6.4    Output Measurement-Feedback HH-Control for a General Class of Nonlinear Systems

In this section, we look at the output measurement-feedback problem for a more general class of nonlinear systems. For this purpose, we consider the following class of nonlinear systems defined on a manifold XnXRn containing the origin in coordinates x = (x1, …, xn):

Σg:{˙x=F(x,w,u);x(t0)=x0z=Z(x,u)y=Y(x,w)Σg:x˙=F(x,w,u);x(t0)=x0z=Z(x,u)y=Y(x,w)

(6.52)

where all the variables have their previous meanings, while F:X×W×UXF:X×W×UX is the state dynamics function, Z:X×UsZ:X×URs is the controlled output function and Y:X×WmY:X×WRm is the measurement output function. Moreover, the functions F (., ., .), Z(., .) and YY (., .) are smooth Cr, r ≥ 1 functions of their arguments, and the point x = 0 is a unique equilibrium-point for the system Σg such that F (0, 0, 0) = 0, Z(0, 0) = 0, YY (0, 0) = 0. The following assumptions will also be adopted in the sequel.

Assumption 6.4.1 The linearization of the function Z(x, u) is such that

rank(D12)rank(Zu(0,0))=p.rank(D12)rank(Zu(0,0))=p.

The control action to Σg is to be provided by a controller of the form (6.3). Motivated by the results of Section 6.1, we can conjecture a dynamic-controller of the form:

Σgce:{˙ζ=F(ζ,w,u)+˜G(ζ)(yY(ζ,w))u=˜α2(ζ)Σgce:{ζ˙=F(ζ,w,u)+G˜(ζ)(yY(ζ,w))u=α2˜(ζ)

(6.53)

where ζ is an estimate of x and ˜G(.)G˜(.) is the output-injection gain matrix which is to be determined.

Again, since w is not directly available, we use its worst-case value in equation (6.53). Denote this value by w=˜α1(x)w=α˜1(x), where ˜α1(.)α˜1(.) is determined according to the procedure explained in Chapter 5. Then, substituting this in (6.53) we obtain

{˙ζ=F(ζ,˜α1(ζ),˜α2(ζ))+˜G(ζ)(yY(ζ,˜α(ζ))u=˜α2(ζ).{ζ˙=F(ζ,α˜1(ζ),α˜2(ζ))+G˜(ζ)(yY(ζ,α˜(ζ))u=α˜2(ζ).

(6.54)

Now, let ˜xe=(xζ)x˜e=(xζ) so that the closed-loop system (6.52) and (6.54) is represented by

{˙˜xe=Fe(˜xe,w)z=Ze(˜xe){x˜˙e=Fe(x˜e,w)z=Ze(x˜e)

(6.55)

where

Fe(˜xe,w)=(F(x,w,˜α2(ζ))F(ζ,˜α1(ζ),˜α2(ζ))+˜G(ζ)Y(x,w)˜G(ζ)Y(ζ,˜α1(ζ))),Ze(˜xe)=Z(x,˜α2(ζ)).Fe(x˜e,w)=(F(x,w,α˜2(ζ))F(ζ,α˜1(ζ),α˜2(ζ))+G˜(ζ)Y(x,w)G˜(ζ)Y(ζ,α˜1(ζ))),Ze(x˜e)=Z(x,α˜2(ζ)).

The objective then is to render the above closed-loop system dissipative with respect to the supply-rate s(w,Z)=12(γ2w2Ze(xe)2)s(w,Z)=12(γ2w2Ze(xe)2) with some suitable storage-function and an appropriate output-injection gain matrix ˜G(.)G˜(.). To this end, we make the following assumption.

Assumption 6.4.2 Any bounded trajectory x(t) of the system (6.52)

˙x(t)=F(x(t),0,u(t))x˙(t)=F(x(t),0,u(t))

satisfying

Z(x(t),u(t))0Z(x(t),u(t))0

for all t ≥ t0, is such that limt→∞ x(t) = 0.

Then we have the following proposition.

Proposition 6.4.1 Consider the system (6.55) and suppose Assumptions 6.4.1, 6.4.2 hold. Suppose also the system

˙ζ=F(ζ,˜α1(ζ),0)˜G(ζ)Y(ζ,˜α1(ζ))ζ˙=F(ζ,α˜1(ζ),0)G˜(ζ)Y(ζ,α˜1(ζ))

(6.56)

has a locally asymptotically-stable equilibrium-point at ζ = 0, and there exists a locally defined C1 positive-definite function ˜U:O×O+,U˜:O×OR+, ˜U(0)U˜(0) = 0, O ⊂ X, satisfying the dissipation inequality

˜U˜xe(˜xe)Fe(˜xe,w)+12Ze(˜xe)212γ2w20U˜x˜e(x˜e)Fe(x˜e,w)+12Ze(x˜e)212γ2w20

(6.57)

for w = 0. Then, the system (6.55) has a locally asymptotically-stable equilibrium-point at ˜xe=0.x˜e=0.

