81
DRAFT July 11, 2016 DRAFT Two results on Asymptotic Behaviour of Random Walks in Random Environment by Jeremy Voltz A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Mathematics University of Toronto c Copyright 2016 by Jeremy Voltz DRAFT July 11, 2016 DRAFT

University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Two results on Asymptotic Behaviour of Random Walks in RandomEnvironment

by

Jeremy Voltz

A thesis submitted in conformity with the requirementsfor the degree of Doctor of PhilosophyGraduate Department of Mathematics

University of Toronto

c© Copyright 2016 by Jeremy Voltz

DRAFT

July 11, 2016

DRAFT

Page 2: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Abstract

Two results on Asymptotic Behaviour of Random Walks in Random Environment

Jeremy Voltz

Doctor of Philosophy

Graduate Department of Mathematics

University of Toronto

2016

In the first chapter of this thesis, we consider a model of directed polymer in 1 + 1 dimensions in

a product-type random environment ω(t, x) = b(t)F (x), where the fields F and b are i.i.d., with F (x)

continuous, symmetric and bounded, and b(t) = ±1 with probabilty 1/2. Thus ω can be viewed as

the field F oscillating in time. We consider directed last-passage percolation through this random field;

namely, we investigate the behavior of the length n polymer path with maximal action, where the action

of a path is simply the sum of the environment variables it moves through.

We prove a law of large numbers for the maximal action of the path from the origin to a fixed

endpoint (n, bαnc), and investigate the limiting shape function a(α). We prove that this shape function

is non-linear, and has a corner at α = 0, thus indicating that this model does not belong to the KPZ

universality class. We conjecture that this shape function has a linear piece near α = 0.

With probability tending to 1, the maximizing path with free endpoint will localize on an edge with

F values far from each other. Under an assumption on the arrival time to this localization site, we prove

that the path endpoint and the centered action of the path, both rescaled by n−2/3, converge jointly to

a universal law, given by the maximizer and value of a functional on a Poisson point process.

In the second chapter, we consider a class of multidimensional random walks in random environment,

where the environment is of the type p0 + γξ, with p0 a deterministic, homogeneous environment with

underlying drift, and ξ an i.i.d. random perturbation. Such environments were considered by Sabot in

[30], who finds a third-order expansion in the perturbation for the non-null velocity (which is guaranteed

to exist by Sznitman and Zerner’s LLN [34]). We prove that this velocity is an analytic function of the

perturbation, by applying perturbation theory techniques to the Markov operator for a certain chain in

the space of environments.

DRAFT

ii

DRAFT

Page 3: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

To Padre.

DRAFT

iii

DRAFT

Page 4: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Acknowledgements

To my advisor Kostya Kahnin, I want to thank you for the time and energy you gave me. Ever the

optimist, you kept me motivated and helped me work through countless difficulties. You were generous

with your time and your ideas. Your intuition was invaluable to my understanding, and you always

inspired confidence. It has been a pleasure working with you.

I owe the second chapter of this thesis to Dima Dolgopyat, who I was very fortunate to work with.

Visiting you in Maryland was a great experience, and I learned a great deal. You offered so much of

your time and guidance, and were always willing to think through things and explain difficult concepts.

It was inspiring to have the opportunity to work with you.

The probability graduate course, taught by the members of my thesis committee Jeremy Quastel and

Balint Virag, inspired me to study the field. Thank you both for your support and discussion over the

years. The probability retreats were great fun, and inspired a sense of camaraderie in the department

that I was lucky to be a part of. I first spent time with Almut Burchard at such a retreat, and it was

such a pleasure expanding the circle of probability. I am thankful to her for attending my defense.

Many thanks to Jemima Merisca for guiding me through the graduation process with aplomb, taking

care of so much behind the scenes. Ever helpful, ever understanding, you made things as smooth as can

be.

Ida Bulat was my first introduction to the department, as she was for many graduate students.

She was a pure spirit of kindness and compassion, and made coming to U of T a warm and inviting

experience. She will be missed.

To my fellow graduate student friends in the department, thank you for the companionship and fun.

Special thanks to Jordan Bell and Andrew Stewart for giving your time to help towards the end.

Last year I had to leave my thesis to take care of my mother, and the year after her death was

filled with legal and emotional turmoil. Coming back to research was very difficult. I owe so much

to the incredible emotional support of my close friends (shumps), and the exceptional characters of

Countermeasure. I also want to thank my father and my sister Jennifer for sharing the weight, and

being family. Jennifer, I could never have done what you did. Thank you for all the hard work.

I owe special thanks to Tara Park, for pushing me to dive back in, meeting me at the library every

day for months, giving me the drive to finish. I couldn’t have done this without you.

And most importantly, I owe everything to my wife Amie Everett. I love you.

DRAFT

iv

DRAFT

Page 5: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Contents

1 A Zero Temperature Directed Polymer in a Product-Type Environment 1

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Notation and Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2.1 The Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.3 The shape function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.3.1 Existence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.3.2 Continuity of Φ on ∂K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.3.3 Properties of the Shape Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.4 The maximizing path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1.4.1 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

1.4.2 The ending edge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.3 The arrival time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

1.5 Distributional Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

1.6 Non-linearity of the Shape Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

1.7 Proof of Theorem 1.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

1.7.1 Basics of Poisson Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

1.7.2 Proof of Theorem 1.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

1.8 Questions / Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

1.9 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

1.9.1 Probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2 Analyticity of the Effective Velocity for Ballistic RWRE 45

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

2.2 Background and Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.2.1 The Configuration Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

DRAFT

v

DRAFT

Page 6: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

2.2.2 The Configuration Markov Chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2.3 The Operator on a Complex Banach Space . . . . . . . . . . . . . . . . . . . . . . . . . . 52

2.3.1 The metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

2.3.2 The Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

2.4 Proof of Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

2.4.1 Quasi-compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

2.4.2 Differentiability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

2.4.3 The Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

2.5 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

2.5.1 Perturbation Theory for Quasi-compact Operators on a Banach Space . . . . . . . 70

Bibliography 73

DRAFT

vi

DRAFT

Page 7: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Chapter 1

A Zero Temperature Directed

Polymer in a Product-Type

Environment

1.1 Introduction

A directed polymer in 1 + d dimensions is a chain of length n in the lattice N0 × Zd which is directed

in the first coordinate. That is, it can be modeled by a graph (t, γ(t))nt=0, where γ(t) is a nearest

neighbor path in Zd (starting at the origin).

For each point in the lattice (t, x), we assign a random variable ω(t, x), and a polymer measure

on paths which inherits its randomness from the random field. This polymer measure is canonically

constructed as a Gibbs measure, by assigning to paths of length n a random action depending on the

random field ω,

Aωn(γ) =n∑i=1

ω(i, γ(i))

and then defining the probability measure

Pωn,β(γ) =

1

Zωn (β)exp(βAωn(γ)

)P(γ),

where P(·) is the uniform measure on nearest neighbor paths of length n in Zd, where β ≥ 0 is the

inverse temperature of the system, and where Zωn,β is the partition function, the normalizing constant

DRAFT

1

DRAFT

Page 8: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

such that Pωn,β(·) is a probability measure. That is,

Zωn,β = P(

exp(βAωn(γ)

)).

The random environment ω is distributed by a probability measure P on the space of environments,

Ω. Often this measure is taken to be product measure on the lattice, so that the random variables ω(t, x)

are independent and identically distributed with some distribution, usually taken to be mean zero with

all finite exponential moments.

For this system, the real number β controls how strongly the environment influences the behavior of

the polymer. When β = 0 (temperature is infinite), the polymer measure is just the uniform measure and

the environment is ignored. When β →∞ (temperature is zero), the polymer measure concentrates on

the path with maximum action, γn. This model is called the (oriented) last passage percolation model,

or the corner growth model (usually rotating the lattice 45 degrees and considering the maximizing

up-right path).

In this β →∞ regime (and others), much work is done to characterize the exponents ζ and χ, which

describe the order of the maximizing path’s endpoint and the order of the fluctuations of the path action,

respectively. That is, calling γn = argmaxγ Aωn(γ),

∣∣γn(n)∣∣ ∼ nζ

and ∣∣Aωn(γn)− E(Aωn(γn))∣∣ ∼ nχ

for typical environments ω. (For the case of finite β, one considers the fluctuations of log(Zωn,β

)instead

of maxγ Aωn(γ).)

It is conjectured that for β → ∞, for each dimension d ≥ 1, these exponents ζ and χ are universal,

that is, they do not depend on the distribution of ω(t, x), as long as they are i.i.d. In dimension 1,

values for these exponents have been calculated for some specific distributions of ω(t, x). Johannson

in [19] considers exponentially and geometrically distributed ω(t, x), and proves that χ = 1/3 with

Tracy-Widom law fluctuations. For a Poissonized version of the corner growth model in dimension 1,

first studied by Hammersley in [13], Baik, Deift, and Johansson show χ = 1/3 with Tracy-Widom law

fluctuations, and ζ = 2/3 is proven by Johansson in [20]. Thus, this model is said to satisfy the so called

“KPZ relation”,

χ = 2ζ − 1,

DRAFT

July 11, 2016

DRAFT

Page 9: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.1. Introduction 3

DRAFT

a somewhat mysterious but often conjectured relation in many random growth models.

Recently, Chatterjee [5] proved the “KPZ relation” for first passage percolation on the lattice Zd for all

d ≥ 2, which considers the (self-avoiding) path which minimizes the Hamiltonian given above for positive

ω(t, x), assuming the exponents χ and ζ exist in a certain sense. In [1], Auffinger and Damron make

simplifications and weaken Chatterjee’s hypotheses, allowing his results to apply to directed polymers

with finite β (see [2]).

In the following, the paths we consider γ are “lazy” in Z (d = 1), so that∣∣γ(i)− γ(i+ 1)

∣∣ ≤ 1, and

we consider a product-type environment,

ω(t, x) = F (x) · b(t),

where F (x) are i.i.d. random variables with continuous, symmetric distribution supported on an

interval [−c, c], and where b(t) are i.i.d. −1 or 1 with probability 1/2, independent of F (x). The

field b(t) can be thought of as an oscillating field. Note that in this model, first-passage (minimal

action) and last-passage (maximal action) are distributionally equivalent.

We consider paths from the origin to a point (t, x), and call the the maximal action

A(t, x) = maxγ:(0,0)→(t,x)

Aωt (γ).

Using standard techniques, we can prove a law of large numbers, so that n−1A(n, [αn]) converges to a

shape function a(α) P-a.s, α ∈ [−1, 1], with the following properties.

Theorem 1.1.1.

1. The function a(α) is a continuous, concave function on [−1, 1], symmetric across 0,

2. a(0) = c,

3. a(α) has a corner at 0, i.e. a′(0−) = −a′(0+) > 0,

4. For any α ∈ (0, 1), a(α) > c(1− α). Namely, a(α) is non-linear on [0, 1].

We also conjecture that

Conjecture 1.1.2. a(α) has a linear piece near the origin.

Note that the corner at 0 heuristically rules out KPZ behavior, and implies informally that χ = ζ

(see [2]), which is the behavior we conjecture.

DRAFT

July 11, 2016

DRAFT

Page 10: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

4 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

To explore the maximizing path with a free endpoint, γn, we note that a path γ can accumulate a

large action A(γ) =∑ni=1 ω(i, γ(i)) by localizing on an edge x, x+ 1 where F (x) and F (x+ 1) are far

away from each other (near the boundary of the support of F ). That is, by localizing on edges with small

“discrepancy”, dx = 2c −∣∣F (x+ 1)− F (x)

∣∣. The locations of discrepancies and their values, scaled by

n−2/3 and n1/3 respectively, converge to a Poisson point process.

Theorem 1.1.3. For each n ∈ N, we introduce the rescaled point process µn as a random measure:

µn =∑k∈Z

δ(n−2/3k,n1/3dk)

As n → ∞, the distribution of µn converges weakly in the vague topology to the distibution of µ, the

Poisson process with driving measure ν absolutely continuous w.r.t. Lebesgue measure with density given

by:

dν(dt× dy)

dt× dy=

p′(0+)

2 y, y > 0

0, y ≤ 0.

,

where p is the density of d0.

For the free endpoint maximizing path, we show that the endpoint location `n = γn(n) is order n2/3,

and that the path localizes on an edge with small discrepancy, a record value, of order n−1/3. Calling

τn the arrival time of γn to this ending edge `n, we show that with high probability, the path γn does

not leave `n. We conjecture the following, and assume it for our final result.

Assumption 1.1.4 (Bounded Arrival Condition). There exists a κ > 1 such that

P(τn > κ|`n|) −→ 0.(1.1)

Note that this assumption implies conjecture 1.1.2, that the shape function is linear near the origin.

With this assumption, after normalizing by n2/3, we are able to prove that the probability distributions

of the centered maximal action A(γn)− cn, and the endpoint `n, converge to a universal limit law.

Theorem 1.1.5. Under assumption 1.1.4,

(1.2)

(`nn2/3

,A(γn)− cn

n2/3

)d−→ (X, g(X,Y )),

DRAFT

July 11, 2016

DRAFT

Page 11: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.2. Notation and Preliminaries 5

DRAFT

where (X,Y ) = argmax(X,Y )∈supp(µ) g(X,Y ), where µ is the PPP described in theorem 1.1.3, and where

g(x, y) = −a′(0+)|x| − y/2.

In section 1.3 we prove the existence of the shape function and items (i)-(iii) of theorem 1.1.1.

The proof of (iv) is given in section 1.6. In section 1.4 we prove that the path localizes on a record

discrepancy with probability tending to 1. In section 1.5 we prove theorem 1.1.5, and section 1.7 is the

proof of theorem 1.1.3.

1.2 Notation and Preliminaries

We will consider environments ω(t, x), t ∈ N, x ∈ Z of the form

ω(t, x) = F (x)b(t),

whereF (x)

∣∣x ∈ Z

are i.i.d with continuous distribution fdm, m Lebesgue measure, supp(f) ⊂ [−c, c]

for some c > 0, and whereb(t)

∣∣ t ∈ N

are i.i.d. with common distribution uniform on −1,+1. The

two fields F and b are taken to be statistically independent. We denote by P the probability distribution

of ω just described on the space Ω = [−c, c]N×Z of environments, endowed with the product σ-algebra.

We require certain conditions on the density f of F , namely, that f is symmetric across 0 with

compact support on a bounded interval [−c, c], positive on the whole interval, and C1, with one sided

derivatives at c and −c. We let F · b denote the whole field of products, ω.

Since we will be concerned with paths through such environments, let us define the space of directed

nearest-neighbor (lazy) paths:

Γ((t1, x1), (t2, x2)

)=γ : [t1, t2] ∩ N0 → Z

∣∣∣ ∣∣γ(i)− γ(i+ 1)∣∣ ≤ 1(1.3)

for i = t1, . . . , t2 − 1, γ(t1) = x1, γ(t2) = x2

.

In the special case that the starting point is the origin, we write

Γ(t, x) = Γ((0, 0), (t, x)

).(1.4)

DRAFT

July 11, 2016

DRAFT

Page 12: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

6 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Leaving any of the parameters as a dot, ·, means that parameter is free to vary. Given any path γ

defined on [a, b] ∩ N0 and a product type environment ω = F · b, we define the action of that path by

the random variable

(1.5) Aω(γ, a, b) =b∑

i=a+1

F (γ(i)) · b(i).

If the domain of γ is clear, we simply write Aω(γ). We are interested in describing the path which

maximizes this action on [0, n]. We denote the maximizing path of length n with free endpoint (which

is unique with probability 1, since ω(t, x) has continuous distribution) by

(1.6) γωn = argmaxγ∈Γ(n,·)

Aω(γ).

Remark 1.2.1. A simple observation is that for any ω = F · b ∈ Ω, the path which maximizes the

above action minimizes the action for ω = F · −b. And since b and −b have the same distribution, the

minimum action is equal in distribution to minus the maximum action.

Denote the maximum action between two points (t1, x1) and (t2, x2) by

Aω(

(t1, x1), (t2, x2))

= maxAω(γ)

∣∣∣ γ ∈ Γ((t1, x1), (t2, x2)

).

When clear, we will usually omit the ω from the above notation.

Similarly to the path space, in the special case that the starting point is the origin, we write

A(t, x) = A((0, 0), (t, x)

).(1.7)

We will be considering the action of “pinned paths”, whose endpoint we fix as a constant multiple

α ∈ [−1, 1] of the length. We obviously have to round values to the nearest integer, and we choose the

floor for positive α. For the model to be symmetric across 0, we take the ceiling for negative α. So we

denote the symmetric rounding function

[x] =

bxc if x ≥ 0

dxe = −b−xc if x < 0

.

DRAFT

July 11, 2016

DRAFT

Page 13: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.2. Notation and Preliminaries 7

DRAFT

1.2.1 The Environment

We note an important observation about the optimal path through an environment ω = F · b.

Remark 1.2.2. If two adjacent values, F (x) and F (x+ 1) are such that one of them is close to c and

the other is close to −c (the boundaries of the support of f), then a path can do very well in the long

run at these two sites, by staying at the c-valued site for all times i where b(i) = +1, and then jumping

to the −c-valued site for all times when b(i) = −1.

It will be important to measure such sites, and denote the optimal path restricted to an edge. For

this we define the discrepancy of an edge as the sum of the distances from each point to the boundary

points of the support of f , −c and c. The smaller this value, the higher the contribution to the action

of a path which stays (optimally) on this edge.

Definition 1.2.3. • For each x ∈ Z, define the discrepancy of the edge x, x+ 1 to be

(1.8) dx = 2c−∣∣F (x+ 1)− F (x)

∣∣

• Define the optimal path restricted to the edge x, x+ 1 by

(1.9) ηx(t) =

x if F (x) ≥ F (x+ 1) and b(t) = 1

x if F (x) ≤ F (x+ 1) and b(t) = −1

x+ 1 otherwise

Note that if dx < c, then necessarily F (x) and F (x + 1) have opposite sign. Since we will focus on

edges with small discrepancies, we will usually be looking at sites where this holds. In this case, we have∣∣F (x)∣∣ +

∣∣F (x+ 1)∣∣ =

∣∣F (x)− F (x+ 1)∣∣ = 2c − dx. With this in mind, we obtain the following useful

inequalities.

2 min(∣∣F (x)

∣∣, ∣∣F (+1)∣∣) ≤ 2c− dx ≤ 2 max

(∣∣F (x)∣∣, ∣∣F (x+ 1)

∣∣),and

c− dx =∣∣F (x)

∣∣+∣∣F (x+ 1)

∣∣− c ≤ min(∣∣F (x)

∣∣, ∣∣F (x+ 1)∣∣).

(1.10)

DRAFT

July 11, 2016

DRAFT

Page 14: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

8 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Using these, we also obtain (assuming opposite sign) that

∣∣F (x) + F (x+ 1)∣∣ = max

(∣∣F (x)∣∣, ∣∣F (x+ 1)

∣∣)−min(∣∣F (x)

∣∣, ∣∣F (x+ 1)∣∣)

=∣∣F (x)

∣∣+∣∣F (x+ 1)

∣∣− 2 min(∣∣F (x)

∣∣, ∣∣F (x+ 1)∣∣)

≤ 2c− dx − (2c− 2dx)

= dx.(1.11)

We will often use the naive bound on the the action of ηx using 1.10,

A(ηx, a, b) ≥ (b− a) min(∣∣F (x)

∣∣, ∣∣F (x+ 1)∣∣)(1.12)

≥ (b− a)(c− dx).(1.13)

For more precise control of the action of ηx, we denote

N+(a, b) = #∣∣∣i ∈ [a+ 1, b]

∣∣ b(i) = +1∣∣∣, and

N−(a, b) = #∣∣∣i ∈ [a+ 1, b]

∣∣ b(i) = −1∣∣∣,(1.14)

and then compute the action of ηx between a and b to be

A(ηx, a, b) = max(F (x), F (x+ 1)

)·N+(a, b)−min

(F (x), F (x+ 1)

)·N−(a, b)

= (b− a)

(c− 1

2dx

)+

(N+(a, b)− b− a

2

)(F (x) + F (x+ 1)

).(1.15)

And as we would expect, the contribution on a large time scale T coming from a path staying on the

edge x, x+ 1 will be T(c− dx

2

), plus fluctuations of order

√T .

