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

LpL^{p}-solutions of QSDE driven by Fermion fields with nonlocal conditions under non-Lipschitz coefficients

Guangdong Jing,  Penghui Wang,  Shan Wang Department of Mathematics, Beijing Institute of Technology, Beijing 100081, China.School of Mathematics, Shandong University, Jinan, 250100, China.School of Mathematics, Shandong University, Jinan, 250100, China.E-mail addresses: [email protected](G.Jing), [email protected](P.Wang), [email protected](S.Wang).

Abstract. The main purpose of this paper is to obtain the existence and uniqueness of LpL^{p}-solution to quantum stochastic differential equation driven by Fermion fields with nonlocal conditions in the case of non-Lipschitz coefficients for p>2p>2. The key to our technique is to make use of the Burkholder-Gundy inequality given by Pisier and Xu and Minkowski-type inequality to iterate instead of the fixed point theorem commonly used for nonlocal problems. Moreover, we also obtain the self-adjointness of the LpL^{p}-solution which is important in the study of optimal control problems.

2020 AMS Subject Classification: 46L51, 47J25, 60H10, 81S25.

Keywords. Fermion feilds, nonlocal conditions, Burkholder-Gundy inequality, Bihari inequality.

1 Introduction

In the present paper, we shall consider the following quantum stochastic differential equation (QSDE for short) introduced by Barnett, Streater and Wilde [4]:

dXt=F(Xt,t)dWt+dWtG(Xt,t)+H(Xt,t)dt,dX_{t}=F(X_{t},t)dW_{t}+dW_{t}G(X_{t},t)+H(X_{t},t)dt, (1.1)

driven by Fermion fields which is closely related to the quantum noise, quantum fields etc. The L2L^{2}-solution of the QSDE (1.1) under Lipschitz coefficients was studied by Barnett, Streater and Wilde [4, 5] and Applebauma-Hudson-Lindsay [1, 2, 18] using Itô-isometry and Itô-formula, respectively. Fermion fields and Boson fields are the two most important quantum fields. The QSDE driven by Boson fields have also been studied extensively [6, 10, 12, 19]. Moreover, these two classes of QSDEs can be unified understood as a same framework of the Hudson and Parthasarathy’s quantum stochastic calculus[9, 22] in noncommutative spaces.

The noncommutative LpL^{p}-space and associated harmonic analysis have been deeply studied in [7, 11, 26, 27, 28, 29] and references therein. In particular, martingale inequalities in the non-commutative case have been greatly developed since Pisier and Xu [23] proved the analogy of the classical Burkholder-Gundy inequality for non-commutative martingales, where they defined the noncommutative HpH^{p} martingale space by introducing the row and column square functions. Subsequently, combined Khintchine inequalities of operator values obtained by Lust-Piquard [20], Junge, Randrianantoanina and Xu et al generalized almost all the martingale inequalities such as Doob maximal inequality and Rosenthal inequalities so on of classical martingale theory to the noncommutative case [13, 14, 16, 15, 17, 24, 25], which lays a foundation for the study of quantum stochastic analysis.

Inspired by Burkholder-Gundy inequality, it is natural to solve the QSDE (1.1) in non-commutative LpL^{p}-spaces. Problems with nonlocal conditions are more applicable to real life problems than problems with traditional local conditions. Analogously, we study the LpL^{p}-solution and the basic properties of solution to QSDE with nonlocal initial conditions for p>2p>2. To state our results, we give some assumptions on the QSDE. Some basic notations of noncommutative filtration {𝒞t}t0\{\mathscr{C}_{t}\}_{t\geq 0} will be introduced in Section 2.

Definition 1.1.

A map X:+Lp(𝒞)X:~{}\mathbb{R}^{+}\rightarrow L^{p}(\mathscr{C}) is said to be adapted if XtLp(𝒞t)X_{t}\in L^{p}(\mathscr{C}_{t}) for each t+t\in\mathbb{R}^{+}. A map F:Lp(𝒞)×+Lp(𝒞)F:~{}L^{p}(\mathscr{C})\times\mathbb{R}^{+}\rightarrow L^{p}(\mathscr{C}) is said to be adapted if F(u,t)Lp(𝒞t)F(u,t)\in L^{p}(\mathscr{C}_{t}), for any t+t\in\mathbb{R}^{+} and uLp(𝒞t)u\in L^{p}(\mathscr{C}_{t}).

It is easy to check that if X:+Lp(𝒞)X:\mathbb{R}^{+}\rightarrow L^{p}(\mathscr{C}) and F:Lp(𝒞)×+Lp(𝒞)F:L^{p}(\mathscr{C})\times\mathbb{R}^{+}\rightarrow L^{p}(\mathscr{C}) are both adapted, so is the map tF(Xt,t)t\mapsto F(X_{t},t).

In the rest of this section, we consider the following equation with nonlocal conditions in the interval [0,T][0,T] for some fixed T>0T>0,

{dXt=F(Xt,t)dWt+dWtG(Xt,t)+H(Xt,t)dt,in[t0,T],Xt0=Z+R(X),\left\{\begin{aligned} dX_{t}&=F(X_{t},t)dW_{t}+dW_{t}G(X_{t},t)+H(X_{t},t)dt,\ {\rm in}\ [t_{0},T],\\ X_{t_{0}}&=Z+R(X),\end{aligned}\right. (1.2)

where F(,),G(,),H(,):Lp(𝒞)×[0,T]Lp(𝒞)F(\cdot,\cdot),\ G(\cdot,\cdot),\ H(\cdot,\cdot):L^{p}(\mathscr{C})\times[0,T]\rightarrow L^{p}(\mathscr{C}), R():Lp(𝒞)Lp(𝒞)R(\cdot):L^{p}(\mathscr{C})\to L^{p}(\mathscr{C}) constitutes the nonlocal condition, and Xt0Lp(𝒞t0)X_{t_{0}}\in L^{p}(\mathscr{C}_{t_{0}}) for fixed p>2p>2.

Definition 1.2.

A stochastic process X():[0,T]Lp(𝒞)X(\cdot):[0,T]\to L^{p}(\mathscr{C}) is called a solution to the equation (1.2) if it satisfies

Xt=Z+R(Xt)+t0tF(Xs,s)𝑑Ws+t0t𝑑WsG(Xs,s)+t0tH(Xs,s)𝑑s,a.s. in[t0,T].X_{t}=Z+R(X_{t})+\int_{t_{0}}^{t}F(X_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s},s)+\int_{t_{0}}^{t}H(X_{s},s)ds,\ \textrm{a.s.\ in}\ [t_{0},T].

Throughout the paper, we shall make use of the following conditions:

Assumption 1.1.

F(,),G(,),H(,)F(\cdot,\cdot),\ G(\cdot,\cdot),\ H(\cdot,\cdot): Lp(𝒞)×[0,T]Lp(𝒞)L^{p}(\mathscr{C})\times[0,T]\rightarrow L^{p}(\mathscr{C}) are operator-valued functions such that

(A1)

F(,),G(,),H(,):Lp(𝒞)×+Lp(𝒞)F(\cdot,\cdot),\ G(\cdot,\cdot),\ H(\cdot,\cdot):L^{p}(\mathscr{C})\times\mathbb{R}^{+}\rightarrow L^{p}(\mathscr{C}) are adapted.

(A2)

For any xLp(𝒞)x\in L^{p}(\mathscr{C}), F(x,),G(x,),H(x,):[0,T]Lp(𝒞)F(x,\cdot),\ G(x,\cdot),\ H(x,\cdot):[0,T]\to L^{p}(\mathscr{C}) are continuous a.s.

(A3)

Osgood condition:  For any x1,x2Lp(𝒞)x_{1},\ x_{2}\in L^{p}(\mathscr{C}) and a.e. t[0,T]t\in[0,T],

F(x1,t)F(x2,t)p2+G(x1,t)G(x2,t)p2+H(x1,t)H(x2,t)p2ρ(x1x2p2).\|F(x_{1},t)-F(x_{2},t)\|_{p}^{2}+\|G(x_{1},t)-G(x_{2},t)\|_{p}^{2}+\|H(x_{1},t)-H(x_{2},t)\|_{p}^{2}\leq\rho(\|x_{1}-x_{2}\|_{p}^{2}).\\

where ρ:++\rho:\mathbb{R}^{+}\to\mathbb{R}^{+} is a continuous non-decreasing function with ρ(0)=0\rho(0)=0, ρ(r)>0\rho(r)>0 for r>0r>0, such that 0+drρ(r)=+\int_{0^{+}}\frac{dr}{\rho(r)}=+\infty.

(A4)

RR is continuous and adapted, and there exists a constant 0<C(R)<10<C(R)<1 such that

R(x1)R(x2)pC(R)x1x2p,x1,x2Lp(𝒞).\|R(x_{1})-R(x_{2})\|_{p}\leq C(R)\|x_{1}-x_{2}\|_{p},\ \forall\ x_{1},\ x_{2}\in L^{p}(\mathscr{C}). (1.3)
Theorem 1.1.

Under the assumptions (A1)-(A4), for p>2p>2, there is a unique continuous adapted LpL^{p}-solution {Xt}tt0\{X_{t}\}_{t\geq t_{0}} to the following quantum stochastic integral equation with nonlocal initial value

Xt=Z+R(Xt)+t0tF(Xs,s)𝑑Ws+t0t𝑑WsG(Xs,s)+t0tH(Xs,s)𝑑s,in[t0,T].X_{t}=Z+R(X_{t})+\int_{t_{0}}^{t}F(X_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s},s)+\int_{t_{0}}^{t}H(X_{s},s)ds,\ {\rm{in}}\ [t_{0},T]. (1.4)
Remark 1.1.

There are many efforts to study the solution to the QSDE (1.1) by many mathematicians.

  • (1).

    By using Itô-isometry, Barnett, Streater and Wilde [4, 5] considered the L2L^{2}-solution to the QSDE (1.1).

  • (2).

    The original study [1] on the following QSDE

    dXt=F(Xt,t)dAt+dAtG(Xt,t)+H(Xt,t)dt,in[0,T],dX_{t}=F(X_{t},t)dA_{t}+dA^{*}_{t}G(X_{t},t)+H(X_{t},t)dt,\ \textrm{in}\ [0,T], (1.5)

    is to consider the solution in the weak sense, i.e., XtX_{t} is called the solution to the initial value problem of the QSDE, if

    Xtu=(Z+0tF(Xs,s)𝑑As+0t𝑑AsG(Xs,s)+0tH(Xs,s)𝑑s)u,X_{t}u=\left(Z+\int_{0}^{t}F(X_{s},s)dA_{s}+\int_{0}^{t}dA^{*}_{s}G(X_{s},s)+\int_{0}^{t}H(X_{s},s)ds\right)u,

    for any uD(Xt)u\in D(X_{t}), where XtX_{t} can be considered as an unbounded operator densely defined on the Hilbert space Λ(L2(+))\Lambda(L^{2}(\mathbb{R}^{+})) with the domain D(Xt)D(X_{t}).

  • (3).

    The martingale inequality and Burkholder-Gundy inequality of Pisier and Xu have been used by Dirksen [8] to study the LpL^{p}-solution for p>2p>2 to QSDE with respect to any normal LpL^{p}-martingale without the drift term 0tH(Xs,s)𝑑s\int_{0}^{t}H(X_{s},s)ds. The reason why the Burkholder-Gundy inequality can be used directly is because, without the drift term 0tH(Xs,s)𝑑s\int_{0}^{t}H(X_{s},s)ds, the solution of the QSDE is a martingale, and hence by the Lipschiz condition on p([0,t]){\cal H}^{p}([0,t]), one can have existence and uniqueness of the solution.

  • (4)

    Compared with the result of Dirksen [8], the QSDE (1.1) that we consider has the drift term. The reason why we can deal with the drift term is because we obtain the estimation in L2(0,T;Lp(𝒞T))L^{2}(0,T;L^{p}({\mathscr{C}}_{T})) with the help of Burkholder-Gundy inequality and Minkowski inequality.

