164
For other titles published in this series, go to www.springer.com/series/7219 Surveys and Tutorials in the Applied Mathematical Sciences Series Editors S.S. Antman, J.E. Marsden, L. Sirovich Volume 5

Surveys and Tutorials in the Applied Mathematical Sciences

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Surveys and Tutorials in the Applied Mathematical Sciences

For other titles published in this series, go to www.springer.com/series/7219

Surveys and Tutorials in the Applied Mathematical Sciences

Series Editors

S.S. Antman, J.E. Marsden, L. Sirovich

Volume 5

Page 2: Surveys and Tutorials in the Applied Mathematical Sciences

Surveys and Tutorials in the Applied Mathematical SciencesVolume 5

EditorsS.S. Antman, J.E. Marsden, L. Sirovich

Mathematics is becoming increasingly interdisciplinary and developing stronger interactions withfields such as biology, the physical sciences, and engineering. The rapid pace and development ofthe research frontiers has raised the need for new kinds of publications: short, up-to-date, readabletutorials and surveys on topics covering the breadth of the applied mathematical sciences. Thevolumes in this series are written in a style accessible to researchers, professionals, and graduatestudents in the sciences and engineering. They can serve as introductions to recent and emergingsubject areas and as advanced teaching aids at universities. In particular, this series provides anoutlet for material less formally presented and more anticipatory of needs than finished texts ormonographs, yet of immediate interest because of the novelty of their treatments of applications,or of the mathematics being developed in the context of exciting applications. The series will oftenserve as an intermediate stage of publication of materials which, through exposure here, will befurther developed and refined to appear later in one of Springer’s more formal series in appliedmathematics.

Page 3: Surveys and Tutorials in the Applied Mathematical Sciences

Jack Xin

An Introduction to Frontsin Random Media

Page 4: Surveys and Tutorials in the Applied Mathematical Sciences

© Springer Science+Business Media, LLC 2009

subject to proprietary rights. Printed on acid-free paper

permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York,

connection with any form of information storage and retrieval, electronic adaptation, computer

The use in this publication of trade names, trademarks, service marks, and similar terms, even if they

NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in

software, or by similar or dissimilar methodology now known or hereafter developed is forbidden.

are not identified as such, is not to be taken as an expression of opinion as to whether or not they are

All rights reserved. This work may not be translated or copied in whole or in part without the written

Jack XinDepartment of MathematicsUniversity of California, Irvine

6Irvine, CA 92 [email protected]

ISBN 978-0-387-87682-5 e-ISBN 978-0-387-87683-2 DOI 10.1007/978-0-387-87683-2

Mathematics Subject Classification (2000): 60H15, 60H30, 76M30, 76M50, 76M45

Library of Congress Control Number: 2009926482

Springer is part of Springer Science+Business Media (www.springer.com)

Editors:S.S. AntmanDepartment of MathematicsandInstitute for Physical Science

and TechnologyUniversity of MarylandCollege ParkMD [email protected]

J.E. MarsdenControl and Dynamical

System, 107-81California Institute

of TechnologyPasadena, CA [email protected]

L. SirovichLaboratory of Applied

MathemaicsDepartment of

Bio-Mathematical SciencesMount Sinai School of MedicineNew York, NY [email protected]

Springer Dordrecht Heidelberg London New York

Page 5: Surveys and Tutorials in the Applied Mathematical Sciences

Dedicated with love to my wife, Lily,my sons, Spencer and Grant,and my parents, Dingding and Ningyuan

Page 6: Surveys and Tutorials in the Applied Mathematical Sciences

Preface

This book aims to give a user-friendly tutorial of an interdisciplinary research topic(fronts or interfaces in random media) to senior undergraduates and beginning grad-uate students with basic knowledge of partial differential equations (PDE) and prob-ability. The approach taken is semiformal, using elementary methods to introduceideas and motivate results as much as possible, then outlining how to pursue rigor-ous theorems, with details to be found in the references section.

Since the topic concerns both differential equations and probability, and proba-bility is traditionally a quite technical subject with a heavy measure-theoretic com-ponent, the book strives to develop a simplistic approach so that students can graspthe essentials of fronts and random media and their applications in a self-containedtutorial.

The book introduces three fundamental PDEs (the Burgers equation, Hamilton–Jacobi equations, and reaction–diffusion equations), analysis of their formulas andfront solutions, and related stochastic processes. It builds up tools gradually, so thatstudents are brought to the frontiers of research at a steady pace.

A moderate number of exercises are provided to consolidate the concepts andideas. The main methods are representation formulas of solutions, Laplace meth-ods, homogenization, ergodic theory, central limit theorems, large-deviation princi-ples, variational principles, maximum principles, and Harnack inequalities, amongothers. These methods are normally covered in separate books on either differentialequations or probability. It is my hope that this tutorial will help to illustrate how tocombine these tools in solving concrete problems.

The three basic equations go from their constant-coefficient forms, well studiedin graduate textbooks, to their full stochastic glory with space–time-dependent ran-dom coefficients. However, they are all connected to Hamilton–Jacobi equations.The reaction–diffusion equations are classified. The KPP (Kolmogorov–Petrovsky–Piskunov) fronts are discussed in detail because of their connections with the Hamil-tonian dynamics in classical mechanics and their elegant analysis. The recent math-ematical advance in solving a long-standing turbulent combustion problem is pre-sented for KPP fronts. The non-KPP reaction–diffusion fronts are associated witha Hamiltonian resembling that in special relativistic mechanics. The mechanical

vii

Page 7: Surveys and Tutorials in the Applied Mathematical Sciences

viii Preface

connections of reaction–diffusion fronts and KPP solution methods in the spirit ofLagrangian and Eulerian perspectives are explored.

The scope of the book goes from exact solutions of scalar deterministic PDEs(Burgers, Hamilton–Jacobi, reaction–diffusion equations) to the asymptotic solu-tions of their stochastic counterparts. The reader will come to appreciate new ran-dom phenomena step by step, and learn that exact solutions are harder and harder tocome by, while asymptotic ones are more accessible.

The first chapter of the book discusses fronts in homogeneous media, and thesecond chapter is on fronts in periodic media. These two chapters serve as an intro-duction to differential equations, their front solutions and representation formulas,and homogenization methods.

The last three chapters introduce stochastic equations, representation formu-las and asymptotics, stochastic homogenization, variational methods, and large-deviation methods for analyzing random fronts.

The first three chapters are adaptations of the author’s 2000 SIAM review articlewith the inclusion of new results. The remaining two chapters are based on recentresults on stochastic homogenization of Hamilton–Jacobi equations and KPP frontsin random media.

Acknowledgments

I would like to thank George Papanicolaou for his encouragement and interest in mywork on random fronts over the years. I am grateful to James Nolen and Janek Wehrfor many helpful conversations during the preparation of the book. I am thankful forthe support of the National Science Foundation and the constructive comments ofthe anonymous referees.

Jack XinIrvine, California

February, 2009

Page 8: Surveys and Tutorials in the Applied Mathematical Sciences

Contents

1 Fronts in Homogeneous Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Traveling Fronts of Burgers and Hamilton–Jacobi Equations . . . . . . 51.2 Traveling Fronts of Reaction–Diffusion Equations . . . . . . . . . . . . . . . 71.3 Variational Principles of Front Speeds . . . . . . . . . . . . . . . . . . . . . . . . . 111.4 Random Variables and Stochastic Processes . . . . . . . . . . . . . . . . . . . . 131.5 Noisy Burgers Fronts and the Central Limit Theorem . . . . . . . . . . . . 181.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2 Fronts in Periodic Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.1 Periodic Media and Homogenization . . . . . . . . . . . . . . . . . . . . . . . . . . 232.2 Reaction–Diffusion Traveling Fronts in Periodic Media . . . . . . . . . . . 252.3 Existence of Traveling Waves and Front Propagation . . . . . . . . . . . . . 282.4 KPP Fronts and Periodic Homogenization of HJ Equations . . . . . . . . 352.5 Fronts in Multiscale Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442.6 Variational Principles, Speed Bounds, and Asymptotics . . . . . . . . . . . 482.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3 Fronts in Random Burgers Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.1 Main Assumptions and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.2 Hopf–Cole Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563.3 Asymptotic and Probabilistic Preliminaries . . . . . . . . . . . . . . . . . . . . . 583.4 Asymptotic Reductions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 593.5 Front Probing and Central Limit Theorem . . . . . . . . . . . . . . . . . . . . . . 633.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations 694.1 Convex Hamilton–Jacobi and Variational Formulas . . . . . . . . . . . . . . 704.2 Subadditive Ergodic Theorem and Homogenization . . . . . . . . . . . . . . 734.3 Unbounded Hamiltonians: Breakdown of Homogenization . . . . . . . . 784.4 Normal and Accelerated Fronts in Random Flows . . . . . . . . . . . . . . . 834.5 Central Limit Theorems and Front Fluctuations . . . . . . . . . . . . . . . . . 86

ix

Page 9: Surveys and Tutorials in the Applied Mathematical Sciences

x Contents

4.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5 KPP Fronts in Random Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935.1 KPP Fronts in Spatially Random Shear Flows . . . . . . . . . . . . . . . . . . . 935.2 KPP Fronts in Temporally Random Shear Flows . . . . . . . . . . . . . . . . . 1085.3 KPP Fronts in Spatially Random Compressible Flows . . . . . . . . . . . . 1175.4 KPP Fronts in Space–Time Random Incompressible Flows . . . . . . . . 1235.5 Stochastic Homogenization of Viscous HJ Equations . . . . . . . . . . . . . 1325.6 Generalized Fronts, Reactive Systems, and Geometric Models . . . . . 1365.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

Page 10: Surveys and Tutorials in the Applied Mathematical Sciences

Chapter 1Fronts in Homogeneous Media

Front propagation and interface motion occur in many scientific areas such as chem-ical kinetics, combustion, biology, transport in porous media, and industrial deposi-tion processes. In spite of these different applications, the basic phenomena can allbe modeled using nonlinear partial differential equations or systems of such equa-tions. Since the pioneering work of Kolmogorov, Petrovsky and Piskunov (KPP)[134] and Fisher [91] in 1937 on traveling fronts of the reaction–diffusion equa-tions, the field has gone through enormous growth and development. However, stud-ies of fronts in heterogeneous media have been more recent. Heterogeneities arealways present in natural environments, such as fluid convection effects in combus-tion (wind factor in spreading of forest fires), inhomogeneous porous structures intransport of solutes, noise effects in biology, and deposition processes.

It is both a fundamental and a practical problem to understand how hetero-geneities influence the characteristics of front propagation such as front speeds,front profiles, and front locations. Our goal here is to give a tutorial of recent resultson front propagation in heterogeneous (especially random) media in a coherent andmotivating manner. It is not the intention of the book to give a complete survey, andso the cited references will cover only a portion of the vast literature.

Depending on the applications, three prototype equations appear, they are con-servation laws, Hamilton–Jacobi equations (HJ), and reaction–diffusion (RD) equa-tions. A class of problems of fronts in heterogeneous media arises in solute transportthrough porous media, an important subject in groundwater and environmental sci-ence. When solutes (ions) migrate inside porous media, some of them tend to attachonto the surface of minerals or colloids due to the existence of nonneutralized elec-tric charges at the surface or inside these minerals [158]. This surface effect is calledadsorption, which often creates a retardation on the movement of solute substance.The transport equation for the concentration of a one-species solute is based on theconservation of mass. When the adsorption reaches equilibrium, which often hap-pens in a much shorter time than the time scale of solute migration, one arrives atthe following conservation law [39, 71] for the solute concentration C:

(ωC +ρψ(C))t = ∇ · (θD∇C−vC), (1.1)

© Springer Science + Business Media, LLC 2009

1J. Xin, An Introduction to Fronts in Random Media, Surveys and Tutorials in the AppliedMathematical Sciences 5, DOI: 10.1007/978-0-387-87683-2_1,

Page 11: Surveys and Tutorials in the Applied Mathematical Sciences

2 1 Fronts in Homogeneous Media

where D is the pore-scale dispersion (viscosity) matrix, v is the incompressiblewater-flow velocity, ω is the total porosity, and ρ = (1−ω)ρs, ρs are the densi-ties of solid particles (minerals and colloids). The function ψ = ψ(C) is called thesorption isotherm. For example, the Freundlich isotherm is of the form ψ(C) = κCp,p ∈ (0,1), where κ represents the spatial distribution of sorption sites. Due to theheterogeneous nature of the porous media, both v and κ are functions of the spatialvariable x. The lack of detailed field information (uncertainties) naturally leads tostatistical modeling. Both v and k are treated as random processes [253, 38, 39, 197].If D = 0 (inviscid regime), a change of variable u = ωC+ρψ(C) converts (1.1) intothe standard form

ut +∇ · f(x,u) = 0, (1.2)

where f is a vector stochastic flux function. The celebrated Burgers equation resultsif the flux function is scalar and equal to u2/2. Since solutions of the inviscid equa-tion (1.2) may become discontinuous [141], we shall view them as weak solutions

times the viscous version of the conservation laws will be analyzed for ease of rep-resentation of fronts and connections with physical modeling and reaction–diffusionequations. Many other equations of the form (1.2) arise in hydrology and multiphaseflows, for example the Richards equation [195, 196, 88] and the Buckley–Leverettequation [118], to name just a few.

In manufacturing of nanomaterials and computer chips, atoms are deposited ontoa substrate; then they may wander, diffuse, and stick to form a complicated land-scape (interface) in the growth process of thin films. In continuum modeling ofinterface growth, a quadratic HJ equation, known as the Kardar, Parisi, and Zhang(KPZ) equation [13, Chapter 6],

ht = ν∇2h+λ2|∇h|2 +η(x, t), (1.3)

describes the evolution of the interface height, where ν > 0 is the diffusion constant,λ is a growth constant (positive if adding material to the interface), and η is a space–time uncorrelated (white) noise reflecting the random fluctuations in the depositionprocess. The gradient of h satisfies a viscous stochastic Burgers equation. In model-ing of turbulent combustion of premixed flames, another HJ equation, known as theG-equation, is proposed [156, 238] and extensively studied for thin flame fronts inadvecting fluid [194, 126, 127]:

Gt +v ·∇xG = sL |∇xG|, (1.4)

where the scalar function G is the level set function of the flame front, v is a stochas-tic flow field, and sL is the (laminar) flame speed in the absence of the flow. Thelevel set G(x, t) = G0, represents a moving interface; G > G0 is the burnt (hot) re-gion, G < G0 is the unburnt (cold) region. Equation (1.4) amounts to saying that thenormal velocity of the front is equal to sL + v ·n, where n is the normal directionpointing to the unburnt region.

obtained from zero viscosity limit of viscous solutions (D ↓ 0). In this tutorial, some-

Page 12: Surveys and Tutorials in the Applied Mathematical Sciences

1 Fronts in Homogeneous Media 3

Figure 1.1 Example of a reactive front in a random medium. Planar laser-induced fluorescencelight sheet image of an experimental arsenous-acid/iodate autocatalytic expanding front in randomcapillary wave flow of broad scales. Field view 14 cm × 14 cm, with the ratio U of root meansquare (rms) magnitude of flow velocity v and laminar front speed sL equal to 650. The arsenous-acid/iodate reaction takes place in an aqueous solution and has two advantages over gas reactions:(1) it allows small density changes across the reaction front, (2) it permits large rms values of vwhere front normal velocity still depends on local flow and curvature. Both properties are helpfulfor comparing experimental findings with predictions from the HJ-type front models such as theG-equation. The capillary wave flow is achieved in a thin layer of liquid in a vertically vibratedtray (20 cm on each side). At large amplitude, the flow field becomes random and develops a broadrange of spatial and temporal scales. In the experiment, the flow has zero ensemble mean and isisotropic and quasi-two-dimensional. For details of the experimental setup, see [112]. We observethe anisotropic and multiscale features of the front. Image used with the permission of Dr. PaulRonney.

The G-equation ignores chemical kinetics, front width, and diffusion. The first-principle-based model is the reaction–diffusion-advection equation (or a system ofsuch equations)

ut +v ·∇xu = D ∇2u+1τ

f (u), (1.5)

where u is the concentration of reactant, D is the diffusion constant, and τ is thetime scale of reaction. As we shall show in Chapter 5, for the KPP reaction f (u) =u(1− u), the large time front speed from compactly supported initial data is givenby averaging a KPZ equation with random advection (η replaced by −v ·∇xh in(1.3)). On the other hand, the G-equation and its variants are able to approximatefront speeds for ignition-type reactions in combustion when the reaction time τ andfront width O(

√D) are much smaller than the time and length scales in v.

Page 13: Surveys and Tutorials in the Applied Mathematical Sciences

4 1 Fronts in Homogeneous Media

0 20 40 60 80 100 1200

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

x

u

Figure 1.2 Sketch of traveling fronts U(x− ct) moving at constant speed in a homogeneousmedium.

Figure 1.1 shows a laser-induced fluorescence light sheet image of an experi-mental arsenous-acid/iodate autocatalytic front in random capillary wave flows. Thereactive front spreads out nonuniformly and takes on a fractal shape. A KPP frontspreading is similar, as we shall analyze later. In particular, the front speed vari-ational characterization allows one to estimate and compute the spreading rate inrandom media. A new theme associated with fronts in heterogeneous media is theunderstanding of multiple scales and their interaction. We illustrate how to applyhomogenization (upscaling) ideas to front problems in periodic and random media.Basic ideas of homogenization theory explained through concrete examples serveas useful guides.

We begin in Chapter 1 with the scalar prototype equations in homogeneous mediaand explain the basic properties of front solutions. A traveling-front solution in ahomogeneous medium is a solution of the form U(p · x− ct), where p is a unitvector, c is a constant speed independent of p, and U is the front profile. Sincethe front speed and profile are the same in each direction p, it is convenient toconsider the traveling front U(x− ct) in the case of one spatial dimension. Thescalar conservation laws are partial differential equations (PDEs) of the form

ut + f (u)x = ν uxx, (x, t) ∈ R× (0,∞), (1.6)

where f is a smooth nonlinear function, the flux function; and ν ≥ 0 is the viscosityparameter. The HJ equations are:

ut +H(ux) = νuxx, x ∈ R, (1.7)

Page 14: Surveys and Tutorials in the Applied Mathematical Sciences

1.1 Traveling Fronts of Burgers and Hamilton–Jacobi Equations 5

where H is the Hamiltonian and is nonlinear in ux (momentum). The R-D equationsare

ut = duxx + f (u), (1.8)

where d > 0 is the diffusion coefficient and f (u) is the reaction nonlinearity. Fig-ure 1.2 shows a right-moving RD front or a front of a viscous conservation law.

All three equations are well documented in graduate textbooks on PDEs [80,226]. We shall look at these three nonlinear equations in terms of traveling-frontsolutions, variational characterizations, front existence and stability, front speed se-lection, and general solution formulas. Many of these properties carry over to het-erogeneous fronts.

1.1 Traveling Fronts of Burgers and Hamilton–Jacobi Equations

For the viscous Burgers equation

ut +uux = νuxx, ν > 0, x ∈ R, (1.9)

we seek a front solution

u(x, t) = U(x− ct) = U(ξ ), ξ = x− ct

connecting the equilibria of zero and one (see Figure 1.2). Upon substitution, theprofile U(ξ ) satisfies a second-order ordinary differential equation (ODE), whichcan be solved exactly under the boundary conditions U(−∞) = 1 and U(+∞) = 0.Details are left as an exercise at the end of the chapter (Section 1.6). The solution is

U(ξ ) =1

1+ expξ/2ν , (1.10)

where x can be shifted by any constant x0 ∈ R. The front (1.10) moves to the rightat speed 1/2 without changing its shape and is called a traveling front.

If the initial data u(x,0) =U(x)+V (x), V (x) give a smooth function with enoughdecay at spatial infinities, a classical result [121] says that solution u(x, t) eventu-ally converges to U(x− t/2 + x0) uniformly in x for a constant x0. The constant x0depends on the integral (mass) of the initial perturbation V (x). In fact, the Burgersequation conserves the total mass

∫R u(x, t)dx. For a bounded and decaying V (x),

there is a unique value x0 such that∫

R[u(x,0)−U(x+ x0)]dx = 0;

hence by conservation of mass we have∫

R

[u(x, t)−U

(x− t

2+ x0

)]dx = 0, ∀t > 0.

Page 15: Surveys and Tutorials in the Applied Mathematical Sciences

6 1 Fronts in Homogeneous Media

If V (x) is also small, then x0 is small. Taylor expanding the above equality at t = 0shows that ∫

R[u(x,0)−U(x)−U ′(x)x0]dx≈ 0,

implying

x0 ≈∫

R[u(x,0)−U(x)]dx =

RV dx. (1.11)

So for a small perturbation, x0 is approximately the total mass of the initial pertur-bation. In the limit ν ↓ 0, Uν converges to a step function (shock front) that alsotravels to the right at speed 1/2.

The simplest traveling-front solutions of HJ (1.7) are the linear functions

u(x, t) = px−H(p)t, (1.12)

where p is a nonzero wave number. This is seen by direct substitution in (1.7). Theycome from spatial integrals of constant solutions of a scalar conservation law. Itslevel set u(x, t) = constant is a line moving at constant speed. In multiple spatialdimensions, (1.7) extends to p · x−H(p)t.

The level set, u = constant represents a hyperplane moving at speed H(p) inthe unit direction p. In one spatial dimension, if p = 1, H(p) = p2/2 as in theBurgers equation, the HJ front moves to the right with speed 1

2 . Clearly, there mayexist other traveling-front solutions. For example, if H(u) = u2/2, the integral of theBurgers traveling front −∫ ∞

x−t/2 U(ξ )dξ is an HJ traveling front, where U is givenby (1.10). For small ν , this solution is approximately a moving cone, or the functionx1(−∞,0)(x), with 1A being the indicator function of the set A. One may view sucha traveling front as consisting of two solutions of the form (1.12) with two valuesof p. The traveling front (1.12) is truly a planar solution, a building block of morecomplicated solutions.

The Burgers equation (1.9) has a closed-form solution formula, or (1.9) is in-tegrable. By the substitution u = −2νϕx/ϕ , we obtain the linear heat equationϕt = νϕxx for ϕ and the well-known Hopf–Cole formula [237] in terms of the heatkernel:

u(x, t) =(∫ ∞

−∞

x−ηt

exp−(2ν)−1G(η)dη)(∫ ∞

−∞exp−(2ν)−1G(η)dη

)−1

,

(1.13)where

G(η) = G(η ;x, t) =∫ η

0u(η ′,0)dη ′+(2t)−1(x−η)2. (1.14)

For a convex Hamiltonian H with superlinear growth, lim|p|→∞ H(p)/p = +∞,the Legendre transform

L(p) = supq∈R

(p ·q−H(q)) (1.15)

defines the convex Lagrangian L, which also grows superlinearly at large |p|. Sup-pose the initial datum u(x,0) = g(x) is a Lipschitz continuous function such that

Page 16: Surveys and Tutorials in the Applied Mathematical Sciences

1.2 Traveling Fronts of Reaction–Diffusion Equations 7

|g(x)−g(y)| ≤ Lp |x− y| for some constant Lp and all x,y ∈ R, with the inviscid HJsolution in the limit of ν ↓ 0 being given by the Hopf formula [80]

u(x, t) = infy∈R

[tL

(x− y

t

)+g(y)

], t > 0, (1.16)

which is Lipschitz continuous in R× [0,∞).The Hopf solution (1.16) is almost everywhere differentiable in (x, t), where it

satisfies the HJ equation and its initial condition [80]. Because the Lagrangian L hassuperlinear growth and g has at most linear growth, the infimum in (1.16) is attainedat a finite point y ∈ R.

The solution formula of the inviscid Burgers equation (or a convex conser-vation law) can be derived from the Hopf formula. For front initial data u(x,0)with enough decay at plus infinity, define h(x) = −∫ +∞

x u(x′,0)dx′ and w(x, t) =−∫ +∞

x u(x′, t)dx′. Then w solves the HJ equation wt + f (wx) = 0, w(x,0) = h(x),and the Hopf formula for w gives the solution formula for u = wx. A variant of thisrepresentation is called the Lax–Oleinik formula [141, 80], which is also the inviscidlimit (ν ↓ 0) of (1.6); see [80].

1.2 Traveling Fronts of Reaction–Diffusion Equations

The traveling-front solutions to reaction–diffusion equations (RD) (1.8) are specialsolutions of the form u = U(x− ct)≡U(ξ ), where c is the front speed and U is thefront profile that connects 0 and 1. Substituting this form into (1.8) with d = 1, weobtain

Uξ ξ + cUξ + f (U) = 0, (1.17)

with boundary conditions U(−∞) = 0 and U(∞) = 1. Since u is a concentration ora temperature, we also impose the physical condition U(ξ )≥ 0.

Note that we could also have an RD front with U(−∞) = 1 and U(+∞) = 0 bysimply changing variables x→−x, c→−c in (1.17), which is invariant. However,for a convex conservation law or the Burgers equation, the front must connect one atminus infinity to zero at plus infinity. Reversing the boundary conditions at infinitywill produce the so-called rarefaction (expansion) wave. This is because (1.6) mod-els compressible fluids (air). The front describes air compression physically, and isformed only under a pressure difference (or the pressure is higher at the inlet of apiston than the pressure inside). In contrast, the RD equation (1.8) models a flame iff is a suitable nonlinearity (type 4 or type 5 below, line–circle or line–star in Figure1.3). A flame moves from a hot (u = 1) region to a cold (u = 0) region, irrespectiveof whether the hot region is on the left or on the right side. This is an interesting dif-ference between (1.6) and (1.8). The other is that the front speed of (1.6) is explicit,thanks to the conservation of mass, while that of (1.8) is typically implicit due tothe lack of an invariant quantity in the evolution. The similarity of the fronts of (1.6)and (1.8) is that they all carry information on speed and profile.

Page 17: Surveys and Tutorials in the Applied Mathematical Sciences

8 1 Fronts in Homogeneous Media

0 0.2 0.4 0.6 0.8 1−0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

u

f

type 1type 2type 3type 4type 5

Figure 1.3 Sketch of five types of reactive nonlinearities: type 1 (u(1− u), KPP-Fisher) solidline; type 2 (u2(1− u), higher-order KPP–Fisher) dotted line; type 3 (bistable) line–dot; type 4(Arrhenius) line–circle; type 5 (ignition) line–star.

To be specific, for the study of traveling fronts, we will be concerned with thefollowing five types of nonlinearities:

1. f (u) = u(1− u): the Kolmogorov–Petrovsky–Piskunov (KPP) [134] or Fishernonlinearity [91];

2. f (u) = um(1− u), m an integer ≥ 2: the mth-order Fisher nonlinearity (calledthe Zeldovich nonlinearity if m = 2);

3. f (u) = u(1−u)(u−µ), µ ∈ (0,1): the bistable nonlinearity;4. f (u) = e−(E/u)(1− u), E > 0: the Arrhenius combustion nonlinearity or com-

bustion nonlinearity with activation energy E but no ignition temperature cutoff;5. f (u) = 0 ∀u ∈ [0,θ ]∪1, f (u) > 0 ∀u ∈ (θ ,1), f (u) Lipschitz continuous: the

combustion nonlinearity with ignition temperature θ .

Types 1 and 2 come from chemical kinetics (for example, from autocatalyticreactions [142]), with type 2 being the high-order generalization of type 1. Type3 comes from biological applications (such as FitzHugh–Nagumo systems [9]) andalso more recently from phase field models of solidification in material science [90].Types 4 and 5 appear in the study of premixed flames in combustion science [28,238]. Types 1, 2, and 4 are nonnegative and can be recovered as a limit of type 5 asθ tends to zero.

If we look at the graphs of f (u) for the five types in Figure 1.3, we see that theydiffer near u = 0 and behave similarly near u = 1. The type-1 nonlinearity has apositive slope at u = 0. The type-2 nonlinearity has zero slope (and derivatives up

Page 18: Surveys and Tutorials in the Applied Mathematical Sciences

1.2 Traveling Fronts of Reaction–Diffusion Equations 9

to order m− 1 for m ≥ 2). The type-4 nonlinearity has an exponentially small tailnear zero, so all derivatives at zero vanish. Type 5 is identically zero for an intervalu∈ [0,θ ], i.e., there is no reaction below ignition temperature. Type 3 has a negativeslope at u = 0, then goes down to a negative minimum, goes up and through anintermediate zero µ , then up to its positive maximum, and finally comes back to itsthird zero at u = 1. Type 3 is the only one that changes sign. Its total area

∫ 10 f (u)du

is positive if µ ∈ (0, 1

2

), zero if µ = 1

2 , and negative if µ > 12 . If we gradually deform

the curve of f = f (u) near u = 0 from above the u-axis (type 1) to below (type 3),we can experience all five types of nonlinearities.

The boundary value problem (1.17) can be thought of as a nonlinear eigenvalueproblem with eigenvalue c and eigenfunction U . It is convenient to perform a phaseplane analysis by writing (1.17) as a first-order system of ODEs,

Uξ = V, Vξ =−cV − f (U). (1.18)

Now we are looking for a trajectory in the phase plane that goes from (0,0) to(1,0). Multiplying both sides of (1.17) by Uξ and integrating over ξ ∈R, we obtain(assuming µ ∈ (

0, 12

)in the case of a nonlinearity of type 3)

c =−∫ 1

0 f (U)dU∫RU2

ξ dξ< 0.

The linearized system about U = 0 is

ddξ

(UV

)=

(0 1

− f ′(0) −c

)(UV

).

The eigenvalues of this 2×2 matrix are given by

λ1,2 =−c±

√c2−4 f ′(0)2

.

In the case of a type-1 nonlinearity, if c2 ≥ 4 f ′(0) or c ≤ c∗1 ≡ −2√

f ′(0), then(0,0) is an unstable node. In the type-3 case, since f ′(0) < 0, it follows that (0,0) isa saddle. In either case, a similar linearization at (1,0) shows that (1,0) is always asaddle, thanks to f ′(1) < 0. Since there is a family of unstable directions going outof an unstable node, and only one direction in or out of a saddle, one can show byisolating the flows in a triangular region in the first quadrant of the U-V plane thatthere is a connecting trajectory for each c≤ c∗1 for type 1, and a unique connectingtrajectory for type 3. Moreover, Uξ > 0 always holds, thanks to the trajectory beingin the first quadrant. Since the ODE system (1.18) is autonomous, U is unique onlyup to a constant translation of ξ .

We shall call the front solution corresponding to c = c∗1 the critical front. The crit-ical front moves at the slowest speed in absolute value, and its asymptotic behavioras |ξ | → ∞ is [8]

Page 19: Surveys and Tutorials in the Applied Mathematical Sciences

10 1 Fronts in Homogeneous Media

U(ξ ) =

1−Ce−βξ +O(e−2βξ ), ξ →+∞,

(A−Bξ )e−c∗1ξ/2 +O(ξ 2e−c∗1ξ ), ξ →−∞,(1.19)

where A, B, and C are positive constants and 2β = −c∗1−√

(c∗1)2−4 f ′(1) > 0. Incontrast, the faster fronts with c < c∗1 have exponential decay O

(econst·ξ )

as |ξ | →∞because the two roots λ1,2 at (0,0) are both simple. The faster fronts decay moreslowly than the critical fronts as U → 0.

For the type-3 cubic polynomial, Huxley [212] solved (1.17)–(1.18) exactly:

U(ξ ) =1

1+ e−ξ/√

2, c =

√2(

µ− 12

)(1.20)

for µ ∈ (0, 1

2

]. If µ ∈ [ 1

2 ,1], one simply switches c to −c and ξ to −ξ .

For the remaining three types, f ′(0) = 0, and so (0,0) has an unstable and aneutral direction. More delicate analysis is required. In the type-2 case, one canshow that there are a center manifold near (0,0) and a connecting trajectory from thecenter manifold to the saddle at (1,0) for each c < c∗m < 0. If c = c∗m, the connectiongoes from the unstable manifold at (0,0) to the saddle. For a type-2 nonlinearity withm = 2, the critical front profile approaches zero at the rate O(e−c∗2ξ ) as ξ →−∞,while the profiles of faster fronts approach zero at an algebraic rate O(ξ−1) [32].This is different from the KPP case (1.19). The absolute values of c∗m decrease withincreasing m.

For type-4 and type-5 nonlinearities, a different method using degree theory on fi-nite intervals to construct approximate solutions, then taking their infinite line limit,is much more expedient and robust; see [28, 155]. We will explain this method indetail in the coming sections on fronts in periodic media. The results of [28] and[155] show in particular that in the case of a type-4 nonlinearity, a continuum fam-ily of traveling-front solutions exists, one for each c ≤ c∗0 < 0, just as for type 1and type 2. However, type 5 is different from type 4 in that for each given ignitiontemperature θ > 0, there is a unique c∗θ such that a corresponding front profile Uexists and is unique up to a constant translation in ξ . We see that type 5 is just liketype 3. See [89, 228] for a phase-plane justification of the result.

The degree approach is very effective in proving the existence of traveling wavesin multiple dimensions. For example, the existence of traveling fronts in channeldomains R×Ω , Ω is a bounded domain with Lipschitz continuous boundary in Rn,n ≥ 1. The front moves along the channel and has the form U(x− ct,y), y ∈ Ω ,due to the y-dependent coefficient in the equation (equation (2.19)); see [27, 31] onthe existence theory for all five types of nonlinearities. The other related methodfor existence of traveling fronts is the Conley index theory [226]; see [99] for itsapplication to the existence of fronts in a channel domain.

The next question is the asymptotic stability of traveling fronts in large time. Thestability means that if the initial data are prescribed in the form u0(x) = Uc(x) +u′(x), where Uc(x) is a front profile corresponding to the speed c and u′(x) is asmooth and spatially decaying perturbation, then u(x, t) converges to Uc(x+ct +ξ0)in a proper function space as t → ∞ for some constant ξ0. The reason we have

Page 20: Surveys and Tutorials in the Applied Mathematical Sciences

1.3 Variational Principles of Front Speeds 11

a constant translation in the definition can be seen as follows. Due to the spatialtranslational invariance of the original equation, we have a family of traveling frontsUc(x− ct + x0) for each allowable wave speed c. Let us take u′(x) = Uc(x + x0)−Uc(x), which is a perturbation with spatial decay. Now for initial data u(x,0) =Uc(x)+u′(x) = Uc(x + x0), the solution for later time is just Uc(x + ct + x0), whichdoes not converge to Uc(x+ct) unless x0 = 0. In the case of a continuum of speeds,we can also take u′(x) =Uc′(x)−Uc(x), c′ 6= c, and the later-time solution is Uc′(x−c′t), again not converging to Uc(x− ct) as t → ∞. Even the wave speed is different.The convergence to a translated front is similar to Burgers’ equation except that thetranslated amount is implicit, while for (1.6) it is the total mass of the initial (small)perturbation (1.11).

These simple examples show that it is a subtle problem to establish asymptoticstability, especially in the case of multiple speeds. Much turns out to depend on therate of decay of the initial perturbations as |ξ | → ∞. Intuitively, the tiny amount ofperturbation in the far field takes a long time to crawl into a front from its tails; how-ever, its effect is crucial, since asymptotic stability concerns large-time behavior.Asymptotic stability for noncritical fronts based on the spectral theory of linearizedoperators can be found in [212]. For asymptotic stability of critical fronts with min-imal speeds, see [131] on KPP nonlinearity, [42] on Ginzburg–Landau (GL) non-linearity f (u) = u(1− u2), [98] on both GL and KPP, and [75] on more generalparabolic equations. For the global asymptotic stability and critical front selectionbased on analysis using maximum principles, see [8, 9, 89, 90, 124], the originalpaper [134] with the initial data being the indicator function of the negative line,and for a probabilistic analysis of the KPP equation, see [40, 95, 160].

Large-time asymptotic stability is also proved for traveling fronts in channel do-mains; see [205, 15], where in particular the critical front stability problem of KPPfor type-2 and type-4 nonlinearities is resolved.

1.3 Variational Principles of Front Speeds

Since the speeds of the RD fronts are in general unknown in closed form, the varia-tional characterization is an invaluable way of estimating them. For general contin-uously differentiable nonlinearity f (u) such that

f (0) = f (1) = 0, f (u) > 0, u ∈ (0,1), f ′(0) > 0, f ′(1) < 0, (1.21)

the min-max variational principle for the minimum speed was first established [107]:

|c∗|= infρ

supu∈(0,1)

ρ ′(u)+

f (u)ρ(u)

, (1.22)

where ρ is any continuously differentiable function on [0,1] such that

ρ(u) > 0, u ∈ (0,1), ρ(0) = 0, ρ ′(0) > 0. (1.23)

Page 21: Surveys and Tutorials in the Applied Mathematical Sciences

12 1 Fronts in Homogeneous Media

The formula (1.22) is based on the phase-plane construction of the fronts. Under(1.21), for each allowable c, there is a connection from an unstable node to a saddle,and the front profile is strictly monotone. Let u = u(x−ct) = u(ξ ) connect u = 1 andu = 0 from left to right, so that c≥ c∗ > 0. Then uξ ξ + cuξ + f (u) = 0, u(−∞) = 1and u(∞) = 0. Now regard uξ as a function of u by defining p = p(u) =−uξ > 0 atu = u(ξ ). The function p(u) is a solution of

p(u)p′(u)− cp(u)+ f (u) = 0, (1.24)

with p(0) = 0, p(1)= 0, and p(u) > 0 on (0,1). The expression inside the supremumof (1.22) is just what we find from (1.24) on writing c in terms of u, p, and p′(u). Ifp(u) is not a solution of (1.24), then (u, p(u)) can represent a curve connecting thenode and the saddle in the phase plane, but with the flow field at the curve pointingtoward the solution curve of (1.24). This geometric information translates into theinequality that c is no less than the supremum in (1.22) for some ρ satisfying (1.23).It follows that any allowable c, in particular c∗, is no less than the min-max of (1.22).The equality is attained by ρ = p∗(u) corresponding to the speed c∗.

It follows from (1.22) that

2√

f ′(0)≤ |c∗| ≤ 2√

L, L = supu∈(0,1)

f (u)u

, (1.25)

which gives the well-known KPP minimal speed 2√

f ′(0) if L is achieved at u = 0.To see (1.25), we take ρ(u) = au, a > 0. Then |c∗| ≤ a + L/a. Minimizing over aestablishes the upper bound. The lower bound is easily deduced by restricting thesupremum to those functions u in a small neighborhood of zero. The formula (1.22)was used further in [107] to find the exact minimal speed for f (u) = u(1− u)(1 +νu): |c∗|= 2 if −1≤ ν ≤ 2, |c∗|= (ν +2)/

√2ν if ν ≥ 2.

Recently, a general variational speed formula was found [17] for any f such thatf (0) = f (1) = 0. Let f be any of the five types of nonlinearity, and assume that amonotone front exists. Then the minimum (or unique) speed c∗ is given by

c2∗ = sup

(2

∫ 10 f gdu

∫ 10 (−g2/g′)du

), (1.26)

where the supremum is over all positive decreasing functions g ∈ (0,1) for whichthe integrals exist. Moreover, the maximizer exists if |c∗| > 2

√f ′(0). The formula

(1.26) appears to be the first variational result in such generality. The formula holdsfor f changing signs in (0,1). The Huxley formula (1.20) is recovered by puttingg(u) = ((1− u)/u)1−2µ . A similar variational formula for f in (1.21) without theconstraint f ′(1) < 0 is established in [16].

The proof of (1.26) is elementary, and it uses (1.24) again. Let g = g(u) be anypositive function on (0,1) such that h =−g′(u) > 0. Multiplying (1.24) by g(u) andintegrating over u ∈ [0,1], we have after integration by parts the equality

Page 22: Surveys and Tutorials in the Applied Mathematical Sciences

1.4 Random Variables and Stochastic Processes 13

∫ 1

0f gdu = c

∫ 1

0pgdu−

∫ 1

0

12

hp2 du. (1.27)

For positive c, g, and h, the function ϕ(p) = cpg− 12 hp2 has its maximum at p =

cg/h, and so ϕ(p)≤ c2g2/2h. It follows that

c2 ≥ 2∫ 1

0 f gdu∫ 1

0 (g2/h)du≡ I(g), (1.28)

which implies (setting c = c∗ if c is nonunique) that c2∗ is no less than the supremumof (1.26). Next, we show that the equality holds for a function g. Notice that thecondition p = cg/h is solvable in g and gives an expression for the maximizer g,

g = exp−

∫ u

u0

cp−1 du

, (1.29)

with u0 ∈ (0,1). Clearly, g is positive and decreasing, with g(1) = 0, since p ∼O((1−u)) for u∼ 1. At u = 0, however, g diverges, since the exponent goes to +∞.A natural choice for g now is p = p∗(u) if we verify that the two integrals are finitein I(g). For nonlinearities of types 2, 4, and 5, p∗ approaches zero exponentially and

p∗ ∼ c+√

c2−4 f ′(0)2

u≡ mu.

Thus, near u = 0, g∼ u−c/m, and f g and g2/h diverge at most like u1−c/m. The inte-grals of I(g) are finite if m/c > 1

2 . This condition holds if f ′(0)≤ 0, which is indeedtrue for types 2, 4, and 5, and also for f in (1.21) if c2∗ > 4 f ′(0). If c2∗ = 4 f ′(0),which is the case for type 1, the maximizer does not exist. However, choosing thetest function g(u) = a(2−a)ua−2 with a ∈ (0,1), we calculate

∫ 10 (g2/h)du = 1. In-

tegration by parts twice shows that as a→ 0, I(g) = 2(2−a)a∫ 1

0 f ua−2 du→ 4 f ′(0).The proof is complete.

1.4 Random Variables and Stochastic Processes

In this section, we give a brief introduction to random variables and stochastic pro-cesses as a preparation for later chapters. We shall follow [41, 72, 206] for basicdefinitions and concepts in probability and stochastic processes below, and refer to[55, 206] for more examples and applications. First, a probability space is a triple(Ω ,F ,P), where (1) Ω is a set of outcomes (sample space), a subset of which iscalled an event; ∅ is the impossible event; (2) F is a collection of events includ-ing Ω and ∅ with the property that F is closed under complementation, countableintersections, and unions; (3) P is a probability function that assigns probability toevents in F . A probability is a nonnegative set function defined on F with values

Page 23: Surveys and Tutorials in the Applied Mathematical Sciences

14 1 Fronts in Homogeneous Media

in [0,1] such that (a) (normalization) P(Ω) = 1, P(∅) = 0; (b) (additivity) for anyfinite or countable disjoint collection Bk of sets in F , P(∪kBk) = ∑k P(Bk).

As a finite sample space example, consider throwing a die. There are six possibleoutcomes, denoted by ωi, i = 1, . . . ,6. The set of all outcomes is the sample spaceΩ = ω1, . . . ,ω6. A subset of Ω , an event, is A = ω2,ω4,ω6. Suppose we didN die experiments, and event A happened Na times. The probability of event A isP(A) = limN→∞ Na/N. For a fair die, P(A) = 1

2 . An infinite sample space exampleis Ω = (0,1), F is the collection of all open subsets and their countable unions inR, P is the Lebesque measure with P((a,b]) = b−a for all a < b. See [72, 55] formore examples.

A real-valued function X defined on Ω is a random variable. For example, an in-dicator function of a set A ∈F is a random variable (r.v.). The distribution functionof an r.v. is F(x) = P(X ≤ x), which is nondecreasing and right continuous in x, andF(+∞) = 1, F(−∞) = 0. If F(x) is absolutely continuous, the density function isf (x) = F ′(x). The expectation of X is

E[X ]≡ µ =∫ +∞

−∞x f (x)dx,

and the variance of X is

Var[X ] = E[(X−µ)2]≡ σ 2,

where σ is the standard deviation.A unit Gaussian r.v. (or a standard normal r.v.) is described by the density func-

tionf (x) = (2π)−1 exp−x2/2,

with µ = 0, σ = 1. The density function of a uniformly distributed r.v. on (0,1) isthe indicator function of the interval (0,1), with µ = 1/2, σ2 = 1/12.

Random variables X1,X2, . . . ,Xn are independent if for any ai ∈R (i = 1,2, . . . ,n),

P(∩ni=1ω : Xi(ω)≤ ai) =

n

∏i=1

P(ω : Xi(ω)≤ ai).

Likewise, events A1, . . . ,An are independent if

P(∩ni=1 Ai) =

n

∏i=1

P(Ai).

The countably many random variables X1,X2, . . . , are independent if for every n≥ 2,the random variables X1,X2, . . . ,Xn are independent.

Concerning the limit behavior of a sequence of random variables X1,X2, . . . ,Xn, . . . ,a few commonly used modes of convergence are as follows:

• (cp1) Convergence with probability one (almost sure convergence) if there existsan r.v. X such that

Page 24: Surveys and Tutorials in the Applied Mathematical Sciences

1.4 Random Variables and Stochastic Processes 15

P(ω ∈Ω : lim

n→∞|Xn(ω)−X(ω)|= 0

)= 1. (1.30)

• (msc) Mean-square convergence if E(X2i )≤C for a constant C, and

limn→∞

E(|Xn−X |2) = 0. (1.31)

• (cp) Convergence in probability if

limn→∞

P(ω ∈Ω : |Xn(ω)−X(ω)| ≥ ε) = 0, ∀ε > 0. (1.32)

• (cl) Convergence in law if there exists an r.v. such that

limn→∞

FXn(x) = FX (x) (1.33)

at all continuous points of FX (the distribution function of X). We shall denotethe limit in this sense of convergence or equivalence in the sense of distributionsas law= or d=.

General inference relations are cp1 (msc) =⇒ cp =⇒ cl.The fundamental laws of probability on a sequence of independent identically

distributed (iid) random variables Xi can be stated as follows. Suppose E|Xi| < ∞,and set µ = E(Xi), σ2 = Var(Xi) ∈ (0,∞). The strong law of large numbers refersto the convergence of

Sn

n= ∑n

i=1 Xi

n→ µ (1.34)

in the sense of (wp1) and (msc). The weak law of large numbers refers to (1.34) inthe sense of (cp). The central limit theorem (CLT)

Sn−nµσ√

n→ N(0,1) (1.35)

holds in law (cl), where N(0,1) is the unit Gaussian.A sequence of random variables X1, X2, . . . ,Xn, . . . occurring at discrete times

t1 < t2 < · · · < tn is called a discrete stochastic process, with joint distributionFXi1 ,Xi2 ,...,Xi j

as its probability law. The process is called Gaussian if all joint dis-tributions are Gaussian.

A continuous stochastic process X(t) = X(t,ω), t ∈ I an interval of R, over theprobability space (Ω ,A,P), is a function of two variables X : I×Ω → R, where Xis an r.v. for each t; for each ω , we have a sample path (a realization) or trajectoryof the process.

A few basic quantities are µ(t) = E(X(t,ω)), σ2(t) = Var(X(t,ω)), and thecovariance function

C(s, t) = E((X(s,ω)−µ(s))(X(t,ω)−µ(t))),

Page 25: Surveys and Tutorials in the Applied Mathematical Sciences

16 1 Fronts in Homogeneous Media

for s 6= t. The covariance measures how correlated the two random variables are attwo points of “time” s and t.

The standard Wiener Process (Brownian motion) is a Gaussian process W (t,ω),t ≥ 0, with independent increment, and

W (0) = 0 w.p. 1, E(W (t)) = 0, Var(W (t)−W (s)) = t− s, (1.36)

for all s∈ [0, t]. It follows that for any t0 < t1 < · · ·< tn, the random variables W (tk)−W (tk−1) are independent normally distributed with mean zero and E[(W (tk)−W (tk−1))2] = tk − tk−1. In particular, the covariance function C(s, t) is equal tomin(s, t).

A few useful properties of Wiener process are (1) (regularity) almost surely, thesample path W (t,ω) is Holder continuous with exponent less than 1

2 , hence nowheredifferentiable; (2) (scaling) the process W (t) = tW (1/t) if t > 0; W (0) = 0 is also aWiener process; (3) (large-time behavior) W (t,ω) oscillates and grows sublinearlyat large t; its upper and lower envelopes obey the law of the iterated logarithm, thatis, almost surely in ω ,

limsupt→∞

W (t)√2t log(log t)

= 1, (1.37)

liminft→∞

W (t)√2t log(log t)

=−1. (1.38)

The sublinear growth of W (t) is a consequence of its independent increments andthe law of large numbers. If t is a positive integer, write

W (t) = (W (t)−W (t−1))+(W (t−1)−W (t−2))+ · · ·+(W (2)−W (1))+(W (1)−W (0))

as a sum of iid random variables with mean zero and variance one, and so W (t)/t →0 almost surely. For general t, the same limit follows as |W (t)−W ([t])| ≤ Cω bycontinuity and independent increment properties of W , where [t] is the integral partof t.

A stochastic process is stationary if all joint distributions are translation-invariant.A Gaussian process is stationary if only its covariance function is translation-invariant.

