Continuously Broken Ergodicity

Embed Size (px)

Citation preview

  • 7/29/2019 Continuously Broken Ergodicity

    1/11

    Continuously broken ergodicity

    John C. MauroScience and Technology Division, Corning Incorporated, SP-TD-01-1, Corning, New York 14831

    Prabhat K. GuptaDepartment of Materials Science and Engineering, Ohio State University, Columbus, Ohio 43210

    Roger J. LoucksScience and Technology Division, Corning Incorporated, SP-TD-01-1, Corning, New York 14831 and

    Department of Physics and Astronomy, Alfred University, One Saxon Drive, Alfred, New York 14802

    Received 6 December 2006; accepted 28 March 2007; published online 14 May 2007

    A system that is initially ergodic can become nonergodic, i.e., display broken ergodicity, if the

    relaxation time scale of the system becomes longer than the observation time over which properties

    are measured. The phenomenon of broken ergodicity is of vital importance to the study of many

    condensed matter systems. While previous modeling efforts have focused on systems with a sudden,

    discontinuous loss of ergodicity, they cannot be applied to study a gradual transition between

    ergodic and nonergodic behavior. This transition range, where the observation time scale is

    comparable to that of the structural relaxation process, is especially pertinent for the study of glass

    transition range behavior, as ergodicity breaking is an inherently continuous process for normal

    laboratory glass formation. In this paper, we present a general statistical mechanical framework for

    modeling systems with continuously broken ergodicity. Our approach enables the directcomputation of entropy loss upon ergodicity breaking, accounting for actual transition rates between

    microstates and observation over a specified time interval. In contrast to previous modeling efforts

    for discontinuously broken ergodicity, we make no assumptions about phase space partitioning or

    confinement. We present a hierarchical master equation technique for implementing our approach

    and apply it to two simple one-dimensional landscapes. Finally, we demonstrate the compliance of

    our approach with the second and third laws of thermodynamics. 2007 American Institute of

    Physics. DOI: 10.1063/1.2731774

    I. INTRODUCTION

    One of the most prevalent, and often unstated, assump-

    tions in statistical mechanics is that of ergodicity, which as-serts the equivalence of time and ensemble averages of the

    thermodynamic properties of a system. Ergodicity implies

    that, given enough time, a system will explore all allowable

    points of phase space. The term broken ergodicity, intro-

    duced by Bantilan and Palmer1

    in 1981, denotes the loss of

    ergodicity that is common in many systems such as spin and

    structural glasses.

    The question of ergodicity, and that of thermodynamic

    equilibrium, is really a question of time scale. We may think

    of an experiment as having two relevant time scales: an in-

    ternal scale int on which the dynamics of the system occur,

    and an external scale ext on which properties are mea-sured. The internal time scale is essentially a relaxation time

    over which a system loses memory of its preceding states,

    whereas the external time scale defines a measurement win-

    dow over which the system is observed. The measurement

    can be made either directly by a human observer or by using

    an instrument accessing times not available to an unaided

    human. A system is in thermal equilibrium if all of the rel-

    evant relaxation processes have taken place, while the re-

    maining slower processes are essentially frozen on the ex-

    ternal time scale.2

    The ratio of internal to external time scales defines the

    Deborah number3

    of an experiment,

    D = intext

    , 1

    named in honor of the prophetess Deborah, who in the Old

    Testament sings, the mountains flowed before the

    Lord Judges 5:5 . The implication is that mountains,while basically static on a human time scale, do in fact flow

    on a geologic time scale inaccessible to mere mortals. For a

    human observer, the phenomenon of continental drift is de-

    scribed by a large Deborah number D1, intext , wherethe system visits only a small subset of the available points

    in phase space during the external i.e., observation time.Such scant sampling of phase space is insufficient for deter-

    mining a long-time average of properties, and hence the sys-tem lacks ergodicity. On the other hand, many phenomena

    such as relaxations in gases and liquids occur on a time scale

    too fast for direct human observation. These phenomena are

    described by a very small Deborah number D1, intext , which is indicative of an ergodic system. Since the

    system can explore a greater portion of phase space during

    the observation time, the measured property values over extare effectively equal to those in the long-time limit.

    The observation of ergodic behavior thus depends on

    both the internal relaxation time of a system and the external

    THE JOURNAL OF CHEMICAL PHYSICS 126, 184511 2007

    0021-9606/2007/126 18 /184511/11/$23.00 2007 American Institute of Physics126, 184511-1

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

    http://dx.doi.org/10.1063/1.2731774http://dx.doi.org/10.1063/1.2731774http://dx.doi.org/10.1063/1.2731774http://dx.doi.org/10.1063/1.2731774
  • 7/29/2019 Continuously Broken Ergodicity

    2/11

    observation time: a system is ergodic when D1 and non-

    ergodic when D1. The issue of ergodicity, while inherently

    relative, is nevertheless of great physical importance: the

    properties we experience are those we measure on our own

    finite time scale, not in the long-time limit of ergodicity. It is

    thus important for statistical mechanical models to account

    for broken ergodicity in order to predict the observed prop-

    erties of nonergodic systems.

    But what about the process by which ergodicity is lost?The breakdown of ergodicity can occur either discontinu-

    ously or continuously. A discontinuous breakdown of ergod-

    icity is caused by a sudden decomposition of the phase space

    into a set of mutually inaccessible components. An example

    of discontinuous ergodicity breaking is glass formation by

    sudden quenching from a melt; here, the system experiences

    an abrupt loss of thermal energy, thereby trapping it in a

    subset of the overall phase space. On the other hand, a glass

    can be formed by gradual cooling from a melt. In this case,

    the glass-forming system experiences a continuous break-

    down of ergodicity as it gradually becomes confined to a

    subset of the phase space. Our definition of continuously

    broken ergodicity is distinct from what some researchers callweakly broken ergodicity,

    46in that we assume that the re-

    laxation time is finite but exceeds the observation time.

