This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

[1,3]\fnmStephen S.-T. \surYau

1]\orgdivDepartment of Mathematical Sciences, \orgnameTsinghua University, \orgaddress\cityBeijing, \postcode100084, \countryChina

2]\orgdivSchool of Mathematics, \orgnameRenmin University of China, \orgaddress\cityBeijing, \postcode100872, \countryChina

3]\orgnameBeijing Institute of Mathematical Sciences and Applications, \orgaddress\cityBeijing, \postcode101408, \countryChina

On the Convergence Analysis of Yau-Yau Nonlinear Filtering Algorithm: from a Probabilistic Perspective

\fnmZeju \surSun [email protected]    \fnmXiuqiong \surChen [email protected]    [email protected] [ [ [
Abstract

At the beginning of this century, a real time solution of the nonlinear filtering problem without memory was proposed in [1, 2] by the third author and his collaborator, and it is later on referred to as Yau-Yau algorithm. During the last two decades, a great many nonlinear filtering algorithms have been put forward and studied based on this framework. In this paper, we will generalize the results in the original works and conduct a novel convergence analysis of Yau-Yau algorithm from a probabilistic perspective. Instead of considering a particular trajectory, we estimate the expectation of the approximation error, and show that commonly-used statistics of the conditional distribution (such as conditional mean and covariance matrix) can be accurately approximated with arbitrary precision by Yau-Yau algorithm, for general nonlinear filtering systems with very liberal assumptions. This novel probabilistic version of convergence analysis is more compatible with the development of modern stochastic control theory, and will provide a more valuable theoretical guidance for practical implementations of Yau-Yau algorithm.

keywords:
nonlinear filtering, DMZ equation, Yau-Yau algorithm, convergence analysis, stochastic partial differential equation
pacs:
[

MSC Classification]60G35, 93E11, 60H15, 65M12

1 Introduction

Filtering is an important subject in the field of modern control theory, and has wide applications in various scenarios such as signal processing [3][4], weather forecast [5][6], aerospace industrial [7][8] and so on. The core objective of filtering problem is pursuing accurate estimation or prediction to the state of a given stochastic dynamical system based on a series of noisy observations [9][10]. For practical implementations, it is also necessary that the estimation or prediction to the state can be computed in a recursive and real-time manner.

In the filtering problems we consider in this paper, the evolution of state processes, as well as the noisy observations, is governed by the following system of stochastic differential equations,

{dXt=f(Xt)dt+g(Xt)dVt,X0=ξ,dYt=h(Xt)dt+dWt,Y0=0,t[0,T],\left\{\begin{aligned} &dX_{t}=f(X_{t})dt+g(X_{t})dV_{t},\quad X_{0}=\xi,\\ &dY_{t}=h(X_{t})dt+dW_{t},\quad Y_{0}=0,\end{aligned}\right.\quad t\in[0,T], (1)

in the filtered probability space (Ω,,{t}t=0T,P)(\Omega,\mathcal{F},\{\mathcal{F}_{t}\}_{t=0}^{T},P), where T>0T>0 is a fixed termination; X={Xt:0tT}dX=\{X_{t}:0\leq t\leq T\}\subset\mathbb{R}^{d} is the state process we would like to track; Y={Yt:0tT}dY=\{Y_{t}:0\leq t\leq T\}\subset\mathbb{R}^{d} is the noisy observation to the state process XX; {Vt:0tT}\{V_{t}:0\leq t\leq T\} and {Wt:0tT}\{W_{t}:0\leq t\leq T\} are mutually independent, t\mathcal{F}_{t}-adapted, dd-dimensional standard Brownian motions; ξ\xi is a d\mathbb{R}^{d}-valued random variable with probability density function σ0(x)\sigma_{0}(x), which is independent of VtV_{t} and WtW_{t}; f:ddf:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d}, g:dd×dg:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d\times d} and h:ddh:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} are sufficiently smooth vector- or matrix- valued functions.

Mathematically, for a given test function φ:d\varphi:\mathbb{R}^{d}\rightarrow\mathbb{R}, the optimal estimation of φ(Xt)\varphi(X_{t}) based on the historical observations up to time tt is the conditional expectation E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}], where 𝒴t:=σ{Ys:0st}\mathcal{Y}_{t}:=\sigma\{Y_{s}:0\leq s\leq t\} the σ\sigma-algebra generated by historical observations. Such estimation is ‘optimal’ in the sense that,

E[φ(Xt)|𝒴t]=arg minU is 𝒴t-measurableE[(φ(Xt)U)2].E[\varphi(X_{t})|\mathcal{Y}_{t}]=\mathop{\text{arg min}}\limits_{U\text{ is }\mathcal{Y}_{t}\text{-measurable}}E[(\varphi(X_{t})-U)^{2}]. (2)

Therefore, the main task of filtering problem can be specified into finding efficient algorithms to numerically compute the conditional expectation E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}], or equivalently, the conditional probability distribution P[Xt|𝒴t]P[X_{t}\in\cdot|\mathcal{Y}_{t}] or the conditional probability density function (if exists).

With some regularity assumptions on the coefficients f,g,hf,g,h in the system (1), the conditional probability measure P[Xt|𝒴t]P[X_{t}\in\cdot|\mathcal{Y}_{t}] is absolutely continuous with respect to the Lebesgue measure in d\mathbb{R}^{d}, and the conditional probability density function σ(t,x)\sigma(t,x), (without considering a normalization constant), can be described by the well-known DMZ equation, which is named after three researchers: T. E. Duncan [11], R. E. Mortensen [12] and M. Zakai [13], who derived the equation independently in the late 1960s.

The DMZ equation satisfied by the unnormalized conditional probability density function σ(t,x)\sigma(t,x) is a second-order stochastic partial differential equation, and does not possess an explicit-form solution in general. In their works [1] and [2], the third author and his collaborator propose a two-stage algorithm framework to compute the solution of the DMZ equation numerically in a memoryless and real-time manner. Later on, this algorithm is referred to as Yau-Yau algorithm.

The basic idea of Yau-Yau algorithm is that the heavy computation burden of numerically solving a Kolmogorov-type partial differential equation (PDE) can be done off-line, and in the meanwhile, the on-line procedure only consists of the basic computations such as multiplication by the exponential function of the observations. In the framework of Yau-Yau algorithm, various kinds of methods to solving the Kolmogorov-type PDEs, such as spectral methods [14][15], proper orthogonal decomposition [16], tensor training [17], etc, are proposed and applied in specific examples of nonlinear filtering problems. Numerical results in the previous works mentioned above show that Yau-Yau algorithm can provide accurate and real-time estimations to the state process of very general nonlinear filtering problems in low and medium-high dimensional space.

In the original works [1] and [2], the convergence analysis of Yau-Yau algorithm is conducted pathwisely. For regular paths of observations with some boundedness conditions, it is proved that the numerical solution provided by Yau-Yau algorithm will converge to the exact solution of DMZ equation both pointwisely and in L2L^{2}-sense, as the size of time-discretization step tends to zero, while in the works on the practical implementations of Yau-Yau algorithm, such as [14] and [15], the convergence analysis mainly focuses on the capability of numerical methods to approximate the solution of Kolmogorov-type PDEs arising in Yau-Yau algorithms.

In this paper, we will revisit the convergence analysis of Yau-Yau algorithm from a probabilistic perspective. Instead of considering the convergence results pathwisely, we prove that the solution of Yau-Yau algorithm will converge to the exact solution of DMZ equation in expectation, and also, after the normalization procedure, the approximated solution to the filtering problem provided by Yau-Yau algorithm will converge to the conditional expectation E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}].

The advantage of this probabilistic perspective is that for a theoretically rigorous convergence analysis, instead of regularity assumptions on observation paths, we only need to make assumptions on the coefficients f,g,hf,g,h of the filtering system (1) and the test function φ\varphi, which are verifiable off-line in advance for practitioners. In the meanwhile, as shown in the main results of this paper in Section 3, those assumptions we need are in fact quite general, and it is straightforward to check that the most commonly used test functions φ(x)=xi\varphi(x)=x_{i} and φ(x)=xixj\varphi(x)=x_{i}x_{j}, x=(x1,,xd)dx=(x_{1},\cdots,x_{d})^{\top}\in\mathbb{R}^{d}, corresponding to the conditional mean and conditional covariance matrix, as well as the linear Gaussian systems (with f(x)=Fxf(x)=Fx, g(x)Γg(x)\equiv\Gamma, and h(x)=Hxh(x)=Hx, F,H,Γd×dF,H,\Gamma\in\mathbb{R}^{d\times d}), satisfy all the assumptions.

Moreover, to the best of the authors’ knowledge, most of the theoretical analysis of PDE-based filtering algorithm mainly deals with convergence results with respect to the DMZ equation. In such probabilistic perspective we consider here in this paper, however, it is natural and convenient to make a step forward and discuss the approximation capability of Yau-Yau algorithm to the normalized conditional expectation and conditional probability distribution. In this way, we will provide a thorough convergence analysis of the Yau-Yau algorithm for filtering problems.

The organization of this paper is as follows. Section 2 serves as preliminaries, in which we will summarize some basic concepts of filtering problems and the main procedure of Yau-Yau algorithm. The main theorems in this paper will be stated in Section 3, together with a sketch of the proofs. In the next four sections, we will provide the detailed proofs of the lemmas and theorems. We first focus on the properties of the exact solution of DMZ equation in Section 4 and Section 5, and then deal with the approximated solutions given by Yau-Yau algorithm in Section 6 and Section 7. Finally, Section 8 is a conclusion.

2 Preliminaries

In this section, we would like to briefly summarize the theory of nonlinear filtering, including the change-of-measure approach to deriving the DMZ equation, as well as the main idea and procedures of Yau-Yau algorithm.

In the change-of-measure approach to deriving the DMZ equation corresponding to the filtering system (1), we first introduce a series of reference probability measures {P~t:0tT}\{\tilde{P}_{t}:0\leq t\leq T\}, absolutely continuous to the original probability measure PP with Radon derivatives given by

ZtdP~tdP|t=exp(0th(Xs)𝑑Ws120t|h(Xs)|2𝑑s),t[0,T].Z_{t}\triangleq\frac{d\tilde{P}_{t}}{dP}\biggr{|}_{\mathcal{F}_{t}}=\exp\biggl{(}-\int_{0}^{t}h(X_{s})^{\top}dW_{s}-\frac{1}{2}\int_{0}^{t}|h(X_{s})|^{2}ds\biggr{)},\ t\in[0,T]. (3)

According to Girsanov’s theorem, as long as the process {Zt:0tT}\{Z_{t}:0\leq t\leq T\} defined in (3) is a martingale, then under the reference probability measure P~T\tilde{P}_{T}, the observation process {Yt:0tT}\{Y_{t}:0\leq t\leq T\} is a standard Brownian motion which is independent of the state process XX.

We also introduce the process {Z~t:0tT}\{\tilde{Z}_{t}:0\leq t\leq T\}, Z~t=Zt1\tilde{Z}_{t}=Z_{t}^{-1}, to be the inverse of ZtZ_{t}, which is also a Radon derivative and can be expressed by the stochastic integral with respect to YY as follows:

Z~t=Zt1=dPdP~t|t=exp(0th(Xs)𝑑Ys120t|h(Xs)|2𝑑s),t[0,T].\tilde{Z}_{t}=Z_{t}^{-1}=\frac{dP}{d\tilde{P}_{t}}\biggr{|}_{\mathcal{F}_{t}}=\exp\biggl{(}\int_{0}^{t}h(X_{s})^{\top}dY_{s}-\frac{1}{2}\int_{0}^{t}|h(X_{s})|^{2}ds\biggr{)},\ t\in[0,T]. (4)

Therefore, for any t\mathcal{F}_{t}-measurable, integrable random variable UtU\in\mathcal{F}_{t}, its expectation with respect to measure PP can be computed by

E[U]=E~[Z~tU],E[U]=\tilde{E}\left[\tilde{Z}_{t}U\right], (5)

where E~\tilde{E} means the expectation is taken under the probability measure P~T\tilde{P}_{T}.

As an extension of Bayesian formula in the context of continuous-time stochastic processes, the following Kallianpur-Striebel formula allows us to express and calculate the solution of filtering problem, E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}], by a ratio of conditional expectations under P~T\tilde{P}_{T}:

E[φ(Xt)|𝒴t]=E~[Z~tφ(Xt)|𝒴t]E~[Z~t|𝒴t],t[0,T].E[\varphi(X_{t})|\mathcal{Y}_{t}]=\frac{\tilde{E}\left[\tilde{Z}_{t}\varphi(X_{t})|\mathcal{Y}_{t}\right]}{\tilde{E}\left[\tilde{Z}_{t}|\mathcal{Y}_{t}\right]},\ t\in[0,T]. (6)

Since the denominator E~[Z~t|𝒴t]\tilde{E}\left[\tilde{Z}_{t}|\mathcal{Y}_{t}\right] in (6) is independent of the test function φ\varphi, people often refer to the nominator, E~[Z~tφ(Xt)|𝒴t]\tilde{E}\left[\tilde{Z}_{t}\varphi(X_{t})|\mathcal{Y}_{t}\right], as the unnormalized conditional expectation of φ(Xt)\varphi(X_{t}). The corresponding measure-valued stochastic process {ρt:0tT}\{\rho_{t}:0\leq t\leq T\} defined by

ρt(A):=E~[Z~t1A|𝒴t],At,t[0,T],\rho_{t}(A):=\tilde{E}\left[\tilde{Z}_{t}1_{A}|\mathcal{Y}_{t}\right],\ \forall\ A\in\mathcal{F}_{t},\ t\in[0,T], (7)

is also referred to as unnormalized conditional probability measure, and we also denote the unnormalized conditional expectation by

ρt(φ):=E~[Z~tφ(Xt)|𝒴t],φ is a test function,t[0,T],\rho_{t}(\varphi):=\tilde{E}\left[\tilde{Z}_{t}\varphi(X_{t})|\mathcal{Y}_{t}\right],\ \varphi\text{ is a test function},\ t\in[0,T], (8)

With sufficient regularity assumptions on the coefficients f,g,hf,g,h and test function φ\varphi, the evolution of ρt(φ)\rho_{t}(\varphi) is governed by the following well-known DMZ equation:

ρt(φ)=ρ0(φ)+0tρs(φ)𝑑s+j=1d0tρs(hjφ)𝑑Ysj,t[0,T].\rho_{t}(\varphi)=\rho_{0}(\varphi)+\int_{0}^{t}\rho_{s}(\mathcal{L}\varphi)ds+\sum_{j=1}^{d}\int_{0}^{t}\rho_{s}(h_{j}\varphi)dY_{s}^{j},\ t\in[0,T]. (9)

where

=12i,j=1daij(x)2xixj+i=1dfi(x)xi\mathcal{L}=\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}+\sum_{i=1}^{d}f_{i}(x)\frac{\partial}{\partial x_{i}} (10)

is a second-order elliptic operator with a(x):=(aij(x))1i,jd=g(x)g(x)a(x):=(a^{ij}(x))_{1\leq i,j\leq d}=g(x)g(x)^{\top}.

If the stochastic measures ρt\rho_{t}, t[0,T]t\in[0,T], are almost surely absolutely continuous to the Lebesgue measure in d\mathbb{R}^{d}, and the density functions (or the Radon derivatives) σ(t,x)\sigma(t,x), as well as the derivatives of σ(t,x)\sigma(t,x), is square-integrable, then σ(t,x)\sigma(t,x) is the solution to the following equation, (which is also referred to as the DMZ equation):

{dσ(t,x)=σ(t,x)dt+j=1dhj(x)σ(t,x)dYtj,t[0,T],σ(0,x)=σ0(x),\left\{\begin{aligned} &d\sigma(t,x)=\mathcal{L}^{*}\sigma(t,x)dt+\sum_{j=1}^{d}h_{j}(x)\sigma(t,x)dY_{t}^{j},\ t\in[0,T],\\ &\sigma(0,x)=\sigma_{0}(x),\end{aligned}\right. (11)

which is a second-order stochastic partial differential equation with

()=12i,j=1d2xixj(aij)i=1dxi(fi).\mathcal{L}^{*}(\star)=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}\star)-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}\star). (12)

the adjoint operator of \mathcal{L}.

In this case, the unnormalized conditional expectation ρt(φ)\rho_{t}(\varphi) can be expressed as

ρt(φ)=E~[Z~tφ(Xt)|𝒴t]=dφ(x)σ(t,x)𝑑x,φ is a test function,t[0,T],\rho_{t}(\varphi)=\tilde{E}\left[\tilde{Z}_{t}\varphi(X_{t})|\mathcal{Y}_{t}\right]=\int_{\mathbb{R}^{d}}\varphi(x)\sigma(t,x)dx,\ \varphi\text{ is a test function},\ t\in[0,T], (13)

and the (normalized) conditional expectation E[φ(Xt)|𝒴t]E\left[\varphi(X_{t})|\mathcal{Y}_{t}\right] can be calculated by

E[φ(Xt)|𝒴t]=E~[Z~tφ(Xt)|𝒴t]E~[Z~t|𝒴t]=dφ(x)σ(t,x)𝑑xdσ(t,x)𝑑x.E\left[\varphi(X_{t})|\mathcal{Y}_{t}\right]=\frac{\tilde{E}\left[\tilde{Z}_{t}\varphi(X_{t})|\mathcal{Y}_{t}\right]}{\tilde{E}\left[\tilde{Z}_{t}|\mathcal{Y}_{t}\right]}=\frac{\int_{\mathbb{R}^{d}}\varphi(x)\sigma(t,x)dx}{\int_{\mathbb{R}^{d}}\sigma(t,x)dx}. (14)

Because the solution of (11) does not have a closed form for general nonlinear filtering systems, efficient numerical methods must be proposed, so that we can get a good approximation to the conditional expectation E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}] through the equation (14).

At the beginning of this century, the third author and his collaborator proposed a two-stage algorithm to numerically solve the DMZ equation (11) in a memoryless and real-time manner, which is often referred to as Yau-Yau algorithm. Here, we would like to briefly introduce the basic idea and main procedure of this algorithm.

Firstly, if we consider the exponential transformation

w(t,x):=exp(h(x)Yt)σ(t,x),t[0,T],w(t,x):=\exp\left(-h^{\top}(x)Y_{t}\right)\sigma(t,x),\ t\in[0,T], (15)

then the function w(t,x)w(t,x) satisfies the robust DMZ equation

wt=12i,j=1daij(x)2wxixj+i=1dFi(t,x)wxi+J(t,x)w(t,x),\frac{\partial w}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial^{2}w}{\partial x_{i}\partial x_{j}}+\sum_{i=1}^{d}F_{i}(t,x)\frac{\partial w}{\partial x_{i}}+J(t,x)w(t,x), (16)

where the stochastic differential terms in the original DMZ equation (11) are eliminated and

Fi(t,x)=j=1d(aijxj+aijk=1dYtkhkxj)fi(x),i=1,,d,F_{i}(t,x)=\sum_{j=1}^{d}\biggl{(}\frac{\partial a^{ij}}{\partial x_{j}}+a^{ij}\sum_{k=1}^{d}Y_{t}^{k}\frac{\partial h_{k}}{\partial x_{j}}\biggr{)}-f_{i}(x),\quad i=1,\cdots,d,
J(t,x)=\displaystyle J(t,x)= 12i,j=1d2aijxixj+i,j,k=1dYtkhkxjaijxi\displaystyle\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}a^{ij}}{\partial x_{i}\partial x_{j}}+\sum_{i,j,k=1}^{d}Y_{t}^{k}\frac{\partial h_{k}}{\partial x_{j}}\frac{\partial a^{ij}}{\partial x_{i}} (17)
+12i,j=1daij(k=1dYtk2hkxixj+k=1dl=1dYtkYtlhkxihlxj)\displaystyle+\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}\biggl{(}\sum_{k=1}^{d}Y_{t}^{k}\frac{\partial^{2}h_{k}}{\partial x_{i}\partial x_{j}}+\sum_{k=1}^{d}\sum_{l=1}^{d}Y_{t}^{k}Y_{t}^{l}\frac{\partial h_{k}}{\partial x_{i}}\frac{\partial h_{l}}{\partial x_{j}}\biggr{)}
i=1dfixii,j=1dYtjhjxifi(x)12|h|2\displaystyle-\sum_{i=1}^{d}\frac{\partial f_{i}}{\partial x_{i}}-\sum_{i,j=1}^{d}Y_{t}^{j}\frac{\partial h_{j}}{\partial x_{i}}f_{i}(x)-\frac{1}{2}|h|^{2}

are stochastic functions that depend on the specific value of observation YtY_{t} at time tt.

