Hybrid Switching Diffusions Properties and Applications
Hybrid Switching Diffusions Properties and Applications
Random Media
Signal Processing
and Applied Probability
and Image Synthesis (Formerly:
Applications of Mathematics)
Mathematical Economics and Finance
Stochastic Optimization
Stochastic Control
Stochastic Models in Life Sciences 63
Edited by B. Rozovskiı̆
G. Grimmett
Managing Editors
Boris Rozovskiı̆ Geoffrey Grimmett
Division of Applied Mathematics Centre for Mathematical Sciences
Brown University University of Cambridge
182 George St Wilberforce Road
Providence, RI 02912 Cambridge CB3 0WB
USA UK
[email protected] [email protected]
ISSN 0172-4568
ISBN 978-1-4419-1104-9 e-ISBN 978-1-4419-1105-6
DOI 10.1007/978-1-4419-1105-6
Springer New York Dordrecht Heidelberg London
To my parents Yulan Zhong and Changming Zhu and my wife Lijing Sun,
with love
Chao Zhu
Contents
Preface xi
Conventions xv
2 Switching Diffusion 27
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Switching Diffusions . . . . . . . . . . . . . . . . . . . . . . 27
2.3 Regularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.4 Weak Continuity . . . . . . . . . . . . . . . . . . . . . . . . 38
2.5 Feller Property . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.6 Strong Feller Property . . . . . . . . . . . . . . . . . . . . . 52
2.7 Continuous and Smooth Dependence on the Initial Data x . 56
2.8 A Remark Regarding Nonhomogeneous Markov Processes . 65
2.9 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
vii
viii Contents
3 Recurrence 69
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.2 Formulation and Preliminaries . . . . . . . . . . . . . . . . 70
3.2.1 Switching Diffusion . . . . . . . . . . . . . . . . . . . 70
3.2.2 Definitions of Recurrence and Positive Recurrence . 72
3.2.3 Preparatory Results . . . . . . . . . . . . . . . . . . 72
3.3 Recurrence and Transience . . . . . . . . . . . . . . . . . . 78
3.3.1 Recurrence . . . . . . . . . . . . . . . . . . . . . . . 78
3.3.2 Transience . . . . . . . . . . . . . . . . . . . . . . . . 82
3.4 Positive and Null Recurrence . . . . . . . . . . . . . . . . . 85
3.4.1 General Criteria for Positive Recurrence . . . . . . . 85
3.4.2 Path Excursions . . . . . . . . . . . . . . . . . . . . 89
3.4.3 Positive Recurrence under Linearization . . . . . . . 89
3.4.4 Null Recurrence . . . . . . . . . . . . . . . . . . . . 93
3.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
3.6 Proofs of Several Results . . . . . . . . . . . . . . . . . . . . 100
3.7 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4 Ergodicity 111
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
4.2 Ergodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.3 Feedback Controls for Weak Stabilization . . . . . . . . . . 119
4.4 Ramifications . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.5 Asymptotic Distribution . . . . . . . . . . . . . . . . . . . . 129
4.6 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7 Stability 183
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
7.2 Formulation and Auxiliary Results . . . . . . . . . . . . . . 184
7.3 p-Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
7.3.1 Stability . . . . . . . . . . . . . . . . . . . . . . . . . 188
7.3.2 Auxiliary Results . . . . . . . . . . . . . . . . . . . . 193
7.3.3 Necessary and Sufficient Conditions for p-Stability . 201
7.4 Stability and Instability of Linearized Systems . . . . . . . 203
7.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.6 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
A Appendix 355
A.1 Discrete-Time Markov Chains . . . . . . . . . . . . . . . . . 355
A.2 Continuous-Time Markov Chains . . . . . . . . . . . . . . . 358
A.3 Fredholm Alternative and Ramification . . . . . . . . . . . 362
A.4 Martingales, Gaussian Processes, and Diffusions . . . . . . . 366
A.4.1 Martingales . . . . . . . . . . . . . . . . . . . . . . . 366
A.4.2 Gaussian Processes and Diffusion Processes . . . . . 369
A.5 Weak Convergence . . . . . . . . . . . . . . . . . . . . . . . 371
A.6 Hybrid Jump Diffusion . . . . . . . . . . . . . . . . . . . . . 376
A.7 Miscellany . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
A.8 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
References 379
Index 392
Preface
xi
xii Preface
regularity, the Feller property, the strong Feller property, and continuous
and smooth dependence on initial data of the solutions of the associated
stochastic differential equations with switching.
A large part of this work is concerned with the stability of switching
diffusion processes. Here stability is meant in the broad sense including
both weak and strong stability. That is, we focus on both a “neighbor-
hoods of infinity” and neighborhoods of equilibrium points; the stability
corresponding to the former is referred to as weak stability, whereas that
of the latter is stability in the usual sense. In studying deterministic dy-
namic systems, researchers used Lagrange stability to depict systems that
are ultimately uniformly bounded. Treating stochastic systems, one would
still hope to adopt such a notion. Unfortunately, the boundedness excludes
many important cases. Thus, we replace this boundedness by a weaker
notion known as recurrence (returning to a prescribed compact region in
finite time). When the expected returning time is finite, we have so-called
positive recurrence. A crucial question is: Under what conditions will the
systems be recurrent (resp. positive recurrent). We demonstrate that pos-
itive recurrence implies ergodicity and provide criteria for the existence
of invariant distributions together with their representation. In addition
to studying asymptotic properties of the systems in the neighborhood of
infinity, we examine the behavior of the systems at the equilibria. Also con-
sidered are invariance principles and stability of differential equations with
random switching but without diffusions.
Because the systems are rarely solvable in closed form, numerical meth-
ods become a viable alternative. We construct algorithms to approximate
solutions of such systems with state-dependent switching, and provide suf-
ficient conditions for convergence for numerical approximations to the in-
variant measures.
In real-world applications, hybrid systems encountered are often of large
scale and complex leading to intensive-computation requirement. Reduc-
tion of computational complexity is thus an important issue. To take this
into consideration, we consider time-scale separation in hybrid switching-
diffusion and hybrid-jump-diffusion models. Several issues including recur-
rence of processes with switching having multiple-weak-connected ergodic
classes, two-time-scale modeling of stochastic volatility, and weak conver-
gence analysis of systems with fast and slow motions involving additional
Poisson jumps are studied in details.
This book is written for applied mathematicians, probabilists, systems
engineers, control scientists, operations researchers, and financial analysts
among others. The results presented in the book are useful to researchers
working in stochastic modeling, systems theory, and applications in which
continuous dynamics and discrete events are intertwined. The book can be
served as a reference for researchers and practitioners in the aforementioned
areas. Selected materials from the book may also be used in a graduate-level
course on stochastic processes and applications.
Preface xiii
This book project could not have been completed without the help and
encouragement of many people. We are deeply indebted to Wendell Flem-
ing and Harold Kushner, who introduced the wonderful world of stochastic
systems to us. We have been privileged to have the opportunity to work
with Rafail Khasminskii on a number of research projects, from whom we
have learned a great deal about Markov processes, diffusion processes, and
stochastic stability. We express our special appreciation to Vikram Krish-
namurthy, Ruihua Liu, Yuanjin Liu, Xuerong Mao, John Moore, Qingshuo
Song, Chenggui Yuan, Hanqin Zhang, Qing Zhang, and Xunyu Zhou, who
have worked with us on various projects related to switching diffusions. We
also thank Eric Key, Jose Luis Menaldi, Richard Stockbridge, and Fubao Xi
for many useful discussions. This book, its contents, its presentation, and
exposition have benefited a great deal from the comments by Ruihua Liu,
Chenggui Yuan, Hanqin Zhang, and by several anonymous reviewers, who
read early versions of the drafts and offered many insightful comments. We
thank the series editor, Boris Rozovsky, for his encouragement and con-
sideration. Our thanks also go to Springer senior editor, Achi Dosanjh, for
her assistance and help, and to the production manager, and the Springer
professionals for their work in finalizing the book. During the years of our
study, the research was supported in part by the National Science Founda-
tion, and the National Security Agency. Their continuing support is greatly
appreciated.
xv
Glossary of Symbols
xvii
xviii Glossary of Symbols
1.1 Introduction
This book focuses on switching diffusion processes involving both con-
tinuous dynamics and discrete events. Before proceeding to the detailed
study, we address the following questions. Why should we study such hy-
brid systems? What are typical examples arising from applications? What
are the main properties we wish to study? This introductory chapter pro-
vides motivations of our study, delineates switching diffusions in a simple
way, presents a number of application examples, and gives an outline of the
entire book.
1.2 Motivation
Owing to their wide range of applications, hybrid switching diffusions, also
known as switching diffusion systems, have become more popular recently
and have drawn growing attention, especially in the fields of control and
optimization. Because of the presence of both continuous dynamics and
discrete events, such systems are capable of describing complex systems
and their inherent uncertainty and randomness in the environment. The
formulation provides more opportunity for realistic models, but adds more
difficulties in analyzing the underlying systems. Resurgent efforts have been
devoted to learning more about the processes and their properties. Much
of the study originated from applications arising in control engineering,
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 1
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_1,
© Springer Science + Business Media, LLC 2010
2 1. Introduction and Motivation
tinuous dynamics and discrete events are intertwined. For example, one
of the early efforts of using such hybrid models for financial applications
can be traced back to [5], in which both the appreciation rate and the
volatility rate of a stock depend on a continuous-time Markov chain. In the
simplest case, a stock market may be considered to have two “modes” or
“regimes,” up and down, resulting from the state of the underlying econ-
omy, the general mood of investors in the market, and so on. The rationale
is that in the different modes or regimes, the volatility and return rates
are very different. The introduction of hybrid models makes it possible to
describe stochastic volatility in a relatively simple manner (simpler than
the so-called stochastic volatility models). Another example is concerned
with a wireless communication network. Consider the performance analysis
of an adaptive linear multiuser detector in a cellular direct-sequence code-
division multiple-access wireless network with changing user activity due
to an admission or access controller at the base station. Under certain con-
ditions, an associated optimization problem leads to a switching diffusion
limit; see [168]. To summarize, a centerpiece in the applications mentioned
above is a two-component Markov process: a continuous component and a
discrete-event component.
Long-time behavior of such hybrid systems is one of the major concerns
in stochastic processes, systems theory, control, and optimization. Nowa-
days, there is a fairly well-known theory for stability of diffusion processes.
The basic setup of the stability study and the foundational work were
originated in the work of Khasminskii [83] and Kushner [97]. The origin
of the study may be traced back to Kac and Krasovskii [79] for differen-
tial equations perturbed by Markov chains. Comprehensive treatment of
Markov processes is in Dynkin [38], whereas that of diffusion processes
may be found in Gihman and Skorohod [55], Liptser and Shiryayev [110],
and Stroock and Varadhan [153] and references therein. Concerning jump
processes and Markov chains, we mention the work of Chung [28], Cox
and Miller [29], and Doob [33]. Davis’s piecewise deterministic viewpoint
adds a new twist to Markov models. Ethier and Kurtz [43] examine weak
convergence of Markov processes and provide an in-depth study on char-
acterization of the limit process. Chen’s book [23] presents an approach
from the angle of coupling methods. Recently, stability of diffusion pro-
cesses with Markovian switching have received much attention (see, e.g.,
[6, 7, 92, 116, 119, 183, 187, 190] and the references therein). Some of
the recent developments in Markov modulated switching diffusions (where
the Markov chain and the Brownian motion are independent) are found
in Mao and Yuan [120]. Accommodating the many applications, although
these systems are more realistically addressing the demands, the nontradi-
tional setup makes the analysis of switching diffusions more difficult. To be
able to effectively treat these systems, it is of foremost importance to have
a thorough understanding of the underpinning of the systems. This is the
main theme of the current work.
4 1. Introduction and Motivation
Suppose that initially, the process is at (X(0), α(0)) = (x, 1). The discrete
event process sojourns in discrete state 1 for a random duration; during this
period, the continuous component evolves according to the diffusion process
specified by the drift and diffusion coefficients associated with discrete state
1 until a jump of the discrete component takes place. At random moment τ 1 ,
a jump to discrete state 3 occurs. Then the continuous component evolves
according to the diffusion process whose drift and diffusion coefficients are
determined by discrete event 3. The process wanders around in the third
plan until another random jump time τ2 . At τ2 , the system switches to the
second parallel plan and follows another diffusion with different drift and
diffusion coefficients and so on.
Discrete-event State 3
6X(τ1 )
? X(τ4 )
X(τ2 ) 6
Discrete-event State 2
? Discrete-event State 1
X(τ3 ) X(0) = x, α(0) = 1
At a first glance, one may feel that the process is not much different from
a diffusion because the switching is a finite-state process. Nevertheless,
even if the switching is a finite-state Markov chain independent of the
Brownian motion (subsequently referred to as Markov-modulated switching
1.4 Examples of Switching Diffusions 5
diffusions), the switching diffusion is still much harder to handle. The main
reason is the coupling and interactions due to the switching process make
the analysis much more difficult. For example, when we study recurrence
or stability, we have to deal with systems of coupled partial differential
equations.
One of the main features of this book is that a large part of it deals with
the switching component depending on the continuous component. This is
often referred to as a state-dependent switching process in what follows;
more precise statements about which will be made later. In this case, the
analysis is much more difficult than that of the case in which the switching
is independent of the continuous states. As shown in Chapter 2, some of the
time-honored characteristics such as continuous dependence of initial data
become highly nontrivial; even Feller properties are not straightforward to
obtain.
for nearly three decades. The original problem is concerned with linear
deterministic systems. In the 1980s, to bridge the gap between minimal
realization theory and to treat the problem of finding lower-order approxi-
mations, a particular realization was first introduced in the seminal paper
by Moore [129], where the problem was studied with principal component
analysis of linear systems. The term “balanced” was used because the real-
izations have a certain symmetry between the input and the output maps
characterized by the controllability and observability Grammians. Owing
to their importance and their wide range of applications, balanced realiza-
tions have attracted much attention. One of the applications areas is model
reduction. Asymptotic stability of the reduced-order systems was studied
in [129], and error bounds between the reduced-order model and the origi-
nal system were obtained in [58] in terms of the associated singular values.
There have been substantial extensions of the theory to time-varying linear
systems. Key existence results concerning balanced realizations were con-
tained in [148] and [157]. Subsequent work can be found in [66, 75, 133, 144]
and references therein.
We generalize the ideas to include random switching to represent random
perturbations under a stochastic environment. We suppose that for each in-
stant t, the state consists of a pair (X(t), α(t)) representing the continuous
state component and discrete event component, respectively. Let α(t) be a
continuous-time Markov chain with finite state space M = {1, 2, . . . , m0 }
and generator
m0
X
Q = (qij ) ∈ Rm0 ×m0 such that qij ≥ 0 for i 6= j, and qij = 0. (1.4)
j=1
each i ∈ M, A(i, ·), B(i, ·), and C(i, ·) are bounded and continuously dif-
ferentiable matrix-valued functions with suitable dimensions. Consider the
following system
d
X(t) = A(α(t), t)X(t) + B(α(t), t)u(t), X(t0 ) = x0 ,
dt (1.5)
y(t) = C(α(t), t)X(t),
where the state X(t) ∈ Rn×1 , input u(t) ∈ Rr×1 , and output y(t) ∈ Rm0 .
Note that because a finite-state Markov chain is used, effectively, (1.5) can
be written as
m0
X m0
X
d
X(t) = A(i, t)X(t)I{α(t)=i} + B(i, t)u(t)I{α(t)=i} , X(t0 ) = x0 ,
dt i=1 i=1
m0
X
y(t) = C(i, t)X(t)I{α(t)=i} ,
i=1
For each t ≥ 0 and i ∈ M, a realization (A(i, t), B(i, t), C(i, t)) is said to
be uniformly completely controllable if and only if there is a δ > 0 such
that for some positive Lc (δ) and Uc (δ),
and Φ(t, λ, i) is the state transition matrix (see [2, p. 349]) of the equation
dz(t)
= A(i, t)z(t).
dt
For each t ≥ 0 and i ∈ M, a realization (A(i, t), B(i, t), C(i, t)) is said
to be uniformly completely observable if and only if there is a δ > 0 such
that for some positive Lo (δ) and Uo (δ),
such that
Note that regime switching is one of the main features (fully degenerate
regime-switching diffusion with the second order term missing in the sys-
tem) of such systems compared with the traditional setup. To study the
behavior of such systems, it is crucial to have a thorough understanding of
the switching diffusion processes. For modeling and analysis of the balanced
realizations, we refer the reader to [111] for further reading.
Example 1.3. Continuing with the setup of the hybrid linear system
P ε = I + εQ, (1.13)
bn − E{π(θn )}
π
vn = √ , n ≥ n0 , v µ (t) = vn for t ∈ [nµ, nµ + µ). (1.18)
µ
We have shown in [168] that (v µ (·), θ µ (·)) converges weakly to (v(·), θ(·))
satisfying the switching diffusion equation
where w(·) is a standard Brownian motion and Σ(θ) is the almost sure limit
given by
n+m−1 n+m−1
1 X X
lim (Xk+1 (θ) − EXk+1 (θ))(Xk1 +1 (θ) − EXk1 +1 (θ))0
µ→0 n
k1 =m k=m
= Σ(θ) a.s.
(1.20)
Note that for each θ, Σ(θ) is an S × S deterministic matrix and in fact,
n+m−1 n+m−1
1 X X
E {(Xk+1 (θ) − EXk+1 (θ))(Xk1 +1 (θ) − EXk1 +1 (θ))0 }
n
k1 =m k=m
= Σ(θ) as µ → 0.
(1.21)
Because of the regime switching, the system is qualitatively different from
the existing literature on stochastic approximation methods; see Kushner
and Yin [104].
is satisfied for some δ > 0, and that all the functions r(t, i), bm (t, i),
σmn (t, i) are measurable and uniformly bounded in t. Inequality (1.25)
is in the sense of positive definiteness for symmetric matrices; that is,
σ(t, i)σ 0 (t, i) − δI is a positive definite matrix.
Suppose that the initial market mode α(0) = i0 . Consider an agent with
an initial wealth x0 > 0. Denote by x(t) the total wealth of the agent at
time t ≥ 0. Assuming that the trading of shares takes place continuously
and that there are no transaction cost and consumptions, then one has
(see, e.g., [181, p.57])
n X d
o
dx(t) = r(t, α(t))x(t) + b m (t, α(t)) − r(t, α(t)) u m (t) dt
m=1
X d X d
+ σmn (t, α(t))um (t)dwn (t),
n=1 m=1
x(0) = x > 0, α(0) = i ,
0 0
(1.26)
where um (t) is the total market value of the agent’s wealth in the mth asset,
m = 0, 1, . . . , d, at time t. We call u(·) = (u1 (·), . . . , ud (·))0 a portfolio of
the agent. Note that once u(·) is determined, u0 (·), the asset in the bank
14 1. Introduction and Motivation
We assume that w(·), N (·), and α(·) are mutually independent. Note that
compared with the traditional jump-diffusion processes, the coefficients in-
volved in (1.30) all depend on an additional switching process–the Markov
chain α(t).
In the context of risk theory, X(t) can be considered as the surplus of
the insurance company at time t, x0 is the initial surplus, f (t, X(t), α(t))
represents the premium rate (assumed to be ≥ 0), g(γ, X(t), α(t)) is the
amount of the claim if there is one (assumed to be ≤ 0), and the diffu-
sion is used to model additional uncertainty of the claims and/or premium
incomes. Similar to the volatility in stock market models, σ(·, ·, i) repre-
sents the amount of oscillations or volatility in an appropriate sense. The
model is sufficiently general to cover the traditional as well as the diffusion-
perturbed ruin models. It may also be used to represent security price in
16 1. Introduction and Motivation
finance (see [124, Chapter 3]). The process α(t) may be viewed as an envi-
ronment variable dictating the regime. The use of the Markov chain results
from consideration of a general trend of the market environment as well as
other economic factors. The economic and/or political environment changes
lead to the changes of surplus regimes resulting in markedly different be-
havior of the system across regimes.
Defining a centered Poisson measure and applying generalized Ito’s rule,
we can obtain the generator of the jump-diffusion process with regime
switching, and formulate a related martingale problem. Instead of a sin-
gle process, we have to deal with a collection of jump-diffusion processes
that are modulated by a continuous-time Markov chain. Suppose that λ
is positive such that λ∆ + o(∆) represents the probability of a switch of
regime in the interval [t, t + ∆), and π(·) is the distribution of the jump.
Then the generator of the underlying process can be written as
∂
GF (t, x, ι) = + L F (t, x, ι)
∂t
Z
+ λ[F (t, x + g(γ, x, ι), ι) − F (t, x, ι)]π(dγ) (1.31)
Γ
+Q(t)F (t, x, ·)(ι), for each ι ∈ M,
where
1 2 ∂2 ∂
LF (t, x, ι) = σ (t, x, ι) 2 F (t, x, ι) + f (t, x, ι) F (t, x, ι),
2 m0
∂x ∂x
X X
Q(t)F (t, x, ·)(ι) = qι` (t)F (t, x, `) = qι` (t)[F (t, x, `) − F (t, x, ι)].
`=1 `6=ι
(1.32)
Of particular interest is the case that the switching process and the diffu-
sion vary at different rates. To reduce the amount of computational effort,
we propose a two-time-scale approach. By concentrating on time-scale sep-
arations, we treat two cases in Chapter 12. In the first one, the regime
switching is significantly faster than the dynamics of the jump diffusions,
whereas in the second case, the diffusion vary an order of magnitude faster
than the other processes. As shown, averaging plays an essential role in
these problems.
Example 1.7. Consider the process Y (·) = (X(·), α(·)) that has two com-
ponents, the diffusion component X(·) and the pure jump component α(·).
The state space of the process is S × M, where S is the unit circle and
M = {1, . . . , m0 } is a state space with finitely many elements for the jump
process. By identifying the endpoints 0 and 1, let x ∈ [0, 1] be the co-
ordinates in S. We assume that the generator of the process (X(t), α(t))
is of the form (1.31) with the jump part missing (i.e., λ = 0) and with
(x, i) ∈ [0, 1] × M.
1.4 Examples of Switching Diffusions 17
for i = 1, . . . , m0 , and g(x) = (g1 (x), . . . , gm0 (x)) is the initial distribution
for Y (t). The existence and properties of such switching diffusion processes
can be found in [55, Section 2.2]. Suppose that all conditions of [46, The-
orem 16, p. 82] are satisfied. Then the system of forward equations has a
unique solution.
Suppose that qij (x) > 0 for each i 6= j. Then the transition density
p(x, t) converges exponentially fast to
for some K > 0 and γ > 0. The estimate is the well-known spectrum gap
condition. This spectrum gap condition helps us to study related asymp-
totic properties of the system.
Example 1.8. This example is concerned with a switching diffusion pro-
cess with a state-dependent switching component. We compare the trajec-
tories of a linear stochastic system with drift and diffusion coefficients given
by
f (x) = 0.11x and σ(x) = 0.2x (1.33)
and another regime-switching linear system with α(t) ∈ {1, 2}, and the
drift and diffusion coefficients given by
What happens if the actual model is the one with switching, but we mis-
modeled the system so that instead of the switching model, we used a
simple linear stochastic differential equation model with drift and diffusion
coefficients given by (1.33)? If in fact, the system parameters are really
given by (1.34), and if the random environment is included in the model,
then we will see a significant departure of the model from that of the true
system. This can be seen from the plots in Figure 1.2.
To get further insight from these plots, imagine that we are encountering
the following scenarios. Suppose that the above example is for modeling the
stock price of a particular equity. Suppose that we modeled the stock price
by using the simple geometric Brownian motion model with return rate
11% and the volatility 0.2 as given in (1.33). However, the market is really
subject to random environment perturbation. For the bull market, the rates
are as before, but corresponding to the bear market, the return rate and
the volatility are given by −20% and 1, respectively. The deviation of the
plot in Figure 1.2 shows that if one uses a simple linear SDE model, one
cannot capture the market behavior.
Example 1.9. This example is concerned with a controlled switching dif-
fusion model. We first give a motivation of the study and begin the discus-
sion on Markov decision processes. Consider a real-valued process X(·) =
{X(t) : t ≥ 0} and a feedback control u(·) = {u(t) = u(X(t)) : t ≥ 0} such
that u(t) ∈ Γ for t ≥ 0 with Γ being a compact subset of an Euclidean
space. Γ denotes the control space. A vast literature is concerned with a
certain optimal control problem. That is, one aims to find a feedback u(·)
so that an appropriate objective function is minimized.
Extending the idea to switching diffusion processes, we consider the fol-
lowing pair of processes (X(t), α(t)) ∈ Rr × M, where M = {1, . . . , m0 }.
Let U be the control space or action space, which is a compact set of
an Euclidean space. Suppose that b(·, ·, ·) : Rr × M × U 7→ Rr , and
σ(·, ·, ·) : Rr × M × U 7→ Rr×r are appropriate functions satisfying cer-
Q(x) = (qij (x)) ∈ Rm0 ×m0 satisfies that for
tain regularity conditions, and P
m0
each x, qij (x) ≥ 0 for i 6= j, j=1 qij (x) = 0 for each i ∈ M. For each
i ∈ M and suitable smooth function h(·, i), define an operator
1
Lh(x, i) = ∇h0 (x, i)b(x, i, u) + tr[∇2 h(x, i)σ(x, i, u)σ 0 (x, i, u)]
2
m0
X (1.35)
+ qij (x)h(x, j).
j=1
1.4 Examples of Switching Diffusions 19
11
10
8
x(t)
4
0 50 100 150 200 250 300 350
t
6
x(t)
1
0 50 100 150 200 250 300 350
t
FIGURE 1.2. Comparisons of sample paths for a linear diffusion with that of a
regime-switching diffusion.
20 1. Introduction and Motivation
dXi (t) = γ(α(t))Xi (t) − Xi3 (t) − β(α(t))(Xi (t) − X(t)) dt
(1.36)
+ σii (X(t), α(t))dwi (t),
1.5 Outline of the Book 21
where
1 X̀
X(t) = Xj (t),
` j=1 (1.37)
X(t) = (X1 (t), X2 (t), . . . , X` (t))0 ,
and γ(i) > 0 and β(i) > 0 for i ∈ M. Moreover, the transition rules of α(t)
are specified by
As shown in the examples above, all the systems considered involve hy-
brid switching diffusions. To have a better understanding of each of the
problems and the properties of the corresponding processes, it is important
that we have a thorough understanding of the switching diffusion process.
This is our objective in this book.
2.1 Introduction
This chapter provides an introduction to switching diffusions. First the
definition of switching diffusion is given. Then with a short review of the
existence and uniqueness of the solution of associated stochastic differential
equations, weak continuity, Feller, and strong Feller properties are estab-
lished. Also given here are the definition of regularity and criteria ensuring
such regularity. Moreover, smooth dependence on initial data is presented.
The rest of the chapter is arranged as follows. After this short introduc-
tory section, Section 2.2 presents the general setup for switching processes.
Section 2.3 is concerned with regularity. Section 2.4 deals with weak con-
tinuity of the pair of process (X(t), α(t)). Section 2.5 proceeds with Feller
properties. Section 2.6 goes one step further to obtain strong Feller prop-
erties. Section 2.7 presents smooth dependence properties of solutions of
the switching diffusions. Section 2.8 gives remarks on how nonhomogeneous
cases in which both the drift and diffusion coefficients depend explicitly on
time t can be handled. Finally, Section 2.9 provides additional notes and
remarks.
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 27
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_2,
© Springer Science + Business Media, LLC 2010
28 2. Switching Diffusion
and for i 6= j,
Lf (x, ι) = ∇f 0 (x, ι)b(x, ι) + tr(∇2 f (x, ι)A(x, ι)) + Q(x)f (x, ·)(ι)
Xr r
∂f (x, ι) 1 X ∂ 2 f (x, ι)
= bi (x, ι) + aij (x, ι) (2.4)
i=1
∂xi 2 i,j=1 ∂xi ∂xj
+Q(x)f (x, ·)(ι),
where ∇f (x, ι) and ∇2 f (x, ι) denote the gradient and Hessian of f (x, ι)
with respect to x, respectively,
m0
X
Q(x)f (x, ·)(ι) = qιj f (x, j), and
j=1
Note that the evolution of the discrete component α(·) can be represented
by a stochastic integral with respect to a Poisson random measure (see,
e.g., [52, 150]). Indeed, for x ∈ Rr and i, j ∈ M with j 6= i, let ∆ij (x) be
the consecutive (with respect to the lexicographic ordering on M × M),
left-closed, right-open intervals of the real line, each having length qij (x).
Define a function h : Rr × M × R 7→ R by
m0
X
h(x, i, z) = (j − i)I{z∈∆ij (x)} . (2.5)
j=1
That is, with the partition {∆ij (x) : i, j ∈ M} used and for each i ∈ M,
if z ∈ ∆ij (x), h(x, i, z) = j − i; otherwise h(x, i, z) = 0. Then (2.3) is
equivalent to Z
dα(t) = h(X(t), α(t−), z)p(dt, dz), (2.6)
R
where p(dt, dz) is a Poisson random measure with intensity dt×m(dz), and
m is the Lebesgue measure on R. The Poisson random measure p(·, ·) is
independent of the Brownian motion w(·).
Similar to the case of diffusions, with the L defined in (2.4), for each
f (·, ı) ∈ C 2 , ı ∈ M, a result known as the generalized Itô lemma (see
[17, 120] or [150]) reads
Z t
f (X(t), α(t)) − f (X(0), α(0)) = Lf (X(s), α(s))ds + M1 (t) + M2 (t),
0
(2.7)
where
Z t
M1 (t) = ∇f (X(s), α(s)), σ(X(s), α(s))dw(s) ,
Z0 t Z
M2 (t) = f (X(s), α(0) + h(X(s), α(s), z))
0 R
−f (X(s), α(s)) µ(ds, dz),
and
µ(ds, dz) = p(ds, dz) − ds × m(dz)
is a martingale measure.
In view of the generalized Itô formula,
Z t
Mf (t) = f (X(t), α(t)) − f (X(0), α(0)) − Lf (X(s), α(s))ds (2.8)
0
see [120] or [150] for details. In what follows, for convenience (with multi-
index notation used, for instance) and emphasis on the x-dependence, we
often use Dx f (x, α) and ∇f (x, α) interchangeably to represent the gradi-
ent. In addition, for two vectors x and y with appropriate dimensions, we
use x0 y and x, y interchangeably to represent their inner product. To pro-
ceed, we present the existence and uniqueness of solutions to a system of
stochastic differential equations associated with switching diffusions first.
In what follows, for Q(x) : Rr → Rm0 ×m0 , we say it satisfies the q-
property, if Q(x) = (qij (x)), for all x ∈ Rr , qij (x) is Borel measurable for
all i, j ∈ M andPx ∈ Rr ; qij (x) is uniformly bounded; qij (x) ≥ 0 for j 6= i
and qii (x) = − j6=i qij (x) for all x ∈ Rr .
Note that an alternative definition of the q-property can be devised. The
boundedness assumption can be relaxed. However, for our purpose, the
current setup seems to be sufficient. The interested reader could find the
desired information through [28].
and that for every integer N ≥ 1, there exists a positive constant M N such
that for all t ∈ [0, T ], i ∈ M and all x, y ∈ Rr with |x| ∨ |y| ≤ MN ,
Then there exists a unique solution (X(t), α(t)) to the equation (2.2) with
given initial data in which the evolution of the jump process is specified by
(2.3).
Remark 2.2. For brevity, the detailed proof is omitted. Instead, we make
the following remarks. There are a number of possible proofs. For example,
2.2 Switching Diffusions 31
the existence can be obtained as in [150, pp. 103-104]. Viewing the switch-
ing diffusion as a special case of a jump-diffusion process (see the stochastic
integral representation of α(t) in (2.6)), one may prove the existence and
uniqueness using [77, Section III.2]. Another possibility is to use a mar-
tingale problem formulation together with utilization of truncations and
stopping times as in [72, Chapter IV]. In Chapter 5, we present numerical
approximation algorithms for solutions of switching diffusions, and show
the approximation algorithms converge weakly to the switching diffusion
of interest by means of a martingale problem formulation. Then using Lip-
schitz continuity and the weak convergence, we further obtain the strong
convergence of the approximations. As a byproduct, we can obtain the ex-
istence and uniqueness of the solution. The verbatim proof can be found in
[172]; see Remark 5.10 for further explanations. While most proofs of the
uniqueness take two different solutions with the same initial data and show
their difference should be 0 by using Lipschitz continuity and Gronwall’s
inequality, it is possible to consider the difference of the two solutions with
different initial data whose difference is arbitrarily small. In this regard,
the uniqueness can be derived from Proposition 2.30. Earlier work using
such an approach may be found in [130].
Note that (2.10) is the linear growth condition and (2.11) is the local Lip-
schitz condition. These conditions for the usual diffusion processes (without
switching) are used extensively in the literature. Theorem 2.1 provides ex-
istence and uniqueness of solutions of (2.2) with (2.3) specified. To proceed,
we obtain a moment estimate on X(t). This estimate is used frequently in
subsequent development. As its diffusion counter part, the main ingredient
of the proof is the use of Gronwall’s inequality.
Using the Hölder inequality and the linear growth condition, detailed com-
32 2. Switching Diffusion
putations lead to
Z t p
Ex,i sup b(X(s), α(s))ds
0≤t≤T1 0
Z T1
p−1 p
≤T |b(X(s), α(s))| ds
0Z
T1
p−1 p
≤ KT (1 + |X(s)| )ds
Z0 T1 h i
≤ c1 (T, p) Ex,i 1 + sup |X(u)|p ds,
0 0≤u≤s
γ γ
Step 2. Note that sup0≤t≤T |X(t)| = sup0≤t≤T |X(s)| for any γ > 0.
Thus we obtain from Hölder’s inequality and Step 1 that for any 1 ≤ p < 2,
1/2
p 2p
Ex,i sup |X(t)| ≤ Ex,i sup |X(t)| ≤ C(T, x, p) < ∞.
0≤t≤T 0≤t≤T
Proof. It is well known that the sample paths of the discrete component
α(·) are right continuous with left limits. Hence it remains to show that
the same is true for the continuous component X(·). To this end, let 0 ≤
s < t ≤ T with T being any fixed positive number, and consider
Z t Z t
X(t) − X(s) = b(X(u), α(u))du + σ(X(u), α(u))dw(u).
s s
Using Lemma 2.3 and [47, Theorem 4.6.3], detailed computations lead to
4 2
Ex,i |X(t) − X(s)| ≤ C |t − s| ,
2.3 Regularity
It follows from Theorem 2.1 that if the coefficients satisfy the linear growth
and the local Lipschitz condition, then the solution (X(t), α(t)) of (2.2)–
(2.3) is defined for all t > 0. Nevertheless, the linear growth condition puts
34 2. Switching Diffusion
It is obvious to see that the process (X x,α (t), αx,α (t)) is regular if and
only if for any 0 < T < ∞,
P{ sup |X x,α (t)| = ∞} = 0. (2.16)
0≤t≤T
That is, a process is regular if and only if it does not blow up in finite
time. Thus, (2.16) can be used as an alternate definition. Nevertheless, it
is handy to use equation (2.15) to delineate the regularity.
In what follows, we take up the regularity issue of a regime-switching
diffusion when its coefficients do not satisfy the linear growth condition.
The following theorem gives sufficient conditions for regularity, which is
based on “local” linear growth and Lipschitz continuity.
Theorem 2.7. Suppose that for each i ∈ M, both the drift b(·, i) and the
diffusion coefficient σ(·, i) satisfy the linear growth and Lipschitz condition
in every bounded open set in Rr , and that there is a nonnegative function
V (·, ·) : Rr × M 7→ R+ that is twice continuously differentiable with respect
to x ∈ Rr for each i ∈ M such that there is an γ0 > 0 satisfying
Hence we have
Theorem 2.8. Suppose that for each i ∈ M, both the drift b(·, i) and
the diffusion coefficient σ(·, i) satisfy the linear growth and Lipschitz con-
dition in every bounded open set in Rr , and that there is a nonnegative and
bounded function V (·, ·) : Rr × M 7→ R+ that is not identically zero, that
for each i ∈ M, V (x, i) is twice continuously differentiable with respect to
x ∈ Rr such that there is a γ1 > 0 satisfying
Then the process (X(t), α(t)) is not regular. In particular, for any ε > 0,
we have
Px0 ,` {β∞ < κ + ε} > 0,
where β∞ := limn→∞ βn , (x0 , `) ∈ Rr × M satisfying V (x0 , `) > 0, and
sup V (x, i)
1 (x,i)∈Rr ×M
κ= log .
γ1 V (x0 , `)
Remark 2.9. If the coefficients of (2.2)–(2.3) satisfy both the local Lip-
schitz condition (2.11) and linear growth condition (2.10), then by The-
orem 2.1, there is a unique solution (X x,α (t), αx,α (t)) to (2.2)–(2.3) for
all t ≥ 0. The regime-switching diffusion (X x,α (t), αx,α (t)) is thus regu-
lar. Alternatively, we can use Theorem 2.7 to verify this. In fact, detailed
computations show that the function
2
V (x, i) = (|x| + 1)r/2 , (x, i) ∈ Rr × M
satisfies all conditions in Theorem 2.7. Thus the desired assertion follows.
In particular, we can verify that V (·, i) is twice continuously differentiable
for each i ∈ M and that LV (x, i) ≤ γ0 V (x, i) for some positive constant
γ0 and all (x, i) ∈ Rr × M. Thus for any t > 0, a similar argument as in
the proof of Theorem 2.7 leads to
V (x, i)eγ0 t
Px,i {βn < t} ≤ ,
Vn
where Vn = (1 + n2 )r/2 . Therefore for any t > 0 and ε > 0, there exists an
N ∈ N such that
Px,i {βn < t} < ε, for all n ≥ N, (2.19)
uniformly for x in any compact set F ⊂ Rr and i ∈ M.
Example 2.10. Consider a real-valued regime-switching diffusion
dX(t) = X(t) (b(α(t)) − a(α(t))X 2 (t))dt + σ(α(t))dw(t) , (2.20)
where w(·) is a one-dimensional standard Brownian motion, and α(·) ∈
M = {1, 2, . . . , m0 } is a jump process with appropriate generator Q(x).
Clearly, the coefficients of (2.20) do not satisfy the linear growth condition
if not all a(i) = 0 for i ∈ M. We claim that if a(i) > 0 for each i ∈ M,
then (2.20) is regular. To see this, we apply Theorem 2.7 and consider
V (x, i) := |x|2 , (x, i) ∈ R × M.
Clearly, V is nonnegative and satisfies
lim V (x, i) = ∞, for each i ∈ M.
|x|→∞
2.3 Regularity 37
and
b(1) = 3, a(1) = 2, σ(1) = 1
b(2) = 10, a(1) = 1, σ(1) = −1.
Figure 2.1 plots a sample path of (2.20) with initial condition (x, α) =
(2.5, 1).
there is a unique solution x(t) to (1.1) on t ≥ 0, and the solution will remain
in Rn+ almost surely; that is, x(t) ∈ Rn+ for any t ≥ 0 with probability 1.
38 2. Switching Diffusion
4.5
X(t)
4 α(t)
3.5
3
X(t) and α(t)
2.5
1.5
0.5
0
0 2 4 6 8 10
t
FIGURE 2.1. A sample path of switching diffusion (2.20) with initial condition
(x, α) = (2.5, 1).
Theorem 2.12. Suppose that the conditions of Theorem 2.1 are satisfied
with the modification Q(x) = Q that generates a Markov chain independent
2.4 Weak Continuity 39
of the Brownian motion. Then the process Y (t) = (X(t), α(t)) is continuous
in probability and also continuous in mean square.
We proceed first to establish the mean square convergence above. Note that
where
and (1, . . . , m0 )0 ∈ Rm0 is a column vector. Because the Markov chain α(t)
is independent of the Brownian motion w(·) (Q is a constant matrix), it is
well known that
Z t+∆
χ(t + ∆) − χ(t) − χ(s)Qds
t
40 2. Switching Diffusion
is a martingale; see Lemma A.5 in this book or [176, Lemma 2.4]. It follows
that Z
t+∆
Et χ(t + ∆) − χ(t) − χ(s)Qds = 0,
t
where Et denotes the conditional expectation on the σ-algebra
Ft = {(X(u), α(u)) : u ≤ t}.
It then follows that
Z t+∆
χ(s)Qds = O(∆) a.s. (2.28)
t
Thus, we obtain
Et χ(t + ∆) = χ(t) + O(∆) a.s. (2.29)
In view of this structure, we see that
χ(t + ∆) − χ(t) = (χ1 (t + ∆) − χ1 (t), . . . , χm0 (t + ∆) − χm0 (t))
with χi (·) given by (2.27). This together with (2.29) implies
Et [χi (t + ∆) − χi (t)]2
= Et [I{α(t+∆)=i} − I{α(t)=i} ]2
(2.30)
= Et I{α(t+∆)=i} − 2I{α(t)=i} Et I{α(t+∆)=i} + I{α(t)=i}
= O(∆) a.s.
E[α(t + ∆) − α(t)]2
= E|[χ(t + ∆) − χ(t)](1, . . . , m0 )0 |2
≤ KE|χ(t + ∆) − χ(t)|2 (2.31)
m0
X
≤K EEt [χi (t + ∆) − χi (t)]2
i=1
≤ K∆ → 0 as ∆ → 0.
From the next to the last line to the last line above, we have used (2.30).
By combining (2.26) and (2.31), we obtain that (2.25) leads to
E|Y (t + ∆) − Y (t)|2 → 0 as ∆ → 0.
The mean square continuity has been established. Then the desired con-
tinuity in probability follows from Tchebyshev’s inequality. 2
We next generalize the above result and allow the switching process to
be x-dependent. The result is presented next.
2.5 Feller Property 41
Theorem 2.13. Suppose that the conditions of Theorem 2.1 are satis-
fied. Then the process Y (t) = (X(t), α(t)) is continuous in probability and
continuous in mean square.
Proof. Step 1 and Step 2 are the same as before. We only point out the
main difference as compared to Theorem 2.12.
Consider the function h(x, α) = I{α=i} for each i ∈ M. Because h is
independent of x, it is readily seen that Lh(x, α) = Q(x)h(x, ·)(α). Conse-
quently,
h Z t+∆ i
Et h(X(t + δ), α(t + ∆)) − h(X(t), α(t)) − Lh(X(s), i)ds = 0.
t
However,
h Z t+∆ i
Et h(X(t + ∆), α(t + ∆)) − h(X(t), α(t)) − Lh(X(s), α(s))ds
h t
= Et h(X(t + ∆), α(t + ∆)) − h(X(t), α(t))
Z t+∆ i
− Q(X(s))h(X(s), ·)(α(s))ds
t Z t+∆ X
h m0 i
= Et I{α(t+∆)=i} − I{α(t)=i} − qij (X(s))I{α(s)=i} ds .
t j=1
With (2.32) at our hands, we proceed with the rest of Step 3 and Step 4 in
the proof of Theorem 2.12. The desired result follows. 2
we say that X x (·) satisfies the Feller property or X x (·) is a Feller process
if for any t ≥ 0 and any bounded and continuous function g(·), u(t, x) =
Ex g(X(t)) is continuous with respect to x and limt↓0 u(t, x) = g(x); see [38,
57] and many references therein for the pioneering work on Feller properties
for the diffusions. For switching-diffusion processes, do the Feller and the
strong Feller properties carry over? This section is concerned with Feller
property, whereas the next section deals with strong Feller property.
