• Keine Ergebnisse gefunden

Transport and aging in glassy systems

N/A
N/A
Protected

Academic year: 2022

Aktie "Transport and aging in glassy systems"

Copied!
163
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Transport and Aging in Glassy Systems

Transport und Altern in glasartigen Systemen

Dissertation

zur Erlangung des akademischen Grades des Doktors der Naturwissenschaften (Dr. rer. nat.)

an der Universit¨at Konstanz Fachbereich Physik

vorgelegt von

B ERND R INN

Dissertation der Universit¨at Konstanz Datum der m ¨undlichen Pr ¨ufung: 9.11.2001 Referenten: Priv. Doz. Dr. P. Maaß

Prof. Dr. P. Nielaba Prof. Dr. R. Schilling

(2)

Special edition of the book with the ISBN: 3-89825- 366-X

dissertation.de - Verlag im Internet GmbH Pestalozzistr. 9

10 625 Berlin

URL: http://www.dissertation.de

(3)

Kurzzusammenfassung

Thema dieser Arbeit ist die Untersuchung der Gleichgewichts– und Nichtgleichge- wichts–Dynamik, die von thermich aktivierten Sprungprozessen in ungeordneten Energielandschaften herr ¨uhrt. Modelle, welche von der Sprungbewegung eines “Zu- standspunktes” in einer ungeordneten komplexen Energielandschaft ausgehen, wer- den seit l¨angerem zur Beschreibung der spezifischen Eigenschaften von glasartigen Materialien verwendet (siehe z. B. [Gol69]).

Im ersten Teil der Arbeit wird eine Theorie des effektiven Mediums f ¨ur die Be- schreibung des dispersiven Transports in amorphen Materialien entwickelt, welche die Einbeziehung asymmetrischer ¨Ubergangsraten erlaubt. Die vorgeschlagene Effek- tive–Medien–Theorie wird auf Ein–Teilchen–Systeme und Systeme mit Harter–Kugel–

Abstoßung angewandt. Bei den Ein–Teilchen–Systemen werden beispielhaft die Dif- fusionskoeffizienten im Frequenzraum berechnet und mit numerischen Ergebnissen aus Monte–Carlo Simulationen verglichen. F ¨ur verd ¨unnte Gittergase mit Harter–Ku- gel–Wechselwirkung berechnen wir die Leitf¨ahigkeit f ¨ur verschiedene Temperaturen und Konzentrationen und zeigen, daß aus der Effektiven–Medien–Theorie Skalenei- genschaften f ¨ur die frequenzabh¨angige Leitf¨ahigkeit folgen, wie sie f ¨ur ionenleitende Gl¨aser gemessen worden sind.

Im zweiten Teil der Arbeit untersuchen wir das Nichtgleichgewichts–Ph¨anomen der physikalischen Alterung in glasartigen Materialien. Dieses wird bei vielen ver- schiedenartigen Systemen beobachtet, wenn diese auf experimentell zug¨anglichen Zeitskalen keinen Gleichgewichtszustand erreichen. Dabei gehen wir von einem asymptotisch exponentiellen Abfall der Energiezustandsdichte aus. Ein solcher Ab- fall, der das Auftreten einer Glas ¨ubergangstemperatur impliziert, wird nach der sta- tistischen Theorie der Extremalwerte f ¨ur die Verteilung der tiefsten Energien in einer komplexen Energielandschaft erwartet und konnte auch in Molekulardynamiksimu- lationen von Gl¨asern nachgewiesen werden. In einer idealisierten Beschreibung be- trachten wir eine streng exponentielle Zustandsdichte. In Energielandschaften mit dieser Zustandsdichte untersuchen wir Sprungprozesse und das mit diesen verbun- dene Zerfallen der Korrelationen. Die Korrelationen zerfallen umso langsamer, je mehr Zeit verstrichen ist seit dem Abk ¨uhlen des Systems unter die Glas ¨ubergangs- temperatur. Dieses Ph¨anomen bezeichnet man als physikalisches Altern. Wir zeigen, daß das Modell ein ¨uberraschend reiches Szenario verschiedener Skaleneigenschaften wie sublineares Altern, ultra–langsames Altern, ein verallgemeinertes Skalenregime und Vielfachskalierung besitzt. Wir formulieren ein (allgemeiner anwendbares) “Kon- zept des partiellen Gleichgewichtes”, das wir f ¨ur unser Modell mathematisch exakt behandeln k ¨onnen und das es uns erlaubt, die Skaleneigenschaften des Modells kor- rekt vorherzusagen. Schließlich zeigen wir, daß entgegen der naiven Erwartung eine

“Mean–Field–Behandlung”, bei der die Energielandschaft zeitlich schnell fluktuiert, selbst in hochdimensionalen Energielandschaften zu qualitativ anderen Ergebnissen f ¨uhrt.

