1: \documentclass[aps,amssymb,twocolumn]{revtex4}
2: %\input epsf
3: \usepackage{graphics}
4:
5: \def\epsfig{\special}
6: \def\D {{\rm D}}
7: \def\d {{\rm d}}
8: \def\vector#1{{{\underline {#1}}}}
9: \def\avector#1{{\overline{\underline {#1}}}}
10: \def\tensor#1{{\underline{\underline {#1}}}}
11: \def\atensor#1{{\overline{\underline{\underline {#1}}}}}
12: \def\div {{\rm div}}
13: \def\tr{{\rm tr}}
14: \def\dd{{\xi}}
15: \def\ss{{\sigma}}
16: \def\dpl{\gamma}
17: \def\ddpl{\dot\dpl}
18: \def\tdpl{\tensor\epsilon^{pl}}
19: \def\dtdpl{\dot\tensor\epsilon^{pl}}
20: \def\dtd{\dot\tensor\epsilon}
21: \def\din{\epsilon^{in}}
22: \def\tdin{\tensor\epsilon^{in}}
23: \def\del{\epsilon^{el}}
24: \def\tdel{\tensor\epsilon^{el}}
25: \def\sdef{\tau}
26: \def\sigp{\zeta}
27:
28: \begin{document}
29: %\bibliographystyle{prsty}
30: \title {A dynamical approach to glassy materials}
31: \author{Ana\"el Lema\^{\i}tre}
32: \affiliation{Department of physics, University of California, Santa Barbara,
33: California 93106, U.S.A.}
34: \affiliation{CEA --- Service de Physique de l'\'Etat Condens\'e,
35: Centre d'\'Etudes de Saclay, 91191 Gif-sur-Yvette, France}
36: \date{\today}
37: \begin{abstract}
38: Typical properties of glassy materials are shown to be captured by a mean-field
39: free-volume theory. Relaxation processes are supposed to be free-volume activated,
40: and different entropy barriers are associated with density relaxation
41: and shear motion.
42: Free-volume time logarithmic relaxation, Kohlrausch-Williams-Watts,
43: and power law viscosity result from
44: the non-linear dynamics of spatially averaged quantities.
45: The exponents associated with these phenomena are related to a single parameter
46: of the theory.
47: The theory also accounts for coexistence of jamming transitions and non-linear rheology.
48: %The study in complemented with the analysis of loss and storage moduli in the linear
49: %and non-linear regime.
50: \end{abstract}
51: \pacs{05.90.+m,83.10.Gr,83.60.-a,62.20.-x}
52:
53:
54: \maketitle
55:
56: %%
57: %%
58:
59: \section{Introduction}
60:
61: %(Recent characterization of aging have been on PMMA
62: %by low frequency dielectric measurements~\cite{bellon00}
63: %and on microgel pastes~\cite{cloitre00}
64: %Also in the context of solid friction~\cite{bureau02})
65:
66: The emergence of slow modes of relaxations is a major signature
67: of the glassy behavior.~\cite{goetze92}
68: The first consequence of large equilibration time is that the properties
69: of a glass evolve on experimental timescales, the system ages.
70: Despite the huge amount of work that has been devoted to glassy systems,
71: a general understanding of slow relaxation processes and of their
72: relation to macroscopic rheology is still lacking.
73: For example, experimental and numerical data converge to a stretched exponential
74: shape of relaxation phenomena, the so-called Kohlrausch-Williams-Watts
75: (KWW) relaxation,~\cite{ediger96,larson99,angell00}
76: while mode coupling theory predicts power law behavior.~\cite{goetze92}
77: When it comes to extend these results to non-linear rheology,
78: the situation is even more difficult from the theoretical point of view,
79: since the system is driven far from equilibrium.
80: At an applied strain rate, power law viscosities or
81: the emergence of a yield stress, are common features
82: of complex fluids.~\cite{larson99}
83: %and can be roughly summarized as the so-called
84: %Herschel-Bulkeley relation $\sigma=\sigma_y+\dot\epsilon^n$ between stress
85: %and strain rate.
86: MD simulations of Lennard-Jones systems~\cite{berthier01,yamamoto97,yamamoto98}
87: and recent experiments on colloidal glasses~\cite{bonn02a,bonn02b}
88: are also consistent with such properties.
89:
90: Recent theoretical approaches to non-linear
91: rheology~\cite{sollich97,sollich98,hebraux98,berthier00,fielding00}
92: have called on ideas originating in mode coupling theory,
93: or trap models of spin glasses.~\cite{goetze92,bouchaud96,monthus96}
94: These approaches, however, fail to provide a description of non-linear
95: rheology that is consistent with experimental observations;~\cite{bonn02a,bonn02b}
96: moreover, these models, based only on a phase-space picture of glassiness,
97: do not lead to a clear identification of real-space physical mechanism
98: that are responsible for an observed macroscopic behavior.
99: Some other attempts have been directed towards phenomenological
100: approaches,~\cite{derec01} but those are still unsatisfactory,
101: and do not offer a microscopic picture, even heuristic, of the undergoing physical
102: mechanisms.
103:
104: The current work lies at the convergence point of several recent theoretical
105: ideas in various fields.
106: Firstly, it is based on the idea by Liu, Nagel and coworkers,
107: that some unique mechanism lies
108: behind the profound similarities displayed by structural glassy materials,
109: be they colloidal glasses, granular materials, foams,\dots~\cite{liu98}
110: %The observed similarity of the force distribution between various structural systems
111: %indicates that the structure of a material at low temperature is essentially
112: %dominated by large forces between molecules: relaxation processes are
113: %therefore expected to be controlled by the force network which acts
114: %as a mechanical skeleton.~\cite{ohern01}
115: Drawing on this idea, the fundemental assumption underlying the
116: current work is that macroscopic
117: properties of dense materials results from major {\it dynamical} properties
118: of structural rearrangement of the contact network.
119:
120: Another important element of the current work comes from the study
121: of density relaxation
122: in granular materials.~\cite{knight95,boutreux97,nowak98}
123: Under vertical tapping, granular materials present a time-logarithmic relaxation
124: of the occupied volume; this is in strong analogy with free-volume relaxation,
125: commonly observed in glasses, and of central importance for aging.~\cite{larson99}
126: However, in the case of tapped granular materials,
127: density relaxation cannot result from thermal activation
128: since there is no thermal bath:
129: the logarithmic density relaxation has been
130: explained by introducing an equation of motion for free-volume,
131: which characterizes the internal state of the granular material.
132:
133: Finally, this work uses a theory introduced recently by Falk and Langer
134: to account for elasto-plastic transition in amorphous solids.~\cite{falk98,falk00}
135: The so-called shear transformation zone (STZ) theory,
136: originates from ideas by Argon, Spaepen and others
137: to describe creep in metallic
138: alloys;~\cite{spaepen77,argon79a,argon79b,spaepen81,argon83}
139: it has been primarily developed in the framework of fracture mechanics,
140: and provides a general scheme for jamming as resulting from a structuration
141: of the material by creation of arrangements oriented along principal directions
142: of an applied stress.
143: STZ theory calls to free-volume arguments, but in fact,
144: free-volume remains a parameter and enters constants of the theory.
145:
146: These ideas lead me to propose a general approach to the rheology of
147: dense materials, leading to a simple set of constitutive equations.~\cite{lemaitre01b}
148: The purpose of the this article is to elaborate the theoretical arguments
149: underlying these equations a to detail some of their consequences.
150:
151: Two types of internal state variables are used: free-volume,
152: which is an isotropic property of a molecular packing, and populations of arrangements,
153: which are related to the distortion of the contact network between molecules
154: (hence introduce anisotropy).
155: Those variables evolves dynamically by free-volume activated processes:
156: two types of rearrangements are associated to compaction and shear motion.
157: Free-volume enters activation factors that control both types
158: of elementary rearrangements, and it is also a dynamical quantity.
159:
160: KWW relaxation has been shown to result from
161: the existence of a distribution of timescales;~\cite{larson99,angell00}
162: and it seems that this is often thought to be a necessary requirement.
163: The first contribution of the current work is to show that it is not necessary,
164: and that KWW relaxation also result from a more simple mechanism:
165: %shear motion and free-volume relaxation result from different types
166: %of collective rearrangements, hence are activated processes with
167: %different entropy barrier heights;
168: %KWW results directly from
169: the interplay between free-volume logarithmic relaxation and
170: free-volume activated shear deformation.
171:
172: The second contribution of this work is to establish a relation
173: between KWW relaxation, time-logarithmic density relaxation,
174: and power law viscosity of a strongly sheared glassy material.
175: A single parameter of the theory determines these three types of phenomena.
176: Power law viscosity will be shown to result from
177: the existence of shear induced dilatancy.
178: I propose one of the simplest set of equations that can account
179: for those three properties is a single framework.
180:
181: The third contribution of this work is to present a theoretical framework
182: which accounts for two coexisting mechanisms for jamming.
183: In mode-coupling theory, jamming results from an entropy crisis:
184: at low temperature, the system supports any applied stress because
185: no motion of molecules is allowed.
186: Resistance to shear is therefore an intrinsic property of a material,
187: determined by the thermodynamic parameters of the system,
188: and results from a phase transition.
189: In STZ theory, jamming results from a structuration
190: of the contact network:
191: jamming requires some amount of plastic deformation
192: to induce the creation of anisotropic structures that
193: can support a static stress.
194: Those two mechanisms are shown here to be incorporated in a unique theory.
195: The difference between hard and soft glassy materials
196: will be identified with the cases when one or the other process dominates.
197: The interplay between those two mechanisms
198: lead to a novel interpretation of brittleness and ductility,
199: and provide age-dependent dynamical yield criteria.
200:
201: \section{Constitutive equations}
202: \subsection{Preliminary settings}
203: %\subsubsection{A mean-field approach}
204:
205: The main assumption made in this work is that the out-of-equilibrium
206: dynamics of a glass can be modeled in the framework of a mean-field approximation.
207: By mean-field, I mean that spatially averaged quantities account
208: for the local state of the material.
209: Constitutive equations are proposed that relate the average shear stress,
210: to the average shear strain.
211: The internal state of the material is determined by free-volume
212: and by densities of arrangements (to be defined further).
213: %Since the distribution of voids in the material
214: %is characterized by the average free-volume only,
215: %an additional Ansatz will be required for this distribution.
216:
217: %\subsubsection{Stress, strains and inherent states}
218:
219: The forces in the material are supposed to be characterized by a stress tensor,
220: which is written
221: $$
222: \left(
223: \matrix{ -P& \sigma\cr
224: \sigma &-P\cr}
225: \right)
226: $$
227: where $P$ is the pressure, and $\sigma$ the shear stress.
228: The deformation tensor is defined accordingly, and composed of shear,
229: and of isotropic deformations, directly related to dilatancy.
230: The average free-volume per molecule is denoted $v_f$,
231: and is related to the total volume $V$ by,
232: $$
233: V=N v_f + N v^{\rm rcp}
234: \quad,
235: $$
236: where $v^{\rm rcp}$ is the volume per molecule of the material in random close packing,
237: and $N$ is the number of molecules.
238:
239: Under strong shear, soft glassy materials, like clays, foams, or granular materials,
240: can achieve large deformations; the evolution of the system
241: is more similar to the flow of a liquid than to the deformation of a solid.
242: In this work I present a theory of deformation that is expected to hold
243: for large deformations of amorphous solids or soft glassy materials.
244: Since there is a long-standing matter of debate in the mechanical community
245: about linear versus non-linear deformations,~\cite{lubliner90}
246: it seems necessary to clarify this issue.
247: I show here that the notion of inherent states~\cite{stillinger84}
248: offers a simple physical interpretation of small and large deformations.
249:
250: At low temperature, a supercooled liquid or a glassy material evolves in a complex
251: energy landscape which can be partitioned in domains of influence of local minima.
252: For almost all geometrical configuration of the molecules, a unique inherent state
253: is defined as the local minimum to which the systems relaxes
254: if suddenly quenched to zero temperature,~\cite{stillinger84}
255: and in the absence of forcing.
256: When a material is forced, a constant stress or strain pulls the system
257: through the energy landscape.
258: Each geometrical configuration reached lies in the domain associated with
259: an instantaneous inherent state.
260: The time series of instantaneous inherent states provides a coarse-grained
261: description of the deformation of the material, and is taken here as the definition
262: of irreversible, plastic, deformation.
263:
264: The instantaneous inherent state is a natural reference state for
265: a given molecular configuration, and deviation from the instantaneous inherent state
266: defines an {\it inherent (shear) deformation} $\epsilon^{\rm in}$.
267: Starting from a known initial configuration, the total shear deformation,
268: $\epsilon= \epsilon^{\rm in} + \gamma$, results both from inherent deformation,
269: and from the {\it flow} of inherent states in the phase space.
270: %, determined by the rate of plastic deformation $\dot\gamma$.
271: By definition, the deformation $\gamma$ is measured between two inherent states:
272: it is not a state variable, since it is defined only from the knowledge
273: of a configuration designed as ``initial''. $\epsilon^{\rm in}$, however, is a state
274: variable, since it is well defined for any instantaneous configuration of the system.
275: Moreover, $\epsilon^{\rm in}$ is expected to be small, since it
276: measures the deformation between two nearby configurations in phase space;
277: elasticity results from a harmonic approximation for every local minimum:
278: it permits to write the stress as given by a Hooke law,
279: $\sigma=\mu\,\epsilon^{\rm in}$.
280: The equation that governs $\sigma$ can finally be written:
281: \begin{equation}
282: \dot\sigma = \mu\,(\dot\epsilon-\dot\gamma)
283: \quad.
284: \label{eqn:sigma:0}
285: \end{equation}
286: The rate of plastic deformation $\dot\gamma$ will come out of a statistical analysis
287: of hopping motion in the phase-space.
288: It will be written as a function of the state variables of the system,
289: $\sigma$, $v_f$, an others, to define the constitutive law of the material.
290:
291: \subsection{Assumptions}
292: \subsubsection{Free-volume}
293:
294: This work relies on the idea that, in dense materials,
295: macroscopic deformation results from free-volume activated rearrangements.
296: This is expected because temperature is so low that thermal activation
297: becomes irrelevant, and activated processes are controlled by entropic fluctuations
298: of the free-volume.
299: A rearrangements occurs if sufficient space exists in the neighborhood
300: of a given point; the transition probabilities are thus
301: determined by the size distribution of voids.
302: In order to work at a mean-field level, the volume distribution is characterized
303: by the single scalar value $v_f$; in order to write activation factors,
304: it is necessary to provide an Ansatz for the distribution of voids in the material.
305:
306: In the original works by Cohen and Turnbull on free-volume theory,
307: it was argued that the voids in a material are given by
308: a Poisson distribution.~\cite{cohen59,turnbull61,turnbull70}
309: In their picture, the motion of a molecule in a glassy material is permitted only
310: if it can move into a hole of size, say $v_1$.
311: The probability to find a hole larger than $v_1$, is proportional to $\exp(-v_1/v_f)$:
312: this leads to estimate a diffusion constant $D\propto\exp(-v_1/v_f)$,
313: which is a basis of free-volume approaches.
314: Important differences, however, exist between various
315: uses and interpretations of free-volume theory.
316:
317: Falk and Langer use free-volume activation factors in the definition
318: of STZ theory.~\cite{falk00}
319: Their argument borrowed from original ideas by Edwards and coworkers
320: for granular materials:~\cite{edwards89,mehta89,edwards94}
321: the only extensive variable is supposed to be the volume $V$;
322: the number of states available to the system
323: is roughly proportional to $(v_f/h)^N$, where $h$ is an arbitrary constant
324: with dimension of a volume; the entropy is defined as,
325: $$
326: S(V,N) \simeq N\,\ln\left({v_f\over h}\right) = N\,\ln\left({V- N v^{\rm rcp}\over N h}\right)
327: $$
328: and the intensive quantity, analogous to temperature, is $\chi$:
329: $$
330: {1\over\chi}\equiv {\partial S\over\partial V} = {1\over v_f}
331: \quad.
332: $$
333: Free-volume thus enters activation factors of the form $\exp(-v_1/v_f)$.
334:
335: In STZ theory, this discussion turns out to be somewhat formal
336: since free-volume is taken constant:
337: $v_f$ is absorbed in the phenomenological constants of the resulting equations.
338: It also leads to complicated expressions for transition rates as functions
339: of the applied stress (because $v_1$ is supposed to be a function of $\sigma$).
340: %At this stage, however, there is no profound reason to argue of the precise
341: %analytical expression given for those rates and I expect the physics of dense
342: %materials to come out from rather general features.
343:
344: However, free-volume should vary when the material dilates or contracts.
345: This idea has been considered recently by several authors in the analysis
346: of density relaxation of granular materials subjected
347: to vertical tapping.~\cite{knight95,boutreux97,nowak98}
348: In these works, free-volume is understood as a purely dynamical quantity,
349: and relaxes by the motion of single grains in holes created by volume fluctuations.
350: A Poisson distribution of volume fluctuations is assumed at all times:
351: this leads to an equation of motion of the form,
352: \begin{equation}
353: \dot v_f = -R_1\,\exp\left[{-{v_1\over v_f}}\right]
354: \quad.
355: \label{eqn:vf:0}
356: \end{equation}
357: The activation factor $\exp\left[{-{v_1/v_f}}\right]$ is the probability that
358: a hole of size larger than $v_1$ is present at a given point in the material;
359: $R_1$ is determined by the tapping frequency.
360:
361: The current work borrows from those different lines of thought.
362: Drawing on the works initiated on granular compaction, free-volume is a dynamical
363: quantity; voids are expected to redistribute fast enough in the material;
364: fast means, faster than free-volume activated collective rearrangements themselves.
365: This is justified because redistribution of free-volume requires very tiny
366: displacements of the molecules, and voids, considered as
367: particles, are expected to diffuse faster than molecules themselves:
368: around a void, there is by definition an excess of free-volume.
369:
370: There is an important difference, however, from the picture used
371: in~\cite{knight95,boutreux97,nowak98}
372: where compaction results from the motion of a single grain in a hole.