    In a landmark 1982 paper, Palmer7

    thoroughly develops

    the concept of broken ergodicity and proposes a statistical

    mechanical framework with which to treat nonergodic sys-

    tems and the discontinuous breakdown of ergodicity. In the

    Palmer approach, the phase space of a nonergodic system

    is divided into a set of disjoint regions or components ,

    where

    =. 2

    The components are assumed to meet the conditions of con-

    finement and internal ergodicity. Confinement indicates that

    transitions are not allowed between components; in other

    words, if a system has its phase point in a particular compo-

    nent at a time t, there is negligible probability of the

    system leaving that component during a specified observa-

    tion time tobs, i.e., in the time interval t, t+ tobs . The condi-tion of internal ergodicity states that ergodicity holds within

    a given component.

    The Palmer approach thus assumes a clear separation of

    intra- and intercomponent relaxation time scales: intracom-

    ponent relaxation i.e., among the various microstates within

    a component occurs on a time scale component much shorterthan the observation time D 1 , while intercomponent

    transitions occur on a time scale system much longer than tobs D 1 . The assumption that

    component tobs system 3

    allows components to be in internal equilibrium while the

    system as a whole is not. As a result, ergodic statistical me-

    chanics can be applied within each of the individual compo-

    nents, and the overall properties of the nonergodic system

    can be computed using a suitable average over the individual

    components.

    This has important implications when computing the en-

    tropy of a nonergodic system. Before proceeding, it is useful

    to review some common definitions of entropy, following the

    terminology of Goldstein and Lebowitz:8

    1 The thermodynamic entropy of Clausius, given by

    dS=dE

    T, 4

    where E is internal energy and T is absolute tempera-

    ture. This definition of entropy is applicable only for

    equilibrium systems and reversible processes.

    2 The Boltzmann entropy,

    SB = kB ln, 5

    applicable to the microcanonical ensemble, where kB is

    Boltzmanns constant and is the volume of phase

    space i.e., the number of microstates visited by a sys-tem to yield a given macrostate.

    3 The Gibbs or statistical entropy,

    S

    = kBi

    pi ln pi , 6

    applicable to the canonical ensemble, where pi gives

    the probability of occupying microstate i, and the sum-

    mation is over all microstates. Here, the probabilities

    are computed based on an ensemble average of all pos-

    sible realizations of a given system. With the assump-

    tion of ergodicity, the ensemble-averaged values of piare equal to the time-averaged values of pi.

    While there is no disagreement about the definition of

    entropy for equilibrium systems, the definition of entropy for

    nonequilibrium systems is not established except for the caseof microcanonical isolated systems where there is a general

    agreement that Boltzmanns definition is valid and consistent

    with the second law.812

    For canonical, nonergodic systems, the proper definition

    of entropy is a highly contentious topic. For glassy systems,

    fortunately, there is a way out. As discussed by Palmer,7

    glass can be described as an ensemble of ergodic compo-

    nents of the phase space. With the conditions of confinement

    and internal ergodicity, we can apply the Gibbs definition of

    entropy within each individual component. The resulting en-

    tropy of the system, as derived by Palmer7

    and provided as

    Eq. 12 in Sec. II of this paper, is simply a weighted sum of

    the Gibbs entropies of the individual components. In thismanner, the entropy of Palmer can account for the broken

    ergodicity of the glassy state.

    One particular point of confusion in the glass science

    community relates to the application of Eq. 4 for thermo-dynamic entropy in conjunction with differential scanning

    calorimetry DSC experiments to compute the entropy of asystem on both cooling and heating through the glass

    transition.13

    This results in two main findings: a there is nochange in entropy at the laboratory glass transition, and b

    glass has a positive residual entropy at absolute zero. We

    argue that these results are incorrect since the glass transition

    is not a reversible process14

    and hence Eq. 4 cannot be

    184511-2 Mauro, Gupta, and Loucks J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    3/11

    applied. Moreover, the ergodic formula for Gibbs entropy,

    which approximately equals the thermodynamic entropy,15,16

    cannot be used since it cannot account for the defining fea-

    ture of the glass transition, viz., the breakdown of ergodicity.

    The glass transition, by definition, occurs when tobs =system i.e., a Deborah number of unity .13 The only reason we ob-

    serve the glassy state at all is that the relaxation time scale of

    the system becomes much longer than the observation time

    scale. Hence, the concept of an observation time is central tothe very definition of glass, and the glass transition can only

    occur if there is a finite observation time. We will show in

    this paper that when a finite observation time is considered,

    the glass transition entails a loss of entropy rather than a

    freezing in of entropy.

    Moreover, we argue that the current belief that glass has

    a finite residual entropy at absolute zero is in direct violation

    of Plancks statement of the third law of thermodynamics,

    which demands that the entropy of any classical system

    must vanish at zero temperature.17

    Some researchers claim

    that glass is somehow exempt from obeying the third law of

    thermodynamics since it is a nonequilibrium material.13

    However, their result from DSC measurements of heat ca-

    pacity that glass has a finite residual entropy at absolute

    zero is based on application of equilibrium, reversible ther-

    modynamics to the nonequilibrium glassy system. Hence,

    their very calculation of entropy is at odds with their argu-

    ment that the third law does not apply to glass. We argue that

    glass is not exempt from the third law and show later in this

    paper that application of Palmers definition of entropy al-

    ways leads to zero entropy at zero temperature since there is

    no thermal energy to allow for transitions between mi-

    crostates .

