cond-mat0602198/cm.tex
1: %\documentclass[prb,preprint,showpacs]{revtex4}
2: 
3: \documentclass[prb,tighten]{revtex4}
4: 
5: %\documentclass[pre,twocolumn,showpacs]{revtex4}
6: 
7: %\documentclass[pre,preprint,draft,showpacs,tighten]{revtex4}
8: 
9: %\usepackage[spanish]{babel}
10: 
11: %\usepackage[cp850]{inputenc}
12: 
13: %\usepackage[latin1]{inputenc}
14: 
15: %\usepackage{showkeys}% Muestra etiquetas de formulas, figuras, referencias,...
16: 
17: % Include figure files
18: 
19: %\usepackage{dcolumn}% Align table columns on decimal point
20: 
21: %\usepackage{bm}% bold math
22: 
23: %\psdraft
24: 
25: 
26: 
27: 
28: 
29: 
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31: 
32: \usepackage{epsfig}
33: 
34: \usepackage{graphicx}
35: 
36: 
37: 
38: %TCIDATA{OutputFilter=Latex.dll}
39: 
40: %TCIDATA{LastRevised=Friday, December 23, 2005 14:49:30}
41: 
42: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
43: 
44: 
45: 
46: 
47: 
48: \begin{document}
49: 
50: 
51: 
52: \title{Mass and heat fluxes for a binary granular mixture at low-density}
53: \author{Vicente Garz\'{o}}
54: \email[E-mail: ]{vicenteg@unex.es}
55: \address{Departamento de F\'{\i}sica, Universidad de Extremadura, E-06071
56: Badajoz, Spain}
57: \author{Jos\'e Mar\'{\i}a Montanero}
58: \email[E-mail: ]{jmm@unex.es}
59: \address{Departamento de Electr\'onica e Ingenier\'{\i}a Electromec\'anica,
60: Universidad de Extremadura, E-06071 Badajoz, Spain}
61: \author{James W. Dufty}
62: \email[E-mail: ]{dufty@phys.ufl.edu}
63: \address{Department of Physics, University of Florida, Gainesville, Florida
64: 32611}
65: 
66: 
67: 
68: \begin{abstract}
69: 
70: The Navier--Stokes order hydrodynamic equations for a low density
71: granular mixture obtained previously from the Chapman--Enskog
72: solution to the Boltzmann equation are considered further. The six
73: transport coefficients associated with mass and heat flux in a
74: binary mixture are given as functions of the mass ratio, size
75: ratio, composition, and coefficients of restitution. Their
76: quantitative variation across this parameter set is demonstrated
77: using low order Sonine polynomial approximations to solve the
78: exact integral equations. The results are also used to quantify
79: the violation of the Onsager reciprocal relations for a granular
80: mixture. Finally, the stability of the homogeneous cooling state
81: is discussed.
82: 
83: \end{abstract}
84: 
85: 
86: 
87: \pacs{ 05.20.Dd, 45.70.Mg, 51.10.+y, 47.50.+d}
88: \date{\today}
89: \maketitle
90: 
91: \draft
92: 
93: 
94: 
95: \section{Introduction}
96: \label{sec1}
97: 
98: The relevance and context of a hydrodynamic description for
99: granular gases remains a controversial topic. At sufficiently low
100: density the origin of hydrodynamics can be studied from Boltzmann
101: kinetic theory. \cite{GS95,BP04} To obtain hydrodynamics from the
102: Boltzmann kinetic equation the essential assumption is the
103: existence of a ``normal'' solution, defined to be one for which
104: all space and time dependence occurs through the macroscopic
105: hydrodynamic fields. \cite{CC70} This solution, together with the
106: macroscopic balance equations, leads to a closed set of
107: hydrodynamic equations for these fields. The Chapman--Enskog
108: method provides a constructive means to obtain an approximation to
109: such a solution for states whose spatial gradients are not too
110: large. In this general context, the study of hydrodynamics for
111: granular gases is the same as that for normal gases.
112: 
113: 
114: 
115: The details of the Chapman--Enskog method have been carried out
116: for a one granular component gas to obtain the Navier--Stokes
117: order hydrodynamic equations, together with exact integral
118: equations determining the transport coefficients occurring in
119: these equations. \cite{BDKS98} These integral equations have been
120: solved approximately using Sonine polynomial expansions and
121: compared to both Direct Monte Carlo simulation of the Boltzmann
122: equation and molecular dynamics simulation of the gas.
123: \cite{BRMC99} Good agreement is obtained even for relatively
124: strong degrees of dissipation. The results support the formal
125: theoretical analysis and the claim that hydrodynamics is not
126: limited to the quasi-elastic limit.
127: 
128: 
129: The analysis for multicomponent granular gases is much more
130: complicated than for a one component gas. Most of the previous
131: attempts \cite{JM89} were made for {\em nearly} elastic spheres
132: where the equipartition of energy can be considered as an
133: acceptable assumption. In addition, according to this level of
134: approximation, the inelasticity is only accounted for by the
135: presence of a sink term in the energy balance equation, so that
136: the expressions for the transport coefficients are the same as
137: those obtained for normal fluids. \cite{MCK83} However, the
138: failure of energy equipartition in multicomponent granular gases
139: \cite{GD99} has been also confirmed by computer simulations
140: \cite{computer} and even observed in real experiments. \cite{exp}
141: Although the possibility of nonequipartition was already pointed
142: out many years ago, \cite{JM87} it has not been until recently
143: that a systematic study of the effect of nonequipartition on
144: transport has been carefuly analyzed. In this context, Garz\'o and
145: Dufty \cite{GD02} have carried out a derivation of the
146: Navier--Stokes hydrodynamic equations for a binary mixture at
147: low-density that accounts for nonequipartition of granular energy.
148: These equations and associated transport coefficients provide a
149: somewhat more stringent test of the analysis since the parameter
150: space is much larger. There are now many more transport
151: coefficients, given as functions of the three independent
152: coefficients of restitution, size ratio, mass ratio, and
153: composition. As in the one component case, explicit expressions
154: for the transport coefficients requires also to consider Sonine
155: polynomial expansions. The accuracy of this approach has been
156: confirmed by comparison with Monte Carlo simulations of the
157: Boltzmann equation in the cases of the shear viscosity \cite{MG03}
158: and the tracer diffusion \cite{GM04} coefficients. Exceptions to
159: this agreement are extreme mass or size ratios and strong
160: dissipation, although these discrepancies between theory and
161: simulation diminish as one considers more terms in the Sonine
162: polynomial approximation. \cite{GM04}
163: 
164: 
165: 
166: 
167: Since the dependence of the shear viscosity coefficient on the
168: parameters of the mixture (masses, sizes, concentration,
169: coefficients of restitution) has been widely studied in a previous
170: work, \cite{MG03}  a primary objective here is to demonstrate the
171: variation of the six transport coefficients associated with the
172: mass and heat flux in this parameter space, using the same Sonine
173: polynomial approximation as was found applicable for the one
174: component gas. To set the context for these quantitative results
175: two qualitative and potentially confusing issues are briefly noted
176: at the outset.
177: 
178: 
179: 
180: There is some ambiguity regarding the hydrodynamic temperature in
181: a mixture since temperatures for each species can be defined in
182: addition to the global temperature. What should be the
183: hydrodynamic fields? For normal gases the answer is clear. These
184: are set by the slow variables (at large space and time scales)
185: associated with conserved quantities. In addition to species
186: number and momentum, only the total kinetic energy is conserved,
187: so only one global temperature occurs as a hydrodynamic field. The
188: energy is no longer conserved for granular gases, but it remains a
189: slow variable if the cooling rate is not too large. Thus, in this
190: case as well, only the global temperature should appear among the
191: hydrodynamic fields.
192: 
193: 
194: Nevertheless, the species temperatures play a new and interesting
195: secondary role. For a normal gas, there is a rapid velocity
196: relaxation in each fluid cell to a local equilibrium state on the
197: time scale of a few collisions. Subsequently, the equilibration
198: among cells occurs via the hydrodynamic equations. In each cell
199: the species velocity distributions are characterized by the
200: species temperatures. These are approximately the same due to
201: equipartition, and the hydrodynamic relaxation occurs for the
202: single common temperature. \cite{CC70} A similar rapid velocity
203: relaxation occurs for granular gases in each small cell, but to a
204: universal state different from local equilibrium and one for which
205: equipartition no longer occurs. Hence, the species temperatures
206: $T_{i}$ are different from each other and from the overall
207: temperature $T$ of the cell. \cite{GD99} Nevertheless, the time
208: dependence of all temperatures is the same in this and subsequent
209: states, $T_{i}(t)=\gamma _{i}T(t)$. This implies that the species
210: temperatures do not provide any new dynamical degree of freedom.
211: They still characterize the shape of the partial velocity
212: distributions and affect the quantitative averages calculated with
213: these distributions. The transport coefficients for granular
214: mixtures therefore have new quantitative effects arising from the
215: time independent temperature ratio $\gamma _{i}$ for each species.
216: \cite{GD02} This dependence is illustrated here as well.
217: 
218: 
219: In some earlier works, \cite{JM87,GDH05} additional equations for
220: each species temperature have been included among the hydrodynamic
221: set. This is possible since the overall temperature is determined
222: from these by $T(t)=\sum_{i}x_{i}T_{i}(t)$, where $ x_{i}$ is the
223: concentration of species $i$. However, this is an unnecessary
224: complication, describing additional kinetics beyond hydrodynamics
225: that is relevant only on the time scale of a few collisions. As
226: described above (and supported by molecular dynamics simulations
227: \cite{DHGD02}) the dynamics of all temperatures quickly reduces to
228: that of $T(t)$. The remaining time independent determination of
229: the $\gamma _{i}$ follows directly from the condition that the
230: cooling rates of all temperatures be the same.
231: 
232: 
233: 
234: A second confusing issue is the context of the Navier--Stokes
235: equations considered here. The derivation of the Navier--Stokes
236: order transport coefficients does not limit their application to
237: weak inelasticity. For this reason the results reported below
238: include a domain of both weak and strong inelasticity, $0.5\leq
239: \alpha \leq 1$. The Navier--Stokes hydrodynamic equations
240: themselves may or may not be limited with respect to inelasticity,
241: depending on the particular states considered. The derivation of
242: these equations by the Chapman--Enskog method assumes that
243: relative changes in the hydrodynamic fields over distances of the
244: order of a mean free path are small. For normal gases this can be
245: controlled by the initial or boundary conditions. It is more
246: complicated for granular gases. In some cases (e.g., steady states
247: such as the simple shear flow problem \cite{SGD04}) the boundary
248: conditions imply a relationship between the coefficient of
249: restitution and some hydrodynamic gradient -- the two cannot be
250: chosen independently. Consequently, there are examples for which
251: the Navier--Stokes approximation is never valid or is restricted
252: to the quasi-elastic limit. \cite{SGD04} However, the transport
253: coefficients characterizing the Navier--Stokes order hydrodynamic
254: equations are well-defined functions of $\alpha$, regardless of
255: the applicability of those equations (e.g., the coefficients of a
256: Taylor series for some function $f(\alpha,x)$ in powers of $x$ may
257: be defined for all $\alpha$ at each order, even when the series
258: cannot be truncated at that order).
259: 
260: 
261: 
262: The plan of the paper is as follows. First, in Sec.\ \ref{sec2}
263: the hydrodynamic equations and associated fluxes to Navier--Stokes
264: order are recalled, and the expressions for the transport
265: coefficients for heat and mass transport are given to leading
266: Sonine polynomial approximation. The elastic limit is discussed to
267: aid in the interpretation of these expressions. Next, in Sec.\
268: \ref{sec3} the results for these six coefficients are illustrated
269: for a common coefficient of restitution  and same size ratio as
270: functions of $\alpha $ at a composition $x_{1}=0.2$ for several
271: values of the mass ratio. With the exception of thermal
272: conductivity, the deviations from normal gas values are largest at
273: small $\alpha $ and large mass ratio. The usual Onsager relations
274: among these coefficients for normal gases is then noted and tested
275: for the granular gas in Sec.\ \ref{sec4}. Since the underlying
276: basis for these relations (time reversal symmetry) no longer holds
277: for granular systems, the expected violation is demonstrated as a
278: function of $\alpha $ for the same conditions. The stability of a
279: special homogeneous solution to the mixture hydrodynamic equations
280: is then studied, and some comments on the implications of the
281: instability found are offered. Finally, the results are summarized
282: and discussed in the last section.
283: 
284: 
285: 
286: \section{Boltzmann kinetic theory for the mass and heat fluxes}
287: \label{sec2}
288: 
289: 
290: 
291: We consider a binary mixture of \emph{inelastic}, smooth, hard
292: spheres of masses $m_{1}$ and $m_{2}$, and diameters $\sigma _{1}$
293: and $\sigma _{2}$. The inelasticity of collisions among all pairs
294: is characterized by three independent constant coefficients of
295: normal restitution $\alpha _{11}$, $ \alpha _{22}$, and $\alpha
296: _{12}=\alpha _{21}$, where $\alpha _{ij}$ is the coefficient of
297: restitution for collisions between particles of species $i$ and
298: $j$. \ It is assumed that the density of each species is
299: sufficiently low that their velocity distribution functions are
300: accurately described by the coupled set of {\em inelastic}
301: Boltzmann kinetic equations. The precise form of these equations
302: is given in Ref.\ \onlinecite{GD02} and will not be required here.
303: These equations imply the exact macroscopic balance equations for
304: the particle number density of each species, $n_{i}\left(
305: \mathbf{r},t\right) $, flow velocity $\mathbf{u}\left(
306: \mathbf{r},t\right) $, and temperature, $T\left(
307: \mathbf{r},t\right) $, \cite{GD02}
308: \begin{equation}
309: D_{t}n_{i}+n_{i}\nabla \cdot \mathbf{u}+\frac{\nabla \cdot
310: \mathbf{j}_{i}}{ m_{i}}=0\;,\hspace{0.3in}i=1,2  \label{2.01}
311: \end{equation}
312: \begin{equation}
313: D_{t}\mathbf{u}+\rho ^{-1}\nabla \mathsf{P}=0\;,  \label{2.02}
314: \end{equation}
315: \begin{equation}
316: D_{t}T-\frac{T}{n}\sum_{i}\frac{\nabla \cdot
317: \mathbf{j}_{i}}{m_{i}}+\frac{2}{ 3n}\left( \nabla \cdot
318: \mathbf{q}+\mathsf{P}:\nabla \mathbf{u}\right) =-\zeta T\;.
319: \label{2.03}
320: \end{equation}
321: In the above equations, $D_{t}=\partial _{t}+\mathbf{u}\cdot
322: \nabla $ is the material derivative, $\rho =m_{1}n_{1}+m_{2}n_{2}$
323: is the total mass density, $\mathbf{j}_{i}$ is the particle number
324: flux for species $i$, $\mathbf{q}$ is the heat flux, $\mathsf{P}$
325: is the pressure tensor, and $\zeta$ is the cooling rate. For the
326: two component mixture considered here there are six independent
327: fields, $n_{1},$ $n_{2},$ $T,$ $\mathbf{u}$. To obtain a closed
328: set of hydrodynamic equations, expressions for ${\bf j}_{i}$,
329: $\mathbf{q}$, $ \mathsf{P}$, and $\zeta $ must be given in terms
330: of these fields. Such expressions are called ``constitutive
331: equations''. It is convenient to give these constitutive equations
332: in terms of a different set of experimentally more accessible
333: fields, $x_{1},$ $p,$ $T,$ $\mathbf{u}$, where $
334: x_{1}=n_{1}/\left( n_{1}+n_{2}\right) $ is the composition of
335: species $1$, and $p=\left( n_{1}+n_{2}\right) T$ is the
336: hydrostatic pressure. This is simply a change of variables, so
337: that Eqs.\ (\ref{2.01})--(\ref{2.03}) become
338: \begin{equation}
339: D_{t}x_{1}+\frac{\rho }{n^{2}m_{1}m_{2}}\nabla \cdot
340: \mathbf{j}_{1}=0\;, \label{2.04}
341: \end{equation}
342: \begin{equation}
343: D_{t}p+p\nabla \cdot \mathbf{u}+\frac{2}{3}\left( \nabla \cdot
344: \mathbf{q}+ \mathsf{P}:\nabla \mathbf{u}\right) =-\zeta p,
345: \label{2.05}
346: \end{equation}
347: \begin{equation}
348: D_{t}\mathbf{u}+\rho ^{-1}\nabla \mathsf{P}=0\;,  \label{2.06}
349: \end{equation}
350: \begin{equation}
351: D_{t}T-\frac{T}{n}\sum_{i}\frac{\nabla \cdot
352: \mathbf{j}_{i}}{m_{i}}+\frac{2}{3n}\left( \nabla \cdot
353: \mathbf{q}+\mathsf{P}:\nabla \mathbf{u}\right) =-\zeta T\;.
354: \label{2.07}
355: \end{equation}
356: 
357: 
358: 
359: The constitutive equations up to the Navier--Stokes order have
360: been obtained from the Boltzmann equation in Ref.\
361: \onlinecite{GD02} with the results
362: \begin{equation}
363: \mathbf{j}_{1}=-\left( \frac{m_{1}m_{2}n}{\rho }\right) D\nabla
364: x_{1}-\frac{ \rho }{p}D_{p}\nabla p-\frac{\rho }{T}D^{\prime
365: }\nabla T,\quad \mathbf{j}_{2}=-\mathbf{j}_{1},  \label{2.1}
366: \end{equation}
367: \begin{equation}
368: \mathbf{q}=-T^{2}D^{\prime \prime }\nabla x_{1}-L\nabla p-\lambda
369: \nabla T, \label{2.2}
370: \end{equation}
371: \begin{equation}
372: P_{k\ell }=p\delta _{k\ell }-\eta \left( \nabla _{\ell
373: }u_{k}+\nabla _{k}u_{\ell }-\frac{2}{3}\delta _{k\ell }\nabla
374: \cdot \mathbf{u}\right) , \label{2.3}
375: \end{equation}
376: \begin{equation}
377: \label{2.2.1} \zeta =\zeta _{0}+\mathcal{O}(\nabla^2)
378: \end{equation}
379: The transport coefficients $\{D,D_{p},D^{\prime},D^{\prime \prime
380: },L,\lambda ,\eta \}$ verify a set of coupled linear integral
381: equations which can be solved approximately by using the leading
382: terms in a Sonine polynomial expansion. This solution provides
383: explicit expressions for the transport coefficients in terms of
384: the coefficients of restitution and the parameters of the mixture
385: (masses, sizes, and composition). The above expressions for mass
386: and heat fluxes can be defined in a variety of equivalent ways
387: depending on the choice of driving forces used. For systems with
388: elastic collisions, the specific set of gradients contributing to
389: each flux is restricted by fluid symmetry, Onsager's relations
390: (time reversal invariance), and the form of entropy production.
391: \cite{GM84} In this case, one usual representation leads to the
392: mass and heat fluxes proportional to $(\nabla \mu_{i})_T$ and
393: $\nabla T$, where $\mu_i$ (defined below) is the chemical
394: potential per unit mass. However, for \emph{inelastic} systems
395: only fluid symmetry holds and so there is more flexibility in
396: representing the fluxes and identifying the corresponding
397: transport coefficients. In particular, a third contribution
398: proportional to $\nabla p$ appears in both fluxes. Some care is
399: required in comparing transport coefficients in different
400: representations using different independent gradients for the
401: driving forces.
402: 
403: 
404: 
405: The cooling rate to lowest order in the gradients is $\zeta =\zeta
406: _{0}(x_{1},p,T)$. There are no contributions to first order in the
407: gradient for the low density Boltzmann equation. The general form
408: including second order gradient contributions is displayed in
409: Appendix A. These second order terms have been calculated for a
410: one component gas \cite{BDKS98} and found to be very small. Here,
411: these second order contributions to the cooling rate will be
412: neglected.
413: 
414: 
415: 
416: Substitution of the Navier--Stokes constitutive equations, (\ref
417: {2.1})-(\ref{2.3}), into the exact balance equations,
418: (\ref{2.04})-(\ref{2.07} ), gives the Navier--Stokes hydrodynamic
419: equations for a binary mixture
420: \begin{equation}
421: D_{t}x_{1}=\frac{\rho }{n^{2}m_{1}m_{2}}\nabla \cdot \left(
422: \frac{m_{1}m_{2}n }{\rho }D\nabla x_{1}+\frac{\rho }{p}D_{p}\nabla
423: p+\frac{\rho }{T}D^{\prime }\nabla T\right) \;,  \label{2.4}
424: \end{equation}
425: \begin{eqnarray}
426: \left( D_{t}+\zeta \right) p+\frac{5}{3}p\nabla \cdot \mathbf{u}
427: &=&\frac{2}{ 3}\nabla \cdot \left( T^{2}D^{\prime \prime }\nabla
428: x_{1}+L\nabla
429: p+\lambda\nabla T\right)   \nonumber \\
430: &&+\frac{2}{3}\eta \left( \nabla _{\ell }u_{k}+\nabla _{k}u_{\ell
431: }-\frac{2}{ 3}\delta _{k\ell }\nabla \cdot \mathbf{u}\right)
432: \nabla _{\ell }u_{k}, \label{2.5}
433: \end{eqnarray}
434: \begin{eqnarray}
435: \left( D_{t}+\zeta \right) T+\frac{2}{3}p\nabla \cdot \mathbf{u}
436: &=&-\frac{T }{n}\frac{m_{2}-m_{1}}{m_{1}m_{2}}\nabla \cdot \left(
437: \frac{m_{1}m_{2}n}{ \rho }D\nabla x_{1}+\frac{\rho }{p}D_{p}\nabla
438: p+\frac{\rho
439: }{T}D^{\prime}\nabla T\right)   \nonumber \\
440: &&+\frac{2}{3n}\nabla \cdot \left( T^{2}D^{\prime \prime }\nabla
441: x_{1}+L\nabla p+\lambda \nabla T\right)   \nonumber \\
442: &&+\frac{2}{3n}\eta \left( \nabla _{\ell }u_{k}+\nabla _{k}u_{\ell
443: }-\frac{2 }{3}\delta _{k\ell }\nabla \cdot \mathbf{u}\right)
444: \nabla _{\ell }u_{k}, \label{2.6}
445: \end{eqnarray}
446: \begin{equation}
447: D_{t}u_{\ell }+\rho ^{-1}\nabla _{\ell }p=\rho ^{-1}\nabla
448: _{k}\eta \left( \nabla _{\ell }u_{k}+\nabla _{k}u_{\ell
449: }-\frac{2}{3}\delta _{k\ell }\nabla \cdot \mathbf{u}\right) \;.
450: \label{2.7}
451: \end{equation}
452: For the chosen set of fields $n=p/T$ and $\rho =p\left[ \left(
453: m_{1}-m_{2}\right) x_{1}+m_{2}\right] /T$. These equations are
454: exact to second order in the spatial gradients for a low density
455: Boltzmann gas.
456: 
457: 
458: 
459: \subsection{Mass flux}
460: \label{subsec2.1}
461: 
462: 
463: The mass flux contains three transport coefficients, the diffusion
464: coefficient $D$, the pressure diffusion coefficient $D_{p}$, and
465: the thermal diffusion coefficient $D^{\prime }$. Explicit
466: expressions for these were obtained in Ref.\ \onlinecite{GD02}
467: using a first Sonine approximation. Dimensionless forms are
468: defined by
469: 
470: \begin{equation}
471: D=\frac{\rho T}{m_{1}m_{2}\nu _{0}}D^{\ast },\quad
472: D_{p}=\frac{nT}{\rho \nu _{0}}D_{p}^{\ast },\quad D^{\prime
473: }=\frac{nT}{\rho \nu _{0}}D^{\prime}{}^{\ast }.  \label{2.8}
474: \end{equation}
475: Here, $\nu _{0}=\sqrt{\pi }n\sigma _{12}^{2}v_{0}$,
476: $\sigma_{12}=(\sigma_1+\sigma_2)/2$, and  $v_{0}=\sqrt{
477: 2T(m_{1}+m_{2})/m_{1}m_{2}}$ is a thermal velocity defined in
478: terms of the temperature $T$ of the mixture. The explicit forms
479: are then
480: \begin{equation}
481: D^{\ast }=\left[ \left( \frac{\partial }{\partial
482: x_{1}}x_{1}\gamma _{1}\right) _{p,T}+\left( \frac{\partial \zeta
483: ^{\ast }}{\partial x_{1}} \right) _{p,T}\left( 1-\frac{\zeta
484: ^{\ast }}{2\nu ^{\ast }}\right) D_{p}^{\ast }\right] \left( \nu
485: ^{\ast }-\frac{1}{2}\zeta ^{\ast }\right)^{-1},  \label{2.9}
486: \end{equation}
487: \begin{equation}
488: D_{p}^{\ast }=x_{1}\left[ \gamma _{1}-\frac{\mu (1+\delta )}{1+\mu
489: \delta } \right] \left( \nu ^{\ast }-\frac{3}{2}\zeta ^{\ast
490: }+\frac{\zeta ^{\ast 2}}{ 2\nu ^{\ast }}\right) ^{-1},
491: \label{2.10}
492: \end{equation}
493: \begin{equation}
494: D^{\prime }{}^{\ast }=-\frac{\zeta ^{\ast }}{2\nu ^{\ast
495: }}D_{p}^{\ast }. \label{2.11}
496: \end{equation}
497: In these equations,
498: \begin{equation}
499: \gamma _{1}=\frac{T_{1}}{T}=\frac{\gamma }{1+x_{1}(\gamma
500: -1)},\quad \gamma _{2}=\frac{T_{2}}{T}=\frac{1}{1+x_{1}(\gamma
501: -1)},  \label{2.11a}
502: \end{equation}
503: where $\mu = m_{1}/m_{2}$ is the mass ratio, $\mu
504: _{ij}=m_{i}/(m_{i}+m_{j}),$ $\delta = x_{1}/x_{2}$, and $\gamma
505: =T_{1}/T_{2}$. The detailed forms for the temperature ratio
506: $\gamma ,$ dimensionless collision rate $\nu ^{\ast }$, and
507: dimensionless cooling rate $\zeta ^{\ast }$ are given in Appendix
508: A. Since $\mathbf{j}_{1}=-\mathbf{j}_{2}$ and $\nabla
509: x_{1}=-\nabla x_{2}$, it is expected that $D^{\ast }$ should be
510: symmetric with respect to interchange of particles $1$ and $2$
511: while $D_{p}^{\ast }$ and $D^{\prime }{}^{\ast}$ should be
512: antisymmetric. This can be easily verified by noting that $
513: x_{1}\gamma _{1}+x_{2}\gamma _{2}=1$.
514: 
515: 
516: 
517: In the case of elastic collisions $\alpha _{ij}=1$, $\zeta ^{\ast
518: }=0$, $ \gamma =1$, and Eqs.\ (\ref{2.9})--(\ref{2.11}) become
519: \begin{equation}
520: D^{\ast }=\frac{3}{8}\frac{1+\delta }{1-\mu _{12}(1-\delta
521: )},\quad D_{p}^{\ast }=x_{1}\frac{(1-\mu )}{1+\mu \delta }D^{\ast
522: }, \quad D^{\prime }{}^{\ast }=0. \label{2.11e}
523: \end{equation}
524: These coincide with known results obtained for elastic collisions
525: in the first Sonine approximation. \cite{CC70} Recently, it has
526: been shown that the estimate given by the first Sonine
527: approximation for the diffusion coefficient $D$ (in the very
528: dilute concentration limit $x_{1}\rightarrow 0$) compares quite
529: well with Monte Carlo simulations, \cite{GM04} except for the
530: cases in which the gas particles are much heavier and/or much
531: larger than impurities. For these extreme cases, the second Sonine
532: approximation to $D$ improves the accuracy of the kinetic theory
533: results. \cite{GM04}
534: 
535: 
536: 
537: 
538: \subsection{Heat flux}
539: \label{subsec2.2}
540: 
541: 
542: The heat flux requires going up to the second Sonine
543: approximation. The transport coefficients $D^{\prime \prime }$,
544: $L$, and $\lambda $ appearing in the heat flux (\ref{2.2}) are
545: given by \cite{GD02}
546: 
547: \begin{equation}
548: D^{\prime \prime }=-\frac{5}{2}\frac{n}{(m_{1}+m_{2})\nu
549: _{0}}\left[ \frac{ x_{1}\gamma _{1}^{3}}{\mu _{12}}d_{1}^{\prime
550: \prime }+\frac{x_{2}\gamma _{2}^{3}}{\mu _{21}}d_{2}^{\prime
551: \prime }-\left( \frac{\gamma _{1}}{\mu _{12}}-\frac{\gamma
552: _{2}}{\mu _{21}}\right) D^{\ast }\right] ,  \label{2.12}
553: \end{equation}
554: \begin{equation}
555: L=-\frac{5}{2}\frac{T}{(m_{1}+m_{2})\nu _{0}}\left[
556: \frac{x_{1}\gamma _{1}^{3}}{\mu _{12}}\ell _{1}+\frac{x_{2}\gamma
557: _{2}^{3}}{\mu _{21}}\ell _{2}-\left( \frac{\gamma _{1}}{\mu
558: _{12}}-\frac{\gamma _{2}}{\mu _{21}} \right) D_{p}^{\ast }\right]
559: ,  \label{2.13}
560: \end{equation}
561: \begin{equation}
562: \lambda =-\frac{5}{2}\frac{nT}{(m_{1}+m_{2})\nu _{0}}\left[ \frac{
563: x_{1}\gamma _{1}^{3}}{\mu _{12}}\lambda _{1}+\frac{x_{2}\gamma
564: _{2}^{3}}{\mu_{21}}\lambda _{2}-\left( \frac{\gamma _{1}}{\mu
565: _{12}}-\frac{\gamma _{2}}{ \mu _{21}}\right) D^{\prime }{}^{\ast
566: }\right] ,  \label{2.14}
567: \end{equation}
568: where the expressions for the (dimensionless) Sonine coefficients
569: $\{d_{i}^{\prime \prime },\ell _{i},\lambda _{i}\}$ are displayed
570: in Appendix \ref{appA}. In Eqs.\ (\ref{2.12})--(\ref{2.14}) it is
571: understood that the coefficients $D^{\ast }$, $D_{p}^{\ast }$, and
572:  $D^{\prime \ast }$ are given by
573:  Eqs.\ (\ref{2.9})--(\ref{2.11}), respectively (first
574: Sonine approximation). As expected, our results show that
575: $D^{\prime \prime }$ is antisymmetric with respect to the change
576: $1\leftrightarrow 2$ while $L$ and $ \lambda $ are symmetric.
577: Consequently, in the case of mechanically equivalent particles
578: ($m_{1}=m_{2}\equiv m$, $\sigma _{1}=\sigma _{2}\equiv \sigma$,
579: $\alpha _{ij}\equiv \alpha $), the coefficient $D^{\prime \prime}$
580: vanishes.
581: 
582: 
583: 
584: An equivalent representation is given in terms of the heat flow
585: $\mathbf{J}_{q}$ defined as
586: \begin{equation}
587: \mathbf{J}_{q}\equiv
588: \mathbf{q}-\frac{5}{2}T\sum_{i}\frac{\mathbf{j}_{i}}{
589: m_{i}}=\mathbf{q}-\frac{5}{2}T\frac{m_{2}-m_{1}}{m_{1}m_{2}}\mathbf{j}_{1},
590: \label{2.15}
591: \end{equation}
592: where in the second equality use has been made of the requirement
593: $\mathbf{j} _{1}=-\mathbf{j}_{2}$. The difference between
594: $\mathbf{q}$ and $\mathbf{J} _{q}$ is a heat flow due to
595: diffusion. In addition, for elastic collisions, $ \mathbf{J}_{q}$
596: is the flux conjugate to the temperature gradient in the form of
597: the entropy production where the contribution coming from the mass
598: flux couples only to the gradient of the chemical potentials. The
599: thermal conductivity in a mixture is generally measured in the
600: absence of diffusion, i.e., when $\mathbf{j}_{1}=\mathbf{0}$. To
601: identify this coefficient, we have to express $\mathbf{J}_{q}$ in
602: terms of $\mathbf{j}_{1}$, $\nabla T$, and $\nabla p$. The
603: corresponding coefficient of $\nabla T$ defines the thermal
604: conductivity. \cite{GM84} According to Eq.\ (\ref{2.1}), the
605: gradient of mole fraction $\nabla x_{1}$ is
606: \begin{equation}
607: \nabla x_{1}=-\frac{\rho }{m_{1}m_{2}nD}\mathbf{j}_{1}-\frac{\rho
608: ^{2}}{ m_{1}m_{2}p}\frac{D^{\prime }}{D}\nabla T-\frac{\rho
609: ^{2}}{m_{1}m_{2}np} \frac{D_{p}}{D}\nabla p.  \label{2.16}
610: \end{equation}
611: The expression of $\mathbf{J}_{q}$ is obtained by substituting
612: Eq.\ (\ref {2.2}) into Eq.\ (\ref{2.15}) and eliminating $\nabla
613: x_{1}$ by using the identity (\ref{2.16}). Thus, the heat flow is
614: given by
615: \begin{equation}
616: \mathbf{J}_{q}=-\kappa \nabla T+\frac{\rho
617: p}{m_{1}m_{2}n^{2}}\kappa_{\text{T}}\mathbf{j}_{1}-L_{p}\nabla p,
618: \label{2.17}
619: \end{equation}
620: where
621: \begin{equation}
622: \kappa =\lambda -\frac{\rho ^{2}T^{2}}{m_{1}m_{2}p}\frac{D^{\prime \prime
623: }D^{\prime }}{D},  \label{2.18}
624: \end{equation}
625: \begin{equation}
626: \kappa _{\text{T}}=T\frac{D^{\prime \prime
627: }}{D}-\frac{5}{2}\frac{n}{\rho } (m_{2}-m_{1}),  \label{2.19}
628: \end{equation}
629: \begin{equation}
630: L_{p}=L-\frac{\rho ^{2}T}{n^{2}m_{1}m_{2}}\frac{D^{\prime \prime }D_{p}}{D}.
631: \label{2.20}
632: \end{equation}
633: As in the elastic case, the coefficient $\kappa $ is the thermal
634: conductivity while $\kappa _{\text{T}}$ is called the thermal-diffusion
635: factor or Dufour coefficient. There is a \emph{new} contribution
636: proportional to $\nabla p$ not present in the elastic case that defines the
637: transport coefficient $L_{p}$.
638: 
639: 
640: For elastic collisions, $L_{p}=0$ \cite{comment} and the
641: expressions derived here for $\kappa $ and $\kappa _{\text{T}}$
642: coincide with those obtained for a gas-mixture of elastic hard
643: spheres. \cite{CC70} Furthermore, in the case of mechanically
644: equivalent particles, the Dufour coefficient $\kappa _{\text{T}}$
645: vanishes as expected and the heat flow (\ref {2.17}) can be
646: written as
647: \begin{equation}
648: \mathbf{J}_{q}=-\overline{\kappa }\nabla T-\overline{\mu }\nabla n,
649: \label{2.21}
650: \end{equation}
651: where
652: \begin{equation}
653: \overline{\kappa }=\kappa +nL_{p}=\frac{25}{32}\left( \frac{mT}{\pi }\right)
654: ^{1/2}\sigma ^{-2}\frac{1+c}{\nu _{\kappa }-2\zeta ^{\ast }},  \label{2.22}
655: \end{equation}
656: \begin{equation}
657: \overline{\mu }=TL_{p}=\frac{75}{32}\frac{T}{n}\left(
658: \frac{mT}{\pi }\right) ^{1/2}\sigma ^{-2}\zeta ^{\ast }\left(
659: \frac{2}{3}\frac{1+c}{\nu _{\kappa }-2\zeta ^{\ast
660: }}+\frac{1}{3}\frac{c}{\zeta ^{\ast }}\right) \left(2\nu _{\kappa
661: }-3\zeta ^{\ast }\right) ^{-1},  \label{2.23}
662: \end{equation}
663: \begin{equation}
664: \nu _{\kappa }=\frac{1}{3}(1+\alpha )\left[
665: 1+\frac{33}{16}(1-\alpha )+\frac{ 19-3\alpha }{1024}c\right] ,
666: \label{2.24}
667: \end{equation}
668: \begin{equation}
669: c=\frac{32(1-\alpha )(1-2\alpha ^{2})}{81-17\alpha +30\alpha
670: ^{2}(1-\alpha )} ,\quad \zeta ^{\ast }=\frac{5}{12}(1-\alpha
671: ^{2})\left( 1+\frac{3}{32} c\right) .  \label{2.25}
672: \end{equation}
673: Note that in writing Eq.\ (\ref{2.21}) use has been made of the
674: relation $ \nabla p=n\nabla T+T\nabla n$. Equations
675: (\ref{2.21})--(\ref{2.25}) are the same as those obtained for the
676: one component granular gas. \cite{BDKS98} This confirms the
677: relevant known limiting cases for the granular mixture results
678: described here.
679: 
680: 
681: \section{Transport coefficients}
682: \label{sec3}
683: 
684: 
685: 
686: The transport coefficients depend on many parameters: $\left\{
687: x_{1},T,m_{1}/m_{2},\sigma _{1}/\sigma _{2}, \alpha _{11}, \alpha
688: _{22}, \alpha _{12}\right\} $. This complexity exists in the
689: elastic limit as well, so the primary new feature is the
690: dependence on the coefficients of restitution  $ \alpha _{ij}$
691: being different from unity. To illustrate the differences between
692: granular and normal gases the transport coefficients are
693: normalized to their values in the elastic limit. Then, the
694: dependence on the overall temperature scales out. Also, only the
695: simplest case of a common coefficient of restitution ($\alpha
696: _{11}=\alpha _{22}=\alpha _{12}\equiv \alpha $) and common size
697: $\omega \equiv \sigma _{1}/\sigma _{2}=1$ is considered. This
698: reduces the parameter set to three quantities:
699: $\{m_{1}/m_{2},x_{1},\alpha \}$.
700: 
701: 
702: In Figs.\ \ref{fig1}--\ref{fig6}, we plot the above transport
703: coefficients as functions of the coefficient of restitution
704: $\alpha $ for $x_{1}=0.2$, and several values of the mass ratio
705: $\mu$. It is understood that all coefficients have been reduced
706: with respect to their elastic values, except in the cases of
707: $D^{\prime }$ and $L_{p}$ since both coefficients vanish for
708: elastic collisions. In these latter two cases, we have considered
709: the reduced coefficients $D^{\prime }{}^{\ast }$ defined by Eq.\
710: (\ref{2.9}) and $ L_{p}^{\ast }=-\left[ (5/2)T\nu
711: _{0}/(m_{1}+m_{2})\right]^{-1}L_{p}$. Figure \ref{fig1} shows the
712: mutual diffusion coefficient as a function of $ \alpha $ for three
713: mass ratios $\mu=0.5,1,$ and $4$. There is a monotonic increase of
714: the coefficient with decreasing $\alpha $ in all cases. Moreover,
715: that effect increases as the mass of the dilute species increases.
716: This is consistent with an observed singular behavior in the
717: extreme case of tracer diffusion for a massive particle.
718: \cite{SD01} The velocity distributions in a granular gas are no
719: longer Maxwellian, \cite{GD99} and the difference is measured by
720: the coefficients $c_{i}$ that appear in the expressions for the
721: transport coefficients. \cite{GD02} The dashed curves in Fig.\
722: \ref{fig1} correspond to $c_{1}=c_{2}\rightarrow 0$, the Maxwell
723: limit. In this case it is seen that the distortion of the
724: Maxwellian is not very important for this transport coefficient
725: (see however, discussion of Fig.\ \ref{fig4} below). Figure
726: \ref{fig2} shows that the pressure diffusion coefficient has a
727: very similar behavior. The thermal diffusion coefficient vanishes
728: in the elastic limit and remains small and slightly negative when
729: the dilute species has small mass ratio, as illustrated in Fig.\
730: \ref{fig3}. However as the mass ratio becomes large it becomes
731: large and positive for strong dissipation. The effect of different
732: species temperatures is also shown on this graph. The dashed
733: curves correspond to setting $\gamma =T_{1}/T_{2}\rightarrow 1$.
734: This is seen to yield large errors, particularly in the case of
735: large mass ratio (temperature differences are greater for
736: mechanically different particles), indicating the real
737: quantitative effect of two different species temperatures in
738: granular gases.
739: \begin{figure}[tbp]
740: \begin{center}
741: \resizebox{7.5cm}{!}{\includegraphics{fig1new.eps}}
742: \end{center}
743: \caption{Plot of the reduced mutual diffusion coefficient
744: $D(\protect\alpha )/D(1)$ as a function of the coefficient of
745: restitution $\protect\alpha$ for $x_1=0.2$, $\protect\omega=1$ and
746: $\protect\mu=0.5$ (a), $\protect\mu=1$ (b), and $\protect\mu=4$
747: (c). The dashed lines correspond to the approximation
748: $c_1=c_2=0$.} \label{fig1}
749: \end{figure}
750: \begin{figure}[tbp]
751: \begin{center}
752: \resizebox{7.5cm}{!}{\includegraphics{fig2new.eps}}
753: \end{center}
754: \caption{Plot of the reduced pressure diffusion coefficient
755: $D_p(\protect \alpha)/D_{p}(1)$ as a function of the coefficient
756: of restitution $\protect \alpha$ for $x_1=0.2$, $\protect\omega=1$
757: and $\protect\mu=0.5$ (a), $ \protect\mu=2$ (b), and
758: $\protect\mu=4$ (c).} \label{fig2}
759: \end{figure}
760: \begin{figure}[tbp]
761: \begin{center}
762: \resizebox{7.5cm}{!}{\includegraphics{fig3.eps}}
763: \end{center}
764: \caption{Plot of the reduced thermal diffusion coefficient
765: $D^{^{\prime }\ast }(\protect\alpha )$ as a function of the
766: coefficient of restitution $ \protect\alpha $ for $x_{1}=0.2$,
767: $\protect\omega =1$ and $\protect\mu =0.5$ (a), $\protect\mu =2$
768: (b), and $\protect\mu =4$ (c). The dashed lines correspond to
769: $\protect\gamma =1$.} \label{fig3}
770: \end{figure}
771: 
772: 
773: The thermal conductivity is shown in Fig.\  \ref{fig4}, with the
774: same monotonic increase with increasing dissipation. The dashed
775: lines ($c_i=0$) indicate a significant effect of the distortion of
776: the reference distribution function from its Maxwellian form.
777: Presumably, this is due to the fact that the thermal conductivity
778: $\kappa$ depends on a higher velocity moment than the mutual
779: diffusion coefficient $D$ and is more sensitive to the larger
780: distortions at higher velocities. We also observe that there is
781: little mass dependence when the dilute species is lighter.
782: However, when the dilute species is more massive there is a
783: significant decrease in the thermal conductivity, opposite to the
784: case of diffusion. In contrast, the Dufour coefficient does have a
785: dependence on the mass ratio more like diffusion (Fig.\
786: \ref{fig5}). Finally, the coefficient $L_{p}^{\ast }$ is
787: illustrated in Figure \ref{fig6}. Again there is weak dependence
788: on the mass ratio until it becomes larger for the dilute species.
789: 
790: 
791: In summary, the mass and heat flux transport coefficients for a
792: granular mixture differ significantly from those for a normal gas
793: mixture even at moderate dissipation. In most cases (thermal
794: conductivity is an exception) the differences increase with
795: decreasing $\alpha $, depend weakly on the mass ratio when the
796: dilute species ($x_1/x_2<1$) is lighter than the excess species
797: ($m_{1}/m_{2}\leq 1$) but increase significantly in the opposite
798: case ($m_{1}/m_{2}>1$).
799: 
800: \begin{figure}[tbp]
801: \begin{center}
802: \resizebox{7.5cm}{!}{\includegraphics{fig4new.eps}}
803: \end{center}
804: \caption{Plot of the reduced thermal conductivity coefficient
805: $\protect\kappa (\protect\alpha)/\protect\kappa(1)$ as a function
806: of the coefficient of restitution $\protect\alpha$ for $x_1=0.2$,
807: $\protect\omega=1$ and $\protect \mu=0.5$ (a), $\protect\mu=1$
808: (b), and $\protect\mu=4$ (c). The dashed lines correspond to the
809: approximation $c_1=c_2=0$.} \label{fig4}
810: \end{figure}
811: \begin{figure}[tbp]
812: \begin{center}
813: \resizebox{7.5cm}{!}{\includegraphics{fig5new.eps}}
814: \end{center}
815: \caption{Plot of the reduced Dufour coefficient
816: $\protect\kappa_T(\protect \alpha)/\protect\kappa_{T}(1)$ as a
817: function of the coefficient of restitution $\protect\alpha$ for
818: $x_1=0.2$, $\protect\omega=1$ and $\protect \mu=0.5$ (a),
819: $\protect\mu=2$ (b), and $\protect\mu=4$ (c).} \label{fig5}
820: \end{figure}
821: \begin{figure}[tbp]
822: \begin{center}
823: \resizebox{7.5cm}{!}{\includegraphics{fig6.eps}}
824: \end{center}
825: \caption{Plot of the reduced coefficient $L_p^*(\protect\alpha)$
826: as a function of the coefficient of restitution $\protect\alpha$
827: for $x_1=0.2$, $ \protect\omega=1$ and $\protect\mu=0.5$ (a),
828: $\protect\mu=1$ (b), and $ \protect\mu=4$ (c).} \label{fig6}
829: \end{figure}
830: 
831: 
832: \section{Onsager's reciprocal relations}
833: \label{sec4}
834: 
835: 
836: 
837: In the usual language of the linear irreversible thermodynamics
838: for ordinary fluids, \cite{GM84} the constitutive equations for
839: the mass flux (\ref{2.1}) and heat flow (\ref{2.17}) is written
840: \begin{equation}
841: \mathbf{j}_{i}=-\sum_{i}L_{ij}\left( \frac{\nabla \mu
842: _{j}}{T}\right) _{T}-L_{iq}\frac{\nabla T}{T^{2}}-C_{p}\nabla p,
843: \label{4.1}
844: \end{equation}
845: \begin{equation}
846: \mathbf{J}_{q}=-L_{qq}\nabla T-\sum_{i}L_{qi}\left( \frac{\nabla
847: \mu _{i}}{T} \right) _{T}-C_{p}^{\prime }\nabla p,  \label{4.2}
848: \end{equation}
849: where
850: \begin{equation}
851: \left( \frac{\nabla \mu _{i}}{T}\right) _{T}=\frac{1}{m_{i}}\nabla
852: \ln (x_{i}p),  \label{4.3}
853: \end{equation}
854: $\mu _{i}$ being the chemical potential per unit mass. Here, the
855: coefficients $L_{ij}$ are the so-called Onsager phenomenological
856: coefficients. For \emph{normal} fluids, Onsager showed \cite{GM84}
857: that time reversal invariance of the underlying microscopic
858: equations of motion implies important restrictions on the above
859: set of transport coefficients
860: \begin{equation}
861: L_{ij}=L_{ji},\quad L_{iq}=L_{qi},\quad C_{p}=C_{p}^{\prime }=0.
862: \label{4.4}
863: \end{equation}
864: The first two symmetries are called reciprocal relations as they
865: relate transport coefficients for different processes. The last
866: two are statements that the pressure gradient does not appear in
867: any of the fluxes even though it is admitted by symmetry. Even for
868: a one component fluid, Onsager's theorem is significant as it
869: leads to Fourier's law for the heat flow rather than (\ref{2.21}),
870: i.e. $\overline{\mu }=0$. Since there is no time reversal symmetry
871: for granular fluids, Eqs. (\ref{4.4}) cannot be expected to apply.
872: However, since explicit expressions for all transport coefficients
873: are at hand, the quantitative extent of the violation can be
874: explored.
875: 
876: 
877: To make connection with the previous sections it is first
878: necessary to transform Eqs.\ (\ref{4.1})--(\ref{4.3}) to the
879: variables $x_{1},$ $p,$ $T.$ Since $\nabla x_{1}=-\nabla x_{2}$,
880: Eq.\ (\ref{4.3}) implies
881: \begin{equation}
882: \frac{(\nabla \mu _{1})_{T}-(\nabla \mu _{2})_{T}}{T}=\frac{n\rho
883: }{\rho _{1}\rho _{2}}\left[ \nabla x_{1}+\frac{n_{1}n_{2}}{n\rho }
884: (m_{2}-m_{1})\nabla \ln p\right] .  \label{4.5}
885: \end{equation}
886: The coefficients $L_{ij}$ then can be easily obtained in terms of
887: those of the previous sections. The result is
888: \begin{equation}
889: L_{11}=-L_{12}=-L_{21}=\frac{m_{1}m_{2}\rho _{1}\rho _{2}}{\rho
890: ^{2}}D,\quad L_{1q}=\rho TD^{\prime },  \label{4.6}
891: \end{equation}
892: \begin{equation}
893: L_{q1}=-L_{q2}=\frac{T^{2}\rho _{1}\rho _{2}}{n\rho }D^{\prime
894: \prime }- \frac{5}{2}\frac{T\rho _{1}\rho _{2}}{\rho
895: ^{2}}(m_{2}-m_{1})D,\quad L_{qq}=\lambda -\frac{5}{2}\rho
896: \frac{m_{2}-m_{1}}{m_{1}m_{2}}D^{\prime }, \label{4.7}
897: \end{equation}
898: \begin{equation}
899: C_{p}\equiv \frac{\rho}{p} D_{p}-\frac{\rho _{1}\rho _{2}}{p\rho
900: ^{2}}(m_{2}-m_{1})D, \label{4.8}
901: \end{equation}
902: \begin{equation}
903: C_{p}^{\prime }\equiv
904: L-\frac{5}{2}\frac{T}{p}\frac{m_{2}-m_{1}}{m_{1}m_{2}}C_{p}-
905: \frac{n_{1}n_{2}}{np\rho }T^{2}(m_2-m_1)D^{\prime \prime}.
906: \label{4.9}
907: \end{equation}
908: The Onsager's relation $L_{12}=L_{21}$ is already evident since
909: the diffusion coefficient $D$ is symmetric under the change
910: $1\leftrightarrow 2$ , as discussed following Eq.\ (\ref{2.11a})
911: 
912: 
913: 
914: Imposing Onsager's relation $L_{1q}=L_{q1}$ yields
915: \begin{equation}
916: D^{\prime \prime }=\frac{5}{2}\frac{n}{T\rho
917: }(m_{2}-m_{1})D+\frac{n\rho ^{2} }{T\rho _{1}\rho _{2}}D^{\prime
918: },  \label{4.10}
919: \end{equation}
920: while the condition $C_{p}=C_{p}^{\prime }=0$ leads to the
921: following additional requirements
922: \begin{equation}
923: D_{p}=\frac{\rho _{1}\rho _{2}}{\rho ^{3}}(m_{2}-m_{1})D,
924: \label{4.11}
925: \end{equation}
926: \begin{equation}
927: \frac{5}{2}\frac{T}{p}(m_{1}-m_{2})\left[ \frac{n_{1}n_{2}}{\rho
928: ^{2}}(m_{2}-m_{1})D- \frac{\rho }{m_{1}m_{2}}D_{p}\right]
929: =pL-\frac{n_{1}n_{2}}{n\rho } T^{2}(m_{2}-m_{1})D^{\prime \prime
930: }.  \label{4.12}
931: \end{equation}
932: Since the relations (\ref{4.10})--(\ref{4.12}) involve transport
933: coefficients that have been determined in the first Sonine
934: approximation, we restrict here our discussion to this level of
935: approximation. In this case, $ d_{i}^{\prime \prime }=\ell
936: _{i}=\lambda _{i}=0$ so that Onsager's theorem, Eqs.\
937: (\ref{4.10})--(\ref{4.12}), gives the conditions
938: \begin{equation}
939: P(\alpha _{ij})\equiv \left[ \gamma _{1}-1+\mu (1-\gamma
940: _{2})\right] \frac{ D^{\ast }}{\mu }+\frac{1}{5}\frac{(1+\delta
941: )(1+\mu \delta )}{\mu \delta } \frac{\zeta ^{\ast }}{\nu ^{\ast
942: }}D_{p}^{\ast }=0,  \label{4.13}
943: \end{equation}
944: \begin{equation}
945: Q(\alpha _{ij})\equiv D_{p}^{\ast }-x_{1}\frac{(1-\mu )}{(1+\mu
946: \delta )} D^{\ast }=0,  \label{4.14}
947: \end{equation}
948: \begin{equation}
949: R(\alpha _{ij})\equiv \frac{1+\mu }{\mu }\left( \gamma _{1}-\mu
950: \gamma _{2}\right) Q(\alpha _{ij})=0.  \label{4.15}
951: \end{equation}
952: 
953: 
954: In the elastic limit, the reduced coefficients $D_{p}^{\ast }$ and
955: $D^{\ast } $ are given by Eq.\ (\ref{2.11}) and these conditions
956: are verified. Also, for mechanically equivalent particles with
957: arbitrary $\alpha $, $\gamma _{i}=1$ and $D_{p}^{\ast }=0$ so that
958: $P(\alpha )=Q(\alpha )=R(\alpha )=0$. Nevertheless, beyond these
959: limit cases, Onsager's relations do not apply (as expected). At
960: this macroscopic level the origin of this failure is due to the
961: cooling of the reference state as well as the occurrence of
962: different kinetic temperatures for both species. Figures
963: \ref{fig7}, \ref{fig8}, and \ref{fig9} show the dependence of the
964: quantities $P$, $Q$, and $R$, respectively, on the (common)
965: coefficient of restitution $\alpha_{ij}\equiv \alpha$ for mass
966: ratios $\mu =0.5,2,$ and $4$.
967: \begin{figure}[tbp]
968: \begin{center}
969: \resizebox{7.5cm}{!}{\includegraphics{fig7.eps}}
970: \end{center}
971: \caption{Plot of $P(\protect\alpha)$ as a function of
972: $\protect\alpha$ for $ x_1=0.2$, $\protect\omega=1$ and
973: $\protect\mu=0.5$ (a), $\protect\mu=2$ (b), and $\protect\mu=4$
974: (c).} \label{fig7}
975: \end{figure}
976: \begin{figure}[tbp]
977: \begin{center}
978: \resizebox{7.5cm}{!}{\includegraphics{fig8.eps}}
979: \end{center}
980: \caption{Plot of $Q(\protect\alpha)$ as a function of
981: $\protect\alpha$ for $ x_1=0.2$, $\protect\omega=1$ and
982: $\protect\mu=0.5$ (a), $\protect\mu=2$ (b), and $\protect\mu=4$
983: (c).} \label{fig8}
984: \end{figure}
985: \begin{figure}[tbp]
986: \begin{center}
987: \resizebox{7.5cm}{!}{\includegraphics{fig9.eps}}
988: \end{center}
989: \caption{Plot of $R(\protect\alpha)$ as a function of
990: $\protect\alpha$ for $ x_1=0.2$, $\protect\omega=1$ and
991: $\protect\mu=0.5$ (a), $\protect\mu=2$ (b), and $\protect\mu=4$
992: (c).} \label{fig9}
993: \end{figure}
994: 
995: 
996: 
997: \section{Linearized hydrodynamic equations and stability}
998: \label{sec5}
999: 
1000: 
1001: 
1002: In contrast to normal fluids, the Navier--Stokes hydrodynamic
1003: equations\ (\ref {2.4})--(\ref{2.7}) have non-trivial solutions
1004: even for spatially homogeneous states,
1005: \begin{equation}
1006: \label{5.0}
1007: \partial _{t}x_{1H}=0=\partial _{t}u_{H\ell },
1008: \end{equation}
1009: \begin{equation}
1010: \left[ \partial _{t}+\zeta \left( x_{1H},T_{H},p_{H}\right)
1011: \right] T_{H}=0, \hspace{0.3in}\left[ \partial _{t}+\zeta \left(
1012: x_{1H},T_{H},p_{H}\right) \right] p_{H}=0.  \label{5.1}
1013: \end{equation}
1014: where the subscript $H$ denotes the homogeneous state. Since the
1015: dependence of the cooling rate $\zeta \left(
1016: x_{1H},T_{H},p_{H}\right)$ on $x_{1H},T_{H},p_{H}$ is known (see
1017: Appendix A), these first order nonlinear equations can be solved
1018: for the time dependence of the homogeneous state. The result is
1019: the familiar Haff's law cooling law for $T(t)$ at constant
1020: density. \cite{BP04}. As discussed above, each species temperature
1021: also has the same time dependence but each with a different value
1022: \cite{GD99} In this section, the hydrodynamics for small initial
1023: spatial perturbations of this homogeneous cooling state (HCS) is
1024: discussed. For normal fluids such perturbations decay in time
1025: according to the hydrodynamic modes of diffusion (shear, thermal,
1026: mass) and damped sound propagation.  The analysis is for fixed
1027: coefficients of restitution different from unity in the long
1028: wavelength limit. It will be seen here that the corresponding
1029: modes for a granular fluid are then quite different from those for
1030: a normal fluid. An alternative analysis with fixed long wavelength
1031: and coefficient of restitution  approaching unity leads to usual
1032: normal fluid modes. Thus, the nature of hydrodynamic modes is non
1033: uniform with respect to the inelasticity and the wavelength of the
1034: perturbation.
1035: 
1036: 
1037: Let $\delta y_{\alpha }(\mathbf{r},t)=y_{\alpha
1038: }(\mathbf{r},t)-y_{H\alpha }(t)$ denote the deviation of
1039: $\{x_{1},\mathbf{u},T,p\}$ from their values in the HCS. If the
1040: initial spatial perturbation is sufficiently small, then for some
1041: initial time interval these deviations will remain small and the
1042: hydrodynamic equations (\ref{2.4})--(\ref{2.7}) can be linearized
1043: with respect to $\delta y_{\alpha }(\mathbf{r},t)$. This leads to
1044: partial differential equations with coefficients that are
1045: independent of space but which depend on time since the HCS is
1046: cooling. As in the one component case, \cite{BDKS98,G05} this time
1047: dependence can be eliminated through a change in the time and
1048: space variables, and a scaling of the hydrodynamic fields.  We
1049: introduce the following dimensionless space and time variables:
1050: \begin{equation}
1051: \tau =\int_{0}^{t}dt^{\prime }\nu_{H}(t^{\prime }),\quad
1052: \mathbf{r}^{\prime }= \mathbf{r/}\ell _{H},  \label{5.2}
1053: \end{equation}
1054: where $v_{0H}=\sqrt{2T_{H}(m_{1}+m_{2})/m_{1}m_{2}}$ is the
1055: thermal velocity introduced above, $\ell_{H}=1/\sqrt{\pi }
1056: n_{H}\sigma _{12}^{2}$ is an effective mean free path, and
1057: $\nu_{H}(t)=v_{0H}(t)/\ell _{H}$ is the effective collision
1058: frequency. The dimensionless time scale is therefore an average
1059: number of collisions up to the time $t$. A set of Fourier
1060: transformed dimensionless variables are then defined by
1061: \begin{equation}
1062: \delta y_{\mathbf{k}\alpha }(\tau )=\int
1063: d\mathbf{r}^{\prime}\;e^{-i\mathbf{k} \cdot
1064: \mathbf{r}^{\prime}}\delta y_{\alpha }(\mathbf{r}^{\prime },
1065: \tau), \label{5.3}
1066: \end{equation}
1067: \begin{equation}
1068: \rho_{\mathbf{k}}(\tau )=\frac{\delta x_{1\mathbf{k}}(\tau
1069: )}{x_{1H}},\quad \mathbf{w}_{\mathbf{k}}(\tau )=\frac{\delta
1070: \mathbf{u}_{\mathbf{k}}(\tau )}{ v_{0H}(\tau )},  \label{5.4}
1071: \end{equation}
1072: \begin{equation}
1073: \theta _{\mathbf{k}}(\tau )=\frac{\delta T_{\mathbf{k}}(\tau
1074: )}{T_{H}(\tau )} ,\quad \Pi _{\mathbf{k}}(\tau )=\frac{\delta
1075: p_{\mathbf{k}}(\tau )}{ p_{H}(\tau )}.  \label{5.5}
1076: \end{equation}
1077: In terms of these variables the linearized hydrodynamic equations
1078: for the set $\{\rho _{\mathbf{k}},\mathbf{w}_{\mathbf{k}},\theta
1079: _{\mathbf{k}},\Pi _{ \mathbf{k}}\}$ separate into a degenerate
1080: pair of equations for the two transverse velocity components
1081: $w_{\mathbf{k}\perp }$ (orthogonal to $ \mathbf{k}$)
1082: \begin{equation}
1083: \left( \frac{\partial }{\partial \tau }-\frac{\zeta ^{\ast
1084: }}{2}+\eta ^{\ast }k^{2}\right) w_{\mathbf{k}\perp }=0,
1085: \label{5.6}
1086: \end{equation}
1087: and a coupled set of equations for $\rho _{\mathbf{k}},\theta
1088: _{\mathbf{k} },\Pi _{\mathbf{k}}$, and the longitudinal velocity
1089: component $w_{\mathbf{k} ||}$ (parallel to $\mathbf{k}$)
1090: \begin{equation}
1091: \frac{\partial \delta z_{\mathbf{k}\alpha }(\tau )}{\partial \tau
1092: }=\left( M_{\alpha \beta }^{(0)}+ikM_{\alpha \beta
1093: }^{(1)}+k^{2}M_{\alpha \beta }^{(2)}\right) \delta
1094: z_{\mathbf{k}\beta }(\tau ),  \label{5.7}
1095: \end{equation}
1096: where now $\delta z_{\mathbf{k}\alpha }(\tau )$ denotes the four
1097: variables $ \left( \rho _{\mathbf{k}},\theta _{\mathbf{k}},\Pi
1098: _{\mathbf{k}},w_{\mathbf{k }||}\right) $. The matrices in this
1099: equation are
1100: \begin{equation}
1101: M^{(0)}=\left(
1102: \begin{array}{cccc}
1103: 0 & 0 & 0 & 0 \\
1104: -x_{1}\left( \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right)
1105: _{T,p} &
1106: \frac{\zeta ^{\ast }}{2} & -\zeta ^{\ast } & 0 \\
1107: -x_{1}\left( \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right)
1108: _{T,p} &
1109: \frac{\zeta ^{\ast }}{2} & -\zeta ^{\ast } & 0 \\
1110: 0 & 0 & 0 & \frac{\zeta ^{\ast }}{2}
1111: \end{array}
1112: \right) ,  \label{5.8}
1113: \end{equation}
1114: \begin{equation}
1115: M^{(1)}=\left(
1116: \begin{array}{cccc}
1117: 0 & 0 & 0 & 0 \\
1118: 0 & 0 & 0 & -\frac{2}{3} \\
1119: 0 & 0 & 0 & -\frac{5}{3} \\
1120: 0 & 0 & -\frac{1}{2}\frac{\mu _{12}}{x_{1}\mu +x_{2}} & 0
1121: \end{array}
1122: \right) ,  \label{5.9}
1123: \end{equation}
1124: \begin{equation}
1125: M^{(2)}=\left(
1126: \begin{array}{cccc}
1127: -\frac{1}{2}\frac{\mu x_{1}+x_{2}}{1+\mu }D^{\ast } &
1128: -\frac{1}{2x_{1}}\frac{ \mu x_{1}+x_{2}}{1+\mu }D^{\prime
1129: }{}^{\ast } & -\frac{1}{2x_{1}}\frac{\mu
1130: x_{1}+x_{2}}{1+\mu }D_{p}^{\ast } & 0 \\
1131: -x_{1}\left( \frac{2}{3}D^{\prime \prime }{}^{\ast }-\frac{1-\mu
1132: }{2(1+\mu )} D^{\ast }\right)  & \frac{1-\mu }{2(1+\mu )}D^{\prime
1133: }{}^{\ast }-\frac{2}{3} \lambda ^{\ast } & -\frac{2}{3}L^{\ast
1134: }+\frac{1-\mu }{2(1+\mu )}D_{p}^{\ast
1135: } & 0 \\
1136: -\frac{2}{3}x_{1}D^{\prime \prime }{}^{\ast} & -\frac{2}{3}
1137: \lambda ^{\ast } & -\frac{2}{3}L^{\ast } & 0 \\
1138: 0 & 0 & 0 & -\frac{4}{3}\eta ^{\ast }
1139: \end{array}
1140: \right) .  \label{5..10}
1141: \end{equation}
1142: In these equations, $x_{i}=n_{iH}/n_{H}$, $\zeta ^{\ast }=\zeta
1143: _{H}/\nu _{H} $ and the reduced transport coefficients $D^{\ast
1144: }$, $ D_{p}^{\ast }$, and $D^{\prime }{}^{\ast }$ are given by
1145: Eqs.\ (\ref{2.9})--(\ref{2.11} ), respectively. Moreover, the
1146: reduced Navier--Stokes transport coefficients are
1147: \begin{equation}
1148: \eta ^{\ast }=\frac{\nu _{H}\eta }{\rho _{H}v_{0H}^{2}},
1149: \label{5.11}
1150: \end{equation}
1151: \begin{equation}
1152: D^{\prime \prime }{}^{\ast }=\frac{\nu _{H}T_{H}D^{\prime \prime
1153: }}{ n_{H}v_{0H}^{2}},\quad L^{\ast }=\frac{\nu
1154: _{H}L}{v_{0H}^{2}},\quad \lambda ^{\ast }=\frac{\nu _{H}\lambda
1155: }{n_{H}v_{0H}^{2}},  \label{5.12}
1156: \end{equation}
1157: where $\rho_H=m_1n_{1H}+m_2n_{2H}$.
1158: 
1159: It is instructive to consider first the solutions to these
1160: equations in the extreme long wavelength limit, $k=0$. In this
1161: case, the eigenvalues or hydrodynamic modes are given by
1162: \begin{equation}
1163: s_{\perp }=\frac{1}{2}\zeta ^{\ast },\hspace{0.3in}s_{n}=\left(
1164: 0,0,-\frac{1}{2} \zeta ^{\ast },\frac{1}{2}\zeta ^{\ast }\right),
1165: \label{5.14}
1166: \end{equation}
1167: where $s_n$ refers to the longitudinal modes. Two of the
1168: eigenvalues are positive, corresponding to growth of the initial
1169: perturbation in time. Thus, some of the solutions are unstable.
1170: The two zero eigenvalues represent marginal stability solutions,
1171: while the negative eigenvalue gives stable solutions. For general
1172: initial perturbations all modes are excited. These modes
1173: correspond to evolution of the fluid due to uniform perturbations
1174: of the HCS, i.e. a global change in the HCS parameters. The
1175: unstable modes are seen to arise from the initial perturbations
1176: $w_{\mathbf{k}\perp }(0)$ or $w_{\mathbf{k}||}(0)$. The marginal
1177: modes correspond to changes in the composition at fixed pressure,
1178: density, and velocity, and to changes in $\Pi_{{\bf
1179: k}}-\theta_{{\bf k}}$ at constant composition and velocity. The
1180: decaying mode corresponds to changes in the temperature or
1181: pressure for $\Pi_{{\bf k}}=\theta_{{\bf k}}$. The unstable modes
1182: may appear trivial since they are due entirely to the
1183: normalization of the fluid velocity by the time dependent thermal
1184: velocity. However, this normalization is required by the scaling
1185: of the entire set of equations to obtain time independent
1186: coefficients.
1187: 
1188: At finite wave vectors, these instabilities give rise to real
1189: growth of spatial perturbations. The linear growth of the
1190: transverse modes is simply given by
1191: \begin{equation}
1192: w_{\mathbf{k}\perp }(\tau )=w_{\mathbf{k}\perp }(0)\exp
1193: (\frac{1}{2}\zeta ^{\ast }-\eta ^{\ast }k^{2})\tau .  \label{5.23}
1194: \end{equation}
1195: The instability for the two shear modes is removed at sufficiently
1196: large $ k>k_{\perp }^{\text{c}}$, where
1197: \begin{equation}
1198: k_{\perp }^{\text{c}}=\left( \frac{\zeta ^{\ast }}{2\eta ^{\ast
1199: }}\right) ^{1/2}.  \label{5.25}
1200: \end{equation}
1201: The wave vector dependence of the remaining four modes is more
1202: complex. This is illustrated in Fig.\ \ref{fig10} showing the real
1203: parts of the modes $ s\left( k\right) $ for $\alpha _{ij}=0.9,$
1204: $\sigma _{1}/\sigma _{2}=1$, $ x_{1}=0.2$, and $m_{1}/m_{2}=4$.
1205: The $k=0$ values are those of (\ref{5.14}), corresponding to six
1206: hydrodynamic modes with two different degeneracies. The shear mode
1207: degeneracy remains at finite $k$ but the other is removed at any
1208: finite $k$. At sufficiently large $k$ a pair of real modes become
1209: equal and become a complex conjugate pair at all larger wave
1210: vectors, like a sound mode. The smaller of the unstable modes is
1211: that associated with the longitudinal velocity, which couples to
1212: the scalar hydrodynamic fields. It becomes negative at a wave
1213: vector smaller than that of Eq.\ (\ref{5.25}) and gives the
1214: threshold for development of spatial instabilities.
1215: 
1216: 
1217: 
1218: The results obtained here for the mixture show no new surprises
1219: relative to the earlier work for a one component gas,
1220: \cite{BDKS98,G05} with only the addition of the stable mass
1221: diffusion mode. Of course, the quantitative features can be quite
1222: different since there are additional degrees of freedom with the
1223: parameter set $\left\{ x_{1},T,m_{1}/m_{2},\sigma _{1}/\sigma
1224: _{2},\alpha _{ij}\right\} .$ Also, the manner in which these
1225: linear instabilities are enhanced by the nonlinearities may be
1226: different from that for the one component case.
1227: \begin{figure}[tbp]
1228: \begin{center}
1229: \resizebox{7.5cm}{!}{\includegraphics{fig10.eps}}
1230: \end{center}
1231: \caption{Dispersion relations for $\protect\alpha =0.9$,
1232: $x_{1}=0.2$, $ \protect\omega =1$ and $\protect\mu =4$.}
1233: \label{fig10}
1234: \end{figure}
1235: 
1236: 
1237: 
1238: \section{Summary and Discussion}
1239: \label{sec6}
1240: 
1241: 
1242: The Navier--Stokes order hydrodynamic equations have been
1243: discussed for a low density granular binary mixture. The form of
1244: the momentum flux is the same as for a one component gas, with
1245: only the value of the viscosity changed (see Appendix \ref{appA}).
1246: Since the dependence of the viscosity on the parameters of the
1247: mixture has been widely explored in a previous paper, \cite{MG03}
1248: attention has been focused here on the mass and heat fluxes and
1249: their associated transport coefficients. There is no phenomenology
1250: involved as the equations and the transport coefficients have been
1251: derived systematically from the inelastic Boltzmann equation by
1252: the Chapman-Enskog procedure. Consequently, there is no \emph{a
1253: priori} limitation on the degree of inelasticity, size and mass
1254: ratios, or composition. For practical purposes, the integral
1255: equations determining the transport coefficients have been solved
1256: by truncated expansions in Sonine polynomials. This is expected to
1257: fail at extreme values of size or mass ratio, \cite{MG03,GM04} but
1258: the results are quite accurate otherwise.
1259: 
1260: The hydrodynamic equations are the same as for a normal gas,
1261: except for a sink in the energy equation due to granular cooling,
1262: and additional transport coefficients in the mass and heat flux
1263: constitutive equations. The latter arise because the usual
1264: restrictions of irreversible thermodynamics no longer apply. These
1265: restrictions include Onsager reciprocal relations among various
1266: transport coefficients, and the extent to which these are violated
1267: has been demonstrated in Section 4. It has been verified that the
1268: results described here reduce to those for a normal mixture in the
1269: elastic limit, \cite{CC70} and to those for a one component
1270: granular gas \cite{BDKS98} when the species are mechanically
1271: identical.
1272: 
1273: As is the case for a normal gas, the hydrodynamic fields include
1274: only the global temperature even though two species temperatures
1275: can be defined. For a normal gas the species temperatures rapidly
1276: approach the global temperature due to equipartition. For the
1277: granular gas the species temperatures approach different values,
1278: but with the same time dependence as the global temperature. The
1279: transport coefficients have an additional dependence on the
1280: composition due to this time independent ratio of species
1281: temperatures. This has been illustrated in Fig.\ \ref{fig3} for
1282: the thermal diffusion coefficient, where the effect is seen to be
1283: large for large mass ratio (implying very different species
1284: temperatures. Generally, it was seen that the deviation due to
1285: inelasticity is enhanced for greater mechanical differences
1286: between the species.
1287: 
1288: The linear equations for small perturbations of the special
1289: homogeneous cooling solution to the hydrodynamic equations were
1290: obtained and discussed. In order to characterize the solutions in
1291: terms of modes, it is necessary to introduce dimensionless fields
1292: so that the time dependence of the reference HCS is eliminated.
1293: The resulting equations exhibit a long wavelength instability for
1294: three of the modes. This is quite similar to the case of a one
1295: component granular gas, \cite{BDKS98,G05} and in fact the same
1296: modes are unstable here. The additional diffusion mode for two
1297: species behaves as for a normal fluid. The consequences of this
1298: instability for a binary mixture were not studied here. This
1299: entails an analysis of the dominant nonlinearities which has not
1300: been performed as yet. Since there are additional degrees of
1301: freedom would be interesting to see if the density clustering that
1302: occurs for a one component system is more complex here (e.g.,
1303: species segregation).
1304: 
1305: \acknowledgments
1306: 
1307: Partial support of the Ministerio de Ciencia y Tecnolog\'{\i}a
1308: (Spain) through Grant No.  FIS2004-01399 (partially financed by
1309: FEDER funds) in the case of V.G. and ESP2003-02859 (partially
1310: financed by FEDER funds) in the case of J.M.M. is acknowledged. V.
1311: G. also acknowledges support from the European Community's Human
1312: Potential Programme HPRN-CT-2002-00307 (DYGLAGEMEM).
1313: 
1314: 
1315: \appendix
1316: \section{Some explicit expressions}
1317: \label{appA}
1318: 
1319: 
1320: 
1321: Navier--Stokes hydrodynamics retains terms up through second order
1322: in the gradients. As a scalar, the cooling rate has the most
1323: general form at this order given by
1324: \begin{eqnarray}
1325: \zeta  &=&\zeta _{0}+\zeta _{u}\nabla \cdot \mathbf{u}+\zeta
1326: _{x}\nabla ^{2}x_{1}+\zeta _{T}\nabla ^{2}T+\zeta _{p}\nabla
1327: ^{2}p+\zeta _{TT}\left( \nabla T\right) ^{2}+\zeta _{xx}\left(
1328: \nabla x_{1}\right) ^{2}\nonumber\\
1329: & &  +\zeta _{pp}\left( \nabla p\right) ^{2}+\zeta _{Tx}\left(
1330: \mathbf{\nabla }T\right) \cdot \left( \mathbf{\nabla }x_{1}\right)
1331: +\zeta _{Tp}\left( \mathbf{\nabla }T\right) \cdot \left(
1332: \mathbf{\nabla } p\right) +\zeta _{px}\left( \mathbf{\nabla
1333: }p\right) \cdot \left( \mathbf{ \nabla }x_{1}\right) \nonumber\\
1334: & & +\zeta _{uu}\left( \nabla _{i}u_{j}\right) \left( \nabla
1335: _{i}u_{j}\right) .  \label{a01}
1336: \end{eqnarray}
1337: For a low density gas the first order gradient term vanishes,
1338: \cite{BDKS98} $\zeta _{u}=0$. However, for higher densities $\zeta
1339: _{u}$ is different from zero. \cite{GD99a} The second order terms
1340: have been left implicit in equations (\ref{2.5}) and (\ref{2.6})
1341: for the temperature and the pressure. As noted in the text, these
1342: second order terms have been calculated for a one component fluid
1343: \cite{BDKS98} and found to be very small relative to corresponding
1344: terms from the fluxes. Consequently, they have been neglected in
1345: the linearized hydrodynamic equations (\ref{5.7}).
1346: 
1347: 
1348: \subsection{Mass flux parameters}
1349: 
1350: 
1351: 
1352: The transport coefficients are expressed in terms of a number of
1353: dimensionless parameters. For completeness, they are listed here.
1354: In general they depend on the reference distribution functions in
1355: the Chapman--Enskog expansion, which are not Maxwellians.
1356: \cite{GD99} The deviation of the reference distributions from
1357: their Maxwellian forms is measured by the cumulants $c_i$.
1358: \cite{GD99} As shown in Sec.\ \ref{sec3}, while the influence of
1359: these coefficients on the transport coefficients is not quite
1360: important in the case of the coefficients associated with the mass
1361: flux (see Fig.\ \ref{fig1}, for example), not happens the same in
1362: the case of the heat flux (see Fig.\ \ref{fig4}, for example)
1363: where the influence of $c_i$ is not negligible for strong
1364: dissipation. However, in order to offer a simplified theory the
1365: parameters given in this Appendix have neglected the corrections
1366: due to the cumulants $c_i$. The full expressions for the transport
1367: coefficients can be found in Ref.\ \onlinecite{GD02}.
1368: 
1369: 
1370: The temperature ratio $\gamma =T_{1}/T_{2}$ is determined from the
1371: condition
1372: \begin{equation}
1373: \label{n1} \zeta_1^*=\zeta_2^*=\zeta^*,
1374: \end{equation}
1375: where the dimensionless cooling rate, $\zeta_i ^{\ast}=\zeta_i
1376: /\nu _{0}$ ($\nu _{0}$ is the average frequency defined below
1377: (\ref{2.8})) is
1378: \begin{eqnarray}
1379: \zeta_1^{\ast } &=&\frac{2}{3}\sqrt{2\pi }\left( \frac{\sigma
1380: _{1}}{\sigma _{12}}\right) ^{2}x_{1}\theta _{1}^{-1/2}\left(
1381: 1-\alpha _{11}^{2}\right)
1382: \nonumber \\
1383: &&+\frac{4}{3}\sqrt{\pi }x_{2}\mu _{21}\left( \frac{1+\theta
1384: }{\theta }\right) ^{1/2}\left( 1+\alpha _{12}\right) \theta
1385: _{2}^{-1/2}\left[ 2-\mu _{21}\left( 1+\alpha _{12}\right) \left(
1386: 1+\theta \right) \right]. \label{a03}
1387: \end{eqnarray}
1388: The expression for $\zeta_2^*$ can be easily obtained by
1389: interchanging $1\leftrightarrow 2$. Here, $\theta _{1}=1/(\mu
1390: _{21}\gamma _{1})$, $\theta _{2}=1/(\mu _{12}\gamma _{2})$, $\mu
1391: _{ij}=m_{i}/(m_{i}+m_{j})$, $\delta =x_{1}/x_{2}$, and $\theta
1392: =\theta _{1}/\theta _{2}=\mu /\gamma $. The temperature ratios
1393: $\gamma_i$ are related to the temperature ratio $\gamma$ through
1394: Eqs.\ (\ref{2.11a}). In the quasielastic limit ($\alpha_{ij}\ll
1395: 1$), the temperature ratio $\gamma$ has the simple form
1396: \begin{eqnarray}
1397: \gamma  &\rightarrow&1+\frac{1}{2\mu _{12}\mu _{21}}\left\{ \left(
1398: \mu _{12}x_{1}-\mu
1399: _{21}x_{2}\right) \left( 1-\alpha _{12}\right) \right.   \nonumber \\
1400: &&\left. +\frac{1}{\sqrt{2}}\left[ \left( \frac{\sigma
1401: _{22}}{\sigma _{12}} \right) ^{2}x_{2}\sqrt{\mu _{12}}\left(
1402: 1-\alpha _{22}\right) -\left( \frac{ \sigma _{11}}{\sigma
1403: _{12}}\right) ^{2}x_{1}\sqrt{\mu _{21}}\left( 1-\alpha
1404: _{11}\right) \right] \right\} .  \label{a02}
1405: \end{eqnarray}
1406: In this limit, the temperature ratio is a linear function of the
1407: hydrodynamic field $x_{1}$. The dimensionless frequency $\nu
1408: ^{\ast}$ appearing in the expressions of the transport
1409: coefficients associated with the mass flux is
1410: \begin{equation}
1411: \nu ^{\ast }=\frac{4}{3}\mu _{21}\frac{1+\mu \delta }{1+\delta
1412: }\left( \frac{ 1+\theta }{\theta }\right) ^{1/2}\theta
1413: _{2}^{-1/2}(1+\alpha _{12}). \label{a04}
1414: \end{equation}
1415: The cooling rate $\zeta ^{\ast}$ and frequency $\nu ^{\ast }$ are
1416: also functions of the hydrodynamic field $x_{1}$. However, all the
1417: parameters above are independent of the temperature and density.
1418: 
1419: 
1420: 
1421: \subsection{Heat and momentum flux parameters}
1422: 
1423: 
1424: 
1425: As in the last section effects due to the distortion of the
1426: reference Maxwellian are neglected. In the case of the heat flux
1427: $\mathbf{q}$, Eq.\ ( \ref{2.2}), the transport coefficients
1428: $D^{\prime \prime }$, $L$, and $ \lambda $ are given by Eqs.\
1429: (\ref{2.12})--(\ref{2.14}), respectively. By using matrix
1430: notation, the (dimensionless) Sonine coefficients $ d_{i}^{\prime
1431: \prime }$, $\ell _{i}$, and $\lambda _{i}$ verify the coupled set
1432: of six equations \cite{GD02}
1433: \begin{equation}
1434: \Lambda _{\sigma \sigma ^{\prime }}X_{\sigma ^{\prime }}=Y_{\sigma
1435: }. \label{a1}
1436: \end{equation}
1437: where $X_{\sigma ^{\prime }}$ is the column matrix
1438: \begin{equation}
1439: \mathbf{X}=\left(
1440: \begin{array}{c}
1441: d_{1}^{\prime \prime } \\
1442: d_{2}^{\prime \prime } \\
1443: \ell _{1} \\
1444: \ell _{2} \\
1445: \lambda _{1} \\
1446: \lambda _{2}
1447: \end{array}
1448: \right) ,  \label{a2}
1449: \end{equation}
1450: and $\Lambda _{\sigma \sigma ^{\prime }}$ is the matrix
1451: \begin{equation}
1452: \Lambda =\left(
1453: \begin{array}{cccccc}
1454: \nu _{11}-\frac{3}{2}\zeta ^{\ast } & \nu _{12} & -\left(
1455: \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right) _{p,T} & 0 &
1456: -\left( \frac{\partial
1457: \zeta ^{\ast }}{\partial x_{1}}\right) _{p,T} & 0 \\
1458: \nu _{21} & \nu _{22}-\frac{3}{2}\zeta ^{\ast } & 0 & -\left(
1459: \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right) _{p,T} & 0 &
1460: -\left( \frac{\partial
1461: \zeta ^{\ast }}{\partial x_{1}}\right) _{p,T} \\
1462: 0 & 0 & \nu _{11}-\frac{5}{2}\zeta ^{\ast } & \nu _{12} & -\zeta
1463: ^{\ast } & 0
1464: \\
1465: 0 & 0 & \nu _{21} & \nu _{22}-\frac{5}{2}\zeta ^{\ast } & 0 &
1466: -\zeta ^{\ast }
1467: \\
1468: 0 & 0 & \zeta ^{\ast }/2 & 0 & \nu _{11}-\zeta ^{\ast } & \nu _{12} \\
1469: 0 & 0 & 0 & \zeta ^{\ast }/2 & \nu _{21} & \nu _{22}-\zeta ^{\ast
1470: }
1471: \end{array}
1472: \right) .  \label{a3}
1473: \end{equation}
1474: The column matrix $Y_{\sigma }$ has the elements
1475: \begin{equation}
1476: Y_{1}=D^{\ast }\left( \tau _{12}-\frac{\zeta ^{\ast }}{x_{1}\gamma
1477: _{1}^{2}} \right) -\frac{1}{\gamma _{1}^{2}}\left( \frac{\partial
1478: \gamma _{1}}{
1479: \partial x_{1}}\right) _{p,T},\hspace{0.3in}Y_{2}=-D^{\ast }\left( \tau
1480: _{21}-\frac{\zeta ^{\ast }}{x_{2}\gamma _{2}^{2}}\right)
1481: -\frac{1}{\gamma _{2}^{2}}\left( \frac{\partial \gamma
1482: _{2}}{\partial x_{1}}\right) _{p,T}, \label{a5}
1483: \end{equation}
1484: \begin{equation}
1485: Y_{3}=D_{p}^{\ast }\left( \tau _{12}-\frac{\zeta ^{\ast
1486: }}{x_{1}\gamma _{1}^{2}}\right) ,\hspace{0.3in}Y_{4}=-D_{p}^{\ast
1487: }\left( \tau _{21}-\frac{ \zeta ^{\ast }}{x_{2}\gamma
1488: _{2}^{2}}\right) ,  \label{a7}
1489: \end{equation}
1490: \begin{equation}
1491: Y_{5}=-\frac{1}{\gamma _{1}}+D^{\prime }{}^{\ast }\left( \tau
1492: _{12}-\frac{ \zeta ^{\ast }}{x_{1}\gamma _{1}^{2}}\right)
1493: ,\hspace{0.3in}Y_{6}=-\frac{1}{ \gamma _{2}}-D^{\prime }{}^{\ast
1494: }\left( \tau _{21}-\frac{\zeta ^{\ast }}{ x_{2}\gamma
1495: _{2}^{2}}\right) .  \label{a9}
1496: \end{equation}
1497: The dimensionless collision integrals $\tau _{12}$, $\nu _{11}$,
1498: and $\nu _{12}$ are given, respectively, by
1499: \begin{eqnarray}
1500: \tau _{12} &=&\frac{4}{3}\sqrt{\frac{\mu _{21}}{2}}\left(
1501: \frac{\sigma _{1}}{ \sigma _{12}}\right) ^{2}\gamma
1502: _{1}^{-3/2}(1-\alpha _{11}^{2}) \nonumber
1503: \label{a11} \\
1504: &&+\frac{4}{15}\mu _{21}^{-1}\gamma _{1}^{-4}(1+\alpha
1505: _{12})\left( \theta _{1}+\theta _{2}\right) ^{-1/2}\left( \theta
1506: _{1}\theta _{2}\right) ^{-3/2}\left( \frac{x_{2}}{x_{1}}A-\gamma
1507: B\right) ,
1508: \end{eqnarray}
1509: \begin{eqnarray}
1510: \label{a12}
1511:  \nu _{11} &=&\frac{16}{15}x_{1}\left( \frac{\sigma
1512: _{1}}{\sigma _{12}} \right) ^{2}(2\theta _{1})^{-1/2}(1+\alpha
1513: _{11})\left[ 1+\frac{33}{16}
1514: (1-\alpha _{11})\right]   \nonumber \\
1515: &&+\frac{2}{15}x_{2}\mu _{21}(1+\alpha _{12})\left( \frac{\theta
1516: _{1}}{ \theta _{2}(\theta _{1}+\theta _{2})}\right) ^{3/2}\left(
1517: E-5\frac{\theta _{1}+\theta _{2}}{\theta _{1}}A\right) ,
1518: \end{eqnarray}
1519: \begin{equation}
1520: \label{a13}
1521:  \nu _{12}=-\frac{2}{15}x_{2}\frac{\mu _{21}^{2}}{\mu
1522: _{12}}(1+\alpha _{12})\left( \frac{\theta _{1}}{\theta _{2}(\theta
1523: _{1}+\theta _{2})}\right) ^{3/2}\left( F+5\frac{\theta _{1}+\theta
1524: _{2}}{\theta _{2}}B\right) .
1525: \end{equation}
1526: In the above equations we have introduced the quantities
1527: \cite{misprints}
1528: \begin{eqnarray}
1529: \label{a14} A &=&5(2\beta _{12}+\theta _{2})+\mu _{21}(\theta
1530: _{1}+\theta _{2})\left[ 5(1-\alpha _{12})-2(7\alpha _{12}-11)\beta
1531: _{12}\theta _{1}^{-1}\right]
1532: \nonumber \\
1533: &&+18\beta _{12}^{2}\theta _{1}^{-1}+2\mu _{21}^{2}\left( 2\alpha
1534: _{12}^{2}-3\alpha _{12}+4\right) \theta _{1}^{-1}(\theta
1535: _{1}+\theta _{2})^{2}-5\theta _{2}\theta _{1}^{-1}(\theta
1536: _{1}+\theta _{2})
1537: \end{eqnarray}
1538: \begin{eqnarray}
1539: \label{a15} B &=&5(2\beta _{12}-\theta _{1})+\mu _{21}(\theta
1540: _{1}+\theta _{2})\left[ 5(1-\alpha _{12})+2(7\alpha _{12}-11)\beta
1541: _{12}\theta _{2}^{-1}\right]
1542: \nonumber \\
1543: &&-18\beta _{12}^{2}\theta _{2}^{-1}-2\mu _{21}^{2}\left( 2\alpha
1544: _{12}^{2}-3\alpha _{12}+4\right) \theta _{2}^{-1}(\theta
1545: _{1}+\theta _{2})^{2}+5(\theta _{1}+\theta _{2})
1546: \end{eqnarray}
1547: \begin{eqnarray}
1548: \label{a16} E &=&2\mu _{21}^{2}\theta _{1}^{-2}(\theta _{1}+\theta
1549: _{2})^{2}\left( 2\alpha _{12}^{2}-3\alpha _{12}+4\right) (5\theta
1550: _{1}+8\theta _{2})
1551: \nonumber \\
1552: &&-\mu _{21}(\theta _{1}+\theta _{2})\left[ 2\beta _{12}\theta
1553: _{1}^{-2}(5\theta _{1}+8\theta _{2})(7\alpha _{12}-11)+2\theta
1554: _{2}\theta
1555: _{1}^{-1}(29\alpha _{12}-37)-25(1-\alpha _{12})\right]   \nonumber \\
1556: &&+18\beta _{12}^{2}\theta _{1}^{-2}(5\theta _{1}+8\theta
1557: _{2})+2\beta
1558: _{12}\theta _{1}^{-1}(25\theta _{1}+66\theta _{2})  \nonumber \\
1559: &&+5\theta _{2}\theta _{1}^{-1}(11\theta _{1}+6\theta
1560: _{2})-5(\theta _{1}+\theta _{2})\theta _{1}^{-2}\theta
1561: _{2}(5\theta _{1}+6\theta _{2})
1562: \end{eqnarray}
1563: \begin{eqnarray}
1564: \label{a17} F &=&2\mu _{21}^{2}\theta _{2}^{-2}(\theta _{1}+\theta
1565: _{2})^{2}\left( 2\alpha _{12}^{2}-3\alpha _{12}+4\right) (8\theta
1566: _{1}+5\theta _{2})
1567: \nonumber \\
1568: &&-\mu _{21}(\theta _{1}+\theta _{2})\left[ 2\beta _{12}\theta
1569: _{2}^{-2}(8\theta _{1}+5\theta _{2})(7\alpha _{12}-11)-2\theta
1570: _{1}\theta
1571: _{2}^{-1}(29\alpha _{12}-37)+25(1-\alpha _{12})\right]   \nonumber \\
1572: &&+18\beta _{12}^{2}\theta _{2}^{-2}(8\theta _{1}+5\theta
1573: _{2})-2\beta
1574: _{12}\theta _{2}^{-1}(66\theta _{1}+25\theta _{2})  \nonumber \\
1575: &&+5\theta _{1}\theta _{2}^{-1}(6\theta _{1}+11\theta
1576: _{2})-5(\theta _{1}+\theta _{2})\theta _{2}^{-1}(6\theta
1577: _{1}+5\theta _{2})
1578: \end{eqnarray}
1579: Here, $\beta _{12}=\mu _{12}\theta _{2}-\mu _{21}\theta _{1}$. The
1580: corresponding expressions for $\tau _{21}$, $\nu _{22}$, and $\nu
1581: _{21}$ can be inferred from Eqs.\ (\ref{a11})--(\ref{a17}) by
1582: interchanging $ 1\leftrightarrow 2$. For elastic collisions, the
1583: expressions (\ref{a11})--( \ref{a17}) reduce to those obtained for
1584: hard sphere mixtures. \cite{CC70bis}
1585: 
1586: 
1587: The solution to Eq.\ (\ref{a1}) is
1588: \begin{equation}
1589: X_{\sigma }=\left( \Lambda ^{-1}\right) _{\sigma \sigma ^{\prime
1590: }}Y_{\sigma ^{\prime }}.  \label{a18.0}
1591: \end{equation}
1592: This relation provides an explicit expression for the coefficients
1593: $ d_{i}^{\prime \prime }$, $\ell _{i}$, and $\lambda _{i}$ in
1594: terms of the coefficients of restitution  and the parameters of
1595: the mixture. Their explicit forms are
1596: \begin{eqnarray}
1597: \label{a18} d_1^{\prime \prime}&=&\frac{1}{\Delta}\left\{2\left[2
1598: \nu_{12}Y_2-Y_1(2\nu_{22}-3\zeta^*)\right]\left[\nu_{12}\nu_{21}-\nu_{11}\nu_{22}
1599: +2(\nu_{11}+\nu_{22})\zeta^*-4\zeta^{*2}\right]\right.\nonumber\\
1600: && +2\left( \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right)
1601: _{p,T}(Y_3+Y_5)\left[2\nu_{12}\nu_{21}+2\nu_{22}^2-\zeta^*(
1602: 7\nu_{22}-6\zeta^{*})\right]\nonumber\\
1603: & & \left.-2\nu_{12}\left( \frac{\partial \zeta ^{\ast }}{\partial
1604: x_{1}}\right)_{p,T}(Y_4+Y_6)\left(2\nu_{11}+2\nu_{22}-7\zeta^*\right)\right\},
1605: \end{eqnarray}
1606: \begin{eqnarray}
1607: \label{a19} \ell_1&=&\frac{1}{\Delta}\left\{-2Y_3\left[2
1608: (\nu_{12}\nu_{21}-\nu_{11}\nu_{22})\nu_{22}+\zeta^*(7\nu_{11}\nu_{22}-5\nu_{12}\nu_{21}+2\nu_{22}^2
1609: -6\nu_{11}\zeta^*-7\nu_{22}\zeta^*+6\zeta^{*2})\right]\right.\nonumber\\
1610: &&
1611: +2Y_4\nu_{12}\left[2\nu_{12}\nu_{21}-2\nu_{11}\nu_{22}+2\zeta^*(\nu_{11}+\nu_{22})
1612: -\zeta^{*2}\right]\nonumber\\
1613: & &
1614: \left.+2Y_5\zeta^*\left[2\nu_{12}\nu_{21}+\nu_{22}(2\nu_{22}-7\zeta^*)+6\zeta^{*2}\right]
1615: -2\nu_{12}\zeta^*Y_6\left[2(\nu_{11}+\nu_{22})-7\zeta^*\right]
1616: \right\},
1617: \end{eqnarray}
1618: \begin{eqnarray}
1619: \label{a20} \lambda_1&=&\frac{1}{\Delta}\left\{-Y_3\zeta^*\left[2
1620: \nu_{12}\nu_{21}+\nu_{22}(2\nu_{22}-7\zeta^*)+6\zeta^{*2}\right]+\nu_{12}\zeta^*
1621: Y_4\left[2(\nu_{11}+\nu_{22})-7\zeta^*\right] \right. \nonumber\\
1622: &
1623: &-Y_5\left[4\nu_{12}\nu_{21}(\nu_{22}-\zeta^*)+2\nu_{22}^2(5\zeta^*-2\nu_{11})+2\nu_{11}
1624: (7\nu_{22}\zeta^*-6\zeta^{*2})+5\zeta^{*2}(6\zeta^*-7\nu_{22})\right]\nonumber\\
1625: & & \left.
1626: +\nu_{12}Y_6\left[4\nu_{12}\nu_{21}+2\nu_{11}(5\zeta^*-2\nu_{22})+\zeta^*(10\nu_{22}-
1627: 23\zeta^*)\right]\right\},
1628: \end{eqnarray}
1629: where
1630: \begin{equation}
1631: \label{a21}
1632: \Delta=\left[4(\nu_{12}\nu_{21}-\nu_{11}\nu_{22})+6\zeta^*(\nu_{11}+\nu_{22})-9\zeta^{*2}\right]
1633: \left[\nu_{12}\nu_{21}-\nu_{11}\nu_{22}+2\zeta^*(\nu_{11}+\nu_{22})-4\zeta^{*2}\right].
1634: \end{equation}
1635: The expressions for $d_2^{\prime \prime}$, $\ell_2$, and
1636: $\lambda_2$ can be obtained from Eqs.\ (\ref{a18})--(\ref{a20}) by
1637: setting $1\leftrightarrow 2$. From the above expressions one can
1638: easily get the transport coefficients $D^{\prime }$, $L$ and
1639: $\lambda $ from Eqs.\ (\ref{2.12})--(\ref{2.14}), respectively.
1640: They are functions of $x_{1}$ but independent of temperature and
1641: pressure.
1642: 
1643: 
1644: 
1645: The pressure tensor $P_{k,\ell }$ is given by
1646: \begin{equation}
1647: P_{k\ell }=p\delta _{k\ell }-\eta \left( \nabla _{\ell
1648: }u_{k}+\nabla _{k}u_{\ell }-\frac{2}{3}\delta _{k\ell }\nabla
1649: \cdot \mathbf{u}\right) , \label{a22}
1650: \end{equation}
1651: where $\eta $ is the shear viscosity coefficient. Its expression
1652: can be written as \cite{GD02}
1653: \begin{equation}
1654: \eta =\frac{nT}{\nu _{0}}\left( x_{1}\gamma _{1}^{2}\eta
1655: _{1}+x_{2}\gamma _{2}^{2}\eta _{2}\right) ,  \label{a22.1}
1656: \end{equation}
1657: with
1658: \begin{equation}
1659: \eta _{1}=\frac{2\gamma _{2}(2\lambda _{22}-\zeta ^{\ast
1660: })-4\gamma _{1}\lambda _{12}}{\gamma _{1}\gamma _{2}\left[ \zeta
1661: ^{\ast 2}-2\zeta ^{\ast }(\lambda _{11}+\lambda _{22})+4(\lambda
1662: _{11}\lambda _{22}-\lambda _{12}\lambda _{21})\right] },
1663: \label{a23}
1664: \end{equation}
1665: \begin{equation}
1666: \eta _{2}=\frac{2\gamma _{1}(2\lambda _{11}-\zeta ^{\ast
1667: })-4\gamma _{2}\lambda _{21}}{\gamma _{1}\gamma _{2}\left[ \zeta
1668: ^{\ast 2}-2\zeta ^{\ast}(\lambda _{11}+\lambda _{22})+4(\lambda
1669: _{11}\lambda _{22}-\lambda _{12}\lambda _{21})\right] }.
1670: \label{a24}
1671: \end{equation}
1672: The dimensionless quantities $\lambda _{ij}$ are given by
1673: \cite{GD02,MG03}
1674: \begin{eqnarray}
1675: \lambda _{11} &=&\frac{16}{5\sqrt{2}}x_{1}\left( \frac{\sigma
1676: _{1}}{\sigma _{12}}\right) ^{2}\theta _{1}^{-1/2}\left[
1677: 1-\frac{1}{4}(1-\alpha _{11})^{2}
1678: \right]   \nonumber  \label{a25} \\
1679: &&+\frac{8}{15}x_{2}\mu _{21}(1+\alpha _{12})\theta
1680: _{1}^{3/2}\theta _{2}^{-1/2}\left[ 6\theta _{1}^{-2}(\mu
1681: _{12}\theta _{2}-\mu _{21}\theta
1682: _{1})(\theta _{1}+\theta _{2})^{-1/2}\right.   \nonumber \\
1683: &&\left.+\frac{3}{2}\mu _{21}\theta _{1}^{-2}(\theta _{1}+\theta
1684: _{2})^{1/2}(3-\alpha _{12})+5\theta _{1}^{-1}(\theta _{1}+\theta
1685: _{2})^{-1/2}\right],
1686: \end{eqnarray}
1687: \begin{eqnarray}
1688: \lambda _{12} &=&\frac{8}{15}x_{2}\frac{\mu _{21}^{2}}{\mu
1689: _{12}}(1+\alpha _{12})\theta _{1}^{3/2}\theta _{2}^{-1/2}\left[
1690: 6\theta _{2}^{-2}(\mu _{12}\theta _{2}-\mu _{21}\theta
1691: _{1})(\theta _{1}+\theta
1692: _{2})^{-1/2}\right.   \nonumber  \label{a26} \\
1693: &&\left.+\frac{3}{2}\mu _{21}\theta _{2}^{-2}(\theta _{1}+\theta
1694: _{2})^{1/2}(3-\alpha _{12})-5\theta _{2}^{-1}(\theta _{1}+\theta
1695: _{2})^{-1/2}\right].
1696: \end{eqnarray}
1697: The corresponding expressions for $\lambda _{22}$ and $\lambda
1698: _{21}$ can be inferred from Eqs.\ (\ref{a23}) and (\ref{a24}) by
1699: interchanging $ 1\leftrightarrow 2$.
1700: 
1701: The program for calculating the cooling rates, the temperature
1702: ratio and the transport coefficients of the binary mixture can be
1703: obtained on request from the authors.
1704: 
1705: 
1706: \begin{thebibliography}{99}
1707: 
1708: \bibitem{GS95}A. Goldshtein and M. Shapiro, ``Mechanics of collisional motion of granular materials.
1709: Part 1. General hydrodynamic equations,'' J. Fluid Mech. {\bf
1710: 282}, 41 (1995); J. J. Brey, J. W. Dufty, and A. Santos,
1711: ``Dissipative dynamics for hard spheres,'' J. Stat. Phys. {\bf
1712: 87}, 1051 (1997).
1713: 
1714: \bibitem{BP04}N. V. Brilliantov and T. P\"oschel, {\em Kinetic Theory
1715: of Granular Gases} (Oxford University Press, Oxford, 2004).
1716: 
1717: 
1718: 
1719: \bibitem{CC70}S. Chapman and T. G. Cowling, {\em The Mathematical Theory of Nonuniform Gases}
1720: (Cambridge University Press, Cambridge, 1970).
1721: 
1722: 
1723: \bibitem{BDKS98}J. J. Brey, J. W. Dufty, C. S. Kim, and A. Santos,
1724: ``Hydrodynamics for granular flow at low density,'' Phys. Rev. E
1725: {\bf 58}, 4638 (1998); J. J. Brey and D. Cubero, ``Hydrodynamic
1726: transport coefficients of granular gases,''  in {\em Granular
1727: Gases} in {\em Lectures Notes in Physics}, edited by T. P\"oschel
1728: and S. Luding (Springer, Berlin, 2001), p. 59.
1729: 
1730: 
1731: \bibitem{BRMC99} See for instance, J. J. Brey, M. J. Ruiz-Montero,
1732: and D. Cubero, ``On the validity of linear hydrodynamics for
1733: low-density granular flows described by the Boltzmann equation,''
1734: Europhys. Lett. {\bf 48}, 359 (1999); J. J. Brey, M. J.
1735: Ruiz-Montero, D. Cubero, and R. Garc\'{\i}a-Rojo, ``Self-diffusion
1736: in freely evolving granular gases,'' Phys. Fluids {\bf 12}, 876
1737: (2000); V. Garz\'o and J. M. Montanero, ``Transport coefficients
1738: of a heated granular gas,'' Physica A {\bf 313}, 336 (2002); J. M.
1739: Montanero, A. Santos, and V. Garz\'o, ``DSMC evaluation of the
1740: Navier-Stokes shear viscosity of a granular fluid,'' in {\em
1741: Rarefied Gas Dynamics 24}, edited by M. Capitelli (AIP Conference
1742: Proceedings, 72, 2005), p. 803.
1743: 
1744: 
1745: 
1746: \bibitem{JM89}J. T. Jenkins and F. Mancini,
1747: ``Kinetic theory for binary mixtures of smooth, nearly elastic
1748: spheres,'' Phys. Fluids A {\bf 1}, 2050 (1989); P. Zamankhan,
1749: ``Kinetic theory for multicomponent dense mixtures of slightly
1750: inelastic spherical particles,''  Phys. Rev. E {\bf 52}, 4877
1751: (1995); B. Arnarson and J. T. Willits, ``Thermal diffusion in
1752: binary mixtures of smooth, nearly elastic spheres with and without
1753: gravity,'' Phys. Fluids {\bf 10}, 1324 (1998); J. T. Willits and
1754: B. Arnarson, ``Kinetic theory of a binary mixture of nearly
1755: elastic disks,'' Phys. Fluids {\bf 11}, 3116 (1999); M. Alam, J.
1756: T. Willits, B. Arnarson, and S. Luding, ``Kinetic theory of a
1757: binary mixture of nearly elastic disks with size and mass
1758: disparity ,'' Phys. Fluids {\bf 14}, 4085 (2002); B. Arnarson and
1759: J. T. Jenkins, ``Binary mixtures of inelastic hard spheres:
1760: Simplified constitutive theory,'' Phys. Fluids {\bf 16}, 4543
1761: (2004.)
1762: 
1763: \bibitem{MCK83}M. L\'opez de Haro, E. G. D. Cohen, and J. M.
1764: Kincaid, ``The Enskog theory for multicomponent mixtures. I.
1765: Linear response theory,'' J. Chem. Phys. {\bf 78}, 2746 (1983).
1766: 
1767: \bibitem{GD99}V. Garz\'o and J. W. Dufty, ``Homogeneous cooling
1768: state for a granular mixture,'' Phys. Rev. E  {\bf 60}, 5706
1769: (1999).
1770: 
1771: 
1772: \bibitem{computer}See for instance, J. M. Montanero and V. Garz\'o,
1773: ``Monte Carlo simulation of the homogeneous cooling state for a
1774: granular mixture,'' Granular Matter {\bf 4}, 17 (2002); A. Barrat
1775: and E. Trizac, ``Lack of energy equipartition in homogeneous
1776: heated binary granular mixtures,'' Granular Matter {\bf 4}, 57
1777: (2002); R. Clelland and C. M. Hrenya, ``Simulations of a
1778: binary-sized mixture of inelastic grains in rapid shear flow,''
1779: Phys. Rev. E {\bf 65}, 031301 (2002); S. R. Dahl, C. M. Hrenya, V.
1780: Garz\'o, and J. W. Dufty, ``Kinetic temperatures for a granular
1781: mixture,'' Phys. Rev. E {\bf 66}, 041301 (2002); R. Pagnani, U. M.
1782: B. Marconi, and A. Puglisi, ``Driven low density granular
1783: mixtures,'' Phys. Rev. E {\bf 66}, 051304 (2002); D. Paolotti, C.
1784: Cattuto, U. M. B. Marconi, and A. Puglisi,``Dynamical properties
1785: of vibrofluidized granular mixtures,''  Granular Matter {\bf 5},
1786: 75 (2003); P. Krouskop and J. Talbot, ``Mass and size effects in
1787: three-dimensional vibrofluidized granular mixtures,'' Phys. Rev. E
1788: {\bf 68}, 021304 (2003); H. Wang, G. Jin, and Y. Ma, ``Simulation
1789: study on kinetic temperatures of vibrated binary granular
1790: mixtures,'' Phys. Rev. E {\bf 68}, 031301 (2003); J. J. Brey, M.
1791: J. Ruiz-Montero, and F. Moreno, ``Energy partition and segregation
1792: for an intruder in a vibrated granular system under gravity,''
1793: Phys. Rev. Lett. {\bf 95}, 098001 (2005).
1794: 
1795: 
1796: \bibitem{exp}R. D. Wildman and D. J. Parker, `` Coexistence of two granular
1797: temperatures in binary vibrofluidized beds,'' Phys. Rev. Lett.
1798: {\bf 88}, 064301 (2002); K. Feitosa and N. Menon, ``Breakdown of
1799: energy equipartition in a 2D binary vibrated granular gas,'' Phys.
1800: Rev. Lett. {\bf 88}, 198301 (2002).
1801: 
1802: 
1803: \bibitem{JM87}J. Jenkins and F. Mancini, ``Balance laws and constitutive relations
1804: for plane flows of a dense bianry mixture of smooth, nearly
1805: elastic, circular disks,'' J. Appl. Mech. {\bf 54}, 27 (1987).
1806: 
1807: \bibitem{GD02}V. Garz\'o and J. W. Dufty, ``Hydrodynamics for a granular binary
1808: mixture at low density,'' Phys. Fluids {\bf 14}, 1476 (2002).
1809: 
1810: \bibitem {MG03}J. M. Montanero and V. Garz\'o, `Shear viscosity for a heated granular binary mixture
1811: at low-density,'' Phys. Rev. E {\bf 67}, 021308 (2003).
1812: 
1813: 
1814: \bibitem{GM04}V. Garz\'o and J. M. Montanero, ``Diffusion of impurities in a
1815: granular gas,'' Phys. Rev. E {\bf 69}, 021301 (2004).
1816: 
1817: 
1818: \bibitem{GDH05}L. Huilin, D. Gidaspow, and E. Manger, ``Kinetic theory of
1819: fluidized binary granular mixtures,'' Phys. Rev. E {\bf64}, 061301
1820: (2001); M.F. Ramahan, J. Naser, and P. J. Witt, ``An unequal
1821: temperature kinetic theory: description of granular flow with
1822: multiple particle classes,'' Powder Tech. {\bf 138}, 82 (2003); J.
1823: E. Galvin, S. R. Dahl, and C. M. Hrenya, ``On the influence of the
1824: equipartition-of-energy assumption in kinetic theories for
1825: rapidly-flowing, granular mixtures,'' J. Fluid Mech. {\bf 528},
1826: 207 (2005).
1827: 
1828: 
1829: \bibitem{DHGD02}S. R. Dahl, C. M. Hrenya, V. Garz\'o, and J. W.
1830: Dufty, ``Kinetic temperatures for a granular mixture,'' Phys. Rev.
1831: E {\bf 66}, 041301 (2002).
1832: 
1833: 
1834: 
1835: \bibitem{SGD04}A. Santos, V. Garz\'o, and J. Dufty, ``Inherent rheology
1836: of a granular fluid in uniform shear flow,'' Phys. Rev. E {\bf
1837: 69}, 061303 (2004).
1838: 
1839: 
1840: \bibitem{GM84}S. R. de Groot and P. Mazur, {\em Nonequilibrium Thermodynamics} (Dover, New York, 1984).
1841: 
1842: 
1843: 
1844: \bibitem{comment} Note that this coefficient is not strictly zero in
1845: the approximation used here due to the fact that while the set
1846: $\{d_{i}'', \ell_{i}, \lambda_{i}\}$ has been evaluated in the
1847: second Sonine approximation, the coefficients $D^*$, $D_p^*$, and
1848: $D'^*$ have been estimated in the first Sonine approximation.
1849: However, the magnitude of $L_p$ is negligible.
1850: 
1851: \bibitem{SD01}A. Santos and J. W. Dufty, ``Critical behavior of a heavy
1852: particle in a granular fluid,'' Phys. Rev. Lett. {\bf 86}, 4823
1853: (2001); ``Nonequilibrium phase transition for a heavy particle in
1854: a granular fluid,'' Phys. Rev. E {\bf 64}, 051305 (2001).
1855: 
1856: 
1857: \bibitem{G05}V. Garz\'o, ``Instabilities in a free granular fluid described
1858: by the Enskog equation,'' Phys. Rev. E {\bf 72}, 021106 (2005).
1859: 
1860: \bibitem{GD99a}V. Garz\'o and J. W. Dufty, ``Dense fluid transport for inelastic hard spheres,''
1861: Phys. Rev. E {\bf 59}, 5895 (1999).
1862: 
1863: \bibitem{misprints}Some misprints occur in the expressions given
1864: in Ref.\ \onlinecite{GD02} for these quantities. The expressions
1865: given here correct such results.
1866: 
1867: \bibitem{CC70bis}See Ref.\ \onlinecite{CC70}, Chap. 9, Eqs.\ (9.6,1)--(9.6,10).
1868: 
1869: 
1870: 
1871: 
1872: 
1873: \end{thebibliography}
1874: 
1875: 
1876: 
1877: 
1878: \end{document}
1879: