1. Introduction
In comparison with classical mechanics which considers macroscopic objects both in deterministic and stochastic settings, quantum mechanics, concerned with physical phenomena at atomic and subatomic scales, inherently incorporates randomness. In particular, the squared absolute value of the wave function of a quantum mechanical particle is interpreted as a probability density function, mixed quantum states are built of pure ones using randomisation, and the latter is also present in the model of quantum measurement [
39,
51]. However, in contrast to scalar-valued classical probability measures [
41], quantum probability describes statistical properties of quantum variables by using quantum states in the form of density operators [
16,
30] on the same Hilbert space where those variables act as linear operators.
The noncommutativity and canonical commutation structures, coming from the operator-valued nature of quantum variables and quantum states, give rise to specific features of quantum probability such as the absence of a classical joint probability distribution and conditional expectations for a set of noncommuting quantum variables (whereas an individual self-adjoint operator has a well-defined marginal distribution). Furthermore, since the microscopic realm is less amenable to manipulation by conventional macroscopic tools (unlike, for example, the coin tossing as a manageable “random number generator” for thought and practical experiments in classical probability theory), its natural time evolution makes statistical properties of quantum systems particularly tied to their dynamics. While an isolated quantum system undergoes a reversible evolution (according to a one-parameter unitary group generated by the system Hamiltonian), a more realistic open dynamics scenario [
6] involves interaction of the system with its environment which can include other quantum or classical systems, measuring devices and external quantum fields (quantised electromagnetic radiation).
The interaction of an open quantum system with its surroundings is accompanied by energy exchange between the subsystems and gives rise to dissipative effects. This provides a way to influence statistical and dynamic properties of such systems by arranging them into quantum networks [
14] and varying the energy and coupling parameters of the resulting interconnections, which can consist of a quantum plant coupled in a measurement-based or coherent (that is, measurement-free) fashion to a controller or observer. This paradigm is used in quantum control [
3,
5,
8] which develops systematic methods for achieving stability, robustness with respect to unmodelled dynamics, and optimality in the sense of relevant performance criteria for quantum systems and their applications, for example, in quantum optics [
11,
50] and quantum information processing [
32].
An important part in these developments belongs to quantum filtering, which, similarly to its classical predecessor (see for example, [
23,
28]), is concerned with mean square optimal real time estimation of the internal variables of the quantum plant using a measurement-based quantum Kalman filter [
2,
4,
9] or a qualitatively different coherent quantum observer [
31,
45]. The latter does not process classical observations and, in contrast to the classical filters, does not compute the conditional expectations of the plant variables (this conditioning and related Bayesian approaches are not applicable in the noncommutative quantum case as mentioned above). Instead, the coherent quantum filter is driven by the output quantum fields of the plant so as to produce a quantum process, whose performance, as an estimator of the plant variables, can be optimised in the sense of minimising the mean square value of the “estimation error” by varying the parameters of the Hamiltonian and coupling operators of the filter.
The optimisation problems, arising in coherent quantum filtering and control (with the latter considering more complicated feedback interconnections of a quantum plant and a quantum controller) involve physical realisability (PR) constraints [
22,
40] which originate from the canonical commutation structures of quantum dynamic variables, the energetics of open quantum systems, and unitarity in the augmented system-environment evolution. These features of open quantum dynamics are incorporated in the Hudson-Parthasarathy quantum stochastic calculus [
19,
35] (see also [
15]) which is often employed as a unified modelling framework in quantum filtering and control. In this approach, the external fields at the input of a quantum system of interest are represented by a multichannel quantum Wiener process with noncommuting components, which act on a symmetric Fock space and drive the quantum stochastic differential equation (QSDE) governing the system dynamics. In contrast to the classical SDEs with a standard Wiener process [
24], the QSDE reflects the unitarity of the augmented system-field evolution, and its drift and dispersion terms (as well as the generator of Markovian quantum dynamics) are specified by the Hamiltonian and coupling operators. These operators (together with a scattering matrix representing the photon exchange between the fields [
35]) describe the energetics of the quantum system and its interaction with the environment and are usually functions (for example, polynomials or Weyl quantization integrals [
10]) of the system variables. A particular form of this dependence and the commutation structure of the system variables affect tractability of the quantum system.
In particular, a quadratic dependence of the system Hamiltonian and a linear dependence of the system-field coupling operators on the quantum mechanical position-momentum variables [
29] lead to linear QSDEs for open quantum harmonic oscillators (OQHOs) [
9,
11] which play the role of building blocks in linear quantum control theory [
8,
22,
34,
38,
54,
55]. The dynamics of such systems are relatively well understood and are similar to the classical linear SDEs in a number of respects, including the preservation of Gaussian nature of quantum states [
20,
36] in the case of vacuum input fields. However, the coherent quantum analogue [
33] of the classical linear-quadratic-Gaussian (LQG) control problem [
1,
25] for OQHOs is complicated by the above mentioned PR constraints on the state-space matrices of the quantum controller and by the impossibility to take advantage of the classical estimation-actuation separation principle and classical conditional expectations with their variational properties [
28].
