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

Towards Global Optimal Control via Koopman Lifts

Mario E. Villanueva [email protected]    Colin N. Jones [email protected]    Boris Houska [email protected] School of Information Science and Technology, ShanghaiTech, China Automatic Control Laboratory, EPFL, Switzerland
Abstract

This paper introduces a framework for solving time-autonomous nonlinear infinite horizon optimal control problems, under the assumption that all minimizers satisfy Pontryagin’s necessary optimality conditions. In detail, we use methods from the field of symplectic geometry to analyze the eigenvalues of a Koopman operator that lifts Pontryagin’s differential equation into a suitably defined infinite dimensional symplectic space. This has the advantage that methods from the field of spectral analysis can be used to characterize globally optimal control laws. A numerical method for constructing optimal feedback laws for nonlinear systems is then obtained by computing the eigenvalues and eigenvectors of a matrix that is obtained by projecting the Pontryagin-Koopman operator onto a finite dimensional space. We illustrate the effectiveness of this approach by computing accurate approximations of the optimal nonlinear feedback law for a Van der Pol control system, which cannot be stabilized by a linear control law.

keywords:
Pontryagin’s maximum principle, Koopman operators, symplectic geometry, optimal feedback control
thanks: MEV and BH acknowledge financial support via ShanghaiTech U. Grant F-0203-14-012. CJ would like to acknowledge support received from the Swiss National Science Foundation under the RISK project (Risk Aware Data-Driven Demand Response), grant number 200021 175627.

, ,

1 Introduction

During the last decades algorithms and software for computing local solutions of nonlinear optimal control problems have reached a high level of maturity [4, 9, 35], enabling its deployment in industrial applications [29]. By now, such local solutions can be computed within the milli- and microsecond range. Moreover, auto-generated implementations of real-time local optimal control solvers run on embedded hardware systems [24, 18]. In contrast to these developments in local optimal control, algorithms for locating global minimizers of non-convex optimal control problems, can hardly be claimed to be ready for widespread industrial application. There are at least two reasons for this. Firstly, generic global optimization methods can often only be applied to problems of modest dimensions. And secondly, their run-time usually exceeds the run-time of local solvers by orders of magnitude. Nevertheless, in the last years there have been a number of promising developments in the field of global optimal control, which are reviewed next.

Dynamic Programming (DP) [3], which proceeds by approximately solving the Hamilton-Jacobi-Bellman (HJB) equation [13], is a historically important method able to find globally optimal feedback laws. In [23] and [15] tailored discretization grids for DP implementations have been developed that can successfully solve nonlinear optimal control problems, as long as the dimension of the state of the system is small. For higher dimensional state-spaces these methods are, however, not applicable as the complexity of storing and processing the value functions during the DP recursion grows exponentially with the number of states. Other deterministic methods for global optimization problems involving nonlinear differential equations are based on Branch-and-Bound (BB) [7, 11, 22, 32] and its variants, including α\alpha-BB methods [27, 8]. These BB algorithms have in common that they can effectively solve problems with a small number of decision variables. Unfortunately, implementations of branch-and-bound search easily run out of memory in higher dimensional spaces due to exponentially growing search trees. In particular, as the discretization of control inputs typically leads to a large number of optimization variables, BB-methods are usually not suited for solving optimal control problems.

An alternative to Branch-and-Bound are the so-called Branch-and-Lift (BL) methods [16]. In contrast to BB, these methods never discretize the control inputs directly. Instead, these methods branch over orthogonal projections of the control function in lower dimensional subspaces of increasing dimension, until an ϵ\epsilon-suboptimal global control input is found. A rigorous analysis of the mathematical properties of such BL methods can be found in [17]. However, in the context of developing practical implementations of Branch-and-Lift one faces numerical challenges. In particular, these methods require the computation of accurate enclosures of moment-constrained reachable sets of nonlinear differential equations. At the current state of research, the lack of generic methods for computing such moment-constrained reachable set enclosures limits the applicability of Branch-and-Lift—currently, some problems have been solved successfully with BL by using enclosure propagation methods that are tailored to particular applications [12, 16].

