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

Lifting of Volterra processes:
optimal control in UMD Banach spaces

Giulia di Nunno†,∗ and Michele Giordano
June 22nd 2023
Abstract.

We study a stochastic control problem for a Volterra-type controlled forward equation with past dependence obtained via convolution with a deterministic kernel. To be able to apply dynamic programming to solve the problem, we lift it to infinite dimensions and we formulate a UMD Banach-valued Markovian problem, which is shown to be equivalent to the original finite-dimensional non-Markovian one. We characterize the optimal control for the infinite dimensional problem and show that this also characterizes the optimal control for the finite dimensional problem.

Department of Mathematics, University of Oslo, P:O: Box 1053 Blindern, N-0316 Oslo, Email: giulian@math.uio.no.Department of Business and Management Science, NHH Norwegian School of Economics, Helleveien 30, N-5045 Bergen.Department of Mathematics, University of Oslo, P:O: Box 1053 Blindern, N-0316 Oslo, Email: michelgi@math.uio.no

Keywords: Backward stochastic integral equation; Dynamic programming principle;
Hamilton Jacobi Bellman; Optimal control; UMD Banach space; Markovian Lift;
MSC 2020: 60H10; 60H20; 93E20; 35R15; 49L20; 91B70;

1. Introduction

We intend to minimize a performance functional of the form

J(t,x,u)=𝔼[tTF(τ,Xτu,uτ)𝑑τ+G(XTu)],J(t,x,u)=\mathbb{E}\left[\int_{t}^{T}F(\tau,X^{u}_{\tau},u_{\tau})d\tau+G(X^{u}_{T})\right], (1.1)

where t[0,T]t\in[0,T], and xx is given in the controlled Volterra-type dynamics of the process XτuX_{\tau}^{u}:

Xu(τ)\displaystyle X^{u}(\tau) =x(τ)+tτK(τs)[β(s,Xu(s))+σ(s,Xu(s))R(s,Xu(s),u(s))]𝑑s\displaystyle=x(\tau)+\int_{t}^{\tau}K(\tau-s)\Big{[}\beta(s,X^{u}(s))+\sigma(s,X^{u}(s))R(s,X^{u}(s),u(s))\Big{]}ds
+tτK(τs)σ(s,Xu(s))𝑑W(s).\displaystyle\quad+\int_{t}^{\tau}K(\tau-s)\sigma(s,X^{u}(s))dW(s). (1.2)

Here x:[0,T]x:[0,T]\longrightarrow\mathbb{R}, β:[0,T]×\beta:[0,T]\times\mathbb{R}\longrightarrow\mathbb{R}, σ:[0,T]×\sigma:[0,T]\times\mathbb{R}\longrightarrow\mathbb{R}, R:[0,T]××𝒰R:[0,T]\times\mathbb{R}\times\mathcal{U}\longrightarrow\mathbb{R}, and the convolution kernel K:[0,T]+K:[0,T]\longrightarrow\mathbb{R}^{+} are all measurable mappings on which additional hypothesis will be stated later on, and u:[0,T]×Ω𝒰u:[0,T]\times\Omega\longrightarrow\mathcal{U}\subset\mathbb{R} is an admissible control. Also, WW is a real-valued Brownian motion on a complete filtered probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}).

Stochastic control problems as (1.1)-(1.2) appear, e.g. when studying optimal advertising strategies (see e.g. [15] for the case of Volterra dynamics in (1.2) and [17, 16, 20] for the case of delay). Other cases of applications are found in electrodynamics [24] and in epidemiology [25]. When dealing with such problems, one cannot directly apply a dynamic programming principle (DPP) in view of the non Markovianity of the framework. While in some particular cases is still possible to derive the DPP also for Volterra forward dynamics (see [1, 18, 5]), most authors approached the general problem by means of a maximum principle (see, e.g., [12, 29, 30, 3, 2] and references therein). Even though the maximum principle approach might seem practical, one usually has to impose regularity conditions on both the drift and volatility which are not always easy to satisfy. In this paper, thanks to the developments on the lift theory for Volterra processes (see [1, 9, 8, 7, 15, 5]) we aim to move the stochastic control problem (1.1)-(1.2) to an infinite dimensional UMD-Banach setting and solve the newly formulated problem by means of DPP.

The main purpose of this lifting approach is to recover the Markov property for the forward process (1.2) which, in turns, allows us to derive a DPP in terms of the Hamilton-Jacobi-Bellman (HJB) equations. In fact, one can show that solving the lifted problem is equivalent to solving the original one, with the fundamental difference that, by moving to an infinite dimensional setting, we work in a Markovian framework. Focusing on Markovian lifts, we assume that the kernel KK can be represented as K(t)=g,𝒮tνY×YK(t)=\langle g,\mathcal{S}_{t}^{*}\nu\rangle_{Y\times Y^{*}}, for 𝒮t\mathcal{S}_{t}^{*} a uniformly continuous semigroup acting on a Banach space YY^{*}, νY\nu\in Y^{*}, gYg\in Y with YY the pre dual of YY^{*} and pairing ,Y×Y\langle\cdot,\cdot\rangle_{Y\times Y^{*}}. Examples of kernels that satisfy this condition can be found both in [9, 8, 15] and in the last section of this paper.

Our goal is to find u^\hat{u} such that, for all t[0,T]t\in[0,T],

J(t,x,u^)=infu𝔸J(t,x,u),J(t,x,\hat{u})=\inf_{u\in\mathds{A}}J(t,x,u), (1.3)

with J(t,x,u)J(t,x,u) as in (1.1) and for uu belonging to some admissible control set 𝔸\mathds{A} defined as

𝔸={u:[0,T]×Ω𝒰,s.t. u is predictable},\mathds{A}=\left\{u:[0,T]\times\Omega\longrightarrow\mathcal{U},\ \text{s.t. $u$ is predictable}\right\},

where 𝒰\mathcal{U} is a closed convex subset of \mathbb{R} and the information flow is associated to the Brownian motion in (1.2). Our approach consists in formulating a new infinite-dimensional Banach-valued optimization problem that can be shown to be equivalent to (1.1)-(1.2) (in the sense that the optimal control u^\hat{u} and optimal value J(t,x,u^)J(t,x,\hat{u}), t[0,T]t\in[0,T] are the same of the original one) and then solve such infinite dimensional optimization problem, which is Markovian. The solution is achieved exploiting Malliavin calculus for unconditional martingale differences (UMD) Banach spaces.

The Markovian lift to the infinite dimensional setting that we present here was originally introduced in [9], and then developed in [8] for the multi-dimensional case, and in [7] for a Lévy drivers. Our work can be seen as a generalization of the case presented in [1], in which a kernel KK that can be expressed as the Laplace transform of a measure is considered, and where the performance functional (1.1) is of linear-quadratic type. Our work differs from [1] as we are able to consider a broader class of kernels and performance functionals thanks to the nature of the lift we apply.

The present work introduces an element of novelty also with respect to infinite dimensional stochastic control. Indeed, we consider a setting which is different from both the ones presented in [14] and [22]. In [14] the authors consider a Hilbert valued forward controlled process, whereas in [22] the forward process has values in a general Banach space, but with a volatility term σ\sigma not depending on XX. Here we are able to take general volatility dynamics for the forward process and this will require to work in Banach spaces of the UMD type so to be able to apply Malliavin calculus techniques.

In the context of optimal control for lifted process, we also mention [5]. There, the authors follow an approach close to the one presented here. However, we remark that, in our framework, we are able to consider a wider class of kernel KK thanks to the nature of our lift, which allows us to work in UMD Banach spaces instead of Hilbert spaces. On the other side, in [5] the authors consider a forward equation driven by a Lévy process instead of a Brownian motion (like (1.2)). While the lift theory for Lévy-driven forward processes is available (see [7]), the optimal control of an infinite-dimensional Lévy-driven forward equation in the present setting is a topic for future research.

This paper is structured as follows: in Section 2 we present some preliminary results both on the Gâteaux differentiability in general Banach spaces and on the lift for Volterra processes. We recall the essentials on UMD Banach spaces and some results of Malliavin calculus in this framework. In Section 3 we give an existence and continuity result for the forward equation, and in Section 4 we introduce the backward equation and the Hamiltonian function associated with the lifted optimization problem. Here we present a solution method via HJB equations. To conclude, in Section 5 we present a problem of optimal consumption where we obtain a characterization of the optimal control uu via DPP.

2. Some preliminary results

We recall some useful notions and results that are used throughout the paper. Then we show how the Markovian lift is performed in the present context of stochastic control (1.1)-(1.2). Lastly, we introduce UMD Banach spaces and state some crucial results for Malliavin calculus in this setting. We refer to [14] for the results on Gâteaux derivatives and Banach spaces, to [9, 8, 7, 15] for the ones concerning Markovian lifts, and to [19, 23, 26, 21] for the results concerning UMD Banach spaces.

2.1. The class of Gâteaux differentiable functions

For a mapping F:UVF:U\longrightarrow V, with U,VU,V two Banach spaces, we say that the directional derivative at uUu\in U in the direction hUh\in U is defined as

F(u;h):=lims0F(u+sh)F(u)s,\nabla F(u;h):=\lim_{s\rightarrow 0}\frac{F(u+sh)-F(u)}{s},

whenever the limit exists in the topology of VV. The mapping FF is said to be Gâteaux differentiable at the point uu if it has directional derivative at uu in every direction and there exists an element F(u)\nabla F(u) in L(U,V)L(U,V) such that F(u;h)=F(u)h\nabla F(u;h)=\nabla F(u)h for every hUh\in U. We call F(u)\nabla F(u) the Gâteaux derivative at uu.

Definition 2.1.

A mapping F:UVF:U\longrightarrow V belongs to 𝒢1(U;V)\mathcal{G}^{1}(U;V) if it is continuous, Gâteaux differentiable for all u𝒰u\in\mathcal{U} and F:UL(U;V)\nabla F:U\longrightarrow L(U;V) is strongly continuous, i.e. the map F()h:UV\nabla F(\cdot)h:U\longrightarrow V is continuous for every hUh\in U.

Remark 2.2.

Let U,V,ZU,V,Z be three Banach spaces and F𝒢1(U,V)F\in\mathcal{G}^{1}(U,V). If G𝒢1(V,Z)G\in\mathcal{G}^{1}(V,Z), then G(F)𝒢1(U,Z)G(F)\in\mathcal{G}^{1}(U,Z) and (G(F))(u)=G(F(u))F(u)\nabla(G(F))(u)=\nabla G(F(u))\nabla F(u).

We also introduce the partial directional derivative for a mapping F:U×VZF:U\times V\longrightarrow Z, with UU, VV, ZZ Banach spaces as

uF(u,v;h):=lims0F(u+sh,v)F(u,v)h,\nabla_{u}F(u,v;h):=\lim_{s\rightarrow 0}\frac{F(u+sh,v)-F(u,v)}{h},

with u,hUu,h\in U, vVv\in V and the limit taken in the topology of ZZ. We say that FF is partially Gâteaux differentiable with respect to uu at (u,v)U×V(u,v)\in U\times V if there exists uF:U×VL(U,Z)\nabla_{u}F:U\times V\longrightarrow L(U,Z) such that uF(u,v;h)=uF(u,v)h\nabla_{u}F(u,v;h)=\nabla_{u}F(u,v)h for all hUh\in U.

Definition 2.3.

We say that F:U×VZF:U\times V\longrightarrow Z belongs to the class 𝒢1,0(U×V;Z)\mathcal{G}^{1,0}(U\times V;Z) if it is continuous, Gâteaux differentiable with respect to uu, for all (u,v)U×V(u,v)\in U\times V and uF:U×VL(U,Z)\nabla_{u}F:U\times V\longrightarrow L(U,Z) is strongly continuous.

For FF depending on additional arguments, the definition above can be easily generalized.

Lemma 2.4.

Given U,V,ZU,V,Z three Banach spaces, a continuous map F:U×VZF:U\times V\longrightarrow Z belongs to 𝒢1,0(U×V,Z)\mathcal{G}^{1,0}(U\times V,Z) provided the following conditions hold:

  1. (1)

    The partial directional derivatives uF(u,v;h)\nabla_{u}F(u,v;h) exist at every point (u,v)U×V(u,v)\in U\times V and in every directoin hUh\in U.

  2. (2)

    For every (u,v)(u,v) the mapping huF(u,v;h)h\longmapsto\nabla_{u}F(u,v;h) is continuous from UU to ZZ.

  3. (3)

    For every hh, the mapping uF(,;h):U×VZ\nabla_{u}F(\cdot,\cdot;h):U\times V\longrightarrow Z is continuous.

We are going to use the following parameter depending contraction principle to study the regular dependence of the solution of stochastic differential equations on their initial data.

Proposition 2.5.

Let U,V,ZU,V,Z be Banach spaces and let F:U×V×ZUF:U\times V\times Z\longrightarrow U a continuous mapping satisfying

|F(u1,v,z)F(u2,v,z)|α|u1u2|,|F(u_{1},v,z)-F(u_{2},v,z)|\leq\alpha|u_{1}-u_{2}|,

for some α[0,1)\alpha\in[0,1) and every u1,u2Uu_{1},u_{2}\in U, vVv\in V, zZz\in Z. Let ϕ(v,z)\phi(v,z) denote the unique fixed point of the mapping F(,v,z):UUF(\cdot,v,z):U\longrightarrow U. Then ϕ:V×ZU\phi:V\times Z\longrightarrow U is continuous. If, in addition F𝒢1,1,0(U×V×Z,U)F\in\mathcal{G}^{1,1,0}(U\times V\times Z,U), then ϕ𝒢1,0(V×Z,U)\phi\in\mathcal{G}^{1,0}(V\times Z,U) and

vϕ(v,z)=uF(ϕ(v,z),v,z)vϕ(v,z)+vF(ϕ(v,z),v,z).\nabla_{v}\phi(v,z)=\nabla_{u}F\big{(}\phi(v,z),v,z\big{)}\nabla_{v}\phi(v,z)+\nabla_{v}F(\phi(v,z),v,z).

2.2. Lift approach to optimal control

In the sequel, we exploit an infinite dimensional lift to reformulate the optimization problem (1.1)-(1.2) in an infinite dimensional setting. Our first step is to rewrite XuX^{u} in (1.2) in terms of a process 𝒵u\mathcal{Z}^{u} with values in a Banach space, using the lift procedure presented in [9]. Notice that we do not actually work in the affine framework of [9], but the approach presented here is actually a particular case of the one introduced in [7].

Definition 2.6.

Let YY be a Banach space with dual YY^{*} and denote with ,Y×Y\langle\cdot,\cdot\rangle_{Y\times Y^{*}} the pairing between YY and YY^{*}. We say that a kernel KLloc2(+,)K\in L^{2}_{loc}(\mathbb{R}_{+},\mathbb{R}) is liftable if there exist gYg\in Y, νY\nu\in Y^{*} and a uniformly continuous semigroup 𝒮t\mathcal{S}_{t}^{*}, t[0,T]t\in[0,T] with generator 𝒜\mathcal{A}^{*}, acting on YY^{*}, such that

  • K(t)=g,𝒮tνY×YK(t)=\langle g,\mathcal{S}_{t}^{*}\nu\rangle_{Y\times Y^{*}}

  • 𝒮tνY\mathcal{S}_{t}^{*}\nu\in Y^{*} for all t>0t>0

  • 0t𝒮sνY2𝑑s<\int_{0}^{t}\|\mathcal{S}_{s}^{*}\nu\|^{2}_{Y^{*}}ds<\infty for all t>0t>0.

For notational simplicity we write ,\langle\cdot,\cdot\rangle for ,Y×Y\langle\cdot,\cdot\rangle_{Y\times Y^{*}} when no confusion arises.

From now on, we make the following assumption:

Hypothesis 2.7.

The kernel KK in (1.2) is liftable.

We rewrite XuX^{u} as

Xu(τ)\displaystyle X^{u}(\tau) =x(τ)+tτK(τs)[β(s,Xu(s))+σ(s,Xu(s))R(s,Xu(s),u(s))]𝑑s\displaystyle=x(\tau)+\int_{t}^{\tau}K(\tau-s)\Big{[}\beta(s,X^{u}(s))+\sigma(s,X^{u}(s))R(s,X^{u}(s),u(s))\Big{]}ds
+tτK(τs)σ(s,Xu(s))𝑑W(s)\displaystyle\quad+\int_{t}^{\tau}K(\tau-s)\sigma(s,X^{u}(s))dW(s)
:=x(τ)+tτK(τs)𝑑Vu(s),\displaystyle:=x(\tau)+\int_{t}^{\tau}K(\tau-s)dV^{u}(s),

where

dVu(s):=[β(s,Xu(s))+σ(s,Xu(s))R(s,Xu(s),u(s))]ds+σ(s,Xu(s))dW(s).dV^{u}(s):=\Big{[}\beta(s,X^{u}(s))+\sigma(s,X^{u}(s))R(s,X^{u}(s),u(s))\Big{]}ds+\sigma(s,X^{u}(s))dW(s). (2.1)

Defining ζ\zeta as an element in YY^{*} such that x(τ)=:g,𝒮τζx(\tau)=:\langle g,\mathcal{S}_{\tau}^{*}\zeta\rangle we can now rewrite (1.2) as follows:

Xu(τ)\displaystyle X^{u}(\tau) =x(τ)+tτK(τs)𝑑Vu(s)\displaystyle=x(\tau)+\int_{t}^{\tau}K(\tau-s)dV^{u}(s)
=g,𝒮τζ+tτg,𝒮τsν𝑑Vu(s)\displaystyle=\langle g,\mathcal{S}_{\tau}^{*}\zeta\rangle+\int_{t}^{\tau}\langle g,\mathcal{S}_{\tau-s}^{*}\nu\rangle dV^{u}(s)
=g,𝒮τζ+tτ𝒮τsν𝑑Vu(s)\displaystyle=\Bigg{\langle}g,\mathcal{S}_{\tau}^{*}\zeta+\int_{t}^{\tau}\mathcal{S}_{\tau-s}^{*}\nu dV^{u}(s)\Bigg{\rangle}
=:g,𝒵τu,\displaystyle=:\langle g,\mathcal{Z}_{\tau}^{u}\rangle, (2.2)

where 𝒵τu:=𝒮τζ+tτ𝒮τsν𝑑Vu(s)\mathcal{Z}_{\tau}^{u}:=\mathcal{S}_{\tau}^{*}\zeta+\int_{t}^{\tau}\mathcal{S}_{\tau-s}^{*}\nu dV^{u}(s). One can then check that 𝒵τu\mathcal{Z}_{\tau}^{u} follows the dynamics:

𝒵τu=𝒮tζ+tτ𝒜𝒵su𝑑s+tτν𝑑Vu(s),\mathcal{Z}_{\tau}^{u}=\mathcal{S}_{t}^{*}\zeta+\int_{t}^{\tau}\mathcal{A}^{*}\mathcal{Z}_{s}^{u}ds+\int_{t}^{\tau}\nu dV^{u}(s), (2.3)

In fact, we have that

tτ𝒜𝒵su𝑑s\displaystyle\int_{t}^{\tau}\mathcal{A}^{*}\mathcal{Z}^{u}_{s}ds =tτ𝒜[𝒮sζ+ts𝒮svν𝑑Vu(v)]𝑑s\displaystyle=\int_{t}^{\tau}\mathcal{A}^{*}\left[\mathcal{S}_{s}^{*}\zeta+\int_{t}^{s}\mathcal{S}^{*}_{s-v}\nu dV^{u}(v)\right]ds
=e𝒜τζe𝒜tζ+tτvτ𝒜e𝒜(sv)ν𝑑s𝑑Vu(v)\displaystyle=e^{\mathcal{A}^{*}\tau}\zeta-e^{\mathcal{A}^{*}t}\zeta+\int_{t}^{\tau}\!\!\!\int_{v}^{\tau}\mathcal{A}^{*}e^{\mathcal{A}^{*}(s-v)}\nu dsdV^{u}(v)
=e𝒜τζe𝒜tζ+tτe𝒜(τv)ν𝑑Vu(v)tτν𝑑Vu(v)\displaystyle=e^{\mathcal{A}^{*}\tau}\zeta-e^{\mathcal{A}^{*}t}\zeta+\int_{t}^{\tau}e^{\mathcal{A}^{*}(\tau-v)}\nu dV^{u}(v)-\int_{t}^{\tau}\nu dV^{u}(v)
=𝒵τue𝒜tζtτν𝑑Vu(v),\displaystyle=\mathcal{Z}_{\tau}^{u}-e^{\mathcal{A}^{*}t}\zeta-\int_{t}^{\tau}\nu dV^{u}(v),

and, rearranging the terms we obtain (2.3).

Remark 2.8.

By defining B(t,Xu(t),u(t)):=β(s,Xu(s))σ(s,Xu(s))R(s,Xu(s),u(s))B(t,X^{u}(t),u(t)):=\beta(s,X^{u}(s))\sigma(s,X^{u}(s))R(s,X^{u}(s),u(s)), and exploiting (2.2), we actually get that the function x(τ)=g,𝒮tζx(\tau)=\langle g,\mathcal{S}_{t}^{*}\zeta\rangle is given by the expression

x(τ)\displaystyle x(\tau) =𝔼[Xu(τ)tτK(τs)B(s,Xu(s),u(s))𝑑s]\displaystyle=\mathbb{E}\left[X^{u}(\tau)-\int_{t}^{\tau}K(\tau-s)B(s,X^{u}(s),u(s))ds\right]
=g,𝔼[𝒵τutτ𝒮τsνB(s,Xu(s),u(s))𝑑s]\displaystyle=\left\langle g,\mathbb{E}\left[\mathcal{Z}^{u}_{\tau}-\int_{t}^{\tau}\mathcal{S}_{\tau-s}^{*}\nu B(s,X^{u}(s),u(s))ds\right]\right\rangle

Set 𝒵τu,g:=g,𝒵τu=Xu(τ)\mathcal{Z}_{\tau}^{u,g}:=\langle g,\mathcal{Z}_{\tau}^{u}\rangle=X^{u}(\tau), and plug (2.1) into (2.3), then we can rewrite (2.3) in differential notation as

d𝒵τu=𝒜𝒵τudτ+ν(β(τ,𝒵τu,g)dτ+σ(τ,𝒵τu,g)[R(τ,𝒵τu,g,uτ)dτ+dWτ]),d\mathcal{Z}_{\tau}^{u}=\mathcal{A}^{*}\mathcal{Z}_{\tau}^{u}d{\tau}+\nu\Big{(}\beta({\tau},\mathcal{Z}_{\tau}^{u,g})d\tau+\sigma({\tau},\mathcal{Z}_{\tau}^{u,g})\big{[}R({\tau},\mathcal{Z}_{\tau}^{u,g},u_{\tau})d{\tau}+dW_{\tau}\big{]}\Big{)}, (2.4)

with 𝒵tu=e𝒜tζ:=ζt\mathcal{Z}^{u}_{t}=e^{\mathcal{A}^{*}t}\zeta:=\zeta_{t}. We also rewrite (2.3) as

d𝒵τu=𝒜𝒵τudτ+νβg(τ,𝒵τu)dτ+νσg(τ,𝒵τu)[Rg(τ,𝒵τu,uτ)dτ+dWτ],d\mathcal{Z}_{\tau}^{u}=\mathcal{A}^{*}\mathcal{Z}_{\tau}^{u}d{\tau}+\nu\beta^{g}({\tau},\mathcal{Z}_{\tau}^{u})d\tau+\nu\sigma^{g}({\tau},\mathcal{Z}_{\tau}^{u})[R^{g}({\tau},\mathcal{Z}_{\tau}^{u},u_{\tau})d\tau+dW_{\tau}], (2.5)

where

βg(s,𝒵su)\displaystyle\beta^{g}(s,\mathcal{Z}_{s}^{u}) :=β(s,𝒵su,g)=β(s,g,Zsu)=β(s,X(s)),\displaystyle:=\beta(s,\mathcal{Z}_{s}^{u,g})=\beta(s,\langle g,Z_{s}^{u}\rangle)=\beta(s,X(s)),
σg(s,𝒵su)\displaystyle\sigma^{g}(s,\mathcal{Z}_{s}^{u}) :=σ(s,𝒵su,g)=σ(s,g,Zsu)=σ(s,X(s)),\displaystyle:=\sigma(s,\mathcal{Z}_{s}^{u,g})=\sigma(s,\langle g,Z_{s}^{u}\rangle)=\sigma(s,X(s)),
Rg(s,𝒵su,us)\displaystyle R^{g}(s,\mathcal{Z}_{s}^{u},u_{s}) :=R(s,𝒵su,g,us)=R(s,g,Zsu,us)=R(s,X(s),us).\displaystyle:=R(s,\mathcal{Z}_{s}^{u,g},u_{s})=R(s,\langle g,Z_{s}^{u}\rangle,u_{s})=R(s,X(s),u_{s}).

We are going to discuss existence and uniqueness results for equation (2.4) in Section 3.1.

Remark 2.9.

We point out that the term

0τνσg(s,𝒵s)𝑑Ws,\int_{0}^{\tau}\nu\sigma^{g}(s,\mathcal{Z}_{s})dW_{s}, (2.6)

in (2.5) can be regarded in two different ways. On the one hand, it can be seen as the element of YY^{*}:

ν(0τσg(s,𝒵s)𝑑Ws),\nu\left(\int_{0}^{\tau}\sigma^{g}(s,\mathcal{Z}_{s})dW_{s}\right),

where the integration of 0τσg(s,𝒵s)𝑑Ws=0τσ(s,X(s))𝑑Ws\int_{0}^{\tau}\sigma^{g}(s,\mathcal{Z}_{s})dW_{s}=\int_{0}^{\tau}\sigma(s,X(s))dW_{s} is done on \mathbb{R} and then lifted to YY^{*} by multiplying it by ν\nu. On the other hand, by writing (2.6) as

0τσg(s,𝒵s)d(νWs),\int_{0}^{\tau}\sigma^{g}(s,\mathcal{Z}_{s})d(\nu W_{s}),

we have that νWs\nu W_{s} can be considered as a cylindrical Wiener process on ν:={νx,x}\mathbb{H}^{\nu}:=\left\{\nu x,\ x\in\mathbb{R}\right\}, which is a Hilbert space with the scalar product ,ν:=νY,\langle\cdot,\cdot\rangle_{\mathbb{H}^{\nu}}:=\|\nu\|_{Y^{*}}\langle\cdot,\cdot\rangle_{\mathbb{R}}. In this case we also see that νY\mathbb{H}^{\nu}\subsetneq Y^{*}.

In section 3.1 we are going to provide sufficient conditions that guarantee the existence of a solution of (2.3)-(2.4). Due to the nature of the lift and identification (2.2), this will, in turn, provide sufficient conditions also for the existence of a solution to (1.2).

Remark 2.10.

From [9, 8] we see that we could perform the lift also under weaker hypothesis, by taking a subspace ZYZ\subset Y with their relative duals YZY^{*}\subset Z^{*} such that:

  • ZZ and YY are Banach spaces ZYZ\subset Y and ZZ embeds continuously into YY.

  • The semigroup 𝒮\mathcal{S}^{*} with generator 𝒜\mathcal{A}^{*} acts in a strongly continuous way on YY^{*} and ZZ^{*} with respect to the respective norm topologies.

  • The map 𝒵𝒮t𝒵\mathcal{Z}\longmapsto\mathcal{S}^{*}_{t}\mathcal{Z} is weak-* continuous on YY^{*} and on ZZ^{*} for every t0t\geq 0.

  • The pre-adjoint operator of 𝒜\mathcal{A}^{*}, generates a strongly continuous semigroup on ZZ with respect to the respective norm topology (but not necessarily on YY).

In this case every kernel of the form K(t)=g,𝒮tνK(t)=\langle g,\mathcal{S}_{t}^{*}\nu\rangle with νZ\nu\in Z^{*} and StνYS_{t}^{*}\nu\in Y^{*} is liftable. While this setting would allow to work with a wider class of kernels, we would not be able to formulate the HJB equations. This is due to the fact that, when considering a kernel K(t)=g,𝒮tνK(t)=\langle g,\mathcal{S}_{t}^{*}\nu\rangle with νZ\nu\in Z^{*}, some of the inner products in the definition of the Hamilton-Jacobi-Bellman equation (3.16), would not be well defined. Being the goal of this work a control problem, we restrict ourselves to the case νY\nu\in Y^{*}, as originally stated.

In a similar fashion to what we did for (1.2), recalling that Xu(τ):=𝒵τu,gX^{u}(\tau):=\mathcal{Z}_{\tau}^{u,g}, we rewrite the performance functional (1.1) so to make its dependence from the lifted process 𝒵τu\mathcal{Z}^{u}_{\tau} explicit:

J(t,x,u)\displaystyle J(t,x,u) =𝔼[tTF(τ,𝒵τu,g,uτ)𝑑τ+G(𝒵Tu,g)]\displaystyle=\mathbb{E}\left[\int_{t}^{T}F(\tau,\mathcal{Z}^{u,g}_{\tau},u_{\tau})d\tau+G(\mathcal{Z}^{u,g}_{T})\right]
:=𝔼[tTFg(τ,𝒵τu,uτ)𝑑τ+Gg(𝒵Tu)]:=Jg(t,ζ,u),\displaystyle:=\mathbb{E}\left[\int_{t}^{T}F^{g}(\tau,\mathcal{Z}_{\tau}^{u},u_{\tau})d\tau+G^{g}(\mathcal{Z}_{T}^{u})\right]:=J^{g}(t,\zeta,u), (2.7)

where the functions F:[0,T]××𝒰F:[0,T]\times\mathbb{R}\times\mathcal{U}\longrightarrow\mathbb{R}, G:G:\mathbb{R}\longrightarrow\mathbb{R} are lifted to the functions

Fg\displaystyle F^{g} :[0,T]×Y×𝒰,\displaystyle:[0,T]\times Y^{*}\times\mathcal{U}\longrightarrow\mathbb{R},
Gg\displaystyle G^{g} :Y\displaystyle:Y^{*}\longrightarrow\mathbb{R}

where YY^{*} is the Banach space associated to the liftable kernel, see Definition 2.6. The lifted maps FgF^{g} and GgG^{g} are

Fg(,𝒵τu,)\displaystyle F^{g}(\cdot,\mathcal{Z}^{u}_{\tau},\cdot) :=F(,𝒵τu,g,)=F(,g,𝒵τu,)=F(,Xu(τ),),\displaystyle:=F(\cdot,\mathcal{Z}^{u,g}_{\tau},\cdot)=F(\cdot,\langle g,\mathcal{Z}^{u}_{\tau}\rangle,\cdot)=F(\cdot,X^{u}(\tau),\cdot),
Gg(𝒵τu)\displaystyle G^{g}(\mathcal{Z}^{u}_{\tau}) :=G(𝒵τu,g)=G(g,𝒵τu)=G(Xu(τ)).\displaystyle:=G(\mathcal{Z}^{u,g}_{\tau})=G(\langle g,\mathcal{Z}^{u}_{\tau}\rangle)=G(X^{u}(\tau)).

The stochastic optimal control problem (1.1)-(1.3) is then lifted to

Jg(t,ζ,u^)=infu𝔸Jg(t,ζ,u)=infu𝔸𝔼[tTFg(τ,𝒵τu,u)𝑑τ+Gg(𝒵Tu)],J^{g}(t,\zeta,\hat{u})=\inf_{u\in\mathds{A}}J^{g}(t,\zeta,u)=\inf_{u\in\mathds{A}}\mathbb{E}\left[\int_{t}^{T}F^{g}(\tau,\mathcal{Z}_{\tau}^{u},u)d\tau+G^{g}(\mathcal{Z}_{T}^{u})\right], (2.8)

where the process 𝒵u\mathcal{Z}^{u} takes values in the Banach space YY^{*}, and where the dynamics for the controlled process are given by (2.3)-(2.4). Notice that, while the performance functional has not changed, we write JgJ^{g} instead of JJ in order to highlight the dependence on 𝒵tu\mathcal{Z}^{u}_{t} instead of Xu(t)X^{u}(t), as underneath there is a passage from finite to infinite dimensions. Indeed, this change of notation embodies a crucial change of framework from a finite to an infinite dimensional setting, allowing us to move from functions β\beta, σ\sigma, RR, FF and GG taking values from \mathbb{R} to new functions βg\beta^{g}, σg\sigma^{g}, RgR^{g}, FgF^{g} and GgG^{g} that now take values from YY^{*}. This lift allows us to consider a new optimization problem, written on a space which is not the original one. Nonetheless, we have that J(t,x,u)=Jg(t,ζ,u)J(t,x,u)=J^{g}(t,\zeta,u) for t[0,T]t\in[0,T], u𝒰u\in\mathcal{U}. Also, being gg fixed and only depending on the kernel representation, finding the pair (u^,𝒵u^)(\hat{u},\mathcal{Z}^{\hat{u}}) that minimizes (2.5)-(2.8) is equivalent to finding the pair (u^,Xu^)(\hat{u},X^{\hat{u}}) that solves (1.1)-(1.2).

2.3. UMD Banach spaces

In the sequel we use techniques of Malliavin calculus on the space YY^{*}. For this, we assume:

Hypothesis 2.11.

The space YY^{*} is a unconditional martingale differences (UMD) Banach space.

For convenience we report here below the essentials on UMD Banach spaces.

Definition 2.12.

Let (Mn)n=1N(M_{n})_{n=1}^{N} be a Banach-space valued martingale, the sequence dn=Mn+1Mnd_{n}=M_{n+1}-M_{n} is called the martingale difference sequence associated with (Mn)n=1N(M_{n})_{n=1}^{N}. A Banach space EE is said to be a UMDpUMD_{p} (1<p<)(1<p<\infty), space if there exists a constant β\beta such that for all EE-valued LpL^{p}-martingale difference sequences (dn)n=1N(d_{n})_{n=1}^{N} we have

𝔼n=1Nϵndnpβp𝔼n=1Ndnp,\mathbb{E}\left\|\sum_{n=1}^{N}\epsilon_{n}d_{n}\right\|^{p}\leq\beta^{p}\mathbb{E}\left\|\sum_{n=1}^{N}d_{n}\right\|^{p},

where ϵn\epsilon_{n}\in\mathbb{R} for all nn and |ϵn|=1.|\epsilon_{n}|=1. Thanks to [26] we also know that, if a Banach space EE is UMDpUMD_{p} for some 1<p<1<p<\infty, then EE is a UMDpUMD_{p} Banach space for all p(1,)p\in(1,\infty), and we simply call it a UMD Banach space.

In the context of stochastic analysis in Banach spaces, martingale difference sequences provide a substitute for orthogonal sequences. In the following parts, we will see that this hypothesis is not very restrictive, as the UMD Banach spaces include all Hilbert spaces, LqL^{q} spaces for q(1,)q\in(1,\infty), reflective Sobolev spaces and many others, thus allowing us to consider a wide class of liftable kernels. In our framework, the process 𝒵u\mathcal{Z}^{u} takes values in a UMD Banach space whenever we consider, for example, a shift operator or a quasi-exponential kernel or a kernel that can be expressed as the Laplace transform of a measure with density in Lq([0,))L^{q}([0,\infty)), q(1,)q\in(1,\infty).

Assuming that YY^{*} is UMD, allows us to define the Malliavin derivative operator DD on Lp(Ω,Y)L^{p}(\Omega,Y^{*}). From [23, Proposition 2.5], we know that DD is a closed operator and we denote with 𝔻1,p(Y)\mathbb{D}^{1,p}(Y^{*}) the closure of the domain.

For the results on UMD Banach spaces exploited in the following parts, we refer to [28] for the BDG inequality, [27] for the Fubini Theorem and to [23] for general Malliavin calculus results. In this framework we will also use a Clark-Okone formula for UMD Banach spaces (see [21]) and the following chain rule linking the Malliavin derivative and the Gâteaux derivative (see [23])

Proposition 2.13.

Let EE be a UMD Banach space and let p(1,)p\in(1,\infty). Suppose that φ𝒢1(E,E)\varphi\in\mathcal{G}^{1}(E,E). If F𝔻1,p(E)F\in\mathbb{D}^{1,p}(E), then φ(F)𝔻1,p(E)\varphi(F)\in\mathbb{D}^{1,p}(E) with

D(φ(F))=φ(F)DF.D(\varphi(F))=\nabla\varphi(F)DF.

3. The optimal control problem

We are now interested in solving the lifted optimal control problem (2.8), where the process 𝒵\mathcal{Z} follows the controlled dynamics given by

{d𝒵τu=𝒜𝒵τudτ+νβg(τ,𝒵τu)dτ+νσg(τ,𝒵τu)Rg(τ,𝒵τu,uτ)dτ+νσg(τ,𝒵τu)dWτ,𝒵tu=ζt.\begin{cases}d\mathcal{Z}^{u}_{\tau}=&\mathcal{A}^{*}\mathcal{Z}_{\tau}^{u}d\tau+\nu\beta^{g}(\tau,\mathcal{Z}^{u}_{\tau})d\tau+\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau})R^{g}(\tau,\mathcal{Z}^{u}_{\tau},u_{\tau})d\tau\\ &+\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau})dW_{\tau},\\ \mathcal{Z}^{u}_{t}=&\zeta_{t}.\end{cases} (3.1)

For our results to hold, we add some Hypothesis on RR, which directly translates into hypothesis on RgR^{g}.

Hypothesis 3.1.

R:[0,T]××𝒰R:[0,T]\times\mathbb{R}\times\mathcal{U}\longrightarrow\mathbb{R} is measurable and R(τ,x,u)KR\|R(\tau,x,u)\|_{\mathbb{R}}\leq K_{R} for a suitable positive constant KR>0K_{R}>0 and every τ[0,T]\tau\in[0,T], xx\in\mathbb{R}, u𝒰u\in\mathcal{U}.

In order to find the optimal value J(u^)J(\hat{u}), we associate the following partially coupled system of forward-backward equations

{d𝒵τ=𝒜𝒵τdτ+νβg(τ,𝒵τ)dτ+νσg(τ,𝒵τ)dWτ,τ[t,T],𝒵t=ζt,dpτ=(τ,𝒵τ,qτ)dτ+qτνdWτ,τ[t,T],pT=G(𝒵T),\begin{cases}d\mathcal{Z}_{\tau}&=\mathcal{A}^{*}\mathcal{Z}_{\tau}d\tau+\nu\beta^{g}(\tau,\mathcal{Z}_{\tau})d\tau+\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau})dW_{\tau},\ \ \tau\in[t,T],\\ \mathcal{Z}_{t}&=\zeta_{t},\\ dp_{\tau}&=-\mathcal{H}(\tau,\mathcal{Z}_{\tau},q_{\tau})d\tau+q_{\tau}\nu dW_{\tau},\quad\tau\in[t,T],\\ p_{T}&=G(\mathcal{Z}_{T}),\end{cases} (3.2)

to (3.1). Here above :[0,T]×Y×Y\mathcal{H}:[0,T]\times Y^{*}\times Y^{**}\longrightarrow\mathbb{R} is the Hamiltonian function defined as

(t,z,ξ)=infu𝒰[Fg(t,z,u)+ξνRg(t,z,u)].\mathcal{H}(t,z,\xi)=\inf_{u\in\mathcal{U}}\left[F^{g}(t,z,u)+\xi\nu R^{g}(t,z,u)\right]. (3.3)

Notice that the control uu only appears in the Hamiltonian functional. The solution of the backward equation is denoted by (p,q)(p,q). We often write (pτ,qτ)=(p(τ,t,ζt),q(τ,t,ζt))(p_{\tau},q_{\tau})=(p(\tau,t,\zeta_{t}),q(\tau,t,\zeta_{t})), τ[t,T]\tau\in[t,T] when we want to emphasize the dependence of pp and qq on the parameter ζt\zeta_{t} at time tt. Analogously, when we want to emphasize the dependence of 𝒵\mathcal{Z} on the initial value ζt\zeta_{t} at time tt, we write 𝒵τ=𝒵(τ,t,ζt)\mathcal{Z}_{\tau}=\mathcal{Z}(\tau,t,\zeta_{t}).

Define now

v(t,z):=p(t,t,z),v(t,z):=p(t,t,z), (3.4)

with pp the solution to the backward SDE in (3.2). In the sequel we show that Jg(t,ζt,u^)J^{g}(t,\zeta_{t},\hat{u}) in (2.8) is such that

Jg(t,ζt,u^)=v(t,ζt)J^{g}(t,\zeta_{t},\hat{u})=v(t,\zeta_{t}) (3.5)

and that the optimal control u^\hat{u} can be retrieved explicitly via a verification theorem once v(t,ζt)v(t,\zeta_{t}) is known. In order to achieve (3.5) we proceed as follow. First we study the forward equation 𝒵(τ,t,ζt)\mathcal{Z}(\tau,t,\zeta_{t}) in Section 3.1, then we study the backward equation (p(τ,t,ζt),q(τ,t,ζt))(p(\tau,t,\zeta_{t}),q(\tau,t,\zeta_{t})) in Section 3.2 and there we prove the crucial identification:

qt=zv(t,𝒵t)νσg(t,𝒵t),q_{t}=\nabla_{z}v(t,\mathcal{Z}_{t})\nu\sigma^{g}(t,\mathcal{Z}_{t}), (3.6)

(see Proposition 3.11). In Section 3.3 we provide an approach to find v(t,z)v(t,z) through HJB equations and at last, in Section 3.4, we prove (3.5) and we provide a characterization of the optimal control u^\hat{u}.

Notice that, for (3.6) to hold, the backward process p(,t,z)p(\cdot,t,z) has to be differentiable with respect to zz. This can be obtained by showing that 𝒵tu\mathcal{Z}_{t}^{u} is differentiable with respect to the initial condition ζt\zeta_{t}, and by assuming the following:

Hypothesis 3.2.

Let us assume that

  • 1)

    There exists L1>0L_{1}>0 such that

    |(t,z,ξ1)(t,z,ξ2)|L1ξ1ξ2Y|\mathcal{H}(t,z,\xi_{1})-\mathcal{H}(t,z,\xi_{2})|\leq L_{1}\|\xi_{1}-\xi_{2}\|_{Y^{**}}

    for every t[0,T]t\in[0,T], zYz\in Y^{*} and ξ1,ξ2Y\xi_{1},\xi_{2}\in Y^{**}.

  • 2)

    For all t[0,T]t\in[0,T], 𝔼[tT|(s,0,0)|2𝑑s]<\mathbb{E}\left[\int_{t}^{T}|\mathcal{H}(s,0,0)|^{2}ds\right]<\infty.

  • 3)

    For every t[0,T]t\in[0,T] we have (t,,)𝒢1,1(Y×Y)\mathcal{H}(t,\cdot,\cdot)\in\mathcal{G}^{1,1}(Y^{*}\times Y^{**}).

  • 4)

    There exist L2>0L_{2}>0 and m0m\geq 0 such that

    |z(t,z,ξ)h|L2hY(1+zY)m(1+ξY)|\nabla_{z}\mathcal{H}(t,z,\xi)h|\leq L_{2}\|h\|_{Y^{*}}(1+\|z\|_{Y^{*}})^{m}(1+\|\xi\|_{Y^{**}})

    for every t[0,T]t\in[0,T], z,hYz,h\in Y^{*} and ξY\xi\in Y^{**}.

  • 5)

    Gg𝒢1(Y)G^{g}\in\mathcal{G}^{1}(Y^{*}) and there exists L3>0L_{3}>0 such that, for every z1,z2Yz_{1},z_{2}\in Y^{*}

    |Gg(z1)Gg(z2)|L3z1z2Y.|G^{g}(z_{1})-G^{g}(z_{2})|\leq L_{3}\|z_{1}-z_{2}\|_{Y^{*}}.

Further details on the continuous dependence on ζt\zeta_{t} of the forward equation can be found in Section 3.1, while we refer to Section 3.2 for details on the differentiability of p(,t,z)p(\cdot,t,z) with respect to zz.

Remark 3.3.

The identification (3.6), in the case where σg\sigma^{g} does not depend on 𝒵\mathcal{Z} in (3.2), can be proven following [22] and dropping the UMD hypothesis on YY^{*}. In our case though, being σg\sigma^{g} dependent on 𝒵t\mathcal{Z}_{t}, we need to exploit some Malliavin calculus techniques on Banach spaces, and thus assume that YY^{*} is UMD.

3.1. On the lifted forward equation

In this section we study the lifted forward equation 𝒵t\mathcal{Z}_{t} in (3.2). In particular, we prove that it admits a unique Markovian solution and we study its continuous dependence from the initial parameter ζt\zeta_{t}. We thus take

{d𝒵τ=𝒜𝒵τdτ+νβg(τ,𝒵τ)dτ+νσg(τ,𝒵τ)dWτ,τ[t,T],𝒵t=ζt,\begin{cases}d\mathcal{Z}_{\tau}&=\mathcal{A}^{*}\mathcal{Z}_{\tau}d\tau+\nu\beta^{g}(\tau,\mathcal{Z}_{\tau})d\tau+\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau})dW_{\tau},\quad\tau\in[t,T],\\ \mathcal{Z}_{t}&=\zeta_{t},\end{cases} (3.7)

we recall that 𝒜\mathcal{A}^{*} is the generator of a uniformly continuous semigroup on the Banach space Y.Y^{*}. We assume the following:

Hypothesis 3.4.

Suppose that

  • i)

    βg:[0,T]×Y\beta^{g}:[0,T]\times Y^{*}\longrightarrow\mathbb{R} is continuous and, for all t[0,T]t\in[0,T] z1,z2Yz_{1},z_{2}\in Y^{*}, there exists a constant L1>0L_{1}>0 such that

    |νβg(t,z1)νβg(t,z2)|L1z1z2Y,|\nu\beta^{g}(t,z_{1})-\nu\beta^{g}(t,z_{2})|\leq L_{1}\|z_{1}-z_{2}\|_{Y^{*}},

    the map νβg:[0,T]×YY\nu\beta^{g}:[0,T]\times Y^{*}\longrightarrow Y^{*} is measurable. Moreover, for all t[0,T]t\in[0,T] and zYz\in Y^{*},

    |νβg(t,z)|2L2(1+zY2),|\nu\beta^{g}(t,z)|^{2}\leq L_{2}(1+\|z\|_{Y^{*}}^{2}),

    for some constant L2>0L_{2}>0.

  • ii)

    σg:[0,T]×Y\sigma^{g}:[0,T]\times Y^{*}\longrightarrow\mathbb{R} is such that, for every vYv\in Y^{**} the map νσgv:[0,T]×Y\nu\sigma^{g}v:[0,T]\times Y^{*}\longrightarrow\mathbb{R} is measurable, es𝒜νσg(t,z)L2(Y)e^{s\mathcal{A}^{*}}\nu\sigma^{g}(t,z)\in L^{2}(Y^{*}) for every s>0s>0, t[0,T]t\in[0,T] and zYz\in Y^{*}, and

    νσg(t,z)L(Y)2L3(1+zY2),\|\nu\sigma^{g}(t,z)\|^{2}_{L(Y^{*})}\leq L_{3}(1+\|z\|_{Y^{*}}^{2}),

    for some constant L3>0L_{3}>0.

    Moreover, for s>0s>0, t[0,T]t\in[0,T], z1,z2Yz_{1},z_{2}\in Y^{*} there exists a constant L4>0L_{4}>0 such that

    νσg(t,ζ1)νσg(t,z2)L2(Y)L4z1z2Y,\|\nu\sigma^{g}(t,\zeta_{1})-\nu\sigma^{g}(t,z_{2})\|_{L^{2}(Y^{*})}\leq L_{4}\|z_{1}-z_{2}\|_{Y^{*}},
  • iii)

    For every s>0s>0, t[0,T]t\in[0,T], νβg(t,)𝒢1(Y,Y)\nu\beta^{g}(t,\cdot)\in\mathcal{G}^{1}(Y^{*},Y^{*}).

Our first result is the following:

Proposition 3.5.

Assume Hypothesis 3.4 holds. For every p[2,)p\in[2,\infty), we have that:

  • i)

    The map (t,z)𝒵(,t,z)(t,z)\longmapsto\mathcal{Z}(\cdot,t,z) is in 𝒢0,1([0,T]×Y;Lp(Ω;C([0,T];Y)))\mathcal{G}^{0,1}([0,T]\times Y^{*};L^{p}(\Omega;C([0,T];Y^{*}))).

  • ii)

    For every hYh\in Y^{*} the partial directional derivative process z𝒵(τ,t,z)h\nabla_{z}\mathcal{Z}(\tau,t,z)h, τ[0,T]\tau\in[0,T] solves \mathbb{P}-a.s. the equation

    z𝒵(τ,t,z)h\displaystyle\nabla_{z}\mathcal{Z}(\tau,t,z)h =e(τt)𝒜h+tτe(τs)𝒜zνβg(s,𝒵(s,t,z))z𝒵(s,t,z)h𝑑s\displaystyle=e^{(\tau-t)\mathcal{A}^{*}}h+\int_{t}^{\tau}e^{(\tau-s)\mathcal{A}^{*}}\nabla_{z}\nu\beta^{g}(s,\mathcal{Z}(s,t,z))\nabla_{z}\mathcal{Z}(s,t,z)h\ ds
    +tτz(e(τs)𝒜νσg(s,𝒵(s,t,z)))z𝒵(s,t,z)h𝑑Ws,τ[t,T],\displaystyle+\int_{t}^{\tau}\nabla_{z}\left(e^{(\tau-s)\mathcal{A}^{*}}\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))\right)\nabla_{z}\mathcal{Z}(s,t,z)h\ dW_{s},\quad\tau\in[t,T],
    z𝒵(τ,t,z,)h\displaystyle\nabla_{z}\mathcal{Z}(\tau,t,z,)h =h,τ[0,t).\displaystyle=h,\quad\tau\in[0,t).
  • iii)

    z𝒵(τ,t,z)hLp(Ω;C([0,T];Y))chY\|\nabla_{z}\mathcal{Z}(\tau,t,z)h\|_{L^{p}(\Omega;C([0,T];Y^{*}))}\leq c\|h\|_{Y^{*}} for some constant cc.

We also find that

  • iv)

    (3.7) admits a unique adapted solution 𝒵Lp(Ω,C([t,T]);Y)\mathcal{Z}\in L^{p}(\Omega,C([t,T]);Y^{*}).

Moreover, we have the following estimate

𝒵pp:=𝔼[supτ[t,T]𝒵τYp]C(1+ζtYp),\|\mathcal{Z}\|_{p}^{p}:=\mathbb{E}\left[\sup_{\tau\in[t,T]}\|\mathcal{Z}_{\tau}\|_{Y^{*}}^{p}\right]\leq C(1+\|\zeta_{t}\|_{Y^{*}}^{p}), (3.8)

where CC is a constant depending only on p,T,Lp,T,L, where L:=max{L1,L2,L3,L4}L:=\max\{L_{1},L_{2},L_{3},L_{4}\} and M:=supτ[t,T]eτ𝒜L2(Y)M:=\sup_{\tau\in[t,T]}\|e^{\tau\mathcal{A}^{*}}\|_{L^{2}(Y^{*})}.

Proof.

The proof is inspired by [10, Theorem 7.4] and [14, Proposition 3.2] The main difference with our work are the spaces at play. Consider the map

Φ(𝒵,t,z)τ:Lp(Ω;C([0,T];Y))×[0,T]×YLp(Ω;C([0,T];Y))\Phi(\mathcal{Z},t,z)_{\tau}:L^{p}(\Omega;C([0,T];Y^{*}))\times[0,T]\times Y^{*}\longrightarrow L^{p}(\Omega;C([0,T];Y^{*}))

defined as

Φ(𝒵,t,z)τ\displaystyle\Phi(\mathcal{Z},t,z)_{\tau} :=e𝒜(τt)z+0τ𝟙[t,T](s)e𝒜(τs)νβg(s,𝒵s)𝑑s\displaystyle:=e^{\mathcal{A}^{*}(\tau-t)}z+\int_{0}^{\tau}\mathds{1}_{[t,T]}(s)e^{\mathcal{A}^{*}(\tau-s)}\nu\beta^{g}(s,\mathcal{Z}_{s})ds
+0τ𝟙[t,T](s)e𝒜(τs)νσg(s,𝒵s)𝑑Ws\displaystyle+\int_{0}^{\tau}\mathds{1}_{[t,T]}(s)e^{\mathcal{A}^{*}(\tau-s)}\nu\sigma^{g}(s,\mathcal{Z}_{s})dW_{s}
:=S0(𝒵,t,z)τ+S1(𝒵,t,z)τ+S2(𝒵,t,z)ττ[0,T].\displaystyle:=S_{0}(\mathcal{Z},t,z)_{\tau}+S_{1}(\mathcal{Z},t,z)_{\tau}+S_{2}(\mathcal{Z},t,z)_{\tau}\quad\tau\in[0,T].

We want to show that Φ\Phi is a contraction with respect to the first variable. We notice that

S1(𝒵,t,z)p\displaystyle\|S_{1}(\mathcal{Z},t,z)\|^{p} Mp𝔼[(0Tνβg(s,𝒵s)Y𝑑s)p]\displaystyle\leq M^{p}\mathbb{E}\left[\left(\int_{0}^{T}\|\nu\beta^{g}(s,\mathcal{Z}_{s})\|_{Y^{*}}ds\right)^{p}\right]
Tp1Mp𝔼0T𝔼[0Tνβg(s,𝒵s)Yp𝑑s]\displaystyle\leq T^{p-1}M^{p}\mathbb{E}\int_{0}^{T}\mathbb{E}\left[\int_{0}^{T}\|\nu\beta^{g}(s,\mathcal{Z}_{s})\|_{Y^{*}}^{p}ds\right]
2p/21Tp1MpLp𝔼[0T(1+𝒵sYp)𝑑s]\displaystyle\leq 2^{p/2-1}T^{p-1}M^{p}L^{p}\mathbb{E}\left[\int_{0}^{T}(1+\|\mathcal{Z}_{s}\|_{Y^{*}}^{p})ds\right]
2p/21(TLM)p(1+𝒵p),\displaystyle\leq 2^{p/2-1}(TLM)^{p}(1+\|\mathcal{Z}\|_{p}),

and

S2(𝒵,t,z)p\displaystyle\|S_{2}(\mathcal{Z},t,z)\|^{p} supτ[0,T]𝔼[0τe(τs)𝒜νσg(s,𝒵s)𝑑W(s)Yp]\displaystyle\leq\sup_{\tau\in[0,T]}\mathbb{E}\left[\left\|\int_{0}^{\tau}e^{(\tau-s)\mathcal{A}^{*}}\nu\sigma^{g}(s,\mathcal{Z}_{s})dW(s)\right\|_{Y^{*}}^{p}\right]
MpCp/2LTp/212p/21𝔼[0T(1+𝒵sYp)𝑑s]\displaystyle\leq M^{p}C_{p/2}LT^{p/2-1}2^{p/2-1}\mathbb{E}\left[\int_{0}^{T}(1+\|\mathcal{Z}_{s}\|^{p}_{Y^{*}})ds\right]
MpCp/2L(2T)p/21(T+𝒵p),\displaystyle\leq M^{p}C_{p/2}L(2T)^{p/2-1}(T+\|\mathcal{Z}\|_{p}),

where we used the linear growth conditions on βg\beta^{g} and σg\sigma^{g} and the Burkholder-Davis-Gundy inequality for UMDUMD Banach spaces (see [28]). We thus have showed that Φ(,t,z)\Phi(\cdot,t,z) is a well defined mapping. Now, taking 𝒵1\mathcal{Z}_{1} and 𝒵2\mathcal{Z}_{2} arbitrary processes in YY^{*}, then

Φ(𝒵1,t,z)Φ(𝒵2,t,z)p\displaystyle\|\Phi(\mathcal{Z}_{1},t,z)-\Phi(\mathcal{Z}_{2},t,z)\|_{p} S1(𝒵1,t,z)S1(𝒵2,t,z)p+S2(𝒵1,t,z)S2(𝒵2,t,z)p\displaystyle\leq\|S_{1}(\mathcal{Z}_{1},t,z)-S_{1}(\mathcal{Z}_{2},t,z)\|_{p}+\|S_{2}(\mathcal{Z}_{1},t,z)-S_{2}(\mathcal{Z}_{2},t,z)\|_{p}
:=I1+I2.\displaystyle:=I_{1}+I_{2}.

With computations similar to the ones above, exploiting the Lipschitz condition on βg\beta^{g} and σg\sigma^{g} (see Hypothesis 3.4, i) and ii)), one finds that

I1p(TML)p𝒵1𝒵2pp,I_{1}^{p}\leq(TML)^{p}\|\mathcal{Z}_{1}-\mathcal{Z}_{2}\|_{p}^{p},

and

I2pCp/2(ML)pTp/2𝒵1𝒵2pp.I_{2}^{p}\leq C_{p/2}(ML)^{p}T^{p/2}\|\mathcal{Z}_{1}-\mathcal{Z}_{2}\|_{p}^{p}.

Summing up, we have that

Φ(𝒵1)Φ(𝒵2)pLM(Tp+Cp/2Tp/2)1/p𝒵1𝒵2p.\|\Phi(\mathcal{Z}_{1})-\Phi(\mathcal{Z}_{2})\|_{p}\leq LM(T^{p}+C_{p/2}T^{p/2})^{1/p}\|\mathcal{Z}_{1}-\mathcal{Z}_{2}\|_{p}.

This means that Φ(𝒵,t,z)\Phi(\mathcal{Z},t,z) is a contraction only for t[0,T]t\in[0,T] when TT satisfies

LM(Tp+Cp/2Tp/2)1/p<1.LM(T^{p}+C_{p/2}T^{p/2})^{1/p}<1. (3.9)

Condition (3.9) on TT can be easily removed by considering the equation on intervals [0,T~][0,\tilde{T}], [T~,2T~][\tilde{T},2\tilde{T}],…, where T~\tilde{T} satisfies (3.9). Thanks to the fixed point theorem we find that (2.5) admits a unique solution. We conclude that (3.8) holds by applying Gronwall’s Lemma with arguments in line with [10, Theorem 7.4 (iii)]. Notice now that, being Φ(,t,z)\Phi(\cdot,t,z) a contraction uniformly with respect to t[0,T]t\in[0,T], zYz\in Y^{*}, by Proposition 2.5 we obtain ii)ii) if

Φ𝒢1,0,1(Lp(Ω;C([0,T];Y))×[0,T]×Y,Lp(Ω;C([0,T];Y))).\Phi\in\mathcal{G}^{1,0,1}\left(L^{p}(\Omega;C([0,T];Y^{*}))\times[0,T]\times Y^{*},L^{p}(\Omega;C([0,T];Y^{*}))\right).

This is verified by a (slight modification) of Lemma 2.4. Indeed we notice that Φ\Phi is differentiable in zz. For more details we refer to [10, 14]. ∎

Remark 3.6.

We notice that 𝒵\mathcal{Z} is Markovian (see e.g. [13, Theorem 1.157]).

Corollary 3.7.

Assume Hypothesis 3.1 and 3.4 hold. Then (3.1) admits a unique solution.

Proof.

Thanks to the boundedness of RR, one can apply the Girsanov Theorem, see e.g. [10, Theorem 10.14], and proceed like in the proof of Proposition (3.5). ∎

3.2. On the backward equation

In this section we study the backward equation

pτ=Gg(𝒵T)+τT(s,𝒵s,qs)𝑑sτTqsν𝑑Ws,p_{\tau}=G^{g}(\mathcal{Z}_{T})+\int_{\tau}^{T}\mathcal{H}(s,\mathcal{Z}_{s},q_{s})ds-\int_{\tau}^{T}q_{s}\nu dW_{s}, (3.10)

introduced in (3.2). We study existence and uniqueness of a solution as well as its continuous dependence from the parameter ζt\zeta_{t}. Later on we will exploit (3.10) to prove (3.6), as well as show that the optimal value Jg(t,z,u^)J^{g}(t,z,\hat{u}) for the optimization problem (2.5) - (2.8) is achieved for Jg(t,ζt,u^)=v(t,ζt)=p(t,t,ζt)J^{g}(t,\zeta_{t},\hat{u})=v(t,\zeta_{t})=p(t,t,\zeta_{t}).

We observe that the following a priori estimate for the pair process (p,q)(p,q) holds (see [22] and [14, Proposition 4.3] ):

𝔼[supτ[t,T]|pτ|2]+𝔼[tTqτY2𝑑τ]c𝔼[tT|(τ,0,0)|2𝑑τ]+c𝔼[|Gg(𝒵T)|2],\mathbb{E}\Bigg{[}\sup_{\tau\in[t,T]}|p_{\tau}|^{2}\Bigg{]}+\mathbb{E}\Bigg{[}\int_{t}^{T}\|q_{\tau}\|^{2}_{Y^{**}}d\tau\Bigg{]}\leq c\mathbb{E}\Bigg{[}\int_{t}^{T}|\mathcal{H}(\tau,0,0)|^{2}d\tau\Bigg{]}+c\mathbb{E}\Bigg{[}|G^{g}(\mathcal{Z}_{T})|^{2}\Bigg{]},

where cc is a constant depending on TT and L:=max{L1,L2,L3}L:=\max\{L_{1},L_{2},L_{3}\}, where LiL_{i}, i=1,..,3i=1,..,3 are the coefficients in Hypothesis 3.2.

Proposition 3.8.

Assume that Hypotheses 3.2 and 3.4 hold true. Then (3.10) admits a unique solution (p,q)L2(Ω,C[0,T];Y)×L2(Ω,L2[0,T];L2(Y))(p,q)\in L^{2}(\Omega,C[0,T];Y^{*})\times L^{2}(\Omega,L^{2}[0,T];L^{2}(Y^{*})) such that the map

z(p(,,z),q(,,z))  belongs to  𝒢1(Lη(Ω;C([0,T];Y)),𝒦cont([0,T]))z\longmapsto(p(\cdot,\cdot,z),q(\cdot,\cdot,z))\text{ \emph{ belongs to } }\mathcal{G}^{1}(L^{\eta}(\Omega;C([0,T];Y^{*})),\mathcal{K}_{cont}([0,T]))

for η=(m+1)(m+2)\eta=\ell(m+1)(m+2), where 𝒦cont([0,T])\mathcal{K}_{cont}([0,T]) is the space of adapted processes (p,q)(p,q) taking values in ×Y\mathbb{R}\times Y^{**} such that pp has continuous paths and

𝔼[supτ[0,T]|pτ|2]+𝔼[τTqsY2𝑑s]<.\mathbb{E}\left[\sup_{\tau\in[0,T]}|p_{\tau}|^{2}\right]+\mathbb{E}\left[\int_{\tau}^{T}\|q_{s}\|_{Y^{**}}^{2}ds\right]<\infty.

Moreover, for every 2\ell\geq 2.

(𝔼[supt[0,T]|zp(t,z)h|])1/ChY(1+zY(m+1)2)\left(\mathbb{E}\left[\sup_{t\in[0,T]}|\nabla_{z}p(t,z)h|^{\ell}\right]\right)^{1/\ell}\leq C\|h\|_{Y^{*}}\left(1+\|z\|_{Y^{*}}^{(m+1)^{2}}\right)
Proof.

See [22] Proposition 4.2. ∎

Still aiming to prove (3.6), we provide yet another crucial result that links the directional derivative of 𝒵\mathcal{Z} to its Malliavin derivative.

Proposition 3.9.

Assume that Hypothesis 3.4 holds. Then for almost all s,τs,\tau such that tsτ<Tt\leq s\leq\tau<T we have that

Ds𝒵(τ,t,z)=z𝒵(τ,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z)),a.s.D_{s}\mathcal{Z}(\tau,t,z)=\nabla_{z}\mathcal{Z}(\tau,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z)),\ \mathbb{P}-a.s. (3.11)

moreover

Ds𝒵(T,t,z)=z𝒵(T,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z)),a.s. for almost all s.D_{s}\mathcal{Z}(T,t,z)=\nabla_{z}\mathcal{Z}(T,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z)),\ \mathbb{P}-a.s.\text{ for almost all }s. (3.12)
Proof.

Thanks to Proposition 3.5, for every s[0,T]s\in[0,T] and every direction hYh\in Y^{*}, the directional derivative process z𝒵(τ,s,z)h\nabla_{z}\mathcal{Z}(\tau,s,z)h, τ[s,T)\tau\in[s,T) solves \mathbb{P}-a.s. the equation

z𝒵(τ,t,z)h\displaystyle\nabla_{z}\mathcal{Z}(\tau,t,z)h =e(τt)𝒜h+tτe(τs)𝒜zνβg(s,𝒵(s,t,z))z𝒵(s,t,z)h𝑑s\displaystyle=e^{(\tau-t)\mathcal{A}^{*}}h+\int_{t}^{\tau}e^{(\tau-s)\mathcal{A}^{*}}\nabla_{z}\nu\beta^{g}(s,\mathcal{Z}(s,t,z))\nabla_{z}\mathcal{Z}(s,t,z)hds
+tτz(e(τs)𝒜νσg(s,𝒵(s,t,z)))z𝒵(s,t,z)h𝑑Ws,τ[t,T],\displaystyle+\int_{t}^{\tau}\nabla_{z}\left(e^{(\tau-s)\mathcal{A}^{*}}\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))\right)\nabla_{z}\mathcal{Z}(s,t,z)hdW_{s},\quad\tau\in[t,T],
z𝒵(τ,t,z)h\displaystyle\nabla_{z}\mathcal{Z}(\tau,t,z)h =h,τ[0,t),\displaystyle=h,\quad\tau\in[0,t),

Given vYv\in Y^{*} and t[0,s]t\in[0,s], we can replace zz by 𝒵(s,t,z)\mathcal{Z}(s,t,z) and hh by νσg(s,𝒵(s,t,z))v\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))v in the previous equation, since 𝒵(s,t,z)\mathcal{Z}(s,t,z) is s\mathcal{F}_{s} measurable. Note now that

𝒵(η,s,𝒵(s,t,z))=𝒵(η,t,z)a.s.,\mathcal{Z}(\eta,s,\mathcal{Z}(s,t,z))=\mathcal{Z}(\eta,t,z)\quad\mathbb{P}-a.s.,

for η[s,T)\eta\in[s,T), as a consequence of the uniqueness of the solution of (3.7). This yields

z\displaystyle\nabla_{z} 𝒵(τ,𝒵(s,t,z))νσg(s,𝒵(s,t,z))v=e(τs)𝒜νσg(s,𝒵(s,t,z))v\displaystyle\mathcal{Z}(\tau,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))v=e^{(\tau-s)\mathcal{A}^{*}}\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))v
+sτe(τη)𝒜zνβg(η,𝒵(η,t,z))z𝒵(η,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z))v𝑑η\displaystyle+\int_{s}^{\tau}e^{(\tau-\eta)\mathcal{A}^{*}}\nabla_{z}\nu\beta^{g}(\eta,\mathcal{Z}(\eta,t,z))\nabla_{z}\mathcal{Z}(\eta,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))vd\eta
+sτz(e(τη)𝒜νσg(η,𝒵(η,t,z)))z𝒵(η,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z))v𝑑Wη,\displaystyle+\int_{s}^{\tau}\nabla_{z}(e^{(\tau-\eta)\mathcal{A}^{*}}\nu\sigma^{g}(\eta,\mathcal{Z}(\eta,t,z)))\nabla_{z}\mathcal{Z}(\eta,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))vdW_{\eta},

for τ[s,T)\tau\in[s,T), \mathbb{P}-a.s. This shows that the process

{z𝒵(τ,t,𝒵(s,t,z,))νσg(s,𝒵(s,t,z))v}tsτ<T,\left\{\nabla_{z}\mathcal{Z}(\tau,t,\mathcal{Z}(s,t,z,))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))v\right\}_{t\leq s\leq\tau<T},

is a solution of the equation

Qs,τ\displaystyle Q_{s,\tau} =e(τs)𝒜νσg(s,𝒵s)v+sτe(τη)𝒜zνβg(η,𝒵η)Qs,η𝑑η\displaystyle=e^{(\tau-s)\mathcal{A}^{*}}\nu\sigma^{g}(s,\mathcal{Z}_{s})v+\int_{s}^{\tau}e^{(\tau-\eta)\mathcal{A}^{*}}\nabla_{z}\nu\beta^{g}(\eta,\mathcal{Z}_{\eta})Q_{s,\eta}d\eta
+sτz(e(τη)𝒜νσg(η,𝒵η))Qs,ηdWη,\displaystyle+\int_{s}^{\tau}\nabla_{z}(e^{(\tau-\eta)\mathcal{A}^{*}}\nu\sigma^{g}(\eta,\mathcal{Z}_{\eta}))Q_{s,\eta}dW_{\eta},

where Qs,τ:=Ds𝒵τvQ_{s,\tau}:=D_{s}\mathcal{Z}_{\tau}v. The thesis now follows from the uniqueness property, as proved e.g. in [14] Proposition 3.5. To complete the proof of (3.12), we take a sequence τnT\tau_{n}\uparrow T such that (3.11) holds for every τn\tau_{n}, and we let nn\rightarrow\infty (see [14]). The result follows from the regularity properties of D𝒵D\mathcal{Z} and z𝒵\nabla_{z}\mathcal{Z}, as well as the closedness of the operator DD on UMD Banach spaces. ∎

In this framework, using the results presented in [21] and [23] we find that:

Proposition 3.10.

Assume Hypotheses 3.2 - 3.4 Then for a.a. s,τs,\tau such that tsτTt\leq s\leq\tau\leq T we have that

Dsp(τ,t,z)\displaystyle D_{s}p(\tau,t,z) =zp(τ,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z))a.s.,\displaystyle=\nabla_{z}p(\tau,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))\quad\mathbb{P}-a.s., (3.13)
Dsq(τ,t,z)\displaystyle D_{s}q(\tau,t,z) =zq(τ,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z))a.s..\displaystyle=\nabla_{z}q(\tau,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))\quad\mathbb{P}-a.s.. (3.14)

Moreover, for a.a. s[t,T]s\in[t,T],

q(s,t,z)=zp(s,s,𝒵(s,t,z))νσg(s,𝒵(s,t,z))a.s.q(s,t,z)=\nabla_{z}p(s,s,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z))\ \mathbb{P}-a.s. (3.15)
Proof.

The proof follows the same arguments as [14, Proposition 5.6] though the spaces at play are different. Indeed, the main tools are provided in Proposition 3.9. So, thanks to the extension of Malliavin calculus to UMD Banach spaces, and the chain rule linking Malliavin derivative and Gateaux derivative (see Proposition 2.13), the result is secured. ∎

Finally, the next result provides the proof of (3.6).

Proposition 3.11.

Assume that Hypothesis 3.2 and 3.4 hold true. Then the function v(t,z):=p(t,t,z)v(t,z):=p(t,t,z) in (3.4) is continuous and for every t[0,T]t\in[0,T], v(t,)v(t,\cdot) belongs to 𝒢1(Y,)\mathcal{G}^{1}(Y^{*},\mathbb{R}) and there exists C>0C>0 such that

|zv(t,z)h|ChY(1+zY(m+1)2).|\nabla_{z}v(t,z)h|\leq C\|h\|_{Y^{*}}(1+\|z\|_{Y^{*}}^{(m+1)^{2}}).

Moreover we have that

q(s,t,z)=zv(t,𝒵(s,t,z))νσg(s,𝒵(s,t,z)).q(s,t,z)=\nabla_{z}v(t,\mathcal{Z}(s,t,z))\nu\sigma^{g}(s,\mathcal{Z}(s,t,z)).
Proof.

The first part is a corollary of Proposition 3.8. The second is derived from (3.15). ∎

3.3. The HJB equation

Formally define

t[f](x)\displaystyle\mathcal{L}_{t}[f](x) :=12Trace(Gg(t,x)Gg(t,x)2f(x))+𝒜x+νβg(t,x),f(x)Y×Y.\displaystyle:=\frac{1}{2}\text{Trace}(G^{g}(t,x)G^{g}(t,x)^{*}\ \nabla^{2}f(x))+\langle\mathcal{A}^{*}x+\nu\beta^{g}(t,x),\nabla f(x)\rangle_{Y^{*}\times Y^{**}}.

We can consider the Hamilton-Jacobi-Bellman equation associated with the control problem (2.5) - (2.8), which is given by

{wt(t,z)=tw(t,z)(t,z,w(t,z)νσg(t,z)),w(T,z)=Gg(z).\begin{cases}\frac{\partial w}{\partial t}(t,z)&=-\mathcal{L}_{t}w(t,z)-\mathcal{H}(t,z,\nabla w(t,z)\nu\sigma^{g}(t,z)),\\ w(T,z)&=G^{g}(z).\end{cases} (3.16)

A solution of this equation provides a way to compute v(t,z)v(t,z) in (3.4) by PDE methods (see e.g. [6]). The connection between (3.16) and (3.4) is actually detailed in the forthcoming Theorem 3.13 by means of the forward backward system (3.2). Later on, in Theorem 3.18, we shall see how w(t,z)w(t,z) is connected with the optimal performance, see (3.5). Thus, we are interested in finding mild solutions to the previous equation, which we are going to defined soon. This problem was tackled in [14] (for a Hilbert space) in the case of a general νσg\nu\sigma^{g}, and in [22] (for a Banach space) in the case of a constant νσg\nu\sigma^{g}. Our result is then extending the on of [22].

Let 𝒵(τ,t,ζt)\mathcal{Z}(\tau,t,\zeta_{t}) be a solution to (3.7), with 𝒜\mathcal{A}^{*}, βg\beta^{g} and σg\sigma^{g} satisfying Hypotheses 3.2 - 3.4. We recall that this solution is a YY^{*}-valued Markov process (see Remark 3.6). We can then define the transition semigroup on continuous and bounded functions φ:Y\varphi:Y^{*}\longrightarrow\mathbb{R} as

Pt,τ[φ](z)=𝔼[φ(𝒵(τ,t,z))].P_{t,\tau}[\varphi](z)=\mathbb{E}\left[\varphi(\mathcal{Z}(\tau,t,z))\right].

Moreover, we have that this semigroup is also well defined on continuous functions φ:Y\varphi:Y^{*}\longrightarrow\mathbb{R} with polynomial growth with respect to zz.

Definition 3.12.

A function w:[0,T]×Yw:[0,T]\times Y^{*}\longrightarrow\mathbb{R} is a mild solution of the Hamilton-Jacobi-Bellman equation (3.16) if:

  • For every t[0,T]t\in[0,T] w(t,)𝒢1(Y)w(t,\cdot)\in\mathcal{G}^{1}(Y^{*}), ww is continuous and (t,z)w(t,z)(t,z)\longmapsto w(t,z) is measurable from [0,T]×Y[0,T]\times Y^{*} with values in YY^{**}

  • For every t[0,T]t\in[0,T], there exists C>0C>0 such that |w(t,z)|C(1+zYj)|w(t,z)|\leq C(1+\|z\|^{j}_{Y^{*}}) and |zw(t,z)h|ChY(1+zYk)|\nabla_{z}w(t,z)h|\leq C\|h\|_{Y^{*}}(1+\|z\|^{k}_{Y^{*}}), with z,hYz,h\in Y^{*} and jj and kk positive integers.

  • The following equality holds.

    w(t,z)=Pt,T[Gg](z)+tTPt,τ[(τ,,w(τ,)νσg(t,))](z)𝑑τ,t[0,T],zY.w(t,z)=P_{t,T}[G^{g}](z)+\int_{t}^{T}P_{t,\tau}[\mathcal{H}(\tau,\cdot,\nabla w(\tau,\cdot)\nu\sigma^{g}(t,\cdot))](z)d\tau,\quad t\in[0,T],z\in Y^{*}.

In order to prove that there exists a unique solution of (3.16) we need once again the forward-backward system (3.2):

{d𝒵τ=𝒜𝒵τdτ+νβg(τ,𝒵τ)+νσg(τ,𝒵τ)dWτ,τ[t,T]𝒵t=ζtdpt=(τ,𝒵τ,qτ)dτ+qτνdWτ,τ[t,T]pT=Gg(𝒵T)\begin{cases}d\mathcal{Z}_{\tau}&=\mathcal{A}^{*}\mathcal{Z}_{\tau}d\tau+\nu\beta^{g}(\tau,\mathcal{Z}_{\tau})+\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau})dW_{\tau},\qquad\tau\in[t,T]\\ \mathcal{Z}_{t}&=\zeta_{t}\\ dp_{t}&=-\mathcal{H}(\tau,\mathcal{Z}_{\tau},q_{\tau})d\tau+q_{\tau}\nu dW_{\tau},\qquad\tau\in[t,T]\\ p_{T}&=G^{g}(\mathcal{Z}_{T})\end{cases}
Theorem 3.13.

Assume that GgG^{g} and \mathcal{H} satisfy Hypothesis 3.2 and that Hypothesis 3.4 hold true. Then there exists a unique mild solution of the Hamilton-Jacobi-Bellman equation (3.16) given by

w(t,ζt)=v(t,ζt)w(t,\zeta_{t})=v(t,\zeta_{t}) (3.17)

where (𝒵,p,q)(\mathcal{Z},p,q) is the solution of (3.2) and v(t,z)=p(t,t,z)v(t,z)=p(t,t,z), see (3.4).

Proof.

The proof is based on arguments similar to [22, Theorem 6.2] and [14, Theorem 6.2], though adapted to the current framework. Note that the main difference with [14] is the nature of the spaces considered. ∎

3.4. Solving the optimal control problem

As we have proven the identification (3.6) to be true (see Proposition 3.11), we can finally move to the study of the optimal control problem (2.5) - (2.8). As anticipated, we want to show that the optimal value

infu𝔸Jg(t,ζt,u)=Jg(t,ζt,u^)=v(t,ζt),\inf_{u\in\mathds{A}}J^{g}(t,\zeta_{t},u)=J^{g}(t,\zeta_{t},\hat{u})=v(t,\zeta_{t}),

where we have defined v(t,ζt)=p(t,t,ζt)v(t,\zeta_{t})=p(t,t,\zeta_{t}) in (3.4), (p,q)(p,q) solve the backward stochastic differential equation (3.2) and a solution of v(t,ζt)v(t,\zeta_{t}) can be obtained through the HJB equation (3.16) (see Theorem 3.13). We define the, possibly empty, set

Γ(τ,z,ξ)={u𝒰:Fg(τ,z,u)+ξνRg(τ,z,u)=(τ,z,ξ)},\Gamma(\tau,z,\xi)=\left\{u\in\mathcal{U}:F^{g}(\tau,z,u)+\xi\nu R^{g}(\tau,z,u)=\mathcal{H}(\tau,z,\xi)\right\}, (3.18)

where τ[t,T]\tau\in[t,T], zYz\in Y^{*}, ξY\xi\in Y^{**}.

Hypothesis 3.14.

We notice that, intuitively, Γ(t,z,ξ)\Gamma(t,z,\xi) represents the set of cotrols that allow us to obtain the minimum in the Hamiltonian (3.3). We will thus assume that for all τ[t,T]\tau\in[t,T], zYz\in Y^{*}, ξY\xi\in Y^{**}, Γ(τ,z,ξ)\Gamma(\tau,z,\xi)\neq\emptyset

Remark 3.15.

Thanks to the Filippov theorem (see [4]), being Γ(τ,z,ξ)\Gamma(\tau,z,\xi) non empty for all τ[t,T]\tau\in[t,T], zYz\in Y, ξY\xi\in Y^{**}, there exists a Borel measurable map Γ0:[0,T]×Y×Y𝒰\Gamma_{0}:[0,T]\times Y^{*}\times Y^{**}\longrightarrow\mathcal{U} such that, for t[0,T]t\in[0,T], zYz\in Y^{*} and ξY\xi\in Y^{**}, Γ0(t,z,ξ)Γ(t,z,ξ)\Gamma_{0}(t,z,\xi)\in\Gamma(t,z,\xi).

Proposition 3.16.

Assume that Hypothesis 3.1 - 3.2 and 3.4 and hold true, and let vv be defined in (3.4) and uu in 𝔸\mathds{A}. Then for all t[0,T]t\in[0,T] and zYz\in Y^{*}, we have that

Jg(t,z,u)v(t,z).J^{g}(t,z,u)\geq v(t,z).
Proof.

Let uu in 𝔸\mathds{A} and take 𝒵τu\mathcal{Z}_{\tau}^{u} to be a solution of (3.1) corresponding to the control uu (for the existence of such a solution see Corollary 3.7). Define

Wτu=Wτ+tττRg(s,𝒵su,us)𝑑s,τ[0,T].W_{\tau}^{u}=W_{\tau}+\int_{t\vee\tau}^{\tau}R^{g}(s,\mathcal{Z}_{s}^{u},u_{s})ds,\quad\tau\in[0,T].

We notice that 𝒵τu\mathcal{Z}_{\tau}^{u} solves the equation

{d𝒵τu=𝒜𝒵τudτ+νβg(τ,𝒵τu)dτ+νσg(τ,𝒵τu)dWτu,𝒵t=ζt,\begin{cases}d\mathcal{Z}^{u}_{\tau}=&\mathcal{A}^{*}\mathcal{Z}_{\tau}^{u}d\tau+\nu\beta^{g}(\tau,\mathcal{Z}^{u}_{\tau})d\tau+\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau})dW^{u}_{\tau},\\ \mathcal{Z}_{t}=&\zeta_{t},\end{cases}

and, being RR and thus RgR^{g} bounded, we can find a probability u\mathbb{P}^{u} equivalent to \mathbb{P} such that WuW^{u} is a Wiener process on \mathbb{R} and thus νWu\nu W^{u} is a cylindrical Wiener process with values in YY^{*} (see [10, Theorem 7.2 (iii)]). We consider the backward equation w.r.t. u\mathbb{P}^{u} for the unknowns (pτu,qτu)(p_{\tau}^{u},q_{\tau}^{u}), τ[t,T]\tau\in[t,T] given by

pτu+τTqsuν𝑑Wsu=Gg(𝒵Tu)+τT(s,𝒵su,qsu)𝑑s.p_{\tau}^{u}+\int_{\tau}^{T}q_{s}^{u}\nu dW_{s}^{u}=G^{g}(\mathcal{Z}_{T}^{u})+\int_{\tau}^{T}\mathcal{H}(s,\mathcal{Z}_{s}^{u},q_{s}^{u})ds. (3.19)

By taking (pu,qu)(p^{u},q^{u}) at τ=t\tau=t in (3.19), we get that p(t,t,ζt)p(t,t,\zeta_{t}) depends only on t,ζt,βg,σg,Ggt,\zeta_{t},\beta^{g},\sigma^{g},G^{g} and \mathcal{H}. With the same approach as in [22, Proposition 5.5], one obtains immediately that τTqsuν𝑑Wsu\int_{\tau}^{T}q_{s}^{u}\nu dW^{u}_{s} is actually a u\mathbb{P}^{u}-martingale.

By recalling that v(t,z):=p(t,t,z)v(t,z):=p(t,t,z) (see (3.4)) and taking expectation with respect to the original probability \mathbb{P}, we obtain that

v(t,ζt)=𝔼[Gg(𝒵Tu)]+𝔼[tT(s,𝒵su,qsu)qsuνR(s,𝒵su,us)ds].v(t,\zeta_{t})=\mathbb{E}[G^{g}(\mathcal{Z}_{T}^{u})]+\mathbb{E}\left[\int_{t}^{T}\mathcal{H}(s,\mathcal{Z}_{s}^{u},q_{s}^{u})-q_{s}^{u}\nu R(s,\mathcal{Z}_{s}^{u},u_{s})ds\right].

Adding and subtracting 𝔼[tTFg(s,𝒵su,us)𝑑s]\mathbb{E}\left[\int_{t}^{T}F^{g}(s,\mathcal{Z}_{s}^{u},u_{s})ds\right], we arrive at

v(t,ζt)=Jg(t,ζt,u)+𝔼[tT(s,𝒵su,qsu)qsuνRg(s,𝒵su,us)Fg(s,𝒵su,us)ds].v(t,\zeta_{t})=J^{g}(t,\zeta_{t},u)+\mathbb{E}\left[\int_{t}^{T}\mathcal{H}(s,\mathcal{Z}_{s}^{u},q_{s}^{u})-q_{s}^{u}\nu R^{g}(s,\mathcal{Z}_{s}^{u},u_{s})-F^{g}(s,\mathcal{Z}_{s}^{u},u_{s})ds\right]. (3.20)

Noticing that

(s,𝒵su,qsu)qsuνRg(s,𝒵su,us)Fg(s,𝒵su,us)0,\mathcal{H}(s,\mathcal{Z}_{s}^{u},q_{s}^{u})-q_{s}^{u}\nu R^{g}(s,\mathcal{Z}_{s}^{u},u_{s})-F^{g}(s,\mathcal{Z}_{s}^{u},u_{s})\leq 0,

by definition of \mathcal{H} in (3.3), we conclude. ∎

Corollary 3.17.

Let t[0,T]t\in[0,T] and ζtY\zeta_{t}\in Y^{*}. If Jg(t,ζt,u)=v(t,ζt)J^{g}(t,\zeta_{t},u^{*})=v(t,\zeta_{t}) then uu^{*} is optimal for the control problem starting from ζt\zeta_{t} at time tt. Assume Hypothesis 3.14 holds and take Γ0(τ,z,ξ)\Gamma_{0}(\tau,z,\xi) be the Borel measurable map defined in Remark 3.15. Then, an admissible control satisfying

u^τ=Γ0(τ,𝒵τu^,qτu^),a.s. for a.e. τ[t,T]\hat{u}_{\tau}=\Gamma_{0}(\tau,\mathcal{Z}_{\tau}^{\hat{u}},q_{\tau}^{\hat{u}}),\quad\mathbb{P}-a.s.\text{ for a.e. }\tau\in[t,T]

is optimal and Jg(t,z,u^)=v(t,z)J^{g}(t,z,\hat{u})=v(t,z).

Proof.

The proof is in line with the one in [22, Corollary 5.6]. ∎

Theorem 3.18.

Assume that Hypothesis 3.1, 3.2, 3.4 and 3.14 hold true. For all admissible controls uu in 𝔸\mathds{A}, we have that

Jg(t,ζt,u)v(t,ζt),{J^{g}}(t,\zeta_{t},u)\geq v(t,\zeta_{t}),

and the equality holds true if and only if

u^τΓ(τ,𝒵τu,v(τ,𝒵τu)νσg(τ,𝒵τu))a.s. for a.a. τ[t,T].\hat{u}_{\tau}\in\Gamma(\tau,\mathcal{Z}^{u}_{\tau},\nabla v(\tau,\mathcal{Z}_{\tau}^{u})\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau}))\quad\mathbb{P}-a.s.\text{ for a.a. }\tau\in[t,T]. (3.21)

Moreover, let us denote by Γ0(τ,z,ξ)\Gamma_{0}(\tau,z,\xi) be the measurable selection of Γ(t,z,ξ)\Gamma(t,z,\xi) defined in Remark 3.15. A control satisfying the feedback law, defined as:

uτ=Γ0(τ,𝒵τu,v(τ,𝒵τu)νσg(τ,𝒵τu))a.s. for a.a. τ[t,T]u_{\tau}=\Gamma_{0}(\tau,\mathcal{Z}^{u}_{\tau},\nabla v(\tau,\mathcal{Z}^{u}_{\tau})\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau}^{u}))\quad\mathbb{P}-a.s.\text{ for a.a. }\ \tau\in[t,T] (3.22)

is optimal. Define the closed loop equation:

{𝒵~τ=[𝒜𝒵~τ+νβg(τ,𝒵~τ)+νσg(τ,𝒵~τ)Rg(τ,𝒵~τ,Γ0(v(τ,𝒵~τ)νσg(τ,𝒵~τ)))]dτ+νσg(τ,𝒵~τ)dWτ,τ[t,T]𝒵~t=ζt.\begin{cases}\widetilde{\mathcal{Z}}_{\tau}=&\left[\mathcal{A}^{*}\widetilde{\mathcal{Z}}_{\tau}+\nu\beta^{g}(\tau,\widetilde{\mathcal{Z}}_{\tau})+\nu\sigma^{g}(\tau,\widetilde{\mathcal{Z}}_{\tau})R^{g}\Bigg{(}\tau,\widetilde{\mathcal{Z}}_{\tau},\Gamma_{0}\Big{(}\nabla v(\tau,\widetilde{\mathcal{Z}}_{\tau})\nu\sigma^{g}(\tau,\widetilde{\mathcal{Z}}_{\tau})\Big{)}\Bigg{)}\right]d\tau\\ &+\nu\sigma^{g}(\tau,\widetilde{\mathcal{Z}}_{\tau})dW_{\tau},\ \ \tau\in[t,T]\\ \widetilde{\mathcal{Z}}_{t}=&\zeta_{t}.\end{cases} (3.23)

Then (3.23) admits a weak solution which is unique in law, and the corresponding pair (u,𝒵~u)(u,\widetilde{\mathcal{Z}}^{u}) is optimal.

For more details about the definition of feedback law and closed loop equation in the case of optimal control for Hilbert spaces we refer to [13, Section 2.5].

Proof.

Using (3.20) and Proposition 3.10 we can rewrite v(t,ζt)v(t,\zeta_{t}) as

v(t,ζt)=\displaystyle v(t,\zeta_{t})= Jg(t,ζt,u)+tT[(τ,𝒵τu,v(τ,𝒵τu)νσg(τ,𝒵τu))\displaystyle J^{g}(t,\zeta_{t},u)+\int_{t}^{T}\left[\mathcal{H}(\tau,\mathcal{Z}^{u}_{\tau},\nabla v(\tau,\mathcal{Z}^{u}_{\tau})\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau}))\right.
v(τ,𝒵τu)νσg(τ,𝒵τu))Rg(τ,𝒵tu,uτ)Fg(τ,𝒵τu,uτ)]dτ.\displaystyle\left.-\nabla v(\tau,\mathcal{Z}^{u}_{\tau})\nu\sigma^{g}(\tau,\mathcal{Z}^{u}_{\tau}))R^{g}(\tau,\mathcal{Z}_{t}^{u},u_{\tau})-F^{g}(\tau,\mathcal{Z}^{u}_{\tau},u_{\tau})\right]d\tau.

The proof of the first statement now follows from Corollary 3.17. The closed loop equation can be solved in the weak sense via a Girsanov change of measure. Recall that (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) is the probability space on which the Wiener process (Wτ)τ0(W_{\tau})_{\tau\geq 0} in (3.2) is defined. Define (W^τ)τ0(\hat{W}_{\tau})_{\tau\geq 0} as

W^t:=Wt0tRg(τ,𝒵~τu,Γ0(τ,𝒵~τ,v(s,𝒵~τ)νσg(τ,𝒵~τ)))𝑑τ.\hat{W}_{t}:=W_{t}-\int_{0}^{t}R^{g}\Bigg{(}\tau,\tilde{\mathcal{Z}}_{\tau}^{u},\Gamma_{0}\Big{(}\tau,\tilde{\mathcal{Z}}_{\tau},\nabla v(s,\tilde{\mathcal{Z}}_{\tau})\nu\sigma^{g}(\tau,\tilde{\mathcal{Z}}_{\tau})\Big{)}\Bigg{)}d\tau.

Due to the Girsanov theorem there exists a probability ^\hat{\mathbb{P}} on Ω\Omega such taht W^τ\hat{W}_{\tau} is a Wiener process. We then notice that νW\nu W and νW^\nu\hat{W} are cylindrical Wiener processes with values in YY^{*}, and that the closed loop equation (3.23) can be rewritten under ^\hat{\mathbb{P}} as

{d𝒵~τu=𝒜𝒵~τudτ+νβg(τ,𝒵~τu)dτ+νσg(τ,𝒵~τu)dW^τ,𝒵t=ζt.\begin{cases}d\tilde{\mathcal{Z}}^{u}_{\tau}=&\mathcal{A}^{*}\tilde{\mathcal{Z}}_{\tau}^{u}d\tau+\nu\beta^{g}(\tau,\tilde{\mathcal{Z}}^{u}_{\tau})d\tau+\nu\sigma^{g}(\tau,\tilde{\mathcal{Z}}^{u}_{\tau})d\hat{W}_{\tau},\\ \mathcal{Z}_{t}=&\zeta_{t}.\end{cases}

Then, thanks to Proposition 3.5, we have a unique solution to this new process related to the probability ^\hat{\mathbb{P}} and the Wiener process W^τ\hat{W}_{\tau}, which implies that also the closed loop equation (3.23) always admits a solution in the weak sense. Thanks to Hypothesis 3.14, we know that Γ(τ,𝒵τu,v(τ,𝒵τu)νσg(τ,𝒵τu))\Gamma(\tau,\mathcal{Z}_{\tau}^{u},\nabla v(\tau,\mathcal{Z}_{\tau}^{u})\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau}^{u})) is non empty and thus, by the Filippov theorem, a measurable selection Γ0(τ,𝒵τu,v(τ,𝒵τu)νσg(τ,𝒵τu))\Gamma_{0}(\tau,\mathcal{Z}_{\tau}^{u},\nabla v(\tau,\mathcal{Z}_{\tau}^{u})\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau}^{u})) of Γ\Gamma exists and the optimal control can be obtained. This proof is in line with [22, Theorem 5.7] and [14, Theorem 7.2]. ∎

Remark 3.19.

Notice that, having solved the lifted optimization problem, thanks to (2.2)-(2.7), we have also solved the original problem (1.1)-(1.2). Indeed, we have that a control u^\hat{u} which is optimal for (2.8) where the dynamics for the forward process are given by 𝒵τu\mathcal{Z}^{u}_{\tau} in (3.1) is also optimal for the original problem (1.1), as Jg(t,ζt,u)=J(t,x,u)J^{g}(t,\zeta_{t},u)=J(t,x,u) by definition and Xu(t)=g,𝒵tuX^{u}(t)=\langle g,\mathcal{Z}_{t}^{u}\rangle.

4. A problem of optimal consumption

A cash flow admits consumption with rate cc according to the forward dynamics

Xtc=x(t)+0tK(tτ)μ(τ,Xτc)𝑑τ+0tK(tτ)σ(τ,Xτc)(R(cτ)dτ+dWτ),X^{c}_{t}=x(t)+\int_{0}^{t}K(t-\tau)\ \mu(\tau,X^{c}_{\tau})d\tau+\int_{0}^{t}K(t-\tau)\ \sigma(\tau,X_{\tau}^{c})(-R(c_{\tau})d\tau+dW_{\tau}), (4.1)

where x:[0,T]x:[0,T]\longrightarrow\mathbb{R}, K(tτ)=tτK(t-\tau)=\sqrt{t-\tau}, μ\mu and σ\sigma satisfy Hypothesis 3.4, and RR satisfy Hypothesis 3.1. In this case we lift K(t):=tK(t):=\sqrt{t} on Lq()×Lp()L^{q}(\mathbb{R})\times L^{p}(\mathbb{R}) for q(1,2)q\in(1,2) and pp such that 1p+1q=1\frac{1}{p}+\frac{1}{q}=1 by considering

e𝒜t:Lp()Lp()e^{\mathcal{A}^{*}t}:L^{p}(\mathbb{R})\longrightarrow L^{p}(\mathbb{R})

the left shift semigroup defined as (e𝒜tf)(s)=f(st)(e^{\mathcal{A}^{*}t}f)(s)=f(s-t), for all fLp()f\in L^{p}(\mathbb{R}), g(x)=12x𝟙[0,1](x)g(x)=\frac{1}{2\sqrt{x}}\mathds{1}_{[0,1]}(x), ν(x)=𝟙[1,0](x)\nu(x)=\mathds{1}_{[-1,0]}(x). Then we have that

K(t)=g,e𝒜tνLp()×Lq()=12x𝟙[0,1](x)𝟙[1,0](xt)𝑑x=0tdx2x=t.K(t)=\langle g,e^{\mathcal{A}^{*}t}\nu\rangle_{L^{p}(\mathbb{R})\times L^{q}(\mathbb{R})}=\int_{\mathbb{R}}\frac{1}{2\sqrt{x}}\mathds{1}_{[0,1]}(x)\mathds{1}_{[-1,0]}(x-t)dx=\int_{0}^{t}\frac{dx}{2\sqrt{x}}=\sqrt{t}.

We consider a classical optimal control problem given by the maximization of the performance functional

J(t,x,c)=𝔼[tTF(τ,Xτc,cτ)𝑑τ+G(XTc)],J(t,x,c)=\mathbb{E}\left[\int_{t}^{T}F(\tau,X_{\tau}^{c},c_{\tau})d\tau+G(X_{T}^{c})\right], (4.2)

for some functions F(τ,Xτc,cτ)=F(cτ):=a1cτ2F(\tau,X_{\tau}^{c},c_{\tau})=F(c_{\tau}):=-a_{1}c_{\tau}^{2} and G(Xtc):=a2XTcG(X_{t}^{c}):=a_{2}X_{T}^{c} (a1,a2>0a_{1},a_{2}\in\mathbb{R}_{>0}) satisfying Hypothesis 3.2. Linear-quadratic performance functionals such as (4.2) appear, for example, when considering optimal advertising problems (see e.g. [16, 17] and [15]). In this case we have that the stochastic control problem can be reformulated in YY^{*} with forward dynamics given by

𝒵τc=ζ0+0τ𝒜𝒵sc+νμg(s,𝒵sc)ds+νσg(s,𝒵sc)(R(cs)ds+dWs),\mathcal{Z}_{\tau}^{c}=\zeta_{0}+\int_{0}^{\tau}\mathcal{A}^{*}\mathcal{Z}_{s}^{c}+\nu\mu^{g}(s,\mathcal{Z}_{s}^{c})ds+\nu\sigma^{g}(s,\mathcal{Z}_{s}^{c})(-R(c_{s})ds+dW_{s}), (4.3)

with ζt=e𝒜tζ\zeta_{t}=e^{\mathcal{A}^{*}t}\zeta such that x(t)=g,ζtx(t)=\langle g,\zeta_{t}\rangle for all t[0,T]t\in[0,T]. The goal is to minimize

Jg(0,ζ0,c)=𝔼[0Ta1cτ2dτ+a2g,𝒵Tc].J^{g}(0,\zeta_{0},c)=\mathbb{E}\left[\int_{0}^{T}-a_{1}c_{\tau}^{2}d\tau+a_{2}\langle g,\mathcal{Z}^{c}_{T}\rangle\right]. (4.4)

In this case the Hamiltonian functional (3.3) is given by

(t,𝒵,ξ)=infc𝒰[a1c2ξνR(c)]\mathcal{H}(t,\mathcal{Z},\xi)=\inf_{c\in\mathcal{U}}[-a_{1}c^{2}-\xi\nu R(c)] (4.5)

and the forward-backward system is

{d𝒵τ=𝒜𝒵τdτ+νμg(s,𝒵τ)dτ+νσg(s,𝒵s)dWτ,τ[0,T],𝒵0=ζ0,dpτ=(τ,𝒵τc,qτ)dτ+qτνdWτ,τ[t,T],pT=a2g,𝒵T.\begin{cases}d\mathcal{Z}_{\tau}&=\mathcal{A}^{*}\mathcal{Z}_{\tau}d\tau+\nu\mu^{g}(s,\mathcal{Z}_{\tau})d\tau+\nu\sigma^{g}(s,\mathcal{Z}_{s})dW_{\tau},\quad\tau\in[0,T],\\ \mathcal{Z}_{0}&=\zeta_{0},\\ dp_{\tau}&=-\mathcal{H}(\tau,\mathcal{Z}_{\tau}^{c},q_{\tau})d\tau+q_{\tau}\nu dW_{\tau},\quad\tau\in[t,T],\\ p_{T}&=a_{2}\langle g,\mathcal{Z}_{T}\rangle.\end{cases} (4.6)

In particular, using (4.5), we have that

pt=a2g,𝒵TtTinfc𝒰(a1c2qsνR(c))ds+qsνdWss[t,T].p_{t}=a_{2}\langle g,\mathcal{Z}_{T}\rangle-\int_{t}^{T}\inf_{c\in\mathcal{U}}\left(-a_{1}c^{2}-q_{s}\nu R(c)\right)ds+q_{s}\nu dW_{s}\ s\in[t,T].

We thus get that the set Γ\Gamma defined in (3.18) is

Γ(t,𝒵,ξ)={c𝒰:a1c2ξνR(c)=(t,𝒵,c)},\Gamma(t,\mathcal{Z},\xi)=\Big{\{}c\in\mathcal{U}:-a_{1}c^{2}-\xi\nu R(c)=\mathcal{H}(t,\mathcal{Z},c)\Big{\}}, (4.7)

and thus the optimal uτu_{\tau} can be characterized by Theorem 3.18 as

cτ=Γ0(τ,𝒵τc,v(τ,𝒵τc)νσg(τ,𝒵τc))a.s. for a.a. τ[0,T],c_{\tau}=\Gamma_{0}(\tau,\mathcal{Z}^{c}_{\tau},\nabla v(\tau,\mathcal{Z}_{\tau}^{c})\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau}^{c}))\quad\ \mathbb{P}-a.s.\text{ for a.a. }\tau\in[0,T], (4.8)

for a certain function Γ0\Gamma_{0} such that Γ0(t,𝒵,ξ)Γ(t,𝒵,ξ)\Gamma_{0}(t,\mathcal{Z},\xi)\in\Gamma(t,\mathcal{Z},\xi). In this case the HJB equations (3.16) become

{vt(t,ζ0)=tv(t,𝒵(t,0,ζ0))infc𝒰[Fg(t,𝒵(t,0,ζ0),ct)v(t,𝒵(t,0,ζ0))νσ(t,𝒵(t,0,ζ0))ct],v(T,ζ0)=a2X(0).\begin{cases}\frac{\partial v}{\partial t}(t,\zeta_{0})&=-\mathcal{L}_{t}v(t,\mathcal{Z}(t,0,\zeta_{0}))\\ &\quad-\inf_{c\in\mathcal{U}}[F^{g}(t,\mathcal{Z}(t,0,\zeta_{0}),c_{t})-\nabla v(t,\mathcal{Z}(t,0,\zeta_{0}))\nu\sigma(t,\mathcal{Z}(t,0,\zeta_{0}))c_{t}],\\ v(T,\zeta_{0})&=a_{2}X(0).\end{cases} (4.9)

Where, we remind, 𝒵(τ,0,ζ0)=𝒵τ\mathcal{Z}(\tau,0,\zeta_{0})=\mathcal{Z}_{\tau}, τ[t,T]\tau\in[t,T], 𝒵(0)=ζ0\mathcal{Z}(0)=\zeta_{0} and

t[v](t,t,z)\displaystyle\mathcal{L}_{t}[v](t,t,z) :=12Trace(Gg(v(t,z)Gg(v(t,z))2v(t,t,z))\displaystyle:=\frac{1}{2}\text{Trace}(G^{g}(v(t,z)G^{g}(v(t,z))^{*}\ \nabla^{2}v(t,t,z))
+𝒜v(t,t,z)+νβg(t,v(t,t,z)),v(t,t,z)Lq([0,))×Lp([0,)).\displaystyle\quad+\langle\mathcal{A}^{*}v(t,t,z)+\nu\beta^{g}(t,v(t,t,z)),\nabla v(t,t,z)\rangle_{L^{q}([0,\infty))\times L^{p}([0,\infty))}.

For more details about solving the HJB equation (4.9) we refer to [15]. Now, thanks to Theorem 3.13 we have that

Theorem 4.1.

Equation (4.9) has a unique mild solution vv. If the cost is given by (4.4), then for all admissible couples (c,z)(c,z) we have that J(t,z,c)v(t,z)J(t,z,c)\geq v(t,z), and the equality holds if and only if c^Γ(t,𝒵,ξ)\hat{c}\in\Gamma(t,\mathcal{Z},\xi) in (4.7), characterized as (4.8). Vice versa, if (4.8) holds, then

𝒵~τ\displaystyle\tilde{\mathcal{Z}}_{\tau} =𝒜𝒵~τ+νμg(τ,𝒵τc)dτ\displaystyle=\mathcal{A}^{*}\tilde{\mathcal{Z}}_{\tau}+\nu\mu^{g}(\tau,\mathcal{Z}_{\tau}^{c})d\tau
+νσg(τ,𝒵τc)(R(Γ0(τ,𝒵~τc,v(τ,𝒵~τc)νσg(τ,𝒵~τc)))dτ+dWτ),τ[0,T],\displaystyle\quad+\nu\sigma^{g}(\tau,\mathcal{Z}_{\tau}^{c})\Big{(}-R(\Gamma_{0}(\tau,\tilde{\mathcal{Z}}^{c}_{\tau},\nabla v(\tau,\tilde{\mathcal{Z}}_{\tau}^{c})\nu\sigma^{g}(\tau,\tilde{\mathcal{Z}}_{\tau}^{c})))d\tau+dW_{\tau}\Big{)},\quad\tau\in[0,T],

with initial condition 𝒵~0=ζ0\tilde{\mathcal{Z}}_{0}=\zeta_{0}, admits a weak solution, which is unique in law, and the corresponding pair (c,𝒵~c)(c,\tilde{\mathcal{Z}}^{c}) is optimal.

Remark 4.2.

This gives us a characterization of the optimal control for the lifted problem (4.4) and thus also for (4.2): in fact in our case we have that Xt=1,𝒵tX_{t}=\langle 1,\mathcal{Z}_{t}\rangle. Thanks to this we are able to find the optimal process Xu^X^{\hat{u}}. Moreover, we note that, the optimal control for the lifted problem and the optimal control for the original problem coincide., This allows us to retrieve the optimal pair (u^,Xu^)(\hat{u},X^{\hat{u}}) for the optimal control problem (4.2). Lastly we notice that the HJB equations (4.9), gives us the optimal value Jg(0,ζ0,u^)J^{g}(0,\zeta_{0},\hat{u}), which in turn gives us the optimal value of J(t,x,u^)J(t,x,\hat{u}), thanks to (2.7).

Remark 4.3.

Inspired by [11], we could also have considered the kernel K(t)=1t+εK(t)=\frac{1}{t+\varepsilon}, ε>0\varepsilon>0 in (4.1). In this case we can take the space of Lp([0,))L^{p}([0,\infty)) of LpL^{p} functions on [0,)[0,\infty) and its dual Lq([0,))L^{q}([0,\infty)) of measures with density in LqL^{q}, where 1p+1q=1\frac{1}{p}+\frac{1}{q}=1 and p>1p>1. We have that

K(t)=1t+ε=g,𝒮tνLp([0,))×Lq([0,)),K(t)=\frac{1}{t+\varepsilon}=\langle g,\mathcal{S}_{t}^{*}\nu\rangle_{L^{p}([0,\infty))\times L^{q}([0,\infty))},

where g=ν=exε/2g=\nu=e^{-x\varepsilon/2}, and 𝒮t=etx\mathcal{S}_{t}^{*}=e^{-tx} i.e. K(t)K(t) is the Laplace trasnform of exεe^{-x\varepsilon}. We notice that kernel is liftable (see Definition 2.6) and that we are in a UMD Banach space. It is clear that eεx/2e^{-\varepsilon x/2} is actually in Lp([0,))L^{p}([0,\infty)) for all p1p\geq 1. In particular we can also take p=q=2p=q=2 and work on the Hilbert space L2([0,))L^{2}([0,\infty)).

Acknowledgment

We would like to thank Anton Yurchenko-Tytarenko and Dennis Schroers for the nice input on kernel decompositions. The research leading to these results is within the project STORM: Stochastics for Time-Space Risk Models, receiving founding from the Research Council of Norway (RCN). Project number: 274410.

References

  • [1] E. Abi Jaber, E. Miller, and H. Pham. Linear-Quadratic control for a class of stochastic Volterra equations: solvability and approximation. The annals of probability, 31:2244–2274, 2021.
  • [2] N. Agram and B. Øksendal. Malliavin Calculus and Optimal Control of Stochastic Volterra Equations. Journal of Optimization Theory and Applications, 167:1070–1094, 2015.
  • [3] N. Agram, B. Øksendal, and S. Yakhlef. Optimal Control of Forward-Backward Stochastic Volterra Equations. Non-linear Partial Differential Equations, Mathematical Physics, and Stochastic Analysis: The Helge Holden Anniversary Volume, pages 3–36, 2009.
  • [4] J.P. Aubin and H. Frankowska. Set-valued analysis. Modern Birkhäuser Classics, 1990.
  • [5] S. Bonaccorsi and F. Confortola. Optimal control for stochastic Volterra equations with multiplicative Lévy noise. Nonlinear differential equations and applications, pages 1–26, 2020.
  • [6] P. Cannarsa and G. Da Prato. Second-order Hamilton–Jacobi equations in infinite dimensions. SIAM Journal on Control and Optimization, 29(2):474–492, 1991.
  • [7] C. Cuchiero and G. Di Nunno. Notes - Markovian lifts. to appear, 2019.
  • [8] C. Cuchiero and J. Teichman. Markovian lifts of positive semidefinite affine Volterra-type processes. Decisions in Economics and Finance, 42:407–448, 2019.
  • [9] C. Cuchiero and J. Teichman. Generalized Feller processes and Markovian lifts of stochastic Volterra processes: the affine case. Journal of Evolution Equations, pages 1–48, 2020.
  • [10] G. Da Prato and J. Zabczyk. Stochastic equations in infinite dimensions. Encyclopedia of Mathematics and its Applications. Cambridge University Press, 2 edition, 2014.
  • [11] G. Di Nunno, A. Fiacco, and E. Hove Karlsen. On the approximation of Lévy driven Volterra processes and their integrals. Journal of Mathematical Analysis and Applications, 476:120–148, 2019.
  • [12] G. Di Nunno and M. Giordano. Stochastic Volterra equations with time-changed Lévy noise and maximum principles. Annals of Operations Research, 2023.
  • [13] G. Fabbri, F. Gozzi, and A. Swiech. Stochastic Optimal Control in Infinite Dimension Dynamic Programming and HJB Equations. Springer, 2017.
  • [14] M. Fuhrman and G. Tessitore. Nonlinear Kolmogorov equations in infinite dimensional spaces: the backward stochastic differential equations approach and applications to optimal control. The Annals of Probability, 30:1397–1465, 2002.
  • [15] M. Giordano and A. Yurchenko-Tytarenko. Optimal control in linear stochastic advertising models with memory. ArXiv, 2021.
  • [16] F. Gozzi and C. Marinelli. Stochastic optimal control of delay equations arising in advertising models. In Stochastic Partial Differential Equations and Applications, Lecture Notes in Pure and Applied Mathematics, page 133–148. Chapman and Hall/CRC, 2005.
  • [17] F. Gozzi, C. Marinelli, and S. Savin. On controlled linear diffusions with delay in a model of optimal advertising under uncertainty with memory effects. Journal of optimization theory and applications, 142(2):291–321, 2009.
  • [18] C. Hernandez and D. Possamaï. A unified approach to well-posedness of type-I backward stochastic Volterra integral equations. ArXiv, 2020.
  • [19] T. Hytönen, J. Van Neerven, M. Veraar, and L. Weis. Analysis in Banach Spaces, volume I: Martingales and Littlewood-Paley Theory. Springer, Cham, 2016.
  • [20] C. Li and W. Zhen. Stochastic optimal control problem in advertising model with delay. Journal of Systems Science and Complexity, 33:968–987, 2020.
  • [21] J. Maas and J. Van Neerven. A Clark-Ocone formula in UMD Banach spaces. Electronic communications in probability, 2008.
  • [22] F. Masiero. Stochastic optimal control problems and parabolic equations in Banach spaces. SIAM Journal on Control and Optimization, 47:251–300, 2008.
  • [23] M. Pronk and M. Veraar. Tools for Malliavin calculus in UMD Banach spaces. Potential Analysis, 40:307–344, 2014.
  • [24] J. Prüss. Evolutionary integral equations and applications, volume 87. Birkhäuser, 2013.
  • [25] M. Saeedian, M. Khalighi, N. Azimi-Tafreshi, G.R. Jafari, and M. Ausloos. Memory effects on epidemic evolution: The susceptible-infected-recovered epidemic model. Physical Review E, 95(2):022409, 2017.
  • [26] J. Van Neerven. Stochastic evolution equations. Lecture notes, 2007.
  • [27] J. Van Neerven and M. Veraar. On the stochastic Fubini theorem in infinite dimensions. In In Stochastic partial differential equations and applications—VII, volume 245 of Lect, 2005.
  • [28] J. Van Neerven, M. Veraar, and L. Weis. Stochastic integration in UMD Banach spaces. The Annals of Probability, 35, 2007.
  • [29] J. Yong. Backward Stochastic Volterra Integral Equations and some Related Problems. Stochastic Processes and their Applications, 116:770–795, 2006.
  • [30] J. Yong. Well-Posedness and Regularity of Backward Stochastic Volterra Integral Equations. Probability Theory and Related Fields, 142:21–77, 2007.