373: Here, borrowing on the argument of STZ theory,
374: compaction results from elementary rearrangements, involving several
375: molecules at a mesoscopic scale.
376: Probabilities of transitions are estimated by entropic arguments.
377: Due to the redistribution of voids, the number of states available
378: to the system is supposed to be proportional to $(v_f/h)^N$ at all times:
379: the fast redistribution of voids allows the system to
380: realize a quasi-equilibrium thermodynamic ensemble,
381: determined by an intensive variable, $v_f$.
382:
383: The existence of an intensive variable, associated to
384: the fluctuations of molecular configurations, is supported
385: by the observation of an effective fluctuation-dissipation theorem
386: in sheared fluids.~\cite{berthier01}
387: For reasons of clarity, and to be specific,
388: I reserve the word temperature for the intensive quantity associated
389: with the fluctuations of energy, or for the specific kinetic energy;%~\cite{langer_priv}
390: therefore, I will not comply with to the recent use of ``effective temperature''
391: for something that is neither thermodynamic nor granular temperature.
392: Nevertheless, the observation of an effective fluctuation-dissipation theorem
393: indicates that such an intensive quantity exists, without identifying
394: what this quantity actually is.
395: Here, it is identified to $v_f$ and is associated with entropic fluctuations of
396: molecular configurations at a mesoscopic scale.
397:
398: The important novelty introduced in the current work it that
399: $v_f$ enters activation factors and is also a dynamical quantity.
400: This is in weak analogy with granular temperature, which is a temperature,
401: and evolves dynamically.
402: But granular temperature couples with properties of the material only as far as
403: it determines a collision frequency,~\cite{savage79,lemaitre01a}
404: while free-volume enters exponential prefactors in transformation rates.
405: Another important element of the current theory is that the dynamics of
406: free-volume results both from density relaxation by activated rearrangements
407: -- which exists in the absence of forcing --
408: and from dilatancy induced by the macroscopic shear flow
409: -- which occurs when the system is driven out-of-equilibrium.
410: The non-linear coupling with the mean-flow will be addressed later;
411: I will first continue the description of rearrangement processes.
412:
413: \subsubsection{Two types of activated processes}
414: The typical elementary compaction process,
415: corresponds to the system going through a saddle point in phase space.
416: The rate of such transformations is controlled by free-volume fluctuations
417: and this saddle is characterized by a typical activation volume $v_1$,
418: which depends on the shape of the molecules,
419: and on the details of microscopic interactions.
420: The activated hopping of the system through this type of saddle
421: permits density relaxation,
422: and leads to an equation of motion for $v_f$ of the form~(\ref{eqn:vf:0}).
423:
424: Other types of elementary rearrangements occur in the material,
425: in particular leading to elementary shear motion.
426: Before presenting a more elaborate scheme for macroscopic shear deformation,
427: let me use a very minimal argument:
428: shear strain is supposed to be proportional to shear stress: $\dot\gamma = D\sigma$;
429: where $D$ is proportional to a free-volume activation factor.
430:
431: The current approach relies on the remark that the relative motion of molecules
432: that permit compaction and shear are different;
433: different types of collective rearrangements,
434: should require different activation volumes.
435: In phase space, shear and compaction correspond to two types of saddle points,
436: and those saddle points are entropic barriers of different heights.
437: The activation volume for an elementary shear is denoted $v_0$,
438: and {\it a priori} differs from $v_1$. This leads to
439: \begin{equation}
440: \dot\gamma \propto \exp\left[{-{v_0\over v_f}}\right]\,\sigma
441: \quad.
442: \label{eqn:dotgamma:0}
443: \end{equation}
444:
445: One major purpose of this work is to study the consequences of
446: introducing complexity, not through a non-trivial distribution of
447: timescales,~\cite{angell00,schlesinger84}
448: but through the existence of two types of entropy barriers of unequal heights.
449: The set of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:vf:0}), and~(\ref{eqn:dotgamma:0})
450: will be shown to account for logarithmic relaxation of $v_f$
451: and for a KWW relaxation of the stress after a strain increment.
452:
453:
454: \subsubsection{Shear deformation}
455:
456: \paragraph{Shear transformation zones.}
457: Equation~(\ref{eqn:dotgamma:0}) accounts for a free-volume dependent viscosity:
458: the material described is a liquid (in fact a non-Newtonian liquid).
459: It is however, a very rough description of shear motion.
460:
461: The theoretical description of shear deformation that I present
462: is directly inspired by STZ theory.~\cite{falk98,falk00}
463: The first assumption on which STZ theory is based is that
464: elementary shear deformation does not occur just anywhere in the material,
465: but in zones, where a rearrangement is permitted by the local configuration of
466: molecules.
467: A shear transformation zone (STZ) is thus defined as a locus within a material,
468: where an elementary shear is possible.
469: Another essential assumption is that STZ's are two-state systems.
470: This can be understood as follows:
471: An elementary arrangement corresponds to the opening and closing of some
472: contacts between molecules; once such a rearrangement has occurred
473: somewhere in the material, these molecules cannot shear further in the same direction, but they can shear backwards.
474: A symmetry is thus introduced by shearing motion, and (at least) two types
475: of arrangements can be identified that are transformed into one another,
476: as pictured here:
477: \begin{center}
478: \unitlength = 0.0011\textwidth
479: \begin{picture}(100,60)(0,0)
480: \put(50,40){\makebox(0,0){\large$R_+$}}
481: \put(50,0){\makebox(0,0){\large$R_-$}}
482: \put(-50,0){\resizebox{200\unitlength}{!}{\includegraphics{stz.ps}}}
483: \end{picture}
484: \end{center}
485: Note that an STZ is not supposed to contain only four molecules,
486: but it is a mesoscopic object that accounts for the structure of several
487: neighboring molecules (say of order 10).
488: To simplify the analysis, only one pair of types of arrangements is considered,
489: aligned along the principal directions of the stress tensor.
490: Macroscopic motion results from the statistics of elementary rearrangements,
491: and reads:
492: \begin{equation}
493: \dot\gamma = {\cal A}_0\,(R_+\,n_+-R_-\,n_-)
494: \label{eqn:stz:0}
495: \end{equation}
496: where $n_\pm$ denote the number density of arrangements, and where
497: $R_\pm$ denote the rate of transition $\pm\to\mp$.
498:
499: Those rates result from activation processes determined by free-volume,
500: but also by force fluctuations.
501: A STZ shears if there is sufficient free-volume at its location,
502: and if the local bias of the force network triggers the shear in the appropriate
503: direction.
504: In this work, force and volume fluctuations are supposed to be uncorrelated;
505: the probabilities associated with those two types of fluctuations factorize:
506: $R_\pm = R^v(v_f) R_\pm^\sigma(\sigma)$.
507: From the preceding discussion, $R^v(v_f) = \exp(-v_0/v_f)$.
508: In order to estimate the factors $R_\pm^\sigma$ associated with the fluctuations
509: of the force network, let me first remark that $R_\pm^\sigma(\sigma)$ is a positive,
510: increasing function of $\sigma$.
511: Moreover, an elementary shear is expected to be triggered by {\em large forces}.
512: The distributions of static forces in glassy systems have been shown to
513: share strong similarities with the force distribution in
514: a granular material,~\cite{ohern01} in which case, large forces are
515: distributed exponentially.~\cite{coppersmith96,radjai98a,radjai98b}
516: This justifies an exponential dependency of $R_\pm^\sigma$ on $\sigma$:
517: $$
518: R_\pm = R_0\,\exp\left[{-{v_0\over v_f}}\right]\,\exp\left[{\pm{\sigma\over\bar\mu}}\right]
519: \quad,
520: $$
521: with $R_0$, the update frequency of microscopic processes.
522: Note however, that the exponential form for the activation factors $R_\pm^\sigma$
523: plays no role in the following, since these factors will be linearized.
524: The parameter $\bar\mu$ measures the typical stress that must be overcome
525: to trigger a rearrangement.
526: In hard-sphere systems, it is expected to be proportional
527: to the pressure $P$. In general, $\bar\mu$ is essentially the energy
528: required to break a bond between two molecules, times the density of bonds.
529: Here, $\bar\mu$ will be taken constant, either because $P$ fixed,
530: or because it depends on some interaction potential.
531: At some point, $\bar\mu$ will be set to unity, which fixes the unit of forces.
532:
533: \paragraph{Dynamics of arrangements.}
534:
535: Following~\cite{falk98,falk00}, the equation of motion for the populations $n_\pm$ is written,
536: \begin{equation}
537: \label{eqn:stz:npm}
538: \dot n_\pm = R_\mp n_\mp - R_\pm n_\pm +\sigma\,\ddpl\,({\cal A}_{c} - {\cal A}_{a}\,n_\pm)
539: \quad.
540: \end{equation}
541: The first two terms of the rhs account for the transitions between both
542: types of STZ. The last term accounts for the renewal of
543: molecular configuration by the macroscopic flow:
544: STZ constantly appear and disappear during plastic deformation.
545: The renewal of configurations is supposed to be proportional to
546: the work of plastic deformation, $\sigma\,\ddpl$, which
547: is thus a common factor of creation and annhilitation terms.
548:
549: \paragraph{Shear-induced dilatancy.}
550: In this work, the dynamics of rearrangements is coupled to free-volume dynamics.
551: Part of this coupling results from the factor $R^v(v_f)$ that controls
552: shear transformation, but the macroscopic flow is also expected
553: to influence the relaxation of free-volume.
554:
555: Free-volume dynamics is derived in analogy with the dynamics
556: of the populations $n_\pm$:
557: plastic deformations corresponds to the evolution of the system in the phase-space
558: along trajectories that do not necessarily permit a minimization of free-volume.
559: In real space, plastic deformation requires the constant renewal of local
560: arrangements, and new arrangements are local configurations of large entropy;
561: by definition, their density is not optimized.
562: Therefore, the renewal of configurations by the shear flow leads to dilatancy.
563: In experiments on suspensions, for example, the system is fluidized by
564: applying a strong shear which rejuvenates the glassy structure.~\cite{cloitre00}
565:
566: In order to quantify the average dilatancy induced by the shear flow,
567: I assume that a fraction of the work of plastic deformation, $\sigma\,\dot\gamma$
568: in used for enthalpy production, $P \dot v_f$.
569: The equation for $v_f$ is:
570: \begin{equation}
571: \dot v_f = - R_1\,\exp\left[-{v_1\over v_f}\right] + {{\cal A}_v\over P}\,\sigma\,\dot\gamma
572: \label{eqn:vf:1}
573: \end{equation}
574: This equation is an essential element of the current theory.
575: The non-linear coupling with the macroscopic flow shows up when the system
576: is strongly driven out-of-equilibrium.
577: The competition between self-relaxation
578: and shear induced dilatancy
579: will be shown to lead to a power law viscosity.
580:
581: \subsection{Equations of motion}
582:
583: From~(\ref{eqn:stz:npm}), the equation of motion for the total density,
584: $n_T=n_++n_-$, of arrangements reads,
585: $$
586: \dot n_T = \sigma\,\ddpl\,( 2\,{\cal A}_{c} - {\cal A}_{a}\,n_T)
587: \quad.
588: $$
589: The creation and destruction terms account for an entropic mixing
590: of molecular configurations.
591: The previous equation shows that this mixing leads to a convergence of
592: $n_T$ to an asymptotic value, $n_\infty=2\,{\cal A}_{c}/{\cal A}_{a}$;
593: therefore, $n_T=n_\infty$ defines a state of maximal entropy
594: of molecular configurations.
595: For this reason, I do not expect $n_T$ to be different from $n_\infty$ unless
596: in very special circumstances, {\it e.g.} if some level of crystallization
597: exists in the material.
598: In particular, an amorphous solid is usually produced by a rapid quench
599: of a liquid at low temperature: the system comes from a state of high entropy,
600: therefore, in the resulting glass, $n_T$ is expected to achieve its asymptotic value.
601: In this work, I assume that $n_T=n_\infty$ at all times.
602: The dynamics of local arrangements is thus determined only by the bias $n_--n_+$
603: between the populations of STZ.
604:
605: Equations~(\ref{eqn:stz:0}) and~(\ref{eqn:stz:npm})
606: are written in a more suitable form by introducing the variable
607: $$
608: \Delta = {n_--n_+\over n_\infty}
609: \quad,
610: $$
611: and parameters,
612: $\epsilon_0 = {{\cal A}_0\,{\cal A}_c/{\cal A}_a}$,
613: $\mu_0 = {\cal A}_0\,{\cal A}_c$, and $E_0 = 2\epsilon_0\,R_0$.
614: It is also convenient to introduce $\kappa=v_1/v_0$, $E_1=R_1/v_0$
615: and $\alpha = {\cal A}_v/(P v_0)$, and to rescale free-volume as,
616: $$
617: \chi = {v_f\over v_0}
618: \quad.
619: $$
620:
621: The complete set of constitutive equations is finally:
622: \begin{eqnarray}
623: \label{eqn:stzdil:1}
624: \ddpl&=&E_0\,\exp\left[{-{1\over\chi}}\right]\,
625: \left(\sinh\left({\sigma\over\bar\mu}\right)-\Delta\,\cosh\left({\sigma\over\bar\mu}\right)\right)\\
626: \label{eqn:stzdil:2}
627: \dot\Delta&=&{\ddpl\over\epsilon_0}\,\left(1- \mu_0\,\sigma\,\Delta\right)\\
628: \label{eqn:stzdil:3}
629: \dot\chi &=& -E_1 \exp\left[-{\kappa\over\chi}\right] + \alpha\,\sigma\,\ddpl
630: \quad.
631: \end{eqnarray}
632: Those equations involve two state variables, $\Delta$ and $\chi$ which
633: play very different roles.
634: $\Delta$ evolves only when the material is sheared, therefore accounts
635: for some sort of structural memory induced by deformation imposed on the material.
636: On the contrary, $\chi$ evolves spontaneously, and relaxes towards 0.
637: Since this relaxation is slow, $\chi$ also encodes some sort of memory,
638: but a very different type of memory than $\Delta$,
639: a memory which constantly evolves as the system ages.
640:
641: Note that the asymptotic value for the relaxation of $\chi$ (which here is 0)
642: may, in fact, be temperature dependent;
643: this question is eluded in the current work:
644: it seems a very reasonable assumption for materials like suspensions
645: if not for glasses below the glass transition, because the logarithmic
646: slow-down of density relaxation around $\kappa$ prevents it to reach
647: asymptotic values far below ($\chi_\infty<<\kappa$).
648: The parameters involved in the theory should depend on pressure and temperature,
649: and this dependency results from particularities of a given material;
650: for example the update frequencies $R_1$ and $R_0$ are expected to
651: be proportional to $\sqrt{T}$ (where $T$ is the temperature)
652: in a dense hard-sphere material.~\cite{lemaitre01a}
653: Such questions, however, are not the focus of this study;
654: in this article, the values of the parameters are constant and can
655: be understood as a given thermodynamical state of the system.
656:
657: \section{Interlude}
658: Equations~(\ref{eqn:stzdil:1}-\ref{eqn:stzdil:3}) present non-linear couplings
659: between shear motion and density relaxation.
660: Before studying the complete set of equations in various tests,
661: I start with two simple cases, when one or the other process dominates.
662: I show here that
663: jamming transitions are captured by equations~(\ref{eqn:stzdil:1}-\ref{eqn:stzdil:2}),
664: and free-volume relaxation by equation~(\ref{eqn:stzdil:3}).
665:
666: \subsection{Free-volume relaxation}
667: In the absence of shear stress, $\sigma=0$ and $\dot\gamma=0$;
668: the system~(\ref{eqn:stzdil:1}-\ref{eqn:stzdil:3}) reduces to the single equation:
669: $$
670: \dot \chi = -E_1 \exp\left[-{\kappa\over\chi}\right]
671: \quad,
672: $$
673: which accounts for density relaxation.
674:
675: At short time, this process is dominated by the initial value, $\chi_0$:
676: $$
677: \dot \chi \simeq -E_1 \exp\left[-{\kappa\over\chi_0}\right]
678: \quad;
679: $$
680: this defines a time scale for short-time free-volume relaxation:
681: $$
682: \tau_\chi = {\chi_0\over E_1}\;\exp\left[{\kappa\over\chi_0}\right]
683: \quad.
684: $$
685:
686: The long time behavior is estimated from the integral expression:
687: $$
688: E_1 t = \int_{\chi(t)}^{\chi(0)} \;\exp\left[{\kappa \over \chi}\right] \d \chi \simeq
689: \kappa\,\exp\left[{\kappa\over \chi(t)}\right]
690: $$
691: At long time, the integral is dominated by the small values of $\chi$,
692: {\it i.e.} by the ultimate value reached, $\chi(t)$. Therefore,
693: \begin{equation}
694: \chi(t) \simeq {\kappa\over \log(E_1 t/\kappa)}
695: \quad.
696: \label{eqn:vf:relax}
697: \end{equation}
698: Free-volume relaxes logarithmically with time:
699: in the absence of any forcing, the system ages.
700:
701: \subsection{Jamming}
702: In order to isolate the STZ mechanism for jamming,
703: I now assume that free-volume is constant, at some value $\chi$;
704: the evolution of the system is governed by
705: equations~(\ref{eqn:stzdil:1}) and~(\ref{eqn:stzdil:2}) only,
706: and a constant stress $\sigma$ is applied.
707:
708: These equations admit two stationary solutions:
709: a jammed state for which, $\dot\epsilon=\dot\gamma=0$, and
710: $$
711: \Delta = \tanh\left({\sigma\over\bar\mu}\right)
712: \quad;
713: $$
714: a steady deformation regime, for which,
715: $$
716: \Delta = {1\over\mu_0 \sigma}
717: \quad,
718: $$
719: and,
720: $$
721: \dot\epsilon=E_0\,\exp\left[{-{1\over\chi}}\right]\,
722: \left(\sinh\left({\sigma\over\bar\mu}\right)-{1\over\mu_0 \sigma}\,\cosh\left({\sigma\over\bar\mu}\right)\right)
723: \quad.
724: $$
725: The latter solution is unstable at small stresses, in which case, the material jams.
726: For large stresses, the jammed state loose stability, and steady deformation results.
