geometry’s fundamental role in the stability of stochastic...

108
Geometry’s Fundamental Role in the Stability of Stochastic Differential Equations by David P. Herzog A Dissertation Submitted to the Faculty of the Department of Mathematics In Partial Fulfillment of the Requirements For the Degree of Doctor of Philosophy In the Graduate College The University of Arizona 2011

Upload: others

Post on 10-Apr-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

Geometry’s Fundamental Role in the Stability

of Stochastic Differential Equations

by

David P. Herzog

A Dissertation Submitted to the Faculty of the

Department of Mathematics

In Partial Fulfillment of the RequirementsFor the Degree of

Doctor of Philosophy

In the Graduate College

The University of Arizona

2 0 1 1

Page 2: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

2

THE UNIVERSITY OF ARIZONA

GRADUATE COLLEGE

As members of the Dissertation Committee, we certify that we have read the disser-tation prepared by David P. Herzog entitled

Geometry’s Fundamental Role in the Stability of Stochastic Differential Equations

and recommend that it be accepted as fulfilling the dissertation requirement for theDegree of Doctor of Philosophy.

Date: April 18, 2011Jan Wehr

Date: April 18, 2011Rabindra N. Bhattacharya

Date: April 18, 2011Thomas G. Kennedy

Date: April 18, 2011Joseph C. Watkins

Final approval and acceptance of this dissertation is contingent uponthe candidate’s submission of the final copies of the dissertation tothe Graduate College.

I hereby certify that I have read this dissertation prepared undermy direction and recommend that it be accepted as fulfilling thedissertation requirement.

Date: April 18, 2010Jan Wehr

Page 3: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

3

Statement by Author

This dissertation has been submitted in partial fulfillment of re-quirements for an advanced degree at The University of Arizona andis deposited in the University Library to be made available to bor-rowers under rules of the Library.

Brief quotations from this dissertation are allowable without spe-cial permission, provided that accurate acknowledgment of source ismade. Requests for permission for extended quotation from or repro-duction of this manuscript in whole or in part may be granted by thehead of the major department or the Dean of the Graduate Collegewhen in his or her judgment the proposed use of the material is in theinterests of scholarship. In all other instances, however, permissionmust be obtained from the author.

Signed: David P. Herzog

Page 4: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

4

Dedication

To Charles, Phyllis, and Brenda.

Page 5: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

5

Acknowledgments

There are many who deserve to be acknowledge for influencing me and my work. My

advisor, Professor Jan Wehr, is certainly on top of the list. Let it be said that I feel

lucky to have stumbled into a course on stochastic differential equations taught by

him. At a time when I was academically adrift, his lucid and energetic lectures re-

instilled my passion for analysis so much that I started working with him the following

semester. At that time, however, my passion was in place but my skills were not. I

am most of all thankful that my advisor allowed me to ignore my ignorance and do

mathematics research anyway.

I am grateful for having Professor Rabindra Bhattacharya, Professor Thomas

Kennedy, and Professor Joseph Watkins serve on my defense committee. They have

all met with me countless times to discuss various aspects of this and prior work.

I must further acknowledge that, due to his convenient office location, Professor

Watkins and I spoke almost every day. Although our conversations were primarily

non-mathematical, we did have quite a few useful discussions on the support theorem

[SV72], control theory, and convergence theorems relating to this work.

I would like to acknowledge fruitful conversations with both Professor Krzysztof

Gawedzki and Professor Martin Hairer. Although I have never met Professor Gawedzki

in person, it was a pleasure collaborating with him via email this past summer. I

was lucky to run into Professor Hairer in Japan last September, and I am thankful

that he pointed me to the work [JK85]. It has been a pleasure learning a sliver of

geometric control theory.

I am thankful that Professor Jerzy Zabczyk referred us to the work [Sch93]. This

paper has certainly helped us mind through the construction of several Lyapunov

functions.

Page 6: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

6

Table of Contents

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

Chapter 1. Introduction and History . . . . . . . . . . . . . . . . . . 91.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91.2. History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Chapter 2. Stability of Stochastic Differential Equations . . . 182.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.2. Absence or Presence of Explosions . . . . . . . . . . . . . . . . . . . . 19

2.2.1. Absence of Explosions . . . . . . . . . . . . . . . . . . . . . . 192.2.2. Presence of Explosions . . . . . . . . . . . . . . . . . . . . . . 25

2.3. Markov Processes and Invariant Measures . . . . . . . . . . . . . . . 292.4. Uniqueness of Invariant Probability Measures . . . . . . . . . . . . . 362.5. Geometric Control Theory . . . . . . . . . . . . . . . . . . . . . . . . 392.6. Geometric Ergodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

Chapter 3. Proof of Main Theorem . . . . . . . . . . . . . . . . . . . 613.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613.2. Lyapunov Coverings . . . . . . . . . . . . . . . . . . . . . . . . . . . 623.3. Lyapunov Regions and Scaling . . . . . . . . . . . . . . . . . . . . . . 65

3.3.1. Lyapunov Regions . . . . . . . . . . . . . . . . . . . . . . . . 653.3.2. Scaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.4. Stage 1: A (Strong) Lyapunov Covering . . . . . . . . . . . . . . . . 733.4.1. Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733.4.2. Lyapunov Functions . . . . . . . . . . . . . . . . . . . . . . . 74

3.5. Stage 2: Gluing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 823.6. Uniqueness of µ and Geometric Ergodicity . . . . . . . . . . . . . . . 903.7. Explosive Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

Chapter 4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

Page 7: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

7

List of Figures

Figure 1.1. Phase portrait for the ODE (1.1) with n = 4. The only solutionsthat are unstable in time begin in D3 = z3 > 0. The rest approach theequilibrium point z = 0 as t→∞. For general n ≥ 2, a similar picture isvalid except that the unstable solutions begin in Dn−1 = zn−1 > 0. . . 10

Figure 1.2. 105 small heavy particles for Stokes time τ = 10−2 . . . . . . . 14

Figure 3.1. Cartoon of Lemma 3.4 . . . . . . . . . . . . . . . . . . . . . . . 64Figure 3.2. The region U1 (in blue) for n = 4. The pink represents the

distance to the rays α(k) defined below. . . . . . . . . . . . . . . . . . . 66Figure 3.3. The region U2 for n = 4. It covers the rays α(k) for all k ∈ Z

and also overlaps U1 by the choice of ηn. . . . . . . . . . . . . . . . . . . 67Figure 3.4. The region U3 with n = 4. With the choice of ηn, U3 overlaps U2. 68Figure 3.5. The region U4(0) with n = 4. Although it is hard to see, there is

a tiny space that still needs to be covered. . . . . . . . . . . . . . . . . . 69Figure 3.6. The region U5(0) for n = 4 . . . . . . . . . . . . . . . . . . . . . 70

Page 8: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

8

Abstract

We study dynamical systems in the complex plane under the effect of constant noise.

We show for a wide class of polynomial equations that the ergodic property is valid

in the associated stochastic perturbation if and only if the noise added is in the

direction transversal to all unstable trajectories of the deterministic system. This has

the interpretation that noise in the “right” direction prevents the process from being

unstable: a fundamental, but not well-understood, geometric principle which seems

to underlie many other similar equations. In view of [Has80, JK85, Jur97, MT93b,

RB06, SV72], the result is proven by using Lyapunov functions and geometric control

theory.

Page 9: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

9

Chapter 1

Introduction and History

1.1 Introduction

The main purpose of this dissertation is to study dynamical systems under the effect

of noise. More precisely, we investigate families of stochastic differential equations

(SDEs) that are slight perturbations of deterministic differential equations. For fixed

n ≥ 2, the equationdz(t)

dt= (z(t))n; z(0) = z ∈ C (1.1)

is the primary focus. In particular, we find the maximal class of parameters (κ1, κ2) ∈

C2 such that the associated SDE:

dz(t) = (z(t))n dt+ κ1 dW(1)(t) + κ2 dW

(2)(t) (1.2)

has the ergodic property. In other words, we find all (κ1, κ2) ∈ C2 such that

1. For all initial conditions z ∈ C, solutions of (1.2) exist for all finite times t ≥ 0.

2. There exists a unique steady-state distribution µ to which the dynamics con-

verges in the long-time regardless of the initial condition.

It is important to point out that W (1)(t) and W (2)(t) are indeed independent

standard REAL Wiener processes defined on a probability space (Ω,F , P ). The

infinitesimals κ1 dW(1)(t) and κ2 dW

(2)(t) thus represent independent “kicks” in the

directions of κ1 and κ2, respectively. The reason we allow noise in this form is that

it will permit us to obtain and state the full results in terms of the geometry of the

deterministic system (1.1). Specifically, one has the phase portrait (see Figure 1.1)

of (1.1).

Page 10: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

10

- 3 - 2 -1 0 1 2 3

- 3

- 2

-1

0

1

2

3

Figure 1.1. Phase portrait for the ODE (1.1) with n = 4. The only solutions thatare unstable in time begin in D3 = z3 > 0. The rest approach the equilibriumpoint z = 0 as t → ∞. For general n ≥ 2, a similar picture is valid except that theunstable solutions begin in Dn−1 = zn−1 > 0.

From this, it is intuitively clear that to obtain the ergodic property, we must at

least require noise in the direction transversal to the rays

Rn−1(k) =

arg(z) =

2πk

n− 1

,

for all k ∈ Z. If, for example, κ2 = 0 and κ1 is such that κn−11 6= 0 ∈ R: for some

k ∈ Z, solutions that begin in Rn−1(k) cannot leave Rn−1(k). Thus if g is a primitive

(n−1)st root of unity, for some j ∈ Z the process x(t) := gjz(t) > 0 evolves according

to the real-valued SDE:

dx(t) = (x(t))n dt+ gjκ1 dW(1)(t), (1.3)

Page 11: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

11

which, by way of Feller’s test [Dur96], is seen to have a positive probability of reaching

infinity in finite time. Using this, we hence have a candidate for the permissible class

of (κ1, κ2) ∈ C2:

Definition 1.4. We say the pair (κ1, κ2) ∈ C2 is transversal to Dn−1 if either κ1

and κ2 are linearly independent over R or the set κn−11 , κn−1

2 contains a non-real

number.

It seems plausible that within the class of parameters transversal toDn−1, equation

(1.2) should have the ergodic property. After all, if a solution starts in the set Dn−1

such noise guarantees the process must exit. In view of the trajectory plot (Figure

1.1), the stable dynamics should then take over. We are; however, reminded the

effect noise can have on a well-behaved system. For example, it is shown in [Sch93]

that there are asymptotically stable systems in R2 such that when any amount of

constant noise is added, solutions of the stochastic perturbation starting anywhere

reach infinity in finite time almost surely. Thus the noise that initially helps the

process z(t) out of Dn−1 could in principle guide it back to Dn−1, or find an alternate

route to infinity. A partial argument in this work is that the example given in [Sch93]

is an exception, as the dynamics is tailored to specification. Outside the realm of

such examples, we suggest there are no surprises. In particular, we prove:

Theorem 1.5. For all n ≥ 2, equation (1.2) has the ergodic property if and only if

(κ1, κ2) ∈ C2 is transversal to Dn−1.

This theorem serves also as an illustration of the difference between the stability of

SDEs in one and higher dimensions. As mentioned earlier, the real-valued counterpart

equation (1.3) has solutions which reach infinity in finite time. This is primarily

because noise cannot moderate the instability by “pushing” the process off of the real

axis and onto a stable region. It therefore seems that the dimension of the instability

as a sub-manifold in the ambient space plays a fundamental role. This is exemplified in

Page 12: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

12

[BHW11] where the stochastic dynamics has a critical pair of parameters (α1, α2) ∈ R2

such that if α1 < α2, the ergodic property holds and if α2 < α1, there are solutions

which reach infinity in finite time. If α1 < α2, the deterministic system has a single

isolated unstable trajectory. When, however, α2 < α1 the unperturbed dynamics

yields an entire open sub-manifold of unstable initial conditions of the state space.

Although there appears to be a fundamental geometric principle underlying the

stability of stochastic differential equations, we are far from a general understanding

of this. For instance, to show Theorem 1.5 for the innocuous family of equations (1.2),

many careful so-called Lyapunov estimates like those performed in [GHW10, Sch93]

are required. Additionally, for certain values of (κ1, κ2) ∈ C2, deep theorems, e.g.

Hormander’s theorem [Hor67] and the support theorem [SV72], are employed. This

is not to say that general results cannot be proven; rather, it is reminder that it is easy

to go beyond the scope of existing theory, outside of which there is little guidance.

To effectively study SDEs with locally-Lipschitz coefficients like the system (1.2),

the most difficult issue to resolve is usually that of global existence. Unfortunately,

there is no known general theorem that can be immediately applied in this setting to

conclude this. There are guiding principles, however. See, for example, the classical

treatment in [Has80], or the more general prescription in the series of works [MT92,

MT93a, MT93b, MT09]. All operate under the assumption that there exists a certain

test function, called a Lyapunov function, which helps prove existence. Consequently,

we must exhibit such a function, a task easier said than done. With the system (1.2) in

mind, here we propose an algorithmic procedure to produce a Lyapunov function for

an SDE. We do not claim this is a general result; however, these methods have been

useful in many instances where existence is non-trivial [BHW11, GHW10, Sch93]. An

additional benefit of this procedure is that we are easily able to infer the existence of

a steady state distribution µ.

After moderating the above, we must settle the question of uniqueness of µ. If

κ1 and κ2 span the entire complex plane over R, uniqueness can be immediately

Page 13: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

13

concluded using classical methods from partial differential equations [Has80]. This

follows intuitively since the process defined by equation (1.2) is Markovian and is, by

the non-degeneracy of the pair (κ1, κ2), supported everywhere in C. Thus, regardless

of where the process begins, it “mixes” well-enough so that in the long-time the

dynamics is unique. If, on the other hand, κ1 = c κ2 for some c ∈ R, uniqueness

of µ no longer follows by the same methods. Using similar ideas, one can establish

uniqueness by proving smoothness of the transition measures P (z, t, · ) of z(t) via

Hormander’s theorem [Hor67] and by showing that processes originating from distinct

initial states still “mix” with perhaps less strength than before. The latter is done

by using methods from control theory via the support theorem [SV72].

Before we proceed onto the main body of this work, we first give a brief history

as to how this project originated.

1.2 History

The primary motivation of this work is to use experiences with systems such as

(1.2) to not only generate new mathematical understanding, but also apply learned

techniques to equations in order to gain insight into other scientific disciplines. With

this motivation in place, it is thus natural to begin with a specific application in

mind, as equations born here are not only interesting but also appear to exhibit a

wide range of behaviors. It is not surprising then that the family (1.2) originated in

a similar fashion, which we now describe.

To this day, fully understanding turbulence remains a challenge. One way to

attack this problem is to study how the fluid transports small particles. For example,

if the particle acts as a simple tracer of the flow, we have the following relation:

y(t) = v(t, y(t)), (1.6)

where y(t) ∈ Rd is the particle’s position at time t ≥ 0 and v is fluid velocity field.

If, however, y(t) ∈ Rd has inertia it is subject to frictional forces. In particular, y(t)

Page 14: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

14

now evolves according to the Newtonian equation:

y(t) = −1

τ(y(t)− v(t, y(t))) , (1.7)

where the constant τ > 0 is called the Stokes time. One interest in equation (1.7)

as opposed to (1.6) is the presence of spatial inhomogeneities in the distribution of

particles when the mass of the particle is much larger than that of the carrier fluid.

In particular, we have the image (Figure 1.2) courtesy of J. Bec [Bec05]. Thus the

small but heavy particles, called inertial particles, separate or cluster in an irregular

manner in the flow over time.

Figure 1.2. 105 small heavy particles for Stokes time τ = 10−2

We can capture this phenomenon by considering the dynamics of the particle

separation ρ(t) = δ y(t) which obeys the linearized equation:

ρ(t) = −1

τ[ρ(t)− (ρ · ∇)v(t, y(t))] , (1.8)

and we may assume to good approximation [BCH07]:

∇jvi(t, y(t)) = Sij(t),

Page 15: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

15

where S(t) is a matrix-valued white noise with covariance structure:

E[Sij(s)S

kl (t)

]= Dik

jl δ(s− t),

where the constants Dikjl are chosen such that the covariance is isotropic and non-

negative. Specifically, we set

Dikjl = Aδikδ

jl +B(δijδ

kl + δilδ

kj )

where A,B ∈ R are such that A ≥ |B| and A+(d+1)B ≥ 0. Under these assumptions,

equation (1.8) becomes the following linear SDE:

ρ(t) = −1

τ[ρ(t)− S(t)ρ(t)] , (1.9)

which can be interpreted using invariably the Ito or Stratonovich conventions. Writing

this in the first-order form:

ρ(t) =1

τχ(t)

χ(t) = −1

τχ(t) + S(t)ρ(t),

we study the joint process p(t) = (ρ(t), χ(t)).

To understand how particles cluster or separate over time, it is convenient to

introduce the quantity (assuming it exists):

λ = limT→∞

1

TEp(0) [ln(|p(T )|)] ,

which is called the (top) Lyapunov exponent of the process p(t). In dimensions

d ≥ 2, certain ergodic properties of p(t) are assumed in [BCH07, WM03] to vali-

date formulas for λ which are used to extract information on particle clustering. In

[GHW10], we prove these formulas using similar techniques to those described in the

previous section. Indispensable components of these arguments are the substitutions:

x(t) =ρ(t) · χ(t)

|ρ(t)|2, y(t) =

ρ(1)(t)χ(2)(t)− ρ(2)(t)χ(1)(t)

|ρ(t)|2

Page 16: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

16

in dimension d = 2 and

x(t) =ρ(t) · χ(t)

|ρ(t)|2, y(t) =

√|ρ(t)|2|χ(t)|2 − (ρ(t) · χ(t))2

|ρ(t)|2

in dimensions d ≥ 3. Using the complex variable z(t) = x(t) + iy(t), the process z(t)

evolves in C for d = 2 and in H+ = Im(z) > 0 for d ≥ 3 according to the equation:

dz(t) = −1

τ

((z(t))2 + z(t)− iτA(d− 2)

2 Im(z)

)dt (1.10)

+√A+ 2B dW (1)(t) + i

√AdW (2)(t),

where W (1)(t) and W (2)(t) are independent standard Wiener processes. When d = 2,

the term i τA(d−2)2 Im(z)

is absent from the expression above.

Using these equations, as done in [GHW10] one can effectively prove the assumed

ergodic properties in [Bec05, BCH07] by doing so for the slightly modified version of

(1.10)

dz(t) = (z(t))2 dt+ ε1 dW(1)(t) + iε2 dW

(2)(t), (1.11)

where ε1 ≥ 0 and ε2 > 0 are positive constants. Note that the above relation certainly

falls within the class of equations (1.2) with noise (κ1, κ2) transversal to D1 = R>0.

Assuming one has not seen the phase portrait of the associated ODE for n = 2,

the fact that z(t) satisfies the ergodic property is rather surprising as it has some

comparable features of its real-valued, highly unstable, counterpart equation:

dx(t) = (x(t))2 dt+ ε dW (t).

Both have coefficients which are polynomial of degree two, hence grow at infinity

relatively fast, and both are one-dimensional equations in some sense. As emphasized

before, the difference is really in the geometry of the phase portrait of the non-random

dynamical system. This is precisely why one conjectures the same stability to hold

for the family (1.2) within the class of noise (κ1, κ2) ∈ C2 transversal to Dn−1. In

this work, we provide a short testament to this.

Page 17: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

17

The layout of the dissertation is as follows. In Chapter 2, we highlight methods

that are used to infer or disprove the ergodic property for time-homogeneous stochas-

tic differential equations in Rd. It is possible to operate more generally under the

assumption that the state space is a manifold, but we prefer to use Rd since such

generality is not necessary to conclude the main results for the system (1.2). Sections

2.1-2.3 provide standard techniques, while Sections 2.4-2.5 illustrate some methods for

proving uniqueness of invariant probability measures which are more esoteric. Section

2.6 provides sufficient conditions under which one can quantify a rate of convergence

to the equilibrium µ. In Chapter 3, we prove Theorem 1.5 and note moreover that if

(κ1, κ2) is transversal to Dn−1, by the results of Section 2.6, we may also prove that

the transition measures approach the limiting distribution µ exponentially fast in the

total variation norm.

Page 18: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

18

Chapter 2

Stability of Stochastic Differential Equations

2.1 Introduction

In this dissertation, we determine if the ergodic property is valid for possibly degener-

ate stochastic differential equations. The goal of this chapter is to thus familiarize the

reader with some techniques that can be used to prove or disprove such stability, both

in this work and in general. We do not promise what follows to be a comprehensive

study; rather, we hope to illustrate methods that were useful in our efforts.

Since the results of this chapter can be applied to a wider family of equations than

(1.2), we consider the more general time-homogeneous stochastic system:

x(t)− x(0) =

∫ t

0

b(x(s)) ds+

∫ t

0

σ(x(s)) dW (s), (2.1)

which will be written equivalently using differentials as:

dx(t) = b(x(t)) dt+ σ(x(t)) dW (t).

Denoting the set of d×d matrices with real entries by Md(R), unless stated otherwise

we make the following assumptions on equation (2.1):

(A1) b : Rd → Rd, σ : Rd →Md(R) are smooth functions on their respective spaces.

