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

Path-dependent Hamilton-Jacobi equations
with super-quadratic growth in the gradient
and the vanishing viscosity method

Erhan Bayraktar Department of Mathematics, University of Michigan, Ann Arbor, MI 48109, United States [email protected]  and  Christian Keller Department of Mathematics, University of Central Florida, Orlando, FL 32816, United States [email protected]
Abstract.

The non-exponential Schilder-type theorem in Backhoff-Veraguas, Lacker and Tangpi [Ann. Appl. Probab., 30 (2020), pp. 1321–1367] is expressed as a convergence result for path-dependent partial differential equations with appropriate notions of generalized solutions. This entails a non-Markovian counterpart to the vanishing viscosity method.

We show uniqueness of maximal subsolutions for path-dependent viscous Hamilton-Jacobi equations related to convex super-quadratic backward stochastic differential equations.

We establish well-posedness for the Hamilton-Jacobi-Bellman equation associated to a Bolza problem of the calculus of variations with path-dependent terminal cost. In particular, uniqueness among lower semi-continuous solutions holds and state constraints are admitted.

Key words and phrases:
Large deviations, vanishing viscosity method, path-dependent partial differential equations, minimax solutions, Dini solutions, calculus of variations, optimal control, state constraints, backward stochastic differential equations, nonsmooth analysis
2010 Mathematics Subject Classification:
60H30; 60F10, 35F21, 35K10, 49J22
The research of the first author was supported in part by NSF-grant DMS-2106556. The research of the second author was supported in part by NSF-grant DMS-2106077.

1. Introduction

Backhoff-Veraguas, Lacker and Tangpi [BLT] derived a non-exponential Schilder-type theorem, which they used to obtain new limit theorems for backward stochastic differential equations (BSDEs) and, in the Markovian case, for the corresponding partial differential equations (PDEs). They posed the question whether it is possible to have a corresponding PDE result in the non-Markovian case as well. Purpose of our work is to provide an answer to this question.

We establish well-posedness for (second order) path-dependent viscous Hamilton-Jacobi equations and for (first-order) path-dependent Hamilton-Jacobi-Bellman (HJB) equations with possibly super-quadratic growth in the gradient. Together with a modification of the Schilder type theorem in [BLT], we obtain a non-Markovian vanishing viscosity result for path-dependent PDEs and thereby address the mentioned open problem in [BLT].

The notions of solutions for our path-dependent PDEs are in the spirit of contingent solutions for PDEs (see, e.g., [Frankowska89AMO]), also known as Dini solutions (see, e.g., [BardiCapuzzoDolcetta] and [Vinter]) or minimax solutions (see, e.g., [SubbotinBook]).

In the context of contingent or Dini solutions for first-order standard PDEs related to Bolza problems, [DM-F00] and also [PQ02] are very close to our approach. More recent works in this direction are [Misztela14], [BernisBettiol18COCV] and [BernisBettiol20chapter]. Regarding the possible use of viscosity solution techniques, we refer the reader to the remarks on p. 1202 in [BettiolVinter17SICON]. In particular, fast growth in the gradient, discontinuity of the Lagrangian, and extended real-valued lower semi-continuous terminal data (to allow right-end point constraints in optimal control problems) cause non-trivial issues. For example, solutions of HJB equations can be expected then to be only lower semi-continuous.

In the second-order case, the only works we are aware of that use contingent-type solutions are [SubbotinaEtAl85] and [Subbotina06], where Isaacs equations corresponding to Markovian stochastic differential games with drift control and bounded control spaces are studied. However, similar constructions (stochastic or Gaussian derivatives) are also used in [Haussmann92MOR] and [Haussmann94SICON].

In the context of first-order path-dependent PDEs, [KaiseEtAl18PPDE] is most relevant. It deals with a calculus of variations problem involving a path-dependent terminal cost and the related path-dependent HJB equation. The setting is very close to our problem (DOC) below. The main difference is that in [KaiseEtAl18PPDE] the terminal cost is required to be Lipschitz continuous, which leads to Lipschitz continuity of the corresponding value function and also makes it possible in [KaiseEtAl18PPDE] to develop a viscosity solution theory. In our work, we require only continuity resp. lower semi-continuity of the terminal cost, which is one of the reasons why we establish a Dini resp. minimax solution theory. For the current state of the art for first-order path-dependent PDEs and for further relevant works, see [GLP21AMO] and the references therein.

In the literature [EKTZ11, K14, ETZ_I, ETZ_II, PhamZhang14SICON, EkrenZhang16PUQR, RTZ14overview, Ren16AAP, Ren17Stoch, Ekren17SPA, RTZ17, cosso18, RTZ20SICON, RR20SIMA] on viscosity solutions of second-order path-dependent PDEs, the Hamiltonian is required to grow at most linearly in the gradient (the same condition is also needed in [CossoRusso19Osaka], where a notion of strong-viscosity solutions is used). Overcoming this restriction for any notion of generalized solutions has been a longstanding open problem. By proving wellposedness of maximal (Dini) subsolutions for a class of second-order path-dependent PDEs with quadratic and even super-quadratic growth in the gradient, we establish first results related to this problem.

Non-Markovian large deviation problems and their connections to path-dependent PDEs are also studied in [MRTZ16]. In contrast to our work, in [MRTZ16] only the (limiting) rate function is characterized as a solution of a (first-order) path-dependent PDE. Moreover, the terminal condition is required to be Lipschitz continuous whereas we need only continuity.

2. Setup

2.1. Notation

Let Ω=C([0,T],d)\Omega=C([0,T],{\mathbb{R}}^{d}). The canonical process on Ω\Omega is denoted by XX, i.e., X(t,ω)=ω(t)X(t,\omega)=\omega(t) for each (t,ω)[0,T]×Ω(t,\omega)\in[0,T]\times\Omega. Let 𝔽={t}t[0,T]{\mathbb{F}}=\{{\mathcal{F}}_{t}\}_{t\in[0,T]} be the (raw) filtration generated by XX. Given a probability measure PP on (Ω,T)(\Omega,{\mathcal{F}}_{T}), denote by 𝔽P={tP}t[0,T]{\mathbb{F}}^{P}=\{{\mathcal{F}}_{t}^{P}\}_{t\in[0,T]} the PP-completion of the right-limit of 𝔽{\mathbb{F}}.

We equip Ω\Omega with the supremum norm \lVert{\cdot}\rVert_{\infty} and [0,T]×Ω[0,T]\times\Omega with the pseudo-metric 𝐝\mathbf{d}_{\infty} defined by

𝐝((t1,ω1),(t2,ω2)):=|t1t2|+sups[0,T]|ω1(st1)ω2(st2)|.\displaystyle\mathbf{d}_{\infty}((t_{1},\omega_{1}),(t_{2},\omega_{2})):=\left|{t_{1}-t_{2}}\right|+\sup\nolimits_{s\in[0,T]}\left|{\omega_{1}(s\wedge t_{1})-\omega_{2}(s\wedge t_{2})}\right|.

Continuity and semi-continuity of functions defined on Ω\Omega (resp. [0,T]×Ω[0,T]\times\Omega) are to be understood with respect to \lVert{\cdot}\rVert_{\infty} (resp. 𝐝\mathbf{d}_{\infty}). Note that semi-continuous functions on [0,T]×Ω[0,T]\times\Omega are 𝔽{\mathbb{F}}-progressive. From now on, we write l.s.c. (resp. u.s.c.) instead of lower semi-continuous (resp. upper semi-continuous).

With slight abuse of notation, we also use the notation \lVert{\cdot}\rVert_{\infty} to express the sup-norm for functions belonging to other function spaces.

We often identify vectors with constant functions, e.g., given a map h:Ωh:\Omega\to{\mathbb{R}}, a vector zdz\in{\mathbb{R}}^{d}, and a path ωΩ\omega\in\Omega, we write h(ω+z)h(\omega+z) instead of h(ω+z 1[0,T])h(\omega+z\,{\mathbf{1}}_{[0,T]}).

Given (t0,x0,n)[0,T]×Ω×(t_{0},x_{0},n)\in[0,T]\times\Omega\times{\mathbb{N}}, denote by Pt0,x0,nP_{t_{0},x_{0},n} be the unique probability measure on (Ω,T)(\Omega,{\mathcal{F}}_{T}) such that nX|[t0,T]\sqrt{n}X|_{[t_{0},T]} is a dd-dimensional standard (Pt0,x0,n,𝔽)(P_{t_{0},x_{0},n},{\mathbb{F}})-Wiener process starting at x0(t0)x_{0}(t_{0}) and that Pt0,x0,n(X|[0,t0]=x0|[0,t0])=1P_{t_{0},x_{0},n}(X|_{[0,t_{0}]}=x_{0}|_{{\color[rgb]{0,0,0}[0,t_{0}]}})=1. We write 𝔼t0,x0,n{{\mathbb{E}}}_{t_{0},x_{0},n} for the corresponding expected value. Moreover, Pt0,x0:=Pt0,x0,1P_{t_{0},x_{0}}:=P_{t_{0},x_{0},1} and 𝔽t0,x0,n:=𝔽Pt0,x0,n{\mathbb{F}}^{t_{0},x_{0},n}:={\mathbb{F}}^{P_{t_{0},x_{0},n}}.

As space of controls, the set b\mathcal{L}_{b} of all bounded 𝔽{\mathbb{F}}-progressive processes from [0,T]×Ω[0,T]\times\Omega to d{\mathbb{R}}^{d} is used (whereas in [BLT] the controls are 𝔽P0,0{\mathbb{F}}^{P_{0,0}}-progressive).

We denote by dom\mathrm{dom} the effective domain of an extended real-valued function.

2.2. Data

Let h:Ω{}h:\Omega\to{\mathbb{R}}\cup\{\infty\} and :[0,T]×d{}\ell:[0,T]\times{\mathbb{R}}^{d}\to{\mathbb{R}}\cup\{\infty\} be measurable functions. We use the following hypotheses for \ell.

(H1) The function =(t,a)\ell=\ell(t,a) satisfies the Tonelli-Nagumo condition, i.e., there is a function ϕ:[0,)\phi:[0,\infty)\to{\mathbb{R}} bounded from below with ϕ(r)/r\phi(r)/r\to\infty as rr\to\infty such that (t,a)ϕ(|a|)\ell(t,a)\geq\phi(\left|{a}\right|) on [0,T]×d[0,T]\times{\mathbb{R}}^{d}. Moreover, (t,)\ell(t,\cdot) is l.s.c., proper, and convex for every t[0,T]t\in[0,T].

(H2) 0T(t,0)𝑑t<\int_{0}^{T}\ell(t,0)\,dt<\infty.

These hypotheses are nearly identical with the corresponding condition (TI) for the Lagrangian in [BLT] (where it is denoted by gg). In some of our main results, we will, in addition to (H1) and (H2), also assume that \ell is continuous and finite-valued. In those cases, (TI) is satisfied as pointed out in [BLT].

2.3. The optimal control problems and HJB equations

Let nn\in{\mathbb{N}}. The value for our stochastic optimal control problem (SOCn)(\text{SOC}_{n}) with initial data (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega is given by

vn(t0,x0):=infabt0𝔼t0,x0,n[t0T(t,a(t))𝑑t+h(X+Aa)],\displaystyle v_{n}(t_{0},x_{0}):=\inf_{a\in\mathcal{L}_{b}^{t_{0}}}{{\mathbb{E}}}_{t_{0},x_{0},n}\left[\int_{t_{0}}^{T}\ell(t,a(t))\,dt+h(X+A^{a})\right],

where bt0={ab:a|[0,t0)=0}\mathcal{L}_{b}^{t_{0}}=\{a\in\mathcal{L}_{b}:a|_{[0,t_{0})}=0\} and AaA^{a} is a continuous process on [0,T]×Ω[0,T]\times\Omega defined by Aa(t):=0ta(s)𝑑sA^{a}(t):=\int_{0}^{t}a(s)\,ds. The terminal value problem involving the corresponding HJB equation is

(TVP nn) (t+12nxx)u(t,x)+infad[axu(t,x)+(t,a)]=0in [0,T)×Ω,u(T,x)=h(x)on Ω.\begin{split}\left(\partial_{t}+\frac{1}{2n}\partial_{xx}\right)u(t,x)+\inf_{a\in{\mathbb{R}}^{d}}\left[a\cdot\partial_{x}u(t,x)+\ell(t,a)\right]&=0\quad\text{in $[0,T)\times\Omega$},\\ u(T,x)&=h(x)\quad\text{on $\Omega$.}\end{split}
Remark 2.1.

If (H1) and (H2) hold and hh is bounded, then supn1vn<\sup_{n\geq 1}\lVert{v_{n}}\rVert_{\infty}<\infty.

The value for our deterministic optimal control problem (DOC)(\text{DOC}) with initial data (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega is given by

v0(t0,x0):=infx𝒳1,1(t0,x0)[t0T(t,x(t))𝑑t+h(x)],\displaystyle v_{0}(t_{0},x_{0}):=\inf_{x\in\mathcal{X}^{1,1}(t_{0},x_{0})}\left[\int_{t_{0}}^{T}\ell(t,x^{\prime}(t))\,dt+h(x)\right],

where

𝒳1,1(t0,x0):={xΩ:x|[0,t0]=x0|[0,t0] and x|(t0,T)W1,1(t0,T;d)}.\displaystyle\mathcal{X}^{1,1}(t_{0},x_{0}):=\left\{x\in\Omega:\,x|_{[0,t_{0}]}=x_{0}|_{[0,t_{0}]}\text{ and }x|_{{\color[rgb]{0,0,0}(t_{0},T)}}\in W^{1,1}(t_{0},T;{\mathbb{R}}^{d})\right\}.

Here, W1,p(t0,T;d)W^{1,p}(t_{0},T;{\mathbb{R}}^{d}), p[1,]p\in[1,\infty], is the Sobolev space of all xLp(t0,T;d)x\in L^{p}(t_{0},T;{\mathbb{R}}^{d}) that have a weak derivative xLp(t0,T;d)x^{\prime}\in L^{p}(t_{0},T;{\mathbb{R}}^{d}). The terminal value problem involving the corresponding HJB equation is

(TVP) tu(t,x)+infad[axu(t,x)+(t,a)]=0in [0,T)×Ω,u(T,x)=h(x)on Ω.\begin{split}\partial_{t}u(t,x)+\inf_{a\in{\mathbb{R}}^{d}}\left[a\cdot\partial_{x}u(t,x)+\ell(t,a)\right]&=0\quad\text{in $[0,T)\times\Omega$},\\ u(T,x)&=h(x)\quad\text{on $\Omega$.}\end{split}

3. Notions of solutions of path-dependent HJB equations

We call a function u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\} non-anticipating if u(t,x)=u(t,x(t))u(t,x)=u(t,x(\cdot\wedge t)) for every (t,x)[0,T]×Ω(t,x)\in[0,T]\times\Omega. Note that whenever a function on [0,T]×Ω[0,T]\times\Omega is l.s.c. or u.s.c. (with respect to 𝐝\mathbf{d}_{\infty}), then it is automatically non-anticipating.

3.1. Dini solutions

Given a non-anticipating function u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\}, we define the lower and upper Dini derivative

du(t0,x0)(1,a)\displaystyle d_{-}u(t_{0},x_{0})(1,a) :=lim¯δ0u(t0+δ,x0(t0)+Aa(t0)Aa(t0))u(t0,x0)δ,\displaystyle:=\varliminf_{\delta\downarrow 0}\frac{u(t_{0}+\delta,x_{0}(\cdot\wedge t_{0})+{\color[rgb]{0,0,0}A^{a}(\cdot\vee t_{0})-A^{a}(t_{0})})-u(t_{0},x_{0})}{\delta}{\color[rgb]{0,0,0},}
d+u(t0,x0)(1,a)\displaystyle d_{+}u(t_{0},x_{0})(1,a) :=lim¯δ0u(t0+δ,x0(t0)+Aa(t0)Aa(t0))u(t0,x0)δ\displaystyle:=\varlimsup_{\delta\downarrow 0}\frac{u(t_{0}+\delta,x_{0}(\cdot\wedge t_{0})+{\color[rgb]{0,0,0}A^{a}(\cdot\vee t_{0})-A^{a}(t_{0})})-u(t_{0},x_{0})}{\delta}

at points (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega in direction (1,a)×d(1,a)\in{\mathbb{R}}\times{\mathbb{R}}^{d}. Here, in unison with the process AaA^{a} in Subsection 2.3, Aa(t)=atA^{a}(t)=at.

The following (path-dependent) notion of Dini semi-solutions is motivated by the notion of (contingent) solutions used in Theorem 4.1 of [DM-F00] for HJB equations related to Bolza problems. Our notion is also related to the infinitesimal version of minimax solutions for path-dependent Isaacs equations used in [Lukoyanov01JAMM].

Definition 3.1.

Let u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\}.

(i) We call uu a Dini supersolution of (TVP) if uu is l.s.c., u(T,)hu(T,\cdot)\geq h, and, for every (t0,x0)dom(u)(t_{0},x_{0})\in{\color[rgb]{0,0,0}\mathrm{dom}(u)} with t0<Tt_{0}<T,

(3.1) infad[du(t0,x0)(1,a)+(t0,a)]0.\displaystyle\inf_{a\in{\mathbb{R}}^{d}}\left[d_{-}u(t_{0},x_{0})(1,a)+\ell(t_{0},a)\right]\leq 0.

(ii) We call uu a Dini subsolution of (TVP) if uu is u.s.c., u(T,)hu(T,\cdot)\leq h, and, for every (t0,x0)dom(u)(t_{0},x_{0})\in\mathrm{dom}(u) with t0<Tt_{0}<T,

(3.2) infad[d+u(t0,x0)(1,a)+(t0,a)]0.\displaystyle\inf_{a\in{\mathbb{R}}^{d}}\left[d_{+}u(t_{0},x_{0})(1,a)+\ell(t_{0},a)\right]\geq 0.

We call uu a maximal Dini subsolution of (TVP) if uu is a Dini subsolution of (TVP) and, for every Dini subsolution vv of (TVP), we have uvu\geq v.

Example 3.2.

Let d=1d=1, h=.1Kch=\infty.{\mathbf{1}}_{K^{c}}, where K:={tt1/2}ΩK:=\{t\mapsto t^{1/2}\}\subset\Omega, and \ell be defined by (t,a)=|a|3/2\ell(t,a)=\left|{a}\right|^{3/2}. Then v0v_{0} satisfies

v0(t0,x0)={t0T23/2t3/4𝑑t=21/2(T1/4t01/4)if x0|[0,t0](t)=t1/2,otherwise.\displaystyle v_{0}(t_{0},x_{0})=\begin{cases}\int_{t_{0}}^{T}2^{-3/2}\,t^{-3/4}\,dt=2^{1/2}\left(T^{1/4}-t_{0}^{1/4}\right)&\text{if $x_{0}|_{[0,t_{0}]}(t)=t^{1/2}$,}\\ \infty&\text{otherwise.}\end{cases}

Note that, for each (t0,x0)dom(v0)(t_{0},x_{0})\in\mathrm{dom}(v_{0}) with t0<Tt_{0}<T and each aa\in{\mathbb{R}}, we have dv0(t0,x0)(1,a)=d_{-}v_{\color[rgb]{0,0,0}0}(t_{0},x_{0})(1,a)=\infty for which the constant perturbation involving aa is responsible. (To obtain a finite value for our Dini derivative, we would need to permit the non-constant perturbation tt1/2t\mapsto t^{1/2}.) Thus infa[dv0(t0,x0)(1,a)+|a|3/2]0\inf_{a\in{\mathbb{R}}}[d_{-}v_{\color[rgb]{0,0,0}0}(t_{0},x_{0})(1,a)+\left|{a}\right|^{3/2}]\leq 0 is never satisfied. Hence, in our example there does not exist a Dini supersolution of (TVP). This justifies the need for an appropriate weaker notion of solution (see Subsection 3.2).

Given a non-anticipating function u:[0,T]×Ωu:[0,T]\times\Omega\to{\mathbb{R}}, we define the upper stochastic Dini derivative

d+1,2u(t0,x0)(1,a,n1Id)\displaystyle d_{+}^{1,2}u(t_{0},x_{0})(1,a,n^{-1}I_{d})
:=lim¯δ0𝔼t0,x0,n[u(t0+δ,X+Aa(t0)Aa(t0))u(t0,x0)]δ.\displaystyle\qquad:=\varlimsup_{\delta\downarrow 0}\frac{{{\mathbb{E}}}_{t_{0},x_{0},n}\left[u(t_{0}+\delta,X+{\color[rgb]{0,0,0}A^{a}(\cdot\vee t_{0})-A^{a}(t_{0})})-u(t_{0},x_{0})\right]}{\delta}.

at points (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega in direction (1,a,n1Id)×d×d×d(1,a,n^{-1}I_{d})\in{\mathbb{R}}\times{\mathbb{R}}^{d}\times{\mathbb{R}}^{d\times d} (cf. [SubbotinaEtAl85, Subbotina06, Haussmann92MOR, Haussmann94SICON]).

The following notion of subsolutions for second order path-dependent PDEs is motivated by the minimax solutions used in [SubbotinaEtAl85, Subbotina06] in a Markovian framework.

Definition 3.3.

Let u:[0,T]×Ωu:[0,T]\times\Omega\to{\mathbb{R}}. We call uu a Dini subsolution of (TVP nn) if uu is u.s.c., u(T,)hu(T,\cdot)\leq h, and, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega,

(3.3) infad[d+1,2u(t0,x0)(1,a,n1Id)+(t0,a)]0.\displaystyle\inf_{a\in{\mathbb{R}}^{d}}\left[d^{1,2}_{+}u(t_{0},x_{0})(1,a,n^{-1}I_{d})+\ell(t_{0},a)\right]\geq 0.

We call uu a maximal Dini subsolution of (TVP nn) if uu is a Dini subsolution of (TVP nn) and, for every Dini subsolution vv of (TVP nn), we have uvu\geq v.

Remark 3.4.

In our specific setting, the use of Dini type semiderivatives such as those introduced in this section suffices. This motivates us to call our generalized solutions Dini solutions. For more general data, path-dependent counterparts of contingent derivatives such as the Clio derivatives in [AubinHaddad02PPDE] need to be utilized. Corresponding generalized solutions would be called contingent solutions.

3.2. Minimax solutions

Here, we introduce a weaker notion of solution, which is an adjustment of the infinitesimal notion of minimax solutions in [BK18JFA]. It is also motivated by the notion of (l.s.c. contingent) solutions used in Theorem 5.1 of [DM-F00]. The problem in Example 3.2, which partially motivated this weaker notion, is overcome by allowing non-constant perturbations (see also the notion of contingent solutions in [Carja12SICON] that are defined via contingent derivatives with function-valued directions).

Definition 3.5.

Let u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\} be non-anticipating.

(i) We call uu a minimax supersolution of (TVP) if uu is l.s.c., u(T,)hu(T,\cdot)\geq h, and, for every (t0,x0)dom(u)(t_{0},x_{0})\in{\color[rgb]{0,0,0}\mathrm{dom}(u)} with t0<Tt_{0}<T,

(3.4) infx𝒳1,1(t0,x0)lim¯δ0[u(t0+δ,x)u(t0,x0)+t0t0+δ(s,x(s))𝑑s]δ10.\displaystyle\inf_{x\in\mathcal{X}^{1,1}(t_{0},x_{0})}\varliminf_{\delta\downarrow 0}\left[u(t_{0}+\delta,x)-u(t_{0},x_{0})+\int_{t_{0}}^{t_{0}+\delta}\ell(s,x^{\prime}(s))\,ds\right]\delta^{-1}\leq 0.

(ii) We call uu an l.s.c. minimax subsolution of (TVP) if uu is l.s.c., u(T,)hu(T,\cdot)\leq h, and, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega and every (t,x)((t0,T]×𝒳1,1(t0,x0))dom(u)(t,x)\in\left((t_{0},T]\times\mathcal{X}^{1,1}(t_{0},x_{0})\right)\cap\mathrm{dom}(u) with t0t(s,x(s))𝑑s<\int_{t_{0}}^{t}\ell(s,x^{\prime}(s))\,ds<\infty,

(3.5) lim¯δ0[u(tδ,x)u(t,x)tδt(s,x(s))𝑑s]δ10.\displaystyle\varliminf_{\delta\downarrow 0}\left[u(t-\delta,x)-u(t,x)-\int_{t-\delta}^{t}\ell(s,x^{\prime}(s))\,ds\right]\delta^{-1}\leq 0.

(iii) We call uu an l.s.c. minimax solution of (TVP) if uu is a minimax super- and an l.s.c. minimax subsolution of (TVP).

3.3. Consistency with classical solutions

First, we provide the definitions for path derivatives. The first-order ones are due to Kim [KimBook] and the second-order ones are due to Dupire [dupirefunctional]. Our presentation follows [ETZ_I] and [Lukoyanov03] .

Definition 3.6.

Let u:[0,T]×Ωu:[0,T]\times\Omega\to{\mathbb{R}}.

(i) We write uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega) if uC([0,T]×Ω,)u\in C([0,T]\times\Omega,{\mathbb{R}}) and if there are functions tuC([0,T]×Ω,)\partial_{t}u\in C([0,T]\times\Omega,{\mathbb{R}}) and xuC([0,T]×Ω,d)\partial_{x}u\in C([0,T]\times\Omega,{\mathbb{R}}^{d}) called first-order path derivatives such that, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega, every x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}) and every t(t0,T]t\in(t_{0},T], we have

u(t,x)u(t0,x0)=t0t[tu(s,x)+x(s)xu(s,x)]𝑑s.\displaystyle u(t,x)-u(t_{0},x_{0})=\int_{t_{0}}^{t}\left[\partial_{t}u(s,x)+x^{\prime}(s)\cdot\partial_{x}u(s,x)\right]\,ds.

(ii) We write uC1,2([0,T]×Ω)u\in C^{1,2}([0,T]\times\Omega) if uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega) with corresponding first-order path derivatives tu\partial_{t}u and xu\partial_{x}u and if there is a function xxuC([0,T]×Ω,d×d)\partial_{xx}u\in C([0,T]\times\Omega,{\mathbb{R}}^{d\times d}) called second-order path derivative such that, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega, every probability measure PP on T{\mathcal{F}}_{T} such that XX is a (P,𝔽)(P,{\mathbb{F}})-Itô-semimartingale after time t0t_{0} with bounded characteristics and with P(X|[0,t0]=x0|[0,t0])=1P(X|_{[0,t_{0}]}=x_{0}|_{[0,t_{0}]})=1, and every t(t0,T]t\in(t_{0},T], we have

u(t,X)u(t0,x0)\displaystyle u(t,X)-u(t_{0},x_{0}) =t0ttu(s,X)ds+t0txu(s,X)dX(s)\displaystyle=\int_{t_{0}}^{t}\partial_{t}u(s,X)\,ds+\int_{t_{0}}^{t}\partial_{x}u(s,X)\cdot dX(s)
+t0t12xxu(s,X):dX(s),P-a.s.\displaystyle\qquad\qquad+\int_{t_{0}}^{t}\frac{1}{2}\partial_{xx}u(s,X):d\langle X(s)\rangle,\quad\text{$P$-a.s.}

Here, X()\langle X(\cdot)\rangle is the quadratic variation of X|[t0,T]×ΩX|_{[t_{0},T]\times\Omega} and, given matrices MM, Nd×dN\in{\mathbb{R}}^{d\times d}, M:NM:N is the trace of MNMN^{\top}.

Remark 3.7.

If uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega), then its first-order path derivatives are uniquely determined. If, in addition, uC1,2([0,T]×Ω)u\in C^{1,2}([0,T]\times\Omega), then its second-order path-derivative is uniquely determined as well. We refer to Section 2.3 of [ETZ_I] for more details.

Definition 3.8.

Let u:[0,T]×Ωu:[0,T]\times\Omega\to{\mathbb{R}}.

(i) We call uu a classical subsolution (resp. classical supersolution, classical solution) of (TVP) if uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega), u(T,)hu(T,\cdot)\leq h (resp. u(T,)hu(T,\cdot)\geq h, u(T,)=hu(T,\cdot)=h), and, for every (t,x)[0,T)×Ω(t,x)\in[0,T)\times\Omega,

tu(t,x)+infad[axu(t,x)+(t,a)](resp. =) 0.\displaystyle\partial_{t}u(t,x)+\inf_{a\in{\mathbb{R}}^{d}}\left[a\cdot\partial_{x}u(t,x)+\ell(t,a)\right]\geq\,\text{(resp.~{}$\leq$, $=$)}\,0.

(ii) We call uu a classical subsolution (resp. classical supersolution, classical solution) of (TVP nn) if uC1,2([0,T]×Ω)u\in C^{1,2}([0,T]\times\Omega) (resp. u(T,)hu(T,\cdot)\geq h, u(T,)=hu(T,\cdot)=h) and, for every (t,x)[0,T)×Ω(t,x)\in[0,T)\times\Omega,

tu(t,x)+12nxxu(t,x)+infad[axu(t,x)+(t,a)](resp. =) 0.\displaystyle\partial_{t}u(t,x)+\frac{1}{2n}\partial_{xx}u(t,x)+\inf_{a\in{\mathbb{R}}^{d}}\left[a\cdot\partial_{x}u(t,x)+\ell(t,a)\right]\geq\,\text{(resp.~{}$\leq$, $=$)}\,0.

The following result follows immediately from Definitions 3.1, 3.3, and 3.8.

Proposition 3.9 (Consistency of Dini solutions with classical solutions).

Assume that \ell is continuous.

(i) Let uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega). Then uu is a classical subsolution (resp. classical supersolution, classical solution) of (TVP) if and only if uu is a Dini subsolution (resp. Dini supersolution, Dini solution of (TVP).

(ii) Let uC1,2([0,T]×Ω)u\in C^{1,2}([0,T]\times\Omega). Then uu is a classical subsolution of (TVP nn) if and only if uu is a Dini subsolution of (TVP nn).

Proposition 3.10 (Partial consistency of l.s.c. minimax solutions with classical solutions).

Assume that \ell is continuous and real-valued. Let uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega).

(a) If uu is an l.s.c. minimax subsolution of (TVP), then uu is a classical subsolution of (TVP).

(b) If uu is a classical supersolution of (TVP), then uu is a minimax supersolution of (TVP).

Remark 3.11.

The converse of Proposition 3.10 (b) cannot expected to be valid in general because the infimum over 𝒳1,1(t0,x0)\mathcal{X}^{1,1}(t_{0},x_{0}) in (3.4) can be strictly less than the corresponding infimum over d{\mathbb{R}}^{d}. For similar reasons, we cannot expect the converse of Proposition 3.10 (a) to be valid in general.

Proof of Proposition 3.10.

Part (b) follows immediately from Definition 3.5 (i) and Definition 3.8 (i). It remains to prove part (a). To this end, fix (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega and assume that uu is an l.s.c. minimax subsolution of (TVP). Fix t(t0,T]t\in(t_{0},T]. Fix ada\in{\mathbb{R}}^{d}. Then, for any x()𝒳1,1(t0,x0)x(\cdot)\in\mathcal{X}^{1,1}(t_{0},x_{0}) with a continuous derivative xx^{\prime} that satisfies x(t)=ax^{\prime}(t)=a, we have

0\displaystyle 0 lim¯δ0[u(tδ,x)u(t,x)tδt(s,x(s))𝑑s]δ1\displaystyle\geq\varliminf_{\delta\downarrow 0}\left[u(t-\delta,x)-u(t,x)-\int_{t-\delta}^{t}\ell(s,x^{\prime}(s))\,ds\right]\delta^{-1}
=lim¯δ0[(tδttu(s,x)+xu(s,x)x(s)ds)tδt(s,x(s))𝑑s]δ1\displaystyle=\varliminf_{\delta\downarrow 0}\left[-\left(\int_{t-\delta}^{t}\partial_{t}u(s,x)+\partial_{x}u(s,x)\cdot x^{\prime}(s)\,ds\right)-\int_{t-\delta}^{t}\ell(s,x^{\prime}(s))\,ds\right]\delta^{-1}
=tu(t,x)xu(t,x)a(t,a).\displaystyle=-\partial_{t}u(t,x)-\partial_{x}u(t,x)\cdot a-\ell(t,a).

Since uC1,1([0,T]×Ω)u\in C^{1,1}([0,T]\times\Omega), \ell is continuous, and aa was arbitrary in d{\mathbb{R}}^{d}, we have

infad[tu(t0,x0)+xu(t0,x0)a+(t0,a)]0,\displaystyle\inf_{a\in{\mathbb{R}}^{d}}\left[\partial_{t}u(t_{0},x_{0})+\partial_{x}u(t_{0},x_{0})\cdot a+\ell(t_{0},a)\right]\geq 0,

i.e., uu is classical subsolution of (TVP). ∎

4. Main results

Theorem 4.1.

Assume (H1) with ϕ(r)=|r|p\phi(r)=\left|{r}\right|^{p} for some p>1p>1 and (H2). Let hh be continuous and bounded. Then (vn)(v_{n}) converges to v0v_{0} uniformly on compacta and v0v_{0} is continuous. Moreover, v0v_{0} is the unique l.s.c. minimax solution of (TVP) that is bounded from below. If, in addition, \ell is continuous and finite-valued, then we have the following:

(i) For each nn\in{\mathbb{N}}, the function vnv_{n} is the unique bounded maximal Dini subsolution of (TVP nn)

(ii) The function v0v_{0} is the unique bounded maximal Dini subsolution of (TVP).

Proof.

The uniform convergence of (vn)(v_{n}) to v0v_{0} on compacta and the continuity of v0v_{0} are proven in Section 5. Theorem 4.3 characterizes v0v_{0} as the unique minimax solution resp. as the unique bounded maximal Dini subsolution of (TVP). Theorem 7.2 addresses the remaining part, i.e., wellposedness of (TVP nn). ∎

Remark 4.2.

Well-posedness of (TVP nn) requires hh to be only u.s.c. and bounded (see Theorem 7.2). The corresponding result in the Markovian case treated in [BLT] is of similar strength (well-posedness holds for maximal viscosity supersolutions of the corresponding viscous Hamilton-Jacobi equations, which is due to [DrapeauMainberger16EJP]).

Theorem 4.3.

Assume (H1).

(a) Let hh be l.s.c., proper, and bounded from below. Then the value function v0v_{0} is the unique l.s.c. minimax solution of (TVP) that is bounded from below.

(b) Assume (H2). Let \ell be continuous and finite-valued. Let hh be u.s.c. and bounded. Then v0v_{0} is the unique maximal bounded Dini subsolution of (TVP).

Proof.

See Section 8. ∎

5. Proof of the convergence result

Consider the semicontinuous envelopes vv_{\ast} and vv^{\ast} defined by

v(t0,x0)\displaystyle v_{\ast}(t_{0},x_{0}) :=supδ>0,ninf(t,x)Oδ(t0,x0),mnvm(t,x),\displaystyle:=\sup\limits_{\begin{subarray}{c}\delta>0,\\ n\in{\mathbb{N}}\end{subarray}}\quad\inf\limits_{\begin{subarray}{c}(t,x)\in O_{\delta}(t_{0},x_{0}),\\ m\geq n\end{subarray}}\quad v_{m}(t,x),
v(t0,x0)\displaystyle v^{\ast}(t_{0},x_{0}) :=infδ>0,nsup(t,x)Oδ(t0,x0),mnvm(t,x)\displaystyle:=\inf\limits_{\begin{subarray}{c}\delta>0,\\ n\in{\mathbb{N}}\end{subarray}}\quad\sup\limits_{\begin{subarray}{c}(t,x)\in O_{\delta}(t_{0},x_{0}),\\ m\geq n\end{subarray}}\quad v_{m}(t,x)

for every (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega. Here, Oδ(t0,x0)O_{\delta}(t_{0},x_{0}) is the open δ\delta-neighborhood of (t0,x0)(t_{0},x_{0}) in ([0,T]×Ω,𝐝)([0,T]\times\Omega,\mathbf{d}_{\infty}).

First, we establish an auxiliary result.

Lemma 5.1.

Assume (H1). Let tnt0t_{n}\to t_{0} in [0,T][0,T] as nn\to\infty. Consider a probability space (Ω~,~,P~)(\tilde{\Omega},\tilde{{\mathcal{F}}},\tilde{P}). Let (an)(a_{n}) be a sequence in L1=L1([0,T]×Ω~,dtdP~;d)L^{1}=L^{1}([0,T]\times\tilde{\Omega},dt\otimes d\tilde{P};{\mathbb{R}}^{d}) that converges weakly to some aL1a\in L^{1}. Then 𝔼P~t0T(t,a(t))𝑑tlim¯n𝔼P~tnT(t,an(t))𝑑t{{\mathbb{E}}}^{\tilde{P}}\int_{t_{0}}^{T}\ell(t,a(t))\,dt\leq\varliminf_{n}{{\mathbb{E}}}^{\tilde{P}}\int_{t_{n}}^{T}\ell(t,a_{n}(t))\,dt.

Proof.

We follow the lines of the proof of the corresponding deterministic closure theorem 10.8.ii in [Cesari]. First, note that, by lower semi-continuity and convexity of \ell, for every δ>0\delta>0, we have 𝔼P~t0+δT(t,a(t))𝑑tlim¯n𝔼P~t0+δT(t,an(t))𝑑t{{\mathbb{E}}}^{\tilde{P}}\int_{t_{0}+\delta}^{T}\ell(t,a(t))\,dt\leq\varliminf_{n}{{\mathbb{E}}}^{\tilde{P}}\int_{t_{0}+\delta}^{T}\ell(t,a_{n}(t))\,dt (more details can be found the proof of Lemma A.1 of [BLT]). Next, fix ε>0{\varepsilon}>0. Since \ell is bounded from below, there is a δ0>0\delta_{0}>0 independent from nn such that, for all δ(0,δ0)\delta\in(0,\delta_{0}), we have 𝔼P~tnt0+δ(t,an(t))𝑑t>ε{{\mathbb{E}}}^{\tilde{P}}\int_{t_{n}}^{t_{0}+\delta}\ell(t,a_{n}(t))\,dt>-{\varepsilon} and thus lim¯n𝔼P~tnT(t,an(t))𝑑t+ε𝔼P~t0+δT(t,a(t))𝑑t\varliminf_{n}{{\mathbb{E}}}^{\tilde{P}}\int_{t_{n}}^{T}\ell(t,a_{n}(t))\,dt+{\varepsilon}\geq{{\mathbb{E}}}^{\tilde{P}}\int_{t_{0}+\delta}^{T}\ell(t,a(t))\,dt. Again, as \ell is bounded from below, either the right-hand side of the previous inequality converges to 𝔼P~t0T(t,a(t))𝑑t{{\mathbb{E}}}^{\tilde{P}}\int_{t_{0}}^{T}\ell(t,a(t))\,dt as δ0\delta\to 0 or otherwise the left-hand side equals \infty. This concludes the proof as ε{\varepsilon} was chosen arbitrarily. ∎

The following two statements adapt Theorem 2.2 in [BLT] to our slightly more general setting.

Lemma 5.2.

Assume (H1) with ϕ(r)=|r|p\phi(r)=\left|{r}\right|^{p} for some p>1p>1 and (H2). Let hh be l.s.c. and bounded from below. Then v0vv_{0}\leq v_{\ast}.

Proof.

Let (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega. It suffices to consider the case v(t0,x0)<v_{\ast}(t_{0},x_{0})<\infty. Let (tn,xn)n(t_{n},x_{n})_{n} be a sequence in [0,T]×Ω[0,T]\times\Omega that converges to (t0,x0)(t_{0},x_{0}) in 𝐝\mathbf{d}_{\infty} and that satisfies v(t0,x0)=lim¯nvn(tn,xn)v_{\ast}(t_{0},x_{0})=\varliminf_{n}v_{n}(t_{n},x_{n}). Let (an)(a_{n}) be a sequence in b\mathcal{L}_{b} such that each ana_{n} belongs to btn\mathcal{L}_{b}^{t_{n}} and is an n1n^{-1}-minimizer of (SOCn)(\text{SOC}_{n}) with initial data (tn,xn)(t_{n},x_{n}). Then there exists a subsequence (θk)k:=(tnk,xnk,nk)k(\theta_{k})_{k}:=(t_{n_{k}},x_{n_{k}},n_{k})_{k} of (tn,xn,n)n(t_{n},x_{n},n)_{n} with

v(t0,x0)\displaystyle v_{\ast}(t_{0},x_{0}) =limk𝔼θk[tnkT(t,ank(t))𝑑t+h(X+Aank)]\displaystyle=\lim_{k}{{\mathbb{E}}}_{\theta_{k}}\left[\int^{T}_{t_{n_{k}}}\ell(t,a_{n_{k}}(t))\,dt+h(X+A^{a_{n_{k}}})\right]

and, for all kk\in{\mathbb{N}},

v(t0,x0)1𝔼θktnkT(t,ank(t))𝑑t+𝔼θkh(X+Aank)v(t0,x0)+1.\displaystyle v_{\ast}(t_{0},x_{0})-1\leq{{\mathbb{E}}}_{\theta_{k}}\int_{t_{n_{k}}}^{T}\ell(t,a_{n_{k}}(t))\,dt+{{\mathbb{E}}}_{\theta_{k}}\,h(X+A^{a_{n_{k}}})\leq v_{\ast}(t_{0},x_{0})+1.

Since \ell and hh are bounded from below, we can assume that

v1=limk𝔼θktnkT(t,ank(t))𝑑t and v2=limk𝔼θkh(X+Aank)\displaystyle v^{1}=\lim_{k}{{\mathbb{E}}}_{\theta_{k}}\int_{t_{n_{k}}}^{T}\ell(t,a_{n_{k}}(t))\,dt\text{ and }v^{2}=\lim_{k}{{\mathbb{E}}}_{\theta_{k}}\,h(X+A^{a_{n_{k}}})

for some v1v^{1}, v2v^{2}\in{\mathbb{R}} with v(t0,x0)=v1+v2v_{\ast}(t_{0},x_{0})=v^{1}+v^{2} (cf. Theorem 11.1.i and its proof in [Cesari]). As supk𝔼θktnkT(t,ank(t))𝑑t<\sup_{k}{{\mathbb{E}}}_{\theta_{k}}\int_{t_{n_{k}}}^{T}\ell(t,a_{n_{k}}(t))\,dt<\infty and ank|[0,tnk]=0a_{n_{k}}|_{[0,t_{n_{k}}]}=0 for all kk\in{\mathbb{N}}, one can proceed nearly exactly as in the proof of Lemma A.1 in [BLT] to show that the probability measures (Pθk(Aank)1)k(P_{\theta_{k}}\circ(A^{a_{n_{k}}})^{-1})_{k} are tight. Let us point out the differences to [BLT]. Thanks to our additional requirement that the function ϕ\phi from Hypothesis (H1) satisfies ϕ(r)=|r|p\phi(r)=\left|{r}\right|^{p} for some p>1p>1, we can estimate 0T|ank(t)|p𝑑t\int_{0}^{T}\left|{a_{n_{k}}(t)}\right|^{p}\,dt instead of 0T|ank(t)|𝑑t\int_{0}^{T}\left|{a_{n_{k}}(t)}\right|\,dt (cf. with the first displayed equation in the proof of Lemma A.1 in [BLT]) and thus we can invoke Lemma 2 in [Zheng85] to obtain tightness (cf. also with the proof of Lemma 3.13 in [TanTouzi13]) instead of using the Aldous tightness criterion (Theorem 16.11 in [Kallenberg2nd]). Let us also note that the sequence (Pθk)(P_{\theta_{k}}) is weakly convergent because for each ηCb(Ω)\eta\in C_{b}(\Omega),

𝔼θkη=𝔼0,0η(xnk(tnk)+1nk(XX(tnk))𝟏[tnk,T])η(x0(t0))\displaystyle{{\mathbb{E}}}_{\theta_{k}}\,\eta={{\mathbb{E}}}_{0,0}\,\eta\left(x_{n_{k}}(\cdot\wedge t_{n_{k}})+\frac{1}{\sqrt{n_{k}}}\,{\color[rgb]{0,0,0}(X-X(t_{n_{k}})){\mathbf{1}}_{[t_{n_{k}},T]}}\right)\to\eta(x_{0}(\cdot\wedge t_{0}))

as kk\to\infty, i.e., a sequence of copies of XX converges in distribution to the constant x0(t0)x_{0}(\cdot\wedge t_{0}). Consequently, the probability measures (Pθk(Aank,X)1)k(P_{\theta_{k}}\circ(A^{a_{n_{k}}},X)^{-1})_{k} are tight. Thus, by Skorohod’s representation theorem, there exists a probability space (Ω¯,¯,P¯)(\bar{\Omega},\bar{{\mathcal{F}}},\bar{P}) with a sequence of Ω×Ω\Omega\times\Omega-valued random variables (A¯nk,X¯nk)k(\bar{A}_{n_{k}},\bar{X}_{n_{k}})_{k} that satisfies P¯(A¯nk,X¯nk)1=Pθk(Aank,X)1\bar{P}\circ(\bar{A}_{n_{k}},\bar{X}_{n_{k}})^{-1}=P_{\theta_{k}}\circ(A^{a_{n_{k}}},X)^{-1} for each k{\color[rgb]{0,0,0}k}\in{\mathbb{N}} and that converges (after passing to a subsequence), P¯\bar{P}-a.s., to some random variable (A¯0,X¯0)(\bar{A}_{0},\bar{X}_{0}). Next, define a sequence (a¯k)(\bar{a}_{k}) of d{\mathbb{R}}^{d}-valued processes on [0,T]×Ω¯[0,T]\times\bar{\Omega} by a¯k(t):=ank(t,X¯nk)\bar{a}_{k}(t):=a_{n_{k}}(t,\bar{X}_{n_{k}}). Again as in the proof of Lemma A.1 in [BLT], one can deduce that (a¯k)(\bar{a}_{k}) is equiabsolutely integrable and thus has a subsequential weak limit in L1([0,T]×Ω¯,dtdP¯;d)L^{1}([0,T]\times\bar{\Omega},dt\otimes d\bar{P};{\mathbb{R}}^{d}) that we denote by a¯0\bar{a}_{0} and that satisfies, by Lemma 5.1,

𝔼P¯t0T(t,a¯0(t))𝑑tlim¯i𝔼P¯tnkiT(t,a¯ki(t))𝑑t=lim¯i𝔼θkitnkiT(t,anki(t))𝑑t\displaystyle{{\mathbb{E}}}^{\bar{P}}\int_{t_{0}}^{T}\ell(t,\bar{a}_{0}(t))\,dt\leq\varliminf_{i}{{\mathbb{E}}}^{\bar{P}}\int_{t_{n_{k_{i}}}}^{T}\ell(t,\bar{a}_{k_{i}}(t))\,dt=\varliminf_{i}{{\mathbb{E}}}_{\theta_{k_{i}}}\int_{t_{n_{k_{i}}}}^{T}\ell(t,a_{n_{k_{i}}}(t))\,dt

as well as A¯0(t)=0ta¯0(s)𝑑s\bar{A}_{0}(t)=\int_{0}^{t}\bar{a}_{0}(s)\,ds, P¯\bar{P}-a.s., for every t[0,T]t\in[0,T]. Moreover, A¯0|[0,t0]=0\bar{A}_{0}|_{[0,t_{0}]}=0, P¯\bar{P}-a.s. Hence, together with hh being l.s.c. and X¯0\bar{X}_{0}=x0(t0)x_{0}(\cdot\wedge t_{0}), ¯\bar{{{\mathbb{P}}}}-a.s., we have

(5.1) v(t0,x0)=v1+v2𝔼P¯[t0T(t,a¯0(t))dt+h(x0(t0)+A¯0)].\displaystyle v_{\ast}(t_{0},x_{0})=v^{1}+v^{2}\geq{{\mathbb{E}}}^{\bar{P}}\left[\int_{t_{0}}^{T}\ell(t,\bar{a}_{0}(t))\,dt+h(x_{0}(\cdot\wedge t_{0})+\bar{A}_{0})\right].

To conclude the proof, it suffices to note that there exists some x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}) such that the right-hand side (5.1) is greater than or equal to t0T(t,x(t))dt+h(x\int_{t_{0}}^{T}\ell(t,x^{\prime}(t))\,dt+h(x) (cf. also Remark 2.6 of [HaussmannLepeltier90SICON]). ∎

Lemma 5.3.

Assume (H1) and (H2). Let hh be u.s.c. and bounded. Then vv0v^{\ast}\leq v_{0}.

Proof.

It suffices to follow the arguments of the first paragraph of the proof of Theorem 2.2 in [BLT] and make the obvious adjustments. For the convenience of the reader, we quickly go over it. Fix (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega and aL1(0,T;Rd)a\in L^{1}(0,T;R^{d}) with 0T(t,a(t))𝑑t<\int_{0}^{T}\ell(t,a(t))\,dt<\infty. Given NN\in{\mathbb{N}}, define aN(t):=a(t)a^{N}(t):=a(t) for |a(t)|N\left|{a(t)}\right|\leq N and aN(t):=(N/|a(t)|)a(t)a^{N}(t):=(N/\left|{a(t)}\right|)\,a(t) for |a(t)|>N\left|{a(t)}\right|>N. Next, consider a sequence (tn,xn)(t_{n},x_{n}) that converges to (t0,x0)(t_{0},x_{0}) in [0,T]×Ω[0,T]\times\Omega and satisfies v(t0,x0)=lim¯nvn(tn,xn)v^{\ast}(t_{0},x_{0})={\color[rgb]{0,0,0}\varlimsup_{n}}v_{n}(t_{n},x_{n}). Then

v(t0,x0)lim¯n𝔼tn,xn,n[tnT(t,aN(t))dt+h(X+AaN(tn)AaN(tn))]\displaystyle v^{\ast}(t_{0},x_{0})\leq\varlimsup_{n}{{\mathbb{E}}}_{t_{n},x_{n},n}\left[\int_{t_{n}}^{T}\ell(t,a^{N}(t))\,dt+h(X+A^{a^{N}}{\color[rgb]{0,0,0}(\cdot\vee t_{n})}-A^{a^{N}}(t_{n}))\right]
lim¯ntnT(t,aN(t))𝑑t\displaystyle\leq\varlimsup_{n}\int_{t_{n}}^{T}\ell(t,a^{N}(t))\,dt
+lim¯n𝔼0,0h(xn(tn)+1n(XX(tn)𝟏[tn,T])+AaN(tn)AaN(tn))\displaystyle\,+{\color[rgb]{0,0,0}\varlimsup_{n}{{\mathbb{E}}}_{0,0}\,h\left(x_{n}(\cdot\wedge t_{n})+\frac{1}{\sqrt{n}}\,{\color[rgb]{0,0,0}(X-X(t_{n}){\mathbf{1}}_{[t_{n},T]}})+A^{a^{N}}{\color[rgb]{0,0,0}(\cdot\vee t_{n})}-A^{a^{N}}(t_{n})\right)}
t0T(t,aN(t))dt+h(x0(t0)+AaN(t0)AaN(t0))\displaystyle\leq\int_{t_{0}}^{T}\ell(t,a^{N}(t))\,dt+h(x_{0}(\cdot\wedge t_{0})+A^{a^{N}}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a^{N}}(t_{0}))

as hh is u.s.c. as well as bounded and 0T(t,aN(t))𝑑t<\int_{0}^{T}\ell(t,a^{N}(t))\,dt<\infty, which follows from 0T(t,a(t))𝑑t<\int_{0}^{T}\ell(t,a(t))\,dt<\infty, (H2), and the convexity of (t,)\ell(t,\cdot) (cf. the first displayed equation after (38) in [BLT]). Finally, letting NN\to\infty concludes the proof as in [BLT]. ∎

Remark 5.4 (No Lavrentiev phenomenon).

Assume (H1) and (H2). Let hh be continuous and bounded. Then

v0(t0,x0)=infx𝒳1,(t0,x0)[t0T(t,x(t))𝑑t+h(x)],\displaystyle v_{0}(t_{0},x_{0})=\inf_{x\in\mathcal{X}^{1,\infty}(t_{0},x_{0})}\left[\int_{t_{0}}^{T}\ell(t,x^{\prime}(t))\,dt+h(x)\right],

where

𝒳1,(t0,x0):={xΩ:x|[0,t0]=x0|[0,t0] and x|[t0,T]W1,(t0,T;d)}.\displaystyle\mathcal{X}^{1,\infty}(t_{0},x_{0}):=\left\{x\in\Omega:\,x|_{[0,t_{0}]}=x_{0}|_{[0,t_{0}]}\text{ and }x|_{[t_{0},T]}\in W^{1,\infty}(t_{0},T;{\mathbb{R}}^{d})\right\}.

This result follows from the proofs of Lemmata 5.2 and 5.3 but with each Pt0,x0,nP_{t_{0},x_{0},n}, nn\in{\mathbb{N}}, replaced by the unique probability measure under which X=x(t0)X=x(\cdot\wedge t_{0}) a.s. For a more direct proof, it suffices to slightly modify the proof of Proposition 4.1 in [ButtazzoBelloni95] (this result is due to [DeArcangelis89]), where the case h0h\equiv 0 is treated.

Proof of the first conclusion of Theorem 4.1.

By Lemmata 5.2 and 5.3, v0vvv0v_{0}\leq v_{\ast}\leq v^{\ast}\leq v_{0}. Thus (vn)(v_{n}) converges to v0v_{0} uniformly on compacta and v0v_{0} is continuous (cf. Lemmata V.1.5 and V.1.9 of [BardiCapuzzoDolcetta] and keep Remark 2.1 in mind). ∎

6. Connections to BSDEs

We present parts of the theory of (convex) superquadratic BSDEs from [DrapeauEtAl13AOP] that are relevant for our work. Fix (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega, t1[t0,T]t_{1}\in[t_{0},T], nn\in{\mathbb{N}}, and a t1{\mathcal{F}}_{t_{1}}-measurable random variable ξ:Ω{}\xi:\Omega\to{\mathbb{R}}\cup\{\infty\}. Consider the BSDE

(6.1) dY(t)=supad[aZ(t)(t,a)]dt+Z(t)dX(t),on [t0,t1]Pt0,x0,n-a.s.,Y(t1)=ξ.\begin{split}dY(t)&=-\sup_{a\in{\mathbb{R}}^{d}}[a\cdot Z(t)-\ell(t,a)]\,dt+Z(t)\,dX(t),\,\text{on $[t_{0},t_{1}]$, $P_{t_{0},x_{0},n}$-a.s.,}\\ Y(t_{1})&=\xi.\end{split}
Definition 6.1.

A pair (Y,Z)(Y,Z) is a supersolution of (6.1) if

  • Y:[t0,t1]×ΩY:[t_{0},t_{1}]\times\Omega\to{\mathbb{R}} is a càdlàg and 𝔽t0,x0,n{\mathbb{F}}^{t_{0},x_{0},n}-adapted process,

  • Z:[t0,t1]×ΩdZ:[t_{0},t_{1}]\times\Omega\to{\mathbb{R}}^{d} is an 𝔽t0,x0,n{\mathbb{F}}^{t_{0},x_{0},n}-predictable process with

    𝔼t0,x0,nt0t1|Z(t)|2𝑑t<,{{\mathbb{E}}}_{t_{0},x_{0},n}\int_{t_{0}}^{t_{1}}\left|{Z(t)}\right|^{2}\,dt<\infty,
  • (t,ω)[t0tZ(r)𝑑X(r)](ω)(t,\omega)\mapsto\left[\int_{t_{0}}^{t}Z(r)\,dX(r)\right](\omega), [t0,t1]×Ω[t_{0},t_{1}]\times\Omega\to{\mathbb{R}}, is an (𝔽t0,x0,n,t0,x0,n)({\mathbb{F}}^{t_{0},x_{0},n},{{\mathbb{P}}}_{t_{0},x_{0},n})-supermartingale, and,

  • for every tt, s[t0,t1]s\in[t_{0},t_{1}] with tst\leq s, we have, Pt0,x0,nP_{t_{0},x_{0},n}-a.s.,

    Y(s)\displaystyle Y(s) Y(t)+ts{supad[aZ(r)(r,a)]}𝑑r+tsZ(r)𝑑X(r),\displaystyle\geq Y(t)+\int_{t}^{s}\left\{-\sup_{a\in{\mathbb{R}}^{d}}[a\cdot Z(r)-\ell(r,a)]\right\}\,dr+\int_{t}^{s}Z(r)\,dX(r),
    Y(t1)\displaystyle Y(t_{1}) ξ.\displaystyle\geq\xi.

A pair (Y,Z)(Y,Z) is a minimal supersolution of (6.1) if it is a supersolution of (6.1) and, for every supersolution (Y~,Z~)(\tilde{Y},\tilde{Z}) of (6.1), we have YY~Y\leq\tilde{Y}, dtdPt0,x0,ndt\otimes dP_{t_{0},x_{0},n}-a.e.

Proposition 6.2.

Assume (H1) and (H2). Let ξ\xi be bounded. Then (6.1) has a unique minimal supersolution (Y,Z)(Y,Z) and we write

(6.2) t,t1t0,x0,n(ξ):=Yt,t[t0,t1].\displaystyle\mathcal{E}^{t_{0},x_{0},n}_{t,t_{1}}(\xi):=Y_{t},\quad t\in[t_{0},t_{1}].
Proof.

By the proof of Theorem 3.4 in [DrapeauEtAl16AIHP] and by Theorem A.1 in [DrapeauEtAl16AIHP], the set of supersolutions of (6.1) is non-empty. Thus we can apply Theorem 4.17 in [DrapeauEtAl13AOP], which concludes the proof. ∎

From now on, let T=1T=1 in this section for the sake of simplicity. Fix (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega. Recall that bt0={ab:a|[0,t0)=0}\mathcal{L}_{b}^{t_{0}}=\{a\in\mathcal{L}_{b}:a|_{[0,t_{0})}=0\}. Given aba\in\mathcal{L}_{b}, define Aa:[0,T]×ΩdA^{a}:[0,T]\times\Omega\to{\mathbb{R}}^{d} by Aa(t,ω):=0ta(s,ω)𝑑sA^{a}(t,{\omega}):=\int_{0}^{t}a(s,{\omega})\,ds. We also identify a control aba\in\mathcal{L}_{b} with the function ΩL0(Ω,T)\Omega\to L^{0}(\Omega,{\mathcal{F}}_{T}), ωa(ω)\omega\mapsto a(\omega), where [a(ω)](t)=a(t,ω)[a(\omega)](t)=a(t,\omega). Recall that X(t,ω)=ω(t)X(t,{\omega})={\omega}(t), (t,ω)[0,T]×Ω(t,{\omega})\in[0,T]\times\Omega, and that 𝔼t0,x0=𝔼Pt0,x0{{\mathbb{E}}}_{t_{0},x_{0}}={{\mathbb{E}}}^{P_{t_{0},x_{0}}}.

For the next statement, we borrow the following constructions from [BLT]. Given t[0,1)t\in[0,1), ω1\omega_{1}, ω2Ω\omega_{2}\in\Omega with ω2(0)=0\omega_{2}(0)=0, a path ω1tω2Ω\omega_{1}\odot_{t}\omega_{2}\in\Omega is defined by

(6.3) (ω1tω2)(s):=ω1(st)+1tω2(st1t)𝟏[t,1](s)\displaystyle(\omega_{1}\odot_{t}\omega_{2})(s):=\omega_{1}(s\wedge t)+\sqrt{1-t}\,\omega_{2}\left(\frac{s-t}{1-t}\right){\mathbf{1}}_{[t,1]}(s)

(note that in [BLT] \otimes is used instead of \odot) and a path ω1(t)Ω\omega_{1}^{(t)}\in\Omega is defined by

ω1(t)(s):=11t[ω1(t+s(1t))ω1(t)].\displaystyle\omega_{1}^{(t)}(s):=\frac{1}{\sqrt{1-t}}\,[\omega_{1}(t+s(1-t))-\omega_{1}(t)].

Moreover, we use the function (t0):[0,1]×d{}\ell^{(t_{0})}:[0,1]\times{\mathbb{R}}^{d}\to{\mathbb{R}}\cup\{\infty\} defined by

(6.4) (t0)(t,a):=(1t0)(t0+t(1t0),a1t0).\displaystyle\ell^{(t_{0})}(t,a):=(1-t_{0})\,\ell\left(t_{0}+t(1-t_{0}),\frac{a}{\sqrt{1-t_{0}}}\right).

Also set (ω11ω2):=ω1(\omega_{1}\odot_{1}\omega_{2}):=\omega_{1}, ω1(1)0\omega_{1}^{(1)}\equiv 0, and (1)0\ell^{(1)}\equiv 0.

Lemma 6.3.

Assume (H1) and (H2). Let hh be bounded. Then

(6.5) v1(t0,x0)=supaˇb𝔼0,0[(h)(x0t0[X+Aaˇ(X)])01(t0)(t,aˇ(t,X))𝑑t].\begin{split}{\color[rgb]{0,0,0}-v_{1}(t_{0},x_{0})}=\sup_{\check{a}\in\mathcal{L}_{b}}{{\mathbb{E}}}_{0,0}\left[(-h)(x_{0}\odot_{t_{0}}[X+A^{\check{a}(X)}])-\int_{0}^{1}\ell^{(t_{0})}(t,\check{a}(t,X))\,dt\right].\end{split}

The calculations in our proof are essentially the same as those in the proof of Lemma 5.1 in [BLT]. Although of similar nature, the corresponding statements are different. This permits a more elementary proof in our case.

Proof of Lemma 6.3.

One should keep in mind, that X=x0t0X(t0)X=x_{0}\odot_{t_{0}}X^{(t_{0})}, Pt0,x0P_{t_{0},x_{0}}-a.s. We only consider the case t0<1t_{0}<1 as otherwise (6.5) is clearly satisfied.

Step 1. Fix abt0a\in\mathcal{L}_{b}^{t_{0}}. Note that

𝔼t0,x0[t01(t,a(t,x0t0X(t0)))𝑑t]=𝔼0,0[t01(t,a(t,x0t0X))𝑑t]\displaystyle{{\mathbb{E}}}_{t_{0},x_{0}}\left[\int_{t_{0}}^{1}\ell(t,a({\color[rgb]{0,0,0}t},x_{0}\odot_{t_{0}}X^{(t_{0})}))\,dt\right]={{\mathbb{E}}}_{0,0}\left[\int_{t_{0}}^{1}\ell(t,a({\color[rgb]{0,0,0}t},x_{0}\odot_{t_{0}}X))\,dt\right]
=𝔼0,0[(1t0)01(t0+t(1t0),a(t0+t(1t0),x0t0X))𝑑t]\displaystyle={{\mathbb{E}}}_{0,0}\left[(1-t_{0})\int_{0}^{1}\ell(t_{0}+t(1-t_{0}),a(t_{0}+t(1-t_{0}),x_{0}\odot_{t_{0}}X))\,dt\right]
=𝔼0,0[01(t0)(t,aˇ(t,X))𝑑t],\displaystyle={{\mathbb{E}}}_{0,0}\left[\int_{0}^{1}\ell^{(t_{0})}(t,\check{a}(t,X))\,dt\right],

where aˇb\check{a}\in\mathcal{L}_{b} is defined by

aˇ(t,ω):=1t0a(t0+t(1t0),x0t0ω).\displaystyle\check{a}(t,\omega):=\sqrt{1-t_{0}}\,a(t_{0}+t(1-t_{0}),x_{0}\odot_{t_{0}}\omega).

Also note that

𝔼t0,x0[(h)(x0t0[X+Aa(x0t0X(t0))](t0))]\displaystyle{{\mathbb{E}}}_{t_{0},x_{0}}\left[(-h)(x_{0}\odot_{t_{0}}[X+A^{a(x_{0}\odot_{t_{0}}X^{(t_{0})}{\color[rgb]{0,0,0})}}]^{(t_{0})})\right]
=𝔼0,0[(h)(x0t0[X+(Aa(x0t0X))(t0)])]\displaystyle={{\mathbb{E}}}_{0,0}\left[(-h)(x_{0}\odot_{t_{0}}[X+(A^{a(x_{0}\odot_{t_{0}}X)})^{(t_{0})}])\right]
=𝔼0,0[(h)(x0t0[X+Aaˇ])]\displaystyle={{\mathbb{E}}}_{0,0}\left[(-h)(x_{0}\odot_{t_{0}}[X+A^{\color[rgb]{0,0,0}\check{a}}])\right]

because, for every t[0,1]t\in[0,1],

(Aa(x0t0X))(t0)(t)=11t0t0t0+t(1t0)a(s,x0t0X)𝑑s=0taˇ(s,X)𝑑s.\displaystyle(A^{a(x_{0}\odot_{t_{0}}X)})^{(t_{0})}{\color[rgb]{0,0,0}(t)}=\frac{1}{\sqrt{1-t_{0}}}\int_{t_{0}}^{t_{0}+t(1-t_{0})}a(s,x_{0}\odot_{t_{0}}X)\,ds=\int_{0}^{t}\check{a}(s,X)\,ds.

Consequently, the left-hand side of (6.5) is less than or equal to the right-hand side of (6.5).

Step 2. Fix aˇb\check{a}\in\mathcal{L}_{b}. Define abt0a\in\mathcal{L}_{b}^{t_{0}} by

a(t,ω):=11t0aˇ(tt01t0,ω(t0)) 1[t0,1]×{x0|[0,t0]}(t,ω|[0,t0]).\displaystyle a(t,\omega):=\frac{1}{\sqrt{1-t_{0}}}\,\check{a}\left(\frac{t-t_{0}}{1-t_{0}},\omega^{(t_{0})}\right)\,{\mathbf{1}}_{[t_{0},1]\times\{x_{0}|_{[0,t_{0}]}\}}(t,\omega|_{[0,t_{0}]}).

Going over the calculations in Step 1 backward, one can deduce that the right-hand side of (6.5) is less than or equal to the left-hand side of (6.5). ∎

The following statement together with Theorem 4.1 provide a non-Markovian non-linear Feynman-Kac formula that connects maximal Dini subsolutions of path-dependent PDEs with minimal supersolutions of convex superquadratic BSDEs, for which we use the notation (6.2).

Theorem 6.4.

Assume (H1) and (H2). Let \ell be continuous and finite-valued. Let hh be bounded. Fix (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega and nn\in{\mathbb{N}}. Then vn(t0,x0)=𝔼t0,x0,n[t0,Tt0,x0,n(h)]v_{n}(t_{0},x_{0})=-{{\mathbb{E}}}_{t_{0},x_{0},n}[\mathcal{E}^{t_{0},x_{0},n}_{t_{0},T}(-h)]. Moreover, for Pt0,x0,nP_{t_{0},x_{0},n}-a.e. ωΩ\omega\in\Omega and every t[t0,T]t\in[t_{0},T], we have vn(t,ω)=t,Tt0,x0,n(h)(ω)v_{n}(t,\omega)=-\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h)(\omega).

Proof.

We prove the theorem only for the case n=1n=1 and T=1T=1.

We will employ expressions ρf\rho^{f}, ff being a measurable function from [0,1]×d[0,1]\times{\mathbb{R}}^{d} to {}{\mathbb{R}}\cup\{\infty\} that satisfies (H1) and (H2) in place of \ell, from Section 2 of [BLT], which are defined by

ρf(ξ):=supQ𝒬𝔼Q[ξ01f(s,aQ(s))𝑑s],ξ:Ω bounded and T-measurable,\displaystyle\color[rgb]{0,0,0}\rho^{f}(\xi):=\sup_{Q\in\mathcal{Q}}{{\mathbb{E}}}^{Q}\left[\xi-\int_{0}^{1}f(s,a^{Q}(s))\,ds\right],\,\text{$\xi:\Omega\to{\mathbb{R}}$ bounded and ${\mathcal{F}}_{T}$-measurable,}

where 𝒬\mathcal{Q} is the set of all probability measures on (Ω,T)(\Omega,{\mathcal{F}}_{T}) that are absolutely continuous with respect to P0,0P_{0,0} and aQ:[0,1]×Ωda^{Q}:[0,1]\times\Omega\to{\mathbb{R}}^{d} is the unique 𝔽{\mathbb{F}}-progressive process with 01|aQ(s)|2𝑑s<\int_{0}^{1}\left|{a^{Q}(s)}\right|^{2}\,ds<\infty, P0,0P_{0,0}-a.s., that satisfies

dQ=exp(01aQ(s)𝑑X(s)1201|aQ(s)|2𝑑s)dP0,0\displaystyle\color[rgb]{0,0,0}dQ=\exp\left(\int_{0}^{1}a^{Q}(s)\,dX(s)-\frac{1}{2}\int_{0}^{1}\left|{a^{Q}(s)}\right|^{2}\,ds\right)\,dP_{0,0}

(Section 2 of [BLT]).

First note that the right-hand side of (6.5) equals ρ(t0)[(h)(x0t0X)]\rho^{\ell^{(t_{0})}}[(-h)(x_{0}\odot_{t_{0}}X{\color[rgb]{0,0,0})}] according to (BBD) in [BLT]. Moreover, by a straight-forward adjustment of the proof of Lemma 5.1 in [BLT], ρ(t)[(h)(ωtX)]=Y(t,ω)\rho^{\ell^{(t)}}[(-h)(\omega\odot_{t}X{\color[rgb]{0,0,0})}]={\color[rgb]{0,0,0}Y(t,\omega)} for every t[t0,T]t\in[t_{0},T] and Pt0,x0P_{t_{0},x_{0}}-a.e. ωΩ\omega\in\Omega. In this context, note that Lemma 5.1 and Lemma A.2, both in [BLT], formally require continuity of hh, which, however, is not necessary (cf. the corresponding material in [DrapeauEtAl16AIHP]). Using Lemma 6.3, we can deduce that v1(t,)=Y(t)v_{1}(t,\cdot)={\color[rgb]{0,0,0}-Y(t)}, Pt0,x0P_{t_{0},x_{0}}-a.s., for every t[t0,T]t\in[t_{0},T]. ∎

7. The second order HJB equations

Lemma 7.1.

Let nn\in{\mathbb{N}}. Assume (H1). Let =(t,a)\ell=\ell(t,a) be continuous in tt. A bounded u.s.c. function u:[0,T]×Ωu:[0,T]\times\Omega\to{\mathbb{R}} is a Dini subsolution of (TVP nn) if and only if u(T,)hu(T,\cdot)\leq h and, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega, t(t0,T]t\in(t_{0},T], and ada\in{\mathbb{R}}^{d},

(7.1) u(t0,x0)𝔼t0,x0,n[t0t(s,a)ds+u(t,X+Aa(t0)Aa(t0))].\displaystyle u(t_{0},x_{0})\leq{{\mathbb{E}}}_{t_{0},x_{0},n}\left[\int_{t_{0}}^{t}\ell(s,a)\,ds+u(t,X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))\right].
Proof.

We adapt the proof of Theorem V.3 in [Subbotina06], which is situated in a Markovian context with bounded control spaces, to our non-Markovian setting with the unbounded control space d{\mathbb{R}}^{d}.

(i) Let uu be a Dini subsolution of (TVP nn). Fix (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega. Put 𝔼=𝔼t0,x0,n{{\mathbb{E}}}={{\mathbb{E}}}_{t_{0},x_{0},n} and P=Pt0,x0,nP=P_{t_{0},x_{0},n}. Assume that there are t1(t0,T]t_{1}\in(t_{0},T] and ada\in{\mathbb{R}}^{d}, and ε>0{\varepsilon}>0 such that

(7.2) 𝔼[u(t1,X+Aa(t0)Aa(t0))u(t0,x0)+t0t1(s,a)ds]<ε.\displaystyle{{\mathbb{E}}}\left[u(t_{1},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))-u(t_{0},x_{0})+\int_{t_{0}}^{t_{1}}\ell(s,a)\,ds\right]<-{\varepsilon}.

Consider the set SS of all 𝔽{\mathbb{F}}-stopping times τ\tau that satisfy t0τt1t_{0}\leq\tau\leq t_{1}, PP-a.s., and

(7.3) 𝔼[u(τ,X+Aa(t0)Aa(t0))u(t0,x0)+t0τ(s,a)ds]𝔼[τ]t0t1t0(ε).\displaystyle{{\mathbb{E}}}\left[u(\tau,X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))-u(t_{0},x_{0})+\int_{t_{0}}^{\tau}\ell(s,a)\,ds\right]\geq\frac{{{\mathbb{E}}}[\tau]-t_{0}}{t_{1}-t_{0}}\cdot(-{\varepsilon}).

We equip SS with an order \preccurlyeq defined by τ1τ2\tau_{1}\preccurlyeq\tau_{2} if and only if τ1τ2\tau_{1}\leq\tau_{2}, PP-a.s. Now, let S~\tilde{S} be a totally ordered non-empty subset of SS. Note that there exists a sequence (τk)(\tau_{k}) in S~\tilde{S} such that 𝔼[τk]supτS~𝔼[τ]{{\mathbb{E}}}[\tau_{k}]\uparrow\sup_{\tau\in\tilde{S}}{{\mathbb{E}}}[\tau] as kk\to\infty. Since S~\tilde{S} is totally ordered and 𝔼{{\mathbb{E}}} is linear, (τk)(\tau_{k}) is increasing and converges to τ~:=supkτk\tilde{\tau}:=\sup_{k}\tau_{k}. We have τ~S\tilde{\tau}\in S because

𝔼[u(τ~,X+Aa(t0)Aa(t0))+t0τ~(s,a)ds]\displaystyle{{\mathbb{E}}}\left[u(\tilde{\tau},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{t_{0}}^{\tilde{\tau}}\ell(s,a)\,ds\right]
lim¯k𝔼[u(τk,X+Aa(t0)Aa(t0))+t0τk(s,a)ds]\displaystyle\geq\varlimsup_{k}{{\mathbb{E}}}\left[u(\tau_{k},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{t_{0}}^{\tau_{k}}\ell(s,a)\,ds\right]
u(t0,x0)+𝔼[τ~]t0t1t0(ε)\displaystyle\geq u(t_{0},x_{0})+\frac{{{\mathbb{E}}}[\tilde{\tau}]-t_{0}}{t_{1}-t_{0}}\cdot(-{\varepsilon})

thanks to (7.3) as well as to uu being u.s.c and bounded. Hence, by Zorn’s lemma, SS has a maximal element τ0\tau_{0}. We show that τ0=t1\tau_{0}=t_{1}, PP-a.s. To this end, assume that P(τ0<t1)>0P(\tau_{0}<t_{1})>0. Define a set-valued function MM on hypou|[t0,t1]×Ω={(t,ω,y)[t0,t1]×Ω×:yu(t,ω)}\color[rgb]{0,0,0}\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}=\{(t,\omega,y)\in[t_{0},t_{1}]\times\Omega\times{\mathbb{R}}:\,y\leq u(t,\omega)\} by

M(t,ω,y):={δ[0,t1t0]:t+δt1 and\displaystyle M(t,\omega{\color[rgb]{0,0,0},y}):=\Biggl{\{}\delta\in[0,{\color[rgb]{0,0,0}t_{1}}-t_{0}]:\,t+\delta\leq{\color[rgb]{0,0,0}t_{1}}\text{ and}
𝔼t,ω,n[u(t+δ,X+Aa(t)Aa(t))y+tt+δ(s,a)ds]δt1t0(ε)}.\displaystyle\quad{{\mathbb{E}}}_{t,\omega,n}\left[u(t+\delta,X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t)}-A^{a}(t))-{\color[rgb]{0,0,0}y}+\int_{t}^{t+\delta}\ell(s,a)\,ds\right]\geq\frac{\delta}{t_{1}-t_{0}}\cdot(-{\varepsilon})\Biggr{\}}.

The sets M(t,ω,y)M(t,\omega{\color[rgb]{0,0,0},y}) are non-empty because they contain 0 and they are compact because, for any sequence (δk)(\delta_{k}) in [0,t1t][0,{\color[rgb]{0,0,0}t_{1}}-t] that converges to some δ\delta, we have

𝔼t,ω,n[u(t+δ,X+Aa(t)Aa(t))+tt+δ(s,a)ds]\displaystyle{{\mathbb{E}}}_{t,\omega,n}\left[u(t+\delta,X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t)}-A^{a}(t))+\int_{t}^{t+\delta}\ell(s,a)\,ds\right]
lim¯k𝔼t,ω,n[u(t+δk,X+Aa(t)Aa(t))+tt+δk(s,a)ds]\displaystyle\qquad\geq\varlimsup_{k}{{\mathbb{E}}}_{t,\omega,n}\left[u(t+\delta_{k},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t)}-A^{a}(t))+\int_{t}^{t+\delta_{k}}\ell(s,a)\,ds\right]

as uu is u.s.c. and bounded. Moreover, for every sequence (sk,ωk,yk)k(s_{k},\omega_{k}{\color[rgb]{0,0,0},y_{k}})_{k} that belongs to hypou|[t0,t1]×Ω{\color[rgb]{0,0,0}\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}} and that converges to some (t,ω,y)(t,\omega{\color[rgb]{0,0,0},y}) with respect to the metric ((s~1,ω~1,y~1),(s~2,ω~2,y~2))|s~1s~2|+ω~1ω~2+|y~1y~2|((\tilde{s}_{1},\tilde{\omega}_{1}{\color[rgb]{0,0,0},\tilde{y}_{1}}),(\tilde{s}_{2},\tilde{\omega}_{2}{\color[rgb]{0,0,0},\tilde{y}_{2}}))\mapsto\left|{\tilde{s}_{1}-\tilde{s}_{2}}\right|+\lVert{\tilde{\omega}_{1}-\tilde{\omega}_{2}}\rVert_{\infty}{\color[rgb]{0,0,0}+\left|{\tilde{y}_{1}-\tilde{y}_{2}}\right|} and hence also with respect to ((s~1,ω~1,y~1),(s~2,ω~2,y~2))𝐝((s~1,ω~1),(s~2,ω~2))+|y~1y~2|{\color[rgb]{0,0,0}((\tilde{s}_{1},\tilde{\omega}_{1}{\color[rgb]{0,0,0},\tilde{y}_{1}}),(\tilde{s}_{2},\tilde{\omega}_{2}{\color[rgb]{0,0,0},\tilde{y}_{2}}))\mapsto\mathbf{d}_{\infty}((\tilde{s}_{1},\tilde{\omega}_{1}),(\tilde{s}_{2},\tilde{\omega}_{2}))+\left|{\tilde{y}_{1}-\tilde{y}_{2}}\right|}, every sequence (δk)(\delta_{k}) in [0,t1t0][0,{\color[rgb]{0,0,0}t_{1}}-t_{0}] with δkM(sk,ωk,yk)\delta_{k}\in M(s_{k},\omega_{k}{\color[rgb]{0,0,0},y_{k}}) has a subsequential limit that belongs to M(t,ω,y)M(t,\omega,{\color[rgb]{0,0,0}y}) because, for every subsequential limit δ\delta of (δk)(\delta_{k}) (there is at least one), we have, by possibly passing to a subsequence,

𝔼t,ω,n[u(t+δ,X+Aa(t)Aa(t))+tt+δ(s,a)ds]\displaystyle{{\mathbb{E}}}_{t,\omega,n}\left[u(t+\delta,X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t)}-A^{a}(t))+\int_{t}^{t+\delta}\ell(s,a)\,ds\right]
=𝔼0,0[u((t+δ,ω(t)+1n(XX(t))𝟏[t,T])+Aa(t)Aa(t))\displaystyle={\color[rgb]{0,0,0}{{\mathbb{E}}}_{0,0}}\Biggl{[}u\left((t+\delta,\omega(\cdot\wedge t)+\frac{1}{\sqrt{n}}{\color[rgb]{0,0,0}(X-X(t)){\mathbf{1}}_{[t,T]}}{\color[rgb]{0,0,0})}+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t)}-A^{a}(t)\right)
+tt+δ(s,a)ds]\displaystyle\qquad\qquad+\int_{t}^{t+\delta}\ell(s,a)\,ds\Biggr{]}
lim¯k𝔼0,0[u((sk+δk,ωk(sk)+1n(XX(sk))𝟏[sk,T])\displaystyle\geq\varlimsup_{k}{\color[rgb]{0,0,0}{{\mathbb{E}}}_{0,0}}\Biggl{[}u\Biggl{(}(s_{k}+\delta_{k},\omega_{k}(\cdot\wedge s_{k})+\frac{1}{\sqrt{n}}{\color[rgb]{0,0,0}(X-X(s_{k})){\mathbf{1}}_{[s_{k},T]}}{\color[rgb]{0,0,0})}
+Aa(sk)Aa(sk))+sksk+δk(s,a)ds]\displaystyle\qquad\qquad\qquad\qquad+A^{a}{\color[rgb]{0,0,0}(\cdot\vee s_{k})}-A^{a}(s_{k})\Biggr{)}+\int_{s_{k}}^{s_{k}+\delta_{k}}\ell(s,a)\,ds\Biggr{]}
=lim¯k𝔼sk,ωk,n[u(sk+δk,X+Aa(sk)Aa(sk))+sksk+δk(s,a)ds].\displaystyle=\varlimsup_{k}{{\mathbb{E}}}_{s_{k},\omega_{k},n}\left[u(s_{k}+\delta_{k},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee s_{k})}-A^{a}(s_{k}))+\int_{s_{k}}^{s_{k}+\delta_{k}}\ell(s,a)\,ds\right].

Hence, MM considered as a map from hypou|[t0,t1]×Ω{\color[rgb]{0,0,0}\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}} equipped here with the metric ((s1,ω1,y1),(s2,ω2,y2))|s1s2|+ω1ω2+|y1y2|((s_{1},\omega_{1}{\color[rgb]{0,0,0},y_{1}}),(s_{2},\omega_{2}{\color[rgb]{0,0,0},y_{2}}))\mapsto\left|{s_{1}-s_{2}}\right|+\lVert{\omega_{1}-\omega_{2}}\rVert_{\infty}{\color[rgb]{0,0,0}+\left|{y_{1}-y_{2}}\right|} into the set of all compact non-empty subsets of [0,t1t0][0,{\color[rgb]{0,0,0}t_{1}}-t_{0}] is u.s.c. (Theorem 2.2 on p. 31 in [Kisielewicz91DI]) and thus measurable (Theorem 2.1 on p. 29 in [Kisielewicz91DI]), i.e., all sets of the form {(t,ω,y)hypou|[t0,t1]×Ω:M(t,ω,y)E}\{(t,\omega,y)\in\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}:\,M(t,\omega,y)\cap E\neq\emptyset\}, EE being a closed subset of [0,t1t0][0,t_{1}-t_{0}], belong to (hypou|[t0,t1]×Ω)\mathcal{B}(\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}) (see p. 41 in [Kisielewicz91DI]). Consequently, by Theorem 3.13 on p. 49 in [Kisielewicz91DI], there exists a (hypou|[t0,t1]×Ω){\color[rgb]{0,0,0}\mathcal{B}(\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega})}-measurable selector δ(,,)\delta(\cdot,\cdot{\color[rgb]{0,0,0},\cdot}) of MM that satisfies |t1t0δ(t,ω,y)|=dist(t1t0,M(t,ω,y))\left|{{\color[rgb]{0,0,0}t_{1}}-t_{0}-\delta(t,\omega{\color[rgb]{0,0,0},y})}\right|=\mathrm{dist}({\color[rgb]{0,0,0}t_{1}}-t_{0},M(t,\omega{\color[rgb]{0,0,0},y})) for every (t,ω,y)hypou|[t0,t1]×Ω(t,\omega{\color[rgb]{0,0,0},y})\in{\color[rgb]{0,0,0}\mathrm{hypo}\,u|_{[t_{0},t_{1}]\times\Omega}}. Therefore, the map

ωδ[ω]\displaystyle\omega\mapsto\delta[\omega] :=δ(τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0),\displaystyle:=\delta(\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}({\color[rgb]{0,0,0}(\cdot\vee t_{0})\wedge}\tau_{0}(\omega))-A^{a}(t_{0}){\color[rgb]{0,0,0},}
u(τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0)))\displaystyle\qquad\qquad{\color[rgb]{0,0,0}u(\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0}))})

is τ0{\mathcal{F}}_{\tau_{0}}-measurable. Also note that, since, by (3.3) and the continuity of (,a)\ell(\cdot,a), every set M(t,ω,u(t,ω))M(t,\omega{\color[rgb]{0,0,0},u(t,\omega)}) contains a strictly positive element, we have the inequality |t1t0δ(t,ω,u(t,ω))|<|t1t0|\left|{{\color[rgb]{0,0,0}t_{1}}-t_{0}-\delta(t,\omega{\color[rgb]{0,0,0},u(t,\omega)})}\right|<\left|{{\color[rgb]{0,0,0}t_{1}}-t_{0}}\right| and hence δ(t,ω,u(t,ω))>0\delta(t,\omega{\color[rgb]{0,0,0},u(t,\omega)})>0. Thus, by Lemma V.3 in [Subbotina06], τ~0:=τ0+δ[]\tilde{\tau}_{0}:=\tau_{0}+\delta[\cdot] is an 𝔽{\mathbb{F}}-stopping time with τ~0>τ0\tilde{\tau}_{0}>\tau_{0} on {τ0<t1}\{\tau_{0}<t_{1}\}. Finally, since

𝔼[u(τ~0,X+Aa(t0)Aa(t0))+t0τ~0(s,a)ds]\displaystyle{{\mathbb{E}}}\left[u(\tilde{\tau}_{0},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{t_{0}}^{\tilde{\tau}_{0}}\ell(s,a)\,ds\right]
=Ω(𝔼[u(τ~0,X+Aa(t0)Aa(t0))+τ0τ~0(s,a)ds|τ0](ω)\displaystyle=\int_{\Omega}\Biggl{(}{{\mathbb{E}}}\left[u(\tilde{\tau}_{0},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{\tau_{0}}^{\tilde{\tau}_{0}}\ell(s,a)\,ds\Bigg{|}{\mathcal{F}}_{\tau_{0}}\right](\omega)
+t0τ0(ω)(s,a)ds)P(dω)\displaystyle\qquad\qquad+\int_{t_{0}}^{\tau_{0}(\omega)}\ell(s,a)\,ds\Biggr{)}\,P(d\omega)
=Ω(𝔼τ0(ω),ω,n[u(τ~0(ω),X+Aa(t0)Aa(t0))+τ0(ω)τ~0(ω)(s,a)ds]\displaystyle=\int_{\Omega}\Biggl{(}{{\mathbb{E}}}_{\tau_{0}(\omega),\omega,n}\left[u(\tilde{\tau}_{0}({\omega}),X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{\tau_{0}(\omega)}^{\tilde{\tau}_{0}(\omega)}\ell(s,a)\,ds\right]
+t0τ0(ω)(s,a)ds)P(dω)\displaystyle\qquad\qquad+\int_{t_{0}}^{\tau_{0}(\omega)}\ell(s,a)\,ds\Biggr{)}\,P(d\omega)
=Ω(𝔼τ0(ω),ω+Aa((t0)τ0(ω))Aa(t0),n\displaystyle=\int_{\Omega}\Biggl{(}{{\mathbb{E}}}_{\tau_{0}(\omega),\omega+A^{a}{\color[rgb]{0,0,0}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))}-A^{a}(t_{0}),n}
[u(τ~0(ω),X+Aa(τ0(ω))Aa(τ0(ω)))+τ0(ω)τ~0(ω)(s,a)ds]\displaystyle\qquad\qquad\Biggl{[}u(\tilde{\tau}_{0}({\omega}),X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee\tau_{0}(\omega))}-A^{a}(\tau_{0}(\omega)))+\int_{\tau_{0}(\omega)}^{\tilde{\tau}_{0}(\omega)}\ell(s,a)\,ds\Biggr{]}
+t0τ0(ω)(s,a)ds)P(dω)\displaystyle\qquad\qquad\qquad\qquad+\int_{t_{0}}^{\tau_{0}(\omega)}\ell(s,a)\,ds\Biggr{)}\,P(d\omega)

and since, for each ωΩ\omega\in\Omega, thanks to δ[ω]M(τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0)),u(τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0)))\delta[\omega]\in M(\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0})){\color[rgb]{0,0,0},u(\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0}))}) and thanks to the definition of MM,

𝔼τ0(ω),ω+Aa((t0)τ0(ω))Aa(t0),n[u(τ~0(ω),X+Aa(τ0(ω))Aa(τ0(ω)))\displaystyle\color[rgb]{0,0,0}{{\mathbb{E}}}_{\tau_{0}(\omega),\omega+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0}),n}\Biggl{[}u(\tilde{\tau}_{0}({\omega}),X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee\tau_{0}(\omega))}-A^{a}(\tau_{0}(\omega)))
+τ0(ω)τ~0(ω)(s,a)ds]\displaystyle\color[rgb]{0,0,0}\qquad\qquad\qquad\qquad+\int_{\tau_{0}(\omega)}^{\tilde{\tau}_{0}(\omega)}\ell(s,a)\,ds\Biggr{]}
=𝔼τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0),n\displaystyle\color[rgb]{0,0,0}={{\mathbb{E}}}_{\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0}),n}
[u(τ0(ω)+δ[ω],X+Aa(τ0(ω))Aa(τ0(ω)))\displaystyle\qquad\Biggl{[}u(\tau_{0}({\omega})+\delta[\omega],X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee\tau_{0}(\omega))}-A^{a}(\tau_{0}(\omega)))
+τ0(ω)τ0(ω)+δ[ω](s,a)ds]\displaystyle\color[rgb]{0,0,0}\qquad\qquad\qquad\qquad+\int_{\tau_{0}(\omega)}^{{\tau}_{0}(\omega)+\delta[\omega]}\ell(s,a)\,ds\Biggr{]}
u(τ0(ω),ω(τ0(ω))+Aa((t0)τ0(ω))Aa(t0))+δ[ω]t1t0(ε)\displaystyle\color[rgb]{0,0,0}\geq u(\tau_{0}(\omega),\omega(\cdot\wedge\tau_{0}(\omega))+A^{a}((\cdot\vee t_{0})\wedge\tau_{0}(\omega))-A^{a}(t_{0}))+\frac{\delta[\omega]}{t_{1}-t_{0}}\cdot(-{\varepsilon})
=u(τ0(ω),ω+Aa(t0)Aa(t0))+δ[ω]t1t0(ε),\displaystyle\color[rgb]{0,0,0}=u(\tau_{0}(\omega),\omega+A^{a}(\cdot\vee t_{0})-A^{a}(t_{0}))+\frac{\delta[\omega]}{t_{1}-t_{0}}\cdot(-{\varepsilon}),

we have

𝔼[u(τ~0,X+Aa(t0)Aa(t0))+t0τ~0(s,a)ds]\displaystyle{\color[rgb]{0,0,0}{{\mathbb{E}}}\left[u(\tilde{\tau}_{0},X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))+\int_{t_{0}}^{\tilde{\tau}_{0}}\ell(s,a)\,ds\right]}
Ω(u(τ0(ω),ω+Aa(t0)Aa(t0))δ[ω]t1t0ε+t0τ0(ω)(s,a)ds)P(dω)\displaystyle\geq\int_{\Omega}\Biggl{(}u(\tau_{0}(\omega),\omega+A^{a}({\color[rgb]{0,0,0}\cdot\vee t_{0}})-A^{a}(t_{0}))-\frac{\delta[\omega]}{t_{1}-t_{0}}{\varepsilon}+\int_{t_{0}}^{\tau_{0}(\omega)}\ell(s,a){\color[rgb]{0,0,0}\,ds}\Biggr{)}\,P(d\omega)
u(t0,x0)+𝔼[δ[]+τ0]t0t1t0(ε),\displaystyle\geq u(t_{0},x_{0})+\frac{{{\mathbb{E}}}\left[\delta[\cdot]+\tau_{0}\right]-t_{0}}{t_{1}-t_{0}}\cdot(-{\varepsilon}),

where the last inequality follows from (7.3) with τ=τ0\tau=\tau_{0}. We can conclude that τ~0S\tilde{\tau}_{0}\in S. Since τ~0τ0\tilde{\tau}_{0}\geq\tau_{0} and P({τ~0>τ0}{τ0<t1})>0P(\{\tilde{\tau}_{0}>\tau_{0}\}\cap\{\tau_{0}<t_{1}\})>0, we have τ0τ~0\tau_{0}\preccurlyeq\tilde{\tau}_{0} but not τ0=τ~0\tau_{0}=\tilde{\tau}_{0}, PP-a.s., i.e., we have a contradiction to τ0\tau_{0} being the maximal element in SS. Hence, τ0=t1\tau_{0}=t_{1}, PP-a.s. However, this in turn contradicts (7.2), which concludes the proof of this direction.

(ii) Showing the remaining direction is straight-forward. ∎

Theorem 7.2.

Assume (H1) and (H2). Fix nn\in{\mathbb{N}}. Let hh be u.s.c. and bounded. Let \ell be continuous and finite-valued. Then vnv_{n} is the unique bounded maximal Dini subsolution of (TVP nn). Moreover, each bounded Dini subsolution of (TVP nn) is dominated from above by vnv_{n} even if we dispense with the assumption that hh is u.s.c.

Proof.

(i) Existence and regularity: First, we show that vnv_{n} is bounded and u.s.c. Boundedness of hh, (H1), and (H2) yield boundedness of vnv_{n}. To establish upper semi-continuity, we will use \odot defined by (6.3) and the notation (6.4). For this reason, we assume that T=1T=1 and n=1n=1 but this assumption is not restrictive. We continue by fixing a pair (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega and considering a sequence (tk,xk)k1(t_{k},x_{k})_{k\geq 1} in [0,T]×Ω[0,T]\times\Omega that converges to (t0,x0)(t_{0},x_{0}) and satisfies limkvn(tk,xk)=lim¯(t,x)(t0,x0)vn(t,x)\lim_{k}v_{n}(t_{k},x_{k})=\varlimsup_{(t,x)\to(t_{0},x_{0})}v_{n}(t,x). We distinguish between two cases.

Case 1: t0<1t_{0}<1. Fix an ε<0{\varepsilon}<0 such that t0+ε<1t_{0}+{\varepsilon}<1. We can assume, without loss of generality, that tk<t0+εt_{k}<t_{0}+{\varepsilon} for all kk\in{\mathbb{N}}. Let aba\in\mathcal{L}_{b}. Then the map (see (6.4))

(s,t,ω)(s)(t,a(t,ω))=(1s)(s+t(1s),a(t,ω)1s)\displaystyle(s,t,\omega)\mapsto\ell^{(s)}(t,a(t,\omega))=(1-s)\,\ell\left(s+t(1-s),\frac{a(t,\omega)}{\sqrt{1-s}}\right)

from [0,t0+ε]×[0,T]×Ω[0,t_{0}+{\varepsilon}]\times[0,T]\times\Omega to {\mathbb{R}} has a real upper bound due to \ell being continuous and aa being bounded. Note that (t,x)(xtω)(t,x)\mapsto(x\odot_{t}\omega), [0,T)×ΩΩ[0,T)\times\Omega\to\Omega, is continuous for every ωΩ\omega\in\Omega with ω(0)=0\omega(0)=0. Thus, by Lemma 6.3,

limkvn(tk,xk)lim¯k𝔼0,0,n[h(xktk[X+Aa])+0T(tk)(t,a(t))𝑑t]\displaystyle\lim_{k}v_{n}(t_{k},x_{k})\leq{\color[rgb]{0,0,0}\varlimsup_{k}{{\mathbb{E}}}_{0,0,n}\left[h(x_{k}\odot_{t_{k}}[X+A^{a}])+\int_{0}^{T}\ell^{(t_{k})}(t,a(t))\,dt\right]}
𝔼0,0,n[h(x0t0[X+Aa])+0T(t0)(t,a(t))𝑑t].\displaystyle\leq{{\mathbb{E}}}_{0,0,n}\left[h(x_{0}\odot_{t_{0}}[X+A^{a}])+\int_{0}^{T}{\color[rgb]{0,0,0}\ell^{(t_{0})}(t,a(t))}\,dt\right].

Since aba\in\mathcal{L}_{b} was arbitrary, we can deduce after again invoking Lemma 6.3 that limkvn(tk,xk)vn(t0,x0)\lim_{k}v_{n}(t_{k},x_{k})\leq v_{n}(t_{0},x_{0}).

Case 2: t0=1t_{0}=1. Then, proceeding similarly as in Case 1 but using the constant control (t,ω)a(t,ω)=0(t,\omega)\mapsto a(t,\omega)=0, we have

limkvn(tk,xk)lim¯k𝔼0,0,n[h(xktk[X+Aa])+0T(tk)(t,0)𝑑t]\displaystyle\color[rgb]{0,0,0}\lim_{k}v_{n}(t_{k},x_{k})\leq{\color[rgb]{0,0,0}\varlimsup_{k}{{\mathbb{E}}}_{0,0,n}\left[h(x_{k}\odot_{t_{k}}[X+A^{a}])+\int_{0}^{T}\ell^{(t_{k})}(t,0)\,dt\right]}
h(x0)=vn(t0,x0).\displaystyle\color[rgb]{0,0,0}\leq h(x_{0})=v_{n}(t_{0},x_{0}).

We can conclude that vnv_{n} is u.s.c.

Next, we establish the subsolution property via the BSDE connection in Theorem 6.4. To this end, we use the notation (6.2) and fix (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega, t[t0,T]t\in[t_{0},T], and ada\in{\mathbb{R}}^{d}. Note first that, by Proposition 3.6 (1) in [DrapeauEtAl13AOP],

(7.4) t0,Tt0,x0,n(h)=t0,tt0,x0,n(t,Tt0,x0,n(h)).\displaystyle\mathcal{E}^{t_{0},x_{0},n}_{t_{0},T}(-h)=\mathcal{E}^{t_{0},x_{0},n}_{t_{0},t}\left(\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h)\right).

Since vnv_{n} is bounded and, by Theorem 6.4, vn(t,)=t,Tt0,x0,n(h)v_{n}(t,\cdot)=-\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h), Pt0,x0,nP_{t_{0},x_{0},n}-a.s., we can apply Theorem 3.4 in [DrapeauEtAl16AIHP] (see also Lemma A.2 in [BLT]) together with (3) in [BLT] to deduce that, with Qt0,x0,na:=t0,x0,n(X+Aa(t0)Aa(t0))1Q^{a}_{t_{0},x_{0},n}:={{\mathbb{P}}}_{t_{0},x_{0},n}\circ(X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))^{-1},

𝔼t0,x0,n[t0,tt0,x0,n(t,Tt0,x0,n(h))]𝔼Qt0,x0,na[t,Tt0,x0,n(h)t0t(s,a)𝑑s]\displaystyle{{\mathbb{E}}}_{t_{0},x_{0},n}\left[\mathcal{E}^{t_{0},x_{0},n}_{t_{0},t}\left(\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h)\right)\right]\geq{{\mathbb{E}}}^{Q^{a}_{t_{0},x_{0},n}}\left[\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h)-\int_{t_{0}}^{t}\ell(s,a)\,ds\right]
=𝔼t0,x0,n[t,Tt0,x0,n(h)(X+Aa(t0)Aa(t0))t0t(s,a)ds].\displaystyle={{\mathbb{E}}}_{t_{0},x_{0},n}\left[\mathcal{E}^{t_{0},x_{0},n}_{t,T}(-h)(X+A^{a}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{a}(t_{0}))-\int_{t_{0}}^{t}\ell(s,a)\,ds\right].

Thus, by (7.4) and Theorem 6.4, vnv_{n} satisfies (7.1) in place of uu. Hence, by Lemma 7.1, vnv_{n} is a Dini subsolution of (TVP nn).

(ii) Uniqueness: Let uu be a bounded Dini subsolution of (TVP nn). Without loss of generality, let n=1n=1. Fix (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega and ε>0{\varepsilon}>0. It suffices to show that u(t0,x0)vn(t0,x0)+εu(t_{0},x_{0})\leq{\color[rgb]{0,0,0}v_{n}}(t_{0},x_{0})+{\varepsilon}. By (a slight modification) of (36) and (BBD), both in [BLT], there exists an abt0a\in\mathcal{L}_{b}^{t_{0}} with a(t,ω)=i=1mai(ω).1[si1,si)(t)a(t,\omega)=\sum_{i=1}^{m}a_{i}(\omega).{\mathbf{1}}_{[s_{i-1},s_{i})}(t), where t0=s0<s1<<sm=Tt_{0}=s_{0}<s_{1}<\cdots<s_{m}=T and each ai:Ωda_{i}:\Omega\to{\mathbb{R}}^{d} is si1{\mathcal{F}}_{s_{i-1}}-measurable, such that 𝔼~t0,x0[t0T(t,a(t))𝑑t+h(X)]vn(t0,x0)+ε\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[\int_{t_{0}}^{T}\ell(t,a(t))\,dt+h(X)\right]\leq v_{n}(t_{0},x_{0})+{\varepsilon}. Here, for any (s,x)[t0,T]×Ω(s,x)\in[t_{0},T]\times\Omega, P~s,x:=Ps,x(X~s,x)1\tilde{P}_{s,x}:=P_{s,x}\circ(\tilde{X}^{s,x})^{-1}, 𝔼~s,x:=𝔼P~s,x\tilde{{{\mathbb{E}}}}_{{\color[rgb]{0,0,0}s,x}}:={{\mathbb{E}}}^{\tilde{P}^{s,x}}, and X~s,x:[0,T]×Ωd\tilde{X}^{s,x}:[0,T]\times\Omega\to{\mathbb{R}}^{d} is the unique solution of X~s,x(t)=X~s,x(s)+sta(r,X~s,x)𝑑r+X(t)X(s)\tilde{X}^{s,x}(t)=\tilde{X}^{s,x}(s)+\int_{s}^{t}a(r,\tilde{X}^{s,x})\,dr+X(t)-X(s), t[s,T]t\in[s,T], with initial condition X~s,x|[0,s]=x|[0,s]\tilde{X}^{s,x}|_{[0,s]}=x|_{[0,s]}. Next, note that a1X~t0,x0=a1x0=:a~1da_{1}\circ\tilde{X}^{t_{0},x_{0}}=a_{1}\circ x_{0}=:\tilde{a}_{1}\in{\mathbb{R}}^{d} and X+Aa~1(t0)Aa~1(t0)=X~t0,x0X+A^{\tilde{a}_{1}}{\color[rgb]{0,0,0}(\cdot\vee t_{0})}-A^{\tilde{a}_{1}}(t_{0})=\tilde{X}^{t_{0},x_{0}} on [0,s1][0,s_{1}], Pt0,x0P_{t_{0},x_{0}}-a.s. Thus, by Lemma 7.1,

u(t0,x0)\displaystyle u(t_{0},x_{0}) 𝔼t0,x0[t0s1(t,a~1)𝑑t+u(s1,X~t0,x0)]\displaystyle\leq{{\mathbb{E}}}_{t_{0},x_{0}}\left[\int_{t_{0}}^{s_{1}}\ell(t,\tilde{a}_{1})\,dt+u(s_{1},\tilde{X}^{t_{0},x_{0}})\right]
=𝔼~t0,x0[t0s1(t,a(t))𝑑t+u(s1,X)].\displaystyle=\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[\int_{t_{0}}^{s_{1}}\ell(t,a(t))\,dt+u(s_{1},X)\right].

Also note that, for Pt0,x0P_{t_{0},x_{0}}-a.e. ωΩ\omega\in\Omega, we have, with x1=X~t0,x0(ω)x_{1}=\tilde{X}^{t_{0},x_{0}}(\omega), a2X~s1,x1=a2x1da_{2}\circ\tilde{X}^{s_{1},x_{1}}=a_{2}\circ x_{1}\in{\mathbb{R}}^{d} and X+Aa2x1(s1)Aa2x1(s1)=X~s1,x1X+A^{a_{2}\circ x_{1}}{\color[rgb]{0,0,0}(\cdot\vee s_{1})}-A^{a_{2}\circ x_{1}}(s_{1})=\tilde{X}^{s_{1},x_{1}} on [0,s2][0,s_{2}], Ps1,x1P_{s_{1},x_{1}}-a.s., and thus, by Lemma 7.1,

u(s1,x1)\displaystyle u(s_{1},x_{1}) 𝔼s1,x1[s1s2(t,a~2)𝑑t+u(s2,X~s1,x1)]\displaystyle\leq{{\mathbb{E}}}_{s_{1},x_{1}}\left[\int_{s_{1}}^{s_{2}}\ell(t,\tilde{a}_{2})\,dt+u(s_{2},\tilde{X}^{s_{1},x_{1}})\right]
=𝔼~s1,x1[s1s2(t,a(t))𝑑t+u(s2,X)].\displaystyle=\tilde{{{\mathbb{E}}}}_{s_{1},x_{1}}\left[\int_{s_{1}}^{s_{2}}\ell(t,a(t))\,dt+u(s_{2},X)\right].

Therefore

𝔼~t0,x0[u(s1,X)]\displaystyle\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[u(s_{1},X)\right] Ω𝔼~s1,x1[s1s2(t,a(t))𝑑t+u(s2,X)]|x1=ωP~t0,x0(dω)\displaystyle\leq\int_{\Omega}\tilde{{{\mathbb{E}}}}_{s_{1},x_{1}}\left[\int_{s_{1}}^{s_{2}}\ell(t,a(t))\,dt+u(s_{2},X)\right]\Bigg{|}_{x_{1}=\omega}\,\tilde{P}_{t_{0},x_{0}}(d\omega)
=𝔼~t0,x0(𝔼~t0,x0[s1s2(t,a(t))𝑑t+u(s2,X)|s1])\displaystyle=\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left(\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[\int_{s_{1}}^{s_{2}}\ell(t,a(t))\,dt+u(s_{2},X)\Bigg{|}{\mathcal{F}}_{s_{1}}\right]\right)
=𝔼~t0,x0[s1s2(t,a(t))𝑑t+u(s2,X)].\displaystyle=\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[\int_{s_{1}}^{s_{2}}\ell(t,a(t))\,dt+u(s_{2},X)\right].

Repeating this procedure and noting that u(T,)hu(T,\cdot)\leq h yields

u(t0,x0)𝔼~t0,x0[t0T(t,a(t))𝑑t+h(X)]vn(t0,x0)+ε,\displaystyle u(t_{0},x_{0})\leq\tilde{{{\mathbb{E}}}}_{t_{0},x_{0}}\left[\int_{t_{0}}^{T}\ell(t,a(t))\,dt+h(X)\right]\leq v_{n}(t_{0},x_{0})+{\varepsilon},

which concludes the proof. ∎

8. The first order HJB equation

Lemma 8.1.

Assume (H1). An l.s.c. function u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\} bounded from below is a minimax supersolution of (TVP) if and only if u(T,)hu(T,\cdot)\geq h and, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega and t(t0,T]t\in(t_{0},T], there exists an x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}) such that

(8.1) u(t0,x0)t0t(s,x(s))𝑑s+u(t,x).\displaystyle u(t_{0},x_{0})\geq\int_{t_{0}}^{t}\ell(s,x^{\prime}(s))\,ds+u(t,x).
Proof.

(i) Let uu be a minimax supersolution of (TVP). Assume that there exist a (t0,x0)dom(u)(t_{0},x_{0})\in\mathrm{dom}(u) with t0<Tt_{0}<T and a t1(t0,T]t_{1}\in(t_{0},T] such that, for all x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}), u(t1,x)u(t0,x0)+t0t1(s,x(s))𝑑s>0u(t_{1},x)-u(t_{0},x_{0})+\int_{t_{0}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds>0. Then there is an ε>0{\varepsilon}>0 such that

(8.2) u(t1,x)u(t0,x0)+t0t1(s,x(s))𝑑s>ε\displaystyle u(t_{1},x)-u(t_{0},x_{0})+\int_{t_{0}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds>{\varepsilon}

for all x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}) because infx𝒳1,1(t0,x0)[t0t1(s,x(s))𝑑s+u(t1,x)]\inf_{x\in\mathcal{X}^{1,1}(t_{0},x_{0})}\left[\int_{t_{0}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds+u(t_{1},x)\right] has a minimizer in {}{\mathbb{R}}\cup\{\infty\} (this can be shown exactly as the corresponding statement for the non-path-dependent counterpart Theorem 11.1.i in Section 11.1 of [Cesari] and its extension to unbounded domains, which is treated in Section 11.2 of [Cesari]). Next, consider the (non-void) set SS of all (t,x)[t0,t1)×𝒳1,1(t0,x0)(t,x)\in[t_{0},t_{1})\times\mathcal{X}^{1,1}(t_{0},x_{0}) for which u(t,x)u(t0,x0)+t0t(s,x(s))𝑑stt0t1t0εu(t,x)-u(t_{0},x_{0})+\int_{t_{0}}^{t}\ell(s,x^{\prime}(s))\,ds\leq\frac{t-t_{0}}{t_{1}-t_{0}}\cdot{\varepsilon} holds. Let s~:=sup{t:(t,x)S for some x}\tilde{s}:=\sup\{t:(t,x)\in S\text{ for some $x$}\}. This supremum is attained. To see this, consider a sequence (sn,xn)n(s_{n},x_{n})_{n} in SS with sns~s_{n}\to\tilde{s}. Since supnt0sn(s,xn(s))𝑑s<\sup_{n}\int_{t_{0}}^{s_{n}}\ell(s,x_{n}^{\prime}(s))\,ds<\infty and uu is bounded from below, one can show as in Sections 11.1 and 11.2 of [Cesari] that there is a subsequence (snk,xnk)k(s_{n_{k}},x_{n_{k}})_{k} such that (xnk)(x_{n_{k}}) converges to some x~\tilde{x} in 𝒳1,1(t0,x0)\mathcal{X}^{1,1}(t_{0},x_{0}), i.e., xnkx~+xnkx~L1(t0,T;d)0\lVert{x_{n_{k}}-\tilde{x}}\rVert_{\infty}+\lVert{x_{n_{k}}^{\prime}-\tilde{x}^{\prime}}\rVert_{L^{1}(t_{0},T;{\mathbb{R}}^{d})}\to 0. Lower semi-continuity of uu together with Theorem 10.8.ii in [Cesari] yield

u(s~,x~)+t0s~(s,x~(s))𝑑s\displaystyle u(\tilde{s},\tilde{x})+\int_{t_{0}}^{\tilde{s}}\ell(s,\tilde{x}^{\prime}(s))\,ds lim¯ku(snk,xnk)+lim¯kt0snk(s,xnk(s))𝑑s\displaystyle\leq\varliminf_{k}u(s_{n_{k}},x_{n_{k}})+\varliminf_{k}\int_{t_{0}}^{s_{n_{k}}}\ell(s,x_{n_{k}}^{\prime}(s))\,ds
s~t0t1t0ε+u(t0,x0).\displaystyle\leq\frac{\tilde{s}-t_{0}}{t_{1}-t_{0}}\cdot{\varepsilon}+u(t_{0},x_{0}).

Thus (s~,x~)dom(u)(\tilde{s},\tilde{x})\in\mathrm{dom}(u) and, by (8.2), s~<t1\tilde{s}<t_{1}. Hence, by (3.4), there is a δ>0\delta>0 and an x𝒳1,1(s~,x~)x\in\mathcal{X}^{1,1}(\tilde{s},\tilde{x}) such that u(s~+δ,x)u(s~,x~)+s~s~+δ(s,x(s))𝑑sδt1t0εu(\tilde{s}+\delta,x)-u(\tilde{s},\tilde{x})+\int_{\tilde{s}}^{\tilde{s}+\delta}\ell(s,x^{\prime}(s))\,ds\leq\frac{\delta}{t_{1}-t_{0}}\cdot{\varepsilon} and s~+δ<t1\tilde{s}+\delta<t_{1}. Thus (s~+δ,x)S(\tilde{s}+\delta,x)\in S, which is a contradiction to the maximality of s~\tilde{s}.

(ii) Fix (t0,x0)dom(u)(t_{0},x_{0})\in\mathrm{dom}(u) with t0<Tt_{0}<T. Suppose that, for every (t~0,x~0)[t0,T)×𝒳1,1(t0,x0)(\tilde{t}_{0},\tilde{x}_{0})\in[t_{0},T)\times\mathcal{X}^{1,1}(t_{0},x_{0}) for every t(t~0,T]t\in(\tilde{t}_{0},T], there exists an x𝒳1,1(t~0,x~0)x\in\mathcal{X}^{1,1}(\tilde{t}_{0},\tilde{x}_{0}) such that (8.1) holds with (t0,x0)(t_{0},x_{0}) replaced by (t~0,x~0)(\tilde{t}_{0},\tilde{x}_{0}). Then one can proceed similarly as in the proof of Lemma 3.6 in [BK18JFA] to show that there exists a sequence (xn)(x_{n}) in 𝒳1,1(t0,x0)\mathcal{X}^{1,1}(t_{0},x_{0}) and an increasing sequence (An)(A_{n}) of finite subsets of [t0,T][t_{0},T] whose union AA is dense in [t0,T][t_{0},T] such that, for each nn\in{\mathbb{N}} and every tAnt\in A_{n}, (8.1) holds with xx replaced by xnx_{n}. Thus suptA,nt0t(s,xn(s))𝑑su(t0,x0)+c<\sup_{t\in A,n\in{\mathbb{N}}}\int_{t_{0}}^{t}\ell(s,x_{n}^{\prime}(s))\,ds\leq u(t_{0},x_{0})+c<\infty, where c-c is a lower bound of uu. Note that we are in a similar situation as in part (i) of this proof and thus it can be shown in the same way that there is a subsequence (xnk)(x_{n_{k}}) of (xn)(x_{n}) that converges to some x~\tilde{x} in 𝒳1,1(t0,x0)\mathcal{X}^{1,1}(t_{0},x_{0}). Now, fix t(t0,T]t\in(t_{0},T] and a sequence (sk)(s_{k}) in AA with skAnks_{k}\in A_{n_{k}} for each kk and with skts_{k}\to t as kk\to\infty. By Theorem 10.8.ii in [Cesari] and lower semi-continuity of uu,

u(t0,x0)\displaystyle u(t_{0},x_{0}) lim¯k[t0sk(s,xnk(s))ds+u(sk,xnk)]t0t(s,x~(s)ds+u(t,x~),\displaystyle\geq\varliminf_{k}\left[\int_{t_{0}}^{s_{k}}\ell(s,x_{n_{k}}^{\prime}(s))\,ds+u(s_{k},x_{n_{k}})\right]\geq\int_{t_{0}}^{t}\ell(s,\tilde{x}^{\prime}(s)\,ds+u(t,\tilde{x}),

i.e., there is an x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}) such that, for every t[t0,T]t\in[t_{0},T], (8.1) holds. From this point, (3.4) follows easily. ∎

Lemma 8.2.

Assume (H1). An l.s.c. function u:[0,T]×Ω{}u:[0,T]\times\Omega\to{\mathbb{R}}\cup\{\infty\} is an l.s.c. minimax subsolution of (TVP) if and only if u(T,)hu(T,\cdot)\leq h and, for every (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega, t(t0,T]t\in(t_{0},T], and x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}), we have

(8.3) u(t0,x0)t0t(s,x(s))𝑑s+u(t,x).\displaystyle u(t_{0},x_{0})\leq\int_{t_{0}}^{t}\ell(s,x^{\prime}(s))\,ds+u(t,x).
Proof.

(i) Let uu be an l.s.c. minimax subsolution of (TVP). For the sake of a contradiction, assume that there exist a (t0,x0)[0,T)×Ω(t_{0},x_{0})\in[0,T)\times\Omega, a t1(t0,T]t_{1}\in(t_{0},T], an x𝒳1,1(t0,x0)x\in\mathcal{X}^{1,1}(t_{0},x_{0}), and an ε>0{\varepsilon}>0 such that

(8.4) u(t0,x0)u(t1,x)t0t1(s,x(s))𝑑s>ε\displaystyle u(t_{0},x_{0})-u(t_{1},x)-\int_{t_{0}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds>{\varepsilon}

as well as (t1,x)dom(u)(t_{1},x)\in\mathrm{dom}(u) and tt1(s,x(s))𝑑s<\int_{t}^{t_{1}}\ell(s,x^{\prime}(s))\,ds<\infty for all t[t0,t1]t\in[t_{0},t_{1}]. Put

s~:=inf{t(t0,t1]:u(t,x)u(t1,x)tt1(s,x(s))𝑑st1tt1t0ε}.\displaystyle\tilde{s}:=\inf\left\{t\in(t_{0},t_{1}]:u(t,x)-u(t_{1},x)-\int_{t}^{t_{1}}\ell(s,x^{\prime}(s))\,ds\leq\frac{t_{1}-t}{t_{1}-t_{0}}\cdot{\varepsilon}\right\}.

We show that this infimum is attained. To this end, consider a sequence (sn)(s_{n}) in (t0,t1](t_{0},t_{1}] with sns~s_{n}\to\tilde{s} and u(sn,x)u(t1,x)snt1(s,x(s))𝑑st1snt1t0εu(s_{n},x)-u(t_{1},x)-\int_{s_{n}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds\leq\frac{t_{1}-s_{n}}{t_{1}-t_{0}}\cdot{\varepsilon}. Then

u(s~,x)s~t1(s,x(s))𝑑slim¯nu(sn,x)limnsnt1(s,x(s))𝑑s\displaystyle u(\tilde{s},x)-\int_{\tilde{s}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds\leq\varliminf_{n}u(s_{n},x)-\lim_{n}\int_{s_{n}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds
lim¯n[u(sn,x)snt1(s,x(s))𝑑s]t1s~t1t0ε+u(t1,x),\displaystyle\leq\varliminf_{n}\left[u(s_{n},x)-\int_{s_{n}}^{t_{1}}\ell(s,x^{\prime}(s))\,ds\right]\leq\frac{t_{1}-\tilde{s}}{t_{1}-t_{0}}\cdot{\varepsilon}+u(t_{1},x),

i.e., s~\tilde{s} is a minimum and (s~,x)dom(u)(\tilde{s},x)\in\mathrm{dom}(u). Moreover, by (8.4), t0<s~t1t_{0}<\tilde{s}\leq t_{1}. Finally, by (3.5), there is a δ(0,s~t0]\delta\in(0,\tilde{s}-t_{0}] such that u(s~δ,x)u(s~,x)s~δs~(s,x(s))𝑑sδt1t0εu(\tilde{s}-\delta,x)-u(\tilde{s},x)-\int_{\tilde{s}-\delta}^{\tilde{s}}\ell(s,x^{\prime}(s))\,ds\leq\frac{\delta}{t_{1}-t_{0}}\cdot{\varepsilon}. Hence, u(s~δ,x)u(t1,x)s~δt1(s,x(s))𝑑st1(s~δ)t1t0εu(\tilde{s}-\delta,x)-u(t_{1},x)-\int_{\tilde{s}-\delta}^{t_{1}}\ell(s,x^{\prime}(s))\,ds\leq\frac{t_{1}-(\tilde{s}-\delta)}{t_{1}-t_{0}}\cdot{\varepsilon}, which is a contradiction to the minimality of s~\tilde{s}.

(ii) Showing the remaining direction is straight-forward. ∎

Proof of Theorem 4.3.

(a) Establishing the lower semi-continuity of v0v_{0} is quite standard. It is very similar to the proof of Lemma 5.2 and actually slightly easier as no probability is involved (cf. also Proposition 3.1 in [DM-F00] for the non-path-dependent case). To deduce that v0v_{0} is an l.s.c. minimax solution, it suffices to apply the dynamic programming principle with the existence of a minimizer for (DOC)(\text{DOC}) (cf.  Theorem 11.1.i in [Cesari]) together with Lemmata 8.1 and 8.2. Finally, we can apply Lemmata 8.1 and 8.2 again to obtain a comparison principle between l.s.c. minimax subsolutions and minimax supersolutions, from which uniqueness follows.

(b) Taking Remark 5.4 into account, one can see that the proof follows essentially from the content of Section 7. The measures t0,x0,n{{\mathbb{P}}}_{t_{0},x_{0},n} need to be replaced by the Dirac measures under which X=x(t0)X=x(\cdot\wedge t_{0}) a.s. for each (t0,x0)[0,T]×Ω(t_{0},x_{0})\in[0,T]\times\Omega and one should note that the domains of the controls are different ([0,T][0,T] here vs. [0,T]×Ω[0,T]\times\Omega in Section 7). Moreover, the BSDE argument in part (i) of the proof of Theorem 7.2 needs to be replaced by the deterministic dynamic programming principle. ∎

9. Conclusion

The main contributions of this paper are a non-Markovian vanishing viscosity result for path-dependent PDEs (PPDEs) that corresponds to the non-exponential Schilder theorem in [BLT], well-posedness for new notions of generalized solutions of PPDEs that can have quadratic or even super-quadratic growth in the gradient, and a non-Markovian Feynman-Kac formula for convex superquadratic BSDEs.

We want to emphasize that, here, control-theoretic methods, or equivalently (in our case), results from the theory of large deviations have been applied to obtain stability results for PPDEs (corresponding PDEs results have been obtained in [BLT] in the Markovian case). Of great interest would be research that investigates the opposite direction, i.e., to establish stronger results (in particular, suitable stability results) for PPDEs with (only) quadratic growth in the gradient in order to derive non-Markovian large deviation results similarly as it has been successfully done in the Markovian case via PDEs.

References