Instead of solving equation (11) directly, we will mainly focus on the robust DMZ equation (16), especially the corresponding initial-boundary value (IBV) problems in a closed ball BR:={xd:|x|R}B_{R}:=\{x\in\mathbb{R}^{d}:|x|\leq R\}, with a given radius R>0R>0.

{uRt=12i,j=1daij(x)2uRxixj+i=1dFi(t,x)uRxi+J(t,x)uR(t,x),t[0,T],uR(0,x)=σ0(x)𝒮R(x),xBR,uR(t,x)=0,(t,x)[τk1,τk]×BR.\left\{\begin{aligned} &\frac{\partial u_{R}}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial^{2}u_{R}}{\partial x_{i}\partial x_{j}}+\sum_{i=1}^{d}F_{i}(t,x)\frac{\partial u_{R}}{\partial x_{i}}+J(t,x)u_{R}(t,x),\ t\in[0,T],\\ &u_{R}(0,x)=\sigma_{0}(x)\cdot\mathscr{S}_{R}(x),\ x\in B_{R},\\ &u_{R}(t,x)=0,\ (t,x)\in[\tau_{k-1},\tau_{k}]\times\partial B_{R}.\end{aligned}\right. (18)

where 𝒮R(x)\mathscr{S}_{R}(x) is a CC^{\infty} function supported in BRB_{R} and satisfies

𝒮R(x)={1,|x|R1R0,|x|R\mathscr{S}_{R}(x)=\left\{\begin{aligned} &1,\quad|x|\leq R-\frac{1}{R}\\ &0,\quad|x|\geq R\end{aligned}\right. (19)

and 0𝒮R(x)10\leq\mathscr{S}_{R}(x)\leq 1 for R1R|x|RR-\frac{1}{R}\leq|x|\leq R, so that the initial value is compatible with the boundary condition in (18). And from now on, we would like to drop the subscript, RR, in the notation uR(t,x)u_{R}(t,x) for the simplicity of notations, and use u(t,x)u(t,x) to denote the solution to the IBV problem (19).

Let 0=τ0<τ1<<τK=T0=\tau_{0}<\tau_{1}<\cdots<\tau_{K}=T be a uniform partition of the time interval [0,T][0,T], with τkτk1=δ=TK\tau_{k}-\tau_{k-1}=\delta=\frac{T}{K}, k=1,,Kk=1,\cdots,K. On each time interval [τk1,τk][\tau_{k-1},\tau_{k}], consider the IBV problem of the following parabolic equation

{ukt=12i,j=1daij(x)2ukxixj+i=1dFi(τk1,x)ukxi+J(τk1,x)uk(t,x),(t,x)(τk1,τk]×BR,uk(τk1,x)=uk1(τk1,x),xBR,uk(t,x)=0,(t,x)[τk1,τk]×BR,\left\{\begin{aligned} &\frac{\partial u_{k}}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial^{2}u_{k}}{\partial x_{i}\partial x_{j}}+\sum_{i=1}^{d}F_{i}(\tau_{k-1},x)\frac{\partial u_{k}}{\partial x_{i}}+J(\tau_{k-1},x)u_{k}(t,x),\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad(t,x)\in(\tau_{k-1},\tau_{k}]\times B_{R},\\ &u_{k}(\tau_{k-1},x)=u_{k-1}(\tau_{k-1},x),\ x\in B_{R},\\ &u_{k}(t,x)=0,\ (t,x)\in[\tau_{k-1},\tau_{k}]\times\partial B_{R},\end{aligned}\right. (20)

with the value of coefficients F(t,x)F(t,x) and J(t,x)J(t,x) frozen at the left point t=τk1t=\tau_{k-1} and initial value u0(τ0,x):=σ0(x)u_{0}(\tau_{0},x):=\sigma_{0}(x).

With another exponential transformation given by

u~k(t,x)=exp(h(x)Yτk1)uk(t,x),t[τk1,τk],\tilde{u}_{k}(t,x)=\exp\left(h^{\top}(x)Y_{\tau_{k-1}}\right)u_{k}(t,x),\ t\in[\tau_{k-1},\tau_{k}], (21)

the newly-constructed function u~k\tilde{u}_{k} satisfies

{u~kt=12i,j=1d2xixj(aiju~k(t,x))i=1dxi(fiu~k(t,x))12|h|2u~k(t,x),(t,x)(τk1,τk]×BR,u~k(τk1,x)=exp(h(x)Yτk1)uk(τk1,x),xBRu~k(t,x)=0,(t,x)[τk1,τk]×BR\left\{\begin{aligned} &\frac{\partial\tilde{u}_{k}}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}\tilde{u}_{k}(t,x))\\ &\quad-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}\tilde{u}_{k}(t,x))-\frac{1}{2}|h|^{2}\tilde{u}_{k}(t,x),\ (t,x)\in(\tau_{k-1},\tau_{k}]\times B_{R},\\ &\tilde{u}_{k}(\tau_{k-1},x)=\exp\biggl{(}h^{\top}(x)Y_{\tau_{k-1}}\biggr{)}u_{k}(\tau_{k-1},x),\ x\in B_{R}\\ &\tilde{u}_{k}(t,x)=0,\ (t,x)\in[\tau_{k-1},\tau_{k}]\times\partial B_{R}\end{aligned}\right. (22)

After the two exponential transformations (15) and (21), the function we would like to use to approximate the unnormalized conditional probability density function σ(τk,x)\sigma(\tau_{k},x) at time t=τkt=\tau_{k} is given by

σ(τk,x)exp(h(x)(YτkYτk1))u~k(τk,x)=u~k+1(τk,x),k=1,,K.\sigma(\tau_{k},x)\approx\exp\biggl{(}h^{\top}(x)(Y_{\tau_{k}}-Y_{\tau_{k-1}})\biggr{)}\tilde{u}_{k}(\tau_{k},x)=\tilde{u}_{k+1}(\tau_{k},x),\ k=1,\cdots,K. (23)

and the value for approximating the conditional expectation E[φ(Xt)|𝒴t]E[\varphi(X_{t})|\mathcal{Y}_{t}] is given by

E[φ(Xt)|𝒴t]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x,k=1,,K.E[\varphi(X_{t})|\mathcal{Y}_{t}]\approx\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx},\ k=1,\cdots,K. (24)

The main idea of the Yau-Yau algorithm is that the problem of solving the DMZ equation satisfied by the unnormalized probability density function σ(t,x)\sigma(t,x) can be separated into two parts. The computationally expensive part of solving the (IBV) problems of parabolic equation (22) can be done off-line, because it is a deterministic Kolmogorov-type PDE which is independent of observations, and at least, the corresponding semi-group, {St:t[0,T]}\{S_{t}:t\in[0,T]\}, can be analyzed and approximated off-line. When new observation comes, the remaining task is only about calculating exponential transformations and numerical integrals. The framework of Yau-Yau algorithm is shown in Algorithm 1.

Algorithm 1 The Two-Stage Framework of Yau-Yau Algorithm
1:  Initialization: Input the terminal time TT, the radius RR of closed ball BRB_{R}, the number of time-discretization steps KK, the initial distribution of state process σ0(x)\sigma_{0}(x), the test function φ(x)\varphi(x), and the initial observation Y0=0Y_{0}=0. Let δ=TK\delta=\frac{T}{K} be the time-discretization step size and {0=τ0<τ1<<τK=T}\{0=\tau_{0}<\tau_{1}<\cdots<\tau_{K}=T\} be a uniform partition of [0,T][0,T] with τkτk1=δ\tau_{k}-\tau_{k-1}=\delta. Initialize u~1(0,x)=σ0(x)\tilde{u}_{1}(0,x)=\sigma_{0}(x).
2:  Off-Line Algorithm: Solve the IBV problem of Kolmogorov-type partial differential equation (22) in closed ball BRB_{R}, and determine or approximate the corresponding semi-group {St:t[0,T]}\{S_{t}:t\in[0,T]\}.
3:  On-Line Algorithm:
4:  for k=1k=1 to KK do
5:     Obtain u~k(τk,x)\tilde{u}_{k}(\tau_{k},x) from the Off-Line Algorithm
u~k(τk,x)=Sτkτk1u~k(τk1,x).\tilde{u}_{k}(\tau_{k},x)=S_{\tau_{k}-\tau_{k-1}}\tilde{u}_{k}(\tau_{k-1},x).
6:     Renew the initial value of the partial differential equation satisfied by u~k+1(x,t)\tilde{u}_{k+1}(x,t),
u~k+1(τk,x)=exp[h(x)(YτkYτk1)]u~k(τk,x).\tilde{u}_{k+1}(\tau_{k},x)=\exp\left[h^{\top}(x)(Y_{\tau_{k}}-Y_{\tau_{k-1}})\right]\tilde{u}_{k}(\tau_{k},x).
7:     Compute the approximated conditional expectation:
BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x.\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}.
8:  end for

In the next section, we will give a mathematically rigorous interpretation of the approximation results (23) and (24) from a probabilistic perspective. In particular, we only need assumptions on the test function and the coefficients of the filtering systems to derive the convergence result. These assumptions are also easy to verify off-line before the observations come, and therefore, this convergence analysis will provide a guidance for practitioners to determine the parameters in the implementations of Yau-Yau algorithm for practical use.

3 Main Results

In this section, we would like to state the main result in this paper and also provide a sketch of the proof.

Firstly, besides the smoothness and regularity requirements which guarantee the existence of conditional expectation and the existence of the solution to the DMZ equation, let us further introduce four particular assumptions on the coefficients of the system, the initial distribution and the test function.

For the state equation in the filtering system (1), the drift term f:ddf:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} is assumed to be Lipschitz, and the the diffusion term g:dd×dg:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d\times d}, together with a(x)=g(x)g(x)a(x)=g(x)g(x)^{\top}, is assumed to have bounded partial derivatives up to second order, i.e.,

(A1)L>0,s.t.|f(x)f(y)|L|xy|,|a(x)|L,|aij(x)xk|L,|2aij(x)xkxl|L,x,yd,i,j,k,l=1,,d.\displaystyle\text{{(A1)}: }\begin{aligned} &\exists\ L>0,\ s.t.\ |f(x)-f(y)|\leq L|x-y|,\\ &|a(x)|\leq L,\ \biggl{|}\frac{\partial a^{ij}(x)}{\partial x_{k}}\biggr{|}\leq L,\ \biggl{|}\frac{\partial^{2}a^{ij}(x)}{\partial x_{k}\partial x_{l}}\biggr{|}\leq L,\ \forall\ x,y\in\mathbb{R}^{d},\ i,j,k,l=1,\cdots,d.\end{aligned}

Assumption (A1) guarantees that the state equation of (1) has a strong solution in [0,T][0,T], and especially, the state equation for linear filter satisfies this assumption.

Also, in order to conduct energy estimations for the (stochastic) partial differential equations, we would like to assume that the diffusion term in the state equation is nondegenerate, in the sense that

(A2):For each xd, there exists a continuous function λ(x)>0,s.t.i,j=1daij(x)ζiζjλ(x)|ζ|2,ζ=(ζ1,,ζd)d.\displaystyle\text{{(A2):}}\quad\begin{aligned} &\text{For each $x\in\mathbb{R}^{d}$, there exists a continuous function $\lambda(x)>0$},\\ &s.t.\ \sum_{i,j=1}^{d}a^{ij}(x)\zeta_{i}\zeta_{j}\geq\lambda(x)|\zeta|^{2},\ \forall\ \zeta=(\zeta_{1},\cdots,\zeta_{d})^{\top}\in\mathbb{R}^{d}.\end{aligned}

For the initial distribution σ0\sigma_{0}, we would like to assume that it is smooth enough and possesses finite high-order moments:

(A3):d|x|2nσ0(x)𝑑x<,n.\displaystyle\text{{(A3):}}\ \int_{\mathbb{R}^{d}}|x|^{2n}\sigma_{0}(x)dx<\infty,\ \forall\ n\in\mathbb{N}.

Assumption (A3) is satisfied by commonly-used light-tailed distributions such as Gaussian distributions. In fact, in the following convergence analysis, we only require Assumption (A3) to hold for sufficiently large n1n\geq 1, rather than for all nn\in\mathbb{N}.

Finally, it is assumed that the test function φ:d\varphi:\mathbb{R}^{d}\rightarrow\mathbb{R} is at most polynomial growth:

(A4): L>0,m,s.t.|φ(x)|L(1+|x|2m),xd,\displaystyle\text{{(A4): }}\exists\ L>0,\ m\in\mathbb{N},\ s.t.\ |\varphi(x)|\leq L(1+|x|^{2m}),\quad\forall\ x\in\mathbb{R}^{d},

which is satisfied by most of the commonly-used test functions, such as those correspond to the conditional mean and covariance matrix.

Based on the above assumptions (A1) to (A4), the main result in this paper is stated as follows:

Theorem 1.

Fix a terminal time T>0T>0 and the filtering system (1) with smooth coefficients. If Assumptions (A1) to (A4) hold, then for every ϵ>0\epsilon>0, there exists R>0R>0, and δ>0\delta>0, such that we can conduct the Yau-Yau algorithm in the closed ball BR={xd:|x|R}B_{R}=\{x\in\mathbb{R}^{d}:|x|\leq R\} and the uniform partition of [0,T][0,T]: 0=τ0<τ1<<τK=T0=\tau_{0}<\tau_{1}<\cdots<\tau_{K}=T with δ=τkτk1\delta=\tau_{k}-\tau_{k-1} for k=1,,Kk=1,\cdots,K, and the numerical solution {u~k+1(τk,x):k=1,,K}\{\tilde{u}_{k+1}(\tau_{k},x):k=1,\cdots,K\} approximates the exact solution of the filtering problems at each time step t=τkt=\tau_{k} well in the sense of mathematical expectation, i.e.,

E|E[φ(Xτk)|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|<ϵ, 1kK.E\biggl{|}E[\varphi(X_{\tau_{k}})|\mathcal{Y}_{\tau_{k}}]-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}<\epsilon,\ \forall\ 1\leq k\leq K. (25)

Here in this section, we provide a sketch of the proof of Theorem 25, in which the main idea of the proof is illustrated. The detailed proofs of those key estimations here will be given in order in the next four sections.

A Sketch of the Proof of Theorem 25.

Let {Z~t:0tT}\{\tilde{Z}_{t}:0\leq t\leq T\} be the Radon derivative Zt=dPdP~t|tZ_{t}=\left.\frac{dP}{d\tilde{P}_{t}}\right|_{\mathcal{F}_{t}} defined in (4). And therefore, for every integrable, t\mathcal{F}_{t}-measurable random variable UU, we have

E[U]=E~[Z~tU].E[U]=\tilde{E}[\tilde{Z}_{t}U]. (26)

According to the properties of conditional expectations, the expectation of the approximation error of Yau-Yau algorithm can be estimated as follows:

E\displaystyle E |E[φ(Xτk)|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|\displaystyle\biggl{|}E[\varphi(X_{\tau_{k}})|\mathcal{Y}_{\tau_{k}}]-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}
=E~[Z~τk|E[φ(Xτk)|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle=\tilde{E}\biggl{[}\tilde{Z}_{\tau_{k}}\biggl{|}E[\varphi(X_{\tau_{k}})|\mathcal{Y}_{\tau_{k}}]-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}\biggr{]}
=E~[Z~τk|E~[φ(Xτk)Z~τk|𝒴τk]E~[Z~τk|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle=\tilde{E}\biggl{[}\tilde{Z}_{\tau_{k}}\biggl{|}\frac{\tilde{E}[\varphi(X_{\tau_{k}})\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}{\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}\biggr{]}
=E~[E~[Z~τk|𝒴τk]|E~[φ(Xτk)Z~τk|𝒴τk]E~[Z~τk|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle=\tilde{E}\biggl{[}\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]\biggl{|}\frac{\tilde{E}[\varphi(X_{\tau_{k}})\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}{\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}\biggr{]}
E~[E~[Z~τk|𝒴τk]|E~[φ(Xτk)Z~τk|𝒴τk]E~[Z~τk|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xE~[Z~τk|𝒴τk]|]\displaystyle\leq\tilde{E}\biggl{[}\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]\biggl{|}\frac{\tilde{E}[\varphi(X_{\tau_{k}})\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}{\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}\biggr{|}\biggr{]}
+E~[E~[Z~τk|𝒴τk]|BRφ(x)u~k+1(τk,x)𝑑xE~[Z~τk|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle\quad+\tilde{E}\biggl{[}\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]\biggl{|}\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]}-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}\biggr{]}
E~[|E~[φ(Xτk)Z~τk|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑x|]\displaystyle\leq\tilde{E}\biggl{[}\biggl{|}\tilde{E}[\varphi(X_{\tau_{k}})\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]-\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
+E~[BR|φ(x)|u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|E~[Z~τk|𝒴τk]BRu~k+1(τk,x)𝑑x|]\displaystyle\quad+\tilde{E}\biggl{[}\frac{\int_{B_{R}}|\varphi(x)|\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggl{|}\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]-\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
=E~[|dφ(x)σ(τk,x)𝑑xBRφ(x)u~k+1(τk,x)𝑑x|]\displaystyle=\tilde{E}\biggl{[}\biggl{|}\int_{\mathbb{R}^{d}}\varphi(x)\sigma(\tau_{k},x)dx-\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
+E~[BR|φ(x)|u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|dσ(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle\quad+\tilde{E}\biggl{[}\frac{\int_{B_{R}}|\varphi(x)|\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggl{|}\int_{\mathbb{R}^{d}}\sigma(\tau_{k},x)dx-\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
I1+I2.\displaystyle\triangleq I_{1}+I_{2}.

where we use the fact that u~k+1(τk,x)\tilde{u}_{k+1}(\tau_{k},x) is 𝒴τk\mathcal{Y}_{\tau_{k}}-measurable and for integrable, 𝒴τk\mathcal{Y}_{\tau_{k}}-measurable random variable VV,

E~[Z~τkV]=E~[E~[Z~τk|𝒴τk]V].\tilde{E}[\tilde{Z}_{\tau_{k}}V]=\tilde{E}[\tilde{E}[\tilde{Z}_{\tau_{k}}|\mathcal{Y}_{\tau_{k}}]V]. (27)

Therefore, the remaining task for us is to estimate the two error terms I1I_{1} and I2I_{2}, and to show that I1I_{1} and I2I_{2} can be arbitrarily small with sufficiently large R>0R>0 and sufficiently small δ>0\delta>0.

Firstly, for the estimation of

I1=E~[|dφ(x)σ(τk,x)𝑑xBRφ(x)u~k+1(τk,x)𝑑x|],I_{1}=\tilde{E}\biggl{[}\biggl{|}\int_{\mathbb{R}^{d}}\varphi(x)\sigma(\tau_{k},x)dx-\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}, (28)

we would like to utilize an intermediate function σR(t,x)\sigma_{R}(t,x), (t,x)[0,T]×BR(t,x)\in[0,T]\times B_{R}, which is the solution of IBV problem of the DMZ equation (11) and will be introduced in (43) in Section 4. And we have

I1\displaystyle I_{1}\leq E~|x|R|φ(x)|σ(τk,x)𝑑x+E~[|BRφ(x)σ(τk,x)𝑑xBRφ(x)σR(τk,x)𝑑x|]\displaystyle\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(\tau_{k},x)dx+\tilde{E}\biggl{[}\biggl{|}\int_{B_{R}}\varphi(x)\sigma(\tau_{k},x)dx-\int_{B_{R}}\varphi(x)\sigma_{R}(\tau_{k},x)dx\biggr{|}\biggr{]}
+E~[|BRφ(x)σR(τk,x)𝑑xBRφ(x)u~k+1(τk,x)𝑑x|]\displaystyle+\tilde{E}\biggl{[}\biggl{|}\int_{B_{R}}\varphi(x)\sigma_{R}(\tau_{k},x)dx-\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
\displaystyle\leq E~|x|R|φ(x)|σ(τk,x)𝑑x+E~BR|φ(x)||σ(τk,x)σR(τk,x)|𝑑x\displaystyle\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(\tau_{k},x)dx+\tilde{E}\int_{B_{R}}|\varphi(x)|\cdot|\sigma(\tau_{k},x)-\sigma_{R}(\tau_{k},x)|dx
+E~BR|φ(x)||σR(τk,x)u~k+1(τk,x)|𝑑x\displaystyle+\tilde{E}\int_{B_{R}}|\varphi(x)|\cdot|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx
\displaystyle\leq E~|x|R|φ(x)|σ(τk,x)𝑑x+L(1+R2m)E~BR|σ(τk,x)σR(τk,x)|𝑑x\displaystyle\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(\tau_{k},x)dx+L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma(\tau_{k},x)-\sigma_{R}(\tau_{k},x)|dx
+L(1+R2m)E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x.\displaystyle+L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx. (29)

For the estimation of

I2=E~[BR|φ(x)|u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|dσ(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|],I_{2}=\tilde{E}\biggl{[}\frac{\int_{B_{R}}|\varphi(x)|\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggl{|}\int_{\mathbb{R}^{d}}\sigma(\tau_{k},x)dx-\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}, (30)

since in the closed ball BRB_{R}, |φ(x)|L(1+R2m)|\varphi(x)|\leq L(1+R^{2m}),

BR|φ(x)|u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑xL(1+R2m)BRu~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x=L(1+Rm).\frac{\int_{B_{R}}|\varphi(x)|\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\leq L(1+R^{2m})\frac{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}=L(1+R^{m}). (31)

and thus,

I2\displaystyle I_{2} L(1+R2m)E~[|dσ(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle\leq L(1+R^{2m})\tilde{E}\biggl{[}\biggl{|}\int_{\mathbb{R}^{d}}\sigma(\tau_{k},x)dx-\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]} (32)
L(1+R2m)(E~|x|Rσ(τk,x)𝑑x+E~[|BRσ(τk,x)𝑑xBRσR(τk,x)𝑑x|])\displaystyle\leq L(1+R^{2m})\biggl{(}\tilde{E}\int_{|x|\geq R}\sigma(\tau_{k},x)dx+\tilde{E}\biggl{[}\biggl{|}\int_{B_{R}}\sigma(\tau_{k},x)dx-\int_{B_{R}}\sigma_{R}(\tau_{k},x)dx\biggr{|}\biggr{]}\biggr{)}
+L(1+R2m)E~[|BRσR(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|]\displaystyle\quad+L(1+R^{2m})\tilde{E}\biggl{[}\biggl{|}\int_{B_{R}}\sigma_{R}(\tau_{k},x)dx-\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx\biggr{|}\biggr{]}
L(1+R2m)(E~|x|Rσ(τk,x)𝑑x+E~BR|σ(τk,x)σR(τk,x)|𝑑x)\displaystyle\leq L(1+R^{2m})\biggl{(}\tilde{E}\int_{|x|\geq R}\sigma(\tau_{k},x)dx+\tilde{E}\int_{B_{R}}|\sigma(\tau_{k},x)-\sigma_{R}(\tau_{k},x)|dx\biggr{)}
+L(1+R2m)E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x.\displaystyle\quad+L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx.

Combining (29) and (32), we have

E\displaystyle E |E[φ(Xτk)|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|I1+I2\displaystyle\biggl{|}E[\varphi(X_{\tau_{k}})|\mathcal{Y}_{\tau_{k}}]-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}\leq I_{1}+I_{2} (33)
E~|x|R|φ(x)|σ(τk,x)𝑑x+L(1+Rm)E~|x|Rσ(τk,x)𝑑x\displaystyle\leq\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(\tau_{k},x)dx+L(1+R^{m})\tilde{E}\int_{|x|\geq R}\sigma(\tau_{k},x)dx
+2L(1+R2m)E~BR|σ(τk,x)σR(τk,x)|𝑑x\displaystyle\quad+2L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma(\tau_{k},x)-\sigma_{R}(\tau_{k},x)|dx
+2L(1+R2m)E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x.\displaystyle\quad+2L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx.

According to Theorem 2 in Section 4, for every nn\in\mathbb{N}, there exists C1>0C_{1}>0, which depends on dd, mm, nn, LL, TT, such that

E~|x|Rσ(τk,x)𝑑xC11+R2nd|x|2nσ0(x)𝑑x\displaystyle\tilde{E}\int_{|x|\geq R}\sigma(\tau_{k},x)dx\leq\frac{C_{1}}{1+R^{2n}}\int_{\mathbb{R}^{d}}|x|^{2n}\sigma_{0}(x)dx (34)
E~|x|R|φ(x)|σ(τk,x)𝑑xC11+R2nd|x|2(m+n)σ0(x)𝑑x\displaystyle\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(\tau_{k},x)dx\leq\frac{C_{1}}{1+R^{2n}}\int_{\mathbb{R}^{d}}|x|^{2(m+n)}\sigma_{0}(x)dx

Therefore, for every ϵ>0\epsilon>0, with Assumption (A3) for the initial distribution σ0\sigma_{0}, as long as we take n>mn>m, there exists R1>0R_{1}>0, such that

E~\displaystyle\tilde{E} |x|R1|φ(x)|σ(τk,x)𝑑x+L(1+R1m)E~|x|R1σ(τk,x)𝑑x\displaystyle\int_{|x|\geq R_{1}}|\varphi(x)|\sigma(\tau_{k},x)dx+L(1+R_{1}^{m})\tilde{E}\int_{|x|\geq R_{1}}\sigma(\tau_{k},x)dx (35)
C1(1+R12m)1+R12nd|x|2nσ0(x)𝑑x+C11+R12nd|x|2(m+n)σ0(x)𝑑x<ϵ3\displaystyle\leq\frac{C_{1}(1+R_{1}^{2m})}{1+R_{1}^{2n}}\int_{\mathbb{R}^{d}}|x|^{2n}\sigma_{0}(x)dx+\frac{C_{1}}{1+R_{1}^{2n}}\int_{\mathbb{R}^{d}}|x|^{2(m+n)}\sigma_{0}(x)dx<\frac{\epsilon}{3}

According to Theorem 3 in Section 5, there exists C2>0C_{2}>0, which depends on d,n,L,Td,n,L,T, such that

E~BR|σ(τk,x)σR(τk,x)|𝑑xC21+R2n.\tilde{E}\int_{B_{R}}|\sigma(\tau_{k},x)-\sigma_{R}(\tau_{k},x)|dx\leq\frac{C_{2}}{1+R^{2n}}. (36)

Therefore, as long as n>mn>m, there exists R2>0R_{2}>0, such that

2L(1+R22m)E~BR2|σ(τk,x)σR2(τk,x)|𝑑x2C2L(1+R22m)1+R22n<ϵ3\displaystyle 2L(1+R_{2}^{2m})\tilde{E}\int_{B_{R_{2}}}|\sigma(\tau_{k},x)-\sigma_{R_{2}}(\tau_{k},x)|dx\leq\frac{2C_{2}L(1+R_{2}^{2m})}{1+R_{2}^{2n}}<\frac{\epsilon}{3} (37)

Let us choose R=max{R1,R2}R=\max\{R_{1},R_{2}\}, and for this particular RR, according to Theorem 5 in Section 7, there exists a time step δ>0\delta>0, such that

E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x<ϵ6L(1+R2m).\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx<\frac{\epsilon}{6L(1+R^{2m})}. (38)

and thus,

2L(1+R2m)E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x2L(1+R2m)ϵ6L(1+R2m)<ϵ3.2L(1+R^{2m})\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx\leq\frac{2L(1+R^{2m})\epsilon}{6L(1+R^{2m})}<\frac{\epsilon}{3}. (39)

Take (35)\eqref{eq:35}, (37)\eqref{eq:37} and (39)\eqref{eq:39} back to (33), and we obtain the desired result, that is, we have found R>0R>0 and δ>0\delta>0, such that

E|E[φ(Xτk)|𝒴τk]BRφ(x)u~k+1(τk,x)𝑑xBRu~k+1(τk,x)𝑑x|<ϵ.E\biggl{|}E[\varphi(X_{\tau_{k}})|\mathcal{Y}_{\tau_{k}}]-\frac{\int_{B_{R}}\varphi(x)\tilde{u}_{k+1}(\tau_{k},x)dx}{\int_{B_{R}}\tilde{u}_{k+1}(\tau_{k},x)dx}\biggr{|}<\epsilon. (40)

4 Estimation of the density outside the ball BRB_{R}

In this section, we will provide an estimation of the value of the unnormalized conditional probability density σ(t,x)\sigma(t,x) outside a ball BRdB_{R}\subset\mathbb{R}^{d}, with R1R\gg 1 large enough.

Especially, we will show that almost all the density of σ(t,x)\sigma(t,x) is contained in the closed ball BRB_{R}, and the estimations (34) in the proof of Theorem 25 in Section 3 holds with Assumptions (A1) to (A4).

Theorem 2.

With Assumptions (A1) to (A4), there exists a constant C>0C>0 which only depends on TT, LL, dd, mm and nn, such that

sup0tTE~|x|Rσ(t,x)𝑑xC1+R2nd|x|2nσ0(x)𝑑x,\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}\sigma(t,x)dx\leq\frac{C}{1+R^{2n}}\int_{\mathbb{R}^{d}}|x|^{2n}\sigma_{0}(x)dx, (41)

and

sup0tTE~|x|R|φ(x)|σ(t,x)𝑑xC1+R2nd|x|2(m+n)σ0(x)𝑑x\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}|\varphi(x)|\sigma(t,x)dx\leq\frac{C}{1+R^{2n}}\int_{\mathbb{R}^{d}}|x|^{2(m+n)}\sigma_{0}(x)dx (42)

holds for all R>0R>0.

Proof of Theorem 2.

We first consider the following IBV problem on the ball BRB_{R}:

{dσR(t,x)=σR(t,x)dt+j=1dhj(x)σR(t,x)dYtj,(t,x)(0,T]×BR;σR(0,x)=σ0,R(x)σ0(x)𝒮R(x),xBR;σR(t,x)=0,(t,x)[0,T]×BR.\left\{\begin{aligned} &d\sigma_{R}(t,x)=\mathcal{L}\sigma_{R}(t,x)dt+\sum_{j=1}^{d}h_{j}(x)\sigma_{R}(t,x)dY_{t}^{j},\ (t,x)\in(0,T]\times B_{R};\\ &\sigma_{R}(0,x)=\sigma_{0,R}(x)\triangleq\sigma_{0}(x)\cdot\mathscr{S}_{R}(x),\ x\in B_{R};\\ &\sigma_{R}(t,x)=0,\ (t,x)\in[0,T]\times\partial B_{R}.\end{aligned}\right. (43)

where 𝒮R(x)\mathscr{S}_{R}(x) is the CC^{\infty} function defined in (19), such that the initial value is compatible with the boundary conditions.

Let ψ(x)=log(1+|x|2n)\psi(x)=\log\left(1+|x|^{2n}\right) and define

Φ(t)=BReψ(x)σR(t,x)𝑑x.\Phi(t)=\int_{B_{R}}e^{\psi(x)}\sigma_{R}(t,x)dx. (44)

Then, according to the IBV problem (43) satisfied by the function σR(t,x)\sigma_{R}(t,x), we have

dΦ(t)=\displaystyle d\Phi(t)= [12i,j=1dBR2xixj[(aij(x))σR(t,x)]eψ(x)dx\displaystyle\biggl{[}\frac{1}{2}\sum_{i,j=1}^{d}\int_{B_{R}}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}\left[\left(a^{ij}(x)\right)\sigma_{R}(t,x)\right]e^{\psi(x)}dx (45)
i=1dBRxi(fi(x)σR(t,x))eψ(x)dx]dt\displaystyle-\sum_{i=1}^{d}\int_{B_{R}}\frac{\partial}{\partial x_{i}}\left(f_{i}(x)\sigma_{R}(t,x)\right)e^{\psi(x)}dx\biggr{]}dt
+j=1d[BReψ(x)hj(x)σR(t,x)𝑑x]dYtj\displaystyle+\sum_{j=1}^{d}\biggl{[}\int_{B_{R}}e^{\psi(x)}h_{j}(x)\sigma_{R}(t,x)dx\biggr{]}dY_{t}^{j}
\displaystyle\triangleq [I1(t)I2(t)]dt+j=1dI3,j(t)dYtj.\displaystyle\left[I_{1}(t)-I_{2}(t)\right]dt+\sum_{j=1}^{d}I_{3,j}(t)dY_{t}^{j}.

By the Gauss-Green formula, we have

I1(t)=12i,j=1d[\displaystyle I_{1}(t)=\frac{1}{2}\sum_{i,j=1}^{d}\biggl{[} BReψ(x)(ψ(x)xiψ(x)xj+2ψ(x)xixj)aij(x)σR(t,x)𝑑x\displaystyle\int_{B_{R}}e^{\psi(x)}\left(\frac{\partial\psi(x)}{\partial x_{i}}\frac{\partial\psi(x)}{\partial x_{j}}+\frac{\partial^{2}\psi(x)}{\partial x_{i}\partial x_{j}}\right)a^{ij}(x)\sigma_{R}(t,x)dx (46)
BRxj(eψ(x)ψ(x)xiaij(x)σR(t,x))𝑑x\displaystyle-\int_{B_{R}}\frac{\partial}{\partial x_{j}}\left(e^{\psi(x)}\frac{\partial\psi(x)}{\partial x_{i}}a^{ij}(x)\sigma_{R}(t,x)\right)dx
+BRxi(eψ(x)xj[aij(x)σR(t,x)])dx]\displaystyle+\int_{B_{R}}\frac{\partial}{\partial x_{i}}\left(e^{\psi(x)}\frac{\partial}{\partial x_{j}}\left[a^{ij}(x)\sigma_{R}(t,x)\right]\right)dx\biggr{]}
=12i,j=1d\displaystyle=\frac{1}{2}\sum_{i,j=1}^{d} BReψ(x)(ψ(x)xiψ(x)xj+2ψ(x)xixj)aij(x)σR(t,x)𝑑x\displaystyle\int_{B_{R}}e^{\psi(x)}\left(\frac{\partial\psi(x)}{\partial x_{i}}\frac{\partial\psi(x)}{\partial x_{j}}+\frac{\partial^{2}\psi(x)}{\partial x_{i}\partial x_{j}}\right)a^{ij}(x)\sigma_{R}(t,x)dx
BR𝔐1(t,x)𝔫𝑑S+BR𝔐2(t,x)𝔫𝑑S,\displaystyle-\int_{\partial B_{R}}\vec{\mathfrak{M}}_{1}(t,x)\cdot\vec{\mathfrak{n}}dS+\int_{\partial B_{R}}\vec{\mathfrak{M}}_{2}(t,x)\cdot\vec{\mathfrak{n}}dS,

where 𝔫\vec{\mathfrak{n}} is the unit outward normal vector of BR\partial B_{R}, dSdS denotes the measure on BR\partial B_{R},

𝔐i(t,x)=(𝔐i,1(t,x),,𝔐i,d(t,x)),i=1,2,\vec{\mathfrak{M}}_{i}(t,x)=\left(\mathfrak{M}_{i,1}(t,x),\cdots,\mathfrak{M}_{i,d}(t,x)\right),\quad i=1,2, (47)

and

𝔐1,j(t,x)=12i=1deψ(x)ψ(x)xiaij(x)σR(t,x),j=1,,d,\displaystyle\mathfrak{M}_{1,j}(t,x)=\frac{1}{2}\sum_{i=1}^{d}e^{\psi(x)}\frac{\partial\psi(x)}{\partial x_{i}}a^{ij}(x)\sigma_{R}(t,x),\quad j=1,\cdots,d, (48)
𝔐2,i(t,x)=12j=1deψ(x)xj[aij(x)σR(t,x)],i=1,,d.\displaystyle\mathfrak{M}_{2,i}(t,x)=\frac{1}{2}\sum_{j=1}^{d}e^{\psi(x)}\frac{\partial}{\partial x_{j}}\left[a^{ij}(x)\sigma_{R}(t,x)\right],\quad i=1,\cdots,d.

Since σR(t,x)0\sigma_{R}(t,x)\equiv 0, (t,x)[0,T]×BR\forall(t,x)\in[0,T]\times\partial B_{R}, 𝔐1,j(t,x)0\mathfrak{M}_{1,j}(t,x)\equiv 0 on [0,T]×BR[0,T]\times\partial B_{R} and

BR𝔐1(t,x)𝔫𝑑S=0.\int_{\partial B_{R}}\vec{\mathfrak{M}}_{1}(t,x)\cdot\vec{\mathfrak{n}}dS=0. (49)

Moreover, we have

σR=(σRx1,,σRxd)=c𝔫,onBR,\nabla\sigma_{R}=\left(\frac{\partial\sigma_{R}}{\partial x_{1}},\cdots,\frac{\partial\sigma_{R}}{\partial x_{d}}\right)=-c\ \vec{\mathfrak{n}},\quad\text{on}\ \partial B_{R}, (50)

where c(x)>0c(x)>0 is a continuous function on BR\partial B_{R}, because σR0\sigma_{R}\geq 0 and σR|BR0\sigma_{R}|_{\partial B_{R}}\equiv 0.

Therefore,

𝔐2(t,x)𝔫=\displaystyle\vec{\mathfrak{M}}_{2}(t,x)\cdot\vec{\mathfrak{n}}= 𝔐2(t,x)σR\displaystyle-\vec{\mathfrak{M}}_{2}(t,x)\cdot\nabla\sigma_{R} (51)
=\displaystyle= eψ(x)12i,j=1daij(x)σRxjσRxi\displaystyle-e^{\psi(x)}\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial\sigma_{R}}{\partial x_{j}}\frac{\partial\sigma_{R}}{\partial x_{i}}
eψ(x)σR12i,j=1dxjaij(x)σRxi\displaystyle-e^{\psi(x)}\sigma_{R}\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial}{\partial x_{j}}a^{ij}(x)\frac{\partial\sigma_{R}}{\partial x_{i}}
=\displaystyle= eψ(x)12i,j=1daij(x)σRxjσRxi0,onBR,\displaystyle-e^{\psi(x)}\frac{1}{2}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial\sigma_{R}}{\partial x_{j}}\frac{\partial\sigma_{R}}{\partial x_{i}}\leq 0,\quad\text{on}\ \partial B_{R},

where the last inequality holds because a(x)=g(x)g(x)a(x)=g(x)g(x)^{\top} is positive semi-definite. Thus,

I1(t)12i,j=1dBReψ(x)(ψ(x)xiψ(x)xj+2ψ(x)xixj)aij(x)σR(t,x)𝑑x.I_{1}(t)\leq\frac{1}{2}\sum_{i,j=1}^{d}\int_{B_{R}}e^{\psi(x)}\left(\frac{\partial\psi(x)}{\partial x_{i}}\frac{\partial\psi(x)}{\partial x_{j}}+\frac{\partial^{2}\psi(x)}{\partial x_{i}\partial x_{j}}\right)a^{ij}(x)\sigma_{R}(t,x)dx. (52)

Similarly,

I2(t)=i=1dBRfi(x)σR(t,x)eψ(x)ψ(x)xi𝑑x.\displaystyle I_{2}(t)=-\sum_{i=1}^{d}\int_{B_{R}}f_{i}(x)\sigma_{R}(t,x)e^{\psi(x)}\frac{\partial\psi(x)}{\partial x_{i}}dx. (53)

Therefore,

dΦ(t)(BR𝔉(x)eψ(x)σR(t,x)𝑑x)dt+j=1d(BRhj(x)eψ(x)u(t,x)𝑑x)dYtj,d\Phi(t)\leq\left(\int_{B_{R}}\mathfrak{F}(x)e^{\psi(x)}\sigma_{R}(t,x)dx\right)dt+\sum_{j=1}^{d}\left(\int_{B_{R}}h_{j}(x)e^{\psi(x)}u(t,x)dx\right)dY_{t}^{j}, (54)

where

𝔉(x)=12i,j=1d(ψ(x)xiψ(x)xj+2ψ(x)xixj)aij(x)+i=1dfi(x)ψ(x)xi.\displaystyle\mathfrak{F}(x)=\frac{1}{2}\sum_{i,j=1}^{d}\left(\frac{\partial\psi(x)}{\partial x_{i}}\frac{\partial\psi(x)}{\partial x_{j}}+\frac{\partial^{2}\psi(x)}{\partial x_{i}\partial x_{j}}\right)a^{ij}(x)+\sum_{i=1}^{d}f_{i}(x)\frac{\partial\psi(x)}{\partial x_{i}}. (55)

Since ψ(x)=log(1+|x|2n)\psi(x)=\log\left(1+|x|^{2n}\right), then

ψxi=2n|x|2n2xi1+|x|2n,2ψxixj=4nxixj|x|2n4((n1)|x|2n)(1+|x|2n)2+2n|x|2n2δij1+|x|2n,\frac{\partial\psi}{\partial x_{i}}=\frac{2n|x|^{2n-2}x_{i}}{1+|x|^{2n}},\ \frac{\partial^{2}\psi}{\partial x_{i}\partial x_{j}}=\frac{4nx_{i}x_{j}|x|^{2n-4}\left((n-1)-|x|^{2n}\right)}{\left(1+|x|^{2n}\right)^{2}}+\frac{2n|x|^{2n-2}\mathscr{\delta}_{ij}}{1+|x|^{2n}}, (56)

where δij\mathcal{\delta}_{ij} is the Kronecker’s symbol, with δij=1\mathcal{\delta}_{ij}=1, if i=ji=j, and δij=0\mathcal{\delta}_{ij}=0 otherwise.

Notice that

|ψxi|2n,|2ψxixj|4n2+2n.\left|\frac{\partial\psi}{\partial x_{i}}\right|\leq 2n,\ \left|\frac{\partial^{2}\psi}{\partial x_{i}\partial x_{j}}\right|\leq 4n^{2}+2n. (57)

With the assumption that |aij(x)|L|a^{ij}(x)|\leq L, we have

|𝔉(x)|d2(4n2+n)L+2ni=1d|fi(x)xi||x|2n21+|x|2n,xd.|\mathfrak{F}(x)|\leq d^{2}(4n^{2}+n)L+2n\sum_{i=1}^{d}\frac{|f_{i}(x)x_{i}|\cdot|x|^{2n-2}}{1+|x|^{2n}},\ \forall x\in\mathbb{R}^{d}. (58)

Because f(x)f(x) is Lipschitz continuous according to (A1).

|f(x)|L|x|+|f(0)|,xd.|f(x)|\leq L|x|+|f(0)|,\quad\forall x\in\mathbb{R}^{d}. (59)

Therefore,

|𝔉(x)|\displaystyle|\mathfrak{F}(x)| d2(4n2+n)L+2n|x|2n21+|x|2ni=1d|fi(x)|2+|xi|22\displaystyle\leq d^{2}(4n^{2}+n)L+\frac{2n|x|^{2n-2}}{1+|x|^{2n}}\sum_{i=1}^{d}\frac{|f_{i}(x)|^{2}+|x_{i}|^{2}}{2} (60)
d2(4n2+n)L+n|x|2n+n|x|2n2|f(x)|21+|x|2n\displaystyle\leq d^{2}(4n^{2}+n)L+\frac{n|x|^{2n}+n|x|^{2n-2}|f(x)|^{2}}{1+|x|^{2n}}
d2(4n2+n)L+n(L2+1)+n|f(0)|2+2nL|f(0)|,xd.\displaystyle\leq d^{2}(4n^{2}+n)L+n(L^{2}+1)+n|f(0)|^{2}+2nL|f(0)|,\ \forall x\in\mathbb{R}^{d}.

Let us denote by M(n,d,L)M(n,d,L) the above upper bound of |𝔉(x)||\mathfrak{F}(x)|:

M(n,d,L):=d2(4n2+n)L+n(L2+1)+n|f(0)|2+2nL|f(0)|,M(n,d,L):=d^{2}(4n^{2}+n)L+n(L^{2}+1)+n|f(0)|^{2}+2nL|f(0)|, (61)

which is a constant that depends on nn, dd and LL, but does not depend on RR.

Take expectation with respect to the reference probability measure P~\tilde{P}, we obtain

ddtE~Φ(t)M(n,d,L)E~Φ(t).\frac{d}{dt}\tilde{E}\Phi(t)\leq M(n,d,L)\tilde{E}\Phi(t). (62)

Here we use the fact that YtY_{t} is a Brownian motion with respect to P~\tilde{P}.

According to the Gronwall’s inequality, we have

sup0tTE~BR(1+|x|2n)σR(t,x)𝑑x\displaystyle\sup\limits_{0\leq t\leq T}\tilde{E}\int_{B_{R}}\left(1+|x|^{2n}\right)\sigma_{R}(t,x)dx eM(n,d,L)TBR(1+|x|2n)σ0,R(x)𝑑x\displaystyle\leq e^{M(n,d,L)T}\int_{B_{R}}\left(1+|x|^{2n}\right)\sigma_{0,R}(x)dx (63)
eM(n,d,L)TBR(1+|x|2n)σ0(x)𝑑x.\displaystyle\leq e^{M(n,d,L)T}\int_{B_{R}}\left(1+|x|^{2n}\right)\sigma_{0}(x)dx.

Let RR tends to infinity, and we have

sup0tTE~d(1+|x|2n)σ(t,x)𝑑xeM(n,d,L)Td(1+|x|2n)σ0(x)𝑑x.\sup\limits_{0\leq t\leq T}\tilde{E}\int_{\mathbb{R}^{d}}\left(1+|x|^{2n}\right)\sigma(t,x)dx\leq e^{M(n,d,L)T}\int_{\mathbb{R}^{d}}\left(1+|x|^{2n}\right)\sigma_{0}(x)dx. (64)

Therefore,

sup0tTE~|x|Rσ(t,x)𝑑x\displaystyle\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}\sigma(t,x)dx 11+R2nsup0tTE~|x|R(1+|x|2n)σ(t,x)𝑑x\displaystyle\leq\frac{1}{1+R^{2n}}\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}\left(1+|x|^{2n}\right)\sigma(t,x)dx (65)
11+R2nsup0tTE~d(1+|x|2n)σ(t,x)𝑑x\displaystyle\leq\frac{1}{1+R^{2n}}\sup\limits_{0\leq t\leq T}\tilde{E}\int_{\mathbb{R}^{d}}\left(1+|x|^{2n}\right)\sigma(t,x)dx
eM(n,d,L)T1+R2nd(1+|x|2n)σ0(x)𝑑x.\displaystyle\leq\frac{e^{M(n,d,L)T}}{1+R^{2n}}\int_{\mathbb{R}^{d}}\left(1+|x|^{2n}\right)\sigma_{0}(x)dx.

Moreover, with condition (A4),

sup0tTE~\displaystyle\sup\limits_{0\leq t\leq T}\tilde{E} |x|R|φ(x)|σ(t,x)𝑑xsup0tTE~|x|R(1+|x|2m)σ(t,x)𝑑x\displaystyle\int_{|x|\geq R}|\varphi(x)|\sigma(t,x)dx\leq\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}\left(1+|x|^{2m}\right)\sigma(t,x)dx (66)
11+R2nsup0tTE~|x|R(1+|x|2n)(1+|x|2m)σ(t,x)𝑑x\displaystyle\leq\frac{1}{1+R^{2n}}\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}\left(1+|x|^{2n}\right)\left(1+|x|^{2m}\right)\sigma(t,x)dx
11+R2nsup0tTE~|x|R2(1+|x|2(m+n))σ(t,x)𝑑x\displaystyle\leq\frac{1}{1+R^{2n}}\sup\limits_{0\leq t\leq T}\tilde{E}\int_{|x|\geq R}2\left(1+|x|^{2(m+n)}\right)\sigma(t,x)dx
21+R2nsup0tTE~d(1+|x|2(m+n))σ(t,x)𝑑x\displaystyle\leq\frac{2}{1+R^{2n}}\sup\limits_{0\leq t\leq T}\tilde{E}\int_{\mathbb{R}^{d}}\left(1+|x|^{2(m+n)}\right)\sigma(t,x)dx
2eM(n+m,d,L)T1+R2nd(1+|x|2(m+n))σ0(x)𝑑x.\displaystyle\leq\frac{2e^{M(n+m,d,L)T}}{1+R^{2n}}\int_{\mathbb{R}^{d}}\left(1+|x|^{2(m+n)}\right)\sigma_{0}(x)dx.

5 Approximation of σ(t,x)\sigma(t,x) by the IBV problem in BRB_{R}

With the estimation in Theorem 2, because almost all the density of σ(t,x)\sigma(t,x) is contained in the closed ball BRB_{R} for RR large enough, it is natural to think about approximating σ(t,x)\sigma(t,x) by the solution, σR(t,x)\sigma_{R}(t,x), to the corresponding initial-boundary value (IBV) problem (43) of DMZ equation in the ball BRB_{R}.

It will be rigorously proved in this section that, for RR large enough, σ(t,x)\sigma(t,x) can be approximated well by σR(t,x)\sigma_{R}(t,x) defined in (43), and in particular, the estimation (36) holds in the proof of Theorem 25 in Section 3.

The main result in this section is stated as follows:

Theorem 3.

With Assumptions (A1) to (A4), there exists a constant C>0C>0 which only depends on TT, nn, dd and LL, such that

sup0tTE~BR|σ(t,x)σR(t,x)|𝑑xC1+Rn\sup\limits_{0\leq t\leq T}\tilde{E}\int_{B_{\sqrt{R}}}|\sigma(t,x)-\sigma_{R}(t,x)|dx\leq\frac{C}{1+R^{n}} (67)

holds for large enough RR (for example, R>5R>5), where σR(t,x)\sigma_{R}(t,x) is the solution of the IBV problem (43).

Proof of Theorem 3.

For each R>0R>0, consider the auxiliary function

ϕ(x)=log(1+Rn(1(1|x|2nR2n)2)),xBR,\phi(x)=\log\left(1+R^{n}\left(1-\left(1-\frac{|x|^{2n}}{R^{2n}}\right)^{2}\right)\right),\quad x\in B_{R}, (68)

and

ψ(x)=eϕ(x)eϕ(R),xBR.\psi(x)=e^{-\phi(x)}-e^{-\phi(R)},\quad x\in B_{R}. (69)

Define v(t,x)=σ(t,x)σR(t,x)v(t,x)=\sigma(t,x)-\sigma_{R}(t,x), (t,x)[0,T]×BR(t,x)\in[0,T]\times B_{R}. Then, according to the maximum principle for SPDEs (cf. [18], for example), we have v(t,x)0v(t,x)\geq 0, for all (t,x)[0,T]×BR(t,x)\in[0,T]\times B_{R} and a.s. P~\tilde{P}. Let Φ(t)\Phi(t) be the stochastic process defined by

Φ(t)=BRψ(x)v(t,x)𝑑x.\Phi(t)=\int_{B_{R}}\psi(x)v(t,x)dx. (70)

Since v(t,x)v(t,x) is the solution to the SPDE

dv(t,x)=[12i,j=1d2xixj(aij(x)v(t,x))\displaystyle dv(t,x)=\biggl{[}\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}(x)v(t,x)) i=1dxi(fi(x)v(t,x))]dt\displaystyle-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}(x)v(t,x))\biggr{]}dt (71)
+j=1dhj(x)v(t,x)dYtj,\displaystyle+\sum_{j=1}^{d}h_{j}(x)v(t,x)dY_{t}^{j},

the \mathbb{R}-valued stochastic process Φ(t)\Phi(t) satisfies

dΦ(t)=\displaystyle d\Phi(t)= 12(i,j=1dBRψ(x)2xixj(aij(x)v(t,x))𝑑x)dt\displaystyle\frac{1}{2}\biggl{(}\sum_{i,j=1}^{d}\int_{B_{R}}\psi(x)\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}(x)v(t,x))dx\biggr{)}dt (72)
(i=1dBRψ(x)xi(fi(x)v(t,x))𝑑x)dt\displaystyle-\biggl{(}\sum_{i=1}^{d}\int_{B_{R}}\psi(x)\frac{\partial}{\partial x_{i}}(f_{i}(x)v(t,x))dx\biggr{)}dt
+j=1d(BRhj(x)ψ(x)v(t,x)𝑑x)dYtj.\displaystyle+\sum_{j=1}^{d}\biggl{(}\int_{B_{R}}h_{j}(x)\psi(x)v(t,x)dx\biggr{)}dY_{t}^{j}.

According to the Gauss-Green formula, we have

dΦ(t)=\displaystyle d\Phi(t)= 12(i,j=1dBRaij(x)2ψxixjv(t,x)𝑑x)dt\displaystyle\frac{1}{2}\left(\sum_{i,j=1}^{d}\int_{B_{R}}a^{ij}(x)\frac{\partial^{2}\psi}{\partial x_{i}\partial x_{j}}v(t,x)dx\right)dt
+(i=1dBRψxifi(x)v(t,x)𝑑x)dt\displaystyle+\left(\sum_{i=1}^{d}\int_{B_{R}}\frac{\partial\psi}{\partial x_{i}}f_{i}(x)v(t,x)dx\right)dt
+j=1d(BRhj(x)ψ(x)v(t,x)𝑑x)dYtj\displaystyle+\sum_{j=1}^{d}\left(\int_{B_{R}}h_{j}(x)\psi(x)v(t,x)dx\right)dY_{t}^{j}
+[BR𝔐1(t,x)𝔫𝑑S+BR𝔐2(t,x)𝔫𝑑S]dt,\displaystyle+\biggl{[}-\int_{\partial B_{R}}\vec{\mathfrak{M}}_{1}(t,x)\cdot\vec{\mathfrak{n}}dS+\int_{\partial B_{R}}\vec{\mathfrak{M}}_{2}(t,x)\cdot\vec{\mathfrak{n}}dS\biggr{]}dt,

where, as in the proof of Theorem 2,

𝔐i(t,x)=(𝔐i,1(t,x),,𝔐i,d(t,x)),i=1,2,\vec{\mathfrak{M}}_{i}(t,x)=\left(\mathfrak{M}_{i,1}(t,x),\cdots,\mathfrak{M}_{i,d}(t,x)\right),\quad i=1,2, (73)
𝔐1,j(t,x)\displaystyle\mathfrak{M}_{1,j}(t,x) =12i=1dψxiaij(x)v(t,x),j=1,,d,\displaystyle=\frac{1}{2}\sum_{i=1}^{d}\frac{\partial\psi}{\partial x_{i}}a^{ij}(x)v(t,x),\ j=1,\cdots,d, (74)
𝔐2,i(t,x)\displaystyle\mathfrak{M}_{2,i}(t,x) =ψ(12j=1dxj(aij(x)v(t,x))fi(x)v(t,x)),i=1,,d,\displaystyle=\psi\biggl{(}\frac{1}{2}\sum_{j=1}^{d}\frac{\partial}{\partial x_{j}}(a^{ij}(x)v(t,x))-f_{i}(x)v(t,x)\biggr{)},\ i=1,\cdots,d,

𝔫\vec{\mathfrak{n}} denotes the outward normal vector of the boundary BR\partial B_{R} and dSdS denotes the measure on BR\partial B_{R}.

Notice that ψ|BR0\psi|_{\partial B_{R}}\equiv 0 and

ψxj=eϕ(x)ϕxj.\frac{\partial\psi}{\partial x_{j}}=-e^{-\phi(x)}\frac{\partial\phi}{\partial x_{j}}. (75)

Moreover,

ϕxi=2Rn(1|x|2nR2n)2n|x|2n2xiR2n1+Rn(1(1|x|2nR2n)2),\frac{\partial\phi}{\partial x_{i}}=\frac{2R^{n}\biggl{(}1-\frac{|x|^{2n}}{R^{2n}}\biggr{)}\frac{2n|x|^{2n-2}x_{i}}{R^{2n}}}{1+R^{n}\biggl{(}1-\biggl{(}1-\frac{|x|^{2n}}{R^{2n}}\biggr{)}^{2}\biggr{)}}, (76)

and therefore,

ϕxi|BR=0=ψxi|BR,i=1,,d.\frac{\partial\phi}{\partial x_{i}}\biggr{|}_{\partial B_{R}}=0=\frac{\partial\psi}{\partial x_{i}}\biggr{|}_{\partial B_{R}},\ i=1,\cdots,d. (77)

Hence,

dΦ(t)=\displaystyle d\Phi(t)= 12(BReϕ(x)v(t,x)i,j=1daij(x)(2ϕ(x)xixj+ϕxiϕxj)dx)dt\displaystyle\frac{1}{2}\left(\int_{B_{R}}e^{-\phi(x)}v(t,x)\sum_{i,j=1}^{d}a^{ij}(x)\left(-\frac{\partial^{2}\phi(x)}{\partial x_{i}\partial x_{j}}+\frac{\partial\phi}{\partial x_{i}}\frac{\partial\phi}{\partial x_{j}}\right)dx\right)dt (78)
(BRi=1deϕ(x)v(t,x)fi(x)ϕxidx)dt\displaystyle-\biggl{(}\int_{B_{R}}\sum_{i=1}^{d}e^{-\phi(x)}v(t,x)f_{i}(x)\frac{\partial\phi}{\partial x_{i}}dx\biggr{)}dt
+j=1d(BRhj(x)ψ(x)v(t,x)𝑑x)dYtj\displaystyle+\sum_{j=1}^{d}\left(\int_{B_{R}}h_{j}(x)\psi(x)v(t,x)dx\right)dY_{t}^{j}

Take expectation with respect to the probability measure P~\tilde{P}, and we have

dE~Φ(t)dt=\displaystyle\frac{d\tilde{E}\Phi(t)}{dt}= 12E~(BRψ(x)v(t,x)i,j=1daij(x)(2ϕ(x)xixj+ϕxiϕxj)dx)\displaystyle\frac{1}{2}\tilde{E}\left(\int_{B_{R}}\psi(x)v(t,x)\sum_{i,j=1}^{d}a^{ij}(x)\left(-\frac{\partial^{2}\phi(x)}{\partial x_{i}\partial x_{j}}+\frac{\partial\phi}{\partial x_{i}}\frac{\partial\phi}{\partial x_{j}}\right)dx\right) (79)
E~(BRi=1dψ(x)v(t,x)fi(x)ϕxidx)\displaystyle-\tilde{E}\biggl{(}\int_{B_{R}}\sum_{i=1}^{d}\psi(x)v(t,x)f_{i}(x)\frac{\partial\phi}{\partial x_{i}}dx\biggr{)}
+eϕ(R)12E~(BRv(t,x)i,j=1daij(x)(2ϕ(x)xixj+ϕxiϕxj)dx)\displaystyle+e^{-\phi(R)}\frac{1}{2}\tilde{E}\left(\int_{B_{R}}v(t,x)\sum_{i,j=1}^{d}a^{ij}(x)\left(-\frac{\partial^{2}\phi(x)}{\partial x_{i}\partial x_{j}}+\frac{\partial\phi}{\partial x_{i}}\frac{\partial\phi}{\partial x_{j}}\right)dx\right)
eϕ(R)E~(BRi=1dv(t,x)fi(x)ϕxidx)\displaystyle-e^{-\phi(R)}\tilde{E}\biggl{(}\int_{B_{R}}\sum_{i=1}^{d}v(t,x)f_{i}(x)\frac{\partial\phi}{\partial x_{i}}dx\biggr{)}

For xBRx\in B_{R}, |xi||x|R|x_{i}|\leq|x|\leq R, i=1,,di=1,\cdots,d, and together with the Lipschitz conditions for f(x)f(x),

|ϕxi(x)|4nRn|x|2n2|xi|R2n(1+|x|2nRn)4n,xBR;\left|\frac{\partial\phi}{\partial x_{i}}(x)\right|\leq\frac{4nR^{n}|x|^{2n-2}|x_{i}|}{R^{2n}\biggl{(}1+\frac{|x|^{2n}}{R^{n}}\biggr{)}}\leq 4n,\ \forall\ x\in B_{R}; (80)
|fi(x)ϕxi|\displaystyle\biggl{|}f_{i}(x)\frac{\partial\phi}{\partial x_{i}}\biggr{|} (|f(0)|+L|x|)|ϕxi|\displaystyle\leq(|f(0)|+L|x|)\biggl{|}\frac{\partial\phi}{\partial x_{i}}\biggr{|} (81)
4n|f(0)|+4nLRn|x|2n1|xi|R2n(1+|x|2nRn)\displaystyle\leq 4n|f(0)|+4nL\frac{R^{n}|x|^{2n-1}|x_{i}|}{R^{2n}\biggl{(}1+\frac{|x|^{2n}}{R^{n}}\biggr{)}}
4n(|f(0)|+L),xBR.\displaystyle\leq 4n(|f(0)|+L),\ \forall\ x\in B_{R}.

Also, according to direct computations,

2ϕxixj=\displaystyle\frac{\partial^{2}\phi}{\partial x_{i}\partial x_{j}}= 8n|x|2n4xixj(R2n(n1)(2n1)|x|2n)(R3n+2R2n|x|2n|x|4n)(R3n+2R2n|x|2n|x|4n)2\displaystyle\frac{8n|x|^{2n-4}x_{i}x_{j}(R^{2n}(n-1)-(2n-1)|x|^{2n})(R^{3n}+2R^{2n}|x|^{2n}-|x|^{4n})}{(R^{3n}+2R^{2n}|x|^{2n}-|x|^{4n})^{2}} (82)
16n2|x|2n2xixj(R2n|x|2n2|x|4n2)(R2n|x|2n)(R3n+2R2n|x|2n|x|4n)2\displaystyle-\frac{16n^{2}|x|^{2n-2}x_{i}x_{j}(R^{2n}|x|^{2n-2}-|x|^{4n-2})(R^{2n}-|x|^{2n})}{(R^{3n}+2R^{2n}|x|^{2n}-|x|^{4n})^{2}}
+4nδij|x|2n2(R2n|x|2n)R3n+2R2n|x|2n|x|4n.\displaystyle+\frac{4n\delta_{ij}|x|^{2n-2}(R^{2n}-|x|^{2n})}{R^{3n}+2R^{2n}|x|^{2n}-|x|^{4n}}.

where δij\delta_{ij} is the Kronecker’s symbol. Thus,

|2ϕxixj|8n(3n2)+16n2+4n,xBR.\biggl{|}\frac{\partial^{2}\phi}{\partial x_{i}\partial x_{j}}\biggr{|}\leq 8n(3n-2)+16n^{2}+4n,\ \forall x\in B_{R}. (83)

We would like to remark that the estimation in (83) is quite rough. Each term on the right-hand side of (83) corresponds to one term on the right-hand side of (82), and the purpose is just to show the second-order derivatives are also bounded by a constant independent of RR.

Notice that eϕ(R)=11+Rne^{-\phi(R)}=\frac{1}{1+R^{n}}. Together with the bounded condition for aij(x)a^{ij}(x), we have

dE~Φ(t)dtC1E~Φ(t)+C11+RnE~BRv(t,x)𝑑x\frac{d\tilde{E}\Phi(t)}{dt}\leq C_{1}\tilde{E}\Phi(t)+\frac{C_{1}}{1+R^{n}}\tilde{E}\int_{B_{R}}v(t,x)dx (84)

where C1>0C_{1}>0 is a constant which depends on n,d,Ln,d,L, but does not depend on RR.

According to Theorem 2, the integral

E~BRv(t,x)𝑑xE~BRσ(t,x)𝑑xE~dσ(t,x)𝑑x,\tilde{E}\int_{B_{R}}v(t,x)dx\leq\tilde{E}\int_{B_{R}}\sigma(t,x)dx\leq\tilde{E}\int_{\mathbb{R}^{d}}\sigma(t,x)dx, (85)

which is also bounded by a constant independent of RR, thus,

dE~Φ(t)dtC1E~Φ(t)+C21+Rn.\frac{d\tilde{E}\Phi(t)}{dt}\leq C_{1}\tilde{E}\Phi(t)+\frac{C_{2}}{1+R^{n}}. (86)

where C2>0C_{2}>0 is a constant which depends on T,n,d,LT,n,d,L.

By Gronwall’s inequality,

E~Φ(t)C31+Rn\tilde{E}\Phi(t)\leq\frac{C_{3}}{1+R^{n}} (87)

where C3>0C_{3}>0 is a constant which depends on TT, nn, dd and LL.

On the other hand, for R>5R>5 and for all xBRx\in B_{\sqrt{R}}, i.e., |x|R|x|\leq\sqrt{R},

ϕ(x)[0,log(31Rn)],ψ(x)ψ(R)=Rn3Rn111+Rn1316=16.\phi(x)\in\biggl{[}0,\log\left(3-\frac{1}{R^{n}}\right)\biggr{]},\quad\psi(x)\geq\psi(\sqrt{R})=\frac{R^{n}}{3R^{n}-1}-\frac{1}{1+R^{n}}\geq\frac{1}{3}-\frac{1}{6}=\frac{1}{6}. (88)

Then,

E~Φ(t)=E~BRψ(x)v(t,x)𝑑xE~BRψ(x)v(t,x)𝑑x16E~BRv(t,x)𝑑x.\tilde{E}\Phi(t)=\tilde{E}\int_{B_{R}}\psi(x)v(t,x)dx\geq\tilde{E}\int_{B_{\sqrt{R}}}\psi(x)v(t,x)dx\geq\frac{1}{6}\tilde{E}\int_{B_{\sqrt{R}}}v(t,x)dx. (89)

Combining (87) and (89), we obtain that, for all R1R\gg 1,

E~BR|σ(t,x)σR(t,x)|𝑑x=E~BRv(t,x)𝑑x6C31+Rn.\tilde{E}\int_{B_{\sqrt{R}}}|\sigma(t,x)-\sigma_{R}(t,x)|dx=\tilde{E}\int_{B_{\sqrt{R}}}v(t,x)dx\leq\frac{6C_{3}}{1+R^{n}}. (90)

6 Regularity of the Approximated Function uk(t,x)u_{k}(t,x)

In this section, we will discuss the regularity of uk(t,x)u_{k}(t,x), t[0,T]t\in[0,T], which is the solution of a series of coefficient-frozen equations (20).

The main purpose of this section is to show that under mild conditions, the recursively defined functions uk(t,x)u_{k}(t,x) will not explode in the finite time interval [0,T][0,T], even if the time-discretization step δ0\delta\rightarrow 0, in the sense that the L2L^{2}-norm of uk(τk,x)u_{k}(\tau_{k},x) (k=1,,Kk=1,\cdots,K) is square integrable with respect to the probability measure P~\tilde{P}, and the expectations, E~BR|uk(τk,x)|2𝑑x\tilde{E}\int_{B_{R}}|u_{k}(\tau_{k},x)|^{2}dx, are uniformly bounded for k=1,,Kk=1,\cdots,K.

As shown in the next section, this following theorem is an essential intermediate result for the convergence analysis of this time-discretization scheme.

Theorem 4.

Let {uk(t,x):τk1tτk}k=1K\{u_{k}(t,x):\tau_{k-1}\leq t\leq\tau_{k}\}_{k=1}^{K} be the solution to the IBV problem of the coefficients-frozen equation (20). Then, with Assumptions (A1) to (A4), the L2L^{2}-norm of uk(τk,x)u_{k}(\tau_{k},x) is square-integrable with respect to the probability measure P~\tilde{P}, and we have

E~BR|uk(τk,x)|2𝑑xC<,k=1,,K,\tilde{E}\int_{B_{R}}|u_{k}(\tau_{k},x)|^{2}dx\leq C<\infty,\ \forall\ k=1,\cdots,K, (91)

where C>0C>0 is a constant that depends on dd, TT, RR, LL, but is uniform in k=1,,Kk=1,\cdots,K.

In the proof of Theorem 4, we will consider another exponential transformation given by

σk(t,x)=exp(h(x)Yτk1)uk(t,x),t[τk1,τk],k=1,,K.\sigma_{k}(t,x)=\exp(h^{\top}(x)Y_{\tau_{k-1}})u_{k}(t,x),\ t\in[\tau_{k-1},\tau_{k}],\ k=1,\cdots,K. (92)

Direct computation implies that σk(t,x)\sigma_{k}(t,x) is the solution of

{σk(t,x)t=12i,j=1d2xixj(aij(x)σk(t,x))i=1dxi(fi(x)σk(t,x))12|h(x)|2σk(t,x),(t,x)[τk1,τk]×BR,σk(τk1,x)=exp(h(x)Yτk1)uk1(τk1,x),xBRσk(t,x)=0,(t,x)[τk1,τk]×BR,\left\{\begin{aligned} &\frac{\partial\sigma_{k}(t,x)}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}(x)\sigma_{k}(t,x))-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}(x)\sigma_{k}(t,x))\\ &\qquad-\frac{1}{2}|h(x)|^{2}\sigma_{k}(t,x),\ (t,x)\in[\tau_{k-1},\tau_{k}]\times B_{R},\\ &\sigma_{k}(\tau_{k-1},x)=\exp\left(h^{\top}(x)Y_{\tau_{k-1}}\right)u_{k-1}(\tau_{k-1},x),\ x\in B_{R}\\ &\sigma_{k}(t,x)=0,\ (t,x)\in[\tau_{k-1},\tau_{k}]\times\partial B_{R},\end{aligned}\right. (93)

and recursively, we can rewrite the initial value in (93) by

σk(τk1,x)=exp(h(x)(Yτk1Yτk2))σk1(τk1,x),k=2,,K.\sigma_{k}(\tau_{k-1},x)=\exp\left(h^{\top}(x)(Y_{\tau_{k-1}}-Y_{\tau_{k-2}})\right)\sigma_{k-1}(\tau_{k-1},x),\ k=2,\cdots,K. (94)

Under the reference probability measure P~\tilde{P}, {Yt:0tT}\{Y_{t}:0\leq t\leq T\} is a Brownian motion and

YτkYτk1𝒩(0,δId),k=1,,K,Y_{\tau_{k}}-Y_{\tau_{k-1}}\sim\mathcal{N}(0,\delta I_{d}),\ k=1,\cdots,K, (95)

with Idd×dI_{d}\in\mathbb{R}^{d\times d} the dd-dimensional identity matrix. We would like to study the regularity of σk(t,x)\sigma_{k}(t,x) first, utilizing the Markov property of YY, and then derive the regularity results for uk(t,x)u_{k}(t,x).

For the sake of discussing the regularity of σk(t,x)\sigma_{k}(t,x) in a recursive manner, we need the following lemma which describes the relationship between σk(τk1,x)\sigma_{k}(\tau_{k-1},x) and σk1(τk1,x)\sigma_{k-1}(\tau_{k-1},x) from (94).

Lemma 1.

For k=2,,Kk=2,\cdots,K, let σk(t,x)\sigma_{k}(t,x), t[τk1,τk]t\in[\tau_{k-1},\tau_{k}] be the solution of (93). The end-point values σk(τk1,x)\sigma_{k}(\tau_{k-1},x) and σk1(τk1,x)\sigma_{k-1}(\tau_{k-1},x) satisfy (94). Let us denote by L4(BR)L^{4}(B_{R}) the space of quartic-integrable functions in BRB_{R}. Assume that σk1(τk1,)L4(BR)\sigma_{k-1}(\tau_{k-1},\cdot)\in L^{4}(B_{R}), and the L4L^{4}-norm, σk1(τk1,)L4\|\sigma_{k-1}(\tau_{k-1},\cdot)\|_{L^{4}}, is quartic integrable with respect to P~\tilde{P}, i.e.,

E~BRσk14(τk1,x)𝑑x<\tilde{E}\int_{B_{R}}\sigma_{k-1}^{4}(\tau_{k-1},x)dx<\infty (96)

then σk(τk1,)L4(BR)\sigma_{k}(\tau_{k-1},\cdot)\in L^{4}(B_{R}), its L4L^{4}-norm, σk(τk1,)L4\|\sigma_{k}(\tau_{k-1},\cdot)\|_{L^{4}} is quartic integrable with respect to P~\tilde{P}, and for sufficiently small time-discretization step size δ=τkτk1\delta=\tau_{k}-\tau_{k-1}, we have

E~BRσk4(τk1,x)𝑑x(1+Cδ)E~BRσk14(τk1,x)𝑑x\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k-1},x)dx\leq(1+C\delta)\tilde{E}\int_{B_{R}}\sigma_{k-1}^{4}(\tau_{k-1},x)dx (97)

where CC is a constant that depends on dd and RR.

Proof of Lemma 1.

According to the expression (94) and the definition of σk1\sigma_{k-1} on [τk2,τk1][\tau_{k-2},\tau_{k-1}], because of the Markov property of YY, exp(h(x)(Yτk1Yτk2))\exp\left(h^{\top}(x)(Y_{\tau_{k-1}}-Y_{\tau_{k-2}})\right) is independent of σk1(τk1,x)\sigma_{k-1}(\tau_{k-1},x).

Because the observation function hh is assumed to be smooth enough, and BRB_{R} is a bounded domain in d\mathbb{R}^{d}, there exists a constant MM, which may depend on RR, such that the maximum of the absolute value of hh, together with its partial derivatives up to order mm, is bounded above by MM.

Therefore, by Fubini’s theorem,

E~BRσk4(τk1,x)𝑑x\displaystyle\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k-1},x)dx =E~BRexp(4h(x)(Yτk1Yτk2))σk14(τk1,x)𝑑x\displaystyle=\tilde{E}\int_{B_{R}}\exp\left(4h^{\top}(x)(Y_{\tau_{k-1}}-Y_{\tau_{k-2}})\right)\sigma_{k-1}^{4}(\tau_{k-1},x)dx (98)
=BRE~exp(4h(x)(Yτk1Yτk2))E~σk14(τk1,x)𝑑x.\displaystyle=\int_{B_{R}}\tilde{E}\exp\left(4h^{\top}(x)(Y_{\tau_{k-1}}-Y_{\tau_{k-2}})\right)\tilde{E}\sigma_{k-1}^{4}(\tau_{k-1},x)dx.

Next, let us estimate the expectations of functions of normal random variable ξ:=Yτk1Yτk2\xi:=Y_{\tau_{k-1}}-Y_{\tau_{k-2}} arising in the above expressions, for small time-discretization step δ\delta.

In fact, because ξ𝒩(0,δId)\xi\sim\mathcal{N}(0,\delta I_{d}), we have

E~exp(4h(x)ξ)=j=1dE~e4hj(x)ξj=j=1d(12πδe4hj(x)zez22δ𝑑z)\tilde{E}\exp\left(4h(x)^{\top}\xi\right)=\prod_{j=1}^{d}\tilde{E}e^{4h_{j}(x)\xi_{j}}=\prod_{j=1}^{d}\left(\int_{\mathbb{R}}\frac{1}{\sqrt{2\pi\delta}}e^{4h_{j}(x)z}e^{-\frac{z^{2}}{2\delta}}dz\right) (99)

In the bounded domain BRB_{R},

12πδe4hj(x)zez22δ𝑑z\displaystyle\int_{\mathbb{R}}\frac{1}{\sqrt{2\pi\delta}}e^{4h_{j}(x)z}e^{-\frac{z^{2}}{2\delta}}dz =12πe4hj(x)δzez22𝑑z\displaystyle=\int_{\mathbb{R}}\frac{1}{\sqrt{2\pi}}e^{4h_{j}(x)\sqrt{\delta}z}e^{-\frac{z^{2}}{2}}dz (100)
=e8hj2(x)δ12πe12(z4hj(x)δ)2𝑑z\displaystyle=e^{8h_{j}^{2}(x)\delta}\int_{\mathbb{R}}\frac{1}{\sqrt{2\pi}}e^{-\frac{1}{2}(z-4h_{j}(x)\sqrt{\delta})^{2}}dz
=e8hj2(x)δe8M2δ.\displaystyle=e^{8h_{j}^{2}(x)\delta}\leq e^{8M^{2}\delta}.

Therefore, for δ1\delta\ll 1 (for example δ116M2d\delta\leq\frac{1}{16M^{2}d}),

E~exp(4h(x)ξ)e8dM2δ1+16dM2δ.\tilde{E}\exp\left(4h(x)^{\top}\xi\right)\leq e^{8dM^{2}\delta}\leq 1+16dM^{2}\delta. (101)

Thus,

E~BRσk4(τk1,x)𝑑x(1+16dM2δ)E~BRσk14(τk1,x)𝑑x.\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k-1},x)dx\leq(1+16dM^{2}\delta)\tilde{E}\int_{B_{R}}\sigma_{k-1}^{4}(\tau_{k-1},x)dx. (102)

Now, we are ready to give the proof of Theorem 4.

Proof of Theorem 4.

The idea of this proof is to study the regularity of σk(t,x)\sigma_{k}(t,x), recursively, and then obtain the regularity of uk(t,x)u_{k}(t,x) based on the relationship (94).

In fact, according to the Cauchy-Schwartz inequality,

E~BR|uk(τk,x)|2𝑑x\displaystyle\tilde{E}\int_{B_{R}}|u_{k}(\tau_{k},x)|^{2}dx =E~BRexp(2h(x)Yτk1)σk2(τk,x)𝑑x\displaystyle=\tilde{E}\int_{B_{R}}\exp\left(-2h^{\top}(x)Y_{\tau_{k-1}}\right)\sigma_{k}^{2}(\tau_{k},x)dx
E~[exp(2Mj=1d|Yτk1,j|)BRσk2(τk,x)𝑑x]\displaystyle\leq\tilde{E}\biggl{[}\exp\left(2M\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\right)\int_{B_{R}}\sigma_{k}^{2}(\tau_{k},x)dx\biggr{]}
(E~exp(4Mj=1d|Yτk1,j|))12(E~(BRσk2(τk,x)𝑑x)2)12\displaystyle\leq\biggl{(}\tilde{E}\exp\left(4M\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\right)\biggr{)}^{\frac{1}{2}}\biggl{(}\tilde{E}\biggl{(}\int_{B_{R}}\sigma_{k}^{2}(\tau_{k},x)dx\biggr{)}^{2}\biggr{)}^{\frac{1}{2}}
C1(E~exp(4Mj=1d|Yτk1,j|))12(E~BRσk4(τk,x)𝑑x)12\displaystyle\leq C_{1}\biggl{(}\tilde{E}\exp\left(4M\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\right)\biggr{)}^{\frac{1}{2}}\biggl{(}\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx\biggr{)}^{\frac{1}{2}}

with C1>0C_{1}>0, a constant depending only on RR.

Under the reference probability measure P~\tilde{P}, {Yt:0tT}\{Y_{t}:0\leq t\leq T\} is a standard dd-dimensional Brownian motion, and therefore, the expectation

E~exp(4Mj=1d|Yτk1,j|)\tilde{E}\exp\left(4M\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\right)

is bounded.

Hence, it remains to show that there exists a constant C2>0C_{2}>0, such that,

E~BRσk4(τk,x)𝑑xC2<,\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx\leq C_{2}<\infty, (103)

holds uniformly for k=1,,Kk=1,\cdots,K.

In the time interval [τk1,τk][\tau_{k-1},\tau_{k}], σk(t,x)\sigma_{k}(t,x) is the solution to (93). According to the regularity results of parabolic partial differential equations, we have

BRσk4(τk,x)𝑑xeC4δBRσk4(τk1,x)𝑑x,k=1,,K.\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx\leq e^{C_{4}\delta}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k-1},x)dx,\ \forall\ k=1,\cdots,K. (104)

where C4C_{4} is a constant which depends on the coefficients of the filtering system. The techniques in the proof of (104) is standard, and the proof of a counterpart, in which L2L^{2}-norm (instead of L4L^{4}-norm) is considered, can be found in the textbook [19]. We also provide a detailed proof in the Appendix, for the readers’ convenience and in order to keep this paper self-contained.

Thus, with the result in Lemma 1, there exists C5,C6>0C_{5},C_{6}>0, such that for small enough δ\delta,

E~BRσk4(τk,x)𝑑x\displaystyle\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx eC4δE~BRσk4(τk1,x)𝑑xeC4δ(1+C5δ)E~BRσk14(τk1,x)𝑑x\displaystyle\leq e^{C_{4}\delta}\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k-1},x)dx\leq e^{C_{4}\delta}(1+C_{5}\delta)\tilde{E}\int_{B_{R}}\sigma_{k-1}^{4}(\tau_{k-1},x)dx (105)
(1+C6δ)E~BRσk14(τk1,x)𝑑x.\displaystyle\leq(1+C_{6}\delta)\tilde{E}\int_{B_{R}}\sigma_{k-1}^{4}(\tau_{k-1},x)dx.

Inductively, we have

E~BRσk4(τk,x)𝑑x\displaystyle\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx (1+C6δ)kBRσ04(x)𝑑x\displaystyle\leq(1+C_{6}\delta)^{k}\int_{B_{R}}\sigma_{0}^{4}(x)dx (106)
(1+C6δ)TδBRσ04(x)𝑑xeC6TBRσ04(x)𝑑x.\displaystyle\leq(1+C_{6}\delta)^{\frac{T}{\delta}}\int_{B_{R}}\sigma_{0}^{4}(x)dx\leq e^{C_{6}T}\int_{B_{R}}\sigma_{0}^{4}(x)dx.

Thus, we have proved the boundedness of E~BRσk4(τk,x)𝑑x\tilde{E}\int_{B_{R}}\sigma_{k}^{4}(\tau_{k},x)dx, and also, the result of Theorem 4 holds. ∎

7 Convergence Analysis of the Time Discretization Scheme

This section serves to show that the solution uk(t,x)u_{k}(t,x) of the coefficient-frozen equations (20) can approximate the solution u(t,x)u(t,x) of the original robust DMZ equation (18) well, if the time-discretization step size δ\delta is small enough.

Also, we will show in this section that, after the exponential transformation exp(h(x)Yτk)\exp(h^{\top}(x)Y_{\tau_{k}}), the L1L^{1}-norm of the difference between the unnormalized densities σR(τk,x)\sigma_{R}(\tau_{k},x) (defined by (43)) and u~k+1(τk,x)\tilde{u}_{k+1}(\tau_{k},x) (defined by (22)) still converges to zero, as δ0\delta\rightarrow 0. In particular, the estimation (38) holds in the proof of Theorem 25 in Section 3.

Theorem 5.

Fix R>0R>0. With Assumptions (A1) to (A4), we can use the solution uk(t,x)u_{k}(t,x) of equation (20) to approximate the solution u(t,x)u(t,x) of equation (18). In particular, for every ϵ>0\epsilon>0, there exists a constant δ>0\delta>0, such that

E~BR|σR(τk,x)u~k+1(τk,x)|𝑑x=E~BReh(x)Yτk|u(τk,x)uk(τk,x)|𝑑x<ϵ,\tilde{E}\int_{B_{R}}|\sigma_{R}(\tau_{k},x)-\tilde{u}_{k+1}(\tau_{k},x)|dx=\tilde{E}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}\left|u(\tau_{k},x)-u_{k}(\tau_{k},x)\right|dx<\epsilon, (107)

holds for every k=1,,Kk=1,\cdots,K.

Proof of Theorem 5.

Since ff is globally Lipschitz, hC2(BR)h\in C^{2}(B_{R}), and BRB_{R} is a bounded domain, there exists a constant M0>0M_{0}>0, such that the absolute value of each component in f(x)f(x) and h(x)h(x), as well as there first and second order derivatives, are dominated by MM in the ball BRB_{R}, i.e.,

maxxBR{max1id|fi(x)|,\displaystyle\max\limits_{x\in B_{R}}\biggl{\{}\max\limits_{1\leq i\leq d}|f_{i}(x)|, max1id|hi(x)|,max1i,jd|fixj|,\displaystyle\max\limits_{1\leq i\leq d}|h_{i}(x)|,\max\limits_{1\leq i,j\leq d}\biggl{|}\frac{\partial f_{i}}{\partial x_{j}}\biggr{|}, (108)
max1i,jd|hixj|,max1i,j,kd|2fixjxk|,max1id|2hixjxk|}M0.\displaystyle\max\limits_{1\leq i,j\leq d}\biggl{|}\frac{\partial h_{i}}{\partial x_{j}}\biggr{|},\max\limits_{1\leq i,j,k\leq d}\biggl{|}\frac{\partial^{2}f_{i}}{\partial x_{j}\partial x_{k}}\biggr{|},\max\limits_{1\leq i\leq d}\biggl{|}\frac{\partial^{2}h_{i}}{\partial x_{j}\partial x_{k}}\biggr{|}\biggr{\}}\leq M_{0}.

Let BR,t+={xBR:u(t,x)uk(t,x)0}B_{R,t}^{+}=\{x\in B_{R}:u(t,x)-u_{k}(t,x)\geq 0\}. According to the technical Lemma 4.1 in [2], we have

ddtBR,t+(u(t,x)uk(t,x))𝑑x=BR,t+t(u(t,x)uk(t,x))𝑑x,\frac{d}{dt}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx=\int_{B_{R,t}^{+}}\frac{\partial}{\partial t}(u(t,x)-u_{k}(t,x))dx, (109)

for almost all t[0,T]t\in[0,T].

Then, according to equations (18) and (20) satisfied by u(t,x)u(t,x) and uk(t,x)u_{k}(t,x) in [τk1,τk][\tau_{k-1},\tau_{k}],

ddtBR,t+(u(t,x)uk(t,x))𝑑x\displaystyle\frac{d}{dt}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx (110)
=BR,t+t(u(t,x)uk(t,x))𝑑x\displaystyle=\int_{B_{R,t}^{+}}\frac{\partial}{\partial t}(u(t,x)-u_{k}(t,x))dx
=12BR,t+i,j=1daij(x)2xixj(uuk)dx+BR,t+i=1dFi(τk1,x)xi(uuk)dx\displaystyle=\frac{1}{2}\int_{B_{R,t}^{+}}\sum_{i,j=1}^{d}a^{ij}(x)\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(u-u_{k})dx+\int_{B_{R,t}^{+}}\sum_{i=1}^{d}F_{i}(\tau_{k-1},x)\frac{\partial}{\partial x_{i}}(u-u_{k})dx
+BR,t+J(τk1,x)(u(t,x)uk(t,x))𝑑x\displaystyle\quad+\int_{B_{R,t}^{+}}J(\tau_{k-1},x)(u(t,x)-u_{k}(t,x))dx
+BR,t+i=1d(Fi(t,x)Fi(τk1,x))uxidx+BR,t+(J(t,x)J(τk1,x))u(t,x)𝑑x.\displaystyle\quad+\int_{B_{R,t}^{+}}\sum_{i=1}^{d}(F_{i}(t,x)-F_{i}(\tau_{k-1},x))\frac{\partial u}{\partial x_{i}}dx+\int_{B_{R,t}^{+}}(J(t,x)-J(\tau_{k-1},x))u(t,x)dx.

Because u(t,x)=uk(t,x)0u(t,x)=u_{k}(t,x)\equiv 0 on the boundary BR\partial B_{R}, and BR,t+BR{xBR:u(t,x)uk(t,x)=0}\partial B_{R,t}^{+}\subset\partial B_{R}\cup\{x\in B_{R}:u(t,x)-u_{k}(t,x)=0\}, we have (uuk)|BR,t+=0(u-u_{k})|_{\partial B_{R,t}^{+}}=0 and (uuk)|BR,t+=c(x)𝔫\nabla(u-u_{k})|_{\partial B_{R,t}^{+}}=-c(x)\vec{\mathfrak{n}} with 𝔫\vec{\mathfrak{n}} the outward normal vector of BR,t+B_{R,t}^{+} and c(x)0c(x)\geq 0 on BR,t+\partial B_{R,t}^{+}. Thus, the first three terms on the right-hand side of (110) can be estimated by

12\displaystyle\frac{1}{2} BR,t+i,j=1daij2xixj(uuk)dx+BR,t+i=1dFi(τk1,x)xi(uuk)dx\displaystyle\int_{B_{R,t}^{+}}\sum_{i,j=1}^{d}a^{ij}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(u-u_{k})dx+\int_{B_{R,t}^{+}}\sum_{i=1}^{d}F_{i}(\tau_{k-1},x)\frac{\partial}{\partial x_{i}}(u-u_{k})dx (111)
+BR,t+J(τk1,x)(u(t,x)uk(t,x))𝑑x\displaystyle+\int_{B_{R,t}^{+}}J(\tau_{k-1},x)(u(t,x)-u_{k}(t,x))dx
=12BR,t+i,j=1d2aijxixj(uuk)dxBR,t+i=1dFi(τk1,x)xi(uuk)dx\displaystyle=\frac{1}{2}\int_{B_{R,t}^{+}}\sum_{i,j=1}^{d}\frac{\partial^{2}a^{ij}}{\partial x_{i}\partial x_{j}}(u-u_{k})dx-\int_{B_{R,t}^{+}}\sum_{i=1}^{d}\frac{\partial F_{i}(\tau_{k-1},x)}{\partial x_{i}}(u-u_{k})dx
+BR,t+J(τk1,x)(u(t,x)uk(t,x))𝑑x\displaystyle\quad+\int_{B_{R,t}^{+}}J(\tau_{k-1},x)(u(t,x)-u_{k}(t,x))dx
12BR,t+i,j=1daijxi(uuk)xj(uuk)dS\displaystyle-\frac{1}{2}\int_{\partial B_{R,t}^{+}}\sum_{i,j=1}^{d}a^{ij}\frac{\partial}{\partial x_{i}}(u-u_{k})\frac{\partial}{\partial x_{j}}(u-u_{k})dS
12BR,t+(uuk)i,j=1daijxi𝔫jdS+BR,t+(uuk)i=1dFi(τk1,x)𝔫idS\displaystyle\quad-\frac{1}{2}\int_{\partial B_{R,t}^{+}}(u-u_{k})\sum_{i,j=1}^{d}\frac{\partial a^{ij}}{\partial x_{i}}\vec{\mathfrak{n}}_{j}dS+\int_{\partial B_{R,t}^{+}}(u-u_{k})\sum_{i=1}^{d}F_{i}(\tau_{k-1},x)\vec{\mathfrak{n}}_{i}dS
(d2L2+C(d,L,M0)(1+j=1d|Yτk1,j|)2)BR,t+(uuk)𝑑x,\displaystyle\leq\biggl{(}\frac{d^{2}L}{2}+C(d,L,M_{0})\biggl{(}1+\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\biggr{)}^{2}\biggr{)}\int_{B_{R,t}^{+}}(u-u_{k})dx,
C1(1+j=1d|Yτk1,j|)2BR,t+(uuk)𝑑x\displaystyle\leq C_{1}\biggl{(}1+\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\biggr{)}^{2}\int_{B_{R,t}^{+}}(u-u_{k})dx

where we use the fact that a(x)=g(x)g(x)a(x)=g(x)g(x)^{\top} is positive semi-definite and the definition of Fi(t,x)F_{i}(t,x) and J(t,x)J(t,x) in (17); C(d,L,M0)C(d,L,M_{0}) and C1C_{1} are constants which depend only on d,L,M0d,L,M_{0}; and dSdS denotes the measure on BR,t+\partial B_{R,t}^{+}.

Also, by the definition of Fi(t,x)F_{i}(t,x) and J(t,x)J(t,x) in (17), we have the following estimation of the differences

|F(t,x)F(τk1,x)|C2|YtYτk1|,\displaystyle|F(t,x)-F(\tau_{k-1},x)|\leq C_{2}|Y_{t}-Y_{\tau_{k-1}}|, (112)
|J(t,x)J(τk1,x)|C3(1+j=1d(|Yt,j|+|Yτk1,j|))|YtYτk|,xBR,\displaystyle|J(t,x)-J(\tau_{k-1},x)|\leq C_{3}\biggl{(}1+\sum_{j=1}^{d}(|Y_{t,j}|+|Y_{\tau_{k-1},j}|)\biggr{)}|Y_{t}-Y_{\tau_{k}}|,\ \forall\ x\in B_{R},

where C2C_{2} and C3C_{3} are constants which only depends on d,L.M0d,L.M_{0}.

Hence,

ddtBR,t+\displaystyle\frac{d}{dt}\int_{B_{R,t}^{+}} (u(t,x)uk(t,x))dxC1(1+j=1d|Yτk1,j|)2BR,t+(uuk)𝑑x\displaystyle(u(t,x)-u_{k}(t,x))dx\leq C_{1}\biggl{(}1+\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\biggr{)}^{2}\int_{B_{R,t}^{+}}(u-u_{k})dx (113)
+C2|YtYτk1|BR|u(t,x)|𝑑x\displaystyle+C_{2}|Y_{t}-Y_{\tau_{k-1}}|\int_{B_{R}}|\nabla u(t,x)|dx
+C3(1+j=1d(|Yt,j|+|Yτk1,j|))|YtYτk|BR|u(t,x)|𝑑x\displaystyle+C_{3}\biggl{(}1+\sum_{j=1}^{d}(|Y_{t,j}|+|Y_{\tau_{k-1},j}|)\biggr{)}|Y_{t}-Y_{\tau_{k}}|\int_{B_{R}}|u(t,x)|dx

holds for almost all t[τk1,τk]t\in[\tau_{k-1},\tau_{k}] and almost surely, where C1C_{1}, C2C_{2} and C3C_{3} are constants which depend on the coefficients of the system.

Under the reference probability distribution P~\tilde{P}, the observation process {Yt:0tT}\{Y_{t}:0\leq t\leq T\} is a standard dd-dimensional Brownian motion, and therefore,

j=1d(Yt,jYτk1,j)N(0,d(tτk1)).\displaystyle\sum_{j=1}^{d}(Y_{t,j}-Y_{\tau_{k-1},j})\sim N(0,d(t-\tau_{k-1})). (114)

Let ΩM1={ω:sup0tTj=1d|Yt,j(ω)|M1}\Omega_{M_{1}}=\left\{\omega:\sup\limits_{0\leq t\leq T}\sum_{j=1}^{d}|Y_{t,j}(\omega)|\leq M_{1}\right\} be the event which represents the observation process YtY_{t} is not severely abnormal, and 1A()1_{A}(\cdot) is the indicator function of the set AA.

For a fixed M1>0M_{1}>0, let us first take the expectation with respect to P~\tilde{P} on the event Ω1,M1\Omega_{1,M_{1}} for both sides of (113), and we have

ddtE~\displaystyle\frac{d}{dt}\tilde{E} [1ΩM1BR,t+(u(t,x)uk(t,x))𝑑x]\displaystyle\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]}
\displaystyle\leq C1E~[1ΩM1(1+j=1d|Yτk1,j|)2BR,t+(uuk)𝑑x]\displaystyle\ C_{1}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\biggl{(}1+\sum_{j=1}^{d}|Y_{\tau_{k-1},j}|\biggr{)}^{2}\int_{B_{R,t}^{+}}(u-u_{k})dx\biggr{]}
+C2E~[1ΩM1|YtYτk1|BR|u|𝑑x]\displaystyle+C_{2}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}|Y_{t}-Y_{\tau_{k-1}}|\int_{B_{R}}|\nabla u|dx\biggr{]}
+C3E~[1ΩM1(1+j=1d(|Yt,j|+|Yτk1,j|))|YtYτk|BR|u|𝑑x]\displaystyle+C_{3}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\biggl{(}1+\sum_{j=1}^{d}(|Y_{t,j}|+|Y_{\tau_{k-1},j}|)\biggr{)}|Y_{t}-Y_{\tau_{k}}|\int_{B_{R}}|u|dx\biggr{]}
\displaystyle\leq C1(1+M1)2E~[1ΩM1BR,t+(u(t,x)uk(t,x))𝑑x]\displaystyle\ C_{1}(1+M_{1})^{2}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]}
+C2E~[1ΩM1|YtYτk1|BR|u|𝑑x]\displaystyle+C_{2}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}|Y_{t}-Y_{\tau_{k-1}}|\int_{B_{R}}|\nabla u|dx\biggr{]}
+C3(1+2M1)E~[1ΩM1|YtYτk1|BR|u|𝑑x]\displaystyle+C_{3}(1+2M_{1})\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}|Y_{t}-Y_{\tau_{k-1}}|\int_{B_{R}}|u|dx\biggr{]}
\displaystyle\leq C1(1+M1)2E~[1ΩM1BR,t+(u(t,x)uk(t,x))𝑑x]\displaystyle\ C_{1}(1+M_{1})^{2}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]}
+C2(E~|YtYτk1|2)12(E~[1ΩM1(BR|u|𝑑x)2])12\displaystyle+C_{2}\biggl{(}\tilde{E}|Y_{t}-Y_{\tau_{k-1}}|^{2}\biggr{)}^{\frac{1}{2}}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\left(\int_{B_{R}}|\nabla u|dx\right)^{2}\biggr{]}\right)^{\frac{1}{2}}
+C3(1+2M1)(E~|YtYτk1|2)12(E~[1ΩM1(BR|u|𝑑x)2])12\displaystyle+C_{3}(1+2M_{1})\left(\tilde{E}|Y_{t}-Y_{\tau_{k-1}}|^{2}\right)^{\frac{1}{2}}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\left(\int_{B_{R}}|u|dx\right)^{2}\biggr{]}\right)^{\frac{1}{2}}
=\displaystyle= C1(1+M1)2E~[1ΩM1BR,t+(u(t,x)uk(t,x))𝑑x]\displaystyle\ C_{1}(1+M_{1})^{2}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]}
+C2d12(tτk1)12(E~[1ΩM1(BR|u|𝑑x)2])12\displaystyle+C_{2}d^{\frac{1}{2}}(t-\tau_{k-1})^{\frac{1}{2}}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\left(\int_{B_{R}}|\nabla u|dx\right)^{2}\biggr{]}\right)^{\frac{1}{2}}
+C3(1+2M1)d12(tτk1)12(E~[1ΩM1(BR|u|𝑑x)2])12.\displaystyle+C_{3}(1+2M_{1})d^{\frac{1}{2}}(t-\tau_{k-1})^{\frac{1}{2}}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\left(\int_{B_{R}}|u|dx\right)^{2}\biggr{]}\right)^{\frac{1}{2}}.

Here, the second inequality holds because of the property of the event ΩM1\Omega_{M_{1}}, the third inequality holds according to the Cauchy-Schwartz inequality and the last equality holds because YtY_{t} is a normal distributed random vector.

On the event ΩM1\Omega_{M_{1}}, the observation process {Yt:0tT}\{Y_{t}:0\leq t\leq T\} is bounded. Therefore, according to the regularity results of parabolic partical differential equations (cf. [19], Section 7.1, Theorem 6), the integrals BR|u|𝑑x\int_{B_{R}}|\nabla u|dx and BR|u|𝑑x\int_{B_{R}}|u|dx are also bounded for almost every t[0,T]t\in[0,T], as long as fC1(BR)f\in C^{1}(B_{R}) and hC2(BR)h\in C^{2}(B_{R}). Thus,

ddtE~[1ΩM1\displaystyle\frac{d}{dt}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}} BR,t+(u(t,x)uk(t,x))dx]\displaystyle\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]} (115)
C4E~[1ΩM1BR,t+(u(t,x)uk(t,x))𝑑x]+C5(tτk1)12,\displaystyle\leq C_{4}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{+}}(u(t,x)-u_{k}(t,x))dx\biggr{]}+C_{5}(t-\tau_{k-1})^{\frac{1}{2}},

where C4,C5>0C_{4},C_{5}>0 are constants which depend on d,L,M0,M1,Td,L,M_{0},M_{1},T.

Similarly, we also have the estimation for the integral on the set BR,t={xBR:u(t,x)uk(t,x)0}B_{R,t}^{-}=\{x\in B_{R}:u(t,x)-u_{k}(t,x)\leq 0\}:

ddtE~[1ΩM1\displaystyle\frac{d}{dt}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}} BR,t(u(t,x)uk(t,x))dx]\displaystyle\int_{B_{R,t}^{-}}(u(t,x)-u_{k}(t,x))dx\biggr{]} (116)
C4E~[1ΩM1BR,t(u(t,x)uk(t,x))𝑑x]+C5(tτk1)12,\displaystyle\leq C_{4}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R,t}^{-}}(u(t,x)-u_{k}(t,x))dx\biggr{]}+C_{5}(t-\tau_{k-1})^{\frac{1}{2}},

and thus

ddtE~[1ΩM1\displaystyle\frac{d}{dt}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}} BR|u(t,x)uk(t,x)|dx]\displaystyle\int_{B_{R}}|u(t,x)-u_{k}(t,x)|dx\biggr{]} (117)
C4E~[1ΩM1BR|u(t,x)uk(t,x)|𝑑x]+2C5(tτk1)12.\displaystyle\leq C_{4}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|u(t,x)-u_{k}(t,x)|dx\biggr{]}+2C_{5}(t-\tau_{k-1})^{\frac{1}{2}}.

Therefore,

ddt(eC4(tτk1)\displaystyle\frac{d}{dt}\biggl{(}e^{-C_{4}(t-\tau_{k-1})} E~[1ΩM1BR|u(t,x)uk(t,x)|dx])\displaystyle\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|u(t,x)-u_{k}(t,x)|dx\biggr{]}\biggr{)} (118)
2C5eC4(tτk1)(tτk1)12,\displaystyle\leq 2C_{5}e^{-C_{4}(t-\tau_{k-1})}(t-\tau_{k-1})^{\frac{1}{2}},

and

E~[\displaystyle\tilde{E}\biggl{[} 1ΩM1BR|u(t,x)uk(t,x)|dx]\displaystyle 1_{\Omega_{M_{1}}}\int_{B_{R}}|u(t,x)-u_{k}(t,x)|dx\biggr{]} (119)
eC4(tτk1)(E~[1ΩM1BR|u(τk1,x)uk(τk1,x)|𝑑x]+43C5(tτk1)32).\displaystyle\leq e^{C_{4}(t-\tau_{k-1})}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|u(\tau_{k-1},x)-u_{k}(\tau_{k-1},x)|dx\biggr{]}+\frac{4}{3}C_{5}(t-\tau_{k-1})^{\frac{3}{2}}\right).

Notice that uk(τk1,x)uk1(τk1,x)u_{k}(\tau_{k-1},x)\equiv u_{k-1}(\tau_{k-1},x) by definition. Inductively, we have

E~[\displaystyle\tilde{E}\biggl{[} 1ΩM1BR|u(τk,x)uk(τk,x)|dx]\displaystyle 1_{\Omega_{M_{1}}}\int_{B_{R}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]} (120)
eC4δ(E~[1ΩM1BR|u(τk1,x)uk1(τk1,x)|𝑑x]+43C5δ32)\displaystyle\leq e^{C_{4}\delta}\left(\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|u(\tau_{k-1},x)-u_{k-1}(\tau_{k-1},x)|dx\biggr{]}+\frac{4}{3}C_{5}\delta^{\frac{3}{2}}\right)
eC4kδE~[1ΩM1BR|σ0(x)σ0(x)|𝑑x]+43C5δ32i=1keC4(i1)δ\displaystyle\leq e^{C_{4}k\delta}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|\sigma_{0}(x)-\sigma_{0}(x)|dx\biggr{]}+\frac{4}{3}C_{5}\delta^{\frac{3}{2}}\sum_{i=1}^{k}e^{C_{4}(i-1)\delta}
43C5δ32keC4kδC6δ12.\displaystyle\leq\frac{4}{3}C_{5}\delta^{\frac{3}{2}}ke^{C_{4}k\delta}\leq C_{6}\delta^{\frac{1}{2}}.

where C6C_{6} is a constant which depends on d,L,M0,M1,Td,L,M_{0},M_{1},T.

Also, for the value we are concerned with in (107),

E~[\displaystyle\tilde{E}\biggl{[} 1ΩM1BReh(x)Yτk|u(τk,x)uk(τk,x)|dx]\displaystyle 1_{\Omega_{M_{1}}}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]} (121)
eM0M1E~[1ΩM1BR|u(τk,x)uk(τk,x)|𝑑x]C6eM0M1δ12\displaystyle\leq e^{M_{0}M_{1}}\tilde{E}\biggl{[}1_{\Omega_{M_{1}}}\int_{B_{R}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]}\leq C_{6}e^{M_{0}M_{1}}\delta^{\frac{1}{2}}

On the event ΩM1c={ω:sup0tTj=1d|Yt,j(ω)|>M1}\Omega_{M_{1}}^{c}=\{\omega:\sup_{0\leq t\leq T}\sum_{j=1}^{d}|Y_{t,j}(\omega)|>M_{1}\}, let

Y¯Tsup0tTj=1d|Yt,j|,\overline{Y}_{T}\triangleq\sup_{0\leq t\leq T}\sum_{j=1}^{d}|Y_{t,j}|,

then,

E~\displaystyle\tilde{E} [1ΩM1cBReh(x)Yτk|u(τk,x)uk(τk,x)|𝑑x]\displaystyle\biggl{[}1_{\Omega_{M_{1}}^{c}}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]} (122)
E~[1Ω1,M1cY¯TM1exp(M0j=1d|Yτk,j|)BR|u(τk,x)uk(τk,x)|𝑑x]\displaystyle\leq\tilde{E}\biggl{[}1_{\Omega_{1,M_{1}}^{c}}\frac{\overline{Y}_{T}}{M_{1}}\exp\left(M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\int_{B_{R}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]}
1M1(E~[Y¯T2exp(2M0j=1d|Yτk,j|)])12(E~(BR|u(τk,x)uk(τk,x)|𝑑x)2)12\displaystyle\leq\frac{1}{M_{1}}\left(\tilde{E}\biggl{[}\overline{Y}_{T}^{2}\exp\left(2M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\biggr{]}\right)^{\frac{1}{2}}\left(\tilde{E}\left(\int_{B_{R}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\right)^{2}\right)^{\frac{1}{2}}
C7M1(E~ξ2)12(E~BR|u(τk,x)|2𝑑x+E~BR|uk(τk,x)|2𝑑x)12.\displaystyle\leq\frac{C_{7}}{M_{1}}\left(\tilde{E}\xi^{2}\right)^{\frac{1}{2}}\left(\tilde{E}\int_{B_{R}}|u(\tau_{k},x)|^{2}dx+\tilde{E}\int_{B_{R}}|u_{k}(\tau_{k},x)|^{2}dx\right)^{\frac{1}{2}}.

where C7>0C_{7}>0 is a constant which is related to the volume of the dd-dimensional ball BRB_{R}, and ξ\xi is the random variable given by

ξ=Y¯Texp(M0j=1d|Yτk,j|),\xi=\overline{Y}_{T}\exp\left(M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right), (123)

and

E~ξ2=E~[Y¯T2exp(2M0j=1d|Yτk,j|)](E~Y¯T4)12(E~exp(4M0j=1d|Yτk,j|))12\displaystyle\tilde{E}\xi^{2}=\tilde{E}\biggl{[}\overline{Y}_{T}^{2}\exp\left(2M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\biggr{]}\leq\left(\tilde{E}\overline{Y}_{T}^{4}\right)^{\frac{1}{2}}\left(\tilde{E}\exp\left(4M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\right)^{\frac{1}{2}} (124)

According to the Burkholder-Davis-Gundy inequality (cf. [20], Chapter 3, Theorem 3.28, for example), there exists C8>0C_{8}>0, such that

E~Y¯T4C8E~j=1d|YT,j|43C8dT2.\tilde{E}\overline{Y}_{T}^{4}\leq C_{8}\tilde{E}\sum_{j=1}^{d}|Y_{T,j}|^{4}\leq 3C_{8}dT^{2}. (125)

and also, because Yτk,jY_{\tau_{k},j} are normal random variables, the expectation of

exp(4M0j=1dYτk,j)\exp\left(4M_{0}\sum_{j=1}^{d}Y_{\tau_{k},j}\right)

is bounded.

For the value E~BR|u(τk,x)|2𝑑x\tilde{E}\int_{B_{R}}|u(\tau_{k},x)|^{2}dx, because

u(t,x)=exp(j=1dhj(x)Yt,j)σ(t,x),u(t,x)=\exp\left(-\sum_{j=1}^{d}h_{j}(x)Y_{t,j}\right)\sigma(t,x), (126)

then

E~BR|u(τk,x)|2𝑑x\displaystyle\tilde{E}\int_{B_{R}}|u(\tau_{k},x)|^{2}dx =E~BRexp(2j=1dhj(x)Yτk,j)σ2(τk,x)𝑑x\displaystyle=\tilde{E}\int_{B_{R}}\exp\biggl{(}-2\sum_{j=1}^{d}h_{j}(x)Y_{\tau_{k},j}\biggr{)}\sigma^{2}(\tau_{k},x)dx (127)
E~[exp(2M0j=1d|Yτk,j|)BRσ2(τk,x)𝑑x]\displaystyle\leq\tilde{E}\biggl{[}\exp\left(2M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\int_{B_{R}}\sigma^{2}(\tau_{k},x)dx\biggr{]}
(E~exp(4M0j=1d|Yτk,j|))12(E~(BR|σ(τk,x)|2𝑑x)2)12.\displaystyle\leq\left(\tilde{E}\exp\left(4M_{0}\sum_{j=1}^{d}|Y_{\tau_{k},j}|\right)\right)^{\frac{1}{2}}\left(\tilde{E}\left(\int_{B_{R}}|\sigma(\tau_{k},x)|^{2}dx\right)^{2}\right)^{\frac{1}{2}}.

Notice that σ(t,x)\sigma(t,x) is the solution to the stochastic partial differential equation

dσ(t,x)=σ(t,x)dt+j=1dhjσ(t,x)dYt,j.d\sigma(t,x)=\mathcal{L}^{*}\sigma(t,x)dt+\sum_{j=1}^{d}h_{j}\sigma(t,x)dY_{t,j}. (128)

and the boundedness of

E~(BR|σ(τk,x)|2𝑑x)2\tilde{E}\left(\int_{B_{R}}|\sigma(\tau_{k},x)|^{2}dx\right)^{2} (129)

follows from the regularity theory of stochastic partial differential equation.

In the monograph [21], the authors provided a similar regularity result, and proved that E~BR|σ(τk,x)|2𝑑x\tilde{E}\int_{B_{R}}|\sigma(\tau_{k},x)|^{2}dx is bounded by the initial values. Here in our case, we will prove that there exists C9>0C_{9}>0, such that

E~(BR|σ(τk,x)|2𝑑x)2C9(BR|σ0(x)|2𝑑x)2.\tilde{E}\left(\int_{B_{R}}|\sigma(\tau_{k},x)|^{2}dx\right)^{2}\leq C_{9}\left(\int_{B_{R}}|\sigma_{0}(x)|^{2}dx\right)^{2}. (130)

The detailed proof of (130) can be found in the Appendix.

Therefore, we have

E~BR|u(τk,x)|2𝑑xC10,\tilde{E}\int_{B_{R}}|u(\tau_{k},x)|^{2}dx\leq C_{10}, (131)

where C10>0C_{10}>0 is a constant that does not depend on δ\delta or M1M_{1}.

Furthermore, as we have discussed in the previous section, E~BR|uk(τk,x)|2𝑑x\tilde{E}\int_{B_{R}}|u_{k}(\tau_{k},x)|^{2}dx is also bounded above, and thus, we have

E~[\displaystyle\tilde{E}\biggl{[} 1ΩM1cBReh(x)Yτk|u(τk,x)uk(τk,x)|dx]C11M1,\displaystyle 1_{\Omega_{M_{1}}^{c}}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]}\leq\frac{C_{11}}{M_{1}}, (132)

where C11C_{11} is a constant which does not depend on M1M_{1} or δ\delta.

In summary, for each ϵ>0\epsilon>0, there exists M1>0M_{1}>0, such that

C11M1<ϵ2,\frac{C_{11}}{M_{1}}<\frac{\epsilon}{2}, (133)

and for this particular M1M_{1}, there exists δ>0\delta>0, such that

C6eM0M1δ12<ϵ2,C_{6}e^{M_{0}M_{1}}\delta^{\frac{1}{2}}<\frac{\epsilon}{2}, (134)

Therefore, for every k=1,,Kk=1,\cdots,K,

E~\displaystyle\tilde{E} BReh(x)Yτk|u(τk,x)uk(τk,x)|𝑑x\displaystyle\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx (135)
=E~[1Ω1,M1BReh(x)Yτk|u(τk,x)uk(τk,x)|𝑑x]\displaystyle=\tilde{E}\biggl{[}1_{\Omega_{1,M_{1}}}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]}
+E~[1Ω1,M1cBReh(x)Yτk|u(τk,x)uk(τk,x)|𝑑x]\displaystyle\quad+\tilde{E}\biggl{[}1_{\Omega_{1,M_{1}}^{c}}\int_{B_{R}}e^{h^{\top}(x)Y_{\tau_{k}}}|u(\tau_{k},x)-u_{k}(\tau_{k},x)|dx\biggr{]}
C6eM0M1δ12+C10M1<ϵ.\displaystyle\leq\ C_{6}e^{M_{0}M_{1}}\delta^{\frac{1}{2}}+\frac{C_{10}}{M_{1}}<\epsilon.

8 Conclusion

In this paper, we provide a novel convergence analysis of Yau-Yau algorithm from a probabilistic perspective. With very liberal assumptions only on the coefficients of the filtering systems and the initial distributions (without assumptions on particular paths of observations), we can prove that Yau-Yau algorithm can provide accurate approximations with arbitrary precision to a quite broad class of statistics for the conditional distribution of state process given the observations, which includes the most commonly used conditional mean and covariance matrix. Therefore, the capability of Yau-Yau algorithm to solve very general nonlinear filtering problems is theoretically verified in this paper.

In the process of deriving this probabilistic version of the convergence results, we study the properties of the exact solution, {σ(t,x):0tT}\{\sigma(t,x):0\leq t\leq T\}, to the DMZ equation and the approximated solution {u~k+1(τk,x):1kK}\{\tilde{u}_{k+1}(\tau_{k},x):1\leq k\leq K\}, given by Yau-Yau algorithm, respectively.

For the exact solution σ(t,x)\sigma(t,x) of the DMZ equation, we have shown in Section 4 and Section 5 that most of the density of σ(t,x)\sigma(t,x) will remain in the closed ball BRB_{R}, and σ(t,x)\sigma(t,x) can be approximated well by the corresponding initial-boundary value problem of DMZ equation in BRB_{R}. This result also implies that it is very unlikely for the state process to reach infinity within finite terminal time.

For the approximated solution u~k+1(τk,x)\tilde{u}_{k+1}(\tau_{k},x) given by Yau-Yau algorithm, we have first proved in Section 6 that u~k+1(τk,x)\tilde{u}_{k+1}(\tau_{k},x), which evolves in a recursive manner, will not explode in finite time interval, even if the time-discretization step δ0\delta\rightarrow 0. And then, in Section 7, the convergence of u~k+1(τk,x)\tilde{u}_{k+1}(\tau_{k},x) is proved and the convergence rate is also estimated to be δ\sqrt{\delta}.

It is clear that the properties of exact solutions and approximated solutions, which we have proved in this paper, highly rely on the nice properties of Brownian motion and Gaussian distributions, especially the Markov and light-tail properties. On the one hand, Brownian motion and Gaussian distribution are up to now, among the most commonly used objects in the mathematical modeling of many areas of applications, and can describe most scenarios in practice. On the other hand, for those systems driven by non-Markov or heavy-tailed processes, minimum mean square criteria, together with the conditional expectations (if exist), may not result in a satisfactory estimation of the state process. In this case, the studies of estimations based on other criteria, such as maximum a posteriori (MAP) [22][23][24], will be a promising direction.

Finally, in this paper, we only consider filtering systems and conduct convergence analysis in time interval [0,T][0,T] with a fixed finite terminal time TT. It is also interesting to study the behavior of the DMZ equation and the approximation capability of Yau-Yau algorithm in the case where the terminal time TT\rightarrow\infty, especially for filtering systems with further stable assumptions. We will continue working on how to combine the existing studies on filter stability, such as [25][26], with our techniques developed in this paper, and hopefully, obtain some convergence results of Yau-Yau algorithm for the whole time line (0,)(0,\infty).

Declarations

Funding

This work is supported by National Natural Science Foundation of China (NSFC) grant (12201631) and Tsinghua University Education Foundation fund (042202008).

Conflict of interest/Competing interests

The authors have no competing interests to declare that are relevant to the content of this article.

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Data, Materials and/or Code availability

Not applicable.

Author contribution

All authors contributed to the study conception and design. The first draft of the manuscript was written by Zeju Sun and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Appendix A Regularity Results of Parabolic Partial Differential Equation and Stochastic Evolution Equation

In this appendix, we will provide a detailed proof of the regularity results of the parabolic partial differential equation and the stochastic evolution equation.

For the purpose of deriving (104) and (130), the regularity results is slightly different from standard ones considered in square-integrable functional spaces.

Theorem 6.

Let σ(t,x)\sigma(t,x) be the solution of the following IBV problem:

{σ(t,x)t=12i,j=1d2xixj(aij(x)σ(t,x))i=1dxi(fi(x)σ(t,x))12|h(x)|2σ(t,x),(t,x)[0,T]×BR,σ(0,x)=σ0(x),xBRσ(t,x)=0,(t,x)[0,T]×BR,\left\{\begin{aligned} &\frac{\partial\sigma(t,x)}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}(x)\sigma(t,x))-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}(x)\sigma(t,x))\\ &\qquad-\frac{1}{2}|h(x)|^{2}\sigma(t,x),\ (t,x)\in[0,T]\times B_{R},\\ &\sigma(0,x)=\sigma_{0}(x),\ x\in B_{R}\\ &\sigma(t,x)=0,\ (t,x)\in[0,T]\times\partial B_{R},\end{aligned}\right. (136)

where BR={xd:|x|R}B_{R}=\{x\in\mathbb{R}^{d}:|x|\leq R\} is the ball in d\mathbb{R}^{d} with radius RR; a:dd×da:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d\times d}, f:ddf:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d}, h:ddh:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} are smooth enough functions. Assume that the matrix-valued function a(x)a(x) is uniformly positive definite, i.e., there exists λ>0\lambda>0, such that

i,j=1daij(x)ξiξjλ|ξ|2,xBR,ξd.\sum_{i,j=1}^{d}a^{ij}(x)\xi_{i}\xi_{j}\geq\lambda|\xi|^{2},\ \forall\ x\in B_{R},\ \xi\in\mathbb{R}^{d}. (137)

If the initial value σ0(x)\sigma_{0}(x) is quartic-integrable in BRB_{R}, then there exists a constant C>0C>0, which depends on the coefficients of the system, such that

BRσ4(T,x)𝑑xeCTBRσ04(x)𝑑x.\int_{B_{R}}\sigma^{4}(T,x)dx\leq e^{CT}\int_{B_{R}}\sigma_{0}^{4}(x)dx. (138)
Remark 1.

In fact, Assumption (A2) in the main text will imply the coercivity condition (137). This is because the closed ball BRB_{R} is a compact set of d\mathbb{R}^{d}, and the continuous function λ(x)\lambda(x) in Assumption (A2) will map BRB_{R} to a compact set. Therefore, there exists λ>0\lambda>0, such that λ(x)λ>0\lambda(x)\geq\lambda>0, for all xBRx\in B_{R}.

Proof.

Let us define

f~i(x)=fi(x)j=1daij(x)x,i=1,,d.\tilde{f}_{i}(x)=f_{i}(x)-\sum_{j=1}^{d}\frac{\partial a^{ij}(x)}{\partial x},\ i=1,\cdots,d. (139)

Then the parabolic equation (136) can be written in a divergence form

σ(t,x)t=12i,j=1dxi(aij(x)xjσ(t,x))i=1dxi(f~i(x)σ(t,x))12|h(x)|2σ(t,x).\frac{\partial\sigma(t,x)}{\partial t}=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial}{\partial x_{i}}\left(a^{ij}(x)\frac{\partial}{\partial x_{j}}\sigma(t,x)\right)-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(\tilde{f}_{i}(x)\sigma(t,x))-\frac{1}{2}|h(x)|^{2}\sigma(t,x). (140)

Hence,

ddt\displaystyle\frac{d}{dt} BRσ4(t,x)𝑑x=BR4σ3(t,x)σt𝑑x\displaystyle\int_{B_{R}}\sigma^{4}(t,x)dx=\int_{B_{R}}4\sigma^{3}(t,x)\frac{\partial\sigma}{\partial t}dx (141)
=BR2σ3i,j=1dxi(aijxjσ)dxBR4σ3i=1dxi(f~iσ)dxBR2σ4|h|2𝑑x\displaystyle=\int_{B_{R}}2\sigma^{3}\sum_{i,j=1}^{d}\frac{\partial}{\partial x_{i}}\left(a^{ij}\frac{\partial}{\partial x_{j}}\sigma\right)dx-\int_{B_{R}}4\sigma^{3}\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(\tilde{f}_{i}\sigma)dx-\int_{B_{R}}2\sigma^{4}|h|^{2}dx
=6BRσ2i,j=1daijσxiσxjdx+12BRi=1df~iσ3σxidx2BRσ4|h|2𝑑x\displaystyle=-6\int_{B_{R}}\sigma^{2}\sum_{i,j=1}^{d}a^{ij}\frac{\partial\sigma}{\partial x_{i}}\frac{\partial\sigma}{\partial x_{j}}dx+12\int_{B_{R}}\sum_{i=1}^{d}\tilde{f}_{i}\sigma^{3}\frac{\partial\sigma}{\partial x_{i}}dx-2\int_{B_{R}}\sigma^{4}|h|^{2}dx
6λBRσ2|σ|2𝑑x+12BRi=1df~iσ2λ(λσσxi)dx2BRσ4|h|2𝑑x\displaystyle\leq-6\lambda\int_{B_{R}}\sigma^{2}|\nabla\sigma|^{2}dx+12\int_{B_{R}}\sum_{i=1}^{d}\frac{\tilde{f}_{i}\sigma^{2}}{\sqrt{\lambda}}\cdot\left(\sqrt{\lambda}\sigma\frac{\partial\sigma}{\partial x_{i}}\right)dx-2\int_{B_{R}}\sigma^{4}|h|^{2}dx
6λBRσ2|σ|2𝑑x+12BRi=1d(f~i2σ42λ+λ2σ2|σxi|2)dx2BRσ4|h|2𝑑x\displaystyle\leq-6\lambda\int_{B_{R}}\sigma^{2}|\nabla\sigma|^{2}dx+12\int_{B_{R}}\sum_{i=1}^{d}\biggl{(}\frac{\tilde{f}_{i}^{2}\sigma^{4}}{2\lambda}+\frac{\lambda}{2}\sigma^{2}\biggl{|}\frac{\partial\sigma}{\partial x_{i}}\biggr{|}^{2}\biggr{)}dx-2\int_{B_{R}}\sigma^{4}|h|^{2}dx
BR(6λi=1df~i22|h|2)σ4(t,x)𝑑x.\displaystyle\leq\int_{B_{R}}\biggl{(}\frac{6}{\lambda}\sum_{i=1}^{d}\tilde{f}_{i}^{2}-2|h|^{2}\biggr{)}\sigma^{4}(t,x)dx.

In the bounded domain BRB_{R}, there exists a constant C>0C>0, such that

|6λi=1df~i22|h|2|C.\biggl{|}\frac{6}{\lambda}\sum_{i=1}^{d}\tilde{f}_{i}^{2}-2|h|^{2}\biggr{|}\leq C. (142)

Thus,

ddtBRσ4(t,x)𝑑xCBRσ4(t,x)𝑑x,t[0,T],\frac{d}{dt}\int_{B_{R}}\sigma^{4}(t,x)dx\leq C\int_{B_{R}}\sigma^{4}(t,x)dx,\ t\in[0,T], (143)

and by Gronwall’s inequality, we have

BRσ4(T,x)𝑑xeCTBRσ04(x)𝑑x.\int_{B_{R}}\sigma^{4}(T,x)dx\leq e^{CT}\int_{B_{R}}\sigma_{0}^{4}(x)dx. (144)

Theorem 7.

Consider the IBV problem of stochastic partial differential equation given by

{dσ(t,x)=σ(t,x)dt+j=1dhj(x)σ(t,x)dYt,j,t[0,T]σ(t,x)=0,(t,x)[0,T]×BR,σ(0,x)=σ0(x),xBR.\left\{\begin{aligned} &d\sigma(t,x)=\mathcal{L}^{*}\sigma(t,x)dt+\sum_{j=1}^{d}h_{j}(x)\sigma(t,x)dY_{t,j},\ t\in[0,T]\\ &\sigma(t,x)=0,\ (t,x)\in[0,T]\times\partial B_{R},\\ &\sigma(0,x)=\sigma_{0}(x),\ x\in B_{R}.\end{aligned}\right. (145)

where Y={Yt:0tT}Y=\{Y_{t}:0\leq t\leq T\} is a standard dd-dimensional Brownian motion in the filtered probability space (Ω,,{t}0tT,P)\left(\Omega,\mathcal{F},\{\mathcal{F}_{t}\}_{0\leq t\leq T},P\right); BR={xd:|x|R}B_{R}=\{x\in\mathbb{R}^{d}:|x|\leq R\} is the ball in d\mathbb{R}^{d} with radius RR, and

()=12i,j=1d2xixj(aij(x))i=1dxi(fi(x)).\mathcal{L}^{*}(\star)=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a^{ij}(x)\star)-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(f_{i}(x)\star). (146)

Assume that the coefficients aa, ff, hh are smooth enough and the Assumption (A2) holds for the matrix-valued function a(x)a(x), which implies that a(x)a(x) is uniformly positive definite in BRB_{R}, i.e., there exists λ>0\lambda>0, such that

i,j=1daij(x)ξiξjλ|ξ|2,xBR,ξd.\sum_{i,j=1}^{d}a^{ij}(x)\xi_{i}\xi_{j}\geq\lambda|\xi|^{2},\ \forall\ x\in B_{R},\ \xi\in\mathbb{R}^{d}. (147)

If the initial value σ0(x)\sigma_{0}(x) is square-integrable in BRB_{R}, then there exists a constant C>0C>0, which depends on TT, RR and the coefficients of the system, such that

E(BR|σ(T,x)|2𝑑x)2C(BR|σ0(x)|2𝑑x)2.E\left(\int_{B_{R}}|\sigma(T,x)|^{2}dx\right)^{2}\leq C\left(\int_{B_{R}}|\sigma_{0}(x)|^{2}dx\right)^{2}. (148)
Proof.

Let us define

f~i(x)=fi(x)j=1daij(x)x,i=1,,d.\tilde{f}_{i}(x)=f_{i}(x)-\sum_{j=1}^{d}\frac{\partial a^{ij}(x)}{\partial x},\ i=1,\cdots,d. (149)

Then the stochastic partial differential equation in (145) can be rewritten in divergence form:

dσ(t,x)=12i,j=1dxi(aijσxj)i=1dxi(f~iσ)+j=1dhjσdYt,j.d\sigma(t,x)=\frac{1}{2}\sum_{i,j=1}^{d}\frac{\partial}{\partial x_{i}}\biggl{(}a^{ij}\frac{\partial\sigma}{\partial x_{j}}\biggr{)}-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(\tilde{f}_{i}\sigma)+\sum_{j=1}^{d}h_{j}\sigma dY_{t,j}. (150)

Let

Φ(t)=BRσ2(t,x)𝑑x,t[0,T],\Phi(t)=\int_{B_{R}}\sigma^{2}(t,x)dx,\ t\in[0,T], (151)

then according to Itô’s formula,

dΦ(t)=(BR(2σσ+σ2|h|2)𝑑x)dt+j=1d(BR2hjσ2𝑑x)dYt,j\displaystyle d\Phi(t)=\left(\int_{B_{R}}(2\sigma\mathcal{L}^{*}\sigma+\sigma^{2}|h|^{2})dx\right)dt+\sum_{j=1}^{d}\left(\int_{B_{R}}2h_{j}\sigma^{2}dx\right)dY_{t,j} (152)

and

dΦ2(t)\displaystyle d\Phi^{2}(t) =2(BRσ2𝑑x)(BR(2σσ+σ2|h|2)𝑑x)dt\displaystyle=2\left(\int_{B_{R}}\sigma^{2}dx\right)\left(\int_{B_{R}}(2\sigma\mathcal{L}^{*}\sigma+\sigma^{2}|h|^{2})dx\right)dt (153)
+2Φ(t)j=1d(BR2hjσ2𝑑x)dYt,j+j=1d(BR2hjσ2𝑑x)2dt\displaystyle\quad+2\Phi(t)\sum_{j=1}^{d}\left(\int_{B_{R}}2h_{j}\sigma^{2}dx\right)dY_{t,j}+\sum_{j=1}^{d}\left(\int_{B_{R}}2h_{j}\sigma^{2}dx\right)^{2}dt

After taking expectations, we have

ddtEΦ2(t)\displaystyle\frac{d}{dt}E\Phi^{2}(t) =ddtE(BRσ2(t,x)𝑑x)2\displaystyle=\frac{d}{dt}E\left(\int_{B_{R}}\sigma^{2}(t,x)dx\right)^{2} (154)
=E[2(BRσ2dx)(BR(2σσ+σ2|h|2)dx)\displaystyle=E\biggl{[}2\left(\int_{B_{R}}\sigma^{2}dx\right)\left(\int_{B_{R}}(2\sigma\mathcal{L}^{*}\sigma+\sigma^{2}|h|^{2})dx\right)
+j=1d(BR2hjσ2dx)2]\displaystyle\qquad+\sum_{j=1}^{d}\left(\int_{B_{R}}2h_{j}\sigma^{2}dx\right)^{2}\biggr{]}

Notice that

BR2σσ𝑑x\displaystyle\int_{B_{R}}2\sigma\mathcal{L}^{*}\sigma dx =BRσi,j=1dxi(aijσxj)dxBR2σi=1dxi(f~iσ)dx\displaystyle=\int_{B_{R}}\sigma\sum_{i,j=1}^{d}\frac{\partial}{\partial x_{i}}\biggl{(}a^{ij}\frac{\partial\sigma}{\partial x_{j}}\biggr{)}dx-\int_{B_{R}}2\sigma\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(\tilde{f}_{i}\sigma)dx (155)
=BRi,j=1daijσxiσxjdx+2BRi=1df~iσσxidx\displaystyle=-\int_{B_{R}}\sum_{i,j=1}^{d}a^{ij}\frac{\partial\sigma}{\partial x_{i}}\frac{\partial\sigma}{\partial x_{j}}dx+2\int_{B_{R}}\sum_{i=1}^{d}\tilde{f}_{i}\sigma\frac{\partial\sigma}{\partial x_{i}}dx
λBR|σ|2𝑑x+2BRi=1df~iσλ(λσxi)dx\displaystyle\leq-\lambda\int_{B_{R}}|\nabla\sigma|^{2}dx+2\int_{B_{R}}\sum_{i=1}^{d}\frac{\tilde{f}_{i}\sigma}{\sqrt{\lambda}}\cdot\biggl{(}\sqrt{\lambda}\frac{\partial\sigma}{\partial x_{i}}\biggr{)}dx
λBR|σ|2𝑑x+BR1λi=1df~i2σ2dx+λBR|σ|2𝑑x.\displaystyle\leq-\lambda\int_{B_{R}}|\nabla\sigma|^{2}dx+\int_{B_{R}}\frac{1}{\lambda}\sum_{i=1}^{d}\tilde{f}_{i}^{2}\sigma^{2}dx+\lambda\int_{B_{R}}|\nabla\sigma|^{2}dx.

Hence,

ddtE(BRσ2(t,x)𝑑x)2\displaystyle\frac{d}{dt}E\left(\int_{B_{R}}\sigma^{2}(t,x)dx\right)^{2} 2E[(BRσ2𝑑x)(BR(1λ|f~|2+|h|2)σ2𝑑x)]\displaystyle\leq 2E\biggl{[}\biggl{(}\int_{B_{R}}\sigma^{2}dx\biggr{)}\biggl{(}\int_{B_{R}}\biggl{(}\frac{1}{\lambda}|\tilde{f}|^{2}+|h|^{2}\biggr{)}\sigma^{2}dx\biggr{)}\biggr{]} (156)
+E[j=1d(BR2hjσ2𝑑x)2]\displaystyle\quad+E\biggl{[}\sum_{j=1}^{d}\left(\int_{B_{R}}2h_{j}\sigma^{2}dx\right)^{2}\biggr{]}

In the bounded domain BRB_{R}, there exists M>0M>0, such that

1λ|f~(x)|2+|h(x)|2M,|hj(x)|M,xBR.\frac{1}{\lambda}|\tilde{f}(x)|^{2}+|h(x)|^{2}\leq M,\ |h_{j}(x)|\leq M,\ \forall\ x\in B_{R}. (157)

Thus,

ddtE(BRσ2(t,x)𝑑x)2(2M+4dM2)E(BRσ2(t,x)𝑑x)2\frac{d}{dt}E\left(\int_{B_{R}}\sigma^{2}(t,x)dx\right)^{2}\leq(2M+4dM^{2})E\left(\int_{B_{R}}\sigma^{2}(t,x)dx\right)^{2} (158)

According to Gronwall’s inequality,

E(BRσ2(T,x)𝑑x)2e(2M+4dM2)T(BRσ02(x)𝑑x)2,E\left(\int_{B_{R}}\sigma^{2}(T,x)dx\right)^{2}\leq e^{(2M+4dM^{2})T}\left(\int_{B_{R}}\sigma_{0}^{2}(x)dx\right)^{2}, (159)

which is the desired result. ∎

References

  • \bibcommenthead
  • Yau and Yau [2000] Yau, S.-T., Yau, S.S.-T.: Real time solution of nonlinear filtering problem without memory i. Mathematical research letters 7(6), 671–693 (2000)
  • Yau and Yau [2008] Yau, S.-T., Yau, S.S.-T.: Real time solution of the nonlinear filtering problem without memory ii. SIAM Journal on Control and Optimization 47(1), 163–195 (2008)
  • Candy [2016] Candy, J.V.: Bayesian Signal Processing: Classical, Modern, and Particle Filtering Methods, (2016)
  • Roth et al. [2017] Roth, M., Hendeby, G., Fritsche, C., Gustafsson, F.: The ensemble kalman filter: a signal processing perspective. EURASIP Journal on Advances in Signal Processing 2017, 1–16 (2017)
  • Galanis et al. [2006] Galanis, G., Louka, P., Katsafados, P., Pytharoulis, I., Kallos, G.: Applications of kalman filters based on non-linear functions to numerical weather predictions. Ann. Geophys 24, 2451–2460 (2006)
  • Chen and Yu [2014] Chen, K., Yu, J.: Short-term wind speed prediction using an unscented kalman filter based state-space support vector regression approach. Applied energy 113, 690–705 (2014)
  • Ichard [2015] Ichard, C.: Random media and processes estimation using non-linear filtering techniques: application to ensemble weather forecast and aircraft trajectories. PhD thesis, Université de Toulouse, Université Toulouse III-Paul Sabatier (2015)
  • Sun et al. [2019] Sun, J., Blom, H.A., Ellerbroek, J., Hoekstra, J.M.: Particle filter for aircraft mass estimation and uncertainty modeling. Transportation Research Part C: Emerging Technologies 105, 145–162 (2019)
  • Jazwinski [2007] Jazwinski, A.H.: Stochastic Processes and Filtering Theory. Courier Corporation, (2007)
  • Bain [2009] Bain, A.: Fundamentals of Stochastic Filtering, 1st ed. 2009. edn. Springer, New York, NY (2009)
  • Duncan [1967] Duncan, T.E.: Probability densities for diffusion processes with applications to nonlinear filtering theory and detection theory. PhD thesis, Stanford University, Stanford, California (May 1967)
  • Mortensen [1966] Mortensen, R.E.: Optimal control of continuous-time stochastic systems. Technical report, California Univ. Berkeley Electronics Research Lab (1966)
  • Zakai [1969] Zakai, M.: On the optimal filtering of diffusion processes. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete 11(3), 230–243 (1969)
  • Luo and Yau [2013] Luo, X., Yau, S.S.-T.: Hermite spectral method to 1-d forward kolmogorov equation and its application to nonlinear filtering problems. IEEE Transactions on Automatic Control 58(10), 2495–2507 (2013)
  • Dong et al. [2020] Dong, W., Luo, X., Yau, S.S.-T.: Solving nonlinear filtering problems in real time by legendre galerkin spectral method. IEEE Transactions on Automatic Control 66(4), 1559–1572 (2020)
  • Wang et al. [2019] Wang, Z., Luo, X., Yau, S.S.-T., Zhang, Z.: Proper orthogonal decomposition method to nonlinear filtering problems in medium-high dimension. IEEE Transactions on Automatic Control 65(4), 1613–1624 (2019)
  • Li et al. [2022] Li, S., Wang, Z., Yau, S.S.-T., Zhang, Z.: Solving nonlinear filtering problems using a tensor train decomposition method. IEEE Transactions on Automatic Control (2022)
  • Chekroun et al. [2016] Chekroun, M.D., Park, E., Temam, R.: The stampacchia maximum principle for stochastic partial differential equations and applications. Journal of Differential Equations 260(3), 2926–2972 (2016)
  • Evans [2010] Evans, L.C.: Partial Differential Equations, 2nd ed. edn. American Mathematical Society, (2010)
  • Karatzas [1998] Karatzas, I.: Brownian Motion and Stochastic Calculus, 2nd edn. New York: Springer, (1998)
  • Rozovsky [2018] Rozovsky, B.L.: Stochastic Evolution Systems Linear Theory and Applications to Non-Linear Filtering, 2nd edn. Cham : Springer, (2018)
  • Godsill et al. [2001] Godsill, S., Doucet, A., West, M.: Maximum a posteriori sequence estimation using monte carlo particle filters. Annals of the Institute of Statistical Mathematics 53, 82–96 (2001)
  • Saha et al. [2012] Saha, S., Mandal, P.K., Bagchi, A., Boers, Y., Driessen, J.N.: Particle based smoothed marginal map estimation for general state space models. IEEE transactions on signal processing 61(2), 264–273 (2012)
  • Kang et al. [2023] Kang, J., Salmon, A., Yau, S.S.-T.: Log-concave posterior densities arising in continuous filtering and a maximum a posteriori algorithm. SIAM Journal on Control and Optimization 61(4), 2407–2424 (2023)
  • Atar [1998] Atar, R.: Exponential stability for nonlinear filtering of diffusion processes in a noncompact domain. Annals of Probability 26(4), 1552–1574 (1998)
  • Ocone and Pardoux [1996] Ocone, D., Pardoux, E.: Asymptotic stability of the optimal filter with respect to its initial condition. SIAM Journal on Control and Optimization 34(1), 226–243 (1996)