Startseite An optimal control of a risk-sensitive problem for backward doubly stochastic differential equations with applications
Artikel Öffentlich zugänglich

An optimal control of a risk-sensitive problem for backward doubly stochastic differential equations with applications

  • Dahbia Hafayed und Adel Chala EMAIL logo
Veröffentlicht/Copyright: 17. Januar 2020
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

In this paper, we are concerned with an optimal control problem where the system is driven by a backward doubly stochastic differential equation with risk-sensitive performance functional. We generalized the result of Chala [A. Chala, Pontryagin’s risk-sensitive stochastic maximum principle for backward stochastic differential equations with application, Bull. Braz. Math. Soc. (N. S.) 48 2017, 3, 399–411] to a backward doubly stochastic differential equation by using the same contribution of Djehiche, Tembine and Tempone in [B. Djehiche, H. Tembine and R. Tempone, A stochastic maximum principle for risk-sensitive mean-field type control, IEEE Trans. Automat. Control 60 2015, 10, 2640–2649]. We use the risk-neutral model for which an optimal solution exists as a preliminary step. This is an extension of an initial control system in this type of problem, where an admissible controls set is convex. We establish necessary as well as sufficient optimality conditions for the risk-sensitive performance functional control problem. We illustrate the paper by giving two different examples for a linear quadratic system, and a numerical application as second example.

MSC 2010: 93E20; 60H30; 60G20

1 Introduction

In this paper, we aim, using Pontryagin’s maximum principle, to prove necessary as well as sufficient optimality conditions, for a risk-sensitive control problem associated with dynamics driven by a backward doubly stochastic differential equation (BDSDE in short). We solve the problem by using the approach developed by Djehiche, Tembine and Tempone in [5]; see also [2]. Our contribution can be summarized as follows: Djehiche et al. [5] have established a stochastic maximum principle for a class of risk-sensitive mean-field type control problems, where the distribution enters only through the mean of a state process, which means that the drift, diffusion, and terminal cost functions depend on the state, the control and the means of the state process. Their work extends the results of Lim and Zhou [10] to risk-sensitive control problems for dynamics that are non-Markovian and without mean-field term.

Nonlinear BDSDEs have been introduced by Pardoux and Peng [12]. They have considered a new kind of BSDE that is a class of BDSDEs with two different directions of stochastic integrals, i.e., the equations involve both a standard (forward) stochastic Itô integral dWt and a backward stochastic Itô integral dBt. More precisely, they dealt with the following BDSDE:

(1.1){-dYt=f(t,Yt,Zt)dt+g(t,Yt,Zt)dBt-ZtdWt,YT=ξ.

Here we regard (1.1) as a backward equation in two different aspects. The first is with respect to B for which the time variable is reversed. The second is with respect to W, which is forward in time while the boundary condition is given at the terminal instead of the initial time. They proved that if f and g are uniform Lipschitz, then (1.1) has a unique solution (Yt,Zt) in the interval [0,T], for any Zt being square integrable, and the terminal condition ξ is a TW-measurable and square integrable random variable. They also showed that BDSDEs can produce a probabilistic representation for solutions to some quasi-linear stochastic partial differential equations. Since this is the first existence and uniqueness result, many papers have been devoted to existence and/or uniqueness results under weaker assumptions; see [12] and the references therein. Further, the control processes are assumed to be adapted to a subfiltration of the filtration generated by both forward and backward filtration. In the risk-sensitive control problem, the system is governed by the nonlinear BDSDE

{dyt=-f(t,yt,zt,vt)dt-g(t,yt,zt,vt)dBt+ztdWt,yT=ξ,

where f, g are given maps, B=(Bt)t0 and W=(Wt)t0 are two mutually independent standard Brownian motion processes, zt is square integrable and the terminal condition ξ is a TW-measurable and square integrable random variable. The control variable v=(vt) is called admissible control. We denote by 𝒰 the class of all admissible controls.

We define the criterion to be minimized, with initial risk-sensitive functional cost, as follows:

Jθ(v)=𝔼[expθ{Ψ(y0)+0Tl(t,yt,zt,vt)𝑑t}],

where Ψ and l are given maps and (yt,zt) is the trajectory of the system controlled by vt.

A control u is called optimal if it solves Jθ(u)=infv𝒰Jθ(v).

The existence of an optimal solution for this problem has been solved to achieve the objective of this paper, and we have established necessary as well as sufficient optimality conditions for this model. In this paper, we generalize the results obtained by Chala in [2] to the BDSDE. The idea here is to reformulate in the first step the risk-sensitive control problem in terms of an augmented state process and terminal payoff problem. An intermediate stochastic maximum principle (SMP in short) is then obtained by applying the SMP of [1, 8] for a loss functional without running cost; for the same particular cases see [7]. Then we transform the intermediate adjoint processes to a simpler form, using the fact that the set of controls is convex. Then we establish necessary as well as sufficient optimality conditions by using the Logarithmic transform established by El Karoui and Hamadène [6]. More precisely, the method of Lim and Zhou [10] shows in fact that it is enough to use a generic square integrable martingale to transform the pair (p1,q1) into the adjoint process (p~1(t),0), where the process p~1(t) is still a square integrable martingale, which would mean that p~1(t)=p~1(T), and is equal to the constant 𝔼[p~1(T)]. But this generic martingale need not be related to the adjoint process p(t) as in Lim and Zhou [10]. Instead, it will be part of the adjoint equation associated with the risk-sensitive SMP (see Proposition 4.2 below).

We note that necessary as well as sufficient optimality conditions for risk-sensitive controls, where the systems are governed by a stochastic differential equation (SDE in short), have been studied by Lim and Zhou [10]. We also note that necessary as well as sufficient optimality conditions for stochastic controls, where the systems are governed by a nonlinear forward stochastic differential equation with jumps, have been studied by Shi and Wu in [14] in the case where the set of admissible controls is convex, and in [15] in the general case with an application to finance. Furthermore, the systems governed by a mean-field SDE have been studied by Djehiche et al. in [5]. For some general case, we can note the paper of Khallout and Chala [9] with the fully coupled forward backward stochastic differential equation, and where the admissible control set could be convex.

Our goal in this paper is to establish a set of necessary as well as sufficient optimality conditions for BDSDE with risk-sensitive performance cost. The backward doubly problem under consideration is not simple extension from the mathematical point of view, but also provides interesting models in many applications such as mathematical economics or optimal control. The proof of our main result is based on the spike variation method based on Theorem 3.2 as a preliminary step for the risk-neutral control problem.