(A2) W (t) = (W (1)(t),W (2)(t), . . . ,W (d)(t)) is a d-dimensional standard Wiener pro-

cess defined on a probability space (Ω,F , P ) to which we attach the filtration

Ftt≥0 where for t ≥ 0, Ft is the sigma algebra generated by (W (t) : s ≤ t).

(A3) The stochastic integral∫σ dW is defined in the Ito sense.

(A4) The initial condition x(0) = x ∈ Rd is deterministic.

Page 19: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

19

By the dimensions of b, σ, andW (t), any solution x(t) of (2.1) is a random process that

evolves in Rd. The first problem we will address is that of existence and uniqueness

of solutions of (2.1). To this end, in the following section we show how to estimate

the time in which the process x(t) leaves every bounded domain in Rd.

Before we proceed further, let us first fix some notation. We use Br(x) ⊂ Rd

to denote the open ball centered at x ∈ Rd of radius r > 0. For x ∈ Rd, let Px

be the probability law of the process x(t) determined by (2.1) with x(0) = x and

let Ex be its corresponding expectation. We use B([0,∞)) and B(Rd) to denote the

Borel sigma-algebra of subsets on [0,∞) and Rd, respectively. For A ∈ B([0,∞)),

U ∈ B(Rd), and k ∈ N, let Ck1 (A× U) be the set of functions Φ : A× U → R which

are once continuously differentiable on A and k times continuously differentiable on

U , let Ck0 (U) denote the set of functions Φ : U → R which are k times continuously

differentiable on U and compactly supported in U , and let Ck(U) be the set functions

Φ : U → R which are k times continuously differentiable on U .

2.2 Absence or Presence of Explosions

2.2.1 Absence of Explosions

In order to prove the ergodic property holds, one must first show that solutions exist

regardless of the initial point x ∈ Rd, and are unique in some sense, for all finite

times t ≥ 0. To this effect, there is a general existence and uniqueness theorem for

stochastic differential equations which we state without proof.

Theorem 2.2 (Existence and Uniqueness). Let b and σ satisfy the following addi-

tional condition:

(B1) |b(x)− b(y)|+ |σ(x)− σ(y)| ≤ C|x− y|; x, y ∈ Rd

for some positive constant C. Then for every x(0) = x ∈ Rd, there exists an al-

most surely continuous solution x(t) of equation (2.1) which is defined for all finite

Page 20: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

20

times t ≥ 0. Morover, x(t) is adapted to the filtration Ftt≥0 and unique up to

equivalence.1

Proof. See [Øks03].

Although interesting, as stated this theorem certainly cannot be applied in the

general setting (2.1) since the bound (B1) is not necessarily satisfied. This is the

case for our system (1.2) and many other systems where one expects global stability.

For a simple example, consider the real-valued stochastic differential equation:

dx(t) = −(x(t))3 + dW (t). (2.3)

It is clear that for every initial condition x(0) = x 6= 0 ∈ R the dynamics, on average,

is directed inward towards the origin, yet b(x) = −x3 is not globally Lipschitz. In

his influential work [Has80], Has′minskiı stresses that “It is therefore of paramount

importance to find other, broader conditions for the existence and uniqueness of

the solution of equation [(2.1)]” [Has80]. In the very same text, he provided such

conditions that were phrased in terms of test functions and these methods are still

widely used to gauge the stability of diffusion processes such as x(t). See, for example,

[EM01, GHW10, MSH02, MT93b, Sch93]. We now outline Has′minskiı’s approach

since it serves us well throughout this work.

Under assumptions (A1)-(A4), one can always define continuous solutions x(t)

of equation (2.1) until the (random) time in which the process leaves every bounded

domain in Rd [Has80]. To see this formally, for n ∈ N choose smooth functions b(n)

and σ(n) on Rd such that on Bn(0):

b(n)(x) = b(x)

σ(n)(x) = σ(x),

1Two solutions x1(t) and x2(t) of equation (2.1) are equivalent if P x1(t) = x2(t) for all t ≥ 0 =1.

Page 21: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

21

and b(n) and σ(n) satisfy (B1). Thus by Theorem 2.2, for each fixed n ∈ N, for all

initial conditions x ∈ Rd there exists an almost surely continuous process x(n)(t) such

that:

dx(n)(t) = b(n)(x(n)(t)) dt+ σ(n)(x(n)(t)) dW (t),

for all t ≥ 0. Moreover, x(n)(t) is unique up to equivalence and adapted to the Wiener

filtration Ft. For m,n ∈ N, we define stopping times

ξ(m)n = inf

t>0

x(m)(t) ∈ Bn(0)c

,

and one can show (see [Dyn65]) that for m,m′ ≥ n, x ∈ Rd:

ξ(m)n = ξ(m′)

n Px - a.s.

Thus for n ∈ N let ξn = ξ(n)n . Moreover it follows that for m,m′ ≥ n and x ∈ Rd:

Px

supt≥0|x(m)(t ∧ ξn)− x(m′)(t ∧ ξn)| = 0

= 1.

Thus we may define a process x(t) by:

x(τ) = x(n)(τ), whenever τ < ξn

and we note that for all n ∈ N the equation:

dx(t ∧ ξn) = b(x(t ∧ ξn)) dt+ σ(x(t ∧ ξn)) dW (t)

holds. Let ξ be the increasing limit of ξn as n→∞. If we can prove that

Px ξ <∞ = 0

for all x ∈ Rd, then for all initial conditions x ∈ Rd we have a unique solution x(t) of

equation (2.1) which is defined and continuous for all finite times t ≥ 0. Moreover,

x(t) is adapted to Ftt≥0. To this end, our goal is to estimate the time ξ, called the

explosion time of the process x(t).

Page 22: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

22

To prove ergodicity, we hope that for all x ∈ Rd:

Px ξ <∞ = 0;

in which case, we call the process x(t) non-explosive. To prove x(t) is non-explosive,

we require the following lemma due to Dynkin.

Lemma 2.4 (Dynkin’s Formula). Let Φ ∈ C21([0,∞)×Rd) and ξn(t) = t∧ ξn. Then

for t ≥ 0 and x ∈ Rd we have:

Ex [Φ(ξn(t), x(ξn(t)))]− Φ(0, x) = Ex

[∫ ξn(t)

0

LΦ(u, x(u)) du

](2.5)

where

LΦ(t, x) =∂Φ(t, x)

∂t+

d∑i=1

b(i)(x)∂Φ(t, x)

∂x(i)+

1

2

d∑i,j=1

(σσ∗)(ij)(x)∂2Φ(t, x)

∂x(i)∂x(j). (2.6)

Proof. Using the discussion above, we may apply Ito’s lemma to the process Φ(t, x(t))

to obtain

Φ(ξn(t), x(ξn(t)))−Φ(0, x(0)) =

∫ ξn(t)

0

LΦ(u, x(u)) du+ bounded martingale. (2.7)

Since the martingale starts at 0, after taking expectations Ex of both sides of equation

(2.7) we have the result.

The above relation (2.5) is perhaps the most beautiful in all of stochastic dif-

ferential equations. On the left-hand side, we have a random process x(t). On the

right-hand side, we have a partial differential operator L. Such a relation provides

just a hint of the intimate connection between the probabilistic theory of stochastic

differential equations and the classical theory of partial differential equations. As we

shall see, both points of view provide equally-valuable insights into the other.

Because the operator L plays a fundamental role in this work, we adopt the

common nomenclature and call it the generator of the process x(t). The choice of

Page 23: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

23

this terminology will become more transparent in the next section when we discuss

Markov processes.

We now use Lemma 2.4 as a means by which to verify existence and uniqueness

in (2.1) when (B1) fails. The intuition behind what follows is to insert a suitable

function Φ into equation (2.5). For simplicity, assume that Φ := Φ(x) ∈ C2(Rd) is

only a function of the spatial variables. To utilize relation (2.5), Φ should approach

infinity as |x| → ∞, so, without loss of generality, we can assume Φ ≥ 0. This often

called “norm-like” property is to assure that Φ hits infinity when the process x(t)

does. If one can then control the right-hand side of (2.5) using properties of LΦ(x),

non-explosivity of the process x(t) should follow.

Theorem 2.8. Let Φ ∈ C2(Rd) be a non-negative function and suppose

Φ(x)→∞ as |x| → ∞,

and there exist positive constants C,D such that

LΦ(x) ≤ CΦ(x) +D for all x ∈ Rd.

Then the process x(t) is non-explosive, i.e., for all x(0) = x ∈ Rd, Px ξ <∞ = 0.

Proof. Let x(0) = x ∈ Rd and define a function Ψ(t, x) = e−Ct (Φ(x) +D/C).

Choose N ∈ N sufficiently large so that Φ(y) ≥ 1 for all |y| ≥ N . By Lemma 2.4, we

have for all n ≥ N :

Ex [Ψ(ξn(t)), x(ξn(t)))]−Ψ(0, x) = Ex

[∫ ξn(t)

0

LΨ(u, x(u)) du

]

= Ex

[∫ ξn(t)

0

−CΨ(u, x(u)) + e−CtLΦ(x(u)) du

]

≤ Ex

[∫ ξn(t)

0

−CΨ(u, x(u)) + CΨ(u, x(u)) du

]= 0.

Page 24: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

24

Thus

Ex [Ψ(ξn(t)), x(ξn(t)))] ≤ Ψ(0, x). (2.9)

Estimating the left-hand side of (2.9), we see that

Ex [Ψ(ξn(t)), x(ξn(t)))] ≥ Ex[1ξn≤tΨ(ξn(t)), x(ξn(t))

]≥ e−Ct inf

|y|=nΦ(y)Px ξn ≤ t .

Thus we have for n ≥ N

Px ξn ≤ t ≤ eCtΨ(0, x)

inf |y|=n Φ(y).

Letting n → ∞, we have that Px ξ ≤ t = 0 for all t ≥ 0. Hence Px ξ <∞ = 0.

Since x(0) = x ∈ Rd was arbitrary, this finishes the proof.

Let us return to equation (2.3). Recall that the coefficients do not satisfy the

assumptions of Theorem 2.2, but it is intuitively clear that the resulting process

is non-explosive. Let us verify this rigorously using the previous theorem. Define

Φ(x) = x2 and note that L = ∂t − x3∂x + 12∂xx. Thus Φ(x)→∞ as |x| → ∞ and

LΦ(x) = −2x4 + 1 (2.10)

≤ x2 + 1 (2.11)

= 1 · Φ(x) + 1.

Hence, by Theorem 2.8, the process defined by (2.3) is non-explosive.

This theorem, although clearly useful, does not tell the whole story. First, no

where does it instruct one on how to obtain the function Φ. It was very simple to find

one for the system (2.3); but as the dynamics of the general equation (2.1) becomes

increasingly complex, so does discovering Φ. Second: from equation (2.10) to (2.11)

we have thrown away valuable information about the solution. In taking a careful

look at relation (2.5) with Φ(x) = x2 and the bound (2.10), we see that, in essence,

Page 25: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

25

the process x(t) should return quickly to a large ball about the origin. As we shall

see, this will be of central importance in proving the existence of and convergence to

a steady state since any solution x(t) of equation (2.1) is a strong Markov process.

2.2.2 Presence of Explosions

In the previous subsection, we provided a means by which to verify the process x(t)

is non-explosive. Here, we do the contrary, i.e., we find sufficient conditions to prove:

Px ξ <∞ > 0 for some x ∈ Rd. (2.12)

If relation (2.12) is satisfied we say the process x(t) is explosive. In a similar manner

to before, one can prove explosivity of x(t) using test functions.

Theorem 2.13. Suppose that Φ ∈ C2(Rd) is a bounded non-negative function such

that

LΦ(x) ≥ CΦ(x) for all x ∈ Rd,

for some C > 0. Then for all x0 ∈ Rd such that Φ(x0) > 0, we have:

Px0 ξ <∞ > 0. (2.14)

Remark 2.15. We can actually be more specific than the estimate (2.14) and prove

for all ε > 0

P

ξ(x0) <

1

Cln

(supy∈Rd Φ(y)

Φ(x0)

)+ ε

> 0,

whenever Φ(x0) > 0. Here, we recall ξ(x0) is the explosion time for the process x(t)

with x(0) = x0.

Proof. Let x0 ∈ Rd be such that Φ(x0) > 0. Upon setting Ψ(t, x) = e−CtΦ(x),

Dynkin’s formula implies:

Ex0 [Ψ(ξn(t), x(ξn(t)))] ≥ Φ(x0).

Page 26: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

26

Estimating the left-hand side above, we have:

Ex0 [Ψ(ξn(t), x(ξn(t)))] ≤ supy∈Rd

Φ(y)Ex0

[e−Cξn(t)

],

hence:

E[e−Cξ

(x0)n (t)

]≥ Φ(x0)

supy∈Rd Φ(y).

Since the bound above holds for all n, by the dominated convergence theorem we

have

E[e−Cξ

(x0)(t)]≥ Φ(x0)

supy∈Rd Φ(y). (2.16)

If for some ε > 0,

P

ξ(x0) <

1

Cln

(supy∈Rd Φ(y)

Φ(x0)

)+ ε

= 0,

by splitting the left-hand side of the bound (2.16) into:

E[e−Cξ

(x0)(t)]

= E[1t≤ξx0e

−Ct]+ E[1t>ξ(x0)e

−Cξx0]

and setting

t =1

Cln

(supy∈Rd Φ(y)

Φ(x0)

)+ ε,

we violate the bound (2.16). Note that this finishes the proof.

This is a very interesting theorem, but it displays a similar weakness to Theorem

2.8. Indeed, there is no instruction manual one can follow to produce a test function

Φ ∈ C2(Rd) which is non-negative and bounded such that

LΦ(x) ≥ CΦ(x) for all x ∈ Rd,

for some C > 0. In [BHW11], we were able to exhibit such a function, but it requires

both the correct intuition and challenging estimates.

Let us now consider equation (1.2) in the case when (κ1, κ2) ∈ C2 is not transversal

to Dn−1. We recall this implies for some k ∈ Z, solutions of (1.2) starting in Rn−1(k),

Page 27: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

27

when rotated by an appropriate (n− 1)st root of unity, evolve on R>0. Determining

explosion in such a case is far simpler than constructing a test function like Φ above.

This is because in R we have Feller’s test which we now discuss.

For the remainder of this subsection, we assume x(t) defined by equation (2.1)

evolves on R>0 such that there exists ε > 0 such that σ2(x) ≥ ε for all x ≥ 0. We will

illustrate how one proves:

Px ξ <∞ > 0 for some x > 0. (2.17)

Instead of working directly with x(t), we define a strictly increasing function

φ(x) =

∫ x

0

exp

(∫ y

0

−2b(y)

σ2(y)dy

)dy

and use the induced process y(t) given by:

y(t) = φ(x(t)).

The benefit of working with y(t) is seen by noting:

Lφ(x) = b(x)φ′(x) +σ2(x)

2φ′′(x) = 0.

In particular by Ito’s formula, the process y(t) is a local martingale. For c > 0,

letting:

ξc+ = inft>0x(t) > c

ξc+ = inft>0y(t) > c,

ξ0 = inft>0x(t) = 0

ξ0 = inft>0y(t) = 0,

and

ξ+ = limc↑∞

ξc+

ξ+ = limc↑∞

ξc+,

we have:

Page 28: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

28

Theorem 2.18 (Feller’s Test). If ψ(x) = 1/(φ′(x)σ2(x)):

(a) Px ξc+ < ξ0 > 0 for x ∈ (0, c) if∫ c

0

ψ(x) (φ(c)− φ(x)) dx <∞.

(b) Px ξ+ < ξ0 > 0 for x ∈ (0,∞) if∫ ∞0

ψ(x) (φ(∞)− φ(x)) dx <∞.

Consequently,

Corollary 2.19. Pxξ <∞ > 0 for x ∈ (0,∞) if∫ ∞0

ψ(x) (φ(∞)− φ(x)) dx <∞. (2.20)

Proof of Corollary 2.20. Note that this follows immediately from (b) since for

x ∈ (0,∞) we have:

Px ξ <∞ ≥ Px ξ+ < ξ0 > 0.

Proof of Theorem 2.18. We shall only prove (a) as (b) follows similarly. To prove

(a), let c > 0 and fix x ∈ (0, c). We claim that it suffices to show that

f(x) =

∫ φ−1(x)

0

ψ(y) (x− φ(y)) dy

is a strictly increasing function on (0, φ(c)) such that

g(t, y(t)) = e−tf(y(t))

is a local martingale. To see this, suppose that f(φ(c)) < ∞. Let yn ≥ y = φ(x) be

such that yn ↑ φ(c) as n→∞. Define τn = ξ0 ∧ ξyn+. We then have:

0 < f(y) = Ey [g(τn, y(τn)))]

≤ f(φ(c))Ey

[1ξyn+<ξ0e

−ξyn+

]↓ f(φ(c))Ey

[1ξφ(c)+<ξ0e

−ξφ(c)+

],

Page 29: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

29

as n → ∞ which clearly implies that 0 < Py

ξφ(c)+ < ξ0

= Px ξc+ < ξ0, as

required. Thus we have left to show that f(x) is strictly increasing on (0, φ(c)) such

that g(t, y(t)) is a local martingale. For this we refer the reader to pages 216-218 in

[Dur96].

Hence to prove that the ergodic property is not satisfied when (κ1, κ2) ∈ C is

not transversal to Dn−1, we will use part (b) of Feller’s test to conclude the process

obtained by rotating by an appropriate (n − 1)st root of unity is explosive. The

benefit of this method is that the issue of explosion reduces to proving the integral

in (b) is finite, which is far simpler than constructing a test function on R2 like the

one in Theorem 2.13.

2.3 Markov Processes and Invariant Measures

Perhaps one of the most important properties of solutions of stochastic differential

equations is that they are strongly Markovian. That is; imprecisely, the future only

depends on the past through the present instant in time, even when one considers

bounded stopping times. For this section, we assume that the solution x(t) of (2.1)

is non-explosive and note the following two lemmata:

Lemma 2.21 (Markov Property). Let Φ : Rd → R be a bounded, Borel measurable

function. We have for s, t ≥ 0:

Ex [Φ(x(s+ t)) | Fs] = Ey [Φ(x(t))] |y=x(s).

Proof. See [Øks03].

Lemma 2.22 (Strong Markov Property). Let Φ : Rd → R be a bounded, Borel

measurable function and υ be an almost surely bounded stopping time with respect to

Ftt≥0. Then for t ≥ 0

Ex [Φ(x(υ + t)) | Fυ] = Ey [Φ(x(t))] |y=x(υ)

Page 30: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

30

where Fυ is the sigma algebra generated by (W (s ∧ υ), s ≥ 0).

Proof. See [Øks03].

Thus, using the words of McKean, the process x(t) “begins afresh” at bounded stop-

ping times υ [McK05].

There is a whole theory of Markov processes outside of the realm of stochastic

differential equations. Although extremely important for this work, we will not delve

too deeply into this well-established area; rather, we will view the solution x(t) of

equation (2.1) as a process that has the (strong) Markov property and use this in-

formation to study x(t). At times during this dissertation, however, the reader will

notice that one can operate purely under the assumption that the process x(t) is

Markovian.

Pedagogically, introducing the Markov property as stated above is probably not

the best way to initially think of such processes. Often, it is much more informative

to define transition kernels

P (x, t, A) := Px x(t) ∈ A , (2.23)

for x ∈ Rd, t ≥ 0, and A ∈ B(Rd) where B(Rd) denotes the Borel sigma algebra

of subsets on Rd. We note that, by the Markov property, we have the so-called

Chapman-Kolmogorov equations:

P (x, s+ t, A) =

∫RdP (x, s, dy)P (y, t, A), (2.24)

for all s, t ≥ 0, A ∈ B(Rd). We will often use relation (2.24) and not Lemma 2.22.

The Chapman-Kolmogorov equations coupled with the use of transition kernels

help connect probabilistic notions to functional analysis. To see this, we define for

t ≥ 0 operators Pt which act on bounded measurable functions Φ : Rd → R and on

Page 31: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

31

finite Borel measures µ on Rd in the following way:

PtΦ(x) =

∫P (x, t, dy)Φ(y)

µPt(A) =

∫µ(dy)P (y, t, A), A ∈ B(Rd). (2.25)

It occurs regularly that the above can be defined for a more general class of functions

and/or measures; in which case, we interpret the expressions PtΦ and µPt in the same

way. By (2.24), the family Ptt≥0 forms a semigroup on B(Rd), the set of bounded

Borel measurable functions, and a (dual) semigroup on finite Borel measures. We

now record some of its properties which we will use later without further comment.

Proposition 2.26. Let Φ(x) be a bounded Borel measurable function. We have:

1. PtΦ(x) ≥ 0 if Φ(x) ≥ 0.

2. PtC = C for constants C.

Proof. This is an easy exercise.

We now consider the quantity:

LΦ(x) = limt↓0

PtΦ(x)− Φ(x)

t.

The set of functions Φ : Rd → R for which the limit exists point-wise on Rd is called

the domain of L. We denote the domain of L by DL.

Definition 2.27. L is called the generator of the Markov process x(t).

In the previous section, we were sloppy in calling L the generator of x(t). L is

more like an special version of L, as we see by the following.

Proposition 2.28. DL ⊃ C20(Rd). Moreover, for Φ ∈ C2

0(Rd) we have

LΦ(x) = LΦ(x),

where L is as before.

Page 32: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

32

Proof. This follows from Dynkin’s formula and the fact that x(t) is non-explosive.