The following corollary of theorem 1.1.3 guarantees the existence of good “candidate sites” for the

best path, while also guaranteeing that such points are well separated with high probability. Recall from

theorem 1.1.3 that µn =∑k∈Z δ(n−2/3k,n1/3dk).

Corollary 1.2.4. For any ε > 0 and any a > 0, there are positive constants N , b, and δ such that for

all n > N ,

P(µn(R) ≥ 1 and µn

(Rδ(x, y)

)≤ 1 for all (x, y) ∈ R

)> 1− ε,(1.16)

DRAFT

July 11, 2016

DRAFT

Page 15: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.2. Notation and Preliminaries 9

DRAFT

where R = (−a, a)× (0, b) and Rδ(x, y) = (x− δ, x+ δ)× (y − δ, y + δ) ∩R .

Proof. First we prove this fact for the limiting Poisson point process µ with driving measure ν described

in theorem 1.1.3. Given a, we can find b such that P(µ(R) ≥ 1) > 1−ε, which is clear from the density of

ν. Inside this event, we condition on the number of points in the rectangle, NR ≥ 1. Then the points are

independent and identically distributed in R with distribution ν(·)/ν(R) (see [26] for basics of poisson

processes). Thus for any two such points, the probability of both of them being within a given rectangle

Rδ ⊂ R is less than or equal to(ν(Rδ)ν(R)

)2

. And thus the probability of any two points being within a

given rectangle is bounded by NR(NR − 1)ν(Rδ)2/ν(R)2.

Now, for each point (j δ2 , kδ2 ) ∈ R, j, k ∈ Z, center a square Rj,kδ := Rδ(j

δ2 , k

δ2 ). This covers R

such that if all Rj,kδ contain no more than 1 point, then no two points are within any Rδ(x, y) for any

(x, y) ∈ R. Then using a union bound, we have

P(µ(Rδ(x, y)) ≥ 2 for some (x, y) ∈ R

)= P

(µ(Rj,kδ ) ≥ 2 for some

(jδ

2, kδ

2

)∈ R

)

≤ #

∣∣∣∣δ2Z2 ∩R∣∣∣∣E[NR(NR − 1)

]max

(x,y)∈R

(ν(Rδ(x, y))

ν(R)

)2

.

For fixed a, b > 0, #∣∣∣ δ2Z2 ∩R

∣∣∣ is a constant times δ−2, and since the density of ν is bounded in R,

max(x,y)∈R ν(Rδ(x, y)) is a constant times δ2. Thus we can make the probability arbitrarily small by

taking δ small.

Finally, by Theorem 4.2 in [22], vague convergence of µn to µ implies that the joint distribution ofµn

(Rj,kδ

)(j δ2 ,k

δ2 )∈R

converges to the joint distribution of

µ(Rj,kδ

)(j δ2 ,k

δ2 )∈R

.

Remark 1.2.5. Note that the driving measure ν in theorem 1.1.3 is stationary in the first dimension,

so that we can take a more general rectangle R = [a1, a2]× [0, b] in the corollary, and b only depends on

the length a2 − a1.

DRAFT

July 11, 2016

DRAFT

Page 16: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

10 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

1.3 The shape function

In this section we show the basic law of large numbers result for the minimal action along a given

direction. Namely, we show the existence of a deterministic function in the plane

Φ(x, y) = limn→∞

n−1A([nx], [ny]

).

This result, along with several nice properties of the limiting function, follows a classical subadditivity

argument. We follow closely the argument in [28] and [31]. We will then investigate the notable properties

of the one dimensional shape function mentioned in section 1.2, a(α) = Φ(1, α).

1.3.1 Existence

Define the cone

K =

(x, y) ∈ R2∣∣∣ 0 ≤ |y| ≤ x.

Theorem 1.3.1. For the environment distribution P described in section 1.2, there exists a deterministic

function Φ : K → [0, c] such that

Φ(x, y) = limn→∞

n−1A([nx], [ny]

)P-almost surely.(1.17)

For (x1, y1) and (x2, y2) in K, the function Φ is

• Super-additive:

Φ(x1, y1) + Φ((x1, y1), (x1 + x2, y1 + y2)

)≤ Φ(x1 + x2, y1 + y2)

• Concave:

tΦ(x1, y1) + (1− t)Φ(x2, y2) ≤ Φ(tx1 + (1− t)x2, ty1 + (1− t)y2)

• Homogenous:

Φ(x, y) = k−1Φ(kx, ky) for any k > 0.

Φ is also continuous on K and symmetric in y. Quantitatively, Φ(x,±x) = 0 and Φ(x, 0) = c for all

DRAFT

July 11, 2016

DRAFT

Page 17: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.3. The shape function 11

DRAFT

x ∈ (0,∞).

Proof. The basis of the proof is the superadditivity of the path actions, which is clear from construction,

for points (x1, y1), (x2, y2) ∈ K ∩ Z2:

A(x1, y1) +A((x1, y1), (x1 + x2, y1 + y2)

)≤ A(x1 + x2, y1 + y2).(1.18)

From the construction of the model, we have the following bound. For points (x1, y1), (x2, y2) ∈

K ∩ Z2, we have

∣∣∣A((x1, y1), (x1 + x2, y1 + y2))∣∣∣ ≤ cx2,(1.19)

and so superadditivity gives the following, which we will use throughout the proof instead of the mono-

tonicity present in the corner growth model. For (x1, y1), (x2, y2) ∈ K ∩ Z2,

A(x1, y1) ≤ A(x1 + x2, y1 + y2) + cx2.(1.20)

Following Lin Hao’s dissertation, [28], and Timo Seppalainen’s lecture notes on the corner growth

model, [31], we first consider integer points, then extend to rational points, and finally all real numbers.

First, suppose that (x, y) ∈ K ∩ Z2. Define Xm,n = A((mx,my), (nx, ny)

)for 0 ≤ m < n. Applying

Kingman’s subadditive ergodic theorem 1.9.3 to −Xm,n, we have that

n−1X0,n = n−1A(xn, yn)

tends almost surely to a limit, Φ(x, y), for any (x, y) ∈ K∩Z2. For any k ∈ N, homogeneity is immediate.

For superadditivity, we look at

n−1A(nx1, ny1) + n−1A((nx1, ny1), (nx1 + nx2, ny1 + ny2)

)≤ n−1A(nx1 + nx2, ny1 + ny2).(1.21)

The first and last terms converge almost surely to Φ(x1, y1) and Φ(x1 + x2, y1 + y2) respectively. The

middle term is equal in distribution to A(nx2, ny2), and thus converges almost surely to Φ(x2, y2)

along a subsequence (lemma 1.9.2). Taking the limit of this inequality along this subsequence gives

superadditivity of the shape function Φ on integer points in K.

DRAFT

July 11, 2016

DRAFT

Page 18: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

12 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

We also have that for any (x, y) ∈ K ∩ Z2,

Φ(x,±x) = 0,(1.22)

0 ≤ Φ(x, y) ≤ cx,(1.23)

and so (1.23) and superadditivity gives a form of monotonicity for (x1, y1), (x2, y2) ∈ K ∩ Z2 for Φ:

Φ(x1, y1) ≤ Φ(x1 + x2, y1 + y2).(1.24)

Non-negativity of the shape function in (1.23) follows from the fact that Φ(x, y) = limn→∞ n−1EX0,n,

and by noting that given an arbitrary path from the origin to (nx, ny) ∈ K ∩ Z2, the expectation of the

action of the path is 0, implying that the max action over all such paths, X0,n, has expectation greater

than or equal to 0.

Remark 1.3.2. In fact, we can see that (1.22) holds on all of ∂K, not just at integer points. The

existence of the limit (1.17) is immediate for (x, y) ∈ ∂K, since for (x, y) 6= (0, 0), the limit on the

right hand side is taken along a (non-random) subsequence of either n−1A(n, n) or n−1A(n,−n), both

of which converge to 0 almost surely by the law of large numbers Thus Φ is identically 0 on ∂K.

Now, suppose (x, y) ∈ K ∩ Q2. To extend the proof to rational points, we choose q ∈ N such that

(qx, qy) ∈ Z2, and define

Φ(x, y) = q−1Φ(qx, qy).

This definition is independent of the choice of q by homogeneity for integers. With this definition, we

have homogeneity for rational k > 0, superadditivity for points in K ∩Q2, as well as (1.22), (1.23), and

(1.24) for points in K ∩Q2.

To show equation (1.17) for rational points (x, y) ∈ K ∩ Q2, given any n ∈ N, write n = mq + r,

r ∈ 0, . . . , q − 1. Call Rx = [nx] − mqx = [rx] and Ry = [ny] − mqy = [ry]. Then∣∣Ry∣∣ ≤ Rx for

(x, y) ∈ K ∩Q2, and thus (Rx, Ry) ∈ K ∩ Z2.

Then by (1.20), we have

A([nx], [ny]) ≥ A([nx]−Rx, [ny]−Ry)− cRx,(1.25)

DRAFT

July 11, 2016

DRAFT

Page 19: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.3. The shape function 13

DRAFT

and thus, dividing by n and taking n→∞, we have

lim infn→∞

n−1A([nx], [ny]) ≥ limn→∞

mq

n

A(mqx,mqy)

mq− cRx

n(1.26)

= Φ(x, y).(1.27)

For the other direction, we have nx + qx − rx = (m + 1)qx, and similarly for y, and thus we

have [nx] + qx − Rx = (m + 1)qx. In order to use superadditivity as above, we need to know that

(qx−Rx, qy−Ry) ∈ K. Without loss of generality, we can take y ≥ 0, since [ry] = −[−ry]. So, we need

to check that [rx]− [ry] ≤ qx− qy. If x− y is an integer, then so is rx− ry. In this case, it is easy to

check that rx− ry = [rx]− [ry], and so the inequality follows, since r < q.

Otherwise, note that

[rx]− [ry] ≤ [rx− ry] + 1 < r(x− y) + 1 ≤ (q − 1)(x− y) + 1.(1.28)

If q(x − y) ≥ (q − 1)(x − y) + 1, then [rx] − [ry] ≤ qx − qy as needed. Otherwise, since q(x − y) is an

integer and (q − 1)(x − y) is not by the assumption that x − y is not an integer, q(x − y) must be the

unique integer between (q− 1)(x− y) and (q− 1)(x− y) + 1. And since [rx]− [ry] is also an integer less

than (q − 1)(x− y) + 1 by (1.28), we have the inequality.

So, (1.20) gives

A([nx], [ny]) ≤ A([nx] + qx−Rx, [ny] + qx−Ry) + c(qx−Rx),(1.29)

and so similarly to above, dividing by n and taking a limit, we obtain the other sided inequality

lim supn→∞

n−1A([nx], [ny]) ≤ Φ(x, y).(1.30)

This finishes the existence of the limit for rational points (x, y) ∈ K ∩Q2.

Finally, as in [28], for all (x, y) ∈ K, extend

Φ(x, y) = sup

Φ(u, v)∣∣∣ (u, v) ∈ K ∩Q2, (x− u, y − v) ∈ K

.(1.31)

This agrees with the case of rational (x, y) by (1.24). Superadditivity is straightforward to show from

the definition, and (1.22), (1.23), and thus (1.24) are immediate.

DRAFT

July 11, 2016

DRAFT

Page 20: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

14 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

For arbitrary (x, y) ∈ K, we have homogeneity for rational k > 0, since

kΦ(x, y) = supkΦ(u, v)

∣∣∣ (u, v) ∈ K ∩Q2, (x− u, y − v) ∈ K

(1.32)

= sup

Φ(ku, kv)∣∣∣ (u, v) ∈ K ∩Q2, (x− u, y − v) ∈ K

(1.33)

= sup

Φ(u, v)∣∣∣ (u, v) ∈ K ∩Q2, (kx− u, ky − v) ∈ K

(1.34)

= Φ(kx, ky).(1.35)

For general k > 0, simply choose k1 < k < k2 rational, and use (1.24):

Φ(k1x, k1y) ≤ Φ(kx, ky) ≤ Φ(k2x, k2y).

Use homogoneity for k1 and k2 and then take the limits k1, k2 → k.

Now, let 0 ≤ t ≤ 1 and let (x1, y1), (x2, y2) ∈ K. By superadditivity and homogeneity, we have

concavity:

tΦ(x1, y1) + (1− t)Φ(x2, y2) = Φ(tx1, ty1) + Φ((1− t)x2, (1− t)y2)(1.36)

≤ Φ(tx1 + (1− t)x2, ty1 + (1− t)y2),(1.37)

and a finite concave function on an open set is continuous, so that we have continuity in K, the interior

of K. Though we know the value along the boundary, Φ(x,±x) = 0, we will obtain continuity on the

diagonals later.

To compute the limit (1.17) for general (x, y) ∈ K, we approximate by rationals. Let 0 < x1 < x < x2

and y1 < y < y2 be rational such that (x1, y1), (x2, y2) ∈ K. To be able to use subadditivity to

compare A([nx], [ny]) and A([nx1], [ny1]), we will need to have that ([nx]− [nx1], [ny]− [ny1]) ∈ K and

([nx2]− [nx], [ny2]− [ny]) ∈ K.

Since (x, y) ∈ K, then we can choose the rational points such that 2(y − y1) ≤ (x − x1) and

2(y2− y) ≤ (x2− x). Taking n large enough so that n(y− y1) and n(y2− y) are greater than 1, we then

DRAFT

July 11, 2016

DRAFT

Page 21: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.3. The shape function 15

DRAFT

have that n(x− x1)− n(y − y1) > 1 and n(x2 − x)− n(y2 − y) > 1, so that

n(x− x1)− n(y − y1) > 1(1.38)

=⇒[(nx− nx1)− (ny − ny1)

]≥ 1(1.39)

=⇒ [nx− nx1]− [ny − ny1] ≥ 1(1.40)

=⇒ [nx]− [nx1] ≥ [ny]− [ny1],(1.41)

and the similar argument with n(x2 − x)− n(y2 − y) > 1.

Then subadditivity gives

A([nx], [ny]) ≥ A([nx1], [ny1])− c([nx]− [nx1])(1.42)

and

A([nx], [ny]) ≤ A([nx2], [ny2]) + c([nx2]− [nx]).(1.43)

Dividing by n and taking the limit, we obtain

lim infn→∞

A([nx], [ny]) ≥ Φ(x1, y1)− c(x− x1)(1.44)

and

lim supn→∞

A([nx], [ny]) ≤ Φ(x2, y2) + c(x2 − x).(1.45)

Continuity of Φ on K implies that we can take (x1, y1) and (x2, y2) to (x, y) to obtain (1.17) for

general (x, y) ∈ K. And by remark (1.3.2), we see have (1.17) on all of K.

We also prove a simple bound for Φ:

Lemma 1.3.3.

Φ(x, y) ≤ yD + c(x− |y|),(1.46)

where D = E∣∣F (0)

∣∣.

DRAFT

July 11, 2016

DRAFT

Page 22: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

16 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Proof. For any (x, y) ∈ K∩Z2, taking y > 0, the action-maximizing path must move through every F (·)-

site between 0 and y, and the maximum action that can be achieved through these sites is∑yk=1

∣∣F (k)∣∣.

And we can bound the remaining steps by c(x− y). Therefore, for points (x, y) ∈ K∩Z2 with y > 0, we

have

A(x, y) ≤y∑k=1

∣∣F (k)∣∣+ c(x− y),(1.47)

and so we have for any (x, y) ∈ K, y > 0,

n−1A([nx], [ny]

)≤

[ny]∑k=1

∣∣F (k)∣∣+ c([nx]− [ny]).(1.48)

Taking limits as n → ∞ and using the distributional symmetry with respect to y of the model, we

obtain, for (x, y) ∈ K,

Φ(x, y) ≤ |y|D + c(x− |y|).(1.49)

1.3.2 Continuity of Φ on ∂K

In this section, we prove that

Lemma 1.3.4. Φ is continuous on ∂K.

Proof. Fix ε > 0. We first bound #

∣∣∣∣Γ(n, ⌊(1− ε)n⌋)∣∣∣∣, the number of paths from the origin to(n,⌊(1− ε)n

⌋).

We can represent each step in a path by a choice of +1, −1, or 0 depending on whether the path moves

up, down, or remains where it is respectively. Thus we can represent Γ(n,⌊(1− ε)n

⌋)by the set of

sequences s = (s1, s2, . . . , sn) ∈ −1, 0, 1n∣∣∣∣∣∣n∑i=1

si =⌊(1− ε)n

⌋.Since

⌊(1− ε)n

⌋= n − dεne, calling v the number of down steps in a sequence (of which there can be

at most⌊dεne/2

⌋), then there must be dεne − 2v steps which remain at a site, and the rest of the steps

must be up steps. We can bound the size of the set of paths by summing over the number of down steps,

of which there are(nv

)ways of placing them, and then

(n−vdεne−2v

)ways of placing the steps which remain

DRAFT

July 11, 2016

DRAFT

Page 23: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.3. The shape function 17

DRAFT

on a site. So, the size of the set of paths is equal to

bdεne/2c∑v=0

(n

v

)(n− vdεne − 2v

).

Using the fact that for 1 ≤ k ≤ n, (n

k

)≤(ne

k

)k,

we can bound the number of paths very naively by

(⌊dεne/2

⌋+ 1) max

0≤v≤bdεne/2c

(n

v

) max0≤v≤bdεne/2c

(n− vdεne − 2v

)≤(⌊dεne/2

⌋+ 1)( n

dεne

)2

≤ εn(ne

εn

)2dεne

.

We can take this less than or equal to enI(ε) for all n, where I(ε) = Cε log(1/ε) with some constant

C > 2.

Now, fix any path γ in Γ(n,⌊(1− ε)n

⌋). Its action A(γ) is the sum of

⌊(1− ε)n

⌋i.i.d. random

variables X1, X2, . . . , Xb(1−ε)nc corresponding to the up steps of γ, and then dεne non-independent

random variables corresponding to the other steps, whose sum we call Z. We can bound this second

group naively by cdεne. Then for λ > 0,

P(A(γ) > λn(1− ε) + cdεne

)≤ P

b(1−ε)nc∑i=1

Xi > λn(1− ε)

≤ exp

−C ′nλ2(1− ε)

(1.50)

by Hoeffding’s inequality [16].

Finally, we can control the maximum action of paths in Γ(n,⌊(1− ε)n

⌋)using a union bound,

P(A(n,⌊(1− ε)n

⌋)> λn(1− ε) + cdεne

)≤ #

∣∣∣∣Γ(n, ⌊(1− ε)n⌋)∣∣∣∣ · exp−C ′nλ2(1− ε)

≤ exp

n[I(ε)− C ′λ2(1− ε)

].

DRAFT

July 11, 2016

DRAFT

Page 24: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

18 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Since I(ε)→ 0 as ε→ 0, we can take λ = λ(ε) such that I(ε)−C ′λ2(1− ε) < 0 and such that λ→ 0

as ε→ 0. For this choice of λ, we then have that

limn→∞

1

nA(n,⌊(1− ε)n

⌋)= Φ(1, 1− ε) ≤ λ(1− ε) + cε,

and this tends to 0 as ε tends to 0. We can extend continuity to all points on ∂K by homogeneity.

1.3.3 Properties of the Shape Function

Due to the homogeneity condition in Theorem 1.3.1, if suffices to simply look at the one-dimensional

shape function,

a(α) := Φ(1, α).(1.51)

We know that a : [−1, 1] → [0, c] is continuous on [−1, 1], concave on [−1, 1], satisfies a(1) = 0 (from

remark 1.3.2), and is an even function. We can also prove one more quantitative result:

Lemma 1.3.5. a(0) = c.

Proof. By corollary 1.2.4, for any ε > 0, we can take a, b large enough such that the event

An =∃ x ∈ N s.t. 0 < x < an2/3 and 0 < dx < bn−1/3

(1.52)

has probability greater than 1− ε for all n. Taking n large enough so that n1/3 >> 2a, we can guarantee

x << n/2, so that the path described below can be constructed.

On the event An, let γn,x move balistically up to site x, optimally stay on the edge x, x+1 (following

ηx as defined in 1.9) as long as possible, and then move balistically back down to 0. That is,

γn,x(i) =

i for 0 ≤ i ≤ x

ηx(i) for x < i < n− x

n− i for n− x ≤ i ≤ n

.

