Skip to main content

Data-Driven Stabilization of Nonlinear Systems via Taylor’s Expansion

  • Chapter
  • First Online:
Hybrid and Networked Dynamical Systems

Part of the book series: Lecture Notes in Control and Information Sciences ((LNCIS,volume 493))

  • 206 Accesses

Abstract

Lyapunov’s indirect method is one of the oldest and most popular approaches to model-based controller design for nonlinear systems. When the explicit model of the nonlinear system is unavailable for designing such a linear controller, finite-length off-line data is used to obtain a data-based representation of the closed-loop system, and a data-driven linear control law is designed to render the considered equilibrium locally asymptotically stable. This work presents a systematic approach for data-driven linear stabilizer design for continuous-time and discrete-time general nonlinear systems. Moreover, under mild conditions on the nonlinear dynamics, we show that the region of attraction of the resulting locally asymptotically stable closed-loop system can be estimated using data.

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 109.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 139.99
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. Ahmadi, A.A., El Khadir, B.: Learning dynamical systems with side information. In: Conference on Learning for Dynamics and Control, vol. 120, pp. 718–727 (2020)

    Google Scholar 

  2. Berberich, J., Köhler, J., Müller, M.A., Allgöwer, F.: Linear tracking MPC for nonlinear systems—part II: the data-driven case. IEEE Trans. Autom. Control 67(9), 4406–4421 (2022)

    Article  MathSciNet  Google Scholar 

  3. Bisoffi, A., De Persis, C., Tesi, P.: Trade-offs in learning controllers from noisy data. Syst. Control Lett. 154, 104985(1)–104985(8) (2021)

    Google Scholar 

  4. Bisoffi, A., De Persis, C., Tesi, P.: Data-driven control via Petersen’s lemma. Automatica 145, 110537(1)–110537(14) (2022)

    Google Scholar 

  5. Blanchini, F., Miani, S.: Set-Theoretic Methods in Control. In: Systems & Control: Foundations & Applications. Birkhäuser Boston (2008)

    Google Scholar 

  6. Bobiti, R., Lazar, M.: Automated sampling-based stability verification and DOA estimation for nonlinear systems. IEEE Trans. Autom. Control 63(11), 3659–3674 (2018)

    Article  MathSciNet  Google Scholar 

  7. Chesi, G.: Estimating the domain of attraction for non-polynomial systems via LMI optimizations. Automatica 45(6), 1536–1541 (2009)

    Article  MathSciNet  Google Scholar 

  8. Chesi, G.: LMI techniques for optimization over polynomials in control: a survey. IEEE Trans. Autom. Control 55(11), 2500–2510 (2010)

    Article  MathSciNet  Google Scholar 

  9. Colbert, B.K., Peet, M.M.: Using trajectory measurements to estimate the region of attraction of nonlinear systems. In: IEEE Conference on Decision and Control, pp. 2341–2347. Miami, FL, USA (2018)

    Google Scholar 

  10. Dai, T., Sznaier, M.: Nonlinear data-driven control via state-dependent representations. In: IEEE Conference on Decision and Control, pp. 5765–5770. Austin, TX, USA, 13–15 Dec 2021

    Google Scholar 

  11. De Persis, C., Tesi, P.: Formulas for data-driven control: stabilization, optimality, and robustness. IEEE Trans. Autom. Control 65(3), 909–924 (2020)

    Article  MathSciNet  Google Scholar 

  12. Folland, G.B.: Higher-order derivatives and Taylor’s formula in several variables

    Google Scholar 

  13. Fraile, L., Marchi, M., Tabuada, P.: Data-driven stabilization of SISO feedback linearizable systems (2021). arXiv:2003.14240

  14. Guo, M., De Persis, C., Tesi, P.: Data-driven stabilization of nonlinear polynomial systems with noisy data. IEEE Trans. Autom. Control 67(8), 4210–4217 (2022)

    Article  MathSciNet  Google Scholar 

  15. Guo, M., De Persis, C., Tesi, P.: Learning control for polynomial systems using sum of squares relaxations. In: IEEE Conference on Decision and Control, pp. 2436–2441. Jeju Island, Republic of Korea (2020)