As we shall see, even though Φ may not be in the domain of L, we can still relate

LΦ(x) with PtΦ(x) is some fashion. This will be useful later when we extract a

convergence rate to equilibrium. For now, however, we discuss what we have been

referring to as the “long-time” behavior of the process x(t).

We ideally hope there exists a Borel probability measure µ such that:

limt→∞

P (x, t, A) = µ(A), (2.29)

for all x ∈ Rd, A ∈ B(Rd). This would be perfect as we see that (1) there is a

limiting distribution and (2) this distribution does not depend on the starting point

x(0) = x ∈ Rd. As noted above, this is ideal since (1) a limiting distribution may not

exist and (2) even if one does exist, there may be many.

For practical purposes, proving the limit above exists is, in general, difficult. If

one yields the strength of the limit (2.29) and proves the Cesaro mean

1

t

∫ t

0

P (x, s, · ) ds (2.30)

has limit points in the weak topology as t→∞, we can still recover, on average, the

long-time behavior of the process.

As one should expect, there is a general property that these limits have in common.

For illustrative purposes, suppose that there exists a sequence of times tn ↑ ∞ as

n→∞ such that for some x ∈ Rd the sequence of measures:

1

tn

∫ tn

0

P (x, s, · ) ds

converges weakly to some Borel probability measure µx. Suppose moreover that Pt

is weak Feller, i.e., Pt maps bounded continuous functions to bounded continuous

Page 33: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

33

functions. We then have for all bounded continuous functions f on Rd:∫f(y)µxPt(y) =

∫f(y)

∫µx(dy)P (y, t, dy)

=

∫Ptf(y)µx(dy)

= limn→∞

1

tn

∫Ptf(y)

∫ tn

0

P (x, s, dy) ds

= limn→∞

1

tn

∫ tn

0

Ps+tf(x) ds

= limn→∞

1

tn

[∫ tn

0

Puf(x) du+

∫ tn+t

tn

Puf(x) du−∫ t

0

Puf(x) du

]=

∫f(y)µx(dy).

From this, we infer µxPt(A) = µx(A) for all A ∈ B(Rd), i.e., µx is invariant under the

action of the semigroup Ptt≥0. This leads us to the following definition.

Definition 2.31. Let µ be a Borel measure such that for all t ≥ 0

µPt = µ.

We call µ an invariant measure. If µ(Rd) < ∞, then it can be normalized to a

probability measure ν which also has νPt = ν. We call ν an invariant probability

measure.

An invariant probability measure is precisely the notion of limiting behavior we

desire. We now give a necessary and sufficient condition for the existence of such a

measure.

Theorem 2.32. Suppose that Pt is weak Feller. Then there exists an invariant prob-

ability measure if and only if for some x ∈ Rd:

limr→∞

lim inft→∞

1

t

∫ t

0

P (x, s, Br(0)c) ds = 0. (2.33)

Page 34: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

34

Proof. Let µ be an invariant probability measure. Suppose that for all x ∈ Rd we

have:

limr→∞

lim inft→∞

1

t

∫ t

0

P (x, s, Br(0)c) ds = p(x) > 0.

Then by Tonelli,

0 = limr→∞

lim inft→∞

1

t

∫ t

0

µ(Br(0)c) ds

= limr→∞

lim inft→∞

1

t

∫ t

0

∫µ(dy)P (y, s, Br(0)c) ds

=

∫µ(dy) lim

r→∞lim inft→∞

1

t

∫ t

0

P (y, s, Br(0)c) ds

=

∫µ(dy)p(y) > 0.

This proves one direction. Suppose conversely that for some x ∈ Rd we have

limr→∞

lim inft→∞

1

t

∫ t

0

P (x, s, Br(0)c) ds = 0.

Note that this implies there exists a sequence of times tn ↑ ∞ as n→∞ such that:

limr→∞

limn→∞

1

tn

∫ tn

0

P (x, s, Br(0)c) ds = 0,

where the limit as r →∞ is uniform in n. Thus the sequence of probability measures:

νn( · ) =1

tn

∫ tn

0

P (x, s, · ) ds,

is tight. By Prokhorov’s theorem, νn( · ) is weakly compact and hence has a subse-

quence νnk that converges to some probability measure ν. By the same computation

above, we see that νPt = ν. This finishes the proof.

It follows that under the assumptions (A1)-(A4) the semi-group Pt is weak Feller

[Has80]. We thus want to verify condition (2.33) for all x ∈ Rd. We will use a test

function to do this.

Page 35: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

35

Theorem 2.34. Let Φ ∈ C2(Rd) be a non-negative function such that

LΦ(x)→ −∞ as |x| → ∞.

Then condition (2.33) is satisfied for all x ∈ Rd and hence, there exists an invariant

probability measure for Ptt≥0.

Proof. Let x ∈ Rd. Note that by Lemma 2.4 we have:

Ex [Φ(x(ξn(t)))]− Φ(x) = Ex

[∫ ξn(t)

0

LΦ(x(s)) ds

](2.35)

Moreover, for r > 0 sufficiently large:

LΦ(x(s)) ≤ sup|x|>r

LΦ(x) · 1|x(s)|>r + supx∈Rd

LΦ(x)

≤ −cr · 1|x(s)|>r + d, (2.36)

for some cr, d > 0 such that cr → ∞ as r → ∞. Combining (2.35) with (2.36) we

obtain:

crEx

[∫ ξn(t)

0

1x(s)|>r ds

]≤ d · t+ Φ(x).

By nonexplosivity of the process x(t), ξn(t)→ t as n→∞ almost surely. Thus

1

t

∫ t

0

P (x, s, Br(0)c) ds ≤ d

cr+

Φ(x)

crt.

This implies the result after taking the lim inf as t→∞ and then taking r →∞.

Remark 2.37. We emphasize that Theorem 2.34 assumes the process x(t) is non-

explosive. One can verify the hypotheses of both Theorem 2.8 and Theorem 2.34

simultaneously by proving there exists a smooth function Φ : Rd → [0,∞) such that

(C1) Φ(x)→∞ as |x| → ∞,

(C2) LΦ(x)→ −∞ as |x| → ∞.

Page 36: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

36

We will do this later for the system (1.2), except that we prove a more explicit

form of (C2). This form is only needed to prove exponential convergence to the

invariant probability measure. Before diving into this, we first handle the question of

uniqueness.

2.4 Uniqueness of Invariant Probability Measures

In the previous section, we introduced invariant probability measures and discussed

how they help describe the process x(t) for large times t > 0. By the proofs of

Theorem 2.32 and Theorem 2.34, if we exhibit a smooth test function Φ : Rd → [0,∞)

such that (C1) and (C2) are satisfied: for all x ∈ Rd, the sequence of measures

µT ( · ) :=1

T

∫ T

0

P (x, s, · ) ds, (2.38)

has limit points in the weak topology as T → ∞. These limit points are invariant

probability measures and, since they possibly depend on the initial condition x(0) =

x ∈ Rd, there could be many such measures. Using the Markov property, we find

sufficient conditions to prove there is only one.

In sole pursuit of uniqueness, assume throughout this section that the process x(t)

is non-explosive and has invariant probability measures. We let M denote the set

of all such measures. It is easy to see that M is convex and consequently, to prove

M = µ it is enough to showM has only one extremal point2. This follows from

Choquet’s Theorem [Cho69] which asserts that if ν ∈ M, there exists a probability

measure aν supported in the extremal points E(M) of M such that

ν =

∫E(M)

µ daν(µ).

Using this, we focus our discussion on extremal points ofM which we call extremal

invariant probability measures. To conclude that E(M) = µ, we require the

following:

2A point µ in a convex set M is extremal if whenever µ = λµ1 + (1 − λ)µ2 for some λ ∈ (0, 1)and µ1, µ2 ∈M, then µ1 = µ2 = µ.

Page 37: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

37

Theorem 2.39. Distinct extremal invariant probability measures are mutually sin-

gular.

Proof. This follows from Birkoff’s Ergodic Theorem. For a discussion in our context,

see Theorem 5.1 of M. Hairer’s notes [Hai08].

From this, we easily obtain:

Corollary 2.40. Suppose that for all µ1, µ2 ∈ E(M) we have

supp(µ1) ∩ supp(µ2) 6= ∅.

Then M = µ, i.e., there is one and only one invariant probability measure.

To prove uniqueness later on, we will use the corollary above. To gain traction on

the supports of extremal invariant probability measures, we will use ideas in [AK87].

The key ingredient in this work is the use of the support theorem [SV72] which

provides an intimate and accessible connection between the process x(t) and control

theory. More precisely, let us assume that the generator L of the process x(t) can be

written in the form:

L =∂

∂t+X0 +

r∑j=1

X2j ,

where Xj is a smooth vector field on Rd for all j = 0, 1, . . . , r. Consider the family of

ordinary differential equations:

dx(t)

dt= X0(x(t)) +

r∑j=1

uj(t)Xj(x(t)), (2.41)

where for all j = 1, 2, . . . , r, uj : [0,∞) → R is a piecewise constant mapping with

at most finitely many discontinuities. We call such u(t) = (u1(t), u2(t), . . . , ur(t))

admissible controls and we denote the class of all admissible controls by U . For

u ∈ U fixed, let ϕ(x, u, t) be the maximal right solution of equation (2.41) passing

Page 38: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

38

through x at time t = 0. Define for x ∈ Rd, T > 0 the sets:

A(x, T ) =⋃u∈U

y ∈ Rd : ϕ(x, u, T ) = y

A(x,≤ T ) =

⋃0<t≤T

A(x, t)

A+(x) =⋃t>0

A(x, t).

Thus, in words, A(x, T ), A(x,≤ T ), and A+(x) are respectively the accessible points

starting from x ∈ Rd through the trajectories (2.41) for all u ∈ U at exactly time

t = T , some time 0 < t ≤ T , and some positive time t > 0. It is useful to note that

for T > 0, the A(x,≤ T ) are nested, while the A(x, T ) need not be.

In view of [SV72], by non-explosivity of the process x(t) with x(0) = x ∈ Rd

defined by equation (2.1), we can determine its support by studying the accessibility

sets A(x, T ).

Theorem 2.42 (Stroock-Varadhan 1972). For all x ∈ Rd and T > 0:

supp(P (x, T, · )) = A(x, T ).

Proof. See [SV72].

In the spirit of the previous result, Arnold and Kliemann [AK87] introduce the

notion of an invariant control set, defined below, to find an expression for supp(µ)

where µ ∈ E(M) in terms of the positive orbits A+(x).

Definition 2.43. A set C 6= ∅ ⊂ Rd is an invariant control set for the system

(2.41) if:

A+(x) = C for all x ∈ C,

and C is maximal with respect to inclusion.

By Proposition 1.1 of [AK87], for all µ ∈ E(M), there exists an invariant control

set C such that

supp(µ) = C.

Page 39: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

39

Thus:

Corollary 2.44. For all µ ∈ E(M), there exists x ∈ Rd such that

supp(µ) = A+(x).

Remark 2.45. To verify the hypotheses of Corollary 2.40 and hence prove uniqueness

of the invariant probability measure, it is enough to show that for all x, y ∈ Rd:

A+(x) ∩ A+(y) 6= ∅.

The benefit of using the above is that it provides the means to use techniques from

control theory which are extremely tractable. Moreover, we do not really need to

know much about the sets A+(x). It many cases, however, it is still difficult to

uncover these sets. As we shall see in the next section, geometric ideas are fruitful in

this regard.

2.5 Geometric Control Theory

In hopes of understanding the accessibility sets A(x, T ), A(x,≤ T ), and A+(x) for the

control system (2.41), we find it convenient to use geometric control theory. Most of

what follows is in the book by Jurdjevic [Jur97] as well as his joint works with Kupka

[JK81, JK85]. By Remark 2.45, to conclude uniqueness of the invariant probability

measure, it suffices to show that for all x, y ∈ Rd, A+(x) ∩ A+(y) 6= ∅. However, we

aim at extracting a convergence rate to this equilibrium. Consequently, we will need a

solid grasp on the sets A(x,≤ T ) and A(x, T ) for x ∈ Rd and T > 0 as well. To mesh

with the geometrical setting in these works, we will slightly adjust some previously

used notation. We begin with a definition.

Definition 2.46. Let F be a collection of smooth vector fields on Rd. We call F a

polysystem.

Page 40: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

40

Let F be a polysystem. For Y ∈ F , let exp(tY )(x) denote the maximal right inte-

gral curve of Y passing through x ∈ Rd at time t = 0. For T > 0, we define AF (x, T )

to be the set of all points y ∈ Rd such that there exist vector fields Y1, Y2, . . . , Yk ∈ F

and corresponding times t1, t2, . . . , tk ≥ 0 such that t1 + t2 + · · ·+ tk = T and

y = exp(tkYk) exp(tk−1Yk−1) · · · exp(t1Y1)(x).

For x ∈ Rd and T > 0, let

AF (x,≤ T ) =⋃

0<t≤T

AF (x, t)

A+F (x) =

⋃t>0

AF (x, t),

Relating this setup with the control system (2.41), we note that, by definition, if we

let

F =

X0 +

r∑j=1

ujXj : u = (u1, u2, . . . , ur) ∈ Rr

,

then for all x ∈ Rd and T > 0:

AF (x, T ) = A(x, T ),

AF (x,≤ T ) = A(x,≤ T ),

A+F (x) = A+(x),

where A(x, T ), A(x,≤ T ) and A+(x) were defined in the previous section. The benefit

of using geometric ideas is it allows us to modify the polysystem F without changing

the accessibility sets too much. With this in mind, we introduce an equivalence

relation ∼ on polysystems.

Definition 2.47. Two polysystems F1 and F2 are equivalent, denoted by F1 ∼ F2,

if for all x ∈ Rd, T > 0

AF1(x,≤ T ) = AF2(x,≤ T ).

Page 41: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

41

It is easy to see that ∼ is an equivalence relation. Starting from an initial polysys-

tem F , the idea is to find operations on this family of vector fields such that, when

performed, we stay within the class of equivalent polysystems. To this end, we have

the following theorem.

Theorem 2.48. Suppose that F, F1, and F2 are polysystems such that F ∼ F1 and

F ∼ F2. Then F ∼ (F1 ∪ F2).

Before we prove the theorem, we need an important proposition which is a conse-

quence of the existence, uniqueness, and smoothness theorem of ordinary differential

equations. For further information, we refer the reader to John M. Lee’s book [Lee03].

Proposition 2.49. Let U ⊂ Rd and F a polysystem. For T > 0 define

AF (U,≤ T ) =⋃x∈U

AF (x,≤ T ).

Then for all x ∈ Rd and S, T > 0

AF (AF (x,≤ S),≤ T ) ⊂ AF (x,≤ S + T ).

Proof. Let y ∈ AF (AF (x,≤ S),≤ T ). Thus there exist Y1, Y2, . . . , Yk ∈ F and times

t1, t2, . . . , tk ≥ 0 such that t1 + t2 + · · ·+ tk = s ≤ T and

y = exp(tkYk) exp(tk−1Yk−1) · · · exp(t1Y1)(y),

for some y ∈ AF (x,≤ S). By definition, there exists a sequence xj ∈ AF (x,≤ S) such

that xj → y as j →∞. By the existence, uniqueness, and smoothness theorem from

ordinary differential equations since Y1, . . . , Yk are smooth, for j sufficiently large, the

sequence:

yj := exp(tkYk) exp(tk−1Yk−1) · · · exp(t1Y1)(xj) ∈ AF (x,≤ S + T )

is defined. Moreover, yj → y as j →∞. This finishes the proof.

Page 42: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

42

Proof of Theorem 2.48. Let x ∈ Rd and T > 0. The inclusion

AF (x,≤ T ) = AF1(x,≤ T ) ⊂ AF1∪F2(x,≤ T )

is clear. To prove the reverse inclusion, we will show thatAF1∪F2(x,≤ T ) ⊂ AF (x,≤ T ).

Let y ∈ AF1∪F2(x,≤ T ). Thus there exist Y1, Y2, . . . , Yk ∈ F1 ∪F2 and corresponding

times t1, t2, . . . , tk ≥ 0 such that t1 + t2 + · · ·+ tk = s ≤ T and

y = exp(tkYk) exp(tk−1Yk−1) · · · exp(t1Y1)(x).

Let y0 = x. For j = 1, 2, . . . , k, define inductively

yj = exp(tjYj)(yj−1).

By the equivalence assumptions, yj ∈ AF (yj−1,≤ tj) for j = 1, 2, . . . , k. Hence y1 ∈

AF (x,≤ t1) and consequently by the previous proposition,

y2 ∈ AF (y1,≤ t2) ⊂ AF (AF (x,≤ t1),≤ t2) ⊂ AF (x,≤ t1 + t2).

Iterating this procedure yields the result since t1 + t2 + · · ·+ tk ≤ T .

Let now

Sat(F ) =⋃F∼F

F .

We call Sat(F ) the saturate of F . As a consequence of the previous theorem, we

have:

Corollary 2.50.

F ∼ Sat(F ).

Proof. Fix x ∈ Rd and T > 0. Since F ∼ F , the inclusionAF (x,≤ T ) ⊂ ASat(F )(x,≤ T )

is clear. We prove ASat(F )(x,≤ T ) ⊂ AF (x,≤ T ). Let y ∈ ASat(F )(x,≤ T ). Then there

exist Y1, . . . , Yk ∈ Sat(F ) and times t1, . . . , tk ≥ 0 such that t1 + · · ·+ tk = s ≤ T and

y = exp(tkYk) · · · exp(t1Y1)(x).

Page 43: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

43

Note that for j = 1, 2, . . . , k, Yj ∈ Fj for some polysystem Fj ∼ F . By the previous

theorem inductively we have:

F ∼k⋃j=1

Fj.

Thus, in particular, y ∈ AF (x,≤ T ).

For a polysystem F , Sat(F ) serves as an enlargement of F with which we can

work and still deduce properties of AF (x,≤ T ) for x ∈ Rd and T > 0. To determine

more polysystems than F that are equivalent to F , we have the following theorem:

Theorem 2.51. F is equivalent to the closed convex hull CF of the family

F = λY : 0 ≤ λ ≤ 1, Y ∈ F .

Here the closure is in the C∞-topology on compact subsets of Rd.

To prove this assertion, we first need two propositions.

Proposition 2.52. Let F be the closure of a polysystem F in the C∞-topology on

compact subsets of Rd. Then for all x ∈ Rd and T > 0:

AF (x, T ) ⊂ AF (x, T ).

Proof. Recall that AF (x, T ) is the set of points that can be reached from x using

trajectorties in F at exactly time T . Fix x ∈ Rd, T > 0, and let y ∈ AF (x, T ). Then

there exist Y1, Y2, . . . , Yk ∈ F and times t1, t2, . . . , tk ≥ 0 such that t1+t2+· · ·+tk = T

and

y = exp(tkYk) · · · exp(t1Y1)(x).

All we must show is that y ∈ AF (x, T ). For j = 1, 2, . . . , k, let Y nj ∈ F be such that

limn→∞ Ynj = Yj. One can show, see Theorem 4 in Chapter 3 of [Jur97], that this

implies for fixed x0 ∈ Rd and j = 1, 2, . . . , k

exp(tYj)(x0) = limn→∞

exp(tY nj )(x0),

Page 44: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

44

uniformly in t ∈ [0, T ]. From this, we see that for j = 1, 2, . . . , k

exp(tjYj) · · · exp(t1Y1)(x) ∈ AF

(x,

j∑l=1

tl

).

Taking j = k, we obtain the result.

Proposition 2.53. Suppose that Y1, Y2, . . . , Yk ∈ F . Then for all λ1, λ2, . . . , λk ∈

[0, 1] such that∑

j λj = 1, and all x ∈ Rd

exp

(T

(∑j

λjYj

))(x) ∈ AF (x, T ),

for all T > 0 for which the left-hand side is defined.

Proof. This proof of this is long but not hard. See Theorem 7 of Chapter 3 in

[Jur97].

Using the previous two propositions, we now prove Theorem 2.51.

Proof of Theorem 2.51. Fix x ∈ Rd and T > 0. The inclusion

AF (x,≤ T ) ⊂ ACF (x,≤ T )

is clear. Hence we show that

ACF (x,≤ T ) ⊂ AF (x,≤ T ).

To see this, we first let Co(F ) be the convex hull of the set

F = λY : 0 ≤ λ ≤ 1, Y ∈ F.

Note that, by Proposition 2.52, we have the inclusion

ACF (x, T ) ⊂ ACo(F )(x, T ),

Page 45: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

45

since CF = Co(F ). Hence ACF (x,≤ T ) ⊂ ACo(F )(x,≤ T ). Moreover, by Proposition

2.53, ACo(F )(x,≤ T ) ⊂ AF (x,≤ T ). Hence it is enough to show that

AF (x,≤ T ) ⊂ AF (x,≤ T ).

Let y ∈ AF (x,≤ T ). Thus there exist Y1, Y2, . . . , Yk ∈ F , λ1, λ2, . . . , λk ∈ [0, 1] and

times t1, t2, . . . , tk ≥ 0 such that t1 + t2 + . . .+ tk = s ≤ T and

y = exp(tkλkYk) exp(tk−1λk−1Yk−1) · · · exp(t1λ1Y1)(x).

But note that taking sj = λjtj for j = 1, 2, . . . , k we have s1 + s2 + · · ·+ sk ≤ T and

y = exp(skYk) · · · exp(s1Y1)(x).