727: The two solutions exchange stability at the yield stress $\sigma_y$,
728: which is the solution of:
729: $$
730: \tanh\left({\sigma_y\over\bar\mu}\right) = {1\over\mu_0 \sigma_y}
731: \quad.
732: $$
733:
734: \subsection{Various limits}
735:
736: In the following, I will consider several limiting cases of the general constitutive
737: equations~(\ref{eqn:stzdil:1}-\ref{eqn:stzdil:3}).
738:
739: In particular, the exponential dependency of $\ddpl$ as a function of $\sigma$
740: will not be studied, although this non-linearity might be
741: important in some cases.~\cite{saulnier02,bureau02}
742: Activation factors involving $\sigma$ will thus be linearized;
743: this is valid under the assumption that $\sigma<<\bar\mu$;
744: $\bar\mu$ is then incorporated into constants by taking it to unity.
745: It leads to:
746: \begin{eqnarray}
747: \label{eqn:stzdil:lin:1}
748: \ddpl&=&E_0\,\exp\left[{-{1/\chi}}\right]\,\left(\sigma-\Delta\right)\\
749: \label{eqn:stzdil:lin:2}
750: \dot\Delta&=&{\ddpl\over\epsilon_0}\,\left(1- \mu_0\,\sigma\,\Delta\right)\\
751: \label{eqn:stzdil:lin:3}
752: \dot \chi &=& -E_1 \exp\left[-{\kappa/\chi}\right]+\alpha\,\sigma\ddpl
753: \quad.
754: \end{eqnarray}
755:
756: I will also consider an {\em isotropic (liquid) limit},
757: where the structural anisotropy, measured by $\Delta$ plays no role.
758: This is the case either at timescales when the dynamics of $\Delta$ is too slow,
759: $1/\epsilon_0\to0$ (and $\Delta=0$) or because the typical stress $1/\mu_0$ vanishes.
760: In this case, the rheology is determined by a single state variable:
761: \begin{eqnarray}
762: \label{eqn:stzdil:iso:1}
763: \ddpl&=&E_0\,\exp\left[{-{1/\chi}}\right]\,\sigma\\
764: \label{eqn:stzdil:iso:2}
765: \dot \chi &=& -E_1 \exp\left[-{\kappa/\chi}\right]+\alpha\,\sigma\ddpl
766: \quad.
767: \end{eqnarray}
768: The sets of equations~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3})
769: and~(\ref{eqn:stzdil:iso:1}-\ref{eqn:stzdil:iso:2})
770: will be helpful to emphasize the roles of $\Delta$ versus $\chi$.
771: Many aspects of glassiness will be shown to be captured by
772: equations~(\ref{eqn:stzdil:iso:1}) and~(\ref{eqn:stzdil:iso:2}).
773: Jamming, however, is not captured at this level of simplification,
774: and requires the existence of the variable $\Delta$.
775:
776: \subsection{Experimental protocol}
777:
778: A word about the preparation of the sample:
779: at time $t=0$, the system is quenched from a highly
780: mixed state, in which, $\Delta=0$ and $\chi$ is large.
781: The sample then ages during a waiting time, $t_w$ without any forcing.
782: After this relaxation, some test is performed on the system.
783:
784: Two main experimental procedures consist in forcing the stress,
785: or forcing the deformation.
786: If a stress, $\sigma$, is applied, the plastic deformation $\gamma(t)$
787: is determined by constitutive equations,
788: and the total deformation is, $\epsilon = \gamma + \sigma/\mu$.
789: If the deformation, $\epsilon(t)$, is forced, the evolution
790: of $\sigma$ is given by equation~(\ref{eqn:sigma:0}),
791: and couples to constitutive equations.
792:
793: If a small perturbation is applied to the sample,
794: the response of the system results from first order terms in the
795: linearized constitutive equations~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}),
796: or~(\ref{eqn:stzdil:iso:1}-\ref{eqn:stzdil:iso:2}).
797: %The non-linear terms, which introduce a coupling between free-volume relaxation
798: %and shear motion are expected to be negligible.
799: In relaxation processes, however, the linear approximation
800: is not always valid at long times,~\cite{yoshino98,derec01}
801: because, the activation factors can become small,
802: possibly smaller than non-linear terms.
803: In the current theory, all non-linear terms involve the work of plastic
804: deformation $\sigma\dot\gamma$. Therefore, the linear analysis is valid
805: so long as this work is small compared to relaxation processes.
806:
807: \section{Stress relaxation after a strain increment}
808: Stress relaxation is monitored after a small strain step.
809: The step can be understood as a ramp in the deformation, which occurs
810: on a very small time interval around $t_w$.
811: The strain $\epsilon$ determines the initial value of
812: the stress, $\sigma(t_w^+)=\mu\,\epsilon$.
813: During the relaxation, $\dot\epsilon=0$;
814: the subsequent evolution of the stress is governed by $\dot\sigma=-\mu\dot\gamma$,
815: and accounts for the plastic adaptation of the material to the imposed strain.
816:
817: The relaxation modulus, $G(t)=\sigma(t)/\sigma(t_w)$ is the correlator associated
818: with rearrangement processes: it measures the correlation between the
819: initial state of the material, and its state at any following time, $t$.
820: Assuming that the diffusion of molecules is dominated by collective rearrangements,
821: stress relaxation can then be compared with the decay of
822: self-intermediate scattering functions, which measure the spatial
823: decorrelation resulting from the diffusion of particles in the dense medium.~\cite{boon91}
824: The long-time behavior of correlation functions is one of the major
825: properties of glassy materials.~\cite{goetze92}
826: (See also recent reviews by Debenedetti and Stillinger,~\cite{debenedetti01}
827: or Angell {\it et al}.~\cite{angell00}).
828:
829: %For a small strain step, the non-linear term in~(\ref{eqn:stzdil:lin:3})
830: %or~(\ref{eqn:stzdil:iso:2}) is negligible;
831: %at long times, the logarithmic relaxation of $\chi$ is given by~(\ref{eqn:vf:relax}),
832: %and controls the timescale of stress relaxation.
833:
834: \subsection{Isotropic limit}
835:
836: In the isotropic limit, the constitutive equations
837: are~(\ref{eqn:stzdil:iso:1}) and~(\ref{eqn:stzdil:iso:2}).
838: Stress relaxation is governed by the linearized equations:
839: \begin{eqnarray*}
840: \dot\sigma &=& -\mu\,E_0\,\exp\left[-{1\over\chi}\right]\sigma\\
841: \dot\chi &=& -E_1 \exp\left[-{\kappa\over\chi}\right]
842: \quad.
843: \end{eqnarray*}
844: To determine the relaxation modulus, $G=\sigma/\sigma(t_w)$,
845: the only relevant initial condition is the initial value of $\chi$,
846: when the strain is applied; this value is also determined by the initial free-volume
847: $\chi_0$ at time $t=0$ and by the waiting time $t_w$.
848: Moreover, either $E_1$ or $\mu\,E_0$ can be incorporated in the time scale,
849: therefore, only two effective parameters determine all the possible outcomes
850: of those equations: $\kappa$, and $\mu\,E_0/E_1$. While $\kappa$
851: measures the relative heights of entropy barriers, $\mu\,E_0/E_1$ measures
852: the relative update frequencies of shear and compaction processes.
853:
854:
855: \subsubsection{Long time relaxation}
856:
857: At long time, the relaxation of $\chi$ is given by~(\ref{eqn:vf:relax}),
858: and stress relaxation verifies:
859: $$
860: \dot\sigma = -\mu\,E_0\,\exp\left[-{1\over\chi}\right]\sigma
861: \simeq -\mu\,E_0\, \left({E_1\,t\over\kappa}\right)^{-1/\kappa}\sigma
862: \quad.
863: $$
864:
865: The response modulus $G(t) = \sigma(t)/\sigma(t_w)$ is,
866: $$
867: G(t) \simeq \exp\left[{A\,\left(t_w^\beta-t^\beta\right)}\right]
868: $$
869: with an exponent $\beta$ which is directly related
870: to the ratio $\kappa$ between the heights of entropy barriers:
871: $$
872: \beta = 1-{1\over\kappa}\quad {\rm and}\quad A={\mu\, E_0\over\beta}
873: \,\left({E_1\over\kappa}\right)^{-1/\kappa}
874: \quad.
875: $$
876:
877: The validity of linear approximation is checked by considering the time-dependency
878: of the work of plastic deformations $\sigma\dot\gamma$;
879: this term should be compared with the first term is free-volume relaxation,
880: $$
881: E_1\,\exp\left[-\kappa/\chi\right]\propto {\kappa\over t}
882: \quad.
883: $$
884: The stress relaxation, leads to the following evolution for the work of plastic
885: deformations:
886: $$
887: \sigma\dot\gamma \propto t^{-{1/\kappa}}\;\exp\left[-2At^\beta\right]
888: \quad.
889: $$
890: If $\kappa>1$, or $\beta>0$, the relaxation of $\sigma\dot\gamma$ is dominated
891: by the exponential factor;
892: if $\kappa<1$, the stress goes to a constant, but $t^{-{1/\kappa}}$ decays
893: faster than $1/t$.
894: In all cases, the non-linear term decays faster than the linear term,
895: which justifies the linear approximation at all times.
896:
897: For $\beta>0$, {\it i.e.} $\kappa>1$, the stress undergoes KWW relaxation.
898: This property results directly from the existence of two
899: types of saddle points of unequal heights,
900: when hopping motion is controlled by a logarithmically relaxing free-volume.
901: Measured values of exponent $\beta$ in glassy materials,
902: ranging from $0.2$ to $0.5$~\cite{larson99,angell00}
903: are consistent with, $\kappa=v_1/v_0\in[1.25,2]$.
904: The ratio $\kappa$ is related to the shape of molecules,
905: and to the details of their interactions.
906: There is no {\it a priori} reason, from a phase space point of view,
907: to assume a particular value for $\kappa$. In this work I will systematically
908: study both cases in order to draw a complete picture of the phenomenology
909: described by the proposed constitutive equations.
910:
911: If $\kappa<1$, then $\beta<0$, and
912: the previous calculation shows that the stress assumes a non-vanishing
913: asymptotic value after relaxation.
914: In this case, $A<0$, and
915: the stress relaxation towards its ultimate constant value can
916: be further estimated as
917: a power law:
918: \begin{eqnarray*}
919: G(t)
920: &\simeq& \exp\left[ |A|\,\left({t^{-|\beta|}}-{t_w^{-|\beta|}}
921: \right)\right]\\
922: &\simeq& \exp\left[-|A|\,t_w^{-|\beta|}\right]\;\left(1+|A|\,t^{-|\beta|}\right)
923: \quad.
924: \end{eqnarray*}
925: %The existence of a non-vanishing stress at long times is due to the fast
926: %free-volume relaxation, faster than elementary shear motion, on a logarithmic scale:
927: %the voids in the material disappear faster than plastic deformation is permitted,
928: %the system is trapped in a state where some residual shear exists.
929: Interpreting $G$ as a correlation function, it also means that the material
930: never completely decorrelates from its initial state.
931: This non-ergodicity is an indication of the glass transition as understood
932: in mode-coupling theory~\cite{goetze92};
933: it can be characterized by the asymptotic value $G(t\to\infty)$,
934: the Edwards-Anderson parameter:
935: $$
936: G(t\to\infty) \simeq \exp\left[{-|A|\,t_w^{-|\beta|}}\right]
937: \quad,
938: $$
939: which increases with the age of the sample.
940:
941: Such a dependency of the Edwards-Anderson parameter
942: of the age of the sample has been observed experimentally by Bonn and coworkers
943: on a colloidal glass at very low concentrations.~\cite{bonn99}
944:
945: \subsubsection{Short time response}
946:
947: At short time, if free-volume can be assumed large or constant,
948: the exponential term $\exp\left[-{1/\chi}\right]$ does not vary very much,
949: and stress undergoes an exponential relaxation:
950: $$
951: G(t) =
952: \exp\left[-\mu\,E_0\,\exp\left[-{1\over\chi_0}\right]\;\left(t-t_w\right)\right]
953: \quad.
954: $$
955: This relaxation occurs on a time scale of order,
956: $$
957: \tau_\sigma = {1\over\mu\,E_0}\,\exp\left[{1\over\chi_0}\right]
958: \quad,
959: $$
960: which should be compared with the time scale of free-volume relaxation, $\tau_\chi$.
961: If $\tau_\sigma>>\tau_\chi$, free-volume relaxes faster than stress,
962: and the long time, free-volume controlled, stress relaxation can be observed.
963: If $\tau_\sigma<<\tau_\chi$, the stress relaxes exponentially before free-volume
964: achieves sufficiently low values to slow down stress relaxation.
965: The crossover between exponential and KWW relaxation is determined by,
966: $\tau_\sigma\sim\tau_\chi$ or,
967: $$
968: {1\over\mu\,E_0}\,\exp\left[{1\over\chi_0}\right] \sim
969: {\chi_0\over E_1}\;\exp\left[{\kappa\over\chi_0}\right]
970: \quad.
971: $$
972: For a large initial free-volume, this reduces to,
973: $$
974: E_1 \sim \mu\,E_0\,\chi_0
975: \quad;
976: $$
977: KWW relaxation can be observed for a larger value of $E_1$,
978: or for sufficiently small initial value of the free-volume, $\chi(t_w)$,
979: (that is for sufficiently large waiting time $t_w$).
980:
981: In order to look at the short and long time response of the system,
982: equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1}) and~(\ref{eqn:stzdil:iso:2})
983: are integrated numerically with a Runge-Kutta algorithm with adaptive timestep.
984: The system is prepared in a dilute state, $\chi=\chi_0$, which is allowed to relax
985: during time $t_w$, in the absence of any forcing ($\sigma([0,t_w])=0$).
986: At time $t_w$, a strain $\epsilon=10^{-3}$ is applied and produces an
987: elastic response,
988: $\sigma(t_w^+)=\mu\epsilon$. The dynamics of free-volume and stress are then monitored.
989:
990: \paragraph{Crossover stress.}
991: In a first series of tests, $t_w$ is set to 0,
992: and the relaxation spectra are displayed for different values of the parameter $E_1$.
993: Parameters $E_0$, $\mu$ and $\alpha$ are set to 1; $\kappa=2$ and $\chi_0=20$.
994: The results are displayed figure~\ref{fig:relax1}.
995: For the first value $E_1=1$, only exponential relaxation is observed
996: because all the stress has been released before $\chi$ entered the logarithmic
997: relaxation.
998: The value $E_1=20$ corresponds to the crossover $\mu\,E_0\,\chi_0$:
999: KWW relaxation is observed at later stages of the relaxation,
1000: with a non-vanishing values of $G$.
1001: %Figure~\ref{fig:relax1} shows that the plateau in the relaxation spectrum
1002: %is enhanced by a high value of $E_1$.
1003: %In fact, since one of the constants $E_1$ and $\mu E_0$ can be eliminated
1004: %by a rescaling of time,
1005: %the qualitative dynamics depends on the ratio $\mu E_0/E_1$ only.
1006: %The time to crossover is determined by free-volume relaxation.
1007: For a given $\chi_0$,
1008: the stress release during the exponential part of the relaxation is
1009: proportional to $\mu\,E_0$, hence inversely proportional to $E_1$.
1010: The plateau in the relaxation spectrum can be observed only
1011: for sufficiently large values of $E_1/(\mu\,E_0)$ so that $G$ is still observable
1012: by the time $\chi$ enters the logarithmic relaxation.
1013:
1014: These relaxation curves are very similar to those observed in
1015: MD simulations,~\cite{kob95,angell00,debenedetti01}
1016: or in experiments;~\cite{megen94}
1017: the general appearance of those curves depends sensibly
1018: on the parameters considered.
1019: From a practical point of view, it means that it depends
1020: on the time window that is available, numerically or experimentally.
1021: \begin{figure}
1022: \narrowtext
1023: \begin{center}
1024: \unitlength = 0.005\textwidth
1025: \begin{picture}(100,110)(5,0)
1026: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{relax1.eps}}}
1027: %\put(0,0){\line(1,0){100}}
1028: %\put(0,0){\line(0,1){100}}
1029: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1030: \put(5,50){\makebox(0,0){\large $G$}}
1031: \put(90,2){\makebox(0,0){\large $\log_{10}(t)$}}
1032: \end{picture}
1033: \end{center}
1034: \caption{\label{fig:relax1}
1035: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1036: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain $\epsilon=10^{-3}$.
1037: Parameters are $E_0=\alpha=\mu=1$, and $\kappa=2$.
1038: The initial stress is set to $\sigma(t_w)=\mu\epsilon=10^{-3}$,
1039: and $\chi_0=20$, $t_w=0$.
1040: Top: $1/\chi$ as a function of $\log_{10}(t)$
1041: for $E_1 = 1, 20, 40, 80, 100$ from right to left.
1042: Bottom: relaxation spectra with $E_1$ increasing from left to right.
1043: }
1044: \end{figure}
1045:
1046: %Such curves are reminiscent of intermediate scattering functions as
1047: %displayed in many works on colloidal suspensions~\cite{vanmegenm}
1048: %or in recent experiments on Laponite.~\cite{bonn98}
1049:
1050: \paragraph{Aging}
1051: The crossover between exponential relaxation can also be studied by varying
1052: the initial free-volume. This amounts to varying the waiting time $t_w$
1053: for a given $\chi(t=0)$.
1054: In order to observe a plateau in the relaxation spectrum, parameter $E_1$ is set to 20,
1055: and parameters $E_0$, $\mu$ and $\alpha$ are set to 1; $\kappa=2$.
1056: In order to obtain an exponential relaxation for short waiting times,
1057: and to observe a wide variety of behavior, the initial value of $\chi$ is set to 100.
1058: The results are displayed figure~\ref{fig:relax2}. With this choice of parameter,
1059: the curves resemble experimental data.~\cite{angell00,gopal99,hebraud97,megen98,cipelletti00}
1060:
1061: Such curves strongly depend on the available experimental window, and
1062: on the width of the crossover region which is unobservable in a log-lin plot for
1063: smaller values of $E_1/(\mu\,E_0)$.
1064: To emphasize this point,
1065: another set of relaxation spectra is presented figure~\ref{fig:relax}
1066: for values of the parameters: $E_0=\alpha=\mu=1$, $E_1=8$, and $\kappa=2$.
1067: The initial condition is $\chi(0)=100$ and waiting times are:
1068: $t_w = 0, 5, 10, 15, 50, 100$. The relaxation spectra are presented in log-lin,
1069: log-log, and log-log(log) plot; the crossover between exponential and KWW relaxation
1070: is observable around values of the relaxation modulus
1071: of order $10^{-5}$, which is featureless in a log-lin plot.
1072: \begin{figure}
1073: \narrowtext
1074: \begin{center}
1075: \unitlength = 0.005\textwidth
1076: \begin{picture}(100,110)(5,0)
1077: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{relax2.eps}}}
1078: %\put(0,0){\line(1,0){100}}
1079: %\put(0,0){\line(0,1){100}}
1080: %\put(100,0){\line(1,0){100}}
1081: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1082: \put(5,50){\makebox(0,0){\large $G$}}
1083: \put(90,2){\makebox(0,0){\large $\log_{10}(t-t_w)$}}
1084: \end{picture}
1085: \end{center}
1086: \caption{\label{fig:relax2}
1087: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1088: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain $\epsilon=10^{-3}$.
1089: Parameters are $E_0=\alpha=\mu=1$, $E_1=20$, and $\kappa=2$.
1090: Initial conditions are obtained by a quench at time $t=0$ from a dilute state
1091: ($\sigma(0)=0$ and $\chi(0)=100$) followed by density relaxation without forcing during
1092: time $t_w$; at time $t_w$ the stress is set to $\sigma(t_w)=\mu\epsilon=10^{-3}$.
1093: Top: $1/\chi$ as a function of $\log_{10}(t-t_w)$
1094: with $t_w = 0, 5, 10, 15, 50, 100$ from bottom to top.
1095: Bottom: relaxation spectra for increasing $t_w$ from left to right.
1096: }
1097: \end{figure}
1098: \begin{figure}
1099: \narrowtext
1100: \begin{center}
1101: \unitlength = 0.005\textwidth
1102: \begin{picture}(100,110)(5,0)
1103: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{relax.eps}}}
1104: %\put(0,0){\line(1,0){100}}
1105: \put(5,95){\makebox(0,0){\large $G$}}
1106: \put(5,50){\makebox(0,0){\large $\log_{10}(G)$}}
1107: \put(90,2){\makebox(0,0){\large $\log_{10}(t-t_w)$}}
1108: \end{picture}
1109: \end{center}
1110: \caption{\label{fig:relax}
1111: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1112: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain $\epsilon=10^{-3}$.
1113: Parameters are $E_0=E_1=\alpha=\mu=1$, and $\kappa=1.5$.
1114: Initial conditions are obtained by a quench at time $t=0$ from a dilute state
1115: ($\sigma(0)=0$ and $\chi(0)=10$) followed by density relaxation without forcing during
1116: time $t_w$; at time $t_w$ the stress is set to $\sigma(t_w)=\mu\epsilon=10^{-3}$.
1117: Top: $1/\chi$ as a function of $\log_{10}(t-t_w)$
1118: with $t_w = 0, 5, 10, 50, 100$ from bottom to top.
1119: Bottom: relaxation spectra for increasing $t_w$ from bottom to top.
1120: There is no inconsistency in the large logarithmic scale for $\log_{10}(G)$:
1121: the scale is determined up to the reduced parameter $A$.
1122: Inset: $\log_{10}(-\log_{10}(G))$ as a function of $\log_{10}(t-t_w)$,
1123: with $t_w$ increasing from top to bottom.
1124: The spectra with the longest waiting time are already in the KWW regime.
1125: }
1126: \end{figure}
1127:
1128:
1129: \paragraph{Ergodic to non-ergodic transition.}
1130: I now consider the dependency of relaxation spectra on the parameter $\kappa$.
1131: This parameter is of primary importance in the current approach and
1132: is directly related to the existence of entropy barriers of unequal heights
1133: in the phase space.
1134: The parameters are taken to be, $E_0=\alpha=\mu=1$, $E_1=10$,
1135: the initial condition is $\chi(0)=100$, and the waiting
1136: is set to $t_w=10$. This choice is primarily motivated to obtain easily
1137: observable features on the relaxation spectrum.
1138: The results are displayed figure~\ref{fig:relax3}.
1139: When $\kappa$ goes to 1, the system presents a transition to non-ergodic behavior.
1140: For $\kappa<1$, the asymptotic value of the stress is no longer vanishing.
1141: At the point $\kappa=1$, the long time relaxation of $\sigma$ verifies:
1142: $$
1143: \dot\sigma = -\mu\,E_0\,\exp\left[-{1\over\chi}\right]\sigma
1144: \simeq -\mu\,{E_0\over E_1\,t}\sigma
1145: \quad,
1146: $$
1147: whence an asymptotic power law relaxation:
1148: $$
1149: G(t) \simeq \left({t\over t_w}\right)^{-{\mu\,E_0/ E_1}}
1150: \quad.
1151: $$
1152:
1153:
1154: \begin{figure}
1155: \narrowtext
1156: \begin{center}
1157: \unitlength = 0.005\textwidth
1158: \begin{picture}(100,110)(5,0)
1159: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{relax3.eps}}}
1160: %\put(0,0){\line(1,0){100}}
1161: %\put(0,0){\line(0,1){100}}
1162: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1163: \put(5,50){\makebox(0,0){\large $G$}}
1164: \put(90,2){\makebox(0,0){\large $\log_{10}(t-t_w)$}}
1165: \end{picture}
1166: \end{center}
1167: \caption{\label{fig:relax3}
1168: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1169: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain $\epsilon=10^{-3}$.
1170: Parameters are $E_0=\alpha=\mu=1$ and $E_1=10$; $\kappa$ varies.
1171: Initial conditions are obtained by a quench at time $t=0$ from a dilute state
1172: ($\sigma(0)=0$ and $\chi(0)=100$) followed by density relaxation without
1173: forcing during time $t_w$;
1174: at time $t_w$ the stress is set to $\sigma(t_w)=\mu\epsilon=10^{-3}$.
1175: Top: $1/\chi$ as a function of $\log_{10}(t-t_w)$
1176: with $\kappa = 1, 1.2, 1.4, 1.6, 1.8, 2$ from top to bottom.
1177: Bottom: relaxation spectra for increasing $\kappa$ from top to bottom.
1178: }
1179: \end{figure}
1180:
1181:
1182: \subsection{General case}
1183:
1184: KWW relaxation is also achieved with the complete set of
1185: equations~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}).
1186: At time $t=0$, $\Delta(0)=0$, and no constraint is applied until time $t_w$:
1187: $\sigma([0,t_w]) = 0$.
1188: During the ramp, the deformation of the material is essentially elastic,
1189: no rearrangement occurs, hence, $\Delta(t_w^+)=0$, while the imposed
1190: strain determines the initial value the stress: $\sigma(t_w^+)=\mu\epsilon$.
1191:
1192: During the ensuing relaxation, $\dot\sigma=-\mu\dot\gamma$ and,
1193: for a small deformation, $\dot\Delta\simeq\dot\gamma/\epsilon_0$.
1194: It results that, at all times, the quantity $\sigma+\mu\,\epsilon_0\,\Delta$
1195: is constant:
1196: $$
1197: \sigma+\mu\,\epsilon_0\,\Delta = \epsilon\, U_0
1198: \quad.
1199: $$
1200: Here, $U_0=\mu$,
1201: but this expression may vary, depending on the preparation of a sample.
1202: To treat initial values of $\sigma$ and $\Delta$ in the greatest generality,
1203: the initial value for the difference $\sigma-\Delta$ is denoted $\epsilon\,S_0$
1204: (and here, $S_0=\mu$).
1205:
1206: At long times, the dynamics of $\Delta$ and $\sigma$ verify,
1207: \begin{eqnarray*}
1208: \dot\sigma-\dot\Delta &\simeq& -\left(\mu+{1\over\epsilon_0}\right)\,E_0\,
1209: \exp\left[-{1\over\chi}\right]\left(\sigma-\Delta\right)\\
1210: &\simeq& -\left(\mu+{1\over\epsilon_0}\right)
1211: \,E_0\,\left({E_1\,t\over\kappa}\right)^{-1/\kappa}\left(\sigma-\Delta\right)
1212: \end{eqnarray*}
1213: whence,
1214: $$
1215: \sigma-\Delta \simeq \epsilon\,S_0\,\exp\left[{A'\,(t_w^\beta-t^\beta)}\right]
1216: $$
1217: with,
1218: $$
1219: \beta = 1-{1\over\kappa}\quad {\rm and}\quad
1220: A'=\left(\mu+{1\over\epsilon_0}\right)\,{E_0\over\beta}\,
1221: \left({E_1\over\kappa}\right)^{-1/\kappa}
1222: \quad.
1223: $$
1224: The evolution in time of the quantity $\sigma-\Delta$
1225: is governed by the same equation as the dynamics of $\sigma$ in the isotropic limit.
1226: Therefore, all the previous discussion presented in the isotropic limit pertains
1227: to the general case, up to a constant term, which is fixed by the conservation
1228: of the quantity $\sigma+\mu\,\epsilon_0\,\Delta$.
1229: The dynamics of $\Delta$ and $\sigma$ follow.
1230:
1231: In particular, the evolution of $\Delta$ reads,
1232: $$
1233: \Delta(t) \propto \epsilon\;{U_0-S_0\exp\left[{A'\,(t_w^\beta-t^\beta)}\right]\over 1+\mu\epsilon_0}
1234: \quad,
1235: $$
1236: and the relaxation modulus verifies,
1237: $$
1238: G(t) \propto {U_0+\mu\epsilon_0\,S_0\,\exp\left[{A'\,(t_w^\beta-t^\beta)}\right]\over 1+\mu\epsilon_0}
1239: \quad.
1240: $$
1241: The isotropic expression for $G(t)$ is recovered by taking $\epsilon_0\to\infty$.
1242: Note that the relaxation of the stress could lead to a negative value if $U_0<0$,
1243: which depends on the preparation of the sample.
1244:
1245: For $\beta>0$ ({\it i.e.} $\kappa>1$), the stress undergoes
1246: a KWW relaxation towards a non-vanishing value, which
1247: depends on the properties of the material and on the preparation of the sample.
1248: For $\beta<0$ power law relaxation is observed towards a non-vanishing value
1249: of the stress.
1250: %Such relaxation towards a negative value of the stress has been observed in
1251: %recent experiments....
1252:
1253: In the current work, two origins are therefore identified for
1254: the existence of a non-vanishing stress at long times:
1255: the creation of mechanical anisotropy of the contact network (jamming),
1256: or the existence of a ergodic to non-ergodic transition
1257: for the hopping motion in phase space (freezing).
1258: Recent observations by Bonn {\it et al} indicate that for a given material,
1259: jamming could not be identified only to a loss of ergodicity,
1260: and that some amount of jamming could coexist
1261: with liquid-like behavior.~\cite{bonn02a,bonn02b}
1262: Such features are allowed in the current theory and show the importance
1263: of the STZ mechanism for soft glassy materials.
1264:
1265: The validity of the linear approximation is checked by considering the
1266: long times relaxation of the rate of plastic deformation:
1267: $$
1268: \dot\gamma \propto t^{-{1/\kappa}}\;\exp\left[{-A'\,t^\beta}\right]
1269: $$
1270: and the work of plastic deformation decays like $\dot\gamma$ since $\sigma$ goes to a constant;
1271: the non-linear term in equation~(\ref{eqn:stzdil:lin:3})
1272: is negligible at all times.
1273:
1274: % ???: \subsection{Response to a large strain}
1275: \section{Response to a constant strain rate}
1276:
1277: \subsection{\label{sec:steady} Steady plastic flow}
1278: When a constant strain rate $\dot\epsilon$ is applied,
1279: the material is driven strongly out-of-equilibrium;
1280: the response of the system involves non-linear terms.
1281: I study here the stationary states of a sheared material, in which, $\dot\gamma=\dot\epsilon$.
1282: In the first sections, I make a very primitive stability analysis of those stationary
1283: states, by assuming that $\dot\gamma=\dot\epsilon$ is fixed at all times, and that the
1284: stress is given by the constitutive relation~(\ref{eqn:stzdil:lin:1})
1285: or~(\ref{eqn:stzdil:iso:1}); this allows to decouple equation~(\ref{eqn:sigma:0}).
1286:
1287: The existence of an unstability in the complete set of equations accounts
1288: for the emergence of stick-slip motion in boundary lubrication.~\cite{lemaitre01a}
1289: This question will be briefly discussed in section~\ref{sec:hopf},
1290: but raises so many issues that it will be tackled thoroughly in another publication;
1291: for the sake of completeness,
1292: the calculation of the Hopf criterion is given in appendix~\ref{app:hopf}.
1293:
1294: \subsubsection{Isotropic limit}
1295: In the limit $\mu_0\to\infty$ (no yield stress), the relation between
1296: shear rate and shear stress reads (from~(\ref{eqn:stzdil:iso:1})):
1297: $$
1298: \sigma = {\dot\epsilon \over E_0}\, \exp\left[{1\over\chi}\right]
1299: \quad.
1300: $$
1301: Eliminating $\sigma$ between equations~(\ref{eqn:stzdil:iso:1}) and~(\ref{eqn:stzdil:iso:2}),
1302: yields:
1303: \begin{equation}
1304: \dot\chi = -E_1 \exp\left[{-{\kappa\over\chi}}\right] + \alpha {\dot\epsilon^2\over E_0}\, \exp\left[{1\over\chi}\right]
1305: \label{eqn:chidyn}
1306: \end{equation}
1307: which is positive iff,
1308: $$
1309: {\alpha\dot\epsilon^2\over E_0\,E_1} > e^{-{\kappa+1\over\chi}}
1310: \quad.
1311: $$
1312: In the stationary flow, the relation between shear strain and volume reads,
1313: $$
1314: \dot\epsilon = \sqrt{E_0\,E_1\over\alpha}
1315: \,\exp\left[{-{\kappa+1\over2\chi}}\right]
1316: \quad.
1317: $$
1318: For future use, I introduce the critical rate of deformation,
1319: $$
1320: \dot\epsilon^* = \sqrt{E_0\,E_1\over\alpha}
1321: \quad.
1322: $$
1323: \begin{figure}
1324: \narrowtext
1325: \begin{center}
1326: \unitlength=0.005\textwidth
1327: \begin{picture}(110,60)(8,-8)
1328: \put(13,40){\makebox(0,0){\large$\dot\epsilon^*$}}
1329: \put(10,48){\makebox(0,0){\large$\dot\epsilon$}}
1330: \put(85,-13){\makebox(0,0){\large$\chi$}}
1331: \put(10,-13){\resizebox{90\unitlength}{!}{\includegraphics{graph.ps}}}
1332: \end{picture}
1333: \end{center}
1334: \caption{\label{fig:vf}
1335: Bifurcation diagram for equation~(\protect\ref{eqn:chidyn}).
1336: $\dot\epsilon$ is drawn as a function of $\chi$,
1337: and the strain rate $\dot\epsilon$ is fixed.
1338: For $\dot\epsilon> \dot\epsilon^*$,
1339: no solution exists, $\dot\chi>0$, the system blows up;
1340: for $\dot\epsilon < \dot\epsilon^*$ the system admits one stable solution.}
1341: \end{figure}
1342: The bifurcation diagram for $\chi$ is presented figure~\ref{fig:vf}.
1343: If $\dot\epsilon> \dot\epsilon^*$, then $\dot\chi>0$ at all times:
1344: free-volume diverges, the system blows up.
1345: If $\dot\epsilon < \dot\epsilon^*$, the equation for $\chi$ admits one solution:
1346: $$
1347: \chi = -{\kappa+1\over 2\,\left(\log{\dot\epsilon}-\log{\dot\epsilon^*}\right)}
1348: \quad.
1349: $$
1350: The existence of a stationary value for the free-volume indicates
1351: that for any finite rate of deformation, aging stops.
1352: %This has already been observed in experiments.
1353: In fact, this is not completely exact if a vanishingly small shear rate
1354: was applied, such that the associated steady free-volume is so small,
1355: that it is not accessible at reasonable timescales.
1356: As a consequence, for very small $\dot\gamma$,
1357: the apparent viscosity saturates to a constant value. This is a common feature
1358: of complex liquids (see {\it e.g.}~\cite{larson99}).
1359:
1360: In the steady state the value of the shear stress is,
1361: $$
1362: \sigma= {\dot\epsilon^{n}\over E_0}\,\left({\alpha\over E_0 E_1}\right)^{{n-1\over2}}
1363: \quad,
1364: $$
1365: with,
1366: $$
1367: n={\kappa-1\over\kappa+1}
1368: \quad.
1369: $$
1370: The system behaves like a power law fluid;
1371: the exponent $n$ of the power law viscosity is
1372: related to the exponent $\beta$ of KWW relaxation by,
1373: $$
1374: n = {\beta\over 2-\beta}
1375: \quad;
1376: $$
1377: those two exponents are directly related to the slope $\kappa$ of $1/\chi$
1378: as a function of $\ln t$.
1379:
1380: The assumptions underlying the current approach
1381: do not permit to evaluate the activation volumes associated with density relaxation
1382: and shear motion. However, the single parameter $\kappa$ determines two
1383: exponents associated with the linear and non-linear response of a strained material.
1384:
1385: The existence of such a relation constitutes a important result of the current theory,
1386: although the form given to this relation depends on non-generic features:
1387: how non-linear dilatancy is introduced in the equation of motion for free-volume;
1388: how the update frequencies (here taken constant) are written.
1389: In a hard-sphere material, for example, collision frequencies
1390: depend on free-volume as $1/\chi$, and the relation between $\sigma$ and
1391: $\dot\epsilon$ is expected to be modified by logarithmic corrections.
