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

Koopman Bilinearization of Nonlinear Control Systems

Wei Zhang, and Jr-Shin Li W. Zhang is with the Department of Electrical and Systems Engineering, Washington University, St. Louis, MO 63130, USA [email protected]J.-S. Li is with the Department of Electrical and Systems Engineering, Washington University, St. Louis, MO 63130, USA [email protected]
Abstract

Koopman operators, since introduced by the French-born American mathematician Bernard Koopman in 1931, have been employed as a powerful tool for research in various scientific domains, such as ergodic theory, probability theory, geometry, and topology, with widespread applications ranging from electrical engineering and machine learning to biomedicine and healthcare. The current use of Koopman operators mainly focuses on the characterization of spectral properties of ergodic dynamical systems. In this paper, we step forward from unforced dynamical systems to control systems and establish a systematic Koopman control framework. Specifically, we rigorously derive a differential equation system governing the dynamics of the Koopman operator associated with a control system, and show that the resulting system is a bilinear system evolving on an infinite-dimensional Lie group, which directly leads to a global bilinearization of control-affine systems. Then, by integrating techniques in geometric control theory with infinite-dimensional differential geometry, this further offers a two-fold benefit to controllability analysis: the characterization of controllability for control-affine systems in terms of de Rham differential operators and the extension of the Lie algebra rank condition to systems defined on infinite-dimensional Lie groups. To demonstrate the applicability, we further adopt the established framework to develop the Koopman feedback linearization technique, which, as one of the advantages, waives the controllability requirement for the systems to be linearized by using the classical feedback linearization technique. In addition, it is worth mentioning that a distinctive feature of this work is the maximum utilization of the intrinsic algebraic and geometric properties of Koopman operators, instead of spectral methods and the ergodicity assumption of systems, which further demonstrate a significant advantage of our work and a substantial difference between the presented and existing research into Koopman operators.

Index Terms:
Geometric control theory, Koopman operators, nonlinear systems, Lie groups.

I Introduction

Koopman operators, named after the French-born American mathematician Bernard Osgood Koopman, was originally introduced in 1931 for the purpose of adopting Hilbert space methods, state-of-the-art techniques just coming into prominence at that time, to investigate spectral properties of Hamiltonian systems [1]. Since then, the potential of Koopman operators has started to be widely recognized and extensively explored in numerous scientific domains. First and foremost, Koopman operators lie in the repertoire of the most powerful tools in ergodic theory, especially for the characterization of ergodicity, recurrence, mixing, and topological entropy of (measure-preserving) dynamical systems [2, 3, 4]. These seminal research works further opened up various interdisciplinary applications of Koopman operators, notably, the study of geodesic flows on Riemannian manifolds in geometry and topology [5, 6], the analysis of statistical properties of Markov processes in probability theory [7, 8], and the construction of Diophantine approximation in number theory [9, 10]. Recently, a large body of research into Koopman operators have been concentrating on the computational, data-driven, and learning aspects, particularly, for inferring temporal evolutions of dynamical systems from measurement data by using spectral-type methods, e.g., Arnoldi, vector Prony, and dynamic mode decomposition methods [12, 11, 16]. Dominant works in this line of research include spectral decompositions of fluid mechanical systems [12, 13, 14] in mechanical engineering, stability analysis of power systems in electrical engineering [15, 11, 16], cognitive classification and seizure detection in biomedicine and healthcare, respectively [17], and multi-modal learning and prediction in biostatistics [18].

Although Koopman operators have demonstrated their capabilities in the study of (unforced) dynamical systems, their roles in control theory for analyzing and manipulating dynamical systems with external inputs still largely remain as an unexplored avenue. To fill in this gap, in this paper, we rigorously establish a systematic Koopman control framework, particularly for tackling control tasks involving nonlinear systems, which in turn simultaneously expands the repertoire of techniques in geometric control theory and the scope of Koopman operator theory. In particular, we start from the rigorous derivation of operator differential equations governing the dynamics of the Koopman operators, referred to as Koopman systems, associated with nonlinear systems by leveraging the theory of semigroups in functional analysis. Built upon this and inspired by the lifting property of Koopman operators that model the dynamics of nonlinear systems on finite-dimensional manifolds by linear transformations on infinite-dimensional vector spaces, we define the notation of Koopman bilinearization for control-affine systems, which then leads to the characterization of controllability for such systems in terms of de Rham differential operators. Independently, we also show that the resulting bilinear Koopman systems evolve on some infinite-dimensional Lie groups, and then investigate their controllability properties by integrating techniques in geometric control theory and infinite-dimensional differential geometry. In particular, this gives rise to an extension of the Lie algebra rank condition (LARC) to systems defined on infinite-dimensional Lie groups. The established Koopman control framework is then adopted in the development of the Koopman feedback linearization technique, which not only demonstrates the applicability of this Koopman control framework but also carries out an extension of the classical feedback linearization technique, e.g., the controllability requirement for the systems to be feedback linearized is waived.

It is worth mentioning that, in this work, the leverage of Koopman operators does not require any measure-preserving property or ergodicity of the corresponding dynamical systems, and the development of the Koopman control framework does not involve spectral methods, either. This unique feature not only highlights the major advantage of our approach but also indicates a substantial difference between the presented and existing works on Koopman operators. In addition, owing to the infinite-dimensional nature of Koopman systems and their intimacy with finite-dimensional systems, this work further sheds light on a general framework of analysis and control of infinite-dimensional systems by using finite-dimensional methods.

The paper is organized as follows. In Section II, we define Koopman operators from the perspective of dynamical systems theory, and then launch a detailed investigation into their algebraic and analytic properties, which provide the tools for the derivation of the differential equations governing the dynamics of Koopman systems in a coordinate-free way in Section III. The focus of Section IV-A, the main section of the paper, is on the study of Koopman systems associated with control systems. In particular, we introduce the Koopman bilinearization technique for control-affine systems, which leads to the de Rham differential operator characterization of controllability for such systems. In parallel, controllability of bilinear Koopman systems are also systematically studied, leading an extension of the LARC to systems defined on infinite-dimensional Lie groups. At last, in Section V, the developed framework is adopted to extend the classical feedback linearization technique from the perspective of Koopman systems.

II Koopman Operators of Dynamical Systems

In this section, we will briefly review the Koopman operator theory from the dynamical systems viewpoint. The emphasis will be placed on investigating the analytical properties of Koopman operators as linear operators for preparing the operator-theoretic analysis of control systems in the following sections. In addition, it is worth mentioning that, unlike many existing literatures on Koopman operator theory, our approach do not require any ergodicity or measure-preserving property of dynamical systems.

II-A Flows of dynamical systems

Given a dynamical system evolving on a finite-dimensional smooth manifold MM, defined by the ordinary differential equation

ddtx(t)=f(x(t)),\displaystyle\frac{d}{dt}x(t)=f\big{(}x(t)\big{)}, (1)

with ff a smooth vector field on MM, we suppose that ff is complete, namely, the solution of the differential equation starting from any point in MM exists for all tt\in\mathbb{R}. This property holds, e.g., in the cases that ff is compactly supported, MM is compact, or MM is a Lie group with ff a left- or -right invariant vector field [19].

Under the completeness assumption, the vector field ff generates a smooth left action of the additive group \mathbb{R} on MM, called the (global) flow or one-paramater group action of ff, given by the smooth map Φ:×MM\Phi:\mathbb{R}\times M\rightarrow M sending (t,x)(t,x) to the solution of the system in (1) at the time tt starting from the initial condition xx. As a group action, the flow Φ\Phi satisfies

Φ(t,Φ(s,x))=Φ(s+t,x),Φ(0,x)=x\displaystyle\Phi(t,\Phi(s,x))=\Phi(s+t,x),\quad\Phi(0,x)=x (2)

for all s,ts,t\in\mathbb{R} and xMx\in M, and further gives rises to two collections of maps:

  • Φt:MM\Phi_{t}:M\rightarrow M given by xΦ(t,x)x\mapsto\Phi(t,x) for each tt\in\mathbb{R}. The completeness of ff and uniqueness of solutions of the ordinary differential equation in (1) implies that Φt\Phi_{t} is a diffeomorphism with Φt1=Φt\Phi^{-1}_{t}=\Phi_{-t}, and we also refer to Φt\Phi_{t} as the time-tt flow of the system. Moreover, the properties of Φ\Phi in (2) are equivalent to the group laws: ΦtΦs=Φs+t\Phi_{t}\circ\Phi_{s}=\Phi_{s+t} and Φ0=IdM\Phi_{0}={\rm Id}_{M}, where IdM{\rm Id}_{M} is the identity map on MM. In another word, the family of maps {Φt}t\{\Phi_{t}\}_{t\in\mathbb{R}} is a subgroup of the diffeomorohism group of MM.

  • Φx:M\Phi^{x}:\mathbb{R}\rightarrow M given by tΦ(t,x)t\mapsto\Phi(t,x) for each xMx\in M. Then, Φx\Phi^{x} is exactly the integral curve of ff passing through xx. Infinitesimally, this simply means that Φx\Phi^{x} satisfies the differential equation in (1) for any xMx\in M as

    tΦ(t,x)=f(Φ(t,x))\displaystyle\frac{\partial}{\partial t}\Phi(t,x)=f(\Phi(t,x))

    or equivalently, in the coordinate-free representation

    Φ(t)=fΦt\displaystyle\Phi_{*}\Big{(}\frac{\partial}{\partial t}\Big{)}=f\circ\Phi_{t} (3)

    where Φ=dΦ:T(×M)TM\Phi_{*}=d\Phi:T(\mathbb{R}\times M)\rightarrow TM is the pushforward (differential) of Φ\Phi, and T(×M)T(\mathbb{R}\times M) and TMTM are the tangent bundles of ×M\mathbb{R}\times M and MM, respectively. The fact of Φt\Phi_{t} being a diffeomorohism guarantees that Φ\Phi_{*} is globally well-defined and smooth.

II-B Koopman operators

In practice, knowledge about a dynamical system may only be acquired from some observables of the system. Following the notations used in the previous section, the uniqueness of solutions of the system in (1) indicates that the state variable x(t)x(t) completely represents the dynamics of the system, and hence data collected from observables of the system must be in the form of functions of x(t)x(t). Mathematically, this implies that observables of the system are just functions defined on the state-space of the system. To be consistent with the smoothness of the state-space manifold MM and the vector field ff governing the system dynamics, we focus on observables in C(M)C^{\infty}(M), the space of real-valued smooth functions defined on MM. Naturally, the flow on MM generated by the dynamics of the system in (1) induces a flow on the space of observables C(M)C^{\infty}(M) of the system. This then opens up the possibility of understanding the dynamics of the system by using the flow on the space of its observables, which also motivates the introduction of Koopman operators.

II-B1 Koopman operators and Koopman groups

Formally, for each tt\in\mathbb{R}, we define the Koopman operator UtU_{t} of the system as the composition operator with the time-tt flow of the system, that is, Uth=hΦtU_{t}h=h\circ\Phi_{t} for any hC(M)h\in C^{\infty}(M). Because the function composition operation is linear, UtU_{t} is a linear operator on C(M)C^{\infty}(M). Together with the fact that the composition hΦth\circ\Phi_{t} of the smooth functions hh and Φt\Phi_{t} is also a smooth function in C(M)C^{\infty}(M), UtU_{t} maps C(M)C^{\infty}(M) into itself. Moreover, as tt varying, the group laws satisfied by the time-tt flow Φt\Phi_{t} are also inherited by the Koopman operator UtU_{t}, i.e.,

U0=IandUs+t=UsUt\displaystyle U_{0}=I\quad\text{and}\quad U_{s+t}=U_{s}U_{t} (4)

for any s,ts,t\in\mathbb{R}, where II denotes the identity map on C(M)C^{\infty}(M). In another word, the Koopman operator UtU_{t} gives rise to a flow on C(M)C^{\infty}(M), and the “vector field”, or equivalently the dynamical system, on C(M)C^{\infty}(M) generating this flow is of great interest and importance in this paper, which will be fully investigated from the analytical perspective in Section III. On the other hand, implied by the invertibility of Φt\Phi_{t} with the inverse Φt\Phi_{-t}, each UtU_{t} is also invertible with the inverse given by UtU_{-t}, that is, the map hhΦth\mapsto h\circ\Phi_{-t}. Therefore, the family of operators {Ut}t\{U_{t}\}_{t\in\mathbb{R}} indeed defines a one-parameter subgroup of the group of invertible linear operators from C(M)C^{\infty}(M) onto itself under the group operation of operator compositions, and it will be referred to as the Koopman group associated with the dynamical system in (1).

II-B2 Analytical properties of Koopman operators

To investigate the analytical properties of the Koopman operator UtU_{t}, it is inevitable to leverage the topology on the observable space C(M)C^{\infty}(M). To motivate the idea, we recall that in the case of MM being an open subspace of the Euclidean space n\mathbb{R}^{n}, the topology on C(M)C^{\infty}(M) generated by the family of seminorms

hK,α=supxK|αxαh(x)|\displaystyle\|h\|_{K,\alpha}=\sup_{x\in K}\Big{|}\frac{\partial^{\alpha}}{\partial x^{\alpha}}h(x)\Big{|}

for any multi-index αn\alpha\in\mathbb{N}^{n} and compact subset KMK\subset M, called the CC^{\infty}-topology on C(M)C^{\infty}(M), makes C(M)C^{\infty}(M) a locally convex topological vector space [20]. For a general finite-dimensional smooth manifold MM, because MM is locally homeomorphic to an Euclidean space, the aforementioned construction of CC^{\infty}-topology for C(M)C^{\infty}(M) will be still valid if the differentiation operations in those seminorms α,K\|\cdot\|_{\alpha,K} are well-defined on MM. In particular, a suitable replacement for the partial derivatives is a covariant derivative D:C(M)C(TM)D:C^{\infty}(M)\rightarrow C^{\infty}(T^{*}M) on MM, where TMT^{*}M denotes the cotangent bundle of MM, and DD can always be globally defined in a coordinate-free way, e.g, by using the Levi-Civita connection of a Riemannian metric on MM. Another consequence of MM being a manifold is that it is a σ\sigma-compact locally compact Hausdorff space [19], and hence MM has an open cover consisting of an increasing sequence of precompact open sets {Ui}i\{U_{i}\}_{i\in\mathbb{N}} such that U¯iUi+1\overline{U}_{i}\subset U_{i+1} for all ii\in\mathbb{N}, where U¯m\overline{U}_{m} denotes the closure of UmU_{m} (with respect to the topology of MM) [21]. As a result, every compact subset of MM is contained in at least one of such open set UmU_{m}. Integrating these observations, the CC^{\infty}-topology on C(M)C^{\infty}(M) is generated by the countable family of seminorms

hi,j=supxU¯i|Djh(x)|,\displaystyle\|h\|_{i,j}=\sup_{x\in\overline{U}_{i}}\Big{|}D^{j}h(x)\Big{|}, (5)

where Dj:C(M)C(jTM)D^{j}:C^{\infty}(M)\rightarrow C^{\infty}(\otimes^{j}T^{*}M), playing the role of the jthj^{\rm th} power (composition with itself) of DD, is the covariant derivative on the sections of the tensor bundle jTM\otimes^{j}T^{*}M induced by DD, and jTM\otimes^{j}T^{*}M denotes the jj times tensor of TMT^{*}M. In another word, C(M)C^{\infty}(M) is a Frechét space. Note that although different choices of the open cover {Ui}i\{U_{i}\}_{i\in\mathbb{N}} and covariant derivative DD result in different families of seminorms, but they generate the same CC^{\infty}-topology on C(M)C^{\infty}(M).

From the analytical viewpoint, a major use of the topology on a vector space is to evaluate continuity, equivalently boundedness, of linear operators. For example, in the case of the CC^{\infty}-topology on C(M)C^{\infty}(M), the covariant derivative DjD^{j}, which is also an order jj partial differential operator on MM, is bounded linear operator as Djhi,k=hi,k+j\|D^{j}h\|_{i,k}=\|h\|_{i,k+j} for all i,j,ki,j,k\in\mathbb{N}. This lays the foundation for the Koopman linearization method proposed in the following sections, where we use the Lie derivative as the differential operator DD. The task now is to show that the Koopman operator UtU_{t} is also bounded.

Proposition 1.

Given a smooth dynamical system defined on a smooth manifold MM as in (1), the associated Koopman operator Ut:C(M)C(M)U_{t}:C^{\infty}(M)\rightarrow C^{\infty}(M) is a bounded linear operator under the CC^{\infty}-topology on C(M)C^{\infty}(M).

Proof.

To show the boundedness, equivalently, continuity of UtU_{t} under the topology generated by the family of seminorms in (5), it suffices to prove that for any i,ji,j\in\mathbb{N}, there exist i1,,in,j1,,jni_{1},\dots,i_{n},j_{1},\dots,j_{n}\in\mathbb{N} and C>0C>0 such that Uthi.jCsupik,jkhik,jk\|U_{t}h\|_{i.j}\leq C\sup_{i_{k},j_{k}}\|h\|_{i_{k},j_{k}} [20].

The chain rule for covariant derivatives yields

Uthi,j\displaystyle\|U_{t}h\|_{i,j} =supxU¯i|Djh(Φt(x))|\displaystyle=\sup_{x\in\overline{U}_{i}}\big{|}D^{j}h(\Phi_{t}(x))\big{|}
=supxU¯i|k=0jDkh(Φt(x))pk(Φt(x),,DkΦt(x))|\displaystyle=\sup_{x\in\overline{U}_{i}}\big{|}\sum_{k=0}^{j}D^{k}h(\Phi_{t}(x))p_{k}(\Phi_{t}(x),\dots,D^{k}\Phi_{t}(x))\big{|}

for some polynomial functions pkp_{k} in kk variables, k=0,,jk=0,\dots,j. Let Ck=supxU¯ipk(Φt(x),,DkΦt(x))C_{k}=\sup_{x\in\overline{U}_{i}}p_{k}(\Phi_{t}(x),\dots,D^{k}\Phi_{t}(x)), then it leads to

Uthi,j\displaystyle\|U_{t}h\|_{i,j} k=0jCksupxU¯i|Dkh(Φt(x))|\displaystyle\leq\sum_{k=0}^{j}C_{k}\sup_{x\in\overline{U}_{i}}\big{|}D^{k}h(\Phi_{t}(x))\big{|}
=k=0jCksupyΦ(U¯i)|Dkh(y)|.\displaystyle=\sum_{k=0}^{j}C_{k}\sup_{y\in\Phi(\overline{U}_{i})}\big{|}D^{k}h(y)\big{|}.

Because Φt\Phi_{t} is smooth and Ui¯\overline{U_{i}} is compact, Φi(Ui¯)\Phi_{i}(\overline{U_{i}}) is also compact and contained in some UlU_{l}, which leads to the desired result

Uthi,j\displaystyle\|U_{t}h\|_{i,j}\leq k=0jCksupxU¯l|Dkh(x)|=k=0jCksupxU¯lh(x)l,k\displaystyle\sum_{k=0}^{j}C_{k}\sup_{x\in\overline{U}_{l}}\big{|}D^{k}h(x)\big{|}=\sum_{k=0}^{j}C_{k}\sup_{x\in\overline{U}_{l}}\|h(x)\|_{l,k}
Csupk=0,,jhl,k\displaystyle\leq C\sup_{k=0,\dots,j}\|h\|_{l,k}

by setting C=jmaxCkC=j\max C_{k}. ∎

Recall that all bounded linear transformations on C(M)C^{\infty}(M) form a group Aut(C(M)){\rm Aut}(C^{\infty}(M)), called the automorphism group of C(M)C^{\infty}(M). Then, Proposition 1 can also be equivalently presented in algebraic terminologies as follows.

Corollary 1.

The Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}} is a one parameter subgroup of the automorphism group Aut(C(M)){\rm Aut}(C^{\infty}(M)) of C(M)C^{\infty}(M).

This algebraic characterization of Koopman operators in Corollary 1 opens up the possibility to associate infinitesimal generators, conceptually derivatives with respect to the time variable tt, to Koopman operators, which then naturally govern the dynamics of ordinary differential equation systems defined on Aut(C(M)){\rm Aut}(C^{\infty}(M)). The focus of the next section will be on the study of these Koopman operator-induced systems, with the emphasis on the relationship between nonlinear systems on MM and their Koopman systems on Aut(C(M)){\rm Aut}(C^{\infty}(M)).

III Koopman Linearization of Nonlinear Dynamical Systems

Koopman operators are defined to be associated with dynamical systems, and this immediately brings to the generic question of how the dynamics of a system reflects on the associated Koopman operator. To gain some insight into this question, taking the system in (1) as an example, we first observe from (3) that the vector field ff governing the system dynamics can be recovered from the flow Φ\Phi by taking the time derivative. Together with the fact that the Koopman operator UtU_{t} associated with the system is the composition operator with respect to the time-tt flow Φt\Phi_{t}, this observation gives a clue about the intimacy between the system dynamics and the time derivative of UtU_{t}, which will be fully explored, mainly from the control-theoretical perspective, in this section. To be more specific, we will derivative a differential equation system governing the dynamics of the Koopman operator UtU_{t}, towards the goal of investigating control-related properties of the system by using the associated Koopman system.

III-A Infinitesimal generators of Koopman groups

As motivated below Corollary 1, the differential equation system governing the dynamics of the Koopman operator UtU_{t} can be obtained from the infinitesimal generator of the Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}}. To guarantee the existence of the infinitesimal generator of {Ut}t\{U_{t}\}_{t\in\mathbb{R}}, it is required that {Ut}t\{U_{t}\}_{t\in\mathbb{R}} is strongly continuous, i.e., UtU0U_{t}\rightarrow U_{0} as t0t\rightarrow 0 in the strong operator topology, equivalently, UthU0h=hU_{t}h\rightarrow U_{0}h=h as t0t\rightarrow 0 for all hC(M)h\in C^{\infty}(M) in the CC^{\infty}-topology [22].

Lemma 1.

For any dynamical system defined on a smooth manifold MM as in (1), the associated Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}} is a strongly continuous one parameter group of continuous linear operators on C(M)C^{\infty}(M).

Proof.

Let Φ:×MM\Phi:\mathbb{R}\times M\rightarrow M and UtU_{t} denote the flow and Koopman operator of the system, respectively, and pick any hC(M)h\in C^{\infty}(M). Because the CC^{\infty}-topology on C(M)C^{\infty}(M) is generated by the family of seminorms i,j\|\cdot\|_{i,j} on MM, UthhU_{t}h\rightarrow h is equivalent to Uthhi,j0\|U_{t}h-h\|_{i,j}\rightarrow 0 for all i,ji,j\in\mathbb{N}. To prove the seminorm convergence, by the definition of i,j\|\cdot\|_{i,j}, we have

limt0Uthhi,j\displaystyle\lim_{t\rightarrow 0}\|U_{t}h-h\|_{i,j} =limt0supxU¯i|Dj(h(Φt(x))h(x))|\displaystyle=\lim_{t\rightarrow 0}\sup_{x\in\overline{U}_{i}}\big{|}D^{j}\big{(}h(\Phi_{t}(x))-h(x)\big{)}\big{|}
supxU¯i|limt0Dj(h(Φt(x))h(x))|\displaystyle\leq\sup_{x\in\overline{U}_{i}}\big{|}\lim_{t\rightarrow 0}D^{j}\big{(}h(\Phi_{t}(x))-h(x)\big{)}\big{|}
supxU¯i|Dj(h(limt0Φt(x))h(x))|,\displaystyle\leq\sup_{x\in\overline{U}_{i}}\big{|}D^{j}\big{(}h(\lim_{t\rightarrow 0}\Phi_{t}(x))-h(x)\big{)}\big{|}, (6)

where the second and third inequalities follow from the lower semicontinuity of the supremum function and uniform continuity of DjhD^{j}h on the compact set U¯i\overline{U}_{i} as a section of jTM\otimes^{j}T^{*}M, respectively. Next, in (6), as a function of tt, Φt(x)=Φ(t,x)\Phi_{t}(x)=\Phi(t,x) is also continuous, and hence limtΦt(x)=Φ(0,x)=x\lim_{t\rightarrow}\Phi_{t}(x)=\Phi(0,x)=x holds for all xU¯jx\in\overline{U}_{j}, giving the desired convergence limt0Uthhi,j=0\lim_{t\rightarrow 0}\|U_{t}h-h\|_{i,j}=0. ∎

Using more elementary terminology, Lemma 1 simply means that the operator-valued function Aut(C(M))\mathbb{R}\rightarrow{\rm Aut}(C^{\infty}(M)) given by tUtt\mapsto U_{t} is continuous. Therefore, similar to the basic calculus, it is possible to take the derivative of UtU_{t} with respect to tt, that is, conceptually,

ddt|t=t0Ut\displaystyle\frac{d}{dt}\Big{|}_{t=t_{0}}U_{t} =limtt0UtUt0tt0=limtt0Ut0(Utt0I)tt0\displaystyle=\lim_{t\rightarrow t_{0}}\frac{U_{t}-U_{t_{0}}}{t-t_{0}}=\lim_{t\rightarrow t_{0}}\frac{U_{t_{0}}(U_{t-t_{0}}-I)}{t-t_{0}}
=Ut0lims0UsIs=Ut0ddt|t=0Ut,\displaystyle=U_{t_{0}}\lim_{s\rightarrow 0}\frac{U_{s}-I}{s}=U_{t_{0}}\frac{d}{dt}\Big{|}_{t=0}U_{t}, (7)

where the second equality follows from the group law for the Koopman operator. Moreover, this also implies that the derivative of UtU_{t} at any time tt is completely determined by its derivative at time 0, which is known as the infinitesimal generator of the Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}}. Technically, because the continuity of {Ut}t\{U_{t}\}_{t\in\mathbb{R}} holds in the strong operator topology, it is required that the derivative is valid in the same topology, meaning, the limit

ddt|t=0Uth=limt0Uthht\displaystyle\frac{d}{dt}\Big{|}_{t=0}U_{t}h=\lim_{t\rightarrow 0}\frac{U_{t}h-h}{t} (8)

converges for all hC(M)h\in C^{\infty}(M) in the CC^{\infty}-topology.

Theorem 1.

The infinitesimal generator of the Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}} associated with the dynamical system ddtx(t)=f(x(t))\frac{d}{dt}x(t)=f(x(t)) evolving on MM is the Lie derivative f\mathcal{L}_{f} with respect to the vector field fC(TM)f\in C^{\infty}(TM).

Proof.

Pick any hC(M)h\in C^{\infty}(M), by considering hh as a 0-tensor field on MM, Uth=hΦt=ΦthU_{t}h=h\circ\Phi_{t}=\Phi^{*}_{t}h is just the pull back of hh by the smooth function Φt\Phi_{t} so that

fh=limt0Φthht=limt0Uthht=ddt|t=0Uth\displaystyle\mathcal{L}_{f}h=\lim_{t\rightarrow 0}\frac{\Phi_{t}^{*}h-h}{t}=\lim_{t\rightarrow 0}\frac{U_{t}h-h}{t}=\frac{d}{dt}\Big{|}_{t=0}U_{t}h (9)

following from the definition of Lie derivatives. The fact fhC(M)\mathcal{L}_{f}h\in C^{\infty}(M) then implies that the the limit in (9) holds in the CC^{\infty}-topology on C(M)C^{\infty}(M), concluding the proof. ∎

In general, in a topological vector space that is not a Banach space, e.g. the Frechét space C(M)C^{\infty}(M) concerned in this paper, equicontinuity of a semigroup of operators in the time variable tt is usually a requirement, in addition to strong continuity, to guarantee the infinitesimal generator to be densely defined on its domain [22]. Unfortunately, Koopman groups generally do not satisfy the equicontintuity condition, for if so, then given any seminorm i,j\|\cdot\|_{i,j}, there is a seminorm k,l\|\cdot\|_{k,l} such that Uthi,jhk,l\|U_{t}h\|_{i,j}\leq\|h\|_{k,l} for all tt\in\mathbb{R} and hC(M)h\in C^{\infty}(M). For example, consider the system ddtx(t)=1\frac{d}{dt}x(t)=1 defined on \mathbb{R}, whose flow is given by Φ(t,x)=x+t\Phi(t,x)=x+t so that Uth(x)=h(x+t)U_{t}h(x)=h(x+t) for all hC()h\in C^{\infty}(\mathbb{R}). However, in the case of h(x)=exh(x)=e^{x}, Uth(x)=ex+tU_{t}h(x)=e^{x+t} escapes every compact subset of \mathbb{R} when tt\rightarrow\infty, and hence UtU_{t} fails to be equicontinuous. Even in such a case, Theorem 1 implies that the infinitesimal generator of the Koopman group is still defined on the whole space C(M)C^{\infty}(M), not only a dense subset of C(M)C^{\infty}(M), since the Lie derivative is always well-defined for all smooth function on MM, which definitely exhibits one of the most outstanding features of Koopman operators.

III-B Coordinate-free Koopman linearization

A recapitulation of the derivation in (7) that the derivative of the Koopman operator UtU_{t} at any time is completely determined by the infinitesimal generator f\mathcal{L}_{f} in a linear way sheds light on the construction of a linear differential equation system on Aut(C(M)){\rm Aut}(C^{\infty}(M)) to describe the dynamics of the Koopman operator.

Theorem 2.

Given a dynamical system defined on a smooth manifold MM as in (1), that is,

ddtx(t)=f(x(t)),\frac{d}{dt}x(t)=f(x(t)),

the Koopman operator UtU_{t} associated with the system satisfies a linear differential equation system on Aut(C(M)){\rm Aut}(C^{\infty}(M)), given by,

ddtUt=Utf=fUt.\displaystyle\frac{d}{dt}U_{t}=U_{t}\mathcal{L}_{f}=\mathcal{L}_{f}U_{t}. (10)
Proof.

Pick any hC(M)h\in C^{\infty}(M), the differential equation in (10) follows from computing the derivative of UthU_{t}h with respect to tt in two ways. At first, the chain rule yields

ddtUth\displaystyle\frac{d}{dt}U_{t}h =ddt(hΦt)=d(hΦt)(ddt)=dhdΦ(t)\displaystyle=\frac{d}{dt}(h\circ\Phi_{t})=d(h\circ\Phi_{t})\Big{(}\frac{d}{dt}\Big{)}=dh\circ d\Phi\Big{(}\frac{\partial}{\partial t}\Big{)}
=dh(f)Φt=fhΦt=Utfh,\displaystyle=dh(f)\circ\Phi_{t}=\mathcal{L}_{f}h\circ\Phi_{t}=U_{t}\mathcal{L}_{f}h,

where we use the definition of the flow in (3) and the fact of the Lie derivative f\mathcal{L}_{f} acting on the 0-tensor, i.e., smooth function, hh as the directional derivative fh=dh(f)\mathcal{L}_{f}h=dh(f) [19].

Alternatively, by using the antisymmetry of Lie derivative actions on vector fields as ff=[f,f]=0\mathcal{L}_{f}f=[f,f]=0, where [,][\cdot,\cdot] denotes the Lie bracket of vector fields, we have dΦt(f)=fΦtd\Phi_{t}(f)=f\circ\Phi_{t}, and then it follows

fUth=d(Uth)(f)=d(hΦt)(f)\displaystyle\mathcal{L}_{f}U_{t}h=d(U_{t}h)(f)=d(h\circ\Phi_{t})(f)
=dhdΦt(f)=dh(f)Φt=Utfh\displaystyle=dh\circ d\Phi_{t}(f)=dh(f)\circ\Phi_{t}=U_{t}\mathcal{L}_{f}h

as desired. ∎

As indicated by the proof of Theorem 2, the system in (10) governing the dynamics of the Koopman operator holds in the strong operator topology induced by the CC^{\infty}-topology on CC^{\infty} and we refer to this system as the Koopman system associated with the system in (1). Although proved computationally, the commutativity of UtU_{t} and f\mathcal{L}_{f} in the Koopman system in (10) also follows abstractly from the general facts in functional analysis that the infinitesimal generator of a strongly continuous semigroup of everywhere defined operators is always densely defined and commutes with every operator in the semigroup [22].

Remark 1 (Coordinate-free global Koopman linearization).

Notice that because the Lie derivative f\mathcal{L}_{f} is a linear operator, to be more specific, a first-order partial differential operator, the Koopman system in (10) is always a linear system regardless of the nonlinearity of the associated system in (1). Together with that the representation of the Koopman system does not involve the coordinates of MM, the linearization procedure from the nonlinear system to the associated Koopman system must hold globally. As a summary, Koopman systems give rise to coordinate-free global linearization of nonlinear system.

To gain some intuition into the mechanism hidden behind the Koopman system, we recall from linear systems theory that, for any time-invariant linear system defined on n\mathbb{R}^{n}, say ddtx(t)=Ax(t)\frac{d}{dt}x(t)=Ax(t) with x(t)nx(t)\in\mathbb{R}^{n} and An×nA\in\mathbb{R}^{n\times n}, the transition matrix etAe^{tA} of the system satisfies [23]:

  • The group laws: etA|t=0=Ie^{tA}|_{t=0}=I, the nn-by-nn identity matrix, and e(s+t)A=esAetAe^{(s+t)A}=e^{sA}e^{tA} for any s,ts,t\in\mathbb{R}.

  • The matrix differential equation

    ddtetA=AetA=etAA.\displaystyle\frac{d}{dt}e^{tA}=Ae^{tA}=e^{tA}A. (11)

In addition, as a fundamental solution, any solution of the system can be generated by the transition matrix as x(t)=etAx0x(t)=e^{tA}x_{0} for some x0nx_{0}\in\mathbb{R}^{n}, equivalently,

ddtetAx0=AetAx0,\displaystyle\frac{d}{dt}e^{tA}x_{0}=Ae^{tA}x_{0}, (12)

and etAe^{tA} admits the power series expansion for the exponential function as

etA=k=0tkk!Ake^{tA}=\sum_{k=0}^{\infty}\frac{t^{k}}{k!}A^{k}

with the convergence radius \infty, i.e., the series converges for any An×nA\in\mathbb{R}^{n\times n} [23].

A comparison between the properties satisfied by the Koopman operator UtU_{t} and the transition matrix etAe^{tA} reviewed above, particularly the operator differential equation in (10) and the matrix differential equation in (11), immediately gives an interpretation of UtU_{t} from the perspective of linear systems theory: the Koopman operator UtU_{t} is the fundamental solution, playing the role of the “transition matrix” as for a finite-dimensional time-invariant linear system, of the infinite-dimensional linear system on C(M)C^{\infty}(M)

ddt(Uth)=f(Uth),\displaystyle\frac{d}{dt}(U_{t}h)=\mathcal{L}_{f}(U_{t}h), (13)

which can be considered as the infinite-dimensional analogue of the system in (12). Then, similarly, the power series expansion

Ut=exp(tf)=k=0tkk!fkU_{t}=\exp(t\mathcal{L}_{f})=\sum_{k=0}^{\infty}\frac{t^{k}}{k!}\mathcal{L}_{f}^{k}

also holds and converges in the strong operator topology with the convergence radius \infty so that any solution of the system in (13) admits the representation Uth=exp(tf)hU_{t}h=\exp(t\mathcal{L}_{f})h for some hC(M)h\in C^{\infty}(M), which in turn gives rise to a functional calculus definition of the Koopman operator UtU_{t} as the exponential of the Lie derivative operator f\mathcal{L}_{f}.

On the other hand, by using the coordinates on MM, the measurement data of the system generated by the observable hC(M)h\in C^{\infty}(M) is given by y(t)=h(x(t))y(t)=h(x(t)). The data can be further represented in terms of the Koopman operator UtU_{t} as y(t)=h(Φt(x0))=Uth(x(0))y(t)=h(\Phi_{t}(x_{0}))=U_{t}h(x(0)), and hence the equation in (13) exactly recovers the dynamics of the data in the form of the differential equation

ddty(t)=fh(x(t))=dh(f)(x(t)),\displaystyle\frac{d}{dt}y(t)=\mathcal{L}_{f}h(x(t))=dh(f)(x(t)), (14)

where we use the fact that the Lie derivative f\mathcal{L}_{f} acts on the 0-tensor hC(M)h\in C^{\infty}(M) as the directional derivative along the direction ff. In the special case of M=nM=\mathbb{R}^{n}, the vector field ff can be identified with an nn-tuple of smooth functions on n\mathbb{R}^{n}, and so is dhdh, regarded as the gradient h\nabla h of hh. Then, the equation in (14) reduces to the form of the chain rule

ddty(t)=hf(x(t)),\frac{d}{dt}y(t)=\nabla h\cdot f(x(t)),

with “\cdot” denoting the Euclidean inner product on n\mathbb{R}^{n}, which coincides with the system governing the dynamics of the output of the system in (1) derived by using the techniques in the classical nonlinear systems theory [24]. In another word, the system in (13), equivalently, the action of the Koopman system in (10) on the space of observables C(M)C^{\infty}(M), basically gives rise to a dynamical system theoretic coordinate-free representation of the chain rule for differentiation.

IV Geometric Control of Koopman Systems on Lie Groups

After the detailed investigation on Koopman systems in Section III, we will move one step further in this section to analyze the dynamic behavior of Koopman systems driven by control inputs. In particular, the focus will be placed on the relationship between control-affine systems and their associated Koopman control systems in terms of control-related properties, and this will lead to an operator-theoretic controllability condition for control-affine systems. However, the techniques to be developed are not constrained to examine controllability. To demonstrate the generalizability, we will establish a systematic Koopman feedback linearization framework parallel to the classical feedback linearization method by leveraging the linearity of Koopman systems.

IV-A Koopman operators associated with control systems

A control system evolving on a smooth manifold MM has the form

ddtx(t)=F(t,x(t),u(t)),\displaystyle\frac{d}{dt}x(t)=F(t,x(t),u(t)), (15)

where uu is the control input, generally assumed to take values in m\mathbb{R}^{m}. In particular, for any given the control input uu, FF is a time-dependent smooth vector field on MM. As discussed in Section II, the Koopman operator associated with a dynamical system needs to be defined through the flow of the system. However, for a system whose dynamics is governed by a time-dependent vector as the system in (15), it might not generate a well-defined flow, even locally, as defined in Section II-A, because different integral curves starting at the same point but at different times might follow different paths. To avoid this technical difficulty, we always set the starting time of the system to be 0 and treat uu as intermediate variables. Moreover, for any choice of the control input uu, we also assume the vector field FF to be complete so that the system generates a global flow.

Under the aforementioned technical setups, we can define the Koopman operator associated with the system in (17) in a similar way as before. To this end, let Φ:×MM\Phi:\mathbb{R}\times M\rightarrow M denote the flow of the system driven by some control inputs uiu_{i}, i=1,,mi=1,\dots,m, then Φ(t,x)M\Phi(t,x)\in M is the state of the system at time tt starting from xMx\in M at time 0. The Koopman operator associated with the system is defined similarly as Ut:C(M)C(M)U_{t}:C^{\infty}(M)\rightarrow C^{\infty}(M) given by hhΦth\mapsto h\circ\Phi_{t} for any observable hC(M)h\in C^{\infty}(M) of the system. Note that, in this case, although not explicitly expressed in the formulas, the Koopman operator UtU_{t}, as well as the flow Φt\Phi_{t}, contains the control inputs uiu_{i}, i=1,,mi=1,\dots,m applied to the system as the intermediate variables. Moreover, for any choice of the control inputs, all of the properties of Koopman operators discussed in Section II still hold, which are summarized in the following proposition.

Proposition 2.

Given a control system evolving on a smooth manifold MM as in (15), the associated Koopman operator Ut:C(M)C(M)U_{t}:C^{\infty}(M)\rightarrow C^{\infty}(M) satisfies the following:

  1. 1.

    The Koopman operator UtU_{t} is continuous in the CC^{\infty}-topology on C(M)C^{\infty}(M).

  2. 2.

    The Koopman group {Ut}t\{U_{t}\}_{t\in\mathbb{R}} is a strongly continuous one parameter subgroup of Aut(C(M)){\rm Aut}(C^{\infty}(M)).

  3. 3.

    The Koopman operator UtU_{t} satisfies the differential equation system on Aut(C(M)){\rm Aut}(C^{\infty}(M)), given by,

    ddtUt=FUt=UtF\displaystyle\frac{d}{dt}U_{t}=\mathcal{L}_{F}U_{t}=U_{t}\mathcal{L}_{F} (16)

    with the initial condition U0=IU_{0}=I, the identity of the group Aut(C(M)){\rm Aut}(C^{\infty}(M)).

Proof.

Driven by any control inputs uiu_{i}, i=1,,mi=1,\dots,m, Φ(,x):M\Phi(\cdot,x):\mathbb{R}\rightarrow M gives the solution of the system in (17) passing through the point xMx\in M at time 0, and hence satisfies the groups laws as in (2) following from the uniqueness of solutions for ordinary differential equations. As a result, {Ut}t\{U_{t}\}_{t\in\mathbb{R}} is a one parameter group of linear operators from C(M)C^{\infty}(M) to itself. The continuity of UtU_{t}, strong continuity of the group {Ut}t\{U_{t}\}_{t\in\mathbb{R}}, and the derivation of the system in (16) follow from the same proofs as Proposition 1, Lemma 1, and Theorem 2, respectively. ∎

Although the form of the Koopman system in (16) associated with a control system appears to be identical to the one in (10) associated with a dynamical system without control input, these two Koopman systems are distinct from several aspects. For example, in the control system case as in (16), we particularly specify the initial condition U0=IU_{0}=I to emphasize that the starting time can only be chosen to be 0, which is necessary to guarantee the Koopman system in (16) to be well-defined as discussed in the beginning of Section IV-A. More importantly, the Koopman system in (16) actually inherits the control inputs from the vector field FF in the Lie derivative F\mathcal{L}_{F}, and hence is a control system defined on Aut(C(M)){\rm Aut}(C^{\infty}(M)) sharing the same control inputs with the associated control system in (15) defined on MM. Consequently, it is natural to ask for controllability of the Koopman system in (16), and particularly its relationship with the associated systems in (15) in terms of control-related properties, which will be fully explored by using control-affine systems in the next section.

IV-B Gelfand duality between bilinear Koopman and control-affine systems

Nonlinear systems in the control-afffine form are natural models for describing the dynamics of control systems arising from numerous practical and scientific applications in diverse fields, ranging from quantum physics and robotics to neuroscience. In the most general form, a (smooth) control-affine system defined on a smooth manifold MM can be represented as

ddtx(t)=f(x(t))+i=1mui(t)g(x(t)),\displaystyle\frac{d}{dt}x(t)=f(x(t))+\sum_{i=1}^{m}u_{i}(t)g(x(t)), (17)

where ff and gig_{i}, i=1,,mi=1,\dots,m are smooth vector fields on MM, and ui:u_{i}:\mathbb{R}\rightarrow\mathbb{R} are the control inputs. The control-linear form of the system in (17) leads to a particularly neat representation of the associated Koopman system.

Theorem 3 (Koopman bilinearization).

Given a control-affine system defined on a smooth manifold MM as in (17), the associated Koopman system is a bilinear system defined on Aut(C(M)){\rm Aut}(C^{\infty}(M)), given by,

ddtUt\displaystyle\frac{d}{dt}U_{t} =Utf+i=1mui(t)Utgi\displaystyle=U_{t}\mathcal{L}_{f}+\sum_{i=1}^{m}u_{i}(t)U_{t}\mathcal{L}_{g_{i}}
=fUt+i=1mui(t)giUt.\displaystyle=\mathcal{L}_{f}U_{t}+\sum_{i=1}^{m}u_{i}(t)\mathcal{L}_{g_{i}}U_{t}. (18)
Proof.

The result directly follows from the property 3) in Proposition 2 by choosing F=f+i=1muigiF=f+\sum_{i=1}^{m}u_{i}{g_{i}} and the linearity of the Lie derivative and Koopman operator. ∎

Similar to the systems in (10) and (16), the Koopman system in (18) holds in the sense of the strong operator topology, i.e.,

ddtUth\displaystyle\frac{d}{dt}U_{t}h =Utfh+i=1mui(t)Utgih\displaystyle=U_{t}\mathcal{L}_{f}h+\sum_{i=1}^{m}u_{i}(t)U_{t}\mathcal{L}_{g_{i}}h
=fUth+i=1mui(t)giUth,\displaystyle=\mathcal{L}_{f}U_{t}h+\sum_{i=1}^{m}u_{i}(t)\mathcal{L}_{g_{i}}U_{t}h,

for any hC(M)h\in C^{\infty}(M), because the infinitesimal generator ddtUt\frac{d}{dt}U_{t} is derived in this topology as shown in Section III-A.

Remark 2.

The Koopman system in (18) requires to be interpreted with great caution: the commutativity between the Koopman operator UtU_{t} and the Lie derivative operator f+i=1muig\mathcal{L}_{f+\sum_{i=1}^{m}u_{i}g} by no means implies the commutativity between UtU_{t} and f\mathcal{L}_{f} as well as UtU_{t} and gi\mathcal{L}_{g_{i}} for each i=1,,mi=1,\dots,m. Taking f\mathcal{L}_{f} for instance, as indicated by the proof of Theorem 2, because f+i=1muig\mathcal{L}_{f+\sum_{i=1}^{m}u_{i}g} is the infinitesimal generator of UtU_{t}, equivalently the vector field f+i=1muigf+\sum_{i=1}^{m}u_{i}g generates the flow of the system, the commutativity between UtU_{t} and f\mathcal{L}_{f} requires f(f+i=1muigi)=i=1muifgi=0\mathcal{L}_{f}(f+\sum_{i=1}^{m}u_{i}g_{i})=\sum_{i=1}^{m}u_{i}\mathcal{L}_{f}g_{i}=0, for which a sufficient condition is fgi=[f,gi]=0\mathcal{L}_{f}g_{i}=[f,g_{i}]=0 for all i=1,,mi=1,\dots,m. This in turn gives rise to a Koopman operator-theoretic characterization of the general fact in differential geometry that two smooth vector fields on a smooth manifold is commutative, that is, their Lie bracket vanishes, if and only if one of them is invariant under the flow of the other one [19].

Remark 2 further sheds light on the intimacy between Koopman operators and Lie brackets of vector fields. It is well-known in geometric control theory that, for a control affine system as in (17), the iterative Lie brackets of, or more specifically, the Lie algebra generated by, the drift and control vector fields determine the controllability of the system [25, 26, 27, 28, 29]. This strongly inspires the study of controllability of control-affine systems through the associated Koopman systems and vice versa. To this end, it is required to investigate Koopman systems further from the perspective of geometric control theory.

IV-B1 Koopman systems on Lie groups

A remarkable feature of Koopman systems associated with control-affine systems is the bilinearity as proved in Theorem 3. Right or left invariant control systems defined on Lie groups constitute a large class of bilinear systems, and have been widely studied in geometric control theory [30, 31, 32]. Fortunately, the Koopman system associated with a control-affine system as in (18) falls into this category of bilinear systems, but it is generally an infinite-dimensional system, which presents a challenge to our study. In particular, the challenge mainly comes from the infinite-dimension of the state-space Lie group Aut(C(M)){\rm Aut}(C^{\infty}(M)) of the Koopman system in (18).

To avoid some technical topological complications for the purpose of illuminating the main idea, we impose the compactness condition that that the state-space manifold MM of the control-affine system in (17) is closed, that is, compact, connected, and without boundary. As a result, any smooth vector field on MM are complete and hence generates a global flow [19]. Moreover, the family of seminorms in (5) generating the CC^{\infty}-topology on C(M)C^{\infty}(M) are reduced to the form

hk=supxM|Dkh(x)|,\|h\|_{k}=\sup_{x\in M}|D^{k}h(x)|,

which then coincides with the Whitney CC^{\infty}-topology so that pointwise addition and multiplication of smooth functions are continuous [33]. This further leads to the continuity of the map Diff(M)\mathbb{R}\rightarrow{\rm Diff}(M) given by tΦtt\mapsto\Phi_{t} for any flow Φ\Phi on MM with MM denoting the group of diffeomorphisms from MM onto MM [33]. The use of these results is to identify the Lie algebra of Aut(C(M)){\rm Aut}(C^{\infty}(M)), for analyzing controllability of the Koopman system in (18) defined on Aut(C(M)){\rm Aut}(C^{\infty}(M)).

Lemma 2.

Let MM be a closed manifold, then Aut(C(M)){\rm Aut}(C^{\infty}(M)) is a Lie group with the Lie algebra consisting of Lie derivatives with respect to smooth vector fields on MM under the Whitney CC^{\infty}-topology on C(M)C^{\infty}(M).

Proof.

The continuity of pointwise addition and multiplication implies that C(M)C^{\infty}(M) is a commutative seminormed CC^{*}-algebra. Then, the application of Gelfand–Naimark theorem implies that Aut(C(M)){\rm Aut}(C^{\infty}(M)) is isomorphic to Diff(M){\rm Diff}(M) [34]. To identify the Lie algebra of Diff(M){\rm Diff}(M), it is equivalent to identify the tangent space at the identity map idM{\rm id}_{M}. To this end, we choose a Riemannian metric on MM. Because MM is compact, the Riemannian exponential map exp:TMM\exp:TM\rightarrow M is globally well-defined and a local homeomorphism following from dexpp(0)=idTpMd\exp_{p}(0)={\rm id}_{T_{p}M}, giving the normal coordinates of MM [35]. Therefore, by varying pp in MM, we obtain that C(M,TM)C^{\infty}(M,TM), the space of smooth sections of the tangent bundle TMTM of MM, is tangent to C(M,M)C^{\infty}(M,M), the space of smooth maps from MM to MM, at idM{\rm id}_{M}. Because Diff(M){\rm Diff}(M) is an open submanifold of C(M,M)C^{\infty}(M,M) in the Whitney CC^{\infty}-topology, C(M,TM)C^{\infty}(M,TM) is also the tangent space of Diff(M){\rm Diff}(M) at idM{\rm id}_{M}. Note that C(M,TM)C^{\infty}(M,TM) is just the space 𝔛(M)\mathfrak{X}(M) of smooth vector fields on MM. In addition, 𝔛(M)\mathfrak{X}(M) is a Lie algebra under the Lie bracket operation, and hence the Lie algebra of Diff(M){\rm Diff}(M). Together with the naturality of Lie derivatives, that is, [f,g]=[f,g]\mathcal{L}_{[f,g]}=[\mathcal{L}_{f},\mathcal{L}_{g}] for any f,g𝔛(M)f,g\in\mathfrak{X}(M), the space of Lie derivatives is a Lie algebra isomorphic to 𝔛(M)\mathfrak{X}(M). Since Lie derivatives are derivations on C(M)C^{\infty}(M), we conclude that the Lie algebra of Aut(C(M)){\rm Aut}(C^{\infty}(M)) is exactly the space of Lie derivatives. ∎

Computationally, given a coordinate chart (x1,,xn)(x_{1},\dots,x_{n}) of MM defined on an open subset UMU\subseteq M, any vector field f𝔛(M)f\in\mathfrak{X}(M) admits the local coordinate representation f=i=1nfixif=\sum_{i=1}^{n}f_{i}\frac{\partial}{\partial x_{i}} for some fiC(U)f_{i}\in C^{\infty}(U). Correspondingly, for any observable hC(M)h\in C^{\infty}(M) of the system in (17), we have

fh=dh(f)=i=1nfihxi\displaystyle\mathcal{L}_{f}h=dh(f)=\sum_{i=1}^{n}f_{i}\frac{\partial h}{\partial x_{i}}

and hence the local coordinate representation of the Lie derivative operator f\mathcal{L}_{f} is identical to that of f𝔛(M)f\in\mathfrak{X}(M) as

f=i=1nfixi.\displaystyle\mathcal{L}_{f}=\sum_{i=1}^{n}f_{i}\frac{\partial}{\partial x_{i}}. (19)

This in turn gives a coordinate-dependent argument for the isomorphism between the Lie algebras of smooth vectors fields and Lie derivatives on MM. From a different perspective, we notice that any partial differential operator of homogenous order 1 on MM has the local representation in the form of (19). Consequently, the Lie algebra of Aut(C(M)){\rm Aut}(C^{\infty}(M)) can also be identified with the space of homogeneous first order partial differential operators on MM.

Remark 3 (System-theoretic Gelfand duality).

As indicated in the proof of Lemma 2, Diff(M){\rm Diff}(M) is related to Aut(C(M)){\rm Aut}(C^{\infty}(M)) by the Gelfand representation. On the dynamical systems level, the time-tt flow of the control-affine system in (17) is always an element of Diff(M){\rm Diff}(M) for any choice of the control inputs uiu_{i}, i=1,,mi=1,\dots,m as defined in Section II-A. Therefore, the associated Koopman system defined on Aut(C(M)){\rm Aut}(C^{\infty}(M)) in (18) associated with it can be interpreted as the ”Gelfrand representation” of the system in (17) on the space C(M)C^{\infty}(M) of observables, which further reveals the Gelfand duality between control-affine systems and bilinear Koopman systems in the algebraic sense.

IV-B2 Controllability of Koopman systems

Because controllability of a control-affine system is determined by the Lie algebra generated by its drift and control vector fields, the aforementioned identification of these vector fields with the corresponding Lie derivative operators, governing the dynamics of the associated Koopman system, opens up the possibility to study controllability of control-affine systems by utilizing their associated bilinear Koopman systems. In particular, the interpretation of Lie derivatives as partial differential operators gives rise to the follows operator-theoretic characterization of controllability for control-affine systems.

Theorem 4.

A control-affine system defined on a closed manifold MM as in (17) is controllable if and only if the associated Koopman system as in (18) acts on the space of observables C(M)C^{\infty}(M) by the de Rham differential operator d:C(M)C(M,TM)d:C^{\infty}(M)\rightarrow C^{\infty}(M,T^{*}M), where C(M,TM)C^{\infty}(M,T^{*}M) denotes the space of smooth sections of the cotangent bundle TMT^{*}M of MM, i.e., the space of differential 1-forms on MM.

Proof.

According to the Lie algebra rank condition (LARC), the system in (17) is controllable on MM if and only if the Lie algebra generated by the vector fields ff and gig_{i}, i=1,,mi=1,\dots,m, evaluated at each point xMx\in M, is the tangent space TxMT_{x}M of MM at the point xx [27]. The isomorphism between the Lie algebras generated by smooth vector fields and Lie derivative operators revealed in the proof of Lemma 2 then implies that, for any vector field v𝔛(M)v\in\mathfrak{X}(M), the tangent vector v(x)TxMv(x)\in T_{x}M belongs to the Lie algebra generated by f\mathcal{L}_{f} and gi\mathcal{L}_{g_{i}}, i=1,,mi=1,\dots,m, evaluated at xx. Recall the action of Lie derivatives on smooth functions as vh(x)=dh(v)|x=dhx(v(x))\mathcal{L}_{v}h(x)=dh(v)|_{x}=dh_{x}(v(x)), the directional derivative of hh at xx along the direction of v(x)v(x), the above observation implies that there exist control inputs ui(t)u_{i}(t), i=1,,mi=1,\dots,m such that ddtUth(x)=vUth(x)=d(Uth)x(v(x))\frac{d}{dt}U_{t}h(x)=\mathcal{L}_{v}U_{t}h(x)=d(U_{t}h)_{x}(v(x)). Because v(x)TxMv(x)\in T_{x}M is arbitrary, we obtain ddtUth=d(Uth)\frac{d}{dt}U_{t}h=d(U_{t}h). In addition, note that Ut:C(M)C(M)U_{t}:C^{\infty}(M)\rightarrow C^{\infty}(M) is invertible, which particularly implies the subjectivity of UtU_{t} so that UthC(M)U_{t}h\in C^{\infty}(M) can be arbitrarily chosen as well. Therefore, we arrive at the conclusion ddtUt=d\frac{d}{dt}U_{t}=d. ∎

The de Rham differential is globally defined for any smooth manifold. Therefore, Theorem 4 also gives rise to a global characterization of controllability for a control-affine system in terms of differential operators, contrary to the Lie algebra rank condition (LARC) which involves the local examination of the Lie algebra generated by the drift and control vector fields of the system at every point in the state-space manifold of the system.

To generalize Theorem 4 to uncontrollable systems, we notice that a system is always controllable on its controllable submanifold. In particular, for a control-affine system as in (17), the Frobenius theorem further implies that the controllable submanifold is a maximal integral manifold of the involutive distribution, that is, the Lie subalgebra of the vector fields, generated by the drift and control vector fields of the system, and the collection of all controllable submanifolds of the system starting from different initial conditions forms a foliation of the state-space manifold. Then, the proof of Theorem 4 can be directly adopted by restricting to the controllable submanifold of the system in (17).

Corollary 2.

Consider the control-affine system in (17) defined on the manifold MM. Let NMN\subseteq M be the controllable submanifold of the system and ι:NM\iota:N\hookrightarrow M denote the inclusion. Then, the associated Koopman system in (18) acts on the space of observables C(M)C^{\infty}(M) by ιd\iota^{*}d, where dd denotes the de Rham differential operator on MM and ι\iota^{*} is the pullback of ι\iota.

Proof.

The condition that NN is the controllable submanifold of the system in (17) implies that the Lie subalgebra of 𝔛(M)\mathfrak{X}(M) generated by the vector fields ff and gig_{i}, governing the dynamics of the system, coincides with TxNT_{x}N evaluated at each xNx\in N. As a result, for any v𝔛(N)v\in\mathfrak{X}(N), there are control inputs uiu_{i}, i=1,,mi=1,\dots,m such that ddtUth(x)=v(x)Uth(x)=d(Uth)x(v(x))=d(Uth)ι(x)(ιv(x))=ιd(Uth)x(v(x))\frac{d}{dt}U_{t}h(x)=\mathcal{L}_{v(x)}U_{t}h(x)=d(U_{t}h)_{x}(v(x))=d(U_{t}h)_{\iota(x)}(\iota_{*}v(x))=\iota^{*}d(U_{t}h)_{x}(v(x)) by the definition of the pullback [19], where ι\iota_{*} denotes the pushforward of ι\iota. Because hC(M)h\in C^{\infty}(M) and v𝔛(N)v\in\mathfrak{X}(N) are arbitrary, we conclude ddtUt=ιd\frac{d}{dt}U_{t}=\iota^{*}d as desired. ∎

On the other hand, the proof of Theorem 4 further sheds light on the controllability analysis of Koopman systems through their actions on the observables of the corresponding control-affine systems. To put forward this idea, we now turn our attention to controllability of the Koopman system on Aut(C(M)){\rm Aut}(C^{\infty}(M)). However, it is so unfortunate that Koopman systems are almost never controllable on Aut(C(M)){\rm Aut}(C^{\infty}(M)), even though the corresponding control-affine systems controllable on MM as shown in the following example.

Example 1 (Uncontrollability of Koopman systems).

consider the control-affine system

ddtx(t)=i=1nui(t)ei(x(t))\displaystyle\frac{d}{dt}x(t)=\sum_{i=1}^{n}u_{i}(t)e_{i}(x(t))

defined on an nn-dimensional smooth manifold MM, where ei𝔛(M)e_{i}\in\mathfrak{X}(M), i=1,,ni=1,\dots,n form a global frame for MM restricting to the coordinate vector fields ei=xie_{i}=\frac{\partial}{\partial x_{i}} in a coordinate chart. Then, the system is clearly controllable on MM by the LARC, since as a frame for MM, they satisfy span{e1(p),,en(p)}=TpM{\rm span}\{e_{1}(p),\dots,e_{n}(p)\}=T_{p}M for all pMp\in M. However, because of ei=xie_{i}=\frac{\partial}{\partial x_{i}}, [ei,ej]=0[e_{i},e_{j}]=0, equivalently [ei,ej]=[ei,ej]=0[\mathcal{L}_{e_{i}},\mathcal{L}_{e_{j}}]=\mathcal{L}_{[e_{i},e_{j}]}=0, holds for all i,j=1,,ni,j=1,\dots,n. Correspondingly, the Lie algebra generated by ei\mathcal{L}_{e_{i}}, i=1,,ni=1,\dots,n is just span{e1,,en}{\rm span}\{\mathcal{L}_{e_{1}},\dots,\mathcal{L}_{e_{n}}\}, which is a proper Lie subalgebra of 𝔛(M)\mathfrak{X}(M), only consisting of “constant vector fields”, i.e, vector fields in the form of f=i=1nfixif=\sum_{i=1}^{n}f_{i}\frac{\partial}{\partial x_{i}} with all of fif_{i} constant functions on MM. In the terminology of algebra, it means that the Lie algebra generated by ei\mathcal{L}_{e_{i}}, i=1,,ni=1,\dots,n fails to be a module over C(M)C^{\infty}(M) so that Koopman operators generated by flows of “nonconstant vector fields” on MM are not in the reachable set of the associated Koopman system

ddtUt=i=1nui(t)eiUt,\displaystyle\frac{d}{dt}U_{t}=\sum_{i=1}^{n}u_{i}(t)\mathcal{L}_{e_{i}}U_{t}, (20)

which in turn implies uncontrollability of the Koopman system on Aut(C(M)){\rm Aut}(C^{\infty}(M)).

Elaborating uncontrollability of the Koopman system in (20) from a purely control-theoretic perspective, the trouble that the Lie algebra generated by ei\mathcal{L}_{e_{i}} fails to be a C(M)C^{\infty}(M)-module is caused by the independence of the control inputs uiu_{i} from the state-space MM of the corresponding control-affine system, meaning, uiu_{i} are function defined on \mathbb{R} solely instead of ×M\mathbb{R}\times M as in the case of control of partial differential equation systems. However, this is definitely not the only obstacle to controllability of Koopman systems. Another one follows from the definition of Koopman operators as composition operators with flows of vector fields on MM, which restricts the reachable set of the Koopman system to elements in Aut(C(M)){\rm Aut}(C^{\infty}(M)), equivalently Diff(M){\rm Diff}(M), generated by flows of vector fields. In general, not every diffeomorphism on MM, even in arbitrary small neighborhood of the identity map idM{\rm id}_{M} of MM in Diff(M){\rm Diff}(M) in the CC^{\infty}-topology, comes from a flow of some vector field. The most notable example was provided by the American mathematician John Milnor in 1984 [36]: The function Φ(θ)=θ+πn+εsin2(nθ)\Phi(\theta)=\theta+\frac{\pi}{n}+\varepsilon\sin^{2}(n\theta) is a diffeomorphism of the unit circle 𝕊1\mathbb{S}^{1} for any nn\in\mathbb{N} and ε>0\varepsilon>0, and particularly, by choosing large enough nn and small enough ε\varepsilon, ff can be arbitrarily close to idM{\rm id}_{M}. The points in 𝕊1\mathbb{S}^{1} of the form θk=kπ/n\theta_{k}=k\pi/n with kk\in\mathbb{Z} are 2n2n-periodic, i.e., f2n(θk)=θkf^{2n}(\theta_{k})=\theta_{k}, where f2nf^{2n} denotes the 2n2n-fold composition of ff, but other points are all aperiodic. This fact prevents ff from being a flow of a vector field on 𝕊1\mathbb{S}^{1}, and further indicates that, in general, the exponential map from the Lie algebra 𝔛(M)\mathfrak{X}(M) to the Lie group Diff(M){\rm Diff}(M) is not a local diffeomorphism, actually neither locally injective nor locally surjective, which demonstrates a significant difference between infinite-dimensional and finite-dimensional Lie groups. Moreover, this also disables the application of LARC to examine controllability of Koopman systems.

Fortunately, by leveraging the algebraic structure of the Lie group Diff(M){\rm Diff}(M), it is still possible to characterize the controllable submanifold of the Koopman system defined on Aut(C(M)){\rm Aut}(C^{\infty}(M))

Theorem 5 (Controllable submanifolds of Koopman systems).

Given a Koopman system defined on the Lie group Aut(C(M)){\rm Aut}(C^{\infty}(M)) as in (18), that is,

ddtUt=fUt+i=1mui(t)giUt\frac{d}{dt}U_{t}=\mathcal{L}_{f}U_{t}+\sum_{i=1}^{m}u_{i}(t)\mathcal{L}_{g_{i}}U_{t}

associated to a control-affine system defined on MM as in (17), the controllable submanifold of the Koopman system is the identity component of the Lie subgroup of Aut(C(M)){\rm Aut}(C^{\infty}(M)) whose Lie algebra is the Lie subalgebra of 𝔛(M)\mathfrak{X}(M) generated by f\mathcal{L}_{f} and gi\mathcal{L}_{g_{i}}, i=1,,ni=1,\dots,n.

Proof.

Let 𝔤\mathfrak{g} denotes the Lie subalgebra of 𝔛(M)\mathfrak{X}(M) generated by f\mathcal{L}_{f} and gi\mathcal{L}_{g_{i}}, i=1,,ni=1,\dots,n, and GG be the Lie subgroup of Aut(C(M)){\rm Aut}(C^{\infty}(M)) with the Lie algebra 𝔤\mathfrak{g}. Then, by applying piecewise constant control inputs, it can be shown that elements in Aut(C(M)){\rm Aut}(C^{\infty}(M)) of the form exp(v1)exp(vk)\exp(\mathcal{L}_{v_{1}})\cdots\exp(\mathcal{L}_{v_{k}}) with vi𝔤\mathcal{L}_{v_{i}}\in\mathfrak{g} for all i=1,,ki=1,\dots,k are in the reachable set of the Koopman system in (18). Note that all such elements form a normal subgroup of G0G_{0}, the identity component of GG, but G0G_{0} is a simple group [37], i.e., does not contain any normal subgroups in addition to the trivial group and itself. Therefore, these elements constitute the whole G0G_{0}. Moreover, because every state in the reachable set of a control-affine system can be reached by applying piecewise constant control inputs [28], we conclude that G0G_{0} is the controllable submanifold of the Koopman system in (18). ∎

It is worth mentioning that the proof of Theorem 5 completely relies on the algebraic structure of the Lie algebra generated by the Lie derivative operators f\mathcal{L}_{f} and gi\mathcal{L}_{g_{i}}, equivalently, the vector fields f,gi𝔛(M)f,g_{i}\in\mathfrak{X}(M), and more specifically, it is independent of whether the system in (18) is a Koopman system or not. Therefore, Theorem 5 gives rise to a universal characterization of controllable submanifolds for bilinear systems defined on the Lie group Aut(C(M)){\rm Aut}(C^{\infty}(M)), which can be viewed as the infinite-dimensional analogue of the LARC for bilinear systems evolving on Lie groups.

To present the result in the most general setting, we temporally relax the compactness condition on the manifold MM. In this case, the Lie algebra of the Lie group Diff(M){\rm Diff}(M) becomes 𝔛c(M)\mathfrak{X}_{c}(M), the space of compactly support smooth vector fields on MM. Technically, to guarantee that the map Diff(M)\mathbb{R}\rightarrow{\rm Diff}(M) given by tΦtt\mapsto\Phi_{t} is smooth for any flow Φ\Phi of a vector field in 𝔛c(M)\mathfrak{X}_{c}(M), we equip Diff(M){\rm Diff}(M) with the final topology generated by smooth curves from \mathbb{R} to C(M,M)C^{\infty}(M,M), the space of all smooth maps form MM to MM. This topology is generally weaker than the Whitney CC^{\infty}-topology, but coincides with it when MM is compact [38]. Algebraically, the isomorphism between Aut(C(M)){\rm Aut}(C^{\infty}(M)) and Diff(M){\rm Diff}(M), as well as their Lie algebras, still holds. Moreover, if MM is a manifold with boundary, then we also require vector fields in 𝔛c(M)\mathfrak{X}_{c}(M) to be vanishing on the boundary.

Corollary 3 (Infinite-dimensional LARC).

Let MM be a smooth manifold, possibly noncompact and with boundary, and

ddtA(t)=fA(t)+i=1mui(t)giA(t)\displaystyle\frac{d}{dt}A(t)=\mathcal{L}_{f}A(t)+\sum_{i=1}^{m}u_{i}(t)\mathcal{L}_{g_{i}}A(t)

be a bilinear system defined on the Lie group Aut(C(M)){\rm Aut}(C^{\infty}(M)), where f,gi𝔛c(M)f,g_{i}\in\mathfrak{X}_{c}(M) for all i=1,,mi=1,\dots,m. Then, the controllable submanifold of the system is the identity component of the Lie subgroup of Aut(C(M)){\rm Aut}(C^{\infty}(M)) whose Lie algebra is the Lie subalgebra of 𝔛c(M)\mathfrak{X}_{c}(M) generated by f\mathcal{L}_{f} and gi\mathcal{L}_{g_{i}}, i=1,,mi=1,\dots,m.

Proof.

The proof is identical to that of Theorem 5. ∎

V Koopman Feedback Linearization

We have launched a detailed investigation into Koopman systems and their relationship with the dynamical systems that they are associated with. In particular, for control-affine systems, the bilinear form of the associated Koopman systems greatly benefits the study of the original systems in a coordinate-free way, e.g., the characterize of controllability in terms of the de Rham differential as shown in Theorem 4. In general, it is no doubt that linear systems are preferable from various aspects. Therefore, the focus of this section will be on the search of appropriate coordinates by using Koopman systems under which the corresponding control-affine systems can be represented as linear systems. In addition, we will also indicate the relation and distinction between the proposed linearization and the well-known feedback linearization. As the prerequisite, we will first explore how to change the coordinates of control-affine systems by using the associated Koopman systems.

V-A Koopman change of coordinates

Owing to the action of Koopman operators on the space of observables, for the purpose of changing the coordinate representation of a system, it is natural that the new coordinates must compose of the observables of the system. In addition, in the case of the state-space manifold MM of the system having dimension n>1n>1, every coordinate chart of MM must contain nn components to guarantee MM to be locally homeomorphic to n\mathbb{R}^{n}. Therefore, the use of Koopman operators to change the coordinates of the system requires the extension of the action of Koopman operators from C(M)C^{\infty}(M) to C(M,n)C^{\infty}(M,\mathbb{R}^{n}), the space of n\mathbb{R}^{n}-valued smooth functions defined on MM, where n\mathbb{R}^{n} is treated as a smooth manifold under its usual topology and smooth functions from MM to n\mathbb{R}^{n} are understood in the usual sense as smooth functions between two smooth manifolds. We also refer to such functions in C(M,n)C^{\infty}(M,\mathbb{R}^{n}) as observables and their values in n\mathbb{R}^{n} as outputs.

Note that n\mathbb{R}^{n}, as a manifold, is parallelizable since the set of coordinate vector fields {x1,,x1}\{\frac{\partial}{\partial x_{1}},\dots,\frac{\partial}{\partial x_{1}}\} restricts to a basis of TpnT_{p}\mathbb{R}^{n} at any pnp\in\mathbb{R}^{n} [19]. As a direct consequence of this result, every elements hC(M,n)h\in C^{\infty}(M,\mathbb{R}^{n}) admits a representation as nn-tuple h=(h1,,hn)h=(h_{1},\dots,h_{n})^{\prime} with each component hiC(M)h_{i}\in C^{\infty}(M) and ‘\prime’ denoting the transpose. Equivalently, we reveal an isomorphism between C(M,n)C^{\infty}(M,\mathbb{R}^{n}) and (C(M))n\big{(}C^{\infty}(M)\big{)}^{n} as C(M)C^{\infty}(M)-algebras, where (C(M))n\big{(}C^{\infty}(M)\big{)}^{n} denotes the direct sum of nn copies of C(M)C^{\infty}(M), which in turn indicates the component-wise action of the Koopman operator UtU_{t} on C(M,n)C^{\infty}(M,\mathbb{R}^{n}) as

Uth=Ut[h1hr]=[Uth1Uthn]C(M,n)U_{t}h=U_{t}\left[\begin{array}[]{c}h_{1}\\ \vdots\\ h_{r}\end{array}\right]=\left[\begin{array}[]{c}U_{t}h_{1}\\ \vdots\\ U_{t}h_{n}\end{array}\right]\in C^{\infty}(M,\mathbb{R}^{n})

so that UtAut(C(M,n))U_{t}\in{\rm Aut}(C^{\infty}(M,\mathbb{R}^{n})). In particular, in the case that the system is in the control-affine form and controllable on MM, the component-wise action can then be extended from the Koopman operator UtU_{t} to its infinitesimal generator ddtUt\frac{d}{dt}U_{t} as

ddtUth=[ddtUth1ddtUthr]=[d(Uth1)d(Uthn)],\frac{d}{dt}U_{t}h=\left[\begin{array}[]{c}\frac{d}{dt}U_{t}h_{1}\\ \vdots\\ \frac{d}{dt}U_{t}h_{r}\end{array}\right]=\left[\begin{array}[]{c}d(U_{t}h_{1})\\ \vdots\\ d(U_{t}h_{n})\end{array}\right],

To elucidate the mechanism of ddtUth\frac{d}{dt}U_{t}h, we calculate its action on vector fields on MM by using local coordinates. Specifically, let (x1,,xn)(x_{1},\dots,x_{n}) and (y1,,yn)(y_{1},\dots,y_{n}) are coordinate charts on MM and n\mathbb{R}^{n}, respectively, then ddtUth\frac{d}{dt}U_{t}h has the local representation

ddtUth=i=1nd(Uthi)yi=i=1nj=1nUthixjdxjyi,\displaystyle\frac{d}{dt}U_{t}h=\sum_{i=1}^{n}d(U_{t}h_{i})\otimes\frac{\partial}{\partial y_{i}}=\sum_{i=1}^{n}\sum_{j=1}^{n}\frac{\partial U_{t}h_{i}}{\partial x_{j}}dx_{j}\otimes\frac{\partial}{\partial y_{i}},

which exactly gives the Jacobian matrix, equivalently the coordinate representation of the differential, of the function UthU_{t}h, where \otimes denotes the tensor product of vector bundles, to be more specific, the tensor product between the cotangent bundle TMT^{*}M of MM and the tangent bundle TnT\mathbb{R}^{n} of n\mathbb{R}^{n}. Let f=i=1nfixif=\sum_{i=1}^{n}f_{i}\frac{\partial}{\partial x_{i}} be a smooth vector field on MM, then the action of ddtUt\frac{d}{dt}U_{t} on ff is given by

ddtUth(f)\displaystyle\frac{d}{dt}U_{t}h(f) =i=1n(j=1nUthixjdxj)(fkxk)yi\displaystyle=\sum_{i=1}^{n}\Big{(}\sum_{j=1}^{n}\frac{\partial U_{t}h_{i}}{\partial x_{j}}dx_{j}\Big{)}\Big{(}f_{k}\frac{\partial}{\partial x_{k}}\Big{)}\otimes\frac{\partial}{\partial y_{i}}
=i=1n(j=1nfjUthixj)yi,\displaystyle=\sum_{i=1}^{n}\Big{(}\sum_{j=1}^{n}f_{j}\frac{\partial U_{t}h_{i}}{\partial x_{j}}\Big{)}\frac{\partial}{\partial y_{i}}, (21)

where we use the duality between TMT^{*}M and TMTM as dxj(xk)=δjkdx_{j}(\frac{\partial}{\partial x_{k}})=\delta_{jk} with δjk=1\delta_{jk}=1 if j=kj=k and δjk=0\delta_{jk}=0 otherwise for all j,k=1,,nj,k=1,\dots,n. Note that the equation in (21) coincides with dUth(f)dU_{t}h(f), the differential of UthU_{t}h acting on ff [19], which then gives rise to the following conclusion.

Proposition 3.

Given a controllable control-affine system evolving on a smooth manifold MM. Then, for any observable hC(M,n)h\in C^{\infty}(M,\mathbb{R}^{n}) of the system, the associated Koopman system acting on hh gives the pushforward of smooth vector fields on MM by the flow of the system output.

Proof.

This is largely an exercise in decoding terminologies. We first assume that the system is controllable on MM. As shown in (21), for any f𝔛(M)f\in\mathfrak{X}(M), ddtUth(f)\frac{d}{dt}U_{t}h(f) defines a section d(Uth)(f)d(U_{t}h)(f) of the pullback bundle (Uth)Tn(U_{t}h)^{*}T\mathbb{R}^{n} over MM, that is, the pushforward of ff by UthC(M,n)U_{t}h\in C^{\infty}(M,\mathbb{R}^{n}), denoted by (Uth)f(U_{t}h)_{*}f [39]. Furthermore, let Φt\Phi_{t} denote the time-tt flow of the system, then Uth=hΦtU_{t}h=h\circ\Phi_{t} exactly gives the follow of the output generated by the observable hh. In another words, (Uth)f=(hΦt)f(U_{t}h)_{*}f=(h\circ\Phi_{t})_{*}f is the pushforward of ff by the flow of the system output as desired.

In the case that the controllable submanifold of the system is NMN\subset M, by Corollary 2, we have ddtUth(f)=ιd(Uth)(f)=d(ιUth)(f)\frac{d}{dt}U_{t}h(f)=\iota^{*}d(U_{t}h)(f)=d(\iota^{*}U_{t}h)(f), where ι:NM\iota:N\hookrightarrow M is the inclusion of NN into MM and we use the commutativity between the de Rham differential and pullback maps. However, because the system is controllable only on NN, the flow Φt\Phi_{t} of the system completely lies in NN so that the system output is given by Uth=hΦt=hΦtι=ι(Uth)U_{t}h=h\circ\Phi_{t}=h\circ\Phi_{t}\circ\iota=\iota^{*}(U_{t}h). Therefore, the result follows the same proof as in the controllable case above. ∎

Specifically, for pursuing the goal of changing coordinates, it is of particular interest to apply the pushforward map ddtUth=(Uth)\frac{d}{dt}U_{t}h=(U_{t}h)_{*} to the vector field governing the dynamics of the system ddtx(t)=f(x(t))+i=1mu(t)g(x(t))\frac{d}{dt}x(t)=f(x(t))+\sum_{i=1}^{m}u(t)g(x(t)). This yields

ddtUth(f+i=1muigi)=(Uth)(f+i=1muigi)\displaystyle\frac{d}{dt}U_{t}h\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}=(U_{t}h)_{*}\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}
=(hΦt)(f+i=1muigi)=h[(Φt)(f+i=1muigi)]\displaystyle=(h\circ\Phi_{t})_{*}\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}=h_{*}\Big{[}(\Phi_{t})_{*}\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}\Big{]}
=h[(f+i=1muigi)Φt]=h(f+i=1muigi)Φt\displaystyle=h_{*}\Big{[}\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}\circ\Phi_{t}\Big{]}=h_{*}\Big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\Big{)}\circ\Phi_{t} (22)

where (Φt)(f+i=1muigi)=(f+i=1muigi)Φt(\Phi_{t})_{*}\big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\big{)}=\big{(}f+\sum_{i=1}^{m}u_{i}g_{i}\big{)}\circ\Phi_{t} follows from the fact that the vector field f+i=1muigif+\sum_{i=1}^{m}u_{i}g_{i} is invariant under its own flow Φt\Phi_{t}.

Theorem 6.

Given any control-affine system, the action of the associated Koopman system on an observable that restricts to a local diffeomorphism on the controllable submanifold of the system defines a coordinate transformation for the system.

Proof.

We adopt the notations above that MM and NN denote the state-space manifold and controllable submanifold of the system, respectively, and hC(M,n)h\in C^{\infty}(M,\mathbb{R}^{n}) is the observable of the system satisfying the required assumption. Then, for any state x(t)Nx(t)\in N, there is a neighborhood UU in NN containing x(t)x(t) diffeomorphic to h(U)nh(U)\subseteq\mathbb{R}^{n}. On the other hand, because NN is the controllable submanifold of the system, the restrictions of the drift and control vector fields ff and gig_{i} to NN are tangent to NN, i.e., are not only vector fields along NN but also vector fields on NN. In the following, we identify hh, ff and gig_{i} with their restrictions on NN, where the system evloves. The diffeomorphism between UU and h(U)h(U) then implies that hfh_{*}f and hgh_{*}g, induced by the action of the associated Koopman system on hh as shown in (22), are well-defined vector fields on h(U)h(U), not only sections of the pullback bundle hTh(U)h^{*}Th(U). This immediately enables the representation of the system in the coordinates of h(U)h(U). Specifically, under the coordinate y=h(x)y=h(x) on h(U)h(U) induced by the observable hC(M,n)h\in C^{\infty}(M,\mathbb{R}^{n}), the flow of the system is exactly hΦt=Uthh\circ\Phi_{t}=U_{t}h, and hence the derivation in (22) leads to the representation

ddty(t)=hf(y(t))+i=1mui(t)hgi(y(t))\displaystyle\frac{d}{dt}y(t)=h_{*}f(y(t))+\sum_{i=1}^{m}u_{i}(t)h_{*}g_{i}(y(t)) (23)

of the system the system ddtx(t)=f(x(t))+i=1mu(t)g(x(t))\frac{d}{dt}x(t)=f(x(t))+\sum_{i=1}^{m}u(t)g(x(t)) in this coordinate on h(U)h(U), concluding the proof. ∎

It is worth to emphasize again that in the system in (23), hh, ff and gig_{i} should be interpreted as their restrictions on the controllable submanifold where the original system ddtx(t)=f(x(t))+i=1mui(t)gi(x(t))\frac{d}{dt}x(t)=f(x(t))+\sum_{i=1}^{m}u_{i}(t)g_{i}(x(t)) evolves on. More importantly, this further illustrates a distinctive feature as well as a great advantage of the Koopman change of coordinates. Conventionally, an observable hC(M,n)h\in C^{\infty}(M,\mathbb{R}^{n}) is required to be a diffeomorphism to generate a change of coordinates for the system on the state-space manifold MM. However, by leveraging the associated Koopman system, the change of coordinates for the system can be localized to a controllable submanifold NN with the requirement for hh relaxed to be a local diffeomorphism when restricted to NN. This localization is exactly enabled by the composition with the system flow Φt\Phi_{t} in the Koopman operator, for Φt\Phi_{t} is defined on the space where the system evolves as revealed in the proof of Proposition 3.

V-B Koopman feedback linearization

It is no doubt that linear systems are generally preferable, as an application of the Koopman change of coordinates, this section focuses on the case that a control-affine system as in (17) can be transformed to a linear system in the coordinates induced by the associated Koopman system. In particular, recalling the transformed system in (23), if it is a linear system, then hfh_{*}f is a linear vector field and hgih_{*}g_{i}, i=1,,mi=1,\dots,m are all constant vector fields. This imposes extremely strong conditions on the original system, e.g., the pushforward of the Lie bracket h[gi,gj]h_{*}[g_{i},g_{j}] between any pair of control vector fields gig_{i} and gjg_{j} has to vanish identically because the pushforward operation is natural in the sense of h[gi,gj]=[hgi,hgi]=0h_{*}[g_{i},g_{j}]=[h_{*}g_{i},h_{*}g_{i}]=0 [19], where hgih_{*}g_{i} and hgjh_{*}g_{j} are constant vector fields and hence have trivial Lie bracket. For the purpose of looking for less restrictive conditions, we first note that, in the transformed system in (17), the new coordinates are actually constructed from the output function hh in the form of y(t)=h(x(t))y(t)=h(x(t)), and this inspires the adoption of the feedback linearization technique. The idea of this technique is to apply feedback control laws in the form of ui(t)=αi(x)+j=1mβij(x)vj(t)u_{i}(t)=\alpha_{i}(x)+\sum_{j=1}^{m}\beta_{ij}(x)v_{j}(t) for some αi,βijC(M)\alpha_{i},\beta_{ij}\in C^{\infty}(M) and all i,j=1,,mi,j=1,\dots,m so that the forced system

ddtx(t)=[f(x(t))+i=1m\displaystyle\frac{d}{dt}x(t)=\big{[}f(x(t))+\sum_{i=1}^{m} αi(x(t))gi(x(t))]\displaystyle\alpha_{i}(x(t))g_{i}(x(t))\big{]}
+i=1mj=1mβijvj(t)gi(x(t))\displaystyle+\sum_{i=1}^{m}\sum_{j=1}^{m}\beta_{ij}v_{j}(t)g_{i}(x(t)) (24)

can be linearized by using a coordinate system determined by the observable hh. However, it has been shown that feedback linearizability immediately implies controllability, i.e., controllability is a necessary condition to guarantee the system to be feedback linearizable [27]. In the sequel, we will integrate the feedback linearization technique into the Koopman framework, and as a major advantage, the controllability condition can be dropped due to the fact that the change of coordinates generated by the associated Koopman system can be localized to the controllable submanifold of the system as discussed at the end of Section V-A.

Specifically, let NN be the controllable submanifold of the system and ι:NM\iota:N\hookrightarrow M the inclusion, then we assume that there is an observable h=(h1,,hl)C(M,l)h=(h_{1},\dots,h_{l})\in C^{\infty}(M,\mathbb{R}^{l}) satisfying the following conditions in a neighborhood UNU\subseteq N of a point pNp\in N:

  • for each i=1,,li=1,\dots,l and j=1,,mj=1,\dots,m, there is some integer rir_{i} such that gifk(hjι)=0\mathcal{L}_{g_{i}}\mathcal{L}^{k}_{f}(h_{j}\circ\iota)=0 for all k=1,,ri2k=1,\dots,r_{i}-2.

  • the matrix of functions in C(M)C^{\infty}(M)

    R=[g1fr11(h1ι)gmfr11(h1ι)g1fr21(h2ι)gmfr21(h2ι)g1frl1(hlι)gmfrl1(hlι)]\displaystyle R=\left[\begin{array}[]{ccc}\mathcal{L}_{g_{1}}\mathcal{L}^{r_{1}-1}_{f}(h_{1}\circ\iota)&\cdots&\mathcal{L}_{g_{m}}\mathcal{L}^{r_{1}-1}_{f}(h_{1}\circ\iota)\\ \mathcal{L}_{g_{1}}\mathcal{L}^{r_{2}-1}_{f}(h_{2}\circ\iota)&\cdots&\mathcal{L}_{g_{m}}\mathcal{L}^{r_{2}-1}_{f}(h_{2}\circ\iota)\\ \vdots&&\vdots\\ \mathcal{L}_{g_{1}}\mathcal{L}^{r_{l}-1}_{f}(h_{l}\circ\iota)&\cdots&\mathcal{L}_{g_{m}}\mathcal{L}^{r_{l}-1}_{f}(h_{l}\circ\iota)\end{array}\right]

    has rank ll.

Then, we say that the system has relative degree (r1,,rl)(r_{1},\dots,r_{l}) at pp.

Proposition 4.

Suppose the observable hC(M,l)h\in C^{\infty}(M,\mathbb{R}^{l}) of the control-affine system in (17) has the relative degree (r1,,rl)(r_{1},\dots,r_{l}) at pNp\in N, then the matrix of 1-forms on MM

[ddtUth1ddtUtfh1ddtUtfrhiddtUth2ddtUtfh2ddtUtfrh2ddtUthlddtUtfhlddtUtfrhl]\displaystyle\left[\begin{array}[]{cccc}\frac{d}{dt}U_{t}h_{1}&\frac{d}{dt}U_{t}\mathcal{L}_{f}h_{1}&\cdots&\frac{d}{dt}U_{t}\mathcal{L}_{f}^{r}h_{i}\\ \frac{d}{dt}U_{t}h_{2}&\frac{d}{dt}U_{t}\mathcal{L}_{f}h_{2}&\cdots&\frac{d}{dt}U_{t}\mathcal{L}_{f}^{r}h_{2}\\ \vdots&\vdots&&\vdots\\ \frac{d}{dt}U_{t}h_{l}&\frac{d}{dt}U_{t}\mathcal{L}_{f}h_{l}&\cdots&\frac{d}{dt}U_{t}\mathcal{L}_{f}^{r}h_{l}\end{array}\right] (29)

has rank ll evaluated at every points in a neighborhood UU of pp in NN, where UtU_{t} denote the associated Koopman operator and r=max{r1,,rl}r=\max\{r_{1},\dots,r_{l}\}.

Proof.

According to Corollary 2 that ddtUt\frac{d}{dt}U_{t} acts on observables by ιd\iota^{*}d, we have ddtUtfkhi=ιdfkhi=dιfkhi=d(fkhiι)\frac{d}{dt}U_{t}\mathcal{L}_{f}^{k}h_{i}=\iota^{*}d\mathcal{L}_{f}^{k}h_{i}=d\iota^{*}\mathcal{L}_{f}^{k}h_{i}=d(\mathcal{L}_{f}^{k}h_{i}\circ\iota), where we use the commutativity of the pullback and de Rham differential. We now claim fhι=f(hι)\mathcal{L}_{f}h\circ\iota=\mathcal{L}_{f}(h\circ\iota). To see this, we use the definition of Lie derivatives f(hι)=d(hι)(f)=dhdι(f)\mathcal{L}_{f}(h\circ\iota)=d(h\circ\iota)(f)=dh\circ d\iota(f). Because NN is the controllable submanifold of the system, f|N𝔛(N)f|_{N}\in\mathfrak{X}(N) holds, which implies dι(f)=f|Nd\iota(f)=f|_{N}, yielding the claim. By induction on the order kk of the successive Lie derivatives, we obtain ddtUtfkhi=dfk(hiι)\frac{d}{dt}U_{t}\mathcal{L}_{f}^{k}h_{i}=d\mathcal{L}_{f}^{k}(h_{i}\circ\iota).

We now consider the ll-by-mm block matrix GG with the (i,j)(i,j)-block GijG_{ij} the rir_{i}-by-rr matrix whose (α,β)(\alpha,\beta)-entry is given by

ddtUtα1\displaystyle\frac{d}{dt}U_{t}\mathcal{L}^{\alpha-1} hi(fβ1gj)=dfα1(hiι)(fβ1gj)\displaystyle h_{i}(\mathcal{L}_{f}^{\beta-1}g_{j})=d\mathcal{L}_{f}^{\alpha-1}(h_{i}\circ\iota)(\mathcal{L}_{f}^{\beta-1}g_{j})
=d(fα1hi)(fβ1gj)ι\displaystyle=d(\mathcal{L}_{f}^{\alpha-1}h_{i})(\mathcal{L}_{f}^{\beta-1}g_{j})\circ\iota
=k=0β1(1)k(β1k)fβ1kgjfα1+khiι\displaystyle=\sum_{k=0}^{\beta-1}(-1)^{k}{\beta-1\choose k}\mathcal{L}_{f}^{\beta-1-k}\mathcal{L}_{g_{j}}\mathcal{L}_{f}^{\alpha-1+k}h_{i}\circ\iota

where we use the facts f|N,gi|N𝔛(N)f|_{N},g_{i}|_{N}\in\mathfrak{X}(N) and the product rule of Lie derivatives. Let UNU\subseteq N be a neighborhood of xx as in the definition of relative degree, then in UU the matrix GijG_{ij} satisfies

ddtUtα1\displaystyle\frac{d}{dt}U_{t}\mathcal{L}^{\alpha-1} hi(fβ1gj)\displaystyle h_{i}(\mathcal{L}_{f}^{\beta-1}g_{j})
={0,if α+β<ri+1(1)β1giλfri1hiι,if α+β=ri+1\displaystyle=\begin{cases}0,\quad\text{if }\alpha+\beta<r_{i}+1\\ (-1)^{\beta-1}\mathcal{L}_{g_{i}}\lambda_{f}^{r_{i}-1}h_{i}\circ\iota,\quad\text{if }\alpha+\beta=r_{i}+1\end{cases}

Correspondingly, the upper left triangular block of GijG_{ij} consists of only 0, and the antidiagonal elements are equal to (1)β1giλfri1hi(-1)^{\beta-1}\mathcal{L}_{g_{i}}\lambda_{f}^{r_{i}-1}h_{i}. Therefore, rearranging the rows of GG results in a block triangular matrix with each of the diagonal blocks equal to the matrix RR. Because AA is full rank by the definition of the relative degree, the desired result follows. ∎

An another interpretation of Proposition 4 is that the matrix in (29) has full row rank, equivalently, the row vectors are linearly independent in UNU\subseteq N. This particularly implies that (ddtUth1,,ddtUtfr11h1,,ddtUthl,,ddtUtfrl1hl)(\frac{d}{dt}U_{t}h_{1},\dots,\frac{d}{dt}U_{t}\mathcal{L}_{f}^{r_{1}-1}h_{1},\dots,\frac{d}{dt}U_{t}h_{l},\dots,\frac{d}{dt}U_{t}\mathcal{L}_{f}^{r_{l}-1}h_{l}) qualifies as a coordinate chart of NN on UU provided r1++rl=dimNr_{1}+\cdots+r_{l}={\rm dim}\,N, which is exactly the desired coordinates linearizing the nonlinear control-affine system in (24), that is, the system in (17) steered by a feedback control law.

Theorem 7 (Koopman feedback linearization).

Given a control-affine system defined on MM as in (17), and let NMN\subseteq M be the controllable submanifold and h=(h1,,hl)C(M,l)h=(h_{1},\dots,h_{l})\in C^{\infty}(M,\mathbb{R}^{l}) be an observable of the system. If the system has relative degree (r1,,rl)(r_{1},\dots,r_{l}) at pNp\in N such that r1++rl=nr_{1}+\dots+r_{l}=n, the dimension of NN, then the action of the associated Koopman system on the observable ϕ=(h1,,fr11h1,,hl,,fri1hl)C(M,n)\phi=(h_{1},\dots,\mathcal{L}_{f}^{r_{1}-1}h_{1},\dots,h_{l},\dots,\mathcal{L}_{f}^{r_{i}-1}h_{l})\in C^{\infty}(M,\mathbb{R}^{n}) gives a linear system defined on and neighborhood UNU\subseteq N of pp in the form of

ddtUtϕ=AUtϕ+Bv(t)\displaystyle\frac{d}{dt}U_{t}\phi=AU_{t}\phi+Bv(t)

where

A=[A1Al]n×n,B=[b1bl]n×mA=\left[\begin{array}[]{ccc}A_{1}&&\\ &\ddots&\\ &&A_{l}\end{array}\right]\in\mathbb{R}^{n\times n},\ B=\left[\begin{array}[]{ccc}b_{1}&&\\ &\ddots&\\ &&b_{l}\end{array}\right]\in\mathbb{R}^{n\times m}

and v(t)=(v1(t),,vl(t))lv(t)=(v_{1}(t),\dots,v_{l}(t))^{\prime}\in\mathbb{R}^{l} with

Ai=[0100001000010000]ri×ri,bi=[0001]riA_{i}=\left[\begin{array}[]{ccccc}0&1&0&\cdots&0\\ 0&0&1&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&1\\ 0&0&0&\cdots&0\end{array}\right]\in\mathbb{R}^{r_{i}\times r_{i}},\ b_{i}=\left[\begin{array}[]{c}0\\ 0\\ \vdots\\ 0\\ 1\end{array}\right]\in\mathbb{R}^{r_{i}}

and vi(t)=j=1muj(t)Utgjfri1(hiι)+Utfri(hiι)v_{i}(t)=\sum_{j=1}^{m}u_{j}(t)U_{t}\mathcal{L}_{g_{j}}\mathcal{L}_{f}^{r_{i}-1}(h_{i}\circ\iota)+U_{t}\mathcal{L}_{f}^{r_{i}}(h_{i}\circ\iota), respectively, for all i=1,,li=1,\dots,l.

Proof.

We first recall that the action of the associated Koopman system ddtUtϕ\frac{d}{dt}U_{t}\phi gives the differential of UtϕU_{t}\phi, then by Proposition 4, the assumption r1++rl=nr_{1}+\dots+r_{l}=n implies that ϕ\phi restricts to a diffeomorphism on a neighborhood UNU\subseteq N of xx to n\mathbb{R}^{n}. This implies that ϕ\phi gives rise to a coordinate chart of NN on UU. Next, to compute the coordinate representation of the system in (17) in this chart, pick any qUq\in U as the initial condition of the system x(0)=qx(0)=q and tt small enough so that the trajectory of the system stays in UU, then for any i=1,,li=1,\dots,l and k=0,,ri1k=0,\dots,r_{i}-1, we have Uthi(q)=h(x(t))U_{t}h_{i}(q)=h(x(t)) and

ddtUtfkhi(p)\displaystyle\frac{d}{dt}U_{t}\mathcal{L}_{f}^{k}h_{i}(p) =ddtfkhi(x(t))=dfkhi(ddtx(t))\displaystyle=\frac{d}{dt}\mathcal{L}_{f}^{k}h_{i}(x(t))=d\mathcal{L}_{f}^{k}h_{i}\Big{(}\frac{d}{dt}x(t)\Big{)}
=dfkhi(f+j=1mujgj)(x(t))\displaystyle=d\mathcal{L}_{f}^{k}h_{i}\big{(}f+\sum_{j=1}^{m}u_{j}g_{j}\big{)}(x(t))
=fk+1hi(x(t))+j=1mujgjfkhi(x(t)),\displaystyle=\mathcal{L}_{f}^{k+1}h_{i}(x(t))+\sum_{j=1}^{m}u_{j}\mathcal{L}_{g_{j}}\mathcal{L}_{f}^{k}h_{i}(x(t)),

Because NN is the controllable submanifold such that x(t)UNx(t)\in U\subseteq N can be arbitrary, we obtain

ddtUtfkhiι\displaystyle\frac{d}{dt}U_{t}\mathcal{L}_{f}^{k}h_{i}\circ\iota =Utfk+1hiι+j=1mujUtgjfkhiι,\displaystyle=U_{t}\mathcal{L}_{f}^{k+1}h_{i}\circ\iota+\sum_{j=1}^{m}u_{j}U_{t}\mathcal{L}_{g_{j}}\mathcal{L}_{f}^{k}h_{i}\circ\iota,

in which gjfkhiι=0\mathcal{L}_{g_{j}}\mathcal{L}_{f}^{k}h_{i}\circ\iota=0 unless k=ri1k=r_{i}-1 by the relative degree assumption and ι:NM\iota:N\hookrightarrow M denotes the inclusion. With the choice of v(t)v(t) in the theorem statement, the desired result follows. ∎

In the classical nonlinear systems theory, the notion of relative degree, and hence feedback linearization, is defined on a neighborhood of a point in the state-space of a system. Moreover, a feedback linearizable system is necessarily controllable, otherwise, the system can at most be partially linearized [27]. However, the integration of the associated Koopman system localizes the feedback linearization technique to the controllable submanifold of the system by effectively eliminating the uncontrollable states of the system, which in turn greatly expands the scope of feedback linearization.

VI Conclusion

In this paper, we develop a Koopman control framework for analyzing control systems by using Koopman operators. Based on the detailed investigation into the algebraic and analytical properties of Koopman operators, instead of spectral methods and ergodicity assumptions, we rigorously derive the infinite-dimensional differential equation system governing the dynamics of the Koopman operator, referred to as the Koopman system, associated with a control system. Especially, for a control-affine system, the associated Koopman system gives rise to a global bilinearizaton of the system. In addition, we further reveal the Gelfand duality between control-affine and bilinear Koopman systems, which greatly benefits controllability analysis of these systems. In part, it leads to the characterization of controllability for control-affine systems in terms of de Rham differenal operators. On the other hand, by leveraging techniques in infinite-dimensional geometry, we show that bilinear Koopman systems are defined on infinite-dimensional Lie groups and carry out an extension of LARC to such infinite-dimensional systems. The developed framework is then adopted in the context of feedback linearization, with the emphasis on the localization nature of Koopman systems. As a consequence, the scope of the classical feedback linearization technique is extended by including uncontrollable systems. This work not only greatly expands the repertoire of tools in geometric control theory, but also sheds light on novel finite-dimensional approaches to the study of infinite-control dimensional control systems, through the bridge of the Gelfrand duality between finite-dimensional systems and the associated infinite-dimensional Koopman systems.

References

  • [1] B. O. Koopman, “Hamiltonian systems and transformation in hilbert space,” in Proceedings of National Academy of Sciences, vol. 17, pp. 315–318, 1931.
  • [2] P. Walters, An Introduction to Ergodic Theory, vol. 79 of Graduate Texts in Mathematics. Springer-Verlag New York, 1 ed., 1982.
  • [3] K. E. Petersen, Ergodic Theory. Cambridge Studies in Advanced Mathematics, Cambridge University Press, 1983.
  • [4] R. M. né, Ergodic Theory and Differentiable Dynamics, vol. 8 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. Springer Berlin, Heidelberg, 1987.
  • [5] E. Hopf, “Ergodic theory and the geodesic flow on surfaces of constant negative curvature,” Bulletin of the American Mathematical Society, vol. 77, pp. 863–877, November 1971.
  • [6] A. Katok and B. Hasselblatt, Introduction to the Modern Theory of Dynamical Systems, vol. 54 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, 1995.
  • [7] I. Karatzas and S. E. Shreve, Brownian Motion and Stochastic Calculus, vol. 2 of Graduate Texts in Mathematics. Springer New York, NY, 2 ed., 1998.
  • [8] J. L. Doob, Stochastic Processes. Wiley Classics Library, Wiley-Interscience, 1991.
  • [9] I. P. Cornfeld, S. V. Fomin, and Y. G. Sinai, Ergodic Theory, vol. 245 of Grundlehren der mathematischen Wissenschaften. Springer New York, NY, 1982.
  • [10] M. Einsiedler and T. Ward, Ergodic Theory with a view towards Number Theory, vol. 259 of Graduate Texts in Mathematics. Springer-Verlag London, 2011.
  • [11] Y. Susuki, I. Mezić, F. Raak, and T. Hikihara, “Applied koopman operator theory for power systems technology,” Nonlinear Theory and Its Applications, pp. 430–459, 10 2016.
  • [12] C. W. Rowley, I. Mezić, S. Bagheri, P. Schlatter, and D. S. Henningson, “Spectral analysis of nonlinear flows,” Journal of Fluid Mechanics, vol. 641, pp. 115–127, December 2009.
  • [13] M. B. R. Mohr and I. Mezić, “Applied koopmanism,” Chaos: An Interdisciplinary Journal of Nonlinear Science, vol. 22, no. 4, 2012.
  • [14] I. Mezić, “Analysis of fluid flows via spectral properties of the koopman operator,” Annual Review of Fluid Mechanics, vol. 45, no. 1, pp. 357–378, 2013.
  • [15] Y. Susuki, I. Mezić, and T. Hikihara, “Coherent swing instability of power grids,” Journal of Nonlinear Science, vol. 21, pp. 403–439, Jun 2011.
  • [16] F. Raak, Y. Susuki, I. Mezić, and T. Hikihara, “On koopman and dynamic mode decompositions for application to dynamic data with low spatial dimension,” in 2016 IEEE 55th Conference on Decision and Control (CDC), (Las Vegas, NV, USA), pp. 6485–6491, December 2016.
  • [17] W. Zhang, Y.-C. Yu, and J.-S. Li, “Dynamics reconstruction and classification via koopman features,” Data Mining and Knowledge Discovery, vol. 33, pp. 1710–1735, 2019.
  • [18] S. Qian, C.-A. Chou, and J.-S. Li, “Deep multi-modal learning for joint linear representation of nonlinear dynamical systems,” Scientific Reports, vol. 12, no. 1, p. 12807, 2022.
  • [19] J. M. Lee, Introduction to Smooth Manifolds, vol. 218 of Graduate Texts in Mathematics. Springer New York, NY, 2 ed., 2012.
  • [20] H. H. Schaefer and M. P. Wolff, Topological Vector Spaces, vol. 3 of Graduate Texts in Mathematics. Springer New York, NY, 2 ed., 1999.
  • [21] J. Munkres, Topology. Pearson College Div, 2000.
  • [22] K. Yosida, Functional Analysis, vol. 8 of Classics in Mathematics. Springer Berlin, Heidelberg, 1995.
  • [23] R. Brockett, Finite Dimensional Linear Systems, vol. 7 of Decision and Control, Electronic & Electrical Engineering Research Studies. John Wiley & Sons, Inc., 1970.
  • [24] H. Khalil, Nonlinear Systems. Pearson, 3 ed., 2001.
  • [25] R. Brockett, “Nonlinear systems and differential geometry,” Proceedings of the IEEE, vol. 64, no. 1, pp. 61–72, 1976.
  • [26] H. Hermes, “Lie algebras of vector fields and local approximation of attainable sets read more: http://epubs.siam.org/doi/abs/10.1137/0316047,” SIAM Journal on Control and Optimization, vol. 33, no. 16, pp. 715–727, 1978.
  • [27] A. Isidori, Nonlinear Control Systems. Communications and Control Engineering, Springer London, 3 ed., 1995.
  • [28] V. Jurdjevic, Geometric Control Theory, vol. 52 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1996.
  • [29] R. Brockett, “The early days of geometric nonlinear control,” Automatica, vol. 50, no. 9, pp. 2203 – 2224, 2014.
  • [30] R. Brockett, “System theory on group manifolds and coset spaces,” SIAM Journal on Control, vol. 10, no. 2, pp. 265–284, 1972.
  • [31] V. Jurdjevic and H. Sussmann, “Control systems on lie groups,” Journal of Differential Equations, vol. 12, no. 2, pp. 313 – 329, 1972.
  • [32] R. Brockett, “Lie theory and control systems defined on spheres,” SIAM Journal on Applied Mathematics, vol. 25, no. 2, pp. 213–225, 1973.
  • [33] M. Golubitsky and V. Guillemin, Stable Mappings and Their Singularities, vol. 14 of Graduate Texts in Mathematics. Springer New York, NY, 1973.
  • [34] T. W. Palmer, Banach Algebras and the General Theory of *-Algebras Volume 2, vol. 79 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, 2001.
  • [35] M. P. do Carmo, Riemannian Geometry. Mathematics: Theory & Applications, Birkhäuser, Boston, MA, 1992.
  • [36] J. Milnor, Remarks on infinite-dimensional Lie groups. Netherlands: North-Holland, 1984.
  • [37] W. Thurston, “Foliations and groups of diffeomorphisms,” Bulletin of the American Mathematical Society, vol. 80, pp. 304–307, 1974.
  • [38] A. Kriegl and P. W. Michor, The Convenient Setting of Global Analysis, vol. 53 of Mathematical Surveys and Monographs. American Mathematical Socity, 1997.
  • [39] J. Jost, Riemannian Geometry and Geometric Analysis. Universitext, Springer Cham, 7 ed., 2017.