Then on this event we can bound the scaled optimal action between c and the scaled action of this path,

(1.53) n−1A(γn,x) ≤ n−1A(n, 0) ≤ c.

DRAFT

July 11, 2016

DRAFT

Page 25: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.3. The shape function 19

DRAFT

Since the actions of the beginning and ending pieces of γn,x are bounded below by −2c(x+ 1), and

since A(ηx, x, n− x− 1) ≥ (n− 2x− 1)(c− dx) by 1.12, it is easy to see that on An for n large enough,

n−1A(γn,x) can be taken as close as we wish to c. And so, since An has probability as close to 1 as we

wish, 1.53 gives n−1A(n, 0)p→ c.

Remark 1.3.6. Since a(0) = c is a global maximum for a, concavity tells us that a is non-decreasing

on [−1, 0] and non-increasing on [0, 1].

An important question when looking at the shape function is its behaviour at 0. For a concave

function, right and left sided derivatives exist in the interior of the domain. Applying the bound in

(1.46), one immediately obtains

Lemma 1.3.7. The right sided derivative of the shape function a at 0, a′(0+) = limh→0+a(h)−ch , is

negative. Thus the shape function a has a corner at 0.

A possible candidate for the shape function would be simply linear on [0, 1], namely a(α) = c(1−α).

Indeed, this limit is easily achieved by considering the naive path γnaiven which moves ballistically up to

the best discrepancy in [0, bαnc], stays there optimally as long as it can, and then moves ballistically to

bαnc at time n. Formally, calling the optimal site rn,α = argminx∈[0,bαnc) dx, set

γnaiven (i) =

i if i ≤ rn,α

ηrn,α(i) if rn,α < i ≤ n− bαnc+ rn,α

i−(n− bαnc

)if n− bαnc+ rn,α < i ≤ n

.

Using equation 1.12, the law of large numbers, and the fact that drn,α → 0 a.s., it is easy to see that

A(γnaiven )

n→ c(1− α) almost surely.

However, it turns out that this is not the behavior of the shape function, which in fact lies above the

straight line. We relegate the proof of the following important theorem to section 1.6.

Theorem 1.3.8. For any α ∈ (0, 1), a(α) > c(1− α). Namely, the shape function a is non-linear on

[0, 1].

After numerical analysis, as mentioned in the introduction, we conjecture that there is a linear piece

of the shape function near 0

DRAFT

July 11, 2016

DRAFT

Page 26: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

20 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

../images/plot.pdf

1.4 The maximizing path

1.4.1 Basic properties

We will often bound the action of a path in terms of the discrepancies it moves through. We would

like to say that whenever there is a sign change of b(·), we can bound the action on such two sites by

2c −minx∈Range(γ) dx. But to avoid over-counting such actions (since there can be two sign changes in

only three steps), we index pairs only by the even numbers. We then have the following simple lemma.

Lemma 1.4.1. For any path γ in Γ(a, b),

A(γ, a, b) ≤ c(b− a)−∑i∈(a,b)i even

1b(i) 6= b(i+ 1)

· minx∈Range(γ)

dx.

We now prove some basic properties of the maximizing path γn.

Lemma 1.4.2. For any ε > 0, there exist constants N , k1, k2, and B such that for all n > N ,

1. P(

maxx∈Range(γn)|x| ∈(k1n

2/3, k2n2/3))

> 1− ε

2. P(∃ x ∈ Range(γn) such that dx < Bn−1/3

)> 1− ε.

Proof. From corollary 1.2.4, for n large enough, there exists a point x such that |x| < an2/3 and

dx < bn−1/3 with probability greater than 1 − ε. Inside this event, we consider the path which moves

DRAFT

July 11, 2016

DRAFT

Page 27: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.4. The maximizing path 21

DRAFT

ballistically up to x and then remains optimally on the edge x, x+ 1. We can lower bound its action

by

−c|x|+A(ηx, |x|, n) > −c(an2/3) +(n− |x|

)(c− dx)(1.54)

> −c(an2/3) + (n− an2/3)(c− bn−1/3)(1.55)

> cn− (2ca+ b)n2/3.(1.56)

Thus, the maximal path γn must have action greater than or equal to this quantity. Using this fact

and lemma 1.4.1, we can give an upper bound on the smallest discrepancy the path must reach.

cn− (2ca+ b)n2/3 ≤ cn−∑

i∈(0,n)i even

1b(i) 6= b(i+ 1)

· minx∈Range(γn)

dx.(1.57)

By the law of large numbers, for any 0 < λ < 1/4, we can take this to be less than

cn− (λn) minx∈Range(γn)

dx

with probability as close to 1 as we wish, for all n large enough. Then inside this event, we have that

minx∈Range(γn)

dx < (2ac+ b)λ−1n−1/3.

This gives the second statement of the lemma. We will now show the first statement, the bounds on the

path.

From theorem 1.1.3, there exists k1 small enough such that, for all n large enough, the probability

of the event µn

([−k1, k1]× [0, (2ac+ b)λ−1]

)= 0

is as close to 1 as we wish (Obviously k1 < a chosen above). So on this event, γn must reach a minimum

distance of k1n2/3 from the origin, by the bound above on minx∈Range(γn) dx.

Finally, we upper bound the range of γn. Any path with range outside of (−k2n2/3, k2n

2/3) moves

through all sites in either [1, k2n2/3] or [−k2n

2/3,−1]. Using the bounds in 1.10, we can bound the action

DRAFT

July 11, 2016

DRAFT

Page 28: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

22 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

of any such path from above by

(1.58) max

k2n2/3∑

i=1

∣∣F (i)∣∣,−k2n

2/3∑i=−1

∣∣F (i)∣∣+ (n− k2n

2/3)c.

The law of large numbers gives that with probability close to 1, for all n large enough, we can take this

less than

cn−(c− E

∣∣F (0)∣∣− ε)k2n

2/3.

Thus, choosing k2 large enough such that

k2

(c− E

∣∣F (0)∣∣− ε) > 2ca+ b,

a path outside of (−k2n2/3, k2n

2/3) has smaller action than the ballistic path to x inside the events we

have been working, and thus cannot have maximal action.

1.4.2 The ending edge

We now define the ending edge of the maximizing path γn. To do this, we consider the time that this

ending edge is reached, which we call τn. That is

Definition 1.4.3.

τn = min

t ∈ [0, n]

∣∣∣∣#∣∣∣γn([t, n])∣∣∣ ≤ 2

.

Note that this is well defined for any path, and that unless γn is identically 0 on [0, n], there are two

neighbors in γn([τn, n]

). Call the smaller neighbor of this edge `n. (Recall that we index discrepancies

by the smaller vertex.) Namely,

Definition 1.4.4.

`n = min γn([τn, n]

).

Note that, by maximality of the action, γn(t) = η`n(t) for all t ∈ [τn, n]. As noted in remark 1.2.2,

the maximizing path γn will tend to localize on sites with small discrepancy. In fact, the path will

tend to end on edges with record discrepancies, as we show with the lemma 1.4.6. But first, a lemma

controlling the probability of neighbors being close to the edge of their support.

DRAFT

July 11, 2016

DRAFT

Page 29: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.4. The maximizing path 23

DRAFT

Lemma 1.4.5. For any m ∈ N, there exists a constant λ = λ(m) > 0 such that for any ε > 0 and for

any integers a, b with b− a > 2m,

P

m∏k=−m

(c−

∣∣F (x+ k)∣∣) > λε

b− a∀x ∈ [a+m, b−m]

> 1− ε.(1.59)

Proof.

P

∃ x ∈ [a+m, b−m] such thatm∏

k=−m

(c−

∣∣F (x+ k)∣∣) ≤ λε

b− a

(1.60)

≤ P

⋃x∈[a+m,b−m]

m∏

k=−m

(c−

∣∣F (x+ k)∣∣) ≤ λε

b− a

(1.61)

≤ (b− a− 2m+ 1) P

m∏k=−m

(c−

∣∣F (x+ k)∣∣) ≤ λε

b− a

(1.62)

= (b− a− 2m+ 1)

∫∫∫x1···x2m+1≤ λε

b−ag(x1) · · · g(x2m+1) dx1 · · · dx2m+1,(1.63)

where here g is the pdf of c −∣∣F (0)

∣∣, and by assumption is supported on [0, c], continuously dif-

ferentiable on [0, c] with finite derivatives at the endpoints, and in particular bounded. And since the

Lebesgue measure of x1 · · ·x2m+1 ≤ λεb−a inside the cube [0, c]2m+1 is bounded by a constant times

λεb−a , by choosing λ small enough we can take the probability in (1.60) less than ε.

Lemma 1.4.6.

(1.64) P

(`n = argmin

x∈Range(γn)

dx

)→ 1.

Additionally, γn does not leave this site with the best discrepancy once there. That is,

(1.65) P(τn = min

t ∈ [0, n]

∣∣ γn(t) ∈ `n, `n + 1)→ 1.

Proof. First, we consider the event such that Range(γn)⊂ [−kn2/3, kn2/3], which, by lemma 1.4.2, we

can take to have probability arbitrarily close to 1 for all n large enough (k as prescribed by the lemma).

Call r = argminx∈Range(γn) dx. By the second part of the lemma, we can assume that dr < Bn−1/3 for

some B > 0. And by corollary 1.2.4, we can also assume that for all x ∈ Range(γn), dx − dr > δn−1/3

for some δ > 0. Again, the intersection of these events can be taken with probability arbitrarily close to

1 for n large enough.

DRAFT

July 11, 2016

DRAFT

Page 30: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

24 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Now, consider the set of pairs with a given small discrepancy

S =x, x+ 1 ∈ [−kn2/3, kn2/3)

∣∣∣ dx <= B′n−1/3,

where B′ is a constant larger than 4B. Lemma 1.4.5 (with m = 3) allows us to work inside the event

that S consists only of disjoint pairs. The lemma also gives, calling the neighbors of this set

∂S =x ∈ [−kn2/3, kn2/3] \ S

∣∣∣ x− 1, x+ 1 ∩ S 6= ∅,

that for all x ∈ ∂S, ∣∣F (x)∣∣ < c− λ,

where λ is some positive constant depending on ε, B, and k and independent of n.

And finally, call the complement of S in the given range (which includes ∂S)

S′ = [−kn2/3, kn2/3] \ S.

Now, consider the event where γn leaves the edge r, r + 1 for some interval of time. Let (t1, t2] be

such an interval, where γn(t1) ∈ r, r + 1 but γn((t1, t2]

)∩ r, r + 1 = ∅. For the path to benefit by

moving away, the action of γn on (t1, t2] must be larger than the action of the path ηr which remains

on the edge r, r + 1 optimally. Now, since r, r + 1 ∈ S, γn(t1 + 1) must be in ∂S. Thus, we have

(c− dr)(t2 − t1) ≤ A(ηr, t1, t2) < A(γn, t1, t2) ≤ c− λ+ c(t2 − t1 − 1),(1.66)

which gives that

t2 − t1 > λd−1r > λB−1n1/3.

So for the path to leave r, r + 1, it must leave for an interval of order n1/3.

Now, we decompose the path on (t1, t2] into two types of intervals. First, we can take intervals [s1, s2]

such that s1 is the step before γn moves into S, and s2 is the last step before it moves back into S′

(or t2). Precisely, γn(s1) and γn(s2 + 1) are in S′ and γn([s1 + 1, s2]

)⊂ S (And most importantly,

γn(s1) ∈ ∂S). Denote the set of such intervals of this type on (t1, t2] by I.

The intervals which remain after considering intervals of this type are intervals of steps for which γn

is contained entirely in S′, whose leftmost step is an entry of γn into S′ from S (hence is a step in ∂S).

The last point in the interval will be two steps before γn moves into S, by construction of the intervals

DRAFT

July 11, 2016

DRAFT

Page 31: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.4. The maximizing path 25

DRAFT

in I (or t2). Call the set of these intervals J . The intervals in I and J form a partition of the points

in (t1, t2].

We now argue that for any interval in I or J , γn cannot have action better than the path ηr which

remains optimally on the record discrepancy site r, thus concluding that inside the intersection of events

we are working on, with measure arbitrarily close to 1 for large enough n, the maximizing path γn will

not leave r, r+ 1 once reaching it. Note that, since both sets in I and J begin with a step in ∂S, the

same argument above applies, so that for any such interval to have action better than that of ηr, the

length of the interval must be at least λB−1n1/3.

First we consider an interval [s1, s2] in J . We use the simple bound from lemma 1.4.1 to say that

the action on this interval of γn is bounded by

c(s2 − s1 + 1)−∑

i∈[s2,s1)i even

1b(i) 6= b(i+ 1)

· minx∈S′

dx(1.67)

< c(s2 − s1 + 1)−(B

B′(s2 − s1 + 1)

)(B′n−1/3

)(1.68)

< (s2 − s1 + 1)(c− dr),(1.69)

since for any ε > 0, for n large enough, we can easily lower bound the sum of indicators∑i∈[a,b)i even

1b(i) 6= b(i+ 1)

by (1 − ε) b−a+1

4 on any interval [a, b] ⊂ [0, n] which satisfies, say, b − a > λB−1n1/3. The proof of this

probabilistic fact can be realized using Hoeffding’s Lemma [16] and a union bound over all such intervals.

And so, we see that the action on [s1, s2] must be smaller than the action of ηr on this interval.

Now we consider an interval [s1, s2] in I. To bound the path in such an interval, we only need to

say that for long enough time periods (s2 − s1 > λB−1n1/3), the action of a path restricted to a single

edge is primarily given by the length of the interval times c minus half the discrepancy of the edge.

By construction, for some x, x + 1 ∈ S, γn simply equals ηx (the optimal path on the edge x, x + 1)

on (s1, s2]. We compare this action to the action of ηr using equation 1.15, equation 1.11, and the

assumption that dx − dr > δn−1/3 for any x ∈ Range(γn).

DRAFT

July 11, 2016

DRAFT

Page 32: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

26 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

A(ηx, s1, s2)−A(ηr, s1, s2)(1.70)

=1

2(s2 − s1)(dr − dx) +

(N+(s1, s2)− s2 − s1

2

)(F (x) + F (x+ 1)−

(F (r) + F (r + 1)

))(1.71)

<1

2(s2 − s1)(dr − dx) +

∣∣∣∣N+(s1, s2)− s2 − s1

2

∣∣∣∣(dx + dr)(1.72)

<

∣∣∣∣N+(s1, s2)− s2 − s1

2

∣∣∣∣((B +B′)n−1/3)− 1

2(s2 − s1)(δn−1/3).(1.73)

Similarly to before, using a union bound and Hoeffding’s inequality, for any ε > 0, for n large enough,

we can work inside the event (with probability as close to 1 as we wish) that∣∣∣N+(a, b)− b−a

2

∣∣∣ < ε(b−a)

for any interval [a, b] ⊂ [0, n] with b− a > λB−1n−1/3. So, taking this ε less than δ2(B+B′) , we have that

for n large enough, A(ηx, s1, s2)−A(ηr, s1, s2) < 0. And so, for any interval in I, the action of the path

which remains at r, r + 1 must be larger than the path which has left.

Thus we have shown that the event in which the action-maximizing path γn leaves r, r+1 is disjoint

from the finite intersection of events which have probability arbitrarily close to 1 for n large enough.

And so this event has probability which tends to 0.

With probability tending to 1, the previous lemma gives that the maximizing path γn hits a record

discrepancy edge and then stays there until time n. Thus, with probability tending to 1, the point

farthest from the origin that γn reaches is `n or `n + 1. Then, combining the above lemma with lemma

1.4.2, we have

Corollary 1.4.7. For any ε > 0, there exist constants N , k1, and k2 such that for all n > N ,

(1.74) P(|`n| ∈

(k1n

2/3, k2n2/3))

> 1− ε.

1.4.3 The arrival time

Recall the arrival time τn of the maximizing path to the ending edge `n, `n + 1 from definition 1.4.3.

By lemma 1.4.6, with probability tending to 1, τn is the first hitting time of the edge `n, `n + 1,

τn = mint ∈ [0, n]

∣∣ γn(t) ∈ `n, `n + 1.

This time τn is an important quantity to understand in terms of the location of the final edge.

DRAFT

July 11, 2016

DRAFT

Page 33: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.4. The maximizing path 27

DRAFT

Numerics suggest that the ratio between them, τn/`n, does not grow to infinity (and in fact, appears

to be centered around a value less than or equal to 2). However, proving this has remained elusive. In

order to prove the main theorem of this paper, we assume this behavior. Namely, we assume

assumption[Bounded Arrival Condition] There exists a κ > 1 such that

P(τn > κ|`n|) −→ 0.(1.75)

assumption

This behavior implies linear behavior of the shape function near the origin, which we show.

Proof that assumption 1.1.4 ⇒ linearity of the shape function near 0. Inside the eventτn ≤ κ|`n|

, the

maximizing path to the point(⌊

2κ|`n|⌋, `n

)is exactly the maximizing path to the point

(⌊κ|`n|

⌋, `n

)which then remains on the edge `n, `n + 1 optimally until time

⌊2κ|`n|

⌋. Thus on this event, the

difference of the maximizing actions to these two points is given by

A(⌊

2κ|`n|⌋, `n

)−A

(⌊κ|`n|

⌋, `n

)= A

(η`n ,

⌊κ|`n|

⌋,⌊2κ|`n|

⌋).

Since

1

κ|`n|A(η`n ,

⌊κ|`n|

⌋,⌊2κ|`n|

⌋) p→ c

and since we are supposing that Pτn ≤ κ|`n|

p→ 1, we have that

1

κ|`n|

[A(⌊

2κ|`n|⌋, `n

)−A

(⌊κ|`n|

⌋, `n

)]p→ c.

But the left hand side converges to

2a

(1

)− a(

1

κ

)in probability, by theorem 1.3.1, and by the fact that `n

p→∞ from So then by uniqueness of the limit,

we must have linearity of a on [0, 1/κ],

2a

(1

)− a(

1

κ

)= c = a(0).

Linearity of the shape function near 0 also appears to be supported by numerics, specifically, linearity

appears to hold on the interval between 0 and approximately 0.6.

DRAFT

July 11, 2016

DRAFT

Page 34: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

28 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

1.5 Distributional Convergence

Define the maximizing path to a point (t, x) by

ξ(t,x) = argmaxξ∈Γ(t,x)

A(ξ).

We consider such a path to the “better” choice between x and x+ 1 at time t,

ξt,x := ξ(t,x)1ηx(t) = x+ ξ(t,x+1)1ηx(t) = x+ 1.

Define the path of length n which is this until time⌊κ|x|

⌋< n, and then is ηx until time n:

γn,x(t) :=

ξbκ|x|c,x(t) for 0 ≤ t ≤

⌊κ|x|

⌋ηx(t) for

⌊κ|x|

⌋≤ t ≤ n

.

Here κ is the arrival ratio given in assumption 1.1.4. Lemma 1.4.6 and assumption 1.1.4 give that γn is

a path of this type with probability tending to 1. Thus, we can say that with probability tending to 1,

`n = argmaxx∣∣∣ bκ|x|c<nA(γn,x)

= argmaxx∣∣∣ bκ|x|c<nBn(x),(1.76)

where Bn(x) is the centered and rescaled action,

Bn(x) := n−2/3(A(γn,x)− cn

).

Of course we need not look over allx∣∣∣ ⌊κ|x|⌋ < n

, since lemma 1.4.2 tells us that the maximizing

path is contained in a ball of order n2/3 with high probability.

Now, theorem 1.3.1 gives that

1

|m|A

(ξ(bκ|m|c,m)

)=

1

|m|A(⌊κ|m|

⌋,m)→ Φ(κ, 1)

in probability as |m| → ∞ (in fact almost surely, but we won’t need this). Note that we have used

distributional symmetry with respect to the second coordinate, as m can be positive or negative. We

DRAFT

July 11, 2016

DRAFT

Page 35: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.5. Distributional Convergence 29

DRAFT

also claim the same about 1|m|A

(ξ(bκ|m|c,m+1)

). Choosing 0 < ε < κ − 1, subadditivity gives for |m|

large enough that

A(⌊

(κ− ε)|m|⌋,m)

+A(

(⌊(κ− ε)|m|

⌋,m), (

⌊κ|m|

⌋,m+ 1)

)≤ A

(⌊κ|m|

⌋,m+ 1

)≤ A

(⌊(κ+ ε)|m|

⌋,m)−A

((⌊κ|m|

⌋,m+ 1), (

⌊(κ+ ε)|m|

⌋,m)

)

. Dividing by |m|, taking a limit, and then using continuity of the shape function Φ taking ε → 0, we

obtain that

1

|m|A

(ξ(bκ|m|c,m+1)

)p→ Φ(κ, 1).