    While the Palmer methodology greatly elucidates the

    impact of broken ergodicity on entropy and other properties,many realistic systems do not lend themselves to neat parti-

    tioning into discrete components that simultaneously satisfy

    the conditions of confinement and internal ergodicity. Also,

    the partitioning itself must depend on the transition rates

    between microstates and therefore on the temperature of the

    system. For a system where temperature evolves with time,

    the partitioning would have to be recomputed at each new

    temperature step.

    In addition to these computational difficulties, there is

    the physical problem that transitions between components

    are never strictly prohibited at any finite temperature; this is

    especially important for Deborah numbers near unity D 1 , where the relaxation is too fast to assume confinement

    and yet too slow to assume internal ergodicity. Consequently,

    the Palmer approach can consider only the case of discon-

    tinuously broken ergodicity, where there are no relaxation

    processes with D1; here, the loss of ergodicity can be

    modeled well by a discontinuous partitioning of phase space

    into multiple components. However, the Palmer approach

    needs modification to model a continuous breaking of ergod-

    icity, such as in normal laboratory glass formation using a

    finite cooling rate. One can argue that the most interesting

    physics occurs during the transition between ergodic and

    nonergodic states, where D1. This regime corresponds di-

    rectly to the glass transition range and is of great scientific

    and technical importance.13

    In this paper, we present a generalization of the Palmer

    approach that accounts for continuously broken ergodicity

    and is thus suitable for a realistic modeling of glass transition

    range behavior and other phenomena with D1. Our meth-odology is based on a hierarchical master equation approach

    that avoids any explicit definition of components and makes

    no assumptions of confinement or internal ergodicity. Ourtechnique can thus be applied to any arbitrary system and for

    any temperature path. We present equations for the configu-

    rational entropy and free energy of a system with continu-

    ously broken ergodicity and show results for two simple one-

    dimensional landscapes.

    II. ENERGY LANDSCAPES AND DISCONTINUOUSLYBROKEN ERGODICITY

    The Palmer approach for discontinuously broken ergod-

    icity involves a partitioning of the phase space into a set of

    components , where each component satisfies the condi-tions of confinement and internal ergodicity.

    7It is useful for

    our ensuing discussion to translate this terminology into the

    energy landscape formalism,18,19

    which offers a powerful ap-

    proach for studying the thermodynamics and kinetics of mo-

    lecular clusters,2024

    biomolecules,23

    supercooled

    liquids,2530

    and structural glasses.23,3134

    The potential energy landscape of an N-particle system

    is a 3N-dimensional hypersurface in the particle configura-

    tion space,

    U= U r1,r2, . . . ,rN , 7

    where r = r1 ,r2 , . . . ,rN R3N are the position vectors of the

    N particles. The underlying assumption is a decoupling of

    the configurational and vibrational thermal components ofenergy. While the potential energy landscape itself is inde-pendent of temperature, the way in which a system explores

    the landscape depends on the available thermal energy. At

    high temperatures there is sufficient thermal energy to enable

    free exploration of the landscape; for a condensed system,

    this corresponds to the case of an ergodic liquid with high

    fluidity. As the system is cooled, the transitions slow down

    and become thermally activated. This leads to continuously

    broken ergodicity as the system gradually becomes trapped

    in a subset of the landscape.

    The U hypersurface itself contains a multitude of local

    minima, each corresponding to a mechanically stable con-

    figuration of the system, termed an inherent structure. Theset of 3N-dimensional hyperspace configurations that drains

    to a particular minimum via steepest descent is known as a

    basin.29

    The study of energy landscapes is facilitated by

    mapping the continuous U hypersurface to a discrete set of

    minima, i.e., basin volumes are mapped to their correspond-

    ing inherent structures. The number of inherent structures

    scales at least exponentially with N.27

    In the energy landscape formalism an individual mi-

    crostate corresponds to a single inherent structure or basin. A

    component in the Palmer approach corresponds to a so-

    called metabasin3539

    in the energy landscape, i.e., a group

    of basins that are mutually accessible at a given temperature

    184511-3 Continuously broken ergodicity J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    4/11

    and for a given observation time. The metabasins themselves

    are separated from each other by large potential energy bar-

    riers such that intermetabasin transitions are highly unlikely.

    Hence, metabasins satisfy the conditions of confinement and

    internal ergodicity.

    The probability of a system being confined in a particu-

    lar metabasin upon cooling is equal to a restricted summa-

    tion of the occupation probabilities of the individual basins

    within ,

    P=i

    pi. 8

    As each basin falls within exactly one metabasin, the sum of

    metabasin probabilities is unity,

    P =

    i

    pi = 1 . 9

    Since ergodicity is maintained within an individual metaba-

    sin, we can define a corresponding intrametabasin Gibbs en-

    tropy as

    S= kBi

    pi

    Pln

    pi

    P, 10

    where from Eq. 8 we know that

    i

    pi

    P= 1 . 11

    Thus, in the Palmer approach see Sec. 5.4 in Ref. 7 , theexpectation value for configurational entropy is a weighted

    average of the Gibbs entropies of the individual metabasins,

    S =

    SP= kB

    ipi ln

    pi

    P. 12

    The corresponding free energy can be computed by

    F =i

    Uipi T S = U T S , 13

    where Ui is the potential energy of inherent structure i. Note

    that this is a Helmholtz free energy, and the Gibbs free en-

    ergy can be computed by

    G =i

    Hipi T S = H T S , 14

    where

    Hi = Ui + PVi 15

    is the zero-temperature enthalpy of inherent structure i,

    which has a system volume of Vi. The enthalpy landscape

    approach is an extension of the potential energy landscape

    formalism for isobaric systems,40

    and is useful for computing

    the volume evolution of systems under constant pressure

    conditions.34

    Equation 12 gives the entropy of a system with discon-

    tinuously broken ergodicity. If broken ergodicity were not

    accounted for, the entropy would be given by direct applica-

    tion of the statistical formula of Eq. 6 . The difference be-

    tween the statistical entropy and the nonergodic glassy en-

    tropy ofEq. 12 is called the complexity of the metabasin

    ensemble7

    and is given by

    I= S S = kB

    P ln P. 16

    The complexity is a measure of the nonergodicity of a sys-

    tem. For a completely nonergodic system trapped in a single

    microstate, each basin itself is a metabasin, and the complex-ity adopts its maximum value of I= S. For an ergodic system,

    all basins are part of the same metabasin, and the complexity

    is zero. The process of ergodicity breaking necessarily results

    in a loss of entropy and increase in complexity as the phasespace divides into mutually inaccessible metabasins. This

    loss of entropy is accompanied by an increase in the free

    energy of the system, as indicated by Eqs. 13 and 14 .

    While the breaking of ergodicity always results in a loss of

    entropy S S and a gain in free energy F F, G

    G , the energy and enthalpy are unaffected U = U and

    H =H .

    While the Palmer approach can be used to compute thesudden entropy loss in systems with discontinuously broken

    ergodicity e.g., glass formation via instantaneous quench-ing , it cannot be used to compute the gradual loss of entropy

    that occurs in systems with continuously broken ergodicity

    e.g., laboratory glass formation using a finite cooling rate .

    III. CONTINUOUSLY BROKEN ERGODICITY

    The limitations of the Palmer approach7

    can be over-

    come by relaxing the assumptions of metabasin confinement

    and internal ergodicity. We consider a nonequilibrium system

    with probability distribution pi t at time t. Suppose that we

    make an instantaneous measurement of the microstate of asystem at time t. The act of measurement causes the system

    to collapse into a single microstate i with probability pi t .

    In the limit of zero observation time, the system is confined

    to one and only one microstate and the observed entropy is

    necessarily zero. However, the entropy becomes positive for

    any finite observation time tobs since transitions between mi-

    crostates are not strictly forbidden except at absolute zero,barring quantum tunneling . What, then, is the entropy of the

    system over the observation window t , t+ tobs ?We can answer this question by following the dynamics

    of a system whose microstate is known at t, the beginning

    of the observation window. Let fi,j t be defined as the con-

    ditional probability of the system occupying microstate j af-ter starting in a known microstate i and subsequently evolv-

    ing for some time t, accounting for the actual transition rates

    between microstates. The conditional probabilities satisfy

    j

    fi,j t = 1 , 17

    for any initial state i and for all time t. Hence, fi,j tobs gives

    the probability of transitioning to microstate j after an initial

    measurement in state i and evolving through an observation

    time tobs.

    While there is no clear agreement on the definition of

    nonequilibrium entropy for a canonical system, a couple of

    184511-4 Mauro, Gupta, and Loucks J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    5/11

    different expressions have been suggested. One is based on a

    relative or conditional entropy and another is time-

    dependent Gibbs entropy.41

    We choose the latter, notwith-

    standing it is not valid for microcanonical systems, for the

    following reasons:

    1 The time-dependent Gibbs entropy is zero in the limitsof tobs0 and T0;

    2 it yields the equilibrium entropy in the long-time limit;

    3 it reproduces the Palmer results for discontinuouslybroken ergodicity; and

    4 several other authors have also considered the time-dependent Gibbs entropy.

    4245

    In our case, the nonequilibrium Gibbs entropy is

    Si tobs = kBj

    fi,j tobs ln fi,j tobs . 18

    This represents the entropy of one possible realization of the

    system and corresponds roughly to Palmers component en-

    tropy of Eq. 10 ; however, we have not made any assump-

    tions about confinement or internal ergodicity. Note thatwhile the above equation is of the same form as the Gibbs

    entropy of Eq. 6 , the value of fi,j tobs above gives theprobability of the system transitioning from an initial state i

    to a final state j within a given observation time tobs. The

    probability pi in the Gibbs formulation represents the ergodic

    limit of tobs; hence, pi represents an ensemble-averaged

    probability of the system occupying a given state and does

    not account for the finite transition time required to visit a

    state, which may be long compared to the observation time

    scale.

    The expectation value of entropy at the end of the ob-

    servation window t , t+ tobs is the weighted sum of entropy

    values for all possible realizations of the system,

    S t,tobs =i

    Si ttobs pi t . 19

    With this approach, there is no need to define components or

    metabasins. By considering all possible configurations of the

    system and the actual transition rates between microstates,

    our approach can be applied to an arbitrary energy landscape

    and for any temperature path. Our approach is thus suitable

    for modeling systems in all regimes of ergodicity: fully er-

    godic, fully confined, and everything in between. We further

    note that the combination of Eq. 17 with Eq. 19 reduces

    to the Palmer equation of entropy, Eq. 12 , for the case ofdiscontinuously broken ergodicity.

    With Eq. 19 , the entropy of the system is zero both inthe limits of tobs0 and T0. The first case tobs0 is in

    agreement with Boltzmanns notion of entropy as the number

    of microstates visited by a system to yield a given

    macrostate:812

    with zero observation time, the system re-

    mains in the initial microstate. The second case T0 is in

    agreement with Plancks statement of the third law,17

    which

    states that the entropy of any classical condensed system

    must vanish at absolute zero. Both limits are equivalent in

    the Palmer model of discontinuously broken ergodicity,7

    where each microstate would have its own separate compo-

    nent metabasin ; no transitions are allowed among the com-

    ponents, so the entropy is necessarily zero.

    For any positive temperature, the limit of tobs yields

    an equilibrated system with complete restoration of ergodic-

    ity. In the Palmer view,7

    this is equivalent to all microstates

    being members of the same component with transitions

    freely allowed among all microstates. Both the Palmer ap-

    proach and ours yield the same result as the Gibbs formula-

    tion of entropy in the limit of t

    , i.e., for a fully ergodic,equilibrated system.