Thus y ∈ AF (x,≤ T ) which finishes the proof.

In what follows, it will be very convenient to use CF to help determine AF (x,≤ T )

for x ∈ Rd and T > 0. There is yet another operation we may perform on F and

remain in the saturate. To describe it, we first need a definition and a remark.

Definition 2.54. Let F be an arbitrary polysystem and ψ : Rd → Rd be a diffeo-

morphism. We call ψ a normalizer of F if for all x ∈ Rd and T > 0:

ψ(AF (ψ−1(x),≤ T )) ⊂ AF (x,≤ T ).

We denote the set of all normalizers of F by Norm(F ).

Remark 2.55. For a diffeomorphism ψ : Rd → Rd and a smooth vector field Y on

Rd we may create another smooth vector field on Rd, which we denote by ψ#(Y ),

defined by

ψ#(Y ) = ψ∗ Y ψ−1,

where ψ∗ is the differential of ψ. The motivation for introducing this operation and

the notion of a normalizer is easily seen by the following theorem.

Page 46: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

46

Theorem 2.56. F is equivalent to

F# = ψ#(Y ) : ψ ∈ Norm(F ), Y ∈ F .

Proof. Fix x ∈ Rd and T > 0. The inclusion AF (x,≤ T ) ⊂ AF#(x,≤ T ) follows by

taking ψ to be the identity map on Rd and realizing the identity is a normalizer.

For the reverse inclusion, let y ∈ AF#(x,≤ T ). Choose Y1, Y2, . . . , Yk ∈ F and

t1, t2, . . . , tk ≥ 0 such that t1 + t2 + . . . tk = s ≤ T and

y = exp(tk(ψk)#(Yk)) exp(tk−1(ψk−1)#(Yk−1)) · · · exp(t1(ψ1)#(Y1))(x), (2.57)

for some ψ1, ψ2, . . . , ψk ∈ Norm(F ). By relation (2.57) and Proposition 2.49, it is

enough to show that for all ψ ∈ Norm(F ), Y ∈ F , and x0 ∈ Rd

exp(tψ#(Y ))(x0) ∈ AF (x0,≤ t),

for all t > 0 for which the left-hand side is defined. Note that we have

exp(tψ#(Y )) = ψ exp(tY ) ψ−1.

Since ψ is a normalizer of F ,

exp(tψ#(Y ))(x0) = ψ(exp(tY )(ψ−1(x0))) ∈ AF (x0,≤ t)

as exp(tY )(ψ−1(x0)) ∈ AF (ψ−1(x0),≤ t).

To determine easily which diffeomorphisms ψ are normalizers, we have:

Lemma 2.58. Let F be an arbitrary polysystem and ψ : Rd → Rd be a diffeomor-

phism. Suppose that for all x ∈ Rd and T > 0, ψ(x), ψ−1(x) ∈ AF (x,≤ T ). Then

ψ ∈ Norm(F ).

Proof. Let x ∈ Rd, T > 0, and ε > 0. We show that

ψ(AF (ψ−1(x),≤ T )) ⊂ AF (x,≤ T + ε).

Page 47: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

47

By Proposition 2.49, since ψ−1(x) ∈ AF (x, ε/2) we have

AF (ψ−1(x),≤ T ) ⊂ AF (AF (x,≤ ε/2),≤ T )

⊂ AF (x,≤ T + ε/2).

Since ψ(y) ∈ AF (y,≤ ε/2) for all y ∈ AF (x,≤ T + ε/2), applying Proposition 2.49

again we have:

ψ(AF (ψ−1(x),≤ T )) ⊂ ψ(AF (x,≤ T + ε/2))

⊂ AF (AF (x,≤ T + ε/2), ε/2)

⊂ AF (x,≤ T + ε).

This finishes the proof of the lemma.

We now stand in perfect position to determine the sets AF (x,≤ T ), but as we

recall from Theorem 2.42, we need some understanding of AF (x, T ) as well. The

transfer mechanism between the two is the next theorem.

Theorem 2.59. Suppose that F is a polysystem such that the span of the Lie algebra

generated by elements of F is the entire tangent space at all points x ∈ Rd. Suppose

moreover that for some x ∈ Rd and some U 6= ∅ open:

AF (x,≤ T ) ⊃ U and x ∈ AF (x,≤ T ) for all T > 0.

Then AF (x, T ) ⊃ U for all T > 0.

Proof. Fix x ∈ Rd and U 6= ∅ so that the assumptions are satisfied. Let y ∈ U ,

T > 0, and pick a small open set V ⊂ U that contains y. By Theorem 2 on p. 68 of

[Jur97], the spanning assumption of the Lie algebra allows us to conclude that

AF (x,≤ S) ⊃ U,

for all S > 0. Therefore for some 0 < T ′ ≤ T there exists a continuous trajectory

γ : [0, T ′] → Rd defined by vector fields of F such that γ(0) = x and γ(T ′) = y.

Page 48: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

48

Suppose that T ′ < T and pick ε > 0 small enough such that ε < T −T ′ and γ(t) ∈ V

for t ∈ [T ′ − ε, T ′]. Since x ∈ AF (x,≤ S) for all S > 0, for any positive numbers

S1 < S2, there exists S ′ > 0 and a curve δ : [0, S ′]→ Rd defined by vector fields of F

such that S1 < S ′ < S2 and δ(0) = δ(S ′) = x. Let S1 = T − T ′ and S2 = T − T ′ + ε

and note that the composite curve

σ(t) =

δ(t), for t ∈ [0, S ′]

γ(t− S ′) for t ∈ [S ′, S ′ + T ′]

has σ(0) = x and σ(T ) ∈ V . This finishes the proof that AF (x, T ) ⊃ U for all

T > 0.

Let us now enjoy the fruits of our labor. We prove of a classical theorem which

we do not need for the main results of this dissertation. Still, however, the proof

illustrates the use of the methods developed in this section.

Theorem 2.60 (Rank Theorem). Let A ∈ Md(R), b ∈ Rd and consider the control

systemdy(t)

dt= Ay(t) + u(t)b,

where u : [0,∞) → R is a piecewise constant mapping with at most finitely many

discontinuities. Let A be the vector field determined by A(x) = Ax and b be the

constant vector field b(x) = b. Consider the polysystem F = A + ub : u ∈ R. If

spanb, Ab, . . . , Ad−1b = Rd, then AF (x, T ) = Rd for all x ∈ Rd and T > 0.

Proof. By Theorem 2.51, for all u ∈ R:

limλ→∞

1

λ(A+ uλb) = ub ∈ Sat(F ).

Since ψu(x)(t) = x + tub is the integral curve of ub with inverse ψu(x)−1(t) =

ψ−u(x)(t), we note that for all u ∈ R, the map ψu(x) := ψu(x)(1) ∈ Norm(Sat(F ))

Page 49: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

49

for all u ∈ R. Hence

(ψu)#(A)(x) = (ψu)∗(A(ψ−1u (x))

= A(x− ub)

= Ax− uAb.

Thus (ψu)#(A) = A− uAb ∈ Sat(F ) for all u ∈ R where uAb is the constant vector

field uAb(x) = uAb. Moreover,

limλ→∞

1

λ(A+ uλAb) = uAb ∈ Sat(F ),

for all u ∈ R. Note by Theorem 2.51, we have

AF (x,≤ T ) ⊃ x+ spanb, Ab.

We can continue this procedure to see that the constant vector fields u2A2b , . . . ,

ud−1Ad−1b ∈ Sat(F ) for all u2, . . . , ud−1 ∈ R. Note that by the observation above

and the spanning assumption, this is sufficient to conclude

AF (x,≤ T ) = Rd,

for all x ∈ Rd and all T > 0. By Theorem 2 on p. 68 of [Jur97], we obtain

AF (x,≤ T ) = Rd,

for all x ∈ Rd and all T > 0. By the previous theorem since x ∈ AF (x,≤ T ) for all

T > 0 and all x ∈ Rd we may conclude

AF (x, T ) = Rd,

for all x ∈ Rd and all T > 0, as required.

It is important to point out that one can prove this theorem without using the

methods of this section. In cases where one cannot infer accessibility properties easily,

geometric control theory allows us to do so with more ease.

Page 50: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

50

2.6 Geometric Ergodicity

Let ‖ · ‖TV be the total variation norm on B(Rd)-measures and suppose that the

Markov process x(t) defined by equation (2.1) with transition kernel P (x, t, A) is

non-explosive and has a unique invariant probability measure µ. In this section, we

provide sufficient conditions under which we can quantify a rate of convergence to

the equilibrium µ. Specifically, we are concerned with the case when this rate is

geometric. In this section, we will thus show that under two minimal assumptions

that the process x(t) has the following property:

Property 2.61. There exists a constant ρ ∈ (0, 1) and a function Ψ : Rd → [0,∞)

such that

‖P (x, t, · )− µ( · )‖TV ≤ Ψ(x)ρt,

for all t ≥ 0 and all x ∈ Rd.

Thus for all fixed x ∈ Rd, the transition measures P (x, t, · ) approach the invariant

probability measure exponentially fast in time in the total variation norm. This is an

extremely strong form of convergence which surprisingly can, in our case, be improved.

If the process x(t) satisfies Property 2.61, we call x(t) exponentially ergodic.

There are many different presentations on exponential ergodicity of Markov pro-

cesses. See, for example, the works of Meyn and Tweedie [MT92, MT93a, MT93b,

MT09] or the treatments in the context of stochastic differential equations of Rey-

Bellet [RB06] and Higham, Mattingly, and Stuart [MSH02]. Appealing partially to

these, we will primarily use the elegant and concise notes of Hairer and Mattingly

[HM08]. The benefit of their work over others is their proof of exponential convergence

is short. Moreover, one can apply their methods as they do to extremely degenerate

(and even infinite dimensional) stochastic differential equations.

Throughout this section, we make the two assumptions below. One should not

think of them as being entirely separate as they are intertwined through the test

Page 51: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

51

function Φ.

Assumption 2.62. There exists a non-negative function Φ ∈ C2(Rd) and positive

constants C,D such that

LΦ(x) ≤ −CΦ(x) +D for all x ∈ Rd.

Assumption 2.63. There exists a distinguished time T0 > 0 such that for all R > 0

sufficiently large, there exists αR ∈ (0, 1) and a probability measure ν such that

infx∈CR

P (x, T0, · ) ≥ αRν( · ),

where CR = x ∈ Rd : Φ(x) ≤ R and Φ is as in Assumption 2.62.

If the sets CR = x : Φ(x) ≤ R for R > 0 are pre-compact, Assumption 2.62

essentially implies that the dynamics is focused in a possibly large, but bounded

region. One can of course guarantee pre-compactness if Φ(x) → ∞ as |x| → ∞,

which is sometimes also a standard assumption. With this idea in place, Assumption

2.63 is only non-standard in the sense in which it is expressed, even though we should

expect an exponentially ergodic Markov process to satisfy it as “mixing” is crucial in

this regard.

To state the main theorem of this section which implies Property 2.61, we intro-

duce a weighted norm ‖ · ‖ on B(Rd)-measureable real-valued functions ϕ:

‖ϕ‖ = supx∈Rd

|ϕ(x)|1 + Φ(x)

,

where ϕ is such that ‖ϕ‖ <∞. The norm ‖ · ‖ comes equipped with a dual norm on

signed B(Rd)-measures ν:

‖ν‖ = sup‖ϕ‖≤1

∫ϕ(x)ν(dx).

We will prove:

Page 52: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

52

Theorem 2.64. Under Assumption 2.62 and Assumption 2.63, there exist constants

ρ ∈ (0, 1) and E > 0 such that

‖P (x, t, · )− µ‖ ≤ Eρt(1 + Φ(x)),

for all t ≥ 0 and all x ∈ Rd.

In many works, the proof of the above is quite non-trivial. To circumambulate

these difficulties, Hairer and Mattingly introduce a family of norms that are slightly

tweaked versions of the norm ‖ · ‖ above. To this end, let γ > 0 and define

‖ϕ‖γ = supx∈Rd

|ϕ(x)|1 + γΦ(x)

,

where ϕ is such that ‖ϕ‖γ < ∞. Of course, like ‖ · ‖ above, the norm ‖ · ‖γ has a

dual norm on signed B(Rd)-measures ν:

‖ν‖γ = sup‖ϕ‖γ≤1

∫ϕ(x)ν(dx).

The goal is to choose the parameter γ > 0 so that a version of Theorem 2.64 holds

in this norm in the discrete setting, for (1) the norms ‖ · ‖ and ‖ · ‖γ on measures

are equivalent and (2) we have a natural time T0 > 0 to define an embedded Markov

chain. In view of this, let xnn∈N be the Markov chain with n-step transitions

P n(x,A) given by:

P n(x,A) := P (x, nT0, A),

for x ∈ Rd and A ∈ B(Rd). To conclude Theorem 2.64, we will show:

Theorem 2.65. Under Assumption 2.62 and Assumption 2.63, there exists ρ ∈ (0, 1)

and γ > 0 such that

‖ν1P − ν2P‖γ ≤ ρ‖ν1 − ν2‖γ, (2.66)

for all B(Rd)-probability measures ν1, ν2 where P = P 1. Moreover,∫Rd

Φ(y)µ(dy) <∞,

where µ is the invariant probability measure.

Page 53: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

53

Before we prove that Theorem 2.65 implies Theorem 2.64, we first relate Assump-

tion 2.62 to a bound on the semigroup Pt.

Lemma 2.67. Suppose that Assumption 2.62 holds. Then for all t ≥ 0

PtΦ ≤ e−CtΦ +D/C.

Proof. Let Ψ(t, x) = eCt(Φ(x)−D/C). Applying Dynkin’s formula, we obtain:

Ex [Ψ(ξn(t), x(ξn(t)))] ≤ Φ(x)−D/C.

Estimating the left-hand side of the above, we note that:

Ex [Ψ(ξn(t), x(ξn(t)))] ≥ Ex[1t<ξnΨ(ξn(t), x(ξn(t)))

]= eCtEx

[1t<ξn(Φ(x(t))−D/C)

]Combining this with the first estimate, we obtain:

Ex[1t<ξn(Φ(x(t))−D/C)

]≤ e−CtΦ(x)− e−CtD/C.

Since the process is non-explosive and the bound above holds for all n, we have:

PtΦ(x) ≤ e−CtΦ(x) + (1− e−Ct)D/C

≤ e−CtΦ(x) +D/C,

for all t ≥ 0 and x ∈ Rd.

Proof that Theorem 2.65 =⇒ Theorem 2.64. We note that since µ is invari-

ant, by (2.66) we have

‖P n(x, · )− µ‖γ ≤ ρn‖P (x, · )− µ‖γ,

for all n ∈ N and all x ∈ Rd. We first show that there exists a constant E ′ > 0

independent of x such that

‖P (x, · )− µ‖γ ≤ E ′(1 + Φ(x)).

Page 54: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

54

Since∫

Φ(y)µ(dy) = C ′ <∞, by the previous lemma we obtain

‖P (x, · )− µ‖γ ≤ sup‖ϕ1‖γ , ‖ϕ2‖γ≤1

∫[|ϕ1(y)|P (x, dy) + |ϕ2(y)|µ(dy)]

≤∫

(1 + γΦ(y))[P (x, dy) + µ(dy)]

≤ 2 + γe−CT0Φ(x) + γD/C + γC ′

≤ E ′[1 + Φ(x)],

for some E ′ > 0 independent of x. We now show that

‖P n(x, · )− µ‖γ ≥ min(1, γ)‖P n(x, · )− µ‖.

To see this, note that

‖P n(x, · )− µ‖γ = sup‖ϕ‖γ≤1

∫ϕ(y)(P n(x, dy)− µ(dy))

sup‖ϕ‖≤1

∫ϕ(y)(P n(x, dy)− µ(dy)) if γ ≥ 1,

γ sup‖ϕ‖≤1

∫ϕ(y)(P n(x, dy)− µ(dy)) if γ ∈ (0, 1)

≥ min(1, γ)‖P n(x, · )− µ‖.

Thus, in particular, we have shown that

‖P n(x, · )− µ‖ ≤ E ′min(1, γ)−1ρn[1 + Φ(x)],

for all x ∈ Rd, n ∈ N. To introduce the continuous paramter t ≥ 0 in the above, for

t ≥ 0 write t = nT0 + δ where δ ∈ [0, T0). By the Chapman-Kolmogorov equations,

Page 55: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

55

we obtain

‖P (x, t, · )− µ‖ = sup‖ϕ‖≤1

∫ϕ(y)(P (x, t, dy)− µ(dy))

= sup‖ϕ‖≤1

∫ϕ(y)

(∫P (x, δ, dx′)P (x′, nT0, dy)− µ(dy)

)= sup

‖ϕ‖≤1

∫P (x, δ, dx′)

∫ϕ(y) (P n(x′, dy)− µ(dy))

=

∫P (x, δ, dx′)‖P n(x′, · )− µ‖

≤∫P (x, δ, dx′)E ′min(1, γ)−1ρn[1 + Φ(x′)]

= E ′min(1, γ)−1ρn [1 + PδΦ(x)]

≤ E ′min(1, γ)−1ρn[1 + e−CδΦ(x) +D/C

]Using this, we can choose ρ ∈ (ρ, 1) and E > 0 independent of x such that

‖P (x, t, ·)− µ‖ ≤ Eρt[1 + Φ(x)],

for all t ≥ 0 and all x ∈ Rd. This finishes the proof.

Sometimes it is useful to know a precise rate of convergence and estimate. For

this, we note:

Corollary 2.68. Under the assumptions of the previous theorem, we have the bound:

‖P (x, t, · )− µ‖ ≤ 1

min(1, γ)ρ(ρ1/T0)t

[2 + γe−CT0Φ(x) + γD/C(1 + e−CT0) + γC ′

],

where C ′ =∫

Φ(y) dµ(y).

Proof. This follows easily from the proof above.

With this implication in place, we now are in position to setup the main compo-

nents of the proof of Theorem 2.66. With the norm ‖ · ‖γ on functions in mind, we

define a metric on Rd:

dγ(x, y) =

0 if x = y;

2 + γΦ(x) + γΦ(y) if x 6= y.

Page 56: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

56

Since Φ ≥ 0, it is easy to show that this is indeed a metric. Moreover, we define a

Lipschitz semi-norm:

|||ϕ|||γ = supx 6=y

|ϕ(x)− ϕ(y)|dγ(x, y)

.

Note that we have the following relationship between |||ϕ|||γ and ‖ϕ‖γ.

Lemma 2.69. Suppose that ‖ϕ‖γ <∞. We have

|||ϕ|||γ = infc∈R‖ϕ+ c‖γ.

Proof. It is easy to see that |||ϕ|||γ ≤ ‖ϕ‖γ for all functions ϕ. Thus by the shift

invariance of ||| · |||, we have |||ϕ|||γ ≤ infc∈R ‖ϕ+ c‖γ. To see the other inequality, let

|||ϕ|||γ = 1 and define

c = infy∈Rd

(1 + γΦ(y)− ϕ(y)).

Since ‖ϕ‖γ <∞, it is easy to see that c ∈ R. Note that

ϕ(x) + c ≤ ϕ(x) + 1 + γΦ(x)− ϕ(x) = 1 + γΦ(x),

and

ϕ(x) + c = infy∈Rd

(ϕ(x) + 1 + γΦ(y)− ϕ(y))

≥ infy∈Rd

(−|ϕ(x)− ϕ(y)|+ 1 + γΦ(y))

≥ infy∈Rd

(−|||ϕ|||γdγ(x, y) + 1 + γΦ(y))

= −1− γΦ(x).

Thus ‖ϕ+c‖γ ≤ 1 = |||ϕ|||c and hence we have infc∈R ‖ϕ+c‖ ≤ |||ϕ|||γ as claimed.

We will now use ||| · |||γ to prove Theorem 2.66.

Proof of Theorem 2.66. We will first prove that∫Φ(y)µ(dy) <∞,

Page 57: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

57

where µ is the unique invariant measure. By the proof of Theorem 2.32, since µ is

unique it can be defined as a limit point in the weak topology of the sequence:

µT ( · ) =1

T

∫ T

0

P (x′, s, · ) ds,

for some x′ ∈ Rd. Thus take a subsequence µTk that converges weakly to µ as k →∞

and note that for all k = 1, 2, . . .∫Φ(y)µTk(dy) =

1

Tk

∫ Tk

0

∫P (x′, s, dy)Φ(y) ds

=1

Tk

∫ Tk

0

PsΦ(x′) ds

≤ 1

TkD′ +D/C,

for some D′ > 0 independent of k. Note that the last inequality follows from Lemma

2.67. Applying Fatou’s lemma, finishes the proof that∫

Φ(y)µ(dy) < ∞. Moreover

we can obtain the precise estimate:

C ′ =

∫Φ(y)µ(dy) ≤ D/C,

To show the contraction property, we first claim that it is enough to prove that

there exist constants ρ ∈ (0, 1) and γ > 0 such that

|||Pϕ|||γ ≤ ρ|||ϕ|||γ.

