-solutions of QSDE driven by Fermion fields with nonlocal conditions under non-Lipschitz coefficients
Abstract. The main purpose of this paper is to obtain the existence and uniqueness of -solution to quantum stochastic differential equation driven by Fermion fields with nonlocal conditions in the case of non-Lipschitz coefficients for . 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 -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]:
(1.1) |
driven by Fermion fields which is closely related to the quantum noise, quantum fields etc. The -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 -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 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 -spaces. Problems with nonlocal conditions are more applicable to real life problems than problems with traditional local conditions. Analogously, we study the -solution and the basic properties of solution to QSDE with nonlocal initial conditions for . To state our results, we give some assumptions on the QSDE. Some basic notations of noncommutative filtration will be introduced in Section 2.
Definition 1.1.
A map is said to be adapted if for each . A map is said to be adapted if , for any and .
It is easy to check that if and are both adapted, so is the map .
In the rest of this section, we consider the following equation with nonlocal conditions in the interval for some fixed ,
(1.2) |
where , constitutes the nonlocal condition, and for fixed .
Definition 1.2.
A stochastic process is called a solution to the equation (1.2) if it satisfies
Throughout the paper, we shall make use of the following conditions:
Assumption 1.1.
: are operator-valued functions such that
- (A1)
-
are adapted.
- (A2)
-
For any , are continuous a.s.
- (A3)
-
Osgood condition: For any and a.e. ,
where is a continuous non-decreasing function with , for , such that .
- (A4)
-
is continuous and adapted, and there exists a constant such that
(1.3)
Theorem 1.1.
Under the assumptions (A1)-(A4), for , there is a unique continuous adapted -solution to the following quantum stochastic integral equation with nonlocal initial value
(1.4) |
Remark 1.1.
There are many efforts to study the solution to the QSDE (1.1) by many mathematicians.
- (1).
-
(2).
The original study [1] on the following QSDE
(1.5) is to consider the solution in the weak sense, i.e., is called the solution to the initial value problem of the QSDE, if
for any , where can be considered as an unbounded operator densely defined on the Hilbert space with the domain .
-
(3).
The martingale inequality and Burkholder-Gundy inequality of Pisier and Xu have been used by Dirksen [8] to study the -solution for to QSDE with respect to any normal -martingale without the drift term . The reason why the Burkholder-Gundy inequality can be used directly is because, without the drift term , the solution of the QSDE is a martingale, and hence by the Lipschiz condition on , one can have existence and uniqueness of the solution.
- (4)
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 -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 be a separable complex Hilbert space. For any , the creation operator defined by , is a bounded operator on with , where . Meanwhile, define the annihilation operator . Then the antisymmetric Fock space over is a Hilbert space. Moreover, define the fermion field on by , where is the conjugation operator (i.e., is antilinear, antiunitary and ). Denote by the von Neumann algebra generated by the bounded operators . For the Fock vacauum , define
(2.1) |
on . Obviously, is a normal faithful state on , and is a quantum probability space by [30].
Now, let , and be the von Neumann subalgebra of generated by then is an increasing family of von Neumann subalgebras of which is the noncommutative analogue of filtration in stochastic analysis. Let
then is a Fermion Brownian motion adapted to the family . For any , define the noncommutative -norm on by
where , then is the completion of , which is the noncommutative -space, abbreviated as . For any interval , set
It is easy to check that is a Banach space with the norm given by
Definition 2.1.
An adapted -processes on is said to be simple if it can be expressed as
(2.2) |
on for and for all .
By [3], the Itô-Clifford integral of any simple adapted -process with respect to Fermion Brownian motion is defined as follows.
Definition 2.2.
If is a simple adapted -processes on , then the Itô-Clifford stochastic integral of over with respect to is
(2.3) |
For , let be the linear space of all simple adapted -processes, i.e.
Then is subspace of whose processes vanish in . It is clear that is a Clifford -martingale for any , i.e. for any . For all , let
and denote by the non-commutative Hardy space the completion of under the norm . For simplicity, we use to represent the set of stochastic processes and . Moreover, we have the following the Burkholder-Gundy inequality (2.4) of Clifford -martingale first established in [23].
Lemma 2.1.
[23, Theorem 4.1] Let . Then, for any and its Itô-Clifford integral
it holds that
(2.4) |
where and are constant depend on .
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 . For any , then the left stochastic integral and right stochastic integral can be well-defined and
By Lemma 2.1 and Lemma 2.2, the Itô-Clifford integral can be defined for any element of , and Burkholder-Gundy inequality holds ture. For any , let be the completion of with the norm
Similarly, is the completion of with the norm
As an application of Minkowski inequality, we have the following result.
Theorem 2.3.
Let . Then, for any ,
(2.5) |
Furthermore, .
Proof.
Remark 2.1.
Actually, the above inequality (2.5) holds for any . 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 -solution to QSDE. Instead, we need the following easy corollary.
Corollary 2.4.
Let . Then, for any , there is positive constant such that
(2.8) |
Moreover, and
(2.9) |
Proof.
Now, we give the parity of each element of . Let the parity operator be automorphism map on von Neumann algebra generated by bounded linear operators on as is in [3, 5, 23].
Definition 2.3.
For any , is said to be odd if , is said to be even if . And, has definite parity if is even or odd.
Furthermore, for any ,
(2.11) |
where denote the even part and the odd part, respectively. More precisely, for any , , where and are even and odd, respectively. Since is isometric on ,
Let denote the algebra of even polynomials in the fields , and let be the -subalgebra of generated by . If is even there is a sequence in such that in , and therefore in . It follows that is also even. Similarly, if is odd in , there is a sequence of odd polynomials in the fields with and thus in , that is, is odd as well. It follows that if in and , then and in .
Lemma 2.5.
[3, Lemma 3.15] Let be Brownian motion. If has definite parity, then
depending on whether is even or odd.
Lemma 2.6.
[21, Theorem 1.8.2 Bihari inequality] Let be a continuous and non-decreasing function vanishing at 0 satisfying . Suppose is a continuous nonnegative function on such that
(2.12) |
where is a nonnegative cinstant and , then
where , is the conver function of . In particular, , then for all .
3 The existence and uniqueness of solutions to QSDE
This section is devoted to proving the existence and uniqueness of -solution to the equation (1.2) with non-Lipschitz coefficients for .
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 is well-defined for any . Let , be fixed. For any non-negative integer , define in inductively by
(3.1) |
where the well-definedness of iteration comes from Banach fixed point theorem and that is a strict contraction.
Firstly, we claim that each , , defines an adapted -continuous process on by induction. By assumption, and are -continuous with respect to and belong to for , then quantum stochastic integral exists for . Furthermore, we can obtain the boundedness of by the continuity on compact sets and easily verify that is continuous: .
Now, if is assumed to be adapted and continuous, then and are adapted, -continuous on and bounded, thus is adapted. For any , by (2.4) and Assumption 1.1,
Now, subtracting from both sides of above inequality and applying Corollary 2.4 and Hölder inequality, we get
which implies that is -continuous on . Hence we have proved our claim by induction.
Step 2. The sequence of iteration is convergent under the given conditions. For any , by Minkowski inequality,
By similar analysis as above, the elementary inequality , Hölder inequality and Osgood conditions of , there is constant such that
where .
Therefore, for any , ,
Since each is -continuous process on for any , is uniformly bounded on . Set , , which is uniformly bounded. Then
Let , . Then,
Denote
Applying Lebesgue dominated convergence theorem, we get
Hence, by Lemma 2.6, one deduces
which implies that is a Cauchy sequence in .
Step 3. is the solution to QSDE (1.2). Since is a Cauchy sequence in , there exists such that for any ,
Thus, for any , there exists such that
It shows that is -continuous and adapted on since is -continuous and adapted.
We shall prove that is the solution to
In fact,
since in for any and is continuous. Similarly,
in . Because for any , the same is true for .
Taking limits on both sides of (3.1), it deduces that
That is, is a -solution to the equation (1.2).
Uniqueness: Suppose that is another adapted -continuous solution with . Then, by (1.2), we obtain again
Furthermore,
Continuing to use the same technique as Step 2 of existence, we can yield that
It follows that, for any ,
This completes the proof. ∎
As described in [1], the Itô product rule holds for any . Based on [6], let
for any , the integral defines a quantum martingale for any . Next, let , we study the properties of the -solutions to QSDE (1.5) with respect to Brownian motion and 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 be adapted processes with . Then, for any , and are -martingales and
(3.2) | |||
Proof.
First, we consider simple adapted -process , then and are -martingales.
Let
be a partition of . Then
Define the martingale difference of as
By Theorem 2.1 of [23], there exists a positive constant such that
(3.3) |
By the canonical anticommutation relation
one has
(3.4) |
According to (2.11), for any , then
(3.5) | ||||
and
(3.6) | ||||
where the above two inequalities are based on Lemma 2.5 and (3.4).
4 The stability of solutions to QSDE with Lipschitz condition
In this section, we shall prove that the -solution to the equation (1.2) is stable, namely, small changes in the initial condition and in the coefficients and lead to small changes in the solution on .
Let the coefficients , of the equation (1.2) satisfy Lipsctitz condition, i.e.
(A3’) For any and a.e. , there exists a constant such that
Let , be the -solution to the equation (1.2) with initial conditions and for any , respectively. That is,
and
Theorem 4.1.
Suppose that assumptions (A1),(A2),(A3’),(A4) hold. With the above notations, for any , there exists such that if , then holds for all .
Proof.
By the directly calculation,
According to the proof of Theorem 3.1 again and , for any , one gains the following estimate
where .
By Gronwall’s inequality,
for all , 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 being replaced respectively by , , for all and be as in the equation (1.2). Assume that in as , uniformly on , in uniformly as on , and in . Furthermore, let be solutions to the equation (1.2) corresponding to and , respectively. Then in uniformly on compact set .
Corollary 4.3.
Suppose that assumptions (A1),(A2),(A3’),(A4) hold. Then the solution to the equation (1.5) is stable on when nonlocal condition 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 -solutions to QSDE (1.1) with nonlocal conditions under non-Lipschitz coefficients for .
According to the description of parity in Section 2, we get the following lemma.
Lemma 5.1.
Let be adapted and satisfy . Suppose further that for each . Then is self-adjoint element of , and .
Proof.
Let denote the self-adjoint part of . Let for , and satisfy . Suppose that are adapted and satisfy the Osgood condition of Assumption 1.1 on , and each is an even function. Set
Evidently, () are even by Lemma 4.1 of [4] for any . Let
It can be seen that satisfies the Osgood conditions and maps self-adjoint elements of into self-adjoint elements of . Then we obtain the self-adjointness of the solutions to QSDE with nonlocal conditions.
Theorem 5.2.
Proof.
Since satisfy the Osgood condition and satisfy the Lipschitz condition as in Assumption 1.1, it follows from Theorem 3.1 that the equation (5.1) admits a unique solution such that
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 ,
(5.2) |
To prove the self-adjointness of , it is sufficient to show that is self-adjoint by induction for all . It is obvious that is self-adjoint since . Assume that is self-adjoint, then . By Lemma 5.1, and are self-adjoint. In addition, and are also self-adjoint. Hence 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 , let denote the -algebra generated by and the solution to the equation (1.2) for , and write for . Since the solution is adapted, i.e. for all , it follows that is a -subalgebra of whenever . Let be the -subalgebra of generated by and . It is clear that and for any .
Next, we denote the algebra generated by field differences. Let denote the -subalgebra of generated by the field differences , and be the -subalgebra of generated by and . Thus, . Then, we get the following Markov property of the adapted solution to the equation (1.2).
Theorem 5.3.
6 Conclusion
Utilizing the Burkholder-Gundy inequality of Clifford -martingale, we obtain the existence, uniqueness and stability of the solutions to QSDEs with nonlocal conditions for . 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 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).