Differentially passive circuits
that switch and oscillate
F.A. Miranda-Villatoro ∗ F. Forni ∗ R. Sepulchre ∗
∗
Department of engineering, University of Cambridge, UK
arXiv:1804.04626v1 [cs.SY] 12 Apr 2018
(e-mail:
[email protected],
[email protected],
[email protected]).
Abstract: The concept of passivity is central to analyze circuits as interconnections of
passive components. We illustrate that when used differentially, the same concept leads to
an interconnection theory for electrical circuits that switch and oscillate as interconnections
of passive components with operational amplifiers (op-amps). The approach builds on recent
results on dominance and p-passivity aimed at generalizing dissipativity theory to the analysis
of non-equilibrium nonlinear systems. Our paper shows how those results apply to basic and
well-known nonlinear circuit architectures. They illustrate the potential of dissipativity theory
to design and analyze switching and oscillating circuits quantitatively, very much like their linear
counterparts.
Keywords: switches, oscillators, differential passivity
1. INTRODUCTION
The concept of passivity originates in circuit theory. It
characterizes circuit elements that can possibly store and
dissipate the energy provided by the environment, but
not the other way around. Passivity is inherently an interconnection concept: passive interconnections of passive
components model passive circuits. Dissipativity theory,
the system theoretic generalization of passivity theory,
has become a cornerstone of system theory. It provides
an interconnection theory to design and analyze stable
dynamical systems. Such systems dissipate the energy
stored internally and provided externally. In short, dissipativity theory is an interconnection theory for Lyapunov stability analysis. In recent years, many researchers
have pointed to the relevance of studying stability incrementally or differentially when addressing questions that
go beyond the stability analysis of isolated equilibria.
Differential stability concepts include contraction theory
(Lohmiller and Slotine, 1998), convergence theory (Pavlov
et al., 2005), or differential Lyapunov theory (Forni and
Sepulchre, 2014). They have proven relevant in a number of areas, most prominently in questions pertaining
to nonlinear observers (Aghannan and Rouchon, 2003),
oscillator synchronization (Stan and Sepulchre, 2007), or
regulation theory (Jouffroy and Fossen, 2010). Differential
dissipativity is to differential stability what dissipativity
is to Lyapunov stability. It was introduced in the recent
papers (Forni et al., 2013; Forni and Sepulchre, 2013; van
Der Schaft, 2013). The present paper aims at illustrating
the potential of differential passivity as an interconnection theory of circuits that can switch an oscillate. In
contrast with classical dissipativity theory, such a theory
must cope with the stability analysis of dynamical systems
that posses multiple equilibria or limit cycles. It is based
⋆ The research leading to these results has received funding from
the European Research Council under the Advanced ERC Grant
Agreement Switchlet n. 670645.
on the concept of p-dominance and p-passivity recently
introduced in Forni and Sepulchre (2017a,b). Intuitively,
the attractors of a p-dominant system are the attractors
of a p-dimensional system: a unique equilibrium for p = 0,
but possibly multiple equilibria for p = 1, and limit cycles
for p = 2. We interpret classical differential dissipativity
theory as an interconnection theory for 0-dominance, that
is, differential stability. The extension to p = 1 and p = 2
is motivated by the analysis of multistability or limit
cycles in interconnected systems. Nonlinear circuit theory
provides a realm of switching and oscillatory behaviors
designed from the simple elements of linear circuit theory interconnected with operational amplifiers (op-amp)
(Clayton and Winder, 2003). Our aim in the present paper
is to illustrate that those building blocks are the natural
building blocks of p-passive circuits. System theoretic tools
have lacked so far for the quantitative analysis and design
of such circuits. Their analysis normally rests on simplifying assumptions, time-scale separation arguments leading
to asymptotic analysis, or reductions to two-dimensional
phase portraits. In contrast, we are aiming at quantitative
and computationally tractable certificates such as those
used in the theory of linear time-invariant systems. Such
tools have made the success of robustness and performance
analysis of linear time-invariant systems. A pillar of passivity theory is the passivity theorem, which states that the
negative feedback interconnection of two passive systems is
passive. It is also well-known that switches and oscillatory
circuits require both positive and negative feedback interconnections of op-amps with passive elements. We stress
in the present paper that such interconnections fall in the
category of the p-passivity theorem, which states that the
negative feedback interconnection of a p1 -passive with a
p2 -passive circuit is p1 + p2 passive.
