Coherent Quantum LQG Controllers with Luenberger Dynamics
Abstract
This paper is concerned with the coherent quantum linear-quadratic-Gaussian control problem of minimising an infinite-horizon mean square cost for a measurement-free field-mediated interconnection of a quantum plant with a stabilising quantum controller. The plant and the controller are multimode open quantum harmonic oscillators, governed by linear quantum stochastic differential equations and coupled to each other and the external multichannel bosonic fields in the vacuum state. We discuss an interplay between the quantum physical realizability conditions and the Luenberger structure associated with the classical separation principle. This leads to a quadratic constraint on the controller gain matrices, which is formulated in the framework of a swapping transformation for the conjugate positions and momenta in the canonical representation of the controller variables. For the class of coherent quantum controllers with the Luenberger dynamics, we obtain first-order necessary conditions of optimality in the form of algebraic equations, involving a matrix-valued Lagrange multiplier.
keywords:
Coherent quantum LQG control, physical realizability, separation principle, Luenberger controller, optimality conditions.1 Introduction
Open quantum harmonic oscillators (OQHOs), described by Hudson-Parthasarathy linear quantum stochastic differential equations (QSDEs) (Hudson & Parthasarathy (1984); Parthasarathy (1992)), are the closest quantum mechanical counterparts of classical linear stochastic systems. However, unlike classical random processes, the dynamic variables of an OQHO are noncommuting self-adjoint operators on an infinite-dimensional Hilbert space, organised similarly to the pairs of conjugate position and momentum operators (Sakurai (1994)). The QSDE, which governs the OQHO, is driven by a quantum Wiener process with noncommuting components on a symmetric Fock space, thus modelling the interaction of the system with an external bosonic quantum field. The energy exchange in this interaction and the self-energy of the OQHO (pertaining to its internal dynamics) are described in terms of the system-field coupling operators and the Hamiltonian, parameterised by coupling and energy matrices. Together with the canonical commutation relations (CCRs) for the system variables, this parameterisation leads to a specific structure of the state-space matrices of the linear QSDE, so that they must satisfy physical realisability (PR) conditions (James, Nurdin & Petersen (2008)) in order to correspond to a quantum oscillator with CCR preservation.
The PR constraints are a significant obstacle to solving coherent quantum feedback control problems, where a given quantum plant is in a measurement-free field-mediated or direct (Zhang & James (2011)) interconnection with a quantum controller, which has to stabilise the closed-loop system and meet optimality or robust performance criteria. One of such settings is the coherent quantum LQG (CQLQG) control problem (Nurdin, James & Petersen (2009)) of minimising an infinite-horizon mean square cost (for the plant variables and the controller output) over stabilising coherent quantum controllers, where both the plant and the controller are OQHOs (for example, with the same number of dynamic variables). Its classical counterpart (Kwakernaak & Sivan (1972)) admits a separation principle, which decomposes the optimal LQG controller into a Kalman filter for updating the conditional expectations of the plant variables, conditioned on the observations, and an actuator using the current plant state estimate, along with a pair of independent algebraic Riccati equations.
However, the CQLQG control problem does not lend itself to this particular combination of classical stochastic filtering and dynamic programming approaches because of the PR constraints mentioned above and the nature of quantum probability (Holevo (2001)). The latter describes the statistical properties of quantum processes in terms of density operators (or quantum states) on the underlying Hilbert space, which are more complicated than the scalar-valued classical probability measures and lead to the absence of classical joint distributions and conditional expectations for noncommutative quantum variables. Also, unlike classical observations, the noncommutative output fields of the quantum plant, which drive the coherent quantum controller, are not accessible to simultaneous measurement. On the other hand, the absence of measurements (which are accompanied by back-action effects and decoherence as the loss of quantum information) is an advantage of coherent quantum control by interconnection compared to the classical observation-actuation paradigm using digital signal processing.
The motivation behind the CQLQG control problem and the issue of obtaining an efficient solution for it explain the recurrent research interest to this problem (and its feedback-free versions on coherent quantum filtering (Miao & James (2012); Vladimirov & Petersen (2013b))) since its formulation in 2009. One of existing approaches to this problem is based on representing it as a constrained covariance control problem and applying variational methods of nonlinear functional analysis (in the form of Frechet differentiation of the mean square cost over the matrix-valued parameters (Vladimirov & Petersen (2013a))) in combination with symplectic geometric and homotopy techniques to the development of optimality conditions and numerical algorithms (Sichani, Vladimirov & Petersen (2017); Vladimirov & Petersen (2021)). Although the CQLQG control problem does not lend itself to a solution obeying the filtering-control separation principle with a Luenberger structure (Luenberger (1966)) (as a predictor-corrector scheme with a gain matrix with respect to an innovation process), the latter was discussed as an additional constraint, combined with the PR conditions, for coherent quantum observers in (Miao & James (2012)).
