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

Optimization of Partially Isolated Quantum Harmonic Oscillator Memory Systems by Mean Square Decoherence Time Criteria

Igor G. Vladimirov1,  Ian R. Petersen2 This work is supported by the Australian Research Council grant DP240101494.1,2School of Engineering, Australian National University, ACT 2601, Canberra, Australia: [email protected], [email protected].
Abstract

This paper is concerned with open quantum harmonic oscillators with position-momentum system variables, whose internal dynamics and interaction with the environment are governed by linear quantum stochastic differential equations. A recently proposed approach to such systems as Heisenberg picture quantum memories exploits their ability to approximately retain initial conditions over a decoherence horizon. Using the quantum memory decoherence time defined previously in terms of a fidelity threshold on a weighted mean-square deviation of the system variables from their initial values, we apply this approach to a partially isolated subsystem of the oscillator, which is not directly affected by the external fields. The partial isolation leads to an appropriate system decomposition and a qualitatively different short-horizon asymptotic behaviour of the deviation, which yields a longer decoherence time in the high-fidelity limit. The resulting approximate decoherence time maximization over the energy parameters for improving the quantum memory performance is discussed for a coherent feedback interconnection of such systems.

I INTRODUCTION

The quantum communication and quantum information processing technologies, which undergo extensive theoretical and practical development, substantially rely on the possibility to manipulate and engineer quantum mechanical systems with novel properties exploiting nonclassical resources. The latter come from the noncommutative operator-valued structure of quantum variables and quantum states reflected in the Heisenberg and Schrödinger pictures of quantum dynamics and quantum probability, and the theory of quantum measurement [7]. One of the aspects of these approaches is that the fragile nature of quantum dynamics and quantum states, involving the atomic and subatomic scales [13], complicates the state initialization for such systems and their isolation from the environment in a controlled fashion. In particular, the controlled isolation is important in the context of quantum computation paradigms [16] where unitary transformations, performed with closed quantum systems [21] between the measurements, play a crucial role.

The absence of dissipation makes the perfectly reversible unitary dynamics of an isolated quantum system (including the preservation of the Hamiltonian and of the system variables or states in the zero-Hamiltonian case) particularly useful for storing quantum information. The storage phase is preceded by the “writing” stage and followed by the “reading” stage which can employ, for example, photonics-based platforms using light-matter interaction effects (see [4, 6, 28, 29, 30] and references therein). However, the ability to retain quantum states or dynamic variables over the course of time, which is required for the storage phase, is corrupted by the unavoidable coupling of the quantum system to its environment such as external fields.

The system-field interaction, even when it is weak, gives rise to quantum noise which makes the system drift away in a dissipative fashion from its initial condition as opposed to the ideal case of isolated zero-Hamiltonian system dynamics. The resulting open quantum systems are modelled by Hudson-Parthasarathy quantum stochastic differential equations (QSDEs) [10, 18] which are linear for open quantum harmonic oscillators (OQHOs) [11, 17, 19, 31] and quasi-linear for finite-level quantum systems [3, 24]. While the latter are particularly relevant for modelling qubit registers in the quantum computing context, continuous variables systems (such as the OQHOs with quantum positions and momenta) are also employed in quantum information theory including the Gaussian quantum channel models [8].

Using the quantum calculus framework, an approach has recently been proposed in [26, 27] for both classes of open quantum systems to their performance as a quantum memory in the storage phase. The system performance index is a decoherence time horizon at which a weighted mean-square deviation of the system variables from their initial values reaches a given fidelity threshold (note that different yet related decoherence measures are also discussed in [25] and references therein). The memory decoherence time (or its high-fidelity asymptotic approximation) can therefore be maximized over the energy and coupling parameters of a quantum network in order to improve its ability to approximately retain the initial conditions.

In the present paper, we follow the approach of [26] to OQHOs as Heisenberg picture quantum memory systems with position-momentum variables, whose internal dynamics and interaction with the environment are governed by linear QSDEs. Using the previously defined memory decoherence time, we apply this approach to a partially isolated subsystem of the oscillator, which is affected by the external fields only indirectly through another subsystem. The partial subsystem isolation from the external fields and the related system decomposition hold under a certain rank condition on the system-field coupling matrix and form an important special case which was not considered in [26]. This setting leads to a qualitatively different short-horizon asymptotic behaviour of the mean-square deviation of the subsystem variables from their initial values and yields a longer decoherence time in the high-fidelity limit. The maximization of the resulting approximate decoherence time over the energy parameters for improving the quantum memory performance is discussed for a coherent feedback interconnection of such systems involving the direct energy coupling and field-mediated coupling [32] between the constituent OQHOs.

The paper is organised as follows. Section II provides a minimum background material on QSDEs for open quantum stochastic systems. Section III specifies the class of OQHOs with position-momentum system variables and linear-quadratic energetics. Section IV considers weighted deviations of the system variables from their initial conditions and discusses a partially isolated subsystem along with the system decomposition. Section V reviews the mean-square quantification of such deviations and the memory decoherence time and studies the short-horizon and high-fidelity asymptotic behaviour of these functionals for the partially isolated subsystem. Section VI solves the approximate decoherence time maximization problem for a partially isolated subsystem of a coherent feedback interconnection of two OQHOs with direct energy and indirect field-mediated coupling.

II QUANTUM STOCHASTIC DYNAMICS

As mentioned above, in comparison with the idealisation of completely isolated quantum dynamics, more realistic settings consider open quantum systems which interact with the environment, such as quantum fields or classical measuring devices. In the framework of the Hudson-Parthasarathy quantum stochastic calculus [10, 18], an open quantum system subject to external quantum fields, as shown in Fig. 1,

openquantumsystemW:=[W1Wm]W:={\scriptsize\begin{bmatrix}W_{1}\\ \vdots\\ W_{m}\end{bmatrix}}Y:=[Y1Yr]Y:={\scriptsize\begin{bmatrix}Y_{1}\\ \vdots\\ Y_{r}\end{bmatrix}}
Figure 1: An open quantum system with vectors WW, YY of input quantum Wiener processes and output fields.

is governed by QSDEs

dX=𝒢(X)dti[X,LT]dW,dY=2DJLdt+DdW\mathrm{d}X=\mathcal{G}(X)\mathrm{d}t-i[X,L^{\mathrm{T}}]\mathrm{d}W,\qquad\mathrm{d}Y=2DJL\mathrm{d}t+D\mathrm{d}W (1)

(the time arguments will often be omitted for brevity), which describe the Heisenberg evolution [21] of the internal and output variables. Here, i:=1i:=\sqrt{-1} is the imaginary unit, and the reduced Planck constant is set to =1\hslash=1 using atomic units for convenience. Also, X:=(Xk)1knX:=(X_{k})_{1\leqslant k\leqslant n}, YY, WW are vectors of internal, output and input dynamic variables, respectively, and L:=(Lk)1kmL:=(L_{k})_{1\leqslant k\leqslant m} is a vector of system-field coupling operators. Unless indicated otherwise, vectors are organised as columns. Accordingly, [X,LT]:=(XjLkLkXj)1jn,1km=XLT(LXT)T[X,L^{\mathrm{T}}]:=(X_{j}L_{k}-L_{k}X_{j})_{1\leqslant j\leqslant n,1\leqslant k\leqslant m}=XL^{\mathrm{T}}-(LX^{\mathrm{T}})^{\mathrm{T}} is the commutator matrix. The system variables X1,,XnX_{1},\ldots,X_{n} and the output field variables Y1,,YrY_{1},\ldots,Y_{r} are time-varying self-adjoint operators on the system-field tensor-product Hilbert space

