630
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

Stability of forced higher-order continuous-time Lur'e systems: a behavioural input-output perspective

&
Pages 1788-1809 | Received 06 Dec 2022, Accepted 28 Jun 2023, Published online: 04 Aug 2023

Abstract

We consider a class of forced continuous-time Lur'e systems obtained by applying nonlinear feedback to a higher-order linear differential equation which defines an input-output system in the sense of behavioural systems theory. This linear system directly relates the input and output signals and does not involve any internal, latent or state variables. A stability theory subsuming results of circle criterion type is developed, including criteria for input-to-output stability and strong integral input-to-output stability, concepts which are very much reminiscent of input-to-state stability and strong integral input-to-state stability, respectively. The methods used in the paper combine ideas from the behavioural approach to systems and control, absolute stability theory and input-to-state stability theory.

1. Introduction

Lur'e systems are feedback connections of linear dynamical systems and static nonlinearities–they form a common and important class of nonlinear control systems. There is a large body of work on the stability properties of these systems, including Carrasco and Heath (Citation2021), Carrasco et al. (Citation2016), Franco et al. (Citation2021), Gilmore et al. (Citation2021), Guiver and Logemann (Citation2020), Guiver et al. (Citation2019), Haddad and Chellaboina (Citation2008), Jayawardhana et al. (Citation2009), Jayawardhana et al. (Citation2011), Khalil (Citation2002), Leonov (Citation2001), Sarkans and Logemann (Citation2015), Sarkans and Logemann (Citation2016b), Vidyasagar (Citation2002), and Yakubovich et al. (Citation2004), usually referred to as absolute stability theory (Carrasco & Heath, Citation2021). Despite originating in the late 1940s, and subject to intensive study since, absolute stability theory is still an active area of research, with recent papers including, for example, Boiko et al. (Citation2022), Drummond et al. (Citation2022), and Guiver and Logemann (Citation2022). Typically, Lur'e systems are considered in a state-space setting or in the functional analytic input-output framework initiated by Sandberg, Zames and other researchers in the 1960s. In much of the literature on state-space theory of Lur'e systems, asymptotic stability properties of unforced Lur'e systems are studied, but the potential effects of persistent external inputs on the asymptotic behaviour of the nonlinear closed-loop dynamics have received relatively little attention (exceptions include Arcak & Teel, Citation2002; Franco et al., Citation2021; Gilmore et al., Citation2021; Guiver & Logemann, Citation2020; Guiver et al., Citation2019; Jayawardhana et al., Citation2009Citation2011; Sarkans & Logemann, Citation2015).

In contrast, here we consider forced Lur'e systems defined by higher-order differential equations of the form (1) P(D)y=Q(D)u+Qe(D)v,u=f(y),(1) where P, Q and Qe are polynomial matrices, D is the operator d/dt (derivative with respect to time), u is an input used for feedback, v is an external input, y is the output and f is a nonlinearity. It is assumed that detP(s)0 and that the rational matrices P1Q and P1Qe are proper. Under these conditions, the linear system (2) P(D)y=Q(D)u+Qe(D)v=(Q(D),Qe(D))(uv)(2) is an input-output system in the sense of the behavioural approach to systems and control (Polderman & Willems, Citation1998, Section 3.3). The systematic investigation of linear differential systems described by polynomial matrices, including models of the form (Equation2), goes back to Rosenbrock (Citation1970) and his work was further developed in algebraic systems theory, see, for example, Blomberg and Ylinen (Citation1983). Textbooks such as Kailath (Citation1980), Kwakernaak and Sivan (Citation1991), and Polderman and Willems (Citation1998) contain detailed discussions of models of the form (Equation2).

It is somewhat surprising that there seems to be hardly any literature on higher-order Lur'e systems, with Brockett and Willems (Citation1965a), Brockett and Willems (Citation1965b), Pendharkar and Pillai (Citation2008), Sarkans and Logemann (Citation2016a), and Willems (Citation1970) being the exceptions we are aware of. The papers (Brockett & Willems, Citation1965aCitation1965b; Pendharkar & Pillai, Citation2008) and Willems (Citation1970, Chapter 6) study stability properties of certain continuous-time higher-order Lur'e systems which are unforced and single-input single-output (that is, in (Equation1), v = 0 and the functions u and y are scalar valued). The paper (Sarkans & Logemann, Citation2016a) develops an input-to-output stability theory for discrete-time higher-order multi-input multi-output Lur'e systems in the presence of forcing.

The current paper develops a theory of continuous-time input-output Lur'e systems (Equation1), similar to the discrete-time framework of Sarkans and Logemann (Citation2016a), which is sufficiently general to accommodate not only input-to-output stability results but also the strong integral input-to-output stability property (not covered in Sarkans & Logemann, Citation2016a). This latter property is the input-output counterpart to strong integral input-to-state stability concept for state-space systems (Chaillet et al., Citation2014; Guiver & Logemann, Citation2020). Whilst it is straightforward to define a natural trajectory concept in the discrete-time case, in the continuous-time setting this is more subtle because it is desirable to allow for non-smooth signals, requiring a suitable notion of weak trajectories.

For Lur'e systems of the form (Equation1), we consider input-to-output stability concepts which are similar in spirit to the well-known state-space notion of input-to-state stability (ISS) (Dashkovskiy et al., Citation2011; Sontag, Citation1989Citation2008) and the state-independent input-to-output stability property (Sontag & Wang, Citation1999). The main result of the paper is reminiscent of the complexified Aizerman conjecture (Gilmore et al., Citation2021; Guiver & Logemann, Citation2020; Hinrichsen & Pritchard, Citation1992Citation2010; Jayawardhana et al., Citation2011; Sarkans & Logemann, Citation2015Citation2016a, Citation2016b): we show that if (Equation1) is stable with v = 0 and f(y)=Ly for all complex matrices L such that LK<r (where the matrix K and scalar r>0 are fixed), then (Equation1) is input-to-output stable for all continuous nonlinearities f for which there exists a strictly increasing continuous comparison function α such that f(ξ)rξα(ξ) for all ξ, that is, stability for all linear complex feedbacks in an open ball implies stability for all nonlinear feedback functions satisfying a similar ‘nonlinear’ ball condition. As a corollary of this result, we obtain a version of the circle criterion (Gilmore et al., Citation2021; Guiver & Logemann, Citation2020; Haddad & Chellaboina, Citation2008; Jayawardhana et al., Citation2009Citation2011; Khalil, Citation2002; Sarkans & Logemann, Citation2015Citation2016aCitation2016b; Vidyasagar, Citation2002). These results comprise the main contribution of the current paper. The methods employed in the paper combine ideas from the behavioural approach to systems and control, absolute stability theory and input-to-state stability theory. We emphasise that our input-to-output stability framework should not be confused with the classical input-output concept of L-stability due to Sandberg and Zames (Desoer & Vidyasagar, Citation1975; Vidyasagar, Citation2002). Some more details on the results in individual sections can be found in the next paragraph.

The paper is organised as follows. In the next section, we will present and discuss a number of preliminaries and introduce some terminology and notation. In Section 3, we collect some ISS results for forced state-space Lur'e systems which will serve as key tools in the analysis of the stability behaviour of forced higher-order Lur'e system of the form (Equation1). Furthermore, Section 3 contains a number of well-posedness criteria for state-space Lur'e systems with non-zero feedthrough. We allow for non-zero feedthrough in the linear component of the state-space Lur'e system because, to enable applications to the input-output setting in Section 5, the state-space framework should be sufficiently general to capture all input-output systems in the sense of behavioural systems theory (Polderman & Willems, Citation1998, Section 3.3) which includes systems with non-zero feedthrough. The inclusion of non-zero feedthrough in the state-space setting gives rise to a system of coupled nonlinear differential and algebraic equations, and requires a careful well-posedness analysis of the associated initial value problem. Section 4 is devoted to the development of results relating to the behaviours and state-space realisations of linear input-output systems of the form (Equation2). To avoid restrictive smoothness assumptions for the forcing function, behaviours are defined as sets of weak trajectories which are defined in terms of distribution theory. A natural characterisation of weak trajectories as solutions of a suitably integrated version of (Equation2) is also provided. The key result on state-space realisations (Theorem 4.6) shows that, under the assumption of properness of P1Q and P1Qe, there exists a state-space realisation of (Equation2) with the property that its full behaviour is isomorphic (in the vector space sense) to the behaviour of (Equation2) (the isomorphism being induced by a certain ‘canonical’ map which enjoys certain useful boundedness properties). Moreover, stabilizability and controllability properties of the realisation correspond nicely to natural conditions in terms of the polynomial matrices P, Q and Qe. In Section 5, we extend the weak trajectory concept to the nonlinear system (Equation1), relate the behaviour of (Equation1) to that of an associated Lur'e system in state-space form (Proposition 5.1), and develop a stability theory for forced input-output Lur'e systems of the form (Equation1) which is inspired by the complexified Aizerman conjecture and the classical circle criterion (Theorem 5.2 and Corollary 5.4). Three examples are discussed in Section 6, and concluding remarks appear in Section 7. Finally, to avoid breaking the flow of the presentation, the proofs of two results in Section 2 are relegated to the Appendix.

2. Notation, terminology and preliminary results

Most mathematical notation we use is standard. Set N:={1,2,3,} and N0:=N{0}, and let Z be the ring of all integers. We denote by R and C the fields of real and complex numbers, respectively. We also define R+:=[0,) and C0:={sC:Res>0}. For KCm×p and r>0, we define the open ball in Cm×p with centre K and radius r: BC(K,r):={MCm×p:MK<r},where the operator norm is induced by the 2-norms in Cp and Cm. For vectors z1,,zlCp, lN, we write col1il(zi):=(z1zl)Clp.

2.1. Polynomial and rational matrices

Next we provide some preliminary material relating to polynomial and rational matrices. We will be brief and refer to Delchamps (Citation1988), Fuhrmann and Helmke (Citation2015), and Kailath (Citation1980) for more details. The ring of polynomials with coefficients in R is denoted by R[s]. For a polynomial matrix PR[s]p×m given by P(s)=j=0kPjsj, where PjRp×m with Pk0, we say that the degree of P is equal to k and write degP=k. The degree of the zero polynomial matrix is defined to be .The ith row degree ri(P) of P is the degree of the polynomial row vector given by the ith row of P, or, equivalently, ri(P)=max1jmdegPij, where PijR[s] is the polynomial in the ith row and jth column of P. Obviously, degP=max1ipri(P). Note that, for a square polynomial matrix PR[s]p×p, degdetPi=1pri(P), and P is said to be row reduced (or row proper) if degdetP=i=1pri(P). A square polynomial matrix PR[s]p×p is called unimodular if detP(s)c for some non-zero constant c, or equivalently, if P has an inverse in R[s]p×p.

The following lemma is well known, see, for example, Fuhrmann and Helmke (Citation2015, Proposition 2.19 and Theorem 2.20) and Kailath (Citation1980, pp. 384).

Lemma 2.1

Let PR[s]p×p be such that detP(s)0.

(1)

There exists unimodular UR[s]p×p such that UP is row reduced.

(2)

The polynomial matrix is row reduced if, and only if, the limit matrix R:=lim|s|diag(sr1(P),,srp(P))P(s)is invertible.

Denoting the entries of P and R by Pij and Rij, respectively, and the leading coefficient of Pij by pij, then Rij=pij if degPij=ri(P) and Rij=0 if degPij<ri(P). In the literature, the matrix R is therefore sometimes referred to as the ‘highest-row-degree coefficient matrix’ of P.

A square polynomial matrix PR[s]p×p is said to be Hurwitz if detP(s)0 for all sC¯0, where C¯0 is the closure of C0 (that is, C¯0 is the closed-right half of the complex plane).

The field of rational functions with coefficients in R is denoted by R(s). For a rational function r=n/d, where n,dR[s], d(s)0, the difference degrelr:=degddegn is said to be the relative degree of r. In particular, the relative degree of the zero rational function is equal to ∞. For a rational matrix RR(s)p×m, the relative degree degrelR of R is defined to be the minimum of the relative degrees of its non-zero entries. According to this definition, the degree of the zero rational matrix is equal to ∞. Note that if R is not the zero matrix, then the relative degree of R is equal to dZ if, and only if, the limit of sdR(s), as |s|, exists in Rp×m and is not equal to 0. We say that R is proper if degrelR0. If the inequality is strict, then R is said to be strictly proper. A rational matrix RR(s)p×m is in the Hardy space H(Cp×m) of all bounded holomorphic functions C0Cp×m if, and only if, R is proper and does not have any poles in the closed right-half plane, in which case we define RH:=supsC0R(s).For a square polynomial matrix PR[s]p×p, we introduce the concept of left degree of P which is defined as follows: degleftP:=min{deg(LP):LR[s]p×ps.t.detL(s)0}.Trivially, degleftPdegP. The concept of left degree seems to be new–at least we were not able to find it in the literature.

Some of the relationships between the various degrees are highlighted in the following result, the proof of which can be found in the Appendix.

Proposition 2.2

Let PR[s]p×p and QR[s]p×m be such that detP(s)0 and set G:=P1Q. The following statements hold.

(1)

degrelGdegPdegQ.

(2)

If UR[s]p×p is unimodular and UP is row reduced, then deg(UP)=degleftP.

(3)

degleftP=min{deg(UP):UR[s]p×punimodular}.

(4)

degleftPdegdetP.

(5)

If Q(s)0, then degleftPdegrelG.

(6)

If degleftP=0, then degdetP=0, that is, P is unimodular.

Obviously, in the scalar case (that is, p = m = 1) equality holds in statement (1). Simple examples show that, in general, the identity degrelG=degPdegQ does not hold. As for statement (2), note that statement (1) of Lemma 2.1 guarantees the existence of a unimodular matrix U such that UP is row reduced.

