• No results found

Conditions on shifted passivity of port-Hamiltonian systems

N/A
N/A
Protected

Academic year: 2021

Share "Conditions on shifted passivity of port-Hamiltonian systems"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Conditions on shifted passivity of port-Hamiltonian systems

Monshizadeh, Nima; Monshizadeh, Pooya; Ortega, Romeo; van der Schaft, Arjan

Published in:

Systems and Control Letters DOI:

10.1016/j.sysconle.2018.10.010

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Final author's version (accepted by publisher, after peer review)

Publication date: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Monshizadeh, N., Monshizadeh, P., Ortega, R., & van der Schaft, A. (2019). Conditions on shifted passivity of port-Hamiltonian systems. Systems and Control Letters, 123, 55-61.

https://doi.org/10.1016/j.sysconle.2018.10.010

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Conditions on Shifted Passivity of Port-Hamiltonian Systems

Nima Monshizadeha, Pooya Monshizadehb, Romeo Ortegac, Arjan van der Schaftb

aEngineering and Technology Institute, University of Groningen, 9747AG, The Netherlands.

bJohann Bernoulli Institute for Mathematics and Computer Science, University of Groningen, 9700 AK, The Netherlands cLaboratoire des Signaux et Systmes, CNRS-SUPELEC, Plateau du Moulon, 91192, Gif-sur-Yvette, France

Abstract

In this paper, we examine the shifted passivity property of port-Hamiltonian systems. Shifted passivity accounts for the fact that in many applications the desired steady-state values of the input and output variables are nonzero, and thus one is interested in passivity with respect to the shifted signals. We consider port-Hamiltonian systems with strictly convex Hamiltonian, and derive conditions under which shifted passivity is guaranteed. In case the Hamiltonian is quadratic and state dependency appears in an affine manner in the dissipation and interconnection matrices, our conditions reduce to negative semidefiniteness of an appropriately constructed constant matrix. Moreover, we elaborate on how these conditions can be extended to the case when the shifted passivity property can be enforced via output feedback, thus paving the path for controller design. Stability of forced equilibria of the system is analyzed invoking the proposed passivity conditions. The utility and relevance of the results are illustrated with their application to a 6th order synchronous generator model as well as a controlled rigid body system.

Keywords: Passivity, shifted passivity, incremental passivity, port-Hamiltonian systems, stability theory

1. Introduction

Passive systems are a class of dynamical systems in which the rate at which the energy flows into the system is not less than the increase in storage. In other words, starting from any initial condition, only a finite amount of energy can be extracted from a passive system. This, together with the invariance under negative feedback interconnection, has promoted passivity as a basic building block for control of dynamical and interconnected systems. Interested readers are referred to [1, 2, 3] for a tutorial account of the applications of passivity in control theory.

Passivity of state-space systems is commonly defined as an input-output property for systems whose desired equilibrium state is the origin and the input and output variables are zero at this equilibrium [4, 1, 2]. If several such systems are interconnected—for instance, a plant with a controller—the origin is an equilibrium point of the overall system whose stability may be assessed using the tools of passivity theory. In many applications, however, the desired equilibrium is not at the origin and the input and output variables of the system take nonzero values at steady-state. A standard procedure to describe the dynamics in these cases is to generate a so-called incremental model with inputs and outputs the deviations with respect to their value at the equilibrium. A natural question that arises is whether passivity of the original system is inherited by its incremental model, a property that we refer in this paper as shifted passivity. Following [5], we use a shifted storage function to address this issue, see also the shaped Hamiltonian in [6]. This shifted function is closely related to the Bregman divergence in convex analysis [7], and to the availability function as used in thermodynamics [8, 9]. A by-product of the construction of shifted storage functions is a passivity property which is uniform for a range of steady-state solutions. This is particularly

Email addresses: n.monshizadeh@rug.nl (Nima Monshizadeh), p.monshizadeh@rug.nl (Pooya Monshizadeh), ortega@lss.supelec.fr (Romeo Ortega), a.j.van.der.schaft@rug.nl (Arjan van der Schaft)

(3)

advantageous in flow networks, distribution, and electrical networks where loads/demands are not precisely known [10, 11, 12, 13, 14]; see also [15, 16, 17] where the term “equilibrium-independent passivity” has been used to refer to a closely related passivity property.

We study in this paper shifted passivity of port-Hamiltonian (pH) systems that, as is well-known, provide an attractive energy-based modeling framework for nonlinear physical systems [18, 19, 20]. The Hamiltonian readily serves as a storage function certifying passivity of a pH system, however, proving its shifted passivity is in general nontrivial. In [5] it is shown that pH systems with convex Hamiltonian are also shifted passive provided the input, dissipation and interconnection matrices are all constant. Conditions for shifted passivity of pH systems with state-dependent matrices have been reported in [6] and [21]. In the former case, quite conservative, integrability conditions, are imposed while the latter ones can be difficult to verify. The main contribution of the present paper is to give compact and easily verifiable conditions—i.e., monotonicity of a suitably defined function—to ensure shifted passivity of pH systems with strictly convex Hamiltonian and state-dependent dissipation and interconnection matrices. Notably, for the case of affine pH systems with quadratic Hamiltonian, our conditions reduce to negative semidefiniteness of a constant matrix. The latter becomes a necessary and sufficient condition for shifted passivity in case the input matrix is nonsingular. Quadratic affine pH systems are relevant in applications such as synchronous generators and a rigid body system as will be demonstrated by the case studies. The proposed conditions are able to verify shifted passivity as well as stability of these systems in an efficient way, namely by checking whether a systematically constructed matrix is negative semidefinite.

The proposed conditions are exploited to certify local and global stability of forced pH systems, i.e., under constant external inputs. An additional contribution of our work is that the proposed conditions provide an estimate of the excess and shortage of passivity that serves as a tool for controller design, see e.g. [22].

The structure of the paper is as follows. The problem formulation is provided in Section 2. The main results are given in Section 3, and are specialized to quadratic affine pH systems in Section 4. Section 5 contains the illustration of the results on a synchronous generator and a rigid body system. The paper closes with conclusions in Section 6.

Notation All functions are assumed to be sufficiently smooth. For mappings H : Rn

→ R and C : Rn

→ Rn

we denote the transposed gradient as ∇H := ∂H ∂x

>

and the transposed Jacobian matrix as ∇C := ∂C ∂x

> . The Jacobian (∇C(·))> is simply denoted by ∇C(·)>. An n × m matrix of zeros is denoted by 0

nm, and the

identity matrix of size n is denoted by In. For a vector x ∈ Rn, we denote its Euclidean norm by kxk. The

set of nonnegative real numbers is denoted by R≥0.

2. Problem Formulation Consider the pH system

˙

x = (J (x) − R(x))∇H(x) + Gu (1a)

y = G>∇H(x), (1b)

with state x ∈ Rn, input u ∈ Rm

, and output y ∈ Rm

. The constant matrix G ∈ Rn×mhas full column rank,

and H : Rn

→ R is the Hamiltonian of the system.1 The matrix J is skew-symmetric, i.e., J (x) + J>(x) = 0,

and

R(x) ≥ R∗, ∀x ∈ Rn (2)

for some constant positive semidefinite matrix R∗. Define the steady-state relation

E := {(x, u) ∈ Rn

× Rm| (J (x) − R(x))∇H(x) + Gu = 0}.

(4)

Fix (x, u) ∈ E and the corresponding output y := G>∇H(x). We are interested in finding conditions under which the mapping (u − u) → (y − y) is passive. We refer to this property as shifted passivity, which is formally defined next:

Definition 1 Consider the pH system (1). Let (x, u) ∈ E and define y := G>∇H(x). The pH system (1) is shifted passive if the mapping (u − u) → (y − y) is passive, i.e., there exists a function H : Rn→ R

≥0 such

that

˙

H = (∇H)>x ≤ (u − u)˙ >(y − y) (3) for all (x, u) ∈ Rn× Rm.

Remark 2 Note that shifted passivity is different from the classical incremental passivity property [23]. In fact, the latter is much more demanding as the word “incremental” refers to two arbitrary input-output pairs of the system, whereas in the former only one input-output pair is arbitrary and the other one is fixed to a constant. Notice that shifted passivity is defined with respect to a given pair (x, u) ∈ E . If (3) holds for all (x, u) ∈ E , then the (shifted) passivity property becomes independent of the steady-state values u and x [15].

3. Main Results

In this section, we provide our main results concerning shifted passivity, stability, and shifted feedback passivity of the pH system (1).

3.1. Shifted passivity

Here, we provide conditions under which the pH system (1) is shifted passive in the sense of Definition 1. Towards this end, we make two assumptions:

Assumption 1 The Hamiltonian H is strictly convex.

Given the strictly convex function H we define the Legendre transform, sometimes called Legendre-Fenchel transform, of H as the function

H∗(p) := max

x∈Rn{x

>p − H(x)}, (4)

where the domain of H∗ is the set of all p for which the expression is well-defined (i.e., the maximum is attained). We list the following properties of the Legendre transform H∗; see e.g. [24], [25].

1. The domain of H∗ is equal to the range of ∇H.2

2. H∗ is strictly convex.

3. H∗∗= H. 3

4. ∇H∗(∇H(x)) = x, for all x.

5. ∇H(∇H∗(p)) = p, for all p in the range of ∇H.

Let F (x) := J (x) − R(x). Leveraging the Legendre transform above, the function F (x) can be restated in terms of co-energy variables s := ∇H(x) as

F (x) = F (∇H∗(s)) =: F (s). (5)

2If p is not in the range of ∇H, then the maximum in (4) for such p does not exist.

3This means that the Legendre transform is an involution, namely the Legendre transform of the Legendre transform of H

(5)

Remark 3 The reason for rewriting the vector field in terms of co-energy variables is mainly technical, and is due to the fact that the gradient of the shifted Hamiltonian (8), which will serve as the underlying storage function, is naturally expressed by co-energy variables. Moreover, we remark that in order to write (5), only the gradient of the Legendre transform, and not the Legendre transform itself, has to be computed. This gradient is equal to the inverse of ∇H as evident from the last two properties of the Legendre transform noted above.

We denote the domain of H∗, which is equal to the range of ∇H, by S. Let s := ∇H(x). We impose the following assumption on F :

Assumption 2 The mapping F verifies

∇(F (s) s ) + ∇(F (s) s )>− 2R≤ 0, ∀s ∈ S. (6)

Note that the choice of R∗ is important for feasibility of (6), and is best to choose the lower bound in (2) as tight as possible. Moreover, notice that Assumption 2 is trivially satisfied in case the matrices R and J in (1) are state-independent. Now, we have the following result:

Proposition 4 Let Assumptions 1 and 2 hold. Then, the pH system (1) is shifted passive, namely ˙

H ≤ (u − u)>(y − y) (7)

is satisfied with

H(x) := H(x) − (x − x)>∇H(x) − H(x). (8) Proof. First, note that H is nonnegative as the Hamiltonian H is (strictly) convex [7, p. 205], [5, Prop. 2]. Substituting (5) into (1) yields

˙

x = F (s)s − F (s) s + G(u − u), (9)

where we have subtracted 0 = F (s) s + G u. Noting that ∇H(x) = ∇H(x) − ∇H(x), the time derivative of H(x) is computed as ˙ H = (∇H)>x = s − s˙ > F (s)s − F (s) s + (y − y)>(u − u) = s − s> F (s) − F (s)s + s − s>F (x) s − s + (y − y)>(u − u) ≤ s − s> F (s) − F (s)s − s − s> R∗ s − s + (y − y)>(u − u), (10) where we used (1b) in the first identity, added and subtracted the term s − s>

F (s)s and used (5) to write the second equality, while the bound is obtained invoking (2). Now, let

M(s) := F (s)s − R∗s. (11)

Then, from (10) we have ˙

H ≤ s − s>

M(s) − M(s) + (y − y)>(u − u). (12)

By (6), we have that ∇M(s) + (∇M(s))>≤ 0, for all s ∈ S, which ensures that the map M(·) is monotone [26]. The proof is completed noting that by monotonicity

s − s>

M(s) − M(s) ≤ 0. (13)

 The storage function in (8) is the availability function used in [5], and, as a function of x and x, is called the Bregman divergence in convex analysis [7].

(6)

Remark 5 To interpret the condition in Assumption 2, note that by adding and subtracting the term F (s)s, the dynamics in (9) can be written as

˙

x = F (s)(s − s) + σ(s) + G(u − u), (14)

coupled with the (multivariable) static nonlinearity σ(s) := (F (s) − F (s))s. Then, condition (6) means that the excess of shifted passivity in (14) with σ = 0, which is induced by the matrix R∗, should not be smaller than the shortage of shifted passivity in the map s → σ.4 In fact, (13), can be restated as

(s − s)>R(s − s) > (s − s)>(σ(s) − σ(s)).

Remark 6 By Assumptions 1 and 2, both the strict convexity and the monotonicity property must hold for the whole sets Rn and S, respectively, which results in “global” shifted passivity of (1). For local shifted passivity,5

we can restrict to a subset X ⊆ Rn, with Assumptions 1 and 2 modified to

1. The Hamiltonian is strictly convex in X ⊆ Rn.

2. Inequality (6) holds for all s ∈ SX := {∇H(x) | x ∈ X },

while R∗ is any matrix satisfying, instead of (2), R(x) ≥ R∗, for all x ∈ X .

We complete this subsection by considering the case where the condition (6) does not hold, which means that the system (1) may not be shifted passive, but it can be rendered shifted passive via output feedback. Lemma 7 Consider the pH system (1) verifying Assumption 1 and such that

∇(F (s) s ) + ∇(F (s) s )>− 2R≤ 2γ GG>,

for some γ ∈ R. Then, the shifted Hamiltonian (8) satisfies the following dissipation inequality ˙

H ≤ (u − u)>(y − y) + γ ky − yk2

.

Proof. The proof is analogous to that of Proposition 4, by adding and subtracting the term γGG>(s − s) in (10), and modifying the map M defined in (11) as fM(s) := M(s) − γGG>s.

 Note that a negative γ proves that the pH system is (output-strictly) shifted passive. On the other hand, a positive γ indicates the shortage of shifted passivity. Notice that the simple proportional controller

u = u − KP(y − y) + v,

with KP ≥ γIm, ensures that the interconnected system is passive from the external input v to output y − y.

Analogously, Lemma 7 can be used to design dynamic passive controllers to stabilize the closed-loop system, see [22] for an application to control of permanent magnet synchronous motors.

3.2. Stability of the forced equilibria

Lyapunov stability of the equilibrium of (1) with u = u, immediately follows from Proposition 4, with the Lyapunov function being the shifted Hamiltonian H. Moreover, asymptotic stability follows by imposing the condition that ˙H is negative definite. As will be shown below, for stability analysis we can in fact drop the assumption that the input matrix is constant. To this end, consider again the pH system (1a), where now G = G(x). Let

G(x) = G(∇H∗(s)) =: G(s). Then, we have the following result:

4For static nonlinearities, shifted passivity can be defined analogous to (3) with the storage function replaced by zero. 5By “local” we mean that there exist open neighborhoods X ⊆ Rnand U ⊆ Rmof (x, u) ∈ X × U such that (7) holds for

(7)

Proposition 8 Consider the pH system ˙

x = (J (x) − R(x))∇H(x) + G(x)u, (15)

with (x, u) ∈ E .Then, we have

1. The equilibrium is asymptotically stable if ∇2H(x) > 0 and there exists  > 0 such that the inequality

∇(F (s) s + G(s)u) + ∇(F (s) s + G(s)u)>− 2R∗≤ −2In, (16)

holds at s = s.6

2. The equilibrium is globally asymptotically stable if the Hamiltonian H is strongly convex and (16) holds for all s ∈ Rn.

Proof. The proof of the first item follows analogously to Proposition 4 by modifying the map M in (11) as F (s)s − R∗s + G(s)u, and noting that the the function H is locally nonnegative and is equal to zero only when x = x [7, p. 205], [5, Prop. 2]. To prove the second statement, it suffices to show that H is radially unbounded. This follows from strong convexity of H noting that [27, Ch.2]

H(x) = H(x) − (x − x)>∇H(x) − H(x) ≥ µ kx − xk2, for some µ > 0.

 Remark 9 The identity matrix in the right hand side of (16) can be replaced by a positive semidefinite matrix C>C, with C ∈ Rm×n, if the equilibrium is “observable” from the input-output pair (u, C∇H(x)), namely if

˙

x = F (x)∇H + G(x)u, C∇H(x) = C∇H(x) =⇒ x = x.

Remark 10 While the inequality (16) with  = 0 does not imply shifted passivity of the pH system with state-dependent input matrix, it ensures that the map u − u to the modified output

ym:= G>(x)∇H(x) = G>(x)(∇H(x) − ∇H(x))

is passive. In case G(x) = G is constant, this amounts to shifted passivity of (1) as expected from Proposition 4.

4. Application to Quadratic Affine Systems

In this section we specialize our results to the case where F (x) = F0+

n

X

i=1

Fixi, (17)

with Fj∈ Rn×n, j = 0, . . . , n, constant and

H(x) =1 2x

>Qx, (18)

with Q ∈ Rn×n being positive definite. We call these systems quadratic affine pH systems.

In order to satisfy (2) and state the global version of our results, we need to assume that R(x) = R0

for some constant matrix R0. This is due to the fact that in the affine case the inequality R(x) ≥ R∗, for

all x ∈ Rn, implies that the matrix R is constant. In Remark 14, we elaborate on how this assumption is

relaxed to obtain local results. Note that, in this case, F0+ F0> = −2R0 ≤ 0 and Fj+ Fj> = 0 for each

j ≥ 1.

(8)

Proposition 11 Consider the quadratic affine pH system (1) with (17) and (18). Fix (x, u) ∈ E and define the n × n constant matrix

B := n X i=1 FiQ x e>i Q −1, (19)

with ei ∈ Rn the i-th element of the standard basis. Then, we have

˙

H ≤ (y − y)>(u − u), (20)

where H is the quadratic shifted Hamiltonian function H(x) := 12(x − x)>Q(x − x), if and only if

B + B>− 2R0≤ 0. (21)

Moreover, in case G ∈ Rn×nand nonsingular, the condition (21) becomes necessary and sufficient for shifted passivity of (1).

