1: \documentclass[runningheads]{cl2emult}
2: \usepackage{psfig}
3: \usepackage{psfrag} % replace labels in figures
4: \usepackage{epsf}
5: \usepackage{makeidx} % allows index generation
6: \usepackage{graphicx} % standard LaTeX graphics tool
7: \usepackage{subeqnar} % subnumbers individual equations
8: \usepackage{multicol} % used for the two-column index
9: \usepackage{citesort}
10: \usepackage{amssymb}
11: \usepackage{amsmath}
12: \usepackage{latexsym}
13: \usepackage{cropmark} % cropmarks for pages without
14: \usepackage{lnp} % placeholder for figures
15: \makeindex % used for the subject index
16:
17: \newcommand{\euler}[1]{{\usefont{U}{eur}{m}{n}#1}}
18: \newcommand{\eulerbold}[1]{{\usefont{U}{eur}{b}{n}#1}}
19: \newcommand{\umu}{\mbox{\euler{\char22}}}
20: \newcommand{\umub}{\mbox{\eulerbold{\char22}}}
21: \begin{document}
22:
23: \frontmatter
24:
25: %\input{Book/Title}
26: %\cleardoublepage
27:
28: %\markboth{Preface}{Preface}
29: %\input{Book/Preface}
30: %\clearpage
31:
32: %\markboth{Contents}{Contents}
33: %\tableofcontents
34: \mainmatter
35: \setcounter{page}{100}
36:
37: %\addcontentsline{toc}{part}{Kinetic Theory and Hydrodynamics}
38: %\part*{Kinetic Theory\\[0.3cm] and\\[0.3cm] Hydrodynamics}
39: %\include{Ernst/Ernst} %FINISHED
40: %\include{Zippelius/Zippelius} %FINISHED (check \stackrel)
41: %\include{Brey/Brey} %FINISHED
42: %\include{Goldhirsch/Goldhirsch} %FINISHED
43: %\include{Brilliantov/PP2} %FINISHED
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: % \begin{document}
46:
47: \title*{{\small In: T. P\"oschel and S. Luding (eds.), {\em Granular Gases}, Lecture Notes in Physics Vol. 564, Springer (Berlin, 2000), p. 100}\vspace*{0.5cm}\\ Granular Gases with Impact-velocity Dependent Restitution Coefficient}
48: \label{BPpage}
49:
50: \toctitle{Granular Gases with impact-velocity dependent restitution coefficient}
51:
52: \titlerunning{Granular Gases with impact-velocity dependent restitution coefficient}
53:
54: \author{Nikolai~V.~Brilliantov\inst{1}$^,$\inst{2} \and Thorsten P\"oschel\inst{2}}
55:
56: \authorrunning{N.\,V.~Brilliantov \and T.~P\"oschel}
57: \tocauthor{N.\,V.~Brilliantov, T.~P\"oschel}
58:
59: \institute{Moscow State University, Physics Department, Moscow 119899, Russia.\\ e-mail:~nbrillia@physik.hu-berlin.de
60: \and Humboldt-Universit\"at, Institut f\"ur Physik,
61: Invalidenstr. 110, D-10115 Berlin, Germany. e-mail:~thorsten@physik.hu-berlin.de, http://summa.physik.hu-berlin.de/$\sim$thorsten}
62:
63: \maketitle % typesets the title of the contribution
64:
65: \begin{abstract}
66: We consider collisional models for granular particles and analyze the
67: conditions under which the restitution coefficient might be a constant.
68: We show that these conditions are not consistent with known collision
69: laws. From the generalization of the Hertz contact law for
70: viscoelastic particles we obtain the coefficient of normal restitution
71: $\epsilon$ as a function of the normal component of the impact velocity
72: $v_{\rm imp}$. Using $\epsilon(v_{\rm imp})$ we describe the
73: time evolution of temperature and of the velocity distribution function of
74: a granular gas in the homogeneous cooling regime,
75: where the particles collide according to the viscoelastic law. We show that for the studied
76: systems the simple scaling hypothesis for the velocity distribution
77: function is violated, i.e. that its evolution is not determined only by
78: the time dependence of the thermal velocity. We observe, that the deviation
79: from the Maxwellian distribution, which we quantify by the value of the second
80: coefficient of the Sonine polynomial expansion of the velocity distribution
81: function, does not depend on time
82: monotonously. At first stage of the evolution it increases on the mean-collision
83: time-scale up to a maximum value and then decays to zero at the second stage,
84: on the time scale corresponding to the evolution of the granular gas
85: temperature. For granular gas in the homogeneous cooling regime we also
86: evaluate the time-dependent self-diffusion coefficient of
87: granular particles. We analyze the time dependence of the mean-square displacement
88: and discuss its impact on clustering. Finally, we discuss
89: the problem of the relevant internal time for the systems of interest.
90: \end{abstract}
91:
92: \section{Introduction}
93:
94: Granular gases, i.e. systems of inelastically colliding particles, are
95: widely spread in nature. They may be exemplified by industrial dust or
96: interterrestrial dust; the behavior of matter in planetary rings is also
97: described in terms of the granular gas dynamics. As compared with common
98: molecular gases, the steady removal of kinetic energy in these systems
99: due to dissipative collisions causes a variety of nonequilibrium phenomena,
100: which have been very intensively studied
101: %experimentally (e.g. \cite{BSHPexper}), numerically
102: (e.g.~\cite{BSHPnumer,McNamara96,GZ93,NoijeErnst:97,EsipovPoeschel:97,NoijeErnstRing,GoldshteinShapiro95,EBEuro,NEBO97,BrilliantovPoeschel:1998d,HuthmannZippelius:97}). In most of these studies the coefficient of
103: restitution, which characterizes the energy lost in the collisions, was assumed
104: to be constant. This approximation, although providing a considerable
105: simplification, and allowing to understand the main effects in granular gas dynamics, is not always justified (see also the paper by Thornton in this book~\cite{ThorntonHere}). Moreover, sometimes it
106: occurs to be too crude to describe even qualitatively the features of
107: granular gases. Here we discuss the properties of granular gases
108: consisting of viscoelastically colliding particles which implies an impact-velocity dependent restitution coefficient. The results are compared with results for gases consisting of particles which interact via a constant restitution coefficient and we see that the natural assumption of viscoelasticity leads to {\em qualitative} modifications of the gas properties.
109:
110: The following problems will be addressed:
111: \begin{enumerate}
112: \item[$\bullet$]Why does the restitution coefficient $\epsilon$ depend on the
113: impact velocity $v_{\rm imp}$?
114:
115: \item[$\bullet$]How does it depend on the impact velocity?
116:
117: \item[$\bullet$]What are the consequences of the dependence of $\epsilon$
118: on $v_{\rm imp}$ on the collective behavior of particles in granular
119: gases? In particular how does $\epsilon=\epsilon\left(v_{\rm imp}\right)$ influence:
120: \begin{itemize}
121: \item the evolution of temperature with time?
122: \item the evolution of the velocity distribution function with time?
123: \item the self-diffusion in granular gases?
124: \end{itemize}
125: \end{enumerate}
126:
127: In what follows we will show that the dependence of the restitution
128: coefficient on the impact velocity is a very basic property of
129: dissipative particle collisions, whereas the assumption of a constant restitution coefficient
130: for the collision of three-dimensional spheres may lead to a physically incorrect dependence of the dissipative force on
131: the compression rate of the colliding particles. From the Hertz
132: collision law and the general relation between the elastic and dissipative
133: forces we deduce the dependence of the restitution coefficient on the
134: impact velocity, which follows purely from scaling considerations.
135: We also give the corresponding relation obtained from rigorous theory.
136: Using the dependence $\epsilon(v_{\rm imp})$ we derive the time
137: dependence of the temperature, the time-evolution
138: of the velocity distribution function and describe self-diffusion in granular gases in the homogeneous cooling regime.
139:
140: \section{Dependence of the restitution coefficient on the impact velocity}
141:
142: The collision of two particles may be characterized by the compression
143: $\xi$ and by the compression rate $\dot{\xi}$, as shown on Fig.~\ref{fig:sketch}. The
144: compression gives rise to the elastic force $F_{\rm el}(\xi)$, while the
145: dissipative force $F_{\rm diss}(\xi,\dot{\xi})$ appears due to the compression
146: rate.
147: \begin{figure}[htbp]
148: \centerline{\psfig{figure=coll.eps,width=5.5cm}}
149: \caption{Sketches of two colliding spheres. The compression $\xi$ is equal to $2R-|\vec{r}_1-\vec{r}_2|$, with $\vec{r}_{1/2}$ being the particles positions.
150: The compression rate is $\dot{\xi}=v_1-v_2$. For simplicity the head on
151: collision of identical spheres is shown.}
152: \label{fig:sketch}
153: \end{figure}
154:
155: If the compression and the compression rate are not very large, one can
156: assume the dependence of the elastic and dissipative
157: force on $\xi$ and $\dot{\xi}$
158: \begin{eqnarray}
159: &&F_{\rm el}(\xi) \sim \xi^{\alpha} \\
160: &&F_{\rm diss}(\xi,\dot{\xi}) \sim \dot{\xi}^{\beta } \xi^{\gamma }\,.
161: \end{eqnarray}
162: The dimension analysis yields the following functional
163: form for the dependence of the restitution coefficient on the impact velocity
164: \cite{rospap}:
165:
166: \begin{equation}
167: \epsilon(v_{\rm imp})=
168: \epsilon \left( v_{\rm imp}^{\frac{ 2 (\gamma -\alpha)}{1+\alpha}+\beta} \right)
169: \end{equation}
170: Therefore, the condition for a constant restitution coefficient imposes the
171: relation between the exponents $\alpha$, $\beta$ and $\gamma$~\cite{Taguchi93,Luding98}:
172:
173: \begin{equation}
174: 2\left(\gamma -\alpha\right)+\beta\,\left ( 1+\alpha\right ) = 0\,.
175: \label{condition}
176: \end{equation}
177: For compressions which do not exceed the plasticity
178: threshold, the particle's material behaves as a viscoelastic medium. Then
179: it may be generally shown~\cite{BSHP1,BSHP2,remark2} that the
180: relation
181:
182: \begin{equation}
183: \label{eldiss}
184: F_{\rm diss}= A\, \dot\xi \frac{\partial}{\partial{\xi}} F_{\rm el}(\xi) \,
185: \end{equation}
186: between the elastic and dissipative force holds, independently
187: on the shape of the bodies in contact, provided three conditions
188: are met~\cite{BrilPoshprep}:
189:
190: \begin{enumerate}
191: \label{item:condistions}
192: \item[(i)] The elastic components of the stress tensor
193: $\sigma_{\rm el}^{ik}$ depend linearly on
194: the components of the deformation tensor $u_{ik}$~\cite{LandauLifshits}.
195: \item [(ii)] The dissipative components of the stress tensor
196: $\sigma_{\rm diss}^{ik}$ depend
197: linearly on the components of the deformation rate tensor $\dot{u}_{ik}$~\cite{LandauLifshits}.
198: \item[(iii)] The conditions of quasistatic motion are provided, i.e.
199: $\dot{\xi} \ll c$, $\tau_{\rm vis} \ll \tau_c$ \cite{BSHP1,BSHP2}
200: (here $c$ is the
201: speed of sound in the material of particles and $\tau_{\rm vis}$ is the
202: relaxation time of viscous processes in its bulk).
203:
204: \end{enumerate}
205: The constant $A$ in Eq.~(\ref{eldiss}) reads~\cite{BSHP1,BSHP2}
206: %??? check Spahn
207: \begin{equation}
208: \label{A}
209: A=\frac13\, \frac{(3\eta_2-\eta_1)^2}{(3\eta_2+2\eta_1)}
210: \left[\frac{(1-\nu^2)(1-2\nu)}{Y \nu^2}\right]\,.
211: \end{equation}
212: where $Y$ and $\nu$ are respectively the Young modulus and the
213: Poisson ratio of the particle material and the viscous constants
214: $\eta_1$, $\eta_2$ relate (linearly) the dissipative stress
215: tensor $\sigma_{\rm diss}^{ik}$ to the deformation rate tensor
216: $\dot{u}_{ik}$ \cite{BSHP1,BSHP2,LandauLifshits}.
