Sunteți pe pagina 1din 16

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO.

10, NOVEMBER 2008 2335

Analytical Approximation Methods for the Stabilizing


Solution of the Hamilton–Jacobi Equation
Noboru Sakamoto, Member, IEEE, and Arjan J. van der Schaft, Fellow, IEEE

Abstract—In this paper, two methods for approximating the Hamilton–Jacobi equation. With these methods, one can cal-
stabilizing solution of the Hamilton–Jacobi equation are proposed culate sub-optimal solutions using a few terms for simple
using symplectic geometry and a Hamiltonian perturbation tech- nonlinearities. Although higher order approximations are
nique as well as stable manifold theory. The first method uses the
fact that the Hamiltonian lifted system of an integrable system is possible to obtain for more complicated nonlinearities, their
also integrable and regards the corresponding Hamiltonian system computations are often time-consuming and there is no guar-
of the Hamilton–Jacobi equation as an integrable Hamiltonian antee that resulting controllers show better performance.
system with a perturbation caused by control. The second method Another approach is through successive approximation, where
directly approximates the stable flow of the Hamiltonian systems the Hamilton–Jacobi equation is reduced to a sequence of first
using a modification of stable manifold theory. Both methods
provide analytical approximations of the stable Lagrangian order linear partial differential equations. The convergence of
submanifold from which the stabilizing solution is derived. Two the algorithm is proven in [24]. In [9] an explicit technique to
examples illustrate the effectiveness of the methods. find approximate solutions to the sequence of partial differential
Index Terms—Hamilton–Jacobi equation, Hamiltonian systems, equation is proposed using the Galerkin spectral method and
nonlinear control theory, perturbation method, stable manifold in [41] the authors propose a modification of the successive
theory, symplectic geometry. approximation method and apply the convex optimization
technique. The advantage of the Galerkin method is that it is
applicable to a larger class of systems, while the disadvantages
I. INTRODUCTION are that it is dependent on how well initial iterate is chosen
and requires the calculation of inner products which can

F OR the analysis and control of linear systems the Riccati


equation plays a fundamental role. In the case of nonlinear
systems the same holds for the Hamilton–Jacobi equation. For
be significantly time-intensive for higher dimensional systems.
The state-dependent Riccati equation approach is proposed in
[20], [29] where a nonlinear function is rewritten in a linear-like
example, an optimal feedback control can be derived from a so- representation. In this method, feedback control is given in
lution of a Hamilton–Jacobi equation [25] and feedback a power series form and has a similar disadvantage to the
controls are obtained by solving one or two Hamilton–Jacobi series expansion technique in that it is useful only for simple
equations [5], [21], [38], [39]. Closely related to optimal control nonlinearities. A technique that employs open-loop controls
and control is the notion of dissipativity, which is charac- and their interpolation is used in [28]. The drawback is that the
terized by a Hamilton–Jacobi inequality (see, e.g., [19], [42]). interpolation of open-loop controls for each point in discretized
Some active areas of research in recent years are the factoriza- state space is time-consuming and the computational cost grows
tion problem [6], [7] and the balanced realization problem [15], exponentially with the state space dimension. A partially related
[36] and the solutions of these problems are again represented research field to approximate solutions of the Hamilton–Jacobi
by Hamilton–Jacobi equations (or, inequalities). Contrary to the equation is the theory of viscosity solutions. It deals with
well-developed theory and computational tools for the Riccati general Hamilton–Jacobi equations for which classical (differ-
equation, which are widely applied, the Hamilton–Jacobi equa- entiable) solutions do not exist. For introductions to viscosity
tion is still an impediment to practical applications of nonlinear solutions see, for instance, [8], [11], [13] and for an application
control theory. to an control problem, see [37]. The finite-element and
In [16], [17], [27], [30] various series expansion tech- finite-difference methods are studied for obtaining viscosity
niques are proposed to obtain approximate solutions of the solutions. They, however, require discretization of state space,
which can be a significant disadvantage.
Manuscript received April 25, 2007; revised November 20, 2007. Current ver- Another direction in the research for the Hamilton–Jacobi
sion published November 05, 2008. This work was supported in part by the Sci-
entist Exchange Program between the Japan Society for the Promotion of Sci-
equation is to study the geometric structure and the proper-
ence (JSPS) and the Netherlands Organisation for Scientific Research (NWO). ties of the equation itself and its exact solutions. The papers
Recommended by Associate Editor A. Loria [38] and [39] give a sufficient condition for the existence of
N. Sakamoto is with the Department of Aerospace Engineering, Nagoya Uni-
versity, Nagoya 464-8603, Japan (e-mail: sakamoto@nuae.nagoya-u.ac.jp).
the stabilizing solution using symplectic geometry. In [35], the
A. J. van der Schaft is with the Institute for Mathematics and Computing Sci- geometric structure of the Hamilton–Jacobi equation is studied
ence, University of Groningen, Groningen 9700 AV, The Netherlands (e-mail: showing the similarity and difference with the Riccati equation.
A.J.van.der.Schaft@math.rug.nl). See also [40] for the treatment of the Hamilton–Jacobi equation
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. as well as recently developed techniques in nonlinear control
Digital Object Identifier 10.1109/TAC.2008.2006113 theory such as the theory of port-Hamiltonian systems. In [32],
0018-9286/$25.00 © 2008 IEEE
2336 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

the solution structure of a nonlinear optimal control problem is solution, is introduced and the geometric theory for the Riccati
investigated using the inverted pendulum as an example. equation is also reviewed. In the beginning of Section IV a key
In this paper, we focus on so-called stationary Hamilton–Ja- observation on integrability for Hamiltonian lifted systems is
cobi equations which are related to, for example, infinite presented. We apply a Hamiltonian perturbation technique (re-
horizon optimal control problems and control problems, viewed in Appendix A) for the system in which the Hamiltonian
and attempt to develop methods to approximate the stabilizing is decomposed into an integrable one and a perturbation Hamil-
solution of the Hamilton–Jacobi equation based on the geo- tonian that is related to the influence of control. By assuming
metric research in [35], [38], and [39]. The main object of that the linearized Riccati equation at the origin has a stabi-
the geometric research on the Hamilton–Jacobi equation is lizing solution, we try to approximate the behaviors on the stable
the associated Hamiltonian system. However, most approx- Lagrangian submanifold. In Section V, an analytical approxi-
imation research papers mentioned above do not explicitly mation algorithm for the stable Lagrangian submanifold is pro-
consider Hamiltonian systems, although it is well-known that posed, using a modification of stable manifold theory. The proof
the Hamiltonian matrix plays a crucial role in the calculation of the main theorem in this section will be given in Appendix B.
of the stabilizing solution for the Riccati equation. One of our In Section VI-A, we address some computational issues. One
purposes in this paper is to fill in this gap. of the eminent features of the approach taken in the paper is that
We will propose two analytical approximation methods for we try to obtain not solutions of the Hamilton–Jacobi equation
obtaining the stabilizing solution of the Hamilton–Jacobi equa- but submanifolds in the extended state space from which the
tion. In the first method, we try to explore the possibility of using solutions are produced by geometric integration (for example,
integrability conditions on the uncontrolled part of the system Poincaré’s lemma). However, only approximations of the
for controller design. Even when one can completely solve the submanifolds are obtained and the integrability condition does
equations of motion for a system with zero input, most non- not hold anymore. We circumvent this difficulty, by obtaining
linear control techniques do not exploit the knowledge because derivatives of the solutions (Section VI-A) or by using integral
once a feedback control is implemented, the system is not inte- expressions of value functions in optimal control problems or
grable anymore. However, within the geometric framework for storage functions in dissipative system theory (Section VI-B).
the Hamilton–Jacobi equation, the effect of control can be con- Also in Section VI-C, we touch on one of the advantages of
sidered as a Hamiltonian perturbation to the Hamiltonian system our analytic approach, by showing that approximate solutions
obtained by lifting the original equations of motion. Here, a cru- can be explicitly obtained as polynomial functions when the
cial property is that if the equations of motion are integrable, system under consideration has only polynomial nonlinearities.
then its lifted Hamiltonian system is also integrable. By using In Section VII-A, we illustrate a numerical example showing
one of the Hamiltonian perturbation techniques (see, e.g., [2], the effectiveness of the proposed methods. Since this is a
[4], [18]) we analyze the behaviors of the Hamiltonian sys- one-dimensional system, one can obtain the rigorous solution,
tems with control effects and try to approximate the Lagrangian which is convenient to see the accuracy and convergence of
submanifold on which the Hamiltonian flow is asymptotically the methods. In Section VII-B, we consider a two-dimensional
stable. problem, an optimal control of a nonlinear spring-mass system,
The second method in this paper takes the approach based in which the spring possesses nonlinear elasticity. Lastly, the
on stable manifold theory (see, e.g., [10], [34]). Using the fact Appendix includes the expositions for the variation of constants
that the stable manifold of the associated Hamiltonian system technique in Hamiltonian perturbation theory, proof of the main
is a Lagrangian submanifold and its generating function corre- theorem in Section V and some formulas of the Jacobi elliptic
sponds to the stabilizing solution, which is shown in [38], and functions used in Section VII-B.
modifying stable manifold theory, we analytically give the solu-
tion sequence that converges to the solution of the Hamiltonian
II. REVIEW OF THE THEORY OF FIRST-ORDER
system on the stable manifold. Thus, each element of the se-
PARTIAL DIFFERENTIAL EQUATIONS
quence approximates the Hamiltonian flow on the stable man-
ifold and the feedback control constructed from each element In this section we outline, by using the symplectic geometric
may serve as an approximation of the desired feedback. It should machinery, the essential parts of the theory of partial differential
be mentioned that computation methods of stable manifolds in equations of the first order.
dynamical systems are being developed and a comprehensive Let us consider a partial differential equation of the form
survey of the recent results in this area can be found in [22].
The proposed method in this paper, however, is different from
the above numerical methods in that it gives analytical expres-
sions of the approximated flows on stable manifolds, which may where is a function of variables, are in-
have considerable potential for control system designs that often dependent variables and with
lead to high dimensional Hamiltonian systems. an unknown function. Since the Hamilton–Jacobi equation in
The organization of the paper is as follows. In Section II, the nonlinear control theory does not explicitly depend on , we
theory of 1st-order partial differential equations is reviewed in did not include it in (PD). The contact geometry handles the
the framework of symplectic geometry, stressing the one-to-one time-varying case (see, e.g., [26]). Let be an dimensional
correspondence between solution and Lagrangian submanifold. space for . We regard the dimensional space
In Section III-A, a special type of solution, called the stabilizing for as the cotangent bundle
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2337

of . is a symplectic manifold with symplectic where with an un-


form . known function, , ,
Let be the natural projection and are all , and is a symmetric matrix for all .
be a hypersurface defined by . Define a submani- We also assume that and satisfy ,
fold and . In what follows, we write , as
, where
is an real matrix and is a symmetric matrix.
The stabilizing solution of (HJ) is defined as follows.
Definition 1: A solution of is said to be the sta-
for a smooth function . Then, is a solution of (PD) if
bilizing solution if and 0 is an asymptotically stable
and only if . Furthermore, is a
equilibrium of the vector field , where
diffeomorphism and is a Lagrangian submanifold because
.
and
It will be important to understand the notion of the stabilizing
solution in the framework of symplectic geometry described in
the previous section. Suppose that we have the stabilizing solu-
Conversely, it is well-known (see, e.g., [1], [31]) that for a La- tion around the origin. Then, the Lagrangian submanifold
grangian submanifold passing through on which corresponding to is
is a diffeomorphism, there exists a neighborhood
of and a function defined on such that

is invariant under the Hamiltonian flow generated by

Therefore, finding a solution of (PD) is equivalent to finding a (3)


Lagrangian submanifold on which
is a diffeomorphism.
Let . To construct such a Lagrangian submanifold To see this invariance, one needs to show that the second equa-
passing through , and hence to obtain a solution de- tion identically holds on , which can be done by taking the
fined on a neighborhood of , it is necessary and sufficient to derivative of (HJ) after replacing with . Note that the
find functions on with right-hand side in the second equation of (3) restricted to
such that , where is the is . The first equation is exactly the
canonical Poisson bracket, and vector field in Definition 1. Therefore, any stabilizing solution is
a generating function of the Lagrangian submanifold on which
(1) is a diffeomorphism and the Hamiltonian flow associated with
is asymptotically stable.
Using these functions, equations , ,
define a Lagrangian submanifold . Note B. Review of the Riccati Equation
that the condition (1) implies, by the implicit function theorem,
It is also useful to see the same picture for the Riccati equa-
that is a diffeomorphism on some neighborhood of .
tion;
Since is the Hamiltonian vector field with Hamil-
tonian , the functions above are first integrals of
. The ordinary differential equations that give the integral
curve of are Hamilton’s canonical equations
which is the linearization of (HJ). A symmetric matrix is said
to be the stabilizing solution of (RIC) if it is a solution of (RIC)
(2) and is stable. The matrix

and therefore, we seek commuting first integrals of (2)


satisfying (1).
is called the Hamiltonian matrix of (RIC) corresponding to the
III. STABILIZING SOLUTION Hamiltonian vector field (3). A necessary and sufficient condi-
tion for the existence of the stabilizing solution [3], [14], [23],
A. The Stabilizing Solution of the Hamilton–Jacobi Equation [33] is that (i) has no eigenvalues on the imaginary axis,
and (ii) the generalized eigenspace for stable eigenvalues
Let us consider the Hamilton–Jacobi equation in nonlinear
satisfies the following complementarity condition;
control theory
2338 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

The condition (i) guarantees that the stable Lagrangian subman- which means that are in involution. Therefore, the
ifold (subspace) does exist while (ii) corresponds to the dif- Hamiltonian system (6) is integrable in the sense of Liouville.
feomorphism assumption of on the Lagrangian submanifold. This means that if one can obtain general solutions of the orig-
More specifically, suppose that the assumptions (i), (ii) are sat- inal system by quadrature, it is also possible for its lifted system.
isfied, then the stabilizing solution exists to (RIC). Take the One may realize that in the analysis of the Hamilton–Jacobi
solution to the Lyapunov equation equation (HJ) Hamilton’s canonical (3) contain the same terms
and set as the Hamiltonian lifting (6) of the plant system. The purpose
of this section is to show that one can exploit this property of
(4) Hamiltonian lifting for approximation of the stabilizing solution
of (HJ).
Assumption 2: The system under control is com-
then it holds that
pletely integrable in the sense that there exist independent
first integrals, and thus a solution for a general
(5) initial condition at is obtained.
Define the perturbation Hamiltonian by
. The Hamiltonian is considered
A nonlinear (Hamilton–Jacobi) extension of (5) is found in [35]. to represent the effect of the control inputs on the integrable
We assume the following throughout the paper. system. We first solve the unperturbed Hamilton’s canonical (6)
Assumption 1: The Riccati equation (RIC) satisfies condi- determined by by means of the Hamilton–Jacobi theory.
tions (i) and (ii), and thus has a stabilizing solution denoted We take the Hamilton–Jacobi approach because it automati-
by . cally produces new canonical variables. It is important to keep
We note that from Assumptiom 1 and (5) one can deduce that working with canonical variables so as not to cause secular
terms in calculations, by which stability analysis may become
unreliable (see, e.g., [18]). The Hamilton–Jacobi equation to
solve (6) is

for all , which shows the invariance of and the linear


(7)
Hamiltonian flow restricted to is asymptotically stable.

Proposition 2: A complete solution of (7) is obtained as


IV. THE HAMILTONIAN PERTURBATION APPROACH
It is well-known that any system described by an ordinary dif-
ferential equation can be represented as a Hamiltonian system
by doubling the system dimension (Hamiltonian lifting). In [12]
this technique is extended to control systems with inputs and '
outputs and is known to be effective for fundamental control
problems such as factorization [6], [7] and model reduction where is the flow of
problems [15]. We first give a useful observation on a Hamil- .
tonian lifted system when the original system is integrable. Proof: The characteristic equation for (7) is
Although it is simple, we did not find this observation in the
literature.
Let the system be completely integrable and
be first integrals. Consider its Hamiltonian
Since the general solution is , ,
lifted system
the independent integrals of the characteristic equation are
. To see this, we note that
(6)

with Hamiltonian . Let for


and . Then,
The general solution of the Hamilton–Jacobi (7) is an ar-
bitrary function of the integrals . We
choose a linear combination of them with constants .

From , by
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2339

a general solution of the lifted unperturbed system (6) is ob- where we have denoted for simplicity and
tained as is a symmetric bilinear map. Noting (11) and
(12), we collect first order terms of and in (10) to get (13).
(8) To solve (13), we set , . Then, we have

or

(9) from which the claim is derived using the inverse transforma-
tion.
We note that the time-dependent transformation Substituting the solution in Proposition 3 into (8) or (9), we
is canonical. In the new coordinates the free motion obtain approximating flows of the Hamiltonian system (3). Now,
(without control) is represented as we wish to select, among them, convergent flows to the origin,
which are the approximations of the flows on the stable La-
grangian submanifold.
With control, the perturbation Hamiltonian is in the coordi- By Assumption 1, it follows that
nates

and , obey Therefore, if we take the initial conditions and satisfying


(stable Lagrangian subspace), then, we have
(10)

We remark that until now no approximation has been made. If


we plug the solution , of (10) into (8) or (9), we get Let us denote quantities in the left-hand side of the above equa-
exact solutions of Hamilton’s canonical (3) for the original con- tion as and . Then, we have the
trol Hamilton–Jacobi equation (HJ) (see, Appendix A). How- following proposition.
ever, it is still difficult to solve (10) and we try to find an ap- Proposition 4: For sufficiently small ,
proximate solution of (10). Using the solution in Assumption 2,
we have

(11)
(12)
(14)
Proposition 3: The linearized equation of (10) is
converge to the origin as .
(13) Proof: This can be verified from (11), (12) and the fact that
is an asymptotically stable matrix.
Moreover, this can be explicitly solved as From Proposition 4, we think of (14) as approximate behav-
iors on the stable Lagrangian submanifold, and thus, (14) can
be regarded as parameterized approximations of the Lagrangian
submanifold. Summarizing, we propose the following method
to approximate the stable Lagrangian submanifold and the sta-
bilizing solution.
Proof: The new Hamiltonian in coordinates is Procedure 1: Solve the uncontrolled system equation
. Form a general solution (8) or (9) of (6) using the solution
H 1 = 012 P T @@x8 (t; x(t; X ))R(x(t;X )) @@x8 (t; x(t; X ))T P:
~ ~
of . Find the stabilizing solution of
(RIC) in Assumption 1. Then,
Thus, we have
q t;X :
+ (8( )) (15)
@H1 T ~
8@ t;x t; X R x t;X )) 2 @ 8~ (t;x(t; X ))T P
= 0 ( ( )) ( (
@P @x @x is a family of approximations of the stable Lagrangian subman-
@ H 1 T = @ 8 T @q (8(t;X ))T ifold. That is, is an ap-
@X @X @x proximation of the derivative of the stabilizing solution.
0 @@x82 @X @ 8 ; R(x(t;X )) @ 8~ P T P
2~
Proof: By eliminating in (14), one can derive (15).
@x Remark IV.1: The set in (15) includes the linearized so-

0 12 @ (p R(@Xx(t;X ))p)
T T lution for . Also, it can be seen that for suffi-
p=(@ 8~ =@x)(t;x(t;X )) P ciently small , each surface in (15) is tangent to , from
2340 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

which one can expect that the performance of the feedback con- B. Approximation Algorithm
trol using (15) is better than that of linear control using
Extracting the linear part in (HJ), (3) can be written as
of (RIC). For a practical method of determining the value of ,
see Section VI-D.

V. THE STABLE MANIFOLD THEORY APPROACH (19)


Using the linear coordinate transformation
A. Approximation of Stable Manifolds
(20)
We consider the following system

(16) where is defined in (4), the linear part of (19) is diagonalized


as

We will assume that the linear part of the equation is separated


in stable and anti-stable directions and , are higher order
terms. (21)
Assumption 3: is an asymptotically stable real matrix
and it holds that , for some constants For (21), Assumption 1 implies Assumption 3 and Assump-
and . tion 4 is satisfied if , and in (HJ) are sufficiently smooth.
Assumption 4: are continuous Thus, we propose the following procedure for parameterized ap-
and satisfy the following. proximation of the stable Lagrangian submanifold.
i) For all , and , Procedure 2:
(i) Construct the sequences (17) for (21) and obtain the se-
quences , in the original coordinates
using (20).
ii) For all , and , (ii) Take a small so as for the convergence of (17) to be
guaranteed for in

where , are continuous and


monotonically increasing on for some constants
.
Furthermore, there exist constants , such that Then,
holds on for .
Let us define the sequences and by (22)

is an approximation of the stable Lagrangian submanifold


(17) and as , where is the stable Lagrangian
submanifold whose existence is assured by Assumption 1
for , and and the results in [38].
Remark V.1: Procedure 1 applies, compared to Procedure 2,
to a smaller class of systems and does not provide a sequential
(18)
method. However, since a nonlinearity is fully taken into ac-
count in Procedure 1, it gives a qualitatively good approximation
with arbitrary . with a large valid range (see, Example VII-A). Nevertheless, one
The following theorem states that the sequences , may wish to obtain better approximations in the Hamiltonian
are the approximating solutions to the exact solution perturbation approach. To this end, we have included the depen-
of (16) on the stable manifold with the property that each ele- dence on in (16), so as to be able to apply Procedure 2 to (10).
ment of the sequences is convergent to the origin. More specifically, one applies the transformation ,
Theorem 5: Under Assumptions 3 and 4, and as in the proof of Proposition 3 to get
are convergent to zero for sufficiently small , that is, ,
as for all Furthermore,
and are uniformly convergent to a solution of
(16) on as . Let and be the limits of
and , respectively. Then, , are the where the higher order terms above are dependent on since
solution on the stable manifold of (16), that is, , is time-dependent. Thus, Procedure 2 can be employed while the
as . 0-th approximation (17) in this case corresponds to the approx-
Proof: See Appendix B. imation in Procedure 1.
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2341

VI. COMPUTATIONAL ISSUES Since the generating function is the minimum value of for
One of the unique features of the approach taken in this paper each which is inside of the radius of convergence of (17), its
is to parameterize a certain -dimensional surface (Lagrangian approximation can be written as
submanifold) in -dimensional space, which is a graph of the
(24)
derivative of the solution. The existence of the solution is guar-
anteed from the integrability property of the surface. For the pur-
pose of the control system design, however, the actual compu- The same computation is possible in the Hamiltonian perturba-
tation of the solution and/or its derivative as a function of is tion approach, when is integrable, by using (14).
necessary. 2) Control Problem: Let us consider the nonlinear
system (23) with disturbances
A. Computation of
In Section IV, the computation for is possible
by eliminating in (14). To obtain an approximation where is a smooth matrix function. The state feedback
of in (22), suppose that is parameterized with control problem is to find a feedback control law
. If one eliminates , from equa- such that the closed loop system is asymptotically stable and has
tions , , the -gain (see, e.g., [39] for definition) from to
the relation is obtained and will serve as an less than or equal to .
approximation of . The elimination of variables in this A sufficient condition for the solvability of the problem
case is, however, not easy to carry out in practice. An effective is that there exists a stabilizing solution to
use of software is required for this purpose. In Section VII,
we interpolate the values of for sample points of to
get the function using MATLAB® commands such as
and .

B. Computation of
and the feedback law is given by
in Procedure 1 and in Procedure 2 are merely approx-
imations of the stable Lagrangian submanifold and do not sat-
isfy the integrability condition. Therefore, it is difficult to get
an approximation of the generating function for the Lagrangian
Procedure 2 can be applied if the linearized problem is
submanifold in a geometric manner. However, since we have an-
solvable and we can construct -th approximation as in (22).
alytical expressions of the approximations, we can write down
From Pontryagin’s minimum principle, one can show that
approximations of the generating function as described below.
1) Optimal Control Problem: Let us consider the following
optimal control problem:

(23)
with the cost function

where
where is a smooth matrix-valued function and
takes the form of, for example, with
smooth , . The optimal feedback control is
given by is the worst disturbance, and is the
solution of the system .
Then, -th approximation for is given, by replacing ,
with , respectively, as
where is the stabilizing solution of the corresponding
Hamilton–Jacobi equation
(25)

where .
By Procedure 2, the -th approximation of the Lagrangian sub- When one designs a feedback control law and only the deriva-
manifold is parameterized as in (22), and the -th approxi- tive of the solution of (HJ) is necessary, we recommend to em-
mation of the optimal feedback can be described with and as ploy the method in Section VI-A. This is because the operations
in (24) or (25) have no effect of approximating the exact solu-
tion and the derivatives of these functions may be less accurate
2342 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

than those obtained by the method inSection VI-A for the same All of these methods are effectively applied using analytical
. The accuracy can be increased by taking larger and the two expressions. We will demonstrate them in the next example.
kinds of approximate derivatives coincide when .
VII. EXAMPLES

C. A Special Case-Polynomial Nonlinearities A. A Numerical Example


Let us consider the 1-dimensional nonlinear optimal control
When (HJ) contains only polynomial nonlinearities, compu-
problem;
tations for and are carried out with elementary func-
tions in the stable manifold theory approach in Section V. In this
case, the higher order terms in (19) are second or higher order
polynomials, and so are , in (17). (26)
The first approximations, corresponding to the linear solu-
tion, are , consisting of exponential and The Hamilton–Jacobi equation for this problem is
trigonometric functions. They are substituted in (17) yielding
also exponential and trigonometric functions since and (27)
are polynomial. The second approximations are obtained by in-
tegrating them after multiplication of the matrix exponential and Hamilton’s canonical equations are
, thus consisting of exponential and trigonometric functions.
This continues for all . Moreover, the integrands in (24) and (28)
(25) are also polynomials of and , and therefore, ’s are
obtained as polynomial functions of .
1) The Hamiltonian Perturbation Method: The Hamiltonian
is split into the integrable and perturbation parts;
D. Determination of Parameters and the Radius of
Convergence
The solution of (26) with the initial condition at
In the perturbation methods, one needs to determine the
without control is obtained from
value of so that (15) gives a good approximation of
in some sense. We propose a practical method of doing that
using the value of Hamiltonian . If is a solution of (HJ), (29)
. Thus, if is an approximation of
with parameter , it may be reasonable to chose so as and is denoted as . The solution of the canonical
to be minimized. equations for corresponding to (8) is
In the stable manifold approach, on the other hand, one needs
to estimate the radius of convergence of the sequence (17).
Since obtaining a theoretical estimation for such a convergence
domain is quite difficult and it tends to be conservative, we pro- where is an arbitrary constant and the last equation is derived
pose a practical method using the values of for each iter- from (29).
ation. If is such that the iteration (22) is convergent, then Based on the linearization of (26), the linearized canonical
is small. However, as grows beyond equations for perturbation that correspond to (10) are
the radius of convergence, the value may rapidly increase. By
looking at this change of for each , one can reasonably es-
timate the radius of convergence.
The radius of convergence in the stable manifold approach
is generally small, meaning that the resultant solution surface The solution of the above equations for the initial condition in
(22) is small around the origin if only positive is used. To the stable Lagrangian subspace of the linearized Riccati equa-
enlarge the domain of the solution, one may try to use negative tion of (27) is
. This, however, is an unstable direction of the flows and
creates a divergent effect. We employ a similar idea to the
above to see how much negative can be substituted in (22).
For a fixed value of , where is the radius of convergence,
where is the stabilizing solution of the Ric-
calculate . Then, for negative , as long as
cati equation and is the closed loop matrix
stays near the exact solution (Lagrangian
(eigenvalue). The family of approximations of the stable La-
submanifold), the value is small. By looking at the growth of
grangian submanifold in Procedure 1 is
this value with respect to , one can see how much negative
can be substituted. If the domain thus obtained is not large
enough, raise and use smaller .
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2343

Fig. 1. Perturbation and Taylor expansion solutions.


Fig. 3.  = 0:42 and extended to the negative time 00:5.

The equations in the new coordinates are

where

We construct the sequences (17) with

and , . From and , the relation of


and is obtained by eliminating , which will be denoted as
. We note that depends on . The approximated
Fig. 2. Integration of error.
feedback functions are .
Figs. 3–5 show the results of calculation for . To guar-
antee the convergence of solution sequence (17), has to be
The feedback function with is shown in Fig. 1. Also, small enough (Theorem 5). If is too large, the sequence is
we showed the result by the Taylor series expansion of order not convergent (compare Figs. 3 and 5). We have estimated the
for the sake of comparison. Since the integrable nonlinearity is radius of convergence using the method in the second paragraph
fully taken into account in this approach, the feedback function of Section VI-D. From Fig. 6, one can see that may
is better approximated in the region further from the origin. be a reasonable estimation.
The value was chosen by the method described If is small and only positive is used in and
in the first paragraph of Section VI-D, which means that the , then the resulting trace in the plane is short, hence,
nonzero value can be thought as an error from the function is defined in a small set around the origin.
the exact solution and its integration may play the role of a norm. Therefore, we substitute negative values in to extend the trace
Fig. 2 shows that takes the minimum value toward the opposite direction. This, however, creates a diver-
at . gent effect on the sequence and this effect becomes smaller as
2) The Stable Manifold Approximation Method: The coordi- increases (see, Fig. 4). We employed the approach in the third
nate transformation that diagonalizes the linear part of (28) is paragraph of Section VI-D to see how much negative time can be
used in (22) to create a larger domain of validity for . From
Fig. 7, one can see that the domain of may be enlarged up
to . If this domain is not large enough, one should raise
and substitute smaller . It can be seen that gives a good
approximation for (domain of validity,
see also Fig. 3), where is from Procedure 2, and that the
2344 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

Fig. 4.  = 0:42 and extended to the negative time 00:8.

( (0;  ); p (0; )).


Fig. 6. Errors H x

Fig. 5.  = 0:6 and extended to the negative time 00:5.

domain of validity for would be


(Fig. 4) which is larger than that of .

B. Optimal Control of a Nonlinear Spring-Mass System


In this example, let us consider an optimal control problem
for a spring-mass system with input :

( ( ) ( ))
Fig. 7. Errors H x t;  ; p t;  .

(30)

1) The Perturbation Approach: Equation (30) with initial


where, is the mass of an object attached to the spring, is the
condition , and no input is integrated as follows:
displacement of the object from rest (at rest, ; the spring
generates no force), and are the linear and nonlinear spring
constants, respectively. Hereafter, we set , for
(32)
the sake of simplicity. The Hamilton–Jacobi equation for this
problem is
where, with
(31) , is the Jacobi elliptic function, and
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2345

Fig. 9. @V=@ x_ with the perturbation method and the second entry of the linear
solution.

Fig. 8. @V=@x with the perturbation method and the first entry of the linear
solution.
2) The Stable Manifold Theory Approach: The associated
Hamiltonian system to (31) is
is the elliptic modulus. is a constant
of integration and can be expressed using and as follows

(35)

Note that , , are functions of , . To express as a


The matrix that diagonalizes the linear part of (35) is
function of , and , we substitute , ,
into (32) and use the addition formulas of the Jacobi p
elliptic functions (see, Appendix C). Thus, in Section IV 1 0 p2p201
0 4(2p2+1)2+4 0
is given as
0 1 0 02p21p201
p p p p
2 201 2 201 1
2 02p22p01201 :

p p p 0 p
2 201 0 p p p
( )( 2+4)
201
2 1 1
The family of approximations of the stable Lagrangian subman- 4(2 2+1) 2 201 2

ifold in Procedure 1
In the new cordinates , (35) is represented as

(36)
is calculated with

(33) where , are in (33), (34) and , are obtained, using

, as
where is the stabilizing solution of

(34)
Now, Procedure 2 can be applied to (36), and sequences (17)
and we have written , .
For the calculation of , it is necessary to differentiate the are transformed into the original coordinates with
Jacobi elliptic functions with respect to the elliptic modulus,
because is differentiated by initial states , and in (32) . Fig. 10 shows the second-order approximation
is dependent on , . We have listed some of the formulas of and the second entry of the linear solution (semi-
required for this calculation in Appendix C. transparent surface). Also, Fig. 11 shows the surfaces repre-
In Figs. 8 and 9, approximations of and with senting with the perturbation and stable manifold
are illustrated with the linear solution . The methods to compare the two methods. The semi-transparent
semi-transparent surfaces represent . It is seen that the surface corresponds to the one with the perturbation method (the
approximate functions are tangent to the linear functions at the same surface in Figs. 9). Figs. 9–11 are drawn from the same di-
origin. rections with the same scales to compare the surfaces. Since the
2346 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

higher order approximations. A one-dimensional example


shows that they are effective in that the optimal feedback is
well approximated and that, compared to the Taylor expansion
method, they give better results especially further from the
equilibrium. An example of a nonlinear spring-mass system
is illustrated to show how they work for a higher dimensional
system.

APPENDIX
The Variation of Constants Technique in Hamiltonian Per-
turbation Theory: We review, in this section, one of the Hamil-
tonian perturbation techniques which is a simple consequence
Fig. 10. @V=@ x_ with the stable manifold method and the second entry of the of the Hamilton–Jacobi theory (see, e.g., [4], [18]).
linear solution.
Let

be the Hamiltonian with the integrable part and the pertur-


bation . By the integrability condition, the Hamilton–Jacobi
equation

(37)

has a complete solution , where


are arbitrary constants. By the canonical coordinate
transformation defined by
Fig. 11. @V=@ x_ with the perturbation and stable manifold methods.
(38)

the unperturbed Hamiltonain in the coordinates becomes


optimal feedback law of this problem does not require ,
0 and the unperturbed equations of motion
the surface for this derivative is not presented.

VIII. CONCLUSION
are converted into
In this paper, we proposed two analytical approxima-
tion approaches for obtaining the stabilizing solution of the
Hamilton–Jacobi equation using a Hamiltonian perturbation
technique and stable manifold theory. The proposed methods By the canonical transformation (38), the new Hamiltonian for
give approximated flows on the stable Lagrangian subman- the perturbed equations of motion is since by (37)
ifold of the associated Hamiltonian system as functions of satisfies
time and initial states. The perturbation approach provides
a set of approximations for the derivative of the stabilizing
solution. On the other hand, in the stable manifold approach,
parametrizations of the stable Lagrangian submanifold are Therefore,
given. Since these methods produce analytical expressions
for approximations, it is possible to compute the solution of
the Hamilton–Jacobi equation using its integral expressions
(Section VI-B). Moreover, in the case of polynomial nonlinear
systems, each approximation step yields the Hamiltonian flows are converted into
with exponential and trigonometric functions in the stable man-
ifold method, providing approximate solutions as polynomial
functions (Section VI-C). In this case, the calculations are all
algebraic, that is, no numerical integration is required and no where, from (38), and .
equations need to be solved. Since these methods focus on the Proof of Theorem 5: From Assumptions 3 and 4, the fol-
stable manifold of the Hamiltonian system, the closed loop lowing inequalities are derived. (In this section, we leave out the
system stability is guaranteed and can be enhanced by taking dependence of and on for the sake of simplicity.)
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2347

• If , then where we have used (39) and (40). Similar calcula-


tions give

(39)
If , then

(40)

• If and for some positive constants


, satisfying , then Thus, (43) for holds with and in (44).
(c) Next, we show that for sufficiently small ,
and are bounded and monotonically in-
creasing sequences and therefore, ,
exists. Furthermore, we show
(41) that , when . Let ,
If and for some positive constants , and . Then, it
, satisfying , then follows that

(42)
(a) First, we show that the limits of sequences (17) and
(18) satisfy (16). By taking limit in (17), we have the
therefore, and if and
integral equations for and
, which is readily verified. We next claim that
the equation

(45)

from which one can see that and satisfy (16). has a solution for sufficiently small . To prove the
(b) For each , and have the claim, define a map by
following estimates;

(43)
where and are the constants defined by
Since
(44)

Indeed, from Assumption 3 and


from which the claim for follows. it follows that for small , and
Let us assume that the claim holds for . is a contraction map in a neighborhood of .
Therefore, has the unique fixed point in ,
which is a solution of (45). We note that when ,
and . It can be shown, in the same
way as the monotonicity proof of and , that
and as long as and ,
which is obvious from (44) and (45). Thus, we have
shown that and are bounded. Therefore,
their limits exist and coincide with since there
is no other solution of (45) in . Because is
the solution of (45), it is clear that as
.
(d) Next, we show that

(46)
(47)
2348 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

where , are the positive sequences defined (e) Lastly, we prove that for sufficiently small ,
by and are monotonically decreasing sequences and
. As a matter of fact,
it can be easily seen, from the definition of the se-
quences, that and for all
if and . However, these
can be verified from
Indeed, for , using (41) and (42), we have

and from the fact that , as . Therefore,


the limits , exist and coincide
with the solution of

Let us assume (46) and (47) for . For , using


(41) and the monotonicity of which has the unique solution .
The Jacobi Elliptic Functions:
Derivation of (32): Let be the solution of
. Then, from
,

where we have set ,


and using (42) and the monotonicity of and is Jacobi’s amplitude function. Thus, we get (32) from
.
Formulas: Differentiations with respect to :
SAKAMOTO AND VAN DER SCHAFT: STABILIZING SOLUTION OF THE HAMILTON–JACOBI EQUATION 2349

Addition formulas: J
[7] J. A. Ball and A. J. van der Schaft, “ -inner-outer factorization
J -spectral factorization, and robust control for nonlinear systems,”
IEEE Trans. Automat. Control, vol. 41, no. 3, pp. 379–392, Mar.
1996.
[8] M. Bardi and I. Capuzzo-Dolcetta, Optimal Control and Viscosity
Solutions of Hamilton–Jacobi–Bellman Equations. Boston, MA:
Birkhäuser, 1997.
[9] R. W. Beard, G. N. Sardis, and J. T. Wen, “Galerkin approximations of
the generalized Hamilton–Jacobi–Bellman equation,” Automatica, vol.
33, no. 12, pp. 2195–2177, 1997.
[10] S.-N. Chow and J. K. Hale, Methods of Bifurcation Theory. New
York: Springer-Verlag, 1982.
[11] M. G. Crandall and P. L. Lions, “Viscosity solutions of Hamilton–Ja-
cobi equations,” Trans. AMS, vol. 277, pp. 1–43, 1983.
[12] P. E. Crouch and A. J. S. van der, Variational and Hamiltonian Control
Systems. New York: Springer-Verlag, 1987.
Differentiation with respect to the elliptic modulus : [13] W. H. Fleming and H. M. Soner, Controlled Markov Processes and
Viscosity Solutions, 2nd ed. New York: Springer, 2005.
[14] B. A. Francis, A Course in H Control Theory. New York: Springer,
1986.
[15] K. Fujimoto and J. M. A. Scherpen, “Nonlinear input-normal
realizations based on the differential eigenstructure of Hankel
operators,” IEEE Trans. Automat. Control, vol. 50, no. 1, pp.
2–18, Jan. 2005.
[16] W. L. Garrard, “Additional results on sub-optimal feedback control of
non-linear systems,” Int. J. Control, vol. 10, no. 6, pp. 657–663, 1969.
[17] W. L. Garrard and J. M. Jordan, “Design of nonlinear automatic flight
control systems,” Automatica, vol. 13, no. 5, pp. 497–505, 1977.
[18] H. Goldstein, Classical Mechanics, 3rd ed. New York: Ad-
dison-Wesley, 2001.
[19] D. Hill and P. Moylan, “The stability of nonlinear dissipative systems,”
IEEE Trans. Automat. Control, vol. 21, no. AC-10, pp. 708–711, Oct.
1976.
[20] Y. Huang and W.-M. Lu, “Nonlinear optimal control: Alternatives to
Hamilton–Jacobi equation,” in Proc. IEEE Conf. Decision Control,
1996, pp. 3942–3947.
where and are the complete elliptic integrals of the [21] A. Isidori and A. Astolfi, “Disturbance attenuation and H -control via
measurement feedback in nonlinear systems,” IEEE Trans. Automat.
first and second kind, respectively, defined by Control, vol. 37, no. 9, pp. 1283–1293, Sep. 1992.
[22] B. Krauskopf, H. M. Osinga, E. J. Doedel, M. E. Henderson, J. Guck-
enheimer, A. Vladimirsky, M. Dellnitz, and O. Junge, “A survey of
methods for computing (un)stable manifolds of vector fields,” Int. J.
Bifurcation Chaos, vol. 15, no. 3, pp. 763–791, 2005.
[23] P. Lancaster and L. Rodman, Algebraic Riccati Equations. Oxford,
U.K.: Oxford Univ. Press, 1995.
[24] R. J. Leake and R.-W. Liu, “Construction of suboptimal control se-
quences,” SIAM J. Control Optim., vol. 5, no. 1, pp. 54–63, 1967.
and zn is Jacobi's zeta function defined by [25] E. B. Lee and L. Markus, Foundations of Optimal Control Theory.
New York: Wiley, 1967.
[26] P. Libermann and C.-M. Marle, Symplectic Geometry and Analytical
Mechanics, D. Reidel, Ed. Dordrecht, The Netherlands: D. Reidel,
1987.
[27] D. L. Lukes, “Optimal regulation of nonlinear dynamical systems,”
SIAM J. Control Optim., vol. 7, no. 1, pp. 75–100, 1969.
[28] J. Markman and I. N. Katz, “An iterative algorithm for solving
ACKNOWLEDGMENT Hamilton–Jacobi type equations,” SIAM J. Sci. Comput., vol. 22, no.
1, pp. 312–329, 2000.
The authors would like to thank anonymous referees and the [29] C. P. Mracek and J. R. Cloutier, “Control designs for the nonlinear
Associate Editor for their valuable comments. benchmark problem via the state-dependent Riccati equation method,”
Int. J. Robust Nonlin. Control, vol. 8, pp. 401–433, 1998.
[30] Y. Nishikawa, N. Sannomiya, and H. Itakura, “A method for subop-
REFERENCES timal design of nonlinear feedback systems,” Automatica, vol. 7, pp.
[1] R. Abraham and J. E. Marsden, Foundations of Mechanics, 2nd ed. 703–712, 1971.
New York: Addison-Wesley, 1979. [31] T. Oshima and H. Komatsu, Partial Differential Equations of the First
[2] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Func- Order. Tokyo, Japan: Iwanami Shoten, 1977.
tions: With Formulas, Graphs, and Mathematical Tables. New York: [32] H. M. Osinga and J. Hauser, “The geometry of the solution set of non-
Dover, 1965. linear optimal control problems,” J. Dynam. Differential Equations,
[3] S. Arimoto, “Optimal feedback control minimizing the effects of noise vol. 18, no. 4, pp. 881–900, 2006.
disturbance,” (in Japanese) Trans. SICE, vol. 2, no. 1, pp. 1–7, 1966. [33] J. E. Potter, “Matrix quadratic solutions,” SIAM J. Appl. Math., vol. 14,
[4] V. I. Arnold, Mathematical Methods of Classical Mechanics, 2nd ed. pp. 496–501, 1966.
Berlin, Germany: Springer-Verlag, 1989. [34] C. Robinson, Dynamical Systems: Stability, Symbolic Dynamics, and
H
[5] J. A. Ball, J. W. Helton, and M. L. Walker, “ control for nonlinear Chaos, 2nd ed. Orlando, FL: CRC Press, 1998.
systems with output feedback,” IEEE Trans. Automat. Control, vol. 38, [35] N. Sakamoto, “Analysis of the Hamilton–Jacobi equation in nonlinear
no. 4, pp. 546–559, Apr. 1993. control theory by symplectic geometry,” SIAM J. Control Optim., vol.
[6] J. A. Ball, M. A. Petersen, and A. van der Schaft, “Inner-outer fac- 40, no. 6, pp. 1924–1937, 2002.
torization for nonlinear noninvertible systems,” IEEE Trans. Automat. [36] J. M. A. Scherpen, “Balancing for nonlinear systems,” Syst. Control
Control, vol. 49, no. 4, pp. 483–492, Apr. 2004. Lett., vol. 21, no. 2, pp. 143–153, 1993.
2350 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 53, NO. 10, NOVEMBER 2008

[37] P. Soravia, “ H control of nonlinear systems: Differential games Arjan J. van der Schaft (M’91–SM’98–F’02) was
and viscosity solutions,” SIAM J. Control Optim., vol. 34, no. 3, pp. born in 1955. He received the B.S. and Ph.D. degrees
1071–1097, 1996. in mathematics from the University of Groningen,
[38] A. J. van der Schaft, “On a state space approach to nonlinear H con- Groningen, The Netherlands, in 1979 and 1983,
trol,” Syst. Control Lett., vol. 16, no. 1, pp. 1–18, 1991. respectively.
L
[39] A. J. van der Schaft, “ -gain analysis of nonlinear systems and non- In 1982 he joined the Department of Applied
linear state feedback H control,” IEEE Trans. Automat. Control, vol. Mathematics, University of Twente, Enschede, The
37, no. 6, pp. 770–784, Jun. 1992. Netherlands, where he was appointed a Full Pro-
L
[40] A. J. S. van der, -Gain and Passivity Techniques in Nonlinear Con- fessor in mathematical systems and control theory in
trol, 2nd ed. Berlin, Germany: Springer, 1999. 2000. In September 2005, he returned to Groningen
[41] A. Wernrud and A. Rantzer, “On approximate policy iteration for as a Full Professor in mathematics. He has served
continuous-time systems,” in Proc. IEEE Conf. Decision Control Eur. as Associate Editor for Systems & Control Letters, the Journal of Nonlinear
Control Conf., 2005, pp. 1453–1458. Science, the SIAM Journal on Control, and the IEEE Transactions on Automatic
[42] J. C. Willems, “Dissipative dynamical systems-Part I, II,” Arch. Rat. Control. Currently, he is Associate Editor for Systems and Control Letters
Mech. Anal., vol. 45, pp. 321–393, 1972. and Editor-at-Large for the European Journal of Control. He is co-author of
System Theoretic Descriptions of Physical Systems (1984), Variational and
Hamiltonian Control Systems (1987), Nonlinear Dynamical Control Systems
Noboru Sakamoto (M’02) received the B.Sc.
degree in mathematics from Hokkaido University,
(1990), L -Gain and Passivity Techniques in Nonlinear Control (2000), and
An Introduction to Hybrid Dynamical Systems (2000). His research interests
Hokkaido, Japan, in 1991 and the M.Sc. and Ph.D. include the mathematical modeling of physical and engineering systems and
degrees in aerospace engineering from Nagoya the control of nonlinear and hybrid systems.
University, Nagoya, Japan, in 1993 and 1996, Dr. van der Schaft served as an Associate Editor of the IEEE TRANSACTIONS
respectively. ON AUTOMATIC CONTROL.
Currently, he is an Associate Professor with the
Department of Aerospace Engineering, Nagoya
University. He held a visiting research position
at the University of Groningen, Groningen, The
Netherlands, in 2005 and 2006. His research inter-
ests include nonlinear control theory, control of chaotic systems, and control
applications for aerospace engineering.

S-ar putea să vă placă și