Skip to main content
Log in

Near optimal \(\mathcal {H}_{\infty }\) performance in the decentralized setting

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

In this paper, we consider the use of a linear periodic controller (LPC) for the control of linear time-invariant (LTI) plants in the decentralized setting with an \(H_{\infty }\)-performance criterion in mind. Here we show that if the plant is centrally stabilizable and detectable, the graph associated with the plant is strongly connected, and a technical condition on the relative degree holds, then we can design a decentralized LPC to provide a level of \(H_{\infty }\) performance as close as desired to that provided by a given (stabilizing) LTI centralized controller; this will be the case even if the plant has an unstable decentralized fixed mode (DFM). This means, in particular, that under the listed conditions, we can design a decentralized LPC to provide near \(H_{\infty }\)-optimal centralized performance.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. For completeness, we’d like to point out that there are specific application problems such as that of controlling a platoon of vehicles or that of networked control, e.g. see [24], in which \(u_i\) depends not only on \(y_i\) but also on a delayed version of the rest of y, but this is beyond the scope of our setup.

  2. Actually, now we only need the system to be centrally stabilizable and detectable.

  3. The set \(\text{ sp } (A + B K C_2 )\) is the set of eigenvalues of \(A + B K C_2\).

  4. We probe in channel i during \([T_i , T_{i+1} )\).

  5. We probe with elements of \(\sqcap _i [k+1]\) during \([\tilde{T}_i, \; \tilde{T}_{i+1} )\).

  6. One plant output is \(\mathcal {F}(P,K_\mathrm{cen}) r\) and the other plant output is \(\mathcal {F}(P,K_\mathrm{dec} (T)) r\).

References

  1. Aghdam AG, Davison EJ (2006) Structural modification of systems using discretization and generalized sampled-data holds. Automatica 42:1935–1941

    Article  MathSciNet  MATH  Google Scholar 

  2. Anderson BDO, Moore JB (1981) Time-varying feedback laws for decentralized control. IEEE Trans Autom Control 26:1133–1139

    Article  MathSciNet  MATH  Google Scholar 

  3. Chen T, Francis BA (1995) Optimal sampled-data control systems. Springer, London

    Book  MATH  Google Scholar 

  4. Corfmat JP, Morse AS (1976) Decentralized control of linear multivariable systems. Automatica 12:479–495

    Article  MathSciNet  MATH  Google Scholar 

  5. Davison EJ, Zhang H (1995) Design of low order compensators for large scale decentralized systems. In: IFAC symposium on large scale systems: theory and applications. London, pp 509–514

  6. Davison EJ, Wang SH (1973) On the stabilization of decentralized control systems. IEEE Trans Autom Control 18:473–478

    Article  MathSciNet  MATH  Google Scholar 

  7. Doyle JC, Glover K, Khargonekar PP, Francis BA (1989) State-space solutions to standard \({\cal{H}}_2 \) and \({\cal{H}}_{\infty } \) control problems. IEEE Trans Autom Control 34:831–847

    Article  MATH  Google Scholar 

  8. Gong Z, Aldeen M (1997) Stabilization of decentralized control systems. Math Syst Estim Control 7:1–16

    MathSciNet  MATH  Google Scholar 

  9. Khargonekar PP, Ozguler AB (1994) Decentralized control and periodic feedback. IEEE Trans Autom Control 39:877–882

    Article  MathSciNet  MATH  Google Scholar 

  10. Kobayashi L, Yoshikawa T (1982) Graph-theoretic approach to controllability and localizability of decentralized control systems. IEEE Trans Autom Control 27:1096–1108

    Article  MathSciNet  MATH  Google Scholar 

  11. Lessard L, Lall S (2011) A state-space solution to the two player decentralized optimal control problem. In: 2011 Allerton conference on communications, control and computing

  12. Miller DE (2006) Near optimal LQR performance for a compact set of plants. IEEE Trans Autom Control 51:1423–1439

    Article  MathSciNet  MATH  Google Scholar 

  13. Miller DE (2016) A comparison of LQR optimal performance in the decentralized and centralized settings. IEEE Trans Autom Control 61:2308–2311

    Article  MathSciNet  MATH  Google Scholar 

  14. Miller DE, Chen T (2002) Simultaneous stabilization with near optimal \(H_{\infty }\) performance. IEEE Trans Autom Control 47:1986–1998

    Article  MathSciNet  MATH  Google Scholar 

  15. Miller DE, Davison EJ (2012) An algebraic characterization of quotient decentralized fixed modes. Automatica 48:1639–1644

    Article  MathSciNet  MATH  Google Scholar 

  16. Miller DE, Davison EJ (2014) Near optimal LQR control in the decentralized setting. IEEE Trans Autom Control 59:327–340

    Article  MATH  Google Scholar 

  17. Miller DE, Qiu L (2005) Near optimal \(H_{\infty }\) tracking for a compact set of plants. In: Proceedings of the IEEE 2005 conference on decision and control, Seville, Spain

  18. Ozguner U, Davison EJ (1985) Sampling and decentralized control systems. In: Proceedings of the American control conference, pp 257–262

  19. Qi X, Salapaka MV, Voulgaris PG, Khammash M (2004) Structured optimal and robust control with multiple criteria: a convex solution. IEEE Trans Autom Control 49:1623–1640

    Article  MathSciNet  MATH  Google Scholar 

  20. Rotkowitz M, Lall S (2006) A characterization of convex problems in decentralized control. IEEE Trans Autom Control 51:274–286

    Article  MathSciNet  MATH  Google Scholar 

  21. Ranganathan T, Miller DE (2015) Near optimal \({\cal{H}}_{\infty }\) performance in the decentralized setting. In: 2015 IEEE conference on control applications, Sydney, Australia, pp 1823–1828

  22. Shah P, Parrilo PA (2010) \(H_2\)-optimal decentralized control over posets: a state space solution for state-feedback. In: 49th IEEE conference on decision and control, Atlanta, Georgia, pp 6722–6727

  23. Swigart J, Lall S (2010) An explicit state-space solution for a decentralized two-player optimal linear-quadratic regulator. In: 2010 American control conference, pp 6385–6390

  24. Voulgaris PG (1999) Control under structural constraints: an input–output approach. In: Lecture notes in control and information sciences, pp 287–305

  25. Vale J, Miller DE (2008) Tracking in the presence of a time-varying uncertain gain. In: Proceedings of the 2008 Allerton conference on communication, control, and computing, Allerton, Illinois

  26. Wang SH (1982) Stabilization of decentralized control systems via time-varying controllers. IEEE Trans Autom Control 27:741–749

    Article  MATH  Google Scholar 

  27. Willems JL (1989) Time-varying feedback for the stabilization of fixed modes in decentralized control. Automatica 25:127–131

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by the Natural Sciences and Engineering Research Council of Canada via a research grant.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel E. Miller.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

Proof of Lemma 1

Fix \(\bar{n}\in \mathbf{N}\) and \(\tilde{h}\in (0,1)\). Let \(t_0\in \mathbf{R}\), \(x_0\in \mathbf{R}\), \(h\in (0,\tilde{h})\), \(\bar{u}\in \mathbf{R}\) and \(\phi \in \mathbf{R}\) be arbitrary, and choose u(t) as indicated. If we solve the plant equation (1) and use the Cauchy–Schwarz inequality, then we obtain

$$\begin{aligned} \left\| x(t)-x_0 \right\|\le & {} \left\| ({\mathrm {e}}^{A(t-t_0)}-I)x_0 \right\| + \int _{t_0}^{t}\left\| e^{A(t-\tau )}B \right\| (\left\| \bar{u} \right\| +|\phi |) \mathrm{d} \tau \\&+\, \left( \int _{t_0}^{t}\left\| e^{A(t-\tau )}E \right\| ^2\mathrm{d}\tau \right) ^{\frac{1}{2}}\Vert r_{[ t_0 , t ]} \Vert _2 , \;\; t \in [ t_0, t_0 + 2 \bar{n}h ) . \end{aligned}$$

Clearly there exists a constant \(\gamma _1>0\) such that

$$\begin{aligned} \left\| x(t)-x_0 \right\| \le \gamma _1h(\left\| x_0 \right\| +\left\| \bar{u} \right\| +|\phi |) + \gamma _1h^{\frac{1}{2}}\left\| r_{[t_0,t_0+2\bar{n}h]} \right\| _2,\;\; t\in [t_0,t_0+2\bar{n}h), \end{aligned}$$
(60)

which yields the second bound.

Using the fact that \(\bar{y}_i(t) = v_i^{\mathrm {T}}y_i(t)\) for \(t\in [t_0,t_0+\bar{n}h)\) we have

$$\begin{aligned} \bar{y}_i(t)= & {} v_i^{\mathrm {T}}C_2^i \left[ e^{A(t-t_0)}x_0+\int _0^{t-t_0}e^{A \tau }B ( \bar{u} + \bar{w}_j \phi ) \mathrm{d}\tau +\int _{t_0}^{t}e^{A(t-\tau )}Er(\tau ){\mathrm {d}\tau }\right] \\= & {} \sum _{k=0}^{\bar{n}}\frac{\bar{C}_2^iA^k(t-t_0)^k}{k!}x_0 +{\mathcal {O}}(h^{\bar{n}+1})x_0 + \sum _{k=0}^{\bar{n}-1}\frac{\bar{C}_2^iA^k(t-t_0)^{k+1}B}{(k+1)!}(\bar{u} +\bar{w}_j\phi )\\&+\,{\mathcal {O}}(h^{\bar{n}+1})(\bar{u}+\phi ) +\underbrace{\bar{C}_2^i\int _{t_0}^{t}e^{A(t-\tau )}Er(\tau )\mathrm{d}\tau }_{=: \mu _1(t)}\\= & {} \left[ \begin{array}{cccc} 1&t-t_0&\cdots&(t-t_0)^{\bar{n}}/{\bar{n}!} \end{array} \right] \times \left\{ \underbrace{ \left[ \begin{array}{c} \bar{C}_2^i \\ \bar{C}_2^iA \\ \vdots \\ \bar{C}_2^iA^{\bar{n}} \end{array} \right] }_ {=: {\mathcal O}_i} x_0 + \underbrace{ \left[ \begin{array}{c} 0 \\ \bar{C}_2^iB \\ \vdots \\ \bar{C}_2^iA^{\bar{n}-1}B \end{array} \right] }_{=: M_i} (\bar{u} +\bar{w}_j\phi ) \right\} \\&+{\mathcal {O}}(h^{\bar{n}+1}) x_0+{\mathcal {O}}(h^{\bar{n}+1})(\bar{u}+\phi )+ \mu _1(t). \end{aligned}$$

Before we proceed further, let us get a bound on the term \( \mu _1(t)\). Using the Cauchy–Schwarz inequality, we have

$$\begin{aligned} \Vert {\mu _1(t)}\Vert \le \left( \int _{t_0}^{t_0 + \bar{n}h} \Vert \bar{C}_2^ie^{A(t-\tau )}E\Vert ^2\mathrm{d}\tau \right) ^{\frac{1}{2}}{\Vert r_{[t_0,t_0+\bar{n}h]}\Vert _2} , \;\; t \in [ t_0 , t_0 + \bar{n}h ] . \end{aligned}$$

Note that

$$\begin{aligned} rel.deg(\bar{C}_2^i(sI-A)^{-1}E) \ge rel.deg(C_2(sI-A)^{-1}E) = \eta _2, \end{aligned}$$

so we see that \(\bar{C}_2 A^i E = 0\) for \(i=0,1,\ldots , \eta _2-2\), so there exists a constant \(\gamma _2\) so that

$$\begin{aligned} \Vert \bar{C}_2^i e^{A \theta } E \Vert \le \gamma _2 \theta ^{\eta _2-1} , \; \theta \in [ 0 , 2 \bar{n}h ] , \end{aligned}$$

which means that there exists a constant \(\gamma _3\) so that

$$\begin{aligned} \left\| \mu _1(t) \right\| = \gamma _3 h^{\eta _2-{\frac{1}{2}}} {\Vert r_{[t_0,t_0+\bar{n}h]}\Vert _2}, \; t \in [ t_0 , t_0 + \bar{n}h ] . \end{aligned}$$

Now we sample \(\bar{y}_i\) in the interval of \([t_0, t_0+\bar{n}h)\) and form \(\bar{\mathcal Y}_i(t_0)\); we see that there exists a constant \(\gamma _4\) and a function \(\mu _2(h)\) so that

$$\begin{aligned} {{\bar{\mathcal Y}}}_i(t_0) = SH(h)[ {\mathcal O}_i x_0 + M_i (\bar{u} +\bar{w}_j\phi )]+ \mu _2(h) \end{aligned}$$

with

$$\begin{aligned} \left\| \mu _2(h) \right\|\le & {} \gamma _4 h^{\bar{n}+1}(\left\| x_0 \right\| +\left\| \bar{u} \right\| +|\phi |)+ \gamma _4 h^{\eta _2-{\frac{1}{2}}}\left\| r_{[t_0,t_0+\bar{n}h]} \right\| _2. \end{aligned}$$
(61)

If we analyse \(\bar{y}_i(t)\) for \(t\in [t_0+\bar{n}h, t_0+2\bar{n}h)\) in a similar fashion and form \({\bar{\mathcal Y}}_i(t_0+\bar{n}h)\), then we see that there exists a constant \(\gamma _5\) and a function \(\mu _3(h)\) so that

$$\begin{aligned} {\bar{\mathcal Y}}_i(t_0+\bar{n}h)= & {} SH(h)[ {\mathcal O}_i x(t_0+\bar{n}h) + M_i (\bar{u} -\bar{w}_j\phi ) ] + \mu _3(h) \end{aligned}$$

with

$$\begin{aligned} \left\| \mu _3(h) \right\|\le & {} \gamma _5 h^{\bar{n}+1}(\left\| x(t_0+\bar{n}h) \right\| +\left\| \bar{u} \right\| +|\phi |) + \gamma _5 h^{\eta _2-{\frac{1}{2}}}{\Vert r_{[t_0+\bar{n}h,t_0+2\bar{n}h]}\Vert _2}.\nonumber \\ \end{aligned}$$
(62)

Using the fact that \(\bar{B}_j = B\bar{w}_j\) and subtracting \({\bar{\mathcal Y}}_i(t_0+\bar{n}h)\) from \({\bar{\mathcal Y}}_i(t_0)\) we obtain

$$\begin{aligned}&{\bar{\mathcal Y}}_i(t_0)- {\bar{\mathcal Y}}_i(t_0+\bar{n}h)-2SH(h) \underbrace{M_i \bar{w}_j}_{= M_{i,j}} \phi \\&\quad =SH(h){\mathcal O}_i [ x_0 - x(t_0+\bar{n}h)] + \mu _2(h)-\mu _3(h) , \end{aligned}$$

so

$$\begin{aligned}&H(h)^{-1} S^{-1} [ {\bar{\mathcal Y}}_i(t_0)- {\bar{\mathcal Y}}_i(t_0+\bar{n}h) ] - 2 M_{i,j} \phi \\&\quad = {\mathcal O}_i [ x_0 - x(t_0+\bar{n}h)] + H(h)^{-1}S^{-1} [ \mu _2(h)-\mu _3(h) ] =: \mu _4 (h). \end{aligned}$$

We can form a bound on \(\left\| x_0 - x(t_0+\bar{n}h) \right\| \) using (60) and bounds on \(\mu _2(h)\) and \(\mu _3(h)\) using (61) and (62), respectively. Using the fact that \(\Vert H(h)^{-1} \Vert = \frac{ \bar{n}!}{h^{\bar{n}}}\) and \(S^{-1} = {\mathcal O} (1)\), we see that there exists a constant \(\gamma _6\) so that

$$\begin{aligned}&\left\| \mu _4(h) \right\| \le \gamma _6 h (\left\| x_0 \right\| +\left\| \bar{u} \right\| +|\phi |)+ \gamma _6 h^{{\frac{1}{2}}}\left\| r_{[t_0,t_0+2\bar{n}h]} \right\| _2 \\&\quad +\, \gamma _6 h^{\eta _2 - \bar{n}-{\frac{1}{2}}}\left\| r_{[t_0,t_0+2\bar{n}h]} \right\| _2, \end{aligned}$$

which provides the first bound. \(\square \)

Proof of Lemma 2

We will model the proof on that of Lemma 2 of [16]. Here our objective is to show the existence of a LPC and we will not attempt to find one of the lowest order.

Since the controller is LPC, it is sufficient to look at what happens on the interval \([kT,(k+1)T)\). In channel i, we partition the state into three sub-states: \(\psi _i^1\), \(\psi _i^2\) and \(\psi _i^3\). We use \(\psi _i^1\) of dimension \({ql_i}\) to store \(\{y_i(kT),y_i(kT+h),\ldots ,y_i(kT+(q-1)h)\}\). The second sub-state \(\psi _i^2\) of dimension \({m_i}\) stores \(\hat{\sqcap }_i[k]\). The last sub-state \(\psi _i^3\) of dimension \(\ell \) only comes into play in channel p: \(\psi _p^3\) stores \(\nu [k]\).

We limit our analysis to the interval of [0, T) as it can easily be extended due to the periodic nature of the controller. Let \(e_i \in \mathbf{R}^q\) denote the ith normal vector. We set

$$\begin{aligned} (L_i^{11},M_i^1)[j] =\left\{ \begin{array}{ll} (0,e_1\otimes I_{r_i}) &{} j=0\\ (I,e_{j+1}\otimes I_{r_i})&{} j=1,\ldots ,q-1.\\ \end{array}\right. \end{aligned}$$

It is clear that for \(j=0,1,\ldots ,q-1\), the vector \( \left[ \begin{array}{c} \psi _i^1[j] \\ y_i(jh) \end{array} \right] \) contains all the elements \(\{y_i(0),y_i(h),\ldots ,y_i(jh)\}\).

Now we turn to the second sub-state. Set

$$\begin{aligned} (L_i^{21},L_i^{22},M_i^2)[j] = \left\{ \begin{array}{ll} (0,I,0) &{} j=0,1,\ldots ,q-2\\ (L_i^{21}[q-1],0, \; M_i^2[q-1]) &{} j=q-1. \\ \end{array}\right. \end{aligned}$$

If we initialize \(\psi _i^2[0]\) to be \(\hat{\sqcap }_i [0]\), then we see that

$$\begin{aligned} \psi _i^2 [j] = \hat{\sqcap }_i [0] , \;\; j = 0,1,\ldots , q-1 . \end{aligned}$$
(63)

From the formula for \(\hat{\sqcap }_i [1]\) given in the controller description, we see that it is a weighted sum of the elements of \(\{ y_i (0) , \; y_i (h) ,\ldots , \; y_i ((q-1)h) \}\); using the fact that these elements are contained in \( \left[ \begin{array}{c} \psi _i^1[q-1] \\ y_i ( ( q-1)h) \end{array} \right] \), it follows that we can define \(L_i^{21}[q-1]\) and \(M_i^2[q-1]\) so that

$$\begin{aligned} \psi _i^2 [q]= & {} L_i^{21}[q-1] \psi _i^1 [ q-1 ] + M_i^2[q-1] y_i ( q-1)h) = \hat{\sqcap }_i [1] . \end{aligned}$$

We conclude that (16) holds.

Now we turn to the third sub-state. With \(q_p =2\bar{n}(l-l_p)\) we set \((L_i^{31},L_i^{33},M_i^3)[\cdot ] = 0\) for \(i =1,\;2,\ldots ,\;p-1\) and

$$\begin{aligned} (L_p^{31},L_p^{33},M_p^3)[j] = \left\{ \begin{array}{ll} (0,I,0) &{} j=0,1,\ldots ,q_p-1\\ (L_p^{31}[q_p],\mathcal {F},M_p^3[q_p]) &{} j=q_p\\ (0,I,0) &{} j=q_p+1,\ldots ,q-1.\\ \end{array}\right. \end{aligned}$$

From the formula given for \(\hat{y}_i (kT)\) in the controller description, we see that \(\hat{y} (0)\) is a weighted sum of the elements of \(\{ y_p (0) , \; y_p (h) , \cdots , \; y_p (q_p h) \} \); using the fact that these elements are contained in \( \left[ \begin{array}{c} \psi _p^1 [q_p ] \\ y_p ( q_p h ) \end{array} \right] \), we can choose \(L_p^{31}[q_p]\) and \(M_p^3[q_p]\) such that

$$\begin{aligned} L_p^{31}[q_p]\psi _p^1[q_p] + M_p^3[q_p]y_p(q_ph) = \hat{y}(0). \end{aligned}$$

So if we initialize \(\psi _p^3[0]=\nu [0]\), we see that

$$\begin{aligned} \psi _p^3[j]=\left\{ \begin{array}{ll} \nu [0],&{}j=0, 1, \cdots ,q_p\\ \nu [1],&{}j=q_p+1, \cdots , q-1 , \end{array}\right. \end{aligned}$$

as required.

At this point, we have defined \(L_i\) and \(M_i\) of \(K_\mathrm{dec}\). It remains to define the time-varying matrices \(Q_i\) and \(R_i\) related to the output. To this end, we define \(\phi _i[j]= \left[ \begin{array}{c} \psi _i[j] \\ y_i(jh) \end{array} \right] \). It is straightforward to verify that \(\phi _i[j]\) contains \(\{\hat{\sqcap }_i[0],y_i(0),\ldots ,y_i(jh)\}\) for \(j=0,1,\ldots , q-1\); for the case of \(i=p\), \(\phi _i[j]\) also contains

$$\begin{aligned} \left\{ \begin{array}{ll} \nu [0],&{}\quad j=0, 1, \ldots ,q_p\\ \nu [1],&{}\quad j=q_p+1, \ldots , q-1 . \end{array}\right. \end{aligned}$$
  1. (i)

    For \(j=0,\;1,\ldots ,\;q_p-1\), the control signal \(u_i(jh)\) equals \(\hat{\sqcap }_i[0]\) plus a linear combination of the elements of \(y_i(0)\). Since \(\hat{\sqcap }_i[0]\) and \(y_i(0)\) are contained in \(\phi _i [j]\), we see that \(u_i(jh)\) is a linear combination of the elements of \(\phi _i[j]\) for \(j=0,1,\ldots ,q_p-1\).

  2. (ii)

    At \(j=q_p\), \(u_i(j) = \hat{\sqcap }_i[0]\) and \(\hat{\sqcap }_i[0]\) is contained in \(\psi _i[j]\), so it is a linear combination of the elements of \(\phi _i(j)\).

  3. (iii)

    Let \(j \in \{ q_p+1,\,\ldots ,\;q-1 \}\). The control signal \(u_i(jh)\), \(i=1,\ldots ,p-1\), equals \(\hat{\sqcap }_i[0]\), which is contained in \(\psi _i[j]\), so it is a linear combination of the elements of \(\phi _i[j]\). The control signal \(u_p(jh)\) equals \(\hat{\sqcap }_p[0]\) plus a linear combination of the elements of \(\sqcap [1] = H \nu [1] + J \hat{y} (0) \); but \(\nu [1]\) lies in \(\phi _p [j]\) and \(\hat{y} (0)\) is a linear combination of the elements of \(\phi _p [j]\), so \(\sqcap [1]\) is a linear combination of the elements of \(\phi _p[j]\).

Using the fact that \(u(t) = u(jh)\) for \(t\in [jh,(j+1)h)\), \(j\in \mathbf{Z}^+ \), we conclude that for \(j=0,1,\ldots ,q-1\), the control signal \(u_i(jh)\) is a linear combination of the elements of \(\phi _i[j]\), which means that \(Q_i\) and \(R_i\) can be defined so that the controller (5) is identical to that of (9)–(14). \(\square \)

Proof of Lemma 3

Suppose that the controller (9)–(14) is applied to the plant (1), and let \(x_0\in \mathbf{R}^n\), \(\nu [0]\in \mathbf{R}^\ell \), \(\hat{\sqcap }[0]\in \mathbf{R}^m\), \(k\in \mathbf{Z}^+\) and \(T>0\) be arbitrary. With the estimate \(\hat{y}(kT)\) of y(kT) given by (10) we see that

$$\begin{aligned} \hat{y}(kT)-y(kT) = \left[ \begin{array}{c} \hat{y}_{1}(kT)-y_{1}(kT) \\ \hat{y}_{2}(kT)-y_{2}(kT) \\ \vdots \\ \hat{y}_{p-1}(kT)-y_{p-1}(kT) \\ 0 \end{array} \right] . \end{aligned}$$

We can use Lemma 1 to derive the estimation error of each individual quantity \([y_i (kT) ]_j \) for \(i\in \{1, 2, \ldots , p-1\}\) and \(j\in \{1, 2, \ldots , l_i\}\). More specifically, extending the error bound provided in (8) to the general case, we have

$$\begin{aligned} \left\| \hat{y}(kT)-y(kT) \right\|= & {} {\mathcal {O}}(T^{1-\delta })\max \limits _{\tau \in [kT,kT+2\bar{n}(l-l_p)h]}\left\| x(\tau ) \right\| +{\mathcal {O}}(T^{1-\delta })\left\| \hat{\sqcap }[k] \right\| \nonumber \\&+\, {\mathcal {O}}\left( T^{{\frac{1}{2}}-\delta }\right) \left\| r_{[kT,kT+2\bar{n}(l-l_p)h)} \right\| _2. \end{aligned}$$
(64)

We now obtain a bound on the first term of the RHS: for \( t \in [kT, kT + 2 \bar{n}( l - l_p) h )\), we have

$$\begin{aligned} x(t)= & {} {\mathrm {e}}^{A(t-kT)}x(kT)+\int _{kT}^{t}e^{A(t-\tau )}Bu(\tau )\mathrm{d}\tau + \int _{kT}^{t}e^{A(t-\tau )}Er(\tau )\mathrm{d}\tau , \end{aligned}$$
(65)

so it is easy to see that

$$\begin{aligned} \left\| x(t) \right\|= & {} {\mathcal {O}}(1)\left\| x(kT) \right\| + \left( \int _{kT}^t{\mathcal {O}}(1)^2\mathrm{d}\tau \right) ^{\frac{1}{2}}\left\| r_k \right\| _2 \\&+\, \int _{kT}^{t}[ {\mathcal {O}}(1)\left\| \hat{\sqcap }[k] \right\| +{\mathcal {O}}({T^\delta })\left\| x(kT) \right\| ] \mathrm{d}\tau \\= & {} {\mathcal {O}}(1)\left\| x(kT) \right\| +{\mathcal {O}}(T)\left\| \hat{\sqcap }[k] \right\| +{\mathcal {O}}\left( T^{{\frac{1}{2}}}\right) \left\| r_k \right\| _2. \end{aligned}$$

Using this we can simplify (64):

$$\begin{aligned} \left\| \hat{y}(kT)-y(kT) \right\| = {\mathcal O}( T^{1 - \delta } ) \Vert x(kT) \Vert + {\mathcal O}( T^{1 - \delta } ) \Vert \hat{\sqcap } [k] \Vert + {\mathcal {O}}\left( T^{{\frac{1}{2}}-\delta }\right) \left\| r_k \right\| _2 , \end{aligned}$$
(66)

which yields (21).

Now let us find a conservative bound on the signal u(t). Observe that

$$\begin{aligned} \left\| u(t)-\hat{\sqcap }[k] \right\|= & {} {\mathcal {O}}(T^{\delta })\left\| y(kT) \right\| +{\mathcal {O}}(T^{\delta })\left\| \sqcap [k+1] \right\| \\= & {} {\mathcal {O}}(T^{\delta })\left\| x(kT) \right\| +{\mathcal {O}}(T^{\delta })\left\| \nu [k+1] \right\| \\&+\,{\mathcal {O}}(T^{\delta })\left\| \hat{y}(kT) \right\| ,\;t\in [kT, (k+1)T). \end{aligned}$$

Using (66) to bound \(\hat{y}(kT)\) and the fact that \(\nu [k+1] = \nu [k]+{\mathcal {O}}(T)\nu [k]+{\mathcal {O}}(T)\hat{y}(kT)\), this simplifies to

$$\begin{aligned} \left\| u(t)-\hat{\sqcap }[k] \right\|= & {} {\mathcal {O}}(T^{\delta })\left\| x(kT) \right\| +{\mathcal {O}}(T^{\delta })\left\| \nu [k] \right\| + {\mathcal {O}}(T)\left\| \hat{\sqcap }[k] \right\| \nonumber \\&+\, {\mathcal {O}}\left( T^{{\frac{1}{2}}}\right) \left\| r_k \right\| _2, \; t\in [kT,(k+1)T), \end{aligned}$$
(67)

which yields (19).

Now we return to the state x(t). Using (67) in (65), we can obtain a bound on \(\Vert x(t) - x(kT) \Vert \):

$$\begin{aligned} \left\| x(t)-x(kT) \right\|= & {} {\mathcal {O}}(T)\left\| x(kT) \right\| + {\mathcal {O}}(T) \left\| \hat{\sqcap }[k] \right\| + {\mathcal {O}}(T^{1+\delta })\left\| \nu [k] \right\| \nonumber \\&+\, {\mathcal {O}}\left( T^{{\frac{1}{2}}}\right) \left\| r_k \right\| _2 \end{aligned}$$
(68)

for \(t \in [kT, (k+1) T ) ,\) which yields Eq. (18).

We now perform a similar analysis as above to obtain a bound on the quantity \(\hat{\sqcap }[k+1]-\sqcap [k+1]\). We first use Lemma 1 to obtain the estimation error of each element of \(\hat{\sqcap }_i[k+1]_j\) for \(i\in \{1,2,\ldots ,p-1\}\) and \(j\in \{1,2,\ldots ,m_i\}\). More specifically, we use the bound in (8) with a minor change reflecting the fact that we are now probing with \(T^\delta \bar{w}_p\sqcap _i[k+1]_j\) rather than \(T^\delta \bar{w}_i[y_i(kT)]_j\). So (8) becomes (for \(i=j=1\)):

$$\begin{aligned}&{\mathcal {O}}(T^{1-\delta })(\left\| x(kT+(2\bar{n}(l-l_p)+1)h) \right\| +\left\| \hat{\sqcap }[k] \right\| )\\&\quad + \,{\mathcal {O}}(T)|\sqcap _1[k+1]_1| + {\mathcal {O}}\left( T^{{\frac{1}{2}}-\delta }\right) \left\| r_k \right\| _2. \end{aligned}$$

As a result, we end up with

$$\begin{aligned} \left\| \hat{\sqcap }[k+1]-\sqcap [k+1] \right\|= & {} {\mathcal {O}}(T^{1-\delta })\left\| \hat{\sqcap }[k] \right\| \nonumber \\&+\, {\mathcal {O}}(T^{1-\delta })\max \limits _{\tau \in [kT+(2\bar{n}(l-l_p)+1)h),(k+1)T]}\left\| x(\tau ) \right\| \nonumber \\&+\,{\mathcal {O}}(T)\left\| \sqcap [k+1] \right\| +{\mathcal {O}}\left( T^{{\frac{1}{2}}-\delta }\right) \left\| r_k \right\| _2. \end{aligned}$$
(69)

But it is easy to see that

$$\begin{aligned} \sqcap [k+1] = {\mathcal {O}}(1)v[k]+{\mathcal {O}}(1)\hat{y}(kT). \end{aligned}$$

After we use the bound on \(\hat{y}(kT)\) given by (66) to simplify this, we substitute the resulting expression for \(\sqcap [k+1]\) into (69) and use (68) to obtain a bound on \(\max \limits _{\tau \in [kT,(k+1)T]}\left\| x(\tau ) \right\| \), yielding

$$\begin{aligned} \left\| \hat{\sqcap }[k+1]-\sqcap [k+1] \right\|= & {} {\mathcal {O}}(T^{1-\delta })\left\| x(kT) \right\| + {\mathcal {O}}(T^{1-\delta }) \left\| \hat{\sqcap }[k] \right\| \\&+\,{\mathcal {O}}(T)\left\| \nu [k] \right\| + {\mathcal {O}}\left( T^{{\frac{1}{2}}-\delta }\right) \left\| r_k \right\| _2, \end{aligned}$$

which yields (20). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ranganathan, T., Miller, D.E. Near optimal \(\mathcal {H}_{\infty }\) performance in the decentralized setting. Math. Control Signals Syst. 30, 12 (2018). https://doi.org/10.1007/s00498-018-0217-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00498-018-0217-1

Keywords

Navigation