In order to mitigate the above-mentioned limitations of existing global optimal control methods, this paper proposes a completely new framework for constructing methods for solving nonlinear infinite horizon optimal control problems. Here, the main idea is to analyze the spectrum of a Koopman operator that is associated with Pontryagin’s differential equation—under the assumption that Pontryagin’s optimality condition [28] is satisfied at the optimal solution. In order to understand the contributions of this paper, it is important to first be aware of the fact that Koopman operators have originally been introduced almost 9090 years ago—in fact, by Koopman himself [19]. This theory has later been extended for more general nonlinear differential equations in [26], where the concept of Koopman mode analysis has been introduced. A more complete analysis of these Koopman mode-based methods from the field of ergodic theory has, however, only appeared much later [1]. Notice that a mathematical analysis of finite dimensional approximations of the Koopman operator can be found in [14] and a variety of related applications of approximate Koopman mode analysis methods can be found in [6]. Moreover, in [20] Koopman mode estimation heuristics for data driven control have been introduced.

Contribution.

The key idea of this paper is to introduce a Koopman operator that is associated with the Pontryagin differential equation of a nonlinear infinite-horizon optimal control problem—in the following we called a Pontryagin-Koopman operator—which can be used to construct globally optimal feedback laws. Section 2 introduces infinite horizon optimal control problems and briefly reviews Pontryagin’s necessary condition for optimality. This review section is followed by a presentation the main contributions of this article, which can be outlined as follows.

  1. 1.

    Section 3 uses ideas from the field of symplectic geometry [2] to characterize the structural properties of Pontryagin-Koopman operators. Section 4 leverages on these structural properties in order to perform a symplectic spectral analysis that finally leads to a complete characterization of globally optimal feedback control laws (see Section 5, Theorem 18). Because this is the first time that the symplecticity of Pontryagin-Koopman operators is analyzed and used to derive practical characterizations of optimal control laws, Theorem 18 can be regarded as the main theoretical contribution of this paper.

  2. 2.

    In order to avoid misunderstandings, it is mentioned clearly that this paper does not claim to solve all the numerical issues regarding the discretization of Pontryagin-Koopman operators that would have to be solved to develop a generic software for global optimal control. However, Section 6 illustrates the practical applicability of the proposed theoretical framework by designing an optimal regulator for a controlled Van der Pol oscillator. Although this implementation is only based on a naive Galerkin projection of the Pontryagin-Koopman operator onto a finite dimensional Legendre basis, this section successfully applies the proposed framework to construct accurate approximations of nonlinear globally optimal feedback laws—in this example, for a system that cannot be stabilized by a linear control law.

The above contributions are relevant for the future of optimal control algorithm development, as they pave the way towards the development of practical procedures for approximating globally optimal control laws of a very general class of nonlinear systems. Therefore, Section 7 does not only conclude the paper, but also outlines the potential and relevance of the proposed framework for future research in control.

Notation.

The distance of a point xnxx\in\mathbb{R}^{n_{x}} to a trajectory ϕ:n\phi:\mathbb{R}\to\mathbb{R}^{n} is denoted by

dist(x,ϕ)=inftxϕ(t).\mathrm{dist}\left(x,\phi\right)=\inf_{t\in\mathbb{R}}\|x-\phi(t)\|\;.

We denote with 𝕃pn\mathbb{L}_{p}^{n} the set of (potentially complex-valued) Lebesgue measurable functions φ:n\varphi:\mathbb{R}^{n}\to\mathbb{C} whose pp-th power of the absolute value is integrable on n\mathbb{R}^{n}. Moreover, we use the notation 𝕎k,pn\mathbb{W}_{k,p}^{n} to denote the Sobolev space of integrable functions on n\mathbb{R}^{n}, whose weak derivatives up to order kk are all in 𝕃pn\mathbb{L}_{p}^{n}. The symbol MM^{\intercal} denotes the Hermitian transpose of the matrix Mn×mM\in\mathbb{C}^{n\times m}. The symbol \nabla denotes the gradient operator. The associated second order derivative operator is denoted by 2=\nabla^{2}=\nabla\nabla^{\intercal}. Last but not least, the support of a function φ:n\varphi:\mathbb{R}^{n}\to\mathbb{C} is denoted by

supp(φ)={xnφ(x)0}.\mathrm{supp}(\varphi)=\{x\in\mathbb{R}^{n}\mid\varphi(x)\neq 0\}\;.

2 Infinite horizon optimal control

This section introduces infinite horizon optimal control problems and briefly summarizes existing methods for analyzing the local stability properties of optimal periodic limit orbits (see Section 2.2). Moreover, Sections 2.3 and 2.4 review Pontryagin’s necessary optimal condition thereby introducing the notation that is used throughout the paper.