Then since both random sequences converge in probability to the same limit, we have

1

|m|A

(ξbκ|m|c,m

)p→ Φ(κ, 1) = κa(κ−1).

Now, define the random function

gn(x) = −a′(0+)|x|n2/3

− dx2n1/3,

and call its maximizer

xn = argmaxx∈Z

gn(x).

Now we compare the random functionals Bn(x) and gn(x). Using equation 1.15 for the second

equality, we have

Bn(x)− gn(x)

=1

n2/3

[A

(ξbκ|x|c,x

)+A

(ηx, κ|x|, n

)− cn− a′(0+)|x|+ dx

2n

]

=|x|n2/3

( 1

|x|A

(ξbκ|x|c,x

)− (a′(0+) + κc)

)+κdx2

+1

|x|

(N+(κ|x|, n)− n− κ|x|

2

)(F (x) + F (x+ 1)

).(1.77)

For x = `n, each of these terms converge to 0 in probability as n→∞. The first because 1|x|A

(ξbκ|x|c,x

)p→

DRAFT

July 11, 2016

DRAFT

Page 36: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

30 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

κa(κ−1) as |x| → ∞, because a′(0+) + κc = κa(κ−1) by the linearity of the shape function on [0, κ−1],

and by lemma 1.9.1, because |`n|p→∞. The second because d`n

p→ 0 by lemma 1.4.6. And the third by

corollary 1.4.7, which gives that `n > kn2/3 for some k with high probability, together with the central

limit theorem. And so we have

Bn(`n)− gn(`n)p→ 0.

Similary we have

Bn(xn)− gn(xn)p→ 0.

To see this, note that the above argument follows for any point of order n2/3 (in the way of corollary

1.4.7) whose discrepancies tend to 0 in probability. This is immediately true of xn by theorem 1.1.3.

(And thus⌊κ|xn|

⌋< n as required to consider Bn(xn) for n large enough.)

We now claim that maximizer of gn(x) over x ∈ Z, xn, is equal to the maximizer of Bn(x), `n, with

probability tending to 1. To see this, define the difference in the maximum value of the functional gn

and the next maximum value by

ζn =

∣∣∣∣∣gn(xn)− argmaxx6=xn

gn(x)

∣∣∣∣∣.An immediate consequence of corollary 1.2.4 is that this random variable is bounded away from 0 with

high probability.

Lemma 1.5.1. For every ε > 0, there is a δ > 0 uniform in n large enough such that

Pζn > δ > 1− ε.(1.78)

If xn is in the domain of Bn (which has probability tending to 1), and if xn 6= `n, then

ζn ≤ gn(xn)− gn(`n)

≤(gn(xn)−Bn(xn)

)+(Bn(`n)− gn(`n)

),

where we have used the fact that Bn(xn) ≤ Bn(`n). By the above lemma, and since both of these terms

tend to 0 in probability as noted above, we have

Proposition 1.5.2.

P`n = xn → 1.

DRAFT

July 11, 2016

DRAFT

Page 37: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.5. Distributional Convergence 31

DRAFT

Now, gn can be viwed as the continuous functional g(x, y) = −a′(0+)|x| − y/2 on the point process

µn defined in theorem 1.1.3, and (xnn−2/3, dxnn

1/3) as the g-maximizing point in µn. We argue that

Lemma 1.5.3. (xnn−2/3, dxnn

1/3) converges in distribution to(X,Y

), where

(X,Y

)= argmax

(X,Y )∈supp(µ)

g(X,Y ).

Here µ is the Poisson Point Process on R× R+ with driving measure ν as described in theorem 1.1.3.

Proof. Call the level set Tr = g−1(r), r < 0, which is a one dimensional triangle in R× R+. Inside the

(Polish, see [22]) space N of point processes in R2 endowed with the vague topology (see 1.7.1), consider

the subset

S =µ 6= 0

∣∣µ(Tr) ≤ 1 for all r < 0.

The poisson process µ described in theorem 1.1.3, which has continuous driving measure, is inside S with

probability 1. This is because the distribution of the points µ in any compact rectangle R is given by N

uniformly distributed points, where N is a poisson random variable with parameter ν(R) (See [26]). It

is easy to check that the probability that two such points lie on the same triangle Tr in R is 0. Let Rn

be an increasing sequence of rectangles, Rn R×R+, and call the event An = µ(Tr) ≤ 1 ∀ Tr ⊂ Rn.

Then A1 ⊂ A2 ⊂ A3 ⊂ · · · , and by continuity of the measure from above,

P(µ ∈ S) = P(∩An) = limP(An) = 1.

Now, given a point process µ in S, the argmax M(µ) = argmax(x,y)∈supp(µ) g(x, y) is a well defined

function from S to R× R+. In fact, it is a continuous function, where S inherits the topology of vague

convergence as described in 1.7.1. To see this, fix an open set O in R×R+, and let µ be a point process

in M−1(O). Then, calling (x, y) = M(µ) ∈ O, we show that there is an open set around µ in M−1(O).

Consider the next maximizing point of µ,(x′, y′

)= argmax(x,y)∈supp(µ)\(x,y) g(x, y). Let B be an open

ball around (x, y) in O which does not contain this next best point(x′, y′

). Define a continuous function

h1 which is positive on the ball B and zero outside of it. Also define another continuous function h2

which is positive on the open region between the x-axis and the triangle through(x′, y′

), Tg(x′,y′), and

zero elsewhere (note that B is contained in the support of h2 by construction). Then the open set in

the vague topology (see 1.7.1)

U =

ν ∈ S

∣∣∣∣ ∫ h1dν > 0

⋂ν ∈ S

∣∣∣∣ ∫ h2dν < 2

DRAFT

July 11, 2016

DRAFT

Page 38: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

32 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

contains µ, since it describes elements of S with one point in the ball B around (x, y) and no other points

until the next best maximizer(x′, y′

). Any such point process has its maximizer in B ⊂ O, and so U

is contained in M−1(O). Thus, U is an open set with respect to the inherited vague topology in S, and

M is continuous. Since µn ∈ S converges weakly to µ in the vague topology, the continuous mapping

theorem (see [3], theorem 2.7, p. 21) gives that M(µn) converges in distribution to M(µ).

Convergence of (xnn−2/3, dxnn

1/3) gives convergence of

(xnn

−2/3, gn(xn))

=(xnn

−2/3, g(xnn−2/3, dxnn

1/3)),

and so in light of the fact that

∣∣∣∣(`nn−2/3, Bn(`n))−(xnn

−2/3, gn(xn))∣∣∣∣ p→ 0,

we have our main theorem:

Theorem 1.5.4.

(1.79)

(`nn2/3

,A(γn)− cn

n2/3

)d−→ (X, g(X,Y )),

where (X,Y ) = argmax(X,Y )∈supp(µ) g(X,Y ), with µ the Poisson process with driving measure ν abso-

lutely continuous w.r.t. Lebesgue measure withdensity given by

dν(dt× dy)

dt× dy=

p′(0+)

2 y, y > 0

0, y ≤ 0

and p the density of d0.

1.6 Non-linearity of the Shape Function

In this section, we prove theorem 1.3.8:

Theorem 1.6.1. For any α ∈ (0, 1), a(α) > c(1− α). Namely, the shape function a is non-linear on

[0, 1].

DRAFT

July 11, 2016

DRAFT

Page 39: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.6. Non-linearity of the Shape Function 33

DRAFT

Proof. Because the shape function is concave on [−1, 1] it suffices to simply check non-linearity at a

single point, say α = 12 . Fix 0 < δ < c

2 . Call `n = argminx∈[0,an2/3) dx the discrepancy minimizer of the

points in [0, an2/3), where a > 0 is an arbitrary positive constant. We will now look at points with small

discrepancy above `n. For simplicity, we take these discrepancies to be independent by only looking at

odd points, In =i ∈ [an2/3, bαnc)

∣∣∣ i is odd

. Define a sequence in the Z ∪ −∞

x1 = supx ∈ In | dx < δ, and xj+1 = supx < xj

∣∣x ∈ In, dx < δ.

This is a non-increasing sequence whose positive elements (if any) are not neighbors. Call kn =

argmaxk∈N xk > 0 the number of points x in In which satisfy dx < δ. If kn > 0, then by defini-

tion, xkn should be the smallest point in In with discrepancy less than δ (if there are any such points

in In, otherwise xkn = −∞), and x1 the largest. Now, consider the event that there are at least order

n of such points. Namely, call pδ = Pd0 < δ, and fix a positive constant λ < αpδ/2. Call the event

Sn = kn ≥ λn. A simple consequence of the law of large numbers is that P(Sn)→ 1.

We now recursively define a sequence of paths whose expected actions inside this event are strictly

increasing. First, define the path

γ0(t) =

t, 0 ≤ t ≤ `n

η`n(t), `n < t ≤ `n + n− banc

t− (n− banc), `n + n− banc < t ≤ n,

(1.80)

namely, the path which moves balistically to `n, remains optimally on the edge `n, `n + 1, and then

moves balistically to bαnc. (Recall definition 1.9 of ηx(·), the path constrained to x, x + 1 which

maximizes action). Similarly to the discussion of the naive path in section 1.3.3, it is easy to calculate

that the action of this path, which we call a0, divided by n converges to the line c(1− α) almost surely.

Since A(a0)/n is bounded for all n,

E[a0

n

]→ c(1− α).

For simplicity, we call m = bλnc. We now define a recursive “borrowing” argument in the event Sn,

where we take steps from the portion of the path constrained to `n, `n + 1 to spend more time at the

sites x1 > x2 > · · · > xm > an2/3, in order to improve the final ballistic portion of the path. We can do

this in a way such that there is an expected positive gain in the action.

DRAFT

July 11, 2016

DRAFT

Page 40: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

34 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Given the path γi−1, i = 1, . . . ,m, define

τi = inft ∈ (an2/3, n]

∣∣∣ γi−1(t) = xi + 1· 1Sn ,

the first (and only) time γi−1 hits the site xi + 1 (which will occur inside the event Sn), and define the

random variables

σi = 1Sn ·

inft > 0

∣∣ ηxi(τi − t) 6= xi + 1− 1 if ηxi(τi) = xi + 1

0 if ηxi(τi) = xi

.(1.81)

The non-negative random variable σi describes the number of steps we will borrow to remain at xi + 1.

The key observation is that if σi is positive, then a path which moves through xi at time τi−σi− 1 and

then remains at xi + 1 on [τi − σi, τi] visits these sites with the correct signs, guaranteeing a positive

contribution to the action at each of the σi+2 steps. We are repeating this process m = bλnc times, and

we want to guarantee that we have enough steps to borrow each time. Since γ0 spends n− bαnc extra

steps on the edge `n, `n + 1, we will be safe if we require σi ≤ n−bαncm . Otherwise, we do no borrowing

at that step. For n large enough, this is satisfied if we take σi ≤ 1−αλ . And note that 1−α

λ > 2, by our

choice of α = 1/2 and our restriction on λ from above. So, for simplicity, we will only allow σi ≤ 2.

Otherwise, if σi > 2, then set σi = 0. That is, redefine σi to be σi1σi ≤ 2.

To denote the borrowing, we define the operator

Φj,k,l(γ)(t) =

γ(k) for k − l ≤ t ≤ k

γ(t+ l) for j − l ≤ t < k − l

γ(t) otherwise

that shifts the portion of γ on [j, k] left l units, and sets γ equal its value at k for the l units to the right

of the shift, up to k.

Let τi = maxt ∈ [0, n]

∣∣ γi−1(t) = `n + 1· 1Sn , the last time γi−1 is on the edge `n, `n + 1 before

moving ballistically towards xi. Then for i = 1, . . . ,m, we recursively define the random paths

γi = Φτi,τi,σi(γi−1) · 1Sn + γi−11Scn .(1.82)

With this definition, in Sn, γi borrows σi steps from the portion of γi−1 that remains on the edge

DRAFT

July 11, 2016

DRAFT

Page 41: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.6. Non-linearity of the Shape Function 35

DRAFT

`n, `n+1, allowing γi to reach xi+1 σi steps earlier and remain there for σi steps, which are beneficial

by the definition of σi. We can compute the difference of actions for subsequent γi’s in Sn.

ai − ai−1 =

−A(η`n∣∣[τi−σi,τi))− τi−1∑k=τi

F (γi−1(k))b(k)

+

τi−σi−2∑k=τi−σi

F (γi−1(k + σi))b(k)

+ F (xi)b(τi − σi − 1) +

τi∑k=τi−σi

F (xi + 1)b(k)

· 1Sn(1.83)

Consider the P-preserving function on environments which flips the sign of each time variable, h :

F · b → F · −b, and consider the function on Sn which switches the values of F (xi) and F (xi + 1) for

the environment ω = F · b

gi : F · b 7−→ F xi↔xi+1 · b,

where

F a↔b(x) =

F (a) if x = b

F (b) if x = a

F (x) otherwise .

The function gi is a measure-preserving bijection from Sn to itself, as it maintains the value of the

discrepancy dxi . And, since the sites x1, . . . , xm can not be neighbors, the function Gk = gkgk−1· · ·g1

is also a measure preserving bijection on Sn for k = 1, . . . ,m, as is Hk := Gk h.

Calling ωi = Hi(ω), note that the path γi on ω is exactly equal to γi on ωi on the interval [τi+1, n]

for any ω ∈ Sn and for each i = 1, . . . ,m, since σj(ω) = σj(ωi) for all j ≤ i. In fact, γi will be the same

path in ω and ωj for all j = i, . . . ,m on [τi+1, n]. Note that the paths will not be equal on (`n, τi+1),

as f switches the signs of b but not of F (`n) and F (`n + 1), so that the optimal path η`n on the edge

`n, `n+1 will disagree in ω and ωi. We can use this to calculate the expectation of terms from equation

DRAFT

July 11, 2016

DRAFT

Page 42: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

36 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

(1.83).

E

τi−1∑k=τi

F (γi−1(k))b(k)

1Sn

=

∫Sn

τi−1∑k=τi

ω(k, γi−1(k))dω

=1

2

∫Sn

τi−1∑k=τi

(ω(k, γi−1(k)) +Hi−1(ω)(k, γi−1(k))

)dω

=1

2

∫Sn

τi−1∑k=τi

F (γi−1(k))(b(k) +−b(k))dω

= 0,(1.84)

since Hi−1(ω) fixes F on γi−1

([τi, τi)

)= [`n + 1, xi] (γi−1 is simply ballistic on this piece). Similarly,

Hi(ω) fixes σi, γi−1 on [τi, τi) (which is ballistic), F on [`n + 1, xi), and switches signs of b, and so

we have

E

τi−σi−2∑k=τi−σi

F (γi−1(k + σi))b(k)

1Sn

= 0.

The key observation now is in the final terms of equation (1.83). By construction of σi, on the event

Sn(i) = σi > 0, all terms below are positive, so that

[F (xi)b(τi − σi − 1) +

τi∑k=τi−σi

F (xi + 1)b(k)

]1Sn(i)

=[∣∣F (xi)

∣∣+∣∣F (xi + 1)

∣∣(σi + 1)]1Sn(i)

≥ (c− dxi)(σi + 2)1Sn(i)

(1.85)

where we have used equation 1.10. Since we can bound the first term in equation (1.83) very simply by

−cσi, we are able to bound E[(ai − ai−1)1Sn ].

E[(ai − ai−1)1Sn ] ≥ E[(−cσi)1Sn + (c− dxi)(σi + 2)1Sn(i)

]≥ E

[(2c− δ(σi + 2)

)1Sn(i)

]≥ (2c− 4δ)P(Sn(i)),(1.86)

DRAFT

July 11, 2016

DRAFT

Page 43: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.7. Proof of Theorem 1.1.3 37

DRAFT

where we are using the fact that σi ≤ 2 by definition. From the definition of σi and the fact that the event

Sn is in the sigma field generated by F (i)ni=an2/3 , it is simple to calculate that P(Sn(i)) = 3

16P(Sn).

Finally, we have

1

nE[A(n, bαnc)] ≥ 1

nE[am1Sn ]

=1

nE[a01Sn ] +

1

n

m∑i=1

E[(ai − ai−1)1Sn ]

≥ 1

nE[a01Sn ] +

m

n

(3

8(c− 2δ)

)P(Sn)

−→ c(1− α) + λ3

8(c− 2δ),(1.87)

since P(Sn)→ 1. And since 1nE[A(n, bαnc)] −→ a(α) and since δ was chosen less than c/2, we have that

a(α) is strictly greater than c(1− α), proving non-linearity.

1.7 Proof of Theorem 1.1.3

We assume that (Xk)k∈Z is a collection of i.i.d. random variables with density f . We assume that f is

symmetric with compact support [−c, c], and that it is continuously differentiable on its support (with

one sided limits at the end points). For every k ∈ Z, we define Yk = 2c − |Xk − Xk+1|. Let p denote

the density of Yk (which we will express later in terms of f). For each n ∈ N, we introduce the rescaled

point process µn as a random measure:

µn =∑k∈Z

δ(n−2/3k,n1/3Yk)

Theorem 1.7.1. As n → ∞, the distribution of µn converges weakly in the vague topology to the

distibution of µ, the Poisson process with driving measure ν absolutely continuous w.r.t. Lebesgue measure

with density given by:

dν(dt× dy)

dt× dy=

p′(0+)

2 y, y > 0

0, y ≤ 0.

1.7.1 Basics of Poisson Approximation

To prove Theorem 1.1.3 we need some terminology and theory on random point processes that we proceed

to describe. This description is based on [22]. We restrict our attention to point processes in R2 since

that is the class of point processes we deal with in this section. The space N consists of all integer-valued

DRAFT

July 11, 2016

DRAFT

Page 44: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

38 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

(nonnegative locally bounded) measures defined on the set B of bounded Borel sets in R2. This set is

equipped with σ-algebra N generated by maps µ 7→ µ(B), B ∈ B. A point process µ is a measurable

map from a probability space (Ω,F ,P) to (N,N ). Its distribution is, of course, the pushforward of P

under µ. A point process µ is called a.s.-simple if with probability 1, all the atoms of µ have weight 1,

which corresponds to the situation where no two points of the point process coincide.

The natural topology on N is vague topology. Its base is given by finite intersections of N-sets of the

form

µ : s <

∫fdµ < t,

where s, t ∈ R, and f is any continuous function with compact support.

For any random point process µ we denote

Bµ = B ∈ B : µ(∂B) = 0 a.s.,

where ∂B denotes the boundary of B.

A collection U of bounded Borel sets in R2 is called a DC(dissecting and covering)-ring if it is a ring

such that for any B ∈ B and any ε > 0, there is a finite cover of B by U-sets of diameter less than ε. A

DC-semiring is a semiring with the same property. An example of a DC-semiring is the collection of all

rectangles [a1, a2)× [b1, b2). The collection of all finite unions of rectangles is a DC-ring.

It turns out that to check the weak convergence to an a.s.-simple point process in vague topology it

is essentially sufficient to check the convergence of the avoidance function that computes the probability

that there is no points inside a given set. The following theorem is a specific case of Theorem 4.7 in [22].

Theorem 1.7.2. Let (µn)n∈N and µ be point processes in R2 and assume that µ is a.s.-simple. Suppose

that U ⊂ Bµ is a DC-ring and I ⊂ Bµ is a DC-semiring. Suppose further that

(1.88) limn→∞

Pµn(U) = 0 = Pµ(U) = 0, U ∈ U ,

and

(1.89) lim supn→∞

Eµn(I) ≤ Eµ(I) <∞, I ∈ I.

Then µn converges to µ weakly in vague topology.