The paper is organized as follows: In Section 2, we give the precise formulation of the problem; we introduce the risk-sensitive model, formulate the problem and give the various assumptions used throughout this paper. In Section 3, we shall study our system of BDSDE, and give the relation between the risk-neutral and risk-sensitive SMP. In Section 4, we give the new method of the transformation of the adjoint processes; the aim of this result will be shown in that section. We give our first main result, the necessary optimality conditions for risk-sensitive control problem under additional hypothesis, and the second result will be given in the following section. In Section 5, we derive our second main result of this paper, which are sufficient optimality conditions for risk-sensitive controls. In Section 6, we improve the quality of the paper by giving two applications to a linear quadratic stochastic control problem, the method which used the Riccati equations is applied in a second example, and we give a numerical example. Section 7 consists of a conclusion and an outlook.

2 Formulation of the problem

Let (Ω,,) be a probability space in which one-dimensional Brownian motions W=(Wt:0tT) and B=(Bt:0tT) are defined, where W and B are two mutually independent standard Brownian motion processes. Let 𝒩 denote the class of -null sets of . For each t[0,T], we define t(W,B)=tWt,TB, where for any process {Lt}t[0,T] one has s,tL=σ{Lr-Ls:srt}𝒩 and tL=0,tL.

Note that the collection {t(W,B):t[0,T]} is neither increasing nor decreasing, and it does not constitute a filtration. We may define the subfiltration (𝒢t)t[0,T] such that 𝒢tt(W,B) for all t[0,T].

Let 2([0,T];) denote the set of one-dimensional jointly measurable random processes {φt:t[0,T]} which satisfy the following conditions:

  1. φ2=𝔼[0T|φt|2𝑑t]<.

  2. φt is t(W,B)-measurable for any t[0,T].

Similarly, we denote by 𝒮2([0,T];) the set of one-dimensional continuous random processes which satisfy the following conditions:

  1. φ𝒮2=𝔼[supt[0,T]|φt|2]<.

  2. φt is t(W,B)-measurable for any t[0,T].

Let T be a positive real number and let 𝒜 be a nonempty subset of .

Definition 2.1.

An admissible control v is a measurable process with values in 𝒜 such that 𝔼0T|vt|2𝑑t<. We denote by 𝒰 the set of all admissible controls.

For any v𝒰, we consider the following BDSDE system:

(2.1){dyt=-f(t,yt,zt,vt)dt-g(t,yt,zt,vt)dBt+ztdWt,yT=ξ,

where f:[0,T]×××𝒜, g:[0,T]×××𝒜 are jointly measurable and such that for any (y,z,v)××𝒜 one has f(,y,z,v)2([0,T]×), g(,y,z,v)2([0,T]×), and zt is square integrable and the terminal condition ξ is a TW-measurable and square integrable random variable.

Note that the integral with respect to (Bt)t[0,T] is a “backward” Itô integral, while the integral with respect to (Wt)t[0,T] is a standard forward Itô integral. These two types of integrals are particular cases of the Itô–Skorohod integral; for more details we refer to [11].

We define the criterion to be minimized, with initial risk-sensitive performance functional cost, as

(2.2)Jθ(v)=𝔼[expθ{Ψ(y0)+0Tl(t,yt,zt,vt)𝑑t}],

where Ψ: and l:[0,T]×××𝒜 are jointly measurable and θ is the risk-sensitive index.

The optimal control problem is to minimize the functional Jθ over 𝒰 if u𝒰 is an optimal control (solution), that is,

(2.3)Jθ(u)=infv𝒰Jθ(v).

We assume the following.

Assumption 2.2.

There exist constants c>0 and 0λ<1 such that for any (w,t)Ω×[0,T] and

(y1,z1,v1),(y2,z2,v2)[0,T]×××𝒜

we have

|f(t,y1,z1,v1)-f(t,y2,z2,v2)|2c(|y1-y2|2+|z1-z2|2+|v1-v2|2),
|g(t,y1,z1,v1)-g(t,y2,z2,v2)|2c(|y1-y2|2+|v1-v2|2)+λ|z1-z2|2.

Theorem 2.3.

For any given admissible control v, we suppose that Assumption 2.2 holds. Then the BDSDE (2.1) has a unique solution (yt,zt)M2([0,T];R)×S2([0,T];R).

Proof.

See [12, Theorem 1.1, p. 212]. ∎

A control that solves problem (2.1)–(2.3) is called optimal. Our objective is to establish risk-sensitive necessary and sufficient optimality conditions, satisfied by a given optimal control, in the form of risk-sensitive SMP.

We also assume the following.

Assumption 2.4.

  1. f, g, l and Ψ are continuously differentiable with respect to (y,z,v).

  2. All derivatives of f, g and l are bounded by C(1+|y|+|z|+|v|).

  3. The derivative of Ψ is bounded by C(1+|y|).

Under the above Assumptions 2.2 and 2.4, for each v𝒰, equation (2.1) has a unique strong solution, and the cost function Jθ is well defined from 𝒰 into .

For more details the reader can see paper of Han, Peng and Wu in [8].

Remark 2.5.

We use the Euclidean norm || in , is a matrix transpose and Tr is the trace of a matrix. All the equalities and inequalities mentioned in this paper are in the sense of dt×d almost surely on [0,T]×Ω.

3 Risk-sensitive stochastic maximum principle of backward doubly type control

The proof of our risk-sensitive stochastic maximum principle necessitates a certain auxiliary state process xt, which is the solution of the following forward SDE:

dxt=l(t,yt,zt,vt)dt,x0=0.

Our control problem of (2.1)–(2.3) is equivalent to

(3.1){infv𝒰𝔼[expθ{Ψ(y0)+xT}]subject todxt=l(t,yt,zt,vt)dt,dyt=-f(t,yt,zt,vt)dt-g(t,yt,zt,vt)dBt+ztdWt,x0=0,yT=ξ.

We require the following condition:

ATθ:=expθ{Ψ(y0)+0Tl(t,yt,zt,vt)𝑑t}.

If we put ΘT=Ψ(y0)+0Tl(t,yt,zt,vt)𝑑t, then the risk-sensitive loss functional is given by

(θ,v):=1θlog[𝔼(expθ{Ψ(y0)+0lT(t,yt,zt,vt)𝑑t})]=1θlog[𝔼(expθΘT)].