This paper is organized as follows. In Section 2, we review some fundamental notations and preliminaries on Fermion fields, and introduce several useful inequalities which are the main techniques of subsequent proof. In section 3 and section 4, we obtain the existence, uniqueness and the stability of LpL^{p}-solution to the QSDE (1.1) and (1.5) with nonlocal conditions. In section 5, we get the self-adjointness and the Markov property of the solution to QSDE.

2 Preliminaries and Burkholder-Gundy inequality

In this section, we introduce the main techniques to solve problems later. We first recall some concepts [3, 4, 5, 23, 32, 33] necessary to the whole article.

Let \mathscr{H} be a separable complex Hilbert space. For any zz\in\mathscr{H}, the creation operator C(z):Λn()Λn+1()C(z):\Lambda_{n}(\mathscr{H})\rightarrow\Lambda_{n+1}(\mathscr{H}) defined by un+1zu\mapsto\sqrt{n+1}\ z\wedgeuu, is a bounded operator on Λ()\Lambda(\mathscr{H}) with C(z)=z\|C(z)\|=\|z\|, where Λ0():=\Lambda_{0}(\mathscr{H}):=\mathbb{C}. Meanwhile, define the annihilation operator A(z)=C(z)A(z)=C^{*}(z). Then the antisymmetric Fock space Λ():=n=0Λn()\Lambda(\mathscr{H}):=\oplus_{n=0}^{\infty}\Lambda_{n}(\mathscr{H}) over \mathscr{H} is a Hilbert space. Moreover, define the fermion field Ψ(z)\Psi(z) on Λ()\Lambda(\mathscr{H}) by Ψ(z):=C(z)+A(Jz)\Psi(z):=C(z)+A(Jz), where J:J:\mathscr{H}\rightarrow\mathscr{H} is the conjugation operator (i.e., JJ is antilinear, antiunitary and J2=1J^{2}=1). Denote by 𝒞\mathscr{C} the von Neumann algebra generated by the bounded operators {Ψ(z):z}\{\Psi(z):z\in\mathscr{H}\}. For the Fock vacauum 𝟙Λ()\mathds{1}\in\Lambda(\mathscr{H}), define

m():=𝟙,𝟙m(\cdot):=\langle\mathds{1},\cdot\mathds{1}\rangle (2.1)

on 𝒞\mathscr{C}. Obviously, mm is a normal faithful state on 𝒞\mathscr{C}, and (Λ(),𝒞,m)(\Lambda(\mathscr{H}),\mathscr{C},m) is a quantum probability space by [30].

Now, let =L2(+)\mathscr{H}=L^{2}(\mathbb{R}^{+}), and 𝒞t\mathscr{C}_{t} be the von Neumann subalgebra of 𝒞\mathscr{C} generated by {Ψ(u):uL2(+),esssuppu[0,t]},\{\Psi(u):u\in L^{2}(\mathbb{R}_{+}),{\rm~{}ess~{}supp}~{}u\subseteq[0,t]\}, then {𝒞t}t0\{\mathscr{C}_{t}\}_{t\geq 0} is an increasing family of von Neumann subalgebras of 𝒞\mathscr{C} which is the noncommutative analogue of filtration in stochastic analysis. Let

Wt=Ψ(χ[0,t])=C(χ[0,t])+A(Jχ[0,t]),t0,W_{t}=\Psi(\chi_{[0,t]})=C(\chi_{[0,t]})+A(J\chi_{[0,t]}),\ t\geq 0,

then {Wt:t+}\{W_{t}:t\in\mathbb{R}^{+}\} is a Fermion Brownian motion adapted to the family {𝒞t:t+}\{\mathscr{C}_{t}:t\in\mathbb{R}^{+}\}. For any 1p<1\leq p<\infty, define the noncommutative LpL^{p}-norm on 𝒞\mathscr{C} by

fp:=m(|f|p)1p=𝟙,|f|p𝟙1p\|f\|_{p}:=m(|f|^{p})^{1\over p}=\langle\mathds{1},|f|^{p}\mathds{1}\rangle^{1\over p}

where |f|=(ff)12|f|=(f^{*}f)^{1\over 2}, then Lp(𝒞,m)L^{p}(\mathscr{C},m) is the completion of (𝒞,p)(\mathscr{C},\|\cdot\|_{p}), which is the noncommutative LpL^{p}-space, abbreviated as Lp(𝒞)L^{p}(\mathscr{C}). For any interval [0,T]+[0,T]\subset\mathbb{R^{+}}, set

C𝔸(0,T;Lp(𝒞T)):={x():\displaystyle C_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})):=\Big{\{}x(\cdot):\ [0,T]Lp(𝒞T)x(t)Lp(𝒞t)\displaystyle[0,T]\to L^{p}(\mathscr{C}_{T})\mid x(t)\in L^{p}(\mathscr{C}_{t})\
andlimstx(s)x(t)p=0,s,t[0,T]}.\displaystyle{\rm and}\ \lim_{s\to t}\|x(s)-x(t)\|_{p}=0,\ \forall\ s,\ t\in[0,T]\Big{\}}.

It is easy to check that C𝔸(0,T;Lp(𝒞T))C_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})) is a Banach space with the norm given by

xC𝔸(0,T;Lp(𝒞T))=maxt[0,T]x(t)p.\|x\|_{C_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T}))}=\max_{t\in[0,T]}\|x(t)\|_{p}.
Definition 2.1.

An adapted LpL^{p}-processes ff on [t0,t][t_{0},t] is said to be simple if it can be expressed as

f=k=0n1f(tk)χ[tk,tk+1)f=\sum_{k=0}^{n-1}f(t_{k})\chi_{[t_{k},t_{k+1})} (2.2)

on [t0,t][t_{0},t] for t0t1tn=tt_{0}\leq t_{1}\leq\cdots\leq t_{n}=t and f(tk)Lp(𝒞tk)f(t_{k})\in L^{p}(\mathscr{C}_{t_{k}}) for all 0kn10\leq k\leq n-1.

By [3], the Itô-Clifford integral of any simple adapted LpL^{p}-process with respect to Fermion Brownian motion WtW_{t} is defined as follows.

Definition 2.2.

If f=kf(tk)χ[tk,tk+1)f=\sum\limits_{k}f(t_{k})\chi_{[t_{k},t_{k+1})} is a simple adapted LpL^{p}-processes on [t0,t][t_{0},t], then the Itô-Clifford stochastic integral of ff over [t0,t][t_{0},t] with respect to WtW_{t} is

t0tf(s)𝑑Ws=k=0n1f(tk)(Wtk+1Wtk).\int_{t_{0}}^{t}f(s)dW_{s}=\sum_{k=0}^{n-1}f(t_{k})(W_{t_{k+1}}-W_{t_{k}}). (2.3)

For p1p\geq 1, let 𝒮𝔸p(+){\cal S}^{p}_{\mathbb{A}}(\mathbb{R}^{+}) be the linear space of all simple adapted LpL^{p}-processes, i.e.

𝒮𝔸p(+):={f:+Lp(𝒞),f is simple and adapted}.{\cal S}^{p}_{\mathbb{A}}(\mathbb{R}^{+}):=\{f:\mathbb{R}^{+}\to L^{p}(\mathscr{C}),\ f\text{ is simple and adapted}\}.

Then 𝒮𝔸p([0,t]){\cal S}^{p}_{\mathbb{A}}([0,t]) is subspace of 𝒮𝔸p(+){\cal S}^{p}_{\mathbb{A}}(\mathbb{R}^{+}) whose processes vanish in (t,)(t,\infty). It is clear that t0tf(s)𝑑Ws\int_{t_{0}}^{t}f(s)dW_{s} is a Clifford LpL^{p}-martingale for any f𝒮𝔸p([0,t])f\in{\cal S}^{p}_{\mathbb{A}}([0,t]), i.e. 𝔼(t0tf(τ)𝑑Wτ𝒞s)=t0sf(τ)𝑑Wτ\mathbb{E}\left(\int_{t_{0}}^{t}f(\tau)dW_{\tau}\mid\mathscr{C}_{s}\right)=\int_{t_{0}}^{s}f(\tau)dW_{\tau} for any t0stt_{0}\leq s\leq t. For all f𝒮𝔸p([0,t])f\in{\cal S}^{p}_{\mathbb{A}}([0,t]), let

fp([0,t]):=max{(0t|f(s)|2𝑑s)12p,(0t|f(s)|2𝑑s)12p},\left\|f\right\|_{\mathcal{H}^{p}([0,t])}:=\max\left\{\left\|\left(\int_{0}^{t}|f(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p},\ \left\|\left(\int_{0}^{t}|f^{*}(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p}\right\},

and denote by the non-commutative Hardy space p([0,t])\mathcal{H}^{p}([0,t]) the completion of 𝒮𝔸p([0,t]){\cal S}^{p}_{\mathbb{A}}([0,t]) under the norm p([0,t])\|\cdot\|_{\mathcal{H}^{p}([0,t])}. For simplicity, we use locp(+)\mathcal{H}^{p}_{loc}(\mathbb{R}^{+}) to represent the set of stochastic processes f:+Lp(𝒞)f:\mathbb{R}^{+}\to L^{p}(\mathscr{C}) and χ[0,t]fp([0,t])\chi_{[0,t]}f\in\mathcal{H}^{p}([0,t]). Moreover, we have the following the Burkholder-Gundy inequality (2.4) of Clifford LpL^{p}-martingale first established in [23].

Lemma 2.1.

[23, Theorem 4.1] Let 2p<2\leq p<\infty. Then, for any flocp(+)f\in\mathcal{H}^{p}_{loc}(\mathbb{R}^{+}) and its Itô-Clifford integral

Xt=0tf(s)𝑑Ws,t0,X_{t}=\int_{0}^{t}f(s)dW_{s},\ t\geq 0,

it holds that

αp1fp([0,t])Xtpβpfp([0,t]),t0,\alpha_{p}^{-1}\|f\|_{\mathcal{H}^{p}([0,t])}\leq\|X_{t}\|_{p}\leq\beta_{p}\|f\|_{\mathcal{H}^{p}([0,t])},\ t\geq 0, (2.4)

where αp\alpha_{p} and βp\beta_{p} are constant depend on pp.

The stochastic integral (2.3) is also called right stochastic integral. Analogously, we can define left stochastic integrals, and have the Burkholder-Gundy inequality with respect to left stochastic integrals.

Lemma 2.2.

[8, Theorem 7.2] Let 1p<1\leq p<\infty. For any flocp(+)f\in\mathcal{H}^{p}_{loc}(\mathbb{R}^{+}), then the left stochastic integral 0t𝑑Wsf(s)\int_{0}^{t}dW_{s}f(s) and right stochastic integral 0tf(s)𝑑Ws\int_{0}^{t}f(s)dW_{s} can be well-defined and

0t𝑑Wsf(s)ppfp([0,t])p0tf(s)𝑑Wsp.\left\|\int_{0}^{t}dW_{s}f(s)\right\|_{p}\simeq_{p}\|f\|_{\mathcal{H}^{p}([0,t])}\simeq_{p}\left\|\int_{0}^{t}f(s)dW_{s}\right\|_{p}.

By Lemma 2.1 and Lemma 2.2, the Itô-Clifford integral can be defined for any element of p([0,t])\mathcal{H}^{p}([0,t]), and Burkholder-Gundy inequality holds ture. For any t0t\geq 0, let L𝔸q(0,t;Lp(𝒞t))L^{q}_{\mathbb{A}}(0,t;L^{p}(\mathscr{C}_{t})) be the completion of 𝒮𝔸p([0,t]){\cal S}^{p}_{\mathbb{A}}([0,t]) with the norm

fL𝔸q(0,t;Lp(𝒞t))=(0tf(s)pq𝑑s)1q.\|f\|_{L^{q}_{\mathbb{A}}(0,t;L^{p}(\mathscr{C}_{t}))}=\left(\int_{0}^{t}\|f(s)\|_{p}^{q}ds\right)^{\frac{1}{q}}.