DRAFT

July 11, 2016

DRAFT

Page 45: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.7. Proof of Theorem 1.1.3 39

DRAFT

1.7.2 Proof of Theorem 1.1.3

Let us take I to be the semiring of rectangles, U to be the ring of finite unions thereof, and check

conditions (1.88) and (1.89) of Theorem 1.7.2.

Take a rectangle I = [a1, a2)× [b1, b2) and write

Eµn(I) =∑

n2/3a1≤k<n2/3a2

PYk ∈ [b1n−1/3, b2n

−1/3)(1.90)

Lemma 1.7.3. For any k, the distribution of Yk has density w.r.t. Lebesgue measure given by

p(y) =

0 y < 0

2f ? f(2c− y) 0 ≤ y ≤ 2c

0 2c < y.

In fact, X0 −X1 is distributed as X0 +X1, which has density f ? f , the convolution of f with itself,

with support [−2c, 2c]. Since this is symmetric, |X0 + X1| is concentrated on [0, 2c] with density at

y ∈ [0, 2c] given by 2f ? f(y). The Lemma follows by composing this density with a reflection and then

a shift to the right by 2c.

Now, call FX and FY the CDF’s of Xk and Yk, respectively. FX has support [−c, c], FY has support

[0, 2c], and since f is continuously differentiable on [−c, c], p is continuously differentiable on [0, 2c]

(convolution maintains the same level of smoothness). Then both FX and FY are twice continuously

differentiable, and so by Taylor’s theorem, for x greater than (but near ) −c,

FX(x) = f(−c)(x+ c) +f ′(−c+)

2(x+ c)2 + o

((x+ c)2

),

and for x ∈ [0, 2c] (near 0)

FY (x) = p(0)x+p′(0+)

2x2 + o(x2) =

p′(0+)

2x2 + o(x2),

and

since p(0) = 0 as the following computation shows:

DRAFT

July 11, 2016

DRAFT

Page 46: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

40 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

p(0) = 2f ? f(2c) =

∫ ∞−∞

f(2c− τ)f(τ)dτ

=

∫ ∞−∞

(f(2c− τ)12c− τ ∈ [−c, c]

)(f(τ)1τ ∈ [−c, c]

)dτ

=

∫ ∞−∞

1τ ∈ [c, 3c] ∩ [−c, c]f(2c− τ)f(τ)dτ = 0

Now for large n, we can use the Lemma to rewrite (1.90) as

Eµn(I) = ([n2/3a2]− [n2/3a1])(FY (b1n

−1/3)− FY (b2n−1/3)

)= ([n2/3a2]− [n2/3a1])

(p′(0+)

2n−2/3(b2 − b1)2 + o(n−2/3)

),

so that

limn→∞

Eµn(I) =p′(0+)

2(a2 − a1)(b22 − b21) = p′(0+)

∫ a2

a1

∫ b2

b1

y dydt = Eµ(I),

and(1.89) holds true.

To prove (1.88), we take arbitrary disjoint rectangles

Ui = [a(i)1 , a

(i)2 )× [b

(i)1 , b

(i)2 ), i = 1, . . . ,m,

define

U =

m⋃i=1

Ui,

and compute

limn→∞

Pµn(U) = 0

.

We quote a Theorem 4.3 from [6]

Theorem 1.7.4. . If X1, . . . , Xn are m-dependent r.v.’s taking values 0 and 1, then for any function

h : Z+ → [−1, 1],

∣∣∣∣∣∣∣Eh n∑i=1

Xi

− ∞∑k=0

e−λλk

k!h(k)

∣∣∣∣∣∣∣ ≤ 6 minλ−1/2, 1

∑i,j:i6=j

cov(Xi, Xj) + (4m+ 1)n∑i=1

p2i

,where pi = PXi = 1, and λ =

∑ni=1 pi.

DRAFT

July 11, 2016

DRAFT

Page 47: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.7. Proof of Theorem 1.1.3 41

DRAFT

We shall apply this theorem to our situation. Notice that

µn(U) =∑k

Zn,k

is a finite sum of 1-dependent r.v.’s

Zn,k = 1(n−2/3k,n1/3Yk)∈U, k ∈ Z

since only k ∈⋃i[n

2/3a(i)1 , n2/3a

(i)2 ) contribute to this sum. Therefore we can take m = 1 in Theo-

rem 1.7.4. Next,

PZn,k = 1 = PYk ∈ n−1/3U(n−2/3k),

where U(x) = y : (x, y) ∈ U. Therefore, for large n,

pn,k = PZn,k = 1 =

∫n−1/3U(n−2/3k)

p′(0+)y

2dy = n−2/3

∫U(n−2/3k)

p′(0+)y

2dy,

so that

λn =∑k∈Z

PZn,k = 1 =∑k

n−2/3

∫U(n−2/3k)

p′(0+)y

2dy → λ, n→∞,

where

λ =

∫R

dx

∫U(x)

p′(0+)y

2dy =

∫U

p′(0+)y

2dxdy.

Since limn→∞ supk pn,k = 0 as n→∞, we also have

limn→∞

∑k∈Z

p2n,k = 0.

Most covariances in the estimate provided by Theorem 1.7.4 are equal to zero in our case. The only

nontrivial contribution comes from

cov(Zn,k, Zn,k+1) ≤ PZn,k = 1, Zn,k+1 = 1

≤ PYk ∈ n−1/3U(n−2/3k), Yk+1 ∈ n−1/3U(n−2/3(k + 1))

≤ PYk ≤ n−1/3b∗, Yk+1 ≤ n−1/3b∗,

DRAFT

July 11, 2016

DRAFT

Page 48: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

42 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

where b∗ = maxi b(i)2 . For large n, the r.h.s. is bounded by

2PXk+1 −Xk ≥ 2c− n−1/3b∗, Xk+1 −Xk+2 ≥ 2c− n−1/3b∗,

where the factor of 2 appears due to the symmetry in |Xk+1−Xk|. Since Xk, Xk+1, and Xk+2 are i.i.d.,

the latter probability equals

∫ c

c−n−1/3b∗

P−c ≤ Xk ≤ xk+1 − 2c+ n−1/3b∗P−c ≤ Xk+2 ≤ xk+1 − 2c+ n−1/3b∗f(xk+1)dxk+1

=

∫ c

c−n−1/3b∗

FX(xk+1 − 2c+ n−1/3b∗)2f(xk+1)dxk+1

=

∫ −c+n−1/3b∗

−cFX(x)2f(x+ 2c− n−1/3b∗)dx,

where in the last line we used a change of variable. Note that for large n, FX(x) in the integral

can be bounded by a constant C times x + c, in light of the Taylor expansion above. So, writing

Mf = supt∈[−c,c] f(t), we can bound the integral by

Mf

∫ −c+n−1/3b∗

−c

(C(x+ c)

)2dx

= Mf

∫ n−1/3b∗

0

(Cx)2dx = O(n−1)

.

Since there are O(n2/3) indices k contributing nonzero covariances we conclude that the total con-

tribution from the covariance term is O(n−1/3). This concludes the proof that the estimate provided by

Theorem 1.7.4 converges to 0 as n → ∞. We now take h(k) = 1k=0, so that Eh(∑Zn,k) is exactly the

avoidance function for measure µn, which completes the proof of Theorem 1.1.3.

1.8 Questions / Future Work

The polymer measure is considered on the set of length n directed polymers, but we can make sense

of the optimal path of infinite length, up to a given record discrepancy. Given a realization of space

variables F (x), call the locations of record discrepancies Yi in increasing distance from the origin.

That is, Y0 = 0, and

Yi+1 = argminy∈Z | dy<dYi

|y|.

DRAFT

July 11, 2016

DRAFT

Page 49: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

1.9. Appendix 43

DRAFT

Consider the set Πk of infinite paths which “end” at the edge Yk, Yk + 1, that is, paths γ : N0 → Z

such that γ(t) = ηYk(t) for all t greater than some τγ > 0. Of course, the action of these paths are

infinite, but even so, we can find the “optimal” such path. We can do this because we can compare any

two paths in Πk as follows.

If γ1(t) = γ2(t) = ηYk(t) for all t > τ , then say γ1 ≤ γ2 if A(γ1, 0, τ) ≤ A(γ2, 0, τ). This defines a

total ordering on Πk. For almost every realization of ω we can choose a unique greatest element, γk.

Conjecture 1.8.1. Define τk = minτ ∈ N

∣∣ γk(t) = ηYk(t) ∀t ≥ τk

. For P1-a.e. F,

Ykτk−→ αcrit

for P2-a.e. b.

An interesting question is what percentage of time this best path spends at the previous discrepancy

before reaching Yk.

Question 1.8.2. Call Lk =t ∈ N

∣∣ γk(t) ∈ Yk−1, Yk−1 + 1

, and define δk = maxLk −minLk. For

a.e. ω, does

δkτk−→ 0?

We also conjecture that this best path stabilizes as k → ∞. More precisely, we conjecture the

following.

Conjecture 1.8.3. For all τ > 0, for all ε > 0, there exists K = K(τ, ε) such that for any k1, k2 > K,

P(γk1

(t) = γk2(t) for all t = 0, . . . , τ

)> 1− ε.

1.9 Appendix

1.9.1 Probability

Here two simple probabilistic lemmas:

Lemma 1.9.1. If Xnp→ X and Yn

p→∞, then XYnp→ X.

Proof. For ε > 0 and δ > 0, there exists an N such that for all n > N ,

P(|Xn −X| > ε

)< δ.

DRAFT

July 11, 2016

DRAFT

Page 50: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

44 Chapter 1. A Zero Temperature Directed Polymer in a Product-Type Environment

DRAFT

Let M be such that for all m > M ,

P(Ym ≤ N) < δ.

Then we can write P(∣∣XYm −X

∣∣ > ε)

as

P(∣∣XYm −X

∣∣ > ε∣∣∣ Ym > N

)P(Ym > N)(1.91)

+P(∣∣XYm −X

∣∣ > ε∣∣∣ Ym ≤ N)P(Ym ≤ N),(1.92)

which is less than 2δ.

Lemma 1.9.2. Suppose Xn converges almost surely to a constant L, and let Yn be a sequence such

that Xnd= Yn for each n. Then Yn converges to L along a subsequence.

Proof. Let cj be any sequence increasing to L. Then by the convergence of Xn, we can choose an

increasing subsequence nj such that for each j,

P(Ynj ≤ cj) = P(Xnj ≤ cj) ≤ 2−j .

The Borel-Cantelli lemma gives that lim inf Ynj ≥ L a.s. Performing the same argument with a sequence

decreasing towards L gives that limYnj = L almost surely.

This is Kingman’s subadditive ergodic theorem [24, 25]:

Theorem 1.9.3. Suppose (Xm,n) is a sequence of random variables indexed by integers 0 ≤ m < n,

such that

1.

Xl,n ≤ Xl,m +Xm,n whenever 0 ≤ l < m < n.

2. The joint distributions of(Xm+1,n+1, 0 ≤ m < n

)are the same as those of

(Xm,n, 0 ≤ m < n

).

3. For each n, E∣∣X0,n

∣∣ <∞ and EX0,n ≥ −cn for some constant c.

Then the limit X = limn→∞1nX0,n exists almost surely and in L1, and

EX = limn→∞

1

nEX0,n = inf

n

1

nEX0,n

Furthermore, if for each k ∈ N, Xnk,(n+1)k is stationary and ergodic, X = EX a.s.

DRAFT

July 11, 2016

DRAFT

Page 51: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Chapter 2

Analyticity of the Effective Velocity

for Ballistic RWRE

2.1 Introduction

For d ≥ 1, the random walk in a random environment (RWRE) on Zd is defined as follows. Let V denote

the set of unit vectors in Zd,e ∈ Zd

∣∣∣ |e| = 1

= ±e1, . . . ,±ed, and call Pdκ the set of uniformly elliptic

probability vectors (p(e))e∈V ,∑p(e) = 1, where p(e) > κ for all e ∈ V for some κ > 0. An environment

ω =(ω(z, ·)

)z∈Zd is an element of Ω = (Pdκ)Z

d

, upon which we put an i.i.d. (product) measure P = ν⊗Zd

,

where ν is a probability measure on Pdκ.

For a fixed environment ω ∈ Ω and z0 ∈ Zd, the “quenched” random walk in ω starting at z0 is a

time homogeneous Markov chain with the following transition probabilities:

Pωz0(Xn+1 = z + e |Xn = z

)= ω(z, e) ∀z ∈ Zd, e ∈ V

and Pωz0(X0 = z0) = 1.

The “annealed” law is the semi-direct product Pz = P× Pωz defined by

Pz(·) =

∫Pωz (·)P(dω).

Under the so called Kalikow’s condition [21], a law of large numbers for multidimensional random

walk in random environment was established by Sznitman and Zerner [34]. Kalikow’s condition is a drift

condition on P, and under it, Xn/n converges P0-almost surely to a non-zero deterministic velocity v.

DRAFT

45

DRAFT

Page 52: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

46 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

To prove this, Sznitman and Zerner constructed renewal times τn in the direction l of drift, where τ1

is essentially the first time such that after τ1, Xn · l does not backtrack past Xτ1 · l.

The environments we consider exhibit strong drift, the so-called “non-nestling” case, where P-a.s. for

all z ∈ Zd, ∑e∈V

ω(z, e) · l ≥ ε

for some ε > 0. Specifically, in this paper we investigate an environment considered by Sabot in [30],

where the random environment is a small i.i.d. perturbation of a fixed homogeneous environment p0,

i.e.

ωγ(z, e) = p0(e) + γξ(z, e),

with γ small (controlling the strength of the perturbation) and ξ random. With a non-zero mean drift,

for γ small enough, this walk satisfies Kalikow’s condition, and Sabot [30] calculates a quantitative

expansion of the limiting velocity to second order for d ≥ 2,

vγ = d0 + γd1 + γ2d2 +O(γ3−ε

),

where d0 =∑e∈V e ·p(e) and d1 =

∑e∈V e ·E[ξ(z, e)]. In dimension d = 1, when d0 = 0, the second order

term d2 has a discontinuity at γ = 0, though in the one dimensional case, the velocity is explicitly known

[32]. Under the stronger assumption that d0 6= 0, Sabot calculates an expansion to third order. Sabot

produces this expansion via Kalikow’s random walk, which is a walk in a non-random environment,

constructed by taking an average of the random environment, weighted by the Green’s function of the

RWRE stopped upon exiting some bounded set (or killed with rate δ < 1).

Under the assumption d0 · e > 0 for e ∈ V, we prove

Theorem 2.1.1. The map γ → vγ · e is analytic in an interval around 0.

The method employed is to consider the often useful “environment viewed from the particle” Markov

chain, but for the space of ξ’s (restricted to a half space), and at the stopping times where the walk

advances in a component of the direction of drift e, under the assumption that d0 · e 6= 0. For small γ,

we analyze the Markov operator L(γ) of this chain as a linear operator on the Banach space of Lipschitz

functions, with a metric chosen to give certain contractive properties. We can extend the operator to

include complex values of γ, and we prove that in the space of linear operators, the map γ 7→ L(γ) is

holomorphic in a neighborhood of zero. Perturbation theory allows us to conclude that the operator has

a spectral gap (quasi-compactness), and that the unique invariant measure for this configuration Markov

DRAFT

July 11, 2016

DRAFT

Page 53: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.2. Background and Notation 47

DRAFT

chain is a holomorphic function of γ. This technique is inspired by Dolgopyat and Liverani in the study

of random walks in a Markovian environment [9].

Recently, Guo [12] recovered Sabot’s results to first order, in the case where the underlying envi-

ronment p0 is random (and not homogeneous). He considers two cases. The first case is when the

underlying environment is balanced, and the perturbed environment satisfies a Kalikow-type condition.

The second case is when the original environment satisfies Sznitman’s (T’) condition [33] (conjectured

to be equivalent to ballisticity). In both cases, Guo proves law of large numbers, and differentiabilty

of γ 7→ vγ at 0. In the balanced case, the results of Lawler [27] give the so-called Einstein relation. It

would be nice to extend analyticity results to the random underlying environment cases considered by

Guo, particularly in the balanced case, when d0 = 0.

2.2 Background and Notation

Denote the canonical basis vectors for the lattice Zd, d ≥ 1, by (e1, . . . , ed), and denote by V the set of

allowable jumps for the nearest neighbor random walk, V =e ∈ Zd

∣∣∣ |e| = 1

= ±e1, . . . ,±ed .

Call the set of uniformly elliptic jump probabilities

Pdκ =

(p(e))e∈V∣∣∣∣∣∣∑e

p(e) = 1, p(e) ∈ (κ, 1] ∀e ∈ V

,with ellipticity constant κ > 0. We fix a p0 ∈ Pdκ, and we assume the following. Calling the drift of p0

d0 =∑e∈V

e · p0(e),

we assume that this drift has a positive component in the e1 direction,

d0 · e1 > 0.

Now, call the set of allowed perturbations at a point

Ξ =

((ξe)e∈V) ∈ [−1, 1]V

∣∣∣∣∣∣∑e

ξe = 0

,and we endow this with the subspace Borel σ-algebra inherited from [−1, 1]V . Let µ be a (non-trivial)

probability measure on this space.

DRAFT

July 11, 2016

DRAFT

Page 54: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

48 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Denote the space of configurations Ω = ΞZd with product measure P = µ⊗Zd

, and consider elements

of Pdκ formed by perturbing p0 by γξ,

pγ,ξ(z, e) = p0(e) + γξ(z, e).

Note that, taking γ0 > 0 small enough so that for all e ∈ V, p0(e)− γ0 > κ, we can ensure that for any

|γ| < γ0, we have pγ,ξ(z, ·) ∈ Pdκ for all z. We also take γ0 small enough so that pγ,ξ has positive drift

for any ξ, i.e., take γ0 small enough so that p0(e1)− γ0 > p0(−e1) + γ0.

We consider the nearest-neighbor random walk on Zd in the environment pγ,ξ, starting from z0, with

law Ppγ,ξz0 and transition probabilities

Ppγ,ξz0

(Xn+1 = z + e |Xn = z

)= pγ,ξ(z, e) ∀z ∈ Zd, e ∈ V.

We denote the annealed measure of the walk by

(2.1) Pγz0(·) = E

(Ppγ,ξz0 (·)

).

The law of large numbers proven by Sznitman and Zerner in [34] holds for random environments

satisfying the so called Kalikow condition, as pγ,ξ does for all |γ| < γ0 and all ξ ∈ Ω, as noted in [30].

Thus there exists a nonzero vγ ∈ Rd, vγ · e1 > 0, such that

(2.2) limn→∞

Xn

n= vγ Pγ

· − a.s.

Instead of the standard environment viewed from the particle, we will consider a Markov chain on a

half space of Ω, along stopping times. But we first need require some notation.

2.2.1 The Configuration Space

We denote slabs of Zd perpendicular to e1 by

Ai,j =z ∈ Zd

∣∣∣ i < z · e1 ≤ j,

where i, j ∈ [−∞,+∞]. We will also consider slabs of configurations, Ωi,j = ΞAi,j . We denote product

measure on slabs by Pi,j = µ⊗Ai,j .

For shorthand, one subscript Aj denotes the left half up to j, A−∞,j , and similarly for Ωj and Pj .

DRAFT

July 11, 2016

DRAFT

Page 55: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.2. Background and Notation 49

DRAFT

Omitting subscripts altogether denotes the entire space Zd, so that Ω = ΞZd , endowed with product

measure P.

Given configurations ξ ∈ Ωi,j and η ∈ Ωj,k, we denote concatenation by

(ξ ? η)(z, ·) =

ξ(z, ·) for z ∈ Ai,j

η(z, ·) for z ∈ Aj,k.

For x ∈ Zd with x · e1 = k, denote the shift operator from Ai,j to Ai−k,j−k taking x to the origin by

(σxξ

)(z, ·) = ξ(z + x, ·).

Note the following properties of the shift, σx+y = σxσy and σx(ξ1 ? ξ2) = σxξ1 ? σxξ2. Product measure

is invariant with respect to this shift.

We will mostly be considering the half space of configurations on the ‘left-half” of Zd, Ω0 = ΞA0 ,

which is a compact product space endowed with the standard product σ-algebra, which we denote by F0

(Denote the slab product sigma-algebras similarly to above). We will be considering a complex Banach

space of functions on Ω0, and study the spectrum of a quasi-compact Markov operator on this space.

2.2.2 The Configuration Markov Chain

We now define a Markov chain on Ω0. The chain is analogous to the environment viewed from the

particle Markov chain, but we are shifting the configurations ξ, as opposed to the whole environment.

We shift only in one direction along hitting times of the next hyperplane, and restrict to the half plane,

so that we are “gluing” a new independent slice of environment at each step.

Precisely, we fix ξ ∈ Ω, and consider the random walk Xn in pγ,ξ, started at the origin. Call the

hitting time of the hyperplane z · e1 = k

Tk = infn ≥ 0 |Xn · e1 = k.

With our assumptions on p0 and γ0 (namely, that p0(e1)− γ0 > p0(−e1) + γ0), for P-a.e. ξ ∈ Ω, the

one dimensional (lazy) random walk Xn · e1 satisfies limn→∞Xn · e1 =∞ (see [32]). And thus for P-a.e.

ξ ∈ Ω, Tk < ∞ for all k ∈ N0. Note that the Tk and X1, . . . , XTk only depends on the environment

restricted to Ak−1, so in what follows, given ξ in Ωk−1 defined only on a half-space, Ppγ,ξz (XTk ∈ A) is

well defined.

DRAFT

July 11, 2016

DRAFT

Page 56: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

50 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Now, call π the projection from Ω to Ω0, π(ξ) = ξ|A0, and for ξ ∈ Ω and Xn the random walk in

pγ,ξ, call

ξn = π(σXTn ξ).

Lemma 2.2.1. (ξn)n∈N0 is a Markov chain on Ω0 under Pγ0 with initial distribution P0 and Markov

operator

Eγ0

[f(ξn+1)

∣∣∣ ξn] =

∫ ∑z:z·e1=1

f(σe(π)(ξ ? η)

)Ppγ,ξn0

(XT1 = z

)P0,1(dη).(2.3)

Proof. Let f0, f1, . . . , fn+1 be bounded measurable functions on Ω0, and let ξ ∈ Ω. Using the strong

Markov property for Xn under Ppγ,ξ0 , and following [4], we have

Epγ,ξ0

[fn+1(ξn+1)fn(ξn) · · · f0(ξ0)

]= E

pγ,ξ0

[fn+1(π(σXTn+1 ξ))fn(π(σXnξ)) · · · f0(π(ξ))

](2.4)

= Epγ,ξ0

[Epγ,ξXTn

[fn+1(π(σXT1 ξ))

]fn(ξn) · · · f0(ξ0)

](2.5)

= Epγ,ξ0

[Rξfn+1(XTn)fn(ξn) · · · f0(ξ0)

],(2.6)

where

Rξf(z) = Epγ,ξz

[f(π(σXT1

ξ))].

We obtain the above statement for Eγ0 as well by performing E-integration. We condition onXT1

, · · · , XTn

and ξ|An (with respect to which ξ0, · · · , ξn are measurable) to obtain

Eγ0

[fn+1(ξn+1)fn(ξn) · · · f0(ξ0)

]=(2.7)

Eγ0

[Eγ

0

[Rξfn+1(XTn)

∣∣∣ XT1 , · · · , XTn , ξ|An

]fn(ξn) · · · f0(ξ0)

].(2.8)

Now, we consider

Eγ0

[Rξfn+1(XTn)

∣∣∣ XT1, · · · , XTn , ξ|An

](2.9)

= Eγ0

∑z:z·e1=1

Ppγ,ξXTn

(XT1

= z)fn+1

(π(σXTn+zξ)

) ∣∣∣ XT1, · · · , XTn , ξ|An

.(2.10)

First, we note that Ppγ,ξXTn

(XT1

= z)

is measurable with respect to ξ|An and XTn , and in fact equals

DRAFT

July 11, 2016

DRAFT

Page 57: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.2. Background and Notation 51

DRAFT

Ppγ,ξn0

(XT1 = z

). Thus we can pull it out of the expectation to obtain

∑z:z·e1=1

Ppγ,ξn0

(XT1

= z)Eγ

0

[fn+1

(π(σXTn+zξ)

) ∣∣∣ XT1, · · · , XTn , ξ|An

].

Now, π(σXTn+zξ) = σzξn ? σXTn+z

(ξ|An,n+1

), and ξ|An,n+1

is independent of σ(XT1 , · · · , XTn , ξ|An).

And so

Eγ0

[fn+1

(π(σXTn+zξ)

) ∣∣∣ XT1, · · · , XTn , ξ|An

]=(2.11) ∫

Ωn,n+1

fn+1

(σzξn ? σ

XTn+zη)Pn,n+1(dη) =

∫Ω0,1

fn+1

(σz(ξn ? η

))P0,1(dη),(2.12)

by translation invariance of product measure. Thus in light of (2.7) we have the Markov property.

Remark 2.2.2. Note that we could define this Markov Chain on this space in the following way. Let

η1, η2, . . . ∈ Ω0,1 be an i.i.d. sequence of P0,1-distributed slabs. Given a configuration ξk ∈ Ω0, consider

the random walk Xkn in the environment pγ,ξk , stopped at the hitting time T k1 of the hyperplane z·e1 = 1,

and define

ξk+1 = σXkTk1

(ξk ? ηk

).

In the next section, we will extend the Markov operator for this chain to a Banach space of functions

with a contractive metric, and for the contraction, we need that the hitting times have exponential tails,

a simple consequence of our assumptions on γ and d0.

Lemma 2.2.3. For any ξ ∈ Ωj, |γ| < γ0, the hitting time Tk, k ≤ j + 1 of the walk Xn with transition

probabilities given by pγ,ξ satisfies

Ppγ,ξ0 (Tk > n) ≤ Ckλn/k,

where C =(pq

)1/2

, and where λ = 1−(√

p−√q)2

, with p = p0(e1)− γ0 and q = p0(−e1) + γ0.

Proof. For fixed γ and ξ, call Xn the walk in the environment pγ,ξ, started at the origin and call

Yn = Xn · e1, the projection to the one dimensional lazy walk along ne1 |n ∈ Z. Via a simple coupling

argument, the hitting time of ke1 for Yn is stochastically dominated by the hitting time of ke1 for the

following stationary simple random walk, Zn. Let P0(Zn + 1 = z + e1 |Zn = z) = p, P0(Zn + 1 =

z − e1 |Zn = z) = q, and P0(Zn + 1 = z |Zn = z) = 1 − p − q. Stochastic domination follows from the

fact that 0 < p < pγ,ξ(z, e1) and pγ,ξ(z,−e1) < q < 1 for any z ∈ Zd, ξ ∈ Ω0, and |γ| < γ0. Note that

p > q by our assumption on γ0.

DRAFT

July 11, 2016

DRAFT

Page 58: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

52 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Notice that for the one dimensional homogenous walk Zn, Tk is equal in distribution to the sum

of k independent copies of T1, the first hitting time of 1. So it suffices to analyze T1. The standard

technique is via martingales, namely, for λ > 0 set

(2.13) Mλn = E0

−λZn − n log

(ϕ(λ)

), where ϕ(λ) = E0

(e−λZ1

).

This is indeed a non-negative martingale, and by the Optional Stopping Theorem (theorem 5.7.5, [10]),

E0

(MλT1

)≤ E0

(Mλ

0

)=⇒ E0

(exp−λZT1

− T1 log(ϕ(λ)

))≤ 1(2.14)

=⇒ E0

(e−T1 log(ϕ(λ))

)≤ eλ.(2.15)

Thus, for ϕ(λ) < 1,

P0(T1 > N) = P0

(e−T1 log(ϕ(λ)) > e−N log(ϕ(λ))

)(2.16)

≤ E0

(e−T1 log(ϕ(λ))+N log(ϕ(λ))

)≤ eλ+N log(ϕ(λ)),(2.17)

and to minimize this tail probability for large N , we can minimize ϕ(λ). Using calculus on ϕ(λ) = e−λp+

eλq+1− (p+q), it is easy to check that λ0 = log(√

p/q)> 0 minimizes ϕ with ϕ(λ0) = 1− (

√p−√q)2,

as required.

Again, since Tk is distributed as the sum of k independent copies of T1, the lemma follows from the

simple union bound P0(Tk > N) ≤ kP0

(T1 >

Nk

).

2.3 The Operator on a Complex Banach Space

This section will lay the foundation for applying perturbation theory to the spectrum of the Markov

operator for the Markov chain of configurations defined above. The Doeblin-Fortet theorem 2.5.4 provides

the framework. Namely, if we consider the Banach space of bounded, Lipschitz functions on Ω0, with a

metric that allows us to prove a certain contractive bound, we can conclude quasi-compactness of the

operator (spectral gap property).

For the following, we consider complex values of γ. Call Br the ball of radius r > 0 around the origin

in C. We now define a metric on Ω0 that will be critical to our procedure.

DRAFT

July 11, 2016

DRAFT

Page 59: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.3. The Operator on a Complex Banach Space 53

DRAFT

2.3.1 The metric

Definition 2.3.1. Given two configurations ξ, ξ′ ∈ Ω0, for λ < θ < 1 (λ given in lemma 2.2.3) and

0 < β, for γ ∈ Bγ0 define the metric by

(2.18) dθ

(ξ, ξ′

)= supz: z·e1≤0

θ−z·e1(1 + ‖z ⊥ e1‖

)−βsupγ∈Bγ0

(maxe∈V

∣∣∣Log(pγ,ξ(z, e)

)− Log

(pγ,ξ′(z, e)

)∣∣∣),where here ‖·‖ denotes the taxicab metric in Zd, z ⊥ e1 denotes the vector (z2, . . . , zd), and Log denotes

the principal branch of the complex logarithm.

Instead of the difference of logarithms, we could have simplified the metric and used the difference

of configurations instead. Note that for γ ∈ Bγ0 , where γ0 is small enough so that p0(e)±γ0 ∈ (κ, 1−κ)

for all e ∈ V, we have that p0(e) + γξ(z, e) ∈ D for any ξ ∈ Ω0, for all z ∈ Zd and e ∈ V, where

D =x+ iy

∣∣x ∈ (κ, 1− κ), |y| < γ0

.

By noting that the derivative of Log(z) is 1/z, which is bounded in D, and that Log is one-to-one in D,

we can find m,M > 0 such that

(2.19) m <

∣∣Log(z)− Log(z′)∣∣

|z − z′|< M for all z, z′ ∈ D.

For γ ∈ Bγ0and taking z and z′ as pγ,ξ(z, e) and pγ,ξ′(z, e), we have that

∣∣∣Log(pγ,ξ(z, e)

)− Log

(pγ,ξ′(z, e)

)∣∣∣ > m|γ|∣∣∣ξ(z, e)− ξ′(z, e)∣∣∣

and ∣∣∣Log(pγ,ξ(z, e)

)− Log

(pγ,ξ′(z, e)

)∣∣∣ < M |γ|∣∣∣ξ(z, e)− ξ′(z, e)∣∣∣.

Taking the supremum over γ ∈ Bγ0 , and calling

(2.20) dθ(ξ, ξ′) = sup

z: z·e1≤0θ−z·e1

(1 + ‖z ⊥ e1‖

)−β(maxe∈V

∣∣∣ξ(z, e)− ξ′(z, e)∣∣∣),we see that

(2.21) m|γ0|dθ(ξ, ξ′) ≤ dθ(ξ, ξ′) ≤M |γ0|dθ(ξ, ξ′).

DRAFT

July 11, 2016

DRAFT

Page 60: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

54 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Thus the two metrics are equivalent, and it is easy to see that the metric dθ generates the Borel product

topology on Ω0 (see [29], Theorem 20.5). So we have the first property in the following lemma.

Lemma 2.3.2.

1. The metric dθ generates the product topology on Ω0.

2. Given configurations ξ, ξ′ in Ω−k, and a strip η ∈ Ω−k,0,

(ξ ? η, ξ′ ? η

)= θkdθ

(ξ, ξ′

).

3. Given x ∈ Zd, x · e1 = 0, we have

(σxξ, σxξ′

)≤(1 + ‖x ⊥ e1‖

)βdθ

(ξ, ξ′

).

Proof. The formula in (2) follows immediately from the notion of concatenation defined above.

For (3), we note that, for x · e1 = 0, we have dγθ(σxξ, σxξ′

)equals

supz: z·e1≤0

θ−z·e1(1 + ‖z ⊥ e1‖

)−β(maxe∈V

∣∣∣Log(pγ,ξ(x+ z, e)

)− Log

(pγ,ξ′(x+ z, e)

)∣∣∣)(2.22)

= supz: z·e1≤0

θ(x−z)·e1(

1 +∥∥(z − x) ⊥ e1

∥∥)−β(maxe∈V

∣∣∣Log(pγ,ξ(z, e)

)− Log

(pγ,ξ′(z, e)

)∣∣∣)(2.23)

≤ supz: z·e1≤0

(1 + ‖z ⊥ e1‖

)β(1 +

∥∥(z − x) ⊥ e1

∥∥)βdγθ(ξ, ξ′).(2.24)

The supremum term on the left, using the triangle inequality, is less than or equal to

supz: z·e1≤0

(1 +

‖x ⊥ e1‖1 +

∥∥(z − x) ⊥ e1

∥∥)β≤ (1 + ‖x ⊥ e1‖)β .

Given a complex-valued function f on Ω0, denote the supremum norm as

|f | = supξ∈Ω0

∣∣f(ξ)∣∣,

DRAFT

July 11, 2016

DRAFT

Page 61: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.3. The Operator on a Complex Banach Space 55

DRAFT

and the Lipschitz norm with respect to dθ,

s(f) = supξ1,ξ2∈Ω0ξ1 6=ξ2

∣∣f(ξ1)− f(ξ2)∣∣

dθ(ξ1, ξ2).

Finally, define the norm

‖f‖ = |f |+ s(f).

Define the space of complex valued functions on Ω0 with finite norm

B =f : Ω0 → C

∣∣ f ∈ F , ‖f‖ <∞.It is well known that this is a complex Banach space, and is the Banach space considered for the

Doeblin-Fortet theorem, 2.5.4.

2.3.2 The Operator

Denote by ΠR the space of nearest neighbor paths from the origin to the hyperplane z ∈ Z | z · e1 = R,

with elements π = (πi)`(π)i=0 , where `(π) denotes the length of the path, π0 = 0, and where for i =

1, . . . , `(π)− 1, πi ∈ AR−1,∣∣πi+1 − πi

∣∣ = 1, and π`(π) · e1 = R. Denote the endpoint π`(π) of the path π

by e(π). Denote the steps by ∆πi = πi+1 − πi.

For a configuration ξ in ΩR−1 and |γ| < γ0, we denote the probability of a path π ∈ ΠR with jump

probabilities pγ,ξ

Pγ,ξ[π] = Ppγ,ξ0

(Xk = πk for k = 0, . . . , `(π)

)(2.25)

=

`(π)−1∏i=0

pγ,ξ(πi,∆πi)(2.26)

=

`(π)−1∏i=0

(p0(∆πi) + γξ(πi,∆πi)

).(2.27)

This definition can be extended to all γ ∈ C, though obviously this no longer represents the probability

of a random walk taking the path π in the environment pγ,ξ.

For γ ∈ Bγ0, f ∈ B, ξ ∈ Ω0, we define a linear operator L(γ) on B:

L(γ)f(ξ) =

∫ ∑π∈Π1

f(σe(π)(ξ ? η)

)Pγ,ξ[π]P0,1(dη).

DRAFT

July 11, 2016

DRAFT

Page 62: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

56 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

For γ ∈ Bγ0∩ R, this coincides with the Markov operator for the Markov chain

(ξn

)in Ω0 defined

in lemma 2.2.1. For such real γ, and for every ξ ∈ Ω0, as noted in section 2.2.2,

∑π∈Π1

Pγ,ξ[π] = Ppγ,ξ0 (T1 <∞) = 1.

This gives that ∣∣L(γ)f∣∣ ≤ |f | for γ ∈ Bγ0

∩ R.

For complex γ,

limγ→0

supξ∈Ω0

z∈A0,e∈V

∣∣p0(e) + γξ(z, e)∣∣

p0(e) + |γ|ξ(z, e)= 1,

so, for 1 < ρ < θλ−1, (where λ is given in lemma 2.2.3 and θ in 2.3.1) we can choose γ0 small enough so

that for all γ ∈ Bγ0, for any π ∈ Π1 and ξ ∈ Ω0 we have

∣∣Pγ,ξ[π]∣∣ ≤ ρ`(π)P|γ|,ξ[π].(2.28)

We can write the operator L(γ) as a series of operators∑∞n=1 Un(γ), where

(2.29) Un(γ)f(ξ) =∑π∈Π1`(π)=n

∫f(σe(π)(ξ ? η)

)Pγ,ξ[π]P0,1(dη)

We use the 2.28 for Un(γ):

∣∣Un(γ)f(ξ)∣∣ ≤ |f | ∑

π∈Π1`(π)=n

∣∣Pγ,ξ[π]∣∣(2.30)

≤ |f |ρn∑π∈Π1`(π)=n

P|γ|,ξ[π](2.31)

= |f |ρnPp|γ|,ξ0

(T1 = n

).(2.32)

Lemma 2.2.3 then gives that

(2.33)∣∣Un(γ)f

∣∣ ≤ C|f |(ρλ)n,

which is summable, since ρλ < 1. This implies that the series for L(γ)f(ξ) is absolutely convergent for

DRAFT

July 11, 2016

DRAFT

Page 63: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.3. The Operator on a Complex Banach Space 57

DRAFT

all f ∈ B, uniformly for all ξ ∈ Ω0 and γ ∈ Bγ0 .

We now show s(L(γ)f

)is bounded for all f ∈ B. We look at s

(Un(γ)f

), and add and subtract the

same term to obtain

∣∣Un(γ)f(ξ1)− Un(γ)f(ξ2)∣∣ ≤∫ ∑

π∈Π1`(π)=n

∣∣∣f(σe(π)ξ1 ? η)− f(σe(π)ξ2 ? η)∣∣∣∣∣Pγ,ξ1 [π]

∣∣P0,1(dη)

+

∫ ∑π∈Π1`(π)=n

∣∣∣f(σe(π)ξ2 ? η)∣∣∣∣∣Pγ,ξ1 [π]− Pγ,ξ2 [π]

∣∣P0,1(dη).(2.34)

To bound the left term of the expression, we note that

∣∣∣f(σe(π)ξ1 ? η)− f(σe(π)ξ2 ? η)∣∣∣ ≤ θ(1 +

∥∥e(π) ⊥ e1

∥∥)βdθ(ξ1, ξ2) s(f)

using the properties of the metric, lemma 2.3.2. Thus the left hand term of (2.34) is less than or equal

to

θdθ(ξ1, ξ2) s(f)∑π∈Π1`(π)=n

(1 + n)β∣∣Pγ,ξ1 [π]

∣∣,(2.35)

and in light of (2.33), the supremum over pairs ξ1 6= ξ2 of this term divided by dθ(ξ1, ξ2) is summable.

For the right hand term of (2.34), we need the following lemma.

Lemma 2.3.3. For γ ∈ Bγ0 ,

dθ(ξ1, ξ2)−1∑π∈Π1`(π)=n

∣∣Pγ,ξ1 [π]− Pγ,ξ2 [π]∣∣ ≤ C(λρ

θ

)n(n+ 1)β+1

for any ξ1, ξ2 ∈ Ω0.

Proof. We first note that Log(z)/(z−1) is holomorphic on C\z ∈ R | z ≤ 0, with a removable singularity

at 1 with value 1. So∣∣Log(z)/(z − 1)

∣∣ is bounded away from zero on the unit disk minus 0. Thus,

there exists a positive constant C such that for B1 \ 0,

|1− z| ≤ C|Log z|.