    Google Scholar 

  16. Guo, M., De Persis, C., Tesi, P.: Data-based stabilizer design and closed-loop analysis of unknown general nonlinear systems (2022). arXiv:2209.01071

  17. Hou, Z., Wang, Z.: From model-based control to data-driven control: survey, classification and perspective. Inf. Sci. 235, 3–35 (2013)

    Article  MathSciNet  Google Scholar 

  18. Hou, Z., Xiong, S.: On model-free adaptive control and its stability analysis. IEEE Trans. Autom. Control 64(11), 4555–4569 (2019)

    Article  MathSciNet  Google Scholar 

  19. Kaiser, E., Kutz, J.N., Brunton, S.L.: Data-driven discovery of Koopman eigenfunctions for control. Mach. Learn.: Sci. Technol. 2(3), 035023(1)–035023(31) (2021)

    Google Scholar 

  20. Liu, W., Sun, J., Wang, G., Bullo, F., Chen, J.: Data-driven self-triggered control via trajectory prediction. IEEE Trans. Autom. Control 68(11), 6951–6958 (2023)

    Article  MathSciNet  Google Scholar 

  21. Luppi, A., De Persis, C., Tesi. P.: On data-driven stabilization of systems with nonlinearities satisfying quadratic constraints. Syst. Control Lett. 163, 105206(1)–105206(11) (2022)

    Google Scholar 

  22. Martin, T., Allgöwer, F.: Data-driven system analysis of nonlinear systems using polynomial approximation (2022). arXiv:2108.11298

  23. Nejati, A., Zhong, B., Caccamo, M., Zamani, M.: Data-driven controller synthesis of unknown nonlinear polynomial systems via control barrier certificates. In: Proceedings of The 4th Annual Learning for Dynamics and Control Conference, vol. 168, pp. 763–776 (2022)

    Google Scholar 

  24. De Persis, C., Rotulo, M., Tesi, P.: Learning controllers from data via approximate nonlinearity cancellation. IEEE Trans. Autom. Control 68(10), 6082–6097 (2023)

    Article  MathSciNet  Google Scholar 

  25. Petersen, I.R.: A stabilization algorithm for a class of uncertain linear systems. Syst. Control Lett. 8(4), 351–357 (1987)

    Article  MathSciNet  Google Scholar 

  26. Shcherbakov, P.S., Topunov, M.V.: Extensions of Petersen’s lemma on matrix uncertainty. IFAC Proc. Vol. 41(2), 11385–11390 (2008)

    Article  Google Scholar 

  27. Shen, Y., Bichuch, M., Mallada, E.: Model-free learning of regions of attraction via recurrent sets. In: IEEE Conference on Decision and Control, pp. 4714–4719. Cancún, Mexico (2022)

    Google Scholar 

  28. Stengle, G.: A nullstellensatz and a positivstellensatz in semialgebraic geometry. Math. Ann. 207, 87–97 (1974)

    Article  MathSciNet  Google Scholar 

  29. Strasser, R., Berberich, J., Allgöwer, F.: Data-driven control of nonlinear systems: beyond polynomial dynamics. In: IEEE Conference on Decision and Control, pp. 4344–4351. Austin, TX, USA (2021)

    Google Scholar 

  30. Tabuada, P., Loong Ma, W., Grizzle, J., Ames, A.D.: Data-driven control for feedback linearizable single-input systems. In: Proceedings of the 56th IEEE Conference on Decision and Control, pp. 6265–6270. Melbourne, VIC, Australia (2017)

    Google Scholar 

  31. Tan, W., Packard, A.: Stability region analysis using polynomial and composite polynomial Lyapunov functions and sum-of-squares programming. IEEE Trans. Autom. Control 53(2), 565–571 (2008)

    Article  MathSciNet  Google Scholar 

  32. Tanaskovic, M., Fagiano, L., Novara, C., Morari, M.: Data-driven control of nonlinear systems: an on-line direct approach. Automatica 75, 1–10 (2017)

    Article  MathSciNet  Google Scholar 

  33. Topcu, U., Packard, A.: Local stability analysis for uncertain nonlinear systems. IEEE Trans. Autom. Control 54(5), 1042–1047 (2009)

    Article  MathSciNet  Google Scholar 

  34. van Waarde, H.J., Camlibel, M.K.: A matrix Finsler’s lemma with applications to data-driven control. In: Proceedings of the 60th IEEE Conference on Decision and Control, pp. 5777–5782. Austin, TX, USA (2021)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Meichen Guo .

Editor information

Editors and Affiliations

Appendix

Appendix

1.1 Petersen’s Lemma

In the section for data-driven controller design, Petersen’s lemma is essential for deriving the sufficient condition characterizing the controller. Due to the space limit, the proof of the lemma is omitted and one may refer to works such as [4, 25, 26] for more details.

Lemma 12.8.1

(Petersen’ s lemma [25]) Consider matrices \(\mathcal G=\mathcal G^{\top }\in \mathbb {R}^{n\times n}\), \(\mathcal M\in \mathbb {R}^{n\times m}\), \(\mathcal M\ne 0\), \(\mathcal N \in \mathbb {R}^{p\times n}\), \(\mathcal N \ne 0\), and a set F defined as

$$\begin{aligned} F =\{\mathcal F\in \mathbb {R}^{m\times p}: \mathcal F^{\top } \mathcal F\preceq \overline{\mathcal F} \}, \end{aligned}$$

where \(\overline{\mathcal F}=\overline{\mathcal F}^{\top }\succeq 0\). Then, for all \(\mathcal F\in F\),

$$\begin{aligned} \mathcal G + \mathcal M \mathcal F \mathcal N + \mathcal N^{\top } \mathcal F^{\top } \mathcal M^{\top } \preceq 0 \end{aligned}$$

if and only if there exists \(\mu >0\) such that

$$\begin{aligned} \mathcal G + \mu \mathcal M \mathcal M^{\top } +\mu ^{-1}\mathcal N^{\top }\overline{\mathcal F}\mathcal N \preceq 0. \end{aligned}$$

1.2 Proof of Lemma 12.5.3

Let \(z=\begin{bmatrix} x^{\top } & u^{\top } \end{bmatrix}^{\top }\). Each element \(f_i(z)\), \(i=1,\dots ,n\) in f(z) can be written as

$$\begin{aligned} f_i(z)= \sum ^{n+m}_{j=1}\frac{\partial f_i}{\partial z_j}(0)z_j+R_i(z). \end{aligned}$$

On the other hand, as shown in [12], the function \(f_i(z)\) can be expressed as

$$\begin{aligned} f_i(z)=\sum ^{n+m}_{j=1}z_j\int _{0}^{1} \frac{\partial f_i}{\partial z_j}(tz)dt. \end{aligned}$$

As a consequence, one can write \(R_i(z)\) as

$$\begin{aligned} R_i(z) & =& \sum _{j=1}^{n+m}z_j \int ^1_0 \frac{\partial f_i}{\partial z_j}(tz) dt - \sum ^{n+m}_{j=1}\frac{\partial f_i}{\partial z_j}(0)z_j\\ & =& \sum _{j=1}^{n+m}z_j \int ^1_0 \left( \frac{\partial f_i}{\partial z_j}(tz) - \frac{\partial f_i}{\partial z_j}(0) \right) dt. \end{aligned}$$

Under Assumption 12.5.2, one has

$$\begin{aligned} \left| \frac{\partial f_i}{\partial z_j}(tz) - \frac{\partial f_i}{\partial z_j}(0) \right| \le L_i\Vert tz\Vert ,~~t\in (0,1). \end{aligned}$$

Then, it holds that

$$\begin{aligned} \left| R_i(z) \right| &\le & \sum _{j=1}^{n+m}|z_j|\int ^1_0 L_i\Vert z\Vert \cdot |t|dt \\ & =& L_i\Vert z\Vert \int ^1_0 |t|dt \cdot \sum _{j=1}^{n+m}|z_j| . \end{aligned}$$

By the fact that \(\int ^1_0 |t|dt=\frac{1}{2}\) and \(|z_1+\cdots z_{n+m}|\le \sqrt{n+m}\Vert z\Vert \), it holds that

$$\begin{aligned} \left| R_i(z) \right| \le \frac{\sqrt{n+m} L_i}{2}\Vert z\Vert ^2. \end{aligned}$$

The proof is complete.

1.3 Proof of Lemma 12.5.6

For the closed-loop system with the controller \(u=Kx\) designed via Theorem 12.4.1, the derivative of the Lyapunov function \(V(x)=x^{\top }P^{-1}x\) satisfies

$$\begin{aligned} \dot{V}(x) = & {} x^{\top }P^{-1}(A+BK)x + x^{\top }(A+BK)^{\top }P^{-1}x+2x^{\top }P^{-1}R(x,Kx) \\ &\le & -wx^{\top }P^{-1}x +2x^{\top }P^{-1}R(x,Kx). \end{aligned}$$

Under Assumption 12.5.2, for all \(x\in \mathbb D\) and \(i=1,\dots ,n\), the bounds on the approximation error can be found as

$$\begin{aligned} |R_i(x,Kx)| \le \frac{\sqrt{m+n} L_i}{2} \Vert (x,Kx)\Vert ^{2}. \end{aligned}$$

Hence, for all \(x\in \mathbb D\), there exists a continuous \(\rho _i(x)\) for each \(i=1,\dots ,n\) such that for each \(x\in \mathbb D\)

$$\begin{aligned} R_i(x,Kx) = & {} \rho _i(x) \Vert (x,Kx)\Vert ^{2}, \\ \rho _i(x)&\in & \left[ -\frac{\sqrt{m+n} L_i}{2}, \frac{\sqrt{m+n} L_i}{2}\right] . \end{aligned}$$

Define \(\rho (x) = [\rho _1(x)~\dots ~\rho _n(x)]^{\top }\). By the definition of polytope [5, Definition 3.21], the vector \(\rho (x)\) belongs to the polytope

$$\begin{aligned} \mathcal H= \{ \varrho : -\bar{h} \preceq \varrho \preceq \bar{h} \}, \end{aligned}$$

where

$$\begin{aligned} \bar{h}=[\bar{h}_1~ \cdots ~\bar{h}_n]^{\top } = \begin{bmatrix} \frac{\sqrt{m+n} L_1}{2} &\cdots & \frac{\sqrt{m+n} L_n}{2} \end{bmatrix}^{\top }. \end{aligned}$$

Denote \(Q_i\) as the ith column of \(P^{-1}\). It holds that

$$\begin{aligned} {} & {} 2x^{\top }P^{-1}R(x,Kx) \\ = & {} 2\begin{bmatrix} x^{\top }Q_1 & \cdots & x^{\top }Q_n \end{bmatrix} \begin{bmatrix} \rho _1(x)\Vert (x,Kx)\Vert ^{2} \\ \vdots \\ \rho _n(x)\Vert (x,Kx)\Vert ^{2} \end{bmatrix}\\ = & {} 2\sum ^n_{i=1} x^{\top }Q_i\rho _i(x)\Vert (x,Kx)\Vert ^{2}\\ = & {} 2\sum ^n_{i=1} x^{\top }Q_i\Vert (x,Kx)\Vert ^{2}\cdot \rho _i(x)\\ & =& 2\begin{bmatrix} x^{\top }Q_1\Vert (x,Kx)\Vert ^{2} & \cdots & x^{\top }Q_n\Vert (x,Kx)\Vert ^{2} \end{bmatrix}\rho (x). \end{aligned}$$

Denote

$$\begin{aligned} \kappa (x)=\begin{bmatrix} x^{\top }Q_1\Vert (x,Kx)\Vert ^{2} & \cdots & x^{\top }Q_n\Vert (x,Kx)\Vert ^{2} \end{bmatrix}. \end{aligned}$$

Then, the derivative of the Lyapunov function satisfies for all \(x\in \mathbb D\)

$$\begin{aligned} \dot{V}(x) \le -wx^{\top }P^{-1}x + 2\kappa (x)\rho (x), \end{aligned}$$

where \(\rho (x)\in \mathcal H\).

1.4 Sum of Squares Relaxation

As solving positive conditions of multivariable polynomials is in general NP-hard, the SOS relaxations are often used to obtain sufficient conditions that are tractable. The SOS polynomial matrices are defined as follows.

Definition 12.8.2

(SOS polynomial matrix [8]) \(M:\mathbb {R}^n\rightarrow \mathbb {R}^{\sigma \times \sigma }\) is an SOS polynomial matrix if there exist \(M_{1},\dots ,M_{k}:\mathbb {R}^n\rightarrow \mathbb {R}^{\sigma \times \sigma }\) such that

$$\begin{aligned} M(x)=\sum ^{k}_{i=1}M_{i}(x)^{\top }M_{i}(x)~~\forall x\in \mathbb {R}^n. \end{aligned}$$
(12.28)

Note that when \(\sigma =1\), M(x) becomes a scalar SOS polynomial.

It is straightforward to see that if a matrix M(x) is an SOS polynomial matrix, then it is positive semi-definite, i.e., \(M(x)\succeq 0~\forall x\in \mathbb {R}^n\). Relaxing the positive polynomial conditions into SOS polynomial conditions makes the conditions tractable and easily solvable by common software.

1.5 Positivstellensatz

In the RoA analysis, we need to characterize polynomials that are positive on a semialgebraic set, and the Positivstellensatz plays an important role in this characterization.

Let \(p_1,\dots ,p_k\) be polynomials. The multiplicative monoid, denoted by \(\mathcal S_M(p_1,\dots ,p_k)\), is the set generated by taking finite products of the polynomials \(p_1,\dots ,p_k\). The cone \(\mathcal S_C(p_1,\dots ,p_k)\) generated by the polynomials is defined as

$$\begin{aligned} {} & {} \mathcal S_C(p_1,\dots ,p_k)\\ & =&\{ s_0 +\sum _{i=1}^j s_iq_i:s_0,\dots ,s_j \text { are SOS polynomials}, q_1,.\dots ,q_j\in \mathcal S_M(p_1,\dots ,p_k) \}. \end{aligned}$$

The ideal \(\mathcal S_I(p_1,\dots ,p_k)\) generated by the polynomials is defined as

$$\begin{aligned} \mathcal S_I(p_1,\dots ,p_k)= \left\{ \sum _{i=1}^k r_ip_i: r_1,\dots ,r_k \text { are polynomials} \right\} . \end{aligned}$$

Stengle’s Positivstellensatz [28] is presented as follows in [8].

Theorem 12.8.3

(Positivstellensatz) Let \(f_1,\dots ,f_k\), \(g_1,\dots ,g_l\), and \(h_1,\dots ,h_m\) be polynomials. Define the set

$$\begin{aligned} \mathcal X = \{x\in \mathbb {R}^n: {} & {} f_1(x)\ge 0, \dots , f_k(x)\ge 0, \\ {} & {} g_1(x)= 0, \dots , g_l(x)= 0,\\ {} & {} \text {and } h_1(x)\ne 0, \dots , h_m(x)\ne 0 \}. \end{aligned}$$

Then, \(\mathcal X = \emptyset \) if and only if

$$\begin{aligned} \exists f \in \mathcal S_C(f_1,\dots ,f_k),~~g \in \mathcal S_I(g_1,\dots ,g_l),~~h \in \mathcal S_M(h_1,\dots ,h_m) \end{aligned}$$

such that

$$\begin{aligned} f(x)+g(x)+h(x)^2=0. \end{aligned}$$

For the subsequent RoA analysis, we will use the following result derived from the Positivstellensatz.

Lemma 12.8.4

Let \(\varphi _1\) and \(\varphi _2\) be polynomials in x. If there exist SOS polynomials \(s_1\) and \(s_2\) in x such that

$$\begin{aligned} -(s_1\varphi _1(x)+s_2\varphi _2(x) +x^{{{\tiny \mathsf T}}}x) \text { is SOS }~\forall x\in \mathbb {R}^n \end{aligned}$$
(12.29)

then the set inclusion condition

$$\begin{aligned} \{x\in \mathbb {R}^n: \varphi _1(x)\ge 0, x\ne 0\} \subseteq \{ x\in \mathbb {R}^n: \varphi _2(x)<0 \} \end{aligned}$$
(12.30)

holds.

Proof

The set inclusion condition (12.30) can be equivalently written as

$$\begin{aligned} \{ x\in \mathbb {R}^n: \varphi _1(x)\ge 0, \varphi _2(x)\ge 0, x\ne 0\} = \emptyset . \end{aligned}$$

By Theorem 12.8.3, we know that this is true if and only if there exist \(\varphi (x)\in \mathcal S_C(\varphi _1,\varphi _2)\) and \(\zeta (x)\in \mathcal S_M(x)\), such that

$$\begin{aligned} \varphi (x) + \zeta (x)^2=0. \end{aligned}$$
(12.31)

Let

$$\begin{aligned} \varphi =s_0+s_1\varphi _1+s_2\varphi _2, \end{aligned}$$

where \(s_j\), \(j=0,1,2\) are SOS polynomials. By the definition of the cone \(\mathcal S_C\), one has that \(\varphi \in \mathcal S_C(\varphi _1,\varphi _2)\). Choosing \(\zeta (x)^2=x^{{{\tiny \mathsf T}}}x\), we write the condition (12.31) as

$$\begin{aligned} s_0+s_1\varphi _1+s_2\varphi _2 +x^{{{\tiny \mathsf T}}}x=0. \end{aligned}$$
(12.32)

As \(s_0=-(s_1\varphi _1+s_2\varphi _2 +x^{{{\tiny \mathsf T}}}x)\) from (12.32), if there exist SOS polynomials \(s_1\) and \(s_2\) such that the SOS condition (12.29) holds, then there exist SOS polynomials \(s_j\), \(j=0,1,2\) such that (12.32) is true, and hence the set inclusion condition (12.30) holds.    \(\square \)

1.6 Proof of Lemma 12.5.8

For the closed-loop system with the controller \(u=Kx\) designed via Theorem 12.4.2, the difference between the Lyapunov functions \(V(x^{+})=(x^{+})^{\top }P^{-1}x^{+}\) and \(V(x)=x^{\top }P^{-1}x\) is

$$\begin{aligned} {} & {} V(x^{+})-V(x) \\ = & {} \left[ (A+BK)x+R(x,Kx) \right] ^{\top } P^{-1} [ (A+BK)x +R(x,Kx) ]-x^{\top } P^{-1} x \\ = & {} x^{\top } \left[ (A+BK)^{\top }P^{-1} (A+BK)-P^{-1} \right] x +2R(x,Kx)^{\top }P^{-1}(A+BK)x \\ {} & {} +R(x,Kx)^{\top } P^{-1}R(x,Kx). \end{aligned}$$

Observe that

$$\begin{aligned} A+BK = \left( -\overline{{\textbf {A}}}^{-1}\overline{{\textbf {B}}} + \overline{{\textbf {A}}}^{-1/2}\Delta \right) ^{\top } \begin{bmatrix} K\\ I \end{bmatrix} \end{aligned}$$

with \(\Delta \Delta ^{\top }\preceq \delta I\). Then, it holds that

$$\begin{aligned} \left\| P^{-1}(A+BK) \right\| \le \left\| - P^{-1}\overline{{\textbf {B}}}^{\top }\overline{{\textbf {A}}}^{-1} \begin{bmatrix} K\\ I \end{bmatrix} \right\| + \sqrt{\delta } \left\| P^{-1} \right\| \left\| \overline{{\textbf {A}}}^{-1/2} \begin{bmatrix} K\\ I \end{bmatrix} \right\| =r_1 . \end{aligned}$$

For any \(\varepsilon >0\), it holds that

$$\begin{aligned} {} & {} 2R(x,Kx)^{\top }P^{-1}(A+BK)x\\ &\le & \varepsilon R(x,Kx)^{\top }R(x,Kx) +\varepsilon ^{-1} \left\| P^{-1}(A+BK) \right\| ^2 x^{\top }x\\ &\le &\varepsilon \Vert R(x,Kx)\Vert ^2 + \varepsilon ^{-1}r_1^2x^{\top }x. \end{aligned}$$

Recall that, by Theorem 12.4.2, \((A+BK)^{\top }P^{-1} (A+BK)-P^{-1}\preceq -w P^{-1}\). Hence, one has that

$$\begin{aligned} V(x^{+})-V(x) = -x^{\top }\left( w P^{-1} - \varepsilon ^{-1} r_1^2I \right) x + \left( \varepsilon + \Vert P^{-1}\Vert \right) \Vert R(x,Kx)\Vert ^2. \end{aligned}$$

Under Assumption 12.5.2,

$$\begin{aligned} \Vert R(x,Kx)\Vert ^2= \sum ^n_{i=1} R_i(x,Kx)^2 \le \sum ^n_{i=1}\frac{(m+n)L_i^2}{4} \Vert (x,Kx)\Vert ^4. \end{aligned}$$

If we write \(\Vert R(x,Kx)\Vert ^2\) as

$$\begin{aligned} \Vert R(x,Kx)\Vert ^2 = \begin{bmatrix} \Vert (x,Kx)\Vert ^{4} & \cdots & \Vert (x,Kx)\Vert ^{4} \end{bmatrix} \begin{bmatrix} \widehat{\rho }_1(x) \\ \vdots \\ \widehat{\rho }_n(x) \end{bmatrix}, \end{aligned}$$

then the scalars \(\widehat{\rho }_i(x)\) are such that \(\widehat{\rho }_i(x)\in \left[ 0,\frac{(m+n)L_i^2}{4} \right] \), \(i=1,\dots ,n\) for all \(x\in \mathbb D\). Defining

$$\begin{aligned} \widehat{\kappa }(x) = \begin{bmatrix} \Vert (x,Kx)\Vert ^{4} & \cdots & \Vert (x,Kx)\Vert ^{4} \end{bmatrix} ~~ \text { and }~~ \widehat{\rho }(x) = \begin{bmatrix} \widehat{\rho }_1(x) & \cdots & \widehat{\rho }_n(x) \end{bmatrix}^{\top } \end{aligned}$$

gives \(\Vert R(x,Kx)\Vert ^2 =\widehat{\kappa }(x)\widehat{\rho }(x)\), and for any \(x\in \mathbb D\) the vector \(\widehat{\rho }(x)\) is contained in the polytope \(\widehat{\mathcal H}\) defined in (12.26).

1.7 Dynamics Used for Data Generation in the Example

The dynamics used for data generation in Sect. 12.6 is the inverted pendulum written as

$$\begin{aligned} \dot{x}_1 &= x_2, \nonumber \\ \dot{x}_2 &= \frac{mgl}{J}\textrm{sin}(x_1) -\frac{r}{J}x_2 +\frac{l}{J}\textrm{cos}(x_1) u, \end{aligned}$$
(12.33)

where \(m=0.1\), \(g=9.8\), \(r=l=J=1\). For the discrete-time case, consider the Euler discretization of the inverted pendulum, i.e.,

$$\begin{aligned} x_1^{+} = & {} x_1 + T_sx_2,\\ x_2^{+} = & {} \frac{T_s g}{l} \sin (x_1) +\left( 1- \frac{T_s r}{ml^2} \right) x_2 + \frac{T_s}{ml^2} \cos (x_1) u, \end{aligned}$$

where \(m=0.1\), \(g=9.8\), \(T_s=0.1\), \(l=1\), and \(r=0.01\).

Rights and permissions

Reprints and permissions

Copyright information

© 2024 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Guo, M., De Persis, C., Tesi, P. (2024). Data-Driven Stabilization of Nonlinear Systems via Taylor’s Expansion. In: Postoyan, R., Frasca, P., Panteley, E., Zaccarian, L. (eds) Hybrid and Networked Dynamical Systems. Lecture Notes in Control and Information Sciences, vol 493. Springer, Cham. https://doi.org/10.1007/978-3-031-49555-7_12

Download citation

  • DOI: https://doi.org/10.1007/978-3-031-49555-7_12

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-031-49554-0

  • Online ISBN: 978-3-031-49555-7

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics