Applications of Nonlinear Control by Meral Altınay - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

i

i tj+1 := ti

i

consider the corresponding finite-dimensional optimization problem

minimize JN ( qN( ·), uN( ·)) and PN( qN( ·)),

( qN( ·), uN( ·)) (

Γ N) × ( U N

C 1 (

i

m, N 0, 1)),

(10)

i=1,..., r

where JN and PN are discrete variants of the objective functionals J and P from (7). Moreover,

Γ N is a correspondingly discrete set Γ

(0, 1) is set of suitable discrete functions

i

i and C 1

m, N

that approximate the trajectories set C 1 m(0, 1). Note that the initial continuous optimization

problem can also be presented in a similar discrete manner. For example, we can introduce

the (Euclidean) spaces of piecewise constant trajectories qN ( ·) and piecewise constant control

functions uN ( ·). As we can see the Banach space C1 n(0, 1) and the Hilbert space L2 m([0, 1]) will be replaced in that case by some appropriate finite-dimensional spaces.

The discrete optimization problem (10) approximates the infinite-dimensional optimization

problem (7). We assume that the set of all weak Pareto optimal solution of the discrete

problem (10) is nonempty. Moreover, similarly to the initial optimization problem (7) we

also assume that the discrete problem (10) is regular. If P( ·) is a convex functional, then the

discrete multiobjective optimization problem (10) is also a convex problem. Analogously to

the continuous case (7) or (8) we also can write the corresponding KKT optimality conditions

for a finite-dimensional optimization problem over the set of variables ( qN( ·), uN( ·)). The

necessary optimality conditions for a discretized problem (10) reduce the finite-dimensional

multiobjective optimization problem to a system of nonlinear equations. This problem can be

solved by some gradient-based or Newton-type methods (see e.g., [24]).

Finally, note that the proposed numerical approach uses the necessary optimality conditions,

namely the KKT conditions, for the discrete variant (10) of the initial optimization problem (7).

It is common knowledge that some necessary conditions of optimality for discrete systems, for

example the discrete version of the classical Pontryagin Maximum Principle, are non-correct

in the absence of some restrictive assumptions. For a constructive numerical treatment of

the discrete optimization problem it is necessary to apply some suitable modifications of

the conventional optimality conditions. For instance, in the case of discrete optimal control

problems one can use so-called Approximate Maximum Principle which is specially designed

for discrete approximations of general OCPs [21].

5. Mechanical example

This section is devoted to a short numerical illustration of the proposed hybrid approach

to mechanical systems. We deal with a practically motivated model that has the following

structure (see Fig. 1).

Let us firstly describe the parameters of the mechanical model under consideration:

index-167_1.jpg

On Optimization Techniques for a Class of Hybrid Mechanical Systems

157

On Optimization Techniques for a Class of Hybrid Mechanical Systems

11

Fig. 1. Mechanical example

q 1 it corresponds to the position of motor.

q 2 is the position of inertia J 2.

J 1, J 2 are the external inertias.

Jm is an inertia of motor.

Bm it corresponds to the friction of the motor.

B 1, B 2 they correspond to the frictions of the inertias J 1, J 2.

k is a constant called the rate or spring constant.

u it corresponds to the torque of motor.

The relations for the kinetic potential energies give a rise to the corresponding Lagrange

dynamics:

K( t) = 1 J

+ 1 J

2 m ˙ q 21

2 2 ˙ q 22

V( t) = 1 k ( q

2

1 − q 2)2

Finally, we have

L( q( t), ˙ q( t)) = 1 J

+ 1 J

1 k ( q

2 m ˙ q 21

2 2 ˙ q 22

2

1 − q 2)2

and the Euler-Lagrange equation with respect to the generalized coordinate q 1 has the

following form

Jm ¨ q 1 + Bm ˙ q 1 − k( q 2( t) − q 1( t)) = u( t) (11)

We now considered the Euler-Lagrange equation with respect to the second generalized

variable, namely, with respect to q 2

d ∂L( q( t), ˙ q( t)) − ∂L( q( t), ˙ q( t)) = −B

dt

˙ q

2 ˙

q 2( t)

2

∂q 2

We get the next relation

J 2 ¨ q 2( t) + B 2 ˙ q 2( t) + k( q 2( t) − q 1( t)) = 0

158

12

Will-be-set-by-IN-TECH

Applications of Nonlinear Control

The redefinition of the states x 1 := q 1, x 2 := ˙ q 1, x 3 := q 2, x 4 := ˙ q 2 with X := ( x 1, x 2, x 3, x 4) T

implies the compact state-space form of the resulting equation:

⎡ ⎤

⎤ ⎡ ⎤ ⎡ ⎤

⎡ ⎤

˙ x

0

1

0

0

1

x 1

0

x 0

⎢ ⎥

1

k −Bm

k

0 ⎥ ⎢

⎥ ⎢ 1 ⎥

⎢ ⎥

˙

˙ x

x

x 0

X := ⎢ 2

J

J

J

2

⎣ ⎥

m

m

m

⎥ ⎢ ⎥ ⎢ Jm

⎢ 2⎥

˙ x ⎦ =

⎣ ⎦ + ⎣ ⎦ u, X 0 := ⎣ ⎦

(12)

3

⎣ 0

0

0

1 ⎦ x 3

0

x 03

˙ x

k

−B

4

0

−k

2

x

0

x 0

J

4

2

J 2

J 2

4

The switching structure of the system under consideration is characterized by an additional

inertia J 1 and the associated friction B 1. The modified energies are given by the expressions:

the kinetic energy:

K( t) = 1 J

+ 1 J

+ 1 J

2 m ˙ q 21

2 1 ˙ q 21

2 2 ˙ q 22

the potential energy:

V( t) = 1 k ( q

2

1 − q 2)2

The function of Lagrange can be evaluated as follows

L( q, ˙ q) = 1 J

+ 1 J

+ 1 J

1 k ( q

2 m ˙ q 21

2 1 ˙ q 21

2 2 ˙ q 22

2

1 − q 2)2

(13)

The resulting Euler-Lagrange equations (with respect to q 1 and to q 2 can be rewritten as

( Jm + J 1) ¨ q 1( t) + ( Bm + B 1) ˙ q 1( t) − k( q 2( t) − q 1( t)) = u( t) (14)

J 2 ¨ q 2( t) + B 2 ˙ q 2( t) + k( q 2( t) − q 1( t)) = 0

Using the notation introduced above, we obtain the final state-space representation of the

hybrid dynamics associated with the given mechanical model:

⎡ ⎤

⎤ ⎡ ⎤ ⎡

˙ x

0

1

0

0

1

x 1

0

⎢ ⎥

−k −( Bm+ B 1)

k

⎥ ⎢ ⎥ ⎢ 1 ⎥

˙

˙ x

0 ⎥ x

X := ⎢ 2

2

⎣ ⎥

Jm+ J 1

Jm+ J 1

Jm+ J 1

⎥ ⎢ ⎥ ⎢ Jm+ J 1 ⎥

˙ x ⎦ =

⎣ ⎦ + ⎣

u

(15)

3

⎣ 0

0

0

1 ⎦ x 3

0

˙ x

k

−B

4

0

−k

2

x

0

J

4

2

J 2

J 2

The considered mechanical system has a switched nature with a state-dependent switching

signal. We put x 4 = 10 for the switching-level related to the additional inertia in the system

(see above).

Our aim is to find an admissible control law that minimize the value of the quadratic costs

functional

t f

I( u( ·)) = 1

XT( t) QX( t) + uT( t) Ru( t) dt −→ min

(16)

2 t 0

u( ·)

The resulting Linear Quadratic Regulator that has the follow form

uopt( t) = −R− 1( t) BT( t) P( t) Xopt( t)

(17)

index-169_1.jpg

On Optimization Techniques for a Class of Hybrid Mechanical Systems

159

On Optimization Techniques for a Class of Hybrid Mechanical Systems

13

where P( t) is a solution of the Riccati equation (see [7] for details)

˙

P( t) = ( AT( t) P( t) + P( t) A( t)) + P( t) B( t) R− 1( t) BT( t) P( t) − Q( t) (18)

with the final condition

P( t f ) = 0

(19)

Let us now present a conceptual algorithm for a concrete computation of the optimal pair

( uopt, Xopt( ·)) in this mechanical example. We refer to [7, 8] for the necessary facts and the

general mathematical tool related to the hybrid LQ-techniques.

Algorithm 1. The conceptual algorithm used:

(0) Select a tswi ∈ 0, t f , put an index j = 0

(1) Solve the Riccati euqation (18) for (15) on the time intervals [0, tswi] ∪ tswi, t f

(2) solve the initial problem (12) for (17)

(3) calculate x 4( tswi) + 10 , if | x 4( tswi) + 10 |∼

= for a prescribed accuracy > 0 then Stop. Else,

increase j = j + 1 , inprove tswi = tswi + Δ t and back to (1)

(4) Finally, solve (15) with the obtained initial conditions(the final conditions for the vector X( tswi) )

computed from (12)

Fig. 2. Components of the optimal trajectories

160

14

Will-be-set-by-IN-TECH

Applications of Nonlinear Control

Finally, let us present the simulation results (figure 2). As we can see, the state x 4 satisfies the

switching condition x 4 + 10 = 0. The computed switching time is equal to tswi = 0.0057 s.

The dynamic behaviour on the second time-interval [0, 50] is presented on the figure (2).

The obtained trajectories of the hybrid states converges to zero. As we can see the dynamic

behaviour of the state vector Xopt( t) generated by the optimal hybrid control uopt( ·) guarantee

a minimal value of the quadratic functional I( ·). This minimal value characterize the specific

control design that guarantee an optimal operation (in the sense of the selected objective) of

the hybrid dynamic system under consideration.

6. Concluding remarks

In this paper we propose new theoretical and computational approaches to a specific class

of hybrid OCPs motivated by general mechanical systems. Using a variational structure

of the nonlinear mechanical systems described by hybrid-type Euler-lagrange or Hamilton

equations, one can formulate an auxiliary problem of multiobjective optimization. This

problem and the corresponding theoretical and numerical techniques from multiobjective

optimization can be effectively applied to numerical solution of the initial hybrid OCP.

The proofs of our results and the consideration of the main numerical concepts are

realized under some differentiability conditions and convexity assumptions. These restrictive

smoothness assumptions are motivated by the "classical" structure of the mechanical hybrid

systems under consideration. On the other hand, the modern variational analysis proceeds

without the above restrictive smoothness assumptions. We refer to [20,21] for theoretical

details. Evidently, the nonsmooth variational analysis and the corresponding optimization

techniques can be considered as a possible mathematical tool for the analysis of discontinuous

(for example, variable structure) and impulsive (nonsmooth) hybrid mechanical systems.

Finally, note that the theoretical approach and the conceptual numerical aspects presented in

this paper can be extended to some constrained OCPs with additional state and/or mixed

constraints. In this case one needs to choose a suitable discretization procedure for the

sophisticated initial OCP and to use the corresponding necessary optimality conditions. It

seems also be possible to apply our theoretical and computational schemes to some practically

motivated nonlinear hybrid and switched OCPs in mechanics, for example, to optimization

problems in robots dynamics.

7. References

[1] R. Abraham, Foundations of Mechanics, WA Benjamin, New York, 1967.

[2] C.D. Aliprantis and K.C. Border, Infinite-Dimensional Analysis, Springer, Berlin, 1999.

[3] V. Azhmyakov and J. Raisch, A gradient-based approach to a class of hybrid optimal

control problems, in: Proceedings of the 2nd IFAC Conference on Analysis and Design of

Hybrid Systems, Alghero, Italy, 2006, pp. 89 – 94.

[4] V. Azhmyakov, S.A. Attia and J. Raisch, On the Maximum Principle for impulsive hybrid

systems, Lecture Notes in Computer Science, vol. 4981, Springer, Berlin, 2008, pp. 30 – 42.

[5] V. Azhmyakov, An approach to controlled mechanical systems based on the

multiobjective technique, Journal of Industrial and Management Optimization, vol. 4, 2008,

pp. 697 – 712

On Optimization Techniques for a Class of Hybrid Mechanical Systems

161

On Optimization Techniques for a Class of Hybrid Mechanical Systems

15

[6] V. Azhmyakov, V.G. Boltyanski and A. Poznyak, Optimal control of impulsive hybrid

systems, Nonlinear Analysis: Hybrid Systems, vol. 2, 2008, pp. 1089 – 1097.

[7] V. Azhmyakov, R. Galvan-Guerra and M. Egerstedt, Hybrid LQ-optimization using

Dynamic Programming, in: Proceedings of the 2009 American Control Conference, St. Louis,

USA, 2009, pp. 3617 – 3623.

[8] V. Azhmyakov, R. Galvan-Guerra and M. Egerstedt, Linear-quadratic optimal control of

hybrid systems: impulsive and non-impulsive models, Automatica, to appear in 2010.

[9] J. Baillieul, The geometry of controlled mechanical systems, in Mathematical Control

Theory (eds. J. Baillieul and J.C. Willems), Springer, New York, 1999, pp. 322 – 354.

[10] A.M. Bloch and P.E. Crouch, Optimal control, optimization and analytical mechanics, in

Mathematical Control Theory (eds. J. Baillieul and J.C. Willems), Springer, New York, 1999,

pp. 268–321.

[11] M.S. Branicky, S.M. Phillips and W. Zhang, Stability of networked control systems:

explicit analysis of delay, in: Proceedings of the 2000 American Control Conference, Chicago,

USA, 2000, pp. 2352 – 2357.

[12] A. Bressan, Impulsive control systems, in Nonsmooth Analysis and Geometric Methods in

Deterministic Optimal Control, (eds. B. Mordukhovich and H.J. Sussmann), Springer, New

York, 1996, pp. 1 – 22.

[13] C. Cassandras, D.L. Pepyne and Y. Wardi, Optimal control of class of hybrid systems,

IEEE Transactions on Automatic Control, vol. 46, no. 3, 2001, pp. 398 – 415.

[14] P.D. Christofides and N.H. El-Farra, Control of Nonlinear and Hybrid Processes, Lecture

notes in Control and Information Sciences, vol. 324, Springer, Berlin, 2005.

[15] F.H. Clarke, Optimization and Nonsmooth Analysis, SIAM, Philadelphia, 1990.

[16] G.P. Crespi, I. Ginchev and M. Rocca, Two approaches toward constrained vector

optimization and identity of the solutions,

Journal of Industrial and Management

Optimization, vol. 1, 2005, pp. 549 – 563.

[17] M. Egerstedt, Y. Wardi and H. Axelsson,

Transition-time optimization for

switched-mode dynamical systems, IEEE Transactions on Automatic Control, vol. 51, no.

1, 2006, pp. 110 – 115.

[18] F.R. Gantmakher, Lectures on Analytical Mechanics (in Russian), Nauka, Moscow, 1966.

[19] D. Liberzon Switching in Systems and Control, Birkhäuser, Boston, 2003.

[20] B.S. Mordukhovich, Variational Analysis and Generalized Differentiation, I: Basic Theory, II

Applications, Springer, New York, 2006.

[21] B.S. Mordukhovich,

Approximation Methods in Problems of Optimization and Optimal

Control, Nauka, Moscow, 1988.

[22] H. Nijmeijer and A.J. Schaft, Nonlinear Dynamical Control Systems, Springer, New York,

1990.

[23] B. Piccoli, Hybrid systems and optimal control, in: Proceedings of the 37th IEEE Conference

on Decision and Control, Tampa, USA, 1998, pp. 13 – 18.

[24] E. Polak, Optimization, Springer, New York, 1997