On Quadratic Differential Forms
On Quadratic Differential Forms
Abstract. This paper develops a theory around the notion of quadratic differential forms in
the context of linear differential systems. In many applications, we need to not only understand
the behavior of the system variables but also the behavior of certain functionals of these variables.
The obvious cases where such functionals are important are in Lyapunov theory and in LQ and H
optimal control. With some exceptions, these theories have almost invariably concentrated on first
order models and state representations. In this paper, we develop a theory for linear time-invariant
differential systems and quadratic functionals. We argue that in the context of systems described by
one-variable polynomial matrices, the appropriate tool to express quadratic functionals of the system
variables are two-variable polynomial matrices. The main achievement of this paper is a description
of the interaction of one- and two-variable polynomial matrices for the analysis of functionals and
for the application of higher order Lyapunov functionals.
Key words. quadratic differential forms, linear systems, polynomial matrices, two-variable poly-
nomial matrices, Lyapunov theory, positivity, spectral factorization, dissipativeness, storage functions
AMS subject classifications. 93A10, 93A30, 93D05, 93D20, 93D30, 93C05, 93C45
PII. S0363012996303062
system variables and their derivatives. We shall see that one-variable polynomials
are the appropriate tool in which to parametrize the model (see also, among others,
[16], [17]) and two-variable polynomials are the appropriate tool for parametrizing
the functionals. Thus, the paper presents an interesting interplay between one- and
two-variable polynomial matrices. Two-variable polynomials turn out to be a very
effective tool for analyzing linear systems with quadratic functionals.
This paper consists of a series of general concepts and questions, combined with
some specific results concerned with Lyapunov stability and with dissipativity, i.e.,
with positivity of (integrals of) quadratic differential forms. As such, the paper aims
at making a contribution to the development of the very useful and subtle notions of
dissipative and lossless (conservative) systems.
In companion papers, these ideas will be applied to LQ and H problems. The
main achievement of this paperthe interaction of one- and two-variable polynomial
matrices for the analysis of functionals and application in higher order Lyapunov
functionsappears to be new. However, seeds of this have appeared previously in the
literature. We mention especially Brocketts early work on path integrals [7], [8] in
addition to classical work on RouthHurwitz-type conditions (see, for example, [6]),
and early work by Kalman [13], [14].
2. Review. In order to make this paper reasonably self-contained, we first in-
troduce some notation and some basic facts from the behavioral approach to linear
dynamical systems. References in which more details can be found include [31], [32],
and [33].
We will deal exclusively with continuous-time real linear time-invariant differential
dynamical systems. Thus, the time axis is R, the signal space is Rq (the number of
variables q, of course, depends on the case at hand), and the behavior B is the solution
set of a system of linear constant coefficient differential equations
d
(2.1) R w=0
dt
in the real variables w1 , w2 , , wq , arranged as the column vector w; R is a real
polynomial matrix with, of course, q columns. The number of rows of R depends,
as do its coefficients, on the particular dynamical system described by (2.1). Hence
we denote this as R Rq [], where denotes the indeterminate. Thus, if R() =
R0 + R1 + + RN N , then (2.1) denotes the system of differential equations
dw dN w
(2.2) R 0 w + R1 + + RN N = 0.
dt dt
For the behavior, i.e., for the solution set of (2.1) or (2.2), it is usually advisable to
consider locally integrable ws as candidate solutions and to interpret the differential
equation in the sense of distributions. However, it is our explicit intention to avoid
mathematical technicalities as much as possible in this paper. In keeping with this,
we assume that the solution set consists of infinitely differentiable functions, even
though many of the results are valid without this assumption. Hence the behavior of
(2.1) is defined as
q d
(2.3) B = w C (R, R ) | R w=0 .
dt
We denote the family of dynamical systems obtained this way by Lq . Hence elements
of Lq are dynamical systems = (R, Rq , B) with time axis R, signal space Rq , and
ON QUADRATIC DIFFERENTIAL FORMS 1705
Here n denotes the dimension of the vector x. The important feature of (2.8) is
that it is an (implicit) differential equation containing derivatives of order at most
one in x and zero in w. We call a state representation state minimal if among all
state representations of B, n is as small as possible. It is possible to prove that
(2.8) is state minimal iff it is state trim (meaning that for all a Rn there exists
(w, x) C (R, Rq+n ) such that x(0) = a) and observable. The dimension of the state
space of a state minimal representation of B Lq is called the McMillan degree of B.
The notion of McMillan degree usually refers to properties of polynomial matrices.
Actually, for the case at hand this correspondence holds in terms of full row rank
kernel representation matrices R or observable image representation matrices M , but
we do not need this correspondence in this paper.
Every system B Lq also admits an input/output representation. By reordering
the components of the vector w, if need be, we can decompose w into
u
(2.9) w=
y
iff LU 1 is biproper.
1708 J. C. WILLEMS AND H. L. TRENTELMAN
The sum in (3.1) ranges over the nonnegative integers and is assumed to be finite,
and k` Rq1 q2 . Such a induces a bilinear differential form (BLDF), that is, the
map
(3.2) L : C (R, Rq1 ) C (R, Rq2 ) C (R, R)
defined by
X dk v T `
dw
(3.3) (L (v, w))(t) := (t) k` (t) .
dtk dt`
k,`
Note that, although is an infinite matrix, all but a finite number of its elements are
zero. We can factor as = N T M , with N and M infinite matrices having a finite
number of rows and all but a finite number of elements equal to zero. This decompo-
sition leads, after premultiplication by [Iq1 Iq1 Iq1 2 ] and postmultiplication by
col[Iq2 Iq2 Iq2 2 ], to the following factorization of :
This decomposition is not unique, but if we take N and M surjective, then their
number of rows is equal to the rank of . The factorization (3.9) is then called a
canonical factorization of . Associated with (3.9), we obtain the following expression
for the BLDF L :
T
d d
(3.10) L (w1 , w2 ) = N w1 M w2 .
dt dt
Also, this decomposition is not unique but if we take M surjective, then M is unique.
We denote this M as and the resulting pair (r , r+ ) by ( , + ). This pair is
called the inertia of . The resulting factorization
(3.12) (, ) = M T () M ()
Note that, among other things, this allows one to associate a differential operator
d d
( dt , dt ) with a QDFthis is one of the key ingredients in LQand variational
problems.
Introduce the star operator ? acting on matrix polynomials by
?
: Rq1 q2 [] Rq2 q1 []; R? () := RT ().
d d
The importance of this operation stems from the fact that M ( dt ) and M ? ( dt ) are
qq
formal adjoints as differential operators. A polynomial matrix M R [] is called
ON QUADRATIC DIFFERENTIAL FORMS 1711
defined as
Z Z +
(3.20) L (v, w) := L (v, w)dt.
R
The notation Q follows readily from this. Furthermore, consider the same integral
over a finite interval [t1 , t2 ]
Z t2
(3.21) L (v, w)dt
t1
R t2
denoted as t1 L . We call this integral independent of path if for any t1 and t2
the result of the integral (3.21) depends only on the values of v and w and (a finite
number of) their derivatives at t = t1 and t = t2 but not on the intermediate path
used to connect these endpoints, assuming, of course that v C (R, Rq1 ) and w
C (R, Rq2 ). R
The questions of when the map L is zero and when path independence holds
are studied next.
Theorem 3.1. Let Rq1 q2 [, ]. Then the following statements are equiva-
lent: R Rt
1. L = 0, equivalently t12 L is independent of path.
2. There exists a Rq1 q2 [, ] such that = , equivalently, such that
d
L = dt L . Obviously is given by
(, )
(3.22) (, ) = .
+
3. = 0, i.e., (, ) = 0.
The same equivalence holds for QDFs. Simply assume Rsqq [, ] and replace
the Ls by Qs in 1 and 2.
Proof. For the proof, see the appendix.
The importance of this theorem is that condition (3) gives a very convenient way
of checking (1) or (2). Path integrals and path independence featured prominently in
Brocketts work in the sixties (see [7], [8]), and indeed some of our results can be viewed
as streamlined versions of this work. Another potentially interesting connection of
the above theorem and our paper with the existing literature is [3] where, in our
notation, (, ) = R()M () is studied, with R and M associated with a kernel
representation (2.1) and an image representation (2.4) of a controllable system. This
defines an intriguing path independent BLDF that can be associated with any
controllable B.
1712 J. C. WILLEMS AND H. L. TRENTELMAN
In this paper we also study the behavior of QDFs evaluated along a differential
behavior B Lq . In order to do so, it is convenient to introduce an equivalence
relation on both the one- and two-variable polynomial matrices modulo a given B
Lq .
B d d
Let D1 , D2 Rq []. Define (D1 = D2 ) : (D1 ( dt ) D2 ( dt ))B = 0. Let
B
1 , 2 Rsqq [, ]. Define (1 = 2 ) : (Q1 (w) = Q2 (w) for all w B). These
equivalencies are easily expressed in terms of a kernel or an image representation of
B.
Proposition 3.2. Let R Rq [] define a kernel representation of B Lq .
B
Then D1 = D2 iff
(3.23) D1 D2 = F R
B
for some F R [] and 1 = 2 iff
D1 M = D2 M
B
and 1 = 2 iff
M T ()1 (, )M () = M T ()2 (, )M ()
B B
(ii) > 0 iff there exists 0 Rsqq [, ] with = 0 and 0 (, ) = DT ()D(),
with (R, D) an observable pair.
Proof. For the proof, see the appendix.
is an observable pair. This is the case, for example, if {0} = ker(K) ker(D + DT )
ker(M ). Indeed, under this condition
Kp+ D + M 2
(D + DT )
has full column rank for all C.
State representations of autonomous systems take a very special form. Indeed, it
is easy to see that B Lq is autonomous iff it admits a state representation of the
form dxdt = Ax, w = Cx. Such state representations are automatically state trim.
If (A, C) is observable, then they are state minimal. It also follows that for every
d B
D Rq [] there exists a matrix H Rn such that D( dt )w = Hx, i.e., every
q
linear differential operator acting on an autonomous B L is B-equivalent to an
instantaneous function of the state. An analogous statement holds, of course, for
QDFs.
Viewed from this perspective, one can regard Theorem 4.3 as being about state
systems and in this sense not very different from classical Lyapunov theorems. The
point of Theorem 4.3 is twofold:
1. It avoids the state construction which algorithmically (and conceptually) is
not always easy in the multivariable case; and
2. It has the usual Lyapunov theory as a special case by applying it to systems
in first order form and using memoryless QDFs. For the sake of completeness,
we record this as a corollary.
Corollary 4.6. Let B be the behavior of dw dt = Aw. Let (, ) = 0 with
B
0 Rqq , 0 = T0 0. Then (, ) = A0 + 0 AT =: 0 . Whence, if
0 = T0 0 and if (A, 0 ) is an observable pair of matrices, B is asymptotically
stable.
Proof. For the proof, see the appendix.
In section 3, we discussed B-positive QDFs. When B is autonomous, it is useful
to consider also a stronger concept. Let B Lq and Rsqq [, ]; we call strongly
B B
B-positive (denoted 0) if 0 and if (w B and Q (w)(0) = 0) imply (w = 0).
B B B
It is easy to see that 0 implies > 0 and that in order for 0, B must be
B
autonomous. In fact, ((Q (w)(0) = 0) (w = 0)) by itself already implies ( 0
B
or 0). Using this notion we arrive at the following refinement of theorem 4.3.
B B
Proposition 4.7. If 0 and < 0, then B is asymptotically stable and
B
0.
ON QUADRATIC DIFFERENTIAL FORMS 1715
(4.5) F T () = F ().
(4.6) Y (, ) = X()
and compute
(, ) Y (, )R() RT ()Y (, )
(4.7) (, ) = .
+
B B B B
Then = . Since any two 1 , 2 such that 1 = and 2 = satisfy 1 = 2 ,
any other solutions of (4.3) and/or (4.6) yield s in (4.7) that are B-equivalent.
Proof. For the proof, see the appendix.
Theorem 4.8 is more than a mouthful and so we illustrate it for ordinary state
space systems dw T
dt = Aw. Let = R
nn
. Then (4.3) becomes
This equation has a constant solution which must be symmetric, X0 = X0T , the
solution of the ordinary Lyapunov equation
(4.9) X0 A + AT0 X = .
1 R()R() R()R()
(4.12) B(, ) =
2 +
as a Lyapunov function. Note that this Lyapunov function can be written directly
from the system parameters, without having to solve linear equations! This fact is
actually well known, even though it is not presented in the vein of providing a higher
order Lyapunov function ([6], [13], [14]).
The two-variable polynomial B defined by (4.12) is called the Bezoutian of R.
B
Note that B (, ) = 12 R()R(); B is R-canonical, but B is not. However, B
is B-equivalent to 12 R()R() + 12 R()R(), which is. If we take this for the
in Theorem 4.8, then the Lyapunov equation yields X = 0. Taking Y = 0 then
also yields the Bezoutian (4.12) as the corresponding (hence R-canonical) Lyapunov
function B.
A close examination of the arguments involved yields the equivalence of the follow-
ing three conditions on a polynomial R of degree n and the corresponding B R[, ]
given by (4.12):
1. R is Hurwitz,
2. B 0 and (R, R? ) is coprime,
3. B (the constant matrix associated with B) has rank n and is 0.
The Lyapunov function (4.12), the Bezoutian, is a very useful one for deriving
various stability tests. It is a classical concept in stability (see [11] for a recent
reference). Let us illustrate its usefulness by deriving the Routh stability test from it.
1718 J. C. WILLEMS AND H. L. TRENTELMAN
Let R R[] be a polynomial of degree n. Decompose R in its even and odd parts as
R() = E0 ( 2 ) + E1 ( 2 ).
Assume for simplicity that R(0) = E0 (0) 0. Rouths stability criterion states that
R is Hurwitz iff all elements of the Routh array E0 (0), E1 (0), . . . , En (0) are positive.
Define Rk () = Ek1 ( 2 ) + Ek ( 2 ) for k = 1, . . . , n, and let Bk be the Bezoutian
associated with Rk . Examining expression (4.12) yields, after a simple calculation
for k = 1, . . . , n (define Bn+1 = 0). Assume that E0 (0), E1 (0), . . . , En (0) are all
positive. Then we obtain (note that B = B1 )
n
X
B(, ) = k k1 k1 Ek ( 2 )Ek ( 2 ),
k=1
where k = Ek1 (0)/E1 (0)E2 (0) . . . Ek (0). Obviously B 0 and has rank n. There-
fore R is Hurwitz. To show the converse, assume that R is Hurwitz. Then E0 (0) > 0.
Also, B 0 and has rank n. Therefore, by (4.13), B2 0 and has rank n 1. Hence
R2 is Hurwitz, and E1 (0) > 0. Now proceed by induction.
The key point thus is that
n
X 2 k1 2
d d
k Ek w
dt2 dtk1
k=1
is a QDF which is well defined and nonnegative definite when the Routh conditions
are satisfied. It has derivative
2 2 !
1 d d
R w R w ,
2 dt dt
d
which is obviously nonnegative definite along solutions of R( dt )w = 0.
Example 4.14. Let E1 , E2 , . . . , EN and O1 , O2 , . . . , ON 0 be two sets of real polyno-
mials, and assume that Rk,` () := Ek ( 2 ) + O` ( 2 ) is Hurwitz for all k = 1, 2, . . . , N
and ` = 1, 2, . . . , N 0 . Then any combination
0
N
X N
X
2
R() = k Ek ( ) + ` O` ( 2 )
k=1 `=1
is also Hurwitz whenever all the k s and ` s are positive. In order to see this, simply
observe that, in the obvious notation, (4.12) yields
0
N X
X N
B(, ) = k ` Bk,` (, )
k=1 `=1
Average nonnegativity states that the net flow of energy going into the system
is nonnegative: the system dissipates energy. Of course, sometimes energy flows into
the system, while at other times it flows out of it. This outflow is due to the fact
that energy is stored. However, because of dissipation, the rate of increase of storage
cannot exceed the supply. This interaction between supply, storage, and dissipation
is now formalized.
Definition 5.3. Let Rsqq [, ] induce the QDF Q . The QDF Q induced
by Rsqq [, ] is said to be a storage function for if
d
(5.1) Q Q .
dt
A QDF Q induced by Rsqq [, ] is said to be a dissipation function for if
Z Z
(5.2) 0 and Q = Q .
The next proposition shows that one can always interpret average positivity by an
instantaneous positivity condition involving the difference between the rate of change
of storage function and the supply rate.
Proposition
R 5.4. The following conditions are equivalent:
1. Q 0,
2. admits a storage function,
3. admits a dissipation function.
Moreover, there is a one-one relation between storage and dissipation functions,
and , respectively, defined by
d
Q (w) = Q (w) Q (w)
dt
equivalently, = , i.e.,
(, ) (, )
(5.3) (, ) = .
+
(5.4) (, ) = M T ()M M ()
discussed earlier in section 3 allows us to introduce a state for the QDF Q . Indeed,
(5.4) induces the dynamical system in image representation
d
(5.5) v=M w.
dt
Note that in (5.5) we are considering w as the latent variable and v as the manifest
one. This is in keeping with the idea that v T M v, the supply rate, is the variable
of interest and that w is a latent variable that explains it. We are hence considering
the behavior of the possible trajectories v. Assume that M has r rows, i.e., that
ON QUADRATIC DIFFERENTIAL FORMS 1721
d
M Rrq []. Thus, (5.5) defines a system B Lr with B = im(M ( dt )) and
d q r
M ( dt ) viewed as a map from C (R, R ) to C (R, R ). Hence this system has a state
representation. Assume that
d
(5.6) x=X w
dt
where, as always, kak2A denotes aT Aa. Note that the factorization (5.4) is not unique.
However, any such factorization is related in a simple way to a canonical one, say, to
(5.9) (, ) = M T () M ()
Thus, considering arbitrary (i.e., not necessarily canonical) factorizations and arbi-
trary (i.e., not necessarily minimal) state representations yields a (rather than the)
state of Q . Thus, the situation with the state is similar to the situation with the
state of a system B Lq .
We have the following R important result.
Theorem 5.5. Let Q 0, and let Rsqq [, ] be a storage function for ,
i.e., . Then is a state function. Let Rsqq [, ] be a dissipation function
for . Then is a state/supply function. In fact, if X Rq [] is a state map for
, then there exist real symmetric matrices P and E such that
Equivalently
T
d d
L (w1 , w2 ) = X w1 PX w2 ,
dt dt
d
T d
M ( dt )w1 M ( dt )w2
L (w1 , w2 ) = d E d
X( dt )w1 X( dt )w2
(5.14) + .
R per
If Q > 0 then and + may be constructed as follows. Let () = H T ()H()
and () = AT ()A() be, respectively, Hurwitz and anti-Hurwitz factorizations of
. Then
(, ) AT ()A()
(5.15) + (, ) =
+
and
(, ) H T ()H()
(5.16) (, ) = .
+
holds for all w B D(R, Rq ). Simply construct an image representation for B, say,
d
(5.18) w=M `.
dt
Upon substituting (5.18) in (5.17), we see that the issue then becomes one of studying
when
Z + Z +
d d
Q M ` dt 0 or Q M ` dt = 0
dt dt
d
for all ` D(R, R ). Since obviously Q (M ( dt )`) = Q0 (`) with
0 (, ) := M T ()(, )M (),
the problem reduces to studying 0 . For example, the existence of a storage function is
established as follows. Without loss of generality, take (5.18) to be an observable image
representation. Then M has a polynomial left inverse M . By Proposition 5.4, there
d
exists 0 such that dt Q0 (`) Q0 (`) for all ` D(R, R ). Now define (, ) :=
T 0 d d
M () (, )M (). Then for w = M ( dt )` we have Q (w) = Q0 (M ( dt )w) =
d
Q0 (`) and Q (w) = Q0 (`), so we obtain dt Q (w) Q (w).
The case that B Lq is neither controllable nor autonomous will be studied
in a later publication. The next comment is relevant to the question of what the
appropriate definition of dissipativity is in that case.
Remark 5.9. Finding an appropriate definition of a dissipative system is an issue
that has attracted considerable attention (see [29], [12], [24], [27]). Of course, this is
at the root of the issues discussed in the present article. Let B Lq . There are many
examples where the instantaneous rate of supply (say, of energy) into the system is
given, not by a static function of the external variables, but by a QDF, Q (w). The
study of supply rates that are themselves dynamic is one of the novel aspects of the
present paper. When would one want to call B dissipative with respect to Q (w)?
Lossless? Conservative? When would one want to say that B absorbs some R of the
supply? The definitions of average nonnegativity for dissipativeness, and Q = 0
for losslessness (= conservativeness), are fully adequate provided that B is controllable
(see Remark 5.8). However, Proposition 5.4 points to another definition which does
not need controllability and which, in the controllable case, reduces to it. Thus, we
arrive at the following definition as the most general: B Lq is said to be dissipative
d
with respect to the supply rate Q if there exists a Q such that dt Q (w) Q (w) for
all w B, and lossless, or conservative, if this holds with equality. The unfortunate
aspect of this definition is its existential nature it shares this notorious feature with
the first and second law of thermodynamics. It does not seem an easy matter in the
noncontrollable case to reduce this to a statement involving only Q , and without
1724 J. C. WILLEMS AND H. L. TRENTELMAN
(, ) = U T ()0 (, )U (),
this A and H. Note that in the lossless case ( = 0) this yields + = , whence
the uniqueness of .
Remark 5.14. It is easy to deduce from the proof of Theorem 5.7 that and +
d
have the following interpretations. Let x = X( dt )w be the state (see 5.6). Let a Rn .
q d
Consider all w D(R, R ) such that (X( dt )w)(0) = a. Denote this set by Ba . By
d
Theorem 5.5 we know that Q is a state function, say, Q (w) = kX( dt )wkK ,
nn
for some symmetric matrix K R . Hence for w Ba we have Q (w)(0) =
aT K a. Similarly Q+ (w)(0) = aT K+ a for some symmetric matrix K+ . Then it can
be shown that
Z +
(5.19) aT K a = sup Q (w)dt
wBa 0
and
Z 0
T
(5.20) a K+ a = inf Q (w)dt .
wBa
For this reason, Q (w)(0) is called the available storage and Q+ (w)(0) the required
supply at t = 0 due to w. In this inf and sup, one keeps the past, respectively, the
future of w fixed.
R +
6. Half-line positivity. In section 5, we studied QDFs for which Q (w)dt
0. The intuitive idea was that this expresses that the net supply (of energy)
is directed into the system: energy is being absorbed and dissipated in the system.
There are, however,
Rt situations where at any moment in time the system has absorbed
energy, i.e., Q (w)( )d 0 for all t R. For example, electrical circuits and
mechanical devices at rest are in a state of minimum energy, and therefore the energy
delivered up to any time is nonnegative. This type of positivity is studied in this
section. It plays a crucial role in H problems.
Definition 6.1. Let Rsqq [, ]. The QDF Q (or simply ) is said to
Rt R0
be half-line nonnegative, denoted by Q 0, if Q (w)dt 0 for all w
Rt R0
D(R, Rq ), and half-line positive, denoted Q > 0, if in addition Q (w)dt = 0
implies w(t) = 0 for t 0.
Note that half-line nonnegativity implies average nonnegativity, and that half-line
positivity implies average positivity.
Write (, ) = M T ()M M () and partition M conform M as
P
(6.1) M=
N
d d
so that (, ) = P T ()P ()N T ()N () and hence Q (w) = kP ( dt )wk2 kN ( dt )wk2 .
In the following, for C, let denote its complex conjugate.
Proposition 6.2. Let Rsqq [, ]. Then
Rt
(i) ( Q 0) ((, ) 0 C, <e() 0)
Rt
(ii) ( Q > 0) ((, ) 0 C, <e() 0 and det() 6= 0).
Proof. For the proof, see the appendix.
As noted before, it immediately follows from the definitions that half-line nonneg-
ativity implies average nonnegativity, etc. Thus, Proposition 5.4 implies the existence
of a storage function. It is the nonnegativity of the storage function that allows us to
conclude the half-line positivity.
Theorem 6.3. Let Rsqq [, ]. Then the following statements are equivalent.
1726 J. C. WILLEMS AND H. L. TRENTELMAN
Rt
1. Q 0,
2. there exists a storage function 0 for ,
3. admits a storage function, and the storage function + defined in Theorem
5.7 satisfies + 0.
Proof. For the proof, see the appendix.
In order to check half-line nonnegativity, one could thus in principle proceed as
follows. Verify that (i, i) 0 for all R, compute + , and check whether
+ 0. In some situations, it is actually possible to verify this condition in a more
immediate fashion; for example, when (, ) = 0 , a constant matrix, with 0 0
(trivial, but that is the case that occurs in standard LQ theory!), or when in (6.1)
P is square and det(P ) 6= 0. Then, under the assumption that a storage function
exists (equivalently: N T (i)N (i) P T (i)P (i) for all R), all storage
functions are actually nonnegative if one of them is nonnegative. In fact, in this case
the following theorem holds.
Theorem 6.4. Let Rsqq [, ]. Assume it is factored as (, ) = P T ()P ()
N ()N () with P square and det(P ) 6= 0. Let X Rq [] be a minimal state
T
map for the B given in image representation by (6.1). The following statements are
equivalent:Rt
1. Q 0,
2. (, ) 0 for all C, <e() 0,
3. N P 1 has no poles in <e() 0 and (i, i) 0 for all R,
4. there exists a storage function 0 for ,
5. there exists a storage function for and every storage function for
satisfies 0,
d
6. there exists a real symmetric matrix K > 0 such that QK (w) := kX( dt )wk2K
is a storage function for ,
7. there exists a storage function for and every real symmetric matrix K such
d
that QK (w) := kX( dt )wk2K is a storage function for satisfies K > 0.
P
Furthermore, if [ N ] is observable, then any of the above statements is equivalent with
30 . P is Hurwitz and (i, i) 0 for all R.
Proof. For the proof, see the appendix.
7. Observability. One of the noticeable features of QDFs is that a number of
interesting systems theory concepts generalize very nicely to QDFs. We have already
seen that the state of a symmetric canonical factorization of functions as the state
of the QDF Q . In this section we introduce observability of a QDF. In a later section
we will discuss duality of QDFs.
For Rq1 q2 [, ] and w1 C (R, Rq1 ) fixed, the linear map w2 7 L (w1 , w2 )
is denoted by L (w1 , ). For w2 C (R, Rq2 ) fixed, the linear map w1 7 L (w1 , w2 )
is denoted by L (, w2 ). The BLDF is called observable if L (w1 , ) and L (, w2 )
determine w1 and w2 uniquely. Equivalently we have the following.
Definition 7.1. Let Rq1 q2 [, ]. We call observable if, for all w1
C (R, Rq1 ) and for all w2 C (R, Rq2 ), we have
L (w1 , ) = 0 w1 = 0
and
L (, w2 ) = 0 w2 = 0.
The following theorem gives necessary and sufficient conditions for observability
purely in terms of the two-variable polynomial matrix and in terms of the (one-
ON QUADRATIC DIFFERENTIAL FORMS 1727
q
for all w D(R, R ). Note that (because of observability) strict positivity implies
positivity, and similarly for the other cases.
These notions of strict positivity involve || which may be difficult to evaluate.
However, it is possible to relate it to any canonical factorization of . This is stated in
the next proposition. For simplicity we state only the case of strict average positivity.
However, completely analogous statements hold for simple strict positivity or for strict
half-line positivity.
1728 J. C. WILLEMS AND H. L. TRENTELMAN
In order to define the matrix T in the general case, we need to take into account
the algebraic multiplicities of the roots i .
Definition 9.2 (general case). Let Rsqq [, ] be observable, det() 6= 0,
and assume that det() has no roots on the imaginary axis. Let 1 , 2 , . . . , k C be
the distinct roots of det() with positive real part, and denote by ni the multiplicity
ON QUADRATIC DIFFERENTIAL FORMS 1729
i 1
nX
j
()(j`) (i )ai,j = 0, (` = 0, 1, . . . , ni 1).
`
j=`
Also, define i,j Cni qnj q by defining its (r, s)th block to be the q q matrix
where (k,`) means taking the kth partial derivative with respect to and the `th with
respect to .
Then we define the Pick matrix of as the matrix T whose (i, j)th block is given
by Ti,j Cni nj , with
1
Ti,j := AT i,j Aj .
i + j i
Pk
Note that the sum i=1 ni of the multiplicities is equal to n := 12 deg det(),
and that T is a complex Hermitian matrix of size n n.
The next theorem is the most refined result of this paper. It shows, on the
one hand, the relation between strict half-line positivity and positivity of a storage
function, and, on the other hand, the relation with the positivity of the Pick matrix
T .
We have seen in Theorem 5.5 that a storage function is a quadratic state function,
d
i.e., Q (w) is of the form xT Kx, K = K T , with x = X( dt )w a minimal state map
for . We call this state function positive definite if K > 0.
Theorem 9.3. Let Rsqq [, ] be observable. The following are equivalent:
Rt
1. QR 0,
2. (a) Q 0,
(b) there exists a storage function that is a positive definite state function,
3. (a) > 0 such that (i, i) ||(i, i) for all R,
(b) T > 0.
Proof. For the proof, see the appendix.
Remark 9.4. It follows from the proof of Theorem 9.3 that half-line nonnegativity
implies that the Pick-type matrix T (see (9.1)) is 0 whenever any set of i s in
the right half of the complex plane and any set of ai s are chosen. It is possible to
prove that if T is nonnegative definite for any such choice for the i s and ai s, then
we have half-line nonnegativity. The remarkable thing about Theorem 9.3 is that it
suffices to evaluate T at the set of special i s and ai s obtained from the singularities
of .
1730 J. C. WILLEMS AND H. L. TRENTELMAN
dx
(10.1) E + F x + Gw = 0,
dt
dz
(10.2) E0 + F 0 z + G0 v = 0,
dt
having the property that for all (w, x) C (R, Rq+n ) satisfying (10.1), and for all
(v, z) C (R, Rq+n ) satisfying (10.2) there holds the following kind of duality in-
volving the state
d T
(10.3) z x = v T w.
dt
In fact, this is an immediate consequence of the following.
d d
Proposition 10.1. Let R( dt )w = 0 and w = M ( dt )` be a minimal kernel
representation and an observable image representation, respectively, of the controllable
system B Lq . Assume that X Rn [] defines a minimal state map for B, i.e.,
ON QUADRATIC DIFFERENTIAL FORMS 1731
d
x = X( dt )` defines a minimal state of B. Then there exists a Z Rn [] defining
d
a minimal state map Z( dt ) for B , such that for all `, `0 C (R, R ), we have
T T
d d d d d
(10.4) Z `0 X ` = RT `0 M `.
dt dt dt dt dt
If we define (, ) := Z T ()X() and (, ) := R()M (), then (10.4) is equiva-
lent to = .
Proof. For the proof, see the appendix.
If a pair of minimal state maps (X, Z) of B and B satisfies (10.4), then we call
it a matched pair of state maps.
We now associate with a QDF a dual one and relate their average nonnegativity
and average positivity. Let Rsqq [, ] and let (, ) = M T () M () be a
symmetric canonical factorization, with
Ir+ 0
= .
0 Ir
Let us assume that M Rrq []. Partition M conformably to as
P
M=
N
so that is written as
(10.5) (, ) = P T ()P () N T ()N ().
Consider the dynamical system B Lr with image representation
d
(10.6) w=M `.
dt
There are a number of integers associated with M that are of interest to us:
(10.7) r+ = rowdim(P ),
(10.8) r = rowdim(N ),
(10.9) m = rank(M ) .
The number r+ corresponds to the number of positive squares in Q , r to the num-
ber of negative squares, while m equals the number of inputs in any input/output
or input/state/output representation of B. Since it is defined by an image repre-
sentation, B is a controllable system and, as such, it admits a dual, B Lr . Let
d
R( dt )w = 0 be a minimal kernel representation of B. Then
d
(10.10) v = RT `0
dt
is an observable image representation for B . Let 0 (, ) := R() RT (). Note
d d
that the QDFs Q (`) = (M ( dt )`)T M ( dt )` and
T
d d
Q0 (`0 ) = RT `0 R T `0
dt dt
are in a sense also dual. Their positivity properties are very much related, as shown
in the following theorem.
Theorem 10.2. Assume that r+ = m. Then
1732 J. C. WILLEMS AND H. L. TRENTELMAN
R R
(i) R Q 0 R Q0 0,
(ii) Q > 0 Q0 < 0, R R
(iii) (assume is observable) Q 0 Q0 0,
R Z) be a matched pair of minimal state maps for B and B . Assume
(iv) Let (X,
that Q 0, and let define a storage function for . By Theorem 5.5,
is a state function, i.e., there exists a real symmetric matrix K such that
T
d d
(10.11) Q (`) = X ` KX `.
dt dt
R 0 .
is a storage function for Q R R
(v) (assume is observable) t Q 0 t Q0 0. Here t Q0 0 is
R
defined as the property that there exists > 0 such that 0 Q0 (w)dt
R
0 Q0 | (w)dt for all w D(R, Rq ) (i.e., half-line positivity over the posi-
tive half-line).
Proof. For the proof, see the appendix.
We close this section by pointing out that it is of interest to generalize the notion
of duality by using, instead of the usual inner product, an inner product that is itself
induced by a QDF. These ramifications are a matter of future research.
11. Conclusions. In this paper we studied two-variable polynomial matrices
and their role in a number of problems in linear system theory. The basic premise set
forward is the following. Dynamic models lead naturally to the study of one-variable
polynomial matrices. By substituting the time derivative for the indeterminate, and
by letting the resulting differential operator act on a variable, one arrives at a dynam-
ical system, which may then be in kernel or in image representation. The study of
quadratic functionals in a variable and its derivative, on the other hand, leads to two-
variable polynomial matrices. Important instances where dynamical systems occur in
conjunction with functionals are, for example, Lyapunov theory, the theory of dissi-
pative systems, and LQ and H control. We developed the former two applications
in the present paper. The latter two will be discussed elsewhere.
Appendix.
Proof of Theorem 3.1. We prove the equivalence of the two statements in (1) at
the end of the proof and proceedR with the first statement by
R running the circle (1)
(3) (2) (1). Assume that L = 0. Then obviously L (v, w)dt = 0 for all
v D(R, Cq1 ) and w D(R, Cq2 ), with L (v, w) in this case (for complex functions)
P k `
defined by k,` ( ddtkv )T k,` ( ddtw` ). Then
Z
v T (i)(i, i)w(i)d = 0
for all v L2 (C, Cq1 ), w L2 (C, Cq2 ) that are Fourier transforms of v D(R, Cq1 ),
and w D(R, Cq2 ). This implies that = 0. Assume to the contrary that there
exist 0 R, a Cq1 , b Cq2 such that aT (i0 , i0 )b 6= 0. Define vN D(R, Cq1 )
ON QUADRATIC DIFFERENTIAL FORMS 1733
for N = 1, 2, . . . , by
2N
ei0 t a |t| 0 ,
2N
(A.1) vN (t) = v(t + 0 ) t < 2N
0 ,
2N 2N
v(t 0 ) t> 0 .
Proof. The if part is obvious. To show the only if part, let r1 have degree n1
and r2 have degree n2 , and assume that they are monic. Consider the term k,` k ` of
k `
(, ). In the quadratic form Q (w1 , w2 ) this term contributes k,` d w 1 d w2
. If w1
dttk dt`
d
satisfies r1 ( dt )w1 = 0, and if k n1 , then the contribution in Q (w1 , w2 ) of k,` k `
is equivalent to that of k,` ( k kn1 r1 ()) ` . Proceeding
P analogously with the `s
and the other terms shows that there exists 0 (, ) = k,` 0k,` k ` with 0k,` = 0
for k n1 or ` n2 such that
ditions can be chosen arbitrarily in the sense that for any values of wm (0), dw dt (0),
m
dnm 1 wm
. . . , dtnm 1 (0) there exist wm Bm having these initial values. It follows that
0 = 0.
Now return to the proof of the case for general q1 , q2 . Bring R1 and R2 in Smith
form, showing that it suffices to prove (A.2) for R = 1 and R2 = 2 with 1 and 2
in Smith form. Let d1 be the (k1 , k1 )th element of 1 and d2 the (k2 , k2 )th element
of 2 . Examine (A.2) and observe that we need to show that the (k1 , k2 )th element
of , k1 k2 , can be written as
Since dk and d` are Hurwitz, d` () and dk () are coprime and hence, by Bezout,
(A.10) has a solution. This then yields a solution of the matrix version.
(ii) Again use the Smith form. Obtain that the difference of two solutions must
satisfy
Hence again using coprimeness of d` () and dk (), there exists a polynomial fk` such
that xk` () = fk` ()d` (). This yields (4.4). To show (4.5), obtain
If R is square and det(R) 6= 0, (4.5) follows by pre- and postmultiplying by (RT ())1
and (R())1 .
(iii) This proof is obvious.
B B
(iv) Let w B and assume that 1 = 2 , i.e., = 0, where = 1 2 . Then
Z t Z t
d
(A.13) Q (w)dt = Q (w)dt = Q (w)(t) Q (w)(0).
0 0 dt
0 (, ) = 00 (, ) + F T (, )R() + RT ()F (, ).
Thus,
Proof of Proposition 5.2. The proof of all three statements is analogous. Therefore
we only give the proof of (i). To prove (), let w D(R, Rq ) and let w be its Fourier
transform. Observe, using Parsevals Theorem, that
Z + Z +
1
(A.14) Q (w)dt = w(i)T (i, i)w(i)d,
2
whence (). To show the converse, as in the proof of Theorem 3.1, we silently switch
from Rq as signal space to Cq . Assume that there exists a Cq and 0 R such that
aT (i0 , i0 )a < 0. Consider the function wN D(R, Cq ) for N = 1, 2, . . ., defined
R +
exactly as vN was in the proof of Theorem 3.1. Next evaluate Q (wN )dt and
observe (using the idea in the proof of Theorem 3.1) that this integral can be made
negative by taking N sufficiently large.
Proof of Proposition 5.4. We will run the circle (3) (2) (1) (3). To see
that (3) (2), assume that is a dissipation function. Then (, ) = (, ),
by Theorem 3.1. Define
(, ) (, )
(A.15) (, ) = .
+
Hence = . Use 0 to conclude that is a storage function. To see
that (2) (1), use and Theorem 3.1 to conclude (1). To see that (1) (3),
use Propositions 5.2 and 5.6 to construct a D such that (, ) = DT ()D().
Observe that (, ) := DT ()D() defines a dissipation function. The one-one
relation between and is given by (A.15).
Proof of Theorem 5.5. By (5.11) it suffices to consider minimal state representa-
tions obtained from a canonical factorization of . Let
d
v=M = w
dt
d
be obtained from such a factorization, and let x = X( dt )w be a minimal state. There
exists a permutation matrix P such that
U
PM = ,
Y
with det(U ) 6= 0 and such that Y U 1 is a matrix of proper rational functions. Denote
d d
u = U ( dt )w. Consider f = F ( dt )w, where F is an arbitrary polynomial matrix. Then
(see section 2) f is a state function, (i.e., there exists a matrix K such that f = Kx)
iff F U 1 is strictly proper and a state/input function (i.e., there exists matrices L, J
such that f = Kx + Ju) iff F U 1 is proper.
We first prove the second part of the theorem, i.e., that every dissipation function
is a state/supply function. Let (, ) = DT ()D() be a dissipation function. Then
(A.16) M T () M () = DT ()D().
Since the left-hand side is proper, so is the right-hand side. This obviously implies
d
that D()U 1 () is proper. Hence D( dt )w is a state/input function and equivalently,
1738 J. C. WILLEMS AND H. L. TRENTELMAN
( + )(, ) = M T () M () DT ()D().
By redefining
T
M () 0 M ()
M T () M () = ,
D() 0 I D()
and observing that a minimal state for the system with image representation v =
d
M ( dt )w is also a minimal state for the system with image representation
d
M ( dt )
v= d w,
D( dt )
it suffices to prove the claim in the lossless case, i.e., when = .
Assume that (, ) = M T () M () is a symmetric canonical factorization of
and that (, ) = N T () N () is a symmetric canonical factorization of .
Postmultiply the identity = by U 1 () to obtain
Assume that Lk k is the term of degree k in the polynomial part of the matrix of
rational functions N ()U 1 (). Using that the right-hand side of (A.18) is proper in
, by equating powers of yields N T () Lk = 0. Express N () as
I
I
N () = N0 N1 . . . NL .
..
I L
T
and use (A.18) to obtain N0 N1 . . . NL Lk = 0. Since the factorization
(, ) = N T () N () is canonical, N0 N1 . . . NL is surjective. Hence
(A.18) yields Lk = 0. This shows that N ()U 1 () is strictly proper and hence that
d
N ( dt )w is a state function as desired. Thus there exists a constant matrix K such
that N () = KX(). This shows that there exists a matrix P such that (5.12) holds.
This completes the proof of the theorem.
Proof of Theorem 5.7. We first prove the second part, the part regarding strong
average positivity. In this case it follows from Proposition 5.3 that (i, i) > 0
for all . Hence by Proposition 5.6, (, ) has a Hurwitz and an anti-Hurwitz
factorization. The associated storage functions, + and , satisfy
d d 2 d
(A.19) (Q+ (w) Q (w)) = kH wk kA wk2 .
dt dt dt
d
Let x = X( dt )w be a minimal state associated with a canonical factorization of .
By Theorem 5.5, there exist real symmetric matrices, say K+ and K , such that for
ON QUADRATIC DIFFERENTIAL FORMS 1739
Z 0
d
(A.20) aT K+ a aT K a = Q+ (w)(0) Q (w)(0) = kH wk2 dt,
dt
whence K+ K , so + .
d
The problem is that there may not be a solution of A( dt )w = 0 for all a such that
d
(X( dt )w)(0) = a. In order to circumvent this difficulty we first prove the statements
of the second part under the additional assumption that || for some > 0, in
addition to the assumption that (i, i) > 0 for all . Next, we modify to
such that these conditions hold for > 0, and, finally, take the limit for 0.
Assume that (i, i) > 0 for all R and || for some > 0. The
system (5.5) allows an I/O representation, in the sense that there exists a permutation
matrix P such that
d
U ( dt )
(A.21) Pv = d w,
Y ( dt )
d d
with det(U ) 6= 0 and G := Y U 1 proper. Let u = U ( dt )w, y = Y ( dt )w. There
exist constant matrices A, B, C, and D such that u, x, and y are related by dx dt =
1 d
Ax + Bu, y = Cx + Du. Since AU is biproper, A( dt )w is of the form F x + Lu with
L nonsingular. Using u = L1 F x and x(0) = a in these equations then results in a
d
solution of A( dt )w = 0. To show that AU 1 is indeed biproper, use Proposition 5.2
to obtain
T
U (i) U (i)
P P T =
Y (i) Y (i)
T
T U (i) U (i)
A (i)A(i) .
Y (i) Y (i)
Since G is proper, AU 1 is proper, by the equality on the left. The inequality on the
right gives biproperness.
Consider a general and define by = + || + I. Then satisfies the
above conditions and hence there exists (in the obvious notation) +
and such
+
that . Observe that for 0 < 1 2 there holds 1 2 and deduce
1740 J. C. WILLEMS AND H. L. TRENTELMAN
from + + +
1 1 2 that 2 1 . Similarly, 2 1 . Consequently 2
+ +
1 1 2 . Prove (using for example the associated matrix representations)
+
that this monotonicity implies the existence of lim0 +
=: 0 and lim0 =: 0 .
We now prove that + +
0 and 0 satisfy 0 and 0 , and subsequently
+
that any storage function of satisfies 0 0 . To prove the first part,
observe that +
and for > 0 and take the limit for 0. To prove
the second part, assume that . Then . Therefore .
+
HT ()H () = (, ) ( + )
(, ).
Since as 0 and 0 as 0, we also have that H converges. Clearly
the limit H0 satisfies (, ) = H0T ()H0 () and must be Hurwitz. The formula
for
+ is treated analogously.
0 (= ) follows. The situation for R
0
Proof of Proposition 6.2: (i) Compute Q (w) for w(t) = et a with <e() > 0
T
and a Cm . This integral equals a ( ,)a
+
. This w is not of compact support, but
an approximation argument can be used to complete the proof of (i). For <e() = 0
the result follows from Proposition 5.2. (ii) is proven similarly.
Proof of Theorem 6.3 : We prove that (3) (2) (1) (3). That (3) (2)
d
is trivial. In order to see that (2) (1), integrate dt Q (w) Q (w) from
to 0. We now prove that (1) (3). Assume first that satisfies the assumptions
(i, i) > 0 for all and || for some > 0. By Theorem 5.7 we then have
(, ) AT ()A()
(A.23) + (, ) = .
+
d d d
This yields dt Q+ (w) = Q (w) kA( dt )wk2 for all w. Let x = X( dt )w be a minimal
d 2
state map of . By Theorem 5.5, Q+ (w) = kX( dt )wkK+ for some real symmetric
matrix K+ . Using this expression in (A.23) and integrating from to 0 yields that,
d d
R0
for all a such that X( dt )w(0) = a and A( dt )w = 0, we have aT K+ a = Q (w)dt.
This integral is 0, so we must have aT K+ a 0 (actually, such w does not have
compact support but, by an approximation argument, the integral cannot be < 0).
As in the proof of Theorem 5.7, it can be shown that for any initial condition a such
w exists. This proves that + 0. Take a general . As in the proof of Theorem 5.7,
first replace by . By applying the previous to , we can conclude that (in the
obvious notation) + 0. Then take the limit for 0.
Proof of Theorem 6.4. We will first run the circle (1) (2) (3) (7) (4)
(1).
(1) (2). This was proven in Proposition 6.2.
(2) (3). We have P T ()P () N T ()N () for C, <e() 0. Assume
that N P 1 has a pole such that <e() 0. Then there exists a vector v 6= 0 such
that P ()v = 0 while N ()v 6= 0. This, however, contradicts the above inequality.
ON QUADRATIC DIFFERENTIAL FORMS 1741
d
(3) (7). Let QK (w) = kX( dt )wk2K be a storage function. We want to show
that K > 0. First we show that N P 1 is a proper rational matrix. We have
P T (i)P (i) N T (i)N (i) 0. Define G := N P 1 . Then for all such that
P (i) is nonsingular, we have GT (i)G(i) I, which shows that G is proper.
There exist matrices A, B, C, and D such that all smooth x, w1 and w2 satisfying
d d
x = Ax + Bw2 , w1 = Cx + Dw2 can be written as x = X( dt )w, w1 = N ( dt )w, and
d 1
w2 = P ( dt )w for some w C (R, R ). Since N P has no poles in <e() 0, the
matrix A can be chosen such that its eigenvalues are in the open left half of the complex
plane. Moreover, we may assume that the pair (C, A) is observable. Let a Rn .
d
Choose w2 = 0, let x satisfy dt x = Ax, x(0) = a, and let w1 = Cx. This shows that
d d
there exists w C (R, R ) such that (X( dt )w)(0) = a and w2 = P ( dt )w = 0. Also,
d d d
X( dt )w L2 [0, ) since A is a Hurwitz matrix. Since w1 = N ( dt )w = CX( dt )w, we
d
also have that N ( dt )w L2 [0, ). Thus we can integrate the dissipation inequality
from 0 to to obtain
2 Z
d d
(A.24) X w (0)
kN ( )wk2 dt.
dt K 0 dt
Put = 1
1+ and conclude that (2) and (3) are equivalent (also in the noncanonical
case). To show that (1) (2), observe that ||(, ) = McT ()Mc () for the special
symmetric canonical factorization of (, ) corresponding to (3.15). Hence statement
(1) of the theorem is actually statement (2) for this special canonical factorization of
. It thus suffices to prove that if (2) holds for one canonical factorization, then it
holds for any. From matrix theory, it follows that two canonical factorizations
are related by M1 () = SM2 (), with S a nonsingular matrix. Hence (2) for M2
implies
Z + Z +
d d
kM1 wk2 dt = kM2 wk2 dt
dt dt
Z +
d
2 kM2 wk2 dt
dt
Z +
2 d
2
kM 1 wk2 dt
kSk dt
and (2) for M1 follows. Obviously this proof can be reversed with M1 playing the role
of M2 . If M2 comes from a noncanonical factorization, then S may not be nonsingular
and the proof goes through (but cannot be reversed).
Proof of Theorem 9.3. The proof is structured as follows. We first prove that (1)
(2). Subsequently, we show that (1) (3) and finally that (3) (1).
Rt
(1) (2). That Q 0 implies (2a) is obvious. In order to prove (2b) we
need the following lemma. Recall that a quadratic state function Q (or simply ),
d
Q (w) = kX( dt )wk2K , is called positive definite if K > 0.
Lemma A.2. Let M Rq [] be observable. Then there exists a positive definite
state function such that
d d
(A.27) Q (w) kM wk2 .
dt dt
for all w C (R, Rq ). As in the proof of Theorem 5.7, it is easily seen that there
d d
exists w 6= 0 such that A( dt )w = 0 and X( dt )w(0) = a (show that AU 1 is biproper,
ON QUADRATIC DIFFERENTIAL FORMS 1743
Since K+ > 0, there exists > 0 such that 1 K+ > C2 I and 1 > C1 . For this we
have
Z 0 Z 0
2 1 2 T
kvk dt kak dt + x0 K+ x0
1744 J. C. WILLEMS AND H. L. TRENTELMAN
n
X
a(t) = k ek t ak
k=1
for all v B D(R, Rq ). Consider the function w : R Rq such that w(0) = 0 for
t 0 and w(t) = w(t) for t 0. Obviously it holds that
Z +
v T wdt = 0
assume that + = dim(L). Then aT Qa > 0 for all 0 6= a L iff aT Q1 a < 0 for
0 6= a L , and aT Qa 0 for all a L iff aT Q1 a 0 for a L .
Proof. Let L = ker(R) = im(M ) with R surjective and M injective. Then
L = im(RT ) = ker(M T ). Furthermore, aT Qa > 0 for 0 6= a L means M T QM > 0.
Consider the relations
MT 1 T
M T QM 0
Q M Q R = ,
RQ1 0 RQ1 RT
MT M T QM 0
Q M Q1 RT = .
R RQM RRT
1746 J. C. WILLEMS AND H. L. TRENTELMAN
The second relation shows that [M Q1 RT ] is nonsingular. The first shows that
in(M T QM ) + in(RQ1 RT ) = in(Q).
Hence M T QM > 0 implies RQ1 RT < 0. To get the case, replace Q by Q + I
and let 0.
Proof of Theorem 10.2. To prove (i) and (ii), combine Proposition 5.2 and Lemma
A.3 in the following way. For R fixed, define L := im(M (i)) = ker(R(i)).
Define Q := . Note that Q1 = as well. Using that M (i) and RT (i) are
injective, we get the equivalence
M T (i) M (i) > 0 R(i) RT (i) < 0
which yields statement (ii). Statement (i) follows from the second assertion of Lemma
A.3, which yields the same equivalence with nonstrict inequalities.
We now prove (iii). Again this can be proven using Lemma A.3, this time with
Q = I. We then get
M T (i)( I)M (i) 0 R(i)( I)1 RT (i) 0.
Using the formula + I = (1 2 )( I)1 , the latter is equivalent with
R(i) RT (i) R(i)RT (i).
This shows (iii).
In order to prove (iv), we need the following lemma.
Lemma A.4. Let (X, Z) be a matched pair of minimal state maps for B and B .
Define subspaces L Rr+2n , M Rr+2n by
d
w w M ( dt )`
(A.35) L := x Rr+2n | ` C (R, R ) x = X( dt d
)` (0) ,
d d
a a dt X( dt )`
T d
v v R ( dt )`0
(A.36) M := b Rr+2n | `0 C (R, R ) b = dt
d d
Z( dt )`0 (0) .
d
z z Z( dt )`0
This implies
0 Im
P 0 0 C D
0 I 0 L im
In
.
0
0 0 I
A B
Here, in fact, equality holds. Indeed, given col (x0 , u0 ), take any x C (R, Rn ) and
u C (R, Rm ) such that x(0) = x0 and u(0) = u0 , and such that dx dt = Ax + Bu,
d d
y = Cx + Du. There exists ` C (R, R ) such that x = X( dt )`, u = U ( dt )`. This
shows that equality holds and that dim(L) = n + m.
We now prove that L = M. For all ` C (R, R ) and `0 C (R, R ) we
have that (10.4) holds. By evaluating this for t = 0, we immediately obtain that
L M. Thus it suffices to show that dim L = n + (r m). This is, however, an
immediate consequence of the fact that the number of inputs of B , m(B ) is equal to
r m. R
We now return to the proof of (iv) of Theorem 10.2. Assume that Q 0 and
let (, ) = X T ()KX(), with K = K T a storage function for , i.e., . In
d d
terms of w = M ( dt )`, x = X( dt )`, dx d d
dt = dt X( dt )` this inequality yields, in particular,
T
w(0) 0 0 w(0)
(A.37) x(0) 0 0 K x(0) 0.
dx dx
dt (0)
0 K 0 dt (0)
Denote the symmetric matrix in (3.16) by Q. Note that (A.37) says that aT Qa 0
for all a L, with L defined by (A.35). Since dim(L) = n + m = n + r+ , which
is exactly the number of positive eigenvalues of Q, it follows from Lemma A.3 that
aT Q1 a 0 for all a L = M. More explicitly,
0 0
(A.38) aT 0 0 K 1 a 0
1
0 K 0
which is the content of (4). To show (v), use (iv) and Theorem 9.3.
REFERENCES
[1] B.D.O. Anderson, Algebraic properties of minimal degree spectral factors, Automatica, 9
(1973), pp. 491500.
1748 J. C. WILLEMS AND H. L. TRENTELMAN
[2] B.D.O. Anderson and S. Vongpanitlerd, Network Analysis and Synthesis, PrenticeHall,
Englewood Cliffs, NJ, 1973.
[3] B.D.O. Anderson and E.I. Jury, Generalized Bezoutian and Sylvester matrices in multivari-
able linear control, IEEE Trans. Automat. Control, 21 (1976), pp. 551556.
[4] D.Z. Arov, Passive linear stationary dynamical systems, Siberian Math. J., 20 (1979), pp. 149
162.
[5] J.A. Ball, J.W. Helton, and J. William, Shift invariant subspaces, passivity, reproducing
kernels and H -optimization, in Contributions to Operator Theory and its Applications,
Oper. Theory Adv. Appl. 35, I. Gohberg, J.W. Helton, and L. Rodman, eds., Birkhauser,
Basel, 1988, pp. 265310.
[6] S. Barnett, Polynomials and Linear Control Systems, Marcel Dekker, New York, 1983.
[7] R.W. Brockett and J.L. Willems, Frequency domain stability criteria, parts 1 and 2, IEEE
Trans. Automat. Control, 10 (1965), pp. 255261; 401413.
[8] R.W. Brockett, Path integrals, Lyapunov functions, and quadratic minimization, in Proc.
4th Allerton Conference on Circuit and System Theory, University of Illinois, Monticello,
IL, 1966, pp. 685698.
[9] F.M. Callier, On polynomial spectral factorization by symmetric extraction, IEEE Trans.
Automat. Control, 30 (1985), pp. 453464.
[10] W.A. Coppel, Linear Systems, Notes in Pure Mathematics 6, Australian National University,
Canberra, 1972.
[11] P.A. Fuhrmann, Algebraic methods in system theory, in Mathematical System Theory, The
Influence of R.E. Kalman, A.C. Antoulas, ed., Springer-Verlag, New York, 1991, pp. 233
265.
[12] D.J. Hill and P.J. Moylan, Stability of nonlinear dissipative systems, IEEE Trans. Automat.
Control, 21 (1976), pp. 708711.
[13] R.E. Kalman, On the Hermite-Fujiwara theorem in stability theory, Quart. Appl. Math., 23
(1965), pp. 279282.
[14] R.E. Kalman, Algebraic characterization of polynomials whose zeros lie in certain algebraic
domains, in Proc. Nat. Acad. of Sci., 64 (1969), pp. 818823.
[15] V.L. Kharitonov, Asymptotic stability of an equilibrium position of a family of systems of
linear differential equations, Differentsialnye Uravneniya, 14 (1978), pp. 20862088.
[16] V. Kucera, Discrete Linear Control, The Polynomial Approach, John Wiley, Chichester, 1979.
[17] H. Kwakernaak, The polynomial approach to H optimal regulation, in H -Control Theory,
Lecture Notes in Math. 1496, E. Mosca and L. Pandolfi, eds., Springer-Verlag, Berlin, 1991,
pp. 141221.
[18] H. Kwakernaak and M. Sebek, Polynomial J-spectral factorization, IEEE Trans. Automat.
Control, 39 (1994), pp. 315328.
[19] G. Meinsma, Frequency Domain Methods in H Control, Ph.D. thesis, University of Twente,
The Netherlands, 1993.
[20] R.J. Minnichelli, J.J. Anagnost, and C.A. Desoer, An elementary proof of Kharitonovs
stability theorem with extensions, IEEE Trans. Automat. Control, 34 (1989), pp. 995998.
[21] A.C.M. Ran and L. Rodman, Factorization of matrix polynomials with symmetries, SIAM
J. Matrix Anal. Appl., 15 (1994), pp. 845864.
[22] P. Rapisarda and J.C. Willems, State maps for linear systems, SIAM J. Control Optim., 35
(1997), pp. 10531091.
[23] J.W. Schumacher, Transformations of linear systems under external equivalence, Linear Al-
gebra Appl., 102 (1988), pp. 133.
[24] H.L. Trentelman and J.C. Willems, The dissipation inequality and the algebraic Riccati
equation, in The Riccati Equation, S. Bittanti, A.J. Laub, and J.C. Willems, eds., Springer-
Verlag, New York, 1991, pp. 197242.
[25] H.L. Trentelman and J.C. Willems, H control in a behavioral context: The full informa-
tion case, IEEE Trans. Automat. Control, to appear.
[26] H.L. Trentelman and J.C. Willems, Every storage function is a state function, Systems
Control Lett., 32 (1997), pp. 249259.
[27] S. Weiland and J.C. Willems, Dissipative systems in a behavioral context, Math. Models
Methods Appl. Sci., 1 (1991), pp. 125.
[28] J.C. Willems, Least squares stationary optimal control and the algebraic Riccati equation,
IEEE Trans. Automat. Control, 16 (1971), pp. 621634.
[29] J.C. Willems, Dissipative dynamical systemsPart I: General theory, Part II: Linear systems
with quadratic supply rates, Arch. Rational Mech. Anal., 45 (1972), pp. 321351; 352393.
[30] J.C. Willems, On the existence of a nonpositive solution to the Riccati equation, IEEE Trans.
Automat. Control, 19 (1974), pp. 592593.
ON QUADRATIC DIFFERENTIAL FORMS 1749
[31] J.C. Willems, From time series to linear system. Part I: Finite dimensional linear time in-
variant systems; Part II: Exact modelling; Part III: Approximate modelling, Automatica,
22 (1986), pp. 561580; 675694; 23 (1987), pp. 87115.
[32] J.C. Willems, Models for dynamics, Dynamics Report, 2 (1989), pp. 171269.
[33] J.C. Willems, Paradigms and puzzles in the theory of dynamical systems, IEEE Trans. Au-
tomat. Control, 36 (1991), pp. 259294.
[34] D.C. Youla and M. Saito, Interpolation with positive-real functions, J. Franklin Inst., 284
(1967), pp. 77108.