2.1 Problem formulation

This paper considers infinite horizon optimal control problems of the form

V(x0)=\displaystyle V(x_{0})= minx,u\displaystyle\min_{x,u} 0[l(x(t),u(t))l]dt\displaystyle\int_{0}^{\infty}\left[l(x(t),u(t))-l^{\star}\right]\,\mathrm{d}t (1)
s.t.\displaystyle\mathrm{s.t.} {t[0,),x˙(t)=f(x(t),u(t))x(0)=x0.\displaystyle\left\{\begin{aligned} &\forall t\in[0,\infty),\\[2.84544pt] &\dot{x}(t)=f(x(t),u(t))\\[2.84544pt] &x(0)=x_{0}\;.\end{aligned}\right.

Here, x:nxx:\mathbb{R}\to\mathbb{R}^{n_{x}} denotes the state trajectory and u:nuu:\mathbb{R}\to\mathbb{R}^{n_{u}} denotes the control input. The initial value x0nxx_{0}\in\mathbb{R}^{n_{x}} is assumed to be given. Throughout this paper the following assumptions are imposed.

Assumption 1.

The function f:nx×nunxf:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{u}}\to\mathbb{R}^{n_{x}} is twice continuously differentiable and globally Lipschitz continuous in xx.

Assumption 2.

The function l:nx×nul:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{u}}\to\mathbb{R} is twice continuously differentiable.

The constant ll^{\star}\in\mathbb{R} in (1) denotes the optimal average cost,

l=limT\displaystyle l^{\star}=\lim_{T\to\infty} minx,u1T0Tl(x(t),u(t))dt\displaystyle\min_{x,u}\frac{1}{T}\int_{0}^{T}l(x(t),u(t))\,\mathrm{d}t
s.t.{t[0,T],x˙(t)=f(x(t),u(t))x(0)=x0,\displaystyle\;\,\mathrm{s.t.}\;\;\left\{\begin{aligned} &\forall t\in[0,T],\\[2.84544pt] &\dot{x}(t)=f(x(t),u(t))\\[2.84544pt] &x(0)=x_{0}\;,\end{aligned}\right.

assuming that this limit exists. Notice that ll^{\star} does not need to be known explicitly in order to solve (1), as adding constant offsets to the objective does not affect the solution of an optimization problem. In this paper, the constant ll^{\star} is merely introduced for mathematical reasons, such that V(x0)V(x_{0}) remains finite and well-defined on infinite horizons under mild regularity assumptions that shall be introduced further below.

2.2 Limit behavior and periodic orbits

In practical instances, one is often interested in whether optimal solutions for the state trajectory of (1) converge to an optimal steady-state or, in more generality, to an optimal periodic limit orbit. These optimal steady states or more general periodic orbits are defined as follows.

Definition 3.

The function (xp,up)𝕎1,1nx×𝕃1nx(x_{\mathrm{p}},u_{\mathrm{p}})\in\mathbb{W}_{1,1}^{n_{x}}\times\mathbb{L}_{1}^{n_{x}} is called an optimal periodic orbit, if there exists a period T>0T>0 such that for all t[0,)t\in[0,\infty)

  1. 1.

    xp(t+T)=xp(t)x_{\mathrm{p}}(t+T)=x_{\mathrm{p}}(t) and up(t+T)=up(t)u_{\mathrm{p}}(t+T)=u_{\mathrm{p}}(t),

  2. 2.

    x˙p(t)=f(xp(t),up(t))\dot{x}_{\mathrm{p}}(t)=f(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t)) for all t[0,T]t\in[0,T], and

  3. 3.

    1T0Tl(xp(t),up(t))dt=l\frac{1}{T}\int_{0}^{T}l(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))\,\mathrm{d}t=l^{\star}.

For the special case that xpx_{\mathrm{p}} and upu_{\mathrm{p}} are constant functions, this pair is called an optimal steady-state.

In order to prepare the following analysis, it is helpful to introduce the shorthands

A(t)\displaystyle A(t) =fx(xp(t),up(t)),B(t)=fu(xp(t),up(t)),\displaystyle=f_{x}(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))\;,\quad B(t)=f_{u}(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))\;, (2)
Q(t)\displaystyle Q(t) =lxx(xp(t),up(t)),R(t)=luu(xp(t),up(t)),\displaystyle=l_{xx}(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))\;,\quad R(t)=l_{uu}(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))\;,
S(t)\displaystyle S(t) =lxu(xp(t),up(t))\displaystyle=l_{xu}(x_{\mathrm{p}}(t),u_{\mathrm{p}}(t))

assuming that (xp,up)(x_{\mathrm{p}},u_{\mathrm{p}}) is an optimal periodic orbit. Here, fxf_{x} and fuf_{u} denote the partial derivatives of ff with respect to xx and uu and an analogous notation is then also used for the mixed second order derivatives of ll. The above matrix-valued functions can be used to construct sufficient conditions under which an optimal periodic orbit is locally stabilizable. Here, one relies on the theory of periodic Riccati differential equations [5] of the form