The Wiener process is Gaussian but not stationary. A well-known stationaryGaussian process is the Ornstein–Uhlenbeck process (O-U), which is defined asa Gaussian process with X(0) a unit Gaussian, E(X(t)) = 0, and the covariancefunction E(XsXt) = e−γ |t−s| for s, t ∈ R, for some constant γ > 0. Another exam-ple is Gaussian white noise, formally W ′(t), the derivative of a Wiener process.One may construct it by passing to the limit on an approximate Gaussian processXh(t) = (W (t +h)−W (t))/h, for h > 0 small. The process Xh has covariance

Page 26: Surveys and Tutorials in the Applied Mathematical Sciences

1.4 Random Variables and Stochastic Processes 17

Ch(s, t) =1h

max(

0,1− |t− s|h

), (1.39)

whose Fourier transform (called spectral density) is

sin2(2πλh)(πλh)2 ≡ F

′h(λ ). (1.40)

In the limit h→ 0, Ch converges to a delta function, and F′h converges to a constant

(flat spectrum or white color). Accordingly, the process Xh converges in some weaksense to the white noise process.

Wiener and O-U processes all belong to a class of Markov processes called dif-fusion processes. Suppose that the multipoint joint distribution function of X(t) hasdensity p(t1,x1; t2,x2; . . . ; tk,xk), and define the conditional probability

P(X(tn+1) ∈ B|X(ti) = xi, i = 1 : n) =∫

B p(t1,x1, . . . , tn,xn; tn+1,y)dy∫p(t1,x1, . . . , tn,xn; tn+1,y)dy

(1.41)

for B any open set of R. The process is Markov if

P(X(tn+1) ∈ B|X(ti) = xi, i = 1 : n) = P(X(tn+1) ∈ B|X(tn) = xn),

and the transition probability is

P(s,x; t,B) =∫

Bp(s,x; t,y)dy,

where p is the transition density. A Markov process with transition density is calleda diffusion process if the following limits exist for any ε > 0:

limt→s+

1t− s

|y−x|>εp(s,x; t,y)dy = 0,

limt→s+

1t− s

|y−x|≤ε(y− x)p(s,x; t,y),dy = a(s,x),

limt→s+

1t− s

|y−x|≤ε(y− x)2 p(s,x; t,y)dy = b2(s,x).

Alternatively, we may write

a(s,x) = limt→s+

1t− s

E(X(t)−X(s)|X(s) = x),

b2(s,x) = limt→s+

1t− s

E((X(t)−X(s))2|X(s) = x). (1.42)

The function a is called the drift coefficient, and b is the diffusion coefficient. Driftand diffusion coefficients indicate the rates of infinitesimal motion of the processover slow (diffusion) and fast (drift) time scales. Using the definitions of Wiener

Page 27: Surveys and Tutorials in the Applied Mathematical Sciences

18 1 Fronts in Homogeneous Media

and O-U processes, one calculates that (a,b) = (0,1) for a Wiener process, and(a,b) = (−γx,2γ) for O-U. Over a small time interval [s, t], using drift–diffusioninformation, we see that O-U is related to W as

X(t)−X(s) =−γX(s)(t− s)+√

2γ(W (t)−W (s)),

or in differential form,dX =−γX dt +

√2γ dW, (1.43)

a stochastic differential equation (SDE), also known as the Langevin equation. Theterm −γXdt introduces damping on Brownian motion. For more discussion of SDEand Brownian motion, see [125].

If the initial distribution of O-U at t = 0 is a Gaussian (normal) r.v. with meanzero and variance ρ > 0, or N (0,ρ), then X is a stationary Gaussian process withcovariance C(t,s) = ρ e−γ |t−s|.

1.5 Noisy Burgers Fronts and the Central Limit Theorem

As an application of stochastic processes to fronts, we study the effects of initialwhite noise perturbations of Burgers fronts. This is a step beyond the determinis-tic localized perturbations. In this class of problems, the noise enters initially andthe governing equation (1.6)–(1.8) remain the same as in classical front stabilityanalysis.

Consider the Burgers equation (1.9) with initial data

u(x,0) =1

1+ expx/(2ν) +V (x), (1.44)

where V (x) is either white noise or a stationary Gaussian process with enough decayof correlations.

Now V (x), being a stationary random process, has no decay at infinities. It turnsout that at time t, the truncated mass of V (x), or the integral of V (x), over the inter-val

[− 12 t, 1

2 t]

plays the role of the whole line integral (1.11) and causes the deviationof front location from the mean position 1

2 t. The “one-half” comes from the unper-turbed front speed, and the interval

[− 12 t, 1

2 t]

resembles the domain of dependencefor the linear wave equation utt − 4−1uxx = 0. The picture behind this is that theperturbation gets sucked into the front from left and right at speed one-half. Let uscalculate formally the front deviation for white noise (formally Wx) as as in (1.11):

x0 = x0(t,ω)∼∫ t/2

−t/2Wx(x)dx = W (t/2)−W (−t/2)

law= Wtlaw=√

tW1, (1.45)

that is,√

t times the unit Gaussian. Thus the front location is

Page 28: Surveys and Tutorials in the Applied Mathematical Sciences

1.5 Noisy Burgers Fronts and the Central Limit Theorem 19

X = X(t,ω) =t2

+ x0(t,ω) law=t2

+√

tW (1). (1.46)

Figure 1.4 (top) illustrates a random front moving according to the law X(t,ω) =ct +W (t,ω), uniformly sampled in time (100 time slices) with the correspondingsuitably scaled velocity (bottom) at the sampled times. The noise term W is a nu-merical approximation of the Wiener process W . The constant c is positive andnonrandom. Compared with the uniform speed motion in Figure 1.2, the front speedin Figure 1.4 is highly oscillatory and random-looking.

The above heuristics are made precise in [233]:

Theorem 1.1. Let u(x, t) be the solution to the initial value problem of the Burgersequation (1.9) and (1.44). Let f be an increasing function of t. Then we have thefollowing:

1. (Front Probing) If f (t)−t/2√t → c ∈ R, then u( f (t), t) converges in distribution to

a random variable equal to zero with probability N (c) and equal to one withprobability 1−N (c), where

N (c) =1√2π

∫ c

−∞exp−y2/2dy

is the unit Gaussian distribution function.Given any positive number ε ∈ (0,1), let us define the left and right endpointsof the interval containing the front as

z−(t) = minx : u(t,x) = 1− ε, z+(t) = maxx : u(t,x) = ε,

and so the front width is z+(t)− z−(t). Then:2. (Front Width) There exists a constant t0 > 0 such that the random variablesz+(t)− z−(t) are tight for t ≥ t0; i.e., for any δ > 0, there exists an M suchthat Prob(z+(t)− z−(t) > M) < δ for all t > t0.

3. (Front Motion) As s→ ∞, there is a constant σ depending only on V (x) (σ = 1for white noise) such that (the same is true for z−)

z+(t)− t/2σ√

tlaw→W (1).

Part 1 is a slightly weaker version of 3, and both substantiate the formal calcu-lation. Part 2 says that the noise does not spread the front width for large time, sononlinearity dominates over the randomness and preserves the coherent structure.The proof uses the Hopf–Cole formula and a Laplace method for stochastic inte-grals [233].

Stability of other wave solutions in Burgers and convex conservation laws canbe found in [85, 235]. If one performs the hyperbolic scaling change of variablesx → x/ε , t → t/ε , ε small, the inviscid convex conservation law (1.6) is invariant.Suppose the unperturbed initial datum is the indicator function 1R±(x), the unit stepfunction supported on the half-line R±. Such data lead to a front (minus sign) or

Page 29: Surveys and Tutorials in the Applied Mathematical Sciences

20 1 Fronts in Homogeneous Media

0 20 40 60 80 100 1200

0.2

0.4

0.6

0.8

1u

x

0 20 40 60 80 100−0.4

−0.2

0

0.2

0.4

fron

t spe

ed

time

Figure 1.4 Sketch of one hundred uniformly sampled time slices of random fronts U(x−X(t,ω)),X(t,ω) = ct + W (t,ω) (top); with the corresponding suitably scaled front speed (bottom). Thenoise W is a numerical approximation of the Wiener process W . The front profile U is nonrandomand invariant in time, and c is a positive nonrandom constant.

a rarefaction wave (plus sign) corresponding to compression and decompressionof gas in a piston. Now we perturb the initial data with white noise, so u(x,0) =1R±(x)+Wx(x). After the scaling change of variables, the initial datum is uε(x,0) =1R±(x) +Wx(x/ε). Write the scaled solution uε(x, t) = vε

x , where vε is the Hopfsolution to the Hamilton–Jacobi equation:

vεt + f (vε

x) = 0, vε(x,0) = x 1R±(x)+ εW (x/ε). (1.47)

Because of the nearly square-root growth of W , the scaled perturbation εW (x/ε)goes to zero for x on any finite interval. In fact, using properties of a Wiener processand the Hopf formula, it can be shown [235] that with probability one, uε convergesin the sense of distributions to the unperturbed solution of (1.6), which is either theunperturbed shock (minus sign) or a rarefaction wave (plus sign). Hence both wavesare stable in the sense of the hyperbolic limit.

The result can be extended for colored noise (stationary Gaussian processes withdecaying correlations) [235]. However, the slower diffusive motion of the front isnot seen in this limit. Likewise, more detailed stability for a rarefaction wave re-quires a large-time asymptotic analysis of u; see [85] for the Burgers case. For a

Page 30: Surveys and Tutorials in the Applied Mathematical Sciences

1.6 Exercises 21

mathematical analysis of random Burgers and KPZ equations in connection withturbulence, see [73, 239].

1.6 Exercises

1. Find the Burgers traveling-front formula (1.10) by letting u = U(x− ct) in theBurgers equation (1.9), deriving a second-order ordinary differential equationfor U and solving it under the boundary condition U(−∞) = 1, U(+∞) = 0.

2. Derive the Hopf–Cole solution formula (1.13)–(1.14) by writing down a so-lution to the heat equation ϕt = νϕxx, then setting u = −2νϕx/ϕ . Find thecorrespondence between the initial data of the Burgers equation and the heatequation.

3. Verify Huxley’s traveling-front formula (1.20) for the bistable reaction–diffusionequation (1.8) with d = 1 and f (u) = u(1− u)(u− µ), µ ∈ (

0, 12

]. Then gen-

eralize the formula to the case d > 0 and study how the diffusion constant dinfluences the solution.

4. Use the conditional expectation formula (1.42) to show that the drift and dif-fusion coefficients of the Wiener process W are equal to (0,1). Likewise, forthe O-U process X(t), derive its drift and diffusion coefficients using the factthat the increment X(t + s)− e−γsX(t) is independent of the past or events inF (X(τ),τ ≤ t).

Page 31: Surveys and Tutorials in the Applied Mathematical Sciences

Chapter 2Fronts in Periodic Media

Fronts or interfaces in periodic media are deterministic problems in between ho-mogeneous media and random media. Much can be learned on how front solutionstransition from monoscale simple solutions in Chapter 1 to multiple-scale solutions.Periodic homogenization and PDE techniques based on maximum principles are es-sential tools for constructing front solutions and analyzing their asymptotics. Weshall observe the close relationship between Hamilton–Jacobi (HJ) and reaction–diffusion (RD) equations, and present the variational principles of front speeds.

2.1 Periodic Media and Homogenization

Multiscale problems are common in applications such as finding the effective con-

media, where one has at least two scales, the large scale of the sample and the smallscale of the embedded inclusions or pores. These two scales normally differ signifi-cantly and render the full resolution of the problem difficult. Therefore, it is of greattheoretical and practical interest to find out how to upscale the collective effect of thesmall scale into the large scale and simplify the problem. When the small scale pos-sesses a periodic structure, the upscale problem has a well-developed theory calledhomogenization. See [18] for a systematic account of the foundational works.

We give here an example of homogenization and use formal asymptotic analysisto illustrate the ideas. Consider a two-point boundary value problem of a second-order ODE with rapidly oscillating periodic coefficients,

(a(ε−1x)uεx)x = f (x), x ∈ [0,1], (2.1)

with boundary condition uε(0) = uε(1) = 0. Here a is a positive smooth functionwith period 1 in y ≡ ε−1x, and f (x) is a bounded continuous function in x. We aregoing to examine the limit of uε as ε → 0, where the large-scale x and small-scaleε−1x are separated. Since there are two separate scales in the problem, it is natural

ductivity of a composite material or the effective permeability for flows in porous

© Springer Science + Business Media, LLC 2009

J. Xin, An Introduction to Fronts in Random Media, Surveys and Tutorials in the Applied 23Mathematical Sciences 5, DOI: 10.1007/978-0-387-87683-2_2,

Page 32: Surveys and Tutorials in the Applied Mathematical Sciences

24 2 Fronts in Periodic Media

to search for a two-scale expansion of the solution in the form

uε ∼ u0(ε−1x)+ εu1(x,ε−1x)+ ε2u2(x,ε−1x)+ · · · , (2.2)

where the y = ε−1x dependence has period 1 also. Substituting the ansatz (2.2) into(2.1), and regarding x and y as independent variables, we have (noting that the xderivative is replaced by the operator ∂x + ε−1∂y)

(∂x + ε−1∂y)(a(y)(∂x + ε−1∂y)(u0 + εu1 + ε2u2 + · · ·)) = f . (2.3)

At the highest order O(ε−2), we have

∂y(a(y)∂yu0) = 0, (2.4)

which has only a y-independent periodic solution. Thus u0 = u0(x). At the next-highest order O(ε−1), we have

∂y(a(y)(∂xu0 +∂yu1)) = 0, (2.5)

which impliesa(y)(∂xu0 +∂yu1) = c(x) (2.6)

for some function c(x). Dividing (2.6) by a and integrating the resulting equationover y ∈ [0,1] yields

ddx

u0 = c(x)〈a−1 〉, (2.7)

where 〈 · 〉 denotes the integral or average over y ∈ [0,1]. At the next order O(1), wehave

∂x(a(y)(∂xu0 +∂yu1))+∂y(a(y)(∂xu1 +∂yu2)) = f . (2.8)

Averaging (2.8) over y ∈ [0,1] gives

∂x⟨a(y)(∂xu0 +∂yu1))

⟩= f ,

which in view of (2.6) is just dc/dx = f . This then becomes, when we insert (2.7),

ddx

(a∗

ddx

u0

)= f , (2.9)

where a∗ = 〈a−1 〉−1 is the harmonic mean of a. Equation (2.9) is the homogenizedequation and is the same type of equation from which we started; however, its co-efficient has been changed to the harmonic mean of the original one in the rapidlyoscillating variable y = ε−1x. Now we have only to solve the large-scale equation(2.9) subject to the same boundary condition, and the small-scale effect has beenbuilt in already.

Rigorous justifications of the above formal asymptotics in any number of dimen-sions are presented in [18] using the energy method and in [78] using the weakconvergence method; see [190] for the first homogenization result in random media

Page 33: Surveys and Tutorials in the Applied Mathematical Sciences

2.2 Reaction–Diffusion Traveling Fronts in Periodic Media 25

(a is a bounded positive random matrix). Equation (2.5) is posed on the periodic do-main in terms of the y variable, and is called the cell problem. Only in one dimensioncan one solve it in closed form; as a result, we know the homogenized coefficientexplicitly. In several dimensions, the corresponding elliptic boundary value problemcan be homogenized, but the homogenized coefficients are not known explicitly ingeneral.

2.2 Reaction–Diffusion Traveling Fronts in Periodic Media

Now let us consider what happens if we let the reaction–diffusion (R-D) fronts dis-cussed in Section 2.1 pass through a medium with periodic structure. If we modelthe medium with a periodic coefficient, then a model equation for R-D fronts is

ut = (a(x)ux)x + f (u), (2.10)

where a(x) is a positive 1-periodic smooth function and f (u) is a nonlinear functionof one of the five types. Since we expect solutions to behave like fronts, we shouldsee them in the large-space and large-time scaling limit. That is, let us consider(2.10) under the change of variables x→ ε−1x, t → ε−1t, for ε small. The rescaledequation is

uεt = ε(a(ε−1x)uε

x)x + ε−1 f (uε), (2.11)

which resembles a homogenization problem except that there is also a singular pref-actor ε−1 in front of the nonlinear term. We realize that there are two scales presentin this problem. One is the width of the front, and the other is the wavelength of theperiodic medium. The first one is easy to capture if we look at the rescaled formof a traveling front in a homogeneous medium, or U(ε−1(x− ct)). The second onecan be built in as in the homogenization ansatz (2.2). Combining the two ideas, wecome up with the following two-scale ansatz for R-D fronts in periodic media:

uε ∼U(ε−1(x− c∗t),ε−1x)+ · · · , (2.12)

where c∗, the average wave speed, plays the role of a∗ in the homogenization exam-ple shown before. Certainly, we impose periodicity in y = ε−1x, and a 0 or 1 far-fieldboundary condition in s = (x− c∗t)/ε .

Substituting (2.12) into (2.11), we find that U as a function of (s,y) satisfies thePDE

(∂s +∂y)(a(y)(∂s +∂y)U)+ c∗Us + f (U) = 0. (2.13)

If (2.13) has a solution under the boundary conditions

U(s, ·) has period 1, U(+∞,y) = 1, U(−∞,y) = 0, (2.14)

the leading term of (2.12) is actually an exact solution! Recalling that the scalingwas just to motivate ourselves, we see that we could have worked with the original

Page 34: Surveys and Tutorials in the Applied Mathematical Sciences

26 2 Fronts in Periodic Media

equation (2.10) to begin with. The exact traveling front then has the functional formU(x− ct,x), and it was first found and constructed in [240].

Comparing (2.2) and (2.12), we see that the two scales of (2.12) are not neces-sarily separate. In fact, they can be arbitrary, while in (2.2), the two scales are vastlyseparate. In this sense, (2.12) is a general two-scale representation. Also for this rea-son, we end up with a PDE cell problem to solve instead of an ODE cell problem.We will see that what makes (2.12) possible is the nonlinearity f (U), and that theextreme cases when the front width is either much larger or much smaller than thewavelength of the medium are simpler.

It is easy to generalize the above form of traveling front to several spatial dimen-sions. Let us consider an R-D equation of the form

ut = ∇x · (a(x)∇xu)+b(x) ·∇xu+ f (u), u|t=0 = u0(x), (2.15)

where

(A1): a(x) = (ai j(x)), x = (x1,x2, . . . ,xn)∈Rn is a smooth positive definite matrixon Rn, 1-periodic in each coordinate xi;

(A2): b(x) = (b j(x)) is a smooth divergence-free vector field, 1-periodic in eachcoordinate xi, with mean zero.

Equations of the form (2.15) appear in the study of premixed flame propaga-tion through turbulent (random) media [56], where u is the temperature of the com-bustible fluid, b(x) is the prescribed turbulent incompressible (divergence-free) fluidvelocity field with zero ensemble mean, f (u) is the Arrhenius reaction term, anda(x) is taken as a constant matrix. Since the fluid velocity b is given as we solve forthe temperature u, the above problem is called passive, and the traveling fronts arecalled passive fronts. In [56], formal asymptotic analysis suggests that u propagateswith an averaged (effective) speed, also called the turbulent flame speed [56, 203].Turbulence refers to complex random flows involving a wide range of spatial andtemporal scales. Let us first consider periodic media to achieve a good preliminaryunderstanding of effective flame speed. In Chapter 5, we shall give a definitive an-swer to front speeds of (2.15) in random (turbulent) media.

Let us fix a unit vector k ∈ RN and look for a traveling wave (front) moving inthis direction with speed c = c(k). The traveling front is of the form

u(x, t) = U(k · x− ct,x), (2.16)

where the front speed c is an unknown constant depending on k, while U , the frontprofile, satisfies as a function of s = k · x− ct and y = x the boundary conditions

U(−∞,y) = 1, U(+∞,y) = 0, U(s, ·) has period 1. (2.17)

Upon substitution into equation (2.15), we obtain the following traveling-front equa-tion for U = U(s,y) and c :

(k∂s +∇y)(a(y)(k∂s +∇y)U)+b(y) · (k∂s +∇y)U + cUs + f (U) = 0. (2.18)

Page 35: Surveys and Tutorials in the Applied Mathematical Sciences

2.2 Reaction–Diffusion Traveling Fronts in Periodic Media 27

The above form of traveling fronts (2.16) in periodic media and the mathematicalstudy of (2.18) were initiated in the author’s work in the early 1990s on bistable andignition nonlinearities [240, 241, 242, 243], where existence and uniqueness areproved under suitable conditions.

A special case of (2.18) is when a is the identity, b(y) = (b1(y′),0), y′ =(y2, . . . ,yN), and k = (1,0, . . . ,0). Such a vector field b is called shear flow. Thenu = U(x1− ct,x′), x′ = (x2, . . . ,xN), and (2.18) reduces to

∆s,y′U +(c+b1(y′))Us + f (U) = 0, (2.19)

a semilinear elliptic PDE.Equation (2.19) appeared earlier ([27] and references therein) as a model of flame

propagation inside an infinite cylinder (s,y′) ∈ R×D for type-5 nonlinearity. Thecylinder has a bounded cross section D, and the boundary condition on y′ is zeroNeumann, so the cylinder boundary is insulated for heat transfer. Existence anduniqueness of solutions to (2.19) is thoroughly studied for nonlinearities of types 1through 5 in [29, 30, 31].

Interestingly, mathematicians were not alone in thinking about traveling fronts inperiodic media. Theoretical biologists have long been interested in R-D fronts sincethe days of Fisher [91] and Hodgkin and Huxley [167]. An interdisciplinary problemof fundamental importance often draws attention and ideas from different scientificcommunities. Indeed, a different notion of traveling front in periodic media wasproposed by biologists [221] in the mid 1980s. A traveling (pulsating) front is asolution u(x, t) satisfying

u(x, t−L · k/c) = u(x+L, t), ∀(x, t),u(x, t)→ 1 as x · k →−∞, (2.20)u(x, t)→ 0 as x · k →+∞,

where L is the (vector) period of the media, c the front speed. The solution re-peats itself in time L · k/c if it is observed at two points a distance L apart in space.Clearly, u(x, t) = U(k · x− ct,x) is such a front. In [221], formal arguments and lin-earizations at the unstable state u = 0 are made to find approximate solutions in onespatial dimension in the case of a (type-1) KPP reaction. However, error estimatesof approximations are not demonstrated.

Interestingly in the late 1970s, about six years earlier than [221], mathematiciansthen working in the former Soviet Union had already developed a probabilistic func-tional integration method [94, 100] to find the KPP minimal speeds in periodic me-dia of any dimensions. In the mid 1980s, this line of work was published in detail inthe West [95, 96]. Though the work was quickly known in the mathematics commu-nity, apparently the authors of [221] were unaware of it, partly because of the lackof communication across scientific and geographical boundaries at the time.

Likewise, [240, 241, 242, 243] were done without knowledge of [221]. The ana-lytical form (2.16)–(2.18) turns out to be more friendly to work with than a propertyof the time-dependent solution (2.20).

Page 36: Surveys and Tutorials in the Applied Mathematical Sciences

28 2 Fronts in Periodic Media

The probabilistic method [94, 100, 96] relies on the large-deviation technique toanalyze the Feynman–Kac representation of KPP solutions. It leads to a variationalformula for KPP minimal front speeds, and also serves as a rigorous justificationof the formal linearization analysis [221]. In the physics literature, the method oflinearization at an unstable state to determine front speeds is known as the marginalstability criterion (MSC) [210]. It originated in the 1950s from the plasma com-munity [43] and was used by physicists in studying pattern selection in the early1980s [63, 140]. Pattern selection refers to the dynamic selection of a front amonga continuum of front solutions from a class of initial data. KPP is one example ofa pattern-forming system in which dynamic selection is called for. In the case ofhomogeneous media, the works [8, 9] established the MSC of the KPP front speed2√

f ′(0) by the PDE method.The probabilistic method [100, 96] proved that MSC also holds for KPP in in-

homogeneous media. We shall discuss this method in conjunction with periodichomogenization of HJ equations in the next section. Its advantage is that it bypassesthe front profile and goes straight to the front speed. Impressively, it was worked outalso for random media in one spatial dimension [100, 96]. PDE methods are morerobust, and can handle more general forms of equations and nonlinearities, thoughthey are traditionally restricted to deterministic media. We shall see in Chapter 5 thatcombining ideas of the large-deviation and PDE methods is a way to handle equa-tion (2.15) in the random setting in arbitrary dimensions and to solve the turbulentfront speed problem [203, 194] for KPP.

2.3 Existence of Traveling Waves and Front Propagation

Let us state the existence results for bistable and ignition fronts [240, 242].

Theorem 2.1. Let T n be the n-dimensional unit torus and ‖ · ‖Hm(T n) the Sobolevnorm of functions on T n with up to m integrable derivatives. Define a =

∫T n a(x)dx,

and assume that conditions (A1) and (A2) hold.

1. If the nonlinearity f (U) is of type 3 (bistable nonlinearity) with µ ∈ (0, 1

2

), there

is a positive number δcr such that if ‖a(x)−a‖Hm(T n) < δcr, ‖b(x)‖Hm(T n) < δcr,m > n+1, then equation (2.18) has a unique classical solution (U,c) such that0 < U < 1, Us < 0 for all (s,y) ∈ R×T n, and c > 0.

2. If the nonlinearity f (U) is of type 5 (combustion nonlinearity with ignition tem-perature), then for all a and b, equation (2.18) has a unique classical front(U,c) satisfying the same properties.

Here uniqueness means that c is uniquely determined by the coefficients (a,b)and the nonlinearity f (U), and U is unique up to a constant translation in s dueto the translation-invariance of equation (2.18). The threshold phenomenon in thebistable case is because the unequal potential wells of the antiderivative of f (u)(which are essentially the driving force behind front motion) can have effectively

Page 37: Surveys and Tutorials in the Applied Mathematical Sciences

2.3 Existence of Traveling Waves and Front Propagation 29

the same depth due to the influence of periodic media. Front speed is zero, andequation (2.15) has a stationary front solution u = u(x). A similar situation occursin the homogeneous case in which the intermediate zero of f (u) is equal to 1

2 .As in homogeneous media, type (1,2,4) front speeds occupy an interval [c∗,∞),

or the speed spectrum is a continuum. More precisely, we have [20] the followingtheorem.

Theorem 2.2. If reaction nonlinearity f is of type (1,2,4), there exists c∗ > 0 suchthat no solution exists to (2.17)–(2.18) if c < c∗, and a monotone decreasing (in s)solution exists to (2.17) if c≥ c∗.

Variational formulas of front speeds of type (1,3,5) will be discussed later.The next problem is to show that under certain conditions on the initial data, the

time-dependent solutions behave like these special traveling-front solutions. Let usfirst state front propagation results for the bistable and ignition reaction [243].

Theorem 2.3 (Front Propagation). Consider the initial value problem for equation(2.15) with initial data 0 ≤ u0(x)≤ 1. Let f be of type 3 with µ ∈ (

0, 12

)or of type

5 with f ′(1) < 0. Assume in the context of type-3 nonlinearity that a traveling wavesolution U(k · x− c(k)t,x) exists for every unit vector k ∈ Rn. Let s ∈ R and let theplane orthogonal to k be S = y ∈ Rn|y = x− (k · x)k, ∀x ∈ Rn.

I. Suppose the initial date are frontlike: u0(x)→ 0 sufficiently fast as k0 ·x→−∞,and u0(x) → 1 sufficiently fast as k0 · x → −∞, uniformly in S(k0), for somek0 ∈ Rn. Then

limt→∞

u(t,sk0t) =

1, s > c(k0),0, s < c(k0).

II. Suppose the initial data are pulselike: for some unit vector k, u0(x)→ 0 suffi-ciently fast as k0 · x →−∞; u0(x) > µ + η , |k · x| < L, for some positive con-stants η and L (θ replacing µ for f of type 5). Then there is a positive numberL0(η) > 0 such that if L≥ L0, then

limt→∞

u(t,skt) =

1, c(k) < s <−c(−k),0, s < c(k) or s >−c(−k).

The existence, uniqueness, and propagation results above are all based on max-imum principles. The idea is to bound from above and below the exact solutionsby simplified comparison functions, then extract asymptotic information. Let us ex-plain the main ingredients below.

Consider equation (2.18) with a type-5 nonlinearity. Our first observation is thatthe three linear terms there do not form a strongly elliptic operator (such as theLaplacian ∆s,y), since the second derivatives are along directions

(ki,0, . . . ,0,yi,0, . . . ,0) ∈ Rn+1, i = 1, . . . ,n,

which do not cover all n+1 directions. The other derivative along direction

Page 38: Surveys and Tutorials in the Applied Mathematical Sciences

30 2 Fronts in Periodic Media

(1,0, . . . ,0) ∈ Rn+1

is the s derivative of U . Hence if c is not equal to zero, we have a parabolic operator(similar to the heat operator ∂t −∆x). This may sound like trouble, since for thestandard heat equation, we cannot pose a boundary value problem in t.

However, what saves us is that the s direction of the infinite cylinder is not char-acteristic, since it is not orthogonal to all the directions (ki,0, . . . ,0,yi,0, . . . ,0). Theother observation is that (2.18) is translation-invariant in s. The loss of ellipticity isabsent in the shear flow case, or equation (2.19).

Now, do we still have a strong maximum principle for the linear operator in(2.18),

Lu = (∇y + k∂s)(a(y)(∇y + k∂s)u)+b(y)T · (∇y + k∂s)u+ cus, (2.21)

even though it is not strongly elliptic?As long as c 6= 0, the answer is yes, thanks to the parabolic maximum princi-

ple and the periodicity in y. Periodicity helps us to overcome the degeneracy! Forclassical maximum principles, we refer to [198, 226].

Now let us take c =−1 for convenience and prove the following result.

Proposition 2.4. Let u be a classical solution of the differential inequality Lu ≤ 0(Lu ≥ 0) on R×T n. If u achieves its minimum (maximum) at (s0,y0) with s0 finite,then u≡ constant.

Proof. We first treat the special case n = 1,k = 1, in which case we have

Lu = (∂s +∂y)(a(y)(∂s +∂y)u)+b(y)(∂s +∂y)u−us.

For the time being, unfold T into R and regard L as an operator on R2. If we makethe change of variables

s′ =1√2(s− y), y′ =

1√2(s+ y),

then∂s =

1√2(∂s′ +∂y′), ∂y =

1√2(−∂s′ +∂y′), ∂s +∂y =

√2∂y′ .

In terms of (s′,y′), Lu becomes

Lu = 2(auy′)y′ −1√2

us′ +(√

2b− 1√2

)uy′ .

Here L is a standard parabolic operator in (s′,y′), elliptic in y′, and parabolic in s′.By the strong maximum principle for parabolic operators, we see that if u attains itsminimum at some finite point (s′0,y

′0), then

u≡ constant if s′ ≤ s′0,

Page 39: Surveys and Tutorials in the Applied Mathematical Sciences

2.3 Existence of Traveling Waves and Front Propagation 31

oru≡ constant if s− y≤ s0− y0.

By the periodicity of u in y, we see that u≡ constant for all s and y. If n≥ 2, wecan always subject y to an orthogonal transform, i.e., y = Qy′, and then Lu becomes

Lu = (k∂s +QT ∇y′)T a(k∂s +QT ∇y′)u+bT · (k∂s +QT ∇y′)u−us

= (Qk∂s +∇y′)T QaQT (Qk∂s +∇y′)u+bT ·QT (Qk∂s +∇y′)u−us.

Choosing Q such that Qk = e1 = (1,0, . . . ,0) and setting a1 = QaQT and b1 = Qb,we have

Lu = (e1∂s +∇y′)T a1(e1∂s +∇y′)u+bT

1 · (e1∂s +∇y′)u−us.

If we make the change of variables

s′ =1√2(s− y′1), z1 =

1√2(s+ y′1), zi =

1√2

y′i, i≥ 2,

then just as in the case n = 1, we have

Lu = 2∇Tz (a1∇zu)− 1√

2us′ +

√2bT

1 ·∇zu− 1√2

uz1 .

By the strong maximum principle for parabolic operators, if u attains its minimumat some finite point P0 = (s′0,z0), then

u = constant if s′ ≤ s′0,

oru = constant if s− y′1 ≤ s0− y′1,0.

In terms of (s,y), this asserts that u is a constant under some hyperplane that is notorthogonal to the s-axis. The periodicity of u in y implies that u≡ constant for all sand y. The proof is complete. ut

Let us outline the two steps of the construction for existence of type-5 solutionsbased on a degree-theoretic approach. In step one, we consider a family of ellipti-cally regularized problems (ε > 0, τ ∈ [0,1]),

εUss +LτU + τ f (U) = 0, (s,y) ∈Ωa = [−a,a]×T n, (2.22)

subject to the boundary conditions U(−a,y) = 1, U(+a,y) = 0. The operator Lτ is Lwith a replaced by 〈a〉(1− τ)+ τa and b replaced by τb, with 〈 · 〉 being the periodaverage.

To remove the translation-invariance of solutions, we must also impose a nor-malization condition: maxy∈T n U(0,y) = θ . By the elliptic maximum principle, weknow that U is bounded between 0 and 1 and that Us < 0. Elliptic regularity also tells

Page 40: Surveys and Tutorials in the Applied Mathematical Sciences

32 2 Fronts in Periodic Media

us that the maximum of ∇U is bounded independently of a and τ . The parameter τlinks the linear problem (τ = 0) with the problem of interest τ = 1.

Consider the space E = C1(Ωa)×R. For (v,c) ∈ E, τ ∈ [0,1], let u = ϕτ(v,c) bethe unique solution of the elliptic boundary value problem

εuss +Lτ u+ τ f (v) = 0

under the same 0 and 1 boundary conditions. Define

hτ(v,c) = maxy∈T n

s=0

ϕτ(v,c).

Then the solution of (2.22) satisfies

u = ϕ1(u,c), h1(u,c) = θ . (2.23)

Define Fτ(u,c) = (ϕτ(u,c),c− hτ(u,c) + θ), τ ∈ [0,1]. Now the existence of thesolution is the same as the fixed-point problem

F1(u,c) = (u,c).

Notice that the mapping (τ,(u,c)) → Fτ(u,c) from [0,1]× E to E is continuousand compact. Due to the a priori bounds on the solutions and their derivatives, theLeray–Schauder degree of the mapping Id−F is well defined on a bounded closedset of the form

D≡ (u,c) ∈ E,‖u‖C1(Ra) ≤ K, |c| ≤ K,where K some constant larger than the bounds of the solutions. This is because thezeros of Id−F cannot occur on the boundary of the set D. The degree is a measureof the number of zeros counting multiplicity, and is invariant under a change ofτ ∈ [0,1]; see [254] for details. If the degree is nonzero, then we have a fixed point.This is easily checked when τ = 0, since (2.23) is explicitly solvable, and we findthat the degree is equal to one.

In step two, we pass to the limit a → ∞ first and then to the limit ε → 0. Tothis end, the main technical work is to bound the wave speed c away from 0 and∞ independently of both parameters. This can be achieved with the help of com-parison principles of wave speeds for the a → ∞ limit, see [241], and the identityc =−∫

R×T n f (U) for the ε → 0 limit; see [242].Thanks to the normalization condition and Us ≤ 0, we have U ≤ θ if s≥ 0. Hence

we have a linear equation for U on s ≥ 0. We can now look for a special decaysolution of the form U = eµsψ(y) with ψ(y) > 0 and µ < 0. This decay solution hasa continuous limit as ε → 0 along a subsequence, and limsupε→0 µ < 0. It followsthat the limiting solution must decay to zero as s→+∞. As s→−∞, monotonicityimplies U(s,y)→U−. It is not hard to show that U− satisfies the elliptic equation(dropping s derivatives from (2.18)

∇y · (a(y)∇yU)+b(y) ·∇yU + f (U) = 0 (2.24)

Page 41: Surveys and Tutorials in the Applied Mathematical Sciences

2.3 Existence of Traveling Waves and Front Propagation 33

under periodic boundary conditions. Since f (U) ≥ 0, the maximum principle im-plies that (2.24) has only constant nonnegative solutions. Thus U− equals either θor 1. In the former case, U ≤ θ , and hence f (U) ≡ 0, for any (s,y). So LU = 0,for all (s,y), and thus U attains its maximum θ at a finite point (0,y?) as a result ofimposing a normalization condition at s = 0.

By the strong maximum principle property of the operator L in Proposition 2.4,U must be identically equal to a constant, which is impossible since it has a zerolimit at s = +∞. We have constructed a desired traveling-front solution with theproperty Us < 0 (strict inequality again follows from the strong maximum principleof L).

The other bonus of the strong maximum principle of L is that the sliding domainmethod [30, 145] applies to show that traveling-front solutions to (2.18) must beunique. The uniqueness means that there is only one value of the wave speed c forany given coefficients (a,b) and nonlinearity f of type 5. Moreover, the profile U isunique up to a constant translation in s, and is strictly monotone in s.

The basic argument to show monotonicity is as follows. First, we compare U(s,y)and its translate Uλ = U(s−λ ,y). For large λ , Uλ is larger than U for those points(s,y) in a bounded cylinder. The bounded cylinder is large enough that U(s,y) isclose to either 0 or 1 outside of it. Then wλ ≡ U(s− λ ,y)−U(s,y) satisfies thedifferential inequality Lwλ ≤ 0 outside of the finite cylinder. The strong maximumprinciple for L implies that wλ > 0 holds at any point. Then we decrease λ to theinfimum value λ0 at which Uλ is no less than U . Now wλ0 ≥ 0.

Again, the strong maximum principle implies that at λ0, U and Uλ must be iden-tical, which is possible only if λ0 = 0, due to the front boundary conditions. Weconclude that U is strictly monotone and actually has positive derivative anywhere(by invoking the minimum principle on the derivative). A similar argument can becarried out for any two profiles to show that they agree up to a constant translation,and for the uniqueness of c, see [241].

The other nonnegative f of type (1,2,4,5) can be approximated by a sequenceof ignition nonlinear functions fθn , where θn ↓ 0. Let χθ (θ < 1

2 ) be a smooth com-pactly supported function such that χθ (u) = 0 if u ≤ θ , and χθ (u) = 1 if u ≥ 2θ .Defining fθ = χθ f will do. The speed cθ is monotone in θ . One may then pass theapproximate type-5 solutions to the limit and verify that the limit remains a frontsolution. The minimal speed c∗ is equal to limn→∞ cθn . One then uses the corre-sponding solution U∗ to prove the existence of other solutions for c > c∗, and theconverging sequence (Uθn ,cθn) to exclude solutions for c < c∗; see [20] for completeproofs. The KPP minimal speed is related to the existence of a positive solution ofthe form eλ sψ(y) to the linearized equation at U = 0, which will be presented laterin a more general context.

For type-3 nonlinearity, new difficulties come up in step two. Because f changessign, there can be many nontrivial periodic solutions of (2.24). As we know, trav-eling fronts may not exist for all a(y) due to the existence of steady states. Theconvenient method for establishing traveling waves in type 3 (the bistable case) is touse the method of continuation [242, 244] and treat a family of problems in whicha(y) is replaced by (1−δ )〈a〉+δa(y).

Page 42: Surveys and Tutorials in the Applied Mathematical Sciences

34 2 Fronts in Periodic Media

We start with δ small and obtain solutions by perturbing the known one-dimensional front. The linearized operator has a simple eigenvalue at zero, and therest of the spectrum is isolated away from zero. The monotonicity of the perturbedsolutions guarantees that the same spectral property of the linearized operator re-mains, and so the perturbation continues on δ . Since each perturbative step relies onthe contraction mapping principle, there is no difficulty as |s| → ∞. Of course, thesame difficulty arises if we want to show that the continuation goes to any value ofδ ∈ [0,1], which we know is false in general.

The continuation method is convenient in that it deals with the problem on theinfinite domain, where estimates of solutions are usually simpler. However, it relieson good spectral properties of the linearized operators. It works for nonlinearity oftype 3, also type 5 if f ′(1) < 0, as well as spatially periodic conservation laws [244].For type 5, the assumption f ′(1) < 0 can be removed by further approximationof nonlinearity. To summarize, the degree-theoretic method and the continuationmethod with the help of maximum principles guarantee the existence of traveling-front solutions as stated.

Let us sketch the proof of statement (I) of Theorem 2.3 in the case of f of type 5,and refer to [243] for the complete proof. The proof is similar for type 3. The ideais to construct subsolutions (supersolutions) using the parabolic maximum princi-ple [198, 226] and the traveling wave solutions. The long-time asymptotics of thesubsolutions (supersolutions) rely on the decay property of solutions of the variable-coefficient linear parabolic equations of the form

ut = ∇ · (a(x)∇u)+b(x) ·∇u, ∇ ·b(x) = 0. (2.25)

The fundamental solution (Green’s function) of (2.25), in turn, has pointwise lowerand upper bounds in terms of heat kernels [169, 84, 186].

First we note that due to the fast convergence of u0 to 1 as k · x → ∞, thereare a number ξ0 > 0 large enough and a positive spatially decaying function q0 =q0(k · x) < (1−θ)/2 such that

U(k · x−ξ0,x)−q0(k · x)≤ u0(x)

on Rn. Now consider the function

ul ≡U(k · x− c(k)t−ξ1(t),x)−q1(t,x),

where ξ1 and q1 will be chosen to satisfy

ξ ′1(t) > 0, ξ1(t) > 0, ξ1(t) = o(t), t → ∞.

We calculate

N[ul ] = ul,t −∇x · (a(x)∇xul)−b(x) ·∇xul − f (ul) (2.26)=−ξ ′1(t)Us−q1,t +∇x · (a(x)∇xq1)+b(x) ·∇xq1 + f (U)− f (U−q1).

Page 43: Surveys and Tutorials in the Applied Mathematical Sciences

2.4 KPP Fronts and Periodic Homogenization of HJ Equations 35

There exists δ ∈ (0,θ) sufficiently small that if q∈ [0, 1−θ

2

]and U ∈ [1−δ ,1], then

f (U)≤ f (U−q).

Since 0≤ q≤ q0 < 1−θ2 , we have for U ∈ [1−δ ,1],

N[ul ]≤−ξ ′1(t)Us−q1,t +∇x · (a(x)∇xq1)+b(x) ·∇xq1. (2.27)

If U ∈ [0,δ ], then f (U) = f (U − q1) = 0, so (2.27) holds with an equality sign. IfU ∈ (δ ,1−δ ), then there exists β > 0 such that Us ≥ β and | f (U)− f (U −q1)| ≤Kq1 for some K > 0. It follows that

N[ul ]≤−ξ ′1β −q1,t +∇x · (a(x)∇xq1)+b(x) ·∇xq1 +Kq1. (2.28)

Let us choose q1 to satisfy the equation

q1,t = ∇x · (a(x)∇xq1)+b(x) ·∇xq1, q1|t=0 = q0(k · x). (2.29)

To make ul a subsolution or N[ul ]≤ 0, we just need to impose the condition

−ξ ′1β +Kq1 ≤ 0, or −ξ ′1β +K‖q1‖L∞(Rn) = 0,

or

ξ ′1 =K‖q1‖L∞(Rn)

β> 0, (2.30)

with ξ1(0) = ξ0 > 0. By our early comments on the fundamental solution of (2.29),‖q1‖L∞ = o(1) as t → ∞. Therefore ξ1(t) = o(t). We have shown that ul is a sub-solution, ul ≤ u. A supersolution can be constructed in a similar way. We concludethat statement (I) holds.

2.4 KPP Fronts and Periodic Homogenization of HJ Equations

Consider the KPP front u(x, t) =U(k ·x−c∗(k)t,x). Then under the hyperbolic scal-ing t → ε−1t, x → ε−1x, the scaled solution uε(x, t) = U((k · x− c∗(k)t)/ε,x/ε)converges to a step function traveling at speed c∗(k) in the direction k. This way offinding the speed by scaling limit can be done directly on the PDE. To fix ideas, letus consider the homogeneous medium in one spatial dimension. The scaled equationis

uεt =

ε2

uεxx + ε−1 f (uε), (2.31)

where we have modified the diffusion constant to 12 for convenience of stochastic

representation. The initial condition is the indicator function 1G0(x), where G0 is an

open interval. Our goal is to recover c∗ =√

2 f ′(0) from (2.31).

Page 44: Surveys and Tutorials in the Applied Mathematical Sciences

36 2 Fronts in Periodic Media

Let c(u) = u−1 f (u). Then equation (2.31) can be regarded as a heat equationwith a time-dependent potential c. The solution has a well-known stochastic repre-sentation formula and the Feynman–Kac formula [96, Chapter 2], [55, Chapter 3]

uε(x, t) = Exg(Xεt )exp

ε−1

∫ t

0c(u(t− s,Xε

s ))ds

, (2.32)

where Xεt = x+

√ε Wt , where Wt is the standard Wiener process. Since 0 < uε ≤ 1,

it follows from the KPP assumption of f that

uε(x, t)≤ Exg(Xεt )exp

ε−1

∫ t

0c(0)ds

= e f ′(0)t/ε P(Xε

t ∈ G0). (2.33)

If we denote the distribution of√

ε Wt by Pε , then it follows from the properties ofthe standard Wiener process that Pε

law→ δx, where δx is the measure with unit massconcentrated at the function identically equal to x. The covariance of Xε − x equalsε min(s, t). There is, however, an exponentially small probability that Xε may escapefrom being close to x. These are called rare events. The large-deviation theory [229]studies the asymptotics of such small probabilities. A family of probability mea-sures Pε on a complete separable metric space X is said to obey the large-deviationprinciple (LDP) with a rate function I(·) if there exists a function I(·) : X → [0,∞]satisfying:

1. 0≤ I(x)≤ ∞, ∀x ∈ X .2. I( ·) is lower semicontinuous.3. For each l < ∞, the set x : I(x)≤ l is compact in X .4. For each closed set C ⊂ X ,

lim supε→0

ε logPε(C)≤− infx∈C

I(x). (2.34)

5. For each open set G⊂ X ,

lim infε→0

ε logPε(G)≥− infx∈G

I(x). (2.35)

If x = 0, the scaled Wiener process Xε(s), s ∈ [0,1], satisfies LDP with the ratefunction I = I(g) defined for any continuous function g on [0,1] with g(0) = 0 as

I(g) =12

∫ 1

0(g′(s))2 ds (2.36)

if g(s) is absolutely continuous with L2 derivative, otherwise I(g) = ∞. See [229,Section 5] for a proof. In view of (2.36), we have

lim supε→0

ε logP(Xεt ∈ G0) = lim sup

ε→0ε logPε(G0) =− inf

ϕ0=xϕt∈G0

∫ t

0|ϕ(s)|2 ds, (2.37)

Page 45: Surveys and Tutorials in the Applied Mathematical Sciences

2.4 KPP Fronts and Periodic Homogenization of HJ Equations 37

which is a minimal action (cost) equal to −d2(x,G0)/2t, where d is the distancefunction. It follows from (2.33) that

limε→0

ε loguε(x, t)≤ f ′(0)t− d2(x,G0)2t

≡V. (2.38)

Clearly,limε→0

uε(x, t) = 0 ∀(x, t) ∈ N ≡ (x, t) : V (x, t) < 0. (2.39)

The function V (x, t) is continuous, and the convergence is uniform on compact sub-sets. Setting V (x, t) = 0 gives the front equation d(x,G0) =

√2 f ′(0)t and the de-

sired front speed c∗ =√

2 f ′(0). One verifies by direct calculation that the functionV (x, t) satisfies the HJ equation

ψt −ψ2x /2− f ′(0) = 0, (2.40)

and initial data ψ(x,0) = 0 if x ∈ G0, ψ(x,0) =−∞ otherwise.It remains to show that uε → 1 if V (x, t) > 0, or that uε(x, t) ≥ 1− λ , on any

compact subset of P = (x, t) : V (x, t) > 0 for any small positive number λ . Weneed a more general Feynman–Kac formula with stopping times installed [96]:

uε(x, t) = Et,x uε(tτ ,Xετ )exp

ε−1

∫ τ

0c(uε(t− s,Xε

s ))ds

= Et,x 1τ=τ1 uε(tτ1 ,Xετ1

)exp

ε−1∫ τ1

0c(uε(t− s,Xε

s ))ds

+Et,x 1τ=τ2 uε(tτ2 ,Xετ2

)exp

ε−1∫ τ2

0c(uε(t− s,Xε

s ))ds

, (2.41)

where τ1, τ2, and τ are given by

τ1 = inf

s : uε(t− s,Xεs )≥ 1−λ

,

τ2 = inf

s : V (t− s,Xεs ) = 0

,

τ = min(τ1,τ2).

A stopping time is a random variable whose value depends only on the informationknown up to this value of time, also called the hitting time (the first time that someevent occurs). The minimum of two stopping times is also a stopping time. See [72,Section 3.1] for more examples. The expectation takes a double subscript to meanthat it acts on the vector trajectory (t − s,Xε

s ). Since c ≥ 0, the first term on theright-hand side of (2.41) is bounded from below by

(1−λ )Eεt,x1τ=τ1 = (1−λ )Pε

t,x(τ = τ1).

For the second expectation term in (2.41), we need to control uε on V = 0 so that itis not too small and the exponential of the integral can balance it out. Then

uε ≥ (1−λ )Pεt,x(τ = τ1)+Pε

t,x(τ = τ2)≥ 1−λ ,

Page 46: Surveys and Tutorials in the Applied Mathematical Sciences

38 2 Fronts in Periodic Media

and we would be done.Note that over s ∈ [0,τ2], we have uε ∈ (0,1−λ ], so c(uε) ≥ minu∈[0,λ ] c(u) ≡

cλ > 0, and the exponential term indeed provides growth of order O(exphcλ /ε),where h ≤ τ2 is a positive number almost surely independent of ε . This is becauseit takes a positive amount of time for ((t − s),Xε

s ) to leave (t,x) where V > 0 toreach a point where V = 0. A lower bound of uε on the interface V = 0 of orderO(exp−δ/ε) for small δ suffices. This lower bound requires V (x, t) to satisfy acondition that for (x, t) ∈ N,

V (x, t) = sup

f ′(0)t−∫ t

0|ϕ(s)|2 ds : ϕ0 = x,ϕt ∈ G0,(t− s,ϕs) ∈ N,s ∈ (0, t)

,

(2.42)for any t > 0. Condition (2.42) says that V (x, t) is the supremum of the action func-tional over the paths in the region of V < 0. The lower bound of uε on V = 0 isobtained by conditioning the stochastic path Xε in formula (2.32) near the optimalpath in region V < 0. The probability of the conditioning is controlled by the ac-tion function in (2.42), whose value over the optimal path is close to zero and socan be bounded from below by −δ . The lower bound on uε on V = 0 of the formO(exp−δ/ε) holds.

Condition (2.42) is valid for the function V in (2.38). Provided that (2.42) con-tinues to hold, the above argument extends to slowly varying media in higher di-mensions, for example when f = f (x,u) > 0 for u ∈ (0,1), f (u,x) < 0 for u < 0and u > 1, fu(x,0) = sup0<u≤1 u−1 f (u,x). See [96] for complete results. The large-deviation method motivates the ansatz

uε ∼ exp−I(x, t)

ε

, (2.43)

and the PDE approach [81, 82, 83] based on the logarithmic change of variablevε =−ε lnuε . Let f (u) = u(1−u). Then the function vε satisfies the equation

vεt =

ε2

vεxx−

12

∣∣vεx∣∣2 + exp

−vε

ε

− f ′(0),

where

vε(x,0) = 0, x ∈ G0; vε(x, t)→+∞, as t ↓ 0+, x ∈ Gc0. (2.44)

The next step is to pass to the limit ε → 0 for vε . Comparison functions and max-imum principles imply that the supremum norm and the Holder norms (with expo-nent α ∈ (0,1)) of vε are bounded in any space–time compact set. Hence vε has auniformly convergent subsequence with limiting function v.

The function v satisfies the variational inequality

min[vt + |vx|2/2+ f ′(0),v

]= 0, x ∈ R, t > 0. (2.45)

This is understood as follows. Fix T > 0. If v≥ 0, then

Page 47: Surveys and Tutorials in the Applied Mathematical Sciences

2.4 KPP Fronts and Periodic Homogenization of HJ Equations 39

vt + |vx|2/2+ f ′(0)≥ 0, (x, t) ∈ R× (0,T ], (2.46)

and on the set v > 0∩R× (0,T ],

vt + |vx|2/2+ f ′(0) = 0, (2.47)

both in the viscosity sense. The viscosity sense in (2.46) means that for each smoothfunction ϕ , if u−ϕ has a local minimum at (x0, t0) ∈ R× (0,T ], then

ϕt(x0, t0)+12|ϕx(x0, t0)|2 + f ′(0)≥ 0. (2.48)

In (2.47), we have in addition that if v−ϕ has a local maximum at (x1, t1) ∈ Rn×(0,T ] and if v(x1, t1) > 0, then

ϕt(x1, t1)+12|ϕx(x1, t1)|2 + f ′(0)≤ 0. (2.49)

To see (2.45), we know that v ≥ 0 by the maximum principle. Equation (2.44)implies the inequality

vεt −

ε2

vεxx +

12|vε

x |2 + f ′(0)≥ 0,

which yields (2.46) as ε → 0 in the viscosity sense. Also on any compact subsetof v > 0, bexp−ε−1vε → 0; hence we have (2.47). The solution to equation(2.47) differs from that of (2.40) by a sign, and the solution to (2.45) can be writtenas v = max(−V,0). In more general slowly varying media, the solution v in thevariational inequality (2.45) with initial condition v = 0 on G0, v = +∞ on Gc

0 admitsa representation in terms of a two-player, zero-sum differential game with stoppingtimes [83, 92], which resembles the action functional in (2.42).

We see from the above analysis that the HJ equation (2.47) or (2.40) carries theinformation on the KPP front speed. Let us exploit this KPP–HJ connection furtherand consider KPP fronts with minimal speed in periodic media by studying theequation

ut =12

n

∑i, j=1

ai j(x)uxi,x j +n

∑i=1

bi(x)uxi + f (x,u), (2.50)

with KPP nonlinearity f (x,u) and initial data g(x) of compact support G0. The prob-lem is solved in [100, 96] by the large-deviation method and a path integral repre-sentation of solutions as we illustrated above.

Again, the nonlinearity f (x,u) can be approximated by c(x) = fu(x,0) times u, sothat the implicit solution formula becomes explicit, and the large-deviation methodyields the long-time front speed. Here we state the result [96].

Theorem 2.5. Let z ∈ Rn. Define the operator

Page 48: Surveys and Tutorials in the Applied Mathematical Sciences

40 2 Fronts in Periodic Media

Lz =12

n

∑i, j=1

ai j(y)(∂yi − zi)(∂y j − z j)+n

∑i=1

bi(∂yi − zi)+ c(y) (2.51)

on 1-periodic functions in y ∈ T n, the n-dimensional unit torus. Let λ = λ (z) be theprincipal eigenvalue of Lz, which can be shown to be convex and differentiable in z.Let H(y) be the Legendre transform of λ ,

H(y) = supz∈Rn

[(y,z)−λ (z)],

y ∈ Rn. The function H(y) is also convex and differentiable. Then for any closedF ⊆y : H(y) > 0, we have limt→∞ u(t, ty) = 0 uniformly in y∈F. For any compactK ⊆ y : H(y) < 0, we have limt→∞ u(t, ty) = 1 uniformly in y ∈ K.

It follows that the asymptotic front speed v = v(e) along the unit direction esatisfies H(ve) = 0. If minRn λ (z) > 0, then the H equation can be solved to yieldthe KPP speed variational formula

v = v(e) = inf(e,z)>0

λ (z)(e,z)

. (2.52)

In fact, λ (z) grows quadratically in z, and so the supremum in the definition of H(y)is achieved. There exists z∗ such that

0 = H(ve) = v(e,z∗)−λ (z∗),

and (e,z∗)> 0 due to λ (z∗)> 0. It follows that v = λ (z∗)/(e,z∗)> 0 and λ (z)(v(e,z∗)−λ (z∗)) = 0≥ λ (z∗)(v(e,z)−λ (z)), implying

λ (z)(e,z)

≥ λ (z∗)(e,z∗)

.

This implies formula (2.52). The assumption minRn λ (z) > 0 holds if the operatorL is self-adjoint or of the form L = ∇ · (a(x)∇·)+b(x) ·∇·, where b is a mean-zeroincompressible velocity.

Instead of going through the large-deviation method, let us follow the spirit of thelogarithmic transform in the PDE approach and derive the same result. First considerequation (2.50) under the scaling x→ ε−1x, t → ε−1t. The rescaled equation reads

uεt =

12

εn

∑i, j=1

ai j(ε−1x)uεxi,x j

+n

∑i=1

bi(ε−1x)uεxi

+ ε−1 f (ε−1x,uε), (2.53)

for which we make the change of variable

uε = expε−1vε. (2.54)

Then vε satisfies the equation

Page 49: Surveys and Tutorials in the Applied Mathematical Sciences

2.4 KPP Fronts and Periodic Homogenization of HJ Equations 41

vεt =

ε2

n

∑i, j=1

ai j(ε−1x)vεxi,x j

+12

n

∑i, j=1

ai j(ε−1x)vεxi

vεx j

+n

∑i=1

bi(ε−1x)vεxi

+f (ε−1x,uε)

uε .

(2.55)The last term is bounded from above by c(ε−1x) = fu(ε−1x,0), which also happensto be the right approximation of the nonlinearity for small values of uε . For locatingthe front or the region where uε is near zero, one can replace the nonlinear term byits linearization at uε equal to zero as we approach the front from the interior wherevε < 0. Then equation (2.55) becomes the periodic homogenization problem of aviscous Hamilton–Jacobi equation.

The periodic homogenization of the inviscid Hamilton–Jacobi equation was firststudied in [148]. Let vε be a solution of

vεt +H(∇vε ,ε−1x) = 0, x ∈ Rn× (0,+∞), (2.56)

with initial data vε(x,0) = v0, where H is periodic in the second variable, say withperiod 1. Under the conditions that H is locally Lipschitz in all variables, H(p,x)→+∞ as |p| →+∞ uniformly in x ∈Rn, u0 is bounded and uniformly continuous, and∇v0 ∈ L∞(Rn); the solution vε converges uniformly on compact sets to the viscositysolution v of the homogenized Hamilton–Jacobi equation

vt +H(∇v) = 0, x ∈ Rn× (0,+∞), (2.57)

where the homogenized Hamiltonian is defined through solving the cell problemstated below.

Theorem 2.6. For each p∈Rn, there exits a unique real number H(p) such that theequation H(p+∇w,y) = H(p) has a 1-periodic viscosity solution w = w(y).

The solution vε has the two-scale expansion

vε ∼ v0(x, t)+ εv1(x,ε−1x, t)+ · · · , (2.58)

implying to leading order upon substitution in (2.56) that

v0,t +H(∇xv0(x, t)+∇yv1(x,y, t)) = 0, (2.59)

which leads to the cell problem in Theorem 2.6, where y = x/ε is the variable, (x, t)are parameters. Equation (2.59) is a nonlinear eigenvalue problem, producing thecell problem in terms of the variable y, and the homogenized equation (2.57) in thevariables (x, t).

The ansatz (2.58) is utilized in the convergence proof of [148]. For generaliza-tions to fully nonlinear first- and second-order equations, see [79], where a weakconvergence method called the perturbed test function method is employed. Such amethod incorporates the above ansatz in the structures of the test functions instead,and can handle equations of first and second order in a unified way.

The homogenized Hamiltonian H is convex if H is in p, but it may lose strictconvexity. One example [148] is that the homogenized Hamiltonian H of the strictly

Page 50: Surveys and Tutorials in the Applied Mathematical Sciences

42 2 Fronts in Periodic Media

convex classical Hamiltonian H(p,x) = p2/2 +V (x) is flat near p = 0. In fact, letV ≤ 0 and maxV = 0. The cell problem reads

12(p+wy)2 +V (y) = H, y ∈ T 1,

which is solvable and gives H ≥ 0 such that

H = 0 if |p| ≤⟨√−2V

⟩,

|p|=⟨√

2H−2V (y)⟩

if |p|>⟨√−2V

⟩, (2.60)

where 〈 · 〉 denotes the average over one period.Now we return to equation (2.55) with c(ε−1x) in place of the last nonlinear

term. Using the above homogenization ansatz (2.58), it is straightforward to derivethe cell problem

H =12

n

∑i, j=1

ai j(y)wyiy j +12

n

∑i, j=1

ai j(y)(pi +wyi)(p j +wy j)+n

∑i=1

bi(pi +wyi)+ c(y),

(2.61)

where we solve for a periodic function w and a real constant H for given p. The ho-mogenized equation is vt −H(∇v) = 0. The cell problem (2.61) can be transformedinto a linear eigenvalue problem with H the principal eigenvalue. To see this, letw = ew > 0. Then (2.61) in terms of w reads

Hw =12

n

∑i, j=1

ai jwyiy j +n

∑i, j=1

ai j piwy j +n

∑i, j=1

bi(piw+ wyi)

+12

n

∑i, j=1

ai j pi p jw+ c(y)w. (2.62)

The right-hand-side operator in (2.62) is just L−p, in view of (2.51). Hence H(−z) =λ (z).

To derive the front speed formula (2.52), consider the Hamilton–Jacobi equation

vt −H(∇v) = 0

with initial condition

v0(x) =

0 if x ∈ G0,

−∞ otherwise.

The Hopf formula is

v(x, t) =− infy∈G0

H∗(

y− xt

), (2.63)

Page 51: Surveys and Tutorials in the Applied Mathematical Sciences

2.4 KPP Fronts and Periodic Homogenization of HJ Equations 43

where H∗ is the Legendre transform of H. The function H(y) in the large-deviationapproach is related to H∗ by

H(y) = sup−z∈Rn

[(y,−z)−λ (−z)] = supz∈Rn

[(−y,z)−H(z)] = H∗(−y).

The points (x, t) where v < 0 or limε→0 uε = 0 then satisfy

H∗(

y− xt

)> 0, ∀y ∈ G0.

Since G0 is compact, we can take both x and t large compared with the size of G0.Then we drop y to get the condition

H∗(−xt

)= H

(xt

)> 0,

implying that the front speed v(e) along direction e satisfies H(v(e)e) = 0.Putting the homogenization ansatz (2.58) into (2.54) shows that for KPP fronts

in periodic media, the solution uε behaves like

uε(t,x) = exp−I(t,x,ε)/ε+ · · · ,I(t,x,ε) = I0(t,x)+ εI1(t,x,x/ε)+ · · · , (2.64)

where I can be regarded as a phase function as in a geometric optics (Wentzel–Kramers–Brillouin (WKB)) ansatz [237]. However, for fronts of type 3 and type 5,the ansatz for uε in the same scaling (x→ ε−1x, t → ε−1t) is

uε(t,x) = U(ϕ(t,x,ε)/ε,x/ε)+ · · · ,ϕ(t,x,ε) = ϕ0(t,x)+ εϕ1(t,x)+ · · · , (2.65)

where ϕ(t,x,ε) is the phase variable. Plugging (2.65) into (2.50), we have

12(∇xϕ0∂s +∇y)(a(y)(∇xϕ0∂s +∇y)U)+b(y) · (∇xϕ0∂s +∇y)U

−ϕ0,tUs + f (U) = 0, (2.66)

where U = U(s,y), s = ϕ(t,x,ε)/ε , y = x/ε . We see that (2.66) is just the traveling-front equation (2.18) with k = ∇xϕ0, and c(k) = −ϕ0,t . Relating them gives theHamilton–Jacobi equation

ϕ0,t + c(∇xϕ0) = 0 (2.67)

for the general front evolution. By uniqueness of c in the case of type (3,5), itis seen from (2.18) that c = c(k) is homogeneous of degree 1 in k, so c(∇xϕ0) =|∇xϕ0|c(νn), where νn = ∇xϕ0/|∇xϕ0| is the normal direction of the level set of ϕ0.The effective Hamiltonian of type-(3,5) nonlinearities is anisotropic and has lineargrowth in |p|. In contrast, the effective Hamiltonian H of KPP in (2.62) is quadraticin p.

Page 52: Surveys and Tutorials in the Applied Mathematical Sciences

44 2 Fronts in Periodic Media

Interestingly, there are mechanical analogies of quadratically and linearly grow-ing Hamiltonians. The Hamiltonian of classical mechanics H = |p|2/2 +V (x) isquadratic in |p|. The Lagrangian of a special relativistic particle of mass m in ascalar potential [103, 139] is

L(q,x) =−mc2√

1−|q|2/c2−V (x), |q| ∈ [0,c],

where c is the speed of light. The corresponding Hamiltonian is

H(p,x) = mc2√

1+ |p|2/c2 +V (x),

which has linear growth in |p|. Bistable and ignition-type fronts belong to the familyof special relativity, while KPP fronts are Newtonian.

2.5 Fronts in Multiscale Media

The study of traveling fronts in heterogeneous media has been an active area ofresearch in recent years. One may find that other equations also have periodicallyvarying traveling waves (2.16), and one may also consider more complicated mediaarising in applications, such as space–time-periodic media, time- or space-almost-periodic media, or more general and complex media. We shall present some of theseextensions here. An interesting trend is that extended front equations become moreand more degenerate if we continue to construct their time dependence explicitly(e.g., constant-speed motion), and that one must use more general dynamic variablesto capture these fronts. More complicated media make more complicated fronts.

Recall the solute transport equation (1.1) in the introduction of Chapter 1. In onespatial dimension, v is a constant, and equation (1.1) simplifies after a rescaling ofconstants to

(u+ k(x)up)t = (D(x)ux)x−ux, (2.68)

where we also make D spatially dependent. We consider the boundary conditionsu(−∞, t) = ul , u(+∞, t) = ur = 0, 0 < ul , representing constant input of solute fromthe left end of a solute-free soil column. Solutions of (2.68) under such boundaryconditions give rise to front solutions.

If k and D are constants, then by making the change of variable v = u+ kup, wecan write (2.68) as a standard conservation law:

vt +( f (v)− (g(v))x)x = 0, x ∈ R. (2.69)

Front solutions v = v(x−ct) are solvable in closed form, and c =( f (ul)− f (ur))/(ul−ur) is the so-called Rankine–Hugoniot relation.

Let us now consider periodic media by supposing k(x) and D(x) to be 1-periodicregular functions. In periodic media, just as in reaction–diffusion equations, travel-

Page 53: Surveys and Tutorials in the Applied Mathematical Sciences

2.5 Fronts in Multiscale Media 45

ing fronts take the form u =U(x−ct,x), which turn out to exist also for conservativeequations such as (2.68) and are asymptotically stable [244, 245]:

Theorem 2.7. Let k(x) and D(x) be smooth positive functions with period 1. If ur =0 < ul, then equation (2.68) admits a Holder continuous traveling wave solution ofthe form u =U(x−st,x)≡U(ξ ,y), ξ = x−st, y = x, U(−∞,y) = ul, U(+∞,y) = 0,and U(ξ , ·) has period 1. Such solutions are unique up to constant translations inξ , and have wave speeds

s =ul

ul + 〈k〉 f (ul)> 0, (2.70)

with 〈k〉 the periodic mean. The wave profile U satisfies

0≤U < ul ∀(ξ ,y); U(ξ1,y)≤U(ξ2,y) ∀ξ1 ≥ ξ2,∀y; Uξ < 0 if U(ξ ,y) > 0.

Assume that the initial condition u0(x) satisfies

0≤ u0(x)≤ ul , u0 ∈ L1(R+); up0 ∈ L1(R+), u0−ul ∈ L1(R−), up

0 −upl ∈ L1(R−).

Let also m(u,x) = u+ k(x)up. Then there exists a unique number x0 such that∫

Rm(u0(x),x)−m(U(x+ x0,x),x)dx = 0 (2.71)

and such thatlimt→∞

‖u(t,x)−U(x− st + x0,x)‖1 = 0. (2.72)

The construction of traveling waves uses the continuation method, and the exis-tence result holds also in several spatial dimensions [244]. The Holder continuityof solutions is a consequence of up being nondifferentiable at u = 0. The expliciteffective wave speed (2.70) is due to the fact that equation (2.68) is conservative.Only the mean value of k contributes to the speed; the rest of the information in kinfluences the wave profile.

The stability proof extends that of [188] and uses L1 contraction of dynamics,as well as a space–time translation invariance of the traveling fronts in the movingframe coordinate. For fronts in another conservative equation (the Richards equa-tion of water infiltration) with more complicated dependence of wave speeds on theperiodic media, see [88].

Space–time-dependent media (flows) arise in combustion [10, 65]. KPP fronts ina periodic flow field with space–time-separated scales are studied in [152], whichconsiders the temperature field of a reacting passive scalar,

T εt +V (x, t,ε−α x,ε−α t) ·∇T ε = εκ∆T ε + ε−1 f (T ε), (2.73)

with compactly supported (in G0) nonnegative initial data, and α ∈ (0,1]. The ve-locity V is bounded and Lipschitz continuous and has periodic dependence on thefast-oscillating scales y ≡ ε−α x,τ ≡ ε−α t. The small parameter ε measures the ra-tio of the front thickness and large scale (dependence on (x, t)) of the velocity field,

Page 54: Surveys and Tutorials in the Applied Mathematical Sciences

46 2 Fronts in Periodic Media

say of order O(1). The effective Hamiltonian H(p,x, t) is defined as a solution ofthe following cell problem: for each (p,x, t) ∈Rn×Rn× (0,+∞) there are a uniquenumber H(p,x, t) and a w(y,τ) ∈C0,1(Rn× (0,+∞)) periodic in both y and τ suchthat

wτ −a(α)κ∆w−κ |p+∇w|2 +V (x, t,y,τ) · (p+∇w) =−H(p,x, t), (2.74)

where a(α) = 0 if α ∈ (0,1), a(α = 1) = 1.The case α = 1 can be derived using an exponential change of variable and a

Hamilton–Jacobi equation as in the last section except that due to the time de-pendence, the wτ term is added. The condition a(α) = 0 in the case α ∈ (0,1)implies the loss of viscosity in the cell problem (2.74), which can be under-stood as follows. Ignore the slow variable (x, t) for now and change the scalingto x = ε−1+α x′. Then the velocity V is V (ε−1x′,ε−1t ′), and the diffusion coefficientbecomes ε3−2α κ ¿ εκ . Hence the diffusion term is too small to be seen at the orderof the cell problem.

The function H is locally Lipschitz continuous, convex in p, and grows quadrat-ically in |p| as |p| → +∞ uniformly in (x, t). The asymptotics of T ε as ε → 0 aregiven by the following theorem.

Theorem 2.8. Let T ε be a solution of (2.73) under the above assumptions. Then asε → 0, T ε → 0 locally uniformly in (x, t) : Z < 0 and T ε → 1 locally uniformlyin the interior of (x, t) : Z = 0, where Z ∈C(Rn× [0,+∞) is the unique viscositysolution of the variational inequality

max(Zt −H(∇Z,x, t)− f ′(0),Z) = 0, (x, t)×Rn× (0,+∞),

with initial data Z(x,0) = 0 in G0 and Z(x,0) =−∞ otherwise. The set Γt = ∂x ∈Rn : Z(x, t) < 0 can be regarded as a front.

Given a space–time-periodic incompressible flow field, the “cell problem” forKPP front speed in the limit t →+∞ is always viscous. To show this, let us consider

ut = ∆u+b(x, t) ·∇u+ f (u), (2.75)

where x ∈ RN , t ∈ R, and f is of KPP type. The N components of the vector fieldb(x, t) := (b1(x, t),b2(x, t), . . . ,bN(x, t)) are smooth and spatially divergence-free,are periodic of period 1 in both x and t, and have mean zero over the period cellQ× (0,1), where Q is the unit cube in RN . Then the KPP large-time minimal frontspeed in direction k (denoted by c∗(k)) is identified by a front propagation (frontspreading) theorem similar to Theorem 2.3. It is given by the variational formula[173]

c∗(k) = infλ>0

µ(λ ,k)/λ ,

where µ(λ ,k) is the principal eigenvalue of the periodic–parabolic operator [117]

Lλ Φ := ∆xΦ +(b−2λk) ·∇xΦ +(λ 2−λb · k + f ′(0)

)Φ−Φt , (2.76)

Page 55: Surveys and Tutorials in the Applied Mathematical Sciences

2.5 Fronts in Multiscale Media 47

defined on spatially–temporally periodic functions Φ(x, t). The eigenvalue problem(2.76) is related to (2.74) at α = 1 by a logarithmic transform. In the case α = 1,taking ε → 0 is the same as t → ∞.

Moreover, there is a traveling-front solution of the form u =U(k ·x−c∗(k)t,x, t)≡U(s,x, t), locally integrable in (s,x, t), periodic in (x, t), U(±∞,x, t) = 0/1, with thecontinuous directional derivatives

Uτ − c∗Us, ki Us +Uyi , i = 1, . . . ,N, and (k∂s +∇y)2 U

and satisfying the traveling-front equation

Uτ − c∗Us = (k∂s +∇y)2 U +b · (k∂s +∇y)U + f (U). (2.77)

Note that (2.77) is an extension of (2.18), and is more degenerate in the sense thatthere are not enough derivatives in (2.77) to ensure continuity (smoothness) of U .The function U has one more dependent variable than u, which does not happenin spatially periodic media. The value c∗ is minimal in that no solutions exist if c∗is replaced by a number c < c∗. Recently, it was proved in [168] that for almostall η ∈ R, we have that U(k · x− c∗(k)t + η ,x, t) satisfies (2.75). Similar existenceresults [173, 168] hold for type (2,4,5). We refer to [168] for further results onexistence of KPP fronts at c > c∗ and continuous KPP fronts.

In view of (2.74) and (2.76), the front speed obtained from the limit t → ∞ ata fixed ε > 0 and that from ε ↓ 0 at a fixed time interval [0,T ] may not agree ingeneral. A numerical study of the difference due to finite ε (finite front thickness)is carried out in [161]. At any finite ε > 0, the cell problem of a front is alwaysviscous, while it is not in the limit ε ↓ 0 when α < 1. This shows the subtlety offront speed upscaling in heterogeneous media. The front speed would be the samefrom either limit in homogeneous media.

Fronts of the form (2.16) persist in space-almost-periodic media [168]. Frontsolutions of the form u = U(k ·x−ct, t) in time-periodic media have been studied in[5] (bistable f ) and [97] (nonnegative f ). Bistable fronts of the form u = U(k · x−ct, t) have been found in time-almost-periodic media. Interestingly, there are alsofronts of the form u = U(k · x +

∫ t0 c(s)ds) where c(s) is almost periodic. The latter

fronts are not reducible to the former. KPP fronts in space-periodic and time-almost-periodic media were studied recently [120], where generalized speed intervals wereproved to exist, and they reduce to singletons in the case of time-periodic media.A more general form of fronts has been introduced recently [21, 159, 219], andproved to exist in various settings [219, 163, 174]. For fronts in periodic media inthe context of discrete models, see [236, 106] and references therein. For frontsand homogenization in the context of free-boundary limits and models, see [46, 47,129, 130, 248, 249]. For fronts in periodically perforated media and fragmentedenvironments among other applications, see [113, 25]. For KPP speeds c∗ undervarious parameter asymptotic limits (diffusion–reaction rates, periods), see [225].For pulselike waves of reaction–diffusion systems in heterogenous media, see [171,252].

Page 56: Surveys and Tutorials in the Applied Mathematical Sciences

48 2 Fronts in Periodic Media

2.6 Variational Principles, Speed Bounds, and Asymptotics

The KPP variational principle (2.52) reduces the front speed problem to the analysisand estimation of the principal eigenvalues of linear advection–diffusion operators,where many classical methods apply. Let us consider (2.75) in space dimension two(N = 2), and scale the velocity field b(x, t) to δb. If δ is small, a perturbation analy-sis of eigenvalues yields the quadratic enhancement law c∗ = c0 +O(δ 2), where c0is the KPP front speed in homogeneous media. The quadratic correction is explicitin the case of shear flow [191, 177]. In the case of spatial shear, write b = (0,b2(x1)),b2 = bx1 . The function b has mean equal to zero and serves as the velocity potential.Then the front speed along the x2 direction has the expansion [191]

c∗ = c0

(1+

12‖b‖2

2 δ 2 +higher-order terms)

, δ ¿ 1. (2.78)

The energy (half of L2 norm square) of b (velocity potential) is the amount of en-hancement to leading order. In the case of bistable nonlinearity, one may performa perturbation analysis of the traveling-front equation (a nonlinear eigenvalue prob-lem) [191]. The interesting finding is that the correction term in (2.78) is the same.By monotonicity of c∗ in terms of f , the value of c∗ from other types of f mustbehave the same. This is the first indication that for front speed c∗, the type of non-linearity does not matter as much as the flow. In other words, there is universality ofc∗ in terms of nonlinearity.

For time-periodic shear flow, let

b2 = b2(x1, t) = ∑m 6=0,l 6=0

bm,leimx1+iωlt .

Then

c∗ = c0

(1+

12

(∑

m>0l>0

|bm,l |2 2m2

m4 + l2ω2

)δ 2 +higher-order terms

). (2.79)

We see that the enhancement decreases with increasing frequency of temporal os-cillations (or as ω increases). The speed slowdown due to temporal oscillations iscalled the speed-bending phenomenon in the combustion literature, and is studiedin various models [10, 128, 65]. It persists in random flows as well [65, 179, 182],which we shall discuss more in Chapter 5. Again formula (2.79) holds for all non-linearities.

In the large δ À 1 regime, consider again spatial shear flow. The eigenvalueanalysis of KPP front speeds [19] shows that c∗(δ )/δ is monotone decreasing inδ À 1 and converges to a positive limit. The limiting value or the linear growth ratedepends on b in an implicit way, and it has a variational formula [115]:

Page 57: Surveys and Tutorials in the Applied Mathematical Sciences

2.6 Variational Principles, Speed Bounds, and Asymptotics 49

limδ→∞

c∗(δ )/δ = supψ∈D1

Ωb(x1)ψ2(x1)dx1, (2.80)

whereD1 =

ψ ∈ H1(Ω) : ‖∇ψ‖2

2 ≤ f ′(0),‖ψ‖2 = 1

,

and Ω is the periodic domain of variable x1. If b has a flat piece near its maximalpoint in Ω , a test function in D1 can be supported near the maximal point, andthe limit equals maxΩ b = ‖b‖∞. If the reaction is fast ( f (u) replaced by r f (u),rÀ 1) or the diffusion constant (equal to one in (2.75)) is made small, the constraint‖∇ψ‖2

2 ≤ f ′(0) is easy to satisfy: again a test function may be localized near themaximal point of b, and so the speed growth rate is close to maxΩ b [12, 24]. Ingeneral,

c∗ = O(δ ), δ À 1, (2.81)

for other nonnegative nonlinearities. Here again we see universal behavior. The lin-ear law (2.81) holds also for time-periodic shear flows [177], and is numericallyobserved for bistable f as well. It is true for more general percolating flows thatcontain at least two infinitely long channels of flow trajectories [59, 132]. The openchannels (streamlines) in the flow are like multiple lanes on the freeway to help thetransport process and speed up the reaction front. The growth exponent is less thanone (sublinear growth) for flows with enough closed streamlines. For example,

c∗(δ ) = O(δ 1/4), δ À 1, (2.82)

holds [12, 184] for the KPP front and cellular flow:

b = (−φx2 ,φx1), φ = cos(πx1)cos(πx2). (2.83)

The proof of the 14 scaling for the KPP front [184] uses the speed variational for-

mula (2.52), boundary layer analysis of cellular flows, and properties of convection-enhanced diffusion [86]. For ignition nonlinearity, (2.82) is supported by numericalsimulations [231]. Moreover, analytical bounds O(δ 1/5)≤ c∗ ≤O(δ 1/4) hold [132].A useful criterion for distinguishing linear and sublinear speed growth is in termsof first integrals of the flow fields [24]. A first integral for a periodic vector field bis a nonzero periodic solution w of the equation b ·∇w = 0. An extension of (2.80)to KPP speed c∗ in a mean-zero divergence-free vector field b(x) is [256]

limδ→∞

c∗(δ ,k)/δ = supw∈DI

T N(b · k)w2(x)dx,

where

DI =

w ∈ H1(T N) : b ·∇w = 0,‖∇w‖22 ≤ f ′(0),‖w‖2 = 1

, (2.84)

where k is the direction of front propagation and T N the N-dimensional unit torus(a unit cube in RN with opposite faces identified).

Page 58: Surveys and Tutorials in the Applied Mathematical Sciences

50 2 Fronts in Periodic Media

It follows that an upper bound is ‖b · k‖∞, and that the limit in (2.84) is nonzeroif

∫T N (b · e)w0 dx 6= 0 for some first integral w0. This is the case for shear flows.In general, suppose such a w0 exists. Then w = (1 + εw0)/‖1 + εw0‖2 ∈ DI for

ε small enough, and the limiting value is to leading order ε∫

T N (b · e)w0 dx, whichis positive if ε is chosen to have the sign of the integral. If

∫T N (b · k)w2 dx ≤ 0 for

all first integrals, then c∗(δ ,k) = o(δ ). The cellular flow (2.83) is an example forwhich

∫T N (b · k)w2 dx = 0 for all k and first integral w.

The front asymptotic enhancement in the sense of limδ→∞ c∗(δ ) = ∞ by periodicincompressible flow has been shown recently [255] to depend on the geometry of theflow and not on nonlinearity f . In particular, front asymptotic enhancement occursfor KPP if and only if it does so for ignition f , and so the phenomenon is universalamong all nonnegative reactions.

A variant of (2.52) holds for partially periodic media where solutions in part ofthe variables are periodic and in the other part are subject to zero Neumann boundaryconditions [20]. There are min–max variational principles of front speeds [109, 116,232] for non-KPP f . In particular, let us state the one for the unique front speedfrom shear flow and bistable/ignition f that has been used in analysis of randomfront speeds [176].

Consider the cylindrical domain x = (x1, x) ∈D =R×Ω , where Ω is a boundeddomain in RN−1, and shear flow b = (b1(x),0). The front moves along x1, u =U(x1 + c∗t, x)), satisfying zero Neumann boundary condition at R1×∂Ω . The ini-tial datum u0 belongs to the set Is. The set Is for bistable f consists of boundedcontinuous functions with limits one and zero at x1 ∼±∞ respectively. For ignitionf , one requires also that u0 decay to zero exponentially at −∞. For u0 ∈ Is, u(x, t)converges to a traveling front at large times [205]. Define the functional as in [116]:

ψ(v) = ψ(v(x))≡ Lv+ f (v)∂x1 v

≡ ∆v+b1(x)∂x1v+ f (v)∂x1 v

. (2.85)

The min–max variational formula [116] for c∗ is

supv∈K

infx∈D

ψ(v(x)) = c(δ ) = infv∈K

supx∈D

ψ(v(x)), (2.86)

where K is the set of admissible functions,

K =

v ∈C2(D) | ∂x1 v > 0, 0 < v(x) < 1,v ∈ Is

.

The proof uses asymptotic stability of traveling fronts and min–max front speedformulations of [232]. Likewise, similar min–max formulas have been derived andstudied for the homogenized Hamiltonian H of HJ in periodic media [57, 104]. In[104], H is computed based on the formula

H(p) = infφ(y)∈C1(T N)

supy

H(p+∇yφ(y),y). (2.87)

Page 59: Surveys and Tutorials in the Applied Mathematical Sciences

2.7 Exercises 51

2.7 Exercises

1. Show that the principal eigenvalue λ (z) of the operator Lz in (2.51) is posi-tive for all z ∈ Rn if (ai j) is the identity matrix and b j(y) is a mean-zero anddivergence-free vector field.

2. Prove by the maximum principle that the homogenized Hamiltonian of KPPnonlinearity H = H(p) defined in (2.62) grows like O(|p|2) for large |p|.

3. Show that for cellular flow (2.83), the integral∫

T N (b ·k)w2 dx is zero for all unitvectors k ∈ R2 and first integral w (b ·∇w = 0).

4. Derive the quadratic speed-enhancement formula (2.78) for bistable and igni-tion fronts with the min–max formula (2.86) in a mean-zero 1-periodic shearflow b = δ (b1(x),0). The fronts move in the x1 direction, x ∈ Rn−1, n ≥ 2. Inthe small-δ regime, define the test function as a perturbation of the traveling-front profile in homogeneous media of the form

v(x) = U(ξ )+δ 2w(ξ , x), (2.88)

whereξ = (1+αδ 2)x1 +δ χ , (2.89)

with α a constant to be determined and χ = χ(x) the mean-zero periodic solu-tion of

−∆xχ = b1.

Choose α properly so that w is uniformly bounded with decay at x1 ∼±∞ (andso v is admissible) and gives the quadratic speed correction in (2.78).

Page 60: Surveys and Tutorials in the Applied Mathematical Sciences

Chapter 3Fronts in Random Burgers Equations

In groundwater and contaminant transport problems arising from environmental sci-ences, fronts typically travel in a spatially inhomogeneous environment because ofthe natural formation of porous structures. Due to lack of field data, the spatiallyinhomogeneous environment is often modeled as a random process. Conservationof mass then leads to a nonlinear scalar conservation law with a random flux,

Ut +( f (U,x,ω))x = 0, (3.1)

or its viscous analogue. Some equations of this form are (1) the Buckley–Leverettequation for two-phase flows [118] and references therein; (2) the contaminanttransport equation [253, 38]; and (3) the Richards equation for infiltration problems[195, 196, 88]. The specific form of the nonlinear and random function f dependson the problem at hand.

One of the fundamental issues discussed in these works is front dynamics inrandom media. In this chapter, we shall analyze a special case in which f is quadraticin U with random multiplicative coefficient, or a Burgers equation with random flux.We find that at large times, front motion obeys the central limit theorem. In otherwords, a front moves at an average deterministic velocity and the fluctuations aroundit obey Gaussian statistics.

3.1 Main Assumptions and Results

We are interested in the long-time behavior of the inviscid Burgers equation with arandom flux,

Ut +(

12

a(x,ω)U2)

x= 0, (3.2)

and for initial data of the front type,

U(x,0) = IR−(x). (3.3)

© Springer Science + Business Media, LLC 2009

J. Xin, An Introduction to Fronts in Random Media, Surveys and Tutorials in the Applied 53Mathematical Sciences 5, DOI: 10.1007/978-0-387-87683-2_3,

Page 61: Surveys and Tutorials in the Applied Mathematical Sciences

54 3 Fronts in Random Burgers Equations

Here IR− denotes the indicator function of the negative real line, and a(x,ω) is astochastic process on the real line satisfying the assumptions A1–A5 stated below.

A1. Stationarity: the finite-dimensional distributions of the process a(x,ω) are in-variant under translations of the variable x.

A2. Positivity: a(x,ω) > 0 with probability one.

A3. Measurability and integrability of the inverse: paths of a are measurable func-tions of x and E [1/a(x)] < ∞. It follows that also

E

[1√a(x)

]def= µ < ∞.

A4. Invariance principle: Let

ξ (x) =∫ x

0

1√a(y)

dy.

Note that ξ (x) < 0 for x < 0. For each x0 > 0, we have(

ξ (tx)−µtxσ√

t

)

|x|≤x0

d→ (Wx)|x|≤x0 , (3.4)

as t → ∞, where W = (Wx)x∈R is the Wiener process and

σ2 = 2∫ +∞

0E

[(1√a(0)

−µ

)(1√a(x)

−µ

)]dx < ∞,

and d→ denotes convergence of processes in law [36]. The finiteness of the last inte-gral is part of the assumption, and σ2 is sometimes called the velocity autocorrela-tion function (of the process 1/

√a).

A5. Regularity: the paths of the process a are Holder continuous with some (pos-itive) exponent. This will be used in the proof of the main theorem to justifytaking the zero-viscosity limit. A well-known probabilistic condition that impliesHolder continuity of sample paths is the Kolmogorov moment condition [202, The-orem 25.2].

A large class of processes for which (3.4) holds is the class of stationary pro-cesses a(x,ω) satisfying the appropriate φ -mixing condition. Here φ is a nonnega-tive function of a positive real variable such that

limt→+∞

φ(t) = 0, (3.5)

Page 62: Surveys and Tutorials in the Applied Mathematical Sciences

3.1 Main Assumptions and Results 55

and the φ -mixing condition says that for any t > 0 and for any s, whenever an eventE1 is in the σ -field generated by the random variables a(x) with −∞≤ x≤ s and anevent E2 is in the σ -field generated by a(x) with s+ t ≤ x≤+∞, we have

|P[E1∩E2]−P[E1]P[E2]| ≤ φ(t)P[E1]. (3.6)

Roughly speaking, because of (3.5), (3.6) expresses a decay of correlations of thevariables a(x). More information on φ -mixing processes can be found in [36], whereit is proved in particular [36, pp. 178–179] that the invariance principle (assumptionA4) holds if

∫ +∞0

√φ(t)dt < ∞.

The Burgers fronts are asymptotically stable for spatially decaying initial pertur-bations [121]. The following main result of the chapter shows that the front structureis also present in the presence of a random flux. Similar results hold for random ini-tial perturbations [233]. Throughout the chapter, the symbol d→ denotes convergencein distribution.

Theorem 3.1 (Random Burgers Fronts). Let 2c = E[a−1/2]−2 denote the squareroot-harmonic mean of the process a(x,ω). Then as t → ∞,

U(αt, t) d→ 0, α > c, (3.7)√

a(αt) U(αt, t) d→√

2c, α < c, (3.8)√a(ct + z

√t)U(ct + z

√t, t) d→ X , (3.9)

where X is a random variable equal to√

2c with probability N(

µ2

σ z)

and equal

to 0 with probability 1−N(

µ2

σ z)

, where N (s) = 1√2π

∫ s−∞ exp−η2/2dη is the

error function.

The first two parts of the theorem say that to leading order, the front speed inthe presence of randomness equals c, or a law of large numbers. The last part of thetheorem says that the front fluctuations around the speed c obey Gaussian statistics,or a central limit theorem. If we view a Burgers front location as a time-dependentrandom variable X(t,ω), then it satisfies the fundamental laws of probability, similarto sums of iid random variables [72].

In the proof of Burgers’ front theorem, we will make use of a regularized equa-tion:

Ut +(

12

a(x) U2)

x= ν

(√a(x)

(√a(x)U

)x

)x, (3.10)

where ν > 0 is a positive parameter that we shall send to zero eventually. It is con-venient to rewrite this equation in terms of the function

u =√

a(x)U. (3.11)

The equation for u becomes

Page 63: Surveys and Tutorials in the Applied Mathematical Sciences

56 3 Fronts in Random Burgers Equations

ut√a(x)

+(u2/2)x = ν(√

a(x)ux

)x. (3.12)

To simplify the last equation, we change the space variable:

ξ =∫ x

0

1√a(x′)

dx′. (3.13)

Since this change of variables depends on the realization of the process a, we obtainin this way a stochastic process ξ (x,ω), which has already been used to state theassumption A4.

The equation for u in the variables (ξ , t) becomes the standard viscous Burgersequation

ut +(u2/2

)ξ = νuξ ξ , (3.14)

with the new initial condition

u(ξ ,0) =√

a(x(ξ ))IR−(ξ ). (3.15)

It is known that the speed of a (shock) front of the Burgers equation is equal toits height divided by two, the Rankine–Hugoniot condition [141]. This then leads toan intuitive explanation. The asymptotic speed of the front arising from our randominitial condition equals one-half of its average height. Calculated in the ξ variable,the speed is

12

limL→∞

1L

∫ 0

−L

√a(x(ξ ))dξ ,

which, after changing the variable of integration to x, gives

12

limL→∞

−x(−L)L

=12

E[a−1/2]−1.

To recover the front speed in the x variable, we divide this value by E[a−1/2] inview of (3.13) and arrive at the speed c in the theorem. A similar but more detailedargument, taking into account fluctuations of the total mass in a finite interval of theinitial data, leads to a heuristic justification of the Gaussian statistics of the frontlocation.

3.2 Hopf–Cole Solutions

The Hopf–Cole formula [237] for u reads

u(ξ , t) =

∫ +∞−∞

ξ−ηt exp

[−G(η ,ξ ,t)

]dη

∫ +∞−∞ exp

[−G(η ,ξ ,t)

]dη

, (3.16)

Page 64: Surveys and Tutorials in the Applied Mathematical Sciences

3.2 Hopf–Cole Solutions 57

where

G(η ,ξ , t) =∫ η

0u(η ′,0)dη ′+

(ξ −η)2

2t= x(η)IR−(η)+

(ξ −η)2

2t

=

E[a−1/2

] + x(η)

)IR−(η)+

(ξ −η)2

2t. (3.17)

The second equality follows by changing the variable to x = x(η ′), as in (3.13),where x denotes the inverse of ξ . We used the fact that the derivative of ξ is 1/

√a.

Define ul = 1/E[a−1/2]. Then clearly,

ul = µ−1 =√

2c. (3.18)

It is convenient to use the suggestive notation ul (the “left state of u”) below.The numerator of (3.16) is equal to

∫ 0

−∞

ξ −ηt

exp[−ulη− x(η)− (2t)−1(ξ −η)2

]dη

+∫ ∞

0

ξ −ηt

exp[−(2t)−1 (ξ −η)2

]dη , (3.19)

which, with the substitution y = ξ −η , becomes

∫ ∞

ξ

yt

exp[−(ξ − y)ul − (2t)−1y2− x(ξ − y)

]dy+

∫ ξ/√

t

−∞ηe−η2/4ν dη

=1t

∫ ∞

ξyexp

[−(ξ − ul

2 t)ul − (2t)−1(y−ult)2− x(ξ − y)2ν

]dy (3.20)

+∫ ξ/

√t

−∞ηe−η2/4ν dη .

Changing variables x′ = y−ult, the numerator becomes

1t

∫ ∞

ξ−ult(x′+ult)exp

[−(ξ − ul

2 t)ul − (2t)−1x′2− x(ξ − x′ −ult)2ν

]dx′

+∫ ξ/

√t

−∞ηe−η2/4ν dη

= ul exp[−

(ξ − ul

2t) ul

]∫ ∞

ξ−ultexp

[−(2t)−1 x′2

2ν− x(ξ − x′ −ult)

]dx′

+ t−1 exp[−

(ξ − ul

2t) ul

]∫ ∞

ξ−ultx′ exp

[−(2t)−1 x′2

2ν− x(ξ − x′ −ult)

]dx′

+∫ ξ/

√t

−∞ηe−η2/4ν dη .

Page 65: Surveys and Tutorials in the Applied Mathematical Sciences

58 3 Fronts in Random Burgers Equations

Finally, let us introduce a new variable η = x′√t and rearrange the order of the terms

to get

∫ ξ√t

−∞ηe−

η24ν dη +

√tule−

ul2ν (ξ− ul

2 t)∫ ∞

ξ−ul t√t

e−η24ν −

x(ξ−√tη−ul t)2ν dη

+ e−ul2ν (ξ− ul

2 t)∫ ∞

ξ−ul t√t

ηe−η24ν −

x(ξ−√tη−ul t)2ν dη

≡ At +Bt +Ct . (3.21)

Likewise, the integral in the denominator equals Btul

+Dt , where Bt is as above and

Dt =√

t∫ ξ/

√t

−∞e−η2/4ν dη . (3.22)

The Hopf–Cole solution formula is put in the form

u =At +Bt +Ct

Bt/ul +Dt, (3.23)

where Bt ,Dt are positive.

3.3 Asymptotic and Probabilistic Preliminaries

Let us first state a Laplace-type asymptotic lemma:

Lemma 3.2. Let ϕλ (u) ∈ C(R1), ϕλ (u) → ϕ(u), uniformly on compact sets of uas λ → ∞, and C1u2 ≤ |ϕλ (u)| ≤ C2u2 for some positive constants Ci, i = 1,2,uniformly in λ →∞. The limiting function ϕ(u) is in C(R1), ϕ(u) < ϕ(u∗), ∀u 6= u∗.Here c0 and C are positive constants. Then for the probability measures µλ withdensities

expλϕλ (u)du∫R1 expλϕλ (u)du

,

we have as λ →+∞,

1. µλd→ δ (u∗), the unit mass at u∗;

2. the expected value Eµλ (u) approaches u∗;3. λ−1 log

∫R1 expλϕλ (u)du→ ϕ(u∗).

The proof is left as an exercise with hints.Let us recall next a probabilistic fact that the stochastic process Wy−y2/2, where

Wy is the standard Wiener process on the line, attains a unique maximum almostsurely [233]. Such a process is also known as Brownian motion with a parabolicshift; see [105] for a detailed study.

Page 66: Surveys and Tutorials in the Applied Mathematical Sciences

3.4 Asymptotic Reductions 59

We shall make use of a consequence of assumption A4. Note that the paths of theprocess ξ (x,ω) are (with probability one) continuous, strictly increasing functionsof x. Therefore, each has a continuous inverse, defining another process x(ξ ,ω).Assumption A4 says that the process ξ (x,ω) satisfies an invariance principle, whichin fact implies that the same is also true about the process x(ξ ). More precisely,

x(tξ )− tξ

µ

µ− 32 σ√

t

|ξ |≤ξ0

d→ (Wξ )|ξ |≤ξ0. (3.24)

The complete proof can be found at [234, Theorem 4.1]. In the sequel, we will usethe following notation for the process x(ξ ) with its mean subtracted:

x(ξ ) = x(ξ )− ξµ

. (3.25)

3.4 Asymptotic Reductions

In the next two propositions we prove that a part of the expression for u goes to zero

will be used in the proof of the theorem, where it will be important that the conver-gence take place uniformly in ν , in the appropriate sense defined below in (3.26).With this in mind, we adopt the following convention about constants: constants in-dependent of ν , but depending on the random parameter ω (i.e., on the realization ofthe random flux), will be denoted by C(ω), or simply by C. Constants independentof both ν and ω will be denoted by c. The actual value of C or c may vary from oneline to another.