The present paper extends these ideas to a class of coherent quantum controllers with Luenberger dynamics. To this end, we use the freedom of assigning an arbitrary nonsingular CCR matrix to the controller variables (without affecting the LQG cost for the closed-loop system), including the negative of the CCR matrix of the plant variables. The latter is achieved by swapping the conjugate positions and momenta in the canonical representation of the quantum variables (or by applying the mirror reflections of (Simon (2000))). With the swapping transformation of the controller, the difference of the plant and controller variables (which corresponds to the plant state estimation error in the case of classical optimal LQG controllers) forms a quantum process with zero one-point CCR matrix. Similarly to the classical case, this difference process conveniently replaces the plant variables in the closed-loop system for coherent quantum controllers of Luenberger type. The latter imposes an additional constraint on the controller matrices in such a way that, together with the PR conditions, the gain matrices of the controller become dependent (through a quadratic constraint) and parameterise the dynamics and output matrices of the controller. This allows a matrix-valued Lagrange multiplier to be used in order to obtain first-order necessary conditions of optimality for this narrower class of coherent quantum controllers in the CQLQG control problem. The resulting optimality conditions involve a pair of coupled algebraic Lyapunov equations (ALEs) with block lower triangular matrices, which can simplify the analysis of their solution.
The paper is organised as follows. Sec. 2 specifies the class of quantum plants with field-mediated coherent quantum feedback. Sec. 3 reviews the PR conditions and parameterization of the closed-loop system in terms of the energy and coupling matrices. Sec. 4 describes the CQLQG control problem. Sec. 5 specifies the swapping transformation for the controller variables. Sec. 6 discusses the class of coherent quantum controllers of Luenberger type. Sec. 7 establishes first-order conditions of optimality for such controllers in the CQLQG control problem using the Lagrange multipliers. Sec. 8 makes concluding remarks.
2 Coherent Quantum Feedback
The CQLQG control setting (Nurdin, James & Petersen (2009)) involves a quantum plant and a coherent quantum controller in the form of multimode OQHOs. They are coupled to each other (see Fig. 1)
through a measurement-free feedback mediated by multichannel bosonic fields organised into column-vectors
(1) |
(the dependence on time is omitted for brevity) and specified below. In this field-mediated interconnection, the plant and controller are also coupled to external bosonic fields modelled by self-adjoint quantum Wiener processes and (with even , ) on symmetric Fock spaces (Hudson & Parthasarathy (1984)) , , respectively. These quantum noises are assembled into vectors
(2) |
where the augmented quantum Wiener process acts on the composite Fock space (with the tensor product of spaces or operators, including the Kronecker product of matrices), and their future-pointing increments have the Ito tables
(3) |
where the transpose applies to vectors or matrices of operators as if the latter were scalars. Here, , , are quantum Ito matrices given by
(4) | ||||
(5) |
with , where is the imaginary unit, and is the identity matrix of order . The matrices , , in (4), (5) (with the subspace of real antisymmetric matrices of order ) specify the CCRs
(6) |
where is the matrix of commutators between linear operators , which form vectors , . The block diagonal structure of in (5) comes from commutativity between the entries of , acting on different Fock spaces.
