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

2\mathcal{H}_{2}-optimal Model Reduction of Linear Quadratic Output Systems in Finite Frequency Range

Umair Zulfiqar [email protected] Zhi-Hua Xiao Qiu-Yan Song Mohammad Monir Uddin Victor Sreeram School of Electronic Information and Electrical Engineering, Yangtze University, Jingzhou, Hubei, 434023, China School of Information and Mathematics, Yangtze University, Jingzhou, Hubei, 434023, China School of Mechatronic Engineering and Automation, Shanghai University, Shanghai, 200444, China Department of Mathematics and Physics, North South University, Dhaka, 1229, Bangladesh Department of Electrical, Electronic, and Computer Engineering, The University of Western Australia, Perth, 6009, Australia
Abstract

Linear quadratic output systems constitute an important class of dynamical systems with numerous practical applications. When the order of these models is exceptionally high, simulating and analyzing these systems becomes computationally prohibitive. In such instances, model order reduction offers an effective solution by approximating the original high-order system with a reduced-order model while preserving the system’s essential characteristics.

In frequency-limited model order reduction, the objective is to maintain the frequency response of the original system within a specified frequency range in the reduced-order model. In this paper, a mathematical expression for the frequency-limited 2\mathcal{H}_{2} norm is derived, which quantifies the error within the desired frequency interval. Subsequently, the necessary conditions for a local optimum of the frequency-limited 2\mathcal{H}_{2} norm of the error are derived. The inherent difficulty in satisfying these conditions within a Petrov-Galerkin projection framework is also discussed. Based on the optimality conditions and Petrov-Galerkin projection, a stationary point iteration algorithm is proposed that enforces three of the four optimality conditions upon convergence. A numerical example is provided to illustrate the algorithm’s effectiveness in accurately approximating the original high-order model within the specified frequency interval.

keywords:
2\mathcal{H}_{2}-optimal, frequency-limited, model order reduction, projection, reduced-order model, quadratic output
journal: ArXiv.org

1 Introduction

This study focuses on a specific class of nonlinear dynamical systems with weak nonlinearity. These systems have linear time-invariant (LTI) state equations but feature quadratic nonlinear terms in the output equation, referred to as linear quadratic output (LQO) systems [1]. They naturally arise in scenarios where it is necessary to observe quantities involving products of state components in either the time or frequency domain. For example, they are used in applications that quantify energy or power, such as the internal energy functional of a system [2] or the objective cost function in optimal quadratic control problems [3]. Additionally, they are used to measure the deviation of a state’s coordinates from a reference point, such as the root mean squared displacement of spatial coordinates around an excitation point, or in stochastic modeling for calculating the variance of a random variable [4].

Consider a LQO system described by the following state and output equations:

G:={x˙(t)=Ax(t)+Bu(t),y(t)=Cx(t)+[x(t)TM1x(t)x(t)TMpx(t)],\displaystyle G:=\begin{cases}\dot{x}(t)=Ax(t)+Bu(t),\\ y(t)=Cx(t)+\begin{bmatrix}x(t)^{T}M_{1}x(t)\\ \vdots\\ x(t)^{T}M_{p}x(t)\end{bmatrix},\end{cases} (1)

wherein x(t)n×1x(t)\in\mathbb{R}^{n\times 1} represents the state vector, u(t)m×1u(t)\in\mathbb{R}^{m\times 1} represents inputs, and y(t)p×1y(t)\in\mathbb{R}^{p\times 1} represents outputs. (A,B,C,M1,,Mp)(A,B,C,M_{1},\cdots,M_{p}) is the state-space realization of GG with An×nA\in\mathbb{R}^{n\times n}, Bn×mB\in\mathbb{R}^{n\times m}, Cp×nC\in\mathbb{R}^{p\times n}, and Min×nM_{i}\in\mathbb{R}^{n\times n}. The state equation in (1) is identical to that of standard LTI systems. However, the output equation in (1) introduces a nonlinearity in the form of the quadratic function of states x(t)TMix(t)x(t)^{T}M_{i}x(t).

Let us denote y1(t)=Cx(t)y_{1}(t)=Cx(t) and y2,i=x(t)TMix(t)y_{2,i}=x(t)^{T}M_{i}x(t). The input-output mapping between u(t)u(t) and y1(t)y_{1}(t) is represented by the following transfer function:

G1(s)\displaystyle G_{1}(s) =C(sIA)1B.\displaystyle=C(sI-A)^{-1}B.

Additionally, the input-output mapping between u(t)u(t) and y2,i(t)y_{2,i}(t) is represented by the following multivariate transfer function:

G2,i(s1,s2)\displaystyle G_{2,i}(s_{1},s_{2}) =BT(s1IA)Mi(s2IA)1B,\displaystyle=B^{T}(s_{1}I-A)^{-*}M_{i}(s_{2}I-A)^{-1}B,

cf. [5].

To ensure high fidelity in the mathematical modeling of complex physical phenomena, it is often necessary to use dynamical systems with very high orders, sometimes exceeding several thousand. This high order nn makes it computationally difficult or even prohibitive to simulate and analyze the model (1). Therefore, it is important to approximate (1) with a reduced-order model (ROM) of much lower order kk (where knk\ll n). This approach simplifies the simulation and analysis processes. The process of creating such a ROM while preserving the key features of the original model is known as model order reduction (MOR); refer to [6, 7, 8, 9, 10] for a deeper understanding of the topic.

Let us denote the kthk^{th}-order ROM of GG as GkG_{k}, characterized by the following state and output equations:

Gk:={xk˙(t)=Akxk(t)+Bku(t),yk(t)=Ckxk(t)+[xk(t)TMk,1xk(t)xk(t)TMk,pxk(t)],\displaystyle G_{k}:=\begin{cases}\dot{x_{k}}(t)=A_{k}x_{k}(t)+B_{k}u(t),\\ y_{k}(t)=C_{k}x_{k}(t)+\begin{bmatrix}x_{k}(t)^{T}M_{k,1}x_{k}(t)\\ \vdots\\ x_{k}(t)^{T}M_{k,p}x_{k}(t)\end{bmatrix},\end{cases} (2)

wherein Ak=WkTAVkk×kA_{k}=W_{k}^{T}AV_{k}\in\mathbb{R}^{k\times k}, Bk=WkTBk×mB_{k}=W_{k}^{T}B\in\mathbb{R}^{k\times m}, Ck=CVkp×kC_{k}=CV_{k}\in\mathbb{R}^{p\times k}, and Mk,i=VkTMiVkk×kM_{k,i}=V_{k}^{T}M_{i}V_{k}\in\mathbb{R}^{k\times k}, satisfying the Petrov-Galerkin projection condition WkTVk=IW_{k}^{T}V_{k}=I. The projection matrices Vkn×kV_{k}\in\mathbb{R}^{n\times k} and Wkn×kW_{k}\in\mathbb{R}^{n\times k} project GG onto a reduced subspace to obtain the ROM GkG_{k}. Various MOR methods differ in how they construct VkV_{k} and WkW_{k}. The choice of VkV_{k} and WkW_{k} depends on the specific characteristics of GG that need to be preserved in GkG_{k}.

Let yk,1(t)=Ckxk(t)y_{k,1}(t)=C_{k}x_{k}(t) and yk,2,i=xk(t)TMk,ixk(t)y_{k,2,i}=x_{k}(t)^{T}M_{k,i}x_{k}(t). The input-output relationships between u(t)u(t) and yk,1(t)y_{k,1}(t), as well as u(t)u(t) and yk,2,i(t)y_{k,2,i}(t), are described by the following transfer functions:

Gk,1(s)\displaystyle G_{k,1}(s) =Ck(sIAk)1Bk,\displaystyle=C_{k}(sI-A_{k})^{-1}B_{k},
Gk,2,i(s1,s2)\displaystyle G_{k,2,i}(s_{1},s_{2}) =BkT(s1IAk)Mk,i(s2IAk)1Bk.\displaystyle=B_{k}^{T}(s_{1}I-A_{k})^{-*}M_{k,i}(s_{2}I-A_{k})^{-1}B_{k}.

Throughout this paper, it is assumed that both AA and AkA_{k} are Hurwitz.

The Balanced Truncation (BT) method, introduced in 1981, is a prominent technique for MOR [11]. This approach retains states that significantly effect energy transfer between inputs and outputs, discarding those with minimal impact as indicated by their Hankel singular values. A notable advantage of BT is its ability to estimate errors in advance of creating the ROM [12]. Moreover, BT preserves the stability of the original system. Originally proposed for LTI systems, BT’s application has broadened to include descriptor systems [13], second-order systems [14], linear time-varying systems [15], parametric systems [16], nonlinear systems [17], and bilinear systems [18]. Additionally, BT has been tailored to maintain specific system properties such as positive realness [19], bounded realness [20], passivity [21], and special structural characteristics [22]. For a detailed overview of the various BT algorithms, refer to the survey [23]. BT has been adapted for LQO systems in [24, 25, 26]. Of these algorithms, only the one proposed in [26] preserves the LQO structure in the ROM.

The 2\mathcal{H}_{2}-optimal MOR problem for standard LTI systems has been thoroughly explored in the literature. Wilson’s conditions, which are necessary conditions for achieving a local optimum of the 2\mathcal{H}_{2} norm of error, are outlined in [27]. The iterative rational Krylov algorithm (IRKA) was the first to apply interpolation theory to meet these conditions [28]. Other algorithms using Sylvester equations and projection have been developed and enhanced for improved robustness in [29] and [30]. Recently, the 2\mathcal{H}_{2}-optimal MOR problem for LQO systems has been addressed in [31], where a Sylvester equation-based algorithm has been proposed to achieve a local optimum upon convergence.

Many MOR problems are inherently frequency-limited, with certain frequency ranges being more important. For example, when creating a ROM for a notch filter, it is crucial to minimize the approximation error near the notch frequency [32]. Similarly, for closed-loop stability, the ROM of a plant must accurately capture the system’s behavior in the crossover frequency region [33, 34]. In interconnected power systems, low-frequency oscillations are essential for small-signal stability studies, so the ROM should accurately represent the behavior within the frequency range of inter-area and inter-plant oscillations [35, 36]. This need has led to frequency-limited MOR, which focuses on achieving high accuracy within specific frequency intervals rather than the entire spectrum [37].

The frequency-limited MOR problem aims to create a ROM that ensures that

G1(jν)Gk,1(jν)andG2,i(jν1,jν2)Gk,2,i(jν1,jν2)\displaystyle||G_{1}(j\nu)-G_{k,1}(j\nu)||\hskip 14.22636pt\textnormal{and}\hskip 14.22636pt||G_{2,i}(j\nu_{1},j\nu_{2})-G_{k,2,i}(j\nu_{1},j\nu_{2})||

are small when ν\nu, ν1\nu_{1}, and ν2\nu_{2} are within the specified frequency range of [0,ω][0,\omega] rad/sec.

BT typically offers an accurate approximation of the original model across the entire frequency spectrum. In [37], BT was adapted to address the frequency-limited MOR problem, resulting in the frequency-limited BT (FLBT) algorithm. However, FLBT does not preserve the stability and a priori error bounds that BT does. The computational aspects of FLBT and efficient methods for handling large-scale systems are discussed in [38]. Additionally, FLBT has been expanded to cover a wider range of systems, including descriptor systems [39], second-order systems [40], and bilinear systems [41]. FLBT has been recently generalized for LQO systems in [42].

The frequency-limited 2\mathcal{H}_{2} norm for LTI systems is defined in [43], with Gramian-based conditions outlined for achieving a local optimum. Generalizations of the iterative rational Krylov algorithm (IRKA) have resulted in the development of the frequency-limited IRKA (FLIRKA) in [44, 45, 46], which approximately fulfills these conditions. This paper examines the frequency-limited 2\mathcal{H}_{2}-optimal MOR problem for LQO systems.

This research work makes several important contributions. First, it introduces the frequency-limited 2\mathcal{H}_{2} norm (2,ω\mathcal{H}_{2,\omega} norm) for LQO systems and shows how to compute it using frequency-limited system Gramians defined in [42]. Second, it derives the necessary conditions for achieving a local optimum of GGk2,ω2||G-G_{k}||_{\mathcal{H}_{2,\omega}}^{2}. Third, it compares these conditions with those for standard 2\mathcal{H}_{2}-optimal model order reduction [31], highlighting that Petrov-Galerkin projection generally cannot achieve a local optimum in the frequency-limited scenario. Fourth, it proposes a stationary point algorithm based on Petrov-Galerkin projection, which meets three of the four necessary conditions for optimality upon convergence. The paper includes a numerical example demonstrating the algorithm’s accuracy within the specified frequency range and showing it outperforms existing methods.

2 Literature Review

In this section, we will briefly explore two key MOR algorithms relevant to LQO systems within the context of the problem at hand. The first is the FLBT [42], and the second is the 2\mathcal{H}_{2}-optimal MOR method [31].

2.1 Frequency-limited Balanced Truncation (FLBT) [42]

FLBT creates the ROM by truncating states that contribute minimally to the input-output energy transfer within the desired frequency range of [0,ω][0,\omega] rad/sec. This is achieved by constructing a frequency-limited balanced realization using frequency-limited Gramians and then truncating the states corresponding to the smallest frequency-limited Hankel singular values.

The frequency-limited controllability Gramian PωP_{\omega} within the desired frequency interval [0,ω][0,\omega] rad/sec is given by

Pω=12πωω(jνIA)1BBT(jνIA)𝑑ν.\displaystyle P_{\omega}=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-1}BB^{T}(j\nu I-A)^{-*}d\nu.

Next, we define FωF_{\omega} as

Fω=12πωω(jνIA)1𝑑ν=jπln(jωIA).\displaystyle F_{\omega}=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-1}d\nu=\frac{j}{\pi}ln(-j\omega I-A).

With this, PωP_{\omega} can be computed by solving the following Lyapunov equation:

APω+PωAT+FωBBT+BBTFω\displaystyle AP_{\omega}+P_{\omega}A^{T}+F_{\omega}BB^{T}+BB^{T}F_{\omega}^{*} =0.\displaystyle=0. (3)

The frequency-limited observability Gramian Qω=Yω+ZωQ_{\omega}=Y_{\omega}+Z_{\omega} within the frequency range [0,ω][0,\omega] rad/sec is defined as

Yω\displaystyle Y_{\omega} =12πωω(jνIA)CTC(jνIA)1𝑑ν,\displaystyle=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-*}C^{T}C(j\nu I-A)^{-1}d\nu,
Zω\displaystyle Z_{\omega} =12πωω(jν1IA)(i=1pMi(12πωω(jν2IA)1BBT\displaystyle=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu_{1}I-A)^{-*}\Bigg{(}\sum_{i=1}^{p}M_{i}\Big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu_{2}I-A)^{-1}BB^{T}
(jν2IA)dν2)Mi)(jν1IA)1dν1\displaystyle\hskip 108.12054pt(j\nu_{2}I-A)^{-*}d\nu_{2}\Big{)}M_{i}\Bigg{)}(j\nu_{1}I-A)^{-1}d\nu_{1}
=12πωω(jν1IA)(i=1pMiPωMi)(jν1IA)1𝑑ν1.\displaystyle=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu_{1}I-A)^{-*}\big{(}\sum_{i=1}^{p}M_{i}P_{\omega}M_{i}\big{)}(j\nu_{1}I-A)^{-1}d\nu_{1}.

YωY_{\omega}, ZωZ_{\omega}, and QωQ_{\omega} can be determined by solving the following Lyapunov equations:

ATYω+YωA+FωCTC+CTCFω\displaystyle A^{T}Y_{\omega}+Y_{\omega}A+F_{\omega}^{*}C^{T}C+C^{T}CF_{\omega} =0,\displaystyle=0,
ATZω+ZωA+i=1p(FωMiPωMi+MiPωMiFω)\displaystyle A^{T}Z_{\omega}+Z_{\omega}A+\sum_{i=1}^{p}\big{(}F_{\omega}^{*}M_{i}P_{\omega}M_{i}+M_{i}P_{\omega}M_{i}F_{\omega}\big{)} =0,\displaystyle=0,
ATQω+QωA+FωCTC+CTCFω+i=1p(FωMiPωMi+MiPωMiFω)\displaystyle A^{T}Q_{\omega}+Q_{\omega}A+F_{\omega}^{*}C^{T}C+C^{T}CF_{\omega}+\sum_{i=1}^{p}\big{(}F_{\omega}^{*}M_{i}P_{\omega}M_{i}+M_{i}P_{\omega}M_{i}F_{\omega}\big{)} =0.\displaystyle=0.