Proposition 3.3. We have

limt→∞

supξ

At

Bt/ul +Dt

d= 0.

Moreover, convergence is uniform in ν in the sense that for every ε > 0, as t → ∞,

P

[supν≤ν0

∣∣∣∣At

Bt/ul +Dt

∣∣∣∣ > ε

]→ 0, (3.26)

for any ν0 > 0. Note that convergence in distribution to 0 is equivalent to conver-gence in probability to 0.

Proof. We have for positive ξ ,

|At | ≤ ν∫ 0

−∞η exp−η2/4νdη = c1 ν ,

at large times, thereby asymptotically simplifying the solutions. These propositions

Page 67: Surveys and Tutorials in the Applied Mathematical Sciences

60 3 Fronts in Random Burgers Equations

and

Dt ≥√

t∫ 0

−∞exp−η2/4νdη =

√ν t c2,

so|At |Dt

≤ c√

νt−1/2, (3.27)

with an absolute constant c. For negative ξ , we restrict the integration in the defi-nition of Bt to the interval 0 ≤ η ≤ 1 and note that since x(u)/

√|u| converges in

distribution to a normal random variable, with probability one there exists an (ω-dependent) constant C such that for all u,

x(u)≤C|u|2/3. (3.28)

Therefore the integral ∫ ∞

ξ−ul t√t

e−η24ν −

x(ξ−√tη−ul t)2ν dη (3.29)

can be bounded below by

e−1/4ν∫ 1

0exp

(− x(ξ −√tη−ult)

)dη ≥ e−1/4ν e

−Cν |ξ−√t−ult|2/3

, (3.30)

for all t and ξ . This implies that for almost all ω and t ≥ 1 (uniformly in ξ ≤ 0), wehave

Bt ≥√

tul exp− ul

(ξ − ul

2t)

e−1

4ν e−Cν (ξ−√t−ult)2/3

≥ ule−ul2ν (ξ− ul

2 t)−Cν |ξ−t−ult|2/3

e−1/4ν .

The last expression clearly goes to ∞ uniformly in ξ ≤ 0 as t → +∞. Since |At | isbounded from above by an absolute constant, it follows that

supξ≤0

|At |Bt

→ 0, (3.31)

for almost all ω . Combining (3.27) and (3.31) ends the proof. utLet us now proceed with our next proposition.

Proposition 3.4. We have

limt→∞

supξ

Ct

Bt/ul +Dt= 0;

the convergence is uniform in ν ∈ (0,ν0) in the sense of (3.26).

Proof (Sketch). We begin by observing that

Page 68: Surveys and Tutorials in the Applied Mathematical Sciences

3.4 Asymptotic Reductions 61

Ct

Bt=

1√tul

e−x(ξ−ul t)

2ν∫ ∞

ξ−ul t√t

ηe[x(ξ−ul t)−x(ξ−ul t−√tη)]

2ν − η24ν dη

e−x(ξ−ul t)

2ν∫ ∞

ξ−ul t√t

e[x(ξ−ul t)−x(ξ−ul t−√tη)]

2ν − η24ν dη

. (3.32)

Changing the variable to y = t−16 η , we obtain

Ct

Bt=

1√tul

t1/3 ∫ ∞ξ−ul t

t2/3ye

t1/32ν

[x(ξ−ul t)−x(ξ−ul t−t2/3y)

t1/3 − y22

]

t1/6∫ ∞

ξ−ul t

t2/3e

t1/32ν

[x(ξ−ul t)−x(ξ−ul t−t2/3y)

t1/3 − y22

] . (3.33)

We shall first consider the values of ξ satisfying

ξ − 23

ult ≤ 0. (3.34)

The stationarity of a implies easily that

x(ξ −ult)− x(ξ −ult− t2/3y)t1/3

d=x(t2/3y)

t1/3 , (3.35)

with equality in law of processes in the variable y∈R. As remarked above, assump-tion A4 implies that the processes

x(ξ −ult)− x(ξ −ult− t2/3y)σ ′t1/3 , (3.36)

where σ ′ = σ ′(µ,σ) is a positive constant, converge in law to the Wiener process,which can be strengthened to almost sure uniform convergence on compact intervalsof y by a change of probability space via the Skorohod representation theorem (see[202, Theorem 86.1]). With the Laplace asymptotic lemma (Lemma 3.2), we seethat as long as

ξ −ultt2/3 →−∞′

the ratio of the two integrals of (3.33) converges in distribution to y0, where y0 de-notes the unique value of y where the function σ ′Wy− y2/2 attains its maximum.Existence and uniqueness of such a point implies that for almost all ω , Ct/Bt con-verges to zero uniformly in ξ satisfying (3.34).

To handle the values of ξ for which

ξ − 23

ult > 0, (3.37)

note that ξ − ul2 t →+∞, uniformly in ξ satisfying (3.37). We will now use an argu-

ment similar to the one used in the proof of Proposition 3.3 to show that for almost

Page 69: Surveys and Tutorials in the Applied Mathematical Sciences

62 3 Fronts in Random Burgers Equations

all ω , the quantity Ct/Bt converges to zero uniformly in ξ satisfying (3.37). Withprobability one, there exists an (ω-dependent) constant C such that (3.28) holds.This, together with subadditivity of the function u 7→ |u|2/3, implies that

x(ξ −√tη−ult)≥−C(|ξ |2/3 + t1/3|η |2/3 +u2/3

l t2/3)

.

Therefore the ξ -dependent part of the integrand can be absorbed into the prefactor:

|Ct | ≤ e−ul2ν ( 5

6 ξ− ul2 t)

∫ ∞

ξ−ul t√t

ηe−η24ν + C

2ν (t1/3|η |2/3+u2/3l t2/3) dη . (3.38)

We now divide the integral in the last formula into two parts, corresponding to |η | ≤1 and |η | ≥ 1. The first integral is clearly bounded above by

ce−ul2ν ( 5

6 ξ− ul2 t)ect2/3

.

The last expression goes exponentially fast to zero, uniformly in ξ satisfying (3.37),since for those ξ ,

56

ξ − ul

2t ≥ 1

18ult. (3.39)

When |η | ≥ 1, we have |η |2/3 ≤ |η |, and consequently, the right-hand side of (3.38)is bounded above by

e−ul2ν ( 5

6 ξ− ul2 t)e

C2ν u2/3

l t2/3∫ ∞

ξ−ul t√t

|η |e− η24ν + C

2ν t1/3|η | dη .

The integral in the above formula can be estimated by first absorbing the factor |η |into the exponential factor (by making C bigger) and then using the identity

Re−

η24ν + C

2ν sη dη =√

4πνeCs

with s = t1/3. We obtain in this way

|Ct | ≤ e−ul2ν ( 5

6 ξ− ul2 t)√

4πνeC2ν u2/3

l t2/3ect1/3

,

where the last expression clearly goes to zero uniformly in ξ satisfying (3.37) (see(3.39)). Since for these ξ , Dt can be uniformly bounded from below by c

√t, where

c > 0 is an absolute constant, the proof is finished. ut

Page 70: Surveys and Tutorials in the Applied Mathematical Sciences

3.5 Front Probing and Central Limit Theorem 63

3.5 Front Probing and Central Limit Theorem

Now let us prove the main theorem of this chapter. The strategy is to study thesolution of the regularized equation (3.10) along rays ξ = αt in the Hopf–Colerepresentation, and then take the limit ν → 0 to infer the same behavior for theBurgers solutions U . It follows from the two preceding propositions that we justneed to study the limiting distribution of Bt

Bt/ul+Dt.

Assume α > c, so we are ahead of the actual front speed. In the ξ coordinate,this means that in the representation (3.21) of Bt , the factor e(−ul/2ν)(ξ−ul/2t) goesexponentially fast to zero, uniformly in ν . We now use the bound (3.28) (true withprobability one for some constant C), and proceeding exactly as in the proof ofProposition 3.4, we have, with probability one,

Re−

η24ν −

x(ξ−√tη−ul t)2ν dη ≤CeCt2/3

. (3.40)

This clearly implies that Bt → 0 almost surely as t→∞. On the other hand, Dt →+∞(at the order of

√t), so

Bt

Bt/ul +Dt→ 0 (3.41)

almost surely, and therefore the analogue of part 1 of the theorem is proven for thesolution uν , ν > 0, solving the regularized equation (3.10). Note that all the aboveconvergence statements, including (3.41), hold uniformly in ν . It follows that

supν≤ν0

|uν(ξ (αt), t)| d→ 0. (3.42)

Thanks to assumption A5, classical results [185, Theorems 13 and 14] and [207]yield that for any given t, except for a set of x consisting of countably many discon-tinuities of the first kind (shocks),

limν→0

uν(x, t) = u0(x, t). (3.43)

Moreover, u0(x, t) is the unique physical weak solution of the (inviscid) Burgersequation, the so-called entropy solution [141]. It follows from (3.42) and (3.43) that

u0(ξ (αt), t) d→ 0.

To prove the same convergence for U , note that by (3.11),

U(αt, t) =1√

a(αt)u0(ξ (αt), t). (3.44)

Since the random variables 1/√

a(αt) have finite second moment uniformly in t(by A1 and A3), the product in (3.44) goes to zero in distribution. In fact, for anyε > 0 and K À 1,

Page 71: Surveys and Tutorials in the Applied Mathematical Sciences

64 3 Fronts in Random Burgers Equations

P

(∣∣∣∣∣1√

a(αt)u0(ξ (αt), t)

∣∣∣∣∣ > ε

)≤ P(|u0(ξ (αt), t)|> ε/K)+P

(1√

a(αt)> K

),

where the first term goes to zero at large t, and the second term can be made as smallas possible for large K by Chebyshev’s inequality. This ends the proof of part one ofthe theorem, which says that we observe the value zero if we are ahead of the front.

Similarly, if α < c, so we are behind the front, then the Bt term grows exponen-tially fast with probability one (due to the exponential prefactor in (3.21)), while Dtgrows at most like

√t (if at all). Hence

Bt

Bt/ul +Dt→ ul , (3.45)

and part two is proven for a positive ν . Just as in the proof of part one, it suffices tonote now that the convergence is uniform in ν , and part two of the theorem follows.In contrast to part one, uν does not converge to zero, and therefore we do not obtainconvergence of U(αt, t). In view of (3.11), we see that U(αt, t) fluctuates as t →∞.So we observe a noisy state in U at the back of the front due to the effect of therandom flux.

Now let us probe the front more precisely at

x = ct + z√

t. (3.46)

We want to find the distribution of

Bt

Bt/ul +Dt

in the limit t → ∞. The Dt integral behaves as√

t times a constant of order O(√

ν).Roughly speaking, Bt is either exponentially large or exponentially small. Accord-ingly, the above ratio is close to ul or 0. This will be seen from the calculation below.Let y ∈ (0,ul) (note that 0 < Bt/

(u−1

l Bt +Dt)

< ul). We have

P

[Bt

u−1l Bt +Dt

≤ y

]= P

[Bt

Dt≤ uly

ul − y

]= P

[logBt√

t− logDt√

t≤

log ulyul−y√t

]

= P

logBt√t−ν

logDt√t≤ ν

log ulyul−y√t

]. (3.47)

Now, ν loguly/(ul−y)√t → 0, and since Dt is of order

√νt, we have ν logDt√

td→ 0 as well.

Both convergence statements hold uniformly in ν in the sense explained in (3.26).The limit of the probability in (3.47) is therefore equal to

limt→∞

P[

ν logBt√t

≤ 0].

Page 72: Surveys and Tutorials in the Applied Mathematical Sciences

3.5 Front Probing and Central Limit Theorem 65

Now write Bt in product form:

Bt = p(t)Bt , (3.48)

where p(t) = e−ul2ν (ξ− ul

2 t)− x(ξ−ul t)2ν and

Bt = ul√

t∫ ∞

ξ−ul t√t

e−η24ν + 1

2ν [x(ξ−ult)−x(ξ−ult−√

tη)] dη . (3.49)

Changing the variable of integration, as in the proof of Proposition 3.4, to y =t−1/6η , we obtain

Bt = t1/6∫ ∞

ξ−ul t

t2/3

et1/32ν

[x(ξ−ul t)−x(ξ−ul t−t2/3y)

t1/3 − y22

]

dy. (3.50)

Notice thatx(ξ −ult)− x(ξ −ult− t2/3y)

µ−3/2σt1/3

converges in distribution to the Wiener process in the variable y, on any finite inter-val of y. Just as before, strengthening this to uniform convergence with a change ofprobability space (a la Skorohod) and applying the Laplace lemma, we find that ast → ∞, the distribution of

t−1/3 log Bt

converges to that of a constant times

supy

(y2

2−Wy

).

It follows thatνt−1/2 log Bt

d→ 0

uniformly in ν , and we just need to study the behavior of νt−12 log p(t).

We have

νt−12 log p(t) =−1

2

[ul(ξ − ul

2 t)√t

+x(ξ −ult)√

t

], (3.51)

where ξ = ξ (ct + z√

t). Since c = 1/2µ2 and ul = 1/µ , substituting (3.46), we getfrom the central limit theorem for ξ in assumption A4 that

ul(ξ (ct + z

√t)− ul

2 t)

√t

d→ z+σ

µ2√

2W1,

where W1 is a Gaussian random variable with mean zero and unit variance. Also,using the central limit theorem for x ((3.24) with b = 1/2µ), we obtain

Page 73: Surveys and Tutorials in the Applied Mathematical Sciences

66 3 Fronts in Random Burgers Equations

x(ξ −ult)√t

d→ σµ2√

2W1.

A further study of the joint distribution of the variables

ul(ξ (ct + z√

t)− ul2 t)√

t

andx(ξ −ult)√

t

in the limit t → ∞ shows that the two-dimensional random variables(

ul(ξ (ct + z√

t)− ul2 t)√

t,

x(ξ −ult)√t

)

converge in distribution to a two-dimensional Gaussian with independent coordi-nates of means z and 0 respectively. The sum

ul(ξ (ct + z

√t)− ul

2 t)

√t

+x(ξ −ult)√

t

converges in distribution to a Gaussian random variable with mean z and varianceσ2/µ4. Hence

P[νt−1/2 log p(t)≤ 0

]→ P

[−1

2(z+

σµ2 W1)≤ 0

]= P

[W1 ≥−µ2

σz]

= N

(µ2

σz)

,

where N (s) = 1/√

2π∫ s−∞ e−s′2/2 ds′ is the error function. The above arguments

can be made uniform in ν > 0, implying that the Burgers solution u0 satisfies

limt→∞

P[u0(ct + z

√t, t)≤ y

]= N

(µ2

σz)

and

limt→∞

P[u0(ct + z

√t, t) > y

]= 1−N

(µ2

σz)

for all z, which can be rephrased as part 3 of the theorem on U .

Page 74: Surveys and Tutorials in the Applied Mathematical Sciences

3.6 Exercises 67

3.6 Exercises

1. Prove the first assertion of the Laplace lemma, Lemma 3.2, by taking thelimit as λ → ∞ of the integral

∫R1 ψ(u)dµλ , where ψ(u) ∈C∞(R1), |ψ(u)| ≤

C(1 + u2)m, for some m > 0. Show that the part of integral over a small neigh-borhood of the maximal point u∗ can be made arbitrarily close to ψ(u∗), and theremaining part of the integral is negligible for large λ . Part two of the Laplacelemma follows by setting ψ(u) = u.

2. Prove part three of the Laplace lemma, Lemma 3.2, by passing to the limit

λ−1 log∫

expλ (ϕλ (u)−ϕ(u∗))du→ 0, λ → ∞,

with a similar integral decomposition as in Exercise 1. Note that over a smallneighborhood of the maximal point u∗, ϕλ (u)− ϕ(u∗) is small for λ largeenough.

3. Define the hitting time

T (b) = infx≥ 0 : ξ (x) = b,

where ξ satisfies assumption A4. Express the probability of deviation of T (tb)from tb/µ in terms of ξ to show that a law of large numbers holds for T (tb):

T (tb)t

p→ bµ

, t → ∞,

where convergence is in probability [72].

4. Write T (tb)− tb/µ in terms of ζ (x) = ξ (x)−µx and apply A4 to show that forany fixed b, as t → ∞,

T (tb)− tb/µµ−3/2σ

√t

d→Wb.

Page 75: Surveys and Tutorials in the Applied Mathematical Sciences

Chapter 4Fronts and Stochastic Homogenization ofHamilton–Jacobi Equations

Hamilton–Jacobi (HJ) equations modeling fronts in spatially random media are first-order nonlinear partial differential equations (PDEs) of the form

ut +H(x,ω,∇xu) = 0, x ∈ RN , N ≥ 1, (4.1)

where H, the Hamiltonian, is random in x (position) and nonlinear in ∇xu (momen-tum). In case of one space dimension (N = 1), a scalar conservation law with arandom flux as in Chapter 3 follows by formally taking the x derivative of (4.1):

vt +(H(x,ω,v))x = 0,

where v = ux. The Hamiltonian becomes the nonlinear flux function. Naturally, frontspeeds in scalar laws are related to HJ solutions. However, simple front solutions ofthe form p · x− c(p) t no longer exist in (4.1) due to the complexity of randomdependence.

A front can be analyzed in a hyperbolic scaling limit. In other words, space andtime are scaled the same. If the space scale is large, such as O(1/ε) for ε ¿ 1,then the time scale should also be O(1/ε). On these space–time scales, a front isan (asymptotic) invariant. If the medium is homogeneous and we scale x = x/ε ,t = t/ε , then the exact front solution is

u(x, t) = p · x−H(p)t =1ε(

p · x−H(p)t).

It follows that the front solution is invariant if we scale u = εu, or u(x, t) =p · x−H(p)t. For inhomogeneous media, one considers the initial value problemu(x,0) = p · x of (4.1) with planar initial data and then recovers the front speedby − limt→∞ u(0, t)/t. Performing the same scaling change of variables for the HJequation (4.1), we have

uεt +H

(x/ε,ω,∇xuε) = 0, (4.2)

© Springer Science + Business Media, LLC 2009

J. Xin, An Introduction to Fronts in Random Media, Surveys and Tutorials in the Applied 69Mathematical Sciences 5, DOI: 10.1007/978-0-387-87683-2_4,

Page 76: Surveys and Tutorials in the Applied Mathematical Sciences

70 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

with initial data uε(x,0) = p · x. The large-t limit problem on u is the same as thesmall-ε limit problem (4.2) of uε , which is known as homogenization. Homogeniza-tion of HJ in the periodic setting was first studied in the 1980s [148]. More recentwork has concentrated on the random setting, which we shall discuss in more detail.

The upshot is that under certain conditions on the Hamiltonian and randomness,solutions of (4.2) converge as ε ↓ 0 to a limiting function u satisfying an averaged(homogenized) HJ equation

ut + H(∇xu) = 0, (4.3)

with the same initial data. In the limiting process, the fast oscillations have beenremoved, and H is a deterministic function even if H is random to begin with. Ho-mogenization is the analogue of the law of large numbers for stochastic PDEs. Thehomogenized equation (4.3) has the exact front solution

u(x, t) = p · x− H(p)t. (4.4)

We observe at (x, t) = (0,1/ε) or (x, t) = (0,1) to obtain

εu(0,1/ε) = uε(0,1)→ u(0,1) =−H(p). (4.5)

Letting T = 1/ε À 1, (4.5) implies

u(0,T )∼−H(p)T, (4.6)

or the front speed exists asymptotically as H(p) in the direction p. Fronts in randommedia are therefore closely connected to the stochastic homogenization of HJ.

In this chapter, we shall discuss sufficient conditions for stochastic homogeniza-tion and related central limit theorems, then show examples in which homogeniza-tion breaks down. The latter (anomalous) regime corresponds to the study of ex-trema of stochastic sequences and processes [143] in probability theory. We shallgive necessary and sufficient conditions for homogenization of the random Hamil-tonians H(x,ω, p) = |p|2/2+V (x,ω) in classical mechanics, and point out the con-sequences of extreme behavior of random media on the anomalous front dynamics.

4.1 Convex Hamilton–Jacobi and Variational Formulas

Let us consider the homogenization problem (4.2) with the tildes removed:

uεt +H (x/ε,ω,∇xuε) = 0, (4.7)

where the superscript ε on u denotes its ε dependence. The major assumptions onthe random Hamiltonian are these:

(A1) Convexity: H = H(x,ω, p) is convex in p for any (x,ω) ∈RN ×Ω . In otherwords, for any p1, p2 ∈ RN and any θ ∈ [0,1], we have for any (x,ω),

Page 77: Surveys and Tutorials in the Applied Mathematical Sciences

4.1 Convex Hamilton–Jacobi and Variational Formulas 71

H(x,ω,θ p1 +(1−θ)p2)≤ θH(x,ω, p1)+(1−θ)H(x,ω, p2).

(A2) Stationarity and ergodicity: H is a stationary and ergodic random field in(x,ω) for any p with respect to the shift (translation) in x.

(A3) Coercivity: There exist positive deterministic constants C1 and C2 such thatfor some exponent α > 1,

C2 (|p|α −1)≤ H(x,ω, p)≤C1(|p|α +1). (4.8)

(A4) Continuity in x: There exists a continuous function m : [0,∞)→ [0,∞) withm(0) = 0 such that

|H(x,ω, p)−H(y,ω, p)| ≤ m(|x− y|)(1+ |p|). (4.9)

When m is a linear function, Lipschitz continuity holds and (A4) will be called(A4)′ below.

A few words are in order for the assumptions. Convexity and coercivity help todefine the Lagrangian function L via the Legendre transform

L(x,ω, p) = supq∈RN

(p ·q−H(x,ω,q)) = maxq∈RN

(p ·q−H(x,ω,q)), (4.10)

which is also convex and coercive. With the Lagrangian L and the Lipschitz regular-ity (A4)′, HJ solutions are represented in terms of a variational path integral formula(the Lax formula [141, 147]):

uε(x, t,ω) = infy∈RN

(g(y)+Sε(x,y,y,ω))

≡ infy∈RN

(g(y)+ inf

ξ∈A

∫ t

0L(ξ (s)/ε, ξ (s),ω)ds

), (4.11)

where the set A consists of all Lipschitz continuous paths ξ (s) joining y = ξ (0)to x = ξ (t), and the function Sε is called the action. The function g is the initialdata of uε and is assumed to be Lipschitz continuous on RN . If H is x-independent(homogeneous media), then the Lax formula reduces to the Hopf formula [80]; see(4.13).

The ergodicity assumption in (A2) provides the averaging mechanism for ho-mogenization. A stronger and more quantitative assumption than ergodicity is “mix-ing” [36, 72], which essentially means fast decay of correlations or independenceof events that occur in vastly separate spatial regions. In case N = 1, x may alsobe thought of as time; then mixing says that the ancient past is nearly independentof the future. The mixing assumption is indirectly made in (A4) for the analysis ofBurgers fronts in Chapter 2. Stationarity is also assumed in Chapter 2, and it makesthe ensemble-averaged statistical quantities equal to constants instead of being x-dependent.

Page 78: Surveys and Tutorials in the Applied Mathematical Sciences

72 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

With stationarity and ergodicity assumptions, H can be generated by the opera-tor τx (shift by x), with H(x,ω, p) = H(τxω, p) for some function H(·, p) convexin p. Moreover, any shift-invariant set of events occurs with probability zero orone. Due to (A2), the averaged Hamiltonian is expected to be deterministic and x-independent. Assumption (A2) has appeared already in stochastic homogenizationof linear elliptic PDEs [190].

The proof of homogenization will first utilize the Lax formula (4.11) to averageout randomness and obtain the homogenized Lagrangian L and then recover H fromthe Legendre transform. The convergence in the mean is given by the followingtheorem [200]:

Theorem 4.1. Assume (A1)–(A4) and that the initial datum g is Lipschitz, and con-sider (x, t) on a compact set D of RN × [δ ,∞). Then for each δ > 0, the quantity uε

as given by the Lax formula (4.11) satisfies

limε↓0

E[

supD|uε(x, t,ω)− u(x, t)|

]= 0, (4.12)

where u is given by the Hopf formula of the homogenized equation (4.3):

u(x, t) = infy∈RN

[g(y)+ tL((x− y)/t)], (4.13)

where L, the homogenized Lagrangian, is both convex and coercive. If Lipschitzcontinuity (A4)′ holds, then uε by the Lax formula satisfies the HJ equation (4.7).

Analogous assumptions in [200] are weaker than what we have stated here in thatthe constants C1–C3 are allowed to be random constants satisfying certain momentconditions, and that coercivity is more general. By the Hopf formula, u satisfies thehomogenized HJ equation (4.3) almost everywhere [80].

Under suitable initial conditions, uε from the Lax formula is the unique vis-cosity solution of the initial value problem of HJ [147], a class of weak solutionsfrom taking the zero viscosity limit as was done for Burgers’ equation in Chapter 2.Compactness and semigroup property of viscosity solutions yields another mode ofconvergence up to the initial time [227]:

Theorem 4.2. Assume (A1)–(A4). Then the unique viscosity solutions of (4.2) frombounded uniformly continuous initial data converge almost surely to those of (4.3)over compact sets of RN × [0,∞).

For a systematic exposition of viscosity solutions, see [147, 80].In the sequel, we shall view Lax and Hopf formulas as generalized HJ solution

formulas, and study the passage to the limit.

Page 79: Surveys and Tutorials in the Applied Mathematical Sciences

4.2 Subadditive Ergodic Theorem and Homogenization 73

4.2 Subadditive Ergodic Theorem and Homogenization

The main part of the homogenization proof via the Lax formula is the application ofthe subadditive ergodic theorem [72] to the convergence of the action functional Sε .Let us prove a theorem:

Theorem 4.3. For fixed x, y ∈ RN, and t > 0, we have

limε↓0

E[|Sε(x,y, t)− tL((x− y)/t)|] = 0, (4.14)

where L is a convex and coercive function on RN.

Proof. In preparation, we first observe a scaling property of Sε :

Sε(x,y, t,ω) = εS1(x/ε,yε, tε,ω). (4.15)

This follows by a change of variables in (4.11), and is left as an exercise.The next observation is that

Sε(x,y, t,ω) d= Sε(x− y,0, t,ω),

by stationarity of the random media. So it suffices to prove a special case of Theorem4.3 that almost surely and in L1 (in the sense of mean (4.14)), the scaled actionfunctional Sε(x,0, t,ω) converges to tL(x/t). To this end, let us consider a doublyindexed random sequence of numbers:

Sm,n = S1(nx,mx,(n−m)t,ω). (4.16)

By the scaling property (4.15) and taking ε = 1/n, we see that Sε(x,0, t,ω) = S0,n/n.The doubly indexed sequence essentially contains information on the “costs” of atrip from mx to nx in time (n−m)t. For the limit S0,n/n to be well behaved, thesequence Sm,n needs to satisfy a subadditive property and finite moment conditions,which are summarized in Kingman’s subadditive ergodic theorem [72, 200]:

Theorem 4.4. Suppose that Sm,n are random variables satisfying the following con-ditions:

1. S0,0 = 0, Sm,n ≤ Sm,k +Sk,n, for m≤ k ≤ n;

2. Sm,m+k,m≥ 0,k ≥ 0 equals Sm+1,m+k+1,m≥ 0,k ≥ 0 in distribution;

3. E[S+0,1] < +∞; αn ≡ E[S0,n] < ∞, ES0,n ≥ γ0n, for a finite constant γ0.

Then the following hold:

1.α = lim

n→∞

αn

n= inf

n≥1

αn

n∈ (−∞,∞);

2. S∞ = limn→∞ S0,n/n exists with probability one;

Page 80: Surveys and Tutorials in the Applied Mathematical Sciences

74 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

3. limn→∞ E[|S0,n/n−S∞|] = 0.

Let us now verify that the sequence Sm,n satisfies the three conditions in King-man’s theorem: subadditivity, stationarity, and finite moments. By definition of Sε ,the action Sε(x,y, t,ω) is the infimum of a collection of Lipschitz paths connectingy to x in time t; hence making a stop in the middle only may increase the cost:

Sε(x,y, t + s,ω) = infz

[Sε(x,z, t,ω)+Sε(z,y, t,ω)]≤ Sε(x,z, t,ω)+Sε(z,y, t,ω),

for any z, implying subadditivity:

Sm,n(ω) = S1(nx,mx,(n−m)t,ω)

≤ S1(kx,mx,(k−m)t,ω)+S1(nx,kx,(n− k)t,ω)= Sm,k(ω)+Sk,n(ω), (4.17)

for any m ≤ k ≤ n. In simple terms, subadditivity says that the direct flight from mto n is cheaper than a one-stop flight, a consequence of the Lax formula.

For stationarity, we again infer from the definition of action (ξ being Lipschitzpath) that

Sm,n(ω) = inf∫ (n−m)t

0L(ξ (s), ξ (s),ω)ds|ξ (0) = mx,ξ ((n−m)t) = nx

= inf∫ (n−m)t

0L(ξ (s)+mx, ξ (s),ω)ds|ξ (0) = 0,ξ ((n−m)t) = (n−m)x.

= inf∫ (n−m)t

0L(ξ (s), ξ (s),τmxω)ds|ξ (0) = 0,ξ ((n−m)t) = (n−m)x

= S0,n−m(τmxω).

Then it follows from the invariance of probability under translation τ that

Sm,m+k(ω) : m,k ≥ 0= S0,k(τmxω) : m,k ≥ 0d= S0,k(τ(m+l)xω : m,k ≥ 0= Sm+l,m+k+l(ω) : m,k ≥ 0. (4.18)

Coercivity and convexity of the Hamiltonian implies that of the Lagrangian [80],so there exists a convex and superlinearly increasing function ψ(p) such that

C2(ψ(p)−1)≤ L(x, p,ω)≤ C1(ψ(p)+1),

for positive constants Ci, i = 1,2. It follows that for any m≤ n,

C2(tψ((x− y)/t)− t)≤ Sε(x,y, t,ω)≤ C1(tψ((x− y)/t)− t),

or

Page 81: Surveys and Tutorials in the Applied Mathematical Sciences

4.2 Subadditive Ergodic Theorem and Homogenization 75

C2((n−m)tψ(x/t)− (n−m)t)≤ S1(nx,mx,(n−m)t,ω)

≤ C1((n−m)tψ(x/t)+(n−m)t).

Hence |S0,n|/n ≤ C3 for a positive constant C3, or S0,n/n is bounded away frominfinity uniformly in n and ω for fixed (x, t), t > 0.

The subadditive ergodic theorem, Theorem 4.4, applies to give

limn→∞

E[S0,n/n] = limn→∞

S0,n/n = limn→∞

S1/n(x,0, t,ω)≡ S(x, t), (4.19)

almost surely and in the mean. The limiting value is τ-invariant, and so is determin-istic by ergodicity of the random media.

Though the above limit (4.19) is established for ε = 1/n, it can be extended toany ε thanks to the Lipschitz continuity of the action Sε uniformly in (ε,ω) overcompact sets of (x, t) (t ≥ δ > 0).

The Lipschitz continuity of the action Sε follows from (A1) and (A3); see [200,Lemma 3.1] for a proof. Note that assumption (A3) implies [200, inequality (3.17)],which is sufficient.

Now write ε−1 = n+ r, r ∈ (0,1). Then

εS1( x

ε,0,

tε,ω

)= εS1(nx+ rx,0,nt + rt,ω) = εnS1/n

(x+

rxn

,0, t +rtn

,ω)

,

and by Lipschitz continuity,

∣∣∣S1/n(

x+rxn

,0, t +rtn

,ω)−S1/n(x,0, t,ω)

∣∣∣≤C4|x|+ t

n,

implying that

limε↓0

εS1( x

ε,0,

tε,ω

)= S(x, t),

almost surely and in the mean.The scaling property of the action Sε leads to the self-similar form of the limiting

function S:

limε→0

Sε(x,0, t,ω) = limε→0

εS1( x

ε,0,

tε,ω

)

= limε→0

(tε)S1( x

tε,0,

ttδ

,ω)

= t limε→0

Sε(x

t,0,1,ω

)

= tS(x

t,1

)≡ tL

(xt

), (4.20)

where L is the homogenized Lagrangian function.For general y 6= 0, we have

Sε(x,y, t,ω = Sε(x− y,0, t,τy/ε)ω) d= Sε(x− y,0, t,ω),

Page 82: Surveys and Tutorials in the Applied Mathematical Sciences

76 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

by shift invariance of probability, and so

limε↓0

E[|Sε(x,y, t,ω)− t L((x− y)/t)|] = 0.

Let us show that L is both convex and coercive. By subadditivity, we have for anya,b ∈ RN , λ ∈ [0,1], that

Sε(λa+(1−λ )b,0,1,ω)≤ Sε(λa,0,λ ,ω)+Sε(λa+(1−λ )b,λa,1−λ ,ω).

Then letting ε ↓ 0 gives, in view of (4.20),

(1−λ )L(b) = limε↓0

Sε(λa+(1−λ )b,λa,1−λ ,ω)

and the inequality

L(λa+(1−λ )b)≤ λ L(a)+(1−λ )L(b).

The convexity of L follows.By coercivity of the Hamiltonian (A3) and the Legendre transform, there exist

positive constants c1 and c0 depending only on α and the constants C1, C2 in (A3)such that

L(

ξε

, ξ ,ω)≥ c1ψ(ξ )− c0t, (4.21)

where ψ(p) = |p|β , β = αα−1 the conjugate exponent to α , and β > 1. It follows

that

Sε(x,y, t,ω) = infξ∈A

∫ t

0L

(ξε

, ξ ,ω)

ds≥ c1tψ((x− y)/t)− c0t. (4.22)

Now letting t = 1, y = 0, and sending ε ↓ 0, we arrive at the inequality

L(x)≥ c1ψ(x)− c0.

A similar upper bound L(x) ≤ c2ψ(x)+ c3 holds for positive constants c2 and c3,and so L satisfies the coercivity condition (A3).

This ends the proof of Theorem 4.3. utThe convergence of uε in (4.11) requires boundedness of minimizers of g(y)+

Sε(x,y, t,ω) in addition to the convergence of Sε in Theorem 4.3. Let us prove alemma:

Lemma 4.5. Fix T > 0. There exists a number R = R(T ) such that for ∀t ∈ (0,T ],and ∀ε > 0,

infy∈RN

[g(y)+Sε(x,y, t,ω)] = inf|y−x|≤R(T )

[g(y)+Sε(x,y, t,ω)] (4.23)

and

Page 83: Surveys and Tutorials in the Applied Mathematical Sciences

4.2 Subadditive Ergodic Theorem and Homogenization 77

infy∈RN

[g(y)+ tL((x− y)/t)] = inf|y−x|≤R(T )

[g(y)+ tL((x− y)/t)]. (4.24)

By coercivity of L, for any set B⊂ RN , we have

infy∈B

[g(y)+ tc1ψ((x− y)/t)]− c0t ≤ infy∈B

[g(y)+Sε(x,y, t,ω)]. (4.25)

It follows by taking y = x that

infy∈RN

[g(y)+Sε(x,y, t,ω)]≤ g(x)+Sε(x,x, t,ω)≤ g(x)+C4t (4.26)

for a positive constant C4. By the superlinear growth of ψ and at most linear growthof g, there exists R such that if |x− y| ≥ RT , then

g(x)+C4t ≤ g(y)+ tc1ψ((x− y)/t)− c0t. (4.27)

Define D = y : |x− y| ≤ RT. Then (4.25) implies

g(x)+C4t ≤ infy∈Dc

[g(y)+ tc1ψ((x− y)/t)]− c0t. (4.28)

Letting B = Dc in (4.25), then combining (4.25), (4.26), and (4.28), we obtain, byapplying the coercive bound on L,

infy∈RN

[g(y)+Sε(x,y, t,ω)]≤ infy∈Dc

[g(y)+Sε(x,y, t,ω)],

which proves (4.23). Inequality (4.24) follows similarly from the coercivity boundsof L. Finally, Theorem 4.1 follows from Theorem 4.3, Lemma 4.5, and Lipschitzcontinuity of Sε uniformly in ε .

Example 4.6. Consider the random Hamiltonian H(x,ω, p) = |p|2n +V (x,ω), p ∈RN , where n is a positive integer, V is a stationary and ergodic random field; |V | ≤Cfor all (x,ω) for some finite deterministic constant C, and V has continuous samplepaths. All assumptions (A1)− (A4) are satisfied.

Example 4.7. The Hamiltonian corresponding to the random flux function of theBurgers equation in Chapter 3 is H(x,ω, p) = a(x,ω)p2/2, where x and p∈R1. Forcoercivity (A3) to hold, a(x,ω) must be bounded uniformly away from zero andinfinity, or there are two positive deterministic constants C1 and C2 such that C1 ≤a(x,ω) ≤ C2. This is more restrictive than the positivity and moment assumptionsin Chapter 3, where a is allowed to be a positive unbounded process such as anexponential of a stationary ergodic Gaussian process (log-normal process).

Page 84: Surveys and Tutorials in the Applied Mathematical Sciences

78 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

4.3 Unbounded Hamiltonians: Breakdown of Homogenization

We see from Example 4.7 above that coercivity assumption (A3) is not optimal andmay omit unbounded processes often used in statistical modeling such as Gaussianand log-normal processes. These processes are widely used because they are intrin-sically Gaussian and have fewer parameters to estimate from data than non-Gaussianprocesses. In this section, we shall analyze examples in which coercivity is not sat-isfied and the Hamiltonian functions are allowed to be unbounded (not uniformlybounded in x).

This turns out to be a rich territory where much more work is needed. Let usillustrate this point for specific examples. For simplicity, we shall consider only theexistence of front speed or the limit

H(p)≡− limε↓0

uε(0,1), (4.29)

where uε is given by the Lax formula for front (affine) data p · x.The first example is the Hamiltonian in classical mechanics:

H(x,ω, p) = |p|2/2+V (x,ω), (4.30)

where V is a stationary and ergodic random field with continuous sample paths andintegrability

E[|V (x)|] < ∞. (4.31)

A precise result is given in the following theorem [74]:

Theorem 4.8. For the random Hamiltonian (4.30), the one-sided almost sure upperbound V (x) ≤ V0 for a deterministic constant V0 is necessary and sufficient for theexistence of the HJ front speed (homogenized Hamiltonian). The limit (4.29) existswith probability one.

In other words, the lower bound of the potential V is not necessary.Let us verify the conditions of the subadditive ergodic theorem, Theorem 4.4.

The Hamiltonian H(x, p) = p2/2+V (x), a sum of kinetic and potential energy, de-scribes a classical particle moving in the field of the potential V . The correspondingLagrangian is L(x,q) = |q|2/2−V (x). Fix x ∈RN and t > 0. The action integral fora particle moving from the origin to the point nx in time nt along the path s 7→ ξ (s),0≤ s≤ nt, equals

Φn =∫ nt

0L

(ξ (s)ξ (s)

)ds.

We study the minimum Sn(t,x) of Φn over all paths ξ such that ξ (0) = 0 and ξ (nt) =nx. Consider Sm,n(t,x), where m and n are nonnegative integers such that m < n,defined as

Sm,n(t,x) = minξ (mt)=mxξ (nt)=nx

∫ nt

mtL

(ξ (t), ξ (t)

)dt, (4.32)

Page 85: Surveys and Tutorials in the Applied Mathematical Sciences

4.3 Unbounded Hamiltonians: Breakdown of Homogenization 79

with the minimum taken over all paths ξ connecting mx to nx in the time (n−m)t.For l < m < n, we have clearly from (4.32) that

S0,m(t,x)+Sm,n(t,x)≥ S0,n(t,x).

Next, for a linear path ξ (s) = sx/t, we obtain

S0,n(t,x)≤Φn =n|x|2t

2−

∫ nt

0V

( sxt

)ds,

which, by the integrability (4.31) of V (x), implies that

E[|S0,n(t,x)|] < +∞.

On the other hand, since the kinetic part of the action integral is positive, the upperbound on V implies that

E[S0,n(t,x)]≥−ntV0.

The family Sm,n thus satisfies assumptions (1)–(3) of the subadditive ergodic theo-rem, Theorem 4.4, implying the existence of the finite limit of S0,n/n. This limitis invariant under translations of the realization of V by vectors proportional tox; hence its value is constant with probability one. Applying the scaling propertyof the action functional as before, one sees that the finite limit of S0,n/n equalsS(x, t) = tL((x− y)/t), where L is convex and is coercive from below:

L(p)≥ |p|22−V0.

So L(p) grows superlinearly at large p. Moreover, the limit of the action holds forSε for any sequence of ε ↓ 0 due to the Lipschitz continuity of Sε , which followsfrom [200, Lemma 3.1]. Interestingly, the conditions of that lemma continue to holdwhen V is bounded from above.

To show that the asymptotic front speed exists, we modify Lemma 4.5 so that theboundedness of minimizers of g(y)+Sε(0,y,1,ω) holds almost surely. Because theright-hand side of inequality (4.26) is now Sε(0,0,ω) =−V (0,ω), the constant R isnow a random constant to ensure that minimization occurs in |y| ≤ R(ω). Then wepass Sε(0,y,1,ω) to its almost sure limit L(−y) and obtain the HJ asymptotic frontspeed

H(p) = infy∈RN

[p · y+ L(−y)]. (4.33)

Now let us consider again a particle moving in a potential V that is a realizationof a stationary random field, but this time the support of the distribution of V (x) isunbounded from above. We also assume that the covariance of V decays sufficientlyfast. We will show that unboundedness of V leads to a different behavior of thesystem, where the homogenization limit (4.29) fails because of the divergence ofthe action functional Sε as ε → 0.

Page 86: Surveys and Tutorials in the Applied Mathematical Sciences

80 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

It is instructive to look at the one-dimensional (N = 1) case, in which elementarycalculations can be done. For simplicity, we take V to be the Ornstein–Uhlenbeckprocess [72], i.e., the Gaussian process with mean zero and the covariance function

E[V (x)V (y)] =12

exp(−|x− y|).

Let us fix x, t > 0. For any n, the trajectory of the particle, starting from the point0 at time t = 0 and arriving at the point nx at time nt, has classical action equal to

Sn = min∫ nt

0

[12

(dudτ

)2

−V (u(τ))

]dτ,

where the minimum is taken over all C2 functions u(τ) such that u(0) = 0 andu(nt) = nx. Our goal is to study the asymptotic behavior of Sn/n as n→ ∞. Clearly,the function u that minimizes the action integral is monotone increasing and satisfiesdudτ 6= 0 for all τ .

We can thus rewrite the action in terms of the minimum over the inverse functionsτ(u):

Sn = min∫ nx

0

[12

(dτdu

)−1

−V (u)dτdu

]du,

where the minimum is over all C2 functions τ(u) such that τ(0) = 0 and τ(nx) = nt.It follows from the principle of energy conservation that the solution of the abovevariational problem satisfies

dτdu

=1√

2[En−V (u)],

where En is the total energy of the particle and therefore

Sn =∫ nx

0

[√En−V (u)

2− V (u)√

2(En−V (u))

]du

and ∫ nx

0

1√2(En−V (u))

du = nt.

In particular, En >V (u) for all u∈ [0,nx] for the particle to be able to reach the pointnx. We rewrite the formula for Sn as follows:

Sn =∫ nx

0

[√En−V (u)

2+

En−V (u)√2(En−V (u))

− En√2(En−V (u))

]du.

The first two terms in the integrand are identical, and their integral can be estimatedby Jensen’s inequality (using concavity of the root function):

Page 87: Surveys and Tutorials in the Applied Mathematical Sciences

4.3 Unbounded Hamiltonians: Breakdown of Homogenization 81

1nx

∫ nx

0

√2(En−V (u))du≤

√1nx

∫ nx

02(En−V (u))du

=

√2En−2

1nx

∫ nx

0V (u)du.

By the ergodicity of the Ornstein–Uhlenbeck process, we have

1nx

∫ nx

0V (u)du→ 0,

with probability 1 as n → ∞, and therefore the integral of the first two terms in theexpression for Sn is bounded by 3

√Ennx for large n. The integral of the third term

is simply −Ennt, so for large n we obtain

−tEn ≤ 1n

Sn ≤ 3x√

En− tEn.

As mentioned above, En≥ supu∈[0,nx]V (u). Since the realizations of the Ornstein–Uhlenbeck process V (u), u≥ 0, are unbounded from above with probability one, itfollows that Sn/−ntEn → 1 with probability 1. In particular, Sn/n is unbounded,which implies that the Hamilton–Jacobi equation with potential V does not homog-enize in the usual sense [227, 200].

Moreover, since for En > supu∈[0,nx]V (u)+ x2/2t2 we have

∫ nx

0

1√2(En−V (u))

du < nt,

it follows from the condition satisfied by En that En ≤ supu∈[0,nx]V (u) + x2/2t2.Consequently, Sn/n is asymptotically equivalent to −tV ∗(nx), where

V ∗(y) = supu∈[0,y]

V (u)

is the running maximum of the Ornstein–Uhlenbeck process. The asymptotic equiv-alence of the two sequences means that their ratio converges to a positive constant.

We will invoke the following classical result about the running maximum of theOrnstein–Uhlenbeck process. It follows from the one-dimensional version of [4,Theorem 6.9.5].

Theorem 4.9. The running maximum of V ∗(y) of the Ornstein–Uhlenbeck processsatisfies the limit theorem

Prob[

V ∗(y)−by

ay≤ x

]→ exp(−e−x)

as y→ ∞, wherea−1

y ∼ by ∼ (2logy)1/2

Page 88: Surveys and Tutorials in the Applied Mathematical Sciences

82 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

with ∼ denoting asymptotic equivalence.

The theorem says that the renormalized random variables V ∗(y) converge in dis-tribution to the double exponential distribution. It follows that

V ∗(y)(2logy)1/2 → 1 as y→ ∞

in probability, implying thatSn

n→−∞

in probability.The standard homogenization limit fails as claimed. Instead, the modified limit

holds in probability:Sn

n(2logn)1/2 →−t.

The divergence of homogenization means in particular that for affine data, thegrowth rate of HJ solutions in time is faster than linear. We shall come back tothis point for front solutions in the next section.

Generalization to other random potentials is straightforward. In the above anal-ysis, we have used only the fact that V is a mean-zero Gaussian process to which[4, Theorem 6.9.5 ] applies. For example, the divergence results hold for stationaryGaussian processes with the covariance function r(τ) satisfying

r(τ) = 1−C|τ|α +o(|τ|α)

as τ → 0 for some α ∈ (0,2], and r(τ) is square-integrable in τ ∈R1. The Ornstein–Uhlenbeck process satisfies this condition with α = 1. Processes that satisfy it withα = 2 have differentiable realizations.

The behavior of a particle in a multidimensional random potential unboundedfrom above is similar to that in one dimension. As before, we study the action Snof the particle that goes from the origin to the point nx in time nt, where x is afixed vector in Rd . Let x∗n denote the point in the ball B(0,n|x|) where V reaches itsmaximum, equal to V ∗

n . We also denote the minimum of V in the same ball by V∗n.To obtain an upper bound on Sn, consider a path that first goes from the origin tox∗n in time δnt, moving with a constant velocity (equal to x∗n/δnt), next spends time(1− 2δ )nt at x∗n, and finally goes from x∗n to nx in the remaining time δnt, movingwith the constant velocity nx− x∗n/δnt. For such paths, we get

Sn ≤ 12

δnt( |x∗n|2

δnt

)2

−δntV∗n− (1−2δ )ntV ∗n +

12

δnt( |nx− x∗n|

δnt

)2

−δntV∗n.

It follows thatSn

n≤ 5

2|x|2δ t

−2δ tV∗n− (1−2δ )tV ∗n .

Page 89: Surveys and Tutorials in the Applied Mathematical Sciences

4.4 Normal and Accelerated Fronts in Random Flows 83

Because the process has mean zero, we have V∗n =−V ∗n in law. Applying the asymp-

totic law of V ∗n [4, Theorem 6.9.5], we have the upper bound

Sn

n≤ 5|x|2

δ t− 1

4tV ∗

n|x|,

for δ small enough. By almost sure logarithmic divergence of V ∗n|x| as n → ∞ (true

for a Gaussian process), we conclude that Sn/n →−∞ almost surely, and the stan-dard homogenization limit again fails. It will be interesting to establish a precisedivergence result as in the case of one spatial dimension. However, the source ofdivergence is the same: dominance of the running maxima of an unbounded randomprocess.

We see through the above case studies that the large space–time effects of ran-domness can take two very different forms:

1. Homogenization: In this case, the behavior of the system on large scales is de-scribed by an effective Lagrangian (or Hamiltonian) that is nonrandom. Thedisorder gets averaged, and the extreme nature of the random medium is tamed.An elementary (and linear) analogue of this phenomenon in classical probabil-ity theory is the strong law of large numbers [72].

2. Domination by finite-volume maxima of the random potential: In this case, ho-mogenization breaks down, and on an arbitrarily large scale, the behavior ofthe system is dictated by the maximum value of the disorder on that scale. Theextreme nature of the random medium prevails. A relevant problem of classi-cal probability theory is the study of extrema of stochastic sequences and pro-cesses [143]. For example, the maximum Mn of n independent unit normal ran-dom variables behaves asymptotically as

√2logn and in particular, diverges as

n → ∞. Divergence of maxima of random fields underlies the breakdown phe-nomenon of stochastic homogenization.

The two types of behavior of random systems are incompatible with each other.For the class of Hamiltonians studied here, the behavior depends on a simple math-ematical criterion: boundedness of the potential from above. The one-sided bound-edness of the spatial potential is the sharp version of the boundedness assumptionin the coercivity condition (A3).

4.4 Normal and Accelerated Fronts in Random Flows

Let us examine more examples of unbounded random Hamiltonians and interpretthe results from the perspective of front speeds in random media.

Example 4.10 (Normal Fronts in Gradient Flows). Consider Hamiltonians of theform H(x,ω, p) = p2/2 + p · b(x,ω), p,x ∈ Rd . Here b(x,ω) = ∇U(x,ω), whereU(x,ω) is a scalar random vector field whose realizations are of class C2, and let usassume that the scalar random field

Page 90: Surveys and Tutorials in the Applied Mathematical Sciences

84 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

V1(x,ω) =−12|b|2(x,ω)

satisfies all the conditions on classical potentials in the last section (clearly, V1 isbounded above). Examples of such fields b include Gaussian random fields withappropriate covariance [4]. The corresponding HJ equation models a front movingin a random gradient (compressible) flow field.

The Lagrangian function of the above H is

L(x,ω,q) =12|q−b(x,ω)|2 =

|q|22−q ·b(x,ω)+

|b(x,ω)|22

. (4.34)

As in the previous section, consider a path ξ (s),0 ≤ s ≤ nt such that ξ (0) = 0 andξ (nt) = nx. For such a path, the contribution from the second term to the Lagrangianintegral is ∫ nt

0b(ξ (s)) · ξ (s)ds = U(nx).

This shows that this term is a null Lagrangian, i.e., that the Lagrangian (4.34)leads to the same Euler–Lagrange equations of motion as the Lagrangian of thepotential system

L1(x,q) =|q|22−V1(x,ω),

where V1(x) = − 12 |b|2(x,ω). Thus Theorem 4.8 applies, and implies that homoge-

nization holds for such Hamiltonians of advection type, even though the flow field isunbounded.

Example 4.11 (Front Divergence in Shear Flows). Let b(x,ω) = (V2(x′,ω),0), x′ =(x2, . . . ,xd), 0 ∈RN−1, N ≥ 2, namely a shear flow in the direction x1. The HJ equa-tion associated with the Hamiltonian H(x,ω, p) = |p|2/2+b(x,ω) · p is

ut +b(x,ω) ·∇xu+ |∇xu|2/2 = 0. (4.35)

Consider a solution that represents a front moving in the x1 direction in the formu(x, t) = x1− 1

2 t +w(x′, t). Then w satisfies the HJ equation

wt + |∇x′w|2/2+V2(x′,ω) = 0, (4.36)

which is in the classical potential form. If V2 obeys the assumptions in the previoussection and is unbounded from above, then −w(x′, t)/t diverges for large time t,and the front speed is not asymptotically constant. Instead, there exists front speedacceleration due to the dominance of running maxima of the process V2. Shear flowis a special incompressible (divergence-free) flow. It is interesting to study othertypes of random incompressible flows that may lead to divergence or anomalousbehavior of front speeds.

The equation of motion for the Hamiltonian H1 = p2/2+b(x)p, with x, p ∈ R1,is

Page 91: Surveys and Tutorials in the Applied Mathematical Sciences

4.4 Normal and Accelerated Fronts in Random Flows 85

ξ ′′(s) = b(ξ (s))b′(ξ (s)).

As discussed earlier, the same equation also arises from another Hamiltonian:H2 = p2/2 +V1(x), where V1(x) = −b2(x)/2. In this sense, the advection modelis equivalent to a potential model, and homogenization of the two models is closelyrelated.

The Lagrangians corresponding to H1 and H2 (their Legendre transforms) arerespectively

L1(x,q) =12(q−b(x))2

andL2(x,q) =

12(q+b2(x)).

That the Hamiltonians H1 and H2 lead to the same equations of motion impliesthat the difference of the corresponding action functionals,

∫ nt

0

[L2(ξ (s), ξ (s))−L1

(ξ (s), ξ (s)

)]ds,

does not depend on the path ξ (s). Taking the linear function ξ (s) = xt s, we obtain

∫ nt

0

[L2

(ξ (s), ξ (s)

)−L1

(ξ (s), ξ (s)

)]ds =

∫ nx

0b(u)du.

Divided by n, this expression converges with probability one to the expected valueof b(0) times x. Consequently, homogenization or breakdown of homogenizationholds for H1 and H2 simultaneously. In a general classical-mechanics Hamiltonian(as well as in a shear flow Hamiltonian), the potential can be unbounded from above.In contrast, the transformed potentials from the gradient flow Hamiltonians are non-positive.

In conclusion, the mechanism behind the absence of homogenization in the ran-dom potential system is that the potential is unbounded from above and that thebehavior of the system is dominated by the large fluctuations in the potential. Incontrast, the Hamiltonian of unbounded random advection of gradient type is equiv-alent to a classical-mechanics Hamiltonian with a potential that though unboundedfrom below, is bounded from above, and so satisfies a homogenization principle. Itis an interesting project to unravel more delicate conditions of stochastic HJ homog-enization for other Hamiltonians in unbounded random media. A related problem isto study Hamiltonians with unbounded temporal fluctuations.

We shall see in the next chapter that analogous effective behavior is present forfront speeds in reaction–diffusion–advection equations with unbounded random ad-vection. So the phenomena we discussed in this chapter may have broader impli-cations for wave propagation in random media. For reaction–diffusion fronts in in-compressible random advection, temporal randomness is found to regularize thedominance of extreme events and promote mixing, and the speed of propagation isasymptotically a deterministic constant [247, 179, 182]. It is conceivable that simi-

Page 92: Surveys and Tutorials in the Applied Mathematical Sciences

86 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

lar results (or homogenization) hold for HJ equations in unbounded time-dependentrandom media under suitable conditions.

4.5 Central Limit Theorems and Front Fluctuations

Homogenization gives the leading-order asymptotic behavior uε(x, t) → u(x, t),where the limiting function u is deterministic, analogous to the law of large numbers.More accurate is a stochastic approximation of uε , the central limit theorem (CLT),which carries information on statistics of fluctuations. We discussed such results inChapter 3 on random Burgers fronts. In this section, we broaden our perspective inthe context of convex HJ equations based on [200, 201], thereby extending CLT tofronts of convex conservation laws and HJs.

Let us consider HJ front solutions to (4.1) that travel with effective speeds

u(x, t, p) = w(x, p)− tH(p), (4.37)

where p ∈ RN , and w(x, p) = p · x + o(|x|). These solutions are associated with theexact front solutions to the homogenized HJ (4.2), because in the scaling limit forhomogenization, εw(x/ε, p)→ p · x. In the examples of this section such solutionscan be found, and

wε(x, p)≡ εw( x

ε, p

)d= p · x+

√εB(x, p)+o(

√ε), (4.38)

for a random field B(x, p) continuous in (x, p). Substituting (4.37) in (4.1) showsthat w(x, p) satisfies the time-independent HJ equation

H(x,∇xw(x, p)) = H(p). (4.39)

If we write w(x, p) = p · x+ w(x, p), then w solves the equation

H(x, p+∇xw) = H(p), (4.40)

which is the corrector equation of homogenization, or the so-called cell problem inperiodic homogenization.

Example 4.12. Let H(x, p) = 12 a(x,ω)p2, x ∈ R1, and a(x,ω) > 0 as in Chapter 3.

Then (4.40) implies H ≥ 0, and

p+ wx =√

2H(a(x,ω))−1/2, (4.41)

or

w = w(0)+∫ x

0

(2H)1/2

a1/2(x′)dx′ − px. (4.42)

Let us select H such that w(x,ω)/x→ 0 almost surely as x→ ∞. It follows that

Page 93: Surveys and Tutorials in the Applied Mathematical Sciences

4.5 Central Limit Theorems and Front Fluctuations 87

p = (2H)1/2⟨

a−1/2⟩

,

or

H(p) =p2

2

⟨a−1/2

⟩−2. (4.43)

Letting p = 1, this gives the asymptotic speed formula of Burgers fronts as in Chap-ter 3. By (4.43), (4.42) reads

w(x) = p⟨

a−1/2⟩−1 ∫ x

0a−1/2(x′)dx′ − px, (4.44)

where without loss of generality we have set the constant term w(0) to zero. Ap-plying the invariance principle (A4 in Chapter 3) to (4.44), we find that w(x, p) isexpressed as

w(x, p) d= px+ pσ⟨

a−1/2⟩−1

Wx +o(√

x), (4.45)

implying (4.38) in the scaled form with B(x, p) being p times a Wiener process in x.Suppose (4.38) holds. Let us formally verify the asymptotic expansion

uε(x, t) d= u(x, t)+√

ε infy∈I(x,t)

[B(x, p(x,y, t))−B(y, p(x,y, t))]+o(√

ε), (4.46)

where I(x, t) is the set of minimizers y of the Hopf formula (4.13) and

p(x,y, t) = ∇L((x− y)/t). (4.47)

Let us find an asymptotic expansion of the action function Sε(x,y, t). For frontsolutions uε(x, t, p), we have

uε(x, t, p) = infy

[wε(y, p)+Sε(x,y, t)], (4.48)

where uε is the solution of uεt + H(x/ε,∇xuε) = 0, uε(x,0) = wε(x, p). If the infi-

mum is attained at some yε(x, t, p), we obtain

Sε(x,yε(x, t, p), t) = uε(x, t, p)−wε(yε(x, t, p), p). (4.49)

Suppose that the function y = yε(x, t, p) is invertible for each (x, t), and denote theinverse function by p = pε(x,y, t), or

yε(x, t, pε(x,y, t)) = y. (4.50)

Since uε converges to u(x, t, p) = x · p− tH(p), equal to

u(x, t, p) = infy

[y · p+ tL((x− y)/t)], (4.51)

the minimizer y satisfies p = p(x,y, t) = ∇L((x−y)/t). We anticipate a CLT expan-sion for pε(x,y, t) (or yε ) to come from that of uε :

Page 94: Surveys and Tutorials in the Applied Mathematical Sciences

88 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

pε(x,y, t) = p(x,y, t)+√

εR(x,y, t)+o(√

ε), (4.52)

for a random process R(x,y, t). It follows from (4.37)–(4.38) that

uε(x, t, p) = x · p− tH(p)+√

εB(x, p)+o(√

ε). (4.53)

Now we write (4.49) in terms of the inverse function pε and expand with the helpof (4.52) and (4.53) as

Sε(x,y, t) = uε(x, t, pε(x,y, t))−wε(y, pε(x,y, t))= [(x− y) · p(x,y, t)− t H(p(x,y, t))]

+√

ε[(x− y)− tH(p(x,y, t))] ·R(x,y, t)

+√

ε[B(x, p(x,y, t))−B(y, p(x,y, t))]+o(√

ε). (4.54)

One checks by the Legendre transform that the first term on the right-hand side of(4.54) equals tL((x− y)/t), and the second term is zero. It follows that

Sε(x,y, t) = tL((x− y)/t)+√

ε [B(x, p(x,y, t))−B(y, p(x,y, t))]+o(√

ε). (4.55)

Substituting (4.55) in the Lax formula (4.11) for a more general solution uε(x, t)leads to the form of CLT expansion (4.46)–(4.47), where y is determined by theinfimum of the sum of the O(1) terms. The cancellation of the second term in (4.54)suggests that CLT expansion of pε is not necessary: a low-order expression of pε =p+o(1) suffices. Because the CLT expansions rely on knowledge of w(x, p), whichmay not exist in general [149], we state below two concrete results adapted from[201].

Theorem 4.13. Suppose H(x, p,ω) = a(x,ω)K(p), x ∈ R1, a(x,ω) is bounded be-tween two deterministic positive constants with probability one, and K :R1 → [0,∞)is a coercive, convex, and continuously differentiable function with K(0) = 0. More-over, a(x,ω) satisfies stationarity and ergodicity for homogenization. The functionK has two branches of inverses, A+ : [0,∞)→ [0,∞), A− : [0,∞)→ (−∞,0], so thatK(A±(λ )) = λ ≥ 0. Let ϕ±(λ ) = E[A±(λ/a(0,ω))]. Define

H(p) = (ϕ+)−1(p)χ(p≥ 0)+(ϕ−)−1(p)χ(p < 0) (4.56)

and

w(x, p) = χ(p≥ 0)∫ x

0A+(H(p)/a(y,ω))dy+ χ(p < 0)

∫ x

0A−(H(p)/a(y,ω))dy.

(4.57)Then

1. w(x, p) and wx(x, p) are continuous in (x, p);2.

limp→∞

Hp(x,wx(x, p)) =±∞,

uniformly in x;

Page 95: Surveys and Tutorials in the Applied Mathematical Sciences

4.5 Central Limit Theorems and Front Fluctuations 89

3. H(x,wx(x, p)) = H(p), for all (x, p).

Assume additionally that H is strictly convex (no degenerate linear pieces), andthat the process 1√

ε (εw(x/ε, p)− p · x) converges to a continuous process B(x, p).Then any finite-dimensional distributions of the process (error of homogenization)

γε(x, t,ω) = ε−1/2(uε(x, t,ω)− u(x, t))

converge to those of the process

γ(x, t) = infy∈I(x,t)

[B(x, p(x,y, t))−B(y, p(x,y, t))], (4.58)

where p = p(x,y, t) = L′((x− y)/t), the prime denoting the derivative and L beingthe Legendre transform of H.

The additional convergence assumption in the above theorem holds in the con-text of the invariance principle under the sufficient mixing conditions of a(x,ω),as in Chapter 3 for Burgers equations. As a result, the asymptotics (4.38) is validand B(x, p) is a Gaussian process in x, implying that the random variable γ(0,1) isGaussian. Hence the front speed fluctuation of convex HJ in positive multiplicativeone-dimensional random media obeys Gaussian statistics.

Next we consider the Hamiltonians of the type in classical mechanics:

H(x, p,ω) = K(p)+V (x,ω), x ∈ R1, (4.59)

where V (x,ω) is a stationary and ergodic process for homogenization, and −∞ <infx V (x,ω) ≤ 0 with probability one; K : R1 → [0,∞) is a continuously differen-tiable strictly convex coercive function with K(0) = 0. So K(p) is monotone in-creasing (decreasing) for p > 0 (p < 0). There are two branches of inverse functionsA±. Set ϕ±(λ ) = E[A±(λ −V (0,ω))], both strictly monotone for λ ≥ 0. Define

H(p) = χ(p≤ ϕ−(0))(ϕ−

)−1 (p)+ χ(p≥ ϕ+(0))(ϕ+)−1 (p), (4.60)

and L its Legendre transform (homogenized Lagrangian). One may check that ifz 6= 0, then L is differentiable at z with value L′(z) ∈ [ϕ−(0),ϕ+(0)].

Theorem 4.14. Assume that the processes

Bε±(x,λ ,ω)≡ ε−1/2

∫ x/ε

0A±(λ −V (y,ω))dy− xϕ±(λ )

)(4.61)

converge in distribution to a continuous process B±(x,λ ) for any (x,λ ) ∈ R1 ×[0,∞). Define

B(x, p) = χ(p≥ ϕ+(0))B+(x,(ϕ+)−1(p))+ χ(p≤ ϕ−(0))B−(x,(ϕ−)−1(p)).(4.62)

Assume also that with probability one,

Page 96: Surveys and Tutorials in the Applied Mathematical Sciences

90 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

limn→∞

1n2

∫ n

−n

dzK′(A±(−V (z,ω))

=±∞. (4.63)

Then any finite-dimensional distributions of the process (error of homogenization)

γε(x, t,ω) = ε−1/2(uε(x, t,ω)− u(x, t)),

converge to those of the process

γ(x, t) = infy∈I(x,t)−x

[B(x, p(x,y, t))−B(y, p(x,y, t))], (4.64)

where p = p(x,y, t) = L′((x− y)/t).

Because of the degeneracy of H(p), namely H(p) = 0 for p ∈ [ϕ−(0),ϕ+(0)],the function w(x, p) is complicated. The condition (4.63) allows one to use onlyw(x, p) for p outside of [ϕ−(0),ϕ+(0)] and a solution formula of the form

w(x, p) =∫ x

0A±(H(p)−V (y,ω))dy

for analysis. This formula is used in (4.61). In the quadratic Hamiltonian case,K(p) = p2/2, condition (4.63) becomes

limn→∞

1n2

∫ n

−n

dz√sup V (z,ω)−V (z,ω)

= +∞,

which is satisfied if V is twice differentiable and so has the asymptotics V (z,ω)−V (z0,ω) = O(|z− z0|2) near a local maximal point z0 that belongs to (−n,n) for nlarge enough.

Extensions of the above results to several spatial dimensions are possible ifthe existence of w(x, p) and the asymptotic property (4.38) are established. Thisis a challenging problem by itself. Some special cases are known, for exampleH(x, p,ω) = H((b(x,ω))−1(p−c(x,ω)), where H is a strictly convex continuouslydifferentiable and coercive function, b is an invertible matrix function, c(x) is a vec-tor function, and both b and c are in gradient form; see [200, 201].

4.6 Exercises

1. Generalize Theorem 4.8 to the random Hamiltonian

H(x, p,ω) = |p|2n +V (x,ω),

where n is a positive integer.

Page 97: Surveys and Tutorials in the Applied Mathematical Sciences

4.6 Exercises 91

2. Suppose that the action function Sε(x,y, t), (x,y, t) ∈ R2N × (0,+∞), is contin-uous in (x,y, t) for any ε and almost all ω , and that for a convex function L, theprocess

Zε(x,y, t,ω) = ε−1/2(Sε(x,y, t,ω)− tL((x− y)/t)

converges to a continuous process Z(x,y, t,ω). Use the Lax formula to showthat finite-dimensional distributions of

γε(x, t,ω) = ε−1/2(uε(x, t,ω)− u(x, t)),

converge to those ofinf

y∈I(x,t)Z(x,y, t),

where I is the set of minimizers in the Hopf formula for u(x, t).

3. Let H(x, p) be convex in p and continuously differentiable in (p,x) ∈ R2N , andL its Legendre transform in p. Suppose that w is a continuously differentiablesolution of

H(x,∇xw(x)) = λ ,

for some constant λ , and that x(t) is a solution of

dxdt

= Hp(x,∇xw(x)),

where Hp is the gradient of H with respect to p. Show that:

a. L(x,Hp(x, p)) = p ·Hp(x, p)−H(x, p).b. ∫ t

0L(x(s), x(s))ds = w(x(t))−w(x(0))−λ t.

c. For any t > 0,w(a)−λ t = inf

y[w(y)+S(a,y, t)] (4.65)

and

S(x(t),x(0), t) =∫ t

0L(x(s), x(s))ds = w(x(t))−w(x(0))−λ t. (4.66)

4. Under the assumptions of Theorem 4.13:

a. Apply (4.65)–(4.66) to show that

S(x,y, t,ω) = supp

[w(x, p,ω)−w(y, p,ω)− tH(p)], (4.67)

where x = x(t,y, p) is the unique solution of

dxdt

= Hp(x,wx(x, p)), (4.68)

Page 98: Surveys and Tutorials in the Applied Mathematical Sciences

92 4 Fronts and Stochastic Homogenization of Hamilton–Jacobi Equations

x(0,y, p) = y.b. Use (4.67) and that limp→±∞ Hp(x,wx(x, p)) =±∞ to show that

limp→±∞

x(t,y, p) =±∞.

c. Deduce from part b above and (4.68) that for each (x,y, t) there exists a psuch that x(t,y, p) = x. Hence (4.67) holds for all (x,y, t).

Page 99: Surveys and Tutorials in the Applied Mathematical Sciences

Chapter 5KPP Fronts in Random Media

In this chapter, we discuss KPP fronts in space and/or time random flows by a com-bination of PDE and probabilistic methods. We first consider fronts in spatial ran-dom shear flows in a channel domain with finite width L. Here randomness appearsin the transverse direction of the front. The variational principle (2.52) under zeroNeumann boundary condition applies to each random speed, and allows a combinedPDE–probabilistic analysis and a computational study of the speed ensemble. Weidentify new random phenomena such as the resonance in speed dependence oncorrelation length of the flows, speed slowdown due to temporal decorrelation ofrandom flows, and speed divergence due to extreme behavior of random media. Forexample, for channel domain width L À 1, front speed diverges due to unboundedrunning maxima of the shear flow, which shares the same source of divergence asstochastic homogenization of Hamilton–Jacobi equations in Chapter 4. We shallpresent front growth or decay laws in random media, in comparison with those ofperiodic media in Chapter 3. Finally, we outline a recent breakthrough in solving theturbulent front speed problem for the KPP model, and study related speed bounds.

5.1 KPP Fronts in Spatially Random Shear Flows

Let us consider KPP front speeds through random shear flows in channel domainsD≡R×Ω , where Ω ⊂Rn−1, n≥ 2, is a bounded simply connected domain with asmooth boundary. We address the enhancement properties of the ensemble-averagedfront speeds, and their dependence on the flow statistics. The materials are based on[175, 178]. Recall the advection–diffusion equation with KPP reaction

ut = ∆x,yu+B ·∇x,yu+ f (u), (5.1)

where t ∈ R+, ∆x,y the n-dimensional Laplacian, (x,y) ∈ D. The vector field B =(b(y,ω),0), where b(y,ω) is a stationary continuous scalar random process in y, hasits ensemble mean equal to zero. The zero Neumann boundary condition is imposed

© Springer Science + Business Media, LLC 2009

J. Xin, An Introduction to Fronts in Random Media, Surveys and Tutorials in the Applied 93Mathematical Sciences 5, DOI: 10.1007/978-0-387-87683-2_5,

Page 100: Surveys and Tutorials in the Applied Mathematical Sciences

94 5 KPP Fronts in Random Media

at ∂Ω : ∂u∂ν = 0, ν the unit outward normal. The KPP speed c∗ variational formula

(2.52) in Chapter 2 simplifies to

c∗ = c∗(ω) = infλ>0

µ(λ ,ω)λ

, (5.2)

where µ(λ ,ω) is the principal eigenvalue with corresponding eigenfunction φ > 0of the problem

Lλ φ = ∆yφ +[λ 2 +λb(y,ω)+ f ′(0)]φ = µ(λ ,ω)φ , y ∈Ω , (5.3)∂φ∂ν

= 0, y ∈ ∂Ω . (5.4)

We shall use the superscript ∗ to denote speed in random media in this chapter,which is different from the speed c∗ in a deterministic medium. The derivation of(5.2)–(5.4) from (2.52) is left as an exercise.

5.1.1 Asymptotics of Averaged Speeds

Let us first study the deterministic case in which b in equation (5.3) is a continuousfunction with

∫Ω b(y)dy = 0. We shall see what quantities appear in estimating the

front speed, then extend them properly to the random setting.

Proposition 5.1. Let χ = χ(y) solve ∆yχ =−b, y ∈Ω , with zero Neumann bound-ary condition, where b ∈ C(Ω) has zero integral over Ω . Then for δ sufficientlysmall, the minimal speed has the expansion

c∗(δ ) = c0 +c0δ 2

2|Ω |∫

Ω|∇χ|2 dy+O(δ 3). (5.5)

We shall give a proof of Proposition 5.1 using variational formulas, and latergeneralize it to the random case. A helpful fact is that the infimum in (5.2) can berestricted to a bounded set independent of b and δ , as stated in the following lemma:

Lemma 5.2. Let b ∈C(Ω) have zero mean over Ω , and let λ0 =√

f ′(0). Then

infλ>0

µ(λ )λ

= inf0<λ≤λ0

µ(λ )λ

. (5.6)

Proof. For each c > 0, we let ρc(λ ) = µ(λ )−λc−λ 2. So if φ > 0 is the eigenfunc-tion defined by (5.3), then ρc(λ ) is the principal eigenvalue defined by the equation

∆yφ +[λb(y)−λc+ f ′(0)]φ = ρc(λ )φ , y ∈Ω .

One can readily verify that ∂λ ρc(λ )|λ=0 = −c < 0. The variational formula (5.2)can be expressed as

Page 101: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 95

0

f’(0)

ρ* (λ )

ρc (λ )

ρ ∗(λ) = −λ2

λ1 λ0 =√

f′(0)

Figure 5.1 Intersecting curves ρc and −λ 2, from [178].

c∗ = infc : ∃λ > 0,λc = µ(λ )= inf

c : ∃λ > 0,ρc(λ ) =−λ 2 . (5.7)

Consider the points where ρc(λ ) = −λ 2. By [31, Proposition 2.1], the continuouscurve λ 7→ ρc(λ ) is convex in λ , for each c > 0. Also, ρc(0) = f ′(0) > 0. Therefore,for a given c > 0, there can be at most two values of λ > 0 such that ρc(λ ) =−λ 2.The line ρ∗(λ ) =−2

√f ′(0)λ + f ′(0) satisfies ρ∗(λ )≥−λ 2, with equality holding

only at one point: λ0 =√

f ′(0). Since ρ∗(0) = ρc(0) and ρc(λ ) is convex and ρ∗ is aline, ρc(λ ) =−λ 2 for some λ > 0 only if ρc(λ1) =−λ 2

1 for some λ1 ∈ (0,λ0]. Thispoint is illustrated in Figure 5.1.1. The solid curve represents the parabola −λ 2.If ρc(λ ) intersects −λ 2, then one of the intersection points must be to the left ofλ0 =

√f ′(0). Therefore, from (5.7),

c : ∃λ > 0,ρc(λ ) =−λ 2 =

c : ∃λ ∈ (0,λ0],ρc(λ ) =−λ 2 .

So we conclude that

c∗(δ ) = inf0<λ

µ(λ )λ

= inf0<λ≤λ0

µ(λ )λ

. (5.8)

This completes the proof utNow we turn to the proof of Proposition 5.1.

Proof (of Proposition 5.1). To estimate c∗(δ ), we bound the principal eigenvalueµ(λ ) using two different representations of µ . First, since L is a self-adjoint opera-tor, we have

Page 102: Surveys and Tutorials in the Applied Mathematical Sciences

96 5 KPP Fronts in Random Media

µ = sup(Lλ ψ,ψ)‖ψ‖2

2, (5.9)

where the supremum is taken over all ψ ∈ H2(Ω) such that ∂ψ∂ν = 0 on ∂Ω . The

other representation is

µ = infψ

supy∈Ω

Lλ ψψ

= infψ

supy∈Ω

(∆ψψ

+λδb+λ 2 + f ′(0))

, (5.10)

where the infimum can be taken over all ψ ∈C1(Ω) such that ∆ψ ∈C(Ω), ψ > 0,and ∂ψ

∂ν = 0 on ∂Ω . This representation follows from the fact that the eigenfunctionφ > 0 lies in the kernel of the self-adjoint operator (Lλ −µ(λ )I) = (Lλ −µ(λ )I)∗.So, if we have the strict inequality

Lλ ψ−µ(λ )ψ = m < 0, (5.11)

then the Fredholm alternative implies that (φ ,m)L2 = 0, a contradiction, since φ > 0,m < 0. Hence

supy∈Ω

Lλ ψψ

≥ µ(λ ). (5.12)

Since Lλ φ = µ(λ )φ , the formula (5.10) follows. Note that we do not require thetest functions ψ to be C2(Ω), only ∆ψ ∈C(Ω). This is important, since we do notwant to require the shear b(y) to be any more regular than b ∈C(Ω).

Let us derive upper and lower bounds for µ(λ ) by choosing test functions ψ as

ψ = 1+λδ χ +λ 2δ 2h, (5.13)

where χ = χ(y) and h = h(y) solve

∆ χ =−b, ∆h =−bχ + k, (5.14)

with zero Neumann boundary conditions at ∂Ω , and k a constant equal to

k =1|Ω |

Ωbχ dy =

1|Ω |

Ω|∇χ |2 dy. (5.15)

We normalize χ and h so that

infx∈Ω

χ(x) = 0, infx∈Ω

h(x) = 0. (5.16)

ThenLλ ψ = λ 2δ 2k +λ 3δ 3bh+(λ 2 + f ′(0))ψ

and(Lλ ψ,ψ)‖ψ‖2

2= λ 2δ 2k

∫ψ∫ψ2 +λ 3δ 3

∫bhψ∫ψ2 +λ 2 + f ′(0). (5.17)

Page 103: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 97

Using the definition of ψ , we see that ψ = ψ2−λδ χψ−λ 2δ 2hψ and∫

Ω ψ dy∫Ω ψ2 dy

= 1−λδ∫

Ω χψ dy∫Ω ψ2 dy

−λ 2δ 2∫

Ω hψ dy∫Ω ψ2 dy

.

Now from (5.9) and (5.17) we have the lower bound

µ(λ )≥ λ 2 + f ′(0)+λ 2δ 2k +R1, (5.18)

with

R1 =− λ 3δ 3∫

Ω ψ2

(k∫

Ωχψ + kλδ

Ωhψ−

Ωbhψ

). (5.19)

By choice of χ ≥ 0 and h≥ 0, we have∫

Ω ψ2 ≥ |Ω | for all δ ≥ 0 and λ > 0. Hence,R1 = O(δ 3) for λ bounded. Returning to the variational formula (5.8), we now havea lower bound on c∗(δ ):

c∗(δ ) = inf0<λ≤λ0

µ(λ )λ

≥ inf0<λ≤λ0

(λ +

f ′(0)λ

+λδ 2k +R1

λ

)

≥ infλ>0

(λ +

f ′(0)λ

+λδ 2k)

+O(δ 3)

= 2√

f ′(0)(1+δ 2k)+O(δ 3)

= c0 +c0δ 2k

2+O(δ 3). (5.20)

To obtain an upper bound on c∗(δ ), we use (5.10) and calculate

Lλ ψψ

=∆ψψ

+λδb+λ 2 + f ′(0) =λ 2δ 2k +λ 3δ 3bh1+λδ χ +λ 2δ 2h

+λ 2 + f ′(0). (5.21)

Since χ ≥ 0 and h≥ 0, we see from (5.10) and (5.21) that

µ(λ )≤ supy∈Ω

Lλ ψψ

≤ λ 2 + f ′(0)+λ 2δ 2k +R2, (5.22)

withR2 = λ 3δ 3‖bh‖∞. (5.23)

The variational formula (5.8) implies

c∗(δ ) = inf0<λ≤λ0

µ(λ )λ

≤ inf0<λ≤λ0

(λ +

f ′(0)λ

+λδ 2k +R2

)

= c0 +c0δ 2k

2+O(δ 3), (5.24)

thus completing the proof. ut

Page 104: Surveys and Tutorials in the Applied Mathematical Sciences

98 5 KPP Fronts in Random Media

When the shear b(y,ω) is a random process, the corresponding minimal speedc∗(δ ) = c∗(δ ,ω) is a random variable for each δ , and we consider how the ex-pectation E[c∗(δ )] scales with the parameter δ by finding an exponent p such thatE[c∗(δ )] = c∗(0)+O(δ p). Each realization of the random process b(y,ω) restrictedto the domain Ω does not necessarily have zero integral over Ω . Nevertheless, eachrealization can be written in the form

b(y,ω) = b(ω)+b1(y,ω), (5.25)

where b(ω) = 〈b(y,ω)〉 is the mean of b over Ω , and b1(y,ω) is the variation aboutthe mean value. For a fixed realization, the minimal speed c∗(δ ) will be affected byboth the scaling of the mean b(ω) and the scaling of the variation b1(y,ω). That is,

c∗(δ ,ω) = c∗0 +δ b(ω)+M(δ ,ω), (5.26)

where the remainder M(δ ,ω) is the enhancement due to the variation b1(y,ω), dif-ferent for each realization. Taking the expectation of both sides of (5.26), we have

E[c∗(δ )] = c∗0 +δE[b(ω)]+E[M(δ ,ω)]. (5.27)

Though for each sample, M(δ ,ω) is O(δ 2) for δ small, we show that E[M(δ )]exhibits the same quadratic scaling for enhancement of averaged front speeds undersuitable moment conditions of the shear.

Theorem 5.3. Let b(y,ω) be a stationary random process in Rn−1 (n≥ 2) such thatsample paths are almost surely continuous and such that

E[‖b‖6∞] < +∞. (5.28)

Then for δ small, the expectation E[c∗(δ )] has the expansion

E[c∗(δ )] = c0 +δE[〈b〉]+ c0δ 2

2|Ω |∫

ΩE[|∇χ |2]dy+O(δ 3), (5.29)

where b(y,ω) = 〈b〉(ω)+b1(y,ω) and χ = χ(y,ω) solves ∆yχ =−b1, y ∈Ω , sub-ject to zero Neumann boundary condition.

Proof. Since the contribution of the integral average 〈b〉 to c∗ is an additive constant,it suffices to consider shear flow b1 and show that it gives the averaged speed

E[c∗(δ )] = c0 +c0δ 2

2|Ω |∫

ΩE[|∇χ|2]dy+O(δ 3). (5.30)

We adapt the proof of Proposition 5.1, noting that in the stochastic case the remain-ders R1 and R2 defined by (5.19) and (5.23) are random and not bounded uniformlyfor all realizations. Instead, we will show that for λ in a bounded interval,

E[|R1|]≤ O(δ 3), E[|R2|]≤ O(δ 3).

Page 105: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 99

To this end, we estimate χ and h, with C denoting a generic positive constant de-pending only on the domain Ω and its dimension. Let χ and h solve (5.14) with zerointegral averages 〈χ〉= 〈h〉= 0. Applying W 2,p estimates [102], we have

‖χ‖W 2,p(Ω) ≤C‖b1‖Lp(Ω) ≤C|Ω |1/p‖b1‖∞ (5.31)

and

‖h‖W 2,p(Ω) ≤C‖b1χ + k‖Lp(Ω)

≤C‖b1‖∞‖χ‖Lp(Ω) +Ck|Ω |1/p

≤C‖b1‖2∞, (5.32)

sincek = 〈|∇χ |2〉 ≤C‖b1‖2

∞. (5.33)

Given α ∈ (0,1), we can choose p > 1 sufficiently large, depending on n, such thatW 2,p(Ω) embeds continuously into C1,α(Ω). It follows that there is a constant C > 0independent of b such that

‖χ‖C1(Ω) ≤C‖b1‖∞, ‖h‖C1(Ω) ≤C‖b1‖2∞. (5.34)

If instead we normalize χ and h by (5.16), then the bounds (5.34) still hold, withdifferent constants. Note that adding a constant to χ and h does not alter the quantity∫

Ω |∇χ|2 that appears in the asymptotic expansion.Now by (5.34), the integrals in R1 are easily bounded as

Ωχψ =

Ωχ +λδ χ2 +λ 2δ 2χh

≤C(‖b1‖∞ +λδ‖b1‖2

∞ +λ 2δ 2‖b1‖3∞).

Similarly,∫

Ωhψ =

Ωh+λδ χh+λ 2δ 2h2

≤C(‖b1‖2

∞ +λδ‖b1‖3∞ +λ 2δ 2‖b1‖4

∞)

and∣∣∣∣∫

Ωb1hψ dy

∣∣∣∣ =∣∣∣∣∫

Ωb1h+λδ

Ωb1hχ +λ 2δ 2

Ωb1h2

∣∣∣∣

≤C(‖b1‖3

∞ +λδ‖b1‖4∞ +λ 2δ 2‖b1‖5

).

Since χ and h are nonnegative, it follows that∫

Ω ψ2 dy ≥ C∫

Ω ψ = C|Ω | > 0,for any realization. So for λ in a bounded interval and δ small, we bound (5.19) by

|R1| ≤Cδ 3λ 3(1+‖b1‖6∞),

Page 106: Surveys and Tutorials in the Applied Mathematical Sciences

100 5 KPP Fronts in Random Media

so that E[|R1|]≤ O(δ 3).To bound E[|R2|], we use the normalization (5.16) and the above estimates:

|R2|= λ 3δ 3‖b1h‖∞ ≤Cλ 3δ 3‖b1‖3∞. (5.35)

Hence E[R2]≤O(δ 3) for λ in a finite interval. Now we return to (5.20) to concludethat

E[c∗(δ )]≥ E[2√

f ′(0)(1+δ 2k)]+O(δ 3) = c0 +c0δ 2E[k]

2+O(δ 3),

since E[k2]≤CE[‖b‖4∞] < ∞.

The opposite inequality follows from (5.24), since E[R2] = O(δ 3) for λ ∈ (0,λ0).Thus formula (5.30) holds. For general b, not necessarily mean zero, we have

E[c∗(δ )] = c0 +δE[〈b〉]+ c0δ 2

2E[k]+O(δ 3)

= c0 +δE[〈b〉]+ c0δ 2

2|Ω |∫

ΩE[|∇χ|2] dy+O(δ 3).

The proof is finished. utIf we choose the Ornstein–Uhlenbeck (O-U) process in Chapter 1 for shear b,

then E[〈b〉] = 0. Let us show below that the O-U process, denoted by X(y,ω), satis-fies the conditions in Theorem 5.3, and so E[c∗(δ )] scales quadratically with δ forδ small.

Corollary 5.4 (Explicit Average Speed Formula).Consider the O-U process b(y,ω) as a solution of the stochastic differential (Ito)

equationdX(y) =−aX(y)dy+ rdW (y), y ∈ [0,L], (5.36)

where W (y,ω) is the standard Wiener process, and X(0,ω) = X0(ω) is a Gaussianrandom variable with mean zero and variance ρ = r2/(2a). Then X(y,ω) satisfiesthe moment conditions in Theorem 5.3. The averaged KPP front speed in the channelR× [0,L] is given by

E[c∗(δ )] = c0 +c0δ 2

2enh+O(δ 3), δ ¿ 1, (5.37)

where

enh =r2

2a

(e−aL

(4

L2a4 −1

3a2

)+

L3a− 4

L2a4 −5

3a2 +4

La3

).

Proof. The O-U process is stationary and Markov. Its sample paths are almost surelyHolder continuous though nowhere differentiable. The process can be written as

Page 107: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 101

b(y,ω) = e−ayb(0,ω)+ r∫ y

0e−a(y−s)dWs(ω). (5.38)

The covariance function of this process is ρe−a|y−s|. Letting g(y,ω) denote the pro-cess

g(y) = eayb(y,ω) = g(0,ω)+ r∫ y

0easdWs(ω), (5.39)

we see that g(y,ω) is a martingale [125]. By Doob’s martingale moment inequality[125], for any p ∈ (1,+∞),

E[

sup0<y<L

|g(y)|p]≤

(p

p−1

)p

E[|g(L)|p]. (5.40)

Since the process b(y,ω) is Gaussian, (5.39) and (5.40) imply that

E[‖b‖6

]≤C E[|b(L)|6] < +∞. (5.41)

Formula (5.37) now applies to the average speed. Notice that

(χx(x))2 =∫ x

0

∫ x

0b1(s)b1(y)dsdy

andE[(χx(x))2] =

∫ x

0

∫ x

0E[b1(s)b1(y)]dsdy. (5.42)

Let us calculate E[b1(s)b1(y)] in terms of E[b(s)b(y)]. Define

g(y) = 〈 f (·,y)〉, or g(s) = 〈 f (s, ·)〉,

so that

E[b1(y)b1(s)] = E[b(s)b(y)]−E[b(s)b]−E[b(y)b]+E[b2],

E[b(s)b] =1L

∫ L

0E[b(y)b(s)] dy = g(s),

E[b2] =1L2

∫ L

0

∫ L

0E[b(s)b(y)] dy ds = 〈g〉.

Thus E[b1(y)b1(s)] = f (s,y)+ 〈g〉−g(y)−g(s). Now, we have

E[(χx(x))2] =∫ x

0

∫ x

0E[b1(s)b1(y)]dsdy

=∫ x

0

∫ x

0f (s,y)+ 〈g〉−g(y)−g(s)dsdy

= x2〈g〉−2x2〈g〉x +∫ x

0

∫ x

0f (s,y)dsdy,

Page 108: Surveys and Tutorials in the Applied Mathematical Sciences

102 5 KPP Fronts in Random Media

where 〈g〉x denotes the average of g over the interval [0,x], for 0 < x ≤ L. Conse-quently, we have

E[〈|χx|2〉] = 1L

∫ L

0

(x2〈g〉−2x2〈g〉x +

∫ x

0

∫ x

0f (s,y)dsdy

)dx. (5.43)

Using the O-U covariance function, we obtain

g(s) =1L

∫ L

0f (s,y)dy =

r2

2a1L

∫ L

0e−a|y−s| dy

=r2

2a

(1− e−as

La+

1− e−a(L−s)

La

)

and

〈g〉x =r2

2a1x

∫ x

0

(1− e−as

La+

1− e−a(L−s)

La

)ds

=r2

2a

(2

La+

1xLa2

(e−ax−1

)+

1xLa2

(e−aL− e−a(L−x)

)).

Letting x = L, we have

〈g〉=r2

2a

(2

La+

2L2a2

(e−aL−1

)).

Similarly, ∫ x

0

∫ x

0f (s,y)dsdy =

r2

2a

(2xa

+2a2

(e−ax−1

)).

Combining the above, we have

E[〈|χx|2〉] = r2

2a

(2L3a

+2

3a2

(e−aL−1

))

− r2

2a2L

∫ L

0

2x2

La+

xLa2

(e−ax−1

)+

xLa2

(e−aL− e−a(L−x)

)dx

+r2

2a1L

∫ L

0

2xa

+2a2

(e−ax−1

)dx

=r2

2a

(e−aL

(4

L2a4 −1

3a2

)+

L3a− 4

L2a4 −5

3a2 +4

La3

).

(5.44)

In view of (5.37), the proof is complete. ut

Page 109: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 103

We see from formula (5.37) that the correction term of E[c∗] diverges as thedomain width L goes to +∞, suggesting that KPP front speed in a Gaussian spatialshear flow on the whole plane R2 is infinite.

Theorem 5.5 (Linear Growth). If the stationary shear process b(y,ω) has almostsurely continuous sample paths and satisfies E[‖b‖∞] < ∞, then the amplified shearfield δb(y,ω) generates the average front speed:

E[|c∗(δ ,ω)|] = O(δ ), δ À 1.

Moreover, limδ→∞ E[|c∗(δ ,ω)|]/δ exists.

Proof. By (2.80), for all ω , |c∗(δ ,ω)|/δ → d∗(ω) as δ → ∞, where d∗ is finite foreach ω . Now recall the upper bound |c∗(δ ,ω)| ≤ |c0|+δ‖b‖∞. Hence for δ > |c0|,we have |c∗(δ ,ω)|/δ ≤ 1+‖b‖∞ ≡ Y , and E(Y ) < ∞. The dominated convergencetheorem implies that

E[ |c∗(δ ,ω)|

δ

]→ E[d∗(ω)]≤ E(Y ).

The proof is finished. utThe O-U process satisfies the required condition for linear average speed growth.By (2.80), the growth rate d∗(ω) is

d∗(ω) = supψ∈D1

Ωb(y,ω)ψ2(y)dy,

whereD1 =

ψ ∈ H1(Ω) : ‖∇ψ‖2

2 ≤ f ′(0),‖ψ‖2 = 1

.

If a realization of b were to have a flat piece near the maximal point of b in Ω , thenchoosing the test function ψ supported near the maximal point within the flat piecewould imply that d∗(ω) equals supΩ b(y,ω).

However, this happens with zero probability for an O-U process because it isoscillatory in every small neighborhood of the maximal point. The distribution ofd∗(ω) appears to be an open problem. However, numerical simulation suggests incertain parameter regimes (next subsection) that d∗(ω) and supΩ b(y,ω) are quitecorrelated. The latter is a running maximum and has a well-studied limit law as thedomain size grows; see Theorem 4.9.

5.1.2 Computing Front Speeds by the Variational Principle

Let us briefly discuss the numerical procedure of computing c∗(ω) and its statis-tics. Let n = 2. For a given λ > 0, we compute the principal eigenvalue µ(λ ) withcorresponding eigenfunction φ = φ(y) > 0, y ∈ [0,L], by solving

Page 110: Surveys and Tutorials in the Applied Mathematical Sciences

104 5 KPP Fronts in Random Media

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−4

−2

0

2

4

6

b(y,

ω)

Figure 5.2 One sample path of the Ornstein–Uhlenbeck process b(y,ω), from [178].

φyy +[λ 2 +λ b(y)+ f ′(0)]φ = µ(λ )φ , y ∈ (0,L),∂φ∂y

= 0, y = 0, L, (5.45)

using a standard second-order finite difference method. Here we suppress the ran-dom parameter ω , since computation is done realization by realization. Denote theuniform partition of the domain by points by yim

i=1 and the numerical solution byφ =

φi

mi=1, where h = L/(m− 1), yi = (i− 1)h, and φi ≈ φ(yi). The discretized

system is

1h2 φi−1 +

(λ 2 +λbi + f ′(0)− 2

h2

)φi +

1h2 φi+1 = µ(λ )φi i = 2, . . . ,m−1,

with second-order approximation of the Neumann boundary conditions. This re-duces to finding the principal eigenvalue of a symmetric tridiagonal matrix, a prob-lem with efficient algorithms in numerical linear algebra such as double-precisionLAPACK routines [6]. Then we compute points on the curve H(λ ) = µ(λ )/λ , andminimize over λ using a Newton’s method with line search.

Our approximation decreases H with each iteration and converges quadraticallyin the region near the infimum. We generate realizations of the shear process b(y,ω)by numerically evaluating the stochastic ODE (5.36) by a second-order-accuratescheme in the parameter h [133]. Figure 5.2 shows an O-U sample path. The O-Uparameters are ρ = 2, a = 4.

To approximate the expectation E[c∗(δ )], we generate N independent realiza-tions (indexed by i = 1, . . . ,N) of the shear and compute the corresponding minimalspeeds c∗i for each δ . Then we compute the average

E[c∗(δ )]≈ E(δ ) = c∗0 +1N

N

∑i=1

Mi(δ ), (5.46)

where Mi(δ ) = c∗i (δ )−c∗0−δ bi. That is, we subtract the linear part due to the meanof the shear being nonzero, as in (5.26). Once we have the averages E(δ ) for each δ ,

Page 111: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 105

0 5 10 15 20 25 30 35 402

2.05

2.1

2.15

2.2

2.25

2.3

2.35

2.4

Covariance parameter α

E[c

* (δ)]

, δ=

1.0

L=1L=2L=3L=4

0 5 10 15 20 25 302

4

6

8

10

12

14

16

Covariance parameter α

E[c

* (δ)]

, δ=

15.0

L=1L=2L=3L=4

Figure 5.3 Effect of covariance parameter α on the speed enhancement for L = 1 (line–circle), 2(line–square), 3 (line–triangle), 4 (line–star) at δ = 1.0 (left) and δ = 15.0 (right), from [178].

we compute the exponents p using the least-squares method to fit a line to a log-logplot of speed versus amplitude. That is, the exponent p is the slope of the best-fit linethrough the data points (log(δ ), log(E[c∗(δ )]− c∗(0))) for each shear amplitude δ .

Besides confirming the quadratic and linear laws of E[c∗], we perform a param-eter study of the dependence of the statistics of c∗ on the correlation length of theflow and domain width. Let us consider the effect of the covariance on the enhance-ment of c∗. The covariance E[b(y)b(s)] is a function of |t| = |s− y|, so we writeV (t) = E[b(y)b(s)]. By choosing r =

√2α3/4, we construct O-U processes with

covariances given byV (t) =

√αe−α|t| (5.47)

By this choice of r, the L2 norm of V (t) remains constant as α changes, so that thetotal energy in the power spectrum of the signal remains constant. Since r2

2a =√

a,we see from equation (5.37) that for fixed L,

limα→+∞

E[〈|χx|2〉] = limα→0+

E[〈|χx|2〉] = 0, (5.48)

and that E[〈|χx|2〉] achieves a maximum for some finite value of α ∈ (0,∞). Thissuggests that there is some optimal α , depending on the domain size L, such thatthe enhancement of E[c∗(δ )] is maximized.

