1,375
Views
1
CrossRef citations to date
0
Altmetric
Review

Mean-field FBSDE and optimal control

&
Pages 235-251 | Received 21 May 2019, Accepted 08 Jul 2020, Published online: 29 Jul 2020

Abstract

We study optimal control for mean-field forward–backward stochastic differential equations with payoff functionals of mean-field type. Sufficient and necessary optimality conditions in terms of a stochastic maximum principle are derived. As an illustration, we solve an optimal portfolio with mean-field risk minimization problem.

2010 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

After the seminal work by Lasry and Lions [Citation1], where they introduced mean-field game theory that is devoted to the analysis of differential games with infinitely many players. Mean-field games attracted a lot of attention and forward/backward stochastic differential equations of mean-field type are used, extensively, as dynamics (see e.g., Huang et al. [Citation2], Xu and Zhang [Citation3] and Xu and Shi [Citation4]). In Huang [Citation5], the author studies a linear–quadratic game with a major player and a large number of minor players. The dynamics of the major player is influenced by an aggregation of all minor players (mean-field coupling) whereas the minor players’ dynamics depend on the control of the major player in addition to their individual controls as well as the mean-field coupling, i.e., a system of partially control-coupled forward stochastic differential equations (SDEs). This work (Ref. [Citation5]) was generalized to the non-linear case in Nourian and Caines [Citation6]. In all previously mentioned works, the authors find ϵ-Nash equilibrium for mean-field games, where each player play a game with the aggregation of the other players (the mass). In the present paper, the setting is different. We consider a mean-field type control problem where the goal is to find an optimal control via stochastic maximum principle. The mass or the laws of state processes are not freezed, they vary with the change of the control. Thus, finding an optimal control will yield optimal laws. Furthermore, in our control problem we consider a controlled partially coupled forward–backward SDE of mean-field type (MF-FBSDE) as dynamics, which is a novel contribution. We have also used the Sobolov space of random measures, introduced in Agram et al. [Citation7–9], in which, the Fréchet derivative with respect to the measure can be taken directly. This is a new approach compared to what is standard in the literature, where the Wasserstein metric space for measures and the lifting technique, introduced by Lions [Citation10], is used to differentiate a function of a measure.

Existence of a fully coupled MF-FBSDE is studied by Carmona and Delarue [Citation11] under Lipschitz assumption on the coefficients but no uniqueness result was proven. Bensoussan et al. [Citation12] prove existence and uniqueness of a fully coupled MF-FBSDE by assuming Lipschitz and monotonicity conditions. Recently, Djehiche and Hamadene [Citation13] prove the same results but under weak monotonicity assumptions and without the non-degeneracy condition on the forward equation.

The purpose of our work is to derive necessary and sufficient optimality conditions in terms of a stochastic maximum principle for a set û of admissible controls which maximize a cost functional of the form J(u)=E[h(X(T),M(T))+ϕ(Y(0),N(0))+0Tf(t,X(t),Y(t),Z(t),M(t),N(t),u(t))dt], with respect to admissible controls u, for some functions f,h,ϕ, under dynamics governed by MF-FBSDEs. More specifically, we consider the coupled system {dX(t)=b(t,X(t),M(t),u(t))dt+σ(t,X(t),M(t),u(t))dB(t),t[0,T],X(0)=x0,{dY(t)=g(t,X(t),Y(t),Z(t)),M(t),N(t),u(t))dt+Z(t)dB(t),t[0,T],Y(T)=ψ(X(T)), for some functions b,σ and a Brownian motion B(t). M(t) and N(t) denote the marginal laws of X and Y, respectively. As an application, we will consider a risk minimization control problem. More precisely, we want to minimize the risk given by Y(0)=E[φ(X(T))] such that E[φ(X(T))] is the convex risk measure by means of backward stochastic differential equations of mean-field type (MF-BSDEs). Let us recall what we mean by the convex risk measure:

Definition 1.1.

A convex risk measure is a map E:Lp(FT)R, p[2,] that satisfies the following properties:

  • (Convexity) E(λφ1+(1λ)φ2)λE(φ1)+(1λ)E(φ2) for all λ[0,1] and all φ1,φ2Lp(FT).

  • (Monotonicity) If φ1φ2, then E(φ1)E(φ2).

  • (Translation invariance) E(φ+a)=E(φ)a for all φLp(FT) and all constants a.

  • E(0)=0.

The construction of risk measures from solutions of BSDEs is given as follows: Assume that M(t):=E[Y(t)] in the driver g(t,y,m,z,n) of the above MF-BSDE and that zg(t,y,E[y],z) is convex for all t. Then E[φ(X(T))]=Y(0) defines a convex risk measure. This shows how crucial is the choice of the functional g. Through this connection, the problem of risk minimization is equivalent to stochastic optimal control of MF-FBSDEs, as shown in Øksendal and Sulem [Citation14], for the non-mean-field case. The rest of the paper is organized as follows. In Section 2, we give some mathematical background. In Section 3, we study a stochastic optimal control of MF-FBSDE where sufficient and necessary optimality conditions are derived. In the last section, we construct a dynamic risk measure by means of MF-BSDE and then we solve an associated risk minimization problem.

2. Generalities

Let B=B(t),t[0,T] be a one-dimensional Brownian motion defined in a complete filtered probability space (Ω,F,F,P). The filtration F={Ft}t0 is assumed to be the P-augmented filtration generated by B.

Definition 2.1.

Let Z be the set of integers.

• Let Mk be the space of random measures μ on R equipped with the norm (2.1) μLk2:=E[R|μ̂(y)|2(1+|y|)key2dy],kZ,(2.1) where μ̂ is the Fourier transform of the measure μ, i.e., μ̂(y):=Reixydμ(x);yR. We endow Mk with the inner product μ,η:=R|μ̂(y)η̂(y)|2(1+|y|)key2 μ,η,yR, μ̂ and η̂ are the Fourier transform of the measures μ and η, respectively. Then (Mk,||·||Mk) is a pre-Hilbert space, for each k. Let M be the union (inductive limit) of Mk,kZ.

• We denote by M0 the set of all deterministic elements of M.

We give some examples:

Example 2.2

(Measures). Let us give some examples of measures in M00 and M0:

  1. Suppose that μ=δx0, the unit point mass at x0R. Then δx0M00 and μ̂(y)=Reixydμ(x)=eix0y,

    and hence μM002=R|eix0y|2ey2dy<.

  2. Suppose dμ(x)=f(x)dx, where fL1(R). Then μM00 and by Riemann–Lebesque lemma, μ̂(y)C0(R), i.e., μ̂ is continuous and μ̂(y)0 when |y|. In particular, |μ̂| is bounded on R and hence μM002=R|μ̂(y)|2ey2dy<.

  3. Suppose that μ is any finite positive measure on R. Then μM00 and |μ̂(y)|Rdμ(y)=μ(R)<, for all y,

    and hence μM002=R|μ̂(y)|2ey2dy<.

  4. Next, suppose x0=x0(ω) is random. Then δx0(ω) is a random measure in M0. Similarly, if f(x)=f(x,ω) is random, then dμ(x,ω)=f(x,ω)dx is a random measure in M0.

We denote by U a nonempty convex subset of R and we denote by UG the set of U-valued G-progressively measurable processes where G:={Gt}t0 with GtFt for all t0; we consider them as the admissible control processes.

We will also use the following spaces:

  • S2 is the set of R-valued F-adapted càdlàg processes X=X(t),t[0,T], such that ||X||S22:=E[supt[0,T]|X(t)|2] < ,

  • L2 is the set of R-valued F-adapted processes Q=Q(t),t[0,T], such that ||QL22:=E[0T|Q(t)|2dt]< .

  • K denotes the set of absolutely continuous functions m:[0,T]M0.

  • K is the set of bounded linear functionals K:M0R equipped with the operator norm ||K||K:=supmM0,||m||M01|K(m)|.

  • SK2 is the set of F-adapted stochastic processes p:[0,T]×ΩK, such that ||p||SK22:=E[supt[0,T]||p(t)||K2]<.

  • LK2 is the set of F-adapted stochastic processes q:[0,T]×ΩK, such that ||q||LK22:=E[0T||q(t)||K2dt]<.

We recall now the notion of differentiability which will be used in the sequel.

Let X,Y be two Banach spaces with norms ||·X,||·Y, respectively, and let F:XY.

  • We say that F has a directional derivative (or Gateaux derivative) at vX in the direction if DwF(v):=limε01ε(F(v+εw)F(v))

exists in Y.
  • We say that F is Fréchet differentiable at vX if there exists a continuous linear map A:XY such that limh0hX1||h||X||F(v+h)F(v)A(h)Y=0, where A(h)=A,h is the action of the liner operator A on h. In this case we call A the gradient (or Fréchet derivative) of F at v and we write A=vF.

  • If F is Fréchet differentiable at v with Fréchet derivative vF, then F has a directional derivative in all directions wX and DwF(v)=vF(w)=vF,w.