The differential analysis in this paper assumes smooth systems. A companion paper shows that an analog framework
exists for non-smooth systems. To account for the lack
of differentiability, differential concepts have then to be
replaced by incremental concepts. Many of the nonlinear
circuits discussed in the present paper have a non-smooth
analog that falls in the category of linear complementarity
systems studied in Miranda-Villatoro et al. (2018). The
rest of the paper is organized as follows. Section 2 provides
a brief summary of the concepts of dominance and ppassivity. Section 3 revisits standard properties of the
operational amplifier its differential passivity properties.
Section 4 revisits the basic architectures of circuits that
switch and oscillate, analyzing those systems as both positive and negative feedback interconnections of operational
amplifiers with passive linear circuits. The discussion in
Section 5 suggests that those architectures are robust and
amenable to regulation.
2. DOMINANCE AND DIFFERENTIAL PASSIVITY
We consider the nonlinear system
ẋ = f (x)
(1)
n
where x ∈ R and f is a smooth vector field. The prolonged
system consists of (1) augmented with the linearized
˙ = ∂f (x)δx, where ∂f (x) denotes the Jacobian
equation δx
linearization of f . By construction δx ∈ Rn (identified with
the tangent space of Rn ). An important notion for this
paper is the inertia (p, 0, n − p) of a symmetric matrix,
meaning that the matrix has p eigenvalues in the open left
half-plane, 0 eigenvalues on the imaginary axis, and n − p
eigenvalues in the right half-plane. The following definition
is taken from (Forni and Sepulchre, 2017a).
Definition 1. A nonlinear system (1) is p-dominant with
rate λ ≥ 0 if there exist a constant symmetric matrix P
with inertia (p, 0, n − p) and ε ≥ 0 for which the prolonged
system satisfies
⊤
˙
˙
0
P
δx
δx
≤0
(2)
P 2λP + εI
δx
δx
for all (x, δx) ∈ Rn × Rn The property is strict if ε > 0. y
Solving (2) is equivalent to finding a uniform solution P
to the linear matrix inequalities ∂f (x)⊤ P + P ∂f (x) +
2λP ≤ −ǫI for all x ∈ Rn . For a linear system ẋ = Ax,
the inequality reduces to (A + λI)⊤ P + P (A + λI) ≤ −ǫI,
which is feasible for linear systems whose eigenmodes
can be split into p dominant modes and n − p transient
modes, separated by the rate λ, (Forni and Sepulchre,
2017b, Proposition 1). For nonlinear systems p-dominance
captures the property that the asymptotic behavior of the
system is p-dimensional. This intuitive characterization is
made precise in the following result (Forni and Sepulchre,
2017a, Corollary 1).
Theorem 1. Let (1) be a strictly p-dominant system with
rate λ ≥ 0. Then, every bounded solution of (1) asymptotically converge to
• a unique fixed point if p = 0;
• a fixed point if p = 1;
• a simple attractor if p = 2, i.e. a fixed point, a set of
fixed points and their connected arcs, or a limit cycle.
In what follows we will study p-dominant systems as interconnections of open systems. We consider open systems
of the form
ẋ = f (x) + Bu , y = Cx
(3)
where u ∈ Rm and y ∈ Rq define the input and the
output to the system, respectively. B and C are matrices
of appropriate dimension. The prolonged system to (3) is
obtained by augmenting (3) with the linearized equations
˙ = ∂f (x)δx + Bδu, δy = Cδx. The following definition
δx
is taken from (Forni and Sepulchre, 2017a).
Definition 2. A nonlinear system (3) is p-passive from u
to y with rate λ ≥ 0 if there exist a constant symmetric
matrix P with inertia (p, 0, n − p) and ε ≥ 0 for which the
prolonged system satisfies
⊤
⊤
˙
˙
δy
0 I
0
P
δy
δx
δx
(4)
≤
I 0 δu
P 2λP + εI
δu
δx
δx
for all (x, δx) ∈ Rn × Rn and all (u, δu) ∈ Rm × Rm . The
property is strict if ε > 0.
y
The concept of p-passivity is related to p-dominance in the
same way as passivity is related to stability. Differential
(Forni and Sepulchre, 2013) or incremental (Pavlov and L,
2008) passivity are synonyms of 0-passivity. For a static
differentiable nonlinearity y = ϕ(u), 0-passivity simply
means monotonicity, that is positivity of its derivative: if
∂ϕ(u) ≥ 0, then δy T δu = (∂ϕ(u)δu)⊤ δu ≥ 0 for all δu.
The following p-passivity theorem is the natural extension
of the classical passivity theorem. It is taken from (Forni
and Sepulchre, 2017a, Theorem 4).
Theorem 2. Let Σ1 and Σ2 be (strictly) p1 and p2 passive,
respectively, from input ui to output yi , i ∈ {1, 2},
both with rate λ ≥ 0, Then, the negative feedback
interconnection
u1 = −y2 + v1 ,
u2 = y1 + v2
of Σ1 and Σ2 is (strictly) (p1 + p2 )-passive from v =
(v1 , v2 ) to y = (y1 , y2 ), with rate λ. The interconnection is
(strictly) (p1 + p2 )- dominant.
We observe that negative feedback preserves p-passivity
only if the two components share a common rate λ. For
linear systems of the form ẋ = Ax + Bu, y = Cx, ppassivity has a useful frequency domain characterization
in terms of the shifted transfer function G(s − λ) =
C(sI − (A + λI))−1 B, as shown by the next theorem from
Miranda-Villatoro et al. (2017).
Theorem 3. A linear system is p-passive with rate λ if and
only if the following two conditions hold,
(1) ℜ {G(jω − λ)} ≥ 0, for all, ω ∈ R ∪ {+∞}.
(2) G(s − λ) has p poles on C+ .
The property is strict if G(s− λ) has p poles in the interior
of C+
3. THE OPERATIONAL AMPLIFIER IS 0-PASSIVE
Figure 1 represents a classical model of operational amplifiers (Karki, 2000; Noseek, 2009): the RC network is
a circuit realization of a first order linear model in parallel with a voltage-controlled current source αVE , where
α ∈ (0, +∞), (Noseek, 2009). The RC network models
the finite bandwidth property of a real device and it is
connected to a static nonlinear element
i = ϕ(V )
(5)
Fig. 1. First order op-amp model with saturation: a)
internal structure, b) symbolic representation.
to account for the bounded range of V0 = x. The static
nonlinear element is a smooth and odd nonlinearity modelled for instance as follows:
i = ϕ(V ) := η sinh(βV )
(6)
where η > 0 and β > 0 are suitable parameters. The
results of this paper hold for any stiffening nonlinearity ϕ
satisfying the following assumption (Stan and Sepulchre,
2007).
Assumption 4. The static nonlinearity ϕ : R → R is an
odd function such that ∂ϕ(y)
∂y ∈ [0, +∞). Furthermore, for
any k > 0, there exists a r > 0 such that
yϕ(y) − ky 2 > 0, for all |y| > r .
(7)
For example, ϕ(y) = y 2n+1 , for n ∈ N, ϕ(y) = arctanh(y)
and ϕ(y) = sinh(y) all satisfy Assumption 4. The firstorder model of the op-amp in Figure 1 has the state-space
model
ẋ = − 1 x − 1 ϕ (x) + α V
E
Σop :
(8)
R C
C0
C0
V = x 0 0
0
which, notably, admits the block diagram representation
of the Lur’e system in Figure 2. The transfer function of
of behaviors, enabled by the interconnection of the op-amp
with suitable linear networks. The circuits in this paper
will only include interconnections of op-amps with linear
stable networks
ż = Az + Bu, y = Cz .
(10)
where z ∈ Rn , u ∈ R, y ∈ R are the state, input, and
output of a generic linear network, respectively. A, B and
C are constant matrices of appropriate dimensions. Such
interconnections always lead to bounded behaviors:
Theorem 5. Suppose that A is a Hurwitz matrix. Under
Assumption 4, the trajectories of the system defined by
(8), (10), and the interconnection rule
VE = ±Cz + Vr u = V0
(11)
are all bounded, for any constant voltage Vr ∈ R.
Proof. Let V : R × Rn → R be the positive definite
function
1
1
V (x, z) = x2 + z ⊤ P z,
(12)
2
2
where P = P ⊤ > 0 satisfies A⊤ P + P A ≤ −Q, for
some Q = Q⊤ > 0. Then, taking η = R01C0 , β = C10 ,
ρ = αkCk + 2kP Bk, ε > 0, and µQ given by the smallest
eigenvalue of Q, we have
V̇ (x, z) = −ηx2 − z ⊤ Qz − βxϕ(x) ± αxCz + 2z ⊤ P Bx
≤ −ηx2 − µQ kzk2 − βxϕ(x) + ρ|x|kzk
ρ
≤ −ηx2 − µQ kzk2 − βxϕ(x) + |x|2 + ρεkzk2 .
ε
µ
Setting ε = 2ρQ yields
ρ
µQ
kzk2 − βxϕ(x) + |x|2 .
V̇ (x, z) ≤ −ηx2 −
2
ε
From (7), there exists r > 0 such that, for all |x| >
µ
r, V̇ (x, z) ≤ −ηx2 − 2Q kzk2. Boundedness of solutions
follows.
In the next sections we will design particular interconnections based on Theorems 1 and 2.
4.2 p-Passivity and interconnections
We consider the interconnection of op-amp (8) and linear
networks of the form (10) typically in positive or negative
feedback, as show in Figure 3. Passivity is a theory of
Fig. 2. Open-loop operational amplifier model
the linear part Σ0
G0 (s) =
1
C0
s+
1
R0 C 0
(9)
is a strictly 0-passive network (by Theorem 3) with rate
λ ∈ [0, R01C0 ). The op-amp is thus given by the negative feedback interconnection of a strictly 0-passive linear
system with a static 0-passive nonlinearity (under Assumption 4). Therefore, by Theorem 2, the closed loop
is a strictly 0-passive system from VE to V0 with rate
λ ∈ [0, R01C0 ). The same representation would hold if the
RC circuit was replaced by any 0-passive network.
4. SWITCHES AND OSCILLATORS VIA FEEDBACK
AMPLIFIERS
4.1 Feedback and boundedness
The great versatility of the op-amp comes from its interconnection properties. The device allows for a wide range
Fig. 3. Feedback loops of a circuit with an operational
amplifier
negative feedback. In order to apply Theorem 2 to positive
feedback interconnections, we consider the reverted output
ȳ = −y and interpret the positive feedback interconnection
VE = +y + Vr as negative feedback interconnection of the
reverted output:
VE = −ȳ + Vr , u = V0 .
(13)
We note that the network in Figure 4 is strictly 0-passive
from u to y = z and strictly 1-passive from u to ȳ = −z.
a
Indeed, define a = RR11R+R
, b = Ra1C1 , and c = 1. Then the
a C1
4.4 2-Passive circuits
By Theorem 1, oscillatory circuits may arise from the
interconnection of the op-amp with strictly 2-passive networks. As an illustration, consider the negative feedback
of the op-amp with the RC network in Figure 6. The RC
Fig. 4. RC network that is both 0-passive and 1-passive
from different ports
transfer functions from u to y reads
cb
G(s) =
(14)
s+a
whereas the transfer function from u to ȳ reads
cb
Ḡ(s) = −
(15)
s+a
By Theorem 3, G(s) is strictly 0-passive for any rate
λ ∈ [0, a). On the other hand, Ḡ(s) is strictly 1-passive
for any rate λ ∈ (a, +∞).
4.3 1-Passive circuits
By Theorem 1, multistable circuits that switch among
several fixed points may arise from the interconnection
of the op-amp with strictly 1-passive networks. As an
illustration, consider the positive feedback of the op-amp
with the RC network in Figure 4. The RC network is
Fig. 5. An op-amp in positive feedback with a passive
network.
strictly 1-passive from u to ȳ = −z with rate λ ∈ (a, +∞),
and takes the role of Σup in Figure 3. The op-amp is 0passive with rate λ ∈ [0, R01C0 ). Thus, for
1
> a,
(16)
R0 C0
the two systems share a common interval for their λ
rates 1 . By Theorem 2 the closed loop system in Figure 5
is strictly 1-passive from Vr to V0 with rate λ ∈ (a, R01C0 ).
Theorem 5 guarantees boundedness of the closed-loop
trajectories for any constant input Vr . Therefore, Theorem
1 guarantees asymptotic convergence of all trajectories
to some fixed point. In particular, taking Vr = 0 for
simplicity, the closed loop is bistable for
∂ϕ(x)
αR1
1
+
,
(17)
<−
∂x x=0
R0
R1 + Ra
which guarantees the existence of three equilibrium points
(two stable nodes and a saddle).
1
(16) requires that the op-amp dynamics is faster than the dynamics
of the external network, as usual in applications.
Fig. 6. RC oscillator circuit.
network admits the state-space representation
#
"
" #
−2 1 0
1
1
1
1 −2 1 , B =
0 , C = [0 0 1] .
A=
R1 C1 0 1 −1
R1 C1 0
The transfer function G(s) from u to y reads
G(s) =
s3 +
1
(R1 C1 )3
6
5
2
R1 C1 s + (R1 C1 )2 s
+
1
(R1 C1 )3
.
(18)
Denoting by pi the i-th pole of G and by βi = |ℜ{pi }|
the magnitude of the real part of the poles of G (without
loss of generality, we assume 0 ≤ β1 ≤ β2 < β3 ),
the RC network
n is strictly 2-passive
from u to y with
o
rate λ ∈ max β2 , β1 +β32 +β3 , β3 . The network is thus
constrained to a negative feedback interconnection, taking
the role of Σdown in Figure 3. For
β1 + β2 + β3
1
(19)
> max β2 ,
R0 C0
3
the op-amp and the RC network share a common interval
for their λ rates. By Theorem 2 the closed loop system in
Figure
to V0 with rate
6 is nthus strictly 2-passive
o
n from Vr o
β1 +β2 +β3
1
λ ∈ max β2 ,
, min β3 , R0 C0 . Theorem 5
3
guarantees boundedness of the closed loop trajectories for
any constant input Vr , thus Theorem 1 guarantees that
the trajectories of the four dimensional closed-loop circuit
all converge to a simple attractor. For R1 = 3.3KΩ and
C1 = 200µF , G(s) has poles p1 = −0.3, p2 = −2.35
and p3 = −4.92. Therefore, strict 2-passivity holds for
λ ∈ (2.52, 4.92). For op-amp parameters α = 0.1, R0 =
x 5
1M Ω, C0 = 15.9nF , ϕ(x) = ( 12
) condition (19) holds.
These specific parameters also ensure that the unique fixed
point at the origin is unstable. Thus, by Theorem 1, every
trajectory converges asymptotically to a limit cycle. The
steady-state is an oscillation, as shown in Figure 7.
5. MIXED FEEDBACK AND MODULATION
The mixed feedback amplifier is a classical device of nonlinear circuit theory (Chua et al., 1987). It combines positive and negative feedback around an operational amplifier
to create nonlinear behaviors. We illustrate this flexibility with the simple system of Figure 8. The two linear
networks correspond to two copies of the RC network in
(2) G(s) is strictly 2-passive with rate λ ∈
for a1 > a2 .
a2 b2 −a1 b1
a2 −a1 , +∞
The conditions above reveal that mixed feedback allows
for both 0-passivity and 2-passivity. The network behavior
can be modulated from monostable to oscillatory via
parameter variations. Figure 9 shows the degree of ppassivity of the closed loop for different values 0Ω <
Ra , Rb < 3KΩ. Indeed, transitions from monostable to
oscillatory regimes are obtained by the variation of one
of the two resistances. As a final illustration, we consider
Fig. 7. Output of the RC circuit of Figure 6, with R1 =
3.3KΩ and C1 = 200µF .
Fig. 8. Mixed feedback with switches.
Figure 4. The behavior of the closed loop is dictated by
the interconnection pattern of the switches Sa and Sb . If
Sa is closed and Sb is open then the network reduces to
the one in Figure 5; the closed loop is 1-passive. If Sa is
open and Sb is closed, then the closed loop is 0-passive.
When both switches are closed, the feedback circuit is not
necessarily 1-passive because the rates of the two networks
may not be compatible. In fact, a suitable selection of the
network parameters lead to richer behaviors. Take Sa and
Sb both closed and define a1 = Ra1C1 , a2 = Rb1C2 , and
bi = ai + Ri1Ci , i = 1, 2. With these data, the upper network
is strictly 1-passive from V0 to −z1 and the lower network
is strictly 0-passive from V0 to z2 , respectively with rates
λup ∈ (b1 , +∞) and λdown ∈ [0, b2 ). A common rate λ can
be found for b1 < min{b2 , R01C0 }, since the op-amp is 0passive with rate λop ∈ [0, R01C0 ). In this case, the feedback
circuit is 1-passive, by Theorem 2. In contrast, Theorem
2 cannot be used for b1 > b2 because of the absence of
a common rate. However, the aggregate transfer function
reads
(a2 − a1 )s + a2 b1 − a1 b2
G(s) =
,
(20)
(s + b1 )(s + b2 )
which has positive real part if and only if there exist
λ ∈ [0, b2 ) ∪ (b1 , +∞) such that
(a2 − a1 )λ < a2 b2 − a1 b1
(21)
(a2 − a1 )λ < a2 b1 − a1 b2 .
(22)
Indeed,
h
−a1 b1
(1) G(s) is strictly 0-passive 2 with rate λ ∈ 0, a2ab22 −a
1
for a2 > a1 and
2
a2 b2 −a1 b1
a2 −a1
> 0;
Indeed, since b1 > b2 , it follows that a1 b1 > a1 b2 and a2 b1 > a2 b2 .
Hence, a2 b2 −a1 b1 < a2 b1 −a1 b2 and conditions (21) and (22) reduce
Fig. 9. p-passivity of the circuit of Figure 8 with both
switches closed, as a function of the resistors Ra and
Rb . Gaps correspond to lack of p-passivity.
parameters R1 = R2 = 3.3KΩ, Ra = Rb = 1KΩ, C1 =
100µF , C2 = 200µF , R0 = 1M Ω, C0 = 15.9nF and α = 1.
For these parameters G(s) in (20) is strictly 2-passive.
When both switches are closed the origin is the only
equilibrium point and is unstable. Hence, by Theorems
1 and 5 we conclude the existence of a limit cycle. Figure
10 shows transitions among different behaviors, driven by
the switches.
6. CONCLUSIONS
We have analyzed basic examples of circuits that switch
and oscillate as interconnections of linear circuits and
operational amplifiers. The approach builds upon dominance theory and p-passivity. The saturated op-amp model
guarantees boundedness of trajectories in closed loop and
allows for positive and negative feedback interconnections
with linear p-passive networks. Specific interconnections
lead to 1- and 2-passive networks, leading to a tractable
analysis of bistability and oscillations in possibly highdimensional models. The stability analysis in this paper
is based on solving linear matrix inequalities very much as
in the stability analysis of linear systems. Such a computational framework suggests many possible extensions to
analyze the performance and robustness of switching and
oscillatory circuits in the same way as for linear systems.
1 b1
to λ < a2ab2 −a
. Moreover, since b1 > b2 it follows that λ < b2 .
2 −a1
From this last observation, together with λ ∈ [0, b2 ) ∪ (b1 , +∞) and
Proposition 3 it follows that G(s) is 0-passive.
Fig. 10. Transitions among different behaviors of circuit of
Figure 8 driven by the switches Sa and Sb ; 0 - open
switch, 1 - closed switch. Recall that for (Sa , Sb ) =
(0, 1) the circuit is 0-passive, for (Sa , Sb ) = (1, 0) the
circuit is 1-passive, and for (Sa , Sb ) = (1, 1) the circuit
is 2-passive.
REFERENCES
Aghannan, N. and Rouchon, P. (2003). An intrinsinc
observer for a class of Lagrangian systems. IEEE
Transactions on Automatic Control, 48(6).
Chua, L.O., Desoer, C.A., and Kuh, E.S. (1987). Linear
and nonlinear circuits. Mc-Graw Hill.
Clayton, G. and Winder, S. (2003). Operational Amplifiers. EDN Series for Design Engineers. Elsevier.
Forni, F. and Sepulchre, R. (2013). On differentially
dissipative dynamical systems. In 9th IFAC Symposium
on Nonlinear Control Systems, 15–20. Toulouse, France.
Forni, F. and Sepulchre, R. (2014). A differential Lyapunov framework for contraction analysis. IEEE Transactions on Automatic Control, 59(3), 614–628.
Forni, F. and Sepulchre, R. (2017a). Differential dissipativity theory for dominance analysis. Submitted to IEEE Transactions on Automatic Control,
http://arxiv.org/abs/1710.01721.
Forni, F., Sepulchre, R., and van der Schaft, A.J. (2013).
On differential passivity of physical systems. In 52rd
IEEE Conference on Decision and Control, 6580–6585.
Florence, Italy.
Forni, F. and Sepulchre, R. (2017b). A dissipativity theorem for p-dominant systems. In 56th IEEE Conference
on Decision and Control. Melbourne, Australia.
Jouffroy, J. and Fossen, T.I. (2010). A tutorial on incremental stability analysis using contraction theory.
Modeling, Identification and Control, 31(3), 93–106.
Karki, J. (2000). Effect of parasitic capacitance in op
amp circuits. Technical Report SLOA013A, Texas
Instruments.
Lohmiller, W. and Slotine, J.J.E. (1998). On contraction
analysis for nonlinear systems. Automatica, 34(6), 683–
696.
Miranda-Villatoro, F.A., Forni, F., and Sepulchre, R.
(2017).
Analysis of Lur’e dominant systems in
the frequency domain. Submitted to Automatica,
http://arxiv.org/abs/1710.01645.
Miranda-Villatoro, F.A., Forni, F., and Sepulchre, R.
(2018).
Dominance analysis of linear complementarity systems.
Submitted to MTNS 2018,
http://arxiv.org/abs/1802.00284.
Noseek, J.A. (2009). Circuit elements, modeling and
equation formulation. In The Circuits and Filters
Handbook: Fundamentals of Circuits and Filters, 13–1,
13–11.
Pavlov, A. and L, M. (2008). Incremental passivity and
output regulation. Systems & Control Letters, 57, 400–
409.
Pavlov, A., Van De Wouw, N., and Nijmeijer, H. (2005).
Convergent systems: analysis and synthesis. In Control
and observer design for nonlinear finite and infinite
dimensional systems, 131–146. Springer.
Stan, G.B. and Sepulchre, R. (2007). Analysis of interconnected oscillators by dissipativity theory. IEEE
Transactions on Automatic Control, 52(2), 256–270.
van Der Schaft, A.J. (2013). On differential passivity. In
9th IFAC Symposium on Nonlinear Control Systems,
21–25. Toulouse, France.