One of the existing approaches to the coherent quantum LQG (CQLQG) control and coherent quantum filtering (CQF) problems employs the Frechet derivatives of the mean square cost (with respect to the state-space matrices subject to the PR constraints) for obtaining the optimality conditions [
44,
45] and for the numerical optimization [
42]. This approach takes into account the quantum nature of the underlying problem only through the PR constraints, being “classical” in all other respects, which has its own advantages from the viewpoint of well-developed conventional optimization methods. However, a disadvantage of this approach is that it is limited to certain parametric classes of linear controllers and observers. In particular, the resulting optimality conditions do not provide information on whether nonlinear quantum controllers or observers can outperform the linear ones for linear quantum plants. For this reason, the coherent quantum control and filtering problems require novel variational methods for their solution, which would be able to operate with sensitivity of the system dynamics and relevant cost functionals to perturbations over wider classes of the Hamiltonian and coupling operators in a “coordinate-free” fashion.
To this end, the present paper (some of its results were briefly announced in [
48]) outlines a fully quantum variational method which allows the sensitivity of the internal and output variables of a nonlinear quantum stochastic system to be investigated with respect to arbitrary (that is, not only linear-quadratic) perturbations of the Hamiltonian and coupling operators. This approach is based on using a
transverse Hamiltonian, defined as an auxiliary time-varying self-adjoint operator which encodes the propagation of such perturbations through the unitary system-field evolution. This leads to an infinitesimal perturbation formula for quantum averaged performance criteria (such as the mean square cost functional) which is applicable to the development of optimality conditions in coherent quantum control and filtering problems over larger classes of controllers and observers. In particular, this approach allows the sensitivity of OQHOs with quadratic performance criteria to be studied with respect to general perturbations of the Hamiltonian and coupling operators. In fact, the transverse Hamiltonian method has already been employed in [
47] in order to establish local sufficiency of linear observers in the mean square optimal CQF problem [
45] for linear quantum systems with respect to varying the Hamiltonian and coupling operators of the observer along linear combinations of the Weyl operators [
10]. Note that our approach is different from [
13] which develops a quantum Hamilton-Jacobi-Bellman principle (for the density operator instead of the dynamic variables) in a measurement-based quantum feedback control problem. We also mention a parallel between the perturbation analysis, discussed in the present paper, and the fluctuation-dissipation theorem [
26].
The paper is organised as follows.
Section 2 specifies the class of quantum stochastic systems under consideration.
Section 3 discusses sensitivity of the internal and output variables of OQHOs to parametric perturbations within the families of quadratic system Hamiltonians and linear system-field coupling operators.
Section 4 returns to general QSDEs and introduces the transverse Hamiltonian associated with arbitrary perturbations of the Hamiltonian and coupling operators of the system. The transverse Hamiltonian is used in
Section 5 for infinitesimal perturbation analysis of system operators.
Section 6 extends the transverse Hamiltonian method to quantum averaged performance criteria.
Section 7 applies this approach to a mean square optimal CQF problem.
Section 8 makes concluding remarks.
We use the following principal notation in the paper. Denoted by is the commutator of linear operators A, B, with a linear superoperator associated with A. This extends to the commutator -matrix for vectors X, Y of operators , , respectively. Vectors are organised as columns unless indicated otherwise, and the transpose acts on matrices of operators as if their entries were scalars. In application to such matrices, is the transpose of the entry-wise operator adjoint . For complex matrices, is the usual complex conjugate transpose . The subspaces of real symmetric, real antisymmetric and complex Hermitian matrices of order n are denoted by , and , respectively, where is the imaginary unit. The real and imaginary parts of a complex matrix are denoted by and . These extend to matrices M with operator-valued entries as and which consist of self-adjoint operators. Positive (semi-) definiteness of matrices and the corresponding partial ordering are denoted by (≽) ≻. Also, and denote the sets of positive semi-definite real symmetric and complex Hermitian matrices of order n, respectively. The tensor product of spaces or operators (in particular, the Kronecker product of matrices) is denoted by ⊗. The tensor product of operators A, B acting on different spaces will sometimes be abbreviated as . The identity matrix of order n is denoted by , and the identity operator on a space H by . The Frobenius inner product of real or complex matrices is denoted by . Denoted by is a weighted Euclidean (semi-)norm of a real vector v specified by a real positive (semi-)definite symmetric matrix K. The quantum expectation of a quantum variable over a density operator extends in an entry-wise fashion to vectors and matrices of such variables.