Note that if ν1, ν2 are Borel probability measures we have for cϕ = infy∈Rd(1+γΦ(y)−

Page 58: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

58

ϕ(y)):

‖ν1P − ν2P‖γ = sup‖ϕ‖γ≤1

∫ϕ(y) (ν1P (dy)− ν2P (dy))

= sup‖ϕ‖γ≤1

∫Pϕ(x)(ν1(dx)− ν2(dx))

≤ sup|||ϕ|||γ≤1

∫Pϕ(x)(ν1(dx)− ν2(dx))

≤ sup|||ϕ|||γ≤1

∫ρϕ(x)(ν1(dx)− ν2(dx))

≤ sup||ϕ+cϕ||γ≤1

∫ρ(ϕ(x) + cϕ − cϕ)(ν1(dx)− ν2(dx))

= sup‖ϕ+cϕ‖γ≤1

∫ρ(ϕ(x) + cϕ)(ν1(dx)− ν2(dx))

≤ ρ‖ν1 − ν2‖γ.

Thus all we have left to show is that there exist constants ρ ∈ (0, 1) and γ > 0 such

that |||Pϕ|||γ ≤ ρ|||ϕ|||γ. By Lemma 2.67, we have:

PT0Φ(x) ≤ e−CT0Φ(x) +D/C.

Set α = e−CT0 ∈ (0, 1) and β = D/C. First suppose that x 6= y are such that

Φ(x) + Φ(y) ≥ R. Then we have:

|Pϕ(x)− Pϕ(y)| ≤ P (|ϕ(x)− ϕ(y)|)

≤ |||ϕ|||γP (dγ(x, y))

= (P (2 + γΦ(x) + γΦ(y)))|||ϕ|||γ

≤ (2 + γαΦ(x) + γαΦ(y) + 2γβ)|||ϕ|||γ.

Let ε > 0 be small enough such that α(ε) = α + ε < 1. We then have the estimate:

|Pϕ(x)− Pϕ(y)| ≤ (2 + 2γβ − 2γεR + γα(ε)Φ(x) + γα(ε)Φ(y))|||ϕ|||γ.

Note that, if we choose R = R(β, ε) > 0 large enough such that:

R > β/ε,

Page 59: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

59

we have the bound

supx 6=y,Φ(x)+Φ(y)≥R

|PΦ(x)− PΦ(y)|dγ(x, y)

≤ ρ1|||ϕ|||γ,

where

ρ1(β, ε, γ) = max 1− γ(Rε− β), α(ε) .

We now suppose that x 6= y are such that Φ(x) + Φ(y) ≤ R. Hence x, y ∈ CR. By

Lemma 2.69, we may assume that ‖ϕ‖γ = 1. Note then we can decompose ϕ into

two functions ϕ = ϕ1 + ϕ2 such that |ϕ1(x)| ≤ 1 and |ϕ2(x)| ≤ γΦ(x) for all x ∈ Rd.

Moreover, by Assumption 2.63, for x ∈ CR we may define another transition kernel:

Q(x, · ) =1

1− αRP (x, · )− αR

1− αRν( · ).

Hence we have

|Pϕ(x)− Pϕ(y)| ≤ (1− αR)|Qϕ1(x)−Qϕ1(y)|+ (1− αR)|Qϕ2(x)−Qϕ2(y)|

≤ 2(1− αR) + γαΦ(x) + γαΦ(y) + 2γαβ.

Hence choose γ > 0 small enough such that

γ <αRαβ

.

Hence we obtain:

supx 6=y,Φ(x)+Φ(y)≤R

|Pϕ(x)− Pϕ(y)|dγ(x, y)

≤ ρ2,

where

ρ2 = max 1− αR + γαβ, α .

Hence we have shown the result for

ρ = max

1− γ(Rε−D/C), e−CT0 + ε, 1− αR + γe−CT0D/C,

Page 60: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

60

where ε, R, γ > 0 are such that

e−CT0 + ε < 1

R > D/(Cε)

γ <αR

e−CT0D/C.

Remark 2.70. We note that if Assumption 2.62 and Assumption 2.63 are satisfied

with Φ such that Φ(x) → ∞ as |x| → ∞, the conclusion of Theorem 2.64 is much

stronger than Property 2.61. This is because the norm ‖ · ‖ which quantifies the

convergence rate is taken over the supremum of a wider class of functions than just

bounded functions. These functions are moreover permitted to approach infinity (of

course no faster than 1 + Φ(x)) as |x| → ∞. In the case of equation 1.2, we will be

able to prove the existence of such a Φ and hence conclude not only Property 2.61,

but also the bound in Theorem 2.64.

With the ideas of Chapter 2 in place, we now proceed onto proving properties

about the system 1.2.

Page 61: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

61

Chapter 3

Proof of Main Theorem

3.1 Introduction

Let n ≥ 2. We now prove Theorem 1.5. To do so, we will show the following two

lemmata:

Lemma 3.1. Suppose that (κ1, κ2) ∈ C2 is transversal to Dn−1, then

1. The process z(t) is non-explosive.

2. There is one and only one invariant probability measure µ.

3. z(t) satisfies Property 2.61.

Lemma 3.2. If (κ1, κ2) ∈ C2 is not transversal to Dn−1, there exists a set of initial

conditions A 6= ∅ ⊂ C such that

Pz ξ <∞ > 0,

for all z ∈ A.

We first attend to Lemma 3.1; the proof of which comprises five sections. Iden-

tifying C with R2 throughout, the first four sections focus entirely on exhibiting a

smooth function Φ : C→ [0,∞) with positive constants C and D such that:

(C1) Φ(z, z)→∞ as |z| → ∞,

(C3) LΦ(z, z) ≤ −CΦ(z, z) +D for all z ∈ C.

By Theorem 2.8 and Theorem 2.34, the existence of Φ proves the process z(t) is

non-explosive and has invariant probability measures. If, in addition, κ1 and κ2 span

Page 62: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

62

the entire complex plane over R, it is not hard to see that the existence of Φ also

guarantees exponential ergodicity of z(t) [RB06]. However, we are also interested in

cases where κ1 and κ2 are linearly dependent over R. Hence different methods are

employed in Section 3.6 to validate parts 2 and 3 of Lemma 3.1.

To prove Lemma 3.2, we spend a short sixth section using Feller’s test as in Section

2.2.2.

3.2 Lyapunov Coverings

The existence of a smooth function Φ : C → [0,∞) that satisfies (C1) and (C3) is

shown by an explicit construction; the first part of which is to find locally-defined test

functions which satisfy local versions of (C1) and (C3) on their respective domains.

If the union of these domains essentially covers C, we have almost, but not quite,

assured stability of the process z(t). To proceed, we need the following terminology:

Definition 3.3. Let U ⊂ C be an unbounded domain with continuous boundary ∂U .

Suppose that ϕ : U → R satisfies:

(I) ϕ ∈ C∞(U).

(II) ϕ(z, z)→∞ as |z| → ∞, z ∈ U .

(III) There exist positive constants c and d such that

Lϕ(z, z) ≤ −cϕ(z, z) + d for all z ∈ U.

We call ϕ a Lyapunov function on U . If there exists a sequence of Lyapunov

functions ϕ1, ϕ2, . . . , ϕk on U1, U2, . . . , Uk respectively such that

C =k⋃j=1

Uj ∪BR(0),

for some R > 0, we call (ϕ1, U1), (ϕ2, U2), . . . , (ϕk, Uk) a Lyapunov covering.

Page 63: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

63

From this definition, it is intuitively clear that the existence of a Lyapunov covering

provides some handle on the behavior of the process z(t). To illustrate this, we prove

the next lemma. Let us use ξn to denote the first exit time of z(t) from Bn(0) and ξ

to denote the finite or infinite limit of ξn as n→∞.

Lemma 3.4. Suppose ϕ is a Lyapunov function on U . Let V ⊂ U be another

unbounded region such that ∂V is continuous and ∂U ∩∂V = ∅. If ξV = inft>0z(t) ∈

V c, then for z ∈ V

Pz ξV < ξ = 1.

Proof. Let ξU = inft>0z(t) ∈ U c and ξn,U(t) = ξn ∧ ξU ∧ t. By Definition 3.3, it is

clear that the much weaker bound Lϕ(z, z) ≤ cϕ(z, z)+d for some c, d > 0 is satisfied

on U . Letting ψ(t, z, z) = e−ct(ϕ(z, z) + dc), for z ∈ V we obtain by Dynkin’s formula

Ez

[ψ(ξn,U(t), z(ξn,U(t)), z(ξn,U(t)))

]− ψ(0, z, z) = Ez

[∫ ξn,U (t)

0

Lψ(s, z(s), z(s)) ds

]≤ 0.

We see that for n sufficiently large:

Ez

[ψ(ξn,U(t), z(ξn,U(t)), z(ξn,U(t)))

]≥ e−ctϕnPz ξn ≤ ξU ∧ t ,

where ϕn = infz∈U∩∂Bn ϕ(z, z) → ∞ as n → ∞. Combining the previous two esti-

mates yields

Pz ξn ≤ ξU ∧ t ≤ ϕ−1n ectψ(0, z, z).

Letting n→∞ we see that Pzξ ≤ ξU ∧t = 0 for all t ≥ 0. Whence Pzξ ≥ ξU = 1.

By path continuity of z(t), PzξU > ξV = 1. Hence Pzξ > ξV = 1 as claimed.

Remark 3.5. Note that the same behavior is true if (III) of Definition 3.3 is replaced

by a much weaker bound. Indeed, suppose U is as in Definition 3.3 and ϕ : U → R

Page 64: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

64

Figure 3.1. Cartoon of Lemma 3.4

U

V z

z H ΞV L

3 4 5 6

-1.0

- 0.5

0.5

1.0

satisfies (I) and (II) of Definition 3.3 on U . If ϕ satisfies the bound

Lϕ(z, z) ≤ cϕ(z, z) + d,

for all z ∈ U for some c, d > 0, then the conclusion of Lemma 3.4 remains valid.

The existence of a Lyapunov covering (ϕj, Ujkj=1 guarantees z(t) cannot leave

C directly through any unbounded region V ⊂ Uj with continuous boundary such

that ∂V ∩ ∂Uj = ∅. As we shall see in our case, it is possible to extract subsets

Vj ⊂ Uj with continuous boundary such that ∂Uj ∩ ∂Vj = ∅ for all j = 1, 2, . . . , k and

(ϕj, Vj)kj=1 is a Lyapnov covering.

Definition 3.6. Suppose that (ϕj, Vj)kj=1 is as above. We call (ϕj, Vj)kj=1 a

strong Lyapunov covering subordinate to (ϕj, Uj)kj=1.

Hence, given the existence of a strong Lyapunov covering (ϕj, Vj)Nj=1, z(t) can-

not exit C directly through any Vj. Moreover

C =N⋃j=1

Vj ∪BR(0),

for some R > 0. Thus it seems likely that z(t) cannot exit C in finite time. However,

it is possible z(t) could oscillate between two or more regions on its way to infinity in

Page 65: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

65

finite time. As we shall see, to eliminate this rare event we must glue our Lyapunov

covering together so that we have a Lyapunov function Φ on all of C. If necessary,

by adding a sufficiently large constant to Φ, we have a nonnegative smooth function

that satisfies (C1) and (C3).

The construction of a Lyapunov function Φ on C naturally splits into two stages:

Stage 1: Existence of a Lyapunov covering from which we can extract a subordinate

strong Lyapunov covering.

Stage 2: Existence of a Lyapunov function Φ on all of C as a glued version of the

Lyapunov covering.

Because the Lyapunov covering presented here involves some non-standard regions

in the complex plane, we will first define, graph, and prove some properties of these

sets separate from their associated Lyapunov functions. This is done to amend the

procedure as a whole. Also, we prove a lemma which allows us to vary the magnitude

of the diffusion present in (1.2). We then proceed onto Stage 1 and Stage 2 in that

order.

3.3 Lyapunov Regions and Scaling

3.3.1 Lyapunov Regions

We now provide the regions of definition for the Lyapunov functions that will follow

in the next section. To this end, for z 6= 0 ∈ C let θz = arg(z). Of course, θz is multi-

valued; however, its multi-valuedness plays no significant role in the arguments. Since

there is some symmetry in (1.2), it is convenient to use g to denote a primitive (n−1)st

root of unity. Moreover, we choose R > 0 sufficiently large which will be made precise

later.

Page 66: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

66

For the first region, we define a constant:

ηn =2(n− 1)

π(e− 1)> 0,

and the reason for its chosen value will be more apparent later in this section.

Region 1.

U1 = BR(0)c ∩z 6= 0 ∈ C : cos((n− 1)θz) ≤ −ηn|z|−1

U1

U1

U1

ΑH0L

ΑH1L

ΑH2 L

ΑH3L

ΑH4L

ΑH5L

-10 - 5 0 5 10

-10

- 5

0

5

10

Figure 3.2. The region U1 (in blue) for n = 4. The pink represents the distance tothe rays α(k) defined below.

Remark 3.7. For |z| large, one should think of U1 as essentially all points where

cos((n− 1)θz) < 0.

The power −1 in |z|−1 allows us to get reasonably close to where cos((n− 1)θz) = 0.

The constant ηn allows us to be precise about how close.

To define the second region, for k ∈ Z we introduce angles

α(k) =π

2(n− 1)+

πk

n− 1,

Page 67: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

67

and let

U2(α(k)) = B2(0) + z 6= 0 ∈ C : θz = α(k) .

We now have:

Region 2.

U2 =⋃k∈Z

U2(α(k)) ∩BR(0)c.

U2

U2

U2

U2

U2

U2

-10 - 5 0 5 10

-10

- 5

0

5

10

Figure 3.3. The region U2 for n = 4. It covers the rays α(k) for all k ∈ Z and alsooverlaps U1 by the choice of ηn.

Region 3.

U3 =z 6= 0 ∈ C : min | sin((n− 1)θz)|, cos((n− 1)θz) ≥ ηn|z|−1

.

Remark 3.8. Hence U3 is similar in some sense to U1. We see that we are within the

realm where cos((n−1)θz) > 0 and can get close to the rays where cos((n−1)θz) = 0

or sin((n− 1)θz) = 0, but only as close as ηn|z|−1 permits.

Page 68: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

68

U3

U3

U3

U3

U3

U3

-15 -10 - 5 0 5 10 15

-15

-10

- 5

0

5

10

15

Figure 3.4. The region U3 with n = 4. With the choice of ηn, U3 overlaps U2.

For the fourth region, we use the variables (x, y) ∈ R2 instead of z = x + iy. We

first define a ground region:

U4(0) = x ≥ 1 ∩|x|−(n−1)/2 ≤ |y| ≤ 2

and rotate it to produce the entire U4.

Region 4.

U4 =⋃k∈Z

(gkU4(0)) ∩BR(0)c.

Remark 3.9. Thus we take U4(0) and rotate it by integer multiples of the angle

2π/(n− 1). We do the same for the next region.

Let

U5(0) = x ≥ 1 ∩|y| ≤ 2|x|−(n−1)/2

Region 5.

U5 =⋃k∈Z

(gkU5(0)) ∩BR(0)c.

Page 69: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

69

U4 H0LU4 H0L

4 6 8 10 12 14

-10

- 5

0

5

10

Figure 3.5. The region U4(0) with n = 4. Although it is hard to see, there is a tinyspace that still needs to be covered.

With these definitions in place, we prove there exists S > 0 such that

C =5⋃j=1

Uj ∪BS(0).

Before we do this, we quantify the word “close” used above in the remarks. For k ∈ Z,

define angles

β(k) =πk

n− 1.

Proposition 3.10. For K,R > 0 and l ∈ Z, consider the sets

V1(K,R, l) =z 6= 0 ∈ C : | cos((n− 1)θz)| = K|z|−1

∩|θz − α(l)| ≤ R−1

V2(K,R, l) =

z 6= 0 ∈ C : | sin((n− 1)θz)| = K|z|−1

∩|θz − β(l)| ≤ R−1

W1(l) = z 6= 0 ∈ C : θz = α(l)

W2(l) = z 6= 0 ∈ C : θz = β(l) .

Page 70: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

70

U5H0L

4 5 6 7 8 9 10

- 0.4

- 0.2

0.0

0.2

0.4

Figure 3.6. The region U5(0) for n = 4

Then there exists R′ > 0 such that for all R ≥ R′ and all l ∈ Z

dist (V1(K,R, l),W1(l)) ≤ Kπ(e− 1)

2(n− 1)

dist (V2(K,R, l),W2(l)) ≤ Kπ(e− 1)

2(n− 1).

Proof. We will only prove there exists R′ > 0 such that for all R ≥ R′:

dist (V1(K,R, 0),W1(0)) ≤ Kπ(e− 1)

2(n− 1),

as the other proofs are nearly identical. By definition, for R′ > 1 sufficiently large

any z ∈ V1(K,R, 0) for R ≥ R′ has |z| ≥ 2K. Fix R′ > 1 such that the prior occurs

Page 71: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

71

and note that for R ≥ R′ and z ∈ V1(K,R, 0) we have the estimates:

||z|eiθz − |z|eiα(0)| = |z||ei(θz−α(0)) − 1|

= |z||θz − α(0)|

∣∣∣∣∣∞∑j=1

in(θz − α(0))n−1

n!

∣∣∣∣∣≤ (e− 1)|z||θz − α(0)|

=e− 1

n− 1|z|∣∣∣(n− 1)θz −

π

2

∣∣∣=

e− 1

n− 1|z|∣∣∣arccos(±K|z|−1)− π

2

∣∣∣=

e− 1

n− 1|z|∣∣arcsin(±K|z|−1)

∣∣=

e− 1

n− 1|z|

∣∣∣∣∣∞∑j=0

((2j)!

22j(j!)2

(±K|z|−1)2j+1

(2j + 1)

)∣∣∣∣∣≤ K

e− 1

n− 1| arcsin(1)|

= Kπ(e− 1)

2(n− 1).

Thus since z′ = |z|eiα(0) ∈ W1(0), we have finished the proof.

Thus the choice of K = ηn now becomes clear. We now are in position to prove:

Proposition 3.11. There exists S > 0 such that

C =5⋃j=1

Uj ∪BS(0).

Proof. Let z 6= 0 ∈ C. There are three cases:

Case 1. cos((n− 1)θz) = 0.

Case 2. cos((n− 1)θz) < 0.

Case 3. cos((n− 1)θz) > 0.

Case 1. Hence θz = α(k) for some k ∈ Z. Thus for |z| ≥ R, z ∈ U2.

Page 72: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

72

Case 2. For |z| ≥ R, either cos((n − 1)θz) ≤ −ηn|z|−1 or cos((n − 1)θz) ∈

(−ηn|z|−1, 0). In the first case, z ∈ U1. In the second case, for |z| sufficiently large,

by the proposition above, z is within distance 1 unit to one of the rays α(k) for some

k ∈ Z. Hence, z ∈ U2.

Case 3. For |z| ≥ R, either cos((n−1)θz) ∈ (0, ηn|z|−1) or cos((n−1)θz) ≥ ηn|z|−1.

In the first case, by the proposition, for |z| sufficiently large, z is within distance 1

unit of one of the rays α(k) for some k ∈ Z. Hence z ∈ U2. In the second case, either

cos((n − 1)θz) ≥ ηn|z|−1 and | sin((n − 1)θz)| ≥ ηn|z|−1, or cos((n − 1)θz) ≥ ηn|z|−1

and | sin((n − 1)θz)| < ηn|z|−1. In the first case, z ∈ U3. In the second case, by the

proposition for |z| sufficiently large, z is within distance 1 unit of one of the rays β(k)

for some k ∈ Z. Since cos((n− 1)θz) > 0, z must be within distance 1 unit of one of

the rays 2πk/(n− 1) for some k ∈ Z. Hence, z ∈ U4 ∪U5 for |z| sufficiently large.

3.3.2 Scaling

It is important to realize that the diffusion process z(t) defined by equation (1.2)

depends on (κ1, κ2) ∈ C2. It is; however, extremely easy to forget about this when

constructing Φ. To emphasize, Φ can (and should) depend on κ1 and κ2. It will be

convenient in the next section to work with diffusion parameters that are sufficiently

small which is why we provide the next lemma. To note that the operator L depends

on κ = (κ1, κ2) we will write Lκ. We have the following:

Lemma 3.12. Suppose that Φ is a Lyapunov function on C corresponding to the

operator Lκ. For all parameters ι = (ι1, ι2) such that ι = λκ for some λ > 0, there

exists a Lyapunov function Ψ corresponding to the operator Lι.

Proof. Let λ > 0 be such that λ = λ−(n+1)/2. Define Ψ(z, z) = Φ(λz, λz). Note

Page 73: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

73

easily that Ψ satisfies properties (I) and (II). To verify (III), let w = λz and note:

LιΨ(z, z) =1

λn−1LκΦ(w, w)

≤ −CΨ(z, z) +D,

for some C,D > 0, as required.

3.4 Stage 1: A (Strong) Lyapunov Covering

In this section we prove the existence of a Lyapunov covering. From what follows, it

is very easy to extract a strong Lyapunov covering, so we will eliminate this minor

detail to focus solely on constructing Lyapunov functions on U1, U2, U3, U4, and U5.

We first define a multitude of constants which depend on the diffusion parameters

κ1 and κ2. In view of Lemma 3.12, we will also choose |κ1|2 + |κ2|2 sufficiently small

so that certain estimates that follow are valid. In what follows, it is perhaps easiest to

skip this part and proceed onto the proofs that ϕ1, ϕ2, ϕ3, ϕ4, and ϕ5 are Lyapunov

functions on their respective domains. One can then refer back to the subsection of

constants when needed.

3.4.1 Constants

Let

m(κ1, κ2) = minj∈Z

(Im(gjκ1)2 + Im(gjκ2)2

).

Note that since the pair (κ1, κ2) is transversal to Dn−1, we have m(κ1, κ2) > 0. We

can now define positive constants δ, d, and f from this:

δ =m(κ1, κ2)

12(2 + 4

n−1

) (1 + 3

8(n− 1)

)d =

m(κ1, κ2)

12(1 + 3

8(n− 1)

) = 2δ +4δ

n− 1

f =m(κ1, κ2)

12= d+

3

8d(n− 1).

Page 74: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

74

Choose |κ1|2 + |κ2|2 > 0 small enough such that

1. δ ∈(0, 1

2

)2. −

(34n− 1

)+ (|κ1|2 + |κ2|2)3

8

(34d+ 1

)< −1

4

3. 234d ≤ 2

4. η4δ/(n−1)n ≥ 1/2.

We can then choose positive constants C1, C2, . . . , C5 independent of κ1 and κ2:

1. C2 = 2 > 1 = C1

2. C3 = 3 > 2 = C2

3. C4 = 25 > 24 = 8C3.

4. 2C5 = 26 < 25 = C4.

We choose R > 0 sufficiently large as in the previous section and moreover to assure

that the regions U2(α(k)) ∩ BcR(0) are disjoint for all k ∈ Z and such that | sin((n−

1)θz)| ≥ 1/2 in U2.

3.4.2 Lyapunov Functions

Lyapunov Function 1. Let ϕ1(z, z) = C1|z|2δ. Then ϕ1 is a Lyapunov function on

U1.

Proof. Since |z| 6= 0 on U1, it follows that ϕ1 ∈ C∞(U1). Moreover, it is clear that

ϕ1 →∞ as |z| → ∞, z ∈ U1. Thus ϕ1 satisfies properties (I) and (II) on U1. To see

Page 75: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

75

that property (III) is valid on U1, note that

1

C1

Lϕ1(z, z) = 2δ|z|2δ−2 Re (zzn) +1

2δ(δ − 1)(κ2

1 + κ22)|z|2δ−4z2

+ δ2(|κ1|2 + |κ2|2)|z|2δ−2 +1

2δ(δ − 1)(κ1

2 + κ22)z2|z|2δ−4

= 2δ|z|2δ Re(zn−1) +1

2δ(δ − 1)(κ2

1 + κ22)|z|2δ−4z2

+ δ2(|κ1|2 + |κ2|2)|z|2δ−2 +1

2δ(δ − 1)(κ1

2 + κ22)z2|z|2δ−4

≤ 2δ|z|2δ Re(zn−1

)+(δ|δ − 1|+ δ2

) (|κ1|2 + |κ2|2

)|z|2δ−2

= 2δ|z|n+2δ−1 cos((n− 1)θz) +(δ|δ − 1|+ δ2

) (|κ1|2 + |κ2|2

)|z|2δ−2

≤ −δηnC1

|z|n−2ϕ1 +d1

C1

for some constant d1 > 0. Thus we have the bound

Lϕ1(z, z) ≤ −δηn|z|n−2ϕ1(z, z) + d1 (3.13)

≤ −δηnϕ(z, z) + d1

on U1 since n ≥ 2 and |z| ≥ 1 on U1.

Lyapunov Function 2. Let ϕ2(z, z) be a function defined on U2 by:

ϕ2(z, z) = C2

(∣∣Im(zn−1)∣∣ 2δn−1 + 2 Re(eiγ(k)z)

), z ∈ U2(α(k)) ∩BR(0)c

where

γ(k) =

π2− π

2(n−1)− πk

n−1for k even

3π2− π

2(n−1)− πk

n−1for k odd.

Then ϕ2 is a Lyapunov function on U2.

Remark 3.14. This is where we use the choice of δ ∈ (0, 1/2).

Proof. By the definition of U2, we have Im(zn−1) 6= 0 and for k ∈ Z the regions

U2(α1(k))∩BR(0)c are disjoint, hence ϕ2 ∈ C∞(U2). For z ∈ U2, write z = eiα(k)r+w

Page 76: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

76

with r > 0 and |w| ≤ 2. Since | sin((n− 1)θz)| ≥ 1/2 in U2 and δ < 1/2, we have:

1

C2

ϕ2(z, z) ≥ 1

2|z|2δ + 2 Re

(ei(γ(k)+α1(k))r + eiγ(k)w

)≥ 1

2|z|2δ + 0− 4.

Thus ϕ2 satisfies (I) and (II) in U2. Since δ ≤ 1, it is easy to check the second

order terms in C−12 Lϕ2(z, z) are bounded above by a constant E2 > 0. After setting

a = 2δ/(n− 1), we then have

1

C2

Lϕ2(z, z) ≤ E2 + a(n− 1) Sign(Im(zn−1))∣∣Im(zn−1)

∣∣a−1Im(z2n−2)

+ 2 Re(eiγ(k)zn)

≤ E2 + a(n− 1)∣∣Im(zn−1)

∣∣a−1 |z|2n−2

+ 2|z|n cos(nθz + γ(k))

≤ E2 + 21/aa(n− 1)|z|n−1+2δ

+ 2|z|n cos(nθz + γ(k)).

Note that as |z| → ∞, z ∈ U2(α(k)), nθz + γ(k)→ π + 2πj for some j ∈ Z. Thus for

some d′2 > 0 we have the bound for z ∈ U2:

1

C2

Lϕ2(z, z) ≤ −|z|n +d′2C2

.

Therefore there exists d2 > 0 such that:

Lϕ2(z, z) ≤ −C2|z|n + d′2

≤ −|z|n−2δϕ2(z, z) + d2 (3.15)

≤ −ϕ2(z, z) + d2,

on U2 since |z| ≥ 1 on U2.

Lyapunov Function 3. Let

ϕ3(z, z) = C3|z|6δ

|Im(zn−1)|4δn−1

.

Then ϕ3 is a Lyapunov function on U3.

Page 77: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

77

Proof. Note that ϕ3 is smooth in U3 since Im(zn−1) 6= 0 in this region. Moreover,

we have the bound

ϕ3(z, z) ≥ C3|z|2δ.

Thus ϕ3 satisfies properties (I) and (II) in U3. Let, for simplicity, 2b = 6δ and

c = 4δ/(n − 1). To verify (III), note that the second order terms in C−13 Lϕ3(z, z)

are bounded above by

E3(|κ1|2 + |κ2|2)|z|2n+2b−4

|Im(zn−1)|c+2 ,

for some constant E3 > 0 independent of κ1, κ2. Thus we have

1

C3

Lϕ3(z, z) ≤ E3(|κ1|2 + |κ2|2)|z|2n+2b−4

|Im(zn−1)|c+2 + 2b|z|2b

|Im(zn−1)|cRe(zn−1

)− c(n− 1) Sign(Im(zn−1))

|z|2b

|Im(zn−1)|c+1 Im(z2n−2

)=

1

C3

ϕ3(z, z)

[E3(|κ1|2 + |κ2|2)

|z|2n−4

|Im(zn−1)|2+ 2bRe(zn−1)

− c(n− 1) Sign(Im(zn−1))Im(z2n−2)

|Im(zn−1)|

]. (3.16)

Note that, by the definition of U3,

Sign(Im(zn−1)) sin((2n− 2)θz) = 2| sin((n− 1)θz)| cos((n− 1)θz)

= 2| sin((n− 1)θz)|| cos((n− 1)θz)|.

We thus have the following estimate after combining the last two terms in (3.16):

1

C3

Lϕ3(z, z) ≤ 1

C3

ϕ3(z, z)[E3(|κ1|2 + |κ2|2)

|z|2n−4

|Im(zn−1)|2

− 2δ| cos((n− 1)θz)||z|n−1]

=1

C3

ϕ3(z, z)[E3(|κ1|2 + |κ2|2)

|z|−2

|sin((n− 1)θz)|2

− 2δ| cos((n− 1)θz)||z|n−1]

=ϕ3(z, z)

C3| sin((n− 1)θz)|[E3(|κ1|2 + |κ2|2)

|z|−2

| sin((n− 1)θz)|− 2δ| sin((n− 1)θz)|| cos((n− 1)θz)||z|n−1

]. (3.17)

Page 78: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

78

Let x = | cos((n− 1)θz)|. Note that for |z| > 2ηn

| sin((n− 1)θz)|| cos((n− 1)θz)| ≥ minx∈

hηn|z|−1,

√1−η2

n|z|−2ix√1− x2

≥ ηn|z|−1√

1− η2n|z|−2

≥√

3

2ηn|z|−1.

Applying this to (3.17), we obtain:

1

C3

Lϕ3(z, z) ≤ ϕ3(z, z)

C3| sin((n− 1)θz)|[E3(|κ1|2 + |κ2|2)η−1

n |z|−1 −√

3ηnδ|z|n−2]

From this we infer that there exists a constant d3 > 0 such that

Lϕ3(z, z) ≤ −√

3ηnδ

2|z|n−2ϕ3(z, z) + d3 (3.18)

≤ −√

3ηnδ

2ϕ3(z, z) + d3,

as |z| ≥ 1 on U3.

Before we proceed onto the fourth and fifth Lyapunov functions, we first need a

lemma:

Lemma 3.19. Let g be a primitive (n− 1)st root of unity and for k ∈ Z let gk · κ =

(gkκ1, gkκ2). Suppose that ϕ is a Lyapunov function on U ⊂ C corresponding to the

operator Lgk·κ for all k ∈ Z. Then ψk(z, z) = ϕ(g−kz, g−kz) is a Lyapunov function

on gkU corresponding to the operator Lgj ·κ for all j ∈ Z.

Proof. Fix k ∈ Z and note that ψk satisfies both (I) and (II) in gkU . To see (III),

we have for w = g−kz and z ∈ gkU :

Lgj ·κψk(z, z) = Lg

j−k·κϕ(w, w)

≤ −Cψk(z, z) +D,

on gkU for some positive constants C,D.

Page 79: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

79

Note that the positive constants C,D in the previous lemma can depend on j ∈ Z

as in the diffusion parameter gj ·κ. Since g is an (n−1)st root of unity, there are only

finitely many distinct diffusion parameters. Hence we may make a uniform choice in

these constants.

In what follows, it is more convenient to prove property (III) in the (x, y) vari-

ables. This is the case because, in U4(0) and U5(0), the variable x plays the dominant

role since y is bounded. We shall thus use the expression:

L = Re(zn)∂

∂x+ Im(zn)

∂y+

1

2(Re(κ1)2 + Re(κ2)2)

∂2

∂x2

+(Re(κ1) Im(κ1) + Re(κ2) Im(κ2))∂2

∂x∂y+

1

2(Im(κ1)2 + Im(κ2)2)

∂2

∂y2.

for the generator L of the process z(t) = x(t) + iy(t).

Lyapunov Function 4. Let d > 0 and define

ϕ4,0(z, z) = ϕ4,0(x, y) = C4xd + |y|d

|y| 34d.

For g, a primitive (n− 1)st root of unity, define

ϕ4(z, z) = ϕ4,0(g−kz, g−kz) if z ∈ gkU4(0).

Then ϕ4 is a Lyapunov function on U4.

Remark 3.20. This is where we use the choice

−(

3

4n− 1

)+ (|κ1|2 + |κ2|2)

3

8

(3

4d+ 1

)< −1

4

Proof. Note by the previous lemma, it suffices to prove that ϕ4,0 is a Lyapunov

function in U4(0) corresponding to the operator Lgk·κ for all k ∈ Z. Note that ϕ4,0

Page 80: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

80

clearly satisfies (I) and (II) in U4(0). Note that we have:

1

C4

Lgk·κϕ4(x, y)

= dxd−1

|y| 34dRe(zn)

+

(−3d

4Sign(y)

xd

|y| 3d4 +1+d

4Sign(y)|y|

d4−1

)Im(zn)

+d(d− 1)

2

(Re(gkκ1)2 + Re(gkκ2)2

) xd−2

|y| 34d

− 3

4d2 Sign(y)

(Re(gkκ1) Im(gkκ1) + Re(gkκ2) Im(gkκ2)

) xd−1

|y| 34d+1

+(Im(gkκ1)2 + Im(gkκ2)2

)(3

8d

(3

4d+ 1

)xd

|y| 34d+2+d

8

(d

4− 1

)|y|

d4−2

)

=xd

|y| 34d

[d

xRe(zn)

+

(−3d

4Sign(y)|y|−1 +

d

4Sign(y)|y|d−1x−d

)Im(zn)

+d(d− 1)

2(Re(gkκ1)2 + Re(gkκ2)2)x−2

−3

4d2 Sign(y)(Re(gkκ1) Im(gkκ1) + Re(gkκ2) Im(gkκ2))x−1|y|−1

+ (Im(gkκ1)2 + Im(gkκ2)2)

(3

8d

(3

4d+ 1

)|y|−2 +

d

8

(d

4− 1

)x−d|y|d−2

)]

=xd

|y| 34d

(− d

(3

4n− 1

)xn−1 + (Im(gkκ1)2 + Im(gkκ2)2)

3

8d

(3

4d+ 1

)|y|−2

+O(xn−1−min(1,d))

)

≤ xd

|y| 34d

(−d

4xn−1 +O(xn−1−min(1,d))

).

Thus there exists a constant d4 > 0 such that

Lϕ4(x, y) ≤ −d8|z|n−1ϕ4(x, y) + d4 (3.21)

≤ −d8ϕ4(x, y) + d4,

Page 81: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

81

as |z| ≥ 1 in U4.

Lyapunov Function 5.

ϕ5,0(z, z) = ϕ5,0(x, y) = C5(6xf − y2xf+n−1).

For g, a primitive (n− 1)st root of unity, define

ϕ5(z, z) = ϕ5,0(g−kz, g−kz), for z ∈ gkU5(0).

Then ϕ5 is a Lyapunov fuction on U5.

Remark 3.22. So far, this is the only point in the argument where we need em-

ploy our assumption that m(κ1, κ2) > 0. Intuitively, since U5 encloses the unstable

trajectories of the ODE system (1.1), this should be expected. The assumption

m(κ1, κ2) > 0 says that if we start on an unstable trajectory, noise kicks us off.

Here, we also use the choice

f =1

12m(κ1, κ2).

Proof. Again we apply the previous lemma and focus our attention on ϕ5,0. Note

clearly ϕ5,0 satisfies (I) in U5,0. Moreover,

1

C5

ϕ5 ≥ 2xf .

Thus ϕ5,0 satisfies (II) in U5. Note that

1

C5

Lgj ·κϕ5(x, y) = (6fxf−1 − (n+ f − 1)y2xn+f−2) Re(zn)

− 2yxn+f−1 Im(zn)

− (Im(gjκ1)2 + Im(gjκ2)2)xn+f−1 +O(xn+f−2)

≤ (6f − (Im gj(κ1)2 + Im(gjκ2)2))xn+f−1 +O(xn+f−2)

≤ −1

2m(κ1, κ2)xn+f−1 +O(xn+f−2).

Page 82: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

82

Thus there exists a constant d5 > 0 such that

Lϕ5(x, y) ≤ −1

4m(κ1, κ2)|z|n−1ϕ5(x, y) + d5 (3.23)

≤ −1

4m(κ1, κ2)ϕ5(x, y) + d5,

as |z| ≥ 1 in U5.

3.5 Stage 2: Gluing

In this section, we finish constructing Φ by piecing together the Lyapunov covering

(ϕ1, U1), (ϕ2, U2), . . . , (ϕ5, U5) exhibited in the previous section. For j = 1, 2, 3, 4,

the idea is to define smooth auxiliary functions

ρj,j+1 : Uj ∪ Uj+1 → [0, 1]

such that ρj,j+1|Uj\Uj∩Uj+1= 0 and ρj,j+1|Uj+1\Uj∩Uj+1

= 1 and argue that

ϕj,j+1(z, z) := ρj,j+1ϕj+1 + (1− ρj,j+1)ϕj (3.24)

is now a Lyapunov function on the larger domain Uj ∪ Uj+1. If this holds, for R > 0

sufficiently large we may then choose a smooth function Φ on C such that

Φ(z, z) =

ϕj,j+1(z, z) for z ∈ Uj ∪ Uj+1

arbitrary smooth for |z| ≤ R− ε,

for fixed 0 < ε < R. By construction, Φ is a Lyapunov function on all of C.

By expression (3.24), we note that properties (I) and (II) are easily satisfied on

Uj ∩ Uj+1 for j = 1, 2, 3, 4. Thus to finish, all we must verify is property (III) on

Uj ∩ Uj+1 for j = 1, 2, 3, 4 and we shall do so in that order.

Patch 1. We first define ρ1,2.

Choose f(z, z) ≥ 0 to be a smooth function on R≥1 := Re(z) ≥ 1 such that

f(z, z) =

1 if Im(z) ≤ 1

0 if Im(z) ≥ 2,

Page 83: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

83

and such that:

∂yf(z, z) < 0 for y ∈ (1, 2)

∂xf(z, z) = 0 for x ≥ 1,

where x = Re(z) and y = Im(z). Choose g(z, z) ≥ 0 to be a smooth function on R≥1

such that

g(z, z) =

0 if Im(z) ≤ −2

1 if Im(z) ≥ −1,

and such that

∂yg(z, z) > 0 for y ∈ (−1,−2)

∂xg(z, z) = 0 for x ≥ 1,

where x = Re(z) and y = Im(z). Hence we define

ρ1,2(z, z) =

f(e−iα(k)z, eiα(k)z) for z ∈ eiα(k)R≥1 if k is even

g(e−iα(k)z, eiα(k)z) for z ∈ eiα(k)R≥1 if k is odd.

Claim 1. ϕ1,2 given by equation (3.24) is a Lyapunov function on U1 ∪ U2.

Remark 3.25. Here we use the choice

C2 = 2 > C1 = 1.

Proof. Recall that it is sufficient to verify property (III) on the intersection U1∩U2.

Note that we have:

Lϕ1,2(z, z) = ρ1,2Lϕ2 + (1− ρ1,2)Lϕ1 + (ϕ2 − ϕ1)(zn∂z(ρ1,2) + zn∂z(ρ1,2))

+1

2(κ2

1 + κ22) ((ϕ2 − ϕ1)∂zz(ρ1,2) + 2∂z(ϕ2 − ϕ1)∂z(ρ1,2))

+ (|κ1|2 + |κ2|2)

((ϕ2 − ϕ1)∂zz(ρ1,2) + ∂z(ϕ2 − ϕ1)∂z(ρ1,2)

+ ∂z(ϕ2 − ϕ1)∂z(ρ1,2)

)+

1

2(κ1

2 + κ22) ((ϕ2 − ϕ1)∂zz(ρ1,2) + 2∂z(ϕ2 − ϕ1)∂z(ρ1,2)) .

Page 84: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

84

It is easy to check that the last four lines of the expression above are O(ρ1(z, z)|z|2δ)

where ρ1 is a bounded smooth function such that ρ1|U1\U1∩U2= ρ1|U2\U1∩U2 = 0. More-

over, by construction of f and g, ρ1(z, z)→ 0 uniformly in |z| as z → (U1 \U1 ∩U2)∪

(U2 \ U1 ∩ U2). Combining this with estimates (3.13) and (3.15), we obtain:

Lϕ1,2(z, z) ≤ −min(1, δηn)|z|n−2ϕ1,2(z, z) + (ϕ2 − ϕ1)(zn∂z(ρ1,2) + zn∂z(ρ1,2))

+O(ρ1(z, z)|z|2δ) + max(d1, d2).

Suppose first that k is even. Since we are only interested in estimating on the inter-

section U1 ∩U2, we may suppose that nθz −α(k)→ (n− 1)α(k) as |z| → ∞ for some

k ∈ Z even. Hence for w = e−iα(k)z we have:

zn∂z(ρ1,2) + zn∂z(ρ1,2) = |z|n(einθz−iα(k)fw(w, w) + e−inθz+iα(k)fw(w, w)

).

Since nθz − α(k)→ (n− 1)α(k) = π/2 + πk (and k is EVEN) as |z| → ∞ and

ifw(w, w)− ifw(w, w) =∂

∂yf(w, w) < 0

where y = Im(w), we have that

Lϕ1,2(z, z) ≤ −min(1, δηn)|z|n−2ϕ1,2(z, z)− 1

2|z|n+2δ |∂y(f(w, w))|

+O(ρ1(z, z)|z|2δ) + max(d1, d2).

Since ∂x(f(w, w)) = 0 where x = Re(w), the term −12|z|n+2δ |∂y(f(w, w))| dominates

the term O(ρ1(z, z)|z|2δ) until we are arbitrarily close to ∂(U1∩U2); in which the first

term −min(1, δηn)|z|n−2ϕ1,2 dominates O(ρ1(z, z)|z|2δ) since ρ1 is smooth an vanishes

on U1 \ U1 ∩ U2 and U2 \ U1 ∩ U2. From this it is clear that there exists a positive

constant d1,2 such that

Lϕ1,2(z, z) ≤ −min(1, δηn)

2ϕ1,2 + d1,2.

If k is odd, we may suppose that nθz − α(k)→ (n− 1)α(k) = π/2 + πk as |z| → ∞

for some k ∈ Z odd. Using the same argument, we realize that for |z| sufficiently

Page 85: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

85

large and w = e−iα(k)z:

zn∂z(ρ1,2) + z∂z(ρ1,2) = |z|n(einθz−iα(k)gw(w, w) + e−inθz+iα(k)gw(w, w)

)≤ −|z|n ∂

∂yg(w, w)

= −|z|n∣∣∣∣ ∂∂yg(w, w)

∣∣∣∣ .The rest of the argument follows through similarly.

Patch 2. We now define ρ2,3.

We will again use the functions f and g defined in Patch 1. Shifting f down

three units and g up three units, let f−3(z, z) = f(z + 3i, z − 3i) and g+3(z, z) =

g(z − 3i, z + 3i). Define

ρ2,3(z, z) =

f−3(e−iα(k)z, eiα(k)z) for z ∈ eiα(k)R≥1 if k is even

g+3(e−iα(k)z, eiα(k)z) for z ∈ eiα(k)R≥1 if k is odd.

Claim 2. ϕ2,3 defined by equation (3.24) is a Lyapunov function on U2 ∪ U3.

Remark 3.26. Here we use the choice

C3 = 3 > C2 = 2.

Proof. Proceeding in a similar fashion as in the proof of Claim 1, we see that by the

estimates (3.15) and (3.18), since 0 < δ < 1/2 we have on U2 ∩ U3:

Lϕ2,3(z, z) ≤ −min

(1,

√3ηnδ

2

)|z|n−2ϕ2,3(z, z) + (ϕ3 − ϕ2)(zn∂z(ρ2,3) + zn∂z(ρ2,3))

+O(ρ2(z, z)|z|2δ) + max(d2, d3),

where ρ2 is a bounded smooth function such that ρ2|U2\U2∩U3 = ρ2|U3\U2∩U3 = 0.

Moreover, by construction of f and g, ρ2(z, z) → 0 uniformly in |z| as z → (U2 \

U2 ∩ U3) ∪ (U3 \ U2 ∩ U3). Using exactly the same argument as before, we can find a

constant d2,3 > 0 such that

Lϕ2,3(z, z) ≤ −min

(1

2,

√3ηnδ

4

)|z|n−2ϕ2,3(z, z) + d2,3

Page 86: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

86

on U2 ∩ U3, as claimed.

Patch 3. We now define ρ3,4.

Let

ρ3,4(z, z) =

f(e−iβ(k)z, eiβ(k)z) for z ∈ eiβ(k)R≥1 if k is even

g(e−iβ(k)z, eiβ(k)z) for z ∈ eiβ(k)R≥1 if k is even.

Note that k is only even since cos((n− 1)θz) > 0 on U3 ∪ U4.

Claim 3. ϕ3,4 defined by equation (3.24) is a Lyapunov function on U3 ∪ U4.

Remark 3.27. Here we use the choice:

C4

234d≥ C4

2= 25/2 > 12 = 4C3 >

2C3

η4δ/(n−1)n

Proof. Applying Lemma 3.19, it suffices to show that ϕ3,4 is a Lyapunov function

on U3 ∩U4(0) for all diffusion parameters gj ·κ for j ∈ Z and κ fixed. Proceeding as

in Claim 1 and Claim 2, by the estimates (3.18) and (3.21) we have:

Lϕ3,4(z, z) ≤ −2c3,4|z|n−2ϕ3,4 + (ϕ4 − ϕ3)(zn∂z(ρ3,4) + zn∂z(ρ3,4))

+O(ρ3(z, z)|z|2δ+4δn−1 ) + max(d3, d4)

on U3 ∩ U4(0) where c3,4 > 0 is a positive constant and ρ3 is a bounded smooth

function on U3 ∪ U4(0) such that ρ3|U3\U3∩U4(0)) = ρ3|U4(0)\U3∩U4(0) = 0. Moreover, by

construction of f and g, ρ3(z, z) → 0 uniformly in |z| as z → (U3 \ U3 ∩ U4(0)) ∪

(U4(0) \U3 ∩U4(0)). Comparing with the proofs of Claim 1 and Claim 2, we see that

the only difference thus far is the asymptotic behavior change in both ϕ3,4 and the O

terms which are, conveniently, both of order 2δ + 4δ/(n − 1). To estimate the term

(ϕ4−ϕ3)(zn∂z(ρ3,4) + zn∂z(ρ3,4)), note that if z ∈ U3∩U4(0) with y = Im(z) > 0 and

Page 87: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

87

|z| sufficiently large we have

zn∂z(ρ3,4) + zn∂z(ρ3,4) = |z|n sin(nθz)∂

∂yf(z, z)

= |z|n(sin((n− 1)θz) cos(θz) + sin(θz) cos((n− 1)θz))∂

∂yf(z, z)

≤ −ηn2|z|n−1

∣∣∣∣ ∂∂y f(z, z)

∣∣∣∣ ,since θz ↓ 0 as |z| → ∞ in this region. If on the other hand y = Im(z) < 0, we have

the same estimate upon replacing f with g. From this point, using the same reasoning

as in Claim 1 and Claim 2, we finish the proof of Claim 3.

Patch 4. We now define ρ4,5. Because U4 ∩U5 is a region that consists of decreasing

strips instead of boxes, this will be the most involved patch.

Define an interpolating function p : [1,∞)× R→ [0, 1] by

pn(x, y) =

1 if |x|n−1

2 |y| ≤ 1

2− |x|n−12 |y| if 1 < |x|n−1

2 |y| < 2

0 if 2 ≤ |x|n−12 |y|.

To smooth the sharp corners of this function, introduce

r(t) =

e−

1t(1−t) if 0 < t < 1

0 otherwise,

and

s(x) =1

N

∫ x

−∞r(t)dt,

where N =∫

R r(t)dt. We then let

ρ4,5,0(x, y) = s(pn(x, y)) on x ≥ 1,

which is now a smooth function on R≥1. Rotating ρ4,5,0(x, y) = ρ4,5,0(z, z) we define

ρ4,5(z, z) = ρ4,5,0(e−iβ(k)z, eiβ(k)z) for z ∈ eβ(k)R≥1 for k even.

Claim 4. ϕ4,5 defined by equation (3.24) is a Lyapunov function on U4 ∪ U5.

Page 88: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

88

Remark 3.28. Here we use the choice

2C5 = 26 < 25 = C4.

Proof. We will again choose |κ1|2 + |κ2|2 smaller than before which is made precise

below. Applying Lemma 3.12 and Lemma 3.19, it suffices to prove that ϕ4,5,0 is a

Lyapunov function on U4(0) ∩ U5(0) for all diffusion parameters gj · κ where j varies

through Z. Thus, using the variables (x, y), we obtain the expression on U4(0)∩U5(0):

Lϕ4,5,0(x, y) = ρ4,5,0Lϕ5,0 + (1− ρ4,5,0)Lϕ4,0

+ Re(zn)(ϕ5,0 − ϕ4,0)∂x(ρ4,5,0) + Im(zn)(ϕ5,0 − ϕ4,0)∂y(ρ4,5,0)

+1

2(Re(κ1)2 + Re(κ2)2)

[(ϕ5,0 − ϕ4,0)∂xx(ρ4,5,0)

+ 2∂x(ϕ5,0 − ϕ4,0)∂x(ρ4,5,0)

]+ (Re(κ1) Im(κ1) + Re(κ2) Im(κ2))

[(ϕ5,0 − ϕ4,0)∂xy(ρ4,5,0)

+ ∂y(ϕ5,0 − ϕ4,0)∂x(ρ4,5,0) + ∂x(ϕ5,0 − ϕ4,0)∂y(ρ4,5,0)

]+

1

2(Im(κ1)2 + Re(κ2)2)

[(ϕ5,0 − ϕ4,0)∂yy(ρ4,5,0)

+ 2∂y(ϕ5,0 − ϕ4,0)∂y(ρ4,5,0)

].

By definition of ρ4,5,0, the term on the third and fourth lines above is O(|x|f−2) on

U4(0) ∩ U5(0) and term on the fifth and sixth lines above is O(|x|f+n−32 ) on U4(0) ∩

U5(0). The last two terms are bounded above by:

E(|κ1|2 + |κ2|2)(n(p)q(p) + n(p))|x|f+n−1,

where E > 0 is a constant independent of κ1 and κ2, n(x) = N−1r(x), q(x) =

|1− 2x|/(x−x2)2 and p = pn(x, y). Hence using the strong bounds (3.21) and (3.23),

Page 89: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

89

there exists a constant c4,5 > 0 independent of κ1 and κ2 such that on U4(0) ∩ U5(0):

Lϕ4,5,0(x, y) ≤ −c4,5m(κ1, κ2)|x|f+n−1 + Re(zn)(ϕ5,0 − ϕ4,0)∂x(ρ4,5,0)

+ Im(zn)(ϕ5,0 − ϕ4,0)∂y(ρ4,5,0)

+E(|κ1|2 + |κ2|2)(n(p)q(p) + n(p))|x|f+n−1

+O(|x|f+n−32 )

≤ −c4,5mn(κ1, κ2)|x|f+n−1 − n(p)(ϕ5,0 − ϕ4,0)|x|n−1

+E(|κ1|2 + |κ2|2)(n(p)q(p) + n(p))|x|f+n−1

+O(|x|f+n−32 ).

Note that for all λ > 0:

|κ1|2 + |κ2|2

m(κ1, κ2)=|λκ1|2 + |λκ2|2

m(λκ1, λκ2).

Thus regardless of how much we decrease the magnitude |κ1|2 + |κ2|2, the ratio above

remains unchanged. Since both n(p) and n(p)q(p) approach 0 as |x|n−12 |y| approaches

1 or 2, the above relation implies there exists ε > 0 independent of both κ1 and κ2

such that

−c4,5m(κ1, κ2) + E(|κ1|2 + |κ2|2)(n(p)q(p) + n(p)) ≤ −c′4,5 < 0,

for all (x, y) such that |x|n−12 |y| ∈ [1, 1 + ε) ∪ (2 − ε, 2]. Moreover it is clear that

choosing |κ1|2 + |κ2|2 sufficiently small we can assure that

−n(p)(ϕ5,0 − ϕ4,0)|x|n−1 + E(|κ1|2 + |κ2|2)(n(p)q(p) + n(p))|x|f+n−1 < 0,

for all (x, y) such that |x|n−12 |y| ∈ [1 + ε, 2− ε]. Note that this finishes the result for

now there exists d4,5 > 0 such that

Lϕ4,5,0(x, y) ≤ −c′4,52ϕ4,5,0(x, y) + d4,5,

as required.

Page 90: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

90

3.6 Uniqueness of µ and Geometric Ergodicity

Now that we have finished constructing a smooth function Φ : C → [0,∞) that

satisfies (C1) and (C3), our goal is to prove that the invariant probability measure µ

is unique and the process z(t) is exponentially ergodic. From what follows, uniqueness

is easily established by Remark 2.45 and the results of Section 2.5. We will thus focus

on showing part 3 of Lemma 3.1. We note that by the existence of Φ : C→ [0,∞) that

satisfies (C1) and (C3) and the results of Section 2.6, it sufficies to show Assumption

2.63 is valid for z(t), i.e., we prove:

Theorem 3.29. There exists a distinguished time T0 > 0 such that for all R > 0

sufficiently large, there exists αR ∈ (0, 1) and a probability measure νR such that

infz∈CR

P (z, T0, · ) ≥ αRνR( · ),

where CR = z ∈ C : Φ(z, z) ≤ R.

We split the proof of the theorem above into two lemmata:

Lemma 3.30. There exists T0 > 0 and non-empty open U ⊂ C such that

supp(P (z, T0, · )) ⊃ U for all z ∈ C.

Lemma 3.31. For all t > 0 and z ∈ C

P (z, t, dw) = p(z, t, w) dw,

where dw is Lebesgue measure on R2 and p is a smooth function on R2× (0,∞)×R2.

Proof that Lemma 3.30 and Lemma 3.31 =⇒ Theorem 3.29. Here we follow

the appendix in [MSH02]. Pick R > 0 large enough so that U ∩ int(CR) 6= ∅. Fix

z∗ ∈ U ∩ int(CR) and δ > 0 such that Bδ(z∗) ⊂ int(CR). By Lemma 3.30, we have:

P (z∗, T0, Bδ(z∗)) > 0.

Page 91: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

91

By Lemma 3.31, we have:

p(z∗, T0, w∗) ≥ 2ε,

for some ε > 0 and w∗ ∈ Bδ(z∗). By Lemma 3.31 again, we obtain:

p(z, T0, w) ≥ ε for all (z, w) ∈ Bε1(z∗)×Bε2(w

∗),

for some ε1, ε2 > 0 where ε2 > 0 is also chosen such that Bε2(w∗) ⊂ CR. Thus for all

z ∈ Bε1(z∗) and A ∈ B(C), we have:

P (z, T0, A) =

∫A

p(z, T0, w) dw

≥∫A∩Bε2 (w∗)

p(z, T0, w) dw

≥ ε λ (A ∩Bε2(w∗))

where λ is Lebegue measure on C. Since CR is compact, by Lemma 3.30 and Lemma

3.31 we have:

infz∈CR

P (z, T0, Bε1(z∗)) ≥ ζ,

for some ζ > 0. Define T0 = 2T0 and note that for all z ∈ CR and A Borel, we have:

P (z, T0, A) =

∫CP (z, T0, dw)P (w, T0, A)

≥∫Bε1 (z∗)

p(z, T0, w)P (w, T0, A) dw

≥ ε λ(A ∩Bε2(w∗))

∫Bε1 (z∗)

p(z, T0, w) dw

= ε λ(A ∩Bε2(w∗))P (z, T0, Bε1(z

∗))

≥ ε ζ λ(A ∩Bε2(w∗))

= ε ζ λ(Bε2(w∗)) ν(A)

where ν(A) = λ(Bε2(w∗))−1λ(A ∩ Bε2(w

∗)). Note that this finishes the proof since ν

is indeed a probability measure.

Page 92: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

92

As we will see, Lemma 3.30 follows from the geometric techniques of Section 2.5

and Lemma 3.31 is a simple consequence of a deep result of Hormander [Hor67]. Both

arguments, however, employ the fact that the the for all z ∈ C, the span of the Lie

algebra generated by the polysystem:

F = Zn + u1κ1 + u2κ2 : u1, u2 ∈ R,

at z ∈ C, where Zn, κ1, κ2 are vector fields on R2 determined by Zn(z) = zn,

κ1(z) = κ1, κ2(z) = κ2, is the whole tangent space. In the first lemma, this is used

to validate one hypothesis of Theorem 2.59. In the latter, it is used to show Lemma

3.31 as stated. With both results in mind, we first show Lemma 3.31 by verifying the

Lie algebra generated by F spans the whole tangent space, which in this case is C,

at all points.

Proof of Lemma 3.31. Note that if κ1 and κ2 are linearly independent over R,

there is nothing to prove since κ1 and κ2 span the tangent space at all points. Suppose

that κ1 = c κ2 for some c ∈ R. Since (κ1, κ2) is transversal to Dn−1 we may assume

κn−11 /∈ R. For vector fields X and Y , we let adX(Y ) = [X, Y ] and for k ∈ N

k ≥ 2, let adkX(Y ) = adk−1X(adX(Y )). Computing Lie brackets (in R2) we obtain

adn κ1(Zn) = n!κn1 , where κn1 is the vector field on R2 determined by n!κn1 (z) = n!κn1 .

Since κ1 and κn1 are linearly independent over R, this finishes the proof.

To prove Lemma 3.30, let us distinguish between two cases; the first of which is

more straightforward than the second.

Case 1. n ≥ 2 is odd or κ1 and κ2 are linearly independent over R.

Case 2. n ≥ 2 is even and κ1 = c κ2 for some c ∈ R.

We will first prove Lemma 3.30 in Case 1, as we now have the techniques to do

so. Moreover, the argument will illustrate the difference between Case 1 and Case 2.

Proof of Lemma 3.30 in Case 1. Suppose first that κ1 and κ2 are linearly inde-

Page 93: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

93

pendent over R. We note that by Theorem 2.51 for all u1, u2 ∈ R

u1κ1 = limλ→∞

1

λ(Zn + λu1κ1) ∈ Sat(F )

u2κ2 = limλ→∞

1

λ(Zn + λu2κ2) ∈ Sat(F ).

Hence for all z ∈ C and all T > 0, by the linearly independence assumption

AF (z,≤ T ) = C.

By Theorem 2 on p. 68 of [Jur97],

AF (z,≤ T ) = C.

Hence by Theorem 2.42 and Theorem 2.59,

supp(P (z, T, · )) = AF (z, T ) = C,

for all z ∈ C, T > 0. Suppose now that n is odd and κ1 = c κ2 for some c ∈ R. We

may suppose without loss of generality that κn−11 /∈ R. Using Theorem 2.51 for all

u1 ∈ R we have:

u1κ1 = limλ→∞

1

λ(Zn + λu1κ1) ∈ Sat(F ).

From this, one can check that for all u1 ∈ R, exp(u1κ1)(z) = z+u1κ1 ∈ Norm(Sat(F )).

Therefore, by Theorem 2.56 exp(u1κ1)#(Zn) ∈ Sat(F ). Computing the vector field

exp(u1κ1)#(Zn), we obtain:

exp(u1κ1)#(Zn)(z) = Zn(z − u1κ1)

=n∑j=0

(n

j

)zn−j(−1)juj1κ

j1.

Thus we determine

limλ→∞

1

λnexp(u1λκ1)#(Zn) = (−1)nun1κ

n1 ∈ Sat(F ),

Page 94: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

94

where κn1 (z) = κn1 . Since n is odd, we infer that u1κn1 ∈ Sat(F ) for all u1 ∈ R. Since

κ1 and κn1 are linearly independent over R, we see that for all z ∈ C and T > 0:

AF (z,≤ T ) = C.

Using the same reasoning as before, we obtain:

supp(P (z, t, · )) = AF (z, T ) = C.

This finishes the proof of Case 1.

We note that in Case 1 of Lemma 3.30, T0 > 0 can be chosen to be any positive

time and U can be chosen to be the whole space. To see what changes in Case 2,

suppose now that n is even and κ1 = c κ2 for some c ∈ R. We may again suppose

without loss of generality that κn−11 /∈ R. Proceeding in exactly the same way as in

the proof of the second part of Case 1, we see that u1κ1 ∈ Sat(F ) and (−1)nun1κn1 ∈

Sat(F ) for all u1 ∈ R. Since n is even, we may only deduce that λκn1 ∈ Sat(F ) for

all λ ≥ 0. Hence for all z ∈ C and T > 0 we may only determine that AF (z,≤ T )

contains

H(z, κ1) = z+ uκ1 + λκn1 : u ∈ R, λ ≥ 0,

which, since κ1 and κn1 are linearly independent over R, is a half-plane that depends

on z. Note that by Theorem 2 on p. 68 of [Jur97], we deduce

AF (z,≤ T ) ⊃ int(H(z, κ1))

for all z ∈ C and all T > 0. Thus by Remark 2.45, this is sufficient to conclude

uniqueness of µ in Case 2, but we must work a little harder to obtain exponential

ergodicity.

In pursuit of the conclusion of Lemma 3.30 in Case 2, there are two problems with

the above. First, H(z, κ1) depends on the initial point z ∈ C. Second, supposing

Page 95: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

95

that we are able to remove this dependence, we still must transfer between the sets

AF (z,≤ T ) and AF (z, T ).

We will be able to get rid the dependence on z in H(z, κ1) in the following way.

We will show that for all ε > 0, there exists a time T ′0 > 0 such that for all initial

conditions z ∈ C, we can use trajectories of F to control a solution starting from z

into Bε(0) for all T ≥ T ′0. It is important to note that the choice of T ′0 > 0 does not

depend upon z, but only on the size ε of the ball. Hence if z(ε, κ1) ∈ ∂Bε(0)∩H(0, κ1)

is such that z(ε, κ1) ⊥ κ1, we obtain for all t > 0 and for all z ∈ C:

AF (z, T ′0 ≤ t) ⊃ H(z(ε, κ1), κ1),

where AF (z, T ′0 ≤ t) is the set of points that can be reached from z using trajectories

in F at some time in the interval [T ′0, T′0 + t].

Although it seems from this that we should be able to deduce that

AF (z, T ′0) ⊃ H(z(ε, κ1), κ1),

it is not immediate. By the proof of Theorem 2.59, what we have left to show is that

z0 ∈ AF (z0,≤ T ) for all T > 0,

for all points z0 ∈ Bε(0) which are images of the trajectories of F that initially guided

us into Bε(0).

Let us now proceed using the ideas above. First note that by separation of vari-

ables for z 6= 0 ∈ C:

(exp(tZn)(z))n−1 = − 1((n− 1)t− zn−1

|z|2n−2

) . (3.32)

Thus if z ∈ Dn−1, exp(tZn)(z) is only defined locally in time. For z 6= 0 otherwise,

however, exp(tZn)(z) is strongly dissipative. More precisely:

Proposition 3.33. For all ε > 0 there exists T1(ε) > 0 such that

|exp(tZn)(z)| ≤ ε for all t ≥ T1(ε) (3.34)

Page 96: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

96

for all |z| ≥ ε such that z /∈ Dn−1.

Proof. For |z| ≥ ε such that z /∈ Dn−1, let w = zn−1/|z|2n−2. Note that |Re(w)| ≤

1/εn−1. Pick then T1(ε) = 2(n−1)εn−1 and note that for t ≥ T0(ε) we have:

|exp(tZn)(z)|2n−2 =1

((n− 1)t− Re(w))2 + Im(w)2

≤ ε2n−2.

In the previous proposition, we only used the vector field Zn. For initial points

z elsewhere besides |z| ≥ ε and z /∈ Dn−1, we will use more of the polysystem F to

control z into the set where |z| ≥ ε and z /∈ Dn−1. This is illustrated in the next two

propositions. Again we recall that κ1 = c κ2 for some c ∈ R and we assume without

loss of generality that κn−11 /∈ R.

Proposition 3.35. For all ε > 0, |z| ≤ ε, and T ′ > 0 there exists u > 0 large enough

such that

|exp(t(Zn + uκ1))(z)| > ε for some t ∈ (0, T ′).

Proof. Let ε > 0, |z| ≤ ε, and T ′ > 0. Suppose to the contrary that for all u > 0,

|exp(t(Zn + uκ1))(z)| ≤ ε for all t ∈ (0, T ′).

We then have the estimate:

|exp(t(Zn + uκ1))(z)− z − uκ1t| ≥ u|κ1|t− 2ε,

for all u > 0, t ∈ (0, T ′). Since exp(t(Zn + uκ1)) is an integral curve, we obtain:

u|κ1|t ≤ 2ε+ |exp(t(Zn + uκ1))(z)− z − uκ1t|

= 2ε+

∣∣∣∣∫ t

0

Zn (exp(s(Zn + uκ1))(z)) ds

∣∣∣∣= 2ε+

∣∣∣∣∫ t

0

(exp(s(Zn + uκ1))(z))n ds

∣∣∣∣≤ 2ε+ εnt,

Page 97: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

97

for all u > 0, t ∈ (0, T ′), a contradiction.

Proposition 3.36. For all ε > 0, u > 0 and |z| ≥ ε such that z ∈ Dn−1, we have:

exp(t(Zn + uκ1))(z) /∈ Dn−1 for some t ∈ (0, T ′),

for all T ′ ≤ Tmax where Tmax > 0 is the maximal time of definition for exp(t(Zn +

uκ1))(z).

Proof. Let ε > 0, u > 0, T ′ ≤ Tmax, and |z| ≥ ε with z ∈ Dn−1. Suppose to the

contrary that (exp(t(Zn + uκ1))(z))n−1 > 0 for all t ∈ (0, T ′) and let g be a primitive

(n − 1)st root of unity. By continuity, there exists j ∈ Z such that gj exp(t(Zn +

uκ1))(z) ∈ R for all t ∈ [0, T ′). But note this implies:

gj exp(t(Zn + uκ1))(z)− gjz = gj∫ t

0

(exp(s(Zn + uκ1))(z))n ds+ gjuκt ∈ R

=

∫ t

0

(gj exp(s(Zn + uκ1))(z)

)nds+ gjuκt ∈ R,

for all t ∈ (0, T ′). In particular, gjκ1 ∈ R, a contradiction.

Let us collect the previous three propositions into a Lemma.

Lemma 3.37. For all ε > 0, there exist a time T ′1 = T ′1(ε) > 1 such that for all

z ∈ C, there exist vector fields Y1, Y2, Y3 ∈ F and times t1, t2 ≥ 0 such that t1 + t2 ≤ 1

and

|exp(tY3) exp(t2Y2) exp(t1Y1)(z)| ∈ (0, ε] for all t ≥ T ′1 − 1

where the last vector field Y3 can always be chosen to be Zn ∈ F . Moreover, for all

t ≥ T ′1 − 1 the path

exp(tY3) exp(t2Y2) exp(t1Y1)(z) ∈ Dcn−1.

Page 98: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

98

Proof. Let ε > 0 and pick T1(ε) such that the first proposition holds. Take T ′1(ε) =

T1(ε) + 1. Thus for all |z| ≥ ε with z /∈ Dn−1, the conclusions of the lemma hold by

taking Y1 = Y2 = Y3 = Zn and t1 = t2 = 0. If |z| ≤ ε + 1, by the second proposition

there exists a u1 > 0 such that

|exp(t1(Zn + u1κ1))(z)| ∈ (ε+ 1, R),

for some t1 ∈ (0, 1/3) and R > ε + 1. If exp(t1(Zn + u1κ1))(z) /∈ Dn−1, let Y1 =

Zn + u1κ1 and Y2 = Y3 = Zn and t2 = 0 and note that the conclusions hold. If

exp(t1(Zn + u1κ1))(z) ∈ Dn−1, let z1 = exp(t1(Zn + u1κ1))(z) and u2 > 0. By the

third lemma, we have

exp(t2(Zn + u2κ1))(z1) /∈ Dn−1 for some t2 ∈ (0, T ′).

for all T ′ > 0 for which T ′ ≤ Tmax. By choosing T ′ < 1/3 smaller if necessary, we

may assure that

|exp(t2(Zn + u2κ1))(z1)| > ε.

Thus we let Y1 = Zn + u1κ1, Y2 = Zn + u2κ1, and Y3 = Zn and note that the

conclusions hold. The only other case we must handle is when |z| ≥ ε + 1 and

z ∈ Dn−1, but this follows easily from the above by replacing z1 with z.

We have now guided any initial point z ∈ C into the closed ball Bε(0) in a very

specific manner. This will be extremely important to show

z0 ∈ AF (z0,≤ T ) for all T > 0,

where z0 ∈ Bε(0)∩Dcn−1 belongs to the image of one of the trajectories defined in the

lemma. Before we proceed further, we first note that we have shown:

Corollary 3.38. For all z ∈ C and t > 0:

AF (z, T ′1 ≤ t) ⊃ H(z(ε, κ1), κ1).

Page 99: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

99

Proof. This follows immediately by the previous lemma and by noting for all z0 ∈

Bε(0) and all T > 0, AF (z0,≤ T ) ⊃ H(z(ε, κ1), κ1).

We now hope to show

z0 ∈ AF (z0,≤ T ) for all T > 0,

where z0 ∈ Bε(0)∩Dcn−1 belongs to the image of one of the trajectories defined in the

lemma. To do this, we require a few more propositions.

Proposition 3.39. For all z 6= 0 such that z /∈ Dn−1:

limt→∞

arg(exp(tZn)(z)n−1

)= [π],

where [θ] is the equivalence class of the angle θ under θ ∼ θ′ iff θ = θ+ 2πk for some

k ∈ Z. If we further assume |z| ≥ ε, then the limit is uniform in the initial condition.

Proof. Using the expression 3.32, we obtain:

arg(exp(tZn)(z)n−1

)= [π] + arg

(1

(n− 1)− zn−1

t|z|2n−2

.

)

Take t → ∞ to obtain the result and note that if |z| ≥ ε, we can take the limit

independent of |z| ≥ ε.

Proposition 3.40. For all ε > 0, there exists a time T2(ε) > 0 such that

exp(tZn)(z) + s κ1 : s ∈ R ∩ 0 = ∅

for all t ≥ T2, and all |z| ≥ ε such that z /∈ Dn−1. In particular, T2 > 0 can be chosen

so that the lines

l(z, t) = exp(tZn)(z) + s κ1 : s ∈ R

intersect the lines gj s : s ∈ R for all j ∈ Z away from the origin for all t ≥ T2

and all |z| ≥ ε, z /∈ Dn−1. Here again g is a primitive (n− 1)st root of unity.

Page 100: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

100

Proof. By the previous proposition, for all δ > 0, we may choose t2 = t2(ε) > 0 such

that for all |z| ≥ ε and z /∈ Dn−1

arg(exp(tZn)(z)) ∈(π + 2πk

n− 1− δ, π + 2πk

n− 1+ δ

)for all t ≥ t2

for some k ∈ Z. Consider the set

S(k, δ) =

arg(z) ∈

(π + 2πk

n− 1− δ, π + 2πk

n− 1+ δ

)and suppose there exists a sequence δj ↓ 0 as j →∞ such that

0 ∈ S(k, δj) + s κ1 : s ∈ R

for all j. Write 0 = z(δj) + s(δj)κ1 where z(δj) ∈ S(k, δ) and note that s(δj) 6= 0

and cannot change sign since z(δj) 6= 0 ∈ S(k, δj) for all δj > 0. But note that this

implies arg(κ1) = − arg(z(δj)/s(δj))→ ±(π+ 2πk)/(n− 1) as j →∞. In particular,

κn−11 ∈ R, a contradiction. Therefore there exists δ′ > 0 sufficiently small such that

0 /∈ S(k, δ′) + s κ1 : s ∈ R

for all k ∈ Z. Note this implies the first result after taking T2 = t2(ε) > 0 where t2 is

chosen so that we are within δ′. The second result follows easily with the same choice

of T2 since κ1 is transversal to Dn−1.

Proposition 3.41. Using the notation in the previous proposition, let n ≥ 4 be even.

For all |z| ≥ ε, z /∈ Dn−1, and all t ≥ T2(ε) there exist gj1 6= gj2

l(z, t) ∩ gj1 s : s > 0 6= ∅ and l(z, t) ∩ gj2 s : s < 0 6= ∅.

Proof. Fix |z| ≥ ε such that z /∈ Dn−1 and t ≥ T2(ε). By the previous proposition,

l(z, t) must intersect the lines gj s : s ∈ R for all j ∈ Z away from the origin. In

particular, this means l(z, t) must intersect the sets gj s : s 6= 0 ∈ R for all j ∈ Z.

Since n ≥ 4, this implies the result since l(z, t) is a line.

Page 101: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

101

Proposition 3.42. Suppose that n = 2. For all |z| ≥ ε such that z /∈ Dn−1 and all

t ≥ T2(ε), the line l(z, t) intersects two trajectories of Zn of opposing direction relative

to l(z, t).

Proof. We know that by Proposition 3.40, l(z, t) must intersect the set s : s 6=

0 ∈ R. Suppose first that l(z, t) ∩ s : s < 0 6= ∅ and that the slope of the line is

positive. Since l(z, t) is transversal to the line s : s ∈ R, the associated trajectory

of Zn starting on s < 0 points to the right of the line l(z, t). We thus need to find

a trajectory of Zn that points to the left of this line. If the line l(z, t) passes through

the first quadrant where |y| > |x|, all trajectories of Zn in this quadrant are strictly

increasing in the imaginary direction and strictly decreasing in the real direction,

hence point to the left of the line. If the line passes through the first quadrant

only where |y| < |x|, the trajectories of Zn in the imaginary direction are strictly

increasing, the rate of which increases as x → ∞, hence the trajectories eventually

point to the left of the line. If l(z, t) ∩ s : s < 0 6= ∅ and the slope of the line

is negative, the associated trajectory starting on s < 0 points to the right of the

line. We thus need to find a trajectory of Zn that points to the left of this line. If

the line l(z, t) passes through the fourth quadrant where |y| > |x|, all trajectories

of Zn in this quadrant are strictly decreasing in the imaginary direction and strictly

decreasing in the real direction, hence point to the left of the line l(z, t). If the line

passes through the fourth quadrant only where |y| < |x|, the trajectories of Zn in

the imaginary direction are strictly decreasing, the rate of which strictly decreases

as x → ∞, hence the trajectories eventually point to the left of the line. If l(z, t) is

vertical, the trajectory of Zn on s < 0 points to the right of the line. Moreover,

there exists a trajectory in the second quadrant that points to the left of l(z, t) since

the trajectories are strictly decreasing in the real direction. This handles all cases

when l(z, t) intersects s < 0. The cases when l(z, t) intersects s > 0 are done

similarly.

Page 102: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

102

Lemma 3.43. For all n ≥ 2 even, |z| ≥ ε such that z /∈ Dn−1, t ≥ T2(ε), z1 ∈ l(z, t),

and all T > 0:

z1 ∈ AF (z1,≤ T ).

Proof. Fix n ≥ 2 even, |z| ≥ ε such that z /∈ Dn−1, t ≥ T2, z1 ∈ l(z, t), and T > 0.

We know that for all t > 0

AF (z1,≤ t) ⊃ H(z1, κ1).

By Theorem 2 on p. 68 of [Jur97], we have for all t > 0:

AF (z1,≤ t) ⊃ int(H(z1, κ1)), (3.44)

where int is the interior. By the previous two propositions, the line l(z, t) intersects

(at least) two trajectories of Zn of opposing direction relative to l(z, t). Hence pick

the direction that opposes the direction of κn1 . We may get arbitrarily close to this

trajectory via 3.44 within any amount of positive time t > 0. Hence we may choose

t > 0 small enough such that we flow opposite of κn1 across the line l(z, t) by or before

time S < T . We realize that from this point, the accessibility set in time ε > 0 or

less must contain z1 for all ε > 0. Note that this finishes the proof.

We note that this finishes the proof of Lemma 3.30 in Case 2.

3.7 Explosive Case

We now handle Theorem 3.2. Under these assumptions, our stochastic differential

equation takes the form

dz(t) = z(t)n dt+ κ dW (t), (3.45)

where κn−1 ∈ R \ 0 and W (t) is a one-dimensional standard Wiener process. For a

primitive (n − 1)st root of unity g, there exists j ∈ Z such that gjκ ∈ R. Hence, if

we let w(t) = gjz(t), we obtain:

dw(t) = w(t)n dt+ gjκdW (t). (3.46)

Page 103: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

103

Rephrasing this, the solution z(t) of equation 3.45 starting from z0 ∈ C explodes if

and only if the solution w(t) of equation 3.46 starting from gjz0 explodes. We will

thus argue from the second equation and prove:

Lemma 3.47. For all x > 0,

Pxξw(t) <∞

> 0,

where ξw(t) is the explosion time of the process w(t).

Proof. Note that since x > 0 and gjκ ∈ R, there exists a real-valued solution x(t)

with x(0) = x of the equation:

dx(t) = x(t)n dt+ gjκ dW (t)

which has the same distribution as w(t) with w(0) = x. Thus it suffices to prove

Pxξx(t) <∞

> 0.

We thus apply Feller’s test as in Section 2.2.2. Let α = gjκ and note that

φ(x) =

∫ x

0

exp

(∫ y

0

−2yn

α2dy

)dy

=

∫ x

0

exp

(−2

yn+1

(n+ 1)α2

)dy,

and

m(x) =1

φ′(x)α2

= α−2 exp

(2

xn+1

(n+ 1)α2

).

It is clear that φ(c) ↑ c∞ ∈ (0,∞) as c→∞. Thus we must prove∫ ∞0

dxm(x)(c∞ − φ(x))dx <∞.

Page 104: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

104

It is clear that the integral∫ 1

0is finite. We shall prove then that there exists δ, C > 0

such that

m(x)(c∞ − φ(x)) ≤ C

x1+δfor all x ≥ 1.

Note by L’Hospital’s rule for δ < n+ 1 we have

limx→∞

x1+δm(x)(c∞ − φ(x)) = limx→∞

(c∞ − φ(x))

x−1−δα2 exp(−2 xn+1

(n+1)α2

)= lim

x→∞

1

2xn−1−δ + α2(1 + δ)x−δ−2

= 0,

which finishes the proof.

Page 105: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

105

Chapter 4

Summary

In this work, we found that the maximal class of (κ1, κ2) ∈ C such that the SDE

(1.2) has the ergodic property consists solely of pairs (κ1, κ2) that are transversal

to Dn−1. Outside the realm of such noise, there are solutions of (1.2) which reach

infinity in finite time with positive probability. In the case when n = 2, the problem

was originally motived by applications to turbulent transport of inertial particles

[GHW10]. For n ≥ 3, the problem was driven by a simple geometric intuition that

noise transversal to all isolated unstable trajectories stabilizes the system as a whole.

This intuition is validated here; however, in [Sch93] this is not the case. The difference

between the two is that the dynamics in [Sch93] “cooked up” to disagree with this

intuition. We feel, therefore, there should be a general class of functions b and σ as

in (2.1) for our intuition to hold. Using the results of this dissertation, one has a

natural place to start to determine such a class.

Page 106: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

106

References

[AK87] Ludwig Arnold and Wolfgang Kliemann. On unique ergodicity for degen-erate diffusions. Stochastics, 21(1):41–61, 1987.

[BCH07] Jeremie Bec, Massimo Cencini, and Rafaela Hillerbrand. Heavy particlesin incompressible flows: the large Stokes number asymptotics. Phys. D,226(1):11–22, 2007.

[Bec05] Jeremie Bec. Multifractal concentrations of inertial particles in smoothrandom flows. Journal of Fluid Mechanics, 528:255–277, 2005.

[BHW11] Jeremiah Birrell, David P. Herzog, and Jan Wehr. Transition from ergodicto explosive behavior in a family of stochastic differential equations withdegenerate noise. http://www.math.arizona.edu/∼dherzog, 2011.

[Cho69] Gustave Choquet. Lectures on analysis. Vol. II: Representation theory.Edited by J. Marsden, T. Lance and S. Gelbart. W. A. Benjamin, Inc.,New York-Amsterdam, 1969.

[Dur96] Richard Durrett. Stochastic calculus. Probability and Stochastics Series.CRC Press, Boca Raton, FL, 1996. A practical introduction.

[Dyn65] E. B. Dynkin. Markov processes. Vols. I, II, volume 122 of Translatedwith the authorization and assistance of the author by J. Fabius, V. Green-berg, A. Maitra, G. Majone. Die Grundlehren der Mathematischen Wis-senschaften, Bande 121. Academic Press Inc., Publishers, New York, 1965.

[EM01] Weinan E and Jonathan C. Mattingly. Ergodicity for the Navier-Stokesequation with degenerate random forcing: finite-dimensional approxima-tion. Comm. Pure Appl. Math., 54(11):1386–1402, 2001.

[GHW10] K. Gawedzki, D.P. Herzog, and J. Wehr. Ergodic properties of a modelfor turbulent dispersion of inertial particles. arXiv:1009.0782v1[math-ph],2010.

[Hai08] Martin Hairer. Ergodic theory for stochastic PDEs. 2008.

[Has80] R. Z. Has′minskiı. Stochastic stability of differential equations, volume 7 ofMonographs and Textbooks on Mechanics of Solids and Fluids: Mechanicsand Analysis. Sijthoff & Noordhoff, Alphen aan den Rijn, 1980. Translatedfrom the Russian by D. Louvish.

Page 107: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

107

[HM08] M. Hairer and Jonathan C. Mattingly. Yet another look at harris’ ergodictheorem for markov chains. arXiv: 0810.2777, 2008.

[Hor67] Lars Hormander. Hypoelliptic second order differential equations. ActaMath., 119:147–171, 1967.

[JK81] V. Jurdjevic and I. Kupka. Control systems on semisimple Lie groupsand their homogeneous spaces. Ann. Inst. Fourier (Grenoble), 31(4):vi,151–179, 1981.

[JK85] V. Jurdjevic and I. Kupka. Polynomial control systems. Math. Ann.,272(3):361–368, 1985.

[Jur97] Velimir Jurdjevic. Geometric control theory, volume 52 of Cambridge Stud-ies in Advanced Mathematics. Cambridge University Press, Cambridge,1997.

[Lee03] John M. Lee. Introduction to smooth manifolds, volume 218 of GraduateTexts in Mathematics. Springer-Verlag, New York, 2003.

[McK05] Henry P. McKean. Stochastic integrals. AMS Chelsea Publishing, Provi-dence, RI, 2005. Reprint of the 1969 edition, with errata.

[MSH02] J. C. Mattingly, A. M. Stuart, and D. J. Higham. Ergodicity for SDEsand approximations: locally Lipschitz vector fields and degenerate noise.Stochastic Process. Appl., 101(2):185–232, 2002.

[MT92] Sean P. Meyn and R. L. Tweedie. Stability of Markovian processes. I.Criteria for discrete-time chains. Adv. in Appl. Probab., 24(3):542–574,1992.

[MT93a] Sean P. Meyn and R. L. Tweedie. Stability of Markovian processes. II.Continuous-time processes and sampled chains. Adv. in Appl. Probab.,25(3):487–517, 1993.

[MT93b] Sean P. Meyn and R. L. Tweedie. Stability of Markovian processes. III.Foster-Lyapunov criteria for continuous-time processes. Adv. in Appl.Probab., 25(3):518–548, 1993.

[MT09] Sean Meyn and Richard L. Tweedie. Markov chains and stochastic stability.Cambridge University Press, Cambridge, second edition, 2009. With aprologue by Peter W. Glynn.

[Øks03] Bernt Øksendal. Stochastic differential equations. Universitext. Springer-Verlag, Berlin, sixth edition, 2003. An introduction with applications.

Page 108: Geometry’s Fundamental Role in the Stability of Stochastic ...orion.math.iastate.edu/dherzog/diss.pdfGeometry’s Fundamental Role in the Stability of Stochastic Differential Equations

108

[RB06] Luc Rey-Bellet. Ergodic properties of Markov processes. In Open quantumsystems. II, volume 1881 of Lecture Notes in Math., pages 1–39. Springer,Berlin, 2006.

[Sch93] M. Scheutzow. Stabilization and destabilization by noise in the plane.Stochastic Anal. Appl., 11(1):97–113, 1993.

[SV72] Daniel W. Stroock and S. R. S. Varadhan. On the support of diffusion pro-cesses with applications to the strong maximum principle. In Proceedingsof the Sixth Berkeley Symposium on Mathematical Statistics and Probabil-ity (Univ. California, Berkeley, Calif., 1970/1971), Vol. III: Probabilitytheory, pages 333–359, Berkeley, Calif., 1972. Univ. California Press.

[WM03] M. Wilkinson and B. Mehlig. Path coalescence transition and its applica-tions. Phys. Rev. E, 68(4):040101, Oct 2003.