    In particular, note that if F is a linear operator, then vF=F for all v.

3. Optimal control problem

Here we denote by M(t):=L(X(t)) the law of X(t) at time t and by N(t):=L(Y(t)) the law of Y(t) at time t. We assume that our system is governed by a coupled system of MF-FBSDE as follows:

The MF-SDE Xu(t)=X(t) is given by (3.1) {dX(t)=b(t,X(t),M(t),u(t))dt+σ(t,X(t),M(t),u(t))dB(t),t[0,T],X(0)=x0,(3.1) for functions σ,b:Ω×[0,T]×R×M0×UR which are supposed to be Ft-measurable and the initial value x0R.

The couple MF-BSDE (Yu(t),Zu(t))=(Y(t),Z(t)) satisfies (3.2) {dY(t)=g(t,X(t),Y(t),Z(t)),M(t),N(t),u(t))dt+Z(t)dB(t),t[0,T],Y(T)=ψ(X(T)),(3.2) where g:Ω×[0,T]×R3×M02×UR is F-adapted and ψ:Ω×R×M0R is FT-measurable.

It follows from the definition of the norm (Equation2.1) that ||L(X(1))L(X(2))||M02E[(X(1)X(2))2], where X(1) and X(2) are random variables that follow the distributions L(X(1)) and L(X(2)), respectively.

Assume that (C is a constant that may change from line to line)

(A1) there exists C > 0, such that

  • for all t[0,T], for all fixed uU, x,xR,m,mM0 |σ(t,x,m,u)σ(t,x,m,u)|+|b(t,x,m,u)b(t,x,m,u)|C(|xx|+||mm||M0).

  • for all t[0,T], for all fixed uU, |σ(t,0,δ0,u)|+|b(t,0,δ0,u)|C,

where δ0 is the distribution law of zero, i.e., the Dirac measure with mass at zero.

(A2) there exists C > 0, such that, for all fixed uU and all knowing X(t)S2 of Equation (Equation3.1) and M(t):=L(X(t))M0, we have

  • for all t[0,T], y,y,z,zR,n,nM0 |g(t,x,y,z,m,n,u)g(t,x,y,z,m,n,u)|C(|yy|+|zz|+||nn||M0).

  • for all t[0,T], |g(t,x,0,0,m,δ0,u)|C.

Proposition 3.1.

Under Assumptions (ACitation1) and (ACitation2), the MF-FBSDE (3.Citation1)–(3.Citation2) admits a unique solution (X,Y,Z)S2×S2×L2.

Since the system is partially coupled i.e., the forward equation does not depend on the solution of the backward one, we can solve the system separately as follows: we first find a solution X(t) of the MF-SDE (Equation3.1) and then we plug it into the backward EquationEquation (3.2), then we solve it.

Our aim is to maximize the performance functional of the form J(u)=E[h(X(T),M(T))+ϕ(Y(0),N(0))+0Tf(t,X(t),Y(t),Z(t),M(t),N(t),u(t))dt], over all admissible controls, for functions f:Ω×[0,T]×R3×M02×UR,h:Ω×R×M0R and ϕ:Ω×R×M0R.

Now, we can define the Hamiltonian H:Ω×[0,T]×R3×M02×U×R2×K×R×KR by (3.3) H(t,x,y,z,m,n,u,p0,q0,p1,λ0,λ1)=f(t,x,y,z,n,u)+p0b(t,x,m,u)+q0σ(t,x,m,u)+λ0g(t,x,y,z,m,n,u)+p1,mλ1,n.(3.3)

Remark 3.2.

For ease of notation we drop the dependence of all variables except for the time t,Φ{σ,f,H,h,g,ϕ}, we write Φ(t),t. Moreover, we will use Φ̂(t):=Φ(t,X̂(t),Ŷ(t),Ẑ(t),M̂(t),N̂(t),û(t))Φˇ(t):=Φ(t,X̂(t),Ŷ(t),Ẑ(t),M̂(t),N̂(t),u(t))].

We assume that

(A3) {σ,f,H,h,g,ϕ,ψ} are continuously differentiable with bounded partial derivatives w.r.t all the variables.