2. Quantum stochastic systems being considered
We consider an open quantum system, which interacts with an external multichannel bosonic field and is equipped with dynamic variables
, evolving in time
and assembled into a vector
. These system variables are assumed to be self-adjoint operators on a composite system-field Hilbert space
. Here,
is the initial space of the system, which provides a domain for
, and
is a symmetric Fock space [
35] for the action of an even number
m of quantum Wiener processes
. The latter are time-varying self-adjoint operators, which model the external fields and are assembled into a vector
. Unlike the classical Brownian motion [
24] in
, the quantum Wiener process
W consists of noncommuting operator-valued components and has a complex Ito matrix
(identified with
) for its future-pointing Ito increments
:
Here, the matrix
specifies the canonical commutation relations (CCRs) for the constituent field processes
:
(with
spanning the subspace of antisymmetric
-matrices), which is an incremental form of the two-point CCRs
These CCRs are closely related to the continuous tensor-product structure of the Fock space [
37] and are complemented by the commutativity between the Ito increments of
W and adapted processes
taken at the same (or an earlier) moment of time:
The adaptedness of quantum processes on the system-field space
is understood with respect to a filtration
, where
and
is the Fock space filtration, so that for any
, the operators
act effectively on
, while
act on the subspace
.
The energetics of the quantum system and its interaction with the external fields is specified by a system Hamiltonian and system-field coupling operators which are time-varying self-adjoint operators, organised as deterministic functions (for example, polynomials with constant coefficients) of the system variables and assembled into a vector . Accordingly, the operators , act on the initial space . Depending on the context, a system operator (a function of the initial system variables ) on the initial space will be identified with its extension to the system-field space .
The system and the fields form a composite quantum system which evolves according to a quantum stochastic flow as described below. This evolution is specified at any time
by a unitary operator
on
governed by the following stochastic Schrödinger equation [
19,
35]:
with initial condition
, so that
captures the internal dynamics of the system and the system-field interaction over the time interval
. Here and in what follows, the subscript 0 marks the initial values of time-varying operators (or vectors or matrices of operators):
,
,
, while the time arguments will often be omitted for the sake of brevity. Also, the units are chosen so that the reduced Planck constant is
. The quantum stochastic differential equation (QSDE) (
6) corresponds to a particular yet important case of the identity scattering matrix, when there is no photon exchange between the fields, and the gauge processes [
35] can be eliminated from consideration. The term
in (
6) can be interpreted as an incremental Hamiltonian of the system-field interaction, while
involves the quantum Ito matrix
from (
1) and counterbalances the Ito correction term
in the differential relation
which describes the preservation of the co-isometry property
for all
. The system variables at time
, as operators on the system-field space
, are the images
of their initial values under the quantum stochastic flow
which maps a system operator
on the initial space
to the operator
on
in (
5). The resulting quantum adapted process
satisfies the following Hudson-Parthasarathy QSDE [
19,
35]:
where use is made of the Hamiltonian and the coupling operators evolved by the flow
from (
8) as
(the flow acts entry-wise on vectors and matrices of operators). Also,
in (
9) is the decoherence superoperator which acts on
as
The second equality in (
11) is applicable to the case when
is a vector of operators on which the superoperator
acts entry-wise. The superoperator
in (
9) is the Gorini-Kossakowski-Sudarshan-Lindblad (GKSL) generator [
12,
27], which is a quantum counterpart of the infinitesimal generators of classical Markov processes [
24]. The identity
, which holds for system operators
in view of (
8) and the unitarity of
, allows the QSDE (
6) to be represented in the Heisenberg picture by using (
10) along with the commutativity (
4) as
In application to the vector
of system variables, the quantum stochastic flow
also acts entry-wise as
in accordance with (
7), (
8), so that the corresponding QSDE (
9) can be represented in vector-matrix form:
where the
n-dimensional drift vector
and the dispersion
-matrix
consist of time-varying self-adjoint operators on
.
The interaction of the system with the input field
W produces the output fields
assembled into a vector
where the system-field unitary evolution is applied to the current input field variables (which reflects the innovation nature of the quantum Wiener process and the continuous tensor product structure of the Fock space mentioned above). The output field
Y satisfies the QSDE
where the matrix
J is given by (
2), and
L is the vector of system-field coupling operators from (
10). The system-field interaction makes the output field
Y different from the input field
W only through the drift vector
in (
16).
The common unitary evolution in (
13), (
15) preserves the commutativity between the system variables and the output field variables in time
:
where the entries of
,
commute as operators on different spaces
,
. By a similar reasoning, the output field
Y inherits the CCR matrix
J from the input quantum Wiener process
W:
where (
3) is used with
. By (
1), (
2), the relation (
18) can also be established as a corollary of the property that
Y in (
16) inherits the quantum Ito matrix
from
W as
.
In view of (
6) (see also (
12)), the quantum stochastic flow
in (
8) depends on the system Hamiltonian
and the system-field coupling operators in
. In turn,
and
are usually functions (such as polynomials or Weyl quantization integrals [
10] in [
49]) of the initial system variables
. The dependence of
,
on
is inherited by
H,
L as functions of
X and (along with a given commutation structure of the system variables) specifies a particular form of the resulting QSDEs (
14), (
16), thus influencing their tractability.
3. Open quantum harmonic oscillators with parametric dependence
An important class of quantum stochastic systems is provided by multimode open quantum harmonic oscillators (OQHOs) [
9] with an even number
n of system variables
(for example, consisting of
conjugate position-momentum pairs [
29]) satisfying the Heisenberg CCRs
(on a dense domain in
), where the CCR matrix
is identified with
and remains unchanged over the course of time
. The Hamiltonian
H of the OQHO is a quadratic function and the system-field coupling operators
in (
10) are linear functions of the system variables,
parameterised by an
energy matrix and a
coupling matrix. In this case, the QSDEs (
14), (
16) take the form
where the drifts
and
depend linearly on the system variables
, with
,
and the dispersion matrices
and
being constant real matrices. This linearity makes some of the dynamic properties of (
21) similar to those of a classical linear stochastic system with a state-space realization quadruple
and the corresponding
-valued transfer function
F on the complex plane:
thus allowing for application of transfer function techniques [
52]. However, in addition to the noncommutative nature of quantum variables, the matrices of coefficients of these QSDEs have a specific parameterization in terms of the energy, coupling and CCR matrices:
Since the energy matrix
R is symmetric, and the CCR matrices
,
J are antisymmetric, the matrices
A,
B,
C satisfy
These equalities pertain to the fulfillment of the CCRs (
19) and the commutativity (
17) at any moment of time and provide necessary and sufficient conditions for physical realizability (PR) [
22,
40] of the linear QSDEs (
21) as an OQHO with the CCR matrix
for the system variables. The first equality in (
24) has the structure of an algebraic Lyapunov equation (ALE) with respect to
, which has a unique solution if and only if the Kronecker sum
is a nonsingular matrix (that is, no two eigenvalues of
A are centrally symmetric about the origin in
). The latter condition holds, for example, when the matrix
A is Hurwitz. For more general open quantum systems (such as anharmonic oscillators whose dynamic variables satisfy the CCRs (
19), while the Hamiltonian and the coupling operators are not necessarily quadratic and linear functions of the system variables), the CCR preservation is secured by the unitary evolution of the system variables in (
13).
By analogy with the state-space realizations of transfer functions in classical linear systems theory [
1,
25], we will use the
input-output map for the OQHO (
21) with the matrix quadruple
:
Now, suppose the energy and coupling matrices
R,
N, which specify the Hamiltonian and the coupling operators in (
20), depend smoothly on an auxiliary parameter
(while the quantum Wiener process
W is independent of
). Then so also do the matrices
A,
B,
C in (
23) and the system and output variables which comprise the vectors
,
(the differentiability of
X,
Y is understood in the weak sense). The corresponding partial derivatives
at any
give rise to adapted quantum processes with self-adjoint operator-valued entries satisfying the QSDEs
(the second of which is, in fact, an ODE
involving the time derivative
of
), with zero initial conditions
,
since
,
do not depend on
. Here,
are the derivatives of the matrices
A,
B,
C from (
23) with respect to the parameter
satisfying the relations
which are obtained by differentiating the PR conditions (
24), with the CCR matrices
,
J being constant. Since the CCR matrices
,
J are antisymmetric (and hence,
), the first equality in (
29) implies that
By assembling the system variables and their parametric derivatives from (
26) to an augmented vector
a combination of the QSDEs (
21), (
27) allows the parametric derivative of the map (
25) to be represented as the input-output map
associated with
Here, the initial condition
, in view of
, is identified with
, and the matrices
,
,
are given by
This can also be obtained by using the transfer functions of the corresponding linear systems (including (
22)) as
for any
which is not an eigenvalue of
A. Here, use is made of a particular case of the block matrix inverse formula [
17]:
The block lower triangular structure of the dynamics matrix
in (
32), (
34) is closely related to the Gateaux derivative of the matrix exponential
in the direction
(see, for example, [
18]):
At any time
, both
and
depend linearly on the matrices
,
from (
28) in view of the following representation of the solutions of the QSDEs (
27), which regards
as a forcing term in the first of these QSDEs:
where
is the unperturbed solution of the first QSDE in (
21) which does not depend on
,
. In the case of linear QSDEs whose coefficients depend smoothly on parameters, the mean square differentiability of their solutions with respect to those parameters can be verified directly by using the closed form (
37) (under certain integrability conditions for the underlying quantum state in terms of relevant moments of the system variables such as
).
An additional insight into the structure of the process
is provided by its cross-commutation relations with
X,
W. In particular,
inherits the commutativity with the Ito increments of the input field
W from that of the underlying system variables in (
4), and so also does
in (
31):
Furthermore, the differentiation of (
19) in
, combined with the identity
for vectors
,
of operators, leads to
whereby the cross-commutation matrix
is symmetric (which holds regardless of a particular structure of the QSDEs (
21) and remains valid for quantum anharmonic oscillators with nonlinear dynamics). This matrix evolves in time (with zero initial condition because
) and has a steady-state value computed below.
Theorem 1.