P˙(t)\displaystyle-\dot{P}(t) =P(t)A(t)+A(t)P(t)+Q(t)(P(t)B(t)+S(t))R(t)1(P(t)B(t)+S(t))\displaystyle=P(t)A(t)+A(t)^{\intercal}P(t)+Q(t)-(P(t)B(t)+S(t))R(t)^{-1}(P(t)B(t)+S(t))^{\intercal} (3)
P(0)\displaystyle P(0) =P(T)0.\displaystyle=P(T)\succ 0\;.

It is well-known [5] that if Assumptions 1 and 2 are satisfied and if RR has full rank, then the existence of a periodic and positive definite function PP satisfying (3) is sufficient to ensure that the solution xx^{\star} of (1) converges to the optimal periodic orbit xpx_{\mathrm{p}} as long as dist(x0,xp)\,\mathrm{dist}\left(x_{0},x_{\mathrm{p}}\right)\, is sufficiently small—that is, if x0x_{0} is in a small neighborhood of the optimal orbit xpx_{\mathrm{p}}. However, this statement is of a rather local nature. This means that, if we wish to understand the global behavior of system (1), an analysis of the periodic Riccati equation is not sufficient. The following sections focus on analyzing global solutions of (1), under the assumption that dist(x0,xp)\,\mathrm{dist}\left(x_{0},x_{\mathrm{p}}\right)\, is not necessarily small.

2.3 Pontryagin’s differential equations

Pontryagin’s maximum principle [28] can be used to derive necessary conditions for the minimizers of (1). The first order variational optimality condition is summarized as follows. Let H:nx×nu×nxH:\mathbb{R}^{n_{x}}\times\mathbb{R}^{n_{u}}\times\mathbb{R}^{n_{x}}\to\mathbb{R} be the Hamiltonian function of (1),

H(x,u,λ)=λf(x,u)+l(x,u).\displaystyle H(x,u,\lambda)=\lambda^{\intercal}f(x,u)+l(x,u)\;. (4)

If Assumptions 1 and 2 hold, HH is, by construction, twice continuously differentiable in all variables. Next, let

u(x,λ)=argmin𝑢H(x,u,λ)\displaystyle u^{\star}(x,\lambda)=\underset{u}{\text{argmin}}\;H(x,u,\lambda) (5)

denote the associated parametric minimizer of HH. At this point, we introduce the following regularity assumption.

Assumption 4.

The parametric minimizer uu^{\star} in (5) satisfies the second order sufficient condition, Huu0H_{uu}\succ 0.

Notice that for the practically relevant special case that ff is affine in uu and ll strongly convex in uu, Assumption (4) always holds whenever Assumptions 1 and 2 are satisfied.

Next, if Assumptions 12, and 4 are satisfied, it is well-known [28] that any minimizer (x,u)(x,u) of (1) necessarily satisfies Pontryagin’s differential equation

x˙(t)\displaystyle\dot{x}(t) =f(x(t),u(x(t),λ(t)))\displaystyle=f(x(t),u^{\star}(x(t),\lambda(t))) (6)
λ˙(t)\displaystyle\dot{\lambda}(t) =xH(x(t),u(x(t),λ(t)),λ(t))\displaystyle=-\nabla_{x}H(x(t),u^{\star}(x(t),\lambda(t)),\lambda(t)) (7)

for a co-state function λ𝕎1nx\lambda\in\mathbb{W}_{1}^{n_{x}}. In the following, we summarize this differential equation in the form

y˙(t)=F(y(t)),\displaystyle\dot{y}(t)=F(y(t))\;, (8)

where y=[x,λ]y=[x^{\intercal},\lambda^{\intercal}]^{\intercal} denotes the stacked state and FF a stacked version of the right-hand side functions of (6) and (7).

2.4 Necessary boundary and limit conditions

Besides Pontryagin’s differential equation (8), optimal solutions of (1) satisfy necessary boundary conditions. First, of course, the initial condition x(0)=x0x(0)=x_{0} must hold. Moreover, the co-state λ\lambda satisfies a necessary limit condition, which can be summarized as follows.

Proposition 5.