Fixing the grid spacing dx = 0.002, we computed the expected value E[c∗(δ )] fora range of α and for L = 1.0,2.0,3.0,4.0. Note that for each α , we must choose theinitial points b0 to have variance E[b2

0] =√

α , so that the process remains stationaryfor each α . Varying the covariance does not affect the order of the scaling in δ . Thatis, in each case the enhancement scales like O(δ 2) for small δ and O(δ ) for large δ .

Figure 5.3 shows the enhancement E[c∗(δ )] for δ = 1,15, and a range of α . Theenhancement peaks at an optimal covariance parameter α .

This resonance effect can be interpreted in terms of V (t) and its Fourier transform(power spectrum)

Page 112: Surveys and Tutorials in the Applied Mathematical Sciences

106 5 KPP Fronts in Random Media

V (w) =

√2

πα

(1+

(wα

)2)−1

. (5.49)

As α → 0, V (w) concentrates at the origin, and so the energy of the shear processis concentrated more in the large-scale spatial modes. The domain Ω , to which theprocess is restricted is bounded, and variations over a length scale that is muchgreater than the diameter of Ω have little effect on the average enhancement of thefront. As a result, E[c∗] decreases as α → 0+. In the other limit α → ∞, V spreadsout so that the energy over any finite band of frequencies goes to zero, causing E[c∗]to decrease as well. Note that V → 0 in L1 as α → ∞, so even though V spreads outmore uniformly as α → 0, the family of processes does not converge to white noise,whose covariance function equals the Dirac delta function.

We also observe that as L increases with δ fixed, the expectation E[c∗(δ )] growssublinearly and the variance Var[c∗(δ )] decreases. The average speed obeys the up-per bound

E[c∗(δ )]≤ c∗0 +δE[ supy∈[0,L]

b(y)]. (5.50)

We modify the diffusion constant to equal to 0.01 (set to 1 in equation (5.1)). Recallfrom (2.80) that in the regime of small diffusion coefficient, the ratio c∗(δ )/δ isclose to supy∈[0,L] b(y) for δ À 1. By Theorem 4.9 and (5.50), the growth of E[c∗(δ )]with respect to L can be no more than O(

√logL).

We observed that at δ = 50, diffusion constant is 0.01, and the mean and varianceof c∗(δ )/δ mimic the mean and variance of supy∈[0,L] b(y) as LÀ 1. In Figure 5.4,we compare E[c∗(δ )] with E[g1(L)] and Var[c∗(δ )] with Var[g1(L)], where

g1(L) = c∗0 +δ supy∈[0,L]

b(y).

The figure shows a close correlation between the speed and the maximum of theshear on [0,L], though the curves are clearly not identical. The tracking of c∗ bythe running maximum of the shear in this regime is another indication of the diver-gence of c∗ as L → +∞. This is very similar to the divergence of homogenizationof HJ equations with classical Hamiltonians and O-U potentials in Chapter 4. Anexplanation and another perspective will be presented in the next section.

Speed Distribution

For a fixed δ = 1 and δ = 14 (corresponding to small and large amplitudes), wecomputed the distributions of the numerically computed values M(δ ). To com-pute these distributions, we partition the range of values into Q disjoint intervals:[x j,x j+1)

Qj=1. Then we let

pdf(x) =1N

N

∑i=1

χ j(Mi(δ ))(x j+1− x j)

if x ∈ [x j,x j+1), (5.51)

Page 113: Surveys and Tutorials in the Applied Mathematical Sciences

5.1 KPP Fronts in Spatially Random Shear Flows 107

0 5 10 15 20 25 30120

140

160

180

200

220

240

Cylinder width (L)

E[g1]

E[c(δ )]

0 5 10 15 20 25 30700

750

800

850

900

950

1000

1050

Cylinder width (L)

Var[g1]

Var[c(δ )]

Figure 5.4 Sublinear growth of ensemble-averaged speed (dot–square) as domain width increases(left); decay of the speed variance (dot–square) as domain width increases (right); δ = 50. Bothare tracked by mean and variance of the upper bound g1 (dot–plus) in (5.50); from [175].

2 3 4 5 6 7 8 9 10 11 120

0.1

0.2

N=20000

2 3 4 5 6 7 8 9 10 11 120

0.1

0.2

c*(δ )

N=100000

Figure 5.5 Convergence of speed-enhancement distribution with increasing number of randomsamples at δ = 14.0; from [178].

where χ j(x) is the characteristic function of the interval [x j,x j+1). The values M(δ )are the enhancement of the minimal speeds due to the variation of the shear afterthe effect of the mean field has been subtracted off. Since a mean-zero shear alwaysenhances the minimal speeds, we expect M(δ ) > 0 for all δ , for all realizations.Figure 5.5 shows the convergence of the distributions at N = 100,000 samples, andQ = 300.

The computation of KPP front speeds in random shear flows inside three-dimensional cylinders (D = R×Ω , with Ω ⊂ R2 a bounded cross section withvarying shapes) has recently been performed [215]. The principal eigenvalue µ ina general-shaped domain Ω is computed by a two-scale finite element method. It isfound that if the area of Ω is kept invariant, then the larger the domain aspect ratio(narrower cross section), the larger the ensemble-averaged front speeds.

Page 114: Surveys and Tutorials in the Applied Mathematical Sciences

108 5 KPP Fronts in Random Media

5.2 KPP Fronts in Temporally Random Shear Flows

In this section, we consider KPP fronts in time-random shear flows and derive thecorresponding variational formula of c∗ based on [179, 247]. In previous formulas,c∗ is associated with the principal eigenvalue of an elliptic operator in case of spatialheterogeneities or a periodic-parabolic operator in case of media with time-periodicvariations. In time-random media, the principal Lyapunov exponent replaces theprincipal eigenvalue and is almost surely deterministic if there is sufficient mixingin the random media. We shall explore the connection of c∗ with a well-studiedtopic called the parabolic Anderson problem in probability, and the homogenizationof the viscous version of the stochastic Hamilton–Jacobi equations in Chapter4.

5.2.1 Lyapunov Exponent, Large Deviation, and Front Spreading

The KPP equation is

ut =12

∆zu+B ·∇zu+ f (u), (5.52)

where u = u(z, t), z = (x,y) ∈R×Rn−1, n≥ 2; f is KPP, with compactly supportedinitial data bounded between 0 and 1; and B = (b(y, t),0, . . . ,0), b(y, t) is a stationaryGaussian process in t, with a deterministic profile in y. More precisely, the functionb(y, t)= b(y, t, ω) is a mean-zero Gaussian random field over (y, t), periodic in y withperiod L for each fixed t, and stationary in t for each fixed y. The field b is definedover the probability space (Ω ,F ,Q) and has covariance function Γ (y1,y2, t1, t2) =EQ[b(y1, t1)b(y2, t2)]. The following assumptions hold on b(y, t):

A1: (Periodicity in y) Let C0,1P (D) denote the space of Lipschitz continuous func-

tions that are periodic on the period cell D = [0,L]n−1. For each ω ∈ Ω , there isa continuous map J(·, ω) : [0,+∞)→C0,1

P (D) such that b(·, t, ω) = J(t, ω).

A2: (Stationarity in t) For each s∈R+ there is a measure-preserving transformationτs : Ω → Ω such that b(y, ·+s, ω) = b(y, ·,τsω). Hence, Γ depends only on y1,y2and |t− s|.

A3: (Ergodicity) The transformation τs is ergodic: if a set A ∈ F is invariant underthe transformation τs, then either Q(A) = 0 or Q(A) = 1.

A4: The field b has mean zero, is almost surely continuous in (y, t), and has uni-formly bounded variance:

E[b(y, t)] = 0, E[b(y, t)2]≤ σ2 ∀y ∈ D, t ≥ 0. (5.53)

A5: (Decay of temporal correlations) The function Γ (r) = supy1,y2Γ (y1,y2,0,r) is

integrable over [0,∞):

Page 115: Surveys and Tutorials in the Applied Mathematical Sciences

5.2 KPP Fronts in Temporally Random Shear Flows 109∫ ∞

0Γ (r)dr = p1 < ∞ (5.54)

for some finite constant p1 > 0. This constant will appear later in estimates of thefront speed.

A6: There is a finite constant p2 > 0 such that

|Γ (y1,y2,s, t)−Γ (y1,y3,s, t)| ≤ p2|y3− y2|Γ (|s− t|).

For example, a field satisfying assumptions A1–A6 might have the form

b(y, t, ω) =N

∑j=1

b j1(y)b

j2(t, ω),

where the functions b j1(y) are deterministic, Lipschitz continuous, and periodic over

D, and the functions b j2(t, ω) are mean-zero stationary Gaussian fields in t.

Before stating the main results, let us define the family of Markov processesassociated with the linear part of the operator in (5.52). For a fixed ω ∈ Ω and foreach z ∈ Rn, t ≥ 0, let Zz,t(s) = (Xz,t(s),Y z,t(s)) ∈ Rn solve the Ito equation

dZz,t(s) = B(Zz,t(s), t− s)ds+dW (s), s ∈ [0, t], (5.55)

with initial condition Zz,t(0) = z = (x,y) ∈ Rn, where W (s) = (W1(s),W2(s)) ∈ Rn

is the n-dimensional Wiener process with W (0) = 0. Because of the shear structureof B, we therefore have

Xz,t(s) = x+∫ s

0b(y+W2(τ), t− τ)dτ +W1(s), (5.56)

Y z,t(s) = y+W2(s).

Let Pz,t denote the corresponding family of measures on C([0, t];Rn). These stochas-tic trajectories will play the role of characteristics in hyperbolic problems or candi-dates for minimizing paths in the Lax formula of HJ analysis.

First, let us define the analogue of the principal eigenvalue in the KPP analysisof Chapter 2 (see (2.51)–(2.52), (2.76)):

Proposition 5.6. Assume that A1–A6 hold for the process b(y, t). There is a set Ω0 ⊂Ω such that Q(Ω0) = 1 and for any ω ∈ Ω0 and any λ ∈ Rn, the limit

µ(λ ,z) = µ(λ ) = limt→∞

1t

logE[e−λ ·(Zz,t (t)−z)

](5.57)

exists uniformly over z ∈ Rn and locally uniformly over λ ∈ Rn. The limit µ(λ ) isa finite constant for all ω ∈ Ω and is independent of z ∈ Rn. Moreover, µ(λ ) ≥ 0,and µ(λ ) is both convex and superlinear: µ(λ )/|λ | →+∞ as |λ | → ∞.

The convexity and coercivity of µ is reminiscent of basic assumptions in the HJhomogenization of Chapter 4. The Legendre transform of µ(λ ) shares the same

Page 116: Surveys and Tutorials in the Applied Mathematical Sciences

110 5 KPP Fronts in Random Media

property and equalsS(c) = sup

λ∈Rn[c ·λ −µ(λ )]. (5.58)

In the language of classical mechanics, the function µ will play the role of theHamiltonian, and S the role of the Lagrangian.

The front speed is bounded from above and below in terms of S.

Theorem 5.7 (Upper bound on front speed). Let b(y, t, ω) satisfy assumptions A1–A6. Let u(z, t, ω) solve (5.52) with initial condition u(z,0, ω) = u0(z), where u0(z)∈[0,1] has compact support and is independent of ω . Then for any closed set F ⊂c∈Rn | S(c)− f ′(0) > 0,

limt→∞

supc∈F

u(ct, t, ω) = 0

uniformly in c ∈ F, for almost every ω ∈ Ω .

The lower bound is given by the following theorem:

Theorem 5.8 (Lower bound on front speed). Let b(y, t, ω) satisfy assumptionsA1–A6. Let u(z, t, ω) solve (5.52) with initial condition u(z,0, ω) = u0(z), whereu0(z) ∈ [0,1] has compact support and is independent of ω . Then for any compactset K ⊂ c ∈ Rn| S(c)− f ′(0) < 0,

limt→∞

infc∈K

u(ct, t, ω) = 1 (5.59)

uniformly in c ∈ K, for almost every ω ∈ Ω .

Putting the upper and lower bounds together, we see that c∗ in the unit directione ∈ Rn satisfies the equation

S(c∗e) = f ′(0), (5.60)

implying, in terms of the “Hamiltonian” µ (following the remarks after Theorem2.5), the front speed variational formula

c∗(e) = infλ ·e>0

µ(λ )+ f ′(0)λ · e . (5.61)

Although the solution u depends on ω ∈ Ω , since B is a random variable over Ω ,the function S(c) and the speeds c∗(e) are independent of ω . They are almost surelyconstant with respect to Q, a consequence of the ergodicity assumption A3. Ergodic-ity follows from mixing or sufficient decay of correlations, so A5 would imply A3 ifΓ were assumed to decay fast enough in large r. The assumption A2 (stationarity) isneeded here so that c∗ becomes constant as in stochastic homogenization in Chap-ter 4. Assumption A6 is a regularity condition, which also appears in analysis ofChapters 3 and 4. The Lipschitz assumption of b in space is needed to define uniquesolutions (characteristics) to the Ito equation (5.55). The periodicity of b(y, t) in yis to provide compactness in the y dimensions. The results with minor modification

Page 117: Surveys and Tutorials in the Applied Mathematical Sciences

5.2 KPP Fronts in Temporally Random Shear Flows 111

extend to fronts in an infinite cylinder with the zero Neumann boundary conditionon the sides of the cylinder, the setting considered [31].

In the infinite cylinder, the compactness property remains, and the process Zz,t(s)just needs to be reflected when it hits the boundary R× ∂Ω . It is also possible totreat the non-Gaussian process b(y, t) in the same framework as long as the processsatisfies a few estimates in the proof [179].

Now let us relate the µ formula of Proposition 5.6 to a stochastic PDE. First, wenote that Zz,t − z is the deviation of Zz,t(s) from the starting point. Without loss ofgenerality, let n = 2. By stationarity, we may set z = (x,y) = (0,0). The expectationterm is

E[e−λ ·Z0,t

]= E

[e−λ1W1,t−λ1 σ

∫ t0 ξ (W1,s,t−s)ds

]·E

[e−λ2W2,t

],

where we have used the independence of W1 and W2. The second factor equalsexpλ 2

2 t/2 by a direct calculation on the Gaussian random variable W2,t (meanzero, variance t). The first factor is the Feynman–Kac formula and equals u(0, t),where u = u(y, t) is the solution of

ut =12

uyy−λ1 b(y, t)u, y ∈ R1,

u(y,0) = e−λ2y. (5.62)

The solutions of (5.62) are invariant in the sense of distributions if (λ1,λ2) →−(λ1,λ2), x →−x, implying that limt→∞ t−1 logu(0, t) = µ(λ ) is an even functionin λ . The function v = logu(y, t) satisfies the viscous Hamilton–Jacobi equation

vt =12

vyy +12

v2y −λ1b(y, t), (5.63)

with linear initial data v(y,0) = −λ2y. The limit limt→∞ t−1v agrees with the ho-mogenized Hamiltonian. We see that µ is a convex function in λ2, being the ho-mogenization limit of a quadratic Hamiltonian with oscillating potential (for fixedλ1). Evenness and convexity of µ in λ2 imply that µ(λ1,0)≤ µ(λ1,λ2), and so thevariational formula (5.61) reads, for e = (1,0),

c∗ = infλ1>0

µ(λ1,0)+ f ′(0)λ1

. (5.64)

At λ2 = 0, the problem (5.62) reduces to the form of the parabolic Andersonmodel, studied extensively in [48, 50, 62, 101]. The asymptotic growth rate of solu-tions

limt→∞

t−1 logu(0, t) = γ, (5.65)

if it exists and is nonrandom, is called the almost sure Lyapunov exponent. In thestudy of the parabolic Anderson model, b is a random potential in both y and t. Re-sults are mostly in the regime of b being white noise in time or independent of time(stationary). In the stationary case b = b(y,ω), the limit (5.65) is closely related tothe geometry of high peaks of the potential b and the corresponding peaks of so-

Page 118: Surveys and Tutorials in the Applied Mathematical Sciences

112 5 KPP Fronts in Random Media

lutions. At large time, the solution exhibits a spatially extremely irregular structureconsisting of islands of high peaks located far from each other. This is reflected inthe fact that first moment of solution E[u(x,0)] grows much faster than individualrealizations u(0, t) as t →+∞. To leading order,

t−1 logu(0, t)∼max|y|≤t

|λ1|b(y,ω), (5.66)

as shown in [49, 101]. The viscous term enters the next-order asymptotics. Hencethere is divergence of front speed c∗ in KPP or viscous HJ due to unbounded run-ning maxima in time-independent Gaussian shear flows, first observed in [247]. Forspatial Gaussian processes (such as O-U) [50], the leading-order exponential growthrate of u has the asymptotic t-dependence

sup|x|≤t

ξ (x)∼√

2Γ (0) log t, t → ∞, a.s., (5.67)

Γ (0) = E[b2]. So

limt→∞

1t√

log tlogu(0, t) = |λ1|

√2Γ (0), a.s., (5.68)

implying front-speed divergence in time. At large t, we have formally

µ(λ1,0)+ f ′(0)λ1

∼ f ′(0)λ1

+λ1

2+ sign(λ1)

√2Γ (0) log t,

or

c∗ = infλ1>0

µ(λ1,0)+ f ′(0)λ1

∼ c0 +√

2Γ (0) log t. (5.69)

The breakdown phenomenon of stochastic homogenization of inviscid quadraticHJ in Chapter 4 extends to viscous quadratic HJ. The KPP front speed c∗ is relatedto stochastic homogenization of viscous HJ. See [135, 150] for the existence of thehomogenization limit for convex and bounded spatially random Hamiltonians, and[136] for convex and bounded space–time random Hamiltonians.

In the time-dependent (white noise) case, the sign change of the potential pro-motes mixing and does not support growth of large peaks in solutions. The almostsure limit (5.65) is finite in both discrete and continuous models [48, 62]. In ourcase, we handle correlated noise in time for y in a bounded domain. The proof of theexistence of µ uses the subadditive ergodic theorem on the logarithm of the infimumand supremum over z of the expectation term in (5.57), then a Harnack inequalityto show that the two limits from inf and sup are the same. The subadditivity followsfrom the Markov property (independent increment) of the Wiener process.

Proof of the KPP front speed depends on large-deviation estimates of the randomvariable

η tz(κt) =

z−Zz,t(κt)κt

, (5.70)

Page 119: Surveys and Tutorials in the Applied Mathematical Sciences

5.2 KPP Fronts in Temporally Random Shear Flows 113

which is the average velocity of a trajectory over the time interval [0,κt]. The needfor the parameter κ ∈ (0,1] results from treating the time-dependence of the fieldb(y, t). Then the work is to show that almost surely, the random variables η t

z(κt)satisfy a large-deviation principle (LDP) with convex and superlinear rate functionS(c) that is nonrandom and independent of z. The following properties are valid:

(i) For each s≥ 0, the set Φ(s) = c ∈ R| S(c)≤ s is compact.(ii) For any δ ,h > 0, there exists t0 > 0 such that for all t > t0

P(d(η t

z(κt),Φ(s)) > δ)≤ e−κt(s−h).

(iii) For any δ ,h > 0, there exists t0 > 0 such that for all t > t0

P(η t

z(κt) ∈Uδ (c))≥ e−κt(S(c)+h). (5.71)

The decoupling of two components of Zz,t(s) in the shear flow and the integralrepresentation of Ito solutions (5.56) are used to prove the LDP [179].

The Feynman–Kac representation for u to be used with the large-deviation esti-mate is

u(z, t) = E[e∫ t

0 ζ (t−s,u(Zz,t (s),t−s))dsu0(Zz,t(t))], (5.72)

oru(z, t) = E

[e

∫ t∧τ0 ζ (t−s,u(Zz,t (s),t−s))dsu(Zz,t(t ∧ τ), t− (t ∧ τ))

], (5.73)

where τ is a stopping time, the expectation is on the diffusion process Zz,t , andζ (u) = f (u)/u. Roughly speaking, for large t, the expectation is supported on thosepaths Zz,t that travel a distance of O(t) and pass through regions where u is small(ζ (u) is nearly maximized) but u(Zz,t(t− (t ∧ τ)), t− (t ∧ τ)) is not too small. Thelower bound is obtained by carefully choosing stopping times and estimating theprobability of such paths.

The large-deviation estimate applied to the Feynman–Kac solution formula thenimplies that for any compact set K ⊂ c ∈ Rn | S(c)− f ′(0) > 0,

liminft→∞

1t

log infc∈K

u(ct, t)≥−maxc∈K

(S(c)− f ′(0)), (5.74)

holds almost surely. The estimate (5.74) is critical for proving the lower bound that

liminft→∞c∈K′

u(ct, t)≥ 1−h,

for any h ∈ (0,1), where K′ is any compact subset of c ∈ Rn : S(c)− f ′(0) < 0.The upper bound is more straightforward, where we replace ζ (u) by f ′(0) in theFeynman–Kac formula (5.72), so that the representation becomes explicit. Then weextract exponential asymptotics by large-deviation estimation (as shown in Chapter2). The lower bound of u uses the Feynman–Kac formula with nontrivial stoppingtimes (5.73) to allow conditioning on the Ito paths Zz,t(s) to pass through where u issmall in a controllable manner.

Page 120: Surveys and Tutorials in the Applied Mathematical Sciences

114 5 KPP Fronts in Random Media

5.2.2 Speed Bounds and Asymptotics

Variational formula (5.61) allows us to bound c∗. Throughout this section we assumethat the shear b(y, t) has the form

b(y, t) =N

∑j=1

b j1(y)b

j2(t), (5.75)

where b j1(y) are Lipschitz continuous and periodic in y, and b j

2(t) are stationary cen-tered Gaussian fields such that assumptions A1–A6 are satisfied. We consider onlyfront propagation in the direction k = (1,0), which is aligned with the direction ofthe shear. The variational formula for the front speed reduces to the one-dimensionaloptimization problem

c∗(k) = infλ1>0

µ(λ1,0)+ f ′(0)λ1

,

where µ(λ1,0) is determined by the limit

µ(λ1,0) =λ 2

12

+ limt→∞

1t

logφ(y, t) (5.76)

andφt =

12

∆yφ −λ1 b(y, t) φ , φ |t=0 ≡ 1. (5.77)

We consider the scaling b(y, t) 7→ δb(y, t) and the resulting enhancement of thecorresponding speed c∗ = c∗(δ ). The analytical bounds below in particular establishthe linear growth law for large δ , extending corresponding results in periodic media(Chapter 2).

Theorem 5.9 (Bounds on c∗). For all δ ≥ 0, c∗(δ ) satisfies the bounds

(i) c∗(δ )≥ c∗(0).

(ii) c∗(δ ) = c∗(0) if b(y, t) = b(t).

(iii) c∗(δ )≤ c∗(0)+δ ∑Nj=1 ‖b j

1‖∞EQ[|b j2|].

(iv) c∗(δ ) ≤ c∗(0)√

1+δ 2 p1, where, p1 =∫ ∞

0 Γ (r)dr is the integral of temporalcorrelation of the shear in assumption A5.

It follows from statement (iv) that when δ is small,

c∗(δ )≤ c∗(0)(

1+δ 2 p1

2

)+O(δ 3),

the upper bound for the quadratic law.

Page 121: Surveys and Tutorials in the Applied Mathematical Sciences

5.2 KPP Fronts in Temporally Random Shear Flows 115

Theorem 5.10 (Linear growth of c∗). The nonrandom constant C∈ [0,+∞) definedby

liminfδ→∞

c∗(δ )δ

= C (5.78)

is equal to zero if and only if b(y, t)≡ b(t).

Let us prove the inequality (iv) above, which uses the Gaussian property of theshear. Recall that the exponential term e−λ ·(Zz,t (t)−z) in the µ formula (5.57) equals

ξ (t,W )≡ e−∫ t

0 λ1b(W2(s)+z,t−s)ds−λ1W1(t)−λ2·W2(t).

For any fixed continuous path W , ξ (t,W ) is lognormal with mean

EQ[ξ (t,W )] = e|λ |2σ2/2e−λ1W1(t)−λ2·W2(t), (5.79)

where

σ2 =∫ t

0

∫ t

0Γ (W (s),W (r),s,r)dsdr (5.80)

≤∫ t

0

∫ t

0supy1,y2

Γ (y1,y2,s,τ)dsdτ

≤ 2∫ √

2t

0

∫ t/√

2

0Γ (r)dr dv

≤ 2√

2p1t, (5.81)

where we have made the change of variables r = s− t, v = s + t. Under the scalingb 7→ |λ |δb, the constant p1 defined in assumption A5 is replaced by p1 7→ |λ |2δ 2 p1.Taking expectation EP[·] of W in the expression of (5.79) shows that

µ(λ )≤ |λ |2/2+√

2|λ |2δ 2 p1,

and so

c∗(δ ) = infλ1>0

µ(λ1,0)+ f ′(0)λ1

≤ infλ1>0

λ1

2+

f ′(0)λ1

+λ 2

1 δ 2 p1

2

= 2√

(1+δ 2 p1) f ′(0)/2 = c∗(0)√

1+δ 2 p1. (5.82)

The upper bound (iv) immediately yields a proof of speed-bending studied in thecombustion literature [65]. Experiments with premixed flames have shown that in-creasing turbulence intensity does not lead to unlimited linear enhancement of theturbulent burning rate [203]. It has been proposed [65] that this “bending” of the tur-bulent burning velocity in high-intensity flows can be explained by a rapid temporaldecorrelation of the flow. The simulation of [65] used a viscous G-equation. For theKPP model, the following upper bound confirms that rapid temporal decorrelationalso leads to sublinear enhancement of the KPP front speed. The agreement of the

Page 122: Surveys and Tutorials in the Applied Mathematical Sciences

116 5 KPP Fronts in Random Media

G-equation and the KPP model predictions identifies the time decorrelation of theflow field as one mechanism of speed-bending among others [65].

Notice that the derivation uses no information about the spatial structure of theflow other than the maximum value ‖b j

1‖∞. As a result, it is likely that the actualspeed may grow more slowly than δ 1/2, or that c∗ eventually decreases with δ , assuggested by numerical experiments of [65].

Corollary 5.11 (Bending of Front Speed). For δ > 0, let b j2(t)N

j=1 be a family of

stationary Gaussian fields on [0,∞) satisfying EQ[b j2(s)b

k2(t)]≤C1e−α j,k|t−s|, where

α j,k > 0 and C1 > 0. Then for the scaled flow bδ (y, t) = ∑Nj=1 δb j

1(y)bj2(δ t),

limsupδ→∞

c∗(δ )√δ

< +∞. (5.83)

Proof. For the flow ∑Nj=1 b j

1(y)bj2(δ t), we have

Γ (r) = supy1,y2

Γ (y1,y2,0,r)

≤∑j,k‖b j

1‖∞‖bk1‖∞EQ[b j

2(0)bk2(δ r)]

≤∑j,k‖b j

1‖∞‖bk1‖∞C1e−α j,kδ |r|, (5.84)

and so

p1 =∫ ∞

0Γ (r)dr ≤ δ−1C1 ∑

j,k

‖b j1‖∞‖bk

1‖∞

α j,k.

The front-bending result (5.83) now follows from part (iv) of Theorem 5.9. utThe quadratic and linear growth laws can be formally derived for KPP front

speeds in a stationary space–time Gaussian white-in-time shear field b(y, t,ω) overRn based on the asymptotic properties of the Lyapunov exponent [247]. Let the co-variance function of b be Γ (t,y) = δ (t)Γ0(|y|). Asymptotic results for the principalLyapunov exponent of the parabolic Anderson problem can be stated as follows[48, 62]. Let v be the solution of

vt = κ∆v−b(y, t)v, x ∈ Rn, κ > 0. (5.85)

Then the almost sure Lyapunov exponent γ(κ) exists as a nonrandom number andobeys (n = 1,2, c1 > 0)

γ(κ)∼ c1κ p, p ∈ (0,1/2), κ ¿ 1, (5.86)

γ(κ)∼ Γ0(0)2

, κ À 1. (5.87)

Moreover, γ(κ) is a monotone increasing continuous function in κ [48]. The resultscan be adapted to equation (5.77) with b 7→ δb. First 1√

a b(y, t/a) = b(y, t) in law,

Page 123: Surveys and Tutorials in the Applied Mathematical Sciences

5.3 KPP Fronts in Spatially Random Compressible Flows 117

for any a > 0. Let t = t ′/a, a = (δλ1)2. Then equation (5.62) becomes

ut ′ =1

2δ 2λ 21

∆u−b(y, t ′)u. (5.88)

It follows that limt→∞ logu(0, t)/t = γ∗ = γ∗(δλ1) such that

γ∗(δλ1)∼ Γ (0)(δλ1)2

2, δλ1 ¿ 1, (5.89)

γ∗(δλ1)∼ b0h(

12δ 2λ 2

1

)(δλ1)2, δλ1 À 1, (5.90)

where h = h(x) = xp, p ∈ (0,1/2).Now we minimize µ(λ1,0)/λ1. For δ ¿ 1, we have

µ(λ1,0)λ1

∼ f ′(0)λ1

+λ1

2+

Γ (0)δλ1

2+higher-order terms,

giving minimum value

c∗ = c∗(0)(

1+12

Γ (0)δ 2)

+higher-order terms. (5.91)

This is the stochastic analogue of quadratic speed enhancement. It is interesting thatin this case, the integral of (5.80) equals Γ (0)t, and (5.91) follows also from the µformula of Proposition 5.6. The linear law in large δ follows from (5.90) and thespeed variational formula [247].

5.3 KPP Fronts in Spatially Random Compressible Flows

In the previous two sections, the randomness in the flow appears in time or in adirection orthogonal to that of front propagation. The front speeds are enhanceddue to shear structure of the flows. In this section, we study the case in which ran-domness is in the direction of front propagation. We consider solutions to the KPPreaction–advection–diffusion equation

ut =12

uxx +b(x)ux + f (u), t > 0, x ∈ R. (5.92)

The initial datum u0(x) ∈ [0,1] is compactly supported. The advection b(x) underour assumptions below will be compressible.

The pioneering work on KPP front speeds in random media was based on thelarge-deviation method [94, 100] in the late 1970s for the KPP equation

ut = uxx + f (x,u), x ∈ R, (5.93)

Page 124: Surveys and Tutorials in the Applied Mathematical Sciences

118 5 KPP Fronts in Random Media

where f (x,u) is a KPP nonlinearity with random stationary ergodic dependence inx, and the initial data are nonnegative and compactly supported. The variationalformula of front speeds was proved [100]; see also [96, Section 7.7] for an exampleof two-state Markov dependence of f in x and the representation of the variationalformula in special functions. The next result along this line was [144], analyzing ad-dimensional (d ≥ 2) lattice KPP equation of the form

ut = ∆u+ξ (x)u(1−u), t > 0, x ∈ Zd , (5.94)

with initial condition u(0,x) = 1 if x = 0, u(0,x) = 0 otherwise. Here

∆ f (x) =1

2d ∑e∈Zd

|e|=1

f (x+ e)− f (x),

the discrete Laplacian. The random variables ξ (x) are independent and identicallydistributed, bounded, and nonnegative. A similar variational front speed formulaholds.

The results in this section are concerned with front speeds in unbounded randomadvection by a combination of the large-deviation approach and the techniques fromanalysis of random walks in random environments. Moreover, we shall show nearlyoptimal asymptotic bounds on the front speeds for strong advection.

For the random drift b(x, ω) : R× Ω → R, we make the following assumptions:

(1) (stationarity) b is a stationary random process on R defined over the probabilityspace (Ω ,F ,Q) with zero mean, EQ[b] = 0.

(2) (ergodicity and regularity) b(·, ω) is almost surely locally Lipschitz continu-ous and translation with respect to x generates an ergodic transformation of thespace Ω .

(3) (moment condition) the process b(x, ω) satisfies

EQ

[sup

x∈[−2,2]|b(x, ω)|

]< ∞. (5.95)

However, we do not assume that the process b is globally bounded or globallyLipschitz continuous.

(4) For some α1,α2 ∈ R,

limsupz→∞

Q(∫ z

0b(s, ω)ds≥ α1

)< 1 (5.96)

and

limsupz→∞

Q(∫ 0

−zb(s, ω)ds≤ α2

)< 1. (5.97)

Assumption (4) is not restrictive. For example, if b(x, ω) is square integrable andsufficiently mixing with respect to shifts in x, then b satisfies an invariance principle

Page 125: Surveys and Tutorials in the Applied Mathematical Sciences

5.3 KPP Fronts in Spatially Random Compressible Flows 119

[36]1

σ√

z

∫ z

0b(x, ω)ds→ N(0,1), Q-a.s., (5.98)

which implies (5.96). In particular, all the above assumptions on b hold for a mean-zero locally Lipschitz continuous Gaussian process with sufficient decay of corre-lation functions, while (5.95) follows from the Borel inequality [4]. The Borel in-equality for Gaussian fields on Rn states that if ‖ f‖ = supy∈M b(y) is almost surelyfinite, M a compact subset of Rn, then EQ[‖ f‖] < ∞; moreover, for any u > 0,

Q(|‖ f‖−E[‖ f‖] |> u)≤ e−u2/2σ2M , (5.99)

where σ2M = supy∈M EQ[b2(y)].

We now present some of the main results of recent work [181] on front propaga-tion and speeds bounds.

Theorem 5.12. Suppose that assumptions (1)–(4) hold. Then there are deterministicconstants c∗− < 0 and c∗+ > 0 such that for any closed set F ⊂ (−∞,c∗−)∪ (c∗+,+∞),

limt→∞

supc∈F

u(ct, t, ω) = 0

for almost every ω ∈ Ω . Also, for any compact set K ⊂ (c∗−,c∗+),

limt→∞

infc∈K

u(ct, t, ω) = 1

for almost every ω ∈ Ω .

The next result describes the effect of scaling the drift b 7→ δb, where δ ∈ [0,∞)is a scaling parameter. The front speed decreases to zero as the flow amplitude in-creases:

Theorem 5.13. The front speed c∗+(δ ) satisfies the lower bound

c∗+(δ )≥ 1C

min(

1,f ′(0)

1+δM

), (5.100)

where C is a universal constant and M = EQ

[supx∈[−2,2] |b(x, ω)|

]. Moreover, for

any p ∈ (0,1), there is a random constant C′ = C′(p, ω) such that

c∗+(δ )≤C′δ−p (5.101)

for all δ > 0. Therefore limsupδ→∞ c∗+(δ ) = 0 holds with probability one. Similarstatements hold for |c∗−(δ )|.

Our analysis of u(x, t) involves large deviations estimates for the associated dif-fusion process Xx(t) in the random environment. From assumption (5.95) and theassumption of stationarity and ergodicity, it follows that almost surely with respect

Page 126: Surveys and Tutorials in the Applied Mathematical Sciences

120 5 KPP Fronts in Random Media

to Q there is a constant k = k(ω) such that |b(x, ω)| ≤ k(1+ |x|) for all x∈R. There-fore, for each ω ∈ Ω fixed, we can define Xx(t) to be the strong solution to the Itoequation:

Xx(t) = x+∫ t

0b(Xx(s))ds+W (t), (5.102)

where W (t) = W (t,ω) is a one-dimensional Wiener process (Brownian motion)defined on (Ω ,F ,P) with W (0) = 0, P-a.s. The large-deviation method for KPPfronts was developed for one-dimensional spatial random media in the late 1970s.Chapter VII of [96] analyzed equation (5.92) under the assumption that b is uni-formly bounded or that randomness appears in the nonnegative reaction f . Solutionsto (5.92) in multiple dimensions with uniformly bounded random coefficients werealso studied recently [150] as an application of stochastic homogenization of viscousHJ. See also [135] for a different method of HJ homogenization. Our results treatedunbounded advection and provided concrete tight bounds on front speeds in the limitof large advection. Let us return to the HJ perspective of KPP to explain why un-boundedness of b does not cause front speed to diverge. The unbounded advectionis a gradient flow, and the corresponding Hamiltonian is H(p,x) = p2/2 + b(x)p.As we discussed in Example 4.10, the HJ asymptotic front speed is finite in spite ofthe unboundedness of b, just as here.

The bounds (5.100)–(5.101) tell us about the rate of “front trapping” by randommedia in the large-advection limit. A few remarks are in order about front speedsand diffusion in random media. When b≡ 0, the solutions to the initial value prob-lem (5.92) develop fronts propagating with speed equal to c∗ = 2

√κ f ′(0), where κ

is the diffusion constant (κ = 1/2 in (5.92)). This suggests that for nonzero b (sta-tionary and ergodic) one might estimate the front speed by c∗ ≈ cκ = 2

√κ f ′(0),

where κ = limt→∞ E[|Xx(t)|2]/t is the Lagrangian way of writing the homogenized(effective) diffusivity of the random medium. For periodic incompressible two-dimensional velocity fields, the ratio c∗/cκ is bounded away from zero and infinityby constants independent of the flow [208]. In the random setting here, however,even when front speeds and κ are both finite, such equivalence does not hold. Thetwo quantities cκ and c∗ may scale quite differently with respect to δ . When b isuniformly bounded, the process Xx(t) is diffusive [190, 214] with effective diffusiv-ity

κ =1

EQ[e−b]EQ[eb]> 0, (5.103)

almost surely with respect to Q. Suppose that the distribution of b is sign-symmetric(i.e., b L=−b). Then EQ[e−b] = EQ[eb], so that the effective diffusivity is

κ =1

(EQ[eb])2 .

In this case, the effective diffusivity (and cκ ) will decrease exponentially fast asthe scaling parameter δ is increased. However, the lower bound in Theorem 5.13shows that the corresponding front speed can decrease no faster than O(δ−1) as δ

Page 127: Surveys and Tutorials in the Applied Mathematical Sciences

5.3 KPP Fronts in Spatially Random Compressible Flows 121

increases. The reason for this difference is that the front speed is determined by largedeviations of the diffusion process Xx(t), which may not be accurately predicted bythe asymptotic behavior of the variance of the process.

Because Xx(t) does not have an explicit formula in terms of b (compare equation(5.102) with equation (5.56)), large-deviation estimates for Xx(t) are derived for thehitting time T s

r , the first time the process hits the point x = r (from the right) startingfrom x = s ≥ r. Then the hitting time estimates are translated into those for theaverage velocity of trajectory Xx(t). For the right-moving front (left-moving frontsare similar), we have the following result:

Proposition 5.14. Suppose b(x, ω) satisfies assumptions (1)–(4). Almost surely withrespect to Q, the following estimates hold. Let v ∈ R, κ ∈ (0,1]. For any closed setG⊂ [(a0)−1,∞), a0 > 0 and deterministic,

limsupt→∞

1κt

logP(

vt−Xvt(κt)κt

∈ G)≤− inf

c∈GcI+

(1c

), (5.104)

and for any open set F ⊂ [(a0)−1,∞),

liminft→∞

1κt

logP(

vt−Xvt(κt)κt

∈ F)≥− inf

c∈FcI+

(1c

). (5.105)

Here the function I+(a) is deterministic and satisfies

(i) I+(a) > 0 for a ∈ (0,a0), a0 ∈ (0,∞].(ii) I+(a) is convex and decreasing in a for a ∈ (0,a0).

(iii) lima→0+ I+(a) = +∞, and lima→(a0)− I+(a) = 0.(iv) If a0 < ∞, then I+(a0) = 0, and I+(a)≥ 0 for a ∈ (a0,∞).

Define nonrandom constants c∗+ > 0 by the equation

(c∗+)I+(1/c∗+) = f ′(0). (5.106)

Then by the properties of I+(a), c∗+ exists uniquely, 1/c∗+ ∈ (0,a0), and cI+(1/c) >f ′(0) for all c > c∗+. The function cI+(1/c) is the analogue of S(c), the Lagrangian(action) in the analysis of shear flows. The large-deviation upper bound above(5.104) gives, via the Feynman–Kac formula,

limt→∞

1t

supc∈F

logu(ct, t)≤ f ′(0)− cI+(

1c

)< 0,

for c ∈ F ⊂ (c∗+,∞). The large-deviation lower bound above (5.105) leads to thelower bound of solution u:

liminft→∞

1t

log infc∈K

u(ct, t)≥−maxc∈K

(cI+

(1c

)− f ′(0)

), (5.107)

almost surely in Q, for any compact set K ⊂ (c∗+,∞).

Page 128: Surveys and Tutorials in the Applied Mathematical Sciences

122 5 KPP Fronts in Random Media

The propagation speed c∗+ is then justified with these bounds of solutions as be-fore. The speed bounds come from analyzing further the I+ function and the hittingtimes of the processes Xx(t) using properties of b and the Wiener process [181].

A related class of random KPP equations arising from the limit of certain inter-acting particle system are of the type

ut = uxx +a(u)+ εb(u)W , (5.108)

where a(u) is a KPP nonlinearity, typically equal to u(1− u); b(u) is a Lipschitzcontinuous function of u; and W = W (x, t) is a space–time white noise. Randomtraveling fronts [166] are investigated in a special form of equation (5.108):

ut = uxx +u−u2 + ε√

u(1−u)W , (5.109)

with the initial datum u0 = I(−∞,a).Equation (5.109) is closely related to the so-called historical process; see [64]

for details. Simply put, the historical process is a measure on the sets of paths overa time period [0, t0], which represent the past history up to time t0 of a cloud ofinfinitesimally small particles whose density at time t is u(x, t). The particles moveaccording to independent Brownian motions, and they give birth and die. There isan excess of births over deaths, which has a size of 1−u. An interesting property ofthis process is that a small collection of particles dies out quickly for large valuesof x for the given initial density u0, which implies that the solution u has compactsupport on the positive x-axis. By symmetry of the solutions, u is also equal to 1beyond a compact interval. The front location is defined as

b(t) = supx ∈ R : u(x, t) > 0. (5.110)

The main results proved in [166] are given in the following theorem:

Theorem 5.15. Consider a solution u of (5.109) with initial data I(−∞,a), a > 0. Withprobability one, 0≤ u≤ 1 for all (x, t). For ε small enough, the solution u behaveslike a moving front with the following properties:

1. (Front speed and shape) With probability one, limt→∞ b(t)/t = c∗ exists and c∗ ∈(0,+∞), and this limit depends on ε . The law of the front profile v(x, t) = u(b(t)+x, t) tends toward a stationary limit as t → ∞.

2. (Front width) Let I(t) = [a(t),b(t)] be the smallest closed interval such that u = 1for x < a(t), and u = 0 if x > b(t). Then with probability one, I(t) is a compactinterval for all t ≥ 0.

Theorem 5.15 seems to be the first on KPP random fronts providing informationabout front shape and width in addition to the front speed. The almost sure finitefront width property is reminiscent of the viscous Burgers front under white noiseperturbation.

Recently, the asymptotic behavior of c∗ in ε has been obtained [58].

Page 129: Surveys and Tutorials in the Applied Mathematical Sciences

5.4 KPP Fronts in Space–Time Random Incompressible Flows 123

Theorem 5.16. Under the conditions of Theorem 5.15, the front speed c∗ satisfiesthe inequalities

liminfε→+∞

ε2c∗(ε)≥ 2 (5.111)

and

2≥ c∗(ε)≥ 2−Klog log(1/ε)

[logε]2(5.112)

for ε ∈ (0,1/10) and some constant K > 0.

Inequality (5.111) says that c∗(ε) > 0 no matter how large ε is; hence there is notrapping of fronts. This is similar to KPP fronts in drift, or inequality (5.100). Nu-merical simulations [68] suggest that (5.111) is an equality, or limε→+∞ ε2 c∗(ε) = 2.This means that the space–time white noise slows down the front when ε is large,similar to KPP front speeds in large drift (5.101) except that the slowdown in (5.109)is faster (O(ε−2)) as ε increases.

5.4 KPP Fronts in Space–Time Random Incompressible Flows

Reaction–diffusion front propagation in incompressible space–time random flows isa fundamental subject in premixed turbulent combustion [56, 251, 203, 153, 246].The challenging mathematical problem is to establish the propagation velocity of thefront (large-time asymptotic spreading rate) using the governing partial differentialequations. Another mathematical problem is to characterize the propagation veloc-ity, the turbulent flame speed [203], in terms of flow statistics. Due to the notoriousclosure problem in turbulence, the turbulent front speed has been approximated byad hoc and formal procedures in the combustion literature, such as various closuresand renormalization group methods [194, 251, 54].

However, these methods are difficult to justify mathematically. In the KPP model,a rigorous mathematical theory of front spreading and a variational formula of speedc∗ can be achieved [182]. In other words, the large-time KPP front speed in space–time random incompressible flows is exactly solvable. We shall outline the main in-gredients of the mathematical theory. Though turbulent combustion problems maybe posed in terms of different models in the literature, the fundamental mathemat-ical issues are the same. Solving KPP provides the first step toward pursuing othermodels.

5.4.1 Eulerian Method of Front Speeds in Random Flows

The KPP reaction–diffusion–advection equation is

∂tu = ∆u+V (x, t, ω) ·∇u+ f (u) ∆= L u+ f (u), (5.113)

Page 130: Surveys and Tutorials in the Applied Mathematical Sciences

124 5 KPP Fronts in Random Media

with smooth, compactly supported, nonnegative initial data u(x,0, ω) = u0(x), 0 ≤u0 ≤ 1. The vector field V (x, t, ω) is defined over a probability space (Ω ,F , P). Wemake the following assumptions:

(1) (Stationarity and ergodicity) The field V is stationary with respect to shifts inx and t: there is a group of measure-preserving transformations τ(x,t) : Ω → Ωsuch that V (x+h, t + r, ω) = V (x, t,τ(h,r)ω), and τ acts ergodically on Ω .

(2) (Regularity) V is locally Holder continuous, almost surely, in the sense that foreach T > 0 there is α = α(ω,T ) ∈ (0,1) such that

‖V (·, ·, ω)‖Cα (Rd×[0,T ]) < ∞ (5.114)

holds for almost every ω ∈ Ω .(3) (Incompressibility) The field V is divergence-free, ∇ ·V = 0, in the sense of

distributions, almost surely with respect to P.(4) (Moment condition) The field V satisfies the bound

V2∆= EP

[sup

(t,x)∈[0,1]×Rd|V (x, t)|2

]< ∞. (5.115)

Condition (4) means that V (x, t, ω) is uniformly bounded in x for each fixed tand ω . However, we do not require that V (x, t, ·) ∈ L∞(Ω), so that V may becomeunbounded as t → ∞. The Holder regularity condition (2) is satisfied by turbulentflows [153, 246] and is a physical assumption for turbulent combustion problems[203, 194, 246].

For almost every ω , there exists a unique classical solution satisfying (5.52). Ourmain result [182] is the following theorem regarding the almost sure asymptoticbehavior of the solution u(x, t, ω) as t → ∞:

Theorem 5.17. There are a convex open set G⊂Rd and a set of full measure Ω0 ⊂Ω , P(Ω0) = 1, such that the following limits hold for all ω ∈ Ω0:

limt→∞

supc∈F

u(ct, t) = 0 (5.116)

for any closed set F ⊂ Rd \ G and

limt→∞

infc∈K

u(ct, t) = 1 (5.117)

for any compact set K ⊂ G.

Thus, the deterministic set ct ∈ Rd | c ∈ ∂G represents the spreading interfacein an asymptotic sense, made precise by (5.116) and (5.117). The set G may becharacterized in the following way. Let φ(x, t, ω)≥ 0 solve the advection–diffusionequation ∂tφ = L φ with initial condition φ(x,0, ω) = φ0(x) ≥ 0, where φ0(x) issmooth, deterministic, and compactly supported.

Page 131: Surveys and Tutorials in the Applied Mathematical Sciences

5.4 KPP Fronts in Space–Time Random Incompressible Flows 125

Theorem 5.18. The limit

µ(λ ) = limt→∞

1t

log∫

Rdeλ ·xφ(x, t, ω)dx = lim

t→∞

1t

log EP[eλ ·xφ(x, t, ω)] (5.118)

exists almost surely with respect to P. Moreover, µ(λ ) is a finite convex function ofλ ∈ Rd, and is superlinear at large λ .

The function µ is the “Hamiltonian,” and now the characterization of G is givenby the “Lagrangian”:

Theorem 5.19. The set G described in Theorem 5.17 is given by

G = c ∈ Rd | S(c)≤ f ′(0), (5.119)

where S(c) = supλ∈Rd (λ · c− µ(λ )) and µ(λ ) is defined as in Theorem 5.18. Itfollows that the asymptotic front speed c∗ in the direction e ∈ Rd is given by thevariational formula

c∗(e) = infλ ·e>0

µ(λ )+ f ′(0)λ · e . (5.120)

For the KPP model, Theorems 5.17 and 5.19 address two open problems in tur-bulent combustion [203]: the existence of a well-defined turbulent front speed andthe precise analytical characterization of the speed. In Theorem 5.18, one may nor-malize φ so that φ is the density for a probability measure on Rd , for each fixed ω ,and the theorem characterizes the asymptotic behavior of the tails of the distribution(large deviations from the mean behavior) almost surely with respect to the measureP on the velocity field. The function S in Theorem 5.19 is the rate function that gov-erns these large deviations. Because S is related to the action functional, it may beviewed as a Lagrangian.