The plant and the controller are endowed with initial Hilbert spaces , and an even number of dynamic variables and , respectively, which are time-varying self-adjoint operators on the space
(7) |
where is the initial plant-controller space. With the same number of dynamic variables assumed for the plant and the controller (this plays an important role in what follows), counts their degrees of freedom. The plant and controller variables are assembled into vectors
(8) |
and satisfy the following CCRs with nonsingular matrices and :
(9) |
In line with the block diagonal structure of , the plant variables commute with the controller variables (considered at the same moment of time): , since these operators act initially (at time ) on different spaces , , and the system-field evolution preserves the one-point CCRs. Accordingly, the output fields and of the plant and the controller in (1) are time-varying self-adjoint operators on the system-field space in (7). The Heisenberg dynamics of the internal and output variables of the plant are described by linear QSDEs
(10) | ||||
(11) |
with given matrices , , , , . The structure of , , , will be specified in Sec. 3. The feedthrough matrix in (11) is formed from conjugate pairs of rows of a permutation matrix of order , so that is even and , with
(12) |
The quantum Ito matrix of the plant output in (11), defined by (similarly to (3)), is computed in terms of (4) as , and its imaginary part
(13) |
specifies the CCRs for the plant output :
(14) |
The QSDE (10) is driven by the external input field (as a quantum plant noise) and the controller output , similar to the actuator signal in classical linear control (Kwakernaak & Sivan (1972)). The QSDE (11) for the plant output resembles the equations for noise-corrupted observations with a “signal” part
(15) |
However, the quantum process differs qualitatively from the classical observations since the output fields are not accessible to simultaneous measurement as noncommuting quantum variables (Holevo (2001)) in view of the relation for all , whose right-hand side vanishes only at or .
The internal and output variables of the coherent quantum controller satisfy the linear QSDEs
(16) | ||||
(17) |
(similar to the plant dynamics (10), (11)), with matrices , , , , , where , in (16) are the gain matrices of the controller with respect to the controller noise and the plant output in (11). Similarly to in (12), the controller feedthrough matrix in (17) is also of full row rank and consists of conjugate pairs of rows of a permutation matrix of order , so that is even and satisfies , along with
(18) |
Accordingly, the quantum Ito matrix of the controller output fields in (17), defined by and computed as in terms of (4), has the imaginary part
(19) |
which, similarly to (14), describes the CCRs for the controller output :
(20) |
In what follows, the matrix (specifying the “amount” of noise in the controller output ) is fixed, while the matrices , , , in (16), (17) can be varied subject to PR constraints of Sec. 3. Similarly to (15), the drift vector
(21) |
in (17) plays the role of a “signal” part of the controller output as a quantum noise-corrupted actuator process.
The QSDEs (10), (11), (16), (17) govern the fully quantum closed-loop system in Fig. 1. By analogy with classical LQG control, the performance of the coherent quantum controller (with the process in (21) corresponding to the actuator signal) is described in Sec. 4 in terms of a mean square cost functional for an auxiliary quantum process
(22) |
where , are given matrices. The entries of are time-varying self-adjoint operators which are linear combinations of the plant variables and the controller output variables from (8), (21) whose relative importance is specified by the weighting matrices , . Similarly to the classical LQG control settings (Kwakernaak & Sivan (1972)), the matrix is of full column rank:
(23) |
so that all the entries of are penalized through in (22) for large mean square values. The matrices , are otherwise free from physical constraints, and their choice is part of the control design specifications. The process in (22) is expressed in terms of the combined vector of the plant and controller variables in (8) and governed by
(24) |
where the QSDE is driven by the quantum Wiener process in (2) on the Fock space . The matrices , , of the closed-loop system (24) are obtained by combining the QSDEs (10), (11), (16), (17) with (21), (22) as
(25) |
similarly to the classical case. While the matrices , in (22) can be arbitrary (subject to (23)), the matrices , of the QSDE in (24) are of specific structure which the fully quantum closed-loop system inherits from the plant and controller (James, Nurdin & Petersen (2008)), as reviewed in the next section.
3 Physical Realizability Constraints
The dynamics of the field-mediated coherent feedback interconnection are specified by the individual Hamiltonians , and the vectors , of operators of coupling of the plant and controller to the external fields and between each other. Here, is the energy matrix of the plant (with the subspace of real symmetric matrices of order ), and , are the matrices of coupling of the plant with the external input field and the controller output , respectively. Similarly, is the energy matrix of the controller, and , are the matrices of coupling of the controller with the external input field and the plant output ; see Fig. 1. These energy and coupling matrices parameterise the plant matrices , , , in (10), (11) and the controller matrices , , , in (16), (17) as
(26) | ||||
(27) | ||||
(28) | ||||
(29) |
with the matrices , given by (13), (19). The special structure of the plant matrices in (26), (27) and the controller matrices in (28), (29) leads to the PR conditions for the plant:
(30) | ||||
(31) |
and similar conditions for the controller (James, Nurdin & Petersen (2008)):
(32) | ||||
(33) |
with the PR constraints (32), (33) on the controller matrices , , , (the matrix is fixed as mentioned before) being the distinctive feature of coherent quantum control formulations.
The matrices , of the closed-loop system (24), expressed through the energy and coupling parameters by substituting (26)–(29) into (25), also satisfy PR conditions:
(34) |
which are similar to (30), (32) and secure the preservation of the CCRs (9). Here, from (5) is the CCR matrix for the combined quantum Wiener process in (2). While the gain matrices , of an arbitrary coherent quantum controller in (28), (29) (related by linear bijections to the coupling matrices , since ) are independent, the matrices , of such a controller are parameterized by the triple as
(35) | ||||
(36) |
(see Vladimirov & Petersen (2013a)). The relations (35), (36) couple the matrices , to , , thus making the stabilization of the closed-loop system and the optimization of the coherent quantum controller (16), (17) qualitatively different from the classical control problems (irrespective of performance criteria). In particular, (36) shows that an “inflow” of the external quantum noise (through a nonzero gain matrix ) is essential in order for such a controller to produce a useful output with a nonzero drift vector in (21). At the same time, due to (18) and the structure of , (satisfying and ) the linear map in (36) is surjective, so that any value of can be achieved by an appropriate choice of (for example, as ).
The PR conditions (32), (33) impose constraints on the controller matrices , , , even if the CCR matrix is not specified. More precisely, if has no centrally symmetric eigenvalues about the origin, and hence, the Kronecker sum is nonsingular, then is recovered from (32) in terms of the vectorization , and its substitution into (36) (assuming that ) makes the controller output matrix a function of , , .
4 CQLQG control problem
Similarly to classical LQG control, the performance of the closed-loop quantum system (24) is described by the infinite-horizon mean square cost
(37) |
(Nurdin, James & Petersen (2009)), where is the Frobenius inner product of matrices (Horn & Johnson (2007)), and
(38) |
The quantum expectation is over the density operator on the system-field space in (7), where is the initial plant-controller quantum state on , and is the vacuum field state on the Fock space . The limits in (37), (38) exist whenever the initial plant and controller variables have finite second moments, , and the closed-loop system is internally stable (the matrix in (25) is Hurwitz). In this case, is the controllability Gramian of the pair : , found uniquely from the ALE
(39) |
Up to the factor of , the cost in (37) is the squared -norm of a strictly proper transfer function with the state-space realization triple . The CCR matrix from (9) does not contribute to (37) since the subspaces , in are orthogonal in the sense of . The unique solution of the ALE (which combines (34), (39)), is the quantum covariance matrix of the invariant zero-mean Gaussian state (Parthasarathy (2010)) for the closed-loop system variables.
The CQLQG control problem (Nurdin, James & Petersen (2009)) is formulated as the minimization
(40) |
of the cost (37) over the controller matrices , , , subject to the PR constraints (32), (33) and the internal stability condition that in (25) is Hurwitz. Although the CCR matrix of the controller variables in this problem is usually fixed, there is a certain freedom in its choice, which is exploited in what follows.
5 Swapping in Controller Variables
For any nonsingular matrix , the transformation
(41) |
of the controller variables and their CCR matrix in (9), with the energy and coupling matrices of the controller in (28), (29) being transformed as , , (where ), does not affect the transfer function of the controller, and hence, the cost in (37) remains unchanged. Indeed, the matrices , in (25), (38) are transformed by (41) as and , whereby remains the same and so also does in (37), thus implying the invariance of . However, (41) can be used in order to assign a given CCR matrix to the controller variables (which is invariant only under the Lie group of symplectic similarity transformations identified with the set of matrices satisfying ). Since the same also applies to the plant variables, there exist nonsingular matrices which convert the nonsingular CCR matrices , to a canonical form:
(42) |
with from (4). The matrix is the CCR matrix for conjugate position-momentum pairs (with commutativity between them) assembled into a vector as
(43) |
so that , or equivalently, for all , where is the Kronecker delta. By swapping the positions and momenta in (43), the vector is transformed as , with and acquires the CCR matrix in view of (42). Therefore, the transformation matrix leads to . This transformation allows the controller variables to be assumed for what follows (without loss of generality, except for the condition ) to have the CCR matrix
(44) |
The same effect can be achieved by the mirror reflections as in (Simon (2000)). The relation (44) leads to commutativity between the entries (taken at the same moment of time) of an auxiliary quantum process
(45) |
which corresponds to the plant state estimation error in the case of classical optimal LQG controllers satisfying the separation principle (Kwakernaak & Sivan (1972)). More precisely, in view of (9), under the condition (44), the one-point CCR matrix of the difference process is zero:
(46) |
Nevertheless, is a substantially quantum process since and also because (46) describes only the one-point CCRs for , which does not prevent the two-point commutator matrix from being nonzero at different moments of time . The one-point CCRs for the processes , take the form
(47) |
where
(48) |
use the augmented vector of system variables from (8).
6 Luenberger Type Controller Dynamics
Consider a class of coherent quantum controllers of Luenberger type (Luenberger (1966)), whose internal dynamics (16) is represented as
(49) |
in accordance with the plant dynamics (10), (11) and the structure of the controller output (17). The Luenberger structure (49) imposes an additional constraint on the controller matrix :
(50) |
In this case, it is convenient to describe the closed-loop system dynamics in terms of the quantum processes from (45) and . Since they are related to the vector in (8) by (48), the matrix in (25) is transformed to a block lower triangular form
(51) |
where the last equality uses the Luenberger structure (50) of the matrix . The matrices , in (25) are transformed to
(52) | ||||
(53) |
The blocks of the matrices , , in (51)–(53) describe the coefficients of the QSDEs for the processes in (45), in (49) and in (24):
(54) | ||||
(55) | ||||
(56) |
The QSDE (54) for the process is autonomous (does not involve ) since the matrix in (51) is block lower triangular. The latter makes the internal stability of the closed-loop system equivalent to the Hurwitz property of the matrices and , as in the classical case. In order for these two conditions to be satisfied, it is necessary that the pair is detectable and is stabilizable. However, in the quantum case being considered,
(57) |
is a function of , obtained by substituting (44) into (36), with
(58) |
As a result, the fulfillment of the classical detectability and stabilizability conditions does not guarantee the existence of controller gain matrices , which make and in (57) Hurwitz, since the Luenberger structure (50), combined with the PR conditions, leads to the following constraint on , .
Theorem 1
From (44), (50) and the second PR conditions (31), (33) for the plant and the controller, it follows that
(60) |
By using (60) and the antisymmetry of the matrices , , along with the first PR condition (30) for the plant, the first PR condition (32) for the controller takes the form
The relation (59) is equivalent to the preservation of the CCRs (46) for the process by the QSDE (54), which can also be seen from the first diagonal -block of the relation obtained by representing (34) in terms of the matrices , , from (47), (51), (52).
Now, by assembling the controller gain matrices , into
(61) |
and using from (5), the condition (59) is represented as
(62) |
where the matrices
(63) |
are associated with the plant gain and feedthrough matrices , , and the controller feedthrough matrix (which are fixed). Completion of the square in (62) yields
(64) |
where
(65) | ||||
(66) |
and use is made of an orthogonal real antisymmetric matrix
(67) |
(so that ), computed with the aid of (4), (5), (13), (63). Similarly to (6), (14), (20), the matrix specifies the joint CCRs for the controller noise and the plant output as . In view of (62), (64), all the pairs satisfying (59) and organised as in (61) are described by the inclusion
(68) |
where, for any given matrix , the set
(69) |
is invariant under the right multiplication of its elements by symplectic matrices (whereby any is converted to since ).
Theorem 2
By Lemma 4 of Appendix A applied to solvability of the equation with from (66) and the nonsingular matrix in (67), the condition (70) implies that the set in (68) is nonempty.
Since the set (whose nonemptiness is guaranteed by (70)) is not an affine subspace, the existence of pairs , which satisfy (68) and make the matrices and in (57) Hurwitz, is a nontrivial open problem. In this regard, the following decompositions of the set (69) can appear to be useful:
(71) |
which are obtained from (64), provided at least one of the conditions or holds (each of them is stronger than (70)). For example, if , application of Lemma 4 shows that the set on the right-hand side of the first equality in (71) is nonempty for any . In this case, the stabilization part of the CQLQG control problem in the class of coherent quantum controllers with Luenberger dynamics is equivalent to finding a matrix such that the matrix in (57) is Hurwitz (provided is stabilizable) and the nonempty set contains a matrix which makes Hurwitz, provided is detectable.
7 Necessary Conditions of Optimality
For the class of coherent quantum controllers with Luenberger dynamics (49), (50), the CQLQG control problem (40) reduces to minimising the mean square cost in (37) over the controller gain matrices , subject to the constraint (59) along with the internal stability condition that and in (57) are Hurwitz. The first-order necessary conditions of optimality for such controllers are those of stationarity for the Lagrange function given by
(72) |
where the matrix is defined by (61), and is a Lagrange multiplier pertaining to the representation (62) of the constraint (59) whose left-hand side is -valued. The LQG cost in (37), which is invariant under the transformation of the system variables in (48), can be computed for any stabilizing Luenberger controller as
(73) |
Here, , are the controllability and observability Gramians for the matrix triple in (51)–(53), satisfying the ALEs
(74) |
and giving rise to the Hankelian
(75) |
which is a diagonalizable matrix whose eigenvalues are the squared Hankel singular values (Kwakernaak & Sivan (1972)). The matrices , , (and related matrices) are split into blocks , block rows and block columns , with , in accordance with the partitioning of the matrices , , into blocks , , (for example, , and ). The block lower triangular structure of the matrix in (51) (with ) allows the ALEs (74) to be represented as
(76) | |||
(77) | |||
(78) | |||
(79) | |||
(80) | |||
(81) |
For any stabilising Luenberger controller, the blocks of are computed by successively solving the ALE (76), the algebraic Sylvester equation (ASE) (77) and the ALE (78). In a similar fashion, the blocks of are obtained by solving the ALE (81), the ASE (80) and the ALE (79) and give rise to an auxiliary matrix
(82) |
Associated with these matrices and the Lagrange multiplier from (72) are self-adjoint operators
(83) | ||||
(84) |
on the Hilbert spaces , (with the Frobenius inner product), respectively. Here, is the sum of “sandwich” operators of the form specified by real matrices , and mapping an appropriately dimensioned real matrix to (so that ). The adjoint of such an operator is , and hence, is self-adjoint whenever the matrices , are both symmetric or both antisymmetric (see Section 7 and Appendix A of (Vladimirov & Petersen (2013a))).
Theorem 3
Under the conditions of Theorem 1, a stabilising coherent quantum controller with Luenberger dynamics (49), (50) is a stationary point of the Lagrange function (72) for the CQLQG control problem (40) if and only if it satisfies
(85) | ||||
(86) |
where the linear operators , are associated by (83), (84) with the Lagrange multiplier and the blocks of the Gramians , and the Hankelian in (74)–(82) for the closed-loop system (54)–(56).
The partial Frechet derivatives of (73) over , , as independent variables are
(87) |
(see (Skelton, Iwasaki & Grigoriadis (1998))). Similarly to (Vladimirov & Petersen (2013a)), the chain rule differentiation of the cost (73) as a composite function and of the independent variables , using (51)–(53), (58), (87) leads to
(88) | ||||
(89) |
Here, use is made of the identities and , along with the relation (58) represented in a sandwich operator form as , where is the matrix transpose operator, so that in view of the antisymmetry of the matrices , and self-adjointness of . By substituting (52), (53) into (88), (89), it follows that
(90) | ||||
(91) |
where use is also made of the identities and in view of (12), (52), (82). The Frechet differentiation of the constraint-related term of the Lagrange function in (72) with respect to in (61) yields
(92) |
where (63)–(67) are used. The corresponding partial Frechet derivatives in , are recovered as the blocks of (92):
(93) | ||||
(94) |
A combination of (90), (91) with (93), (94) and (83), (84) leads to
(95) | ||||
(96) |
The conditions of stationarity (85), (86) are now obtained by equating the Frechet derivatives of the Lagrange function in (95), (96) to zero.
The first-order necessary conditions of optimality for the CQLQG control problem in the class of Luenberger controllers, provided by Theorem 3, form a set of nonlinear algebraic equations for the controller gain matrices , and the Lagrange multiplier . They include (59) and the ALEs (74) which are coupled through (85), (86). These equations for a locally optimal coherent quantum controller can be solved numerically (for example, by using Newton or gradient descent iterative algorithms). Their theoretical analysis (as well as computational aspects) can benefit from the block triangular structure of the ALEs for the Gramians in (76)–(81) (which is a consequence of the Luenberger architecture) and will be discussed elsewhere.
8 Conclusion
In the context of the CQLQG control problem, we have considered a swapping transformation for the controller variables, leading to a difference process with zero one-point CCR matrix. We have discussed the interplay between the quantum PR conditions and the classical Luenberger structure, resulting in an additional quadratic constraint on the controller gain matrices. For the class of coherent quantum controllers with Luenberger dynamics, we have obtained the first-order necessary conditions of optimality, which involve coupled ALEs along with a matrix-valued Lagrange multiplier and “multisandwich” operators on appropriate matrix spaces.
References
- Holevo (2001) A.S.Holevo, Statistical Structure of Quantum Theory, Springer, Berlin, 2001.
- Horn & Johnson (2007) R.A.Horn, and C.R.Johnson, Matrix Analysis, Cambridge University Press, New York, 2007.
- Hudson & Parthasarathy (1984) R.L.Hudson, and K.R.Parthasarathy, Quantum Ito’s formula and stochastic evolutions, Commun. Math. Phys., vol. 93, 1984, pp. 301–323.
- James, Nurdin & Petersen (2008) M.R.James, H.I.Nurdin, and I.R.Petersen, control of linear quantum stochastic systems, IEEE Trans. Automat. Contr., vol. 53, no. 8, 2008, pp. 1787–1803.
- Kwakernaak & Sivan (1972) H.Kwakernaak, and R.Sivan, Linear Optimal Control Systems, Wiley, New York, 1972.
- Luenberger (1966) D.Luenberger, Observers for multivariable systems, IEEE Trans. Automat. Contr., vol. 11, no. 2, 1966, pp. 190–197.
- Miao & James (2012) Z.Miao, and M.R.James, Quantum observer for linear quantum stochastic systems, 51st IEEE Conf. Decision Control, Maui, Hawaii, USA, December 10-13, 2012, pp. 1680–1684.
- Nurdin, James & Petersen (2009) H.I.Nurdin, M.R.James, and I.R.Petersen, Coherent quantum LQG control, Automatica, vol. 45, 2009, pp. 1837–1846.
- Parthasarathy (1992) K.R.Parthasarathy, An Introduction to Quantum Stochastic Calculus, Birkhäuser, Basel, 1992.
- Parthasarathy (2010) K.R.Parthasarathy, What is a Gaussian state?, Commun. Stoch. Anal., vol. 4, no. 2, 2010, pp. 143–160.
- Sakurai (1994) J.J.Sakurai, Modern Quantum Mechanics, Addison-Wesley, Reading, Mass., 1994.
- Sichani, Vladimirov & Petersen (2017) A.Kh.Sichani, I.G.Vladimirov, and I.R.Petersen, A numerical approach to optimal coherent quantum LQG controller design using gradient descent, Automatica, vol. 85, 2017, pp. 314–326.
- Simon (2000) R.Simon, Peres-Horodecki separability criterion for continuous variable systems, Phys. Rev. Lett., vol. 84, no. 12, 2000, pp. 2726–2729.
- Skelton, Iwasaki & Grigoriadis (1998) R.E.Skelton, T.Iwasaki, and K.M.Grigoriadis, A Unified Algebraic Approach to Linear Control Design, Taylor & Francis, London, 1998.
- Vladimirov & Petersen (2013a) I.G.Vladimirov, and I.R.Petersen, A quasi-separation principle and Newton-like scheme for coherent quantum LQG control, Syst. Contr. Lett., vol. 62, no. 7, 2013, pp. 550–559.
- Vladimirov & Petersen (2013b) I.G.Vladimirov, and I.R.Petersen, Coherent quantum filtering for physically realizable linear quantum plants, European Control Conference, IEEE, Zurich, Switzerland, 17-19 July 2013, pp. 2717–2723.
- Vladimirov & Petersen (2021) I.G.Vladimirov, and I.R.Petersen, A homotopy approach to coherent quantum LQG control synthesis using discounted performance criteria, IFAC PapersOnLine, vol. 54, no. 9, 2021, pp. 166–171.
- Zhang & James (2011) G.Zhang, and M.R.James, Direct and indirect couplings in coherent feedback control of linear quantum systems, IEEE Trans. Automat. Contr., vol. 56, no. 7, 2011, 1535–1550.
Appendix A Special Quadratic Equations
Lemma 4
For any nonsingular matrix and any matrix of even order , there exists a matrix satisfying
(97) |
Since is a nonsingular real antisymmetric matrix (and hence, its order is even), it is representable as
(98) |
in terms of a nonsingular matrix , with from (4). In a similar fashion, since is even and , there exists a matrix (singular if so is ) such that
(99) |
Due to the assumption , the last equality in (99) is obtained by padding with zeros to a -matrix and using the partitioning . Since , it follows from (98), (99) that (97) is satisfied, for example, with .
A slight modification of the proof extends Lemma 4 to the case when the order of the matrix is odd.