Systems of Semilinear Parabolic Variational Inequalities with Time-Dependent Convex Obstacles
- 410 Downloads
Abstract
We consider a system of seminlinear parabolic variational inequalities with time-dependent convex obstacles. We prove the existence and uniqueness of its solution. We also provide a stochastic representation of the solution and show that it can be approximated by the penalization method. Our proofs are based upon probabilistic methods from the theory of Markov processes and the theory of backward stochastic differential equations.
Keywords
Semilinear variational inequality Divergence form operator Time-dependent obstacle Penalization method Reflected backward stochastic differential equationMathematics Subject Classification
Primary 35K87 Secondary 60H301 Introduction
The main feature of the paper is that we deal with time-dependent obstacles. In case of single equation, i.e. when \(m=1\), problem (1.1), (1.2) is quite well investigated. For various results on existence, uniqueness and approximation of solutions in case of \(L^2\) data and one or two regular obstacles, i.e. when D has the form \(D(t,x)=\{y\in {\mathbb {R}}:\underline{h}(t,x)\le y\le {\bar{h}}(t,x)\}\) for some regular \(\underline{h} ,{\bar{h}}:E_T\rightarrow {\bar{{\mathbb {R}}}}\) (possibly \(\underline{h}\equiv -\infty \) or \({\bar{h}}\equiv +\infty \)) see the monograph [2, Sects. 2.2, 2.18] and more recent papers [4, 13, 15]. Linear problem of the form (1.1), (1.2) with \(L^2\) data and one irregular barrier is investigated in [20, 26]. For recent results on semilinear problem see [8] (one merely measurable obstacle) and [9] (two measurable obstacles satisfying some separation condition). The problem with two irregular obstacles and \(L^1\) data is investigated in [11].
In case of systems of equations the situation is quite different. To our knowledge, in this case only few partial results exist (see [20, Sect. 1.2] for the existence of solutions of weakly coupled systems and Example 9.3 and Theorem 9.2 in [18, Chap. 2] for the special case \(0\in D(t_2,\cdot )\subset D(t_1,\cdot )\) if \(0\le t_1\le t_2\); see also [14] for existence results concerning a different but related problem). The aim of the present paper is to prove quite general results on existence, uniqueness and approximation of solutions of (1.1), (1.2) in case the data are square integrable and D satisfies some mild regularity assumptions. The case of \(L^1\) data and irregular obstacles is more difficult but certainly deserves further investigation.
- (a)
\(u^i\) are quasi-continuous (with respect to the parabolic capacity determined by \(L_t\)) functions of class \(C([0,T];H)\cap L^2(0,T;H^1_0(E))\) and \(\mu ^i\) are smooth (with respect to the same capacity) measures of finite variation,
- (b)u is a weak solution of the problemsuch that \(u(t,x)\in D(t,x)\) for quasi-every (q.e. for short) \((t,x)\in (0,T]\times E\),$$\begin{aligned} \frac{\partial u}{\partial t}+L_tu=-f_u-\mu ,\quad u(T, \cdot )=\varphi , \quad u(t, \cdot )_{|\partial E}=0,\,\,t \in (0,T), \end{aligned}$$
- (c)for every quasi-continuous function \(h=(h^1,\dots ,h^m)\) such that \(h(t,x)\in D(t,x)\) for q.e \((t,x)\in (0,T]\times E\),$$\begin{aligned} \sum ^m_{i=1}\int ^T_t\int _E(u^i-h^i)\,d\mu ^i\le 0,\quad t\in (0,T). \end{aligned}$$
As in [8, 9, 11], to prove our result we use probabilistic methods. In particular, we rely heavily on the results of our earlier paper [12] devoted to reflected backward stochastic differential equations with time-dependent obstacles and in proofs we use the methods of the theory of Markov processes and probabilistic potential theory. Also note that the first results on multidimensional reflected backward stochastic differential equations were proved in [7] in case D is a fixed convex domain. The results of [7] were generalized in [25] to equations with Wiener–Poisson filtration. For related results with some time-depending domains see the recent paper [22].
2 Preliminaries
For \(x\in {\mathbb {R}}^m\), \(z\in {\mathbb {R}}^{m\times d}\) we set \(|x|^2=\sum ^m_{i=1}|x_i|^2\), \(\Vert z\Vert ^2=\text{ trace }(z^*z)\). By \(\langle \cdot ,\cdot \rangle \) we denote the usual scalar product in \({\mathbb {R}}^m\). Given a Hilbert space \({\mathcal {H}}\) we denote by \([{\mathcal {H}}]^m\) its product equipped with the usual inner product \((u,v)_{[{\mathcal {H}}]^m}=\sum ^m_{i=1}(u^i,v^i)_{{\mathcal {H}}}\) and norm \(\Vert u\Vert _{[{\mathcal {H}}]^m}=((u,u)_{[{\mathcal {H}}]^m})^{1/2}\). We will identify the space \([L^2(0,T;{\mathcal {H}})]^m\) with \(L^2(0,T;[{\mathcal {H}}]^m)\).
Throughout the paper, E is a nonempty bounded connected open subset of \({\mathbb {R}}^d\). \({\bar{E}}\) is the closure of E in \({\mathbb {R}}^d\), \(E^1={\mathbb {R}}\times E\), \(E_T=[0,T]\times E\), \(E_{0,T}=(0,T]\times E\). We set \(H=[L^2(E)]^m\), \(V=[H^1_0(E)]^m\), \(V'=[H^{-1}(E)]^m\), where \(H^{-1}(E)\) is the dual of the Sobolev space \(H^1_0(E)\). For \(u\in V\) we set \(|||\nabla u|||^2_H=\sum ^d_{k=1}\sum ^m_{i=1}\Vert \frac{\partial u^i}{\partial x_k}\Vert ^2_{L^2(E)}\).
The Lebesgue measure on \({\mathbb {R}}^d\) will be denoted by m. By \(m_1\) we denote the Lebesgue measure on \(E^1\).
2.1 Convex Sets and Functions
2.2 Time-Dependent Dirichlet Forms
In the paper by \(\text{ cap }\) we denote the parabolic capacity determined by the form \({\mathcal {E}}\) (for the construction and properties of \(\text{ cap }\) see [23, Sect. 4] or [24, Sect. 6.2]). We will say that some property is satisfied quasi-everywhere (q.e. for short) if it is satisfied except for some Borel subset of \(E^1\) of capacity \(\text{ cap }\) zero. Using \(\text{ cap }\) we define quasi-continuity as in [23]. By [23, Theorem 4.1] each function \(u\in {\mathcal {W}}\) has a quasi-continuous \(m_1\)-version, which we will denote by \({\tilde{u}}\).
Let \(\mu \) be a Borel signed measure on \(E^1\). In what follows \(|\mu |\) stands for the total variation of \(\mu \). By \({\mathcal {M}}_{0,b}(E^1)\) we denote the set of all Borel measures on \(E^1\) such that \(|\mu |\) does not charge sets of zero capacity \(\text{ cap }\) and \(|\mu |(E^1)<\infty \). By \({\mathcal {M}}_{0,b}(E_{0,T})\) we denote the subset of \({\mathcal {M}}_{0,b}(E^1)\) consisting of all measures with support in \(E_{0,T}\).
2.3 Markov Processes
Let \(\mu ,\nu \) be Borel measures on \(E^1\) and E, respectively. In what follows we write \(P_{\mu }(\cdot )=\int _{E^1}P_{s,x}(\cdot )\,d\mu (s,x)\), \(P_{s,\nu }(\cdot )=\int _{E}P_{s,x}(\cdot )\,\nu (dx)\). By \(E_{\mu }\) (resp. \(E_{s,\nu }\)) we denote the expectation with respect to \(P_{\mu }\) (resp. \(P_{s,\nu }\)).
Note that if u is quasi-continuous then it is \({\mathbf {M}}\)-quasi-continuous, i.e. for q.e. \((s,x)\in E^1\), \(P_{s,x}([0,\infty )\ni t\mapsto u({\mathbf {X}}_t)\) is continuous)=1 (see, e.g., [23, p. 298]).
Remark 2.1
- (i)
From (2.6) and the fact that \(p_E(s,x,t,\cdot )\) is strictly positive on E it follows that \(\text{ cap }(\{s\}\times \Gamma )>0\) for any \(s\in (0,T)\) and Borel set \(\Gamma \subset E\) such that \(m(\Gamma )>0\). Hence, if some property holds for q.e. \((s,x)\in (0,T)\times E\) then it holds for m-a.e. \(x\in E\) for every \(s\in (0,T)\).
- (ii)
If \(P_{m_1}(\sigma _{B}<\infty )=0\) then in fact \(P_{s,x}(\sigma _{B}<\infty )=0\) for q.e. \((s,x)\in E^1\). This follows from the fact that the function \((s,x)\mapsto P_{s,x}(\sigma _{B}<\infty )\) is excessive (see remark at the end of [23, p. 294]).
The following proposition is probably well known, but we do not have a reference.
Proposition 2.2
Proof
Let Y be a special \((({\mathcal {F}}_t),P_z)\)-semimartingale on [0, T], i.e. an \((({\mathcal {F}}_t),P_z)\)-semimartingale admitting a (unique) decomposition \(Y_t=Y_0+M_t+B_t\), \(t\in [0,T]\), with \(M_0=B_0=0\), M a local \((({\mathcal {F}}_t),P_z)\)-martingale and B an \(({\mathcal {F}}_t)\)-predictable finite variation process (see, e.g., [27, Sect. III.7]). Recall that the \({\mathcal {H}}^2(P_z)\) norm of Y is defined to be \(\Vert Y\Vert _{{\mathcal {H}}^2(P_z)}=(E_z|Y_0|^2)^{1/2}+(E_z[M]_T)^{1/2} +(E_z|B|_T^2)^{1/2}\), where [M] is the quadratic variation of M and \(|B|_T\) is the variation of B on the interval [0, T]. By \({\mathcal {H}}^2(P_z)\) we denote the space of all special \((({\mathcal {F}}_t),P_z)\)-semimartingales on [0, T] with finite \({\mathcal {H}}^2(P_z)\) norm.
Remark 2.3
- (i)
Let \(z\in E_{0,T}\) and let \(\varphi ,f\) satisfy the assumptions of Proposition 2.2. Then \(M^{[u]}\) of Proposition 2.2 is a martingale under \(P_z\) (see [29, p. 327]), \(E_z|{\tilde{u}}({\mathbf {X}}_0)|^2=|{\tilde{u}}(z)|^2<\infty \), \(E_z[M^{[u],k}_{\cdot \wedge \zeta _{\tau }}]_T<\infty \), \(E_z(\int ^{\zeta _{\tau }}_0|f({\mathbf {X}}_t)|\,dt)^2<\infty \) for q.e. \(z\in E_{0,T}\). Hence \(Y={\tilde{u}}({\mathbf {X}}_{\cdot \wedge \zeta _{\tau }})\) is a semimartingale of class \({\mathcal {H}}^2(P_z)\) for q.e. \(z\in E_{0,T}\).
- (ii)Let \(M=(M^1,\dots ,M^d)\) be the AF of Proposition 2.2. Then M is a martingale under \(P_z\) for q.e. \(z\in E_{0,T}\) (see [24, Sect. 5.1.2]). Setwhere \(\sigma ^{-1}\) is the inverse matrix of \(\sigma \). By Lévy’s theorem and (2.11), for q.e. \(z\in E_{0,T}\) the process \(B=(B^1,\dots ,B^d)\) is under \(P_z\) a d-dimensional standard Brownian motion with respect to \(({\mathcal {F}}_t)_{t\ge 0}\). Finally, note that by (2.10) and (2.13),$$\begin{aligned} B^i_t=\sum ^d_{j=1}\int ^t_0\sigma ^{-1}_{ij}({\mathbf {X}}_{\theta })\,dM^j_{\theta }, \quad t\ge 0,\quad i=1,\dots ,d, \end{aligned}$$(2.13)$$\begin{aligned} M^{[u],k}_t =\sum _{i,j=1}^d\int ^t_0\frac{\partial u^k}{\partial x_i}({\mathbf {X}}_{\theta })\sigma _{ij}({\mathbf {X}}_{\theta })\,dB^j_{\theta },\quad t\ge 0,\quad k=1,\dots ,m. \end{aligned}$$
3 Probabilistic Solutions of the Obstacle Problem
- (A1)
\(\varphi (x)\in D(T,x)\) for \(x\in E\), \(\varphi \in H\),
- (A2)
\(f(\cdot ,\cdot ,0,0)\in L^2(0,T;H)\),
- (A3)\(f:E_T\times {\mathbb {R}}^{m}\times {\mathbb {R}}^{m\times d}\rightarrow {\mathbb {R}}^m\) is a measurable function and there exist \(\alpha ,\beta \ge 0\) such thatfor all \(y_1,y_2\in {\mathbb {R}}^m\) and \(z_1,z_2\in {\mathbb {R}}^{m\times d}\).$$\begin{aligned} |f(t,x,y_1,z_1)-f(t,x,y_2,z_2)| \le \alpha |y_1-y_2|+\beta \Vert z_1-z_2\Vert \end{aligned}$$
- (D1)
The sets D(t, x) are bounded uniformly in \((t,x)\in E_T\) and the mapping \(E_T\ni (t,x)\mapsto D(t,x)\in \mathrm{Conv}\) is continuous.
- (D2)
For some \(\varepsilon >0\) chosen so that Int\(D^{*}(t,x)\ne \emptyset \) for \((t,x)\in E_T\) and some \(f_{*}\in L^{2}(0,T;H)\) and \(\varphi _{*}\in H \) such that \(\varphi _{*}(x)\in D^{*}(T,x)\) for \(x\in E\) there exists a solution \(u_{*}\in {\mathcal {W}}\) of the Cauchy problem (1.5) such that \(u_{*}(t,x)\in D^{*}(t,x)\) for q.e. \((t,x)\in E_{0,T}\).
Remark 3.1
- (D3)There exists a quasi-continuous function \(u_{*}:E_{0,T}\rightarrow {\mathbb {R}}^m\) such that the process \(Y^{*}\) defined as \(Y^{*}_t=u_{*}({\mathbf {X}}_{t\wedge \zeta _{\tau }})\), \(t\in [0,T]\), has the following properties:
- (a)
\(Y^{*}_t\in \text{ Int }D({\mathbf {X}}_{t\wedge \zeta _{\tau }})\) and \(\mathrm{dist}(Y^{*}_{t},\partial D({\mathbf {X}}_{t\wedge \zeta _{\tau }})) \ge \varepsilon \), \(t\in [0,T]\), \(P_{s,x}\)-a.s. for q.e. \((s,x)\in E_{0,T}\),
- (b)
\(Y^{*,i}\), \(i=1,\dots ,m\), is a special semimartingale of class \({\mathcal {H}}^2(P_{s,x})\) for q.e. \((s,x)\in [0,T)\times E\).
- (a)
Condition (D2) is satisfied in the following natural situations.
Example 3.2
Assume that one can find \(r>0\) such that for every \((t,x)\in E_T\), \(B(0,r)\subset D(t,x)\), where B(0, r) denotes the open ball with radius r and center at 0. Then D satisfies (D2). To see this, it suffices to consider the constant function \(u_{*}=(u_{*}^1,\dots ,u_{*}^m)=(0,\dots ,0)\). For instance, the above condition on D is satisfied if \(D(t,x)=\{y\in {\mathbb {R}}^m:\underline{h}^i(t,x)\le y_i\le {\bar{h}}^i(t,x),\,i=1,\dots ,m\}\), where \(\underline{h}^i,{\bar{h}}^i:[0,T]\times {\bar{E}}\rightarrow {\mathbb {R}}\) are continuous functions such that \(\underline{h}^i(t,x)<0<{\bar{h}}^i(t,x)\) for \((t,x)\in [0,T]\times {\bar{E}}\), \(i=1,\dots ,m\).
Example 3.3
Let \(D(t,x)=\{y\in {\mathbb {R}}^m:\underline{h}^i(t,x)\le y_i\le {\bar{h}}^i(t,x),\,i=1,\dots ,m\}\), where \(\underline{h}^i,{\bar{h}}^i:[0,T]\times {\bar{E}}\rightarrow {\mathbb {R}}\) are continuous functions such that \(\underline{h}^i(t,x)<{\bar{h}}^i(t,x)\) for \((t,x)\in [0,T]\times {\bar{E}}\). Assume that there are continuous functions \(\underline{h}^i_0,{\bar{h}}^i_0:[0,T]\times {\bar{E}}\rightarrow {\mathbb {R}}\) such that \(\underline{h}^i(t,x)\le \underline{h}^i_0(t,x)<{\bar{h}}^i_0(t,x)\le {\bar{h}}^i(t,x)\) for \((t,x)\in [0,T]\times {\bar{E}}\), \(\underline{h}^i_0,{\bar{h}}^i_0\in {\mathcal {W}}_{0,T}\) and \((\frac{\partial }{\partial t}+L_t)\underline{h}^i_0,\, (\frac{\partial }{\partial t}+L_t){\bar{h}}^i_0\in L^2(0,T;L^2(E))\), \(i=1,\dots ,m\). Then (D2) is satisfied with \(u^i_{*}=({\bar{h}}^i_0+\underline{h}^i_0)/2\), \(i=1,\dots ,m\). Of course, the last condition for \(\underline{h}^i_0,{\bar{h}}^i_0\) is satisfied if \(\underline{h}^i_0,{\bar{h}}^i_0\in W^{1,2}_{2}((0,T)\times E)\) and \(\frac{\partial a_{ij}}{\partial x_k}\in L^{\infty }((0,T)\times E)\) for \(i,j,k=1,\dots ,d\).
In what follows we are going to show that under assumptions (A1)–(A3), (D1), (D2) there exists a unique solution of the problem (1.1), (1.2). It is convenient to start with probabilistic solutions. Solutions of (1.1), (1.2) in the sense of the definition given in Sect. 1 will be studied in the next section. Note that the definition formulated below is an extension, to the case of systems, of the probabilistic definition adopted in [9, 11] in case of single equation.
Definition 3.4
- (a)
u is quasi-continuous, \(f^i_u\in L^1(E_{0,T})\), \(\mu ^i\in {\mathcal {M}}_{0,b}(E_{0,T})\), \(i=1,\dots ,m\),
- (b)For q.e. \((s,x)\in E_{0,T}\) and \(i=1,\dots ,m\),where \(A^{\mu ^i}\), \(i=1,\dots ,m\), is the continuous additive functional of \({\mathbf {M}}\) associated with \(\mu ^i\) in the sense of (2.7),$$\begin{aligned} u^i({\mathbf {X}}_t)&=\varphi ^i({\mathbf {X}}_{\zeta _{\tau }}) +\int ^{\zeta _{\tau }}_{t\wedge \zeta _{\tau }}f^i_u({\mathbf {X}}_{\theta })\,d\theta +\int ^{\zeta _{\tau }}_{t\wedge \zeta _{\tau }}dA^{\mu ^i}_{\theta } \nonumber \\&\quad -\int ^{\zeta _{\tau }}_{t\wedge \zeta _{\tau }}\sigma \nabla u^i({\mathbf {X}}_{\theta })\,dB_{\theta },\quad t\in [0,T], \quad P_{s,x}\text{-a.s. }, \end{aligned}$$(3.1)
- (c)\(u(t,x)\in D(t,x)\) for q.e. \((t,x)\in E_{0,T}\) and for every quasi-continuous function \(h=(h^1,\dots ,h^m)\) such that \(h(t,x)\in D(t,x)\) for q.e. \((t,x)\in E_{0,T}\) we havefor q.e. \((s,x)\in E_T\).$$\begin{aligned} \int ^{\zeta _{\tau }}_0\langle u({\mathbf {X}}_t)-h({\mathbf {X}}_t),dA^{\mu }_t\rangle \le 0,\quad P_{s,x}\text{-a.s. } \end{aligned}$$(3.2)
Remark 3.5
- (i)
From the fact that u is quasi-continuous and \(u(t,x)\in D(t,x)\) for q.e. \((t,x)\in E_{0,T}\) it follows that \(u({\mathbf {X}}_{t\wedge \zeta _{\tau }})\in D({\mathbf {X}}_{t\wedge \zeta _{\tau }})\), \(t\in [0,T]\), \(P_{s,x}\)-a.s. for q.e. \((s,x)\in E_{0,T}\). To see this, let us set \(B=\{(t,x)\in E_{0,T}:u(t,x)\not \in D(t,x)\}\) and \(\sigma _B=\inf \{t>0:{\mathbf {X}}_t\in B\}\). Since \(\text{ cap }(B)=0\), the set B is \({\mathbf {M}}\)-exceptional, and hence, by 2.1(ii), \(P_{s,x}(\sigma _B<\infty )=0\) for q.e. \((s,x)\in E_{0,T}\). Hence \(P_{s,x}({\mathbf {X}}_{t\wedge \zeta _{\tau }}\in B, t\in (0,T])=0\), which implies that \(u({\mathbf {X}}_{t\wedge \zeta _{\tau }})\in D({\mathbf {X}}_{t\wedge \zeta _{\tau }})\), \(t\in (0,T]\), \(P_{s,x}\)-a.s. for q.e. \((s,x)\in E_{0,T}\). In fact, we can replace (0, T] by [0, T], because \({\mathbf {X}}\) is right-continuous at \(t=0\) and D satisfies (D1).
- (ii)
Conditions (3.1), (3.2) of the above definition say that under \(P_{s,x}\) the pair \((Y_t,Z_t)=(u({\mathbf {X}}_{t\wedge \zeta _{\tau }}),\sigma \nabla u({\mathbf {X}}_{t\wedge \zeta _{\tau }}))\), \(t\in [0,T]\), is a solution of the generalized Markov-type reflected BSDE with final condition \(\varphi \), coefficient f, finite variation process \(A^{\mu }\) and obstacle D.
- (iii)Taking \(t=0\) in (3.1) and then integrating with respect to \(P_{s,x}\) we see that for q.e. \((s,x)\in E_{0,T}\),By this and [10, Proposition 3.4], if \(u\in L^{2}(0,T;H)\) then \(u^i\in C([0,T];L^2(E))\), \(i=1,\dots ,m\).$$\begin{aligned} u^i(s,x)=E_{s,x}\Big (\varphi ^i({\mathbf {X}}_{\zeta _{\tau }}) +\int ^{\zeta _{\tau }}_0f^i_u({\mathbf {X}}_{\theta })\,d\theta +\int ^{\zeta _{\tau }}_0dA^{\mu ^i}_{\theta }\Big ),\quad i=1,\dots ,m. \end{aligned}$$
- (iv)
From continuity of \(A^{\mu ^i}\) and (2.7) one can deduce that \(\mu ^i(\{t\}\times E)=0\) for every \(t\in [0,T]\).
We begin with uniqueness of probabilistic solutions.
Proposition 3.6
Assume (A3), (D1). Then there exists at most one probabilistic solution of OP\((\varphi ,f,D)\).
Proof
Proposition 3.7
- (i)For every \(n\in {\mathbb {N}}\) there exists a unique strong solution \(u_n\in [{\mathcal {W}}]^m\) of the problem$$\begin{aligned} \frac{\partial u_n}{\partial t}+L_tu_n =-f_{u_n}+n(u_n-\Pi _{D(\cdot ,\cdot )}(u_n)),\quad u_n(T)=\varphi . \end{aligned}$$(3.4)
- (ii)There is C depending only on \(\alpha ,\beta ,\Lambda \) and \(\Vert u_{*}\Vert _{[{\mathcal {W}}]^m}\) such that$$\begin{aligned} \sup _{0\le s\le T}\Vert u_n(s)\Vert ^2_H +\int ^T_0|||\nabla u_n(t)|||^2_H\,dt \le C(\Vert \varphi \Vert ^2_H+\Vert f(\cdot ,\cdot ,0,0)\Vert ^2_{L^{2}(0,T;H)}).\nonumber \\ \end{aligned}$$(3.5)
Proof
Theorem 3.8
- (i)
There exists a unique solution \((u,\mu )\) of OP\((\varphi ,f,D)\).
- (ii)\(u^i\in C([0,T];H)\cap L^2(0,T;H^1_0(E))\), \(\mu ^i\in {\mathcal {M}}_{0,b}(E_{0,T})\), \(i=1,\dots ,m\), andwith constant C of Proposition 3.7.$$\begin{aligned} \sup _{0<s\le T}\Vert u(s)\Vert ^2_H +\int ^T_0|||\nabla u(t)|||^2_H\,dt \le C(\Vert \varphi \Vert ^2_H+\Vert f(\cdot ,\cdot ,0,0)\Vert ^2_{L^{2}(0,T;H)})\nonumber \\ \end{aligned}$$(3.7)
- (iii)Let \(u_n\in [{\mathcal {W}}]^m\) be a solution of (3.4). Thenand$$\begin{aligned} \Vert u_n-u\Vert ^2_{L^{2}(0,T;H)}\rightarrow 0, \qquad \int ^T_0|||\nabla (u_n-u)(t)|||^2_H\,dt \rightarrow 0 \end{aligned}$$(3.8)$$\begin{aligned} \mu _n\rightharpoonup \mu \quad \text{ weakly }\,{}^{*} \text{ on } (0,T]\times E. \end{aligned}$$(3.9)
Proof
- (H1)
\(\xi \in D_T\), \(E_{s,x}|\xi |^2<\infty \),
- (H2)
\(E_{s,x}\int _0^T|{\bar{f}}(t,0,0)|^2\,dt<\infty \),
- (H3)\({\bar{f}}:[0,T]\times \Omega \times {\mathbb {R}}^m\times {\mathbb {R}}^{m\times d} \rightarrow {\mathbb {R}}^m\) is measurable with respect to \(Prog\otimes {\mathcal {B}}({\mathbb {R}}^m)\otimes {\mathcal {B}}^{m\times d}\) (Prog denotes the \(\sigma \)-field of all progressive subsets of \([0,T]\times \Omega \)) and there are \(\alpha ,\beta \ge 0\) such that \(P_{s,x}\)-a.s. we havefor all \(y_1,y_2\in {\mathbb {R}}^m\) and \(z_1,z_2\in {\mathbb {R}}^{m\times d}\),$$\begin{aligned} |{\bar{f}}(t,y_1,z_1)-{\bar{f}}(t,y_2,z_2)| \le \alpha |y_1-y_2|+\beta \Vert z_1-z_2\Vert \end{aligned}$$
- (H4)
for each N the mapping \(t\mapsto D_t\cap \{x\in {\mathbb {R}}^d:|x|\le N\}\in \text{ Conv }\) is càdlàg \(P_{s,x}\)-a.s. (with the convention that \(D_T=D_{T-}\)), and there is a semimartingale \(A\in {\mathcal {H}}^2(P_{s,x})\) such that \(A_t\in \text{ Int } D_t\), \(t\in [0,T]\), and \(\inf _{t\le T} \text{ dist }(A_t,\partial D_t)>0\).
It is perhaps worth remarking that in case the operator L is in nondivergence form [for instance, L is the Laplace operator \(\Delta \), or, more generally, \(\frac{\partial a_{ij}}{\partial x_k}\in L^{\infty }((0,T)\times E)\) for \(i,j,k=1,\dots ,d]\), then the process \({\mathbf {X}}\) corresponding to L can be constructed by solving an Itô equation. This allows simplifying some arguments in the proofs of the results presented above, but actually not much. One reason is that we are working in \(L^2\) setting, so even in the case where \(L=\Delta \) the solution of (1.1), (1.2) need not be continuous. On the other hand, since we use a stochastic approach via BSDEs [the basic relation is \(Y=u({\mathbf {X}})\), where Y is the first component of the solution of the corresponding BSDE; see Remark 3.5(ii)], we must know that Y is continuous, and hence that u is quasi-continuous. This in turn requires the introduction of quasi-notions (capacity, quasi-continuity, etc.) and we still have to use some results from the probabilistic potential theory.
4 Variational Solutions
In this section we show that results of Sect. 3 can be translated into results on solutions of (1.1), (1.2) in the sense of the analytical definition formulated below. Solutions in the sense of this definition will be called variational solutions.
Definition 4.1
- (a)
u is quasi-continuous, \(u^i\in C([0,T];H)\cap L^2(0,T;H^1_0(E))\), \(\mu ^i\in {\mathcal {M}}_{0,b}(E_{0,T})\), \(i=1,\dots ,m\),
- (b)for every \(\eta =(\eta ^1,\dots ,\eta ^m)\) such that \(\eta ^i\) is bounded and \(\eta ^i\in {\mathcal {W}}_0\) for \(i=1,\dots ,m\) we havefor all \(t\in [0,T]\), where \(\{a^{(t)}(\cdot ,\cdot ), t\ge 0\}\) is the family of bilinear forms on \(V\times V\) defined as$$\begin{aligned}&(u(t),\eta (t))_H+\int ^T_t\big \langle \frac{\partial \eta }{\partial s}(s),u(s)\big \rangle _{V',V}\,ds +\int ^T_ta^{(s)}(u(s),\eta (s))\,ds\nonumber \\&\qquad =(\varphi ,\eta (T))_H +\int ^T_t(f_u(s),\eta (s))_H\,ds +\int ^T_t\int _E\langle \eta ,d\mu \rangle \end{aligned}$$(4.1)$$\begin{aligned} a^{(t)}(\varphi ,\psi )=\frac{1}{2}\sum ^m_{k=1}\sum ^{d}_{i,j=1}\int _Ea_{ij}(t,x) \frac{\partial \varphi ^k}{\partial x_i}\frac{\partial \psi ^k}{\partial x_j}\,dx. \end{aligned}$$
- (c)\(u(t,x)\in D(t,x)\) for q.e. \((t,x)\in E_{0,T}\), and for every quasi-continuous function \(h=(h^1,\dots ,h^m)\) such that \(h(t,x)\in D(t,x)\) for q.e. \((t,x)\in E_{0,T}\),for all \(t\in (0,T)\).$$\begin{aligned} \int ^T_t\int _E\langle u-h,d\mu \rangle \le 0 \end{aligned}$$(4.2)
Remark 4.2
Remark 4.3
Assume that \(m=1\) and \(D(t,x)=\{y\in {\mathbb {R}}:\underline{h}(t,x)\le y\le {\bar{h}}(t,x)\}\) for some quasi-continuous \(\underline{h},{\bar{h}}:E_T\rightarrow {\mathbb {R}}\) such that \(\underline{h}<{\bar{h}}\). Then (4.2) reduces to the condition \(\int _{E_{0,T}} (u-\underline{h})\,d\mu ^+=\int _{E_{0,T}}({\bar{h}}-u)\,d\mu ^-=0\), where \(\mu ^+\) (resp. \(\mu ^-\)) is the positive (resp. negative) part of the Jordan decomposition of \(\mu \). The case of merely measurable obstacles is discussed in [8, 9, 11].
Proposition 4.4
Assume (A3), (D1). Then there exists at most one variational solution of OP\((\varphi ,f,D)\).
Proof
Theorem 4.5
Under the assumptions of Theorem 3.8 there exists a unique variational solution of OP\((\varphi ,f,D)\).
Proof
Remark 4.6
Notes
Acknowledgments
This research was supported by NCN Grant No. 2012/07/B/ST1/03508.
References
- 1.Aronson, D.G.: Non-negative solutions of linear parabolic equations. Ann. Scuola Norm. Super. Pisa 22, 607–693 (1968)MathSciNetMATHGoogle Scholar
- 2.Bensoussan, A., Lions, J.-L.: Applications of Variational Inequalities in Stochastic Control. North-Holland, Amsterdam (1982)MATHGoogle Scholar
- 3.Blumenthal, R.M., Getoor, R.K.: Markov Processes and Potential Theory. Academic Press, New York (1968)MATHGoogle Scholar
- 4.Chen, Q.: Optimal obstacle control problem for semilinear evolutionary bilateral variational inequalities. J. Math. Anal. Appl. 307, 677–690 (2005)MathSciNetCrossRefMATHGoogle Scholar
- 5.Duvaut, G., Lions, J.-L.: Inequalities in Mechannics and Physics. Springer, Berlin (1976)CrossRefMATHGoogle Scholar
- 6.Fukushima, M., Oshima, Y., Takeda, M.: Dirichlet Forms and Symmetric Markov Processes. Second revised and extended edition. Walter de Gruyter, Berlin (2011)MATHGoogle Scholar
- 7.Gegout-Petit, A., Pardoux, É.: Equations différentielles stochastiques rétrogrades réfléchies dans un convexe. Stoch. Stoch. Rep. 57, 111–128 (1996)CrossRefMATHGoogle Scholar
- 8.Klimsiak, T.: Reflected BSDEs and the obstacle problem for semilinear PDEs in divergence form. Stoch. Process. Appl. 122, 134–169 (2012)MathSciNetCrossRefMATHGoogle Scholar
- 9.Klimsiak, T.: Cauchy problem for semilinear parabolic equation with time-dependent obstacles: a BSDEs approach. Potential Anal. 39, 99–140 (2013)MathSciNetCrossRefMATHGoogle Scholar
- 10.Klimsiak, T.: Semi-Dirichlet forms, Feynman-Kac functionals and the Cauchy problem for semilinear parabolic equations. J. Funct. Anal. 268, 1205–1240 (2015)MathSciNetCrossRefMATHGoogle Scholar
- 11.Klimsiak, T., Rozkosz, A.: Obstacle problem for semilinear parabolic equations with measure data. J. Evol. Equ. 15, 457–491 (2015)MathSciNetCrossRefMATHGoogle Scholar
- 12.Klimsiak, T., Rozkosz, A., Słomiński, L.: Reflected BSDEs in time-dependent convex regions. Stoch. Process. Appl. 125, 571–596 (2015)MathSciNetCrossRefMATHGoogle Scholar
- 13.Kubo, M.: Variational inequalities with time-dependent constraints in \(L^p\). Nonlinear Anal. 73, 390–398 (2010)MathSciNetCrossRefMATHGoogle Scholar
- 14.Kubo, M., Shirakawa, K., Yamazaki, N.: Variational inequalities for a system of elliptic-parabolic equations. J. Math. Anal. Appl. 387, 490–511 (2012)MathSciNetCrossRefMATHGoogle Scholar
- 15.Kubo, M., Yamazaki, N.: Periodic solutions of elliptic-parabolic variational inequalities with time-dependent constraints. J. Evol. Equ. 6, 71–93 (2006)MathSciNetCrossRefMATHGoogle Scholar
- 16.Ladyženskaja, O.A., Solonnikov, V.A., Ural’ceva, N.N.: Linear and Quasi-linear Equations of Parabolic Type. Translations of Mathematical Monographs, vol. 23. American Mathematical Society, Providence, RI (1968)Google Scholar
- 17.Lions, J.-L.: Contrôle optimal de systèmes gouvernés par des équations aux dérivées partielles. Dunod, Paris (1968)MATHGoogle Scholar
- 18.Lions, J.-L.: Quelques méthodes de résolution des problèmes aux limites non linéares. Dunod, Paris (1969)MATHGoogle Scholar
- 19.Menaldi, J.L.: Stochastic variational inequality for reflected diffusion. Indiana Univ. Math. J. 32, 733–744 (1983)MathSciNetCrossRefMATHGoogle Scholar
- 20.Mignot, F., Puel, J.P.: Inéquations d’Evolution Paraboliques avec Convexes dépendant du Temps. Applications aux Inéquations Quasi-Variationnelles d’Evolution. Arch. Ration. Mech. Anal. 64, 59–91 (1997)CrossRefMATHGoogle Scholar
- 21.Monteiro Marques, M.: Differential Inclusions in Nonsmooth Mechanical Problems. Shocks and Dry Friction. Birkhäuser, Basel (1993)CrossRefMATHGoogle Scholar
- 22.Nyström, K., Olofsson, M.: Reflected BSDE of Wiener–Poisson type in time-dependent domains. Stoch. Models 32, 275–300 (2016)MathSciNetCrossRefMATHGoogle Scholar
- 23.Oshima, Y.: Time-dependent Dirichlet forms and related stochastic calculus. Infin. Dimens. Anal. Quant. Probab. Relat. Top. 7, 281–316 (2004)MathSciNetCrossRefMATHGoogle Scholar
- 24.Oshima, Y.: Semi-Dirichlet Forms and Markov Processes. Walter de Gruyter, Berlin (2013)CrossRefMATHGoogle Scholar
- 25.Ouknine, Y.: Reflected backward stochastic differential equations with jumps. Stoch. Stoch. Rep. 65, 111–125 (1998)MathSciNetCrossRefMATHGoogle Scholar
- 26.Pierre, M.: Problemes d’Evolution avec Contraintes Unilatérales et Potentiel Paraboliques. Commun. Partial Differ. Equ. 4, 1149–1197 (1979)CrossRefMATHGoogle Scholar
- 27.Protter, P.: Stochastic Integration and Differential Equations, 2nd edn. Springer, Berlin (2004)MATHGoogle Scholar
- 28.Stannat, S.: The theory of generalized Dirichlet forms and its applications in analysis and stochastics. Mem. Am. Math. Soc., vol. 142(678) (1999)Google Scholar
- 29.Trutnau, G.: Stochastic calculus of generalized Dirichlet forms and applications to stochastic differential equations in infinite dimensions. Osaka J. Math. 37, 315–343 (2000)MathSciNetMATHGoogle Scholar
Copyright information
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.