Вы находитесь на странице: 1из 125

The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 1

The port-Hamiltonian approach to


physical system modeling and control
Arjan van der Schaft, University of Groningen
Hans Zwart, University of Twente

In collaboration with Bernhard Maschke, Romeo Ortega,


Jacquelien Scherpen, Stefano Stramigioli, Alessandro Macchelli,
Peter Breedveld, Dimitri Jeltsema, Morten Dalsmo, Guido
Blankenstein, Damien Eberard, Goran Golo, Ram Pasumarthy,
Javier Villegas, Gerardo Escobar, Guido Blankenstein, Aneesh
Venkatraman, Rostyslav Polyuga ..

Part I : Network Modeling and Analysis


Part II : Control of Port-Hamiltonian Systems
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 2

Part I : Network Modeling and Analysis


1. Motivation and context

2. From passive systems to port-Hamiltonian systems

3. Some properties of port-Hamiltonian systems

4. Intro to distributed-parameter port-Hamiltonian systems

5. Mixed lumped- and distributed-parameter port-Hamiltonian


systems

6. Lumping of distributed-parameter physical systems, and


port-Hamiltonian dynamics on graphs and k -complexes.

7. Model reduction

8. Conclusions of Part I
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 3

From passive systems to port-Hamiltonian systems


A square nonlinear system

ẋ = f (x) + g(x)u, u ∈ Rm
Σ:
y = h(x), y ∈ Rm

where x ∈ Rn are coordinates for an n-dimensional state space X , is


passive if there exists a storage function H : X → R with
H(x) ≥ 0 for every x, such that
Z t2
H(x(t2 )) − H(x(t1 )) ≤ uT (t)y(t)dt
t1

for all solutions (u(·), x(·), y(·)) and times t1 ≤ t2 .


The system is lossless if ≤ is replaced by =.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 4

If H is differentiable then ’passive’ is equivalent to


d
H ≤ uT y
dt
which reduces to (Willems, Hill-Moylan)
∂T H
∂x (x)f (x) ≤ 0

h(x) = g T (x) ∂H
∂x (x)

while in the lossless case ≤ is replaced by =.


In the linear case
ẋ = Ax + Bu
y = Cx
is passive if there exists a quadratic storage function H(x) = 12 xT Qx,
with Q = QT ≥ 0 satisfying the LMIs
AT Q + QA ≤ 0, C = BT Q
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 5

Every linear passive system with storage function H(x) = 21 xT Qx,


satisfying
ker Q ⊂ ker A

can be rewritten as a linear port-Hamiltonian system

ẋ = (J − R)Qx + Bu, J = −J T , R = RT ≥ 0
y = B T Qx,

in which case the storage function H(x) = 21 xT Qx is called the


Hamiltonian.

• Passive linear systems are thus port-Hamiltonian with


non-negative Hamiltonian.

• Conversely every port-Hamiltonian system with


non-negative Hamiltonian is passive.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 6

Mutatis mutandis ’most’ nonlinear lossless systems can be written


as a port-Hamiltonian system

ẋ = J(x) ∂H
∂x (x) + g(x)u

y = g T (x) ∂H
∂x (x)
∂H
with J(x) = −J T (x) and ∂x (x) the column vector of partial
derivatives. Note that
∂H
ẋ = J(x) (x)
∂x
is the internal Hamiltonian dynamics known from physics, which in
classical mechanics can be written as
∂H
q̇ = ∂p (q, p)
ṗ = − ∂H
∂q (q, p)

with the Hamiltonian H the total (kinetic + potential) energy.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 7

Similarly, most nonlinear passive systems can be written as a


port-Hamiltonian system (with dissipation)

ẋ = [J(x) − R(x)] ∂H
∂x (x) + g(x)u

y = g T (x) ∂H
∂x (x)

with R(x) = RT (x) ≥ 0 specifying the energy dissipation

d ∂T H ∂H
H=− (x)R(x) (x) + uT y ≤ uT y
dt ∂x ∂x
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 8

However, in network modeling it is the other way around: one


derives the system in port-Hamiltonian form (and if the Hamiltonian
H ≥ 0 then it is the storage function of a passive system).
The matrix J(x) corresponds to the internal power-conserving
interconnection structure of physical systems due to:

• Basic conservation laws such as Kirchhoff’s laws.


• Powerless constraints; kinematic constraints.
• Transformers, gyrators, exchange between different types of
energy.

The matrix R(x) corresponds to the internal energy dissipation


in the system (due to resistors, damping, viscosity, etc.)
Main message: start with port-Hamiltonian models instead of
passive models. Closer to physical modeling, and capturing more
information than just the energy-balance of passivity.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 9

A bit of port-based network modeling


The passivity framework considers a system component, and its
power-exchange with other system components:

d
H ≤ uT y
dt
The feedback interconnection of two passive systems

d d
dt H1 ≤ uT1 y1 , dt H2 ≤ uT2 y2

u1 = −y2 + v1 , u2 = y1 + v2

leads to an interconnected system that is again passive, since

d
(H1 + H2 ) ≤ uT1 y1 + uT2 y2 = v1T y1 + v2T y2
dt
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 10

The feedback interconnection is a typical example of a


power-conserving interconnection (total power is zero):

uT1 y1 + uT2 y2 + v1T y1 + v2T y2 = 0


In port-based modeling, e.g. bond graphs, one looks at the
system as the power-conserving interconnection of ideal basic
system components: (energy-) storage elements, resistive
elements, transformers, gyrators, constraints, etc.

• The Hamiltonian of the resulting port-Hamiltonian system is


the sum of the energies of the storage elements.

• The J - and B -matrix is determined by the transformers,


gyrators, constraints, and the power-conserving
interconnection.

• The R-matrix is determined by the resistive elements, and the


way they are connected.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 11

An k dimensional storage element is determined by a


k -dimensional state vector x = (x1 , · · · , xk ) and a Hamiltonian
H(x1 , · · · , xk ) (energy storage), defining the lossless system

ẋi = fi , i = 1, · · · , k
∂H
ei = ∂xi (x1 , · · · , xk )
d
Pk
dt H = i=1 fi ei

Such a k - dimensional storage component is written in vector


notation as a port-Hamiltonian system with J = 0, R = 0, and B = I :

ẋ = f
∂H
e = ∂x (x)

The elements of x are called energy variables, those of ∂H


∂x (x)
co-energy variables. Furthermore the elements of f are flow
variables, and of e effort variables.
Note that eT f is power.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 12

Example: The ubiquitous mass-spring-damper system:


Two storage elements:

• Spring Hamiltonian Hs (q) = 21 kq 2 (potential energy)

q̇ = fs = velocity
dHs
es = dq (q) = kq = force

1 2
• Mass Hamiltonian Hm (p) = 2m p (kinetic energy)

ṗ = fm = force
dHm p
em = dp (p) = m = velocity
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 13

interconnected by

fs = em = y, fm = −es + u

(power-conserving since fs es + fm em = uy ) yields the


port-Hamiltonian system
      
∂H
q̇ 0 1 (q, p) 0
  =    ∂q  (q, p) +   u
∂H
ṗ −1 0 ∂p (q, p) 1


h i ∂H (q, p)
y = 0 1  ∂q 
∂H
∂p (q, p)

with
H(q, p) = Hs (q) + Hm (p)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 14

Addition of the damper

dR 1 2
ed = = cfd , R(fd ) = cfd (Rayleigh function)
dfd 2

via the extended interconnection

fs = em = fd = y, fm = es − ed + u

leads to the mass-damper-spring system in port-Hamiltonian


form
         
∂H
q̇ 0 1 0 0 (q, p) 0
  = ( − )  ∂q  +  u
∂H
ṗ −1 0 0 c ∂p (q, p) 1
 
h i ∂H (q, p)
y = 0 1  ∂q 
∂H
∂p (q, p)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 15

Example: Electro-mechanical systems

      
∂H
q̇ 0 1 0 ∂q (q, p, φ) 0
  
 ṗ  = −1 0

∂H
   ∂H
0   ∂p (q, p, φ)
 
0 V,
+  I= (q, p, φ)
    ∂ϕ
ϕ̇ 0 0 − R1 ∂H
∂ϕ (q, p, φ) 1
Coupling electrical/mechanical domain via Hamiltonian H(q, p, φ).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 16

Example: LC circuits. Two inductors with magnetic energies


H1 (ϕ1 ), H2 (ϕ2 ) (ϕ1 and ϕ2 magnetic flux linkages), and capacitor
with electric energy H3 (Q) (Q charge). V denotes the voltage of
the source.

L1 L2
C

ϕ1 Q ϕ2
V

Question: How to write this circuit as a port-Hamiltonian system


in a modular way?
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 17

Hamiltonian equations for the components of the LC-circuit:

Inductor 1 ϕ̇1 = f1 (voltage)


∂H1
(current) e1 = ∂ϕ1

Inductor 2 ϕ̇2 = f2 (voltage)


∂H2
(current) e2 = ∂ϕ2

Capacitor Q̇ = f3 (current)
∂H3
(voltage) e3 = ∂Q

All are port-Hamiltonian systems with J = 0 and g = 1.

If the elements are linear then the Hamiltonians are quadratic, e.g.
ϕ1
H1 (ϕ1 ) = 2L1 1 ϕ21 , and ∂H
∂ϕ1
1
= L1 = current , etc.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 18

Kirchhoff’s interconnection laws in f1 , f2 , f3 , e1 , e2 , e3 , f = V, e = I are


    
−f1 0 0 1 −1 e1
    
−f   0 0 −1 0  e 
 2    2
 =  
−f3  −1 1 0 0  e3 
 
  
e 1 0 0 0 f
Substitution of eqns. of components yields port-Hamiltonian
system  
    ∂H  
ϕ˙1 0 0 −1  ∂ϕ1  1
      
ϕ˙2  = 0 0 ∂H  +   f
   1   ∂ϕ2  0
 
 
Q̇ 1 −1 0 ∂H 0
∂Q
∂H
e = ∂ϕ1

with H(ϕ1 , ϕ2 , Q) := H1 (ϕ1 ) + H2 (ϕ2 ) + H3 (Q) total energy.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 19

However, this class of port-Hamiltonian systems is not closed


under interconnection:

Figure 1: Capacitors and inductors swapped

Interconnection leads to algebraic constraints between the state


variables Q1 and Q2 .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 20

How to model DAEs as port-Hamiltonian systems ?


Intermezzo: what is the appropriate generalization of the
skew-symmetric mapping J ? Answer: Dirac structures.
(’From skew-symmetric mappings to skew-symmetric relations’)
Power is defined by
P = e(f ) =: < e | f >= eT f, (f, e) ∈ V × V ∗ .
where the linear space V is called the space of flows f (e.g.
currents), and V ∗ the space of efforts e (e.g. voltages).
Symmetrized form of power is the indefinite bilinear form ,  on
V × V ∗:
(f a , ea ), (f b , eb )  := < ea | f b > + < eb | f a >,

(f a , ea ), (f b , eb ) ∈ V × V ∗ .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 21

Definition 1 (Weinstein, Courant, Dorfman) A (constant)


Dirac structure is a subspace
D ⊂ V × V∗
such that
D = D⊥ ,
where ⊥ denotes orthogonal complement with respect to the
bilinear form , .
For a finite-dimensional space V this is equivalent to
(i) < e | f > eT f = 0 for all (f, e) ∈ D ,
(ii) dim D = dim V .
For any skew-symmetric map J : V ∗ → V its graph
{(f, e) ∈ V × V ∗ | f = Je}
is a Dirac structure !
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 22

For many systems, especially those with 3-D mechanical


components, the interconnection structure will be modulated by
the energy or geometric variables.
This leads to the notion of non-constant Dirac structures on
manifolds.

Definition 2 Consider a smooth manifold M . A Dirac structure on


M is a vector subbundle D ⊂ T M ⊕ T ∗ M such that for every x ∈ M
the vector space
D(x) ⊂ Tx M × Tx∗ M

is a Dirac structure as before.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 23

Geometric definition of a port-Hamiltonian system

ẋ fx f
H D(x)
∂H ex e
∂x (x)
Figure 2: Port-Hamiltonian system

The dynamical system defined by the DAEs


∂H
(−ẋ(t) = fx (t), (x(t)) = ex (t), f (t), e(t)) ∈ D(x(t)), t∈R
∂x
is called a port-Hamiltonian system.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 24

Particular case is a Dirac structure D(x) ⊂ Tx X × Tx∗ X × F × F ∗


given as the graph of the skew-symmetric map
    
f −J(x) −g(x) e
 x =  x ,
e g T (x) 0 f

leading (fx = −ẋ, ex = ∂H


∂x (x)) to a port-Hamiltonian system as
before
ẋ = J(x) ∂H ∂x (x) + g(x)f, x ∈ X , f ∈ Rm

e = g T (x) ∂H
∂x (x), e ∈ Rm
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 25

Energy-dissipation is included by terminating some of the ports


by static resistive elements

fR = −F (eR ), where eTR F (eR ) ≥ 0, for all eR .

If H ≥ 0 then the system stays passive with respect to the


remaining flows and efforts f and e:
d
H ≤ eT f
dt
This leads, e.g. for linear damping, to input-state-output
port-Hamiltonian systems in the form

 ẋ = [J(x) − R(x)] ∂H (x) + g(x)f
∂x
 e = g T (x) ∂H (x)
∂x

where J(x) = −J T (x), R(x) = RT (x) ≥ 0 are the interconnection and


damping matrices, respectively.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 26

Example: Mechanical systems with kinematic constraints

Constraints on the generalized velocities q̇ :

AT (q)q̇ = 0.

This leads to constrained Hamiltonian equations

∂H
q̇ = ∂p (q, p)
ṗ = − ∂H
∂q (q, p) + A(q)λ + B(q)u
0 = AT (q) ∂H
∂p (q, p)
y = B T (q) ∂H
∂p (q, p)

with H(q, p) total energy, and λ the constraint forces.


By elimination of the constraints and constraint forces one derives
a port-Hamiltonian model without constraints.

Can be extended to general multi-body systems.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 27

Intermezzo: Relation with classical Hamiltonian equations

∂H
ẋ = J (x)
∂x
with constant or ’ integrable’J - matrix admits coordinates
x = (q, p, r) in which
 
0 I 0 q̇ = ∂H∂p (q, p, r)
 
J = −I 0 0 , ∂H

 ṗ = − ∂q (q, p, r)
0 0 0 ṙ = 0
For constant or integrable Dirac structure one gets Hamiltonian
DAEs
q̇ = ∂H∂p (q, p, r, s)
ṗ = − ∂H
∂q (q, p, r, s)
ṙ = 0
∂H
0 = ∂s (q, p, r, s)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 28

Some properties of port-Hamiltonian systems


Port-Hamiltonian systems modeling encodes more
information than energy-balance.

The Dirac structure determines all the Casimir functions


(conserved quantities which are independent of H ).
Example: In the first LC circuit the total flux φ1 + φ2 is a
conserved quantity that is solely determined by the circuit topology.
(In Part II this will be used for set-point control.)
Furthermore, the Dirac structure directly determines the algebraic
constraints.
Example: In the second LC-circuit the state variables Q1 and Q2
are related by
Q1 Q2
=
C1 C2
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 29

Any power-conserving interconnection of


port-Hamiltonian systems is again port-Hamiltonian
• The resulting Hamiltonian is the sum of the Hamiltonians of
the individual systems.

• The Dirac structure is determined by the Dirac structures of


the individual systems, and the way they are interconnected.

• The resistive structure is determined by the resistive structures


of the individual systems.

Conclusion: port-Hamiltonian systems theory provides a


modular framework for modeling and analysis of complex
multi-physics lumped-parameter systems.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 30

Network modeling is prevailing in modeling and simulation of


lumped-parameter physical systems (multi-body systems, electrical
circuits, electro-mechanical systems, hydraulic systems, robotic
systems, etc.), with many advantages:

• Modularity and flexibility. Re-usability (‘libraries’).

• Multi-physics approach.

• Suited to design/control.

Disadvantage of network modeling: it generally leads to a large set


of DAEs, seemingly without any structure.
Port-based modeling and port-Hamiltonian system theory
identifies the underlying structure of network models of
physical systems, to be used for analysis, simulation and
control.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 31

Distributed-parameter port-Hamiltonian systems

fa fb
ea eb
a b
Figure 3: Simplest example: Transmission line

Telegrapher’s equations define the boundary control system


∂Q ∂ ∂ φ(z,t)
∂t (z, t) = − ∂z I(z, t) = − ∂z L(z)
∂φ ∂ ∂ Q(z,t)
∂t (z, t) = − ∂z V (z, t) = − ∂z C(z)
fa (t) = V (a, t), e1 (t) = I(a, t)
fb (t) = V (b, t), e2 (t) = I(b, t)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 32

Transmission line as port-Hamiltonian system


Define internal flows fx = (fE , fM ) and efforts ex = (eE , eM ):

electric flow fE : [a, b] → R


magnetic flow fM : [a, b] → R
electric effort eE : [a, b] → R
magnetic effort eM : [a, b] → R
together with external boundary flows f = (fa , fb ) and boundary
efforts e = (ea , eb ). Define the infinite-dimensional Dirac structure
    

fE 0 ∂z e
  =   E 

fM ∂z 0 eM

   
fa,b eE|a,b
  =  
ea,b eM |a,b
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 33

This defines a Dirac structure on the space of internal flows and


efforts and boundary flows and efforts.
Substituting (as in the lumped-parameter case)

fE = − ∂Q∂t

∂ϕ  x
f = −ẋ
fM = − ∂t


Q ∂H
eE = C = ∂Q
 ∂H
e =
eM = ϕ
= ∂H  x ∂x
L ∂ϕ

with, for example, quadratic energy density

1 Q2 1 ϕ2
H(Q, ϕ) = +
2 C 2 L
we recover the telegrapher’s equations.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 34

Of course, the telegrapher’s equations can be rewritten as the


linear wave equation

∂2Q ∂ ∂I ∂ ∂ φ
∂t2 = − ∂z ∂t = − ∂z ∂t L =

2
∂ 1 ∂φ ∂ 1 ∂ Q 1 ∂ Q
− ∂z L ∂t = ∂z L ∂z C = LC ∂z 2

(if L(z), C(z) do not depend on z ), or similar expressions in φ, I or V .


The same equations hold for a vibrating string, or for a
compressible gas/fluid in a one-dimensional pipe.

Basic question:

Which of the boundary variables fa , fb , ea , eb can be considered to


be inputs, and which outputs ? See Lecture of Hans Zwart.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 35

Example 2: Shallow water equations; distributed-parameter


port-Hamiltonian system with non-quadratic Hamiltonian

The dynamics of the water in an open-channel canal can be


described by      
h v h h
∂t   +   ∂z  = 0

v g v v

with h(z, t) the height of the water at position z , and v(z, t) the
velocity (and g gravitational constant).
This can be written as a port-Hamiltonian system by recognizing
the total energy
b
1
Z
H(h, v) = [hv 2 + gh2 ]dz
2 a
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 36

yielding the co-energy functionsa

eh = ∂H
∂h = 12 v 2 + gh Bernoulli function
∂H
ev = ∂v = hv mass flow

It follows that the shallow water equations can be written, similarly


to the telegraphers equations, as

∂h ∂ ∂H
∂t (z, t) = − ∂z ∂v

∂v ∂ ∂H
∂t (z, t) = − ∂z ∂h

with boundary variables −hv|a,b and ( 21 v 2 + gh)|a,b .


a DanielBernoulli, born in 1700 in Groningen as son of Johann Bernoulli, profes-
sor in mathematics at the University of Groningen and forerunner of the Calculus
of Variations (the Brachistochrone problem).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 37

Paying tribute to history:

Figure 4: Johann Bernoulli, professor in Groningen 1695-1705.

Figure 5: Daniel Bernoulli, born in Groningen in 1700.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 38

We obtain the energy balance


b
d 1
Z
[hv 2 + gh2 ]dz = −(hv)( v 2 + gh)|ba
dt a 2

which can be rewritten as

−v( 21 gh2 )|ba − v( 12 hv 2 + 12 gh2 ))|ba =

velocity × pressure + energy flux through the boundary


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 39

Conservation laws

All examples sofar have the same structure

∂α1 ∂ ∂H ∂
∂t (z, t) = − ∂z ∂α2 = − ∂z β2

∂α2 ∂ ∂H ∂
∂t (z, t) = − ∂z ∂α1 = − ∂z β1

with boundary variables β1 |{a,b} , β2 |{a,b} , corresponding to two


coupled conservation laws:

d
Rb Rb ∂
dt a
α1 = − β
a ∂z 2
= β2 (a) − β2 (b)

d
Rb Rb ∂
dt a
α2 = − β
a ∂z 1
= β1 (a) − β1 (b)

(In the transmission line, α1 and α2 is charge- and flux-density, and


β1 , β2 voltage V and current I , respectively.)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 40

For some purposes it is illuminating to rewrite the equations in


terms of the co-energy variables β1 , β2 :
   2 2
   2 2
 
∂β1 ∂ H ∂ H ∂α1 ∂ H ∂ H ∂β2
2 ∂α1 α2   ∂t  ∂α21 ∂α1 α2   ∂z 
 ∂t  =  ∂α
2
1 = − 
∂β2 ∂ H ∂2H ∂α2 ∂2H ∂2H ∂β1
∂t ∂α2 α1 ∂α22 ∂t ∂α2 α1 ∂α22 ∂z

For the transmission line this yields


    
∂V 1 ∂V
0
 ∂t  = −  C   ∂z 
∂I 1 ∂I
∂t L 0 ∂z

The matrix is called the characteristic matrix, whose eigenvalues


1 1
are the characteristic velocities √LC and − √LC corresponding to
the characteristic eigenvectors (and curves).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 41

For the shallow water equations this yields


    
∂β1 ∂β1
v g
 ∂t  = −    ∂z 
∂β2 ∂β2
∂t h v ∂z

with
1 2
β1 = v + gh, β2 = hv
2
being the Bernoulli function and mass flow, respectively.

This corresponds to two characteristic velocities v ± gh, which
are, like in the transmission line case, of opposite sign (subcritical
or fluvial flow) if
v 2 ≤ gh
Because the Hamiltonian is non-quadratic, and thus the pde’s are
nonlinear, the characteristic curves may intersect, corresponding
to shock waves.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 42

Mixed lumped- and distributed-parameter


port-Hamiltonian systems
Typical example: power-converter connected via a transmission
line to a resistive load or an induction motor:

• The power-converter is a port-Hamiltonian system (with


switching Dirac structure).

• Transmission line is distributed-parameter port-Hamiltonian


system.

• Induction motor is a port-Hamiltonian system, with


Hamiltonian being the electro-mechanical energy.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 43

Power converter connected to the load via transmission line

L D Il
Vl

E S C VC = VL line R

Figure 6: The Boost converter with a transmission line


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 44

Boost power converter

+ L − D

+
+ +
E S C R
− −

Figure 7: Boost circuit with clamping diode

The circuit consists of a capacitor C with electric charge qC , an


inductor L with magnetic flux linkage φL , and a resistive load R,
together with an ideal diode and an ideal switch S , with switch
positions s = 1 (switch closed) and s = 0 (switch open).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 45

The voltage-current characteristics of the ideal diode and switch


are depicted in Figure 8.

iS
6iD 6

vS
-
−vD
-

Figure 8: Voltage-current characteristic of an ideal diode and ideal


switch

The ideal diode thus satisfies the complementarity conditions:

vD iD = 0, vD ≤ 0, iD ≥ 0.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 46

This yields the port-Hamiltonian model


1 2 1 2
(with H(qC , φL ) = 2C qC + 2L φL ):
        
qC
q̇C − R1 1−s ∂H
∂qC = C 0 siD
  =   + E +  
∂H φL
φ̇L s−1 0 ∂φL = L 1 (s − 1)vD

φL
I = L

Assume that the switch and the diode are coupled in the following
sense: if the switch is closed (s = 1) then the diode is open
(iD = 0), while if the switch is open (s = 0), then the diode is
closed (vD = 0). (This means that we disregard the so-called
discontinuous modes.)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 47

In this case we obtain the switching port-Hamiltonian system


      
qC
q̇C − R1 1−s ∂H
= 0
  =    ∂qC C 
+  E
∂H φ
φ̇L s−1 0 ∂φL = L
L
1

 
∂H qC
h i
∂qC = C φL
I = 0 1  φL
=
L
∂H
∂φL = L

A port-Hamiltonian system where the J -matrix depends on the


switch position.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 48

In general, a switching port-Hamiltonian system (without algebraic


equality and inequality constraints) is defined as

∂H
ẋ = F (ρ)z + g(ρ)u, z= ∂x (x)

y = g T (ρ)z

with ρ ∈ {0, 1}p , and

F (ρ) = J(ρ) − R(ρ), J(ρ) = −J T (ρ), R(ρ) = RT (ρ) ≥ 0

Note that the system is passive for every switching sequence:

d ∂T H ∂H
H=− (x)R(ρ) (x) + uT y ≤ uT y
dt ∂x ∂x
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 49

Oriented graphs and Kirchhoff behavior

Figure 9: Kirchhoff

An oriented graph G consists of a finite set N of nodes (vertices)


and a finite set B of branches (edges). To every branch b ∈ B
there corresponds an ordered pair (n, m) ∈ N 2 representing the
starting node n and the final node m of this branch.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 50

An oriented graph is specified by its incidence matrix D , which is


an n̄ × b̄ matrix, n̄ being the number of nodes and b̄ being the
number of branches, with (i, j)−th element dij equal to 1 if the j -th
branch is a branch towards node i, equal to −1 if the j -th branch is
a branch originating from node i, and equal to 0 otherwise.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 51

The node space Λ0 is the vector space of all functions from N to


R. Clearly, Λ0 can be identified with Rn̄ .
The branch space Λ1 is the vector space of all functions from B to
R. Again, Λ1 can be identified with Rb̄ .
For an electrical circuit Λ1 will be the vector space of currents
over the branches in the circuit.
The dual space of the vector space Λ1 will be denoted by Λ1 ; the
vector space of voltages over the branches.
The product < V |I > of a vector of currents I ∈ Λ1 with a vector of
voltages V ∈ Λ1 is the total power over the circuit.
Λ0 is the vector space of potentials over the nodes, with dual
space Λ0 .
(Note: since Λ0 and Λ1 have a canonical basis we can identify them
with their dual spaces Λ0 and Λ1 .)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 52

The incidence matrix D is the matrix representation of a map

∂ : Λ1 → Λ0

called the incidence operatora . The adjoint map of the incidence


operator ∂ is the linear map

d : Λ0 → Λ1

which is called the co-incidence operator. The matrix


representation of the map d is given by D T .
a In the literature this operator is usually called the boundary operator.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 53

Kirchhoff’s current laws (KCL) for a circuit are given as

∂I = 0

while Kirchhoff’s voltage laws (KVL) take the form

V ∈ im d

(The kernel of the incidence operator ∂ is the cycle space of the


graph, the image of the co-incidence operator d is its cut space.)
Kirchhoff’s voltage laws are thus

V = dφ

for some φ ∈ Λ0 (the vector of potentials at every node). In matrix


notation Kirchhoff’s laws are

DI = 0, V = DT φ
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 54

Tellegen’s theorem automatically follows from Kirchhoff’s laws.


Take any current distribution I satisfying KCL, and any voltage
distribution V satisfying KVL. Then
X X X
Vb Ib = (dφ)b Ib = φn (∂I)n = 0
b b n

since I satisfies KCL ∂I = 0.


In particular, Tellegen’s theorem implies that for any actual current
and voltage distribution over the circuit the total power is equal to
zero.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 55

The Kirchhoff behavior BK (G) of a graph G with incidence


operator ∂ is

BK (G) := {(I, V ) ∈ Λ1 × Λ1 | I ∈ ker ∂, V ∈ im d}

It immediately follows that the Kirchhoff behavior defines a Dirac


structure.
D ⊂ V × V ∗ is a Dirac structure if < v ∗ |v >= 0 for every (v, v ∗ ) ∈ D
while dim D = dim V .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 56

Open graphs
An open graph G is obtained from an ordinary graph with set of
nodes N by identifying a subset Ne ⊂ N of external nodes. The
remaining subset Ni := N − Ne are the internal nodes of the open
graph.
Kirchhoff’s current laws now take the form
 
Ie
∂I +  =0
0

with Ie ∈ Λe the vector of external currents entering the graph at


the external nodes (terminals).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 57

In this case we obtain


P P P
b Vb Ib = b (dφ)b Ib =
φn (∂I)n =
n
P P P (1)
ni φni (∂I)ni + ne φne (∂I)ne = − ne φne Ine

since I satisfies the Kirchhoff’s current laws (∂I)ni = 0 and


(∂I)ne = −Ine .
Thus, for open graphs the total power over the graph is not zero
P
but equal to − ne φne Ine (the incoming power at the external
nodes).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 58

Even though the vector of potentials φe at the external nodes is


not uniquely determined by the vector V of voltages, the
P
expression ne φne Ine is uniquely determined.

The freedom in the choice of φ corresponding to the same vector


of voltages V is given by all vectors ψ such that D T ψ = 0. Hence
the freedom in φe is given by all vectors ψe such that for some ψi it
holds that DeT ψe + DiT ψi = 0, where De is the submatrix of D
consisting of the first n̄e rows and Di is the submatrix consisting of
the last n̄ − n̄e rows. For any such ψe we have

ψeT Ie = −ψeT De I = ψiT Di I = 0


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 59

The image of D , or equivalently the kernel of D T , can be


characterized as follows. First
X
T
0 = 11 DI = Ine (2)
ne

Hence the external part of the Kirchhoff behavior of an open graph


is constrained by the obvious fact that all external currents sum up
to zero.
In general, the rank of D is equal to n̄ − kG , where kG denotes the
number of connected components.

  if Gconsists
For example,  of two connected components, then the
11 0
vectors   and   span the kernel of D T . This implies that both
0 11
the sum of the external currents belonging to the first component
as well as the sum of the external currents of the second
component are equal to zero.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 60

It turns out that an open graph G can be closed (in a unique


fashion) to an ordinary graph Ḡ while keeping the same number of
connected components.
Let G be connected. Then add one virtual node n̄, and virtual
edges from this virtual node n̄ to every external node ne ∈ Ne . The
Kirchhoff behavior of this graph Ḡ extends the Kirchhoff behavior
of the open graph G . In fact, to the virtual node n̄ we may
associate an arbitrary potential φn̄ (a ground-potential), and we
P
may rewrite the righthand-side of (1) as (since e Ine = 0)
X X
− (φne − φn̄ )Ine = − Vne Ine
ne ne

where Vne := φne − φn̄ and Ine denote the voltage, respectively
current, over the virtual edge towards the external node ne .
Thus the incoming power is rewritten as a product of external
currents and voltages.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 61

Constitutive relations for every branch.


Simplest case is a resistive relation between Ib and Vb such that
Vb Ib ≥ 0.
In the case of a capacitive relation one defines the charge Qb
together with the electric energy Hb (Qb ). The constitutive relations
are then given by
dHb
Q̇b = −Ib , Vb = (Qb )
dQb
Alternatively, in the case of an inductor one specifies the magnetic
energy Hb (ϕb ), where ϕb denotes the flux, together with the
relations
dHb
ϕ̇b = −Vb , Ib = (ϕb )
dϕb
Substituting these constitutive relations into the Kirchhoff behavior
(and corresponding Dirac structure) defined by the graph results in
a port-Hamiltonian system description.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 62

Interconnection of open graphs

Consider two open graphs G A and G B with respective sets of


external nodes NeA and NeB . Suppose we want to interconnect
these open graphs over a set of shared external nodes

N̄ ⊂ NeA ∩ NeB

This may be done by identifying the shared nodes corresponding to


the two graphs, leading to an interconnected open graph G A k G B
with resulting set of internal nodes

NiA ∪ NiB ∪ N̄

and external nodes

(NeA − N̄ ) ∪ (NeB − N̄ )
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 63

Extension to higher-order networks

An oriented graph G with incidence operator ∂ defines a


1-dimensional complex. Indeed, the sequence
∂ 11
Λ1 → Λ0 → R

satisfies 11 ◦ ∂ = 0.
The ’algebraic-topological invariants’ of this 1-complex (the
so-called Betti numbers) are nothing else than the number of
connected components.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 64

An arbitrary k -complex Λ is specified by a sequence of linear spaces


Λ0 , Λ1 , · · · , Λk , together with a sequence of incidence operators

∂k ∂k−1 ∂ ∂
Λk → Λk−1 → · · · Λ1 →1 Λ0 →0 0

with the property that

∂j−1 ◦ ∂j = 0, j = 1, · · · , k

The vector spaces Λj , j = 0, 1 · · · , k, are called the spaces of


j -chains.
Each Λj is generated by a finite set of j -cells in the sense that Λj is
the set of functions from the set of j -cells to R.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 65

A typical example of a 3-complex is the triangularization of a


3-dimensional manifold, with the j -cells, j = 0, 1, 2, 3, being the sets
of vertices, edges, faces, and tetrahedra, and ∂1 , ∂2 , ∂3 capturing
the incidence structure.
Denoting the dual linear spaces by Λj , j = 0, 1 · · · , k, we have the
following dual sequence
d d d
Λ0 →1 Λ1 →2 Λ2 · · · Λk−1 →
k
Λk

having the analogous property

dj ◦ dj−1 = 0, j = 2, · · · , k

The elements of Λj are called j -cochains.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 66

Consider any k -complex Λ, with k -chains α ∈ Λk and k -cochains


β ∈ Λk . We define, similarly as in the case of a graph (1-complex)
its Kirchhoff behavior as

BK (Λ) := {(α, β) ∈ Λk × Λk |
∂k α = 0, ∃φ ∈ Λk−1 s.t. β = dk φ}

As before, it is immediately seen that BK (Λ) ⊂ Λk × Λk is a Dirac


structure. In particular, it follows that

< β | α >k = 0

for every (α, β) ∈ BK (Λ), where < · | · >k denotes the duality product
between the dual linear spaces Λk and Λk .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 67

Open k -complexes

An open k -complex is obtained by identifying a subset N(k−1)e of


the set of all (k − 1)-cells, called the external (k − 1)-cells.
Define the linear space of functions from this subset of (k − 1)-cells
to R as Λe ⊂ Λk−1 (the space of ’external currents’).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 68

As before, Kirchhoff’s voltage laws remain unchanged

β = dk φ,

while Kirchhoff’s current laws are modified into


 
αe
∂k α +  =0
0

where αe is the vector of external currents associated to the


external (k − 1)-cells. We obtain

< β | α >k =< dk φ | α >k =< φ | ∂k α >k−1 =


 
−αe
<φ|   >k−1 = − < φe | αe >k−1
0

where φe denotes the vector of potentials at the external


(k − 1)-cells.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 69

Similar to graphs it follows that the Kirchhoff current laws for open
k -complexes Dke α = −αe imply certain constraints for the external
’currents’ αe . Indeed, by the fact that

∂k−1 ◦ ∂k = 0

it follows that
D(k−1)e αe = 0

Furthermore, we can uniquely extend the open k -complex to an


ordinary k -complex while keeping the same Betti numbers.
This means again that the external power can be rewritten as a
product of ’external currents’ and ’external voltages’.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 70

Hamiltonian dynamics on a k -complex

There are a number of canonical ways to define ’physical’ dynamics


on k -complexes.
First option: define the Dirac structure

fx = −dk f, fx ∈ Λk , f ∈ Λk−1

e = ∂k ex , ex ∈ Λk , e ∈ Λk−1
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 71

Associate to every k -cell an energy storage, leading to a total


energy storage H(x), where x ∈ Λk denotes the vector of energy
variables, with
∂H
ẋ = −fx , ex = (x)
∂x
Furthermore, associate to every (k − 1)-cell the resistive relation

f = −Re, R = RT ≥ 0

This yields the relaxation dynamics

∂H
ẋ = −dk e = dk R f = −dk R ∂k (x), x ∈ Λk
∂x
with the property that

dH ∂H T ∂H
= −(∂k (x)) R ∂k (x) ≤ 0
dt ∂x ∂x
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 72

For an open complex with external (k − 1)-cells and external


’currents’ Λ̄ ⊂ Λk−1 the definition is modified as follows.
Consider instead
   
f k  
f
fx = −dk  , fx ∈ Λ , ∈ Λk−1 , fb ∈ Λ̄k−1
fb f
   b (3)
e e
  = ∂k ex , ex ∈ Λk ,   ∈ Λk−1 , fb ∈ Λ̄k−1
eb eb

with fb , eb corresponding to the external (k − 1)-cells,


and f, e corresponding to the internal (k − 1)-cells.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 73

∂H
Imposing the same storage relations fx = −ẋ, ex = ∂x (x) and
resistive relations f = −Re we arrive at

ẋ = −drk R ∂kr ∂H b
∂x (x) + dk fb

eb = ∂kb ∂H
∂x (x)
 
h i ∂kr
where we have split dk as dk = drk dbk and ∂k =  .
∂kb
This defines a port-Hamiltonian system with inputs fb and
outputs eb .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 74

Heat transfer on a 2-complex


We will write the heat transfer as a conservation law in terms of
the conservation of internal energy:
The internal energy u of the 2-complex is a 2-cochain, u ∈ Λ2 (with
every component of u denoting the energy of the corresponding
2-cell).
The thermodynamic properties are defined by Gibbs’ relation, and
generated by the entropy function s = s(u) as thermodynamic
potential. Since we consider transformations with constant volume
and without mass transfer, Gibbs’ relation reduces to the definition
of the vector of intensive variables eu , conjugated to the extensive
variables u by
∂s
eu = (u)
∂u
The components of the vector eu ∈ Λ2 are equal to the reciprocal
of the temperature in each 2-cell.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 75

The heat conduction is given by the heat flux

f ∈ Λ1

describing the heat flux through every 1-cell (edge). This flux
arises from thermal non-equilibrium, defined by the fact that the
temperature is varying over the 2-cells.
Its conjugate vector of variables is the thermodynamical driving
force vector
e ∈ Λ1

given as the vector of the differences of the reciprocals of the


temperatures of the 2-cells with common edges (1-cells)

e = ∂2 eu
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 76

By Fourier’s law the heat flux due to thermal non-equilibrium is


expressed as
f = R(eu ) e,

with R(eu ) = RT (eu ) ≥ 0 depending on the heat conduction


coefficients. (Note the sign-difference !). Finally

du
= d2 f
dt
Hence the resulting system is a port-Hamiltonian system (of
relaxation type) defined on the 2-complex, with vector of state
variables x given by the internal energy vector u, and Hamiltonian
s(u).
By the different sign the entropy s(u) satisfies

ds ∂s ∂s
= (∂2 (u))T R(eu )∂2 (u) ≥ 0
dt ∂u ∂u
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 77

expressing the fact that the entropy s(u) is monotonously


increasing.
The exchange of heat through the boundary of the system can be
incorporated by splitting the edges (1-cells) into internal edges with
the resistive relation and external (boundary) edges. This would
lead to
ds ∂s T ∂s
= (∂2 (u)) R(eu ) ∂2 (u) + eb fb
dt ∂u ∂u
with fb , eb denoting the heat flux, respectively, thermodynamical
driving force, through the boundary 1-cells.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 78

Another type of Hamiltonian dynamics is obtained by considering


the relations

fx1 = dk ex2 , fx1 ∈ Λk , ex2 ∈ Λk−1


fx2 = −∂k ex1 , ex1 ∈ Λk , fx2 ∈ Λk−1

defining again a Dirac structure, together with two energy-storage


relations
1
ẋ1 = −fx1 , ex1 = ∂H∂x1 (x1
)
2 ∂H 2 2
ẋ = −fx2 , ex2 = ∂x2 (x )
leading to the oscillatory dynamics
∂H
ẋ1 = −dk ∂x 1 2
2 (x , x )

∂H
ẋ2 = ∂k ∂x 1 2
1 (x , x )

Example 1: (k = 2) Discretized Maxwell’s equations on a


2-dimensional domain.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 79

Example 2: Formation controla

Consider n point masses

ṗi = ui , i = 1, · · · , n
pi
with pi denoting their momenta, and vi = m i
their velocities.
Suppose v̄ is a desired joint velocity vector, and moreover, we want
their position vectors qi converge to a certain desired formation,
e.g. (for n = 3)

|q1 − q2 | = |q2 − q3 | = |q3 − q1 | = 1

This determines a graph structure with incidence matrix D .


a M. Arcak, IEEE TAC, 52, August 2007
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 80

Apply feedback to obtain

q̇i = ξi + v̄
ξ˙i = fi

and use
(D T ⊗ I3 )(q̇) = (D T ⊗ I3 )(ξ)

as an input to a Hamiltonian integrator dynamics defined on the


edges, leading to an input

f := −(D ⊗ I3 )(ψ)

for the dynamics at the nodes.


By designing carefully the Hamiltonian corresponding to the
dynamics on the edges, and by adding damping, one obtains the
desired formation control.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 81

Model reduction of port-Hamiltonian systems


• Network modeling of complex lumped-parameter systems
(circuits, multi-body systems) often leads to high-dimensional
models.

• Structure-preserving spatial discretization of


distributed-parameter port-Hamiltonian systems yields
high-dimensional port-Hamiltonian models.

• Lumped-parameter modeling of systems like MEMS gives


high-dimensional port-Hamiltonian systems.

• Controller systems may be in first instance


distributed-parameter, and need to be discretized to low-order
controllers.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 82

In many cases we want the reduced-order system to be again


port-Hamiltonian:

• Port-Hamiltonian model reduction preserves passivity.

• Port-Hamiltonian model reduction may (approximately)


preserve other balance laws /conservation laws.

• Physical interpretation of reduced-order model.

• Reduced-order system can replace the high-order


port-Hamiltonian system in a larger context.

Thus there is a need for structure-preserving model reduction of


high-dimensional port-Hamiltonian systems.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 83

Controllability analysis

Consider a linear port-Hamiltonian system, written as

ẋ = F Qx + Bu, F := J − R, J = −J T , R = RT ≥ 0
y = B T Qx, Q = QT ≥ 0

Take linear coordinates x = (x1 , x2 ) such that the upper part of


         
ẋ1 F11 F12 Q11 Q12 x1 B1
  =       +  u
ẋ2 F21 F22 Q21 Q22 x2 B2

   
h i Q11 Q12 x1
y = B1T B2T    
Q21 Q22 x2

is the reachability subspace R.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 84

By invariance of R this implies

F21 Q11 + F22 Q21 = 0


B2 = 0

If follows that the dynamics restricted to R is given as

ẋ1 = (F11 Q11 + F12 Q21 )x1 + B1 u


y = B1T Q11 x1

−1
Now solve for Q21 as Q21 = −F22 F21 Q11 . This yields

−1
ẋ1 = (F11 − F12 F22 F21 )Q11 x1 + B1 u
y = B1T Q11 x1

which is again a port-Hamiltonian system.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 85

Observability analysis

Suppose the system is not observable. Then there exist


coordinates x = (x1 , x2 ) such that the lower part is the
unobservability subspace N . By invariance of N it follows that

F11 Q12 + F12 Q22 = 0


B1T Q12 + B2T Q22 = 0

Then the dynamics on the quotient space X /N is

ẋ1 = (F11 Q11 + F12 Q21 )x1 + B1 u


y = B1T Q11 x1 + B2T Q21 x1
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 86

It follow from that F12 = −F11 Q12 Q−1


22 and B T
2 = −B T
1 Q12 Q−1
22 .
Substitution yields

ẋ1 = F11 (Q11 − Q12 Q−1


22 Q21 )x1 + B1 u

y = B1T (Q11 − Q12 Q−1


22 Q21 )x1

which is again a port-Hamiltonian system with Hamiltonian


H̄ = 21 xT1 (Q11 − Q12 Q−1
22 Q21 )x1 .

Remark Note that the Schur complement (Q11 − Q12 Q−1


22 Q21 ) ≥ 0 if
Q ≥ 0.
This suggests two canonical ways for structure-preserving
model reduction.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 87

Conclusions of Part I
• Port-Hamiltonian systems provide a unified framework for
modeling, analysis, and simulation of complex
lumped-parameter multi-physics systems.

• Inclusion of distributed-parameter components.

• Direct lumping of distributed-parameter systems to


finite-dimensional port-Hamiltonian systems.

• Structure-preserving model reduction.

◦ Extensions to thermodynamic systems and chemical reaction


networks.

◦ Further exploration of the network (graph) information.

See www.math.rug.nl/˜arjan, for further info.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 88

The port-Hamiltonian approach to


physical system modeling and control
Part II: Control of Port-Hamiltonian systems
Contents

• Use of passivity for control

• Control by interconnection: set-point stabilization

• The dissipation obstacle

• A state feedback perspective; shaping the Hamiltonian

• New control paradigms

• Model reduction of port-Hamiltonian systems


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 89

Use of passivity for control and beyond


• The storage function can be used as Lyapunov function,
implying some sort of stability for the uncontrolled system.
• The standard feedback interconnection of two passive systems
is again passive, with storage function being the sum of the
individual storage functions.
• Passive systems can be asymptotically stabilized by adding
artificial damping. In fact,
d
H ≤ uT y
dt
together with the additional damping u = −y yields
d
H ≤ − k y k2
dt
proving asymptotic stability provided an observability
condition is met.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 90

Example The Euler equations for the motion of a rigid body


revolving about its center of gravity with one input are

I1 ω̇1 = [I2 − I3 ]ω2 ω3 + g1 u


I2 ω̇2 = [I3 − I1 ]ω1 ω3 + g2 u
I3 ω̇3 = [I1 − I2 ]ω1 ω2 + g3 u,
T
Here ω := (ω1 , ω2 , ω3 ) are the angular velocities around the principal
axes of the rigid body, and I1 , I2 , I3 > 0 are the principal moments
of inertia. The system for u = 0 has the origin as an equilibrium
point. Linearization yields the linear system
   
−1
0 0 0 I1 g1
   
A= 0 0 0  −1
B =  I2 g2  .
  

0 0 0 I3−1 g3

Hence the linearization does not say anything about stabilizability.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 91

Stability and asymptotic stabilization by damping injection


Rewrite the system in port-Hamiltonian form by defining the
angular momenta

p1 = I1 ω̇1 , p2 = I2 ω̇2 , p3 = I3 ω̇3

and defining the Hamiltonian H(p) as the total kinetic energy

1 p21 p22 p23


H(p) = ( + + )
2 I1 I2 I3

Then the system can be rewritten as


        
∂H ∂H
ṗ1 0 −p3 p2 g1
     ∂p1    h i  ∂p1 
ṗ2  =  p3  ∂H  + ∂H 
0 −p1  g2  u, y = g1 g2 g3 

     ∂p2   ∂p2 
∂H ∂H
ṗ3 −p2 p1 0 ∂p3 g3 ∂p3
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 92

Since Ḣ = 0 and H has a minimum at p = 0 the origin is stable.


Damping injection amounts to the negative output feedback
p1 p2 p3
u = −y = −g1 − g2 − g3 = −g1 ω1 − g2 ω2 − g3 ω3 ,
I1 I2 I3

yielding convergence to the largest invariant set contained in


p1 p2 p3
S := {p ∈ R3 | Ḣ(p) = 0} = {p ∈ R3 | g1 + g2 + g3 = 0}
I1 I2 I3

It can be shown that the largest invariant set contained in S is the


origin p = 0 if and only if

g1 6= 0, g2 6= 0, g3 6= 0,

in which case the origin is rendered asymptotically stable (even,


globally).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 93

Beyond control via passivity: What can we do if the


desired set-point is not a minimum of the storage function ??

Recall the proof of stability of an equilibrium (ω1∗ , 0, 0) 6= (0, 0, 0) of


the Euler equations.
The total energy H = 2Ip 2
1
+ 2I2
p 2 + 2I3
p 2 = 1
2 I 1 ω1
2
+ 1
2 I 2 ω2
2
+ 1 2
2 I3 ω3 has a
1 2 3
minimum at (0, 0, 0). Stability of (ω1∗ , 0, 0) is shown by taking as
Lyapunov function a combination of the total energy K and
another conserved quantity, namely the total angular momentum

C = p21 + p22 + p23 = I12 ω12 + I22 ω22 + I32 ω32

This follows from


 
h i 0 −p3 p2

p1 p2 p3 
 p3 0 =0
−p1 
−p2 p1 0
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 94

In general, for any Hamiltonian dynamics

∂H
ẋ = J(x) (x)
∂x
one may search for conserved quantities C , called Casimirs, as
being solutions of
∂T C
(x)J(x) = 0
∂x
d
Then dt C = 0 for every H , and thus also H + C is a candidate
Lyapunov function.
Note that the minimum of H + C may now be different from the
minimum of H .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 95

Control by interconnection: set-point stabilization:


Consider first a lossless Hamiltonian plant system P

ẋ = J(x) ∂H
∂x (x) + g(x)u

y = g T (x) ∂H
∂x (x)

where the desired set-point x∗ is not a minimum of the


Hamiltonian H , while the Hamiltonian dynamics ẋ = J(x) ∂H
∂x (x) does
not possess useful Casimirs.
How to (asymptotically) stabilize x∗ ?
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 96

Control by interconnection:
Consider a controller port-Hamiltonian system

ξ˙ = Jc (ξ) ∂H
∂ξ (ξ) + gc (ξ)uc ,
c
ξ ∈ Xc
C:
yc = g T (ξ) ∂H
∂ξ (ξ)
c

via the standard feedback interconnection

u = −yc , uc = y

u y
P

C
yc uc
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 97

Then the closed-loop system is the port-Hamiltonian system


    
∂H
ẋ J(x) −g(x)gcT (ξ) (x)
 =   ∂x 
ξ˙ gc (ξ)g T (x) Jc (ξ) ∂Hc
∂ξ (ξ)

with state space X × X c , and total Hamiltonian H(x) + Hc (ξ).

Main idea: design the controller system in such a manner


that the closed-loop system has useful Casimirs C(x, ξ) !
This may lead to a suitable candidate Lyapunov function

V (x, ξ) := H(x) + Hc (ξ) + C(x, ξ)

with Hc to-be-determined.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 98

Thus we look for functions C(x, ξ) satisfying


 
h
T
∂ C T
∂ C
i J(x) −g(x)gcT (ξ)
=0
∂x (x, ξ) ∂ξ (x, ξ)

gc (ξ)g T (x) Jc (ξ)

such that the candidate Lya[unov function

V (x, ξ) := H(x) + Hc (ξ) + C(x, ξ)

has a minimum at (x∗ , ξ ∗ ) for some (or a set of) ξ ∗ ⇒ stability.

Remark: The set of such achievable closed-loop Casimirs C(x, ξ)


can be fully characterized.

Subsequently, one may add extra damping (directly or in the


dynamics of the controller) to achieve asymptotic stability.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 99

Example: the ubiquitous pendulum

Consider the mathematical pendulum with Hamiltonian

1 2
H(q, p) = p + (1 − cos q)
2
actuated by a torque u, with output y = p (angular velocity).
Suppose we wish to stabilize the pendulum at a non-zero angle q ∗
and p∗ = 0.
Apply the nonlinear integral control

ξ˙ = uc = y
∂Hc
−u = yc = ∂ξ (ξ)

which is a port-Hamiltonian controller system with Jc = 0.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 100

Casimirs C(q, p, ξ) are found by solving


 
h i 0 1 0 
∂C ∂C ∂C  =0
∂q ∂p ∂ξ  −1 0 −1
0 1 0

leading to Casimirs C(q, p, ξ) = K(q − ξ), and candidate Lyapunov


functions
1 2
V (q, p, ξ) = p + (1 − cos q) + Hc (ξ) + K(q − ξ)
2
with the functions Hc and K to be determined.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 101

For a local minimum, determine K and Hc such that


Equilibrium assignment

∂K ∗
sin q ∗ + ∂z (q − ξ∗) = 0
∂Hc ∗
− ∂K ∗
∂z (q − ξ ) +

∂ξ (ξ ) =0

Minimum condition
 
∂2K ∗ ∂2K ∗
cos q + ∂z 2 (q − ξ ∗ )

0 − ∂z 2 (q − ξ∗)
 
 0 1 0 >0
 
∂2K ∗ ∂2K ∗ ∂ 2 Hc ∗
− ∂z 2 (q − ξ ∗ ) 0 ∂z 2 (q −ξ )+∗
∂ξ 2 (ξ )

Many possible solutions.


The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 102

Example: stabilization of the shallow water equations

The dynamics of the water in a canal can be described by


     
h v h h
∂t   +   ∂z  = 0

v g v v

with h(z, t) the height of the water at position z , and v(z, t) its
velocity (and g the gravitational constant).
Recall that by recognizing the total energy
b b
1 2
Z Z
H(h, v) = Hdz = [hv + gh2 ]dz
a a 2

this can be written (similarly to the telegrapher’s equations) as the


port-Hamiltonian system
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 103

∂h ∂ ∂H
∂t (z, t) = − ∂z ∂v (h, v)

∂v ∂ ∂H
∂t (z, t) = − ∂z ∂h (h, v)

with the 4 boundary variables

hv|a,b
−( 21 v 2 + gh)|a,b

denoting respectively the mass flow and the Bernoulli function


at the boundary points a, b.
(Note that the product hv · ( 21 v 2 + gh) equals power.)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 104

Suppose we want to control the water level h to a desired


height h∗ .
An obvious ’physical’ controller is to add to one side of the canal,
say the right-end b, an infinite water reservoir of height h∗ ,
corresponding to the port-Hamiltonian ’source’ system

ξ˙ = uc
∂Hc
yc = ∂ξ ( = gh∗ )

with Hamiltonian Hc (ξ) = gh∗ ξ , by the feedback interconnection


1 2
uc = y = h(b)v(b), yc = −u = v (b) + gh(b)
2
This yields a closed-loop port-Hamiltonian system with total
Hamiltonian Z l
1 2
[hv + gh2 ]dz + gh∗ ξ
0 2
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 105

By mass balance,
Z b
h(z, t)dz + ξ + c
a

is a Casimir for the closed-loop system. Thus we may take as


Lyapunov function

1
Rb Rb
V (h, v, ξ) := 2 a
2 2
[hv + gh ]dz + gh ξ − gh [ ∗ ∗
a
h(z, t)dz + ξ] + 12 g(b − a)h∗2

1
Rb 2 ∗ 2
= 2 a
[hv + g(h − h ) ]dz

which has a minimum at the desired set-point (h∗ , v ∗ = 0, ξ ∗ )


(with ξ ∗ arbitrary).
Remark Note that the source port-Hamiltonian system is not
passive, since the Hamiltonian Hc (ξ) = gh∗ ξ is not bounded from
below.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 106

An alternative, passive, choice of the Hamiltonian controller system


is to take e.g.
1
Hc (ξ) = gh∗ ξ 2
2
leading to the Lyapunov function
1 b 2 1 ∗
Z
V (h, v, ξ) = ∗ 2
[hv + g(h − h ) ]dz + gh (ξ − 1)2
2 a 2
Asymptotic stability of the equilibrium (h∗ , v ∗ = 0, ξ ∗ = 1) can be
obtained by adding ’damping’, that is, replacing uc = y = h(b)v(b) by
∂V
uc := y − (ξ) = h(b)v(b) − gh∗ (ξ − 1)
∂ξ
leading to (if there is no power flow through the left-end a)
d
V = −gh∗ (ξ − 1)2
dt
(See also the work of Bastin & co-workers for related and more
refined results.)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 107

The dissipation obstacle


Surprisingly, the presence of dissipation R 6= 0 may pose a problem !
C(x) is a Casimir for the Hamiltonian dynamics with dissipation
∂H
ẋ = [J(x) − R(x)] (x), J = J T , R = RT ≥ 0
∂x
iff
∂T C ∂T C ∂C ∂ T C ∂C ∂T C
[J − R] = 0 ⇒ [J − R] =0⇒ R =0⇒ R=0
∂x ∂x ∂x ∂x ∂x ∂x
and thus C is a Casimir iff
∂T C ∂T C
(x)J(x) = 0, (x)R(x) = 0
∂x ∂x
The physical reason for the dissipation obstacle is that by using a
passive controller only equilibria where no energy-dissipation takes
place may be stabilized.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 108

Similarly, if C(x, ξ) is a Casimir for the closed-loop port-Hamiltonian


system then it must satisfy
 
h T T
i R(x) 0
∂ C ∂ C =0
∂x (x, ξ) ∂ξ (x, ξ)

0 Rc (ξ)

implying by semi-positivity of R(x) and Rc (x)


∂T C
∂x (x, ξ)R(x) = 0
∂T C
∂ξ (x, ξ)Rc (ξ) = 0

This is the dissipation obstacle, which implies that one cannot


shape the Lyapunov function in the coordinates that are directly
affected by energy dissipation.
Remark: For shaping the potential energy in mechanical systems
this is not a problem since dissipation enters in the differential
equations for the momenta.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 109

To overcome the dissipation obstacle

Suppose one can find a mapping C : X → Rm , with its (transposed)


Jacobian matrix K T (x) := ∂C
∂x (x) satisfying

[J(x) − R(x)]K(x) + g(x) = 0

Construct now the interconnection and dissipation matrix of an


augmented system as
   
J JK R RK
Jaug :=   , Raug :=  
K T J K T JK K T R K T RK

By construction

[K T (x) | −I]Jaug = [K T (x) | −I]Raug = 0

implying that the components of C are Casimirs for the


Hamiltonian dynamics
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 110

   
∂H
ẋ (x)
  = [Jaug − Raug ]  ∂x 
∂Hc
ξ̇ ∂ξ (ξ)

Furthermore, since [J(x) − R(x)]K(x) + g(x) = 0


 
J −R [J − R]K
Jaug − Raug =  
K T [J − R] K T JK − K T RK

 
J −R −g
=  
[g − 2RK]T K T JK − K T RK

Thus the augmented system is a closed-loop system for a different


output !
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 111

Port-Hamiltonian systems with feedthrough term take the form

ẋ = [J(x) − R(x)] ∂H
∂x (x) + g(x)u

y = (g(x) + 2P (x))T ∂H
∂x (x) + [M (x) + S(x)]u,

with M skew-symmetric and S symmetric, while


 
R(x) P (x)
 ≥0
P T (x) S(x)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 112

The augmented system is thus the feedback interconnection of the


nonlinear integral controller

ξ˙ = uc
∂Hc
yc = ∂ξ (ξ)

with the plant port-Hamiltonian system with modified output with


feedthrough term

ẋ = [J(x) − R(x)] ∂H
∂x (x) + g(x)u

ymod = [g(x) − 2R(x)K(x)]T ∂H


∂x (x) + [−K T
(x)J(x)K(x) + K T
(x)R(x)K(x)]u

Remark: See Jeltsema, Ortega and Scherpen for further


explorations.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 113

Generalization to feedback interconnection with


state-modulation.
Recall that K T (x) := ∂C
∂x (x) is a solution to
[J(x) − R(x)]K(x) + g(x) = 0. This can be generalized to

[J(x) − R(x)]K(x) + g(x)β(x) = 0

with β(x) an m × m design matrix.


The same scheme as above works if we extend the standard
feedback interconnection u = −yc , uc = y to the state-modulated
feedback
u = −β(x)yc , uc = β T (x)y
Note that K(x) is a solution for some β(x) iff

g ⊥ (x)[J(x) − R(x)]K(x) = 0

(In fact, β(x) := −(g T (x)g(x))−1 g T (x)[J(x) − R(x)]K(x) does the job.)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 114

A state feedback perspective: shaping the


Hamiltonian
Restrict (without much loss of generality) to Casimirs of the form

C(x, ξ) = ξj − Gj (x)
It follows that for all time instants

ξj = Gj (x) + cj , cj ∈ R
Suppose that in this way all control state components ξi can be
expressed as function
ξ = G(x)
of the plant state x. Then the dynamic feedback reduces to a
state feedback, and the Lyapunov function H(x) + Hc (ξ) + C(x, ξ)
reduces to the shaped Hamiltonian

H(x) + Hc (G(x))
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 115

A direct state feedback perspective:


Interconnection-Damping Assignment (IDA)-PBC control

A direct way to generate candidate Lyapunov functions Hd is to


look for state feedbacks u = ûIDA (x) such that

∂H ∂Hd
[J(x) − R(x)] (x) + g(x)ûIDA (x) = [Jd (x) − Rd (x)] (x)
∂x ∂x
where Jd and Rd are newly assigned interconnection and damping
structures.
Remark: For mechanical systems IDA-PBC control is equivalent to
the theory of Controlled Lagrangians (Bloch, Leonard, Marsden, .).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 116

For Jd = J and Rd = R (Basic IDA-PBC) this reduces to

∂(Hd − H)
[J(x) − R(x)] (x) = g(x)ûBIDA (x)
∂x
and thus in this case, there exists an ûBIDA (x) if and only if

∂(Hd − H)
g ⊥ (x)[J(x) − R(x)] (x) = 0
∂x
which is the same equation as obtained for stabilization by Casimir
generation with a state-modulated nonlinear integral controller !
Conclusion: Basic IDA-PBC ⇔ State-modulated Control by
Interconnection.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 117

Shifted passivity w.r.t. a controlled equilibrium


(see Jayawardhana, Ortega). Consider a port-Hamiltonian system
∂H
ẋ = F z + gu, z= ∂x (x)
y = gT z
where F = J − R, g are constant, and a controlled equilibrium x0 :
∂H
F z0 + gu0 = 0, z0 = (x0 )
∂x
Define the shifted storage function
∂Hp T
V (x) := Hp (x) − (x − x0 ) (x0 ) − Hp (x0 )
∂x
∂V
Note that ∂x = z − z0 . It follows that
d
dt V = (z − z0 )T ẋ = (z − z0 )T (F z + gu) =
(z − z0 )T F (z − z0 ) + (z − z0 )T g(u − u0 ) + (z − z0 )T (F z0 + gu0 ) ≤ (y − y0 )T (u − u0 )
implying passivity w.r.t. the shifted inputs u − u0 and outputs y − y0 .
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 118

Application to switching control


Consider the port-Hamiltonian model of a power-converter
∂Hp
ẋ = F (ρ)z + g(ρ)E + gl u, z= (x), F (ρ) := J(ρ) − R(ρ)
∂x
with vector of Boolean variables ρ ∈ {0, 1}k , Hp (x) the total stored
electromagnetic energy, and output vector y = glT z .
Let x0 be an equilibrium of the averaged model, that is
∂H
F (ρ0 )z0 + g(ρ0 )E + gl u0 = 0, z0 = (x0 )
∂x
for some ρ0 ∈ [0, 1]k and u0 . Then

ẋ = F (ρ)(z − z0 ) + F (ρ)z0 + g(ρ)E + gl u


= F (ρ)(z − z0 ) + [F (ρ) − F (ρ0 )]z0 + [g(ρ) − g(ρ0 )]E + gl (u − u0 )
+F (ρ0 )z0 + g(ρ0 )E + gl u0
= F (ρ)(z − z0 ) + [F (ρ) − F (ρ0 )]z0 + [g(ρ) − g(ρ0 )]E + gl (u − u0 )
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 119

For many power converters we know that


Pp
F (ρ) − F (ρ0 ) = i=1 Fi (ρi − ρ0i )
Pp
g(ρ) − g(ρ0 ) = i=1 gi (ρi − ρ0i )

and thus
p
X
ẋ = F (ρ)(z − z0 ) + [Fi z0 + gi E](ρi − ρ0i ) + gl (u − u0 )
i=1

Take as Lyapunov/storage function


∂Hp T
V (x) := Hp (x) − (x − x0 ) (x0 ) − Hp (x0 )
∂x
Then
d ∂Hp ∂Hp T T
dt V (x) = [ ∂x (x) − ∂x (x0 )] ẋ = (z − z0 ) ẋ =
Pp
(z − z0 )T F (ρ)(z − z0 ) + i=1 (z − z0 )T [Fi z0 + gi E](ρi − ρ0i ) + (z − z0 )T gl (u − u0 )

with (z − z0 )T F (ρ)(z − z0 ) ≤ 0.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 120

Thus at any time we can choose ρi ∈ {0, 1} such that


d
V (x) ≤ (z − z0 )T gl (u − u0 )
dt
implying passivity of the switched system with respect to the input
vector u − u0 and output vector y − y0 = glT (z − z0 ). As a
consequence, if the converter is terminated on a static resistive
load then the switched converter is (asymptotically) stable around
x0 . Thus the voltage over the resistive load can be stabilized
around any set-point.
This can be immediately generalized to converters connected to a
load via a transmission line (see Part I).
Note that for linear capacitors and inductors we have
1 T 1
Hp (x) = x Qx, V (x) = (x − x0 )T Q(x − x0 )
2 2
(cf. Buisson & co-workers)
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 121

New control paradigms


Example: Energy transfer control

Consider two port-Hamiltonian systems Σi

ẋi = Ji (xi ) ∂H
∂xi (xi ) + gi (xi )ui
i

yi = giT (xi ) ∂H
∂xi (xi ),
i
i = 1, 2

Suppose we want to transfer the energy from the port-Hamiltonian


system Σ1 to the port-Hamiltonian system Σ2 , while keeping the
total energy H1 + H2 constant.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 122

This can be done by using the output feedback


    
u1 0 −y1 y2T y
 =   1
u2 y2 y1T 0 y2

It follows that the closed-loop system is energy-preserving.


However, for the individual energies
d
H1 = −y1T y1 y2T y2 = −||y1 ||2 ||y2 ||2 ≤ 0
dt
implying that H1 is decreasing as long as ||y1 || and ||y2 || are
different from 0. On the other hand,
d
H2 = y2T y2 y1T y1 = ||y2 ||2 ||y1 ||2 ≥ 0
dt
implying that H2 is increasing at the same rate. Has been
successfully applied to energy-efficient path-following control of
mechanical systems (cf. Duindam & Stramigioli).
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 123

Impedance control

Consider a system with two (not necessarily distinct) ports

ẋ = [J(x) − R(x)] ∂H
∂x (x) + g(x)u + k(x)f, x ∈ X , u ∈ Rm
y = g T (x) ∂H
∂x (x) u, y ∈ Rm (4)
e = k T (x) ∂H
∂x (x) f, e ∈ Rm
The relation between the f and e variables is called the
’impedance’ of the (f, e)-port. In Impedance Control (Hogan) one
tries to shape this impedance by using the control port
corresponding to u, y .
Typical application: the (f, e)-port corresponds to the end-point of
a robotic manipulator, while the (u, y)-port corresponds to
actuation.
Basic question: what are achievable impedances of the
(f, e)-port ?
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 124

Conclusions of Part II
• Beyond passivity by port-Hamiltonian systems theory.

• Control by interconnection and Casimir generation, IDA-PBC


control.

• Allows for ’physical’ interpretation of control strategies.


Suggests new control paradigms for nonlinear systems.

◦ Use of passivity generally yields good robustness, but


performance theory is yet lacking.
The port-Hamiltonian approach to physical system modeling and control, Namur, November, 2008 125

THANK YOU !

Вам также может понравиться