:=0𝔉,\mathfrak{H}:=\mathfrak{H}_{0}\otimes\mathfrak{F}, (2)

with the system variables acting initially, at time t=0t=0, on a Hilbert space 0\mathfrak{H}_{0}. The input bosonic field variables W1,,WmW_{1},\ldots,W_{m} are quantum Wiener processes which are time-varying self-adjoint operators on a symmetric Fock space 𝔉\mathfrak{F}. The map 𝒢\mathcal{G} in the drift term of the first QSDE in (1) is the Gorini-Kossakowski-Sudarshan-Lindblad superoperator [5, 14] acting on a system operator ξ\xi (such as a function of the system variables X1,,XnX_{1},\ldots,X_{n}) as

𝒢(ξ):=i[H,ξ]+𝒟(ξ).\mathcal{G}(\xi):=i[H,\xi]+\mathcal{D}(\xi). (3)

This is a quantum counterpart of the infinitesimal generators of classical diffusion processes [12], with 𝒟\mathcal{D} the decoherence superoperator given by

𝒟(ξ):=12([LT,ξ]ΩL+LTΩ[ξ,L]).\mathcal{D}(\xi):=\frac{1}{2}([L^{\mathrm{T}},\xi]\Omega L+L^{\mathrm{T}}\Omega[\xi,L]). (4)

The Hamiltonian HH in (3) and the system-field coupling operators L1,,LmL_{1},\ldots,L_{m} in (1), (4) are self-adjoint operator-valued functions of X1,,XnX_{1},\ldots,X_{n}. The matrix Dr×mD\in{\mathbb{R}}^{r\times m} in (1) consists of rmr\leqslant m conjugate rows of a permutation matrix of order mm, where both rr and mm are even. Thus, DD is a co-isometry: DDT=IrDD^{\mathrm{T}}=I_{r}, where IrI_{r} is the identity matrix of order rr. The meaning of DD is that it selects some of the mm output fields. For example, the case of r=mr=m and D=ImD=I_{m} corresponds to selecting all of the output fields. The vector WW of quantum Wiener processes W1,,WmW_{1},\ldots,W_{m}, which drive the QSDEs (1), has the quantum Ito table

dWdWT=Ωdt,Ω:=Im+iJ=Ω0,\mathrm{d}W\mathrm{d}W^{\mathrm{T}}=\Omega\mathrm{d}t,\qquad\Omega:=I_{m}+iJ=\Omega^{*}\succcurlyeq 0, (5)

where ():=()¯T(\cdot)^{*}:={\overline{(\cdot)}}{}^{\mathrm{T}} is the complex conjugate transpose. Here, J:=ImΩJ:=\mathrm{Im}\Omega is a real antisymmetric matrix of order mm given by

J:=Im/2𝐉,𝐉:=[0110]J:=I_{m/2}\otimes\mathbf{J},\qquad\mathbf{J}:={\begin{bmatrix}0&1\\ -1&0\end{bmatrix}} (6)

and specifying the two-point canonical commutation relations (CCRs)

[W(s),W(t)T]=2imin(s,t)J,s,t0,[W(s),W(t)^{\mathrm{T}}]=2i\min(s,t)J,\qquad s,t\geqslant 0, (7)

where JJ represents its tensor product J𝔉J\otimes\mathcal{I}_{\mathfrak{F}} with the identity operator 𝔉\mathcal{I}_{\mathfrak{F}} on the Fock space 𝔉\mathfrak{F}.

III OPEN QUANTUM HARMONIC OSCILLATORS

The CCRs (7) are typical for unbounded operators of position-momentum type [21] on a dense domain of an infinite-dimensional Hilbert space. Such quantum variables, when they are considered as the system variables X1,,XnX_{1},\ldots,X_{n} with an even

n:=2ν,n:=2\nu, (8)

are provided by the quantum mechanical positions q1,,qνq_{1},\ldots,q_{\nu} and momenta p1,,pνp_{1},\ldots,p_{\nu} which can be implemented as the multiplication and differential operators

X2k1:=qk,X2k:=pk:=iqk,k=1,,ν,X_{2k-1}:=q_{k},\qquad X_{2k}:=p_{k}:=-i\partial_{q_{k}},\qquad k=1,\ldots,\nu, (9)

on the Schwartz space of rapidly decreasing functions [20] on ν{\mathbb{R}}^{\nu}. At any time t0t\geqslant 0, these operators satisfy the one-point CCRs

[X(t),X(t)T]=2iΘ,Θ:=12Iν𝐉.[X(t),X(t)^{\mathrm{T}}]=2i\Theta,\qquad\Theta:=\frac{1}{2}I_{\nu}\otimes\mathbf{J}. (10)

As continuous quantum variables, the conjugate position-momentum pairs (9) are qualitatively different from those in finite-level quantum systems [3, 24] and result from quantizing the classical positions and momenta of Hamiltonian mechanics [1]. They are employed as system variables in OQHOs [11, 17, 19, 31]. For an OQHO, the Hamiltonian HH and the system-field coupling operators L1,,LmL_{1},\ldots,L_{m} are quadratic and linear functions of the system variables specified by an energy matrix R=RTn×nR=R^{\mathrm{T}}\in{\mathbb{R}}^{n\times n} and a coupling matrix Mm×nM\in{\mathbb{R}}^{m\times n} as

H:=12XTRX,L:=MX.H:=\frac{1}{2}X^{\mathrm{T}}RX,\qquad L:=MX. (11)

The general QSDEs (1) then acquire the form of linear QSDEs

dX=AXdt+BdW,dY=CXdt+DdW,\mathrm{d}X=AX\mathrm{d}t+B\mathrm{d}W,\qquad\mathrm{d}Y=CX\mathrm{d}t+D\mathrm{d}W, (12)

where the matrices An×nA\in{\mathbb{R}}^{n\times n}, Bn×mB\in{\mathbb{R}}^{n\times m}, Cr×nC\in{\mathbb{R}}^{r\times n} are computed as

A:=A0+A~,B:=2ΘMT,C:=2DJM,A:=A_{0}+\widetilde{A},\qquad B:=2\Theta M^{\mathrm{T}},\qquad C:=2DJM, (13)

with

A0:=2ΘR,A~:=2ΘMTJM,A_{0}:=2\Theta R,\qquad\widetilde{A}:=2\Theta M^{\mathrm{T}}JM, (14)

in terms of the CCR matrices JJ, Θ\Theta from (6), (10) and the energy and coupling matrices RR, MM from (11). In view of (14), the dynamics matrix AA in (13) depends linearly on RR and quadratically on MM. Moreover, in addition to the noncommutative operator-valued nature of the quantum variables, the specific representation of the coefficients of the QSDEs (12) in terms of the commutation, energy and coupling parameters imposes quantum physical realizability constraints [11] in contrast to classical SDEs [12].

Despite the qualitative differences, OQHOs share some common features with classical linear stochastic systems due to the linearity of the QSDEs (12). In particular, the solution of the first QSDE is given by

X(t)=etAX(0)+Z(t),Z(t):=0te(ts)ABdW(s)X(t)=\mathrm{e}^{tA}X(0)+Z(t),\qquad Z(t):=\int_{0}^{t}\mathrm{e}^{(t-s)A}B\mathrm{d}W(s) (15)

for any t0t\geqslant 0 and consists of the system responses to the initial condition X(0)X(0) and the driving quantum Wiener process WW over the time interval [0,t][0,t], with Z(0)=0Z(0)=0. Here, the quantum process ZZ of nn time-varying self-adjoint operators Z1,,ZnZ_{1},\ldots,Z_{n} is adapted to the filtration of the Fock space 𝔉\mathfrak{F}, and hence, X(0)X(0) and Z(t)Z(t) commute with each other as operators acting on different spaces (0\mathfrak{H}_{0} and 𝔉\mathfrak{F}):

[X(0),Z(t)T]=0,t0.[X(0),Z(t)^{\mathrm{T}}]=0,\qquad t\geqslant 0. (16)