Proof: Set w = 0 in the dissipation-inequality (6.57) and rewrite it as

˙˜U(˜xe(t))=˜Uxe(˜xe)Fe(˜xe,w)=12Z(x,˜α2(ζ))0,U˜˙(x˜e(t))=U˜xe(x˜e)Fe(x˜e,w)=12Z(x,α˜2(ζ))0,

which implies that ˜xe=0.x˜e=0. is stable for (6.55). Further, any bounded trajectory ˜xe=(t)x˜e=(t) of (6.55) with w = 0 resulting in ˙˜U(˜xe(t))=0U˜˙(x˜e(t))=0 for all t ≥ ts for some ts ≥ t0, implies that

Z(x(t),˜α2(ζ(t)))0tts.Z(x(t),α˜2(ζ(t)))0tts.

By Assumption 6.4.2 we have limt→∞ x(t) = 0. Moreover, Assumption 6.4.1 implies that there exists a smooth function u = u(x) locally defined in a neighborhood of x = 0 such that

Z(x,u(t))=0andu(0)=0.Z(x,u(t))=0andu(0)=0.

Therefore, limt→∞ x(t) = 0 and Z(x(t),˜α2(ζ2(t)))=0implylimt˜α2(ζ(t))=0.Z(x(t),α˜2(ζ2(t)))=0implylimtα˜2(ζ(t))=0. Finally, limt˜α2(ζ(t))=0implylimtζ(t)=0limtα˜2(ζ(t))=0implylimtζ(t)=0 since ζ is a trajectory of (6.56). Hence, by LaSalle’s invariance-principle, we conclude that xe = 0 is asymptotically-stable. □

Remark 6.4.1 As a consequence of the above proposition, the design of the injectiongain matrix ˜G(ζ)G˜(ζ) should be such that: (a) the dissipation-inequality (6.57) is satisfied for the closed-loop system (6.55) and for some storage-function ˜U(ζ)U˜(ζ); (b) system (6.56) is asymptotically-stable. If this happens, then the controller (6.53) will solve the MFBNLHICP .

In the next several lines, we present the main result for the solution of the MFBNLHICP for the class of systems (6.52). We begin with the following assumption.

Assumption 6.4.3 The linearization of the output function Y (x, w) is such that

rank(D21)rank(Yw(0,0))=m.rank(D21)rank(Yw(0,0))=m.

Define now the Hamiltonian function K : T XX × WW × ℜm ℜ by

K(x,p,w,y)=pTF(x,w,0)yTY(x,w)+12Z(x,0)212γ2w2.K(x,p,w,y)=pTF(x,w,0)yTY(x,w)+12Z(x,0)212γ2w2.

(6.58)

Then, K is concave with respect to w and convex with respect to y by construction. This implies the existence of a smooth maximizing function ŵ(x, p, y) and a smooth minimizing function y(x, p) defined in a neighborhood of (0, 0, 0) and (0, 0) respectively, such that

(K(x,p,w,y)w)w=ˆw(x,p,y)=0,ˆw(0,0,0)=0,(K(x,p,w,y)w)w=wˆ(x,p,y)=0,wˆ(0,0,0)=0,

(6.59)

(2K(x,p,w,y)w2)(x,p,w,y)=(0,0,0,0)=γ2I,(2K(x,p,w,y)w2)(x,p,w,y)=(0,0,0,0)=γ2I,

(6.60)

(K(x,p,ˆw(x,p,y),y)y)y=y*(x,p)=0,y*(0,0)=0,(K(x,p,wˆ(x,p,y),y)y)y=y(x,p)=0,y(0,0)=0,

(6.61)

(2K(x,p,ˆw(x,p,y),y)y2)(x,p,w,y)=(0,0,0,0)=1γ2D21DT21.(2K(x,p,wˆ(x,p,y),y)y2)(x,p,w,y)=(0,0,0,0)=1γ2D21DT21.

(6.62)

Finally, setting

w**(x,p)=ˆw(x,p,y*(x,p)),w**(x,p)=wˆ(x,p,y(x,p)),

we have the following main result.

Theorem 6.4.1 Consider the system (6.52) and let the Assumptions 6.4.1, 6.4.2, 6.4.3 hold. Suppose there exists a smooth positive-definite solution ˜VV˜ to the HJI-inequality (5.75) and the inequality

K(x,WTx(x),w**(x,WTx(x)),y*(x,WTx(x)))˜H(x,˜VTx(X))<0,K(x,WTx(x),w**(x,WTx(x)),y(x,WTx(x)))H˜(x,V˜Tx(X))<0,

(6.63)

where ˜H(.,.)H˜(.,.) is given by (5.72), has a smooth positive-definite solution W (x) defined in the neighborhood of x = 0 with W (0) = 0. Suppose also in addition that

(i)  W(x)˜V(x)>0x0;W(x)V˜(x)>0x0;

(ii)  the Hessian matrix of

K(x,WTx(x),w**(x,WTx(x)),y*(x,WTx(x)))˜H(x,˜VTx(X))K(x,WTx(x),w**(x,WTx(x)),y(x,WTx(x)))H˜(x,V˜Tx(X))

is nonsingular at x = 0; and

(iii)  the equation

(Wx(x)˜V(x))˜G(x)=yT*(x,WTx(x))(Wx(x)V˜(x))G˜(x)=yT(x,WTx(x))

(6.64)

has a smooth solution ˜G(x).