(4)
(5)

Contents

List of Figures v

Notation vii

1 Introduction 1

1.1 Glassy systems . . . 1

1.2 Relaxation spectra of glasses . . . 3

1.3 Physical aging . . . 4

1.4 Hopping models of glassy systems . . . 9

1.5 Goals of this work . . . 11

I Dispersive transport in disordered energy landscapes 13 2 Hopping in disordered energy landscapes 15 2.1 Equations of motion . . . 15

2.2 Computing transport quantities . . . 17

2.3 Exactly solvable limits of the master equation . . . 18

2.3.1 Short–time limit . . . 18

2.3.2 Long–time limit of the one–dimensional system . . . 18

3 The Effective Medium Approximation 21 3.1 The history of the EMA . . . 21

3.2 The idea of the EMA . . . 21

3.3 The self–consistency equation . . . 22

3.4 Generalization of the EMA to the case of asymmetric rates . . . 26

3.5 Temperature rescaled EMA . . . 27

3.6 The transition rate . . . 28

3.7 Special cases and comparison with Monte–Carlo simulations . . . 30

(6)

3.7.1 Equations and results for a constant DOS . . . 31

3.7.2 Equations and results for a Gamma DOS . . . 31

3.7.3 Equations and results for a bimodal DOS . . . 34

4 AC–conductivity in ion conducting glasses 41 4.1 Experimental findings for the ac–conductivity . . . 41

4.2 Extension of the EMA to dilute hard–core systems . . . 43

4.2.1 Derivation of an effective one–particle master equation . . . 44

4.2.2 Density of States . . . 45

4.3 The self–consistency equation . . . 46

4.4 Results of the EMA . . . 46

II Physical aging in glassy systems 51 5 Some results of the statistical theory of extreme values 53 5.1 Definition of the problem . . . 53

5.2 The convergence to types theorem . . . 54

5.3 The FISHER–TIPPETTtheorem . . . 55

5.4 Maximum Domains of Attraction . . . 56

5.4.1 MDA of the FRECHET´ distribution . . . 57

5.4.2 MDA of the WEIBULLdistribution . . . 59

5.4.3 MDA of the GUMBELdistribution . . . 59

5.5 How to treat sums of PARETOdistributed random variables . . . 60

6 The Quenched Jump Model 63 6.1 Motivation and definition of the model . . . 63

6.2 Existence of a stationary distribution — the glassy phase . . . 65

6.3 Initial condition and the aging function . . . 67

6.4 Precursors of aging for>1. . . 67

6.5 Failure of the EMA to account for aging . . . 69

(7)

Contents

7 A mean–field approximation of the QJM 71

7.1 Idea of the Annealed Jump Model . . . 71

7.2 Mapping of the case with>0onto the trap model . . . 71

7.3 The high–temperature regime0>1 . . . 72

7.4 Scaling: A simple argument . . . 74

7.5 Exact treatment . . . 75

8 The Partial Equilibrium Concept applied to the QJM 79 8.1 Idea of the Partial Equilibrium Concept . . . 79

8.2 Scaling and generalized scaling from simple arguments . . . 80

8.2.1 Limit1 !1 . . . 82

8.2.2 Limit1 !0 . . . 83

8.3 Comparison of the PEC with Monte–Carlo results . . . 84

8.3.1 Scaling properties . . . 84

8.3.2 Full comparison with simulations . . . 85

8.4 Exact evaluation of the partial equilibrium formula . . . 90

8.4.1 First scaling regime:F1(1) . . . 91

8.4.2 Limit1 !1 . . . 91

8.4.3 Limit1 !0 . . . 92

8.4.4 Second scaling regime: F2(2) . . . 93

8.4.5 Limit2 !1 . . . 94

8.4.6 Limit2 !0 . . . 95

8.5 The influence of the form of hopping rates on the scaling behavior . . . . 95

8.6 Randomly distributedparameters and multi–scaling . . . 96

9 Summary and outlook 103

A Calculation of the diffusion coefficient in the trap model I

B ABELian theorems III

C Calculation and limits of the lattice matrix elementG(d)

00

V C.1 Computation ford=1andd=2 . . . V C.2 Computation ford=3 . . . V C.3 Limits . . . VI

(8)