Furthermore, if the system-field quantum state on the space (2) has the form

ρ:=ρ0υ,\rho:=\rho_{0}\otimes\upsilon, (17)

where ρ0\rho_{0} is the initial system state on 0\mathfrak{H}_{0}, and υ\upsilon is the vacuum field state [18] on 𝔉\mathfrak{F}, then ZZ is a Gaussian quantum process (see, for example, [23, Section 3]), which is statistically independent of X(0)X(0) and has zero mean:

𝐄Z(t)=(Tr(υZk(t)))1kn=0.\mathbf{E}Z(t)=(\mathrm{Tr}(\upsilon Z_{k}(t)))_{1\leqslant k\leqslant n}=0. (18)

Here, use is made of the quantum expectation 𝐄ζ:=Tr(ρζ)\mathbf{E}\zeta:=\mathrm{Tr}(\rho\zeta) of an operator ζ\zeta on the system-field space (2) over the quantum state (17), which reduces in (18) to the averaging over the vacuum state υ\upsilon since the operators ZkZ_{k} act on the Fock space 𝔉\mathfrak{F}. In combination with the independence between X(0)X(0) and Z(t)Z(t), (18) implies that

𝐄(X(0)Z(t)T)=𝐄X(0)𝐄Z(t)T=0,\mathbf{E}(X(0)Z(t)^{\mathrm{T}})=\mathbf{E}X(0)\mathbf{E}Z(t)^{\mathrm{T}}=0, (19)

where 𝐄X(0)=(Tr(ρ0Xk(0)))1kn\mathbf{E}X(0)=(\mathrm{Tr}(\rho_{0}X_{k}(0)))_{1\leqslant k\leqslant n}. The above properties of ZZ hold regardless of a particular structure of the initial system state ρ0\rho_{0}.

IV PARTIAL SUBSYSTEM ISOLATION

The process ZZ in (15), caused by the interaction of the OQHO with the external fields, plays the role of a noise which corrupts the ability of the system as a quantum memory to retain the initial condition X(0)X(0). Similarly to [26, 27], we describe the quantum memory performance in terms of a quadratic form

Q(t):=ξ(t)TΣξ(t)Q(t):=\xi(t)^{\mathrm{T}}\Sigma\xi(t) (20)

of the deviation

ξ(t):=X(t)X(0)=αtX(0)+Z(t)\xi(t):=X(t)-X(0)=\alpha_{t}X(0)+Z(t) (21)

of the system variables at time t0t\geqslant 0 from their initial values, where

αt:=etAIn,\alpha_{t}:=\mathrm{e}^{tA}-I_{n}, (22)

with ξ(0)=0\xi(0)=0, Q(0)=0Q(0)=0 and α0=0\alpha_{0}=0. Here, Σ\Sigma is a real positive semi-definite symmetric matrix of order nn factorised as

Σ:=FTF,Fs×n,s:=rankΣn,\Sigma:=F^{\mathrm{T}}F,\qquad F\in{\mathbb{R}}^{s\times n},\qquad s:=\mathrm{rank}\Sigma\leqslant n, (23)

so that the rows of FF specify the coefficients of ss independent linear combinations of the deviations in (21) forming the vector

η(t):=Fξ(t)=FX(t)FX(0)=FαtX(0)+FZ(t),\eta(t):=F\xi(t)=FX(t)-FX(0)=F\alpha_{t}X(0)+FZ(t), (24)

in terms of which the quantum process QQ in (20) is represented as

Q(t):=η(t)Tη(t).Q(t):=\eta(t)^{\mathrm{T}}\eta(t). (25)

As discussed below, the matrix FF can be used for selecting those linear combinations (in particular, a subset) of the system variables, whose dynamics are affected by the external fields only indirectly and thus pertain to a partially isolated subsystem of the OQHO.

Lemma 1

Suppose the coupling matrix MM of the OQHO in (11) satisfies

d:=nrankM>0.d:=n-\mathrm{rank}M>0. (26)

Then for any sds\leqslant d, there exists a full row rank matrix Fs×nF\in{\mathbb{R}}^{s\times n} such that the vector

φ(t):=FX(t)\varphi(t):=FX(t) (27)

of ss time-varying self-adjoint operators on the space (2) satisfies the ODE

𝜑(t)=GX(t),G:=FA0,\mathop{\varphi}^{\centerdot}(t)=GX(t),\qquad G:=FA_{0}, (28)

where A0A_{0} is the matrix from (14). \square

Proof:

Under the condition (26), the columns of the matrix MTn×mM^{\mathrm{T}}\in{\mathbb{R}}^{n\times m} are linearly dependent and span in n{\mathbb{R}}^{n} a proper subspace im(MT)\mathrm{im}(M^{\mathrm{T}}) of dimension rankM\mathrm{rank}M, whose orthogonal complement kerM\ker M has dimension dd. In combination with the nonsingularity of the CCR matrix Θ\Theta in (10), this implies that for any sds\leqslant d, there exists a full row rank matrix Fs×nF\in{\mathbb{R}}^{s\times n} such that the rows of FΘF\Theta are orthogonal to all the columns of MTM^{\mathrm{T}}, that is, FΘMT=0F\Theta M^{\mathrm{T}}=0. By the parameterization of the matrix BB in (13), any such FF satisfies

FB=2FΘMT=0,FB=2F\Theta M^{\mathrm{T}}=0, (29)

and hence, the left multiplication of the first QSDE in (12) by FF yields

dφ=FAXdt+FBdW=FAXdt\mathrm{d}\varphi=FAX\mathrm{d}t+FB\mathrm{d}W=FAX\mathrm{d}t (30)

for the quantum process (27). Since the diffusion term in (30) vanishes, this QSDE is an ODE: φ˙=FAX\dot{\varphi}=FAX, which establishes (28) since

FA=FA0+FBJM=GFA=FA_{0}+FBJM=G (31)

by the first equality in (13) and the representation A~=BJM\widetilde{A}=BJM of A~\widetilde{A} in (14) in terms of BB. ∎

Since rankMmin(m,n)m\mathrm{rank}M\leqslant\min(m,n)\leqslant m, then a sufficient condition for (26) is provided by the inequality n>mn>m. Furthermore, if the null spaces of the matrices FF, GG in (28) satisfy kerFkerG\ker F\subset\ker G and hence, G=NFG=NF for some matrix Ns×sN\in{\mathbb{R}}^{s\times s}, then the ODE in (28) becomes autonomous: φ˙=Nφ\dot{\varphi}=N\varphi, in which case, the dynamics of φ\varphi are completely isolated from the external fields.

More generally, the matrix FF from Lemma 1 can be augmented by a matrix T(ns)×nT\in{\mathbb{R}}^{(n-s)\times n} to a nonsingular (n×n)(n\times n)-matrix:

detS0,S:=[FT]n×n.\det S\neq 0,\qquad S:={\begin{bmatrix}F\\ T\end{bmatrix}}\in{\mathbb{R}}^{n\times n}. (32)

By partitioning the inverse matrix S1S^{-1} into blocks S1n×sS_{1}\in{\mathbb{R}}^{n\times s} and S2n×(ns)S_{2}\in{\mathbb{R}}^{n\times(n-s)} as S1:=[S1S2]S^{-1}:=\begin{bmatrix}S_{1}&S_{2}\end{bmatrix} and introducing a quantum process ψ\psi of nsn-s time-varying self-adjoint operators on (2), so that

X=S1ζ=S1φ+S2ψ,ζ:=[φψ],ψ:=TX,X=S^{-1}\zeta=S_{1}\varphi+S_{2}\psi,\qquad\zeta:={\begin{bmatrix}\varphi\\ \psi\end{bmatrix}},\qquad\psi:=TX, (33)

it follows that the first QSDE in (12) can be decomposed into an ODE and a QSDE with respect to the transformed system variables as

𝜑=a11φ+a12ψ,dψ=(a21φ+a22ψ)dt+bdW,\mathop{\varphi}^{\centerdot}=a_{11}\varphi+a_{12}\psi,\qquad\mathrm{d}\psi=(a_{21}\varphi+a_{22}\psi)\mathrm{d}t+b\mathrm{d}W, (34)

where

a:=SAS1:=[a11a12a21a22]=[GS1GS2TAS1TAS2],b:=TB.a:=SAS^{-1}:={\begin{bmatrix}a_{11}&a_{12}\\ a_{21}&a_{22}\end{bmatrix}}={\begin{bmatrix}GS_{1}&GS_{2}\\ TAS_{1}&TAS_{2}\end{bmatrix}},\qquad b:=TB. (35)

Therefore, the internal dynamics of the OQHO can be represented as a feedback interconnection of linear systems Φ\Phi and Ψ\Psi, with the corresponding vectors φ\varphi, ψ\psi of system variables, where only Ψ\Psi interacts directly with the quantum Wiener process WW, as shown in Fig. 2.