For uU with corresponding solution Xu=X, define, whenever solutions exist, pû=p=(p0,p1) and qû=q=(q0,q1) and λû=λ=(λ0,λ1) by the adjoint equations:

The BSDE for the unknown processes (p0,q0)S2×L2 (3.4) {dp0(t)=xH(t)dt+q0(t)dB(t),t[0,T],p0(T)=xh(T)+λ0(T)xψ(T).(3.4)

The MF-BSDE for the unknown processes (p1,q1)SK2×LK2 (3.5) {dp1(t)=mH(t)dt+q1(t)dB(t),t[0,T],p1(T)=mh(T).(3.5)

The forward SDE λ0S2 (3.6) {dλ0(t)=yH(t)dt+zH(t)dB(t),t[0,T],λ0(0)=yϕ(0),(3.6) and λ1SK2 (3.7) {dλ1(t)=nH(t)dt,t[0,T],λ1(0)=nϕ(0).(3.7)

Remark 3.3.

The real-valued linear system of FBSDE (Equation3.4) and (Equation3.6) have a unique solution by Proposition 3.1 since the coefficients satisfy condition (ACitation3). However, EquationEquation (3.5) is equivalent to the degenerate BSDE p1(t)=mh(T)+tT{mf(s)+p0(s)mb(s)+q0(s)mσ(s)}dstTq1(t)dB(t).

We take conditional expectation to obtain p1(t)=E[mh(T)+tT{mf(s)+p0(s)mb(s)+q0(s)mσ(s)}ds|Ft].

Similarly, a solution for (Equation3.7) is given by λ1(t)=nϕ(0)+0t{nf(s)+λ0(s)ng(s)}ds.

Before stating and proving sufficient and necessary conditions of optimality, we need the following result, which is Lemma 2.3 in Agram and Øksendal [Citation7].

Lemma 3.4.

Suppose that X(t) is an Itô process of the form {dX(t)=θ(t)dt+γ(t)dB(t),t[0,T],X(0)=x0R, where θ,γ are adapted processes.

Then the map M(t):[0,T]M0 is absolutely continuous.

It follows that tM(t) is differentiable for t-a.e. We will in the following use the notation M(t)=ddtM(t).

In fact, it is proven in [Citation7] that if M(t)Mk then M(t)Mk4;kZ.

3.1. Sufficient optimality conditions

We state and prove a type of a verification theorem.

Theorem 3.5.

Suppose that ûUG with corresponding solutions X̂(t),(Ŷ(t),Ẑ(t)), (p0(t),q0(t)),(p1(t),q1(t)),λ0(t),λ1(t) to Equations (3.1), (Equation3.2), (Equation3.4), (Equation3.5), (Equation3.6) and (Equation3.7), respectively. Suppose that x,mh(k,m), y,nϕ(y,n),xψ(x), x,y,z,m,n,uH(·,x,y,z,m,n,u), are concave functions P-a.s for each t[0,T]. Moreover, E[Ĥ(t)|Gt]=maxuUE[Hˇ(t)|Gt],

P-a.s for all t [0,T]. Then û is an optimal control.

Proof.

We show that J(u)J(û)0, for an arbitrary u and a fixed optimal ûUG.

We introduce first the following notation Φ{σ,f,H,h,g,ϕ,M,N,M,N} and t, δΦ(t)=Φˇ(t)Φ̂(t), and δM(t)=δ(ddtM(t))=ddt(δM(t)).

From the definition of the Hamiltonian (Equation3.3), we have δf(t)=δH(t)δb(t)p0(t)δσ(t)q0(t)p1(t),δM(t)λ1(t),δN(t), and (3.8) J(u)J(û)=E[0T{δH(t)δb(t)p0(t)δσ(t)q0(t)p1(t),δM(t)λ1(t),δN(t)}dt+δh(T)+δΦ(0)].(3.8)

We use the concavity of h and ϕ as well as the boundary values of EquationEquations (3.4), (3.5), (3.Citation6) and (Equation3.7) (3.9) δh(T)+δϕ(0)xh(T)δX(T)+mh(T),δM(T)+xϕ(0)δY(0)+nϕ(0),δN(0)=p0(T)δX(T)λ0(T)δX(T)+p1(T),δM(T)+λ0(0)δY(0)+λ1(0),δN(0).(3.9)

Applying Itô formula to p0(t)δX(t),p1(t),δM(t),λ0(t)δY(t) and λ1(t),δN(t), yields the following duality relations: (3.10) E[p0(T)δX(T)]E[λ0(T)xψ(T)]=E[0Tp0(t)δb(t)dt0TδX(t)xH(t)dt]+E[0Tq0(t)δσ(t)dt]E[λ0(T)xψ(T)],(3.10) (3.11) E[p1(T),δM(T)]=E[0Tp1(t),δM(t)dt0TmĤ(t),δM(t)dt],(3.11) (3.12) E[λ0(T)δY(T)]E[λ0(0)δY(0)]=E[0Tλ0(t)δg(t)dt]+E[0TδY(t)yĤ(t)dt]+E[0TδZ(t)zĤ(t)dt].(3.12)

Concavity of ψ gives (3.13) E[λ0(0)δY(0)]=E[λ0(T)δψ(T)]+E[0Tλ0(t)δg(t)dt]E[0TδY(t)yĤ(t)dt]E[0TδZ(t)zĤ(t)dt]E[λ0(T)xψ(T)δX(T)]+E[0Tλ0(t)δg(t)dt]E[0TδY(t)yĤ(t)dt]E[0TδZ(t)zĤ(t)dt],E[λ1(T)δN(T)]E[λ1(0),δN(0)]=E[0Tλ1(t),δN(t)+nĤ(t),δN(t)dt].(3.13)

By the concavity of H, we obtain (3.14) δH(t)xĤ(t)δX(t)+yĤ(t)δY(t)+zĤ(t)δZ(t)+mĤ(t),δM(t)+nĤ(t),δN(t)+uĤ(t)δu(t).(3.14)

Finally, by substituting the derived duality relations (Equation3.10), (Equation3.11), (Equation3.12) and (Equation3.13) in (Equation3.8) and using the estimates (Equation3.9), (Equation3.14), we obtain J(u)J(û)E[0TuĤ(t)δu(t)].

Using the tower property and the fact that u(t) is G-adapted the desired result follows J(u)J(û)E[0TE[uĤ(t)|Gt]δu(t)dt]0, and thus, û is optimal.□

3.2. Necessary optimality conditions

Given an arbitrary but fixed control uUG, we define (3.15) uρ:=û+ρu,ρ[0,1].(3.15)

Note that, the convexity of U and UG guarantees that uρUG,ρ[0,1]. We denote by Xρ:=Xuρ and by X̂:=Xû, the solution processes corresponding to uρ and û, respectively.

For each t0[0,T] and all bounded Gt0-measurable random variables α, the process u(t)=α1(t0,T](t), belongs to UG.

In general, if Kû(t) is a process depending on û, we define the operator D on K by (3.16) DKû(t):=DuKû(t)=ddρKû+ρu(t)|ρ=0,(3.16) whenever the derivative exists.

Define the following derivative processes DX(t):=ddρXû+ρu(t)|ρ=0=X(t),DY(t):=ddρYû+ρu(t)|ρ=0=Y(t),DZ(t):=ddρZû+ρu(t)|ρ=0=Z(t),DN(t):=ddρNû+ρu(t)|ρ=0,DM(t):=ddρMû+ρu(t)|ρ=0,DN(t):=ddρddtMû+ρu(t)|ρ=0,DM(t):=ddρddtMû+ρu(t)|ρ=0, such that (3.17) {dX(t)={xb(t)X(t)+mb(t),DM(t)+ub(t)u(t)}dt+{xσ(t)X(t)+mσ(t),DM(t)+uσ(t)u(t)}dB(t),t[0,T],X(0)=0,(3.17) and (3.18) {dY(t)={xg(t)X(t)+yg(t)Y(t)+zg(t)Z(t)+mg(t),DM(t)+ng(t),DN(t)+ug(t)u(t)}dt+Z(t)dB(t),t[0,T],Y(T)=xψ(T)X(T).(3.18)

Remark 3.6.

Equations (Equation3.17), (Equation3.18) are linear FBSDE with bounded coefficients, then by Proposition 3.1 they have a unique solution.

Theorem 3.7.

Let ûUG be the optimal control and X(t),(Y(t),Z(t)),(p0(t),q0(t)),(p1(t),q1(t)),λ0(t),λ1(t) be the corresponding solutions to the EquationEquations (Equation3.17), (Equation3.18), (Equation3.4), (Equation3.5), (Equation3.6), (Equation3.7), respectively. Then, the following statements are equivalent

  1. ddρJ(û+ρu)|ρ=0=0 for all bounded βUG.

  2. E[uĤ(t)|Gt]=0 for all t[0,T].

Proof

We first prove Theorem 3.7 by assuming (i) and aiming to show (ii) 0=ddρJ(u+ρu)|ρ=0=E[0Tddρfρ(t)|ρ=0dt+p0(T)X(T)λ0(T)xψ(T)X(T)+p1(T),DM(T)+λ0(0)Y(0)+λ1(0),DN(0)]

{we substitute f(t) from Equation (3.3)} =E[0Tddρ{Hρ(t)p0(t)bρ(t)q0(t)σρ(t)λ0(t)gρ(t)p1(t),Mρ(t)+λ1(t),Nρ(t)}|ρ=0dt+p0(T)X(T)λ0(T)X(T)+p1(T),DM(T)+λ0(0)Y(0)+λ1(0),DN(0)], by using the chain rule, we obtain ddρHρ(t)|ρ=0=xH(t)X(t)+yH(t)Y(t)+zH(t)Z(t)+mH(t),DM(t)+nH(t),DN(t)+uH(t)u(t),ddρp0(t)bρ(t)|ρ=0=p0(t)xb(t)X(t)+p0(t)mb(t),DM(t)+p0(t)ub(t)u(t),ddρq0(t)σρ(t)|ρ=0=q0(t)xσ(t)X(t)+q0(t)mσ(t),DM(t)+q0(t)uσ(t)u(t),ddρλ0(t)gρ(t)|ρ=0=λ0(t)xg(t)X(t)+λ0(t)yg(t)Y(t)+λ0(t)zg(t)Z(t)+λ0(t)mg(t),DM(t)+λ0(t)ng(t),DN(t)+λ0(t)ug(t)u(t),ddρp1(t),Mρ(t)|ρ=0=p1(t),DM(t), and ddρλ1(t),Nρ(t)|ρ=0=λ1(t),DN(t).

We apply Itô formula to p0(t)X(t),p1(t),DM(t),λ0(t)Y(t) and λ1(t),DN(t) then we take the expectation, we obtain the following important duality relations: E[p0(T)X(T)]=E[0T{p0(t)xb(t)X(t)+p0(t)mb(t),DM(t)+p0(t)ub(t)u(t)xH(t)X(t)+q0(t)xσ(t)X(t)+q0(t)mσ(t),DM(t)+q0(t)uσ(t)u(t)}dt],E[p1(T),DM(T)]=E[0Tp1(t),DM(t)mH(t),DM(t)dt],E[λ0(T)Y(T)]E[λ0(0)Y(0)]=E[0T{λ0(t)xg(t)X(t)λ0(t)yg(t)Y(t)λ0(t)zg(t)Z(t)λ0(t)mg(t),DM(t)λ0(t)ng(t),DN(t)λ0(t)ug(t)u(t)+yH(t)Y+zH(t)Z(t)}dt],E[λ1(T),DN(T)]E[λ1(0),DN(0)]=E[0T{λ1(t),DN(t)+nH(t),DN(t)}dt].

By substituting the derived duality relations and the partial derivatives of f(t) the desired result follows. This proof can be reversed to prove (ii)(i). We omit the details.□

4. Mean-field risk minimization

4.1. Mean-field dynamic risk measure

In this section, we are interested in a particular class of MF-BSDE of the following form (4.1) {dY(t)=f(t,Y(t),E[Y(t)],Z(t))dt+Z(t)dB(t),t[0,T],Y(T)=ξ,(4.1) where f(t,Y(t),E[Y(t)],Z(t)):=r(t)Y(t)r(t)E[Y(t)]+F(t,Z(t)).

We assume that the generator (y,y¯,z)f(t,Y(t),E[Y(t)],Z(t)):Ω×[0,T]×R×R×RR is F-adapted, uniformly Lipschitz and concave, and the terminal condition ξL2(Ω,FT).

Definition 4.1.

Define Et: (T;ξ)Et(T;ξ) by Et(T;ξ)=Yt(T;ξ),t[0,T], where Yt(T;ξ) is a component of the solution of the MF-BSDE (Equation4.1) with terminal horizon T, terminal condition ξ and driver f. Then Et(T;ξ) is a dynamic risk measure induced by the MF-BSDE (Equation4.1).

We may remark that the driver f depends linearly on Y and its expected value E[Y], and nonlinearly on Z. This is interpreted as a market with interest rates (r(t),r(t)). We can reformulated this as a problem with a driver independent of Y and E[Y] by discounting the financial position ξ. We assume that the instantaneous interest rates r(t) and r(t) are deterministic. We denote by E·, the corresponding discounted risk measure.

Define the discounted process Yr(t):=e0t(r(s)+r(s))dsY(t).

Then Yr with driver Fr(·,t,Z(t)):=e0t(r(s)+r(s))dsF(·,t,e0t(r(s)+r(s))dsZ(t)), and the terminal value ξr:=e0t(r(s)+r(s))dsξ is a part of the solution of the associated BSDE. We obtain also a discounted risk measure accordingly E0(ξ,T)=E0r(e0t(r(s)+r(s))dsξ,T).

This discounted risk measure is translation-invariant because Fr does not depend on Y, we have for ξL2(Ω,FT) and aR, E0(ξ+ae0t(r(s)+r(s))ds,T)=E0r(e0t(r(s)+r(s))dsξ+a,T)=E0r(e0t(r(s)+r(s))dsξ,T)a=E0(ξ,T)a.

Similarly, we can get for each t[0,T], that E(ξ,T)=Er(e0t(r(s)+r(s))dsξ,T) is translation-invariant.

4.2. Optimal portfolio with mean-field risk minimization

Consider a financial market with two investment possibilities:

  1. Safe, or risk-free asset with unit price S0(t)=1,t[0,T].

  2. Risky asset with unit price dS1(t)=S1(t)[b0(t)dt+σ0(t)dB(t)],t[0,T].

Let π(t) be a self-financing portfolio invested in the risky asset at time t. We want to minimize the risk φ(Xπ(T)) of the terminal value of the wealth process Xπ(t) corresponding to a portfolio π which satisfies the linear SDE (4.2) {dXπ(t)=π(t)Xπ(t)[b0(t)dt+σ0(t)dB(t)], t[0,T],Xπ(0)=x0,(4.2) such that φ(Xπ(T))=Yπ(0) where Yπ(t) satisfies a MF-BSDE (4.3) {dYπ(t)= [r0(t)E[Yπ(t)]+F(Z(t))]dtZ(t)dB(t), t[0,T],Yπ(T)= Xπ(T).(4.3)

Here we assume that b0(t), σ0(t), r0(t) are given deterministic functions and F:RR is some given concave function. We want to find π̂UG such that infπUG(Yπ(0))=Yπ̂(0).

Define the Hamiltonian H that corresponds to our problem by H(t,x,z,y¯,π,p0,q0,λ0,λ1)=p0b0πx+q0σ0πx+λ0(r0y¯+F(z))+λ1,y¯.

The couple (p0,q0) solution of the following BSDE {dp0(t)=[p0(t)b0(t)π(t)+q0(t)σ0(t)π(t)]dt+q0(t)dB(t),t[0,T],p0(T)=λ0(T), and (p1,q1) satisfies {dp1(t)=q1(t)dB(t),t[0,T],p1(T)=0.

The equation for λ0 is given by the forward SDE (4.4) {dλ0(t)=zF(Z(t))λ0(t)dB(t),t[0,T],λ0(0)=1,(4.4) and λ1 satisfies {dλ1(t)=r0(t)λ0(t)dt,t[0,T],λ1(0)=0.

The first order necessary optimality condition gives p̂0(t)b0(t)X̂(t)+q̂0(t)σ0(t)X̂(t)=0,

where we denoted by X̂(t)=Xπ̂(t) and so on. Since X̂(t)>0 for all t P-a.s., we obtain (4.5) p̂0(t)b0(t)+q̂0(t)σ0(t)=0,(4.5) which implies {dp̂0(t)=q̂0(t)dB(t)=b0(t)σ0(t)p̂0(t)dB(t),t[0,T],p̂0(T)=λ̂0(T), this together with EquationEquation (Equation4.4), yields p̂0(t)=λ̂0(t), q̂0(t)=zF(Ẑ(t))λ̂0(t).

From (Equation4.5), we get zF(Ẑ(t))=b0(t)σ0(t).

For example, if we choose (4.6) F(z)=12z2.(4.6)

That is Ẑ(t)=b0(t)σ0(t).

Substituting the expression of Ẑ(t) above into the MF-BSDE (Equation4.3), we obtain (4.7) {dŶ(t)=[r0(t)E[Ŷ(t)]12(b0(t)σ0(t))2]dtb0(t)σ0(t)dB(t), t[0,T],Ŷ(T)=X̂(T).(4.7)

Consequently dE[Ŷ(t)]=[r0(t)E[Ŷ(t)]12(b0(t)σ0(t))2]dt, thus (4.8) E[Ŷ(t)]=exp(0tr0(s)ds)[Ŷ(0)+120tb02(s)σ02(s)exp(0sr0(α)dα)ds].(4.8)

Define Γ(t) to be the solution of the linear SDE {dΓ(t)=b0(t)σ0(t)Γ(t)dB(t), t[0,T],Γ(0)=1,

or explicitly (4.9) Γ(t)=exp(0tb0(s)σ0(s)dB(s)120t(b0(s)σ0(s))2ds), t[0,T].(4.9)

By the Girsanov theorem of change of measures, we know that there exists an equivalent local martingale measure QP, such that dQ=Γ(T)dP on FT, with Γ(T)=dQdP is the Radon–Nikodym derivative of Q with respect to P on FT.

Substituting (Equation4.8), (Equation4.9) into (Equation4.7) we have X̂(T)=Ŷ(T)=Ŷ(0)+exp(0tr0(s)ds)[Ŷ(0)+120tb02(s)σ02(s)exp(0sr0(α)dα)]+120T(b0(s)σ0(s))2ds+0Tb0(s)σ0(s)dB(s)=Ŷ(0)+exp(0tr0(s)ds)[Ŷ(0)+120tb02(s)σ02(s)exp(0sr0(α)dα)ds]lnΓ(t).

Taking the expectation but now with respect to the new measure Q, we get (4.10) Ŷ(0)=x0exp(0tr0(s)ds)[Ŷ(0)+120tb02(s)σ02(s)exp(0sr0(α)dα)ds]EQ[lnΓ(T)]=11exp(0tr0(s)ds){x0exp(0tr0(s)ds)[120tb02(s)σ02(s)exp(0sr0(α)dα)ds]E[Γ(T)lnΓ(T)]},(4.10) where E[Γ(T)lnΓ(T)] is the entropy of Q with respect to P.

Since we obtained the optimal value of Ŷ(0), we can get the corresponding optimal terminal wealth X̂(T).

Summarizing, we have the following conclusion:

Theorem 4.2.

Suppose that (Equation4.6) holds. Then the minimal risk of our problem is given by (Equation4.10).

Additional information

Funding

This research was carried out with support of the Norwegian Research Council, within the research project Challenges in Stochastic Control, Information and Applications (STOCONINF), project number 250768/F20; and U.S. Air Force Office of Scientific Research under grant number FA9550-17-1-0259.

References

  • Lasry, J.-M., Lions, P.-L. (2007). Mean field games. Jpn. J. Math. 2(1):229–260. DOI: 10.1007/s11537-007-0657-8.
  • Huang, J., Wang, S., Wu, Z. (2016). Backward mean-field linear–quadratic–Gaussian (LQG) games: Full and partial information. IEEE Trans. Automat. Contr. 61(12):3784–3796. DOI: 10.1109/TAC.2016.2519501.
  • Xu, R., Zhang, F. (2020). ε-Nash mean-field games for general linear–quadratic systems with applications. Automatica. 114:108835. DOI: 10.1016/j.automatica.2020.108835.
  • Xu, R., Shi, J. (2019). ε-Nash mean-field games for linear–quadratic systems with random jumps and applications. Int. J. Control. DOI: 10.1080/00207179.2019.1651940.
  • Huang, M. (2010). Large-population LQG games involving a major player: the Nash certainty equivalence principle. SIAM J. Control Optim. 48(5):3318–3353. DOI: 10.1137/080735370.
  • Nourian, M., Caines, P. E. (2013). ε-Nash mean field game theory for nonlinear stochastic dynamical systems with major and minor agents. SIAM J. Control Optim. 51(4):3302–3331. DOI: 10.1137/120889496.
  • Agram, N., Øksendal, B. (2019). Model uncertainty stochastic mean-field control. Stoch. Anal. Appl. 37(1):36–21. DOI: 10.1080/07362994.2018.1499036.
  • Agram, N., Øksendal, B. (2019). Stochastic control of memory mean-field processes. Appl. Math. Optim. 79(1):181–204. DOI: 10.1007/s00245-017-9425-1.
  • Agram, N., Bachouch, A., Øksendal, B., Proske, F. (2019). Singular control optimal stopping of memory mean-field processes. SIAM J. Math. Anal. 51(1):450–468. DOI: 10.1137/18M1174787.
  • Lions, P. Cours au college de france: Th´eorie des jeux ´a champs moyens (2014).
  • Carmona, R., Delarue, F. (2015). Forward–backward stochastic differential equations and controlled McKean–Vlasov dynamics. Ann. Probab. 43(5):2647–2700. DOI: 10.1214/14-AOP946.
  • Bensoussan, A., Yam, S. C. P., Zhang, Z. (2015). Well-posedness of mean-field type forward–backward stochastic differential equations. Stoch. Process. Appl. 125(9):3327–3354. DOI: 10.1016/j.spa.2015.04.006.
  • Djehiche, B., Hamadene, S. (2019). Mean-field backward–forward stochastic differential equations and nonzero sum stochastic differential games. arXiv preprint arXiv:1904.06193
  • Øksendal, B., Sulem, A. (2015). Risk minimization in financial markets modeled by Ito⁁-L´evy processes. Afr. Math. 26(5–6):939–979. DOI: 10.1007/s13370-014-0248-9.