217:
218: \ From Eq.~(\ref{eldiss}) follows that
219: \begin{equation}
220: \beta=1 \qquad \gamma = \alpha -1 \,.
221: \label{1conditionres}
222: \end{equation}
223:
224:
225: Consider now a collision of three-dimensional spherical particles of
226: radii $R_1$ and $R_2$. The Hertz contact contact law gives for the
227: elastic force~\cite{Hertz.Brill}
228: \begin{equation}
229: \label{Hertzlaw}
230: F_{\rm el} =\rho \, \xi^{3/2} \,,
231: \qquad \rho \equiv \frac{2Y}{3(1-\nu^2)}\sqrt{R^{\rm eff}}\,,
232: \end{equation}
233: where $R^{\rm eff} \equiv R_1R_2/(R_1+R_2)$. With the set of exponents, $\alpha =3/2$, $\beta=1$ and $\gamma = 1/2$,
234: which generally follows from the basic laws of the viscoelastic collision,
235: the condition for the constant restitution coefficient, Eq.~(\ref{condition}), is
236: obviously not satisfied. For spherical particles the restitution coefficient could be constant
237: only for $\gamma = 1/4$; this, however, is not consistent with the collision laws.
238: Instead one obtains the functional dependence
239: %\vspace*{-0.3cm}
240: \begin{equation}
241: \epsilon=\epsilon \left( v_{\rm imp}^{1/5} \right) \,.
242: \label{eps15}
243: \end{equation}
244: Note that this conclusion comes from the general analysis of viscoelastic
245: collisions with no other assumptions needed. Therefore, the dependence of the restitution coefficient on the impact velocity, Eq.~(\ref{eps15}), is a natural property, provided the assumption on viscoelasticity holds true which is the case in a wide range of impact velocities (see discussion on page \pageref{item:condistions}). We want to mention that the functional dependence Eq.~(\ref{eps15}) was already given in \cite{Kuwabara} using heuristic arguments.
246:
247:
248: Rigorous calculations~\cite{TomThor} yield for the dependence of the
249: restitution coefficient on the impact velocity:
250: \begin{equation}
251: \label{epsC1C2}
252: \epsilon=1-C_1 A \kappa^{2/5}v_{\rm imp}^{1/5}
253: +C_2A^2 \kappa^{4/5}v_{\rm imp}^{2/5} \mp \cdots
254: \end{equation}
255: with
256: \begin{equation}
257: \kappa = \left( \frac32 \right)^{3/2}
258: \frac{Y\sqrt{R^{\rm eff}}}{m^{\rm eff} (1-\nu^2)}
259: \end{equation}
260: where $m^{\rm eff}=m_1m_2/(m_1+m_2)$ ($m_{1/2}$ are the masses of the colliding
261: particles). Numerical values for the constants $C_1$ and
262: $C_2$ obtained in~\cite{TomThor} may be also written in a more convenient
263: form~\cite{rospap}:
264:
265: \begin{equation}
266: \label{C1C2}
267: C_1= \frac{ \Gamma(3/5)\sqrt{\pi}}{2^{1/5}5^{2/5} \Gamma(21/10)} =
268: 1.15344, \qquad C_2=\frac{3}{5}C_1^2\, .
269: \end{equation}
270: Although the next-order coefficients of the above expansion $C_3=-0.483582$,
271: $C_4=0.285279$, are now available~\cite{rospap}, we assume that
272: the dissipative constant $A$ is small enough to ignore these high-order
273: terms. (For large $A$ a very accurate Pad\'e approximation for
274: $\epsilon(v_{\rm imp})$ has been proposed recently~\cite{rospap}).
275:
276: \section{Time-evolution of temperature and of the velocity distribution function}
277:
278: We consider a granular gas composed of $N$ identical particles
279: confined in a volume $\Omega$. The particles are assumed to be smooth
280: spheres, so that the collision properties are determined by the normal component of the relative
281: motion only. The gas is supposed to be dilute enough so that one can
282: assume binary collisions (i.e. neglect multiple collisions) and ignore the collision duration as compared with the mean free time in between successive collisions.
283: We assume that the initial velocities of the
284: particles (more precisely the temperature, which we define below) are not
285: very large to assure viscoelastic properties of the collisions, i.e. to avoid plastic deformations and fragmentation.
286: The final velocities are assumed not to be very small which allows to neglect surface forces as adhesion and others. Under these restrictions one can apply the
287: viscoelastic collision model. Furthermore, we assume that dissipation is not
288: large, so that the second-order expansion (\ref{epsC1C2}) for
289: $\epsilon(v_{\rm imp})$ describes the collisions accurately. We analyze the
290: granular gas in the regime of homogeneous cooling, i.e. in the
291: pre-clustering regime, when the gas is homogeneously distributed in space.
292:
293: The impact velocity $v_{\rm imp}$ of colliding smooth spheres,
294: which determines the value of the restitution coefficient according to
295: Eq.~(\ref{epsC1C2}), is given by the normal component of the relative velocity
296: \begin{equation}
297: v_{\rm imp}=\left|\vec{v}_{12} \cdot \vec{e} \right|~~~~\mbox{with}~~~~ \vec{v}_{12}=\vec{v}_1-\vec{v}_2\,.
298: \end{equation}
299: The unit vector $\vec{e} = \vec{r}_{12}/ |\vec{r}_{12}|$ gives the
300: direction of the intercenter vector $\vec{r}_{12}=\vec{r}_{1}-\vec{r}_{2}$ at
301: the instant of the collision.
302:
303: The evolution of the granular gas proceeds by elementary collision events in
304: which the pre-collisional velocities of colliding particles $\vec{v}_1$,
305: $\vec{v}_2$, are converted into after-collisional ones, $\vec{v}_1^*$, $\vec{v}_2^*$,
306: according to the rules
307: \begin{equation}
308: \label{directcoll}
309: \begin{split}
310: %\label{v1v2v1*v2*}
311: \vec{v}_1^*&=\vec{v}_1-
312: \frac12 \left[1+\epsilon(|\vec{v}_{12} \cdot \vec{e} | )\right]
313: (\vec{v}_{12} \cdot \vec{e})\vec{e} \\
314: \vec{v}_2^*&=\vec{v}_2+ \frac12
315: \left[1+\epsilon(|\vec{v}_{12} \cdot \vec{e} | )\right]
316: (\vec{v}_{12} \cdot \vec{e})\vec{e} \,,
317: \end{split}
318: \end{equation}
319: where $\epsilon$ depends on the impact velocity
320: $v_{\rm imp}=|\vec{v}_{12} \cdot \vec{e} |$. Due to the {\em direct} collision
321: (\ref{directcoll}) the population in the velocity phase-space near the points
322: $\vec{v}_1$, $\vec{v}_2$ decreases, while near the points $\vec{v}_1^*$,
323: $\vec{v}_2^*$ it increases. The decrease of the population near
324: $\vec{v}_1$, $\vec{v}_2$ caused by the direct collision, is (partly)
325: counterbalanced by its increase in the {\em inverse} collision, where
326: the after-collisional velocities are $\vec{v}_1$, $\vec{v}_2$ with
327: the pre-collisional ones $\vec{v}_1^{**}$, $\vec{v}_2^{**}$. The
328: rules for the inverse collision read
329: \begin{equation}
330: \label{inversecoll}
331: \begin{split}
332: %\label{v1v2v1*v2*}
333: \vec{v}_1&=\vec{v}_1^{**}-
334: \frac12 \left[ 1+\epsilon(|\vec{v}_{12}^{**} \cdot \vec{e} |) \right]
335: \left( \vec{v}_{12}^{**} \cdot \vec{e} \right)\vec{e} \\
336: \vec{v}_2&=\vec{v}_2^{**}+
337: \frac12 \left[ 1+\epsilon(|\vec{v}_{12}^{**} \cdot \vec{e} |) \right]
338: \left( \vec{v}_{12}^{**} \cdot \vec{e} \right)\vec{e}\,.
339: \end{split}
340: \end{equation}
341: Note that in contrast to the case of $\epsilon ={\rm const}$, the
342: restitution coefficients in the inverse and in the direct collisions
343: are different.
344:
345: The Enskog-Boltzmann equation describes the evolution of the population of particles in the phase space on the mean-field level.
346: The evolution is characterized by the distribution
347: function $f(\vec{r},\vec{v},t)$, which for the force-free case does not
348: depend on $\vec{r}$ and obeys the equation~\cite{NoijeErnst:97,resibua}
349: \begin{multline}
350: \label{collint}
351: \frac{\partial}{\partial t}f\left(\vec{v}_1,t\right) =
352: g_2(\sigma)\sigma^2 \int d \vec{v}_2 \int d\vec{e}
353: \Theta(-\vec{v}_{12} \cdot \vec{e}) |\vec{v}_{12} \cdot \vec{e}|\\
354: \times \left\{\chi f(\vec{v}_1^{**},t)f(\vec{v}_2^{**},t)-
355: f(\vec{v}_1,t)f(\vec{v}_2,t) \right\} \equiv g_2(\sigma)I(f,f)
356: \end{multline}
357: where $\sigma$ is the diameter of the particles.
358: The contact value of the pair distribution function~\cite{CarnahanStarling}
359: \begin{equation}
360: g_2(\sigma)=(2-\eta)/2(1-\eta)^3\,,
361: \end{equation}
362: accounts on the mean-field level for the increasing frequency of collisions due to
363: excluded volume effects with $\eta=\frac16\, \pi n \sigma^3$ being the
364: volume fraction.
365:
366: The first term in the curled brackets in the right-hand side of
367: Eq.~(\ref{collint}) refers to the ``gain'' term for the population in the
368: phase-space near the point $\vec{v}_1$, while the second one is the
369: ``loss'' term. The Heaviside function $\Theta(-\vec{v}_{12} \cdot \vec{e})$ discriminates approaching particles (which do collide) from separating particles (which do not collide), and
370: $|\vec{v}_{12} \cdot \vec{e}|$ gives the length of the collision cylinder.
371: Integration in Eq.~(\ref{collint}) is performed over all
372: velocities $\vec{v}_2$ and interparticle vectors $\vec{e}$ in the direct
373: collision. Equation~(\ref{collint}) accounts also
374: for the inverse collisions via the factor $\chi$, which appears due to
375: the Jacobian of the transformation
376: $\vec{v}_1^{**},\vec{v}_2^{**} \to \vec{v}_1,\vec{v}_2$, and due to the
377: difference between the lengths of the collision cylinders of the
378: direct and the inverse collision:
379: \begin{equation}
380: \chi=
381: \frac{{\cal D}(\vec{v}_1^{**},\vec{v}_2^{**} )}{{\cal D}(\vec{v}_1,\vec{v}_2)}
382: \,
383: \frac{|\vec{v}_{12}^{**} \cdot \vec{e}|}{|\vec{v}_{12} \cdot \vec{e}|}\,.
384: \end{equation}
385:
386: For constant restitution coefficient the factor $\chi$ is a
387: constant
388: \begin{equation}
389: \chi = \frac{1}{\epsilon^2} = {\rm const} \, ,
390: \end{equation}
391: while for $\epsilon=\epsilon\left(v_{\rm imp}\right)$, as given in Eq.~(\ref{epsC1C2}), it reads~\cite{BPMaxw}
392: \begin{equation}
393: \label{chi}
394: \chi = 1 + \frac{11}{5}C_1 A \kappa^{2/5} |\vec{v}_{12} \cdot \vec{e}|^{1/5}
395: +\frac{66}{25}C_1^2 A^2 \kappa^{4/5} |\vec{v}_{12} \cdot \vec{e}|^{2/5}
396: +\cdots
397: \end{equation}
398: \ From Eq.~(\ref{chi}) it follows that $\chi =\chi(|\vec{v}_{12} \cdot \vec{e}|)$.
399: Since the average velocity in granular gases changes with time, such a
400: dependence of $\chi$ means, as we will show below, that $\chi$ and, therefore, the velocity distribution function itself depend explicitly on time.
401: The time dependence of $\chi$ changes drastically the properties of the collision integral
402: and destroys the simple scaling form of the velocity distribution function,
403: which holds for the case of the constant restitution coefficient
404: (e.g.~\cite{EsipovPoeschel:97,NoijeErnst:97}).
405:
406: Nevertheless, some important properties of the collision integral are preserved. Namely, it may be shown that the relation
407: \begin{eqnarray}
408: \label{deraver}
409: &&\!\!\!\!\!\!\frac{d}{dt} \left< \psi(t) \right> =\int d\vec{v}_1 \psi (\vec{v}_1)
410: \frac{\partial}{\partial t} f(\vec{v}_1, t) = g_2(\sigma)
411: \int d\vec{v}_1 \psi (\vec{v}_1) I(f,f) =\\
412: &&\!\!\!\!\!\!\frac{g_2(\sigma)\sigma^2}{2} \int d\vec{v}_1d\vec{v}_2 \int d\vec{e}\,
413: \Theta(-\vec{v}_{12} \cdot \vec{e}) |\vec{v}_{12} \cdot \vec{e}|
414: f(\vec{v}_1, t)f(\vec{v}_2, t) \Delta \left[ \psi( \vec{v}_1)+
415: \psi( \vec{v}_2) \right] \nonumber
416: \end{eqnarray}
417: holds true, where
418: \begin{equation}
419: \left< \psi(t) \right> \equiv \int d\vec{v} \psi (\vec{v}) f(\vec{v}, t)
420: \end{equation}
421: is the average of some function $\psi (\vec{v})$, and
422: \begin{equation}
423: \Delta \psi(\vec{v}_i) \equiv \left[\psi(\vec{v}_i^*)-\psi(\vec{v}_i) \right]
424: \end{equation}
425: denotes the change of $\psi( \vec{v}_i)$ in a direct collision.
426:
427: Now we introduce the temperature of the three-dimensional granular gas,
428: \begin{equation}
429: \label{deftemp1}
430: \frac{3}{2} n T(t)=\int d \vec{v} \frac{m v^2}{2} f(\vec{v},t)\,,
431: \end{equation}
432: where $n$ is the number density of granular particles
433: ($n=N/ \Omega$), and the characteristic velocity $v_0^2(t)$ is related to temperature via
434: \begin{equation}
435: \label{deftemp}
436: T(t)=\frac12 mv_0^2(t)\,.
437: \end{equation}
438: First we try the scaling ansatz
439: \begin{equation}
440: \label{veldis}
441: f(\vec{v}, t)=\frac{n}{v_0^3(t)} \tilde{f}(\vec{c})
442: \end{equation}
443: where
444: $\vec{c} \equiv \vec{v}/v_0(t)$ and following
445: \cite{NoijeErnst:97,GoldshteinShapiro95} assume that deviations from the
446: Maxwellian distribution are not large, so that $\tilde{f}(\vec{c})$ may
447: be expanded into a convergent series with the leading term being the
448: Maxwellian distribution $\phi(c) \equiv \pi^{-3/2} \exp(-c^2)$.
449: It is convenient to use the Sonine polynomials expansion
450: \cite{NoijeErnst:97,GoldshteinShapiro95}
451: \begin{equation}
452: \label{Soninexp}
453: \tilde{f}(\vec{c})=\phi(c) \left\{1 + \sum_{p=1}^{\infty} a_p S_p\left(c^2\right) \right\}
454: \, .
455: \end{equation}
456: These polynomials are orthogonal, i.e.
457: \begin{equation}
458: \label{Soninortog}
459: \int d \vec{c}\, \phi (c) S_p(c^2)S_{p^{\prime}}\!\left(c^2\right)
460: = \delta_{pp^{\prime}}{\cal N}_p\,,
461: \end{equation}
462: where $\delta_{pp^{\prime}}$ is the Kronecker delta and ${\cal N}_p$ is the
463: normalization constant. The first few polynomials read
464: \begin{equation}
465: \label{Soninfewfirst}
466: \begin{split}
467: S_0(x)&=1 \\
468: S_1(x)&=-x^2 +\frac32 \\
469: S_2(x)&=\frac{x^2}{2}-\frac{5x}{2}+\frac{15}{8}\,.
470: \end{split}
471: \end{equation}
472: Writing the Enskog-Boltzmann equation in terms of the scaling variable
473: $\vec{c}_1$, one observes that the factor $\chi$ may not be expressed
474: only in terms of the scaling variable, but it depends also on the
475: characteristic velocity $v_0(t)$, and thus depends on time. Therefore, the
476: collision integral also occurs to be time-dependent. As a result, it is not
477: possible to reduce
478: the Enskog-Boltzmann equation to a pair of equations,
479: one for the time evolution of the temperature and another for the
480: time-independent scaling function, whereas for $\epsilon=$const. the Boltzmann-Enskog equation is separable, e.g.~\cite{NoijeErnst:97,GoldshteinShapiro95,EsipovPoeschel:97}. Formally adopting the approach of Refs.~\cite{NoijeErnst:97,GoldshteinShapiro95} for $\epsilon=$const., one would obtain
481: time-dependent coefficients $a_p$ of the Sonine polynomials expansion.
482: This means that the simple scaling ansatz (\ref{veldis}) is violated
483: for the case of the impact-velocity dependent restitution coefficient.
484:
485: Thus, it seems natural to write the distribution function in the following
486: general form
487:
488: \begin{equation}
489: \label{genscal}
490: f(\vec{v}, t)=\frac{n}{v_0^3(t)}\tilde{f}(\vec{c}, t)
491: \end{equation}
492: with
493: \begin{equation}
494: \label{genSoninexp}
495: \tilde{f}(\vec{c})
496: =\phi(c) \left\{1 + \sum_{p=1}^{\infty} a_p(t) S_p(c^2) \right\}
497: \end{equation}
498: and find then equations for the {\em time-dependent} coefficients $a_p(t)$.
499: Substituting (\ref{genscal}) into the Boltzmann equation (\ref{collint})
500: we obtain
501: \begin{equation}
502: \label{geneqveldis}
503: \frac{\mu_2}{3}
504: \left(3 + c_1 \frac{\partial}{\partial c_1} \right) \tilde{f}(\vec{c}, t) +
505: B^{-1} \frac{\partial}{\partial t} \tilde{f}(\vec{c}, t) =
506: \tilde{I}\left( \tilde{f}, \tilde{f} \right)
507: \end{equation}
508: with
509: \begin{equation}
510: B=B(t) \equiv v_0(t) g_2(\sigma) \sigma^2 n\,.
511: \end{equation}
512: We define the
513: dimensionless collision integral:
514: \begin{multline}
515: \label{dimlcolint}
516: \tilde{I}\left( \tilde{f}, \tilde{f} \right)=\\
517: \int d \vec{c}_2 \int d\vec{e}
518: \Theta(-\vec{c}_{12} \cdot \vec{e}) |\vec{c}_{12} \cdot \vec{e}|
519: \left\{\tilde{\chi} \tilde{f}(\vec{c}_1^{**},t) \tilde{f}(\vec{c}_2^{**},t)-
520: \tilde{f}(\vec{c}_1,t)\tilde{f}(\vec{c}_2,t) \right\}
521: \end{multline}
522: with the reduced factor $\tilde{\chi}$
523: \begin{equation}
524: \label{chiscal}
525: \tilde{\chi} = 1 +
526: \frac{11}{5}C_1 \delta^{\prime} |\vec{c}_{12} \cdot \vec{e}|^{1/5}
527: +\frac{66}{25}C_1^2 \delta^{\prime \, 2} |\vec{c}_{12} \cdot \vec{e}|^{2/5}
528: +\cdots
529: \end{equation}
530: which depends now on time via a quantity
531: \begin{equation}
532: \label{deltaprime}
533: \delta^{\, \prime} (t) \equiv A \kappa^{2/5} \left[2T(t)\right]^{1/10}
534: \equiv \delta \left[2T(t)/T_0 \right]^{1/10}\,.
535: \end{equation}
536: Here $\delta \equiv A \kappa^{2/5}[T_0]^{1/10}$, $T_0$ is the initial
537: temperature, and for simplicity we assume the unit mass, $m=1$.
538: We also define the moments of the dimensionless collision integral
539: \begin{equation}
540: \label{mup}
541: \mu_p \equiv - \int d \vec{c}_1 c_1^p
542: \tilde{I}\left( \tilde{f}, \tilde{f} \right)\ ,
543: \end{equation}
544: so that the second moment describes the rate of the temperature change:
545: \begin{equation}
546: \label{dTdt}
547: \frac{dT}{dt} =-\frac23 BT\mu_2\,.
548: \end{equation}
549: Equation~(\ref{dTdt}) follows from the definitions of the temperature and of the
550: moment $\mu_2$. Note that these moments depend on time, in contrast to the
551: case of the constant restitution coefficient, where these moments are time-independent~\cite{NoijeErnst:97}.
552:
553: Multiplying both sides of Eq.~(\ref{geneqveldis}) with $c_1^p$ and integrating
554: over $d \vec{c}_1$, we obtain
555: \begin{equation}
556: \label{momeq}
557: \frac{\mu_2}{3} p \left< c^p \right> -B^{-1}\sum_{k=1}^{\infty}
558: \dot{a}_k \nu_{kp} = \mu_p
559: \end{equation}
560: where integration by parts has been performed and we define
561: \begin{eqnarray}
562: \label{nukp}
563: \nu_{kp} &\equiv& \int \phi(c) c^p S_k(c^2) d\vec{c} \\
564: \left< c^p \right> &\equiv& \int c^p \tilde{f}(\vec{c}, t) d\vec{c}\,.
565: \label{cpdef}
566: \end{eqnarray}
567: The calculation of $\nu_{kp}$ is straightforward; the first few of these read:
568: $\nu_{22}=0$, $\nu_{24}=\frac{15}{4}$. The odd moments
569: $\left< c^{2n+1} \right> $ vanish, while the even ones
570: $\left< c^{2n} \right> $ may be expressed in terms of $a_k$ with
571: $0 \leq k \leq n$, namely,
572: $\left< c^2 \right> =\frac32 - \frac32 a_1$. On the other hand, from the
573: definition of temperature and of the thermal velocity in Eqs.~(\ref{deftemp}) and (\ref{deftemp1})
574: follows that $\left< c^2 \right> = \frac32$ and thus, $a_1=0$.
575: Similar considerations yield
576: $\left< c^4 \right> = \frac{15}{4}\left( 1 + a_2 \right)$.
577: The moments $\mu_p$ may be expressed in terms of coefficients
578: $a_2, a_3, \cdots$ too; therefore, the system Eq.~(\ref{momeq}) is an infinite (but closed) set of equations for these
579: coefficients.
580:
581: It is not possible to get a general solution of the problem. However, since
582: the dissipative parameter $\delta$ is supposed to be small, the deviations from the
583: Maxwellian distribution are presumably small too. Thus, we assume that one can neglect all
584: high-order terms with $p>2$ in the expansion (\ref{genSoninexp}). Then Eq.~(\ref{momeq})
585: is an equation for the coefficient $a_2$. For $p=2$ Eq.~(\ref{momeq}) converts into an identity,
586: since $\left< c^2 \right>=\frac32$, $a_1=0$, $\nu_{22}=0$ and
587: $\nu_{24}=\frac{15}{4}$. For $p=4$ we obtain
588: \begin{equation}
589: \label{eqa2}
590: \dot{a}_2-\frac43\, B\mu_2 \left(1+a_2 \right)+\frac{4}{15}B\mu_4 =0\,.
591: \end{equation}
592: In Eq.~(\ref{eqa2}) $B$
593: depends on time as
594: \begin{equation}
595: B(t)=(8 \pi)^{-1/2} \tau_c(0)^{-1}[T(t)/T_0]^{1/2}\,,
596: \end{equation}
597: where $\tau_c(0)$ is related to the initial mean-collision time,
598: \begin{equation}
599: \tau_c(0)^{-1}=4 \pi^{1/2}g_2(\sigma) \sigma^2 n T_0^{1/2}\,.
600: \end{equation}
601: The time evolution of the temperature is determined by Eq.~(\ref{dTdt}), i.e. by the
602: time dependence of $\mu_2$.
603:
604: The time-dependent coefficients $\mu_p(t)$ may be expressed in terms of $a_2$ owing to
605: their definition Eq.~(\ref{mup}) and the approximation $\tilde{f}= \phi (c) [ 1+a_2(t)S_2(c^2)]$.
606: We finally obtain:
607: \begin{multline}
608: \label{mupa2}
609: \mu_p=-\frac12 \int d\vec{c}_1\int d\vec{c}_2 \int d\vec{e}
610: \Theta(-\vec{c}_{12} \cdot \vec{e}) |\vec{c}_{12} \cdot \vec{e}| \phi(c_1) \phi(c_2)\\ \times
611: \left\{1+a_2\left[S_2(c_1^2)+S_2(c_2^2) \right] + a_2^2\,S_2(c_1^2)S_2(c_2^2) \right\}
612: \Delta (c_1^p+c_2^p)
613: \end{multline}
614: with the definition of $\Delta (c_1^p+c_2^p)$ given above. Calculations
615: performed up to the second order in terms of the dissipative parameter $\delta$
616: yield~\cite{BPMaxw}:
617: \begin{equation}
618: \label{mu2A}
619: \mu_2= \sum_{k=0}^{2} \sum_{n=0}^2 {\cal A}_{kn} \delta^{\, \prime \,k } a_2^n
620: \end{equation}
621: where
622: \begin{align}
623: \label{A1A6}
624: {\cal A}_{00}&=0; \quad &{\cal A}_{01}&=0; \quad &{\cal A}_{02}&=0 \nonumber\\
625: {\cal A}_{10}&=\omega_0; \quad &{\cal A}_{11}&=\frac{6}{25}\omega_0; \quad &{\cal A}_{12}&=\frac{21}{2500} \omega_0 \\
626: {\cal A}_{20}&=-\omega_1; \quad & {\cal A}_{21}&=-\frac{119}{400}\omega_1; \, \quad &{\cal A}_{22}&=-\frac{4641}{640000}\omega_1\nonumber
627: \end{align}
628: with
629: \begin{eqnarray}
630: \omega_0 &\equiv& 2 \sqrt{2 \pi} 2^{1/10} \Gamma \left (\frac{21}{10} \right)C_1=6.48562 \ldots\\
631: \omega_1 &\equiv& \sqrt{2 \pi} 2^{1/5} \Gamma \left (\frac{16}{5} \right)C_1^2=9.28569 \ldots
632: \end{eqnarray}
633: Similarly
634: \begin{equation}
635: \label{mu4A}
636: \mu_4= \sum_{k=0}^{2} \sum_{n=0}^2 {\cal B}_{kn} \delta^{\, \prime \,k } a_2^n
637: \end{equation}
638: with
639:
640: \begin{align}
641: %\label{B1B3}
642: {\cal B}_{00}&=0; \quad &{\cal B}_{01}&=4\sqrt{2 \pi}; \quad &{\cal B}_{02}&=\frac{1}{8}\sqrt{2 \pi} \nonumber \\
643: \label{B4B6}
644: {\cal B}_{10}&= \frac{56}{10}\omega_0; \quad &{\cal B}_{11}&=\frac{1806}{250}\omega_0; \quad &{\cal B}_{12}&=\frac{567}{12500}\omega_0 \\
645: %\label{B7B9}
646: {\cal B}_{20}&= -\frac{77}{10}\omega_1;\quad &{\cal B}_{21}&=-\frac{149054}{13750}\omega_1;\quad &{\cal B}_{22}&=-\frac{348424}{5500000}\omega_1\nonumber
647: \end{align}
648:
649: Thus, Eqs.~(\ref{dTdt}) and (\ref{eqa2}), together with Eqs.~(\ref{mu2A}) and
650: (\ref{mu4A}) form a closed set to find the time evolution of the temperature and
651: the coefficient $a_2$. We want to stress an important difference for the time evolution
652: of temperature for the case of the impact-velocity dependent restitution coefficient,
653: as compared to that of a constant restitution coefficient.
654: In the former case it is coupled to the time evolution of the coefficient $a_2$, while
655: in the latter case there is no such coupling since $a_2={\rm const}$.
656: This coupling may lead in to a rather peculiar time-dependence of
657: the temperature.
658:
659: Introducing the reduced temperature $u(t) \equiv T(t)/T_0$ we recast the set of equations
660: (\ref{dTdt}) and (\ref{eqa2}) into the following form:
661: \begin{multline}
662: \label{genseteq1}
663: \dot{u}+\tau_0^{-1}u^{8/5}\left( \frac53 +\frac25 a_2+\frac{7}{500}a_2^2 \right)\\
664: -\tau_0^{-1}q_1 \delta \, u^{17/10}
665: \left( \frac53 +\frac{119}{240}a_2 +\frac{1547}{128000}a_2^2 \right) =0
666: \end{multline}
667: \begin{equation}
668: \label{genseteq2}
669: \dot{a}_2-r_0u^{1/2}\mu_2 \left(1 +a_2 \right) + \frac15 r_0u^{1/2}\mu_4=0\,,
670: \end{equation}
671: where we introduce the characteristic time
672: \begin{equation}
673: \label{tau0}
674: \tau_0^{-1}=\frac{16}{5} q_0 \delta \cdot \tau_c(0)^{-1}
675: \end{equation}
676: with
677: \begin{eqnarray}
678: q_0&=&2^{1/5}\Gamma(21/10)C_1/8=5^{-2/5}\sqrt{\pi}\Gamma(3/5)/8=0.173318\ldots \\
679: r_0 &\equiv& \frac{2}{3 \sqrt{2 \pi}} \tau_c(0)^{-1}\\
680: q_1 &\equiv& 2^{1/10} (\omega_1/\omega_0) =1.53445 \ldots
681: \end{eqnarray}
682: As shown below the characteristic time $\tau_0$ describes the time evolution of the temperature.
683: To obtain these equations we use the
684: expressions for $\mu_2(t)$, $B(t)$, and for the coefficients ${\cal A}_{nk}$. Note that the
685: characteristic time $\tau_0$ is $\delta^{-1} \gg 1$ times larger than the collision
686: time $ \sim \tau_c(0)$.
687:
688: We will find the solution to these equations as expansions in terms of the small
689: dissipative parameter $\delta$ ($\delta^{\, \prime}(t) = \delta \cdot 2^{1/10}u^{1/10}(t)$):
690: \begin{eqnarray}
691: \label{expudel}
692: &&u=u_0+ \delta \cdot u_1 +\delta^2 \cdot u_2 +\cdots \\
693: \label{expadel}
694: &&a_2=a_{20}+\delta \cdot a_{21}+\delta^2 \cdot a_{22} +\cdots
695: \end{eqnarray}
696: Substituting Eqs.~(\ref{mu2A},\ref{mu4A},\ref{expudel},\ref{expadel}) into
697: Eqs.~(\ref{genseteq1},\ref{genseteq2}), one can solve these equations perturbatively,
698: for each order of $\delta$. The solution of the order of ${\cal O}(1)$ reads for
699: the coefficient $a_2(t)$ \cite{BPMaxw}:
700: \begin{equation}
701: \label{a20tsmall}
702: a_{20}(t) \approx a_{20}(0)e^{-4t\left/\left(5 \tau_{\rm E}(0)\right)\right.}\,,
703: \end{equation}
704: where $\tau_{\rm E} =\frac32 \tau_c$ is the Enskog relaxation time, so that
705: $a_{20}(t)$ vanishes for $t \sim \tau_0$. This refers to the relaxation of an initially non-Maxwellian velocity distribution to the Maxwellian distribution. Note that the relaxation occurs within few collisions per particle, similarly to the relaxation of common molecular gases.
706:
707: We now assume that the initial distribution is Maxwellian, i.e., that $a_{20}(0)=0$ for
708: $t=0$. Then the deviation from the Maxwellian distribution originates from the
709: inelasticity of the interparticle collisions. For the case of $a_{20}(0)=0$ the
710: solution of the order of ${\cal O}(1)$ for the reduced temperature reads
711: \begin{equation}
712: \label{T(t)del1}
713: \frac{T(t)}{T_0}=u_0(t)= \left(1 + \frac{t}{\tau_0} \right)^{-5/3}\,,
714: \end{equation}
715: which coincides with the time-dependence of the temperature obtained previously
716: using scaling arguments~\cite{TomThor} (up to a constant $\tau_0$ which may
717: not be determined by scaling arguments).
718:
719: The solution for $a_2(t)$ in linear approximation with respect to $\delta$ reads
720: \begin{equation}
721: \label{a21gensolLi}
722: a_2(t)=\delta \cdot a_{21}(t)=-\frac{12}{5}w(t)^{-1}
723: \left\{ {\rm Li} \left[ w(t) \right]-{\rm Li} \left[ w(0) \right] \right\}
724: \end{equation}
725: where
726: \begin{equation}
727: \label{w(t)}
728: w(t) \equiv \exp \left[ \left(q_0 \delta \right)^{-1} \left(1+t/\tau_0 \right)^{1/6} \right]\,.
729: \end{equation}
730: and with the logarithmic Integral
731: \begin{equation}
732: {\rm Li}(x)=\int\limits_0^x\frac{1}{\ln(t)} dt\,.
733: \end{equation}
734: For $t \ll \tau_0$ the coefficient $a_2(t)$
735: (\ref{a21gensolLi}) reduces to
736: \begin{equation}
737: \label{a21tsmallsol}
738: a_{2}(t)=
739: - \delta \cdot h \left( 1- e^{-4t\left/\left(5\tau_{\rm E}(0)\right)\right.} \right)
740: \end{equation}
741: where
742: \begin{equation}
743: h \equiv 2^{1/10} \left( {\cal B}_{10}-5{\cal A}_{10}\right)/16 \pi
744: =(3/10)\Gamma(21/10)2^{1/5}C_1=0.415964\,.
745: \end{equation}
746: As it follows from Eq.~(\ref{a21tsmallsol}), after a transient
747: time of the order of few collisions per particle, i.e. for
748: $\tau_{\rm E}(0) < t \ll \tau_0$, $a_{2}(t)$ saturates at the
749: ``steady-state''- value $-h\, \delta=-0.415964\,\delta$, i.e. it changes only slowly on the time-scale $\sim \tau_c(0)$. On the other hand, for $t \gg \tau_0$ one obtains
750: \begin{equation}
751: \label{a21tlargesol}
752: a_{2}(t) \simeq
753: - \delta \cdot h \left( t/\tau_0 \right)^{-1/6}
754: \end{equation}
755: so that $a_{2}(t)$ decays to zero on a
756: time-scale $\sim \tau_0$, i.e. slowly in the collisional time-scale
757: $\sim \tau_c(0) \ll \tau_0$. The velocity distribution thus tends asymptotically
758: to the Maxwellian distribution. For this regime the first-order correction
759: for the reduced temperature, $u_1(t)$, reads~\cite{BPMaxw}:
760: \begin{equation}
761: \label{u1asympsol}
762: u_1(t)=\left( \frac{12}{25}h+2q_1 \right)(t/\tau_0)^{-11/6}=
763: 3.26856\,(t/\tau_0)^{-11/6}\,,
764: \end{equation}
765: where we used the above results for the constants $h$ and $q_1$. From the last
766: equation one can see how the coupling between the temperature evolution and the
767: evolution of the velocity distribution influences the evolution of temperature. Indeed, if there
768: were no such coupling, there would be no coupling term in Eq.~(\ref{genseteq1}), and
769: thus no contribution from $\frac{12}{25}h$ to the prefactor of $u_1(t)$ in Eq.~(\ref{u1asympsol}). This would noticeably change the time behavior of $u_1(t)$.
770: On the other hand, the leading term in the time dependence of temperature,
771: $u_0(t)$, is not affected by this kind of coupling.
772:
773: In Fig.~\ref{fig:a2} and Fig.~\ref{fig:T} we show the time dependence of the coefficient
774: $a_2(t)$ of the Sonine polynomial expansion and of the temperature of the
775: granular gas. The analytical findings are compared with the numerical
776: solution of the system (\ref{genseteq1},\ref{genseteq2}). As one can see from the
777: figures the analytical theory reproduces fairly well the
778: numerical solution for the case of small $\delta$.
779: \begin{figure}[htbp]
780: \centerline{\psfig{figure=linLi.eps,width=5.5cm}\psfig{figure=linLilong.eps,width=5.5cm}}
781: \centerline{\psfig{figure=a2scale.eps,width=5.5cm}}
782: \caption{Time dependence of the second coefficient of the Sonine polynomial
783: expansion $a_2(t)$. Time is given in units of the mean collisional time
784: $\tau_c(0)$. (Left): $a_2 \times 1000$ (solid lines) for
785: $\delta =0.001, 0.005, 0.01, 0.015$
786: %???? Please, check the last number, 0.015 !!!!!!!!!}
787: (top to bottom) together with the linear approximation (dashed lines); (Right): the same as (left) but for larger times; (Middle): $-a_2(t)$ over time (log-scale) for $\delta =0.03, 0.01, 0.003, 0.001$ (top to bottom) together with the power-law asymptotics $\sim t^{-1/6}$.}
788: \label{fig:a2}
789: \end{figure}
790: %Fig.2 %%%%% Yours Fig.1 -> Fig.2a, Fig.2 -> Fig.2b, Fig.9 -> Fig.2c %%%%%%
791:
792: \begin{figure}[htbp]
793: \centerline{\psfig{figure=TdecayA.eps,width=5.5cm}\psfig{figure=TdecayB.eps,width=5.5cm}}
794:
795: \centerline{\psfig{figure=TdecayAlog.eps,width=5.5cm}}
796: \caption{Time-evolution of the reduced temperature, $u(t)=T(t)/T_0$.
797: The time is given in units of mean collisional time
798: $\tau_c(0)$. Solid line: numerical solution, short-dashed:
799: $u_0(t)=(1+t/\tau_0)^{-5/3}$ (zero-order theory),
800: %??? {\bf Please, check the power {-5/3} !!!!! in your simulations !!!!!!!!!}
801: long-dashed: $u(t)=u_0(t)+\delta \, u_1(t)$ (first-order theory).
802: (Left): for $\delta=0.05, 0.1$ (top to bottom); (Right): $\delta=0.15, 0.25$ (top to bottom);
803: (Middle): the same as (Left) but log-scale and larger ranges.}
804: \label{fig:T}
805: \end{figure}
806: %Fig.3 %%%%%%%% Yours Fig.4 -> Fig.3a, Fig.5 -> Fig.3b, Fig.6 ->Fig.3c %%%%%%%%%
807: %%%
808:
809: As it follows from Fig.~\ref{fig:a2} (where the time is given in collisional units),
810: for small $\delta$ the following scenario of evolution of the velocity
811: distribution takes place for a force-free granular gas.
812: The initial Maxwellian distribution evolves to a non-Maxwellian distribution, with
813: the discrepancy between these two characterized by the second coefficient of the
814: Sonine polynomials expansion $a_2$. The deviation from the Maxwellian distribution
815: (described by $a_2$) quickly grows, until it saturates after a few collisions per
816: particle at a ``steady-state'' value. At this instant the deviation from the Maxwellian
817: distribution is maximal, with the value $a_2 \approx - 0.4 \delta$ (Fig.~\ref{fig:a2}a).
818: This refers to the first ``fast'' stage of the evolution, which takes place on a
819: mean-collision time-scale $\sim \tau_c(0)$. After this maximal deviation is reached,
820: the second ``slow'' stage of the evolution starts. At this stage $a_2$ decays to zero on
821: the ``slow'' time scale $\tau_0 \sim \delta^{-1} \tau_c(0) \gg \tau_0(0)$, which
822: corresponds to the time scale of the temperature evolution (Fig.~\ref{fig:a2}b); the decay of the
823: coefficient $a_2(t)$ in this regime occurs according to a power law $ \sim t^{-1/6}$
824: (Fig.~\ref{fig:a2}c). Asymptotically the
825: Maxwellian distribution would be achieved, if the clustering process did not occur.
826:
827: Fig.~\ref{fig:T} illustrates the significance of the first-order correction $u_1(t)$ in the time-evolution
828: of temperature. This becomes more important as the dissipation parameter $\delta$
829: grows (Figs.~\ref{fig:T}a,b). At large times the results of the first-order theory
830: (with $u_1(t)$ included) practically coincide with the numerical results, while
831: zero-order theory (without $u_1(t)$) demonstrates noticeable deviations (Fig.~\ref{fig:T}c).
832:
833: For larger values of $\delta$ the linear theory breaks down. Unfortunately, the
834: equations obtained for the second order approximation ${\cal O}(\delta^2)$ are
835: too complicated to be treated analytically. Hence, we studied them only numerically
836: (see Fig.~\ref{fig:a2time}). As compared to the case of small $\delta$, an additional intermediate regime in
837: the time-evolution of the velocity distribution is observed. The first ``fast'' stage
838: of evolution takes place, as before, on the time scale of few collisions per particle,
839: where maximal deviation from the Maxwellian distribution is achieved (Fig.~\ref{fig:a2time}). For
840: $\delta \geq 0.15$ these maximal values of $a_2$ are positive. Then, on the second
841: stage (intermediate regime), which continues $10-100$ collisions, $a_2$ changes its
842: sign and reaches a maximal negative deviation. Finally, on the third, slow stage,
843: $a_2(t)$ relaxes to zero on the slow time-scale $\sim \tau_0$, just as for
844: small $\delta$. In Fig.~\ref{fig:a2time} we show the first stage of the
845: time evolution of $a_2(t)$ for systems with large $\delta$. At a certain value of the dissipative parameter $\delta$ the behavior changes qualitatively, i.e. the system then reveals another time scale as discussed above.
846: \begin{figure}[htbp]
847: \centerline{\psfig{figure=a2.01-02.eps,width=6cm}}
848: \caption{Time dependence of the second coefficient of the Sonine polynomial expansion $a_2(t) \times 100 $.
849: %??? {\bf Please, check the factor 100, my be this should be 10 as far as I can remember, that I saw.}
850: Time is given in units of mean collisional time $\tau_c(0)$. $\delta=0.1, 0.11, 0.12, \ldots, 0.20$ (bottom to top).}
851: \label{fig:a2time}
852: \end{figure}
853: %Fig.4 %%%%% Yours Fig.10 -> Fig.4 %%%%%%
854: %%
855:
856: Figure~\ref{fig:a2evo} shows the numerical solution of Eqs.~(\ref{genseteq1}) and (\ref{genseteq2}) for the second Sonine coefficient $a_2(t)$ as a function of time. One can clearly distinguish the different stages of evolution of the velocity distribution function. A more detailed investigation of the evolution of the distribution function for larger dissipation is subject of present research~\cite{BPMaxw}.
857:
858: \begin{figure}[htbp]
859: \centerline{\psfig{figure=a2.60.eps,width=5cm}\psfig{figure=a2.1200.eps,width=5cm}}
860: \centerline{\psfig{figure=a2.12000.eps,width=5cm}\psfig{figure=a2.1200000.eps,width=5cm}}
861: \caption{The second Sonine coefficient $a_2$ for $\delta=0.16$ over time. The numerical solutions of Eqs.~(\ref{genseteq1}) and (\ref{genseteq2}) show all stages of evolution discussed in the text.}
862: \label{fig:a2evo}
863: \end{figure}
864:
865: The interesting property of the granular gases in the regime of homogeneous
866: cooling is the overpopulation of the high-velocity tails in the velocity
867: distribution~\cite{EsipovPoeschel:97}, which has been shown for granular gases consisting of particles which interact via a constant restitution coefficient,
868: $\epsilon ={\rm const}$. How does the velocity dependence of the
869: restitution coefficient as it appears for viscoelastic spheres influence this effect? We observe, that for the
870: case of $\epsilon=\epsilon(v_{\rm imp})$ the functional form (i.e. the exponential
871: overpopulation~\cite{EsipovPoeschel:97}) persists, but it decreases with time on the
872: ``slow'' time-scale $\sim \tau_0$. Namely we obtain for the velocity distribution
873: for $c \gg 1$ \cite{BPMaxw}:
874: \begin{equation}
875: \label{veldiscgg1}
876: \tilde{f}(\vec{c}, t) \sim
877: \exp \left[ - \frac{b}{ \delta}\,c\left(1+\frac{t}{\tau_0} \right)^{1/6} \right]\, .
878: \end{equation}
879: where
880: $b=\sqrt{\pi/2}\left(16q_0/5\right)^{-1}=2.25978\ldots$, which holds for
881: $t \gg \tau_c(0)$.
882: Again we see that the distribution tends asymptotically to the Maxwellian distribution,
883: since the overpopulation vanishes as $t \to \infty$.
884:
885: Using the temperature and the velocity distribution of a granular gas as were derived in this section, one can
886: calculate the kinetic coefficients. In the next section we consider the simplest one --
887: the self-diffusion coefficient.
888:
889: \section{Self-diffusion in granular gases of viscoelastic particles}
890:
891: In the simplest case diffusion of particles occurs when there are
892: density gradients in the system.
893: The diffusion coefficient $D$ relates the flux of particles $\vec{J} $ to the
894: density gradient $\nabla n$ according to a linear relation, provided the gradients are
895: not too large:
896: \begin{equation}
897: \label{Jdiff}
898: \vec{J}= -D \nabla n\,.
899: \end{equation}
900: The coefficient $D$ also describes the statistical average of the migration of a single particle. For {\em equilibrium}
901: 3D-systems the mean-square displacement of a particle reads
902: \begin{equation}
903: \left\langle \left( \Delta r(t) \right)^2 \right\rangle_{\rm eq}
904: = 6\,D\,t \,,
905: \label{Dgen}
906: \end{equation}
907: where $\left< \cdots \right>_{\rm eq}$ denotes the {\em equilibrium}
908: ensemble averaging. For {\em nonequilibrium} systems, such as granular gases,
909: one should consider the time-dependent diffusion coefficient $D(t)$ and the corresponding
910: generalization of Eq.~(\ref{Dgen}):
911: \begin{equation}
912: \left\langle \left( \Delta r(t) \right)^2 \right\rangle =
913: 6\, \int^t D(t^{\prime}) dt^{\prime}\,,
914: \label{Dgengen}
915: \end{equation}
916: where $\langle \cdots \rangle$ denotes averaging over the nonequilibrium ensemble.
917: If the migration of a particle occurs in a uniform system composed of particles of the
918: same kind, this process is called ``self-diffusion''. Correspondingly, the kinetic
919: coefficient $D$ is called self-diffusion coefficient.
920:
921: To find the mean-square displacement, one writes
922: \begin{equation}
923: \left\langle \left( \Delta r(t) \right)^2 \right\rangle =
924: \left< \int_0^{t}\vec{v}(t^{\prime})dt^{\prime}
925: \int_0^t \vec{v}(t^{\prime \prime}) dt^{\prime \prime} \right> \,
926: \label{delR}
927: \end{equation}
928: and encounters then with the velocity autocorrelation function
929:
930: $$
931: K_v(t^{\prime}, t)
932: \equiv \left< \vec{v}(t^{\prime}) \vec{v}(t^{\prime \prime}) \right>
933: $$
934: which should be evaluated in order to obtain the mean-square displacement and
935: the self-diffusion coefficient.
936:
937: To calculate $K_v(t^{\prime}, t)$ we use the approximation of uncorrelated
938: successive binary collision, which is valid for moderately dense systems, and
939: an approach based on the formalism of the pseudo-Liouville operator
940: ${\cal L}$~\cite{pseudo}. The pseudo-Liouville operator is defined as
941: \begin{equation}
942: i{\cal L}=\sum_j \vec{v}_j \cdot \frac{\partial}{\partial \vec{r}_j}
943: +\sum_{i<j}\, \hat{T}_{\mbox{\footnotesize\em ij}}\,.
944: \label{L}
945: \end{equation}
946: The first sum in (\ref{L}) refers to the free streaming of the
947: particles (the ideal part) while the second sum refers to the particle
948: interactions which are described by the binary collision
949: operators~\cite{Chandler}
950: \begin{equation}
951: \hat{T}_{\mbox{\footnotesize\em ij}}\!=\sigma^{2}\!\! \int\!\! d\vec{e}\,
952: \Theta \left(- \vec{v}_{\mbox{\footnotesize\em ij}} \cdot \vec{e}\, \right)\!
953: |\vec{v}_{\mbox{\footnotesize\em ij}} \cdot \vec{e}\, |
954: \delta\! \left( \vec{r}_{\mbox{\footnotesize\em ij}}- \sigma \vec{e}
955: \right)\!\!
956: \left(\hat{b}_{\mbox{\footnotesize\em ij}}^{\vec{e}}-1 \right) \,,
957: \label{Tij}
958: \end{equation}
959: where $\Theta(x)$ is the Heaviside function. The operator
960: $\hat{b}_{\mbox{\footnotesize\em ij}}^{{\vec{e}}}$ is defined as
961: \begin{equation}
962: \hat{b}_{\mbox{\footnotesize\em ij}}^{\vec{e}} f \left (\vec{r}_{i},
963: \vec{r}_{j}, \vec{v}_{i},\vec{v}_{j} \cdots \right)=f \left
964: (\vec{r}_{i}, \vec{r}_{j},
965: \vec{v}^{*}_{i},\vec{v}^{*}_{j} \cdots \right) \, ,
966: \end{equation}
967: where $f$ is some function of the dynamical variables and $\vec{v}^{*}_{i}$ and $\vec{v}^{*}_{j}$ are the postcollisional velocities from Eq.~(\ref{directcoll}).
968: The pseudo-Liouville operator gives the time derivative of any
969: dynamical variable $B$ (e.g.~\cite{resibua}):
970: \begin{equation}
971: \frac{d}{dt} B\left( \left\{ \vec{r}_i, \vec{v}_i \right\}, t \right)=
972: i{\cal L}\, B\left( \left\{ \vec{r}_i, \vec{v}_i \right\}, t \right)\,.
973: \label{derA}
974: \end{equation}
975: Therefore, the time evolution of $B$
976: reads ($t>t^{\prime}$)
977: \begin{equation}
978: B\left( \{ \vec{r}_i, \vec{v}_i \}, t \right)=
979: e^{i{\cal L} (t-t^{\,\prime} )}
980: B\left( \{ \vec{r}_i, \vec{v}_i \}, t^{\,\prime} \right)\,.
981: \label{evolA}
982: \end{equation}
983: With Eq.~(\ref{evolA})
984: the time-correlation function reads
985: \begin{equation}
986: \left< B(t^{\prime})B(t) \right >=
987: \int d\Gamma \rho(t^{\prime}) B(t^{\prime}) e^{i{\cal L} (t-t^{\prime})}
988: B(t^{\prime})\,,
989: \label{evolAA}
990: \end{equation}
991: where $\int d\Gamma$ denotes integration over all degrees of freedom
992: and $\rho(t^{\prime})$ depends on temperature $T$, density $n$, etc.,
993: which change on a time-scale $t \gg \tau_c$.
994:
995: Now we assume that
996: \begin{enumerate}
997: \item[(i)] the coordinate part and the velocity part of the distribution
998: function $\rho(t)$ factorize, and
999: \item [(ii)] the molecular chaos hypothesis is valid.
1000: \end{enumerate}
1001: This suggests the following form of the distribution function:
1002: \begin{equation}
1003: \rho(t)=\rho(\vec{r}_1, \ldots, \vec{r}_N)\cdot f(\vec{v}_1,t) \ldots f(\vec{v}_N,t)\,.
1004: \label{molchaos}
1005: \end{equation}
1006: In accordance with the
1007: molecular chaos assumption the sequence of the
1008: successive collisions occurs without correlations. If the variable $B$
1009: does not depend on the positions of the particles, its
1010: time-correlation function decays exponentially~\cite{Chandler1}:
1011:
1012: \begin{equation}
1013: \left\langle B(t^{\prime})B(t) \right \rangle =
1014: \left< B^2 \right>_{t^{\prime}} e^{-\left.\left|t-t^\prime\right|\right/\tau_B(t^\prime)}
1015: ~~~~\left(t > t^{\prime}\right) \,.
1016: \label{AAexp}
1017: \end{equation}
1018: where $\langle \cdots \rangle_{t^\prime}$ denotes the averaging with
1019: the distribution function taken at time $t^{\prime}$. The relaxation
1020: time $\tau_B$ is inverse to the initial slope of the autocorrelation
1021: function~\cite{Chandler1}, as it may be found from the time derivative
1022: of $\left\langle B(t^{\prime} )B(t) \right \rangle$ taken at
1023: $t=t^{\prime}$. Equations~(\ref{evolAA}) and (\ref{AAexp}) then yield
1024: \begin{equation}
1025: -\tau_B^{-1}(t^{\prime})=\int\!\!d\Gamma \rho(t^{\prime}) B {i{\cal L} }B /
1026: \left\langle B^2 \right \rangle_{t^{\prime}}=
1027: \frac{\left\langle B i{\cal L} B \right
1028: \rangle_{t^{\prime}}}{\left\langle B^2 \right \rangle_{t^{\prime}}}\,.
1029: \label{ALA}
1030: \end{equation}
1031: The relaxation time $\tau_B^{-1}(t^{\prime})$ depends on time via
1032: the distribution function $\rho(t^{\prime})$ and varies on the
1033: time-scale $t \gg \tau_c$.
1034:
1035: Let $B(t)$ be the velocity of some particle, say $\vec{v}_1(t)$. Then
1036: with $3T(t)=\left\langle v^2 \right \rangle_t$,
1037: Eqs.~(\ref{AAexp}) and (\ref{ALA}) (with Eqs.~(\ref{L}) and (\ref{Tij})) read~\cite{BrilliantovPoeschel:1998d}
1038: \begin{equation}
1039: \left\langle \vec{v}_1 (t^{\prime})\cdot \vec{v}_1(t)\right\rangle =
1040: 3T(t^{\prime}) e^{-|t-t^{\prime}|/\tau_v(t^{\prime})}
1041: \label{vvexp}
1042: \end{equation}
1043: \begin{equation}
1044: -\tau_v^{-1}(t^\prime)=
1045: (N-1) \frac{\left< \vec{v}_1 \cdot \hat{T}_{12}
1046: \vec{v}_1\right>_{t^{\prime}}}{\left< \vec{v}_1 \cdot \vec{v}_1 \right>_{t^{\prime}}}\,.
1047: \label{vTv}
1048: \end{equation}
1049: To obtain Eq.~(\ref{vTv}) we take into account that ${\cal L}_0
1050: \vec{v}_1=0$, $\hat{T}_{\mbox{\footnotesize\em ij}}\,\vec{v}_1=0$ (for $i\neq 1$) and the identity of the particles.
1051:
1052: Straightforward calculation yields for the case of a constant
1053: restitution coefficient:
1054: \begin{equation}
1055: \tau_v^{-1}(t)=\frac{\epsilon +1}{2}\frac83 n \sigma^2 g_2(\sigma)
1056: \sqrt{\pi T(t)} =\frac{\epsilon +1}{2} \tau_E^{-1}(t)\,,
1057: \label{tEt}
1058: \end{equation}
1059: where $\tau_E(t)=\frac32\,\tau_c(t)$ is the Enskog relaxation time
1060: \cite{resibua}. Note that according to Eq.~(\ref{tEt}),
1061: $\tau_v =\frac{2}{1+\epsilon} \tau_E > \tau_E$, i.e.,
1062: the velocity correlation time for inelastic collisions exceeds that of
1063: elastic collisions. This follows from partial suppression of the
1064: backscattering of particles due to inelastic losses in their normal
1065: relative motion, which, thus, leads to more stretched particle trajectories,
1066: as compared to the elastic case.
1067:
1068: Similar (although somewhat more complicated) computations may be
1069: performed for the system of viscoelastic
1070: particles yielding
1071: \begin{equation}
1072: \tau_v^{-1}(t)=
1073: \tau_E^{-1}(t) \left[1+\frac{3}{16}a_2(t)-4q_0 \, \delta \,u^{1/10}(t) \right]\,,
1074: \label{tauva}
1075: \end{equation}
1076: where $q_0=0.173318$ has been already introduced and $a_2(t)$, $u(t)$ are the
1077: same as defined above.
1078: To obtain Eq.~(\ref{tauva}) we neglect terms of the order
1079: of ${\cal O}(a_2^2)$,
1080: ${\cal O}(\delta^2)$ and ${\cal O}(a_2\, \delta)$.
1081:
1082: Using the velocity correlation function one writes
1083: \begin{equation}
1084: \left\langle \left( \Delta r(t) \right)^2 \right\rangle
1085: =2 \int_0^t dt^\prime 3T (t^\prime) \int_{t^\prime}^t
1086: dt^{\prime\prime} e^{-|t^{\prime\prime}-t^\prime|/\tau_v(t^\prime)}\,.
1087: \label{Difvel}
1088: \end{equation}
1089: On the short-time scale $t \sim \tau_c$, $T (t^\prime)$ and
1090: $\tau_v(t^\prime)$ may be considered as constants. Integrating in
1091: (\ref{Difvel}) over $t^{\prime\prime}$ and equating with
1092: (\ref{Dgengen}) yields for $t \gg \tau_c \sim \tau_v$ the
1093: time-dependent self-diffusion coefficient
1094: \begin{equation}
1095: D(t)= T(t) \tau_v(t)\,.
1096: \label{Difviatau}
1097: \end{equation}
1098:
1099: Substituting the dependencies for $u(t)=T(t)/T_0$ and $a_2(t)$ as functions of time, which has been derived in
1100: the previous section, we obtain the time dependence of the coefficient of
1101: self-diffusion $D(t)$. For $t \gg \tau_0$ this may be given in an explicit form:
1102: \begin{equation}
1103: \frac{D(t)}{D_0} \simeq \left(\frac{t}{\tau_0} \right)^{-5/6} \!\!
1104: +\delta \left(4q_0+q_1+\frac{21}{400}h \right) \left(\frac{t}{\tau_0} \right)^{-1}\,,
1105: \label{Diffinal}
1106: \end{equation}
1107: where the constants $q_0$, $q_1$ and $h$ are given above. Hence,
1108: the prefactor in the term proportional to $\delta$ reads
1109: $\left(4q_0+q_1+\frac{21}{400}h \right)=2.24956$, and $D_0$ is the initial Enskog value
1110: of the self-diffusion coefficient
1111: \begin{equation}
1112: D_0^{-1} = \frac83 \, \pi^{1/2} n g_2(\sigma) \sigma^2 T_0^{-1/2}\,.
1113: \end{equation}
1114: Correspondingly, the mean-square displacement reads
1115: asymptotically for $t \gg \tau_0 $:
1116: \begin{equation}
1117: \left\langle \left( \Delta r(t) \right)^2 \right\rangle \sim
1118: t^{1/6}+ b \, \delta\, \log t + \ldots\,,
1119: \label{dRasym}
1120: \end{equation}
1121: where $b$ is some constant.
1122: This dependence holds true for times
1123: \begin{equation}
1124: \tau_c(0)\,\delta^{-1} \ll t \ll \tau_c(0)\, \delta^{-11/5}\,.
1125: \label{ineq}
1126: \end{equation}
1127: The
1128: first inequality in Eq.~(\ref{ineq}) follows from the condition $\tau_0 \ll t$, while the
1129: second one follows from the condition $\tau_c(t) \ll
1130: \tau_0$, which means that temperature changes are slow on the collisional time-scale. For the constant restitution coefficient one obtains
1131: \begin{equation}
1132: T(t)/T_0=\left[ 1+\gamma_0 t/\tau_c(0) \right]^{-2}\,,
1133: \end{equation}
1134: where $\gamma_0 \equiv \left(1-\epsilon^2\right)/6$ \cite{GZ93,NEBO97}. Thus, using
1135: Eqs.~(\ref{tEt}) and (\ref{Difviatau}) one obtains for the
1136: mean-square displacement in this case
1137: \begin{equation}
1138: \label{drcons}
1139: \left< \left( \Delta r(t) \right)^2
1140: \right> \sim \log t \, .
1141: \end{equation}
1142: As it follows from Eqs.~(\ref{dRasym}) and (\ref{drcons}) the impact-velocity
1143: dependent restitution coefficient, Eq.~(\ref{epsC1C2}), leads to a significant
1144: change of the long-time behavior of the mean-square displacement of
1145: particles in the laboratory-time. Compared to its logarithmically
1146: weak dependence for the constant restitution coefficient, the
1147: impact-velocity dependence of the restitution coefficient
1148: gives rise to a considerably faster increase of this quantity with time,
1149: according to a power law.
1150:
1151: One can also compare the dynamics of the system in its
1152: inherent-time scale. First we consider the average cumulative number of
1153: collisions per particle ${\cal N}(t)$ as an inherent measure for time
1154: (e.g.~\cite{McNamara96,EBEuro}).
1155: It may be found by integrating
1156: $d{\cal N}=\tau_c(t)^{-1}dt$ \cite{NEBO97}. For a constant restitution
1157: coefficient $\epsilon$ one obtains ${\cal N}(t) \sim \log t$, while for the impact-velocity
1158: dependent $\epsilon\left( v_{\rm imp} \right)$ one has ${\cal N}(t) \sim t^{1/6}$. Therefore, the temperature and the mean-square
1159: displacement behave in these cases as\\
1160:
1161: \begin{center}
1162: %\begin{table}
1163: %\centering
1164: %\caption{xx}
1165: \renewcommand{\arraystretch}{1.4}
1166: \setlength\tabcolsep{5pt}
1167: \begin{tabular}{l|l}
1168: \hline
1169: $\epsilon={\rm const}$ & $\epsilon=\epsilon\left( v_{\rm imp} \right)$\\
1170: \hline\noalign{\smallskip}
1171: $T({\cal N}) \sim e^{-2(1-\epsilon^2){\cal N}}$ & $T({\cal N}) \sim {\cal N}^{-10} $ \\
1172: \noalign{\smallskip}
1173: \hline
1174: \noalign{\smallskip}
1175: $\left< \left( \Delta r({\cal N}) \right)^2\right> \sim {\cal N}$ & $\left< \left( \Delta r({\cal N}) \right)^2\right> \sim {\cal N}$ \\
1176: \noalign{\smallskip}
1177: \hline
1178: \end{tabular}
1179: %\label{Tab1a}
1180: %\end{table}
1181: \end{center}
1182: \vspace{0.1cm}
1183:
1184: If the number of collisions per particle ${\cal N}(t)$ would be the relevant
1185: quantity specifying the stage of the granular gas evolution, one would
1186: conjecture that the dynamical behavior of a granular gas with a
1187: constant $\epsilon$ and velocity-dependent $\epsilon$ are identical,
1188: provided an ${\cal N}$-based time-scale is used. Whereas in equilibrium systems the number of collisions is certainly an appropriate measure of time, in nonequilibrium systems this value has to be treated with more care. As a trivial example may serve a particle bouncing back and forth between two walls, each time it hits a wall it loses part of its energy: If one describes this system using a ${\cal N}$-based time, one would come to the conclusion that the system conserves its energy, which is certainly not the proper description of the system. According to our
1189: understanding, therefore, the number of collision is not an appropriate time scale to describe physical reality. Sometimes, it may be even misleading.
1190:
1191: Indeed, as it was shown in Ref.~\cite{EBEuro}, the value of
1192: ${\cal N}_c$, corresponding to a crossover from the linear regime of
1193: evolution (which refers to the homogeneous cooling state) to the nonlinear
1194: regime (when clustering starts) may differ by orders of magnitude,
1195: depending on the restitution coefficient and on the density of the granular
1196: gas. Therefore, to analyze the behavior of a granular gas, one can try
1197: an alternative inherent time-scale, ${\cal T}^{-1} \equiv T(t)/T_0$
1198: which is based on the gas temperature. Given two
1199: systems of granular particles at the same density and the same initial
1200: temperature $T_0$, consisting of particles colliding with constant and
1201: velocity-dependent restitution coefficient, respectively, the time
1202: ${\cal T}$ allows to compare directly their evolution. A strong
1203: argument to use a temperature-based time has been given by Goldhirsch
1204: and Zanetti~\cite{GZ93} who have shown that as the temperature decays, the
1205: evolution of the system changes from a linear regime to a nonlinear one.
1206: Recent numerical results of Ref.~\cite{EBEuro}
1207: also support our assumption: It was shown that while ${\cal N}_c$
1208: differs by more than a factor of three for two different systems,
1209: the values of ${\cal T}_c$, (defined, as ${\cal T}_c=T({\cal N}_c)/T_0$) are very close~\cite{EBEuro}. These arguments show that one could consider ${\cal T}$ as a
1210: relevant time-scale to analyze the granular gas evolution.
1211:
1212: With the temperature decay $T({\cal N})/T_0 \sim e^{-2 \gamma_0{\cal N}}$ for a constant
1213: restitution coefficient and $T({\cal N})/T_0 \sim {\cal N}^{-10}$ for the
1214: impact-velocity dependent one, we obtain the following dependencies:\\
1215: \begin{center}
1216: %\begin{table}
1217: %\centering
1218: %\caption{xx}
1219: \renewcommand{\arraystretch}{1.4}
1220: \setlength\tabcolsep{5pt}
1221: \begin{tabular}{l|l}
1222: \hline\noalign{\smallskip}
1223: $\epsilon={\rm const}$ & $\epsilon=\epsilon\left( v_{\rm imp} \right)$\\
1224: \noalign{\smallskip}
1225: \hline
1226: \noalign{\smallskip}
1227: $T \sim \frac{1}{ {\cal T} }$ & $T \sim \frac{1}{ {\cal T} } $ \\
1228: \noalign{\smallskip}
1229: $\left< \left( \Delta r({\cal T}) \right)^2\right> \sim \log {\cal T}$ & $\left< \left( \Delta r({\cal T}) \right)^2 \right> \sim {\cal T}^{1/10}$ \\
1230: \noalign{\smallskip}
1231: \hline
1232: \end{tabular}
1233: %\label{Tab2a}
1234: %\end{table}
1235: \end{center}
1236: \vspace{0.1cm}
1237:
1238: This shows that in the temperature-based time-scale, in which the cooling
1239: of both systems is synchronized, the mean-square displacement grows logarithmically
1240: slow for the case of constant restitution coefficient and much faster, as a power
1241: law, for the system of viscoelastic particles with $\epsilon = \epsilon(v_{\rm imp})$.
1242: Thus, we conclude that clustering may be retarded for the latter system.
1243:
1244: \section{Conclusion}
1245: In conclusion, we considered kinetic properties of granular gases composed of
1246: viscoelastic particles, which implies the
1247: impact-velocity dependence of the restitution coefficient. We found that
1248: such dependence gives rise to some new effects in granular gas dynamics:
1249: (i) complicated, non-monotonous time-dependence of the coefficient $a_2$ of
1250: the Sonine polynomial expansion, which describes the deviation of the
1251: velocity distribution from the Maxwellian and (ii) enhanced spreading of
1252: particles, which depends on time as a power law, compared to a logarithmically
1253: weak dependence for the systems with a constant $\epsilon$.
1254:
1255: The Table below compares the properties of granular gases consisting of particles interacting via a constant coefficient of restitution $\epsilon ={\rm const}$ and consisting of viscoelastic particles where the collisions are described using an impact velocity dependent restitution coefficient $\epsilon = \epsilon(v_{\rm imp})$:\\
1256: \begin{center}
1257: %\begin{table}
1258: %\centering
1259: %\caption{xx}
1260: \renewcommand{\arraystretch}{1.4}
1261: \setlength\tabcolsep{5pt}
1262: \begin{tabular}{l|l}
1263: \hline\noalign{\smallskip}
1264: $\epsilon={\rm const}$ & $\epsilon=\epsilon\left( v_{\rm imp} \right)$\\
1265: \noalign{\smallskip}
1266: \hline
1267: \noalign{\smallskip}
1268: $\epsilon $ is a model parameter & $\epsilon=1-C_1A \kappa^{2/5}v_{\rm imp}^{1/5}+\cdots $ \\
1269: &$C_1=1.15396$, $C_2=\frac35 C_1^2$, $\ldots$ \\
1270: &$\kappa=\kappa(Y, \nu, m, R)$ \\
1271: &$A=A(\eta_1, \eta_2, Y, \nu)$ \\
1272: &all quantities are defined via parameters \\[-0.15cm]
1273: &of the particle material $Y$, $\nu$, $\eta_{1/2}$ \\[-0.15cm]
1274: &and their mass and radius. \\
1275: \noalign{\smallskip}
1276: \hline
1277: \noalign{\smallskip}
1278: \multicolumn{2}{c}{\bf Small parameter} \\
1279: $1-\epsilon^2$ -- does not depend & $\delta =A\kappa^{2/5}T_0^{1/10}$ -- depends \\[-0.15cm]
1280: on the state of the system & on the initial temperature $T_0$. \\
1281: \noalign{\smallskip}
1282: \hline
1283: \noalign{\smallskip}
1284: \multicolumn{2}{c}{\bf Temperature}\\
1285: $T=T_0 \left( 1+t/\tau_0^{\prime}\right)^{-2}$ & $T=T_0\left( 1+t/\tau_0\right)^{-5/3}$ \\
1286: \noalign{\smallskip}
1287: \hline
1288: \noalign{\smallskip}
1289: \multicolumn{2}{c}{\bf Velocity distribution}\\
1290: $f(\vec{v}, t)=\frac{n}{v_0^3(t)} \tilde{f}(\vec{c})$
1291: & $f(\vec{v}, t)=\frac{n}{v_0^3(t)}\tilde{f}(\vec{c}, t) $ \\
1292: $\tilde{f}(\vec{c},t)=\phi(c) \left\{1 + \sum_{p=1}^{\infty} a_p S_p(c^2) \right\}$ &
1293: $\tilde{f}(\vec{c})=\phi(c) \left\{1 + \sum_{p=1}^{\infty} a_p(t) S_p(c^2) \right\}$ \\
1294: $a_2 ={\rm const.} $ & $a_2=a_2(t)$ -- is a (complicated) \\[-0.15cm]
1295: & function of time. \\
1296: \noalign{\smallskip}
1297: \hline
1298: \noalign{\smallskip}
1299: \multicolumn{2}{c}{\bf Self-diffusion}\\
1300: $\left< \left( \Delta r(t) \right)^2\right> \sim \log t $ & $\left< \left( \Delta r(t) \right)^2 \right> \sim t^{1/6} $ \\
1301: \noalign{\smallskip}
1302: \hline
1303: \end{tabular}
1304: %\label{Tab3a}
1305: %\end{table}
1306: \end{center}
1307: \vspace{0.1cm}
1308:
1309:
1310: \section*{Acknowledgements}
1311:
1312: We thank M. H. Ernst and I. Goldhirsch for valuable discussions.
1313:
1314: \begin{thebibliography}{99}
1315: \addcontentsline{toc}{section}{References}
1316: \bibitem{BSHPnumer}
1317: %\bibitem{BSHPtheor}
1318: %\bibitem{theo1}
1319: %J.~T.~Jenkins, {\em Rapid flow of granular materials},
1320: %in: R.~J.~Knops and A.~A.~Lacey (eds.), {\em Non--classical
1321: %continuum mechanics}, Cambridge University Press, (Cambridge, 1987);
1322: Y. Du, H. Li, and L. P. Kadanoff, Phys. Rev. Lett. {\bf 74}, 1268 (1995);
1323: T. Zhou and L. P. Kadanoff, Phys. Rev. E {\bf 54}, 623 (1996);
1324: %H. M. Jaeger, S. R. Nagel, and R. P. Behringer, Rev. Mod. Phys {\bf 68}, 1259 (1996);
1325: %S. Succi, J. Wang, and Y-H. Qian, J. Mod. Phys. C {\bf 8}, 999 (1997);
1326: %\bibitem{Kadanoff2}
1327: %F. Bridges, K. D. Supulver, D. N. C. Lin, R. Knight, and M. Zafra, ICARUS {\bf 123}, 422 (1996);
1328: A. Goldshtein, M. Shapiro, and C. Gutfinger, J. Fluid Mech. {\bf 316}, 29 (1996);
1329: A. Goldshtein, V. N. Poturaev, and I. A. Shulyak, Izvestiya Akademii Nauk SSSR, Mechanika Zhidkosti i Gaza {\bf 2}, 166 (1990);
1330: J. T. Jenkins and M. W. Richman, Arch. Part. Mech. Materials {\bf 87}, 355 (1985);
1331: V. Buchholtz and T. P\"oschel, Granular Matter {\bf 1}, 33 (1998);
1332: E. L. Grossman, T. Zhou, and E. Ben-Naim, Phys. Rev. E {\bf 55}, 4200 (1997);
1333: %\bibitem{Ernst1}
1334: F. Spahn, U. Schwarz, and J. Kurths, Phys. Rev. Lett. {\bf 78}, 1596 (1997);
1335: S. Luding and H. J. Herrmann, Chaos {\bf 9}, 673 (1999);
1336: S. Luding and S. McNamara, Gran. Matter {\bf 1}, 113 (1998);
1337: T. P. C. van Noije, M. H. Ernst, and R. Brito, Phys. Rev. E {\bf 57}, R4891 (1998);
1338: %\bibitem{Ernst2}
1339: J. A. C. Orza, R. Brito, T. P. C. van Noije, and M. H. Ernst,
1340: Int. J. Mod. Phys. C {\bf 8}, 953 (1997);
1341: %\bibitem{SelaGoldhirsh:98}
1342: N. Sela and I. Goldhirsch, J. Fluid. Mech., {\bf 361}, 41 (1998);
1343: N. Sela, I. Goldhirsch, and H. Noskowicz, Phys. Fluids {\bf 8}, 2337 (1996);
1344: I. Goldhirsch, M.-L. Tan, and G. Zanetti, J. Sci. Comp. {\bf 8}, 1 (1993);
1345: %\bibitem{Brey}
1346: T. Aspelmeier, G. Giese, and A. Zippelius, Phys. Rev. E {\bf 57}, 857 (1997);
1347: J. J. Brey and D. Cubero, Phys. Rev. E {\bf 57}, 2019 (1998);
1348: J. J. Brey, J. W. Dufty, C. S. Kim, and A. Santos, Phys. Rev. E {\bf 58}, 4648 (1998);
1349: J. J. Brey, M. J. Ruiz-Montero, and F. Moreno, Phys. Rev. E {\bf 55}, 2846 (1997);
1350: J. J. Brey, D. Cubero, and M. J. Ruiz-Montero, Phys. Rev. E {\bf 59}, 1256 (1999);
1351: %\bibitem{Kumaran99}
1352: V. Kumaran,
1353: Phys.Rev.E, {\bf 59}, 4188, (1999);
1354: %\bibitem{GarzoDufty99}
1355: V. Garzo and J. W. Dufty, Phys. Rev. E, {\bf 59}, 5895, (1999).