φ\varphiψ\psiΦ\PhiΨ\PsiWW
Figure 2: A schematic representation of (34) as an interconnection of systems Φ\Phi, Ψ\Psi, where Φ\Phi is affected by the external fields WW only through Ψ\Psi which interacts with WW.

With a slight abuse of notation, the transfer functions of these systems are computed in terms of the matrices (35) as

Φ(u)\displaystyle\Phi(u) =(uIsa11)1a12,\displaystyle=(uI_{s}-a_{11})^{-1}a_{12}, (36)
Ψ(u)\displaystyle\Psi(u) =[Ψ1(u)Ψ2(u)]=(uInsa22)1[a21b]\displaystyle=\begin{bmatrix}\Psi_{1}(u)&\Psi_{2}(u)\end{bmatrix}=(uI_{n-s}-a_{22})^{-1}\begin{bmatrix}a_{21}&b\end{bmatrix} (37)

and relate the Laplace transforms

φ^(u)\displaystyle\widehat{\varphi}(u) :=0+eutφ(t)dt,ψ^(u):=0+eutψ(t)dt,\displaystyle:=\int_{0}^{+\infty}\mathrm{e}^{-ut}\varphi(t)\mathrm{d}t,\qquad\widehat{\psi}(u):=\int_{0}^{+\infty}\mathrm{e}^{-ut}\psi(t)\mathrm{d}t, (38)
W^(u)\displaystyle\widehat{W}(u) :=0+eutdW(t)\displaystyle:=\int_{0}^{+\infty}\mathrm{e}^{-ut}\mathrm{d}W(t) (39)

(with the integral in (39) being of the Ito type in contrast to those in (38)) of the quantum processes φ\varphi, ψ\psi, WW for all uu\in\mathbb{C} with sufficiently large Reu>0\mathrm{Re}u>0 for convergence of the integrals as

φ^(u)\displaystyle\widehat{\varphi}(u) =χ1(u)ζ(0)+Φ(u)ψ^(u),\displaystyle=\chi_{1}(u)\zeta(0)+\Phi(u)\widehat{\psi}(u), (40)
ψ^(u)\displaystyle\widehat{\psi}(u) =χ2(u)ζ(0)+Ψ(u)[φ^(u)W^(u)]\displaystyle=\chi_{2}(u)\zeta(0)+\Psi(u){\begin{bmatrix}\widehat{\varphi}(u)\\ \widehat{W}(u)\end{bmatrix}}
=χ2(u)ζ(0)+Ψ1(u)φ^(u)+Ψ2(u)W^(u).\displaystyle=\chi_{2}(u)\zeta(0)+\Psi_{1}(u)\widehat{\varphi}(u)+\Psi_{2}(u)\widehat{W}(u). (41)

Here, χ1\chi_{1}, χ2\chi_{2} are (s×n)(s\times n) and (ns)×n(n-s)\times n-blocks of the transfer function χ(u):=[χ1(u)χ2(u)]=(uIna)1\chi(u):={\small\begin{bmatrix}\chi_{1}(u)\\ \chi_{2}(u)\end{bmatrix}}=(uI_{n}-a)^{-1} associated with the response of the quantum process ζ\zeta in (33) to its initial condition ζ(0)\zeta(0). The equations (40), (41) can be solved for φ^\widehat{\varphi} as

φ^(u)\displaystyle\widehat{\varphi}(u) =(χ1(u)+Γ(u)χ2(u))ζ(0)+Γ(u)Ψ2(u)W^(u)\displaystyle=(\chi_{1}(u)+\Gamma(u)\chi_{2}(u))\zeta(0)+\Gamma(u)\Psi_{2}(u)\widehat{W}(u)
=[χ1(u)+Γ(u)χ2(u)Γ(u)Ψ2(u)][ζ(0)W^(u)],\displaystyle=\begin{bmatrix}\chi_{1}(u)+\Gamma(u)\chi_{2}(u)&\Gamma(u)\Psi_{2}(u)\end{bmatrix}{\begin{bmatrix}\zeta(0)\\ \widehat{W}(u)\end{bmatrix}}, (42)

where Γ\Gamma is an auxiliary transfer function associated with (36), (37) by

Γ(u):=Φ(u)(InsΨ1(u)Φ(u))1.\Gamma(u):=\Phi(u)(I_{n-s}-\Psi_{1}(u)\Phi(u))^{-1}. (43)

The Laplace transform of the quantum process η\eta in (24) is then found by substituting (42) into

η^(u):=0+eutη(t)dt=φ^(u)1uφ(0),\widehat{\eta}(u):=\int_{0}^{+\infty}\mathrm{e}^{-ut}\eta(t)\mathrm{d}t=\widehat{\varphi}(u)-\frac{1}{u}\varphi(0), (44)

where φ(0)=FX(0)\varphi(0)=FX(0) from (27) is the subvector of ζ(0)\zeta(0) in (33). Note that the resulting frequency-domain representation (44) of η\eta does not depend on a particular choice of the matrix TT in (32). In view of (42), the subsystem Φ\Phi with the vector φ\varphi of transformed system variables can be approximately isolated from the external fields WW by making the transfer function ΓΨ2\Gamma\Psi_{2} “small” in a suitable sense. By (43), this requires “smallness” not only of Φ\Phi and Ψ2\Psi_{2}, but also of Ψ1Φ\Psi_{1}\Phi due to the presence of the feedback loop in Fig. 2, which is typical for small-gain theorem arguments (see, for example, [2] and references therein).

V MEAN-SQUARE MEMORY DECOHERENCE TIME

Following [26, 27], for any given time t0t\geqslant 0, we will quantify the “size” of the deviation η(t)\eta(t) in (24) by a mean-square functional

Δ(t):=𝐄Q(t)=Σ,ReΥ(t)=FαtP2+Σ,ReV(t).\Delta(t):=\mathbf{E}Q(t)=\langle\Sigma,\mathrm{Re}\Upsilon(t)\rangle=\|F\alpha_{t}\sqrt{P}\|^{2}+\langle\Sigma,\mathrm{Re}V(t)\rangle. (45)

This quantity provides an upper bound on the tail probabilities for the positive semi-definite self-adjoint quantum variable Q(t)Q(t) in (20), (25):

𝐏t([z,+))Δ(t)z,z>Δ(t),\mathbf{P}_{t}([z,+\infty))\leqslant\frac{\Delta(t)}{z},\qquad z>\Delta(t), (46)

where Markov’s inequality [22] is applied to the probability distribution 𝐏t()\mathbf{P}_{t}(\cdot) of Q(t)Q(t) on [0,+)[0,+\infty) obtained by averaging its spectral measure [7] over the quantum state (17). In (45), use is made of the Frobenius norm \|\cdot\| and inner product ,\langle\cdot,\cdot\rangle of matrices [9] along with the one-point second-moment matrix of the process ξ\xi: Υ(t):=𝐄(ξ(t)ξ(t)T)=αtΠαtT+V(t)\Upsilon(t):=\mathbf{E}(\xi(t)\xi(t)^{\mathrm{T}})=\alpha_{t}\Pi\alpha_{t}^{\mathrm{T}}+V(t). The latter is computed in [26] by using (16), (19), (21), (22) and the second-moment matrix

Π:=𝐄(X(0)X(0)T)=P+iΘ,P:=ReΠ\Pi:=\mathbf{E}(X(0)X(0)^{\mathrm{T}})=P+i\Theta,\qquad P:=\mathrm{Re}\Pi (47)

of the initial system variables X1(0),,Xn(0)X_{1}(0),\ldots,X_{n}(0), including their CCR matrix Θ\Theta from (10), and the quantum covariance matrix

V(t):=𝐄(Z(t)Z(t)T)=0tesABΩBTesATds,t0V(t):=\mathbf{E}(Z(t)Z(t)^{\mathrm{T}})=\int_{0}^{t}\mathrm{e}^{sA}B\Omega B^{\mathrm{T}}\mathrm{e}^{sA^{\mathrm{T}}}\mathrm{d}s,\qquad t\geqslant 0 (48)

of the process ZZ in (15), with Ω\Omega the quantum Ito matrix from (5). The matrix V(t)V(t) in (48), which coincides with the controllability Gramian of the pair (A,BΩ)(A,B\sqrt{\Omega}) over the time interval [0,t][0,t], satisfies the initial value problem for the Lyapunov ODE

𝑉(t)=AV(t)+V(t)AT+,t0,V(0)=0,\mathop{V}^{\centerdot}(t)=AV(t)+V(t)A^{\mathrm{T}}+\mho,\qquad t\geqslant 0,\quad V(0)=0, (49)

where

:=BΩBT.\mho:=B\Omega B^{\mathrm{T}}. (50)

From (49), the time derivatives V(k):=dkV/dtkV^{(k)}:=\mathrm{d}^{k}V/\mathrm{d}t^{k} can be computed recursively as V(k+1)=AV(k)+V(k)AT+δk0V^{(k+1)}=AV^{(k)}+V^{(k)}A^{\mathrm{T}}+\delta_{k0}\mho for any k0k\geqslant 0, where δjk\delta_{jk} is the Kronecker delta. In particular, the first three derivatives at t=0t=0 are

V˙(0)\displaystyle\dot{V}(0) =,\displaystyle=\mho, (51)
V¨(0)\displaystyle\ddot{V}(0) =A+AT,\displaystyle=A\mho+\mho A^{\mathrm{T}}, (52)
V˙˙˙(0)\displaystyle\dddot{V}(0) =A2+(AT)2+2AAT.\displaystyle=A^{2}\mho+\mho(A^{\mathrm{T}})^{2}+2A\mho A^{\mathrm{T}}. (53)

In what follows, we will use the short-horizon asymptotic behaviour of (45) described below.

Lemma 2

Suppose the condition (26) of Lemma 1 is satisfied and the matrix FF in (23) is chosen so as to satisfy (29). Then the mean-square deviation functional (45) behaves asymptotically as

Δ(t)=GP2t2+O(t3),ast0+,\Delta(t)=\|G\sqrt{P}\|^{2}t^{2}+O(t^{3}),\qquad{\rm as}\ t\to 0+, (54)

where the coefficient is computed in terms of the matrices GG from (28) and PP from (47). \square

Proof:

From a combination of (45) with (49)–(52), it follows that

Δ(0)\displaystyle\Delta(0) =Σ,ReV(0)=0,\displaystyle=\langle\Sigma,\mathrm{Re}V(0)\rangle=0, (55)
Δ˙(0)\displaystyle\dot{\Delta}(0) =Σ,ReV˙(0)=FB2,\displaystyle=\langle\Sigma,\mathrm{Re}\dot{V}(0)\rangle=\|FB\|^{2}, (56)
Δ¨(0)\displaystyle\ddot{\Delta}(0) =2FAP2+Σ,ReV¨(0)\displaystyle=2\|FA\sqrt{P}\|^{2}+\langle\Sigma,\mathrm{Re}\ddot{V}(0)\rangle
=2(FAP2+FB,FAB).\displaystyle=2(\|FA\sqrt{P}\|^{2}+\langle FB,FAB\rangle). (57)

Now, let the matrix FF satisfy (29). The existence of such an FF is guaranteed by (26) according to Lemma 1 and its proof. Then FV˙(0)FT=FFT=0F\dot{V}(0)F^{\mathrm{T}}=F\mho F^{\mathrm{T}}=0, FV¨(0)FT=F(A+AT)FT=0F\ddot{V}(0)F^{\mathrm{T}}=F(A\mho+\mho A^{\mathrm{T}})F^{\mathrm{T}}=0 and FV˙˙˙(0)FT=F(A2+(AT)2+2AAT)FT=2GGTF\dddot{V}(0)F^{\mathrm{T}}=F(A^{2}\mho+\mho(A^{\mathrm{T}})^{2}+2A\mho A^{\mathrm{T}})F^{\mathrm{T}}=2G\mho G^{\mathrm{T}} in view of (50)–(53), so that the external fields manifest themselves in (24) starting from the third-order term of the Taylor series approximation

𝐜𝐨𝐯(FZ(t))=FV(t)FT=t33GGT+O(t4),ast0+.\mathbf{cov}(FZ(t))=FV(t)F^{\mathrm{T}}=\frac{t^{3}}{3}G\mho G^{\mathrm{T}}+O(t^{4}),\qquad{\rm as}\ t\to 0+. (58)

Accordingly, their contributions through the terms Σ,ReV(k)(0)\langle\Sigma,\mathrm{Re}V^{(k)}(0)\rangle, with k=1,2k=1,2, vanish, thus reducing (56), (57) to

Δ˙(0)=0,Δ¨(0)=2GP2,\dot{\Delta}(0)=0,\qquad\ddot{\Delta}(0)=2\|G\sqrt{P}\|^{2}, (59)

where use is also made of (31). Substitution of (55), (59) into an appropriately truncated Taylor series expansion of Δ\Delta leads to (54). ∎

Due to the special choice of the matrix FF in Lemmas 1 and 2, the leading term in (54) is quadratic (rather than linear) in time. This is a consequence of the partial isolation of the subsystem Φ\Phi from the external fields (see (58) and Fig. 2) and makes the mean-square deviation functional Δ\Delta grow slower (at least at short time horizons) compared to the case of arbitrary matrices FF considered in [26].

From the viewpoint of optimizing a network of OQHOs as a quantum memory system, the minimization (at a suitably chosen time t>0t>0) of the mean-square deviation functional Δ(t)\Delta(t) in (45) over admissible energy and coupling parameters of the network also minimizes the tail probability bounds (46) and provides a relevant performance criterion for such applications. This minimization improves the ability of the OQHO as a quantum memory to retain its initial system variables and is closely related (by duality) to maximizing the decoherence time

τ(ϵ):=inf{t0:Δ(t)>ϵFP2}\tau(\epsilon):=\inf\{t\geqslant 0:\ \Delta(t)>\epsilon\|F\sqrt{P}\|^{2}\} (60)

proposed in [26, 27] as a horizon by which the selected system variables do not deviate too far from their initial values. In accordance with (23), (47), the quantity 𝐄((FX(0))TFX(0))=Σ,Π=FP2\mathbf{E}((FX(0))^{\mathrm{T}}FX(0))=\langle\Sigma,\Pi\rangle=\|F\sqrt{P}\|^{2} in (60) provides a reference scale, with respect to which the dimensionless parameter ϵ>0\epsilon>0 specifies a relative error threshold for FX(t)FX(t) to approximately reproduce FX(0)FX(0). Since the columns of the matrix FF in (23) are linearly dependent if s<ns<n, we assume that

FP0F\sqrt{P}\neq 0 (61)

in order to avoid the trivial case τ(ϵ)=0\tau(\epsilon)=0. In the partial isolation setting of Lemmas 1 and 2, where FF satisfies (29) (as opposed to the case of FB0FB\neq 0 studied in [26, 27]), the asymptotic behaviour of (60) in the high-fidelity limit (of small values of ϵ\epsilon) is as follows.