1392:
1393: Molecular dynamics simulations of sheared fluid are consistent with power law rheologies
1394: for $n$ equal to 0.1~\cite{berthier00,berthier01} or 0.2~\cite{yamamoto97,yamamoto98}
1395: Recent measurement of power law rheology have also been
1396: obtained on a colloidal glass,
1397: leading to values of $n$ ranging from 0.1 to 0.35.~\cite{bonn02a,bonn02b}
1398: A value of $\beta$ of order $0.4$ corresponds to $n=0.25$, which seems
1399: in reasonable agreement with the currently available data.
1400:
1401: The theory of structural rearrangement described here,
1402: is expected to present some degree of universality,
1403: and hold, in a given form, for wide classes of materials.
1404: What this model shows is that there is an intimate relation between
1405: time-logarithmic density relaxation, KWW relaxation,
1406: and power law rheology of a dense material, and that this relation may depend
1407: only on generic features for a given class of materials.
1408: Therefore the existence of a relation between $\beta$ and $n$
1409: could be tested by considering families of materials.
1410: The experimental or numerical evaluation of such a relation
1411: appears as an important mean of characterization of the coupling between
1412: two major modes of deformation.
1413:
1414: \subsubsection{General case}
1415:
1416: The constitutive equations are now~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}).
1417: In the steady state,
1418: $$
1419: \Delta = {1\over\mu_0\sigma}
1420: \quad;
1421: $$
1422: plugging expression~(\ref{eqn:stzdil:lin:1}) for the plastic deformation
1423: in equation~(\ref{eqn:stzdil:lin:3}) leads to the following relation between $\chi$
1424: and $\sigma$ in the steady state:
1425: $$
1426: \exp\left[{1-\kappa\over\chi}\right]={\alpha\,E_0\over E_1}\,\left(\sigma^2-{1\over\mu_0}\right)
1427: %\sigma = \sqrt{{1\over\mu_0}+{E_1\over\alpha E_0}\,\exp\left[{1-\kappa\over\chi}\right]}}
1428: \quad.
1429: $$
1430: Using~(\ref{eqn:stzdil:lin:3}) again,
1431: $\dot\epsilon$ can then be written as a function of $\sigma$:
1432: $$
1433: \dot\epsilon =
1434: {E_1\over\alpha}\;
1435: \left(\alpha\,E_0\over E_1\right)^{{\kappa\over\kappa-1}}
1436: \;{1\over\sigma}\;\left(\sigma^2-{1\over\mu_0}\right)^{{\kappa\over\kappa-1}}
1437: \quad.
1438: $$
1439: For large $\sigma$ or $\dot\epsilon$, the system behaves as a power law fluid,
1440: with the same exponent $n$ as in the isotropic limit.
1441: However, $\dot\epsilon$ vanishes for a non-zero value of $\sigma$,
1442: which determines the yield stress:
1443: $$
1444: \sigma_y = {1\over\sqrt{\mu_0}}
1445: \quad.
1446: $$
1447:
1448: The following expression can also be obtained for $\dot\epsilon$ as a function of $\chi$:
1449: $$
1450: \dot\epsilon = {
1451: \sqrt{E_0\,\mu_0}\,E_1\;\exp\left[-{\kappa/ 2\chi}\right]
1452: \over
1453: \sqrt{\alpha}\;
1454: \sqrt{E_0\alpha\,\exp\left[{\kappa/\chi}\right]
1455: + E_1\mu_0\,\exp\left[{1/\chi}\right]}
1456: }
1457: \quad,
1458: $$
1459: which is an increasing function of $\chi$, and which has a maximum for $\chi\to\infty$:
1460: $$
1461: \dot\epsilon^*=
1462: {\sqrt{E_0\,\mu_0}\,E_1
1463: \over
1464: \sqrt{\alpha}\;
1465: \sqrt{E_0\alpha+ E_1\mu_0}
1466: }
1467: $$
1468: For $\dot\epsilon>\dot\epsilon^*$, like in the isotropic case,
1469: the shear deformation leads to a constant increase of free-volume,
1470: and the system blows up.
1471:
1472: \subsection{Transients}
1473:
1474: When a shear rate is suddenly imposed on a material,
1475: the transient dynamics often lead to a first rise of the stress before
1476: a dynamical yield stress is reached at which plastic deformations begin;
1477: once plastic deformation occur, the stress relaxes towards its strain-rate
1478: dependent asymptotic value.~\cite{larson99}
1479: These features are observed in sheared bulk solids~\cite{hassan95}
1480: but also in sheared thin films,~\cite{drummond00,drummond01,persson00}
1481: granular materials~\cite{losert00}
1482: or in solid friction.~\cite{bureau02}
1483: Another generic property of transient regimes is that
1484: the dynamical yield stress depends on the age of the material:
1485: during aging, a material strengthens.
1486:
1487: These phenomena are captured by the proposed equations.
1488: %The interplay of the dynamics of free-volume and of arrangements during
1489: %transient regimes leads to a rather diverse phenomenology.
1490: Strain softening and strengthening result essentially from the equation of motion
1491: for free-volume and are captured by the isotropic equations.
1492:
1493: Figure~\ref{fig:bump} present some examples of transient dynamics
1494: obtained by numerical integration of
1495: equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1}-\ref{eqn:stzdil:iso:2}).
1496: Parameters are set to $E_0=E_1=\mu=\alpha=1$.
1497: The initial value for the free-volume (at time $t=0$) is set to $\chi(0)=10$,
1498: and the system relaxes without forcing during a time $t_w$.
1499: At time $t_w$, a constant strain rate is applied.
1500: To prevent the blow-up of the system,
1501: $\dot\epsilon$ must be smaller than $\dot\epsilon^*$,
1502: which equals $1$ with the chosen parameters;
1503: for this reason, the strain rate is set to $\dot\epsilon=0.9$.
1504: Two different values of the parameter $\kappa$ have been used to explore
1505: the two important cases when $\kappa$ is larger or smaller than 1.
1506: The stress is displayed as a function of the total time $t$
1507: since preparation of the sample.
1508: \begin{figure}
1509: \narrowtext
1510: \begin{center}
1511: \unitlength = 0.005\textwidth
1512: \begin{picture}(100,110)(5,0)
1513: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{bumps.eps}}}
1514: %\put(0,0){\line(1,0){100}}
1515: %\put(0,0){\line(0,1){100}}
1516: %\put(100,0){\line(1,0){100}}
1517: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1518: \put(5,50){\makebox(0,0){\large $\sigma$}}
1519: \put(90,2){\makebox(0,0){\large $t$}}
1520: \end{picture}
1521: \end{center}
1522: \caption{\label{fig:bump}
1523: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1524: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain rate $\dot\epsilon=0.9$.
1525: Parameters are $E_0=E_1=\alpha=\mu=1$, and values of $\kappa$:
1526: $\kappa=0.8$ (solid lines) and $\kappa=1.2$ (dashed lines).
1527: No stress is applied during the initial density relaxation of the material
1528: from, $\chi(0)=10$.
1529: At various times, $t_w$, the strain rate is suddenly applied and the ensuing
1530: dynamics of $1/\chi$ (top) and $\sigma$ (bottom) are displayed.
1531: The strong glass $\kappa<1$ presents larger of the dynamical yield,
1532: which result from the smaller values of $\chi$ reached.
1533: }
1534: \end{figure}
1535:
1536: A dynamical yield stress clearly emerges for a sufficiently aged material,
1537: and increases with time. The yielding of the material is accompanied by a sudden
1538: rise of the free-volume allowing transformation rates to display non-vanishing values.
1539:
1540: %When a constant strain rate is imposed on the material,
1541: %the stress starts to increase linearly with time,
1542: %so long as the plastic deformation rate is negligible
1543: %compared to the elastic deformation, or $\dot\epsilon>>\dot\gamma$.
1544: %This vanishing value of $\dot\gamma$ is due to the very small
1545: %free-volume, hence the small prefactor $\exp[-1/\chi]$
1546: %in the constitutive relation. During the ramp, free-volume continues
1547: %to decay, so long as the relaxation is not balance by dilatancy.
1548: %Free-volume starts to increase when,
1549: %$$
1550: %\sigma^2(t)={E_1\over \alpha\,E_0}\exp\left[{1-\kappa\over\chi}\right]
1551: %\quad.
1552: %$$
1553: %Since, $\sigma(t) = \mu\,(\epsilon-\gamma)\simeq \mu\,\dot\epsilon\,(t-t_w)$,
1554: %this occurs at a time $t_y$ such that
1555: %$$
1556: %t_y - t_w \simeq \sqrt{E_1^{1/\kappa}\over \alpha\,E_0} \, t^{-\beta/2}
1557: %\simeq \sqrt{E_1^{1/\kappa}\over \alpha\,E_0} \, t_w^{-\beta/2}
1558: %\quad.
1559: %$$
1560: %The increase of free-volume triggers the plastic flow, further increase of
1561: %free-volume,\dots, and the process of acceleration occurs of a very small
1562: %time interval. Therefore, the value of the stress that inverts free-volume
1563: %dynamics is almost equal to the yield stress itself:
1564: %$$
1565: %\sigma_y^d\simeq \sqrt{E_1\over \alpha\,E_0} \, t^{-\beta/2}
1566: %\quad.
1567: %$$
1568:
1569: These transients result from a rather complex interplay between the dynamics
1570: of free-volume and shear transformation. The complete study of this process
1571: may lead to many developments that would divert this work from the main questions
1572: that I would like to address, and I prefer to leave the discussion here.
1573:
1574:
1575: \subsection{\label{sec:hopf} Stick-slip instability}
1576:
1577: The stability analysis of equations~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3})
1578: and~(\ref{eqn:sigma:0}) shows the existence of a Hopf bifurcation which is responsible
1579: for oscillatory response to a given strain rate.
1580: This leads to some understanding of stick-slip motion in sheared
1581: lubricants~\cite{thompson92,yoshizawa93,demirel96a,demirel96b,drummond00,drummond01}
1582: or in granular materials.~\cite{thompson91,nasuno97,nasuno98}
1583: In this work, I mention this instability for the sake of completeness,
1584: but it raises many questions, and will require a more thorough consideration.
1585: %Since this question raises many issues, I prefer to stop the discussion here,
1586: %and discuss stick-slip motion in another publication.%~\cite{lemaitre02b}
1587: The calculation of the Hopf criterion of bifurcation is given
1588: in appendix~\ref{app:hopf}.
1589:
1590: In the isotropic limit,
1591: the steady plastic deformation is stable iff $\mu>\mu_{\rm hopf}$, with:
1592: $$
1593: \mu_{\rm hopf} = {E_1\over E_0}\,{1-\kappa\over(\kappa+1)^2}\;
1594: \left({\alpha\,\dot\epsilon^2\over E_0\,E_1}\right)^{{\kappa-1\over\kappa+1}}\;
1595: \ln\left[{\alpha\,\dot\epsilon^2\over E_0\,E_1}\right]^2
1596: \quad.
1597: $$
1598: For $\kappa>1$, $\mu_{\rm hopf}<0$: there is no Hopf bifurcation.
1599: For $\kappa<1$, $\mu_{\rm hopf}>0$ on the interval $[0,\dot\epsilon^*]$,
1600: and the steady state becomes unstable iff $\mu<\mu_{\rm hopf}$.
1601: This criterion is represented figure~\ref{fig:hopf:iso}.
1602: For driving strain rates above $\dot\epsilon^*$, the system looses stability,
1603: because there is no stationary solution for $\chi$.
1604: Below $\dot\epsilon^*$, the Hopf criterion separates a domain of stability,
1605: above $\mu_{\rm hopf}$ where the system reaches steady plastic deformation,
1606: and and domain below this curve where stick-slip motion is achieved.
1607: \begin{figure}
1608: \narrowtext
1609: \begin{center}
1610: \unitlength=0.005\textwidth
1611: \begin{picture}(110,60)(8,-8)
1612: \put(6,40){\makebox(0,0){\large$\mu_{\rm hopf}$}}
1613: \put(68,-13){\makebox(0,0){\large$\dot\epsilon^*$}}
1614: \put(85,-13){\makebox(0,0){\large$\dot\epsilon$}}
1615: \put(10,-13){\resizebox{90\unitlength}{!}{\includegraphics{hopfiso.ps}}}
1616: \end{picture}
1617: \end{center}
1618: \caption{\label{fig:hopf:iso}
1619: Phase diagram for equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1620: and~(\ref{eqn:stzdil:iso:2}), and $\kappa<1$.
1621: The solid line denotes the curve $\mu_{\rm hopf}$ below
1622: which stick-slip motion occurs. $\mu_{\rm hopf}$ vanishes for
1623: $\dot\epsilon=\dot\epsilon^*$ (dashed line)
1624: where steady plastic regime disappears, leading
1625: the system to break-up.
1626: }
1627: \end{figure}
1628:
1629: The introduction of the variable $\Delta$ in the set of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3})
1630: slightly modifies this picture (see appendix~\ref{app:hopf}).
1631: Two limit curves $\mu_{\rm hopf}^\pm$ emerge, which delimit the unstable
1632: domain in the plane $(\dot\epsilon,\mu)$.
1633: These curves are displayed figure~\ref{fig:hopf:lin} along with the previous curves $\mu_{\rm hopf}$ of the isotropic limit.
1634: The unstable domain is a subspace of what it is in the isotropic limit,
1635: and the system present a creeping zone, close to $\dot\epsilon=0$,
1636: when steady sliding is stable for very small values of the strain rate.
1637: \begin{figure}
1638: \narrowtext
1639: \begin{center}
1640: \unitlength = 0.005\textwidth
1641: \begin{picture}(90,90)(0,0)
1642: \put(10,5){\resizebox{70\unitlength}{!}{\includegraphics{hopf.eps}}}
1643: \put(5,75){\makebox(0,0){\large $\mu$}}
1644: \put(70,2){\makebox(0,0){\large $\dot\epsilon$}}
1645: \end{picture}
1646: \end{center}
1647: \caption{\label{fig:hopf:lin}
1648: Phase diagram for equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}) and $\kappa<1$.
1649: Parameters used are, $E_0=E_1=\alpha=\epsilon_0=\mu_0=1$, and $\kappa=0.8$.
1650: The solid line denotes the Hopf criterion $\mu_{\rm hopf}$ in the isotropic
1651: limit. The dashed lines indicates the Hopf criterion $\mu_{\rm hopf}^\pm$
1652: with the complete set of equations.
1653: }
1654: \end{figure}
1655:
1656: Some examples of stick-slip motion are displayed figure~\ref{fig:stickslip},
1657: with the set of parameters used in the phase diagram of figure~\ref{fig:hopf:lin}:
1658: $E_0=E_1=\alpha=1$, and $\kappa=0.8$ in the isotropic limit (solid lines)
1659: and $\epsilon_0=\mu_0=1$ for the complete set of equations (dashed lines).
1660: The fast bursts of plastic deformation are accompanied by sudden
1661: jumps in the free-volume.
1662: $\Delta$ does not play a major role for the definition of the dynamical
1663: yield stress, as seen from the almost identical values of the maximum stress;
1664: $\Delta$, however, strongly influences smaller values of the stress as it shifts
1665: upwards the unstable fixed point;
1666: this also modifies the periodicity of stick-slip motion.
1667: \begin{figure}
1668: \narrowtext
1669: \begin{center}
1670: \unitlength = 0.005\textwidth
1671: \begin{picture}(100,110)(5,0)
1672: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{stickslip.eps}}}
1673: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1674: \put(5,50){\makebox(0,0){\large $\sigma$}}
1675: \put(90,2){\makebox(0,0){\large $t$}}
1676: \end{picture}
1677: \end{center}
1678: \caption{ \label{fig:stickslip}
1679: Solid lines:
1680: Numerical integration of equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1})
1681: and~(\ref{eqn:stzdil:iso:2}) for a fixed strain rate $\dot\epsilon=0.1$.
1682: Parameters are $E_0=E_1=\alpha=\mu=1$, and $\kappa=0.8$.
1683: Initial value of the free-volume is $\chi=10$.
1684: The regime of steady plastic deformation is unstable and leads to stick-slip motion.
1685: Fast relaxations of the stress result from sudden dilatancy of the material.
1686: Dashed lines: Same for equations~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}) with the same parameters and, $\epsilon_0=\mu_0=1$.
1687: }
1688: \end{figure}
1689:
1690: % ???: bifurcation diagram
1691: %\subsection{Stick-slip motion}
1692: %\subsubsection{Over-damped stick-slip}
1693: %\subsubsection{Hopf analysis}
1694: %The introduction of the equation of motion for $\Delta$ slightly changes this picture.
1695:
1696: \section{Response to a constant stress}
1697: In a creep test, a material is tested by forcing a constant stress.
1698: Since $\sigma$ is constant, the elastic deformation is $\epsilon=\sigma/\mu$
1699: at all times, and the plastic deformation is determined by the constitutive equations.
1700:
1701: %The preparation of the sample is the same as previously:
1702: %until time $t_w$, no constraint is applied to the material
1703: %which undergoes density relaxation.
1704: %At time $t_w$ a constant stress is suddenly applied, and the elastic deformation
1705: %adapts in a very small time interval while no plastic deformation has yet occurred.
1706:
1707: \subsection{Linear response}
1708:
1709: \subsubsection{\label{sec:lin:iso} Isotropic limit}
1710: A small constant stress $\sigma$ is applied at time $t=t_w$.
1711: The plastic deformation of the material is determined
1712: by equations~(\ref{eqn:stzdil:iso:1}-\ref{eqn:stzdil:iso:2}),
1713: for a fixed $\sigma$.
1714: At large times, the rate of plastic deformation reads
1715: $$
1716: \dot\epsilon=\dot\gamma \simeq E_0\, \left({E_1\,t\over \kappa}\right)^{-1/\kappa}\sigma
1717: \quad.
1718: $$
1719: This lead to a power law variation of the compliance, $J=\gamma/\sigma$:
1720: $$
1721: J(t) \propto A\;\left(t^\beta-t_w^\beta\right)
1722: \quad.
1723: $$
1724: If $\kappa>1$, or $\beta>0$, the compliance diverges at long times;
1725: if $\kappa<1$, or $\beta<0$, the compliance saturates at long times:
1726: $$
1727: J(\infty) = |A|\,t_w^{-|\beta|}
1728: \quad,
1729: $$
1730: the system jams. This jamming is purely entropic, and does not result
1731: from a structural anisotropy of the force network.
1732: Note that the asymptotic value of the compliance
1733: decays with the increasing age of the sample.
1734:
1735: The validity of the linear approximation must be checked
1736: by considering the evolution of the work of plastic deformation:
1737: $$
1738: \sigma\,\dot\gamma\propto t^{-{1/\kappa}}
1739: \quad;
1740: $$
1741: the linear approximation is justified so long as this term is negligible,
1742: when compared to the first term in free-volume relaxation~(\ref{eqn:stzdil:iso:2}),
1743: which decays as $\kappa/t$.
1744: When $\kappa<1$, the work $\sigma\,\dot\gamma$ decays faster than $\kappa/t$,
1745: and therefore, the linear approximation is valid at all times;
1746: when $\kappa>1$, however, this approximation fails when:
1747: $$
1748: \alpha\,E_0\,\left({E_1\,t\over\kappa}\right)^{-1/\kappa}\sigma^2\sim \kappa/t
1749: \quad,
1750: $$
1751: or,
1752: $$
1753: t \sim t^{\rm n.l.} = \kappa\,\left({E_1^{1/\kappa}\over \alpha\,E_0\,\sigma^2}\right)^{\kappa\over\kappa-1}
1754: \quad.
1755: $$
1756: The time $t^{\rm n.l.}$ is a typical time after which the linear
1757: approximation fails; this time diverges when $\sigma\to0$.
1758: It will appear clearer in the following that, when $\kappa>1$,
1759: free-volume presents a non-vanishing stationary value for any small applied stress.
1760: This asymptotic value is determined by the balance between density relaxation
1761: and shear induced dilatancy. The failure of the linear approximation, for $\kappa>1$,
1762: is due to the
1763: existence of a non-zero stationary value of $\chi$ for any applied stress.
1764:
1765: \subsubsection{\label{sec:lin:lin} General case}
1766:
1767:
1768: In the general case, the constitutive equations
1769: are~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}).
1770: The constant applied stress deforms the material elastically,
1771: and at time $t_w^+$, no plastic rearrangement has yet occurred.
1772: Note that the initial value of the state variable $\Delta$
1773: depends on the preparation of the sample.
1774: In the experiments by Cloitre {\it et al},
1775: for example, the sample is prepared by applying a strong shear stress,
1776: above the yield stress during a rather long time.~\cite{cloitre00}
1777: In this case, the initial value of $\Delta$ is expected
1778: to be non-vanishing, and this leads to their observation of strain recovery.
1779: For the sake of generality, I will consider that the initial value of $\Delta$
1780: is non-vanishing: $\Delta(t_w^+)=\Delta_0$.
1781: In this section, I will assume that this value of $\Delta$ is small enough
1782: so that the linear approximation is still valid.
1783: When plastic rearrangements occur in the material,
1784: they induce a structuration of the material, and
1785: the variable $\Delta$ starts to evolve as $\dot\Delta\simeq\dot\gamma/\epsilon_0$.
1786:
1787: The plastic rate of deformation evolves as,
1788: $$
1789: \dot\gamma \simeq E_0\, \left({E_1\,t\over\kappa}\right)^{-1/\kappa}(\sigma-\Delta)
1790: \quad,
1791: $$
1792: and $\Delta$ as,
1793: $$
1794: \dot\Delta \simeq {E_0\over\epsilon_0}\,\left({E_1\,t\over\kappa}\right)^{-1/\kappa}
1795: \,(\sigma-\Delta)
1796: \quad.
1797: $$
1798:
1799: From this equation, the relaxation of the variable $\sigma-\Delta$ is easily
1800: obtained:
1801: $$
1802: \sigma-\Delta(t) \simeq (\sigma-\Delta_0)\,\exp\left[{A''\,(t_w^\beta-t^\beta)}\right]
1803: \quad,
1804: $$
1805: with
1806: $$
1807: A'' = {E_0\over\beta \epsilon_0}{\left({E_1\over\kappa}\right)^{-1/\kappa}}
1808: \quad.
1809: $$
1810: When $\kappa>1$, $\sigma-\Delta$ undergoes a KWW relaxation towards 0:
1811: the material relaxes towards a jammed state in which the stress is supported
1812: by structural anisotropy.
1813: When $\kappa<1$, $\sigma-\Delta$ saturates to a non-vanishing value:
1814: at long times, a fraction of the stress is supported by
1815: the bias between arrangements, measured by $\Delta$, but another fraction
1816: is supported by the loss of ergodicity resulting from the
1817: freezing of elementary shear processes.
1818:
1819: From equation $\dot\Delta\simeq\dot\gamma/\epsilon_0$, the compliance reads,
1820: $$
1821: J(t) \simeq \epsilon_0\,\left(1-{\Delta_0\over\sigma}\right)\,
1822: (1 - \exp\left[{A''\,(t_w^\beta-t^\beta)}\right])
1823: \quad.
1824: $$
1825: The deformation relaxes to a constant value, which depends on the preparation of
1826: the sample; the relaxation is either a KWW or a power law
1827: relaxation depending on $\kappa$.
1828: Note that the power law variation of the compliance that was found
1829: in the isotropic limit is recovered when $A''\to0$;
1830: for finite $A''$, this power law behavior shows up at short times,
1831: when $\Delta$ is still far from saturation.
1832: %The crossover between those regimes occurs when $t^\beta\simeq t_w^\beta-1/A''$.
1833:
1834: \subsection{Jamming}
1835:
1836: In this section I start by considering the complete set of equations~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3}),
1837: and in particular, I recall how the dynamics of $\Delta$ lead to jamming.
1838: Then I show how this leads to an effective equation for free-volume which
1839: is valid in the isotropic limit or in the general case.
1840:
1841: A constant stress $\sigma$ is applied.
1842: The dynamics of $\Delta$ is governed by the following equation:
1843: \begin{equation}
1844: \dot\Delta = {E_0\over\epsilon_0}\,\exp\left[-{1\over\chi}\right]\,
1845: \left(\sigma-\Delta\right)\;\left(1-\mu_0\,\sigma\,\Delta\right)
1846: \quad.
1847: \label{eqn:deltadyn}
1848: \end{equation}
1849: Since the normalized free-volume $\chi$ enters this equation only through
1850: a positive factor, the time-dependent value of $\chi$ does not change
1851: the stability analysis performed at any fixed free-volume.
1852: The bifurcation diagram for $\Delta$ is drawn figure~\ref{fig:delta}.
1853: \begin{figure}
1854: \narrowtext
1855: \begin{center}
1856: \unitlength=0.005\textwidth
1857: \begin{picture}(110,60)(8,-8)
1858: \put(12,40){\makebox(0,0){\large$\Delta$}}
1859: \put(85,-13){\makebox(0,0){\large$\sigma$}}
1860: \put(10,-13){\resizebox{90\unitlength}{!}{\includegraphics{delta.ps}}}
1861: \end{picture}
1862: \end{center}
1863: \caption{\label{fig:delta}
1864: Bifurcation for the dynamics of the variable $\Delta$ as defined by
1865: equation~(\protect\ref{eqn:deltadyn}). The intersection between the two branches
1866: of solutions define the yield stress $\sigma_y$.
1867: For $\sigma<\sigma_y$, the stable solution is the jammed state, while it is the
1868: plastic flow for $\sigma>\sigma_y$.~\protect\cite{falk98}
1869: }
1870: \end{figure}
1871:
1872: The dynamics of the variable $\Delta$ leads to jamming for any stress smaller
1873: than the yield stress~\cite{falk98}
1874: $$
1875: \sigma_y = {1\over\sqrt{\mu_0}}
1876: \quad,
1877: $$
1878: in which case, $\Delta = \sigma$ and $\dot\epsilon=0$.
1879: Above the yield stress, the steady shear flow becomes a stable solution,
1880: in which case,
1881: $$
1882: \Delta = {1\over\mu_0\,\sigma}
1883: \quad,
1884: $$
1885: and
1886: \begin{equation}
1887: \dot\epsilon =
1888: {E_0}\,\exp\left[-{1\over\chi}\right]\,\left(\sigma-{1\over\mu_0\,\sigma}\right)
1889: \quad.
1890: \label{eqn:eps:creep}
1891: \end{equation}
1892:
1893: If $\sigma<\sigma_y$, the systems jams,
1894: and the second term in the equation of motion for the free-volume vanishes.
1895: Free-volume continues to relax logarithmically in time while the system does not deform.
1896:
1897: When $\sigma>\sigma_y$, the system presents a non-zero plastic flow,
1898: and the evolution of free-volume is governed by the following asymptotic equation:
1899: \begin{equation}
1900: \dot\chi = -E_1 \exp\left[{-{\kappa\over\chi}}\right]
1901: +\alpha {E_0}\,\exp\left[-{1\over\chi}\right]\,\left(\sigma^2-{1\over\mu_0}\right)
1902: \quad.
1903: \label{eqn:chi:creep}
1904: \end{equation}
1905: At the level of this equation, the jamming resulting from STZ's
1906: shows up only through the term $1/\mu_0$ in equation~(\ref{eqn:chi:creep}),
1907: which shifts the square of the applied stress.
1908: The stability analysis pertains essentially to the isotropic limit.
1909: From this point on, I will consider the response to a constant stress
1910: as governed by equations~(\ref{eqn:eps:creep}) and (\ref{eqn:chi:creep})
1911: (which amounts to neglect the influence of $\Delta$ during transients).
1912:
1913: \subsection{Brittleness and ductility}
1914: I now study the free-volume dynamics which result from equation~(\ref{eqn:chi:creep}).
1915: The material dilates iff $\sigma>\sigma_y^*$, with:
1916: $$
1917: \sigma_y^* = \sqrt{{1\over\mu_0} +
1918: {E_1\over\alpha \,E_0} \exp\left[{{1-\kappa\over\chi}}\right]}
1919: $$
1920: and contracts otherwise. The resulting bifurcation diagram for $\chi$
1921: is drawn figure~\ref{fig:chi:1} and~\ref{fig:chi:2} for $\kappa>1$ and $\kappa<1$
1922: respectively.
1923:
1924: \subsubsection{Soft glasses}
1925:
1926: For $\kappa>1$, free-volume dynamics admit a stationary solution for
1927: any $\sigma\in[\sigma_y,\sigma^*]$, with,
1928: $$
1929: \sigma^* = \sqrt{{1\over\mu_0} + {E_1\over\alpha \,E_0}}
1930: \quad.
1931: $$
1932: For values of the applied stress in the interval $[\sigma_y,\sigma^*]$,
1933: the material undergoes steady plastic deformation: it is ductile.
1934: The shear deformation reads,
1935: $$
1936: \dot\epsilon =
1937: {E_1\over\alpha}\;
1938: \left(\alpha\,E_0\over E_1\right)^{{\kappa\over\kappa-1}}
1939: \;{1\over\sigma}\;\left(\sigma^2-{1\over\mu_0}\right)^{{\kappa\over\kappa-1}}
1940: \quad,
1941: $$
1942: and the strain rate behaves as a power of the stress for large stresses.
1943: When the applied stress varies in the interval, $\sigma\in[\sigma_y,\sigma^*]$,
1944: the strain rate can display a whole interval of values
1945: $$
1946: \dot\epsilon\in\left[0,E_0\,\left(\sigma^*-{1\over\mu_0\,\sigma^*}\right)\right]
1947: \quad.
1948: $$
1949: For larger values of $\sigma$, above $\sigma^*$, free-volume diverges, leading
1950: to the break-up of the material.
1951:
1952: \begin{figure}
1953: \narrowtext
1954: \begin{center}
1955: \unitlength=0.005\textwidth
1956: \begin{picture}(110,60)(8,-8)
1957: \put(10,48){\makebox(0,0){\large$\sigma$}}
1958: \put(13,40){\makebox(0,0){\large$\sigma^*$}}
1959: \put(13,-1){\makebox(0,0){\large${1\over\mu_0}$}}
1960: \put(85,-13){\makebox(0,0){\large$\chi$}}
1961: \put(10,-13){\resizebox{90\unitlength}{!}{\includegraphics{chi1.ps}}}
1962: \end{picture}
1963: \end{center}
1964: \caption{Bifurcation diagram for the free-volume dynamics defined by
1965: equation~(\protect\ref{eqn:chi:creep}), for $\kappa>1$.
1966: For any value of the stress
1967: in the interval $[\sigma_y,\sigma^*]$, there exist a stationary value of $\chi$.
1968: The applied stress stops aging, and the material undergoes steady
1969: plastic deformation.
1970: Above $\sigma^*$, free-volume diverges leading to the break-up of the material.
1971: }
1972: \label{fig:chi:1}
1973: \end{figure}
1974:
1975: The resulting dynamics for the compliance and for the free-volume are presented figure~\ref{fig:compl:soft} for different values of the driving stress $\sigma$.
1976: The first part of the dynamics is determined by the initial value of
1977: free-volume and the corresponding timescale.
1978: As easily seen on the bifurcation diagram~(\ref{fig:chi:1}), for small stresses,
1979: an initially large free-volume decays, and slows down the plastic deformation.
1980: This results in a plateau for the dynamics of the compliance; a larger times, free-volume converges
1981: towards a constant value, and the dynamics of $J$ resumes a steady increase.
1982:
1983: These curves are to compare with measurements of particle diffusion:
1984: the particle diffusion in a dense material is expected to be dominated
1985: by collective processes; the diffusive constant is then $\propto\exp[-1/\chi]$,
1986: and the average motion of a particle in a glassy medium is essentially
1987: governed by the same equation as $J$.
1988: The features displayed of figure~\ref{fig:compl:soft}
1989: are amazingly similar to the curves found in
1990: several recent experiments.~\cite{megen98,cloitre00}
1991: \begin{figure}
1992: \narrowtext
1993: \begin{center}
1994: \unitlength = 0.005\textwidth
1995: \begin{picture}(100,110)(5,0)
1996: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{compl-soft.eps}}}
1997: \put(5,95){\makebox(0,0){\large $1/\chi$}}
1998: \put(5,50){\makebox(0,0){\large $\log_{10}(J)$}}
1999: \put(90,2){\makebox(0,0){\large $\log_{10}(t)$}}
2000: \end{picture}
2001: \end{center}
2002: \caption{
2003: Numerical integration of equations~(\ref{eqn:stzdil:iso:1})
2004: and~(\ref{eqn:stzdil:iso:2}) for constant stresses $\sigma$, and for $\kappa>1$.
2005: Parameters are $E_0=E_1=\alpha=\mu=1$, and $\kappa=2$;
2006: for these parameter, $\sigma^*=1$.
2007: Initial value of the free-volume is $\chi=100$ and $t_w=0$.
2008: Top: $1/\chi$ as a function of time for
2009: $\sigma = 0.2, 0.4, 0.6, 0.8, 1$ from top to bottom; for $\sigma=1$, and for
2010: all values above, free-volume diverges.
2011: Bottom: $\log_{10}(J)$ as a function of time for $\sigma$
2012: increasing from bottom to top.
2013: }
2014: \label{fig:compl:soft}
2015: \end{figure}
2016:
2017: If free-volume diverges, just before break-up,
2018: the current equations indicate
2019: that the strain rate should display a behavior of the form,
2020: $$
2021: \dot\epsilon = E_0\,\left(\sigma-{1\over\mu_0\,\sigma}\right)
2022: \quad.
2023: $$
2024: However, this expression must be taken with caution,
2025: since the strongly out-of-equilibrium behavior of a material during break-up,
2026: when $\chi$ is large, is beyond the scope of the current theory.
2027: In particular, assumption has been made that the distribution
2028: of voids is close to an equilibrium state.
2029: During break-up, the distribution of free-volume is expected to display strong
2030: heterogeneities leading to the nucleation of fractures.
2031:
2032: \subsubsection{Hard glasses}
2033:
2034: For $\kappa<1$, and for any $\sigma<\sigma^*$, $\dot\chi<0$:
2035: free-volume relaxes to 0. In this case, the applied stress does not stop aging,
2036: shear deformation is vanishingly small: the material creeps.
2037: At long time, the linear term dominates free-volume relaxation: the calculation
2038: of linear response that has been presented in section~\ref{sec:lin:iso} is valid,
2039: and indicates that compliance saturates: the system jams.
2040: In general, part of this jamming is supported by the structuration of the material,
2041: which lowers the value of the stress that enters equation~(\ref{eqn:chi:creep});
2042: the remaining stress is supported by entropic freezing of the rearrangement dynamics.
2043:
2044: The sign of free-volume relaxation can be inverted only for larger stresses,
2045: above $\sigma_y^*(\chi)$; it this case, the response depends sensitively on
2046: the value of $\sigma$ and on the state of the material at the time $t_w$
2047: when stress is applied:
2048: a larger stress is required to break an older material.
2049: The stationary state is unstable: when $\sigma$ is large enough
2050: to invert the sign of free-volume dynamics, the material breaks up.
2051: In this case, there are two types of response: creep for low stress, and break-up
2052: for large values of the stress. The material is brittle.
2053: \begin{figure}
2054: \narrowtext
2055: \begin{center}
2056: \unitlength=0.005\textwidth
2057: \begin{picture}(110,60)(8,-8)
2058: \put(10,40){\makebox(0,0){\large$\sigma$}}
2059: \put(13,20){\makebox(0,0){\large$\sigma^*$}}
2060: \put(85,-13){\makebox(0,0){\large$\chi$}}
2061: \put(10,-13){\resizebox{90\unitlength}{!}{\includegraphics{chi2.ps}}}
2062: \end{picture}
2063: \end{center}
2064: \caption{Bifurcation diagram for the free-volume dynamics defined by
2065: equation~(\protect\ref{eqn:chi:creep}), for $\kappa<1$.
2066: For all $\sigma<\sigma^*$, free-volume continues relaxation towards 0;
2067: shear deformation occurs at a vanishingly small rate; the material creeps.
2068: The sign of free-volume relaxation can be changed only by applying a stress at least
2069: larger than $\sigma^*$; the required stress depends on the age of the material.
2070: The steady solution is always unstable: large stresses lead to break up.}
2071: \label{fig:chi:2}
2072: \end{figure}
2073:
2074: The resulting dynamics for free-volume and for the compliance are
2075: displayed figure~\ref{fig:compl:hard}. The saturation of the compliance
2076: is clearly seen for small stresses while the divergence occurs at
2077: a value at $\sigma=\sigma_y^*(\chi)$.
2078: The shape of these curves changes drastically in a very small interval of $\sigma$
2079: around $\sigma_y^*(\chi)$ because of long transient to escape from
2080: the unstable fixed point. These features are commonly observed in creep tests.~\cite{hassan95,larson99}
2081: \begin{figure}
2082: \narrowtext
2083: \begin{center}
2084: \unitlength = 0.005\textwidth
2085: \begin{picture}(100,110)(5,0)
2086: \put(10,5){\resizebox{90\unitlength}{!}{\includegraphics{compl-hard.eps}}}
2087: \put(5,95){\makebox(0,0){\large $1/\chi$}}
2088: \put(5,50){\makebox(0,0){\large $J$}}
2089: \put(90,2){\makebox(0,0){\large $t$}}
2090: \end{picture}
2091: \end{center}
2092: \caption{
2093: Numerical integration of equations~(\ref{eqn:stzdil:iso:1})
2094: and~(\ref{eqn:stzdil:iso:2}) for constant stresses $\sigma$.
2095: Parameters are $E_0=E_1=\alpha=\mu=1$, and $\kappa=0.8$.
2096: Initial value of the free-volume is $\chi=1$ and $t_w=0$.
2097: For these values of the parameters, and of initial conditions,
2098: the critical value of $\sigma$ where the free-volume diverges is,
2099: $\sigma_y^*\simeq 1.105$.
2100: Top: $1/\chi$ as a function of time for
2101: $\sigma = 0.5, 1, 1.05, 1.075, 1.1, 1.15, 1.2$ from top to bottom.
2102: Bottom: $J$ as a function of time for $\sigma$ increasing from bottom to top.
2103: }
2104: \label{fig:compl:hard}
2105: \end{figure}
2106:
2107:
2108: \subsubsection{Jamming and freezing}
2109:
2110: Materials with $\kappa>1$ are liable to display ductile behavior,
2111: but can break; those with $\kappa<1$ either creep or break-up,
2112: and could be classified as unconditionally brittle.
2113: The study of STZ equations one a one-dimensional
2114: elasto-plastic decohesion has shown the importance of dynamics in ductile to brittle
2115: transition.~\cite{lobkovsky98}
2116: The current analysis slightly contradicts those results,
2117: and indicates that two major types of materials
2118: can be identified, which display different qualitative behavior:
2119: this is in agreement
2120: with the common idea that some materials are ductile, some other brittle.
2121:
2122: In fact, for hard materials ($\kappa<1$), the existence of a strong entropic crisis
2123: precludes dynamics of rearrangements to come into play: intrinsic
2124: properties of the material completely dominate the dynamical process of deformation.
2125: The existence of a dynamically driven transition between ductile and brittle
2126: behavior is still expected for soft materials ($\kappa>1$) for which the dynamics
2127: or arrangements can fully play its role.
2128: The interplay between dynamics and intrinsic glassy properties of material,
2129: as displayed in the current approach, seems a promising perspective.
2130:
2131: As far as fracture mechanics is concerned, this picture indicates that
2132: during the breaking of a hard glass, the motion of molecules in a material
2133: should be essentially localized around the surface of a crack,
2134: where there is some free-volume,
2135: whereas in the bulk, plastic deformation can occur only in places where a very large
2136: stress in concentrated. This is apparently related to a
2137: theoretical approach of fracture mechanics,
2138: recently introduced by Brener and Spatschek,~\cite{brener02}
2139: where the dynamics of the fracture is controlled by diffusion
2140: of molecules along the surface.
2141: For soft glassy materials, on the contrary, the propagation
2142: of a fracture is expected to induce more easily observable
2143: plastic deformation in the bulk of the material.
2144:
2145: %In the case of hard materials, ($\kappa<1$), the current model predicts that
2146: %there is either creep motion with a very small $\dot\epsilon$, depending
2147: %on the age of the system, or the system breaks up, and the rate of deformation achieves
2148: %the preceding value.
2149:
2150: %\subsubsection{Creep test}
2151: %I am now in position to draw a complete picture of a creep test.
2152: %This is to compare {\it e.g.} with the experimental results obtained
2153: %by Hassan and Boyce on PMMA.~\cite{hassan95}
2154:
2155: \section{Conclusion}
2156:
2157: The current work relies on the introduction of dynamical equations for
2158: the intensive quantity associated with entropic fluctuations
2159: of molecular configurations.
2160: In this work, this quantity is identified to free-volume,
2161: but other physical pictures could be given.
2162: For example, free-volume is related to the average number of contacts per molecule;
2163: in a foam, the deformation of the bubbles can be accounted for a effective
2164: free-volume that is determined by the packing of effective overlapping
2165: particles as those used in the simulations by Durian,~\cite{durian95,durian97}
2166: even though the relative proportions
2167: of a fluid mixture is kept constant; in a colloidal suspension, the existence
2168: of electrostatic screening contributes to an effective size of the particle
2169: that depends on the packing structure.
2170: In fact, the assumption made in this work is rather general,
2171: and many situation can be expected to lead to a dynamically
2172: driven effective free-volume; this may explain the ubiquity of glassy behavior.
2173:
2174: %The current approach might be seen as the opposite point of view
2175: %of Berthier {\it et al}~\cite{berthier00}
2176: %who introduced two temperatures to account for multiple timescales.
2177: %But here, the analysis comes out from a real-space picture of rearrangements:
2178: %there is no reasons whatsoever to assume that different types of rearrangements
2179: %should be governed by a unique barrier height and,
2180: %the multiplicity of timescales is directly provided by this simple remark,
2181: %and quite minimal assumption.
2182:
2183: The theoretical approach to glassy materials can be summarized as:
2184: how ergodicity breaks? or what ergodicity breaking is sufficient to account
2185: for slow modes of relaxation. The introduction of a whole distribution of timescales
2186: is a very general way to account for non-ergodicity.
2187: Here I assume that a much weaker ergodicity breaking is sufficient,
2188: and this is directly supported by the observation that fluctuation-dissipation
2189: theorem holds in a weak (but, not so weak) form.
2190: This indicates that the underlying
2191: distributions of state variables should not (in fact) be too weird,
2192: and that an effective width accounts for its essential features.
2193: However, since the system is out-of-equilibrium,
2194: this width must, at least, become a dynamical quantity.
2195: The equation proposed appear as a very minimal first order closure
2196: of the dynamics of an underlying distribution of voids.
2197:
2198: The set of curves displayed in figure~\ref{fig:relax3}
2199: is very reminiscent of the temperature
2200: dependency of relaxation spectra.~\cite{larson99} Obviously, the parameters
2201: of the theory should depend on temperature and other thermodynamical parameters,
2202: but here I purposely do not intend to address this question.
2203: There are (at least) two reasons for that: firstly, because
2204: glassy behavior is observed in many systems,
2205: in particular in suspensions or in granular materials,
2206: where thermodynamical temperature is not a relevant parameter.
2207: Secondly, because it seems to me more interesting to extract all possible conclusions
2208: from the simple equations proposed without trying to incorporate them into a more
2209: specific setup. I do think that an effective dynamically driven free-volume
2210: should emerge from microscopic dynamics under minimal assumptions.
2211: Therefore, it seems to me important to try to get
2212: some fundamental understanding of the consequences of such assumptions for a fixed
2213: set of parameters, without complicating the discussion.
2214: The dependency of the parameters of this theory on thermodynamic quantities
2215: is definitely an important issue to address in the future, but it first requires
2216: to have some intuition about the outcome of those equations.
2217: This work is intended to provide a grid of possible behavior displayed
2218: by simple equations, and I hope it can be helpful for further developments.
2219:
2220: Finally, one major interest of the current work is, in my opinion, that is gives
2221: a unified picture where different mechanisms for jamming coexist.
2222: This coexistence of jamming with non-linear rheology has been observed
2223: experimentally,
2224: but could not be captured by earlier model of glassy rheology.
2225: Moreover, it opens the way towards a description of glassy materials
2226: where thermodynamics (that determine the parameters of constitutive equations)
2227: and dynamics coexists.
2228: %Here, jamming results either from a phase transition,
2229: %or from the dynamics of a structural arrangements.
2230: %These simple equations also permit to clarify
2231: %the relative importance of solid-like or liquid-like features
2232: %and the importance of force and time scales.
2233: %Soft glassy materials are thus expected
2234: %to present some features of a yield stress and some of a power law fluid.
2235: %The observed behavior of a given material is expected to depend drastically
2236: %on the position of the experimental accessible window of forcing parameters
2237: %relatively to the internal characteristics of the material.
2238:
2239: \acknowledgements
2240:
2241: This work has primarily benefited from the support and encouragements of
2242: Jean Carlson and Jim Langer who received me among their people
2243: in the busy heights of UCSB;
2244: I acknowledge the opportunity they gave me to pursue my ideas,
2245: and their continuous interest for my work.
2246: I have enjoyed useful discussions with them, and also with others:
2247: Ralph Archuleta, Pascal Favreau, Anthony Foglia, Delphine Gourdon,
2248: Jacob Israelachvili, Daniel Lavall\'ee, Jean-Christophe Nave.
2249: I also would like to thank Daniel Bonn for giving me early versions of his papers,
2250: which were helpful in the writing of this manuscript.
2251:
2252: This work was supported by the W. M. Keck Foundation,
2253: and the NSF Grant No. DMR-9813752,
2254: and EPRI/DoD through the Program on Interactive Complex Networks.
2255:
2256: \thanks
2257:
2258: \appendix
2259: \section{\label{app:hopf} Hopf bifurcation at constant strain rate}
2260:
2261: \subsection{Isotropic limit}
2262: In order to carry out the stability analysis of the
2263: system~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1}),~(\ref{eqn:stzdil:iso:2}),
2264: it is convenient to introduce
2265: the variable,
2266: $$
2267: \phi = \exp\left[-{1\over\chi}\right]
2268: \quad.
2269: $$
2270:
2271: The system~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:iso:1}),~(\ref{eqn:stzdil:iso:2})
2272: reduces to two equations for $\sigma$ and $\chi$,
2273: \begin{eqnarray*}
2274: \dot\sigma &=& \mu\,\dot\epsilon-\mu\,E_0\,\exp\left[-{1\over\chi}\right]\sigma\\
2275: \dot\chi &=& -E_1 \exp\left[-{\kappa\over\chi}\right] + \alpha\,E_0\,\exp\left[-{1\over\chi}\right]\,\sigma^2
2276: \end{eqnarray*}
2277: and the jacobian of this dynamical system reads:
2278: $$
2279: \left(
2280: \matrix{
2281: -\mu\,E_0\,\phi &
2282: -\mu\,E_0\,\sigma\,\phi\,\log[\phi]^2\cr
2283: 2\,E_0\,\alpha\,\sigma\,\phi&
2284: \left(-\kappa\,E_1\phi^\kappa + E_0\,\alpha \,{\sigma }^2\,\phi\right)\log[\phi]^2\cr
2285: }
2286: \right)
2287: \quad.
2288: $$
2289:
2290: In the steady state, the variable $\sigma$ verifies,
2291: $$
2292: \sigma^2 = {E_1\over\alpha\,E_0}\,\phi^{\kappa-1}
2293: $$
2294: and $\phi$ takes on the value,
2295: $$
2296: \phi_{\rm s}=\left({\alpha\,\dot\epsilon^2\over E_0\,E_1}\right)^{1\over\kappa+1}
2297: \quad.
2298: $$
2299: For all $\kappa>0$, $\phi_{\rm s}$ is a strictly increasing function of $\dot\epsilon$.
2300: In terms of $\phi_{\rm s}(\dot\epsilon)$, the eigenvalues $\lambda$ of the jacobian verify:
2301: \begin{eqnarray*}
2302: \lambda^2 +
2303: \lambda\;\left(
2304: \mu\,E_0\,\phi_{\rm s}
2305: + {E_1\,(\kappa-1)}\,\phi_{\rm s}^\kappa
2306: \;\ln\left[\phi_{\rm s}\right]^2
2307: \right)
2308: &+&\\
2309: {\mu\,E_0\,E_1\,(\kappa+1)}\;\phi_{\rm s}^{\kappa+1}\,\ln\left[\phi_{\rm s}\right]^2
2310: &=&0
2311: \quad,
2312: \end{eqnarray*}
2313: and the Hopf bifurcation is determined by a critical value of $\mu$ as a function
2314: of $\phi_{\rm s}$:
2315: $$
2316: \mu_{\rm hopf} = {E_1\over E_0}\,(1-\kappa)\,\phi_{\rm s}^{\kappa-1}\,\ln[\phi_{\rm s}]^2
2317: $$
2318: or,
2319: $$
2320: \mu_{\rm hopf} = {E_1\over E_0}\,{1-\kappa\over(\kappa+1)^2}\;
2321: \left({\alpha\,\dot\epsilon^2\over E_0\,E_1}\right)^{{\kappa-1\over\kappa+1}}\;
2322: \ln\left[{\alpha\,\dot\epsilon^2\over E_0\,E_1}\right]^2
2323: \quad.
2324: $$
2325:
2326: \subsection{General case}
2327: In the general case, the system~(\ref{eqn:sigma:0}),~(\ref{eqn:stzdil:lin:1}-\ref{eqn:stzdil:lin:3})
2328: involves one additional variable ($\Delta$), and the bifurcation analysis is somewhat more complicated.
2329: In the steady state, the relation between strain rate and free-volume can be rewritten as,
2330: $$
2331: \dot\epsilon = {
2332: \sqrt{E_0\,\mu_0}\,E_1\;\phi^{{\kappa/2}}
2333: \over
2334: \sqrt{\alpha}\;
2335: \sqrt{E_0\alpha\,\phi^{-\kappa}
2336: + E_1\mu_0\,\phi^{-1}}
2337: }
2338: \quad.
2339: $$
2340: Since this expression is an strictly increasing function of $\phi$,
2341: it is possible to carry out the analysis in terms of $\phi=\phi_{\rm s}$ as an effective
2342: parameter. When the strain rate varies between $0$ and $\dot\epsilon^*$,
2343: $\phi$ varies on the interval $[0,1]$. The values above $1$ are meaningless
2344: because there is no solution to the equations of motion.
2345:
2346: The Hopf criterion can be written in the form,
2347: $$
2348: {\cal H}=A\,\mu^2+B\,\mu+C=0
2349: $$
2350: %\begin{widetext}
2351: with:
2352: \begin{eqnarray*}
2353: A &=& {E_0}^2\,\alpha \,\epsilon_0\,{\phi }^2\,
2354: \Big( \mu_0\,\left( 2\,E_0\,\alpha \,\phi +
2355: E_1\,\mu_0\,{\phi }^{\kappa } \right) \Big.\\
2356: &+& \Big.
2357: \alpha \,\epsilon_0\,\left( E_0\,\alpha \,\kappa \,\phi +
2358: E_1\,\left( 1 + \kappa \right) \,\mu_0\,{\phi }^{\kappa } \right) \,
2359: {\log (\phi )}^2 \Big)\\
2360: B &=& E_0\,E_1\,{\phi }^{1 + \kappa }\,
2361: \Big( {\mu_0}^2\,
2362: \left( 2\,E_0\,\alpha \,\phi+ E_1\,\mu_0\,{\phi }^{\kappa } \right)
2363: + \alpha \,\epsilon_0\,\mu_0\,\Big.\\
2364: \Big.&&\times \left(
2365: E_0\,\alpha \,\left( 2\,\kappa-3 \right) \,\phi +
2366: 2\,E_1\,\left( \kappa -1 \right) \,\mu_0\,{\phi }^{\kappa }
2367: \right)\, {\log (\phi )}^2 \Big.\\
2368: \Big.&+&
2369: {\alpha }^2\,{\epsilon_0}^2\,\left( \kappa -1 \right) \,
2370: \left( E_0\,\alpha \,\kappa \,\phi +
2371: E_1\,\left( 1 + \kappa \right) \,\mu_0\,{\phi }^{\kappa } \right) \,
2372: {\log (\phi )}^4 \Big)\\
2373: C &=& {E_1}^2\,\left(\kappa -1 \right) \,\mu_0\,{\phi }^{2\,\kappa }\,
2374: \left( E_0\,\alpha \,\phi + E_1\,\mu_0\,{\phi }^{\kappa } \right) \,
2375: {\log (\phi )}^2\, \\
2376: &&\times\left( \mu_0 + \alpha \,\epsilon_0\,\left(\kappa -1 \right) \,{\log (\phi )}^2 \right)
2377: \end{eqnarray*}
2378: %\end{widetext}
2379: The quantity ${\cal H}$ has the opposite sign of the real part
2380: of the complex eigenvalues of the jacobian: the steady state is stable,
2381: iff ${\cal H}>0$.
2382:
2383: For $\kappa>1$, and for any $\phi\in[0,1]$, $C$ is positive: the two values of $\mu$,
2384: solutions of ${\cal H}=0$ have the same sign. This sign can change only
2385: at a point where $C$ vanishes, that is $\phi=0$ or $1$. The sign shared by these
2386: solutions on the interval $\phi\in[0,1]$ can then be evaluated by taking the limit
2387: value of the coefficient $B$ close to 0 or 1: it is positive.
2388: Therefore, the equation ${\cal H}=0$ has no real positive solution for $\kappa>1$:
2389: there is no Hopf instability. It can be checked that there is no other instability,
2390: therefore, for $\kappa>1$, the system is stable.
2391:
2392:
2393: I am now considering the case when $\kappa<1$.
2394: To get a first idea of the phase diagram, in the parameter space $\{\mu,\phi\}$,
2395: the quantity ${\cal H}$ can be evaluated on the line $\mu=0$: ${\cal H}(\mu=0,\phi)=C$.
2396: On this line, the positiveness of ${\cal H}$ is then equivalent to,
2397: $$
2398: {\log (\phi )}^2 > {\mu_0\over\alpha \,\epsilon_0\,\left(1-\kappa\right) }
2399: \quad,
2400: $$
2401: which requires that:
2402: $$
2403: \phi < \phi_{\rm hopf}^-(\mu=0) =
2404: \exp\left[-\sqrt{\mu_0\over\alpha \,\epsilon_0\,\left( 1-\kappa\right) }\right]
2405: \quad.
2406: $$
2407:
2408: For $\phi>\phi_{\rm hopf}^-$ the coefficient $C$ is negative, hence, equation ${\cal H}=0$
2409: has a single positive solution.
2410: On the interval $[0,\phi_{\rm hopf}^-]$ this equation admits two solutions which are positive
2411: because $B$ is negative close to $\phi=0$: the solution $\mu_{\rm hopf}^-$ which vanishes
2412: at the point $\phi_{\rm hopf}^-$ and the solution $\mu_{\rm hopf}^+$ which exists on the whole
2413: interval $[0,1]$. The unstability occurs for $\mu\in[\mu_{\rm hopf}^-,\mu_{\rm hopf}^+]$.
2414: At small shear rates, or small $\phi$, the criterion ${\cal H}$ is dominated by:
2415: %\begin{eqnarray*}
2416: %A &\simeq& (E_0\,\alpha \,\epsilon_0)^2\,{\phi }^2\,
2417: % \left( E_0\,\alpha \,\kappa \,\phi +
2418: % E_1\,\left( 1 + \kappa \right) \,\mu_0\,{\phi }^{\kappa } \right) \,
2419: % {\log (\phi )}^2\\
2420: %B &\simeq& E_0\,E_1\,{\alpha }^2\,{\epsilon_0}^2\,{\phi }^{1 + \kappa }\,
2421: %\left(\kappa-1\right)\,
2422: % \left( E_0\,\alpha \,\kappa \,\phi+
2423: % E_1\,\left( 1 + \kappa \right) \,\mu_0\,{\phi }^{\kappa } \right) \,
2424: % {\log (\phi )}^4\\
2425: %C &\simeq& {E_1}^2\,\left(\kappa-1\right)^2\,\alpha \,\epsilon_0\,\mu_0\,{\phi }^{2\,\kappa }\,
2426: % \left( E_0\,\alpha \,\phi + E_1\,\mu_0\,{\phi }^{\kappa } \right) \,
2427: % {\log (\phi )}^4
2428: %\end{eqnarray*}
2429: \begin{eqnarray*}
2430: A &\simeq& E_0^2\,E_1\,\alpha^2\,\epsilon_0^2\,\mu_0\,
2431: \left( 1 + \kappa \right) \,{\phi }^{2+\kappa }\,{\log (\phi )}^2\\
2432: B &\simeq& E_0\,E_1^2\,{\alpha }^2\,{\epsilon_0}^2\,\mu_0\,\left(\kappa^2-1\right)\,
2433: {\phi }^{1+2\kappa } \, {\log (\phi )}^4\\
2434: C &\simeq& {E_1}^3\,\alpha \,\epsilon_0\,\mu_0^2\,\left(\kappa-1\right)^2\,
2435: {\phi }^{3\,\kappa }\,{\log (\phi )}^4
2436: \end{eqnarray*}
2437: whence,
2438: \begin{eqnarray*}
2439: {B\over A} &\simeq& {E_1\over E_0}\,(\kappa-1)
2440: {\phi }^{\kappa-1} \, {\log (\phi )}^2\\
2441: {C\over A} &\simeq& {E_1\,\mu_0\over E_0^2\,\alpha\,\epsilon_0}\,
2442: {\left(\kappa-1\right)^2\over\kappa+1}\,
2443: {\phi }^{2\,\kappa-2}\,{\log (\phi )}^2\\
2444: {C\over B} &\simeq& {\mu_0\over E_0\,\alpha\,\epsilon_0}\,
2445: {\kappa-1\over\kappa+1}\,{\phi }^{\kappa-1}
2446: \end{eqnarray*}
2447: The solution $\mu_{\rm hopf}^+$ can then be identified to:
2448: $$
2449: \mu_{\rm hopf}^+ \simeq -{B\over A}\simeq
2450: {E_1\over E_0}\,(1-\kappa) {\phi }^{\kappa-1} \, {\log (\phi )}^2
2451: \quad,
2452: $$
2453: close to $\phi=0$.
2454: The same expression was found in the isotropic limit,
2455: but here, the relation between $\phi$ and the shear rate,
2456: $\dot\epsilon$ is different.
2457: The solution $\mu_{\rm hopf}^-$ can be expanded as,
2458: $$
2459: \mu_{\rm hopf}^- \simeq -{C\over B}\simeq
2460: {\mu_0\over E_0\,\alpha\,\epsilon_0}\,
2461: {1-\kappa\over\kappa+1}\,{\phi }^{\kappa-1}
2462: $$
2463: %\bibliography{thermo}
2464:
2465: \begin{thebibliography}{100}
2466:
2467: \bibitem{goetze92}
2468: W. G\"otze and L. Sj\"ogren, Rep. Prog. Phys. {\bf 55}, 241 (1992).
2469:
2470: \bibitem{ediger96}
2471: M.~D. Ediger, C.~A. Angell, and S.~R. Nagel, J. Phys.: Condens. Matter {\bf
2472: 100}, 13200 (1996).
2473:
2474: \bibitem{larson99}
2475: R. Larson, {\em The structure and rheology of complex fluids} (Oxford
2476: University Press, New York, 1999).
2477:
2478: \bibitem{angell00}
2479: C.~A. Angell {\it et~al.}, J. Appl. Phys. {\bf 88}, 3113 (2000).
2480:
2481: \bibitem{berthier01}
2482: L. Berthier and J.-L. Barrat, Phys. Rev. E {\bf 63}, 012503 (2001).
2483:
2484: \bibitem{yamamoto97}
2485: R. Yamamoto and A. Onuki, Europhysics Letters {\bf 40}, 61 (1997).
2486:
2487: \bibitem{yamamoto98}
2488: R. Yamamoto and A. Onuki, Phys. Rev. E {\bf 58}, 3515 (1998).
2489:
2490: \bibitem{bonn02a}
2491: D. Bonn {\it et~al.}, Laponite: aging and shear rejuvenation of a colloidal
2492: glass, in preparation, 2002.
2493:
2494: \bibitem{bonn02b}
2495: D. Bonn {\it et~al.}, Rheology of soft glassy materials, in preparation, 2002.
2496:
2497: \bibitem{sollich97}
2498: P. Sollich, F. Lequeux, P. H\'ebraud, and M.~E. Cates, Phys. Rev. Lett. {\bf
2499: 78}, 2020 (1997).
2500:
2501: \bibitem{sollich98}
2502: P. Sollich, Phys. Rev. E {\bf 58}, 738 (1998).
2503:
2504: \bibitem{hebraux98}
2505: P. H\'ebraux and F. Lequeux, Phys. Rev. Lett. {\bf 81}, 2934 (1998).
2506:
2507: \bibitem{berthier00}
2508: L. Berthier, J.-L. Barrat, and J. Kurchan, Phys. Rev. E {\bf 61}, 5464
2509: (2000).
2510:
2511: \bibitem{fielding00}
2512: S. Fielding, P. Sollich, and M. Cates, J. Rheol. {\bf 44}, 323 (2000).
2513:
2514: \bibitem{bouchaud96}
2515: J. Bouchaud, L. Cugliandolo, J. Kurchan, and M. M\'ezard, Physica A {\bf 226},
2516: 243 (1996).
2517:
2518: \bibitem{monthus96}
2519: C. Monthus and J.-P. Bouchaud, J. Phys. A {\bf 29}, 3847 (1996).
2520:
2521: \bibitem{derec01}
2522: C. Derec, A. Ajdari, and F. Lequeux, Eur. Phys. J. E {\bf 4}, 355 (2001).
2523:
2524: \bibitem{liu98}
2525: A.~J. Liu and S.~R. Nagel, Nature {\bf 396}, 21 (1998).
2526:
2527: \bibitem{ohern01}
2528: C.~S. O'Hern, S.~A. Langer, A.~J. Liu, and S.~R. Nagel, Phys. Rev. Lett. {\bf
2529: 86}, 111 (2001).
2530:
2531: \bibitem{knight95}
2532: J.~B. Knight {\it et~al.}, Phys. Rev. E {\bf 51}, 3957 (1995).
2533:
2534: \bibitem{boutreux97}
2535: T. Boutreux and P.~G. de~Gennes, Physica A {\bf 244}, 59 (1997).
2536:
2537: \bibitem{nowak98}
2538: E.~R. Nowak {\it et~al.}, Phys. Rev. E {\bf 57}, 1971 (1998).
2539:
2540: \bibitem{falk98}
2541: M.~L. Falk and J.~S. Langer, Phys. Rev. E {\bf 57}, 7192 (1998).
2542:
2543: \bibitem{falk00}
2544: M.~L. Falk and J.~S. Langer, M.R.S. Bulletin {\bf 25}, 40 (2000).
2545:
2546: \bibitem{spaepen77}
2547: F. Spaepen, Acta Metall. {\bf 25}, 407 (1977).
2548:
2549: \bibitem{argon79a}
2550: A. Argon, Acta Metall. {\bf 27}, 47 (1979).
2551:
2552: \bibitem{argon79b}
2553: A. Argon and H. Kuo, Mater. Sc. Eng. {\bf 39}, 101 (1979).
2554:
2555: \bibitem{spaepen81}
2556: F. Spaepen and A. Taub, in {\em Physics of Defects}, {\em Les Houches
2557: Lectures, Session XXXV}, edited by J.-P. Balian and M. Kleman (North-Holland,
2558: Amsterdam, 1981), p.\ 133.
2559:
2560: \bibitem{argon83}
2561: A. Argon and L. Shi, Acta Metall. {\bf 31}, 499 (1983).
2562:
2563: \bibitem{lemaitre01b}
2564: A. Lema\^{\i}tre, cond-mat/0108442, 2001.
2565:
2566: \bibitem{dieterich78}
2567: J.~H. Dieterich, Pure Appl. Geophys. {\bf 116}, 790 (1978).
2568:
2569: \bibitem{dieterich79}
2570: J.~H. Dieterich, J. Geophys. Res. {\bf 84}, 2161 (1979).
2571:
2572: \bibitem{rice83}
2573: J. Rice and A. Ruina, J. Appl. Mech. {\bf 105}, 343 (1983).
2574:
2575: \bibitem{radjai98a}
2576: F. Radjai, D.~E. Wolf, M. Jean, and J.-J. Moreau, Phys. Rev. Lett. {\bf 80},
2577: 61 (1998).
2578:
2579: \bibitem{radjai98b}
2580: F. Radjai and D.~E. Wolf, Granular Matter {\bf 1}, 3 (1998).
2581:
2582: \bibitem{lemaitre01a}
2583: A. Lema\^{\i}tre, cond-mat/0107422, 2001.
2584:
2585: \bibitem{schlesinger84}
2586: M.~F. Schlesinger and E. Montroll, Proc. Nat. Acad. Sc. US {\bf 81}, 1280
2587: (1984).
2588:
2589: \bibitem{langer99}
2590: J.~S. Langer, Phys. Rev. E {\bf 62}, 1351 (1999).
2591:
2592: \bibitem{langer01}
2593: J.~S. Langer, Phys. Rev. E {\bf 64}, 011504 (2001).
2594:
2595: \bibitem{saulnier02}
2596: F. Saulnier, E. Raphael, and P.-G. de~Gennes, Dewetting of thin polymer films
2597: near the glass transition, cond-mat/0201528, 2002.
2598:
2599: \bibitem{lubliner90}
2600: J. Lubliner, {\em Plasticity Theory} (Macmillan, New York, 1990).
2601:
2602: \bibitem{stillinger84}
2603: F.~H. Stillinger and T.~A. Weber, Science {\bf 225}, 983 (1984).
2604:
2605: \bibitem{cohen59}
2606: M.~H. Cohen and D. Turnbull, J. Chem. Phys. {\bf 31}, 1164 (1959).
2607:
2608: \bibitem{turnbull61}
2609: D. Turnbull and M.~H. Cohen, J. Chem. Phys. {\bf 34}, 120 (1961).
2610:
2611: \bibitem{turnbull70}
2612: D. Turnbull and M.~H. Cohen, J. Chem. Phys. {\bf 52}, 3038 (1970).
2613:
2614: \bibitem{edwards89}
2615: S.~F. Edwards and R.~B.~S. Oakeshott, Physica A {\bf 157}, 1080 (1989).
2616:
2617: \bibitem{mehta89}
2618: A. Mehta and S.~F. Edwards, Physica A {\bf 157}, 1091 (1989).
2619:
2620: \bibitem{edwards94}
2621: S. Edwards, in {\em Granular Matter: An Interdisciplinary Approach}, edited by
2622: A. Mehta (Springer-Verlag, New-York, 1994), pp.\ 121--140.
2623:
2624: \bibitem{savage79}
2625: S.~B. Savage, J. Fluid. Mech. {\bf 92}, 53 (1979).
2626:
2627: \bibitem{coppersmith96}
2628: S. Coppersmith {\it et~al.}, Phys. Rev. E {\bf 53}, 4673 (1996).
2629:
2630: \bibitem{cloitre00}
2631: M. Cloitre, R. Borrega, and L. Leibler, Phys. Rev. Lett. {\bf 85}, 4819
2632: (2000).
2633:
2634: \bibitem{bureau02}
2635: L. Bureau, T. Baumberger, and C. Caroli, Rheological aging and rejuvenation in
2636: solid friction contacts, cond-mat/0202245, 2002.
2637:
2638: \bibitem{yoshino98}
2639: H. Yoshino, Phys. Rev. Lett. {\bf 81}, 1493 (1998).
2640:
2641: \bibitem{boon91}
2642: J.-P. Boon and S. Yip, in {\em Molecular Hydrodynamics}, edited by McGraw-Hill
2643: (Dover, New York, 1991).
2644:
2645: \bibitem{debenedetti01}
2646: P.~G. Debenedetti and F.~H. Stillinger, Nature {\bf 400}, 259 (2001).
2647:
2648: \bibitem{bonn99}
2649: D. Bonn {\it et~al.}, langmuir {\bf 15}, 7534 (1999).
2650:
2651: \bibitem{kob95}
2652: W. Kob and H.~C. Andersen, Phys. Rev. E {\bf 52}, 4134 (1995).
2653:
2654: \bibitem{megen94}
2655: W. van Megen and S. Underwood, Phys. Rev. E {\bf 49}, 4206 (1994).
2656:
2657: \bibitem{gopal99}
2658: A. Gopal and D. Durian, J. Coll. Int. Sc. {\bf 213}, 169 (1999).
2659:
2660: \bibitem{hebraud97}
2661: P. H\'ebraud, F. Lequeux, J. Munch, and D. Pine, Phys. Rev. Lett. {\bf 78},
2662: 4657 (1997).
2663:
2664: \bibitem{megen98}
2665: W. van Megen, T. Mortensen, S. Williams, and J. {M\"uller}, Phys. Rev. E {\bf
2666: 58}, 6073 (1998).
2667:
2668: \bibitem{cipelletti00}
2669: L. Cipelletti, S. Manley, R. Ball, and D. Weitz, Phys. Rev. Lett. {\bf 84},
2670: 2275 (2000).
2671:
2672: \bibitem{hassan95}
2673: O.~A. Hassan and M.~C. Boyce, Pol. Engineer. and Science {\bf 35}, 331
2674: (1995).
2675:
2676: \bibitem{drummond00}
2677: C. Drummond and J. Israelachvili, Macromolecules {\bf 33}, 4910 (2000).
2678:
2679: \bibitem{drummond01}
2680: C. Drummond and J. Israelachvili, Phys. Rev. E {\bf 63}, 041506 (2001).
2681:
2682: \bibitem{persson00}
2683: Persson, in {\em "Sliding Friction : Physical Principles and Applications
2684: (Nanoscience and Technology)"}, edited by "" ("Springer Verlag", "", 2000).
2685:
2686: \bibitem{thompson92}
2687: P.~A. Thompson, G.~S. Grest, and M.~O. Robbins, Phys. Rev. Lett. {\bf 68},
2688: 3448 (1992).
2689:
2690: \bibitem{yoshizawa93}
2691: H. Yoshizawa and J. Israechvili, J. Phys.: Condens. Matter {\bf 97}, 11300
2692: (1993).
2693:
2694: \bibitem{demirel96a}
2695: A.~L. Demirel and S. Granick, Phys. Rev. Lett. {\bf 77}, 2261 (1996).
2696:
2697: \bibitem{demirel96b}
2698: A.~L. Demirel and S. Granick, Phys. Rev. Lett. {\bf 77}, 4330 (1996).
2699:
2700: \bibitem{thompson91}
2701: P.~A. Thompson and G.~S. Grest, Phys. Rev. Lett. {\bf 67}, 1751 (1991).
2702:
2703: \bibitem{nasuno97}
2704: S. Nasuno, A. Kudrolli, and J. Gollub, Phys. Rev. Lett. {\bf 79}, 949 (1997).
2705:
2706: \bibitem{losert00}
2707: S.~N. W.~Losert, J.-C.~G\'eminard and J. Gollub, Phys. Rev. E {\bf 61}, 4060
2708: (2000).
2709:
2710: \bibitem{lobkovsky98}
2711: A.~E. Lobkovsky and J.~S. Langer, Phys. Rev. E {\bf 58}, 1568 (1998).
2712:
2713: \bibitem{brener02}
2714: E.~A. Brener and R. Spatschek, Fast crack propagation by surface diffusion,
2715: submitted, 2002.
2716:
2717: \bibitem{durian95}
2718: D. Durian, Phys. Rev. Lett. {\bf 75}, 4780 (1995).
2719:
2720: \bibitem{durian97}
2721: D. Durian, Phys. Rev. E {\bf 55}, 1739 (1997).
2722:
2723: \end{thebibliography}
2724:
2725: \end{document}
2726: