12.07.2015 Views

Port-Hamiltonian systems: network modeling and control of ...

Port-Hamiltonian systems: network modeling and control of ...

Port-Hamiltonian systems: network modeling and control of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

then it follows that (4) with inputs u = τ <strong>and</strong> outputs y = ˙q is a passive (in fact, lossless)state space system with storage function H(q, p) − C ≥ 0 (see e.g. [61, 20, 47] for the generaltheory <strong>of</strong> passive <strong>and</strong> dissipative <strong>systems</strong>). Since the energy is only defined up to a constant,we may as well as take as potential energy the function P (q) − C ≥ 0, in which case the totalenergy H(q, p) becomes nonnegative <strong>and</strong> thus itself is the storage function.System (4) is an example <strong>of</strong> a <strong>Hamiltonian</strong> system with collocated inputs <strong>and</strong> outputs,which more generally is given in the following form˙q = ∂H∂p (q, p) , (q, p) = (q 1, . . . , q k , p 1 , . . . , p k )ṗ = − ∂H∂q (q, p) + B(q)u, u ∈ Rm , (8)y = B T (q) ∂H∂p (q, p) (= BT (q) ˙q), y ∈ R m ,Here B(q) is the input force matrix, with B(q)u denoting the generalized forces resultingfrom the <strong>control</strong> inputs u ∈ R m . The state space <strong>of</strong> (8) with local coordinates (q, p) is usuallycalled the phase space. In case m < k we speak <strong>of</strong> an underactuated system. If m = k <strong>and</strong> thematrix B(q) is everywhere invertible, then the <strong>Hamiltonian</strong> system is called fully actuated.Because <strong>of</strong> the form <strong>of</strong> the output equations y = B T (q) ˙q we again obtain the energybalancedHdt (q(t), p(t)) = uT (t)y(t) (9)Hence if H is non-negative (or, bounded from below), any <strong>Hamiltonian</strong> system (8) is a losslessstate space system. For a system-theoretic treatment <strong>of</strong> the <strong>Hamiltonian</strong> <strong>systems</strong> (8),especially if the output y can be written as the time-derivative <strong>of</strong> a vector <strong>of</strong> generalizedconfiguration coordinates, we refer to e.g. [8, 43, 44, 10, 36].A major generalization <strong>of</strong> the class <strong>of</strong> <strong>Hamiltonian</strong> <strong>systems</strong> (8) is to consider <strong>systems</strong>which are described in local coordinates asẋ = J(x) ∂H∂x(x) + g(x)u, x∈ X , u ∈ Rmy = g T (x) ∂H∂x (x), y ∈ Rm (10)Here J(x) is an n × n matrix with entries depending smoothly on x, which is assumed to beskew-symmetricJ(x) = −J T (x), (11)<strong>and</strong> x = (x 1 , . . . , x n ) are local coordinates for an n-dimensional state space manifold X .Because <strong>of</strong> (11) we easily recover the energy-balance dH dt (x(t)) = uT (t)y(t), showing that (10)is lossless if H ≥ 0. We call (10) with J satisfying (11) a port-<strong>Hamiltonian</strong> system withstructure matrix J(x) <strong>and</strong> <strong>Hamiltonian</strong> H ([24, 30, 25]). Note that (8) (<strong>and</strong> hence (4)) is aparticular case [ <strong>of</strong> (10) ] with x = (q, p), [ <strong>and</strong>]J(x) being given by the constant skew-symmetricmatrix J = 0 Ik−I k 0, <strong>and</strong> g(q, p) = 0B(q) .As an important mathematical note, we remark that in many examples the structurematrix J will satisfy the “integrability” conditionsn∑[J lj (x) ∂J ik(x) + J li (x) ∂J kj(x) + J lk (x) ∂J ]ji(x) = 0, i, j, k = 1, . . . , n (12)∂x l ∂x l ∂x ll=14


In this case we may find, by Darboux’s theorem (see e.g. [60]) around any pointx 0 where the rank <strong>of</strong> the matrix J(x) is constant, local coordinates ˜x = (q, p, s) =(q 1 , . . . , q k , p 1 , . . . , p k , s 1 , . . . s l ), with 2k the rank <strong>of</strong> J <strong>and</strong> n = 2k + l, such that J in thesecoordinates takes the form⎡J = ⎣0 I k 0−I k 0 00 0 0⎤⎦ (13)The coordinates (q, p, s) are called canonical coordinates, <strong>and</strong> J satisfying (11) <strong>and</strong> (12) iscalled a Poisson structure matrix. In such canonical coordinates the equations (10) take theform˙q = ∂H∂p (q, p, s) + g q(q, p, s)uṗ = − ∂H∂q (q, p, s) + g p(q, p, s)uṡ = g s (q, p, s)u (14)y = g T q(q, p, s)∂H∂q (q, p, s) + gT p (q, p, s)∂H∂p (q, p, s) + gT s(q, p, s)∂H (q, p, s)∂swhich is, apart from the appearance <strong>of</strong> the variables s, very close to the st<strong>and</strong>ard <strong>Hamiltonian</strong>form (8). In particular, if g s = 0, then the variables s are merely an additional set <strong>of</strong> constantparameters.Although traditionally <strong>Hamiltonian</strong> <strong>systems</strong> arise from the Euler-Lagrange equations <strong>of</strong>motion (which are usually derived from variational principles) the point <strong>of</strong> departure for thetheory <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> is different. Indeed, port-<strong>Hamiltonian</strong> <strong>systems</strong> arise systematicallyfrom <strong>network</strong> models <strong>of</strong> physical <strong>systems</strong>. In <strong>network</strong> models <strong>of</strong> complex physical<strong>systems</strong> the overall system is seen as the interconnection <strong>of</strong> energy-storing elements via basicinterconnection (balance) laws as Newton’s third law or Kirchh<strong>of</strong>f’s laws, as well as powerconservingelements like transformers, kinematic pairs <strong>and</strong> ideal constraints, together withenergy-dissipating elements. The basic point <strong>of</strong> departure for the theory <strong>of</strong> port-<strong>Hamiltonian</strong><strong>systems</strong> is to formalize the basic interconnection laws together with the power-conservingelements by a geometric structure, <strong>and</strong> to define the <strong>Hamiltonian</strong> as the total energy storedin the system. Indeed, for the (restricted) form <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> given above thestructure matrix J(x) <strong>and</strong> the input matrix g(x) may be directly associated with the <strong>network</strong>interconnection structure, while the <strong>Hamiltonian</strong> H is just the sum <strong>of</strong> the energies <strong>of</strong> all theenergy-storing elements; see our papers [30, 24, 32, 31, 50, 52, 27, 46, 58]. In particular,<strong>network</strong> models <strong>of</strong> complex physical <strong>systems</strong> formalized within the (generalized) bond graphlanguage ([41, 7]) can be shown to immediately lead to port-<strong>Hamiltonian</strong> <strong>systems</strong>; see e.g.[19].Example 2.1 (LCTG circuits). Consider a <strong>control</strong>led LC-circuit (see Figure 1) consisting<strong>of</strong> two inductors with magnetic energies H 1 (ϕ 1 ), H 2 (ϕ 2 ) (ϕ 1 <strong>and</strong> ϕ 2 being the magnetic fluxlinkages), <strong>and</strong> a capacitor with electric energy H 3 (Q) (Q being the charge). If the elementsare linear then H 1 (ϕ 1 ) = 12L 1ϕ 2 1 , H 2(ϕ 2 ) = 12L 2ϕ 2 2 <strong>and</strong> H 3(Q) = 12C Q2 . Furthermore let V = udenote a voltage source. Using Kirchh<strong>of</strong>f’s laws one immediately arrives at the dynamical5


L 1 L 2Cϕ 1Qϕ 2VFigure 1: Controlled LC-circuitequations⎡ ⎤˙Q⎣ ˙ϕ 1⎦ =˙ϕ 2y = ∂H∂ϕ 1⎡⎡ ⎤0 1 −1⎣−1 0 0 ⎦ ⎢⎣1 0 0} {{ }J⎤∂H∂Q∂H ⎥∂ϕ 1∂H∂ϕ 2⎡ ⎤0⎦ + ⎣1⎦ u (15)0(= current through first inductor)with H(Q, ϕ 1 , ϕ 2 ) := H 1 (ϕ 1 )+H 2 (ϕ 2 )+H 3 (Q) the total energy. Clearly the matrix J is skewsymmetric,<strong>and</strong> since J is constant it trivially satisfies (12). In [31] it has been shown that inthis way every LC-circuit with independent elements can be modelled as a port-<strong>Hamiltonian</strong>system. Furthermore, also any LCTG-circuit with independent elements can be modelledas a port-<strong>Hamiltonian</strong> system, with J determined by Kirchh<strong>of</strong>f’s laws <strong>and</strong> the constitutiverelations <strong>of</strong> the transformers T <strong>and</strong> gyrators G.✷Example 2.2 (Actuated rigid body). Consider a rigid body spinning around its center<strong>of</strong> mass in the absence <strong>of</strong> gravity. The energy variables are the three components <strong>of</strong> the bodyangular momentum p along the three principal axes: p = (p x , p y , p z ), <strong>and</strong> the energy is thekinetic energy()H(p) = 1 p 2 x+ p2 y+ p2 z,2 I x I y I zwhere I x , I y , I z are the principal moments <strong>of</strong> inertia. Euler’s equations describing the dynamicsare⎡ ⎤⎡ ⎤ ⎡⎤ ∂Hp˙x 0 −p z p y∂p x⎣p˙y⎦ = ⎣ p z 0 −p x⎦ ⎢ ∂H ⎥⎣ ∂p y ⎦ + g(p)u (16)p˙z −p y p x 0 ∂H} {{ } ∂p zJ(p)It can be checked that the skew-symmetric matrix J(p) satisfies (12). (In fact, J(p) is thecanonical Lie-Poisson structure matrix on the dual <strong>of</strong> the Lie algebra so(3) corresponding tothe configuration space SO(3) <strong>of</strong> the rigid body.) In the scalar input case the term g(p)u6


denotes the torque around an axis with coordinates g = (b x b y b z ) T , with correspondingcollocated output given asy = b xp xI x+ b yp yI y+ b zp zI z, (17)which is the velocity around the same axis (b x b y b z ) T .Example 2.3. A third important class <strong>of</strong> <strong>systems</strong> that naturally can be written as port-<strong>Hamiltonian</strong> <strong>systems</strong>, is constituted by mechanical <strong>systems</strong> with kinematic constraints. Consideras before a mechanical system with k degrees <strong>of</strong> freedom, locally described by k configurationvariables q = (q 1 , . . . , q k ). Suppose that there are constraints on the generalizedvelocities ˙q, described asA T (q) ˙q = 0, (18)with A(q) a r × k matrix <strong>of</strong> rank r everywhere (that is, there are r independent kinematicconstraints). Classically, the constraints (18) are called holonomic if it is possible to find newconfiguration coordinates q = (q 1 , . . . , q k ) such that the constraints are equivalently expressedas˙q k−r+1 = ˙q n−r+2 = · · · = ˙q k = 0 , (19)in which case one can eliminate the configuration variables q k−r+1 , . . . , q k , since the kinematicconstraints (19) are equivalent to the geometric constraintsq k−r+1 = c k−r+1 , . . . , q k = c k , (20)for certain constants c k−r+1 , . . . , c k determined by the initial conditions. Then the systemreduces to an unconstrained system in the remaining configuration coordinates (q 1 , . . . , q k−r ).If it is not possible to find coordinates q such that (19) holds (that is, if we are not able tointegrate the kinematic constraints as above), then the constraints are called nonholonomic.The equations <strong>of</strong> motion for the mechanical system with Lagrangian L(q, ˙q) <strong>and</strong> constraints(18) are given by the Euler-Lagrange equations [35]ddt( ∂L∂ ˙q)− ∂L∂q= A(q)λ + B(q)u, λ ∈ R r , u ∈ R mA T (q) ˙q = 0 (21)where B(q)u are the external forces (<strong>control</strong>s) applied to the system, for some k × m matrixB(q), while A(q)λ are the constraint forces. The Lagrange multipliers λ(t) are uniquelydetermined by the requirement that the constraints A T (q(t)) ˙q(t) = 0 have to be satisfied forall t.Defining as before (cf. (3)) the generalized momenta the constrained Euler-Lagrangeequations (21) transform into constrained <strong>Hamiltonian</strong> equations (compare with (8)),˙q = ∂H (q, p)∂pṗ = − ∂H (q, p) + A(q)λ + B(q)u∂qy = B T (q) ∂H (q, p)∂p(22)0 = A T (q) ∂H∂p (q, p) 7✷


with H(q, p) = 1 2 pT M −1 (q)p + P (q) the total energy. The constrained state space is thereforegiven as the following subset <strong>of</strong> the phase space:X c = {(q, p) | A T (q) ∂H (q, p) = 0} (23)∂pOne way <strong>of</strong> proceeding with these equations is to eliminate the constraint forces, <strong>and</strong> to reducethe equations <strong>of</strong> motion to the constrained state space. In [49] it has been shown that thisleads to a port-<strong>Hamiltonian</strong> system (10). Furthermore, the structure matrix J c <strong>of</strong> the port-<strong>Hamiltonian</strong> system satisfies the integrability conditions (12) if <strong>and</strong> only if the constraints(18) are holonomic. (In fact, if the constraints are holonomic then the coordinates s as in(13) can be taken to be equal to the “integrated constraint functions” q k−r+1 , . . . , q k <strong>of</strong> (20),<strong>and</strong> the matrix g s as in (14) is zero.)An alternative way <strong>of</strong> approaching the system (22) is to formalize it directly as an implicitport-<strong>Hamiltonian</strong> system, as will be sketched in Section 4.2.2 Basic properties <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong>As allude to above, port-<strong>Hamiltonian</strong> <strong>systems</strong> naturally arise from a <strong>network</strong> <strong>modeling</strong> <strong>of</strong>physical <strong>systems</strong> without dissipative elements, see our papers [24, 30, 25, 32, 31, 26, 50, 48,52, 27, 46]. Recall that a port-<strong>Hamiltonian</strong> system is defined by a state space manifold Xendowed with a triple (J, g, H). The pair (J(x), g(x)) , x ∈ X , captures the interconnectionstructure <strong>of</strong> the system, with g(x) <strong>modeling</strong> in particular the ports <strong>of</strong> the system. This isvery clear in Example 2.1, where the pair (J(x), g(x)) is determined by Kirchh<strong>of</strong>f’s laws, theparadigmatic example <strong>of</strong> a power-conserving interconnection structure, but it naturally holdsfor other physical <strong>systems</strong> without dissipation as well. Independently from the interconnectionstructure, the function H : X → R defines the total stored energy <strong>of</strong> the system. Furthermore,port-<strong>Hamiltonian</strong> <strong>systems</strong> are intrinsically modular in the sense that a power-conserving interconnection<strong>of</strong> a number <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> again defines a port-<strong>Hamiltonian</strong> system,with its overall interconnection structure determined by the interconnection structures <strong>of</strong> thecomposing individual <strong>systems</strong> together with their power-conserving interconnection, <strong>and</strong> the<strong>Hamiltonian</strong> just the sum <strong>of</strong> the individual <strong>Hamiltonian</strong>s (see [52, 46, 11]).As we have seen before, a basic property <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> is the energybalancingproperty dH dt (x(t)) = uT (t)y(t). Physically this corresponds to the fact that theinternal interconnection structure is power-conserving (because <strong>of</strong> skew-symmetry <strong>of</strong> J(x)),while u <strong>and</strong> y are the power-variables <strong>of</strong> the ports defined by g(x), <strong>and</strong> thus u T y is theexternally supplied power.From the structure matrix J(x) <strong>of</strong> a port-<strong>Hamiltonian</strong> system one can directly extractuseful information about the dynamical properties <strong>of</strong> the system. Since the structure matrixis directly related to the <strong>modeling</strong> <strong>of</strong> the system (capturing the interconnection structure)this information usually has a direct physical interpretation.A very important property which may be directly inferred from the structure matrix is theexistence <strong>of</strong> dynamical invariants independent <strong>of</strong> the <strong>Hamiltonian</strong> H, called Casimir functions.Consider the set <strong>of</strong> p.d.e.’s∂ T C(x)J(x) = 0, x ∈ X , (24)∂x8


Example 2.5 (Example 2.2 continued). The quantity 1 2 p2 x + 1 2 p2 y + 1 2 p2 z (total angularmomentum) is a Casimir function.For a further discussion <strong>of</strong> the dynamical properties <strong>of</strong> <strong>Hamiltonian</strong> <strong>systems</strong> (especially ifJ satisfies the integrability conditions (12)) we refer to the extensive literature on this topic,see e.g. [1, 23].2.3 <strong>Port</strong>-<strong>Hamiltonian</strong> <strong>systems</strong> with dissipationEnergy-dissipation is included in the framework <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> (10) by terminatingsome <strong>of</strong> the ports by resistive elements. Indeed, consider instead <strong>of</strong> g(x)u in (10) aterm[g(x) gR (x) ] [ ]u= g(x)u + gu R (x)u R (28)R<strong>and</strong> extend correspondingly the output equations y = g T (x) ∂H∂x(x) to⎡[ ] g T (x) ∂Hy∂x (x) ⎤= ⎣⎦ (29)y RgR T ∂H(x)∂x (x)Here u R , y R ∈ R mr denote the power variables at the ports which are terminated by staticresistive elementsu R = −F (y R ) (30)where the resistive characteristic F : R mr → R mrsatisfiesy T RF (y R ) ≥ 0, y R ∈ R mr (31)(In many cases, F will be derivable from a so-called Rayleigh dissipation function R : R mr → Rin the sense that F (y R ) = ∂R∂y R(y R ).) In the sequel we concentrate on port-<strong>Hamiltonian</strong> <strong>systems</strong>with ports terminated by linear resistive elementsu R = −Sy R (32)for some positive semi-definite symmetric matric S = S T ≥ 0. Substitution <strong>of</strong> (32) into (28)leads to a model <strong>of</strong> the formẋ = [J(x) − R(x)] ∂H∂x(x) + g(x)uy = g T (x) ∂H∂x (x) (33)where R(x) := g R (x)SgR T (x) is a positive semi-definite symmetric matrix, depending smoothlyon x. In this case the energy-balancing property (8) takes the formdHdt (x(t)) = uT (t)y(t) − ∂T H∂H∂x(x(t))R(x(t))∂x (x(t))≤u T (t)y(t).(34)showing that a port-<strong>Hamiltonian</strong> system is passive if the <strong>Hamiltonian</strong> H is bounded frombelow. We call (33) a port-<strong>Hamiltonian</strong> system with dissipation. Note that in this case two10


Furthermore, E is a voltage source. The dynamical equations <strong>of</strong> motion can be written asthe port-<strong>Hamiltonian</strong> system with dissipation⎡ ⎤⎡⎣˙qṗ˙Q⎤⎛⎡⎦ = ⎝⎣y 1 = ∂H∂p = ˙qy 2 = 1 ∂HR0 1 0−1 0 00 0 0⎤⎡⎦ − ⎣0 0 00 c 00 0 1/R⎤⎞⎦⎠⎢⎣(39)⎡ ⎤0⎡0⎤∂Q = I + ⎣ 1 ⎦ F + ⎣001/R⎦ Ewith p the momentum, R the resistance <strong>of</strong> the resistor, I the current through the voltagesource, <strong>and</strong> the <strong>Hamiltonian</strong> H being the total energyH(q, p, Q) = 12m p2 + 1 2 k(q − ¯q)2 + 12C(q) Q2 , (40)with ¯q denoting the equilibrium position <strong>of</strong> the spring. Note that F ˙q is the mechanical power,<strong>and</strong> EI the electrical power applied to the system. In the application as a microphone thevoltage over the resistor will be used (after amplification) as a measure for the mechanicalforce F .Example 2.7. ([Ortega et al. [39]]) A permanent magnet synchronous motor can be writtenas a port-<strong>Hamiltonian</strong> system with dissipation (in a rotating reference, i.e. the dq frame) forthe state vector⎡ ⎤⎡⎤i dL d 0 0x = M ⎣ i q⎦ , M = ⎣ 0 L q 0 ⎦ (41)ω0 0jn pthe magnetic flux linkages <strong>and</strong> mechanical momentum (i d , i q being the currents, <strong>and</strong> ω theangular velocity), L d , L q stator inductances, j the moment <strong>of</strong> inertia, <strong>and</strong> n p the number<strong>of</strong> pole pairs. The <strong>Hamiltonian</strong> H(x) is given as H(x) = 1 2 xT M −1 x (total energy), whilefurthermore J(x), R(x) <strong>and</strong> g(x) are determined as⎡⎤0 L 0 x 3 0J(x) = ⎣ −L 0 x 3 0 −Φ q0⎦ ,0 Φ q0 0⎡⎤ ⎡R S 0 01 0 0R(x) = ⎣ 0 R S 0 ⎦ , g(x) = ⎣ 0 1 00 0 00 0 − 1n pwith R S the stator winding resistance, Φ q0 a constant term due to interaction <strong>of</strong> the permanentmagnet <strong>and</strong> the magnetic material in the stator, <strong>and</strong> L 0 := L d n p /j. The three inputs are thestator voltage (v d , v q ) T <strong>and</strong> the (constant) load torque. Outputs are i d , i q <strong>and</strong> ω.12⎤⎦∂H∂q∂H∂p∂H∂Q⎥⎦(42)


In some cases the interconnection structure J(x) may be actually varying, depending onthe mode <strong>of</strong> operation <strong>of</strong> the system, as exemplified by the following simple dc-to-dc powerconverter with a single switch. See for a further treatment <strong>of</strong> power converters in this context[39].Example 2.8. Consider the ideal boost converter given in Figure 3.Ls = 1Es = 0CRFigure 3: Ideal boost converterThe system equations are given as] [ẋ1ẋ 2=([ ] [ ]) [ ∂H0 −s 0 0 ∂x−1s 0 0 1/R ∂H∂x 2][ 1+ E0]y = ∂H∂x 1(43)with x 1 the magnetic flux linkage <strong>of</strong> the inductor, x 2 the charge <strong>of</strong> the capacitor, <strong>and</strong>H(x 1 , x 2 ) = 12L x2 1 + 12C x2 2 the total stored energy. The internal interconnection structure[ ] [ ]0 0 0 −1matrix J is either or, depending on the ideal switch being in position0 0 1 0s = 0 or s = 1.3 Control <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong>The aim <strong>of</strong> this section is to discuss a general methodology for port-<strong>Hamiltonian</strong> <strong>systems</strong> (withor without dissipation) which exploits their <strong>Hamiltonian</strong> properties in an intrinsic way, seee.g. [33, 34, 39, 47]. An expected benefit <strong>of</strong> such a methodology is that it leads to physicallyinterpretable <strong>control</strong>lers, which possess inherent robustness properties. Future research isaimed at corroborating these claims.We have already seen that port-<strong>Hamiltonian</strong> <strong>systems</strong> are passive if the <strong>Hamiltonian</strong> H isbounded from below. Hence in this case we can use all the results from the theory <strong>of</strong> passive<strong>systems</strong>, such as asymptotic stabilization by the insertion <strong>of</strong> damping by negative output feedback,see e.g. [47]. The emphasis in this section is however on the somewhat complementaryaspect <strong>of</strong> shaping the energy <strong>of</strong> the system, which directly involves the <strong>Hamiltonian</strong> structure<strong>of</strong> the system, as opposed to the more general passivity structure.13


3.1 Control by interconnectionConsider a port-<strong>Hamiltonian</strong> system (33)P :ẋ = [J(x) − R(x)] ∂H∂x(x) + g(x)uy = g T (x) ∂H∂x (x) x ∈ X (44)regarded as a plant system to be <strong>control</strong>led. In the previous section we have seen that manyphysical <strong>systems</strong> can be modelled in this way, <strong>and</strong> that the defining entities J(x), g(x), R(x)<strong>and</strong> H have a concrete physical interpretation. Furthermore, if H is bounded from belowthen (44) is a passive system.Recall the well-known result that the st<strong>and</strong>ard feedback interconnection <strong>of</strong> two passive <strong>systems</strong>again is a passive system; a basic fact which can be used for various stability <strong>and</strong> <strong>control</strong>purposes. In the same vein we can consider the interconnection <strong>of</strong> the plant (44) with anotherport-<strong>Hamiltonian</strong> systemC :˙ξ = [J C (ξ) − R C (ξ)] ∂H C∂ξ(ξ) + g C(ξ)u Cy C = gC T (ξ) ∂H C∂ξ (ξ) ξ ∈ X C (45)regarded as the <strong>control</strong>ler system, via the st<strong>and</strong>ard feedback interconnectionu = −y C + eu C = y + e C(46)with e, e C external signals inserted in the feedback loop. The closed-loop system takes theform[ẋ ] [J(x) −g(x)gC T = ((ξ)] [ ]R(x) 0−)˙ξ g C (ξ)g T (x) J C (ξ) 0 R C (ξ)} {{ } } {{ }J cl (x,ξ)R cl (x,ξ)[ ∂H∂x (x)] [ ] [ ]g(x) 0 e(47)∂H C∂ξ (ξ) +0 g C (ξ) e C[ ] [ ] [y g(x) 0∂H∂x=(x) ]y ∂H C 0 g C (ξ) C∂ξ (ξ)which again is a port-<strong>Hamiltonian</strong> system, with state space given by the product space X ×X C ,total <strong>Hamiltonian</strong> H(x) + H C (ξ), inputs (e, e C ) <strong>and</strong> outputs (y, y C ). Hence the feedbackinterconnection <strong>of</strong> any two port-<strong>Hamiltonian</strong> <strong>systems</strong> results in another port-<strong>Hamiltonian</strong>system; just as in the case <strong>of</strong> passivity.It is <strong>of</strong> interest to investigate the Casimir functions <strong>of</strong> the closed-loop system, especiallythose relating the state variables ξ <strong>of</strong> the <strong>control</strong>ler system to the state variables x <strong>of</strong> theplant system. Indeed, from a <strong>control</strong> point <strong>of</strong> view the <strong>Hamiltonian</strong> H is given while H Ccan be assigned. Thus if we can find Casimir functions C i (ξ, x), i = 1, · · · , r, relating ξ to xthen by the Energy-Casimir method the <strong>Hamiltonian</strong> H + H C <strong>of</strong> the closed-loop system maybe replaced by the <strong>Hamiltonian</strong> H + H C + H a (C 1 , · · · , C r ), thus creating the possibility <strong>of</strong>14


obtaining a suitable Lyapunov function for the closed-loop system.In particular, let us consider Casimir functions <strong>of</strong> the formξ i − G i (x) , i = 1, . . . , dim X C = n C (48)That means (see (35)) that we are looking for solutions <strong>of</strong> the p.d.e.’s (with e i denoting thei-th basis vector)[] ⎡ J(x) − R(x) −g(x)g− ∂T C T G (ξ) ⎤i∂x (x) eT ⎣⎦ = 0ig C (ξ)g T (x) J C (ξ) − R C (ξ)or written out∂ T G i∂x (x) [J(x) − R(x)] − gi C (ξ)gT (x) = 0∂ T G iwith ∂T G i∂x∂x (x)g(x)gT C (ξ) + J i C (ξ) − Ri C (ξ) = 0 (49)denoting as before the gradient vector(∂Gi∂x 1, . . . , ∂G i∂x n), <strong>and</strong> g i C , J i C , Ri C denotingthe i-th row <strong>of</strong> g C , J C , respectively R C .Suppose we want to solve (49) for i = 1, . . . , ¯n , with ¯n ≤ n C (possibly after permutation<strong>of</strong> ξ 1 , . . . , ξ nC ). with ¯J C (ξ), ¯R C (ξ) the ¯n × ¯n left-upper submatrices <strong>of</strong> J C , respectively R C .The following proposition has been shown in [47].Proposition 3.1. The functions ξ i − G i (x), i = 1, . . . , ¯n ≤ n C , satisfy (49) (<strong>and</strong> thus areCasimirs <strong>of</strong> the closed-loop port-<strong>control</strong>led <strong>Hamiltonian</strong> system (47) for e = 0, e c = 0) if <strong>and</strong>only if G = (G 1 , . . . , G¯n ) T satisfies∂ T G ∂G∂x(x)J(x)∂x (x) = ¯J C (ξ)R(x) ∂G∂x (x) = 0¯R C (ξ) = 0(50)∂ T G∂x (x)J(x) = ˜g C (ξ) g T (x)with ¯J C (ξ), ¯R C (ξ) the ¯n × ¯n left-upper submatrices <strong>of</strong> J C , respectively R C .In particular, we conclude that the functions ξ i −G i (x), i = 1, . . . ¯n, are Casimirs <strong>of</strong> (47) fore = 0, e C = 0, if <strong>and</strong> only if they are Casimirs for both the internal interconnection structureJ cl (x, ξ) as well as for the dissipation structure R cl (x, ξ). Hence, as in (38), it follows directlyhow the closed-loop port-<strong>control</strong>led <strong>Hamiltonian</strong> system with dissipation (47) for e = 0, e C = 0reduces to a system any multi-level set {(x, ξ) |ξ i = G i (x) + c i , i = 1, . . . , ¯n}, by restrictingboth J cl <strong>and</strong> R cl to this multi-level set.Example 3.2. [56] Consider the “plant” system[ ] [ ] ⎡ ⎤∂H[ ]˙q 0 1∂q⎢ ⎥ 0=⎣ ⎦ + uṗ −1 0 ∂H 1⎡ ⎤ ∂py = [ 0 1 ] ∂H∂q⎢ ⎥⎣ ⎦∂H∂p15(51)


with q the position <strong>and</strong> p being the momentum <strong>of</strong> the mass m, in feedback interconnection(u = −y C + e, u C = y) with the <strong>control</strong>ler system (see Figure 4)k cm ckmebFigure 4: Controlled mass⎡⎣˙∆q cṗ c˙∆q⎤⎡⎦ = ⎣0 1 0−1 −b 10 −1 0⎡⎤⎦⎢⎣∂H C∂∆q c∂H C∂p c⎤⎡+ ⎣⎥⎦001⎤⎦ u C(52)y C = ∂H C∂∆q∂H C∂∆qwhere ∆q c is the displacement <strong>of</strong> the spring k c , ∆q is the displacement <strong>of</strong> the spring k, <strong>and</strong> p cis the momentum <strong>of</strong> the mass m c . The plant <strong>Hamiltonian</strong> is H(p) = 12m p2 , <strong>and</strong> the <strong>control</strong>ler<strong>Hamiltonian</strong> is given as H C (∆q c , p c , ∆q) = 1 2 ( p2 cm c+ k(∆q) 2 + k c (∆q c ) 2 ). The variable b > 0is the damping constant, <strong>and</strong> e is an external force. The closed-loop system possesses theCasimir functionC(q, ∆q c , ∆q) = ∆q − (q − ∆q c ), (53)implying that along the solutions <strong>of</strong> the closed-loop system∆q = q − ∆q c + c (54)with c a constant depending on the initial conditions. With the help <strong>of</strong> LaSalle’s Invarianceprinciple it can be shown that restricted to the invariant manifolds (54) the system is asymptoticallystable for the equilibria q = ∆q c = p = p c = 0.✷Let us next consider the special case ¯n = n C , in which case we wish to relate all the<strong>control</strong>ler state variables ξ 1 , . . . , ξ nC to the plant state variables x via Casimir functionsξ 1 − G 1 (x), . . . , ξ nC − G nC (x). Denoting G = (G 1 , . . . , G nC ) T this means that G should satisfy(see (50))∂ T G ∂G∂x(x)J(x)∂x (x) = J C (ξ)R(x) ∂G∂x (x) = 0 = R C(ξ)(55)∂ T G∂x (x)J(x) = g C (ξ) g T (x)In this case the reduced dynamics on any multi-level setL C = {(x, ξ)|ξ i = G i (x) + c i , i = 1, . . . n C } (56)16


can be immediately recognized. Indeed, the x-coordinates also serve as coordinates for L C .Furthermore, the x-dynamics <strong>of</strong> (47) with e = 0, e C = 0 is given asẋ = [J(x) − R(x)] ∂H∂x (x) − g(x)gT C (ξ)∂H C(ξ). (57)∂ξUsing the second <strong>and</strong> the third equality <strong>of</strong> (55) this can be rewritten as( )∂H ∂Gẋ = [J(x) − R(x)] (x) +∂x ∂x (x)∂H C∂ξ (ξ) . (58)<strong>and</strong> by the chain-rule property for differentiation this reduces to the port-<strong>Hamiltonian</strong> systemẋ = [J(x) − R(x)] ∂H s(x), (59)∂xwith the same interconnection <strong>and</strong> dissipation structure as before, but with shaped <strong>Hamiltonian</strong>H s given byH s (x) = H(x) + H C (G(x) + c). (60)An interpretation <strong>of</strong> the shaped <strong>Hamiltonian</strong> H s in terms <strong>of</strong> energy-balancing is the following.Since R C (ξ) = 0 by (55) the <strong>control</strong>ler <strong>Hamiltonian</strong> H C satisfies dH Cdt= u T C y C. Hence alongany multi-level set L C given by (56), invariant for the closed loop port-<strong>Hamiltonian</strong> system(47) for e = 0, e C = 0dH sdt= dH dt + dH Cdt= dH dt − uT y (61)since u = −y C <strong>and</strong> u C = y. Therefore, up to a constant,H s (x(t)) = H(x(t)) −∫ t0u T (τ)y(τ)dτ, (62)<strong>and</strong> the shaped <strong>Hamiltonian</strong> H s is the original <strong>Hamiltonian</strong> H minus the energy suppliedto the plant system (44) by the <strong>control</strong>ler system (45) (modulo a constant; depending on theinitial states <strong>of</strong> the plant <strong>and</strong> <strong>control</strong>ler).Remark 3.3. Note that from a stability analysis point <strong>of</strong> view (62) can be regarded asan effective way <strong>of</strong> generating c<strong>and</strong>idate Lyapunov functions H s from the <strong>Hamiltonian</strong> H.(Compare with the classical construction <strong>of</strong> Lur’e functions.)Example 3.4. A mechanical system with damping <strong>and</strong> actuated by external forces u ∈ R mis described as a port-<strong>Hamiltonian</strong> system[ ˙q=ṗ]([ ] [ ] [0 I k 0 0 ) ∂H] [ ]∂q 0−+ u−I k 0 0 D(q)B(q)(63)y = B T (q) ∂H∂p∂H∂pwith x = [ q p ], where q ∈ R k are the generalized configuration coordinates, p ∈ R k thegeneralized momenta, <strong>and</strong> D(q) = D T (q) ≥ 0 is the damping matrix. In most cases the<strong>Hamiltonian</strong> H(q, p) takes the formH(q, p) = 1 2 pT M −1 (q)p + P (q) (64)17


where M(q) = M T (q) > 0 is the generalized inertia matrix,12 pT M −1 (q)p = 1 2 ˙qT M(q) ˙q isthe kinetic energy, <strong>and</strong> P (q) is the potential energy <strong>of</strong> the system. Now consider a generalport-<strong>Hamiltonian</strong> <strong>control</strong>ler system (45), with state space R m . Then the equations (55) forG = (G 1 (q, p), . . . , G m (q, p)) T take the form∂ T G ∂G∂q∂p − ∂T G∂p∂G∂q = J C (ξ)D(q) ∂G∂p = 0(65)∂ T G∂por equivalentlyJ C = 0,= 0, ∂T G∂q= g C (ξ)B T (q)∂G∂p = 0,gT C(ξ)B(q) = ∂G (q) (66)∂qNow let g C (ξ) be the m × m identity matrix. Then there exists a solution G =(G 1 (q), . . . , G m (q)) to (66) if <strong>and</strong> only if the columns <strong>of</strong> the input force matrix B(q) satisfythe integrability conditions∂B il∂q j(q) = ∂B jl∂q i(q), i, j = 1, . . . k, l = 1, . . . m (67)Hence, if B(q) satisfies (67), then the closed-loop port-<strong>Hamiltonian</strong> system (47) for the <strong>control</strong>lersystem (45) with J C = 0 admits Casimirs ξ i − G i (q), i = 1, . . . , m, leading to a reducedport-<strong>Hamiltonian</strong> system[ [ ] [ ]˙q 0 Ik 0 0= ( − )ṗ]−I k 0 0 D(q)y = B T (q) ∂Hs∂pfor the shaped <strong>Hamiltonian</strong>⎡⎢⎣∂H s∂q∂H s∂p⎤⎥⎦ +[ ] 0eB(q)H s (q, p) = H(q, p) + H C (G 1 (q) + c 1 , . . . , G m (q) + c m ) (69)If H(q, p) is as given in (64), thenH s (q, p) = 1 2 pT M −1 (q)p + [P (q) + H C (G 1 (q) + c 1 , · · · , G m (q) + c m )] (70)(68)<strong>and</strong> the <strong>control</strong> amounts to shaping the potential energy <strong>of</strong> the system, see [59, 38].✷3.2 Passivity-based <strong>control</strong> <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong>In the previous section we have seen how under certain conditions the feedback interconnection<strong>of</strong> a port-<strong>Hamiltonian</strong> system having <strong>Hamiltonian</strong> H (the “plant”) with another port-<strong>Hamiltonian</strong> system with <strong>Hamiltonian</strong> H C (the “<strong>control</strong>ler”) leads to a reduced dynamicsgiven by (see (59))ẋ = [J(x) − R(x)] ∂H s(x) (71)∂x18


for the shaped <strong>Hamiltonian</strong> H s (x) = H(x) + H C (G(x) + c), with G(x) a solution <strong>of</strong> (55).From a state feedback point <strong>of</strong> view the dynamics (71) could have been directly obtained bya state feedback u = α(x) such thatg(x)α(x) = [J(x) − R(x)] ∂H C(G(x) + c) (72)∂xIndeed, such an α(x) is given in explicit form asα(x) = −gC(G(x) T + c) ∂H C(G(x) + c) (73)∂ξA state feedback u = α(x) satisfying (72) is customarily called a passivity-based <strong>control</strong> law,since it is based on the passivity properties <strong>of</strong> the original plant system (44) <strong>and</strong> transforms(44) into another passive system with shaped storage function (in this case H s ).Seen from this perspective we have shown in the previous section that the passivity-basedstate feedback u = α(x) satisfying (72) can be derived from the interconnection <strong>of</strong> the port-<strong>Hamiltonian</strong> plant system (44) with a port-<strong>Hamiltonian</strong> <strong>control</strong>ler system (45). This fact hassome favorable consequences. Indeed, it implies that the passivity-based <strong>control</strong> law definedby (72) can be equivalently generated as the feedback interconnection <strong>of</strong> the passive system(44) with another passive system (45). In particular, this implies an inherent invarianceproperty <strong>of</strong> the <strong>control</strong>led system: the plant system (71), the <strong>control</strong>ler system (60), as wellas any other passive system interconnected to (71) in a power-conserving fashion, may changein any way as long as they remain passive, <strong>and</strong> for any perturbation <strong>of</strong> this kind the <strong>control</strong>ledsystem will remain stable.The implementation <strong>of</strong> the resulting passivity-based <strong>control</strong> u = α(x) is a somewhat complexissue. In cases <strong>of</strong> analog <strong>control</strong>ler design the interconnection <strong>of</strong> the plant port-<strong>Hamiltonian</strong>system (44) with the port-<strong>Hamiltonian</strong> <strong>control</strong>ler system (45) seems to be the logical option.Furthermore, in general it may be favorable to avoid an explicit state feedback, but insteadto use the dynamic output feedback <strong>control</strong>ler (45). On the other h<strong>and</strong>, in some applicationsthe measurement <strong>of</strong> the passive output y may pose some problems, while the state feedbacku = α(x) is in fact easier to implement, as illustrated in the next example.Example 3.5 (Example 3.4 continued). The passivity-based <strong>control</strong> u = α(x) resultingfrom (66) is given by (assuming g C to be the identity matrix <strong>and</strong> B(q) to satisfy (67))u i = − ∂H C∂ξ i(G(q) + c), i = 1, . . . m (74)This follows from (73), <strong>and</strong> can be directly checked by substituting (74) into (63) <strong>and</strong> using theequality B(q) = ∂G∂q(q). Comparing the implementation <strong>of</strong> the state feedback <strong>control</strong>ler (74)with the implementation <strong>of</strong> the port-<strong>Hamiltonian</strong> <strong>control</strong>ler system based on the measurement<strong>of</strong>y = B T (q) ∂H∂p (q, p) = BT (q) ˙q (75)one may note that the measurement <strong>of</strong> the generalized velocities (75) is in some cases (forexample in a robotics context) more problematic than the measurement <strong>of</strong> the generalizedpositions (74).✷19


Remark 3.6. On the other h<strong>and</strong>, the <strong>control</strong> <strong>of</strong> the port-<strong>Hamiltonian</strong> plant system (44) byinterconnection with the port-<strong>Hamiltonian</strong> <strong>control</strong>ler system (45) allows for the possibility <strong>of</strong>inserting an asymptotically stabilizing damping not directly in the plant but instead in the<strong>control</strong>ler system, cf. Example 3.2.In the rest <strong>of</strong> this section we concentrate on the passivity-based (state feedback) <strong>control</strong>u = α(x). The purpose is to more systematically indicate how a port-<strong>Hamiltonian</strong> system withdissipation (44) may be asymptotically stabilized around a desired equilibrium x ∗ in two steps:I Shape by passivity-based <strong>control</strong> the <strong>Hamiltonian</strong> in such a way that it has a strictminimum at x = x ∗ . Then x ∗ is a (marginally) stable equilibrium <strong>of</strong> the <strong>control</strong>led system.II Add damping to the system in such a way that x ∗ becomes an asymptotically stableequilibrium <strong>of</strong> the <strong>control</strong>led system.As before, we shall concentrate on Step I. Therefore, let us consider a port-<strong>Hamiltonian</strong>system with dissipation (44) with X the n-dimensional state space manifold. Suppose wewish to stabilize the system around a desired equilibrium x ∗ , assigning a closed-loop energyfunction H d (x) to the system which has a strict minimum at x ∗ (that is, H d (x) > H d (x ∗ ) forall x ≠ x ∗ in a neighbourhood <strong>of</strong> x ∗ ). DenoteH d (x) = H(x) + H a (x), (76)where the to be defined function H a is the energy added to the system (by the <strong>control</strong> action).We have the followingProposition 3.7. [39, 34] Assume we can find a feedback u = α(x) <strong>and</strong> a vector functionK(x) satisfyingsuch that[J(x) − R(x)] K(x) = g(x)α(x) (77)(i)∂K i∂x j(x) = ∂K j∂x i(x),i, j = 1, . . . , n(ii) K(x ∗ ) = − ∂H∂x (x∗ )(78)(iii)∂K∂x (x∗ ) > − ∂2 H(x ∗ )∂x 2with ∂K∂x the n×n matrix with i-th column given by ∂K i∂x (x), <strong>and</strong> ∂2 H(x ∗ ) denoting the Hessian∂x 2matrix <strong>of</strong> H at x ∗ . Then the closed-loop system is a <strong>Hamiltonian</strong> system with dissipationẋ = [J(x) − R(x)] ∂H d(x) (79)∂xwhere H d is given by (76), with H a such thatK(x) = ∂H a(x) (80)∂xFurthermore, x ∗ is a stable equilibrium <strong>of</strong> (79).20


A further generalization is to use state feedback in order to change the interconnectionstructure <strong>and</strong> the resistive structure <strong>of</strong> the plant system, <strong>and</strong> thereby to create more flexibilityto shape the storage function for the (modified) port-<strong>control</strong>led <strong>Hamiltonian</strong> systemto a desired form. This methodology has been called Interconnection-Damping AssignmentPassivity-Based Control (IDA-PBC) in [39, 40], <strong>and</strong> has been succesfully applied to a number<strong>of</strong> applications. The method is especially attractive if the newly assigned interconnection<strong>and</strong> resistive structures are judiciously chosen on the basis <strong>of</strong> physical considerations, <strong>and</strong>represent some “ideal” interconnection <strong>and</strong> resistive structures for the physical plant. For anextensive treatment <strong>of</strong> IDA-PBC we refer to [39, 40].4 Implicit port-<strong>Hamiltonian</strong> <strong>systems</strong>From a general <strong>modeling</strong> point <strong>of</strong> view physical <strong>systems</strong> are, at least in first instance, <strong>of</strong>tendescribed as DAE’s, that is, a mixed set <strong>of</strong> differential <strong>and</strong> algebraic equations. This stemsfrom the fact that in many <strong>modeling</strong> approaches the system under consideration is naturallyregarded as obtained from interconnecting simpler sub-<strong>systems</strong>. These interconnections ingeneral, give rise to algebraic constraints between the state space variables <strong>of</strong> the sub-<strong>systems</strong>;thus leading to implicit <strong>systems</strong>. While in the linear case one may argue that it is <strong>of</strong>tenrelatively straightforward to eliminate the algebraic constraints, <strong>and</strong> thus to reduce the systemto an explicit form, in the nonlinear case such a conversion from implicit to explicit form isusually fraught with difficulties. Indeed, if the algebraic constraints are nonlinear they neednot be analytically solvable (locally or globally). More importantly perhaps, even if they areanalytically solvable, then <strong>of</strong>ten one would prefer not to eliminate the algebraic constraints,because <strong>of</strong> the complicated <strong>and</strong> physically not easily interpretable expressions for the reducedsystem which may arise.Therefore it is important to extend the framework <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong>, assketched in the previous sections, to the context <strong>of</strong> implicit <strong>systems</strong>. In order to give thedefinition <strong>of</strong> an implicit port-<strong>Hamiltonian</strong> system (with dissipation) we first consider the notion<strong>of</strong> a Dirac structure, formalizing the concept <strong>of</strong> a power-conserving interconnection, <strong>and</strong>generalizing the notion <strong>of</strong> a structure matrix J(x) as encountered before.4.1 Power-conserving interconnectionsLet us return to the basic setting <strong>of</strong> passivity, starting with a finite-dimensional linear space<strong>and</strong> its dual, in order to define power. Thus, let F be an l-dimensional linear space, <strong>and</strong>denote its dual (the space <strong>of</strong> linear functions on F) by F ∗ . The product space F × F ∗ isconsidered to be the space <strong>of</strong> power variables, with power defined byP =< f ∗ |f >, (f, f ∗ ) ∈ F × F ∗ , (81)where < f ∗ |f > denotes the duality product, that is, the linear function f ∗ ∈ F ∗ acting onf ∈ F. Often we call F the space <strong>of</strong> flows f, <strong>and</strong> F ∗ the space <strong>of</strong> efforts e, with the power<strong>of</strong> an element (f, e) ∈ F × F ∗ denoted as < e|f >.Remark 4.1. If F is endowed with an inner product structure , then F ∗ can be naturallyidentified with F in such a way that < e|f >=< e, f >, f ∈ F, e ∈ F ∗ ≃ F.21


Example 4.2. Let F be the space <strong>of</strong> generalized velocities, <strong>and</strong> F ∗ be the space <strong>of</strong> generalizedforces, then < e|f > is mechanical power. Similarly, let F be the space <strong>of</strong> currents, <strong>and</strong> F ∗be the space <strong>of</strong> voltages, then < e|f > is electrical power.There exists on F × F ∗ a canonically defined symmetric bilinear form< (f 1 , e 1 ), (f 2 , e 2 ) > F×F ∗:=< e 1 |f 2 > + < e 2 |f 1 > (82)for f i ∈ F, e i ∈ F ∗ , i = 1, 2. Now consider a linear subspaceS ⊂ F × F ∗ (83)<strong>and</strong> its orthogonal complement with respect to the bilinear form F×F ∗ on F ×F ∗ , denotedasS ⊥ ⊂ F × F ∗ . (84)Clearly, if S has dimension d, then the subspace S ⊥ has dimension 2l − d.(F × F ∗ ) = 2l, <strong>and</strong> F×F ∗ is a non-degenerate form.)(Since dimDefinition 4.3. [9, 12, 11] A constant Dirac structure on F is a linear subspace D ⊂ F × F ∗such thatD = D ⊥ (85)It immediately follows that the dimension <strong>of</strong> any Dirac structure D on an l-dimensionallinear space is equal to l. Furthermore, let (f, e) ∈ D = D ⊥ . Then by (82)0 =< (f, e), (f, e) > F×F ∗= 2 < e|f > . (86)Thus for all (f, e) ∈ D we obtain< e | f >= 0. (87)Hence a Dirac structure D on F defines a power-conserving relation between the powervariables (f, e) ∈ F × F ∗ .Remark 4.4. The condition dim D = dim F is intimately related to the usually expressedstatement that a physical interconnection can not determine at the same time both the flow<strong>and</strong> effort (e.g. current <strong>and</strong> voltage, or velocity <strong>and</strong> force).Constant Dirac structures admit different matrix representations. Here we just list anumber <strong>of</strong> them, without giving pro<strong>of</strong>s <strong>and</strong> algorithms to convert one representation intoanother, see e.g. [11].Let D ⊂ F × F ∗ , with dim F = l, be a constant Dirac structure. Then D can be representedas1. (Kernel <strong>and</strong> Image representation, [11, 50]).D = {(f, e) ∈ F × F ∗ |F f + Ee = 0} (88)for l × l matrices F <strong>and</strong> E satisfying(i) EF T + F E T = 0(ii)rank [F .E] = l(89)22


Equivalently,D = {(f, e) ∈ F × F ∗ |f = E T λ, e = F T λ, λ ∈ R l } (90)2. (Constrained input-output representation, [11]).D = {(f, e) ∈ F × F ∗ |f = Je + Gλ, G T e = 0} (91)for an l × l skew-symmetric matrix J, <strong>and</strong> a matrix G such that ImG ={f|(f, 0) ∈ D}. Furthermore, KerJ = {e|(0, e) ∈ D}.3. (Hybrid input-output representation, [6]).Let D be given as in (88). Suppose rank F = l 1 (≤ l). Select l 1 independent columns <strong>of</strong>F , <strong>and</strong> group them into a matrix F 1 . Write[(possibly] [after]permutations) F = [F 1 .F 2 ]f<strong>and</strong>, correspondingly E = [E 1 .E 2 1 e1], f =f 2 , e =e 2 . Then the matrix [F 1 .E 2 ]can be shown to be invertible, <strong>and</strong>D ={( ) ( ) ∣( ) ( )}f1 e1 ∣∣∣ f1 e1f 2 ,e 2 e 2 = Jf 2] −1 ]with J := −[F 1 .E[F 2 2 .E 1 skew-symmetric.4. (Canonical coordinate representation, [9]).There exist linear coordinates (q, p, r, s) for F such in these coordinates <strong>and</strong> dual coordinatesfor F ∗ , (f, e) = (f q , f p , f r , f s , e q , e p , e r , e s ) ∈ D if <strong>and</strong> only if⎧⎨⎩f q = e p , f p = −e qf r = 0, e s = 0Example 4.5. Kirchh<strong>of</strong>f’s laws are an example <strong>of</strong> (88), taking F the space <strong>of</strong> currents <strong>and</strong>F ∗ the space <strong>of</strong> voltages.Given a Dirac structure D on F, the following subspaces <strong>of</strong> F, respectively F ∗ , are <strong>of</strong> importance(92)(93)G 1 := {f ∈ F | ∃e ∈ F ∗ s.t. (f, e) ∈ D}P 1 := {e ∈ F ∗ | ∃f ∈ F s.t. (f, e) ∈ D}(94)The subspace G 1 expresses the set <strong>of</strong> admissible flows, <strong>and</strong> P 1 the set <strong>of</strong> admissible efforts. Itfollows from the image representation (90) thatG 1 = Im E TP 1 = Im F T (95)23


4.2 Implicit port-<strong>Hamiltonian</strong> <strong>systems</strong>From a <strong>network</strong> <strong>modeling</strong> perspective a (lumped-parameter) physical system is naturallydescribed by a set <strong>of</strong> (possibly multi-dimensional) energy-storing elements, a set <strong>of</strong> energydissipatingor resistive elements, <strong>and</strong> a set <strong>of</strong> ports (by which interaction with the environmentcan take place), interconnected to each other by a power-conserving interconnection, see Figure5.resistiveelementsenergystoringelementspowerconservinginterconnectionportsFigure 5: Implicit port-<strong>Hamiltonian</strong> system with dissipationHere the power-conserving interconnection also includes power-conserving elements like(in the electrical domain) transformers, gyrators, or (in the mechanical domain) transformers,kinematic pairs <strong>and</strong> kinematic constraints.Associated with the energy-storing elements are energy-variables x 1 , · · · , x n , being coordinatesfor some n-dimensional state space manifold X , <strong>and</strong> a total energy H : X → R. The powerconservinginterconnection is formalized in first instance (see later on for the non-constantcase) by a constant Dirac structure D on the finite-dimensional linear space F := F S ×F R ×F P ,with F S denoting the space <strong>of</strong> flows f S connected to the energy-storing elements, F R denotingthe space <strong>of</strong> flows f R connected to the dissipative (resistive) elements, <strong>and</strong> F P the space<strong>of</strong> external flows f P which can be connected to the environment. Dually, we write F ∗ =FS ∗ × F R ∗ × F P ∗ , with e S ∈ FS ∗ the efforts connected to the energy-storing elements, e R ∈ FR∗the efforts connected to the resistive elements, <strong>and</strong> e P ∈ FP ∗ the efforts to be connected tothe environment <strong>of</strong> the system.The flow variables <strong>of</strong> the energy-storing elements are given as ẋ(t) = dxdteffort variables <strong>of</strong> the energy-storing elements as ∂H(t), t ∈ R, <strong>and</strong> the∂H∂x(x(t)) (implying that =dHdt(x(t)) is the increase in energy). In order to have a consistent sign convention for energyflow we putf S = −ẋe S = ∂H∂x (x) (96)Similarly, restricting to linear resistive elements as in (32), the flow <strong>and</strong> effort variablesconnected to the resistive elements are related asf R = −Re R (97)24


for some matrix R = R T ≥ 0.Substitution <strong>of</strong> (96) <strong>and</strong> (97) into the Dirac structure D leads to the following geometricdescription <strong>of</strong> the dynamics(f S = −ẋ, f R = −Re R , f P , e S = ∂H∂x (x), e R, e P ) ∈ D (98)We call (98) an implicit port-<strong>Hamiltonian</strong> system (with dissipation), defined with respect tothe constant Dirac structure D, the <strong>Hamiltonian</strong> H <strong>and</strong> the resistive structure R.An equational representation <strong>of</strong> an implicit port-<strong>Hamiltonian</strong> system is obtained by takinga matrix representation <strong>of</strong> the Dirac structure D as discussed in the previous subsection. Forexample, in kernel representation the Dirac structure on F = F S × F R × F P may be given asD = {(f S , f R , f P , e S , e R , e P ) |F S f S + E S e S + F R f R + E R e R + F P f P + E P e P = 0}(99)for certain matrices F S , E S , F R , E R , F P , E P satisfying(i) E S F T S + F SE T S + E RF T R + F RE T R + E P F T P + F P E T P = 0(ii)rank][F S .F R .F P .E S .E R .E P = dim F(100)Then substitution <strong>of</strong> (96) <strong>and</strong> (97) into (99) yields the following set <strong>of</strong> differential-algebraicequations for the implicit port-<strong>Hamiltonian</strong> systemF S ẋ(t) = E S∂H∂x (x(t)) − F RRe R + E R e R + F P f P + E P e P , (101)Different representations <strong>of</strong> the Dirac structure D lead to different representations <strong>of</strong> theimplicit port-<strong>Hamiltonian</strong> system, <strong>and</strong> this freedom may be exploited for simulation <strong>and</strong>analysis.Actually, for many purposes this definition <strong>of</strong> port-<strong>Hamiltonian</strong> system is not generalenough, since <strong>of</strong>ten the Dirac structure is not constant, but modulated by the state variablesx. In this case the matrices F S , E S , F R , E R , F P , E P in the kernel representation depend(smoothly) on x, leading to the implicit port-<strong>Hamiltonian</strong> systemF S (x(t))ẋ(t) = E S (x(t)) ∂H∂x (x(t)) − F R(x(t))Re R (t)+E R (x(t))e R (t) + F P (x(t))f P (t) + E P (x(t))e P (t), t ∈ R(102)withE S (x)F T S (x) + F S(x)E T S (x) + E R(x)F T R (x) + F R(x)E T R (x)+ E P (x)FP T (x) + F P (x)EP T (x) = 0, ∀x ∈ X[]rank F S (x).F R (x).F P (x).E S (x).E R (x).E P (x) = dim F(103)25


Remark 4.6. Strictly speaking the flow <strong>and</strong> effort variables ẋ(t) = −f S (t), respectively∂H∂x (x(t)) = e S(t), are not living in a constant linear space F S , respectively FS ∗ , but instead inthe tangent spaces T x(t) X , respectively co-tangent spaces Tx(t) ∗ X , to the state space manifoldX . This is formalized in the definition <strong>of</strong> a non-constant Dirac structure on a manifold; seethe references [9, 12, 11, 47].It can be checked that the definition <strong>of</strong> a port-<strong>Hamiltonian</strong> system as given in (33) is a specialcase <strong>of</strong> (102), see [47]. By the power-conservation property <strong>of</strong> a Dirac structure (cf. (87)) itfollows directly that any implicit port-<strong>Hamiltonian</strong> system satisfies the energy-balancedHdt∂H(x(t)) = == −e T R (t)Re R(t) + e T P (t)f P (t),(104)as was the case for an (explicit) port-<strong>Hamiltonian</strong> system (33).The algebraic constraints that are present in the implicit system (102) are expressed bythe subspace P 1 , <strong>and</strong> the <strong>Hamiltonian</strong> H. In fact, since the Dirac structure D is modulatedby the x-variables, also the subspace P 1 is modulated by the x-variables, <strong>and</strong> thus the effortvariables e S , e R <strong>and</strong> e P necessarily satisfy(e S , e R , e P ) ∈ P 1 (x), x ∈ X , (105)or, because <strong>of</strong> (95),e S ∈ Im FS T (x), e R ∈ Im FR T (x), e P ∈ Im FP T (x). (106)The second <strong>and</strong> third inclusions entail the expression <strong>of</strong> e R <strong>and</strong> e P in terms <strong>of</strong> the othervariables, while the first inclusion determines, since e S = ∂H∂x(x), the following algebraicconstraints on the state variables∂H∂x (x) ∈ Im F S T (x). (107)Remark 4.7. Under certain non-degeneracy conditions the elimination <strong>of</strong> the algebraic constraints(107) for an implicit port-<strong>Hamiltonian</strong> system (98) can be shown to result in anexplicit port-<strong>Hamiltonian</strong> system.The Casimir functions C : X → R <strong>of</strong> the implicit system (102) are determined by the subspaceG 1 (x). Indeed, necessarily (f S , f R , f P ) ∈ G 1 (x), <strong>and</strong> thus by (95)f S ∈ Im E T S (x), f R ∈ Im E T R(x), f P ∈ Im E T P (x). (108)Since f S = ẋ(t), the first inclusion yields the flow constraintsẋ(t) ∈ Im E T S (x(t)), t ∈ R. (109)dCThus C : X → R is a Casimir function ifdt (x(t)) = ∂T C∂x(x(t))ẋ(t) = 0 for all ẋ(t) ∈Im ES T (x(t)). Hence C : X → R is a Casimir <strong>of</strong> the implicit port-<strong>Hamiltonian</strong> system (98) ifit satisfies the set <strong>of</strong> p.d.e.’s∂C∂x (x) ∈ Ker E S(x) (110)26


Remark 4.8. Note that C : X → R satisfying (110) is a Casimir function <strong>of</strong> (98) in a strongsense: it is a dynamical invariant ( dCdt (x(t)) = 0) for every port behavior <strong>and</strong> every resistiverelation (97).Example 4.9. [11, 51] The constrained <strong>Hamiltonian</strong> equations (22) can be viewed as animplicit port-<strong>Hamiltonian</strong> system, with respect to the Dirac structure D, given in constrainedinput-output representation (91) byD = {(f S , f P , e S , e P )|0 = A T (q)e S , e P = B T (q)e S ,−f S =[ 0 Ik−I k 0] [e S +0A(q)] [λ +0B(q)]f P , λ ∈ R r }(111)In this case, the algebraic constraints on the state variables (q, p) are given as0 = A T (q) ∂H (q, p) (112)∂pwhile the Casimir functions C are determined by the equations∂ T C∂q (q) ˙q = 0, for all ˙q satisfying AT (q) ˙q = 0. (113)Hence, finding Casimir functions amounts to integrating the kinematic constraints A T (q) ˙q =0. In particular, if the kinematic constraints are holonomic, <strong>and</strong> thus can be expressed as(19), then ¯q k−r+1 , · · · , ¯q k generate all the Casimir functions. ✷Remark 4.10. For a proper notion <strong>of</strong> integrability <strong>of</strong> non-constant Dirac structures, generalizingthe integrability conditions (12) <strong>of</strong> the structure matrix J(x), we refer e.g. to [11].In principle, the theory presented before for explicit port-<strong>Hamiltonian</strong> <strong>systems</strong> can be directlyextended, mutatis mut<strong>and</strong>is, to implicit port-<strong>Hamiltonian</strong> system. In particular, the st<strong>and</strong>ardfeedback interconnection <strong>of</strong> an implicit port-<strong>Hamiltonian</strong> system P with port variables f P , e P(the “plant”) with another implicit port-<strong>Hamiltonian</strong> system with port variables f C P , eC P (the“<strong>control</strong>ler”) is readily seen to result in a closed-loop implicit port-<strong>Hamiltonian</strong> system withport variables. Furthermore, as in the explicit case, the <strong>Hamiltonian</strong> <strong>of</strong> this closed-loop systemis just the sum <strong>of</strong> the <strong>Hamiltonian</strong> <strong>of</strong> the plant port-<strong>Hamiltonian</strong> system <strong>and</strong> the <strong>Hamiltonian</strong><strong>of</strong> the <strong>control</strong>ler port-<strong>Hamiltonian</strong> system. Finally, the Casimir analysis for the closed-loopsystem can be performed along the same lines as before.5 Distributed-parameter port-<strong>Hamiltonian</strong> <strong>systems</strong>From a <strong>modeling</strong> <strong>and</strong> <strong>control</strong> point <strong>of</strong> view it is very desirable to be able to include distributedparametercomponents into the <strong>Hamiltonian</strong> description <strong>of</strong> complex physical <strong>systems</strong>. However,in extending the <strong>Hamiltonian</strong> theory as for instance exposed in [37] to distributedparameter<strong>control</strong> <strong>systems</strong> a fundamental difficulty arises in the treatment <strong>of</strong> boundary conditions.Indeed, the treatment <strong>of</strong> infinite-dimensional <strong>Hamiltonian</strong> <strong>systems</strong> in the literatureseems mostly focussed on <strong>systems</strong> with infinite spatial domain, where the variables go to zer<strong>of</strong>or the spatial variables tending to infinity, or on <strong>systems</strong> with boundary conditions such thatthe energy exchange through the boundary is zero. On the other h<strong>and</strong>, from a <strong>control</strong> <strong>and</strong>27


interconnection point <strong>of</strong> view it is essential to be able to describe a distributed-parametersystem with varying boundary conditions inducing energy exchange through the boundary,since in many applications interaction with the environment (e.g. actuation or measurement)takes place through the boundary <strong>of</strong> the system. Clear examples are the telegraph equations(describing the dynamics <strong>of</strong> a transmission line), where the boundary <strong>of</strong> the system is describedby the behavior <strong>of</strong> the voltages <strong>and</strong> currents at both ends <strong>of</strong> the transmission line,or a vibrating string (or, more generally, a flexible beam), where it is natural to consider theevolution <strong>of</strong> the forces <strong>and</strong> velocities at the ends <strong>of</strong> the string. Furthermore, in both examplesit is obvious that in general the boundary exchange <strong>of</strong> power (voltage times current in thetransmission line example, <strong>and</strong> force times velocity for the vibrating string) will be non-zero,<strong>and</strong> that in fact one would like to consider the voltages <strong>and</strong> currents or forces <strong>and</strong> velocitiesas additional boundary variables <strong>of</strong> the system, which can be interconnected to other <strong>systems</strong>.Also for numerical integration <strong>and</strong> simulation <strong>of</strong> complex distibuted-parameter <strong>systems</strong> itis essential to be able to describe the complex system as the interconnection or coupling <strong>of</strong>its sub<strong>systems</strong> via their boundary variables; for example in the case <strong>of</strong> coupled fluid-soliddynamics.From a mathematical point <strong>of</strong> view, it is not obvious how to incorporate non-zero energyflow through the boundary in the existing <strong>Hamiltonian</strong> framework for distributed-parameter<strong>systems</strong>. The problem is already illustrated by the <strong>Hamiltonian</strong> formulation <strong>of</strong> e.g. theKorteweg-de Vries equation (see e.g. [37]). Here for zero boundary conditions a Poissonbracket can be formulated with the use <strong>of</strong> the differential operator , since by integrationby parts this operator is obviously skew-symmetric. However, for boundary conditions correspondingto non-zero energy flow the differential operator is not skew-symmetric anymore(since after integrating by parts the remainders are not zero ).In [54, 28, 29] we have provided a framework to overcome this fundamental problem byusing the notion <strong>of</strong> an infinite-dimensional Dirac structure. The infinite-dimensional Diracstructure employed in these papers has a specific form by being defined on certain spaces <strong>of</strong>differential forms on the spatial domain <strong>of</strong> the system <strong>and</strong> its boundary, <strong>and</strong> making use <strong>of</strong>Stokes’ theorem. Its construction emphasizes the geometrical content <strong>of</strong> the physical variablesinvolved, by identifying them as differential k-forms, for appropriate k. This framework hasbeen used [54] for a port-<strong>Hamiltonian</strong> representation <strong>of</strong> the telegrapher’s equations <strong>of</strong> an idealtransmission line, Maxwell’s equations on a bounded domain with non-zero Poynting vectorat its boundary, a vibrating string with traction forces at its ends, <strong>and</strong> planar beam models[17, 18]. Furthermore the framework has been extended to cover Euler’s equations for an idealfluid on a domain with permeable boundary, see also [53].Throughout, let Z be an n-dimensional smooth manifold with smooth (n−1)-dimensionalboundary ∂Z, representing the space <strong>of</strong> spatial variables.Denote by Ω k (Z), k = 0, 1, · · · , n, the space <strong>of</strong> exterior k-forms on Z, <strong>and</strong> by Ω k (∂Z), k =0, 1, · · · , n − 1, the space <strong>of</strong> k-forms on ∂Z. (Note that Ω 0 (Z), respectively Ω 0 (∂Z), is thespace <strong>of</strong> smooth functions on Z, respectively ∂Z.) Clearly, Ω k (Z) <strong>and</strong> Ω k (∂Z) are (infinitedimensional)linear spaces (over R). Furthermore, there is a natural pairing between Ω k (Z)<strong>and</strong> Ω n−k (Z) given by∫< β|α >:=Zβ ∧ α (∈ R) (114)with α ∈ Ω k (Z), β ∈ Ω n−k (Z), where ∧ is the usual wedge product <strong>of</strong> differential formsyielding the n-form β ∧ α. In fact, the pairing (114) is non-degenerate in the sense that if28ddx


β|α >= 0 for all α, respectively for all β, then β = 0, respectively α = 0.Similarly, there is a pairing between Ω k (∂Z) <strong>and</strong> Ω n−1−k (∂Z) given by∫< β|α >:= β ∧ α (115)∂Zwith α ∈ Ω k (∂Z), β ∈ Ω n−1−k (∂Z). Now let us define the linear spaceF p,q := Ω p (Z) × Ω q (Z) × Ω n−p (∂Z), (116)for any pair p, q <strong>of</strong> positive integers satisfyingp + q = n + 1, (117)<strong>and</strong> correspondingly let us defineE p,q := Ω n−p (Z) × Ω n−q (Z) × Ω n−q (∂Z). (118)Then the pairing (114) <strong>and</strong> (115) yields a (non-degenerate) pairing between F p,q <strong>and</strong> E p,q(note that by (117) (n − p) + (n − q) = n − 1). As before, symmetrization <strong>of</strong> this pairingyields the following bilinear form on F p,q × E p,q with values in R:≪ ( f 1 p , f 1 q , f 1 b , e1 p, e 1 q, e 1 b),(f2p , f 2 q , f 2 b , e2 p, e 2 q, e 2 b)≫:=∫Zwhere for i = 1, 2[e1p ∧ f 2 p + e 1 q ∧ f 2 q + e 2 p ∧ f 1 p + e 2 q ∧ f 1 q]+∫f i p ∈ Ωp (Z), f i q ∈ Ωq (Z)e i p ∈ Ω n−p (Z), e i p ∈ Ω n−q (Z)∂Z[e1b∧ fb 2 + e2 b ∧ f b1 ](119)(120)f i b ∈ Ωn−p (∂Z), e i b ∈ Ωn−q (∂Z)The spaces <strong>of</strong> differential forms Ω p (Z) <strong>and</strong> Ω q (Z) will represent the energy variables <strong>of</strong> twodifferent physical energy domains interacting with each other, while Ω n−p (∂Z) <strong>and</strong> Ω n−q (∂Z)will denote the boundary variables whose (wedge) product represents the boundary energyflow. For example, in Maxwell’s equations (Example 5.4) we will have n = 3 <strong>and</strong> p = q = 2;with Ω p (Z) = Ω 2 (Z), respectively Ω q (Z) = Ω 2 (Z), being the space <strong>of</strong> electric field inductions,respectively magnetic field inductions, <strong>and</strong> Ω n−p (∂Z) = Ω 1 (∂Z) denoting the electric <strong>and</strong>magnetic field intensities at the boundary, with product the Poynting vector.Theorem 5.1. Consider F p,q <strong>and</strong> E p,q given in (116), (118) with p, q satisfying (117), <strong>and</strong>bilinear form ≪, ≫ given by (119). Define the following linear subspace D <strong>of</strong> F p,q × E p,qD = {(f p , f q , f b , e p , e q , e b ) ∈ F p,q × E p,q |[ ] [fp 0 (−1)=r ] [ ]d ep,f q d 0 e q[ ] [ ] [ ]fb 1 0 ep|∂Z=e b 0 −(−1) n−q } (121)e q|∂Zwhere | ∂Z denotes restriction to the boundary ∂Z, <strong>and</strong> r := pq + 1. Then D = D ⊥ , that is,D is a Dirac structure.29


The pro<strong>of</strong> <strong>of</strong> Theorem 5.1 relies on Stokes’ theorem, <strong>and</strong> the infinite-dimensional Diracstructure defined in Theorem 5.1 is therefore called a Stokes-Dirac structure.The definition <strong>of</strong> a distributed-parameter <strong>Hamiltonian</strong> system with respect to a Stokes-Dirac structure can now be stated as follows. Let Z be an n-dimensional manifold withboundary ∂Z, <strong>and</strong> let D be a Stokes-Dirac structure. Consider furthermore a <strong>Hamiltonian</strong>density (energy per volume element)H : Ω p (Z) × Ω q (Z) × Z → Ω n (Z) (122)resulting in the total energy∫H := H ∈ R (123)ZRecall, see (114), that there exists a non-degenerate pairing between Ω p (Z) <strong>and</strong> Ω n−p (Z),respectively between Ω q (Z) <strong>and</strong> Ω n−q (Z). This means that Ω n−p (Z) <strong>and</strong> Ω n−q (Z) can beregarded as dual spaces to Ω p (Z), respectively Ω q (Z) (although strictly contained in theirfunctional analytic duals). Let now α p , ∂α p ∈ Ω p (Z), α q , ∂α q ∈ Ω q (Z). Then under weaksmoothness conditions on H∫H(α p + ∂α p , α q + ∂α q ) = H (α p + ∂α p , α q + ∂α q , z) =∫Z∫H (α p , α q , z) + [δ p H ∧ ∂α p + δ q H ∧ ∂α q ] + higher order terms in ∂α p , ∂α q (124)Zfor certain differential formsδ p H ∈ Ω n−p (Z)δ q H ∈ Ω n−q (Z)Z(125)Furthermore, from the non-degeneracity <strong>of</strong> the pairing between Ω p (Z) <strong>and</strong> Ω n−p (Z), respectivelybetween Ω q (Z) <strong>and</strong> Ω n−q (Z), it immediately follows that these differential forms areuniquely determined. This means that (δ p H, δ q H) ∈ Ω n−p (Z) × Ω n−q (Z) can be regardedas the (partial) variational derivatives (see e.g. [37]) <strong>of</strong> H at (α p , α q ) ∈ Ω p (Z) × Ω q (Z).Throughout this paper we shall assume that the <strong>Hamiltonian</strong> H admits variational derivativessatisfying (124).Now consider a time-function(α p (t), α q (t)) ∈ Ω p (Z) × Ω q (Z), t ∈ R, (126)<strong>and</strong> the <strong>Hamiltonian</strong> H(α p (t), α q (t)) evaluated along this trajectory. It follows that at anytime t[dHdt∫Z= δ p H ∧ ∂α p∂t + δ qH ∧ ∂α ]q(127)∂tThe differential forms ∂αp∂trepresent the generalized velocities <strong>of</strong> the energy variablesα p , α q . They are connected to the Stokes-Dirac structure D by setting∂t , ∂αqf p = − ∂αp∂tf q = − ∂αq∂t(128)30


(again the minus sign is included to have a consistent energy flow description). Since theright-h<strong>and</strong> side <strong>of</strong> (127) is the rate <strong>of</strong> increase <strong>of</strong> the stored energy H, we sete p = δ p He q = δ q H(129)Definition 5.2. The distributed-parameter port-<strong>Hamiltonian</strong> system with n-dimensionalmanifold <strong>of</strong> spatial variables Z, state space Ω p (Z) × Ω q (Z) (with p + q = n + 1), Stokes-Dirac structure D given by (121), <strong>and</strong> <strong>Hamiltonian</strong> H, is given as (with r = pq + 1)[− ∂αp∂t− ∂αq∂t][fbe b]==[ 0 (−1) r dd 0] [δp Hδ q H[ ] [ ]1 0 δp H| ∂Z0 −(−1) n−q δ q H| ∂Z](130)By the power-conserving property <strong>of</strong> any Dirac structure it immediately follows that forany (f p , f q , f b , e p , e q , e b ) in the Stokes-Dirac structure D∫∫[e p ∧ f p + e q ∧ f q ] + e b ∧ f b = 0 (131)Z∂ZHence by substitution <strong>of</strong> (128), (129) <strong>and</strong> using (127) we obtainProposition 5.3. Consider the distributed parameter port-<strong>Hamiltonian</strong> system (130). ThendHdt∫∂Z= e b ∧ f b , (132)expressing that the increase in energy on the domain Z is equal to the power supplied to thesystem through the boundary ∂Z.The system (130) can be called a (nonlinear) boundary <strong>control</strong> system in the sense <strong>of</strong> e.g.[14]. Indeed, we could interpret f b as the boundary <strong>control</strong> inputs to the system, <strong>and</strong> e b asthe measured outputs (or the other way around).Example 5.4 (Maxwell’s equations). Let Z ⊂ R 3 be a 3-dimensional manifold withboundary ∂Z, defining the spatial domain, <strong>and</strong> consider the electromagnetic field in Z. Theenergy variables are the electric field induction 2-form α p = D ∈ Ω 2 (Z):D = 1 2 D ij(t, z)dz i ∧ dz j (133)<strong>and</strong> the magnetic field induction 2-form α q = B ∈ Ω 2 (Z) :B = 1 2 B ij(t, z)dz i ∧ dz j (134)The corresponding Stokes-Dirac structure (n = 3, p = 2, q = 2) is given as (cf. (121))[ ] [ ] [ ] [ ] [ ] [ ]fp 0 −d ep fb 1 0 ep|∂Z=, =f q d 0 e q e b 0 1 e q|∂Z(135)31


Usually in this case one does not start with the definition <strong>of</strong> the total energy (<strong>Hamiltonian</strong>)H, but instead with the co-energy variables δ p H, δ q H, given, respectively, as the electric fieldintensity E ∈ Ω 1 (Z) :E = E i (t, z)dz i (136)<strong>and</strong> the magnetic field intensity H ∈ Ω 1 (Z) :H = H i (t, z)dz i (137)They are related to the energy variables through the constitutive relations <strong>of</strong> the medium (ormaterial equations)∗D = ɛE∗B = µH(138)with the scalar functions ɛ(t, z) <strong>and</strong> µ(t, z) denoting the electric permittivity, respectivelymagnetic permeability, <strong>and</strong> ∗ denoting the Hodge star operator (corresponding to a Riemannianmetric on Z), converting 2-forms into 1-forms. Then one defines the <strong>Hamiltonian</strong> Has∫1H = (E ∧ D + H ∧ B), (139)2Z<strong>and</strong> one immediately verifies that δ p H = E, δ q H = H.Nevertheless there are other cases (corresponding to a nonlinear theory <strong>of</strong> the electromagneticfield, such as the Born-Infeld theory, see e.g. [21]) where one starts with a more general<strong>Hamiltonian</strong> H = ∫ Zh, with the energy density h(D, B) being a more general expressionthan 1 2 (ɛ−1 ∗ D ∧ D + µ −1 ∗ B ∧ B).Assuming that there is no current in the medium Maxwell’s equations can now be writtenas (see [21])∂D∂t= dH∂B∂t= −dE(140)Explicitly taking into account the behavior at the boundary, Maxwell’s equations on a domainZ ⊂ R 3 are then represented as the port-<strong>Hamiltonian</strong> system with respect to the Stokes-Diracstructure given by (135), as[ −∂D∂t− ∂B∂t][fbe b]==[ 0 −dd 0[ ]δD H| ∂Zδ B H| ∂Z] [δD Hδ B H](141)Note that the first line <strong>of</strong> (140) is nothing else than (the differential version <strong>of</strong>) Ampère’s law,while the second line <strong>of</strong> (140) is Faraday’s law. Hence the Stokes-Dirac structure in (140),(141) expresses the basic physical laws connecting D, B, H <strong>and</strong> E.The energy-balance (132) in the case <strong>of</strong> Maxwell’s equations takes the formdHdt∫∂Z= δ B H ∧ δ D H =∫∂Z∫H ∧ E = −with E ∧ H a 2-form corresponding to the Poynting vector (see [21]).32∂ZE ∧ H (142)


Example 5.5 (Vibrating string). Consider an elastic string subject to traction forces atits ends. The spatial variable z belongs to the interval Z = [0, 1] ⊂ R. Let us denote byu(t, z) the displacement <strong>of</strong> the string. The elastic potential energy is a function <strong>of</strong> the straingiven by the 1-formα q (t) = ɛ(t, z)dz = ∂u (t, z)dz (143)∂zThe associated co-energy variable is the stress given by the 0-formσ = T ∗ α q (144)with T the elasticity modulus <strong>and</strong> ∗ the Hodge star operator. Hence the potential energy isthe quadratic functionU(α q ) =∫ 10σα q =∫ 10T ∗ α q ∧ α q =∫ 10T( ) ∂u 2dz (145)∂z<strong>and</strong> σ = δ q U.The kinetic energy K is a function <strong>of</strong> the kinetic momentum defined as the 1-formα p (t) = p(t, z)dz (146)given by the quadratic functionK(α p ) =∫ 10p 2dz (147)µThe associated co-energy variable is the velocity given by the 0-formv = 1 µ ∗ α p = δ p K (148)In this case the Dirac structure is the Stokes-Dirac structure for n = p = q = 1, with anopposite sign convention leading to the equations (with H := U + K)[ ] [ ] [ ]− ∂αp∂t0 −d δp H=− ∂αq −d 0 δ∂tq H[fbe b]=[ 1 00 1or, in more down-to-earth notation∂p∂t= ∂σ∂z = ∂ ∂z(T ɛ)( )∂ɛ∂t= ∂v∂z = ∂ 1∂z µ p] [ ]δp H| ∂Zδ q H| ∂Z(149)(150)f b = v| {0,1}33


with boundary variables the velocity <strong>and</strong> stress at the ends <strong>of</strong> the ( string. ) Of course, bysubstituting ɛ = ∂u∂zinto the 2nd equation <strong>of</strong> (150) one obtains ∂ ∂u∂z ∂t − p µ= 0, implyingthatp = µ ∂u∂t+ µf(t) (151)for some function f, which may be set to zero. Substitution <strong>of</strong> (151) into the first equation<strong>of</strong> (150) then yields the wave equationµ ∂2 u∂t 2 = ∂ (T ∂u )∂z ∂zThis framework can be extended to general beam models, see [17, 18].(152)6 Conclusions <strong>and</strong> future researchThe theory presented in this paper opens up the way for many analysis, simulation <strong>and</strong><strong>control</strong> problems. Its potential for set-point regulation has already received some attention(see [33, 34, 39, 40, 47]), while the extension to tracking problems is wide open. In this contextwe also like to refer to some recent work concerned with the shaping <strong>of</strong> the Lagrangian, see e.g.[5]. Also, the <strong>control</strong> <strong>of</strong> mechanical <strong>systems</strong> with nonholonomic kinematic constraints can befruitfully approached from this point <strong>of</strong> view, see e.g. [15], as well as the modelling <strong>and</strong> <strong>control</strong><strong>of</strong> multi-body <strong>systems</strong>, see [32, 27, 58]. The framework <strong>of</strong> port-<strong>Hamiltonian</strong> <strong>systems</strong> seemsperfectly suited to theoretical investigations on the topic <strong>of</strong> impedance <strong>control</strong>; see already[56] for some initial results in this direction. Furthermore, the connection with multi-modal(hybrid) <strong>systems</strong>, corresponding to port-<strong>Hamiltonian</strong> <strong>systems</strong> with varying interconnectionstructure, needs further investigations. Some applications <strong>of</strong> the framework <strong>of</strong> distributedparameterport-<strong>Hamiltonian</strong> <strong>systems</strong> have been already reported in [53, 42]; see also [55] forrelated work on smart structures. Finally, current research is concerned with the spatialdiscretization <strong>of</strong> port-<strong>Hamiltonian</strong> distributed-parameter <strong>systems</strong> to finite-dimensional port-<strong>Hamiltonian</strong> <strong>systems</strong> [16], with direct applications towards simulation <strong>and</strong> <strong>control</strong> <strong>of</strong> such<strong>systems</strong>.References[1] R.A. Abraham & J.E. Marsden, Foundations <strong>of</strong> Mechanics (2nd edition), Reading, MA:Benjamin/Cummings, 1978.[2] V. I. Arnold, B.A. Khesin, Topological Methods in Hydrodynamics, Springer Verlag, AppliedMathematical Sciences 125, New York, 1998.[3] G. Blankenstein, Implicit <strong>Hamiltonian</strong> Systems; Symmetry <strong>and</strong> Interconnection, PhDthesis, University <strong>of</strong> Twente, The Netherl<strong>and</strong>s, November 2000.[4] G. Blankenstein, A.J. van der Schaft, “Symmetry <strong>and</strong> reduction in implicit generalized<strong>Hamiltonian</strong> <strong>systems</strong>”, Rep. Math. Phys., 47, pp. 57–100, 2001.34


[5] A. Bloch, N. Leonard & J.E. Marsden, “Matching <strong>and</strong> stabilization by the method <strong>of</strong><strong>control</strong>led Lagrangians”, in Proc. 37th IEEE Conf. on Decision <strong>and</strong> Control, Tampa, FL,pp. 1446-1451, 1998.[6] A.M. Bloch & P.E. Crouch, “Representations <strong>of</strong> Dirac structures on vector spaces <strong>and</strong>nonlinear LC circuits”, Proc. Symposia in Pure Mathematics, Differential Geometry <strong>and</strong>Control Theory, G. Ferreyra, R. Gardner, H. Hermes, H. Sussmann, eds., Vol. 64, pp.103-117, AMS, 1999.[7] P.C. Breedveld, Physical <strong>systems</strong> theory in terms <strong>of</strong> bond graphs, PhD thesis, University<strong>of</strong> Twente, Faculty <strong>of</strong> Electrical Engineering, 1984[8] R.W. Brockett, “Control theory <strong>and</strong> analytical mechanics”, in Geometric Control Theory,(eds. C. Martin, R. Hermann), Vol. VII <strong>of</strong> Lie Groups: History, Frontiers <strong>and</strong>Applications, Math. Sci. Press, Brookline, pp. 1-46, 1977.[9] T.J. Courant, “Dirac manifolds”, Trans. American Math. Soc., 319, pp. 631-661, 1990.[10] P.E. Crouch & A.J. van der Schaft, Variational <strong>and</strong> <strong>Hamiltonian</strong> Control Systems, Lect.Notes in Control <strong>and</strong> Inf. Sciences 101, Springer-Verlag, Berlin, 1987.[11] M. Dalsmo & A.J. van der Schaft, “On representations <strong>and</strong> integrability <strong>of</strong> mathematicalstructures in energy-conserving physical <strong>systems</strong>”, SIAM J. Control <strong>and</strong> Optimization,37, pp. 54-91, 1999.[12] I. Dorfman, Dirac Structures <strong>and</strong> Integrability <strong>of</strong> Nonlinear Evolution Equations, JohnWiley, Chichester, 1993.[13] G. Escobar, A.J. van der Schaft & R. Ortega, “A <strong>Hamiltonian</strong> viewpoint in the modelling<strong>of</strong> switching power converters”, Automatica, Special Issue on Hybrid Systems, 35, pp.445-452, 1999.[14] H.O. Fattorini, ”Boundary <strong>control</strong> <strong>systems</strong>”, SIAM J. Control, 6, pp. 349-385, 1968.[15] K. Fujimoto, T. Sugie, “Stabilization <strong>of</strong> a class <strong>of</strong> <strong>Hamiltonian</strong> <strong>systems</strong> with nonholonomicconstraints via canonical transformations”, Proc. European Control Conference’99, Karlsruhe, 31 August - 3 September 1999.[16] G. Golo, V. Talasila, A.J. van der Schaft, “Approximation <strong>of</strong> the telegrapher’s equations”,Proc. 41st IEEE Conf. Decision <strong>and</strong> Control, Las Vegas, Nevada, December 2002.[17] G. Golo, V. Talasila, A.J. van der Schaft, “A <strong>Hamiltonian</strong> formulation <strong>of</strong> the Timoshenkobeam”, Mechatronics 2002, pp. 838–847, Enschede, 24-26 June 2002.[18] G. Golo, A.J. van der Schaft, S.Stramigioli, “<strong>Hamiltonian</strong> formulation <strong>of</strong> planar beams”,Proceedings 2nd IFAC Workshop on Lagrangian <strong>and</strong> <strong>Hamiltonian</strong> Methods for NonlinearControl, Eds. A. Astolfi, F. Gordillo, A.J. van der Schaft, pp. 169–174, Sevilla, 2003.[19] G. Golo, A. van der Schaft, P.C. Breedveld, B.M. Maschke, “<strong>Hamiltonian</strong> formulation <strong>of</strong>bond graphs”, Nonlinear <strong>and</strong> Hybrid Systems in Automotive Control Eds. R. Johansson,A. Rantzer, pp. 351–372, Springer London, 2003.35


[20] D.J. Hill & P.J. Moylan, “Stability <strong>of</strong> nonlinear dissipative <strong>systems</strong>,” IEEE Trans. Aut.Contr., AC-21, pp. 708-711, 1976.[21] R.S. Ingarden, A. Jamiolkowski, Classical Electrodynamics, PWN-Polish Sc. Publ.,Warszawa, Elsevier, 1985.[22] A. Isidori, Nonlinear Control Systems (2nd Edition), Communications <strong>and</strong> Control EngineeringSeries, Springer-Verlag, London, 1989, 3rd Edition, 1995.[23] J.E. Marsden & T.S. Ratiu, Introduction to Mechanics <strong>and</strong> Symmetry, Texts in AppliedMathematics 17, Springer-Verlag, New York, 1994.[24] B.M. Maschke & A.J. van der Schaft, “<strong>Port</strong>-<strong>control</strong>led <strong>Hamiltonian</strong> <strong>systems</strong>: Modellingorigins <strong>and</strong> system-theoretic properties”, in Proc. 2nd IFAC NOLCOS, Bordeaux, pp.282-288, 1992.[25] B.M. Maschke & A.J. van der Schaft, “System-theoretic properties <strong>of</strong> port-<strong>control</strong>led<strong>Hamiltonian</strong> <strong>systems</strong>”, in Systems <strong>and</strong> Networks: Mathematical Theory <strong>and</strong> Applications,Vol. II, Akademie-Verlag, Berlin, pp. 349-352, 1994.[26] B.M. Maschke & A.J. van der Schaft, “Interconnection <strong>of</strong> <strong>systems</strong>: the <strong>network</strong>paradigm”, in Proc. 35th IEEE Conf. on Decision <strong>and</strong> Control, Kobe, Japan, pp. 207-212,1996.[27] B.M. Maschke & A.J. van der Schaft, “Interconnected Mechanical Systems, Part II:The Dynamics <strong>of</strong> Spatial Mechanical Networks”, in Modelling <strong>and</strong> Control <strong>of</strong> MechanicalSystems, (eds. A. Astolfi, D.J.N. Limebeer, C. Melchiorri, A. Tornambe, R.B. Vinter),pp. 17-30, Imperial College Press, London, 1997.[28] B.M. Maschke, A.J. van der Schaft, “<strong>Port</strong> <strong>control</strong>led <strong>Hamiltonian</strong> representation <strong>of</strong> distributedparameter <strong>systems</strong>”, Proc. IFAC Workshop on Lagrangian <strong>and</strong> <strong>Hamiltonian</strong>methods for nonlinear <strong>control</strong>, Princeton University, Editors N.E. Leonard, R. Ortega,pp.28-38, 2000.[29] B.M. Maschke, A.J. van der Schaft, “<strong>Hamiltonian</strong> representation <strong>of</strong> distributed parameter<strong>systems</strong> with boundary energy flow”, Nonlinear Control in the Year 2000. Eds. A. Isidori,F. Lamnabhi-Lagarrigue, W. Respondek, Lect. Notes Control <strong>and</strong> Inf. Sciences, vol. 258,Springer-Verlag, pp. 137-142, 2000.[30] B.M. Maschke, A.J. van der Schaft & P.C. Breedveld, “An intrinsic <strong>Hamiltonian</strong> formulation<strong>of</strong> <strong>network</strong> dynamics: non-st<strong>and</strong>ard Poisson structures <strong>and</strong> gyrators”, J. FranklinInstitute, vol. 329, no.5, pp. 923-966, 1992.[31] B.M. Maschke, A.J. van der Schaft & P.C. Breedveld, “An intrinsic <strong>Hamiltonian</strong> formulation<strong>of</strong> the dynamics <strong>of</strong> LC-circuits, IEEE Trans. Circ. <strong>and</strong> Syst., CAS-42, pp. 73-82,1995.[32] B.M. Maschke, C. Bidard & A.J. van der Schaft, “Screw-vector bond graphs for thekinestatic <strong>and</strong> dynamic <strong>modeling</strong> <strong>of</strong> multibody <strong>systems</strong>”, in Proc. ASME Int. Mech.Engg. Congress, 55-2, Chicago, U.S.A., pp. 637-644, 1994.36


[33] B.M. Maschke, R. Ortega & A.J. van der Schaft, “Energy-based Lyapunov functionsfor forced <strong>Hamiltonian</strong> <strong>systems</strong> with dissipation”, in Proc. 37th IEEE Conference onDecision <strong>and</strong> Control, Tampa, FL, pp. 3599-3604, 1998.[34] B.M. Maschke, R. Ortega, A.J. van der Schaft & G. Escobar, “An energy-based derivation<strong>of</strong> Lyapunov functions for forced <strong>systems</strong> with application to stabilizing <strong>control</strong>”, in Proc.14th IFAC World Congress, Beijing, Vol. E, pp. 409-414, 1999.[35] J.I. Neimark & N.A. Fufaev, Dynamics <strong>of</strong> Nonholonomic Systems, Vol. 33 <strong>of</strong> Translations<strong>of</strong> Mathematical Monographs, American Mathematical Society, Providence, RhodeIsl<strong>and</strong>, 1972.[36] H. Nijmeijer & A.J. van der Schaft, Nonlinear Dynamical Control Systems, Springer-Verlag, New York, 1990.[37] P.J. Olver, Applications <strong>of</strong> Lie Groups to Differential Equations, Springer-Verlag, secondedition, 1993.[38] R. Ortega, A. Loria, P.J. Nicklasson & H. Sira-Ramirez, Passivity-based Control <strong>of</strong> Euler-Lagrange Systems, Springer-Verlag, London, 1998.[39] R. Ortega, A.J. van der Schaft, B.M. Maschke & G. Escobar, “Interconnection <strong>and</strong>damping assignment passivity-based <strong>control</strong> <strong>of</strong> port-<strong>control</strong>led <strong>Hamiltonian</strong> <strong>systems</strong>”,Automatica, vol. 38, pp. 585–596, 2002.[40] R. Ortega, A.J. van der Schaft, I. Mareels, & B.M. Maschke, “Putting energy back in<strong>control</strong>”, Control Systems Magazine, 21, pp. 18–33, 2001.[41] H. M. Paynter, Analysis <strong>and</strong> design <strong>of</strong> engineering <strong>systems</strong>, M.I.T. Press, MA, 1960.[42] H. Rodriguez, A.J. van der Schaft, R. Ortega, “On stabilization <strong>of</strong> nonlinear distributedparameter port-<strong>control</strong>led <strong>Hamiltonian</strong> <strong>systems</strong> via energy-shaping”, Proc. 40th IEEEConf. on Decision <strong>and</strong> Control, Orl<strong>and</strong>o, Florida, pp.131–136, 2001.[43] A.J. van der Schaft, System theoretic properties <strong>of</strong> physical <strong>systems</strong>, CWI Tract 3, CWI,Amsterdam, 1984.[44] A.J. van der Schaft, “Stabilization <strong>of</strong> <strong>Hamiltonian</strong> <strong>systems</strong>”, Nonl. An. Th. Math. Appl.,10, pp. 1021-1035, 1986.[45] A.J. van der Schaft, “Implicit <strong>Hamiltonian</strong> <strong>systems</strong> with symmetry”,Rep. Math. Phys.,41, pp. 203–221, 1998.[46] A.J. van der Schaft, “Interconnection <strong>and</strong> geometry”, in The Mathematics <strong>of</strong> Systems<strong>and</strong> Control, From Intelligent Control to Behavioral Systems (eds. J.W. Polderman, H.L.Trentelman), Groningen, 1999.[47] A.J. van der Schaft, L 2 -Gain <strong>and</strong> Passivity Techniques in Nonlinear Control, 2nd revised<strong>and</strong> enlarged edition, Springer-Verlag, Springer Communications <strong>and</strong> Control Engineeringseries, p. xvi+249, London, 2000 (first edition Lect. Notes in Control <strong>and</strong> Inf.Sciences, vol. 218, Springer-Verlag, Berlin, 1996).37


[48] A.J. van der Schaft, M. Dalsmo & B.M. Maschke, “Mathematical structures in the <strong>network</strong>representation <strong>of</strong> energy-conserving physical <strong>systems</strong>”, in Proc. 35th IEEE Conf.on Decision <strong>and</strong> Control, Kobe, Japan, pp. 201-206, 1996.[49] A.J. van der Schaft & B.M. Maschke, “On the <strong>Hamiltonian</strong> formulation <strong>of</strong> nonholonomicmechanical <strong>systems</strong>”, Rep. Math. Phys., 34, pp. 225-233, 1994.[50] A.J. van der Schaft & B.M. Maschke, “The <strong>Hamiltonian</strong> formulation <strong>of</strong> energy conservingphysical <strong>systems</strong> with external ports”, Archiv für Elektronik und Übertragungstechnik, 49,pp. 362-371, 1995.[51] A.J. van der Schaft & B.M. Maschke, “Mathematical <strong>modeling</strong> <strong>of</strong> constrained <strong>Hamiltonian</strong><strong>systems</strong>”, in Proc. 3rd IFAC NOLCOS ’95, Tahoe City, CA, pp. 678-683, 1995.[52] A.J. van der Schaft & B.M. Maschke, “Interconnected Mechanica5a Systems, Part I: Geometry<strong>of</strong> Interconnection <strong>and</strong> implicit <strong>Hamiltonian</strong> Systems”, in Modelling <strong>and</strong> Control<strong>of</strong> Mechanical Systems, (eds. A. Astolfi, D.J.N. Limebeer, C. Melchiorri, A. Tornambe,R.B. Vinter), pp. 1-15, Imperial College Press, London, 1997.[53] A.J. van der Schaft, B.M. Maschke, “Fluid dynamical <strong>systems</strong> as <strong>Hamiltonian</strong> boundary<strong>control</strong> <strong>systems</strong>”, Proc. 40th IEEE Conf. on Decision <strong>and</strong> Control, Orl<strong>and</strong>o, Florida,pp.4497–4502, 2001.[54] A.J. van der Schaft, B.M. Maschke “<strong>Hamiltonian</strong> representation <strong>of</strong> distributed parameter<strong>systems</strong> with boundary energy flow”, Journal <strong>of</strong> Geometry <strong>and</strong> Physics, vol.42, pp.166-194, 2002.[55] K. Schlacher, A. Kugi, “Control <strong>of</strong> mechanical structures by piezoelectric actuators<strong>and</strong> sensors”. In Stability <strong>and</strong> Stabilization <strong>of</strong> Nonlinear Systems, eds. D. Aeyels, F.Lamnabhi-Lagarrigue, A.J. van der Schaft, Lecture Notes in Control <strong>and</strong> InformationSciences, vol. 246, pp. 275-292, Springer-Verlag, London, 1999.[56] S. Stramigioli, From Differentiable Manifolds to Interactive Robot Control, PhD Dissertation,University <strong>of</strong> Delft, Dec. 1998.[57] S. Stramigioli, B.M. Maschke & A.J. van der Schaft, “Passive output feedback <strong>and</strong> portinterconnection”, in Proc. 4th IFAC NOLCOS, Enschede, pp. 613-618, 1998.[58] S. Stramigioli, B.M. Maschke, C. Bidard, “A <strong>Hamiltonian</strong> formulation <strong>of</strong> the dynamics<strong>of</strong> spatial mechanism using Lie groups <strong>and</strong> screw theory”, to appear in Proc. SymposiumCommemorating the Legacy, Work <strong>and</strong> Life <strong>of</strong> Sir R.S. Ball, J. Duffy <strong>and</strong> H. Lipkinorganizers, July 9-11, 2000, University <strong>of</strong> Cambridge, Trinity College, Cambridge, U.K..[59] M. Takegaki & S. Arimoto, “A new feedback method for dynamic <strong>control</strong> <strong>of</strong> manipulators”,Trans. ASME, J. Dyn. Systems, Meas. Control, 103, pp. 119-125, 1981.[60] A. Weinstein, “The local structure <strong>of</strong> Poisson manifolds”, J. Differential Geometry, 18,pp. 523-557, 1983.[61] J.C. Willems, “Dissipative dynamical <strong>systems</strong> - Part I: General Theory”, Archive forRational Mechanics <strong>and</strong> Analysis, 45, pp. 321-351, 1972.38

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!