Theorem 1

Suppose the conditions of Lemma 2 are satisfied together with (61) and

GP0.G\sqrt{P}\neq 0. (62)

Then the memory decoherence time (60) behaves asymptotically as

τ(ϵ)FPGPϵ=:τ^(ϵ),asϵ0+.\tau(\epsilon)\sim\frac{\|F\sqrt{P}\|}{\|G\sqrt{P}\|}\sqrt{\epsilon}=:\widehat{\tau}(\epsilon),\qquad{\rm as}\ \epsilon\to 0+. (63)

\square

Proof:

By (60), the memory decoherence time τ(ϵ)\tau(\epsilon) is a nondecreasing function of the fidelity parameter ϵ\epsilon satisfying

Δ(τ(ϵ))=ϵFP2,\Delta(\tau(\epsilon))=\epsilon\|F\sqrt{P}\|^{2}, (64)

so that τ(ϵ)>0\tau(\epsilon)>0 for any ϵ>0\epsilon>0 due to (61) and since Δ(t)\Delta(t) in (45) is a continuous function of t0t\geqslant 0, with Δ(0)=0\Delta(0)=0. Furthermore,

limϵ0+τ(ϵ)=0.\lim_{\epsilon\to 0+}\tau(\epsilon)=0. (65)

By combining (64), (65) with the asymptotic relation (54) under the condition (62), it follows that ϵFP2GP2τ(ϵ)2\epsilon\|F\sqrt{P}\|^{2}\sim\|G\sqrt{P}\|^{2}\tau(\epsilon)^{2}, as ϵ0+\epsilon\to 0+, which leads to (63). ∎

As a consequence of the partial subsystem isolation, the asymptotic behaviour (63) is qualitatively different and yields a longer decoherence time compared to the case of FB0FB\neq 0 in [26], where τ(ϵ)\tau(\epsilon) is asymptotically linear with respect to ϵ\epsilon. In the framework of (63), the maximization of τ(ϵ)\tau(\epsilon) at a given small value of ϵ\epsilon can be replaced with its “approximate” version

τ^(ϵ)sup.\widehat{\tau}(\epsilon)\longrightarrow\sup. (66)

With FF and PP being fixed, (66) is equivalent to the minimization of the denominator

GP=2FΘRP\|G\sqrt{P}\|=2\|F\Theta R\sqrt{P}\| (67)

and is organised as a convex quadratic optimization problem over the energy matrix RR.

VI MEMORY DECOHERENCE TIME MAXIMIZATION FOR OQHO INTERCONNECTION

Consider the approximate memory decoherence time maximization (66) for a coherent feedback interconnection of two OQHOs from [25, Section 8] and [26], which interact with external input bosonic fields and are coupled to each other through a direct energy coupling and an indirect field-mediated coupling [32]; see Fig. 3.

OQHO1OQHO2W(1)W^{(1)}Y(2)Y^{(2)}W(2)W^{(2)}Y(1)Y^{(1)}
Figure 3: A coherent feedback interconnection of two OQHOs, interacting with external input quantum Wiener processes W(1)W^{(1)}, W(2)W^{(2)} and coupled to each other through a direct energy coupling (represented by a double-headed arrow) and a field-mediated coupling through the quantum Ito processes Y(1)Y^{(1)}, Y(2)Y^{(2)} at the corresponding outputs.

In this setting, the quantum Wiener processes W(1)W^{(1)}, W(2)W^{(2)} of the external fields of even dimensions m1m_{1}, m2m_{2} are in the vacuum states υ1\upsilon_{1}, υ2\upsilon_{2} on symmetric Fock spaces 𝔉1\mathfrak{F}_{1}, 𝔉2\mathfrak{F}_{2}, respectively, so that W(1)W^{(1)}, W(2)W^{(2)} commute with and are statistically independent of each other. The augmented quantum Wiener process W:=[W(1)W(2)]W:={\small\begin{bmatrix}W^{(1)}\\ W^{(2)}\end{bmatrix}} of dimension m:=m1+m2m:=m_{1}+m_{2} on the composite Fock space 𝔉:=𝔉1𝔉2\mathfrak{F}:=\mathfrak{F}_{1}\otimes\mathfrak{F}_{2} has the quantum Ito matrix Ω=[Ω100Ω2]\Omega={\small\begin{bmatrix}\Omega_{1}&0\\ 0&\Omega_{2}\end{bmatrix}} in (5) coming from the individual Ito tables dW(k)dW(k)=TΩkdt\mathrm{d}W^{(k)}\mathrm{d}W^{(k)}{}^{\mathrm{T}}=\Omega_{k}\mathrm{d}t, where J=[J100J2]J={\small\begin{bmatrix}J_{1}&0\\ 0&J_{2}\end{bmatrix}}, Ωk:=Imk+iJk\Omega_{k}:=I_{m_{k}}+iJ_{k} and Jk:=Imk/2𝐉J_{k}:=I_{m_{k}/2}\otimes\mathbf{J}, with the matrix 𝐉\mathbf{J} from (6). The component OQHOs are endowed with initial spaces k\mathfrak{H}_{k} and vectors X(k)X^{(k)} of even numbers nk:=2νkn_{k}:=2\nu_{k} of dynamic variables on the composite system-field space (2), where 0:=12\mathfrak{H}_{0}:=\mathfrak{H}_{1}\otimes\mathfrak{H}_{2}. Accordingly, the vector X:=[X(1)X(2)]X:={\small\begin{bmatrix}X^{(1)}\\ X^{(2)}\end{bmatrix}} of n:=n1+n2n:=n_{1}+n_{2} system variables with ν=ν1+ν2\nu=\nu_{1}+\nu_{2} in (8) for the augmented OQHO satisfies (10) with the CCR matrix

Θ:=[Θ100Θ2]\Theta:={\begin{bmatrix}\Theta_{1}&0\\ 0&\Theta_{2}\end{bmatrix}} (68)

which consists of the individual CCR matrices Θk=12Iνk𝐉\Theta_{k}=\frac{1}{2}I_{\nu_{k}}\otimes\mathbf{J}:

[X(1),X(2)]T=0,[X(k),X(k)]T=2iΘk,k=1,2.[X^{(1)},X^{(2)}{}^{\mathrm{T}}]=0,\qquad[X^{(k)},X^{(k)}{}^{\mathrm{T}}]=2i\Theta_{k},\qquad k=1,2. (69)

In addition to the individual Hamiltonians Hk:=12X(k)RkTX(k)H_{k}:=\frac{1}{2}X^{(k)}{}^{\mathrm{T}}R_{k}X^{(k)} of the OQHOs, parameterized by their energy matrices Rk=RkTnk×nkR_{k}=R_{k}^{\mathrm{T}}\in{\mathbb{R}}^{n_{k}\times n_{k}}, the direct energy coupling (see Fig. 3) contributes the term

H12:=X(1)R12TX(2)=X(2)R21TX(1)H_{12}:=X^{(1)}{}^{\mathrm{T}}R_{12}X^{(2)}=X^{(2)}{}^{\mathrm{T}}R_{21}X^{(1)} (70)

specified by

R12=R21Tn1×n2R_{12}=R_{21}^{\mathrm{T}}\in{\mathbb{R}}^{n_{1}\times n_{2}} (71)

in view of the commutativity in (69). Since there is also an additional indirect coupling of the OQHOs, which is mediated by their output fields Y(1)Y^{(1)}, Y(2)Y^{(2)} of even dimensions r1r_{1}, r2r_{2}, the resulting interconnection is governed by the QSDEs [26]

dX(k)\displaystyle\mathrm{d}X^{(k)} =(AkX(k)+FkX(3k))dt+BkdW(k)+EkdY(3k),\displaystyle=(A_{k}X^{(k)}+F_{k}X^{(3-k)})\mathrm{d}t+B_{k}\mathrm{d}W^{(k)}+E_{k}\mathrm{d}Y^{(3-k)}, (72)
dY(k)\displaystyle\mathrm{d}Y^{(k)} =CkX(k)dt+DkdW(k),k=1,2.\displaystyle=C_{k}X^{(k)}\mathrm{d}t+D_{k}\mathrm{d}W^{(k)},\qquad k=1,2. (73)

The matrices Aknk×nkA_{k}\in{\mathbb{R}}^{n_{k}\times n_{k}}, Bknk×mkB_{k}\in{\mathbb{R}}^{n_{k}\times m_{k}}, Ckrk×nkC_{k}\in{\mathbb{R}}^{r_{k}\times n_{k}}, Eknk×r3kE_{k}\in{\mathbb{R}}^{n_{k}\times r_{3-k}}, Fknk×n3kF_{k}\in{\mathbb{R}}^{n_{k}\times n_{3-k}} are parameterized as

Ak\displaystyle A_{k} =2Θk(Rk+MkTJkMk+NkTJ~3kNk),Bk=2ΘkMkT,\displaystyle=2\Theta_{k}(R_{k}+M_{k}^{\mathrm{T}}J_{k}M_{k}+N_{k}^{\mathrm{T}}\widetilde{J}_{3-k}N_{k}),\quad B_{k}=2\Theta_{k}M_{k}^{\mathrm{T}}, (74)
Ck\displaystyle C_{k} =2DkJkMk,Ek=2ΘkNkT,Fk=2ΘkRk,3k,\displaystyle=2D_{k}J_{k}M_{k},\qquad E_{k}=2\Theta_{k}N_{k}^{\mathrm{T}},\qquad F_{k}=2\Theta_{k}R_{k,3-k}, (75)

and the matrices Dkrk×mkD_{k}\in{\mathbb{R}}^{r_{k}\times m_{k}} consist of rkmkr_{k}\leqslant m_{k} conjugate rows of permutation matrices of orders mkm_{k} with DkΩkDkT=Irk+iJ~kD_{k}\Omega_{k}D_{k}^{\mathrm{T}}=I_{r_{k}}+i\widetilde{J}_{k}, where J~k:=DkJkDkT=Irk/2𝐉\widetilde{J}_{k}:=D_{k}J_{k}D_{k}^{\mathrm{T}}=I_{r_{k}/2}\otimes\mathbf{J}. Here, Mkmk×nkM_{k}\in{\mathbb{R}}^{m_{k}\times n_{k}}, Nkr3k×nkN_{k}\in{\mathbb{R}}^{r_{3-k}\times n_{k}} are the matrices of coupling of the kkth OQHO to its external input field W(k)W^{(k)} and the output Y(3k)Y^{(3-k)} of the other OQHO, respectively. By (72), (73), the composite OQHO in Fig. 3 is governed by the first QSDE in (12) with the matrices

A=[A1F1+E1C2F2+E2C1A2],B=[B1E1D2E2D1B2]A={\begin{bmatrix}A_{1}&F_{1}+E_{1}C_{2}\\ F_{2}+E_{2}C_{1}&A_{2}\end{bmatrix}},\qquad B={\begin{bmatrix}B_{1}&E_{1}D_{2}\\ E_{2}D_{1}&B_{2}\end{bmatrix}} (76)

and, by (74), (75), has the following energy and coupling matrices:

R=R+[0R12R12T0],M=[M1D1TN2D2TN1M2].R=R_{*}+{\begin{bmatrix}0&R_{12}\\ R_{12}^{\mathrm{T}}&0\end{bmatrix}},\qquad M={\begin{bmatrix}M_{1}&D_{1}^{\mathrm{T}}N_{2}\\ D_{2}^{\mathrm{T}}N_{1}&M_{2}\end{bmatrix}}. (77)

Here, the matrix

R:=[R1N1TD2J2M2M1TJ1D1TN2N2TD1J1M1M2TJ2D2TN1R2]R_{*}:={\begin{bmatrix}R_{1}&N_{1}^{\mathrm{T}}D_{2}J_{2}M_{2}-M_{1}^{\mathrm{T}}J_{1}D_{1}^{\mathrm{T}}N_{2}\\ N_{2}^{\mathrm{T}}D_{1}J_{1}M_{1}-M_{2}^{\mathrm{T}}J_{2}D_{2}^{\mathrm{T}}N_{1}&R_{2}\end{bmatrix}} (78)

is associated with the field-mediated coupling between the constituent OQHOs, their coupling to the external fields, and the individual Hamiltonians. Therefore, if the matrices MkM_{k}, NkN_{k}, RkR_{k} are fixed for all k=1,2k=1,2, and hence, so also is RR_{*} in (78), then the energy matrix RR in (77) can only be varied over a proper affine subspace in the subspace of real symmetric matrices of order nn by varying the direct energy coupling matrix R12R_{12}. This leads to the following formulation of the approximate memory decoherence time maximization problem (66):

τ^(ϵ)supoverR12n1×n2.\widehat{\tau}(\epsilon)\longrightarrow\sup\qquad{\rm over}\ R_{12}\in{\mathbb{R}}^{n_{1}\times n_{2}}. (79)

This setting is similar to that in [26], except that we are concerned here with a subsystem of the closed-loop OQHO which is partially isolated from the external fields W(1)W^{(1)}, W(2)W^{(2)}.

Theorem 2

Suppose the OQHO interconnection, described by (68)–(78), and the matrix FF in (23) satisfy the conditions of Theorem 1. Then the direct energy coupling matrix R12R_{12} in (70), (71) solves the problem (79) with the approximate memory decoherence time τ^\widehat{\tau} in (63) for the composite OQHO if and only if

g(R12)+K=0,g(R_{12})+K=0, (80)

where gg is a linear operator acting on a matrix Nn1×n2N\in{\mathbb{R}}^{n_{1}\times n_{2}} as

g(N):=\displaystyle g(N):= Θ1Σ11Θ1NP22+P11NΘ2Σ22Θ2\displaystyle\Theta_{1}\Sigma_{11}\Theta_{1}NP_{22}+P_{11}N\Theta_{2}\Sigma_{22}\Theta_{2}
+Θ1Σ12Θ2NTP12+P12NTΘ1Σ12Θ2,\displaystyle+\Theta_{1}\Sigma_{12}\Theta_{2}N^{\mathrm{T}}P_{12}+P_{12}N^{\mathrm{T}}\Theta_{1}\Sigma_{12}\Theta_{2}, (81)

and Kn1×n2K\in{\mathbb{R}}^{n_{1}\times n_{2}} is an auxiliary matrix which is computed as

K:=2(𝐒(ΘΣΘRP))12K:=2(\mathbf{S}(\Theta\Sigma\Theta R_{*}P))_{12} (82)

in terms of (23), (47), (78), with ()jk(\cdot)_{jk} the appropriate matrix blocks, and 𝐒(γ):=12(γ+γT)\mathbf{S}(\gamma):=\frac{1}{2}(\gamma+\gamma^{\mathrm{T}}) the matrix symmetrizer. \square

Proof:

As mentioned above, (79) reduces to the minimization of the quantity (67), which is equivalent to minimizing the convex quadratic function

f(R12):=12FΘRP2f(R_{12}):=\frac{1}{2}\|F\Theta R\sqrt{P}\|^{2} (83)

over R12n1×n2R_{12}\in{\mathbb{R}}^{n_{1}\times n_{2}}, where the 12\frac{1}{2}-factor is introduced for convenience, and RR is an affine function of R12R_{12} in (77). Therefore, R12R_{12} delivers a global minimum to (83) if and only if it makes the Frechet derivative [20] of ff (on the Hilbert space n1×n2{\mathbb{R}}^{n_{1}\times n_{2}} with the Frobenius inner product ,\langle\cdot,\cdot\rangle) vanish:

f(R12)=0.f^{\prime}(R_{12})=0. (84)

In view of the variational identity δ(ϕ2)=2ϕ,δϕ\delta(\|\phi\|^{2})=2\langle\phi,\delta\phi\rangle, the first variation of (83) with respect to the matrix R12R_{12} takes the form

δf(R12)\displaystyle\delta f(R_{12}) =FΘRP,FΘ(δR)P=ΘFTFΘRP,δR\displaystyle=\langle F\Theta R\sqrt{P},F\Theta(\delta R)\sqrt{P}\rangle=-\langle\Theta F^{\mathrm{T}}F\Theta RP,\delta R\rangle
=(ΘΣΘRP)12,δR12(ΘΣΘRP)21,δR12T\displaystyle=-\langle(\Theta\Sigma\Theta RP)_{12},\delta R_{12}\rangle-\langle(\Theta\Sigma\Theta RP)_{21},\delta R_{12}^{\mathrm{T}}\rangle
=(ΘΣΘRP)12+(ΘΣΘRP)21T,δR12,\displaystyle=-\langle(\Theta\Sigma\Theta RP)_{12}+(\Theta\Sigma\Theta RP)_{21}^{\mathrm{T}},\delta R_{12}\rangle, (85)

where use is also made of the antisymmetry of Θ\Theta in (68) and symmetry of PP in (47) along with (23) and δR=[0δR12δR12T0]\delta R={\small\begin{bmatrix}0&\delta R_{12}\\ \delta R_{12}^{\mathrm{T}}&0\end{bmatrix}} since the matrix RR_{*} in (77) is fixed. From (85), it follows that

f(R12)=2(𝐒(ΘΣΘRP))12=g(R12)+K,-f^{\prime}(R_{12})=2(\mathbf{S}(\Theta\Sigma\Theta RP))_{12}=g(R_{12})+K, (86)

with gg, KK from (81), (82). By (86), (84) is equivalent to (80), thus establishing the latter as a necessary and sufficient condition of optimality for (79). ∎

In view of (86), the linear map gg in (81) is a negative semi-definite self-adjoint operator on n1×n2{\mathbb{R}}^{n_{1}\times n_{2}} related to the second Frechet derivative f′′f^{\prime\prime} of the function ff in (83) by g=f′′g=-f^{\prime\prime}. Also note that the linear equation (80) can be solved by vectorizing [15] the matrix R12R_{12}.

References

  • [1] V.I.Arnold, Mathematical Methods of Classical Mechanics, 2nd Ed., Springer, New York, 1989.
  • [2] C.D’Helon, and M.R.James, Stability, gain, and robustness in quantum feedback networks, Phys. Rev. A, vol. 73, no. 5, 2006, 053803.
  • [3] L.A.D.Espinosa, Z.Miao, I.R.Petersen, V.Ugrinovskii, and M.R.James, Physical realizability and preservation of commutation and anticommutation relations for nn-level quantum systems, SIAM J. Control Optim., vol. 54, no. 2, 2016, pp. 632–661.
  • [4] S.Fu, A.R.R.Carvalho, M.R.Hush, and M.R.James, Cross-phase modulation and entanglement in a compound gradient echo memory, Phys. Rev. A, vol. 93, 2016, 023809.
  • [5] V.Gorini, A.Kossakowski, E.C.G.Sudarshan, Completely positive dynamical semigroups of NN-level systems, J. Math. Phys., vol. 17, no. 5, 1976, pp. 821–825.
  • [6] K.Heshami, D.G.England, P.C.Humphreys, P.J.Bustard, V.M.Acosta, J.Nunn, B.J.Sussman, Quantum memories: emerging applications and recent advances, J. Mod. Opt., vol. 63, no. 20, 2016, pp. 2005–2028.
  • [7] A.S.Holevo, Statistical Structure of Quantum Theory, Springer, Berlin, 2001.
  • [8] A.S.Holevo, Quantum Systems, Channels, Information: A Mathematical Introduction, De Gruyter, Berlin, 2019.
  • [9] R.A.Horn, and C.R.Johnson, Matrix Analysis, Cambridge University Press, New York, 2007.
  • [10] R.L.Hudson, and K.R.Parthasarathy, Quantum Ito’s formula and stochastic evolutions, Commun. Math. Phys., vol. 93, 1984, pp. 301–323.
  • [11] M.R.James, H.I.Nurdin, and I.R.Petersen, HH^{\infty} control of linear quantum stochastic systems, IEEE Trans. Automat. Contr., vol. 53, no. 8, 2008, pp. 1787–1803.
  • [12] I.Karatzas, and S.E.Shreve, Brownian Motion and Stochastic Calculus, 2nd Ed., Springer-Verlag, New York, 1991.
  • [13] L.D.Landau, and E.M.Lifshitz, Quantum Mechanics: Non-relativistic Theory, 3rd Ed., Butterworth-Heinemann, Amsterdam, 1981.
  • [14] G.Lindblad, On the generators of quantum dynamical semigroups, Comm. Math. Phys., vol. 48, 1976, pp. 119–130.
  • [15] J.R.Magnus, Linear Structures, Oxford University Press, New York, 1988.
  • [16] M.A.Nielsen, and I.L.Chuang, Quantum Computation and Quantum Information, Cambridge University Press, Cambridge, 2000.
  • [17] H.I.Nurdin, and N.Yamamoto, Linear Dynamical Quantum Systems, Springer, Netherlands, 2017.
  • [18] K.R.Parthasarathy, An Introduction to Quantum Stochastic Calculus, Birkhäuser, Basel, 1992.
  • [19] I.R.Petersen, Quantum linear systems theory, Open Automat. Contr. Syst. J., vol. 8, 2017, pp. 67–93.
  • [20] M.Reed, and B.Simon, Methods of Modern Mathematical Physics: Functional Analysis, vol. 1, 2nd Ed., Academic Press, San Diego, 1980.
  • [21] J.J.Sakurai, Modern Quantum Mechanics, Addison-Wesley, Reading, Mass., 1994.
  • [22] A.N.Shiryaev, Probability, 2nd Ed., Springer, New York, 1996.
  • [23] I.G.Vladimirov, I.R.Petersen, and M.R.James, Multi-point Gaussian states, quadratic-exponential cost functionals, and large deviations estimates for linear quantum stochastic systems, Appl. Math. Optim., vol. 83, 2021, pp. 83–137 (published 24 July 2018).
  • [24] I.G.Vladimirov, and I.R.Petersen, Moment dynamics and observer design for a class of quasilinear quantum stochastic systems, SIAM J. Control Optim., vol. 60, no. 3, 2022, pp. 1223–1249.
  • [25] I.G.Vladimirov, and I.R.Petersen, Decoherence quantification through commutation relations decay for open quantum harmonic oscillators, Systems & Control Letters, vol. 178, 2023, 105585.
  • [26] I.G.Vladimirov, and I.R.Petersen, Decoherence time in quantum harmonic oscillators as quantum memory systems, 2024 Australian and New Zealand Control Conference (ANZCC), Gold Coast, Australia, 2024, pp. 211–216.
  • [27] I.G.Vladimirov, and I.R.Petersen, Decoherence time control by interconnection for finite-level quantum memory systems, European Journal of Control, 2024, 101054, published online 15 June 2024, https://doi.org/10.1016/j.ejcon.2024.101054.
  • [28] D.F.Walls, and G.J.Milburn, Quantum Optics, Springer, Berlin, 2008.
  • [29] Y.Wang, J.Li, S.Zhang, et al, Efficient quantum memory for single-photon polarization qubits, Nat. Photonics, vol. 13, 2019, pp. 346–351.
  • [30] N.Yamamoto, and M.R.James, Zero-dynamics principle for perfect quantum memory in linear networks, New J. Phys., vol. 16, 2014, 073032.
  • [31] G.Zhang, and Z.Dong, Linear quantum systems: a tutorial, Annual Reviews in Control, vol. 54, 2022, pp. 274–294.
  • [32] 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.