Let Assumptions 12, and 4 hold. Moreover, let (x,u)(x^{\star},u^{\star}) be a primal optimal solution of (1) converging to an optimal periodic orbit (xp,up)𝕎1,1nx×𝕃1nx(x_{\mathrm{p}},u_{\mathrm{p}})\in\mathbb{W}_{1,1}^{n_{x}}\times\mathbb{L}_{1}^{n_{x}} at which the periodic Riccati differential equation (3) admits a positive definite solution. Then, there exists a periodic function λp𝕎1,1nx\lambda_{\mathrm{p}}\in\mathbb{W}_{1,1}^{n_{x}} such that the associated co-state λ\lambda^{\star} necessarily satisfies

limtλ(t)λp(t)=0.\lim_{t\to\infty}\|\lambda^{\star}(t)-\lambda_{\mathrm{p}}(t)\|=0\;.

Notice that Proposition 5 is—at least in very similar versions—well-known in the literature [28, 21]. However, as this result is important for understanding the developments in this paper, the following proof briefly recalls the main argument, why this proposition holds.

Proof. Since (3) admits a positive definite solution, the value function VV in (1) is well-defined and differentiable in a neighborhood of the optimal orbit xx^{\star}. Thus, (1) can be replaced by an equivalent finite-horizon optimal control problem as long as VV is used as a terminal cost. Now, Pontryagin’s principle for optimal control problems with Mayer terms [28] yields the boundary condition

λ(t)=V(x(t))\lambda^{\star}(t)=\nabla V(x^{\star}(t))

for any horizon length t>0t>0. Clearly, since xx^{\star} converges to the optimal periodic orbit xpx_{\mathrm{p}}, λ\lambda^{\star} must converge to the associated dual limit orbit λp(t)=V(xp(t))\lambda_{\mathrm{p}}(t)=\nabla V(x_{\mathrm{p}}(t)). ∎

In the next sections we will see that Proposition 5 implies that the states of Pontryagin’s differential equation for stabilizable systems must evolve along certain stable manifolds. As it shall demonstrated, these manifolds can be characterized using a Koopman mode analysis [26].

3 Symplectic Koopman operators

This section analyzes the symplectic structure of the flow associated with Pontryagin’s differential equation. These symplectic structures are needed to understand the properties of the Pontryagin Koopman operators. In fact, Section 3.2 introduces a symplectic test space—that is, a space of observables—in which the Pontryagin-Koopman operator inherits certain symplecticity properties of its underlying incompressible flow field. Moreover, Section 3.3 leverages on ideas from the field of symplectic geometry [2] in order to work out the symplectic dual of the Pontryagin-Koopman operator. Notice that these developments are the basis for the developments in Section 4, which uses a symplectic spectral analysis in order to characterize globally optimal control laws.

3.1 Symplectic Flows

Let Γt:2nx2nx\Gamma_{t}:\mathbb{R}^{2n_{x}}\to\mathbb{R}^{2n_{x}} denote the flow of Pontryagin’s differential equations such that

ddtΓt(z)=F(Γt(z))andΓ0(z)=z\frac{\mathrm{d}}{\mathrm{d}t}\Gamma_{t}(z)=F(\Gamma_{t}(z))\quad\text{and}\quad\Gamma_{0}(z)=z

for all z2nxz\in\mathbb{R}^{2n_{x}}. If Assumptions 12, and 4 hold, then Γt\Gamma_{t} is a well-defined, continuously differentiable function. In the mathematical literature [2] it is well-known that Γt\Gamma_{t} is a so-called symplectic flow. In order to reveal this symplectic structure it is helpful to introduce the block matrix

Ω=(0II0)2nx×2nx.\Omega=\left(\begin{array}[]{rr}0&I\\ -I&0\end{array}\right)\in\mathbb{R}^{2n_{x}\times 2n_{x}}\;.

Now, the structural properties of Γt\Gamma_{t} can be summarized as follows.

Lemma 6.

Let Assumptions 12, and 4 be satisfied. Then, the function zΓt\frac{\partial}{\partial z}\Gamma_{t} satisfies the equation

t,[zΓt]Ω[zΓt]=Ω,\displaystyle\forall t\in\mathbb{R},\quad\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]=\Omega\;, (9)

that is, Γt\Gamma_{t} is symplectic function.

Proof. Let us first recall that the right-hand side FF of Pontryagin’s differential equation is given by (6) and (7), which can also be summarized as

F=[fHx]withy=[xλ].\displaystyle F=\left[\begin{array}[]{c}f\\ -H_{x}^{\intercal}\end{array}\right]\quad\text{with}\quad y=\left[\begin{array}[]{c}x\\ \lambda\end{array}\right]\;. (14)

This implies that the derivative of FF with respect yy can be written in the form

Fy=(fxfuHuu1HuxfuHuu1fuHxx+HxuHuu1Huxfx+HxuHuu1fu).\displaystyle F_{y}=\left(\begin{array}[]{cc}f_{x}-f_{u}H_{uu}^{-1}H_{ux}&-f_{u}H_{uu}^{-1}f_{u}^{\intercal}\\ -H_{xx}+H_{xu}H_{uu}^{-1}H_{ux}&-f_{x}^{\intercal}+H_{xu}H_{uu}^{-1}f_{u}^{\intercal}\end{array}\right). (17)

Here, we have used Assumptions 1 and 2 to ensure that the second derivatives of HH with respect to xx and uu exist. Moreover, we have used Assumption 4, which implies that HuuH_{uu} is invertible and, as a consequence, that the implicit function theorem holds. In particular, we have

ux=Huu1Huxanduλ=Huu1Huλ=Huu1fu.u_{x}^{\star}=-H_{uu}^{-1}H_{ux}\quad\text{and}\quad u_{\lambda}^{\star}=-H_{uu}^{-1}H_{u\lambda}=-H_{uu}^{-1}f_{u}^{\intercal}\;.

In this form, it becomes clear that the matrix ΩFy\Omega F_{y} is symmetric and we arrive at the intermediate result

ΩFy=FyΩ=FyΩ.\displaystyle\Omega F_{y}=F_{y}^{\intercal}\Omega^{\intercal}=-F_{y}^{\intercal}\Omega\;. (18)

In order to proceed, we write the first order variational differential equation for Γt\Gamma_{t} in the form

ddt[zΓt]=Fy[zΓt]with[zΓ0]=I.\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]=F_{y}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\quad\text{with}\quad\left[\frac{\partial}{\partial z}\Gamma_{0}\right]=I\;. (19)

Now, the main idea of this proof is to show that the function

Δ(t)=[zΓt]Ω[zΓt]Ω\displaystyle\Delta(t)=\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]-\Omega (20)

vanishes, Δ(t)=0\Delta(t)=0. Here, we first have Δ(0)=0\Delta(0)=0 by construction, since [zΓ0]=I\left[\frac{\partial}{\partial z}\Gamma_{0}\right]=I. Moreover, the derivative of Δ\Delta with respect to time is given by

Δ˙(t)\displaystyle\dot{\Delta}(t) =ddt[zΓt]Ω[zΓt]+[zΓt]Ωddt[zΓt]\displaystyle=\frac{\mathrm{d}}{\mathrm{d}t}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]+\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\frac{\mathrm{d}}{\mathrm{d}t}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]
=(19)[zΓt][FyΩ+ΩFy][zΓt]=(18)0\displaystyle\overset{\eqref{eq::SecondVariation}}{=}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\left[F_{y}^{\intercal}\Omega+\Omega F_{y}\right]\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\overset{\eqref{eq::FHamiltonian}}{=}0

Thus, in summary, we must have Δ(t)=0\Delta(t)=0 for all tt\in\mathbb{R}, which yields the statement of this lemma. ∎

The following corollary summarizes two immediate consequences of Lemma 6 which are both equivalent to stating that Γt\Gamma_{t} is an incompressible flow.

Corollary 7.

Let Assumptions 12, and 4 be satisfied. Then, Γt\Gamma_{t} is an incompressible flow; that is, we have

t,det([zΓt])=1.\displaystyle\forall t\in\mathbb{R},\qquad\mathrm{det}\left(\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\right)=1\;. (21)

Moreover, the divergence of the associated vector field FF vanishes, div(F)=F=0\mathrm{div}(F)=\nabla^{\intercal}F=0.

Proof. By taking the determinant on both sides of (9) and using that det(Ω)=1\mathrm{det}(\Omega)=1, we find that

det([zΓt])2=1det([zΓt])=±1.\mathrm{det}\left(\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\right)^{2}=1\quad\Longleftrightarrow\quad\mathrm{det}\left(\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\right)=\pm 1\;.

Since zΓ0=I\frac{\partial}{\partial z}\Gamma_{0}=I and since zΓt\frac{\partial}{\partial z}\Gamma_{t} is continuous, this is only possible if (21) holds. Next, by taking the logarithm on both sides of (21) and differentiating with respect to time, we find111Alternatively, the equation Tr(Fy)=0\mathrm{Tr}(F_{y})=0 can also directly be found by substituting the explicit expression (17) from the proof of Lemma 6.

0\displaystyle 0 =ddtlog(det([zΓt]))\displaystyle=\frac{\mathrm{d}}{\mathrm{d}t}\log\left(\mathrm{det}\left(\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\right)\right)
=Tr([zΓt]1ddt[zΓt])=(19)Tr(Fy)=div(F).\displaystyle=\mathrm{Tr}\left(\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{-1}\frac{\mathrm{d}}{\mathrm{d}t}\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\right)\overset{\eqref{eq::SecondVariation}}{=}\mathrm{Tr}(F_{y})=\mathrm{div}(F)\;.

This completes the proof of the corollary. ∎

3.2 Pontryagin-Koopman operators

The developments from the previous section can be used to analyze the structural properties of the Pontryagin-Koopman operator, which are defined to be the Koopman operator that is associated with the flow Γt\Gamma_{t} of Pontryagin’s differential equation, or, more formally:

Definition 8 (Pontryagin-Koopman Operator).

We use the notation Ut:𝕎1,22nx𝕎1,22nxU_{t}:\mathbb{W}_{1,2}^{2n_{x}}\to\mathbb{W}_{1,2}^{2n_{x}} to denote the Pontryagin Koopman operator of Γt\Gamma_{t}, which is defined such that

Φ𝕎1,22nx,t,UtΦ=ΦΓt\forall\Phi\in\mathbb{W}_{1,2}^{2n_{x}},\ \forall t\in\mathbb{R},\qquad U_{t}\Phi=\Phi\circ\Gamma_{t}

with \circ denoting the composition operator.

It is well known [19, 26] that UtU_{t} is a linear operator satisfying Ut1+t2=Ut1Ut2U_{t_{1}+t_{2}}=U_{t_{1}}U_{t_{2}} for all t1,t2t_{1},t_{2}\in\mathbb{R}. Moreover, UtU_{t} satisfies Ut1=UtU_{t}^{-1}=U_{-t}, which follows by substituting t=t1=t2t=t_{1}=-t_{2} in the previous equation and using that Γ0=id\Gamma_{0}=\mathrm{id}.

In order to introduce a notion of symplecticity in the space of observables of the Pontryagin-Koopman operator, we introduce the bilinear form

ω:𝕎1,22nx×𝕎1,22nx,\omega:\mathbb{W}_{1,2}^{2n_{x}}\times\mathbb{W}_{1,2}^{2n_{x}}\to\mathbb{R}\;,

which is defined as

ω(φ,Φ)=2nxφ(z)ΩΦ(z)dz\omega(\varphi,\Phi)=\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(z)^{\intercal}\Omega\nabla\Phi(z)\,\mathrm{d}z

for all φ,Φ𝕎1,22nx\varphi,\Phi\in\mathbb{W}_{1,2}^{2n_{x}}. Because we have Ω=Ω\Omega^{\intercal}=-\Omega, the bilinear form ω\omega is skew-symmetric,

ω(φ,Φ)=ω(Φ,φ),\omega(\varphi,\Phi)=-\omega(\Phi,\varphi),

and (𝕎1,22nx,ω)\left(\mathbb{W}_{1,2}^{2n_{x}},\omega\right) is a symplectic space [2].

Theorem 9.

Let Assumptions 12, and 4 be satisfied. Now, UtU_{t} is for all tt\in\mathbb{R} a symplectic operator in the space (𝕎1,22nx,ω)\left(\mathbb{W}_{1,2}^{2n_{x}},\omega\right). This means that we have

ω(Utφ,UtΦ)=ω(φ,Φ)\omega(U_{t}\varphi,U_{t}\Phi)=\omega(\varphi,\Phi)

for all φ,Φ𝕎1,22nx\varphi,\Phi\in\mathbb{W}_{1,2}^{2n_{x}}.

Proof. Assumptions 12, and 4 imply that the function Γt\Gamma_{t} and its inverse Γt1=Γt\Gamma_{t}^{-1}=\Gamma_{-t} are both continuously differentiable. This implies in particular that the equivalence

Φ𝕎1,22nx[ΦΓt]𝕎1,22nx\Phi\in\mathbb{W}_{1,2}^{2n_{x}}\quad\Longleftrightarrow\quad\left[\Phi\circ\Gamma_{t}\right]\in\mathbb{W}_{1,2}^{2n_{x}}

holds. Next, the definition of the operator UtU_{t} and the chain rule for differentiation imply that

[UtΦ]=[ΦΓt]=[zΓt](Φ)Γt\displaystyle\nabla[U_{t}\Phi]=\nabla[\Phi\circ\Gamma_{t}]=\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}(\nabla\Phi)\circ\Gamma_{t} (22)

for all Φ𝕎1,22nx\Phi\in\mathbb{W}_{1,2}^{2n_{x}}. Furthermore, because Assumptions 12, and 4 hold, it follows from Lemma 6 that

[zΓt]Ω[zΓt]=Ω.\displaystyle\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]=\Omega\;.