Similarly, L𝔸p(𝒞t;Lq(0,t))L^{p}_{\mathbb{A}}(\mathscr{C}_{t};L^{q}(0,t)) is the completion of 𝒮𝔸p([0,t]){\cal S}^{p}_{\mathbb{A}}([0,t]) with the norm

fL𝔸p(𝒞t;Lq(0,t))=(0t|f(s)|q𝑑s)1qp,t0.\|f\|_{L^{p}_{\mathbb{A}}(\mathscr{C}_{t};L^{q}(0,t))}=\left\|\left(\int_{0}^{t}|f(s)|^{q}ds\right)^{\frac{1}{q}}\right\|_{p},\ t\geq 0.

As an application of Minkowski inequality, we have the following result.

Theorem 2.3.

Let 1qp<1\leq q\leq p<\infty. Then, for any fL𝔸q(0,T;Lp(𝒞T))f\in L^{q}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})),

(0t|f(s)|q𝑑s)1qp(0tf(s)pq𝑑s)1q, 0tT.\left\|\left(\int_{0}^{t}|f(s)|^{q}ds\right)^{\frac{1}{q}}\right\|_{p}\leq\left(\int_{0}^{t}\|f(s)\|_{p}^{q}ds\right)^{\frac{1}{q}},\ 0\leq t\leq T. (2.5)

Furthermore, L𝔸q(0,T;Lp(𝒞T))L𝔸p(𝒞T;Lq(0,T))L^{q}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T}))\subseteq L^{p}_{\mathbb{A}}(\mathscr{C}_{T};L^{q}(0,T)).

Proof.

Let

0=t0t1t2tn=t0=t_{0}\leq t_{1}\leq t_{2}\leq\cdots\leq t_{n}=t

be an equal time partition of [0,t][0,t] where the mesh of the subdivision is l=t/n=tk+1tkl=t/n=t_{k+1}-t_{k}, k=0, 1,,n1k=0,\ 1,\ \cdots,\ n-1. For simple adapted process k0atkχ[tk,tk+1)\sum\limits_{k\geq 0}a_{t_{k}}\chi_{[t_{k},t_{k+1})} of Lp(𝒞)L^{p}(\mathscr{C}) where atkLp(𝒞tk)a_{t_{k}}\in L^{p}(\mathscr{C}_{t_{k}}), and any positive integer nn, one has

(k=0n1|atk|q(tk+1tk))1qp=l1q(k=0n1|atk|q)1qp=l1qk=0n1|atk|qpq1q.\left\|\left(\sum_{k=0}^{n-1}|a_{t_{k}}|^{q}(t_{k+1}-t_{k})\right)^{\frac{1}{q}}\right\|_{p}=l^{\frac{1}{q}}\left\|\left(\sum_{k=0}^{n-1}|a_{t_{k}}|^{q}\right)^{\frac{1}{q}}\right\|_{p}=l^{\frac{1}{q}}\left\|\sum_{k=0}^{n-1}|a_{t_{k}}|^{q}\right\|_{\frac{p}{q}}^{\frac{1}{q}}. (2.6)

Since pq1\frac{p}{q}\geq 1, by Minkowski inequality [31, Theorem 5.2.2],

k=0n1|atk|qpqk=0n1|atk|qpq,n+.\left\|\sum_{k=0}^{n-1}|a_{t_{k}}|^{q}\right\|_{\frac{p}{q}}\leq\sum_{k=0}^{n-1}\left\||a_{t_{k}}|^{q}\right\|_{\frac{p}{q}},\ n\in\mathbb{N}^{+}. (2.7)

By (2.6) and (2.7), we have

(k=0n1|atk|q(tk+1tk))1qp(k=0n1atkpq(tk+1tk))1q.\displaystyle\left\|\left(\sum_{k=0}^{n-1}|a_{t_{k}}|^{q}(t_{k+1}-t_{k})\right)^{\frac{1}{q}}\right\|_{p}\leq\left(\sum_{k=0}^{n-1}\|a_{t_{k}}\|_{p}^{q}(t_{k+1}-t_{k})\right)^{\frac{1}{q}}.

Since 𝒮𝔸p([0,T]){\cal S}^{p}_{\mathbb{A}}([0,T]) is dense in L𝔸p(𝒞;Lq(0,T))L^{p}_{\mathbb{A}}(\mathscr{C};L^{q}(0,T)),

(0t|f(s)|q𝑑s)1qp(0tf(s)pq𝑑s)1q, 0tT,\left\|\left(\int_{0}^{t}|f(s)|^{q}ds\right)^{\frac{1}{q}}\right\|_{p}\leq\left(\int_{0}^{t}\|f(s)\|^{q}_{p}ds\right)^{\frac{1}{q}},\ 0\leq t\leq T,

for any fL𝔸q(0,T;Lp(𝒞T))f\in L^{q}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})), and L𝔸q(0,T;Lp(𝒞T))L𝔸p(𝒞T;Lq(0,T))L^{q}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T}))\subseteq L^{p}_{\mathbb{A}}(\mathscr{C}_{T};L^{q}(0,T)). ∎

Remark 2.1.

Actually, the above inequality (2.5) holds for any 0<qp<0<q\leq p<\infty. In particular, when p=q, the equality in the above inequality (2.5) holds and is Fubini’s theorem in the noncommutative case.

In general, the Burkholder-Gundy inequality (2.4) cannot be directly used to study LpL^{p}-solution to QSDE. Instead, we need the following easy corollary.

Corollary 2.4.

Let p>2p>2. Then, for any fL𝔸2(0,T;Lp(𝒞T))f\in L^{2}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})), there is positive constant C(p)C(p) such that

0tf(s)𝑑WspC(p)(0tf(s)p2𝑑s)12, 0tT.\left\|\int_{0}^{t}f(s)dW_{s}\right\|_{p}\leq C(p)\left(\int_{0}^{t}\|f(s)\|^{2}_{p}ds\right)^{\frac{1}{2}},\ \ 0\leq t\leq T. (2.8)

Moreover, L𝔸2(0,T;Lp(𝒞T))p([0,T])L^{2}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T}))\subseteq{\cal H}^{p}([0,T]) and

fp([0,t])(0tf(s)p2𝑑s)12, 0tT.\|f\|_{{\cal H}^{p}([0,t])}\leq\left(\int_{0}^{t}\|f(s)\|^{2}_{p}ds\right)^{\frac{1}{2}},\ \ 0\leq t\leq T. (2.9)
Proof.

According to Theorem 2.3, for any fL𝔸2(0,T;Lp(𝒞T))f\in L^{2}_{\mathbb{A}}(0,T;L^{p}(\mathscr{C}_{T})), one has

(0t|f(s)|2𝑑s)12p(0tf(s)p2𝑑s)12.\left\|\left(\int_{0}^{t}|f(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p}\leq\left(\int_{0}^{t}\|f(s)\|_{p}^{2}ds\right)^{\frac{1}{2}}. (2.10)

Since f(s)p=f(s)p\|f(s)\|_{p}=\|f^{*}(s)\|_{p} for any s[0,T]s\in[0,T],

fp([0,t])=max{(0t|f(s)|2𝑑s)12p,(0t|f(s)|2𝑑s)12p}(0tf(s)p2𝑑s)12.\|f\|_{{\cal H}^{p}([0,t])}=\max\left\{\left\|\left(\int_{0}^{t}|f(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p},\ \left\|\left(\int_{0}^{t}|f(s)^{*}|^{2}ds\right)^{\frac{1}{2}}\right\|_{p}\right\}\leq\left(\int_{0}^{t}\|f(s)\|^{2}_{p}ds\right)^{\frac{1}{2}}.

Combined with the Burkholder-Gundy inequality (2.4), we have (2.8) immediately. ∎

Now, we give the parity of each element of Lp(𝒞)L^{p}(\mathscr{C}). Let the parity operator PP be automorphism map on von Neumann algebra 𝒞\mathscr{C} generated by bounded linear operators on Λ()\Lambda(\mathscr{H}) as is in [3, 5, 23].

Definition 2.3.

For any hLp(𝒞)h\in L^{p}(\mathscr{C}), hh is said to be odd if Ph=hPh=-h, hh is said to be even if Ph=hPh=h. And, hh has definite parity if hh is even or odd.

Furthermore, for any 1<p<1<p<\infty,

Lp(𝒞)=Lp(𝒞o)Lp(𝒞e),L^{p}(\mathscr{C})=L^{p}(\mathscr{C}_{o})\oplus L^{p}(\mathscr{C}_{e}), (2.11)

where Lp(𝒞e),Lp(𝒞o)L^{p}(\mathscr{C}_{e}),\ L^{p}(\mathscr{C}_{o}) denote the even part and the odd part, respectively. More precisely, for any hLp(𝒞)h\in L^{p}(\mathscr{C}), h=h+Ph2+hPh2=ho+heh=\frac{h+Ph}{2}+\frac{h-Ph}{2}=h_{o}+h_{e}, where heh_{e} and hoh_{o} are even and odd, respectively. Since PP is isometric on Lp(𝒞)L^{p}(\mathscr{C}),

max{hop,hep}hphop+hep.\max\left\{\|h_{o}\|_{p},\|h_{e}\|_{p}\right\}\leq\|h\|_{p}\leq\|h_{o}\|_{p}+\|h_{e}\|_{p}.

Let \mathscr{E} denote the algebra of even polynomials in the fields {Ψ(u):u}\{\Psi(u):u\in\mathscr{H}\}, and let 𝒞e\mathscr{C}_{e} be the WW^{*}-subalgebra of 𝒞\mathscr{C} generated by \mathscr{E}. If hLp(𝒞)h\in L^{p}(\mathscr{C}) is even there is a sequence {hn}\{h_{n}\} in \mathscr{E} such that hnhh_{n}\rightarrow h in Lp(𝒞)L^{p}(\mathscr{C}), and therefore hnhh_{n}^{*}\rightarrow h^{*} in Lp(𝒞)L^{p}(\mathscr{C}). It follows that hh^{*} is also even. Similarly, if gg is odd in Lp(𝒞)L^{p}(\mathscr{C}), there is a sequence {gn}\{g_{n}\} of odd polynomials in the fields with gngg_{n}\rightarrow g and thus gngg_{n}^{*}\rightarrow g^{*} in Lp(𝒞)L^{p}(\mathscr{C}), that is, gg^{*} is odd as well. It follows that if h=hh=h^{*} in Lp(𝒞)L^{p}(\mathscr{C}) and h=he+hoh=h_{e}+h_{o}, then he=heh_{e}=h_{e}^{*} and ho=hoh_{o}=h_{o}^{*} in Lp(𝒞)L^{p}(\mathscr{C}).

Lemma 2.5.

[3, Lemma 3.15] Let {Wt}tt0\{W_{t}\}_{t\geq t_{0}} be Brownian motion. If hLp(𝒞t0)h\in L^{p}(\mathscr{C}_{t_{0}}) has definite parity, then

h(Wt2Wt1)=±(Wt2Wt1)h,t0t1t2,h(W_{t_{2}}-W_{t_{1}})=\pm(W_{t_{2}}-W_{t_{1}})h,\quad t_{0}\leq t_{1}\leq t_{2},

depending on whether hh is even or odd.

Lemma 2.6.

[21, Theorem 1.8.2 Bihari inequality] Let ρ:[0,+)[0,+)\rho:[0,+\infty)\to[0,+\infty) be a continuous and non-decreasing function vanishing at 0 satisfying 0+drρ(r)=\int_{0^{+}}\frac{dr}{\rho(r)}=\infty. Suppose u(t)u(t) is a continuous nonnegative function on [t0,T][t_{0},T] such that

u(t)u0+t0tϕ(r)ρ(u(r))𝑑r,t0tT,u(t)\leq u_{0}+\int_{t_{0}}^{t}\phi(r)\rho(u(r))dr,\ t_{0}\leq t\leq T, (2.12)

where u0u_{0} is a nonnegative cinstant and ϕ:[t0,T]+\phi:[t_{0},T]\to\mathbb{R}^{+}, then

u(t)U1(U(u0)+t0tϕ(r)𝑑r),t0tT,u(t)\leq U^{-1}\left(U(u_{0})+\int_{t_{0}}^{t}\phi(r)dr\right),\ t_{0}\leq t\leq T,

where U(t)=t0t1ρ(r)𝑑rU(t)=\int_{t_{0}}^{t}\frac{1}{\rho(r)}dr, U1U^{-1} is the conver function of UU. In particular, u0=0u_{0}=0, then u(t)=0u(t)=0 for all t0tTt_{0}\leq t\leq T.

3 The existence and uniqueness of solutions to QSDE

This section is devoted to proving the existence and uniqueness of LpL^{p}-solution to the equation (1.2) with non-Lipschitz coefficients for p>2p>2.

Theorem 3.1.

Let Assumption 1.1 hold. Then the equation (1.2) admits a unique solution X()C𝔸(t0,T;Lp(𝒞T))X(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T})).

Proof.

We shall deal with the existence and uniqueness separately.
Existence: The proof of the existence is divided into three steps.

Step 1. The iteration {Xt(n)}n0\{X_{t}^{(n)}\}_{n\geq 0} is well-defined for any t[t0,T]t\in[t_{0},T]. Let T>t0T>t_{0}, t0tTt_{0}\leq t\leq T be fixed. For any non-negative integer nn, define Xt(n)X_{t}^{(n)} in Lp(𝒞)L^{p}(\mathscr{C}) inductively by

Xt(n+1)=Z+R(Xt(n+1))+t0tF(Xs(n),s)𝑑Ws+t0t𝑑WsG(Xs(n),s)+t0tH(Xs(n),s)𝑑s,X_{t}^{(n+1)}=Z+R(X_{t}^{(n+1)})+\int_{t_{0}}^{t}F(X_{s}^{(n)},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s}^{(n)},s)+\int_{t_{0}}^{t}H(X_{s}^{(n)},s)ds, (3.1)

where the well-definedness of iteration comes from Banach fixed point theorem and that RR is a strict contraction.

Firstly, we claim that each Xt(n)X_{t}^{(n)}, n1n\geq 1, defines an adapted LpL^{p}-continuous process on [t0,T][t_{0},T] by induction. By assumption, F(Z,s),G(Z,s)F(Z,s),\ G(Z,s) and H(Z,s)H(Z,s) are LpL^{p}-continuous with respect to ss and belong to Lp(𝒞s)L^{p}(\mathscr{C}_{s}) for t0sTt_{0}\leq s\leq T, then quantum stochastic integral Xt(1)X_{t}^{(1)} exists for t[t0,T]t\in[t_{0},T]. Furthermore, we can obtain the boundedness of Xt(1)X_{t}^{(1)} by the continuity on compact sets and easily verify that tXt(1)t\mapsto X_{t}^{(1)} is continuous: [t0,T]Lp(𝒞)[t_{0},T]\rightarrow L^{p}(\mathscr{C}).

Now, if Xt(n)X_{t}^{(n)} is assumed to be adapted and continuous, then F(Xt(n),t),G(Xt(n),t)F(X_{t}^{(n)},t),~{}G(X_{t}^{(n)},t) and H(Xt(n),t)H(X_{t}^{(n)},t) are adapted, LpL^{p}-continuous on [t0,T][t_{0},~{}T] and bounded, thus Xt(n+1)X_{t}^{(n+1)} is adapted. For any t1,t2[t0,T]t_{1},\ t_{2}\in[t_{0},T], by (2.4) and Assumption 1.1,

Xt1(n+1)Xt2(n+1)p\displaystyle\|X_{t_{1}}^{(n+1)}-X_{t_{2}}^{(n+1)}\|_{p} R(Xt1(n+1))R(Xt2(n+1))p+t1t2F(Xs(n),s)𝑑Wsp\displaystyle\leq\|R(X_{t_{1}}^{(n+1)})-R(X_{t_{2}}^{(n+1)})\|_{p}+\left\|\int_{t_{1}}^{t_{2}}F(X_{s}^{(n)},s)dW_{s}\right\|_{p}
+t1t2𝑑WsG(Xs(n),s)p+t1t2H(Xs(n),s)𝑑sp\displaystyle\indent+\left\|\int_{t_{1}}^{t_{2}}dW_{s}G(X_{s}^{(n)},s)\right\|_{p}+\left\|\int_{t_{1}}^{t_{2}}H(X_{s}^{(n)},s)ds\right\|_{p}
C(R)Xt1(n+1)Xt2(n+1)p+C(p)F(Xs(n),s)p([t1,t2])\displaystyle\leq C(R)\|X_{t_{1}}^{(n+1)}-X_{t_{2}}^{(n+1)}\|_{p}+C(p)\|F(X_{s}^{(n)},s)\|_{\mathcal{H}^{p}([t_{1},t_{2}])}
+C(p)G(Xs(n),s)p([t1,t2])+t1t2H(Xs(n),s)𝑑sp.\displaystyle\indent+C(p)\|G(X_{s}^{(n)},s)\|_{\mathcal{H}^{p}([t_{1},t_{2}])}+\left\|\int_{t_{1}}^{t_{2}}H(X_{s}^{(n)},s)ds\right\|_{p}.

Now, subtracting C(R)Xt1(n+1)Xt2(n+1)pC(R)\|X_{t_{1}}^{(n+1)}-X_{t_{2}}^{(n+1)}\|_{p} from both sides of above inequality and applying Corollary 2.4 and Hölder inequality, we get

(1C(R))Xt1(n+1)Xt2(n+1)p\displaystyle(1-C(R))\|X_{t_{1}}^{(n+1)}-X_{t_{2}}^{(n+1)}\|_{p}
C(p)(t1t2F(Xs(n),s)p2𝑑s)12+C(p)(t1t2G(Xs(n),s)p2𝑑s)12+C(T)(t1t2H(Xs(n),s)p2𝑑s)12,\displaystyle\leq C(p)\left(\int_{t_{1}}^{t_{2}}\|F(X_{s}^{(n)},s)\|_{p}^{2}ds\right)^{\frac{1}{2}}+C(p)\left(\int_{t_{1}}^{t_{2}}\|G(X_{s}^{(n)},s)\|_{p}^{2}ds\right)^{\frac{1}{2}}+C(T)\left(\int_{t_{1}}^{t_{2}}\|H(X_{s}^{(n)},s)\|_{p}^{2}ds\right)^{\frac{1}{2}},

which implies that tXt(n+1)t\mapsto X_{t}^{(n+1)} is LpL^{p}-continuous on [t0,T][t_{0},T]. Hence we have proved our claim by induction.

Step 2. The sequence of iteration is convergent under the given conditions. For any t[t0,T]t\in[t_{0},T], by Minkowski inequality,

Xt(n+1)Xt(n)p\displaystyle\|X_{t}^{(n+1)}-X_{t}^{(n)}\|_{p}
R(Xt(n+1))R(Xt(n))p+t0t(F(Xs(n),s)F(Xs(n1),s))𝑑Wsp\displaystyle\indent\leq\|R(X_{t}^{(n+1)})-R(X_{t}^{(n)})\|_{p}+\left\|\int_{t_{0}}^{t}(F(X_{s}^{(n)},s)-F(X_{s}^{(n-1)},s))dW_{s}\right\|_{p}
+t0t𝑑Ws(G(Xs(n),s)G(Xs(n1),s))p+t0t(H(Xs(n),s)H(Xs(n1),s))𝑑sp.\displaystyle\indent\indent+\left\|\int_{t_{0}}^{t}dW_{s}(G(X_{s}^{(n)},s)-G(X_{s}^{(n-1)},s))\right\|_{p}+\left\|\int_{t_{0}}^{t}(H(X_{s}^{(n)},s)-H(X_{s}^{(n-1)},s))ds\right\|_{p}.

By similar analysis as above, the elementary inequality (a+b+c)23(a2+b2+c2)(a+b+c)^{2}\leq 3(a^{2}+b^{2}+c^{2}), Hölder inequality and Osgood conditions of F,G,HF,\ G,\ H, there is constant C(p,R,T)C(p,R,T) such that

Xt(n+1)Xt(n)p2\displaystyle\|X_{t}^{(n+1)}-X_{t}^{(n)}\|_{p}^{2}
3(1C(R))2(C2(p)t0tF(Xs(n),s)F(Xs(n1),s)p2ds\displaystyle\indent\leq\frac{3}{(1-C(R))^{2}}\Bigg{(}C^{2}(p)\int_{t_{0}}^{t}\|F(X_{s}^{(n)},s)-F(X_{s}^{(n-1)},s)\|_{p}^{2}ds
+C2(p)t0tG(Xs(n),s)G(Xs(n1),s)p2ds+C2(T)t0tH(Xs(n),s)H(Xs(n1),s)p2ds)\displaystyle\indent\indent+C^{2}(p)\int_{t_{0}}^{t}\|G(X_{s}^{(n)},s)-G(X_{s}^{(n-1)},s)\|_{p}^{2}ds+C^{2}(T)\int_{t_{0}}^{t}\|H(X_{s}^{(n)},s)-H(X_{s}^{(n-1)},s)\|_{p}^{2}ds\Bigg{)}
C(p,R,T)t0t(F(Xs(n),s)F(Xs(n1),s)p2+G(Xs(n),s)G(Xs(n1),s)p2\displaystyle\indent\leq C(p,R,T)\int_{t_{0}}^{t}\Big{(}\|F(X_{s}^{(n)},s)-F(X_{s}^{(n-1)},s)\|_{p}^{2}+\|G(X_{s}^{(n)},s)-G(X_{s}^{(n-1)},s)\|_{p}^{2}
+H(Xs(n),s)H(Xs(n1),s)p2)ds\displaystyle\indent\indent+\|H(X_{s}^{(n)},s)-H(X_{s}^{(n-1)},s)\|_{p}^{2}\Big{)}ds
C(p,R,T)t0tρ(Xs(n)Xs(n1)p2)𝑑s,\displaystyle\indent\leq C(p,R,T)\int_{t_{0}}^{t}\rho\left(\|X_{s}^{(n)}-X_{s}^{(n-1)}\|_{p}^{2}\right)ds,

where C(p,R,T)=3(1C(R))2max{C2(p),C2(T)}C(p,R,T)=\frac{3}{(1-C(R))^{2}}\max\{C^{2}(p),\ C^{2}(T)\}.

Therefore, for any n,k1n,\ k\geq 1, t[t0,T]t\in[t_{0},T],

Xt(n+k)Xt(n)p2C(p,R,T)t0tρ(Xs(n+k1)Xs(n1)p2)𝑑s.\|X_{t}^{(n+k)}-X_{t}^{(n)}\|_{p}^{2}\leq C(p,R,T)\int_{t_{0}}^{t}\rho\left(\|X_{s}^{(n+k-1)}-X_{s}^{(n-1)}\|_{p}^{2}\right)ds.

Since each Xt(n)X_{t}^{(n)} is LpL^{p}-continuous process on [t0,T][t_{0},T] for any n+n\in\mathds{N}^{+}, Xt(n)p\|X_{t}^{(n)}\|_{p} is uniformly bounded on [t0,T][t_{0},T]. Set un,k(t)=sups[t0,t]Xs(n+k)Xs(n)p2u_{n,k}(t)=\sup\limits_{s\in[t_{0},t]}\|X_{s}^{(n+k)}-X_{s}^{(n)}\|_{p}^{2}, t[t0,T]t\in[t_{0},T], which is uniformly bounded. Then

un,k(t)C(p,R,T)t0tρ(un1,k(s))𝑑s.u_{n,k}(t)\leq C(p,R,T)\int_{t_{0}}^{t}\rho(u_{n-1,k}(s))ds.

Let vn(t)=supkun,k(t)v_{n}(t)=\sup\limits_{k}u_{n,k}(t), t[t0,T]t\in[t_{0},T]. Then,

0vn(t)C(p,R,T)t0tρ(vn1(s))𝑑s.0\leq v_{n}(t)\leq C(p,R,T)\int_{t_{0}}^{t}\rho(v_{n-1}(s))ds.

Denote

α(t)=lim supn+vn(t),t0tT.\alpha(t)=\limsup_{n\to+\infty}v_{n}(t),\ t_{0}\leq t\leq T.

Applying Lebesgue dominated convergence theorem, we get

0α(t)C(p,R,T)t0tρ(α(s))𝑑s,t0tT.0\leq\alpha(t)\leq C(p,R,T)\int_{t_{0}}^{t}\rho(\alpha(s))ds,\ t_{0}\leq t\leq T.

Hence, by Lemma 2.6, one deduces

α(t)=0,t0tT,\alpha(t)=0,\ t_{0}\leq t\leq T,

which implies that {Xt(n)}n0\{X_{t}^{(n)}\}_{n\geq 0} is a Cauchy sequence in Lp(𝒞)L^{p}(\mathscr{C}).

Step 3. X()C𝔸(t0,T;Lp(𝒞T))X(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T})) is the solution to QSDE (1.2). Since {Xt(n)}n0\{X_{t}^{(n)}\}_{n\geq 0} is a Cauchy sequence in Lp(𝒞)L^{p}(\mathscr{C}), there exists XtLp(𝒞)X_{t}\in L^{p}(\mathscr{C}) such that for any t[t0,T]t\in[t_{0},T],

limnXt(n)Xtp=0.\lim_{n\to\infty}\|X^{(n)}_{t}-X_{t}\|_{p}=0.

Thus, for any ε>0\varepsilon>0, there exists δ>0\delta>0 such that

Xt1Xt2p\displaystyle\|X_{t_{1}}-X_{t_{2}}\|_{p} =Xt1Xt1(n)+Xt1(n)Xt2(n)+Xt2(n)Xt2p\displaystyle=\|X_{t_{1}}-X_{t_{1}}^{(n)}+X_{t_{1}}^{(n)}-X_{t_{2}}^{(n)}+X_{t_{2}}^{(n)}-X_{t_{2}}\|_{p}
Xt1Xt1(n)p+Xt1(n)Xt2(n)p+Xt2(n)Xt2p\displaystyle\leq\|X_{t_{1}}-X_{t_{1}}^{(n)}\|_{p}+\|X_{t_{1}}^{(n)}-X_{t_{2}}^{(n)}\|_{p}+\|X_{t_{2}}^{(n)}-X_{t_{2}}\|_{p}
<εasn,t1,t2[t0,T]satisfying|t1t2|<δ.\displaystyle<\varepsilon\quad\textrm{as}\ n\to\infty,\ \forall\ t_{1},t_{2}\in[t_{0},T]\ \textrm{satisfying}\ |t_{1}-t_{2}|<\delta.

It shows that XtX_{t} is LpL^{p}-continuous and adapted on [t0,T][t_{0},T] since Xt(n)X_{t}^{(n)} is LpL^{p}-continuous and adapted.

We shall prove that {Xt}tt0\{X_{t}\}_{t\geq t_{0}} is the solution to

Xt=Z+R(Xt)+t0tF(Xs,s)𝑑Ws+t0t𝑑WsG(Xs,s)+t0tH(Xs,s)𝑑s,a.s. in[t0,T].X_{t}=Z+R(X_{t})+\int_{t_{0}}^{t}F(X_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s},s)+\int_{t_{0}}^{t}H(X_{s},s)ds,\ \textrm{a.s.\ in}\ [t_{0},T].

In fact,

t0tF(Xs(n),s)𝑑Wst0tF(Xs,s)𝑑Wsp2\displaystyle\left\|\int_{t_{0}}^{t}F(X_{s}^{(n)},s)dW_{s}-\int_{t_{0}}^{t}F(X_{s},s)dW_{s}\right\|_{p}^{2} C2(p)F(Xs(n),s)F(Xs,s)p([t0,t])2\displaystyle\leq C^{2}(p)\|F(X_{s}^{(n)},s)-F(X_{s},s)\|^{2}_{\mathcal{H}^{p}([t_{0},t])}
C2(p)t0tF(Xs(n),s)F(Xs,s)p2𝑑s\displaystyle\leq C^{2}(p)\int_{t_{0}}^{t}\|F(X_{s}^{(n)},s)-F(X_{s},s)\|_{p}^{2}ds
C2(p)t0tρ(Xs(n)Xsp2)𝑑s,\displaystyle\leq C^{2}(p)\int_{t_{0}}^{t}\rho\left(\|X_{s}^{(n)}-X_{s}\|_{p}^{2}\right)ds,
0,asn,\displaystyle\rightarrow 0,\ \textrm{as}\ n\to\infty,

since Xs(n)XsX_{s}^{(n)}\to X_{s} in Lp(𝒞)L^{p}(\mathscr{C}) for any s[t0,T]s\in[t_{0},T] and ρ\rho is continuous. Similarly,

t0t𝑑WsG(Xs(n),s)t0t𝑑WsG(Xs,s)andt0tH(Xs(n),s)𝑑st0tH(Xs,s)𝑑s\int_{t_{0}}^{t}dW_{s}G(X_{s}^{(n)},s)\rightarrow\int_{t_{0}}^{t}dW_{s}G(X_{s},s)~{}{\rm and}~{}\int_{t_{0}}^{t}H(X_{s}^{(n)},s)ds\rightarrow\int_{t_{0}}^{t}H(X_{s},s)ds

in Lp(𝒞)L^{p}(\mathscr{C}). Because Xs(n)XsX_{s}^{(n)}\rightarrow X_{s} for any s[t0,T]s\in[t_{0},T], the same is true for R(Xs(n))R(Xs)R(X_{s}^{(n)})\rightarrow R(X_{s}).
Taking limits on both sides of (3.1), it deduces that

Xt=\displaystyle X_{t}= limnXt(n+1)\displaystyle\lim_{n\rightarrow\infty}X_{t}^{(n+1)}
=\displaystyle= limn(Z+R(Xt(n+1))+t0tF(Xs(n),s)𝑑Ws+t0t𝑑WsG(Xs(n),s)+t0tH(Xs(n),s)𝑑s)\displaystyle\lim_{n\rightarrow\infty}\left(Z+R(X^{(n+1)}_{t})+\int_{t_{0}}^{t}F(X_{s}^{(n)},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s}^{(n)},s)+\int_{t_{0}}^{t}H(X_{s}^{(n)},s)ds\right)
=\displaystyle= Z+R(Xt)+t0tF(Xs,s)𝑑Ws+t0t𝑑WsG(Xs,s)+t0tH(Xs,s)𝑑s,in[t0,T].\displaystyle Z+R(X_{t})+\int_{t_{0}}^{t}F(X_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s},s)+\int_{t_{0}}^{t}H(X_{s},s)ds,\ \textrm{in}\ [t_{0},T].

That is, {Xt}tt0\{X_{t}\}_{t\geq t_{0}} is a LpL^{p}-solution to the equation (1.2).

Uniqueness: Suppose that Yt,t[t0,T]Y_{t},\ t\in[t_{0},T] is another adapted LpL^{p}-continuous solution with Yt0=Z+R(Y)Y_{t_{0}}=Z+R(Y). Then, by (1.2), we obtain again

Yt=Z+R(Yt)+t0tF(Ys,s)𝑑Ws+t0t𝑑WsG(Ys,s)+t0tH(Ys,s)𝑑s,a.s. in[t0,T].Y_{t}=Z+R(Y_{t})+\int_{t_{0}}^{t}F(Y_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(Y_{s},s)+\int_{t_{0}}^{t}H(Y_{s},s)ds,\ \textrm{a.s.\ in}\ [t_{0},T].

Furthermore,

XtYtp\displaystyle\|X_{t}-Y_{t}\|_{p} R(Xt)R(Yt)p+t0t(F(Xs,s)F(Ys,s))𝑑Wsp\displaystyle\leq\|R(X_{t})-R(Y_{t})\|_{p}+\left\|\int_{t_{0}}^{t}(F(X_{s},s)-F(Y_{s},s))dW_{s}\right\|_{p}
+t0t𝑑Ws(G(Xs,s)G(Ys,s))p+t0t(H(Xs,s)H(Ys,s))𝑑sp.\displaystyle\indent+\left\|\int_{t_{0}}^{t}dW_{s}(G(X_{s},s)-G(Y_{s},s))\right\|_{p}+\left\|\int_{t_{0}}^{t}(H(X_{s},s)-H(Y_{s},s))ds\right\|_{p}.

Continuing to use the same technique as Step 2 of existence, we can yield that

XtYtp2C(p,R,T)t0tρ(XsYsp2)𝑑s,t0tT.\|X_{t}-Y_{t}\|_{p}^{2}\leq C(p,R,T)\int_{t_{0}}^{t}\rho\left(\|X_{s}-Y_{s}\|_{p}^{2}\right)ds,\ t_{0}\leq t\leq T.

It follows that, for any t[t0,T]t\in[t_{0},T],

XtYtp=0, a.s.\|X_{t}-Y_{t}\|_{p}=0,\ \textrm{ a.s.}

This completes the proof. ∎

As described in [1], the Itô product rule dA(χ[0,t))dA(χ[0,t))=dtdA(\chi_{[0,t)})dA^{*}(\chi_{[0,t)})=dt holds for any t0t\geq 0. Based on [6], let

ξt=α1A(uχ[0,t))+α2A(uχ[0,t))\xi_{t}=\alpha_{1}A(u\chi_{[0,t)})+\alpha_{2}A^{*}(u\chi_{[0,t)})

for any t0t\geq 0, the integral 0tf(s)𝑑ξs\int_{0}^{t}f(s)d\xi_{s} defines a quantum martingale for any f𝒮𝔸p(+)f\in{\cal S}_{\mathbb{A}}^{p}(\mathbb{R}^{+}). Next, let A(t):=A(χ[0,t))A(t):=A(\chi_{[0,t)}), we study the properties of the LpL^{p}-solutions to QSDE (1.5) with respect to Brownian motion A(t)A(t) and A(t)A^{*}(t) on the basis of martingale inequalities. From Lemma 2.5 and the canonical anticommutation relation, we can deduce the following martingale inequalities.

Theorem 3.2.

Let f:[0,T]Lp(𝒞)f:\ [0,T]\to L^{p}(\mathscr{C}) be adapted processes with p2p\geq 2. Then, for any t[0,T]t\in[0,T], 0tf(s)𝑑A(s)\int_{0}^{t}f(s)dA(s) and 0t𝑑A(s)f(s)\int_{0}^{t}dA^{*}(s)f(s) are LpL^{p}-martingales and

0tf(s)𝑑A(s)pβp(0tf(s)p2𝑑s)12,\displaystyle\left\|\int_{0}^{t}f(s)dA(s)\right\|_{p}\leq\beta_{p}\left(\int_{0}^{t}\|f(s)\|_{p}^{2}ds\right)^{\frac{1}{2}}, (3.2)
0t𝑑A(s)f(s)pβp(0tf(s)p2𝑑s)12.\displaystyle\left\|\int_{0}^{t}dA^{*}(s)f(s)\right\|_{p}\leq\beta_{p}\left(\int_{0}^{t}\|f(s)\|_{p}^{2}ds\right)^{\frac{1}{2}}.
Proof.

First, we consider simple adapted LpL^{p}-process f𝒮𝔸p([0,T])f\in{\cal S}^{p}_{\mathbb{A}}([0,T]), then 0tf(s)𝑑A(s)\int_{0}^{t}f(s)dA(s) and 0t𝑑A(s)f(s)\int_{0}^{t}dA^{*}(s)f(s) are LpL^{p}-martingales.

Let

0=t0t1t2tn=t0=t_{0}\leq t_{1}\leq t_{2}\leq\ldots\leq t_{n}=t

be a partition of [0,t][0,t]. Then

Q(t)=0tf(s)𝑑A(s)=k=0n1f(tk)(A(tk+1)A(tk)),Q(t)=\int_{0}^{t}f(s)dA(s)=\sum_{k=0}^{n-1}f(t_{k})\left(A(t_{k+1})-A(t_{k})\right),
Q(tk)=i=0k1f(ti)(A(ti+1)A(ti)),k.Q(t_{k})=\sum_{i=0}^{k-1}f(t_{i})\left(A(t_{i+1})-A(t_{i})\right),\ k\in\mathds{N}.

Define the martingale difference of Q(t)Q(t) as

dQk=Q(tk+1)Q(tk)=f(tk)(A(tk+1)A(tk)),k+.dQ_{k}=Q(t_{k+1})-Q(t_{k})=f(t_{k})(A(t_{k+1})-A(t_{k})),\ k\in\mathds{N}^{+}.

By Theorem 2.1 of [23], there exists a positive constant βp\beta_{p} such that

0tf(s)𝑑A(s)pβpmax{(k0|dQk|2)12p,(k0|dQk|2)12p}.\left\|\int_{0}^{t}f(s)dA(s)\right\|_{p}\leq\beta_{p}\max\left\{\left\|\left(\sum_{k\geq 0}|dQ_{k}|^{2}\right)^{\frac{1}{2}}\right\|_{p},\quad\left\|\left(\sum_{k\geq 0}|dQ^{*}_{k}|^{2}\right)^{\frac{1}{2}}\right\|_{p}\right\}. (3.3)

By the canonical anticommutation relation

A(t)A(t)+A(t)A(t)=t,t0,A(t)A^{*}(t)+A^{*}(t)A(t)=t,\ t\geq 0,

one has

(A(t)A(s))(A(t)A(s))ts,(A(t)A(s))(A(t)A(s))ts, 0stT.(A(t)-A(s))(A^{*}(t)-A^{*}(s))\leq t-s,\ (A^{*}(t)-A^{*}(s))(A(t)-A(s))\leq t-s,\ 0\leq s\leq t\leq T. (3.4)

According to (2.11), f=fe+fof=f_{e}+f_{o} for any fLp(𝒞)f\in L^{p}(\mathscr{C}), then

k0|dQk|2=\displaystyle\sum_{k\geq 0}|dQ_{k}|^{2}= k=0n1(A(tk+1)A(tk))f(tk)f(tk)(A(tk+1)A(tk))\displaystyle\sum_{k=0}^{n-1}\left(A^{*}(t_{k+1})-A^{*}(t_{k})\right)f^{*}(t_{k})f(t_{k})\left(A(t_{k+1})-A(t_{k})\right) (3.5)
=\displaystyle= k=0n1(fe(tk)fo(tk))(A(tk+1)A(tk))(A(tk+1)A(tk))(fe(tk)fo(tk))\displaystyle\sum_{k=0}^{n-1}\left(f^{*}_{e}(t_{k})-f^{*}_{o}(t_{k})\right)\left(A^{*}(t_{k+1})-A^{*}(t_{k})\right)\left(A(t_{k+1})-A(t_{k})\right)\left(f_{e}(t_{k})-f_{o}(t_{k})\right)
\displaystyle\leq k=0n1(fe(tk)fo(tk))(tk+1tk)(fe(tk)fo(tk))\displaystyle\sum_{k=0}^{n-1}\left(f^{*}_{e}(t_{k})-f^{*}_{o}(t_{k})\right)(t_{k+1}-t_{k})\left(f_{e}(t_{k})-f_{o}(t_{k})\right)
=\displaystyle= 0t|fe(s)fo(s)|2𝑑s,\displaystyle\int_{0}^{t}|f_{e}(s)-f_{o}(s)|^{2}ds,

and

k0|dQk|2=\displaystyle\sum_{k\geq 0}|dQ^{*}_{k}|^{2}= k=0n1f(tk)(A(tk+1)A(tk))(A(tk+1)A(tk))f(tk)\displaystyle\sum_{k=0}^{n-1}f(t_{k})(A(t_{k+1})-A(t_{k}))(A^{*}(t_{k+1})-A^{*}(t_{k}))f^{*}(t_{k}) (3.6)
\displaystyle\leq k=0n1f(tk)f(tk)(tk+1tk)\displaystyle\sum_{k=0}^{n-1}f(t_{k})f^{*}(t_{k})(t_{k+1}-t_{k})
=\displaystyle= 0t|f(s)|2𝑑s,\displaystyle\int_{0}^{t}|f^{*}(s)|^{2}ds,

where the above two inequalities are based on Lemma 2.5 and (3.4).

Substituting (3.5) and (3.6) into the right side of (3.3) and applying Theorem 2.3, we get

0tf(s)𝑑A(s)p\displaystyle\left\|\int_{0}^{t}f(s)dA(s)\right\|_{p} βpmax{(0t|f(s)|2𝑑s)12p,(0t|fe(s)fo(s)|2𝑑s)12p}\displaystyle\leq\beta_{p}\max\left\{\left\|\left(\int_{0}^{t}|f^{*}(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p},\ \left\|\left(\int_{0}^{t}|f_{e}(s)-f_{o}(s)|^{2}ds\right)^{\frac{1}{2}}\right\|_{p}\right\}
βpmax{(0tf(s)p2𝑑s)12,(0tfe(s)fo(s)p2𝑑s)12}.\displaystyle\leq\beta_{p}\max\left\{\left(\int_{0}^{t}\|f^{*}(s)\|_{p}^{2}ds\right)^{\frac{1}{2}},\ \left(\int_{0}^{t}\|f_{e}(s)-f_{o}(s)\|_{p}^{2}ds\right)^{\frac{1}{2}}\right\}.

On the other hand, for any f𝒮𝔸p([0,T])f\in{\cal S}^{p}_{\mathbb{A}}([0,T]) and any s[0,T]s\in[0,T],

fe(s)fo(s)p2f(s)p,f(s)p=f(s)p.\|f_{e}(s)-f_{o}(s)\|_{p}\leq 2\|f(s)\|_{p},\ \|f^{*}(s)\|_{p}=\|f(s)\|_{p}.

Thus we have

0tf(s)𝑑A(s)pβp(0tf(s)p2𝑑s)12, 0tT.\left\|\int_{0}^{t}f(s)dA(s)\right\|_{p}\leq\beta_{p}\left(\int_{0}^{t}\|f(s)\|_{p}^{2}ds\right)^{\frac{1}{2}},\ 0\leq t\leq T. (3.7)

Similarly, one has

0t𝑑A(s)f(s)pβp(0tf(s)p2𝑑s)12, 0tT.\left\|\int_{0}^{t}dA^{*}(s)f(s)\right\|_{p}\leq\beta_{p}\left(\int_{0}^{t}\|f(s)\|_{p}^{2}ds\right)^{\frac{1}{2}},\ 0\leq t\leq T. (3.8)

Finally, since the general adapted LpL^{p}-processes can be approximated by simple processes, (3.2) can be directly obtained from (3.7) and (3.8). ∎

Clearly, by virtue of Theorem 3.1 and Theorem 3.2, the following result holds.

Corollary 3.3.

Let Assumption 1.1 hold. Then there is a unique solution X()C𝔸(t0,T;Lp(𝒞T))X(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T})) to the equation (1.5) with nonlocal condition Xt0=Z+R(X)Lp(𝒞t0)X_{t_{0}}=Z+R(X)\in L^{p}(\mathscr{C}_{t_{0}}) on [t0,T][t_{0},T].

4 The stability of solutions to QSDE with Lipschitz condition

In this section, we shall prove that the LpL^{p}-solution to the equation (1.2) is stable, namely, small changes in the initial condition and in the coefficients F,G,HF,\ G,\ H and RR lead to small changes in the solution on [t0,T][t_{0},T].

Let the coefficients F,GF,\ G, HH of the equation (1.2) satisfy Lipsctitz condition, i.e.
(A3’) For any x1,x2Lp(𝒞)x_{1},\ x_{2}\in L^{p}(\mathscr{C}) and a.e. t[0,T]t\in[0,T], there exists a constant L>0L>0 such that

F(x1,t)F(x2,t)p2+G(x1,t)G(x2,t)p2+H(x1,t)H(x2,t)p2Lx1x2p2.\|F(x_{1},t)-F(x_{2},t)\|_{p}^{2}+\|G(x_{1},t)-G(x_{2},t)\|_{p}^{2}+\|H(x_{1},t)-H(x_{2},t)\|_{p}^{2}\leq L\|x_{1}-x_{2}\|_{p}^{2}.

Let {Xt}tt0\{X_{t}\}_{t\geq t_{0}}, {Yt}tt0\{Y_{t}\}_{t\geq t_{0}} be the LpL^{p}-solution to the equation (1.2) with initial conditions Xt0=Z+R(X)X_{t_{0}}=Z+R(X) and Yt0=Z+R(Y)Y_{t_{0}}=Z^{{}^{\prime}}+R(Y) for any Xt0,Yt0Lp(𝒞t0)X_{t_{0}},\ Y_{t_{0}}\in L^{p}(\mathscr{C}_{t_{0}}), respectively. That is,

Xt=Z+R(Xt)+t0tF(Xs,s)𝑑Ws+t0t𝑑WsG(Xs,s)+t0tH(Xs,s)𝑑s, a.s. in[t0,T],X_{t}=Z+R(X_{t})+\int_{t_{0}}^{t}F(X_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(X_{s},s)+\int_{t_{0}}^{t}H(X_{s},s)ds,\ \textrm{ a.s.\ in}\ [t_{0},T],

and

Yt=Z+R(Yt)+t0tF(Ys,s)𝑑Ws+t0t𝑑WsG(Ys,s)+t0tH(Ys,s)𝑑s, a.s. in[t0,T].Y_{t}=Z^{\prime}+R(Y_{t})+\int_{t_{0}}^{t}F(Y_{s},s)dW_{s}+\int_{t_{0}}^{t}dW_{s}G(Y_{s},s)+\int_{t_{0}}^{t}H(Y_{s},s)ds,\ \textrm{ a.s.\ in}\ [t_{0},T].
Theorem 4.1.

Suppose that assumptions (A1),(A2),(A3’),(A4) hold. With the above notations, for any ε>0\varepsilon>0, there exists δ>0\delta>0 such that if ZZp<δ\|Z-Z^{{}^{\prime}}\|_{p}<\delta, then XtYtp<ε\|X_{t}-Y_{t}\|_{p}<\varepsilon holds for all t0tTt_{0}\leq t\leq T.

Proof.

By the directly calculation,

XtYtp\displaystyle\|X_{t}-Y_{t}\|_{p} ZZp+R(Xt)R(Yt)p+t0t(F(Xs,s)F(Ys,s))𝑑Wsp\displaystyle\leq\|Z-Z^{\prime}\|_{p}+\|R(X_{t})-R(Y_{t})\|_{p}+\left\|\int_{t_{0}}^{t}(F(X_{s},s)-F(Y_{s},s))dW_{s}\right\|_{p}
+t0t𝑑Ws(G(Xs,s)G(Ys,s))p+t0t(H(Xs,s)H(Ys,s))𝑑sp.\displaystyle\indent+\left\|\int_{t_{0}}^{t}dW_{s}(G(X_{s},s)-G(Y_{s},s))\right\|_{p}+\left\|\int_{t_{0}}^{t}(H(X_{s},s)-H(Y_{s},s))ds\right\|_{p}.

According to the proof of Theorem 3.1 again and (A3)(A3^{\prime}), for any t[t0,T]t\in[t_{0},T], one gains the following estimate

XtYtp2\displaystyle\|X_{t}-Y_{t}\|_{p}^{2}\leq 4(1C(R))2ZZp2+C(p,R,T)(t0tF(Xs,s)F(Ys,s)p2ds\displaystyle\frac{4}{(1-C(R))^{2}}\|Z-Z^{{}^{\prime}}\|_{p}^{2}+C^{\prime}(p,R,T)\Bigg{(}\int_{t_{0}}^{t}\|F(X_{s},s)-F(Y_{s},s)\|_{p}^{2}ds
+t0tG(Xs,s)G(Ys,s)p2ds+t0tH(Xs,s)H(Ys,s)p2ds)\displaystyle+\int_{t_{0}}^{t}\|G(X_{s},s)-G(Y_{s},s)\|_{p}^{2}ds+\int_{t_{0}}^{t}\|H(X_{s},s)-H(Y_{s},s)\|_{p}^{2}ds\Bigg{)}
\displaystyle\leq 4(1C(R))2ZZp2+C(p,T,R,L)t0tXsYsp2𝑑s,\displaystyle\frac{4}{(1-C(R))^{2}}\|Z-Z^{{}^{\prime}}\|_{p}^{2}+C(p,T,R,L)\int_{t_{0}}^{t}\|X_{s}-Y_{s}\|_{p}^{2}ds,

where C(p,T,R,L)=C(p,R,T)L=4(1C(R))2max{C2(p),C2(T)}LC(p,T,R,L)=C^{\prime}(p,R,T)L=\frac{4}{(1-C(R))^{2}}\max\{C^{2}(p),C^{2}(T)\}L.

By Gronwall’s inequality,

XtYtp24(1C(R))2eC(p,T,R,L)(tt0)ZZp2,\|X_{t}-Y_{t}\|_{p}^{2}\leq\frac{4}{(1-C(R))^{2}}e^{C(p,T,R,L)(t-t_{0})}\|Z-Z^{{}^{\prime}}\|_{p}^{2},

for all t[t0,T]t\in[t_{0},T], and the desired result is obtained. ∎

In a similar manner, we establish stability theorems for the connection between coefficient convergence and solution convergence under Lipschitz condition.

Theorem 4.2.

Let assumptions (A1),(A2),(A3’),(A4) hold with F,G,H,RF,\ G,\ H,\ R being replaced respectively by Fn,GnF_{n},\ G_{n}, Hn,RnH_{n},\ R_{n}, for all n=1, 2,n=1,\ 2,\cdots and WtW_{t} be as in the equation (1.2). Assume that FnF,GnG,HnHF_{n}\rightarrow F,\ G_{n}\rightarrow G,\ H_{n}\rightarrow H in Lp(𝒞)L^{p}(\mathscr{C}) as nn\to\infty, uniformly on Lp(𝒞)×[t0,T]L^{p}(\mathscr{C})\times[t_{0},T], RnRR_{n}\rightarrow R in Lp(𝒞)L^{p}(\mathscr{C}) uniformly as nn\to\infty on Lp(𝒞)L^{p}(\mathscr{C}), and ZnZZ_{n}\rightarrow Z in Lp(𝒞t0)L^{p}(\mathscr{C}_{t_{0}}). Furthermore, let X(),Xn()C𝔸(t0,T;Lp(𝒞T))X(\cdot),\ X_{n}(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T})) be solutions to the equation (1.2) corresponding to F,G,H,R,ZF,\ G,\ H,\ R,\ Z and Fn,Gn,Hn,Rn,ZnF_{n},\ G_{n},\ H_{n},\ R_{n},\ Z_{n}, respectively. Then Xn(t)X(t)X_{n}(t)\rightarrow X(t) in Lp(𝒞)L^{p}(\mathscr{C}) uniformly on compact set [t0,T][t_{0},T].

Likewise, by Lemma 3.2 and Corollary 3.3, we could get the following result.

Corollary 4.3.

Suppose that assumptions (A1),(A2),(A3’),(A4) hold. Then the solution X()C𝔸(t0,T;Lp(𝒞T)X(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T}) to the equation (1.5) is stable on [t0,T][t_{0},T] when nonlocal condition Xt0=Z+R(X)Lp(𝒞t0)X_{t_{0}}=Z+R(X)\in L^{p}(\mathscr{C}_{t_{0}}) and the coefficients change slightly, respectively.

5 The Self-adjointness and Markov Property

In this section, we consider the self-adjointness and Markov property of LpL^{p}-solutions to QSDE (1.1) with nonlocal conditions under non-Lipschitz coefficients for p>2p>2.

According to the description of parity in Section 2, we get the following lemma.

Lemma 5.1.

Let F:[0,t]Lp(𝒞)F:[0,t]\rightarrow L^{p}(\mathscr{C}) be adapted and satisfy 0tF(s)p2𝑑s<\int_{0}^{t}\|F(s)\|_{p}^{2}ds<\infty. Suppose further that F(s)=F(s)Lp(𝒞e)F(s)=F(s)^{*}\in L^{p}(\mathscr{C}_{e}) for each s[0,t]s\in[0,t]. Then 0tF(s)𝑑Ws\int_{0}^{t}F(s)dW_{s} is self-adjoint element of Lp(𝒞)L^{p}(\mathscr{C}), and 0tF(s)𝑑Ws=0t𝑑WsF(s)\int_{0}^{t}F(s)dW_{s}=\int_{0}^{t}dW_{s}F(s).

Proof.

It is sufficient to consider the case that F(s)F(s) is simple with values in \mathscr{E} for any s[0,t]s\in[0,t]. Since FF is simple, F(s)=k=0n1F(tk)χ[tk,tk+1)(s)F(s)=\sum\limits_{k=0}^{n-1}F(t_{k})\chi_{[t_{k},t_{k+1})}(s) and

0tF(s)𝑑Ws=k=0n1F(tk)(Wtk+1Wtk),\int_{0}^{t}F(s)dW_{s}=\sum_{k=0}^{n-1}F(t_{k})(W_{t_{k+1}}-W_{t_{k}}),

where {tk}k=0n\{t_{k}\}_{k=0}^{n} is a partition of [0,t][0,t]. On the other hand, F(s)=F(s)F(s)=F(s)^{*} and WsW_{s} is hermitian,

(0tF(s)𝑑Ws)\displaystyle\left(\int_{0}^{t}F(s)dW_{s}\right)^{*} =(k=0n1F(tk)(Wtk+1Wtk))\displaystyle=\left(\sum_{k=0}^{n-1}F(t_{k})(W_{t_{k+1}}-W_{t_{k}})\right)^{*}
=k=0n1(Wtk+1Wtk)F(tk)\displaystyle=\sum_{k=0}^{n-1}\left(W_{t_{k+1}}-W_{t_{k}}\right)^{*}F(t_{k})^{*}
=k=0n1(Wtk+1Wtk)F(tk)\displaystyle=\sum_{k=0}^{n-1}\left(W_{t_{k+1}}-W_{t_{k}}\right)F(t_{k})
=0t𝑑WsF(s).\displaystyle=\int_{0}^{t}dW_{s}F(s).

By Lemma 2.5,

(Wtk+1Wtk)F(tk)=F(tk)(Wtk+1Wtk),F(tk)Lp(𝒞e).\left(W_{t_{k+1}}-W_{t_{k}}\right)F(t_{k})=F(t_{k})\left(W_{t_{k+1}}-W_{t_{k}}\right),\ \forall\ F(t_{k})\in L^{p}(\mathscr{C}_{e}).

Then 0t𝑑WsF(s)=0tF(s)𝑑Ws\int_{0}^{t}dW_{s}F(s)=\int_{0}^{t}F(s)dW_{s}. Since 0tF(s)p2𝑑s<\int_{0}^{t}\|F(s)\|_{p}^{2}ds<\infty, 0tF(s)𝑑Ws\int_{0}^{t}F(s)dW_{s} is self-adjoint element in Lp(𝒞)L^{p}(\mathscr{C}) by virtue of Corollary 2.4. ∎

Let Lp(𝒞)saL^{p}(\mathscr{C})_{sa} denote the self-adjoint part of Lp(𝒞)L^{p}(\mathscr{C}). Let Fi,Gi:,F_{i},\ G_{i}:\mathbb{R}\to\mathbb{R}, for i=1,2i=1,2, and satisfy |t||Fi^(t)|𝑑t<,|t||G^i(t)|𝑑t<\int|t||\widehat{F_{i}}(t)|dt<\infty,\ \int|t||\widehat{G}_{i}(t)|dt<\infty. Suppose that Fi,Gi:Lp(𝒞)saLp(𝒞)F_{i},\ G_{i}:L^{p}(\mathscr{C})_{sa}\to L^{p}(\mathscr{C}) are adapted and satisfy the Osgood condition of Assumption 1.1 on Lp(𝒞)L^{p}(\mathscr{C}), and each FiF_{i} is an even function. Set

F~i(h)=Fi(ho),G~i(h)=G(he),hLp(𝒞)sa.\widetilde{F}_{i}(h)=F_{i}(h_{o}),\ \widetilde{G}_{i}(h)=G(h_{e}),\ \forall\ h\in L^{p}(\mathscr{C})_{sa}.

Evidently, F~i(h),G~i(h)\widetilde{F}_{i}(h),\ \widetilde{G}_{i}(h) (i=1,2i=1,2) are even by Lemma 4.1 of [4] for any hLp(𝒞)sah\in L^{p}(\mathscr{C})_{sa}. Let

Φ~i=F~i+G~i,i=1,2.\widetilde{\Phi}_{i}=\widetilde{F}_{i}+\widetilde{G}_{i},\ i=1,2.

It can be seen that Φ~i\widetilde{\Phi}_{i} satisfies the Osgood conditions and maps self-adjoint elements of Lp(𝒞)L^{p}(\mathscr{C}) into self-adjoint elements of Lp(𝒞e)L^{p}(\mathscr{C}_{e}). Then we obtain the self-adjointness of the solutions to QSDE with nonlocal conditions.

Theorem 5.2.

Let Φ~1,Φ~2\widetilde{\Phi}_{1},\ \widetilde{\Phi}_{2} be as above. Let H~:\widetilde{H}:\mathbb{R}\to\mathbb{R} satisfy |t||H~^(t)|𝑑t<\int|t||\widehat{\widetilde{H}}(t)|dt<\infty, and H~:Lp(𝒞)saLp(𝒞)\widetilde{H}:L^{p}(\mathscr{C})_{sa}\rightarrow L^{p}(\mathscr{C}) be adapted and satisfy assumption (A3)(A3) on Lp(𝒞)L^{p}(\mathscr{C}) in Assumption 1.1. Furthermore, R~:Lp(𝒞)saLp(𝒞)sa\widetilde{R}:L^{p}(\mathscr{C})_{sa}\rightarrow L^{p}(\mathscr{C})_{sa} satisfies assumption (A4)(A4) in Assumption 1.1. Then, for any Z=ZZ=Z^{*}, there is a unique self-adjoint, adapted, LpL^{p}-continuous solution {Xt}tt0\{X_{t}\}_{t\geq t_{0}} to the following QSDE

dXt=Φ~1(Xt)dWt+dWtΦ~2(Xt)+H~(Xt)dtdX_{t}=\widetilde{\Phi}_{1}(X_{t})dW_{t}+dW_{t}\widetilde{\Phi}_{2}(X_{t})+\widetilde{H}(X_{t})dt (5.1)

on [t0,T][t_{0},T] with Xt0=Z+R~(X)Lp(𝒞t0)X_{t_{0}}=Z+\widetilde{R}(X)\in L^{p}(\mathscr{C}_{t_{0}}) provided 0<C(R~)<10<C(\widetilde{R})<1.

Proof.

Since Φ~1,Φ~2,H~\widetilde{\Phi}_{1},\ \widetilde{\Phi}_{2},\ \widetilde{H} satisfy the Osgood condition and R~\widetilde{R} satisfy the Lipschitz condition as in Assumption 1.1, it follows from Theorem 3.1 that the equation (5.1) admits a unique solution X()C𝔸(t0,T;Lp(𝒞T))X(\cdot)\in C_{\mathbb{A}}(t_{0},T;L^{p}(\mathscr{C}_{T})) such that

Xt=Z+R~(Xt)+t0tΦ~1(Xs)𝑑Ws+t0t𝑑WsΦ~2(Xs)+t0tH~(Xs)𝑑s,a.s. in[t0,T].X_{t}=Z+\widetilde{R}(X_{t})+\int_{t_{0}}^{t}\widetilde{\Phi}_{1}(X_{s})dW_{s}+\int_{t_{0}}^{t}dW_{s}\widetilde{\Phi}_{2}(X_{s})+\int_{t_{0}}^{t}\widetilde{H}(X_{s})ds,\ \textrm{a.s.\ in}\ [t_{0},T].

Next, it is enough to prove the self-adjointness of the solution to the equation (5.1). We can define the following equation inductively with Xt0=Z+R~(X)X_{t_{0}}=Z+\widetilde{R}(X),

Xt(n+1)=Z+R~(Xt(n+1))+t0tΦ~1(Xs(n))𝑑Ws+t0t𝑑WsΦ~2(Xs(n))+t0tH~(Xs(n))𝑑s.X_{t}^{(n+1)}=Z+\widetilde{R}(X_{t}^{(n+1)})+\int_{t_{0}}^{t}\widetilde{\Phi}_{1}(X_{s}^{(n)})dW_{s}+\int_{t_{0}}^{t}dW_{s}\widetilde{\Phi}_{2}(X_{s}^{(n)})+\int_{t_{0}}^{t}\widetilde{H}(X_{s}^{(n)})ds. (5.2)

To prove the self-adjointness of XtX_{t}, it is sufficient to show that Xt(n+1)X_{t}^{(n+1)} is self-adjoint by induction for all n0n\geq 0. It is obvious that Xt(1)X_{t}^{(1)} is self-adjoint since Z=ZZ=Z^{*}. Assume that Xt(n)X_{t}^{(n)} is self-adjoint, then Φ~i(Xs(n))Lp(𝒞e)Lp(𝒞s)sa\widetilde{\Phi}_{i}(X_{s}^{(n)})\in L^{p}(\mathscr{C}_{e})\cap L^{p}(\mathscr{C}_{s})_{sa}. By Lemma 5.1, t0tΦ~1(Xs(n))𝑑Ws\int_{t_{0}}^{t}\widetilde{\Phi}_{1}(X_{s}^{(n)})dW_{s} and t0t𝑑WsΦ~2(Xs(n))\int_{t_{0}}^{t}dW_{s}\widetilde{\Phi}_{2}(X_{s}^{(n)}) are self-adjoint. In addition, 0tH(Xs(n))𝑑s\int_{0}^{t}H(X^{(n)}_{s})ds and R~(Xs(n+1))\widetilde{R}(X_{s}^{(n+1)}) are also self-adjoint. Hence Xt(n+1)X_{t}^{(n+1)} is self-adjoint. ∎

This result of self-adjoint of solutions is the basis of studying optimal control problem of QSDE. Apart from this, we also obtain the following Markov property of solutions to the equation (1.2) under non-Lipschitz coefficients consistent with Theorem 2.2, Corollary 2.3 and Corollary 2.4 of [5].

For any interval I[t0,)\textit{I}\subseteq[t_{0},\infty), let 𝒜I\mathscr{A}_{\textit{I}} denote the WW^{*}-algebra generated by 𝟙\mathds{1} and the solution XtX_{t} to the equation (1.2) for tIt\in\textit{I}, and write 𝒜s\mathscr{A}_{s} for 𝒜[s,s]\mathscr{A}_{[s,s]}. Since the solution XtX_{t} is adapted, i.e. XtLp(𝒞t)X_{t}\in L^{p}(\mathscr{C}_{t}) for all tt0t\geq t_{0}, it follows that 𝒜I\mathscr{A}_{\textit{I}} is a WW^{*}-subalgebra of 𝒞t\mathscr{C}_{t} whenever I[t0,t]\textit{I}\subseteq[t_{0},t]. Let 𝒜~I=𝒜Iβ(𝒜I)\mathscr{\tilde{A}}_{\textit{I}}=\mathscr{A}_{\textit{I}}\vee\beta(\mathscr{A}_{\textit{I}}) be the WW^{*}-subalgebra of 𝒞\mathscr{C} generated by 𝒜I\mathscr{A}_{\textit{I}} and β(𝒜I)\beta(\mathscr{A}_{\textit{I}}). It is clear that β(𝒜~I)=𝒜~\beta(\mathscr{\tilde{A}}_{\textit{I}})=\mathscr{\tilde{A}} and 𝒜~s𝒞s\mathscr{\tilde{A}}_{s}\subseteq\mathscr{C}_{s} for any st0s\geq t_{0}.

Next, we denote the algebra generated by field differences. Let s\mathscr{F}_{s} denote the WW^{*}-subalgebra of 𝒞\mathscr{C} generated by the field differences {WtWs:t0st}\{W_{t}-W_{s}:\ t_{0}\leq s\leq t\}, and 𝒜~ss\mathscr{\tilde{A}}_{s}\vee\mathscr{F}_{s} be the WW^{*}-subalgebra of 𝒞\mathscr{C} generated by 𝒜~s\mathscr{\tilde{A}}_{s} and s\mathscr{F}_{s}. Thus, β(𝒜~ss)=𝒜~ss\beta(\mathscr{\tilde{A}}_{s}\vee\mathscr{F}_{s})=\mathscr{\tilde{A}}_{s}\vee\mathscr{F}_{s}. Then, we get the following Markov property of the adapted solution {Xt}tt0\{X_{t}\}_{t\geq t_{0}} to the equation (1.2).

Theorem 5.3.

Let assumption 1.1 hold and {Xt}tt0\{X_{t}\}_{t\geq t_{0}} be an adapted, unique, continuous LpL^{p}-solution to the equation (1.2). Then XsLp(𝒜~ss)X_{s}\in L^{p}(\mathscr{\tilde{A}}_{s}\vee\mathscr{F}_{s}) for all t0stt_{0}\leq s\leq t. Moreover, the process {Xt}tt0\{X_{t}\}_{t\geq t_{0}} is a Markov process in the following sense: for any st0s\geq t_{0} and fLp(𝒜~[s,)),f\in L^{p}(\mathscr{\tilde{A}}_{[s,\infty)}), one has

m(f|𝒜~[t0,s])=m(f|𝒜~s),m(f|\mathscr{\tilde{A}}_{[t_{0},s]})=m(f|\mathscr{\tilde{A}}_{s}), (5.3)

where m(|)m(\cdot|\mathscr{B}) denotes the conditional expectation with respect to the subalgebra \mathscr{B} of 𝒞\mathscr{C}.

The proof of Theorem 5.3 is similar to Theorem 2.2 of [5]. Furthermore, the result for the solution of the equation (1.5) also holds.

6 Conclusion

Utilizing the Burkholder-Gundy inequality of Clifford LpL^{p}-martingale, we obtain the existence, uniqueness and stability of the solutions to QSDEs with nonlocal conditions for p>2p>2. In addition, the acquisition of the self-adjoint solution pave the way for the next study on the optimal control problems of QSDE.

References

  • [1] Applebaum, D.B. R.L. Hudson, Fermion Itô’s formula and stochastic evolutions, Comm. Math. Phys. 96(4), 473-496 (1984).
  • [2] D.B. Applebaum, R.L. Hudson, Fermion diffusions, J. Math. Phys. 25(4), 858-861 (1984).
  • [3] C. Barnett, R.F. Streater, I.F. Wilde, The Itô-Clifford integral, J. Funct. Anal. 48(2), 172-212 (1982).
  • [4] C. Barnett, R.F. Streater, I.F. Wilde, The Itô-Clifford integral. II. Stochastic differential equations, J. Lond. Math. Soc. 27(2), 373-384 (1983).
  • [5] C. Barnett, R.F. Streater, I.F. Wilde, The Itô-Clifford integral. III. The Markov property of solutions to stochastic differential equations, Comm. Math. Phys. 89(1), 13-17 (1983).
  • [6] C. Barnett, R.F. Streater, I.F. Wilde, Quasifree quantum stochastic integrals for the CAR and CCR, J. Funct. Anal. 52(1), 19-47 (1983).
  • [7] E.A. Carlen, P. Krée, On martingale inequalities in non-commutative stochastic analysis, J. Funct. Anal. 158(2), 475-508 (1998).
  • [8] S. Dirksen, Noncommutative stochastic integration through decoupling, J. Math. Anal. Appl. 370(1), 200-223 (2010).
  • [9] R.L. Hudson, The early years of quantum stochastic calculus, Commun. Stoch. Anal. 6(1), 111-123 (2012).
  • [10] R.L. Hudson, J.M. Lindsay, A noncommutative martingale representation theorem for non-Fock quantum Brownian motion, J. Funct. Anal. 61(2), 202-221 (1985).
  • [11] G. Hong, T. Mei, John-Nirenberg inequality and atomic decomposition for noncommutative martingales, J. Funct. Anal. 263(4), 1064-1097 (2012).
  • [12] R.L. Hudson, K.R. Parthasarathy, Quantum Itô’s formula and stochastic evolutions, Comm. Math. Phys. 93(3), 301-323 (1984).
  • [13] M. Junge, Doob’s inequality for noncommutative martingales, J. Reine Angew. Math. 549, 149-190 (2002).
  • [14] M. Junge, Q. Xu, Noncommutative Burkholder/Rosenthal inequalities, Ann. Probab. 31, 948-995 (2003).
  • [15] M. Junge, Q. Xu, On the best constants in some non-commutative martingale inequalities, Bull. London Math. Soc. 37(2), 243-253 (2005).
  • [16] M. Junge, Q. Xu, Noncommutative Burkholder/Rosenthal inequalities. II: Applications, Israel J. Math. 167(1), 227-282 (2008).
  • [17] M. Junge, Q. Xu, Noncommutative maximal ergodic theorems, J. Amer. Math. Soc. 20(2), 385-439 (2008).
  • [18] J.M. Lindsay, Fermion martingales, Probab. Theory Related Fields. 71(2), 307-320 (1986).
  • [19] J.M. Lindsay, I.F. Wilde, On non-Fock boson stochastic integrals, J. Funct. Anal. 65(1), 76-82 (1986).
  • [20] F. Lust-Piquard, G. Pisier, Noncommutative Khintchine and Paley inequalities, Ark. Mat. 29(2), 241-260 (1991).
  • [21] X. Mao, Stochastic Differential Equations and Their Applications, Horwood Publishing, Chichester (1997).
  • [22] K.R. Parthasarathy, An introduction to quantum stochastic calculus, Monographs in Mathematics, (1992).
  • [23] G. Pisier, Q. Xu, Non-Commutative Martingale Inequalities, Comm. Math. Phys. 189, 667-698 (1997).
  • [24] N. Randrianantoanina, A weak type inequality for non-commutative martingales and applications, Proc. Lond. Math. Soc. 91(2), 509-542 (2005).
  • [25] N. Randrianantoanina, Conditioned square functions for noncommutative martingales, Ann. Probab. 35(3), 1039-1070 (2007).
  • [26] I.E. Segal, A Non-Commutative Extension of Abstract Integration, Ann. of Math. 57(3), 401-457 (1953).
  • [27] I.E. Segal, Tensor algebras over Hilbert spaces. I, Trans. Amer. Math. Soc. 81, 106-134 (1956).
  • [28] I.E. Segal, Tensor algebras over Hilbert spaces. II, Ann. of Math. 63(1), 160-175 (1956).
  • [29] I.E. Segal, Algebraic integration theory, Bull. Amer. Math. Soc. 71, 419-489 (1965).
  • [30] M. Takesaki, Conditional expectation in von Neumann algebras, J. Funct. Anal. 9, 306-321 (1972).
  • [31] Q. Xu, T.N. Bekjan, Z. Chen, Introduction to Operator Algebra and Non-commutative LpL^{p} Space, vol. 134, Sciencep, Beijing, (2010).
  • [32] I.F. Wilde, The free fermion field as a Markov field, J. Funct. Anal. 15, 12-21 (1974).
  • [33] I.F. Wilde, Quantum martingales and stochastic integrals, Quantum probability and applications, III (Oberwolfach, 1987), Lecture Notes in Math. 1303, 363-373 (1988).