We show that such properties carry over to the switching diffusions, but
the proofs are in fact nontrivial. The difficulty stems from the consideration
of x-dependent generator Q(x) for the discrete component α(t). The clas-
sical arguments for Feller and strong Feller properties of diffusions will not
work here. We need to prove that the function u(t, x, i) = Ex,i f (X(t), α(t))
is continuous with respect to the initial data (x, i) for any t ≥ 0 and
limt↓0 u(t, x, i) = f (x, i) for any bounded and continuous function f . By
virtue of Proposition 2.4 and the boundedness and continuity of f , we have
limt↓0 u(t, x, i) = Ex,i f (X(0), α(0)) = f (x, i). Note also that u(0, x, i) is au-
tomatically continuous by the continuity of f . Because M = {1, . . . , m0 } is
a finite set, it is enough to show that u(t, x, i) is continuous with respect to
x for any t > 0. In this section, we establish the Feller property for switch-
ing diffusions under Lipschitz continuity and linear growth conditions. This
section is motivated by the recent work [161]. The discussion of the strong
Feller property is deferred to Section 2.6.
We first present the following lemma. In lieu of the local Lipschitz con-
dition, we assume a global Lipschitz condition holds henceforth without
specific mentioning. Also, we assume for the moment that the discrete
component α(·) is generated by a constant Q.
Lemma 2.14. Assume the conditions of Theorem 2.1 hold with the mod-
ification of the local Lipschitz condition replaced by a global Lipschitz con-
dition. Moreover, suppose that α(·) is generated by Q(x) = Q and that α(·)
is independent of the Brownian motion w(·). Then for any fixed T > 0, we
have
2
2
E sup X xe,α (t) − X x,α (t) ≤ C |e
x − x| , (2.33)
0≤t≤T
where C is a constant depending only on T and the global Lipschitz and the
linear growth constant K.
2
Z t 2
e − X(t) 2e
X(t) ≤ 3 |∆x| + 3 [b(X(s), α(s)) − b(X(s), α(s))]ds
0
Z t 2
+3 e e(s)) − b(X(s), α(s))]dw(s) .
[σ(X(s), α
0
(2.34)
Using the assumption that b(·, j) is Lipschitz continuous for each j ∈ M,
we have
e
b(X(s), α(s)) − b(X(s), α(s))
m0
X
= e
[b(X(s), j) − b(X(s), j)]I{α(s)=j}
j=1
e
≤ K X(s) − X(s) ,
where in the above and hereafter, K is a generic positive constant not de-
e, or t whose exact value may change in different appearances.
pending on x, x
Consequently, we have from the Hölder inequality that
" Z 2
#
t
E sup e
[b(X(s), α(s)) − b(X(s), α(s))]ds
0≤t≤T1 0
Z T1 2
≤ TE e
b(X(s), α(s)) − b(X(s), α(s)) ds (2.35)
Z0 T1 2
≤ KT e
E X(s) − X(s) ds,
0
e
σ(X(s), e(s)) − σ(X(s), α
α e
e(s)) ≤ K X(s) − X(s) ,
e, x ∈ Rr , we have
and for α, ` ∈ M and x
|σ(e
x, α) − σ(x, α)| + |b(e
x, α) − b(x, α)| + |qα` (e
x) − qα` (x)|
≤ K |e
x − x| .
(2.38)
2.5 Feller Property 45
P (2.40)
where for each i ∈ M, qi (x) := −qii (x) = q
j6=i ij (x), τ i is a sequence of
stopping times defined by:
τ0 = 0,
and for i = 0, 1, . . .,
τi+1 := inf {t > τi : r(t) 6= r(τi )} ,
and n(T ) = max {n ∈ N : τn ≤ T }. Note that if n(T ) = 0, then
( Z )
T
x,α α
pT (Z (·), r (·)) = exp − [qα (Z(s)) − m0 + 1] ds .
0
Lemma 2.16. Assume the conditions of Theorem 2.1 and (A2.1). Then
for any T > 0 and (x, α) ∈ Rr × M, we have
Lemma 2.17. Assume the conditions of Lemma 2.16. Then for any T >
e, x ∈ Rr , and α ∈ M, we have
0, x
n
Y n
Y n−1
ck − dk ≤ n max {ck , dk } max {|ck − dk |} . (2.41)
k=1,...,n k=1,...,n
k=1 k=1
e
Applying (2.41) to pT (Z(·), r(·)) and pT (Z(·), r(·)), we have
e
E pT (Z(·), r(·)) − pT (Z(·), r(·))
" ∞
X
≤ exp {(m0 − 1)T } E (n + 1)K n I{n(T )=n}
( n=0
RT RT
e
× max e− τn
qr(τn ) (Z(s))ds
− e− τn
qr(τn ) (Z(s))ds
,
i=0,1,...,n−1
R τi+1
e i+1 ))e− e
qr(τi ) (Z(s))ds
qr(τi )r(τi+1 ) (Z(τ τi
)#
R τi+1
−qr(τi )r(τi+1 ) (Z(τi+1 ))e− τi qr(τi ) (Z(s))ds
.
R τi+1
−qr(τi )r(τi+1 ) (Z(τi+1 ))e− τi qr(τi ) (Z(s))ds
h Rτ R τi+1 i
e
e i+1 )) e− τii+1 qr(τi ) (Z(s))ds
≤ qr(τi )r(τi+1 ) (Z(τ − e− τi qr(τi ) (Z(s))ds
h i Rτ
e i+1 )) − qr(τ )r(τ ) (Z(τi+1 )) e− τii+1 qr(τi ) (Z(s))ds
+ qr(τi )r(τi+1 ) (Z(τ i i+1
Z τi+1
≤K e
qr(τi ) (Z(s)) − qr(τi ) (Z(s)) ds + K Z(τe i+1 ) − Z(τi+1 )
τi n o
≤ K(T + 1) sup Z(s) e − Z(s) : 0 ≤ s ≤ T .
e
E pT (Z(·), r(·)) − pT (Z(·), r(·))
" ∞
X
≤E exp {(m0 − 1)T } (n + 1)K n I{n(T )=n}
n=0 #
n o
×K(T + 1) sup e − Z(s) : 0 ≤ s ≤ T
Z(s)
∞
X
≤ exp {(m0 − 1)T } (T + 1)(n + 1)K n+1
n=0 n o 2
1/2
×E 1/2
I{n(T )=n} E e
sup Z(s) − Z(s) : 0 ≤ s ≤ T .
As noted in the proof of Lemma 2.16, n(T ) is a Poisson process with rate
m0 − 1. Hence
[(m0 − 1)T ]n
E I{n(T )=n} = P {n(T ) = n} = exp {−(m0 − 1)T } .
n!
48 2. Switching Diffusion
Also, because the generator of r(t) of the auxiliary process (Z(t), r(t)) is a
constant matrix, by virtue of Lemma 2.14, we have
n o 2
E sup Z(s) e − Z(s) : 0 ≤ s ≤ T
2
= E sup Z(s) e − Z(s) : 0 ≤ s ≤ T
2
≤ K |e
x − x| .
X∞
(n + 1)K n/2
√ < ∞,
n=0 n!
e − x, or t. This
e, x, x
and as before, K is a generic constant independent of x
completes the proof of the lemma. 2
With the preparations above, we prove the Feller property for the switch-
ing diffusion process.
Theorem 2.18. Let (X x,α (t), αx,α (t)) be the solution to the system given
by (2.2)–(2.3) with (X(0), α(0)) = (x, α). Assume the conditions of Theo-
rem 2.1 hold. Then for any bounded and continuous function g(·, ·) : Rr ×
M → R, the function u(x, α) = Ex,α g(X(t), α(t)) = Eg(X x,α (t), αx,α (t))
is continuous with respect to x.
Proof. We prove the theorem in two steps. The first step deals with the
special case when Assumption (A2.1) is true. The general case is treated
in Step 2.
Step 1. Assume (A2.1). Then for any fixed t > 0, we have
E [g(X x,α (t), αx,α (t))] = E [g(Z x,α (t), r α (t))pt (Z x,α (·), r α (·))] .
Now g is bounded and continuous, it follows from Lemma 2.16 and the
dominated convergence theorem that
aN N
jk (x, i) := ajk (x, i)φ (x),
bN N
j (x, i) := bj (x, i)φ (x).
50 2. Switching Diffusion
n
1 X N ∂2
LN ϕ(x, i) = ajk (x, i) ϕ(x, i)
2 ∂xj ∂xk
j,k=1
X n m0
X (2.42)
∂
+ bN
j (x, i) ϕ(x, i) + qij (x)ϕ(x, j).
j=1
∂xj j=1
Denote by PN x,i the probability measure for which the associated martingale
problem [153] has operator LN with coefficients aN N
jk (x, i), bj (x, i), and
qij (x), and denote by EN x,i the corresponding expectation. Then by Step 1,
N
(X (t), α(t)) is Feller.
As in (2.14), let βN be the first exit time from the ball B(0, N ). Then
by virtue of the strong uniqueness result in [172], the probabilities Px,i
and PN x,i agree until the moment when the continuous component reaches
the boundary |x| = N . Hence it follows that for any bounded and Borel
measurable function f (·, ·) : Rr × M 7→ R, we have
Ex,i f (X(t), α(t))I{βN >t} = EN
x,i f (X(t), α(t))I{βN >t} . (2.43)
(Alternatively, one can obtain (2.43) by showing that the solution to the
martingale problem with operator L is unique in the weak sense. The weak
uniqueness can be established by using the characteristic function as in the
proof of [176, Lemma 7.18].)
Fix (x0 , α) ∈ Rr × M. By virtue of (2.19) and (2.43), it follows that for
any t > 0 and ε > 0, there exists an N ∈ N sufficiently large and a bounded
neighborhood N (x0 ) of x0 (N (x0 ) = {x : |x − x0 | ≤ N }) such that for any
(x, i) ∈ N (x0 ) × M, we have
= 2 kgk [1 − PN
x,i {βN > t}]
as desired. 2
Corollary 2.19. Assume that the conditions of Theorem 2.18 hold. Then
the process (X(t), α(t)) is strong Markov.
Proof. This claim follows from [47, Theorem 2.2.4], Lemma 2.4, and The-
orem 2.18. 2
In Step 2 of the proof of Theorem 2.18, we used a truncation device and
defined X N (t). As a digression, using such a truncation process, we obtain
the following result as a by-product. It enables us to extend the existence
and uniqueness to more general functions satisfying only local linear growth
and Lipschitz conditions.
Proposition 2.20. Under the conditions of Theorem 2.7, equation (2.2)
together with (2.3) has a unique solution a.s.
where (x, i) = (X(0), α(0)), N ∈ N satisfying N > |x|, βN is the first exit
time from the bounded ball B(0, N ) as defined in (2.14), and
Z r∧βN
M1 (r ∧ βN ) = ∇u(t − s, X(s), α(s)), σ(X(s), α(s))dw(s) .
Z0 r∧βN Z
M2 (r ∧ βN ) = [u(t − s, X(s), α(0) + h(X(s), α(s), z))
0 R
−u(t − s, X(s), α(s))] µ(ds, dz),
∂u
= Lu, t>0
∂t
has a unique fundamental solution p(x, i, t, y, j) (as a function of t, x, i),
which is positive, jointly continuous in t, x, and y, and satisfies
( )
2
θ (−r+|θ|)/2 −c |y − x|
Dx p(x, i, t, y, j) ≤ Ct exp , (2.49)
t
∂θ ∂ |θ|
Dxθ = := .
∂xθ ∂xθ11 ∂xθ22 · · · ∂xθrr
To see that p is the transition probability density of the process (X(t), α(t)),
consider an arbitrary bounded and continuous function φ(x, i), (x, i) ∈
Rr × M and define
m0 Z
X
Φ(t, x, i) := p(x, i, t, y, j)φ(y, j)dy.
j=1 Rr
Lemma 2.23. Assume the process (X(t), α(t)) is strong Feller. Denote
U := D × J ⊂ Rr × M, where D is a nonempty bounded open subset of
Rr . Then for any t > 0 and every bounded real Borel measurable function
f (·, ·) on U , the functions
F (x, i) := Ex,i I{τU >t} f (X(t), α(t)) ,
G(x, i) := Ex,i [f (X(t ∧ τU ), α(t ∧ τU ))] ,
Theorem 2.24. Assume (A2.2)–(A2.4) hold. Then the process (X(t), α(t))
possesses the strong Feller property.
It follows that
= 2 kf k [1 − PN
x,i {βN > t}],
2.6 Strong Feller Property 55
where k·k is the essential sup norm. Fix some (x0 , i) ∈ Rr ×M. Because the
process (X(t), α(t)) is regular, by (2.15), βN → ∞ a.s. Px0 ,i as N → ∞.
Therefore for any positive number ε > 0, we can choose some N sufficiently
large so that
ε
1 − PN
x0 ,i {βN > t} = 1 − Px0 ,i {βN > t} < . (2.51)
12 kf k
Also by Lemma 2.23, the function x 7→ PNx,i {βN > t} is continuous. Hence
there exists some δ1 > 0 such that whenever |x − x0 | < δ1 we have
N
1 − PN
x,i {βN > t} ≤ 1 − Px0 ,i {βN > t}
+ PN N
x0 ,i {βN > t} − Px,i {βN > t} (2.52)
ε
< .
6 kf k
Note also that by virtue of Lemma 2.22, there exists some δ2 > 0 such that
EN N
x,i f (X(t), α(t)) − Ex0 ,i f (X(t), α(t)) < ε/3, if |x − x0 | < δ2 . (2.53)
+ EN N
x,i f (X(t), α(t)) − Ex0 ,i f (X(t), α(t))
< ε.
We end this section with a brief discussion of strong Feller processes and
L-harmonic functions. More properties of L-harmonic functions are treated
in Chapter 3. As in [52], for any U = D × J, where D ⊂ Rr is a nonempty
domain, and J ⊂ M, a Borel measurable function u : U 7→ R is said to be
L-harmonic in U if u is bounded in compact subsets of U and that for all
56 2. Switching Diffusion
We treat each of the terms above separately. Choose η = ∆γ0 with γ0 > 2
and partition the interval [0, T ] by η. We obtain
Z T
E e
|b(X(s), e
e(s)) − b(X(s),
α α(s))|2 ds
0
bT /ηc−1 Z kη+η
X
=E e
|b(X(s), e
e(s)) − b(X(s),
α α(s))|2 ds
k=0 kη
bT /ηc−1 h
X Z kη+η
≤ KE e
|b(X(s), e
e(s)) − b(X(ηk),
α e(s))|2 ds
α (2.57)
k=0 Z kη
kη+η
+ KE e
|b(X(ηk), e
e(s)) − b(X(ηk),
α α(s))|2 ds
Zkηkη+η i
+ KE e
|b(X(ηk), e
α(s)) − b(X(s), α(s))|2 ds .
kη
Likewise, we can deal with the term on the last line of (2.57), and obtain
Z kη+η
E e
|b(X(ηk), e
α(ηk)) − b(X(s), α(s))|2 ds ≤ Kη 2 . (2.59)
kη
To treat the term on the next to the last line of (2.57), note that for
k = 0, 1, . . . , bT /ηc − 1,
Z kη+η
E e
|b(X(ηk), e
e(s)) − b(X(ηk),
α α(s))|2 ds
kη
Z kη+η
≤ KE e
|b(X(ηk), e
e(s)) − b(X(ηk),
α e(ηk))|2 ds
α (2.60)
kη
Z kη+η
+KE e
|b(X(ηk), e
e(ηk)) − b(X(ηk),
α α(s))|2 ds.
kη
2.7 Continuous and Smooth Dependence on the Initial Data x 59
Z kη+η
E e
|b(X(ηk), e
e(s)) − b(X(ηk),
α e(ηk))|2 ds
α
kη
Z kη+η
=E e
|b(X(ηk), α e
e(s)) − b(X(ηk), e(ηk))|2 I{e
α e(ηk)} ds
α(s)6=α
kη
Z
X X kη+η
=E e
|b(X(ηk), e
i) − b(X(ηk), j)|2 I{e
α(s)=j} I{e
α(ηk)=i} ds
i∈M j6=i kη
X XZ kη+η
≤ KE e
[1 + |X(ηk)| 2
]I{e
α(ηk)=i}
i∈M j6=i kη
×E[I{e e
α(s)=j} X(ηk), α e(ηk) = i]ds
X Z kη+η
≤ KE e
[1 + |X(ηk)| 2
]I{e
α(ηk)=i}
i∈M kη
hX i
× e
qij (X(ηk))(s − ηk) + o(s − ηk) ds
j6=i
Z kη+η
≤K O(η)ds ≤ Kη 2 .
kη
Z kη+η
E e
|b(X(ηk), e
e(ηk)) − b(X(ηk),
α α(s))|2 ds ≤ Kη 2 . (2.61)
kη
h i
P (α(t + h), α e(t + h)) = (j, i) (α(t), α e
e(t)) = (k, l), (X(t), X(t)) e)
= (x, x
e)h + o(h),
qe(k,l)(j,i) (x, x if (k, l) 6= (j, i),
=
e)h + o(h), if (k, l) = (j, i),
1 + qe(k,l)(k,l) (x, x
(2.62)
where h → 0, and the matrix (e e)) is the basic coupling of ma-
q(k,l)(j,i) (x, x
60 2. Switching Diffusion
X
e x
Q(x, e)fe(k, l) = e)(fe(j, i) − fe(k, l))
q(k,l)(j,i) (x, x
(j,i)∈M×M
X
= x))+ (fe(j, l) − fe(k, l))
(qkj (x) − qlj (e
j
X (2.63)
+ x) − qkj (x))+ (fe(k, j) − fe(k, l))
(qlj (e
j
X
+ x))(fe(j, j) − fe(k, l)),
(qkj (x) ∧ qlj (e
j
for any function fe(·, ·) defined on M × M. Note that for s ∈ [ηk, ηk + η),
e(s) can be written as
α
X
e(s) =
α lI{e
α(s)=l} .
l∈M
Owing to the coupling defined above and noting the transition probabilities
(2.62), for i1 , i, j, l ∈ M with j 6= i and s ∈ [ηk, ηk + η), we have
E[I{α(s)=j} α(ηk) = i1 , α e
e(ηk) = i, X(ηk) = x, X(ηk) =xe]
X
= E[I{α(s)=j} I{e
α(s)=l} α(ηk) = i1 , α
e
e(ηk) = i, X(ηk) = x, X(ηk) e]
=x
l∈M
X
= e)(s − ηk) + o(s − ηk) = O(η).
qe(i1 ,i)(j,l) (x, x
l∈M
(2.64)
By virtue of (2.64), we obtain
Z kη+η
E e
|b(X(ηk), e
e(ηk)) − b(X(ηk),
α α(s))|2 ds
kη
Z kη+η
=E e
|b(X(ηk), e
α(s)) − b(X(ηk), e(ηk))|2 I{α(s)6=αe(ηk)} ds
α
kη
X X Z kη+η
=E e
|b(X(ηk), e
i) − b(X(ηk), j))|2 I{α(s)=j} I{e
α(ηk)=i} ds
i∈M j6=i kη
X XZ kη+η
≤ KE e
[1 + |X(ηk)| 2
]I{e
α(ηk)=i,α(ηk)=i1 }
i,i1 ∈M j6=i kη
× E[α(s) = j α(ηk) = i1 , α e
e(ηk) = i, X(ηk) = x, X(ηk) e]ds
=x
= O(η 2 ).
2.7 Continuous and Smooth Dependence on the Initial Data x 61
Likewise, we obtain
Z T 2
E e
[σ(X(s), α e
e(s)) − σ(X(s), α(s))]dw(s) ≤ Kη. (2.68)
0
Putting (2.67) and (2.68) into φ∆ (t), and noting γ0 > 2, we obtain
η
E sup |φ∆ (t)|2 ≤ K = K∆γ0 −2 → 0 as ∆ → 0. (2.69)
0≤t≤T ∆2
Proposition 2.30. Using the conditions of Lemma 2.14 except Q(x) being
allowed to be x-dependent, the conclusion of Lemma 2.14 continues to hold.
Proof. As before, let (X(t), α(t)) denote the switching diffusion process
e
satisfying (2.2) and (2.3) with initial condition (x, α) and (X(t), e(t)) be the
α
process starting from (e e
x, α) (i.e., (X(0), α(0)) = (x, α)) and (X(0), e(0)) =
α
62 2. Switching Diffusion
(e e − x. Then we
x, α) respectively). Let T > 0 be fixed and denote ∆ = x
e − X(t) = ∆ + A(t) + B(t), and hence
have X(t)
2
e − X(t) 2 2
sup X(t) ≤ 3∆2 + 3 sup |A(t)| + 3 sup |B(t)| ,
t∈[0,T ] t∈[0,T ] t∈[0,T ]
where
Z t
A(t) := e
[b(X(s), α e
e(s)) − b(X(s), α(s))]ds
0Z
t
+ e
[σ(X(s), e
e(s)) − b(X(s),
α α(s))]dw(s) = ∆φ∆ (t),
0
and
Z t
B(t) := e
[b(X(s), α(s)) − b(X(s), α(s))]ds
0Z
t
+ e
[σ(X(s), α(s)) − b(X(s), α(s))]dw(s).
0
Therefore, we have
2
e − X(t) 2
E sup X(t) ≤ K∆2 + o(∆2 ) ≤ K |e
x − x| .
t∈[0,T ]
Corollary 2.31. Assume the conditions of Theorem 2.27. Then X x,α (t)
is continuous in mean square with respect to x.
2.7 Continuous and Smooth Dependence on the Initial Data x 63
Proof of Theorem 2.27. With Lemma 2.28 and Proposition 2.30 at our
hands, we proceed to prove Theorem 2.27. Because b(·, j) is twice continu-
ously differentiable with respect to x, we can write
Z
1 t e
[b(X(s), α(s)) − b(X(s), α(s))]ds
∆ 0 Z Z
1 t 1 d e
= b(X(s) + v(X(s) − X(s)), α(s))dvds
Z∆ t Z
0
1
0 dv
= e
bx (X(s) + v(X(s) − X(s)), α(s))dv Z ∆ (s)ds,
0 0
where Z ∆ (t) is defined in (2.55) and bx (·) denotes the partial derivative of
b(·) with respect to x (i.e., bx = (∂/∂x)b). It follows from Lemma 2.30 that
for any s ∈ [0, T ],
e
X(s) − X(s) → 0
in probability as ∆ → 0. This implies that
Z 1
e
bx (X(s) + v(X(s) − X(s)), α(s))dv → bx (X(s), α(s)) (2.70)
0
and
Z 1
e
σx (X(s) + v(X(s) − X(s)), α(s))dv → σx (X(s), α(s)) (2.71)
0
Corollary 2.32. Under the assumptions of Theorem 2.27, the mean square
derivatives (∂/∂xj )X x,α (t) and (∂ 2 /∂xj ∂xk )X x,α (t), j, k = 1, . . . , n, are
mean square continuous with respect to t.
Proof. As in the proof of Theorem 2.27, we consider only the case when
X(t) is real valued. Also, we use the same notations here as those in the
proof of Theorem 2.27. To see that ζ(t) is continuous in the mean square
sense, we first observe that for any t ∈ [0, T ],
2 2 2
E |ζ(t)| ≤ 2E ζ(t) − Z ∆ (t) + 2E Z ∆ (t) .
It follows from (2.69), the Lipschitz condition, and Lemma 2.30 that
Z 2
2 2 1 t e
E Z ∆ (t) ≤ 3E φ∆ (t) + 3E [b(X(u), α(u)) − b(X(u), α(u))]du
∆ 0
Z 2
1 t e
+ 3E [σ(X(u), α(u)) − σ(X(u), α(u))]dw(u)
∆ 0 Z
t 2
1 e
≤ K + 3t 2E b(X(u), α(u)) − b(X(u), α(u)) du
|∆|Z 0
t
1 e
2
+ 3E 2 σ(X(u), α(u)) − σ(X(u), α(u)) du
|∆| 0 Z t
1 2
≤ K + 3K(T + 1) e − X(u) du
2E X(u)
|∆| 0
≤ C = C(x, T, K).
≤ KC(t − s)2 ,
where K is the Lipschitz constant and C is a constant independent of t, s,
or ∆. Similarly, we can show that
Z 2
1 t e
E σ(X(u), α(u)) − σ(X(u), α(u))dw(u)
∆ s Z
t
1 e
2
(2.77)
=E 2 σ(X(u), α(u)) − σ(X(u), α(u)) du
|∆| s
≤ KC(t − s).
Next, using the same argument as that of Lemma 2.28, we can show that
2
E φ∆ (t) − φ∆ (s) ≤ K(t − s). (2.78)
Thus it follows from (2.75)–(2.78) that
2
E Z ∆ (t) − Z ∆ (s) = O(|t − s|) → 0 as |t − s| → 0
and hence ζ(t) is mean square continuous with respect to t.
Likewise, we can show that (∂ 2 /∂x2 )X x,α (t) is mean square continuous
with respect to t. This concludes the proof. 2
and as ∆ → 0,
In view of the generalized Itô formula, for any f (·, ·, ı) ∈ C 1,2 with ı ∈ M,
and τ1 , τ2 being bounded stopping times such that 0 ≤ τ1 ≤ τ2 a.s., if
f (t, X(t), α(t)) and Lf (t, X(t), α(t)) etc. are bounded on t ∈ [τ1 , τ2 ] with
probability 1, then Dynkin’s formula becomes
Z τ2
Ef (τ2 , X(τ2 ), α(τ2 )) = Ef (τ1 , X(τ1 ), α(τ1 )) + E Lf (s, X(s), α(s))ds.
τ1
(2.84)
Moreover, for each f (·, ·, ι) ∈ Cb1,2 or f (·, ·, ι) ∈ C01,2 ,
2.9 Notes
The connection between generators of Markov processes and martingales is
explained in Ethier and Kurtz [43]. An account of piecewise-deterministic
processes is in Davis [30]. Results on basic probability theory may be found
in Chow and Teicher [27]; the theory of stochastic processes can be found
in Gihman and Skorohod [53], Khasminskii [83], and Liptser and Shiryayev
[110], among others. More detailed discussions regarding martingales and
diffusions are in Elliott [40]; an in-depth study of stochastic differential
equations and diffusion processes is contained in Ikeda and Watanabe [72].
Concerning the existence of solutions to stochastic differential equations
with switching, using Poisson random measures (see Skorohod [150] also
Basak, Bisi, and Ghosh [6] and Mao and Yuan [120]), it can be shown
that there is a unique solution for each initial condition by following the
approach of Ikeda and Watanabe [72] with the appropriate use of the stop-
ping times. However, the Picard iteration method does not work, which is
further explained when we study the numerical solutions (see Chapter 5).
In Section 2.7, we dealt with smoothness properties of solutions of stochas-
tic differential equations with x-dependent switching. It is interesting to
note that even the time-honored concept of well-posedness cannot be eas-
ily obtained. Once x-dependence is added, the difficulty rises considerably.
3
Recurrence
3.1 Introduction
This chapter is concerned with recurrence of switching diffusion processes.
Because practical systems in applications are often in operation for a rel-
atively long time, it is of foremost importance to understand the systems’
asymptotic behavior. By asymptotic behavior, we mean the properties of
the underlying processes in a neighborhood of “∞” and in a neighborhood
of an equilibrium point. Properties concerning a neighborhood of ∞ are
treated in this chapter, whereas stability of an equilibrium point is dealt
with in Chapter 7.
Dealing with dynamic systems, one often wishes to examine if the un-
derlying system is sensitive to perturbations. In accordance with LaSalle
and Lefschetz [107], a deterministic system ẋ = h(t, x) satisfying appropri-
ate conditions, is Lagrange stable if the solutions are ultimately uniformly
bounded. When diffusions are used, one may wish to add probability modi-
fications such as “almost surely” or “in probability” to the aforementioned
uniform boundedness. Nevertheless, for instance, if almost sure bounded-
ness is used, it will exclude many systems due to the presence of Brownian
motion. Thus as pointed out by Wonham [160], such boundedness is inap-
propriate; an alternative notion of stability in a certain weak sense should
be used. In lieu of requiring the system to be bounded, one aims to find
conditions under which the systems return to a prescribed compact region
in finite time. This chapter focuses on weak-sense stability for switching
diffusions. We define recurrence, positive recurrence, and null recurrence;
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 69
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_3,
© Springer Science + Business Media, LLC 2010
70 3. Recurrence
r r
1 X ∂ 2 h(x, i) X ∂h(x, i)
Lh(x, i) = ajk (x, i) + bj (x, i) + Q(x)h(x, ·)(i)
2 ∂xj ∂xk j=1
∂xj
j,k=1
1
= tr(a(x, i)∇2 h(x, i)) + b0 (x, i)∇h(x, i) + Q(x)h(x, ·)(i),
2
(3.1)
where ∇h(·, i) and ∇2 h(·, i) denote the gradient and Hessian of h(·, i),
respectively, b0 (x, i)∇h(x, i) denotes the usual inner product on Rr with z 0
3.2 Formulation and Preliminaries 71
and
P{α(t + ∆) = j|α(t) = i, X(s), α(s), s ≤ t}
(3.4)
= qij (X(t))∆ + o(∆), i 6= j,
where w(t) is a d-dimensional standard Brownian motion, b(·, ·) : Rr ×M 7→
Rr , and σ(·, ·) : Rr × M 7→ Rr×d satisfying σ(x, i)σ 0 (x, i) = a(x, i). Note
that (3.3) depicts the system dynamics and (3.4) delineates the probability
structure of the jump process. Note that if α(·) is a continuous-time Markov
chain independent of the Brownian motion w(·) and Q(x) = Q or Q(x) =
Q(t) (independent of x), then equation (3.3) together with the generator Q
or Q(t) is sufficient to characterize the underlying process. As long as there
is an x-dependence, equation (3.4) is needed in delineating the dynamics
of the switching diffusion.
In this chapter, our study is carried out with the use of the operator L
given in (3.1). Throughout the chapter, we assume that both b(·, i) and
σ(·, i) satisfy the usual local Lipschitz and linear growth conditions for
each i ∈ M and that Q(·) is bounded and continuous. As described in
Theorem 2.1, the system (3.3)–(3.4) has a unique strong solution. In what
follows, denote the solution of (3.3)–(3.4) by (X x,α (t), αx,α (t)) when we em-
phasize the dependence on initial data. To study recurrence and ergodicity
of the process Y (t) = (X(t), α(t)), we further assume that the following
condition (A3.1) holds throughout the chapter. For convenience, we also
put the boundedness and continuity of Q(·) in (A3.1).
τD := inf{t ≥ 0 : X(t) ∈
/ D},
(3.7)
σD := inf{t ≥ 0 : X(t) ∈ D}.
Definition 3.1. Recurrence and positive and null recurrence are defined
as follows.
• Recurrence and Transience. For U := D × J, where J ⊂ M and
D ⊂ Rr is an open set with compact closure, let
x,α
σU = inf{t : (X x,α (t), αx,α (t)) ∈ U }.
A regular process (X x,α (·), αx,α (·)) is recurrent with respect to U if
x,α
P{σU < ∞} = 1 for any (x, α) ∈ D c × M, where D c denotes the
complement of D; otherwise, the process is transient with respect to
U.
• Positive Recurrence and Null Recurrence. A recurrent process with
finite mean recurrence time for some set U = D × J, where J ⊂ M
and D ⊂ Rr is a bounded open set with compact closure, is said to
be positive recurrent with respect to U ; otherwise, the process is null
recurrent with respect to U .
Proof. First, note that from the uniform ellipticity condition (3.5), we have
Set !
2
c= sup |b1 (x, i)(x1 + β)| + 1 + 1.
κ1 (x,i)∈D×M
Now let τD (t) = t∧τD := min{t, τD }. Then we have from Dynkin’s formula
and (3.10) that
is the unique solution of the differential equation (3.13) with boundary con-
dition
Note that Theorem 3.2 implies that Px,i {τV < ∞} = 1. Letting t → ∞,
we obtain by virtue of the bounded convergence theorem that
This shows that f (·, i) ∈ C 2 (D) for each i ∈ M and that f satisfies the
differential equation (3.13).
Step 3. Now assume ∂D is sufficiently smooth, D is compact, and ϕ(·, i)
is an arbitrary continuous function on ∂D for any i ∈ M. We show that the
function u defined in (3.14) is the unique solution of (3.13) with boundary
condition (3.15). Indeed, it follows from Lemma 3.4 that u is L-harmonic in
U . Then we have from Step 2 that u satisfies the differential equation (3.13).
Finally, the boundary condition (3.15) is satisfied by the assumptions that
∂D is sufficiently smooth, D is compact, and that φ is continuous. This
completes the proof of the lemma. 2
Using a similar argument, we can prove the following lemma.
Lemma 3.6. Let U = D × M ⊂ Rr × M, where D ⊂ Rr is a nonempty
open set with compact closure. Suppose that g(·, i) ∈ Cb (D) and f (·, ·) :
D × M 7→ R. Then f solves the boundary value problem
Lf (x, i) = −g(x, i), (x, i) ∈ D × M
f (x, i) = 0, (x, i) ∈ ∂D × M
if and only if
Z τU
f (x, i) = Ex,i g(X(t), α(t))dt, for all (x, i) ∈ D × M.
0
Using Lemmas 3.5 and 3.6, we proceed to prove that if the process
Y (t) = (X(t), α(t)) is recurrent (resp., positive recurrent) with respect
to some “cylinder” D × M ⊂ Rr × M, then it is recurrent (resp., positive
recurrent) with respect to any “cylinder” E ×M ⊂ Rr ×M, where D is any
nonempty domain in Rr with compact closure. These results are proved in
the following two lemmas. To preserve the flow of presentation, the proofs
are postponed to Section 3.6.
Lemma 3.7. Let D ⊂ Rr be a nonempty open set with compact closure.
Suppose that
The following lemma shows that if the process Y (t) = (X(t), α(t)) reaches
the “cylinder” D×M in finite time a.s. Px,i , then it will visit the set D×{`}
in finite time a.s. Px,i for any ` ∈ M. Its proof together with the proof of
Lemma 3.10 is placed in Section 3.6 as well.
Lemma 3.9. Let D ⊂ Rr be a nonempty open set with compact closure
satisfying
With Lemma 3.9, we can now prove that if the process Y (t) = (X(t), α(t))
is positive recurrent with respect to some “cylinder” D × M, then it is pos-
itive recurrent with respect to the set D × {`} ⊂ Rr × M.
Lemma 3.10. Let D ⊂ Rr be a nonempty open set with compact closure
satisfying
Ey,j σD < ∞ for any (y, j) ∈ D c × M. (3.21)
Then for any (x, i) ∈ Rr × M,
Vn := inf V (x, i) → ∞, as n → ∞,
|x|≥n,i∈M (3.23)
LV (x, i) ≤ 0, for all (x, i) ∈ D c × M.
Consequently,
(n) (n)
EV (X(σD (t)), α(σD (t))) ≤ V (x, α).
V (x, α)
P {βn < σD } ≤ → 0.
Vn
Note that
P {σD = ∞} ≤ P {βn < σD } .
Thus we have P {σD < ∞} = 1, as desired. 2
The above result is based on a Liapunov function argument. However,
constructing Liapunov functions is generally difficult. It would be nice if we
could place certain conditions on the coefficients of processes. The following
theorem is an attempt in this direction.
Theorem 3.15. Either one of the following conditions implies that the
process (X(t), α(t)) is recurrent.
(i) There exist constants γ > 0 and ci ∈ R with i ∈ M such that for
(x, i) ∈ {x ∈ Rr : |x| ≥ 1} × M,
m0
X
x0 b(x, i) tr(a(x, i)) x0 a(x, i)x 1
+ + (γ − 2) − qij (x)cj ≤ 0,
|x|2 2|x|2 2|x|4 k − γci j=1
(3.24)
where k is a positive constant sufficiently large so that k − γci > 0
for each i ∈ M.
(ii) There exist a positive constant γ and symmetric and positive definite
matrices Pi for i ∈ M such that for (x, i) ∈ {x ∈ Rr : |x| ≥ 1} × M,
Lemma 3.17. If the process (X(t), α(t)) is recurrent, then for any (x, i) ∈
Rr × M and ε > 0, we have
Proof. Denote B = B(x, ε), B1 = B(x, ε/2), and B2 = B(x, 2ε). Define a
sequence of stopping times by
η1 := inf {t ≥ 0 : X(t) ∈
/ B2 } ;
and for n = 1, 2, . . .,
Note that the process (X(t), α(t)) is recurrent, in particular, with respect
to B1 × {i}. This, together with Theorem 3.2, implies that ηn < ∞ a.s.
Px,i . Thus (3.27) follows. 2
Combining Lemmas 3.16 and 3.17, we obtain the following theorem.
Theorem 3.18. The process (X(t), α(t)) is recurrent if and only if every
point (x, i) ∈ Rr × M is recurrent in the sense that for any ε > 0,
3.3.2 Transience
We first argue that if the process (X(t), α(t)) is transient, then the norm
of the continuous component |X(t)| → ∞ a.s. as t → ∞, and vice versa.
Then we provide two criteria for transience in terms of mean sojourn time
and Liapunov functions, respectively. If the coefficients b(x, i) and σ(x, i)
for i ∈ M are linearizable in x, we also obtain easily verifiable conditions
for transience.
Theorem 3.20. The process (X(t), α(t)) is transient if and only if
lim |X(t)| = ∞ a.s. Px,α for any (x, α) ∈ Rr × M.
t→∞
sup V (x, i) ≤ 0,
(x,i)∈∂D×M
Proof. Assuming the process (X(t), α(t)) is regular, we need to show that
it is transient. Fix (y, `) ∈ D c × M with V (y, `) > 0. Define the stopping
y,`
times σD = σD , βn = βny,` , and βn ∧σD ∧t as in the proof of Theorem 3.14,
with n > (n0 ∨ |y|), where n0 is an integer such that D ⊂ {x : |x| < n0 }
and βny,` = inf{t : |X y,` (t)| = n}. By virtue of Dynkin’s formula and (3.32),
V (y, `)
P(An ) ≥ , for any n ≥ n0 .
M
Note that βn ≤ βn+1 implies An ⊃ An+1 . This, together with the regularity
yields that
∞
\
An = lim An = {σD > t} .
n→∞
n=n0
Therefore,
V (y, `)
P {σD > t} ≥ .
M
Finally, by letting t → ∞, we obtain that
P {σD = ∞} > 0.
Thus the process (X(t), α(t)) is transient. This completes the proof. 2
To proceed, we focus on linearizable (in the continuous component) sys-
tems. In addition to condition (A3.1), we also assume that the following
condition holds.
∇W (x, i) = −β|x|β−2 x,
∇2 W (x, i) = −β |x|β−2 I + (β − 2)|x|β−4 xx0 .
Note that
Pd
0
x b(i)x
d
1 X x0 σj0 (i)σj (i)x x0 b0 (i) + b(i) + j=1 σj0 (i)σj (i) x
+ =
|x|2 2 j=1
|x|2 2|x|2
Xd
1
≥ λmin b(i) + b0 (i) + σj0 (i)σj (i) .
2 j=1
(3.36)
3.4 Positive and Null Recurrence 85
where in the last step, we used the fact that β < 0 and condition (3.34).
Hence the second equation in (3.32) of Theorem 3.23 is valid. Thus Theo-
rem 3.23 implies that the process (X(t), α(t)) is not recurrent. 2
(n)
Note that the function V is nonnegative, hence we have Ex,i σD (t) ≤
V (x, i). Meanwhile, because the process Y (t) = (X(t), α(t)) is regular, it
(n)
follows that σD (t) → σD (t) a.s. as n → ∞, where σD (t) = σD ∧ t. By
virtue of Fatou’s lemma, we obtain
Now the argument after equation (3.11) in the proof of Theorem 3.2 yields
that Ex,i σD ≤ V (x, i) < ∞. Then Lemma 3.10 implies that Ex,i σU =
Ex,i σD×{`} < ∞. Since (x, i) ∈ D c × M is arbitrary, we conclude that Y (t)
is positive recurrent with respect to U .
Step 2: Show that u(x, i) := Ex,i σD is the smallest positive solution of
(3.39). To this end, let n0 be defined as before, that is, a positive integer
(n)
sufficiently large so that D ⊂ {|x| < n0 }. For n ≥ n0 , set σD = σD ∧ βn .
(n) (n+1)
Clearly, we have σD ≤ σD for all n ≥ n0 . Then the regularity of the
(n)
process Y (t) implies that σD % σD a.s. as n → ∞. Hence the monotone
convergence theorem implies that as n → ∞,
(n)
Ex,i σD % Ex,i σD . (3.41)
Note that Ex,i σD < ∞ from Step 1. Meanwhile, Lemma 3.6 implies that
(n)
the function un (x, i) = Ex,i σD solves the boundary value problem
Using Harnack’s inequality for L-elliptic systems of equations (see [3, 25],
and also [158] for general references on elliptic systems), it can be shown by
a slight modification of the well-known arguments (see, e.g., [56, pp. 21–22])
that the sum of a convergent series of positive L-harmonic functions is also
an L-harmonic function. Hence we conclude that u(x, i) is twice continu-
ously differentiable and satisfies equation (3.39). To verify that u(x, i) is
the smallest positive solution of (3.39), let w(x, i) be any positive solution
(n)
of (3.39). Note that un (x, i) = Ex,i σD satisfies the boundary conditions
un (x, i)|x∈∂D = 0,
un (x, i)||x|=n = 0, i ∈ M.
Then the functions un (x, i) − w(x, i) for i ∈ M are L-harmonic and satisfy
un (x, i) − w(x, i) = 0 for (x, i) ∈ ∂D × M and un (x, i) − w(x, i) < 0 for
(x, i) ∈ {|x| = n} × M. Hence it follows from the maximum principle
for L-elliptic system of equations [138, p. 192] that un (x, i) ≤ w(x, i) in
(Dc ∩ {|x| < n}) × M for all n ≥ n0 . Letting n → ∞, we obtain u(x, i) ≤
w(x, i), as desired.
Step 3: Show that there exists a nonnegative function V satisfying the
conditions of the theorem if the process Y (t) = (X(t), α(t)) is positive
recurrent with respect to the domain U = D×{`}. Then Ex,i σD < ∞ for all
(x, i) ∈ D c ×M and consequently equation (3.43) and Harnack’s inequality
for L-elliptic system of equations [3, 25] imply that the bounded monotone
increasing sequence un (x, i) converges uniformly on every compact subset
of Dc × M. Moreover, its limit u(x, i) satisfies the equation Lu(x, i) = −1
for each i ∈ M. Therefore the function V (x, i) := u(x, i) satisfies equation
(3.38). This completes the proof of the theorem. 2
Theorem 3.26. A necessary and sufficient condition for positive recur-
rence with respect to a domain U = D × {`} ⊂ Rr × M is: For each i ∈ M,
there exists a nonnegative function V (·, i) : D c 7→ R such that V (·, i) is
twice continuously differentiable and that for some γ > 0,
Proof. Necessity: This part follows immediately from the necessity of The-
orem 3.25 with γ = −1.
88 3. Recurrence
(n) (n)
Ex,i V (X(σD (t)), α(σD (t))) − V (x, i)
Z σD
(n)
(t)
(n)
= Ex,i LV (X(s), α(s))ds ≤ −γEx,i σD (t).
0
(n)
Hence we have by the nonnegativity of the function V that Ex,i σD (t) ≤
V (x, i)/γ. Meanwhile, the regularity of the process Y (t) = (X(t), α(t)) im-
(n)
plies that σD (t) → σD (t) a.s. as n → ∞, where σD (t) = σD ∧ t. Therefore
Fatou’s lemma leads to Ex,i σD (t) ≤ V (x, i)/γ. Moreover, from the proof
of Theorem 3.25, σD (t) → σD a.s. as t → ∞. Thus we obtain
1
Ex,i σD ≤ V (x, i)
γ
by applying Fatou’s lemma again. Then Lemma 3.10 implies that Ex,i σU =
Ex,i σD×{`} < ∞. Because (x, i) ∈ D c × M is arbitrary, we conclude that
Y (t) is positive recurrent with respect to U . This completes the proof of
the theorem. 2
Example 3.27. Let us continue our discussion from Example 1.1. Con-
sider (1.1) and assume that for each α ∈ M = {1, 2, . . . , m} and i, j =
1, 2, . . . , n with j 6= i, aii (α) > 0 and aij (α) ≥ 0. Then for each i = 1, . . . , n
and α ∈ M, we have
n
X
−aii (α)x2i + ri (α) + aji (α) xi − bi (α)
j=1
Pn 2
ri (α) + j=1 aji (α)
≤ b i (α).
− bi (α) := K
4aii (α)
Denote
Denote
ρ := max {ρi (α), i = 1, . . . , n, α = 1, . . . , m} . (3.45)
Then the solution x(t) to (1.1) is positive recurrent with respect to the
domain
Eρ := x ∈ Rn+ : 0 < xi < ρ, i = 1, 2, . . . , n .
We refer the reader to [189] for a verbatim proof.
EV (X(τ1 ), α(τ1 ))
P( sup V (X(t), α(t)) ≥ γ) ≤ , for γ > 0,
τ1 ≤t≤σ1 γ
EV (X(τ1 ), α(τ1 ))
E(σ1 − τ1 ) ≤ ,
γ
(3.46)
where γ is as given in Theorem 3.26. The idea can be continued in the
following way for k ≥ 1. Define
and
σk+1 = min{t ≥ τk+1 : X(t) ∈ D0 }.
Then we can obtain estimates of E(σk+1 − τk+1 ), which gives us the ex-
pected difference of (k + 1)st return time and exit time.
Let
m0 d
1X 0
X
0
β := −πµ = − πi b(i) + b (i) + σj (i)σj (i) .
2 i=1 j=1
Xd
γ
LV (x, i) = (1 − γci ) x0 σj0 (i)|x|γ−2 Iσj (i)x
2 j=1
+x0 σj0 (i)(γ − 2)|x|γ−4 xx0 σj (i)x
+(1 − γci )γ|x|γ−2 x0 b(i)x + o(|x|γ )
X
− qij (x)|x|γ γ(cj − ci )
j6=i !
1 X d
x0 σj0 (i)σj (i)x (x0 σj0 (i)x)2
γ
= γ(1 − γci )|x| + (γ − 2)
2
j=1
|x|2 |x|4
x0 b(i)x X cj − c i
+ − q ij (x) + o(1) ,
|x|2 1 − γci
j6=i
(3.49)
with o(1) → 0 as |x| → ∞. Note that
d
x0 b(i)x 1 X x0 σj0 (i)σj (i)x
+
|x|2 2 j=1 |x|2
d (3.50)
1 X
0 0
≤ λmax b(i) + b (i) + σj (i)σj (i) = µi .
2 j=1
b(x, i) x
= bi + o(1),
|x| |x|
σ(x, i) x (3.52)
= σi + o(1),
|x| |x|
(a) V (x, i) ≥ 0 for all (x, i) ∈ D c × M, and for some positive constant k,
L(V − Rn W )(x, i) ≤ k,
(V − Rn W )(x, i)|x∈∂D ≥ 0,
(V − Rn W )(x, i)|x∈Γn ≥ 0.
That is,
1
EσD ∧ βn ≥ [Rn W (x, α) − V (x, α)].
k
Therefore, we have from (3.54) that
EσD ∧ βn → ∞, as n → ∞.
Meanwhile, note that σD ≥ σD ∧ βn . We have
EσD = ∞.
Therefore the process (X(t), α(t)) is not positive recurrent by virtue of
Theorem 3.12. 2
3.5 Examples
In this section, we provide several examples to illustrate the results obtained
thus far.
Example 3.33. Suppose that for each x ∈ Rr and each i ∈ M, there exist
positive constants c and γ such that for all x with |x| ≥ c,
x
b0 (x, i) < −γ; (3.58)
|x|
where |x| denotes the norm of x. That is, the drifts are pointed inward.
Then the process Y (t) = (X(t), α(t)) is positive recurrent.
3.5 Examples 95
First note that (3.5) implies that for all x with |x| ≥ r/(γκ1 ), we have
r
X r
X
tr(a(x, i)) = e0j a(x, i)ej ≤ κ−1
1 ≤ γ|x|.
j=1 j=1
By Theorem 3.12, it is enough to prove that the process Y (t) = (X(t), α(t))
is positive recurrent with respect to the domain U := {|x| < %} × {`}
for some ` ∈ M, where % := max{c, r/(γκ1 )}. To this end, consider the
function
1 0
V (x, i) = x x, for each i ∈ M and for all |x| ≥ %.
2
For each i ∈ M, ∇V (·, i) = x and ∇2 V (·, i) = I, where I is the r×r identity
matrix. Thus by the definition of L, we have for all (x, i) ∈ {|x| ≥ %} × M
that
1 x
LV (x, i) = tr(a(x, i)) + b0 (x, i) |x|
2 |x|
1
< γ|x| − γ|x|
2
1 1
= − γ|x| ≤ − κ%.
2 2
Then the conclusion follows from Theorem 3.26 immediately.
Remark 3.34. Suppose that the diffusion component X(t) of the process
Y (t) = (X(t), α(t)) is one-dimensional and that there exist constants c0 > 0
and c1 > 0 such that for each i ∈ M,
< −c , for x > c ,
1 0
b(x, i) (3.59)
>c , for x < −c .
1 0
Then the process Y (t) = (X(t), α(t)) is positive recurrent. In fact, the
conclusion follows immediately if we observe that (3.59) satisfies (3.58).
Alternatively, we can verify this directly by defining the Liapunov function
V (x, i) = |x| for each i ∈ M.
Example 3.35. To illustrate the utility of Theorem 3.26, consider a real-
valued process
and
b(x, 1) = −x, σ(x, 1) = 1, b(x, 2) = x, σ(x, 2) = 1.
Thus (3.60) can be regarded as the result of the following two diffusions:
dX(t) = −X(t)dt + dw(t), (3.61)
and
dX(t) = X(t)dt + dw(t), (3.62)
switching back and forth from one to the other according to the movement
of α(t).
Note that (3.61) is positive recurrent whereas (3.62) is a transient diffu-
sion process. But, the switching diffusion (3.60) is positive recurrent. We
verify these as follows. Consider the Liapunov function V (x, 1) = |x|. Let
L1 be the operator associated with (3.61). Then we have for all |x| ≥ 1,
L1 V (x, 1) = −x sign(x) = −|x| ≤ −1 < 0. It follows from [83, Theorem
3.7.3] that (3.61) is positive recurrent. Recall that the real-valued diffusion
process
dX(t) = b(X(t))dt + σ(X(t))dw(t)
with σ(x) 6= 0 for all x ∈ R, is recurrent if and only if
Z x Z u
b(z)
exp −2 2
dz du → ±∞ (3.63)
0 0 σ (z)
as x → ±∞; see [83, p. 105]. Direct computation shows that (3.62) fails to
satisfy this condition and hence is transient.
Next, we use Theorem 3.26 to demonstrate that the switching diffusion
(3.60) is positive recurrent for appropriate Q. Consider Liapunov functions
7
V (x, 1) = |x|, V (x, 2) = |x|.
3
We have
1 1 7 2 2
LV (x, 1) = −x · sign x + + cos x − 1 |x| ≤ − |x| ≤ − ,
3 4 3 9 9
7 7 1 7 1 1
LV (x, 2) = x · sign x + + sin x 1− |x| ≤ − |x| ≤ − ,
3 3 2 3 9 9
for all |x| ≥ 1. Thus the switching diffusion (3.60) is positive recurrent by
Theorem 3.26.
Example 3.36. To illustrate the result of Corollary 3.29, we consider a
real-valued process given by
dX(t) = b(X(t), α(t))dt + σ(X(t), α(t))dw(t)
P{α(t + ∆) = j|α(t) = i, X(s), α(s), s ≤ t} = q (X(t))∆ + o(∆),
ij
(3.64)
3.5 Examples 97
with the following specifications. The jump component α(t) has three states
and is generated by
2 + |x| 1 + 3x2 2 + |x| 1 + 3x2
− 1 + x2 − 2 + x2 1 + x2 2 + x2
sin x sin x cos x cos x
Q(x)= 1− − − 3 2 + ,
1 + x2 1 + x2 2 + |x| 2 + |x|
1 + cos2 x cos2 x + 1
2 − − 2
2 + x2 2 + x2
and the drift and diffusion coefficients are given by
3 x
b(x, 1) = 3x − 1, b(x, 2) = x + 1, b(x, 3) = −x + ,
2 1 + |x|
p p p
σ(x, 1) = 3 + x2 , σ(x, 2) = 2 − sin x + 2x2 , σ(x, 3) = 3 + 4 + x2 .
Hence associated with (3.64), there are three diffusions
p
dX(t) = (3X(t) − 1)dt + 3 + X 2 (t)dw(t), (3.65)
3 p
dX(t) = X(t) + 1 dt + 2 − sin(X(t)) + 2X 2 (t)dw(t), (3.66)
2
X(t) p
dX(t) = − X(t) dt + 3 + 4 + X 2 (t) dw(t), (3.67)
1 + |X(t)|
switching back and forth from one to another according to the movement
of the jump component α(t).
Note that (3.67) is positive recurrent, whereas (3.65) and (3.66) are tran-
sient diffusions. But due to the stabilization effect of α(t), the switching
diffusion (3.64) is positive recurrent. We verify these as follows. Consider
the Liapunov function V (x) = |x|γ with 0 < γ < 1 sufficiently small; let L1
be the operator associated with the third equation in (3.65). Detailed com-
putation shows that for all |x| ≥ 1, we have L1 V (x) ≤ − 21 γ < 0. Thus it
follows from [83, Theorem 3.7.3] that (3.67) is positive recurrent. Detailed
computations show that (3.65) and (3.66) fail to satisfy (3.63) and hence
these diffusions are transient.
Next we use Corollary 3.29 to demonstrate that the switching diffusion
(3.64) is positive recurrent. In fact, it is readily seen that as x → ∞, the
constants b(i) and σ 2 (i), i = 1, 2, 3, as in (3.52) are given by
3
b(1) = 3, b(2) = , b(3) = −1,
2 (3.68)
σ 2 (1) = 1, σ 2 (2) = 2, σ 2 (3) = 1.
In addition, as |x| → ∞, Q(x) tends to
−3 0 3
b
Q = 1 −3 2 .
0 2 −2
98 3. Recurrence
e
we obtain the stationary distribution π associated with Q
2 6 9
π= , , . (3.69)
17 17 17
Thus Corollary 3.29 implies that (3.64) is positive recurrent; see the sample
path demonstrated in Figure 3.1.
35
X(t)
30 α(t)
25
20
15
X(t) and α(t)
10
−5
−10
−15
0 2 4 6 8 10
t
FIGURE 3.1. Sample path of switching diffusion (3.64) with initial condition
(x, α) = (3, 1).
with (X(t), α(t)) ∈ R2 × {1, 2}. The discrete component α(t) is generated
by
2(1 + cos(x21 )) 2(1 + cos(x21 ))
−3 + 3 + x2 + x2 3−
3 + x21 + x22
Q(x1 , x2 ) =
1
1 2
1
,
1+ p −1 − p
1 + x21 + x22 2
1 + x 1 + x2 2
and
−x1 + 2x2
b(x1 , x2 , 1) = ,
2x2
p
3 + sin x1 + 2 + x22 0
σ(x1 , x2 , 1) = p ,
0 2 − cos x2 + 1 + x21
−3x1 − x2
b(x1 , x2 , 2) = ,
x1 − 2x2
1 1
σ(x1 , x2 , 2) = p .
0 10 + 3 + x21
Associated with the regime-switching diffusion (3.70), there are two diffu-
sions
dX(t) = b(X(t), 1)dt + σ(X(t), 1)dw(t), (3.71)
and
dX(t) = b(X(t), 2)dt + σ(X(t), 2)dw(t), (3.72)
switching back and forth from one to another according to the movement of
the jump component α(t), where w(t) = (w1 (t), w2 (t))0 is a two-dimensional
standard Brownian motion. By selecting appropriate Liapunov functions as
in [83, Section 3.8], or using the criteria in [15], we can verify that (3.71)
is transient and (3.72) is positive recurrent.
Next we use Theorem 3.28 to show that the switching diffusion (3.70)
is positive recurrent owing to the presence of the stabilizing effect of the
discrete component α(t). Note that the matrices b(i) and σj (i), i, j = 1, 2,
and Q b as in condition (A3.2) are
−1 2 −3 −1
b(1) = , b(2) = , σ1 (2) = 0,
0 2 1 −2
0 1 −3 3
σ1 (1) = σ2 (2) = σ20 (1) = b=
, and Q .
0 0 1 −1
100 3. Recurrence
30
20
10
X (t)
0
2
−10
−20
−30
−25 −20 −15 −10 −5 0 5 10 15
X1(t)
FIGURE 3.2. Phase portrait of switching diffusion (3.70) with initial condition
(x, α), where x = [2.5, 2.5]0 and α = 1.
Proof of Lemma 3.7. It suffices to prove the lemma when E∪∂E ⊂ D and
∂E is sufficiently smooth. Fix any (x, i) ∈ E c × M. Let G ⊂ Rr be an open
3.6 Proofs of Several Results 101
15
X1(t)
10 α(t)
0
X (t) and α(t)
−5
1
−10
−15
−20
−25
0 2 4 6 8 10
t
30
X (t)
2
α(t)
20
10
X (t) and α(t)
0
2
−10
−20
−30
0 2 4 6 8 10
t
FIGURE 3.3. Componentwise sample path of switching diffusion (3.70) with ini-
tial condition (x, α), where x = [2.5, 2.5]0 and α = 1.
102 3. Recurrence
and for n = 1, 2, . . .,
It follows from (3.17) and Theorem 3.2 that ςn < ∞ a.s. Px,i for n =
1, 2, . . . Let H := G − E and define u(x, i) := Px,i {X(τH ) ∈ ∂E}. Note
that u(x, j)|x∈∂E = 1 and u(x, j)|x∈∂G = 0 for all j ∈ M. Therefore, it
follows that
m0 Z
X
u(x, i) = Px,i {(X(τH ), α(τH )) ∈ (dy × {j})} u(y, j)
j=1 ∂E
m0 Z
X
+ Px,i {(X(τH ), α(τH )) ∈ (dy × {j})} u(y, j)
j=1 ∂G
and for n = 1, 2, . . .,
Note that the event Ac0 implies that X(τH ) = X(ς1 ) ∈ ∂G. Hence we have
from (3.75) that
Then it follows from the strong Markov property and (3.75) that
( n )
\
Px,i Ak ≤ (1 − δ1 )n+1 .
c
(3.78)
k=0
3.6 Proofs of Several Results 103
Thus, we have
Therefore, we have
∞
X
Ex,i τ Ec = Ex,i σE I{0<σE <ς1 } + Ex,i σE I{ς2n <σE <ς2n+1 }
n=1
∞
X
≤ Px,i [0 < σE < ς1 ]Ex,i ς1 + Px,i [ς2n < σE < ς2n+1 ]Ex,i ς2n+1
n=1
∞
X
≤ (1 − δ1 )n Ex,i ς2n+1 ,
n=0
≤ M1 + M2 + M3 ≤ 2M,
Redefine
ς1 := inf{t ≥ 0 : X(t) ∈ ∂B}, (3.80)
and for n = 1, 2, . . .,
Note that equation (3.19), Theorem 3.2, and Lemma 3.7 imply that ςn < ∞
a.s. Px,i . Set
n o
u(x, i) := Px,i σB×{`} < τB1 .
Redefine
A0 := {α(t) = `, for some t ∈ [0, ς2 )}, (3.83)
and for n = 1, 2, . . .,
Using almost the same argument as in the proof of Lemma 3.7, we obtain
that
( n )
\
c
Px,i (A0 ) ≤ 1 − δ2 , and Px,i Ak ≤ (1 − δ2 )n+1 .
c
(3.85)
k=0
3.6 Proofs of Several Results 105
Thus, we have
As a result,
It follows that
Following almost the same argument as that for the proof of Lemma 3.8, we
can show that Ex,i ς2n ≤ nM for some positive constant M . Consequently,
∞
X
Ex,i σD×{`} ≤ (1 − δ2 )n (n + 1)M < ∞.
n=0
(We use the convention that inf {∅} = ∞.) Then by virtue of the strong
Markov property, we can show
n
sup Px,` {ςn < ∞} ≤ sup Px,` {ς1 < ∞} . (3.86)
x∈B(x0 ,ε) x∈B(x0 ,ε)
Hence we have
∞
X Z ςn+1
R0 (x0 , `, U ) = Ex0 ,` I{ςn <∞} IU (X(t), α(t))dt
n=0 ςn
X∞ Z ς1
= Ex0 ,` I{ςn <∞} EX(ςn ),α(ςn ) IU (X(t), α(t))dt .
n=0 0
Note that
Z ς1 Z T Z ς1
IU (X(t), α(t))dt = IU (X(t), α(t))dt + IU (X(t), α(t))dt ≤ T.
0 0 T
T
where σB(x 0 ,ε)×{`}
:= ς1 = inf {t > T : (X(t), α(t)) ∈ B(x0 , ε) × {`}}. Be-
cause B(x0 , ε) ⊂ B,
n o n o
T T
Px,` σB(x 0 ,ε)×{`}
< ∞ ≤ P x,` σ B×{`} < ∞ ,
n o
T
and supx∈B(x0 ,ε) Px,` σB×{`} < ∞ = 1. Finally, because (X(t), α(t)) is
strong Feller, we obtain by letting ε → 0 that
n o
T
Px0 ,` σB×{`} <∞ =1
3.6 Proofs of Several Results 107
for any T > 0. Thus (3.30) follows. This completes the proof of the propo-
sition. 2
Proof of Theorem 3.20. Sufficiency. This is clear by a contradiction
argument.
Necessity. Assume the process (X(t), α(t)) is transient. Fix any (x, α) ∈
Rr × M. Let ρ ∈ R be sufficiently large such that ρ > 1 ∨ |x|. The process
(X(t), α(t)) is transient, therefore it is transient with respect to the “cylin-
der” B(0, ρ) × M. Thus there exists some (y0 , j0 ) ∈ (Rr − B(0, ρ)) × M
such that
Py0 ,j0 σB(0,ρ) < ∞ < 1. (3.87)
Then using the standard argument (see [15, Theorem 3.2]), we can show
that n o
Px,α lim inf |X(t)| > ρ − 1 = 1.
t→∞
Px,α {|X(t)| → ∞ as t → ∞} = 1,
as desired. 2
Proof of Proposition 3.21. By virtue of Theorem 3.18, it is enough to
show that the point (x0 , `) is transient in the sense that there exists some
ε0 > 0 and a finite time T0 > 0 such that
Because the process (X(t), α(t)) is strong Feller by virtue of Theorem 2.24,
we conclude that there exists a neighborhood E ⊂ D of x0 such that
Z t
inf Ey,` IU (X(t), α(t))dt > 0.
y∈E 0
Hence we have
δ := inf R0 (x, `, U ) > 0. (3.90)
x∈E
108 3. Recurrence
For any ε > 0, in view of the assumption R0 (x0 , `, U ) < ∞, we can choose
some Te0 > 0 such that
Z ∞
Ex0 ,` IU (X(t), α(t))dt < δε
Te0
and hence n e o
T0
Px0 ,` σE×{`} < ∞ < ε.
That is,
n o
T
lim Px0 ,` σE×{`} <∞
T →∞ n o
T
= Px0 ,` σE×{`} < ∞, for every T > 0 = 0.
3.7 Notes
Under general conditions, necessary and sufficient conditions for recurrence,
nonrecurrence, and positive recurrence have been studied in this chap-
ter. We refer the reader to Chapter 2; see also Skorohod [150] for related
3.7 Notes 109
4.1 Introduction
Continuing with the study of basic properties of switching-diffusion pro-
cesses, this chapter is concerned with ergodicity. Many applications in con-
trol and optimization require minimizing an expected cost of certain objec-
tive functions. Treating average cost per unit time problems, we often wish
to “replace” the time-dependent instantaneous measure by a steady-state
(or ergodic) measure. Thus we face the following questions: Do the sys-
tems possess an ergodic property? Under what conditions do the systems
have the desired ergodicity? Significant effort has been devoted to approxi-
mating such expected values by replacing the instantaneous measures with
stationary measures when the time horizon is long enough. To justify such
a replacement, ergodicity is needed. For diffusion processes, we refer the
readers to, for example, [10, 103] among others for the study of ergodic
control problems. In what follows, we study ergodicity and reveal the main
features of the ergodic measures. We carry out our study on ergodicity by
constructing cycles and using induced discrete-time Markov chains.
We consider the two-component process Y (t) = (X(t), α(t)) as in Chap-
ter 3. Let w(t) be a d-dimensional standard Brownian motion, b(·, ·) : Rr ×
M 7→ Rr , and σ(·, ·) : Rr × M 7→ Rr×d satisfying σ(x, i)σ 0 (x, i) = a(x, i).
For t ≥ 0, let X(t) ∈ Rr and α(t) ∈ M such that
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 111
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_4,
© Springer Science + Business Media, LLC 2010
112 4. Ergodicity
and
P{α(t + ∆) = j|α(t) = i, X(s), α(s), s ≤ t}
(4.2)
= qij (X(t))∆ + o(∆), i 6= j.
The rest of the chapter is arranged as follows. Section 4.2 begins with
the discussion of ergodicity. The analysis is carried out by using cycles and
induced Markov chains in discrete time. Then the desired result is obtained
together with the representation of the stationary density. Section 4.3 takes
up the issue of making a switching-diffusion process ergodic by means of
feedback controls. Section 4.4 discusses some ramifications, and Section 4.5
obtains asymptotic normality when the continuous component belongs to
a compact set. Section 4.6 presents some further remarks.
4.2 Ergodicity
In this section, we study the ergodic properties of the process Y (t) =
(X(t), α(t)) under the assumption that the process is positive recurrent
with respect to some bounded domain U = E × {`}, where E ⊂ Rr and
` ∈ M are fixed throughout this section. We also assume that the boundary
∂E of E is sufficiently smooth. Let the operator L satisfy (A3.1). Then
it follows from Theorem 3.12 that the process is positive recurrent with
respect to any nonempty open set.
Let D ⊂ Rr be a bounded open set with sufficiently smooth boundary
∂D such that E ∪ ∂E ⊂ D. Let ς0 = 0 and define the stopping times
ς1 , ς2 , . . . inductively as follows: ς2n+1 is the first time after ς2n at which
the process Y (t) = (X(t), α(t)) reaches the set ∂E × {`} and ς2n+2 is the
first time after ς2n+1 at which the path reaches the set ∂D × {`}. Now we
can divide an arbitrary sample path of the process Y (t) = (X(t), α(t)) into
cycles:
[ς0 , ς2 ), [ς2 , ς4 ), . . . , [ς2n , ς2n+2 ), . . . (4.3)
Figure 4.1 presents a demonstration of such cycles when the discrete com-
ponent α(·) has three states.
The process Y (t) = (X(t), α(t)) is positive recurrent with respect to
E × {`} and hence positive recurrent with respect to D × {`} by Theo-
rem 3.12. It follows that all the stopping times ς0 < ς1 < ς2 < ς3 < ς4 < · · ·
are finite almost surely (a.s.). Because the process Y (t) = (X(t), α(t)) is
positive recurrent, we may assume without loss of generality that Y (0) =
(X(0), α(0)) = (x, `) ∈ ∂D × {`}. It follows from the strong Markov prop-
erty of the process Y (t) = (X(t), α(t)) that the sequence {Yn } is a Markov
chain on ∂D × {`}, where Yn = Y (ς2n ) = (Xn , `), n = 0, 1, . . . Let Pe (x, A)
denote the one-step transition probabilities of this Markov chain; that is,
D E State 3
6
∂D ∂E
D E State 2
∂D ∂E
ς2
ς3
D E State 1
ς1
∂D ?∂E
ς4 ς0
FIGURE 4.1. A sample path of the process Y (t) = (X(t), α(t)) when m0 = 3.
Using the harmonic measure defined and Lemmas 2.2 and 2.3 in Chen and
Zhao [25], relating the kernel and surface area (similar to the solution for
the diffusion process without switching in the form of double-layer poten-
tial given in the first displayed equation in [83, p. 97]) and the harmonic
measure, we can finish the proof of this lemma analogously to that of [83,
Lemma 4.4.1]. The details are omitted. 2
Remark 4.2. Note that
where (X 0,X(0),α(0) (u), α0,X(0),α(0) (u)) denotes the sample path of the pro-
cess (X(·), α(·)) with initial point (X(0), α(0)) at time t = 0, and a similar
definition for (X s,X(s),α(s) (t), αs,X(s),α(s) (t)). When no confusion arises, we
simply write
Proof. Recall that the cycles were defined in (4.3). Let A ∈ B(Rr ) and
i ∈ M. Denote by τ A×{i} the time spent by the path of Y (t) = (X(t), α(t))
in the set (A × {i}) during the first cycle. Set
Z
νb(A, i) := m(dx)Ex τ A×{i} , (4.8)
∂D
Now we claim that for any bounded and continuous function f (·, ·),
m0 Z
X Z Z ς2
f (y, j)b
ν (dy, j) = m(dx)Ex f (X(t), α(t))dt (4.10)
j=1 Rr ∂D 0
holds. In fact, if f is an indicator function, that is, f (y, j)IA×{i} (y, j) for
some A ∈ B(Rr ) and i ∈ M, then from (4.8),
m0 Z
X
IA×{i} (y, j)b
ν (dy, j)
Rr
j=1 Z
= νb(A, i) = m(dx)Ex τ A×{i}
Z ∂D Z ς
2
n
X
f (y, j) = cp IUp (y, j),
p=1
m0 Z
X
Ex,i f (X(t), α(t))b ν (dx, i)
Rr
i=1 Z Z ς2
= m(dx)Ex EX(s),α(s) f (X(t + s), α(t + s))ds
Z∂D Z 0 ς2
= m(dx)Ex f (X(t + s), α(t + s))ds
Z∂D Z 0 ς2
= m(dx)Ex f (X(u), α(u))du
Z∂D Z
0
t+ς2
+ m(dx)Ex f (X(u), α(u))du
Z∂D Zς2t
− m(dx)Ex f (X(u), α(u))du.
∂D 0
Z ς2 +t
g(x) = Ex f (X(u), α(u))du,
ς2
116 4. Ergodicity
we obtain
Z Z ς2 +t
m(dx)Ex f (X(u), α(u))du
∂D Z ς2 Z ς2 +t
= m(dx)Ex EX1 ,` f (X(u + ς2 ), α(u + ς2 ))du
Z∂D Z t ς2
Note that in the above deduction, we used equation (4.6) again. Therefore,
the above two equations and (4.10) yield that
m0 Z
X m0 Z
X
Ex,i f (X(t), α(t))b
ν (dx, i) = f (x, i)b
ν (dx, i).
i=1 Rr i=1 Rr
Then Z !
T
1
Px,i lim f (X(t), α(t))dt = f = 1, (4.13)
T →∞ T 0
Then it follows from equation (4.15) that {ηn } is a strictly stationary se-
quence. Also from equations (4.8) and (4.10), we have
m0 Z
X
Eηn = f (x, i)b
ν (dx, i), (4.17)
i=1 Rr
RT
Then we can decompose 0
f (X(t), α(t))dt into
Z T υ(T )
X Z T
f (X(t), α(t))dt = ηn + f (X(t), α(t))dt, (4.18)
0 n=0 ς2υ(T )
Note that the positive recurrence of the process Y (t) = (X(t), α(t)) implies
that υ(T ) → ∞ as T → ∞. Clearly, υ(T )/(υ(T ) + 1) → 1 almost surely as
T → ∞. Thus, it follows from (4.20) that as T → ∞,
ς2υ(T )
ς2υ(T ) υ(T )
υ(T )
= ς → 1 a.s. (4.21)
ς2υ(T )+2 2υ(T )+2 υ(T ) + 1
υ(T ) + 1
118 4. Ergodicity
υ(T ) 1
→ P m0 a.s. as T → ∞. (4.23)
ς2υ(T ) b(Rr , i)
i=1 ν
by the definition of µ(·, ·). Thus, equation (4.13) holds. This proves (4.13)
if the initial distribution is (4.15).
Let (x, i) ∈ Rr × M. Because the process Y (t) = (X(t), α(t)) is positive
recurrent with respect to the domain D × {`}, we have
h Z i
1 T
Px,i lim f (X(t), α(t))dt = f
T →∞ T 0
h Z i
1 T
= Px,i lim f (X(t), α(t))dt = f .
T →∞ T ς
2
where the last line above follows from the use of the invariant distribution.
The following illustrates that starting from an arbitrary point (x, i) and
arbitrary initial distribution is asymptotically equivalent to starting with
the initial distribution being the stationary distribution. Therefore, (4.13)
holds for all (x, i) ∈ Rr × M. This completes the proof of the theorem. 2
As a consequence of Theorem 4.4, we obtain the following corollary.
Corollary 4.5. Let the assumptions of Theorem 4.4 be satisfied and let
u(t, x, i) be the solution of the Cauchy problem
∂u(t, x, i) = Lu(t, x, i), t > 0, (x, i) ∈ Rr × M,
∂t (4.24)
u(0, x, i) = f (x, i), (x, i) ∈ Rr × M,
with respect to the probability Px,i . Then equation (4.25) follows from the
dominated convergence theorem. 2
X(0) = x, α(0) = α.
and
−2 3 2 −3
B(1) = , B(2) = .
1 −1 1 3
Thus associated with the regime-switching diffusion (4.32), there are two
diffusions
dX(t) = b(X(t), 1)dt + σ(X(t), 1)dw(t), (4.34)
and
dX(t) = b(X(t), 2)dt + σ(X(t), 2)dw(t) (4.35)
switching back and forth from one to another according to the movement
of the discrete component α(t), where w(t) = (w1 (t), w2 (t))0 is a two-
dimensional standard Brownian motion. Assume that the system is observ-
able when the discrete component α(·) is in state 2. Detailed calculations
using the methods in [83, Section 3.8] or [15] allow us to verify that both
(4.34) and (4.35) are transient diffusions; see also the phase portraits in
Figure 4.2 (a) and (b).
4.3 Feedback Controls for Weak Stabilization 123
8
x 10
14
12
10
8
2
x
0
−5 0 5 10 15 20
x1 x 10
8
12
x 10
4
−2
−4
2
x
−6
−8
−10
−12
−14
−2 0 2 4 6 8 10 12 14 16
x1 12
x 10
FIGURE 4.2. Phase portraits of transient diffusions in (4.34) and (4.35) with
initial condition x, where x = [−5, 3]0 .
124 4. Ergodicity
Next we use Theorem 3.24 to verify that the switching diffusion (4.32)
is transient. To this end, we compute the matrices b(i), σj (i) for i, j = 1, 2
and Q b as in condition (A3.2):
√
1 1 0 2 0 0
b(1) = , σ1 (1) = , σ2 (1) = ,
0 2 0 0 0 1
2 1 0 1 0 0
b(2) = , σ1 (2) = , σ2 (2) = ,
−1 3 0 0 1 0
and
−2 2
b=
Q .
1 −1
Thus the stationary distribution is π = (1/3, 2/3) and
Hence Theorem 3.24 implies that the switching diffusion (4.32) is transient.
Figure 4.3 (a) confirms our analysis.
By our assumption, the switching diffusion is observable when the dis-
crete component α(·) is in state 2. Note that
4 −2
B(2) + B 0 (2) =
−2 6
Figure 4.3(b) demonstrates the phase portrait of (4.33) under the feedback
control (4.36). We also demonstrate the component-wise sample path of
(4.33) under the feedback control (4.36) in Figure 4.4(a) and (b).
4.4 Ramifications
Remark on a Tightness Result
Under positive recurrence, we may obtain tightness (or boundedness in the
sense of in probability) of the underlying process. Suppose that (X(t), α(t))
is positive recurrent. We can use the result in Chapter 3 about path excur-
sion (3.46) to prove that for any compact set D (the closure of the open
set D), the set
[
{(X(t), α(t)) : t ≥ 0, X(0) = x, α(0) = α}
x∈D
is tight (or bounded in probability). The idea is along the line of the diffu-
sion counterpart; see [102].
Suppose that for each i ∈ M, there is a Liapunov function V (·, i) such
that minx V (x, i) = 0, V (x, i) → ∞ as |x| → ∞. Let a1 > a0 > 0 and
X(0) = x and α(0) = α. Using the argument in Chapter 3, because recur-
rence is independent of the chosen set, we can work with a fixed ` ∈ M,
and consider the sets
Ba0 = {x ∈ Rr : V (x, `) ≤ a0 },
Ba1 = {x ∈ Rr : V (x, `) ≤ a1 }.
Then Ba0 will be visited by the switching diffusion infinitely often a.s.
Because the switching is positive recurrent, Theorem 3.26 implies that there
is a κ > 0 such that LV (x, i) ≤ −κ for all (x, i) ∈ Bac0 .
Define a sequence of stopping times recursively as
21
x 10
1
−1
X (t)
−2
2
−3
−4
−5
−8 −7 −6 −5 −4 −3 −2 −1 0 1
X1(t) 21
x 10
2
X (t)
2
−2
−4
−6
−25 −20 −15 −10 −5 0 5 10
X1(t)
FIGURE 4.3. Phase portraits of switching diffusion (4.32) and its controlled
system (4.33) under feedback control law (4.36) with initial condition (x, α),
where x = [−5, 3]0 and α = 2.
4.4 Ramifications 127
10
X1(t)
α(t)
5
0
X (t) and α(t)
−5
−10
1
−15
−20
−25
0 2 4 6 8 10
t
8
X (t)
2
α(t)
6
4
X (t) and α(t)
0
2
−2
−4
−6
0 2 4 6 8 10
t
FIGURE 4.4. Componentwise sample path of (4.33) under feedback control law
(4.36) with initial condition (x, α), where x = [−5, 3]0 and α = 2.
128 4. Ergodicity
Occupation Measures
To illustrate another utility of Theorem 4.4, take f (x, i) = I{B×J} (x, i),
the indicator function of the set B × J, where B ⊂ Rr and J ⊂ M. Then
Theorem 4.4 becomes a result regarding the occupation measure. In fact,
we have
Z XZ
1 T
I{B×J} (X(t), α(t))dt → µ(x, i)dx a.s. as T → ∞.
T 0 Bi∈J
Stochastic Approximation
Consider a parameter optimization problem. We wish to find θ∗ , a vector-
valued parameter so that the cost function
Z
1 T b
J(θ) = lim E J(θ, Y (t))dt
T →∞ T 0
for some positive constant C and κ3 given by (4.37), where Ef (X(t), α(t))
is the expectation with respect to the stationary measure
m0 Z
X def
Ef (X(t), α(t)) = f (x, i)µ(x, i)dx = fav . (4.40)
i=1 S
130 4. Ergodicity
m0 Z
X
|f (x, i)|dx < ∞.
i=1 S
Moreover,
≤ C exp(−κ3 (t − s)).
Lemma 4.11. Assume the conditions of Lemma 4.10. Then the following
assertion holds:
Proof. Assume without loss of generality that s ≤ t. Note that from (4.40),
4.5 Asymptotic Distribution 131
by virtue of Lemma 4.10. In the above, from the next to the last line to
the last line, we used C as a generic positive constant whose values may be
different for different appearances. The proof of the lemma is concluded. 2
In what follows, denote
Proof. Define
Z T
1
ζ(T ) = [f (X(t), α(t)) − fav ]dt, and ξ(T ) = √ ζ(T ). (4.45)
0 T
Note that
Z T
1
Eξ(T ) = E √ [f (X(t), α(t)) − fav ]dt
TZ 0
T
1
=√ E[f (X(t), α(t)) − fav ]dt = 0.
T 0
To calculate the asymptotic variance, note that
" Z T #2
1
E √ [f (X(t), α(t)) − fav ]dt
T 0
Z Z
1 T T
=E [f (X(t), α(t)) − fav ][f (X(s), α(s)) − fav ]dsdt
T Z0 Z0
1 T t
= 2E [f (X(t), α(t)) − fav ][f (X(s), α(s)) − fav ]dsdt.
T 0 0
(4.46)
Equation (4.42) and changing variables lead to
Z Z
1 T t
E [f (X(t), α(t)) − fav ][f (X(s), α(s)) − fav ]dsdt
T 0 0
Z Z
1 T t
= ρ(t − s)dsdt
T 0 0
Z TZ T (4.47)
1
= ρ(u)dtdu
T 0 u
Z T
u
= 1− ρ(u)du.
0 T
Choose some 0 < ∆ < 1. Then by virtue of Lemma 4.12, it follows that as
T → ∞,
Z T Z T∆ Z T
u u u
ρ(u)du = ρ(u)du + ρ(u)du
0 T 0 T T ∆ T
Z ∆ Z
T∆ T T T
≤ ρ(u)du + ρ(u)du
T 0 T T∆
→ 0.
Therefore, we have
Z ∞
E[ξ(T )]2 → σav
2
=2 ρ(t)dt as T → ∞. (4.48)
0
4.6 Notes
Many applications of diffusion processes without switching require using
invariant distributions of the underlying processes. One of them is the for-
mulation of two-time-scale diffusions. Dealing with diffusions having fast
and slow components, under the framework of diffusion approximation, it
has been shown that in the limit, the slow component is averaged out with
respect to the invariant measure of the fast component; see Khasminskii
[82] for a weak convergence limit result, and Papanicolaou, Stroock, and
Varadhan [131] for using a martingale problem formulation. In any event, a
crucial step in these references is the use of the invariance measure. In this
chapter, in contrast to the diffusion counterpart, we treat regime-switching
diffusions, where in addition to the continuous component, there are dis-
crete events. Our interest has been devoted to obtaining ergodicity.
For regime-switching diffusions, asymptotic stability for the density of
the two-state random process (X(t), α(t)) was established in [136]; asymp-
totic stability in distribution (or the convergence to the stationary mea-
sures of the switching diffusions) for the process (X(t), α(t)) was obtained
in [6, 183]. Here, we addressed ergodicity for (X(t), α(t)) under different
conditions from those in [6, 136, 183].
Taking into consideration of the many applications in which discrete
events and continuous dynamics are intertwined and the discrete-event
process depends on the continuous state, we allow the discrete compo-
nent α(·) to have x-dependent generator Q(x). Another highlight is that
we obtain the explicit representation of the invariant measure of the process
(X(t), α(t)) by considering certain cylinder sets and by defining cycles ap-
propriately. As a byproduct, we demonstrate a strong law of large numbers
type theorem for positive recurrent regime-switching diffusions. It reveals
that positive recurrence and ergodicity of switching diffusions are equiva-
lent.
In this chapter, we first developed ergodicity for positive recurrent regime-
switching diffusions. Focusing on a compact space, we then obtained asymp-
totic distributions as a consequence of the ergodicity of the process. The
asymptotic distributions are important in treating limit ergodic control
problems as well as in applications of Markov chain Monte Carlo. A crucial
step in the proof is the verification of the centered and scaled process being
a mixing process with exponential mixing rate.
134 4. Ergodicity
5.1 Introduction
As is the case for deterministic dynamic systems or stochastic differential
equations, closed-form solutions for switching diffusions are often difficult
to obtain, and numerical approximation is frequently a viable or possi-
bly the only alternative. Being extremely important, numerical methods
have drawn much attention. To date, a number of works (e.g., [120, 121,
122, 171]) have focused on numerical approximations where the switching
process is independent of the continuous component and is modeled by
a continuous-time Markov chain. In addition to the numerical methods,
approximation to invariant measures and non-Lipschitz data were dealt
with. Nevertheless, it is necessary to be able to handle the coupling and
dependence of the continuous states and discrete events. This chapter is de-
voted to numerical approximation methods for switching diffusions whose
switching component is x-dependent. Section 5.2 presents the setup of the
problem. Section 5.3 suggests numerical algorithms. Section 5.4 establishes
the convergence of the numerical algorithms. Section 5.5 proceeds with a
couple of examples. Section 5.6 gives a few remarks concerning the rates of
convergence of the algorithms and the study on decreasing stepsize algo-
rithms. Finally Section 5.7 concludes the chapter.
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 137
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_5,
© Springer Science + Business Media, LLC 2010
138 5. Numerical Approximation
5.2 Formulation
Let M = {1, . . . , m0 } and consider the hybrid diffusion system
and
P(α(t + ∆) = j|α(t) = i, X(s), α(s), s ≤ t)
(5.2)
= qij (X(t))∆ + o(∆), i 6= j,
where w(·) is an r-dimensional standard Brownian motion, X(t) ∈ Rr ,
b(·, ·) : Rr × M 7→ Rr , σ(·, ·) : Rr × M 7→ Rr×r are appropriate functions
m0 ×m0
satisfying certain regularity conditions, and Q(x)Pm=0 (qij (x)) ∈ R
satisfies that for each x, qij (x) ≥ 0 for i 6= j, j=1 qij (x) = 0 for each
i ∈ M. There is an associated operator for the switching diffusion process
defined as follows. For each i ∈ M and suitable smooth function h(·, i),
define an operator
1
Lh(x, i) = ∇h0 (x, i)b(x, i) + tr[∇2 h(x, i)σ(x, i)σ 0 (x, i)]
m0
2
X (5.3)
+ qij (x)h(x, j).
j=1
Proof. Define U (x) = |x|p and use En to denote the conditional expecta-
tion with respect to the σ-algebra Gn , where Gn was given in (A5.3). Note
that
Taking the expectation on both sides and iterating on the resulting recur-
sion, we obtain
n
X
E|Xn+1 |p ≤ |X0 |p + Kεn + Kε E|Xk |p .
k=0
as desired. 2
Remark 5.3. In view of the estimate above, {Xn : 0 ≤ n ≤ T /ε} is tight
in Rr by means of the well-known Tchebyshev
p inequality. That is, for each
η > 0, there is a Kη satisfying Kη > (1/η) such that
sup E|Xn |2
0≤n≤T /ε
P(|Xn | > Kη ) ≤ ≤ Kη.
Kη2
Proof. First note that by the boundedness and the continuity of Q(·), for
each i ∈ M,
m0
X
E[I{αk+1 =j} − I{αk =i} Gk ]
j=1
X
= E[I{αk+1 =j} Gk ]
j6=i,j∈M
X
= εqij (Xk )I{αk =i} = O(ε)I{αk =i} .
j6=i,j∈M
142 5. Numerical Approximation
= ge(t + s − t)
= ge(s),
and that
Ee
g (s) = O(s).
In the above, we have used the convention that t/ε and (t + s)/ε denote
the integer parts of t/ε and (t + s)/ε, respectively. Inasmuch as
h (t+s)/ε−1
X i
ε ε
E χ (t + s) − χ (t) − [χk+1 − χk ] Ftε = 0,
k=t/ε
The tightness criterion in [102, p. 47] yields that {χε (·)} is tight. Conse-
quently, {αε (·)} is tight. 2
Lemma 5.6. Assume that the conditions of Lemma 5.5 are satisfied. Then
{X ε (·)} is tight in D r ([0, ∞) : Rr ), the space of functions that are right
continuous and have left limits, endowed with the Skorohod topology.
E|X ε (t + s) − X ε (t)|2
(t+s)/ε−1 (t+s)/ε−1 2
X √ X
=Eε b(Xk , αk ) + ε σ(Xk , αk )ξk
k=t/ε k=t/ε
(t+s)/ε−1 (t+s)/ε−1
X X
≤ Kε2 (1 + E|Xk |2 ) + Kε E|σ(Xk , αk )|2 E|ξk |2
k=t/ε k=t/ε
(t+s)/ε−1
X
≤ Kε2 (1 + sup E|Xk |2 )
t/ε≤k≤(t+s)/ε−1
k=t/ε
(t+s)/ε−1
X
+Kε (1 + sup E|Xk |2 )
t/ε≤k≤(t+s)/ε−1
k=t/ε
t+s t
≤O − O(ε) = O(s).
ε ε
(5.11)
In the above, we have used Lemma 5.2 to ensure that
κ
Y h (t+s)/ε−1
X i
E ρι (X ε (tι ), αε (tι )) χε (t+s)−χε (t)− (χk+1 −χk ) = 0. (5.12)
ι=1 k=t/ε
κ
Y
lim E ρι (X ε (tι ), αε (tι ))[χε (t + s) − χε (t)]
ε→0
ι=1
κ
Y
=E e ι ), α
ρι (X(t e(tι ))[χ(t + s) − χ(t)].
ι=1
nε → ∞ as ε → 0 but δε = εnε → 0.
as a base for partition. Then the continuity together with the boundedness
of Q(·) implies that
κ
Y h (t+s)/ε−1
X i
lim E ρι (X ε (tι ), αε (tι )) (χk+1 − χk )
ε→0
ι=1 k=t/ε
κ
Y h (t+s)/ε−1
X i
= lim E ρι (X ε (tι ), αε (tι )) (E(χk+1 Gk ) − χk )
ε→0
ι=1 k=t/ε
κ
Y h (t+s)/ε−1
X lnεX +nε −1 i
= lim E ρι (X ε (tι ), αε (tι )) χk (I + εQ(Xk )− I)
ε→0
ι=1 lnε =t/ε k=lnε
κ
Y h (t+s)/ε−1
X 1
+nε −1
lnεX i
= lim E ρι (X ε (tι ), αε (tι )) δε χk Q(Xlnε ) .
ε→0
ι=1
nε
lnε =t/ε k=lnε
(5.14)
Note that
κ
Y h (t+s)/ε−1
X 1
+nε −1
lnεX i
lim E ρι (X ε (tι ), αε (tι )) δε [χk − χlnε ]Q(Xlnε )
ε→0
ι=1
nε
lnε =t/ε k=lnε
κ
Y
= lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
h (t+s)/ε−1
X 1
lnεX
+nε −1 i
× δε E[χk − χlnε Glnε ]Q(Xlnε )
nε
lnε =t/ε k=lnε
Yκ
= lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
h (t+s)/ε−1
X 1
lnεX
+nε −1 i
× δε Eχlnε [(I + εQ(Xlnε ))k−lnε − I]Q(Xlnε )
nε
lnε =t/ε k=lnε
= 0.
(5.15)
Therefore,
κ
Y h (t+s)/ε−1
X 1
lnεX
+nε −1 i
lim E ρι (X ε (tι ), αε (tι )) δε χk Q(Xlnε )
ε→0
ι=1
nε
lnε =t/ε k=lnε
κ
Y h (t+s)/ε−1
X i
= lim E ρι (X ε (tι ), αε (tι )) δε χlnε Q(Xlnε )
ε→0
ι=1 lnε =t/ε
κ
Y Z t+s
=E e ι ), α
ρι (X(t e(tι )) e
χ(u)Q(X(u))du.
ι=1 t
(5.16)
146 5. Numerical Approximation
Moreover, the limit does not depend on the chosen subsequence. Thus,
κ
Y h Z t+s i
E e e(tι )) χ(t + s) − χ(t) −
ρι (X(tι ), α e
χ(u)Q(X(u))du = 0. (5.17)
ι=1 t
e
e(·) has a generator Q(X(·)).
Therefore, the limit process α
Step 2: For t, s, κ, tι as chosen before, for each bounded and continuous
function ρι (·, i), and for each twice continuously differentiable function with
compact support h(·, i) with i ∈ M, we show that
κ
Y
E e ι ), α
ρι (X(t e + s), α
e(tι ))[h(X(t e(t + s))
ι=1 Z t+s (5.18)
e
− h(X(t), e(t)) −
α e
Lh(X(u), e(u))du] = 0.
α
t
e
is a continuous-time martingale, which in turn implies that (X(·), αe(·)) is
a solution of the martingale problem with operator L defined in (5.3).
To establish the desired result, we work with the sequence (X ε (·), αε (·)).
Again, we use the sequence {nε } as in Step 1. By virtue of the weak con-
vergence and the Skorohod representation, it is readily seen that
κ
Y
E ρι (X ε (tι ), αε (tι ))[h(X ε (t + s), αε (t + s)) − h(X ε (t), αε (t))]
ι=1
κ
Y
→E e ι ), α
ρι (X(t e + s), α
e(tι ))[h(X(t e
e(t + s)) − h(X(t), e(t))]
α
ι=1
(5.19)
as ε → 0. On the other hand, direct calculation shows that
κ
Y
E ρι (X ε (tι ), αε (tι ))[h(X ε (t + s), αε (t + s)) − h(X ε (t), αε (t))]
ι=1
κ
Y
=E ρι (X ε (tι ), αε (tι ))
ι=1
X h
n (t+s)/ε−1
× [h(Xlnε +nε , αlnε +nε ) − h(Xlnε +nε , αlnε )]
lnε =t/ε
io
+[h(Xlnε +nε , αlnε ) − h(Xlnε , αlnε )] .
(5.20)
5.4 Convergence of the Algorithm 147
Step 3: Still use the notation Ξε` defined in (5.13). For the terms on the
last line of (5.20), we have
κ
Y
lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
(t+s)/ε−1
X
× [h(Xlnε +nε , αlnε ) − h(Xlnε , αlnε )]
lnε =t/ε
Yκ
= lim E ρι (X ε (tι ), αε (tι )) (5.21)
ε→0
ι=1
n (t+s)/ε−1 h
X lnεX
+nε −1
× ε∇h0 (Xlnε , αlnε ) b(Xk , αlnε )
lnε =t/ε k=lnε
ε
lnεX
+nε −1 io
+ tr[∇2 h(Xlnε , αlnε )σ(Xk , αlnε )σ 0 (Xk , αlnε )] .
2
k=lnε
1
lnεX+nε −1 o
× [b(Xk , αlnε ) − b(Xlnε , αlnε )]
nε
k=lnε
= 0.
Thus, in evaluating the limit, for lnε ≤ k ≤ lnε + nε − 1, b(Xk , αlnε ) can
be replaced by b(Xlnε , αlnε ).
The choice of nε implies that εlnε → u as ε → 0 yielding εk → u for all
lnε ≤ k ≤ lnε + nε . Consequently, by weak convergence and the Skorohod
representation, we obtain
κ
Y n (t+s)/ε−1
X
def
L1 = lim E ρι (X ε (tι ), αε (tι )) ε∇h0 (Xlnε , αlnε )
ε→0
ι=1 lnε =t/ε
lnεX
+nε −1 o
× b(Xk , αlnε )
k=lnε (5.22)
Yκ
ε ε
= lim E ρι (X (tι ), α (tι ))
ε→0
ι=1
n (t+s)/ε−1
X 1
lnεX
+nε −1 o
× δε ∇h0 (Xlnε , αlnε ) b(Xlnε , αlnε ) .
nε
lnε =t/ε k=lnε
148 5. Numerical Approximation
Thus,
κ
Y
L1 = lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
n (t+s)/ε−1
X o
× δε ∇h0 (Xlnε , αlnε )b(X ε (lδε ), αε (lδε ))
lnε =t/ε
κ
Y nZ t+s o
=E e ι ), α
ρι (X(t e(tι )) e
∇h0 (X(u), e
e(u))b(X(u),
α e(u))du .
α
ι=1 t
(5.23)
In the above, treating such terms as b(X ε (lδε ), αε (lδε )), we can approx-
imate X ε (·) by a process taking finitely many values using a standard
approximation argument (see, e.g., [104, p. 169] for more details).
Similar to (5.23), we also obtain
κ
Y
def
L2 = lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
nε (t+s)/ε−1 lnε +nε −1
X X o
× tr[∇ h(Xlnε , αlnε )σ(Xk , αlnε )σ 0 (Xk , αlnε )]
2
2
lnε =t/ε k=lnε
κ
Y
=E e ι ), α
ρι (X(t e(tι ))
ι=1
Z
n t+s
1 o
× e
tr[∇2 h(X(u), e
e(u))σ(X(u),
α α e
e(u))σ 0 (X(u), e(u))]du .
α
t 2
(5.24)
Step 4: We next examine the terms on the next to the last line of (5.20).
First, again using the continuity, weak convergence, and the Skorohod rep-
resentation, it can be shown that
κ
Y h (t+s)/ε−1
X
lim E ρι (X ε (tι ), αε (tι )) [h(Xlnε +nε , αlnε +nε )
ε→0
ι=1 lnε =t/ε
i
−h(Xlnε +nε , αlnε )]
κ
Y (5.25)
= lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
h (t+s)/ε−1
X i
× [h(Xlnε , αlnε +nε ) − h(Xlnε , αlnε )] .
lnε =t/ε
κ
Y n (t+s)/ε−1
X o
ε ε
lim E ρι (X (tι ), α (tι )) [h(Xlnε , αlnε +nε ) − h(Xlnε , αlnε )]
ε→0
ι=1 lnε =t/ε
κ
Y
= lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
n (t+s)/ε−1
X lnεX +nε −1 o
× [h(Xlnε , αk+1 ) − h(Xlnε , αk )]
lnε =t/ε k=lnε
Y κ
= lim E ρι (X ε (tι ), αε (tι ))
ε→0
ι=1
n (t+s)/ε−1
X lnεX +nε −1 X m0
m0 X h
× E [h(Xlnε , i)I{αk+1 =i}
lnε =t/ε k=lnε i=1 i1 =1
io
−h(Xlnε , i1 )I{αk =i1 } ] Gk .
(5.26)
Note that for k ≥ lnε ,
h i
E [h(Xlnε , i)I{αk+1 =i} − h(Xlnε , i1 )I{αk =i1 } ] Gk
= [h(Xlnε , i)P(αk+1 = i Gk , αk = i1 ) − h(Xlnε , i1 )]I{αk =i1 }
(5.27)
= [h(Xlnε , i)(δi1 i + εqi1 i (Xk )) − h(Xlnε , i1 )]I{αk =i1 }
= εh(Xlnε , i)qi1 i (Xk )I{αk =i1 } .
Using (5.27) in (5.26) and noting the continuity and boundedness of Q(·),
we can replace qi1 i (Xk ) by qi1 i (Xlnε ) yielding the same limit. Then as in
(5.15) and (5.16), replace I{αk =i1 } by I{αε (εlnε )=i1 } , again yielding the same
limit. Thus, we have
κ
Y n (t+s)/ε−1
X o
lim E ρι (X ε (tι ), αε (tι )) [h(Xlnε , αlnε +nε ) − h(Xlnε , αlnε )]
ε→0
ι=1 lnε =t/ε
κ
Y nZ t+s o
=E e ι ), α
ρι (X(t e(tι )) e
Q(X(u))h( e
X(u), ·)(e
α(u))du ,
ι=1 t
(5.28)
150 5. Numerical Approximation
Step 5: Combining Steps 1–4, we arrive at that (X(·), e e(·)), the weak
α
limit of (X ε (·), αε (·)) is a solution of the martingale problem with oper-
ator L defined in (5.3). Using characteristic functions, we can show as in
[176, Lemma 7.18], (X(·), α(·)), the solution of the martingale problem with
operator L, is unique in the sense of in distribution. Thus (X ε (·), αε (·)) con-
verges to (X(·), α(·)) as desired, which concludes the proof of the theorem.
2
In addition, we can obtain the following convergence result as a corollary.
Corollary 5.9. Under the conditions of Theorem 5.8, the sequence of pro-
cesses (X ε (·), αε (·)) converges to (X(·), α(·)) in the sense
Remark 5.10. With a little more effort, we can obtain strong convergence
(in the usual sense used in the numerical solutions of stochastic differential
equations). The steps involved can be outlined as follows. (a) We consider
two sequences of approximations (X ε (·), αε (·)) and (X η (·), αη (·)) with the
same initial data. (b) Define for sufficiently small ε > 0 and η > 0,
Z t Z t
Xe ε (t) = X0 + b(X ε (s), αε (s))ds + σ(X ε (s), αε (s))dw(s),
Z0 t Z0 t
e η
X (t) = X0 + η η
b(X (s), α (s))ds + σ(X η (s), αη (s))dw(s).
0 0
That is, they are two approximations of the solution with the use of different
stepsizes. Then we can show
e ε (t) − X
E sup |X e η (t)|2 → 0 as ε → 0 and η → 0.
t∈[0,T ]
5.5 Examples
Here we present two examples for demonstration. It would be ideal if we
could compare the numerical solutions using our algorithms with the an-
alytic solutions. Unfortunately, due to the complexity of the x-dependent
switching process, closed-form solutions are not available. We are thus con-
tended with the numerical solutions. In both examples, we use the state-
dependent generator Q(x) given by
−5 cos2 x 5 cos2 x
Q(x) = .
10 cos2 x −10 cos2 x
Let
b(x, i) = A(i)x with A(1) = −3.3 and A(2) = −2.7.
We use the constant stepsize algorithm (5.4). Specify the initial conditions
as X0 = 5 and α0 = 1, and use the constant stepsize 0.001. A sample path
of the computed iterations is depicted in Figure 5.1.
4
x
0
0 200 400 600 800 1000 1200
Iterations
5.5
5
x
4.5
3.5
0 200 400 600 800 1000 1200
Iterations
For these numerical examples, we have tried different stepsizes. They all
produced similar sample path behavior as displayed above. For numeri-
cal experiment purposes, we have also tested different functions b(·, ·) and
σ(·, ·).
Z t Z t
X ε (t) = X0 + b(X ε (s), αε (s))ds + σ(X ε (s), αε (s))dw(s). (5.30)
0 0
Note that in (5.31), the first inequality is obtained from the familiar in-
equality (a + b)2 ≤ 2(a2 + b2 ) for two real numbers a and b. The first term
on the right-side of the second inequality follows from Hölder’s inequality,
and the second term is a consequence of the well-known Doob martingale
inequality (see (A.19) in the appendix). To proceed, we treat the drift and
154 5. Numerical Approximation
The first inequality in (5.32) follows from the familiar triangle inequality,
and the second inequality is a consequence of the Lipschitz continuity, the
Cauchy inequality, and the linear growth condition. We now concentrate
on the last term in (5.32). Using discrete iteration, we have
Z T
E [1 + |X ε (s)|2 ]I{α(s)6=αε (s)} ds
0
bt/εc−1
X Z εk+ε
= E [1 + |X ε (s)|2 ]I{α(s)6=αε (s)} ds.
k=0 εk
Define
n−1
X
tn = εk , m(t) = max{n : tn ≤ t},
k=0
Theorem 5.14. Under (A5.1), (A5.2), and (A5.4), (X n (·), αn (·)) con-
verges to (X(·), α(·)) weakly, which is a solution of the martingale problem
with operator L defined in (5.3). Moreover,
5.7 Notes
Numerical methods for stochastic differential equations have been studied
extensively, for example, in [94, 126] among others. A comprehensive study
of the early results is contained in Kloeden and Platen [94]. Accelerated
rates of convergence are given in Milstein [126]. As a natural extension,
further results are considered in Milstein and Tretyakov [127], among oth-
ers. Numerical solutions for stochastic differential equations modulated by
Markovian switching have also been well studied; recent progress for switch-
ing diffusions are contained in [120] and references therein.
Although numerical methods for Markov modulated switching diffusions
have been considered by many researchers, less is known for processes with
continuous-state-dependent switching. In fact, the study of switching diffu-
sions with state-dependent switching is still in its infancy. There are certain
difficulties. For example, the usual Picard iterations cannot be used, which
uses the Lipschitz conditions crucially. When Markovian regime-switching
processes are treated, with the given generator of the switching process,
we can pre-generate the switching process throughout the iterations. How-
ever, in the state-dependent switching case, since the generation of the
x-dependent switching processes is different in every step, we can no longer
5.7 Notes 157
6.1 Introduction
Continuing with the development in Chapter 5, this chapter is devoted to
additional properties of numerical approximation algorithms for switching
diffusions, where continuous dynamics are intertwined with discrete events.
In this chapter, we establish that if the invariant measure exists, under suit-
able conditions, the sequence of iterates obtained using Euler–Maruyama
approximation converges to the invariant measure.
Here for simplicity, the discrete events are formulated as a finite-state
continuous-time Markov chain that can accommodate a set of possible
regimes, across which the dynamic behavior of the systems may be markedly
different. For simplicity, we have chosen to present the result for Q(x) = Q.
One of the motivations for the use of a constant matrix Q is: We may view it
as an approximation to x-dependent Q(x) in the sense of Q(x) = Q + o(1)
as |x| → ∞. This is based on the results in Chapters 3 and 4. Because
positive recurrence implies ergodicity, we could concentrate on a neighbor-
hood of ∞. Then effectively, Q, the constant matrix is the one having the
most contributions to the asymptotic properties in which we are interested.
Thus, we can “replace” Q(x) by Q in the first approximation. At any given
instance, in lieu of a fixed regime, the system parameters can take one of
several possible regimes (configurations). As the Markov chain sojourns in
a given state for a random duration, the system dynamics are governed by
a diffusion process in accordance with the associated stochastic differential
equation. Subsequently, the Markov chain jumps into another state, and
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 159
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_6,
© Springer Science + Business Media, LLC 2010
160 6. Numerical Approximation to Invariant Measures
where
m0
X
Qf (x, ·)(ı) = qı f (x, ), ı ∈ M,
=1
and
A(x, ı) = (aij (x, ı)) = σ(x, ı)σ 0 (x, ı).
As alluded to in the previous chapters, switching diffusions have pro-
vided many opportunities in terms of flexibility. The formulation allows
the mathematical models to have multiple discrete configurations thereby
making them more versatile. However, solving systems of diffusions with
switching is still a challenging task, which often requires using numerical
methods and/or approximation techniques. The Euler–Maruyama scheme
is one of such approaches. In Section 6.2, we derive the tightness of the
approximating sequence. To proceed, an important problem of both theo-
retical and practical concerns is: Whether the sequence of approximation
converges to the invariant measure of the underlying system, provided that
it exists. To answer this question, we derive the convergence to the invari-
ant measures of the numerical approximation in Section 6.3. To obtain the
results requires the convergence of the algorithm under the weak conver-
gence framework. Rather than working with sample paths or numerically
solving systems of Kolmogorov–Fokker–Planck equations, we focus on the
corresponding measures and use a purely probabilistic argument. Why is
such a consideration important from a practical point of view? Suppose
that one considers an ergodic control problem of a hybrid-diffusion sys-
tem with regime switching. Then it is desirable to “replace” the actual
time-dependent measure by an invariant measure. The control problems
often have to be solved using numerical approximations. Because solving
6.2 Tightness of Approximation Sequences 161
√
Xn+1 = Xn + εn b(Xn , αn ) + εn σ(Xn , αn )ξn , (6.3)
Various quantities are as given in the last chapter. Associated with (6.1),
there is a martingale problem formulation. A process (X(t), α(t)) is said to
be a solution of the martingale problem with operator L defined in (6.2) if,
Z t
h(X(t), α(t)) − Lh(X(s), α(s))ds (6.5)
0
Remark 6.1. Note that sufficient conditions for the existence and unique-
ness of the solution of the switching diffusion were provided in Chapter 2.
Here we simply assume them for convenience.
Condition (A6.2) requires the existence of Liapunov functions V (·, α).
Condition (A6.2)(d) is a growth condition on the functions b(x, α) and
σ(x, α). If b(·, α) and σ(·, α) grow at most linearly, and the Liapunov func-
tion is quadratic, this condition is verified. Condition (c) requires the dif-
fusion with regime switching (6.1) to be stable in the sense of Liapunov.
Conditions listed in (A6.2) cover a large class of functions; see the related
comments in [102, 104].
Remark 6.2. A quick glance at the algorithm reveals that (6.3) has a cer-
tain resemblance to a stochastic approximation algorithm, which has been
the subject of extensive research for over five decades since the pioneering
work of Robbins and Monro [139]. The most recent account on the subject
and a state-of-the-art treatment can be found in Kushner and Yin [104] and
references therein. In what follows, we use certain ideas from stochastic ap-
proximation methods to establish the limit of the discretization through
suitable interpolations. Weak convergence methods are used to study the
convergence of the algorithm and the associated invariant measure.
Remark 6.3. As alluded to in Chapter 5, it can be established that
the sequence {Xn } is tight by using the moment estimates together with
Tchebeshev’s inequality. Here we use an alternative approach based on Li-
apunov function methods. Note that this approach can be modified to a
perturbed Liapunov function approach, which can be used to handle cor-
related random processes under suitable conditions. To illustrate the use
of the Liapunov function method for the tightness, we present the result
below together with a proof.
Theorem 6.4. Assume (A6.2) and (A6.3). Then
(i) the iterates generated by (6.3) using decreasing stepsizes satisfy
EV (Xn , αn ) = O(1);
EV (Xn , αn ) = O(1)
Proof. The proof uses Liapunov functions. We concern ourselves with the
proof for algorithm (6.3) only. The results for (6.4) can be obtained simi-
larly.
Henceforth, denote by Fen and Fenα the σ-algebras generated by {ξk , αk :
k < n} and {αn , ξk , αk : k < n}, and denote by En and Eα n the correspond-
ing conditional expectations w.r.t. to Fen and Fenα , respectively. Similarly,
denote Ft = σ{w(u), α(u) : u ≤ t} and Et the conditional expectation
w.r.t. Ft , where w(t) is an r-dimensional standard Brownian motion (hav-
ing independent increments). Note that {ξn } is a sequence of independent
and identically distributed random vectors with 0 mean and covariance I.
We have
Eα
n V (Xn+1 , αn+1 ) − V (Xn , αn )
= Eα
n [V (Xn+1 , αn+1 ) − V (Xn+1 , αn )]
(6.7)
+Eα
n V (Xn+1 , αn ) − V (Xn , αn ).
We proceed to estimate each of the terms after the equality sign above.
Using the smoothness of V (·, α), the independence of ξn with αn , and
the independent increment property, estimates (with details omitted) lead
to
Eα
n V (Xn+1 , αn ) − V (Xn , αn )
= εn ∇V 0 (Xn , αn )b(Xn , αn )
εn (6.8)
+ tr[σ 0 (Xn , αn )∇2 V (Xn , αn )σ(Xn , αn )]
2
+O(ε2n )(1 + V (Xn , αn )).
As for the term on the second line of (6.7), using a truncated Taylor
expansion, we obtain
Eα
n [V (Xn+1 , αn+1 ) − V (Xn+1 , αn )]
= Eα
n [V (Xn , αn+1 ) − V (Xn , αn )]
Z 1 (6.9)
α
+En ∇V 0 (Xn + s∆Xn , αn )(Xn+1 − Xn )ds
0
= εn QV (Xn , ·)(αn ) + O(εn )(1 + V (Xn , αn )).
By virtue of (A6.2),
Eα
n V (Xn+1 , αn+1 ) − V (Xn , αn )
EV (Xn+1 , αn+1 )
n
X n
X
≤ An,−1 EV (X0 , α0 ) + K εk Ank EV (Xk , αk ) + K εk Ank ,
k=0 k=0
(6.10)
where Qn
j=k+1 (1 − λεj ), if n > k,
Ank =
1, otherwise.
Note that
n
X
εk Ank = O(1).
k=0
An application of Gronwall’s inequality implies
n
!
X
EV (Xn , αn ) ≤ K exp εk Ank = O(1).
k=0
Let m(t) be the unique value of n such that tn ≤ t < tn+1 for t ≥ 0. Define
αn = α(tn ). Let X 0 (t) and α0 (t) be the piecewise constant interpolations
of Xn and αn on [tn , tn+1 ), and X n (t) and αn (t) be their shifts. That is,
(A6.4) The process (X(·), α(·)) has a unique invariant measure ν(·).
Denote by ν(x, α; t, ·) the measure of (X(t), α(t)) with initial
condition (X(0), α(0)) = (x, α). As t → ∞, ν(x, α; t, ·) converges
weakly to ν(·) for each (x, α). For any compact Kb ⊂ Rr and for
r
any ϕ ∈ C(R × M),
L∗ µ(x, i) = 0, i ∈ M,
m0 Z
X (6.13)
µ(x, j)dx = 1
j=1
is bounded and continuous for each t > 0 because ϕ(·, ι) ∈ Cb (Rr ) for
each ι ∈ M. Suppose that X(0) ∈ Kc and the distribution with initial
condition (X(0), α(0)) is denoted by ν(0, ·). Then the tightness of X(0)
168 6. Numerical Approximation to Invariant Measures
inasmuch as
m0 Z
X
ν(dx × i) = 1.
i=1
(i) For arbitrary positive integer n0 , ϕ(·, ι) ∈ C(Rr ) for each ι ∈ M, and
for any δ > 0, there exist t0 < ∞ and positive integer N0 such that
for all t ≥ t0 and n ≥ N0 ,
Eϕ(X n (t + δj ), αn (t + δj ), j ≤ n0 )
(6.16)
− Eν ϕ(X(δj ), α(δj ), j ≤ n0 ) < δ.
Proof of Theorem 6.11. Suppose that (6.16) were not true. There would
exist a subsequence {nk } and a sequence snk → ∞ such that
For a fixed T > 0, choose a further subsequence {k` } of {nk }, and the cor-
responding sequence (X k` (·), αk` (·)) such that (X k` (sk` − T ), αk` (sk` − T ))
converges weakly to a random variable (X(0), α(0)). Theorem 6.5 implies
that (X k` (sk` − T + ·), αk` (sk` − T + ·)) converges weakly to (X(·), α(·))
with initial condition (X(0), α(0)). Moreover,
Proof of Theorem 6.5. We can show that Lemma 5.4 continues to hold
and {αn (·)} is tight with the constant stepsize replaced by the decreasing
stepsizes.
170 6. Numerical Approximation to Invariant Measures
N √
Xn+1 = XnN + εn bN (XnN , αn ) + εn σ N (XnN , αn )ξn , (6.19)
where
2
Etm(tn +t) X n,N (t + s) − X n,N (t)
m(t+s+tn )−1 2
X
≤ 2Etm(tn +t) bN (XkN , αk )εk (6.21)
k=m(t+tn )
m(t+s+tn )−1 2
X √
+ 2Etm(tn +t) εk σ N (XkN , αk )ξk ,
k=m(t+tn )
where Etm(tn +t) denotes the conditional expectation on the σ-algebra gen-
erated by {(Xj , αj ) : j ≤ m(tn + t)}. The continuity of bN (·, i) and σ N (·, i)
(for each i ∈ M), the smoothness of q N (·), and the boundedness of XnN
6.4 Proof: Convergence of Algorithm 171
Eα 0
l ξk ξl = E[ξ
0
ξ |α , ξ , α ; j < l]
k l l j j
0, if l 6= k,
=
1, if l = k.
Etm(tn +t) |X n,N (t + s) − X n,N (t)|2 ≤ Etm(tn +t) γ n,N (s), (6.25)
Lemma 6.14 The pair (X N (·), α(·)) is the solution of the martingale prob-
lem with operator LN obtained from L by replacing b(·) and σ(·) with bN (·)
and σ N (·) (defined in (6.20)), respectively.
Proof of Lemma 6.14. We need to verify that (6.5) holds. Without loss
of generality, we work with t ≥ 0. It suffices to show that for each i ∈ M,
any real-valued function h(·, i) ∈ C02 , any T < ∞, 0 ≤ t ≤ T , s > 0,
arbitrary positive integer n0 , bounded and continuous functions ϕj (·, i)
(with j ≤ n0 ), and any sj satisfying 0 < sj ≤ t ≤ t + s,
n0
Y
E ϕj (X N (sj ), α(sj )) h(X N (t + s), α(t + s))
j=1
Z t+s
(6.26)
N N N
− h(X (t), α(t)) − L h(X (u), α(u))du = 0.
t
To obtain (6.26), let us begin with the process (X n,N (·), αn (·)). By virtue
of the weak convergence of (X n,N (·), αn (·)) to (X N (·), α(·)) and the Sko-
rohod representation, as n → ∞,
n0
Y
E ϕj (X n,N (sj ), αn (sj )) h(X n,N (t + s), αn (t + s))
j=1
− h(X n,N (t), αn (t))
n0
Y
→E ϕj (xN (sj ), α(sj )) h(X N (t + s), α(t + s)) − h(X N (t), α(t)) .
j=1
(6.27)
Choose δn and ml such that δn → 0 as n → ∞ and
ml+1 −1
1 X
εj → 1 as n → ∞.
δn j=1
Then
h(X n,N (t + s), αn (t + s)) − h(X n,N (t), αn (t))
X
N N
= [h(Xm , αml+1 ) − h(Xm , αml )]
l∈Ξ
l+1 l+1
(6.29)
X
N N
+ [h(Xm l+1
, αml ) − h(Xm l
, αml )].
l∈Ξ
6.4 Proof: Convergence of Algorithm 173
X
N N
[h(Xm l+1
, αml ) − h(Xm l
, αml )]
l∈Ξ
X
= ∇h0 (Xm
N
l
N
, αml )[Xm l+1
N
− Xm l
] (6.30)
l∈Ξ
1X N N 0 2 N N N
+ [Xm − Xm ] ∇ h(Xm , αml )[Xm − Xm ] + een ,
2 l+1 l l l+1 l
l∈Ξ
where ∇2 h denotes the Hessian of h(·, α) and een represents the error in-
curred from the truncated Taylor expansion. By the continuity of ∇2 h(·, α),
the boundedness of {XnN }, it is readily seen that
n0
Y
lim E ϕj (X n,N (sj ), αn (sj ))(|e
en |) = 0. (6.31)
n→∞
j=1
X
∇h0 (Xm
N
l
N
, αml )[Xm l+1
N
− Xm l
]
l∈Ξ
ml+1 −1
X X
= ∇h0 (Xm
N
l
, α ml ) bN (XkN , αk )εk
l∈Ξ k=ml
ml+1 −1
X X √
+ ∇h0 (Xm
N
l
, α ml ) εk σ N (XkN , αk )ξk .
l∈Ξ k=ml
n0
Y hX
E ϕj (X n,N (sj ), αn (sj )) ∇h0 (Xm
N
l
, α ml )
j=1 l∈Ξ
ml+1 −1
X √ i
× εk σ N (XkN , αk )ξk
k=ml
n0
Y h X
=E ϕj (X n,N (sj ), αn (sj )) Em(tn +t) ∇h0 (Xm
N
l
, α ml )
j=1 l∈Ξ
ml+1 −1
X √ i
× εk σ N (XkN , αk )Eα
k ξk = 0.
k=ml
174 6. Numerical Approximation to Invariant Measures
Note that
n0
Y hX
lim E ϕj (X n,N (sj ), αn (sj )) ∇h0 (Xm
N
l
, α ml )
n→∞
j=1 l∈Ξ
ml+1 −1
X i
× bN (XkN , αk )εk
k=ml
n0
Y (6.32)
n,N n
X 0 N
= lim E ϕj (x (sj ), α (sj )) ∇h (Xm l
, αml )δn
n→∞
j=1 l∈Ξ
ml+1 −1
1 X
× bN (Xm
N
, αk )εk .
δn l
k=ml
Owing to the interpolation, for k ∈ [ml , ml+1 − 1], write αk as αn (u) with
u ∈ [tml , tml+1 −1 ). Then
ml+1 −1
1 X
bN (Xm
N
, αk )εk
δn l
k=ml
m0 ml+1 −1
X 1 X
= bN (Xm
N
, i)I{αk =i} εk
δ
i=1 n
l
k=ml
m0 ml+1 −1
X 1 X
= bN (Xm
N
, i)I{αn (u)=i} εk .
δ
i=1 n
l
k=ml
n0
Y hX ml+1 −1
X i
E ϕj (X n,N (sj ), αn (sj )) ∇h0 (Xm
N
l
, α ml ) bN (XkN , αk )εk
j=1 l∈Ξ k=ml
m0
X n0
Y
= E ϕj (X n,N (sj ), αn (sj ))
i=1 j=1
X
× ∇h0 (X n,N (tml ), i)bN (X n,N (tml ), i)I{αn (u)=i} δn
l∈Ξ
m0
X n0
Y Z t+s
→ E ϕj (X N (sj ), α(sj )) ∇h0 (X N (u), α(u))bN (X N (u), i)
i=1 j=1 t
×I{α(u)=i} du .
6.4 Proof: Convergence of Algorithm 175
Thus
n0
Y
E ϕj (X n,N (sj ), αn (sj ))[Xm
n,N
l+1
n,N
− Xm l
]
j=1
n0
Y hZ t+s i
→E ϕj (X N (sj ), α(sj )) ∇h0 (X N (u), α(u))bN (X N (u), α(u))du .
j=1 t
(6.33)
Next, consider the term involving ∇2 h(·, α) in (6.30). We have
X
N N 0 2 N N N
[Xm l+1
− Xm l
] ∇ h(Xm l
, αml )[Xm l+1
− Xm l
]
l∈Ξ
X−1 h
X−1 ml+1
X ml+1
= εk1 εk bN,0 (XkN1 , αk1 )
l∈Ξ k1 =ml k=ml
×∇ 2 N
h(Xm , αml )bN (XkN , αk ) (6.34)
l
√
+ εk1 εk bN,0 (XkN1 , αk1 )∇2 h(Xm N
l
, αml )σ N,0 (XkN , αk )ξk
√
+ εk εk1 ξk0 1 σ N,0 (XkN1 , αk1 )∇2 h(Xm N
l
, αml )bN (XkN , αk )
√
+ tr[∇2 h(Xm N
l
, αml ) εk εk1 σ N (XkN , αk )ξk ξk0 1 σ N,0 (XkN1 , αk1 ) .
n0
Y X ml+1
X−1 ml+1
X−1
E ϕj (X n,N (sj ), αn (sj )) εk 1 εk
j=1 l∈Ξ k1 =ml k=ml
≤ Kδn → 0 as n → ∞,
since
X ml+1 −1
X
t+s t 1
δn = δ n − = O(1) and εk = O(1).
δn δn δn
l∈Ξ k=ml
176 6. Numerical Approximation to Invariant Measures
X ml+1
X−1 ml+1
X−1 √
E εk1 εk bN,0 (XkN1 , αk1 )∇2 h(Xm
N
l
, α ml )
l∈Ξ k1 =ml k=ml
N
×σ (XkN , αk )ξk
ml+1 −1 ml+1 −1 2
X X X √
≤K εk1 E1/2 εk σ N (XkN , αk )ξk (6.35)
l∈Ξ k1 =ml k=ml
ml+1 −1
X 1 X
≤K δn3/2 εk 1
δn
l∈Ξ k1 =ml
X
≤K δn3/2 → 0 as n → ∞,
l∈Ξ
and
n0 ml+1 −1 ml+1 −1
Y X X
n,N n
E ϕj (X (sj ), α (sj )) εk ξk0 1 σ N,0 (XkN1 , αk1 )
j=1 k1 =ml k=ml (6.36)
2 N
×∇ h(Xm l
, αml )bN (XkN , αk ) → 0 as n → ∞.
n0 ml+1 −1 ml+1 −1
Y X X √
E ϕj (X n,N (sj ), αn (sj )) εk εk1 tr[∇2 h(Xm
N
l
, α ml )
j=1 k1 =ml k=ml
n0 ml+1 −1
Y X
n,N n
= E ϕj (X (sj ), α (sj )) εk tr[∇2 h(Xm
N
l
, αml )σ N (XkN , αk )
j=1 k=ml
× εk [Eα 0
k ξk ξk ]σ
N,0
(XkN , αk )]
Yn0 Z
t+s
→E ϕj (X N (sj ), α(sj )) tr[∇2 h(X N (u), α(u))σ N (X N (u), α(u))
j=1 t
× σ N,0 (X N (u), α(u))]du .
(6.37)
6.4 Proof: Convergence of Algorithm 177
where
H(x) = (h(x, 1), . . . , h(x, m0 ))0 ∈ Rm0 ×1 . (6.40)
Thus Lemma 6.14 is proved. 2
Completion of the proof of the theorem. We show the convergence
of the untruncated process. We have demonstrated that the truncated pro-
cess {(X n,N (·), αn (·))} converges to (X N (·), α(·)). Here we show that the
untruncated sequence {(X n (·), αn (·))} also converges. The basic premise is
the uniqueness of the martingale problem. By letting N → ∞, we obtain
the desired result. The argument is similar to that of [104, pp. 249–250].
We present only the basic idea here.
Let the measures induced by (X(·), α(·)) and (X N (·), α(·)) be P (·) and
N
P (·), respectively. The martingale problem with operator L has a unique
solution (in the sense in distribution) for each initial condition, therefore
P (·) is unique. For any 0 < T < ∞ and |t| ≤ T , P (·) and P N (·) are the
same on all Borel subsets of the set of paths in D((−∞, ∞); Rr × M) with
values in SN × M. By using
P ( sup |X(t)| ≤ N ) → 1 as N → ∞,
|t|≤T
and the weak convergence of X n,N (·) to X N (·), we conclude X n (·) con-
verges weakly to X(·). This leads to the desired result. The proof of the
theorem is completed. 2
6.5 Notes
Chapter 4 provides sufficient conditions for ergodicity of switching diffu-
sions with state-dependent switching. Based on the work Yin, Mao, and Yin
[171], this chapter addresses the ergodicity for the corresponding numerical
algorithms. The main result here is the demonstration of convergence to the
invariant measure of the Euler–Maruyama-type numerical algorithms when
the invariant measure exists. To obtain the result, we have first proved weak
convergence of the algorithms. Here our approach is inspired by Kushner’s
work [101], in which he considered convergence to invariant measures for
systems driven by wideband noise. We have adopted the method in that
reference to treat the numerical approximation problem. Moreover, conver-
gence of numerical algorithms has been proved using ideas from stochastic
approximation (see Kushner and Yin [104]). We have dealt with both al-
gorithms with decreasing stepsizes and constant stepsize.
Here our approach is based on weak convergence methods and we work
with the associated measure. A different approach concentrating on the
associated differential equations is in Mao, Yuan, and Yin [121]. The rate
of convergence of the algorithms may be studied, for example, by means
6.5 Notes 179
7.1 Introduction
Continuing our effort of studying positive recurrence and ergodicity of
switching diffusion processes in Chapters 3 and 4, this chapter focuses
on stability of the dynamic systems described by switching diffusions. For
some of the recent progress in stability analysis, we refer the reader to
[48, 116, 136, 182, 183] and references therein. For treating dynamic sys-
tems in science and engineering, linearization techniques are used most
often. Nevertheless, the nonlinear systems and their linearizations may or
may not share similar asymptotic behavior. A problem of great interest is:
If a linear system is stable, what can we say about the associated nonlinear
systems? This chapter provides a systematic approach for treating such
problems for switching diffusions. We solve these problems using Liapunov
function methods.
The rest of the chapter is arranged as follows. Section 7.2 begins with
the formulation of the problem together with an auxiliary result, which is
used in our stability analysis. Section 7.3 recalls various notions of stability,
and presents p-stability and exponential p-stability results. Easily verifiable
conditions for stability and instability of linearized systems are provided
in Section 7.4. To demonstrate our results, we provide several examples
in Section 7.5. Further remarks are made in Section 7.6 to conclude this
chapter.
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 183
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_7,
© Springer Science + Business Media, LLC 2010
184 7. Stability
+Q(x)g(x, ·)(i)
1
= tr(a(x, i)∇2 g(x, i)) + b0 (x, i)∇g(x, i) + Q(x)g(x, ·)(i),
2
(7.1)
where x ∈ Rr , Q(x) = (qij (x)) isPan m0 × m0 matrix depending on x
satisfying qij (x) ≥ 0 for i 6= j and j∈M qij (x) = 0 for each i ∈ M, and
X
Q(x)g(x, ·)(i) = qij (x)g(x, j)
j∈M
X
= qij (x)(g(x, j) − g(x, i)), i ∈ M,
j∈M
and ∇g(·, i) and ∇2 g(·, i) denote the gradient and Hessian of g(·, i), respec-
tively.
The process (X(t), α(t)) can be described by
and for i 6= j,
the random dynamic system. Hence we assume that the following conditions
are valid throughout this chapter.
(A7.1) The matrix-valued function Q(·) is bounded and continuous.
(A7.2) b(0, α) = 0, and σ(0, α) = 0 for each α ∈ M. Moreover, assume
that σ(x, α) vanishes only at x = 0 for each α ∈ M.
(A7.3) There exists a constant K0 > 0 such that for each α ∈ M, and
for any x, y ∈ Rr ,
|b(x, α) − b(y, α)| + |σ(x, α) − σ(y, α)| ≤ K0 |x − y|. (7.4)
Under these conditions, the system given by (7.2) and (7.3) has a unique
solution; see Chapter 2 for more details. In what follows, a process start-
ing from (x, α) is denoted by (X x,α (t), αx,α (t)) if the emphasis on initial
condition is needed. If the context is clear, we simply write (X(t), α(t)).
To study stability of the equilibrium point x = 0, we first present
the following “nonzero” property, which asserts that almost all the sam-
ple paths of any solution of the system given by (7.2) and (7.3) starting
from a nonzero state will never reach the origin. For diffusion processes,
such a result was established in [83, Section 5.2]; for Markovian regime-
switching processes, similar results were obtained in [116, Lemma 2.1]. In
what follows, we give a proof for switching diffusions with continuous-state-
dependent switching processes. The result is useful, provides us with flexi-
bility for choices of Liapunov functions, and enables us to build Liapunov
functions in a deleted neighborhood of the origin.
Lemma 7.1. Under conditions (A7.1)–(A7.3), we have
P{X x,α (t) 6= 0, t ≥ 0} = 1, for any x 6= 0, α ∈ M, (7.5)
and for any β ∈ R and t > 0
h i
β β
E |X x,α (t)| ≤ |x| eKt , x 6= 0, α ∈ M, (7.6)
Proof. For x 6= 0, for each i ∈ M, define V (x, i) = |x|β for any β ∈ R−{0}.
For any ∆ > 0 and |x| > ∆,
∇V (x, i) = β|x|β−2 x,
2
∇2 V (x, i) = β|x|β−4 |x| I + (β − 2)xx0 .
Let τ∆ be the first exit time from {x ∈ Rr : |x| > ∆}. Denote (X(t), α(t)) =
(X x,α (t), αx,α (t)). Now applying generalized Itô’s formula (see (2.7)) to V ,
we obtain
Z τ∆ ∧t
β β β
|X(τ∆ ∧ t)| = |x| + L |X(u)| du
Z τ∆ ∧t0 (7.8)
β−2 0
+ β |X(u)| X (u)σ(X(u), α(u))dw(u).
0
P {X x0 ,α0 (T ) = 0} > 0.
which would contradict with (7.6) that we just proved. This completes the
proof of the lemma. 2
Remark 7.2. In view of (7.5), we can work with functions V (·, i), i ∈
M that are twice continuously differentiable and are defined on a deleted
neighborhood of 0 in what follows.
To proceed, we present an auxiliary result, namely, the solvability of a
system of deterministic equations. Suppose that Q, an m0 × m0 constant
matrix, is the generator of a continuous-time Markov chain r(t) and that
Q is irreducible.
Remark 7.3. In the above, by the irreducibility, we mean that the system
of equations
νQ = 0
ν1l = 1,
has a unique solution such that ν = (ν1 , . . . , νm0 ) satisfies νi > 0; see
Definition A.7 and discussions there for further details.
Note that if Q is irreducible, the rank of Q is m0 −1. Denote by R(Q) and
N (Q) the range and the null space of Q, respectively. It follows that N (Q)
is one-dimensional spanned by 1l (i.e., N (Q) = span{1l}). As a consequence,
the Markov chain r(t) is ergodic; see, for example, [29]. In what follows,
denote the associated stationary distribution by
Qc = η, (7.11)
188 7. Stability
where Q ∈ Rm0 ×m0 and η ∈ Rm0 are given and c ∈ Rm0 is an unknown
vector. Note that (7.11) is a Poisson equation. The properties of solutions
of (7.11) are provided in Lemma A.12. Basically, it indicates that under
the irreducibility of Q, equation (7.11) has a solution if and only if νη = 0.
Moreover, suppose that c1 and c2 are two solutions of (7.11). Then c1 −c2 =
α0 1l for some α0 ∈ R.
7.3 p-Stability
This section is concerned with stability of the equilibrium point x = 0 for
the system given by (7.2) and (7.3). Adopting the terminologies of [83], we
first present definitions of stability, p-stability, and exponential p-stability.
Then general results in terms of Liapunov functions are provided.
7.3.1 Stability
Definition 7.4. The equilibrium point x = 0 of the system given by (7.2)
and (7.3) is said to be
(i) stable in probability, if for any ε > 0 and any α ∈ M,
Using similar arguments as those of [83, Theorems 5.3.1, 5.4.1, and 5.4.2],
we establish the following three lemmas. The statements are given together
their proofs.
7.3 p-Stability 189
Proof. Let ς > 0 be such that the ball Bς = {x ∈ Rr : |x| < ς} and its
boundary ∂Bς = {x ∈ Rr : |x| = ς} are contained in D. Set
Vς := inf {V (y, j) : y ∈ D \ Bς , j ∈ M} .
Then Vς > 0 by assumption (i). Next, by virtue of assumption (ii) and
Dynkin’s formula, we have
Z τς ∧t
Ex,i V (X(t ∧ τς ), α(t ∧ τς )) = V (x, i) + Ex,i LV (X(s), α(s))ds
0
≤ V (x, i),
where (x, i) ∈ Bς × M and τς is the first exit time from Bς , that is, τς :=
{t ≥ 0 : |X(t)| ≥ ς}. Because V is nonnegative, we further have
Vς P {τς ≤ t} ≤ Ex,i V (X(τς ), α(τς ))I{τς ≤t} ≤ V (x, i).
Note that τς ≤ t if and only if sup0≤u≤t |x(u)| > ς. Therefore it follows
that
V (x, i)
Px,i sup |X(u)| > ς ≤ .
0≤u≤t Vς
Letting t → ∞, we obtain
V (x, i)
Px,i sup |X(t)| > ς ≤ .
0≤t Vς
Finally, the desired conclusion follows from the assumptions that V (0, i) =
0 and V (·, i) is continuous for each i ∈ M. 2
Introduce the notation
x,α
τε,r 0
:= inf{t ≥ 0 : |X x,α (t)| = ε or |X x,α (t)| = r0 }, (7.12)
for any 0 < ε < r0 and any (x, α) ∈ Rr × M with ε < |x| < r0 .
Lemma 7.6. Assume the conditions of Lemma 7.5. If for any sufficiently
small 0 < ε < r0 and any (x, α) ∈ Rr × M with ε < |x| < r0 , we have
x,α
P{τε,r 0
< ∞} = 1, (7.13)
then the equilibrium point x = 0 is asymptotically stable in probability.
190 7. Stability
Px,α {τρ < ∞} ≥ Px,α {τρ,ς < ∞} − Px,α {τς < ∞} ≥ 1 − ε/2.
This implies that Px,α {inf t≥0 |X(t)| ≤ ρ} ≥ 1 − ε/2. Since ρ > 0 can be
arbitrarily small,
Px,α inf |X(t)| = 0 ≥ 1 − ε/2.
t≥0
Now let
A := ω ∈ Ω : τς (ω) = ∞, inf |X(t, ω)| = 0 .
t≥0
If the claim were false, there would exist a B ⊂ A with Px,α (B) > 0 such
that for all ω ∈ B, we would have
Then for any ω ∈ B, there exists a T = T (ω) > 0 such that |X x,α (t, ω)| ≥ θ
for all t ≥ T . Therefore for any ω ∈ B and n ∈ N sufficiently large,
τ1/n (ω) ≤ T , where τ1/n (ω) := inf {t ≥ 0 : |X x,α (t, ω)| ≤ 1/n}. Hence it
follows that
lim τ1/n (ω) ≤ T (ω) < ∞.
n→∞
7.3 p-Stability 191
But this would lead to a contradiction because by virtue of Lemma 7.1, the
equilibrium point 0 is inaccessible with probability 1. Thus it follows that
n o n o
Px,α lim τ1/n = ∞ = 1, or Px,α lim τ1/n < ∞ = 0.
n→∞ n→∞
lim inf V (X x,α (t), αx,α (t)) = 0, for almost all ω ∈ A. (7.16)
t→∞
Note that (x, α) is an arbitrary point in Bδ × M. Thus, for any ε > 0, there
exists a δ > 0 such that
n o
Px,α lim X(t) = 0 ≥ 1 − ε/2, for all (x, α) ∈ Bδ × M.
t→∞
Assertion (i) can be established using almost the same proof as that for
[83, Theorem 3.7.1]. Also (ii) follows by observing that if (7.22) is satisfied,
then we can construct some Liapunov function V (·, ·) satisfying (7.21); see
also a similar argument in the proof of [83, Corollary 3.7.2]. We omit the
details here for brevity.
Theorem 7.9. Assume that for each i ∈ M, the coefficients of the op-
erator L defined in (7.1) satisfy b(·, i) ∈ C 2 and σ(·, i) ∈ C 2 and that
|qij (x)| ≤ K for all x ∈ Rr and some K > 0. Suppose that φ(·, i) ∈ C 2 and
that Dxθ φ(·, i) is Lipschitz continuous for each i ∈ M and |θ| = 2, and that
γ
Dxβ b(x, i) + Dxβ σ(x, i) + Dxθ φ(x, i) ≤ K(1 + |x| ), i ∈ M, (7.23)
where K and γ are positive constants and β and θ are multi-indices with
|β| ≤ 2 and |θ| ≤ 2. Then for any T > 0, the function
u(t, x, i) := Ex,i [φ(X(t), α(t))] = E[φ(X x,i (t), αx,i (t))] (7.24)
Proof. For notational simplicity, we prove the theorem when X(t) is one-
dimensional, the multidimensional case can be handled in a similar man-
ner. Fix (t, x, i) ∈ [0, T ] × Rr × M. Let x e = x + ∆ with 0 < |∆| < 1. As
in the proof of Theorem 2.27, we denote (X(t), α(t)) = (X x,i (t), αx,i (t))
e
and (X(t), e(t)) = (X xe,i (t), αxe,i (t)). By virtue of Theorem 2.27, the mean
α
square derivative ζ(t) = (∂/∂x)X x,i (t) exists and is mean square continu-
ous with respect to x and t.
194 7. Stability
Write
e, i) − u(t, x, i)
u(t, x 1 e
= E[φ(X(t), e(t)) − φ(X(t), α(t))]
α
∆ ∆
1 e e
= E[φ(X(t), e(t)) − φ(X(t),
α α(t))] (7.25)
∆
1 e
+ E[φ(X(t), α(t)) − φ(X(t), α(t))].
∆
Similar to the proof of Lemma 2.28, we can show that
1 2
E sup φ( e
X(t), e
α (t)) − φ( e
X(t), α(t)) → 0, (7.26)
∆2 0≤t≤T
where
e − X(t)
X(t)
Z(t) = . (7.27)
∆
Thus it follows that
1 e
E[φ(X(t), α(t)) − φ(X(t), α(t))] − E [φx (X(t), α(t))ζ(t)]
∆ Z 1
≤E e − X(t)), α(t))dvZ(t) − φx (X(t), α(t))ζ(t)
φx (X(t) + v(X(t)
Z
0
1
≤E e
φx (X(t) + v(X(t) − X(t)), α(t))dv − φx (X(t), α(t)) Z(t)
0
+E |φx (X(t), α(t)) [Z(t) − ζ(t)]|
:= e1 + e2 ,
for all 0 < |∆| < 1. Recall also from the proof of Theorem 2.27 that
e
X(t) → X(t) in probability for any t ∈ [0, T ]. Thus it follows that
2
e − X(t)), α(t)) − φx (X(t), α(t))
E φx (X(t) + v(X(t) → 0,
2
as ∆ → 0. Note that we proved in Corollary 2.32 that E |Z(t)| ≤ K, where
Z(t) is the “difference quotient” defined in (7.27). Then we have from the
Cauchy–Schwartz inequality that
Z 1
e1 = E e − X(t)), α(t))dv − φx (X(t), α(t)) Z(t)
φx (X(t) + v(X(t)
0
Z 1 2
≤ E1/2 e − X(t)), α(t))dv − φx (X(t), α(t))
φx (X(t) + v(X(t)
0
2
×E1/2 |Z(t)|
→ 0 as ∆ → 0.
1 e
E[φ(X(t), α(t)) − φ(X(t), α(t))] − E [φx (X(t), α(t))ζ(t)] → 0. (7.28)
∆
e, i) − u(t, x, i)
u(t, x
− E [φx (X(t), α(t))ζ(t)] → 0 as ∆ → 0.
∆
e, i)
∂u(t, x, i) ∂u(t, x e e − φx (X(t), α(t))ζ(t)
− ≤ E φx (X(t), e(t))ζ(t)
α
∂x ∂x
e − ζ(t))
≤ E φx (X(t), α(t))(ζ(t)
e
+E [φx (X(t), e(t)) − φx (X(t), α(t))]ζ(t) ,
α
where
e,i
x
e = ζ xe,i (t) = ∂X (t) .
ζ(t)
∂x
By virtue of Theorem 2.27, ζ(t) = ∂X(t)/∂x is mean square continuous.
Hence it follows that
2
e − ζ(t)) ≤ E1/2 |φx (X(t), α(t))|2 E1/2 ζ(t)
E φx (X(t), α(t))(ζ(t) e − ζ(t)
→ 0 as x
e → x.
e
E [φx (X(t), e(t)) − φx (X(t), α(t))]ζ(t)
α
2
e 2
≤ E1/2 φx (X(t), e(t)) − φx (X(t), α(t)) E1/2 |ζ(t)|
α
2
e
φx (X(t), e(t)) − φx (X(t), α(t))
α 2
1/2
≤ KE |e
x − x|
e−x
x
→ 0 as x
e → x.
∂ 2 u(t, x, i) γ
≤ K(1 + |x| ).
∂x2
Theorem 7.10. Assume the conditions of Theorem 7.9. Then the function
u defined in (7.24) is continuously differentiable with respect to the variable
t. Moveover, u satisfies the system of Kolmogorov backward equations
∂u(t, x, i)
= Lu(t, x, i), (t, x, i) ∈ (0, T ] × Rr × M, (7.30)
∂t
with initial condition
Proof. First note that by virtue of Proposition 2.4, the process (X(t), α(t))
is càdlàg. Hence the initial condition (7.31) follows from the continuity of
φ. We divide the rest of the proof into several steps.
Step 1. For fixed (x, i) ∈ Rr × M, u(t, x, i) is absolutely continuous with
respect to t ∈ [0, T ]. In fact, for any 0 ≤ s ≤ t ≤ T , we have from Dynkin’s
formula that
Step 2. For any h > 0, we have from the strong Markov property that
Now let g(x, i) := u(t, x, i). Then Theorem 7.9 implies that g(·, i) ∈ C 2 for
each i ∈ M and for some K > 0 and γ0 > 0,
γ
Dxβ g(x, i) ≤ K(1 + |x| 0 ), i ∈ M.
Using the same argument as in the proof of [47, Theorem 5.6.1], we can
show that
Z h
1
Ex,i Lg(X(v), α(v))dv → Lg(x, i) as h ↓ 0. (7.34)
h 0
Therefore,
Ex,i g(X(h), α(h)) − g(x, i)
lim = Lg(x, i).
h↓0 h
But by the definition of g, we have from (7.33) that
u(t + h, x, i) − u(t, x, i)
lim = Lg(x, i) = Lu(t, x, i). (7.35)
h↓0 h
∂u(t, x, i) ∂u(s, x, i)
−
∂x ∂x
= |Ex,i [φx (X(t), α(t))ζ(t)] − Ex,i [φx (X(s), α(s))ζ(s)]|
≤ Ex,i |φx (X(t), α(t))ζ(t) − φx (X(s), α(s))ζ(s)|
≤ Ex,i |[φx (X(t), α(t)) − φx (X(s), α(s))] ζ(t)|
+Ex,i |φx (X(s), α(s)) [ζ(t) − ζ(s)]|
1/2 2 1/2 2
≤ Ex,i |φx (X(t), α(t)) − φx (X(s), α(s))| Ex,i |ζ(t)|
1/2 2 1/2 2
+Ex,i |φx (X(s), α(s))| Ex,i |ζ(t) − ζ(s)|
As we demonstrated before,
1/2 2
Ex,i |φx (X(s), α(s))| ≤ K.
While Corollary 2.32 implies that ζ(t) is mean square continuous with
respect to t. Hence it follows that
1/2 2
Ex,i |ζ(t) − ζ(s)| → 0 as |t − s| → 0.
Meanwhile,
2
Ex,i |φx (X(t), α(t)) − φx (X(s), α(s))|
2
≤ KEx,i |φx (X(t), α(t)) − φx (X(s), α(t))|
2
+KEx,i |φx (X(s), α(t)) − φx (X(s), α(s))|
:= e1 + e2 .
Using Theorem 2.13 or (2.26) and (2.74), detailed computations show that
Z 1 2
e1 ≤ KEx,i φxx (X(s) + v(X(t) − X(s)), α(t))dv(X(t) − X(s))
0
→ 0 as |t − s| → 0.
To treat the term e2 , we assume without loss of generality that t > s and
200 7. Stability
compute
2
e2 = KEx,i |φx (X(s), α(t)) − φx (X(s), α(s))|
m0 X
X 2
= Ex,i |φx (X(s), j) − φx (X(s), i)| I{α(t)=j} I{α(s)=i}
i=1 j6=i
m0 X
X h i
2
= Ex,i |φx (X(s), j) − φx (X(s), i)| I{α(s)=i} Ex,i [I{α(t)=j} |Fs ]
i=1 j6=i
m0 X
X h
2
= Ex,i |φx (X(s), j) − φx (X(s), i)| I{α(s)=i}
i=1 j6=i
i
×qij (X(s))(t − s) + o(t − s)
≤ K(t − s).
Proof. Once again, for notational simplicity, we present the proof for
X(t) being a real-valued process. By virtue of Theorem 7.9, u(t, x, i) :=
p
E X x,i (t) is twice continuously differentiable with respect to x, except
possibly at x = 0. We need only show that the partial derivatives satisfy
(7.38). To this end, similar to the proofs of Theorems 7.9 and 7.10, we
assume x to be a scalar without loss of generality. By virtue of (7.29),
∂u(t, x, α) x,α p−1 x,α ∂X x,α (t)
= pE |X (t)| sgn(X (t)) . (7.39)
∂x ∂x
Then it follows from (7.6) and (2.74) that
x,α
∂u(t, x, α) x,α p−1 ∂X (t)
≤ KE |X (t)|
∂x ∂x
2
2p−2 ∂X x,α (t)
≤ E1/2 |X x,α (t)| E1/2
∂x
2p−2 Kt 1/2 p−1 κ0 t
≤ K(|x| e ) = K |x| e .
It follows from Lemma 7.11 that the functions V (x, i), i ∈ M are twice
continuously differentiable with respect to x except possibly at x = 0.
The equilibrium point 0 is exponentially p-stable, therefore by the defi-
nition of exponential p-stability, there is a β > 0 such that
Z T
V (x, i) ≤ K|x|p exp(−βu)du ≤ K|x|p .
0
Since 0 is an equilibrium point, |A(x, i)| ≤ K|x|2 and b(x, i)| ≤ K|x|.
Consequently, |L|x|p | ≤ K|x|p . An application of Itô’s lemma to g(x) = |x|p
implies that
Z T
E|X x,i (T )|p − |x|p = E L|X x,i (u)|p du
0Z
T
≥ −K E|X x,i (u)|p du = −KV (x, i).
0
Likewise, we can verify the second part of (7.42). Thus the proof is com-
pleted. 2
We end this section with the following results on linear systems. Assume
that the evolution (7.2) is replaced by
d
X
dX(t) = b(α(t))X(t)dt + σj (α(t))X(t)dwj (t), (7.43)
j=1
where b(i) and σj (i) are r × r constant matrices, and wj (t) are independent
one-dimensional standard Brownian motions for i = 1, 2, . . . , m0 , and j =
1, 2, . . . , d. Then we have the following two theorems.
Theorem 7.13. The equilibrium point x = 0 of system (7.43) together
with (7.3) is exponentially p-stable if and only if for each i ∈ M, there is
a function V (·, i) : Rr 7→ R satisfying equations (7.40) and (7.42) for some
constants ki > 0, i = 1, . . . , 4.
Proof. The proof of sufficiency was contained in [116]. However, the neces-
sity follows from Theorem 7.12 because the coefficients of (7.43) and (7.3)
satisfy the conditions of Theorem 7.12. We omit the details here. 2
7.4 Stability and Instability of Linearized Systems 203
Proof. The proof follows from a crucial observation. In this case, because
(7.43) is linear in X(t), X λx,α (t) = λX x,α (t). Using a similar argument as
that for [83, Lemma 6.4.1], we can conclude the proof; a few details are
omitted. 2
b is irreducible and α
Moreover, Q b(t) is a Markov chain with gen-
b
erator Q.
Proof. (a) We first prove that the equilibrium point x = 0 of the system
given by (7.2) and (7.3) is asymptotically stable in probability if (7.46)
holds. For notational simplicity, define the column vector
with
d
X
µi = λmax b(i) + b0 (i) + σj0 (i)σj (i) .
j=1
Also let β := −πµ. Note that β > 0 by (7.46). It follows from assumption
(A7.4) and Lemma A.12 that the equation
b = µ + β1l
Qc
where 0 < γ < 1 is sufficiently small so that 1 − γci > 0 for each i ∈ M. It
is readily seen that for each i ∈ M, V (·, i) is continuous, nonnegative, and
vanishes only at x = 0. Detailed calculations reveal that for x 6= 0, we have
Next, it follows from condition (A7.4) that when |x| and γ are sufficiently
small,
X cj − c i
qij (x)
1 − γci
j6=i
m0
X X ci (cj − ci )
= qij (x)cj + qij (x) γ (7.51)
j=1
1 − γci
j6=i
m
X 0
where o(1) → 0 as |x| → 0. Hence it follows from (7.49) and (7.51) that
when |x| < r0 with r0 and 0 < γ < 1 sufficiently small, we have
m0
X
LV (x, i) ≤ γ(1 − γci )|x|γ µi − qbij cj + o(1) + O(γ) .
j=1
206 7. Stability
and set
m0
X
δ := −πθ = πi θi < 0.
i=1
As in part (a), assumption (A7.4), the definition of δ, and Lemma A.12
b = θ + δ1l has a solution c = (c1 , c2 , . . . , cm )0 ∈
imply that the equation Qc 0
m0
R and
Xd
θi − qbij cj = −δ > 0, i ∈ M. (7.52)
j=1
and
m0
X d
X
0
πi λmin b(i) + b (i) + σj (i)σj0 (i)
i=1 j=1
d
! (7.55)
1 X 2
− λmax (σj (i) + σj0 (i)) > 0,
2 j=1
Corollary 7.19. Let assumption (A7.4) and (7.56) be valid. Then the
equilibrium point x = 0 is asymptotically stable in probability if
m0
X
σi2
πi b i − < 0,
i=1
2
Remark 7.20. As can be seen from Corollary 7.19, if the continuous com-
ponent of the system is one-dimensional, we obtain a necessary and suffi-
cient condition for stability. One question of particular interest is: Will we
be able to obtain a similar condition for a multidimensional counterpart.
For linear systems of stochastic differential equations with constant coef-
ficients without switching, such a condition was obtained in Khasminskii
[83, pp. 220–224]. The main ingredient is the use of the transformations
y = x/|x| and ln |x|. The result is a sharp necessary and sufficient con-
dition. In Mao, Yin, and Yuan [119], inspired by the approach of [83],
Markov-modulated regime-switching diffusions were considered, and neces-
sary and sufficient conditions were obtained for exponential stability. The
main ingredient is the use of a logarithm transformation technique leading
to the derivation of the so-called Liapunov exponent. Such an approach
can be adopted to treat switching diffusions with state-dependent switch-
ing with no essential difficulty. Because in Chapter 8, we will also examine
a related problem for switched ordinary differential equations (a completely
degenerate switching diffusion with the absence of the diffusion terms), we
will not dwell on it here.
7.5 Examples
Example 7.21. To illustrate Theorem 7.17 and Corollary 7.19, we con-
sider a real-valued process given by
dX(t) = b(X(t), α(t))dt + σ(X(t), α(t))dw(t)
P{α(t + ∆) = j|α(t) = i, X(s), α(s), s ≤ t} = q (X(t))∆ + o(∆),
ij
(7.57)
for j 6= i, where the jump process α(t) has three states and is generated by
−3 − sin x cos x + sin x2 1 + sin x cos x 2 − sin x2
x2 x2
Q(x) = 2 −2 − ,
2 + x2 2 + x2
4 − sin x sin2 x −4 + sin x − sin2 x
7.5 Examples 209
3X(t)
dX(t) = (X(t) − X(t) sin X(t))dt − dw(t), (7.58)
1 + X 2 (t)
2
1
dX(t) = X(t) − X(t) sin X (t) dt + X(t) + X(t) sin X(t) dw(t),
3
(7.59)
1
dX(t) = (4X(t) + X(t) sin X(t)) dt + X(t) − X(t) sin2 X(t) dw(t),
2
(7.60)
switching from one to another according to the movement of the jump
process α(t). It is readily seen that as x → 0, the constants bi , σi2 , i = 1, 2, 3,
as in (7.56) are given by
b1 = 1, b2 = 1, b3 = 4
(7.61)
σ12 = 9, σ22 = 1, σ32 = 1.
only if its linear approximation is stable. Hence we can check that the
equilibrium point x = 0 of (7.58) is stable in probability whereas (7.59) and
(7.60) are unstable in probability. Therefore the jump process α(t) could
be considered as a stabilization factor. Note that similar examples were
demonstrated in [116] under the assumptions that the jump component
α(·) is generated by a constant matrix Q and that the Markov chain α(·) is
independent of the Brownian motion w(·). Note also Examples 4.1 and 4.2
in [48] are also concerned with stability of switching systems. Their result
indicates that if the switching takes place sufficiently fast, the system will
be stable even if the individual mode may be unstable. Essentially, it is
related to singularly perturbed systems. Due to the fast variation, there is
a limit system that is an average with respect to the stationary distribution
of the Markov chain and that is stable. Then if the rate of switching is fast
enough the original system will also be stable. Such an idea was also used
in an earlier paper [18].
To illustrate, we plot a sample path of (7.57) in Figure 7.1. For com-
parison, we also demonstrate the sample paths of (7.58), (7.59), and (7.60)
(without switching) in Figures 7.2–7.4, respectively.
3
α(t)
X(t)
2.5
2
X(t) and α(t)
1.5
0.5
0
0 2 4 6 8 10
t
FIGURE 7.1. A sample path of (7.57) with initial condition (x, α) = (1.5, 1).
2.5
X1(t)
1.5
X1(t)
0.5
0
0 2 4 6 8 10
t
1400
X2(t)
1200
1000
800
X2(t)
600
400
200
0
0 2 4 6 8 10
t
15
x 10
10
X (t)
3
9
6
X3(t)
0
0 2 4 6 8 10
t
et log(|x(t)|) − log(|x(0)|)
Z t X n Xn
x2i (s)
= es 2 r i (α(s)) − aij (α(s))xj (s)
0 i=1 |x(s)|! j=1
1 2x2 (s)
+ 1 − i 2 σi2 (α(s)) ds
2 |x(s)|
Z t Z t X n
x2i (s)
+ es log(|x(s)|)ds + es 2 σi (α(s))dwi (s).
0 0 i=1 |x(s)|
Denote Z t
x2i (s)
Mi (t) = es 2 σi (α(s))dwi (s),
0 |x(s)|
whose quadratic variation is
Z t
e2s 4 2
Mi (t), Mi (t) = 4 xi (s)σi (α(s))ds.
0 |x(s)|
et log(|x(t)|) − log(|x(0)|)
Z t X n Xn
x2i (s)
≤ es 2 r i (α(s)) − aij (α(s))xj (s)
0 i=1 |x(s)| ! j=1
1 2x2i (s) 2
+ 1− 2 σi (α(s)) ds
2 |x(s)|
Z t Z t n
nεe−kγ 2s X x4i (s) 2
+ es log(|x(s)|)ds + e 4 σi (α(s))ds
0 0 2 i=1 |x(s)|
θekγ log k
+
Z t ε Xn
s x2i (s) 2
≤ e log(|x(s)|) + 2 bi (α(s)) + σi (α(s))
0 |x(s)|
X n i=1
− aij (α(s))xj (s)
j=1
n
X
εne−kγ s x4i (s)σi2 (α(s)) θekγ log k
+ e −1 4 + .
i=1
2 |x(s)| ε
214 7. Stability
Note that for any t ∈ [0, kγ], s ∈ [0, t], and (x, α) ∈ Rn+ × M, we have
X n 2 Xn
xi 2
log(|x|) + 2 bi (α) + σ i (α) − a ij (α)x j
i=1 |x| j=1
εnes−kγ x4i σi2 (α)
+ −1
2 |x|4
n
1 X 3
≤ log(|x|) + κ − 2 β xi + K
|x| i=1
β
≤ log(|x|) + κ − √ |x| + K ≤ K.
n
Hence it follows that for all 0 ≤ t ≤ kγ with k ≥ k0 (ω), we have
Z t
θekγ log k
et log(|x(t)|) − log(|x(0)|) ≤ Kes ds +
0 ε
t θekγ log k
= K(e − 1) + .
ε
Thus for (k − 1)γ ≤ t ≤ kγ, we have
θekγ log k
log(|x(t)|) ≤ e−t log(|x(0)|) + K(1 − e−t ) +
εe(k−1)γ
θeγ log k
= e−t log(|x(0)|) + K(1 − e−t ) + ,
ε
and hence it follows that
log(|x(t)|) log(|x(0)|) K(1 − e−t ) θeγ log k
≤ + + .
log t et log t log t ε log((k − 1)γ)
Now let k → ∞ (and so t → ∞) and we obtain
log(|x(t)|) θeγ
lim sup ≤ .
t→∞ log t ε
Finally, by sending γ ↓ 0, ε ↑ 1, and θ ↓ 1, we have
log(|x(t)|)
lim sup ≤ 1,
t→∞ log t
as desired. Thus, the solution x(t) of (1.1) satisfies
log(|x(T )|)
lim sup ≤ 1. (7.63)
T →∞ log T
Furthermore, since
log |x(T )| log |x(T )| log T
lim sup = lim sup lim sup
T →∞ T T →∞ log T T →∞ T
log T
≤ lim sup ,
T →∞ T
7.6 Notes 215
log |x(T )|
lim sup ≤ 0 a.s. (7.64)
T →∞ T
7.6 Notes
The framework of this chapter is based on the paper of Khasminskii, Zhu,
and Yin [92]. Some new results are included, e.g., Theorems 7.9 and 7.10,
where we derived Kolmogorov-type backward equations without assuming
nondegeneracy of the diffusion part. These results are interesting in their
own right.
For brevity, we are not trying to cover every angle of the stability analysis.
For example, sufficient conditions for exponential p-stability can be derived
similar to that of [116], and as a result the verbatim proof for the sufficiency
is omitted. Nevertheless, necessary conditions for stability and stability
under linearization are provided. Finally, we note that using linearization to
infer the stability of the associated nonlinear systems should be interesting
owing to its wide range of applications.
8
Stability of Switching ODEs
8.1 Introduction
The main motivational forces for this chapter are the work of Davis [30]
on piecewise deterministic systems, and the work of Kac and Krasovskii
[79] on stability of randomly switched systems. In recent years, growing
attention has been drawn to deterministic dynamic systems formulated as
differential equations modulated by a random switching process. This is
because of the increasing demands for modeling large-scale and complex
systems, designing optimal controls, and carrying out optimization tasks.
In this chapter, we consider stability of such hybrid systems modulated by a
random-switching process, which are “equivalent” to a number of ordinary
differential equations coupled by a switching or jump process.
In this chapter, for random-switching systems, we first obtain sufficient
conditions for stability and instability. Our approach leads to a necessary
and sufficient condition for systems whose continuous component is one-
dimensional. For multidimensional systems, our conditions involve the use
of minimal and maximal eigenvalues of appropriate matrices. The differ-
ence of maximal and minimal eigenvalues results in a gap for stability and
instability. To close this gap, we introduce a logarithm transformation lead-
ing to the continuous component taking values in the unit sphere. This in
turn, enables us to obtain necessary and sufficient conditions for stability.
The essence is the utilization of the so-called Liapunov exponent.
Because the systems we are interested in have continuous components
(representing continuous dynamics) as well as discrete components (repre-
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 217
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_8,
© Springer Science + Business Media, LLC 2010
218 8. Stability of Switching ODEs
For further references on the associated operator (or generator) of the hy-
brid system (8.1)–(8.2), we refer the reader to Chapter 2 of this book; see
also [79] and [150].
To proceed, we need conditions regarding the smoothness and growth of
the functions involved, and the condition that 0 is an equilibrium point of
the dynamic system. We assume the following hypotheses throughout the
chapter.
p
E |X x,α (τδ (t))|
Z τδ (t)
p p−2
= |x| + E p |X x,α (s)| X x,α (s), f (X x,α (s), αx,α (s)) ds
0 Z τδ (t)
p p−1
≤ |x| + pm0 K0 E |X x,α (s)| (1 + |X x,α (s)|)ds
0
Z τδ (t)
p
≤ |x|p + 2pm0 K0 E |X x,α (s)| ds
Z τδ (t) 0
p−1
+ pm0 K0 E |X x,α (s)| I{δ≤|X x,α (s)|<1} ds
0 Z τδ (t)
p x,α p 1
≤ |x| + 2pm0 K0 E |X (s)| + ds.
0 2
p
Applying Gronwall’s inequality to [E |X x,α (τδ (t))| + (1/2)] leads to
x,α p 1 p 1
E |X (τδ (t))| + ≤ |x| + exp(2pm0 K0 t), (8.8)
2 2
or equivalently,
p 1 1
E |X x,α (τδ (t))| ≤ |x|p + exp(2pm0 K0 t) −
2 2
(8.9)
1
≤ |x|p + exp(2pm0 K0 t).
2
Note that we have from (8.6) that
where K is some positive constant, then the conclusion Theorem 8.3 can
be strengthened to the following. For any β ∈ R and any (x, α) ∈ Rr × M
with x 6= 0, we have
E|X x,α (t)|β ≤ |x|β eρt , (8.11)
where ρ is a constant depending only on β, m0 , and the constant K given
in (8.10).
In fact, by virtue of (8.10), we obtain by a slight modification of the
argument in the proof of Theorem 8.3 that
Proof. Choose r0 > 0 such that the ball Br0 = {x ∈ Rr : |x| < r0 } and
its boundary ∂Br0 = {x ∈ Rr : |x| = r0 } are contained in D. Set Vr0 :=
inf {V (x, i) : x ∈ D \ Br0 , i ∈ M}. Then Vr0 > 0 by assumption (i). Next,
assumption (ii) leads to
Z τr0 ∧t
Ex,i V (X(t ∧ τr0 ), α(t ∧ τr0 )) = V (x, i) + Ex,i LV (X(s), α(s))ds
0
≤ V (x, i),
where (x, i) ∈ Br0 × M and τr0 is the first exit time from Br0 ; that is,
τr0 := {t ≥ 0 : |x(t)| ≥ r0 }. Because V is nonnegative, we further have
h i
Vr0 P {τr0 ≤ t} ≤ Ex,i V (X(τr0 ), α(τr0 ))I{τr ≤t} ≤ V (x, i).
0
Note that τr0 ≤ t if and only if sup0≤u≤t |x(u)| > r0 . Therefore it follows
that
V (x, i)
Px,i sup |X(u)| > r0 ≤ .
0≤u≤t V r0
Letting t → ∞, we obtain from assumption (i) that
V (x, i)
Px,i sup |X(t)| > r0 ≤ .
0≤t V r0
Finally, the desired conclusion follows from the assumptions that V (0, i) =
0 and V (·, i) is continuous for each i ∈ M. 2
Proposition 8.6. Assume the conditions of Proposition 8.5. Suppose also
that for each i ∈ M, the function V (·, i) satisfies
where % > 0 and κ(%) is a positive constant. Then the equilibrium point
x = 0 is asymptotically stable in probability.
where Bδ := {x ∈ Rr : |x| < δ}. Now fix some (x, α) ∈ (Bδ − {0}) × M and
let %1 > % > 0 be arbitrary satisfying %1 < |x|. Define
τ% := {t ≥ 0 : |X(t)| ≤ %} ,
τr0 := {t ≥ 0 : |X(t)| ≥ r0 } .
Note that (8.14) implies that Px,α {τr0 < ∞} ≤ ε/2. Hence it follows that
Px,α {τ% < ∞} ≥ Px,α {τ% ∧ τr0 < ∞} − Px,α {τr0 < ∞}
ε (8.16)
≥1− .
2
Now let
τ%1 := {t ≥ τ% : |X(t)| ≥ %1 } .
We use the convention that inf ∅ = ∞. Then for any t > 0, we have
Vb% ε
Px,α {τ%1 < t} ≤ ≤ .
V %1 2
Letting t → ∞, we obtain
ε
Px,α {τ%1 < ∞} ≤ . (8.19)
2
Finally, it follows from (8.16) and (8.19) that
Px,α {τ% < ∞, τ%1 = ∞} ≥ Px,α {τ% < ∞} − Px,α {τ%1 < ∞}
ε ε
≥ 1 − − = 1 − ε.
2 2
This implies that
Px,α lim sup |X(t)| ≤ %1 ≥ 1 − ε.
t→∞
1
µi = λmax (GAi G−1 + G−1 A0i G),
2
where G is as in (8.24). Also let
m0
X
β := −πµ = − πi µ i .
i=1
Note that β > 0 by (8.24). By virtue of condition (A8.4) and Lemma A.12,
the equation
b = µ + β1l
Qc
has a solution c = (c1 , c2 , . . . , cm0 )0 ∈ Rm0 . Thus we have
m0
X
µi − qbij cj = −β, i ∈ M. (8.26)
j=1
where 0 < γ < 1 is sufficiently small so that 1 − γci > 0 for each i ∈ M. It
is readily seen that for each i ∈ M, V (·, i) is continuous, nonnegative, and
vanishes only at x = 0. In addition, since γ > 0 and 1 − γci > 0, we have
LV (x, i) ≤ γ(1 − γci )(λmin (G2 ))γ/2 |x|γ (−β + o(1) + O(γ))
≤ −κ(ε) < 0,
where −1 < γ < 0 is sufficiently small so that 1 − γci > 0 for each i ∈ M.
Similar to the argument in (8.27), we can verify that V (·, i) satisfies (8.21)
for each i ∈ M. Detailed computations as in part (a) show that for any
sufficiently small 0 < ε < r0 ,
Sr = {y ∈ Rr : |y| = 1}.
X 0 (t)Ẋ(t)
ż(t) = . (8.36)
|X(t)|2
In what follows, we deal with a case that the Markov chain is fast varying.
It acts as a “noise,” whereas the X(t) is slowly varying. In the end, the
noise is averaged out and replaced by its stationary distribution. To put
this in a mathematical form, we suppose that there is a small parameter
ε > 0 and
Q0
Q = Qε = ,
ε
232 8. Stability of Switching ODEs
Note that in the above, ζei (t) is a piecewise constant approximation of ζi (t)
with the interpolation intervals [tk , tk+1 ), k = 0, 1, . . . , N .
Lemma 8.11. Suppose that Q0 is irreducible and that
ε = o(T −∆ ) as T → ∞. (8.38)
Then
m0 Z T
1 X X 0 (t)Ai X(t)
[I{αε (t)=i} − πi ]dt → 0 in probability as T → ∞.
T i=1 0 |X(t)|2
Proof. The assertion of Lemma 8.11 will follow immediately if we can show
that for each i ∈ M,
Z
1 T 2
E ζi (t)Iiε (t)dt → 0 as T → ∞.
T 0
8.4 A Sharper Result 233
by the choice of δ∆ .
By means of (8.39) and (8.41), it remains to show that the last term of
(8.39) converges to 0 as T → ∞. For any 0 ≤ t ≤ T , define
hZ t i2
hi (t) = E ζei (s)Iiε (s)ds . (8.42)
0
Then Z t
dhi (t)
=2 E[ζei (s)Iiε (s)ζei (t)Iiε (t)]ds. (8.43)
dt 0
For 0 ≤ t ≤ t2 ,
Z t
E[ζei (s)Iiε (s)ζei (t)Iiε (t)]ds
0Z t2
≤ E1/2 |ζei (s)|2 E1/2 |ζei (t)|2 ds ≤ Kt2 = O(T −∆ ).
0
234 8. Stability of Switching ODEs
Furthermore, Z t
E[ζei (s)Iiε (s)ζei (t)Iiε (t)]ds
tk−1
Z t
≤ E1/2 |ζei (s)|2 E1/2 |ζei (t)|2 ds
tk−1
The next to the last inequality follows from the asymptotic expansions
obtained in, for example, [176, Lemma 5.1].
Using (8.44) and (8.45), we obtain that
dhi (t) T −∆
sup ≤ O(T )O ε + exp − for each i ∈ M.
0≤t≤T dt ε
Because hi (0) = 0, we conclude
T −∆
|hi (T )| ≤ O(T 2 )O ε + exp − .
ε
Thus hi (T )/T 2 → 0 as T → ∞ as long as ε = o(T −∆ ). The lemma then
follows. 2
By virtue of Lemma 8.11, we need only consider the last term in (8.37).
To this end, for each i ∈ M, there is a λi ∈ R such that
Z Z
1 T X 0 (t)Ai X(t) 1 T 0
dt = Y (t)Ai Y (t)dt
T 0 |X(t)|2 T 0 (8.46)
→ λi in probability as T → ∞,
8.4 A Sharper Result 235
where the existence of the limit follows from a slight modification of [61,
p. 344, Theorem 6]. The number λi above is precisely the average of
Y 0 (t)Ai Y (t), known as the mean value of Y 0 (·)Ai Y (·) and denoted by
M [Y 0 Ai Y ] in the literature of almost periodic functions [61, Appendix].
Define
m0
X
λ= πi λ i . (8.47)
i=1
It follows from Lemma 8.11, equations (8.37), (8.46), and (8.47) that
Z T
1 |X(T )| 1 X 0 (t)A(αε (t))X(t)
ln = dt → λ as T → ∞. (8.48)
T |X(0)| T 0 |X(t)|2
1 |X(T )|
ln = λ + o(1), (8.49)
T |X(0)|
Likewise, if λ > 0, we can find 0 < λ2 < λ such that λ + o(1) ≥ λ2 , which
in turn, implies that
where (A(i)y)l denotes the lth component of A(i)y and Q0 is the transpose
of Q. Moreover, for any Γ ⊂ Rr and i ∈ M,
Z
P(y(t) ∈ Γ, α(t) = i) = p(y, i, t)dx.
Γ
Note that the process (Y (t), α(t)) is on the compact set Sr × M. The
existence of the invariant density of (Y (t), α(t)) is thus guaranteed owing
to the compactness. We further assume the uniqueness of the invariant
density, and denote it by (µ(y, i) : i = 1, . . . , m0 ). Then µ(y, i) satisfies
Xm0 Z
∗
L µ(y, i) = 0, µ(y, i)dy = 1.
i=1 Rr
Under this condition, as in the proof similar to [74, Theorem 1], it can be
shown that there is a κ0 > 0 such that
|p(y, i, t) − µ(y, i)| ≤ K exp(−κ0 t). (8.54)
Similar to (8.37), we obtain
Z
z(T ) − z(0) 1 T 0
= Y (t)A(α(t))Y (t)dt
T T Z0
T
1
= Y 0 (t)A(α(t))Y (t)dt.
T 0
238 8. Stability of Switching ODEs
Redefine λ as
m0 Z
X
λ= y 0 A(i)yµ(y, i)dy,
i=1 Rr
0
E (Y 0 (t)A(α(t))Y (t) − λ) (Y 0 (s)A(α(s))Y (s) − λ)
≤ K exp(−κ0 (t − s)).
X 0 m Z
ln |X(t)|
lim =λ= y 0 A(i)yµ(y, i)dy w.p.1. (8.56)
t→∞ t i=1 Sr
dϕ
γ(ϕ, i) = = −[sin ϕ cos ϕa11 (i) + sin2 ϕa12 (i)]
dt (8.57)
+[cos2 ϕa21 (i) + a22 (i) cos ϕ sin ϕ].
∂V
LV (ϕ, i) = γ(ϕ, i) + QV (ϕ, ·)(i), i ∈ M. (8.58)
∂ϕ
Z ϕ h(ϕ)
h(s) (8.64)
q1 c ds + c − c
γ1 (s)
γ2 = 0 ,
h(ϕ)
8.6 Examples 241
8.6 Examples
For simplicity, by a slightly abuse of notation, we call a system that is linear
with respect to the continuous state a linear system or hybrid linear system.
To illustrate, we provide several examples in this section. In addition, we
demonstrate results that are associated with the well-known Hartman–
Grobman theorem for random switching systems.
Example 8.15. Consider a system (8.35) (linear in the variable x) with
the following specifications. The Markov chain α(t) has two states and is
generated by
−3 3
Q= ,
1 −1
242 8. Stability of Switching ODEs
and
−1 2 −3 −1
A1 = A(1) = , A2 = A(2) = .
0 2 1 −2
Thus associated with the hybrid system (8.35), there are two ordinary
differential equations
Ẋ(t) = A1 X(t), (8.67)
and
Ẋ(t) = A2 X(t), (8.68)
switching back and forth from one to another according to the movement
of the jump process α(t). It is readily seen that the eigenvalues of A1 are
−1 and 2. Hence the motion of (8.67) is unstable. Similarly, by computing
the eigenvalues of A2 , the motion of (8.68) is asymptotically stable.
Next we use Corollary 8.9 to show that the hybrid system (8.35) is asymp-
totically stable owing to the presence of the Markov chain α(t). That is,
the Markov chain becomes a stabilizing factor. The stationary distribution
of the Markov chain α(t) is π = (0.25, 0.75), which is obtained by solving
the system of equations πQ = 0 and π1l = 1. The maximum eigenvalues of
A1 + A01 and A2 + A02 are
λmax (A1 + A01 ) = 4.6056, λmax (A2 + A02 ) = −4,
respectively. This yields that
π1 λmax (A1 + A01 ) + π2 λmax (A2 + A02 ) = −1.8486 < 0.
Therefore, we conclude from Corollary 8.9 that the hybrid system (8.35) is
asymptotically stable. The phase portrait Figure 8.1(a) confirms our find-
ings. It is interesting to note the dynamic movements and the interactions
of the continuous and discrete components. To see the difference between
hybrid systems and ordinary differential equations, we also present the
phase portraits of (8.67) and (8.68) in Figure 8.1(b).
Q(x) = Q(x1 , x2 )
−2 − sin x1 cos x21 − sin x22 2 + sin x1 cos(x21 ) + sin(x22 )
= ,
1 − 0.5 sin(x1 x2 ) −1 + 0.5 sin(x1 x2 )
8.6 Examples 243
−1
−0.5 0 0.5 1 1.5 2 2.5 3 3.5
12
10
0
−1 0 1 2 3 4 5 6 7 8
FIGURE 8.1. Comparisons of switching system and the associated ordinary dif-
ferential equations.
244 8. Stability of Switching ODEs
and
0 0 −1 2
A1 = A(1) = , A2 = A(2) = .
3 0 −2 −1
Note that the distinct feature of this example compared with the last one
is that the Q matrix is x dependent, which satisfies the approximation
condition posed in Section 8.3. Associated with the hybrid system (8.35),
there are two ordinary differential equations
and
Ẋ(t) = A2 X(t) (8.71)
switching from one to another according to the movement of the jump
process α(t). Solving (8.70), we obtain
b(t) is π =
and hence the stationary distribution of the Markov chain α
(1/3, 2/3). The maximum eigenvalues of A1 + A01 and A2 + A02 are
Therefore, we conclude from Corollary 8.9 that the hybrid system (8.69)
is asymptotically stable. The phase portrait in Figure 8.2(a) confirms our
results. To delineate the difference between hybrid systems and ordinary
differential equations, we also present the phase portraits of (8.70) and
(8.71) in Figure 8.2(b). The phase portraits reveal the interface of continu-
ous and discrete components. They illustrate the hybrid characteristics in
an illuminating way.
8.6 Examples 245
−2
−4
−2 −1 0 1 2 3 4 5 6
12
10
−2
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2
FIGURE 8.2. Comparisons of switching system and the associated ordinary dif-
ferential equations.
246 8. Stability of Switching ODEs
Q(x1 , x2 ) =
sin x2 − 2 − cos x1 1 − sin x2 1 + cos x1
1 − sin(x21 x2 ) sin(x21 x2 ) − 1 − cos2 (x1 x2 ) cos2 (x1 x2 ) .
0 3 + sin x1 + sin x2 cos x2 −3 − sin x1 − sin x2 cos x2
b and Ai , i ∈ M in
Note that the matrices Q, assumption (A8.4) can be
obtained as follows
−3 1 2
b = 1 −1
Q 0 (8.73)
0 3 −3
and
1 0 −2 1 0 1
A1 = , A2 = , and A3 = .
0 −2 −1 −1 −1 0
(8.74)
For a system of differential equations without switching, the well-known
Hartman–Grobman theorem holds. It indicates that for hyperbolic equilib-
ria, a linear system arising from approximation is topologically equivalent
to the associated nonlinear system, whereas a system with a center is not.
Here we demonstrate that this phenomenon changes a little. We show that
the nonlinear system and its approximation could be equivalent even for
8.7 Notes 247
and compute the maximum eigenvalues of the matrices GAi G−1 + G−1 A0i G
with i = 1, 2, 3,
11
λmax (GA1 G−1 + G−1 A01 G) = ,
4
11
λmax (GA2 G−1 + G−1 A02 G) = − ,
4
and
3
λmax (GA3 G−1 + G−1 A03 G) = .
2
Thus we obtain
3
X 27
πi λmax (GAi G−1 + G−1 A0i G) = − < 0. (8.75)
i=1
28
Hence Theorem 8.8 implies that the hybrid system (8.72) is asymptotically
stable.
It is interesting to note that the linear approximation
Ẋ(t) = A(b
α(t))X(t), (8.76)
8.7 Notes
Consideration of switching ordinary differential equations of the form (8.1)
stems from a wide variety of applications including control, optimization,
estimation, and tracking. For example, in [168], with motivation of using
stochastic recursive algorithms for tracking Markovian parameters such
248 8. Stability of Switching ODEs
0.5
−0.5
−0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4
1.2
0.8
0.6
0.4
0.2
−0.2
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
tions for stability and instability with the aid of nonquadratic Liapunov
functions. Compared with the conditions in [79], our conditions in Theo-
rem 8.8, Corollary 8.9, and Corollary 8.10 are simpler and easier to verify.
Moreover, our results can be applied to more general models.
9
Invariance Principles
9.1 Introduction
In the previous two chapters, we have studied stability of switching dif-
fusions and random switching ordinary differential equations. Continuing
our effort, this chapter is concerned with invariance principles of switching
diffusion processes. This chapter together with the previous two chapters
delineates long-time behavior and gives a complete picture of the switching-
diffusion processes under consideration.
The rest of the chapter is arranged as follows. Section 9.2 begins with the
formulation of the problem. Section 9.3 is devoted to the invariance princi-
ples using sample paths and kernels of the associated Liapunov functions.
Here, Liapunov function-type criteria are obtained first. Then linear (in x)
systems are treated. Section 9.4 switches gears to examine the invariance
using the associated measures. Finally, a few more remarks are made in
Section 9.5 to conclude the chapter.
9.2 Formulation
As in the previous chapters, we use z 0 to denote the transpose of z ∈ R`1 ×`2
with `i ≥ 1, whereas R`×1 is simply written as R` ; 1l = (1, 1, . . . , 1)0 ∈ Rm0
is a column vector with all entries being 1; the Euclidean norm for a row or a
column vector x is denoted by |x|. As usual, I denotes the identity matrix
with suitable dimension. For a matrix A, its trace norm is denoted by
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 251
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_9,
© Springer Science + Business Media, LLC 2010
252 9. Invariance Principles
p
|A| = tr(A0 A). If a matrix A is real and symmetric, we use λmax (A) and
λmin (A) to denote the maximal and minimal eigenvalues of A, respectively,
and set ρ(A) := max {|λmax (A)| , |λmin (A)|}. When B is a set, IB (·) denotes
the indicator function of B.
We work with (Ω, F, P), a complete probability space, and consider a
two-component Markov process (X(t), α(t)), where X(·) is a continuous
component taking values in Rr and α(·) is a jump component taking values
in a finite set M = {1, 2, . . . , m0 }. The process (X(t), α(t)) has a generator
L given as follows. For each i ∈ M and any twice continuously differentiable
function g(·, i),
r r
1 X ∂ 2 g(x, i) X ∂g(x, i)
Lg(x, i) = ajk (x, i) + bj (x, i) + Q(x)g(x, ·)(i),
2 ∂xj ∂xk j=1
∂xj
j,k=1
1
= tr(a(x, i)∇2 g(x, i)) + b0 (x, i)∇g(x, i) + Q(x)g(x, ·)(i),
2
(9.1)
where x ∈ Rr , and Q(x) = (qij (x))Pis an m0 × m0 matrix depending on x
satisfying qij (x) ≥ 0 for i 6= j and j∈M qij (x) = 0 for each i ∈ M,
X
Q(x)g(x, ·)(i) = qij (x)g(x, j)
j∈M
X
= qij (x)(g(x, j) − g(x, i)), i ∈ M,
j∈M,j6=i
and ∇g(·, i) and ∇2 g(·, i) denote the gradient and Hessian of g(·, i), respec-
tively.
The process (X(t), α(t)) can be described by
and
P{α(t + ∆t) = j|α(t) = i, (X(s), α(s)), s ≤ t}
(9.3)
= qij (X(t))∆t + o(∆t), i 6= j,
where w(t) is a d-dimensional standard Brownian motion, b(·, ·) : Rr ×M 7→
Rr , and σ(·, ·) : Rr × M 7→ Rr×d satisfies σ(x, i)σ 0 (x, i) = a(x, i).
Throughout the chapter, we assume that both b(·, i) and σ(·, i) satisfy the
usual Lipschitz condition and linear growth condition for each i ∈ M and
that Q(·) is bounded and continuous. Under these conditions, the system
(9.2)–(9.3) has a unique strong solution; see Chapter 2 and also [77] or
[172] for details. From time to time, we often wish to emphasize the initial
data (X(0), α(0)) = (x, α) dependence of the solution of (9.2)–(9.3), which
is denoted by (X x,α (t), αx,α (t)).
9.3 Invariance (I): A Sample Path Approach 253
It is easy to verify that for any (x, i), (y, j), and (z, l),
(i) d((x, i), (y, j)) ≥ 0 and d((x, i), (y, j)) = 0 if and only if (x, i) = (y, j),
(ii) d((x, i), (y, j)) = d((y, j), (x, i)), and
(iii) d((x, i), (y, j)) ≤ d((x, i), (z, l)) + d((z, l), (y, j)).
Thus d is a distance function of Rr × M. Also if U is a subset of Rr × M,
we define
d((x, i), U ) = inf {d((x, i), (y, j)) : (y, j) ∈ U } . (9.7)
Let M be endowed with the trivial topology. As usual, we denote by d(x, D)
the distance between x ∈ Rr and D ⊂ Rr ; that is,
In addition, we use Px,α and Ex,α to denote the probability and expectation
with (X(0), α(0)) = (x, α), respectively. Then for a fixed U ∈ Rr × M, the
function d(·, U ) is continuous.
254 9. Invariance Principles
then any solution with initial condition (x, i) satisfying x 6= 0 will never
reach the origin almost surely, in other words, the set (Rr − {0}) × M is
invariant with respect to the solutions of (9.2)–(9.5).
Using the terminologies in [47, 83], we recall the definitions of stability
and asymptotic stability of a set. Then general results in terms of the
Liapunov function are provided.
(i) stable in probability if for any ε > 0 and ρ > 0, there is a δ > 0 such
that
Px,i sup d((X(t), α(t)), K) < ρ ≥ 1−ε, whenever d((x, i), K) < δ;
t≥0
is nonempty and bounded, and that for each α ∈ M, V (·, α) is twice con-
tinuously differentiable with respect to x, and
Then
(i) Ker(V ) is an invariant set for the process (X(t), α(t)), and
is a martingale measure with p(dt, dz) being the Poisson random measure
with intensity dt × m(dz) as in (9.5). Taking expectations on both sides of
(9.11) (using a sequence of stopping times and Fatou’s lemma, if necessary
as the argument in Theorem 3.14), it follows from (9.10) that
The last equality above holds because (x0 , i0 ) ∈ Ker(V ). But V is non-
negative, so we must have V (X(t), α(t)) = 0 a.s. for any t ≥ 0. Then we
have ( )
Px0 ,i0 sup V (X(tn ), α(tn )) = 0 = 1,
tn ∈Q+
256 9. Invariance Principles
That is,
Let the initial condition (x, i) ∈ Uδ − Ker(V ) and τ be the first exit time
of the process from Uδ . That is,
where
Z t∧τ
M (t ∧ τ ) = ∇V (X(s), α(s)), σ(X(s), α(s))dw(s)
0ZZ t∧τ
+ V (X(s), i + h(X(s), α(s−), z))
0 R
−V (X(s), α(s)) µ(ds, dz).
For notational simplicity, denote (ξ, `) = (X(τ ), α(τ )). We claim that
Suppose (9.15) were not true. Then there would exist a sequence {xn } ⊂ D
such that limn→∞ V (xn , `) = 0. Because {xn } is bounded, there exists a
subsequence {xnk } such that xnk → x e. Thus by the continuity of V (·, `),
we have
x, `) = lim V (xnk , `) = 0.
V (e
k→∞
where
and τε,r0 is the first exit time from Uε,r0 ; that is,
Proof. Fix any (x, i) ∈ Uε,r0 . By virtue of generalized Itô’s lemma, we have
that for any t ≥ 0,
Theorem 9.5. Assume that there exists a function V satisfying the con-
ditions of Lemma 9.4. Then Ker(V ) is an invariant set for the process
(X(t), α(t)) and Ker(V ) is asymptotically stable in probability.
Proof. Motivated by Mao and Yuan [120, Theorem 5.36], we use similar
ideas and proceed as follows. By virtue of Theorem 9.3, we know that
Ker(V ) is an invariant set for the process (X(t), α(t)) and that Ker(V ) is
stable in probability. Hence it remains to show that
n o
Px,i lim d((X(t), α(t)), Ker(V )) = 0 → 1 as d((x, i), Ker(V )) → 0.
t→∞
Because Ker(V ) is stable in probability, for any ε > 0 and any θ > 0, there
exists some δ > 0 (without loss of generality, we may assume that δ < θ)
such that
ε
Px,i sup d((X(t), α(t)), Ker(V )) < θ ≥ 1 − , (9.17)
t≥0 2
for any (x, i) ∈ Uδ , where Uδ is defined in (9.12). Now fix any (x, α) ∈
Uδ − Ker(V ) and let ρ > 0 be arbitrary satisfying 0 < ρ < d((x, i), Ker(V ))
and choose some % ∈ (0, ρ). Define
where τ%,θ is the first exit time from U%,θ that is defined as
But (9.17) implies that Px,α {τθ < ∞} ≤ ε/2. Note also
Now let
τρ := inf {t ≥ τ% : d((X(t), α(t)), Ker(V )) ≥ ρ} .
We have used the convention that inf {∅} = ∞. For any t ≥ 0, we apply
generalized Itô’s lemma and (9.16) to obtain
Ex,α V (X(τρ ∧ t), α(τρ ∧ t)) ≤ Ex,α V (X(τ% ∧ t), α(τ% ∧ t))
Z τρ ∧t
+ Ex,α LV (X(s), α(s))ds (9.20)
τ% ∧t
where Vb% := sup {V (y, j) : d((y, j), Ker(V )) = %}. Note that τρ < t implies
τ% < t. Hence we further have
h i
Vb% ≥ Ex,α I{τ% <t} I{τρ <t} V (X(τρ ∧ t), α(τρ ∧ t))
h i
= Ex,α I{τρ <t} V (X(τρ ∧ t), α(τρ ∧ t))
h i
= Ex,α I{τρ <t} V (X(τρ ), α(τρ ))
≥ Vρ Px,α {τρ < t} ,
where Vρ := inf {V (y, j) : ρ ≤ d((y, j), Ker(V )) ≤ ρe}, with ρe > 0 being
some constant. Recall that we showed in the proof of Theorem 9.3 that
Vρ > 0. Because V is continuous, we may choose % sufficiently small so
that
Vb% ε
Px,α {τρ < t} ≤ ≤ .
Vρ 2
9.3 Invariance (I): A Sample Path Approach 261
Letting t → ∞, we obtain
ε
Px,α {τρ < ∞} ≤ . (9.22)
2
Finally, it follows from (9.19) and (9.22) that
r
for any (x, α) ∈ R × M.
Let ε > 0 and fix any (x, α) ∈ Rr × M. Then (9.23) implies that there
exists some positive constant β > (R + 2) ∨ d((x, α), Ker(V )) such that
2V (x, α)
inf {V (y, j) : |y| ≥ β, j ∈ M} ≥ . (9.26)
ε
262 9. Invariance Principles
Define
τβ := inf {t ≥ 0 : d((X(t), α(t)), Ker(V )) ≥ 2β} .
For any t ≥ 0, we have by virtue of generalized Itô’s lemma and (9.16) that
We claim that |X(τβ )| ≥ β. If this were not true, it would follow from
(9.25) that for any (y, j) ∈ Ker(V ),
where in the last inequality above, we used the fact that β > R + 2. Then
we have d((X(τβ ), α(τβ )), Ker(V )) ≤ 2β − 1 < 2β. This contradicts the
definition of τβ . Thus we must have |X(τβ )| ≥ β. Then it follows from
(9.27) that
h i
V (x, α) ≥ Ex,α V (X(τβ ), α(τβ ))I{τβ <t}
≥ inf {V (y, j) : |y| ≥ β, j ∈ M} Px,α {τβ < t} ,
c ) 6= ∅,
(ii) Ker(W
c )) = 0 a.s., and
(iii) lim d((X x,α (t), αx,α (t)), Ker(W
t→∞
c ) = {0} × M, then
(iv) if moreover, Ker(W
Proof. This theorem can be proved using the arguments in [117, Theorem
2.1] although some modifications are needed. 2
where b(i), σj (i) are r × r constant matrices and wj (t) are independent
one-dimensional standard Brownian motions for i = 1, 2, . . . , m0 and j =
1, 2, . . . , d.
Note that 0 is an equilibrium point for the system given by (9.30) and
(9.3). As we indicated earlier, it was shown in [92, 116] that the set Rr × M
is invariant with respect to the process (X(t), α(t)).
Theorem 9.8. Assume that the discrete component α(·) is ergodic with
constant generator Q = (qij ) and invariant distribution π = (π1 , . . . , πm0 ) ∈
R1×m0 . Then the equilibrium point x = 0 of the system given by (9.30) and
(9.3)
(i) is asymptotically stable with probability one if
m
X d
X
0
πi λmax b(i) + b (i) + σj (i)σj0 (i) < 0, (9.31)
i=1 j=1
Xm Xd 1 2
0 0 0
πi λmin b(i) + b (i) + σj (i)σj (i) − ρ(σj (i) + σj (i)) > 0.
i=1 j=1
2
(9.32)
264 9. Invariance Principles
Proof. We need only prove assertion (i), because assertion (ii) was con-
sidered in Theorem 7.17 in Chapter 7. For notational simplicity, define the
column vector
µ = (µ1 , µ2 , . . . , µm0 )0 ∈ Rm0
with
Xd
1
µi = λmax b(i) + b0 (i) + σj (i)σj0 (i) .
2 j=1
where 0 < γ < 1 is sufficiently small so that 1 − γci > 0 for each i ∈ M.
It is readily seen that for each i ∈ M, V (·, i) is continuous, nonnegative,
vanishes only at x = 0, and satisfies (9.29). Detailed calculations as in the
proof of Theorem 7.17 in Chapter 7 reveal that for x 6= 0, we have
n x0 b(i)x X cj − c i
LV (x, i) = γ(1 − γci )|x|γ 2
− qij
|x| 1 − γci
j6=i
(9.34)
d
1 X x σj (i)σj (i)x
0 0
(x0 σj0 (i)x)2 o
+ + (γ − 2) .
2 j=1 |x|2 |x|4
Note that
d
x0 b(i)x 1 X x0 σj0 (i)σj (i)x
+
|x|2 2 j=1 |x|2
d
x0 (b0 (i) + b(i))x 1 X x0 σj0 (i)σj (i)x
= + (9.35)
2|x|2 2 j=1 |x|2
X d
1 0 0
≤ λmax b(i) + b(i) + σj (i)σj (i) = µi .
2 j=1
Hence it follows from (9.33)–(9.36) that when 0 < γ < 1 sufficiently small,
we have
n m
X o
LV (x, i) ≤ γ(1 − γci )|x|γ µi − qbij cj + O(γ)
j=1
γ
= γ(1 − γci )|x| − β + O(γ)
β γ c (x, i),
≤ − γ(1 − γci ) |x| := −W
2
Note that Ker(Wc ) = {0} × M. Thus, we conclude from Theorem 9.7 that
the equilibrium point x = 0 is asymptotically stable with probability 1. 2
[97]. We recall that the system is stable with respect to (O1 , O2 , ρ) (or
stable relative to (O1 , O2 , ρ)) if for each i ∈ M and x ∈ O1 implies that
Px,i (X(t) ∈ O2 for all t < ∞) ≤ ρ.
Proposition 9.9. Let Y (t) = (X(t), α(t)) be a switching diffusion process
on Rr × M. For each i ∈ M, let V (x, i) be a nonnegative continuous
function, Om = {x : V (x, i) < m} be a bounded set, and
τ = inf{t : X(t) 6∈ Om }. (9.37)
Suppose that for each i ∈ M, LV (x, i) ≤ −k(x) ≤ 0, where k(·) is contin-
uous in Om . Denote Obm = Om ∩ {x : k(x) = 0}. Suppose that
Px,i sup |X(s) − x| ≥ η → 0 as t → ∞
0≤s≤t
bm with probability at
for any η > 0 and for each x ∈ O m . Then X(t) → O
least 1 − V (x, i)/m.
f relative to an
Remark 9.10. Recall that an η-neighborhood of a set M
f
open set O is an open set Nη (M ) ⊂ O such that
f) = {x ∈ O : |x − y| < η for some y ∈ M
Nη (M f}.
Then Ex,i V (X(t ∧ τm ), α(t ∧ τm )) ≤ V (x, i). Thus the stopped process is a
supermartingale. We have
V (x, i)
Px,i ( sup V (X(t ∧ τm ), α(t ∧ τm )) ≥ λ) ≤ .
0≤t<∞ λ
9.4 Invariance (II): A Measure-Theoretic Approach 267
Define Z τm
T (t, η` ) = Tx,i (t, η` ) = χ(s)ds.
t∧τm
That is, T (t, η` ) is the total time that the process spent in Om − Nη` (O bm )
with t < T (t, η` ), and T (t, η` ) < min(∞, τm ); T (t, η` ) = 0 if τm < t. Using
Lemma 9.11,
where Ix,i (s) is the indicator of the set of (s, ω)s where LV (x, i) ≤ −γ.
Using the nonnegativity of V (·) and taking the limit as t → ∞, we obtain
Ex,i τm < V (x, i)/γ. Thus, Tx,i (t, η` ) < ∞ a.s. and Tx,i (t, η` ) → 0 as t → ∞.
There are only two possibilities. (i) There is a random variable τ (η1 ) < ∞
a.s. such that for all t > τ (η1 ), X(t) ∈ Nη1 (O bm ) with probability at least
1 − V (x, i)/m (this can also be represented as X(t) ∈ Nη1 (O bm ) a.s. relative
b
to Ωm ); (ii) for ω ∈ Ωm , X(t) moves from Nη2 (Om ) back and forth to
bm ) infinitely often in any interval [t, ∞). We demonstrate that
O m − N η1 ( O
the second alternative cannot happen.
268 9. Invariance Principles
Remark 9.13. In the above and henceforth, for simplicity, we use the
phrase “f ∈ Cb ” for a f function defined on Rr × M. In fact, since α(t)
takes values in a finite set M, for each x, f (x, ·) is trivially continuous.
Thus, the requirement of f ∈ Cb can be rephrased as “for each i ∈ M,
f (·, i) ∈ Cb .” In all subsequent development, when we say f ∈ Cb , it is
understood to be in this sense.
for each f ∈ Cb .
e ψ, dx × j) = ψ(dx × j),
(1) m(0,
e + s, ψ, dx × j) = m(t, m(s,
(2) m(t e ψ), dx × j) for any s ∈ (−∞, ∞) and
t ∈ [0, ∞).
(c) for each f (·) ∈ Cb and some weakly bounded sequence {ϕn } converg-
ing weakly to ϕ, we have
m0 Z
X m0 Z
X
f (x, j)m(ϕn , dx × j) → f (x, j)m(ϕ, dx × j).
j=1 j=1
e
Then (i) Ω(ϕ) e
6= ∅, (ii) Ω(ϕ) e
is weakly bounded, (iii) Ω(ϕ) is an invariant
set, and (iv)
m0 Z
X
f (x, j)[m(t, ϕ, dx × j) − ϕ(dx × j)] → 0, as t → ∞.
j=1
for each t ∈ (−∞, ∞) and each f (·, i) ∈ CK for each i ∈ M together with
the weak boundedness of {m(tn + t, ϕ, dx × i)} implies that the restriction
of {m(tn + t, ϕ)} to (K, B((Rr ∩ K) × M)) converges weakly to a measure
ψK (t) for each t ∈ (−∞, ∞).
Step 3. By the weak boundedness of {m(t, ϕ)}, there exists a sequence
ηi → 0, and a sequence of compact sets {Ki } in Rr satisfying Ki ⊂ Ki+1
such that for each j ∈ M,
For k > i,
XZ XZ
f (x, j)ψKi (t, dx × j) = f (x, j)ψKk (t, dx × j). (9.42)
j∈M Ki j∈M Ki
By (9.40)–(9.42), there is a measure ψ(·) such that for all t ∈ (−∞, ∞) and
all f (·) ∈ Cb ,
m0 Z
X m0 Z
X
f (x, j)m(tn + t, ϕ, dx × j) → f (x, j)ψ(t, dx × j).
j=1 j=1
Noting ψ(0) = ψ and by (c) in the assumption of the theorem, for each
f ∈ Cb , s ≥ 0, and t ∈ (−∞, ∞),
m0 Z
X
f (x, j)m(s, m(tn + t, ϕ), dx × j)
j=1
m0 Z
X
→ f (x, j)m(s, ψ(t), dx × j) as n → ∞,
j=1
272 9. Invariance Principles
and
m0 Z
X
f (x, j)m(s, m(tn + t, ϕ, dx × j), dx × j)
j=1
m0 Z
X
= f (x, j)m(0, m(tn + t + s, ϕ), dx × j)
j=1
m0 Z
X
→ f (x, j)m(0, ψ(t + s), dx × j)
j=1
m0 Z
X
= f (x, j)ψ(t + s, dx × j).
j=1
for any f ∈ Cb as n → ∞.
Proof. Fix any t ∈ [0, T ]. For any f ∈ Cb , let h(x, i) = Ex,i f (X(t), α(t)).
Then by virtue of the Feller property, h is continuous and bounded. Note
that
m0 Z
X m0 Z
X
f (x, j)m(t, ϕn , dx × j) = Ex,i f (X(t), α(t))ϕn (dx × j)
j=1 Rr i=1 R
r
m0 Z
X
= h(x, i)ϕn (dx × i).
i=1 Rr
9.4 Invariance (II): A Measure-Theoretic Approach 273
But
m0 Z
X m0 Z
X
f (x, j)m(t, ϕ, dx × j) = Ex,i f (X(t), α(t))ϕ(dx × i)
j=1 Rr i=1 R
r
m0 Z
X
= h(x, i)ϕ(dx × i).
i=1 Rr
Then condition (b) of Theorem 9.16 holds. That is, for any f ∈ Cb , the
function
m0 Z
X
t 7→ f (x, j)m(t, ϕ, dx × j)
j=1 Rr
m0 Z
X m0 Z
X
f (y, j)m(t + s, ϕ, dy × j) − f (y, j)m(t, ϕ, dy × j)
j=1 Rr j=1 Rr
m0 Z
X
= (Ex,i f (X(t + s), α(t + s)) − Ex,i f (X(t), α(t))) ϕ(dx × i)
r
j=1 R
m
X 0 Z
≤ |Ex,i f (X(t + s), α(t + s)) − Ex,i f (X(t), α(t))| ϕ(dx × i)
i=1 K
m0 Z
X
+ |Ex,i f (X(t + s), α(t + s)) − Ex,i f (X(t), α(t))| ϕ(dx × i).
i=1 Kc
m0 Z
X m0 Z
X
f (y, j)m(t + s, ϕ, dy × j) − f (y, j)m(t, ϕ, dy × j)
j=1 Rr j=1 Rr
m0 Z
X
≤ |Ex,i f (X(t + s), α(t + s)) − Ex,i f (X(t), α(t))| ϕ(dx × i)
i=1 K0
m0 Z
X
+ |Ex,i f (X(t + s), α(t + s)) − Ex,i f (X(t), α(t))| ϕ(dx × i)
c
i=1 K0
m0 Z
X X 0 m Z
ε
≤ ϕ(dx × i) + 2 kf k ϕ(dx × i)
i=1 K0
2 i=1 K0
ε ε
≤ + 2 kf k
2 4 kf k
= ε.
P m0 R
This shows that the function t 7→ j=1 Rr
f (y, j)m(t, ϕ, dy × j) is contin-
uous, uniformly in ϕ ∈ S, a weakly bounded set of M. 2
Theorem 9.19. Assume the conditions of Theorem 9.16. Then the fol-
lowing assertions hold.
e
(a) Suppose that F = suppx (Ω(ϕ)) is the support of the x-section of set
e
Ω(ϕ). Then the process X(t) converges to F in probability as t → ∞.
That is,
such that
m0 Z
X
lim sup fe(x, i)m(t, ϕ, dx × i) > 0.
t→∞ Rr
i=1
However, this implies that there is some point x ∈ Rr − O such that ψ(N ×
i) > 0 for all open set N containing x. Hence it follows that (x, i) ∈ supp(ψ),
which leads to a contradiction, because supp(ψ) ⊂ supp(Ω(ϕ)) = F ⊂ O.
Therefore (9.46) must be true and hence the assertion of (a) is established.
To prove (b), by virtue of part (a), X(t) converges in probability to F as
t → ∞. In addition, by the hypothesis, X(t) converges to G in probability
as t → ∞. Thus it follows that F ⊂ G. Meanwhile, Theorem 9.16 implies
e
that Ω(ϕ) is an invariance set. Hence it follows that F = suppx (Ω(ϕ))e ⊂
supp(LI ). As a result, X(t) converges to supp(LI ) in probability as t → ∞.
Remark 9.20. Let the conditions of Theorem 2.1 be satisfied. Then by
virtue of Theorems 2.13 and 2.18, the process (X(t), α(t)) is continuous in
probability and Feller. Suppose that for each i ∈ M, there are nonnegative
functions V (·, i) ∈ C 2 and k(x) ∈ C satisfying
LV (x, i) ≤ −k(x) ≤ 0.
Assume also that
(i) lim inf V (x, i) = ∞, and
|x|→∞ i∈M
XZ
(ii) EV (X(0), α(0)) = V (x, i)ϕ(dx × i) < ∞.
i∈M Rr
1
m(t, ϕ, Knc × M) < .
n
Then for any open set O in Rr containing the set Kn ∩ {x : k(x) = 0}, we
have
1
lim Pϕ (X(t) 6∈ O) < .
t→∞ n
So we conclude from part (b) of Theorem 9.19 that X(t) converges in
probability to the largest support of an invariant set that is contained in
limn Kn ∩ {x : k(x) = 0}.
dX 2 (t) dX(t)
+ f (X(t), α(t)) + g(X(t)) = 0,
dt2 dt
where α(t) is a continuous-time Markov chain taking values in M = {1, 2},
and for each i ∈ M, f (·, i) : R 7→ R and g(·) : R 7→ R are continuously
differentiable functions satisfying for each i ∈ M,
Denote by
Oλ = {(x1 , x2 , i) : V (x1 , x2 , i) < λ}.
It is easily checked that
since f (x1 , i) > 0 for all x1 ∈ R. With (X1 (0), X2 (0), α(0)) = (x1 , x2 , α),
by Dynkin’s formula,
So Ex1 ,x2 ,α V (X1 (t), X2 (t), α(t)) ≤ V (x1 , x2 , α), and it is a supermartin-
gale. It follows that for any t1 ≥ 0,
Px1 ,x2 ,α sup V (X1 (t), X2 (t), α(t)) ≥ λ
t1 ≤t<∞
Ex1 ,x2 ,α V (X1 (t1 ), X2 (t1 ), α(t))
≤ .
λ
By virtue of Proposition 9.9, it can be shown that with probability 1,
relative to
Ωλ = {ω : sup V (X1 (t), X2 (t), α(t)) < λ},
and
P(Ωλ ) ≥ 1 − V (x1 , x2 , i)/λ,
(X1 (t), X2 (t)) → ∂Oλ . It follows that (X1 (t), X2 (t)) → (0, 0) in probability
relative to Ωλ . Therefore, for each η > 0, there is a T < ∞, such that for
all t ≥ T , P(V (X1 (t), X2 (t)) ≥ λ0 ) ≤ η. In particular,
Thus,
η(1 + λ)
Px1 ,x2 ,α sup V (X1 (t), X2 (t), α(t)) ≥ λ0 ≤ .
T ≤t<∞ λ0
Because η and λ0 are arbitrary, (X1 (t), X2 (t)) → (0, 0) almost surely rela-
tive to Ωλ . However, Oλ is also arbitrary and bounded for any 0 < λ. Since
P(Ωλ ) → 1 as λ → ∞, (X1 (t), X2 (t)) → (0, 0) almost surely.
9.5 Notes
For systems running for a long time, it is crucial to learn their long-run
behavior; see [92, 116, 183] for recent progress on stability of such systems.
The rapid progress in natural science, life science, engineering, as well as
in social science demands the consideration of stability of such systems. In
fact, the advent of switching diffusions is largely because of the practical
needs in modeling complex dynamic systems; see [7, 52, 74, 92, 116, 168,
180, 183, 187, 188] for some of the recent studies.
Most works to date has been concentrated on Markov-modulated dif-
fusions, in which the Brownian motion and the switching force are inde-
pendent, whereas less is known for systems with continuous-component
dependent switching processes. As demonstrated in Chapter 2 (see also
[188]), when x-dependent switching diffusions are encountered, even such
properties as continuous and smooth dependence on the initial data are
nontrivial and fairly difficult to establish. Nevertheless, studying such sys-
tems is both practically useful and theoretically interesting. In our recent
work, basic properties such as recurrence, positive recurrence, and ergod-
icity are studied in [187]; stability is treated in [92]; stability of randomly
switching ordinary differential equations is treated in [190].
9.5 Notes 281
10.1 Introduction
To study the positive recurrence and ergodicity, one of the conditions used
in Chapters 3 and 4 is that the states of the switching process belong to
only one ergodic class. In this chapter, we further our study by treating a
more general class of problems. We consider the case that the states of the
discrete event process belong to several “ergodic” classes that are weakly
connected. This notion is made more precise in what follows. A key idea is
the use of two-time-scale formulation; see [176, 177] and many references
therein.
The rest of the chapter is arranged as follows. Section 10.2 begins with
the formulation. Section 10.3 focuses on hybrid diffusions whose discrete
component lives in weakly connected “ergodic” (irreducible) classes. Fi-
nally, the chapter is concluded with additional remarks in Section 10.4.
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 285
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_10,
© Springer Science + Business Media, LLC 2010
286 10. Positive Recurrence: Weakly Connected Ergodic Classes
and
P(α(t + ∆t) = j|α(t) = i, X(s), α(s), s ≤ t)
(10.2)
= qij (X(t))∆t + o(∆t), i 6= j,
where w(t) is a d-dimensional standard Brownian motion, b(·, ·) : Rr ×M 7→
Rr , and σ(·, ·) : Rr × M 7→ Rr×d satisfying
b(x, i) = (bj (x, i)) ∈ Rr , σ(x, i)σ 0 (x, i) = a(x, i) = (ajk (x, i)) ∈ Rr×r ,
with z 0 denoting the transpose of z for z ∈ Rι1 ×ι2 and ι1 , ι2 ≥ 1.
Associated with the process given in (10.1) and (10.2), there is a gener-
ator L0 defined as follows. For each i ∈ M and for any twice continuously
differentiable function g(·, i), let
1
L0 g(x, i) =tr(a(x, i)∇2 g(x, i)) + b0 (x, i)∇g(x, i) + Q(x)g(x, ·)(i),
2
(10.3)
where ∇g(·, i) and ∇2 g(·, i) denote the gradient and Hessian of g(·, i), re-
spectively, and
m0
X
Q(x)g(x, ·)(i) = qij (x)g(x, j)
j=1 (10.4)
X
= qij (x)(g(x, j) − g(x, i)), i ∈ M.
j6=i,j∈M
deals with the situation that weakly connected, multiple ergodic classes
are included. We assume that the discrete jump component is generated by
1e b
Qε = Q + Q, (10.5)
ε
10.3.1 Preliminary
Before proceeding further, let us give the motivation for using such models.
First, we note that Qε is a constant matrix independent of x. The ratio-
nale is similar to those considered in the previous chapters, which can be
considered as linearizing Q(x) at “point of ∞.” We may begin with an x-
dependent matrix, say Qε (x). Then for x large enough, it can be replaced
by a constant Qε . To be more precise, assume
where Qε has the form (10.5). Note that in the previous chapters, a con-
dition similar to (10.6) was used without the ε-dependence, but the cor-
responding constant matrix (the limit at |x| → ∞) is a generator of an
ergodic Markov chain. Here, we are mostly concerned with the case that
the generator could possibly be reducible with several ergodic classes. Nev-
ertheless, the states belonging to different ergodic classes are not completely
separable. They are linked together through weak interaction due to the
presence of the slow part of the generator Q. b
The formulation of Qε being a generator of an ε-dependent Markov chain
for a small parameter ε, stems from an effort of using two-time-scale models
to reduce the complexity of the underlying systems. It has been observed
in [149] that there are natural hierarchical structures in many large-scale
and complex systems. The formulation in (10.5) is an effort to highlight the
different parts of subsystems varying at different rates. It is often possible
to partition the system states into a number of groups so that within each
group, the state transitions take place rapidly, whereas among different
groups, the changes are relatively infrequent. Such scenarios, in fact, appear
in many applications. Thus an effective way is to treat the systems through
decompositions and aggregations. Loosely, one can use the natural scales
shown in the system to aggregate the states in each ergodic class into
288 10. Positive Recurrence: Weakly Connected Ergodic Classes
one state. In this way, the total number of states is much reduced for the
aggregated system. This point of view was exclusively discussed in Yin and
Zhang [176]. To begin, one may not have an ε in the system, but it is
brought into the formulation to highlight the different rates of change so
as to separate the fast and slow motions.
M = M 1 ∪ M2 ∪ · · · ∪ M l , (10.8)
e i is irreducible.
(A10.1) For each i = 1, . . . , l, Q
satisfies
pε (t) = θ(t)ν̃ + O(ε(t + 1) + e−κ0 t/ε )
for some κ0 > 0, and θ(t) = (θ1 (t), . . . , θl (t)) ∈ R1×l satisfies
dθ(t)
= θ(t)Q, θ(0) = p(0)1l.
dt
dΘ(t)
= Θ(t)Q, Θ(0) = I.
dt
Proof. The proofs of (a) and (b) in Lemma 10.1 can be found in [176,
Corollary 6.12, p. 130]. The proof of (c) is based on the martingale averaging
method [176, Theorem 7.4, p. 172]; an outline of the idea is in [4], and a
discrete version of the approximation may be found in [177].
As for (d), using (a) and (b), direct calculation reveals that
Z ∞ 2
−t
E e (I{αε (t)=sij } − νji I{αε (t)=i} )dt
Z 0∞ Z t
= e−t−s O ε(t + 1) + e−κ0 (t−s)/ε dsdt
Z ∞ 0Z s
0
+ e−t−s O ε(t + 1) + e−κ0 (s−t)/ε dtds.
0 0
290 10. Positive Recurrence: Weakly Connected Ergodic Classes
Remark 10.3. In view of the weak convergence result, the limit stochastic
differential equation for (10.9) is
dx = b(X(t), α(t))dt + σ(X(t), α(t))dw, (10.16)
where σ(x, i) is defined in terms of the average in (10.15); that is,
σ(x, i)σ 0 (x, i) = a(x, i).
(A10.2) Q is irreducible. For each i ∈ M, a(x, i) satisfies
κ1 |ξ|2 ≤ ξ 0 a(x, i)ξ, for all ξ ∈ Rr , (10.17)
with some constant κ1 ∈ (0, 1] for all x ∈ Rr .
With the conditions given, using the techniques of Chapter 3, to be more
specific, by virtue of Theorem 3.26, we establish the following assertion.
Proposition 10.4. Under (A10.1) and (A10.2), the switching diffusion
with generator given by (10.14) is positive recurrent.
We now present the result on positive recurrence of the underlying pro-
cess. The main idea is that although the discrete events, described by a
continuous-time Markov chain may have several weakly connected ergodic
classes, when ε is sufficiently small, we still have positive recurrence for the
process (X ε (·), αε (·)). The proof rests upon the use of perturbed Liapunov
function methods, which were first used to treat diffusion approximations
in [131] by Papanicolaou, Stroock, and Varadhan, and later on have been
successfully used in stochastic systems theory (see Kushner [102]), and
stochastic approximation (see Kushner and Yin [104]), among others. The
basic idea is to introduce perturbations of Liapunov functions that are
small in magnitude and that result in the desired cancellation of unwanted
terms.
10.3 Weakly Connected, Multiergodic-Class Switching Processes 291
Theorem 10.5. Assume that conditions (A10.1) and (A10.2) hold. Then
for sufficiently small ε > 0, the process (X ε (·), αε (·)) is positive recurrent.
Remark 10.6. First note that by Proposition 10.5, the process (X(·), α(·))
is positive recurrent. Then it follows that there are Liapunov functions
V (x, i) for i = 1, . . . , l for the limit system (10.16) such that
In Theorem 10.5, the meaning of the property holds for sufficiently small
ε. That is, there exists an ε0 > 0 such that for all 0 < ε ≤ ε0 , the property
holds.
Proof. To prove this result, we begin with the Liapunov function in (10.18)
for the limit system. Choose n0 to be an integer large enough so that D is
contained in the ball {|x| < n0 }. For any (x, i) ∈ D c × M, any t > 0, any
positive integer n > n0 , define
where βn comes from the regularity consideration and is the first exit time
of the process (X(t), α(t)) from the set {ex : |e
x| < n} × M; this is, βn =
inf{t : |X(t)| = n}. For i ∈ M, use the Liapunov function V (x, i) in (10.18)
to define
l
X
V (x, α) = V (x, i)I{α∈Mi } = V (x, i) if α ∈ Mi . (10.20)
i=1
Note that
V (X ε (t), αε (t)) = V (X ε (t), αε (t)), (10.21)
which involves unwanted terms on the right-hand side. To get rid of these
terms, we use methods of perturbed Liapunov functions [104, 176] to aver-
age out the “bad” terms. Note that the process (X ε (t), αε (t)) is Markov.
Thus, for a suitable function ξ(·), Lε ξ(t) can be calculated by
ξ(t + δ) − ξ(t)
Lε ξ(t) = lim Eεt , (10.24)
δ→0 δ
where Eεt denotes the conditional expectation with respect to the σ-algebra
Z ∞
V1ε (x, α, t) = Eεt et−u V x (x, α) b(x, αε (u)) − b(x, αε (u)) du,
Zt ∞
1
V2ε (x, α, t) = Eεt et−u tr[V xx (x, α)(a(x, αε (u)) − a(x, αε (u)))]du,
Z ∞t 2
ε ε t−u b
V3 (x, t) = Et e [QV (x, ·)(αε (u)) − QV (x, ·)(αε (u))]du.
t
(10.25)
To proceed, we first state a lemma.
Lemma 10.7. Assume the conditions of Theorem 10.5. Then the following
assertions hold.
(b) Moreover,
Likewise, we obtain the estimates for V2ε (X ε (t), αε (t), t) and V3ε (X ε (t), t)
in (10.26). This establishes statement (a).
Next, we prove (b). For convenience, introduce a notation
Moreover,
Z t+δ
1
− lim Eεt et−u Γ(X ε (t), αε (t), αε (u))du
δ→0 δ
t (10.32)
ε ε ε ε ε ε
= − V x (X (t), α (t)) b(X (t), α (t)) − b(X (t), α (t)) ,
and
Z ∞
1
lim Eεt [et+δ−u − et−u ]Γ(X ε (t), αε (t), αε (u))du
δ→0 δ
Z t+δ
∞
= et−u Eεt Γ(X ε (t), αε (t), αε (u))du = V1ε (X ε (t), αε (t)).
t
(10.33)
The independence of αε (·) and w(·), (10.28), and (10.31) lead to that for
u ≥ t,
− V x (X ε (t), αε (t))b(X ε (t), sij )]Eεt+δ [I{αε (u)=sij } − νji I{αε (u)∈Mi } ].
(10.34)
Furthermore, we have
Z
1 ∞ t+δ−u ε
lim e Et [Γ(X ε (t + δ), αε (t), αε (u))
δ→0 δ t+δ
and
mi
l X
X 1
tr[(Vx (X ε (t), αε (t))b(X ε (t), sij ))xx a(X ε (t), αε (t))]
2
i=1 j=1 Z
∞ (10.36)
× Eεt et−u I{αε (u)=sij } − νji I{αε (u)=i} du
t
= O(ε)(V (X ε (t), αε (t)) + 1).
It follows that
and
Therefore,
Lε V ε (t) = LV (X ε (t), αε (t)) + O(ε)(V (X ε (t), αε (t)) + 1). (10.40)
Note that through the use of perturbations, the first term on the right-
hand side above involves the limit operator and the Liapunov function of
the limit system, which is crucially important.
For fixed but arbitrary T > 0, we then obtain
In addition,
Because
X3
1
Px,i (τD > t) ≤ [V (x, i) + Ex,i Vιε (x, i, 0)] → 0 as t → ∞,
α1 t ι=1
Write
b 11
Q b 12
Q
b=
Q
b 21
Q b 22
Q
b 11 ∈ R(m0 −m∗ )×(m0 −m∗ ) , Q
so that Q b 22 ∈ Rm∗ ×m∗ , and Q
b 12 and Q
b 21 have
appropriate dimensions. Denote
b 11 1l + Q
Q = ν̃(Q b 12 A),
e Q∗ = diag(Q, 0m ×m ). (10.48)
∗ ∗
where
l
X
Uj = iI{Pi−1 P
aj0 j <U ≤ ij
e aj0 j } ,
e
j0 =1 0 =1
i=1
satisfies
where P (0) (t) = 1l∗ Θ∗ (t)ν∗ with Θ∗ (t) = diag(Θ(t), Im∗ ×m∗ ), where
Θ(t) satisfies
dΘ(t)
= Θ(t)Q, Θ(0) = I.
dt
(c) αε (·) converges weakly to α(·), a Markov chain generated by Q.
(d) For i = 1, . . . , l, j = 1, . . . , mi ,
Z ∞ 2
E e−t (I{αε (t)=sij } − νji I{αε (t)=i} )dt = O(ε),
0
and for i = ∗, j = 1, . . . , m∗ ,
Z ∞ 2
−t
E e I{αε (t)=s∗j } dt = O(ε).
0
(e) (X ε (·), αε (·)) converges weakly to (X(·), α(·)) such that the limit op-
erator is as given in Lemma 10.2.
Remark 10.9. Although a collection of transient states of the discrete
events are included, the limit system is still an average with respect to
the stationary measures of those ergodic classes only. Asymptotically, the
transient states can be discarded because the probabilities go to 0 rapidly.
Theorem 10.10. Assume (A10.2) and (A10.3), and for each i ∈ M, there
is a Liapunov function V (x, i) such that LV (x, i) ≤ −γ for some γ > 0.
Then for sufficiently small ε > 0, the process (X ε (t), αε (t)) with inclusion
of transient discrete events is still positive recurrent.
Idea of Proof. Since the proof is along the line of that of Theorem 10.5,
we only note the differences compared with that theorem. For i = 1, . . . , l,
let V (x, i) be the Liapunov function associated with the limit process given
by Lemma 10.8(v). The perturbed Liapunov function method is used again.
This time, redefine
l
X m∗
X
V (x, α) = V (x, i)I{α∈Mi } + V (x, i)e
ai,j I{α=s∗j } . (10.50)
i=1 i=1
300 10. Positive Recurrence: Weakly Connected Ergodic Classes
Therefore, we can carry out the proof in a similar manner to that of The-
orem 10.5. The details are omitted.
10.4 Notes
This chapter continues our study of positive recurrence for switching diffu-
sions. One of the crucial assumptions in the ergodicity study in the previous
chapters is that the switching component has a single ergodic class. This
chapter takes up the issue that the discrete component may have multi-
ple ergodic classes that are weakly connected. The main idea is the use
of the two-time-scale approach. Roughly, if the discrete events or certain
components of the discrete events change sufficiently quickly, the positive
recurrence can still be guaranteed. The main ingredient is the use of per-
turbed Liapunov function methods.
11
Stochastic Volatility Using
Regime-Switching Diffusions
11.1 Introduction
This chapter aims to model stochastic volatility using regime-switching dif-
fusions. Effort is devoted to developing asymptotic expansions of a system
of coupled differential equations with applications to option pricing under
regime-switching diffusions. By focusing on fast mean reversion, we aim
at finding the “effective volatility.” The main techniques used are singular
perturbation methods. Under simple conditions, asymptotic expansions are
developed with uniform asymptotic error bounds. The leading term in the
asymptotic expansions satisfies a Black–Scholes equation in which the mean
return rate and volatility are averaged out with respect to the stationary
measure of the switching process. In addition, the full asymptotic series is
developed. The asymptotic series helps us to gain insight on the behavior
of the option price when the time approaches maturity. The asymptotic
expansions obtained in this chapter are interesting in their own right and
can be used for other problems in control optimization of systems involving
fast-varying switching processes.
Nowadays, sophisticated financial derivatives such as options are used
widely. The Nobel prize winning Black–Scholes formula provides an im-
portant tool for pricing options on a basic equity. It has encouraged and
facilitated the union of mathematics, finance, computational sciences, and
economics. On the other hand, it has been recognized, especially by practi-
tioners in the financial market, that the assumption of constant volatility,
which is essential in the Black–Scholes formula, is a less-than-perfect de-
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 301
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_11,
© Springer Science + Business Media, LLC 2010
302 11. Stochastic Volatility Using Regime-Switching Diffusions
scription of the real world. To capture the behavior of stock prices and
other derivatives, there has been much effort in taking into account fre-
quent volatility changes. It has been recognized that it is more suitable to
use a stochastic process to model volatility variations. In [70], instead of
the usual GBM (geometric Brownian motion) model, a second stochastic
differential equation is introduced to describe the random environments.
Such a formulation is known as a stochastic volatility model.
What happens if the stochastic volatility undergoes fast mean rever-
sion? To answer this question, in [49] and the subsequent papers [50, 51],
a class of volatility models has recently been studied in details. Under the
setup of mean reversion, two-time-scale methods are used. The rationale is
to identify the important groupings of market parameters. It also reveals
that the Black–Scholes formula is a “first approximation” to such fast-
varying volatility models. Assume that the volatility is a function f (·) of a
fast-varying diffusion that is mean reverting (or ergodic). The mean rever-
sion implies that although rapidly varying, the volatility does not blow up.
By exploiting the time-scale separation, it was shown in [49, 50] that the
“slow” component (the leading term or the zeroth-order outer expansion
term in the approximation) of the option prices can be approximated by a
2
Black–Scholes differential equation with constant volatility f where f is
the average of f 2 (·) with respect to the stationary measure of the “fast”
component. Moreover, using a singular perturbation approach, the next
term (the first-order outer expansion term) in the asymptotic expansion
was also found. For convenience, the volatility was assumed to be driven
by an Ornstein–Uhlenbeck (OU) process in [49]–[51]. The fast mean rever-
sion has been further examined in [90] with more general models. A full
asymptotic series with uniform error bounds was obtained; see also related
diffusion approximation in [131] and asymptotic expansions for diffusion
processes [88, 89].
Along another line, increasing attention has been drawn to modeling,
analysis, and computing using regime-switching models [164], which are
alternatives to the stochastic volatility models mentioned above. They
present an effective way to model stochastic volatility with simpler struc-
tures. The use of the Markov chains is much simpler than the use of a
second stochastic differential equation. This is motivated by the desire
to use regime switching to describe uncertainty and stochastic volatility.
Nowadays, it has been well recognized that due to stochastic volatility, a
phenomena known as the volatility smile arises. There has been an effort
to provide better models to replicate the “smile.” In [164], the authors re-
produced the volatility smile successfully by using regime-switching models
easily. Earlier efforts in modeling and analysis of regime-switching models
can be found in [5, 32, 184] among others.
To some large extent, this chapter is motivated by [49] for extremely fast
mean reversion processes in the driving force of stochastic volatility. Nev-
11.2 Formulation 303
11.2 Formulation
In the rest of the book, r is used as the dimension of the continuous state
variables (i.e., Rr is used as the space for the continuous state variable).
In this chapter, however, to adopt the traditional convention, we use r as
the interest rate throughout. We consider a basic equity, a stock whose
price is given by S(t). Different from the traditional geometric Brownian
motion setup, we assume that the price follows a switching-diffusion model.
Suppose that α(t) is a continuous-time Markov chain with generator Q(t)
and a finite state space M = {1, . . . , m0 }. The price of the stock is a
solution of the stochastic differential equation
where w(·) is a standard Brownian motion independent of α(·), and µ(·) and
σ(·) are the appreciation rate and the volatility rate, respectively. Such a
304 11. Stochastic Volatility Using Regime-Switching Diffusions
has a unique solution ν(t) = (ν1 (t), . . . , νm0 (t)) ∈ R1×m0 satisfying νi (t) ≥
0 for each i ∈ M. Such a nonnegative solution is termed a quasi-stationary
distribution; see [176].
Let
V ε (S, t, i) = ES,i [exp(−r(T − t))h(S ε (T ))]
(11.6)
= E[exp(−r(T − t))h(S ε (T ))|S ε (t) = S, αε (t) = i].
The option price can be characterized by the following system of partial
differential equations,
∂V ε (S, t, i) 1 ∂ 2 V ε (S, t, i)
+ σ 2 (i)S 2
∂t 2 ∂S 2
∂V ε (S, t, i) (11.7)
+ rS − rV ε (S, t, i)
∂S
+ Qε (t)V ε (S, t, ·)(i) = 0, i ∈ M,
which is a generalization of the usual Black–Scholes PDE. Associated with
the above system of PDEs, we define an operator Lε . For i ∈ M, and each
g(·, ·, i) ∈ C 2,1 , let
∂g(S, t, i) 1 2 ∂ 2 g(S, t, i)
Lε g(S, t, i) = + σ (i)S 2
∂t 2 ∂S 2
∂g(S, t, i) (11.8)
+ rS − rg(S, t, i)
∂S
+ Qε (t)g(S, t, ·)(i).
Now the setup of the problem is complete. We proceed to obtain the ap-
proximation of the option price by means of asymptotic expansions.
T −t
τ= .
ε
The ϕk (·) are called regular terms or outer expansion terms, and the ψk (·)
are the boundary layer correction terms (or to be more precise, terminal
layer corrections). In this problem, the terminal layer correction terms are
particularly useful for behavior of the option price near the time of maturity.
We aim to obtain asymptotic expansions of the order n, and derive the
uniform error bounds. For the purposes of error estimates, we need to
calculate a couple of more terms for analysis reasons.
First let us look at Φεn (S, t, i), the regular part of the asymptotic expan-
sions. Substituting it into (11.7), and comparing coefficients of like powers
of εk for 0 ≤ k ≤ n + 1, we obtain
n+1
X (−1)k (ετ )k dk Q(T )
Q(t) = Q(T − ετ ) = + Rn+1 (T − ετ ), (11.12)
k! dtk
k=0
Our task to follow is to construct the sequences {ϕk (S, t)} and {ψk (S, τ )}.
Then (11.15) together with the terminal condition (11.16) has a unique
solution.
Remark 11.3. Note that γ0 (S, t) satisfies the Black–Scholes partial dif-
ferential equation, in which the coefficients are averaged out with respect
to the stationary distributions of the Markov chain. The result reveals that
the Black–Scholes formulation, indeed, is a first approximation to the op-
tion model under regime switching. Thus, the regime-switching model can
be thought of as another stochastic volatility model.
In view of (11.11) and (11.13), ψ0 (S, τ, i) is obtained from the first equa-
tion in (11.13) together with the condition ψ0 (S, 0, i) = 0. It follows that
ψ0 (S, τ, i) = 0 for each i ∈ M.
Equivalently, denote
Then
ϕ1 (S, t) = γ1 (S, t)1l + ϕ01 (S, t), (11.17)
310 11. Stochastic Volatility Using Regime-Switching Diffusions
where
F0 (S, t) = (F0 (S, t, 1), . . . , F0 (S, t, m0 ))0 ∈ Rm0 ,
with
h ∂γ (S, t) σ 2 (i)S 2 ∂γ (S, t) ∂γ0 (S, t) i
0 0
F0 (S, t, i) = − + 2
+ rS − rγ0 (S, t) ,
∂t 2 ∂S ∂S
(11.19)
by Lemma A.12. That is, ϕ1 (S, t) is the sum of a general solution of the
homogeneous equation plus a particular solution verifying the orthogonality
condition.
Note that γ1 (S, t) has not been determined yet. We proceed to obtain
it from the next equation in (11.10). By substituting (11.17) into the next
equation in (11.10) and premultiplying it by ν, we obtain
with
h ∂ϕ0 (S, t, i) σ 2 (i)S 2 ∂ϕ0 (S, t, i) ∂ϕ01 (S, t, i)
1 1
F1 (S, t, i) = − + rS
∂t i 2 ∂S 2 ∂S
− rϕ01 (S, t, i) ,
(11.21)
and
F 1 (S, t) = ν(t)F1 (S, t).
∂ψ1 (S, τ )
= Q(T )ψ1 (S, τ ). (11.22)
∂τ
0
Choose ψ1 (S, 0) = ψ10 (S) such that ψ 1 (S) = νψ10 (S) = 0 where ν = ν(T ).
Here and hereafter, we always use ν to denote ν(T ) for notational simplicity.
11.3 Asymptotic Expansions 311
ψ 1 (S, τ ) = 0.
for each τ ≥ 0.
Note that once ψ10 (S) is chosen, the ϕ1 (S, t) and ψ1 (S, τ ) will be deter-
mined. The choice of ψ10 (S, 0) enables us to obtain the exponential decay
of ψ1 (S, τ ) easily.
It follows that the Cauchy problem given by (11.22) and (11.24) has a
unique solution. We are in a position to derive the exponential decay prop-
erty of ψ1 (S, τ ).
Lemma 11.5. For ψ1 (S, τ ) obtained from the solution of (11.22) and
(11.24), we have that for some K > 0 and κ0 > 0,
The solution of (11.22) and (11.24) yields that for each S ∈ [0, 1] and for
some K > 0,
Furthermore, it is readily seen that the above estimate holds uniformly for
S ∈ [0, 1]. The desired result thus follows. 2
∂ψ1 (S, τ )
sup ≤ K exp(−κ0 τ ),
S∈[0,1] ∂S
∂ 2 ψ1 (S, τ )
sup ≤ K exp(−κ0 τ ).
S∈[0,1] ∂S 2
Proof. Consider
∂ψ1 (S, τ )
U (S, τ ) = and
2
∂S
∂ ψ1 (S, τ )
V (S, τ ) = .
∂S 2
Then we have U and V satisfy
Suppose that we have constructed ϕk−1 (S, t) and ψk−1 (S, τ ) for 1 < k ≤
n + 1. Then similar to ϕ1 (S, t) and ψ1 (S, τ ), we can define ϕk (S, t) and
ψk (S, τ ). Next, write ϕk (S, t) as
With given initial data ψ2 (S, 0), the solution of (11.32) is given by
Z τ
ψ2 (S, τ ) = exp(Q(T )τ )ψ2 (S, 0) + exp(Q(T )(τ − s))R e2 (S, s)ds.
0
(11.34)
The solution of (11.34) can be further expanded as
Z ∞
ψ2 (S, τ ) = 1lνψ2 (S, 0) + 1l e2 (S, s)ds
νR
0 Z ∞
+ [exp(Q(T )τ ) − 1lν]ψ2 (S, 0) − 1lν e2 (S, s)ds
R (11.35)
Z τ τ
In addition,
Z τ
[exp(Q(T )(τ − s)) − 1lν]R e2 (S, s)ds
Z τ
0
≤ Kτ exp(−κ0 τ ).
Thus, ψ2 (S, τ ) will decay to 0 exponentially fast if we choose
Z ∞
0
1lψ 2 (S) = 1lνψ2 (S, 0) = −1l e2 (S, s)ds.
νR (11.36)
0
0
Note that there is only one unknown in (11.36), namely, ψ 2 (S, 0) = ψ 2 (S).
That is, (11.36) enables us to obtain ψ 2 (S, 0) uniquely. Using γ2 (S, T ) =
−ψ 2 (S, 0) together with (11.30) enables us to find the unique solution for
the Cauchy problem (11.30) that satisfies ψ 2 (S, 0). Therefore, ϕ2 (S, t) and
ψ2 (S, τ ) are completely determined. Moreover, the construction ensures
that
sup |ψ2 (S, τ )| ≤ K exp(−κ0 τ ).
S∈[0,1]
In addition, it can be shown that
∂ i ψ2 (S, τ )
sup ≤ K exp(−κ0 τ ), i = 1, 2.
S∈[0,1] ∂S i
Recall that for simplicity, we have used the convention that K and κ0 are
some positive real numbers. Their values may vary. Their precise values are
not important but only the exponential decay property is crucial.
Proceeding in a similar way, we obtain ψk (S, τ ) as follows. Suppose that
ψ1 (S, τ ), . . . , ψk−1 (S, τ ) have been constructed so that ψj (S, τ ) for j =
1, . . . , k − 1 decay exponentially fast together with their first and second
derivatives (∂/∂S)ψj (S, τ ) and (∂ 2 /∂S 2 )ψj (S, τ ), respectively.
Consider
∂ψk (S, τ ) ek (S, τ ).
= Q(T )ψk (S, τ ) + R (11.37)
∂τ
The solution is then given by
Z τ
ψk (S, τ ) = exp(Q(T )τ )ψk (S, 0) + exp(Q(T )(τ − s))R ek (S, s)ds
Z ∞ 0
Choose
Z ∞
0
ψ k (S) := 1lνψk (S, 0) = −1l ek (S, s)ds,
νR
0
0 (11.39)
γk (S, T ) = −ψ k (S),
0
ψk (S, 0) = −ϕk (S, T ) = ψ k (S)1l − ϕk (S, T ).
∂ψk (S, τ )
sup ≤ K exp(−κ0 τ ),
S∈[0,1] ∂S
∂ 2 ψk (S, τ )
sup ≤ K exp(−κ0 τ ).
S∈[0,1] ∂S 2
Proof. Redefine
∂ψk (S, τ )
U (S, τ ) = and
2
∂S
∂ ψk (S, τ )
V (S, τ ) = .
∂S 2
Then U and V satisfy
• ϕ0 (S, t) = γ0 (S, t)1l with γ0 (S, t) being the solution of (11.15) satisfy-
ing the terminal condition (11.16); ψ0 (S, τ, i) = 0 for each i ∈ M;
• ϕ1 (S, t, i) is given by (11.17) with ϕ01 (S, t) being the unique solution
of (11.18) and γ1 (S, t) given by (11.20) with γ1 (S, T ) = 0 and the
terminal layer term ψ1 (S, τ, i) specified in Lemma 11.5;
• ϕk (S, t) is given by (11.28) with ϕ0k (S, t) being the unique solution of
(11.29) and γk (S, t) being the solution of (11.30) with γ(S, T ) given
in Lemma 11.8 and the terminal layer term ψk (S, τ, i) specified in
Lemma 11.8.
Proof. Note that S ∈ [0, 1] and t ∈ [0, T ] for some finite T > 0 so v(·) is
bounded together with its derivatives with respect to S up to the second
order and its derivative with respect to t. In view of (11.2) and (11.3), u(·)
is integrable. Moreover, the local martingale in Lemma 11.1 is in fact a
martingale now. The desired result then follows from Dynkin’s formula. The
proof may also be worked out by means of martingale problem formulation.
2
318 11. Stochastic Volatility Using Regime-Switching Diffusions
Lemma 11.12. Suppose that (A11.1) and (A11.2) hold and for each i ∈
M, eε (·, ·, i) is a suitable function such that
sup |eε (S, t, i)| = O(ε` ) for ` ≤ n + 1. (11.43)
(S,t)∈[0,1]×[0,T ]
Proof. The proof is divided in two steps. In the first step, we obtain an
estimate on Lε eεn+1 (S, t, i), and in the second step, we derive the desired
order estimate.
Step 1. Claim: Lε eεn+1 (S, t, i) = O(εn+1 ). To obtain this, first note that
ε ε
L V (S, t, i) = 0. Thus
where in the last line above, we have used ψ0 (S, τ, i) = 0 for each i ∈ M.
For the outer expansions, we have
n+1
X
Lε εk ϕk (S, t, i)
k=0
n+1
X h ∂ϕ (S, t, i) 1 ∂ 2 ϕk (S, t, i)
k
= εk + σ 2 (i)S 2
∂t 2 ∂S 2
k=0
∂ϕk (S, t, i) Q(t) i
+ rS − rϕk (S, t, i) + ϕk (S, t, ·)(i)
∂S ε
Xn h 2
k ∂ϕk (S, t, i) 1 2 2 ∂ ϕk (S, t, i)
= ε + σ (i)S
∂t 2 ∂S 2
k=0
∂ϕk (S, t, i) i
+ rS − rϕk (S, t, i) + Q(t)ϕk+1 (S, t, ·)(i)
∂S h ∂ϕ
Q(t) n+1 (S, t, i)
+ ϕ0 (S, t, ·)(i) + εn+1
ε 2
∂t i
1 2 2 ∂ ϕn+1 (S, t, i) ∂ϕn+1 (S, t, i)
+ σ (i)S + rS − rϕ n+1 (S, t, i)
2 h ∂S 2 2
∂S
n+1 ∂ϕn+1 (S, t, i) 1 2 2 ∂ ϕn+1 (S, t, i)
=ε + σ (i)S 2
∂t 2 i ∂S
∂ϕn+1 (S, t, i)
+ rS − rϕn+1 (S, t, i) .
∂S
(11.46)
The boundedness of ϕn+1 (S, t, i) and their derivatives up to the order 2
together with (11.46) then lead to
max sup |Lε Φεn+1 (S, t, i)| = O(εn+1 ). (11.47)
i∈M (S,t)∈[0,1]×[0,T ]
and that
n+1
X
εj−1 ψj (S, τ, i)O(tn+1−j )
j=1
n+1
X
≤K εn+1−j tj exp(−κ0 τ ) ≤ Kεn+1 .
j=1
Step 2. Obtain estimate (11.45). Note that the definition of eεn+1 (S, t, i)
and (11.11) yield that eεn+1 (S, T, i) = 0. Because (11.50) holds, it follows
from Lemma 11.12 that
max sup |eεn+1 (S, t, i)| = O(εn+1 ).
i∈M (S,t)∈[0,1]×[0,T ]
Note that
eεn+1 (S, t, i) = eεn (S, t, i) + εn+1 ϕn+1 (S, t, i) + εn+1 ψn+1 (S, τ, i).
(11.51)
The smoothness of ϕn+1 (S, t, i) and the exponential decay of ψn+1 (S, τ, i)
imply that
max sup |εn+1 ϕn+1 (S, t, i) + εn+1 ψn+1 (S, τ, i)| = O(εn+1 ).
i∈M (S,t)∈[0,1]×[0,T ]
11.5 Notes
In this chapter, we have developed asymptotic expansions for a European-
type option price. The essence is the use of two-time-scale formulation to
deal with solutions of systems of parabolic PDEs. The result is based on
the recent work of Yin [166]. The approach we are using is constructive.
Thus it sheds more light on how these approximations can be carried out.
Full asymptotic expansions have been obtained with uniform asymptotic
error bounds for the continuous component belonging to a compact set. If
one is only interested in getting asymptotic expansions with certain fixed
state variables (as in [49, 50]), then one can work with the entire space R
rather than a compact set. In lieu of Qε (t) considered thus far, we may
treat a slightly more complex model with
Q0 (t)
Qε (t) = + Q1 (t),
ε
where both Q0 (·) and Q1 (·) are generators of continuous-time Markov
chains such that Q0 (t) is weakly irreducible. Then we can still obtain
asymptotic expansions. The notation, however, will be a bit more complex
due to the addition of Q1 (t). In view of the work by Il’in, Khasminskii, and
Yin [73, 74], the results of this chapter can be extended to switching dif-
fusions in which the switching process has generator Qε (x, t) that depends
on x as well.
For risk-neutral valuation, it is natural to let r be a constant. Never-
theless, the techniques presented here carry over for the more general α-
dependent process; that is r(t) = r(α(t)). Although the main motivation is
from mathematical finance, the techniques developed here can be used for
other problems involving systems of coupled differential equations where a
fast-varying switching process is a driving force.
12
Two-Time-Scale Switching Jump
Diffusions
12.1 Introduction
This chapter is concerned with jump diffusions involving Markovian switch-
ing regimes. In the models, there are a finite set of regimes or configura-
tions and a switching process that dictates which regime to take at any
given instance. At each time t, once the configuration is determined by
the switching process, the dynamics of the system follow a jump-diffusion
process. It evolves until the next jump takes place. Then the post-jump
location is determined and the process sojourns in the new location follow-
ing the evolution of another jump-diffusion process and so on. The entire
system consists of random switches and jump-diffusive motions.
One of our motivations stems from insurance risk theory. To capture
the features of insurance policies that are subject to economic or political
environment changes, generalized hybrid risk models may be considered. To
reduce the complexity of the systems, time-scale separation may be used.
Under the classical insurance risk model, the surplus U (t) of an insurance
company at t ≥ 0 is given by
U (t) = u + ct − S(t),
where u is the initial surplus, c > 0 is the rate at which the premiums
are received, and S(t), a compound Poisson process, is the total claim in
the duration [0, t]. In [35], Dufresne and Gerber extended the classical risk
model by adding an independent diffusion process so that the surplus is
given by
U (t) = u + ct − S(t) + σw(t),
G.G. Yin and C. Zhu, Hybrid Switching Diffusions: Properties and Applications, 323
Stochastic Modelling and Applied Probability 63, DOI 10.1007/978-1-4419-1105-6_12,
© Springer Science + Business Media, LLC 2010
324 12. Two-Time-Scale Switching Jump Diffusions
Throughout the chapter, we assume that w(·), N (·), and α(·) are mutually
independent. Compared with the traditional jump-diffusion processes, the
coefficients involved in (12.1) all depend on an additional switching process,
namely, the Markov chain α(t).
In the context of risk theory, X(t) can be considered as the surplus of the
insurance company at time t, x is the initial surplus, f (t, X(t), α(t)) repre-
sents the premium rate (assumed to be ≥ 0), g(γ, X(t), α(t)) is the amount
of the claim whenever there is one (assumed to be ≤ 0), and the diffusion
is used to model additional uncertainty of the claims and/or premium in-
comes. Similar to the volatility in stock market models, σ(·, ·, i) represents
the amount of oscillations or volatility in an appropriate sense. The model
is sufficiently general to cover the traditional compound Poisson models as
well as the diffusion perturbed ruin models. It may also be used to repre-
sent security price in finance (see [124, Chapter 3]). The process α(t) may
be viewed as an environment variable dictating the regime. The use of the
Markov chain results from consideration of the general trend of the mar-
ket environment as well as other economic factors. The economic and/or
political environment changes lead to the changes of regime of the surplus,
resulting in markedly different behavior of the system across regimes.
Defining a centered (or compensated) Poisson measure and applying gen-
eralized Itô’s rule, we can obtain the generator of the jump-diffusion process
with regime switching, and formulate a related martingale problem. Instead
of a single process, we have to deal with a collection of jump-diffusion pro-
cesses that are modulated by a continuous-time Markov chain. Suppose that
λ is positive such that λ∆ + o(∆) represents the probability of a switch of
regime in the interval [t, t + ∆), and π(·) is the distribution of the jump.
Then the generator of the underlying process can be written as
∂
GF (t, x, ι) = + L F (t, x, ι)
∂t
Z
+ λ[F (t, x + g(γ, x, ι), ι) − F (t, x, ι)]π(dγ) (12.2)
Γ
+ Q(t)F (t, x, ·)(ι), for each ι ∈ M,
326 12. Two-Time-Scale Switching Jump Diffusions
where
1 2 ∂2 ∂
LF (t, x, ι) = σ (t, x, ι) 2 F (t, x, ι) + f (t, x, ι) F (t, x, ι),
2 m ∂x ∂x
X 0 X
Q(t)F (t, x·)(ι) = aι` (t)F (t, x, `) = aι` (t)[F (t, x, `) − F (t, x, ι)].
`=1 `6=ι
(12.3)
By concentrating on time-scale separations, in this chapter, we treat two
cases. In the first one, the regime switching is significantly faster than the
dynamics of the jump diffusions, whereas in the second case, the diffusion
varies an order of magnitude faster than the switching processes.
The rest of the chapter is arranged as follows. Section 12.2 is devoted
to the case of fast switching. It begins with the precise formulation of the
problem. Then we derive weak convergence results and demonstrate that
the complicated problem can be “replaced” by a limit problem in which the
system coefficients are averaged out with respect to the stationary measures
of the switching process. In Section 12.3, we continue our study for the case
of fast varying diffusions. Again, by means of weak convergence methods,
we obtain a limit system. Section 12.4 gives remarks on specialization and
generalization of the asymptotic results. Section 12.5 gives remarks on nu-
merical approximation for switching-jump-diffusion processes. Section 12.6
concludes the chapter.
e
Both Q(t) b
and Q(t) are generators, where Q(t)/ε e represents the rapidly
b
changing part and Q(t) describes the slowly varying part. The slow and
fast components are coupled through weak and strong interactions in the
sense that the underlying Markov chain fluctuates rapidly within a single
group Mk of states and jumps less frequently from group Mk to Mj for
e
k 6= j. Suppose that the generator Q(t) has the form
e = diag Q
Q(t) e 1 (t), . . . , Q
e l (t) , (12.5)
Note that the last two terms are martingales. The operator of the regime-
switching jump-diffusion process is given by
ε ∂
G F (t, x, ι) = + L F (t, x, ι) + J(t, x, ι) + Qε (t)F (t, x, ·)(ι), (12.9)
∂t
where
Z
J(t, x, ι) = λ[F (t, x + g(γ, x, ι), ι) − F (t, x, ι)]π(dγ).
Γ
328 12. Two-Time-Scale Switching Jump Diffusions
b
Q(t) = diag(ν 1 (t), . . . , ν l (t))Q(t)diag(1
lm1 , . . . , 1lml ), (12.11)
(12.14)
330 12. Two-Time-Scale Switching Jump Diffusions
and the bound holds uniformly in t ∈ [0, T ]. Using the linear growth of
f (t, x, α) and σ(t, x, α) given in (A12.1) together with properties of stochas-
tic integrals, we obtain
Z t
E|X ε (t)|2 ≤ K + K E|X ε (u)|2 du.
0
as desired. 2
Next, we derive the tightness of Y ε (·).
Lemma 12.7. Assume that the conditions of Theorem 12.3 are satisfied.
Then Y ε (·) is tight in D([0, T ] : R × M), the space of functions that are
right-continuous, and have left limits endowed with the Skorohod topology.
Proof. Because αε (·) converges weakly to α(·) [176, p. 172], {αε (·)} is
tight. Therefore, to obtain the tightness of {Y ε (·)}, it suffices to derive the
tightness of {X ε (·)}.
Note that for any t > 0, s > 0, and any δ > 0 with 0 < s ≤ δ, we have
Eεt |X ε (t + s) − X ε (t)|2
Z t+s 2
ε ε ε
≤ KEt |f (u, X (u), α (u))|du
t
Z t+s 2 (12.18)
+ Eεt σ(u, X ε (u), αε (u))dw(u)
Zt t+s Z 2
+ Eεt g(γ, X ε (u− ), αε (u− ))N (du, dγ) ,
t Γ
332 12. Two-Time-Scale Switching Jump Diffusions
where Eεt denotes the conditional expectation with respect to the σ-algebra
generated by {αε (u), X ε (u) : u ≤ t}. Using the argument as in [103, p. 39],
since s ≤ δ,
Z t+s Z 2
ε
Et g(γ, X ε (u− ), αε (u− ))N (du, dγ) ≤ Ks = O(δ).
t Γ
As a result,
lim lim sup E|X ε (t + s) − X ε (t)|2 = 0.
δ→0 ε→0
By virtue of the tightness criterion (see Lemma A.28 of this book, and also
[43, Section 3.8, p. 132] or [102, p. 47] or [16]), the tightness of {X ε (·)}
follows. 2
Lemma 12.8. Assume the conditions of Theorem 12.3 are fulfilled. Sup-
pose that ζ(t, z) defined on [0, T ] × R is a real-valued function that is Lip-
schitz continuous in both variables, and that for each x ∈ Rr , |ζ(t, x)| ≤
K(1 + |x|). Denote
ε
vij (t) = vij (t, αε (t)) with vij (t, α) = I{α=sij } − νji (t)I{α∈Mi } .
Then for any i = 1, . . . , l, j = 1, . . . , mi ,
Z t 2
sup E ζ(u, X ε (u))vij ε
(u, αε (u))du → 0 as ε → 0. (12.20)
0<t≤T 0
Proof. The proof of this lemma is similar to that of Lemma 7.14 in [176].
Pick out 0 < ∆ < 1. For any t ∈ [0, T ], partition [0, t] into subintervals
of equal length ε1−∆ (without loss of generality, assume that `0 = t/ε1−∆
is an integer, otherwise, we can always take its integer part). Denote the
partition boundaries by tk = kε1−∆ for 0 ≤ k ≤ `0 . Define
ζeε (u) = ζ(tk , X ε (tk )), u ∈ [tk , tk+1 ), 0 ≤ k ≤ `0 − 1. (12.21)
In view of the process N (·), αε (·), and w(·), the same argument as in the
proof of the tightness yields
We claim that the term on the second line of (12.23) goes to 0. To see this
(recall that K is a generic positive constant), by using the Cauchy–Schwarz
inequality and the Lipschitz continuity of ζ(·) and (12.22),
Z t 2
E [ζ(u, X ε (u)) − ζeε (u)]vij
ε
(u, αε (u))du
0Z
t
≤T E|ζ(u, X ε (u)) − ζeε (u)|2 du
0
0 −1 Z tk+1
`X
≤K E[(u − tk )2 + |X ε (u) − X ε (tk )|2 ]du
k=0 tk
0 −1 Z tk+1
`X
≤K O(ε1−∆ )du
k=0 tk
→ 0 as ε → 0.
To estimate the term on the last line of (12.23), for each i = 1, . . . , l, and
j = 1, . . . , mi , define
Z t 2
ε
ηij (t) = E ζeε (u)vij
ε
(u, αε (u))du .
0
as desired. 2
To proceed, for each i ∈ M = {1, . . . , l} and each F (·, ·, i) ∈ C01,2 (C01,2
represents the class of functions that have compact support and that are
continuously differentiable with respect to t and twice continuously differ-
entiable with respect to x), consider the operator defined in (12.13). The
next lemma gives the characterization of the limit process as a solution of
a martingale problem.
334 12. Two-Time-Scale Switching Jump Diffusions
Lemma 12.9. Under the conditions of Theorem 12.3, the limit process
{Y (·)} is the solution of the martingale problem with operator G given by
(12.13).
is a martingale. To this end, we show that for any positive integer n0 , any
bounded and continuous functions h` (·), ` ≤ n0 , and any t, s, t` ≥ 0 with
t` ≤ t < t + s ≤ T ,
n0
Y
E h` (X(t` ), α(t` )) F (t + s, X(t + s), α(t + s)) − F (t, X(t), α(t))
`=1
Z !
t+s
− GF (u, X(u), α(u))du = 0.
t
(12.24)
Let us begin with the process Y ε (·). Define
l
X
Fb (t, x, α) = F (t, x, i)I{α∈Mi } for each α ∈ M.
i=1
Clearly, Fb(t, X ε (t), αε (t)) = F (t, X ε (t), αε (t)). Moreover, for each ι ∈ M,
Fb (·, ·, ι) ∈ C01,2 . The function Fb(·) allows us to conveniently use the avail-
able αε (·) process in lieu of the aggregated process αε (·). Consider the
operator G ε defined in (12.9). Because Y ε (·) is a Markov process,
Z t
Fb (t, X ε (t), αε (t)) − Fb(0, x, αε (0)) − G ε Fb (u, X ε (u), αε (u))du
0
is a martingale. Consequently,
n0
Y h
E h` (X ε (t` ), αε (t` )) Fb(t + s, X ε (t + s), αε (t + s)) − Fb (t, X ε (t), αε (t))
`=1 Z
t+s i
− G ε Fb (u, X ε (u), αε (u))du = 0.
t
(12.25)
We proceed to obtain the limit in (12.25) as ε → 0.
12.2 Fast-Varying Switching 335
First, by the weak convergence Y ε (·) to Y (·), the definition of Fb (·), and
the Skorohod representation, as ε → 0,
n0
Y
E h` (X ε (t` ), αε (t` ))[Fb(t + s, X ε (t + s), αε (t + s)) − Fb (t, X ε (u), αε (t))]
`=1
n0
Y
→E h` (X(t` ), α(t` ))[F (t + s, X(t + s), α(t + s)) − F (t, X(t), α(t))].
`=1
(12.26)
The definition of G ε leads to
n0
Y Z t+s
E ε
h` (X (t` ), α (t` )) ε
G Fb(u, X ε (u), αε (u))du
ε
t
`=1
n0
"Z
Y t+s
∂ b
ε ε
=E h` (X (t` ), α (t` )) F (u, X ε (u), αε (u))du
t ∂u
`=1
Z t+s
+ Fbx (u, X (u), α (u))f (u, X ε (u), αε (u))du
ε ε
Zt t+s
+ [Fbxx (u, X ε (u), αε (u))σ 2 (u, X ε (u), αε (u))]du
Zt t+s
+ Qε (u)Fb(u, X ε (u), ·)(αε (u))du
Z t #
t+s
+ J(u, X ε (u), αε (u))du .
t
Note that
n0
Y Z t+s
E h` (X ε (t` ), αε (t` )) Fbx (u, X ε (u), αε (u))f (u, X ε (u), αε (u))du
`=1 t
mi
l X
X n0
Y
= E h` (X ε (t` ), αε (t` ))
i=1 j=1 `=1
Z t+s
× Fbx (u, X ε (u), sij )f (u, X ε (u), sij )I{αε (u)=sij } du
t
mi
l X
X n0
Y
= E h` (X ε (t` ), αε (t` ))
i=1 j=1 `=1
Z t+s
× Fbx (u, X ε (u), sij )f (u, X ε (u), sij )νji (u)I{αε (u)=i} du
t
mi
l X
X n0
Y
+ E h` (X ε (t` ), αε (t` ))
"i=1
Z
j=1 `=1
t+s
× Fbx (u, X ε (u), sij )f (u, X ε (u), sij )
t
#
i
× I{αε (u)=sij } − νj (u)I{αε (u)=i} du .
336 12. Two-Time-Scale Switching Jump Diffusions
n0
Y Z t+s
E ε ε
h` (X (t` ), α (t` )) Fbx (u, X ε (u), sij )f (u, X ε (u), sij )
`=1 t
2
× I{αε (u)=sij } − νji (u)I{αε (u)=i} du
Z t+s
≤ KE Fbx (u, X ε (u), sij )f (u, X ε (u), sij )
t
2
× I{αε (u)=sij } − νji (u)I{αε (u)=i} du
→ 0 as ε → 0.
In view of [176, Lemma 2.4] and similar to [176, Theorem 7.30], it can
be shown that
n0
Y Z t+s
E ε ε
h` (X (t` ), α (t` )) Fbx (u, X ε (u), αε (u))f (u, X ε (u), αε (u))du
`=1 t
mi
l X
X n0
Y
ε→0
−→ E h` (X(t` ), α(t` ))
i=1 j=1 `=1
Z t+s
× Fbx (u, X(u), sij )f (u, X(u), sij )νji (u)I{α(u)=i} du
t
l
X Yn0
= E h` (X(t` ), α(t` ))
i=1
Z t+s
`=1
× Fx0 (u, X(u), i)f (u, X(u), i)I{α(u)=i} du
t Z t+s
n0
Y
0
=E h` (X(t` ), α(t` )) Fx (u, X(u), α(u))f (u, X(u), α(u))du .
`=1 t
(12.27)
12.2 Fast-Varying Switching 337
as ε → 0 and
n0
Y Z t+s
∂ b
E h` (X ε (t` ), αε (t` )) F (u, X ε (u), αε (u))du
t ∂u
`=1
n0
Y Z t+s
∂
→ h` (α(t` ), X(t` )) F (u, α(u), X(u), α(u))du ,
t ∂u
`=1
(12.29)
as ε → 0.
Next, since
e i (u)1lm = 0 for each i = 1, . . . , l,
Q i
Therefore, we have
n0
Y Z t+s
E h` (X ε (t` ), αε (t` )) Qε (u)Fb(u, X ε (u), ·)(αε (u))du
`=1 t
n0
Y Z t+s
=E h` (X ε (t` ), αε (t` )) b Fb(u, X ε (u), ·)(αε (u))du
Q(u)
`=1 t
mi
l X
X n0
Y
= E h` (X ε (t` ), αε (t` ))
i=1 j=1 `=1
Z t+s
× b Fb (u, X ε (u), ·)(sij )νji (u)I{αε (u)=i} du
Q(u)
t
mi
l X
X n0
Y
+ E h` (X ε (t` ), αε (t` ))
"i=1
Z
j=1 `=1
t+s
× b Fb(u, X ε (u), ·)(sij )
Q(u)
t
#
×[I{αε (u)=sij } − νji (u)I{αε (u)=i} ]du .
338 12. Two-Time-Scale Switching Jump Diffusions
mi
l X
X n0
Y
E h` (X ε (t` ), αε (t` ))
i=1 j=1 `=1
Z t+s
× b Fb (u, X ε (u), ·)(sij )νji (u)I{αε (u)=i} du
Q(u)
t
mi
l X
X n0
Y
= E h` (X ε (t` ), αε (t` ))
i=1 j=1 `=1
Z t+s
× b
Q(u)F (u, X ε (u), ·)(i)I{αε (u)=i} du
t
l
X n0
Y Z t+s
→ E h` (X(t` ), α(t` )) Q(u)F (u, X(u), ·)(i)I{α(u)=i} du
i=1 `=1 t
n0
Y Z t+s
=E h` (X(t` ), α(t` )) Q(u)F (u, X(u), ·)(α(u))du ,
`=1 t
as ε → 0.
Arguing along the same line as in the above estimates, we obtain
Z t+s
λ[Fb(u, X ε (u− ) + g(γ, X ε (u− ), αε (u− )), αε (u− ))
t
− Fb(u, X ε (u− ), αε (u− ))]π(dγ)
Xl Xmi Z t+s
= λ[Fb (u, X ε (u− ) + g(γ, X ε (u− ), sij ), sij )
i=1 j=1 t
− Fb(u, X ε (u− ), sij )]π(dγ)[I{αε (u− )=sij } − νji (u− )I{αε (u− )∈Mi } ]
Xl Xmi Z t+s
= λ[Fb (u, X ε (u− ) + g(γ, X ε (u− ), sij ), sij )
i=1 j=1 t
f1 (·), σ(·), and σ1 (·). That is, both the diffusion process z ε (·) and the
jump diffusion X ε (·) are time homogeneous. Treating z ε as a parameter,
the generator for X ε (·) can be written as
Z
G F (x, ι) = L F (x, ι) + λ[F (x + g(γ, x, z ε , ι), ι) − F (x, ι)]π(dγ)
ε ε
Γ
+ QF (x, ·)(ι), for each ι ∈ M,
(12.32)
where
1 2 ∂2 ∂
Lε F (x, ι) = σ (x, ι, z ε ) 2 F (x, ι) + f (x, ι, z ε ) F (x, ι),
2m ∂x ∂x
X 0 X (12.33)
QF (x, ·)(ι) = aι` F (x, `) = aι` [F (x, `) − F (x, ι)].
`=1 `6=ι
Remark 12.10. Part (b) yields that the fast-varying process z ε (·) is a
so-called periodic diffusion; see [10, 85] among others. This condition guar-
antees that there is an invariant density µ(x, z) (for a fixed x). In [85],
under suitable conditions, it was proved that not only an invariant mea-
sure exists, but also asymptotic expansions of the transition density can be
constructed. The choice of periodicity 1 is more or less for convenience; we
could in fact use other positive constants as the periodicity. It seems to be
more instructive to use simpler conditions as in the current setup.
for the last term (12.18), using the estimate, the linear growth in x, for
the first two terms on the right side of (12.18), and by virtue of Gronwall’s
inequality, we can show that
where Z 1
fb(x, α) = f (x, α, z)µ(x, z)dz,
0Z
1
gb(γ, x, α) = g(γ, x, α, z)µ(x, z)dz,
Z 10 (12.35)
σb2 (x, α) = σ 2 (x, α, z)µ(x, z)dz,
0
p
σb(x, α) = σ b2 (x, α).
where v(·) is the Brownian motion given in (12.31). Then the equation for
z ε (t) may be written as
Z τ Z τ
ε ε ε
ze (τ ) = z0 + f1 (X (εu), ze (u))du + σ1 (X ε (εu), zeε (u))de
v (u).
0 0
(12.36)
Note that as ε → 0, τ → ∞. Note also that X ε (ετ ) is slowly varying
whereas zeε (τ ) is fast changing. Intuitively, it tells us that we can treat
X ε (ετ ) as if it were a “constant” in a small interval. Taking x as a fixed
parameter, in what follows, we consider the following fixed-x process (see
[103] for an explanation of the fixed-x process),
Z τ Z τ
zeε,x (τ ) = z0 + f1 (x, zeε,x (u))du + σ1 (x, zeε,x (u))de
v (u),
0 0
representation, (X ε (·), α(·)) converges to (X(·), α(·)) w.p.1 and the conver-
gence is uniform on each bounded set. We proceed to show that the limit
is a solution of the martingale problem with generator Gb given by
where
Z
b α) =
J(x, λ[F (x + gb(γ, x, α), α) − F (x, α)]π(dγ),
Γ
Z t
F (X(t), α(t)) − F (x, α) − b (X(u), α(u))du
GF (12.40)
0
n0
Y
E h` (X ε (t` ))[F (X ε (t + s), α(t + s)) − F (X ε (t), α(t))]
`=1
n0
Y
→E h` (X(t` ))[F (X(t + s), α(t + s)) − F (X(t), α(t))] as ε → 0.
`=1
(12.41)
In addition,
n0
Y
E h` (X ε (t` )) [F (X ε (t + s), α(t + s)) − F (X ε (t), α(t))
`=1 Z t+s
ε ε
− G F (X (u), α(u))du = 0.
t
344 12. Two-Time-Scale Switching Jump Diffusions
n0
Y Z t+s
ε ε ε
E h` (X (t` )) G F (X (u), α(u))du
t
`=1
n0
"Z
Y t+s
∂F (X ε (u), α(u))
=E h` (X ε (t` )) f (X ε (u), α(u), z ε (u))
t ∂x
`=1
2 ε ε
1 ∂ F (X (u), α(u), z (u))
+ σ 2 (X ε (u), α(u), z ε (u))
2Z ∂x2
+ λ[F (X (u) + g(γ, X (u), α(u), z ε (u)), α(u))
ε ε
Γ
− F (X ε (u), α(u))]π(dγ)
#
+ QF (X ε (u), α(u)) du .
(12.42)
Consider the last term in (12.42). Using the weak convergence, the Sko-
rohod representation, and the continuity of the function F (·, α) for each
α ∈ M, it can be shown that as ε → 0,
n0
Y Z t+s
E h` (X ε (t` )) QF (X ε (u), α(u))du
`=1 t
n0
Y Z t+s (12.43)
→E h` (X(t` )) QF (X(u), α(u))du .
`=1 t
Z t+s (t+s)/δε −1
X Z lδε +δε
b ε (u)du = 1 b ε (u)du,
H δε H (12.44)
t δε lδε
l=t/δε
ε
Flδ ε
= {X ε (u), α(u), z ε (u) : 0 ≤ u ≤ lδε }.
12.3 Fast-Varying Diffusion 345
Similarly,
Z tlε +Tε
1
E Eεlδε f (X ε (lδε ), α(εu), zeε (u))]
Tε tlε
∂F (X ε (εu), α(εu)) ∂F (X ε (lδε ), α(εu))
× − du
∂x ∂x
Z tlε +Tε
1
≤K E|X ε (εu) − X ε (lδε )|du
Tε tlε
≤K sup E|X ε (εu) − X ε (lδε )| → 0 as ε → 0.
lδε ≤εu≤lδε +δε
Thus
Z lδε +δε
1 ∂F (X ε (u), α(u))
Eεlδε f (X ε (u), α(u), z ε (u)) du
δε lδε ∂x
Z tlε +Tε
1 ∂F (X ε (lδε ), α(εu))
= Eεlδε f (X ε (lδε ), α(εu), zeε (u)) du + o(1),
Tε tlε ∂x
ε
Flδ ε
, and the independence of α(·) and z ε (·),
Z tlε +Tε
1 ∂F (X ε (lδε ), α(εu))
Eεlδε f (X ε (lδε ), α(εu), zeε (u)) du
Tε ε tl ∂x
X 1 Z tlε +Tε ∂F (X ε (lδε ), j)
= Eεlδε f (X ε (lδε ), j, zeε (u)) I{α(εu)=j} du
Tε tlε ∂x
j∈M
X 1 Z tlε +Tε ∂F (X ε (lδε ), j)
= Eεlδε f (X ε (lδε ), j, zeε (u))
Tε tlε ∂x
i,j∈M
1, if X ε (lδε ) − Xn∆ → 0,
I{X ε (lδε )∈Bn∆ } →
0, otherwise.
In view of (12.36) and the existence of the unique invariant density µ(x, ·)
for each x, we have that
n∆
XX
I{X ε (lδε )∈Bn∆ } I{α(lδε )=i}
i∈M n=1
Z tlε +Tε
1 ∂F (Xn∆ , i)
× Eεlδε f (Xn∆ , i, zeε (u)) du
Tε tlε ∂x
XX n∆
= I{X ε (lδε )∈Bn∆ } I{α(lδε )=i}
i∈M n=1
Z tlε +Tε
Z 1
1 ∂F (Xn∆ , i)
× f (Xn∆ , i, z)Eεlδε I{ez ε,Xn∆ (u)∈dz} du
Tε tlε 0 ∂x
XX n∆
= I{X ε (lδε )∈Bn∆ } I{α(lδε )=i}
i∈M n=1
Z tlε +TεZ 1
1 ∆ ∂F (Xn∆ , i)
∆
× f (Xn∆ , i, z)P(e
z ε,Xn (u) ∈ dz|e
z ε,Xn (tlε ))du
Tε tlε 0 ∂x
XZ 1 ∂F (X(u), i)
→ I{α(u)=i} f (X(u), i, z)µ(X(u), z)dz
∂x
i∈M 0
∂F (X(u), α(u))
= fb(X(u), α(u)) .
∂x
Thus, as ε → 0,
n0
Y Z t+s
∂F (X ε (u), α(u))
E h` (X ε (t` )) f (X ε (u), α(u), z ε (u)) du
t ∂x
`=1
n0
Y Z t+s
b ∂F (X(u), α(u))
→E h` (X(t` )) f (X(u), α(u)) du.
t ∂x
`=1
348 12. Two-Time-Scale Switching Jump Diffusions
and
n0
Y Z t+s Z
ε
E h` (X (t` )) λ[F (X ε (u) + g(γ, X ε (u), α(u), z ε (u)), α(u))
t Γ
`=1
− F (X ε (u), α(u))]π(dγ)
Yn0 Z t+s Z
→E h` (X(t` )) λ[F (X(u) + gb(γ, X(u), α(u)), α(u))
t
`=1 Γ
− F (X(u), α(u))]π(dγ) .
becomes
Z t Z tZ
X(t) = x + f (s, X(s), α(s))ds + g(γ, X(s− ), α(s− ))N (ds, dγ).
0 0 Γ
Example 12.15. We consider the model in (12.7) again with Q(t) e = Q(t)
that is weakly irreducible. Then the limit system is given by Corollary 12.5.
The limit does not involve the Markov switching process. Thus the com-
plexity reduction is more pronounced. Further specification leads to the
classical risk model.
Example 12.16. Consider the system given by (12.31) with σ(t, x, α) ≡ 0.
Then similar to Example 12.14, Theorem 12.3 still holds with the limit
system being a Markovian modulated jump process.
For convenience, with a slight abuse of notation, we omitted the floor func-
tion notation above. Henceforth, for instance, we use t/∆ to denote the
integer part of t/∆ here. We state a result, whose proof is along the line of
martingale averaging. Further details can be found in [173].
Theorem 12.18. Under the conditions in (A12.1), (X ∆ (·), α∆ (·)) con-
verges weakly to (X(·), α(·)) as ∆ → 0 such that (X(·), α(·)) is the solution
of the martingale problem with operator L.
As in Chapter 5, Theorem 12.18 implies that the algorithm we con-
structed is convergent. The limit is nothing but the solution of (12.1). As
in Chapter 5, one of the variations of the algorithm is to use a sequence of
decreasing step sizes. The modifications are as follows. LetP{∆n } be a se-
∞
quence of nonnegative real numbers such that ∆n → 0 and n=0 ∆n = ∞.
352 12. Two-Time-Scale Switching Jump Diffusions
For example, we may take ∆n = 1/n, or ∆n = 1/nγ for some 0 < γ < 1.
Define
n−1
X
tn = ∆l , and αn = α(tn ), n ≥ 0.
l=0
P n,n+1 = (pn,n+1
ij )m0 ×m0 = exp((tn+1 − tn )Q) = exp(∆n Q). (12.50)
It is clear that now the Markov chain αn is not time homogeneous. The
approximate solution for the stochastic differential equation with jumps
and regime switching (12.1) is given by
n n p
X X
Xn+1 = X0 +
f (X , α )∆ + ∆k σ(Xk , αk )ξk
k k k
k=0 k=0
X
+ g(ψj , Xj−1 , αj ), (12.52)
τej ≤n
X0 = x, α0 = α.
With the algorithm proposed, we can then proceed to study its perfor-
mance. The proof is along the same lines as that of the constant stepsize
algorithm. The interested reader is referred to [173]; see also Chapters 5
and 6 of this book for further reading.
12.6 Notes
As a continuation of our study, based on the work Yin and Yang [175], this
chapter has been devoted to two-time-scale jump diffusions with regime
switching. Roughly, the fast-changing driving processes can be treated as
a noise, whose stationary measure does exist. Under suitable conditions,
the slow process is averaged out with respect to the stationary measure of
the fast-varying process. Such an idea has been used in the literature. We
refer the reader to Khasminskii [82], Papanicolaou, Stroock, and Varadhan
[131], Khasminskii and Yin [86, 88, 89], Pardoux and Veretennikov [132],
and references therein for related work in diffusions, and Yin [165] for that
of switching diffusions.
The limit results obtained in this chapter can be useful for applications.
For example, one may use such results in a subsequent study for obtaining
bounds on ruin probability in risk management. It is conceivable that this
12.6 Notes 353
355
356 Appendix A. Appendix
pk,k+1
ij = P(αk+1 = j|αk = i)
= P(αk+1 = j|α0 = i0 , . . . , αk−1 = ik−1 , αk = i),
(n)
pij = P(Xn = j|X0 = i).
Then P (n) = P n . That is, the n-step transition matrix is simply the matrix
P to the nth power. Note that
P
(a) pij ≥ 0, j pij = 1, and
P k − P ≤ c0 λk for k = 1, 2, . . . ,
where P = 1lm0 π, 1lm0 = (1, . . . , 1)0 ∈ Rm0 ×1 , and π = (π1 , . . . , πm0 ) is the
stationary distribution of αk . This implies, in particular,
lim P k = 1lm0 π.
k→∞
.
Define Qc = (P − I ..1lm0 ) ∈ Rm0 ×(m0 +1) . Consider (A.1) together with
Pm 0
the condition F 1lm0 = i=1 Fi = Fb , which may be written as F Qc = Gc
..
where Gc = (G.Fb). Because for each t, (A.9) has a unique solution, it
follows that Qc (t)Q0c (t) is a matrix with full rank; therefore, the equation
F [Qc Q0c ] = Gc Q0c (A.2)
has a unique solution, which is given by F = Gc Q0c [Qc Q0c ]−1 .
pij (t, s) ≥ 0, i, j ∈ M,
X
pij (t, s) = 1, i ∈ M,
j∈M
X
pij (t, s) = pik (ς, s)pkj (t, ς), i, j ∈ M.
k∈M
is a martingale.
is a martingale.
Proof: For a proof, see Yin and Zhang [176, Lemma 2.4]. 2
For any given Q(t) satisfying the q-property, there exists a Markov chain
α(·) generated by Q(t). If Q(t) = Q, a constant matrix, the idea of Ethier
and Kurtz [43] can be utilized for the construction. For time-varying gen-
erator Q(t), we need to use the piecewise-deterministic process approach
as described in Davis [30], to define the Markov chain α(·).
360 Appendix A. Appendix
The discussion below is taken from that of Yin and Zhang [176], which
was originated in the work of Davis [30]. Let 0 = τ0 < τ1 < · · · < τl < · · · be
a sequence of jump times of α(·) such that the random variables τ1 , τ2 − τ1 ,
. . ., τk+1 − τk , . . . are independent. Let α(0) = i ∈ M. Then α(t) = i on
the interval [τ0 , τ1 ). The first jump time τ1 has the probability distribution
Z Z t
P(τ1 ∈ B) = exp qii (s)ds (−qii (t)) dt,
B 0
In general, α(t) = α(τl ) on the interval [τl , τl+1 ). The jump time τl+1 has
the conditional probability distribution
(d) Assume further that Q(t) is continuous in t. Then P (t, s) also satisfies
the backward differential equation
∂P (t, s) = −Q(s)P (t, s), t ≥ s,
∂s (A.7)
P (s, s) = I.
Proof. For (a)–(c), see Yin and Zhang [176, Theorem 2.5]. As for (d), see
[26, p. 402]. 2
Note that frequently, working with s ∈ [0, T ], the backward equations are
written slightly differently by using reversed time τ = T − s. In this case,
the minus sign in (A.7) disappears. Suppose that α(t), t ≥ 0, is a Markov
chain generated by an m0 × m0 matrix Q(t). The notions of irreducibility
and quasi-stationary distribution are given next.
(a) A generator Q(t) is said to be weakly irreducible if, for each fixed
t ≥ 0, the system of equations
ν(t)Q(t) = 0,
m0
X (A.8)
νi (t) = 1
i=1
has a unique solution ν(t) = (ν1 (t), . . . , νm0 (t)) and ν(t) ≥ 0.
the weak irreducibility requires only λ(t) + µ(t) > 0, whereas the irre-
ducibility requires that both λ(t) and µ(t) be positive. Such a definition
is convenient for many applications (e.g., the manufacturing systems men-
tioned in Khasminskii, Yin, and Zhang [91, p. 292]).
Definition A.8 (Quasi-Stationary Distribution). For t ≥ 0, ν(t) is termed
a quasi-stationary distribution if it is the unique solution of (A.8) satisfying
ν(t) ≥ 0.
(a) The homogeneous equation (γI − A)f = 0 has only the zero solution,
in which case γ ∈ ρ(A), the resolvent set of A, (γI −A)−1 is bounded,
and the inhomogeneous equation (γI − A)f = g also has one solution
f = (γI − A)−1 g, for each g ∈ B.
A.3 Fredholm Alternative and Ramification 363
Qc = b − νb1l. (A.11)
Moreover, because α1 (t) has a finite state space, the convergence above
takes place exponentially fast; see [176, Appendix] and references therein.
Thus, h is well defined. We proceed to show that h, in fact, is a solution of
(A.11).
By direct calculation, it is seen that
Z ∞ Z ∞
Qh = (Qνb1l − QP (t)b)dt = − QP (t)bdt
0Z ∞ 0
dP (t) ∞
=− bdt = −P (t)b = −1lνb + b,
0 dt 0
where P (t) = (pij (t)) is the transition matrix satisfying the Kolmogorov
backward equation (d/dt)P (t) = QP (t) with P (0) = I and limt→∞ P (t) =
1lν. Thus, the vector h satisfies equation (A.11) and hence (A.10). By using
the result proved earlier, any solution of (A.10) can be represented by
c = h + γ1l for some γ ∈ R.
Finally, we verify that any solution of (A.10) can be written as c = h 0 +α1l
for α ∈ R, where h0 is the unique solution of (A.10) satisfying νh0 = 0.
In fact, we have shown that h defined in (A.12) solves (A.10) and any
A.3 Fredholm Alternative and Ramification 365
Qa h = bb. (A.13)
It can be shown that Q0a Qa has full rank m0 due to the irreducibility of Q,
and the solution of (A.13) can be represented by
A.4.1 Martingales
Many applications involving stochastic processes depend on the concept of
the martingale. The definition and properties of discrete-time martingales
can be found in Breiman [19, Chapter 5], Chung [28, Chapter 9], and Hall
and Heyde [63] among others. This section provides a brief review.
Discrete-Time Martingales
Definition A.14. Suppose that {Fn } is a filtration, and {Xn } is a se-
quence of random variables. The pair {Xn , Fn } is a martingale if for each
n,
(a) Xn is Fn -measurable;
(d) The Doob inequality (see Hall and Heyde [63, p.15]) states that for
each p > 1,
p
1/p p 1/p
E |Xn | ≤ E max |Xj | ≤ qE1/p |Xn |p ,
1≤j≤n
−1 −1
where p +q = 1.
(e) The Burkholder inequality (see Hall and Heyde [63, p.23]) is: For
1 < p < ∞, there exist constants K1 and K2 such that
n
X p/2 n
X p/2
K1 E yj2 ≤ E|Xn |p ≤ K2 E yj2 ,
j=1 i=j
where Yn = Xn − Xn−1 .
368 Appendix A. Appendix
= f (αn ) = Xn a.s.
Therefore, {Xn , Fn } is a martingale. Note that if M is finite, the bound-
edness of {f (i) : i ∈ M} is not needed.
As explained in Karlin and Taylor [80], one of the widely used ways of
constructing martingales is through the utilization of eigenvalues and eigen-
vectors of a transition matrix. Again, let {αn } be a discrete-time Markov
chain with transition matrix P . Recall that a column vector f is a right
eigenvector of P associated with an eigenvalue λ ∈ C, if P f = λf . Let f
be a right eigenvector of P satisfying E|f (αn )| < ∞ for each n. For λ 6= 0,
define Xn = λ−n f (αn ). Then {Xn } is a martingale.
A.4 Martingales, Gaussian Processes, and Diffusions 369
Continuous-Time Martingales
Next, let us denote the space of Rr -valued continuous functions on [0, T ]
by C([0, T ]; Rr ), and the space of functions that are right-continuous with
left-hand limits endowed with the Skorohod topology by D([0, T ]; Rr ); see
Definition A.18. Consider X(·) = {X(t) ∈ Rr : t ≥ 0}. If for each t ≥ 0,
X(t) is an Rr random vector, we call X(·) a continuous-time stochastic
process and write it as X(t), t ≥ 0, or simply X(t) if there is no confusion.
A process X(·) is adapted to a filtration {Ft }, if for each t ≥ 0, X(t) is an
Ft -measurable random variable; X(·) is progressively measurable if for each
t ≥ 0, the process restricted to [0, t] is measurable with respect to the σ-
algebra B[0, t] × Ft in [0, t] × Ω, where B[0, t] denotes the Borel sets of [0, t].
A progressively measurable process is measurable and adapted, whereas
the converse is not generally true. However, any measurable and adapted
process with right-continuous sample paths is progressively measurable.
Frequently, we need to work with a stopping time for applications. Con-
sider (Ω, F, P ) with a filtration {Ft }. A stopping time τ is a nonnegative
random variable satisfying {τ ≤ t} ∈ Ft for all t ≥ 0.
A stochastic process {X(t) : t ≥ 0} (real- or vector-valued) is a martin-
gale on (Ω, F, P ) with respect to {Ft } if:
holds for all 0 ≤ s ≤ t and for any Borel set A. A slightly more general
definition allows b(·) and σ(·) to be Ft -measurable processes. Nevertheless,
the current definition is sufficient for our purpose.
A.5 Weak Convergence 371
for every bounded and continuous function f (·) on S. Suppose that {Xk }
and X are random variables associated with Pk and P, respectively. The
sequence Xk converges to X weakly if for any bounded and continuous
function f (·) on S, Ef (Xk ) → Ef (X) as k → ∞.
Analogous definitions and results are available for D([0, T ]; F), where F is
a metric space; see Ethier and Kurtz [43] and Billingsley [16] for related
references. Although we frequently work with D([0, T ]; Rr ) in this book, the
following results are often stated with respect to the space D([0, ∞); Rr ).
This enables us to apply them to t ∈ [0, T ] for any T > 0.
e X
P( e X
ek ∈ B) = P(Xk ∈ B), and P( e ∈ B) = P(X ∈ B)
satisfying
ek = X
lim X e a.s.
k→∞
and Z t
f (Y (t)) − f (Y (0)) − Af (Y (s))ds, t ≥ 0
0
are martingales and X(t) and Y (t) have the same distribution for each
t ≥ 0, X(·) and Y (·) have the same distribution on D([0, ∞); Rr ).
dX ε (t)
= F ε (t),
dt
and for each T < ∞, {F ε (t) : 0 ≤ t ≤ T } be uniformly integrable. If the
set of initial values {X ε (0)} is tight, then {X ε (·)} is tight in C([0, ∞); Rr ).
Proof: The proof is essentially in Billingsley [16, Theorem 8.2] (see also
Kushner [102, p. 51, Lemma 7]). 2
Define the notion of “p-lim” and an operator Aε as in Ethier and Kurtz
[43]. Suppose that X ε (·) are defined on the same probability space. Let Ftε
be the minimal σ-algebra over which {X ε (s), ξ ε (s) : s ≤ t} is measurable
and let Eεt denote the conditional expectation given Ftε . Denote
ε
M = f : f is real-valued with bounded support and is
progressively measurable w.r.t. {Ftε }, sup E|f (t)| < ∞ .
t
ε
Let g(·), f (·), f δ (·) ∈ M . For each δ > 0 and t ≤ T < ∞, f = p − limδ f δ
if
sup E|f δ (t)| < ∞,
t,δ
then
lim E|f (t) − f δ (t)| = 0 for each t.
δ→0
The function f (·) is said to be in the domain of Aε ; that is, f (·) ∈ D(Aε ),
and Aε f = g, if
ε
Et f (t + δ) − f (t)
p − lim − g(t) = 0.
δ→0 δ
A.5 Weak Convergence 375
If f (·) ∈ D(Aε ), then Ethier and Kurtz [43] or Kushner [102, p. 39] implies
that Z t
f (t) − Aε f (u)du
0
is a martingale, and
Z t+s
Eεt f (t + s) − f (t) = Eεt Aε f (u)du a.s.
t
Remark A.29. In lieu of (A.24), one may verify the following condition
(see Kurtz [95, Theorem 2.7, p. 10]). Suppose that for each η > 0 and
rational t ≥ 0 there is a compact set Γt,η ⊂ Rr such that
where λ is known as the impulse rate of ψ(·) and/or the jump rate of
N (·, Γ), and π(H) is the jump distribution in the sense that
P(ψ(t) ∈ H|ψ(t) 6= 0, ψ(u), u < t) = π(H).
The values and times of the impulses can be recovered from the integral
Z tZ X
G(t) = γN (ds, dγ) = ψ(s).
0 Γ s≤t
A.7 Miscellany
Suppose that A is an r × r square matrix. Denote the collection of eigenval-
ues of A by Λ. Then the spectral radius of A, denoted by ρ(A), is defined
by ρ(A) = maxλ∈Λ |λ|. Recall that a matrix with real entries is a positive
matrix if it has at least one positive entry and no negative entries. If every
entry of A is positive, we call the matrix strictly positive. Likewise, for a
vector x = (x1 , . . . , xr ), by x ≥ 0, we mean xi ≥ 0 for i = 1, . . . , r; by
x > 0, we mean all entries xi > 0.
By a multi-index ζ, we mean a vector ζ = (ζ1 , . . . , ζr ) with nonnegative
integer components with |ζ| defined as |ζ| = ζ1 + · · · + ζr . For a multi-index
ζ, Dxζ is defined to be
∂ζ ∂ |ζ|
Dxζ = = ; (A.28)
∂xζ ∂x11 . . . ∂xζrr
ζ
The following inequalities are widely used. The first of them is known as
the Gronwall inequality and second one is the so-called generalized Gron-
wall inequality. Both of them can be found in [61, p. 36].
Lemma A.30. If γ ∈ R, β(t) ≥ 0, and ϕ(t) are continuous real-valued
functions for a ≤ t ≤ b, which satisfy
Z t
ϕ(t) ≤ γ + β(s)ϕ(s)ds, t ∈ [a, b],
a
378 Appendix A. Appendix
then Z t
ϕ(t) ≤ γ exp( β(s)ds), t ∈ [a, b].
a
Lemma A.31. Suppose that ϕ(·) and γ(·) are real-valued continuous func-
tions on [a, b], that β(t) ≥ 0 is integrable on [a, b], and that
Z t
ϕ(t) ≤ γ(t) + β(s)ϕ(s)ds, t ∈ [a, b].
a
Then
Z t Z t
ϕ(t) ≤ γ(t) + β(s)γ(s) exp( β(u)du)ds, t ∈ [a, b].
a s
Lemma A.32. LetR ψ : [0, ∞) 7→ Rr×d for some positive integers r and d,
∞
and suppose that E 0 |ψ(s)|2 < ∞. Define
Z t Z t
Z(t) = ψ(s)dw(s) and a(t) = |ψ(s)|2 ds,
0 0
A.8 Notes
One may find a nonmeasure-theoretic introduction to stochastic processes
in Ross [141]. The two volumes by Karlin and Taylor [80, 81] provide
an introduction to discrete-time and continuous-time Markov chains. Ad-
vanced treatments of Markov chains can be found in Chung [28] and Re-
vuz [142]. A book that deals exclusively with finite-state Markov chains is
by Iosifescu [76]. The book of Meyn and Tweedie [125] examines Markov
chains and their stability. Doob’s book [33] gives an introduction to stochas-
tic processes. Gihman and Skorohod’s three-volume work [55] provides a
comprehensive introduction to stochastic processes, whereas Liptser and
Shiryayev’s book [110] presents further topics such as nonlinear filtering.
References
[2] B.D.O. Anderson and J.B. Moore, Optimal Control: Linear Quadratic
Methods, Prentice Hall, Englewood Cliffs, NJ, 1990.
[3] A. Arapostathis, M.K. Ghosh, and S.I. Marcus, Harnack’s inequality for
cooperative weakly coupled elliptic systems, Comm. Partial Differential
Eqs., 24 (1999), 1555–1571.
[6] G.K. Basak, A. Bisi, and M.K. Ghosh, Stability of a random diffusion with
linear drift, J. Math. Anal. Appl., 202 (1996), 604–622.
[7] G.K. Basak, A. Bisi, and M.K. Ghosh, Stability of degenerate diffusions
with state-dependent switching, J. Math. Anal. Appl., 240 (1999), 219–
248.
[8] M.M. Benderskii and L.A. Pastur, The spectrum of the one-dimensional
Schrodinger equation with random potential, Mat. Sb. 82 (1972), 273–284.
[9] M.M. Benderskii and L.A. Pastur, Asymptotic behavior of the solutions of
a second order equation with random coefficients, Teorija Funkeii i Func-
tional’nyi Analiz, 22 (1973), 3–14.
379
380 References
[12] A. Bensoussan and J.L. Menaldi, Hybrid control and dynamic program-
ming, Dynamics Continuous Disc. Impulsive Sys., 3 (1997), 395–442.
[13] B. Bercu, F. Dufour, and G. Yin, Almost sure stabilization for feedback
controls of regime-switching linear systems with a hidden Markov chain,
IEEE Trans. Automat. Control, 54 (2009).
[15] R.N. Bhattacharya, Criteria for recurrence and existence of invariant mea-
sures for multidimensional diffusions, Ann. Probab., 6 (1978), 541–553.
[17] T. Björk, Finite dimensional optimal filters for a class of Ito processes with
jumping parameters, Stochastics, 4 (1980), 167–183.
[18] G.B. Blankenship and G.C. Papanicolaou, Stability and control of stochas-
tic systems with wide band noise, SIAM J. Appl. Math., 34 (1978), 437–
476.
[20] J. Buffington and R.J. Elliott, American options with regime switching,
Int. J. Theoret. Appl. Finance, 5 (2002), 497–514.
[21] P.E. Caines and H.-F. Chen, Optimal adaptive LQG control for systems
with finite state process parameters, IEEE Trans. Automat. Control, 30
(1985), 185–189.
[22] P.E. Caines and J.-F. Zhang, On the adaptive control of jump parameter
systems via nonlinear filtering, SIAM J. Control Optim., 33 (1995), 1758–
1777.
[24] Z.Q. Chen and Z. Zhao, Potential theory for elliptic systems, Ann. Probab.,
24 (1996), 293–319.
[25] Z.Q. Chen and Z. Zhao, Harnack inequality for weakly coupled elliptic
systems, J. Differential Eqs., 139 (1997), 261–282.
[27] Y.S. Chow and H. Teicher, Probability Theory, 3rd ed., Springer-Verlag,
New York, NY, 1997.
[28] K.L. Chung, Markov Chains with Stationary Transition Probabilities, 2nd
ed., Springer-Verlag, New York, NY, 1967.
[29] D.R. Cox and H.D. Miller, The Theory of Stochastic Processes, J. Wiley,
New York, NY, 1965.
[30] M.H.A. Davis, Markov Models and Optimization, Chapman & Hall, Lon-
don, UK, 1993.
[31] D.A. Dawson, Critical dynamics and fluctuations for a mean–field model
of cooperative behavior, J. Statist. Phys., 31 (1983), 29–85.
[32] G.B. Di Masi, Y.M. Kabanov, and W.J. Runggaldier, Mean variance hedg-
ing of options on stocks with Markov volatility, Theory Probab. Appl., 39
(1994), 173–181.
[33] J.L. Doob, Stochastic Processes, Wiley Classic Library Edition, Wiley, New
York, NY, 1990.
[35] F. Dufresne and H.U. Gerber, Risk theory for the compound Poisson pro-
cess that is perturbed by diffusion, Insurance; Math. Economics, 10 (1991),
51–59.
[36] F. Dufour and P. Bertrand, Stabilizing control law for hybrid modes, IEEE
Trans. Automat. Control, 39 (1994), 2354–2357.
[37] N.H. Du, R. Kon, K. Sato, and Y. Takeuchi, Dynamical behavior of Lotka-
Volterra competition systems: Non-autonomous bistable case and the effect
of telegraph noise, J. Comput. Appl. Math., 170 (2004), 399–422.
[38] E.B. Dynkin, Markov Processes, Vols. I and II, Springer-Verlag, Berlin,
1965.
[41] A. Eizenberg and M. Freidlin, On the Dirichlet problem for a class of second
order PDE systems with small parameter. Stochastics Stochastics Rep., 33
(1990), 111–148.
[43] S.N. Ethier and T.G. Kurtz, Markov Processes: Characterization and Con-
vergence, J. Wiley, New York, NY, 1986.
382 References
[44] W.H. Fleming and R.W. Rishel, Deterministic and Stochastic Optimal
Control, Springer-Verlag, New York, NY, 1975.
[45] W.H. Fleming and H.M. Soner, Controlled Markov Processes and Viscosity
Solutions, Springer-Verlag, New York, 1992.
[48] M.D. Fragoso and O.L.V. Costa, A unified approach for stochastic and
mean square stability of continuous-time linear ystems with Markovian
jumping parameters and additive disturbances, SIAM J. Control Optim.,
44 (2005), 1165–1191.
[50] J.P. Fouque, G. Papanicolaou, R.K. Sircar, and K. Solna, Singular pertur-
bations in option pricing, SIAM J. Appl. Math., 63 (2003), 1648–1665.
[52] M.K. Ghosh, A. Arapostathis, and S.I. Marcus, Ergodic control of switching
diffusions, SIAM J. Control Optim., 35 (1997), 1952–1988.
[53] I.I. Gihman and A.V. Skorohod, Introduction to the Theory of Random
Processes, W.B. Saunders, Philadelphia, PA, 1969.
[55] I.I. Gihman and A.V. Skorohod, Theory of Stochastic Processes, I, II, III,
Springer-Verlag, Berlin, 1979.
[59] G.H. Golub and C.F. Van Loan, Matrix Computations, 2nd ed., Johns
Hopkins University Press, Baltimore, MD, 1989.
References 383
[60] M.G. Garroni and J.L. Menaldi, Green Functions for Parabolic Second Or-
der Integro-Differential Equations, Pitman Research Notes in Math. Series,
No. 275, Longman, London, 1992.
[61] J.K. Hale, Ordinary Differential Equations, 2nd ed., R.E. Krieger, Malabar,
FL, 1980.
[62] J.K. Hale and E.P. Infante, Extended dynamical systems and stability the-
ory, Proc. Nat. Acad. Sci. 58 (1967), 405–409.
[63] P. Hall and C.C. Heyde, Martingale Limit Theory and Its Application,
Academic Press, New York, NY, 1980.
[64] F.B. Hanson, Applied Stochastic Processes and Control for Jump-
diffusions: Modeling, Analysis, and Computation, SIAM, Philadelphia, PA,
2007.
[65] Q. He and G. Yin, Invariant density, Liapunov exponent, and almost sure
stability of Markovian-regime-switching linear systems, preprint, 2009.
[68] J.P. Hespanha, A model for stochastic hybrid systems with application to
communication networks, Nonlinear Anal., 62 (2005), 1353–1383.
[69] J.C. Hull, Options, Futures, and Other Derivatives, 3rd ed., Prentice-Hall,
Upper Saddle River, NJ, 1997.
[70] J.C. Hull and A. White, The pricing of options on assets with stochastic
volatilities, J. Finance, 42 (1987), 281–300.
[71] V. Hutson and J.S. Pym, Applications of Functional Analysis and Operator
Theory, Academic Press, London, UK, 1980.
[73] A.M. Il’in, R.Z. Khasminskii, and G. Yin, Singularly perturbed switching
diffusions: rapid switchings and fast diffusions, J. Optim. Theory Appl.,
102 (1999), 555–591.
[74] A.M. Il’in, R.Z. Khasminskii, and G. Yin, Asymptotic expansions of solu-
tions of integro-differential equations for transition densities of singularly
perturbed switching diffusions, J. Math. Anal. Appl., 238 (1999), 516–539.
[75] J. Imae, J.E. Perkins, and J.B. Moore Toward time-varying balanced real-
ization via Riccati equations, Math. Control Signals Syst., 5 (1992), 313–
326.
384 References
[77] J. Jacod and A.N. Shiryayev, Limit Theorems for Stochastic Processes,
Springer-Verlag, New York, NY, 1980.
[79] I.I. Kac and N.N. Krasovskii, On the stability of systems with random
parameters, J. Appl. Math. Mech., 24 (1960), 1225–1246.
[80] S. Karlin and H.M. Taylor, A First Course in Stochastic Processes, 2nd
ed., Academic Press, New York, NY, 1975.
[81] S. Karlin and H.M. Taylor, A Second Course in Stochastic Processes, Aca-
demic Press, New York, NY, 1981.
[84] R.Z. Khasminskii and F.C. Klebaner, Long term behavior of solutions of
the Lotka-Volterra systems under small random perturbations, Ann. Appl.
Probab., 11 (2001), 952–963.
[85] R.Z. Khasminskii and G. Yin, Asymptotic series for singularly perturbed
Kolmogorov-Fokker-Planck equations, SIAM J. Appl. Math. 56 (1996),
1766–1793.
[90] R.Z. Khasminskii and G. Yin, Uniform asymptotic expansions for pricing
European options, Appl. Math. Optim., 52 (2005), 279–296.
[97] H.J. Kushner, Stochastic Stability and Control, Academic Press, New York,
NY, 1967.
[98] H.J. Kushner, The concept of invariant set for stochastic dynamical sys-
tems and applications to stochastic stability, in Stochastic Optimization
and Control, H.F. Karreman Ed., J. Wiley, New York, NY, 1968, 47–57.
[102] H.J. Kushner, Approximation and Weak Convergence Methods for Ran-
dom Processes, with Applications to Stochastic Systems Theory, MIT Press,
Cambridge, MA, 1984.
[104] H.J. Kushner and G. Yin, Stochastic Approximation and Recursive Algo-
rithms and Applications, 2nd ed., Springer-Verlag, New York, NY, 2003.
[105] O.A. Ladyzenskaja, V.A. Solonnikov, and N.N. Ural’ceva, Linear and
Quasi-linear Equations of Parabolic Type, Translations of Math. Mono-
graphs, Vol. 23, Amer. Math. Soc., Providence, RI, 1968.
[106] J.P. LaSalle, The extent of asymptotic stability, Proc. Nat. Acad. Sci., 46
(1960), 365.
[107] J.P. LaSalle and S. Lefschetz, The Stability by Liapunov Direct Method,
Academic Press, New York, NY, 1961.
386 References
[109] J.-J. Liou, Recurrence and transience of Gaussian diffusion processes, Ko-
dai Math. J., 13 (1990), 210–230.
[110] R.S. Liptser and A.N. Shiryayev, Statistics of Random Processes I & II,
Springer-Verlag, New York, NY, 2001.
[111] Y.J. Liu, G. Yin, Q. Zhang, and J.B. Moore, Balanced realizations of
regime-switching linear systems, Math. Control, Signals, Sys., 19 (2007),
207–234.
[112] K.A. Loparo and G.L. Blankenship, Almost sure instability of a class of
linear stochastic systems with jump process coefficients, in Lyapunov Ex-
ponents, Lecture Notes in Math., 1186, Springer, Berlin, 1986, 160–190.
[113] Q. Luo and X. Mao, Stochastic population dynamics under regime switch-
ing, J. Math. Anal. Appl., 334 (2007), 69–84.
[115] X. Mao, Stochastic Differential Equations and Applications, 2nd ed., Hor-
wood, Chichester, UK, 2007.
[121] X.R. Mao, C. Yuan, and G. Yin, Numerical method for stationary dis-
tribution of stochastic differential equations with Markovian switching, J.
Comput. Appl. Math., 174 (2005), 1–27.
[125] S.P. Meyn and R.L. Tweedie, Markov Chains and Stochastic Stability,
Springer-Verlag, London, UK, 1993.
[127] G.N. Milstein and M.V. Tretyakov, Stochastic Numerics for Mathematical
Physics, Springer-Verlag, Berlin, 2004.
[128] C.M. Moller, Stochastic differential equations for ruin probability, J. Appl.
Probab., 32 (1995), 74–89.
[133] J.E. Perkins, U. Helmke, and J.B. Moore, Balanced realizations via gradient
flow techniques, Sys. Control Lett., 14 (1990), 369–379.
[134] L. Perko, Differential Equations and Dynamical Systems, 3rd ed., Springer,
New York, NY, 2001.
[135] S. Peszat and J. Zabczyk, Strong Feller property and irreducibility for
diffusions on Hilbert spaces, Ann. Probab., 23 (1995), 157–172.
[143] W. Rudin, Real and Complex Analysis, 3rd ed., McGraw-Hill, New York,
NY, 1987.
[146] S.P. Sethi and Q. Zhang, Hierarchical Decision Making in Stochastic Man-
ufacturing Systems, Birkhäuser, Boston, 1994.
[148] S. Shokoohi, L.M. Silverman, and P.M. Van Dooren, Linear time variable
systems: Balancing and model reduction, IEEE Trans. Automat. Control,
28 (1983), 810–822.
[152] Q.S. Song and G. Yin, Rates of convergence of numerical methods for
controlled regime-switching diffusions with stopping times in the costs,
SIAM J. Control Optim., 48 (2009), 1831–1857.
[154] D.D. Sworder and J.E. Boyd, Estimation Problems in Hybrid Systems,
Cambridge University Press, Cambridge, UK, 1999.
[155] D.D. Sworder and V.G. Robinson, Feedback regulators for jump parameters
systems with state and control dependent transition rates, IEEE Trans.
Automat. Control, AC-18 (1973), 355–360.
[156] Y. Takeuchi, N.H. Du, N.T. Hieu, and K. Sato, Evolution of predator-prey
systems decribed by a Lotka-Volterra equation under random environment,
J. Math. Anal. Appl., 323 (2006), 938–957.
[158] J.T. Wloka, B. Rowley, and B. Lawruk, Boundary Value Problems for El-
liptic Systems, Cambridge University Press, Cambridge, UK, 1995.
[160] W.M. Wonham, Liapunov criteria for weak stochastic stability, J. Differ-
ential Eqs., 2 (1966), 195–207.
[161] F. Xi, Feller property and exponential ergodicity of diffusion processes with
state-dependent switching, Sci. China Ser. A, 51 (2008), 329–342.
[163] H. Yang and G. Yin, Ruin probability for a model under Markovian switch-
ing regime, in Probability, Finance and Insurance, World Scientific, T.L.
Lai, H. Yang, and S.P. Yung, Eds., 2004, 206–217.
[164] D.D. Yao, Q. Zhang and X.Y. Zhou, A regime-switching model for Euro-
pean options, in Stochastic Processes, Optimization, and Control Theory
Applications in Financial Engineering, Queueing Networks, and Manufac-
turing Systems, H.M. Yan, G. Yin, and Q. Zhang, Eds., Springer, New
York, NY, 2006, 281–300.
[165] G. Yin, On limit results for a class of singularly perturbed switching diffu-
sions, J. Theoret. Probab., 14 (2001), 673–697.
[167] G. Yin and S. Dey, Weak convergence of hybrid filtering problems involving
nearly completely decomposable hidden Markov chains, SIAM J. Control
Optim., 41 (2003), 1820–1842.
[169] G. Yin and V. Krishnamurthy, Least mean square algorithms with Markov
regime switching limit, IEEE Trans. Automat. Control, 50 (2005), 577–593.
[170] G. Yin, R.H. Liu, and Q. Zhang, Recursive algorithms for stock liquidation:
A stochastic optimization approach, SIAM J. Optim., 13 (2002), 240–263.
[172] G. Yin, X. Mao, C. Yuan, and D. Cao, Approximation methods for hybrid
diffusion systems with state-dependent switching diffusion processes: Nu-
merical alogorithms and existence and uniqueness of solutions, preprint,
2007.
[173] G. Yin, Q.S. Song, and Z. Zhang, Numerical solutions for jump-diffusions
with regime switching, Stochastics, 77 (2005), 61–79.
390 References
[174] G. Yin, H.M. Yan, and X.C. Lou, On a class of stochastic optimization algo-
rithms with applications to manufacturing models, in Model-Oriented Data
Analysis, W.G. Müller, H.P. Wynn and A.A. Zhigljavsky, Eds., Physica-
Verlag, Heidelberg, 213–226, 1993.
[175] G. Yin and H.L. Yang, Two-time-scale jump-diffusion models with Marko-
vian switching regimes, Stochastics Stochastics Rep., 76 (2004), 77–99.
[179] G. Yin and C. Zhu, On the notion of weak stability and related issues of
hybrid diffusion systems, Nonlinear Anal.: Hybrid System, 1 (2007), 173–
187.
[181] J. Yong and X.Y. Zhou, Stochastic Controls: Hamiltonian Systems and
HJB Equations, Springer-Verlag, New York, 1999.
[184] Q. Zhang, Stock trading: An optimal selling rule, SIAM J. Control Optim.,
40 (2001), 64–87.
[185] Q. Zhang and G. Yin, On nearly optimal controls of hybrid LQG problems,
IEEE Trans. Automat. Control, 44 (1999) 2271–2282.
[186] X.Y. Zhou and G. Yin, Markowitz mean-variance portfolio selection with
regime switching: A continuous-time model, SIAM J. Control Optim., 42
(2003), 1466–1482.
[188] C. Zhu and G. Yin, On strong Feller, recurrence, and weak stabilization
of regime-switching diffusions, SIAM J. Control Optim., 48 (2009), 2003–
2031.
References 391
393
394 Index