02.08.2013 Views

Research Statement - Carleton University

Research Statement - Carleton University

Research Statement - Carleton University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

RESEARCH STATEMENT<br />

Son Luu Nguyen<br />

My primary research interests lie in the areas of regime-switching systems and game theory<br />

with a large population.<br />

Regime-switching diffusions, also known as hybrid diffusions, have two components. One of<br />

them is the continuous state component as in the usual diffusion, and the other is a switching<br />

component represented by a pure jump process. In recent years, it has been recognized that the<br />

traditional formulation of differential equation models is often inadequate for treating complex<br />

systems in many real-world applications. Due to the inherent complexity, the underlying systems<br />

often involve characters that cannot be represented by the usual dynamic systems, but rather<br />

display interactive behavior of continuous and discrete dynamics. As a result, effort has been<br />

directed to finding alternative models. Along this line, regime-switching diffusions have drawn<br />

increasing attention. The models have been used in such applications as option pricing [29], jump<br />

linear systems in automatic control [30], hierarchical decision making in production planning [38],<br />

estimation in hybrid systems [41], stock liquidation [43], and competitive Lotka-Volterra model in<br />

random environments [47] among others.<br />

With the aforementioned motivations, my research to date mainly focuses on the study of hybrid<br />

systems. One part of my research is on the pathwise convergence rates for numerical solutions<br />

of regime-switching stochastic differential equations. Our results are already interesting for the<br />

traditional stochastic differential equations since they present a new angle of ascertaining the convergence<br />

rates. Another part of my research is on asymptotic properties of Markov-modulated<br />

random sequences which can be considered as discrete-time counterpart of switching diffusion systems.<br />

We obtain the weak convergence and a strong invariance principle for the scaled processes<br />

of a two-component process under suitable conditions.<br />

Recently, I also have interests in mean field game theory with mixed players. In the context of<br />

noncooperative game theory, large population models have been well studied in economics [22, 10],<br />

social science [40], biological science [2], and engineering [12]. In these areas, of particular interest<br />

is the class of games in which each player interacts with the average effect of many others and<br />

individually has negligible effect on the overall population. Such an interaction pattern may be<br />

modeled by mean field coupling, and it has naturally arisen in economics [22, 10], engineering<br />

[19], and public health research [5]. The class of games we consider involves a major player and<br />

many minor players. The major player has a significant role in affecting others, but each minor<br />

player possesses only a weak influence. This kind of interaction modeling is motivated by many<br />

socio-economic problems.<br />

For large population dynamic games, a central issue is the development of low complexity<br />

solutions so that each player may implement a strategy based on local information. We consider<br />

a mean field linear-quadratic-Gaussian (LQG) game with a major player and a large number of<br />

minor players parametrized by a continuum set and construct the set of decentralized strategies<br />

which has an ε-Nash equilibrium property when applied to the large but finite population model.<br />

In the following sections, these lines of research are described in more detail. In addition, future<br />

directions of research are outlined.<br />

1 Convergence of Numerical Solutions of Stochastic Differential Equations<br />

Because of the nonlinearity, closed-form solutions of many important stochastic differential equations<br />

are virtually impossible to obtain. Thus, numerical solutions are a viable alternative. Due to<br />

its simplicity, Euler-Maruyama (E-M) numerical method has been investigated intensively by many<br />

authors and for different modes of convergence. For example, the pointwise strong and pointwise<br />

1


weak convergence can be found in [28]. The pathwise strong convergence is studied in [14]. In [6],<br />

the pathwise weak convergence is considered for a general class of numerical methods including<br />

E-M method, but the rate is not obtained.<br />

In our works in [36] and [37], we investigate a new approach to provide the rate of pathwise weak<br />

convergence (i.e. the convergence of the distributions of the whole trajectories of the numerical<br />

solutions) in the sense of strong invariance principle (see [8]) for the E-M numerical solutions of<br />

SDEs and regime-switching SDEs. Different from the well-known results in numerical solutions of<br />

stochastic differential equations, in lieu of the usually employed Brownian motion increments in<br />

the algorithm, an easily implementable sequence of independent and identically distributed (i.i.d.)<br />

random variables is used. Being easier to implement compared to the construction of Brownian<br />

increments, such an i.i.d. sequence is preferable in the actual computation. To establish the<br />

convergence of the algorithms, using either the Brownian increments or i.i.d. random variables do<br />

not make much difference. Nevertheless, the analysis becomes much more difficult for our study<br />

of rates of convergence because one has to deal with the difference in the almost sure sense of the<br />

Brownian increments and the i.i.d. sequence.<br />

Next, we will further describe our recent work [36, 37] concerning convergence rate for E-M<br />

numerical solutions of SDEs and regime-switching SDEs.<br />

1.1 Convergence Rate for Numerical Solutions of SDEs<br />

We begin with the SDE<br />

dX(t) = f(X(t))dt + σ(X(t))dW (t), X(0) = x0, (1.1)<br />

where W (·) is a standard one-dimensional Brownian motion, and f(·) and σ(·) are real-valued<br />

functions satisfying suitable conditions. A simple numerical scheme for obtaining approximate<br />

solutions to (1.1) is the E-M method using Brownian increments: for k ≥ 0 and step size ε,<br />

x ε k+1 = xε k + εf(xε k ) + σ(xε k )( W ((k + 1)ε) − W (kε) ) , x ε 0 = x0.<br />

In [36], we consider a similar scheme called weak E-M numerical method: for k ≥ 0 and step size ε,<br />

x ε k+1 = xε k + εf(xε k ) + √ εσ(x ε k )ξk+1, x ε 0 = x0, (1.2)<br />

where {ξn} is a sequence of independent and identically distributed (i.i.d.) random variables with<br />

mean 0 and covariance 1.<br />

One of the main advantages of this weak method is that it allows us to use simple random<br />

variables {ξn} instead of Brownian increments. However, the solution X(t) to (1.1) may be defined<br />

in a probability space different from that of the sequence {xε k , k ≥ 0} defined in (1.2).<br />

Assumptions. Assume that f(·) and σ(·) are real-valued Lipschitz functions, and {ξn} is a<br />

sequence of bounded i.i.d. random variables with zero mean and variance 1.<br />

Main Result. We show that one can define in the same probability space as the Brownian motion<br />

W (t) a sequence of i.i.d random variables { ˜ ξn} which have the same distribution as that of {ξn}<br />

such that for any 0 < λ < 1/4,<br />

sup<br />

0≤t≤T<br />

<br />

<br />

ε<br />

X(t) − ˜x (t) λ<br />

= O(ε ) a.s.<br />

where ˜x ε (t) = ˜x ε n for nε ≤ t < (n + 1)ε, and ˜x ε n+1 = ˜xε n + εf(˜x ε n) + √ εσ(˜x ε n) ˜ ξn+1, ˜x ε 0 = x0.<br />

Our approach is based on the Skorohod embedding theorem (see [15]) and the strong invariance<br />

principles. The rate we obtained appear to be sharp. That is, if we use the Skorohod embedding,<br />

it appears that the rate cannot be further improved; see Kiefer [27].<br />

2


1.2 Convergence Rate for Numerical Solutions of Regime-Switching SDEs<br />

In [37], we extend the result in previous section to the following regime-switching diffusion system<br />

and<br />

dX(t) = f(X(t), α(t))dt + σ(X(t), α(t))dW (t), (1.3)<br />

X(0) = x0, α(0) = α0<br />

P ( α(t + ∆t) = j|α(t) = i, α(s), x(s), s ≤ t ) = qij∆t + o(∆t), i, j ∈ M, i ̸= j (1.4)<br />

where W (·) is a d-dimensional standard Brownian motion defined in (Ω, F, P ), x ∈ R d , M =<br />

{1, 2, ..., m}, f(·, ·) : R d × M ↦→ R d , and σ(·, ·) : R d × M ↦→ R d×d are appropriate functions,<br />

and Q = (qij) ∈ R m0×m0 is the generator of the Markov chain satisfying qij ≥ 0 for i ̸= j and<br />

∑ m0<br />

j=1 qij = 0 for each i ∈ M.<br />

To approximate the d-dimensional standard Brownian motion W (·), we use a sequence of i.i.d.<br />

random vectors {ξn, n ≥ 0} taking values in R d . Therefore, to approximate the solution to (1.3)<br />

we propose the following algorithm<br />

x ε n+1 = x ε n + εf(x ε n, α ε n) + √ εσ(x ε n, α ε n)ξn, ε > 0, n = 0, 1, ..., (1.5)<br />

where {ξn, n ≥ 0} is a sequence of R d −valued i.i.d. random vectors and {α ε n, n ≥ 0} is a Markov<br />

chain independent of {ξn, n ≥ 0} with transition probability matrix P = exp(εQ).<br />

Assumptions. We assume the following conditions hold.<br />

(A1) For some K > 0, the functions f and σ satisfy<br />

|f(x, α) − f(y, α)| ≤ K|x − y|, |σ(x, α) − σ(y, α)| ≤ K|x − y| for x, y ∈ R d , α ∈ M.<br />

(A2) {ξn, n ≥ 0} is a sequence of R d -valued i.i.d. random vectors independent of {α ε n, n ≥ 0} such<br />

that Eξ0 = 0 and Eξ0ξ ′ 0 = I. Furthermore, the moment generating function of ξ0 exists, that<br />

is<br />

E[exp⟨h, ξ0⟩] < ∞ for |h| ≤ t0,<br />

where t0 > 0, h ∈ R r , and ⟨h, ξ0⟩ denotes the inner product of h and ξ0.<br />

Main Result. We show that one can construct in the same probability space as the Brownian<br />

motion W (t) a Markov chain {˜α ε n, n ≥ 1} with transition probability matrix P = exp(εQ) and a<br />

sequence of i.i.d. random vectors { ˜ ξn, n ≥ 1} which are independent of {˜α ε n, n ≥ 1} and have the<br />

same distribution as that of {αε n, n ≥ 1} such that for any 0 < λ < 1/4,<br />

sup<br />

0≤t≤T<br />

<br />

<br />

ε<br />

X(t) − ˜x (t) λ<br />

= O(ε ) a.s.<br />

where ˜x ε (t) = ˜x ε n for nε ≤ t < (n + 1)ε, and ˜x ε n+1 = ˜xε n + εf(˜x ε n, ˜α ε n) + √ εσ(˜x ε n, , ˜α ε n) ˜ ξn+1, ˜x ε 0 = x0.<br />

2 Asymptotic Properties for Markov-Modulated Sequences<br />

My recent research in [34, 35] has been encompassing asymptotic properties of the process {X(k, αk)},<br />

where αk is a Markov chain taking values in a finite set M, and for each α ∈ M, {X(k, α)} is the<br />

primary random sequence. Such processes belong to the class of regime-switching models and may<br />

be viewed as the discrete-time counterpart of the switching diffusion systems. The motivation of<br />

our study stems from a wide variety of applications in actuarial science, communication networks,<br />

3


production planning, manufacturing, and financial engineering. Due to the increasing complexity of<br />

the real-world scenarios, one is often forced to deal with large-scale systems. Due to the uncertainty<br />

of the random environment, there is a growing interest in modeling, analysis, and optimization of<br />

large-scale systems using a regime-switching model in addition to the usual dynamic systems.<br />

For the random sequences under consideration, there are two main issues. Firstly, not much<br />

structure of the sequence {X(k, αk)} is known. Secondly, for a large |M|, there are |M| sequences<br />

of {X(k, α)} with α ∈ M to be dealt with and the complexity becomes an important issue.<br />

Among the large number of states of a Markov chain, a typical situation is that transitions<br />

among some of the states are changing rapidly, whereas others are varying slowly. The state space<br />

can often be split into smaller subspaces such that within each subspace the transitions are about<br />

the same rate, and from one subspace to another, the transitions happen relatively rarely. More<br />

precisely, suppose the state space M admits the representation M = M1 ∪ · · · ∪ Ml0<br />

so that Mi,<br />

i = 1, . . . , l0, can be considered as subspaces, where the Mi’s are not isolated. Such a model is<br />

known to have nearly completely decomposable structure [7, 39] in the literature. A systematic<br />

study of the related Markovian models has been taken recently [45].<br />

Using two-time scales in the formulation, we introduce a small parameter ε > 0 into the transition<br />

probabilities so as to highlight the different rates of transitions. Then, we can write the<br />

sequence as X(k, α ε k ). We assume for each α, {X(k, α)} is mixing, and {αε k<br />

} is a discrete Markov<br />

chain with nearly completely decomposable structure. Due to the lack of structure of the multiple<br />

random sequences, the problem is difficult to solve. Using the idea of aggregation, we lump the<br />

states of the Markov chain in each Mi into one state for i = 1, . . . , l0 to get an aggregate process αε k .<br />

Corresponding to this, we consider a new sequence X(k, αε k ). Effectively, we use a single sequence<br />

{X(k, i)} as a representative for the sequences {X(k, α)} for all α ∈ Mi.<br />

We aim to demonstrate that the new sequence leads to a limit under suitable interpolations.<br />

The limit process is a switching diffusion with a well-defined operator. Denote the original state<br />

space and the state space of the limit by M and M, respectively. If |M| ≫ |M|, a substantial<br />

reduction of computational complexity will be achieved when one treats control and optimization<br />

problems.<br />

In the literature, treating discrete sequences, under suitable scaling, one may obtain a diffusion<br />

limit, which is the classical weak sense functional central limit theorem known as Donsker’s<br />

invariance principle. One may also be interested in obtaining the rates of convergence of the scaled<br />

sequence commonly referred to as strong invariance principle results. In what follows, our effort to<br />

obtain asymptotic properties of the underlying processes under suitable scaling will be described<br />

in more detail.<br />

2.1 Weak Convergence<br />

For ε > 0, let αε k be a time-homogeneous Markov chain on a probability space (Ω, F, P ) with state<br />

space M and transition matrix Pε = P + εQ, where P = (pij) is a transition probability matrix<br />

and Q = (qij) is a generator of a continuous-time Markov chain (i.e., pij ≥ 0 and ∑m j=1 pij = 1;<br />

qij ≥ 0 for i ̸= j and ∑m j=1 qij = 0 for each i). Assume that the state space M can be written as<br />

M = {s11, . . . , s1m1 } ∪ · · · ∪ {sl01, . . . , sl0ml 0 } = M1 ∪ · · · ∪ Ml0 ,<br />

with m0 = m1 + · · · + ml0 and P = diag[P 1 , . . . , P l0 ], where the subspace Mi consists of recurrent<br />

states belonging to the ith ergodic class and the transition probability matrix P i is irreducible and<br />

aperiodic, i = 1, . . . , l0.<br />

) the<br />

Let pε k be the probability vector, pε k = (P (αε k = sij)) ∈ R1×m0 , and νi = (νi1, . . . , νimi<br />

stationary distribution corresponding to the transition matrix P i . Assume that pε 0 = p0 and define<br />

4


an aggregated process α ε k of αε k by<br />

α ε k = i if αε k ∈ Mi, i = 1, . . . , l0, α ε (t) = α ε k<br />

for t ∈ [εk, ε(k + 1)).<br />

It is shown in [45] that the aggregated process α ε (·) converges weakly to α(·), which is a continuoustime<br />

Markov chain generated by Q = (q ij) = diag(ν1, . . . , νl0 )Qdiag(1lm1 , . . . , 1lml 0 ) where 1ll =<br />

(1, . . . , 1) ′ ∈ R l×1 .<br />

For each sij ∈ M, let {X(k, sij)} be a wide-sense stationary sequence of R d -valued random<br />

vectors on (Ω, F, P ). Denote X(n, i) = ∑ mi<br />

j=1 νijX(n, sij) for i = 1, ..., l0.<br />

Assumptions. Assume that the sequence {(X(k, s11), . . . , X(k, sl0ml )) : k ∈ Z} is independent of<br />

0<br />

the Markov process {αε k }, and is ϕ-mixing with mixing measure denoted by ϕ(·). Moreover, assume<br />

that there exist δ > 0 and a constant C that does not depend on k, i such that<br />

∞∑<br />

n=0<br />

ϕ(n) δ<br />

1+δ < ∞, EX(k, sij) = 0, E|X(k, sij)| 2(1+δ) ≤ C, k ≥ 0, sij ∈ M.<br />

Main Results. We obtain the following results:<br />

(i) Denote<br />

z ε (t) = √ ε<br />

⌊t/ε⌋−1<br />

∑<br />

n=0<br />

[ ε<br />

X(n, αn) − X(n, α ε n) ] = √ ⌊t/ε⌋−1 ∑<br />

ε<br />

n=0<br />

l0∑ mi ∑<br />

X(n, sij)[1 {αε − ν n=sij} ij 1 {αε n=i}].<br />

i=1 j=1<br />

Then (z ε (·), α ε (·)) converges weakly to (z(·), α(·)) which is the unique solution to the martingale<br />

problem with operator L determined by the initial layer term in the asymptotic expansion of ( ) k;<br />

Pε<br />

see [45].<br />

(ii) Denote zε (t) = √ ε ∑⌊t/ε⌋−1 n=0 X(n, αεn). Then (zε (·), αε (·)) converges weakly to (z(·), α(·))<br />

which is the unique solution to the martingale problem with operator given by<br />

Lf(x, i) = 1<br />

2<br />

d∑<br />

d∑<br />

j1=1 j2=1<br />

where A(i) = (a j1j2 (i)) = EX(0, i)X ′ (0, i) + ∑ ∞<br />

k=1<br />

Qf(x, ·)(i) = ∑ l0<br />

j=1 q ijf(x, j).<br />

2.2 Strong Invariance Principle<br />

a j1j2 ∂<br />

(i) 2f(x, i)<br />

∂xj1∂x + Qf(x, ·)(i), i = 1, . . . , l0,<br />

j2<br />

[<br />

EX(k, i)X ′ (0, i) + EX(0, i)X ′ ]<br />

(k, i)<br />

Let ε > 0 and αε k be a time-homogeneous Markov chain on (Ω, F, P ) with state space M =<br />

{1, 2, . . . , m} and transition matrix<br />

Pε = P + εQ, (2.1)<br />

where P = (pij) is a transition probability matrix and Q = (qij) is a generator of a continuous-time<br />

Markov chain. Suppose that P is irreducible and aperiodic (i.e. l0 = 1 or M = M1) with the<br />

stationary distribution denoted by ν = (ν1, . . . , νm) ∈ R1×m . Denote pε k = (P (αε k = 1), . . . , P (αε k =<br />

m)) and assume that the initial probability p ε 0 is independent of ε, i.e., pε 0<br />

= p0.<br />

Let {(X(k, 1), X(k, 2), . . . , X(k, m)) : k ∈ Z} be an Rm-valued wide-sense stationary ϕ-mixing<br />

sequence on (Ω, F, P ) which is independent of the Markov process {αε k }. Denote<br />

z ε (t) = √ ⌊t/ε⌋−1 ∑<br />

ε<br />

n=0<br />

m∑<br />

i=1<br />

5<br />

X(k, i)[1 {α ε n =i} − νi].<br />

and


Assumptions. Assume that the sequence {X(k, i) : k ∈ Z, i ∈ M} and its the mixing measure<br />

function ϕ(·) satisfy<br />

EX(k, i) = 0, E|X(k, i)| 4 4<br />

−<br />

< ∞, ϕ(n) < Cn 3 (1+β) for some constants C and β > 0.<br />

Main Result. We show that one can construct on a probability space a one-dimensional Brownian<br />

motion W (t) with EW (t) = 0 and E[W (t)] 2 = σ 2 t where σ can be determined explicitly, and<br />

stochastic processes ˜z ε (t) having the same distributions as those of z ε (t) such that<br />

for some θ > 0.<br />

sup<br />

0≤t≤T<br />

<br />

<br />

ε<br />

˜z (t) − W (t) θ<br />

= o(ε ) a.s.<br />

3 LQG Mixed Games with Continuum-Parametrized Minor Players<br />

We consider a mean field LQG game model with a major player and a large number of minor<br />

players which are parametrized by a continuum set. Large population stochastic dynamic games<br />

with mean field coupling have experienced intense investigation in the past decade; see, e.g., [19, 20,<br />

23, 24, 25, 26]. To obtain low complexity strategies, consistent mean field approximations provide<br />

a powerful approach, and in the resulting solution, each agent only needs to know its own state<br />

information and the aggregate effect of the overall population which may be pre-computed off-line.<br />

One may further establish an ε-Nash equilibrium property for the set of control strategies [20].<br />

The LQG model in [18] shows that the presence of the major player causes an interesting<br />

phenomenon called the lack of sufficient statistics. More specifically, in order to obtain asymptotic<br />

equilibrium strategies, the major player cannot simply use a strategy as a function of its current<br />

state and time; for a minor player, it cannot simply use the current states of the major player and<br />

itself. To overcome this lack of sufficient statistics for decision, the system dynamics are augmented<br />

by adding a new state, which approximates the mean field and is driven by the major player’s state.<br />

This additional state enters the obtained decentralized strategy of each player and it captures the<br />

past history of the major player.<br />

A crucial modeling assumption in [18] is that the minor players are from a finite number of<br />

classes labelled by a set {1, . . . , K}, where players in each class share the same set of parameters<br />

in their dynamics and costs. The size of the additional state introduced in [18] depends on the<br />

number of classes so that it provides sub-mean field approximations for different classes, and this<br />

approach becomes invalid when the dynamic parameters are from an infinite set.<br />

In [32], we consider a population of minor players parametrized by an infinite set such as<br />

a continuum, and seek a different approach for mean field approximations. Due to the linear<br />

quadratic structure of the game with a finite number of players, it is plausible to assume that the<br />

limiting mean field is a Gaussian process and may be represented by using the driving noise of the<br />

major player. Eventually, we will justify this argument by showing the consistency of the mean field<br />

approximation. Given the above representation of the limiting mean field, we may approximate<br />

the original problems of the major player and a typical minor player by stochastic control problems<br />

with random coefficients in the dynamics and costs [3]. This further enables the use of powerful<br />

tools from the theory of backward stochastic differential equations [3]. Also, the Gaussian property<br />

of various processes involved will play an important role and we exploit this to develop kernel<br />

representation to reduce the analysis to function spaces [17].<br />

Mean Field Games with Major and Minor Players. Consider the LQG game with a major<br />

player A0 and a population of minor players {Ai, 1 ≤ i ≤ N}. At time t ≥ 0, the states of the<br />

6


players A0 and Ai are, respectively, denoted by x0(t) and xi(t), 1 ≤ i ≤ N. Let (Ω, F, Ft, t ≥ 0, P )<br />

be the underlying filtration. The dynamics of the N + 1 players are given by a system of linear<br />

stochastic differential equations (SDEs)<br />

dx0(t) = [ A0x0(t) + B0u0(t) + F0x (N) (t) ] dt + D0dW0(t), (3.1)<br />

dxi(t) = [ A(θi)xi(t) + B(θi)ui(t) + F (θi)x (N) (t) ] dt + D(θi)dWi(t), (3.2)<br />

where x (N) = 1/N ∑ N<br />

i=1 xi is the mean field term. The initial states {xj(0), 0 ≤ j ≤ N} are measurable<br />

with respect to F0 and have finite second moments. The states x0, xi and controls u0, ui<br />

are, respectively, n and n1 dimensional vectors. The noise processes W0 and Wi are n2 dimensional<br />

independent standard Brownian motions adapted to (Ft)t≥0, which are also independent of<br />

{xj(0), 0 ≤ j ≤ N}. The vector θi ∈ R d is a parameter in the dynamics of player Ai.<br />

For 0 ≤ j ≤ N, denote u−j = (u0, ..., uj−1, uj+1, ..., uN). For positive semi-definite matrix<br />

M ≥ 0, we may write the quadratic form z T Mz as |z| 2 M . The cost function for A0 is given by<br />

∫ T [<br />

J0(u0, u−0) = E x0(t) − χ0(x<br />

0<br />

(N) (t)) 2 Q0 + uT0 (t)R0u0(t) ] dt, (3.3)<br />

where χ0(x (N) (t)) = H0x (N) (t) + η0, Q0 ≥ 0 and R0 > 0. The cost function for Ai, 1 ≤ i ≤ N, is<br />

∫ T [<br />

Ji(ui, u−i) = E xi(t) − χ(x0(t), x<br />

0<br />

(N) (t)) 2 Q + uT i (t)Rui(t) ] dt, (3.4)<br />

where χ(x0(t), x (N) (t)) = Hx0(t) + ˆ Hx (N) (t) + η, Q ≥ 0 and R > 0. The component Hx0 in<br />

the coupling term χ indicates the strong influence of the major player on each minor player. By<br />

contrast, the mutual impact of two minor players is insignificant. We assume that all matrix<br />

or vector parameters (A0, B0, . . . , η, . . . ) in (3.1)-(3.4) are deterministic and have compatible<br />

dimensions.<br />

Assumptions. We introduce the following assumptions:<br />

(A1) The initial states {xj(0), 0 ≤ j ≤ N}, are independent, and there exists a constant C independent<br />

of N such that sup 0≤j≤N E|xj(0)| 2 ≤ C.<br />

(A2) There exists a distribution function F(θ, x) on Rd+n such that the sequence of empirical<br />

distribution functions FN(θ, x) = 1 ∑N N i=1 1 {θi≤θ,Exi(0)≤x}, N ≥ 1, where each inequality holds<br />

componentwise, converges to F(θ, x) weakly, i.e., for any bounded and continuous function<br />

h(θ, x) on Rd+n ,<br />

∫<br />

lim<br />

N→∞ Rd+n ∫<br />

h(θ, x)dFN(θ, x) =<br />

Rd+n h(θ, x)dF(θ, x).<br />

(A3) A(·), B(·), F (·) and D(·) are continuous matrix-valued functions of θ ∈ Θ, where Θ is a<br />

compact subset of R d .<br />

Main Results. We approximate the mean field generated by the minor players by a kernel<br />

representation using the Brownian motion of the major player.<br />

Solve local optimal control problems for both the major player and a representative minor player<br />

via backward stochastic differential equations.<br />

Construct a set of decentralized control strategies based on consistent mean field approximations<br />

and show that these strategies have an ε-Nash equilibrium property.<br />

7


4 Future <strong>Research</strong> Plan<br />

My goal is to be a research mathematician and an educator, and to make life-time commitment<br />

to mathematical research and training the next generation of mathematicians, scientists, and engineers.<br />

My ultimate plan is to make substantial contributions to the research of stochastic systems<br />

and game theory. In the next few years, my research plan is outline below.<br />

4.1 Further Topics in Stochastic Hybrid Systems<br />

The study of stochastic hybrid systems is still in an early stage. A number of problems remain<br />

open. Obtaining convergence rates for other numerical methods like Milstein scheme (see [28, 31])<br />

for regime-switching diffusion processes will have many significant applications, especially in the<br />

more challenging case that the switching component is diffusion-dependent. Next, concerning the<br />

degenerated switching diffusion, can we obtain a strong invariance principle? It appears that the<br />

desired result will be more difficult to obtain compared to the case of traditional diffusion process<br />

(see [16]) since one needs to deal with the Markov dependent component when estimating the<br />

moments. Also, it will be useful and natural to consider the regime-switching model for the system<br />

of many interacting particles in a random environment. Although some stability results in this<br />

direction when the number of particles is finite was obtained in [42, 44], there are still a lot of<br />

open questions on the asymptotic behavior of such systems when the number of particles becomes<br />

large. Attempts in finding answers to these questions present new challenges, which will lead to<br />

many substantial discoveries. Hence, my immediate future work will be devoted to the above<br />

problems. In addition, I will also focus on applications of the theoretical results of hybrid systems<br />

to real world problems including risk theory, financial engineering, stochastic approximation and<br />

stochastic optimization and control for stochastic hybrid systems.<br />

4.2 Mean-Field Game Theory and Stochastic McKean-Vlasov Equations<br />

In the past decade, the mean field games theory has been investigated intensively with many<br />

applications in wireless communication systems, biological science, and economics. However, this<br />

still is a relative new area with many open questions. In the future, I would like to focus on<br />

this topic. In particular, I am interested in the general problems in the mean field games theory<br />

with a large population of minor player and a major player which generalize the LQG cases (see<br />

[18, 19, 20]), numerical solutions for mean-field game and control problems (see [21]), and stochastic<br />

mean-field systems (see [9, 13]). In what follows, two specific problems will be described in more<br />

detail.<br />

Mean-Field Game Problem. In order to obtain low complexity strategies so that each player<br />

only needs to know its own state information, the consistent mean field approximation (see [18,<br />

19, 20]) is a powerful approach. For the general mean field game problems with a major player<br />

and a large population of minor players, this approach leads to the importance of investigating the<br />

limiting behavior of the empirical measure obtained by stochastic interacting particles described<br />

by the following equations<br />

(<br />

dx0(t) = f0 t, x0(t), ε N ) (<br />

x(t) dt + σ0 t, x0(t), ε N )<br />

x(t) dW0(t), (4.1)<br />

dxi(t) = f ( t, x0(t), xi(t), ε N )<br />

N∑<br />

x(t) dt + σ ( t, x0(t), xi(t), ε N x(t) , j) dWj(t), 1 ≤ i ≤ N (4.2)<br />

j=1<br />

where {Wi, 0 ≤ i ≤ N} are n dimensional independent standard Brownian motions, f0, σ0 : R ×<br />

R n × P → R n , f : R × R 2n × P → R n , σ : R × R 2n × P × {1, 2, ..., N} → R n , P is the set of<br />

8


all probability measures on Rn , and εN x(t) = (1/N) ∑N i=1 δxi(t). This is an interesting extension of<br />

traditional stochastic interacting particle system considered in [9] with a major particle.<br />

In [9] (see also [13]), several significant results for the traditional stochastic interacting particle<br />

system are presented. In particular, the limiting process of the sequence {εN x(t) } is characterized<br />

as a weak solution to a stochastic McKean-Vlasov equation. For the above system, we also hope<br />

to characterize the limit of the coupled sequence {(x0(t), ε N x(t)<br />

)} as a weak solution to a coupled<br />

stochastic McKean-Vlasov equation by using martingale problem approach. Moreover, conditions<br />

for the existence and uniqueness for the weak solution, the strong solution, and the regular property<br />

of the paths of the solution to the desired equation may be determined by considering the martingale<br />

problems in a suitable rigged Hilbert space. This study is expected to bring new insights to<br />

stochastic mean field control and game problems with mixed players in our far future plan.<br />

Numerical Solution for Mean Field Stochastic System. Let consider the celebrated mean<br />

field stochastic differential equation (MFSDE) in which the state process involves as follow<br />

(<br />

dX(t) = f t, X(t), Eφ ( X(t) )) (<br />

dt + σ t, X(t), Eφ ( X(t) ))<br />

dW (t) (4.3)<br />

where W (t) is a multi-dimensional Brownian motion defined on a probability space (Ω, F, P ).<br />

The studies of MFSDE are of great theoretical values due to its interesting state law dependence<br />

structure. On the other hand, the MFSDE also possesses broad-range real applications. Such largepopulation<br />

interacting system has already been extensively investigated in many different fields,<br />

such as engineering and socioeconomic science, game theory, finance, etc. Since the MFSDEs are<br />

well-defined and widely-used systems, it is of importance to study the numerical approximations<br />

of X(t) and its probability law µt. In [1] and [4], stochastic particle methods were proposed to<br />

construct approximations for the distribution function Vt of µt. These methods, however, require to<br />

generate too many stochastic processes if one expects to get an accurate approximation. Therefore,<br />

we intend to investigate another method to find a numerical approximation of Vt and X(t) with<br />

reduced computational effort. It is expected that by combining both the MFSDE of X(t) and the<br />

Fokker-Planck-Kolmogorov equation of Vt, we will be able to obtain the desired method.<br />

I am keen on continuing these and other inter-disciplinary projects in the directions described<br />

above. One of my primary goal at this stage is to keep broadening my horizons to as many areas<br />

of mathematics as possible.<br />

References<br />

[1] F. Antonelli, and A. Kohatsu-Higa, Rate of convergence of a particle method to the solution of the<br />

McKean-Vlasov equation. Ann. Appl. Probab., 12 (2002), 423–476.<br />

[2] C.T. Bauch and D.J.D. Earn, Vaccination and the theory of games, Proc. Natl. Acad. Sci. U.S.A., 101<br />

(2004), 13391–13394.<br />

[3] J. M. Bismut. Linear quadratic optimal stochastic control with random coefficients. SIAM J. Control<br />

Optim., 14 (1976), 419–444.<br />

[4] M.Bossy, and D. Talay, A stochastic particle method for the McKean-Vlasov and the Burgers equation.<br />

Math. Comp., 66 (1997), 157–192.<br />

[5] R. Breban, R. Vardavas, and S. Blower, Mean-field analysis of an inductive reasoning game: Application<br />

to influenza vaccination, Phys. Rev. E, 76 (2007), 031127.<br />

[6] B. Charbonneau, Y. Svyrydov and P.F. Tupper, Weak convergence in the Prokhorov metric of methods<br />

for stochastic differentia equations, IMA J. Numer. Anal., 30 (2010), 579–594.<br />

[7] P.J. Courtois, Decomposability: Queueing and Computer System Applications, Academic Press, New<br />

York, NY, 1977.<br />

[8] M. Csörgö and P. Revesz, Strong Approximations in Probability and Statistics, Academic Press, New<br />

York-London, 1981.<br />

9


[9] D. Dawson, and J. Vaillancourt, Stochastic McKean-Vlasov equations. NoDEA Nonlinear Differential<br />

Equations Appl. 2 (1995), 199–229.<br />

[10] G. M. Erickson, Differential game models of advertising competition, European J. Oper. Res., 83 (1995),<br />

431–438.<br />

[11] S.N. Ethier and T.G. Kurtz, Markov Processes: Characterization and Convergence, J. Wiley, New York,<br />

NY, 1986.<br />

[12] D. Fudenberg and D.K. Levine, The Theory of Learning in Games, MIT Press, Cambridge, MA, 1998.<br />

[13] J. Gärtner, On the McKean-Vlasov limit for interacting diffusions, Math. Nachr., 137 (1988), 197–248.<br />

[14] I. Gyöngy, A note on Euler’s approximation, Potential Anal., 8 (1998), 205–216.<br />

[15] P. Hall and C.C. Heyde, Martingale Limit Theory and Its Application, Academic Press, New York,<br />

1980.<br />

[16] J.A. Heunis, Strong invariance principle for singular diffusions, Stochastic Process. Appl., 104 (2003),<br />

57–80.<br />

[17] T. Hida and M. Hitsuda. Gaussian Processes, AMS, Providence, RI, 1993.<br />

[18] M. Huang. Large-population LQG games involving a major player: the Nash certainty equivalence<br />

principle. SIAM J. Control Optim., 48 (2010), 3318–3353.<br />

[19] M. Huang, P.E. Caines, and R. P. Malhame, Individual and mass behaviour in large population stochastic<br />

wireless power control problems: Centralized and Nash equilibrium solutions, in Proc. 42nd IEEE<br />

Conf. Dec. Contr., Maui, HI, 2003, 98–103.<br />

[20] M. Huang, P. E. Caines and R. P. Malhamé. Large-population cost-coupled LQG problems with nonuniform<br />

agents: Individual-mass behavior and decentralized ϵ-Nash equilibria. IEEE Trans. Autom. Control,<br />

52 (2007), 1560–1571.<br />

[21] H. Kushner and P. Dupuis, Numerical Methods for Stochastic Control Problems in Continuous Time,<br />

Springer-Verlag, 2001.<br />

[22] V.E. Lambson, Self-enforcing collusion in large dynamic markets, J. Econ. Theory, 34 (1984), 282–291.<br />

[23] J.-M. Lasry and P.-L. Lions. Jeux à champ moyen. I. Le cas stationnaire, C. R. Math. Acad. Sci. Paris,<br />

343 (2006), 619–625.<br />

[24] J.-M. Lasry and P.-L. Lions. Jeux à champ moyen. II. Horizon fini et contrle optimal, C. R. Math.<br />

Acad. Sci. Paris, 343 (2006), 679–684.<br />

[25] J.-M. Lasry and P.-L. Lions. Mean field games, Japan. J. Math., 2 (2007), 229–260.<br />

[26] T. Li and J.-F. Zhang. Asymptotically optimal decentralized control for large population stochastic<br />

multiagent systems, IEEE Trans. Automat. Control, 53 (2008), 1643–1660.<br />

[27] J. Kiefer, On the deviations in the Skorokhod-Strassen approximation scheme, Z. Wahrsch. Verw.<br />

Gebiete, 13 (1969), 321–332.<br />

[28] P.E. Kloeden and E. Platen, Numerical Solution of Stochastic Differential Equations, Springer-Verlag,<br />

1992.<br />

[29] X. Mao, A. Truman, and C. Yuan, Euler-Maruyama approximations in mean-reverting stochastic<br />

volatility model under regime-switching, J. Appl. Math. Stoch. Anal., (2006), Article ID 80967.<br />

[30] M. Mariton, Jump Linear Systems in Automatic Control, Marcel Dekker, New York, 1990.<br />

[31] G.N. Milstein and M.V. Tretyakov, Stochastic Numerics for Mathematical Physics, Springer-Verlag,<br />

Berlin, 2004.<br />

[32] S.L. Nguyen and M. Huang, LQG Mixed games with continuum-parametrized minor players, (2011),<br />

submitted.<br />

[33] S.L. Nguyen and G. Yin, Asymptotic properties of hybrid random processes modulated by Markov<br />

chains, to appear in Nonlinear Analysis: Theory, Methods and Applications, 71 (2009), e1638–e1648.<br />

[34] S.L. Nguyen and G. Yin, Asymptotic properties of Markov modulated random sequences with fast and<br />

slow time scales, Stochastics, 82 (2010), 445–474.<br />

[35] S.L. Nguyen and G. Yin, Weak convergence of Markov modulated random sequences, Stochastics, 82<br />

(2010), 521–552.<br />

[36] S.L. Nguyen and G. Yin, Pathwise convergence rate for numerical solutions of stochastic differential<br />

equations, IMA Journal of Numerical Analysis, (2011), in press, doi: 10.1093/imanum/drr025.<br />

[37] S.L. Nguyen and G. Yin, Pathwise convergence rate for numerical solutions of Markovian switching<br />

stochastic differential equations, Nonlinear Analysis: Real World Applications, (2011), to appear.<br />

[38] S.P. Sethi and Q. Zhang, Hierarichical Decision Making in Stochastic Manufacturing Systems,<br />

Birkhäuser, Boston, 1994.<br />

[39] H.A. Simon and A. Ando, Aggregation of variables in dynamic systems, Econometrica, 29 (1961),<br />

111–138.<br />

10


[40] J.M. Smith, Evolution and the Theory of Games, Cambridge <strong>University</strong> Press, Cambridge, UK, 1982.<br />

[41] D.D. Sworder and J.E. Boyd, Estimation Problems in Hybrid Systems, Cambridge <strong>University</strong> Press,<br />

Cambridge, UK, 1999.<br />

[42] F. Xi, and G. Yin, Asymptotic properties of a mean-field model with a continuous-state-dependent<br />

switching process, J. Appl. Probab., 46 (2009), 221–243.<br />

[43] G. Yin, R.H. Liu, and Q. Zhang, Recursive algorithms for stock liquidation: a stochastic optimization<br />

approach, SIAM J. Optim., 13 (2002), 240–263.<br />

[44] G. Yin, G. Zhao, and F. Xi, Mean-Field models involving continuous-state-dependent random switching:<br />

Nonnegativity constraints, moment bounds, and two-time-scale limits, Taiwanese J. Math., 15 (2011),<br />

1783–1805.<br />

[45] G. Yin and Q. Zhang, Discrete-time Markov Chains: Two-time-scale Methods and Applications,<br />

Springer, New York. 2005.<br />

[46] H. Yin, P. G. Mehta, S. P. Meyn and U. V. Shanbhag. Synchronization of coupled oscillator is a game,<br />

Proc. American Control Conference, Baltimore, MD, 1783–1790, June, 2010.<br />

[47] C. Zhu and G. Yin, On competitive Lotka-Volterra model in random environments, J. Math. Anal.<br />

Appl., 357 (2009), 154–170.<br />

School of of Mathematics & Statistics, <strong>Carleton</strong> <strong>University</strong>, 1125 Colonel By<br />

Drive, Ottawa, ON, Canada, K1S 5B6<br />

Email: snguyen@math.carleton.ca<br />

11

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!