D Continuous–Time Monte–Carlo simulations IX D.1 General description . . . IX D.2 Technical aspects . . . X D.3 Calculation ofD^ from simulations . . . X E Summary of formulae of the asymmetric EMA for special DOS XIII E.1 Formulae and results for a constant DOS . . . XIII E.1.1 Exact limites . . . XIII E.1.2 Formulae of the asymmetric EMA . . . XIII E.1.3 Temperature rescaling . . . XIV E.2 Equations and results for a Gamma DOS . . . XVI E.2.1 Exact limites . . . XVI E.2.2 Formulae of the asymmetric EMA . . . XVI E.2.3 Temperature rescaling . . . XVII E.3 Equations and results for a bimodal DOS . . . XVII E.3.1 Exact limites . . . XVII E.3.2 Formulae of the asymmetric EMA . . . XIX E.3.3 Temperature rescaling . . . XX E.4 Formulae and results for the dilute hard–core system . . . XX E.4.1 How to determine the fugacity . . . XX E.4.2 Limits of the self consistency equation . . . XXI E.4.3 D(!)obtained by averaging over the energy . . . XXII E.4.4 Low frequency limit: Relation between(!)andD(!) . . . XXIII

F Number of distinct visited states XXIX

F.1 Trap model (=0) . . . XXIX F.2 One dimension (0<1) . . . XXX F.3 General arguments . . . XXXI F.4 Monte–Carlo results . . . XXXI

G C(t;˜ S)in the limitS!1 XXXIII

Bibliography XXXV

Index XL

(9)

List of Figures

1.1 (a)ViscosityversusTg=Tfor different glasses. (b)Correlation and re- laxation times obtained from different measurements for the organic

glass o–terphenyl. . . 2

1.2 Typical relaxation spectrum of a glass for two temperaturesT1<T2. . . . 3

1.3 Dependence of the loss–part of the dielectric permeability of glycerol on the waiting time. . . 4

1.4 Torque when applying a torsion to polycarbonate. . . 5

1.5 Parametric plot of the relaxation versus the correlation function for a sequence of increasing waiting times for a mean–field spin glass model. 8 3.1 Illustration of the energy levels involved in a jump. . . 28

3.2 Tracer diffusion coefficient for a constant DOS ind=1. . . 32

3.3 Tracer diffusion coefficient for a constant DOS ind=3. . . 33

3.4 Tracer diffusion coefficient for a Gamma DOS ind=1. . . 35

3.5 Tracer diffusion coefficient for a Gamma DOS ind=3. . . 36

3.6 Tracer diffusion coefficient for a bimodal DOS ind=3withq=0:7. . . . 38

3.7 Tracer diffusion coefficient for a bimodal DOS ind=3withq=0:25. . . . 39

4.1 Effective scaling of ac–conductivity data for (AgI)x(AgPO3)1-x. . . 42

4.2 Effective scaling of ac–conductivity data for (Na2O)x(B2O3)1-x. . . 42

4.3 EMA result of the dilute hard–core lattice gas. . . 47

4.4 Scaling of the EMA diffusion coefficient for different temperatures and concentrations (“generic” version). . . 48

4.5 Scaling of the EMA diffusion coefficient for different temperatures and concentrations (BNN version). . . 48

4.6 Scaling of the EMA diffusion coefficient for different temperatures and concentrations (T=cversion). . . 49

4.7 Mastercurve of (Na2O)x(GeO2)1-x glasses compared to the EMA mas- tercurve. . . 49

(10)

6.1 Sketch of a rugged energy landscape of a glas with various metastable

minima. . . 64

6.2 Mean square displacement of the QJM ford=10,=1=3and=1=4. . . 66

6.3 Aging functionC(t+tw;tw)of the QJM from Monte–Carlo simulations for(d;;)=(10;1=4;3=8)and different waiting timestw. . . 68

7.1 Aging function (a)C(t+tw;tw)and (b)1-C(t+tw;tw)from eq. (7.20) (thin lines) and from simulations of the AJM. . . 73

8.1 Jumps out of a valley with energyEminafter timetw. . . 81

8.2 First scaling function for a parameter set with>. . . 86

8.3 First scaling function for a parameter set with<. . . 87

8.4 Scaling functionF2(2)for the same parameter set as in fig. 8.3. . . 88

8.5 C(tw+t;tw) and1-C(tw+t;tw)as a function of the scaling variable from Monte–Carlo simulations and the partial equilibrium formula. . . . 89