2.2. Functions and function spaces

The symbols ★ and denote convolution and composition of functions, respectively. For 1p and 0<τ, let Lp([0,τ),Rn) denote the usual Lebesgue space of functions defined on [0,τ) with values in Rn. The local version of Lp([0,τ),Rn) is denoted by Llocp([0,τ),Rn). As usual, C([0,τ),Rn) is the vector space of continuous functions defined on [0,τ) with values in Rn, and the space of l-times continuously differentiable functions [0,τ)Rn is denoted by Cl([0,τ),Rn). We set C0([0,τ),Rn):=C([0,τ),Rn). Further, for lN, 1q and 0<τ, we define Wlocl,q([0,τ),Rn) to be the space of all xCl1([0,τ),Rn) such that x(l1) is (locally) absolutely continuous and x(l)=(d/dt)x(l1)Llocq([0,τ),Rn). It is convenient to set Wloc0,q([0,τ),Rn):=Llocq([0,τ),Rn). We remark that Wlocl,q([0,τ),Rn) can be identified with the space {xWlocl,q((0,τ),Rn):x,Ddx,,DdlxLlocq([0,τ),Rn)},where Wlocl,q((0,τ),Rn) is the local version of the Sobolev space Wl,q((0,τ),Rn) of regular distributions on (0,τ) and Dd denotes the distributional derivative. Note that Wlocl,q([0,τ),Rn)Wlocl,q((0,τ),Rn), for example, the function x defined by x(t):=t for t0 is in Wlocl,1((0,),R) for all lN0, whilst xWlocl,1([0,),R) if, and only if, l = 0, 1.

The space of real-valued Bohl functions defined on R+ is denoted by B(R+,R), that is, B(R+,R) is the space of all finite linear combinations of functions of the form ttkRe(eζtz), where kN0, ζC and zC. The space of Rn×m-valued Bohl functions can be defined in a similar way, is denoted by B(R+,Rn×m) and can be identified with B(R+,R)n×m. We set B(R+,Rn):=B(R+,Rn×1).

For the convenience of the reader, we recall the definitions of three classes of comparison functions. The sets K and K are defined by K:={αC(R+,R+):α(0)=0 and α is strictly increasing},K:={αK:limsα(s)=},and KL denotes the set of functions β:R+×R+R+ with the following properties: β(,t)K for every tR+, and β(z,) is non-increasing with limtβ(z,t)=0 for every z0. For more details on comparison functions, we refer the reader to Kellett (Citation2014).

Finally, we state a lemma about the potential smoothing effect of convolution operators induced by proper rational matrices. It should be well-known, but we could not find a suitable statement in the literature. For completeness, we provide a proof in the Appendix.

In the following, let δ be the Dirac distribution (delta function) supported at 0.

Lemma 2.3

Let GR(s)p×m be non-zero and proper with inverse Laplace transform denoted by G. Set d:=degrelG and G0:=GG()δB(R+,Rp×m). If uWlocl,q(R+,Rm) for some lN0 and 1q, then the convolution Gu is in Wlocl+d,q(R+,Rp) and (3) (Gu)(l+d)={G0(l)u+G()u(l)+i=0l1G0(l1i)(0)u(i),ifd=0,G(l+d)u+i=0lG(l+d1i)(0)u(i),ifd1.(3)

3. Input-to-state stability for state-space Lur'e systems

In this section, we consider controlled Lur'e systems in state space and record a stability result which will play an important role for the theory of input-output Lur'e systems developed in Section 5. For fixed m,me,pN and variable nN, define Σssn:=Rn×n×Rn×m×Rn×me×Rp×n×Rp×m×Rp×me.With S:=(A,B,Be,C,D,De)Σssn, we associate the following controlled and observed linear state-space system (4) x˙=Ax+Bu+Bev,y=Cx+Du+Dev.(4) Below, the input u will be used for nonlinear output feedback of the form u=f(y) and v will be an an external input to the nonlinear feedback system. Frequently, we will refer to (Equation4) as the system S=(A,B,Be,C,D,De). We let G, given by G(s)=C(sIA)1B+D, denote the transfer function of the linear state-space system (A,B,C,D) (equivalently, of (Equation4) with v = 0).

The behaviour B(S) of S (or of (Equation4)) is the linear subspace of all quadruples (u,v,x,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Wloc1,1(R+,Rn)×Lloc(R+,Rp)which satisfy (Equation4) almost everywhere on R+ (as we are interested in stability properties, we restrict attention to forward time). The elements of B(S) are called trajectories of S.

A matrix KCp×m is said to be a stabilising (complex) output feedback matrix for (A,B,C,D) if 1 is not an eigenvalue of DK and all eigenvalues of the matrix A+BK(IDK)1C have negative real parts, the set of which we denote by Sss(A,B,C,D).

If (A,B) is stabilizable and (C,A) is detectable, then, as is well-known, Sss(A,B,C,D) can be characterised in terms of the transfer function G as follows: (5) Sss(A,B,C,D)={KCm×p:det(IKG(s))0andG(IKG)1H(Cp×m)}.(5) Application of the nonlinear feedback u=f(y) to the state-space system (Equation4), yields the following closed-loop system (6) x˙=Ax+B(fy)+Bev,y=Cx+D(fy)+Dev,(6) which will be denoted by Sf:=(A,B,Be,C,D,De,f), where f:RpRm is continuous. The behaviour B(Sf) of Sf is the set of all triples (v,x,y)Lloc(R+,Rme)×Wloc1,1(R+,Rn)×Lloc(R+,Rp)such that (v,x,y) satisfies (Equation6) for almost every t0. Equivalently, (v,x,y)B(Sf) if, and only if, (fy,v,x,y)B(S). Elements in B(Sf) will also be referred to as trajectories of Sf.

The following stability result will play an important role in Section 5.

Theorem 3.1

Let S=(A,B,Be,C,D,De)Σssn, f:RpRm be continuous, KRm×p and r>0. Assume that BC(K,r)Sss(A,B,C,D), rD(IKD)1<1 and there exists continuous α:R+R+ such that (7) f(ξ)rξα(ξ)ξRp.(7) The following statements hold.

(1)

If αK, then there exist βKL and γK such that, for all (v,x,y)B(Sf), (8) x(t)β(x(0),t)+γ(vL(0,t))t0.(8)

(2)

If αK, then there exist β,ψKL, γ,ϕ,θK and b>0 such that (Equation8) holds for all (v,x,y)B(Sf) with vLb and (9) x(t)ψ(x(0),t)+ϕ(0tθ(v(τ))dτ)t0,(9) is satisfied for all (v,x,y)B(Sf).

Inequalities (Equation8) and (Equation9) say that the Lur'e system (Equation6) is ISS and integral ISS, respectively (Dashkovskiy et al., Citation2011; Sontag, Citation2008). Statement (2) is a criterion for strong integral ISS (Chaillet et al., Citation2014; Guiver & Logemann, Citation2020) of (Equation6). Proofs of statements (1) and (2) can be found in Sarkans and Logemann (Citation2015) and Guiver and Logemann (Citation2020), respectively, both in the special case when D = 0. A careful analysis of these proofs shows that they carry over to the non-zero feedthrough case D0, provided that rD(IDK)1<1. As has been mentioned in the Introduction, we allow for non-zero feedthrough D because, for later purposes (see Section 5), we wish the state-space framework to be sufficiently general to capture all input-output systems in the sense of behavioural systems theory (Polderman & Willems, Citation1998, Section 3.3) including those with non-zero feedthrough.

If KSss(A,B,C,D) and G(s)0, then it is well-known that the largest possible value for r>0 such that BC(K,r)Sss(A,B,C,D) is equal to 1/G(IKG)1H. It follows that if BC(K,r)Sss(A,B,C,D), then the condition rD(IKD)1<1 is only violated if r=1/G(IKG)1H and D(IKD)1=G(IKG)1H (the latter identity means that G(IKG)1 attains its H-norm at ∞). In the (not very interesting) case wherein G(s)0 (and hence, D = 0), we have that BC(K,r)Sss(A,B,C,D) and rD(IKD)1<1 for every r>0, and it follows that the conclusions of Theorem 3.1 hold for every linearly bounded nonlinearity f.

We say that (v,x,y)Lloc(R+,Rme)×Wloc1,1([0,τ),Rn)×Lloc([0,τ),Rp),where 0<τ, is a pre-trajectory of Sf if (v,x,y) satisfies (Equation6) for almost every t[0,τ). The triple (v,x,y) is said to be maximally defined if there does not exist another pre-trajectory (v,x~,y~)Lloc(R+,Rme)×Wloc1,1([0,σ),Rn)×Lloc([0,σ),Rp) such that σ>τ and (x~,y~)|[0,τ)=(x,y). A routine argument based on Zorn's lemma shows that every pre-trajectory of Sf can be extended to a maximally defined pre-trajectory of Sf. The set of all maximally defined pre-trajectories of Sf is denoted by B~(Sf). Obviously, B(Sf)B~(Sf), that is, every trajectory is also a (maximally defined) pre-trajectory.

In Theorem 3.1, the triples (v,x,y) under consideration are assumed to be trajectories that is, elements in B(Sf) (and hence defined on the half line R+). Statement (3) of Proposition 3.2 below shows that, under certain conditions, the assumptions of Theorem 3.1 imply that B~(Sf)=B(Sf). Note that nonzero feedthrough D can cause issues with well-posedness (that is, existence and uniqueness of solutions to initial-value problems associated with (Equation6)), because (Equation6) is a ‘mixture’ of nonlinear differential and algebraic equations if D0 (see Examples (3.3) and (3.4) below). Proposition 3.2 provides natural sufficient condition guaranteeing well-posedness in the presence of nonzero feedthrough D.

Proposition 3.2

Let S=(A,B,Be,C,D,De)Σssn and KRm×p. Let f:RpRm be locally Lipschitz and such that fD:RpRp defined by (10) fD(ξ):=Df(ξ)ξRp(10) is of class C1. The following statements hold.

(1)

If det(IfD(ξ))0 for all ξRp, IDK is invertible and there exist a0 and b[0,1) such that (11) (IDK)1D(f(ξ))a+bξξRp,(11) then, for every vLloc(R+,Rme) and every x0Rn, there exists a unique pair (x,y) such that (v,x,y)B~(Sf) and x(0)=x0.

(2)

If f(ξ)=O(ξ) as ξ, (12) supξRpfD(ξ)<andinfξRp|det(IfD(ξ))|>0,(12) then B~(Sf)=B(Sf), and, for every vLloc(R+,Rme) and every x0Rn, there exists a unique pair (x,y) such that (v,x,y)B(Sf) and x(0)=x0.

(3)

If det(IfD(ξ))0 for all ξRp and there exist r>0 and αK such that BC(K,r)Sss(A,B,C,D), rD(IKD)1<1 and (Equation7) is satisfied, then B~(Sf)=B(Sf) and, for all vLloc(R+,Rme) and x0Rn, there exists a unique pair (x,y) such that (v,x,y)B(Sf) and x(0)=x0.

Note that the conditions involving fD and D(IKD)1=(IDK)1D are trivially satisfied when D = 0.

The following simple examples show that, in general (in the absence of the hypotheses of Proposition 3.2), even if f is globally Lipschitz, we may have B(Sf)B~(Sf), and, for given vLloc(R+,Rme) and x0Rn, the set of pre-trajectories (v,x,y)B~(Sf) satisfying x(0)=x0 may be empty or contain several elements.

Example 3.3

Consider the closed-loop system (Equation6) with A=(1000),B=Be=(01),C=(1,0),D=De=1,and nonlinearity f given by f(ξ)={ξ+1,ξ<2,ξ/2,2ξ2,ξ1,ξ>2.It is easy to check that CeAtB0 and we note that (13) ξf(ξ)=ξDf(ξ)={1,ξ<2,ξ/2,2ξ2,1,ξ>2.(13) Let x0=(a,0)T, aR. If (0,x,y) is a pre-trajectory defined on some interval [0,τ) and such that x(0)=x0, then y satisfies y(t)=CeAtx0+Df(y(t)+0tCeA(ts)Bf(y(s))ds=aet+f(y(t))t[0,τ),and thus (14) y(t)f(y(t))=aett[0,τ).(14)

  1. Invoking (Equation13), we see that (Equation14) does not have a solution for any t0 if |a|>1. Hence, there does not exist any pre-trajectory (0,x,y) such that x(0)=(a,0)T if |a|>1.

  2. Let now a=1/2. In this case, we see that (Equation14),

    • has the unique solution y(t)=et for every t[0,ln2),

    • is solved by y(t)=ξ for every ξ2, when t=ln2,

    • does not have a solution for all t>ln2.

Setting y(t):=et for all t[0,ln2) and x(t):=eAtx0+0teA(ts)Bf(y(s))ds=(et/2(et1)/2)t[0,ln2),we conclude that (0,x,y) is the unique maximally defined pre-trajectory satisfying x(0)=(1/2,0)T. Note that although the pre-trajectory (0,x,y) is bounded on [0,ln2), it cannot be continued to the right beyond ln2.

Example 3.4

Consider the scalar system x˙=x+fy+v,y=x+fy+v,which is a special case of (Equation6) with A = −1 and B=Be=C=D=De=1. With f given by f(ξ)=ξ(1ξ) and x1(t)1/4, y1(t)1/2 and x2(t)={(et1/2)2,0tln2,0,t>ln2,y2(t)=x2(t)={1/2et,0tln2,0,t>ln2,it is easy to check that (0,x1,y1),(0,x2,y2)B(Sf). As x1(0)=1/4=x2(0), we see that there are multiple trajectories of the form (0,x,y) satisfying x(0)=1/4.

Proof

Proof of Proposition 3.2

To start, we define the map F:RpRp by F(ξ):=ξDf(ξ)=ξfD(ξ) for all ξRp which will play a key role in the proof.

To prove statement (1), let x0Rn and vLloc(R+,Rme). It is sufficient to show that F is a C1-diffeomorphism. Indeed, in this case, it follows from (Equation6) that every (v,x,y)B~(Sf) satisfies y=F1(Cx+Dev) and (15) x˙=Ax+Bg(x,v)+Bev,whereg(x,v):=(fF1)(Cx+Dev)(15) on the interval on which (v,x,y) is defined. As f is locally Lipschitz and F1 is of class C1, the map (t,z)g(z,v(t)) is measurable in t for fixed z and locally Lipschitz in z in the sense that, for every compact set ΓRn and every τ>0, there exists L>0 and a set Θ[0,τ] of zero measure such that g(z1,v(t))g(z2,v(t))Lz1z2z1,z2Γ,t[0,τ]∖Θ,where we have used that v is locally essentially bounded. Consequently, by a standard result in the theory of ordinary differential equations, (Equation15) has a unique maximally defined solution x satisfying x(0)=x0, and statement (1) follows.

It remains to prove that F is a C1-diffeomorphism. For this purpose, we define the map H:RpRp by H(ξ)=ξ(IDK)1D(f(ξ)) and show that H is a C1-diffeomorphism. As H(ξ)=(IDK)1F(ξ) for all ξRp, it then follows that F is also a C1-diffeomorphism. Since H(ξ)=(IDK)1F(ξ)=(IDK)1(IfD(ξ))ξRp,we obtain that detH(ξ)0 for all ξRp, and thus, by the local inversion theorem (see, for instance, Ambrosetti and Prodi (Citation1993, Theorem 1.2 in Chapter 2), H is a local C1-diffeomorphism. Furthermore, using that (Equation11) holds with b<1, we see that H(ξ) as ξ, and the global inverse function theorem (see, for example, Ambrosetti & Prodi, Citation1993, Theorem 1.8 in Chapter 3 or Sandberg, Citation1980, Theorem 2) now implies that H is a C1-diffeomorphism.

We proceed to prove statement (2). By Cramer's rule, (F(ξ))1=1det(F(ξ))adjugate(F(ξ))ξRp,which combined with (Equation12) shows that the function ξ(F(ξ))1 is bounded. Thus, by the Hadamard-Levy theorem (see Deimling, Citation1985, Theorem 15.4 or De Marco et al., Citation1994, Theorem 0.2), F is a C1-diffeomorphism, and we can argue as in the proof of statement (1) to show that, for every vLloc(R+,Rme) and every x0Rn, there exists a unique pair (x,y) such that (v,x,y)B~(Sf) and x(0)=x0. As (F1)(ξ)=[F(F1(ξ))]1ξRp,we see that the function ξ(F1)(ξ) is bounded, implying that F1 is globally Lipschitz, and thus (fF1)(ξ)=O(ξ) as ξ. Therefore, there exists c0 such that (16) Bg(z,v(t))+Bev(t)c(1+v(t)+z)zRn,vLloc(R+,Rme),a.e.t0,(16) with g defined as in (Equation15). Let (v,x,y)B~(Sf) and let [0,τ) be the (maximal) interval on which (v,x,y) is defined. We have to show that τ=. By (Equation15) and the variation-of-parameters formula, we have x(t)=eAtx(0)+0teA(ts)(Bg(x(s),v(s))+Bev(s))dst[0,τ).A routine argument based on the Gronwall lemma and (Equation16) shows that x is bounded if τ<, but that is not possible, whence τ=.

Finally, to prove statement (3). It follows from the hypotheses that (Equation11) is satisfied with a = 0 and b=r(IDK)1D. Hence, invoking statement (1), we conclude that for every vLloc(R+,Rme) and every x0Rn, there exists a unique pair (x,y) such (v,x,y)B~(Sf) and x(0)=x0. Let (v,x,y)B~(Sf) with (maximal) interval of definition [0,τ). It remains to show that τ=. Exploiting the hypotheses once more, we see that the map F is radially unbounded and detF(ξ)0 for all ξRp, hence F is a C1-diffeomorphism, and so x satisfies (Equation15). The Lyapunov analysis developed in the proof of the stability result Theorem 3.1 (see Guiver & Logemann, Citation2020; Sarkans & Logemann, Citation2015) shows that x is bounded on [0,τ). As x is a maximally defined solution of (Equation15), we conclude that τ=, and consequently, the triple (v,x,y) is in B(Sf).

4. Linear input-output systems: key concepts and results

Before discussing input-output Lur'e systems in Section 5, we consider a key constituent ingredient, namely, linear input-output systems of the form (Equation2) which relate input and outputs by means of higher-order differential equations. More formally, let Σio be the following subset of R[s]p×p×R[s]p×m×R[s]p×me: (P,Q,Qe)Σio if, and only if, detP(s)0 and both G:=P1Q and Ge:=P1Qe are proper. Let (P,Q,Qe)Σio and set k:=degP, so that P(s)=j=0kPjsjfor suitable matrices PjRp×p, where Pk0. Note that, since G and Ge are proper, degQk and degQek. Consequently, Q(s)=j=0kQjsjandQe(s)=j=0kQejsjfor suitable matrices Qj and Qej. Obviously, if degQ<k or degQe<k, then Qk=0 or Qek=0, respectively.

With (P,Q,Qe)Σio, we associate the following linear input-output system (17) P(D)y=Q(D)u+Qe(D)v=(Q(D),Qe(D))(uv)(17) or, equivalently, (18) j=0kPjy(j)(t)=j=0kQju(j)(t)+j=0kQejv(j)(t)t0,(18) where u is an input available for feedback, v is an external input and y is an output. As we are interested in stability theory, we have chosen R+ as the time domain in (Equation18).

In this section, we shall introduce a suitable weak trajectory concept for (Equation17), define the behaviour of (Equation17) in terms of weak trajectories, analyse the structure of weak trajectories (see Proposition 4.3) and prove the existence of a state-space realisation S of (Equation17) such that the behaviours of S and (Equation17) are isomorphic under a natural isomorphism (see Theorem 4.6).

Obviously, interpreted in the classical sense, (Equation17) and (Equation18) are only meaningful if u, v and y are sufficiently often differentiable. Since it is desirable to allow for discontinuous inputs u and v (step functions, for example), it would be restrictive to impose any smoothness assumptions on u and v. However, in the absence of suitable differentiability properties, it is still possible to make sense of (Equation17) and (Equation18) by using basic ideas from distribution theory. To this end, let wLloc1(R+,Rl), where lN, and let w~ denote the extension by zero of w to all of R. In the following, the function w~ and the regular Rl-valued distribution on R induced by w~ will be identified. Locally integrable functions and the associated regular distributions will not be distinguished notationally. We say that the triple (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp) is a weak trajectory of (Equation17) if there exist diRp, i=0,,k1, such that (19) P(Dd)y~Q(Dd)u~Qe(Dd)v~=i=0k1diDdiδ,(19) where Dd denotes the distributional derivative and δ is the Dirac distribution. The behaviour B(P,Q,Qe) of (Equation17) is defined to be the set of all weak trajectories of (Equation17). It is obvious that B(P,Q,Qe) is a linear subspace of Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp), and any ‘classical trajectory’ (u,v,y) (in the sense that (u,v,y) is sufficiently often differentiable for (Equation18) to hold for almost every t0) is a weak trajectory.

We present a simple example of a weak trajectory which is not classical.

Example 4.1

Consider (Equation17) with P(s)=s+1, Q(s)=s and Qe(s)1. Let τ>0, aR and define y(t):={aet,0t<τ,(aeτ)et,tτ,u(t):={0,0t<τ,1,tτ,v(t):=0t0.Then Ddy~=δτy~ and Ddu~=δτ, where δτ is the Dirac distribution supported at τ. Consequently, P(Dd)y~Q(Dd)u~Qe(Dd)v~=(Dd+1)y~+Ddu~=,showing that (u,v,y) is a weak trajectory. It is obvious that (u,v,y) is not a classical trajectory.

We now briefly explain how the concept of a weak trajectory can be characterised in terms of an integrated version of (Equation17). For this purpose set (Iz)(t):=0tz(τ)dτtR,zLloc1(R,Rn),whereR=R+orR=R,and define polynomial matrices associated with P, Q and Qe as follows: (20) Pw(s):=skP(1/s),Qw(s):=skQ(1/s)andQew(s):=skQe(1/s).(20)

Proposition 4.2

A triple (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp) is a weak trajectory of (Equation17) if, and only if, there exist ciRp, i=0,,k1, such that (21) (Pw(I)y)(t)(Qw(I)u)(t)(Qew(I)v)(t)=i=0k1citia.e.t0.(21)

We remark that in Polderman and Willems (Citation1998) a weak trajectory concept is defined via (Equation21): as the ‘bilateral’ trajectories (trajectories defined on R) are considered in Polderman and Willems (Citation1998), the equation in (Equation21) is required to hold for almost every tR.

Proof

Proof of Proposition 4.2

Let θ:RR denote the Heaviside function, that is, θ(t):={0t<0,1t0.Let (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp) and assume that (Equation21) holds. It then follows that (Pw(I)y~)(t)(Qw(I)u~)(t)(Qew(I)v~)(t)=θ(t)i=0k1citia.e.tR.Taking the kth distributional derivative of this identity results in (Equation19) with di=ck1i(k1i)! for i=0,,k1. Hence, (u,v,y)B(P,Q,Qe)

Conversely, let (u,v,y)B(P,Q,Qe), and so, (Equation19) holds. By a well-known result of distribution theory (see, for example, Zemanian, Citation1987, Theorem 2.6-1), any two primitives of a distribution on R differ by a constant distribution, and consequently, invoking (Equation19), we conclude that there exists γ0Rp such that (22) P0Iy~+j=1kPjDdj1y~Q0Iu~j=1kQjDdj1u~Qe0Iv~j=1kQejDdj1v~=d0θ+i=1k1diDdi1δ+γ0.(22) As the distribution d0θ+i=1k1diDdi1δ and the distribution on the left-hand side of (Equation22) have their supports in R+, the same must apply to the constant distribution γ0, and thus γ0=0. Taking primitives on both sides of (Equation22) and continuing with this process, we obtain that (Pw(I)y~)(t)(Qw(I)u~)(t)(Qew(I)v~)(t)=i=0k1diθ(t)(k1i)!tk1ia.e.tR.Restricting attention to the half-line R+, we see that (Equation21) holds with ci=dk1i/(i!).

Proposition 4.3 below shows that B(P,Q,Qe) and kerP(D) are closely related, where kerP(D) is defined in the classical sense, that is, kerP(D):={zB(R+,Rp):j=0kPjz(j)(t)=0t0}.Furthermore, we set kerweakP(D):={yLloc(R+,Rp):(0,0,y)B(P,Q,Qe)},and note that a function yLloc(R+,Rp) is in kerweakP(D) if, and only if, there exist diRp, i=0,,k1, such that P(Dd)y~=i=0k1diDdiδ.

Let G and Ge denote the inverse Laplace transforms (impulse responses) of G=P1Q and Ge=P1Qe, respectively. We note that the entries of G and Ge are elements in B(R+,R)+Rδ.

Proposition 4.3

Let (P,Q,Qe)Σio. The following statements hold.

(1)

dimkerP(D)=degdetP.

(2)

(u,v,Gu+Gev)B(P,Q,Qe) for all (u,v)Lloc(R+,Rm)×Lloc(R+,Rem).

(3)

(u,v,y)B(P,Q,Qe) if, and only if, yGuGevkerP(D).

(4)

kerweakP(D)=kerP(D).

(5)

If the triple (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp) is such that (23) support(P(Dd)y~Q(Dd)u~Qe(Dd)v~){0},(23) then (u,v,y)B(P,Q,Qe).

Proof.

Statement (1), sometimes referred to as Chrystal's theorem, is valid for all PR[s]p×p such that detP(s)0. Its proof (which is based on the Smith canonical form for polynomial matrices) can be found in, for example, Gohberg et al. (Citation1982, Theorem S1.6) and Polderman and Willems (Citation1998, Theorem 3.2.16).

To establish statements (2)–(4), let DIS+ denote the convolution algebra of real distributions on R with support contained in R+ and define P:=P(Dd)δ(DIS+)p×p,Q:=Q(Dd)δ(DIS+)p×m,Qe:=Qe(Dd)δ(DIS+)p×me.Since detP(s)0, the inverse P1 exists and is a matrix of real rational functions. Consequently, the inverse Laplace transform of P1, denoted by P1, is in (DIS+)p×p and is the inverse of P (with respect to convolution). The Laplace transforms of P1, P1Q(DIS+)p×m and P1Qe(DIS+)p×me are equal to P1, P1Q and P1Qe, respectively. Obviously, P1Q and P1Qe extend G and Ge to R, respectively. Now P1Q and P1Qe are proper real rational matrices and so the entries of P1Q and P1Qe are are of the form b~+, where bB(R+,R), aR, and, as before, ~ denotes the extension of the function by zero to all of R.

We proceed to prove statement (2), let (u,v)Lloc(R+,Rm)×Lloc(R+,Rme) and set y:=Gu+Gev. Then y~=(P1Q)u~+(P1Qe)v~, whence Py~Qu~Qev~=0. As P(Dd)y~=Py~, Q(Dd)u~=Qu~ and Qe(Dd)v~=Qev~, we conclude that P(Dd)y~Q(Dd)u~Qe(Dd)v~=0,showing that (u,v,y)B(P,Q,Qe).

To prove statement (3), let (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rme)×Lloc(R+,Rp) and assume that yGuGevkerP(D). Then (0,0,yGuGev)B(P,Q,Qe), and as (u,v,Gu+Gev)B(P,Q,Qe) by statement (2), linearity of the behaviour implies that (u,v,y)=(0,0,yGuGev)+(u,v,Gu+Gev)B(P,Q,Qe).Conversely, let us assume that (u,v,y)B(P,Q,Qe). Then there exist diRp, i=1,,k1, such that P(Dd)y~Q(Dd)u~Qe(Dd)v~=i=0k1diδ(i),or, equivalently, Py~Qu~Qev~=i=0k1diδ(i),from which it follows that (24) y~(P1Q)u~(P1Qe)v~=P1(i=0k1diδ(i)).(24) The entries of P1 are real rational functions and thus the entries of P1 are of the form (25) b~+i=0aiδ(i),wherebB(R+,R),aiRandai0for at most finitely many iN0.(25) Consequently, the components of P1(i=0k1diδ(i)) are also of the form (Equation25). But as the LHS of (Equation24) is locally essentially bounded, the non-regular distributional part of P1(i=0k1diδ(i)) must be equal to 0. Therefore, setting z:=yGuGev and noting that z~=y~(P1Q)u~+(P1Qe)v~, we conclude that zB(R+,Rp)C(R+,Rp). As Pz~=i=0k1diδ(i), or equivalently, P(Dd)z~=i=0k1diδ(i), it follows that (P(D)z)(t)=0 for all t>0, and, by continuity, this identity extends to [0,), showing that (P(D)z)(t)=0 for all t0, completing the proof of statement (3).

To establish statement (4), we note that if ykerP(D), then, trivially, yLloc(R+,Rp) and (0,0,y)B(P,Q,Qe), whence ykerweakP(D). Conversely, if ykerweakP(D), then (0,0,y)B(P,Q,Qe), and statement (3) implies that ykerP(D).

Finally, we prove statement (5). Let (u,v,y)Lloc(R+,Rm)×Lloc(R+,Rem)×Lloc(R+,Rp) be such that (Equation23) holds. By a well-known result from distribution theory (see, for example, Zemanian, Citation1987, Theorem 3.5-2), there exist lN0 and diRp, i=0,,l, such that P(Dd)y~Q(Dd)u~Qe(Dd)v~=i=0ldiδ(i),whereδ(i):=Ddiδ.Without loss of generality, we may assume that lk. We need to show that dk+i=0 for i=0,,lk. As in the proof of Proposition 4.2, by taking primitives on both sides of the above identity and repeating this process k-times, we obtain (26) Pw(I)y~Qw(I)u~Qew(I)v~g=i=0lkdk+iδ(i),(26) where Pw, Qw and Qew are given by (Equation20) and g(t):={i=0k1di(k1i)!tk1it00t<0.The distribution on the left-hand side of (Equation26) is regular and therefore i=0lkdk+iδ(i)=0, implying that dk+i=0 for i=0,,lk and completing the proof of statement (5).

We remark that statements (2)–(4) of Proposition 4.3 can also be proved by combining Proposition 4.2 with suitable modifications of arguments in Polderman and Willems (Citation1998), see Polderman and Willems (Citation1998, proofs of Theorems 3.2.15, 3.3.13 and 3.3.19).

Corollary 4.4

Let (P,Q,Qe)Σio.

(1)

If UR[s]p×p is unimodular, then B(UP,UQ,UQe)=B(P,Q,Qe).

(2)

If degleftP=0, then degdetP=0 (that is, P is unimodular), G(s)G(), Ge(s)Ge() and B(P,Q,Qe)={(u,v,G()u+Ge()v):(u,v)Lloc(R+,Rm)×Lloc(R+,Rem)}.

Proof.

Trivially, for unimodular U, we have G=(UP)1(UQ), Ge=(UP)1(UQe) and kerP(D)=ker(UP)(D), and thus, statement (1) is an immediate consequence of statement (3) of Proposition 4.3. To prove statement (2), we make use of statement (6) of Proposition 2.2, which implies that P is unimodular. Consequently, G=P1Q and Ge=P1Qe are polynomial matrices, which, combined with the properness of G and Ge, shows that G(s)G() and Ge(s)Ge(). An application of statement (1) with U=P1 yields that B(P,Q,Qe)=B(I,G(),G()). Since, trivially, B(I,G(),G())={(u,v,G()u+Ge()v):(u,v)Lloc(R+,Rm)×Lloc(R+,Rem)},the claim follows.

The next corollary follows from Lemma 2.3 and statement (3) of Proposition 4.3. Not surprisingly, the corollary shows that if (u,v,y)B(P,Q,Qe), then any smoothness properties of (u,v) are inherited by y and the smoothness enjoyed by y is enhanced by strict properness of G and Ge.

Corollary 4.5

Let (P,Q,Qe)Σio, assume that Q(s)0 and let d and de be the relative degrees of G and Ge, respectively. If (u,v,y)B(P,Q,Qe) with (u,v)Wlocl,q(R+,Rm)×Wlocle,q(R+,Rme), l,leN0 and 1q, then yWlocl0,q(R+,Rp), where l0:=min{l+d,le+de}.

Note that, as Q(s)0, we have d<, and so l0<. The assumption Q(s)0 avoids the occurrence of the uninteresting scenario in which the Lur'e system (Equation1) ‘degenerates’ into the linear system P(D)y=Qev.

For the rest of the paper we set :=degleftP,and define J(u,v,y):=col0i1((yGuGev)(i)(0))Rℓp(u,v,y)B(P,Q,Qe).Whilst, in general, point evaluations for the functions u, v or y and their derivatives do not make sense, the function yGuGev is in C(R+,Rp) (by statement (3) of Proposition 4.3) and so the definition of J(u,v,y) is meaningful. In the following, J(u,v,y) will play the role of an initial-value vector for the trajectory (u,v,y).

In order to apply the state-space results of Section 3 in the stability analysis of input-output Lur'e systems (see Section 5), we will prove the existence of suitable state-space realisations of the linear input-output system (P,Q,Qe). To this end, we define an n-dimensional realisation of (P,Q,Qe)Σio to be a state-space system S:=(A,B,Be,C,D,De)Σssn such that G(s)=P1(s)Q(s)=C(sIA)1B+D and Ge(s)=P1(s)Qe(s)=C(sIA)1Be+De. A realisation of minimal dimension is said to be a minimal realisation.

The following theorem is the main result of this section. It guarantees the existence of a realisation S of (P,Q,Qe)Σio such that the behaviours B(S) and B(P,Q,Qe) are isomorphic and the corresponding isomorphism has certain useful boundedness properties.

Theorem 4.6

Let (P,Q,Qe)Σio, define D:=G()=lim|s|G(s),De:=Ge()=lim|s|Ge(s)and set n:=degdetP. The following statements hold.

(1)

There exists an n-dimensional realisation S:=(A,B,Be,C,D,De) of (P,Q,Qe) such that (C,A) is observable, (27) kerP(D)={yz:zRn},whereyz(t):=CeAtzfor allt0,(27) and (28) ker(CCACA1)={0}.(28)

(2)

The map ι:RnkerP(D),zyz is a vector space isomorphism and there exists b>0 such that (29) ι1(w)bcol(w(0),w(0),,w(1)(0))wkerP(D).(29)

(3)

For every (u,v,x,y)B(S), the triple (u,v,y) is in B(P,Q,Qe), the map λ:B(S)B(P,Q,Qe),(u,v,x,y)(u,v,y)is a vector space isomorphism, and (30) x(0)bJ(u,v,y)(u,v,x,y)B(S),(30) where b is the constant from (Equation29).

(4)

Assume that Q(s),Qe(s)0 and let d and de denote the relative degrees of G and Ge, respectively. There exists c>0 such that, for every (u,v,x,y)B(S) with (u,v)Wlocd,q(R+,Rm)×Wlocde,q(R+,Rme), where 1q, the function y is in Wloc,q(R+,Rp) and (31) x(0)bJ(u,v,y)c(i=0d1u(i)(0)+i=0de1v(i)(0)+i=01y(i)(0)).(31)

(5)

Under the additional assumption that there exists LCm×p such that P(I+DL)QL is Hurwitz, the pair (A,B) is stabilizable.

(6)

Under the additional assumption that P and Q are left coprime, the pair (A,B) is controllable and S is a minimal realisation of (P,Q,Qe).

In the ‘extreme’ case wherein =0, it follows from statement (6) of Proposition 2.2 that P is unimodular, hence n = 0, and the conclusions of Theorem 4.6 hold trivially. Indeed, the 0-dimensional realisation S=(0,0,0,0,D,De) is controllable and observable, kerP(D)={0}, J(u,v,y)=0 for all (u,v,y)B(P,Q,Qe), B(S)={(u,v,0,Du+Dev):(u,v)Lloc(R+,Rm)×Lloc(R+,Rem)},and so, invoking statement (2) of Corollary 4.4, B(S)={(u,v,0,y):(u,v,y)B(P,Q,Qe)}.

By statement (4) of Proposition 2.2, n. If =n, then (Equation28) follows immediately from the observability of (C,A). But typically, we will have <n (for example, if Pk is invertible and p>1) and in that case (Equation28) provides additional information. Note that in statement (4), d0 and de0 as a consequence of statement (5) of Proposition 2.2. If d=0 (de=0), then the first (second) sum on the RHS of (Equation31) is defined to be equal to 0.

Proof

Proof of Theorem 4.6

Parts of the proof are similar to arguments used in the proof of the corresponding result in the discrete-time case (see Sarkans & Logemann, Citation2016a, Theorem 3.2). However, there are some significant differences (including properties which are not covered by Sarkans & Logemann, Citation2016a, Theorem 3.2) and we will focus on these. As pointed out in the above commentary, it may be assumed, without loss of generality, that 1. We proceed in two steps.

Step 1. In this step, we assume that P is row reduced and the row degrees ρi:=ri(P), where i=1,,p, satisfy ρ1ρ2ρp. It follows from statement (2) of Proposition 2.2 that =degP=k. Let 1pp be the unique integer such that ρi1 for all i=1,,p and ρi=0 for all i=p+1,,p. Since P is row reduced, we have that i=1pρi=degdetP=n. Following the argument in Sarkans and Logemann (Citation2016a), it can be shown that there exists an n-dimensional realisation S:=(A,B,Be,C,D,De) of (P,Q,Qe) such that (C,A) is observable and Ψ(s)[sIA,(B,Be)]=[P(s)C,Q(s)P(s)D,Qe(s)P(s)De],where Ψ(s):=(blockdiag1ip(sρi1,sρi2,,s,1)0)R[s]p×n.We remark that the so-called observer-form realisation (Kailath, Citation1980, Section 6.4) plays a key role in the construction of the realisation S.

To establish (Equation27), it is sufficient to show that (32) kerP(D){yz:zRn}.(32) Indeed, invoking Proposition 4.3, the dimension of kerP(D) is equal to n, and so, if (Equation32) holds, then kerP(D)={yz:zRn}, which is (Equation27).

To prove (Equation32), let wkerP(D). By Proposition 4.3, w is a Bohl function and, hence, has a Laplace transform, denoted w^. Taking the Laplace transform of the equality P(D)w=0, and using standard properties of the Laplace transform, gives P(s)w^(s)=j=1kPji=0j1sj1iw(i)(0)=l=0k1hl(w)sl,where hl:kerP(D)Rp is the linear map given by (33) hl(w):=j=l+1kPjw(jl1)(0)=j=lk1Pj+1w(jl)(0),0lk1.(33) Consequently, since detP(s)0, there exists zRn such that w=ι(z)=yz if, and only if, P(s)C(sIA)1z=l=0k1hl(w)sl.As Ψ(s)=P(s)C(sIA)1, we see that there exists zRn such that w=yz if, and only if, (34) (blockdiag1ip(sρi1,sρi2,,s,1)0)z=l=0k1hl(w)sl.(34) In the following, let hil denote the ith component of hl for 1ip. As in Sarkans and Logemann (Citation2016a), it can be shown that (35) hil=0for alli=1,,pandl=0,,k1such that ρi1<l(35) and (36) hil=0for alli=p+1,,pandl=0,,k1.(36) Setting σ1:=0 and σi:=j=1i1ρj,i=2,,p,the ith component of the left-hand side of (Equation34) is given by l=1ρisρilzσi+l for i=1,,p, where zj denotes the jth component of z. Consequently, (Equation34) can be written in the form l=1ρisρilzσi+l=j=0k1hij(w)sj,i=1,,p.We conclude that (Equation34) has a unique solution z=col(z1,,zn)Rn which is given by (37) zσi+l=hiρil(w),1ip,1lρi,(37) where we have used (Equation35) and (Equation36). We have now established (Equation32), and so, as has already been pointed out, (Equation27) follows. It is a straightforward consequence of (Equation27) that the map ι:RnkerP(D),zyz is an isomorphism. Invoking (Equation33) and (Equation37) shows that there exists b>0 such that ι1(w)=zbcol(w(0),w(0),,w(k1)(0))wkerP(D),which is (Equation29). To establish (Equation28), let zker(CCACAk1)and set w=yz. Then w(j)(0)=CAjz=0 for all j=0,,k1, and so, by (Equation29), z=ι1(w)=0. This completes the proof of statements (1) and (2).

To prove statement (3), let (u,v,x,y)B(S). Then, by (Equation27) and the variation-of-parameters formula, yGuGev=yx(0)kerP(D),where G and Ge are the inverse Laplace transforms (impulse responses) of P1Q and P1Qe, respectively. Appealing to Proposition 4.3, we obtain that (u,v,y)B(P,Q,Qe), showing that λ indeed maps into B(P,Q,Qe). Furthermore, an application of (Equation29) with z=x(0) and w=yGuGev yields x(0)=ι1(yGuGev)bJ(u,v,y)(u,v,x,y)B(S),which establishes (Equation30). It is obvious that λ is linear and it follows from the observability of (C,A) that λ is injective. As for surjectivity, let (u,v,y)B(P,Q,Qe). By Proposition 4.3, yGu+GevkerP(D), and so there exists zRn such that yz=ι(z)=yGu+Gev. Defining x(t):=eAtz+0teA(tτ)(Bu(τ)+Bev(τ))dτt0,it is clear that (u,v,x,y)B(S), whence (u,v,y) is the image of (u,v,x,y) under λ, completing the proof of statement (3).

To establish statement (4), let (u,v,x,y)B(S) with (u,v)Wlockd,q(R+,Rm)×Wlockde,q(R+,Rme). By statement (3), we have that (u,v,y)B(P,Q,Qe) and Corollary 4.5 guarantees that yWlock,q(R+,Rp). The existence of constant c>0 such that (Equation31) holds follows from statement (3) and Lemma 2.3.

The remaining statements (5) and (6) can be proved as in Sarkans and Logemann (Citation2016a, Proof of Theorem 3.2).

Step 2. Let us now remove the assumption that P is row reduced. Appealing to statement (1) of Lemma 2.1 and statement (2) of Proposition 2.2, we see that there exists unimodular U0R[s]p×p such that U0P is row reduced and deg(U0P)=degleftP=. Let TRp×p be a product of suitable row-switching transformations such that r1(TU0P)r2(TU0P)rp(TU0P). Obviously, U:=TU0 is unimodular, UP is row reduced, deg(UP)=deg(U0P)= and r1(UP)r2(UP)rp(UP), and so Step 1 can be applied with (P,Q,Qe) replaced by (UP,UQ,UQe). Since, trivially, any state-space realisation of (UP,UQ,UQe) is also a realisation of (P,Q,Qe), kerP(D)=ker(UP)(D) and B(P,Q,Qe)=B(UP,UQ,UQe) (by Corollary 4.4), we conclude that statements (1)–(6) hold, completing the proof.

Whilst Theorem 4.6 has some overlap with Willems (Citation1983, Theorem 5.1), we emphasise that, for our purposes, Theorem 4.6 is more appropriate than Willems (Citation1983, Theorem 5.1). In particular, it contains key results relevant in Section 5 which are not included in Willems (Citation1983), for example, (Equation29), (Equation30) and statements (4)–(6).

5. Stability of input-to-output Lur'e systems

Throughout this section, we let P, Q and Qe be as in Section 4, that is, (P,Q,Qe)Σio. Furthermore, let G and Ge be the inverse Laplace transforms (impulse responses) of G=P1Q and Ge=P1Qe, respectively, and let f:RpRm be a continuous function.

Application of the feedback u=f(y) to (Equation17) results in (38) P(D)y=Q(D)(fy)+Qe(D)v,(38) or, equivalently, j=0kPjy(j)(t)=j=0kQj(fy)(j)(t)+j=0kQejv(j)(t)t0,where Pk0. We say that (Equation38) is a (higher-order) input-output Lur'e system. Occasionally, it will be convenient to refer to (Equation38) as (P,Q,Qe,f).

In this section, we develop a stability theory for the input-output Lur'e system (Equation38). We first extend the weak trajectory concept introduced in Section 4 to the nonlinear system (Equation38) and define the associated behaviour. Then, in Proposition 5.1 below, we relate the behaviour of (Equation38) to that of a corresponding state-space Lur'e system, using results from Section 4. Theorem 3.1 and Proposition 5.1 provide the basis for the proof of Theorem 5.2, the main result of the section and the paper.

A pair (v,y)Lloc(R+,Rme)×Lloc(R+,Rp) is a weak trajectory of (Equation38) if (fy,v,y) is a weak trajectory of (Equation17). The behaviour B(P,Q,Qe,f) of (Equation38) is defined to be the set of all weak trajectories of (Equation38). Consequently, (v,y)B(P,Q,Qe,f) if, and only if, (fy,v,y)B(P,Q,Qe).

The next proposition relates the behaviour B(P,Q,Qe,f) of the nonlinear input-output system (Equation38) to the behaviour B(Sf) of the nonlinear state-space system (Equation6), where S=(A,B,Be,C,D,De) is the realisation of (P,Q,Qe) having the properties guaranteed by Theorem 4.6. Recall that :=degleftP.

Proposition 5.1

Let (P,Q,Qe)Σio and let d and de denote the relative degrees of G and Ge, respectively. Furthermore, let f:RpRm be continuous, let S be the realisation of (P,Q,Qe) guaranteed to exist by Theorem 4.6, and define the map fD:RpRp by (Equation10). The following statements hold.

(1)

For every (v,x,y)B(Sf), the pair (v,y) is in B(P,Q,Qe,f), the map λf:B(Sf)B(P,Q,Qe,f) given by (v,x,y)(v,y),is a bijection, and there exists b>0 such that (39) x(0)bJ(fy,v,y)(v,x,y)B(Sf).(39)

(2)

Assume that d = 0, Qe(s)0, f is of class Cl for some lN and the map IfD is injective with det(IfD(ξ))0 for all ξRp, and set le:=max{0,lde}. If (v,x,y)B(Sf) with vWlocle,q(R+,Rme), where 1q, then yWlocl,q(R+,Rp). In particular, when l=, there exist b, c>0 such that, for all (v,x,y)B(Sf) with vWlocde,q(R+,Rme), (40) x(0)bJ(fy,v,y)c(i=0de1v(i)(0)+i=01((fy)(i)(0)+y(i)(0))).(40)

(3)

Assume that d1 and Q(s),Qe(s)0. If f is of class Cl and (v,y)B(P,Q,Qe,f) with vWlocle,q(R+,Rme), where l,leN0, 1q, then yWlocl0,q(R+,Rp), where l0:=min{l+d,le+de}. In particular, when f is of class Cd, then there exist b, c>0 such that, for all (v,x,y)B(Sf) with vWlocde,q(R+,Rme), yWloc,q(R+,Rp) and (41) x(0)bJ(fy,v,y)c(i=0de1v(i)(0)+i=0d1(fy)(i)(0)+i=01y(i)(0))).(41)

Note that d0 and de0 as a consequence of statement (5) of Proposition 2.2. If de=0, then the first sum on the RHS of (Equation40) is defined to be equal to 0. A similar convention applies to (Equation41).

Proof

Proof of Proposition 5.1

To prove statement (1), let (v,x,y)B(Sf). Then (fy,v,x,y)B(S) and so, by Theorem 4.6, (fy,v,y)B(P,Q,Qe), which in turn implies that (v,y)B(P,Q,Qe,f). Consequently, λf maps into B(P,Q,Qe,f). To show surjectivity of λf, let (v,y)B(P,Q,Qe,f). Then (fy,v,y)B(P,Q,Qe) and so, invoking Theorem 4.6, there exists a function x:R+Rn such that (fy,v,x,y)B(S), and hence, (v,x,y)B(Sf). This shows that (v,y) is the image of (v,x,y) under λf. In a similar way, injectivity λf follows also from Theorem 4.6, as does the existence of a constant b>0 such that (Equation39) holds.

We proceed to prove statement (2). Set F:=IfD and note that the hypotheses on f guarantee (via the local inverse function theorem) that F(Rp) is open and F1:F(Rp)Rp is of class Cl. Let (v,y)B(P,Q,Qe,f) with vWlocle,q(R+,Rme). By Proposition 4.3, y=w+G(fy)+Gev for suitable wkerP(D)B(R+,Rp). Setting ζ:=w+G0(fy)+Gev,whereG0:=GG()δ=GB(R+,Rp×m),we have that y=F1ζ.As a consequence of Lemma 2.3, GevWle+de,q(R+,Rp). The function fy is locally essentially bounded, and so, G0(fy) is in Wloc1,q(R+,Rp), and it follows that ζWloc1,q(R+,Rp), and (42) ζ=w+G0(fy)+G0(0)(fy)+(Gev).(42) Since F1 is of class Cl, it follows that y=((F1)ζ)ζ and yWloc1,q(R+,Rp). Furthermore, fyWloc1,q(R+,Rm), and so ζWloc1,q(R+,Rp). Consequently, yWloc1,q(R+,Rp), or, equivalently, yWloc2,q(R+,Rp), and, for almost every t0, y(t)=[(F1)(ζ(t))](ζ(t))ζ(t)+((F1)(ζ(t)))ζ(t),where we note that, for each t0, (F1)(ζ(t)) is a linear map from Rp to Rp×p. We also note that (fy)=(fy)y is in Wloc1,q(R+,Rm), and so, by (Equation42), ζWloc1,q(R+,Rp). We conclude that yWloc1,q(R+,Rp), or, equivalently, yWloc3,q(R+,Rp). Repeating this argument shows that yWlocl,q(R+,Rp). The inequality (Equation40) is now an immediate consequence of statement (4) of Theorem 4.6.

The proof of statement (3) is similar to that of statement (2), but now F = I, as D=G()=0, and so the argument becomes simpler.

Proposition 5.1 shows, in particular, that B(P,Q,Qe,f) (the set of weak trajectories of the input-output Lur'e system) and B(Sf) (the set of trajectories of the state-space Lur'e system) are equally ‘rich’.

To formulate the main result of the paper, it is convenient to define Sio(P,Q):={KCm×p:P(s)Q(s)Kis Hurwitz and(PQK)1Q is proper}.The elements of Sio(P,Q) are called stabilising (complex) feedback matrices for the linear input-output system given by (P,Q). If rk(P(s),Q(s))=p for all sC¯0, then Sio(P,Q) can be expressed in terms of G=P1Q as follows: (43) Sio(P,Q)={KCp×m:det(IKG(s))0 andG(IKG)1H(Cp×m)}.(43) Note that the RHS of (Equation43) is identical to that of (Equation5).

The next theorem is the main stability result of this paper.

Theorem 5.2

Let (P,Q,Qe)Σio, f:RpRm be continuous, KRm×p and r>0. Assume that BC(K,r)Sio(P,Q), rD(IKD)1<1, where D:=G(), and that there exists continuous α:R+R+ such that (44) f(ξ)rξα(ξ)ξRp.(44) The following statements hold.

(1)

If αK, then there exist βKL and γK such that, for all (v,y)B(P,Q,Qe,f), (45) y(t)β(J(fy,v,y),t)+γ(vL(0,t))a.e.t0.(45)

(2)

If αK, then there exist β,ψKL, γ,ϕ,θK and b>0 such that (Equation45) holds for all (v,y)B(P,Q,Qe,f) with vLb and (46) y(t)ψ(J(fy,v,y),t)+ϕ(0tθ(v(τ))dτ)a.e.t0(46) is satisfied for all (v,y)B(P,Q,Qe,f).

Before we prove Theorem 5.2, we provide some commentary and state a corollary. The stability properties described by (Equation45) and (Equation46) are the input-output counterparts of the state-space concepts of ISS and integral ISS, respectively, and it would be natural to refer to them as input-to-output stability–not be confused with the classical input-output concept of L-stability (Desoer & Vidyasagar, Citation1975; Vidyasagar, Citation2002)–and integral input-to-output stability, respectively. Adopting this terminology, statement (2) then guarantees ‘small signal’ input-to-output stability and integral input-to-output stability, or strong integral input-to-output stability, for short.

Theorem 5.2 is reminiscent of the complex Aizerman conjecture (Guiver & Logemann, Citation2020; Hinrichsen & Pritchard, Citation1992Citation2010; Jayawardhana et al., Citation2011; Sarkans & Logemann, Citation2015Citation2016a, Citation2016b) in the sense that the assumption of stability for all linear feedback gains in the complex ball BC(K,r) guarantees stability of the nonlinear Lur'e system for every nonlinearity satisfying the ‘nonlinear’ ball condition (Equation44) with αK or αK.

Remark 5.3

We note that if KSio(P,Q), then the largest r>0 such that BC(K,r)Sio(P,Q) is given by r=1/G(IKG)1H, provided that Q(s)0 (or, equivalently, G(s)0), and (Equation44) can be expressed in form of the following ‘nonlinear’ small-gain condition: G(IKG)1Hf(ξ)ξ1ρ(ξ)ξξRp,ξ0,where ρ is in K or K.

Under suitable regularity assumptions on f and v, the term J(fy,v,y) on the RHS of (Equation45) and (Equation46) can be replaced by a term involving the norms of the individual derivatives v(i)(0), y(i)(0) and (fy)(i)(0), where i=0,,1, see statements (2) and (3) of Proposition 5.1.

Next, we state a corollary which provides a circle criterion for the stability conditions (Equation45) and (Equation46). Recall that a square rational matrix H is said to be positive real if H(s)+H(s) is positive semi-definite for all complex numbers sC0 which are not poles of H, where H(s):=(H(s)), the Hermitian transposition of H(s). More information and details on matrix-valued positive-real functions can be found, for example, in Guiver et al. (Citation2017).

The classical circle criterion for absolute stability is usually formulated in an input/output-operator setting or in state-space terms. It guarantees L2 or Lyapunov stability, respectively (see, for example, Desoer & Vidyasagar, Citation1975, Theorem 10, p. 140 or Khalil, Citation2002, Theorem 7.1, p. 265) for all nonlinearities satisfying a certain sector-condition provided that the transfer function of the linear system satisfies a suitable positive-real condition. The nomenclature circle criterion stems from the fact that in the single-input single-output (SISO) case the positive-real condition admits a graphical characterisation involving circles in the complex plane.

Corollary 5.4

Let (P,Q,Qe)Σio, f:RpRm be continuous, K1,K2Rm×p, assume that det(IK1G(s))0 and set H:=(IK2G)(IK1G)1. Assume further that H is positive real, that H()+H() is positive definite and that there exists continuous α:R+R+ such that (47) f(ξ)K1ξ,f(ξ)K2ξα(ξ)ξξRp,(47) where , denotes the usual Euclidean inner product in Rp. The following statements hold.

(1)

If αK, then the conclusions of statement (1) of Theorem 5.2 hold.

(2)

If αK, then the conclusions of statement (2) of Theorem 5.2 hold.

Corollary 5.4 can be derived from Theorem 5.2 by arguments similar to those used in the proof of Sarkans and Logemann (Citation2016a, Corollary 4.6) and, therefore, we do not go into details.

In the single-input single-output case (m = p = 1), where K1=k1 and K2=k2 are scalars with k1<k2, the following inequalities (48) k1ξ2+α1(|ξ|)|ξ|ξf(ξ)k2ξ2α2(|ξ|)|ξ|ξR,(48) may seem more natural than the sector condition (Equation47). In (Equation48), α1 and α2 are functions from R+ to R+. The inequalities in (Equation48) mean that the graph of f is ‘sandwiched’ between the straight lines k1ξ and k2ξ and is bounded away from these lines by αj(|ξ|). A graphical illustration is shown in Figure .

Figure 1. Illustrative sector condition (Equation48) when m = p = 1 with 0<k1<k2. The straight lines have slopes ki.

Figure 1. Illustrative sector condition (Equation48(48) k1ξ2+α1(|ξ|)|ξ|≤ξf(ξ)≤k2ξ2−α2(|ξ|)|ξ|∀ξ∈R,(48) ) when m = p = 1 with 0<k1<k2. The straight lines have slopes ki.

To link the conditions (Equation47) and (Equation48), the following simple result on comparison functions is useful.

Lemma 5.5

If h:(0,)(0,) is continuous and such that lim infsh(s)>0, then there exists αK such that α(s)h(s) for all s>0. Furthermore, if  lim infsh(s)=, then αK.

Although the proof of the above lemma is not difficult, it is, for completeness and for the convenience of the reader, included in the Appendix.

The corollary below provides the desired link between (Equation47) and (Equation48).

Corollary 5.6

Let k1 and k2 be real scalars with k1<k2, let α1,α2:R+R+ be continuous and such that αj(s)>0 for all s>0, j = 1, 2, and (49) lj:=lim infsαj(s)>0,j=1,2(49) If (Equation48) holds, then there exists αK such that (f(ξ)k1ξ)(f(ξ)k2ξ)α(|ξ|)|ξ|ξR.Furthermore, if lj= for j = 1, 2, then αK.

Proof.

Arguments similar to those used in the proof of Sarkans and Logemann (Citation2015, Corollary 3.13) show that if (Equation48) holds, then (f(ξ)k1ξ)(f(ξ)k2ξ)k2k12min(α1(|ξ|),α2(|ξ|))|ξ|ξR.The claim now follows from Lemma 5.5

Proof

Proof of Theorem 5.2

Let S=(A,B,Be,C,D,De) be the state-space realisation of the input-output system (P,Q,Qe) guaranteed to exist by Theorem 4.6. As KSio(P,Q), it is clear that IDK is invertible. Setting L:=K(IDK)1, we have that I+DL=(IDK)1. Therefore, P(I+DL)QL=(PQK)(IDK)1, and so, P(I+DL)QL is Hurwitz. It follows from statement (5) of Theorem 4.6 that (A,B) is stabilizable. Now (C,A) is observable, and a fortiori (C,A) is detectable, implying that (Equation5) holds. Furthermore, rk(P(s),Q(s))=p for all sC¯0 (because otherwise PQK could not be Hurwitz), and thus, (Equation43) is satisfied. Consequently, Sss(A,B,C,D)=Sio(P,Q), whence BC(K,r)Sss(A,B,C,D) by assumption. We conclude that the hypotheses of Theorem 3.1 are satisfied.

To prove statement (1), we first note that, for all (v,x,y)B(Sf), y=CKx+DK(fyKy)+DeKv, where CK:=(IDK)1C, DK:=(IDK)1D and DeK:=(IDK)1De, and thus, CKx(t)+DeKv(t)y(t)rDKy(t)a.e.t0,(v,x,y)B(Sf).By hypothesis, rDK<1, and we conclude that there exists a constant b>0 such that (50) b(x(t)+v(t))y(t)a.e.t0,(v,x,y)B(Sf).(50) Next we apply statement (1) of Theorem 3.1 by which there exist β0KL and γ0K such that, for all (v,x,y)B(Sf) x(t)β0(x(0),t)+γ0(vL(0,t))t0.Combining this with (Equation50) yields, for all (v,x,y)B(Sf) (51) y(t)β1(x(0),t)+γ(vL(0,t))a.e.t0,(51) where β1KL and γK are defined by β1(s,t):=bβ0(s,t) and γ(s):=b(γ0(s)+s) for all s,t0. Using Proposition 5.1, there exists c>0 such that x(0)cJ(fy,v,y)(v,y)B(P,Q,Qe,f),where x is the unique function in Wloc1,1(R+,Rn) such that (v,x,y)B(Sf). Hence, invoking (Equation51), we conclude that, for all (v,y)B(P,Q,Qe,f), y(t)β(J(fy,v,y),t)+γ(vL(0,t))a.e.t0,where βKL is given by β(s,t):=β1(cs,t) for all s,t0, completing the proof of statement (1).

Statement (2) can be proved by a similar argument, with the application of statement (1) of Theorem 3.1 replaced by that of statement (2) of Theorem 3.1.

In the corollary below, we consider the situation wherein condition (Equation44) is only satisfied on the complement of a bounded set. It turns out that some form of stability, reminiscent of practical ISS or ISS with bias (see, for example, Jayawardhana et al., Citation2009Citation2011; Mironchenko, Citation2019), is retained under this weaker assumption.

Corollary 5.7

Let (P,Q,Qe)Σio, f:RpRm be continuous, KRm×p and r>0. If BC(K,r)Sio(P,Q), rD(IKD)1<1, where D:=G(), and there exist αK and a>0 such that f(ξ)rξα(ξ)for allξRpwithξa,then there exist βKL, γK and b>0 such that, for all (v,y)B(P,Q,Qe,f), (52) y(t)β(J(fy,v,y),t)+γ(vL(0,t)+b)a.e.t0.(52)

Proof.

To make use of Theorem 5.2, we introduce a modified nonlinearity f~ defined by f~(ξ):={f(ξ),ifξa,(ξ/a)f((a/ξ)ξ),if0<ξ<a,0,ifξ=0.Note that f~ is continuous and f~(ξ)rξα~(ξ)ξRp,where α~K is given by α~(s):={α(s),ifsa,(s/a)α(a),if0s<a.An application of Theorem 5.2 to the system (P,Q,Qe,f~) shows that there exist βKL and γK such that, for all (v~,y~)B(P,Q,Qe,f~), (53) y~(t)β(J(f~y~,v~,y~),t)+γ(v~L(0,t))a.e.t0.(53) Furthermore, if (v,y)B(P,Q,Qe,f), then (v+d,y)B(P,Q,Qe,f~), where d(t)=f(y(t))f~(y(t)) for all t0. Clearly, d(t)maxξaf(ξ)f~(ξ) for all t0. It now follows from (Equation53) that, for all (v,y)B(P,Q,Qe,f), (Equation52) holds with b:=maxξaf~(ξ)f(ξ).

Finally, we introduce a certain type of initial-value problem for (Equation38). Due to the lack of regularity of weak trajectories, it is not immediately clear how an initial-value problem for (Equation38) can be defined. We will now briefly explain how this can be done. To this end, we introduce the vector space J:={col0i1(w(i)(0)):wkerP(D)}Rℓpand note that dimJ=degdetP, as follows from statement (1) of Proposition 4.3 and statement (1) of Theorem 4.6. Moreover, J(fy,v,y)J for every (v,y)B(P,Q,Qe,f) (because if (v,y)B(P,Q,Qe,f), then (fy,v,y)B(P,Q,Qe), and thus, yG(fy)GevkerP(D) by Proposition 4.3).

The next result shows that, under a suitable local invertibility conditions on the map IDf, the assumptions of Theorem 5.2 ensure that, for every vLloc(R+,Rme) and every ζJ, there exists a unique yLloc(R+,Rp) such that (v,y)B(P,Q,Qe,f) and J(fy,v,y)=ζ.

Proposition 5.8

Imposing the notation and assumptions of Theorem 5.2, assume further that f is locally Lispchitz, (Equation44) holds with αK, and fD defined by (Equation10) is of class C1 with det(IfD(ξ))0 for all ξRp, where D:=G(). Then, for all vLloc(R+,Rme) and ζJ, there exists a unique yLloc(R+,Rp) such that (v,y)B(P,Q,Qe,f) and J(fy,v,y)=ζ.

Trivially, the conditions involving fD are satisfied when D = 0.

Proof

Proof of Proposition 5.8

Let S=(A,B,Be,C,D,De) be the state-space realisation of the input-output system (P,Q,Qe) guaranteed to exist by Theorem 4.6. As has been shown in the proof of Theorem 5.2, BC(K,r)Sss(A,B,C,D), and so, the assumptions of statement (3) of Proposition 3.2 are satisfied.

Let vLloc(R+,Rme) and ζJ. Recalling the notation yz(t)=CeAtz, t0 and zRn, and invoking Theorem 4.6, there exists x0Rn such that (54) col0i1((yx0)(i)(0))=ζ,where=degleftP.(54) By statement (3) of Proposition 3.2, there exists a unique pair (x,y) such that (v,x,y)B(Sf) and x(0)=x0. An application of statement (1) of Proposition 5.1 shows that (v,y)B(P,Q,Qe,f). As y=Cx+D(fy)+Dev=yx0+G(fy)+Gev,we have that yG(fy)Gev=yx0, and so J(fy,v,y)=ζ. To show uniqueness, assume that (v,y~)B(P,Q,Qe,f) satisfies J(fy~,v,y~)=ζ. Another application of statement (1) of Proposition 5.1 shows that there exists x~Wloc1,1(R+,Rn) such that (v,x~,y~)B(Sf) and so y~G(fy~)Gev=yx~(0). Consequently, col0i1((yx~(0))(i)(0))=ζ, and thus, appealing to (Equation54), col0i1((yx~(0))(i)(0))=col0i1((yx0)(i)(0)). This in turn leads to col0i1(CAix~(0))=col0i1(CAix0).It now follows from (Equation28) that x~(0)=x0. But as (x,y) is the unique pair such that (v,x,y)B(Sf) and x(0)=x0, it follows that x~=x and y~=y, completing the proof.

6. Examples

In this section we will illustrate the input-to-output stability results of Section 5 by three examples.

Example 6.1

Consider the resistor-inductor-capacitor (RLC) circuit with a current source shown in Figure  and inspired by the electrical circuit example discussed in Jayawardhana et al. (Citation2011, p. 34). The inductor and capacitor are modelled as linear, time-invariant components, with positive constants L (inductance) and C (capacitance). The resistor is nonlinear with current-voltage characteristic given by the (continuous) function f, that is, IR=f(VR), where IR and VR denote the current and voltage, respectively, associated with the nonlinear resistive element. Taking into account the signs of the potential differences, an application of Kirchoff's voltage law gives that the voltages across all the components are equal, and the following differential equation Cd2dt2V+1LV=ddtf(V)+ddtIe,holds, where V is the voltage and Ie is the current of the external source. Setting y = V, u=IR=f(y) and v=Ie, we arrive at (55) P(D)y=Q(D)(fy)+Qe(D)v,whereP(s):=Cs2+1L,Q(s):=s,Qe(s):=s,(55) which is of the the form (Equation38) with m=me=p=1. As C and L are positive, it it is clear that, in the linear case wherein f(y)=ky, with real gain parameter k, the associated differential equation Cd2dt2y+ddtky+1Ly=0,is asymptotically stable if, and only if, k>0. Setting G:=Q/P, we have that Gk(s):=G(s)1kG(s)=sCs2+ks+1/L,and GkH(C) if, and only if, k>0. As rk(P(s),Q(s))=1 for all sC, we conclude that kSio(P,Q) if, and only if, k>0. Using elementary calculus, it can be shown that GkH=|Gk(±i/CL)|=1/kk>0.Consequently, it follows from statement (1) of Theorem 5.2 and Remark 5.3 that, for every continuous function f:RR satisfying (56) |f(ξ)|k|ξ|α(|ξ|)ξR(56) for some k>0 and αK, there exist βKL and γK such that (Equation45) holds for all (v,y)B(P,Q,Qe,f), where v=Ie and y = V.

Figure 2. RLC circuit of Example 6.1.

Figure 2. RLC circuit of Example 6.1.

To capture the situation when f is a so-called negative resistance element (such as a tunnel-diode see, for example, Khalil, Citation2002, Section 1.2.2), meaning that f(0)=0, lim sup|ξ|0(f(ξ)/ξ)<0, and sign(ξ)f(ξ) as |ξ|, we note that Remark 5.3 and Corollary 5.7 guarantee that, for every continuous function f:RR satisfying (57) |f(ξ)|k|ξ|α(|ξ|)for allξRwith|ξ|a(57) for some a, k>0 and αK, there exist βKL, γK and b>0 such that (Equation52) holds for all (v,y)B(P,Q,Qe,f), where v=Ie and y = V.

Noting that for this example J=R2, it follows from Proposition 5.8 that, for every ζ=(ζ1,ζ2)TR2, there exists a unique weak trajectory (v,y)B(P,Q,Qe,f)) satisfying J(fy,v,y)=ζ. A routine calculation shows that ζ and (y(0),y˙(0))T are related as follows y(0)=ζ1andy˙(0)=ζ2+(v(0)f(y(0))/C.To illustrate our results numerically, we present two simulations. The following model data are common to both: C=1,L=1,k=1,y(0)=y˙(0)=12,and the bounded, periodic and discontinuous forcing term v(t):={at[2m,2m+1],at(2m+1,2(m+1)),mN0where a0 is an amplitude parameter, with a = 0 giving rise to the unforced input-output Lur'e system. We consider two nonlinearities fj:RR given by f1(ξ):=sign(ξ)g1(|ξ|),withg1(ξ):=(k/2)min{mod(ξ,2),1}+(k/2)ξ/(2k),andf2(ξ):=(|ξ|1+|ξ|4e|ξ|),where x denotes the largest integer less or equal to x and mod(x,z):=xzx/z, x,zR, z0 (if x and z are integers, then mod(x,z) is the remainder after division of x by z). The functions f1 and f2 have been chosen somewhat arbitrarily to illustrate a positive- and negative-resistance element, respectively. The function f1 is globally Lipschitz, but not differentiable everywhere. Graphs of the functions f1 and f2 are plotted in Figure (a,b) illustrates that f1 and f2, respectively, satisfy (Equation56) and (Equation57), both for some αK. Furthermore, note that f2 does not satisfy (Equation56). These properties are readily verified mathematically.

Figure 3. Numerical simulation results from Example 6.1. (a) Graphs of fj (b) Graphs of fj(ξ). The dotted straight lines have slope ±k. In both panels, j = 1 is shown in solid line and j = 2 in dashed-dotted line. (c) Outputs y=y1 for specified values of a. (d) Outputs y=y2 for specified values of a.

Figure 3. Numerical simulation results from Example 6.1. (a) Graphs of fj (b) Graphs of fj(ξ)−kξ. The dotted straight lines have slope ±k. In both panels, j = 1 is shown in solid line and j = 2 in dashed-dotted line. (c) Outputs y=y1 for specified values of a. (d) Outputs y=y2 for specified values of a.

Let (v,yj) denote the unique weak trajectory of (Equation55) with f=fj, that is, (v,yj)B(P,Q,Qe,fj) for j=1,2. Graphs of the output trajectories plotting yj against t are shown in Figures (c) (j = 1) and (d) (j = 2). In both figures, the blue lines denote the output trajectory subject to zero forcing, which converges to zero when f=f1, and is seen to boundedly oscillate when f=f2. These trajectories have been plotted for comparison purposes. As expected from the estimate (Equation45), the solutions y1 are bounded and we observe larger deviation from the converging solution of the unforced equation as the magnitude parameter a increases. Similar observations apply to y2, but now deviations from the oscillatory unforced solution are observed–behaviour which is compatible with the practical input-to-output stability estimate (Equation52) via the ‘offset’ term b.

In the next example, we provide an illustration of Corollary 5.4 (the circle criterion).

Example 6.2

The following linear input-output system is considered in Polderman and Willems (Citation1998, Example 3.3.25) (58) P(D)y=Q(D)u,whereP(s):=(s+1)2andQ(s):=4s23s+1.(58) The transfer function G(s)=Q(s)/P(s)=(4s23s+1)/(s+1)2 is proper, but not strictly proper. Applying the feedback u=Q(D)(f(y)+v) to (Equation58) leads to P(D)y=Q(D)f(y)+Q(D)vwhich is of the form (Equation38) with Qe=Q. To apply Corollary 5.4, we set H:=(1k2G)/(1k1G) for k1,k2R, plot the Nyquist diagram of G in Figure , and consider two cases.

Figure 4. Nyquist plot for G in Example 6.2.

Figure 4. Nyquist plot for G in Example 6.2.

Case 1: k1<0=k2. In this case, H is positive real if, and only if, 1/(1k1G) is positive real, or, equivalently, if, and only if, (59) ReG()1k1ωR.(59) An inspection of Figure yields that (Equation59) holds if k1=k=0.6095. Consider a continuous nonlinearity satisfying kξ2+α1(|ξ|)|ξ|f(ξ)ξα2(|ξ|)|ξ|ξR,for continuous functions α1,α2:R+R+, where it is assumed that α1(s)>0 and α2(s)>0 for all s>0 and (Equation49) holds. An application of Corollary 5.6 shows that there exists αK (with αK if lj= for j = 1, 2) such that (f(ξ))f(ξ)α(|ξ|)|ξ|ξR,which is of the form (Equation47) with K1=k and K2=0. If lj= for j = 1, 2, then it follows from statement (1) of Corollary 5.4 that there exist βKL and γK such that (Equation45) holds for all (v,y)B(P,Q,Qe,f). Similarly, if l1< or l2< then statement (2) of Corollary 5.4 shows that there exist β,ψKL, γ,ϕ,θK and b>0 such that (Equation45) holds for all (v,y)B(P,Q,Qe,f) with vLb and (Equation46) is satisfied for every trajectory (v,y)B(P,Q,Qe,f).

Case 2: k1=0<k2. We note that H is positive real if, and only if, 1k2G is positive real, or, equivalently, if, and only if, ReG()1k2ωR.It follows from Figure that the above condition is satisfied for k2=1/4. Consider a continuous nonlinearity satisfying α1(|ξ|)|ξ|f(ξ)ξξ2/4α2(|ξ|)|ξ|ξR,where α1,α2:R+R+ are continuous, α1(s)>0 and α2(s)>0 for all s>0 and (Equation49) holds. Corollaries 5.4 and 5.6 allow us to derive conclusions similar to those in Case 1. We leave the details to the reader.

In the third and final example, we study a simple multivariable system.

Example 6.3

Consider the input-output Lur'e system (60) P(D)y=Q(D)f(y)+Qe(D)v,(60) where P(s):=(s0s(s+1)s)=s(10s+11),Q(s):=(s+10s21),Qe:=Q,and f:R2R2 is continuous. As usual, we set G:=P1Q and thus, G(s)=1s(s+10(2s+1)1).Furthermore, defining L:=(1011),it is a routine exercise to show that I+λLG is positive real if, and only if, λ[0,2λ], where λ:=1+2. It follows from Logemann and Townley (Citation1997, Lemma 3.10) that (61) LG(I+λLG)1H=1λλ(0,λ].(61) Consequently, G(I+λLG)1HL1λλ(0,λ],and, as (G(I+λLG)1)(0)=(λL)1, the above equation yields that G(I+λLG)1H=L1λ=3+5λ2λ(0,λ].As rk(P(s),Q(s))=2 for all sC, we conclude that λLSio(P,Q) for all λ(0,λ]. An application of statement (1) of Theorem 5.2 and Remark 5.3 with K:=λL, λ(0,λ], shows that, for every continuous function f:R2R2 for which there exist λ(0,λ] and αK such that f(ξ)+λLξλL1ξα(ξ)=λ23+5ξα(ξ)ξR2,there exist βKL and γK such that (Equation45) holds for all (v,y)B(P,Q,Qe,f).

Now consider system (Equation60) with P(s) replaced by P~(s):=P(s)L1, that is, P~(D)y=Q(D)f(y)+Qe(D)v,where P~(s):=(s0s2s)=s(10s1)and the polynomial matrices Q and Qe are as before. We set G~:=P~1Q=LG, and so, G~(s)=1s(s+10s1).Setting G~k:=G~(IkG~)1, it follows from (Equation61) that G~kH=1|k|k[λ,0)Obviously, rk(P~(s),Q(s))=2 for all sC, and thus, kISio(P,Q) for all k[λ,0). An application of statement (2) of Theorem 5.2 and Remark 5.3 with K: = kI, k[λ,0), shows that, for every continuous function f:R2R2 for which there exist k[λ,0) and αK such that (62) f(ξ)|k|ξα(ξ)ξR2,(62) there exist β,ψKL, γ,ϕ,θK and b>0 such that (Equation45) holds for all (v,y)B(P~,Q,Qe,f) with bLb and (Equation46) is satisfied for all (v,y)B(P~,Q,Qe,f).

Finally, let us analyse the special case wherein k=λ and f is a saturation nonlinearity of the form fa(ξ):={ξ,ifξa,(a/ξ)ξ,ifξ>a,wherea>0.Then (63) fa(ξ)=fa(ξ)+λξ={2ξ,ifξa,(1+2)ξa,ifξ>a,(63) whence fa(ξ)(1+2)ξαa(ξ)=|k|ξαa(ξ)ξR2,where αaK is given by αa(s):={s/2,if0sa,as/(s+a),ifs>a.Hence (Equation62) holds with f=fa, and thus, for every a>0, there exist comparison functions β,ψKL, γ,ϕ,θK and a constant b>0 such that (Equation45) holds for all (v,y)B(P~,Q,Qe,fa) with bLb and (Equation46) is satisfied for all (v,y)B(P~,Q,Qe,fa). Finally, (Equation63) necessitates that any αK for which (Equation62) holds with f=fa and k=λ satisfies α(ξ)a for all ξR2, that is, α is bounded. Consequently, (Equation62) cannot hold with αK, and therefore, we should not expect the conclusions of statement (1) of Theorem 5.2 to hold in the current scenario.

7. Conclusions

We have studied a class of forced continuous-time Lur'e systems obtained by applying nonlinear feedback to a higher-order linear differential equation which defines an input-output system in the sense of behavioural systems theory. A stability theory has been developed for this class of systems with the underlying stability concepts being input-output versions of the input-to-state stability and strong integral input-to-state properties for nonlinear state-space control systems. Our main results are Theorem 5.2 and Corollary 5.4. Theorem 5.2 is reminiscent of the complexified Aizerman conjecture, and states that if all complex gains in a certain ball are stabilising for the associated unforced, linear input-output system, then stability for the corresponding input-output Lur'e system is ensured for all nonlinearities satisfying the corresponding ‘nonlinear’ ball condition (Equation44). Corollary 5.4 is a novel version of the circle criterion, the hypotheses of which can be checked graphically in the single-input single-output case, although the result is valid in the general multivariable case.

The ISS theory for controlled state-space Lur'e systems developed in Sarkans and Logemann (Citation2015) and Guiver and Logemann (Citation2020) provide key tools for the proof of our main results. For this suitable state-state space realisations are required (or, more accurately, relationships between the behaviours of state-space and input-output Lur'e systems) which can be found in Theorem 4.6 and Proposition 5.1, and are of some independent interest. Whilst these results do show that stability of higher-order input-output Lur'e systems can be resolved by converting to a state-space formulation, in practice this is often unsatisfactory. Indeed, many control systems are naturally specified in input-output form and state-space realisations frequently introduce ‘unphysical’ variables irrelevant to the problem under consideration. The availability of stability criteria formulated in terms of the input-output model, such as Theorem 5.2 and Corollary 5.4, is therefore of key importance.

By way of potential future work, it has recently been commented in Sepulchre et al. (Citation2022) that, roughly, incremental stability concepts are more important than stability notions alone. In fact, the work (Sepulchre et al., Citation2022) notes that incremental stability used to have more prominence in the control theory community than perhaps it currently does, and that attention should refocus on this area. Incremental stability broadly refers to bounding the difference of two arbitrary trajectories of a given system; see, for instance Aminzare and Sontagy (Citation2014), Angeli (Citation2002), and Rüffer et al. (Citation2013). The recent works (Gilmore et al., Citation2020Citation2021; Guiver et al., Citation2019) have shown how many absolute stability criteria generalise to ensure incremental stability for various forced state-space Lur'e systems, and we expect that input-output versions of some of these results could be derived.

Statements and declarations

The Matlab routines used to generate the figures in Section 6 are available from the corresponding author on request. Data sharing is not otherwise applicable to this article as no other datasets were generated or analysed during the current study.

Disclosure statement

The authors have no relevant financial or non-financial interests to disclose.

Additional information

Funding

Chris Guiver's contribution to this work was supported by a Personal Research Fellowship from the Royal Society of Edinburgh (RSE) [REF2168].

References

  • Ambrosetti, A., & Prodi, G. (1993). A primer of nonlinear analysis, Cambridge Studies in Advanced Mathematics, vol. 34, Cambridge University Press.
  • Aminzare, Z., & Sontagy, E. D. (2014). Contraction methods for nonlinear systems: A brief introduction and some open problems. In 53rd IEEE conference on decision and control (pp. 3835–3847). IEEE.
  • Angeli, D. (2002). A Lyapunov approach to the incremental stability properties. IEEE Transactions on Automatic Control, 47(3), 410–421. https://doi.org/10.1109/9.989067
  • Arcak, M., & Teel, A. (2002). Input-to-state stability for a class of Lurie systems. Automatica Journal of IFAC, 38(11), 1945–1949. https://doi.org/10.1016/S0005-1098(02)00100-0
  • Blomberg, H., & Ylinen, R. (1983). Algebraic theory for multivariable linear systems, Mathematics in Science and Engineering, vol. 166, Academic Press Inc.
  • Boiko, I. M., Kuznetsov, N. V., Mokaev, R. N., Mokaev, T. N., Yuldashev, M. V., & Yuldashev, R. V. (2022). On counter-examples to Aizerman and Kalman conjectures. International Journal of Control, 95(4), 906–913. https://doi.org/10.1080/00207179.2020.1830304
  • Brockett, R. W., & Willems, J. L. (1965a). Frequency domain stability criteria. I. IEEE Transactions on Automatic Control, AC-10(3), 255–261. https://doi.org/10.1109/TAC.1965.1098143
  • Brockett, R. W., & Willems, J. L. (1965b). Frequency domain stability criteria. II. IEEE Trans. Automatic Control, AC-10(4), 407–413. https://doi.org/10.1109/TAC.1965.1098198
  • Carrasco, J., & Heath, W. P. (2021). Absolute stability. In J. Webster (Ed.), Wiley encyclopedia of electrical and electronics engineering (pp. 1–12). John Wiley & Sons, Ltd.
  • Carrasco, J., Turner, M. C., & Heath, W. P. (2016). Zames-Falb multipliers for absolute stability: from O'Shea's contribution to convex searches. European Journal of Control, 28, 1–19. https://doi.org/10.1016/j.ejcon.2015.10.003
  • Chaillet, A., Angeli, D., & Ito, H. (2014). Combining iISS and ISS with respect to small inputs: the strong iISS property. IEEE Transactions on Automatic Control, 59(9), 2518–2524. https://doi.org/10.1109/TAC.9
  • Dashkovskiy, S., Efimov, D. V., & Sontag, E. D. (2011). Input to state stability and allied system properties. Automation and Remote Control, 72(8), 1579–1614. https://doi.org/10.1134/S0005117911080017
  • Deimling, K. (1985). Nonlinear functional analysis. Springer-Verlag.
  • Delchamps, D. F. (1988). State space and input-output linear systems. Springer-Verlag.
  • De Marco, G., Gorni, G., & Zampieri, G. (1994). Global inversion of functions: an introduction. Nonlinear Differential Equations and Applications NoDEA, 1(3), 229–248. https://doi.org/10.1007/BF01197748
  • Desoer, C. A., & Vidyasagar, M. (1975). Feedback systems: input-output properties. Academic Press.
  • Doetsch, G. (1950). Handbuch der Laplace-Transformation. Band I. Theorie der Laplace-Transformation. Verlag Birkhäuser.
  • Drummond, R., Guiver, C., & Turner, M. C (2022). Aizerman Conjectures for a class of multivariate positive systems. IEEE Transactions on Automatic Control, 68(8), 5073–5080. https://doi.org/10.1109/TAC.2022.3217740
  • Franco, D., Guiver, C., & Logemann, H. (2021). Persistence and stability for a class of forced positive nonlinear delay-differential systems. Acta Applicandae Mathematicae, 174(1), 1. https://doi.org/10.1007/s10440-021-00414-5
  • Fuhrmann, P. A., & Helmke, U. (2015). The mathematics of networks of linear systems, Universitext, Springer.
  • Gilmore, M. E., Guiver, C., & Logemann, H. (2020). Infinite-dimensional Lur'e systems with almost periodic forcing. Mathematics of Control, Signals, and Systems, 32(3), 327–360. https://doi.org/10.1007/s00498-020-00262-y
  • Gilmore, M. E., Guiver, C., & Logemann, H. (2021). Incremental input-to-state stability for Lur'e systems and asymptotic behaviour in the presence of Stepanov almost periodic forcing. Journal of Differential Equations, 300, 692–733. https://doi.org/10.1016/j.jde.2021.08.009
  • Gohberg, I., Lancaster, P., & Rodman, L. (1982). Matrix polynomials, Computer Science and Applied Mathematics, Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers].
  • Guiver, C., & Logemann, H. (2020). A circle criterion for strong integral input-to-state stability. Automatica, 111, Article 108641. https://doi.org/10.1016/j.automatica.2019.108641
  • Guiver, C., & Logemann, H. (2022). The circle criterion for a class of sector-bounded dynamic nonlinearities. Mathematics of Control, Signals, and Systems, 34(3), 461–492. https://doi.org/10.1007/s00498-022-00324-3
  • Guiver, C., Logemann, H., & Opmeer, M. R. (2017). Transfer functions of infinite-dimensional systems: positive realness and stabilization. Mathematics of Control, Signals, and Systems, 29(4), 1–61. https://doi.org/10.1007/s00498-017-0203-z
  • Guiver, C., Logemann, H., & Opmeer, M. R. (2019). Infinite-dimensional Lur'e systems: Input-to-state stability and convergence properties. SIAM Journal on Control and Optimization, 57(1), 334–365. https://doi.org/10.1137/17M1150426
  • Haddad, W. M., & Chellaboina, V. (2008). Nonlinear dynamical systems and control. Princeton University Press.
  • Hinrichsen, D., & Pritchard, A. J. (1992). Destabilization by output feedback. Differential and Integral Equations, 5(2), 357–386. https://doi.org/10.57262/die/1371043976
  • Hinrichsen, D., & Pritchard, A. J. (2010). Mathematical systems theory I, Texts in Applied Mathematics vol 48, Springer.
  • Jayawardhana, B., Logemann, H., & Ryan, E. P. (2009). Input-to-state stability of differential inclusions with applications to hysteretic and quantized feedback systems. SIAM Journal on Control and Optimization, 48(2), 1031–1054. https://doi.org/10.1137/070711323
  • Jayawardhana, B., Logemann, H., & Ryan, E. P. (2011). The circle criterion and input-to-state stability. IEEE Control Systems Magazine, 31(4), 32–67. https://doi.org/10.1109/MCS.2011.941143
  • Kailath, T. (1980). Linear systems, Prentice-Hall Information and System Sciences Series, Prentice-Hall, Inc.
  • Kellett, C. M. (2014). A compendium of comparison function results. Mathematics of Control, Signals, and Systems, 26(3), 339–374. https://doi.org/10.1007/s00498-014-0128-8
  • Khalil, H. K. (2002). Nonlinear systems (3rd ed.). Prentice Hall.
  • Kwakernaak, H., & Sivan, R. (1991). Modern signals and systems. Prentice-Hall.
  • Leonov, G. A. (2001). Mathematical problems of control theory, Series on Stability, Vibration and Control of Systems. Series A: Textbooks, Monographs and Treatises, vol. 6, World Scientific Publishing Co., Inc.
  • Logemann, H., & Townley, S. (1997). Discrete-time low-gain control of uncertain infinite-dimensional systems. IEEE Transactions on Automatic Control, 42(1), 22–37. https://doi.org/10.1109/9.553685
  • Mironchenko, A. (2019). Criteria for input-to-state practical stability. IEEE Transactions on Automatic Control, 64(1), 298–304. https://doi.org/10.1109/TAC.2018.2824983
  • Pendharkar, I., & Pillai, H. K. (2008). Systems with sector bound nonlinearities: A behavioral approach. Systems & Control Letters, 57(2), 112–122. https://doi.org/10.1016/j.sysconle.2007.06.019
  • Polderman, J. W., & Willems, J. C. (1998). Introduction to mathematical systems theory, Texts in Applied Mathematics, vol. 26. Springer-Verlag.
  • Rosenbrock, H. H. (1970). State-space and multivariable theory. John Wiley & Sons, Inc. [Wiley Interscience Division].
  • Rüffer, B. S., van de Wouw, N., & Mueller, M. (2013). Convergent system vs. incremental stability. Systems & Control Letters, 62(3), 277–285. https://doi.org/10.1016/j.sysconle.2012.11.015
  • Sandberg, I. W. (1980). Global inverse function theorems. IEEE Transactions on Circuits and Systems, 27(11), 998–1004. https://doi.org/10.1109/TCS.1980.1084740
  • Sarkans, E., & Logemann, H. (2015). Input-to-state stability of Lur'e systems. Mathematics of Control, Signals, and Systems, 27(4), 439–465. https://doi.org/10.1007/s00498-015-0147-0
  • Sarkans, E., & Logemann, H. (2016a). Stability of higher-order discrete-time Lur'e systems. Linear Algebra and Its Applications, 506, 183–211. https://doi.org/10.1016/j.laa.2016.05.012
  • Sarkans, E., & Logemann, H. (2016b). Input-to-state stability of discrete-time Lur'e systems. SIAM Journal on Control and Optimization, 54(3), 1739–1768. https://doi.org/10.1137/130939067
  • Sepulchre, R., Chaffey, T., & Forni, F. (2022). On the incremental form of dissipativity. arXiv preprint. https://doi.org/10.48550/arXiv.2208.04879
  • Sontag, E. D. (1989). Smooth stabilization implies coprime factorization. IEEE Transactions on Automatic Control, 34(4), 435–443. https://doi.org/10.1109/9.28018
  • Sontag, E. D. (2008). Input to state stability: Basic concepts and results. In Nonlinear and optimal control theory. Lecture Notes in Math., vol. 1932 (pp. 163–220). Springer.
  • Sontag, E. D., & Wang, Y. (1999). Notions of input to output stability. Systems & Control Letters, 38(4–5), 235–248. https://doi.org/10.1016/S0167-6911(99)00070-5
  • Vidyasagar, M. (2002). Nonlinear systems analysis, Classics in Applied Mathematics, vol. 42, Society for Industrial and Applied Mathematics (SIAM). Reprint of the second (1993) edition.
  • Willems, J. C. (1983). Input-output and state-space representations of finite-dimensional linear time-invariant systems. Linear Algebra and Its Applications, 50, 581–608. https://doi.org/10.1016/0024-3795(83)90070-8
  • Willems, J. L. (1970). Stability theory of dynamical systems. John Wiley & Sons.
  • Yakubovich, V. A., Leonov, G. A., & Gelig, A. K. (2004). Stability of stationary sets in control systems with discontinuous nonlinearities, Series on Stability, Vibration and Control of Systems. Series A: Textbooks, Monographs and Treatises, vol. 14, World Scientific Publishing Co., Inc.
  • Zemanian, A. H. (1987). Distribution theory and transform analysis. Dover Publications, Inc. Reprint of the 1965 original published by McGraw-Hill.

Appendix

Here we provide proofs of Proposition 2.2, and Lemmas 2.3 and 5.5.

Proof

Proof of Proposition 2.2

To prove statement (1), denote the entries of P, Q and G=P1Q by Pij, Qij and Gij, respectively. Choose i{1,,p} and j{1,,m} such that degQij=degQ. As Qij=l=1pPilGlj, we conclude that (A1) degQ=degQijmax1lp(degPildegrelGlj).(A1) Let l0{1,,p} be such that the maximum on the RHS of (EquationA1) is achieved for l=l0. It follows from (EquationA1) that degrel(P1Q)=degrelGdegrelGl0jdegPil0degQdegPdegQ.We proceed to establish statement (2). Let L,UR[s]p×p be such that detL(s)0, U is unimodular and UP is row reduced. It is sufficient to show that deg(LP)deg(UP). Let ρ1,,ρp be the row degrees of UP and set D(s):=diag1jp(sρj).Using statement (2) of Lemma 2.1, we conclude that the limit matrix R:=lim|s|D1(s)U(s)P(s)Rp×pis invertible. Therefore, as LP=(LU1D)(D1UP), we may conclude that (A2) deg(LP)=deg(LU1D).(A2) Furthermore, writing LU1=(v1,,vp),wherevjR[s]p,j=1,,p,we have that L(s)U1(s)D(s)=(sρ1v1(s),,sρpvp(s))As det(L(s)U1(s))0, we see that vj(s)0, j=1,,p, and thus deg(sρjvj(s))ρj,j=1,,p,implying that deg(LU1D)=max1jpdeg(sρjvj(s))max1jpρj=deg(UP).Hence, by (EquationA2), deg(LP)deg(UP), showing that deg(UP)=degleftP.

Statement (3) follows from statement (2) and the fact that there exists a unimodular matrix UR[s]p×p such that UP is row reduced (see statement (1) of Lemma 2.1).

To prove statements (4) and (5), we use statement (1) of Lemma 2.1 which guarantees the existence of unimodular UR[s]p×p such that UP is row reduced. By statement (2), degleftP=deg(UP) and thus, degleftP=deg(UP)=max1jprj(UP)j=1prj(UP)=degdet(UP)=degdetP,showing that statement (4) holds. Moreover, since G=(UP)1(UQ), statement (1) yields that (A3) degrelGdeg(UP)deg(UQ)=degleftPdeg(UQ).(A3) By hypothesis, Q(s)0, and so, U(s)Q(s)0, whence deg(UQ)0. Invoking (EquationA3) shows that degleftPdegrelG, establishing statement (5).

Finally, we proceed to prove statement (6). By statement (3) there exists unimodular UR[s]p×p such that deg(UP)=0. This means that there exists ΓRp×p such that U(s)P(s)Γ. As det(U(s)P(s))0, we conclude that Γ is invertible. Thus, P1=Γ1UR[s]p×p, showing that P is unimodular.

Proof

Proof of Lemma 2.3

Let uWlocl,q(R+,Rm). We may assume that l+d1. By a well-known result on the differentiation of a convolution (see, for instance, Doetsch, Citation1950, Kapitel 2, Section 14, Satz 10), (A4) (Gu)=G0u+G0(0)u+G()u.(A4)

We note that if d1, then G()=0 and G0=G. By differentiating (EquationA4) repeatedly, we conclude that GuWlocl,q(R+,Rp). If d = 1, then (EquationA4) implies that GuWlocl+1,q(R+,Rp). Moreover, if d2, then the initial-value theorem (see, for example, Doetsch, Citation1950, Kapitel 14, Section 2, Satz 4 or Zemanian, Citation1987, Corollary 8.6-1a) guarantees that G0(0)=G0(0)==G0(d2)(0)=0,and repeated differentiation of (EquationA4) shows that GuWlocl+d,q(R+,Rp). Finally, in each case, the process of repeated differentiation leads to (Equation3).

Proof

Proof of Lemma 5.5

Define β:R+R+ by β(0):=0 and β(s):=inft[s,)min(h(t),t)s>0.The function β is continuous and non-decreasing, β(s)>0 and β(s)h(s) for all s>0. Furthermore, if lim infsh(s)=, then β(s) as s. Setting α(s):=ss+1β(s)s0,we have that αK, α(s)<β(s)h(s) for all s>0, and, if lim infsh(s)=, then αK.