Given the pair Pγ,ξ1 [π] and Pγ,ξ2 [π], call M(π) the term with larger modulus, and m(π) the term with

smaller modulus. Then noting that m(π)/M(π) 6= 0, we can bound

DRAFT

July 11, 2016

DRAFT

Page 64: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

58 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

(2.36)∣∣Pγ,ξ1 [π]− Pγ,ξ2 [π]

∣∣ ≤ C∣∣M(π)∣∣∣∣∣∣∣∣Log

(m(π)

M(π)

)∣∣∣∣∣∣.Recall that Log(z) = log|z|+ iArg(z), where Arg(z) ∈ (−π, π]. Since Arg(

∏zi) =

∑Arg(zi)− 2πk,

where k is the integer which minimizes the absolute value of∑

Arg(zi), it is clear that∣∣Arg(

∏zi)∣∣ ≤∣∣∑Arg(zi)

∣∣. Then, from the formula for the logarithm above, for zi 6= 0 we have

∣∣∣∣Log(∏

zi

)∣∣∣∣ ≤∑∣∣Log(zi)∣∣,

since

<(

Log(∏

zi

))=∑<(Log(zi)

)and by the inequality for Arg (which is the imaginary part of Log) given above.

With this, and noting that Arg(pγ,ξ(z, e)

)∈ (−π/2, π/2) by the choice of γ0, we have

∑π∈Π1`(π)=n

∣∣Pγ,ξ1 [π]− Pγ,ξ2 [π]∣∣

≤C∑π∈Π1`(π)=n

∣∣M(π)∣∣ n−1∑i=0

∣∣∣Log(pγ,ξ1(πi,∆πi)

)− Log

(pγ,ξ2(πi,∆πi)

)∣∣∣≤ Cn

∑π∈Π1`(π)=n

∣∣M(π)∣∣ supz:‖z‖≤n,e∈V

∣∣∣Log(pγ,ξ1(πi,∆πi)

)− Log

(pγ,ξ2(πi,∆πi)

)∣∣∣(2.37)

Note that we can bound

(2.38) supz:‖z‖≤n,e∈V

∣∣∣Log(pγ,ξ1(πi,∆πi)

)− Log

(pγ,ξ2(πi,∆πi)

)∣∣∣ ≤ (n+ 1)βθ−ndθ(ξ1, ξ2)

from the definition of the metric, 2.3.1. And∣∣M(π)

∣∣ is less than or equal to the sum∣∣Pγ,ξ1 [π]

∣∣+∣∣Pγ,ξ2 [π]∣∣,

each of which is less than or equal to ρ`(π) times the path probability with real parameter |γ|, as in (2.28).

So we can use lemma 2.2.3 to get that (2.37) is less than or equal to

C ′dθ(ξ1, ξ2)

(ρλ

θ

)n(1 + n)β+1,

as required.

DRAFT

July 11, 2016

DRAFT

Page 65: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 59

DRAFT

The bound given by lemma 2.3.3 is summable, since λρ < θ, and so together with (2.35), we obtain

that for all all f ∈ B, s(Ln(γ)f

)/‖f‖ is bounded and summable in n. This with 2.33 gives that

∥∥Un(γ)∥∥

is bounded by a summable sequence, uniformly for γ ∈ Bγ0. This implies convergence in norm of∑

Un(γ) to L(γ), uniformly for γ ∈ Bγ0 .

It also gives that∥∥L(γ)

∥∥ is bounded, so we have that L(γ)(B) ⊂ B for all γ ∈ Bγ0.

2.4 Proof of Main Result

2.4.1 Quasi-compactness

Now, the form of L(γ)n is clear when L(γ) is a Markov operator, i.e., when γ is real. We prove that this

form holds for complex γ as well.

Lemma 2.4.1. For all n ∈ N, γ ∈ Bγ0, f ∈ B, ξ ∈ Ω0,

L(γ)nf(ξ) =

∫Ω0,n

∑π∈Πn

f(σe(π)(ξ ? η)

)Pγ,ξ?η[π]P0,n(dη)

Proof. Given π ∈ Πi+1, π′ ∈ Πj+1, denote π ?π′ as the concatenation of the two paths, shifting the start

of π′ to the end of π:

(π ? π′

)j

=

πj for j = 1, . . . , `(π)

π′j−`(π) + e(π) for j = `(π) + 1, . . . , `(π) + `(π′).

Then for ξ ∈ Ωi, η ∈ Ωi,j , we have

(2.39) Pγ,ξ[π] · Pγ,σe(π)(ξ?η)[π′] = Pγ,ξ?η[π ? π′].

Inductively, assuming the assertion is true for n, then we can write L(γ)n+1f(ξ) = L(γ)(L(γ)nf(ξ)

)

DRAFT

July 11, 2016

DRAFT

Page 66: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

60 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

as

∫Ω0,1

∑π∈Π1

L(γ)nf(σe(π)ξ ? η

)Pγ,ξ[π]P0,1(dη)(2.40)

=

∫Ω0,1

∑π∈Π1

∫Ω0,n

∑π′∈Πn

f(σe(π′)(σe(π)(ξ ? η) ? η′)

)Pγ,σe(π)(ξ?η)?η′ [π

′]P0,n(dη′)Pγ,ξ[π]P0,1(dη)(2.41)

(2.42)

We can write σe(π)(ξ ? η) ? η′ as σe(π)(ξ ? η ? σ−e(π)η′), where σ−e(π)η′ is now an element of Ω1,n+1,

and since the product measure is translation invariant, we can integrate instead over ηπ = σ−e(π)η′ to

obtain

∫Ω0,1

∑π∈Π1

∑π′∈Πn

∫Ω1,n+1

f(σe(π′)(σe(π)(ξ ? η ? ηπ))

)Pγ,σe(π)(ξ?η?ηπ)[π

′]P1,n+1(dηπ)Pγ,ξ[π]P0,1(dη).

Now, we note that there is a bijection from pairs of paths in Π1×Πn to Πn+1 given above by concatenation

(and reversed by splitting a path in Πn+1 at the first time the path hits the hyperplane z · e1 = 1).

Thus we can sum over π′′ = π?π′ ∈ Πn+1, noting that e(π′′) = e(π)+e(π′). And we can write the double

integral as the integral over the product space Ω0,n+1 with product measure P0,1 × P1,n+1 = P0,n+1,

integrating with respect to η′′ = η′ ? ηπ. Finally, since

Pγ,σe(π)(ξ?η?ηπ)[π′] · Pγ,ξ[π] = Pγ,ξ?η?ηπ [π ? π′] = Pγ,ξ?η′′ [π

′′],

the claim follows.

Now we use this formula to prove a bound on the Lipschitz norm of L(0)kf . Decomposing as in

(2.34), and noting that for γ = 0 the second term is 0, we have

∣∣∣L(0)kf(ξ1)− L(0)kf(ξ1)∣∣∣ ≤θks(f)dθ(ξ1, ξ2)

∑π∈Πk

(1 +

∥∥e(π) ⊥ e1

∥∥)βP0,·[π].(2.43)

We can decompose the sum by the length of the paths, and use lemma 2.2.3 to obtain that this is

less than

θks(f)dθ(ξ1, ξ2)∞∑n=k

Ckλnk (1 + n)β ≤ s(f)dθ(ξ1, ξ2)

(C ′ kθkλβ/k

(1− λ1/k)β

),

where C ′ does not depend on k. The final term tends to 0 as k grows, and so we can take k large enough

DRAFT

July 11, 2016

DRAFT

Page 67: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 61

DRAFT

such that there is an 0 < α < 1 with

(2.44) s(L(0)kf

)≤ αs(f).

And since∣∣L(0)f

∣∣ ≤ |f |, we have that

∥∥∥L(0)kf∥∥∥ ≤ α‖f‖+ (1− α)|f |.(2.45)

As per the discussion following theorem 2.5.2, as (Ω0, dθ) is a compact metric space, and B is as

defined, we can conclude that L(0) is quasi-compact and of diagonal type, provided the spectral radius

of L(0) is 1. To see this, recall that for any real |γ| < γ0, L(γ) is a Markov operator, and 1 is an

eigenvector with eigenvalue 1. So r(L(0)

)≥ 1. To see that r

(L(0)

)≤ 1, as shown above we have that

for any f ∈ B,

s(L(0)kf

)≤ s(f)εk,

where εk → 0 as k →∞ and does not depend on f .

This and the fact that∣∣∣L(0)kf

∣∣∣ ≤ |f | for any f ∈ B gives that

∥∥∥L(0)k∥∥∥ = sup

f∈Bf 6=0

∥∥∥L(0)kf∥∥∥

‖f‖≤ 1 + εk.

Taking the k-th root and the limit as k →∞ gives r(L(0)

)≤ 1.

Lemma 2.4.2. For any f ∈ B, limn→∞ L(0)nf(ξ) = f , where f is a constant function.

Proof. First note that if f ∈ B depends measurably only on ξ(z, ·) for z · e1 > −N for some N ∈ N,

then for n ≥ N , for any π ∈ Πn, f(σe(π)ξ ? η) is independent of ξ, and by shift invariance of P0,n, from

lemma 2.4.1 we then have that L(0)nf(ξ) is a constant.

Now fix an arbitrary f ∈ B, and define fm(ξ) = f(xm ? ξ|A−m,0

), where xm is an arbitrary but fixed

element of Ωm. fm depends only on ξ(z, ·) for z · e1 > −m, and so L(0)nfm(ξ) is a constant for n ≥ m,

namely, it is equal to∫

Ω−n,0fm(η)P−n,0(dη), which is equal to

∫Ω0fm(η)P0(dη).

And ∣∣fm(ξ)− f(ξ)∣∣ ≤ s(f) dθ

(xm ? ξ|A−m,0 , ξ

)≤ C s(f) θm,

DRAFT

July 11, 2016

DRAFT

Page 68: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

62 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

so since∣∣L(0)ng

∣∣ ≤ |g| for any n, we have that for all ξ ∈ Ω0,

limm→∞

L(0)nfm(ξ) = L(0)nf(ξ)

uniformly in n (and ξ). Then we can interchange limits to obtain

limn→∞

L(0)nf(ξ) = limm→∞

limn→∞

L(0)nfm(ξ) = limm→∞

∫Ω0

fm(η)P0(dη).

By the dominated convergence theorem, the right hand side converges to∫

Ω0f(η)P0(dη).

This fact immediately gives that 1 is the only peripheral eigenvalue for L(0), and it is simple.

Then in light of remark 2.5.4 and the decomposition in 2.5.3, we can write

L(0)n = 1⊗ ϕ1 +Nn,(2.46)

where 1 ∈ B is the constant function 1, where ϕ1 is a bounded linear functional such that ϕ1(1) = 1,

L(0)∗ϕ1 = ϕ1, and r(N) < 1.

Calling F = vect1 and H =f ∈ B

∣∣ϕ1(f) = 0

, then we see that B = F ⊕H, as f = ϕ1(f)1 +(f − ϕ1(f)1

). L(0) takes constant functions to themselves, so that L(0)(F ) ⊂ F , and to see that

L(0)(H) ⊂ H, suppose that there is a function f ∈ H such that L(0)f = C1 for some some constant

C. Since L(0)C1 = C1, this implies that C = L(0)nf = Nnf (since we are assuming ϕ1(f) = 0), which

contradicts r(N) < 1 unless f ≡ 0.

Thus as in definition 2.5.5, 1 is a dominating simple eigenvalue for L(0).

2.4.2 Differentiability

Lemma 2.4.3. The map γ 7→ L(γ) is holomorphic in the disc Bγ0.

Proof. By theorem 2.5.7, it suffices to show that for any ϕ ∈ B∗ and f ∈ B, γ → ϕ(L(γ)f

)is holomorphic

in Bγ0 . Recalling that L(γ) =∑Un(γ) as in (2.29), we showed in section 2.3.2 that this series converges

in norm uniformly for γ ∈ Bγ0. Convergence in norm implies weak convergence, so that ϕ(L(γ)f) =∑

ϕ(Un(γ)f). And a consequence of Morera’s theorem is that a series of holomorphic functions in a

domain D is holomorphic on compact subsets of D if the series is uniformly convergent on compact

subset of D (7.3 in [11]).

Thus it suffices to show that γ → Un(γ) is holomorphic in norm for γ ∈ Bγ0, n ∈ N

DRAFT

July 11, 2016

DRAFT

Page 69: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 63

DRAFT

From here on, fix γ ∈ Bγ0 and call

Dγ,ξ[π] :=(Pγ,ξ[π]

)−1 d

dγPγ,ξ[π] =

`(π)−1∑i=0

ξ(πi,∆πi)

p0(∆πi) + γξ(πi,∆πi).(2.47)

We now focus on controlling

∣∣∣∣∣Pγ+h,ξ[π]− Pγ,ξ[π]

h−Dγ,ξ[π]Pγ,ξ[π]

∣∣∣∣∣,both in terms of the supremum over ξ, and in terms of the difference between two configurations ξ1 and

ξ2. Note that

Pγ+h,ξ[π]

Pγ,ξ[π]=

`(π)−1∏i=0

p0(∆πi) + (γ + h)ξ(πi,∆πi)

p0(∆πi) + γξ(πi,∆πi)(2.48)

=

`(π)−1∏i=0

(1 + h

ξ(πi,∆πi)

p0(∆πi) + γξ(πi,∆πi)

),(2.49)

and call ai = ξ(πi,∆πi)p0(∆πi)+γξ(πi,∆πi)

. Then factoring out∣∣Pγ,ξ[π]

∣∣, we first control

(2.50)

∣∣∣∣∣∣∣(

Pγ+h,ξ[π]Pγ,ξ[π]

)− 1

h−Dγ,ξ[π]

∣∣∣∣∣∣∣ =

∣∣∣∣∣∣∏n−1i=0 (1 + hai)− 1

h−n−1∑i=0

ai

∣∣∣∣∣∣.

For notational simplicity, we denote this difference term by

Aγ,ξh [π] =

∏n−1i=0 (1 + hai)− 1

h−n−1∑i=0

ai.

To control this term, we use the multi-binomial theorem, which gives

n∏i=1

(1 + bi) = 1 +∑

1≤i1≤n

bi1 +∑

1≤i1<i2≤n

bi1bi2 +∑

1≤i1<i2<i3≤n

bi1bi2bi3+(2.51)

· · ·+∑

1≤i1<i2<···<in≤n

bi1bi2 · · · bin .(2.52)

DRAFT

July 11, 2016

DRAFT

Page 70: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

64 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Then we have that

Aγ,ξh [π] = h∑

1≤i1<i2≤n

ai1ai2 + h2∑

1≤i1<i2<i3≤n

ai1ai2ai3+(2.53)

· · ·+ hn−1∑

1≤i1<i2<···<in≤n

ai1ai2 · · · ain .(2.54)

This tends to 0 as h tends to 0 uniformly for all ξ ∈ Ω0 and π ∈ Π1 with `(π) = n, since |ai| ≤ κ−1

for any ξ ∈ Ω0. So for f ∈ B, calling the operator

Vn(γ)f(ξ) =

∫ ∑π∈Π1`(π)=n

f(σe(π)(ξ ? η)

)Pγ,ξ[π]Dγ,ξ[π]P0,1(dη),

we have that, for fixed n,

|f |−1

∣∣∣∣∣Un(γ + h)f − Un(γ)f

h− Vn(γ)f

∣∣∣∣∣(2.55)

tends to 0 as h tends to 0 uniformly for f ∈ B.

For the Lipschitz norm, fix ξ1 and ξ2, and call a1i = ξ1(πi,∆πi)

p0(∆πi)+γξ1(πi,∆πi)and a2

i = ξ2(πi,∆πi)p0(∆πi)+γξ2(πi,∆πi)

(which depend on π).

We need to show that supf 6=0‖f‖−1s

((h−1

(Un(γ + h)− Un(γ)

)− Vn

)f

)tends to 0 as h tends to

0, or that

∣∣∣∣∣∣∣∣∣∫ ∑

π∈Π1`(π)=n

[f(σe(π)ξ1?η)Pγ,ξ1 [π]Aγ,ξ1h [π]− f(σe(π)ξ2?η)Pγ,ξ2 [π]Aγ,ξ2h [π]

]P0,1(dη)

∣∣∣∣∣∣∣∣∣is bounded by ‖f‖dθ(ξ1, ξ2) times something which tends to 0 as h tends to 0, independent of f , ξ1, and

ξ2. To do this, we decompose similarly to (2.34) and bound this by the sum of two terms,

∫ ∑π∈Π1`(π)=n

∣∣∣f(σe(π)ξ1 ? η)− f(σe(π)ξ2 ? η)∣∣∣∣∣∣Pγ,ξ1 [π]Aγ,ξ1h [π]

∣∣∣P0,1(dη)

+

∫ ∑π∈Π1`(π)=n

∣∣∣f(σe(π)ξ2 ? η)∣∣∣∣∣∣Pγ,ξ1 [π]Aγ,ξ1h [π]− Pγ,ξ2 [π]Aγ,ξ2h [π]

∣∣∣P0,1(dη).(2.56)

Just as in (2.35), we see that we can bound the first term by s(f)dθ(ξ1, ξ2) times something which tends

to 0 independently of ξ1 and ξ2 as h tends to 0.

DRAFT

July 11, 2016

DRAFT

Page 71: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 65

DRAFT

For the second term, we can bound this by |f | times

∑π∈Π1`(π)=n

∣∣Pγ,ξ1 [π]− Pγ,ξ2 [π]∣∣∣∣∣Aγ,ξ1h [π]

∣∣∣+∣∣Pγ,ξ2 [π]

∣∣∣∣∣Aγ,ξ1h [π]−Aγ,ξ2h [π]∣∣∣

As seen by lemma 2.3.3, the difference of path probabilities in the first term is bounded by a constant

times dθ(ξ1, ξ2) as required.

Finally, we control the second term. For a given path π, we can bound∣∣∣Aγ,ξ1h [π]−Aγ,ξ2h [π]

∣∣∣ by

|h|∑

1≤i1<i2≤n

∣∣∣a1i1a

1i2 − a

2i1a

2i2

∣∣∣+ |h|2∑

1≤i1<i2<i3≤n

∣∣∣a1i1a

1i2a

1i3 − a

2i1a

2i2a

2i3

∣∣∣+· · ·+ |h|n−1

∑1≤i1<i2<···<in≤n

∣∣∣a1i1a

1i2 · · · a

1in − a

2i1a

2i2 · · · a

2in

∣∣∣.(2.57)

We control each of these terms inductively, by first noting that

(2.58) sup‖z‖≤n,e∈V

∣∣ξ1(z, e)− ξ2(z, e)∣∣ ≤ c(n)dθ(ξ1, ξ2)

by 2.21, where c(n) is a constant that depends on n (and γ0, θ, and β) but not on ξ1 and ξ2. We can then

obtain a bound for∣∣∣a1i − a2

i

∣∣∣, which by multiplying by∣∣∣(p0(∆πi) + γξ1(πi,∆πi)

)(p0(∆πi) + γξ2(πi,∆πi)

)∣∣∣(which is greater than or equal to κ2), we have then that κ−2

∣∣∣a1i − a2

i

∣∣∣ is less than or equal to

∣∣∣ξ1(πi,∆πi)(p0(∆πi) + γξ1(πi,∆πi)

)− ξ2(πi,∆πi)

(p0(∆πi) + γξ1(πi,∆πi)

)∣∣∣.(2.59)

And by adding and subtracting the mixed term and noting that for ‖z‖ ≤ n and e ∈ V,

∣∣∣(p0(e) + γξ1(z, e))−(p0(e) + γξ2(z, e)

)∣∣∣ ≤ c(n)|γ0|dθ(ξ1, ξ2),(2.60)

we obtain that∣∣∣a1i − a2

i

∣∣∣ is less than or equal to dθ(ξ1, ξ2) times a constant independent of ξ1 and ξ2.

Now, for k ≥ 2 we control individual terms

∣∣∣a1i1a

1i2 · · · a

1ik− a2

i1a2i2 · · · a

2ik

∣∣∣

DRAFT

July 11, 2016

DRAFT

Page 72: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

66 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

by adding and subtracting a common term to obtain that this is less than or equal to

∣∣∣a1i1

∣∣∣∣∣∣a1i2 · · · a

1ik− a2

i2 · · · a2ik

∣∣∣+∣∣∣a1i1 − a

2i1

∣∣∣∣∣∣a2i2 · · · a

2ik

∣∣∣ ≤(2.61)

κ−1∣∣∣a1i2 · · · a

1ik− a2

i2 · · · a2ik

∣∣∣+ Cdθ(ξ1, ξ2)κ−(k−1).(2.62)

We have a recurrence relation on the number of terms, and induction gives that for any indices i1, . . . , ik

and j1, . . . , jk, we have

∣∣∣a1i1a

1i2 · · · a

1ik− a2

i1a2i2 · · · a

2ik

∣∣∣ ≤ Cκ−kdθ(ξ1, ξ2).(2.63)

Finally, from (2.57) we obtain

∑π∈Π1`(π)=n

∣∣Pγ,ξ2 [π]∣∣∣∣∣Aγ,ξ1h [π]−Aγ,ξ2h [π]

∣∣∣ ≤(2.64)

dθ(ξ1, ξ2)∑π∈Π1`(π)=n

∣∣Pγ,ξ2 [π]∣∣ n∑k=2

|h|k−1Cκ−k

(n

k

),(2.65)

with the final term tending to 0 in h as required.