    In the next section, we describe a hierarchical master

    equation approach for implementing the above formalism.

    IV. HIERARCHICAL MASTER EQUATIONAPPROACH

    The master equation approach is a useful technique for

    modeling the dynamics of nonequilibrium statistical me-

    chanical systems.46,47

    The approach involves constructing a

    set of coupled rate equations, with one equation for each

    available microstate in the system. For a system with mi-crostates, labeled i ,j 1 , 2 , . . . , , the set of master equa-

    tions is given by

    dpi

    dt=

    ji

    Wjipj Wijpi , 20

    where pi denotes the probability of occupying state i, Wji is

    the transition rate from state j to state i, and Wij is the rate

    of reverse transition. The occupation probabilities are subject

    to the constraint

    i

    pi = 1 21

    for all times t. Assuming fixed rate parameters Wij, the

    master equation dynamics of Eq. 20 always follow the re-laxation of a system toward an equilibrium state e.g., during

    isothermal relaxation . For an isolated, ergodic system the

    Gibbs entropy computed by Eq. 6 increases with time,48

    as

    required by the second law for a spontaneous, irreversible

    process such as relaxation.

    A recent application of the master equation approach by

    Mauro and Varshneya33

    considers the problem of glass tran-

    sition in an energy landscape. Rather than starting in a non-equilibrium state and following the dynamics of a system as

    it relaxes toward equilibrium, Mauro and Varshneya consider

    a liquid system initially at equilibrium. Departure into the

    nonequilibrium glassy regime is computed by solving a set

    of master equations, where the transition rates are functions

    of an arbitrary cooling path, Wij T t . As the system is

    cooled, the relaxation time becomes longer than the observa-

    tion i.e., simulation time scale. At low temperatures theoccupation probabilities pi are effectively frozen as the sys-

    tem becomes trapped in local regions of the energy land-

    scape. The macroscopic properties of the system at any point

    in time can be computed with the weighted average,

    184511-5 Continuously broken ergodicity J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    6/11

    A T t =i

    Aipi T t , 22

    where Ai denotes the specific property value associated with

    configuration i. For example, Eq. 22 can be used to com-

    pute volume-temperature diagrams of glass-forming systemsfor different cooling paths.

    34

    However, Eq. 6 cannot be used in combination with the

    Mauro-Varshneya approach to compute the entropy of a

    glassy system, as it does not account for the broken ergodic-

    ity inherent in the glassy state.4956

    The definition of glass

    itself has within it the assumption of an observation time that

    is shorter than the relaxation time scale of the system;13

    hence, glass is inherently a nonergodic system trapped in a

    subset of the overall phase space available during the finite

    observation time. Since Boltzmanns definition of entropy

    considers only the microstates visited by a system during the

    time of observation,812

    application of Eq.

    6

    will lead to an

    artificially high value of entropy. Likewise, Eq. 12 cannotbe used since it assumes discontinuously broken ergodicity,

    and the glass transition under a finite cooling rate involves a

    continuous loss of ergodicity.

    Therefore, we propose the following hierarchical master

    equation algorithm to compute the dynamics of the system

    accounting for continuously-broken ergodicity.

    1 Choose an appropriate initial state for the system. In the

    Mauro-Varshneya approach,33

    the initial configuration

    is chosen to be an equilibrium liquid at the melting

    temperature Tm ,

    pi 0 =1Q

    exp UikBTm , 23where Q is the partition function,

    Q =i

    exp Ui

    kBTm. 24

    By starting the system in equilibrium at Tm, we are able

    to account for all thermal history effects on the final

    nonequilibrium state.

    2 Compute the master equation dynamics according tothe Mauro-Varshneya approach,

    33

    dp i t

    dt=

    ji

    Wji T t pj t ji

    Wij T t pi t . 25

    This gives the ensemble-averaged probability of occu-

    pying each of the various basins in the energy land-

    scape at any point in time.

    3 For any time t when we wish to compute entropy orfree energy, select a particular basin i. Construct a new

    set of master equations for the conditional probabilities

    fi,j,

    dfi,j t

    dt = kjWkj T t + t fi,k t

    kj

    Wjk T t + t fi,j t , 26

    using an initial condition of fi,i 0 =1 for the chosen

    basin and fi,ji 0 =0 for all other basins. Compute the

    dynamics of the system for exactly the observation

    time, tobs. The output from this step is fi,j tobs , the con-

    ditional probability of the system reaching any basin j

    after starting in basin i and evolving for exactly the

    observation time. Note that the time scale of fi,j t is

    shifted by t relative to pi t .

    FIG. 1. One-dimensional fragile landscape with nine basins.

    FIG. 2. Evolution of glassy and equilibrium potential energies with respect

    to a time and b temperature for the one-dimensional fragile landscape ofFig. 1. We consider linear cooling from 500 to 300 K with a total cooling

    time of 1.0 s.

    184511-6 Mauro, Gupta, and Loucks J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    7/11

    4 Compute the entropy of the simulation in step 3 usingEq. 18 .

    5 Repeat steps 3 and 4, starting from each of the indi-

    vidual basins, i 1 , 2 , . . . , .

    6 Once all values of Si are computed, the overall entropyof the system is given by Eq. 19 , where the values of

    pi t are taken from step 2.

