Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Angiotensin II, dopamine and nitric oxide. An asymmetrical neurovisceral interaction between brain and plasma to regulate blood pressure

  • Received: 10 June 2019 Accepted: 24 July 2019 Published: 26 July 2019
  • Vital functions, such as blood pressure, are regulated within a framework of neurovisceral integration in which various factors are involved under normal conditions maintaining a delicate balance. Imbalance of any of these factors can lead to various pathologies. Blood pressure control is the result of the balanced action of central and peripheral factors that increase or decrease. Special attention for blood pressure control was put on the neurovisceral interaction between Angiotensin II and the enzymes that regulate its activity as well as on nitric oxide and dopamine. Several studies have shown that such interaction is asymmetrically organized. These studies suggest that the neuronal activity related to the production of nitric oxide in plasma is also lateralized and, consequently, changes in plasma nitric oxide influence neuronal function. This observation provides a new aspect revealing the complexity of the blood pressure regulation and, undoubtedly, makes such study more motivating as it may affect the approach for treatment.

    Citation: I. Banegas, I. Prieto, A.B. Segarra, M. Martínez-Cañamero, M. de Gasparo, M. Ramírez-Sánchez. Angiotensin II, dopamine and nitric oxide. An asymmetrical neurovisceral interaction between brain and plasma to regulate blood pressure[J]. AIMS Neuroscience, 2019, 6(3): 116-127. doi: 10.3934/Neuroscience.2019.3.116

    Related Papers:

    [1] Jinheng Liu, Kemei Zhang, Xue-Jun Xie . The existence of solutions of Hadamard fractional differential equations with integral and discrete boundary conditions on infinite interval. Electronic Research Archive, 2024, 32(4): 2286-2309. doi: 10.3934/era.2024104
    [2] Qingcong Song, Xinan Hao . Positive solutions for fractional iterative functional differential equation with a convection term. Electronic Research Archive, 2023, 31(4): 1863-1875. doi: 10.3934/era.2023096
    [3] Hongyu Li, Liangyu Wang, Yujun Cui . Positive solutions for a system of fractional $ q $-difference equations with generalized $ p $-Laplacian operators. Electronic Research Archive, 2024, 32(2): 1044-1066. doi: 10.3934/era.2024051
    [4] Xiaoli Wang, Peter Kloeden, Meihua Yang . Asymptotic behaviour of a neural field lattice model with delays. Electronic Research Archive, 2020, 28(2): 1037-1048. doi: 10.3934/era.2020056
    [5] David Cheban, Zhenxin Liu . Averaging principle on infinite intervals for stochastic ordinary differential equations. Electronic Research Archive, 2021, 29(4): 2791-2817. doi: 10.3934/era.2021014
    [6] Mufit San, Seyma Ramazan . A study for a higher order Riemann-Liouville fractional differential equation with weakly singularity. Electronic Research Archive, 2024, 32(5): 3092-3112. doi: 10.3934/era.2024141
    [7] Limin Guo, Weihua Wang, Cheng Li, Jingbo Zhao, Dandan Min . Existence results for a class of nonlinear singular $ p $-Laplacian Hadamard fractional differential equations. Electronic Research Archive, 2024, 32(2): 928-944. doi: 10.3934/era.2024045
    [8] Chungen Liu, Yuyou Zhong . Infinitely many periodic solutions for ordinary $ p(t) $-Laplacian differential systems. Electronic Research Archive, 2022, 30(5): 1653-1667. doi: 10.3934/era.2022083
    [9] Xu Liu, Jun Zhou . Initial-boundary value problem for a fourth-order plate equation with Hardy-Hénon potential and polynomial nonlinearity. Electronic Research Archive, 2020, 28(2): 599-625. doi: 10.3934/era.2020032
    [10] Milena Dimova, Natalia Kolkovska, Nikolai Kutev . Global behavior of the solutions to nonlinear Klein-Gordon equation with critical initial energy. Electronic Research Archive, 2020, 28(2): 671-689. doi: 10.3934/era.2020035
  • Vital functions, such as blood pressure, are regulated within a framework of neurovisceral integration in which various factors are involved under normal conditions maintaining a delicate balance. Imbalance of any of these factors can lead to various pathologies. Blood pressure control is the result of the balanced action of central and peripheral factors that increase or decrease. Special attention for blood pressure control was put on the neurovisceral interaction between Angiotensin II and the enzymes that regulate its activity as well as on nitric oxide and dopamine. Several studies have shown that such interaction is asymmetrically organized. These studies suggest that the neuronal activity related to the production of nitric oxide in plasma is also lateralized and, consequently, changes in plasma nitric oxide influence neuronal function. This observation provides a new aspect revealing the complexity of the blood pressure regulation and, undoubtedly, makes such study more motivating as it may affect the approach for treatment.


    Time inconsistency in dynamic decision making is often observed in social systems and daily life. Motivated by practical applications, especially in mathematical economics and finance, time-inconsistency control problems have recently attracted considerable research interest and efforts attempting to seek equilibrium, instead of optimal controls. At a conceptual level, the idea is that a decision made by the controller at every instant of time is considered as a game against all the decisions made by the future incarnations of the controller. An "equilibrium" control is therefore one such that any deviation from it at any time instant will be worse off. The study on time inconsistency by economists can be dated back to Stroz [1] and Phelps [2,3] in models with discrete time (see [4] and [5] for further developments), and adapted by Karp [6,7], and by Ekeland and Lazrak [8,9,10,11,12,13] to the case of continuous time. In the LQ control problems, Yong [14] studied a time-inconsistent deterministic model and derived equilibrium controls via some integral equations.

    It is natural to study time inconsistency in the stochastic models. Ekeland and Pirvu [15] studied the non-exponential discounting which leads to time inconsistency in an agent's investment-consumption policies in a Merton model. Grenadier and Wang [16] also studied the hyperbolic discounting problem in an optimal stopping model. In a Markovian systems, Björk and Murgoci [17] proposed a definition of a general stochastic control problem with time inconsistent terms, and proposed some sufficient condition for a control to be solution by a system of integro-differential equations. They constructed some solutions for some examples including an LQ one, but it looks very hard to find not-to-harsh condition on parameters to ensure the existence of a solution. Björk, Murgoci and Zhou [18] also constructed an equilibrium for a mean-variance portfolio selection with state-dependent risk aversion. Basak and Chabakauri [19] studied the mean-variance portfolio selection problem and got more details on the constructed solution. Hu, Jin and Zhou [20,21] studied the general LQ control problem with time inconsistent terms in a non-Markovian system and constructed an unique equilibrium for quite general LQ control problem, including a non-Markovian system.

    To the best of our knowledge, most of the time-inconsistent problems are associated with the control problems though we use the game formulation to define its equilibrium. In the problems of game theory, the literatures about time inconsistency is little [22,23]. However, the definitions of equilibrium strategies in the above two papers are based on some corresponding control problems like before. In this paper, we formulate a general stochastic LQ differential game, where the objective functional of each player include both a quadratic term of the expected state and a state-dependent term. These non-standard terms each introduces time inconsistency into the problem in somewhat different ways. We define our equilibrium via open-loop controls. Then we derive a general sufficient condition for equilibrium strategies through a system of forward-backward stochastic differential equations (FBSDEs). An intriguing feature of these FBSDEs is that a time parameter is involved; so these form a flow of FBSDEs. When the state process is scalar valued and all the coefficients are deterministic functions of time, we are able to reduce this flow of FBSDEs into several Riccati-like ODEs. Comparing to the ODEs in [20], though the state process is scalar valued, the unknowns are matrix-valued because of two players. Therefore, such ODEs are harder to solve than those of [20]. Under some more stronger conditions, we obtain explicitly an equilibrium strategy, which turns out to be a linear feedback. We also prove that the equilibrium strategy we obtained is unique.

    The rest of the paper is organized as follows. The next section is devoted to the formulation of our problem and the definition of equilibrium strategy. In Section 3, we apply the spike variation technique to derive a flow of FBSEDs and a sufficient condition of equilibrium strategies. Based on this general results, we solve in Section 4 the case when the state is one dimensional and all the coefficients are deterministic. The uniqueness of such equilibrium strategy is also proved in this section.

    Let $ T > 0 $ be the end of a finite time horizon, and let $ (W_t)_{0\le t\le{T}} = (W_t^1, ..., W_t^d)_{0\le t\le{T}} $ be a $ d $-dimensional Brownian motion on a probability space $ (\Omega, \mathcal{F}, \mathbb{P}) $. Denote by $ (\mathcal{F}_t) $ the augmented filtration generated by $ (W_t) $.

