Skip to main content

Adaptive Control for Weakly Minimum Phase Linear Infinite-Dimensional Systems in Hilbert Space Using a Zero Filter

  • Chapter
  • First Online:
Variational Analysis and Aerospace Engineering

Part of the book series: Springer Optimization and Its Applications ((SOIA,volume 116))

  • 1516 Accesses

Abstract

Given a linear continuous-time infinite-dimensional plant on a Hilbert space and disturbances of known waveform but unknown amplitude and phase, we show that there exists a stabilizing direct model reference adaptive control law with persistent disturbance rejection and robustness properties. The plant is described by a closed, densely defined linear operator that generates a continuous semigroup of bounded operators on the Hilbert space of states. For this paper, the plant will be weakly minimum phase, i.e., there will be a finite number of unstable zeros with real part equal to zero. All other zeros will be exponentially stable.

The central result will show that all errors will converge to a prescribed neighborhood of zero in an infinite-dimensional Hilbert space even though the plant is not truly minimum phase. The result will not require the use of the standard Barbalat’s lemma which requires certain signals to be uniformly continuous. This result is used to determine conditions under which a linear infinite-dimensional system can be directly adaptively controlled to follow a reference model. In particular we examine conditions for a set of ideal trajectories to exist for the tracking problem. Our principal result will be that the direct adaptive controller can be compensated with a zero filter for the unstable zeros which will produce the desired robust adaptive control results even though the plant is only weakly minimum phase. Our results are applied to adaptive control of general linear infinite-dimensional systems described by self-adjoint operators with compact resolvent.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations. Springer, Berlin (1983)

    Book  MATH  Google Scholar 

  2. D’Alessandro, D.: Introduction to Quantum Control and Dynamics. Chapman & Hall, London (2008)

    MATH  Google Scholar 

  3. Kothari, D., Nagrath, I.: Modern Power System Analysis. McGraw-Hill, New York (2003)

    Google Scholar 

  4. Curtain, R., Pritchard, A.: Functional Analysis in Modern Applied Mathematics. Academic, London (1977)

    MATH  Google Scholar 

  5. Cannarsa, P., Coron, J.-M. (eds.): Control of Partial Differential Equations (Lecture Notes in Mathematics 2048/C.I.M.E. Foundation Subseries). Springer, Heidelberg (2012)

    Google Scholar 

  6. Troltzsch, F.: Optimal Control of Partial Differential Equations. American Mathematical Society, Providence, RI (2010)

    Book  MATH  Google Scholar 

  7. Balas, M., Erwin, R.S., Fuentes, R.: Adaptive control of persistent disturbances for aerospace structures. Proceedings of the AIAA Guidance, Navigation and Control Conference, Denver (2000)

    Google Scholar 

  8. Fuentes, R., Balas, M.: Direct adaptive rejection of persistent disturbances. J. Math. Anal. Appl. 251, 28–39 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  9. Fuentes, R., Balas, M.: Disturbance accommodation for a class of tracking control systems. Proceedings of the AIAA Guidance, Navigation and Control Conference, Denver, Colorado (2000)

    Google Scholar 

  10. Fuentes, R., Balas, M.: Robust model reference adaptive control with disturbance rejection. Proceedings of the American Control Conference, Anchorage (2002)

    Google Scholar 

  11. Balas, M., Gajendar, S., Robertson, L.: Adaptive tracking control of linear systems with unknown delays and persistent disturbances (or Who You Callin’ Retarded?). Proceedings of the AIAA Guidance, Navigation and Control Conference, Chicago, IL, August 2009

    Google Scholar 

  12. Balas, M., Frost, S.: Adaptive model tracking for distributed parameter control of linear infinite-dimensional systems in Hilbert space. Proceedings of AIAA SCITECH, Boston, July 2013

    Google Scholar 

  13. Balas, M., Frost, S.: Adaptive regulation in the presence of persistent disturbances for linear infinite-dimensional systems in Hilbert space: conditions for almost strict dissipativity. Proceedings of the European Control Conference, Linz, July 2015

    Google Scholar 

  14. Balas, M., Frost, S.: Robust adaptive model tracking for distributed parameter control of linear infinite-dimensional systems in Hilbert space. IEEE/CAA J Automat Sin 1(3), 294–301 (2014)

    Article  Google Scholar 

  15. Kailath, T.: Linear Systems, pp. 448–449. Prentice-Hall (1980)

    Google Scholar 

  16. Wen, J.: Time domain and frequency domain conditions for strict positive realness”. IEEE Trans. Autom. Control 33(10), 988–992 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  17. Renardy, M., Rogers, R.: An Introduction to Partial Differential Equations. Springer, Berlin (1993)

    MATH  Google Scholar 

  18. Kato, T.: Perturbation Theory for Linear Operators. Springer, Berlin (1980)

    MATH  Google Scholar 

  19. Popov, V.M.: Hyperstability of Control Systems. Springer, Berlin (1973)

    Book  MATH  Google Scholar 

  20. Snell, A.: Decoupling of nonminimum phase plants and application to flight control. AIAA Guidance Navigation and Control Conference, Monterey, CA, August 2002

    Google Scholar 

  21. Balas, M.: Trends in large space structure control theory: fondest hopes, wildest dreams. IEEE Trans Automatic Control, AC-27, No. 3 (1982)

    Google Scholar 

  22. Balas, M., Fuentes, R.: A non-orthogonal projection approach to characterization of almost positive real systems with an application to adaptive control. Proceedings of the American Control Conference, Boston (2004)

    Google Scholar 

  23. Antsaklis, P., Michel, A.: A Linear Systems Primer. Birkhauser, Basel (2007)

    MATH  Google Scholar 

  24. Balas, M., Frost, S.: Distributed parameter direct adaptive control using a new version of the Barbalat-Lyapunov stability result in Hilbert Space. Proceedings of AIAA Guidance, Navigation and Control Conference, Boston, MA, 2013

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mark J. Balas .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Proofs of Lemmae 1, 2, and Theorem 1

Proof of Lemma 1

Consider \( \begin{array}{c}{P}_1^2=\left(B{(CB)}^{-1}C\right)\left(B{(CB)}^{-1}C\right)\\ {}=B{(CB)}^{-1}C\equiv {P}_1\end{array} \) .

Hence P 1 is a projection.

Clearly, \( R\left({P}_1\right)\subseteq R(B) \) and \( z=Bu\in R(B) \) which implies \( \begin{array}{c}{P}_1z\kern0.5pt\,{=}\kern-0.5pt\left(\!B{(CB)}^{-1}C\!\right)\!Bu\\ {}=Bu=z\in R\left({P}_1\right)\end{array} \).

Therefore \( R\left({P}_1\right)=R(B) \).

Also \( N\left({P}_1\right)=N(C) \) because \( N(C)\subseteq N\left({P}_1\right) \) and \( z\in N\left({P}_1\right) \) implies that \( {P}_1z\,{=}\,0 \) which implies that \( C{P}_1z=CB{(CB)}^{-1}Cz=0 \) or \( N\left({P}_1\right)\subseteq N(C) \).

So P 2 is a projection onto R(B) along N(C) but \( {P}_2^{*}\ne {P}_2 \) so it is not an orthogonal projection in general. We have \( X=R\left({P}_1\right)\oplus N\left({P}_1\right) \); hence \( X=R(B)\oplus N(C). \)

Since \( {b}_i\in D(A) \), we have \( R(B)\subset D(A) \).

Consequently \( D(A)=\left(R(B)\cap D(A)\right)\oplus \left(N(C)\cap D(A)\right)=R(B)\oplus \left(N(C)\cap D(A)\right) \).

The projection P 1 is bounded since its range is finite-dimensional, and the projection P 2 is bounded because \( \left\Vert {P}_2\right\Vert \le 1+\left\Vert {P}_1\right\Vert <\infty . \)

This completes the proof of Lemma 1.

Proof of Lemma 2

Since X is separable, we can let \( N(C)=\overline{sp}{\left\{{\theta}_k\right\}}_{k=1}^{\infty } \) be an orthonormal basis.

Define \( {W}_2:X\to {l}_2 \) by \( {W}_2x\equiv \left[\begin{array}{c} \left({\theta}_1,{P}_2x\right) \\ {} \left({\theta}_2,{P}_2x\right) \\ {} \left({\theta}_3,{P}_2x\right) \\ {} \dots \\ {} \end{array}\right] \).

Note that \( {\left\Vert {W}_2x\right\Vert}^2={\displaystyle \sum_{k=1}^{\infty }{\left|\left({\theta}_k,{P}_2x\right)\right|}^2}={\left\Vert {P}_2x\right\Vert}^2<\infty \) which implies \( {W}_2x\in {l}_2 \).

So W 2 is a bounded linear operator, and an isometry of W 2 N(C) into l 2.

Consequently \( {W}_2{W}_2^{*}=I \) on N(C).

Then we have \( {W}_2^{*}{W}_2={P}_2 \) and the retraction: \( {z}_2={W}_2{P}_2x\in {l}_2 \).

Also \( {W}_2^{*}{z}_2={W}_2^{*}\left({W}_2{P}_2x\right)={P}_2x \).

Now, using \( x={P}_1x+{P}_2x \) from Lemma 1, we have

$$ \begin{aligned}\overset{.}{y}&=C{P}_1\overset{.}{x}\\ &=C{P}_1A\left({P}_1x+{P}_2x\right)+C{P}_1Bu\\ &=C\left(B{(CB)}^{-1}C\right) AB{(CB)}^{-1}y+C\left(B{(CB)}^{-1}C\right)A\left({W}_2^{*}{z}_2\right)+C\left(B{(CB)}^{-1}C\right)Bu\\ &={\overline{A}}_{11}y+{\overline{A}}_{12}{z}_2+CBu\end{aligned} $$

and

$$ \begin{aligned}{\overset{.}{z}}_2&={W}_2{P}_2\overset{.}{x}\\ &=W{P}_2\left[A\left({P}_1x+{P}_2x\right)+Bu\right]\\ &={W}_2{P}_2A\left(B{(CB)}^{-1}y+{W}_2^8{z}_2\right)+{W}_2{P}_2Bu\\ &={W}_2\left(I-B{(CB)}^{-1}B\right) AB{(CB)}^{-1}y+{W}_2\left(I-B{(CB)}^{-1}B\right)A{W}_2^{*}{z}_2\\ &={\overline{A}}_{21}y+{\overline{A}}_{22}{z}_2.\vspace*{-3pt}\end{aligned} $$

This yields the normal form (11).

Choose \( W\equiv \left[\begin{array}{c} C \\ {} {W}_2{P}_2 \end{array}\right] \) which is a bounded linear operator. Then W has a bounded inverse explicitly stated as \( {W}^{-1}\equiv \left[\begin{array}{cc} B{(CB)}^{-1} & {W}_2^{*} \end{array}\right] \).

This gives

$$ \begin{aligned}W{W}^{-1}&=\left[\begin{array}{cc} CB{(CB)}^{-1} & C{W}_2^{*} \\ {} {W}_2{P}_2B{(CB)}^{-1} & {W}_2{P}_2{W}_2^{*} \end{array}\right]\\ {}&=\left[\begin{array}{cc} {I}_m & 0 \\ {} 0 & {W}_2{W}_2^{*} \end{array}\right]=\left[\begin{array}{cc} {I}_m & 0 \\ {} 0 & I \end{array}\right]=I\end{aligned}\vspace*{-5pt} $$

because \( R\left({W}_2^{*}\right)\subseteq N(C) \).

Furthermore, \( {W}^{-1}W={P}_1+{W}_2^{*}{W}_2{P}_2={P}_1+{P}_2=I \) because \( {W}_2{W}_2^{*}=I \) on N(C).

Also direct calculation yields

$$ \left\{\begin{array}{l}\overline{B}\equiv WB=\left[\begin{array}{c} CB \\ {} {W}_2{P}_2B \end{array}\right]=\left[\begin{array}{c} CB \\ {} 0 \end{array}\right]\\ {}\overline{C}\equiv C{W}^{-1}=\left[\begin{array}{cc} CB{(CB)}^{-1} & C{W}_2^{*} \end{array}\right]=\left[\begin{array}{cc} {I}_m & 0 \end{array}\right]\\ {}\ \overline{A}\equiv WA{W}^{-1}=\left[\begin{array}{cc} CAB{(CB)}^{-1} & CA{W}_2^{*} \\ {} {W}_2{P}_2 AB{(CB)}^{-1} & {W}_2{P}_2A{P}_2{W}_2^{*} \end{array}\right]\end{array}\right. $$

This completes the proof of Lemma 2.

Proof of Theorem 1

Define \( {\overline{S}}_1\equiv {W}^{-1}{S}_1=\left[\begin{array}{c} {\overline{S}}_a \\ {} {\overline{S}}_b \end{array}\right]\ \mathrm{and}\ {\overline{H}}_1\equiv W{H}_1=\left[\begin{array}{c} {\overline{H}}_a \\ {} {\overline{H}}_b \end{array}\right] \).

From (10), we obtain

$$ \left\{\begin{array}{l}\overline{A}{\overline{S}}_1+\overline{B}{S}_2={\overline{S}}_1{L}_m+{\overline{H}}_1\\ {}\overline{C}{\overline{S}}_1={H}_2\end{array}\right. $$

where \( \left(\overline{A},\overline{B},\overline{C}\right) \) is the normal form (11).

From this we obtain

\( \left\{\begin{array}{l}{\overline{S}}_a={H}_2\\ {}{S}_2={(CB)}^{-1}\left[{H}_2{L}_m+{\overline{H}}_a-\left({\overline{A}}_{11}{H}_2+{\overline{A}}_{12}{\overline{S}}_b\right)\right]\\ {}{\overline{A}}_{22}{\overline{S}}_b-{\overline{S}}_b{L}_m={\overline{H}}_b-{\overline{A}}_{21}{H}_2\end{array}\right. \).

We can rewrite the last of these equations as

\( \left(\lambda I-{\overline{A}}_{22}\right){\overline{S}}_b-{\overline{S}}_b\left(\lambda I-{L}_m\right)={\overline{A}}_{21}{H}_2-{\overline{H}}_b\equiv \overline{H} \) for all complex λ.

Now assume that L m is simple and therefore provides a basis of eigenvectors \( {\left\{{\emptyset}_k\right\}}_{k=1}^L\ \mathrm{f}\mathrm{o}\mathrm{r}\ {\Re}^L \). This is not essential but will make this part of the proof easier to understand. The proof can be done with generalized eigenvectors and the Jordan form. So we have

$$ \left({\lambda}_kI-{\overline{A}}_{22}\right){\overline{S}}_b{\phi}_k-{\overline{S}}_b\underset{=0}{\underbrace{\left({\lambda}_kI-{L}_m\right){\phi}_k}}={\overline{A}}_{21}{H}_2-{\overline{H}}_b\equiv \overline{H} $$

which implies that

$$ {\overline{S}}_b{\phi}_k={\left({\lambda}_kI-{\overline{A}}_{22}\right)}^{-1}\overline{H}{\phi}_k $$

because \( {\lambda}_{\mathrm{k}}\in \sigma \left({L}_m\right)\subset \rho \left({\overline{A}}_{22}\right) \).

Thus we have \( {\overline{S}}_bz={\displaystyle \sum_{k=1}^L{\alpha}_k{\left({\lambda}_kI-{\overline{A}}_{22}\right)}^{-1}\overline{H}{\phi}_k}\forall z={\displaystyle \sum_{k=1}^L{\alpha}_k{\phi}_k\in {\Re}^L} \).

Since \( {\lambda}_{\mathrm{k}}\in \sigma \left({L}_m\right)\subset \rho \left({\overline{A}}_{22}\right) \), all \( {\left({\lambda}_kI-{\overline{A}}_{22}\right)}^{-1} \) are bounded operators.

Also \( \overline{H}\equiv {\overline{A}}_{21}{H}_2-{\overline{H}}_b \) is a bounded operator on L.

Therefore \( {\overline{S}}_b \) is a bounded linear operator, and this leads to S 1 also bounded linear.

If we look at the converse statement and let \( {\lambda}_{*}\in \sigma \left({L}_m\right)\cap \sigma \left({\overline{A}}_{22}\right)=\emptyset \).

Then there exists \( {\phi}_{*}\ne 0 \) such that \( \left({\lambda}_{*}I-{\overline{A}}_{22}\right){\overline{S}}_b{\phi}_{*}-{\overline{S}}_b\underset{=0}{\underbrace{\left({\lambda}_{*}I-{L}_m\right){\phi}_{*}}}=\left({\lambda}_{*}I-{\overline{A}}_{22}\right){\overline{S}}_b{\phi}_{*}=\overline{H} \).

In this case three things can happen when \( {\lambda}_{*}\in \sigma \left({\overline{A}}_{22}\right) \):

  1. (1)

    \( \left({\lambda}_{*}I-{\overline{A}}_{22}\right) \) can fail to be one to one so multiple solutions of \( {\overline{S}}_b \) will exist

  2. (2)

    \( R\left({\lambda}_{*}I-{\overline{A}}_{22}\right) \) can fail to be all of X so no solutions \( {\overline{S}}_b \) may occur, or

  3. (3)

    \( {\left({\lambda}_{*}I-{\overline{A}}_{22}\right)}^{-1} \) can fail to be a bounded operator so solutions \( {\overline{S}}_b \) may be unbounded.

In all cases these three alternatives lead to a lack of unique bounded operator solutions for S 1.

The proof of Theorem 1 is complete.

Appendix 2: Proof of Theorem 2

From (15) and Pazy Cor 2.5 p. 107 [1], we have a well-posed system in (16) where A c is a closed operator, densely defined on \( D\left({A}_C\right)\subseteq X \) and generates a C 0 semigroup on X, and all trajectories starting in D(A C ) will remain there. Hence we can differentiate signals in D(A C ).

Consider the positive definite function,

$$ V\equiv \frac{1}{2}\left(Pe,e\right)+\frac{1}{2}\mathrm{t}\mathrm{r}\left[\varDelta G{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right] $$
(21)

where \( \varDelta G(t)\equiv G(t)-{G}^{\ast } \) and P satisfies (13).

Taking the time derivative of (21) (this can be done \( \forall e\in D\left({A}_C\right) \)) and substituting (2a) into the result yields \( \overset{.}{V}=\frac{1}{2}\left[\left(P{A}_c\;e,e\right)+\left(e,P{A}_ce\right)\right]+\left(PBw,e\right)+\mathrm{t}\mathrm{r}\left[\varDelta \overset{.}{G}{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]+\left(Pe,v\right);w\equiv \varDelta Gz \).

Invoking the equalities in Definition 2 of strict dissipativity, using x T y = tr[yx T], and substituting (16) into the last expression, we get (with \( \left\langle {e}_y,w\right\rangle \equiv {e}_y^{*}w \)),

$$ {\fontsize{9}{11}\selectfont\left\{\begin{array}{ll}\overset{.}{V}&=\mathrm{R}\mathrm{e}\left(P{A}_ce,e\right)+\left\langle {e}_y,w\right\rangle -a\cdot \mathrm{t}\mathrm{r}\left[G{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]-\underset{\left\langle {e}_y,w\right\rangle }{\underbrace{\mathrm{tr}\left({e}_y{z}^T\varDelta {G}^T\right)}}+\left(Pe,v\right)\\ {}&\le -{q}_{\min }{\left\Vert e\right\Vert}^2-a\cdot \mathrm{t}\mathrm{r}\left[\left(\varDelta G+{G}^{\ast}\right){\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]+\left(Pe,v\right)\\ {}&\le -\left({q}_{\min }{\left\Vert e\right\Vert}^2+a\cdot \mathrm{t}\mathrm{r}\left[\varDelta G{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right)+a\cdot \left|\mathrm{t}\mathrm{r}\left[{G}^{\ast }{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right|+\left|\left(Pe,v\right)\right|\\ {}&\le -\left(\frac{2{q}_{\min }}{p_{\max }}\cdot \frac{1}{2}\left(Pe,e\right)+2a\cdot \frac{1}{2}\mathrm{t}\mathrm{r}\left[\varDelta G{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right)+a\cdot \left|\mathrm{t}\mathrm{r}\left[{G}^{\ast }{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right|+\left|\left(Pe,v\right)\right|\\ {}&\le -2aV+a\cdot \left|\mathrm{t}\mathrm{r}\ \left[{G}^{\ast }{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right|+\left|\left(Pe,v\right)\right|\end{array}\right. } $$

Now, using the Cauchy-Schwarz inequality

$$ \left|\mathrm{t}\mathrm{r}\left[{G}^{\ast }{\gamma}^{-1}\varDelta {G}^{\mathrm{T}}\right]\right|\le {\left\Vert {G}^{*}\right\Vert}_2{\left\Vert \varDelta G\right\Vert}_2 $$

and

$$ \left|\left(Pe,v\right)\right|\le \left\Vert {P}^{\frac{1}{2}}\nu \right\Vert\ \left\Vert {P}^{\frac{1}{2}}e\right\Vert =\sqrt{\left(P\nu, v\right)}\cdot \sqrt{\left(Pe,e\right)} $$

We have

$$ \begin{array}{l}\overset{.}{V}+2aV\le a\cdot {\left\Vert {G}^{*}\right\Vert}_2{\left\Vert \varDelta G\right\Vert}_2+\sqrt{p_{\max }}\left\Vert \nu \right\Vert \sqrt{\left(Pe,e\right)}\\ {}\le a\cdot {\left\Vert {G}^{*}\right\Vert}_2{\left\Vert \varDelta G\right\Vert}_2+\left(\sqrt{p_{\max }}{M}_{\nu}\right)\sqrt{\left(Pe,e\right)}\\ {}\le \left(a{\left\Vert {G}^{*}\right\Vert}_2+\sqrt{p_{\max }}{M}_{\nu}\right)\sqrt{2}\underset{V^{\frac{1}{2}}}{\underbrace{{\left[\frac{1}{2}\left(Pe,e\right)+\frac{1}{2}{\left\Vert \varDelta G\right\Vert}_2^2\right]}^{\frac{1}{2}}}}\end{array} $$

Therefore,

$$ \frac{\overset{.}{V}+2aV}{V^{\frac{1}{2}}}\le \left(a{\left\Vert {G}^{*}\right\Vert}_2+\sqrt{p_{\max }}{M}_{\nu}\right)\sqrt{2} $$

Now, using the identity \( \mathrm{t}\mathrm{r}\left[ABC\right]=\mathrm{t}\mathrm{r}\left[ CAB\right] \),

$$ \begin{array}{ll}{\left\Vert {G}^{*}\right\Vert}_2&\equiv {\left[\mathrm{t}\mathrm{r}\left({G}^{*}{\gamma}^{-1}{\left({G}^{*}\right)}^T\right)\right]}^{\frac{1}{2}}={\left[\mathrm{t}\mathrm{r}\left({\left({G}^{*}\right)}^T{G}^{*}{\gamma}^{-1}\right)\right]}^{\frac{1}{2}}\\ {}&\le {\left[{\left(\mathrm{t}\mathrm{r}\left({\left({G}^{*}\right)}^T{G}^{*}{\left({G}^{*}\right)}^T{G}^{*}\right)\right)}^{{}^{\frac{1}{2}}}{\left(\mathrm{t}\mathrm{r}\left({\gamma}^{-1}{\gamma}^{-1}\right)\right)}^{\frac{1}{2}}\right]}^{\frac{1}{2}}\\ {}&={\left[\mathrm{t}\mathrm{r}\left({G}^{*}{\left({G}^{*}\right)}^T\right)\right]}^{\frac{1}{2}}{\left[\mathrm{t}\mathrm{r}\left({\gamma}^{-1}\right)\right]}^{\frac{1}{2}}\\ {}&\le \frac{M_{\nu }}{a{M}_G}\cdot {M}_G=\frac{M_{\nu }}{a}\end{array} $$

which implies that

$$ \frac{\overset{.}{V}+2aV}{V^{\frac{1}{2}}}\le \left(1+\sqrt{p_{\max }}\right){M}_{\nu}\sqrt{2} $$
(22)

From

$$ \begin{aligned}\frac{d}{dt}\left(2{e}^{at}{V}^{\frac{1}{2}}\right)&={e}^{at}\frac{\overset{.}{V}+2aV}{V^{\frac{1}{2}}}\\ {}&\le {e}^{at}\left(1+\sqrt{p_{\max }}\right){M}_{\nu}\sqrt{2}\end{aligned} $$

Integrating this expression we have \( {e}^{at}V{(t)}^{1/2}-V{(0)}^{1/2}\le \frac{\left(1+\sqrt{p_{\max }}\right)\;{M}_{\nu }}{a}\left({e}^{at}-1\right) \).

Therefore,

$$ V{(t)}^{1/2}\le V{(0)}^{1/2}{e}^{- at}+\frac{\left(1+\sqrt{p_{\max }}\right)\;{M}_{\nu }}{a}\left(1-{e}^{- at}\right) $$
(23)

The function V(t) is a norm function of the state e(t) and matrix G(t). So, since V(t)1/2 is bounded for all t, then e(t) and G(t) are bounded. We also obtain the following inequality:

$$ \sqrt{p_{\min }}\left\Vert e(t)\right\Vert \le V{(t)}^{1/2} $$

Substitution of this into (23) gives us an exponential bound on state e(τ):

$$ \left\Vert e(t)\right\Vert \le \frac{e^{- at}}{\sqrt{p_{\min }}}V{(0)}^{1/2}+\frac{\left(1+\sqrt{p_{\max }}\right)\;{M}_{\nu }}{a\sqrt{p_{\min }}}\left(1-{e}^{- at}\right) $$
(24)

Taking the limit superior of (24), we have

$$ \overline{\underset{\tau \to \infty }{ \lim }}\left\Vert e(t)\right\Vert \le \frac{\left(1+\sqrt{p_{\max }}\right)\;}{a\sqrt{p_{\min }}}{M}_{\nu}\equiv {R}_{*} $$
(25)

And the proof is complete.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Balas, M.J., Frost, S.A. (2016). Adaptive Control for Weakly Minimum Phase Linear Infinite-Dimensional Systems in Hilbert Space Using a Zero Filter. In: Frediani, A., Mohammadi, B., Pironneau, O., Cipolla, V. (eds) Variational Analysis and Aerospace Engineering. Springer Optimization and Its Applications, vol 116. Springer, Cham. https://doi.org/10.1007/978-3-319-45680-5_2

Download citation

Publish with us

Policies and ethics