin wave function acquires an additional geometric phase (Berry phase) besides the usual electromagnetic phase in the static magnetic fields.
Loss $`et`$ $`al`$ studied the persistent current in an $`ideal`$ one-dimensional ring in the presence of a static inhomogeneous magnetic field by examining the coupling between spin and orbital motion through the Zeeman interaction. Using the imaginary time path integral method, they showed that the spin wave function accumulates the Berry phase when the spin of an electron traversing an AB ring adiabatically follows a textured inhomogeneous magnetic filed with a tilt angle and this phase leads to persistent equilibrium charge current.
In this paper we investigate the persistent charge current of the quasi-one-dimensional $`disordered`$ rings (see the inset in Fig. 2) and compare it with the results of Loss $`et`$ $`al`$ We compute the disorder averaged persistent current with the help of the diagrammatic technique and find that the average current is given by the sum of the term which oscillates as a function of the Berry phase and the term which does not depend on the the Berry phase. We show that the latter term is negligibly small compared with the former term in the adiabatic regime, so that we will be able to clearly observe the Berry phase dependent persistent current experimentally.
We begin by considering a quasi-one-dimensional ring of circumference $`L_x=2\pi r`$ and volume $`V=L_xL_yL_z`$. The ring is embedded in an static inhomogeneous magnetic field $`𝐁`$. For a spin-$`1/2`$ electrons of mass $`m`$ and charge $`e`$, the system may be described by the Hamiltonian
$`={\displaystyle \frac{1}{2m}}\left[𝐩{\displaystyle \frac{e}{c}}𝐀^{em}(𝐫)\right]^2+u(𝐫){\displaystyle \frac{1}{2}}g\mu _B𝐁(𝐫)\mathbf{}𝝈,`$ (1)
where $`𝐩`$, $`𝐫`$, $`g`$, $`\mu _B`$, and $`\mathrm{}𝝈/2`$ are the momentum, position, $`g`$ factor, Bohr magneton, and spin, respectively. The operator $`u(𝐫)`$ represents the spin-independent random impurity potential and, $`𝐀^{em}`$ is the electromagnetic gauge potential, with $`𝐁=\times 𝐀^{em}`$ relating it to the magnetic fields. In the following we specialize in the case of inhomogeneous magnetic fields with constant magnitude $`B`$, and we have parametrized $`𝐁`$ in terms of the spherical polar angles $`\chi `$ and $`\eta `$ so that it has Cartesian components $`B(\mathrm{sin}\chi (𝐫)\mathrm{cos}\eta (𝐫),\mathrm{sin}\chi (𝐫)\mathrm{sin}\eta (𝐫),\mathrm{cos}\chi (𝐫))`$, with the angles $`\chi `$ and $`\eta `$ being smooth functions of position.
Using the Green’s function, the thermal average of the persistent current is given by
$$I=\frac{d\epsilon }{2\pi i}\underset{𝐤}{}\underset{\alpha =\pm 1}{}I_xf(\epsilon )\left[G_{\alpha ,\alpha }^A(𝐤,𝐤,\epsilon )G_{\alpha ,\alpha }^R(𝐤,𝐤,\epsilon )\right],$$
(2)
where $`\alpha =\pm 1`$ is the spin index and $`f(\epsilon )`$ is the Fermi-Dirac distribution function. The current vertex is given by $`I_x=e\mathrm{}k_x/mL`$, where $`k_x`$ is the $`x`$ component of the wave number vector. Moreover $`G^{R(A)}`$ is the retarded (advanced) Green’s function and it denotes the exact Green’s function before impurity averaging. To determine the canonical disorder-averaged persistent current, we follow the procedure described in Refs. \[6$``$8\] The chemical potential is set to $`\mu =\mu _0+\delta \mu `$ and we expand Eq. (2) to first order in $`\delta \mu `$. Then we obtain the averaged persistent current at a fixed particle number as follows:
$$I(\mathrm{\Phi }^{em})\frac{\mathrm{\Delta }}{2}\frac{}{\mathrm{\Phi }^{em}}\left\{\underset{\alpha =\pm 1}{}\delta N_\alpha (\mu _0,\mathrm{\Phi }^{em})\right\}^2,$$
(3)
where $`\mathrm{\Phi }^{em}`$ is the electromagnetic flux and $`\mathrm{\Delta }`$ is the mean level spacing. The fluctuation of the particle number $`(_\alpha \delta N_\alpha )^2`$ is related to the two-point correlator of the density of states $`\rho _\alpha `$
$$\left\{\underset{\alpha =\pm 1}{}\delta N_\alpha (\mu _0,\mathrm{\Phi }^{em})\right\}^2=V^2_{\mathrm{}}^{\mathrm{}}𝑑\epsilon _1_{\mathrm{}}^{\mathrm{}}𝑑\epsilon _2f(\epsilon _1)f(\epsilon _2)\underset{\alpha ,\alpha ^{}}{}K_{\alpha ,\alpha ^{}}(\epsilon _1,\epsilon _2),$$
(4)
where
$`K_{\alpha ,\alpha ^{}}(\epsilon _1,\epsilon _2)`$ $`=`$ $`\rho _\alpha (\epsilon _1)\rho _\alpha ^{}(\epsilon _2)\rho _\alpha (\epsilon _1)\rho _\alpha ^{}(\epsilon _2)`$ (5)
$``$ $`{\displaystyle \frac{1}{2\pi ^2V^2}}{\displaystyle 𝑑𝐫_1𝑑𝐫_2\text{Re}\left[G_{\alpha ,\alpha }^R(𝐫_1,𝐫_1;\epsilon _1)G_{\alpha ^{},\alpha ^{}}^A(𝐫_2,𝐫_2;\epsilon _2)G_{\alpha ,\alpha }^R(𝐫_1,𝐫_1;\epsilon _1)G_{\alpha ^{},\alpha ^{}}^A(𝐫_2,𝐫_2;\epsilon _2)\right]}.`$ (6)
In order to calculate $`K_{\alpha ,\alpha ^{}}`$, we use the diagrammatic perturbation method which is a powerful tool in the theory of disordered system. In the language of diagrammatics, the dominant contribution to $`K_{\alpha ,\alpha ^{}}`$ arises from the exchange of the two Cooperon (particle-particle) ladders between two closed loops (see Fig. 1). On the other hand, the Diffuson process gives no contribution to the average current because it has no dependence of the magnetic flux $`\mathrm{\Phi }^{em}`$. Thus we need only the Cooperon process in order to calculate the average persistent current. As usual, we are now interested only in the field dependence resulting from interfering paths. Therefore we ignore the less important field dependence of the averaged single Green’s function. The two-point correlator then translates into the following expression:
$`K_{\alpha ,\alpha ^{}}(\epsilon _1,\epsilon _2)={\displaystyle \frac{1}{2\pi ^2V^2\mathrm{}^2}}\text{Re}{\displaystyle 𝑑𝐱_1𝑑𝐱_2𝒞_{\alpha ,\alpha ^{}}(𝐱_1,𝐱_2;\epsilon _1\epsilon _2)𝒞_{\alpha ,\alpha ^{}}(𝐱_2,𝐱_1;\epsilon _1\epsilon _2)}.`$ (7)
In this equation we have used the definition of the particle-particle pair propagator
$`𝒞_{\alpha ,\alpha ^{}}(𝐱_1,𝐱_2;\epsilon _1\epsilon _2)={\displaystyle \frac{2\pi \rho (0)}{\mathrm{}}}G_{\alpha ,\alpha }^R(𝐱_2,𝐱_1;\epsilon _1)G_{\alpha ^{},\alpha ^{}}^A(𝐱_2,𝐱_1;\epsilon _2),`$ (8)
where $`\rho (0)`$ is the density of states (per unit volume and spin) at the Fermi surface. Following Ref. we have evaluated the pair propagator $`𝒞_{\alpha ,\alpha ^{}}`$ by using the quasi-classical Feynman path-integral method and obtained
$`𝒞_{\alpha ,\alpha ^{}}(𝐱_1,𝐱_2;t_1,t_2)=\theta (t_2t_1){\displaystyle _{𝐑(t_2)=𝐱_2}^{𝐑(t_1)=𝐱_1}}𝒟𝐑\mathrm{exp}\left(i𝒮_{\alpha ,\alpha ^{}}^{eff}\left[𝐑\right]\right),`$ (9)
where the effective action is given by
$`𝒮_{\alpha ,\alpha ^{}}^{eff}\left[𝐑\right]={\displaystyle \frac{i}{4D}}{\displaystyle _{t_2}^{t_1}}𝑑t\left|\dot{𝐑}\right|^2{\displaystyle \frac{2e}{\mathrm{}}}{\displaystyle _{t_2}^{t_1}}𝑑t\dot{𝐑}𝐀^{em}\left(𝐑(t)\right){\displaystyle \frac{g\mu _BB}{2\mathrm{}}}(t_2t_1)(\alpha \alpha ^{})\gamma _\alpha \left[𝐑\right]\gamma _\alpha ^{}\left[𝐑\right].`$ (10)
Here the Berry phase $`\gamma _\alpha \left[𝐑\right]`$ which arises from adiabatic approximation for the spin (dynamical Zeeman) propagator is of the form
$`\gamma _\alpha \left[𝐑\right]={\displaystyle _{t_1}^{t_2}}𝑑t\dot{𝐑}𝐀_\alpha ^g(𝐑),`$ (11)
where $`𝐀_\alpha ^g(𝐑)`$ is the spin dependent geometric gauge potential and is given by $`𝐀_\alpha ^g(𝐑)=(\alpha /2)\eta \left[\mathrm{cos}\chi 1\right]`$. It should be note that the expression Eq. (9) is only valid in the adiabatic regime in which the spin of the electron adiabatically follows the local direction of the non-uniform magnetic field. This adiabaticity requires that the precession frequency $`\omega _B=g\mu _BB/2\mathrm{}`$ is large compared to the reciprocal of the diffusion time $`\tau _d=L_x^2/D`$ ($`D`$ is the diffusion constant) around the ring, i.e., $`\omega _B\tau _d1`$, or equivalently $`BB_c2E_{Th}/g\mu _B`$
The path-integral representation for the pair propagator can be transformed into a differential equation,
$`\left[{\displaystyle \frac{}{t^{}}}+D\left\{i{\displaystyle \frac{}{𝐱^{}}}𝐀_\alpha \left(𝐱^{}\right)𝐀_\alpha ^{}\left(𝐱^{}\right)\right\}^2i\omega _B(\alpha \alpha ^{})\right]𝒞_{\alpha ,\alpha ^{}}(𝐱_1,𝐱_2;t_1,t_2)=\delta \left(𝐱_1𝐱_2\right)\delta \left(t_1t_2\right),`$ (12)
which is characterized by the presence of the spin-dependent gauge potential $`𝐀_\alpha (e/h)𝐀^{em}+𝐀_\alpha ^g`$. We now evaluated the pair propagator for the case of the metallic rings, having height $`L_z`$ and thickness $`L_y`$. The magnetic field was assumed to have constant magnitude, and not to vary appreciably either radially across the wall or parallel to the ring axis. Thus the gauge potential $`𝐀_\alpha `$ can be replaced by its tangential component $`A_\alpha ^\varphi `$ ($`\varphi `$ is the azimuthal angle). This can be expressed as $`2\pi rA_\alpha ^\varphi =\mathrm{\Phi }_\alpha =\mathrm{\Phi }^{em}/\mathrm{\Phi }_0+\alpha \mathrm{\Phi }^g`$, where $`\mathrm{\Phi }^{em}`$ is the electromagnetic flux through the area $`\pi r^2`$ and the geometric flux is given by
$`\mathrm{\Phi }^g={\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}𝑑\varphi \left[\mathrm{cos}\chi (\varphi )1\right]_\varphi \eta (\varphi ).`$ (13)
The geometric flux take a simple form in the case of a cylindrically symmetric texture ($`\eta =\varphi `$), i.e., $`\mathrm{\Phi }^g=(\mathrm{cos}\chi 1)/2`$, where $`\chi `$ is the constant tilt angle away from the cylinder axis. According to the standard method, we can show that the pair propagator $`𝒞_{\alpha ,\alpha ^{}}`$ is explicitly given by
$`𝒞_{\alpha ,\alpha ^{}}(𝐱_1,𝐱_2;t_1,t_2)`$ $`=`$ $`{\displaystyle \frac{1}{V}}{\displaystyle \underset{𝐤}{}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}𝒞_{\alpha ,\alpha ^{}}(𝐤,\omega )e^{i𝐤(𝐱_1𝐱_2)i\omega (t_1t_2)},`$ (14)
$`𝒞_{\alpha ,\alpha ^{}}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{1}{i(\omega \varpi _{\alpha ,\alpha ^{}})+{\displaystyle \frac{1}{\tau _\phi }}+D𝐤_{}^2+{\displaystyle \frac{D}{r^2}}\left[n_x(\mathrm{\Phi }_\alpha +\mathrm{\Phi }_\alpha ^{})\right]^2}},`$ (15)
where $`𝐤=(2\pi n_x/L_x,𝐤_{})`$, $`𝐤_{}=(2\pi n_y/L_y,2\pi n_z/L_z)`$ ($`n_x,n_y,n_z`$ are integer) and $`\varpi _{\alpha ,\alpha ^{}}\omega _B(\alpha \alpha ^{})`$. For convenience we have incorporated the effect of dephasing $`1/\tau _\phi =D/L_\phi ^2`$ ($`L_\phi `$ is the phase coherence length) directly into the pair propagator. Note that for diagonal component of the pair propagator $`𝒞_{\alpha ,\alpha }`$, the Zeeman terms cancel, i.e., $`\varpi _{\alpha ,\alpha }=0`$.
In the following we shall calculate the diagonal part ($`\alpha ^{}=\alpha `$) and the off-diagonal part ($`\alpha ^{}=\alpha `$) of the average current separately, namely,
$`I(\mathrm{\Phi }^{em})I^D(\mathrm{\Phi }^{em})+I^{OD}(\mathrm{\Phi }^{em}).`$ (16)
The diagonal part is attributed to the interference between same spin states. If the thickness and the height of the ring is small compared with the length $`L_\phi `$, then the integration over $`𝐤_{}`$ should be replaced by a summation with only the term corresponding to $`𝐤_{}=\mathrm{𝟎}`$ retained. Substituting the Fourier transform of Eq. (15) into Eq. (7) and using the Matsubara method, therefore, we obtain the diagonal part of the average current
$`I^D(\mathrm{\Phi }^{em})={\displaystyle \frac{\mathrm{\Delta }}{2\pi \beta }}{\displaystyle \underset{\alpha =\pm 1}{}}\text{Re}{\displaystyle \frac{}{\mathrm{\Phi }^{em}}}{\displaystyle \underset{\nu _{\mathrm{}}>0}{}}{\displaystyle \underset{n_x=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\nu _{\mathrm{}}}{\left\{\nu _{\mathrm{}}+{\displaystyle \frac{\mathrm{}}{\tau _\phi }}+4\pi ^2E_{Th}\left(n_x+2\mathrm{\Phi }_\alpha \right)^2\right\}^2}},`$ (17)
where $`\beta =1/k_BT`$ and $`E_{Th}=\mathrm{}D/L_x^2`$ is the Thouless energy and $`\nu _{\mathrm{}}=2\pi \mathrm{}/\beta `$ ($`\mathrm{}`$ is an integer) is the boson Matsubara frequency, respectively. In the limit of zero temperature, the Matsubara sum turns into an integral $`_\nu 2\pi /\beta 𝑑\nu `$, which is easily evaluated. This yields for the diagonal part of the averaged persistent current at $`T=0`$ as
$`I^D(\mathrm{\Phi }^{em})={\displaystyle \frac{I_0}{M}}{\displaystyle \underset{\alpha =\pm 1}{}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\mathrm{exp}\left(n{\displaystyle \frac{L_x}{L_\phi ^B}}\right)\mathrm{sin}\left(4\pi n\mathrm{\Phi }_\alpha \right).`$ (18)
$`I_0=e\upsilon _F/L_x`$ is the current carried by a single electron state in an ideal one-dimensional ring and $`M=k_F^2V/L_x`$ is the effective channel number, where $`\upsilon _F`$ is the Fermi velocity and $`k_F=m\upsilon _F/\mathrm{}`$ is the Fermi wave number. In Eq. (18) the magnetic dephasing length $`L_\phi ^B`$ is given by $`1/(L_\phi ^B)^2=1/(L_\phi )^2+1/(L^B)^2`$ and $`L^B=\sqrt{3}\mathrm{\Phi }_0/(2\pi L_y\left|B_z\right|)`$ accounts for the dephasing effect of the magnetic field penetrating into the sample region, where $`B_z`$ is the $`z`$ component of $`𝐁`$. Therefore we will observe the persistent current oscillation with somewhat reduced amplitude. It is important to note that unlike the case of the one dimensional ring the inhomogeneous magnetic fields leads to the dephasing effect to the persistent current.
For the off-diagonal part of the persistent current which results from the interference between opposite spin states, same procedure as above yields
$`I^{OD}(\mathrm{\Phi }^{em})`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }}{\pi \beta }}\text{Re}{\displaystyle \frac{}{\mathrm{\Phi }^{em}}}{\displaystyle \underset{\nu _{\mathrm{}}>0}{}}{\displaystyle \underset{n_x=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\nu _{\mathrm{}}}{\left\{\nu _{\mathrm{}}+{\displaystyle \frac{\mathrm{}}{\tau _\phi }}+4\pi ^2E_{Th}\left(n_x+2{\displaystyle \frac{\mathrm{\Phi }^{em}}{\mathrm{\Phi }_0}}\right)^2\right\}^2+\left(g\mu _BB\right)^2}}`$ (19)
$`=`$ $`{\displaystyle \frac{I_0}{M}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}i_n\mathrm{sin}\left(4\pi n{\displaystyle \frac{\mathrm{\Phi }^{em}}{\mathrm{\Phi }_0}}\right),`$ (20)
with
$`i_n`$ $`=`$ $`2n\xi {\displaystyle _0^{\sqrt{\sqrt{\mathrm{\Gamma }^2+1}\mathrm{\Gamma }}}}𝑑y\left({\displaystyle \frac{1y^4}{2y^2}}\mathrm{\Gamma }\right)\left[\mathrm{cos}(n\xi y)+{\displaystyle \frac{1}{y^2}}\mathrm{sin}(n\xi y)\right]\mathrm{exp}\left(n{\displaystyle \frac{\xi }{y}}\right),`$ (21)
where $`\mathrm{\Gamma }D/[2\omega _B(L_\phi ^B)^2]`$ and $`\xi \sqrt{\mathrm{}\omega _B/E_{Th}}`$. Note that the off-diagonal term does not depend on the Berry phase. In contrast to the diagonal term, we can expect that this term gives negligibly small contributions to the average current in the adiabatic regime because the Cooperon pole has a non-zero Zeeman term. Figure 2 shows the $`B`$ dependence of $`I^D`$ and $`I^{OD}`$ of quasi-one-dimensional rings ($`L_x=11\mu `$m and $`L_y=2.0\mu `$m) in the case of the cylindrically symmetric texture. As expected above, $`I^{OD}`$ is much smaller than $`I^D`$ in the adiabatic regime, $`BB_c10`$Gauss. Therefore the average persistent current is given by
$`I(\mathrm{\Phi }^{em})I^D(\mathrm{\Phi }^{em})={\displaystyle \frac{I_0}{2M}}{\displaystyle \underset{\alpha =\pm 1}{}}{\displaystyle \frac{\mathrm{sin}\left(4\pi \mathrm{\Phi }_\alpha \right)}{\mathrm{cosh}\left({\displaystyle \frac{L_x}{L_\phi ^B}}\right)\mathrm{cos}\left(4\pi \mathrm{\Phi }_\alpha \right)}}`$ (22)
which is a periodic function of $`\mathrm{\Phi }^{em}`$ and $`\mathrm{\Phi }^g`$ with period $`\mathrm{\Phi }_0/2`$ and $`1/2`$, respectively. These periods are half the value of the case of a single one-dimensional ring. As the half flux periodicity of the electromagnetic flux in disordered rings, that of geometric flux is ascribed to the ensemble averaging for fixed particle number: averaging eliminates the first Fourier component of the current although the second components which results from the interference between time-reversed trajectories survives. If the magnetic field is homogeneous then the Berry phase vanishes, i.e., $`\mathrm{\Phi }^g=0`$. In this case the expression Eq. (22) is reduced to the well-known result.
In summary, we have investigated the persistent current of disordered rings in the inhomogeneous magnetic field. Halving of the geometric flux period with respect to the single ballistic one-dimensional ring is obtained in the canonical disorder-ensemble average. We have also showed that the existence of magnetic field which penetrates the sample causes the dephasing effect.
Finally we point out that the phenomena investigated in this paper will also appear in doubly-connected chaotic billiards. We expect that these systems will open a new area to explore the effect of $`chaos`$ on the Berry phase.
The author wishes to thank B. A. Friedman for helpful comments.
.
|
no-problem/9908/cond-mat9908035.html
|
ar5iv
|
text
|
# The Fermi surfaces of Metallic Alloys and the Oscillatory Magnetic Coupling between Magnetic Layers separated by such Alloy Spacers
## 1 Introduction
Many random alloys such as Cu<sub>(1-x)</sub>Ni<sub>x</sub>, Cu<sub>(1-x)</sub>Au<sub>x</sub> are metals and therefore have Fermi surfaces in a well defined sense. Moreover these Fermi surfaces determine many of the properties of these scientifically interesting and technologically important class of materials. Thus, it would be useful to know what these Fermi surfaces are like and how they evolve with changing concentration. Unfortunately the classic probes of the Fermi surface such as the measurement of the de Haas van Alphen (dHvA) oscillations work only if the quasi-particles can complete a cycle along the Landau orbit between two scattering events associated by deviations of the crystal potential from periodicity. As it happens this physical requirement of long quasi-particle life times translates into very small, $``$ppm, concentration of impurities and hence no dHvA signal is expected for the concentrated alloys of interest. This leaves, until recently, two dimensional Angular Correlation of (Positron) Annihilation Radiation (2d ACAR) and Compton Scattering (SC) studies, which do not require long quasi-particle life times, as the only source of reasonably direct quantitative information about the Fermi Surfaces of Random Alloys. Our aim here is to argue that measurements of the oscillatory coupling between magnetic layers across random alloy spacers can also provide such information.
In short, what one measures is the exchange coupling $`J_{12}(L)`$ between magnetic layers 1 and 2 separated by a non magnetic spacer layer of thickness $`L`$ made of a random metallic alloy. A schematic picture of the experiment configuration is shown in Fig. 1 and the exchange interaction $`J_{12}`$ is defined in terms of the magnetic interaction energy $`\delta E_{12}(L)`$:
$$\delta E_{12}(L)=J_{12}(L)𝐌_1𝐌_2$$
(1)
where $`𝐌_1`$, $`𝐌_2`$ are the average magnetization of magnetic layers 1 and 2 respectively. As was discoverer by Parkin et al, and has been observed for a vast variety of systems, $`J_{12}(L)`$ oscillates between being Ferromagnetic, $`J_{12}<0`$ and Antiferromagnetic, $`J_{12}>0`$ as a function of the separation $`L`$. In fact most experiments seem to be consistent with the formula
$$J_{12}(L)=\frac{1}{L^2}\underset{\nu }{}A_\nu \mathrm{cos}(Q_\nu L+\varphi _\nu )e^{L/\mathrm{\Lambda }_\nu }$$
(2)
where each contribution $`\nu =1,2,\mathrm{}`$ is characterized by the period $`P_\nu =2\pi /Q_\nu `$, amplitude $`A_\nu `$, phase $`\varphi _\nu `$ and coherence length $`\mathrm{\Lambda }_\nu `$ which is infinite for pure metal spacer but is finite if the spacer is a random metallic alloy. Our discussion will focus on the astonishing fact that $`Q_\nu `$ is a quantitative measure of a geometrical feature of the Fermi surface of the infinite (bulk) spacer metal. In fact, most theories predict the form in eq. (1) asymptotically for large $`L`$ and $`Q_\nu `$ turns out to be an extremal caliper vector, connecting two opposite points on the Fermi surface, in the direction perpendicular to the plane of the magnetic layers, that is to say in the growth direction of the multilayer system. This result is well confirmed for a large number of pure metal spacers. In the next two sections we review the agreement between the predicted and measured evolution of $`Q_\nu `$ with concentration $`x`$ for random metallic alloys. In section 4 we will discuss the exponential damping factor and in section 5 we present an asymptotic theory for the amplitudes and the phases $`Q_\nu `$.
The interest in the above interlayer magnetic coupling has arisen in the wake of its discovery, because of its connection with the technologically very important Giant Magneto-resistance (GMR) phenomenon. From this point of view the problem is largely solved. The physical mechanism understood to be the planar defect analogue of the RKKY interaction between point like magnetic defects in metals, enhanced by confinement. In this contribution we wish to emphasize an other aspect of the problem. Namely, we shall explore the possibility of using the oscillatory coupling phenomenon as a new probe of the Fermi surface in transition metal alloys. Clearly, compared with 2d-ACAR the measurement of these oscillations are simple and cheap and hence the prospects of such project are bright. However, before the full power of the method can be assessed both the experimental technique and the theoretical framework used to interpret the data will require further scrutiny. In what follows we will present a few initial steps in the direction of the latter.
## 2 Evolution of the Fermi surface of Cu<sub>(1-x)</sub>Ni<sub>x</sub> alloys with concentration.
The FCC Cu<sub>(1-x)</sub>Ni<sub>x</sub> alloys are one of the best known examples for which the rigid band model completely misconstrues the nature of the changes in the electronic structure on alloying. Thus a direct experimental study of the Fermi surface is of fundamental interest from the point of view of the electronic structure of metallic alloys.
Oscillations for this alloy systems have been observed in three separate experiments. Parkin et al and Bobo et al studied a Co/Cu<sub>(1-x)</sub>Ni<sub>x</sub>/Co system with (111) growth direction, while Okunoet al studied the same system for the (110) orientation. The relatively long oscillation period ($``$ 10Å) observed in all of these experiments is believed to correspond to neck caliper vectors of the Cu-like Fermi surface of Cu<sub>(1-x)</sub>Ni<sub>x</sub> alloys for $`x0.4`$. In particular in the case of (110) direction the caliper vector is the diameter of the neck itself, while in the case of (111) orientation it spans the neck in an angle of 19.47<sup>o</sup> with respect to the neck plane. In fig. 2 all the extremal vectors for the (100), (110) and (111) directions are shown. The ones we already mentioned for the (110) and (111) directions are the $`Q_{(110)}^{(2)}`$ and $`Q_{(111)}^{(1)}`$ respectively. In the refs. we have calculated the concentration dependence of these extremal vectors using the KKR-CPA electronic structure method and have compared the predicted oscillation periods with the experimental ones. We present that result also in fig. 3 for both the (110) and (111) orientations.
The decrease with concentration of the neck diameter, resulting in an increase of the oscillation period, is in qualitative agreement with the Rigid Band Model. Indeed, that decrease comes from the fact that an electron-like neck of the Fermi surface such as the one we are looking for is expected to shrink when electrons are removed from the system (for instance by increasing the Ni concentration) but the calculated as well as the observed shrinkage is more gentle than the predicted from the Rigid Band Model. The excellent agreement between our calculation and the experiment as illustrated in fig. 3 is one more striking example of the success of the CPA theory in binary alloy systems. The relatively simple Fermi surface of the Cu<sub>(1-x)</sub>Ni<sub>x</sub> binary alloy system ($`x<0.5`$) makes it easy to examine whether the evolution of the Fermi surface with alloying is in agreement with OMC measurements, and as is illustrated in fig. 3 that agreement is excellent. In the next section a binary alloy system with much more complicated Fermi surface is examined, namely the Cr<sub>1-x</sub>V<sub>x</sub> binary alloy.
## 3 Which piece of the Fermi surface drives the long period oscillations across Cr<sub>(1-x)</sub>V<sub>x</sub> spacers?
Although the long period oscillation for Fe/Cr/Fe ($``$18 Å) was the first example of Oscillatory coupling discovered, until recently the origin of that long period oscillation was the subject of an open debate. In the literature, the most popular candidate parts of the Fermi surface for being relevant are the electron-like lens of the Fermi surface and the N point centered hole-like pocket. A further point of interest is that the long period oscillation appears to be unaffected by the orientation of the specimen, at least for the (100), (110) and (211) directions, leading to the conclusion that the coupling comes from a fairly isotropic region of the Fermi surface. The extra complication that makes it extremely difficult to associate the oscillation period with an area of the Fermi surface comes from the fact that the Cr Fermi surface is fairly complicated with high degree of nesting features arising from the d-states. Its rather difficult even to enumerate all the extremal vectors of the Fermi surface and of course the procedure of just comparing the experimental periods with the modes of the Fourier transform of a total energy calculation totally fails in this case. Of course the panacea would be to calculate the amplitudes of the individual oscillatory terms and reveal which terms are dominant. Such calculations have been done using semi-empirical Tight Binding methods, and although the size of the associated period is significantly smaller than the observed, they conclude that the origin of the oscillation is the N hole-like pocket for both the (100) and (110) orientations. Unfortunately, other authors are drawing different conclusions.
In ref. we suggested that the evolution of the Fermi surface with alloying in the Cr spacer could give conclusive answer to the debate. Indeed, there are experiments on the OMC across Cr<sub>(1-x)</sub>V<sub>x</sub> as well as Cr<sub>(1-x)</sub>Mn<sub>x</sub> spacers for poly-crystalline samples with (110) predominant orientation. The idea is that if the origin of the oscillation is an electron-pocket then that pocket should shrink as V is added enhancing the size of the oscillation period. If on the other hand, the origin is a hole-like pocket the period should decrease when V is added. The opposite apply in the case of alloying with Mn. What the experiment shows is a monotonic decreases of the period with V concentration and increase with Mn concentration, which is consistent with the source of oscillation being a hole-like pocket. Of course the RBM is not quantitatively correct in general, but it serves as a good qualitative picture. Of course, our calculation is based on the KKR-CPA method and someone expects the agreement to be beyond the qualitative level. In fig. 4 the experimental period is shown as a function of the V concentration along with the theoretical periods predicted from various extremal vectors of the alloy Fermi surface. As we see the period predicted from the N-hole-like pocket is the only one which agrees quantitatively with experiment. Thus, our results strongly suggest that the source of the oscillations for both pure Cr as well as Cr<sub>(1-x)</sub>V<sub>x</sub> spacers is the N-hole-like pocket of the Fermi surface. For this particular case the rigid band model seem to agree quantitatively with the more accurate CPA result, as has been shown by Koelling who used that model to draw similar conclusion to ours for the Cr spacer long period oscillation.
The lattice mismatch at the interfaces of the sandwich structures, is one of the factors that could probably affect the agreement between theory and experiment in comparing the sizes of the oscillation periods to the sizes of the extremal vectors of the bulk Fermi surface. In the case of Cr as well as Cr<sub>(1-x)</sub>V<sub>x</sub> spacers, that lattice mismatch is not important since the lattice constants of Cr, V, and Fe are very close to each other, but in other cases like for example Fe/Mo/Fe or Fe/Cr<sub>1-x</sub>Mo<sub>x</sub>/Fe alloy spacers the effect of lattice mismatch might be large enough to be ignored. Thus for instance in ref. , we argue that it could be the explanation for the discrepancy between the Rigid Band Model and the significant decrease with concentration been observed experimentally for Fe/Cr<sub>1-x</sub>Mo<sub>x</sub>/Fe systems. The Rigid Band Model is been proved to be correct for the isoelectronic Cr, Mo and their alloys and is not consistent with the decrease in the oscillation period with concentration. We argue in ref, that the size of the effect of lattice mismatch is enough to explain such a behavior, although someone needs to know the exact geometry of the sandwich structure to draw a conclusive evidence.
Having established the relation of the N-hole pocket and the OMC across Cr and Cr<sub>(1-x)</sub>V<sub>x</sub> spacers the OMC is proven to be a powerful experimental technique for studying the geometry of that pocket and how it evolves with concentration. In particular the N-hole pocket ellipsoid appears to grow isotropically with V concentration as is shown in fig. 5. In that figure the periods predicted form the 3 smallest N pocket extremal vectors for the (100), (110) and (211) directions are shown as functions of V concentration. That does not appear to be the case in recent 2d-ACAR experiments where a rotation of the N-hole pocket is observed with increasing concentration of Vanadium. The only experimental technique apart from 2-d ACAR which could resolve this extremely delicate feature of the Fermi surface appears to be the OMC.
The long period oscillation across Cr<sub>(1-x)</sub>V<sub>x</sub> spacers is an example in which the alloy theory gives a conclusive answer, by continuity to an outstanding problem concerning the pure metal spacer system.
## 4 The exponential damping due to disorder scattering.
As we have already mentioned an exponential damping of the OMC is present in the case of disordered binary alloy spacers. The characteristic length of the damping, i.e. the quantity $`\mathrm{\Lambda }_\nu `$ in the eq. (2) is related to the coherence length of the quasi-particle states at the endpoints of the extremal vector, which is a measure of the mean free path for these states. A convenient quantity to describe the electronic structure of substitutionally disordered systems such as the random binary alloys is the so called Bloch Spectral Function (BSF) $`A_B(𝐤,E)`$ which is the number of states per Energy and wave length. In the case of pure metals that function is simply a sum of delta functions either as function of $`E`$ at a constant wavevector $`𝐤`$, or as a function of the wavevector $`𝐤`$ for constant value of the energy. For constant $`E=E_f`$ where $`E_f`$ is the Fermi energy the positions of the peaks in k-space define the Fermi surface of the metal. The k-space representation is still a good description of the electronic structure of the alloy although strictly speaking there is periodicity only on the average. In terms of the BSF, the fundamental difference is the peak lowering and broadening being Lorentzian-like. Thus a Fermi surface for the alloy is still defined through the position of the peaks but these peaks have a finite width, the inverse of which defines the coherence lengths we mentioned above. In a simple theory for the OMC in the case of disordered spacer in ref we showed that
$$\frac{1}{\mathrm{\Lambda }_\nu }=\mathrm{\Gamma }_\nu ^{(+)}\mathrm{\Gamma }_\nu ^{()}$$
(3)
where $`\mathrm{\Gamma }^{(+)}`$, $`\mathrm{\Gamma }^{()}`$ are the widths of the BSF peaks along the direction of the extremal vector with $`(+)`$ and $`()`$ labeling the two peaks at the end-points of the extremal vector. Of course the Fermi surface is well defined if the size of $`\mathrm{\Gamma }_\nu ^{(+)}`$ $`\mathrm{\Gamma }_\nu ^{()}`$ is small compared to $`Q_\nu `$, i.e. the size of the extremal vector itself. In that case of course, $`\mathrm{\Lambda }_\nu `$ is large compared to the oscillation period $`P_\nu `$, i.e. no damping is observed within the first few oscillation periods.
The obvious question in the light of the above discussion is of course how broad are the spectral functions for the cases of extremal vectors we considered above for the Cu<sub>(1-x)</sub>Ni<sub>x</sub> and Cr<sub>(1-x)</sub>V<sub>x</sub> Fermi surfaces? In fig 6 we show the BSF along these extremal vectors. We see that for Cu<sub>(1-x)</sub>Ni<sub>x</sub> at concentrations of the order $`x0.5`$, where an electronic topological transition (ETT) takes place, in order the widths become comparable with the size of the extremal vectors. On the other hand for Cr<sub>(1-x)</sub>V<sub>x</sub> the extremal wave vector size is always very large compared to the width of the peaks. The BSF peaks for the N-hole pocket appear to be the sharpest for the whole Fermi surface. Although ETT occurs in other parts of the Fermi surface, the N-hole pockets are robust in alloying, with only its total volume increased as more V is added. Thus, in both Cu<sub>(1-x)</sub>Ni<sub>x</sub> ($`x0.4`$) and Cr<sub>(1-x)</sub>V<sub>x</sub> for the whole range of V concentration, no significant damping is expected agreement with the experiments on these two systems.
It would be interesting if an alloy spacer system was found with a strong topological transition happening in the area of the extremal vector at some value of the concentration $`x_o`$. The exponential damping of the OMC for values of $`x`$ close to $`x_o`$ could be measured and the dramatic prediction of the theory is that no oscillation would be observed for the concentration where the ETT occurs. Such an experiment would be a direct observation of the ETT. Moreover, the characteristic length of the damping for $`x`$ close to but not equal to $`x_o`$ would be a direct measurement of the coherence length at particular points of the Fermi surface, i.e. the endpoints of the extremal vector. Of course such measurements require further refinement of the experimental techniques used for measuring the OMC.
## 5 Calculations of the amplitudes $`A_\nu `$ and phases $`\varphi _\nu `$.
The asymptotic analysis leading to a theory of the amplitudes $`A_\nu `$ and the phases $`\varphi _\nu `$ has been carried out within a number of approaches to the problem of calculating the distortions in the electronic structure of the spacer by the magnetic layers. In particular, it was carried out for the powerfull, fully first principles screened KKR method. Quantitative calculation of the asymptotic formulas are compared with the results of full total energy calculations and experiments in fig. 7. Similar calculations for alloy spacers are in progress.
As can be seen in fig. 7 the agreement between our calculated amplitudes and phases with both theoretical and experimental results for the OMC across Co/Cu/Co is remarkable for all the orientations. In particular there is very good quantitative agreement between our results and Lang et al in the amplitudes as seen in fig. 7a for the (100) orientation. Furthermore, there is excellent agreement with the experiments of refs. concerning the phases and periods, i.e. the positions of AF peaks, as seen in the inset of fig. 7b and in fig. 7c. There is only agreement in the order of magnitude with experiments concerning the size of the interaction. For example, for the (100) orientation Johnson et al have measured 0.4 mJ/m<sup>2</sup> for an average Cu thickness of 6.7 ML, while we find 3.3 mJ/m<sup>2</sup> and -1.0 mJ/m<sup>2</sup> for 6 and 7 ML respectively. For the (111) Johnson et al measured 1.1 mJ/m<sup>2</sup> for 4 ML of Cu thickness, while our calculated value for the same thickness is 2.1 mJ/m<sup>2</sup>. The agreement with experiments concerning the amplitudes, although it is restricted to the order of magnitude so far, is at least a convincing evidence that our models have picked up the effect which is also seen by experiments.
In conclusion, we note that the asymptotic formulas such as the one we derived in may play a role in interpreting OMC experiments similar to that of the Lifshitz-Kosevich semi-classical formula, used to deducing the Fermi surface geometry from the measurement of the de Haas van Alphen oscillations.
|
no-problem/9908/cond-mat9908487.html
|
ar5iv
|
text
|
# References
Abstract
We study in this paper the localization of light and the dielectric properties of thin metal-dielectric composites at the percolation threshold and around a resonant frequency where the conductivities of the two components are of the same order. In particular, the effect of the loss in metallic components are examined. To this end, such systems are modelized as random $`LC`$ networks, and the local field distribution as well as the effective conductivity are determined by using two different methods for comparison: an exact resolution of Kirchoff equations, and a real space renormalization group method. The latter method is found to give the general behavior of the effective conductivity but fails to determine the local field distribution. It is also found that the localization still persists for vanishing losses. This result seems to be in agreement with the anomalous absorption observed experimentally for such systems.
Introduction
Thin metal-dielectric films have been shown experimentally to exhibit an anomalous high absorption in the visible, near infrared and microwave regimes and near the percolation threshold\[1-4\]. This effect was interpreted as cluster plasmon absorption . Naturally, these mixtures should show an absorption since the dielectric constant of the metallic component is complex and the effective dielectric constant of the whole system should then also be complex particularly at the percolation threshold where
$$ϵ_{eff}=\sqrt{ϵ_mϵ_d}$$
(1)
The indices $`m,d`$ and $`eff`$ stand respectively for the metal, dielectric and effective medium. From this equation, both real and imaginary parts of the dielectric constant of the metallic component contribute to the optical absorption of the system. Furthermore, a vanishing metallic absorption naturally should lead to a vanishing absorption of the whole system since a superconductor-dielectric mixture should be non dissipative. Therefore, this description of the effective properties of such medium may not be sufficient to explain the behavior experimentally observed for such systems \[1-4\].
Indeed, since the light has a wave behavior, the backscattering and the interference effects can strongly affect its propagation through a disordered system and its localization is enhanced by disorder as well as by absorption . Therefore, the localization properties of the electromagnetic field in such systems where both absorption and disorder are present, can be a good tool for explaining the anomalous behavior observed in such thin composite films, particularly at the classical percolation threshold which is a transition point of the effective $`dc`$ conductivity of the system from non-conductor to conductor due to the appearance of a continuous path of the conducting region through the sample. The classical percolation threshold in 2D disordered bonds corresponds to a concentration of the metallic bonds $`p_c=0.5`$ (note that some composites carry current even below the classical percolation threshold due to the fact that tunnelling through disconnected (dispersed) metallic regions can give some virtually connected percolating clusters ).
On the other hand, it has recently been found numerically in such films, giant local field fluctuations at the percolation threshold and for frequencies close to a resonant one $`\omega _{res}`$ where the conductivities of the two components are of the same order ($`\left|\sigma _m\right|=\left|\sigma _d\right|`$). High local field fluctuations have also been found both in fractal 2D films , 3D rough surfaces and non-linear Raman scattering . In both systems the electromagnetic modes were found to be localized, apparently, due to such fluctuations. Furthermore, Brouers et al. showed for the metal-dielectric films that the local field distribution is asymptotically log-normal. However, from the electromagnetic field theory investigated recently by Sarychev et al. , we can easily deduce that the high strengths of the current (or equivalently the high local field intensities in this case) behave as the inverse of the local transmission of light. We deduce, then, that the local transmission has also a log-normal distribution in such films. Therefore, if we use the analogy between the electric field in Helmoltz equation and the electronic wave-function in Schrödinger equation , the local transmission is equivalent to the electronic conductance at zero temperature where a log-normal distribution is a signature of localization .
This is the aim of the present paper where we study the localization and absorption properties of such films modelized by a square $`RLC`$ network. The local field and the effective conductivity are calculated by using two different methods for comparison with the results of : an exact resolution of the Kirchoff equations for such network which we call Exact Method (EM) from now on, and a Real Space Renormalization Group (RSRG) self-similar scheme . The degree of localization is measured by means the inverse participation ratio (IPR) applied to the electric field while the absorption is deduced from the real part of the effective conductivity. We compare in a first step the field distribution obtained by the two methods and then examine the effect of the loss in the metallic component on the localization as well as the absorption at the percolation threshold.
Method of the calculations
The conductivity of metallic and dielectric grains is related to their dielectric constant by
$$\sigma _{m,d}=\frac{id\omega ϵ_{m,d}}{4\pi }$$
(2)
where $`d`$ is the film thickness, $`\omega `$ the field frequency and $`ϵ_{m,d}`$ the dielectric constants respectively of the metal and dielectric components assumed to be homogeneous spheres. Here the film thickness and the size of the components must be smaller than the light wavelength in order to neglect the magnetic field variation. The dielectric constant of the insulator is real while the metallic one is complex and, from Eq.(2), its imaginary part (absorption) is related to the real part of the conductivity.
When the light frequency is large compared to the relaxation frequency, composite metal dielectric films can be modelled by 2D resistor networks . The effective properties of these networks have already been extensively studied during the last two decades \[5,6,23-25\]. In the $`RLC`$ picture, and if the frequency is smaller than the plasmon frequency $`\omega _p`$, capacitors $`C`$ stand for the dielectric grains with a conductivity $`\sigma _C=iC\omega `$ and a concentration $`1p`$ while inductances $`L`$ represent the metallic grains with a conductivity $`\sigma _L=(iL\omega +R)^1`$ ($`R`$ being the loss), and a concentration $`p`$ deposited or evaporated over the substrate. Therefore we can take in this case, without loss of generality, $`L=C=\omega _{res}=1`$ (the resonant frequency $`\omega _{res}`$ corresponds to the frequency where metallic and dielectric conductivities have the same magnitude for small losses, i.e., $`C\omega _{res}=1/L\omega _{res}`$ ). $`L`$ and $`C`$ are constants near the resonant frequency, this assumption can be generalized to any frequency $`\omega `$ (smaller than the plasmon frequency) normalized to $`\omega _{res}`$. The metallic conductivity for small losses becomes
$$\sigma _L=\frac{1}{i\omega +R}=(i+\frac{R}{\omega })/\omega $$
(3)
while the dielectric conductivity is
$$\sigma _C=i\omega $$
(4)
The first method (EM), used for the calculation of the local field distribution and the effective conductivity, consists in solving exactly Kirchoff equations for the corresponding 2D square resistor network. This implies the use of $`M^2`$ x $`M^2`$ matrices (where $`M`$ is the size of the square lattice), which are impossible to handle numerically for large samples (for memory and computational time consuming reasons). However, we take advantage of the sparce configuration of such matrices and their organization in blocks in their diagonal region. The diagonalization and inversion of such matrices, then, is obtained simply from the diagonalization and inversion of the constituent blocks. This reduces considerably the memory and the computational time consuming. We note that this method provides exactly the same results on the effective conductivity as the Frank and Lobb method but calculates also the local field distribution through the lattice which cannot be done by the other method. However, the time consuming remains large with this method (in particular when averaging over a large number of configurations), and the maximum size we reach by this method is $`256`$ x $`256`$ (which is sufficient for our statistical treatment).
This is one of the reasons for using also the RSRG method which is much less computational time consuming. This method, extensively described in previous works (see ), consists in a transformation of the 2D square lattice into Weatstone bridges in $`x`$ and $`y`$ directions (see Fig.1). Each bridge is transformed into an equivalent admittance, and after a number of steps the lattice is reduced into two equivalent admittances following these directions. It is then easy to calculate by this transformation the effective conductivity while the local field distribution can be obtained by the inverse procedure starting from the effective admittances already calculated. Although this method is an approximation, it has been shown to give values of the effective conductivity near the percolation threshold very close to the exact ones for 2D composites and critical exponents not far away from the known values of the percolation theory . Furthermore, this method uses only few matrices of $`M`$x$`M`$ for sample sizes $`M`$x$`M`$ which reduces considerably the computational memory in comparison with the other methods. We can, then, easily reach sizes of $`1024`$x$`1024`$ with the same computer configuration as for the previous method. However, the first method (EM) is also needed in the present work since the validity of RSRG in calculating the local field has not been checked before.
As discussed in the previous section, it seems that the localization properties of the optical waves in thin metal-dielectric composites is an interesting way to explain the anomalous absorption observed near the percolation threshold. This classical threshold corresponds to the appearance of an infinite metallic channel which, in 2D disordered systems, is reached for an equal probability of the two components . One of the useful quantities to study the localization in electronic systems is the inverse participation ratio (IPR) . By analogy with the quantum counter part, the local electric field in Helmoltz equation plays the role of the electronic wave function in the Shrödinger equation and the IPR becomes
$$IPR=\frac{\mathrm{\Sigma }_i\left|E_i\right|^4}{\left(\mathrm{\Sigma }_i\left|E_i\right|^2\right)^2}$$
(5)
where $`E_i`$ denotes the local electric field at site $`i`$. The IPR has been defined for the electronic waves in order to measure the spatial extend of the dominant eigenstates and to characterize the electronic states in disordered materials . Therefore, for the electromagnetic eigenmodes this quantity will behave as
$`IPR=O(M^d)forextendedeigenmodes,`$ (6)
$`IPR=O(M^0)forstronglylocalizedeigenmodes.`$ (7)
Here $`d`$ denotes the Euclidian dimension of the system ($`d=2`$ in this case) and $`M`$ the size of the system. Thus in the case of purely extended eigenmodes, the field has a significant strength over the whole surface of the film and the denominator will be $`M^4\left|E\right|^4`$ while the numerator behaves as $`M^2\left|E\right|^4`$ (by assuming the field constant) leading to a decay of the IPR as $`M^2`$. In the case of strongly localized eigenmodes, the more significant field strengths are located in a limited area of average size $`M_c^2`$ where $`M_c`$ is the localization length. It is then obvious that the IPR remains constant outside this area. Therefore we can estimate the degree of localization of the light from the power-law decay exponent of the IPR which varies from $`0`$ (for strongly localized eigenmodes) to $`2`$ (for purely extended eigenmodes). This exponent is determined by the slope of the variation of the IPR as a function the system size in log-log scale. It also measures the correlation fractal dimension ($`D_2`$) of the local field . From now on we will call this exponent the slope of the IPR.
Results and Discussion
In this section we consider a film of size $`256`$x$`256`$. As discussed above, the RSRG method is a good approximation for the calculation of the effective conductivity and the critical exponents. However, this agreement cannot be generalized to any other quantity. In particular, the value of the effective conductivity at the percolation threshold is due only to the appearance of an infinite metallic cluster channel in the sample no matter how the bonds are distributed over the sample . However, the distribution of bonds can affect sensitively the local field distribution which is the main quantity to measure the localization properties of light in this system. Thus, some particular arrangements of the bonds yield a local field intensity ($`\left|E\right|^2`$) of the order of $`R^2`$ by the RSRG method. Therefore, in the limit of vanishing losses the field intensity will diverge which is unphysical. This divergence can limit the validity of the RSRG method for vanishing losses.
In Figs.2 we compare the distributions of the local field intensity obtained by the RSRG method to those of the EM for sample sizes $`256`$x$`256`$, for two different losses and two different frequencies in order to check the validity of RSRG in calculating the local field. The RSRG distributions seem to be wider, with very large field strengths, than for the EM distributions (supporting the previous discussion on the divergence of the field) which seem to be perfectly log-normal for any loss and frequency. Furthermore, other peaks of small field strengths appear in the distribution obtained by RSRG particularly for small losses and at the resonant frequency $`\omega =1`$, while for higher losses and different frequencies, these peaks move to larger field strengths and overlap with the main one contributing to the broadening of the distribution. This behavior affects strongly the IPR calculations since the broadening of the distribution means an increase of the localization . However, the additional peaks appearing for small losses in the RSRG method should contribute only slightly to the IPR since they are many orders smaller than the main peak. Therefore, although the ditributions of field for RSRG are clearly different from those of the EM, it seems that the results for $`\omega =1`$ provide the best fit for the localization properties of these films. Therefore, we will restrict ourselves to the resonant frequency ($`\omega =1`$) from now on.
In Figs.3, we show, by using both RSRG and EM methods, the real part of the conductivity averaged over $`50`$ samples of size $`256`$x$`256`$ (Fig.3a) as well as the slope of the IPR (Fig.3b) as a function of the loss parameter $`R`$ for $`\omega =1`$. It seems that the conductivity tends to vanish as a power-law for small losses with an exponent close to $`1`$ while it tends to saturate for larger losses with strong fluctuations in the region of the loss between $`10^3`$ and $`10^6`$ (while for smaller or larger losses the conductivity seems to be self averaged). The behavior for small losses is in agreement with the theoretical predictions. Indeed, for vanishing losses both dielectric and metallic components of the film are non-dissipative and then the whole system becomes non-dissipative. Furthermore, from Eqs. (1-4), the real part of the effective conductivity should behave as
$$Real(\sigma _{eff})=\frac{1}{2}\frac{R}{\omega }$$
(8)
which seems to be well fitted in Fig.3a. Although a difference is shown between the two methods, they follow qualitatively the same behavior for small losses. However, as expected from Figs.2, the IPR shows a delocalization (for the RSRG method) for increasing loss while the EM method yields the inverse situation which is the expected behavior (for the same configuration, an increase of the loss corresponds to an increase of the absorption which means a localization ). Therefore, this confirms the failure of this renormalization group method in describing the localization properties of the system, due to the large field strengths obtained by this method which broaden the distributions shown in Fig.2. On the other hand, the IPR slope obtained by the EM method seems to saturate at the value $`1.3`$ in the region of small losses (see Fig.3b) indicating that the eigenmodes remain localized even for vanishing losses. Indeed, this localization is due only to the disorder in the conductivity (Anderson like localization). The disorder here does not come from the strength of the conductivity (since the conductivities of the two components have the same strength) but from its phase which takes randomly two values: $`+\frac{\pi }{2}`$ and $`\frac{\pi }{2}`$ for vanishing losses. We also see from this Figure, in the region of loss between $`10^3`$ and $`10^6`$, a non monotonic behavior of the IPR (EM). This is due to the strong fluctuations observed in this region. Therefore, this region should show a different statistical behavior.
Conclusion
In this article we have studied, by using the RSRG method and also the EM method, the localization and absorption properties of the electromagnetic field in a thin semicontinuous metal-dielectric film for a characteristic frequency $`\omega _{res}=1`$ at the classical percolation threshold. It seems that the real part of the effective conductivity for RSRG agrees qualitatively with the exact calculations (EM) and give the expected power-law behavior for vanishing loss. While this method (RSRG) fails in giving the right behavior for the IPR, due to the large field strengths provided in comparison with the EM method. On the other hand, it seems that the IPR saturates at $`1.3`$ for vanishing losses in agreement with the anomalous experimental results observed for such films. We can explain the anomalous absorption observed by a confinement of the light at the percolation threshold. We also found that the conductivity strongly fluctuates in the region of the loss between $`10^3`$ and $`10^6`$ while it is self averaging for larger or smaller losses. An extensive study of the statistical properties of the conductivity is then needed. Indeed, may be in that region the logarithm of the conductivity is self averaged in this region. Furthermore, it should be interesting to study these effective properties around the classical threshold where interesting features can occur. It is also important to examine these effects for a partially ordered sample since realistic films show a local arrangement. All these interesting features will be the subject of a forthcoming investigation.
Acknowledgments
Two of us (L.Z and N.Z) would like to acknowledge the hospitality of the ICTP during the progress of this work. This work was done within the framework of the Abdus Salam International Centre for Theoretical Physics, Trieste, Italy. Financial support from the Arab Fund is acknowledged.
Figure Captions
Fig.1 The real space renormalization group for a square network.
Fig.2 The distribution of the local electric field intensity $`\mathrm{log}(\left|\mathrm{E}\right|^2)`$ for $`\omega =1`$ : a) $`R=10^1`$, b) $`R=10^6`$, and $`\omega =1/8`$ : c) $`R=10^1`$, d) $`R=10^6`$. Solid curves correspond to the EM calculations and dashed curves to RSRG method.
Fig.3 a) The real part of the effective conductivity in a log-log plot and b) the slope of the IPR in a semi-log plot as a function of the loss parameter $`R`$ for $`\omega =1`$. Open squares correspond to the RSRG method and solid squares to the EM method. The conductivity is averaged over 50 samples and the IPR is calculated for only one configuration.
|
no-problem/9908/quant-ph9908088.html
|
ar5iv
|
text
|
# Comment on “Validity of Feynman’s prescription of disregarding the Pauli principle in intermediate states”
## Abstract
In a recent paper Coutinho, Nogami and Tomio \[Phys. Rev. A 59, 2624 (1999)\] presented an example in which, they claim, Feynman’s prescription of disregarding the Pauli principle in intermediate states of perturbation theory fails. We show that, contrary to their claim, Feynman’s prescription is consistent with the exact solution of their example.
PACS numbers: 03.65.Ge, 03.65.Pm, 11.15.Bt, 12.39.Ba
Feynman’s prescription of disregarding the exclusion principle in intermediate states of perturbation theory is based in his observation that all virtual processes that violate that principle (formally) cancel out order by order in perturbation theory. However, Coutinho, Nogami and Tomio have recently presented an example in which Feynman’s prescription seems to fail. They calculated (in second order perturbation theory) the energy shift $`WE(\lambda )E(0)`$ of the ground state of the one-particle sector of the one-dimensional bag-model caused by the potential $`V(x)=\lambda x`$. The calculation was performed using both prescriptions, i.e., either excluding virtual transitions to occupied states, in accordance with the Pauli principle (method I), or including such transitions, as suggested by Feynman (method II). It turns out that the result depends on the prescription used; in the massless case, for instance, $`W_\mathrm{I}0`$ and $`W_{\mathrm{II}}=0`$. In view of this discrepancy, Coutinho et al. suggested that one should abandon Feynman’s prescription and remain faithful to the Pauli principle in every step of the calculation.
The purpose of this Comment is to show that the opposite alternative is the correct one. More precisely, we show that $`W_{\mathrm{II}}`$ agrees (while $`W_\mathrm{I}`$ does not) with the exact value of $`W`$ in the massless case.
To calculate $`W`$ we need to solve the Dirac equation
$$(H_0+\lambda x)\psi (x)=ϵ\psi (x),$$
(1)
where $`H_0`$ is the Hamiltonian of the bag model (see for its definition). With the conventions of ref. , Eq. (1) can be written explicitly in the massless case as
$$\left(\begin{array}{cc}\lambda xϵ& _x\\ _x& \lambda xϵ\end{array}\right)\left(\begin{array}{c}u\\ v\end{array}\right)=\left(\begin{array}{c}0\\ 0\end{array}\right)(|x|a),$$
(2)
with $`u`$ and $`v`$ subject to the boundary conditions $`u(\pm a)=v(\pm a)`$. Defining $`w_\pm =u\pm iv`$, we can rewrite (2) as
$$_xw_\pm i(\lambda xϵ)w_\pm =0,$$
(3)
whose general solution is
$$w_\pm (x)=C_\pm e^{\pm i\left(\frac{\lambda }{2}x^2ϵx\right)}.$$
(4)
The boundary conditions on $`u`$ and $`v`$ turn into $`w_+(\pm a)=iw_{}(\pm a)`$. This gives two relations between $`C_+`$ and $`C_{}`$:
$`C_+`$ $`=`$ $`iC_{}e^{i\left(\lambda a^2+2ϵa\right)},`$ (5)
$`C_+`$ $`=`$ $`iC_{}e^{i\left(\lambda a^22ϵa\right)}.`$ (6)
Dividing (5) by (6) and solving for $`ϵ`$ gives
$$ϵ=(2n+1)\frac{\pi }{4a}(n=0,\pm 1,\pm 2,\mathrm{}).$$
(7)
Since the energy levels do not depend on $`\lambda `$, the energy shift $`W`$ is zero. This is precisely the result obtained by Coutinho et al. using Feynman’s prescription in perturbation theory.
The author acknowledges the financial support from FAPESP.
|
no-problem/9908/astro-ph9908014.html
|
ar5iv
|
text
|
# Classical Cepheid Pulsation Models. III. The Predictable Scenario.
## 1 INTRODUCTION
This is the third paper of a series devoted to the pulsational properties of classical Cepheids predicted by nonlinear pulsation models. In paper I (Bono, Marconi & Stellingwerf 1999) we presented both numerical and physical assumptions adopted for constructing nonlinear, convective models, and discussed the approach to limit cycle stability of both fundamental and first overtone pulsators. By adopting the periods and the modal stability predicted by these models, paper II (Bono et al. 1999a) was focused on the Period-Luminosity (PL), the Period-Color (PC), and on the Period-Luminosity-Color (PLC) relations at selected stellar chemical compositions. Moreover, Bono, Caputo, & Marconi (1998) have already discussed the satisfactory agreement between empirical Cepheid radii and the corresponding theoretical predictions based on this new theoretical scenario.
However, nonlinear models do provide many more predictions, such as the pulsational amplitudes and the morphology of both light and velocity curves, which can be directly compared with the observed Cepheid properties. Even though some of these properties were already investigated on the basis of linear (Chiosi, Wood, & Capitanio 1993; Simon & Young 1997; Saio & Gautschy 1998; Alibert et al. 1999) and nonlinear (Carson & Stothers 1988; Moskalik, Buchler, & Marom 1992) envelope models, several empirical evidence such as the change of luminosity and velocity amplitudes across the instability strip (Sandage & Tammann 1971; Pel 1978; Cogan 1980; Fernie 1990) or the constant amplitude ratio in different photometric bands (Freedman 1988; Tanvir 1997) have not been enlightened by theoretical insights yet. In this paper we supply a comprehensive analysis of pulsation predictions to stimulate further investigations by others working in this field and to discuss the comparison between these theoretical predictions and observational data available in the current literature.
In section 2 we discuss the instability boundaries, we compare them with previous results and discuss the dependence of the boundaries on the adopted chemical composition as well as on the assumptions about the mass-luminosity (ML) relation. In Section 3 we supply suitable analytical relations for fundamental and first overtone pulsators connecting the period to stellar masses, luminosities and effective temperatures. Luminosity, radius, gravity, temperature, and radial velocity amplitudes for fundamental pulsators are discussed in section 4, while first overtone pulsation amplitudes are outlined in section 5. In these two sections we also compare theoretical observables with both photometric and spectroscopic measurements available in the literature. Finally, the main findings of this investigation are summarized in §6, together with some hints on future developments of the present theoretical framework.
Appendix A gives input parameters and theoretical observables for all the computed pulsation models, while Appendix B presents their predicted light and velocity curves together with a brief discussion on the variation of the morphology across the instability strip.
## 2 Modal instability and periods
Limiting amplitude, nonlinear, convective models represent a necessary and sufficient condition to assess the modal stability and, in turn, to evaluate the edges of the instability strip for each given stellar structure. In the case of Cepheid variables, one is dealing with intermediate-mass stars which cross the instability region at different luminosities which depend on stellar mass and on chemical composition. The pulsational investigation thus requires suitable assumptions about the ML relation for constraining the predicted luminosity of the model for each given value of the stellar mass. For this purpose we adopted evolutionary predictions presented by Castellani, Chieffi & Straniero (1992, hereinafter CCS92), which were computed for suitable ranges of stellar masses and chemical compositions and under the canonical assumption of no convective core overshooting. Figure 1 shows the comparison between the ML relation adopted in our computations and the relations adopted in linear investigations by Chiosi et al. (1993, hereinafter CWC93) and by Alibert et al. (1999, hereinafter ABHA99). Data plotted in the top panel for Z=0.02 show that the quoted evolutionary models are in excellent agreement. As we will discuss later on, a maximum difference of the order of $`\mathrm{\Delta }logL0.1`$ has indeed negligible effects on the derived pulsational scenario. The same figure in the bottom panel shows the dependence of the ML relation on the adopted chemical composition. Note that our models assume Y=0.28 for Z=0.02 and $`\mathrm{\Delta }`$Y/$`\mathrm{\Delta }`$Z=2.5 for other metallicities, in close similarity with the assumptions made by the quoted authors. In particular, CWC93 adopted Y=0.25, Z=0.004; Y=0.25, Z=0.008; and Y=0.30, Z=0.016, while ABHA99 adopted Y=0.25, Z=0.004; Y=0.25, Z=0.01; and Y=0.28, Z=0.02. Once again we find a rather satisfactory agreement between the different predictions, with a slight enhanced dependence of ABHA99 relations, to be possibly ascribed to different input physics adopted in ABHA99 evolutionary models. However, data plotted in Figure 1 clearly show that we are facing a rather well-established evolutionary scenario.
On the basis of the adopted ML relation we explored the pulsation stability of models at 5, 7, 9, and 11 M with the effective temperature $`T_e`$ as a free parameter to be moved by steps of 100 degrees across the HR diagram. With this procedure the location of the instability boundaries was evaluated with an accuracy of $`\pm `$50 K. The input physics used for constructing both linear and nonlinear envelope models was already described in paper I and in paper II as well as in Bono et al. (1998). Here we only mention that we adopted OPAL opacities (Iglesias & Rogers 1996) for temperatures greater than $`10^4`$ K, and Alexander & Ferguson (1994) molecular opacities for lower temperatures. In addition an independent set of models was computed for the same values of stellar masses but by increasing the predicted luminosities by $`\mathrm{\Delta }logL/L_{}=0.25`$. According to CWC93, this luminosity shift allows us to explore an evolutionary scenario which accounts for a mild convective core overshooting and therefore to investigate the effects of the mass/luminosity ratio on the pulsation instability. Input parameters and theoretical results for all models characterized by a stable nonlinear limit cycle in the fundamental and/or in the first overtone are presented in Appendix A, Tables 1-4.
Results concerning the fundamental edges of the instability strip are shown in Figure 2 for the two assumed ML relations and the three different stellar metallicities. Data in this figure give the plain evidence that the instability edges depend mainly on luminosity. Noncanonical models -i.e. pulsation models constructed by adopting a ML relation based on mild overshooting evolutionary models- suggest that the red edges at the highest luminosities present a sudden shift toward hotter effective temperatures, with a sharp decrease in the width of the instability strip. We find that both canonical and noncanonical boundaries can be described over a quite large luminosity range by analytical relations of the type $`logT_e=\alpha +\beta \times logL/L_{}`$. The analytic fits for the red edges were derived by neglecting the noncanonical point at $`M=11M_{}`$, while the blue edge relation for Z=0.004 was derived by neglecting the canonical point at $`M=5M_{}`$. Coefficients and errors of both blue and red edge relations for the three different chemical compositions are given in Table 5.
Moreover, data displayed in Figure 2 strongly support the empirical evidence originally pointed out by Pel & Lub (1978) and by Fernie (1990) that Galactic Cepheids present a ”wedge-shaped” and not a ”rectangular-shaped” instability strip. A similar behavior for Magellanic Cepheids was originally brought out by Martin, Warren, & Feast (1979) and more recently by Caldwell & Laney (1991). This suggests that the narrowing at the lower luminosities is an intrinsic feature of the fundamental instability strip which does not depend on metallicity. Figure 2 shows the additional evidence that at each given luminosity an increase in the metal content shifts the instability strip toward cooler effective temperatures. The dependence of PL, PC, and PLC relations on this effect was thoroughly described in paper II and therefore it is not discussed further.
Figure 3 shows the comparison between our fundamental edges and similar predictions obtained by CWC93 and by ABHA99 from linear, nonadiabatic, convective models. We find that at the lowest metallicity our predicted blue boundaries are located between the two linear ones, with a slope which appears in good agreement with both CWC93 and ABHA99 results. However, linear red edges appear much steeper than is predicted by nonlinear models. Such a discrepancy is not surprising, since the linear approach does not account for the coupling between pulsation and convection, and therefore can hardly predict the modal stability of models located close to the cool edge. To cope with such a difficulty, linear red edge estimates by CWC93 and by ABHA99 were not fixed at the effective temperature where the fundamental growth rate attains vanishing values but, more or less tentatively, at the effective temperature where the growth rate attains its maximum value (Chiosi et al. 1992; ABHA99). Present nonlinear results suggest that such an assumption is far from being adequate, a conclusion further supported by the evidence that ABHA99 predictions fail to reproduce the already discussed empirical evidence for a ”wedge-shaped” instability strip. The difference in the blue edges provided by CWC93 and ABHA99 also suggests that linear predictions depend on the numerical and physical assumptions adopted for constructing pulsation models.
Moreover, one may notice that an increase in the metal content causes a flattening in the slope of the nonlinear blue boundary, at variance with the predicted linear behavior. Such an occurrence is probably connected with the evidence that a metallicity increase shifts the blue edge toward cooler effective temperatures. As a consequence, metal-rich models show a stronger dependence on the nonlinear effects introduced by the coupling between pulsation and convection than metal-poor ones. The reader interested in a detailed discussion on the dependence of modal stability and pulsation properties on these effects is referred to paper I.
Figure 4 shows the location in the HR diagram of first overtone unstable models we found for the three selected chemical compositions in the lower portion of the explored luminosity range. In this region for the two lower metallicities we derived the following linear analytical relations for first overtone blue boundaries:
$$\mathrm{log}T_{e}^{}{}_{}{}^{B}(Z=0.004)=0.047(\pm 0.005)\mathrm{log}L+3.954(\pm 0.002)\sigma =0.002$$
$$\mathrm{log}T_{e}^{}{}_{}{}^{B}(Z=0.008)=0.067(\pm 0.004)\mathrm{log}L+4.011(\pm 0.001)\sigma =0.001$$
where $`L`$ is the luminosity (solar units), $`T_e`$ is the effective temperature (K), and $`\sigma `$ is the standard deviation.
## 3 Pulsation relations
During the last few years several authors (CWC93; Simon & Young 1997; Saio & Gautschy 1998) have used linear, nonadiabatic models to derive pulsational relation connecting periods to stellar masses and luminosities. In paper I (see Fig. 53) we already compared theoretical periods based on both linear and nonlinear models. We found that nonlinear periods of high-mass Cepheids are from 1% to 10% shorter than the linear ones when moving from the blue to the red edge of the instability strip. This difference can introduce an uncertainty of the order of 0.08 mag in the comparison between theoretical and empirical PL relations. This finding supports the relevance of precise pulsation relations.
Fortunately enough, we find that at each given chemical composition the present nonlinear values of canonical and noncanonical fundamental periods follow with very good accuracy a linear relation connecting $`logP`$ to the logarithms of the pulsator mass, luminosity, and effective temperature. The coefficients and the errors of these analytical relations are given in Table 6. Figure 5 shows how accurately these relations fit the periods of the computed models. It is now possible to estimate the effects of uncertainties in the adopted ML relation. Going back to Figure 1 we find that at fixed luminosity -the physical parameter governing the instability boundaries- current canonical ML relations predict quite similar mass values. In fact, the difference at $`logL/L_{}=3.5`$ between CCS92 and CWC93 is $`\delta logM\pm 0.01`$, while between CCS92 and ABHA99 is $`\delta logM\pm 0.02`$. This difference causes a period uncertainty of 2% and 4% respectively and, in turn, of only few hundreds of magnitude on the distances obtained by adopting the theoretical PL relation. On the other hand, the mass difference between canonical and noncanonical ML relations is of the order of $`\delta logM\pm 0.07`$ which implies a 14% period uncertainty and, in turn, an uncertainty of about 0.15 mag on the distance modulus. However, this uncertainty affects absolute but not relative distance determinations due to the systematic difference between the two different evolutionary frameworks over the whole period range.
Data in Figure 5 shows that for each given effective temperature an increase in metallicity, in spite of the decrease in luminosity, causes a small increase in the pulsation period of the order of 0.03 dex when moving from Z=0.004 to Z=0.02. However, an increase in metallicity has the major effect of shifting the instability strip toward lower effective temperatures and therefore longer periods. As already discussed in Paper II, such a shift plays a key role in governing the dependence of the PL relation on metallicity. This shift in temperature also provides a plain explanation of the empirical evidence originally pointed out by Gascoigne (1969, 1974) and more recently by Sasselov et al. (1997) that Cepheids in the Small Magellanic Cloud (SMC) are, for a given period, bluer than Cepheids in the LMC. Alternatively, we predict that at fixed luminosity the period distribution of SMC Cepheids should be shifted toward shorter periods when compared with LMC Cepheids. The cumulative period distributions of Cepheids in LMC and in SMC recently derived by the EROS project suggest a similar trend (Marquette 1998), but we still lack a detailed comparison between the period distribution of both LMC and SMC Cepheids at fixed magnitude bins.
In the explored range of luminosities we found a relatively small number of first overtone pulsators. This mode is indeed unstable only in low-mass ($`57M_{}`$) and therefore low-luminosity models. Owing to the mild dependence of first overtone periods on metallicity we derived a single pulsational relation which does not account for the metallicity dependence:
$`logP=10.763(\pm 0.002)+0.672(\pm 0.002)logL3.305(\pm 0.039)\mathrm{log}T_e\sigma =0.002`$
where symbols have their usual meaning.
## 4 Pulsational amplitudes: fundamental pulsators
### 4.1 Luminosity
Dating back to the seminal investigation by Sandage & Tammann (1971, hereinafter ST71), the luminosity amplitude has been recognized as a key parameter for assessing the pulsation properties of classical Cepheids. If amplitudes, as suggested by the quoted authors (see also Sandage 1972), depend on the distance from the edges of the instability strip, this parameter could be used to properly locate a Cepheid within the strip, and in turn to use a Period-Luminosity-Amplitude relation for estimating distances. However, empirical evidence concerning the behavior of the luminosity amplitude inside the strip, the so-called ”amplitude mapping”, turned out to be quite controversial. On the basis of Cepheid samples in four different galaxies, ST71 suggested that in the period range from $`logP0.40`$ to $`0.86`$ and for $`logP>1.3`$ the largest luminosity amplitudes are attained close to the blue edge, while the trend is reversed for Cepheids with periods ranging from $`logP=0.86`$ to $`logP=1.3`$. A similar conclusion for Galactic Cepheids was reached by Cogan (1980), who suggested to move the cut-off period from $`logP=1.3`$ to 1.1.
On the other hand, Madore (1976) suggested, on the basis of empirical relations provided by Fernie (1970), that classical Cepheids attain the largest amplitudes close to the red edge, whereas Butler (1976,1978) found an opposite trend among MC Cepheids. More recently Fernie (1990) investigated a large sample of Galactic Cepheids suggesting that for a given period the largest amplitudes are possibly attained close to the center of the instability strip, even though they are not tightly correlated with the position in the strip.
Figure 6 shows bolometric amplitudes of both canonical and noncanonical models as a function of period for the various given masses and for the three selected metallicities. In this figure each sequence of models depicts the predicted amplitudes across the instability strip. A glance at the data plotted in Figure 6 discloses that for the three explored metallicities the pulsators in the short-period range (i.e. with lower masses and luminosities) attain their maximum amplitude at the blue edge, whereas toward longer periods the amplitudes present a ”bell-shaped” distribution.
Such a behavior is similar to the trend found in fundamental RR Lyrae pulsators (Bono et al. 1997, hereinafter BCCM97): at the luminosities where the fundamental blue boundary falls in a region where the first overtone is already unstable, the amplitude of fundamental pulsators steadly decreases from the blue to the red edge, whereas at the luminosities in which only the fundamental pulsators attain a stable nonlinear limit cycle the pulsation amplitudes show a ”bell-shaped” variation from the hot to the cool edge of the instability strip. However, we notice that in the large majority of cases the maximum amplitude is attained only few hundred degrees from the blue boundary, and then the amplitude steadly decreases from the blue to the red over a large portion of the strip. Therefore theoretical results appear in reasonable agreement with the empirical evidence brought out by ST71 and by Cogan (1980).
Another feature disclosed by data plotted in Figure 6 is that an increase in the metal content generally causes a decrease in the maximum amplitudes. This trend supplies a sound support to the empirical evidence originally suggested by Arp & Kraft (1961) and then confirmed by van Genderen (1978) on the basis of a large sample of short-period Galactic and Magellanic Cepheids. The only exceptions to this rule are the 7 $`M_{}`$ (canonical) for Z=0.02 and the 9 $`M_{}`$ (noncanonical) models for Z=0.008 which show a larger bolometric amplitude when compared with more metal-poor models of the same mass value. This peculiarity could be due to a nonlinear behavior of the thermal properties of the outermost layers caused by the shock propagation close to the phases of maximum amplitude (see paper I). In fact, as we will discuss later on, the radial velocity amplitudes of these models do not show this peculiar behavior.
The sequence of canonical models at 7 $`M_{}`$ presents a ”double-peaked” distribution with two maxima located close to the blue and to the red edge. Interestingly enough, these models attain the minimum amplitude at $`logP1.03`$, in remarkable agreement with the minimum at $`logP=1.05\pm 0.03`$ in the empirical distribution of Fourier parameters $`\varphi _{21}`$ and $`R_{21}`$ found in LMC Cepheids by MACHO (Welch et al. 1997) and EROS projects (Beaulieu & Sasselov 1997). The physical mechanism(s) which govern the appearance of such a phenomenon and the dependence on chemical composition will be discussed in a forthcoming paper (Bono et al. 1999b). Finally we note that the period-amplitude variation of this sequence supports the local minimum in B amplitudes found in this period range by van Genderen (1978) for LMC Cepheids. This finding is very encouraging since the Bailey diagram (luminosity amplitude vs. period) does not depend on distance modulus and on color-temperature relation, and presents a negligible dependence on bolometric correction scale and on reddening.
Figure 7 shows the comparison in the Bailey diagram between pulsation amplitudes in the V band of Galactic Cepheids collected by Fernie et al. (1995) and theoretical predictions for Z=0.008 and Z=0.02. The choice of metallicities follows the results of a recent spectroscopic investigation by Fry & Carney (1997), which supports the evidence that Galactic calibrating Cepheids cover this metallicity range. Bolometric amplitudes were transformed into V amplitudes by adopting the bolometric corrections provided by Castelli, Gratton, & Kurucz (1997a,b). We assumed $`M_{bol}()=4.62`$. Theoretical V amplitudes show the same variation of bolometric amplitudes across the instability strip and are in satisfactory agreement with empirical values both at short and at long-periods. The only exception are the models at $`M=7M_{}`$, Z=0.02 and $`M=9M_{}`$, Z=0.008 whose V amplitudes are 20-25% higher than the observed ones. However, it is noteworthy that the luminosity amplitudes predicted by nonlinear radiative models are at least a factor of two larger than those predicted by convective ones (see Figure 1 in Bono & Marconi 1998).
The period-amplitude behavior of Galactic Cepheids was extensively investigated by van Genderen (1974) who found that the envelope line of B amplitudes attains a constant value for $`0.5logP1.0`$ and then a rapid increase up to $`logP=1.15`$. Toward longer periods the envelope line attains once again a constant value up to $`logP=1.5`$ and then a steady decrease (see data plotted in his Figure 1). This empirical evidence was subsequently confirmed by Laney & Stobie (1993) by adopting infrared J, H, and K amplitudes for 51 Galactic Cepheids. Theoretical predictions plotted in Figures 6 and 7 seem to support this behavior, but to confirm this evidence further models at solar chemical composition in the period range $`0.7logP1.0`$ are needed.
### 4.2 Amplitude ratio
The agreement between theory and observations on the behavior of the luminosity amplitude inside the instability strip suggested to test the empirical assumption that the amplitude ratio between the I (Cousins) and the V (Jonhson) bands is equal to 0.6 mag (Tanvir 1997). This assumption is generally adopted (Freedman 1988) for deriving the I-band light curve on the basis of the shape of the V-band light curve and thus for reducing the number of individual measurements necessary for estimating the mean I magnitude. Amplitude ratios in different photometric bands are also interesting because as demonstrated by Coulson & Caldwell (1989) they can be adopted for discovering companions to Cepheids (see also Laney & Stobie 1993).
Figure 8 shows the predicted $`A_I/A_V`$ ratio as a function of the pulsation period for canonical, fundamental pulsators at the three adopted metallicities. Luminosity amplitudes in the I band were estimated by adopting the color-temperature relations provided by Castelli et al. (1997a,b). Data plotted in this figure show that the mean $`A_I/A_V`$ ratio ranges from $`0.64\pm 0.03`$ at Z=0.004 and Z=0.008 to $`0.65\pm 0.02`$ at Z=0.02. This result suggests that, within the intrinsic dispersions, the $`A_I/A_V`$ ratio is in average constant and slightly higher than the empirical value currently adopted. However, this finding mainly depends on stellar atmosphere models, and indeed some numerical experiments performed by artificially enhancing both the surface gravity and the effective temperature variations along the pulsation cycle caused an increase in the $`A_I/A_V`$ ratio smaller than 10%.
### 4.3 Radial velocity
Even though radial velocity amplitudes are available only for a limited number of Galactic Cepheids, this observable can supply useful information on the dynamical behavior across the instability strip. Data listed in Tables 2 and 4 show that the behavior of radial velocity curves are largely correlated to the already discussed light curves. The main difference is found in $`5M_{}`$ canonical models at Z=0.004 and Z=0.008, which show very large $`\mathrm{\Delta }u`$ values when compared with more massive models. Since the bolometric amplitudes do not present this behavior, it turns out that in these models the bolometric amplitude over a full pulsation cycle is governed by temperature variations more than by radius variations.
The top panel of Figure 9 shows the comparison between empirical pulsational velocity amplitudes for Galactic Cepheids provided by Bersier et al. (1994), Bersier & Burki (1996) and by Bersier (1999, private communication) and theoretical predictions for Z=0.02 and Z=0.008. The bottom panel of Figure 9 shows a similar comparison but with empirical data collected by Cogan (1980). In the Cogan sample we identified Cepheid variables by means of a cross-identification with the database on Galactic Cepheids provided by Fernie et al. (1995) and among them we selected the objects presenting accurate velocity amplitudes. Pulsational velocities plotted in Figure 9 were derived from the empirical radial velocity amplitudes by adopting a projection factor of 1.36, as suggested by Bersier & Burki (1996). However, we notice that in the literature quite different values were suggested, with the additional evidence that the projection factor could change when moving from short to long-period Cepheids (Gieren et al. 1989) as well as along the pulsation cycle (Sabbey et al. 1995). We adopted this value since its effect is marginal when compared to observational errors affecting non homogeneous spectroscopic measurements plotted in the bottom panel. In fact, we estimated that the difference in the velocity amplitude among objects included both in the Bersier et al. and in the Cogan samples ranges from few percent to more than 20%.
Figure 9 shows that theoretical predictions appear in reasonable agreement with observations over the whole period range. In particular, one may notice that toward longer periods -$`logP=1.5÷1.6`$\- the velocity amplitudes decrease to approximately 50 km/sec, in agreement with theoretical predictions. However, from $`logP=1.1`$ to $`logP=1.5`$ our computations predict pulsators with smaller velocity amplitudes that are marginally supported by current empirical estimates. We can hardly assess whether this discrepancy is due to a selection effect or because theoretical amplitudes are too small close to the instability boundaries. Even by accounting for the quoted spread in metallicity the discrepancy between theory and observations is still present. In fact, models at Z=0.008 predict both high and small velocity amplitudes in the period range $`logP=1.1÷1.5`$.
However, on a very general ground, data plotted in Figure 9 show that limiting amplitude models which include a proper treatment of the pulsation/convection interaction seem to solve the long-standing discrepancy between theoretical and observed velocity amplitudes. In fact, velocity amplitudes across the instability strip predicted by nonlinear radiative models are systematically larger than the observed ones (Carson & Stothers 1988; Moskalik et al. 1992).
### 4.4 Radius, gravity and temperature
Empirical estimates of radius, gravity and temperature amplitudes are unfortunately rather scanty. In the following we will refer to the last comprehensive investigations on this subject provided by Pel (1978, 1980) and based on a large set of photometric data in the Walraven system. We find that theoretical fractional radius variations -$`\mathrm{\Delta }R/R_{ph}`$\- for Z=0.02 and Z=0.008 and for periods shorter than 11 d show -in agreement with empirical data- a decrease when moving from the blue to the red edge. However, predicted values appear somehow smaller than the observed ones, since the observed $`\mathrm{\Delta }R/R_{ph}`$ values range from more than 20% close to the blue edge to 5% at the red edge, whereas theoretical predictions range from 10-15% close to the blue edge to 5% at the red edge. The origin of such a discrepancy cannot be firmly established since we lack an estimate of the observational uncertainties as well as data for long-period Cepheids. Moreover, Pel’s results relyed on old atmosphere models (Kurucz 1975) which could considerably affect estimation of these parameters.
Theoretical and observed effective gravity amplitudes -$`\mathrm{\Delta }g_e`$\- appear in reasonable qualitative agreement. The bulk of the empirical estimates by Pel (1978) for bright Cepheids with $`P<11`$ d is around $`\mathrm{\Delta }g_e`$ 0.4, and range roughly from 0.22 to 0.9. The theoretical ones attain similar values and range from 0.3 to 0.8. Obviously, spectroscopic investigations can supply tighter constraints on the accuracy of theoretical predictions.
The agreement we found for gravity amplitudes applies also to fractional temperature variations. Empirical estimates by Pel (1978) range approximately from $`\mathrm{\Delta }\mathrm{\Theta }_e`$=0.06 to 0.22, while the theoretical ones from $`\mathrm{\Delta }\mathrm{\Theta }_e`$=0.07 to 0.18, where $`\mathrm{\Theta }_e`$ = $`5040/T_e`$. This qualitative agreement is further supported by recent spectroscopic investigations on the temperature amplitudes in a small sample of Galactic Cepheids collected by Bersier, Burki, & Kurucz (1997). Since the thermal behavior along the pulsation cycle strongly depends on the coupling between pulsation and convection, such qualitative agreement in the short-period range supports, within current observational uncertainties, the treatment adopted to account for the convective transport.
## 5 Pulsational amplitudes: first overtone pulsators
In Figure 10 from top to bottom are shown bolometric and radial velocity amplitudes, fractional radius and temperature variations for first overtone pulsators. Data plotted in this figure show that in canonical models an increase in the metal content causes a decrease in the pulsation amplitudes (see also paper I). Unfortunately, the small number of unstable first overtone pulsators does not allow us to find out a clear trend concerning the change of the pulsation amplitudes from the blue to the red edge of the instability strip. Canonical, metal-poor models at $`5M_{}`$ show a linear decrease in the bolometric amplitude and in the fractional temperature variation, while amplitudes of more metal-rich and noncanonical models present across the instability strip the characteristic ”bell-shaped” variation we already found for first overtone RR Lyrae models (BCCM97).
In order to supply a useful theoretical framework for first overtone identification among field stars, the light and velocity curves of all first overtone models are presented in appendix A. Data plotted in Figures 10-21 show that the shape of both light and velocity curves of first overtone models are almost sinusoidal and resemble the empirical light curves for s-Cepheids recently observed by the EROS project in the bar of the LMC (Beaulieu et al. 1995). However, it is worth noting that the shape of the light curve of canonical models at $`M=5M_{}`$ and Z=0.004 are not sinusoidal, and indeed the rising branch is steeper than the decreasing branch. Moreover both canonical and noncanonical models located in the middle of the instability strip and with period ranging from P=1.934 d to P=3.811 d also show a well defined bump just before the luminosity maximum. This feature is also present in models at Z=0.008 with periods ranging from P=2.076 d to P=3.816 d.
Empirical light curve Fourier parameters $`\varphi _{21}`$ of s-Cepheids show a sudden jump close to P=3.2 d when plotted as a function of the pulsation period. This abrupt change was explained as a $`2:1`$ resonance between the first and the fourth overtone (Antonello & Poretti 1986; Petersen 1989). Current radiative hydrodynamical models fail to reproduce this empirical behavior, and in particular the appearance around the quoted period of the bump along the light curve (Kienzle et al. 1999, and references therein). A detailed analysis of this phenomenon is beyond the scope of this investigation. However, the evidence that the bump along the light curve of first overtone pulsators is located in the right period range suggests that nonlinear, convective models can shed new light on this long-standing problem.
CORAVEL radial velocity amplitudes for Galactic first overtone pulsators have been recently collected by Bersier et al. (1994) and by Kienzle et al. (1999, and references therein). Theoretical predictions appear once again in reasonable agreement with empirical data, since close to $`logP=0.35`$ both data sets attain values of the order of 20-25 km/sec. The same agreement is found for fractional temperature variations, and indeed both empirical and theoretical observables for Galactic Cepheids (Bersier et al. 1997) with period shorter than $`logP=0.5`$ range approximately from $`\mathrm{\Delta }T_e/T_e`$=0.05 to 0.09. However, we do not put forward the comparison between theory and observations since first overtone models need to be extended to lower stellar masses before firm conclusions on the behavior of this mode within the instability strip can be reached.
## 6 Summary and conclusions
In this paper we presented the large set of pulsational properties predicted by our classical Cepheid, limiting amplitude, nonlinear, convective models constructed by adopting three different assumptions on chemical composition. We first discussed the location of the instability strip in the H-R digram, and showed that the instability edges predicted by nonlinear models are substantially different from the linear ones. This difference is mainly caused by the nonlinear effects of the coupling between pulsation and convection which is not included in linear, nonadiabatic, convective models. For each given metallicity we find that both blue and red boundaries of fundamental pulsators appear fairly independent of the adopted ML relation, and indeed they can be approximated over a quite large luminosity range by an analytical relation between $`logL`$ and $`logT_e`$.
The occurrence of first overtone pulsators in the lower mass range is also discussed and the linear analytical relations for the blue boundaries of metal-poor structures -Z=0.004, Z=0.008- are provided as well. We also found that predicted fundamental periods, at fixed chemical composition, can be nicely fitted by an analytical relation connecting the logarithmic period to $`logM`$, $`logL`$ and $`logT_e`$ independently of the assumption about the ML relation. These findings supply straightforward theoretical support to the use of Cepheid PL relation for estimating distances, since the topology of the instability strip presents, at fixed chemical composition, a negligible dependence on the ML relation. A similar conclusion, though based on empirical evidence, was reached by Tanvir (1997, and references therein).
Theoretical predictions concerning luminosity, radius, velocity, gravity, and temperature amplitudes are discussed in connection with observational data available in the current literature. As a whole, we found a rather satisfactory agreement, within the present large observational uncertainties, between theory and observations. The exhaustive discussion on theoretical observables presented in this investigation was mainly aimed at stimulating further detailed comparison with spectroscopic and photometric data which can supply tight constraints on the adequacy and consistency of the input physics adopted for constructing nonlinear pulsation models.
We also mention that Cepheids for which estimates of both stellar mass and effective temperature (or color) are available, such as Cepheids in stellar clusters (Bono & Marconi 1997) or in binary systems (Böhm-Vitense et al. 1998, and references therein), can supply useful suggestions on the accuracy of predicted periods, and in turn on the ML relation which governs these objects. In fact, together with Cepheids in LMC and SMC stellar clusters which are a fundamental laboratory for studying their evolutionary and pulsational properties, current empirical estimates suggest that the incidence of binaries among field Cepheids ranges from 30% (Evans 1992) to more than 50% (Szabados & Pont 1998).
We conclude that the nonlinear theoretical framework appears to be the only approach which can supply a reliable description of the pulsation behavior in radial pulsators. Moreover, it turns out that limiting amplitude calculations are a fundamental requirement for estimating on a firm basis both the modal stability and the intrinsic properties to be compared with observed variable stars (see also Ya’ari & Tuchman 1999).
We are particularly grateful to D. Bersier for sending us Cepheid radial velocity data in electronic form and for new data in advance of publication as well as for a detailed reading of an early draft of this paper. We are also grateful to D. Alves, F. Caputo, D. Laney and N. Panagia for many insightful discussions on Cepheid properties. In addition, we acknowledge an anonymous referee for some useful remarks that served to improve the paper. This work was supported, in part, by the Ministero dell’Università e della Ricerca Scientifica e Tecnologica -Cofinanziamento 98- under the project ”Stellar Evolution”. Partial support by Agenzia Spaziale Italiana is also acknowledged.
## Appendix A Predicted observables
In order to provide a useful theoretical framework which accounts for the systematic properties of Cepheids to be compared with actual properties of observed variables, Tables 1-4 summarize the observables of all limiting amplitude, nonlinear, convective models we computed.
Tables 1 and 2 list both input parameters and pulsational amplitudes for first overtone and fundamental pulsators constructed by adopting a ML relation based on evolutionary models which neglect the convective core overshooting. In particular, columns 1, 2, and 3 report the stellar mass (solar units), the logarithmic luminosity (solar units), and the static effective temperature (K) adopted for each model. From left to right the other quantities listed in these tables are: 4) nonlinear period (d); 5) mean radius (solar units); 6) fractional radius oscillation i.e. $`\mathrm{\Delta }R/R_{ph}=(R^{max}R^{min})/R_{ph}`$ where $`R_{ph}`$ is the photospheric radius; 7) radial velocity amplitude (km $`s^1`$) i.e. $`\mathrm{\Delta }u=u^{max}u^{min}`$; 8) bolometric amplitude (mag.) i.e. $`\mathrm{\Delta }M_{bol}=M_{bol}^{max}M_{bol}^{min}`$; 9) logarithmic amplitude of static gravity i.e. $`\mathrm{\Delta }logg_s=logg_s^{max}logg_s^{min}`$; 10) logarithmic amplitude of effective gravity i.e. $`\mathrm{\Delta }logg_{eff}=logg_{eff}^{max}logg_{eff}^{min}`$ where $`g_{eff}=GM/R^2+du/dt`$; 11) temperature amplitude (K) i.e. $`\mathrm{\Delta }T=T^{max}T^{min}`$ where $`T`$ is the temperature of the outer boundary; 12) effective temperature amplitude (K) i.e. $`\mathrm{\Delta }T_e=T_e^{max}T_e^{min}`$ where $`T_e`$ is derived from the surface luminosity. With the exception of the effective gravity, the quantities reported in column 4 to 11 are referred to the surface zone. The mean effective gravity of the outermost layers was estimated by adopting the procedure suggested by Bono, Caputo & Stellingwerf (1994). The temperature amplitudes listed in column 10 and 11 have been rounded to the nearest 50 K. At fixed stellar mass, luminosity level and chemical composition the temperature of the blue (red) edges can be estimated by increasing (decreasing) the effective temperature of the hottest (coolest) unstable model by 50 K.
Tables 3 and 4 report the same quantities of Tables 1 and 2 but refer to envelope models constructed by adopting a ML relation typical of evolutionary models which account for a mild convective core overshooting (see paper I for details).
## Appendix B Light and velocity curves
The morphology of both light and velocity curves play a key role for assessing Cepheid pulsation modes and also for constraining the pulsation behavior of these variables across the instability strip. Figures 11-34 show the light (left panel) and velocity (right panel) variations throughout two consecutive cycles. Solid and dashed lines refer to fundamental and first overtone pulsators respectively. Figures 11-22 show the Cepheid sequences constructed by adopting a canonical ML relation, while figures 23-34 the sequences based on a noncanonical one. However, figures 11-14 as well as 23-26 ($`M/M_{}`$=5,7,9,11; Y=0.25; Z=0.004) are published in the printed edition, while figures 15-18 and 27-30 ($`M/M_{}`$=5,7,9,11; Y=0.25; Z=0.008), together with figures 19-22 and 31-34 ($`M/M_{}`$=5,7,9,11; Y=0.28; Z=0.02) are only available in the on-line edition.
The first extensive nonlinear investigations on the shape of theoretical light and velocity curves of Cepheids date back to Christy (1966, 1975 and references therein), Stobie (1969) and more recently to Carson & Stothers (1988) and Moskalik et al. (1992). Light and velocity curves presented in Figures 11-34, when compared with similar predictions available in the literature, do not show spurious features such as spikes and/or ripples throughout the pulsation cycle (see also paper I). In this context it is worth mentioning that Christy (1975) on the basis of leading physical arguments, suggested that the main Cepheid features can be correlated to the Christy parameter $`P/(R/R_{})`$. Even though Christy’s predictions were based on nonlinear, radiative, small amplitude models, current convective models support his findings concerning the systematics of Cepheid properties. In fact we found, in agreement with Christy, that the transition from first overtone to fundamental pulsators takes place close to $`P/(R/R_{})0.1`$ Moreover, our models show that the appearance of the bump along the decreasing branch takes place close to $`P/(R/R_{})0.12`$ and falls at earlier pulsation phases when moving toward longer period Cepheids. Christy predicted a similar behavior but for slightly larger $`P/(R/R_{})`$ values.
Unfortunately our models do not reach very long periods ($`P>130`$ d) and therefore we cannot assess on firm basis whether Cepheids in this region of the instability strip are characterized by an irregular behavior and by very large amplitudes. Finally, we mention that our models do not present RV Tauri characteristics, i.e. alternations of deep and shallow minima in both light and velocity curves in the period range $`2540`$ d, as suggested by Moskalik & Buchler (1991) and by Moskalik et al. (1992). Since these calculations were performed by adopting similar ML relations, we suspect that this difference is caused by different assumptions on the energy transport mechanism (radiative versus convective).
|
no-problem/9908/cond-mat9908379.html
|
ar5iv
|
text
|
# The 6-vertex model of hydrogen-bonded crystals with bond defects
## 1 Introduction
The 6-vertex model on a square lattice describes hydrogen-bonded crystals in two dimensions. Historically, it was Slater who first considered the evaluation of the residual entropy of ice, a hydrogen-bonded crystal, under the assumptions that (i) there is one hydrogen atom on each lattice edge, and (ii) there are always two hydrogen atoms near, and away from, each lattice site (the ice rule). Under these assumptions there are 6 possible hydrogen configurations at each site, and one is led to a 6-vertex model. The exact residual entropy of the “square” ice, i.e., ice on the square lattice, was obtained by Lieb , which gives rise to a numerical number surprisingly close to the experimental residual entropy of real ice (in three dimensions). The 6-vertex model is therefore an accurate description of hydrogen-bonded crystals.
In real hydrogen-bonded crystals, however, there exist bonding defects . One way through which bond defects can occur is caused by the double-well potential seen by hydrogen atoms between two lattice sites. When two hydrogens occupy the two off-center potential wells along a given lattice edge, the assumption (i) above is broken, albeit the ice rule (ii) is still intact. This leads to the fact that one, two or zero hydrogen atoms can be present on a lattice edge. Indeed, one of us has considered this possibility in a percolation model of supercooled water. The same model has later been considered by Attard and Batchelor who analyzed it using series analyses. More recently, Attard reformulated the problem as a 14-vertex model with Bjerrum bond defects, and analyzed the 14-vertex model using an independent bond approximation. Here, using a somewhat different mapping, we establish the exact equivalence of the 6-vertex model with bond defects with an 8-vertex model in an external field. As a result, we are able to analyse the exact solution in a particular parameter subspace. We also discuss the general solution of the 6-vertex model with bond defects on the Bethe lattice.
## 2 Equivalence with an 8-vertex model
Consider a square lattice $``$ of $`N`$ sites under periodic boundary conditions so that there are $`2N`$ lattice edges. The lattice is hydrogen-bonded with defects such that there can be one, two or zero hydrogen atoms on each lattice edge. As the hydrogen atoms are placed off-center on the edges, we place two Ising spins $`\sigma ,\sigma ^{}`$ on each lattice edge such that $`\sigma =1`$ denotes that the site is occupied by a hydrogen and $`\sigma =1`$ the site is empty. However, the ice rule dictates that the sum of the 4 Ising spins $`\sigma _1,\sigma _2,\sigma _3,\sigma _4`$ surrounding a square lattice sites must vanish. There are altogether six ice-rule configurations as shown in Fig. 1.
More generally, we consider the ice-rule model with weights
$$\omega _1=\omega _2=a,\omega _3=\omega _4=b,\omega _5=\omega _6=c,$$
(1)
where $`\omega _i`$ is the weight of the $`i`$th configuration shown in Fig. 1. Denote the weights (1) by $`\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)`$ where the subscripts are indexed as shown and $`\omega _1=\omega (1,1,1,1)`$, etc. The weight $`\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)`$ satisfies the “spin reversal” symmetry
$$\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)=\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)$$
(2)
and vanishes except for the six weights given in (1). To each lattice edge containing two spins $`\sigma `$ and $`\sigma ^{}`$, introduce an edge factor $`E(\sigma ,\sigma ^{})`$ to reflect the effect of bond defects. Then, the partition function of interest is
$$Z=\underset{\sigma =\pm 1}{}\underset{\mathrm{vertices}}{}\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)\underset{\mathrm{edges}}{}E(\sigma ,\sigma ^{}).$$
(3)
In the 6-vertex model without bond defects we have $`E(\sigma ,\sigma ^{})=(1\sigma \sigma ^{})/2`$ so that there is precisely one hydrogen on each lattice edge. The model considered by Attard is described by
$`E(\sigma ,\sigma ^{})`$ $`=`$ $`w_+,\sigma =\sigma ^{}=1`$ (4)
$`=`$ $`w_{},\sigma =\sigma ^{}=1`$
$`=`$ $`1,\sigma =\sigma ^{}.`$
The percolation model of is equivalent to a special case of (4) with $`w_+=w_{}=e^{2K}`$ and
$$E(\sigma ,\sigma ^{})=e^{K(\sigma \sigma ^{}+1)}=e^K(\mathrm{cosh}K)(1+z\sigma \sigma ^{}),$$
(5)
where $`z=\mathrm{tanh}K`$. For our purposes, we shall restrict our considerations to the percolation model (5).
Attard and Batchelor adopted an arrow representation for the hydrogen configurations which, due to the occurrence of defects, led to a 14-vertex model with Bjerrum defects. A weak-graph transformation is then carried out for the 14-vertex model. Here, we expand the partition function directly. Substituting (5) into (3) and expanding the second product over the edges of $``$, we obtain an expansion of $`2^{2N}`$ terms. To each term in the expansion we associate a bond graph by drawing bonds on those edges corresponding to the $`z`$ factors contained in the term. This leads to a 16-vertex model on $``$. Besides an overall Boltzman factor $`(e^K\mathrm{cosh}K)^{2N}`$, the 16-vertex model has vertex weights
$$W=\underset{\sigma _1\sigma _2\sigma _3\sigma _4}{}\left(\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)(\sqrt{z}\sigma _i)\right),$$
(6)
where the product is taken over those incident edges with bonds. The symmetry relation (2) now implies that $`W=0`$ whenever there are an odd number of incident bonds, and the 16-vertex model becomes an 8-vertex model.
Using the bond configurations of the 8-vertex model shown in Fig. 2, it is straightforward to deduce using (6) the following vertex weights:
$`W_1`$ $`=`$ $`{\displaystyle \omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)}=2(a+b+c)`$
$`W_2`$ $`=`$ $`z^2{\displaystyle \sigma _1\sigma _2\sigma _3\sigma _4\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)}=2z^2(a+b+c)`$
$`W_3`$ $`=`$ $`W_4=z{\displaystyle \sigma _2\sigma _4\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)}=2z(ab+c)`$
$`W_5`$ $`=`$ $`W_6=z{\displaystyle \sigma _1\sigma _2\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)}=2z(a+bc)`$
$`W_7`$ $`=`$ $`W_8=z{\displaystyle \sigma _2\sigma _3\omega (\sigma _1,\sigma _2,\sigma _3,\sigma _4)}=2z(abc),`$ (7)
where $`W_i`$ is the vertex weight of the $`i`$th vertex and we have used the fact that in the non-vanishing $`\omega `$’s we have $`\sigma _1\sigma _2\sigma _3\sigma _4=1`$. Now, in an 8-vertex model configurations, vertices 3 and 4, 5 and 6, and 7 and 8, always occur in pairs and/or in even numbers. Therefore we can conveniently replace the relevant weights by their absolute values and arrive at, after dividing all weights by a common factor $`z`$,
$`W_1`$ $`=`$ $`z^1(a+b+c)`$
$`W_2`$ $`=`$ $`z(a+b+c)`$
$`W_3`$ $`=`$ $`W_4=|ab+c|`$
$`W_5`$ $`=`$ $`W_6=|a+bc|`$
$`W_7`$ $`=`$ $`W_8=|abc|.`$ (8)
The vertex weights (8) describes an 8-vertex model in an external electric field $`h=(\mathrm{ln}z)/2`$ in both the vertical and horizontal directions .
More generally, if we allow different values of $`w_+=w_{}=w_i,i=1,2`$ in (4) and (5) for the horizontal and vertical edges respectively, and write $`h=(\mathrm{ln}z_1)/2`$ and $`v=(\mathrm{ln}z_2)/2`$, where $`z_i=(w_i1)/(w_i+1),i=1,2`$, then one arrives at an 8-vertex model with weights
$`W_1`$ $`=`$ $`e^{h+v}(a+b+c)`$
$`W_2`$ $`=`$ $`e^{hv}(a+b+c)`$
$`W_3`$ $`=`$ $`e^{hv}|ab+c|`$
$`W_4`$ $`=`$ $`e^{vh}|ab+c|`$
$`W_5`$ $`=`$ $`W_6=|a+bc|`$
$`W_7`$ $`=`$ $`W_8=|abc|.`$ (9)
In ensuing discussions we shall consider the general model (9).
## 3 The free-fermion solution
The free-fermion model is defined as a particular case of the eight-vertex model in which the vertex weights satisfy the relation
$$W_1W_2+W_3W_4=W_5W_6+W_7W_8,$$
(10)
a condition equivalent to the consideration of a noninteracting many-fermion system in an $`S`$-matrix formulation of the 8-vertex model . In the present case the free-fermion condition (10) is satisfied when either $`a=0`$ or $`b=0`$ . Without the loss of generality, we consider $`a=0`$ and $`b>c`$.
The closed expression for the free energy of the free-fermion model is well-known and after some algebraic manipulation using results of , we obtain
$$\beta f=\underset{N\mathrm{}}{lim}\frac{1}{N}\mathrm{ln}Z=\mathrm{ln}(b+c)+\frac{1}{4\pi }_0^{2\pi }𝑑\varphi \mathrm{ln}(A+Q^{1/2}),$$
(11)
where
$`Q`$ $`=`$ $`[\mathrm{sinh}(2v+2h)+k^2\mathrm{sinh}(2v2h)+2k\mathrm{cosh}2v\mathrm{cos}\varphi ]^2+4k^2\mathrm{sin}\varphi ^2,`$
$`A`$ $`=`$ $`\mathrm{cosh}(2v+2h)+k^2\mathrm{cosh}(2v2h)+2k\mathrm{sinh}2v\mathrm{cos}\varphi `$ (12)
with $`k=(bc)(b+c)`$.
The critical condition of the free-fermion model is given by
$$W_1+W_2+W_3+W_4=2\mathrm{max}\{W_1,W_2,W_3,W_4\},$$
(13)
where
$`W_1`$ $`=`$ $`e^{h+v}(b+c)`$
$`W_2`$ $`=`$ $`e^{hv}(b+c)`$
$`W_3`$ $`=`$ $`e^{hv}(bc)`$
$`W_4`$ $`=`$ $`e^{vh}(bc).`$ (14)
Thus, the $`h`$-$`v`$ plane is divided into four regions depending on which vertex $`1`$, $`2`$, $`3`$ or $`4`$ has the largest weight. Denoting the four regions by $`I`$, $`II`$, $`III`$ and $`IV`$ respectively, as shown in Fig. 3, the critical condition (13) can be rewritten as
$`{\displaystyle \frac{b}{c}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{tanh}v+e^{2h}}{1e^{2h}\mathrm{tanh}v}},\text{region}I,`$
$`{\displaystyle \frac{b}{c}}`$ $`=`$ $`{\displaystyle \frac{1e^{2h}\mathrm{tanh}v}{\mathrm{tanh}v+e^{2h}}},\text{region}II,`$
$`{\displaystyle \frac{b}{c}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{tanh}ve^{2h}}{1+e^{2h}\mathrm{tanh}v}},\text{region}III,`$
$`{\displaystyle \frac{b}{c}}`$ $`=`$ $`{\displaystyle \frac{1+e^{2h}\mathrm{tanh}v}{\mathrm{tanh}ve^{2h}}},\text{region}IV.`$ (15)
The critical condition (15) is plotted in Fig. $`4a4d`$ for four specific values of $`b/c`$. Generally, the free energy exhibits a logarithmic singularity at the phase boundaries with exponents $`\alpha =\alpha ^{}=0`$ . When $`b=c`$ for which $`Q`$ is a complete square, however, we have
$`\beta f`$ $`=`$ $`\mathrm{max}\{\mathrm{ln}W_1,\mathrm{ln}W_2\}`$ (16)
$`=`$ $`\mathrm{ln}(2b)+|h+v|`$
and the phase boundary $`h+v=0`$ separates the two frozen states $`W_1`$ and $`W_2`$.
## 4 The Bethe-like lattice
Now, we consider the 6-vertex model with bond defects on a Bethe-like lattice with plaquettes as shown in Fig. 5. The study of systems in the Bethe-like lattices is an alternative approach to the usual mean-field theory. The main features of the model under investigation will be obtained by studying the properties of the free energy.
A free energy in a region deep inside a Bethe-like lattice must be carefully defined. It cannot be obtained by directly evaluating the logarithm of the partition function in which the contribution from the outside of this region is not negligible and as result the system exhibits an unusual type of phase transitions without long-range order . Recently, a method for the surface independent free energy calculation is presented . The free energy $`f_{\mathrm{}}`$ per plaquette of our model is expressed as
$$\beta f_{\mathrm{}}=\underset{n\mathrm{}}{lim}\frac{1}{2}\left(\mathrm{ln}Z_n3\mathrm{ln}Z_{n1}\right)$$
(17)
where $`Z_n`$ and $`Z_{n1}`$ is the partition functions of the 6-vertex model with bond defect on the Bethe-like lattice consist of $`n`$ and $`n1`$ generations respectively.
The calculation on the Bethe-like lattice is based on a recursion method . When the tree is cut at the central plaquette, it is separate into 4 branches, each of which contains 3 branches. Then the partition function of interest (3) can be written as follows:
$$Z_n=\left(\frac{\omega +1}{2}\right)^{N_b^{(n)}}\underset{\{\sigma _{0i}\}}{}\omega (\sigma _{01},\sigma _{02},\sigma _{03},\sigma _{04})g_n(\sigma _{01})g_n(\sigma _{02})g_n(\sigma _{03})g_n(\sigma _{04}),$$
(18)
where $`N_b^{(n)}=2(3^n1)`$ is the number of bonds, $`n`$ is the number of generations and $`g_n(\sigma _{0i})`$ is in fact the partition function of a branch nearest to the $`0i`$ site.
Each branch, in turn, can be cut along any site of the first generation. then the expressions for $`g_n(\sigma _{0i})`$ can therefore be written in the form:
$$g_{n+1}(\sigma _{01})=\underset{\{\sigma _{1i}\}}{}(1+z\sigma _{01}\sigma _{13})\omega (\sigma _{11},\sigma _{12},\sigma _{13},\sigma _{14})g_n(\sigma _{11})g_n(\sigma _{12})g_n(\sigma _{14})$$
(19)
After dividing $`g_n()`$ by $`g_n(+)`$, we obtain a recursion relation for $`x_n=g_n()/g_n(+)`$. Let us consider the case when series of solution of recursion relation converge to a stable point at $`n\mathrm{}`$, namely,
$$\underset{n\mathrm{}}{lim}x_n=x.$$
We obtain the following equation:
$$x=\frac{(1z)x^2+(1+z)x}{(1+z)x^2+(1z)x}.$$
(20)
We are now in a position to compute the free energy per plaquette of our model. Using Eqs. (17), (18), (19) and (20), the expression for the free energy functional can be written as
$`\beta f_{\mathrm{}}`$ $`=`$ $`\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{2}}\left(N_b^{(n)}3N_b^{(n1)}\right)\mathrm{ln}{\displaystyle \frac{\omega +1}{2}}`$
$`+`$ $`\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{2}}\left[\mathrm{ln}\mathrm{\Phi }(x_{(n1)})+\mathrm{ln}\mathrm{\Psi }(x_{(n)})3\mathrm{ln}\mathrm{\Psi }(x_{(n1)})\right]`$
where
$$\mathrm{\Psi }(x)=2(a+b+c)x^2\text{and}\mathrm{\Phi }(x)=(a+b+c)^4x^4[(1+z)x+1z]^4$$
It easy to see that $`N_b^{(n)}3N_b^{(n1)}=4`$ for all $`n`$. Thus the free energy per plaquette can be finally written as
$$\beta f_{\mathrm{}}=\mathrm{ln}\frac{(\omega +1)^2}{2}(a+b+c)+\mathrm{ln}\frac{(1+z)x+1z}{2}.$$
(22)
Together with the expression (20) for $`x`$ it gives the free energy per plaquette of the 6-vertex model with bond defects.
The equation of state (20) always has $`x=1`$ as fixed point solution. In this case the free energy per plaquette is
$$\beta f_{\mathrm{}}=\mathrm{ln}\frac{(w+1)^2}{2}(a+b+c)$$
(23)
In the case of $`a=b=c=1`$, we recover the result obtained by Attard and Batchelor for the six-vertex model with bond defect in the mean-field (independent vertex) framework,
$$\beta f_{\mathrm{}}=\mathrm{ln}\frac{3}{2}(w+1)^2,$$
which reduces to Pauling’s estimate when $`w=0`$ .
For the specific heat, we obtain
$`C`$ $`=`$ $`{\displaystyle \frac{d}{dT}}\left[T^2{\displaystyle \frac{d}{dT}}(\beta f_{\mathrm{}})\right]`$ (24)
$`=`$ $`{\displaystyle \frac{2w}{(w+1)^2}}(\mathrm{ln}w)^2+{\displaystyle \frac{ab(\mathrm{ln}a/b)^2+ac(\mathrm{ln}a/c)^2+bc(\mathrm{ln}b/c)^2}{(a+b+c)^2}},`$
or, explicitly,
$$C=\frac{1}{T^2}\frac{2K^2}{(\mathrm{cosh}K/T)^2}+\frac{\epsilon _{ab}^2e^{\epsilon _{ab}/T}+\epsilon _{ac}^2e^{\epsilon _{ac}/T}+(\epsilon _{ab}\epsilon _{ac})^2e^{(\epsilon _{ab}+\epsilon _{ac})/T}}{T^2(1+e^{\epsilon _{ab}/T}+e^{\epsilon _{ac}/T})^2}$$
(25)
where $`\epsilon _{ab}=\epsilon _a\epsilon _b,\epsilon _{ac}=\epsilon _a\epsilon _c`$ and
$$a=e^{\epsilon _a/T},b=e^{\epsilon _b/T},c=e^{\epsilon _c/T}.$$
The specific heat versus $`T/K`$ for $`a=b`$ is plotted in Fig. 6.
## 5 Summary and discussion
In this paper we have considered a 6-vertex model of hydrogen-bonded crystals with bond defects. We have established the exact equivalence of the 6-vertex model with bond defects with an 8-vertex model in an external electric field. Using this equivalence we solve exactly our model in the free-fermion subspace. We also obtain the exact solution of the 6-vertex model with bond defects on a Bethe-like lattice.
In , one of us has used the percolation representation of the hydrogen-bonded model to argue that the specific heat of the system will increase as the temperature decreases from very high temperature. Figure 6 indeed shows such behavior. However, the specific heat in Fig. 6 does not diverge. It is of interest to calculate the specific heat for the square or diamond lattices by Monte Carlo method. We expect to find singular behavior of specific heat in such systems.
ACKNOWLEDGMENTS
This work was partly supported by the National Science Council of the Republic of China (Taiwan) under grant number NSC 89-2112-M-001-005. The work of FYW is supported in part by the National Science Foundation grant DMR-9614170. FYW thanks C. K. Hu for the hospitality at the Academia Sinica and T. K. Lee for the hospitality at the National Center for Theoretical Sciences where this work is completed.
|
no-problem/9908/cond-mat9908023.html
|
ar5iv
|
text
|
# Non-exponential kinetic behaviour of confined water
## Abstract
We present the results of molecular dynamics simulations of SPC/E water confined in a realistic model of a silica pore. The single-particle dynamics have been studied at ambient temperature for different hydration levels. The confinement near the hydrophilic surface makes the dynamic behaviour of the liquid strongly dependent on the hydration level. Upon decrease of the number of water molecules in the pore we observe the onset of a slow dynamics due to the “cage effect”. The conventional picture of a stochastic single-particle diffusion process thus looses its validity.
The normal limit of supercooling and the possibility of vitrification of water is the subject of a longstanding scientific debate . At $`T_H236`$ K homogeneous nucleation, due to the presence of impurities, drives bulk supercooled water into its solid, crystalline phase. Experiments sustain the hypothesis that the amorphous phase can be connected to the normal liquid phase through a reversible thermodynamic path . Recent molecular dynamics (MD) simulations of bulk supercooled SPC/E water also showed a kinetic glass transition as predicted by mode coupling theory (MCT) at a critical temperature $`T_CT_S`$ , where $`T_S`$ is the singular temperature of water , which is $`T_S=228K`$ or $`49`$ degrees below the temperature of maximum density.
A comparison of the behaviour of the bulk liquid with the same liquid in a confined environment is highly interesting for the development of both biological and industrial applications. Understanding how the dynamics of liquid water is perturbed by the interaction with hydrophilic or hydrophobic substrates at various levels of hydration lays the physical basis for the prediction of stability and enzymatic activity of biologically important macromolecules . Many experimental studies have been performed on confined and interfacial water. They all showed evidence of a substantial degree of slowing down of water when confined in the proximity of a polar surface . Particularly significant in this context are the indications of a transition of adsorbed water to a glassy state, which is supposedly driven by the protein surface . MD calculations found a reduced self diffusion coefficient of water in contact with solid hydrophilic surfaces, when compared to bulk water . Moreover, a recent simulation study of water close to the surface of a protein evidenced a typical spectral glassy anomaly, the so called boson peak , which the authors related to protein-solvent coupling. These and other studies suggest the possibility of a common underlying molecular mechanism for the slowing down of the single particle dynamics of interfacial water. Nevertheless, many questions concerning the substrate-induced perturbation of the dynamics of water are still unanswered. Some of these questions can be addressed by MD simulation techniques due to their power as a very versatile microscopic tool.
In this letter, we present the results of MD simulations of SPC/E water confined in a cylindrical silica pore. The confining system has been carefully designed to mimic the pores of real Vycor glass. We chose this particular system, because Vycor is characterized by a quite sharp distribution of pore sizes with a small average diameter ($`40\pm 5`$ Å), because the pore size does not depend on the hydration level (no swelling occurs), and because the surface is strongly hydrophilic. For the same reasons, this system has been studied extensively by a variety of experiments . Here, we focus on the role of hydration. In particular, we try to understand the extent to which the changes in the system dynamics upon lowering the hydration level in a hydrophilic silica pore are equivalent to supercooling the bulk phase . To our knowledge, this is the first time that the dynamics of water in a realistic pore model is calculated in such detail. The results presented here can help to shed light on the microscopic mechanism governing the dynamics of water close to a hydrophilic surface.
The MD simulations started from a silica cube containing a cylindrical cavity. In the system preparation step a vitreous silica cubic cell was first created by melting a $`\beta `$-cristobalite crystalline structure and then quenching the liquid down to ambient temperature . Inside the glass, which is formed by this process, a cylindrical cavity with a radius of $`R=20`$ Å is created by appropriate removal of atoms. Next, all those remaining silicon atoms which are bonded to less than four oxygens are removed. With this procedure two kinds of oxygen atoms are obtained, nonbridging oxygens (NBOs), bonded to only one silicon ion, and bridging oxygens (BOs), bonded to at least two silicons. The NBOs are saturated with hydrogen atoms, similar to the experimental process before hydration , which leads to free hydroxyl groups on the pore surface. With this process, we create a cavity with a rough surface, whose geometry, size, and microscopic structure are typical for a pore in Vycor glass. The block of silica provides a rigid matrix for the simulation of SPC/E water molecules which are then introduced into the pore. Water interacts with the atoms of the rigid matrix by means of an empirical potential model , where different fractional charges are assigned to BOs ($`0.629|e|`$), NBOs ($`0.533|e|`$), silicon ($`+1.283|e|`$) and surface hydrogen atoms ($`+0.206|e|`$). Additional Lennard-Jones potential functions between the oxygen sites of water and BOs and NBOs complement the water-Vycor interaction . Since the simulations are time-consuming, we have used the shifted force method with a cutoff at $`9`$ Å. We did check, however, that the use of a larger cutoff or Ewald summations does not change the trend of the obtained results .
There is experimental evidence that the average density of water confined in real Vycor glass at full hydration is $`11\%`$ less than the density of bulk water at ambient conditions . In order to match this value 2700 water molecules are needed in our pore, this number defines our full hydration. We have conducted simulations in the NVE ensemble at ambient temperature ($`T=298K`$) for four different hydrations of the pore. Simulations with 500, 1000, 1500, 2000 and 2600 water molecules correspond, based on the above consideration, to hydration levels of 19%, 37%, 56%, 74% and 96%, respectively. The timestep for the integration of the molecular trajectories is $`1`$ fs. Thermal equilibrium was achieved through the use of a Berendsen thermostat ; the equilibration process has been carefully monitored via the time dependence of the potential energy. Further details of the simulation can be found in ref. , where static properties of the system are studied, and where it is shown that the interaction model used here leads to satisfactory agreement with the experimental site-site pair correlation functions.
We now discuss some of the most significant single particle dynamical properties of water from our simulations for different hydration levels at ambient temperature. In Fig. 3 we show the mean square displacement (MSD) of the center of mass motion along the non-confined z-direction. After an initial ballistic diffusion, water molecules in the bulk phase at ambient temperature (not shown) enter the Brownian diffusive regime (characterized by a slope of 1 in the log-log plot). Our data show at intermediate times an obvious flattening of the MSD curve, which increases with decreasing hydration level. This is the signature of the so-called cage effect in the system, which describes the trapping of molecules in the cage formed by their nearest neighbors at intermediate times after the initial ballistic regime. Once the cage has relaxed, molecules enter the normal diffusive regime. In the inset of Fig. 3 we report the self diffusion coefficient, $`D`$, as calculated from the slope of the MSD in the diffusive regime. A substantial decrease of $`D`$ with decreasing hydration level is evident. All values of $`D`$ are substantially lower than the one of SPC/E bulk water at the same temperature.
Figure 3 shows the single particle intermediate scattering function (ISF), $`F_S(Q,t)`$, calculated for the center of mass motion at the maximum of the oxygen-oxygen structure factor for all hydration levels investigated. As the hydration is lowered, the ISF displays an increasingly pronounced shoulder around 1 ps, which is the characteristic of a two-step relaxation process. The long time tail is highly non-exponential for all hydration levels . We are able to fit a combination of the Kohlrausch-William-Watts (KWW) relaxation function plus a gaussian term
$$F_S(Q,t)=\left[1A(Q)\right]\mathrm{exp}\left[\left(t/\tau _s\right)^2\right]+A(Q)\mathrm{exp}\left[\left(t/\tau _l\right)^\beta \right]$$
(1)
only to the ISF at the lowest hydration level. $`A(Q)=e^{a^2q^2/3}`$ is the Debye-Waller factor arising from the cage effect. $`a`$ is the amplitude of the vibration of the molecule inside the cage created by the potential barrier of the nearest neighbours. $`\tau _s`$ and $`\tau _l`$ are, respectively, the short and the long relaxation times. The fitted curve is shown as the dashed line in Fig. 3. For bulk water, this function fitted the center of mass ISF for several temperatures . For the shown fit we obtain a cage radius $`a0.44`$ Å, which is similar to the radius obtained for bulk supercooled water, $`a0.5`$ Å. The lower radius obtained for confined water may be due to the slightly higher density of water close to the surface . The short relaxation time $`\tau _s0.14`$ ps is again comparable to the bulk value $`\tau _s0.2`$ ps. The long time tail is characterized by $`\beta =0.35`$ and $`\tau _l=356`$ $`ps`$.
For the higher hydration levels we have calculated the ISF separately for the first two layers of molecules close to the surface and for the remaining molecules. This choice is based on the shape of the density profile, which shows a double layer of water molecules close to the surface with density higher than the average . This double layer extends up to $`5`$ Å from the surface. The inset of Fig. 3 shows, for $`N_W=2600`$ (96% hydration), the two contributions. While the molecules in the adsorbed layer relax very slowly, the decay of the ISF of the remaining molecules is much faster and very well described by eq.1 (bottom curves). From the fit we extract $`\beta =0.71`$, $`\tau _l=0.5`$ $`ps`$, $`\tau _s=0.17`$ ps, and $`a=0.54`$ Å. While the $`a`$, $`\tau _s`$ and $`\tau _l`$ are similar to the values found for bulk water at ambient conditions, the $`\beta `$ differs from the value, $`1`$, of the bulk. It turns out that the behaviour of molecules belonging to the first hydration layers change as a complete surface coverage is achieved. In fact for $`N_W=500`$ the surface coverage is not complete and patches of water molecules are visible along the pore surface. A stretched exponential function is able to account also for the dynamics of clusters in a frozen environment . In Fig. 3 $`F_S(Q,t)`$ at 19 % hydration is shown for several values of $`Q`$ ranging from 0.5 to 4 $`\mathrm{\AA }^1`$. This figure very clearly demonstrates the overshooting of the ISF around 1 ps, which was also visible in Fig. 3.
The maximum can be observed both in the $`z`$ direction along the pore axis and in the $`xy`$ plane, as is visible in the inset. The intermediate maximum of the ISF has been related to the so-called boson peak (BP) . The BP is an excess of vibrational modes present in many glasses at frequencies around $`1`$ THz. When this glassy anomaly appears in a liquid phase, it is usually considered as a precursor to the actual glass transition. We also performed a shell analysis of the dynamic behaviour at the higher hydration levels, and we found in these cases that the contribution to the BP in our simulation comes only from water molecules which are not in the first layer. This fact, and the fact that our substrate is a rigid framework, is an indication that the BP is a feature of liquid water, which is not induced by the substrate dynamics.
Experimental signatures of our observations were detected in some confined hydrogen bonded complex liquids . An analysis of quasi-elastic neutron scattering data of the water / Vycor system is consistent with a highly non-exponential relaxation behaviour . The BP for the water / Vycor system has been detected recently at energies around $`3.5`$ meV .
In summary, we presented MD results concerning the single particle dynamics of liquid water confined in a silica pore. We found evidence of glassy behaviour already at room temperature. On lowering the hydration level of water inside the pore, the MSD flattens at intermediate times due to a cage effect. Correspondingly, the ISF displays a two step relaxation behaviour with a highly non-exponential slow relaxation. Such behaviour is typical for a glass forming liquid approaching its glass transition point. At the lowest hydration level, a KWW function could be fitted to the ISF. In this particular system, all phenomena are strongly influenced by the substrate. Specifically the interaction with the hydrophilic surface seems to drive the liquid closer to the glass transition point. This interaction leads to a significant slowing down of water molecules close to the substrate. The strong distortion of the hydrogen bond network close to the surface will lead to a strong variation of the cage structure from one molecule to the other, which, in contrast to bulk water, might mask the simple behaviour suggested by mode coupling theory. The dynamic results moreover support the notion of two quite distinct subsets of molecules. One is bound directly to the substrate surface, the other consists of the remaining water molecules in the pore center. Near the substrate surface, the local water density is slightly higher , and the dynamics is severely slowed down at all hydration levels, whereas the retardation of the slow relaxation process is much less pronounced for the inner water shells, whose contribution to the slow relaxation in the ISF could be fitted to a KWW law. The shouldering of the relaxation law is indeed determined by the propagation of the perturbation of the substrate to the water molecules inside the pore. With the increase of the hydration level the effect is overthrown by the increasing number of molecules located far from the surface.
The confinement also leads to the formation of an overshooting related to the BP at the lowest hydration level. Since the BP is found typically in strong glass formers , the appearance of the intermediate maximum in the ISF at the lowest hydration would favor the hypothesis of a conversion to a more strong glass-forming behaviour of water as the level of hydration is lowered.
\***
P.G. wishes to thank S.-H. Chen for all the interesting discussions on this subject.
|
no-problem/9908/cond-mat9908462.html
|
ar5iv
|
text
|
# Landau quantization and equatorial states on a surface of a nanosphere.
## Abstract
The Landau quantization for the electron gas on a surface of sphere is considered. We show that in the regime of strong fields the lowest energy states are those with magnetic quantum numbers $`m`$ of order of $`\mathrm{\Phi }/\mathrm{\Phi }_0`$, the number of magnetic flux quanta piercing the sphere. For the electron gas of low density (semiconducting situation), it leads to the formation of the electronic stripe on the equator of the sphere in high fields.
The electronic properties of cylindrical and spherical nanosize objects attract much of the theoretical interest last years. This interest is mostly related to the physics of carbon macromolecules and, particularly, to the transport properties of carbon nanotubes. One meets the spherical nanosize objects in the studies of the nonlinear optical response in composite materials , of simple metal clusters and of the photonic crystals on the base of synthetic opals . In the most of these studies, the coating of the nano-sphere is characterized by an effective dielectric function. It was noted however that this approach should be revised if the coating of a sphere has a width of a few monolayers, which limit is allowed by modern technologies.
In the recent papers we considered the electron gas on a sphere. We showed that various correlation functions in such gas exhibit maxima when the electrons are at the points-antipodes (north and south poles). The exact solution was found for a problem in the uniform magnetic field and the limits of weak and high fields were investigated. In the high-field regime the formation of Landau levels was shown. The complexity of the special functions describing the exact solution, however, complicates the analysis of the physical picture in the high-field regime. In this paper we explore qualitative arguments, supported by the numerical calculations, to clarify the issue. We find that the minimum energy to the Hamiltonian is provided by the electronic states located near the equator of the sphere. For low densities of the gas one thus expects, that the high field pushes the electrons towards the equator and forms an electronic ring there.
We consider the electron gas moving within a thin layer on a surface of the sphere of radius $`r_0`$. We assume the Hamiltonian of the form $`=\frac{^2}{2m_e}+U(r),`$ with $`m_e`$ the electron mass. The chemical potential $`\mu `$ defines the total number of electrons $`N`$ (with one projection of spin) and the areal density $`\nu =N/(4\pi r_0^2)`$. The confining quantum-well potential $`U(r)`$ restricts the radial motion within the thin layer $`\delta rr_0`$. We are interested in the case $`\delta r<\nu ^{1/2}`$, when the first excited state of the radial motion lies above the chemical potential. Then one can ignore the radial component of the wave function and put $`r=r_0`$ in the remaining angular part of the Hamiltonian. In the absence of the magnetic field we have $`_\mathrm{\Omega }^{(0)}=(2m_er_0^2)^1\mathrm{\Delta }_\mathrm{\Omega }`$ with $`\mathrm{\Delta }_\mathrm{\Omega }`$ being the angular part of the Laplacian. The solutions to this Hamiltonian are the spherical harmonics $`Y_{lm}`$ and the spectrum is that of a free rotator model :
$`\mathrm{\Psi }^{(0)}(\theta ,\varphi )=r_0^1Y_{lm}(\theta ,\varphi ),E_l^{(0)}=(2m_er_0^2)^1l(l+1).`$
In the presence of the uniform magnetic field $`𝐁`$ directed towards a north pole of the sphere ($`\theta =0`$) we choose the gauge of the vector potential as $`𝐀=\frac{1}{2}(𝐁\times 𝐫)`$. Then the angular part $`_\mathrm{\Omega }`$ of the Hamiltonian $`=\frac{1}{2m_e}(i+e𝐀)^2+U(r)`$ acquires the form
$$_\mathrm{\Omega }=(2m_er_0^2)^1[\mathrm{\Delta }_\mathrm{\Omega }+2ip\frac{}{\varphi }+p^2\mathrm{sin}^2\theta ]$$
(1)
The ruling parameter here is $`p=\pi Br_0^2/\mathrm{\Phi }_0`$ with the magnetic flux quantum $`\mathrm{\Phi }_0=210^{15}`$ T$``$m<sup>2</sup>. For a sphere of radius $`r_0=100`$nm one has $`p=1`$ at the field $`B600`$Oe. The solutions of (1) are given by the oblate (angular) spheroidal functions and were analyzed in to some detail.
In the weak-field regime, $`p1`$, the jumps in the static magnetic susceptibility $`\chi `$ at half-integer $`p`$ were demonstrated. The amplitude of these jumps is parametrically larger than the Pauli spin contribution and decreases with the increase of $`p`$. It was shown that the weak-field regime ends at $`p^2p_c^2=2\sqrt{N}`$. For $`r_0100`$nm (the case of opals) and in the metallic situation, e.g., at the densities $`\nu 10^{14}`$cm<sup>-2</sup>, we have $`N10^5`$ and $`p_c30`$.
On the other hand, at the lower (semiconducting) densities, $`\nu 10^{10}`$cm<sup>-2</sup>, we have $`N10`$. Formally in this case $`p_c3`$, i.e. the field is not small already at $`p`$ of order of unity. The jumps in the susceptibility were predicted in at the assumption $`\sqrt{N}1`$ which is violated in the latter case of lower $`\nu `$. At the same time the experimentally accessible fields of order of $`6`$T result in $`p100`$ for the spheres with $`r_0100`$nm, i.e. we come into the strong field regime.
For strong fields, $`p\mathrm{}`$, one observes the eventual formation of the Landau levels (LL). The spherical geometry brings into the problem some peculiarities which were partly discussed in . First is the incomplete restructuring of the spectrum into the LL scheme. This restructuring takes place only for levels with initial momentum $`l`$ lower than $`p`$, and the field remains weak for the levels with $`l>p^2`$. Secondly, the field-induced two-well potential $`p^2\mathrm{sin}^2\theta `$ in (1) localizes the electron states with moderate magnetic quantum numbers $`m`$ ($`|m|p`$) near the poles $`\theta =0`$ and $`\theta =\pi `$. As a result, the correlations within one hemisphere only survive. Specifically, if an electron was initially in the northern hemisphere, then the probability to find it in the southern hemisphere is exponentially small.
At the same time, the spherical geometry produces yet another effect which is to be discussed here. The effective potential in the strong-field regime can be written as
$`U_{eff}(\theta )`$ $`=`$ $`{\displaystyle \frac{p\omega _c}{4}}\left({\displaystyle \frac{m/p}{\mathrm{sin}\theta }}+\mathrm{sin}\theta \right)^2`$ (2)
with the cyclotron frequency $`\omega _c=eB/m_e`$. At small negative $`m`$ we have two minima of $`U_{eff}(\theta )`$ near $`\theta _0=\mathrm{arcsin}(\sqrt{|m|/p})`$ and $`\theta _0=\pi \mathrm{arcsin}(\sqrt{|m|/p})`$ where we expand
$$U_{eff}(\theta _0+\theta )(\omega _cp\mathrm{cos}^2\theta _0)\theta ^2$$
(3)
Rescaling here $`\theta x/\sqrt{2p|\mathrm{cos}\theta _0|}`$ we arrive at the quantum oscillator problem of the form
$`{\displaystyle \frac{\omega _c|\mathrm{cos}\theta _0|}{2}}\left(d^2/dx^2+x^2\right),`$
i.e. the well-known Landau quantization. As long as $`|\mathrm{cos}\theta _0|1`$, the wave-functions are extended at the scale $`|\theta \theta _0|p^{1/2}`$. The possibility of quantum tunneling between $`\theta _0`$ and $`\pi \theta _0`$ produces the exponentially small splitting between the states centered at these points.
Thus we see that the energy levels are labeled by two quantum numbers, magnetic number $`m`$ and LL number $`n`$, with approximate double degeneracy for given $`m,n`$.
This simple picture becomes inadequate, when $`|m|p`$ and $`\theta _0\pi /2`$. In this case the harmonic potential in (3) weakens, which makes necessary the consideration of the fourth-order terms in the expansion. We have in this case
$$U_{eff}(\theta +\pi /2)\frac{\omega _cp}{4}\theta ^4,|m|=p$$
(4)
Rescaling now $`\theta xp^{1/3}`$ we arrive at the following Schrödinger equation :
$`{\displaystyle \frac{\omega _c}{4p^{1/3}}}\left(d^2/dx^2+x^4\right)\psi =E\psi ,`$
The solution of the last equation apparently is not known . For our purposes it suffices to note that the energy scales as $`\omega _c/p^{1/3}\omega _c`$ and the wave-functions on the equator extend on the scale $`|\theta \pi /2|p^{1/3}`$. In addition, we have no situation with two-well potential now and the energy levels are separated by the same scale $`\omega _c/p^{1/3}`$. The crossover between Eq.(3) and Eq.(4) takes place at $`1|m|/pp^{2/3}`$.
With the further increase of $`|m|`$, at $`|m|>p`$, the minimum value of $`U_{eff}`$ is found at $`\theta =\pi /2`$ and increases rapidly with $`|m|`$. In this case the energy levels $`E_{lm}\omega _c(m+p)^2/p`$ and thus lie well above those with $`|m|<p`$.
We illustrate these qualitative results by the numerical calculations. We found the spectrum of Eq.(1) by diagonalizing $`_\mathrm{\Omega }`$ in the basis of Legendre polynomials $`P_{m+n}^m(\mathrm{cos}\theta )`$ with $`0n200`$ . The results are shown in the Fig. 1. One verifies that the “equatorial” states with $`|m|p1`$ provide the minimum eigenvalues to the Hamiltonian. This dependence of the energy level scheme on $`m`$ is probably of minor importance, if we consider the situation of a metal, when the chemical potential lies well above the bottom of the conduction band ($`\mu \omega _c`$). In this case the number of electrons $`N`$ on the sphere is expected to exceed the number of flux quanta $`p`$. Then several Landau levels are occupied (several lines in the Fig. 1) and the electrons are distributed upon the whole sphere.
At the same time, the semiconducting coating of the sphere can lead to a different result. Indeed, if the cyclotron frequency $`\omega _c`$ is enough high, then the “polar” states with small negative $`m`$ are poorly occupied. Meanwhile, the “equatorial” states with the energies $`\omega _cp^{1/3}`$ are occupied to a larger extent. This produces the effective ring on the equator of the sphere. The criterion for this phenomenon is $`Np`$ or, equivalently, $`B\mathrm{\Phi }_0\nu `$. Note that the latter inequalities correspond to the partially filled lowest Landau level in the usual planar geometry.
We see that in the spherical geometry of the electron gas the states with higher $`|m|`$ possess the lower energy. Our situation is thus opposite to the one discussed for the quantum Hall edge states. Nevertheless both problems have a common ingredient, the linear-in-$`m`$ spectrum for a given $`n`$ (in our case in two domains, $`|m|<p`$ and $`|m|>p`$ ). Having effectively a case of one spatial dimension, we can consider the interaction effects as well. The problem however has a certain subtlety which is described below.
In certain cases one may hope to ignore the interaction between the states belonging to different “Landau levels” (different curves in the Fig. 1). Considering now the lowest LL, one observes familiar branches of right- and left-going fermions, $`|m|p`$ and $`|m|p`$, with the negative and positive “Fermi velocities” $`v_F=dE_{lm}/dm`$, respectively. The absolute values of $`v_F`$ for left and right movers are different. This point alone makes it difficult to pass to a bosonization description with one scalar field for both movers, and the notion of the chiral Luttinger liquid appears. A thorough consideration of the latter problem is beyond the scope of this study.
In conclusion, we considered the Landau quantization for the electron gas on a surface of sphere. The exact solution of this problem involves complicated functions, which are not very instructive for the analysis of states with large magnetic numbers $`m`$ for the electron motion. We elucidate the role of the “equatorial” states with large $`m`$ both analitically and numerically. These states are lower in energy, thus the electronic stripe on the equator can be realized for the semiconducting coating of the sphere in high magnetic fields.
I thank S.G. Romanov, A.G. Yashenkin, K. Hansen, V.A. Kulbachinskii for useful discussions and communications. The financial support from the Russian State Program for Statistical Physics (Grant VIII-2), grant FTNS 99-1134 and grant INTAS 97-1342 is gratefully acknowledged.
|
no-problem/9908/math-ph9908013.html
|
ar5iv
|
text
|
# The connection of Monge-Bateman equations with ordinary differential equations and their generalisation
## 1 Introduction
The famous Monge equation, which has been quoted in textbooks for more than 150 years, has the form:
$$\lambda _t=\lambda \lambda _x$$
Its implicit solution:
$$x\lambda t=f(\lambda )$$
reminds one of the solution of the equation of free motion with constant velocity $`\lambda `$ with initial value for the coordinate $`x_0=f(\lambda )`$ at $`t=0`$.
This similarity is not accidental and in the present note we want to draw attention to the fact of the connection of the Monge equation and the generalised it Bateman equation :
$$(\frac{\lambda _t}{\lambda _x})_t(\frac{\lambda _t}{\lambda _x})(\frac{\lambda _t}{\lambda _x})_x=0$$
(1)
with the ordinary differential equation $`\ddot{X}=0`$. This fact allow us to introduce partial differential equations connected with each ordinary differential equation of second order (this limitation is inessential) which possess the same integrable properties as the initial ODE.
This result can be generalised to any system of equations of second order, leading to a generalisation of the Monge and Bateman type of equations to the multidimensional case. The hydrodynamic system of D.B.Fairlie is the simplest example of such a generalisation.
## 2 The main assertion and its proof
Assertion
Let $`\ddot{X}=F(\dot{X},X;t)`$ be an ordinary differential equation, where $`F`$ is an arbitrary function of its arguments and $`\dot{}`$ denotes differentiation with respect to the independent argument $`t`$. Then the equation in partial derivatives:
$$(\frac{\lambda _t}{\lambda _X})_t(\frac{\lambda _t}{\lambda _X})(\frac{\lambda _t}{\lambda _X})_X=F(\frac{\lambda _t}{\lambda _X},X;t)$$
(2)
is exactly integrable (implicitly) simultaneously with the initial ordinary differential equation.
We would like to prove this assertion from both sides. First by using the known solution of the initial ordinary differential equation and secondly by explicit exchange of variables in it directly.
### 2.1 The first proof
The general solution of an ordinary differential equation of second order depends upon two arbitrary constants $`(c^1,c^2)`$ which we will consider as functions of two arguments $`\lambda (X,t)`$. So we have an implicit definition of the function $`\lambda `$:
$$X=X(t;c^1(\lambda ),c^2(\lambda ))$$
(3)
By we denote the derivative of $`X`$ with respect to the argument $`\lambda `$:
$$X^{}=X_{c^1}c_\lambda ^1+X_{c^2}c_\lambda ^2$$
It is obvious that dot and prime differentiations are mutually commutative. We obtain in consequence after differentiation of (3) with respect to its independent arguments $`X`$ and $`t`$:
$$1=X^{}\lambda _X(X_{c_1}c_1^{}+X_{c_2}c_2^{})\lambda _X,0=X_t+X^{}\lambda _t$$
or
$$X_t=\frac{\lambda _t}{\lambda _X}$$
Differentiation of the last equation once more by the arguments $`X,t`$ leads to the result:
$$X_{tt}+X_t^{}\lambda _t=(\frac{\lambda _t}{\lambda _X})_t,X_t^{}\lambda _X=(\frac{\lambda _t}{\lambda _X})_X$$
(4)
Eliminating $`X_t^{}`$ from the last two equalities and taking into account that under differentiation $`X_{tt}`$ $`c^1,c^2`$ remain constant we arrive at the generalised Bateman equation (2). This way from a given solution to equation it satisfied is decribed in each manual.
The generalized Monge equation may be obtain from (2) after making the identification $`\frac{\lambda _t}{\lambda _X}\lambda `$ and takes the form:
$$\lambda _t\lambda \lambda _X=F(\lambda ,X;t)$$
(5)
### 2.2 The second proof
Let us present the solution of equation of the second order from the assertion in the form $`Q(X,t)=\mathrm{constant}`$. Then for derivatives $`\dot{X}`$ ( with the help of the theorem of differentiation of implicit functions) we obtain:
$$\dot{X}=\frac{Q_t}{Q_X}$$
Repeating this operation once more we come to the generalized Bateman equation from this assertion (with the obvious exchange $`Q\lambda `$).
The corollary of the results of last two subsections may be summarised in:
Proposition (equivalent to the main assertion):
Each equation of second order can be presented in Monge-Bateman form (2). If the general solution of an ODE may be presented in explicit form, then the general solution of (5) is given by (3).
## 3 Multidimensional generalisation
Suppose we are given a set of $`(n1)`$ arbitrary functions $`X^iX^i(c^\alpha ;t)`$, each one depending on $`2(n1)`$ variables $`c^\alpha `$ and a single ”time” variable $`t`$. All $`2(n1)`$ independent variables $`c^\alpha `$ may be expressed implicitly in terms of a system of $`2(n1)`$ equations:
$`x^i=X^i(c^\alpha ;t),\dot{x}^i={\displaystyle \frac{X^i(c^\alpha ;t)}{t}}`$ (6)
After substitution of these values into explicit expressions for second derivatives we arrive at a system of ordinary differental equations of second order (as described in textbooks):
$$x_{tt}^i=\frac{^2X^i}{t^2}(x,x_t;t)$$
Now let us consider $`2(n1)`$ values $`c^\alpha `$ as a functions of $`(n1)`$ functions $`\rho ^s\rho ^s(x;t)`$. Differentiating the first equation of (6) first with respect to $`x_j`$ and secondly with respect to $`t`$ we obtain respectively:
$$\delta _{ij}=X_{\rho ^s}^i\rho _{x_j}^s,0=X_{\rho ^s}^i\rho _t^s+X_t^i$$
From the last equations we obtain immediately:
$$\{X_{\rho ^s}^i\}=J^1(\rho ;x),X_t=J^1\rho _t$$
where $`J(\rho ;x)`$ is the Jacobian matrix. Further differentiation of the second equation with respect to $`x_i,t`$ arguments leads to the result:
$$X_{\rho ^s,t}^i\rho _{x_i}^s=(J^1\rho _t)_{x_i},X_{tt}^i+X_{\rho ^s,t}^i\rho _t^s=(J^1\rho _t)t$$
Eliminating the matrix $`\{X_{\rho ^s,t}^i\}`$ from the last system we obtain finally:
$`\tau _t^s{\displaystyle \tau ^r\tau _{x_r}^s}=X_{tt}^s(x,\tau ;t)`$ (7)
where $`\tau =J^1\rho _t`$.
In the case of a noninteracting system $`X_{tt}^s=0`$ (7) goes over to the so called hydrodynamical system introduced and solved by D.B.Fairlie . In connection with this, the above implicit solution of the hydrodynamical system arises after treating the constants of motion in the trajectories of $`n`$ free moving particles as functions of $`n`$ arbitrary functions $`\rho `$. Thus $`n`$ equations:
$$x^s=f^s(\rho )t+g^s(\rho )$$
define in implicit form the solution of the $`n`$ dimensional Bateman-Monge equations. If we choose $`f^s=\rho ^s`$ we reproduce the original form of the solution of the hydrodynamical system proposed in .
## 4 Outlook
The results contained in formulae (4),(5) and (7) are so simple and clear that they don’t demand additional comments. It is possibile that they have been discovered long ago, but the author has not seen them in the literature.
## Acknowledgements.
The author gratefully thanks D.B.Fairlie and A.V.Razumov for discussions in the process of working on this paper and important comments.
Author is indebted to the Center for Research on Engenering and Applied Sciences (UAEM, Morelos, Mexico) for its hospitality and Russian Foundation of Fundamental Researches (RFFI) GRANT N 98-01-00330 for partial support.
|
no-problem/9908/cond-mat9908090.html
|
ar5iv
|
text
|
# Quantum transport through ballistic cavities: soft vs. hard quantum chaos
\[
## Abstract
We study transport through a two-dimensional billiard attached to two infinite leads by numerically calculating the Landauer conductance and the Wigner time delay. In the generic case of a mixed phase space we find a power law distribution of resonance widths and a power law dependence of conductance increments apparently reflecting the classical dwell time exponent, in striking difference to the case of a fully chaotic phase space. Surprisingly, these power laws appear on energy scales below the mean level spacing, in contrast to semiclassical expectations.
\]
Advances in the fabrication of semiconductor heterostructures and metal films have made it possible to produce two dimensional nanostructures with a very low amount of disorder . At low temperatures, scattering of the electrons happens mostly at edges of the structures with the electrons moving ballistically between collisions with the boundary. Theoretical and experimental investigations have shown that the spectral and transport properties of such quantum coherent cavities, commonly called “billiards”, depend strongly on the nature of their classical dynamics. In particular, integrable and chaotic systems were found to behave quite differently .
Generic billiards are neither integrable nor ergodic , but have a mixed phase space with regions of regular as well as chaotic dynamics . Their dynamics is much richer than in either of the extreme cases, as phase space has a hierarchical structure at the boundary of regular and chaotic motion. In particular, this leads to a trapping of chaotic trajectories close to regular regions with a probability $`P(t)t^\beta `$ for $`t>t_0`$, to be trapped longer than a time $`t`$, with $`t_0`$ of the order of a few traversal times . The exponent $`\beta >1`$ depends on system and parameters with typically $`\beta 1.5`$ . This power-law trapping in mixed systems is in contrast to the typical exponentially decaying staying probability of fully chaotic systems (see Fig. 1).
Recently, it was shown semiclassically employing the diagonal approximation that the variance of conductance increments (for a small dc bias voltage) over small energy intervals $`\mathrm{\Delta }E`$ grows as
$$\mathrm{\Delta }g^2(\mathrm{\Delta }E)[g(E+\mathrm{\Delta }E)g(E)]^2_E|\mathrm{\Delta }E|^\beta ,$$
(1)
for mixed systems if $`\beta <2`$. This is in strong contrast to an increase as $`(\mathrm{\Delta }E)^2`$ in the case of fully chaotic systems . The semiclassical approximation requires $`\mathrm{\Delta }E`$ to be larger than the mean level spacing $`\mathrm{\Delta }`$, corresponding to the picture that quantum mechanics can follow the classical power law trapping at most until the Heisenberg time $`t_H=h/\mathrm{\Delta }`$ . In the semiclassical approximation the graph of $`g`$ vs. $`E`$ has the statistical properties of fractional Brownian motion with a fractal dimension $`D=2\beta /2`$ . Fractal conductance fluctuations have indeed been found in experiments on gold wires and semiconductor nanostructures and numerically for the quantum separatrix map .
In this Letter, we numerically study quantum transport through a simple cavity, the cosine billiard (see insets of Fig. 1). Although we observe completely different behavior for the mixed and fully chaotic cases in the semiclassical regime of many (45) transmitting modes, we find in the mixed case no indication of fractality or fractional Brownian motion behavior of the graph $`g`$ vs. $`E`$. This is the first surprise, as it is in contrast to the above mentioned semiclassical , experimental , and numerical works. Instead, the conductance is characterized by narrow isolated resonances, with the classical exponent $`\beta `$ appearing in a power law distribution of resonance widths smaller than the mean level spacing. This leads to a scaling of $`\mathrm{\Delta }g^2(\mathrm{\Delta }E)`$ in agreement with the semiclassically derived Eq. (1), however, only on scales below the mean level spacing. This surprising result contradicts the semiclassical intuition that quantum mechanics may mimic classical properties at most until the Heisenberg time corresponding to energy scales above the mean level spacing. At present, there is no explanation for these numerical results. They show that even with a detailed (semiclassical) knowledge of the universal chaotic regime as well as the integrable case at hand we are just at the beginning of understanding the quantum properties of generic Hamiltonian systems.
The cosine billiard is defined by two hard walls at $`y=0`$ and $`y(x)=W+(M/2)(1\mathrm{cos}(2\pi x/L))`$, for $`0xL`$, with two semi-infinite perfect leads of width $`W`$ attached to the openings of the billiard at $`x=0`$ and $`x=L`$ (see insets of Fig. 1). By changing the parameter ratios $`W/L`$ and $`M/L`$ the stability of periodic orbits associated with the billiard can be changed, allowing a transition from a mixed to a predominantly chaotic phase space. Note, that in the mixed case the leads couple to the chaotic part of phase space only.
The $`S`$-matrix of the system has been calculated by the recursive Green’s function method after expanding the two-dimensional wave function in terms of local transverse energy eigenfunctions . In the numerical calculations, it was checked that a sufficient number of modes in the expansion in transverse eigenmodes was kept and that the lattice constant in $`x`$-direction was sufficiently small. For a given energy $`E_F=\mathrm{}^2k_F^2/2m`$, $`N`$ modes in the leads are transmitting, with $`k_FW/\pi N`$. We turn from the $`S`$-matrix to the experimentally relevant conductance at small dc bias voltage using the Landauer formula, $`G=e^2/h\text{Tr}(tt^{})`$, where $`t`$ is the transmission matrix. Spectral information is contained in the Wigner-Smith time delay $`\tau =i\mathrm{}\text{Tr}(S^{}dS/dE)/2N`$, where $`2N`$ is the dimension of the $`S`$-matrix. All energies in this paper are given in units of $`\mathrm{}^2\pi ^2/(2mW^2)`$.
Figure 2 shows the dimensionless conductance $`g=G(h/e^2)`$ and the Wigner-Smith time delay $`\tau `$ \[in units of $`2mW^2/(\mathrm{}\pi ^2)`$\] for parameters corresponding to a mixed phase space ($`W/L=0.18`$, $`M/L=0.11`$) and a chaotic phase space with no apparent stability island ($`W/L=0.36`$, $`M/L=0.22`$) for $`N=45`$ transmitting modes. The differences are quite dramatic. For the fully chaotic case, both quantities are smooth functions of energy and in good agreement with semiclassical theory (see below). While the average values are comparable, many sharp isolated resonances on top of a smooth background are visible in the mixed case , also in contrast to the semiclassically predicted fractional Brownian motion. The simple explanation that these narrow resonances are related to quantum tunneling into the islands of regular motion does not apply here, as the phase space volume of stable islands is about 5%, while the narrow resonances (below the mean level spacing) make up about 18% of all states associated with the billiard. This roughly corresponds to the phase space volume around the stable islands where trapping of chaotic trajectories occurs.
In order to analyze the narrow resonances in the mixed case, it is convenient to examine the Wigner-Smith time delay. Each resonance in the time delay has the Breit-Wigner shape, characterized by a width $`\mathrm{\Gamma }_i`$ and a height $`\tau _i`$ situated at an energy $`E_i`$ on top of a smooth background. We find our data well described by
$$\tau (E)=\underset{i}{}\tau _i\frac{\mathrm{\Gamma }_i^2/4}{(EE_i)^2+\mathrm{\Gamma }_i^2/4}+\tau _{\text{smooth}}(E),$$
(2)
with $`\tau _{\text{smooth}}(E)E^{1/2}`$. Since the phase shift through a resonance is $`2\pi `$, width and height are related by $`\tau _i\mathrm{\Gamma }_i=2/N`$. The energy was initially sampled on an equidistant grid and subsequently refined in order to resolve the sharp resonances. Only resonances with a $`\mathrm{\Gamma }10^3`$, i.e. much smaller than the initial grid are lost. As a result we can numerically construct the cumulative distribution $`N(\mathrm{\Gamma })`$ of resonance widths, corresponding to the probability of finding a resonance smaller than $`\mathrm{\Gamma }`$ (Fig. 3). The distribution is very broad, spanning 5 orders of magnitude, and is approximately a power law $`N(\mathrm{\Gamma })a\mathrm{\Gamma }^r`$, with $`r0.35`$, over a wide range below the mean level spacing $`\mathrm{\Delta }=0.176`$.
The consequences of this broad distribution of resonance widths for the variances of conductance and time delay increments are studied now. For $`\mathrm{\Delta }E\mathrm{\Delta }`$ correlations between different isolated resonances ($`\mathrm{\Gamma }\mathrm{\Delta }`$) do not contribute to the variance $`\mathrm{\Delta }g^2`$, which then is governed by the distribution of resonance widths. Each resonance is reflected in the conductance,
$$g(E)=g_{\text{smooth}}(E)+\underset{i=1}{\overset{N_R}{}}\delta g_i(E),$$
(3)
where $`\delta g_i(E)`$ is a function of the width $`\mathrm{\Gamma }_i`$ and the typical height $`\stackrel{~}{g}_i`$. $`N_R`$ is the number of resonances with $`\mathrm{\Gamma }_i<\mathrm{\Delta }`$ in the energy interval $`E_BE_A`$ over which we take the average. The variance of the increments of a single resonance is given by
$`[\delta g_i(E+\mathrm{\Delta }E)\delta g_i(E)]^2_E=`$ (7)
$`{\displaystyle \frac{\stackrel{~}{g}_i^2\mathrm{\Gamma }_i^{}}{E_BE_A}}\{\begin{array}{ccc}b_i(\mathrm{\Delta }E/\mathrm{\Gamma }_i)^2& ,& \mathrm{\Delta }E\mathrm{\Gamma }_i\\ 1& ,& \mathrm{\Delta }E\mathrm{\Gamma }_i\end{array},`$
which defines $`\stackrel{~}{g}_i`$ and where $`b_i`$ is a numerical factor of order unity. Since the distribution of widths is very broad, the strong inequalities are almost always fulfilled in the sum over resonances. Splitting this sum into resonances smaller and larger than $`\mathrm{\Delta }E`$, we get
$$\mathrm{\Delta }g^2(\mathrm{\Delta }E)\frac{1}{E_BE_A}\left(\underset{\mathrm{\Gamma }_i<\mathrm{\Delta }E}{}\stackrel{~}{g}_i^2\mathrm{\Gamma }_i^{}+\underset{\mathrm{\Gamma }_i>\mathrm{\Delta }E}{}b_i^{}\stackrel{~}{g}_i^2\frac{(\mathrm{\Delta }E)^2}{\mathrm{\Gamma }_i}\right).$$
(8)
Replacing the sums by integrals over the density of widths $`n(\mathrm{\Gamma })ar\mathrm{\Gamma }^{r1}`$ and neglecting the weak fluctuations of $`g_i`$ and $`b_i`$ as compared to $`\mathrm{\Gamma }_i`$, we can estimate for small $`\mathrm{\Delta }E`$,
$$\mathrm{\Delta }g^2(\mathrm{\Delta }E)\stackrel{~}{g}^2^{}\frac{N_R}{E_BE_A}a|\mathrm{\Delta }E|^{1+r}.$$
(9)
where $`\mathrm{}`$ stands for the average over isolated resonances. A power law distribution of resonances thus leads to a power law increase of the variance of conductance increments with the exponent given by $`1+r`$.
Fig. 4 a) shows the variances of the conductance increments. On scales smaller than the minimum resonance width the variance increases quadratically, as expected. On larger scales we find the power law Eq. (9). This result coincides with the semiclassically derived Eq. (1) with $`r=\beta 1`$, however, only on scales below the mean level spacing. At present, there is no explanation why the classical exponent $`\beta `$ appears on such small energy scales. Remarkably, on scales above the mean level spacing the correlation energy for the conductance fluctuations is given by the Weisskopf width $`\mathrm{\Gamma }_W2`$, as in the fully chaotic case (see below).
The variance of increments of the time delay are shown in Fig. 4 b). Since the variance $`\mathrm{\Delta }\tau ^2(\mathrm{\Delta }E)`$ measures the square of the resonance peak height in the time delay, in the mixed case, they are completely dominated by the sharpest resonance, once the energy exceeds the minimum resonance width. Thus, in contrast to fully chaotic systems, in mixed systems the scale of the correlations of the time delay is the smallest resonance width and not the Weisskopf width $`\mathrm{\Gamma }_W`$.
For comparison in Fig. 4 we also show the results for $`\mathrm{\Delta }g^2(\mathrm{\Delta }E)`$ and $`\mathrm{\Delta }\tau ^2(\mathrm{\Delta }E)`$ for the fully chaotic case. They are characterized by single scales $`\mathrm{\Gamma }_g4.8`$ and $`\mathrm{\Gamma }_\tau 3.8`$ and are in good agreement with semiclassical results . The chaotic case can also be described by the random matrix theory (RMT) , whose results coincide with the cited semiclassical ones for $`N1`$ . In the absence of direct processes, random matrix theory predicts a single correlation scale, known as Weisskopf correlation width, $`\mathrm{\Gamma }_W=\mathrm{\Delta }/2\pi _cT_c`$, where the sum runs over all channels $`c`$ with transmission probability $`T_c`$ . Approximating $`_cT_c`$ by twice the average dimensionless conductance we obtain $`\mathrm{\Gamma }_W4.2`$, in agreement with the numerical values within the statistical accuracy. Before concluding, it is worthwhile to stress that depending on $`N`$ and the coupling to the leads, quantum chaotic scattering can also exhibit isolated resonances. Their width distribution, however, follows a $`\chi ^2`$-distribution with $`N`$ degrees of freedom , rather than power law.
In conclusion, we have shown that generic Hamiltonian systems, which have regular as well as chaotic phase space regions, differ drastically in the Landauer conductance and Wigner time delay from fully chaotic systems. We find many isolated narrow resonances with a power law distribution of their widths accompanied by a power law increase of the variance of conductance increments. Both power laws appear to be connected to the classical power law trapping, surprisingly they only appear on scales below the mean level spacing. Similar unexplained power laws are found in recent studies using quantum graphs modeling a mixed phase space (see inset of Fig. 4 a) . Further research on the quantum signatures of classically mixed systems is urgently needed.
R.K. thanks I. Guarneri for many stimulating discussions. We thank L. Hufnagel, F. Steinbach, and M. Weiss for the inset of Fig. 4 a). This work was supported by the DFG and the ITP at UCSB (B.H.), CNPq and PRONEX (C.H.L.), and the ICTP in Trieste (B.H and C.H.L.).
Note added.—The authors of Ref. have informed us that there are quantum graphs where the classical and quantum exponents do not agree.
|
no-problem/9908/quant-ph9908003.html
|
ar5iv
|
text
|
# Quantum Clone and States Estimation for 𝑛-state System
## Abstract
We derive a lower bound for the optimal fidelity for deterministic cloning a set of $`n`$ pure states. In connection with states estimation, we obtain a lower bound about average maximum correct states estimation probability.
PACS numbers: 03.67.-a, 03.65.Bz, 89.70.+c
Quantum no-cloning theorem has prohibited cloning and estimating an arbitrary quantum state exactly by any physical means in a consequence of linearity of quantum theory. The unitarity of quantum theory does not allow to clone (identify) no-orthogonal states though orthogonal states can be cloned (identified) perfectly . However, clone and estimation of quantum states with a limited degree of success are always possible. Universal quantum cloning machine (UQCM) acts on any unknown quantum state and produce optimal approximate copies. This machine is called universal because it produces copies that are state-independent. State-dependent quantum cloning machines is designed to clone states belonging to a finite set and may be divided into two main categories: deterministic , probabilistic and hybrid . Deterministic state-dependent cloning machine generates approximate clones with probability 1. Deterministic exact clone violates the no-cloning theorem, thus perfectly clone must be probabilistic. Probabilistic quantum cloning machines can clone states perfectly, though the success probability cannot be unit all the time. It is shown that a set of non-orthogonal states can be probabilistically cloned if and only if the states are linearly independent. Hybrid clone interpolates between deterministic and probabilistic ones, that is, the copies (not exact) are better than those in deterministic clone, but the success probability (less than 1) is greater than probabilistic exact clone. Universal quantum states estimation were considered in Ref. , given $`M`$ independent realizations. What’s more, we have discussed general states discrimination strategies for state-dependent system.
Optimal results for two-state deterministic clone have been obtained in Ref. . In this letter we consider deterministic clone for a set of $`n`$ pure states $`\left\{|\psi _i,i=1,2,\mathrm{},n\right\}`$. When $`|\psi _i`$ are non-orthogonal, they cannot be cloned perfectly. What we require is that the final states should be most similar as the target states, that is, the fidelity between the final and target states should be optimal. We derive a lower bound for the optimal fidelity of the cloning machine. Applying it to states estimation, we obtain the lower bound about average maximum correct identification probability in deterministic states estimation.
A quantum state-dependent cloning device is a quantum machine which performs a prescribed unitary transformation on an extended input which contains $`M`$ original states in system $`A`$ and $`NM`$ blank states in system $`B`$ with $`N`$ output copies. The unitary evolution transfers states as follows
$$U|\psi _i^M_A|\mathrm{\Sigma }^{NM}_B=|\alpha _i\text{,}$$
(1)
where $`|\psi _i^M_A=|\psi _i_1\mathrm{}|\psi _i_M`$ are the $`M`$ original states, $`|\mathrm{\Sigma }^{NM}_B`$ are the blank states and $`|\alpha _i`$ are the output cloned states. The $`n\times n`$ inter-inner-products of Eq. (1) yield the matrix equation<sup>*</sup><sup>*</sup>*We notice the preserving inner-product property of unitary transformation, that is, if two sets of states $`\{|\varphi _1,|\varphi _2,\mathrm{},|\varphi _n\}`$ and $`\{|\stackrel{~}{\varphi }_1,|\stackrel{~}{\varphi }_2,\mathrm{},|\stackrel{~}{\varphi }_n\}`$ satisfy the condition $`\varphi _i\varphi _j=\stackrel{~}{\varphi }_i\stackrel{~}{\varphi }_j`$, there exists a unitary operate $`U`$ to make $`U|\varphi _i=|\stackrel{~}{\varphi }_i`$ ($`i=1,2,\mathrm{},n`$).
$$X^{(M)}=\stackrel{~}{X}\text{,}$$
(2)
where $`n\times n`$ matrices $`\stackrel{~}{X}=\left[\alpha _i|\alpha _j\right]`$, $`X^{(M)}=\left[\psi _i|\psi _j^M\right]`$.
We require a figure of merit to characterize how closely our copies $`|\alpha _i`$ resemble exact copies $`|\psi _i^N`$. Denoting the priori probability of the state $`|\psi _i^M`$ by $`\eta _i`$, one interesting measure of the final states is the global fidelity introduced by Bru$`\beta `$ et al. , which is defined formally as
$$F=\underset{i=1}{\overset{n}{}}\eta _i\left|\alpha _i|\psi _i^N\right|^2.$$
(3)
As a criterion for optimality of the state-dependent cloner, the unitary evolution $`U`$ should maximize the global fidelity $`F`$ of $`n`$ final states $`|\alpha _i`$ with respect to the perfect cloned states $`|\psi _i^N`$. We focus here on the global fidelity since it has an important interpretation in connection with states estimation .
Now the remained problem is to find the maximum value of the fidelity $`F`$, which means optimal clone. It is equivalent to the problem of maximizing $`F`$ under the condition of Eq. (2). This problem is a nonlinear programming and fairly difficult to solve. Nevertheless a lower bound of the optimal fidelity could still be derived by adopting an auxiliary function $`F^{^{}}`$, which is defined as
$$F^{^{}}=\underset{i=1}{\overset{n}{}}\eta _i\left|\psi _i^N\alpha _i\right|.$$
(4)
Such function also describes how closely our output copies resemble exact copies. There exists a bound between $`F`$ and $`F^{^{}}`$ (see below, inequality (9)), therefore a bound for $`F`$ may be obtained by optimizing $`F^{^{}}`$.
We find that the optimal output states $`|\alpha _i`$ must lie in the subspace spanned by the exact clones $`|\psi _i^N`$. This conclusion may be easily come to with the method of Lagrange Multipliers (please refer to , where $`n=2`$) and here we omit the proof.
If a set of states $`|\stackrel{~}{\alpha }_i`$ fulfil Eq. (2), that is, $`X^{\left(M\right)}=\stackrel{~}{X}=\left[\stackrel{~}{\alpha }_i|\stackrel{~}{\alpha }_j\right]`$, there must exist a unitary transformation $`V`$ satisfies $`V|\stackrel{~}{\alpha }_i=|\alpha _i`$, thus we can vary $`V`$ to optimize $`F^{^{}}`$ with chosen states $`|\stackrel{~}{\alpha }_i`$. Suppose $`\left|\psi _i^N\alpha _i\right|=\lambda _i\psi _i^N\alpha _i`$ with $`\lambda _i\left\{\pm 1\right\}`$ in the optimal situation (the determination of $`\lambda _i`$ will be shown in later part), the optimal $`F^{^{}}`$ is
$$F_{opt}^{^{}}=\underset{V}{\mathrm{max}}F^{^{}}=\underset{V}{\mathrm{max}}\left|\underset{i=1}{\overset{n}{}}\eta _i\lambda _i\psi _i^N\left|V\right|\stackrel{~}{\alpha }_i\right|.$$
(5)
Choose $`n`$ orthogonal states $`|\chi _i`$ which span a space $``$ and the space spanned by $`|\psi _i^N`$ is a subspace of $``$We consider space $`\left\{|\psi _i^N,i=1,2,\mathrm{},n\right\}`$ may be a subspace of $``$ since $`|\psi _i`$ may be linear-dependent and cannot span a $`n`$-dimensional Hilbert space.. Set $`|\stackrel{~}{\alpha }_i=\underset{j=1}{\overset{n}{}}a_{ij}|\chi _j`$, $`|\psi _i^N=\underset{j=1}{\overset{n}{}}b_{ij}|\chi _j`$ on the orthogonal bases $`|\chi _i`$, $`i=1,2,\mathrm{},n`$, we get
$$F_{opt}^{^{}}=\underset{V}{\mathrm{max}}\left|tr\left(\eta \lambda BVA^+\right)\right|=\underset{V}{\mathrm{max}}\left|tr(VO)\right|=tr\sqrt{O^+O},$$
(6)
where $`A=\left[a_{ij}\right]`$, $`B=\left[b_{ij}\right]`$, $`\eta =diag(\eta _1,\eta _2,\mathrm{},\eta _n)`$, $`\lambda =diag(\lambda _1,\lambda _2,\mathrm{},\lambda _n)`$, $`O=A^+\eta \lambda B`$. We have used the freedom in $`V`$ to make the inequality as tight as possible. To do this we have recalled that $`\underset{V}{\mathrm{max}}\left|tr(VO)\right|=tr\sqrt{O^+O}`$, where $`O`$ is any operator and the maximum is achieved only by those $`V`$ such that
$$VO=e^{i\nu }\sqrt{O^+O},$$
(7)
where $`\nu `$ is arbitrary. Generally, we choose $`\nu =0`$.
As we require above, $`\lambda _i`$ should satisfy $`\lambda _i\psi _i^N\left|V\right|\stackrel{~}{\alpha }_i0`$. This condition can be represented as $`\chi _i\left|\lambda BVA^+\right|\chi _i0`$, which means the diagonal elements of matrix $`\lambda BVA^+`$ should be positive. Since $`\lambda _i\left\{\pm 1\right\}`$, a simple method to determine $`\lambda _i`$ is to enumerate the $`2^n`$ possible results of $`\lambda =diag(\lambda _1,\lambda _2,\mathrm{},\lambda _n)`$ and verify which one fulfils above inequality. With a chosen basis $`|\chi _i`$, matrix $`A`$, $`B`$ can be given by equations $`A^+A=X^{\left(M\right)}`$ and $`B^+B=X^{\left(N\right)}`$ respectively, $`V`$ can be represented with parameters $`\lambda _i`$, thus above postcalculation method can determine matrix $`\lambda `$ and then give the maximum $`F_{opt}^{^{}}`$. According to Eq. (6), we obtain a tight upper bound for the function $`F^{^{}}`$,
$$F^{^{}}tr\sqrt{B^+\lambda \eta X^{\left(M\right)}\eta \lambda B}.$$
(8)
The fidelity $`F`$ of the cloning machine is constrained by the following inequality
$`F`$ $`=`$ $`\left({\displaystyle \underset{i=1}{\overset{n}{}}}\eta _i\left|\alpha _i|\psi _i^N\right|^2\right)\left({\displaystyle \underset{i=1}{\overset{n}{}}}\eta _i\right)`$ (9)
$``$ $`\left({\displaystyle \underset{i=1}{\overset{n}{}}}\eta _i\left|\alpha _i|\psi _i^N\right|\right)^2=\left(F^{^{}}\right)^2,`$ (10)
where the equation is met if and only if $`\left|\alpha _i|\psi _i^N\right|`$ are constant. Obviously $`F`$ is not always optimal even if $`F^{^{}}`$ is optimal. However optimal $`F`$ should be greater than or equal to $`\left(F_{opt}^{^{}}\right)^2`$. When $`n=2`$ and $`\eta _1=\eta _2`$, equation in Ineq. (9) is satisfied and gives the optimal results, which has been provided in Ref. .
State-dependent clone has a close connection with states estimation in the limit as $`N\mathrm{}`$. Given infinite copies of $`n`$ non-orthogonal states, we can discriminate them exactly with probability 1. On the other hand, if we can discriminate $`n`$ states, we can obtain infinite copies. There are two ways in which an attempt to discriminate between non-orthogonal states; it can give either an erroneous or an inconclusive result . In the following we will consider a strategy without inconclusive results using above results in the limit as $`N\mathrm{}`$. In fact, since the optimal output states $`|\alpha _i`$ lie in the subspace spanned by the exact clones $`|\psi _i^N`$, Eq. (1) may be rewritten as
$$U|\psi _i^M|\mathrm{\Sigma }^{NM}=\underset{j=1}{\overset{n}{}}c_{ij}|\psi _j^N\text{,}$$
(11)
where $`c_{ij}=\psi _j^N\alpha _i`$. If $`N\mathrm{}`$, $`\left\{|\psi _j^N,j=1,2,\mathrm{},n\right\}`$ are orthogonal. After the evolution, the cloning system is measured and if $`|\psi _j^{\mathrm{}}`$ is obtained, the original state is estimated as $`|\psi _j^M`$. The states estimation is correct with probability $`\left|c_{ii}\right|^2`$ when $`j=i`$. If $`ji`$, errors occur with probability $`\underset{ji}{}\left|c_{ij}\right|^2`$. The inter-inner products of Eq. (10) give the matrix equation in the limit $`N\mathrm{}`$,
$$X^{(M)}EE^+=0,$$
(12)
where $`E=\left[c_{ij}\right]`$. The diagonal elements is corresponding to the probabilities of correct states estimation while non-diagonal elements to those of error. This equation describes the bound between the maximum probabilities of correct discrimination and those of incorrect one. In fact, this result is a special case of that we have derived in . In Ref. , we have consider two possible ways in which an attempt to discriminate between non-orthogonal states can fail, by giving either an erroneous or an inconclusive result. Above strategy just gives an erroneous result with some probability. Our principal result in Ref. is the matrix inequality which prescribes the bound among the probabilities of correct, error and inconclusive discrimination results. Such bound may have intriguing implications for quantum communication theory and cryptography since it offers a potential eavesdropper increased flexibility by a compromise between inconclusive and erroneous results.
An important optimality criterion of the states estimation is the average maximum correct probability, that is, $`P=_i\eta _i\left|c_{ii}\right|^2=F`$ in the limit $`N\mathrm{}`$It is the reason why we choose the definition of $`F`$ as that in Eq. (3).. In this situation $`|\psi _j^N`$ are orthogonal, thus matrix $`B=I_n`$. Applying Eq. (8) and (9), we obtain
$$P=\underset{i}{}\eta _i\left|c_{ii}\right|^2\left(tr\sqrt{\lambda \eta X^{(M)}\eta \lambda }\right)^2.$$
(13)
Such $`F`$ is not always optimal bound of the average maximum probability of correct states estimation, however, the optimal one is always greater than $`\left(tr\sqrt{\lambda \eta X^{(M)}\eta \lambda }\right)^2`$.
We note that above bound about $`F`$ and $`P`$ have the meaning in average. They describe the optimality approach to the final states we can reach in average of the $`n`$ initial states and does not mean the best for each initial state. However, since we do not know which one the initial state is in the clone or estimation process, such average may be the most important value to describe the efficiencies of cloning (estimating) machines.
In summary, we have derived a lower bound for the optimal fidelity for the state-dependent quantum clone. In connection with states estimation, we obtained the matrix inequality which describes the bound between the maximum probabilities of correct discrimination and those of incorrect one. A lower bound about average maximum probability of correct identification has also been presented. Our results give some bounds which the optimal cloner and states estimation can be better than in average, however, we have not found a limit which optimal cloner can reach at most. It is still an open question needed to be explored.
ACKNOWLEDGMENTS: This work was supported by the National Natural Science Foundation of China.
* Electronic address: [email protected]
Electronic address: [email protected]
|
no-problem/9908/cond-mat9908429.html
|
ar5iv
|
text
|
# Anomalous microwave response of high-temperature superconducting thin-film microstrip resonator in weak dc magnetic fields
## I Introduction
In general, it is expected that the surface impedance of high-temperature superconducting (HTS) thin films will increase as a function of applied dc or mw field. This nonlinearity was believed to be mainly contributed by weak links and vortex dynamics in the HTS films . Besides, other sources of nonlinearity, such as intrinsic pairbreaking, local heating of grain boundaries and thermal switch of superconducting grains, were also introduced to explain the diverged experimental data.
Recently, several anomalous microwave responses, i.e. surface resistance R<sub>s</sub> and (or) surface reactance X<sub>s</sub> change non-monotonically with applied mw or dc magnetic field, were observed . Choudhury $`et`$ $`al`$. reported an anomalous field dependence of surface resistance R<sub>s</sub> of a YBCO suspended line resonator in the presence of a weak, perpendicular dc magnetic field at 10K. They found a sharp decrease of R<sub>s</sub> at small H<sub>dc</sub> ($``$5 Gauss) with subsequent increase at higher fields. Hein $`et`$ $`al`$. observed a correlated reduction in R<sub>s</sub> and X<sub>s</sub> of epitaxial YBCO films in both dc and mw magnetic fields of the scale 20 mT at frequencies f=8.5 and 77 GHz and temperatures T=77 and 4.2K, respectively. The field effects on R<sub>s</sub> and X<sub>s</sub> were found to correlate within the framework of the two-fluid model (TFM). The reduction in R<sub>s</sub> and X<sub>s</sub> was attributed to magnetic field-induced suppression of the spin-flip scattering rate. Similar effects were also reported by Kharel and his colleagues. Besides correlated changes in R<sub>s</sub> and X<sub>s</sub>, uncorrelated non-monotonic field dependence was also observed in their measurement of YBCO coplanar resonators at 8.0GHz and different temperatures (15, 35 and 75K). As far as we know, the origins of these anomalous microwave responses are still pending and the experimental data are very limited. More experimental investigations are needed to give enough clues for learning the mechanisms underneath the anomalous effects.
In this work, we employed a microstrip resonator technique to measure R<sub>s</sub> and X<sub>s</sub> of YBCO thin films in dc magnetic fields H<sub>dc</sub> up to 200 Gauss at 77K. It was observed that R<sub>s</sub> decreases as H<sub>dc</sub> raises from zero and this is followed by a monotonically increase of R<sub>s</sub> at higher H<sub>dc</sub>. Correlatively, X<sub>s</sub> show a similar behaivior. The effects were investigated at different mw amplitudes and frequencies in different dc field alignments. The experimental data have been examined with several theoretical models.
## II Sample preparation
The double-sided YBCO thin films were prepared by on-axis pulsed-laser deposition (PLD) technique on polished LaAlO<sub>3</sub> (100) substrates with the size of 15$`\times `$10$`\times `$0.5 mm<sup>3</sup>. The films are 400nm thick and show a strong preferential orientation with the c-axis perpendicular to the film surface. The AFM results show that the grain sizes of both faces of the double-sided films are around 1$`\mu `$m$`\times `$1$`\mu `$m and the root-mean-square (rms) surface roughness of both faces in a 50$`\mu `$m$`\times `$50$`\mu `$m area are similar, around 170 nm while their mean roughness are also similar, which are around 130 nm. The dc critical current density J<sub>c</sub> is in excess of 10<sup>6</sup>A/cm<sup>2</sup> at 77K and the transition temperature T<sub>c</sub> is around 90K with a narrow transition width, 0.5$``$1K. Employing a dielectric resonant cavity , the surface resistances of the films were measured at 10.66GHz, 77K before patterning. The resultant R<sub>s</sub> are around 1 m$`\mathrm{\Omega }`$ and the difference of the surface resistance between the two sides is smaller than 5%.
The design of the microstrip resonator was done with the help of a commercial full wave electromagnetic simulator, IE3D, and the quasi-TEM analytical formulations for microstrip resonator. The resonator geometry is chosen so that it has a characteristic impedance of 50$`\mathrm{\Omega }`$. The meandering microstrip has a width of 169 $`\mu `$m and is coupled with the outer circuits by a capacitive gap of 500 $`\mu `$m. The traditional wet etching method was employed to fabricate the microstrip resonator. We have checked the geometry of the resonator after the etching process and found the actual strip width is around 4$`\mu `$m narrower than the designed value.
## III Experimental results
The microwave responses of the resonators were measured at 77K using a vector network analyzer and the dc magnetic field was applied by a copper solenoid. There is no remanent field in the measurements since all the experimental set-ups are made of Teflon except the copper solenoid. The changes in surface impedance associated with the application of dc magnetic field are extracted from the measured quantities as follow :
$$\mathrm{\Delta }\mathrm{Z}_\mathrm{s}(\mathrm{H}_{\mathrm{dc}})=\mathrm{Z}_\mathrm{s}(\mathrm{H}_{\mathrm{dc}})\mathrm{Z}_\mathrm{s}(0)=\mathrm{\Gamma }[\mathrm{\Delta }\mathrm{f}_{3\mathrm{d}\mathrm{B}}(\mathrm{H}_{\mathrm{dc}})\mathrm{\Delta }\mathrm{f}_{3\mathrm{d}\mathrm{B}}(0)+i2(\mathrm{f}_0(0)\mathrm{f}_0(\mathrm{H}_{\mathrm{dc}}))]$$
(1)
where $`\mathrm{\Gamma }`$ is a geometric factor determined from the sample dimensions, f<sub>0</sub> and $`\mathrm{\Delta }`$f<sub>-3dB</sub> are the resonance frequency and the $``$3dB (half power) bandwidth of the resonance peak, respectively. The loaded quality factor Q<sub>L</sub> of the resonator can be easily obtained from f<sub>0</sub> and $`\mathrm{\Delta }`$f<sub>-3dB</sub>, Q<sub>L</sub>=f<sub>0</sub>/$`\mathrm{\Delta }`$f<sub>-3dB</sub>. We replace Q<sub>L</sub> with unloaded quality factor because the coupling is very weak (insertion loss$`>`$30dB).
Figure 1(a) shows the typical data for $`\mathrm{\Delta }`$f<sub>-3dB</sub>, which is proportional to R<sub>s</sub>, of the fundamental resonant peak as a function of applied dc field H<sub>dc</sub>. The input mw power was fixed at $``$10dBm and the sample was initially cooled in a zero-field state where the earth magnetic field ($``$0.3 Gauss) was neglected. We do not give absolute values of R<sub>s</sub> and X<sub>s</sub> because $`\mathrm{\Gamma }`$ is difficult to be determined for the current in the microstrip is highly non-uniform . Fortunately, only the functional dependence of $`\mathrm{\Delta }`$Z<sub>s</sub> and the ratio of $`\mathrm{\Delta }`$R<sub>s</sub> and $`\mathrm{\Delta }`$X<sub>s</sub> are of importance to learn the mechanisms of microwave nonlinearity and they can be determined without the geometric factor. For the case H<sub>dc</sub> was applied parallel to the $`c`$-axis of the film (namely, perpendicular to the surface of the film), R<sub>s</sub> drops dramatically as H<sub>dc</sub> is increased from zero. After passing through a minimum, R<sub>s</sub> begin to increase monotonically when H<sub>dc</sub> is further raised. The minimum depth is 23% of the R<sub>s</sub> value in zero dc field and the crossover field (H$`{}_{}{}^{}{}_{\mathrm{c}}{}^{}`$) is 4.5 Gauss, which is of the same order of that reported by Choudhury $`et`$ $`al`$. . The shift of the resonant frequency, f<sub>0</sub>, shown in Figure 1(b) reveals that the change of X<sub>s</sub> is correlated with that of R<sub>s</sub>. The resonant peak initially shifts to higher frequency (X<sub>s</sub> decreases, correspondingly) and then shifts down. The H$`{}_{}{}^{}{}_{\mathrm{c}}{}^{}`$ value determined from X<sub>s</sub>(H<sub>dc</sub>) is the same as that from R<sub>s</sub>(H<sub>dc</sub>) within 1.5 Gauss.
The general features observed above were presented again for H<sub>dc</sub>$``$$`c`$ while the crossover field H$`{}_{}{}^{}{}_{\mathrm{dc}}{}^{}`$ is of the order 80$``$90 Gauss. The difference between H$`{}_{}{}^{}{}_{\mathrm{dc}}{}^{}`$ and H$`{}_{}{}^{}{}_{\mathrm{dc}}{}^{}`$ may come from the serious demagnetizing effect and the anisotropy of the HTS film. The insets in Fig.1 show the curves normalised with H$`{}_{}{}^{}{}_{\mathrm{dc}}{}^{}`$ and H$`{}_{}{}^{}{}_{\mathrm{dc}}{}^{}`$, respectively. The two curves almost collapse to one. The implication of this result will be discussed below.
Similar effects are observed as the input power is varied. Fig.2 shows the $`\mathrm{\Delta }`$f<sub>-3dB</sub> versus H<sub>dc</sub> with different orientations when the input mw power is fixed at $``$20dBm. No shift of the crossover field H<sub>c</sub>, which represents H<sub>dc</sub>$``$$`c`$ or H<sub>dc</sub>$``$$`c`$, at different mw power levels is observed within the accuracy of the measurement. It means that H<sub>c</sub> is independent of the amplitude of the input microwave signal. However, the decrease in R<sub>s</sub> is more prominent than that of P=$``$10dBm. The minimum depth in this case is as large as 43% of the zero-dc field R<sub>s</sub> value.
The same measurements were also carried out for different resonant modes (second and third harmonics), shown in Fig.3(a). The field dependencies of R<sub>s</sub> for different resonant modes are qualitatively the same. R<sub>s</sub> first decreases with H<sub>dc</sub> and then increase after H<sub>dc</sub> reaches the crossover field H<sub>c</sub>. The crossover fields H<sub>c</sub> measured from the second and third modes are almost the same as that from the fundamental mode. Thus H<sub>c</sub> is also independent of the frequency of the input mw signals. As is well known, the frequency dependence of R<sub>s</sub> is an important key in the determination of the microwave loss mechanism. To shed some light on the microwave loss mechanism in the applied dc magnetic field with different magnitudes, we chose three typical H<sub>dc</sub>: zero dc field, H<sub>c</sub> and a field larger than H<sub>c</sub> to plot the $``$3dB bandwidth as a function of the harmonic numbers, as shown in Fig.3(b). Nearly linear frequency dependencies of R<sub>s</sub> are found for all the three values of H<sub>dc</sub> and even the slopes of the three curves are almost independent of H<sub>dc</sub>. The similarity in the frequency dependence of R<sub>s</sub> would suggest the microwave loss mechanisms at the three dc fields are essentially the same as each others. The linear frequency dependence rules out the loss mechanisms with f<sup>2</sup> dependence, such as two fluid model , BCS theory and weakly coupled grain model . And it implies that the hysteretic losses due to pinning and nucleation of Josephson fluxons may play an important role in the loss mechanism.
Although the loss mechanism is almost same in our measurements, the effect of H<sub>dc</sub> on the loss mechanism is totally different for the case H<sub>dc</sub>$`<`$ H<sub>c</sub> and H<sub>dc</sub>$`>`$H<sub>c</sub>. To illustrate the effect of the applied dc field, we adopt an impedance plane analysis, which has been proven to be a powerful approach in distinguishing between various nonlinear mechanisms of superconductors, in terms of r-parameter . Here the dc field-induced changes in the surface impedance are characterised by a dimensionless r<sub>H</sub> which is defined as the ratio of the change of surface resistance and reactance with varing H<sub>dc</sub> at fixed input mw powers, i.e., r<sub>H</sub> = $`\mathrm{\Delta }`$R<sub>s</sub>(H<sub>dc</sub>)/$`\mathrm{\Delta }`$X<sub>s</sub>(H<sub>dc</sub>). Figure 4 demonstrates $`\mathrm{\Delta }`$f<sub>-3dB</sub> versus f<sub>0</sub> dependencies at P<sub>mw</sub>=$``$10dBm for varied H<sub>dc</sub> values and for different orientations of the applied dc magnetic field. The r<sub>H</sub>-parameter can be easily extracted from the slopes of the curves (r<sub>H</sub> is one-half of the slope). The obtained value for r<sub>H</sub> is about 0.6 if H<sub>dc</sub> is below H<sub>c</sub>, while r<sub>H</sub>$``$0.1 if the dc field is above H<sub>c</sub>. The transition of r<sub>H</sub> happens within a narrow field range at the crossover dc field. The prominent difference in the values of r<sub>H</sub> implies that the dc field below H<sub>c</sub> plays a different role with that above H<sub>dc</sub> in determining the surface impedance. It is noted that the r<sub>H</sub>$``$0.1 at H<sub>dc</sub>$`>`$ H<sub>c</sub> is consistent with the r<sub>H</sub> values (0.1 - 0.2) reported by other groups . The impedance plane analysis at P<sub>mw</sub>=$``$20dBm presents a similar behavior. It shound be stressed that Fig.4 is qualitatively different from the similar analysis given in Fig.5 of Ref. 17. This suggests that the anomalous response reported by Hein $`et`$ $`al`$. and what observed in this work may come from different origins.
In addition, Fig.4 shows that for different orientations of H<sub>dc</sub> ( H<sub>dc</sub>$``$$`c`$ and H<sub>dc</sub>$``$$`c`$ ) the two sets of data points essentially coincide with each other on the same curve. As shown in the insets of Figure 1, the R<sub>s</sub>(H<sub>dc</sub>) and X<sub>s</sub>(H<sub>dc</sub>) curves with different H<sub>dc</sub> orientations can also be normalised to coincide with each other. Combining these observations together, we conclude that the effects of H<sub>dc</sub> on the surface impedance are essentially independent on the orientations of the applied dc magnetic field.
Three samples were measured in this work. Though quantitative differences are found, the features of the data are essentially the same as mentioned above. The quantitative diferences may come from the differences on growth, deposition and patterning of the films.
## IV Discussions
The results show that H<sub>c</sub> is independent of both the amplitude and frequency of the applied microwave signal. It implies that H<sub>c</sub> may be related to some characteristic parameters of the YBCO films. Since the effects of dc magnetic field change dramatically at H<sub>c</sub>, one could naturally expect that the observed H<sub>c</sub> in the measurements is simply a manifestation of H<sub>c1</sub>, the lower critical field of the superconducting YBCO films. Below H<sub>c</sub>, the external dc field causes a decrease in R<sub>s</sub> for some reasons. Above H<sub>c</sub>, the dc fluxons begin to penetrate into the samples and produce additional loss with a large increase of the surface impedance. This may give an qualitative description on some features of the anomalous effect mentioned above. However, we can not definitely clarify the relationship of H<sub>c</sub> to H<sub>c1</sub> while the origin of the drop in R<sub>s</sub> is still not known. The conception, which says the observed H<sub>c</sub> is a manifestation of H<sub>c1</sub>, needs to be further studied.
Recently, several groups proposed that magnetic field-induced recovery of superconductivity might account for the anomalous microwave response . Magnetic impurities are likely to be present in most HTS materials. The interaction between localised magnetic moments of magnetic impurities and cooper pairs which are in the singlet state destroys the pair correlation and is accompanied by spin-flip scattering . An external magnetic field forces the localised magnetic moments to align, frustrates the spin-flip scattering, and leads to a reduction of pair breaking. According to two-fluid model the increase of pair electrons can lead to the decrease of R<sub>s</sub> and X<sub>s</sub>. This mechanism has been proven to be effective in explaining some of the experimental results . However, it is obvious that the feature of Fig.4 is qualitatively different from that of the similar analysis given in Fig.5 of Ref.. So it is doubtful that the anomalous response we observed comes from the same origin as that in Ref.. In addition, a phenomenological description is proposed in terms of this mechanism . Simulation shows that quantitative fitting of the experimental data with this model requires a large increase of pair density ($`>`$50%) with a relatively small change of external field (80$``$90 Gauss). So the applicability of this mechanism to the present experimental results is questionable.
A phenomenological model of the nonlinear microwave response of a superconducting weak link was proposed by Velichko to describe the effect of both dc and mw magnetic fields . The results show that the value of Z<sub>s</sub>(H) of both “the non-shunted” and “the shunted WL” can fall with increased H under certain conditions. However, the observed correlated decrease in R<sub>s</sub> and X<sub>s</sub> still can not be explained by the model. For “the non-shunted WL”, R<sub>s</sub> and X<sub>s</sub> initially increase as H raises from zero and then decrease with elevated magnetic field, which is inconsistent with our experimental observation. For “the shunted WL”, a decrease in R<sub>s</sub> is accompanied by an increase of X<sub>s</sub>. This is also inconsistent with our data. As shown by Herd $`et`$ $`al`$., however, if a network of weak links with a distribution of the I<sub>c</sub>R<sub>n</sub> products was considered, the decrease of X<sub>s</sub> with H can be expected . It is not sure whether a network of “the shunt WL” can account for the decrease of both R<sub>s</sub> and X<sub>s</sub>.
Similar effects were once observed by Thompson $`et`$ $`al`$ . in the measurements of ac hysteretic losses in Nb<sub>3</sub>Ge material. With fixed ac amplitude, the ac loss was observed initially decrease and then increase with elevated dc-bias magnetic field after passing through a minimum. The decrease of ac losses was explained in terms of Abrikosov vortex-antivortex annihilation within the frame of critical state model . One of the central features of this theory lies in an ac amplitude-dependent H<sub>c</sub> at which ac loss reaches its minimum. While this feature was observed in Thompson’s experiment, it is opposite to our observation. As we mentioned above, H<sub>c</sub> is almost independent of the amplitude of the input mw signal.
So far the observed anomalous microwave response can not be explained consistently by any of the proposed models. It is generally expected that both of the intrinsic and extrinsic effects in the HTS thin films may have their contributions to the microwave loss. To get a clear understanding of the anomalous effect, all these effects should be taken into account. The high value of R<sub>s</sub> ($``$ 1m$`\mathrm{\Omega }`$ at 10.66 GHz, 77K) shows that the films used in this work are highly granular and most likely consist of a microbridge-type weak links. The weak links may play an important role in the anomalous effects though the mechanism is not known at present. As we know, granularity has been ruled out as a general source of the anomaly observed in their experiments since their films showed well-established qualities and excellent power handling capabilities. This can also explain why Fig.4 is qulitatively different from the Fig.5 in Ref.17. This suggest that the anomalous effects may have more than one source.
## V Conclusions
We have observed an anomalous microwave response in HTS thin-film microstrip resonators. The main results of this work are summarized as follows: (a) The surface resistance and reactance show a correlated non-monotonic behavior in the presence of a weak dc magnetic field. The R<sub>s</sub> drops as H<sub>dc</sub> raises from zero, after passing through a minimum, it increases gradually as H<sub>dc</sub> is further increased. The X<sub>s</sub> presents a similar behavior; (b) The crossover field H<sub>c</sub> , at which R<sub>s</sub> reaches its minimum, is independent of the amplitude and frequency of the input microwave signal within the ranges of measurements; (c) The qualitative effects of H<sub>dc</sub> on the surface impedance are essentially independent of dc field orientation, microwave amplitude and frequency, while prominently different at H<sub>dc</sub>$`<`$H<sub>c</sub> (r<sub>H</sub>$``$0.6) and at H<sub>dc</sub>$`>`$H<sub>c</sub> (r<sub>H</sub>$``$0.1); (d) At a fixed H<sub>dc</sub> and P<sub>mw</sub>, the frequency dependence of the surface resistance is almost linear. This linear frequency dependence and even the slope of R<sub>s</sub>-f are unchanged at varied H<sub>dc</sub> magnitudes ($`<`$H<sub>c</sub>, $`=`$H<sub>c</sub> and $`>`$H<sub>c</sub>). Several mechanisms were examined and they can not give a satisfactory explanation on the results. The origin of the observed anomaly is still not clear. Further efforts should be addressed to understand it fully.
|
no-problem/9908/astro-ph9908134.html
|
ar5iv
|
text
|
# HS 1603+3820: a bright 𝑧ₑₘ=2.51 quasar with a very rich heavy element absorption spectrum
## 1 Introduction
Quasars often exhibit narrow heavy element absorption lines near their emission redshift. One possible explanation for the origin of these systems is that they arise in clouds of matter associated with galaxies in the clusters surrounding the quasars. Alternate possibility is that they originate in the clouds that are physically associated with the quasars themselves. It has been shown that both these scenarios may be true (e.g. Ellingson et al. ell1994 (1994); Hamann et al. 1997a ).
Clusters of absorbers of any type are of particular interest. On one hand, clustering properties of intervening systems depend strongly on the type of absorbers, while on the other hand complexes of associated absorbers give a unique opportunity to analyze the properties of their hosts and of the quasar emission.
The stumbling block in such studies is the fact that bright high redshift quasars with rich absorption are rare (for examples of such systems see, e.g., Morris et al. mor1986 (1986); Foltz et al. fol1987 (1987); Petitjean et al. pet1994 (1994); Hamann et al. 1997a ; Lespine & Petitjean les1997 (1997); Petitjean & Srianand pet1999 (1999)). There are only a few quasars known with more than just a handful of associated absorption systems which are bright enough to perform high resolution spectral analysis.
In this paper we present a bright $`z_{\mathrm{em}}=2.51`$ quasar HS 1603+3820 ($`\alpha _{1950}=\text{16:03:07.7},\delta _{1950}=\text{+38:20:07}`$) with a very rich metal absorption spectrum. It was discovered during the course of the Hamburg/CfA Bright Quasar Survey (Hagen et al. hag1995 (1995); Dobrzycki et al. dob1996 (1996)). Quasar candidates in the survey are selected from the objective prism spectra taken with the Hamburg Schmidt telescope at Calar Alto, Spain. Follow-up low resolution spectroscopy is performed with the 1.5-m Tillinghast reflector with the FAST spectrograph at Fred Lawrence Whipple Observatory on Mount Hopkins, Arizona.
The discovery spectrum and other basic information on HS 1603+3820 will be presented in the forthcoming survey paper (Engels et al. eng1999 (1999)). In this paper we present a more detailed analysis of the MMT spectrum of HS 1603+3820. Its very unusual properties prompted us to make the quasar available to interested researchers prior to the publication of the survey.
To our knowledge, there are no radio or X-ray sources near the position of HS 1603+3820.
The finding chart for HS 1603+3820 is shown on Fig. 1. HS 1603+3820 was selected for followup studies because of its unusually high brightness ($`B=15.9`$) for a high-$`z`$ quasar and because the discovery spectrum hinted that there could be absorption systems in the vicinity of the emission lines.
Observations performed with the Multiple Mirror Telescope<sup>1</sup><sup>1</sup>1MMT is a joint facility of the Smithsonian Institution and the University of Arizona. revealed a very rich heavy element absorption spectrum. Several metal systems, both intervening and associated, are present, including a couple of systems with $`z_{\mathrm{abs}}>z_{\mathrm{em}}`$. A unique feature is a rich complex of at least five Civ absorbers near the emission redshift of the quasar.
This paper is organized as follows. In Sect. 2 we present the MMT observations of HS 1603+3820. In Sect. 3 we present the metal absorption systems. In Sect. 4 we discuss the of Civ absorbers at $`z_{\mathrm{abs}}z_{\mathrm{em}}`$ and present arguments that they likely are associated with the quasar. We summarize our results in Sect. 5.
## 2 Observations
Spectra presented here were obtained during two nights, April 12-13, 1997, at the Multiple Mirror Telescope (MMT) on Mt. Hopkins, Arizona. We used the Blue Channel spectrograph, with the 1200 l/mm grating in first order, 1.25$`\times `$3 arcsec slit and the 3k$`\times `$1k Loral CCD, binned by two in the spatial direction. Weather and seeing were very good during both nights. The wavelength range of 3530–5050 Å, which contains the Ly-$`\alpha `$ forest part of the HS 1603+3820 spectrum, was observed for the total of 3300 seconds during two nights. On the second night another 1200 second exposure was obtained for the wavelength range containing the Civ emission line, 4840–6350 Å. Spectra have 0.495 Å/pixel binning and spectral resolution of $``$3 pixels. The S/N ratio per resolution element in the “Ly-$`\alpha `$ spectrum” varies from $``$40 to $``$100, and in the “Civ spectrum” it is roughly uniform at the level of 45–50.
Continua were fitted to the spectra and the absorption line lists were generated following the procedure similar to outlined in Scott et al. (1999a , 1999b ), where one can also find the analysis of the Ly-$`\alpha `$ forest spectrum of HS 1603+3820. Interested researchers can access the line lists (as well as finding chart and the spectra – both plots and digital versions) on the World Wide Web at http://hea-www.harvard.edu/QEDT/Papers/hs1603/.
Full spectra can be seen in Figs. 2 and 3. Fig. 4 shows in detail the vicinity of the Ly-$`\alpha `$ and Civ emission lines.
It has to be noted that the continuum level in the region near 4250 Å, which occurs near the peak of the Ly-$`\alpha `$ emission line, is quite uncertain, making the emission redshift of the quasar somewhat uncertain, since an important part of the line is seriously affected by complex absorption. (Please note that this absorption is not related to Broad Absorption Line phenomenon – the absorption lines affecting the Ly-$`\alpha `$ emission profile are not Ly-$`\alpha `$.) We adopted the emission redshift of the quasar of 2.51, keeping in mind the well known fact that broad emission lines of quasars can be blueshifted by as much as $``$1000 km s<sup>-1</sup> with respect to “true” quasar redshift, as determined from narrow forbidden lines (see e.g. Espey esp1993 (1993)). It should be noted (and it will be discussed in a little more detail below) that there are absorption lines at redshift as high as 2.55, but they clearly are on the red wing of the emission lines, well past the peak emission.
## 3 Heavy element absorption systems
As expected in a quasar with $`z_{\mathrm{em}}=2.51`$, there are numerous absorption features in the Ly-$`\alpha `$ forest part of the spectrum. The Ly-$`\alpha `$ forest of HS 1603+3820 has been included in the large sample of Scott et al. (1999b ). In the present paper we concentrate on the heavy element absorption lines, presence of which is clearly seen on top of the emission lines and redwards of the Ly-$`\alpha `$ emission.
Metal systems in the spectrum were searched for in two ways. First, an attempt was made to interactively identify all absorption lines redwards of the Ly-$`\alpha `$ emission line. Second, heavy element line searching code METALS, written by Jill Bechtold and kindly provided to us by the author, was applied. In all, thirteen heavy element absorption systems were found, though the reality of two of them is somewhat shaky. Notes on individual systems follow.
$`z_{\mathrm{abs}}=1.8882`$: This is the first of the two uncertain systems. Its identification relies primarily on the presence of a doublet which has the wavelength ratio matching the ratio for the Civ doublet. Though this match is indeed very good, both lines are possible blends with lines from other systems. Two other identified lines, Nv(1242) and Feii(1608), are also blends with lines from other systems. The wavelength corresponding to the Ly-$`\alpha `$ line lies outside of the spectrum.
$`z_{\mathrm{abs}}=1.9650,2.0703,2.1762`$: All these systems are unambiguously identified, showing well resolved Ly-$`\alpha `$, Civ, and other lines.
$`z_{\mathrm{abs}}=2.4189,2.4270,2.4367,2.4420,2.4523`$: These are the confirmed systems that have their Civ lines in the complex in the blended region near 5320 Å. See Sect. 4 below for a more detailed discussion of the complex.
$`z_{\mathrm{abs}}=2.4794`$: This is a very strong system, just below the emission redshift of the quasar. As many as seventeen absorption lines belonging to this system were identified. The Siii(1526) line lies inside the Civ complex near 5320 Å, but is easily recognizable. Very strong Ly-$`\alpha `$ and Civ lines are prominent on Fig. 4. The Ly-$`\alpha `$ absorption line, which is somewhat asymmetric on the red wing, is in fact a blend of (at least) two lines: very strong Ly-$`\alpha `$ proper, and a weaker line, which appears to be a member of another metal system(s).
$`z_{\mathrm{abs}}=2.5114`$: This is the second of the two uncertain systems. If it is real, it is almost exactly a $`z_{\mathrm{abs}}=z_{\mathrm{em}}`$ system. It has six lines at correct wavelengths, including three Siii lines. The Ly-$`\alpha `$ line lies in the middle of the blended region near 4250 Å. There is, however, no significant Civ absorption. Dotted marker on the lower panel of Fig. 4 shows the expected location of the Civ doublet; the 5$`\sigma `$ rest equivalent width threshold at this location is 0.07 Å. Also, there are two other issues that make the identification of this system uncertain. First, the system appears to contain a somewhat unusual combination of lines. Second, the ratios of line strengths of the Siii lines appear to be incorrect (Morton mor1991 (1991)). Acknowledging that especially the second of these problems casts doubts on the reality of this system, we hesitate, however, to entirely discard it, since the identified Siii lines may be blended with other lines, not to mention the fact that the environment in the vicinity of the quasar central engine is likely to be highly unusual.
$`z_{\mathrm{abs}}=2.5374,2.5541`$: These are the two systems with $`z_{\mathrm{abs}}>z_{\mathrm{em}}`$ seen in the spectrum of HS 1603+3820. Both systems have Ly-$`\alpha `$, Civ, and other metal lines clearly resolved. See below for discussion.
## 4 Complex of metal systems at $`z_{\mathrm{abs}}z_{\mathrm{em}}`$: intervening or associated?
The most interesting feature in the spectrum of HS 1603+3820 is the large number of systems with $`z_{\mathrm{abs}}z_{\mathrm{em}}`$. Nine of the metal systems are within 8000 km s<sup>-1</sup> of the quasar emission redshift. Of these, the most spectacular is a complex of at least five Civ absorption systems, all within 3000 km s<sup>-1</sup> from one another. Other systems may be present in the complex, but limited resolution of our data does not allow us to claim that at a reasonable level of confidence.
The average ejection velocity of the complex, i.e. displacement from the redshift of the quasar in velocity units, is $``$6500 km s<sup>-1</sup>. In principle, large ejection velocities would indicate that these systems are intervening, i.e. are not associated physically with the quasar. On the other hand, convincing arguments were presented (Hamann et al. 1997b , Richards et al. ric1999 (1999)) that absorbers with much larger ejection velocities could be intrinsic to the quasar.
Of the three methods for distinguishing between intrinsic and intervening systems (see Barlow & Sargent bar1997 (1997), Hamann et al. 1997a , and references therein) – details of shapes of absorption lines, temporal variability of line strengths, and partial covering of quasars by the absorbers – only the third can be applied in our case. The first is precluded by insufficient resolution of our data. The second was made impossible by decommissioning of the MMT, since we could not observe the quasar again with the same telescope/detector combination, which is a standard approach in such a case. We can, however, attempt to estimate whether at least one of the systems in the complex obscures the emission source in the quasar entirely. If it does not, it is a strong indication that the cloud in which the absorption system originates is physically associated with the quasar.
The Civ doublet at $`z_{\mathrm{abs}}=2.4189`$, which has $`v_{\mathrm{ej}}8000`$ km s<sup>-1</sup>, is at the high ejection velocity end of the Civ complex. The lines from this doublet fulfill all the conditions needed for the quasar covering factor criterion: they are satisfactorily well resolved, they are markedly broader than the spectrum resolution, and they are not contaminated by absorption lines from any other known systems. If these two lines are indeed resolved, we find the Civ(1548)/Civ(1550) ratio in this system to be close to 1, which indicates that the system is optically thick. At the same time, the lines do not reach zero intensity in their bottoms, which suggests that the cloud in which the lines originate does not obscure the entire continuum emitting source. That in turn suggests that this absorption system is intrinsic to the quasar. The lower ejection velocity systems in the complex also do not reach zero intensity, which makes it reasonable to assume that they are also associated with the quasar (even though we cannot unambiguously establish whether they are saturated).
There are three (or four, if one includes the uncertain $`z_{\mathrm{abs}}=z_{\mathrm{em}}`$ system) other heavy element systems with ejection velocities lower than the complex. Two of these systems are clearly infalling: even if the emission redshift of the quasar is indeed underestimated by as much as 1000 km s<sup>-1</sup>, these systems still have $`z_{\mathrm{abs}}>z_{\mathrm{em}}`$; their infall velocities are $``$2000 and $``$3600 km s<sup>-1</sup> with respect to adopted emission redshift of the quasar.
Arguments presented above suggest that all eight (or nine) systems are physically associated with the quasar. If this is the case, then all systems need to be considered as one large associated absorption complex, with dispersion of $``$4300 km s<sup>-1</sup>.
It appears that the absorption originates in clouds in the immediate vicinity of the quasar. The ejection velocities and the velocity dispersion of the complex (both of the order of thousands of km s<sup>-1</sup>) appear to be too high to interpret the absorbers as originating in the halos of galaxies in the quasar host cluster.
## 5 Summary
The combination of high redshift, brightness, richness of the heavy element absorption – and associated absorption in particular – in the spectrum of HS 1603+3820 is truly unique. Eleven confirmed Civ systems in the spectrum of HS 1603+3820 ranks among the richest known. The York catalog (York et al. yor1991 (1991), Richards et al. ric1999 (1999)) contains only 10 other quasars with a greater number of absorbers. Of these quasars, only two (Q0958+551 and Q1225+317) are of comparable brightness, but both are at considerably lower redshifts.
HS 1603+3820 is even more spectacular when associated absorption spectrum is concerned. York’s catalog contains only three quasars which have eight or more systems within 8000 km s<sup>-1</sup> of the emission redshift (Q1037–270, Q1511+091 and Q1556+335). All three are much fainter; among QSOs brighter than 16 mag no object comes close to HS 1603+3820.
We can hypothesize that since none of the absorbers in the complex appears to be drastically different than the other ones they all may have been produced by a single event in the quasar’s past, and that they may represent the velocity dispersion of shells of matter ejected during this event. On the other hand, the spectrum also contains an absorber which is much stronger than the systems from the complex, as well as two infalling absorbers, which indicate that the environment of HS 1603+3820 is more complicated.
We stress that this paper presents an analysis based on the observations in relatively low resolution. Our conclusions are by necessity mostly qualitative since the resolution is inadequate for performing detailed studies of chemical composition or velocity structure of the individual systems. High resolution studies of the complex will be of special interest, since it is heavily blended in our data and it is very likely that high resolution spectrum will reveal more absorption systems. HS 1603+3820 is very bright for a $`z=2.51`$ quasar and is therefore an excellent candidate for such observations.
###### Acknowledgements.
We would like to thank T. Aldcroft, J. Bechtold, D. Dobrzycka, M. Elvis, S. Mathur, J. Scott, and A. Siemiginowska for helpful discussions, A. Milone for assistance at the MMT, and F. Drake for providing good working environment for this project. T. Aldcroft and J. Bechtold wrote computer codes that were used in the analysis. A.D. acknowledges support from NASA Contract No. NAS8-39073 (Chandra X-Ray Observatory Center). The Hamburg Quasar Survey is supported by the DFG through grants Re 353/11 and Re 353/22.
|
no-problem/9908/physics9908051.html
|
ar5iv
|
text
|
# Creation of an Ultracold Neutral Plasma published in Physical Review Letters 83, 4776 (1999)
## Abstract
We report the creation of an ultracold neutral plasma by photoionization of laser-cooled xenon atoms. The charge carrier density is as high as $`2\times 10^9`$ cm<sup>-3</sup>, and the temperatures of electrons and ions are as low as $`100`$mK and $`10\mu `$K, respectively. Plasma behavior is evident in the trapping of electrons by the positive ion cloud when the Debye screening length becomes smaller than the size of the sample. We produce plasmas with parameters such that both electrons and ions are strongly coupled.
The study of ionized gases in neutral plasma physics spans temperatures ranging from $`10^{16}`$ K in the magnetosphere of a pulsar to $`300`$ K in the earth’s ionosphere . At lower temperatures the properties of plasmas are expected to differ significantly. For instance, three-body recombination which is prevalent in high temperature plasmas, should be suppressed . If the thermal energy of the particles is less than the Coulomb interaction energy, the plasma becomes strongly coupled, and the usual hydrodynamic equations of motion and collective mode dispersion relations are no longer valid . Strongly coupled plasmas are difficult to produce in the laboratory and only a handful of examples exist , but such plasmas do occur naturally in astrophysical systems.
In this work we create an ultracold neutral plasma with an electron temperature as low as $`T_e=100`$ mK, an ion temperature as low as $`T_i=10\mu `$K, and densities as high as $`n=2\times 10^9`$cm<sup>-3</sup>. We obtain this novel plasma by photoionization of laser-cooled xenon atoms. Within the experimentally accessible ranges of temperatures and densities both components can be simultaneously strongly coupled. A simple model describes the evolution of the plasma in terms of the competition between the kinetic energy of the electrons and the Coulomb attraction between electrons and ions. A numerical calculation accurately reproduces the data.
Photoionization and laser-cooling have been used before in plasma experiments. Photoionization in a $`600`$K Cs vapor cell produced a plasma with $`T_e2000`$K , and a strongly coupled non-neutral plasma was created by laser-cooling magnetically trapped Be<sup>+</sup> ions .
A plasma is often defined as an ionized gas in which the charged particles exhibit collective effects . The length scale which divides individual particle behavior and collective behavior is the Debye screening length $`\lambda _D`$. It is the distance over which an electric field is screened by redistribution of electrons in the plasma, and is given by $`\lambda _D=\sqrt{ϵ_0k_BT_e/e^2n}`$. Here, $`ϵ_0`$ is the electric permittivity of vacuum, $`k_B`$ is the Boltzmann constant, and $`e`$ is the elementary charge. An ionized gas is not a plasma unless the Debye length is smaller than the size of the system . In our experiment, the Debye length can be as low as $`500`$nm, while the size of the sample is $`\sigma 200\mu `$m. The condition $`\lambda _D<\sigma `$ for creating a plasma is thus easily fulfilled.
The atomic system we use is metastable xenon in the $`6s[3/2]_2`$ state. This state has a lifetime of $`43`$s and can be treated as the ground state for laser-cooling on the transition at $`882`$nm to the $`6p[5/2]_3`$ state . The metastable atoms are produced in a discharge and subsequently decelerated using the Zeeman slowing technique. The atoms are then collected in a magneto-optical trap and further cooled with optical molasses to approximately $`10\mu `$K. We characterize the cold neutral atoms by optical absorption imaging . This measurement provides the density and size of the atomic cloud and the number of atoms in the sample. Typically we prepare a few million atoms at a density of $`2\times 10^{10}`$ atoms/cm<sup>3</sup>. Their spatial distribution is Gaussian with a rms radius $`\sigma 200\mu `$m.
We partially ionize the cold atom sample via two photon excitation. A pulse of light from the cooling laser at $`882`$ nm populates the $`6p[5/2]_3`$ level. Green photons ($`\lambda =514`$ nm) from a pulsed dye laser, pumped by a frequency-tripled pulsed Nd:YAG laser, then excite atoms to states at or above the ionization potential.
The energy difference, $`\mathrm{\Delta }E`$, between the photon energy and the ionization potential is distributed between the electrons and ions. Because of the large ion to electron mass ratio, all except $`4\times 10^6\mathrm{\Delta }E`$ is given to the electrons. Equipartition of energy between ions and electrons requires tens of ms . We vary $`\mathrm{\Delta }E/k_B`$ in a controlled manner between $`0.11000`$ K by changing the green laser frequency. The bandwidth of the laser of $`0.07`$ cm<sup>-1</sup> sets the lower limit.
By adjusting the pulse energy of the green laser, we control the number of atoms photoionized. For the highest energy available, $`1`$ mJ in a $`10`$ns pulse, we can produce up to $`2\times 10^5`$ ions, which corresponds to a peak density of $`2\times 10^9`$ cm<sup>-3</sup>. The number of atoms photoionized varies linearly with the laser intensity. Although the ionization fraction is low ($`10`$%), the charged particles show no evidence of interactions with the neutral atoms. This is to be expected because the mean free path for neutral-charged particle collisions is much greater than the sample size .
For detection of charged particles, an external electric field is applied to direct ions towards a microchannel plate detector and electrons towards a single channel electron multiplier. The efficiencies are 50% for ions and 85% for electrons. The magnitude of the applied electric field is calibrated through field ionization of Rydberg atoms .
In each cycle of the experiment the atoms are first laser cooled and an electric field of approximately $`0.005`$V/cm is applied. The atoms are then photoionized, and after about $`500`$ns of time of flight a pulse of electrons arrives at the detector (see Fig. 1). If the green laser energy is high enough, the first peak develops a tail, and a second peak appears when the electric field is linearly increased a few microseconds later. On this time scale the ions are essentially stationary. About $`300\mu `$s after the electric field ramp is applied, they are detected on the microchannel plates.
A simple model (Fig. 2) explains the experimental data. The charge distribution is everywhere neutral immediately after photoionization. Due to the initial kinetic energy of the electrons ($`\mathrm{\Delta }E`$), the electron cloud begins expanding. The resulting local charge imbalance creates an internal electric field which produces a Coulomb potential energy well for electrons. If the well never becomes deeper than the initial kinetic energy, all the electrons escape. This corresponds to the uppermost curve (i.e. lowest laser intensity) in Fig. 1. If enough atoms are photoionized, however, only an outer shell of electrons escapes, and the well becomes deep enough to trap the rest. Electrons in the well redistribute their energy through collisions within $`10100`$ns . As charges are promoted to energies above the trap depth they leave the well. This explains the tail of the first peak in the electron signal. During this process of evaporation the potential well depth increases. Evaporation eventually slows and remaining electrons are held until an applied electric field overcomes the trapping potential. They appear as the second peak in Fig. 1.
This description suggests that for a given $`\mathrm{\Delta }E`$ there is a threshold number of positive ions required for trapping electrons. The data show such behavior in a plot of the fraction of electrons trapped versus the number of photoions produced (Fig. 3a). As $`\mathrm{\Delta }E`$ increases, more positive charges are required to produce the trapping effect.
At the trapping threshold, after all the electrons have left, the potential well depth equals the initial kinetic energy of the electrons. From this relation one can calculate the number of positive ions at the threshold, $`N^{}=\mathrm{\Delta }E/U_0`$. Here, $`U_0=\sqrt{2/\pi }e^2/4\pi \epsilon _0\sigma `$, and $`\sigma `$ is the rms radius of the Gaussian spatial distribution of positive ions.
The data agree with this simple calculation. For a wide range of electron kinetic energies, the onset of trapping occurs at $`N=N^{}`$, as shown in Fig. 3b. Scaling the number of ions produced by $`N^{}`$ shows that all data fall on a universal curve . A numerical simulation which approximates the initial velocity, $`v=\sqrt{2\mathrm{\Delta }E/m}`$, as directed radially outward and integrates the equations of motion for the electrons, reproduces this behavior.
The model described above (Fig. 2) implies that the temperature of the electrons is $`T_e`$$`\stackrel{<}{}`$$`\mathrm{\Delta }E/k_B`$. This is confirmed by the numerical simulation. In principle, the energy distribution of the trapped electrons can also be determined from the shape of the second peak in Fig. 1. However, the analysis is complicated because the trap depth increases as electrons are removed. Also, rethermalization is fast and the temperature changes on the timescale of the electric field ramp.
The initial ion temperature is easily estimated. For excitation close to the ionization threshold, the energy imparted to the ions from photoionization is negligible compared to the kinetic energy of the atoms. Therefore the minimum initial temperature is $`10\mu `$K. For large $`\mathrm{\Delta }E`$ the temperature approaches $`4\times 10^6\mathrm{\Delta }E/k_B`$, which is $`4`$mK for $`\mathrm{\Delta }E=1000`$K. Although the equilibration time is on the order of tens of ms, collisions with the energetic electrons are expected to approximately double the ion temperature within a $`\mu `$s.
The threshold condition $`N=N^{}`$ is mathematically equivalent to $`\lambda _D=\sigma `$. In this context, one can interpret $`\lambda _D`$ as the displacement of electrons from their equilibrium positions when their energy in the local internal electric field in the plasma equals their kinetic energy . If $`\lambda _D>\sigma `$, the electrons are free to escape to infinity. If $`\lambda _D<\sigma `$, electrons are trapped in the ion cloud by the internal field and a plasma is formed.
After the untrapped fraction of electrons has escaped, the cloud as a whole is no longer strictly neutral. But as mentioned above, electrons escape most easily from the edges of the spatial distribution, and for $`N>N^{}`$ the center of the cloud is well described as a neutral plasma. This behavior is also seen in the numerical simulation.
The only significant effect of the residual charge im-
balance is a Coulomb expansion of the cloud that occurs on a long time scale of many microseconds. This limits the time available for studying the highest density conditions. The expansion also decreases the potential well depth, allowing formerly trapped electrons to escape. In a plasma with 5000 ions and 10% charge imbalance, half of the initially trapped electrons escape in about 100 $`\mu `$s. The expansion is slowed compared to what would be observed for a bare cloud of positive charges, however. A bare cloud of 5000 ions, initially with $`\sigma =200`$ $`\mu `$m, expands to twice its radius in a few microseconds, reducing the well depth by a factor of two.
Phenomena similar to the electron trapping observed here are seen in traditional plasmas. For instance, recombination can often occur at containment walls and leads to net charge diffusion from the center of the plasma. The mobility of the electrons is larger than that of the ions, but their motion is retarded by local internal electric fields which develop from any charge imbalance. This leads to ambipolar diffusion in which electrons and ions migrate at equal rates.
In our ultracold plasma the thermal energy of the charged particles can be less than the Coulomb interaction energy between nearest neighbors, making it strongly coupled. The situation is characterized quantitatively by the electron and ion Coulomb coupling parameters $`\mathrm{\Gamma }_e=(e^2/4\pi \epsilon _0a)/k_BT_e`$ and $`\mathrm{\Gamma }_i=\text{e}^{a/\lambda _D}\mathrm{\Gamma }_eT_e/T_i`$ where $`a=(4\pi n/3)^{1/3}`$ is the Wigner-Seitz radius. The exponential term in the expression for $`\mathrm{\Gamma }_i`$ is due to the shielding of the ion-ion interaction by electrons . When $`\mathrm{\Gamma }>1`$, many-body spatial correlations exist and phase transitions such as crystallization of the sample may occur. Systems with $`\mathrm{\Gamma }_e>1`$ are sometimes called non-Debye plasmas because $`\lambda _D<a`$.
For the densities and temperatures accessed in the experiment, we can prepare a plasma in which both electrons and ions are initially strongly coupled: $`\mathrm{\Gamma }_e=10`$ and $`\mathrm{\Gamma }_i=1000`$. To our knowledge such a system has never been created before.
This novel plasma is well suited for a wide range of experiments. Plasma oscillations , which have a frequency $`f_p=\sqrt{ne^2/mϵ_0}/2\pi `$ of up to $`400`$MHz, can be used to probe the density distribution of the system. Magnetic confinement may greatly extend the plasma lifetime, and because of the low sample temperature, the required field should be small. The thermalization and evaporative cooling of electrons, and the temperature of the ions require further study.
We can also look for three-body electron-ion recombination. At higher temperatures, the rate for this process scales as $`T^{9/2}`$, and an extrapolation to the present experimental conditions yields a recombination time of nanoseconds (for $`T_e=1`$K and $`n=2\times 10^8`$cm<sup>-3</sup>). The long lifetime we observe ($`100\mu `$s) is the first clear indication that this theory, as well as an extension to $`T1`$K , breaks down for the temperatures studied here.
The initial kinetic energy of the electrons can be reduced to $`10`$ mK by using a laser with a bandwidth equal to the Fourier transform limit of a 10 ns pulse. One may be able to decrease this energy even further by exciting below the ionization limit. In this case one creates a dense gas of highly excited cold Rydberg atoms for which many-body interactions can cause a phase transition to a plasma-like state . In preliminary experiments we have observed the formation of free electrons and ions in such a system. This will be the subject of future experiments.
The technique to produce ultracold plasmas demonstrated in this work is applicable to any atom that can be laser-cooled, and other atoms may offer experimental advantages. With alkali systems one can attain higher initial densities, and alkaline earth ions have accessible optical transitions.
To summarize, we have accessed a new region in the parameter space of neutral plasmas by photoionizing a cloud of laser cooled atoms. Conditions were realized in which both electrons and ions are strongly coupled. Experimentally, the initial plasma properties are easily controlled and the evolution of the system is described with an uncomplicated model.
S. Kulin acknowledges funding from the Alexander-von-Humboldt foundation, and T. C. Killian is supported by a NRC postdoctoral fellowship. This work was supported by ONR.
|
no-problem/9908/cond-mat9908171.html
|
ar5iv
|
text
|
# DELINEATION OF THE NATIVE BASIN IN CONTINUUM MODELS OF PROTEINS
Chains of beads on the cubic or square lattices, with some effective interactions between the beads, often serve as simple models of proteins (see for instance ref.). A more realistic modelling, however, requires considering off-lattice systems. Simple off-lattice heteroplymers have been discussed recently by Iori, Marinari, and Parisi, Irback et al., Klimov and Thirumalai, and by the present authors. The purpose to use such models is to understand the basic mechanism of folding to the native state. In lattice models, the native state is usually non-degenerate and it coincides with the ground state of the system. Delineating boundaries of the native basin in off-lattice systems, however, is difficult, especially when the number of degrees of freedom is large , yet it is essential for studies of almost all equilibrium and dynamical properties of proteins. For instance, stability of a protein is determined by estimating the equilibrium probability to stay in the native basin: the temperature at which this probability is $`\frac{1}{2}`$ defines the folding temperature, $`T_f`$. The native basin consists of the native state and its immediate neighborhood, as shown schematically in Figure 1, and it should not be confused with the whole folding funnel. The latter involves a much larger set of conformations which are linked kinetically to the native state.
In most studies, such as in Ref., the size of a basin is declared by adopting a reasonable but ad hoc cutoff bound. Systematic approaches, however, are needed and will be presented here. The task of delineating of the native basin is facilitated by introducing the concept of a distance between two conformations $`a`$ and $`b`$, $`\delta _{ab}`$. The distance should be defined in a way that excludes effects of an overall translation or rotation. There are two definitions of $`\delta _{ab}`$, for a sequence of $`N`$ monomers, that we shall use. The first one is:
$$\delta _{ab}^2=\text{min}\frac{1}{N}\underset{i=1}{\overset{N}{}}|\stackrel{}{r}_i^a\stackrel{}{r}_i^b|^2,$$
(1)
where $`\stackrel{}{r}_i^a`$ denotes the position of monomer $`i`$ in conformation $`a`$ and the minimization is performed over translations, rotations and reflections. In practice, we put chain $`a`$ over chain $`b`$ by overlapping the two centers of mass, and then we find the optimal rotation of $`b`$ which minimizes $`\delta _{ab}`$.
The second is
$$\delta _{ab}^2=\frac{1}{N^23N+2}\underset{ij,j\pm 1}{}(|\stackrel{}{r}_i^a\stackrel{}{r}_j^a||\stackrel{}{r}_i^b\stackrel{}{r}_j^b|)^2.$$
(2)
The first definition is closer to an intuitive understanding of the distance between shapes whereas the second is easier to compute, especially in three-dimensional situations. Nevertheless both are expected to be physically equivalent.
In this paper, we develop two techniques for determining the basin of the native state. In the first approach, we generate an image of the phase space by sampling it by low energy trajectories that start from the native state and continue by displacing the monomers in a way that increases the distance away from the native state while preserving the connectedness of the chain. For each trajectory, a dependence of the potential energy on the distance away from the native state is obtained and locations of the saddle points are determined. The average distance to a first encountered saddle point, $`<\delta _s>`$, may be considered as characterizing the size of the basin.
In the second technique, which we find to be more statistically reliable and more automatic in its implementation, we adopt a variant of de Gennes’s idea of the ”ant in a labirynth”. Specifically, we characterize the geometry of the native basin by monitoring random shape distortions of the heteropolymer in the basin. The distortions are induced by diffusive-like displacements of individual beads. The method of the shape distortion is implemented by first placing the polymer in an initial conformation, $`a(0)`$, which usually coincides with the native state. Subsequently, one performs random displacements of individual beads in the chain, through a Monte Carlo routine, and the conformation at time $`t`$ acquires a shape $`a(t)`$. The process is characterized by determining the evolution of a mean square distance, $`<\delta ^2>_t=<\delta _{a(0)a(t)}^2>`$ between $`a(t)`$ and $`a(0)`$. The focus is on short time behavior and the average is over many trajectories that start from the same $`a(0)`$. The characteristic size of the basin, $`\delta _c`$ is obtained by studying features in $`<\delta ^2>_t`$ as described later. In order to scale the walls of the basin one needs to make the polymer ’crawl’ up these walls without involving any kinetic energy. Thus the process does not correspond to any ’real’ evolution in time, as defined e.g. through the molecular dynamics. As opposed to the first technique, the fluctuations in the shape of the polymer are understood to be due to coupling to a heat bath.
Models
In order to illustrate the two techniques, we consider a two-dimensional version of the model introduced by Iori, Marinari and Parisi. The Hamiltonian is given by
$$H=\underset{ij}{}\{k(d_{i,j}d_0)^2\delta _{i,j+1}+4[\frac{C}{d_{i,j}^{12}}\frac{A_{ij}}{d_{i,j}^6}]\},$$
(3)
where $`i`$ and $`j`$ range from 1 to the number of beads, $`N`$, which in our model is equal to 16. The distance between the beads, $`d_{i,j}`$ is defined as $`|\stackrel{}{r}_i\stackrel{}{r}_j|`$, where $`\stackrel{}{r}_i`$ denotes the position of bead $`i`$, and is measured in units of the standard Lennard-Jones length parameter $`\sigma `$. The harmonic term in the Hamiltonian couples the adjacent beads along the chain. The remaining terms represent the Lennard-Jones potential. In $`A_{ij}`$ is chosen as $`A_{ij}=A_0+\sqrt{ϵ}\eta _{ij}`$, where $`A_0`$ is constant, $`\eta _{ij}`$’s are Gaussian variables with zero mean and unit variance; $`ϵ`$ controls the strength of the quenched disorder. The case of $`\eta _{ij}=0`$ and $`A_0=C`$ would correspond to a homopolymer. Our choice for the values of $`A_{ij}`$ is that all of the $`A_{ij}`$’s are positive, which corresponds to attraction. We measure the energy in units of $`C`$ and consider $`k`$ to be equal to 25 in units of $`C/\sigma ^2`$. Smaller values of $`k`$ may violate the self-avoidence of the chain.
We focus on two 16-monomer sequences in two dimensions which were characterized in detail previously. One of them, $`G`$, is a good folder and the other, $`R^{}`$, is a bad folder. We used two criteria for the quality of folding: 1) based on the evaluation of the specific heat and structural susceptibitity and 2) based on the location of the folding temperature, $`T_f`$, relative to the temperature at which the onset of the glassy effects takes place . $`G`$ has been designed as an off-lattice Go-like sequence and the $`A_{ij}`$’s are taken to be 1 for native contacts and 0 for non-native ones. The target conformation is on lattice and is shown in Figure 1a. This target is the same as for the lattice sequence $`R`$ of Ref.. The ground state of $`G`$ has approximately the shape of the target - note that the interaction between two beads forming a contact also slightly affects other beads. The bad folder $`R^{}`$ is constructed following the rank-ordering procedure which is an off-lattice analog of what was done in references. We choose the equlibrium interbead distance, $`d_0`$ of $`2^{1/6}`$ and $`1.16`$ (the latter value is determined by the average of $`A_{ij}`$ over all couplings, see Ref.) for $`G`$ and $`R^{}`$ respectively. The target conformation is chosen initially to be the same as in Figure 1a. The most strongly attractive Lennard-Jones interactions are assigned to the nine native contacts. They are enhanced by making the corresponding $`A_{ij}`$ bigger than one. The remaining couplings have $`A_{ij}`$’s which are smaller than 1. Rank ordering of the contacts generates good folders among lattice models. In the off-lattice models, however, the non-native Lennard-Jones interactions, due to their long range nature, overconstrain the system and frustrate it leading to a deterioration of folding properties. The values of $`A_{ij}`$ for $`R^{}`$ are listed in Ref.. The resulting native state of $`R^{}`$ is shown in Figure 1b.
Figures 2c and 2d show the distributions of distances $`\delta `$ for local energy minima of sequences $`G`$ and $`R^{}`$ respectively. The native states and the local energy minima have been obtained by multiple quenches from random conformations. Note that the sequence $`G`$ has a much fewer number of the local energy minima compared to sequence $`R^{}`$ and the energy gap to the first excited local minimum is significantly larger. In each case, there is one minimum which has the closest geometrical distance, $`\delta _{min}`$ to the native state. $`\delta _{min}`$ is thus an upper bound for the size of the native basin (0.5 for $`G`$ and 0.3 for $`R^{}`$).
Trajectories in the energy landscape
The first technique is implemented by creating the trajectories from the native state at $`T=0`$ in a stochastic way. The bead positions are displaced randomly and the moves are accepted if they increase the geometrical distance to the native state and keep the distance between nearest beads to be in the interval $`1<d_{i,i+1}<1.1d_0`$. In addition one has to keep the interaction energy for any pair of monomers in sequence $`R^{}`$ to be negative. For $`G`$, however, the non-native pairs interact only repulsively. In order to minimize the repulsion as much as possible we explore only those trajectories which keep the distances between the monomers of non-native contacts sufficiently large. We choose the minimal distance between the beads of non-native contacts, $`d_m`$, to be $`d_m=1.5`$ (the choice of 1.6 for $`d_m`$ yields similar results) and reject trajectories which lead to a monotonic increase in energy even after entering into the region of overall positive energies. Smaller values of $`d_m`$ usually lead to trajectories with unreasonably large positive energies. Substantially larger values of $`d_m`$ generate short trajectories which terminate on a conformation which does not allow for a further increase in the distance.
Figure 3 shows typical trajectories for $`G`$ and $`R^{}`$. For each trajectory, we define the postion of the saddle point $`\delta _s`$. This point shows as a local maximum on the energy vs. $`\delta `$ curve. The average value, $`<\delta _s>`$, defines the basin size. Averaging over 50 trajectories we obtain $`<\delta _s>=0.17\pm 0.02`$ and $`0.12\pm 0.01`$ for sequences $`G`$ and $`R^{}`$ respectively.
It should be noted that the technique described here is simple conceptually and easy to implement for a small number of trajectories but its resulting statistics on the basin size are inherently poor since there is no convenient control over the choice of important trajectories. Following patterns in the force field deterministically is a possible improvement but another stochastic approach, described below, combines simplicity with reliability of the results.
Fluctuations in the shape of the polymer
In order to compute $`<\delta ^2>_t`$ we update the monomer positions randomly within circles of radius of 0.01 (the choice of 0.02 yield similar results). We assume that the system is in contact with a heat bath corresponding to temperature $`T`$ which provides a controlling device. Figure 4 shows the dependence of $`<\delta ^2>_t`$ on $`t`$ for sequence $`G`$. We observe that there are, in general, three regimes of behavior of $`<\delta ^2>_t`$. For sequence G, all of the three regimes appear below $`T_c`$=0.19 and are shown, in Figure 4, for $`T`$=0.05 and 0.15.
In the first regime, I, corresponding to very short time scales during which merely several percents of a linear size of the basin are covered, one has a power law
$$<\delta ^2>_tt^{\nu _0}.$$
(4)
Pliszka and Marinari have demonstrated that $`\nu _0`$ is sensitive to the details of the Hamiltonian. We find that $`\nu _0`$ is equal to $`0.96\pm 0.02`$ and $`0.94\pm 0.02`$ for sequences G and R’ respectively. Both values are close to 1 and stay the same for all $`T`$’s (even for $`T`$ up to 50) which suggests a simple diffusive behavior.
In the second regime, II, the plot of $`<\delta ^2>_t`$ vs. $`t`$ acquires a $`T`$-dependence. Sommelius has focused on this regime and has postulated that the behavior here appears to follow a power law, at least approximately. The corresponding exponent $`\nu `$ is then found to depend on $`T`$ in a way that relates to characteristic temperatures of the system. We have found, however, that this power law is not robust – the effective exponent depends on the size of the steps in which one implements distortions. More importantly, the spatial extent of this regime is necessarily limited, and it would probably remain so even for very long heteropolymers.
In the third regime, III, observed only below a critical temperature $`T_c`$, $`<\delta ^2>_t`$ saturates at a constant value as explained in detail in Ref. . Above $`T_c`$, the shape distortion ceases to be confined to the native basin and the type II behavior continues to take place, as illustrated by the data points corresponding to $`T=0.45`$ in Figure 4. The limiting saturation value of $`\sqrt{<\delta ^2>}`$ at $`T_c`$ defines a characteristic basin size, $`\delta _c`$, that can be used, e.g., when deciding if a folding took place if the system started to evolve from and unfolded state. Naturally, $`T_c`$ is a measure of the folding temperature $`T_f`$ which characterizes thermodynamic stability of the system.
Figure 5 switches from the logarithmic scale of Figure 4 to the linear scale in wich the transition between the staturation and lack of saturation in $`<\delta ^2>_t`$ shows in a more convincing way. Figure 5 compares the time dependence of $`<\delta ^2>_t`$ to that of $`<\delta ^2>_t`$ for $`G`$ at $`T_c`$ and demonstrates that the actual choice of the definition of the distance has small effect on the results. For sequence $`G`$ we have $`\delta _c=0.2\pm 0.02`$ (the second definition of the distance yields $`\delta _c^{}=0.17\pm 0.02`$) For sequence $`R^{}`$ we find that $`T_c0.09`$ and $`\delta _c=0.09\pm 0.02`$. Within the error bars, $`\delta _c`$ is close to $`<\delta _s>`$ defined by the first approach which is also indicated in Figure 3. For both sequences, the values of $`\delta _c`$ and $`<\delta _s>`$ are smaller than $`\delta _{min}`$. Thus the saturation value of the distance to the native state at $`T=T_c`$ may indeed serve as a measure of size of the native basin $`\delta _c`$, i.e $`\delta _c=[<\delta ^2>_{sat}(T=T_c)]^{1/2}`$ determines the true boundary of the native basin.
The fact that $`\delta _c`$ for $`G`$ is found to be bigger than for $`R^{}`$ indicates bigger stability of $`G`$ relative to $`R^{}`$. This also correlates well with the higher value of $`T_c`$ which in turn suggests that $`T_c`$ is a measure of $`T_f`$ – the folding temperature. This interpretation is confirmed by comparing the values of $`T_c`$ to $`T_f`$ obtained by calculating the probability to be within the cutoff distance, $`\delta _c`$, away from the native state and by studying positions of the peaks in the structural susceptibility. These studies yield $`T_f`$ of 0.24 $`\pm `$ 0.03 and $``$0.12 for $`G`$ and $`R^{}`$ respectively. Both of these values are close to $`T_c`$ obtained from the shape distortion. It should be noted that the calculation of $`T_c`$ by monitoring stochastic shape distortions is significantly less involving computationally. We have also used this technique to determine the sizes of basins of low lying local energy minima. The corresponding values of $`\delta _c`$ are found to be smaller than for the native state.
We now consider, following Struglia , the basin of attraction for random initial conformations which are subsequently quenched in the steepest descent fashion. The basin of attraction is defined in terms of a distance at which the probability to fall onto the native state is bigger than or equal to $`p_c`$. The corresponding basin size will be denoted by $`\delta _f`$. In Ref. , $`p_c`$ was taken to be equal to $`\frac{1}{2}`$. In our studies, we took $`p_c=1`$ and determined the basin of attraction for sequences $`G`$ and $`R^{}`$ through a standard quenching procedure. We considered 200 trajectories and obtained $`\delta _f0.55`$ and $`0.35`$ for sequences $`G`$ and $`R^{}`$ respectively. These values exceed not only our estimate of $`\delta _c`$ but also that of the minimal distance between the native state and the nearest minimum $`\delta _{min}`$. The values of $`\delta _f`$ would become even larger for a $`p_c`$ that was less than 1. This emphasizes the point that the procedure used by Struglia probably delineates the folding funnel and not the native basin itself.
In summary, we have explored the native basins of two off-lattice sequences by monitoring the shape distortion and by exploring the saddle points of the trajectories from the native state. We have come out with computationally simple methods to delineate the boundaries of the basins and to estimate the folding temperature. The bad and good folders are found to have native basins which are comparable in size even though the structure of their folding funnels must be very different.
We thank T. X. Hoang for discussions. This work was supported by Komitet Badan Naukowych (Poland; Grant number 2P03B-025-13).
|
no-problem/9908/cond-mat9908466.html
|
ar5iv
|
text
|
# Magnetomechanics of mesoscopic wires
## I Introduction
The electrical conductance in a ballistic wire with dimensions comparable to the Fermi wavelength increases in steps of $`G_0=2e^2/h`$ as the cross section increases. This conductance quantization is observable at room temperature in metallic nanowires formed by pressing two pieces of metal together, into a metallic contact. When the two pieces are separated the contact is stretched into a nanowire, a wire of nanometer dimensions. Several experiments varying this principle have been performed, e.g. using scanning tunneling microscopy, mechanically controlled break junctions or just plain macroscopic wires . Most nanowire experiments have been performed on metals, however conductance quantization have been seen in Bismuth at 4K. Since Bismuth has a Fermi wavelength $`\lambda _F=26`$ nm , these semi-metal “nanowires” are larger than the metallic nanowires.
The stepwise variation of conductance in such a mesoscopic wire is accompanied by an abrupt change of the force in the wire . Using a free electron model, neglecting all atomic structure of the wire, it has been shown that the size of the electronic contribution to the force fluctuations are comparable to the experimentally found values and that the qualitative behavior, i.e. the abrupt change that accompanies the conductance steps, is the same.
In the wire the transverse motion of the electrons give rise to quantized modes $`\alpha `$ of energy $`E_\alpha `$. In the simplest version of the Landauer formalism, a mode is considered fully transmitting, open, if $`E_F>E_\alpha `$ and closed otherwise. Each open mode contributes an amount $`e^2/h`$ to the conductance, if modes with different spin are considered separately. When the wire is elongated, the cross section decreases, more and more modes are pushed above the Fermi level and closed, thus decreasing the conductance stepwise. This has been shown in two dimensions and in three dimensions.
It has been suggested that the conductance and the mechanical force in a nanowire can be controlled by an applied driving voltage. This effect originates from the injection of additional electrons with voltage dependent energy, because of the different chemical potentials of the two reservoirs. A relatively large applied voltage is needed so one will have to worry about heating in this case.
The eigenenergies of the transverse motion can be affected by an external magnetic field, $`B`$, perpendicular to the cross section of the wire. This will show in the conductance and in the force as a function of $`B`$. The effect of a magnetic field on the conductance has been considered in ref. . To use an external magnetic field is an equilibrium method to control the number of transporting channels, without significant risk of relaxation.
Because of band bending, due to the small size of the wire, the eigenenergies will have to be corrected. This can, however, be taken care of by introducing an effective Fermi energy in the wire, $`\stackrel{~}{E}_F`$. Assuming that the number of electrons (per unit volume) is constant, $`\stackrel{~}{E}_F`$ can be determined selfconsistently and will vary with wire length and magnetic field.
In this paper we present force calculations for different applied magnetic fields and wire lengths, using a free electron model. We take into account the effect of band bending, adjusting the Fermi energy in the wire. In order to resolve any effect for moderate magnetic fields, a low cyclotron effective mass (which enters in the cyclotron frequency) is needed, which can be found in semi-metals. Metals are less favorable since because of a larger cyclotron effective mass (larger Fermi energy) we would need a larger magnetic field in order to resolve any effect. For numerical estimates we have used values for Bismuth, a typical semi-metal. For Bismuth also the spin splitting is important since it has a large spectroscopic spin splitting factor $`g`$.
## II Model
We consider a cylindrical ballistic wire of length $`L`$ with circular cross section and a parabolic confining potential,
$$\omega (r)=\frac{\omega _0^2m^{}r^2}{2}E_F\frac{r^2}{R^2},$$
(1)
using cylindrical coordinates $`(r,\varphi ,z)`$ and where $`m^{}`$ is the effective electron mass. The wire is along the $`z`$-direction. The last equality in Eq. 1 defines $`\omega _0`$. In this equation $`E_F`$ is the zero $`B`$-field bulk value, yielding a magnetic field independent confining potential. We assume that the volume $`V=\pi R^2L`$ of the wire is kept constant during elongation, which makes $`R`$ and $`L`$ mutually dependent.
With the above confining potential and an applied magnetic field along the wire the Schrödinger equation has been solved . If also spin is included the eigenenergies are
$`E_\alpha `$ $`=`$ $`\mathrm{}\left({\displaystyle \frac{\omega _c^2}{4}}+\omega _0^2\right)^{1/2}n+{\displaystyle \frac{1}{2}}l\mathrm{}\omega _c+sg\mu _BB`$ (2)
$`n`$ $`=`$ $`2m+|l|+1`$ (3)
$`m`$ $`=`$ $`0,1,2,\mathrm{}`$ (4)
$`l`$ $`=`$ $`0,\pm 1,\pm 2,\mathrm{}`$ (5)
$`s`$ $`=`$ $`\pm 1/2`$ (6)
$`\alpha `$ $`=`$ $`\{m,l,s\},`$ (7)
where $`\omega _c=eB/m^{}`$ is the cyclotron frequency, $`\mu _B`$ is the Bohr magneton and $`sg\mu _B`$ is the magnetic moment associated with the electronic spin.
Since our system is open the electronic contribution to the force in the wire is given by the derivative of the grand potential $`\mathrm{\Omega }=E\mu N`$ with respect to elongation. Here $`E`$ is the total energy of the electrons in the wire, $`\mu `$ the chemical potential and $`N`$ the number of electrons in the wire. If the Fermi energy $`E_F`$ is much higher than the thermal energy (as in metals or at low temperature) we have $`\mu E_F`$. The grand potential is then
$$\mathrm{\Omega }(E_F)=\underset{\alpha }{}\frac{4}{3}L\sqrt{\frac{2m^{}}{\pi ^2\mathrm{}^2}}(E_FE_\alpha )^{3/2},$$
(8)
where the sum is over all open modes. The force in the wire is given by
$$F=\frac{\delta \mathrm{\Omega }}{\delta L},$$
(9)
which in general has to be calculated numerically.
The magnetic field affects the system primarily by splitting the otherwise degenerate eigenenergies of the conduction modes, Eq. 2. Since then the conduction modes will open one by one this will cause more structure in the force and conductance when displayed as a function of wire length. Subsequently, when applying an external magnetic field we will see the (clearest) effect when the highest open level or the lowest closed level goes through the Fermi level (whichever happens first). If we do not adjust the Fermi energy for band bending, but use the bulk Fermi energy for zero magnetic field, one can analytically calculate the $`B`$-field needed, when keeping the wire at a specific length. The least favorable situation would be on the middle of a conduction step.
## III Results and discussion
We have used numerical values for Bismuth, a typical semi-metal with $`E_F=25`$ meV. Bismuth has an anisotropic Fermi surface resulting in different effective masses in different directions, between $`0.009m_e1.8m_e`$. The cyclotron effective mass is in the range $`0.009m_e0.13m_e`$. Assuming an isotropic Fermi surface and an quadratic dispersion relation, both effective masses are the same, for $`E_F=25`$ meV $`m^{}=0.07m_e`$. The spectroscopic splitting factor, $`g`$, can be as high as 260, or one order of magnitude smaller depending on the direction of the magnetic field. With $`g=20`$ the spin splitting is roughly of the same order as the Landau level distance, and becomes dominant for $`g`$ as large as 200. We have used $`g=20`$. The wire volume was kept constant at $`30000`$ nm<sup>3</sup>.
To find the effective Fermi energy of the wire we have adjusted the value in order to keep the number of electrons constant, with a tolerance of $`10^4\%`$.
Figure 1 shows the force in the wire as a function of wire length for different magnetic fields. For non-zero fields the force curves show more structure since now the eigenenergies of the conduction channels are non-degenerate and close one by one, each time resulting in a sharp change of the force.
The force and conductance for two particular magnetic fields, $`B=0`$ and $`B=2.5`$ T, are shown in Fig. 2. Each step in the conductance is accompanied by an abrupt change in the force. We also show the corresponding picture for the simplest possible case, when we use the bulk value of the Fermi energy, $`E_F`$, in Fig. 3. In this case the force is one order of magnitude smaller then in the more realistic case with $`\stackrel{~}{E}_F`$. This is because the effective Fermi energy has to be larger then the bulk value in order to keep the number of electrons per unit volume in the wire constant in spite of the quantization of levels. Also the conduction modes close much later in the $`\stackrel{~}{E}_F`$-case than in the more simple case when the wire is elongated. The reason for this is that the effective Fermi energy, as a function of wire length, follows the eigenenergies before intercepting it and closing the channel.
On the middle of the second conduction step ($`G=3G_0`$, $`n=2`$) the circumstances are least favorable to see the effect of the magnetic field. For the case with the zero $`B`$-field bulk value of the Fermi energy ($`L=19.8`$nm) we have analytically calculated that one needs $`B=2.4`$ T, to see the highest open level go through the Fermi energy, thus giving a sharp change in the force as well as in the conductance. For higher conduction modes one will see the effect for smaller fields, since the splitting is proportional to $`l`$, whos absolute maximum is equal to $`n`$.
In Fig. 4 we see the force and the conductance as a function of magnetic field for a fixed wire length, $`L=54.6`$ nm. This is for the case with an effective wire Fermi energy and the length corresponds to the middle of the second conduction step ($`G=3G_0`$, $`n=2`$). We see that we need about 1.3 T before the highest open level goes through the Fermi surface showing us the articulate effect of the magnetic field. In the lower part of the same figure we also see the effective Fermi energy (thick line) and the eigenenergies of the second conduction steps. Notice how the Fermi level increases with the eigenenergy before it intercepts. These variations are however small compared to the overall magnitude of the Fermi energy.
So far we have used the spectroscopic splitting factor $`g=20`$. In Fig. 5 we show the force as a function of length for $`B=1`$ T for different $`g`$-factors: $`g=0,2,20`$ and 200. For $`g=0`$ there is no spin splitting, but we still see more structure than for $`B=0`$ (cf. Fig. 1). This is due to the breaking of the degeneracy into the Landau levels. With increasing $`g`$-factor the spin splitting becomes larger and larger, however whatever the size of the spin-splitting is: more structure in the force appears with an applied magnetic field.
Also the Fermi energy of the bulk will be affected by the magnetic field, due to the de Haas-van Alphen effect. In the case when an effective Fermi energy, $`\stackrel{~}{E_F}`$, is used this does not affect the results since the bulk Fermi energy does not enter into the calculations. When adjusting the bulk Fermi energy for de Haas-van Alphen effect, in the more simple case shown in Fig. 3, there is no significant change on the force. We have also studied the influence of a moderate applied voltage (in the mV-range) but have seen no significant effect.
For metals the Fermi energy is in the eV-range demanding a much higher magnetic fields to resolve results similar to those for Bismuth above. Since the size of the splitting is proportional to the number of open channels, having more channels will decrease the magnetic field needed. So if we design the circumstances to be more favorable, i.e. more open channels and close to a conduction step a moderate magnetic field will be enough to make an eigenenergy go through the Fermi level, thus giving an effect in the force and in the conductance.
## IV Conclusion
Using a free electron model we have shown that the force in a mesoscopic wire can be affected by an external magnetic field parallel to the wire. With a magnetic field present the degenerate eigenenergies of the conduction modes split and become conducting, open, at different elongations resulting in more force fluctuations with increasing wire length. At fixed wire length we propose that an external magnetic field is an equilibrium method that can be used to affect the force as well as the conductance in mesoscopic wires.
I wish to thank Robert Shekhter for valuable discussions. Financial support from the Swedish NFR is gratefully acknowledged.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.