    Let $ \mathbb{S}^n $ be the set of symmetric $ n\times n $ real matrices; $ L_{\mathcal{F}}^2(\Omega, \mathbb{R}^l) $ be the set of square-integrable random variables; $ L_{\mathcal{F}}^2(t, T;\mathbb{R}^n) $ be the set of $ \{\mathcal{F}_s\}_{s\in[t, T]} $-adapted square-integrable processes; and $ L_{\mathcal{F}}^2(\Omega; C(t, T;\mathbb{R}^n)) $ be the set of continuous $ \{\mathcal{F}_s\}_{s\in[t, T]} $-adapted square-integrable processes.

    We consider a continuous-time, $ n $-dimensional nonhomogeneous linear controlled system:

    $ dX_s = [A_sX_s+B_{1,s}'u_{1,s}+B_{2,s}'u_{2,s}+b_s]ds+\sum\limits_{j = 1}^d[C_s^jX_s+D_{1,s}^ju_{1,s}+D_{2,s}^ju_{2,s}+\sigma_s^j]dW_s^j, \\ X_0 = x_0. $ (2.1)

    Here $ A $ is a bounded deterministic function on $ [0, T] $ with value in $ \mathbb{R}^{n\times n} $. The other parameters $ B_1, B_2, C, D_1, D_2 $ are all essentially bounded adapted processes on $ [0, T] $ with values in $ \mathbb{R}^{l\times n}, \mathbb{R}^{l\times n}, \mathbb{R}^{n\times n}, \mathbb{R}^{n\times l}, \mathbb{R}^{n\times l} $, respectively; $ b $ and $ \sigma^j $ are stochastic processes in $ L^2_{\mathcal{F}}(0, T;\mathbb{R}^n) $. The processes $ u_i\in L^2_{\mathcal{F}}(0, T;\mathbb{R}^l), \ i = 1, 2 $ are the controls, and $ X $ is the state process valued in $ \mathbb{R}^{n} $. Finally, $ x_0\in\mathbb{R}^{n} $ is the initial state. It is obvious that for any controls $ u_i\in L^2_{\mathcal{F}}(0, T;\mathbb{R}^l), \ i = 1, 2 $, there exists a unique solution $ X\in L^2_{\mathcal{F}}(\Omega, C(0, T;\mathbb{R}^n)) $.

    As time evolves, we need to consider the controlled system starting from time $ t\in[0, T] $ and state $ x_t\in L^2_{\mathcal{F}_t}(\Omega; \mathbb{R}^n) $:

    $ dX_s = [A_sX_s+B_{1,s}'u_{1,s}+B_{2,s}'u_{2,s}+b_s]ds+\sum\limits_{j = 1}^d[C_s^jX_s+D_{1,s}^ju_{1,s}+D_{2,s}^ju_{2,s}+\sigma_s^j]dW_s^j, \\ X_t = x_t. $ (2.2)

    For any controls $ u_i\in L^2_{\mathcal{F}}(0, T;\mathbb{R}^l), \ i = 1, 2 $, there exists a unique solution $ X^{t, x_t, u_1, u_2}\in L^2_{\mathcal{F}}(\Omega, C(0, T;\mathbb{R}^n)) $.

    We consider a two-person differential game problem. At any time $ t $ with the system state $ X_t = x_t $, the $ i $-th ($ i = 1, 2 $) person's aim is to minimize her cost (if maximize, we can times the following function by $ -1 $):

    $ Ji(t,xt;u1,u2)=12EtTt[Qi,sXs,Xs+Ri,sui,s,ui,s]ds+12Et[GiXT,XT]12hiEt[XT],Et[XT]λixt+μi,Et[XT] $ (2.3)

    over $ u_1, u_2\in L_{\mathcal{F}}^2(t, T;\mathbb{R}^l) $, where $ X = X^{t, x_t, u_1, u_2} $, and $ \mathbb{E}_t[\cdot] = \mathbb{E}[\cdot|\mathcal{F}_t] $. Here, for $ i = 1, 2 $, $ Q_i $ and $ R_i $ are both given essentially bounded adapted process on $ [0, T] $ with values in $ \mathbb{S}^n $ and $ \mathbb{S}^l $, respectively, $ G_i, h_i, \lambda_i, \mu_i $ are all constants in $ \mathbb{S}^n $, $ \mathbb{S}^n $, $ \mathbb{R}^{n\times n} $ and $ \mathbb{R}^n $, respectively. Furthermore, we assume that $ Q_i, R_i $ are non-negative definite almost surely and $ G_i $ are non-negative definite.

    Given a control pair $ (u_1^*, u_2^*) $. For any $ t\in[0, T), \epsilon > 0 $, and $ v_1, v_2\in L_{\mathcal{F}_t}^2(\Omega, \mathbb{R}^l) $, define

    $ ut,ϵ,vii,s=ui,s+vi1s[t,t+ϵ),s[t,T], i=1,2. $ (2.4)

    Because each person at time $ t > 0 $ wants to minimize his/her cost as we claimed before, we have

    Definition 2.1. Let $ (u_1^*, u_2^*)\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^l)\times L_{\mathcal{F}}^2(0, T;\mathbb{R}^l) $ be a given strategy pair, and let $ X^* $ be the state process corresponding to $ (u_1^*, u_2^*) $. The strategy pair $ (u_1^*, u_2^*) $ is called an equilibrium if

    $ limϵ0J1(t,Xt;ut,ϵ,v11,u2)J1(t,Xt;u1,u2)ϵ0, $ (2.5)
    $ limϵ0J2(t,Xt;u1,ut,ϵ,v22)J2(t,Xt;u1,u2)ϵ0, $ (2.6)

    where $ u_i^{t, \epsilon, v_i}, i = 1, 2 $ are defined by (2.4), for any $ t\in[0, T) $ and $ v_1, v_2\in L_{\mathcal{F}_t}^2(\Omega, \mathbb{R}^l) $.

    Remark. The above definition means that, in each time $ t $, the equilibrium is a static Nash equilibrium in a corresponding game.

    Let $ (u_1^*, u_2^*) $ be a fixed strategy pair, and let $ X^* $ be the corresponding state process. For any $ t\in[0, T) $, as a similar arguments of Theorem 5.1 in pp. 309 of [24], defined in the time interval $ [t, T] $, there exist adapted processes $ (p_i(\cdot; t), (k_i^j(\cdot; t)_{j = 1, 2, ..., d}))\in L_{\mathcal{F}}^2(t, T;\mathbb{R}^n)\times(L_{\mathcal{F}}^2(t, T;\mathbb{R}^n))^d $ and $ (P_i(\cdot; t), (K_i^j(\cdot; t)_{j = 1, 2, ..., d}))\in L_{\mathcal{F}}^2(t, T;\mathbb{S}^n)\times(L_{\mathcal{F}}^2(t, T;\mathbb{S}^n))^d $ for $ i = 1, 2 $ satisfying the following equations:

    $ {dpi(s;t)=[Aspi(s;t)+dj=1(Cjs)kji(s;t)+Qi,sXs]ds+dj=1kji(s;t)dWjs,s[t,T],pi(T;t)=GiXThiEt[XT]λiXtμi, $ (3.1)
    $ {dPi(s;t)={AsPi(s;t)+Pi(s;t)As+Qi,s+dj=1[(Cjs)Pi(s;t)Cjs+(Cjs)Kji(s;t)+Kji(s;t)Cjs]}ds+dj=1Kji(s;t)dWji,s[t,T],Pi(T;t)=Gi, $ (3.2)

    for $ i = 1, 2 $. From the assumption that $ Q_i $ and $ G_i $ are non-negative definite, it follows that $ P_i(s; t) $ are non-negative definite for $ i = 1, 2 $.

    Proposition 1. For any $ t\in[0, T), \epsilon > 0 $, and $ v_1, v_2\in L_{\mathcal{F}_t}^2(\Omega, \mathbb{R}^l) $, define $ u_i^{t, \epsilon, v_i}, i = 1, 2 $ by (2.4). Then

    $ J1(t,Xt;ut,ϵ,v11,u2)J1(t,Xt;u1,u2)=Ett+ϵt{Λ1(s;t),v1+12H1(s;t)v1,v1}ds+o(ϵ), $ (3.3)
    $ J2(t,Xt;u1,ut,ϵ,v22)J2(t,Xt;u1,u2)=Ett+ϵt{Λ2(s;t),v2+12H2(s;t)v2,v2}ds+o(ϵ), $ (3.4)

    where $ \Lambda_i(s; t) = B_{i, s}p_i(s; t)+\sum_{j = 1}^d(D_{i, s}^j)'k_i^j(s; t)+R_{i, s}u_{i, s}^* $ and $ H_i(s; t) = R_{i, s}+\sum_{j = 1}^d(D_{i, s}^j)'P_i(s; t)D_{i, s}^j $ for $ i = 1, 2 $.

    Proof. Let $ X^{t, \epsilon, v_1, v_2} $ be the state process corresponding to $ u_i^{t, \epsilon, v_i}, i = 1, 2 $. Then by standard perturbation approach (cf. [20,25] or pp. 126-128 of [24]), we have

    $ Xt,ϵ,v1,v2s=Xs+Yt,ϵ,v1,v2s+Zt,ϵ,v1,v2s,s[t,T], $ (3.5)

    where $ Y\equiv Y^{t, \epsilon, v_1, v_2} $ and $ Z\equiv Z^{t, \epsilon, v_1, v_2} $ satisfy

    $ {dYs=AsYsds+dj=1[CjsYs+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)]dWjs,s[t,T],Yt=0, $ (3.6)
    $ {dZs=[AsZs+B1,sv11s[t,t+ϵ)+B2,sv21s[t,t+ϵ)]ds+dj=1CjsZsdWjs,s[t,T],Zt=0. $ (3.7)

    Moreover, by Theorem 4.4 in [24], we have

    $ Et[sups[t,T)|Ys|2]=O(ϵ),Et[sups[t,T)|Zs|2]=O(ϵ2). $ (3.8)

    With $ A $ being deterministic, it follows from the dynamics of $ Y $ that, for any $ s\in[t, T] $, we have

    $ Et[Ys]=stEt[AsYτ]dτ=stAsEt[Yτ]dτ. $ (3.9)

    Hence we conclude that

    $ Et[Ys]=0s[t,T]. $ (3.10)

    By these estimates, we can calculate

    $ Ji(t,Xt;ut,ϵ,v11,ut,ϵ,v22)Ji(t,Xt;u1,u2)=12EtTt[<Qi,s(2Xs+Ys+Zs),Ys+Zs>+<Ri,s(2ui+vi),vi>1s[t,t+ϵ)]ds+Et[<GiXT,YT+ZT>]+12Et[<Gi(YT+ZT),YT+ZT>]<hiEt[XT]+λiXt+μi,Et[YT+ZT]>12<hiEt[YT+ZT],Et[YT+ZT]>=12EtTt[<Qi,s(2Xs+Ys+Zs),Ys+Zs>+<Ri,s(2ui+vi),vi>1s[t,t+ϵ)]ds+Et[<GiXThiEt[XT]λiXtμi,YT+ZT>+12<Gi(YT+ZT),YT+ZT>]+o(ϵ). $ (3.11)

    Recalling that $ (p_i(\cdot; t), k_i(\cdot; t)) $ and $ (P_i(\cdot; t), K_i(\cdot; t)) $ solve, respectively, BSDEs (3.1) and (3.2) for $ i = 1, 2 $, we have

    $ Et[<GiXThiEt[XT]λiXtμi,YT+ZT>]=Et[<pi(T;t),YT+ZT>]=Et[Ttd<pi(s;t),Ys+Zs>]=EtTt[<pi(s;t),As(Ys+Zs)+B1,sv11s[t,t+ϵ)+B2,sv21s[t,t+ϵ)><Aspi(s;t)+dj=1(Cjs)kji(s;t)+Qi,sXs,Ys+Zs>+dj=1<kji(s;t),Cjs(Ys+Zs)+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)>]ds=EtTt[<Qi,sXs>+B1,spi(s;t)+dj=1(Dj1,s)kji(s;t),v11s[t,t+ϵ)+B2,spi(s;t)+dj=1(Dj2,s)kji(s;t),v21s[t,t+ϵ)]ds $ (3.12)

    and

    $ Et[12<Gi(YT+ZT),YT+ZT>]=Et[12<Pi(T;t)(YT+ZT),YT+ZT>]=Et[Ttd<Pi(s;t)(Ys+Zs),Ys+Zs>]=EtTt{<Pi(s;t)(Ys+Zs),As(Ys+Zs)+B1,sv11s[t,t+ϵ)+B2,sv21s[t,t+ϵ)>+<Pi(s;t)[As(Ys+Zs)+B1,sv11s[t,t+ϵ)+B2,sv21s[t,t+ϵ)],Ys+Zs><[AsPi(s;t)+Pi(s;t)As+Qi,s+dj=1((Cjs)Pi(s;t)Cjs+(Cjs)Kji(s;t)+Kji(s;t)Cjs)](Ys+Zs),Ys+Zs>+dj=1<Kji(s;t)(Ys+Zs),Cjs(Ys+Zs)+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)>+dj=1<Kji(s;t)[Cjs(Ys+Zs)+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)],Ys+Zs>+dj=1<Pi(s;t)[Cjs(Ys+Zs)+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)],Cjs(Ys+Zs)+Dj1,sv11s[t,t+ϵ)+Dj2,sv21s[t,t+ϵ)>}ds=EtTt[<Qi,s(Ys+Zs),Ys+Zs>+dj=1<Pi(s;t)[Dj1,sv1+Dj2,sv2],Dj1,sv1+Dj2,sv2>1s[t,t+ϵ)]ds+o(ϵ) $ (3.13)

    Combining (3.11)-(3.13), we have

    $ Ji(t,Xt;ut,ϵ,v11,ut,ϵ,v22)Ji(t,Xt;u1,u2)=EtTt[12<Ri,s(2ui+vi),vi>1s[t,t+ϵ)+B1,spi(s;t)+dj=1(Dj1,s)kji(s;t),v11s[t,t+ϵ)+B2,spi(s;t)+dj=1(Dj2,s)kji(s;t),v21s[t,t+ϵ)+12dj=1<Pi(s;t)[Dj1,sv1+Dj2,sv2],Dj1,sv1+Dj2,sv2>1s[t,t+ϵ)]ds+o(ϵ). $ (3.14)

    Take $ i = 1 $, we let $ v_2 = 0 $, then $ u_2^{t, \epsilon, v_2} = u_2^* $, from (3.14), we obtain

    $ J1(t,Xt;ut,ϵ,v11,u2)J1(t,Xt;u1,u2)=EtTt{R1,su1+B1,sp1(s;t)+dj=1(Dj1,s)kj1(s;t),v11s[t,t+ϵ)+12[R1,s+dj=1(Dj1,s)P1(s;t)Dj1,s]v1,v1}ds=Ett+ϵt{<Λ1(s;t),v1>+12<H1(s;t)v1,v1>}ds+o(ϵ). $ (3.15)

    This proves (3.3), and similarly, we obtain (3.4).

    Because of $ R_{i, s} $ and $ P_i(s; t), i = 1, 2 $ are non-negative definite, $ H_i(s; t), \ i = 1, 2 $ are also non-negative definite. In view of (3.3)-(3.4), a sufficient condition for an equilibrium is

    $ EtTt|Λi(s;t)|ds<+,limstEt[Λi(s;t)]=0 a.s. t[0,T],i=1,2. $ (3.16)

    By an arguments similar to the proof of Proposition 3.3 in [21], we have the following lemma:

    Lemma 3.1. For any triple of state and control processes $ (X^*, u_1^*, u_2^*) $, the solution to BSDE (3.1) in $ L^2(0, T;\mathbb{R}^n)\times (L^2(0, T;\mathbb{R}^n))^d $ satisfies $ k_i(s; t_1) = k_i(s; t_2) $ for a.e. $ s\ge\max\{t_1, t_2\}, \; i = 1, 2 $. Furthermore, there exist $ \rho_i\in L^2(0, T;\mathbb{R}^l) $, $ \delta_i\in L^2(0, T;\mathbb{R}^{l\times n}) $ and $ \xi_i\in L^2(\Omega; C(0, T;\mathbb{R}^n)) $, such that

    $ Λi(s;t)=ρi(s)+δi(s)ξi(t),i=1,2. $ (3.17)

    Therefore, we have another characterization for equilibrium strategies:

    Proposition 2. Given a strategy pair $ (u_1^*, u_2^*)\in L^2(0, T;\mathbb{R}^l)\times L^2(0, T;\mathbb{R}^l) $. Denote $ X^* $ as the state process, and $ (p_i(\cdot; t), (k_i^j(\cdot; t)_{j = 1, 2, ..., d}))\in L_{\mathcal{F}}^2(t, T;\mathbb{R}^n)\times(L_{\mathcal{F}}^2(t, T;\mathbb{R}^n))^d $ as the unique solution for the BSDE (3.1), with $ k_i(s) = k_i(s; t) $ according to Lemma 3.1 for $ i = 1, 2 $ respectively. For $ i = 1, 2 $, letting

    $ Λi(s,t)=Bi,spi(s;t)+dj=1(Dj,s)k(s;t)j+Ri,sui,s,s[t,T], $ (3.18)

    then $ u^* $ is an equilibrium strategy if and only if

    $ Λi(t,t)=0,a.s.,a.e.t[0,T],i=1,2. $ (3.19)

    Proof. From (3.17), we have $ \Lambda_1(s; t) = \rho_1(s)+\delta_1(s)\xi_1(t) $. Since $ \delta_1 $ is essentially bounded and $ \xi_1 $ is continuous, we have

    $ \lim\limits_{\epsilon\downarrow0}\mathbb{E}_t\left[{1\over\epsilon}\int_t^{t+\epsilon}|\delta_1(s)(\xi_1(s)-\xi_1(t))|ds\right] \le c\lim\limits_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[|\xi_1(s)-\xi_1(t)|]ds = 0, $

    and hence

    $ \lim\limits_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[\Lambda_1(s;t)]ds = \lim\limits_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[\Lambda_1(s;s)]ds. $

    Therefore, if (3.19) holds, we have

    $ \lim\limits_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[\Lambda_1(s;t)]ds = \lim\limits_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[\Lambda_1(s;s)]ds = 0. $

    When $ i = 2 $, we can prove (3.19) similarly.

    Conversely, if (3.16) holds, then $ \lim_{\epsilon\downarrow0}{1\over\epsilon}\int_t^{t+\epsilon}\mathbb{E}_t[\Lambda_i(s; s)]ds = 0, i = 1, 2 $ leading to (3.19) by virtue of Lemma 3.4 of [21].

    The following is the main general result for the stochastic LQ differential game with time-inconsistency.

    Theorem 3.2. A strategy pair $ (u_1^*, u_2^*)\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^l)\times L_{\mathcal{F}}^2(0, T;\mathbb{R}^l) $ is an equilibrium strategy pair if the following two conditions hold for any time $ t $:

    (i) The system of SDEs

    $ {dXs=[AsXs+B1,su1,s+B2,su2,s+bs]ds+dj=1[CjsXs+Dj1,su1,s+Dj2,su2,s+σjs]dWjs,X0=x0,dp1(s;t)=[Asp1(s;t)+dj=1(Cjs)kj1(s;t)+Q1,sXs]ds+dj=1kj1(s;t)dWjs,s[t,T],p1(T;t)=G1XTh1Et[XT]λ1Xtμ1,dp2(s;t)=[Asp2(s;t)+dj=1(Cjs)kj2(s;t)+Q2,sXs]ds+dj=1kj2(s;t)dWjs,s[t,T],p2(T;t)=G2XTh2Et[XT]λ2Xtμ2, $ (3.20)

    admits a solution $ (X^*, p_1, k_1, p_2, k_2) $;

    (ii) $ \Lambda_i(s; t) = R_{i, s}u_{i, s}^*+B_{i, s}p_i(s; t)+\sum_{j = 1}^d(D_{i, s}^j)'k_i^j(s; t), i = 1, 2 $ satisfy condition (3.19).

    Proof. Given a strategy pair $ (u_1^*, u_2^*)\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^l)\times L_{\mathcal{F}}^2(0, T;\mathbb{R}^l) $ satisfying (i) and (ii), then for any $ v_1, v_2\in L_{\mathcal{F}_t}^2(\Omega, \mathbb{R}^l) $, define $ \Lambda_i, H_i, i = 1, 2 $ as in Proposition 1. We have

    $ limϵ0J1(t,Xt;ut,ϵ,v11,u2)J1(t,Xt;u1,u2)ϵ=limϵ0Ett+ϵt{<Λ1(s;t),v1>+12<H1(s;t)v1,v1>}dsϵlimϵ0Ett+ϵt<Λ1(s;t),v1>dsϵ=0, $ (3.21)

    proving the first condition of Definition 2.1, and the proof of the second condition is similar.

    Theorem 3.2 involve the existence of solutions to a flow of FBSDEs along with other conditions. The system (3.20) is more complicated than system (3.6) in [20]. As declared in [20], "proving the general existence for this type of FBSEs remains an outstanding open problem", it is also true for our system (3.20).

    In the rest of this paper, we will focus on the case when $ n = 1 $. When $ n = 1 $, the state process $ X $ is a scalar-valued rocess evolving by the dynamics

    $ dX_s = [A_sX_s+B_{1,s}'u_{1,s}+B_{2,s}'u_{2,s}+b_s]ds+[C_sX_s+D_{1,s}u_{1,s}+D_{2,s}u_{2,s}+\sigma_s]'dW_s, \\ X_0 = x_0, $ (3.22)

    where $ A $ is a bounded deterministic scalar function on $ [0, T] $. The other parameters $ B, C, D $ are all essentially bounded and $ \mathcal{F}_t $-adapted processes on $ [0, T] $ with values in $ \mathbb{R}^l, \mathbb{R}^d, \mathbb{R}^{d\times l} $, respectively. Moreover, $ b\in L_{\mathcal{F}}^2(0, T;\mathbb{R}) $ and $ \sigma\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^d) $.

    In this case, the adjoint equations for the equilibrium strategy become

    $ {dpi(s;t)=[Aspi(s;t)+(Cs)ki(s;t)+Qi,sXs]ds+ki(s;t)dWs,s[t,T],pi(T;t)=GiXThiEt[XT]λiXtμi, $ (3.23)
    $ {dPi(s;t)=[(2As+|Cs|2)Pi(s;t)+2CsK(s;t)+Qi,s]ds+Ki(s;t)dWs,s[t,T],Pi(T;t)=Gi, $ (3.24)

    for $ i = 1, 2 $. For convenience, we also state here the $ n = 1 $ version of Theorem 3.2:

    Theorem 3.3. A strategy pair $ (u_1^*, u_2^*)\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^l)\times L_{\mathcal{F}}^2(0, T;\mathbb{R}^l) $ is an equilibrium strategy pair if, for any time $ t\in[0, T) $,

    (i) The system of SDEs

    $ {dXs=[AsXs+B1,su1,s+B2,su2,s+bs]ds+[CsXs+D1,su1,s+D2,su2,s+σs]dWs,X0=x0,dp1(s;t)=[Asp1(s;t)+(Cs)k1(s;t)+Q1,sXs]ds+k1(s;t)dWs,s[t,T],p1(T;t)=G1XTh1Et[XT]λ1Xtμ1,dp2(s;t)=[Asp2(s;t)+(Cs)k2(s;t)+Q2,sXs]ds+k2(s;t)dWs,s[t,T],p2(T;t)=G2XTh2Et[XT]λ2Xtμ2, $ (3.25)

    admits a solution $ (X^*, p_1, k_1, p_2, k_2) $;

    (ii) $ \Lambda_i(s; t) = R_{i, s}u_{i, s}^*+B_{i, s}p_i(s; t)+(D_{i, s})'k_i(s; t), i = 1, 2 $ satisfy condition (3.19).

    The unique solvability of (3.25) remains a challenging open problem even for the case $ n = 1 $. However, we are able to solve this problem when the parameters $ A, B_1, B_2, C, D_1, D_2, b, \sigma, Q_1, Q_2, R_1 $ and $ R_2 $ are all deterministic functions.

    Throughout this section we assume all the parameters are deterministic functions of $ t $. In this case, since $ G_1, G_2 $ have been also assumed to be deterministic, the BSDEs (3.24) turns out to be ODEs with solutions $ K_i\equiv0 $ and $ P_i(s; t) = G_ie^{\int_s^T(2A_u+|C_u|^2)du}+\int_s^Te^{\int_s^T(2A_u+|C_u|^2)du}Q_{i, v}dv $ for $ i = 1, 2 $. Hence, the equilibrium strategy will be characterized through a system of coupled Riccati-type equations.

    As in classical LQ control, we attempt to look for a linear feedback equilibrium strategy pair. For such purpose, motivated by [20], given any $ t\in[0, T] $, we consider the following process:

    $ pi(s;t)=Mi,sXsNi,sEt[Xs]Γi,sXt+Φi,s,0tsT,i=1,2, $ (4.1)

    where $ M_i, N_i, \Gamma_i, \Phi_i $ are deterministic differentiable functions with $ \dot{M}_i = m_i, \dot{N}_i = n_i, \dot{\Gamma}_i = \gamma_i $ and $ \dot\Phi_i = \phi_i $ for $ i = 1, 2 $. The advantage of this process is to separate the variables $ X_s^*, \mathbb{E}_t[X_s^*] $ and $ X_t^* $ in the solutions $ p_i(s; t), i = 1, 2 $, thereby reducing the complicated FBSDEs to some ODEs.

    For any fixed $ t $, applying Ito's formula to (4.1) in the time variable $ s $, we obtain, for $ i = 1, 2 $,

    $ dp_i(s;t) = \{M_{i,s}(A_sX_s^*+B_{1,s}'u_{1,s}^*+B_{2,s}'u_{2,s}^*+b_s)\\+m_{i,s}X_s^*-N_{i,s}\mathbb{E}_t[A_sX_s^*+B_{1,s}'u_{1,s}^*+B_{2,s}'u_{2,s}^*+b_s] \\ \qquad -n_{i,s}\mathbb{E}_t[X_s^*]-\gamma_{i,s}X_t^*+\phi_{i,s}\}ds+M_{i,s}(C_sX_s^*+D_{1,s}u_{1,s}^*+D_{2,s}u_{2,s}^*+\sigma_s)'dW_s. $ (4.2)

    Comparing the $ dW_s $ term of $ dp_i(s; t) $ in (3.25) and (4.2), we have

    $ ki(s;t)=Mi,s[CsXs+D1,su1,s+D2,su2,s+σs],s[t,T],i=1,2. $ (4.3)

    Notice that $ k(s; t) $ turns out to be independent of $ t $.

    Putting the above expressions (4.1) and (4.3) of $ p_i(s; t) $ and $ k_i(s; t), i = 1, 2 $ into (3.19), we have

    $ R_{i,s}u_{i,s}^*+B_{i,s}[(M_{i,s}-N_{i,s}-\Gamma_{i,s})X_s^*+\Phi_{i,s}]+D_{i,s}'M_{i,s}[C_sX_s^*+D_{1,s}u_{1,s}^*+D_{2,s}u_{2,s}^*+\sigma_s]\\ = 0, \;s\in[0,T], $ (4.4)

    for $ i = 1, 2 $. Then we can formally deduce

    $ ui,s=αi,sXs+βi,s,i=1,2. $ (4.5)

    Let $ M_s = {{{\text{diag}}}}(M_{1, s}I_l, M_{2, s}I_l), N_s = {{{\text{diag}}}}(N_{1, s}I_l, N_{2, s}I_l), \Gamma_s = {{{\text{diag}}}}(\Gamma_{1, s}I_l, \Gamma_{2, s}I_l), \Phi_s = {{{\text{diag}}}}(\Phi_{1, s}I_l, \Phi_{2, s}I_l) $, $ R_s = {{{\text{diag}}}}(R_{1, s}, R_{2, s}), B_s = \left(B1,sB2,s\right), D_s = \left(D1,s,D2,s\right) $, $ u_s^* = \left(u1,su2,s\right), \alpha_s = \left(α1,sα2,s\right) $ and $ \beta_s = \left(β1,sβ2,s\right) $. Then from (4.4), we have

    $ Rsus+[(MsNsΓs)Xs+Φs]Bs+MsDs[CsXs+Ds(αsXs+βs)+σs]=0,s[0,T] $ (4.6)

    and hence

    $ αs=(Rs+MsDsDs)1[(MsNsΓs)Bs+MsDsCs], $ (4.7)
    $ βs=(Rs+MsDsDs)1(ΦsBs+MsDsσs). $ (4.8)

    Next, comparing the $ ds $ term of $ dp_i(s; t) $ in (3.25) and (4.2) (we supress the argument $ s $ here), we have

    $ M_i[AX^*+B'(\alpha X^*+\beta)+b]+m_iX^*-N_i\{A\mathbb{E}_t[X^*]+\\ B'\mathbb{E}_t[\alpha X^*+\beta]+b\} -n_i\mathbb{E}_t[X^*]-\gamma_iX_t^*+\phi_i \\ = -[A(M_iX^*-N_i\mathbb{E}_t[X^*]-\Gamma_iX_t^*+\Phi_i)+M_iC'(CX^*+D(\alpha X^*+\beta)+\sigma)]. $ (4.9)

    Notice in the above that $ X^* = X_s^* $ and $ \mathbb{E}_t[X^*] = \mathbb{E}_t[X_s^*] $ due to the omission of $ s $. This leads to the following equations for $ M_i, N_i, \Gamma_i, \Phi_i $:

    $ \left\{ ˙Mi=(2A+|C|2)MiQi+Mi(B+CD)(R+MDD)1[(MNΓ)B+MDC],s[0,T],Mi,T=Gi; \right. $ (4.10)
    $ {˙Ni=2ANi+NiB(R+MDD)1[(MNΓ)B+MDC],s[0,T],Ni,T=hi; $ (4.11)
    $ {˙Γi=AΓi,s[0,T],Γi,T=λi; $ (4.12)
    $ {˙Φi={A[B(MN)+CDM](R+MDD)1B}Φi(MiNi)bMiCσ[(MiNi)B+MiCD](R+MDD)1MDσ,s[0,T],Φi,T=μi. $ (4.13)

    Though $ M_i, N_i, \Gamma_i, \Phi_i, i = 1, 2 $ are scalars, $ M, N, \Gamma, \Phi $ are now matrices because of two players. Therefore, the above equations are more complicated than the similar equations (4.5)–(4.8) in [20]. Before we solve the equations (4.10)–(4.13), we first prove that, if exist, the equilibrium constructed above is the unique equilibrium. Indeed, we have

    Theorem 4.1. Let

    $ L1={X(;):X(;t)L2F(t,T;R),supt[0,T]E[supst|X(s;t)|2]<+} $ (4.14)

    and

    $ L2={Y(;):Y(;t)L2F(t,T;Rd),supt[0,T]E[Tt|X(s;t)|2ds]<+}. $ (4.15)

    Suppose all the parameters $ A, B_1, B_2, C, D_1, D_2, b, \sigma, Q_1, Q_2, R_1 $ and $ R_2 $ are all deterministic.

    When $ (M_i, N_i, \Gamma_i, \Phi_i), i = 1, 2 $ exist, and for $ i = 1, 2 $, $ (p_i(s; t), k_i(s; t))\in\mathcal{L}_1\times \mathcal{L}_2 $, the equilibrium strategy is unique.

    Proof. Suppose there is another equilibrium $ (X, u_1, u_2) $, then the BSDE (3.1), with $ X^* $ replaced by $ X $, admits a solution $ (p_i(s; t), k_i(s), u_{i, s}) $ for $ i = 1, 2 $, which satisfies $ B_{i, s}p_i(s; s)+D_{i, s}'k_i(s)+R_{i, s}u_{i, s} = 0 $ for a.e. $ s\in[0, T] $. For $ i = 1, 2 $, define

    $ ˉpi(s;t)pi(s;t)[Mi,sXsNi,sEt[Xs]Γi,s+Φi,s], $ (4.16)
    $ ˉki(s;t)ki(s)Mi,s(CsXs+D1,su1,s+D2,su2,s+σs), $ (4.17)

    where $ k_i(s) = k_i(s; t) $ by Lemma 3.1.

    We define $ p(s; t) = {{{\text{diag}}}}(p_1(s; t)I_l, p_2(s; t)I_l) $, $ \bar{p}(s; t) = {{{\text{diag}}}}(\bar{p}_1(s; t)I_l, \bar{p}_2(s; t)I_l) $, and $ u = \left(u1,su2,s\right) $. By the equilibrium condition (3.19), we have

    $ 0 = \left(B1,sp1(s;s)+D1,sk1(s)+R1,su1,sB2,sp2(s;s)+D2,sk2(s)+R2,su2,s\right) \\ = p(s;s)B_s+\left(D1,sk1(s)D2,sk2(s)\right)+R_su_s \\ = [\bar{p}(s;s)+X_s(M_s-N_s-\Gamma_s)+\Phi_s]B_s+\left(D1,sˉk1(s)D2,sˉk2(s)\right) +\\M_sD_s'(C_sX_s+D_su_s+\sigma_s)+R_su_s \\ = \bar{p}(s;s)B_s+\left(D1,sˉk1(s)D2,sˉk2(s)\right) +X_s[(M_s-N_s-\Gamma_s)B_s+M_sD_s'C_s]+\Phi_sB_s+M_sD_s'\sigma_s \\ +(R_s+M_sD_s'D_s)u_s. $ (4.18)

    Since $ R_s+M_sD_s'D_s $ is invertible, we have

    $ us=(Rs+MsDsDs)1{ˉp(s;s)Bs+(D1,sˉk1(s)D2,sˉk2(s))+Xs[(MsNsΓs)Bs+MsDsCs]+ΦsBs+MsDsσs}, $ (4.19)

    and hence for $ i = 1, 2 $,

    $ d\bar{p}_i(s;t) = dp_i(s;t)-d[M_{i,s}X_s-N_{i,s}\mathbb{E}_t[X_s]-\Gamma_{i,s}+\Phi_{i,s}] \\ = -[A_sp_i(s;t)+C_{s}'k_i(s)+Q_{i,s}X_s]ds+k_i'(s)dW_s-d[M_{i,s}X_s-N_{i,s}\mathbb{E}_t[X_s]-\Gamma_{i,s}X_t+\Phi_{i,s}] \\ = -\bigg\{A_s\bar{p}_i(s;t)+C_s'\bar{k}_i(s)+A_s(M_{i,s}X_s-N_{i,s}\mathbb{E}_t[X_s]-\Gamma_{i,s}X_t+\Phi_{i,s}) \\ \ +C_s'M_{i,s}(C_sX_s+D_{1,s}u_{1,s}+D_{2,s}u_{2,s}+\sigma_s)\bigg\}ds \\ \ +[\bar{k}_i(s)-M_{i,s}(C_sX_s+D_{1,s}u_{1,s}+D_{2,s}u_{2,s}+\sigma_s)]'dW_s \\ -\bigg\{M_{i,s}[A_sX_s+B_s'u_s+b_s]+m_{i,s}X_s-N_{i,s}(A_s\mathbb{E}_t[X_s]+B_s'\mathbb{E}_t[u_s]+b_s) \\ \ -n_{i,s}\mathbb{E}_t[X_s]-\gamma_{i,s}X_t+\phi_{i,s}\bigg\}ds \\ -M_{i,s}[C_sX_s+D_su_s+\sigma_s]'dW_s \\ = -\Bigg\{A_s\bar{p}_i(s;t)+C_s'\bar{k}_i(s)-M_{i,s}(B_s'+C_s'D_s)(R_s+M_sD_s'D_s)^{-1}\left[B_s\bar{p}(s;s)+\left(D1,sˉk1(s)D2,sˉk2(s)\right)\right] \\ \qquad N_{i,s}B_s'(R_s+M_sD_s'D_s)^{-1}\mathbb{E}_t\left[B_s\bar{p}(s;s)+\left(D1,sˉk1(s)D2,sˉk2(s)\right)\right]\Bigg\}ds+\bar{k}_i(s)'dW_s, $ (4.20)

    where we suppress the subscript $ s $ for the parameters, and we have used the equations (4.10)–(4.13) for $ M_i, N_i, \Gamma_i, \Phi_i $ in the last equality. From (4.16) and (4.17), we have $ (\bar{p}_i, \bar{k}_i)\in\mathcal{L}_1\times\mathcal{L}_2 $. Therefore, by Theorem 4.2 of [21], we obtain $ \bar{p}(s; t)\equiv0 $ and $ \bar{k}(s)\equiv0 $.

    Finally, plugging $ \bar{p}\equiv\bar{k}\equiv0 $ into $ u $ of (4.19), we get $ u $ being the same form of feedback strategy as in (4.5), and hence $ (X, u_1, u_2) $ is the same as $ (X^*, u_1^*, u_2^*) $ which defined by (4.5) and (3.25).

    The solutions to (4.12) is

    $ Γi,s=λieTsAtdt, $ (4.21)

    for $ i = 1, 2 $. Let $ \tilde{N} = N_1/ N_2 $, from (4.11), we have $ \dot{\tilde{N}} = 0 $, and hence

    $ ˜Nh1h2,N2h2h1N1. $ (4.22)

    Equations (4.10) and (4.11) form a system of coupled Riccati-type equations for $ (M_1, M_2, N_1) $:

    $ {˙M1=[2A+|C|2+BΓ(R+MDD)1(B+DC)]M1Q1+(B+DC)(R+MDD)1M(B+DC)M1BN(R+MDD)1(B+DC)M1,M1,T=G1;˙M2=[2A+|C|2+BΓ(R+MDD)1(B+DC)]M2Q2+(B+DC)(R+MDD)1M(B+DC)M2BN(R+MDD)1(B+DC)M2,M2,T=G2;˙N1=2ANi+NiB(R+MDD)1[(MNΓ)B+MDC],N1,T=h1. $ (4.23)

    Finally, once we get the solution for $ (M_1, M_2, N_1) $, (4.13) is a simple ODE. Therefore, it is crucial to solve (4.23).

    Formally, we define $ \tilde{M} = {M_1\over M_2} $ and $ J_1 = {M_1\over N_1} $ and study the following equation for $ (M_1, \tilde{M}, J_1) $:

    $ {˙M1=[2A+|C|2+BΓ(R+MDD)1(B+DC)]M1Q1+(B+DC)(R+MDD)1M(B+DC)M1BN(R+MDD)1(B+DC)M1,M1,T=G1;˙˜M=(Q1M1Q2M1˜M)˜M,˜MT=G1G2;˙J1=[|C|2CD(R+MDD)1M(B+DC)+BΓ(R+MDD)1DC+Q1M1]J1CD(R+MDD)1Mdiag(Il,h2h1˜MIl)B,J1,T=G1h1, $ (4.24)

    where $ M = {{{\text{diag}}}}(M_1I_l, {M_1\over\tilde M}I_l), N = {{{\text{diag}}}}({M_1\over J_1}I_l, {h_2\over h_1}{M_1\over J_1}I_l) $ and $ \Gamma = {{{\text{diag}}}}(\lambda_1e^{\int_s^T A_tdt}I_l, \lambda_2e^{\int_s^T A_tdt}I_l) $.

    By a direct calculation, we have

    Proposition 3. If the system (4.24) admits a positive solution $ (M_1, \tilde{M}, J_1) $, then the system (4.23) admits a solution $ (M_1, M_2, N_1) $.

    In the following, we will use the truncation method to study the system (4.24). For convenienc, we use the following notations:

    $ ab=max{a,b},a,bR, $ (4.25)
    $ ab=min{a,b},a,bR. $ (4.26)

    Moreover, for a matrix $ M\in\mathbb{R}^{m\times n} $ and a real number $ c $, we define

    $ (Mc)i,j=Mi,jc,1im,1jn, $ (4.27)
    $ (Mc)i,j=Mi,jc,1im,1jn. $ (4.28)

    We first consider the standard case where $ R-\delta{I}\succeq0 $ for some $ \delta > 0 $. We have

    Theorem 4.2. Assume that $ R-\delta{I}\succeq0 $ for some $ \delta > 0 $ and $ G\ge h > 0 $. Then (4.24), and hence (4.23) admit unique solution if

    (i) there exists a constant $ \lambda\ge0 $ such that $ B = \lambda D'C $;

    (ii) $ \frac{|C|^2}{2l}D'D-(\lambda+1)D'CC'D\succeq0 $.

    Proof. For fixed $ c > 0 $ and $ K > 0 $, consider the following truncated system of (4.24):

    $ {˙M1=[2A+|C|2+BΓ(R+M+cDD)1(B+DC)]M1Q1+(B+DC)(R+M+cDD)1(M+cK)(B+DC)M1B(N+cK)(R+M+cDD)1(B+DC)M1,M1,T=G1;˙˜M=(Q1M1cQ2M1c˜MK)˜M,˜MT=G1G2;˙J1=λ(1)J1CD(R+M+cDD)1(M+cK)diag(Il,h2h1(˜MK)Il)B,J1,T=G1h1, $ (4.29)

    where $ M_{c}^+ = {{{\text{diag}}}}((M_1\vee0)I_l, {{M_1\vee0}\over\tilde M\vee c}I_l) $, $ N_c^+ = {{{\text{diag}}}}({{M_1\vee0}\over J_1\vee c}I_l, {h_2\over h_1}{{M_1\vee0}\over J_1\vee c}I_l) $ and

    $ λ(1)=|C|2CD(R+M+cDD)1(M+cK)(B+DC)+BΓ(R+M+cDD)1DC+Q1M1c. $ (4.30)

    Since $ R-\delta I\succeq0 $, the above system (4.29) is locally Lipschitz with linear growth, and hence it admits a unique solution $ (M_1^{c, K}, \tilde{M}^{c, K}, J_1^{c, K}) $. We will omit the superscript $ (c, K) $ when there is no confusion.

    We are going to prove that $ J_1\ge1 $ and that $ M_1, \tilde{M}\in[L_1, L_2] $ for some $ L_1, L_2 > 0 $ independent of $ c $ and $ K $ appearing in the truncation functions. We denote

    $ λ(2)=(2A+|C|2+BΓ(R+M+cDD)1(B+DC))(B+DC)(R+M+cDD)1(M+cK)(B+DC)B(N+cK)(R+M+cDD)1(B+DC). $ (4.31)

    Then $ \lambda^{(2)} $ is bounded, and $ M_1 $ satisfies

    $ ˙M1+λ(2)M1+Q1=0,M1,T=G1. $ (4.32)

    Hence $ M_1 > 0 $. Similarly, we have $ \tilde{M} > 0 $.

    The equation for $ \tilde{M} $ is

    $ {˙˜M=(Q1M1c˜MQ2M1c(˜MK)˜M,˜MT=G1G2; $ (4.33)

    hence $ \tilde{M} $ admits an upper bound $ L_2 $ independent of $ c $ and $ K $. Choosing $ K = L_2 $ and examining again (4.33), we deduce that there exists $ L_1 > 0 $ independent of $ c $ and $ K $ such that $ \tilde{M}\ge L_1 $. Indeed, we can choose $ L_1 = \min_{0\le t\le T}{Q_{1, t}\over Q_{2, t}}\wedge{G1\over G_2} $ and $ L_2 = \max_{0\le t\le T}{Q_{1, t}\over Q_{2, t}}\vee{G1\over G_2} $. As a result, choosing $ c < L_1 $, the terms $ M_c^+ $ can be replaced by $ M = {{{\text{diag}}}}(M_1I_l, {M_1\over\tilde M}I_l) $, respectively, in (4.29) without changing their values.

    Now we prove $ J\ge1 $. Denote $ \tilde{J} = J_1-1 $, then $ \tilde{J} $ satisfies the ODE:

    $ ˙˜J=λ(1)˜J[λ(1)+CD(R+MDD)1(MK)diag(Il,h2h1˜MIl)B]=λ(1)˜Ja(1), $ (4.34)

    where

    $ a^{(1)} = \lambda^{(1)}+C'D(R+MD'D)^{-1}(M\wedge K){{{\text{diag}}}}(I_l,{h_2\over h_1}\tilde{M}I_l)B \\ = |C|^2-(\lambda+1)C'D(R+MD'D)^{-1}(M\wedge K)D'C+C'D\Gamma(R+MD'D)^{-1}(M\wedge K)D'C++{Q_1\over M_1\vee c} \\ +C'D(R+MD'D)^{-1}(M\wedge K){{{\text{diag}}}}(I_l,{h_2\over h_1}\tilde{M}I_l)D'C \\ \ge|C|^2-(\lambda+1)C'D(R+MD'D)^{-1}MD'C+C'D\Gamma(R+MD'D)^{-1}(M\wedge K)D'C++{Q_1\over M_1\vee c} \\ = tr\left\{(R+MD'D)^{-1}{|C|^2+Q_1/(M_1\vee c)\over2l}(R+MD'D)\right\}-(\lambda+1)tr\{(R+MD'D)^{-1}D'CC'DM\} \\ = tr\left\{(R+MD'D)^{-1}H\right\} $ (4.35)

    with $ H = {|C|^2+Q_1/(M_1\vee c)\over2l}(R+D'DM)-(\lambda+1)D'CC'DM $.

    When $ c $ is small enough such that $ R-cD'D\succeq0 $, we have

    $ Q1M1c(R+MDD)Q1L2DD. $ (4.36)

    Hence,

    $ H(|C|22lDD(λ+1)DCCD)M0, $ (4.37)

    and consequently $ a^{(1)}\ge tr\{(R+MD'D)^{-1}H\}\ge0 $. We then deduce that $ \tilde{J}\ge0 $, and hence $ J_1\ge1 $. The boundness of $ M_1 $ can be proved by a similar argument in the proof of Theorem 4.2 in [20].

    Similarly, for the singular case $ R\equiv0 $, we have

    Theorem 4.3. Given $ G_1\ge h_1\ge1, R\equiv0 $, if $ B = \lambda D'C $ and $ |C|^2-(\lambda+1)C'D(D'D)^{-1}D'C\ge0 $, then (4.24) and (4.23) admit a unique positive solution.

    Concluding the above two theorems, we can present our main results of this section:

    Theorem 4.4. Given $ G_1\ge h_1\ge1 $ and $ B = \lambda D'C $. The (4.23) admits a unique positive solution $ (M_1, M_2, N_1) $ in the following two cases:

    (i) $ R-\delta I\succeq0 $ for some $ \delta > 0 $, $ \frac{|C|^2}{2l}D'D-(\lambda+1)D'CC'D\succeq0 $;

    (ii) $ R\equiv0 $, $ |C|^2-(\lambda+1)C'D(D'D)^{-1}D'C\ge0 $.

    Proof. Define $ p_i(s; t) $ and $ k_i(s; t) $ by (4.1) and (4.3), respectively. It is straightforward to check that $ (u_1^*, u_2^*, X^*, p_1, p_2, k_1, k_2) $ satisfies the system of SDEs (3.25). Moreover, in the both cases, we can check that $ \alpha_{i, s} $ and $ \beta_{i, s} $ in (4.5) are all uniformly bounded, and hence $ u_i^*\in L_{\mathcal{F}}^2(0, T;\mathbb{R}^l) $ and $ X^*\in L^2(\Omega; C(0, T;\mathbb{R})) $.

    Finally, denote $ \Lambda_i(s; t) = R_{i, s}u_{i, s}^*+p_i(s; t)B_{i, s}+(D_{i, s})'k_i(s; t), i = 1, 2 $. Plugging $ p_i, k_i, u_i^* $ define in (4.1), (4.3) and (4.5) into $ \Lambda_i $, we have

    $ \Lambda_i(s;t) = R_{i,s}u_{i,s}^*+(M_{i,s}X_s^*-N_{i,s}\mathbb{E}_t[X_s^*]-\Gamma_{i,s}X_t^*+\Phi_{i,s})B_{i,s} +\\ M_{i,s}D_{i,s}'[C_sX_s^*+D_{1,s}u_{1,s}^*+D_{2,s}u_{2,s}^*+\sigma_s] $ (4.38)

    and hence,

    $ \Lambda(t;t) \triangleq \left(Λ1(t;t)Λ2(t;t)\right) \\ = (R_t+M_tD_t'D_t)u_t^*+M_t(B_t+D_t'C_t)X_t^* -N_tB_t\mathbb{E}_t[X_t^*]-\Gamma_tB_tX_t^*+(\Phi_tB_t+M_tD_t'\sigma_t) \\ = -[(M_t-N_t-\Gamma_t)B_t+M_tD_t'C_t]X_t^*-(\Phi_tB_t+M_tD_t'\sigma_t) \\ \quad+M_t(B_t+D_t'C_t)X_t^*-N_tB_tX_t^*-\Gamma_tB_tX_t^*+(\Phi_tB_t+M_tD_t'\sigma_t) \\ = 0. $ (4.39)

    Therefore, $ \Lambda_i $ satisfies the seond condition in (3.19).

    We investigate a general stochastic linear-quadratic differential game, where the objective functional of each player include both a quadratic term of the expected state and a state-dependent term. As discussed in detail in Björk and Murgoci [17] and [18], the last two terms in each objective functional, respectively, introduce two sources of time inconsistency into the differential game problem. That is to say, the usual equilibrium aspect is not a proper way when the players at 0 cannot commit the players at all intermediate times to implement the decisions they have planed. With the time-inconsistency, the notion "equilibrium" needs to be extended in an appropriate way. We turn to adopt the concept of equilibrium strategy between the players at all different times, which is at any time, an equilibrium "infinitesimally" via spike variation. By applying the spike variation technique, We derive a sufficient condition for equilibrium strategies via a system of forward-backward stochastic differential equation. The unique solvability of such FBSDEs remains a challenging open problem.

    For a special case, when the state is one-dimensional and the coefficients are all deterministic, the equilibrium strategy will be characterized through a system of coupled Riccati-type equations. At last, we find an explicit equilibrium strategy, which is also proved be the unique equilibrium strategy.

    The research of the first author was partially supported by NSFC (No.12171426), the Natural Science Foundation of Zhejiang Province (No. Y19A010020) and the Fundamental Research Funds for the Central Universities (No. 2021FZZX001-01). The research of the second author was partially supported by NSFC (No. 11501325, No.71871129) and the China Postdoctoral Science Foundation (Grant No. 2018T110706, No.2018M642641). The authors would like to thank sincerely the referees and the associate editor for their helpful comments and suggestions.

    The authors declare there is no conflicts of interest.



    Conflict of interest



    The authors declare no conflicts of interest.

    [1] Oparil S, Acelajado MC, Bakris GL, et al. (2018) Hypertension. Nat Rev Dis Primers 4: 18014. doi: 10.1038/nrdp.2018.14
    [2] Prieto I, Villarejo AB, Segarra AB, et al. (2014) Brain, heart and kidney correlate for the control of blood pressure and water balance: role of angiotensinases. Neuroendocrinology 100: 198–208. doi: 10.1159/000368835
    [3] Prieto I, Segarra AB, Martinez-Canamero M, et al. (2017) Bidirectional asymmetry in the neurovisceral communication for the cardiovascular control: New insights. Endocr Regul 51: 157–167. doi: 10.1515/enr-2017-0017
    [4] Prieto I, Segarra AB, de Gasparo M, et al. (2018) Divergent profile between hypothalamic and plasmatic aminopeptidase activities in WKY and SHR. Influence of beta-adrenergic blockade. Life Sci 192: 9–17.
    [5] Parati G, Ochoa JE, Lombardi C, et al. (2013) Assessment and management of blood-pressure variability. Nat Rev Cardiol 10: 143–155. doi: 10.1038/nrcardio.2013.1
    [6] Banegas I, Prieto I, Vives F, et al. (2006) Brain aminopeptidases and hypertension. J Renin Angiotensin Aldosterone Syst 7: 129–134. doi: 10.3317/jraas.2006.021
    [7] Banegas I, Prieto I, Vives F, et al. (2009) Asymmetrical response of aminopeptidase A and nitric oxide in plasma of normotensive and hypertensive rats with experimental hemiparkinsonism. Neuropharmacology 56: 573–579. doi: 10.1016/j.neuropharm.2008.10.011
    [8] Banegas I, Prieto I, Segarra AB, et al. (2011) Blood pressure increased dramatically in hypertensive rats after left hemisphere lesions with 6-hydroxydopamine. Neurosci Lett 500: 148–150. doi: 10.1016/j.neulet.2011.06.025
    [9] Banegas I, Prieto I, Segarra AB, et al. (2017) Bilateral distribution of enkephalinase activity in the medial prefrontal cortex differs between WKY and SHR rats unilaterally lesioned with 6-hydroxydopamine. Prog Neuropsychopharmacol Biol Psychiatry 75: 213–218. doi: 10.1016/j.pnpbp.2017.02.015
    [10] Prieto I, Segarra AB, Villarejo AB, et al. (2019) Neuropeptidase activity in the frontal cortex of Wistar-Kyoto and spontaneously hypertensive rats treated with vasoactive drugs: a bilateral study. J Hypertens 37: 612–628. doi: 10.1097/HJH.0000000000001884
    [11] Cuspidi C, Tadic M, Grassi G, et al. (2018) Mancia G. Treatment of hypertension: The ESH/ESC guidelines recommendations. Pharmacol Res 128: 315–321.
    [12] Ramírez-Sánchez M, Prieto I, Wangensteen R, et al. (2013) The renin-angiotensin system: new insight into old therapies. Curr Med Chem 20: 1313–1322. doi: 10.2174/0929867311320100008
    [13] Ferdinand KC, Balavoine F, Besse B, et al. (2019) Efficacy and safety of Firibastat, a first-in-class brain aminopeptidase A inhibitor, in hypertensive overweight patients of multiple ethnic origins: A Phase 2, open-label, multicenter, dose-titrating study. Circulation 140: 138–146. doi: 10.1161/CIRCULATIONAHA.119.040070
    [14] Gao J, Marc Y, Iturrioz X, et al. (2014) A new strategy for treating hypertension by blocking the activity of the brain renin-angiotensin system with aminopeptidase A inhibitors. Clin Sci (Lond) 127: 135–148. doi: 10.1042/CS20130396
    [15] Keck M, De Almeida H, Compère D, et al. (2019) NI956/QGC006, a potent orally active, brain-penetrating aminopeptidase a inhibitor for treating hypertension. Hypertension 73: 1300–1307. doi: 10.1161/HYPERTENSIONAHA.118.12499
    [16] Wright JW, Mizutani S, Murray CE, et al. (1990) Aminopeptidase-induced elevations and reductions in blood pressure in the spontaneously hypertensive rat. J Hypertens 8: 969–974. doi: 10.1097/00004872-199010000-00013
    [17] Wright JW, Jensen LL, Cushing LL, et al. (1989) Leucine aminopeptidase M-induced reductions in blood pressure in spontaneously hypertensive rats. Hypertension 13: 910–915. doi: 10.1161/01.HYP.13.6.910
    [18] Zini S, Masdehors P, Lenkei Z, et al. (1997) Aminopeptidase A: distribution in rat brain nuclei and increased activity in spontaneously hypertensive rats. Neuroscience 78: 1187–1193. doi: 10.1016/S0306-4522(96)00660-4
    [19] Banegas I, Prieto I, Segarra AB, et al. (2017) Study of the neuropeptide function in Parkinson's disease using the 6-Hydroxydopamine model of experimental Hemiparkinsonism. AIMS Neuroscience 4: 223–237. doi: 10.3934/Neuroscience.2017.4.223
    [20] Kondo K, Ebihara A, Suzuki H, et al. (1981) Role of dopamine in the regulation of blood pressure and the renin--angiotensin-aldosterone system in conscious rats. Clin Sci (Lond) 61: 235s–237s. doi: 10.1042/cs061235s
    [21] Zeng C, Jose PA (2011) Dopamine receptors: important antihypertensive counterbalance against hypertensive factors. Hypertension 57: 11–17. doi: 10.1161/HYPERTENSIONAHA.110.157727
    [22] Förstermann U, Sessa WC (2012) Nitric oxide synthases: regulation and function. Eur Heart J 33: 829–837. doi: 10.1093/eurheartj/ehr304
    [23] Calver A, Collier J, Vallance P (1993) Nitric oxide and the control of human vascular tone in health and disease. Eur J Med 2: 48–53.
    [24] Venturelli M, Pedrinolla A, Boscolo Galazzo I, et al. (2018) Impact of nitric oxide bioavailability on the progressive cerebral and peripheral circulatory impairments during aging and Alzheimer's disease. Front Physiol 9: 169. doi: 10.3389/fphys.2018.00169
    [25] Steinert JR, Chernova T, Forsythe ID (2010) Nitric oxide signaling in brain function, dysfunction, and dementia. Neuroscientist 16: 435–452. doi: 10.1177/1073858410366481
    [26] Maia-de-Oliveira JP, Trzesniak C, Oliveira IR, et al. (2012) Nitric oxide plasma/serum levels in patients with schizophrenia: a systematic review and meta-analysis. Braz J Psychiatry 34: S149–155.
    [27] Nakano Y, Yoshimura R, Nakano H, et al. (2010) Association between plasma nitric oxide metabolites levels and negative symptoms of schizophrenia: a pilot study. Hum Psychopharmacol 25: 139–144. doi: 10.1002/hup.1102
    [28] Shabeeh H, Khan S, Jiang B, et al. (2017) Blood pressure in healthy humans is regulated by neuronal no synthase. Hypertension 69: 970–976. doi: 10.1161/HYPERTENSIONAHA.116.08792
    [29] Chen ZQ, Mou RT, Feng DX, et al. (2017) The role of nitric oxide in stroke. Med Gas Res 7: 194–203. doi: 10.4103/2045-9912.215750
    [30] Zheng R, Qin L, Li S, et al. (2014) CT perfusion-derived mean transit time of cortical brain has a negative correlation with the plasma level of nitric oxide after subarachnoid hemorrhage. Acta Neurochir (Wien) 156: 527–533. doi: 10.1007/s00701-013-1968-6
    [31] Taffi R, Nanetti L, Mazzanti L, et al. (2008) Plasma levels of nitric oxide and stroke outcome. J Neurol 255: 94–98. doi: 10.1007/s00415-007-0700-y
    [32] Li S, Wang Y, Jiang Z, et al. (2018) Impaired cognitive performance in endothelial nitric oxide synthase knockout mice after ischemic stroke: A pilot study. Am J Phys Med Rehabil 97: 492–499. doi: 10.1097/PHM.0000000000000904
    [33] Bernard C (1878) Etude sur la physiologie du coeur. La science experimentale: 316–366. Available from: https://gallica.bnf.fr/ark:/12148/bpt6k69992b/f316.image.
    [34] Thayer JF, Lane RD (2009) Claude Bernard and the heart-brain connection: further elaboration of a model of neurovisceral integration. Neurosci Biobehav Rev 33: 81–88. doi: 10.1016/j.neubiorev.2008.08.004
    [35] Segarra AB, Prieto I, Banegas I, et al. (2012) Asymmetrical effect of captopril on the angiotensinase activity in frontal cortex and plasma of the spontaneously hypertensive rats: expanding the model of neuroendocrine integration. Behav Brain Res 230: 423–427. doi: 10.1016/j.bbr.2012.02.039
    [36] Segarra AB, Prieto I, Banegas I, et al. (2013) The brain-heart connection: frontal cortex and left ventricle angiotensinase activities in control and captopril-treated hypertensive rats-a bilateral study. Int J Hypertens 2013: 156179.
    [37] Segarra AB, Banegas I, Prieto I, et al. (2016) Brain asymmetry and dopamine: beyond motor implications in Parkinson's disease and experimental hemiparkinsonism. Rev Neurol 63: 415–421.
    [38] de Jager L, Amorim EDT, Lucchetti BFC, et al. (2018) Nitric oxide alterations in cardiovascular system of rats with Parkinsonism induced by 6-OHDA and submitted to previous exercise. Life Sci 204: 78–86. doi: 10.1016/j.lfs.2018.05.017
    [39] Alexander N, Kaneda N, Ishii A, et al. (1990) Right-left asymmetry of tyrosine hydroxylase in rat median eminence: influence of arterial baroreflex nerves. Brain Res 523: 195–198. doi: 10.1016/0006-8993(90)91487-2
    [40] Hersh LB (1985) Characterization of membrane-bound aminopeptidases from rat brain:identification of the enkephalin-degrading aminopeptidase. J Neurochem 44: 1427–1435. doi: 10.1111/j.1471-4159.1985.tb08779.x
    [41] Wright JW, Harding JW (1997) Important roles for angiotensin III and IV in the brain renin-angiotensin system. Brain Res Rev 25: 96–124. doi: 10.1016/S0165-0173(97)00019-2
  • This article has been cited by:

    1. Wei Ji, Closed-loop and open-loop equilibrium of a class time-inconsistent linear-quadratic differential games, 2024, 53, 0020-7276, 635, 10.1007/s00182-024-00895-2
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5730) PDF downloads(641) Cited by(2)

Figures and Tables

Figures(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog