This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Data-driven MPC with stability guarantees using extended dynamic mode decompositionthanks: K. Worthmann gratefully acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 507037103

Lea Bold1, Lars Grüne2, Manuel Schaller1, and Karl Worthmann1
(1Optimization-based Control Group, Technische Universität Ilmenau, Germany
2Chair of Applied Mathematics, University of Bayreuth, Germany
July 2024 )

Abstract: For nonlinear (control) systems, extended dynamic mode decomposition (EDMD) is a popular method to obtain data-driven surrogate models. Its theoretical foundation is the Koopman framework, in which one propagates observable functions of the state to obtain a linear representation in an infinite-dimensional space. In this work, we prove practical asymptotic stability of a (controlled) equilibrium for EDMD-based model predictive control, in which the optimization step is conducted using the data-based surrogate model. To this end, we derive novel bounds on the estimation error that are proportional to the norm of state and control. This enables us to show that, if the underlying system is cost controllable, this stabilizablility property is preserved. We conduct numerical simulations illustrating the proven practical asymptotic stability.

1 Introduction

Model Predictive Control (MPC; [9]) is a well-established feedback control technique. In each iteration, an optimal control problem is solved, and a first portion of the optimal control is applied [4]. This process is then repeated at the successor time instant after measuring (or estimating) the resulting state of the system. The popularity of MPC is mainly due to its solid mathematical foundation and the ability to cope with nonlinear constrained multi-input systems. In the optimization step, it is, however, necessary to predict the cost functional and/or constraints along the flow of the underlying system, which requires a model, e.g., based on first principles.

Due to recent progress in data-driven methods, there are several works considering MPC and other model-based controllers using data-driven surrogate models. A popular approach is based on extended dynamic mode decomposition (EDMD [37]) as an approximation technique in the Koopman framework. The key idea is to lift a nonlinear (control) system to a linear, but infinite-dimensional one and, then, employ EDMD to generate a data-driven finite-dimensional approximation [24]. Convergence of EDMD in the infinite-data limit was shown in [14]. Generally speaking, the Koopman framework can be utilized for data-driven predictions of so-called observables (quantities of interest, e.g., the stage cost in MPC) along the flow of the dynamical (control) system. For control systems there are two popular approaches: The first seeks a linear surrogate and is widely called (e)DMDc [28, 13]. The second approach yields a bi-linear representation [36] and performs particularly well for systems with direct state-control coupling. For this approach also finite-data error bounds for ordinary and stochastic differential equations with i.i.d. and ergodic sampling were recently shown in [30, 23].

In [19], an LQR-based approach to control unconstrained systems by means of a linear surrogate model using Taylor arguments is proposed. The performance was further assessed in [18] using a simulation study. Recently, robust control of bi-linear Koopman models with guarantees was proposed in [32] or, using Lyapunov-based arguments, in [31, 22]. However, without rigorously linking the analysis to verifiable error bounds. EDMD-based surrogate models were further applied in the prediction step of MPC [25, 13] and [40] for a robust tube-based approach. Simulation-based case studies can be found in [39] for Koopman-based MPC and in [12] for the bi-linear approach. Whereas many of the proposed approaches are shown to perform well in examples, no rigorous guarantees for closed-loop stability of Koopman-based MPC are given.

The main contribution of this work is threefold. First, we propose and prove novel error bounds, which are proportional to the distance from the desired set point rather than uniform in the state, building upon the error bounds derived in [23]. Second, we show that cost controllability (roughly speaking asymptotic null controllability in terms of the stage costs, see [4] for details), i.e., a key property to establish asymptotic stability in MPC without terminal conditions, is preserved using the EDMD-based surrogat. Third, we establish semi-global practical asymptotic stability of the original system if the feedback law is computed using the data-driven surrogate model only. To this end, we recall a key result from [9] on practical asymptotic stability for numerical approximations and verify the assumptions based on the novel proportional error bounds and the maintained cost controllability.

The manuscript is organized as follows. In Section 2, we recap EDMD within the Koopman framework. Then, we introduce MPC, before we derive the novel proportional error bound and provide the problem formulation. In Section 4, we present our main results, i.e., the preservation of cost controllability for the EDMD-based surrogate and practical asymptotic stability of the EDMD-based MPC closed loop. Then, we illustrate our findings by means of a simulation study. Finally, conclusions are drawn in Section 6.

Notation: We use the following comparison functions: α𝒞(0,0)\alpha\in\mathcal{C}(\mathbb{R}_{\geq 0},\mathbb{R}_{\geq 0}) is said to be of class 𝒦\mathcal{K} if it is strictly increasing with α(0)=0\alpha(0)=0 and of class 𝒦\mathcal{K}_{\infty} if it, in addition, grows unboundedly. A function δ𝒞(0,0)\delta\in\mathcal{C}(\mathbb{R}_{\geq 0},\mathbb{R}_{\geq 0}) is of class \mathcal{L} if it is strictly decreasing with limtδ(t)=0\lim_{t\to\infty}\delta(t)=0. Moreover, β𝒞(02,0)\beta\in\mathcal{C}(\mathbb{R}_{\geq 0}^{2},\mathbb{R}_{\geq 0}) is said to be of class 𝒦\mathcal{KL} if β(,t)𝒦\beta(\cdot,t)\in\mathcal{K} and β(r,)\beta(r,\cdot)\in\mathcal{L} hold. For integers nmn\leq m, we set [n:m]:=[n,m][n:m]:=[n,m]\cap\mathbb{Z}. The ii-th standard unit vector in n\mathbb{R}^{n} is denoted by eie_{i}, i[1:n]i\in[1:n]. For a matrix A=(aij)n×mA=(a_{ij})\in\mathbb{R}^{n\times m}, AF2=i=1nj=1maij2\|A\|_{F}^{2}=\sum_{i=1}^{n}\sum_{j=1}^{m}a_{ij}^{2} denotes the squared Frobenius norm. For a set XX, we denote the interior by int(X)\operatorname{int}(X).

2 Koopman-based prediction and control

In this section, we recap the basics of surrogate modeling of nonlinear control systems within the Koopman framework. The underlying idea is to exploit an identity between the nonlinear flow and a linear, but infinite-dimensional operator. Then, a compression of this operator onto a finite-dimensional subspace is approximated by extended dynamic mode decomposition (EDMD) using finitely many samples of the system.

First, we consider the autonomous dynamical system governed by the nonlinear ordinary differential equation (ODE)

x˙(t)=g0(x(t)),\displaystyle\dot{x}(t)=g_{0}(x(t)), (2.1)

with locally-Lipschitz continuous map g0:nxnxg_{0}:\mathbb{R}^{n_{x}}\rightarrow\mathbb{R}^{n_{x}}. For initial condition x(0)=x^nxx(0)=\hat{x}\in\mathbb{R}^{n_{x}}, we denote the unique solution of System (2.1) at time t[0,)t\in[0,\infty) by x(t;x^)x(t;\hat{x}). We consider the ODE (2.1) on a compact and non-empty set 𝕏nx\mathbb{X}\subset\mathbb{R}^{n_{x}}. Then, to avoid technical difficulties in this introductory section, forward invariance of the set 𝕏\mathbb{X} w.r.t. the dynamics (2.1) is assumed, i.e., x(t;x^)𝕏x(t;\hat{x})\in\mathbb{X}, t0t\geq 0, holds for all x^𝕏\hat{x}\in\mathbb{X}. This may be ensured, e.g., by some inward-pointing condition and guarantees existence of the solution on [0,)[0,\infty). Then, the Koopman semigroup (𝒦t)t0(\mathcal{K}^{t})_{t\geq 0} of bounded linear operators is defined by the identity

(𝒦tφ)(x^)=φ(x(t;x^))t0,x^𝕏,φL2(𝕏,),(\mathcal{K}^{t}\varphi)(\hat{x})=\varphi(x(t;\hat{x}))\quad\forall\,t\geq 0,\hat{x}\in\mathbb{X},\varphi\in L^{2}(\mathbb{X},\mathbb{R}), (2.2)

see, e.g., [20, Prop. 2.4] or [17, Chapter 7]. Here, the real-valued functions φ\varphi are called observables. The identity (2.2) states that, instead of evaluating the observable φ\varphi at the solution of the nonlinear system (2.1) emanating from initial state x^\hat{x} at time tt, one may also apply the linear, infinite-dimensional Koopman operator 𝒦t\mathcal{K}^{t} to the observable φ\varphi and, then, evaluate 𝒦tφ\mathcal{K}^{t}\varphi at x^\hat{x}.

Since the flow of System (2.1) is continuous, (𝒦t)t0(\mathcal{K}^{t})_{t\geq 0} is a strongly-continuous semigroup of bounded linear operators. Correspondingly, we can define the, in general, unbounded infinitesimal generator \mathcal{L} of this semigroup by

φ:=limt0𝒦tφφtφD(),\displaystyle\mathcal{L}\varphi:=\lim_{t\searrow 0}\frac{\mathcal{K}^{t}\varphi-\varphi}{t}\qquad\forall\,\varphi\in D(\mathcal{L}), (2.3)

where the domain D()D(\mathcal{L}) consists of all L2L^{2}-functions, for which the above limit exists. Using this generator, we may formulate the equivalent evolution equation for Φ(t)=𝒦tφ=φ(x(t;))\Phi(t)=\mathcal{K}^{t}\varphi=\varphi(x(t;\cdot))

Φ˙(t)=Φ(t),Φ(0)=φ.\displaystyle\dot{\Phi}(t)=\mathcal{L}\Phi(t),\qquad\Phi(0)=\varphi. (2.4)

Next, we recap the extension of the Koopman approach to control-affine systems, i.e., systems governed by the dynamics

x˙(t)=g0(x(t))+i=1ncgi(x(t))ui(t),\dot{x}(t)=g_{0}(x(t))+\sum_{i=1}^{n_{c}}g_{i}(x(t))u_{i}(t), (2.5)

where the control function uLloc([0,),nc)u\in L^{\infty}_{\operatorname{loc}}([0,\infty),\mathbb{R}^{n_{c}}) serves as an input and the input maps gi:nxnxg_{i}:\mathbb{R}^{n_{x}}\to\mathbb{R}^{n_{x}}, i[0:nc]i\in[0:n_{c}], are locally Lipschitz continuous. A popular approach to obtain a data-based surrogate model is DMDc [28] or c [13], where one seeks a linear control system. In this paper, we pursue an alternative bi-linear approach, which exploits the control-affine structure of system (2.5) and was – to the best of our knowledge – proposed by [36, 35]. This approach shows a superior performance for systems with state-control coupling [2, 6]. For the flow of the control system (2.5) with constant control input uu, the Koopman operator 𝒦ut\mathcal{K}^{t}_{u} is defined analogously to (2.2). A straightforward computation shows that its generator preserves control affinity, i.e.,

u=0+i=1ncui(ei0)\displaystyle\mathcal{L}^{u}=\mathcal{L}^{0}+\sum\nolimits_{i=1}^{n_{c}}u_{i}(\mathcal{L}^{e_{i}}-\mathcal{L}^{0}) (2.6)

holds for uncu\in\mathbb{R}^{n_{c}}, where 0\mathcal{L}^{0} and ei\mathcal{L}^{e_{i}}, i[1:nc]i\in[1:n_{c}], are the generators of the Koopman semigroups corresponding to the constant controls u0u\equiv 0 and ueiu\equiv e_{i}, i[1:nc]i\in[1:n_{c}], respectively. For general control functions uLloc([0,),nc)u\in L^{\infty}_{\operatorname{loc}}([0,\infty),\mathbb{R}^{n_{c}}), one can now state the respective abstract Cauchy problem analogously to (2.4) replacing the generator \mathcal{L} by its time-varying counterpart u(t)\mathcal{L}^{u(t)} defined by (2.6), see [23] for details.

The success of the Koopman approach in recent years is due to its linear nature such that the compression of the Koopman operator or its generator (2.6) to a finite-dimensional subspace – called dictionary – leads to matrix representations. Being finite-dimensional objects, these matrices can then be approximated by a finite amount of data. Let the dictionary 𝕍:=span({ψk:k[1:M]})\mathbb{V}:=\operatorname{span}(\{\psi_{k}:k\in[1:M]\}) be the MM-dimensional subspace spanned by the chosen observables ψk\psi_{k}. We denote the L2L^{2}-orthogonal projection onto 𝕍\mathbb{V} by P𝕍P_{\mathbb{V}}. Further, using dd i.i.d. data points x1,,xd𝕏x_{1},\ldots,x_{d}\in\mathbb{X}, the (M×d)(M\times d)-matrices

X:=((ψ1(x1):ψM(x1))||(ψ1(xd):ψM(xd))) and Y:=(((0ψ1)(x1):(0ψM)(x1))||((0ψ1)(xd):(0ψM)(xd)))\displaystyle X:=\left(\left.\left(\begin{smallmatrix}\psi_{1}(x_{1})\\ :\\ \psi_{M}(x_{1})\end{smallmatrix}\right)\right|\ldots\left|\left(\begin{smallmatrix}\psi_{1}(x_{d})\\ :\\ \psi_{M}(x_{d})\end{smallmatrix}\right)\right.\right)\quad\text{ and }\quad Y:=\left(\left.\left(\begin{smallmatrix}(\mathcal{L}^{0}\psi_{1})(x_{1})\\ :\\ (\mathcal{L}^{0}\psi_{M})(x_{1})\end{smallmatrix}\right)\right|\ldots\left|\left(\begin{smallmatrix}(\mathcal{L}^{0}\psi_{1})(x_{d})\\ :\\ (\mathcal{L}^{0}\psi_{M})(x_{d})\end{smallmatrix}\right)\right.\right)

are defined, where (0ψk)(xj)=ψk(xj)g0(xj)(\mathcal{L}^{0}\psi_{k})(x_{j})=\nabla\psi_{k}(x_{j})^{\top}g_{0}(x_{j}) holds for k[1:M]k\in[1:M] and j[1:d]j\in[1:d]. Then, the empirical estimator of the compressed Koopman generator P𝕍0|𝕍P_{\mathbb{V}}\mathcal{L}^{0}|_{\mathbb{V}} is given by

d0:=argminLM×MLXYF2.\displaystyle{\mathcal{L}}^{0}_{d}:=\operatorname{arg}\min_{{L}\in\mathbb{R}^{M\times M}}\|{L}X-Y\|_{F}^{2}.

We have to repeat this step for ei\mathcal{L}^{e_{i}}, i[1:nc]i\in[1:n_{c}], based on the identity

(eiψk)(xj)=ψk(xj)(g0(xj)+gi(xj))(\mathcal{L}^{e_{i}}\psi_{k})(x_{j})=\nabla\psi_{k}(x_{j})^{\top}\left(g_{0}(x_{j})+g_{i}(x_{j})\right)

to construct the data-driven approximation of u\mathcal{L}^{u} according to (2.6). Consequently, for φ𝕍\varphi\in\mathbb{V} and control function uLloc([0,t],nc)u\in L^{\infty}_{\operatorname{loc}}([0,t],\mathbb{R}^{n_{c}}), a data-driven predictor is given as the solution of the linear time-varying Cauchy problem (2.4), where the unbounded operator \mathcal{L} is replaced by du(t)\mathcal{L}_{d}^{u(t)}. The convergence of this estimator was shown in [14] if both the dictionary size and the number of data points goes to infinity. Finite-data bounds typically split the error into two sources: A projection error stemming from the finite dictionary and an estimation error resulting from a finite amount of data. A bound on the estimation error for control systems was derived in [23], where, in addition to i.i.d. sampling of ODEs, also SDEs and ergodic sampling, i.e. sampling along one sufficiently-long trajectory, were considered. A full approximation error bound for control systems was provided in [30] using a dictionary of finite elements. We provide an error bound tailored to the sampled-data setting used in this work in Subsection 3.1.

3 Proportional error bound for EDMD-based MPC and problem formulation

We consider the discrete-time control system given by

x+=f(x,u)x^{+}=f(x,u) (3.1)

with nonlinear map f:nx×ncnxf:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{c}}\rightarrow\mathbb{R}^{n_{x}}. Then, for initial state x^nx\hat{x}\in\mathbb{R}^{n_{x}} and sequence of control values (u(k))k0(u(k))_{k\in\mathbb{N}_{0}}, xu(n;x^)x_{u}(n;\hat{x}) denotes the solution at time n0n\in\mathbb{N}_{0}, which is recursively defined by (3.1) and xu(0;x^)=x^x_{u}(0;\hat{x})=\hat{x}. In the following, f(0,0)=0f(0,0)=0 is assumed, i.e., the origin is a controlled equilibrium for u=0u=0. After reviewing the basics of model predictive control, we derive a sampled-data representation of the continuous-time dynamics (2.5) and the corresponding abstract Cauchy problem, i.e., (2.4) with u(t)\mathcal{L}^{u(t)} including its -based surrogate in Subsection 3.1. Then, we provide the problem formulation in Subsection 3.2.

We impose state and control constraints using the compact sets 𝕏nx\mathbb{X}\subset\mathbb{R}^{n_{x}} and 𝕌nc\mathbb{U}\subset\mathbb{R}^{n_{c}} with (0,0)int(𝕏×𝕌)(0,0)\in\operatorname{int}(\mathbb{X}\times\mathbb{U}), respectively. Next, we define admissibility of a sequence of control values.

Definition 1.

A sequence of control values (u(k))k=0N1𝕌(u(k))_{k=0}^{N-1}\subset~{}\mathbb{U} of length NN is said to be admissible for state x^𝕏\hat{x}\in\mathbb{X}, if xu(k;x^)𝕏x_{u}(k;\hat{x})\in\mathbb{X} holds for all k[1:N]k\in[1:N]. For x^𝕏\hat{x}\in\mathbb{X}, the set of admissible control sequences is denoted by 𝒰N(x^)\mathcal{U}_{N}(\hat{x}). If, for u=(u(k))k0u=(u(k))_{k\in\mathbb{N}_{0}}, (u(k))k=0N1𝒰N(x^)(u(k))_{k=0}^{N-1}\in\mathcal{U}_{N}(\hat{x}) holds for the restriction of uu for all N0N\in\mathbb{N}_{0}, we write u𝒰(x^)u\in\mathcal{U}_{\infty}(\hat{x}).

We introduce the quadratic stage cost :𝕏×𝕌0\ell:\mathbb{X}\times\mathbb{U}\rightarrow\mathbb{R}_{\geq 0},

(x,u):=xQ2+uR2:=xQx+uRu,\ell(x,u):=\|x\|_{Q}^{2}+\|u\|_{R}^{2}:=x^{\top}Qx+u^{\top}Ru, (3.2)

for symmetric and positive definite matrices Qnx×nxQ\in\mathbb{R}^{n_{x}\times n_{x}} and Rnc×ncR\in\mathbb{R}^{n_{c}\times n_{c}}. Next, based on Definition 1, we introduce the MPC Algorithm, where we tacitly assume existence of an optimal sequence of control values in Step (2) along the MPC closed-loop dynamics and full-state measurement.

Algorithm 2 (Model Predictive Control with horizon NN).

At each time n0n\in\mathbb{N}_{0}:

  1. (1)

    Measure the state x(n)𝕏x(n)\in\mathbb{X} and set x^:=x(n)\hat{x}:=x(n).

  2. (2)

    Solve the optimization problem

    uargminu𝒰N(x^)JN(x^,u):=k=0N1(xu(k;x^),u(k))u^{\star}\!\in\!\operatorname{argmin}_{u\in\mathcal{U}_{N}(\hat{x})}\ J_{N}(\hat{x},u):=\!\sum_{k=0}^{N-1}\ell(x_{u}(k;\hat{x}),u(k))

    subject to xu(0;x^)=x^x_{u}(0;\hat{x})=\hat{x} and the dynamics xu(k+1;x^)=f(xu(k;x^),u(k))x_{u}(k+1;\hat{x})=f(x_{u}(k;\hat{x}),u(k)), k[0:N2]k\in[0:N-2].

  3. (3)

    Apply the feedback value μN(x(n)):=u(0)𝕌\mu_{N}(x(n)):=u^{\star}(0)\in\mathbb{U}.

Overall, Algorithm 2 yields the MPC closed-loop dynamics

xμN+=f(xμN,μN(xμN)),x^{+}_{\mu_{N}}=f(x_{\mu_{N}},\mu_{N}(x_{\mu_{N}})), (3.3)

where the feedback law μN\mu_{N} is well defined at x^\hat{x} if 𝒰N(x^)\mathcal{U}_{N}(\hat{x})\neq\emptyset holds. We emphasize that this condition holds if, e.g., 𝕏\mathbb{X} is controlled forward invariant and refer to [1] and [5] for sufficient condition to ensure recursive feasibility without requiring controlled forward invariance of 𝕏\mathbb{X} (and without terminal conditions) for discrete and continuous-time systems, respectively. The closed-loop solution resulting from the dynamics (3.3) is denoted by xμN(n;x^)x_{\mu_{N}}(n;\hat{x}), where xμN(0;x^)=x^x_{\mu_{N}}(0;\hat{x})=\hat{x} holds. Moreover, we define the (optimal) value function VN:𝕏0{}V_{N}:\mathbb{X}\rightarrow\mathbb{R}_{\geq 0}\cup\{\infty\} as VN(x):=infu𝒰N(x)JN(x,u)V_{N}(x):=\inf_{u\in\mathcal{U}_{N}(x)}J_{N}(x,u).

3.1 Proportional error bound for sampled-data systems

We consider the nonlinear continuous-time control system given by (2.5). Equidistantly discretizing the time axis [0,)[0,\infty), i.e., using the partition k=0[kΔt,(k+1)Δt)\bigcup_{k=0}^{\infty}[k\Delta t,(k+1)\Delta t) with sampling period Δt>0\Delta t>0, and using a (piecewise) constant control function on each sampling interval, i.e., u(t)u^𝕌ncu(t)\equiv\hat{u}\in\mathbb{U}\subset\mathbb{R}^{n_{c}} on [kΔt,(k+1)Δ)[k\Delta t,(k+1)\Delta), we generate the discrete-time system

x+=f(x^,u^):=0Δtg0(x(t;x^,u))+i=1ncgi(x(t;x^,u))ui(t)dt.x^{+}=f(\hat{x},\hat{u}):=\!\int_{0}^{\Delta t}g_{0}(x(t;\hat{x},u))+\sum_{i=1}^{n_{c}}g_{i}(x(t;\hat{x},u))u_{i}(t)\,\mathrm{d}t. (3.4)

We emphasize that the drift g0g_{0} does not exhibit an offset independently of the state variable xx in view of our assumption f(0,0)=0=g0(0)f(0,0)=0=g_{0}(0). We define the vector-valued observable

Ψ(x)=(ψ1(x),,ψM(x))=(1,x1,,xnx,ψnx+2(x),,ψM(x)),\displaystyle\begin{split}\Psi(x)&=\begin{pmatrix}\psi_{1}(x),\ldots,\psi_{M}(x)\end{pmatrix}\\ &=\begin{pmatrix}1,x_{1},\ldots,x_{n_{x}},\psi_{n_{x}+2}(x),\ldots,\psi_{M}(x)\end{pmatrix},\end{split} (3.5)

where ψ1(x)1\psi_{1}(x)\equiv 1, ψk+1(x)=xk\psi_{k+1}(x)=x_{k}, k[1:nx]k\in[1:n_{x}], and ψk𝒞1(nx,)\psi_{k}\in\mathcal{C}^{1}(\mathbb{R}^{n_{x}},\mathbb{R}), k[nx+2:M]k\in[n_{x}+2:M], are locally-Lipschitz continuous functions satisfying ψk(0)=0\psi_{k}(0)=0 and (Dψk)(0)=0(D\psi_{k})(0)=0. Hence, Ψ:𝕏M\Psi:\mathbb{X}\to\mathbb{R}^{M} is Lipschitz continuous with constant LΨL_{\Psi} such that Ψ(x)Ψ(0)LΨx\|\Psi(x)-\Psi(0)\|\leq L_{\Psi}\|x\| holds. A straightforward calculation then shows (P𝕍0|𝕍)k,10(P_{\mathbb{V}}\mathcal{L}^{0}|_{\mathbb{V}})_{k,1}\equiv 0, k[1:M]k\in[1:M], which we impose for the data-driven approximation to ensure consistency, i.e., that f(0,0)=g0(0)=0f(0,0)=g_{0}(0)=0 is preserved. For gig_{i}, i[1:nc]i\in[1:n_{c}], the first (constant) observable enables us to approximate components of the control maps, which do not depend on the state xx, separately.

In this note, we make use of the following Assumption 3, which ensures that no projection error occurs. This assumption is common in systems and control when the Koopman framework is used, see, e.g., [29, 13]. The construction of suitable dictionaries ensuring this assumption is discussed in [3, 15]. A condition ensuring this invariance is provided, e.g., in [7, Theorem 1], where even a method for the construction of a suitable dictionary is discussed.

Assumption 3 (Invariance of 𝕍\mathbb{V}).

For any φ𝕍\varphi\in\mathbb{V}, the relation φ(x(Δt;,u))𝕍\varphi(x(\Delta t;\cdot,u))\in\mathbb{V} holds for all u(t)u^𝕌ncu(t)\equiv\hat{u}\in\mathbb{U}\subset\mathbb{R}^{n_{c}}.

We note that if this invariance assumption does not hold, and in order mitigate the projection error, subspace identification methods may be employed to (approximately) ensure invariance of the dictionary, i.e., the space spanned by the choosen observables, see, e.g., [11, 16].

Next, we deduce an error bound adapted to our sampled-data setting. Assumption 3 implies that the compression of the generator coincides with its restriction onto 𝕍\mathbb{V}, i.e., P𝕍u|𝕍=u|𝕍P_{\mathbb{V}}\mathcal{L}^{u}|_{\mathbb{V}}=\mathcal{L}^{u}|_{\mathbb{V}}. Thus, for u𝕌u\in\mathbb{U}, the Koopman operator is the matrix exponential of the generator, i.e., 𝒦uΔt=eΔtu\mathcal{K}^{\Delta t}_{u}=e^{\Delta t\mathcal{L}^{u}} holds.

Proposition 4.

Suppose that Assumption 3 holds. For every error bound ε>0\varepsilon>0 and probabilistic tolerance δ(0,1)\delta\in(0,1), there is an amount of data d0d_{0}\in\mathbb{N} such that with probability 1δ1-\delta, the error bound

eΔtu|𝕍eΔtduε\displaystyle\big{\|}e^{\Delta t\mathcal{L}^{u}|_{\mathbb{V}}}-e^{\Delta t\mathcal{L}_{d}^{u}}\big{\|}\leq\varepsilon (3.6)

holds for all dd0d\geq d_{0} and all u𝕌u\in\mathbb{U} for the Koopman operator 𝒦uΔt=eΔtu\mathcal{K}^{\Delta t}_{u}=e^{\Delta t\mathcal{L}^{u}}.

Proof.

For g(t)=etu|𝕍etdug(t)=e^{t\mathcal{L}^{u}|_{\mathbb{V}}}-e^{t\mathcal{L}_{d}^{u}}, we have

g(t)=u|𝕍etu|𝕍u|𝕍etduduetdu=u|𝕍g(t)+(u|𝕍du)(etduetu|𝕍).\displaystyle g^{\prime}(t)=\mathcal{L}^{u}|_{\mathbb{V}}e^{t\mathcal{L}^{u}|_{\mathbb{V}}}\mp\mathcal{L}^{u}|_{\mathbb{V}}e^{t\mathcal{L}_{d}^{u}}-\mathcal{L}_{d}^{u}e^{t\mathcal{L}_{d}^{u}}=\mathcal{L}^{u}|_{\mathbb{V}}g(t)+(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u})\Big{(}e^{t\mathcal{L}_{d}^{u}}\mp e^{t\mathcal{L}^{u}|_{\mathbb{V}}}\Big{)}.

Since g(0)=0g(0)=0, we have g(t)=0Δtg(s)dsg(t)=\int_{0}^{\Delta t}g^{\prime}(s)\,\mathrm{d}s. Then, plugging in the expression for g(s)g^{\prime}(s), the triangle inequality yields

g(t)0tβg(s)ds+α(t)\displaystyle\|g(t)\|\leq\int_{0}^{t}\beta\|g(s)\|\,\mathrm{d}s+\alpha(t)

with the constant β=u|𝕍+(u|𝕍du)\beta=\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|+\|(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u})\| and

α(t)=(u|𝕍du)0tesu|𝕍dsΔt(u|𝕍du)u|𝕍(eΔtu|𝕍1)=:cΔt\displaystyle\alpha(t)=\|(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u})\|\int_{0}^{t}\|e^{s\mathcal{L}^{u}|_{\mathbb{V}}}\|\,\mathrm{d}s\leq\frac{\Delta t\cdot\|(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u})\|}{\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}\Big{(}e^{\Delta t\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}-1\Big{)}=:c_{\Delta t}

for all t(0,Δt]t\in(0,\Delta t], where u|𝕍\mathcal{L}^{u}_{*}|_{\mathbb{V}} maximizes u|𝕍\|\mathcal{L}^{u}|_{\mathbb{V}}\| w.r.t. the compact set 𝕌\mathbb{U}. Then, Gronwall’s inequality with α(t)\alpha(t) replaced by cΔtc_{\Delta t} yields

g(Δt)\displaystyle\|g(\Delta t)\| cΔt(1+0Δtβe(Δtt)βdt)=cΔteΔtβ.\displaystyle\leq c_{\Delta t}\Big{(}1+\int_{0}^{\Delta t}\beta e^{(\Delta t-t)\beta}\,\mathrm{d}t\Big{)}=c_{\Delta t}e^{\Delta t\beta}.

Invoking [30, Theorem 3] yields, for any ε~>0\tilde{\varepsilon}>0, a sufficient amount of data d0d_{0}\in\mathbb{N} such that u|𝕍duε~\|\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}^{u}_{d}\|\leq\tilde{\varepsilon} holds for all u𝕌u\in\mathbb{U} and dd0d\geq d_{0}. Hence, setting ε~\tilde{\varepsilon} such that the inequality

Δtε~u|𝕍(eΔtu|𝕍1)eΔt(u|𝕍+ε~)ε\frac{\Delta t\cdot\tilde{\varepsilon}}{\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}\Big{(}e^{\Delta t\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}-1\Big{)}e^{\Delta t\left(\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|+\tilde{\varepsilon}\right)}\leq\varepsilon (3.7)

holds and using the definitions of β\beta and cΔtc_{\Delta t} ensures Inequality (3.6). Since the left hand side is monotonically increasing in ε~\tilde{\varepsilon} and zero for ε~=0\tilde{\varepsilon}=0, this is always possible, which completes the proof. ∎

We briefly quantify the sufficient amount of data d0d_{0} in view of the dictionary size MM and the parameters ε\varepsilon and δ\delta. First, by a standard Chebychev inequality, one obtains the dependency d0M2/ε2δd_{0}\sim\nicefrac{{M^{2}}}{{\varepsilon^{2}\delta}}, cf. [30, 23]. This can be improved in reproducing kernel Hilbert spaces, where the dictionary is given by feature maps given by the kernel evaluated at the samples. Here a scaling depending logarithmically on δ\delta was shown in [27, Proposition 3.4] using Hoeffding’s inequality, see also [26]. In the latter reference, invariance conditions were discussed, which may allow to relax Assumption 3. Otherwise, only bounds on the projection error w.r.t. the L2L_{2}-norm are available [30], which does not yield pointwise bounds.

For the discrete-time dynamics (3.4), we get the identity

f(x^,u^)=PxeΔtu^|𝕍Ψ(x^)\displaystyle f(\hat{x},\hat{u})=P_{x}e^{\Delta t\mathcal{L}^{\hat{u}}|_{\mathbb{V}}}\Psi(\hat{x}) (3.8)

resulting from sampling with zero-order hold in view of Assumption 3, where Px:MnxP_{x}:\mathbb{R}^{M}\to\mathbb{R}^{n_{x}} is the projection onto the first nxn_{x} components. Further, based on the bi-linear -based surrogate model of Subsection 2 for dd data points, we define the data-driven surrogate model

fε(x^,u^)=PxeΔtdu^Ψ(x^).\displaystyle f^{\varepsilon}(\hat{x},\hat{u})=P_{x}e^{\Delta t\mathcal{L}^{\hat{u}}_{d}}\Psi(\hat{x}). (3.9)

Next, we derive a novel error bound that is proportional to the norm of the state and the control and, thus, ensures that the error becomes small close to the origin.

Proposition 5.

Let LΨL_{\Psi} be the Lipschitz constant of Ψ\Psi on the set 𝕏\mathbb{X}. Then, for every error bound ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], the inequality

f(x,u)fε(x,u)ε(LΨx+Δtc~u)\displaystyle\|f(x,u)-f^{\varepsilon}(x,u)\|\leq\varepsilon\left(L_{\Psi}\|x\|+\Delta t\cdot\tilde{c}\|u\|\right) (3.10)

holds for all x𝕏x\in\mathbb{X} and u𝕌u\in\mathbb{U} with some constant c~\tilde{c} if (3.6) holds provided {f(x,u),fε(x,u)}𝕏\{f(x,u),f^{\varepsilon}(x,u)\}\subset\mathbb{X}.

Proof.

By local Lipschitz continuity of Ψ\Psi, 0int(𝕏)0\in\operatorname{int}(\mathbb{X}) and Px1\|P_{x}\|\leq 1 we compute

f(x,u)fε(x,u)\displaystyle\|f(x,u)-f^{\varepsilon}(x,u)\| =Px[eΔtu|𝕍eΔtdu][Ψ(x)±Ψ(0)]\displaystyle=\big{\|}P_{x}[e^{\Delta t\mathcal{L}^{u}|_{\mathbb{V}}}-e^{\Delta t\mathcal{L}^{u}_{d}}][\Psi(x)\pm\Psi(0)]\big{\|}
εΨ(x)Ψ(0)LΨεx+(eΔtu|𝕍eΔtdu)Ψ(0)=:h(Δt).\displaystyle\leq\underbrace{\varepsilon\|\Psi(x)-\Psi(0)\|}_{\leq L_{\Psi}\varepsilon\|x\|}+\big{\|}\underbrace{(e^{\Delta t\mathcal{L}^{u}|_{\mathbb{V}}}-e^{\Delta t\mathcal{L}_{d}^{u}})\Psi(0)}_{=:h(\Delta t)}\big{\|}.

Then, Taylor series expansion of h(Δt)=h(0)+Δth(ξ)h(\Delta t)=h(0)+\Delta t\cdot h^{\prime}(\xi), ξ[0,Δt]\xi\in[0,\Delta t], with h(0)=0h(0)=0 leads to the representation

h(Δt)Δt=\displaystyle\text{\scriptsize$\frac{h(\Delta t)}{\Delta t}$}= (eξu|𝕍u|𝕍±eξduu|𝕍eξdudu)Ψ(0)\displaystyle(e^{\xi\mathcal{L}^{u}|_{\mathbb{V}}}\mathcal{L}^{u}|_{\mathbb{V}}\pm e^{\xi\mathcal{L}_{d}^{u}}\mathcal{L}^{u}|_{\mathbb{V}}-e^{\xi\mathcal{L}_{d}^{u}}\mathcal{L}_{d}^{u})\Psi(0)
=\displaystyle= (eξu|𝕍eξdu)u|𝕍Ψ(0)+eξdu(u|𝕍du)Ψ(0).\displaystyle(e^{\xi\mathcal{L}^{u}|_{\mathbb{V}}}-e^{\xi\mathcal{L}_{d}^{u}})\mathcal{L}^{u}|_{\mathbb{V}}\Psi(0)+e^{\xi\mathcal{L}_{d}^{u}}\left(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u}\right)\Psi(0).

For a sufficient amount of data d0d_{0}\in\mathbb{N}, we have maxi[1:n]ei|𝕍deiε¯\max_{i\in[1:n]}\|\mathcal{L}^{e_{i}}|_{\mathbb{V}}-\mathcal{L}^{e_{i}}_{d}\|\leq\bar{\varepsilon}. Then, the second summand can be estimated by

[eξu|𝕍eξdu+eξu|𝕍](u|𝕍du)Ψ(0)c0ε¯u\displaystyle\left[\|e^{\xi\mathcal{L}^{u}|_{\mathbb{V}}}-e^{\xi\mathcal{L}_{d}^{u}}\|+\|e^{\xi\mathcal{L}^{u}|_{\mathbb{V}}}\|\right]\left\|\left(\mathcal{L}^{u}|_{\mathbb{V}}-\mathcal{L}_{d}^{u}\right)\Psi(0)\right\|\leq c_{0}\bar{\varepsilon}\|u\|

with c0:=eΔtu|𝕍+εc_{0}:=e^{\Delta t\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}+\varepsilon with ε\varepsilon from Proposition 4, where u|𝕍\mathcal{L}^{u}_{*}|_{\mathbb{V}} maximizes u|𝕍\|\mathcal{L}^{u}|_{\mathbb{V}}\| w.r.t. the compact set 𝕌\mathbb{U} and we have used that the contributions of 0\mathcal{L}^{0} and d0\mathcal{L}^{0}_{d} cancel out thanks to Ψ(0)\Psi(0) and the control value acts as a factor. The same argument yields u|𝕍Ψ(0)u|𝕍u\|\mathcal{L}^{u}|_{\mathbb{V}}\Psi(0)\|\leq\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|\|u\|. Combining the derived estimates yields the assertion, i.e., Inequality (3.10) with c~:=eΔtu|𝕍+ε0+u|𝕍\tilde{c}:=e^{\Delta t\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|}+\varepsilon_{0}+\|\mathcal{L}^{u}_{*}|_{\mathbb{V}}\|. ∎

In [32], a bound of the form (3.10) was assumed in the lifted space, i.e., without the projector PxP_{x}. Therein, the bound was used to construct a feedback controller achieving robust local stability using a finite gain argument. However, the bound was not established, but rather assumed – in addition to the invariance in Assumption 3.

3.2 Problem statement

We will leverage the error bound of Proposition 5 to provide a stability result when using the surrogate dynamics fεf^{\varepsilon} in Step (2) of the MPC Algorithm 2 to stabilize the original system. The main result shows that, if the nominal MPC controller is asymptotically stabilizing, the data-based controller with fεf^{\varepsilon} ensures convergence to a neighborhood of the origin, whose size depends on ε\varepsilon, i.e., practical asymptotic stability.

Definition 6 (Practical asymptotic stability).

For ε>0\varepsilon>0, let μNε\mu_{N}^{\varepsilon} be the feedback law defined in Algorithm 2 with f=fεf=f^{\varepsilon}, where admissibility of control sequences at x^\hat{x}, i.e., u𝒰Nε(x^)u\in\mathcal{U}_{N}^{\varepsilon}(\hat{x}), is defined w.r.t. the tightened set 𝕏ε(0)\mathbb{X}\ominus\mathcal{B}_{\varepsilon}(0). Let A𝕏ε(0)A\subset\mathbb{X}\ominus\mathcal{B}_{\varepsilon}(0) be given such that 𝒰Nε(x^)\mathcal{U}_{N}^{\varepsilon}(\hat{x})\neq\emptyset for all x^A\hat{x}\in A. Then, the origin is said to be semi-globally practically asymptotically stable (PAS) on AA if there exists β𝒦\beta\in\mathcal{K}\mathcal{L} such that for each r>0r>0 and R>rR>r there is ε0>0\varepsilon_{0}>0 such that for each x^A\hat{x}\in A with x^R\|\hat{x}\|\leq R and all ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] such that (3.10) holds, the solution xμNε(,x^)x_{\mu_{N}^{\varepsilon}}(\cdot,\hat{x}) of

xμNε(n+1)=f(xμNε(n),μNε(xμNε(n)))\displaystyle x_{\mu_{N}^{\varepsilon}}(n+1)=f(x_{\mu_{N}^{\varepsilon}}(n),\mu_{N}^{\varepsilon}(x_{\mu_{N}^{\varepsilon}}(n))) (3.11)

with xμNε(0)=x^x_{\mu_{N}^{\varepsilon}}(0)=\hat{x} satisfies xμNε(n;x^)Ax_{\mu_{N}^{\varepsilon}}(n;\hat{x})\in A and

xμNε(n;x^)max{β(x^,n),r}n0.\displaystyle\|x_{\mu_{N}^{\varepsilon}}(n;\hat{x})\|\leq\max\{\beta(\|\hat{x}\|,n),r\}\qquad\forall\,n\in\mathbb{N}_{0}.

The incorporation of the Pontryagin difference 𝕏ε(0)\mathbb{X}\ominus\mathcal{B}_{\varepsilon}(0) in the admissibility of control sequences for the surrogate model ensures that the original system evolves in the compact set 𝕏\mathbb{X}, i.e., that every optimal control function is, in particular, admissible for the original system in view of the error bound of Proposition 4. In the following section, we will show that the error bound shown in Proposition 5 and cost-controllability of the original dynamics imply practical asymptotic stability of the closed-loop using EDMD-based MPC.

4 Practical asymptotic stability of surrogate-based MPC

In this section, we prove our main result, i.e., practical asymptotic stability of the data-based MPC Algorithm 2 using the surrogate fεf^{\varepsilon} as defined in (3.9) to stabilize the original system with ff given by (3.4) or, equivalently, (3.8).

We follow the line of reasoning outlined in [9, Section 11.5]. To this end, we recall [9, Theorem 11.10] regarding stability for perturbed solutions in Proposition 7, which is a key tool for our analysis. We define

VNε(x^):=infu𝒰Nε(x^)k=0N1(xuε(k;x^),u(k))V^{\varepsilon}_{N}(\hat{x}):=\inf_{u\in\mathcal{U}_{N}^{\varepsilon}(\hat{x})}\sum_{k=0}^{N-1}\ell(x^{\varepsilon}_{u}(k;\hat{x}),u(k))

where xuε(0;x^)=x^x_{u}^{\varepsilon}(0;\hat{x})=\hat{x} and xuε(k+1;x^)=fε(xuε(k;x^),u(k))x_{u}^{\varepsilon}(k+1;\hat{x})=f^{\varepsilon}(x_{u}^{\varepsilon}(k;\hat{x}),u(k)) for k[0:N2]k\in[0:N-2].

Proposition 7.

Consider the MPC-feedback law μNε\mu_{N}^{\varepsilon} of Algorithm 2 with f=fεf=f^{\varepsilon}, where fεf^{\varepsilon} satisfies Condition (3.10) and let S𝕏S\subset\mathbb{X} be a forward-invariant set w.r.t. fε(,μNε())f^{\varepsilon}(\cdot,\mu^{\varepsilon}_{N}(\cdot)). Further, let the following assumptions hold:

(i) There is ε0>0\varepsilon_{0}>0 and α(0,1]\alpha\in(0,1] such that for all ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] the relaxed dynamic programming inequality

VNε(x)α(x,μNε(x))+VNε(fε(x,μNε(x)))\displaystyle V_{N}^{\varepsilon}(x)\geq\alpha\ell(x,\mu^{\varepsilon}_{N}(x))+V_{N}^{\varepsilon}(f^{\varepsilon}(x,\mu^{\varepsilon}_{N}(x)))

holds on SS. In addition, there exist α1,α2,α3𝒦\alpha_{1},\alpha_{2},\alpha_{3}\in\mathcal{K}_{\infty} such that

α1(x)VNε(x)α2(x)and(x,u)α3(x)\displaystyle\alpha_{1}(\|x\|)\leq V_{N}^{\varepsilon}(x)\leq\alpha_{2}(\|x\|)\ \ \mathrm{and}\ \ \ell(x,u)\geq\alpha_{3}(\|x\|)

hold for all xSx\in S, ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], and u𝕌u\in\mathbb{U}.

(ii) VNεV^{\varepsilon}_{N} is uniformly continuous and fεf^{\varepsilon} is uniformly continuous in uu on closed balls B¯ρ(0)\overline{B}_{\rho}(0), i.e., there is ε0\varepsilon_{0} such that, for each ρ>0\rho>0, there exists ωV,ωf𝒦\omega_{V},\omega_{f}\in\mathcal{K}:

|VNε(x)VNε(y)|\displaystyle|V^{\varepsilon}_{N}(x)-V^{\varepsilon}_{N}(y)| ωV(xy),\displaystyle\leq\omega_{V}(\|x-y\|),
fε(x,u)fε(y,u)\displaystyle\|f^{\varepsilon}(x,u)-f^{\varepsilon}(y,u)\| ωf(xy)u𝕌\displaystyle\leq\omega_{f}(\|x-y\|)\qquad\forall\,u\in\mathbb{U}

for all x,yB¯ρ(0)Sx,y\in\overline{B}_{\rho}(0)\cap S and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}]. Then the exact closed-loop system with perturbed feedback μNε\mu_{N}^{\varepsilon} defined in (3.11) is semiglobally practically asymptotically stable on A=SA=S in the sense of Definition 6.

We first verify the condition of Proposition 7 considering uniform continuity of the surrogate model.

Lemma 8.

Let ε0>0\varepsilon_{0}>0 be given. Then, fεf^{\varepsilon} is uniform continuous in uu with ωf(r)=cLΨr\omega_{f}(r)=cL_{\Psi}r, c=c(ε0)c=c(\varepsilon_{0}), i.e.,

fε(x,u)fε(y,u)cLΨxy\displaystyle\|f^{\varepsilon}(x,u)-f^{\varepsilon}(y,u)\|\leq cL_{\Psi}\|x-y\| (4.1)

holds for all x,y𝕏x,y\in\mathbb{X}, u𝕌u\in\mathbb{U}, and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] provided that the error bound (3.6) is satisfied.

Proof.

The error bound (3.6) and Px1\|P_{x}\|\leq 1 imply

fε(x,u)fε(y,u)\displaystyle\|f^{\varepsilon}(x,u)-f^{\varepsilon}(y,u)\| =PxeΔtdu(Ψ(x)Ψ(y))\displaystyle=\|P_{x}e^{\Delta t\mathcal{L}^{u}_{d}}(\Psi(x)-\Psi(y))\|
eΔtdueΔtu|𝕍Ψ(x)Ψ(y)\displaystyle\leq\|e^{\Delta t\mathcal{L}^{u}_{d}}\mp e^{\Delta t\mathcal{L}^{u}|_{\mathbb{V}}}\|\cdot\|\Psi(x)-\Psi(y)\|
(ε0+eΔtu|𝕍)LΨxy,\displaystyle\leq(\varepsilon_{0}+\|e^{\Delta t\mathcal{L}^{u}_{*}|_{\mathbb{V}}}\|)L_{\Psi}\|x-y\|,

where u|𝕍\mathcal{L}^{u}_{*}|_{\mathbb{V}} maximizes u|𝕍\|\mathcal{L}^{u}|_{\mathbb{V}}\| w.r.t. the compact set 𝕌\mathbb{U}. This completes the proof with c:=ε0+eΔtu|𝕍c:=\varepsilon_{0}+\|e^{\Delta t\mathcal{L}^{u}_{*}|_{\mathbb{V}}}\|. ∎

Using the novel proportional error bound of Proposition 5 we rigorously show that cost controllability as defined in [4] and [38] for continuous- and discrete-time systems, respectively, is inherited by the EDMD-based surrogate model. Cost controllability links stabilizability with the stage cost employed in MPC, see, e.g., [10, 38]. The only additional requirement is that optimal control sequences have to be admissible also for the surrogate model. While this may be a severe restriction close to the boundary of the set 𝕏ε(0)\mathbb{X}\ominus\mathcal{B}_{\varepsilon}(0), it is typically satisfied on a suitably chosen sub-level set of the optimal value function VNV_{N} in view of the finite prediction horizon NN.

Proposition 9.

Let the error bound (3.6) hold with ε>0\varepsilon>0 and the stage cost be given by (3.2). Suppose existence of a monotonically increasing and bounded sequence (Bk)k(B_{k})_{k\in\mathbb{N}}\subset\mathbb{R} and a set S𝕏ε(0)S\subseteq\mathbb{X}\ominus\mathcal{B}_{\varepsilon}(0) such that the growth bound

Vk(x^)Jk(x^,u^)Bk(x^)k\displaystyle V_{k}(\hat{x})\leq J_{k}(\hat{x},\hat{u})\leq B_{k}\ell^{\star}(\hat{x})\qquad\forall\,k\in\mathbb{N} (4.2)

with (x^):=infu𝕌(x^,u)\ell^{\star}(\hat{x}):=\inf_{u\in\mathbb{U}}\ell(\hat{x},u) holds for all x^S\hat{x}\in S and some u^=u^(x^)𝒰N(x^)𝒰Nε(x^)\hat{u}=\hat{u}(\hat{x})\in\mathcal{U}_{N}(\hat{x})\cap\mathcal{U}_{N}^{\varepsilon}(\hat{x}). Then, there exists a monotonically increasing and bounded sequence (Bkε)k(B_{k}^{\varepsilon})_{k\in\mathbb{N}}\subset\mathbb{R} such that Inequality (4.2) holds for VkεV_{k}^{\varepsilon} and JkεJ_{k}^{\varepsilon} instead of VkV_{k} and JkJ_{k}, respectively. Moreover, we have BkεBkB_{k}^{\varepsilon}\rightarrow B_{k} for ε0\varepsilon\rightarrow 0, kk\in\mathbb{N}.

Proof.

Let x~()\tilde{x}(\cdot) and x(){x}(\cdot) denote the trajectories generated by x~(n+1)=fε(x~(n),u^n)\tilde{x}(n+1)=f^{\varepsilon}(\tilde{x}(n),\hat{u}_{n}) and x(n+1)=f(x(n),u^n)x(n+1)=f(x(n),\hat{u}_{n}), n0n\in\mathbb{N}_{0}, with x~(0)=x^=x(0)\tilde{x}(0)=\hat{x}=x(0), respectively. Set λ¯=max{|λ|:λeigenvalueofRorQ}\bar{\lambda}=\max\{|\lambda|:\lambda\ \mathrm{eigenvalue\ of}\ R\ \mathrm{or}\ Q\} and 0<λ¯=min{|λ|:λeigenvalueofRorQ}0<\underline{\lambda}=\min\{|\lambda|:\lambda\ \mathrm{eigenvalue\ of}\ R\ \mathrm{or}\ Q\}. Then, we have

(x~(n),u^n)\displaystyle\ell(\tilde{x}(n),\hat{u}_{n}) =(x~(n)x(n))+x(n)Q2+u^nR2\displaystyle=\|(\tilde{x}(n)-x(n))+x(n)\|^{2}_{Q}+\|\hat{u}_{n}\|^{2}_{R} (4.3)
λ¯x~(n)x(n)2+2λ¯x~(n)x(n)x(n)+(x(n),u^n),\displaystyle\leq\bar{\lambda}\|\tilde{x}(n)-x(n)\|^{2}+2\bar{\lambda}\|\tilde{x}(n)-x(n)\|\|x(n)\|+\ell(x(n),\hat{u}_{n}),

If (3.6) holds, then Proposition 5 yields the bound (3.10) on the difference of ff and fεf^{\varepsilon}. Thus, we may estimate the term en+1:=x~(n+1)x(n+1)e_{n+1}:=\|\tilde{x}(n+1)-x(n+1)\| by

en+1\displaystyle e_{n+1} =fε(x~(n),u^n)±f(x~(n),u^n)f(x(n),u^n)\displaystyle=\|f^{\varepsilon}(\tilde{x}(n),\hat{u}_{n})\pm f(\tilde{x}(n),\hat{u}_{n})-f(x(n),\hat{u}_{n})\|
ε(LΨx~(n)x(n)+Δtc~u^n)+Lfen\displaystyle\leq\varepsilon\left(L_{\Psi}\|\tilde{x}(n)\mp x(n)\|+\Delta t\tilde{c}\|\hat{u}_{n}\|\right)+L_{f}e_{n}
=εc¯(x(n)+u^n)+den\displaystyle=\varepsilon\cdot\bar{c}\left(\|x(n)\|+\|\hat{u}_{n}\|\right)+de_{n}

with c¯:=max{LΨ,Δtc~}\bar{c}:=\max\{L_{\Psi},\Delta t\tilde{c}\} and d:=Lf+εLΨd:=L_{f}+\varepsilon L_{\Psi}. Hence,

en2\displaystyle e_{n}^{2} 4ε2c¯2(x(n1)2+u^n12)+2d2en12\displaystyle\leq 4\varepsilon^{2}\bar{c}^{2}(\|x(n-1)\|^{2}+\|\hat{u}_{n-1}\|^{2})+2d^{2}e_{n-1}^{2}
4ε2c¯2λ¯i=0n1(2d2)n1i(x(i),u^i),\displaystyle\leq\frac{4\varepsilon^{2}\bar{c}^{2}}{\underline{\lambda}}\sum_{i=0}^{n-1}(2d^{2})^{n-1-i}\ell(x(i),\hat{u}_{i}),
enx(n)\displaystyle e_{n}\|x(n)\| εc¯x(n1)2+u^n12+2x(n)22+den1x(n)\displaystyle\leq\varepsilon\bar{c}\text{\footnotesize$\frac{\|x(n-1)\|^{2}+\|\hat{u}_{n-1}\|^{2}+2\|x(n)\|^{2}}{2}$}+de_{n-1}\|x(n)\|
εc¯2λ¯i=0n1dn1i((x(i),u^i)+(x(n)))\displaystyle\leq\frac{\varepsilon\bar{c}}{2\underline{\lambda}}\sum_{i=0}^{n-1}d^{n-1-i}\Big{(}\ell(x(i),\hat{u}_{i})+\ell^{\star}(x(n))\Big{)}

Summing up the resulting inequalities for (x~(n),u^n)\ell(\tilde{x}(n),\hat{u}_{n}) over n[1:N1]n\in[1:N-1] and using that the first summands in J~N(x^,u^)\tilde{J}_{N}(\hat{x},\hat{u}) and VN(x^)V_{N}(\hat{x}) coincide, we get

V~N(x^)\displaystyle\tilde{V}_{N}(\hat{x}) J~N(x^,u^)(4.3)JN(x^,u^)+λ¯(n=1N1en2+2enx(n))\displaystyle\leq\tilde{J}_{N}(\hat{x},\hat{u})\stackrel{{\scriptstyle\eqref{eq:cost_controllability:proof:1}}}{{\leq}}J_{N}(\hat{x},\hat{u})+\bar{\lambda}\left(\sum_{n=1}^{N-1}e_{n}^{2}+2e_{n}\|x(n)\|\right)
(BN+εc¯λ¯2λ¯c1+ε24c¯2λ¯λ¯c2)(x^)=:BNε(x^)\displaystyle\leq\left(B_{N}+\varepsilon\frac{\bar{c}\bar{\lambda}}{2\underline{\lambda}}c_{1}+\varepsilon^{2}\frac{4\bar{c}^{2}\bar{\lambda}}{\underline{\lambda}}c_{2}\right)\ell^{\star}(\hat{x})=:B_{N}^{\varepsilon}\ell^{\star}(\hat{x})

with constants c1=n=1N1dn1Bn+dN1BNc_{1}=\sum_{n=1}^{N-1}d^{n-1}B_{n}+d^{N-1}B_{N} and c2=n=1N1(2d2)n1Bnc_{2}=\sum_{n=1}^{N-1}(2d^{2})^{n-1}B_{n}, where we have invoked the imposed cost controllability multiple times. ∎

Finally, invoking our findings on cost controllability, we verify the remaining conditions of Proposition 7 to show the main result.

Theorem 10 (PAS of EDMD-based MPC).

Let the error bound (3.6), ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], for some ε0>0\varepsilon_{0}>0, Assumption 3 and cost controllability of the dynamics (3.4) and the stage cost (3.2), i.e., Condition (4.2), hold. Let the prediction horizon NN be chosen such that α(0,1)\alpha\in(0,1) holds with

α=αN:=1(B2ω)(BN1)i=3N(Bi1)i=2NBi(B2ω)i=3N(Bi1)\alpha=\alpha_{N}:=1-\frac{(B_{2}-\omega)(B_{N}-1)\prod_{i=3}^{N}(B_{i}-1)}{\prod_{i=2}^{N}B_{i}-(B_{2}-\omega)\prod_{i=3}^{N}(B_{i}-1)} (4.4)

and ω=1\omega=1.111The performance index or degree of suboptimality αN\alpha_{N} was proposed in [8] and [10, Theorem 5.4] and updated to (4.4) in [38]. Further, let S𝕏ε0(0)S\subset\mathbb{X}\ominus\mathcal{B}_{\varepsilon_{0}}(0) contain the origin in its interior and η>0\eta>0 be chosen such that, for all x^S\hat{x}\in S, an optimal control function u𝒰Nε(x^)u^{\star}\in\mathcal{U}_{N}^{\varepsilon}(\hat{x}) exists satisfying xuε(k;x^)𝕏ε+η(0)x^{\varepsilon}_{u^{\star}}(k;\hat{x})\in\mathbb{X}\ominus\mathcal{B}_{\varepsilon+\eta}(0), k[0:N1]k\in[0:N-1]. Then the EDMD-based MPC controller ensures semi-global practical asymptotic stability of the origin w.r.t. ε\varepsilon on the set SS.

Proof.

First, we show condition (i) of Proposition 7 for the system dynamics (3.9). To this end, note that the lower bound on the optimal value function can be inferred by

VNε(x^)=infu𝒰Nε(x^)JNε(x^,u)infu𝕌(x^,u)=x^Q2λ¯x^2\displaystyle V_{N}^{\varepsilon}(\hat{x})=\!\!\!\inf_{u\in\mathcal{U}_{N}^{\varepsilon}(\hat{x})}\!\!J_{N}^{\varepsilon}(\hat{x},u)\geq\!\inf_{u\in\mathbb{U}}\ell(\hat{x},u)=\|\hat{x}\|_{Q}^{2}\geq\underline{\lambda}\|\hat{x}\|^{2}

with λ¯>0\underline{\lambda}>0 defined as in the proof of Proposition 9. Then, defining αε0\alpha^{\varepsilon_{0}} analogously (4.4) using the sequence (Bnε0)n=2N(B_{n}^{\varepsilon_{0}})_{n=2}^{N} instead and invoking limε00Bnε0=Bn\lim_{\varepsilon_{0}\searrow 0}B_{n}^{\varepsilon_{0}}=B_{n} yields αε0(α,1)\alpha^{\varepsilon_{0}}\in(\alpha,1) for sufficiently small ε0\varepsilon_{0}. This ensures the relaxed Lyapunov inequality for all VNεV_{N}^{\varepsilon}, ε(0,ε0]\varepsilon\in(0,\varepsilon^{0}] by applying [8, Theorem 5.2]. Further, the upper bound on the value function VNε(x^)V_{N}^{\varepsilon}(\hat{x}) directly follows from the imposed (and preserved) cost controllability. Hence, we established the value function VNεV_{N}^{\varepsilon} as a Lyapunov function for the closed loop of the surrogate dynamics fεf^{\varepsilon}.

It remains to show |VNε(y1)VNε(y2)|Ly1y2|V^{\varepsilon}_{N}(y_{1})-V^{\varepsilon}_{N}(y_{2})|\leq L\|y_{1}-y_{2}\| for all y1,y2Sy_{1},y_{2}\in S and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], i.e., uniform continuity of VNεV^{\varepsilon}_{N} with ωV(r)=Lr\omega_{V}(r)=Lr, for some L>0L>0. Then, the condition of Proposition 7 hold and the assertion follows.

In combination with the uniform continuity of fεf^{\varepsilon} proven in Lemma 8, the assumption xuε(k;x^)𝕏ε+η(0)x^{\varepsilon}_{u^{\star}}(k;\hat{x})\in\mathbb{X}\ominus\mathcal{B}_{\varepsilon+\eta}(0) for all k[0:N1]k\in[0:N-1] implies the existence of η^>0\hat{\eta}>0 such that, for each x^S\hat{x}\in S, the respective optimal control u𝒰Nε(x^)u^{\star}\in\mathcal{U}_{N}^{\varepsilon}(\hat{x}) remains admissible for all initial values from η^(x^)\mathcal{B}_{\hat{\eta}}(\hat{x}). Then, VNεV_{N}^{\varepsilon} is uniformly bounded on SS. This immediately shows the assertion for y1,y2Sy_{1},y_{2}\in S with y1y2>η^\|y_{1}-y_{2}\|>\hat{\eta}, see, e.g., [1] for a detailed outline of the construction. Hence, it remains to show the assumption for y1,y2Sy_{1},y_{2}\in S satisfying y1y2η^\|y_{1}-y_{2}\|\leq\hat{\eta}. Based on our assumption that an optimal sequence of control values exists, for every y2𝕏y_{2}\in\mathbb{X} there is u2𝒰Nε(y2)u_{2}^{\star}\in\mathcal{U}_{N}^{\varepsilon}(y_{2}) such that VNε(y2)=JN(y2,u2)V^{\varepsilon}_{N}(y_{2})=J_{N}(y_{2},u_{2}^{\star}). Then, invoking admissibility of u2u_{2}^{\star} for y1y_{1}, uniform Lipschitz continuity of fε(,u)f^{\varepsilon}(\cdot,u) on SS in ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] and u𝕌u\in\mathbb{U}, we get

VNε(y1)VNε(y2)\displaystyle V^{\varepsilon}_{N}(y_{1})-V^{\varepsilon}_{N}(y_{2}) JNε(y1,u2)JNε(y2,u2)\displaystyle\leq J_{N}^{\varepsilon}(y_{1},u_{2}^{\star})-J_{N}^{\varepsilon}(y_{2},u_{2}^{\star})
=k=0N1xu2ε(k;y1)xu2ε(k;y2)Q2+2xu2ε(k;y2)Q(xu2ε(k;y1)xu2ε(k;y2))\displaystyle=\sum_{k=0}^{N-1}\|x_{u_{2}^{\star}}^{\varepsilon}(k;y_{1})-x_{u_{2}^{\star}}^{\varepsilon}(k;y_{2})\|_{Q}^{2}+2x_{u_{2}^{\star}}^{\varepsilon}(k;y_{2})^{\top}Q(x_{u_{2}^{\star}}^{\varepsilon}(k;y_{1})-x_{u_{2}^{\star}}^{\varepsilon}(k;y_{2}))
λ¯c¯[c¯η^+2Nxu2ε(k;y2))]y1y2\displaystyle\leq\bar{\lambda}\bar{c}\bigg{[}\bar{c}\hat{\eta}+2N\|x_{u_{2}^{\star}}^{\varepsilon}(k;y_{2})\|\bigg{)}\bigg{]}\|y_{1}-y_{2}\|

with c¯:=k=0N1(c(ε0)LΨ)k\bar{c}:=\sum_{k=0}^{N-1}(c(\varepsilon_{0}){L}_{\Psi})^{k} for all y1,y2𝕏y_{1},y_{2}\in\mathbb{X}. Then, using that xu2ε(k;y2)\|x_{u_{2}^{\star}}^{\varepsilon}(k;y_{2})\| is uniformly bounded on the compact set 𝕏\mathbb{X}, we have derived VNε(y1)VNε(y2)Ly1y2V^{\varepsilon}_{N}(y_{1})-V^{\varepsilon}_{N}(y_{2})\leq L\|y_{1}-y_{2}\|. Analogously,

VNε(y2)VNε(y1)JN(y2,u1)JN(y1,u1)Ly1y2\displaystyle V^{\varepsilon}_{N}(y_{2})-V^{\varepsilon}_{N}(y_{1})\leq J_{N}(y_{2},u_{1}^{\star})-J_{N}(y_{1},u_{1}^{\star})\leq L\|y_{1}-y_{2}\|

on SS. Combining both inequalities yields the assertion. ∎

The assumption that the minimum exists may be completely dropped and is only imposed to streamline the presentation, see, e.g., [9, p. 59] for details. The imposed (technical) condition w.r.t. η>0\eta>0 can, e.g., be ensured by choosing a sufficiently small sub-level set {xS:VNε(x)a}\{x\in S:V_{N}^{\varepsilon}(x)\leq a\} such that xuε(k)𝕏ε+η(0)x_{u^{\star}}^{\varepsilon}(k)\notin\mathbb{X}\ominus\mathcal{B}_{\varepsilon+\eta}(0) for some k[1:N1]k\in[1:N-1] yields a contradiction in view of the quadratic penalization of that state in the stage cost and the assumed bound aa on the sub-level set – similar to the construction used in [1].

The assumed bound (3.6) of Theorem 10 and cost controllability of the original system are the key ingredients for PAS of EDMD-based MPC. In Proposition 4 we proved that such a bound can be guaranteed with probability 1δ1-\delta. This allows to also deduce PAS with probability 1δ1-\delta. Increasing the number of samples can then be used to either increase the confidence (that is, to reduce δ\delta), or reduce ε\varepsilon. The latter allows to shrink the set of PAS, i.e., reduce the radius r>0r>0 in Definition 6.

5 Numerical simulations

In this section we conduct numerical simulations to validate practical asymptotic stability of the origin for EDMD-based MPC as rigorously shown in Theorem 10.

First, we consider the van-der-Pol oscillator given by

(x˙1(t)x˙2(t))=(x2(t)μ(1x12(t))x2(t)x1(t)+u(t))\displaystyle\begin{pmatrix}\dot{x}_{1}(t)\\ \dot{x}_{2}(t)\end{pmatrix}=\begin{pmatrix}x_{2}(t)\\ \mu(1-x_{1}^{2}(t))x_{2}(t)-x_{1}(t)+u(t)\end{pmatrix} (5.1)

for μ=0.1\mu=0.1. Since the linearization at the origin is controllable, cost controllablility holds for the quadratic stage cost (3.2), see, e.g., [38]. We consider the ODE (5.1) as a sampled-data system with zero-order hold as introduced in (3.4), where the integrals are numerically solved using the Runge-Kutta-Fehlberg method (RK45) with step-size control (Python function scipy.integrate.solve_ivp). For the approximation of the Koopman operator on the set 𝕏=[2,2]2\mathbb{X}=[-2,2]^{2}, EDMD as described in Section 2 is used. As dictionary of observables we choose all nxn_{x}-variate monomials of degree less or equal than three, resulting in a dictionary size of M=10M=10. The step size is set to Δt=0.05\Delta t=0.05.

First, we inspect the open-loop error of the EDMD-based surrogate for a random but fixed control sequence uu and different numbers of data points d{10,50,100,1000,10000}d\in\{10,50,100,1000,10000\}, cf. Figure 1, which shows the average norm of the error for 100100 initial conditions distributed uniformly over the set 𝕏=[2,2]2\mathbb{X}=[-2,2]^{2}. As to be expected from Proposition 5, the open-loop error decreases for increased number of samples.

Refer to caption
Figure 1: Averaged error of the EDMD-based solution for different number of data points for fixed random control sequence.

Next, we inspect the MPC closed-loop while imposing the constraints 5u(k)5-5\leq u(k)\leq 5 for k[0:N1]k\in[0:N-1] and x(k)𝕏x(k)\in\mathbb{X} for k[0:N]k\in[0:N], respectively. We compare the closed-loop performance resulting from nominal MPC denoted by xμNx_{\mu_{N}} as defined in (3.3) and EDMD-based MPC xμNεx_{\mu_{N}^{\varepsilon}} defined in (3.11) for λ{0.05,0.25}\lambda\in\{0.05,0.25\} and optimization horizons N{30,50}N\in\{30,50\}. The Koopman approximation is performed using EDMD with d=10000d=10000 i.i.d. data points. For small control penalization parameter λ=0.05\lambda=0.05, the norm of the closed-loop state corresponding to nominal MPC decays until the precision 101210^{-12} of the optimization solver is reached. As to be expected, this decay is faster for a longer prediction horizon. As proven in Theorem 10, the EDMD-based surrogate only enjoys practical asymptotic stability. More precisely, increasing the horizon only increases the convergence speed, but does not lead to a lower norm at the end of the considered simulation horizon.

In Figure 2, we illustrate the decrease of the optimal value function along the closed-loop trajectories. The observed stagnation indicates that the bottleneck is the approximation quality of the EDMD-based surrogate. The behavior is qualitatively very similar to the norm of the solution. Moreover, we observe a strict decrease of the value function over time. This is not the case for the EDMD-based MPC, for which we only have practical asymptotic stability of the origin. Correspondingly, VN(xμNε(;x^))V_{N}(x_{\mu_{N}^{\varepsilon}}(\cdot;\hat{x})) only decreases outside of a neighboorhood of the origin.

Refer to caption
Figure 2: Optimal value functions along the closed-loop of system (5.1) for nominal MPC (black) and EDMD-based MPC (gray) for horizons N=30N=30 (solid) and N=50N=50 (dashed) for λ=0.25\lambda=0.25.

The next example is taken from [21], where the parameter values can be found. Here, x˙(t)=f(x(t),Q)\dot{x}(t)=f(x(t),Q) describes an exothermic reaction that converts reactant AA to product BB and is given by

x˙(t)\displaystyle\dot{x}(t) =(FVr(CA0CA)k0eERTrCA2FVr(TA0Tr)ΔHρCpk0eERTrCA2+QρCpVr)\displaystyle=\begin{pmatrix}\frac{F}{V_{r}}(C_{A0}-C_{A})-k_{0}e^{\frac{-E}{RT_{r}}}C_{A}^{2}\\ \frac{F}{V_{r}}(T_{A0}-T_{r})-\frac{\Delta H}{\rho C_{p}}k_{0}e^{\frac{-E}{RT_{r}}}C_{A}^{2}+\frac{Q}{\rho C_{p}V_{r}}\end{pmatrix} (5.2)

with state x=(CA,Tr)2x=(C_{A},T_{r})^{\top}\in\mathbb{R}^{2}, where CAC_{A} is the concentration of AA, TrT_{r} the reactor temperature, and the control input QQ is the heat supplied to the reactor. Since we want to stabilize the controlled steady state xs=(CAs,Trs)=(1.907,300.6287)x^{s}=(C_{As},T_{rs})^{\top}=(1.907,300.6287)^{\top} (Qs=0kJ/hrQ^{s}=0\ \mathrm{kJ/hr}), we consider the shifted dynamics, for which is origin is a steady state.

For EDMD, we use d=1000d=1000 i.i.d. data points xix^{i} drawn from the state-constrained set 𝕏=[0.5,0.5]×[20,30]\mathbb{X}=[-0.5,0.5]\times[-20,30] and propagate them by Δt=102\Delta t=10^{-2} time units for control input u0=0u_{0}=0 and u1=1000u_{1}=1000, respectively. The dictionary consists of observables {1,x1,x2,x12,x22,e1/x1,e1/x2}\{1,x_{1},x_{2},x_{1}^{2},x_{2}^{2},e^{\nicefrac{{1}}{{x_{1}}}},e^{\nicefrac{{1}}{{x_{2}}}}\}. We consider the respective OCP subject to 𝕌=[10000,10000]\mathbb{U}=[-10000,10000] with weighting λ=106\lambda=10^{-6} and P=diag(102,1)P=\mathrm{diag}(10^{2},1) for control and state.

Figure 3 shows the numerical simulations emanating from the initial condition x0=(0.5,18)x_{0}=(0.5,-18)^{\top}. The decay in norm of the closed-loop state corresponding to the EDMD-based surrogate stagnates around 10210^{-2}, i.e., practical asymptotic stability can be observed in this example, too. For the considered horizons, the decreasing behavior in the beginning until the point of stagnation is reached is similar to that of nominal MPC. The fact that the convergence stagnates earlier for larger NN is not unexpected, because ωV\omega_{V} in Proposition 7(ii) may deteriorate since larger NN may render the optimal values more sensitive w.r.t. the initial condition.

0112233445510110^{-1}10210^{-2}10510^{-5}10810^{-8}101110^{-11}time tt
00.50.5111.51.52210210^{-2}10110^{-1}10010^{0}10110^{1}time ttN=5N=5N=10N=10N=15N=15N=20N=20N=30N=30N=40N=40
Figure 3: MPC closed loop xμN(;x0)\|x_{\mu_{N}}(\cdot;x_{0})\| (left) and xμNε(;x0)\|x_{\mu_{N}}^{\varepsilon}(\cdot;x_{0})\| (right, EDMD: d=1000d=1000) for system dynamics (5.2) for different horizons NN.

6 Conclusions

We proved practical asymptotic stability of data-driven MPC for nonlinear systems using EDMD embedded in the Koopman framework. To this end, we established a novel bound on the estimation error, which scales proportional to the norm of the state and the control. The underlying idea of imposing a certain structure in EDMD and, then, deriving proportional bounds was also key in follow-up work for controller design using the Koopman generator [33] and operator [34]. Then, we showed that cost controllability of the original model is preserved for the proposed data-based surrogate. Last, we provided two numerical examples to illustrate our findings and, in particular, the practical asymptotic stability of the origin.

References

  • [1] A. Boccia, L. Grüne, and K. Worthmann. Stability and feasibility of state constrained MPC without stabilizing terminal constraints. Systems & Control letters, 72:14–21, 2014.
  • [2] D. Bruder, X. Fu, and R. Vasudevan. Advantages of bilinear Koopman realizations for the modeling and control of systems with unknown dynamics. IEEE Robotics Automat. Lett., 6(3):4369–4376, 2021.
  • [3] S. L. Brunton, B. W. Brunton, J. L. Proctor, and J. N. Kutz. Koopman invariant subspaces and finite linear representations of nonlinear dynamical systems for control. PloS one, 11(2):e0150171, 2016.
  • [4] J.-M. Coron, L. Grüne, and K. Worthmann. Model predictive control, cost controllability, and homogeneity. SIAM Journal on Control and Optimization, 58(5):2979–2996, 2020.
  • [5] W. Esterhuizen, K. Worthmann, and S. Streif. Recursive feasibility of continuous-time model predictive control without stabilising constraints. IEEE Control Systems Letters, 5(1):265–270, 2020.
  • [6] C. Folkestad and J. W. Burdick. Koopman NMPC: Koopman-based learning and nonlinear model predictive control of control-affine systems. In IEEE International Conference on Robotics and Automation (ICRA), pages 7350–7356, 2021.
  • [7] D. Goswami and D. A. Paley. Bilinearization, reachability, and optimal control of control-affine nonlinear systems: A Koopman spectral approach. IEEE Trans. Automat. Control, 67(6):2715–2728, 2021.
  • [8] L. Grüne. Analysis and design of unconstrained nonlinear MPC schemes for finite and infinite dimensional systems. SIAM Journal on Control and Optimization, 48(2):1206–1228, 2009.
  • [9] L. Grüne and J. Pannek. Nonlinear model predictive control. Springer Cham, 2017.
  • [10] L. Grüne, J. Pannek, M. Seehafer, and K. Worthmann. Analysis of unconstrained nonlinear MPC schemes with time varying control horizon. SIAM Journal on Control and Optimization, 48(8):4938–4962, 2010.
  • [11] M. Haseli and J. Cortés. Learning Koopman eigenfunctions and invariant subspaces from data: Symmetric subspace decomposition. IEEE Transactions on Automatic Control, 67(7):3442–3457, 2021.
  • [12] M. Kanai and M. Yamakita. Linear model predictive control with lifted bilinear models by Koopman-based approach. SICE Journal of Control, Measurement, and System Integration, 15(2):162–171, 2022.
  • [13] M. Korda and I. Mezić. Linear predictors for nonlinear dynamical systems: Koopman operator meets model predictive control. Automatica, 93:149–160, 2018.
  • [14] M. Korda and I. Mezić. On convergence of extended dynamic mode decomposition to the Koopman operator. Journal of Nonlinear Science, 28(2):687–710, 2018.
  • [15] M. Korda and I. Mezić. Optimal construction of Koopman eigenfunctions for prediction and control. IEEE Transactions on Automatic Control, 65(12):5114–5129, 2020.
  • [16] A. Krolicki, D. Tellez-Castro, and U. Vaidya. Nonlinear dual-mode model predictive control using Koopman eigenfunctions. In 61st IEEE Conference on Decision and Control (CDC), pages 3074–3079, 2022.
  • [17] A. Lasota and M. C. Mackey. Chaos, fractals, and noise: stochastic aspects of dynamics. Springer New York, 2013.
  • [18] X. Ma, B. Huang, and U. Vaidya. Optimal quadratic regulation of nonlinear system using Koopman operator. In Proceedings of the 2019 IEEE American Control Conference (ACC), pages 4911–4916, 2019.
  • [19] G. Mamakoukas, M. L. Castano, X. Tan, and T. D. Murphey. Derivative-based Koopman operators for real-time control of robotic systems. IEEE Transactions on Robotics, 37(6):2173–2192, 2021.
  • [20] A. Mauroy, Y. Susuki, and I. Mezić. Koopman operator in systems and control. Springer Cham, 2020.
  • [21] A. Narasingam and J. S.-I. Kwon. Koopman lyapunov-based model predictive control of nonlinear chemical process systems. AIChE Journal, 65(11):e16743, 2019.
  • [22] A. Narasingam, S. H. Son, and J. S.-I. Kwon. Data-driven feedback stabilisation of nonlinear systems: Koopman-based model predictive control. International Journal of Control, 96(3):770–781, 2023.
  • [23] F. Nüske, S. Peitz, F. Philipp, M. Schaller, and K. Worthmann. Finite-data error bounds for Koopman-based prediction and control. Journal of Nonlinear Science, 33:14, 2023.
  • [24] S. E. Otto and C. W. Rowley. Koopman operators for estimation and control of dynamical systems. Annual Review of Control, Robotics, and Autonomous Systems, 4:59–87, 2021.
  • [25] S. Peitz, S. E. Otto, and C. W. Rowley. Data-driven model predictive control using interpolated Koopman generators. SIAM Journal on Applied Dynamical Systems, 19(3):2162–2193, 2020.
  • [26] F. M. Philipp, M. Schaller, K. Worthmann, S. Peitz, and F. Nüske. Error analysis of kernel EDMD for prediction and control in the Koopman framework. Preprint arXiv:2312.10460, 2024.
  • [27] F. M. Philipp, M. Schaller, K. Worthmann, S. Peitz, and F. Nüske. Error bounds for kernel-based approximations of the Koopman operator. Applied and Computational Harmonic Analysis, 71:101657, 2024.
  • [28] J. L. Proctor, S. L. Brunton, and J. N. Kutz. Dynamic mode decomposition with control. SIAM J. Appl. Dynam. Syst., 15(1):142–161, 2016.
  • [29] J. L. Proctor, S. L. Brunton, and J. N. Kutz. Generalizing Koopman theory to allow for inputs and control. SIAM Journal on Applied Dynamical Systems, 17(1):909–930, 2018.
  • [30] M. Schaller, K. Worthmann, F. Philipp, S. Peitz, and F. Nüske. Towards reliable data-based optimal and predictive control using extended DMD. IFAC-PapersOnLine, 56(1):169–174, 2023.
  • [31] S. H. Son, A. Narasingam, and J. S.-I. Kwon. Development of offset-free Koopman Lyapunov-based model predictive control and mathematical analysis for zero steady-state offset condition considering influence of Lyapunov constraints on equilibrium point. Journal of Process Control, 118:26–36, 2022.
  • [32] R. Strässer, J. Berberich, and F. Allgöwer. Control of bilinear systems using gain-scheduling: Stability and performance guarantees. In 62nd IEEE Conference on Decision and Control (CDC), pages 4674–4681, 2023.
  • [33] R. Strässer, M. Schaller, K. Worthmann, J. Berberich, and F. Allgöwer. Koopman-based feedback design with stability guarantees. IEEE Transactions on Automatic Control, 2024.
  • [34] R. Strässer, M. Schaller, K. Worthmann, J. Berberich, and F. Allgöwer. SafEDMD: A certified learning architecture tailored to data-driven control of nonlinear dynamical systems. Preprint arXiv:2402.03145, 2024.
  • [35] A. Surana. Koopman operator based observer synthesis for control-affine nonlinear systems. In 55th IEEE Conference on Decision and Control (CDC), pages 6492–6499, 2016.
  • [36] M. O. Williams, M. S. Hemati, S. T. Dawson, I. G. Kevrekidis, and C. W. Rowley. Extending data-driven Koopman analysis to actuated systems. IFAC-PapersOnLine, 49(18):704–709, 2016.
  • [37] M. O. Williams, I. G. Kevrekidis, and C. W. Rowley. A data–driven approximation of the Koopman operator: Extending dynamic mode decomposition. Journal of Nonlinear Science, 25:1307–1346, 2015.
  • [38] K. Worthmann. Stability analysis of unconstrained receding horizon control schemes. PhD thesis, University of Bayreuth, 2011. https://epub.uni-bayreuth.de/id/eprint/273/1/Dissertation_KarlWorthmann.pdf.
  • [39] S. Yu, C. Shen, and T. Ersal. Autonomous driving using linear model predictive control with a Koopman operator based bilinear vehicle model. IFAC-PapersOnLine, 55(24):254–259, 2022.
  • [40] X. Zhang, W. Pan, R. Scattolini, S. Yu, and X. Xu. Robust tube-based model predictive control with Koopman operators. Automatica, 137:110114, 2022.