The quantity µ(λ ) has another interpretation in terms of exponential growth ofPDE solutions, or the almost sure (principal) Lyapunov exponent. It can be formu-lated as either an initial value problem of the related PDE as we have done in Section5.2 or a terminal value problem below. Consider the function ϕ∗(x,τ; t, ω), whichsolves the terminal value problem (τ ∈ (0, t)):

∂τ ϕ∗+∆ϕ∗ − (V (x,τ)−2λ ) ·∇ϕ∗+(|λ |2−λ ·V (x,τ)

)ϕ∗ = 0, (5.121)

with linear terminal data ϕ∗(x, t; t, ω) ≡ 1, x ∈ Rd . Then ϕ∗(x,0; t, ω) grows expo-nentially in t with a rate equal to µ(λ ):

Theorem 5.20. If ϕ∗(x,τ; t, ω) solves (5.121) with terminal data ϕ∗(x, t, ω) ≡ 1,then for any r > 0,

limt→∞

sup|x|≤rt

∣∣∣∣1t

logϕ∗(x,0; t, ω)−µ(λ )∣∣∣∣ = 0 (5.122)

holds almost surely with respect to the measure P.

Page 132: Surveys and Tutorials in the Applied Mathematical Sciences

126 5 KPP Fronts in Random Media

The function µ(λ ) is related to the effective Hamiltonian that arises fromstochastic homogenization of “viscous” Hamilton–Jacobi equations [135, 136, 150].The convex function µ(λ ) in (5.118) is equal to an effective Hamiltonian H(λ ).To see this, define the function η∗(x,τ; t, ω) = eλ ·yϕ∗(x,τ; t, ω), which satisfies∂τ η + L ∗η∗ = 0 for τ < t and terminal data η∗(x, t; t, ω) = eλ ·y. Here L ∗η∗ =∆xη∗ −∇ · (V η∗) denotes the adjoint operator. For ε > 0 and T > 0, define

ζ ε(x,τ;T, ω) = ε logη∗(ε−1x,ε−1τ;ε−1T, ω).

Then ζ ε solves the Hamilton–Jacobi equation

∂τ ζ ε + ε∆ζ ε + |∇ζ ε |2−V( x

ε,

τε, ω

)·∇ζ ε = 0, τ ∈ [0,T ), (5.123)

with terminal data ζ ε(x,T ;T, ω) = λ · x. For a velocity field V (x,τ, ω) that is uni-formly bounded in τ (i.e., V ∈ L∞(Ω ;L∞(Rd))), the result of [136] implies that asε → 0, the function ζ ε converges locally uniformly to a function ζ 0(x,τ; t) thatsolves an effective Hamilton–Jacobi equation ∂τ ζ 0(z,τ; t)+ H(∇ζ 0) = 0 with thesame terminal data. The effective Hamiltonian H(λ ) is a deterministic function. Inparticular, by choosing T = 1, we see that

H(λ ) = limε→0

ζ ε(0,0;1, ω) = limε→0

ε logη∗(0,0;ε−1, ω) = µ(λ )

holds almost surely with respect to P.Theorem 5.20 extends this connection to the case of velocity fields V (x, t) that

are not uniformly bounded in t, a case not covered by [135, 136, 150]. As discussedin Section 5.2, the time randomness helps mixing and homogenization of HJ, so onewould anticipate that our results also hold if spatial boundedness of V is removed.We shall show a formal calculation of c∗ for a space–time Gaussian and white-in-time velocity field V later.

Because of the rather general form of V in several space dimensions, neitherthe explicit representation nor a useful hitting time formulation is available for an-alyzing the Ito paths of the associated diffusion process (the stochastic character-istics). This makes proving the large-deviation principle (LDP) based on Ito so-lutions difficult, and is perhaps the reason why progress has been slow since theone-dimensional KKP front result appeared in the 1970s [94, 100, 96]. A methodof proving LDP based on analyzing Ito solutions is Lagrangian. A way to handle amore general form of advection V is the new Eulerian approach that we developedin [182].

The first step is to use the Krylov–Safonov–Harnack inequality [137] to establishcontinuity estimates of the solution. Though the constants appearing in the Krylov–Safonov–Harnack inequality may be arbitrarily bad, they are well behaved “on av-erage.” We use this observation and the subadditive ergodic theorem to establishalmost sure behavior of the tails of the linearized KPP equation.

The tails of these solutions in the large-time limit contain the information on the“Lagrangian” function S (the large-deviation rate function). To apply this property

Page 133: Surveys and Tutorials in the Applied Mathematical Sciences

5.4 KPP Fronts in Space–Time Random Incompressible Flows 127

to the solution of the nonlinear equation and recover front speed c∗, we constructsubsolutions (supersolutions) and use the comparison principle to bound solutionsinstead of estimating with Ito solutions and the Feynman–Kac formula.

Since the proof relies only on the Krylov–Safonov–Harnack inequality and thecomparison principle, it applies readily to a large class of operators L . In fact, allof the proofs may be modified slightly to treat the case that the diffusion is also arandom process. For example, a variant of Theorems 5.17–5.20 holds in the casethat u is governed by an equation of the form

∂tu = ∇ · (A(x, t, ω)∇u)+V (x, t, ω) ·∇u+ f (u), (5.124)

where A(x, t, ω) = Ai j(x, t, ω) is a random positive definite matrix function and uni-formly C1,α .

Let us walk through the main estimates. We begin with the Krylov–Safonov–Harnack inequality [137]. Let θ > 1 and R≤ 2 be two constants, and 0≤ ξ (x, t)≤1. Suppose η is an integrable nonnegative distribution solution of ∂tη −L η +ξ (x, t)η = 0 in Q(θ ,R). Suppose ‖V‖L∞(Q(θ ,R)) ≤ 1. Then there exists a constantKo > 0 depending only on θ and the dimension such that

inf|x|≤R/2

η(x,θR2)≥ Koη(0,R2).

The inequality allows one to deduce the continuity (the relation between the min-imum and maximum of a function at different points over a parabolically scaleddomain Q). We apply it to the logarithm of the KPP solution logu(x, t). The maxi-mum principle ensures that u ∈ (0,1) for all (x, t).

Define ξ (x, t, ω) = f (u(x, t, ω))/u(x, t, ω). The KPP equation (5.52) is writtenas

∂tu = ∆u+V (x, t, ω) ·∇u+ξ (x, t, ω)u, (5.125)

where ξ (x, t, ω) ∈ [0, f ′(0)] is ready for the Harnack analysis.The result is that if γ(t)≥ 0 is any nondecreasing function satisfying

limsupt→∞

γ(t)t≤ ε,

then for any c ∈ Rd ,

liminft→∞

1t

(log inf

|z|≤γ(t)u(ct + z, t)− log sup

y∈Bδ (c(t−γ(t))u(y, t− γ(t))

)

≥−C(1+ |c|+δ )2ε(1+V2) (5.126)

and

Page 134: Surveys and Tutorials in the Applied Mathematical Sciences

128 5 KPP Fronts in Random Media

limsupt→∞

1t

(log sup

|z|≤γ(t)u(ct + z, t)− log inf

y∈Bδ (c(t+γ(t))u(y, t + γ(t))

)

≤C(1+ |c|+δ )2ε(1+V2) (5.127)

with probability one. Here, V2 is defined in (5.115) and C = C(θ) is a constant.Inequalities (5.126)–(5.127) lead to the tail estimates of the linearized solution. Forδ > 0, x ∈Rd , and t ≥ s≥ 0, let φ(y, t;x,s) = φ(y, t;x,s, ω) satisfy the linear part ofthe KPP equation

∂tφ = ∆yφ +V ·∇φ (5.128)

for t > s with the initial condition

φ(y,s;x,s, ω) =

1 if y ∈ Bδ (x),0 otherwise,

at time t = s, where δ > 0 is a fixed parameter. Define the family of functions

φ−(y, t;x,s) = infy′∈Bδ (y)

φ(y′, t;x,s). (5.129)

It follows from the maximum principle that for any x,y,z ∈ Rd and r < s < t, wehave

φ−(z, t;x,r)≥ φ−(y,s;x,r)φ−(z, t;y,s). (5.130)

For c ∈ Rd fixed, define the random process qm,n(ω) = logφ−(cm,m;cn,n, ω) in-dexed by integers m,n, 0 ≤ m < n. We observe that qm,n is stationary and superad-ditive:

qm,n ≥ qm,k +qk,n, ∀m < k < n,

qm+r,n+r(ω) = qm,n(τ(cr,r)ω). (5.131)

One can show that E[|q0,n|] < ∞ for all n. Then the subadditive ergodic theoremapplies to −qm,n, and we have that

−S(c) ∆= limn→∞

1n

q0,n = supn>0

1n

q0,n ≤ 0 (5.132)

exists almost surely and is nonrandom. The convexity of S follows from the subaddi-tivity as shown in Chapter 4. By continuity estimates (5.126)–(5.127), the infimumin (5.129) may be replaced by the supremum for an error under control. More pre-cisely, if γ(t) ≥ 0 is any nondecreasing function satisfying limsupt→∞ γ(t)/t ≤ ε ,then for any c ∈ Qd (d- dimensional rational vectors),

limsupt→∞

1t

log sup|z|≤γ(t)

φ(ct + z, t;0,0)≤C(1+ |c|+δ )2ε(1+V2)−S(c) (5.133)

and

Page 135: Surveys and Tutorials in the Applied Mathematical Sciences

5.4 KPP Fronts in Space–Time Random Incompressible Flows 129

liminft→∞

1t

log inf|z|≤γ(t)

φ(ct + z, t;0,0)≥−C(1+ |c|+δ )2ε(1+V2)−S(c) (5.134)

hold with probability one. The large-deviation principle of φ follows, that for anyopen set G⊂ Rd ,

liminft→∞

1t

log infz∈tG

φ(z, t;0,0, ω)≥− infc∈Go

S(c), (5.135)

and for any closed set F ⊂ Rd ,

limsupt→∞

1t

log supz∈tF

φ(z, t;0,0, ω)≤− infc∈F

S(c), (5.136)

with probability one. The upper bound of the solution follows from the large-deviation upper bound (5.136) and the comparison inequality

u(y, t)≤ et f ′(0)φ(y, t;0,0),

by containing the support of initial data in a ball Bδ (0). So

limt→∞

supc∈F

u(ct, t) = 0,

for any compact set of c such that S(c) > f ′(0). As before, to show convergence ofthe KPP solution to one, we prove that the lower bound

liminft→∞

1t

log infc∈K

u(ct, t)≥−maxc∈K

(S(c)− f ′(0)) (5.137)

holds with probability one, for any compact set K ⊂ c ∈ Rd | S(c)− f ′(0) > 0.This is done by using φ(y, t;0,0) and its localizations as subsolutions, then extract-ing logarithmic asymptotics based on the large-deviation lower bound (5.135). Thelower bound (5.137) and further construction of comparison functions of KPP solu-tions leads to the convergence of u(ct, t) to one if S(c) < f ′(0).

5.4.2 Speed Bounds and Asymptotics

By working with PDE characterization (5.121) of the Lyapunov exponent µ andthe variational formula (5.120) of c∗, we obtain the lower and upper bounds of c∗in terms of statistics of V and the front speed c0 in the absence of advection. ThePDE formulation (Eulerian method) is convenient for handling the divergence-freecondition of V .

Proposition 5.21. Suppose V is divergence-free and of mean zero: E[V ( j)] = 0 forj = 1, . . . ,d. The front speed c∗ satisfies the upper bound

Page 136: Surveys and Tutorials in the Applied Mathematical Sciences

130 5 KPP Fronts in Random Media

c∗(e)≤ c0 +EP[‖V‖L∞x ],

implying at most linear growth in δ À 1 if V is scaled according to V 7→ δV .If V (x, t) is uniformly bounded, then c∗ also satisfies the lower bound

c∗(e)≥ c0.

The above proposition extends similar bounds in the time-random shear flowsas well as bounds for deterministic periodic flows. Numerical computation of c∗ inrandomly perturbed cellular flows [180] suggests that c∗ ∼ O(δ p) at large δ mayoccur for any exponent p ∈ (0,1), when V is scaled according to V 7→ δV . So theabove bounds are optimal in time-random incompressible flows.

The other type of bound on c∗ for δV with Gaussian statistics in time is (iv) ofTheorem 5.9, namely c∗ ≤ c0

√1+δ 2 p1, where p1 is the integral of the correlation

function. We next give an extension of such bounds for the nonshear space–time-random flows.

The key assumption is that V is white in time. The following calculation is formalbut illustrative. It uses the Lagrangian method (Feynman–Kac formula) to involvethe second-order statistics of V , a change-of-measure Girsanov formula, and theproperties of Wiener processes. A velocity field that is white noise in time could beincorporated rigorously through a term of the form V ·∇udW in the original equa-tion (5.52), where denotes the Stratonovich integral [125]. Although this scenariodoes not fall within our assumptions on V , the following derivation demonstratesthe difficulty in estimating c∗ when the velocity V is correlated in time.

Proposition 5.22. Suppose that V has the form

V (x, t, ω) = ∑k

Xk(x)Fk(t, ω), (5.138)

where Xk(x) are periodic or almost periodic divergence-free fields and Fk arewhite-noise processes in time, so that the covariance matrix function is

Γi j = Γi j(x1,x2, t1− t2) = EP[V (i)(x1, t1)V ( j)(x2, t2)]≤ p1δ0(t1− t2)Ai j(x1,x2),

where δ0 is the standard delta function centered at zero, and p1 is a constant. As-sume that the KPP front speeds are given by (5.120). Then c∗ ≤ c0

√1+C2 p1, where

C2 depends only on the dimension d and f ′(0).

Proof. The Feynman–Kac formula for ϕ∗ of equation (5.121) gives

ϕ∗(x,0) = E[e−λ ·∫ t

0 V (Zλ ,s)ds]

e|λ |2 t ,

where Zλ is the diffusion process obeying the Ito equation

dZλ (s) = (V (Zλ ,s)−2λ )ds+√

2dW (s), s ∈ [0, t],

Page 137: Surveys and Tutorials in the Applied Mathematical Sciences

5.4 KPP Fronts in Space–Time Random Incompressible Flows 131

Zλ (0) = z, W (s) = W i(s)di=1 a d-dimensional Wiener process. Changing measure

by the Girsanov theorem [125, Theorem 5.1] yields the following representationof ϕ∗:

E[

exp−λ

√2 ·W (t)+

√2

d

∑i=1

∫ t

0V (i)(Wz(r),r)dW (i)(r)

− 12

∫ t

0‖V (Wz(s),s)‖2 ds

], (5.139)

where Wz(s) = z+W (s) and E is expectation with respect to W . It follows that

ϕ∗ ≤ E[

exp−λ

√2 ·W (t)+

√2

d

∑i=1

∫ t

0V (i)(Wz(r),r)dW (i)(r)

]

and

EPϕ∗ ≤ E[e−λ

√2·W (t)EP

[exp

√2

d

∑i=1

∫ t

0V (i)(Wz(r),r)dW (i)(r)

]]. (5.140)

Notice that inside the inner expectation (with Wz(r) fixed), the sum of stochasticintegrals is a linear combination of Gaussian variables. In other words, the innerexpectation is over a log-normal variable, and so

EPϕ∗ ≤ E[

exp−λ

√2 ·W (t)

+∫ t

0

∫ t

0∑i j

Γi j(W (s),W (τ),s,τ)dW (i)(s)dW ( j)(τ)]

. (5.141)

Since V is white in time, e.g., Γi j = Ai j(x1,x2)p1δ0(t1−t2), the integral in (5.141)is bounded from above by p1C1

∫ t0 ‖dW (s)‖2. The right-hand-side expectation of

(5.141) is bounded from above by

E[

exp∫ t

0p1C1‖dW (s)‖2−

√2λ ·dW (s)

]

=N

∏j=1

d

∏l=1

E[

exp

p1 C1

(dW (l)(s)

)2−√

2λ (l)dW (l)]

, (5.142)

where dW (l) is the Wiener increment over an interval of length t/N. We have usedindependence of Wiener increments in each component and among components.The last expression of (5.142) can be calculated explicitly, and equals, on taking thelimit N → ∞,

exp|λ |2t + p1dC1t.It follows that

Page 138: Surveys and Tutorials in the Applied Mathematical Sciences

132 5 KPP Fronts in Random Media

µ = limt→∞

1t

EP logϕ∗ ≤ limt→∞

1t

logEPϕ∗ ≤ |λ |2 +C1d p1, (5.143)

orc∗ ≤ 2

√f ′(0)+C1d p1 = c0

√1+C2 p1. (5.144)

This completes the proof. utInequality (5.144) implies that rapid temporal decorrelation can reduce speed

enhancement, the so-called speed-bending phenomenon. We remark that the physi-cal mechanisms contributing to the speed-bending in a combustion process may bemuch more complicated and depend on whether the process is in gaseous or liquidphase and on activation energy, among others. Here we have identified time decor-relation as one mechanism for KPP fronts.

If V is Gaussian but nonwhite in time, then p1δ0 in the upper bound of the covari-ance matrix function is replaced by a nonnegative L1 function with integral equal top1. The estimate of the right-hand-side expectation of (5.141) will be more compli-cated. It will be interesting to obtain a similar result as in the random-in-time shearflow case.

5.5 Stochastic Homogenization of Viscous HJ Equations

The study of KPP front speeds in Section 5.4 led to the homogenization of viscousHJs with quadratic Hamiltonian (5.123). A more general homogenization problemis to analyze the ε ↓ 0+ limit of the solutions of

uεt = εσ2∆ uε +H(x/ε, t/ε,ω,∇uε), (t,x) ∈ (0,∞)×Rd , (5.145)

where σ > 0 is a constant (viscosity), uε(x,0) = g(x), with g(x) a uniformly Lip-schitz continuous function. The Hamiltonian H = H(x, t,ω, p) is a stationary er-godic process in (x, t) and is convex in p. Convergence of (5.145) is proved understructural assumptions of H in [135, 136]. In [150], the spatially random case istreated, allowing a degenerate viscous term and using a different method. Both re-sults will be briefly discussed below.

To illustrate ideas, let us consider the spatially random case H = H(x,ω, p). Themain assumptions on H are quite similar to those in Section 4.1 on inviscid HJs:

(H0) Convexity, stationarity, and ergodicity (see (A1)–(A2) in Section 4.1); the func-tion H(x,ω, p) is equal to H(p,τxω), where τx is translation by x;

(H1) Coercivity: there exist constants 1 < α ≤ β , c1,c2 > 0 such that

c1(|p|α −1)≤ H(p,ω)≤ c2(|p|β +1)

holds for all (p,ω). The Legendre transform is well defined and gives a similarLagrangian L(q,ω) = supp[p ·q−H(p,ω)].

Page 139: Surveys and Tutorials in the Applied Mathematical Sciences

5.5 Stochastic Homogenization of Viscous HJ Equations 133

(H2) Uniform continuity of H(p,τxω) in x, which implies that of L(q,τxω).

It follows from convexity that uε has a variational representation. Let ε = 1, andlet C be the set of all bounded maps c : [0,∞)×Rd → Rd . Consider the diffusionprocess (Ito equation)

x(t) = x+∫ t

0c(s,x(s))ds+

√2σW (t), (5.146)

where W is the standard Wiener process. Denote by Qcx the corresponding probabil-

ity measure on the continuous sample path space of x(t). Then

u(x, t,ω) = supc∈C

EQcx

(g(x(t))−

∫ t

0L(x(s),ω,c(s,x(s)))ds

). (5.147)

Rescaling time t → t/ε , the scaled solution is (x = x/ε)

uε(x, t,ω) = supc∈C

EcQx

(g(εx(t/ε))− ε

∫ t

0L(x(s),ω,c(s,x(s)))ds

). (5.148)

The next step is to pass to ε ↓ 0 in the variational formula (5.148) and recoverthe Hopf–Lax formula in the limit with the help of stationarity, ergodicity, and con-tinuity. By stationarity, it suffices to consider the limit of solutions at x = 0. Becauseσ > 0, the diffusion process (5.146) has an invariant measure for a special class ofc and permits averaging by ergodic theory. When restricting the control c and per-forming averaging, one gets a lower bound of liminfuε(0, t). Such a class of controlfunctions is in the form c = b(τxω), for some bounded function b. For given b, theprobability associated with the diffusion process (5.146) has a density Φ , which isa solution in the distribution sense of the equation

∇ · (bΦ) = σ2∆Φ . (5.149)

The averaging with respect to b and x(t) occurs over large time, and it follows from(5.146) that almost surely in ω ,

limε↓0

εx(t/ε) = limε↓0

ε∫ t/ε

0b(τx(s)ω)ds = tE[b(ω)Φ(ω)]≡ tm(b,Φ) (5.150)

and

limε↓0

ε∫ t/ε

0L(b(τx(s)ω),τx(s)ω)ds = tE[L(b(ω),ω)Φ(ω)]≡ th(b,Φ), (5.151)

where expectation is over ω ∈Ω .Then the variational formula (5.148) gives the lower bound

liminf uε(0, t)≥ sup(b,Φ)∈Te

(g(tm(b,Φ))− th(b,Φ)) = supy∈Rd

(g(y)− tL(y/t)), (5.152)

Page 140: Surveys and Tutorials in the Applied Mathematical Sciences

134 5 KPP Fronts in Random Media

where Te denotes the space of pairs (b,Φ) satisfying the equation (5.149) and

L = L(q) = inf(b,Φ)∈TeE[bΦ ]=q

h(b,Φ) = inf(b,Φ)∈TeE[bΦ ]=q

E[L(b(ω),ω)Φ(ω)]. (5.153)

We leave as an exercise to show that L is convex in q. The lower bound in (5.152)is the Hopf–Lax formula of the homogenized HJ with Hamiltonian H, which is theLegendre transform of L:

H = supq∈Rd

(p ·q− L) = sup(b,Φ)∈Te

E[p ·b(ω)−L(b(ω),ω)Φ(ω)]. (5.154)

The formula (5.154) is further used to find an upper bound of uε . It is first put intoa dual variational form in which the constraint (5.149) is replaced by an additionallayer of minimization

H = supφ

supb

infψ

E [[p ·b(ω)−L(b(ω),ω)+Abψ(ω)]φ(ω)] , (5.155)

where Ab is the adjoint operator of the one in the constraint (5.149), Ab = σ 2∆ +b(ω) ·∇. Formula (5.155) is easy to see by noting that for any φ ,

infψ

E[φAbψ(ω)] =−∞,

unless (b,ψ) ∈ Te, which is preferred by supb, since the corresponding value ofinfψ is zero. The triple-layer min–max optimization in (5.155) is convenient forexpressing H in terms of H:

H = supφ

infψ

supb

E [[p ·b(ω)−L(b(ω),ω)+Abψ(ω)]φ(ω)]

= supφ

infψ

supb

E[[(p+∇ψ) ·b(ω)−L(b(ω),ω)+σ2∆ψ ]φ(ω)

]

= supφ

infψ

E[[H(p+∇ψ(ω),ω)+σ2∆ψ]φ(ω)

]

= infψ

supφ

E[[H(p+∇ψ(ω),ω)+σ2∆ψ]φ(ω)

]

= infψ

esssupω

[H(p+∇ψ(ω),ω)+σ2∆ψ ], (5.156)

where the last equality follows from 0 ≤ φ ∈ L1(Ω). Hence for any small numberδ > 0, there is a function ψδ such that

H(p+∇ψδ (ω),ω)+σ2∆ψδ ≤ H +δ . (5.157)

An upper solution is

uε(x, t) = p · x+ t(H +δ )+ εψδ (x/ε,ω), (5.158)

Page 141: Surveys and Tutorials in the Applied Mathematical Sciences

5.5 Stochastic Homogenization of Viscous HJ Equations 135

for the essentially linear initial condition uε(x,0) = p · x+ εψδ (x/ε,ω). Additionalwork [135] shows that εψδ (x/ε,ω)→ 0 for bounded x. It follows from (5.158) that

limsupε→0

uε(x, t)≤ p · x+ tH,

removing arbitrarily small δ > 0. An extension is possible for more general initialdata under certain conditions of the Hamiltonian H [135, Sections 6 and 7]). Moreprecisely, a main convergence result is given by the following theorem [135]:

Theorem 5.23. Let u(x, t) = supy(g(y)− tL((y− x)/t)). Assume that (H0)–(H2)hold and that (H3) there exists a function ν(δ ), ν → 0 as δ → 0, such that for|x| ≤ δ , ∀ω ∈Ω and a positive constant C:

H(p,x,ω)≥ (1+ν(δ ))H((1+ν(δ ))−1 p,ω)−Cν(δ ).

Then with probability 1,

limε→0

|uε(x, t,ω)−u(x, t)|= 0,

uniformly in (x, t) ∈ Rd × [0,∞).

The method above based on the ergodic theorem and min–max formulas devel-oped in [135] extends to space–time random media [136] under slightly modifiedconditions (H0)–(H3), where spatial translation τx is replaced by space–time trans-lation τx,t in (H2), and boundedness of x is replaced by that in (x, t). The space Te ismodified to

Te = (b,Φ) : Φt +∇ · (bΦ) = σ2∆Φ,where derivatives are generated by the translation τx,t on Ω and Φ is a probabilitydensity. The homogenized Lagrangian L is given by the same formula (5.153), andthe homogenized Hamiltonian H by (5.154). Under similar conditions of H, the ho-mogenization of inviscid (σ = 0) HJ in space–time ergodic random media is provedin [213] by combining a subadditive ergodic theorem (Section 4.2) and continuityestimates.

Another approach to homogenization of spatially random viscous HJs is basedon a subadditive ergodic theorem and uniform gradient estimates [150]. The subad-ditive ergodic theorem applies to H that grows faster than quadratically in p. WhenH has slower growth but is still coercive in the sense that H → ∞ uniformly as|p| →∞, then H is perturbed (penalized) to H +η |p|m, for m > 2 and a small η > 0.The solutions of the perturbed HJ equation are homogenized first, then one passes tothe limit η → 0. It is shown in [150] that such a double limit of penalized solutionsagrees with the homogenization limit of the unperturbed HJ solutions (η = 0). Theapproach covers a wider class of HJ equations with degenerate viscosity of the form

uεt − ε tr(A(x,x/ε,ω)∇2uε)+H(∇uε ,uε ,x,x/ε,ω) = 0, (5.159)

Page 142: Surveys and Tutorials in the Applied Mathematical Sciences

136 5 KPP Fronts in Random Media

with bounded uniformly continuous initial datum u0. Here, A and H are stationaryand ergodic in y and ω , and A is a nonnegative matrix such that for any vectorξ ∈ Rd , there is a positive constant Λ such that

0≤ (Aξ ,ξ )≤Λ |ξ |2,

where (·, ·) denotes inner product in Rd . The key assumptions on H are as follows:

(A1) H = H(p,r,x,y,ω) is convex in p;

(A2) H is coercive in that:H(p,r,x,y,ω)→+∞,

uniformly in r,x,y,ω as |p| →+∞.

The main result is the following [150]:

Theorem 5.24. Under (A1)–(A2), stationarity and ergodicity of A and H, and non-negativity and boundedness of A, the solution uε of the HJ equation (5.159) con-verges to u in C(Rd × [0,T ]) almost surely in ω . Here u satisfies the homogenized(deterministic) equation

ut + H(∇u, u,x) = 0, (x, t) ∈ Rd × (0,T ],

with the same initial datum u0. The homogenized Hamiltonian H is convex.

We refer to [150] for details of the proof, including various approximations anderror estimates.

Though the above homogenization results treat general convex nonlinearity ordegenerate viscosity (diffusion), they do not allow unbounded random media, asseen from the coercivity condition (H1) or (A2). The methods in Sections 5.2 and 5.3allow unbounded random media, yet for quadratic Hamiltonians. Homogenizationof general convex viscous HJs with unbounded space–time randomness (withoutcoercivity) remains an interesting issue for further research.

5.6 Generalized Fronts, Reactive Systems, and GeometricModels

In this section we discuss random fronts in non-KPP equations, systems of equa-tions, and geometric models. The aim is to introduce a number of unsolved problemsfor future research based on recent progress.

Page 143: Surveys and Tutorials in the Applied Mathematical Sciences

5.6 Generalized Fronts, Reactive Systems, and Geometric Models 137

5.6.1 Generalized Front Speeds and Central Limit Theorem

When reactive nonlinearity is non-KPP, one cannot find front speeds by analyzingsolutions near zero. Instead, an estimate and control of the entire transition fromzero to one must be available. This then relies on the dynamical properties of solu-tions. The KPP fronts are called pulled fronts (pulled by the unstable state u = 0),while the non-KPP fronts are pushed fronts (both regions where u ≈ 0 and u ≈ 1contribute); see [210] for an overview of related physical literature. The notion ofgeneralized transition front (GTF) in heterogeneous media has been recently pro-posed and studied for non-KPP reactions [159, 218, 21, 23, 162, 163, 174, 172],extending the known constant-speed traveling fronts. We shall give a definition ofGTF below, in the context of bistable and ignition fronts (type 3 and type 5) inspatial random media.

Consider the scalar RD equation

ut = uxx + f (x,u), (5.160)

where the spatial variation of f is arbitrary (aperiodic, nonergodic), and the initialcondition u(x,0) is a profile connecting one (at −∞) to zero (at +∞). A GTF is aglobal solution u(x, t) for all t ∈ R such that 0 < u < 1, and there is a continuousfunction (an interface) X(t) such that for any ε > 0, there is a finite distance Nεindependent of t such that for all t ∈ R,

u(x, t) > 1− ε, ∀x < X(t)−Nε ; u(x, t) < ε, ∀x > X(t)+Nε . (5.161)

This definition is adapted from the more general one in [21] and is reminiscent ofthe front-probing asymptotics in Chapters 2 and 3. It implies a finite width and speedof a frontlike global solution. Alternative definitions [159, 218] require a global-in-time solution and a certain continuity and invariance of its shape.

A solution u = U(x, t,ω) is called a random traveling wave [218] if it is a globalsolution such that 0 < U < 1, limx→+∞ U(x,0,ω) = 0, limx→−∞ U(x,0,ω) = 1, andthere exists a function X(t,ω) such that

U(x, t,ω) = U(x−X(t,ω),0,τX(t,ω)ω), (5.162)

where τ is the spatial translation. The function U(x,0,ω) generates the randomtraveling wave, and plays the role of traveling-front profile in the case of periodicmedia in Chapter 2, where X(t,ω) also reduces to ct.

Now let us turn to equation (5.160) with spatially random reaction. Suppose thatf (x,u) = g(x,ω) f0(u), where g(x,ω) is a stationary ergodic process, almost surelyuniformly Lipschitz in x, and is bounded by two deterministic constants

0 < gmin ≤ g(x,ω)≤ gmax.

For bistable f , one also requires that almost surely in ω ,

Page 144: Surveys and Tutorials in the Applied Mathematical Sciences

138 5 KPP Fronts in Random Media

∫ 1

0inf

x(g(x,ω) f0(u))du≥ εb > 0,

for a positive constant εb > 0. Both conditions help to prevent front pinning bythe random media. It is proved [174] that a GTF exists with the properties thatUt > 0, and there is a continuous increasing function X(t), the interface, satisfying(1) U(X(t), t) = θ ∈ (0,1), (2) shape invariance (5.162) holds almost surely in ω ,and (3)

limR→+∞

supt∈R

supx>R

U(x+X(t), t,ω) = 0,

limR→+∞

inft∈R

infx<−R

U(x+X(t), t,ω) = 1,

almost surely in ω . The GTF is proved [162] to be unique up to a constant translationin x and stable with perturbations decaying exponentially fast in time. Moreover,

limt→∞

X(t)/t = c∗ (5.163)

almost surely, where c∗ ∈ (cmin,cmax)⊂ (0,+∞) is the finite large-time front speed.The GTF satisfies all the definitions above. It has an invariant shape and finite widthfor all time, extending known traveling fronts in homogeneous and periodic media.A similar GTF is studied for a free boundary model in [45, 163].

Analogous to Burgers and HJ fronts in Chapters 3 and 4, the GTF speed fluctua-tions around c∗t is proved to be Gaussian under a condition of sufficient mixing ofthe random media [172]. See also [230] for a related study on the asymptotic tail be-havior of a semilinear heat equation with a random source at the origin. The mixingcondition means that the events related to the random media in the past (x ∼ −∞)and in the future (x∼+∞) are close to being independent; see Chapter 3 and [172]for precise mathematical characterizations. The main result is given in the followingtheorem [172]:

Theorem 5.25. Consider the GTFs of equation (5.160) for bistable or ignition-typenonlinearity f . Assume that the random process g(x,ω) is sufficiently mixing. Theneither (A) there is a positive constant κ0 > 0 such that

limt→+∞

P(

X(t,ω)− c∗tκ0√

t< α

)= N(α), ∀α ∈ R, (5.164)

where

N(α) =1√2π

∫ c

−∞exp

−y2

2

dy,

the unit Gaussian distribution function; or (B)

limt→+∞

P(

X(t,ω)− c∗t√t

> α)

= 0, ∀α ∈ R. (5.165)

In case (A), the invariance principle holds:

Page 145: Surveys and Tutorials in the Applied Mathematical Sciences

5.6 Generalized Fronts, Reactive Systems, and Geometric Models 139

X(nt,ω)− c∗ntκ0√

nlaw→W (t), t ∈ [0,1], (5.166)

where W (t) is the standard Wiener process. An example exists for case A.

The ideas of proofs for (5.163) and Theorem 5.25 are as follows. Let Tn be therandom times at which the interface X(t,ω) reaches the integer points x = n. SinceX is increasing in t, its inverse T (x,ω) is well defined such that x = X(T (x,ω),ω),Tn = T (n,ω), n = 1,2, . . . . It follows from (5.162) that

U(x+ y,T (y,ω),ω) = U(x,0,τyω), ∀y.

So as the front passes through the point y, the front profile is statistically invariant.Hence the sequence of increments ∆Tn = Tn+1 − Tn is stationary and satisfies thelaw of large numbers (ergodic theorem [41, Section 6.5]):

limn→+∞

Tn

n= lim

n→∞

1n

n−1

∑j=1

∆Tj =1c∗

,

which is another form of (5.163). To prove (5.164)–(5.166), it suffices to demon-strate that

limn→+∞

E

[∣∣∣∣Tn−n/c∗√

n

∣∣∣∣2]

= σ2, (5.167)

for some constant σ ≥ 0, and if σ > 0, then the family of processes

Zn(x) =Txn−nx/c∗

σ√

nlaw→W (x), (5.168)

for x ∈ [0,R] and any positive R. The difficulty is that the increments ∆Tn are corre-lated in a complicated way by the nonlinear PDE (5.160). What comes to the rescueis the stability property of the GTF, which implies that the interfacial motion de-pends primarily on the local environment, and only weakly on the distant past andfuture. In other words, the interfacial motion forgets its past and ∆Tn has enoughdecay of correlations at large times, and the Gaussian statistics of GTF speed fluc-tuation follows, except in a degenerate situation (σ = 0).

It is not known whether CLT is true for KPP fronts, though its quadratic viscousHJ approximation obeys CLT, as shown in Chapter 4. The limit law (5.163) is thenon-KPP analogue of the KPP front speeds in a one-dimensional spatial randommedium [94, 100], which served as a fundamental and inspiring first step in thestudy of RD fronts in random media. In the KPP case, (5.163) holds without theknowledge of GTF. In fact, GTF is not known to exist for KPP, type-2 and type-4reactions. It remains to study GTF in random flows in multiple space dimensionsand compare with KPP fronts.

Page 146: Surveys and Tutorials in the Applied Mathematical Sciences

140 5 KPP Fronts in Random Media

5.6.2 Fronts in Reaction–Diffusion Systems

The RD systems in combustion or autocatalytic reactions [238, 28, 33, 53] are ofthe following form:

ut = ∆xu+ v f (u), x ∈ Rd ,

vt = Le−1∆xv− v f (u), (5.169)

where Le > 0 is called the Lewis number. If (5.169) models premixed flame frontsin a one-step exothermic chemical reaction of the form A → B, then u is the tem-perature of the reacting mixture and v is the mass fraction of the reactant A. Thefunction f takes the Arrhenius form e−E/T , with activation energy constant E > 0.

If d = 1, then adding the two equations shows that u+v satisfies the heat equationand hence is forever equal to one if this is so arranged at t = 0. Replacing v by 1−uin the first equation of (5.169), we find a scalar R-D equation of type 4, and type 5then arises as we introduce a temperature cutoff θ .

For existence and uniqueness of traveling fronts of the form (u,v) = (U(p · x−ct),V (p · x− ct)), p a given unit vector, see [28, 37, 52, 155]. It is well known thatif d is much larger or smaller than 1, then fronts are unstable; see [14, 26, 123, 223]and references therein. Intuitively, the very distinct diffusion constants cause thefront to develop spatial–temporal scales as a way of keeping balance. With Le > 1,fronts oscillate in time, and with Le < 1, they generate transverse spatial oscillationsin two or three dimensions. The scales continue to grow with Le, and eventually thesolutions are chaotic.

When (5.169) models isothermal autocatalytic reactions of the form A + mB →(m+1)B, m≥ 1, with rate law proportional to vum, where v and u are the concentra-tions of the reactant A and the catalyst B, the function f is now f (u) = vum. Againwhen Le = 1, we recover a scalar R-D equation of type 1 if m = 1, of type 2 if m≥ 2.Existence and dynamics of fronts are discussed in [33, 34, 35, 52, 53, 93]. Similarly,if Le is sufficiently far away from one, fronts are unstable and can be chaotic; see[119, 154, 164].

Scalar R-D equations of type 3 come from the FitzHugh–Nagumo (FHN) systemin mathematical biology,

ut = ∆x u+u−u3− v, x ∈ R,

vt = ε(u− γv), (5.170)

where γ > 0, and ε > 0 is a small parameter. In the limit ε → 0, (5.170) reduces toa bistable scalar R-D equation; see [89, 122, 167, 170, 211].

The front problem of KPP systems in a shear flow was studied recently [111].The system of equations is (T temperature, and Y concentration of reactant)

Tt +u(y)Tx = ∆x,yT + f (T )Y,

Yt +u(y)Yx = Le−1∆x,yY − f (T )Y, (5.171)

Page 147: Surveys and Tutorials in the Applied Mathematical Sciences

5.6 Generalized Fronts, Reactive Systems, and Geometric Models 141

where (x,y) ∈ R×Ω , u(y) is Holder continuous,∫

Ω u(y)dy = 0, and f is of KPPtype: f is continuously differentiable, and

f (0) = 0 < f (s)≤ f ′(0)s, f ′(s)≥ 0, ∀s > 0, f (+∞) = +∞.

In particular f (T ) = T . The boundary conditions are zero Neumann on R× ∂Ω .The traveling fronts are solutions of the form T = T (x− ct,y), Y = Y (x− ct,y), sothat (T ,Y )(+∞,y) = (0,1), (T ,Y )(−∞,y) = (1,0), uniformly in y ∈ Ω . The mainfinding is contained in the following theorem [111]:

Theorem 5.26. For any Le > 0, the minimal KPP front speed c∗ is same as thatin the case of the unit Lewis number. If and only if c ≥ c∗, a traveling front T =T (x− ct,y), Y = Y (x− ct,y) exists such that T > 0, 0 < Y < 1, T is bounded.

The interesting fact is that the KPP minimal speed is independent of Le > 0. Thiswas first pointed out in [33] for homogeneous media. In [53], c∗ is also proved tobe the large-time asymptotic speed selected by compactly supported initial data ofT (x,0) (Y (x,0) = 1) in the absence of flow.

It would be very interesting to confirm the Le independence of c∗ for more gen-eral flows, especially random flows.

The existence of traveling fronts is studied in [22, 110] when the boundary con-ditions contains a heat loss, e.g., ∂

∂n T + σT = 0 on R×Ω , n the unit normal di-rection, and the minimal speed c∗ = c∗(σ) obeys a variational principle. The proofsare based on topological degree theory and PDE estimates. When f is non-KPP,traveling fronts are studied in the perturbative regime in which Le is near one; see[69, 70] for the ignition nonlinearity.

5.6.3 Geometric Models and Huygens Fronts

One way to model motion of an interface (a curve in R2, a surface in R3) is toprescribe a motion law in terms of its normal velocity, denoted by Vn. The simplestlaw is that Vn = c, a constant. We shall represent the interface by a constant level setof a scalar function G = G(x, t); see [187] and references therein for level set theory,numerical methods, and applications.

In the level set formulation, the normal velocity satisfies Vn = −Gt/|∇xG|. Forexample, an expanding circle (sphere) can be represented as G = |x|2−t2 = 0, x∈R2

(R3). The inside part is G < 0, and the outside part is G > 0. The derivatives of thelevel function are Gt =−2t, ∇G = 2x, |∇G|= 2|x|. When restricted to the level G =0, we have |∇G|= 2|x|= 2t, and so −Gt/|∇xG|= 1 gives the unit normal velocitypointing from the inside (G < 0) to the outside (G > 0). The outward-pointing unitnormal is n = ∇G/|∇G|, which is simply x/|x| in our example. In general, we shallconsider an expanding closed curve or surface, as shown in the experimental pictureof Chapter 1. A geometric quantity is the mean curvature defined as the divergenceof the normal,

Page 148: Surveys and Tutorials in the Applied Mathematical Sciences

142 5 KPP Fronts in Random Media

κ = ∇ ·n = ∇ ·∇G/|∇G|, (5.172)

which is positive (negative) if the interface is locally convex (concave). The meancurvature is κ = 1/|x| (κ = 2/|x|) for a circle (sphere) of finite radius.

The next-simplest law is Vn = g(x, t), g being a given space–time function. Incombustion, g(x, t) is chosen as the sum of a constant laminar speed −sL and thelocal fluid velocity v(x, t) along the normal n of a thin flame [238]. So

Vn =− Gt

|∇xG| =−sL +v(x, t) ·n =−sL +v ·∇G|∇G| ,

which is just the Hamilton–Jacobi equation

Gt +v(x, t) ·∇G = sL|∇G|, (5.173)

known as the G-equation [156, 238].The G = 0 isocontour represents implicitly the reaction surface of a moving

flame front whose width is infinitesimal. Such fronts are also called Huygens fronts;their dynamics depend only on the local environment. The Hamiltonian function of(5.173) is H = H(p,x, t) = sL|p| − v(x, t) · p. Recall the remark at the end of Sec-tion 2.4; the asymptotic HJ equation for bistable and ignition fronts has a linearlygrowing (relativistic) Hamiltonian H(p,x, t) = O(|p|) at large |p| for fixed (x, t),which is also the case for the G-equation (5.173). In contrast, the KPP Hamiltonianis quadratic (classical) in p, H = O(|p|2). From the Hamiltonian perspective, theG-equation models the bistable and ignition fronts better than the KPP fronts.

The G-equation has been widely adopted in the combustion literature. Variousanalytical and numerical approximations on front speeds in periodic and randomflows are based on it; see [126, 127, 251, 224, 11, 204, 54, 194, 1, 2, 51], amongmany others. A comparative study of front speeds in the G-equation and the KPPequation is conducted in [76], where examples of shear flow with nonzero meanshow that the front speeds from the G-equation are less than the KPP speeds. Theamount of discrepancy may vary depending on the alignment of the mean flowand the shear flow. In [77], the validity of Huygens fronts and the G-equation isstudied for piecewise linear reactions and linear incompressible flows. In spite ofthe differences between Huygens fronts and RD fronts, theoretical estimates [251]from the G-equation using a formal renormalization group (RG) method have beenfound to match the empirical speed growth laws for the aqueous autocatalytic chem-ical reaction fronts [222]. The flows generated in such experiments range fromcapillary-wave flow, Taylor–Couette flow, to vibrating-grid turbulence [222]. For theG-equation (5.173) with space–time random velocity field v, the (turbulent) frontspeed sT is defined as sL〈|∇G|〉, where the bracket is the ensemble average. Theformula for sT by RG analysis [251] is

UT = exp(U/UT )p (5.174)

for p = 2, where UT = sT /sL, and U is the ratio of the root-mean-square amplitudeof v and sL. For large U , (5.174) implies that UT is approximately U/

√lnU .

Page 149: Surveys and Tutorials in the Applied Mathematical Sciences

5.6 Generalized Fronts, Reactive Systems, and Geometric Models 143

In terms of establishing the almost sure existence of sT from stochastic homog-enization methods, one observes immediately that the G-equation violates the coer-civity condition (H1) or (A2) or similar assumptions in Chapter 4. For a mean-zero-velocity field v with larger amplitude than sL, we have that H(p,x, t), although stillconvex in p, does not grow to +∞ as |p| → +∞. The current mathematical theorydoes not apply. Stochastic homogenization of the G-equation remains an interestingtopic for future research.

The G-equation (5.173) clearly ignores diffusion, as suggested by the viscous HJof the KPP equation (5.123). A generalized G-equation has been proposed [193] tomodel fronts in gaseous combustion systems. Some physical length scales are intro-duced into (5.173). A characteristic length for flame is called the Markstein lengthLm = l f α(Le,E), where l f is flame thickness, and α is a nondimensional numberdepending on the Lewis number Le and the activation energy E, [193, (1.13)]. Thegeneralized motion law [193] is derived by replacing sL in (5.173) by

sL = sL(1+Lmκ)+Lmn ·∇v ·n, (5.175)

where the correction terms contribute to flame-stretching due to flame geometry andflows. The mean curvature κ is a second-order term similar to diffusion. In fact, itis equal to ∆dist f , where dist f is the signed distance function to the front.

So the generalized G-equation is a viscous (parabolic) stochastic HJ equationworthy of analysis and qualitative comparison with (5.173) in terms of the propertiesof their (turbulent) front speeds sT .

Geometric models have also been proposed for the study of phase boundary mo-tion through a heterogeneous material (a matrix with precipitates) [60, 61]. Themotion law is

Vn = f (x)− cκ ,

which becomes, in the level set formulation,

ht =− f (x)|∇h|+ c|∇h|∇ · (∇h/|∇h|), (5.176)

where f (x) is the jump in energy density across the interface and c is a nonnega-tive constant. The term f − cκ is a thermodynamic force driving the phase bound-ary [61]. If f is a constant plus noise, the model (5.176) also appears in statisticalmechanics [192] and the study of dislocation loops in materials [108]. In case ofdislocation, the function f (x) is equal to a constant except on point defects or inclu-sions. The defects may be periodically or randomly distributed. If f (x) is periodicand has a fixed sign (hence coercive), then periodic homogenization (scale c to εcand f to f (x/ε)) is studied in the spirit of [148, 79] for both c = 0 [60] and c > 0[61]. In the case c > 0 and f does not change sign, the homogenization problemis treated in [151] for both periodic and almost-periodic media. The homogenizedmotion is Vn = f (n) for some continuous function f , implying an anisotropic ge-ometric law. See [60] for variational formulas of f in the case c = 0, and [61] foran explicit formula of f in the large-c limit. If f changes sign, front-trapping mayoccur [60, 61, 67]. The existence and uniqueness of pulsating traveling fronts with

Page 150: Surveys and Tutorials in the Applied Mathematical Sciences

144 5 KPP Fronts in Random Media

constant speeds are proved [66] if f changes sign and is small enough and if c > 0.The stochastic homogenization problem of (5.176) remains largely open.

5.7 Exercises

1. Show that the Lyapunov exponent µ(λ ) at (λ1,0), λ1 ∈R, for the time-randomshear flow b = b(y, t,ω) satisfies the inequality µ(λ1,0) ≥ λ 2

1 /2. Then deducefrom the speed variational formula that c∗(δ ), the front speed in scaled shearflow δb(y, t), is no less than c∗(0), the speed in the absence of flow.

2. Make the change of variable ψ = log(φ) in the equation (5.77) and derive aviscous quadratic HJ equation for ψ . Then show that c∗(δ ) = c∗(0) if and onlyif b = b(t).

3. Derive the linear upper bound (iii) of Theorem 5.9 by the Feynman–Kac for-mula

φ = E[eλδ

∫ t0 b(W (s),t−s)ds

].

4. Verify that if the shear flow b is a mean-zero Gaussian process, then for anyfixed continuous path W ∈C([0,1],R2), the random variable

ξ (t,W )≡ e−∫ t

0 λ1b(W2(s)+z,t−s)ds−λ1W1(t)−λ2W2(t)

is lognormal with mean equal to

E[ξ (t,W )] = e|λ |2σ2/2e−λ1W1(t)−λ2W2(t),

whereσ 2 =

∫ t

0

∫ t

(W (s),W (r),s,r

)dsdr,

where Γ is the covariance function of the process b.

Page 151: Surveys and Tutorials in the Applied Mathematical Sciences

References

1. M. Abel, A. Celani, D. Vergni, A. Vulpiani, Front propagation in laminar flows, Phys.Review E 64, 046307 (2001), 1–12.

2. M. Abel, M. Cencini, D. Vergni, A. Vulpiani, Front speed enhancement in cellular flows12(2) (2002), 481–488.

3. R. Abraham, J.E. Marsden, Foundations of Mechanics, 2nd edition, Addison-WesleyPublishing Co, Reading, MA, 1978.

4. R. Adler, The Geometry of Random Fields, John Wiley and Sons, 1980.5. N. Alikakos, P. Bates, X. Chen, Traveling waves in a time periodic structure and a

singular perturbation problem, Transactions of AMS 351 (1999), 2777–2805.6. E. Anderson, et. al., LAPACK Users’ Guide, Third Edition, Society for Industrial and

Applied Mathematics, 1999.7. V. Arnold, Mathematical Method of Classical Mechanics, Springer-Verlag, 1978.8. D. G. Aronson and H. F. Weinberger, Multidimensional nonlinear diffusion arising in

population genetics, Adv. in Math. 30 (1978), 33–76.9. D. Aronson and H. Weinberger, Nonlinear diffusion in population genetics, combustion,

and nerve propagation. Lecture Notes in Mathematics 446, pp. 5–49, Springer-Verlag,1975.

10. W. Ashurst, Flow-frequency effect upon Huygens front propagation, Combust. TheoryModelling 4 (2000), 99–105.

11. W. Ashurst, G.I. Sivashinsky, On flame propagation through periodic flow fields, Com-bust. Sci. Technol. 80 (1991), 159.

12. B. Audoly, H. Berestycki, and Y. Pomeau, Reaction-diffusion en ecoulement station-naire rapide, Note C. R. Acad. Sci. Paris Ser. II 328 (2000), 255–262.

13. A.-L. Barabasi and H. E. Stanley, Fractal Concepts in Surface Growth, Cambridge Univ.Press, Cambridge, 1995.

14. G. I. Barenblatt, Y. B. Zeldovich and A. G. Istranov, On the diffusional-thermal stabilityof a laminar flame, Zh. Prikl. Mekh. Tekh. Fiz. 4 (1962), 21–26.

15. J. Bebernes, C. Li, Y. Li, Travelling fronts in cylinders and their stability, Rocky Moun-tain J. of Math, 27(1) (1997), 23–150.

16. R. D. Benguria and M. C. Depassier, Variational characterization of the speed of prop-agation of fronts for the nonlinear diffusion equation, Comm. Math. Phys. 175 (1996),221–227.

17. R. D. Benguria and M. C. Depassier, Speed of fronts of the reaction-diffusion equation,Phys. Rev. Lett. 77 (1996), 1171–1173.

18. A. Bensoussan, J.-L. Lions and G. Papanicolaou, Asymptotic Analysis for PeriodicStructures. Studies in Mathematics and its Applications, Vol. 5, North-Holland, Publ.,1978.

145

Page 152: Surveys and Tutorials in the Applied Mathematical Sciences

146 References

19. H. Berestycki, The influence of advection on the propagation of fronts in reaction-diffusion equations, in Proceedings of the NATO ASI Conference, Cargese, France,H. Berestycki and Y. Pomeau, eds., Kluwer, Dordrecht, the Netherlands, 2003.

20. H. Berestycki, F. Hamel, Front propagation in periodic excitable media, Comm. Pureand Appl. Math, 55 (2002), 949–1032.

21. H. Berestycki, F. Hamel, Generalized traveling waves for reaction-diffusion equations,Contemp. Math 446, Amer. Math. Soc., pp. 101–123, 2007.

22. H. Berestycki, F. Hamel, A. Kiselev, L. Ryzhik, Quenching and propagation in KPPreaction-diffusion equations with a heat loss, Arch. Ration. Mech. Anal 178 (2005),57–80.

23. H. Berestycki, F. Hamel, H. Matano, Bistable travelling waves around an obstacle,preprint, 2008.

24. H. Berestycki, F. Hamel, and N. Nadirashvili, Elliptic eigenvalue problems with largedrift and applications to nonlinear propagation phenomena, Comm. Math. Phys, 253(2005), 451–480.

25. H. Berestycki, F. Hamel, L. Roques, Analysis of the periodically fragmented environ-ment model: 2-biological invasions and pulsating travelling fronts, J. Math. Pures Appl.84 (2005), 1101–1146.

26. H. Berestycki and B. Larrouturou, Quelque aspect mathematique de la propagation desflames premelangee, p. 65, Pitman Research Notes in Mathematics Series, H. Brezisand J.-L. Lions eds., Vol. X, 1991.

27. H. Berestycki, B. Larrouturou and P.-L. Lions, Multi-dimensional travelling-wave solu-tions of a flame propagation model. Arch. Rat. Mech. Anal. 111 (1990), 33–49.

28. H. Berestycki, B. Nicolaenco and B. Scheurer, Traveling wave solution to combustionmodels and their singular limits, SIAM J. Math. Anal. 16 (1985), 1207–1242.

29. H. Berestycki and L. Nirenberg, Some qualitative properties of solutions of semilin-ear elliptic equations in cylindrical domains, Analysis etc., ed., P. Rabinowitz, et al.,Academic Pr., 1990, pp. 115–164.

30. H. Berestycki and L. Nirenberg, On the method of moving planes and the slidingmethod, Bol. da Soc. Brasileira de Matematica, 22 (1991), 1–37.

31. H. Berestycki and L. Nirenberg, Traveling Fronts in Cylinders, Annales de l’IHP, Anal-yse non lineaire 9 (1992), 497–572.

32. J. Billingham and D. J. Needham, A note on the properties of a family of traveling wavesolutions arising in cubic autocatalysis, Dynamics Stabil. Systems 6 (1991), 33–49.

33. J. Billingham and D. J. Needham, The development of traveling waves in quadratic andcubic autocatalysis with unequal diffusion rates, I. Permanent form traveling waves,Phil. Trans. R. Soc. Lond. A 334 (1991), 1–24.

34. J. Billingham and D. J. Needham, The development of traveling waves in quadratic andcubic autocatalysis with unequal diffusion rates, II. The initial value problem with animmobilized or nearly immobilized autocatalyst., Phil. Trans. Roy. Soc. Lond. A 336(1991), 497–539.

35. J. Billingham and D. J. Needham, The development of traveling waves in quadraticand cubic autocatalysis with unequal diffusion rates, III. Large time development inquadratic autocatalysis, Quart. Appl. Math. 1 (1992), 343–372.

36. P. Billingsley, Convergence of Probability Measures, Wiley, 1968.37. A. Bonnet, B. Larrouturou and L. Sainsaulieu, On the stability of multiple steady planar

flames when the Lewis number is less than one, Phys. D 69 (1993), 345–352.38. W. Bosma and S. van der Zee, Transport of reacting solute in a one-dimensional chem-

ically heterogeneous porous media, Water Resources Research 29 (1993), 117–131.39. W. Bosma, S. van der Zee and C. J. van Duijn, Plume development of a nonlinearly

adsorbing solute in heterogeneous porous formations, Water Resources Research 32(1996), 1569–1584.

40. M. Bramson, Convergence of solutions of the Kolmogorov equations to traveling waves,Mem. of AMS 285 (1983).

Page 153: Surveys and Tutorials in the Applied Mathematical Sciences

References 147

41. L. Breiman, Probability, Classics in Applied Mathematics, SIAM, 1992.42. J. Bricmont and A. Kupiainen, Stability of moving fronts in the Ginzburg–Landau equa-

tions, Comm. Math. Phys. 159 (1994), 287–318.43. R. J. Briggs, Electron-Stream Interaction with Plasmas, MIT press, Cambridge, 1964.44. N. F. Britton, Reaction-Diffusion Equations and Their Applications to Biology, London,

Academic Press, 1986.45. L. Caffarelli, A. Mellet, Flame propagation in one-dimensional stationary ergodic me-

dia, Math. Models Methods Appl Sci 17 (2007), 155–169.46. L. Caffarelli, K. Lee, A. Mellet, Singular limit and homogenization for flame propaga-

tion in periodic excitable media, Arch. Rat. Mech. Analy. 172(2) (2004), 153–190.47. L. Caffarelli, K. Lee, A. Mellet, Homogenization and flame propagation in periodic

excitable media: the asymptotic speed of propagation, Comm. Pure Appl. Math 59(4)(2006), 501–525.

48. R. Carmona, S. Molchanov, Parabolic Anderson problem and intermittency. Mem.Amer. Math. Soc. 108 (1994), no. 518, viii+125.

49. R. Carmona, S. Molchanov, Stationary parabolic Anderson model and intermittency,Prob. Theory Rel. Fields 102 (1995), 433–453.

50. R. Carmona, L. Koralov, S. Molchanov, Asymptotics for the almost sure Lyapunov ex-ponent for the solution of the parabolic Anderson problem, Random Oper. and Stoch.Equ. 9(1) (2001), 77–86.

51. M. Cencini, A. Torcini, D. Vergni, A. Vulpiani, Thin front propagation in steady andunsteady cellular flows, Phys. of Fluids 15(3) (2003), 670–688.

52. X. Chen, Y. Qi, Sharp estimates on minimum traveling wave speed of reaction diffusionsystems modeling autocatalyst, SIAM J. Math Analysis 39 (2007), 437–448.

53. X. Chen, Y. Qi, Propagation of local disturbances in reaction diffusion systems model-ing quadratic autocatalysis, SIAM J. Appl. Math 69(1) (2008), 273–282.

54. M. Chertkov, V. Yakhot, Hyugens front propagation in turbulent fluids, Phys. Rev. Lett80 (1998), 2837–2841.

55. A. Chorin, O. Hald, Stochastic Tools in Mathematics and Science, Vol. 1, Surveys andTutorials in the Applied Mathematical Sciences, Springer, 2006.

56. P. Clavin and F. A. Williams, Theory of premixed-flame propagation in large-scale tur-bulence, J. Fluid Mech. 90 (1979), 598–604.

57. M. Concordel, Periodic homogenization of Hamilton-Jacobi equations I: additive eigen-values and variational formula, Indiana Univ math J 45 (1996), 1095–1117.

58. J. Conlon, C. Doering, On Traveling Waves for the Stochastic FKKP Equation, Jour.Stat Physics 120(3-4) (2005), 421–477.

59. P. Constantin, A. Kiselev, A. Oberman, and L. Ryzhik, Bulk burning rate in passive-reactive diffusion, Arch. Ration. Mech. Anal. 154 (2000), 53–91.

60. B. Craciun and K. Bhattacharya, Homogenization of a Hamilton-Jacobi equation as-sociated with the geometric motion of an interface, Proceedings of the Royal SocietyEdinburgh A. 133 (2003), 773–805.

61. B. Craciun and K. Bhattacharya, Effective motion of a curvature sensitive interfacethrough a heterogeneous medium, Inter. Free. Bound. 6 (2004), 151–173.

62. M. Cranston and T. Mountford, Lyapunov exponent for the parabolic Anderson modelin Rd , Journal of Functional Analysis 236 (2006), 78–119.

63. G. Dee, J.S. Langer, Propagating Pattern Selection, Phys. Rev Lett 50 (1983), 383–387.64. D. A. Dawson and E. Perkins, Historical processes, Mem. Amer. Math. Soc. 93 (1991).65. B. Denet, Possible role of temporal correlations in the bending of turbulent flame veloc-

ity, Combust. Theory Modelling 3 (1999), 585–589.66. N. Dirr, G. Karali, A. Yip, Pulsating Wave for Mean Curvature Flow in Inhomogeneous

Medium, European J. of Applied Mathematics 19 (2008), 661–699.67. N. Dirr, A. Yip, Pinning and De-Pinning Phenomena in Front Propagation in Hetero-

geneous Medium, Interfaces and Free Boundaries, 8 (2006), 79–109.

Page 154: Surveys and Tutorials in the Applied Mathematical Sciences

148 References

68. C. Doering, C. Mueller, P. Smereka, Interacting particles, the stochastic FKPP equationand duality, Phys. A 325 (2003), 243–259.

69. A. Ducrot, Multi-dimensional combustion waves for Lewis number close to one, Math.Methods Appl. Sci. 20 (2007), 291–304.

70. A. Ducrot, M. Marion, Two-dimensional traveling wave solutions of a system modellingnear equi-diffusional flames, Nonlinear Anal 61 (2005), 1105–1134.

71. C. J. van Duijn and P. Knabner, Solute transport in porous media with equilibrium andnonequilibrium multiple sites adsorption, travelling waves, J. Reine Angew. Math. 415(1991), 1–49.

72. R. Durrett, Probability: Theory and Examples, 2nd ed., Wadsworth and Brooks, 1995.73. Weinan E, Y. Sinai, New results in mathematical and statistical hydrodynamics, Russian

Math. Surveys 55(4) (2000), 635–666.74. Weinan E, J. Wehr, J. Xin, Breakdown of homogenization for the random Hamilton-

Jacobi equations, Comm. Math Sci, 6(1) (2008), 189–197.75. J-P. Eckmann and C.E. Wayne, The nonlinear stability of front solutions for parabolic

pde’s, Comm. Math. Phys. 161 (1994), 323–334.76. P. Embid, A. Majda and P. Souganidis, Comparison of turbulent flame speeds from com-

plete averaging and the G-equation, Phys. Fluids 7(8) (1995), 2052–2060.77. P. Embid, A. Majda and P. Souganidis, Examples and Counterexamples for Huygens

Principle in Premixed Combustion, Combustion Sci and Tech. 120 (1996), 73–303.78. L. C. Evans, Weak Convergence Methods for Nonlinear PDE’s, NSF-CBMS Regional

Conference Lectures, 1988.79. L. C. Evans, Periodic homogenization of certain fully nonlinear partial differential

equations, Proc. Royal Soc. Edingburgh, Section A 127 (1992), 65–689.80. L. C. Evans, Partial Differential Equations, Graduate Studies in Mathematics, American

Mathematical Society, Providence, R.I, 1998.81. L. C. Evans and P. E. Souganidis, A PDE approach to geometric optics for certain

semilinear parabolic equations, Indiana Math. J., 38 (1989), 141–172.82. L. C. Evans and P. E. Souganidis, A PDE approach to certain large deviation problems

for systems of parabolic equations, Ann. Inst. H. Poincare, Anal. Nonlineare, 6 (1989),Suppl., 229–258.

83. L. C. Evans and P. E. Souganidis, Differential games and representation formulas forsolutions of Hamilton–Jacobi–Isaacs equations, Indiana Math. J. 33 (1984), 773–797.

84. E. B. Fabes and D. W. Stroock, A new proof of Moser’s parabolic Harnack inequalityusing the old ideas of Nash, Arch. Rat. Mech. Anal. 96 (1986), 327–338.

85. H. Fan, Large time behavior of elementary waves of Burgers equation under white noiseperturbations, Comm. Partial Differential Equations, 20(9-10) (1995), 1699–1723.

86. A. Fannjiang and G. Papanicolaou, Convection enhanced diffusion for periodic flows,SIAM J. Appl. Math, 54 (1994), 333–408.

87. A. Fannjiang and G. Papanicolaou, Convection enhanced diffusion for random flows,Journal of Statistical Physics 88 (1997), 1033–1076.

88. G. Fennemore and J. Xin, Wetting fronts in one dimensional periodically layered soils,SIAM J. Applied Math. 58 (1998), 387–427.

89. P. C. Fife, Mathematical Aspects of Reacting and Diffusing Systems, Lecture Notes inBiomath, 28, Springer, New York, 1979.

90. P. C. Fife, Dynamics of Internal Layers and Diffusive Interfaces, CBMS-NSF RegionalConf. Ser. in Appl. Math., SIAM, 1988.

91. R. A. Fisher, The wave of advance of advantageous genes, Ann. Eugenics 7 (1937),355–369.

92. W. H. Fleming and H. M. Soner, Controlled Markov Processes and Viscosity Solutions,Applications of Mathematics, 25, Springer-Verlag, 1993.

93. S. Focant and Th. Galley, Existence and stability of propagating fronts for an autocat-alytic reaction-diffusion system, Univ. de Paris-Sud preprint, 97–38, 1997.

Page 155: Surveys and Tutorials in the Applied Mathematical Sciences

References 149

94. M. Freidlin, Propagation of a concentration wave in the presence of random motionassociated with the growth of a substance, Soviet Math. Dokl. 20 (1979), 503–507.

95. M. Freidlin, Limit theorems for large deviations and reaction-diffusion equations. An-nals Prob. 13 (1985), 639–675.

96. M. Freidlin, Functional Integration and Partial Differential Equations. Annals of Math-ematics Studies 109, Princeton University Press, 1985.

97. G. Frejacques, Traveling waves in infinite cylinders with time-periodic coefficients, Ph.DThesis, Universite Aix-Marseille, 2006.

98. Th. Gallay, Local Stability of critical fronts in nonlinear parabolic pde’s, Nonlinearity7 (1994), 741–764.

99. R. Gardner, Existence of multidimensional traveling wave solutions of an initial bound-ary value problem, J. Diff. Eqn. 61 (1986), 335–379.

100. J. Gartner and M. I. Freidlin, On the propagation of concentration waves in periodicand random media, Soviet Math. Dokl. 20 (1979), 1282–1286.

101. J. Gartner, W. Konig, S. Molchanov, Almost sure asymptotics for the continuousparabolic Anderson model, Probab. Theory Relat. Fields, 118 (2000), 547–573.

102. D. Gilbarg, N. Trudinger, Elliptic Partial Differential Equations of Second Order,Springer-Verlag, 2nd edition, 1983.

103. H. Goldstein, Classical Mechanics, Addison-Wesley Publishing Co, 1950.104. D. Gomes, M. Oberman, Computing the Effective Hamitonian Using a Variational Ap-

proach, SIAM J. Control. Optim. 43(3) (2004), 792–812.105. P. Groeneboom, Brownian Motion with a Parabolic Drift and Airy Functions, Prob. Th.

Rel. Fields 81 (1989), 79–109.106. J-S Guo, F. Hamel, Front propagation for discrete periodic monostable equations, Math.

Ann, 335 (2006), 489–525.107. K. P. Hadeler and F. Rothe, Traveling fronts in nonlinear diffusion equations, J. Math.

Biology 2 (1975), 251–263.108. T. Halpin-Healy, Y.C. Zhang, Kinetic roughening phenomena, stochastic growth, di-

rected polymers and all that—Aspects of multidisciplinary statistical mechanics, Phys.Report 254 (1995), 215–415.

109. F. Hamel, Formules min-max pour les vitesses d’ondes progressives multidimension-nelles, Ann. Fac. Sci. Toulouse 8 (1999), 259–280.

110. F. Hamel, L. Ryzhik, Non-adiabatic KPP fronts with arbitrary Lewis number, Nonlin-earity 18 (2005), 2881–2902.

111. F. Hamel, L. Ryzhik, Traveling fronts for the thermal-diffusive system with arbitraryLewis numbers, Comm. PDE, 2009, to appear.

112. B. G. Haslam and P. D. Ronney, Fractal properties of propagating fronts in a stronglystirred fluid, Phys. Fluids 7 (1995), 1931–1937.

113. S. Heinze, Wave solutions to reaction-diffusion systems in perforated domains, Z. Anal.Anwendungen 20 (2001), 661–670.

114. S. Heinze, The Speed of Travelling Waves for Convective Reaction-Diffusion Equations,Preprint 84, Max-Planck-Institut fur Mathematik in den Naturewissenschaften, Leipzig,Germany, 2001.

115. S. Heinze, Large Convection Limits for KPP Fronts, preprint 21, Max-Planck-Institutfur Mathematik in den Naturewissenschaften, Leipzig, Germany, 2005.

116. S. Heinze, G. Papanicolaou, A. Stevens, Variational principles for propagation speedsin inhomogeneous media, SIAM J. Applied Math. 62(1), (2001), 129–148.

117. P. Hess, Periodic–Parabolic Boundary Value Problems and Positivity, vol. 247 of PitmanResearch Notes in Mathematics, Longman Scientific & Technical, Harlow, 1991.

118. H. Holden and N. H. Risbero, Stochastic Properties of the Scalar Buckley-LeverettEquation, SIAM J. Appl. Math.. 51(5) (1991), 1472–1488.

119. D. Horvath, V. Petrov, S. Scott, and K. Schowalter, Instabilities in propagating reaction-diffusion fronts, J. Chem. Phys. 98 (1993), 6332–6343.

Page 156: Surveys and Tutorials in the Applied Mathematical Sciences

150 References

120. J. Huang, W. Shen, Speeds of Spread and Propagation for KPP Models in Time Almostand Space Periodic Media, preprint, 2008.

121. A. M. Ilin, and O. A. Oleinik, Behavior of the solution of the Cauchy problem for certainquasilinear equations for unbounded increase of the time, AMS Transl. (2) 42 (1964),19–23.

122. C. Jones, Stability of traveling wave solutions of the FitzHugh–Nagumo system, Trans.AMS 286 (1984), 431–469.

123. G. Joulin and P. Clavin, Linear stability analysis of nonadiabatic flames: diffusion-thermal model, Combust. and Flame 35 (1979), 139–153.

124. J. Kanel, Stabilization of solutions of the Cauchy problem for equations encountered incombustion theory, Mat. Sbornik 59 (1962), 245–288.

125. I. Karatzas, S. Shreve, Brownian Motion and Stochastic Calculus, Springer-Verlag,1991.

126. A. R. Kerstein, W. T. Ashurst, and F. A. Williams, Field equations for interface propa-gation in an unsteady homogeneous flow field, Phys. Rev. A 37 (1988), 2728.

127. A. R. Kerstein and W. T. Ashurst, Propagation rate of growing interfaces in stirredfluids, Phys. Rev. Lett. 68 (1992), 934.

128. B. Khouider, A. Bourlioux, and A. Majda, Parameterizing turbulent flame speed-Part I:Unsteady shears, flame residence time and bending, Combust. Theory Model. 5 (2001),295–318.

129. I. Kim, Homogenization of free boundary velocities, Arch. Ration. Mech. Anal. 185(2007), 69–103.

130. I. Kim, A. Mellet, Homogenization of a Hele-Shaw type problem in periodic and ran-dom media, to appear in Arch. Ration. Mech. Anal.

131. K. Kirchgassner, On the nonlinear dynamics of traveling fronts, J. Diff. Eqs. 96 (1992),256–278.

132. A. Kiselev and L. Ryzhik, Enhancement of the traveling front speeds in reaction-diffusion equations with advection, Ann. Inst. H. Poincare Anal. Non Lineaire 18(2001), 309–358.

133. P. Kloeden, E. Platen, Numerical Solution of Stochastic Differential Equations,Springer-Verlag, Berlin, 1992.

134. A. N. Kolmogorov, I. G. Petrovsky and N. S. Piskunov, Etude de l’equation de la dif-fusion avec croissance de la quantite de matiere et son application a un probleme bi-ologique, Moskow Univ. Math. Bull. 1 (1937), 1–25.

135. E. Kosygina, F. Rezakhanlou, and S.R.S. Varadhan, Stochastic homogenization ofHamilton-Jacobi-Bellman equations. Comm. Pure Appl. Math. 59(10) (2006), 1489–1521.

136. E. Kosygina and S. R. S. Varadhan, Homogenization of Hamilton-Jacobi-Bellman equa-tions with respect to time-space shifts in a stationary ergodic medium, Comm. PureAppl. Math, 61(6) (2008), 816–847.

137. N. V. Krylov, M. V. Safonov, A property of the solutions of parabolic equations withmeasurable coefficients. (English translation) Izv. Akad. Nauk SSSR Ser. Mat., 16(1)(1981), 151–164.

138. J. Krug, H. Spohn, Kinetic roughening of growing surfaces, In: Solids Far Away fromEquilibrium, C. Godreche ed, Cambridge University Press, 1991, pp. 479–588.

139. L. D. Landau, E. M. Lifschitz, The Classical Theory of Fields, Oxford, Pergamon, 1960.140. J. S. Langer, H. Muller-Krumbhaar, Mode selection in a dendritelike nonlinear systems,

Phys. Rev. A 27 (1983), 499.141. P. D. Lax, Hyperbolic systems of conservation laws and the mathematical theory of

shock waves, Philadelphia SIAM Regional Conf. Ser. in Applied Math., No. 11 (1973).142. J. Leach, D. Needham, Matched Asymptotic Expansions in Reaction-Diffusion Theory,

Springer Monographs in Mathematics, 2003.

Page 157: Surveys and Tutorials in the Applied Mathematical Sciences

References 151

143. M. R. Leadbetter, G. Lindgren, H. Rootzn, Extremes and related properties of ran-dom sequences and processes, Springer Series in Statistics, New York, Springer-Verlag,1983.

144. T.-Y. Lee and F. Torcaso, Wave propagation in a lattice KPP equation in random media,Ann. Probab. 26 (1998), 1179–1197.

145. C. Li, Monotonicity and symmetry on bounded and unbounded domains, I & II, Comm.PDE, 16(2-3) (1991), 491–526; 16(4-5) (1991), 585–615.

146. T. M. Liggett, An improved subadditive ergodic theorem, Annals of Probability 13(1985), 1279–1285.

147. P.-L. Lions, Generalized Solutions of Hamilton-Jacobi equations, Pitman ResearchNotes in Mathematics, No. 69, Pitman Advanced Publishing Program, Boston, 1982.

148. P.-L. Lions, G. C. Papanicolaou and S. R. S. Varadhan, Homogenization of Hamilton–Jacobi equations, unpublished preprint, circa 1986.

149. P.-L. Lions and P. E. Souganidis, Correctors for the homogenization of Hamilton-Jacobiequations in the stationary ergodic setting, Comm. Pure Appl. Math. 56(10) (2003),1501–1524.

150. P.-L. Lions and P. E. Souganidis, Homogenization of viscous Hamilton-Jacobi equationsin stationary ergodic media, Comm. Partial Diff. Eqn. 30 (2005), 335–375.

151. P.-L. Lions and P. E. Souganidis, Homogenization of degenerate second order PDE inperiodic and almost periodic environments and applications, Ann Inst. H. Poincare,Analyse Non Lineaire 22 (2005), 667–677.

152. A. Majda and P. E. Souganidis, Large scale front dynamics for turbulent reaction-diffusion equations with separated velocity scales, Nonlinearity 7 (1994), 1–30.

153. A. Majda and P. E. Souganidis, Flame fronts in a turbulent combustion model withfractal velocity fields, Comm. Pure Appl. Math. LI (1998), 1337–1348.

154. A. Malevanets, A. Careta and R. Kapral, Biscale chaos in propagating front, Phys. Rev.E 52 (1995), 4724.

155. M. Marion, Qualitative properties of a nonlinear system for laminar flames withoutignition temperature, Nonlinear Anal., TMA, 11 (1983), 1269–1292.

156. G. Markstein, Nonsteady Flame Propagation, Pergamon Press, Oxford, 1964.157. J. E. Marsden, Lectures on Mechanics, London Math Soc Lecture Notes, 174, Cam-

bridge University Press, Cambridge, 1992.158. G. de Marsily, Quantitative Hydrogeology—Groundwater Hydrology for Engineers,

Academic Press, 1986.159. H. Matano, Traveling waves in spatially inhomogeneous diffusive media, a talk at Inst.

H. Poincare, 2002, among other conference presentations.160. H. McKean, Application of Brownian motion to the equation of KPP, Comm. Pure Ap-

plied Math. 28 (1975), 323–331.161. R. McLaughlin, J. Zhu, The Effect of Finite Front Thickness on the Enhanced Speed of

Propagation, Combus. Sci. Tech., 129 (1997), 89–112.162. A. Mellet, J. Nolen, L. Ryzhik, J.-M. Roquejoffre, Stability of Generalized Transition

Fronts, Comm. in PDE, 2009, to appear.163. A. Mellet, J.-M. Roquejoffre, Construction d’ondes generalise pour les models 1-d

scalaire a temperature d’ignition, preprint, 2008.164. M. Metcalf, J. Merkin and S. Scott, Oscillating wave fronts in isothermal chemical

systems with arbitrary powers of autocatalysis, Proc. Roy. Soc. Lond. A 447 (1994),155–174.

165. J. Mierczynski, W. Shen, Exponential separation and principal Lyapunov expo-nent/spectrum for random/nonautonomous parabolic equations, J. Differential Equa-tions 191 (2003), 175–205.

166. C. Mueller, R. Sowers, Random Traveling Waves for the KPP equation with Noise, J.Functional Analysis 128 (1995), 439–498.

167. J. D. Murray, Mathematical Biology, 2nd ed., Biomath texts, 19, 1993.

Page 158: Surveys and Tutorials in the Applied Mathematical Sciences

152 References

168. G. Nagin, Equations de reaction-diffusion et propagation en milieu heterogene, Thesede Doctorat, Universite Paris VI- Pierre et Marie Curie, June 2008.

169. J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer. J. Math. 80(1958), 931–954.

170. S. Nii, Stability of the traveling multiple-front (multiple-back) wave solutions of theFitzHugh–Nagumo equations, SIAM J. Math. Anal. 28 (1997), 1094–1112.

171. Y. Nishiura, Y. Oyama, K. Ueda, Dynamics of traveling pulses in heterogeneous mediaof jump type, Hokkaido Math Journal 16(1) (2007), 207–242.

172. J. Nolen, An Invariance principle for random traveling waves in one dimension,preprint, 2009.

173. J. Nolen, M. Rudd, J. Xin, Existence of KPP fronts in spatially-temporally periodicadvection and variational principle for propagation speeds, Dynamics of PDE 2(1)(2005), 1–24.

174. J. Nolen, L. Ryzhik, Traveling waves in a one-dimensional random medium, Ann Inst.H. Poincare, Analyse Non Lineaire, to appear, 2009.

175. J. Nolen, J. Xin Variational Principle Based Computation of KPP Average Front Speedsin Random Shear Flows, Methods and Applications of Analysis 11(3) (2004), 389–398.

176. J. Nolen, J. Xin, Min-Max Variational Principle and Front Speeds in Random ShearFlows, Methods and Applications of Analysis 11(4) (2004), 635–644.

177. J. Nolen, J. Xin, Reaction-diffusion front speeds in spatially-temporally periodic shearflows, SIAM J. Multiscale Modeling and Simulation 1(4) (2003), 554–570.

178. J. Nolen, J. Xin, A Variational Principle Based Study of KPP Minimal Front Speeds inRandom Shears, Nonlinearity 18 (2005), 1655–1675.

179. J. Nolen, J. Xin, Variational principle of KPP front speeds in temporally random shearflows with applications, Comm. Math. Phys. 269 (2007), 493–532.

180. J. Nolen, J. Xin, Computing Reactive Front Speeds in Random Flows by VariationalPrinciple, Physica D 237 (2008), 3172–3177.

181. J. Nolen, J. Xin, KPP Fronts in a One Dimensional Random Drift, Discrete and Contin-uous Dynamical Systems-B 11(2) (2009), 421–442.

182. J. Nolen, J. Xin, Asymptotic Spreading of KPP Reactive Fronts in Incompressible Space-Time Random Flows, Ann Inst. H. Poincare, Analyse Non Lineaire, to appear, 2009.

183. A. Novikov, G. Papanicolaou and L. Ryzhik, Boundary layers for cellular flows at highPeclet numbers, Comm. Pure and Appl. Mathematics LXIII, (2005), 867–922.

184. A. Novikov, L. Ryzhik, Bounds on the speed of propagation of the KPP fronts in acellular flow, Arch. Rational Mech. Anal. 184 (2007), 23–48.

185. O. A. Oleinik, Discontinuous solutions of the nonlinear differential equations, UspehiMat. Nauk (N.S.) 12(3) (1957) 3–73; AMS Translations, pp. 95–172, 1963.

186. H. Osada, Diffusion processes with generators of generalized divergence form, J. Math.Kyoto Univ 27(4)(1987), 597–619.

187. S. Osher, R. Fedkiw, Level Set Methods and Dynamic Implicit Surfaces, Applied MathSci, 153, Springer, New York, 2003.

188. S. Osher and J. Ralston, Stability of traveling waves in convective porous medium equa-tion, Comm. Pure Appl. Math. 35 (1982), 737–749.

189. G. Papanicolaou, Diffusion in Random Media, Surveys in Applied Mathematics, editedby J. B. Keller, D. McLaughlin, and G. Papanicolaou, Plenum Press, (1995), pp. 205–255.

190. G. Papanicolaou and S. Varadhan, Boundary value problems with rapidly oscillatingrandom coefficients, Colloq. Math. Soc. Janos Bolyai 27, Random Fields, Esztergom,Hungary (1979), Amsterdam, North Holland, 1982, pp. 835–873.

191. G. Papanicolaou and J. Xin, Reaction-diffusion fronts in periodically layered media, J.Stat. Physics 63(5/6) (1991), 915–931.

192. K. Park, I. M. Kim, Dynamics of an interface driven through random media: The effectof spatially correlated noise, J. Phys. Soc. Japan 72 (2003), 111–116.

Page 159: Surveys and Tutorials in the Applied Mathematical Sciences

References 153

193. N. Peters, A spectral closure for premixed turbulent combustion in the flamelet regime,J. Fluid Mech. 242 (1992), 611–629.

194. N. Peters, Turbulent Combustion, Cambridge University Press, Cambridge, 2000.195. J. R. Phillip, Theory of Infiltration, Adv. in Hydrosciences 5 (1969), 213–305.196. J. R. Phillip, Issues in Flow and Transport in Heterogeneous Porous Media, Transport

in Porous Media 1 (1986), 319–338.197. M. Postel, J. Xin, A numerical study of fronts in random media using a reactive solute

transport model, Computational Geoscience 1 (1997), 251–270.198. M. Protter and H. Weinberger, Maximum Principles in Differential Equations, Prentice-

Hall, 1967.199. J. Qian, Two approximations for Effective Hamiltonians Arising from Homogenization

of Hamilton-Jacobi Equations, CAM report 03-39, UCLA, August, 2003.200. F. Rezakhanlou, J. Tarver, Homogenization for Stochastic Hamilton-Jacobi Equations,

Arch. Rational Mech. Anal, 151 (2000), 277–309.201. F. Rezakhanlou, Central Limit Theorem for Stochastic Hamilton-Jacobi Eqautions,

Comm. Math. Phys. 211 (2000), 413–438.202. L. C. Rogers, D. Williams, Diffusion, Markov Processes and Martingales, Vol. 1, Foun-

dations, John Wiley and Sons, 1994.203. P. Ronney, Some Open Issues in Premixed Turbulent Combustion, Lecture Notes in

Physics, 449, Springer-Verlag, Berlin, 1995, pp. 1–22.204. P. D. Ronney, B. G. Haslam, and N. O. Rhys, Front propogation rates in randomly

stirred media, Phys. Rev. Lett. 74 (1995), 3804–3807.205. J.-M. Roquejoffre, Eventual monotonicity and convergence to traveling fronts for the

solutions of parabolic equations in cylinders, Ann Inst. H. Poincare, Analyse Nonlin-eare 14 (1997), 499–552.

206. S. Ross, Introduction to Probability Models, 9th edition, Academic Press, 2007.207. B. L. Rozdestvenskii, The Cauchy Problem for Quasilinear Equations, Dokl. Akad.

Nauk SSSR 115 (1957), pp. 454–457; AMS Translations, Series 2 42 (1964), 25–30.208. L. Ryzhik and A. Zlatos, KPP pulsating front speed-up by flows, Comm. Math Sci 5

(2007), 575–593.209. H. Rund, The Hamilton-Jacobi theory in the calculus of variations: its role in mathe-

matics and physics, Van Nostrand Co, London, 1966.210. Wm. van Saarloos, Front propagation into unstable states, Physics Reports 386 (2003),

29–222.211. B. Sandstede, Stability of N-fronts bifurcating from a twisted heteroclinic loop and an

application to the FitzHugh–Nagumo equation, SIAM. J. Math. Anal. 29 (1998), 183–207.

212. D. H. Sattinger, On the stability of waves of nonlinear parabolic systems, Advances inMath. 22 (1976), 312–355.

213. R. Schwab, Stochastic Homogenization of Hamilton-Jacobi Equations in Stationary Er-godic Spatial-Temporal Media, preprint (2008), Indiana Univ Math J, to appear.

214. S. Schumacher, Diffusions with random coefficients, in: Particle Systems, Random Me-dia and Large Deviations (R. Durrett Ed.), American Mathematical Society, Providence:1985.

215. L. Shen, J. Xin, A. Zhou, Finite Element Computation of KPP Front Speeds in RandomShear Flows in Cylinders, SIAM J. Multiscale Modeling and Simulation 7(3) (2008),1029–1041.

216. W. Shen, Traveling waves in time almost periodic structures governed by bistable non-linearities, I. stability and uniqueness, J. Diff. Eqn. 159 (1999), 1–55.

217. W. Shen, Traveling waves in time almost periodic structures governed by bistable non-linearities, II. existence, J. Diff. Eqn. 159 (1999), 55–101.

218. W. Shen, Traveling Waves in Diffusive Random Media, J. Dynamics Diff. Eqs. 16(4)(2004), 1011–1060.

Page 160: Surveys and Tutorials in the Applied Mathematical Sciences

154 References

219. W. Shen, Traveling waves in time dependent bistable media, Diff Int. Eqn. 19(3) (2006),241–278.

220. W. Shen, Variational Principle for Spatial Spreading Speeds and Generalized WaveSolutions in Time Almost Periodic and Space Periodic KPP Models, preprint, 2008.

221. N. Shigesada, K. Kawasaki and E. Teramoto, Traveling periodic waves in heterogeneousenvironments, Theor. Population Biology 30 (1986), 143–160.

222. S. Shy, R. Jang, P. Ronney, Laboratory Simulation of Flamelet and Distributed Modelsfor Premixed Turbulent Combustion Using Aqueous Autocatalytic Reactions, Combus-tion Science and Technology, 113-114 (1996), 329–340.

223. G. I. Sivashinsky, Instabilities, pattern formation, and turbulence in flames, Ann. Rev.Fluid Mech. 15 (1983), 179–199.

224. G. I. Sivashinsky, Cascade-renormalization theory of turbulent flame speed, Combust.Sci. Technology 62 (1988), 77–96.

225. M. E. Smaily, Pulsating traveling fronts: Asymptotics and homogenization regimes, Eu-ropean J. Appl. Math, to appear.

226. J. Smoller, Shock Waves and Reaction-Diffusion Equations, Graduate Text in Mathe-matics, Vol. 258, Springer-Verlag, 1983.

227. P. E. Souganidis, Stochastic homogenization of Hamilton-Jacobi equations and someapplications, Asymptotic Analysis 20 (1999), 1–11.

228. K. Uchiyama, The behavior of some nonlinear diffusion equations for large time, J.Math. Kyoto Univ. 18 (1978), 453–508.

229. S. R. S. Varadhan, Large Deviations and Applications, CBMS-NSF Regional Confer-ence Series in Applied Math., 46, SIAM, 1984.

230. S. R. S. Varadhan, N. Zygouras, Behavior of the solution of a random semilinear heatequation, Comm. Pure Appl. Math. 61 (2008), 1298–1329.

231. N. Vladimirova, P. Constantin, A. Kiselev, O. Ruchayskiy, L. Ryzhik, Flame enhance-ment and quenching in fluid flows, Combust. Theory Model. 7 (2003), 487–508.

232. A. I. Volpert, V. A. Volpert, Traveling wave solutions of parabolic systems, Translationsof Math Monographs, 140, American Math Soc, 1994.

233. J. Wehr, J. Xin, White Noise Perturbation of the Viscous Shock Fronts of the BurgersEquation, Commun. Math. Phys. 181 (1996), 183–203.

234. J. Wehr, J. Xin, Front speed in a Burgers equation with random flux, J. Stat. Phys.88(3/4) (1997), 843–871 .

235. J. Wehr, J. Xin, Scaling Limits of Waves in Convex Scalar Conservation Laws underRandom Initial Perturbations, J. Stat. Physics 122(2) (2006), 361–370.

236. H. Weinberger, On spreading speed and traveling waves for growth and migration mod-els in a periodic habitat, J. Math. Biol 45 (2002), 511–548.

237. G. B. Whitham, Linear and Nonlinear Waves, Wiley and Sons, 1979.238. F. Williams, Combustion Theory, Benjamin Cummings, Menlo Park, CA, second edi-

tion, 1985.239. W. Woyczynski, Burgers-KPZ Turbulence–Gottingen Lectures, Lecture Notes in Math-

ematics, 1700, Springer, 1998.240. J. Xin, Existence and stability of travelling waves in periodic media governed by a

bistable nonlinearity, J. Dynamics Diffl. Eqs. 3 (1991), 541–573.241. J. Xin, Existence and uniqueness of travelling wave solutions in a reaction-diffusion

equation with combustion nonlinearity, Indiana Math. J. 40 (1991), 985–1008.242. J. Xin, Existence of planar flame fronts in convective-diffusive periodic media, Arch.

Rat. Mech. Anal. 121 (1992), 205–233.243. J. Xin, Existence and nonexistence of traveling waves and reaction-diffusion front prop-

agation in periodic media, J. Stat. Phys. 73 (1993), 893–926.244. J. Xin, Existence of multidimensional traveling waves in transport of reactive solutes

through periodic porous media, Arch. Rat. Mech. Anal. 128 (1994), 75–103.245. J. Xin, Stability of traveling waves in a solute transport equation, J. Diffl. Eqs., 135

(1997), 269–298.

Page 161: Surveys and Tutorials in the Applied Mathematical Sciences

References 155

246. J. Xin, Front propagation in heterogeneous media, SIAM Review, 42(2) (2000), 161–230.

247. J. Xin, KPP front speeds in random shears and the parabolic Anderson problem, Meth-ods and Applications of Analysis, 10(2) (2003), 191–198.

248. J. Xin, A. Peirce, J. Chadam, P. Ortoleva, Reactive flows in layered porous media, I.Homogenization of free boundary problems, Asymptotic Analysis 11 (1995), 31–54.

249. J. Xin, A. Peirce, J. Chadam, P. Ortoleva, Reactive flows in layered porous media, II. TheShape Stability of the Reaction Interface, SIAM J. Applied Math, 53/2 (1993), 319–339.

250. J. Xin, J. Zhu, Quenching and propagation of bistable reaction-diffusion fronts in mul-tidimensional periodic media, Physica D 81 (1995), 94–110.

251. V. Yakhot, Propagation velocity of premixed turbulent flames, Comb. Sci. Tech. 60(1988), 191–241.

252. X. Yuan, T. Teramoto, Y. Nishiura, Heterogeneity-induced defect bifurcation and pulsedynamics for a three-component reaction-diffusion system, Physical Review E, 75,036220 (2007), 1–12.

253. S. van der Zee and W. H. van Riemsdijk, Transport of reactive solute in spatially vari-able soil systems, Water Resources Research 23 (1987), 2059–2069.

254. E. Zeidler, Nonlinear Functional Analysis, Springer-Verlag, 1984.255. A. Zlatos, Pulsating front speed-up and quenching of reaction by fast advection, Non-

linearity 20 (2007), 2907–2921.256. A. Zlatos, Sharp asymptotics for KPP pulsating front speed-up and diffusion enhance-

ment by flows, to appear in Arch. Ration. Mech. Anal, 2009.

Page 162: Surveys and Tutorials in the Applied Mathematical Sciences

Index

Action function, 91Action integral, 78Affine data, 78Almost periodic media, 47Ansatz

homogenization, 42Wentzel–Kramers–Brillouin, 43

Asymptotic front speed, 40Asymptotic reduction, 59Asymptotic stability, 10

Biology, 1Boundedness, 83

one-sided bounded potential, 83Brownian motion, 18, 120

parabolic shift, 58Burgers equation, 2, 6

Burgers solution, 66random Burgers front, 55random flux, 53viscous, 56

Central limit theorem, 15, 53, 55, 63, 86, 137Chemical kinetics, 1, 8Coefficient

diffusion, 17drift, 17homogenized, 25periodic, 25

Combustion, 1Conservation law, 1, 53

random flux, 53viscous, 5

Continuation method, 34, 45Contraction mapping principle, 34Convergence

almost sure, 14, 72, 136

in law, 15, 55, 59in probability, 15, 59, 67mean square, 15

Covariance function, 16, 108Critical fronts, 10, 11

Degree theory, 10, 31Density

function, 14transition, 17

Deposition process, 1Double exponential distribution, 82

Ensemble-averaged front speeds, 93Ergodicity, 71, 108, 119, 124, 132Eulerian method, 123Expectation, 14

Feynman–Kac formula, 28, 36, 113, 130Fisher, 27FitzHugh–Nagumo system, 8, 140Flows

cellular, 49compressible, 117gradient, 83incompressible, 49, 123, 129random, 83shear, 48, 93, 113space-time random, 93

Fourier transform, 17Fractal shape, 4Front, 53

acceleration, 83anomalous behavior, 70divergence in shear flows, 84fluctuations, 53generalized, 137

157

Page 163: Surveys and Tutorials in the Applied Mathematical Sciences

158 Index

Huygens, 141location, 1monotonicity, 33motion, 19probing, 19, 63profile, 1, 7, 10propagation, 1pulsating, 27random, 19shape, 122speed, 1, 7, 122speed bending, 116, 132speed bounds, 114speed distribution, 106speed enhancement, 50speed reduction, 119speed selection, 5spreading, 108, 124trapping, 120, 143width, 19, 122

G-equation, 2, 142generalized, 143

Geometric models, 141Girsanov formula, 130Growth

linear, 7, 105, 115sublinear, 16, 49, 116superlinear, 7

Hamilton–Jacobi equations, 1, 20, 42, 43, 69stochastic homogenization, 70, 132, 135viscous, 41, 132

Hamiltonian, 5, 69, 125coercive, 71convex, 6, 71, 134homogenized, 41, 50random, 70unbounded, 78

Heterogeneous media, 1Hitting time, 121Hodgkin and Huxley, 27Homogenization, 4, 23, 70

breakdown, 78cell problem, 25, 41, 42periodic, 28, 41running maximum and divergence, 83stochastic, 69, 126, 132

Hopfformula, 7, 91solution, 7

Hopf–Cole formula, 6, 56Huxley formula, 10Hyperbolic scaling, 19, 69

Invarianceprinciple, 54, 59space-time translation, 45translational, 11

Kardar–Parisi–Zhang equation, 2Kolmogorov–Petrovsky–Piskunov

equation, 1front, 35, 39, 47, 93, 123system, 140

Krylov–Safonov–Harnack inequality, 126

Lagrangian, 7, 78, 125, 126convex, 6homogenized, 72Newtonian mechanics, 44relativistic mechanics, 44

Langevin equation, 18Laplace method, 19, 58, 67Large deviation, 28, 36, 43, 108, 112, 119, 129

rate function, 36, 113Law of large numbers, 16, 55Lax–Oleinik formula, 7, 71Legendre transform, 6, 43, 71, 134Level set, 2, 6, 141Limit

homogenization, 83, 112interacting particle system, 122inviscid, 7passage of, 72

Marginal stability criterion, 28Maximum principles, 11, 30, 51Minimal speed, 11, 141Moment condition, 118, 124Multiple scales, 4Multiscale media, 44

Noisecolored, 20white, 16

Nonlinear eigenvalue problem, 41Normal velocity, 2

Parabolic Anderson problem, 108, 116Periodic media, 23, 43Perturbation, 6

decaying, 55random, 55

Phase plane, 12Phase variable, 43Porous media, 1Principal

eigenvalue, 42, 94

Page 164: Surveys and Tutorials in the Applied Mathematical Sciences

Index 159

Lyapunov exponent, 108, 116Probability

conditional, 17function, 13space, 13transition, 17

Random media, 4, 28, 53, 70, 73extreme behavior, 70front speeds in, 83unbounded, 85, 136

Random process, 2diffusion, 17, 119, 133Gaussian, 16Markov, 17mixing, 55Ornstein–Uhlenbeck, 16, 80, 100running maximum, 81stationary, 54Wiener, 16, 36, 61

Random variables, 13Gaussian, 14independent, 14

Rankine–Hugoniot condition, 56Reaction nonlinearity, 5, 29, 137Reaction–diffusion

advection equations, 3, 123equations, 1, 7

Regularity, 54, 124

Scienceboundaries, 27communities, 27environmental, 1, 53interdisciplinary, 27

Sliding doman method, 33Spreading rate, 4, 125Stationarity, 54, 108, 119, 124, 132Stopping time, 37, 113Subadditive ergodic theorem, 73, 126, 135

Temporal decorrelation, 108, 132Time scale, 1, 3Traveling fronts, 1, 4, 5, 7, 11, 27, 43, 47

random, 137Turbulent

combustion, 2, 123, 125front speed, 28, 123, 125, 143

Variance, 14Variational

formula, 70, 133inequality, 38min-max principles, 11, 50, 135principles, 11, 48, 103speed formula, 12, 28, 40, 94

Viscosity solutions, 39, 72