Then the controller (6.54) solves the MFBNLHICP for the system.

Proof: Let Q(x)=W(x)˜V(x) and define

˜S(x,w)Qx[F(x,w,0)G(x)Y(x,w)]+˜H(x,˜VTx,w,0)˜H(x,VTx)

where ˜H(.,.,.,.) is as defined in (5.69). Then,

˜S(x,w)=WxF(x,w,0)yT*(x,WTx)Y(x,w)˜VxF(x,w,0)+˜H(x,˜VTx,w,0)˜H(x,˜VTx)=WxF(x,w,0)yT*(x,WTx)Y(x,w)+12Z(x,0)12γ2w˜H(x,˜VTx)=K(x,WTx,w,y*(x,WTx))˜H(x,˜VTx)K(x,WTx,w**(x,WTx),y*(x,WTx))˜H(x,˜VTx)=xTϒ(x)x

where ϒ(.) is some smooth matrix function which is negative-definite at x = 0.

Now, let

˜U(xe)=Q(xζ)+˜V(x).

Then, by construction, ˜U is such that the conditions (a), (b) of Remark 6.4.1 are satisfied. Moreover, if ˜G is chosen as in (6.64), then

˜UxeF(xe,w)+12Ze(xe)212w2=Qx(xζ)[F(x,w,˜α2(ζ))F(ζ,˜α1(ζ),˜α2(ζ)˜G(ζ)Y(x,w)+˜G(ζ)Y(ζ,˜α1(ζ))]+˜VxF(x,w,˜α2(ζ))+12Z(x,˜α2(ζ))212γ2w2Qx(xζ)[F(x,w,˜α(ζ))F(ζ,˜α1(ζ),˜α2(ζ))˜G(ζ)Y(x,w)+˜G(ζ)Y(ζ,˜α(ζ))+˜H(x,˜VTx,w,˜α(ζ))˜H(x,VTx).

If we denote the right-hand-side (RHS) of the above inequality by L(x, ζ, w), and observe that this function is concave with respect to w, then there exists a unique maximizing function ˜w(x,ζ) such that

(L(x,ζ,W)w)w=˜w(x,ζ)=0,˜w(0,0)=0,

for all (x, ζ, w) in the neighborhood of (0, 0, 0). Moreover, in this neighbohood, it can be shown (it involves some lengthy calculations) that L(x,ζ,˜w(x,ζ)) can be represented as

L(x,ζ,˜w(x,ζ))=(xζ)TR(x,ζ)(xζ)

for some smooth matrix function R(., .) such that R(0, 0) = ϒ(0). Hence, R(., .) is locally negative-definite about (0, 0), and thus the dissipation-inequality

˜UxeF(xe,w)+12Ze(xe)212w2L(x,ζ,˜w(x,ζ))0

is satisfied locally. This guarantees that condition (a) of the remark holds.

Finally, setting w=˜α1(x) in the expression for ˜S(x,w), we get

0>S(x,˜α1(x))Qx(F(x,α1(x),0)G(x)Y(x,α1(x))),

which implies that Q(x) is a Lyapunov-function for (6.56), and consequently the condition (b) also holds. □

Remark 6.4.2 Notice that solving the MFBNLHICP involves the solution of two HJI-inequalities (5.75), (6.63) together with a coupling condition (6.64). This agrees well with the linear theory [92, 101, 292].

6.4.1    Controller Parametrization

In this subsection, we consider the controller parametrization problem for the general class of nonlinear systems represented by the model Σg. For this purpose, we introduce the following additional assumptions.

Assumption 6.4.4 The matrix DT11D11γ2I is negative-definite for some γ > 0, where D11=Zw(0,0).

Define similarly the following

˜H(x,p,w,u)=pTF(x,w,u)+12Z(x,w,u)212γ2w2

(6.65)

r11(x)=(2˜H(x,˜Vx(x),w,u)w2)w=˜α1(x),u=˜α2(x)

(6.66)

r12(x)=(2˜H(x,˜Vx(x),w,u)uw)w=˜α1(x),u=˜α2(x)

(6.67)

r21(x)=(2˜H(x,˜Vx(x),w,u)uw)w=˜α1(x),u=˜α2(x)

(6.68)

r22(x)=(2˜H(x,˜Vx(x),w,u)u2)w=˜α1(x),u=˜α2(x),

(6.69)

where ˜V solves the HJI-inequality (5.75), while ˜α1 and ˜α2 are the corresponding worst-case disturbance and optimal feedback control as defined by equations (5.73), (5.74) in Section 5.6. Set also

˜R(x)=[(11)r11r12(x)r21(x)(1+2)r22(x)]

where 0 < ε1 < 1 and ε2 > 0, and define also ˜K:TX×2(+)×m by

˜K(x,p,w,y)=pTF(x,w+α1(x),0)yTY(x,w+α1(x))+12[w˜α2(x)]T˜R(x)[w˜α2(x)].

Further, let the functions w1(x, p, y) and y1(x, p) defined in the neighborhood of (0, 0, 0) and (0, 0) respectively, be such that

(˜K(x,p,w,y)w)w=w1=0,w1(0.0.0)=0,

(6.70)

(˜K(x,p,w1(x,p,y),y)w)y=y1=0,y1(0,0)=0.

(6.71)

Then we make the following assumption.

Assumption 6.4.5 There exists a smooth positive-definite function ˜Q(x) locally defined in a neighborhood of x = 0 such that the inequality

Y2(x)=˜K(x,˜QTx(x),w1(x,˜QTx(x),y1(x,˜QTx(x)),y1(x,˜QTx(x)))<0

(6.72)

is satisfied, and the Hessian matrix of the LHS of the inequality is nonsingular at x = 0.

Now, consider the following family of controllers defined by

Σcq:{˙ξ=F(ξ,˜α1(ξ),˜α2(ξ)+c(q))+G(ξ)(yY(ξ,˜α1(ξ)))+ˆg1(ξ)c(q)+ˆg1(ξ)d(q)˙q=a(q,yY(ξ,˜α1(ξ)))u=˜α2(ξ)+c(q),

(6.73)

where ξX and q ∈vX are defined in the neighborhoods of the origin in X and ℜv respectively, while G(.) satisfies the equation

˜Qx(x)G(x)=y1(x,˜QTx(x))

(6.74)

with ˜Q(.) satisfying Assumption 6.4.5. The functions a(., .), c(.) are smooth, with a(0, 0) = 0, c(0) = 0, while ˆg1(.),ˆg2(.) and d(.) are Ck, k ≥ 1 of compatible dimensions.

Define also the following Hamiltonian function J : ℜ2n+v ×ℜ2n+v ×ℜr →ℜ by

J(xa,pa,w)=pTaFa(xa,w)+12[w˜α1(x)α2(ξ)+c(q)˜α2(x)]T˜R(x)[w˜α1(x)˜α2(ξ)+c(q)˜α2(x)],

where

xa=[xξq],Fa(xa,w)=[F(x,w,˜α2(ξ)+c(q))˜F(ξ,q)+G(ξ)Y(x,w)+ˆg1(ξ)c(q)+ˆg2(ξ)d(q)a(q,Y(x,w)Y(ξ,˜α1))]

and

˜F(ξ,q)=F(ξ,˜α1,˜α2(ξ)+c(q))G(ξ)Y(ξ,˜α1(ξ),˜α2(ξ)+c(q)).

Note also that, since

(2J(xa,pa,w)w2)(xa,pa,w)=(0,0,0)=(11)(GT11D11γ2I),

there exists a unique smooth solution w2(xa, pa) defined on a neighborhood of (xa, pa) = (0, 0) satisfying

(2J(xa,pa,w)w2)(xa,pa,w)=(0,0,0)=(1ε1)(GT11D11γ2I),

The following proposition then gives the parametrization of the set of output measurement-feedback controllers for the system Σg.

Proposition 6.4.2 Consider the system Σg and suppose the Assumptions 6.4.1-6.4.5 hold. Suppose the following also hold:

(H1) there exists a smooth solution ˜V to the HJI-inequality (5.75), i.e.,

Y1(x):=˜H(x,˜VTx(x),˜α1(x),˜α2(x))0

for all x about x = 0;

(H2) there exists a smooth real-valued function M(xa) locally defined in the neighborhood of the origin in2n+v which vanishes at xa = col(x, x, 0) and is positive everywhere, and is such that

Y3(xa):J(xa,MTxa(xa),w2(xa,MTxa(xa))<0

and vanishes at xa = (x, x, 0).

Then the family of controllers gq solves the MFBNLHICP for the system Σg.

Proof: Consider the Lyapunov-function candidate

Ψ2(xa)=˜V(x)+M(xa)

which is positive-definite by construction. Along the trajectories of the closed-loop system (6.1), (6.73):

Σgcq:{˙xa=Fa(xa,w)z=Z(x,˜α2(ξ)+c(q)),

(6.75)

and by employing Taylor-series approximation of J(xa,MTxa,w2(xa,MTxa(xa)))about w = w2(., .), we have

dΨ2dt+12Z(x,˜α2(ξ)+c(q))212γ2w2=Y1(x)+Y3(xa)+12[w˜α1(x)˜α2(ξ)+c(q)˜α2(x)]T[ε1r11(x)00ε2r22(x)][w˜α1(x)˜α2(ξ)+c(q)˜α2(x)]+12ww2(xa,MTxa(xa))2Γ(xa)+o(w˜α1(x)˜α2(ξ)+c(q)˜α2(x))+o(ww2(xa,MTxa(xa))3)

(6.76)

where

Γ(xa)=(2J(xa,MTxa(xa),w)w2)w=w2(xa,MTxa(xa))

and Γ(0)=(1ε1)(DT11D11γ2I). Further, setting w = 0 in the above equation, we get

dΨ2dt=12Z(x,˜α2(ξ)+c(q))2+Y1(x)+Y3(xa)+12[˜α1(x)˜α2(ξ)+c(q)˜α2(x)]T[ε1r11(x)00ε2r22(x)][˜α1(x)˜α2(ξ)+c(q)˜α2(x)]+12w2(xa,MTxa(xa))2Γ(xa)+(˜α1(x)˜α2(ξ)+c(q)˜α2(x))+o(w2(xa,MTxa(xa))3)

which is negative-semidefinite near xa = 0 by hypothesis and Assumption 6.4.1 as well as the fact that r11(x) < 0, r22(x) > 0 about xa = 0. This proves that the equilibrium-point xa = 0 of (6.75) is stable.

To conclude asymptotic-stability, observe that any trajectory (x(t), ξ(t), q(t)) satisfying

dΨ2dt(x(t),ξ(t),q(t))=0tts

(say!), is necessarily a trajectory of

˙x(t)=F(x,0,˜α2(ξ)+c(q)),

such that x(t) is bounded and Z(x, 0, ˜α2(ξ) + c(q)) = 0 for all tts. Furthermore, by hypotheses H1 and H2 and Assumption 6.4.1, ˙Ψ2(x(t),ξ(t),q(t))=0 for all tts, implies x(t) = ξ(t) and q(t) = 0 for all tts. Consequently, by Assumption 6.4.2 we have limt→∞ x(t) = 0, limt→∞ ξ(t) = 0, and by LaSalle’s invariance-principle, we conclude asymptotic-stability. Finally, integrating the expression (6.76) from t = t0 to t = t1, starting at x(t0), ξ(t0), q(t0), and noting that Y1(x) and Y (xa) are nonpositive, it can be shown that the closed-loop system has locally ℒ2-gain ≤ γ or the disturbance-attenuation property. □

The last step in the parametrization is to show how the functions ˆg1(.),ˆg2(.),d(.) can be selected so that condition H2 in the above proposition can be satisfied. This is summarized in the following theorem.

Theorem 6.4.2 Consider the nonlinear system Σg and the family of controllers (6.73). Suppose the Assumptions 6.4.1-6.4.5 and hypothesis H1 of Proposition 6.4.2 holds. Suppose the following also holds:

(H3) there exists a smooth positive-definite function L(q) locally defined on a neighborhood of q = 0 such that the function

Y4(q,w)=Lq(q)a(q,Y(0,w))+12[wc(q)]T˜R(0)[wc(q)]

is negative-definite at w = w3(q) and its Hessian matrix is nonsingular at q = 0, where w3(.) is defined on a neighborhood of q = 0 and is such that

(Y4(q,w)w)w=w3(q)=0,w3(0)=0.

Then, if ˆg1(.),ˆg2(.) are selected such that

˜Qx(x)ˆg1(x)=˜Qx(x)(Fu(x,0,0)G(x)Yu(x,0,0))+βT(x,0,0)r12(x)(1+2)˜αT2(x)r22(x)

and

˜Qx(x)ˆg2(x)=YT(x,˜α1(x)+β(x,0,0)),

respectively, where

β(x,ξ,q)=w2(xa,[˜Qx(xξ)˜Qx(xξ)Lq(q)])

and let d(q)=LTq(q). Then, each of the family of controllers (6.73) solves the MFBNLHICP for the system (6.52).

Proof: It can easily be shown by direct substitution that the function M(xa)=˜Q(xξ)+L(q) satisfies the hypothesis (H2) of Proposition 6.4.2. □

6.5    Static Output-Feedback Control for Affine Nonlinear Systems

In this section, we consider the static output-feedback (SOFB) stabilization problem (SOFBP) with disturbance attenuation for affine nonlinear systems. This problem has been extensively considered by many authors for the linear case (see [113, 121, 171] and the references contained therein). However, the nonlinear problem has received very little attention so far [45, 278]. In [45] some sufficient conditions are given in terms of the solution to a Hamilton-Jacobi equation (HJE) with some side conditions, which when specialized to linear systems, reduce to the necessary and sufficient conditions given in [171].

In this section, we present new sufficient conditions for the solvability of the problem in the general affine nonlinear case which we refer to as a “factorization approach” [27]. The sufficient conditions are relatively easy to satisfy, and depend on finding a factorization of the state-feedback solution to yield the output-feedback solution.

We begin with the following smooth model of an affine nonlinear state-space system defined on an open subset Xn without disturbances:

Σa:{˙x=f(x)+g(x)u;x(t0)=x0y=h(x)

(6.77)

where xX is the state vector, u ∈ ℜp is the control input, and y ∈ ℜm is the output of the system. The functions f : XV(X ), and g:Xn×p,h:Xm are smooth C functions of x. We also assume that x = 0 is an equilibrium point of the system (6.77) such that f(0) = 0, h(0) = 0.

The problem is to find a SOFB controller of the form

u=K(y)

(6.78)

for the system, such that the closed-loop system (6.1)-(6.78) is locally asymptotically-stable at x = 0.

For the nonlinear system (6.77), it is well known that [112, 175], if there exists a C1 positive-definite local solution V : ℜn → ℜ+ to the following Hamilton-Jacobi-Bellman equation (HJBE):

Vx(x)f(x)12Vx(x)g(x)gT(x)VTx(x)+12hT(x)h(x)=0,V(0)=0

(6.79)

(or the inequality with “ ≤ ″), then the optimal control law

u*=gT(x)VTx(x)

(6.80)

locally asymptotically stabilizes the equilibrium point x = 0 of the system and minimizes the cost functional:

J1(u)=12t0(y2+u2)dt.

(6.81)

Moreover, if V > 0 is proper (i.e., the set Ωc={x|0V(x)a} is compact for all a > 0), then the above control law (6.80) globally asymptotically stabilizes the equilibrium point x = 0. The aim is to replace the state vector in the above control law by the output vector or some function of it, with some possible modifications to the control law. In this regard, consider the system (6.77) and suppose there exists a C1 positive-definite local solution to the HJE (6.79) and a C0 function F : ℜm → ℜp such that:

Fh(x)=gT(x)VTx(x),

(6.82)

where “◦” denotes the composition of functions. Then the control law

u=F(y)

(6.83)

solves the SOFBP locally. It is clear that if the function F exists, then

u=F(y)=Fh(x)=gT(x)VTx(x).

This simple observation leads to the following result.

Proposition 6.5.1 Consider the nonlinear system (6.77) and the SOFBP. Assume the system is locally zero-state detectable and the state-feedback problem is locally solvable with a controller of the form (6.80). Suppose there exist C0 functions F1:YmU,ϕ1:X0Xp,η1:X(non-positive), η1[t0,) such that

F1h(x)=gT(x)VTx(x)+ϕ1(x)

(6.84)

V(x)g(x)ϕ1(x)=η1(x).

(6.85)

Then:

(i)  the SOFBP is locally solvable with a feedback of the form

u=F1(y);

(6.86)

(ii)  if the optimal cost of the state-feedback control law (6.80) is J*SFB(u), then the cost JSOFB (u) of the SOFB control law (6.86) is given by

J1,SOBF(u)=J*1,SFB(u)+t0η1(x)dt.

Proof: Consider the closed-loop system (6.77), (6.86):

˙x=f(x)+g(x)F1(h(x)).

Differentiating the solution V > 0 of (6.79) along the trajectories of this system, we get upon using (6.84), (6.85) and (6.79):

˙V=Vx(x)(f(x)+g(x)F1(h(x))=Vx(x)(f(x)g(x)gT(x)VTx(x)+g(x)ϕ1(x))=Vx(x)f(x)Vxg(x)gT(x)VTx(x)+Vx(x)g(x)ϕ1(x)=12Vx(x)g(x)gT(x)VT(x)12hT(x)h(x)+η1(x)0.

(6.87)

This shows that the equilibrium point x = 0 of the closed-loop system is Lyapunov-stable. Moreover, by local zero-state detectability of the system, we can conclude local asymptotic-stability. This establishes (i).

To prove (ii), integrate the last expression in (6.87) with respect to t from t = t0 to , and noting that by asymptotic-stability of the closed-loop system, V (x(∞)) = 0, this yields

J1(u*):=12t0(y2+u*2)dt=V(x0)+t0η1(x)dt

Since V (x0) is the total cost of the policy, the result now follows from the fundamental theorem of calculus. □

Remark 6.5.1 Notice that if Σa is globally detectable, V > 0 is proper, and the functions ϕ1(.) and η1(.) are globally defined, then the control law (6.86) solves the SOFBP globally.

A number of interesting corollaries can be drawn from Proposition 6.5.1. In particular, the condition (6.85) in the proposition can be replaced by a less stringent one given in the following corollary.

Corollary 6.5.1 Take the same assumptions as in Proposition 6.5.1. Then, condition (6.85) can be replaced by the following condition

ϕT1(x)F1h(x)0

(6.88)

for all xX1 a neighborhood of x = 0.

Proof: Multiplying equation (6.84) by ϕT1(x) from the left and substituting equation (6.85) in it, the result follows by the non-positivity of η1. □

Example 6.5.1 We consider the following scalar nonlinear system

˙x=x3+12x+uy=2x.

Then the HJBE (6.79) corresponding to this system is

(x3+12x)Vx(x)12V2x(x)+x2=0,

and it can be checked that the function V (x) = x2 solves the HJE. Moreover, the state-feedback problem is globally solved by the control law u = −2x. Thus, clearly,u=2y, and hence the function F1(y) = −y solves the SOFBP globally.

In the next section we consider systems with disturbances.

6.5.1    Static Output-Feedback Control with Disturbance-Attenuation

In this section, we extend the simple procedure discussed above to systems with disturbances. For this purpose, consider the system (6.77) with disturbances [113]:

Σa2:{˙x=f(x)+g1(x)w+g2(x)uy=h2(x)z=[h1(x)u]

(6.89)

where all the variables are as defined in the previous sections and wW2[t0,). We recall from Chapter 5 that the state-feedback control (6.80) where V : N → ℜ+ is the smooth positive-definite solution of the Hamilton-Jacobi-Isaacs equation (HJIE):

Vx(x)f(x)+12Vx(x)[1γ2g1(x)gT1(x)g2(x)gT2(x)]VTx(x)+12hT1(x)h1(x)=0,V(0)=0,

(6.90)

minimizes the cost functional:

J2(u,w)=12t0(z(t)2γ2w(t)2)dt

and renders the closed-loop system (6.89), (6.80) dissipative with respect to the supply-rate s(w,z)=12(γ2w2z2) and hence achieves ℒ2-gain ≤ γ. Thus, our present problem is to replace the states in the feedback (6.78) with the output y or some function of it. More formally, we define the SOFBP with disturbance attenuation as follows.

Definition 6.5.1 (Static Output-Feedback Problem with Disturbance Attenuation (SOFBPDA)). Find (if possible!) a SOFB control law of the form (6.78) such that the closed-loop system (6.89), (6.78) has locally L2-gain from w to z less or equal γ > 0, and is locally asymptotically-stable with w = 0.

The following theorem gives sufficient conditions for the solvability of the problem.

Theorem 6.5.1 Consider the nonlinear system (6.89) and the SOFBPDA. Assume the system is zero-state detectable and the state-feedback problem is locally solvable in NX with a controller of the form (6.80) and V > 0 a smooth solution of the HJIE (6.90). Suppose there exist C0 functions F2:YmU,ϕ2:X1Xp,η2:X1X (non-positive),η2[t0,) such that the conditions

F2h2(x)=gT2(x)VTx(x)+ϕ2(x),

(6.91)

Vx(x)g2(x)ϕ2(x)=η2(x),

(6.92)

are satisfied. Then:

(i)  the SOFBP is locally solvable with the feedback

u=F2(y);

(6.93)

(ii)  if the optimal cost of the state-feedback control law (6.80) is J*2,SFB(u,w)=V(x0), then the cost J2,SOFB(u) of the SOFBPDA control law (6.93) is given by

J2,SOFB(u)=J*2,SFB(u)+t0η2(x)dt.

Proof: (i) Differentiating the solution V > 0 of (6.90) along a trajectory of the closed-loop system (6.89) with the control law (6.93), we get upon using (6.91), (6.92) and the HJIE (6.90)

˙V=Vx(x)[f(x)+g1(x)w+g2(x)F2(h2(x))]=Vx(x)[f(x)+g1(x)wg2(x)gT2(x)VTx(x)+g2(x)ϕ(x)]=Vx(x)f(x)+Vx(x)g1(x)wVx(x)g2(x)gT2(x)VT(x)+Vx(x)g2(x)ϕ2(x)12Vx(x)g2(x)gT2(x)VT(x)12hT1(x)h1(x)+12γ2w2γ22wgT1(x)VTx(x)γ2212(γ2w2z2).

(6.94)

Integrating now from t = t0 to t = T we have the dissipation inequality

V(x(T))V(x0)Tt012(γ2w2z2)dt,

and therefore the system has ℒ2-gain ≤ γ. In addition, with w ≡ 0, we get

˙V12z2,

and therefore, the closed-loop system is locally stable. Moreover, the condition ˙V0 for all ttc, for some tc, implies that z ≡ 0 for all ttc, and consequently, limt→∞ x(t) = 0 by zero-state detectability. Finally, by the LaSalle’s invariance-principle, we conclude asymptotic-stability. This establishes (i).

(ii) Using similar manipulations as in (i) above, it can be shown that

˙V=12(γ2w2z2)γ22wgT1(x)VTx(x)γ22+η2(x)

integrating the above expression from t = t0 to ∞, substituting w = w and rearranging, we get

J2(u,w)=V(x0)+t0η2(x)dt.

We consider another example.

Example 6.5.2 For the second order system

˙x1=x1x2˙x2=x21+x2+u+wy=[x1x1+x2]Tz=[x2u]T,

the HJIE (6.90) corresponding to the above system is given by

x1x2Vx1+(x21+x2)Vx2+(1γ2)γ2V2x2+12x22=0.

Then, it can be checked that for γ=25, the function V(x)=12(x21+x22), solves the HJIE. Consequently, the control law u = −x2 stabilizes the system. Moreover, the function F2(y) = y1y2 clearly solves the equation (6.91) with ϕ2(x) = 0. Thus, the control law u = y1y2 locally stabilizes the system.

We can specialize the result of Theorem 6.5.1 above to the linear system:

Σl:{˙x=Ax+B1w+B2u,x(t0)=0y=c2xz=[c1xu]

(6.95)

where A ∈ ℜn×n, B1n×r, B2 ∈ ℜn×p, C2 ∈ ℜs×n, and C1 ∈ ℜ(spn are constant matrices. We then have the following corollary.

Corollary 6.5.2 Consider the linear system (6.95) and the SOFBPDA. Assume (A, B2) is stabilizable, (C1, A) is detectable and the state-feedback problem is solvable with a controller of the form u=BT1Px, where P > 0 is the stabilizing solution of the algebraic-Riccati equation (ARE):

ATP+PA+P[1γ2B1BT1B2BT2]P+CT1C1=0.

In addition, suppose there exist constant matrices Γ2p×m,Φ2p×n,0Λ2n×n such that the conditions

Γ2C2=BT2P+Φ2,

(6.96)

PB2Φ2=Λ2,

(6.97)

are satisfied. Then:

(i)  the SOFBP is solvable with the feedback

u=Γ2y;

(6.98)

(ii)  if the optimal cost of the state-feedback control law (6.80) is J*2,SFBl(u,w)=12xT0Px0, then the cost JSOFB2(u,w) of the SOFBDA control law (6.93) is given by

J*2,SOFBl(u,w)=J*2,SFBl(u,w)xT0Λ2x0.

In closing, we give a suggestive adhoc approach for solving for the functions F1, F2, for of the SOFB synthesis procedures outlined above. Let us represent anyone of the equations (6.84), (6.91) by

Fh2(x)=α(x)+ϕ(x)

(6.99)

where α(.) represents the state-feedback control law in each case. Then if we assume for the time-being that ϕ is known, then F can be solved from the above equation as

F=αh12+ϕh12

(6.100)

provided h12 exists at least locally. This is the case if h2 is injective, or h2 satisfies the conditions of the inverse-function theorem [230]. If neither of these conditions is satisfied, then multiple or no solution may occur.

Example 6.5.3 Consider the nonlinear system

˙x1=x31+x2x1+w˙x2=x1x2+uy=x2.

Then, it can be checked that the function V(x)=12(x21+x22) solves the HJIE (6.90) corresponding to the state-feedback problem for the system. Moreover,

α=gT2(x)VTx(x)=x2,h12(x)=h2(x)=x2.

Therefore, with ϕ=0,F(x)=αh12(x)=x2orF(y)=y.

We also remark that the examples presented are really simple, because it is otherwise difficult to find a closed-form solution to the HJIE. It would be necessary to develop a computational procedure for more difficult problems.

Finally, again regarding the existence of the functions F1, F2, from equation (6.99) and applying the Composite-function Theorem [230], we have

DxF(h2(x))Dxh2(x)=Dxα(x)+Dxϕ(x).

(6.101)

This equation represents a system of nonlinear equations in the unknown Jacobian DxF. Thus, if Dh2 is nonsingular, then the above equation has a unique solution given by

DxF(h2(x))=(Dxα(x)+Dxϕ(x))(Dxh2(x))1.

(6.102)

Moreover, if h12 exists, then F can be recovered from the Jacobian, DF (h2), by composition. However, Dh2 is seldom nonsingular, and in the absence of this, the generalized-inverse can be used. This may not however lead to a solution. Similarly, h12 may not usually exist, and again one can still not rule out the existence of a solution F. More investigation is still necessary at this point. Other methods for finding a solution using for example Groebner basis [85] and techniques from algebraic geometry are also possible.

6.6    Notes and Bibliography

The results of Sections 6.1, 6.3 of the chapter are based mainly on the valuable papers by Isidori et al. [139, 141, 145]. Other approaches to the continuous-time output-feedback problem for affine nonlinear systems can be found in the References [53, 54, 166, 190, 223, 224], while the results for the time-varying and the sampled-data measurement-feedback problems are discussed in [189] and [124, 213, 255, 262] respectively. The discussion on controller parametrization for the affine systems is based on Astolfi [44] and parallels that in [190]. It is also further discussed in [188, 190, 224], while the results for the general case considered in Subsection 6.4.1 are based on the Reference [289]. In addition, the results on the tracking problem are based on [50].

Similarly, the robust control problem is discussed in many references among which are [34, 147, 192, 208, 223, 265, 284]. In particular, the Reference [145] discusses the robust output-regulation (or tracking or servomechanism) problem, while the results on the reliable control problem are based on [285].

Finally, the results of Section 6.5 are based on [27], and a reliable approach using TSfuzzy models can be found in [278].

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset