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

Instabilizability Conditions for Continuous-Time Stochastic Systems Under Control Input Constraints

Ahmet Cetinkaya    \IEEEmembershipMember, IEEE    Masako Kishida    \IEEEmembershipSenior Member, IEEE This work was supported by JST ERATO HASUO Metamathematics for Systems Design Project (No. JPMJER1603) and by JSPS KAKENHI Grant Number 20K14771.Ahmet Cetinkaya and Masako Kishida are with the National Institute of Informatics, Tokyo, 101-8430, Japan. (e-mail: c[email protected], [email protected]).
Abstract

In this paper, we investigate constrained control of continuous-time linear stochastic systems. We show that for certain system parameter settings, constrained control policies can never achieve stabilization. Specifically, we explore a class of control policies that are constrained to have a bounded average second moment for Ito-type stochastic differential equations with additive and multiplicative noise. We prove that in certain settings of the system parameters and the bounding constant of the control constraint, divergence of the second moment of the system state is inevitable regardless of the initial state value and regardless of how the control policy is designed.

{IEEEkeywords}

Stochastic systems, constrained control, linear systems

1 Introduction

\IEEEPARstart

Stabilization under control input constraints is an important research problem due to its wide applicability to systems with actuator saturation. The works [1, 2] describe the challenges of this problem and provide comprehensive discussions of the important results. A key result on this problem is an impossibility result: linear deterministic systems with strictly unstable system matrices cannot be globally stabilized if the norm of the control input is constrained to stay below a constant threshold [3, 4].

There is a rapidly growing interest in exploring control input constraints for stochastic systems. For instance, [5, 6, 7, 8] proposed stochastic model predictive controllers with control constraints; [9] and [10] developed reinforcement learning control frameworks with constraints, [11] investigated fuzzy controllers for stochastic systems with actuator saturation. Constrained control of nonlinear stochastic systems was investigated by [12] and [13], and moreover, [14] explored control constraints in stochastic networked control systems.

The work [15] presented an impossibility result for constrained control of discrete-time stochastic systems. It was shown there that if the control input of a strictly unstable discrete-time stochastic system is subject to hard norm-constraints, then the second moment of the state always diverges under nonvanishing and unbounded additive stochastic process noise. A common approach to overcome the difficulties in the stabilization of strictly unstable systems is to consider probabilistic constraints instead of hard deterministic constraints. However, it was shown in [16] that under certain conditions, stabilization of a discrete-time linear stochastic system is impossible even under probabilistic constraints.

The scope of the impossibility results provided in the abovementioned articles covers discrete-time stochastic systems with additive noise. In this paper, we are motivated to expand this scope by addressing two issues. First, we want to know if similar impossibility results can be obtained for continuous-time stochastic systems. Secondly, we want to investigate the effects of both additive and multiplicative noise terms. Handling multiplicative noise terms is important, since such terms can characterize parametric uncertainties in the system (see [17, 18]). As our main contribution, we identify the scenarios where stabilization of a continuous-time stochastic system (with both additive and multiplicative noise) is not possible under probabilistically-constrained control policies. Specifically, we consider control policies that have bounded time-averaged second moments. This class of control policies encapsulate many types of controllers with (probabilistic or deterministic) control constraints. We obtain conditions on the bounding value of the control constraint, under which the second moment of the state diverges regardless of the controller choice and regardless of the initial state value.

Our analysis for the continuous-time systems with additive and multiplicative noise has a few key differences from that for the discrete-time case with additive-only noise provided in [15, 16]. First, in our case, we handle Ito-type stochastic differential equations with state-dependent noise terms characterizing multiplicative Wiener noise. In addition, we develop a form of reverse Gronwall’s inequality to obtain lower bounds on functions with superlinear growth. Through our analysis, we observe that combination of additive and multiplicative noise can make systems harder to stabilize. Even systems that have Hurwitz-stable system matrices can be impossible to stabilize with constrained controllers under the combination of additive and multiplicative noise.

We organize the rest of the paper as follows. In Section 2, we describe the constrained control problem. Then in Sections 3 and 4, we provide our impossibility results for constrained control of continuous-time stochastic systems. Finally, we present numerical examples in Section 5 and conclude the paper in Section 6.

Notation: We denote the Euclidean norm by \|\cdot\|, the trace operator by tr()\mathrm{tr}(\cdot), and the maximum eigenvalue of a Hermitian matrix Hn×nH\in\mathbb{C}^{n\times n} by λmax(H)\lambda_{\max}(H). We use H12H^{\frac{1}{2}} to represent the unique nonnegative-definite Hermitian square root of a nonnegative-definite Hermitian matrix Hn×nH\in\mathbb{C}^{n\times n}, satisfying H12H12=HH^{\frac{1}{2}}H^{\frac{1}{2}}=H and (H12)=H12(H^{\frac{1}{2}})^{*}=H^{\frac{1}{2}}. The identity matrix in n×n\mathbb{R}^{n\times n} is denoted by InI_{n}. The notations []\mathrm{\mathbb{P}}[\cdot] and 𝔼[]\mathbb{E}[\cdot] respectively denote the probability and expectation on a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) with sample space Ω\Omega and σ\sigma-algebra \mathcal{F}. We consider a continuous-time filtration {t}t0\{\mathcal{F}_{t}\}_{t\geq 0} with t1t2\mathcal{F}_{t_{1}}\subseteq\mathcal{F}_{t_{2}}\subseteq\mathcal{F} for t1t2t_{1}\leq t_{2}. Throughout the paper {W(t)=[W1(t),,W(t)]T}t0\{W(t)=[W_{1}(t),\ldots,W_{\ell}(t)]^{\mathrm{T}}\in\mathbb{R}^{\ell}\}_{t\geq 0} denotes the Wiener process. Here, for every i{1,,}i\in\{1,\ldots,\ell\}, the process {Wi(t)}t0\{W_{i}(t)\in\mathbb{R}\}_{t\geq 0} is t\mathcal{F}_{t}-adapted; {Wi(t)}t0\{W_{i}(t)\in\mathbb{R}\}_{t\geq 0}, i{1,,}i\in\{1,\ldots,\ell\}, are independent processes. Moreover, c¯\overline{c} denotes the complex conjugate of a complex number cc\in\mathbb{C}, and Re(c)\mathrm{Re}(c) denotes its real part. We use CC^{*} to denote the complex conjugate transpose of a complex matrix Cn×mC\in\mathbb{C}^{n\times m}, that is, (C)=i,jCj,i¯(C^{*}){}_{i,j}=\overline{C_{j,i}}, i{1,,m}i\in\{1,\ldots,m\}, j{1,,n}j\in\{1,\ldots,n\}. Given a vector vnv\in\mathbb{R}^{n}, and indices i,j{1,,n}i,j\in\{1,\ldots,n\}, iji\leq j, we define vi:jji+1v_{i:j}\in\mathbb{R}^{j-i+1} as vi:j[vi,,vj]Tv_{i:j}\triangleq[v_{i},\ldots,v_{j}]^{\mathrm{T}}.

2 Constrained Control of Continuous-Time Linear Stochastic Systems

Consider the continuous-time linear stochastic system described by the Ito-type stochastic differential equation

dx(t)\displaystyle\mathrm{d}x(t) =(Ax(t)+Bu(t))dt+[Ψ(x(t)),D]dW(t),\displaystyle=(Ax(t)+Bu(t))\mathrm{d}t+\left[\Psi(x(t)),\,D\right]\mathrm{d}W(t), (1)

for t0t\geq 0, where x(t)nx(t)\in\mathbb{R}^{n} is the state with deterministic initial value x(0)=x0x(0)=x_{0}, u(t)mu(t)\in\mathbb{R}^{m} is the control input, and moreover, {W(t)}t0\{W(t)\in\mathbb{R}^{\ell}\}_{t\geq 0} is the Wiener process.

The matrices An×nA\in\mathbb{R}^{n\times n} and Bn×mB\in\mathbb{R}^{n\times m} are called system and input matrices, respectively. Moreover, Ψ(x(t))n×1\Psi(x(t))\in\mathbb{R}^{n\times\ell_{1}} and Dn×2D\in\mathbb{R}^{n\times\ell_{2}} (with 1+2=\ell_{1}+\ell_{2}=\ell) are noise matrices. The matrix-valued function Ψ:nn×1\Psi\colon\mathbb{R}^{n}\to\mathbb{R}^{n\times\ell_{1}} characterizes the effects of multiplicative noise and it is given by

Ψ(x)\displaystyle\Psi(x) =[C1x,C2x,,C1x],\displaystyle=[C_{1}x,C_{2}x,\ldots,C_{\ell_{1}}x], (2)

where Cin×nC_{i}\in\mathbb{R}^{n\times n}, i{1,,1}i\in\{1,\ldots,\ell_{1}\}. The matrix Dn×2D\in\mathbb{R}^{n\times\ell_{2}} in (1) is used for characterizing the effects of additive noise.

Notice that W1:1()W_{1:\ell_{1}}(\cdot) enters in the dynamics as multiplicative noise and W1+1:()W_{\ell_{1}+1:\ell}(\cdot) enters as additive noise, since [Ψ(x(t)),D]dW(t)=Ψ(x(t))dW1:1(t)+DdW1+1:(t)\left[\Psi(x(t)),\,D\right]\mathrm{d}W(t)=\Psi(x(t))\mathrm{d}W_{1:\ell_{1}}(t)+D\mathrm{d}W_{\ell_{1}+1:\ell}(t).

In this paper, we are interested in a stabilization problem. Since the Wiener process W1+1:()W_{\ell_{1}+1:\ell}(\cdot) enters in the dynamics in an additive way, the state and its moments cannot converge to 0 regardless of the control input, unless D=0D=0. For this reason, asymptotic stabilization is not possible and a weaker notion of stabilization is needed. In this paper, we consider the bounded second-moment stabilization notion, where the control goal is to achieve supt0𝔼[x(t)2]<\sup_{t\geq 0}\mathbb{E}[\|x(t)\|^{2}]<\infty.

We consider a stochastic constraint such that the time-averaged 22nd moment of u(t)u(t) is bounded by u^0\hat{u}\geq 0, i.e.,

1t0t𝔼[u(τ)2]dτ\displaystyle\frac{1}{t}\int_{0}^{t}\mathbb{E}[\left\|u(\tau)\right\|^{2}]\mathrm{d}\tau u^,t0.\displaystyle\leq\hat{u},\quad t\geq 0. (3)

This constraint is a relaxation of other types of control constraints, i.e., the satisfaction of (3) does not necessarily imply satisfaction of other constraints. Note on the other hand that norm-constraints (e.g., u(t)u¯\|u(t)\|\leq\overline{u} or u(t)u¯\|u(t)\|_{\infty}\leq\overline{u}), time-averaged norm constraints (e.g., 1t0tu(τ)dτu¯\frac{1}{t}\int_{0}^{t}\|u(\tau)\|\mathrm{d}\mathrm{\tau}\leq\overline{u}), as well as first- and second-moment constraints (e.g., 𝔼[u(t)]u¯\mathbb{E}[\|u(t)\|]\leq\overline{u} or 𝔼[u(t)2]u¯\mathbb{E}[\|u(t)\|^{2}]\leq\overline{u}) all satisfy (3) for certain values of u^\hat{u}.

Remark 2.1

The structure of (3) is motivated by the networked control problem of a plant with a remotely located controller. In this problem, control commands uC(t)u_{\mathrm{C}}(t) transmitted from the controller are subject to packet losses, and the plant sets its input u(t)u(t) to 0 if there is a packet loss (and to uC(t)u_{\mathrm{C}}(t) otherwise). The actuator at the plant side has a hard constraint requiring uC(t)2u¯C\|u_{\mathrm{C}}(t)\|^{2}\leq\overline{u}_{\mathrm{C}} for t0t\geq 0. With randomness involved in packet losses, the plant input u(t)u(t) actually satisfies (3) with u^<u¯C\hat{u}<\overline{u}_{\mathrm{C}} (see Section IV.D of [16] for the specific form of u^\hat{u}). Even though the actuator may be powerful (u¯C\overline{u}_{\mathrm{C}} is large), unstable noisy plants in certain scenarios cannot be stabilized if there are very frequent packet losses, because in such cases u^\hat{u} is much smaller than u¯C\overline{u}_{\mathrm{C}}, and the controller is unable to provide inputs with sufficient average energy to the plant. \triangleleft

For given A,B,Ψ,DA,B,\Psi,D, our goal is to find a threshold for u^\hat{u}, below which stabilization of (1) becomes impossible and the second moment 𝔼[x(t)2]\mathbb{E}[\|x(t)\|^{2}] diverges regardless of the controller design.

The following lemmas are used in the derivation of our main result in Section 3. The first lemma is related to the bounding value u^\hat{u} of the control constraint (3). The second lemma is an extension of Gronwall’s lemma (see, e.g., [19]), where the key condition involves a linear term and the result provides a lower bound instead of an upper bound.

Lemma 2.2

Let u^[0,)\hat{u}\in[0,\infty), κ(0,)\kappa\in(0,\infty) be scalars that satisfy u^<κ\hat{u}<\kappa. Then 𝒬{q>1:u^<κ/q}\mathcal{Q}\triangleq\left\{q>1\colon\hat{u}<\kappa/q\right\} is non-empty.

Proof 2.3.

Let q~2κ/(u^+κ)\tilde{q}\triangleq 2\kappa/(\hat{u}+\kappa). Since u^<κ\hat{u}<\kappa, we have 2κ>u^+κ2\kappa>\hat{u}+\kappa, which implies q~>1\tilde{q}>1. Moreover, since u^<κ\hat{u}<\kappa, we have κ/q~=(u^+κ)/2>u^\kappa/\tilde{q}=(\hat{u}+\kappa)/2>\hat{u}. As both q~>1\tilde{q}>1 and u^<κ/q~\hat{u}<\kappa/\tilde{q} hold, we have q~Q\tilde{q}\in Q, implying that 𝒬\mathcal{Q}\neq\emptyset.

Lemma 2.4.

Given scalars c0,c1c_{0},c_{1}\in\mathbb{R} and ϕ>0\phi>0, suppose that h:[0,)h\colon[0,\infty)\to\mathbb{R} is a continuous function that satisfies

h(t)\displaystyle h(t) c0+c1t+ϕ0th(τ)dτ,t0.\displaystyle\geq c_{0}+c_{1}t+\phi\int_{0}^{t}h(\tau)\mathrm{d}\tau,\quad t\geq 0. (4)

Then we have

h(t)\displaystyle h(t) c0eϕt+(c1/ϕ)(eϕt1),t0.\displaystyle\geq c_{0}e^{\phi t}+(c_{1}/\phi)(e^{\phi t}-1),\quad t\geq 0. (5)

Moreover, if (4) holds with equality, then (5) holds with equality.

Proof 2.5.

Let g(s)c0+c1s+ϕ0sh(τ)dτg(s)\triangleq c_{0}+c_{1}s+\phi\int_{0}^{s}h(\tau)\mathrm{d}\tau for s0s\geq 0. Since hh is a continuous function, by fundamental theorem of calculus, we have dg(s)ds=c1+ϕh(s)\frac{\mathrm{d}g(s)}{\mathrm{d}s}=c_{1}+\phi h(s). Note that (4) implies h(s)g(s)h(s)\geq g(s). Furthermore, since ϕ>0\phi>0,

d(g(s)eϕs)ds=dg(s)dseϕsϕeϕsg(s)\displaystyle\frac{\mathrm{d}(g(s)e^{-\phi s})}{\mathrm{d}s}=\frac{\mathrm{d}g(s)}{\mathrm{d}s}e^{-\phi s}-\phi e^{-\phi s}g(s)
=c1eϕs+ϕeϕs(h(s)g(s))c1eϕs.\displaystyle\quad=c_{1}e^{-\phi s}+\phi e^{-\phi s}(h(s)-g(s))\geq c_{1}e^{-\phi s}. (6)

By integrating both left- and far right-hand sides of the inequality (6) over the interval [0,t][0,t], we get g(t)eϕtg(0)c1ϕ(1eϕt)g(t)e^{-\phi t}-g(0)\geq\frac{c_{1}}{\phi}(1-e^{-\phi t}). By using this inequality and g(0)=c0g(0)=c_{0}, we obtain g(t)c0eϕt+c1ϕ(eϕt1)g(t)\geq c_{0}e^{\phi t}+\frac{c_{1}}{\phi}(e^{\phi t}-1) for t0t\geq 0, which implies (5), since h(t)g(t)h(t)\geq g(t). Finally, if (4) holds with equality, then h(s)=g(s)h(s)=g(s) for s0s\geq 0, and thus, (6) holds with equality, which implies that (5) holds with equality.

3 Conditions for Impossibility of Stabilization

In this section, we present our main result, which provides conditions on the control constraint (3), under which the stochastic system (1) is impossible to be stabilized.

Theorem 3.1.

Consider the stochastic system (1). Assume that there exist a nonnegative-definite Hermitian matrix Rn×n{0}R\in\mathbb{C}^{n\times n}\setminus\{0\} and a scalar ϕL>0\phi_{\mathrm{L}}>0 such that

ATR+RA+i=11CiTRCi\displaystyle A^{\mathrm{T}}R+RA+\sum_{i=1}^{\ell_{1}}C_{i}^{\mathrm{T}}RC_{i} ϕLR,\displaystyle\geq\phi_{\mathrm{L}}R, (7)
tr(DTRD)\displaystyle\mathrm{tr}\left(D^{\mathrm{T}}RD\right) >0.\displaystyle>0. (8)

If the control policy is t\mathcal{F}_{t}-adapted and satisfies (3) with

u^\displaystyle\hat{u} <{ϕLtr(DTRD)/βU,ifβU0,,otherwise,\displaystyle<\begin{cases}\phi_{\mathrm{L}}\mathrm{tr}\left(D^{\mathrm{T}}RD\right)/\beta_{\mathrm{U}},&\mathrm{if}\,\,\beta_{\mathrm{U}}\neq 0,\\ \infty,&\mathrm{otherwise},\end{cases} (9)

where βUλmax(BTRB)\beta_{\mathrm{U}}\triangleq\lambda_{\max}(B^{\mathrm{T}}RB), then the second moment of the state diverges, that is,

limt𝔼[x(t)2]\displaystyle\lim_{t\to\infty}\mathbb{E}[\left\|x(t)\right\|^{2}] =,\displaystyle=\infty, (10)

for any initial state x0nx_{0}\in\mathbb{R}^{n}.

Proof 3.2.

Let V(x)xTRxV(x)\triangleq x^{\mathrm{T}}Rx. As a first step, we will show limt𝔼[V(x(t))]=.\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty. Let ΛATR+RA+i=11CiTRCi\Lambda\triangleq A^{\mathrm{T}}R+RA+\sum_{i=1}^{\ell_{1}}C_{i}^{\mathrm{T}}RC_{i}. It follows from Ito formula (see Section 4.2 of [20]) that

dV(x(t))\displaystyle\mathrm{d}V(x(t)) =tr(DTRD)dt+xT(t)Λx(t)dt\displaystyle=\mathrm{tr}(D^{\mathrm{T}}RD)\mathrm{d}t+x^{\mathrm{T}}(t)\Lambda x(t)\mathrm{d}t
+(xT(t)RBu(t)+uT(t)BTRx(t))dt\displaystyle\,\,\quad+(x^{\mathrm{T}}(t)RBu(t)+u^{\mathrm{T}}(t)B^{\mathrm{T}}Rx(t))\mathrm{d}t
+i=11xT(t)(CiTR+RCi)x(t)dWi(t)\displaystyle\,\,\quad+\sum_{i=1}^{\ell_{1}}x^{\mathrm{T}}(t)(C_{i}^{\mathrm{T}}R+RC_{i})x(t)\mathrm{d}W_{i}(t)
+i=12(xT(t)Rdi+diTRx(t))dWi+1(t),\displaystyle\,\,\quad+\sum_{i=1}^{\ell_{2}}(x^{\mathrm{T}}(t)Rd_{i}+d_{i}^{\mathrm{T}}Rx(t))\mathrm{d}W_{i+\ell_{1}}(t), (11)

where dind_{i}\in\mathbb{R}^{n}, i{1,,2}i\in\{1,\ldots,\ell_{2}\}, denote the columns of matrix DD. Under an t\mathcal{F}_{t}-adapted control policy, {x(t)}t0\{x(t)\}_{t\geq 0} is t\mathcal{F}_{t}-adapted. Thus, by Theorem 3.2.1 of [20], we have 𝔼[0txT(τ)(CiTR+RCi)x(τ)dWi(τ)]=0\mathbb{E}[\int_{0}^{t}x^{\mathrm{T}}(\tau)(C_{i}^{\mathrm{T}}R+RC_{i})x(\tau)\mathrm{d}W_{i}(\tau)]=0 and 𝔼[0t(xT(τ)Rdi+diTRx(τ))dWi+1(τ)]=0\mathbb{E}[\int_{0}^{t}(x^{\mathrm{T}}(\tau)Rd_{i}+d_{i}^{\mathrm{T}}Rx(\tau))\mathrm{d}W_{i+\ell_{1}}(\tau)]=0. As a result, it follows from (11) that

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))]
=𝔼[V(x(0))]+tr(DTRD)t+𝔼[0txT(τ)Λx(τ)dτ]\displaystyle\,=\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t+\mathbb{E}\left[\int_{0}^{t}x^{\mathrm{T}}(\tau)\Lambda x(\tau)\mathrm{d}\tau\right]
+𝔼[0t(xT(τ)RBu(τ)+uT(τ)BTRx(τ))dτ],\displaystyle\,\quad+\mathbb{E}\left[\int_{0}^{t}(x^{\mathrm{T}}(\tau)RBu(\tau)+u^{\mathrm{T}}(\tau)B^{\mathrm{T}}Rx(\tau))\mathrm{d}\tau\right], (12)

for t0t\geq 0. Next, we change the order of expectation and integration in (12) by using Fubini’s theorem [21] and obtain

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))]
=𝔼[V(x(0))]+tr(DTRD)t+0t𝔼[xT(τ)Λx(τ)]dτ\displaystyle\,=\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t+\int_{0}^{t}\mathbb{E}[x^{\mathrm{T}}(\tau)\Lambda x(\tau)]\mathrm{d}\tau
+0t𝔼[xT(τ)RBu(τ)+uT(τ)BTRx(τ)]dτ.\displaystyle\,\quad+\int_{0}^{t}\mathbb{E}[x^{\mathrm{T}}(\tau)RBu(\tau)+u^{\mathrm{T}}(\tau)B^{\mathrm{T}}Rx(\tau)]\mathrm{d}\tau. (13)

In what follows, we use (13) to show limt𝔼[V(x(t))]=\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty, separately for two cases: βU=0\beta_{\mathrm{U}}=0 and βU>0\beta_{\mathrm{U}}>0.

First, consider the case where βU=λmax(BTRB)=0\beta_{\mathrm{U}}=\lambda_{\max}(B^{\mathrm{T}}RB)=0. In this case, we have R12B=0R^{\frac{1}{2}}B=0, and hence RB=0RB=0. Furthermore, (7) implies 𝔼[xT(τ)Λx(τ)]ϕL𝔼[xT(τ)Rx(τ)]=ϕL𝔼[V(x(τ))]\mathbb{E}[x^{\mathrm{T}}(\tau)\Lambda x(\tau)]\geq\phi_{\mathrm{L}}\mathbb{E}[x^{\mathrm{T}}(\tau)Rx(\tau)]=\phi_{\mathrm{L}}\mathbb{E}[V(x(\tau))]. As a consequence, we obtain from (13) that 𝔼[V(x(t))]𝔼[V(x(0))]+tr(DTRD)t+ϕL0t𝔼[V(x(τ))]dτ\mathbb{E}[V(x(t))]\geq\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t+\phi_{\mathrm{L}}\int_{0}^{t}\mathbb{E}[V(x(\tau))]\mathrm{d}\tau for all t0t\geq 0. Therefore, we can use Lemma 2.4 with c0=𝔼[V(x(0))]c_{0}=\mathbb{E}[V(x(0))], c1=tr(DTRD)c_{1}=\mathrm{tr}(D^{\mathrm{T}}RD), ϕ=ϕL\phi=\phi_{\mathrm{L}}, and h(t)=𝔼[V(x(t))]h(t)=\mathbb{E}[V(x(t))] to obtain

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))] 𝔼[V(x(0))]eϕLt\displaystyle\geq\mathbb{E}[V(x(0))]e^{\phi_{\mathrm{L}}t}
+(tr(DTRD)/ϕL)(eϕLt1).\displaystyle\quad+(\mathrm{tr}(D^{\mathrm{T}}RD)/\phi_{\mathrm{L}})(e^{\phi_{\mathrm{L}}t}-1). (14)

Notice that ϕL\phi_{\mathrm{L}} is positive, and hence, limteϕLt=\lim_{t\to\infty}e^{\phi_{\mathrm{L}}t}=\infty. Moreover, tr(DTRD)\mathrm{tr}(D^{\mathrm{T}}RD) is positive by the assumption (8). As a result, (14) implies limt𝔼[V(x(t))]=\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty.

Next, we will show that limt𝔼[V(x(t))]=\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty holds also for the case where βU>0\beta_{\mathrm{U}}>0. For this case let κϕLtr(DTRD)/βU\kappa\triangleq\phi_{\mathrm{L}}\mathrm{tr}\left(D^{\mathrm{T}}RD\right)/\beta_{\mathrm{U}} and 𝒬{q>1:u^<κ/q}\mathcal{Q}\triangleq\left\{q>1\colon\hat{u}<\kappa/q\right\}. By Lemma 2.2, we have 𝒬\mathcal{Q}\neq\emptyset.

Now let q^𝒬\hat{q}\in\mathcal{Q} and define γ(q^)q^/ϕL\gamma(\hat{q})\triangleq\hat{q}/\phi_{\mathrm{L}}. The scalars γ1/2(q^)\gamma^{1/2}(\hat{q}) and γ1/2(q^)\gamma^{-1/2}(\hat{q}) are well-defined since γ(q^)>0\gamma(\hat{q})>0. Moreover, since RR is a nonnegative-definite Hermitian matrix, we have 0zTRz0\leq z^{\mathrm{T}}Rz for any znz\in\mathbb{R}^{n}. Using this inequality with z=γ1/2(q)x(τ)+γ1/2(q)Bu(τ)z=\gamma^{-1/2}(q)x(\tau)+\gamma^{1/2}(q)Bu(\tau), we get

0\displaystyle 0 (γ1/2(q)x(τ)+γ1/2(q)Bu(τ))TR\displaystyle\leq\left(\gamma^{-1/2}(q)x(\tau)+\gamma^{1/2}(q)Bu(\tau)\right)^{\mathrm{T}}R
(γ1/2(q)x(τ)+γ1/2(q)Bu(τ))\displaystyle\quad\cdot\left(\gamma^{-1/2}(q)x(\tau)+\gamma^{1/2}(q)Bu(\tau)\right)
=γ1(q^)xT(τ)Rx(τ)+xT(τ)RBu(τ)\displaystyle=\gamma^{-1}(\hat{q})x^{\mathrm{T}}(\tau)Rx(\tau)+x^{\mathrm{T}}(\tau)RBu(\tau)
+uT(τ)BTRx(τ)+γ(q^)u1T(τ)BTRBu(τ),\displaystyle\quad+u^{\mathrm{T}}(\tau)B^{\mathrm{T}}Rx(\tau)+\gamma(\hat{q})u_{1}^{\mathrm{T}}(\tau)B^{\mathrm{T}}RBu(\tau),

which implies

xT(τ)RBu(τ)+uT(τ)BTRx(τ)\displaystyle x^{\mathrm{T}}(\tau)RBu(\tau)+u^{\mathrm{T}}(\tau)B^{\mathrm{T}}Rx(\tau)
γ1(q^)xT(τ)Rx(τ)γ(q^)uT(τ)BTRBu(τ).\displaystyle\,\geq-\gamma^{-1}(\hat{q})x^{\mathrm{T}}(\tau)Rx(\tau)-\gamma(\hat{q})u^{\mathrm{T}}(\tau)B^{\mathrm{T}}RBu(\tau). (15)

It then follows from (13) together with 𝔼[xT(τ)Λx(τ)]ϕL𝔼[V(x(τ))]\mathbb{E}[x^{\mathrm{T}}(\tau)\Lambda x(\tau)]\geq\phi_{\mathrm{L}}\mathbb{E}[V(x(\tau))] and (15) that

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))] 𝔼[V(x(0))]+tr(DTRD)t\displaystyle\geq\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t
+(ϕLγ1(q^))0t𝔼[V(x(τ))]dτ\displaystyle\quad+(\phi_{\mathrm{L}}-\gamma^{-1}(\hat{q}))\int_{0}^{t}\mathbb{E}[V(x(\tau))]\mathrm{d}\tau
γ(q^)0t𝔼[uT(τ)BTRBu(τ)]dτ.\displaystyle\quad-\gamma(\hat{q})\int_{0}^{t}\mathbb{E}[u^{\mathrm{T}}(\tau)B^{\mathrm{T}}RBu(\tau)]\mathrm{d}\tau. (16)

Since γ(q^)>0\gamma(\hat{q})>0, we have γ(q^)<0-\gamma(\hat{q})<0. Thus, by using uT(τ)BTRBu(τ)λmax(BTRB)u(τ)2=βUu(τ)2u^{\mathrm{T}}(\tau)B^{\mathrm{T}}RBu(\tau)\leq\lambda_{\max}(B^{\mathrm{T}}RB)\|u(\tau)\|^{2}=\beta_{\mathrm{U}}\|u(\tau)\|^{2} with (16), we obtain

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))] 𝔼[V(x(0))]+tr(DTRD)t\displaystyle\geq\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t
+(ϕLγ1(q^))0t𝔼[V(x(τ))]dτ\displaystyle\quad+(\phi_{\mathrm{L}}-\gamma^{-1}(\hat{q}))\int_{0}^{t}\mathbb{E}[V(x(\tau))]\mathrm{d}\tau
γ(q^)βU0t𝔼[u(τ)2]dτ.\displaystyle\quad-\gamma(\hat{q})\beta_{\mathrm{U}}\int_{0}^{t}\mathbb{E}[\|u(\tau)\|^{2}]\mathrm{d}\tau. (17)

Now, since the control policy satisfies (3), we have 0t𝔼[u(τ)2]dτu^t\int_{0}^{t}\mathbb{E}[\|u(\tau)\|^{2}]\mathrm{d}\tau\leq\hat{u}t, and hence, it follows from (17) that

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))] 𝔼[V(x(0))]+tr(DTRD)t\displaystyle\geq\mathbb{E}[V(x(0))]+\mathrm{tr}(D^{\mathrm{T}}RD)t
+(ϕLγ1(q^))0t𝔼[V(x(τ))]dτ\displaystyle\quad+(\phi_{\mathrm{L}}-\gamma^{-1}(\hat{q}))\int_{0}^{t}\mathbb{E}[V(x(\tau))]\mathrm{d}\tau
γ(q^)βUu^t,t0.\displaystyle\quad-\gamma(\hat{q})\beta_{\mathrm{U}}\hat{u}t,\quad t\geq 0. (18)

Let c0𝔼[V(x(0))]c_{0}\triangleq\mathbb{E}[V(x(0))], c1tr(DTRD)γ(q^)βUu^c_{1}\triangleq\mathrm{tr}(D^{\mathrm{T}}RD)-\gamma(\hat{q})\beta_{\mathrm{U}}\hat{u}, ϕϕLγ1(q^)\phi\triangleq\phi_{\mathrm{L}}-\gamma^{-1}(\hat{q}), and h(t)𝔼[V(x(t))]h(t)\triangleq\mathbb{E}[V(x(t))]. By definition of 𝒬\mathcal{Q}, we have q^>1\hat{q}>1, which implies 11/q^>01-1/\hat{q}>0. This inequality and ϕL>0\phi_{\mathrm{L}}>0 imply ϕ=ϕLγ1(q^)=ϕL(11/q^)>0\phi=\phi_{\mathrm{L}}-\gamma^{-1}(\hat{q})=\phi_{\mathrm{L}}(1-1/\hat{q})>0. Thus, by Lemma 2.4, we obtain

𝔼[V(x(t))]\displaystyle\mathbb{E}[V(x(t))] c0eϕt+(c1/ϕ)(eϕt1).\displaystyle\geq c_{0}e^{\phi t}+(c_{1}/\phi)(e^{\phi t}-1). (19)

Since VV is a nonnegative-definite function, we have c00c_{0}\geq 0. Next, we show c1>0c_{1}>0. By definition of 𝒬\mathcal{Q}, we have u^<ϕLtr(DTRD)/(βUq^)\hat{u}<\phi_{\mathrm{L}}\mathrm{tr}\left(D^{\mathrm{T}}RD\right)/(\beta_{\mathrm{U}}\hat{q}). Noting that γ(q^)=q^/ϕL\gamma(\hat{q})=\hat{q}/\phi_{\mathrm{L}}, this inequality implies γ(q^)βUu^<tr(DTRD)\gamma(\hat{q})\beta_{\mathrm{U}}\hat{u}<\mathrm{tr}\left(D^{\mathrm{T}}RD\right). Thus, c1=tr(DTRD)γ(q^)βUu^>0c_{1}=\mathrm{tr}(D^{\mathrm{T}}RD)-\gamma(\hat{q})\beta_{\mathrm{U}}\hat{u}>0. Now, since c00c_{0}\geq 0, c1>0c_{1}>0, and ϕ>0\phi>0 hold, (19) implies limt𝔼[V(x(t))]=\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty.

Finally, since Rn×n{0}R\in\mathbb{C}^{n\times n}\setminus\{0\} (i.e., R0R\neq 0), the nonnegative-definite Hermitian matrix RR has at least one eigenvalue strictly larger than 0. Thus, λmax(R)>0\lambda_{\max}(R)>0. Consequently, V(x(t))λmax(R)x(t)2V(x(t))\leq\lambda_{\max}(R)\|x(t)\|^{2} implies x(t)2(1/λmax(R))V(x(t))\|x(t)\|^{2}\geq\left(1/\lambda_{\max}(R)\right)V(x(t)) for t0t\geq 0. Hence, (10) follows from limt𝔼[V(x(t))]=\lim_{t\to\infty}\mathbb{E}\left[V(x(t))\right]=\infty.

Theorem 3.1 provides sufficient conditions under which the system (1) is instabilizable and the second moment of the state diverges regardless of the controller design and the initial state value. Condition (7) in Theorem 3.1 quantifies the instability of the uncontrolled (u(t)0u(t)\equiv 0) system, and the term tr(DTRD)\mathrm{tr}\left(D^{\mathrm{T}}RD\right) in (8) represents the effect of additive noise characterized with the matrix DD. If there is no multiplicative noise (i.e., Ci=0C_{i}=0 for i{1,,1}i\in\{1,\ldots,\ell_{1}\}), then (7) requires AA to be strictly unstable. On the other hand, when there is multiplicative noise, (7) may hold even if AA is Hurwitz-stable. Notice also that RR is a nonnegative-definite matrix and it may have 0 as an eigenvalue. This property is essential in our analysis, since it allows us to deal with the cases where some of the states are diverging, while the others are stable.

Theorem 3.1 implies that if conditions (7), (8) hold, then it is not possible to stabilize the system by using control inputs with too small average second moments as in (3). The impossibility threshold on the average second moment of control input u(t)u(t) is characterized in (9). If λmax(BTRB)=0\lambda_{\max}(B^{\mathrm{T}}RB)=0, then this threshold value becomes infinity indicating that stabilization is impossible regardless of the input constraint. We note that the case λmax(BTRB)=0\lambda_{\max}(B^{\mathrm{T}}RB)=0 represents the situation, where u(t)u(t) does not have any effect on xT(t)Rx(t)x^{\mathrm{T}}(t)Rx(t).

Remark 3.3 (Instability conditions).

The structure of condition (7) is similar to those of stability/instability conditions provided in [22] for stochastic systems with multiplicative noise. In particular, when specialized to linear systems, Corollary 4.7 of [22] yields an instability condition based on existence of a positive-definite matrix Pn×nP\in\mathbb{R}^{n\times n} and a scalar ψ>0\psi>0 such that ATP+PA+i=11CiTPCiψPA^{\mathrm{T}}P+PA+\sum_{i=1}^{\ell_{1}}C_{i}^{\mathrm{T}}PC_{i}\geq\psi P. Notice that for systems with only multiplicative noise, a positive-definite matrix PP is required to show global instability. In our setting, a nonnegative-definite matrix RR is sufficient, because there is also additive noise and (8) guarantees that this noise can make the projection of the state on unstable modes of the uncontrolled system take a nonzero value even if the initial state is zero. Moreover, under the condition (9), 𝔼[xT(t)Rx(t)]\mathbb{E}[x^{\mathrm{T}}(t)Rx(t)] diverges, which in turn implies divergence of the second moment of the state, as shown in the proof of Theorem 3.1. \triangleleft

Remark 3.4 (Numerical approach).

We note that linear matrix inequalities can be used for checking the conditions of Theorem 3.1. First of all, for a given ϕL\phi_{\mathrm{L}}, condition (7) is linear in RR. Similarly, (8) is a linear inequality of RR. Note that (8) also guarantees that R0R\neq 0. Moreover, the inequality (9) can be transformed into β¯u^<ϕLtr(DTRD)\overline{\beta}\hat{u}<\phi_{\mathrm{L}}\mathrm{tr}\left(D^{\mathrm{T}}RD\right) and BTRBβ¯Im,B^{\mathrm{T}}RB\leq\overline{\beta}I_{m}, which are linear in RR and β¯0\overline{\beta}\geq 0, for a given ϕL\phi_{\mathrm{L}}. If the abovementioned inequalities are satisfied with R=R~R=\widetilde{R}, then R=cR~R=c\widetilde{R} with any c>0c>0 also satisfies them. To restrict the solutions, we can impose an additional constraint tr(R)=1\mathrm{tr}(R)=1. In our numerical method, we iterate over a set of candidate values of ϕL\phi_{\mathrm{L}} and utilize linear matrix inequality solvers (for each value of ϕL\phi_{\mathrm{L}}) to check the conditions of Theorem 3.1. \triangleleft

Remark 3.5 (Partial constraints).

Theorem 3.1 can be extended to handle partial input constraints. Consider dx(t)=(Ax(t)+Bu(t)+Fν(t))dt+[Ψ(x(t)),D]dW(t)\mathrm{d}x(t)=(Ax(t)+Bu(t)+F\nu(t))\mathrm{d}t+\left[\Psi(x(t)),D\right]\mathrm{d}W(t), where u(t)u(t) is constrained as in (3) and ν(t)\nu(t) is unconstrained. If (7)–(9) and RF=0RF=0 hold, then it is impossible to achieve stabilization of this modified system. The proof is similar to that of Theorem 3.1, as RF=0RF=0 implies that V(x(t))=xT(t)Rx(t)V(x(t))=x^{\mathrm{T}}(t)Rx(t) is not affected by ν()\nu(\cdot), and hence (11) holds. \triangleleft

3.1 Tightness of the result for scalar systems

Theorem 3.1 provides a tight bound for u^\hat{u} in (9) for scalar systems with a scalar state and a scalar constrained input.

Consider (1) with scalars A,B,DA,B,D and scalar-valued function Ψ(x)C1x\Psi(x)\triangleq C_{1}x such that 2A+C12>02A+C_{1}^{2}>0, B0B\neq 0, and D0D\neq 0. In this case, conditions (7) and (8) hold with R=1R=1 and ϕL=2A+C12\phi_{\mathrm{L}}=2A+C_{1}^{2}. Thus, Theorem 3.1 implies that if the control policy satisfies (3) with u^<(2A+C12)D2/B2\hat{u}<(2A+C_{1}^{2})D^{2}/B^{2}, then the system is impossible to stabilize regardless of the initial state x0x_{0}. This bound is tight, because, as shown in the following result, stability can be achieved when u^=(2A+C12)D2/B2\hat{u}=(2A+C_{1}^{2})D^{2}/B^{2}.

Proposition 3.6.

Consider (1) with scalars A,B,DA,B,D\in\mathbb{R} and scalar-valued function Ψ(x)C1x\Psi(x)\triangleq C_{1}x. Suppose 2A+C12>02A+C_{1}^{2}>0, B0B\neq 0, D0D\neq 0, and x0=0x_{0}=0. Then feedback control policy u(t)=Kx(t)u(t)=Kx(t) with K=(2A+C12)/BK=-(2A+C_{1}^{2})/B can achieve stabilization (i.e., supt0𝔼[x2(t)]<\sup_{t\geq 0}\mathbb{E}[x^{2}(t)]<\infty) and satisfies (3) with u^=(2A+C12)D2/B2\hat{u}=(2A+C_{1}^{2})D^{2}/B^{2}.

Proof 3.7.

Let Λ2(A+BK)+C12\Lambda\triangleq 2(A+BK)+C_{1}^{2} and ϕ(x,t)eΛtx2\phi(x,t)\triangleq e^{-\Lambda t}x^{2}. By Ito formula (Section 4.2 of [20]),

dϕ(x(t),t)\displaystyle\mathrm{d}\phi(x(t),t) =2eΛtC1x2(t)dW1(t)+2eΛtDx(t)dW2(t)\displaystyle=2e^{-\Lambda t}C_{1}x^{2}(t)\mathrm{d}W_{1}(t)+2e^{-\Lambda t}Dx(t)\mathrm{d}W_{2}(t)
+eΛtD2dt.\displaystyle\quad+e^{-\Lambda t}D^{2}\mathrm{d}t. (20)

Now, since {x(t)}t0\{x(t)\}_{t\geq 0} is an t\mathcal{F}_{t}-adapted process, we obtain 𝔼[0t2eΛτC1x2(τ)dW1(τ)]=0\mathbb{E}[\int_{0}^{t}2e^{-\Lambda\tau}C_{1}x^{2}(\tau)\mathrm{d}W_{1}(\tau)]=0 and 𝔼[0t2eΛτDx(τ)dW2(τ)]=0\mathbb{E}[\int_{0}^{t}2e^{-\Lambda\tau}Dx(\tau)\mathrm{d}W_{2}(\tau)]=0, by using Theorem 3.2.1 of [20]. Thus, with x0=0x_{0}=0, (20) implies 𝔼[ϕ(x(t),t)]=𝔼[ϕ(x(0),0)]+0teΛτD2dτ=(D2/Λ)(eΛt1)\mathbb{E}[\phi(x(t),t)]=\mathbb{E}[\phi(x(0),0)]+\int_{0}^{t}e^{-\Lambda\tau}D^{2}\mathrm{d}\tau=-(D^{2}/\Lambda)(e^{-\Lambda t}-1). Therefore, for all t0t\geq 0, we have 𝔼[x2(t)]=eΛt𝔼[ϕ(x(t),t)]=(D2/Λ)(1eΛt)D2/Λ\mathbb{E}[x^{2}(t)]=e^{\Lambda t}\mathbb{E}[\phi(x(t),t)]=-(D^{2}/\Lambda)(1-e^{\Lambda t})\leq-D^{2}/\Lambda, where the last inequality follows from Λ=(2A+C12)<0\Lambda=-(2A+C_{1}^{2})<0. As a consequence, supt0𝔼[x2(t)]D2/Λ<\sup_{t\geq 0}\mathbb{E}[x^{2}(t)]\leq-D^{2}/\Lambda<\infty, which implies that stability is achieved. Moreover, we have 𝔼[u2(t)]=K2𝔼[x2(t)](2A+C12)D2/B2\mathbb{E}[u^{2}(t)]=K^{2}\mathbb{E}[x^{2}(t)]\leq(2A+C_{1}^{2})D^{2}/B^{2} for all t0t\geq 0, showing that control input constraint (3) is satisfied with u^=(2A+C12)D2/B2\hat{u}=(2A+C_{1}^{2})D^{2}/B^{2}.

Proposition 3.6 handles the case where 2A+C12>02A+C_{1}^{2}>0. With similar analysis, we can also show that if 2A+C12<02A+C_{1}^{2}<0, then 𝔼[x2(t)]\mathbb{E}[x^{2}(t)] stays bounded even without control (i.e., u(t)0u(t)\equiv 0). Furthermore, if 2A+C12=02A+C_{1}^{2}=0, B0B\neq 0, then a state-feedback control policy u(t)=Kx(t)u(t)=Kx(t) with BK<0BK<0 achieves stabilization, and moreover, for x0=0x_{0}=0, it guarantees the bound 𝔼[u2(t)](1/2)D2K/B\mathbb{E}[u^{2}(t)]\leq-(1/2)D^{2}K/B for all t0t\geq 0. This bound can be made arbitrarily small by choosing small |K||K|. If 2A+C12=02A+C_{1}^{2}=0, B=0B=0, then the system is uncontrollable and 𝔼[x2(t)]\mathbb{E}[x^{2}(t)] grows unboundedly unless D=0D=0.

3.2 Existence of instability-inducing noise matrices

The following proposition complements Theorem 3.1. It shows that if A,C1,,C1A,C_{1},\ldots,C_{\ell_{1}} satisfy (7), then for any BB and u^\hat{u}, there exists a noise matrix DD that satisfies both (8) and (9). Thus, by Theorem 3.1, the system with that noise matrix is impossible to be stabilized under constraint (3).

Proposition 3.8.

Assume that there exist a nonnegative-definite Hermitian matrix Rn×n{0}R\in\mathbb{C}^{n\times n}\setminus\{0\} and a scalar ϕL>0\phi_{\mathrm{L}}>0 that satisfy (7). Then for any Bn×mB\in\mathbb{R}^{n\times m} and u^[0,)\hat{u}\in[0,\infty), there exists Dn×2D\in\mathbb{R}^{n\times\ell_{2}} such that both (8) and (9) hold.

Proof 3.9.

By the spectral theorem for Hermitian matrices (see Theorem 2.5.6 of [23]), the nonnegative-definite Hermitian matrix RR can be written as R=Ξdiag(μ1,,μn)ΞR=\Xi\mathrm{diag}(\mu_{1},\ldots,\mu_{n})\Xi^{*}, where μ1,,μn0\mu_{1},\ldots,\mu_{n}\geq 0 are the eigenvalues of RR and Ξn×n{0}\Xi\in\mathbb{C}^{n\times n}\setminus\{0\} is a unitary matrix. Let ξ1,,ξnn{0}\xi_{1},\ldots,\xi_{n}\in\mathbb{C}^{n}\setminus\{0\} denote the columns of Ξ\Xi. We have R=iμiξiξiR=\sum_{i}\mu_{i}\xi_{i}\xi_{i}^{*}. Since R0R\neq 0, at least one eigenvalue of RR is strictly larger than 0. Let i~min{i{1,,n}:μi>0}\tilde{i}\triangleq\min\{i\in\{1,\ldots,n\}\colon\mu_{i}>0\}. Now, let ξi~,j\xi_{\tilde{i},j}\in\mathbb{C} denote the jjth entry of vector ξi~\xi_{\tilde{i}}. Since ξi~0\xi_{\tilde{i}}\neq 0, at least one entry of ξi~\xi_{\tilde{i}} is nonzero. We define j~min{j{1,,n}:ξi~,j0}\tilde{j}\triangleq\min\{j\in\{1,\ldots,n\}\colon\xi_{\tilde{i},j}\neq 0\}. Now let α>0\alpha>0 and D[d1,,d2]D\triangleq[d_{1},\ldots,d_{\ell_{2}}], where dind_{i}\in\mathbb{R}^{n}, i{1,,2}i\in\{1,\ldots,\ell_{2}\}, are columns of DD. We let di=0d_{i}=0 for i1i\neq 1, and set the entries of d1nd_{1}\in\mathbb{R}^{n} as

d1,j\displaystyle d_{1,j} ={(α+(u^βU/ϕL))1/2μi~1/2|ξi~,j~|,ifj=j~,0,otherwise,\displaystyle=\begin{cases}\frac{(\alpha+(\hat{u}\beta_{\mathrm{U}}/\phi_{\mathrm{L}}))^{1/2}}{\mu_{\tilde{i}}^{1/2}|\xi_{\tilde{i},\tilde{j}}|},&\mathrm{if}\,\,j=\tilde{j},\\ 0,&\mathrm{otherwise},\end{cases} (21)

where βU=λmax(BTRB)\beta_{\mathrm{U}}=\lambda_{\max}(B^{\mathrm{T}}RB). Since di=0d_{i}=0 for i1i\neq 1, we have

tr(DTRD)=tr(i=1nμiDTξiξiD)\displaystyle\mathrm{tr}(D^{\mathrm{T}}RD)=\mathrm{tr}\left(\sum_{i=1}^{n}\mu_{i}D^{\mathrm{T}}\xi_{i}\xi_{i}^{*}D\right)
=i=1nμitr(DTξiξiD)=i=1nμij=12(djTξiξidj)\displaystyle\quad=\sum_{i=1}^{n}\mu_{i}\mathrm{tr}(D^{\mathrm{T}}\xi_{i}\xi_{i}^{*}D)=\sum_{i=1}^{n}\mu_{i}\sum_{j=1}^{\ell_{2}}\left(d_{j}^{\mathrm{T}}\xi_{i}\xi_{i}^{*}d_{\mathrm{j}}\right)
=i=1nμid1Tξiξid1=d1T(i=1nμiξiξi)d1.\displaystyle\quad=\sum_{i=1}^{n}\mu_{i}d_{1}^{\mathrm{T}}\xi_{i}\xi_{i}^{*}d_{1}=d_{1}^{\mathrm{T}}\left(\sum_{i=1}^{n}\mu_{i}\xi_{i}\xi_{i}^{*}\right)d_{1}. (22)

Moreover, since i=1nμiξiξiμi~ξi~ξi~\sum_{i=1}^{n}\mu_{i}\xi_{i}\xi_{i}^{*}\geq\mu_{\tilde{i}}\xi_{\tilde{i}}\xi_{\tilde{i}}^{*}, it follows from (22) that tr(DTRD)μi~d1Tξi~ξi~d1\mathrm{tr}(D^{\mathrm{T}}RD)\geq\mu_{\tilde{i}}d_{1}^{\mathrm{T}}\xi_{\tilde{i}}\xi_{\tilde{i}}^{*}d_{1}. By using this inequality and noting that μi~>0\mu_{\tilde{i}}>0, we obtain from (21) that

tr(DTRD)μi~(j=1nd1,jξi~,j~)(j=1nξ¯i~,j~d1,j)\displaystyle\mathrm{tr}(D^{\mathrm{T}}RD)\geq\mu_{\tilde{i}}\left(\sum_{j=1}^{n}d_{1,j}\xi_{\tilde{i},\tilde{j}}\right)\left(\sum_{j=1}^{n}\overline{\xi}_{\tilde{i},\tilde{j}}d_{1,j}\right)
=μi~ξi~,j~ξ¯i~,j~d1,j~2=μi~|ξi~,j~|2α+(u^βU/ϕL)μi~|ξi~,j~|2\displaystyle\quad=\mu_{\tilde{i}}\xi_{\tilde{i},\tilde{j}}\overline{\xi}_{\tilde{i},\tilde{j}}d_{1,\tilde{j}}^{2}=\mu_{\tilde{i}}|\xi_{\tilde{i},\tilde{j}}|^{2}\frac{\alpha+(\hat{u}\beta_{\mathrm{U}}/\phi_{\mathrm{L}})}{\mu_{\tilde{i}}|\xi_{\tilde{i},\tilde{j}}|^{2}}
=α+(u^βU/ϕL),\displaystyle\quad=\alpha+(\hat{u}\beta_{\mathrm{U}}/\phi_{\mathrm{L}}), (23)

which implies (8), as α>0\alpha>0, u^0\hat{u}\geq 0, βU0\beta_{\mathrm{U}}\geq 0, and ϕL>0\phi_{\mathrm{L}}>0. If βU=0\beta_{\mathrm{U}}=0, then the inequality (9) holds directly, as u^<\hat{u}<\infty. If βU0\beta_{\mathrm{U}}\neq 0, then the inequality (23) implies tr(DTRD)>u^βU/ϕL\mathrm{tr}(D^{\mathrm{T}}RD)>\hat{u}\beta_{\mathrm{U}}/\phi_{\mathrm{L}}, which in turn implies (9).

4 Special Setting with Only Additive Noise

In this section, we are interested in a special setting, where Ψ(x(t))=0\Psi(x(t))=0 in (1). In this setting, the system does not face multiplicative noise and it is only subject to additive noise.

We first present an improvement of the numerical approach presented in Remark 3.4. Then we show that eigenstructure of AA can be used to obtain new instability conditions that are easier to check compared to Theorem 3.1. The eigenstructure-based analysis was previously considered only for discrete-time systems in [16]. Here, we show that continuous-time systems also allow a similar approach.

4.1 Candidate values of ϕL\phi_{\mathrm{L}} in instability analysis

Remark 3.4 provides a numerical approach for checking instability conditions of Theorem 3.1. This approach is based on iterating over a set of candidate values of ϕL\phi_{\mathrm{L}} and checking the feasibility of certain linear matrix inequalities. In the general case with both additive and multiplicative noise, we do not have prespecified bounds for the candidate value set. However, in the case of only additive noise (Ψ(x(t))=0\Psi(x(t))=0, i.e., Ci=0C_{i}=0 in (2)), the candidate values of ϕL\phi_{\mathrm{L}} can be restricted to belong to the set (0,2ϑmax(A)](0,2\vartheta_{\max}(A)], where we define ϑmax:n×n\vartheta_{\max}\colon\mathbb{C}^{n\times n}\to\mathbb{R} as

ϑmax(A)\displaystyle\vartheta_{\max}(A) max{Re(λ):λspec(A)}\displaystyle\triangleq\max\{\mathrm{Re}(\lambda)\colon\lambda\in\mathrm{spec}(A)\} (24)

with spec(A)\mathrm{spec}(A)\subset\mathbb{C} denoting the set of eigenvalues of AA. It is sufficient to choose the values of ϕL\phi_{\mathrm{L}} from the set (0,2ϑmax(A)](0,2\vartheta_{\max}(A)], as it is not possible to satisfy (7) with ϕL>2ϑmax(A)\phi_{\mathrm{L}}>2\vartheta_{\max}(A) and R0R\neq 0, as shown in the following proposition. Here, we note that with Ci=0C_{i}=0, i{1,,1}i\in\{1,\ldots,\ell_{1}\}, the inequality (7) reduces to ATR+RAϕLRA^{\mathrm{T}}R+RA\geq\phi_{\mathrm{L}}R.

Proposition 4.1.

Let An×nA\in\mathbb{C}^{n\times n}. For every nonnegative-definite Hermitian matrix Rn×n{0}R\in\mathbb{C}^{n\times n}\setminus\{0\}, there exists yn{0}y\in\mathbb{C}^{n}\setminus\{0\} such that R12y0R^{\frac{1}{2}}y\neq 0 and y(AR+RA)y2ϑmax(A)yRyy^{*}(A^{\mathrm{*}}R+RA)y\leq 2\vartheta_{\max}(A)y^{*}Ry, where ϑmax(A)\vartheta_{\max}(A) is defined in (24).

Proof 4.2.

It follows from Lemma A.1 of [16] that R12Aν^=λ^R12ν^R^{\frac{1}{2}}A\hat{\nu}=\hat{\lambda}R^{\frac{1}{2}}\hat{\nu}, where λ^\hat{\lambda}\in\mathbb{C} is an eigenvalue of AA and ν^n{0}\hat{\nu}\in\mathbb{C}^{n}\setminus\{0\} is a generalized eigenvector of AA that satisfies R12ν^0R^{\frac{1}{2}}\hat{\nu}\neq 0. Let yν^y\triangleq\hat{\nu}. We have R12y0R^{\frac{1}{2}}y\neq 0. Moreover, y(AR+RA)y=λ^¯yRy+λ^yRy=2Re(λ^)yRyy^{*}(A^{\mathrm{*}}R+RA)y=\overline{\hat{\lambda}}y^{*}Ry+\hat{\lambda}y^{*}Ry=2\mathrm{Re}(\hat{\lambda})y^{*}Ry. Therefore, y(AR+RA)y2ϑmax(A)yRy,y^{*}(A^{\mathrm{*}}R+RA)y\leq 2\vartheta_{\max}(A)y^{*}Ry, since Re(λ^)ϑmax(A)\mathrm{Re}(\hat{\lambda})\leq\vartheta_{\max}(A).

4.2 Instability conditions based on the eigenstructure of AA

Even with the improvement discussed in the previous subsection, checking feasibility of the linear matrix inequalities mentioned in Remark 3.4 can be computationally costly. In this subsection, we show that the eigenstructure of the system matrix AA can be used to derive instability conditions that can be checked numerically efficiently.

Let r{1,,n}r\in\{1,\ldots,n\} denote the number of distinct eigenvalues of AA and let λ1,λ2,,λr\lambda_{1},\lambda_{2},\ldots,\lambda_{r}\in\mathbb{C} with λiλj\lambda_{i}\neq\lambda_{j} denote those eigenvalues. Moreover, for every i{1,,r}i\in\{1,\ldots,r\}, let ni{1,,n}n_{i}\in\{1,\ldots,n\} represent the geometric multiplicity of the eigenvalue λi\lambda_{i}. Eigenvalues of AA are also the eigenvalues of ATA^{\mathrm{T}} with the same multiplicities. Thus, for every λi\lambda_{i}, there exists nin_{i} number of vectors vi,jnv_{i,j}\in\mathbb{C}^{n} such that

ATvi,j\displaystyle A^{\mathrm{T}}v_{i,j} =λivi,j,j{1,,ni}.\displaystyle=\lambda_{i}v_{i,j},\quad j\in\{1,\ldots,n_{i}\}. (25)

These vectors vi,1,,vi,niv_{i,1},\ldots,v_{i,n_{i}} are called the left-eigenvectors of AA associated with the eigenvalue λi\lambda_{i}. We remark that if λi\lambda_{i} is a complex eigenvalue (i.e., λi\lambda_{i}\notin\mathbb{R}), then the complex conjugate λi¯\overline{\lambda_{i}} is also an eigenvalue of AA, and moreover,

vi,jA\displaystyle v_{i,j}^{*}A =λi¯vi,j,j{1,,ni},\displaystyle=\overline{\lambda_{i}}v_{i,j}^{*},\quad j\in\{1,\ldots,n_{i}\}, (26)

where vi,jv_{i,j}^{\mathrm{*}} is the complex conjugate transpose of vi,jnv_{i,j}\in\mathbb{C}^{n}. In the instability conditions presented below, we use the eigenvalues λi\lambda_{i}, and the left-eigenvectors vi,jnv_{i,j}\in\mathbb{C}^{n}, j{1,,ni}j\in\{1,\ldots,n_{i}\}, i{1,,r}i\in\{1,\ldots,r\}. Furthermore, we define

{(i,j):Re(λi)>0,vi,jDDTvi,j>0,\displaystyle\mathcal{I}\triangleq\{(i,j)\colon\mathrm{\mathrm{Re}(\lambda_{i})}>0,\,\,v_{i,j}^{*}DD^{\mathrm{T}}v_{i,j}>0,
j{1,,ni},i{1,,r}}.\displaystyle\quad\,\,\,\,\,\,\,\,j\in\{1,\ldots,n_{i}\},\,i\in\{1,\ldots,r\}\}.
Corollary 4.3.

Consider the linear stochastic system (1) where Ψ(x)=0\Psi(x)=0. Suppose \mathcal{I}\neq\emptyset. If the control policy is t\mathcal{F}_{t}-adapted and satisfies (3) with

u^\displaystyle\hat{u} <max(i,j)φi,j,\displaystyle<\max_{(i,j)\in\mathcal{I}}\varphi_{i,j}, (27)

where

φi,j\displaystyle\varphi_{i,j} {2Re(λi)vi,jDDTvi,j/βUi,j,ifβUi,j0,,otherwise,\displaystyle\triangleq\begin{cases}2\mathrm{Re}(\lambda_{i})v_{i,j}^{*}DD^{\mathrm{T}}v_{i,j}/\beta_{\mathrm{U}}^{i,j},&\mathrm{if}\,\,\beta_{\mathrm{U}}^{i,j}\neq 0,\\ \infty,&\mathrm{otherwise},\end{cases}

and βUi,jλmax(BTvi,jvi,jB)\beta_{\mathrm{U}}^{i,j}\triangleq\lambda_{\max}(B^{\mathrm{T}}v_{i,j}v_{i,j}^{*}B), (i,j)(i,j)\in\mathcal{I}, then the second moment of the state diverges (i.e., (10) holds) for any initial state x0nx_{0}\in\mathbb{R}^{n}.

Proof 4.4.

By (25) and (26), for each (i,j)(i,j)\in\mathcal{I}, we obtain ATvi,jvi,j+vi,jvi,jA=λivi,jvi,j+λi¯vi,jvi,j=2Re(λi)vi,jvi,j.A^{\mathrm{T}}v_{i,j}v_{i,j}^{*}+v_{i,j}v_{i,j}^{*}A=\lambda_{i}v_{i,j}v_{i,j}^{*}+\overline{\lambda_{i}}v_{i,j}v_{i,j}^{*}=2\mathrm{Re}(\lambda_{i})v_{i,j}v_{i,j}^{*}. Since vi,jvi,jn×nv_{i,j}v_{i,j}^{*}\in\mathbb{C}^{n\times n} is a nonnegative-definite Hermitian matrix, we have (7) with R=vi,jvi,jR=v_{i,j}v_{i,j}^{*} and ϕL=Re(λi)\phi_{\mathrm{L}}=\mathrm{Re}(\lambda_{i}). Moreover, by the definition of \mathcal{I}, we have tr(DTvi,jvi,jD)=vi,jDDTvi,j>0\mathrm{tr}(D^{\mathrm{T}}v_{i,j}v_{i,j}^{*}D)=v_{i,j}^{*}DD^{\mathrm{T}}v_{i,j}>0 for (i,j)(i,j)\in\mathcal{I}. Thus, (8) holds with R=vi,jvi,jR=v_{i,j}v_{i,j}^{*}. Now, since (7) and (8) both hold for each (i,j)(i,j)\in\mathcal{I}, it follows from Theorem 3.1 by setting βU=βUi,j\beta_{\mathrm{U}}=\beta_{\mathrm{U}}^{i,j} that under control policies satisfying (3) with u^<φi,j\hat{u}<\varphi_{i,j}, the second moment of the state diverges. Finally, (27) implies that there exists (i~,j~)(\tilde{i},\tilde{j})\in\mathcal{I} such that u^<φi~,j~\hat{u}<\varphi_{\tilde{i},\tilde{j}}, implying divergence.

5 Numerical Examples

In this section, we illustrate our results on two example continuous-time stochastic systems.

Example 1

Consider the continuous-time stochastic system described by (1) and (2) with

A=[011a],B=[01],D=d[11],\displaystyle A=\left[\begin{array}[]{rr}0&1\\ -1&a\end{array}\right],\,\,B=\left[\begin{array}[]{c}0\\ 1\end{array}\right],\,\,D=d\left[\begin{array}[]{c}1\\ 1\end{array}\right], (34)
1=2,C1=[00c10],C2=[000c2],\displaystyle\ell_{1}=2,\,\,C_{1}=\left[\begin{array}[]{cc}0&0\\ c_{1}&0\end{array}\right],\,\,C_{2}=\left[\begin{array}[]{cc}0&0\\ 0&c_{2}\end{array}\right], (39)

where a,c1,c2,da,c_{1},c_{2},d\in\mathbb{R} are scalar coefficients. The case with a0a\geq 0 corresponds to the linearized dynamics of the forced Van der Pol oscillator (see Section 5.3.1 of [24]).

In each row of Table 1, we consider a different setting for a,c1,c2,da,c_{1},c_{2},d\in\mathbb{R}. Our goal is to obtain ranges of u^\hat{u} such that the system cannot be stabilized with constrained control policies satisfying (3) with u^\hat{u} chosen in the given range. To obtain the range for each parameter setting, we apply Theorem 3.1. In addition, when there is no multiplicative noise (i.e., c1=c2=0c_{1}=c_{2}=0), we also apply Corollary 4.3.

Setting # aa c1c_{1} c2c_{2} dd Applied Result Range of u^\hat{u}
1 0.5-0.5 22 22 11 Theorem 3.1 [0,7.4)[0,7.4)
2 1.51.5 22 22 11 Theorem 3.1 [0,10.8)[0,10.8)
3 1.51.5 22 0 11 Theorem 3.1 [0,8.4)[0,8.4)
4 1.51.5 0 22 11 Theorem 3.1 [0,4.5)[0,4.5)
5 1.51.5 0 0 11 Theorem 3.1 [0,1)[0,1)
Corollary 4.3 [0,0.75)[0,0.75)
6 1.51.5 0 0 22 Theorem 3.1 [0,4.1)[0,4.1)
Corollary 4.3 [0,3)[0,3)
Table 1: Ranges of u^\hat{u} for which stabilization is impossible.

Setting 1 in Table 1 represents the case where AA is a Hurwitz-stable matrix. Notice that when AA is Hurwitz-stable, without multiplicative noise, the second moment of the state of the uncontrolled system would remain bounded. However, as Table 1 indicates, under multiplicative noise (with parameters c1=c2=2c_{1}=c_{2}=2), the system is unstable under any constrained control that satisfies (3) with u^[0,7.4)\hat{u}\in[0,7.4).

Settings 2–6 in Table 1 represent different scenarios where AA is strictly unstable. In each of those settings, different noise parameters are considered. The main observation is that systems with increased noise levels are harder to stabilize with constrained controllers.

Settings 5 and 6 in Table 1 correspond to the cases where there is no multiplicative noise. In those cases Corollary 4.3 can be applied. Notice that Corollary 4.3 provides smaller ranges for u^\hat{u} compared to Theorem 3.1. This shows that Corollary 4.3 is conservative. We note that the main advantage of Corollary 4.3 is that its conditions can be checked faster than those of Theorem 3.1.

For checking conditions of Corollary 4.3, we can speedily compute max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} in (27). In particular, the computation yields the analytical expression

max(i,j)φi,j\displaystyle\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} ={d2(2a)a,a[0,2),4d2(12a1),a2,\displaystyle=\begin{cases}d^{2}(2-a)a,&a\in[0,2),\\ 4d^{2}(\frac{1}{2}a-1),&a\geq 2,\end{cases} (40)

which is also shown in Fig. 1. Given aa and dd, if u^<max(i,j)φi,j\hat{u}<\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} (i.e., the value is below the surface in Fig. 1), then Corollary 4.3 implies that stabilization of (1) is impossible with a control policy that satisfies (3) with that particular u^\hat{u}. Notice that max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} is a quadratic function of dd, and thus, for larger values of dd, the value of max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} becomes larger. This result is intuitive in the sense that stabilization becomes harder under stronger noise. On the other hand, max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} depends on aa in a nonlinear nonmonotonic way. It follows from (40) that for a[0,2)a\in[0,2), the maximum of max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} is achieved when a=1a=1. For a2a\geq 2, max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} increases as aa increases.

Refer to caption
Figure 1: The value of max(i,j)φi,j\max_{(i,j)\in\mathcal{I}}\varphi_{i,j} in (27). If (3) holds with u^<max(i,j)φi,j\hat{u}<\max_{(i,j)\in\mathcal{I}}\varphi_{i,j}, then stabilization is impossible.

Example 2

Consider (1) and (2) with

A\displaystyle A =[01003ζ2002ζ000102ζ00],B=D=[00100001],\displaystyle=\left[\begin{array}[]{cccc}0&1&0&0\\ 3\zeta^{2}&0&0&2\zeta\\ 0&0&0&1\\ 0&-2\zeta&0&0\end{array}\right],\,\,B=D=\left[\begin{array}[]{cc}0&0\\ 1&0\\ 0&0\\ 0&1\end{array}\right],
1\displaystyle\ell_{1} =1,C1=I4.\displaystyle=1,\,\,C_{1}=I_{4}.

This system is a noisy version of the uncoupled, linearized, and normalized dynamics that describes a satellite’s motion in the equatorial plane, as provided in [25]. The scalar ζ\zeta is the angular velocity of the equatorial orbit along which the system is linearized and the control input u(t)2u(t)\in\mathbb{R}^{2} is the vector of thrusts applied to the satellite in the equatorial plane. We consider the control input constraint (3) with the average second moment bound u^\hat{u}.

We check feasibility of the linear matrix inequalities discussed in Remark 3.4 to assess the conditions of Theorem 3.1 for different values of ζ\zeta and u^\hat{u}. When ζ=0.1\zeta=0.1, the conditions of Theorem 3.1 hold for u^[0,1.7)\hat{u}\in[0,1.7). Thus the system is instabilizable under the control constraint (3) with those values of u^\hat{u}. On the other hand, with ζ=1\zeta=1, the corresponding instabilizability range is obtained as u^[0,1.1)\hat{u}\in[0,1.1).

6 Conclusion

We have investigated the constrained control problem for linear stochastic systems with additive and multiplicative noise terms. We have shown that in certain scenarios, stabilization is impossible to achieve with control policies that have bounded time-averaged second moments. In particular, we have obtained conditions, under which the second moment of the system state diverges regardless of the controller design and regardless of the initial state. Moreover, we have showed the tightness of our results for scalar systems and provided extensions for partially-constrained control policies and additive-only noise settings.

References

  • [1] S. Tarbouriech, G. Garcia, J. M. G. da Silva Jr., and I. Queinnec, Stability and stabilization of linear systems with saturating actuators. Springer, 2011.
  • [2] A. Saberi, A. A. Stoorvogel, and P. Sannuti, Internal and External Stabilization of Linear Systems with Constraints. Springer, 2012.
  • [3] E. D. Sontag, “An algebraic approach to bounded controllability of linear systems,” Int. J. Control, vol. 39, no. 1, pp. 181–188, 1984.
  • [4] H. J. Sussmann, E. D. Sontag, and Y. Yang, “A general result on the stabilization of linear systems using bounded controls,” IEEE Trans. Automat. Control, vol. 39, no. 12, pp. 2411–2425, 1994.
  • [5] P. Hokayem, E. Cinquemani, D. Chatterjee, F. Ramponi, and J. Lygeros, “Stochastic receding horizon control with output feedback and bounded controls,” Automatica, vol. 48, no. 1, pp. 77–88, 2012.
  • [6] M. Korda, R. Gondhalekar, F. Oldewurtel, and C. N. Jones, “Stochastic MPC framework for controlling the average constraint violation,” IEEE Trans. Automat. Control, vol. 59, no. 7, pp. 1706–1721, 2014.
  • [7] L. Hewing and M. N. Zeilinger, “Scenario-based probabilistic reachable sets for recursively feasible stochastic model predictive control,” IEEE Control Syst. Lettr., vol. 4, no. 2, pp. 450–455, 2019.
  • [8] C. Mark and S. Liu, “Stochastic MPC with distributionally robust chance constraints,” IFAC-PapersOnLine, vol. 53, no. 2, pp. 7136–7141, 2020.
  • [9] M. A. Pereira and Z. Wang, “Learning deep stochastic optimal control policies using forward-backward SDEs,” in Robotics: Science and Systems, 2019.
  • [10] H. Wang, T. Zariphopoulou, and X. Y. Zhou, “Reinforcement learning in continuous time and space: A stochastic control approach,” J. Mach. Learn. Res., vol. 21, no. 198, pp. 1–34, 2020.
  • [11] W.-J. Chang, Y.-W. Lin, Y.-H. Lin, C.-L. Pen, and M.-H. Tsai, “Actuator saturated fuzzy controller design for interval type-2 Takagi-Sugeno fuzzy models with multiplicative noises,” Processes, vol. 9, no. 5, 2021.
  • [12] Z. G. Ying and W. Q. Zhu, “A stochastically averaged optimal control strategy for quasi-Hamiltonian systems with actuator saturation,” Automatica, vol. 42, no. 9, pp. 1577–1582, 2006.
  • [13] H. Min, S. Xu, B. Zhang, and Q. Ma, “Output-feedback control for stochastic nonlinear systems subject to input saturation and time-varying delay,” IEEE Trans. Automat. Control, vol. 64, no. 1, pp. 359–364, 2018.
  • [14] P. K. Mishra, D. Chatterjee, and D. E. Quevedo, “Sparse and constrained stochastic predictive control for networked systems,” Automatica, vol. 87, pp. 40–51, 2018.
  • [15] D. Chatterjee, F. Ramponi, P. Hokayem, and J. Lygeros, “On mean square boundedness of stochastic linear systems with bounded controls,” Syst. Control Lettr., vol. 61, pp. 375–380, 2012.
  • [16] A. Cetinkaya and M. Kishida, “Impossibility results for constrained control of stochastic systems,” IEEE Trans. Automat. Control, (to appear), 2021. https://doi.org/10.1109/TAC.2021.3059842.
  • [17] L. El Ghaoui, “State-feedback control of systems with multiplicative noise via linear matrix inequalities,” Syst. Control Lettr., vol. 24, no. 3, pp. 223–228, 1995.
  • [18] R. Khasminskii, Stochastic Stability of Differential Equations. Springer, 2011.
  • [19] A. C. King, J. Billingham, and S. R. Otto, Differential Equations: Linear, Nonlinear, Ordinary, Partial. Cambridge University Press, 2003.
  • [20] B. Øksendal, Stochastic Differential Equations. Springer, 2005.
  • [21] P. Billingsley, Probability and Measure. Wiley, 2012.
  • [22] X. Mao, Stochastic Differential Equations and Applications. Elsevier, 2007.
  • [23] R. A. Horn and C. R. Johnson, Matrix Analysis. Cambridge University Press, 1985.
  • [24] M. Lakshmanan and S. Rajaseekar, Nonlinear Dynamics: Integrability, Chaos and Patterns. Springer, 2012.
  • [25] T. E. Fortmann and K. L. Hitz, An Introduction to Linear Control Systems. Marcel Dekker Inc., 1977.