Suppose the matrix A of the OQHO (21) in (23) is Hurwitz. Then there exist the following limits
where the matrices , are found as unique solutions of the ALEs
Proof. In view of the CCRs (
19), the commutator matrix for the vector
in (
31) is organised as
where the blocks
,
and
consist of time-varying skew self-adjoint operators on the system-field space
. By using the quantum Ito lemma [
19,
35] and the bilinearity of the commutator along with the QSDE (
33) and the commutativity (
38), it follows that (
43) satisfies the QSDE
which (similarly to [
22], Eq. (
56)]) reduces to the ODE
with the initial condition
in view of
. Hence, at any time
, the matrix
is an imaginary antisymmetric matrix which can be represented as
with time-varying matrices
and
. Substitution of (
46) into (
44) and using (
34) leads to the ODEs
with zero initial conditions
,
in accordance with the corresponding blocks in (
45). Now, the matrix
is symmetric at any time
due to (
39). This also follows from (
47) in view of the symmetry (
30). Hence,
, and the ODE (
48) takes the form
If the matrix
A is Hurwitz, then the solutions of the ODEs (
47), (
49) converge to their unique steady-state values which (slightly abusing notation) satisfy the ALEs (
41), (
42) and specify the limits (
40). □
The relations (
28)–(
36) and Theorem 1 provide infinitesimal perturbation analysis for sensitivity of the internal and output variables of the OQHO to the matrices
R,
N. Under the perturbations of
R,
N, the Hamiltonian
H and the coupling operators in
L, given by (
20), remain in the corresponding classes of quadratic and linear functions of the system variables. Accordingly, the above analysis is not applicable to more general perturbations (for example, higher order polynomials of the system variables) and is restricted to linear QSDEs, so that an alternative approach is needed in the general case.
4. Transverse Hamiltonian
For the general quantum stochastic system, described in
Section 2 and governed by (
14), (
16), we will now consider its response to arbitrary perturbations in the system Hamiltonian
and the system-field coupling operators in
. More precisely, suppose they are perturbed in directions
,
as
Here,
and the entries of the
m-dimensional vector
are self-adjoint operators on the initial system space
, and
is a small real-valued parameter as before, so that
and
. In what follows, the derivative
is taken at
. The perturbations
,
in (
50) are also assumed to be functions of the initial system variables
, and these dependencies describe
as functions of
under the
unperturbed flow (
8). For example, in the case of OQHOs considered in
Section 3, the perturbations
K,
M, which are caused by the perturbations in the energy and coupling matrices
R,
N, inherit the structure of the Hamiltonian as a quadratic function and the coupling operators as linear functions of the system variables in (
20), respectively:
Returning to the general case, we will avoid at this stage technical assumptions on
,
in (
50), so that the calculations, carried out below for arbitrary perturbations, should be regarded as formal. Since the operators
,
completely specify the dynamics of the unitary operator
in (
6) which determines the evolution of the system and output field variables, the response of the latter to the perturbations (
50) of
,
reduces to that of
. Therefore, the propagation of the initial perturbations
,
of the operators
,
through the subsequent unitary system-field evolution can be described in terms of the operator
which satisfies
since
does not depend on
. The smoothness of dependence of
on the parameter
is analogous to the corresponding property of solutions of classical SDEs (whose drift and dispersion satisfy suitable regularity conditions [
24,
43]) and holds at least in the case of linear QSDEs discussed in
Section 3. The following theorem is closely related to Stone’s theorem on generators of one-parameter unitary groups [
53].
Theorem 2.
For any time , the operator in (53), associated with the unitary evolution from (6), can be represented as
Here, is a self-adjoint operator on the system-field space , which satisfies zero initial condition and is governed by the QSDE
Furthermore, depends linearly on the initial perturbations , of the Hamiltonian and coupling operators in (50) through their unperturbed evolutions in (51).
Proof. The differentiation of both sides of the unitarity relation
with respect to the parameter
at
leads to
, which implies self-adjointness of the operator
thus establishing (
54). Now, consider the time evolution of
. To this end, the differentiation of (
6) with respect to
yields
By left multiplying both sides of (
57) by
and recalling (
56), it follows that
where use is made of the evolved perturbations (
51). By a similar reasoning, a combination of (
12) with (
56) leads to
where use is also made of (
58) along with the quantum Ito product rules [
35] and (
1), (
4). It now follows from (
56), (
58)–(
60) that
which establishes (
55). The linear dependence of
on
,
follows from the integral representation
of the QSDE (
55) and the property that the evolved perturbations
K,
M in (
51) depend linearly on
,
, respectively. □
In view of Theorem 2, for any fixed but otherwise arbitrary time
, the relation (
54), represented as
has the structure of isolated quantum dynamics in fictitious time
, where
plays the role of a Hamiltonian pertaining to the perturbation of the unitary operator
. In order to reflect this property, we will refer to the time-varying operator
Q as the
transverse Hamiltonian associated with the perturbations
K,
M of the system Hamiltonian
H and the system-field coupling operators in
L. The computation of
Q is illustrated by the following two examples.
Example 1
In the absence of perturbation to the system-field coupling, when the vector
in (
50) consists of zero operators, and hence, so also does
M in (
51), the transverse Hamiltonian in (
61) reduces to
Moreover, if the system is isolated, that is,
, then (
6) reduces to the ODE
which leads to
with the Hamiltonian being preserved in time:
for all
. In this case of isolated system dynamics, (
62) takes the form
Here, use is made of Hadamard’s lemma [
29] along with an entire function
of a complex variable (with
by continuity) which plays a role in the solution of nonhomogeneous linear ODEs with constant coefficients and constant forcing terms [
18]. The Gateaux derivative (
53) of (
63) can be represented by using an operator version of (
35) as
which provides an alternative verification of (
54) for this particular case, with
Q given by (
64). ▲
Example 2
For the OQHO of
Section 3, substitution of the perturbations (
52) into (
61) leads to the transverse Hamiltonian
which depends linearly on the matrices
,
and in a quadratic fashion on the past history of the system variables. The latter are given by the unperturbed equation (
37). The last term
in (
66) comes from the relation
(following from (
1)) and the identity
which holds for any matrix
in view of the CCRs (
19). ▲
5. Infinitesimal perturbation analysis of system operators
Since the transverse Hamiltonian
in (
56), (
61) encodes the propagation of the initial perturbations of the Hamiltonian and coupling operators in (
50) through the unitary system-field evolution over the time interval
, it provides a tool for infinitesimal perturbation analysis of general system operators. The following theorem is concerned with an extended setting which, in addition to (
50), allows for a smooth dependence of a system operator
on the parameter
, so that an appropriate infinitesimal perturbation in it is specified by an operator
on the initial space
, with
also being a function of the initial system variables
.
Theorem 3.
For any self-adjoint system operator on the initial space , which smoothly depends on the same scalar parameter ϵ as in (50) and is evolved by the flow (8), the derivative of its evolved version with respect to ϵ can be represented as
at any time , where is the transverse Hamiltonian from Theorem 2. Here, the operator satisfies the QSDE
with zero initial condition , where is the unperturbed GKSL generator from (9), and χ is a linear superoperator given by
Proof. By using the Leibniz product rule together with (
8), (
53), (
54) and the self-adjointness of
, it follows that
which establishes (
67), with the initial condition
being inherited by
from
. We will now obtain a QSDE for the process
. To this end, by combining the quantum Ito lemma with the bilinearity of the commutator, and using the QSDEs (
9), (
55), it follows that
Here, the quantum Ito product rules are applied together with the commutativity (
4) between adapted processes and the Ito increments of the quantum Wiener process
W. In the second and third of the equalities (
71), use is also made of the relations
and
for appropriately dimensioned adapted processes
,
with self-adjoint operator-valued entries. These relations are combined with the identity
in the fourth equality of (
71). The QSDE (
68) is now obtained my multiplying both sides of (
71) by
i and using the superoperator
from (
69). □
As can be seen from (
70), the process
in (
67) is the response of
to the initial perturbation
of the system operator under the unperturbed flow (
8) and satisfies the QSDE (
9):
In contrast to
, the operator
in (
67) describes the response of the flow
itself to the perturbations (
50) of the Hamiltonian and coupling operators and will be referred to as the
derivative process for the system operator
. Accordingly, the term
in the drift of the QSDE (
68) is the derivative process for
. At any given time
, the last equality in (
67) is organised as the right-hand side of the Heisenberg ODE (in fictitious time
) of an isolated quantum system with the state space
and the Hamiltonian
. While
(evolved by the unperturbed flow) depends linearly on
, the derivative process
depends linearly on the perturbations
,
of the Hamiltonian
and the coupling operators in
through the transverse Hamiltonian
Q and the superoperator
in (
69).
In application to the vectors
X,
L of the system variables and the system-field coupling operators, the QSDEs (
68), (
72) and the definition (
69) lead to
where
F,
G are the unperturbed drift vector and the dispersion matrix from (
14). In (
73), use is also made of the absence
of initial perturbations in the system variables, whereby
for any
. Furthermore, in (
74), we have used the relations
in view of (
50), (
51). Since the matrix
J and the quantum Wiener process
W do not depend on the parameter
, the derivative of the output field
Y of the system in (
16) evolves according to the ODE
where
is governed by the QSDE (
74).
In particular, for the OQHO of
Section 3 with the perturbations (
52), the QSDEs (
73)–(
75) lead to the relations (
27), (
28) which were obtained in
Section 3 using more elementary techniques. However, the latter are limited to quadratic perturbations of the Hamiltonian and linear perturbations of the coupling operators in (
52), whereas the transverse Hamiltonian approach allows the system response to be investigated for general perturbations of these operators. Therefore, this approach can be used for the development of optimality conditions in quantum control and filtering problems for larger classes of controllers and observers.
6. Sensitivity of infinite-horizon quantum averaged functionals
Similarly to optimal control of classical time invariant stochastic systems [
1,
25], suppose the infinite-horizon performance of the quantum system being considered is described by a cost functional
which (whenever it exists) leads to the same Cesaro limit
and is to be minimised in optimal quantum control settings. Here, the quantum expectation
is taken over the tensor product
of the initial quantum state
of the system and the vacuum state [
35]
in the Fock space
for the external bosonic fields. This expectation is applied in (
76) to a quantum
criterion processZ, specified as
by a function
. The latter is extended to the noncommutative system variables and the coupling operators (with
L being in a bijective correspondence with the drift vector
of the output field in (
16) since
in view of (
2)) so as to make
a self-adjoint operator for any
. Such an operator-valued extension of
f is straightforward in the case of polynomials and can be carried out through the Weyl quantization [
10] for more general functions. In the coherent quantum linear-quadratic-Gaussian (CQLQG) control and filtering problems [
33,
44,
45], the function
f in (
78) is a positive semi-definite quadratic form, in which case, the minimization of (
76) provides an infinite-horizon mean square optimality criterion.
Now, if the Hamiltonian and coupling operators of the quantum system are perturbed according to (
50), then application of the transverse Hamiltonian
Q from Theorems 2, 3 leads to
where
is the derivative process for
Z. Note that, despite
, the operator
can be nonzero due to the dependence of
Z in (
78) on
L. Assuming the existence and interchangeability of appropriate limits, (
79) leads to the following perturbation formula for the cost functional
in (
76):
The right-hand side of (
80) is a linear functional of the perturbations
,
, which describes the corresponding (formal) Gateaux derivative of
in the direction
. Therefore, the quantum system is a stationary point of the performance criterion (
76) with respect to a subspace
of perturbations
in (
50), if
is contained by the null space of the linear functional
in (
80):
This inclusion provides a first-order necessary condition of optimality in the quantum control problem of minimising (
76) over a manifold of the Hamiltonian and coupling operators with the local tangent space
.
While
in (
80) reduces to averaging over the invariant quantum state of the unperturbed system (provided certain integrability conditions are satisfied together with the existence of and weak convergence to the invariant state), the computation of
is less straightforward. Due to the product structure (
77) of the system-field state
(with the external fields being in the vacuum state
), the martingale part of the QSDE (
68) does not contribute to the time derivative
where
is the derivative process for
, and the superoperator
from (
69) is applied to the criterion process
Z in (
78).
The relation (
82) is a complicated integro-differential equation (IDE). However, this IDE admits an efficient solution, for example, in the case when the system is an OQHO, and the function
f in (
78) is a polynomial. In this case, due to the structure of the GKSL generator of the OQHO,
is also a polynomial in the system variables of the same degree, thus leading to a linear relation (with constant coefficients) between the derivative processes
,
and to algebraic closedness in the IDE (
82). Therefore, for stable OQHOs (with a Hurwitz matrix
A) and a polynomial criterion process
Z, the computation of
(and verification of the stationarity condition (
81)) reduces to averaging over the unperturbed invariant state (which is unique and Gaussian [
36]). These calculations are exemplified below.
Example 3
Consider the OQHO of
Section 3, described by (
19)–(
23) with a Hurwitz matrix
A. Suppose the criterion process
Z in (
78) is a quadratic polynomial of the system variables
and the system-field coupling operators
:
where
is a given weighting matrix, partitioned into blocks
,
,
, and
is an auxiliary matrix which involves the coupling matrix
N from (
20). Then, in view of
, it follows from (
83), (
52) that
The second equality in (
83) allows the quantum average of the corresponding derivative process
in (
79) to be represented as
Here,
is the expectation of the derivative process
for
and hence, satisfies the following IDE, similar to (
82):
where the superoperator
from (
69) is applied to
entry-wise as
with
the
ℓth row of the CCR matrix
. By substituting the Hamiltonian and coupling operators of the OQHO from (
20) into the GKSL generator
in (
9), (
11) and using the CCRs (
19) together with the state-space matrices (
23), it follows that
Substitution of (
89) into (
87) reduces the IDE to a Lyapunov ODE:
Since the matrix
A is Hurwitz, the unperturbed OQHO has an invariant state which is Gaussian [
36] with zero mean
and covariance matrix
, where
is a unique solution of the ALE
Therefore, under appropriate integrability conditions for
, the solution of the Lyapunov ODE (
90) has a limit
which is a unique solution of the ALE
where
can be computed by averaging
over the invariant Gaussian state of the OQHO. Therefore, by assembling (
85), (
86) into (
80), it follows that
where the limit also reduces to averaging over the invariant Gaussian state and is expressed in terms of
as
The relations (
84), (
88), (
92)–(
94) allow the Gateaux derivative
of the quadratic cost functional (
76), specified by (
83), to be computed through mixed moments of the system variables
X and the perturbations
K,
M over the invariant zero-mean Gaussian quantum state whose covariance matrix is found from (
91). In particular, if
K,
M are polynomials of the system variables
, then these moments can be calculated in terms of the covariances by using the Isserlis-Wick theorem [
21]. Alternatively, the perturbations
K,
M can be trigonometric polynomials, that is, linear combinations of unitary Weyl operators [
10]
associated with the system variables, or more generally, represented as the Weyl quantization integrals
similar to those in [
49], (Eq.
32). Here,
,
are countably additive measures of finite variation on the
-algebra
of Borel subsets of
, which take values in
and
, respectively, and satisfy the Hermitian property
and
for any
, thus ensuring that
K and the entries of
M in (
96) are self-adjoint operators in view of the second equality in (
95). Now, the Weyl CCRs
for all
(with their infinitesimal Heisenberg form given by (
19)) imply that
, whereby
Since the quantum expectation commutes with the differential operator on the right-hand side of (
97), and the quasi-characteristic function [
7] of the invariant zero-mean Gaussian state is given by
, then
which can also be obtained by using quantum Price’s theorem [
46]. A combination of the second equality from (
96) with (
98) leads to
The computation of the limit in (
93) in the Weyl quantization framework is now completed by substituting (
99) into (
94). In this framework, the matrix
, which is needed for finding
from the ALE (
92) for the last term in (
93), is computed by averaging (
88) in a similar fashion. ▲
7. Mean square optimal coherent quantum filtering problem
We will now outline an application of the transverse Hamiltonian approach of
Section 4,
Section 5 and
Section 6 to a quantum filtering problem for the open quantum system, described in
Section 2 and referred to as a
quantum plant. The plant has the Hamiltonian
H, the vector
L of system-field coupling operators, the input field
W, the vector
X of internal variables and the output field
Y governed by the QSDEs (
14), (
16). Although the plant is not necessarily an OQHO, we assume that the plant variables
are organised as position-momentum pairs and satisfy the CCRs (
19).
Suppose the quantum plant is cascaded in a measurement-free field-mediated fashion with another open quantum system, playing the role a
coherent quantum observer, which is driven by the plant output
Y and another quantum Wiener process
of an even dimension
on a different bosonic Fock space
; see Fig.
Figure 1.
The observer is endowed with its own initial Hilbert space
, dynamic variables
with a CCR matrix
(so that
for any
, similarly to (
19)), and an
-dimensional output field
:
Since the observer is driven by the plant output
Y (together with the quantum noise
), the observer output
acquires, over the course of time, quantum statistical correlations with the plant variables and can be used for estimating the latter in a mean square optimal fashion as specified below. To this end, we denote the observer Hamiltonian by
and the vectors of operators of coupling of the observer with the plant output
Y and with the quantum Wiener process
by
respectively. The Hamiltonian
and the coupling operators
,
are functions of the dynamic variables
of the observer and hence, commute with the plant variables and functions thereof, including
H,
L. The plant and observer form a composite quantum stochastic system, whose vector
of dynamic variables satisfies the CCRs
and is driven by an augmented quantum Wiener process
with the following Ito table:
Here, ℧ is the Ito matrix of the quantum Wiener process
of the observer, which is defined similarly to
from (
1), (
2) as
The quantum feedback network formalism [
14] allows the Hamiltonian
of the plant-observer system and its vector
of operators of coupling with
to be computed as
Hence, the internal and output variables
and
of the observer in (
100) are governed by the QSDEs
Here,
is the GKSL generator of the plant-observer system, and
is the corresponding decoherence superoperator, in accordance with (
11), (
103), (
105). In (
106), use is also made of the partial decoherence superoperator
which acts on the observer variables as
in view of (
102)–(
105). Now, consider a coherent quantum filtering (CQF) problem formulated as the minimisation of the mean square discrepancy
between
r linear combinations of the plant variables of interest and the entries of the drift part
of the observer output
in (
107) as specified by given weighting matrices
and
, with
,
. The criterion process
Z in (
108) is similar to that in (
83):
with
X being the only subvector of
from (
102) which is present in (
109). The minimization in (
108) is over the observer Hamiltonian
and the vector
of the observer-plant coupling operators in (
101) as functions of the observer variables
. This mean square optimal CQF problem extends [
45] in that we do not restrict attention to linear observers even if the plant is an OQHO and
depends linearly on the observer variables
. Note that the Hamiltonian
H of the plant and its coupling
L to the input quantum noise
W are fixed, and so also is the coupling
of the observer to the input quantum noise
; see
Figure 1. If
,
are perturbed in the directions
consisting of self-adjoint operators representable as functions of the initial observer variables
, then the corresponding perturbations of the plant-observer Hamiltonian and coupling operators in (
105) are
By applying Theorem 2 and using (
103), (
110), (
111), it follows that the corresponding transverse Hamiltonian
Q for the plant-observer system satisfies the QSDE
where use is also made of
(following from (
1) and the commutativity
). Then the Gateaux derivative
of the cost functional in (
108) is given by (
80), where
in view of (
109), (
111) and similarly to (
85). In accordance with (
82), the expectation of the derivative process
satisfies the IDE
where, in view of (
112), the superoperator
in (
69) is given by
In particular, if both the plant and the unperturbed observer are OQHOs with Hurwitz dynamics matrices (while the perturbations in (
110) are not necessarily linear-quadratic), then
can be found (in a form similar to (
93)) along the lines of Example 3 of
Section 6 due to the criterion process
Z in (
108), (
109) being a quadratic function of the plant and observer variables in
from (
102).
The transverse Hamiltonian method, outlined above, was used in [
47] for showing that, in the mean square optimal CQF problem for linear quantum plants, those observers, which are locally optimal in the class of linear quantum observers, cannot be improved locally (in the sense of the first-order optimality conditions (
81)) by varying the Hamiltonian and coupling operators of the observer along linear combinations of the Weyl operators associated with the observer variables.