Proof. The “if ” part follows by verifying the conditions of Proposition 4. In this case, H∗(p) = 1 2p >Q−1p, ∇H(s) = Q−1s, s = Qx, and F (s) = F0+ n X i=1 Fi(e>i Q −1s). (22) Hence, F (s)s = (F0+ n X i=1 Fi(e>i Q −1s))Q x = F 0Qx + n X i=1 FiQxe>i Q −1s

and we obtain that

∇(F (s)s) =

n

X

i=1

FiQxe>i Q−1.

Then, using (19), condition (6) takes the form

∇(F (s)s) + ∇(F (s)s)>− 2R0= B + B>− 2R0≤ 0.

To prove the “only if” part, suppose that (20) is satisfied. This implies that the pH system linearized around the equilibrium x = x is passive with the Hamiltonian H(x) = 12x>Qx serving as a storage function [20, Ch. 11.3]. It is easy to verify that the resulting linear time-invariant system admits the form

˙

x = (J (x) − R0)Qx + BQx + Gu (23a)

y = G>Qx. (23b)

Therefore, we obtain that [20, Ch. 4.1]

Q J (x) − R0+ BQ + Q J(x) − R0+ B

> Q ≤ 0,

which reduces to Q(−2R0+ B + B>)Q ≤ 0. Clearly, the latter inequality is equivalent to (21).

To complete the proof, it suffices to show that the nonlinear quadratic affine pH system (1), (17), (18), with G being nonsingular, is shifted passive only if inequality (21) holds. Again shifted passivity of (1) implies passivity of (23), and by Kalman-Yakubovich-Popov Lemma [20, Ch. 4.1], there exists a symmetric matrix X such that

X J (x) − R0+ BQ + Q J(x) − R0+ B

>

X ≤ 0, XG = QG. (24)

The fact that G is nonsingular yields X = Q, which results in (21) as before.

(9)

Remark 12 The condition in Proposition 11 can equivalently be stated in terms of the co-energy variables s = ∇H(x), which in certain cases decreases the computational effort. To this end, note that by (22), the function F (s) can be written in the affine form:

F (s) = F0+ n X i=1 Fisi, where Fi:=P n

j=1Fj(Q−1)ij. Hence, the matrix B in (19) can be equivalently written as

B = n X i=1 FiQx e>i = n X i=1 Fis e>i .

Since H is strongly convex, and noting the condition (21), Proposition 8 yields the following result: Corollary 13 Consider the quadratic affine pH system (15), (17), (18), with G(x) = G0+P

n i=1Gixi,

where Gj∈ Rn×m, j = 0, . . . , n, are constant. Then the equilibrium ¯x is globally asymptotically stable if

e B + eB>− 2R0< 0, where e B := B + n X i=1 Giu e>i Q −1.

Remark 14 Analogous to the previous section, local variations of Proposition 11 and Corollary 13 can be obtained by restricting x to a domain X ∈ Rn with x ∈ X . In that case, the matrix R

0 in (21) is replaced

by R∗, where R(x) ≥ R∗≥ 0 for all x ∈ X . 5. Examples

In this section, we apply the proposed method to two physical systems. Both systems are affine and have quadratic Hamiltonian.

Example 1 Synchronous generator (6th-order model) connected to a resistor

The state variables of the six-dimensional model of the synchronous generator comprise of the stator fluxes on the dq axes ψd ∈ R, ψq ∈ R, rotor fluxes ψr ∈ R3 (the first component of ψr corresponds to the field

winding and the remaining two to the damper windings), and the angular momentum of the rotor p. The Hamiltonian H (total stored energy of the synchronous generator) is the sum of the magnetic energy of the generator and the kinetic energy of the rotating rotor. More precisely, the Hamiltonian takes the form H(x) = 12x>Qx with x =ψd ψq ψr p > and Q =L −1 0 51 015 m−1  > 0 , L =       Ld 0 kLaf d kLakd 0 0 Lq 0 0 −kLakq kLaf d 0 Lf f d Lakd 0 kLakd 0 Lakd Lkkd 0 0 −kLakq 0 0 Lkkq       ,

where m ∈ R is the total moment of inertia of the turbine and the rotor. Note that the elements of the inductance matrix L are all constant parameters, see [28, 29] for more details. The system dynamics is then given by the pH system [29]

˙

x = (J (x) − R)∇H(x) + GVf τ

 ,

(10)

with J (x) =         022 023 −ψq ψd  032 033 031 ψq −ψd  013 0         , R =             r 0 0 r  023 021 032   Rf 0 0 0 Rkd 0 0 0 Rkq   031 012 013 d             > 0 , G =       021 021   1 0 0   031 0 1       ,

where Vf represents the rotor field winding voltage, τ is the mechanical torque, r is the summation of

the load and stator resistances, Rf, Rkd, Rkq denote the rotor resistances, and d corresponds to the static

friction. We can rewrite the system as

Q−1˙s = (J (s) − R)s + GVf τ



, (25)

where s = Qx = [Id Iq Ir ω]>. Here Id ∈ R, Iq ∈ R are the components of the stator current on the dq

axes, and Ir∈ R3 and ω ∈ R are the currents and angular velocity of the rotor, respectively. Note that

J (s) =       022 023 vJ(s) 032 033 031 vJ>(s) 013 0       , vJ(s) := " −LqIq+ LakqIkq LdId+ Laf dIf+ LakdIkd # .

Let Vf = Vf and τ = τ , for some constant vectors Vf and τ . Through straightforward calculations, and

using Remark 12, the condition (21) reads as         −2r ω(Ld− Lq) 0 0 kωLakq − ¯IqLd ω(Ld− Lq) −2r kωLaf d kωLakd 0 I¯dLd 0 kωLaf d −2Rf 0 0 − ¯IqLaf d 0 kωLakd 0 −2Rkd 0 − ¯IqLakd kωLakq 0 0 0 −2Rkq − ¯IdLakq − ¯IqLd I¯dLd − ¯IqLaf d − ¯IqLakd − ¯IdLakq −2d         ≤ 0 , (26)

where ¯Id, ¯Iq, ¯Ir, ω are the associated values of s at the equilibrium of (25), i.e.

(J (s) − R)s + GVf τ

 = 0 ,

with s = [ ¯Id I¯q I¯r ω]>. Hence, by Proposition 11, (25) is shifted passive if (26) holds. Moreover, by

Corollary 13, if (26) holds with strict inequality, then the equilibrium x = Q−1s is globally asymptotically stable.

Note that the general method proposed in [6] is based on augmenting the Hamiltonian with an extra term which is a function of the input and the states. However, this method relies on integrability conditions that are not satisfied for τ 6= 0 [29], which is the case of interest in practice. The stability condition (26) is consistent with those of [30, 31]7, obtained using suitable, yet ad-hoc, algebraic calculations. We note that Corollary 13 is valid for a general quadratic affine pH-system, and the condition (26) is obtained here in a systematic manner, namely by verifying the negative definiteness test in (21). Moreover, if (26) does not hold, then in view of Lemma 7, one can investigate the possibility of designing suitable proportional, PI, or more generally dynamic (input-strictly) passive controllers rendering the equilibrium globally asymptotically stable.

7Notice that there is a typo in [31] as the term ω(L

(11)

x y

z

Figure 1: A rigid body with three axes of rotation

Example 2 Controlled rigid body under constant disturbances

The equations for the angular momentum of a rigid body (see Figure 1) with external torque u ∈ R3 and

disturbance d ∈ R3 reads as [19] ˙ p = J (p)∇H(p) + u + d (27) y = ∇H(p) , (28) where p =px py pz > , J (p) =   0 −pz py pz 0 −px −py px 0   ,

and the Hamiltonian is given by H(p) = 12p>M p, with

M =   mx 0 0 0 my 0 0 0 mz  > 0 .

Here, mx∈ R, my∈ R, and mz∈ R are the principal moments of inertia. Consider a constant disturbance

d =dx dy dz

>

and a proportional controller u = −Ry with R = diag (rx, ry, rz) > 0. We can rewrite

the system as M ˙ω = J (ω) − Rω +   dx dy dz   . (29) where J (ω) =   0 −ωzmz ωymy ωzmz 0 −ωxmx −ωymy ωxmx 0   , and ω = M−1p =ωx ωy ωz >

is the vector of angular velocities around the axes x, y, and z. The point (ωx, ωy, ωz) is an equilibrium of the system (29) satisfying

J (ω) − R ω +   dx dy dz  = 0 , with ω =ωx ωy wz >

. Through straightforward calculations, the condition (21) (with strict inequality) reads as   −2rx ωz(my− mx) ωy(mx− mz) ωz(my− mx) −2ry ωx(mz− my) ωy(mx− mz) ωx(mz− my) −2rz  < 0 . (30)

(12)

Hence, by Corollary 13, if (30) holds for the equilibrium point (ωx, ωy, ωz), then global asymptotic stability

is guaranteed. In the case that there is disturbance actuating only on one axis, e.g. dy = dz = 0 (without

loss of generality), the equilibrium (ωx, ωy, ωz) = (drx

x, 0, 0) is globally asymptotically stable if

r2xryrz> dx(mz− my) 2 2 . 6. Conclusion

We have examined the shifted passivity property of pH systems with convex Hamiltonian by proposing conditions in terms of the monotonicity of suitably constructed functions, under which the property is ver-ified. We have leveraged these conditions to study (global) asymptotic stability of forced equilibria of the system. As we observed, for quadratic affine pH systems, shifted passivity and (global) asymptotic stability are guaranteed if an appropriately constructed constant matrix is negative semidefinite. We demonstrated the applicability and usefulness of the results on a 6th order synchronous generator model and a controlled rigid body system. Future works include attempting to reduce possible conservatism in the stability condi-tions as well as investigating the conneccondi-tions of the proposed results to contraction and differential passivity [32, 33]. Another direction for future research is robustness analysis of the proposed stability conditions in the presence of time-varying disturbances [34].

7. References

[1] A. van der Schaft, L2-Gain and Passivity Techniques in Nonlinear Control. 3rd Revised and Enlarged Edition (1st

edition 1996, 2nd edition 2000), Springer Communications and Control Engineering series, Springer-International, 2017. [2] R. Ortega, J. A. L. Perez, P. J. Nicklasson, and H. Sira-Ramirez, Passivity-based Control of Euler-Lagrange Systems:

Mechanical, Electrical and Electromechanical Applications. Springer Science & Business Media, 2013.

[3] H. Bai, M. Arcak, and J. Wen, Cooperative Control Design: A Systematic, Passivity–based Approach. New York, NY: Springer, 2011.

[4] J. Willems, “Dissipative dynamical systems Part I : General Theory,” Archive for Rational Mechanics and Analysis, vol. 45, pp. 321–351, 1972.

[5] B. Jayawardhana, R. Ortega, E. Garc´ıa-Canseco, and F. Casta˜nos, “Passivity of nonlinear incremental systems: Applica-tion to PI stabilizaApplica-tion of nonlinear RLC circuits,” Systems & Control Letters, vol. 56, no. 9, pp. 618–622, 2007. [6] B. Maschke, R. Ortega, and A. J. van der Schaft, “Energy-based Lyapunov functions for forced Hamiltonian systems with

dissipation,” IEEE Transactions on Automatic Control, vol. 45, no. 8, pp. 1498–1502, 2000.

[7] L. M. Bregman, “The relaxation method of finding the common point of convex sets and its application to the solution of problems in convex programming,” USSR Computational Mathematics and Mathematical Physics, vol. 7, no. 3, pp. 200–217, 1967.

[8] A. A. Alonso and B. E. Ydstie, “Stabilization of distributed systems using irreversible thermodynamics,” Automatica, vol. 37, no. 11, pp. 1739–1755, 2001.

[9] J. H. Keenan, “Availability and irreversibility in thermodynamics,” British Journal of Applied Physics, vol. 2, no. 7, p. 183, 1951.

[10] J. T. Wen and M. Arcak, “A unifying passivity framework for network flow control,” in INFOCOM 2003. Twenty-Second Annual Joint Conference of the IEEE Computer and Communications. IEEE Societies, vol. 2. IEEE, 2003, pp. 1156– 1166.

[11] C. D. Persis and N. Monshizadeh, “Bregman storage functions for microgrid control,” IEEE Transactions on Automatic Control, vol. PP, no. 99, pp. 1–1, 2017.

[12] S. Trip, M. B¨urger, and C. De Persis, “An internal model approach to (optimal) frequency regulation in power grids with time-varying voltages,” Automatica, vol. 64, pp. 240–253, 2016.

[13] N. Monshizadeh and C. De Persis, “Agreeing in networks: Unmatched disturbances, algebraic constraints and optimality,” Automatica, vol. 75, pp. 63–74, 2017.

[14] J. Wei and A. van der Schaft, “Load balancing of dynamical distribution networks with flow constraints and unknown in/outflows,” Systems & Control Letters, vol. 62, no. 11, pp. 1001–1008, 2013.

[15] G. H. Hines, M. Arcak, and A. K. Packard, “Equilibrium-independent passivity: A new definition and numerical certifi-cation,” Automatica, vol. 47, no. 9, pp. 1949 – 1956, 2011.

[16] M. B¨urger, D. Zelazo, and F. Allg¨ower, “Duality and network theory in passivity-based cooperative control,” Automatica, vol. 50, no. 8, pp. 2051–2061, 2014.

[17] J. W. Simpson-Porco, “Equilibrium-independent dissipativity with quadratic supply rates,” arXiv preprint

(13)

[18] A. van der Schaft and B. Maschke, “The Hamiltonian formulation of energy conserving physical systems with external ports,” AEU. Archiv f¨ur Elektronik und ¨Ubertragungstechnik, vol. 49, no. 5-6, pp. 362–371, 1995.

[19] A. van der Schaft and D. Jeltsema, Port-Hamiltonian Systems Theory: An Introductory Overview. Now Publishers

Incorporated, 2014.

[20] A. van der Schaft and B. Maschke, “Port-Hamiltonian systems on graphs,” SIAM Journal on Control and Optimization, vol. 51, no. 2, pp. 906–937, 2013.

[21] J. Ferguson, R. H. Middleton, and A. Donaire, “Disturbance rejection via control by interconnection of port-Hamiltonian systems,” in Decision and Control (CDC), 2015 IEEE 54th Annual Conference on. IEEE, 2015, pp. 507–512.

[22] R. Ortega, N. Monshizadeh, P. Monshizadeh, D. Bazylev, and A. Pyrkin, “Permanent magnet synchronous motors are globally asymptotically stabilizable with PI current control,” Automatica, vol. 98, pp. 296–301, 2018.

[23] C. A. Desoer and M. Vidyasagar, Feedback Systems: Input-Output Properties. SIAM, 2009.

[24] V. I. Arnol’d, Mathematical Methods of Classical Mechanics. Springer Science & Business Media, 2013, vol. 60. [25] R. T. Rockafellar, Convex analysis. Princeton university press, 2015.

[26] E. K. Ryu and S. Boyd, “Primer on monotone operator methods,” Applied and Computational Mathematics, vol. 15, no. 1, pp. 3–43, 2016.

[27] Y. Nesterov, Introductory Lectures on Convex Optimization: A Basic Course. Springer Science & Business Media, 2013, vol. 87.

[28] R. Ortega, A. Stankovi´c, and P. Stefanov, “A passivation approach to power systems stabilization,” IFAC Proceedings Volumes, vol. 31, no. 17, pp. 309–313, 1998.

[29] S. Fiaz, D. Zonetti, R. Ortega, J. Scherpen, and A. van der Schaft, “A port-Hamiltonian approach to power network modeling and analysis,” European Journal of Control, vol. 19, no. 6, pp. 477–485, 2013.

[30] S. Y. Caliskan and P. Tabuada, “Compositional transient stability analysis of multimachine power networks,” IEEE Transactions on Control of Network Systems, vol. 1, no. 1, pp. 4–14, 2014.

[31] A. van der Schaft and T. Stegink, “Perspectives in modeling for control of power networks,” Annual Reviews in Control, vol. 41, pp. 119–132, 2016.

[32] F. Forni and R. Sepulchre, “A differential Lyapunov framework for contraction analysis,” IEEE Transactions on Automatic Control, vol. 59, no. 3, pp. 614–628, 2014.

[33] F. Forni, R. Sepulchre, and A. Van Der Schaft, “On differential passivity of physical systems,” in Decision and Control

(CDC), 2013 IEEE 52nd Annual Conference on. IEEE, 2013, pp. 6580–6585.

[34] E. D. Sontag, “Input to state stability: Basic concepts and results,” in Nonlinear and optimal control theory. Springer, 2008, pp. 163–220.

Referenties

GERELATEERDE DOCUMENTEN

“Evolutie van de mens” wordt gehouden door Dr.. John de

In section 5 stability and passivity properties applied to the algorithm are discussed whereas in section 6 a technique to preserve passivity of the reduced models is

The geometric formulation of port-Hamiltonian systems motivates a model reduction approach for general port-Hamiltonian systems (possibly also includ- ing the algebraic

dit nu echt heeft gezegd of niet: de uitspraak zal met een korreltje zout genomen moeten worden, maar geeft wel aan dat er in hogere maat- schappelijke kringen een enigszins

In Section 7 (IT and distance learning in K-12 education) the potential of IT for distance learning in pri- mary and secondary education has been explored with particular

From the moral foundation, we derive the proper benchmark to judge institutional functioning from a moral perspective: an institution is just when it sufficiently

Linear model of TPB &amp; NAM variables and demographic variables as predictors and stewardship intention as outcome variable, for with 95% BCa confidence intervals, standard errors

Hence, in this study we investigate porous materials within a single layer, taking fluence, which expresses the energy density at the peak of an ideal beam with a Gaussian