    7 Repeat steps 26 for all times of interest.

    The above approach is a direct extension of the tech-

    nique of Mauro and Varshneya.33

    Steps 1 and 2 above are

    identical to the Mauro-Varshneya approach, while steps 36

    account for continuously broken ergodicity. Whereas thepi t values computed in step 2 give the ensemble-averagedoccupation probabilities of the various microstates, step 3

    considers a particular instance of the system. By using the

    initial condition fi,i 0 =1 for one particular basin, step 3 es-

    sentially constitutes a measurement of system, causing it to

    collapse into a single microstate. The probability of collaps-

    ing into a particular microstate i is equal to the probability of

    the system sampling that microstate at the time of measure-

    ment, pi t . Equivalently, pi t can be considered a quench

    probability, i.e., the probability of the system becoming

    trapped in inherent structure i upon an instantaneous quench

    to absolute zero.

    The solution to the set of master equations in step 3gives the conditional probabilities of occupying all of the

    various j microstates after starting in microstate i and propa-

    gating for exactly the observation time tobs. No assumptions

    are made regarding which states are accessible versus inac-

    cessible, since the dynamics are computed using the actual

    transition rates Wjk. In this manner, our approach represents

    a generalization of the Palmer approach for broken ergodic-

    ity, wherein microstates must be either accessible or inacces-

    sible from each other, with no intermediate classification.7

    Also, use of the master equations in step 3 allows us to avoid

    partitioning into metabasins; in this way, our technique is

    generally applicable to any energy landscape and for any

    temperature path.The output of steps 35 is a set of Gibbs entropies Si for

    each of the possible configurations of the system, accounting

    for the actual observation time and transition rates. The ex-

    pectation value of the glassy entropy is the sum of these Sivalues weighted by their corresponding quench probabilities

    pi t as shown in Eq. 19 . Free energy can then be evalu-

    ated using Eq. 13 or 14 .

    V. RESULTS AND DISCUSSION

    In this section we apply our technique to two simple test

    cases. First, let us consider the one-dimensional fragile land-

    FIG. 3. Evolution of glassy, statistical, and equilibrium entropy values with

    respect to a time and b temperature for the system in Fig. 2. The obser-vation time is 0.01 s.

    FIG. 4. Evolution of glassy, statistical, and equilibrium free energies with

    respect to a time and b temperature for the system in Fig. 2.

    184511-7 Continuously broken ergodicity J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    8/11

    scape of Mauro and Varshneya,33 depicted in Fig. 1, whichcontains nine basins of varying energy with no clear division

    into metabasins. The transitions between basins are thermally

    activated and governed by standard transition state theory

    details are provided in Ref. 33 . Direct transitions are al-

    lowed between all pairs of basins and are not limited to the

    immediately adjacent basins. Figure 2 shows the evolution of

    potential energy for this system, starting in equilibrium at

    500 K and linearly cooling to 300 K over 1 s. Figure 2 a

    plots the glassy and equilibrium values of potential energy

    with respect to time, and Fig. 2 b plots the same values with

    respect to temperature.

    Figure 3 shows a similar plot for the glassy, statistical,and equilibrium values of entropy. The glassy entropy S is

    computed using Eq. 19 accounting for continuously broken

    ergodicity; the statistical entropy S is computed with Eq. 6with the assumption of ergodicity. At high temperatures the

    relaxation time of the system is much shorter than the obser-

    vation time, and the value of entropy corresponds to exactly

    that of the equilibrium liquid. As the system is cooled, the

    relaxation slows and the system departs from equilibrium

    i.e., undergoes a glass transition . Figure 3 shows that thedeparture of the nonergodic glassy and ergodic statistical

    entropies from equilibrium exhibits markedly different be-

    haviors. Using the statistical formulation of Eq. 6 , the glass

    transition corresponds to a freezing of the occupation prob-

    abilities and hence a freezing of the entropy. However, this

    does not account for the loss of ergodicity at the glass tran-

    sition and results in the nonergodic state always having a

    higher entropy than the corresponding ergodic state. In real-

    ity, the glass transition must correspond to a loss of entropy,

    since the loss of ergodicity limits the number of states that

    can be visited during a finite observation time. Whereas the

    ergodic formula predicts a large residual entropy of the glass

    at absolute zero, in violation of Planks statement of the third

    law,17

    the nonergodic formalism of Sec. IV correctly predicts

    zero entropy at absolute zero.

    The corresponding evolution of free energy is shown inFig. 4. In both the nonergodic glassy and ergodic statistical

    cases, the glass transition is a nonspontaneous process lead-

    ing to an increase in free energy with respect to the equilib-

    rium supercooled liquid. In the statistical case, this is due to

    a freezing of the potential energy at a higher level than that

    of the supercooled liquid i.e., the potential energy of the

    glass corresponds to that of the liquid at the glass transition

    temperature . This is also true for the nonergodic glassy free

    energy, but there is a much greater increase in free energy

    due to the loss of entropy upon ergodicity breaking. With the

    ergodic formulation of Eq. 6 , the glass is exactly the same

    macrostate as its corresponding liquid at the glass transition

    FIG. 5. Relaxation of glassy and statistical a entropies and b free ener-gies in the long-time limit toward equilibrium values.

    FIG. 6. Impact of observation time on a entropy and b free energyassuming a fixed cooling rate.

    184511-8 Mauro, Gupta, and Loucks J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    9/11

    temperature; accounting for loss of ergodicity, as in Sec. IV,

    the glass is represented as a different unstable macrostaterelative to the liquid. A thorough discussion of this topic is

    the subject of a forthcoming paper by Gupta and Mauro.57

    Figure 5 plots the isothermal relaxation of the noner-godic glassy and ergodic statistical entropies over long time.

    Whereas both achieve the same equilibrium value in the limit

    of long time, the approach to equilibrium is markedly differ-

    ent. With the ergodic formulation of Eq. 6 , relaxation is

    characterized by a decrease in entropy as it approaches equi-

    librium; accounting for the glasss broken ergodicity, relax-

    ation is characterized by an increase in entropy as ergodicity

    is gradually restored over the increasing observation time.

    Relaxation to the equilibrium state is a spontaneous process

    in both cases since, as shown in Fig. 5 b , the free energy

    decreases monotonically.

    The impact of observation time on entropy and free en-

    ergy is shown in Fig. 6. As expected, the glass transition

    occurs at a higher temperature for shorter observation times.

    In the limit of zero temperature, the systems have the same

    entropy S = 0 and free energy F = U , regardless of the

    observation time.

    We now consider a second simple landscape, shown inFig. 7. In contrast to the previous landscape, here there is a

    clear division into two metabasins, labeled A and B in the

    figure. Metabasin A has three degenerate inherent structures,

    and metabasin B has six; the inherent structures in B are at a

    potential energy H higher than those in metabasin A. The

    transition energy among the inherent structures is H/ 2 within

    both metabasins. The transition energy from metabasin B to

    metabasin A is four times as large. This effectively sets up

    two relaxation modes: fast relaxation within a metabasin

    and slow relaxation between metabasins. For this land-

    scape, we plot all quantities in nondimensional units of H,

    kB, and 1/, where is vibrational frequency.

    FIG. 7. Model potential energy landscape with two metabasins. Metabasin

    A has three degenerate inherent structures, and metabasin B has six.

    FIG. 8. Relaxation of a entropy and b free energy for the two-metabasinenergy landscape in Fig. 7 after quenching from T=4 to T=2 .

    FIG. 9. Relaxation of a entropy and b free energy for the two-metabasinenergy landscape in Fig. 7 after quenching from T=4 to T= 1/ 4.

    184511-9 Continuously broken ergodicity J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    10/11

    Figure 8 plots the relaxation of the entropy and free en-

    ergy of a glass after quenching from a nondimensional tem-perature of T=4 to T=2, as a function of normalized obser-

    vation time. Relaxation proceeds as expected, with

    monotonically increasing entropy and decreasing free en-

    ergy. However, the relaxation takes on a different character

    in Fig. 9, which plots the evolution of entropy and free en-

    ergy after quenching from T=4 to T=1/4. While the free

    energy still decreases monotonically as the system ap-

    proaches equilibrium in agreement with the second law , theentropy actually passes through a maximum during relax-

    ation and then approaches equilibrium asymptotically from

    above. The maximum in entropy following this quench is

    due to a crossover of the system from metabasin B to me-

    tabasin A. While equilibrium statistical mechanics favorsmetabasin B at the higher temperatures T=4 and T= 2 due

    to its higher degeneracy, metabasin A is preferred at the low

    temperature T= 1 / 4 due to its lower energy states. Hence,

    after the quench from T=4 to T=1/4, the system relaxes

    from metabasin B to metabasin A and passes through a point

    of maximum spreading between the two metabasins.

    This crossover point corresponds to a maximum in the

    intermetabasin entropy, which we define as

    Sinter = kBPA ln PA kBPB ln PB, 27

    where PA and PB are the total probabilities of occupying

    metabasins A and B. Note that for the purposes of this dis-

    cussion, the partitioning of the landscape into the two me-

    tabasins A and B is kept constant, regardless of the observa-tion time. Figure 10 plots the relaxation of the

    intrametabasin and intermetabasin components of entropy

    for both quenches, starting with an observation time that al-

    lows for transitions within the metabasins but not the

    transition between metabasins i.e., zero intermetabasin en-tropy . Whereas the intermetabasin entropy increases mono-

    tonically after the quench to T=2, it passes through a maxi-

    mum after the quench to T=1/4 due to the crossover effect.

    In both cases the relaxation is spontaneous monotonicallydecreasing free energy, in agreement with the second law of

    thermodynamics and irreversible dU/dtTdS/dt, asshown in Fig. 11 .

    VI. CONCLUSIONS

    We have presented a statistical mechanical framework

    for treating systems with continuously broken ergodicity.

    Our approach is a generalization of Palmers approach7

    for

    discontinuously broken ergodicity and relaxes the assump-

    tions of component confinement and internal ergodicity. It

    accounts for the actual transition rates among microstates

    and enables the direct computation of entropy loss at the

    glass transition. We have implemented our approach using a

    hierarchical master equation technique that builds on the

    work of Mauro and Varshneya.33

    Unlike the traditional mas-

    FIG. 10. Relaxation of intra- and intermetabasin entropies after quenching

    from T=4 to a T=2 and b T= 1/ 4.

    FIG. 11. Plots of dU/dt and T dS/dt after quenching from T=4 to a T=2 and b T= 1/ 4.

    184511-10 Mauro, Gupta, and Loucks J. Chem. Phys. 126, 184511 2007

    Downloaded 14 May 2007 to 199.197.135.1. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

  • 7/29/2019 Continuously Broken Ergodicity

    11/11

    ter equation approach, which assumes ergodicity, our tech-

    nique is consistent with Boltzmanns definition of entropy

    and the third law of thermodynamics. Using a simple one-

    dimensional glass former, we demonstrated the impact of

    observation time on glass transition range behavior. Depend-

    ing on the energy landscape and cooling path, a crossover

    effect may be observed whereby entropy passes through a

    maximum during the relaxation process. In all cases relax-

    ation toward equilibrium is a spontaneous and irreversibleprocess, in compliance with the second law.

    ACKNOWLEDGMENT

    The authors gratefully acknowledge valuable conversa-

    tions with Arun K. Varshneya.

    1F. T. Bantilan, Jr. and R. G. Palmer, J. Phys. F: Met. Phys. 11, 261

    1981 .2

    R. P. Feynman, Statistical Mechanics Westview, New York, 1972 .3

    M. Reiner, Phys. Today 17 1 , 62 1964 .4

    C. Monthus and J.-P. Bouchaud, J. Phys. A 29, 3847 1996 .5

    G. Bel and E. Barkai, Phys. Rev. Lett. 94, 240602 2005 .6

    G. Bel and E. Barkai, J. Phys.: Condens. Matter 17, S4287 2005 .7 R. G. Palmer, Adv. Phys. 31, 669 1982 .8

    S. Goldstein and J. L. Lebowitz, Physica D 193, 53 2004 .9

    J. L. Lebowitz, Phys. Today 46 9 , 32 1993 .10

    J. L. Lebowitz, Physica A 194, 1 1993 .11

    J. L. Lebowitz, Physica A 263, 516 1999 .12

    J. L. Lebowitz, Rev. Mod. Phys. 71, S346 1999 .13

    A. K. Varshneya, Fundamentals of Inorganic Glasses, 2nd ed. Society of

    Glass Technology, Sunderland, UK, 2006 .14

    R. J. Speedy, Mol. Phys. 80, 1105 1993 .15

    J. Jckle, Philos. Mag. B 44, 533 1981 .16

    J. Jckle, Physica B & C 127, 79 1984 .17

    H. B. Callen, Thermodynamics and an Introduction to Thermostatistics,

    2nd ed. Wiley, New York, 1985 .18

    F. H. Stillinger and T. A. Weber, Phys. Rev. A 25, 978 1982 .19

    F. H. Stillinger and T. A. Weber, Science 225, 983 1984 .20

    D. J. Wales, J. Chem. Phys. 101, 3750 1994 .21 M. A. Miller, J. P. K. Doye, and D. J. Wales, J. Chem. Phys. 110, 328 1999 .

    22J. N. Murrell and K. J. Laidler, J. Chem. Soc., Faraday Trans. 2 64, 371

    1968 .

    23D. J. Wales, Energy Landscapes Cambridge University Press, Cam-bridge, 2003 .

    24J. C. Mauro, R. J. Loucks, J. Balakrishnan, and A. K. Varshneya, Phys.

    Rev. A 73, 023202 2006 .25

    P. G. Debenedetti, Metastable Liquids Princeton University Press,Princeton, NJ, 1996 .

    26M. Goldstein, J. Chem. Phys. 51, 3728 1969 .

    27F. H. Stillinger, J. Chem. Phys. 88, 7818 1988 .

    28P. G. Debenedetti, F. H. Stillinger, T. M. Truskett, and C. J. Roberts, J.

    Phys. Chem. B 103, 7390 1999 .29

    P. G. Debenedetti and F. H. Stillinger, Nature London 410, 259 2001 .30 F. H. Stillinger and P. G. Debenedetti, J. Chem. Phys. 116, 3353 2002 .31

    T. F. Middleton and D. J. Wales, Phys. Rev. B 64, 024205 2001 .32

    J. Hernndez-Rojas and D. J. Wales, J. Non-Cryst. Solids 336, 218

    2004 .33

    J. C. Mauro and A. K. Varshneya, J. Am. Ceram. Soc. 89, 1091 2006 .34

    J. C. Mauro and A. K. Varshneya, Am. Ceram. Soc. Bull. 85, 25 2006 .35

    R. A. Denny, D. R. Reichman, and J. P. Bouchaud, Phys. Rev. Lett. 90,

    025503 2003 .36

    B. Doliwa and A. Heuer, J. Phys.: Condens. Matter 15, S849 2003 .37

    B. Doliwa and A. Heuer, Phys. Rev. E 67, 031506 2003 .38

    G. Fabricius and D. A. Stariolo, Physica A 331, 90 2004 .39

    G. A. Appignanesi, J. A. Rodrguez Fris, R. A. Montani, and W. Kob,

    Phys. Rev. Lett. 96, 057801 2006 .40

    J. C. Mauro, R. J. Loucks, and J. Balakrishnan, J. Phys. Chem. B 110,

    5005 2006 .41

    M. C. Mackey and M. Tyran-Kamiska, J. Stat. Phys. 124, 1443 2006 .42 D. Ruelle, Proc. Natl. Acad. Sci. U.S.A. 100, 3054 2003 .43

    D. Ruelle, Phys. Today 57 5 , 48 2004 .44

    D. Daems and G. Nicolis, Phys. Rev. E 59, 4000 1999 .45

    B. Bag, J. Chem. Phys. 119, 4988 2003 .46

    R. Zwanzig, Nonequilibrium Statistical Mechanics Oxford UniversityPress, Oxford, 2001 .

    47B. Gaveau and L. S. Schulman, J. Math. Phys. 37, 3898 1996 .

    48S. Langer, J. P. Sethna, and E. C. Grannan, Phys. Rev. E 41, 2261 1990 .

    49J. H. Gibbs and E. A. DiMarzio, J. Chem. Phys. 28, 373 1958 .

    50G. Adam and J. H. Gibbs, J. Chem. Phys. 43, 139 1965 .

    51W. Gtze and L. Sjgren, J. Phys. C 21, 3407 1988 .

    52D. Thirumalai, R. D. Mountain, and T. R. Kirkpatrick, Phys. Rev. A 39,

    3563 1989 .53

    W. van Megen and S. M. Underwood, J. Phys.: Condens. Matter 6, A181

    1994 .54

    D. L. Stein and C. M. Newman, Phys. Rev. E 51, 5228 1995 .55 S. Ishioka and N. Fuchikami, Chaos 11, 734 2001 .56

    B. Coluzzi, A. Crisanti, E. Marinari, F. Ritort, and A. Rocco, Eur. Phys.

    J. B 32, 495 2003 .57

    P. K. Gupta and J. C. Mauro, J. Chem. Phys. to be published .

    184511-11 Continuously broken ergodicity J. Chem. Phys. 126, 184511 2007