8.6 Ultra–slow aging behavior obtained from the PEC. . . 96

8.7 Illustration of the effect of a bimodal distribution ofvalues. . . 97

8.8 Aging functions when using a bimodal distribution ofvalues. . . 99

8.9 Three scaling regimes in the aging function when using a trimodal dis- tribution ofvalues. . . 100

8.10 Two scaling regimes in the aging function when using a bimodal distri- bution ofvalues. . . 101 E.1 Tracer diffusion coefficient for a constant DOS ind=2. . . XV E.2 Tracer diffusion coefficient for a Gamma DOS ind=2. . . XVIII E.3 Tracer diffusion coefficient for a bimodal DOS ind=1withq=0:7. . . . XXIV E.4 Tracer diffusion coefficient for a bimodal DOS ind=2withq=0:7. . . . XXV E.5 Tracer diffusion coefficient for a bimodal DOS ind=1withq=0:25. . . . XXVI E.6 Tracer diffusion coefficient for a bimodal DOS ind=2withq=0:25. . . . XXVII F.1 Number of distinct visited statesS(tw)for(;) = (1=4;3=8)and dif-

ferent dimensionsd. . . XXXII F.2 Number of distinct visited statesS(tw)for(;) = (1=6;1=4)and dif-

ferent dimensionsd. . . XXXII

(11)

Notation

h i thermal average

disorder average

weighting parameter for the final state energy in the transition rate

a lattice constant

inverse temperature1=T

0 unit of one over energy in Densities of State

b (d)

() frequency–dependent function in EMA self–consistency equation

(t

1

;t

2

) response function of an internal variablem(t1)to an external fieldH(t2)

C(t

1

;t

2

) correlation functionhm(t1)m(t2)iof some (generic) internal variablem(t)

Æ

n;m KRONECKER delta,1forn=m,0otherwise

Æ() DIRACdelta function, defined by

R

1

-1

dxÆ(x)f(x)=f(0)

d spatial dimension

D(t) time–dependent tracer diffusion coefficient 2d1t

h x(t)-x(0)

2

i

D

0 diffusion coefficient fort=0

D

1 diffusion coefficient fort!1

D(!) frequency–dependent tracer diffusion coefficient lim!0D (^ -i!)

^

D(z) LAPLACE transform z2

2d

L[h x(t)-x(0)

2

i](z)

"() dielectric relaxation (permittivity) function

^

f(z) LAPLACE transform of a functionf(t),L[f(t)](z)=

R

1

0 dte

-zt

f(t)

E

0 unit of energy in Densities of State (part I) or ground–state energy (part II)

E

n site energy on sitexn

F() scaling function

F

n;k

(t) effective lattice GREENmatrix of the EMA

2F1

() hypergeometric function

(z;E) Effective transition rate of the EMA

TR

(z;E) temperature rescaled effective transition rate of the EMA

() Gamma function

G

n;k

(t) lattice GREENmatrix

G (d)

00

() scaled lattice matrix element forddimensions

H(t) external field adjoint to the (generic) internal variablem(t)

(12)

H

n;k(t) symmetrized lattice GREENmatrix

j

(st) stationary probability current

fugacity

scaling variable

L number of states (“lattice sites”) in one dimension (linear dimension)

m(t) internal (generic) variable, adjoint to the external fieldH(t)

M(!) mobility

0 unit of hopping rate (thermally activated)

a attempt frequency

m;n,

(E

m

;E

n

) symmetrized transition ratewm;ne-Em, write(Em)if rate does not depend onEn

n occupation number on the lattice with components(n1;:::;nN)

N total number of lattice sites (part I) or total number of hops performed (part II)

! frequency variable

(t

1

;t

1

) relaxation function of an internal variablem(t1)when disturbed by an external fieldH(t1)

C(t

1

;t

2

) aging function

() waiting time distribution density

qc site percolation threshold

qEA EDWARDS–ANDERSON order parameter of some internal variablem, equals hmi2

p

(eq) equilibrium distribution on the lattice with components(p(1eq)

;:::;p

(eq)

N )

() Density of States

dc long–time conductivity

(!) complex conductivity, equal to0(!)+i00(!)

temperature divided by the glass transition temperatureT0

hopping time

x(tw) characteristic relaxation time aftertw

T hopping time (PARETOdistributed random variable)

t

w waiting time

T temperature in units of energy

T

0 glass transition temperature

T

g calorimetric glass transition temperature

w

m;n,

w(E

m

;E

n

) transition rate, writew(Em)if rate does not depend onEn

!x crossover frequency

x

n lattice vector of lattice siten

(13)

Notation

coordination number, equal to2dfor a hypercubic lattice

z LAPLACE variable

z

E energy–dependent LAPLACEvariableze-E

Z partition sum

(14)
(15)

1 Introduction

1.1 Glassy systems

Glassy materials are characterized by the absence of structural long–range transla- tional order (while on short ranges the structure may be similar to a crystal) and the fact that their relaxation time becomes exceedingly long and diverges for any practi- cal purpose when temperature is decreased below the glass transition temperature

T

g. Glassy materials are therefore never in equilibrium on laboratory or even on ge- ological time scales. For an overview see [Bin95, J¨ac86, Won76]. The divergent relax- ation time can be observed in a variety of different quantities as elastic viscosity, mechanical torque relaxation, conductivity spectra, dynamical light scattering or neu- tron scattering measurements. An especially well–known connection is given by the MAXWELL relation,'G1, where G1 is the long–time shear modulus and the shear relaxation time. Examples are shown in fig. 1.1. As one can see, relaxation times as a function of temperature change over many orders of magnitude. This behavior can be described by a VOGEL–TAMMANN–FULCHERlaw

=

0 exp

A T

0

T-T

0

; (1.1)

where VOGEL–TAMMANN–FULCHERtemperaturesT0, which are determined from fits of experimental data, have been found to be smaller thanTgand to be rather close to the KAUZMANN temperatureTK. (AtT=TKthe entropy of the glass would be equal to the entropy of the crystal, if the difference of glass and crystal in the specific heats atTgwould be extrapolated to temperaturesT<Tg(min), see e. g. [Ang97].)

Glassy materials include traditional (anorganic) network–glasses like SiO2or BeF2, organic glasses and polymers like polystyrene, metallic glasses, which are usually very fast cooled alloys like Sn–Cu, and systems where only a subsystem is in a glassy (dis- ordered) state like spin–glasses or orientational glasses. Strictly speaking only a mate- rial that has been prepared by cooling a melt below its glass transition temperature is called “glassy”, while other materials with intrinsic disorder are called “amorphous”.

But in the further text we will not make a strict distinction between these two classes of materials when talking about “glassy materials”.

(16)

Figure 1.1: (a)ViscosityversusTg=Tfor different glasses [Ang76]. Tghas been defined by

(T

g )=10

13Poise.(b)Correlation and relaxation times obtained from different measurements for the organic glass o–terphenyl [Cah91]. The symbols show: NMR spin–lattice correlation time (), spin–spin correlation time (), dielectric relaxation time (+), dynamic Kerr effect (), light scattering () and dielectric relaxation ().

(17)

1.2 Relaxation spectra of glasses

10

−6

10

−3

10

0

10

3

10

6

10

9

10

12

10

15

ν [Hz]

10

1

10

−1

10

2

10

−2

10

0

ε ’’

wing excess

β

fast

boson peak

α slow

T < T

1 2

β [a. u.]

Figure 1.2:Typical relaxation spectrum of a glass for two temperaturesT1<T2.

1.2 Relaxation spectra of glasses

A typical dielectric relaxation spectrum"00() of a glassy material in double logarith- mic representation is shown in fig. 1.2. The spectrum reveals a number of peaks, that may be characterized by their peak–position and their width. Starting from low fre- quencies and approaching higher frequencies the contributions are calledpeak, slow

peak (JOHARI–GOLDSTEIN relaxations), excess wing, fastpeak and Boson peak.

For even higher frequencies, infrared excitations show up. The fastrelaxation can be described in the framework of the mode–coupling theory [G ¨ot91] as “rattling” in a dynamical cage. There is much evidence that the Boson peak is connected to the excess contribution to the Density of States of the phonons in the glass compared to the crystal, and a couple of theoretical explanations have been given, but none is yet commonly accepted. The slowrelaxations are not observed in all glassy materials.

In some cases they can be attributed to some known dynamical process like vibrations or rotations of special functional groups in the glass. It is known experimentally that they show thermally activated behavior.

The most outstanding feature of relaxation spectra of glasses is the large peak that evolves for higher frequencies into the excess wing. A theoretical explanation of its detailed form for a particular material is most challenging and has not yet been achieved. However, on a descriptive level there are some well established facts. The

peak is not symmetric, but the increasing wing is always steeper than the decreas- ing wing. In the time domain this peak corresponds to a non–exponential relaxation.

(18)

There are different formulae in use for therelaxation, the most prominent being the KOHLRAUSCH–WILLIAMS–WATTSfunction (KWW)

(t) =

0 exp

"

-

t

KWW

KWW

#

(1.2) with a KWW exponentKWW<1. Furthermore it is known that thepeak of different materials for different temperatures can be scaled to a very good approximation on a mastercurve by the DIXON–NAGEL scaling1: When is the peak position of the peak,wthe full width at half maximum and"the difference of the permeability at very low and at very high frequencies the spectra can be superimposed on a master- curve with respect to therelaxation and the excess wing when plotting

w -1log

"

00

()

"

versus w-1(1+w-1)log

(DIXON–NAGELscaling):

1.3 Physical aging

Figure 1.3:Dependence of the loss–part of the dielectric permeability"00(!)of glycerol on the waiting timeta[Leh98].

Fluids and crystals reach a stationary state after leaving the transient regime. There- fore, correlations hm(t2)m(t1)iof some internal variable m(t) depend only on the time differencet2-t1 after the initial transient correlations have died out. The same

1When JOHARI–GOLDSTEINrelaxations are absent, both thepeak and the excess wing can be scaled by the DIXON–NAGELscaling.

(19)

1.3 Physical aging

Figure 1.4:Torque when applying a torsion to polycarbonate [OCo97]. All curves collapse to one master curve when the time is rescaled with a proper time scalethat depends on the waiting time.

holds true for response functions(t2;t1)that describe the response of a variable at timet2when at timet1a fieldH0Æ(t1)has been applied.

The situation changes in general when glassy systems are concerned. It has been observed experimentally that response functions like the elastic modulus or the dielec- tric susceptibility of glassy samples depend on thewaiting timetw that has passed since the sample has been quenched below the glass transition temperatureTg. Ex- amples of this dependence are shown in figs. 1.3 and 1.4. This phenomenon is called physical aging(oragingfor short) and the connected dynamics is referred to as “slow dynamics”.

The well known formulae for correlation and response functions usually depend on time–translational invariance and thus have to be generalized in order to describe aging in glassy materials. We will start with correlation functions of an internal vari- ablem(t),

C(t

w +t;t

w

) hm(t

w

+t)m(t

w

)i ; (1.3)

where denotes a disorder average. We assume that the sample cannot reach a stationary state for finite waiting timestw, in which case the correlation function can be written as the sum of a stationary partCst(t)and a partCag(t+tw;tw)that contains the physical age of the system,

C(t

w +t;t

w

) = Cst(t)+Cag(tw+t;tw): (1.4a) Typically the aging partCag(t+tw;tw)is governed for large waiting times by a scaling

(20)

functionFC t=x(tw)

with some characteristic time scalex(tw),

Cag(tw+t;tw) FC

t

x(tw)

for tw !1: (1.4b)

Fig. 1.4 shows an experimental example, where this scaling has been successfully ap- plied to a response function. Note that eq. (1.4b) is not the most general form: In prin- ciple it is possible that there is not only one characteristic time scale (and associated scaling regime), but that there are many characteristic time scalesx;1<x;2<:::<x;n

and that different scaling regimes show up when the correlation function is plotted versust=x;i for different x;i. This behavior is called multi–scaling. In fact, multi–

scaling recently has been conjectured on the basis of analytical results for the Langevin dynamics of mean-field spin glass models [Cug94, Bou98].

Often the time scalexis proportional to the waiting time itself andCag(t+tw;tw)

F

C (t=t

w

), which is callednormal aging. Another (observed) possibility issubaging;

in this casex increases slower thantw. In any casexwill increase withtw, leading to a slower and slower decrease of the correlations. Therefore, there is a subtle point when performing the limitt ! 1 in the correlation function. For any finite waiting time this limit will vanish, leading to

lim

tw!1

lim

t!1 C(t

w +t;t

w ) = F

C

(1) = 0: (1.5a)

However, if eqs. (1.4) holds and the limits are taken in reversed order, we will get lim

t!1

lim

t

w

!1 C(t

w +t;t

w ) = F

0

(0) qEA; (1.5b)

where the plateau value qEA is called EDWARDS–ANDERSON order parameter. The two limits eqs. (1.5) belong to different regimes of the two–time correlation function

C(t;t 0

): The stationary regime is reached, when tw ! 1 is taken for some finite value oft. In this limit the correlation function is described by the (time–translational invariant) partCst(t), while from the aging part only the offsetqEA is present. The limitt ! 1of the stationary regime is given by eq. (1.5b). In theaging regimethe limitst;tw ! 1are taken for some fixed value of the scaling variable t=x(tw). Here the correlation function is governed by the (non time–translational invariant) partCag(tw+t;tw), sinceCst(t)has died away. The limit!1of the aging regime leads to eq. (1.5a). When considering large, but finite values of tw and t, the two regimes can be distinguished most easily by the value of the correlation function: In the stationary regime the correlation function starts atCst(0)+qEAand then decreases toqEAwhile in the aging regime it starts atFC(0) =qEAand decreases toFC(1)= 0. Thus, in comprehensive form we may write

C(t

w +t;t

w

) Cst(t)+qEA & qEA (stationary regime); (1.6a)

C(t

w +t;t

w ) F

C

t

x(tw)

. q

EA (aging regime): (1.6b)

(21)

1.3 Physical aging

When formally setting

C(t

1

;t

2

) 0 for t2 < 0;

since at time t= 0 the system has been quenched below Tg, we can generalize the LAPLACE transform to the two–time correlation function (and generally to two–time functions) by

^

C(z;t

w ) =

Z

1

0 dt

0

e -zt

0

C(t

w

;t

w -t

0

); (1.7)

thus leading to a waiting time dependent LAPLACE transform (and thereby the spec- tral density) of the correlation function. When eqs. (1.4) hold, this can asymptotically be written as

^

C(z;t

w )

^

Cst(z)+

1

z G

C

zx(tw)

for tw !1; (1.8)

where the scaling functionGC(zx)is determined byFC(t=x).

Note that sometimes the ansatz x=x(tw+t)is used [Str78]. For normal aging this only affects the form of the aging function, since witht=tw

t

t+t

w

=

1+

;

while for the class of algebraic subaging behaviort=tw with0<<1asymptoti- cally fort;tw !1it holds

t

t+t

w

=

1+t -1+

w

;

leading to the same aging function for the two ans¨atze. Note that for practical pur- poses, where only finite waiting times can be reached, the two ans¨atze nevertheless can give different results.

When disturbing the system by some external fieldH(t)and measuring the adjoint variablehm(t)i, one gains information about theresponse function(t;t0)defined by

hm(t)i = Z

t

-1 dt

0

(t;t 0

)H(t 0

): (1.9)

WhenH(t0)=H0is chosen for0<t0<twand zero otherwise, the relaxation for times

t

w +t>t

wcan be measured,

h m(t

w

+t)i = H

0 Z

t

w

0 dt

0

(t

w +t;t

0

) (t

w +t;t

w )H

0

; (1.10)

where(t;t0)is calledrelaxation function. When applying a harmonic field instead of a constant field, the frequency–dependent susceptibility(!;tw)(or the dielectric function "(!;tw)) is measured instead. To this end the harmonic field is switched on at timet=0when the sample reaches the glassy phase; after a waiting timetw the

(22)

adjoint variablehm(tw+t)iis measured over a few periods and from this measurement the amplitudem0(!;tw)is determined and divided by the amplitude of the external field H0. From this rule of measurement it is clear that the considered frequencies need to have periods2=! that are much shorter thantw. Formally the frequency–

dependent susceptibility can be obtained as a limit of the LAPLACE transform of the time–dependent susceptibility,

(!;t

w

) = lim

!0 +

^

(-i!;t

w

) = lim

!0 +

Z

1

0 dt

0

e (i!-)t

0

(t

w

;t

w -t

0

); (1.11) where, as in the case of the two–time correlation function,(t1;t2)0has been set for

t

2

<0. For both, the susceptibility(!;tw)and the relaxation function(tw+t;tw), an ansatz of the form eq. (1.4) can be made in the limit of large waiting times,

(!;t

w

) st(!)+

1

! G

!x(tw)

; (1.12)

(t

w +t;t

w

) st(t)+F

t

x(tw)

: (1.13)

C(t

w +t;t

w ) (tw

+t;tw +t)-(tw +t;tw

)

Figure 1.5:Parametric plot (with respect tot) of the relaxation versus the correlation function for a sequence of increasing waiting times for a mean–field spin glass model [Bou98]. The dotted line is a limiting curvetw!1.

For a system in the stationary regime the fluctuation–dissipation theorem (FDT)

st(0)-st(t) =

1

T

Cst(0)-Cst(t)

(1.14)

(23)

1.4 Hopping models of glassy systems or in LAPLACEspace

^

st(z) =

1

T

Cst(t=0)+qEA-zC^st(z)

(1.15) holds. In the aging regime, however, this is no longer true. When considering a para- metric plot (with respect to t) of (t;tw) (tw+t;tw+t)-(tw+t;tw) versus

C(t

w +t;t

w

) for some fixed (large but finite) tw, the slope should be constant and equal to-1=Tdue to the FDT. When doing this in aging systems, it turns out that this is the case forC(tw+t;tw) & qEA, i. e. the stationary regime. In the aging regime, i. e. for C(tw+t;tw) . qEA, where the parameter of the plot can be expressed by the scaling variable and thus()=st+F(0)-F()is plotted versusFC(), the slope changes. This can be interpreted as some “effective dynamic temperature”

-1=Teff(tw+t;tw), whereTeffquite generally is higher thanT. An example for a mean–

field spin glass model from [Bou98] is shown in fig. 1.5.

1.4 Hopping models of glassy systems

Transport properties of glassy systems are often associated with a hopping motion in a random energy landscape. For example, for describing the dynamics of simple struc- tural glass formers in the highly supercooled melt beyond the mode–coupling regime, hopping processes are expected to become dominant [Das86, G ¨ot91]. A well estab- lished concept is to view the dynamics of the glass forming liquid as the motion of a point in a high–dimensional configuration space with many potential energy minima of varying depth. The first one to propose this idea was GOLDSTEIN[Gol69], who de- veloped the picture of a potential energy landscapeU(whereU(r1;:::;rN)is plotted versus the3Nconfiguration variables) as a rugged energy landscape with many local minima. He justified the importance of the local minima by stating that:

Glasses at low temperatatures are mechanically stable against small deviations from their equilibrium positions and thus have to be near minima of the poten- tial energy.

Experimentally a finite configuration entropy is found for glasses. Furthermore, the state of a glass depends strongly on the thermal treatment it has received in the transition regime to the glassy phase. GOLDSTEIN concludes that in the part of the potential energy landscape that represents the glassy phase, a large number of minima of different depth are accessible to the glassy system.

At the lowest temperatures transitions from one valley of the potential energy to another one have to be thermally activated (when the possibility of quantum–

mechanical tunneling is neglected).

GOLDSTEINassumes that, although the potential energy is no local quantity, the tran- sitions from one minimum to a near–by minimum in the potential energy landscape

(24)

are “local” in the sense, that only particles in a small region of the sample are consider- ably displaced. He uses a consideration originally put forward by ZWANZIG[Zwa67]

regarding the vibrational spectrum to estimate the range of validity of the hopping approximation in glasses. He argues that only those vibrations can be excited in a visco–elastic sample whose period is small compared to the shear relaxation time.

Consequently the picture of a state point performing vibrations in a deep energy min- imum should be justified if the main contributions to thermodynamic properties stem from frequencies obeying this condition. When assuming a DEBYEform for the spec- trum with cut–off frequencymax'1013Hz and taking into account that frequencies smaller thanmax=10contribute only a few tenths of a percent to the entropy, he con- cludes that (as a very rough estimate) for shear relaxation times10=max'10-12s the picture of vibrations in a deep minimum (and consequently the picture of a hop- ping motion between deep minima) makes sense. When comparing this to typical relaxation times'103s atTg, GOLDSTEIN concludes that the hopping approxima- tion should remain valid even for a supercooled liquids above the calorimetric glass transition temperature. We like to note that the potential energy landscape itselfdoes notdepend on temperature, but that only the kinetic energy of the system is tempera- ture dependent.

The hopping picture of glasses has been applied and extended by several au- thors (see e. g. [B¨as87, Ang95, Sti95]). Considerable work has been done to clarify the structural properties of the potential energy landscape. STILLINGER and WEBER have shown [Sti82] that rather general arguments suggest an increase of the number of minima(N)with the number of particleNas

(N) N!e N

(1.16) for a single–component system in the limitN!1. Attempts have been made only recently to calculate the distribution functions of these energy minima (and other sta- tistical features) both from molecular dynamics simulations [Heu97, Don00, Ang00, Bro00] and theory [Sch98, Mez99] and to model relaxation dynamics in terms of such distributions [Die99].

For spin–glasses, hopping processes in ruggedfree energy landscapes have been proposed [Bou92, Sib97a, Sib98]. The free energy functionals are defined in a space ofpure states where cumulants vanish in the limit of large distances (F[m1;:::;mN] with mean magnetizationsmiin the case of spin–glasses) instead of the configuration space of the particles. In these models the separation of the time scales between the time spent in a trap and the time spent to surmount a barrier can have its origin in en- ergy barriers or in kinematic constraints created by the fact that parts of the state space are totally inaccessible (“entropic barriers”). Since free energy functionals depend on temperature, in these models the form of the rugged energy landscape itself is tem- perature dependent (contrary to the case of the potential energy landscape above).

Within such models, SIBANIand HOFFMANNhave proposed [Sib97a] that aging may be attributed to partial thermal equilibrium within one “valley” of a free energy land- scape. Here theratioof the occurrence of two different states in a valley (or subspace) is assumed to have approximately the same value as in equilibrium, while theabsolute valuewill in general be very different.

Referenzen

ÄHNLICHE DOKUMENTE

The results are formulated and proved using the recent and general theory of time scales via the backward nabla differential operator. Key words: Calculus of Variations;

The results are formulated and proved using the recent and general theory of time scales via the backward nabla differential operator.. Key words: Calculus of Variations;

Theo- rem 1 implies upper semicontinuity of 2, also in this topology, but under t h e condition of strong asymptotic stability of the equilibrium set KO with

And because the equation can be established on a partial set of data, i t can be used, e,g, to predict how many books a famous writer suiLl produce, and

• Receptions are behavioral features and thus are realized by behavior (e.g., a state

Even in the limiting case of perfectly stiff polymers, the appearance of long-range order no longer coincides with the gelation transition as it did for hard cross-linkers, but

Clouds are moving on a short term timescale and have a high influence on the available solar radiation, as they absorb, reflect and scatter parts of the incoming light.2 However,

For example, in the optimal control theory for linear lumped parameter systems with quadratic per- formance indices, the relationships between optimal continuous- time