1356: %%%%%%%%%%%%%%%%%%
1357: \bibitem{McNamara96}
1358: S. McNamara and W. R. Young, Phys. Rev. E {\bf 53}, 5089 (1996).
1359:
1360: \bibitem{GZ93} I.~Goldhirsch and G.~Zanetti, {\em Phys.~Rev.~Lett.}, {\bf 70}, 1619 (1993).
1361:
1362: \bibitem{NoijeErnst:97}
1363: T. P. C. van Noije and M. H. Ernst,
1364: Granular Matter, {\bf 1}, 57 (1998).
1365: %%%%%%%%%%%%%%%%%%%%%%%%%
1366: \bibitem{EsipovPoeschel:97}
1367: S. E. Esipov and T. P\"{o}schel, J. Stat.
1368: Phys., {\bf 86}, 1385 (1997).
1369: \bibitem{NoijeErnstRing}
1370: T. P. C. van Noije, M. H. Ernst and R.Brito,
1371: Physica A {\bf 251}, 266 (1998).
1372: %%%%%%%%%%%%%%%%%%%%%
1373: \bibitem{GoldshteinShapiro95}
1374: A. Goldshtein and M. Shapiro,
1375: J. Fluid. Mech., {\bf 282}, 75 (1995).
1376: %%%%%%%%%%%%%%%%%%%%
1377: \bibitem{EBEuro}R. Brito and M. H. Ernst,
1378: Europhys. Lett. {\bf 43}, 497 (1998).
1379: \bibitem{NEBO97}T. P. C. van Noije and M. H.
1380: Ernst, R. Brito and J. A. G. Orza, Phys. Rev. Lett. {\bf 79}, 411
1381: (1997).
1382: \bibitem{BrilliantovPoeschel:1998d}
1383: N. V. Brilliantov and T.
1384: P\"oschel, Phys. Rev. E {\bf 61}, 1716 (2000).
1385: % Diffusion paper
1386: \bibitem{HuthmannZippelius:97}
1387: M. Huthmann and A. Zippelius, { Phys.
1388: Rev. E.} {\bf 56}, R6275 (1997); S. Luding, M. Huthmann,
1389: S. McNamara and A. Zippelius, { Phys.
1390: Rev. E.} {\bf 58}, 3416 (1998).
1391:
1392: \bibitem{ThorntonHere} C. Thornton, {\em Contact mechanics and coefficients of restitution}, (in this volume, page 184.
1393: %\pageref{Thorntonpage}).
1394:
1395: %\bibitem{SalazarBrenig99}
1396: % J.M.Salazar and L.Brenig,
1397: % Phys.Rev.E, {\bf 59}, 2093 (1999).
1398:
1399:
1400: \bibitem{rospap}
1401: R. Ram\'{\i}rez, T. P\"oschel, N. V. Brilliantov,
1402: and T. Schwager, Phys. Rev. E {\bf 60}, 4465 (1999).
1403:
1404: \bibitem{Taguchi93} Y. Taguchi, Europhys. Lett. {\bf 24}, 203 (1993).
1405:
1406: \bibitem{Luding98} S. Luding, {\em Collisions and Contacts between two particles}, in: H. ~J. Herrmann, J. -P. Hovi, and S. Luding (eds.), {\em Physics of dry granular media - NATO ASI Series E350}, Kluwer (Dordrecht, 1998), p. 285.
1407:
1408: \bibitem{BSHP1}
1409: N. V. Brilliantov, F. Spahn, J.-M. Hertzsch, and T.
1410: P\"oschel, Phys. Rev. E {\bf 53}, 5382 (1996).
1411: \bibitem{BSHP2}
1412: J.-M. Hertzsch, F. Spahn, and N. V. Brilliantov, J. Phys. II (France),
1413: {\bf 5}, 1725 (1995).
1414: \bibitem{remark2}
1415: Derivation of the dissipative force given in~\cite{BSHP1,BSHP2} for colliding
1416: spheres may be straightforwardly generalized to obtain the relation
1417: (\ref{eldiss}) (or (A17) in~\cite{BSHP1,BSHP2}) for colliding bodies of any
1418: shape, provided that displacement field in the bulk of the material
1419: of bodies in contact is a one-valued function of the compression
1420: (see also~\cite{BrilPoshprep}).
1421: \bibitem{BrilPoshprep}
1422: N. V. Brilliantov and T. P\"oschel, in preparation.
1423:
1424: \bibitem{LandauLifshits}
1425: L. D. Landau and E. M. Lifschitz, {\it Theory of Elasticity},
1426: Oxford University Press (Oxford, 1965).
1427: \bibitem{Hertz.Brill}
1428: H. Hertz, J. f. reine u. angewandte Math. {\bf 92}, 156 (1882).
1429:
1430: \bibitem{Kuwabara} G. Kuwabara and K. Kono, Jpn. J. Appl. Phys. {\bf 26}, 1230 (1987).
1431:
1432: \bibitem{TomThor}
1433: T. Schwager and T. P\"oschel, Phys. Rev. E
1434: {\bf 57}, 650 (1998).
1435:
1436: \bibitem{resibua} P. Resibois and M. de Leener, {\em Classical Kinetic
1437: Theory of Fluids} (Wiley, New York, 1977).
1438: \bibitem{CarnahanStarling}
1439: N. F. Carnahan, and K. E. Starling, J. Chem. Phys., {\bf 51}, 635 (1969).
1440:
1441: \bibitem{BPMaxw}
1442: N. V. Brilliantov and T. P\"oschel, Phys. Rev. E, {\bf 61}, 5573 (2000)
1443:
1444: \bibitem{pseudo}
1445: The term ``pseudo'' was initially referred to the
1446: dynamics of systems with singular hard-core potential
1447: \cite{Ernst:69,resibua}.
1448: \bibitem{Chandler} For application to ``ordinary'' fluids, see
1449: \cite{Chandler1} and to granular systems
1450: \cite{NoijeErnstRing,HuthmannZippelius:97}. A rigorous
1451: definition of ${\cal L}$ includes a pre-factor, preventing
1452: successive collisions of the same pair of particles
1453: \cite{Ernst:69,resibua} which, however, does not affect the present
1454: analysis.
1455:
1456: \bibitem{Chandler1}
1457: D. Chandler, {J. Chem. Phys.} {\bf 60}, 3500,
1458: 3508 (1974); B. J. Berne, {\em ibid} {\bf 66}, 2821, (1977);
1459: N. V. Brilliantov and O. P. Revokatov,
1460: Chem. Phys. Lett. {\bf 104}, 444 (1984).
1461:
1462: \bibitem{Ernst:69}
1463: M. H. Ernst, J. R. Dorfman, W. R. Hoegy and J. M.
1464: J. van Leeuwen, Physica A {\bf 45}, 127 (1969).
1465:
1466: \end{thebibliography}
1467:
1468: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1469: %%%%%%%%%%%%% FIGURE CAPTINOS %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1470:
1471:
1472:
1473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1474: %%%%%%%%%%%%%%%%%%%%%% END OF FIGURE CAPTINOS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1475:
1476: %INDEX%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1477: %\clearpage
1478: %\addcontentsline{toc}{section}{Index}
1479: %\flushbottom
1480: %\printindex
1481: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1482:
1483: %\end{document}
1484:
1485:
1486:
1487:
1488:
1489:
1490:
1491:
1492: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1493: %\include{Jenkins/Jenkins} %FINISHED
1494: %\include{Henrique/Henrique} %FINISHED
1495:
1496: %\addcontentsline{toc}{part}{Collisions and One-dimensional Models}
1497: %\part*{Collisions\\[0.3cm] and\\[0.3cm] One-dimensional Models}
1498: %\include{Bridges/Bridges} %FINISHED
1499: %\include{Thornton/Thornton} %FINISHED
1500: %\include{Ramirez/Ramirez}
1501: %\include{Poeschel/PoeschelBrilliantov} %FINISHED
1502: %\addtocontents{toc}{\vfill\newpage}
1503:
1504: %\addcontentsline{toc}{part}{Vibrated Granular Media}
1505: %\addtocontents{toc}{\vspace*{-0.07cm}}
1506: %\part*{Vibrated Granular Media}
1507: %\include{Huntley/Huntley}
1508: %\addtocontents{toc}{\vspace*{-0.07cm}}
1509: %\include{Clement/Clement} %FINISHED
1510: %\addtocontents{toc}{\vspace*{-0.07cm}}
1511: %\include{Falcon/Falcon} %FINISHED
1512: %\addtocontents{toc}{\vspace*{-0.07cm}}
1513: %\include{Farkas/Farkas} %FINISHED
1514: %\addtocontents{toc}{\vspace*{-0.07cm}}
1515: %\include{Goldshtein/goldshtein} %FINISHED
1516: %\addtocontents{toc}{\vspace*{-0.07cm}}
1517:
1518: %\addcontentsline{toc}{part}{Granular Astrophysics}
1519: %\part*{Granular Astrophysics}
1520: %\addtocontents{toc}{\vspace*{-0.07cm}}
1521: %\include{Brahic/Brahic}
1522: %\addtocontents{toc}{\vspace*{-0.07cm}}
1523: %\include{Salo/salo} %FINISHED (confirm e-mail)
1524: %\addtocontents{toc}{\vspace*{-0.07cm}}
1525: %\include{Hanninen/Hanninen} %FINISHED
1526: %\addtocontents{toc}{\vspace*{-0.07cm}}
1527: %\include{Spahn/Spahn2} %FINISHED
1528: %\addtocontents{toc}{\vspace*{-0.07cm}}
1529:
1530: %\addcontentsline{toc}{part}{Towards Dense Granular Systems}
1531: %\addtocontents{toc}{\vspace*{-0.07cm}}
1532: %\part*{Towards Dense Granular Systems}
1533: %\include{Luding/Luding} %FINISHED
1534: %\addtocontents{toc}{\vspace*{-0.07cm}}
1535: %\include{Urbach/Urbach} %FINISHED
1536: %\addtocontents{toc}{\vspace*{-0.07cm}}
1537: %\include{Hong/hong_new} %FINISHED
1538: %\addtocontents{toc}{\vspace*{-0.07cm}}
1539: %\include{Herrmann/honnef99} %FINISHED
1540:
1541: %\include{Book/author_index}
1542:
1543: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1544: %\clearpage
1545: %\addcontentsline{toc}{chapter}{Index}
1546: %\flushbottom
1547: %\printindex
1548: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1549:
1550: \end{document}
1551:
1552: