stochastic dynamics

Upload: michael-pearson

Post on 09-Feb-2018

232 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/22/2019 Stochastic Dynamics

    1/161

    Effective stochastic dynamics in deterministic

    systems: model problems and applications

    by

    Gil Ariel

    A dissertation submitted in partial fulfillment

    of the requirements for the degree of

    Doctor of Philosophy

    Department of Mathematics

    New York University

    September 2006

    Eric Vanden-EijndenAdvisor

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    2/161

    UMI Number: 3234111

    Copyright 2006 by

    Ariel, Gil

    All rights reserved.

    INFORMATION TO USERS

    The quality of this reproduction is dependent upon the quality of the copy

    submitted. Broken or indistinct print, colored or poor quality illustrations and

    photographs, print bleed-through, substandard margins, and improper

    alignment can adversely affect reproduction.

    In the unlikely event that the author did not send a complete manuscript

    and there are missing pages, these will be noted. Also, if unauthorized

    copyright material had to be removed, a note will indicate the deletion.

    UMIUMI Microform 3234111

    Copyright 2006 by ProQuest Information and Learning Company.

    All rights reserved. This microform edition is protected against

    unauthorized copying under Title 17, United States Code.

    ProQuest Information and Learning Company

    300 North Zeeb Road

    P.O. Box 1346

    Ann Arbor, Ml 48106-1346

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    3/161

    Gil Ariel

    All Rights Reserved, 2006

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    4/161produced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    5/161

    to Orit, Guy, Mia,

    and who ever may join us next

    iv

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    6/161

    Acknowledgements

    First and foremost, I would like to thank my advisor, Eric Vanden-

    Eijnden. Working with Eric, who has a rare skill to find the interesting

    mathem atics in physical problems has been a rewarding experience. His

    guidance is evident in every aspect of this work.

    I would also like to thank the faculty at the Courant institute, in partic

    ular Jonathan Goodman, Robert Kohn, Charles Newman, Ragu Varadhan

    and Lai-Sang Young for always being willing to explain, talk and advise.

    I am also grateful to my fellow students: An drea Barreiro, M atthias

    Heymann, Jose Koiller, Stan Mintchev, Junyoep Park, Haiping Shen and

    Paul Wright for making my time at Courant pleasant and helping with all

    those little things we know we should know, but are nevertheless not quite

    sure of.

    Finally, I would like to than k my family. My wife, Orit, who followed

    me to New York and now to Texas, and with great love built our new home.

    Our son Guy, who grew to become a little boy and provided ample enjoy

    able distractions, and last but not least Mia, who came just in time to be

    included.

    v

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    7/161

    Contents

    Dedicat ion ........................................................................................................ iv

    Acknowledgements ....................................................................................... v

    List of Figures ................................................................................................. ix

    List of T a b le s ................................................................................................. xi

    1 Introd uction 1

    2 The Kac-Zw anzig m odel 13

    2.1 Introduction ......................................................................................... 13

    2.2 Th e Kac-Zwanzig model: Formulation

    and elementary propert ies ............................................................... 15

    2.3 A stro ng limit th e o re m ...................................................................... 19

    2.3.1 Introduction ............................................................................. 19

    2.3.2 Main results and ex am ple s .................................................. 21

    vi

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    8/161

    2.3.3 Proof of the th e o re m ............................................................. 27

    2.3.4 Conclusion and generalizations ......................................... 49

    2.4 Testing Transition State Theory on Kac-Zwanzig Model . . . . 53

    2.4.1 Introduction .............................................................................. 53

    2.4.2 The m od el ................................................................................. 58

    2.4.3 M etastability in K ac -Z w an zig ............................................. 61

    2.4.4 Numerical ex pe rim en ts .......................................................... 65

    2.4.5 Tran sition rate s for the limiting dynamics ......................... 70

    2.4.6 Transition state th e o ry .......................................................... 82

    2.4.7 Local dynamics around the hyperbolic point:

    Why does VTS T works while naive TS T does not? . . 93

    2.4.8 Concluding re m ar ks ................................................................. 96

    3 Accelerated simulation of a heavy particle in a gas of elastic

    spheres 101

    3.1 Introduction ............................................................................................. 101

    3.2 The m od el ................................................................................................. 105

    3.3 Limiting dynam ics ................................................................................ 109

    3.3.1 Ap proximate Markov nature of colloid-gascollisions . . 112

    3.3.2 The rate of momentum tr a n s fe r............................................ 114

    vii

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    9/161

    3.3.3 Th e limiting equations of m o t i o n .........................................115

    3.3.4 Improved a c c u ra c y ................................................................... 120

    3.4 Th e numerical scheme ......................................................................... 121

    3.4.1 The basic algorithm ................................................................ 123

    3.4.2 Higher order accuracy ............................................................. 127

    3.4.3 Some technical a s p e c ts ............................................................. 128

    3.4.4 Pra ctical considerations ......................................................... 130

    3.5 Numerical e x p erim en ts ..........................................................................131

    3.5.1 Simulation m e th o d ................................................................... 131

    3.5.2 Simulation re su lts ....................................................................... 133

    3.6 Conclusion ................................................................................................. 134

    Bibliography 138

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    10/161

    List of Figures

    2.1 Comparison of the rand om noise, /y(f) (solid line), and the

    limiting one (f) (dashed line). The frequencies are draw n

    from a one sided exponential distribution and N ~ 2000. The

    bath in itial cond itions are cano nical .............................................

    2.2 A semi-log plot of the distrib ution of the waiting times between

    transitions, P [ t s > s], for the case 7 = 10. In figure (a), the

    sharp jump near the origin is due to rapid re-crossings of the

    {xo = 0} plane. In figure (b), the statistic s of trans ition times

    confirms the quasi-Markov hypothesis. The slope of the curve

    is 1.1-10-4, the same as the average tran sition rate. The graph

    diverges from the linear fit near the origin ..................................

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    11/161

    2.3 A semi-log plot of the joint dis tribu tion of successive waiting

    time, -P[tqV + > s], for the case 7 = 10. As expected, th e

    slope of the curve is 5 10- 5 which is about half the average

    transition rate ........................................................................................... 1 0 0

    3.1 A car icature of a colloid moving with velocity v. The large

    sphere collides with all particles whose center of mass is within

    a distance of R + r from its center. Hence, the colloid collides

    with all particles in a volume of vdtn(R + r) 2 ................................. 108

    3.2 Th e figure depic ts a situatio n in which subsequent collisions

    be tween the colloid an d a part icula r gas pa rt icle carries some

    memory: a slow moving particle approaches the colloid (a).

    After this collision, the particle bounces off with almost the

    same speed, only in the oppo site direction (b). After a second

    collision with a fast moving particle (c), the colloid catches up

    with the first particle and collides again (d) ..................................... I l l

    x

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    12/161

    List of Tables

    2.1 Com parison between tran sition rates with two different can

    didate metastable sets.............................................................................. 67

    2.2 Com parison between TS T predictions and simulation results

    for small and large values of 7 ................................................................90

    3.1 Sim ulation results for a system of iV = 40000 gas partic les

    and inverse temperature ( 3 = 1 . The table compares the av

    erage kinetic energy and diffusivity of the colloid as obtained

    using our newly suggested, accelerated method, with values

    predicted by th e lim iting OU process................................................. 134

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    13/161

    Chapter 1

    Introduction

    Deterministic dynamical systems, in particular, Hamiltonian systems, often

    display very com plicated chaotic behav ior when th e number of degrees of free

    dom in the systems is large. It is not surprising th a t the evo lution of certain

    observables in these systems can be approximated by a stochastic process.

    Results in this direction abound in the literature [29, 36, 42, 47, 77]. A typ

    ical example is the extraction of conformational dynamics of biomolecules,

    which is stochastic in nature even though the underling dynamics is Hamil

    ton ian [31, 61]. In this con text, th e development of simple models in which

    the emergence of the stochastic behavior out of the deterministic dynamics

    can be analyzed rigorously is of importance.

    1

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    14/161

    In this thesis we consider two different model examples in which a single

    particle is coupled to a larger, man y pa rt icle system which plays th e role of

    a heat ba th. The dynamics in both cases is deterministic. As a first step we

    prove th a t taking a part icula r lim it, and un der appro priate in itia l conditions,

    the dynamics of the single particle of interest is given by a solution to a

    stochastic equation. Next, we apply the knowledge of the limiting equation

    for the study of some properties of the deterministic dynamics.

    The first example, presented in Chapter 2, considers a variant of the Kac-

    Zwanzig model [24, 25, 76]. In th is model a single dis tinguish ed partic le is

    moving in a given external potential and is coupled to N harmonic oscilla

    tors via linear springs. The system is one-dimensional. The oscillators have

    random frequencies that are chosen independently from a probability distri

    bu tio n. In it ia l co nd itions are chosen rando mly from an inva rian t measure.

    In Section 2.3 we prove that under certain sufficient conditions on the model

    pa rameters , and in any fin ite time segment, th e traje cto ry of th e partic le

    converges strongly (in L 2) to the solution of an effective stochastic equa

    tion. The limiting equation has the form of a generalized Langevin equation

    (GLE) and satisfies a fluctuation-dissipation relation. Strong convergence is

    proved by map ping the in itial co nd itions of the fin ite dimensional system to

    2

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    15/161

    the probability space on which the random noise term of the GLE is defined.

    Bo th canonical and micro-canonical ensembles are considered. We also prove

    convergence for a more general class of initial conditions that are absolutely

    continuous with respect to the above two invariant measures.

    Once the form of the limiting equation is established, it is used for study

    ing properties of the d ynamics a t large, but finite system size. In the second

    half of the chapter, the model is used as a benchmark problem for a rigorous

    analysis of transition state theory (TST) and variational TST (VTST).

    In the particular case in which the distinguished particle moves in a

    double-well potential the dynamics is bistable. Amid the complexity of

    individual trajectories, it is found that that most trajectories remain con

    fined for very long periods of time in well separated regions of phase-space

    whose boundary are loosely determined by the double-well potential, and

    only switch from one region to ano ther occasionally. The confinement is due

    to the presence of dynam ical bottlenecks between these regions. The system

    is then said to display m etastability, and th e regions in which the trajec tories

    remain confined are referred to as meta stable sets. Example of systems dis

    playing m eta stability abound in natu re , with examples arising from physics

    (phase transitions), chemistry (chemical reactions, conformation changes of

    3

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    16/161

    bio-molecules), biology (switches in popula tio n dynamics), and many others.

    In these systems, it is worth verifying whether the dynamics can be approx

    imated by a Markov chain over the state-space of the metastable sets with

    appropriate ra te constants. The m ain question of interest then becomes the

    determination of these rate constants. This question is nontrivial because it

    amounts to understanding the pathways by which the transitions between

    the metastable sets occur. These pathways are usually non-trivial and com

    plicated .

    It is here that our knowledge of the limiting dynamics can be applied

    by ca lculat ing th e ex ac t transit io n ra te s for th e lim iting equation. This is

    done by spectral decomposition of the backward operator characterizing the

    limiting stochastic equation. Eigenvalues and eigenfunctions are found using

    the matched asymptotics method [60].

    One of the earliest attempts to determine transition pathways and rate

    con stan ts is tran sitio n sta te theo ry (TST ) [22, 35, 75]. Under the sole as

    sumption that the dynamics of the system is ergodic with respect to some

    known equilibrium distribution, TST gives the exact average frequency at

    which trajectories cross a given hypersurface or hyperplane which separates

    two metastable sets of interest. This average frequency can be used as a

    4

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    17/161

    first approximation for the frequency of transition between the metastable

    sets. Unfortunately, it was recognized early on th a t this approx ima tion can

    be qui te po or , for not every crossing of th e dividing surface correspond s to

    a transition between the m etastable sets. Indeed, the trajectories can cross

    the dividing surface many times in the course of one transition . As a result,

    the TST prediction for the frequency of transition always overestimates the

    ac tua l frequency, sometimes grossly so. One way to minimize this problem is

    to use the freedom in the choice of dividing surface. The best prediction for

    the frequency from TST is the one corresponding to the dividing surface with

    minimum crossing frequency. In variational transitio n state theory (VTST )

    [35, 51, 65, 70] one minimizes the TST rate over some class of candidate

    separating hyperplanes.

    Unfortunately, VTST (just like TST) is an uncontrolled approximation,

    for it only provides an upper bound on the transition frequency between

    the m etastab le sets. In general, one does not know how sharp th is boun d is.

    Other assumptions beyond TST and VTST are usually difficult to assess too.

    Are successive transitions between the metastable sets well approximated by

    Poisson events (i.e. statistically indepen dent and with exponentially dis

    tributed waiting times) as required for the approximation of the dynamics

    5

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    18/161

    by a Markov chain to hold? How does th is pro perty depend s on th e defin ition

    of the m etastable sets? Etc.

    The questions described above are first studied here through computer

    simulations. M etastable sets are identified and it is shown tha t the dynam ics

    is well app rox imated by a two-s tate Markov process over these sets. We also

    examine the various assumptions required for a consistent use of TST in this

    system.

    We then turn to evaluate the transition rate using TST and VTST and

    compare them to both numerical results and to transition rates obtained

    from the limiting equations. We find th a t the VTST approxim ation for

    the transition rate is exact, rather than just an upper bound, which is not

    generally the case. We explain why the m ethod works for the c urrent model,

    draw necessary conditions for the applicability of VT ST in the way used here,

    and discuss the nature of spurious recrossings of the TST plane in light of

    our findings.

    The motivation for studying TST in the context of the Kac-Zwanzig

    model is its simplicity, which allows a detailed , rigorous analysis. We have

    shown that in this model, VTST given the exact rates of crossing between

    the two me tastable sets. Unfortunately, th e general case is not as simple. In

    6

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    19/161

    part icular, th e only reason why hy perplanes, ra th er th an more complicated

    hypersurfaces were sufficient is the quadratic nature of the interaction with

    the bath . This is obviously an artifact of this particu lar model. Because

    the transition region is not localized at a saddle point, the VTST dividing

    surface is in general a more complicated surface than the lift-up in phase

    space of a hyperplane in configu ration space. It has been proved [64] tha t

    VTST is always exact when minimizing over hypersurfaces in phase space.

    The question of whether minimization over hypersurfaces in configuration

    space (position only, not momentum) is sufficient is still open, and is of

    much practical importance.

    The theoretica l pa rt of proving stong convergence to the limiting equa tion

    poses some inte rest ing op en problems and generalizations as well. The ra te

    at which the solution of the finite system converges to the solution of the

    limiting one depends exponentially on the time length considered. Hence,

    it seems like stron g convergence is too strong . It would be interesting

    to prove a weak type of convergence (in dis tribu tion) for all times. Such a

    statement is not trivial since there is no clear separation between the time

    scale for the dynamics of the distinguished particle and that of the bath.

    The second example considered in this thesis, presented in Chapter 3,

    7

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    20/161

    studies a system of elastic spheres. The dy namical theory of Brownian motion

    has been the subject of extensive research both from the theoretical and

    the num erical points of view. A particularly simplified model is th a t of

    spheres interacting only through elastic collisions. In this model, a large

    heavy spherical particle (colloid) is placed in a gas of smaller lighter ones.

    Due to the collective effect of numerous collisions between colloid and gas,

    the d isplacement an d velocity of the colloid seem erratic an d stochastic. It

    has been proved [15, 34, 62, 63] that in the limit of an infinitely massive

    colloid, and under specific scaling of time and space, the dynamics of the

    colloid can be described by an Ornstein-Uhlenbeck (OU) process

    t

    dx(y) = v( t )dt

    ( i . i )

    dv(t) = av{t)dt + V D d W ( t ) ,K

    where x( t) and v(t) denote the position and velocity of the colloid, respec

    tively, W(t) denotes the Wiener process, a > 0 is the viscosity and D > 0 is

    the diffusion coefficient. Althoug h the limiting eq uation is known, the rate

    of convergence is still undeterm ined. As a result, comparison between the

    dynamics predicted by the limiting OU process and numerical simulations is

    of importance.

    Different attempts for simulating hard sphere systems with disparate

    8

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    21/161

    masses can be found in the litera ture [2, 39], all of which amou nt to advancing

    the system collision by collision [1, 27]. Th e main draw back of this metho d is

    that due to the disparity between the masses of the two constituents, it takes

    an extremely large number of colloid-gas collisions to produce a significant

    change in the colloid position or velocity. In add ition, one also has to resolve

    all collisions between gas partic les themselves. A simple estimation shows

    that the required simulation time is proportional to the ratio between the

    colloid and gas particle masses.

    In the second chapter we present an alternative method for simulating

    the system in three dimensions. Although it is stochastic in natu re, we

    prove th a t th is method ap proximates th e dynamics of th e colloid with in a

    calculated error. The efficiency of the me thod is indepen dent of the ra tio

    be tween th e two masses, which const itutes a big improvem ent over th e full

    simulation method . The a lgorithm is based on the heterogeneous multiscale

    meth od [16, 18] and is co nstructed along the lines of [18, 23, 69]. Th e basic

    idea is to calculate the drift, av, and diffusion terms, DdW/dt , on the fly

    by simulating th e gas for a sh ort tim e segment At . Using these values the

    colloid is advanced by a big time step As according to the limiting equation

    (1 . 1 ).

    9

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    22/161

    The key property of the system used in this Chapter is the gap between

    the timescale on which the gas and the colloid evolve. Because the colloid

    is so much heavier than gas particles, the gas will reach equilibrium before

    x(s) and v(s) change considerably. This mo tivates an approximation of the

    stochastic differential equation 1.1 by a forward Euler scheme with step size

    As

    (x i+i = xt+ ViAs

    ( 1.2 )

    vi+i = Vi - aViAs + y/DAs{,i,

    where are indepen dent random variables with normal distribu tion in M3.

    The idea behind th e algorithm is to evaluate a and D in each step by making

    a short simulation of the gas with fixed v V{. Each Euler step consists of

    three parts:

    1. Simulate the gas for a time segment of length A t while keeping the

    velocity of the colloid is fixed.

    2. Use the statistics o btained from pa rt 1 for evaluating a and D.

    3. Move the colloid according to the forward Euler approximated equa

    tions of motion (1 .2 ).

    Denoting the ratio between the colloid and gas particle masses by E 1, it is

    10

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    23/161

    shown that ifA tis 0 (1 ), i.e., independ ent of E, while As is proportio nal to E,

    then the variance in the error introduced by this approximation is bounded.

    Hence, the simulation is accelerated by a factor E. This is the m ain result of

    Chapter 3.

    The numerical experiments presented here are benchmark examples, sim

    ple enough for demonstra ting th e ad vantag es of th e method. First , we calcu

    lated the viscosity and diffusion coefficients of the colloid in the case E = 100.

    Th e results deviate substan tially from values obtained by Du rr et. al. [15]

    in the limit E >oo are found. The source of the errors are co rrections due

    to correlations and the finite size of gas particles, not taken into account by

    the limiting equation. A second example, in which E 104, demonstrates

    that when E is large enough, the dynamics of the colloid is well described by

    the limiting stochastic equation.

    The new simulation method suggests numerous applications and general

    izations in which the colloid dynamics is more complicated, and the limiting

    equation is more difficult to ob tain or may no t be known. For instance, th e

    case in which the colloid is not a sphere but an ellipsoid poses an enormous

    com putation al challenge. The an gular mom entum of the colloid has to be

    taken into account and calculating collision times is much more complicated.

    11

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    24/161

    It is also interesting to investigate more thoroughly a system in which the

    gas is in a dense regime. In this case the correlation function C{t\ 12) does

    not decay exponentially, and a larger time step A t may be required.

    Another interesting generalization is to the case of two or more colloids.

    When the separation between two colloids is a few times the diameter of gas

    pa rticles , th e collective effect of colloid-gas collisions is to pu sh th e colloids

    closer toge ther [7, 45]. In other words, th e gas induces an effective close

    range attra ction . Because the rang e of attra ctio n depend s on the size of gas

    particles, it vanishes in th e lim it E > oo. This implies th a t th e lim iting

    equations of motion of the two colloids is the same as for a single one (except

    for elastic collisions between th e colloids). However, with a fixed bu t large E,

    the dynamics is completely different. We expect to find th a t und er app ropri

    ate scaling, the two colloids become effectively trapped in a metastable state

    keeping them close for a long time. Hence, the limiting rate for the dynamics

    may be the escape rate o ut of this state. This rate does not depend on E

    expo nentially since fluctu ation s do not vanish even in the limit E oo.

    12

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    25/161

    Chapter 2

    The Kac-Zwanzig model

    2.1 Introduction

    In this Chapter we consider a variant of the Kac-Zwanzig model [24, 25, 76].

    In this model a single distinguished particle is moving in a given external

    potential and is coupled to N free one dimensional free particles via lin

    ear springs. The pa rticles oscillate around th e distinguished one and have

    random frequencies that are chosen independently from a probability distri

    bution. In itia l co nd itions are chosen rand omly from an inva rian t measure.

    In Section 2.2 we introduce the model in detail and derive the Hamiltonian

    equation of motion.

    13

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    26/161

    The re st of the chap ter is devided into two topics. In the first, Section 2.3,

    we prove that with appropriate initial conditions, and in any finite time

    segment, the trajectory of the particle converges strongly (in L2) to the

    solution of an effective stochastic equation. The limiting equa tion has the

    form of a generalized Langevin equation (GLE) and satisfies a fluctuation-

    dissipation relation. Strong convergence is proved by map ping the initial

    conditions of the finite dimensional system to the probability space on which

    the rando m noise term of the GLE is defined. Bo th canonical and micro-

    canonical ensembles are considered. We also prove convergence for a more

    general class of initial conditions that are absolutely continuous with respect

    to the above two invariant measures.

    The second topic is an application of the limiting equation of motion for

    the distiguished particle in a particu lar choice of model param eters. In Sec

    tion 2.4 we use the Kaz-Zwanzig model as a b enchm ark problem for a rigorous

    analysis of transition state theory (TST) and variational TST (VTST).

    In a par ticula r case, in which the distinguished particle moves in a double

    well potential the dynamics is bistable. We begin by studying metastab ility

    in Kac-Zwanzig throu gh num erical simulations. In particula r, we show th at

    by an appro priate choice of metastable sets th e dynamics of th e pa rtic le may

    14

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    27/161

    be ap proximately described by a two state s Markov process. We th en use

    our knowledge of the limiting equation to calculate transition rates for the

    limiting equations. This is done by spectral decom position of the backward

    ope rator characterizing the limiting stochastic equation. Eigenvalues and

    eigenfunctions are found using matched a symptotics [60]. The numerical

    and limiting results will serve in our analysis of the various assumptions

    underlying TST , VSTS and th e predictions of these theories. We find th at

    the VTST approximation for the transition rate is exact, rather than just an

    upp er bound , which is not generally the case. We explain why the m ethod

    works for the current model, draw necessary conditions for the applicability

    of VT ST in the way used here, and discuss the n atu re of spurious recrossings

    of the TST plane in light of our findings.

    2.2 The Kac-Zwanzig model: Formulation

    and elementary properties

    The Kac-Zwanzig model is a system describing the evolution of a distin

    guished particle with unit mass and position x 0, moving in an external po

    tential V (x q) [24, 25, 58, 76]. Th e pa rticle is coupled by a harm onic p otentia l

    15

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    28/161

    to a bath of N particles of mass m* > 0 with positions Zj, i = 1 , . . ., N. The

    system is described by the Hamiltonian

    # ( x , p) = i p i + v ( x 0) + 2 ^ - + Y l ^ X i ~ x ^ ^ t2-1)i = l 1 t = l

    where, for short hand we use the vector notation x = (x0, X \ , - , x ^ ) , P =

    (po.Pi, ,Pn)i and Pi is the momentum associated with X{. The coupling

    constant between the distinguished particle and each oscillator in the bath

    is 7/ N > 0. Th e scaling with N emphasizes the fact that the interaction is

    weak. The equations of mo tion derived from this Ham iltonian are:

    N

    Xq = - V ' ( x q) - ~ - X i )

    - 1 (2 .2)

    X i = U J i { x0 - X i )

    where

    J - z k - (2'3)

    We also assume that the frequencies {(^1) 1=1,.,.,jv are independent and iden

    tically distrib uted (i.i.d.) rando m variables whose distribu tion is absolutely

    continuous to the Lebesgue measure on M.

    Here we proceed informally and postpone the mathematically rigorous

    analysis to Section 2.3. Similar arguments can be found in [1 1 , 29, 32, 58].

    In the derivation presented here we assume that initial conditions for the

    16

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    29/161

    bath are dis tr ib uted according to th e canonica l ensemble, i.e., a t t = 0 bath

    pa rticles are dis tr ib uted according to th eir equilib rium Gibbs measure which

    is given by

    Xi (0) = A f ( x 0(Q), N/ /3at i)

    (2.4)

    P i ( 0 ) = N ( Q , m i / P ) ,

    where a 2)denotes the Gaussian distribution w ith mean m and variance

  • 7/22/2019 Stochastic Dynamics

    30/161

    In the limit N > oo, th e strong law of large numbers implies th a t for any

    fixed t, R N(t) will converge to its average R(t)

    R( t) = lim RnU) = 7 lim -J- V '' cosuit = yElcoswf]. (2.9)N00 N>OON

    In order to ev aluate the rate of convergence, we calculate the second moment

    of Rn. Breaking all double sums into the diagonal and off-diagonal pa rts

    yields

    E[RN( ti)RN(t2)] E[.Rjv(i)]E[i?/v(f2)] = ^ ( i v ) (2-10)

    We conclude that the limiting equation describing the dynamics of the dis

    tinguished particle has the form of a Generalized Langevin equation

    rf 1x 0+ V'{ xQ) + 7 J R( t - T)x0(T)dT = (2 .1 1 )

    where ^(t) is a G aussian process with zero mean a nd covariance function R( t) .

    The equality between the memory kernel R(t) and the covariance function

    of the random noise is a fluctuation-dissipation relation.

    18

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    31/161

    2.3 A strong limit theorem

    2.3.1 Introduction

    Deterministic dynamical systems with a large number of degrees of freedom

    typically display chaotic behavior and it is not surprising that the evolution

    of certain observables in these systems can be approximated by a stochastic

    process. Results in th is direction abound in th e li tera tu re . Most of these

    results, however, are of weak convergence-type, i.e. it is shown th a t the

    evolution of the observables tends to that of a stochastic process in some dis

    tribu tiona l sense. It is more surprising th a t th e evolution of some observables

    in deterministic dynamical systems converges pathwise to that of a stochas

    tic process, since this requires relating the probability space on which this

    stochastic process is defined to the pa ram eters in the system. The present

    chapter offers a result in this direction.

    Consider

    N

    m0x0+ f ( x Q)+ rriiUjf(xo - Xi) = 0

    < R is a C 1 function, and {u i ,mi} i=i...at

    are positive param eters. Eq uation (2.12) is the well-known system introduced

    19

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    32/161

    by Ford, Kac, Mazur and Zwanzig [24, 25, 76] as a toy model to investigate

    several issues in nonequilibrium sta tistic al mechanics. Kac-Zwanzig model

    has been intensely investigated, both in the physics and mathematical liter

    atures [11, 24, 25, 29, 32, 38, 40, 41, 42, 57, 58, 74, 76]. In pa rt icular it is

    known that, with appropriate choices of /(), {w,, n and initial con

    ditions, in the limit as N oo the traje cto ry of the distinguished particle,

    {x0 (t),x 0(t )} , can be approximated by the solution of a integro-differential

    equation with random coefficients, namely

    m 0Xo + f { X Q) + J R ( t - T ) X 0(T)dT = (2.13)where 0 > 0 is a parameter, R : [0, oo) t>R is a ce rta in memory kernel and

    : [0, oo) t>R is a Gaussian ran dom fu nct ion w ith mean zero and covariance

    /?(). Th e typical results, however, are weak convergence theorem s. The

    pu rpose of th is cha pter is to offer a strong er convergence re su lt, namely the

    pa thwise convergence of {xo(t), x 0(t )} toward { X 0(t),Ao(t)} as N>oo.

    The dynamics of a Hamiltonian system is given by a system of determin

    istic differential equations. In many particle systems, solving these equations

    is either impossible or imprac tical, bo th an alytically and numerically. How

    ever, it is often the case that one is not interested in the dynamics of the

    full system, b ut ra the r only in a small part of it. The rest of the system is

    20

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    33/161

    referred to as a bath . It is then reasonable to look for effective equations

    that describe the behavior of the smaller, interesting part, in the presence of

    the ba th. We prove tha t if the b ath initial conditions are distributed accord

    ing to the microcanonical invariant measure (and assuming some sufficient

    conditions on the model parameters), then the solution of the equations of

    motion for a finite N converges strongly (in L 2), in the limit N >oo, to the

    solution of an effective stochastic integro-differential equation describing the

    dynamics of the distinguished particle.

    2.3.2 M ain results and exam ples

    We study a variation of the well-known Ford, Kac, Mazur an d Zwanzig model

    [24, 25, 76], in which a single - distinguished - one dimensiona l partic le is

    coupled to a ba th ofNone dimensional free particles via harmo nic po tentials.

    Similar models were previously studied from several different aspects [11, 24,

    25, 29, 32, 38, 40, 41, 42, 57, 58, 74, 76]. As already n oted above, the model

    is described by the Hamiltonian

    where xo, Po and m 0 denote the position, momentum and mass of the dis

    tinguished particle we are interested in describing, and ay, px and m, denote

    ^ { x i - x o ) 2, (2.14)

    t=i

    N

    21

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    34/161

    the position, momentum and mass of the i-th bath oscillator, i = 1 . . . N .

    For shorthand we use the notation x = (xq,Xi, . . . , xn ) and similarly for

    writing vectors in MN+1. The disting uished particle is placed in an extern al

    potential V{xq). Some restrictions on this potential that guaranty existence

    of solutions are specified later. The coupling between th e pa rticle and each

    oscillator is taken as harmonic, with spring constant 7/ N > 0. Th e scaling

    with N emphasizes the fact that the interaction between the particle and

    each one of the oscillators is weak. The equa tions of mo tion derived from

    this Hamiltonian are (2.12) where

    For fixed N , the Hamiltonian dynamics preserves the total energy and the

    dynamics is restricted to an energy shell {H = E}. Since the to tal energy is

    an extensive variable, we take E = Eq+ N//3 , where Eq= p%/(2mo) + V(x q)

    is the initial energy of the distinguished particle in the absence of the bath.

    The parameter 0 plays the role of an inverse tempera ture. We therefore make

    the following assumptions regarding initial conditions:

    Assumption la: The initial conditions of the distinguished particle

    are fixed to (xo,Po).

    22

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    35/161

    and for the bath,

    Assumption lb: The initial conditions of the bath distributed ac

    cording to the microcanonical equilibrium measure.

    For fixed N this measure is given by

    dtiNtE(x,P) = | ^ )r (2'16)

    where | | is the s tanda rd Euc lidean norm in R2iV+2, dcr(x,p) denotes a surface

    element (Lebesgue measure) on H (x , p) = E, and Z E is a normalization

    constant.

    Louivilles Theorem implies th at the Ham iltonian dynamics preserves vol

    ume in phase space. Hence, the equilibrium distributio n of the m icrocanon

    ical ensemble is invariant under the dynamics (2 .1 2 ).

    We also make the following assumptions regarding the model parameters:

    Assum ption 2: The bat h frequencies, {wj}i=i...Ar are independen t, iden

    tically distrib uted (i.i.d.) rando m variables, with probability density

    function (PDF) p(u>) with respect to (w.r.t.) the Lebesgue measure

    on [0, oo). In add ition , p(u>) is strictly positive in its support and the

    number of spectral gaps is finite, i.e., the support is a finite union of

    segments.

    23

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    36/161

    The assumption on the finite number of spectral gaps can be relaxed at the

    cost of a slower convergence rate. Th is generalization is discussed in Section

    2.3.3. In order to avoid confusion, exp ecta tions w.r.t. initial cond itions are

    denoted Eo[-], while expec tations w.r.t. the frequencies are deno ted =

    J[-]p(uj)duj. Assumption 2 can be relaxed to admit a more general case in

    which both the coupling coefficient 7 and the PDF p(cu) depend on N , but

    converge to a b ound ed integrable function. This general set up is discussed

    in Section 2.3.4.

    Our goal is to derive a stochastic equation that approximates the dy

    namics of the distinguished particle for large N . The effective equation has

    the form of a generalized Langevin equation (2.13), where R(t) = Ew[cos ut\

    is a memo ry kernel and f() is a rando m noise. It is a Gaussian process

    with zero mean and covariance given by a fluctuation-dissipation relation

    Eo,u[(ti)(t2)] = R ( h 2)- If V'{%q) is uniformly Lipshit, then the solution

    of (2.13) is well defined.

    Denote by (xN(t),pN (t)) the full solution of the 2N+ 2 equations of mo

    tion (2 .1 2 ), and by (xq (t),p^ ()) its projec tion on the coordinates describing

    the d istinguished particle. Th e solution of the effective stochastic equation

    (2.13) is denoted (X(t), P(t)). We prove the following strong convergence

    24

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    37/161

    theorem which is the main result of this chapter

    Theorem: Suppose assumptions 1 and 2 hold. Then, fo r any T ). For instance, if the tail of p(uj) is polynomial, C u~ q ) < D u ~ q, then I = (q l ) / (3 q1). I f the tail of p(u>) is exponential or

    better, then Z= 1/3 . In the case of macrocanonical bath init ial condit ions,

    the random noise (t) is a Gaussian process with zero mean and covariance

    m t i ) s ( t2 ) = R { t i - t 2).

    Note th a t th is rate is only an upper bo un d. Improved bo un ds may be ob

    tain ed for specific cases. Th e pro of of strong convergence requires identifi

    cation of a map between the probability space on which the random initial

    conditions are defined, and the probability space on which the effective ran

    dom noise is defined. Path -wise convergence of Xq (t) to xq(Z) is obtained

    using the Gronwall inequality.

    In particular, the theorem implies that the probability that trajectories

    of the distinguished particle given by the finite system (2 .1 2 ) are far from

    25

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    38/161

    that of the effective one (2.13) can be made arbitrarily small, i.e.,

    lim PN oo

    sup (t) X()| > eo< t < T

    = 0 . (2.18)

    Our results improve on previous works in several aspects. Fir st, we con

    sider a wide range of model parameters rather than a particular choice of

    PDF p(u>). Second, our initial conditions are drawn from the microcanon-

    ical invariant measure, and not the canonical one. Since the H amiltonian

    dynamics conserves energy, the microcanonical ensemble is the appropriate

    one for this model. Finally, we prove a stron g type of convergence rath er

    than convergence in distribution or probability. The proof is constructive in

    the sense that it shows explicitly how the random initial conditions blend

    into the noise. W ith arb itrarily large probability, every choice of (x(0), p(0))

    that satisfies #(x (0), p(0)) = Eq + N/ (3 corresponds to a continuous func

    tion () such that the solution of ( 2 .1 2 ) with initial conditions (x(0),p(0))

    approximates the solution of (2.13) with this particular ().

    It is interesting to note that due to pathwise convergence, the result of

    the Theorem holds for all initial distributions for the bath that are absolutely

    continuous with respect to the microcanonical one. Any initial distribution

    on and { p ,} ^ , will induce a probability distribution on the noise

    w(0- This, in turn, induces a probab ility distribu tion on the limiting process

    26

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    39/161

    (t) = lini/v-xxj w(0- Th e limiting noise will generally no t be Gau ssian, and

    will not satisfy a fluctuation-dissipation relation.

    As an example, take P (u ) to be the one sided Lorentz distribution

    P M = 2, ^ > 0 (2.19)7T OL U)

    and p(u>)= 0 otherwise, where a > 0. In a previous pape r [5] we stud ied this

    model both formally and numerically in the case when V(x0) is a double-well

    potential an d th e dynamics is metastab le. In Section 2.3.3 we prove th a t th e

    effective equation for x 0(t) is

    m 0X 0 + V ' ( X 0) + 7 j \ ~ a't- T' X 0( r)dr = ~ ^ t ) , (2.20)

    and (t) is an Ornstein-Uhlenbeck (OU) process at equilibrium with zero

    drift and an exponentially decaying covariance with rate a.

    2 .3 .3 Pr oof o f the theorem

    Although the appropriate invariant measure for the Hamiltonian dynamics

    (2 .1 2 ) is the microcano nical one, it is technically simpler to use the cano ni

    cal invariant measure. Under the canonical ensemble, initial conditions are

    indepen dent G aussian random variables. The ir distribution is given by the

    27

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    40/161

    Gibbs measure

    Xi = N ( x 0,N /pyi ) ,

    ( 2 .21 )

    Pi = N(0,rrii /f3) , i = l . . . N ,

    where N( fi, 1 / a 2) denotes the G aussian distribution with mean n and vari

    ance a 2. Most auth ors studying such models consider only the canonical

    ensemble [29, 32, 38, 40, 41, 42, 57, 58, 74]. We will first prove the Theo

    rem for the case of canonical initial conditions and then prove that the same

    result holds for the microcanonical one.

    Canonical initial conditions

    We begin with a few prelimina ry calculations. Using either variation of pa

    rameters or the Laplace transform, (2.12) can be solved for x ^ .. . x N. Sub

    stituting into the equation for x 0(t) and integrating by parts, the equation

    for Xq (t) can be written as

    (2 .22)

    where

    N

    (2.23)

    and /v(t) is given by

    (2.24)

    28

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    41/161

    Note th a t th e bath in itia l co nd itions appear only in ^ . For th is reason we

    will refer to /v(t) as a ran dom noise term. Changing variables into dimen-

    sionless, centered coordinates

    IQ'Y

    hi = V77 ~~

    9i = ~ P*(0)>V rr i i

    (2.25)

    the noise term r(t) can be written as

    I N= J ^ 2 (hi cosujit + gi sin LUit). (2.26)

    1=1

    The T heorem is proved in three steps. The first considers convergence

    of the memory kernel i?/v(f). For fixed time t, the random variable cosu t is

    bo unded. Hence, th e law of large nu mbers implies th a t Rn(I) converges to

    its average for almost all u>

    N

    lim f?Ar(t) = lim -7- cosujt =7 ECi)cosu t R( t), (2.27)N-+O O N OO N /

    1= 1

    In Lemma 1 we prove that Rn(I) converges to R(t) strongly. The second

    step considers the noise. In Lem ma 2 we prove th a t jv() converges strongly

    to a limiting Gaussian process, (), with zero mean and covariance R( t) .

    The final step considers the position and momentum of the distinguished

    pa rticle. Stron g convergence of x ff(t) to x 0(t) and of (t) to p0(t) is proved

    in Theorem 1. All steps consider a finite time interval 0 < t < T, T

  • 7/22/2019 Stochastic Dynamics

    42/161

    Initial conditions of the b ath have a Gibbs distribution. Hence, hi and i, and the definition of R(t),

    (2.27), yields (2.28). Th e strong convergence is also an imm ediate conse

    quence of Birkoffs ergodic theorem in L 2 [73].

    A similar calculation shows that for large N,

    sup E J i ? ( t ) - i ? ( t ) |4 < I ^ . (2.30)0 < t < T (V

    This is used in the proof of Theorem 1.

    30

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    43/161

    L e m m a 2 : For all T < oo, there exists a constant C independent of T , N

    and p(u>) such that,

    C T 2

    sup E j ( f ) - 6 v ( * ) | 2 ^ (231)0 < t < T A

    where (f) is a Gaussian process with zero mean and covariance R( t) . The

    rate of convergence, I, is determined by the tail of the distribution ofp(u>) as

    detailed before.

    Proof: For fixed N and t, jv() is a linear combination of independent Gaus

    sian rando m variables and is therefore Gaussian itself. Hence, () is also

    Gaussian for fixed t. As a result, all marginals of the form ( (f i) ,. .. ,(&))

    are Gaussian vectors, which implies th a t () is a Gaussian process. It is

    interesting to note that (2.31) implies that jv() converges in distribution

    to (). The ra te of this weaker convergence is always N ~ 1/2 and does not

    depend on the tail of p(u>).

    In order to prove strong convergence, we must identify how the random

    initial conditions hi and gi blend into the limiting process. In other words,

    both th e sequence of processes i( t) , 2^) , and the limiting process (t)

    have to be defined on the same pro bability space. To achieve this we write

    31

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    44/161

    the limiting process as

    m = \ / 7 / p l/ 2(ui)[costutdhu+ sin u>tdgj\ , (2.32)Jo

    where and gw are independent copies of the standard Brownian motion

    (BM) on [0, oo). To see this is indeed (), we use the Ito iso metry to show

    that the average of this process is zero and the covariance is

    We now wish to define the initial conditions, hi and gx,on the same probabil

    ity space of hMand gw. The difficulty is in finding a scaling tha t relates th e

    random variables hi and ^ to parts of h^ and g^. We begin with a few defini

    tions. Let a be a permutation of {1,.. . , N } that rearranges the frequencies

    in increasing order, i.e., ui^i) ). For short

    hand we denote p x = ipf(u>a(i)). We take the maps to b e increasing in the

    sense that

  • 7/22/2019 Stochastic Dynamics

    45/161

    Th e rand om variables are defined similarly using gw instead of h^. The

    motivation for this definition is that for any family of such maps, hi and gl

    are independen t random v ariables with normal distribution. This is easily

    verified using the Ito isometries. We now look for the scaling maps th a t

    would yield strong convergence, and show that for all 0 < t < T,

    for some I > 0. The constant C depends only on the density function p(u>).

    A necessary condition is to have convergence at t 0. Th is simple case will

    provide us with th e requ ired scaling. Substi tu ting in the representation s for

    at and ,

    where x a {u) denotes the indicator function of the set A evaluated at u>.

    Using the Ito isometries yields

    (2.35)

    Eo.0,16v(0) - mi2=27 - 27 L V Pi-Pi-1- (2-37)

    It is clear that we want to find a map ip such that

    (2.38)

    33

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    46/161

    Squaring this expression and using Jense ns inequality yields an up per bound:

    ^(^Ev^wn) < 1. (2.39)

    Equality holds if all the terms Pj P,_i are equal, i.e., Pj = i / N . Hence,

    strong convergence is obtained if

    ( d ' K w ) = P - 1 ( - 0 = P - \ ( 2 . 40)

    rpr l K( i) = V N / p l,2{ u) dh u, (2.41)

    Jp- \

    where P 1denotes the inverse of the distribution function Pw{u < z). There

    fore, initial conditions take the simple representation

    rp,r

    and similarly for gt .

    For t > 0, we have, substituting in the representations for f, N, hi an d

    9i

    Eo.c m ) - ZN{t)I2 = 27 (1 - S N) (2.42)

    with

    x

    5 iv = E 0iW / p 1 2(x) cos(xt)dhx + / p 1 2(x) sin (xt)dgxL v R Jr

    N ( rpi X P E 1V ] / p 1/2(x)dhx cos(uaii)t) + / p 1/2(x )dg x sm(u

  • 7/22/2019 Stochastic Dynamics

    47/161

    Note th a t averaging with respec t to in itia l co nd itions implies averaging over

    the two BM hx and gx, while averaging with respect to u implies averaging

    over mu ltiple samples of Clearly, we need to show th a t Sn > 1 .

    Using the Ito isometries yields

    N f P r 1SV = y ~ ' E u, / p( x) cos( (x - u a(i))t)dx. (2.44)

    i = i J p r - \

    Bounding the cosine by 1 yields an up per bo und , Sn < 1. To get a lower

    boun d, we show th a t for most i, (x oJa{i)) is small, and the cosine is almost

    one. The reason why (x u>a(i)) is not small for all i is due to the tail of

    the density p(uj). We therefore break the sum into two parts: up to kN and

    above. For i = 1 . . . k^ , we use co sx > 1 x 2/2. Let u>k be such that

    Pw[u> < u>k] kN/ N . For i = (kn + 1) N , we bou nd the cosine by 1.

    Hence,

    fcjv pP~l N rW 1I , p{x ) [ ^ ~ i.x - ^ a ( i ) f T2/ ^ ] d x - / p{x)dx

    1 = 1 J p i - l i = k N + l ^ p i - \

    (2.45)

    p - 1

    Noting th a t f p p(x)dx 1 /N yields

    \SN - 1| < ~ P^ E^ X ~ u (i))2(lx+ 2 ( l (2-46)

    Th e main difficulty is in evaluating the exp ecta tion E w(x tu^j))2. Noting

    35

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    48/161

    that for x G [P ^ ),Pi *]

    K , ( x ~ ^ ) 2 < 2 ( Pr ' - ^ - l ) 2 + 2EW( P - 1 - coa { i ) ) \ (2.47)

    we proceed in several steps:

    L e m m a 2 .1 : For all T < oo, there exists a constant C such that for all

    i < kf j, and except fo r a possible fini te number of indices,

    We begin with the second term of (2.47), which has a form similar to the

    setu p of the Kolmogorov-Smirnov statistics. Bre iman [10] gives a proof of

    the following Theorem:

    Let U \ , . . . ,Un denote n independent samples from a random variable, un i

    forml y dis tributed on [0,1]. Let P CT(i ), . . . , Ua(n) denote the same set of sam

    ples arranged in increasing order. Then, the random variable

    L > n = Vnm ax \Uaii) - i / n \ ,i < n

    has a limiting distribution with finite variance.

    This implies that,

    (2.48)

    Proof:

    m ax E ( U. (2.49)

    36

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    49/161

    Noting tha t P(u>lT ')) is uniformally distributed [0,1], i / n = P(Pi : )

    P (v) > 0 for all u>< u>k and for k 1 , . . . , we have

    e u p - 1 - u a{i))2 0j)(w ) ) ]2

    Since p{u ) is strictly positive in its support, the minimum is bounded and

    E ( / r UVW)2 < (2.51)

    If there are sp ectral gaps then (2.51) fails whenever two subsequent samples

    u>cr(i) and u>tr(i+i) are in different conectivity classes of the s upp ort ofp(u>). In

    the latter case, which may happen in a finite set of indices, we have

    (jEu)( i^ _1 - u a{i))2 < + largest gap. (2.52)

    Substituting (2.51) and (2.52) into (2.46) yields,

    +T ?

  • 7/22/2019 Stochastic Dynamics

    50/161

    for some I2> 0 that depend on l\.

    Proof:

    We prove Lemma 2.2 for different types of distributions p(u), and show

    that the rate of convergence of t) to (t) depends on the tail of p(uj).

    Suppose that the tail of the density function p(oj) decreases polynomially,

    i.e.,

    C u~ q < p{uj) < D u~ q, (2.55)

    for some q > 1 and C, D > 0. We use C and D to denote generic constants

    whose values may vary between expressions. Den ote l - k N/ N = N ~ l, I > 0.

    Since p{u>) is strictly positive, then for large enough N we have,

    Cu k q )COk] = / p(u})du> < D qdu = ----- 7^l~q.

    J^k Jwk Q~ k

    (2.57)

    Using the opposite bound yields

    C N h/(q-i) < U k

  • 7/22/2019 Stochastic Dynamics

    51/161

    Hence, for i < k^,

    . r 'N\j = ( ' p(u>)dw > C(P- ' - P ~\ y i > C ( p r ' - p r_ \) N - W i -'), (2.59)v Jpr~\

    which implies that

    |P f1 - < N hqn* - l)- 1. (2.60)

    Similar evaluations can be done with different tails. For instance, if p(u>)

    decreases like e~ax, e~ax or has finite sup port th an th e rate is the sup over

    all qs, I2= h 1 .

    Combining the above bounds we find that with a w~ g tail, the optimal

    rate is obtained with

    With an exponential tail or better,

    l = i . (2.62)

    This concludes the proof of Lemma 2.

    Figure 2.1 compares the random noise at() for N = 2000 obtained by

    drawing random initial conditions hi and gi, with (i) obtained by sampling

    39

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    52/161

    two BMs hw and th a t satisfy the con straints (2.41). The sample is pre

    pared in th e following way: we first draw th e frequencies, {uJi}i=i,..N, and

    then initial conditions hi and from the canonical distribution. For the fre

    quencies we use the one sided exponential distribution p(u>) = e~w for u > 0 ,

    and p(u) = 0 otherwise. We can then comp ute /v(t). The exponen tial distri

    bu tion is convenient since th e an ti-de riva tiv e of e~u^2cosuit, which appears

    in (2.41), is easily calcula ted. We the n app rox imate two BMs hu and that

    satisfy the constraints (2.41) by a linear interpolation between the points

    P _1(i/./V), i = 0 . . . N . This partition becomes finer as N >oo. Hence, the

    approx imation converges to BM by the invariance principle. The limiting

    noise (t) is then obtained through its representation given by (2.32). Since

    the approximation is piece-wise linear, the integration w.r.t. dh^ and dg^

    can be preformed analytically.

    Theorem 1 : For all T < oo, there exists a constant C(T), independent of

    N such that,

    sup E0,w0< t < T

    |xK(t ) - X ( t ) I2+ 1 ^ ( 0 - F ( ( )|2] < ^ 2 , (2.63)

    where the rate of convergence, I, is the same as in Lemma 2.

    This implies that Xq (t), the solution of the full equations of motion (2.12),

    40

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    53/161

    2.5

    0.5

    - 0.5

    - 1.520

    t

    Figure 2.1: Com parison of the rando m noise, jv(f) (solid line), and the

    limiting one (t) (dashed line). The frequencies are drawn from a one sided

    exponential distribution and N = 2000. The ba th initial conditions are

    canonical.

    converges strongly to X ( t ) , the solution of the limiting stochastic equation

    (2.13).

    Proof: Compare the solutions of the limiting equation for (X ,P):

    X = P / m 0

    t (2 -64 )

    P = - f - V ( X ) - ^ f R ( t - T ) P ( T ) d T + - ; ( ( t ) ,m0 mlQJo v/3m0

    41

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    54/161

    (2.65)

    with the one obtained for finite N:

    %o = P o / m o

    Po = - V ' ( x Z ) - f R N(t - r )p g ( r )d r + N(t).m 0 rag J0 V/5m0

    For shorthand, we drop the subscript zero from mo,XQ and Pq for the rest

    of the section. We wish to show th at

    sup E 0,a, [X (t) - ^ { t ) ] 2+ [P(t) ~ pN(t)]2 oo, one has to

    consider the convergence of R ^ ( t ) , iv(t) and (x q (t ) ,p(t )) . However, since

    jRiv(t) does not depend on initial cond itions, th e only difference between the

    microcanonical ensemble and the canonical one considered in Section 2.3.3 is

    the convergence of the noise, #() Once this is established, convergence of

    Xq (t) follows by the exac t same arg um ent d etailed in the proof of Theorem 1,

    (2.79)

    The normalization Zis obtained by taking /( h i) = 1. S etting /( h i) = e lthl

    yields the characteristic function of the marginal h\

    (2.80)

    46

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    59/161

    Section 2.3.3. In this Section we use the s up er sc rip tsm and c to distinguish

    between micro- and canonical initia l co nd itions or noise.

    In the case of Gaussian initial conditions, %(t) is trivially a Gaussian

    process for any N. Hence, the sequence %(t) converges to a Gaussian pro

    cess, c(f). Th e situa tion is not as simple with th e microcanonica l ensemble.

    We recall the definition of jv, (2.26):

    1 NG W = - j = ^ 2 y / T i ( h ^ c o s u } i t + g ^ s i n u i t) . (2.81)

    ' i = l

    For fixed t, it is easy to see th at the characte ristic function of converge,

    as IV > oo to the sam e charac teristic function in th e can onical case, c().

    This is because contributions of order iV- 1 do not sum up to 0(1) due to

    the N ~ 1/2 prefactor in (t). Hence, () converge in distr ibut ion to the

    same limit c(f). The rat e is prop ortiona l to N ~ ^ 2. Th e same holds for

    any linear com bination of marginals a t different times. This implies tha t the

    whole process $() converges in distrib utio n to (), which is Gaussian. For

    this reason, from now on we can drop the ensemble label from the limiting

    noise ().

    In order to prove strong convergence, we need to map the probability

    space of the initial conditions hand gto the one on which () was defined:

    the two BM hu and gu. The idea is to draw independent initial conditions

    47

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    60/161

    according to the Gibbs distribution and th en scale them to the desired energy.

    The dependence between different variables comes from the scaling factor.

    Define

    K ^ h y r ; g? = r f / r , ( 2 .8 2 )

    where

    = v / 2H / N =

    \ 1 = 1Then, with hc{ and g\ given by (2.41), h' f and ghave the same distribution

    as the microcanonical initial conditions hand g. We can therefore drop

    the tilde notation. Substituting into (2.81) yields

    ( 0 = i& W - (2.84)

    An elementary calculation shows that, in the limit N > oo, the average of

    r te nds to 1 while the variance is 0 (N ~ 1). Hence, $() convergence to (t)

    strongly. The dependence of the ra te on N is the same as in Sections 2.3.3.

    To sum up, we have proved the following Theorem:

    Theorem 2: For all T < oo, there exists a constant C(T), independent of

    48

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    61/161

    N such that,

    sup E 0)U0 < t < T

    X, \

    7 t = 7 If the product 'y(N,u>)pn{uj) converge in Li(ui) to a limiting

    function p(u>), then the cosine transform ofp(uj) is bounded and continuous.

    We denote R(t) = J p(u>) cosuitdu). Note tha t p(u/) may not be normalized.

    Assuming that 7 (N,lo)pn(u>) converge also in L 2 (w), and that 7 ( N , u ) < C N

    for some constant C > 0, it is easily shown that R E (t) converges to R(t) in

    L 2 (u>). The proof is the same as in Lemma 1, however, the convergence rate

    may be smaller. Once strong convergence of the covariance function is estab

    lished, stron g convergence of the noise w(t ) and the trajectory ( x ^(t),p$ (t))

    follows. For instance , S tu ar t and W arren [58] suggest the following example

    Pn {u ) = t ^ X [ o , j v ] ( w )

    N (2 .86 )

    7(NM=7r a z + ur

    where a, 7 > 0 and 0 < a

  • 7/22/2019 Stochastic Dynamics

    63/161

    (2.19).

    If we remove the requirement for L\ (u>) convergence, then the model ad

    mits a m uch larger variety of limiting processes. For instance, if y (N , cu)p^r(u)

    is not uniformly bounded, then the covariance function R(t) is not necessar

    ily bounded and continuous. Taking p(u>) = N a / n / ( N 2a -f- u 2) and 7j = N a,

    yields, for any 0 < a < 1/2, i?jv(t) = N ae~N

  • 7/22/2019 Stochastic Dynamics

    64/161

    fortunately, it is often these long time scales, or even the asy mp totic behavior

    of the pa rticle th a t are interesting. An example of this is a case in which

    the external potential V(x0) is metastable [5]. However, it seems reasonable

    that the process (x q,Pq) should converge to the limiting one weakly or in

    distribution uniformly for all times T > 0. This is because the invariant

    measure of the finite system should converge to the invariant measure of the

    limiting one. This problem is beyond the scope of the present m anuscript

    and will be trea ted in a later publication. A proof for a specific case similar

    to (2.86) was given in [42].

    52

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    65/161

    2.4 Testing Transition State Theory on Kac-

    Zwanzig Model

    2.4.1 Introduction

    Deterministic dynamical systems, in particular, Hamiltonian systems, often

    display very complicated chaotic behavior when the number of degrees of

    freedom in the systems is large. Amid the complexity of individual traje c

    tories, it is sometimes the case that most trajectories remain confined for

    very long periods of time in well separated regions of phase-space and only

    switch from one region to ano ther occasionally. The confinement is due to

    the presence of dynam ical bottlenecks between these regions. The system is

    then said to display metastability, and the regions in which the trajectories

    remain confined are referred to as me tastable sets. Example of systems dis

    playing m eta stabil ity abound in natu re , with exam ples ar ising from physics

    (phase transitions), chemistry (chemical reactions, conformation changes of

    bio-molecules), biology (switches in po pula tion dynamics), and many others.

    In these systems, it is worth verifying whether the dynamics can be approx

    imated by a Markov chain over the state-space of the metastable sets with

    approp riate rate constants. The m ain question of interest then becomes the

    53

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    66/161

    determ ination of these rate con stants. This question is nontrivial because it

    amounts to understanding the pathways by which the transitions between

    the m etastab le sets occur. These pathways are usually non-trivial and com

    plicated.

    One of the earliest attempts to determine transition pathways and rate

    con stants is tran sition s tate theory (TST ) [22, 35, 75]. Under the sole as

    sumption that the dynamics of the system is ergodic with respect to some

    known equilibrium distribution, TST gives the exact average frequency at

    which trajectories cross a given hypersurface or hyperplane which separates

    two m etasta ble sets of interest. This average frequency can be used as a

    first approximation for the frequency of transition between the metastable

    sets. Unfortunately, it was recognized early on tha t this approxim ation can

    be qu ite po or , for not every crossing of th e dividing surface corresponds to

    a transition between the meta stable sets. Indeed, the trajectorie s can cross

    the dividing surface many times in the course of one transition. As a result,

    the TST prediction for the frequency of transition always overestimate the

    actua l frequency, sometimes grossly so. One way to minimize this problem

    is to use the freedom in the choice of dividing surface. The best prediction

    for the frequency from TST is the one corresponding to the dividing surface

    54

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    67/161

    with minimum crossing frequency. This idea is at th e core of the so-called

    variational transition state theory (VTST) [35, 51, 65, 70], which aims at

    identifying the dividing surface with minimum crossing rate which is the

    lift-up in phase space of a surface defined in configuration space.

    Unfortunately, VTST (just like TST) is an uncontrolled approximation,

    for it only provides an upper bound on the transition frequency between

    the m etastab le sets. In general, one does not know how sharp this b oun d is.

    Other assumptions beyond TST and VTST are usually difficult to assess too.

    Are successive transitions between the metastable sets well approximated by

    Poisson events (i.e. statistically indepen dent and with exponentially dis

    tributed waiting times) as required for the approximation of the dynamics

    by a Markov chain to hold? How does th is pro perty dep en ds on th e definition

    of the metastable sets? Etc.

    In this section we study a benchmark problem, which is simple enough

    so that many of the assumptions and approximations underlying TST and

    VT ST can be examined. Bu t the m odel is also complex enough to display a

    wide variety of phenomena common to many dynamical systems exhibiting

    me tastability. The problem we consider is a variant of a model originally pro

    posed by Ford, Kac, and Mazur [24, 25] and Zwanzig [76]. It was revisited

    55

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    68/161

    in the context of transition rates in [13, 30, 52, 53, 54, 55] and more re

    cently, from a more analytical point of view in [11, 29, 32, 36, 40, 41, 42, 58].

    The Kac-Zwanzig model is a Hamiltonian system describing the evolution

    of a distinguished particle moving in a double-well external potential and

    coupled to a bath of N free particles via linear springs. The dynamics in

    this model is deterministic. However with app rop riate choice of the pa ram e

    ters, the evolution of the distinguished particle can be captured by a closed

    stochastic differential equation in the limit of infinite bath, N >oo. W ith

    the double-well external potential, the model displays metastability over two

    sets (bistability). Tran sition rate con stants between these sets can be com

    puted exac tly from th e effective stocha stic dynamics in th e lim it N oo.

    The values for these exact transition rate constants can then be compared

    to the predictions of TST and V TST. This comparison is the m ain objective

    of this paper. In particular , we will show tha t the application of TS T w ith a

    naive (but natural) choice of dividing surface based only on the position of

    the distinguished particle leads to a wrong prediction for the transition rate

    constants. This is because the naive dividing surface is crossed many times

    in the course of each transition between the two me tastable sets. However,

    we will also show that if one optimizes over the dividing surface following

    56

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    69/161

    VT ST, these many spurious crossings can be eliminated completely. Hence,

    the correct transition rate constants for the model can be computed within

    VTST. The optimal dividing surface which allows one to do so is then a

    surface whose normal spans all the configurational degrees of freedom in the

    system and n ot only the one associated with the distinguished particle. We

    shall try to explain why this is the case and when a similar success of VTST

    can be expected in other more realistic systems.

    In [52], Poliak et al. app rox imate a generalized Langevin equation , which

    has the same form of the limiting equation in our case, by the Hamiltonian

    dynamics of the Kac-Zwanzig model, and then use TST for obtaining the

    escape rates. In [54, 55], they ob tain th e same rate s from the limiting equa

    tion by a generalization of Kram ers meth od [28]. Although our results are

    similar, th e po int of view is different. Here, the Kac-Zwanzig model is used

    as a platform for analyzing the predictions of TST and VTST and testing

    the und erlying assumptions of these theories. Our results are also all de

    rived from basic principles, and rely on the only (uncontrolled) assumption

    of ergodicity.

    The rem ainder of this section is organ ized as follows. In Section 2.4.2

    we present the model and derive some of its basic properties. Section 2.4.4

    57

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    70/161

    details numerical experiments, discusses some of the properties of the sys

    tem conjectured in Section 2.4.2 and presents results for the transition rate.

    In Section 2.4.5 we derive the effective stochastic differential equation that

    describes the dynamics of the distinguished particle in the limit TV > oo.

    We then calculate trans ition ra tes for the limiting dynamics. This is done by

    spectral decomposition of the backward operator of the stochastic equations.

    Eigenvalues and eigenfunctions are obtaine d using matched asym ptotics. In

    Section 2.4.6 we develop TST and VTST. We find predictions for the tran

    sition rates from these theories. Finally, in Section 2.4.8 we summarize our

    findings and discuss possible generalizations.

    2.4 .2 Th e m odel

    In order to simplify the discussion, for the rest of this section we will restrict

    our analysis to the following particular choice of probabilty density for the

    frequencies

    p(u) = * 1 y r ~ 2 if w > o

    tt 0,2 + ^ 2 (2.87)

    0 otherwise

    where o,* > 0 is a parameter playing the role of a characteristic frequency.

    Unless sta ted otherwise, we will take o,* = 1. Notice th a t all the moments

    58

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    71/161

    of (2.87) are infinite, i.e. u>* ^ E ul = oo, where E denotes expectation with

    respect to (2.87). Hence, using the no tation of the previous sections

    R(t) = qEfcoswf] = 7 e- ^. (2.88)

    In section 2.4.5, we proved that if we choose initial conditions (x(0), p(0))

    so that

    E = N/P, (2.89)

    for some P > 0 playing the role of an inverse temperature, then in the

    limit N >oo, the evolution of the distinguished particle is described by the

    stochastic equations

    f t iX o + V ' { x o ) + 7 j e~(t~T)x0(T)dr= (2-90)

    where (t) is a G aussian process with zero mean a nd covariance function

    7 e- ^. Hence, the noise is an Ornstein-Uhlenbeck process at equilibrium

    which solves the stochastic differential equation

    d= - d t + x/27 dWt , C(O) = A7(0,1 ), (2.91)

    where Wt isa stan da rd Brow nian motion. Using (2.90) and (2.91), the lim-

    59

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    72/161

    iting equation (2.90) can be written as [29, 58]:

    xq = Po

    * Po = s / i s - V \ x q) (2.92)

    s = - s - ^ p Q+y/W~mt

    Hence, with this particu lar choice of model parameters th e limiting dynamics

    satisfy a stochastic differential equation.

    In Section 2.4.5 we will find the following form of (2.92) useful:

    Ldt

    ( \Xq

    Po

    \ S /

    = -K X 7H (x 0, po, s) + x / ^ W t.

    Here H ( x q , P q , s ) = V(x0)+ \ p l+ | s 2,

    / \0

    (2.93)

    (2.94)

    V 1 /

    60

    produced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    73/161

    and K= K s+ K A, with

    /0 0 0

    K s = aaT = 0 0 0

    V0 0 1

    /(2.95)

    0 - 1 0

    K A = 1 0 -V 7

    2.4.3 M etastabi l ity in Kac-Zwanzig

    In order to study metastability, the distinguished particle is placed in a

    double-well poten tial. Unless state d o ther wise, for the rest of this Section

    we take

    When the inverse temperature (3 and the size of the bath N are large

    enough and satisfy 1

  • 7/22/2019 Stochastic Dynamics

    74/161

    instance be taken as

    S - ( N , P , 5 ) = {(x, p) :H(x , p) = N/(3,x0 < 0, and H 0{x0,po) < 5}(2.97)

    S+(N,0 ,6) = {(x,p) :H (x, p) = N/(3,x0 > 0, and H 0{ x0,Po) oo. (2.98)

    Here (Xe is the microcanonical distribution on H(x , p) = E,

    (2.99)

    62

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    75/161

    element (Lebesgue measure) on H(x , p) = E, and 2 ( E ) is a normalization

    constan t. The ergodic assumption can not be proved rigorously for the po

    tential in (2.96), but it will be corroborated by the numerical experiments

    presen ted in section 2.4.41.

    The key property implying metastability over the sets in (2.97) is that

    for every 5 E (0,1], and similarly for the integral over S -( N , (3,6). Note that

    the o rder in which the limits are taken m atters. The validity of (2.100) can

    be checked by direct ca lculation. Indeed, performing first th e in tegrat ion

    over x \ , . . . , xn and p i , . . . , p we have

    where Z 0(N/ /3) = f Ho

  • 7/22/2019 Stochastic Dynamics

    76/161

    stant. In the limit as Noo, this is

    lim [ dtiE=N/0(X;P) = Z q 1 [ e~/3Hodx0dp0 (2 .1 0 2 )N ^ J S + ( N , 0 , 8 ) J H 0< 6

    where Zq = f R2 e~^Hodx0dp0. The B oltzmann-Gibbs probability density func

    tion ZQ1e~/3H is in fact the marginal density for the position and momentum

    of the distinguished particle in the limit of infinite bath, N oo. For every

    5G (0 , 1], the minimum at (x0,po) = (1 , 0 ) is the only minimum that belongs

    to the domain where H 0 < 8. Therefore, (2.100) follows from (2 .1 0 2 ) by

    simple evaluation of this integral by Laplace method. It is im po rtan t to note

    that S are cylindrical sets in R27V+ 2 and n ot small neighborhoods around of

    the energy minima. Due to the high dimensionality of the model, the mass

    of the equilibrium measure is not concentrated in a small volume of phase

    space since the Laplace approximation does not hold when N > (3.

    Equation (2.100) implies that, when 1 -C j3 N , any generic trajectory

    solution of (2.2) spends most of its time in either S+(N, (3,8) or S- (N , (3,8).

    However, under the ergodicity assumption, this trajectory must switch be

    tween S+(N,(3,8) and S-(N,(3,8) infinitely often. W hat are the rate con

    stants of these transitions ? How do they depend on 81 Are they statistically

    independent, with transitio n events Poisson distributed? In othe r words, can

    the dynamics in (2.2) be reduced to a Markov process over S+(N,(3,8) and

    64

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    77/161

    S-(N, /3 ,5) for some suitable choice of 6? These are the questions which we

    shall investigate in the remainder of this paper, first via a series of numer

    ical experiments with (2.2) (section 2.4.4), then using the limiting equation

    in (2.92) (section 2.4.5), and finally within TST and VTST (section 2.4.6).

    2.4 .4 N um erical experim ents

    In this section, we perform a series of numerical experiments with (2.2) to

    investigate when the dynamics can be app roxim ated by a Markov process over

    the two m etastab le sets in (2.97). The questions we are especially interested

    in are:

    1 . W hat are the rate of the transition? How do they depend on the

    para m ete r 7 (interaction strength with the bath) in the model? How

    do they depend on the choice of 5 in (2.97)?

    2. Are successive transitions to a good approximation statistically inde

    pendent? Are th e transit io n tim es in th e sets (2.97) Poisson dis tr ibuted

    with intensity equal to the rate of transition? How do these properties

    depend on 5?

    65

    roduced with permission of the copyright owner. Further reproduction prohibited without permission.

  • 7/22/2019 Stochastic Dynamics

    78/161

    The second question is especially important since it determines when the

    dynamics in (2.2) can be approximated by a Markov process, and how the

    metastable sets have to be chosen in this case to get the correct transition

    rate constants to use in the chain.

    For these experiments, we will take N = 1000 and /? = 7. We will

    also consider two different values of 7: 7 = 1 and 7 = 10. In (2.2), the

    equations of motion of the bath are integrated explicitly, while the equations

    of motion describing the distinguished particle are integrated numerically

    using the Verlet algo rithm [71]. Each time step is made reversible by a

    Tro tter expansion of the time evolution ope rator [27, 67]. In all the results

    reported below we use the double-well potential (2.96), but the integration

    scheme was also checked using the harmonic potential, V(xq) ^Xq, for

    which (2.2) can be integra ted analytically. Initial conditions are chosen once

    from the microcanonical invariant distribution on the energy shell E = N/ /?.

    The integration is performed up to time T = 2 106. T he param eters N , T

    and the step size are chosen so that further increase (in Nand T) or decrease

    (in step size) does not change the average rates considerably. Transition ra tes

    are obtained by counting the number of times the trajectory switches between

    S + and S - . Table 1 deta ils resu lts obtained for

  • 7/22/2019 Stochastic Dynamics

    79/161

    7 5 = 1 (5 ==0 .1

    1 4.1 1 0 4 3.7 1 0 ~ 4

    1 0 3.5 1 0 ~ 4 1 .1 1 0 ~ 4

    Table 2.1: Com parison between tran sition ra tes with two different candida te

    metastable sets.

    5 = 0.1 is arb itrary . Any choice of 0.05 <

  • 7/22/2019 Stochastic Dynamics

    80/161

    between successive transi tio ns, we ex pect th a t it will have an expo nen tia l

    distribution,

    P[tS> s ] = e ~ k s , (2.103)

    for some rate constant k,which is also the average transitio n rate. Figure 2.2

    depicts the distribution of the waiting times between transitions, P [ts > s],

    on a semi-log plot for the case 7 = 10. For 5 = 1 , the graph is not linear

    near the origin. However, for 5 = 0.1, the linear fit is very good, indicating

    th a t transitio ns events have