The frequency-limited Hankel singular values σi\sigma_{i} are defined as

σi=λi(PωQω)fori=1,,n,\displaystyle\sigma_{i}=\sqrt{\lambda_{i}(P_{\omega}Q_{\omega})}\hskip 14.22636pt\textnormal{for}\hskip 14.22636pti=1,\cdots,n,

where λi()\lambda_{i}(\cdot) denotes the eigenvalues. The projection matrices in FLBT are then computed such that WkTPωWk=VkTQωVk=diag(σ1,,σk)W_{k}^{T}P_{\omega}W_{k}=V_{k}^{T}Q_{\omega}V_{k}=\text{diag}(\sigma_{1},\cdots,\sigma_{k}), where σ1,,σk\sigma_{1},\cdots,\sigma_{k} are the kk largest frequency-limited Hankel singular values of GG.

2.2 2\mathcal{H}_{2}-optimal MOR Algorithm (HOMORA) [31]

Let us define the matrices P12P_{12}, PkP_{k}, Y12Y_{12}, YkY_{k}, Z12Z_{12}, ZkZ_{k}, Q12Q_{12}, and QkQ_{k}, which satisfy the following set of linear matrix equations:

AP12+P12AkT+BBkT\displaystyle AP_{12}+P_{12}A_{k}^{T}+BB_{k}^{T} =0,\displaystyle=0,
AkPk+PkAkT+BkBkT\displaystyle A_{k}P_{k}+P_{k}A_{k}^{T}+B_{k}B_{k}^{T} =0,\displaystyle=0,
ATY12+Y12Ak+CTCk\displaystyle A^{T}Y_{12}+Y_{12}A_{k}+C^{T}C_{k} =0,\displaystyle=0,
AkTYk+YkAk+CkTCk\displaystyle A_{k}^{T}Y_{k}+Y_{k}A_{k}+C_{k}^{T}C_{k} =0,\displaystyle=0,
ATZ12+Z12Ak+i=1pMiP12Mk,i\displaystyle A^{T}Z_{12}+Z_{12}A_{k}+\sum_{i=1}^{p}M_{i}P_{12}M_{k,i} =0,\displaystyle=0,
AkTZk+ZkAk+i=1pMk,iPkMk,i\displaystyle A_{k}^{T}Z_{k}+Z_{k}A_{k}+\sum_{i=1}^{p}M_{k,i}P_{k}M_{k,i} =0,\displaystyle=0,
ATQ12+Q12Ak+CTCk+i=1pMiP12Mk,i\displaystyle A^{T}Q_{12}+Q_{12}A_{k}+C^{T}C_{k}+\sum_{i=1}^{p}M_{i}P_{12}M_{k,i} =0,\displaystyle=0,
AkTQk+QkAk+CkTCk+i=1pMk,iPkMk,i\displaystyle A_{k}^{T}Q_{k}+Q_{k}A_{k}+C_{k}^{T}C_{k}+\sum_{i=1}^{p}M_{k,i}P_{k}M_{k,i} =0.\displaystyle=0.

According to [31], the necessary conditions for achieving a local optimum of the (squared) 2\mathcal{H}_{2}-norm of the error, denoted as GGk22||G-G_{k}||_{\mathcal{H}_{2}}^{2}, are described by the following set of equations:

(Y12+2Z12)TP12+(Yk+2Zk)Pk\displaystyle-(Y_{12}+2Z_{12})^{T}P_{12}+(Y_{k}+2Z_{k})P_{k} =0,\displaystyle=0, (4)
P12TMiP12+PkMk,iPk\displaystyle-P_{12}^{T}M_{i}P_{12}+P_{k}M_{k,i}P_{k} =0,\displaystyle=0, (5)
(Y12+2Z12)TB+(Yk+2Zk)Bk\displaystyle-(Y_{12}+2Z_{12})^{T}B+(Y_{k}+2Z_{k})B_{k} =0,\displaystyle=0, (6)
CP12+CkPk\displaystyle-CP_{12}+C_{k}P_{k} =0.\displaystyle=0. (7)

Furthermore, it is shown that these optimality conditions can be met by setting the projection matrices as Vk=P12V_{k}=P_{12} and Wk=(Y12+2Z12)(P12T(Y12+2Z12))1W_{k}=(Y_{12}+2Z_{12})\big{(}P_{12}^{T}(Y_{12}+2Z_{12})\big{)}^{-1}. Starting with an initial guess for the ROM, the projection matrices are iteratively updated until convergence, at which point the optimality conditions (4)-(7) are satisfied.

3 Main Work

In this section, we define the frequency-limited 2\mathcal{H}_{2} norm and establish its connection to the frequency-limited observability Gramian. We then derive the necessary conditions for achieving a local optimum of the (squared) frequency-limited 2\mathcal{H}_{2} norm of the error. Building on these optimality conditions, we present a projection-based iterative algorithm that meets three of the four optimality conditions. The challenge of satisfying the fourth optimality condition within the projection framework is also discussed. Finally, the computational aspects of the proposed algorithm are briefly discussed.

3.1 2,ω\mathcal{H}_{2,\omega} norm Definition

The classical 2\mathcal{H}_{2} norm for LQO systems is defined in the frequency domain as follows:

G2\displaystyle||G||_{\mathcal{H}_{2}} =[trace(12πG1(jν)G1(jν)dν\displaystyle=\Bigg{[}trace\Big{(}\frac{1}{2\pi}\int_{-\infty}^{\infty}G_{1}^{*}(j\nu)G_{1}(j\nu)d\nu
+1(2π)2i=1pG2,i(jν1,jν2)G2,i(jν1,jν2)dν1dν2)]12,\displaystyle\hskip 28.45274pt+\frac{1}{(2\pi)^{2}}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\sum_{i=1}^{p}G_{2,i}^{*}(j\nu_{1},j\nu_{2})G_{2,i}(j\nu_{1},j\nu_{2})d\nu_{1}d\nu_{2}\Big{)}\Bigg{]}^{-\frac{1}{2}},

cf. [5, 31]. The 2\mathcal{H}_{2} norm quantifies the output response’s power to unit white noise across the entire frequency spectrum. However, for the problem at hand, we are only interested in the output response’s power within a specific, limited frequency range. This leads to the definition of the frequency-limited 2\mathcal{H}_{2} norm.

Definition 3.1.

The frequency-limited 2\mathcal{H}_{2} norm of the LQO system within the frequency interval [0,ω][0,\omega] rad/sec is defined as

G2,ω\displaystyle||G||_{\mathcal{H}_{2,\omega}} =[trace(12πωωG1(jν)G1(jν)dν\displaystyle=\Big{[}trace\Big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}G_{1}^{*}(j\nu)G_{1}(j\nu)d\nu
+1(2π)2ωωωωi=1pG2,i(jν1,jν2)G2,i(jν1,jν2)dν1dν2)]12.\displaystyle\hskip 28.45274pt+\frac{1}{(2\pi)^{2}}\int_{-\omega}^{\omega}\int_{-\omega}^{\omega}\sum_{i=1}^{p}G_{2,i}^{*}(j\nu_{1},j\nu_{2})G_{2,i}(j\nu_{1},j\nu_{2})d\nu_{1}d\nu_{2}\Big{)}\Big{]}^{-\frac{1}{2}}.
Proposition 3.2.

The 2,ω\mathcal{H}_{2,\omega} norm is related to the frequency-limited observability Gramian QωQ_{\omega} as follows:

G2,ω=trace(BTQωB).\displaystyle||G||_{\mathcal{H}_{2,\omega}}=\sqrt{trace(B^{T}Q_{\omega}B)}.
Proof.

Observe that

trace(12πωωG1(jν)G1(jν)𝑑ν)\displaystyle trace\Big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}G_{1}^{*}(j\nu)G_{1}(j\nu)d\nu\Big{)}
=trace(BT[12πωω(jνIA)CTC(jνIA)1]Bdν)\displaystyle=trace\Big{(}B^{T}\Big{[}\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-*}C^{T}C(j\nu I-A)^{-1}\Big{]}Bd\nu\Big{)}
=trace(BTYωB).\displaystyle=trace(B^{T}Y_{\omega}B).

Additionally, note that

trace(1(2π)2ωωωωi=1pG2,i(jν1,jν2)G2,i(jν1,jν2)dν1dν2)\displaystyle trace\Big{(}\frac{1}{(2\pi)^{2}}\int_{-\omega}^{\omega}\int_{-\omega}^{\omega}\sum_{i=1}^{p}G_{2,i}^{*}(j\nu_{1},j\nu_{2})G_{2,i}(j\nu_{1},j\nu_{2})d\nu_{1}d\nu_{2}\Big{)}
=trace(BT[12πωω(jν2IA)(i=1pMi(12πωω(jν1IA)1BBT\displaystyle=trace\Bigg{(}B^{T}\Big{[}\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu_{2}I-A)^{-*}\Big{(}\sum_{i=1}^{p}M_{i}\Big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu_{1}I-A)^{-1}BB^{T}
(jν1IA)dν1)Mi)(jν2IA)1dν2]B)\displaystyle\hskip 113.81102pt(j\nu_{1}I-A)^{-*}d\nu_{1}\Big{)}M_{i}\Big{)}(j\nu_{2}I-A)^{-1}d\nu_{2}\Big{]}B\Bigg{)}
=trace(BTZωB).\displaystyle=trace(B^{T}Z_{\omega}B).

Therefore, we have G2,ω=trace(BT(Yω+Zω)B)=trace(BT(Qω)B)||G||_{\mathcal{H}_{2,\omega}}=\sqrt{trace\big{(}B^{T}(Y_{\omega}+Z_{\omega})B\big{)}}=\sqrt{trace\big{(}B^{T}(Q_{\omega})B\big{)}}. ∎

3.2 2,ω\mathcal{H}_{2,\omega} Norm of the Error

Let us define E=GGkE=G-G_{k} with the following state-space equations

E:={xe˙(t)=[x(t)xk(t)]=Aexe(t)+Beu(t),ye(t)=y(t)yk(t)=Cexe(t)+[xe(t)TMe,1xe(t)xe(t)TMe,pxe(t)],\displaystyle E:=\begin{cases}\dot{x_{e}}(t)=\begin{bmatrix}x(t)\\ x_{k}(t)\end{bmatrix}=A_{e}x_{e}(t)+B_{e}u(t),\\ y_{e}(t)=y(t)-y_{k}(t)=C_{e}x_{e}(t)+\begin{bmatrix}x_{e}(t)^{T}M_{e,1}x_{e}(t)\\ \vdots\\ x_{e}(t)^{T}M_{e,p}x_{e}(t)\end{bmatrix},\end{cases}

wherein

Ae\displaystyle A_{e} =[A00Ak],\displaystyle=\begin{bmatrix}A&0\\ 0&A_{k}\end{bmatrix}, Be\displaystyle B_{e} =[BBk],\displaystyle=\begin{bmatrix}B\\ B_{k}\end{bmatrix},
Me,i\displaystyle M_{e,i} =[Mi00Mk,i],\displaystyle=\begin{bmatrix}M_{i}&0\\ 0&-M_{k,i}\end{bmatrix}, Ce\displaystyle C_{e} =[CCk].\displaystyle=\begin{bmatrix}C&-C_{k}\end{bmatrix}. (8)

Let us define Fe,ωF_{e,\omega} as follows

Fe,ω=12πωω(jνIAe)1𝑑ν=jπln(jωIAe).\displaystyle F_{e,\omega}=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A_{e})^{-1}d\nu=\frac{j}{\pi}ln(-j\omega I-A_{e}).

Then the frequency-limited controllability Gramian Pe,ωP_{e,\omega} and the frequency-limited observability Gramian Qe,ω=Ye,ω+Ze,ωQ_{e,\omega}=Y_{e,\omega}+Z_{e,\omega} of realization (Ae,Be,Ce,Me,1,,Me,p)(A_{e},B_{e},C_{e},M_{e,1},\cdots,M_{e,p}) can be determined by solving the following Lyapunov equations:

AePe,ω+Pe,ωAeT+Fe,ωBeBeT+BeBeTFe,ω=0,\displaystyle\hskip 85.35826ptA_{e}P_{e,\omega}+P_{e,\omega}A_{e}^{T}+F_{e,\omega}B_{e}B_{e}^{T}+B_{e}B_{e}^{T}F_{e,\omega}^{*}=0,
AeTYe,ω+Ye,ωAe+Fe,ωCeTCe+CeTCeFe,ω=0,\displaystyle\hskip 88.2037ptA_{e}^{T}Y_{e,\omega}+Y_{e,\omega}A_{e}+F_{e,\omega}^{*}C_{e}^{T}C_{e}+C_{e}^{T}C_{e}F_{e,\omega}=0,
AeTZe,ω+Ze,ωAe+i=1p(Fe,ωMe,iPe,ωMe,i+Me,iPe,ωMe,iFe,ω)=0,\displaystyle A_{e}^{T}Z_{e,\omega}+Z_{e,\omega}A_{e}+\sum_{i=1}^{p}\big{(}F_{e,\omega}^{*}M_{e,i}P_{e,\omega}M_{e,i}+M_{e,i}P_{e,\omega}M_{e,i}F_{e,\omega}\big{)}=0,
AeTQe,ω+Qe,ωAe+Fe,ωCeTCe+CeTCeFe,ω\displaystyle A_{e}^{T}Q_{e,\omega}+Q_{e,\omega}A_{e}+F_{e,\omega}^{*}C_{e}^{T}C_{e}+C_{e}^{T}C_{e}F_{e,\omega}
+i=1p(Fe,ωMe,iPe,ωMe,i+Me,iPe,ωMe,iFe,ω)=0.\displaystyle\hskip 78.24507pt+\sum_{i=1}^{p}\big{(}F_{e,\omega}^{*}M_{e,i}P_{e,\omega}M_{e,i}+M_{e,i}P_{e,\omega}M_{e,i}F_{e,\omega}\big{)}=0.

Let us partition Pe,ωP_{e,\omega}, Ye,ωY_{e,\omega}, Ze,ωZ_{e,\omega}, and Qe,ωQ_{e,\omega} according to (8) as follows:

Pe,ω\displaystyle P_{e,\omega} =[PωP12,ωP12,ωPk,ω],\displaystyle=\begin{bmatrix}P_{\omega}&P_{12,\omega}\\ P_{12,\omega}^{*}&P_{k,\omega}\end{bmatrix}, Ye,ω\displaystyle Y_{e,\omega} =[YωY12,ωY12,ωYk,ω],\displaystyle=\begin{bmatrix}Y_{\omega}&-Y_{12,\omega}\\ -Y_{12,\omega}^{*}&Y_{k,\omega}\end{bmatrix},
Ze,ω\displaystyle Z_{e,\omega} =[ZωZ12,ωZ12,ωZk,ω],\displaystyle=\begin{bmatrix}Z_{\omega}&-Z_{12,\omega}\\ -Z_{12,\omega}^{*}&Z_{k,\omega}\end{bmatrix}, Qe,ω\displaystyle Q_{e,\omega} =[QωQ12,ωQ12,ωQk,ω].\displaystyle=\begin{bmatrix}Q_{\omega}&-Q_{12,\omega}\\ -Q_{12,\omega}^{*}&Q_{k,\omega}\end{bmatrix}.

Additionally, define Fk,ωF_{k,\omega} as

Fk,ω=12πωω(jνIAk)1𝑑ν=jπln(jωIAk).\displaystyle F_{k,\omega}=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A_{k})^{-1}d\nu=\frac{j}{\pi}ln(-j\omega I-A_{k}).

The following linear matrix equations then hold:

AP12,ω+P12,ωAkT+FωBBkT+BBkTFk,ω=0,\displaystyle\hskip 110.96556ptAP_{12,\omega}+P_{12,\omega}A_{k}^{T}+F_{\omega}BB_{k}^{T}+BB_{k}^{T}F_{k,\omega}^{*}=0, (9)
AkPk,ω+Pk,ωAkT+Fk,ωBkBkT+BkBkTFk,ω=0,\displaystyle\hskip 98.16191ptA_{k}P_{k,\omega}+P_{k,\omega}A_{k}^{T}+F_{k,\omega}B_{k}B_{k}^{T}+B_{k}B_{k}^{T}F_{k,\omega}^{*}=0, (10)
ATY12,ω+Y12,ωAk+FωCTCk+CTCkFk,ω=0,\displaystyle\hskip 100.72256ptA^{T}Y_{12,\omega}+Y_{12,\omega}A_{k}+F_{\omega}^{*}C^{T}C_{k}+C^{T}C_{k}F_{k,\omega}=0, (11)
AkTYk,ω+Yk,ωAk+Fk,ωCkTCk+CkTCkFk,ω=0,\displaystyle\hskip 101.00737ptA_{k}^{T}Y_{k,\omega}+Y_{k,\omega}A_{k}+F_{k,\omega}^{*}C_{k}^{T}C_{k}+C_{k}^{T}C_{k}F_{k,\omega}=0, (12)
ATZ12,ω+Z12,ωAk+i=1p(FωMiP12,ωMk,i+MiP12,ωMk,iFk,ω)=0,\displaystyle\hskip 14.79555ptA^{T}Z_{12,\omega}+Z_{12,\omega}A_{k}+\sum_{i=1}^{p}\big{(}F_{\omega}^{*}M_{i}P_{12,\omega}M_{k,i}+M_{i}P_{12,\omega}M_{k,i}F_{k,\omega}\big{)}=0, (13)
AkTZk,ω+Zk,ωAk+i=1p(Fk,ωMk,iPk,ωMk,i+Mk,iPk,ωMk,iFk,ω)=0,\displaystyle\hskip 8.5359ptA_{k}^{T}Z_{k,\omega}+Z_{k,\omega}A_{k}+\sum_{i=1}^{p}\big{(}F_{k,\omega}^{*}M_{k,i}P_{k,\omega}M_{k,i}+M_{k,i}P_{k,\omega}M_{k,i}F_{k,\omega}\big{)}=0, (14)
ATQ12,ω+Q12,ωAk+FωCTCk+CTCkFk,ω\displaystyle A^{T}Q_{12,\omega}+Q_{12,\omega}A_{k}+F_{\omega}^{*}C^{T}C_{k}+C^{T}C_{k}F_{k,\omega}
+i=1p(FωMiP12,ωMk,i+MiP12,ωMk,iFk,ω)=0,\displaystyle\hskip 102.43008pt+\sum_{i=1}^{p}\big{(}F_{\omega}^{*}M_{i}P_{12,\omega}M_{k,i}+M_{i}P_{12,\omega}M_{k,i}F_{k,\omega}\big{)}=0, (15)
AkTQk,ω+Qk,ωAk+Fk,ωCkTCk+CkTCkFk,ω\displaystyle A_{k}^{T}Q_{k,\omega}+Q_{k,\omega}A_{k}+F_{k,\omega}^{*}C_{k}^{T}C_{k}+C_{k}^{T}C_{k}F_{k,\omega}
+i=1p(Fk,ωMk,iPk,ωMk,i+Mk,iPk,ωMk,iFk,ω)=0.\displaystyle\hskip 88.2037pt+\sum_{i=1}^{p}\big{(}F_{k,\omega}^{*}M_{k,i}P_{k,\omega}M_{k,i}+M_{k,i}P_{k,\omega}M_{k,i}F_{k,\omega}\big{)}=0. (16)

Finally, the 2,ω\mathcal{H}_{2,\omega} norm of EE can be expressed as:

E2,ω\displaystyle||E||_{\mathcal{H}_{2,\omega}} =trace(BeTQe,ωBe)\displaystyle=\sqrt{trace(B_{e}^{T}Q_{e,\omega}B_{e})}
=trace(BTQωB2BTQ12,ωBk+BkTQk,ωBk).\displaystyle=\sqrt{trace(B^{T}Q_{\omega}B-2B^{T}Q_{12,\omega}B_{k}+B_{k}^{T}Q_{k,\omega}B_{k})}.
Corollary 3.3.

The expression E2,ω2=G2,ω22G,Gk2,ω+Gk2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}=||G||_{\mathcal{H}_{2,\omega}}^{2}-2\langle G,G_{k}\rangle_{\mathcal{H}_{2,\omega}}+||G_{k}||_{\mathcal{H}_{2,\omega}}^{2} holds, where G,Gk2,ω\langle G,G_{k}\rangle_{\mathcal{H}_{2,\omega}} denotes the 2,ω\mathcal{H}_{2,\omega} inner product of GG and GkG_{k}.

Proof.

The first and last terms in the expression for E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2} are straightforward. The main objective is to demonstrate that the middle term corresponds to the 2,ω\mathcal{H}_{2,\omega} inner product of GG and GkG_{k}. By expanding the definition of the inner product, we can express it as:

G,Gk2,ω\displaystyle\langle G,G_{k}\rangle_{\mathcal{H}_{2,\omega}} =12πωωtrace(G1(jν)Gk,1(jν)dν)\displaystyle=\frac{1}{2\pi}\int_{-\omega}^{\omega}trace\big{(}G_{1}^{*}(j\nu)G_{k,1}(j\nu)d\nu\big{)}
+14π2trace(ωωωωi=1p(G2,i(jν1,jν2)Gk,2,i(jν1,jν2))dν1dν2).\displaystyle+\frac{1}{4\pi^{2}}trace\Big{(}\int_{-\omega}^{\omega}\int_{-\omega}^{\omega}\sum_{i=1}^{p}\big{(}G_{2,i}^{*}(j\nu_{1},j\nu_{2})G_{k,2,i}(j\nu_{1},j\nu_{2})\big{)}d\nu_{1}d\nu_{2}\Big{)}.

Furthermore,

trace(12πωωG1(jν)Gk,1(jν)𝑑ν)\displaystyle trace\Big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}G_{1}^{*}(j\nu)G_{k,1}(j\nu)d\nu\Big{)}
=trace(BT(12πωω(jνIA)CTCk(jνIAk)1𝑑ν)Bk),\displaystyle=trace\Big{(}B^{T}\big{(}\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-*}C^{T}C_{k}(j\nu I-A_{k})^{-1}d\nu\big{)}B_{k}\Big{)},
trace(14π2ωωωωi=1pG2,i(jν1,jν2)Gk,2,i(jν1,jν2)dν1dν2)\displaystyle trace\Big{(}\frac{1}{4\pi^{2}}\int_{-\omega}^{\omega}\int_{-\omega}^{\omega}\sum_{i=1}^{p}G_{2,i}^{*}(j\nu_{1},j\nu_{2})G_{k,2,i}(j\nu_{1},j\nu_{2})d\nu_{1}d\nu_{2}\Big{)}
=trace(BT[14π2ωω(jν1IA)(i=1pMi(ωω(jν2IA)1B\displaystyle=trace\Bigg{(}B^{T}\Big{[}\frac{1}{4\pi^{2}}\int_{-\omega}^{\omega}(j\nu_{1}I-A)^{-*}\Big{(}\sum_{i=1}^{p}M_{i}\big{(}\int_{-\omega}^{\omega}(j\nu_{2}I-A)^{-1}B
BkT(jν2IAk)dν2)Mk,i)(jν1IAk)1]Bk).\displaystyle\hskip 113.81102ptB_{k}^{T}(j\nu_{2}I-A_{k})^{-*}d\nu_{2}\big{)}M_{k,i}\Big{)}(j\nu_{1}I-A_{k})^{-1}\Big{]}B_{k}\Bigg{)}.

The Sylvester equations (11) and (13) can be solved by evaluating the following integrals:

Y12,ω\displaystyle Y_{12,\omega} =12πωω(jνIA)CTCk(jνIAk)1𝑑ν,\displaystyle=\frac{1}{2\pi}\int_{-\omega}^{\omega}(j\nu I-A)^{-*}C^{T}C_{k}(j\nu I-A_{k})^{-1}d\nu,
Z12,ω\displaystyle Z_{12,\omega} =14π2ωω(jν1IA)(i=1pMi(ωω(jν2IA)1B\displaystyle=\frac{1}{4\pi^{2}}\int_{-\omega}^{\omega}(j\nu_{1}I-A)^{-*}\Bigg{(}\sum_{i=1}^{p}M_{i}\Big{(}\int_{-\omega}^{\omega}(j\nu_{2}I-A)^{-1}B
BkT(jν2IAk))Mk,i)(jν1IAk)1dν1;\displaystyle\hskip 113.81102ptB_{k}^{T}(j\nu_{2}I-A_{k})^{-*}\Big{)}M_{k,i}\Bigg{)}(j\nu_{1}I-A_{k})^{-1}d\nu_{1};

cf. [26, 42] Consequently, the 2,ω\mathcal{H}_{2,\omega} inner product between GG and GkG_{k} can be written as

G,Gk2,ω=trace(BT(Y12,ω+Z12,ω)Bk)=trace(BTQ12,ωBk).\displaystyle\langle G,G_{k}\rangle_{\mathcal{H}_{2,\omega}}=trace\big{(}B^{T}(Y_{12,\omega}+Z_{12,\omega})B_{k}\big{)}=trace(B^{T}Q_{12,\omega}B_{k}).

3.3 Optimality Conditions

In this subsection, we present the necessary conditions for achieving a local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}. These optimality conditions require the introduction of several new variables. We begin by defining Z¯12\bar{Z}_{12} and Z¯k\bar{Z}_{k} as the solutions to the following equations:

ATZ¯12+Z¯12Ak+i=1pMiP12,ωMk,i=0,\displaystyle A^{T}\bar{Z}_{12}+\bar{Z}_{12}A_{k}+\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}=0,
AkTZ¯k+Z¯kAk+i=1pMk,iPk,ωMk,i=0.\displaystyle A_{k}^{T}\bar{Z}_{k}+\bar{Z}_{k}A_{k}+\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i}=0.

It is important to note that P12,ωP_{12,\omega}, Pk,ωP_{k,\omega}, Z12,ωZ_{12,\omega} and Zk,ωZ_{k,\omega} can be derived from P12P_{12}, PkP_{k}, Z¯12\bar{Z}_{12} and Z¯k\bar{Z}_{k}, respectively, by restricting the integration limits to [0,ω][0,\omega] rad/sec in their integral definitions. Next, we define P~12\tilde{P}_{12}, P~k\tilde{P}_{k}, Z~12\tilde{Z}_{12}, and Z~k\tilde{Z}_{k} as follows:

P~12\displaystyle\tilde{P}_{12} =P12|P12|0ω=P12P12,ω,\displaystyle=P_{12}\Big{|}_{-\infty}^{\infty}-P_{12}\Big{|}_{0}^{\omega}=P_{12}-P_{12,\omega}, (17)
P~k\displaystyle\tilde{P}_{k} =Pk|Pk|0ω=PkPk,ω,\displaystyle=P_{k}\Big{|}_{-\infty}^{\infty}-P_{k}\Big{|}_{0}^{\omega}=P_{k}-P_{k,\omega}, (18)
Z~12\displaystyle\tilde{Z}_{12} =Z¯12|Z¯12|0ω=Z¯12Z12,ω,\displaystyle=\bar{Z}_{12}\Big{|}_{-\infty}^{\infty}-\bar{Z}_{12}\Big{|}_{0}^{\omega}=\bar{Z}_{12}-Z_{12,\omega}, (19)
Z~k\displaystyle\tilde{Z}_{k} =Z¯k|Z¯k|0ω=Z¯kZk,ω.\displaystyle=\bar{Z}_{k}\Big{|}_{-\infty}^{\infty}-\bar{Z}_{k}\Big{|}_{0}^{\omega}=\bar{Z}_{k}-Z_{k,\omega}. (20)

Additionally, we define VV, WW and LωL_{\omega} as follows:

V\displaystyle V =BkBTZ¯12BkBkTZ¯k+P12TCTCkPkCkCkT\displaystyle=B_{k}B^{T}\bar{Z}_{12}-B_{k}B_{k}^{T}\bar{Z}_{k}+P_{12}^{T}C^{T}C_{k}-P_{k}C_{k}C_{k}^{T}
+P12Ti=1pMiP12,ωMk,iPki=1pMk,iPk,ωMk,i,\displaystyle\hskip 56.9055pt+P_{12}^{T}\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}-P_{k}\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i},
W\displaystyle W =j2π(jνIAk,V),\displaystyle=\frac{j}{2\pi}\mathcal{L}(-j\nu I-A_{k},V),
Lω\displaystyle L_{\omega} =Q12,ωP~12Z~12P12,ω+Qk,ωP~k+Z~kPk,ω+W,\displaystyle=-Q_{12,\omega}^{*}\tilde{P}_{12}-\tilde{Z}_{12}^{*}P_{12,\omega}+Q_{k,\omega}\tilde{P}_{k}+\tilde{Z}_{k}P_{k,\omega}+W^{*},

where (jνIAk,V)\mathcal{L}(-j\nu I-A_{k},V) denotes the Fr’echet derivative of the matrix logarithm ln(jνIAk)ln(-j\nu I-A_{k}) in the direction of the matrix VV, specifically:

(jνIAk,V)\displaystyle\mathcal{L}(-j\nu I-A_{k},V) =01(ν(jνIAkI)+I)1V\displaystyle=\int_{0}^{1}\big{(}\nu(-j\nu I-A_{k}-I)+I\big{)}^{-1}V
(ν(jνIAkI)+I)1dν;\displaystyle\hskip 113.81102pt\big{(}\nu(-j\nu I-A_{k}-I)+I\big{)}^{-1}d\nu;

cf. [47].

We are now ready to state the necessary conditions for a local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}.

Theorem 3.4.

The local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2} must satisfy the following necessary conditions:

(Y12,ω+2Z12,ω)P12,ω+(Yk,ω+2Zk,ω)Pk,ω+Lω\displaystyle-(Y_{12,\omega}+2Z_{12,\omega})^{*}P_{12,\omega}+(Y_{k,\omega}+2Z_{k,\omega})P_{k,\omega}+L_{\omega} =0,\displaystyle=0, (21)
P12,ωMiP12,ω+Pk,ωMk,iPk,ω\displaystyle-P_{12,\omega}^{*}M_{i}P_{12,\omega}+P_{k,\omega}M_{k,i}P_{k,\omega} =0,\displaystyle=0, (22)
(Y12,ω+2Z12,ω)B+(Yk,ω+2Zk,ω)Bk\displaystyle-(Y_{12,\omega}+2Z_{12,\omega})^{*}B+(Y_{k,\omega}+2Z_{k,\omega})B_{k} =0,\displaystyle=0, (23)
CP12,ω+CkPk,ω\displaystyle-CP_{12,\omega}+C_{k}P_{k,\omega} =0.\displaystyle=0. (24)
Proof.

The proof of this theorem is lengthy and complex, so it is provided in the Appendix A. ∎

3.4 Comparison with Local Optimum of E22||E||_{\mathcal{H}_{2}}^{2}

In this subsection, we compare the necessary conditions for the local optima of E22||E||_{\mathcal{H}_{2}}^{2} and E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}. To begin, we provide the expression for E2||E||_{\mathcal{H}_{2}} as presented in [26]. The controllability Gramian PeP_{e} and the observability Gramian Qe=Ye+ZeQ_{e}=Y_{e}+Z_{e} of realization (Ae,Be,Ce,Me,1,,Me,p)(A_{e},B_{e},C_{e},M_{e,1},\cdots,M_{e,p}) can be computed by solving the following Lyapunov equations:

AePe+PeAeT+BeBeT\displaystyle A_{e}P_{e}+P_{e}A_{e}^{T}+B_{e}B_{e}^{T} =0,\displaystyle=0,
AeTYe+YeAe+CeTCe\displaystyle A_{e}^{T}Y_{e}+Y_{e}A_{e}+C_{e}^{T}C_{e} =0,\displaystyle=0,
AeTZe+ZeAe+i=1pMe,iPeMe,i\displaystyle A_{e}^{T}Z_{e}+Z_{e}A_{e}+\sum_{i=1}^{p}M_{e,i}P_{e}M_{e,i} =0,\displaystyle=0,
AeTQe+QeAe+CeTCe+i=1pMe,iPeMe,i\displaystyle A_{e}^{T}Q_{e}+Q_{e}A_{e}+C_{e}^{T}C_{e}+\sum_{i=1}^{p}M_{e,i}P_{e}M_{e,i} =0.\displaystyle=0.

We then partition PeP_{e}, YeY_{e}, ZeZ_{e}, and QeQ_{e} according to (8) as follows:

Pe\displaystyle P_{e} =[PP12P12TPk],\displaystyle=\begin{bmatrix}P&P_{12}\\ P_{12}^{T}&P_{k}\end{bmatrix}, Ye\displaystyle Y_{e} =[YY12Y12TYk],\displaystyle=\begin{bmatrix}Y&-Y_{12}\\ -Y_{12}^{T}&Y_{k}\end{bmatrix},
Ze\displaystyle Z_{e} =[ZZ12Z12TZk],\displaystyle=\begin{bmatrix}Z&-Z_{12}\\ -Z_{12}^{T}&Z_{k}\end{bmatrix}, Qe\displaystyle Q_{e} =[QQ12Q12TQk].\displaystyle=\begin{bmatrix}Q&-Q_{12}\\ -Q_{12}^{T}&Q_{k}\end{bmatrix}.

The 2\mathcal{H}_{2} norm of EE can be expressed as:

E2\displaystyle||E||_{\mathcal{H}_{2}} =trace(BeTQeBe)\displaystyle=\sqrt{trace(B_{e}^{T}Q_{e}B_{e})}
=trace(BTQB2BTQ12Bk+BkTQkBk).\displaystyle=\sqrt{trace(B^{T}QB-2B^{T}Q_{12}B_{k}+B_{k}^{T}Q_{k}B_{k})}.

The optimality conditions (21)-(24) and (4)-(7) are similar, but there are some important differences. By restricting the integration limit of PeP_{e} and QeQ_{e} to [0,ω][0,\omega] rad/sec, the optimality conditions (22)-(24) can be derived from (5)-(7), respectively. However, the optimality condition (4) does not simplify to (21) by merely limiting the integration range.

Furthermore, from the optimality conditions (5)-(7), we can deduce the optimal selections for Mk,iM_{k,i}, BkB_{k}, and CkC_{k} as:

Mk,i\displaystyle M_{k,i} =Pk1P12TMiP12Pk1,\displaystyle=P_{k}^{-1}P_{12}^{T}M_{i}P_{12}P_{k}^{-1}, (25)
Bk\displaystyle B_{k} =(Yk+2Zk)1(Y12+2Z12)TB,\displaystyle=(Y_{k}+2Z_{k})^{-1}(Y_{12}+2Z_{12})^{T}B, (26)
Ck\displaystyle C_{k} =CP12Pk1.\displaystyle=CP_{12}P_{k}^{-1}. (27)

By restricting the integration limits of PeP_{e} and QeQ_{e} to [0,ω][0,\omega] rad/sec, we can derive the frequency-limited optimal choices for Mk,iM_{k,i}, BkB_{k}, and CkC_{k} from (25)-(27) as follows:

Mk,i\displaystyle M_{k,i} =Pk,ω1P12,ωMiP12,ωPk,ω1,\displaystyle=P_{k,\omega}^{-1}P_{12,\omega}^{*}M_{i}P_{12,\omega}P_{k,\omega}^{-1}, (28)
Bk\displaystyle B_{k} =(Yk,ω+2Zk,ω)1(Y12,ω+2Z12,ω)B,\displaystyle=(Y_{k,\omega}+2Z_{k,\omega})^{-1}(Y_{12,\omega}+2Z_{12,\omega})^{*}B, (29)
Ck\displaystyle C_{k} =CP12,ωPk,ω1.\displaystyle=CP_{12,\omega}P_{k,\omega}^{-1}. (30)

The optimal projection matrices VkV_{k} and WkW_{k} for computing a local optimum of E22||E||_{\mathcal{H}_{2}}^{2} are given by:

Vk\displaystyle V_{k} =P12Pk1,\displaystyle=P_{12}P_{k}^{-1}, Wk\displaystyle W_{k} =(Y12+2Z12)(Yk+2Zk)1.\displaystyle=(Y_{12}+2Z_{12})(Y_{k}+2Z_{k})^{-1}.

In the frequency-limited scenario, by setting:

Vk\displaystyle V_{k} =P12,ωPk,ω1,\displaystyle=P_{12,\omega}P_{k,\omega}^{-1}, Wk\displaystyle W_{k} =(Y12,ω+2Z12,ω)(Yk,ω+2Zk,ω)1,\displaystyle=(Y_{12,\omega}+2Z_{12,\omega})(Y_{k,\omega}+2Z_{k,\omega})^{-1},

we make with the optimal choices for Mk,iM_{k,i}, BkB_{k}, and CkC_{k} as indicated by the optimality conditions (22)-(24). However, with this choice of VkV_{k} and WkW_{k}, determining an optimal AkA_{k} remains elusive. By enforcing the Petrov–Galerkin projection condition WkVk=IW_{k}^{*}V_{k}=I, we ensure

(Y12,ω+2Z12,ω)P12,ω+(Yk,ω+2Zk,ω)Pk,ω=0.\displaystyle-(Y_{12,\omega}+2Z_{12,\omega})^{*}P_{12,\omega}+(Y_{k,\omega}+2Z_{k,\omega})P_{k,\omega}=0.

It is important to note that, generally, LωL_{\omega} does not simplify to zero with this choice of projection matrices. Consequently, this selection introduces a deviation in the optimality condition (21) quantified by LωL_{\omega}. In contrast, in the classical 2\mathcal{H}_{2}-optimal MOR framework, enforcing the Petrov–Galerkin projection condition WkTVk=IW_{k}^{T}V_{k}=I satisfies the optimality condition (4) and achieves a local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}. In summary, it is generally impossible to attain a local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2} within the projection framework. While the optimality conditions (22)-(24) can be precisely met, the optimality condition (21) can only be approximately satisfied.

Up to this point, we have determined the appropriate projection matrices Vk=P12,ωPk,ω1V_{k}=P_{12,\omega}P_{k,\omega}^{-1} and Wk=(Y12,ω+2Z12,ω)(Yk,ω+2Zk,ω)1W_{k}=(Y_{12,\omega}+2Z_{12,\omega})(Y_{k,\omega}+2Z_{k,\omega})^{-1} for the problem at hand. However, these matrices depend on the ROM (Ak,Bk,Ck,Mk,1,,Mk,p)(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}), which is unknown. Therefore, equations (2) and (9)-(14) form a coupled system of equations, expressed as:

(Ak,Bk,Ck,Mk,1,,Mk,p)\displaystyle(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}) =f(P12,ω,Pk,ω,Y12,ω,Yk,ω,Z12,ω,Zk,ω),\displaystyle=f(P_{12,\omega},P_{k,\omega},Y_{12,\omega},Y_{k,\omega},Z_{12,\omega},Z_{k,\omega}),
(P12,ω,Pk,ω,Y12,ω,Yk,ω,Z12,ω,Zk,ω)\displaystyle(P_{12,\omega},P_{k,\omega},Y_{12,\omega},Y_{k,\omega},Z_{12,\omega},Z_{k,\omega}) =g(Ak,Bk,Ck,Mk,1,,Mk,p).\displaystyle=g(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}).

The stationary points of the function

(Ak,Bk,Ck,Mk,1,,Mk,p)=f(g(Ak,Bk,Ck,Mk,1,,Mk,p))\displaystyle(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p})=f\big{(}g(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p})\big{)}

satisfy the optimality conditions (22)-(24). Additionally, by enforcing the Petrov-Galerkin projection condition WkVk=IW_{k}^{*}V_{k}=I, the optimality condition (21) is nearly satisfied, with the deviation quantified by LωL_{\omega}. In the classical 2\mathcal{H}_{2}-optimal MOR scenario, the situation is similar; however, enforcing the Petrov–Galerkin projection condition WkTVk=IW_{k}^{T}V_{k}=I at the stationary points ensures that all the optimality conditions (4)-(7) are fully satisfied.

In the classical 2\mathcal{H}_{2}-optimal MOR case, it is demonstrated that if the reduction matrices are chosen as Vk=P12V_{k}=P_{12} and Wk=Y12+2Z12W_{k}=Y_{12}+2Z_{12} instead of Vk=P12Pk1V_{k}=P_{12}P_{k}^{-1} and Wk=(Y12+2Z12)(Yk+2Zk)1W_{k}=(Y_{12}+2Z_{12})(Y_{k}+2Z_{k})^{-1}, the stationary points satisfy:

Pk\displaystyle P_{k} =I,andYk+2Zk=I.\displaystyle=I,\hskip 14.22636pt\textnormal{and}\hskip 14.22636ptY_{k}+2Z_{k}=I.

Thus, the projection matrices Vk=P12V_{k}=P_{12} and Wk=Y12+2Z12W_{k}=Y_{12}+2Z_{12}, along with the Petrov–Galerkin projection condition WkTVk=IW_{k}^{T}V_{k}=I, satisfy all the optimality conditions (4)-(7). However, using Vk=P12,ωV_{k}=P_{12,\omega} and Wk=Y12,ω+2Z12,ωW_{k}=Y_{12,\omega}+2Z_{12,\omega} along with the Petrov–Galerkin projection condition WkVk=IW_{k}^{*}V_{k}=I does not satisfy any of the optimality conditions (21)-(24).

Theorem 3.5.

If WkFωB=Fk,ωBkW_{k}^{*}F_{\omega}B=F_{k,\omega}B_{k}, CFωVk=CkFk,ωCF_{\omega}V_{k}=C_{k}F_{k,\omega}, and VkMiFωVk=Mk,iFk,ωV_{k}^{*}M_{i}F_{\omega}V_{k}=M_{k,i}F_{k,\omega}, then selecting Vk=P12,ωV_{k}=P_{12,\omega} and Wk=Y12,ω+2Z12,ωW_{k}=Y_{12,\omega}+2Z_{12,\omega}, together with the Petrov–Galerkin projection condition WkVk=IW_{k}^{*}V_{k}=I ensures that:

Pk,ω\displaystyle P_{k,\omega} =I,andYk,ω+2Zk,ω=I,\displaystyle=I,\hskip 14.22636pt\textnormal{and}\hskip 14.22636ptY_{k,\omega}+2Z_{k,\omega}=I, (31)

which in turn satisfies the optimality conditions (22)-(24).

Proof.

The proof is provided in Appendix B. ∎

In general, choosing Vk=P12,ωV_{k}=P_{12,\omega} and Wk=Y12,ω+2Z12,ωW_{k}=Y_{12,\omega}+2Z_{12,\omega} while enforcing the Petrov–Galerkin projection condition WkVk=IW_{k}^{*}V_{k}=I does not meet the conditions WkFωB=Fk,ωBkW_{k}^{*}F_{\omega}B=F_{k,\omega}B_{k}, CFωVk=CkFk,ωCF_{\omega}V_{k}=C_{k}F_{k,\omega}, and VkTMiFωVk=Mk,iFk,ωV_{k}^{T}M_{i}F_{\omega}V_{k}=M_{k,i}F_{k,\omega}. Consequently, the condition (31) is not satisfied, and therefore, optimality conditions (22)-(24) are not fulfilled.

3.5 Algorithm

So far, for simplicity, we have considered VkV_{k} and WkW_{k} as complex matrices in the problem under consideration. However, in practice, using complex projection matrices results in a complex ROM, which is undesirable since most practical dynamical systems are represented by real mathematical models. This issue can be addressed by extending the desired frequency interval to include negative frequencies, i.e., [ω,0][-\omega,0] rad/sec. Additionally, we have assumed that the desired frequency interval starts from 0 rad/sec for simplicity. For any general frequency interval [ω2,ω1][ω1,ω2][-\omega_{2},-\omega_{1}]\cup[\omega_{1},\omega_{2}] rad/sec, the only modification needed is in the computation of FωF_{\omega} and Fk,ωF_{k,\omega} given by:

Fω\displaystyle F_{\omega} =Re(jπln((jω1I+A)1(jω2I+A))),\displaystyle=Re\Big{(}\frac{j}{\pi}ln\big{(}(j\omega_{1}I+A\big{)}^{-1}(j\omega_{2}I+A)\big{)}\Big{)}, (32)
Fk,ω\displaystyle F_{k,\omega} =Re(jπln((jω1I+Ak)1(jω2I+Ak)));\displaystyle=Re\Big{(}\frac{j}{\pi}ln\big{(}(j\omega_{1}I+A_{k}\big{)}^{-1}(j\omega_{2}I+A_{k})\big{)}\Big{)}; (33)

see [48] for more details.

We are now ready to present the algorithm, referred to in this paper as the “Frequency-limited 2\mathcal{H}_{2} Near-optimal Iterative Algorithm (FLHNOIA)”. The algorithm begins with an arbitrary initial guess for the ROM and iteratively updates it until convergence is achieved. Convergence is quantified by the stagnation in the relative change of the state-space matrices of the ROM. In each iteration, Steps (4) and (5) calculate the projection matrices, while Steps (6)-(10) bi-orthogonalize these matrices using the bi-orthogonal Gram–Schmidt method to ensure that WkTVk=IW_{k}^{T}V_{k}=I.

Algorithm 1 FLHNOIA

Input: Full order system: (A,B,C,M1,,Mp)(A,B,C,M_{1},\cdots,M_{p}); Desired frequency interval: [ω1,ω2][\omega_{1},\omega_{2}] rad/sec; Initial guess of ROM: (Ak,Bk,Ck,Mk,1,,Mk,p)(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}); Tolerance: toltol. Output: ROM: (Ak,Bk,Ck,Mk,1,,Mk,p)(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}).

1:  Compute FωF_{\omega} from (32).
2:  while(relative change in (Ak,Bk,Ck,Mk,1,,Mk,p)(A_{k},B_{k},C_{k},M_{k,1},\cdots,M_{k,p}) >> toltol)
3:  Compute Fk,ωF_{k,\omega} from (33).
4:  Solve equations (9)-(14) to compute P12,ωP_{12,\omega}, Pk,ωP_{k,\omega}, Y12,ωY_{12,\omega}, Yk,ωY_{k,\omega}, Z12,ωZ_{12,\omega}, and Zk,ωZ_{k,\omega}.
5:  Set Vk=P12,ωPk,ω1V_{k}=P_{12,\omega}P_{k,\omega}^{-1} and Wk=(Y12,ω+2Z12,ω)(Yk,ω+2Zk,ω)1W_{k}=(Y_{12,\omega}+2Z_{12,\omega})(Y_{k,\omega}+2Z_{k,\omega})^{-1}.
6:  for l=1,,kl=1,\ldots,k do
7:  v=Vk(:,l)v=V_{k}(:,l), v=j=1l(IVk(:,j)Wk(:,j)T)vv=\prod_{j=1}^{l}\big{(}I-V_{k}(:,j)W_{k}(:,j)^{T}\big{)}v.
8:  w=Wk(:,l)w=W_{k}(:,l), w=j=1l(IWk(:,j)Vk(:,j)T)ww=\prod_{j=1}^{l}\big{(}I-W_{k}(:,j)V_{k}(:,j)^{T}\big{)}w.
9:  v=vv2v=\frac{v}{||v||_{2}}, w=ww2w=\frac{w}{||w||_{2}}, v=vwTvv=\frac{v}{w^{T}v}, Vk(:,l)=vV_{k}(:,l)=v, Wk(:,l)=wW_{k}(:,l)=w.
10:  end for
11:  Update the ROM as Ak=WkTAVkA_{k}=W_{k}^{T}AV_{k}, Bk=WkTBB_{k}=W_{k}^{T}B, Ck=CVkC_{k}=CV_{k}, Mk,i=VkTMiVkM_{k,i}=V_{k}^{T}M_{i}V_{k}.
12:  end while
Remark 1.

For evaluating convergence, observing the stagnation of the ROM poles provides a more dependable measure compared to analyzing state-space realizations. This is because 2\mathcal{H}_{2}-optimal MOR techniques often produce ROMs with varied state-space representations but identical transfer functions. Therefore, the stagnation of ROM poles is commonly used as a convergence criterion in 2\mathcal{H}_{2}-optimal MOR algorithms, due to its proven effectiveness [44].

3.6 Computational Aspects

In this subsection, we briefly discuss the efficient implementation of FLHNOIA. Step (1) of FLHNOIA involves the computation of the matrix logarithm FωF_{\omega}, which can become computationally expensive when the order nn of the original model is large. In such cases, Krylov subspace-based methods proposed in [38] can be utilized to approximate FωBF_{\omega}B, CFωCF_{\omega}, and MFωMF_{\omega}. The most computationally intensive task in each iteration is solving the Sylvester equations (9), (11), and (13). The state-space matrices of most high-order dynamical systems are sparse, making these equations “sparse-dense” Sylvester equations, which are commonly encountered in 2\mathcal{H}_{2}-optimal MOR algorithms. A “sparse-dense” Sylvester equation typically has the structure:

𝒜+𝒞+𝒟\displaystyle\mathcal{A}\mathcal{B}+\mathcal{B}\mathcal{C}+\mathcal{D}\mathcal{E} =0,\displaystyle=0,

where the large matrices 𝒜n×n\mathcal{A}\in\mathbb{R}^{n\times n} and 𝒟n×d\mathcal{D}\in\mathbb{R}^{n\times d} (dnd\ll n) are sparse, while the smaller matrices 𝒞r×r\mathcal{C}\in\mathbb{R}^{r\times r} and d×r\mathcal{E}\in\mathbb{R}^{d\times r} are dense. An efficient algorithm for solving this type of Sylvester equation is proposed in [30]. The remaining steps in FLHNOIA involve basic matrix computations and the solution of simple Lyapunov equations, which can be executed with minimal computational cost.

4 Illustrative Example

This section provides an illustrative example to validate the key properties of FLHNOIA. Consider a sixth-order LQO system defined by the following state-space representation:

A\displaystyle A =[92910082192100000010000001000000100000010],\displaystyle=\begin{bmatrix}-9&-29&-100&-82&-19&-2\\ 1&0&0&0&0&0\\ 0&1&0&0&0&0\\ 0&0&1&0&0&0\\ 0&0&0&1&0&0\\ 0&0&0&0&1&0\end{bmatrix},
B\displaystyle B =[100000]T,\displaystyle=\begin{bmatrix}1&0&0&0&0&0\end{bmatrix}^{T},
C\displaystyle C =[000011],\displaystyle=\begin{bmatrix}0&0&0&0&-1&1\end{bmatrix},
M1\displaystyle M_{1} =diag(0.7,0.4,0.1,0.1,0.1,0.1).\displaystyle=diag(0.7,0.4,0.1,0.1,0.1,0.1).

The desired frequency range for this example is [5,6][5,6] rad/sec. To initialize FLHNOIA, the following initial guess is employed:

Ak\displaystyle A_{k} =[0.03400.14000.01240.14000.15790.14240.04840.22750.1438],\displaystyle=\begin{bmatrix}-0.0340&-0.1400&0.0124\\ 0.1400&-0.1579&0.1424\\ 0.0484&-0.2275&-0.1438\end{bmatrix},
Bk\displaystyle B_{k} =[0.15920.20760.1170]T,\displaystyle=\begin{bmatrix}-0.1592&0.2076&0.1170\end{bmatrix}^{T},
Ck\displaystyle C_{k} =[0.15920.20760.0433],\displaystyle=\begin{bmatrix}-0.1592&-0.2076&0.0433\end{bmatrix},
Mk,1\displaystyle M_{k,1} =[0.00300.00290.00040.00290.00320.00440.00040.00440.1763].\displaystyle=\begin{bmatrix}0.0030&0.0029&-0.0004\\ 0.0029&0.0032&0.0044\\ -0.0004&0.0044&0.1763\end{bmatrix}.

FLHNOIA was terminated when the change in eigenvalues of AkA_{k} stagnated, as the change in the ROM’s state-space realization persisted. The resulting final ROM is:

Ak\displaystyle A_{k} =[2.04311.57581.67387.96945.42141.109811.02290.80180.4171],\displaystyle=\begin{bmatrix}-2.0431&-1.5758&1.6738\\ -7.9694&-5.4214&1.1098\\ -11.0229&-0.8018&-0.4171\end{bmatrix},
Bk\displaystyle B_{k} =[0.00000.03530.0366]T,\displaystyle=\begin{bmatrix}0.0000&0.0353&0.0366\end{bmatrix}^{T},
Ck\displaystyle C_{k} =[0.27140.04070.0183],\displaystyle=\begin{bmatrix}0.2714&0.0407&-0.0183\end{bmatrix},
Mk,1\displaystyle M_{k,1} =[68.526456.905952.721256.90593244.33141886.990752.72121886.99071103.3966].\displaystyle=\begin{bmatrix}68.5264&56.9059&-52.7212\\ 56.9059&3244.3314&-1886.9907\\ -52.7212&-1886.9907&1103.3966\end{bmatrix}.

The numerical results below confirm that this ROM effectively satisfies the optimality conditions (22)-(24):

P12,ωTMiP12,ω+Pk,ωMk,iPk,ω2\displaystyle||-P_{12,\omega}^{T}M_{i}P_{12,\omega}+P_{k,\omega}M_{k,i}P_{k,\omega}||_{2} =1.5960×1011,\displaystyle=1.5960\times 10^{-11},
(Y12,ω+2Z12,ω)TB+(Yk,ω+2Zk,ω)Bk2\displaystyle||-(Y_{12,\omega}+2Z_{12,\omega})^{T}B+(Y_{k,\omega}+2Z_{k,\omega})B_{k}||_{2} =2.3583×106,\displaystyle=2.3583\times 10^{-6},
CP12,ω+CkPk,ω2\displaystyle||-CP_{12,\omega}+C_{k}P_{k,\omega}||_{2} =1.3523×109.\displaystyle=1.3523\times 10^{-9}.

Next, a third-order ROM is generated using BT, FLBT, and HOMORA, with the same initial ROM used to initialize HOMORA. Figures 1 and 2 display the relative error on a logarithmic scale within the specified frequency range of 55 to 66 rad/sec. As illustrated, FLBT and FLHNOIA exhibit superior accuracy.

Refer to caption
Figure 1: Relative Error |G1(jν)Gk,1(jν)||G1(jν)|\frac{|G_{1}(j\nu)-G_{k,1}(j\nu)|}{|G_{1}(j\nu)|} within [5,6][5,6] rad/sec
Refer to caption
Figure 2: Relative Error |G2,1(jν)Gk,2,1(jν)||G2,1(jν)|\frac{|G_{2,1}(j\nu)-G_{k,2,1}(j\nu)|}{|G_{2,1}(j\nu)|} within [5,6][5,6] rad/sec

5 Conclusion

This research addresses the problem of 2\mathcal{H}_{2}-optimal MOR within a specified finite frequency range. To measure the output strength within this range, we introduce the frequency-limited 2\mathcal{H}_{2} norm for LQO systems. We derive the necessary conditions for achieving local optima of the squared frequency-limited 2\mathcal{H}_{2} norm of the error and compare these conditions to those of the standard, unconstrained 2\mathcal{H}_{2}-optimal MOR problem. The study highlights the limitations of the Petrov-Galerkin projection method in satisfying all optimality conditions in the frequency-limited context. As a result, we propose a Petrov-Galerkin projection algorithm that meets three out of the four optimality conditions. Numerical experiments are conducted to validate the theoretical results and demonstrate the algorithm’s effectiveness in achieving high accuracy within the specified frequency range.

Appendix A

In this appendix, we present the proof of Theorem 3.4. Throughout the proof, the following properties of the trace operation are utilized repeatedly:

  1. 1.

    Trace of Hermitian: trace(XYZ)=trace(ZYZ)trace(XYZ)=trace(Z^{*}Y^{*}Z^{*}).

  2. 2.

    Circular permutation in trace: trace(XYZ)=trace(ZXY)=trace(YZX)trace(XYZ)=trace(ZXY)=trace(YZX).

  3. 3.

    Trace of addition: trace(X+Y+Z)=trace(X)+trace(Y)+trace(Z)trace(X+Y+Z)=trace(X)+trace(Y)+trace(Z);

cf. [49].

Let us define the cost function JJ as the component of EH2,ω2||E||_{H_{2,\omega}}^{2} that depends on the ROM, expressed as:

J=trace(2BTQ12,ωBk+BkTQk,ωBk).\displaystyle J=trace(-2B^{T}Q_{12,\omega}B_{k}+B_{k}^{T}Q_{k,\omega}B_{k}).

When a small first-order perturbation ΔAk\Delta_{A_{k}} is added to AkA_{k}, JJ changes to J+ΔJAkJ+\Delta_{J}^{A_{k}}. This causes Q12,ωQ_{12,\omega} and Qk,ωQ_{k,\omega} to perturb to Q12,ω+ΔQ12,ωAkQ_{12,\omega}+\Delta_{Q_{12,\omega}}^{A_{k}} and Qk,ω+ΔQk,ωAkQ_{k,\omega}+\Delta_{Q_{k,\omega}}^{A_{k}}, respectively. Consequently, the first-order terms of ΔJAk\Delta_{J}^{A_{k}} are given by:

ΔJAk=trace(2BTΔQ12,ωAkBk+BkTΔQk,ωAkBk).\displaystyle\Delta_{J}^{A_{k}}=trace(2B^{T}\Delta_{Q_{12,\omega}}^{A_{k}}B_{k}+B_{k}^{T}\Delta_{Q_{k,\omega}}^{A_{k}}B_{k}).

Furthermore, it is evident from (15) and (16) that ΔQ12,ωAk\Delta_{Q_{12,\omega}}^{A_{k}} and ΔQk,ωAk\Delta_{Q_{k,\omega}}^{A_{k}} satisfy the following Lyapunov equations:

ATΔQ12,ωAk+ΔQ12,ωAkAk+Q12,ωΔAk+CTCkΔFk,ωAk+i=1p(MiP12,ωMk,iΔFk,ωAk\displaystyle A^{T}\Delta_{Q_{12,\omega}}^{A_{k}}+\Delta_{Q_{12,\omega}}^{A_{k}}A_{k}+Q_{12,\omega}\Delta_{A_{k}}+C^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}+\sum_{i=1}^{p}\big{(}M_{i}P_{12,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}
+FωMiΔP12,ωAkMk,i+MiΔP12,ωAkMk,iFk,ω)=0,\displaystyle\hskip 128.0374pt+F_{\omega}^{*}M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}+M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}F_{k,\omega}\big{)}=0,
AΔP12,ωAk+ΔP12,ωAkAkT+P12,ω(ΔAk)T+BBkT(ΔFk,ωAk)=0,\displaystyle\hskip 81.09052ptA\Delta_{P_{12,\omega}}^{A_{k}}+\Delta_{P_{12,\omega}}^{A_{k}}A_{k}^{T}+P_{12,\omega}(\Delta_{A_{k}})^{T}+BB_{k}^{T}(\Delta_{F_{k,\omega}}^{A_{k}})^{*}=0,
AkTΔQk,ωAk+ΔQk,ωAkAk+(ΔAk)TQk,ω+Qk,ωΔAk\displaystyle A_{k}^{T}\Delta_{Q_{k,\omega}}^{A_{k}}+\Delta_{Q_{k,\omega}}^{A_{k}}A_{k}+(\Delta_{A_{k}})^{T}Q_{k,\omega}+Q_{k,\omega}\Delta_{A_{k}}
+(ΔFk,ωAk)CkTCk+CkTCkΔFk,ωAk+i=1p((ΔFk,ωAk)Mk,iPk,ωMk,i\displaystyle\hskip 18.49411pt+(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}C_{k}+C_{k}^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}+\sum_{i=1}^{p}\big{(}(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}P_{k,\omega}M_{k,i}
+Mk,iPk,ωMk,iΔFk,ωAk+Fk,ωMk,iΔPk,ωAkMk,i+Mk,iΔPk,ωAkMk,iFk,ω)=0,\displaystyle\hskip 18.49411pt+M_{k,i}P_{k,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}+F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}M_{k,i}+M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}M_{k,i}F_{k,\omega}\big{)}=0,
AkΔPk,ωAk+ΔPk,ωAkAkT+ΔAkPk,ω+Pk,ω(ΔAk)T\displaystyle A_{k}\Delta_{P_{k,\omega}}^{A_{k}}+\Delta_{P_{k,\omega}}^{A_{k}}A_{k}^{T}+\Delta_{A_{k}}P_{k,\omega}+P_{k,\omega}(\Delta_{A_{k}})^{T}
+ΔFk,ωAkBkBkT+BkBkT(ΔFk,ωAk)=0,\displaystyle\hskip 170.71652pt+\Delta_{F_{k,\omega}}^{A_{k}}B_{k}B_{k}^{T}+B_{k}B_{k}^{T}(\Delta_{F_{k,\omega}}^{A_{k}})^{*}=0,

where

ΔFk,ωAk=jπ(jνIAk,ΔAk)+o(ΔAk);\displaystyle\Delta_{F_{k,\omega}}^{A_{k}}=\frac{j}{\pi}\mathcal{L}(-j\nu I-A_{k},\Delta_{A_{k}})+o(||\Delta_{A_{k}}||);

cf. [47]. Since we are only concerned with first-order perturbations, the term o(ΔAk)o(||\Delta_{A_{k}}||) will be omitted for the rest of the proof. Now,

trace(BBkT(ΔQ12,ωAk))\displaystyle trace\Big{(}BB_{k}^{T}(\Delta_{Q_{12,\omega}}^{A_{k}})^{*}\Big{)}
=trace((AP12P12AkT)(ΔQ12,ωAk))\displaystyle=trace\Big{(}(-AP_{12}-P_{12}A_{k}^{T})(\Delta_{Q_{12,\omega}}^{A_{k}})^{*}\Big{)}
=trace(AP12(ΔQ12,ωAk)P12AkT(ΔQ12,ωAk))\displaystyle=trace\Big{(}-AP_{12}(\Delta_{Q_{12,\omega}}^{A_{k}})^{*}-P_{12}A_{k}^{T}(\Delta_{Q_{12,\omega}}^{A_{k}})^{*}\Big{)}
=trace(P12T(ATΔQ12,ωAkAkΔQ12,ωAk))\displaystyle=trace\Big{(}P_{12}^{T}(-A^{T}\Delta_{Q_{12,\omega}}^{A_{k}}-A_{k}\Delta_{Q_{12,\omega}}^{A_{k}})\Big{)}
=trace(P12TQ12,ωΔAk+P12TCTCkΔFk,ωAk+i=1p(P12TMiP12,ωMk,iΔFk,ωAk\displaystyle=trace\Big{(}P_{12}^{T}Q_{12,\omega}\Delta_{A_{k}}+P_{12}^{T}C^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}+\sum_{i=1}^{p}\big{(}P_{12}^{T}M_{i}P_{12,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}
+P12TFωTMiΔP12,ωAkMk,i+Fk,ωP12TMiΔP12,ωAkMk,i))\displaystyle\hskip 56.9055pt+P_{12}^{T}F_{\omega}^{T}M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}+F_{k,\omega}P_{12}^{T}M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}\big{)}\Big{)}
=trace(Q12,ωP12ΔAkT+P12TCTCkΔFk,ωAk+i=1p(P12TMiP12,ωMk,iΔFk,ωAk\displaystyle=trace\Big{(}Q_{12,\omega}^{*}P_{12}\Delta_{A_{k}}^{T}+P_{12}^{T}C^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}+\sum_{i=1}^{p}\big{(}P_{12}^{T}M_{i}P_{12,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}
+P12TFωMiΔP12,ωAkMk,i+Fk,ωP12TMiΔP12,ωAkMk,i)).\displaystyle\hskip 56.9055pt+P_{12}^{T}F_{\omega}^{*}M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}+F_{k,\omega}P_{12}^{T}M_{i}\Delta_{P_{12,\omega}}^{A_{k}}M_{k,i}\big{)}\Big{)}.

Similarly, note that:

trace(BkBkTΔQk,ωAk)\displaystyle trace(B_{k}B_{k}^{T}\Delta_{Q_{k,\omega}}^{A_{k}})
=trace((AkPkPkAkT)ΔQk,ωAk)\displaystyle=trace\Big{(}\big{(}-A_{k}P_{k}-P_{k}A_{k}^{T}\big{)}\Delta_{Q_{k,\omega}}^{A_{k}}\Big{)}
=trace(Pk(AkTΔQk,ωAkΔQk,ωAkAk))\displaystyle=trace\Big{(}P_{k}\big{(}-A_{k}^{T}\Delta_{Q_{k,\omega}}^{A_{k}}-\Delta_{Q_{k,\omega}}^{A_{k}}A_{k}\big{)}\Big{)}
=trace(Pk((ΔAk)TQk,ω+Qk,ωΔAk+(ΔFk,ωAk)CkTCk+CkTCkΔFk,ωAk\displaystyle=trace\Bigg{(}P_{k}\Big{(}(\Delta_{A_{k}})^{T}Q_{k,\omega}+Q_{k,\omega}\Delta_{A_{k}}+(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}C_{k}+C_{k}^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}
+i=1p((ΔFk,ωAk)Mk,iPk,ωMk,i+Mk,iPk,ωMk,iΔFk,ωAk\displaystyle\hskip 56.9055pt+\sum_{i=1}^{p}\big{(}(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}P_{k,\omega}M_{k,i}+M_{k,i}P_{k,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}
+Fk,ωMk,iΔPk,ωAkMk,i+Mk,iΔPk,ωAkMk,iFk,ω)))\displaystyle\hskip 56.9055pt+F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}M_{k,i}+M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}M_{k,i}F_{k,\omega}\big{)}\Big{)}\Bigg{)}
=trace(2Qk,ωPk(ΔAk)T+2(ΔFk,ωAk)CkTCkPk\displaystyle=trace\Big{(}2Q_{k,\omega}P_{k}(\Delta_{A_{k}})^{T}+2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}C_{k}P_{k}
+i=1p(2(ΔFk,ωAk)Mk,iPk,ωMk,iPk+Mk,iFk,ωPkMkΔPk,ωAk\displaystyle\hskip 56.9055pt+\sum_{i=1}^{p}\big{(}2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}P_{k,\omega}M_{k,i}P_{k}+M_{k,i}F_{k,\omega}P_{k}M_{k}\Delta_{P_{k,\omega}}^{A_{k}}
+Mk,iPkFk,ωMk,iΔPk,ωAk)).\displaystyle\hskip 56.9055pt+M_{k,i}P_{k}F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}\big{)}\Big{)}.

Therefore:

ΔJAk=trace(2(Q12,ω)P12(ΔAk)T+2Qk,ωPk(ΔAk)T\displaystyle\Delta_{J}^{A_{k}}=trace\Big{(}-2(Q_{12,\omega})^{*}P_{12}(\Delta_{A_{k}})^{T}+2Q_{k,\omega}P_{k}(\Delta_{A_{k}})^{T}
2(ΔFk,ωAk)CkTCP12+2(ΔFk,ωAk)CkTCkPk\displaystyle\hskip 56.9055pt-2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}CP_{12}+2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}C_{k}P_{k}
+i=1p(2MiFωP12Mk,ω(ΔP12,ωAk)2MiP12Fk,ωMk,i(ΔP12,ωAk)\displaystyle\hskip 56.9055pt+\sum_{i=1}^{p}\big{(}-2M_{i}F_{\omega}P_{12}M_{k,\omega}(\Delta_{P_{12,\omega}}^{A_{k}})^{*}-2M_{i}P_{12}F_{k,\omega}^{*}M_{k,i}(\Delta_{P_{12,\omega}}^{A_{k}})^{*}
+Mk,iFk,ωPkMk,iΔPk,ωAk+Mk,iPkFk,ωMk,iΔPk,ωAk\displaystyle\hskip 56.9055pt+M_{k,i}F_{k,\omega}P_{k}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}+M_{k,i}P_{k}F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}
2(ΔFk,ωAk)Mk,i(P12,ω)MiP12+2(ΔFk,ωAk)Mk,iPk,ωMk,iPk)).\displaystyle\hskip 56.9055pt-2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}(P_{12,\omega})^{*}M_{i}P_{12}+2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}P_{k,\omega}M_{k,i}P_{k}\big{)}\Big{)}.

Since

P12,ω=FωP12+P12Fk,ω,\displaystyle P_{12,\omega}=F_{\omega}P_{12}+P_{12}F_{k,\omega}^{*},
Pk,ω=Fk,ωPk+PkFk,ω,\displaystyle P_{k,\omega}=F_{k,\omega}P_{k}+P_{k}F_{k,\omega}^{*},

we have:

ΔJAk=trace(2(Q12,ω)P12(ΔAk)T+2Qk,ωPk(ΔAk)T\displaystyle\Delta_{J}^{A_{k}}=trace\Big{(}-2(Q_{12,\omega})^{*}P_{12}(\Delta_{A_{k}})^{T}+2Q_{k,\omega}P_{k}(\Delta_{A_{k}})^{T}
2(ΔFk,ωAk)CkTCP12+2(ΔFk,ωAk)CkTCkPk\displaystyle\hskip 28.45274pt-2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}CP_{12}+2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}C_{k}^{T}C_{k}P_{k}
+i=1p(2MiP12,ωMk,i(ΔP12,ωAk)+Mk,iPk,ωMk,iΔPk,ωAk\displaystyle\hskip 28.45274pt+\sum_{i=1}^{p}\big{(}-2M_{i}P_{12,\omega}M_{k,i}(\Delta_{P_{12,\omega}}^{A_{k}})^{*}+M_{k,i}P_{k,\omega}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}
2(ΔFk,ωAk)Mk,i(P12,ω)MiP12+2(ΔFk,ωAk)Mk,iPk,ωMk,iPk));\displaystyle\hskip 28.45274pt-2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}(P_{12,\omega})^{*}M_{i}P_{12}+2(\Delta_{F_{k,\omega}}^{A_{k}})^{*}M_{k,i}P_{k,\omega}M_{k,i}P_{k}\big{)}\Big{)};

cf. [50]. Note that

trace(i=1pMiP12,ωMk,i(ΔP12,ωAk))\displaystyle trace\Big{(}\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}(\Delta_{P_{12,\omega}}^{A_{k}})^{*}\Big{)}
=trace((ATZ¯12Z¯12Ak)(ΔP12,ωAk))\displaystyle=trace\Big{(}\big{(}-A^{T}\bar{Z}_{12}-\bar{Z}_{12}A_{k}\big{)}(\Delta_{P_{12,\omega}}^{A_{k}})^{*}\Big{)}
=trace((AΔP12,ωAkΔP12,ωAkAkT)Z¯12)\displaystyle=trace\Big{(}\big{(}-A\Delta_{P_{12,\omega}}^{A_{k}}-\Delta_{P_{12,\omega}}^{A_{k}}A_{k}^{T}\big{)}\bar{Z}_{12}^{*}\Big{)}
=trace(Z¯12P12,ω(ΔAk)T+Z¯12BBkT(ΔFk,ωAk)T).\displaystyle=trace\big{(}\bar{Z}_{12}^{*}P_{12,\omega}(\Delta_{A_{k}})^{T}+\bar{Z}_{12}^{*}BB_{k}^{T}(\Delta_{F_{k,\omega}}^{A_{k}})^{T}\big{)}.

Additionally, note that

trace(i=1pMk,iPk,ωMk,iΔPk,ωAk)\displaystyle trace\Big{(}\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i}\Delta_{P_{k,\omega}}^{A_{k}}\Big{)}
=trace((AkTZ¯kZ¯kAk)ΔPk,ωAk)\displaystyle=trace\Big{(}\big{(}-A_{k}^{T}\bar{Z}_{k}-\bar{Z}_{k}A_{k}\big{)}\Delta_{P_{k,\omega}}^{A_{k}}\Big{)}
=trace((AkΔPk,ωAkΔPk,ωAkAkT)Z¯k)\displaystyle=trace\Big{(}\big{(}-A_{k}\Delta_{P_{k,\omega}}^{A_{k}}-\Delta_{P_{k,\omega}}^{A_{k}}A_{k}^{T}\big{)}\bar{Z}_{k}\Big{)}
=2trace(Z¯kPk,ω(ΔAk)T+Z¯kBkBkT(ΔFk,ωAk)T).\displaystyle=2trace\big{(}\bar{Z}_{k}P_{k,\omega}(\Delta_{A_{k}})^{T}+\bar{Z}_{k}B_{k}B_{k}^{T}(\Delta_{F_{k,\omega}}^{A_{k}})^{T}\big{)}.

Thus,

ΔJAk\displaystyle\Delta_{J}^{A_{k}} =trace(2(Q12,ω)P12(ΔAk)T+2Qk,ωPk(ΔAk)T\displaystyle=trace\Big{(}-2(Q_{12,\omega})^{*}P_{12}(\Delta_{A_{k}})^{T}+2Q_{k,\omega}P_{k}(\Delta_{A_{k}})^{T}
2Z¯12P12,ω(ΔAk)T+2Z¯kPk,ω(ΔAk)T\displaystyle\hskip 56.9055pt-2\bar{Z}_{12}^{*}P_{12,\omega}(\Delta_{A_{k}})^{T}+2\bar{Z}_{k}P_{k,\omega}(\Delta_{A_{k}})^{T}
2BkBTZ¯12ΔFk,ωAk+2BkBkTZ¯kΔFk,ωAk\displaystyle\hskip 56.9055pt-2B_{k}B^{T}\bar{Z}_{12}\Delta_{F_{k,\omega}}^{A_{k}}+2B_{k}B_{k}^{T}\bar{Z}_{k}\Delta_{F_{k,\omega}}^{A_{k}}
2P12TCTCkΔFk,ωAk+2PkCkCkTΔFk,ωAk\displaystyle\hskip 56.9055pt-2P_{12}^{T}C^{T}C_{k}\Delta_{F_{k,\omega}}^{A_{k}}+2P_{k}C_{k}C_{k}^{T}\Delta_{F_{k,\omega}}^{A_{k}}
2i=1pP12TMiP12,ωMk,iΔFk,ωAk\displaystyle\hskip 56.9055pt-2\sum_{i=1}^{p}P_{12}^{T}M_{i}P_{12,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}
+2i=1pPkMk,iPk,ωMk,iΔFk,ωAk)\displaystyle\hskip 56.9055pt+2\sum_{i=1}^{p}P_{k}M_{k,i}P_{k,\omega}M_{k,i}\Delta_{F_{k,\omega}}^{A_{k}}\Big{)}
=trace(2(Q12,ω)P12(ΔAk)T+2Qk,ωPk(ΔAk)T\displaystyle=trace\Big{(}-2(Q_{12,\omega})^{*}P_{12}(\Delta_{A_{k}})^{T}+2Q_{k,\omega}P_{k}(\Delta_{A_{k}})^{T}
2Z¯12P12,ω(ΔAk)T+2Z¯kPk,ω(ΔAk)T2VΔFk,ωAk).\displaystyle\hskip 56.9055pt-2\bar{Z}_{12}^{*}P_{12,\omega}(\Delta_{A_{k}})^{T}+2\bar{Z}_{k}P_{k,\omega}(\Delta_{A_{k}})^{T}-2V\Delta_{F_{k,\omega}}^{A_{k}}\Big{)}.

Recall that

ΔFk,ωAk=j2π(AkjνI,ΔAk).\displaystyle\Delta_{F_{k,\omega}}^{A_{k}}=\frac{j}{2\pi}\mathcal{L}(-A_{k}-j\nu I,-\Delta_{A_{k}}).

By exchanging the trace and integral operations, we obtain

trace(VΔFk,ωAk)=trace(WΔAk);\displaystyle trace(V\Delta_{F_{k,\omega}}^{A_{k}})=-trace(W\Delta_{A_{k}});

cf. [48]. Hence,

ΔJAk=2trace((Q12,ωP12+Qk,ωPkZ¯12P12,ω+Z¯kPk,ω+W)(ΔAk)T).\displaystyle\Delta_{J}^{A_{k}}=2trace\Big{(}\big{(}-Q_{12,\omega}^{*}P_{12}+Q_{k,\omega}P_{k}-\bar{Z}_{12}^{*}P_{12,\omega}+\bar{Z}_{k}P_{k,\omega}+W^{*}\big{)}(\Delta_{A_{k}})^{T}\Big{)}.

Therefore, the gradient of JJ with respect of AkA_{k} is given by

JAk=2(Q12,ωP12+Qk,ωPkZ¯12P12,ω+Z¯kPk,ω+W);\displaystyle\nabla_{J}^{A_{k}}=2\big{(}-Q_{12,\omega}^{*}P_{12}+Q_{k,\omega}P_{k}-\bar{Z}_{12}^{*}P_{12,\omega}+\bar{Z}_{k}P_{k,\omega}+W^{*}\big{)};

cf. [47]. Consequently,

Q12,ωP12+Qk,ωPkZ¯12P12,ω+Z¯kPk,ω+W=0\displaystyle-Q_{12,\omega}^{*}P_{12}+Q_{k,\omega}P_{k}-\bar{Z}_{12}^{*}P_{12,\omega}+\bar{Z}_{k}P_{k,\omega}+W^{*}=0 (34)

is a necessary condition for a local optimum of EH2,ω2||E||_{H_{2,\omega}}^{2}. Moreover, substituting (17)-(20) into (34), it simplifies to

Q12,ωP12,ω+Qk,ωPk,ωZ12,ωP12,ω+Zk,ωPk,ω+Lω=0.\displaystyle-Q_{12,\omega}^{*}P_{12,\omega}+Q_{k,\omega}P_{k,\omega}-Z_{12,\omega}^{*}P_{12,\omega}+Z_{k,\omega}P_{k,\omega}+L_{\omega}=0.

Additionally, since Q12,ω=Y12,ω+Z12,ωQ_{12,\omega}=Y_{12,\omega}+Z_{12,\omega} and Qk,ω=Yk,ω+Zk,ωQ_{k,\omega}=Y_{k,\omega}+Z_{k,\omega}, we arrive at

(Y12,ω+2Z12,ω)P12,ω+(Yk,ω+2Zk,ω)Pk,ω+Lω\displaystyle-(Y_{12,\omega}+2Z_{12,\omega})^{*}P_{12,\omega}+(Y_{k,\omega}+2Z_{k,\omega})P_{k,\omega}+L_{\omega} =0.\displaystyle=0.

By introducing a small first-order perturbation ΔMk,i\Delta_{M_{k,i}} to Mk,iM_{k,i}, JJ is perturbed to J+ΔJMk,iJ+\Delta_{J}^{M_{k,i}}. Consequently, Q12,ωQ_{12,\omega} and Qk,ωQ_{k,\omega} are perturbed to Q12,ω+ΔQ12,ωMk,iQ_{12,\omega}+\Delta_{Q_{12,\omega}}^{M_{k,i}} and Qk,ω+ΔQk,ωMk,iQ_{k,\omega}+\Delta_{Q_{k,\omega}}^{M_{k,i}}, respectively. As a result, the first-order terms of ΔJMk,i\Delta_{J}^{M_{k,i}} are given by

ΔJMk,i=trace(2BTΔQ12,ωMk,iBk+BkTΔQk,ωMk,iBk).\displaystyle\Delta_{J}^{M_{k,i}}=trace(-2B^{T}\Delta_{Q_{12,\omega}}^{M_{k,i}}B_{k}+B_{k}^{T}\Delta_{Q_{k,\omega}}^{M_{k,i}}B_{k}).

Furthermore, it can be easily observed from (15) and (16) that ΔQ12,ωMk,i\Delta_{Q_{12,\omega}}^{M_{k,i}} and ΔQk,ωMk,i\Delta_{Q_{k,\omega}}^{M_{k,i}} satisfy the following Lyapunov equations:

ATΔQ12,ωMk,i+ΔQ12,ωMk,iAk+FωMiP12,ωΔMk,i+MiP12,ωΔMk,iFk,ω=0,\displaystyle A^{T}\Delta_{Q_{12,\omega}}^{M_{k,i}}+\Delta_{Q_{12,\omega}}^{M_{k,i}}A_{k}+F_{\omega}^{*}M_{i}P_{12,\omega}\Delta_{M_{k,i}}+M_{i}P_{12,\omega}\Delta_{M_{k,i}}F_{k,\omega}=0,
AkTΔQk,ωMk,i+ΔQk,ωMk,iAk+Fk,ωΔMk,iPk,ωMk,i+ΔMk,iPk,ωMk,iFk,ω\displaystyle A_{k}^{T}\Delta_{Q_{k,\omega}}^{M_{k,i}}+\Delta_{Q_{k,\omega}}^{M_{k,i}}A_{k}+F_{k,\omega}^{*}\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}+\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}F_{k,\omega}
+Fk,ωMk,iPk,ωΔMk,i+Mk,iPk,ωΔMk,iFk,ω=0.\displaystyle\hskip 85.35826pt+F_{k,\omega}^{*}M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}+M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}F_{k,\omega}=0.

Observe that

trace(BTΔQ12,ωMk,iBk)\displaystyle trace(B^{T}\Delta_{Q_{12,\omega}}^{M_{k,i}}B_{k})
=trace(BBkT(ΔQ12,ωMk,i))\displaystyle=trace\big{(}BB_{k}^{T}(\Delta_{Q_{12,\omega}}^{M_{k,i}})^{*}\big{)}
=trace((AP12P12AkT)(ΔQ12,ωMk,i))\displaystyle=trace\Big{(}\big{(}-AP_{12}-P_{12}A_{k}^{T}\big{)}(\Delta_{Q_{12,\omega}}^{M_{k,i}})^{*}\Big{)}
=trace((ATΔQ12,ωMk,iΔQ12,ωMk,iAk)P12T)\displaystyle=trace\Big{(}\big{(}-A^{T}\Delta_{Q_{12,\omega}}^{M_{k,i}}-\Delta_{Q_{12,\omega}}^{M_{k,i}}A_{k}\big{)}P_{12}^{T}\Big{)}
=trace((FωMiP12,ωΔMk,i+MiP12,ωΔMk,iFk,ω)P12T)\displaystyle=trace\Big{(}\big{(}F_{\omega}^{*}M_{i}P_{12,\omega}\Delta_{M_{k,i}}+M_{i}P_{12,\omega}\Delta_{M_{k,i}}F_{k,\omega}\big{)}P_{12}^{T}\Big{)}
=trace(P12TFωMiP12,ωΔMk,i+Fk,ωP12TMiP12,ωΔMk,i)\displaystyle=trace\big{(}P_{12}^{T}F_{\omega}^{*}M_{i}P_{12,\omega}\Delta_{M_{k,i}}+F_{k,\omega}P_{12}^{T}M_{i}P_{12,\omega}\Delta_{M_{k,i}}\big{)}
=trace((P12,ω)MiFωP12(ΔMk,i)T+(P12,ω)MiP12Fk,ω(ΔMk,i)T)\displaystyle=trace\big{(}(P_{12,\omega})^{*}M_{i}F_{\omega}P_{12}(\Delta_{M_{k,i}})^{T}+(P_{12,\omega})^{*}M_{i}P_{12}F_{k,\omega}^{*}(\Delta_{M_{k,i}})^{T}\big{)}
=trace((P12,ω)MiP12,ω(ΔMk,i)T).\displaystyle=trace\big{(}(P_{12,\omega})^{*}M_{i}P_{12,\omega}(\Delta_{M_{k,i}})^{T}\big{)}.

Additionally, note that

trace(BkTΔQk,ωMk,iBk)\displaystyle trace(B_{k}^{T}\Delta_{Q_{k,\omega}}^{M_{k,i}}B_{k}) =trace(BkBkTΔQk,ωMk,i)\displaystyle=trace\big{(}B_{k}B_{k}^{T}\Delta_{Q_{k,\omega}}^{M_{k,i}}\big{)}
=trace((AkPkPkAkT)ΔQk,ωMk,i)\displaystyle=trace\Big{(}\big{(}-A_{k}P_{k}-P_{k}A_{k}^{T}\big{)}\Delta_{Q_{k,\omega}}^{M_{k,i}}\Big{)}
=trace((AkTΔQk,ωMk,iΔQk,ωMk,iAk)Pk)\displaystyle=trace\Big{(}\big{(}-A_{k}^{T}\Delta_{Q_{k,\omega}}^{M_{k,i}}-\Delta_{Q_{k,\omega}}^{M_{k,i}}A_{k}\big{)}P_{k}\Big{)}
=trace((Fk,ωΔMk,iPk,ωMk,i+ΔMk,iPk,ωMk,iFk,ω\displaystyle=trace\Big{(}\big{(}F_{k,\omega}^{*}\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}+\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}F_{k,\omega}
+Fk,ωMk,iPk,ωΔMk,i+Mk,iPk,ωΔMk,iFk,ω)Pk)\displaystyle+F_{k,\omega}^{*}M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}+M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}F_{k,\omega}\big{)}P_{k}\Big{)}
=trace(Fk,ωΔMk,iPk,ωMk,iPk+ΔMk,iPk,ωMk,iFk,ωPk\displaystyle=trace\big{(}F_{k,\omega}^{*}\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}P_{k}+\Delta_{M_{k,i}}P_{k,\omega}M_{k,i}F_{k,\omega}P_{k}
+Fk,ωMk,iPk,ωΔMk,iPk+Mk,iPk,ωΔMk,iFk,ωPk)\displaystyle+F_{k,\omega}^{*}M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}P_{k}+M_{k,i}P_{k,\omega}\Delta_{M_{k,i}}F_{k,\omega}P_{k}\big{)}
=trace(2Pk,ωMk,iPk,ω).\displaystyle=trace\big{(}2P_{k,\omega}M_{k,i}P_{k,\omega}\big{)}.

Thus, ΔJMk,i\Delta_{J}^{M_{k,i}} becomes

ΔJMk,i=2trace((P12,ωMiP12,ω+Pk,ωMk,iPk,ω)(ΔMk,i)T).\displaystyle\Delta_{J}^{M_{k,i}}=2trace\big{(}(-P_{12,\omega}^{*}M_{i}P_{12,\omega}+P_{k,\omega}M_{k,i}P_{k,\omega})(\Delta_{M_{k,i}})^{T}\big{)}.

Hence, the gradient of JJ with respect to Mk,iM_{k,i} is given by

JMk,i=2(P12,ωMiP12,ω+Pk,ωMk,iPk,ω).\displaystyle\nabla_{J}^{M_{k,i}}=2(-P_{12,\omega}^{*}M_{i}P_{12,\omega}+P_{k,\omega}M_{k,i}P_{k,\omega}).

Therefore,

P12,ωMiP12,ω+Pk,ωMk,iPk,ω=0\displaystyle-P_{12,\omega}^{*}M_{i}P_{12,\omega}+P_{k,\omega}M_{k,i}P_{k,\omega}=0

is a necessary condition for the local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}.

By introducing a small first-order perturbation ΔBk\Delta_{B_{k}} to BkB_{k}, JJ is perturbed to J+ΔJBkJ+\Delta_{J}^{B_{k}}. Consequently, P12,ωP_{12,\omega}, Pk,ωP_{k,\omega}, Q12,ωQ_{12,\omega} and Qk,ωQ_{k,\omega} are perturbed to P12,ω+ΔP12,ωBkP_{12,\omega}+\Delta_{P_{12,\omega}}^{B_{k}}, Pk,ω+ΔPk,ωBkP_{k,\omega}+\Delta_{P_{k,\omega}}^{B_{k}}, Q12,ω+ΔQ12,ωBkQ_{12,\omega}+\Delta_{Q_{12,\omega}}^{B_{k}}, and Qk,ω+ΔQk,ωBkQ_{k,\omega}+\Delta_{Q_{k,\omega}}^{B_{k}}, respectively. As a result, the first-order terms of ΔJBk\Delta_{J}^{B_{k}} are given by

ΔJBk\displaystyle\Delta_{J}^{B_{k}} =trace(2Q12,ωB(ΔBk)T+2Qk,ωBk(ΔBk)T\displaystyle=trace\big{(}-2Q_{12,\omega}^{*}B(\Delta_{B_{k}})^{T}+2Q_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}
2BBk(ΔQ12,ωBk)+BkBkTΔQk,ωBk).\displaystyle\hskip 85.35826pt-2BB_{k}(\Delta_{Q_{12,\omega}}^{B_{k}})^{*}+B_{k}B_{k}^{T}\Delta_{Q_{k,\omega}}^{B_{k}}\big{)}.

It follows from (9)-(16) that ΔP12,ωBk\Delta_{P_{12,\omega}}^{B_{k}}, ΔPk,ωBk\Delta_{P_{k,\omega}}^{B_{k}}, ΔQ12,ωBk\Delta_{Q_{12,\omega}}^{B_{k}}, and ΔQk,ωBk\Delta_{Q_{k,\omega}}^{B_{k}} satisfy the following equations:

AΔP12,ωBk+ΔP12,ωBkAkT+FωB(ΔBk)T+B(ΔBk)TFk,ω=0,\displaystyle A\Delta_{P_{12,\omega}}^{B_{k}}+\Delta_{P_{12,\omega}}^{B_{k}}A_{k}^{T}+F_{\omega}B(\Delta_{B_{k}})^{T}+B(\Delta_{B_{k}})^{T}F_{k,\omega}^{*}=0,
AkΔPk,ωBk+ΔPk,ωBkAkT+Fk,ωΔBkBkT+ΔBkBkTFk,ω\displaystyle A_{k}\Delta_{P_{k,\omega}}^{B_{k}}+\Delta_{P_{k,\omega}}^{B_{k}}A_{k}^{T}+F_{k,\omega}\Delta_{B_{k}}B_{k}^{T}+\Delta_{B_{k}}B_{k}^{T}F_{k,\omega}^{*}
+Fk,ωBk(ΔBk)T+Bk(ΔBk)TFk,ω=0\displaystyle\hskip 85.35826pt+F_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}+B_{k}(\Delta_{B_{k}})^{T}F_{k,\omega}^{*}=0
ATΔQ12,ωBk+ΔQ12,ωBkAk+Fωi=1pMiΔP12,ωBkMk,i+i=1pMiΔP12,ωBkMk,iFk,ω=0\displaystyle A^{T}\Delta_{Q_{12,\omega}}^{B_{k}}+\Delta_{Q_{12,\omega}}^{B_{k}}A_{k}+F_{\omega}^{*}\sum_{i=1}^{p}M_{i}\Delta_{P_{12,\omega}}^{B_{k}}M_{k,i}+\sum_{i=1}^{p}M_{i}\Delta_{P_{12,\omega}}^{B_{k}}M_{k,i}F_{k,\omega}=0
AkTΔQk,ωBk+ΔQk,ωBkAk+Fk,ωi=1pMk,iΔPk,ωBkMk,i+i=1pMk,iΔPk,ωBkMk,iFk,ω=0.\displaystyle A_{k}^{T}\Delta_{Q_{k,\omega}}^{B_{k}}+\Delta_{Q_{k,\omega}}^{B_{k}}A_{k}+F_{k,\omega}^{*}\sum_{i=1}^{p}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}M_{k,i}+\sum_{i=1}^{p}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}M_{k,i}F_{k,\omega}=0.

Note that

trace(BBkT(ΔQ12,ωBk))\displaystyle trace\big{(}BB_{k}^{T}(\Delta_{Q_{12,\omega}}^{B_{k}})^{*}\big{)}
=trace((AP12P12AkT)(ΔQ12,ωBk))\displaystyle=trace\Big{(}\big{(}-AP_{12}-P_{12}A_{k}^{T}\big{)}(\Delta_{Q_{12,\omega}}^{B_{k}})^{*}\Big{)}
=trace((ATΔQ12,ωBkΔQ12,ωBkAk)P12T)\displaystyle=trace\Big{(}\big{(}-A^{T}\Delta_{Q_{12,\omega}}^{B_{k}}-\Delta_{Q_{12,\omega}}^{B_{k}}A_{k}\big{)}P_{12}^{T}\Big{)}
=trace((Fωi=1pMiΔP12,ωBkMk,i+i=1pMiΔP12,ωBkMk,iFk,ω)P12T)\displaystyle=trace\Big{(}\big{(}F_{\omega}^{*}\sum_{i=1}^{p}M_{i}\Delta_{P_{12,\omega}}^{B_{k}}M_{k,i}+\sum_{i=1}^{p}M_{i}\Delta_{P_{12,\omega}}^{B_{k}}M_{k,i}F_{k,\omega}\big{)}P_{12}^{T}\Big{)}
=trace(i=1pMk,iP12TFωMiΔP12,ωBk+i=1pMk,iFk,ωP12TMiΔP12,ωBk)\displaystyle=trace\Big{(}\sum_{i=1}^{p}M_{k,i}P_{12}^{T}F_{\omega}^{*}M_{i}\Delta_{P_{12,\omega}}^{B_{k}}+\sum_{i=1}^{p}M_{k,i}F_{k,\omega}P_{12}^{T}M_{i}\Delta_{P_{12,\omega}^{B_{k}}}\Big{)}
=trace(i=1pMiP12,ωMk,i(ΔP12,ωBk)).\displaystyle=trace\Big{(}\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}(\Delta_{P_{12,\omega}}^{B_{k}})^{*}\Big{)}.

Furthermore,

trace(BkBkTΔQk,ωBk)\displaystyle trace\big{(}B_{k}B_{k}^{T}\Delta_{Q_{k,\omega}}^{B_{k}}\big{)}
=trace((AkPkPkAkT)ΔQk,ωBk)\displaystyle=trace\Big{(}\big{(}-A_{k}P_{k}-P_{k}A_{k}^{T}\big{)}\Delta_{Q_{k,\omega}}^{B_{k}}\Big{)}
=trace((AkTΔQk,ωBkΔQk,ωBkAk)Pk)\displaystyle=trace\Big{(}\big{(}-A_{k}^{T}\Delta_{Q_{k,\omega}}^{B_{k}}-\Delta_{Q_{k,\omega}}^{B_{k}}A_{k}\big{)}P_{k}\Big{)}
=trace((i=1pFk,ωMk,iΔPk,ωBkMk,i+i=1pMk,iΔPk,ωBkMk,iFk,ω)Pk)\displaystyle=trace\Big{(}\big{(}\sum_{i=1}^{p}F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}M_{k,i}+\sum_{i=1}^{p}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}M_{k,i}F_{k,\omega}\big{)}P_{k}\Big{)}
=trace(i=1pMk,iPkFk,ωMk,iΔPk,ωBk+i=1pMk,iFk,ωPkMk,iΔPk,ωBk)\displaystyle=trace\Big{(}\sum_{i=1}^{p}M_{k,i}P_{k}F_{k,\omega}^{*}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}+\sum_{i=1}^{p}M_{k,i}F_{k,\omega}P_{k}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}\Big{)}
=trace(i=1pMk,iPk,ωMk,iΔPk,ωBk).\displaystyle=trace\Big{(}\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}\Big{)}.

Thus,

ΔJBk\displaystyle\Delta_{J}^{B_{k}} =trace(2Q12,ωB(ΔBk)T+2Qk,ωBk(ΔBk)T\displaystyle=trace\Big{(}-2Q_{12,\omega}^{*}B(\Delta_{B_{k}})^{T}+2Q_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}
2i=1pMiP12,ωMk,i(ΔP12,ωBk)T+i=1pMk,iPk,ωMk,iΔPk,ωBk).\displaystyle\hskip 56.9055pt-2\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}(\Delta_{P_{12,\omega}}^{B_{k}})^{T}+\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}\Big{)}.

Additionally,

trace(i=1pMiP12,ωMk,i(ΔP12,ωBk))\displaystyle trace\Big{(}\sum_{i=1}^{p}M_{i}P_{12,\omega}M_{k,i}(\Delta_{P_{12,\omega}}^{B_{k}})^{*}\Big{)}
=trace((ATZ¯12Z¯12Ak)(ΔP12,ωBk))\displaystyle=trace\Big{(}\big{(}-A^{T}\bar{Z}_{12}-\bar{Z}_{12}A_{k}\big{)}(\Delta_{P_{12,\omega}}^{B_{k}})^{*}\Big{)}
=trace((AΔP12,ωBkΔP12,ωBkAkT)Z¯12)\displaystyle=trace\Big{(}\big{(}-A\Delta_{P_{12,\omega}}^{B_{k}}-\Delta_{P_{12,\omega}}^{B_{k}}A_{k}^{T}\big{)}\bar{Z}_{12}^{*}\Big{)}
=trace((FωB(ΔBk)T+B(ΔBk)TFk,ω)Z¯12)\displaystyle=trace\Big{(}\big{(}F_{\omega}B(\Delta_{B_{k}})^{T}+B(\Delta_{B_{k}})^{T}F_{k,\omega}^{*}\big{)}\bar{Z}_{12}^{*}\Big{)}
=trace(Z¯12FωB(ΔBk)T+Fk,ωZ¯12B(ΔBk)T)\displaystyle=trace\big{(}\bar{Z}_{12}^{*}F_{\omega}B(\Delta_{B_{k}})^{T}+F_{k,\omega}^{*}\bar{Z}_{12}^{*}B(\Delta_{B_{k}})^{T}\big{)}
=trace(Z12,ωB(ΔBk)T).\displaystyle=trace\big{(}Z_{12,\omega}^{*}B(\Delta_{B_{k}})^{T}\big{)}.

Similarly,

trace(i=1pMk,iPk,ωMk,iΔPk,ωBk)\displaystyle trace\Big{(}\sum_{i=1}^{p}M_{k,i}P_{k,\omega}M_{k,i}\Delta_{P_{k,\omega}}^{B_{k}}\Big{)}
=trace((AkTZ¯kZ¯kAk)ΔPk,ωBk)\displaystyle=trace\Big{(}\big{(}-A_{k}^{T}\bar{Z}_{k}-\bar{Z}_{k}A_{k}\big{)}\Delta_{P_{k,\omega}}^{B_{k}}\Big{)}
=trace((AkΔPk,ωBkΔPk,ωBkAkT)Z¯k)\displaystyle=trace\Big{(}\big{(}-A_{k}\Delta_{P_{k,\omega}}^{B_{k}}-\Delta_{P_{k,\omega}}^{B_{k}}A_{k}^{T}\big{)}\bar{Z}_{k}\Big{)}
=trace(2Z¯kFk,ωBk(ΔBk)T+2Fk,ωZ¯kBk(ΔBk)T)\displaystyle=trace\Big{(}2\bar{Z}_{k}F_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}+2F_{k,\omega}^{*}\bar{Z}_{k}B_{k}(\Delta_{B_{k}})^{T}\Big{)}
=2trace(Zk,ωBk(ΔBk)T).\displaystyle=2trace\big{(}Z_{k,\omega}^{*}B_{k}(\Delta_{B_{k}})^{T}\big{)}.

Thus,

ΔJBk\displaystyle\Delta_{J}^{B_{k}} =trace(2Q12,ωB(ΔBk)T+2Qk,ωBk(ΔBk)T\displaystyle=trace\big{(}-2Q_{12,\omega}^{*}B(\Delta_{B_{k}})^{T}+2Q_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}
2Z12,ωB(ΔBk)T+2Zk,ωBk(ΔBk)T).\displaystyle\hskip 56.9055pt-2Z_{12,\omega}^{*}B(\Delta_{B_{k}})^{T}+2Z_{k,\omega}B_{k}(\Delta_{B_{k}})^{T}\big{)}.

Therefore, the gradient of JJ with respect to BkB_{k} is

JBk=2(Q12,ωB+Qk,ωBkZ12,ωB+Zk,ωBk).\displaystyle\nabla_{J}^{B_{k}}=2(-Q_{12,\omega}^{*}B+Q_{k,\omega}B_{k}-Z_{12,\omega}^{*}B+Z_{k,\omega}B_{k}).

Hence,

Q12,ωB+Qk,ωBkZ12,ωB+Zk,ωBk\displaystyle-Q_{12,\omega}^{*}B+Q_{k,\omega}B_{k}-Z_{12,\omega}^{*}B+Z_{k,\omega}B_{k}
=(Y12,ω+2Z12,ω)B+(Yk,ω+2Zk,ω)Bk\displaystyle=-(Y_{12,\omega}+2Z_{12,\omega})^{*}B+(Y_{k,\omega}+2Z_{k,\omega})B_{k}
=0\displaystyle=0

is a necessary condition for the local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2}.

First, we reformat the cost function JJ as follows:

J\displaystyle J =trace(2BT(Y12,ω+Z12,ω)Bk+BkT(Yk,ω+Zk,ω)Bk)\displaystyle=trace(-2B^{T}(Y_{12,\omega}+Z_{12,\omega})B_{k}+B_{k}^{T}(Y_{k,\omega}+Z_{k,\omega})B_{k})
=trace(2BTY12,ωBk2BTZ12,ωBk+BkTYk,ωBk+BkTZk,ωBk).\displaystyle=trace(-2B^{T}Y_{12,\omega}B_{k}-2B^{T}Z_{12,\omega}B_{k}+B_{k}^{T}Y_{k,\omega}B_{k}+B_{k}^{T}Z_{k,\omega}B_{k}).

Since

trace(2BTY12,ωBk+BkTYk,ωBk)=trace(2CP12,ωCkT+CkPk,ωCkT),\displaystyle trace(-2B^{T}Y_{12,\omega}B_{k}+B_{k}^{T}Y_{k,\omega}B_{k})=trace(2CP_{12,\omega}C_{k}^{T}+C_{k}P_{k,\omega}C_{k}^{T}),

we can write

J=trace(2CP12,ωCkT+CkPk,ωCkT2BTZ12,ωBk+BkTZk,ωBk);\displaystyle J=trace(2CP_{12,\omega}C_{k}^{T}+C_{k}P_{k,\omega}C_{k}^{T}-2B^{T}Z_{12,\omega}B_{k}+B_{k}^{T}Z_{k,\omega}B_{k});

cf. [50]. By adding a small first-order perturbation ΔCk\Delta_{C_{k}} to CkC_{k}, the cost function JJ is perturbed to J+ΔJCkJ+\Delta_{J}^{C_{k}}. The first-order terms of ΔJCk\Delta_{J}^{C_{k}} are given by:

ΔJCk=trace(2CP12,ω(ΔCk)T+2CkPk,ω(ΔCk)T)\displaystyle\Delta_{J}^{C_{k}}=trace(2CP_{12,\omega}(\Delta_{C_{k}})^{T}+2C_{k}P_{k,\omega}(\Delta_{C_{k}})^{T})

Thus, the gradient of JJ with respect to CkC_{k} is:

JCk=2(CP12,ω+CkPk,ω).\displaystyle\nabla_{J}^{C_{k}}=2(CP_{12,\omega}+C_{k}P_{k,\omega}).

Therefore, the necessary condition for the local optimum of E2,ω2||E||_{\mathcal{H}_{2,\omega}}^{2} is:

CP12,ω+CkPk,ω=0.\displaystyle CP_{12,\omega}+C_{k}P_{k,\omega}=0.

This concludes the proof.

Appendix B

By pre-multiplying (9) with WkW_{k}^{*}, we obtain:

Wk(AP12,ω+P12,ωAkT+FωBBkT+BBkTFk,ω)\displaystyle W_{k}^{*}\big{(}AP_{12,\omega}+P_{12,\omega}A_{k}^{T}+F_{\omega}BB_{k}^{T}+BB_{k}^{T}F_{k,\omega}^{*}\big{)} =0\displaystyle=0
Ak+AkT+Fk,ωBkBkT+BkBkTFk,ω\displaystyle A_{k}+A_{k}^{T}+F_{k,\omega}B_{k}B_{k}^{T}+B_{k}B_{k}^{T}F_{k,\omega} =0.\displaystyle=0.

Given the uniqueness of solution for equation (10), we have Pk,ω=IP_{k,\omega}=I.

Note that Y12,ω+2Z12,ωY_{12,\omega}+2Z_{12,\omega} and Yk,ω+2Zk,ωY_{k,\omega}+2Z_{k,\omega} satisfies the following equations:

AT(Y12,ω+2Z12,ω)+(Y12,ω+2Z12,ω)Ak+FωCCkT+CCkTFk,ω\displaystyle A^{T}(Y_{12,\omega}+2Z_{12,\omega})+(Y_{12,\omega}+2Z_{12,\omega})A_{k}+F_{\omega}CC_{k}^{T}+CC_{k}^{T}F_{k,\omega}^{*}
+2i=1p(FωMiP12,ωMk,i+MiP12,ωMk,iFk,ω)=0,\displaystyle\hskip 83.93553pt+2\sum_{i=1}^{p}\big{(}F_{\omega}^{*}M_{i}P_{12,\omega}M_{k,i}+M_{i}P_{12,\omega}M_{k,i}F_{k,\omega}\big{)}=0, (35)
AkT(Yk,ω+2Zk,ω)+(Yk,ω+2Zk,ω)Ak+Fk,ωCkCkT+CkCkTFk,ω\displaystyle A_{k}^{T}(Y_{k,\omega}+2Z_{k,\omega})+(Y_{k,\omega}+2Z_{k,\omega})A_{k}+F_{k,\omega}C_{k}C_{k}^{T}+C_{k}C_{k}^{T}F_{k,\omega}^{*}
+2i=1p(Fk,ωMk,iPk,ωMk,i+Mk,iPk,ωMk,iFk,ω)=0.\displaystyle\hskip 83.93553pt+2\sum_{i=1}^{p}\big{(}F_{k,\omega}^{*}M_{k,i}P_{k,\omega}M_{k,i}+M_{k,i}P_{k,\omega}M_{k,i}F_{k,\omega}\big{)}=0. (36)

Taking the Hermitian of equation (35) and post-multiplying it by VkV_{k}, we get:

(AkT(Y12,ω+2Z12,ω)+(Y12,ω+2Z12,ω)A+CkTCFω+Fk,ωCkTC\displaystyle\Big{(}A_{k}^{T}(Y_{12,\omega}+2Z_{12,\omega})^{*}+(Y_{12,\omega}+2Z_{12,\omega})^{*}A+C_{k}^{T}CF_{\omega}+F_{k,\omega}^{*}C_{k}^{T}C
+i=1p2Mk,iP12,ωMiFω+i=1p2Fk,ωMk,iP12,ωMi)Vk\displaystyle\hskip 93.89418pt+\sum_{i=1}^{p}2M_{k,i}P_{12,\omega}^{*}M_{i}F_{\omega}+\sum_{i=1}^{p}2F_{k,\omega}^{*}M_{k,i}P_{12,\omega}^{*}M_{i}\Big{)}V_{k}
=AkT+Ak+CkTCkFk,ω+Fk,ωCkTCk\displaystyle=A_{k}^{T}+A_{k}+C_{k}^{T}C_{k}F_{k,\omega}+F_{k,\omega}^{*}C_{k}^{T}C_{k}
+i=1p2Mk,iMk,iFk,ω+i=1p2Fk,ωMk,iMk,i=0.\displaystyle\hskip 93.89418pt+\sum_{i=1}^{p}2M_{k,i}M_{k,i}F_{k,\omega}+\sum_{i=1}^{p}2F_{k,\omega}^{*}M_{k,i}M_{k,i}=0.

With the uniqueness of solutions for equations (10) and (36), we have Pk,ω=IP_{k,\omega}=I and Yk,ω+2Zk,ω=IY_{k,\omega}+2Z_{k,\omega}=I. As a result, the optimality conditions (22)-(24) are met with Vk=P12,ωV_{k}=P_{12,\omega}, and Wk=Y12,ω+2Z12,ωW_{k}=Y_{12,\omega}+2Z_{12,\omega}.

References

  • [1] J. M. Montenbruck, S. Zeng, F. Allgöwer, Linear systems with quadratic outputs, in: 2017 American Control Conference (ACC), IEEE, 2017, pp. 1030–1034.
  • [2] A. Van Der Schaft, Port-hamiltonian systems: An introductory survey, in: International Congress of Mathematicians, European Mathematical Society Publishing House (EMS Ph), 2006, pp. 1339–1365.
  • [3] B. Picinbono, P. Devaut, Optimal linear-quadratic systems for detection and estimation, IEEE Transactions on Information Theory 34 (2) (1988) 304–311.
  • [4] Q. Aumann, S. W. Werner, Structured model order reduction for vibro-acoustic problems using interpolation and balancing methods, Journal of Sound and Vibration 543 (2023) 117363.
  • [5] S. Reiter, S. W. Werner, Interpolatory model order reduction of large-scale dynamical systems with root mean squared error measures, arXiv preprint arXiv:2403.08894 (2024).
  • [6] R. Van Beeumen, K. Van Nimmen, G. Lombaert, K. Meerbergen, Model reduction for dynamical systems with quadratic output, International Journal for Numerical Methods in Engineering 91 (3) (2012) 229–248.
  • [7] P. Benner, V. Mehrmann, D. C. Sorensen, Dimension reduction of large-scale systems, Vol. 45, Springer, 2005.
  • [8] A. Antoulas, D. Sorensen, K. A. Gallivan, P. Van Dooren, A. Grama, C. Hoffmann, A. Sameh, Model reduction of large-scale dynamical systems, in: Computational Science-ICCS 2004: 4th International Conference, Kraków, Poland, June 6-9, 2004, Proceedings, Part III 4, Springer, 2004, pp. 740–747.
  • [9] A. C. Antoulas, R. Ionutiu, N. Martins, E. J. W. ter Maten, K. Mohaghegh, R. Pulch, J. Rommes, M. Saadvandi, M. Striebel, Model order reduction: Methods, concepts and properties, Coupled Multiscale Simulation and Optimization in Nanoelectronics (2015) 159–265.
  • [10] A. C. Antoulas, C. A. Beattie, S. Gugercin, Interpolatory methods for model reduction, SIAM, 2020.
  • [11] B. Moore, Principal component analysis in linear systems: Controllability, observability, and model reduction, IEEE Transactions on Automatic Control 26 (1) (1981) 17–32.
  • [12] D. F. Enns, Model reduction with balanced realizations: An error bound and a frequency weighted generalization, in: The 23rd IEEE Conference on Decision and Control, IEEE, 1984, pp. 127–132.
  • [13] V. Mehrmann, T. Stykel, Balanced truncation model reduction for large-scale systems in descriptor form, in: Dimension Reduction of Large-Scale Systems: Proceedings of a Workshop held in Oberwolfach, Germany, October 19–25, 2003, Springer, 2005, pp. 83–115.
  • [14] T. Reis, T. Stykel, Balanced truncation model reduction of second-order systems, Mathematical and Computer Modelling of Dynamical Systems 14 (5) (2008) 391–406.
  • [15] H. Sandberg, A. Rantzer, Balanced truncation of linear time-varying systems, IEEE Transactions on Automatic Control 49 (2) (2004) 217–229.
  • [16] N. T. Son, P.-Y. Gousenbourger, E. Massart, T. Stykel, Balanced truncation for parametric linear systems using interpolation of gramians: a comparison of algebraic and geometric approaches, Model Reduction of Complex Dynamical systems (2021) 31–51.
  • [17] B. Kramer, K. Willcox, Balanced truncation model reduction for lifted nonlinear systems, in: Realization and Model Reduction of Dynamical Systems: A Festschrift in Honor of the 70th Birthday of Thanos Antoulas, Springer, 2022, pp. 157–174.
  • [18] P. Benner, P. Goyal, M. Redmann, Truncated gramians for bilinear systems and their advantages in model order reduction, Model reduction of Parametrized Systems (2017) 285–300.
  • [19] N. Wong, V. Balakrishnan, Fast positive-real balanced truncation via quadratic alternating direction implicit iteration, IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems 26 (9) (2007) 1725–1731.
  • [20] C. Guiver, M. R. Opmeer, Bounded real and positive real balanced truncation for infinite-dimensional systems, Mathematical Control and Related Fields 3 (1) (2013) 83–119.
  • [21] N. Wong, V. Balakrishnan, C.-K. Koh, Passivity-preserving model reduction via a computationally efficient project-and-balance scheme, in: Proceedings of the 41st Annual Design Automation Conference, 2004, pp. 369–374.
  • [22] A. Sarkar, J. M. Scherpen, Structure-preserving generalized balanced truncation for nonlinear Port-Hamiltonian systems, Systems & Control Letters 174 (2023) 105501.
  • [23] S. Gugercin, A. C. Antoulas, A survey of model reduction by balanced truncation and some new results, International Journal of Control 77 (8) (2004) 748–766.
  • [24] R. Van Beeumen, K. Meerbergen, Model reduction by balanced truncation of linear systems with a quadratic output, in: AIP Conference Proceedings, Vol. 1281, American Institute of Physics, 2010, pp. 2033–2036.
  • [25] R. Pulch, A. Narayan, Balanced truncation for model order reduction of linear dynamical systems with quadratic outputs, SIAM Journal on Scientific Computing 41 (4) (2019) A2270–A2295.
  • [26] P. Benner, P. Goyal, I. P. Duff, Gramians, energy functionals, and balanced truncation for linear dynamical systems with quadratic outputs, IEEE Transactions on Automatic Control 67 (2) (2021) 886–893.
  • [27] D. Wilson, Optimum solution of model-reduction problem, in: Proceedings of the Institution of Electrical Engineers, Vol. 117, IET, 1970, pp. 1161–1165.
  • [28] S. Gugercin, A. C. Antoulas, C. Beattie, 2\mathcal{H}_{2} model reduction for large-scale linear dynamical systems, SIAM journal on matrix analysis and applications 30 (2) (2008) 609–638.
  • [29] Y. Xu, T. Zeng, Optimal 2\mathcal{H}_{2} model reduction for large scale MIMO systems via tangential interpolation, International Journal of Numerical Analysis & Modeling 8 (1) (2011).
  • [30] P. Benner, M. Køhler, J. Saak, Sparse-dense Sylvester equations in 2\mathcal{H}_{2}-model order reduction, Preprint MPIMD/11-11, Max Planck Institute Magdeburg, available from http://www.mpi-magdeburg.mpg.de/preprints/ (Dec. 2011).
  • [31] S. Reiter, I. Pontes Duff, I. V. Gosea, S. Gugercin, 2\mathcal{H}_{2} optimal model reduction of linear systems with multiple quadratic outputs, arXiv preprint arXiv:2405.05951 (2024).
  • [32] R. W. Aldhaheri, Frequency-domain model reduction approach to design iir digital filters using orthonormal bases, AEU-International Journal of Electronics and Communications 60 (6) (2006) 413–420.
  • [33] G. Obinata, B. D. Anderson, Model reduction for control system design, Springer Science & Business Media, 2012.
  • [34] D. F. Enns, Model reduction for control system design, Tech. Rep. NASA-CR-170417, NASA, CA (1985).
  • [35] U. Zulfiqar, V. Sreeram, X. Du, Finite-frequency power system reduction, International Journal of Electrical Power & Energy Systems 113 (2019) 35–44.
  • [36] U. Zulfiqar, V. Sreeram, X. Du, Frequency-limited pseudo-optimal rational Krylov algorithm for power system reduction, International Journal of Electrical Power & Energy Systems 118 (2020) 105798.
  • [37] W. Gawronski, J.-N. Juang, Model reduction in limited time and frequency intervals, International Journal of Systems Science 21 (2) (1990) 349–376.
  • [38] P. Benner, P. Kürschnerrschner, J. Saak, Frequency-limited balanced truncation with low-rank approximations, SIAM Journal on Scientific Computing 38 (1) (2016) A471–A499.
  • [39] M. Imran, A. Ghafoor, Model reduction of descriptor systems using frequency limited Gramians, Journal of the Franklin Institute 352 (1) (2015) 33–51.
  • [40] P. Benner, S. W. Werner, Frequency-and time-limited balanced truncation for large-scale second-order systems, Linear Algebra and its Applications 623 (2021) 68–103.
  • [41] H. R. Shaker, M. Tahavori, Frequency-interval model reduction of bilinear systems, IEEE Transactions on Automatic Control 59 (7) (2013) 1948–1953.
  • [42] Q.-Y. Song, U. Zulfiqar, Z.-H. Xiao, M. M. Uddin, V. Sreeram, Balanced truncation of linear systems with quadratic outputs in limited time and frequency intervals, arXiv preprint arXiv:2402.11445 (2024).
  • [43] D. Petersson, J. Löfberg, Model reduction using a frequency-limited 2\mathcal{H}_{2}-cost, Systems & Control Letters 67 (2014) 32–39.
  • [44] P. Vuillemin, C. Poussot-Vassal, D. Alazard, 2\mathcal{H}_{2} optimal and frequency limited approximation methods for large-scale LTI dynamical systems, IFAC Proceedings Volumes 46 (2) (2013) 719–724.
  • [45] X. Du, K. I. B. Iqbal, M. M. Uddin, A. M. Fony, M. T. Hossain, M. I. Ahmad, M. S. Hossain, Computational techniques for 2\mathcal{H}_{2} optimal frequency-limited model order reduction of large-scale sparse linear systems, Journal of Computational Science 55 (2021) 101473.
  • [46] U. Zulfiqar, X. Du, Q.-Y. Song, V. Sreeram, On frequency-and time-limited 2\mathcal{H}_{2}-optimal model order reduction, Automatica 153 (2023) 111012.
  • [47] N. J. Higham, Functions of matrices: theory and computation, SIAM, 2008.
  • [48] D. Petersson, A nonlinear optimization approach to 2\mathcal{H}_{2}-optimal modeling and control, Ph.D. thesis, Linköping University Electronic Press (2013).
  • [49] K. B. Petersen and M. S. Pedersen, The matrix cookbook, Technical University of Denmark 7 (15) (2008) 510.
  • [50] U. Zulfiqar, V. Sreeram, X. Du, Adaptive frequency-limited 2\mathcal{H}_{2}-model order reduction, Asian Journal of Control 24 (6) (2022) 2807–2823.