When the risk-sensitive index θ is small, the loss functional (θ,u) can be expanded as

𝔼(ΘT)+θ2Var(ΘT)+O(θ2),

where Var(ΘT) denotes the variance of ΘT. If θ<0, the variance of ΘT, as a measure of risk, improves the performance (θ,v), in which case the optimizer is called risk seeker. But when θ>0, the variance of ΘT worsens the performance (θ,v), in which case the optimizer is called risk averse. The risk-neutral loss functional 𝔼(ΘT) can be seen as a limit of the risk-sensitive functional (θ,v) when θ0; for more details see [4].

Next, let us introduce the following notations.

Notation 3.1.

We set X:=(xy), Mt:=(WtBt), dp(t):=(dp1(t)dp2(t)),

F(t,yt,zt,vt):=(l(t,yt,zt,vt)-f(t,yt,zt,vt))andG(t,yt,zt,vt):=(00zt-g(t,yt,zt,vt)).

With these notations the system (3.1) can be rewritten in the following compact SDE form:

(3.2){dXt=F(t,yt,zt,vt)dt+G(t,yt,zt,vt)dMt,X(0T)=(0ξ).

For convenience, we will use the following notations throughout this paper: For ϕ{f,g,l} we define

{ϕ(t)=ϕ(t,yt,zt,vt),ϕ(t)=ϕ(t,yt,zt,vt)-ϕ(t,ytu,ztu,ut),ϕζ(t)=ϕζ(t,yt,zt,vt),ζ=y,z,v,
{H~θ(t)=H~θ(t,xt,yt,zt,vt,p(t),q(t)),H~θ(t)=H~θ(t,xt,yt,zt,vt,p(t),q(t))-H~θ(t,xtu,ytu,ztu,ut,p(t),q(t)),H~ζθ(t)=H~θζ(t,xt,yt,zt,vt,p(t),q(t)),ζ=y,z,v,

and

{Hθ(t)=Hθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t)),Hθ(t)=Hθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))-Hθ(t,ytu,ztu,ut,p~2(t),q~(t)2,Vθ(t),N(t)),Hζθ(t)=Hθζ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t)),ζ=y,z,v,

where vt is an admissible control from 𝒰.

We suppose that Assumptions 2.2 and 2.4 hold. We may combine the SMP for a risk-neutral controlled differential equation of backward doubly stochastic type from [1, 8] with the result of Yong [16] and with augmented state dynamics (x,y,z) to derive the adjoint equation. There exist two unique 𝒢t-adapted pairs of processes (p1,q1) and (p2,q2) which solve the following system of backward SDEs:

(3.3){dp(t)=-A(t)dt+R(t)dMt,(p1(T)p2(0))=θATθ(1-Ψy(y0u)),

with

𝔼[i=12supt[0,T]|pi(t)|2+i=120T|qi(t)|2𝑑t]<,

where

A(t)=(00ly(t)-fy(t))(p1(t)p2(t))+(000-gy(t))(q1(t)q2(t))

and

R(t)=([c]ccq1(t)0p3(t)-q2(t))

such that

p3(t)=-Tr[(lz(t)-fz(t)0-gz(t))(p1(t)q1(t)p2(t)q2(t))].

Let H~θ be the Hamiltonian associated with the optimal state dynamics (xu,yu,zu), and let the two pairs of adjoint process ((p1,q1),(p2,q2)) be given by

(3.4)H~θ(t):=H~θ(t,xtu,ytu,ztu,ut,p(t),q(t))=l(t)p1(t)-f(t)p2(t)-g(t)q2(t).

Theorem 3.2.

We suppose that Assumptions 2.2 and 2.4 hold. If (xu,yu,zu) is an optimal solution of the risk-neutral control problem (3.2), then there exist two pairs of Gt-adapted processes ((p1,q1),(p2,q2)) that satisfy (3.3) such that

(3.5)H~vθ(t,xtu,ytu,ztu,ut,p(t),q(t))(ut-vt)0

for all uA, almost every t[0,T] and P-almost surely, where H~θ(t) is defined in Notation 3.1.

4 Finding the new adjoint equations and necessary optimality conditions

As mentioned, Theorem (3.2) is a good SMP for the risk-neutral control problem of forward-backward doubly type. We follow the same approach used in [2, 9], and suggest a transformation of the adjoint processes (p1,q1) and (p2,q2) in such a way that we can omit the first component (p1,q1) in (3.3) and express the SMP in terms of only one adjoint process which we denote by (p~2,q~2).

By noting that dp1(t)=q1(t)dWt and p1(T)=θATθ, the explicit solution of this backward SDE is

(4.1)p1(t)=θ𝔼[ATθ𝒢t]=θVtθ,

where Vtθ:=𝔼[ATθ𝒢t], 0tT.

In view of (4.1), it would be natural to choose a transformation of (p,q) into an adjoint process (p~,q~), where

p~1(t)=1θVtθp1(t)=1.

We consider the following transform:

(4.2)p~(t)=(p~1(t)p~2(t)):=1θVtθp(t),0tT.

By using (3.3) and (4.2), we have

p~():=(p~1(T)p~2(0))=(1-Ψy(y0u)).

The following properties of the generic martingale Vθ are essential in order to investigate the properties of these new process (p~(t),q~(t)): The process Λθ is the first component of the 𝒢t-adapted pair of processes (Λθ,N) which is the unique solution to the following quadratic backward SDE:

(4.3){dΛtθ=-(l(t)+θ2|N(t)|2)dt+N(t)dMt,ΛTθ=Ψ(y0u),

where 𝔼[0T|N(t)|2𝑑t]<.

Next, we will state and prove the necessary optimality conditions for the system driven by a BDSDE with a risk-sensitive performance functional type. To this end, let us summarize and prove some lemmas that we will use thereafter.

Lemma 4.1.

Suppose that Assumption 2.4 holds. Then

(4.4)𝔼[supt[0,T]|Λtθ|]CT.

In particular, Vθ solves the following linear backward SDE:

(4.5)dVtθ=θN(t)VtθdMt,VTθ=ATθ.

Hence, the process defined on (Ω,F,(Gt)t0,P) by Ltθ, where

(4.6)Ltθ:=VtθV0θ=exp(θ0tN(s)𝑑Ms-θ220t|N(s)|2𝑑s),0tT,

is a uniformly bounded Gt-martingale.

Proof.

First, we prove (4.4). By Assumption 2.4, l and Ψ are bounded by a constant C>0. We have

(4.7)0<e-(1+T)CθATθe(1+T)Cθ.

Therefore, Vθ is a uniformly bounded 𝒢t-martingale satisfying

(4.8)0<e-(1+T)CθVtθe(1+T)Cθ,0tT.

The sufficient conditions of the logarithmic transform established in [6, Proposition 3.1] can be applied in the martingale Vθ as follows:

Vtθ=exp(θΛtθ+θ0tl(s)𝑑s),0tT,

and V0θ=exp(θΛ0θ)=𝔼[ATθ].

It is very easy to see from (4.8) and the boundedness of l that

𝔼[supt[0,T]|Λtθ|]CT,

where CT is a positive constant that depends only on T and the boundedness of l and Ψ.

Second, we find the explicit form of (4.5). Using the second Itô’s formula on

Vtθ=exp(θΛtθ+θ0tl(s)𝑑s),

we get

dVtθ=θN(t)VtθdMt.

Now, we can prove (4.6) by starting from the integral form of (4.5) such that

dVtθ=θN(t)VtθdMt,VTθ=ATθ.

On the other hand, we have

Vtθ=exp(θΛtθ+θ0tl(s)𝑑s).

Using expression (4.3), we can write

Vtθ=exp(θ0tN(s)𝑑Ms-θ220t|N(s)|2𝑑s+θΛ0θ).

Then

Ltθ:=VtθV0θ=exp(θ0tN(s)𝑑Ms-θ220t|N(s)|2𝑑s),0tT.

In view of(4.4), the above equality is a uniformly bounded 𝒢t-martingale.

For more details of this proof the reader can visit the papers of Chala [2, 3]. ∎

Proposition 4.2.

The main risk-sensitive of second adjoint equation for (p~2,q~2) and (Vθ,N) becomes

(4.9){dp~2(t)=-Hyθ(t)dt-Hzθ(t)dWtθ-{q~2(t)+θN2(t)p~2(t)}dBtθ,dVtθ=θN(t)VtθdMt,p~2(T)=-Ψy(y0),VTθ=ATθ.

The solution (p~,q~,Vθ,N) of system (4.9) is unique such that

(4.10)𝔼[supt[0,T]|p~(t)|2+supt[0,T]|Vθ(t)|2+0T(|q~(t)|2+|N(t)|2)𝑑t]<,

where

(4.11)Hθ(t):=Hθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))=l(t)-(f(t)-θN1(t)zt)p~2(t)-g(t)q~2(t).

Proof.

We wish to identify the processes α~ and β~ such that

(4.12)dp~(t)=-α~(t)dt+β~(t)dMt,

where

β~(t):=(β~1(t)β~2(t)):=(β~11(t)β~12(t)β~21(t)β~22(t)).

By applying Itô’s formula to the process p(t)=θVtθp~(t) and using the expression of Vθ in (4.5), we obtain

dp~(t)=-1θVtθ[(00ly(t)-fy(t))(p1(t)p2(t))+(000-gy(t))(q1(t)q2(t))]dt
-θN(t)β~(t)dt+1θVtθ(q1(t)0p3(t)-q2(t))dMt-θp~(t)N(t)dMt.

By combining the above equation with (4.12) and using the relation

p~(t)=1θVtθp(t),

the drift coefficient α~(t) can be written as

α~(t)=(00ly(t)-fy(t))(p~1(t)p~2(t))+(000-gy(t))(q~1(t)q~2(t))+θN(t)β~(t),

and the diffusion term of the process p~(t) as

(4.13)β~(t)=(q~1(t)0p~3(t)-q~2(t))-θp~(t)N(t).

Finally, we obtain

dp~(t)=-[(00ly(t)-fy(t))(p~1(t)p~2(t))+(000-gy(t))(q~1(t)q~2(t))]dt+β~(t)[dMt-θN(t)dt].

It is easily verified that

dp~1(t)=β~1(t)[dMt-θN(t)dt],p~1(T)=1.

In view of (4.6), we may use Girsanov’s theorem [4] to claim that

dp~1(t)=β~1(t)dMtθ,θ-a.s.,p~1(T)=1,

where dMtθ=dMt-θN(t)dt. By using Notation 3.1, dMtθ can be written as

dMtθ=(dWtdBt)-θ(N1(t)N2(t))dt,

which is a θ-Brownian motion, where

dθd|t(W,B):=Ltθ=exp(θ0tN(s)dMs-θ220t|N(s)|2ds),0tT.

But according to (4.6) and (4.7), the probability measures θ and are in fact equivalent. Hence, noting that

p~1(t):=1θVtθp1(t)

is square integrable, we get that

p~1(t)=𝔼θ[p~1(T)𝒢t]=1.

Thus, its quadratic variation is

0T|β~1(t)|2𝑑t=0.

This implies that, for almost every 0tT, β~1(t)=0, θ-a.s. and -a.s. we have

(4.14)dp~(t)=-[(00ly(t)-fy(t))(p~1(t)p~2(t))+(000-gy(t))(q~1(t)q~2(t))]dt+β~(t)dMtθ.

Now replacing (4.13) in (4.14), we obtain

dp~(t)=-[(00ly(t)-fy(t))(p~1(t)p~2(t))+(000-gy(t))(q~1(t)q~2(t))]dt+(β~11(t)β~12(t)β~21(t)β~22(t))(dWtθdBtθ),

where

β~11(t)=q~1(t)-θN1(t)p~1(t),β~12(t)=-θN2(t)p~1(t),β~21(t)=p~3(t)-θN1(t)p~2(t),
β~22(t)=-q~2(t)-θN2(t)p~2(t),p~3(t)=-Tr[(lz(t)-fz(t)0-gz(t))(p~1(t)q~1(t)p~2(t)q~2(t))].

Therefore, the second component p~2(t) given in (4.14) has the form

dp~2(t)=-{ly(t)p~1(t)-fy(t)p~2(t)-gy(t)q~2(t)}dt
-{lz(t)p~1(t)-(fz(t)-θN1(t))p~2(t)-gz(t)q~2(t)}dWtθ-{q~2(t)+θN2(t)p~2(t)}dBtθ,

and the main risk-sensitive second adjoint equation for (p~2,q~2) and (Vθ,N) becomes

{dp~2(t)=-Hyθ(t)dt-Hzθ(t)dWtθ-{q~2(t)+θN2(t)p~2(t)}dBtθ,dVtθ=θN(t)VtθdMt,p~2(T)=-Ψy(y0),VTθ=ATθ.

The solution (p~,q~,Vθ,N) of the system (4.9) is unique such that

𝔼[supt[0,T]|p~(t)|2+supt[0,T]|Vθ(t)|2+0T(|q~(t)|2+|N(t)|2)𝑑t]<,

where

Hθ(t):=Hθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))=l(t)-(f(t)-θN1(t)zt)p~2(t)-g(t)q~2(t).

The proof of this proposition is completed. ∎

Now, we can state the necessary optimality conditions.

Theorem 4.3 (Necessary optimality conditions for risk-sensitive).

Suppose that Assumption 2.2 and Assumption 2.4 hold. If (yu,zu,u) is an optimal solution of the risk-sensitive control problem (2.1)–(2.3). Then there exist two pairs of Gt-adapted processes (Vθ,N), (p~,q~) which satisfy (4.9) and (4.10) such that

(4.15)Hvθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))(ut-vt)0

for all uU, almost every 0tT and P-almost surely.

Proof.

We arrive at a risk-sensitive stochastic maximum principle expressed in terms of the adjoint processes (p~2,q~2) and (Vθ,N) which solve (4.9), where the Hamiltonian H~θ associated with (3.1), given by (3.4) satisfies

(4.16)H~θ(t,xtu,ytu,ztu,ut,p(t),q(t))={θVtθ}Hθ(t,ytu,ztu,ut,p~2(t),q~2(t),Vθ(t),N(t)),

and Hθ is the risk-sensitive Hamiltonian given by (4.11). Hence, since Vθ>0, the variational inequality (3.5) translates into

Hvθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))(ut-vt)0

for all u𝒰, almost every 0tT and -almost surely. This finishes the proof of Theorem 4.3. ∎

5 Sufficient optimality conditions for risk-sensitive performance cost

In this section, we study when the necessary optimality conditions (3.5) become sufficient.

Theorem 5.1 (Sufficient optimality conditions for risk-sensitive).

Assume that the functions Ψ and

(xt,yt,zt,vt)H~θ(t,xt,yt,zt,vt,p(t),q(t))

are convex and that for any vtU, the variable yT=ξ is a one-dimensional FTW-measurable random variable such that E|ξ|2<. Then u is an optimal solution of the control problem (2.1)–(2.3) if it satisfies (3.5).

Proof.

Let u be an arbitrary element of 𝒰 (candidate to be optimal). For any v𝒰, we have

Jθ(v)-Jθ(u)=𝔼[exp(θ{Ψ(y0)+xT})]-𝔼[exp(θ{Ψ(y0u)+xTu})].

By applying the Taylor’s expansion and the convexity of Ψ, we get

Jθ(v)-Jθ(u)𝔼[θexp(θ{Ψ(y0u)+xTu})(xT-xTu)]+𝔼[θexp(θ{Ψ(y0u)+xTu})Ψy(y0u)(y0-y0u)].

It follows from (3.3), that p1(T)=θATθ and p2(0)=-θATθΨy(y0u). Then we have

(5.1)Jθ(v)-Jθ(u)𝔼[p1(T)(xT-xTu)]-𝔼[p2(0)(y0-y0u)].

Applying Itô’s formula to p1(t)(xt-xtu) and p2(t)(yt-ytu) leads to

𝔼[p1(T)(xT-xTu)]=𝔼[0T(l(t,yt,zt,vt)-l(t,ytu,ztu,ut))p1(t)𝑑t]

and

-𝔼[p2(0)(y0-y0u)]=-𝔼[0TH~yθ(t,xtu,ytu,ztu,ut,p(t),q(t))(yt-ytu)𝑑t]
-𝔼[0TH~zθ(t,xtu,ytu,ztu,ut,p(t),q(t))(zt-ztu)𝑑t]
-𝔼[0T(f(t,yt,zt,vt)-f(t,ytu,ztu,ut))p2(t)𝑑t]
-𝔼[0T(g(t,yt,zt,vt)-g(t,ytu,ztu,ut))q2(t)𝑑t].

Putting the two above formulas into (5.1), we get

Jθ(v)-Jθ(u)𝔼[0TH~θ(t,xt,yt,zt,vt,p(t),q(t))𝑑t]-𝔼[0TH~θ(t,xtu,ytu,ztu,ut,p(t),q(t))𝑑t]
-𝔼[0TH~yθ(t,xtu,ytu,ztu,ut,p(t),q(t))(yt-ytu)𝑑t]
(5.2)-𝔼[0TH~zθ(t,xtu,ytu,ztu,ut,p(t),q(t))(zt-ztu)𝑑t].

Since the Hamiltonian H~θ is convex with respect to (y,z,v), we have

𝔼[0TH~θ(t,xt,yt,zt,vt,p(t),q(t))𝑑t]-𝔼[0TH~θ(t,xtu,ytu,ztu,ut,p(t),q(t))𝑑t]
𝔼[0TH~yθ(t,xtu,ytu,ztu,ut,p(t),q(t))(yt-ytu)𝑑t]
+𝔼[0TH~zθ(t,xtu,ytu,ztu,ut,p(t),q(t))(zt-ztu)𝑑t]
+𝔼[0TH~vθ(t,xtu,ytu,ztu,ut,p(t),q(t))(vt-ut)𝑑t].

Then, by using the above inequality in (5.2), we obtain

Jθ(v)-Jθ(u)𝔼[0TH~vθ(t,xtu,ytu,ztu,ut,p(t),q(t))(vt-ut)𝑑t]0.

In virtue of the necessary optimality conditions (3.5), the last inequality implies that Jθ(v)-Jθ(u)0. Thus the theorem is proved. ∎

Remark 5.2.

In virtue of (4.16) there is a relationship between the Hamiltonian with respect to risk-neutral and the Hamiltonian with respect to risk-sensitive. In fact, we have

Jθ(v)-Jθ(u)𝔼[0TθVtθHvθ(t,ytu,ztu,ut,p~2(t),q~2(t),Vθ(t),N(t))(vt-ut)𝑑t]0.

We know that θVtθ>0. Then the above inequality can be rewritten as

Jθ(v)-Jθ(u)𝔼[0THvθ(t,ytu,ztu,ut,p~2(t),q~2(t),Vθ(t),N(t))(vt-ut)𝑑t]0.

In virtue of the necessary optimality conditions (4.15), the last inequality implies that Jθ(v)-Jθ(u)0.

6 Applications: A linear quadratic risk-sensitive control problem

We illustrate the paper by giving two different examples for a linear quadratic system, and a numerical application as second example.

6.1 Example 1

We provide a concrete example of a risk-sensitive backward doubly stochastic LQ problem, give the explicit optimal control and validate our major theoretical results in Theorem 5.1 (risk-sensitive sufficient optimality conditions). First, let the control domain be 𝒜=[-1,1]. Consider the following linear quadratic risk-sensitive control problem

(6.1){infv𝒰𝔼[expθ{120Tvt2𝑑t+12(y0)2}]subject todyt=-(Ayt+Bzt+Cvt+D)dt-(Ayt+Bzt+Cvt+D)dBt+ztdWt,yT=ξ,

where A, B, C, D, A, B, C and D are positive real constants.

Let (yt,zt) be a solution of (6.1) associated with vt. Then there exist two unique 𝒢t-adapted pairs of processes (p1,p2),(q1,q2) of the following forward-backward doubly SDE system (called adjoint equation), according to equation (3.3):

(6.2){dp1(t)=q1(t)dWt,dp2(t)=[Ap2(t)+Aq2(t)]dt+[Bp2(t)+Bq2(t)]dWt-q2(t)dBt,p1(T)=θATθ,p2(0)=-θy0ATθ,

where

ATθ:=expθ{120Tvt2𝑑t+(y0)2}.

We give the Hamiltonian H~θ defined by

H~θ(t):=H~θ(t,xt,yt,zt,vt,p(t),q(t))
=12vt2p1(t)-(Ayt+Bzt+Cvt+D)p2(t)-(Ayt+Bzt+Cvt+D)q2(t).

We have H~vθ(t)=[vtp1(t)-Cp2(t)-Cq2(t)]. Maximizing the Hamiltonian yields

(6.3)ut=1p1u(t)(Cp2u(t)+Cq2u(t)).

We only need to prove that ut is an optimal control of (6.1).

Theorem 6.1 (Risk-sensitive sufficient optimality conditions for a linear quadratic control problem).

Assume that θ>0 and suppose that ut satisfies (6.3), where (p,q) satisfy (6.2). Then ut is the unique optimal control of the above BDSDE of the linear quadratic problem (6.1).

Proof.

From the definition of the functional cost Jθ we have

Jθ(vt)-Jθ(ut)=𝔼[expθ{120Tvt2𝑑t+12(y0)2}]-𝔼[expθ{120Tut2𝑑t+12(y0u)2}].

We put xT=120Tvt2𝑑t and, by applying the Taylor expansion, we have

Jθ(vt)-Jθ(ut)=𝔼[θexpθ{xTu+12(y0u)2}(xT-xTu)]+𝔼[θy0uexpθ{xTu+12(y0u)2}(y0-y0u)].

It follows from (3.3) that p1u(T)=θATθ and p2u(0)=-θy0uATθ. Then we have

(6.4)Jθ(vt)-Jθ(ut)=𝔼[p1u(T)(xT-xTu)]-𝔼[p2u(0)(y0-y0u)].

By applying Itô’s formula to p1u(t)(xt-xtu) and p2u(t)(yt-ytu) and combining them with (6.4), we get

Jθ(vt)-Jθ(ut)=𝔼[0Tvt(vt-ut)p1u(t)𝑑t]+𝔼[0Tut(vt-ut)p1u(t)𝑑t]
-𝔼[0TC(vt-ut)p2(t)udt]-𝔼[0TC(vt-ut)q2u(t)dt].

Because of θ>0, we have (vt-ut)>0. Thus we get the following result:

Jθ(vt)-Jθ(ut)𝔼[0Tut(vt-ut)p1u(t)𝑑t]-𝔼[0TC(vt-ut)p2u(t)𝑑t]-𝔼[0TC(vt-ut)q2u(t)𝑑t].

Then

Jθ(vt)-Jθ(ut)𝔼[0T(utp1u(t)-Cp2u(t)-Cq2u(t))(vt-ut)𝑑t].

By replacing ut with its value in (6.3), we obtain

Jθ(vt)Jθ(ut),

i.e. ut is optimal. This proof is finished ∎

6.2 Example 2

In this section, we apply the risk-sensitive maximum principles obtained in the previous section (Theorem 4.3) to deal with the linear-quadratic risk-sensitive stochastic optimal control problem (2.1)–(2.3) mentioned in Section 2. Our state dynamics is

(6.5){dyt=-(yt+vt)dt-σvtdBt+ztdWt,yT=ξ,

and our functional cost is the following expected exponential-of-integral form:

(6.6)Jθ(v)=𝔼[expθ{0Tl(t,yt,zt,vt)𝑑t}],

where

θ>0,θ1,l(t,yt,zt,vt)=12(vt2+yt2).

We want to minimize (6.6) subject to (6.5) by choosing v over 𝒰. Hence, we may apply Theorem 4.3 to solve our linear-quadratic risk-sensitive stochastic optimal control problem (6.5), (6.6). The Hamiltonian function (4.11) is defined by

Hθ(t):=Hθ(t,yt,zt,vt,p~2(t),q~2(t),Vθ(t),N(t))
(6.7)=12(vt2+yt2)-(yt+vt-θN1(t)zt)p~2(t)-σvtq~2(t).

Let (ytu,ztu,ut) be an optimal solution. The adjoint equation (4.9) can be written by

(6.8){dp~2(t)=[-yt+p~2(t)]dt-θN1(t)p~2(t)dWtθ-[q~2(t)+θN2(t)p~2(t)]dBtθ,p~2(T)=0.

Minimizing the Hamiltonian (6.7), we obtain

(6.9)ut=p~2u(t)+σq~2u(t).

By substituting (6.9) into the BDSDE (6.5), we have

{dytu=-[ytu+p~2u(t)+σq~2u(t)]dt-[σ(p~2u(t)+σq~2u(t))]dBt+ztudWt,yTu=ξ.

Similarly, by substituting equation (6.9) into the BDSDE (6.8) and replacing dWtθ=dWt-θN1(t)dt and dBtθ=dBt-θN2(t)dt, we get

(6.10){dp~2u(t)=[-ytu+(1+θ2(N12(t)+N22(t)))p~2u(t)+θN2(t)q~2u(t)]dt,-[q~2u(t)+θN2(t)p~2u(t)]dBt-θN1(t)p~2u(t)dWt,p~2u(T)=0.

Peng and Shi [13] introduced a type of time-symmetric forward-backward stochastic differential equations (SFBSDE in short), i.e., so-called fully coupled forward-backward doubly stochastic differential equations (FBDSDE in short). Therefore, an optimal solution (p~2u,yu,u) can be obtained by solving the following type of SFBSDE:

(6.11){dytu=-[ytu+p~2u(t)+σq~2u(t)]dt-[σ(p~2u(t)+σq~2u(t))]dBt+ztudWt,dp~2u(t)=[-ytu+(1+θ2(N12(t)+N22(t)))p~2u(t)+θN2(t)q~2u(t)]dt-[q~2u(t)+θN2(t)p~2u(t)]dBt-θN1(t)p~2u(t)dWt,yTu=ξ,p~2u(T)=0.

Unfortunately, in such a system it is difficult to find the explicit solution. To solve this type of SFBSDE (6.11), we use a technique similar to the one used by Yong and Zhou [17]. We conjecture that the solution to (6.11) is related by

(6.12)p~2u(t)=φ(t)ytu+χ(t)

for some deterministic differentiable functions φ(t) and χ(t). Applying Itô’s formula to (6.12) gives

(6.13){dp~2u(t)=[φ(t)ytu-φ(t)ytu-φ(t)(p~2u(t)+σq~2u(t))+χ(t)]dt-φ(t)σ(p~2u(t)+σq~2u(t))dBt+φ(t)ztudWt,p~2u(T)=0.

Putting (6.12) into (6.13), we get

(6.14){dp~2u(t)=[(φ(t)-φ2(t)-φ(t))ytu-φ(t)χ(t)-φ(t)σq~2u(t)+χ(t)]dt-[φ2(t)σytu+φ(t)χ(t)σ+φ(t)σq~2u2(t)]dBt+φ(t)ztudWt,p~2u(T)=0.

On the other hand, after substituting (6.12) into (6.10), we arrive at

(6.15){dp~2u(t)=[((1+θ2(N12(t)+N22(t)))φ(t)-1)ytu+(1+θ2(N12(t)+N22(t)))χ(t)+θN2(t)q~2u(t)]dt-[q~2u(t)+θN2(t)φ(t)ytu+θN2(t)χ(t)]dBt-[θN1(t)φ(t)ytu+θN1(t)χ(t)]dWt,p~2u(T)=0.

Equating the coefficients of (6.14) and (6.15), we have

(6.16)(p~2u(t),q~2u(t))=(φ(t)ytu+χ(t),φ(t)(σφ(t)-θN2(t))ytu+(φ(t)σ-θN2(t))χ(t)1-φ(t)σ2),

where φ(t) is the solution to the Riccati-type equation

(6.17){φ(t)-φ2(t)-2φ(t)(1+12θ2(N12(t)+N22(t)))+1=0,φ(T)=0,

and χ(t) is a solution to the following ordinary differential equation:

(6.18){χ(t)-(φ(t)+1+θ2(N12(t)+N22(t)))χ(t)-(θN2(t)+φ(t)σ)q~2(t)=0,χ(T)=0.

By using the same identification, we get

(6.19)χ(t)=-1θN1(t)φ(t)ztu-φ(t)ytu.

Finally, by (6.9) and (6.16), we can get the optimal control in the following state in feedback form:

(6.20)ut=(1-σθN2(t))1-φ(t)σ2φ(t)ytu+(1-σθN2(t))1-φ(t)σ2χ(t).

Putting (6.19) in (6.20), we get

(6.21)ut=-(1-σθN2(t))(1-φ(t)σ2)θN1(t)φ(t)ztu,

where φ(t) is determined by (6.17).

Theorem 6.2.

We assume that the pair (φ(t),χ(t)) has the solution of system (6.17) and (6.18). Then the optimal control of our linear-quadratic risk-sensitive stochastic optimal control problem (6.5), (6.6) has the state feedback form (6.21).

6.2.1 Solution of the deterministic functions via Riccati equation

To the best of our knowledge, it is very hard to find the explicit solution to a Riccati equation in general. But in our case, we can find the explicit solution of

(6.22){φ(t)-φ2(t)+2φ(t)K+1=0,φ(T)=0,

where we set

(6.23)K=-(1+12θ2(N12(t)+N22(t))).

It is very easy to see that for the discriminate Δ=4(K2+1)>0 holds. We obtain

dt=dφ(t)φ2(t)-2φ(t)K-1.

Then

T-t=12K2+1tT{1(φ(s)-(K+K2+1))-1(φ(s)-(K-K2+1))}𝑑φ(s).

This implies

(T-t)2K2+1=ln|K+K2+1|-ln|K-K2+1|
+ln|φ(t)-(K-K2+1)|-ln|φ(t)-(K+K2+1)|.

Hence,

φ(t)-(K+K2+1)φ(t)-(K-K2+1)=±K+K2+1K-K2+1exp(-(T-t)2K2+1).

This concludes

φ(t)=(K+K2+1)(1+exp(-2(T-t)K2+1))(1+(K+K2+1K-K2+1)exp(-2(T-t)K2+1)).

Substituting the equation 4(K2+1)=Δ, we then have

φ(t)=(K+Δ2)(1+exp(-(T-t)Δ))(1+(K+Δ2K-Δ2)exp(-(T-t)Δ)).

In fact, the Riccati equation has another solution:

φ(t)=(K+Δ2)(1-exp(-(T-t)Δ))(1-(K+Δ2K-Δ2)exp(-(T-t)Δ)).

We must reject this solution because of the choose problem. If we denote

L=K+Δ2K-Δ2,δ1=K+Δ2,δ2=K-Δ2,

then we get

(6.24)φ(t)=δ1+δ2Lexp(-(T-t)Δ)1+Lexp(-(T-t)Δ).

We put

(6.25)α(t)=-(φ(t)+1+θ2(N12(t)+N22(t))),β(t)=-(θN2(t)+φ(t)σ)q~2(t).

We rewrite equation (6.18) as follows:

(6.26){χ(t)+α(t)χ(t)+β(t)=0,χ(T)=0,

The explicit solution to equation (6.26) is

(6.27)χ(t)=[exp(tTα(s)𝑑s)][tT-β(s)exp(tTα(r)𝑑r)ds],

where α(t), β(t) are determined by (6.25).

Corollary 6.3.

The explicit solution of the Riccati equation (6.22) is given by (6.24) and equation (6.26) has an explicit solution given by (6.27), where the constant coefficients K and α(t), β(t) are given by (6.23) and (6.25) respectively.

Corollary 6.4.

We assume that the pair (φ(t),χ(t)) has the unique solution given by (6.24), (6.27). Then the optimal control of the problem (6.5), (6.6) has the state feedback from (6.21), where the constant coefficient K is given by (6.23), and α(t), β(t) are determined by (6.25).

7 Conclusion and further works

This paper contains two main results. The first result is Theorem 4.3, which establishes the necessary optimality condition for the system of BDSDE with risk-sensitive performance, using a scheme very similar to the one by Chala [2]. The second main result, Theorem 5.1, suggests sufficient optimality conditions of BDSDE given in form of risk-sensitive performance. We acknowledge that this result is a good extension of the result established by Chala in [3]. The proof is based on the convexity conditions of the Hamiltonian function and the initial and terminal terms of the performance function. Note that the risk-sensitive control problem studied by Lim and Zhou in [10] is different to ours. By adding the jump diffusion term to our system (this result will be discussed in our next paper), in this case we can compare this result with Shi and Wu in [15]. On the other hand, in the case where the system is governed by mean-field fully coupled FBSDE (this result will be discussed in detail in our next paper), this result will hopefully generalize the result of Djehiche et al. in [5]. The maximum principle of risk-neutral obtained by [1, 8] and Yong [16] are similar to our Theorem 3.2, but the adjoint equations and maximum conditions heavily depend on the risk-sensitive parameter.


Communicated by Vyacheslav L. Girko


Funding statement: This work is partially supported by PRFU project N: C00L03UN070120180002.

Acknowledgements

The authors wish to thank the referees and editors for their valuable comments and suggestions which led to improvements in the document.

References

[1] S. Bahlali and B. Gherbal, Optimality conditions of controlled backward doubly stochastic differential equations, Random Oper. Stoch. Equ. 18 (2010), no. 3, 247–265. 10.1515/rose.2010.014Suche in Google Scholar

[2] A. Chala, Pontryagin’s risk-sensitive stochastic maximum principle for backward stochastic differential equations with application, Bull. Braz. Math. Soc. (N. S.) 48 (2017), no. 3, 399–411. 10.1007/s00574-017-0031-2Suche in Google Scholar

[3] A. Chala, Sufficient optimality condition for a risk-sensitive control problem for backward stochastic differential equations and an application, J. Numer. Math. Stoch. 9 (2017), no. 1, 48–60. 10.1007/s00574-017-0031-2Suche in Google Scholar

[4] A. Chala, D. Hafayed and R. Khallout, The use of Girsanov’s theorem to describe the risk-sensitive problem and application to optimal control, Stochastic Differantiel Equation-basics and Applications, Nova Science Publishers, New York (2018), 111–142. Suche in Google Scholar

[5] B. Djehiche, H. Tembine and R. Tempone, A stochastic maximum principle for risk-sensitive mean-field type control, IEEE Trans. Automat. Control 60 (2015), no. 10, 2640–2649. 10.1109/CDC.2014.7039929Suche in Google Scholar

[6] N. El-Karoui and S. Hamadène, BSDEs and risk-sensitive control, zero-sum and nonzero-sum game problems of stochastic functional differential equations, Stochastic Process. Appl. 107 (2003), no. 1, 145–169. 10.1016/S0304-4149(03)00059-0Suche in Google Scholar

[7] D. Hafayed and A. Chala, A general maximum principle for mean-field forward-backward doubly stochastic differential equations with jumps processes, Random Oper. Stoch. Equ. 27 (2019), no. 1, 9–25. 10.1515/rose-2019-2002Suche in Google Scholar

[8] Y. Han, S. Peng and Z. Wu, Maximum principle for backward doubly stochastic control systems with applications, SIAM J. Control Optim. 48 (2010), no. 7, 4224–4241. 10.1137/080743561Suche in Google Scholar

[9] R. Khallout and A. Chala, A risk-sensitive stochastic maximum principle for fully coupled forward-backward stochastic differential equations with applications, Asian J. Control (2019), 10.1002/asjc.2020. 10.1002/asjc.2020Suche in Google Scholar

[10] A. E. B. Lim and X. Y. Zhou, A new risk-sensitive maximum principle, IEEE Trans. Automat. Control 50 (2005), no. 7, 958–966. 10.1109/TAC.2005.851441Suche in Google Scholar

[11] D. Nualart and E. Pardoux, Stochastic calculus with anticipating integrands, Probab. Theory Related Fields 78 (1988), no. 4, 535–581. 10.1007/BF00353876Suche in Google Scholar

[12] E. Pardoux and S. G. Peng, Backward doubly stochastic differential equations and systems of quasilinear SPDEs, Probab. Theory Related Fields 98 (1994), no. 2, 209–227. 10.1007/BF01192514Suche in Google Scholar

[13] S. Peng and Y. Shi, A type of time-symmetric forward-backward stochastic differential equations, C. R. Math. Acad. Sci. Paris 336 (2003), no. 9, 773–778. 10.1016/S1631-073X(03)00183-3Suche in Google Scholar

[14] J. Shi and Z. Wu, A risk-sensitive stochastic maximum principle for optimal control of jump diffusions and its applications, Acta Math. Sci. Ser. B (Engl. Ed.) 31 (2011), no. 2, 419–433. 10.1016/S0252-9602(11)60242-7Suche in Google Scholar

[15] J. Shi and Z. Wu, Maximum principle for risk-sensitive stochastic optimal control problem and applications to finance, Stoch. Anal. Appl. 30 (2012), no. 6, 997–1018. 10.1080/07362994.2012.727138Suche in Google Scholar

[16] J. Yong, Optimality variational principle for controlled forward-backward stochastic differential equations with mixed initial-terminal conditions, SIAM J. Control Optim. 48 (2010), no. 6, 4119–4156. 10.1137/090763287Suche in Google Scholar

[17] J. Yong and X. Y. Zhou, Stochastic Controls. Hamiltonian Systems and HJB Equations, Appl. Math. (New York) 43, Springer, New York, 1999. Suche in Google Scholar

Received: 2018-05-30
Accepted: 2019-10-25
Published Online: 2020-01-17
Published in Print: 2020-03-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 28.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/rose-2020-2024/html
Button zum nach oben scrollen