So, since |f | and s(f) are bounded by ‖f‖, along with (2.55) we have shown that

limh→0

∥∥∥h−1(Un(γ + h)− Un(γ)

)− Vn(γ)

∥∥∥ = 0.

Of course, Vn(γ) needs to be a bounded linear operator on B. With uniform boundedness in ξ of

Dγ,ξ[π], it is straightforward to check that∣∣Vn(γ)

∣∣/‖f‖ is bounded. To check that s(Vn(γ)f

)‖f‖ is

bounded, we can again decompose as in (2.34), and the details are similar to the above. The only new

detail is bounding

dθ(ξ1, ξ2)−1∑π∈Π1`(π)=n

∣∣Pγ,ξ2 [π]∣∣∣∣Dγ,ξ1 [π]−Dγ,ξ2 [π]

∣∣(2.66)

≤ dθ(ξ1, ξ2)−1∑π∈Π1`(π)=n

n−1∑k=0

∣∣∣a1k − a2

k

∣∣∣,(2.67)

which, by the bound (2.58), is bounded uniformly in ξ1 and ξ2.

DRAFT

July 11, 2016

DRAFT

Page 73: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 67

DRAFT

Thus the perturbation theorem 2.5.6 gives that, for γ in a ball Bγ1 around 0 in the complex plane,

some γ1 < γ0, L(γ) has one dominating simple eigenvalue. That is, there exist λ(γ) ∈ C, v(γ) ∈ B, and

ϕ(γ) ∈ B∗, all holomorphic functions of γ, such that

(2.68) Ln(γ) = λ(γ)nv(γ)⊗ ϕ(γ) +N(γ)n,

where L(γ)v(γ) = λ(γ)v(γ), L(γ)∗ϕ(γ) = λ(γ)ϕ(γ),∣∣λ(γ)

∣∣ ≥ 1−η1, < ϕ(γ), v(γ) >= 1, and where N(γ)

is a bounded linear operator on B which depends holomorphically on γ, such that∥∥N(γ)n

∥∥ ≤ c(1−η2)n

for η2 > η1 > 0.

Now, we know that for real γ, 1 is an eigenvalue of L(γ), i.e. λ(γ) = 1 and v(γ) = 1 for real γ.

Remark 2.4.4. The bounded linear functional ϕ can be extended to the space of bounded continuous

complex-valued functions on Ω0. By the Riesz representation theorem, ϕ(f) =∫fdµγ , where µγ is the

unique L(γ)-invariant probability measure. It is easy to check that µγ << P0.

2.4.3 The Velocity

Now, for ξ ∈ Ω0, and for γ ∈ Bγ1 , call

(2.69) gγ(ξ) =∑π∈Π1

`(π)Pγ,ξ[π] = Epγ,ξ0 (T1).

For real γ, gγ(ξ) = Epγ,ξ0 (T1).

From lemma 2.2.3 and the proof of lemma 2.3.3, it is straightforward to see that gγ ∈ B. Further,

we prove the following.

Lemma 2.4.5. γ 7→ gγ is holomorphic in norm for γ ∈ Bγ0.

Proof. Call

gnγ (ξ) =∑π∈Π1`(π)=n

nPγ,ξ[π],

so that gγ =∑gnγ . Note that gnγ = nUn(γ)1, and thus it is easy to see that

∥∥∥gnγ∥∥∥ is summable, with

a bound uniform for γ ∈ Bγ0 . So then∑gnγ converges uniformly to gγ in norm. Thus we can apply

a similar argument to that at the beginning of the proof of lemma 2.4.3 using theorem 2.5.7, and so it

suffices to check that γ 7→ gnγ is holomorphic for γ ∈ Bγ0 .

Recall the definitions of Vn(γ) and Dγ,ξ[π] from the proof of lemma 2.4.3.

DRAFT

July 11, 2016

DRAFT

Page 74: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

68 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Call

(2.70) Gnγ (ξ) =∑π∈Π1`(π)=n

`(π)Pγ,ξ[π]Dγ,ξ[π].

Note that Gnγ = nVn(γ)1, and is thus in B. Then we have that

∥∥∥∥∥gnγ+h − gnγh

−Gnγ

∥∥∥∥∥ = n

∥∥∥∥∥Un(γ + h)1− Un(γ)1h

− Vn(γ)1

∥∥∥∥∥(2.71)

≤ n

∥∥∥∥∥Un(γ + h)− Un(γ)

h− Vn(γ)

∥∥∥∥∥,(2.72)

which tends to 0 as h→ 0 as shown in lemma 2.4.3.

Lemma 2.4.6. For ξ0 ∈ Ω0, n ∈ N0

L(γ)ngγ(ξ0) = E

[Epγ,ξ0

[Tn+1 − Tn

] ∣∣∣∣ ξ(z, ·) = ξ0(z, ·) for z · e1 ≤ 0

].(2.73)

Proof. For n = 0, this is trivial. For n ≥ 1, using lemma 2.4.1, we have

L(γ)ngγ(ξ0) =

∫Ω0,n

∑π∈Πn

(σe(π)(ξ0 ? η)

)Pγ,ξ0?η[π]P0,n(dη)(2.74)

=

∫Ω0,n

∑π∈Πn

∑π′∈Π1

`(π′)Pγ,σe(π)(ξ0?η)[π′]Pγ,ξ0?η[π]P0,n(dη),(2.75)

and as in (2.39), we have that Pγ,σe(π)(ξ0?η)[π′]Pγ,ξ0?η[π] = Pγ,ξ0?η[π?π′]. For a path π ∈ Πk, call τk[π] =

minj ∈ N

∣∣πj · e1 = k

. Then for π′′ = π ? π′ above, we have τn[π′′] = `(π) and τn+1 = `(π) + `(π′), so

that τn+1[π′′]− τn[π′′] = `(π′). Then we have

∫Ω0,n

∑π′′∈Πn+1

(τn+1[π′′]− τn[π′′]

)Pγ,ξ0?η[π′′]P0,n(dη)(2.76)

=

∫Ω0,n

(Eξ0?ηγ (Tn+1)− Eξ0?ηγ (Tn)

)P0,n(dη)(2.77)

= E

[Epγ,ξ0

[Tn+1 − Tn

] ∣∣∣∣ ξ(z, ·) = ξ0(z, ·) for z · e1 ≤ 0

].(2.78)

For the random walk Xn in an environment pγ,ξ, on the Pγ0 -probability 1 event that limn→∞Xn·e1 =

DRAFT

July 11, 2016

DRAFT

Page 75: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.4. Proof of Main Result 69

DRAFT

∞ ⊂ Tn <∞ ∀n ∈ N, we have that XTn/Tn has the same limit as Xn/n Pγ0 -a.s., and so

vγ · e1 = limn→∞

Xn · e1

n= limn→∞

n

TnPγ

0 − a.s.(2.79)

Then 1vγ ·e1 = limn→∞

Eγ0 (Tn)n , as uniform integrability of Tn/n follows from lemma 2.2.3. And we

can write

Eγ0 (Tn)

n= Eγ

0

1

n

n−1∑k=0

(Tk+1 − Tk)

(2.80)

=

∫Ω0

∫Ω0,∞

1

n

n−1∑k=0

Epγ,ξ?η0

[Tk+1 − Tk

]P0,∞(dη)P0(dξ)(2.81)

=

∫Ω0

1

n

n−1∑k=0

L(γ)kgγ(ξ)

P0(dξ)(2.82)

=

∫Ω0

1

n

n−1∑k=0

< ϕ(γ), gγ > +N(γ)kgγ(ξ)

P0(dξ).(2.83)

From the decay on the norm of N(γ)k in (2.68), we have that

∣∣∣∣∣∣ 1nn−1∑k=0

N(γ)kgγ(ξ)

∣∣∣∣∣∣ ≤ 1

n

−1∑k=0

∥∥∥N(γ)kgγ

∥∥∥ ≤ 1

n

n−1∑k=0

c(1− η)k∥∥gγ∥∥,(2.84)

where c and η are positive constants. And thus, taking the limit n→∞, for γ small enough,

vγ · e1 =1

< ϕ(γ), gγ >.

This is holomorphic by noting that

limh→0

1

h

(< ϕ(γ + h), gγ+h > − < ϕ(γ), gγ >

)(2.85)

= limh→0

< ϕ(γ + h),gγ+h − gγ

h> + <

ϕ(γ + h)− ϕ(γ)

h, gγ >(2.86)

and using the joint continuity of the inner product.

DRAFT

July 11, 2016

DRAFT

Page 76: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

70 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

2.5 Appendix

2.5.1 Perturbation Theory for Quasi-compact Operators on a Banach Space

We follow [15], [23], and [18].

Let (B, ‖·‖) be a Banach space, LB the space of bounded linear operators on B, and T ∈ LB. Denote

the spectral radius of a bounded lineary operator by r(T ) = limn→∞‖Tn‖1/n.

Definition 2.5.1. T is quasi-compact if there are two closed, invariant subspaces F and H such that

B = F ⊕H,

where dimF <∞, each eigenvalue of T |F has modulus r(T ), and r(T |H

)< r(T ).

We include a generalization [14] of the Ionescu Tulcea Marinescu theorem [17] theorem for establishing

quasi-compactness, Theorem XIV.3 in [15]:

Theorem 2.5.2. Let (B, ‖·‖) be a Banach space, |·| a continuous semi-norm on B and T a bounded

linear operator on B, such that

1. T (f ∈ B

∣∣ ‖f‖ ≤ 1

) is totally bounded in (B, |·|),

2. there exists a constant M such that, for f ∈ B, |Tf | ≤M |f |,

3. there exists k ∈ N and real positive numbers r and R such that r < r(T ) and

∥∥∥T kf∥∥∥ ≤ rk‖f‖+R|f |, for f ∈ B.

Then T is quasi-compact and re(T ) ≤ r.

Here re(·) denotes the essential spectral radius.

Let B∗ denote the space of bounded linear functionals on B, and for ϕ ∈ B∗, f ∈ B, call < ϕ, f >=

ϕ(f). For v ∈ B, write v ⊗ ϕ to be the bounded linear operator (v ⊗ ϕ)f =< ϕ, f > v. Denote the

adjoint of T by T ∗.

Combining propositions III.1 and lemma III.3 from [15], we have

Lemma 2.5.3. A quasi-compact operator T is of diagonal type if

supnr(T )

−n∥∥Tn∥∥ <∞.

DRAFT

July 11, 2016

DRAFT

Page 77: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

2.5. Appendix 71

DRAFT

Then there exist an integer s ≥ 1, sequences (λk)sk=1 ∈ Cs, (vk)sk=1 ∈ Bs and (ϕk)sk=1 ∈ (B∗)s, such

that, for k, l = 1, . . . , s and n ≥ 1,

1. |λk| = r(T )

2. (ϕk, vl) = δk,l

3. Tvk = λkvk, T ∗ϕk = λkϕk,

4. Tn =∑sk=1 λ

nkvk ⊗ ϕk +Nn, where r(N) < r(T ).

Remark 2.5.4. The preceding is a generalization of the Doeblin-Fortet theorem [8], which is the case

when the Banach space B is assumed to be the space of continuous bounded complex-valued functions

with bounded Lipschitz constant on a compact metric space. The bounded linear operator T is assumed

to have spectral radius 1, and satisfy hypotheses (2) and (3) of theorem 2.5.2, with M = 1 and r < 1.

Under these assumptions, the theorem holds using Arzela-Ascoli, and additionally the operator T is of

diagonal-type. See [8], [7], and [18].

Following section III.2 of [15], we give the perturbation theorem we use in this paper, along with

a definition (to ensure preserving the number of peripheral eigenvalues under a smooth perturbation).

This theorem follows from the techniques in [23].

Definition 2.5.5. T has s ∈ N dominating simple eigenvalues if there exist closed subspaces F and H

such that

1. B = F ⊕H,

2. T (F ) ⊂ F , T (H) ⊂ H,

3. dimF = s and T |F has s simple eigenvalues λk, k = 1, . . . , s,

4. r(T |H

)< min

|λk|

∣∣ k = 1, . . . , s

.

For an open ball of radius r in C, Br, let H(Br,G) denote the space of holomorphic functions from

Br to the Banach space G.

Theorem 2.5.6. Let(T (γ)

)γ∈Br0

be a collection of bounded linear operators on B such that

1. T (·) ∈ H(Br0 ,B

),

2. T (0) has s dominating simple eigenvalues and r(T (0)

)= 1.

DRAFT

July 11, 2016

DRAFT

Page 78: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

72 Chapter 2. Analyticity of the Effective Velocity for Ballistic RWRE

DRAFT

Then there exists r1 < r0 such that for γ ∈ Br1 , T (γ) has s dominating simple eigenvalues. More

precisely, there exist η2 > η1 > 0 and distinct functions λk(·) ∈ H(Br1 ,C

), vk(·) ∈ H

(Br1 ,B

), ϕ(·) ∈

H(Br1 ,B∗

), k = 1, . . . , s, N(·) ∈ H

(Br1 ,LB

)such that for γ ∈ Br1 and k, l = 1, . . . , s

1. T (γ)vk(γ) = λk(γ)vk(γ), T (γ)∗ϕk(γ) = λk(γ)ϕk(γ),

2. min∣∣λk(γ)

∣∣ ∣∣∣ k = 1, . . . , s≥ 1− η1,

3. < ϕk(γ), vl(γ) >= δk,l,

4. for all n ≥ 1, Tn(γ) =∑sk=1 λk(γ)nvk(γ)⊗ ϕk(γ) +N(γ)n, with

∥∥N(γ)n∥∥ ≤ (1− η2)n.

We also recall the following from [23], pages 139 and 152.

Theorem 2.5.7.

1. γ 7→ f(γ) ∈ B is holomorphic in norm on a domain D of the complex plane if γ 7→< ϕ, f(γ) > is

holomorphic in D for each ϕ ∈ B∗.

2. γ 7→ T (γ) ∈ LB is holomorphic in norm on a domain D of the complex plane if γ 7→< ϕ, T (γ)v >

is holomorphic in γ ∈ D for each v ∈ B and ϕ ∈ B∗.

DRAFT

July 11, 2016

DRAFT

Page 79: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

July 11, 2016

DRAFT

Bibliography

[1] Antonio Auffinger and Michael Damron. A simplified proof of the relation between scaling exponents

in first-passage percolation. The Annals of Probability, 42(3):1197–1211, sep 2011.

[2] Antonio Auffinger and Michael Damron. The scaling relation chi = 2 xi - 1 for directed polymers

in a random environment. arXiv preprint arXiv:1211.0992, page 27, nov 2012.

[3] Patrick Billingsley. Convergence of Probability Measures. John Wiley & Sons, INC, New York, 1999.

[4] E Bolthausen and A S Sznitman. Ten Lectures on Random Media. Oberwolfach Seminars.

Birkhauser Basel, 2002.

[5] Sourav Chatterjee. The universal relation between scaling exponents in first-passage percolation.

page 33, may 2011.

[6] Louis H. Y. Chen. Poisson Approximation for Dependent Trials. The Annals of Probability, 3(3):534–

545, 1975.

[7] Harry Cohn, editor. Doeblin and Modern Probability, volume 149 of Contemporary Mathematics.

American Mathematical Society, Providence, Rhode Island, 1993.

[8] W Doeblin and R Fortet. Sur des chaines a liaisons completes. Bull. Soc. Math. France, 65:132–148,

1937.

[9] Dmitry Dolgopyat and Carlangelo Liverani. Non-perturbative approach to random walk in Marko-

vian environment. Electronic Communications In Probability, 14:245–251, 2009.

[10] R Durrett. Probability: Theory and Examples. Duxbury advanced series. Thomson Brooks/Cole,

2005.

[11] Jane P. Gilman, Irwin Kra, and Rubı E. Rodrıguez. Complex Analysis, volume 245 of Graduate

Texts in Mathematics. Springer New York, New York, NY, 2007.

DRAFT

73

DRAFT

Page 80: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

74 Bibliography

DRAFT

[12] Xiaoqin Guo. Einstein relation for random walks in random environment. The Annals of Probability,

44(1):324–359, jan 2016.

[13] JM Hammersley. A Few Seedlings of Research. Proceedings of the Sixth Berkeley Symposium . . . ,

pages 345–394, 1972.

[14] H Hennion. Sur un Theoreme Spectral et son Application aux Noyaux Lipchitziens. Proc. Amer.

Math. Soc, 118(2):627–634, 1993.

[15] Hubert Hennion and Loıc Herve. Limit Theorems for Markov Chains and Stochastic Properties of

Dynamical Systems by Quasi-Compactness, volume 1766 of Lecture Notes in Mathematics. Springer

Berlin Heidelberg, Berlin, Heidelberg, 2001.

[16] W Hoeffding. Probability Inequalities for Sums of Bounded Random Variables. Journal of the

American statistical . . . , 58(301):13–30, 1963.

[17] C. T. Ionescu Tulcea and G. Marinescu. Theorie Ergodique Pour Des Classes D’Operations Non

Completement Continues. Annals of Mathematics, 52(1):140–147, 1950.

[18] Marius Iosifescu. A basic tool in mathematical chaos theory: Doeblin and Fortet’s ergodic theorem

and Ionescu Tulcea and Marinescu’s generalization. Contemporary Mathematics, 149:111–124, 1993.

[19] Kurt Johansson. Shape fluctuations and random matrices. Communications in mathematical

physics, pages 1–51, 2000.

[20] Kurt Johansson. Transversal fluctuations for increasing subsequences on the plane. Probability

theory and related fields, pages 1–16, 2000.

[21] Steven A Kalikow. Generalized Random Walk in a Random Environment. The Annals of Probability,

9(5):753–768, 1981.

[22] O. Kallenberg. Random Measures. Schriftenreihe. Akad.-Verlag, 1975.

[23] Tosio Kato. Perturbation theory for linear operators. Springer-Verlag, New York, 2 edition, 1995.

[24] J. F. C. Kingman. The ergodic theory of subadditive stochastic processes. Journal of the Royal

Statistical Society. Series B, 30(3):499–510, 1968.

[25] J. F. C. Kingman. Subadditive ergodic theory. The annals of Probability, 1(6):883–909, 1973.

[26] J. F. C. Kingman. Poisson Processes. Oxford University Press, feb 1993.

DRAFT

July 11, 2016

DRAFT

Page 81: University of Torontoblog.math.toronto.edu/GraduateBlog/files/2016/07/ut... · 2016. 7. 11. · DRAFT July 11, 2016 DRAFT Abstract Two results on Asymptotic Behaviour of Random Walks

DRAFT

Bibliography 75

DRAFT

[27] GF Lawler. Weak Convergence of a Random Walk in a Random Environment. Communications in

Mathematical Physics, 87:81–87, 1982.

[28] Hao Lin. Properties of the limit shape for some last passage growth models in random environments.

PhD thesis, UW-Madison, 2011.

[29] J R Munkres. Topology. Topology. Prentice Hall, 2000.

[30] Christophe Sabot. Ballistic random walks in random environment at low disorder. The Annals of

Probability, 32(4):2996–3023, oct 2004.

[31] Timo Seppalainen. Lecture Notes on the Corner Growth Model. 2009.

[32] Fred Solomon. Random walks in random environment. The Annals of Probability, 3, 1975.

[33] Alain-sol Sznitman. An effective criterion for ballistic behavior of random walks in random envi-

ronment. Probability Theory and Related Fields, 122(4):509–544, 2002.

[34] Alain-sol Sznitman and Martin Zerner. A Law of Large Numbers for Random Walks in Random

Environment. The Annals of Probability, 27(4):1851–1869, 1999.

DRAFT

July 11, 2016

DRAFT