Let us multiply this equation with [zΓt]Ω\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\Omega from the left and substitute Ω2=I\Omega^{2}=-I. This yields

[zΓt]Ω[zΓt]Ω[zΓt]=[zΓt].\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]=-\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\;.

Next, we multiply the latter equation with [zΓt]1Ω\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{-1}\Omega from the right and use the relation Ω2=I\Omega^{2}=-I once more, in order to arrive at the equation

[zΓt]Ω[zΓt]=Ω.\displaystyle\left[\frac{\partial}{\partial z}\Gamma_{t}\right]\Omega\left[\frac{\partial}{\partial z}\Gamma_{t}\right]^{\intercal}=\Omega\;. (23)

Thus, by substituting the previous equations, we get

ω(Utφ,UtΦ)=2nx[Utφ](z)Ω[UtΦ](z)dz=(22),(23)2nxφ(Γt(z))ΩΦ(Γt(z))dz.\omega(U_{t}\varphi,U_{t}\Phi)=\int_{\mathbb{R}^{2n_{x}}}\nabla[U_{t}\varphi](z)^{\intercal}\Omega\nabla[U_{t}\Phi](z)\,\mathrm{d}z\overset{\eqref{eq::aux2},\eqref{eq::aux3}}{=}\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(\Gamma_{t}(z))^{\intercal}\Omega\nabla\Phi(\Gamma_{t}(z))\,\mathrm{d}z\,.

In order to further transform this integral, we need to introduce the change of variables z=Γt(z)z^{\prime}=\Gamma_{t}(z). Because Corollary 7 ensures that

|det(zΓt(z))|=1,\left|\mathrm{det}\left(\frac{\partial}{\partial z}\Gamma_{t}(z)\right)\right|=1\;,

this change of variables is volume preserving. Thus, this implies that

ω(Utφ,UtΦ)=2nxφ(z)ΩΦ(z)dz=ω(φ,Φ)\omega(U_{t}\varphi,U_{t}\Phi)=\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(z^{\prime})^{\intercal}\Omega\nabla\Phi(z^{\prime})\,\mathrm{d}z^{\prime}=\omega(\varphi,\Phi)

for all φ,Φ𝕎1,22nx\varphi,\Phi\in\mathbb{W}_{1,2}^{2n_{x}}. Thus, UtU_{t} is a symplectic operator. ∎

3.3 Symplectic duality

In order to prepare the analysis of the properties of the spectrum of symplectic Koopman operators, it is helpful to introduce a notion of duality. However, instead of defining duality in a Hilbert space, we propose to introduce the following notion of symplectic duality.

Definition 10.

Let A:𝕎1,22nx𝕎1,22nxA:\mathbb{W}_{1,2}^{2n_{x}}\to\mathbb{W}_{1,2}^{2n_{x}} be a linear operator. If there exists a linear operator A:𝕎1,22nx𝕎1,22nxA^{\star}:\mathbb{W}_{1,2}^{2n_{x}}\to\mathbb{W}_{1,2}^{2n_{x}} with

ω(φ,AΦ)=ω(Aφ,Φ)\omega(\varphi,A\Phi)=\omega(A^{\star}\varphi,\Phi)

for all φ,Φ𝕎1,22nx\varphi,\Phi\in\mathbb{W}_{1,2}^{2n_{x}}, then we call AA^{\star} the symplectic adjoint operator of AA.

In the following, we consider a linear differential operator L:𝕎1,22nx𝕎1,22nxL:\mathbb{W}_{1,2}^{2n_{x}}\to\mathbb{W}_{1,2}^{2n_{x}} that is given by

LΦ=FΦ\displaystyle L\Phi=F^{\intercal}\nabla\Phi (24)

for all Φ𝕎1,22nx\Phi\in\mathbb{W}_{1,2}^{2n_{x}}. Notice that this operator is found by differentiating the definition of the Pontryagin-Koopman operator UtU_{t} with respect to time, such that L=U˙tL=\dot{U}_{t}.

Lemma 11.

Let Assumptions 12, and 4 hold. The operator LL admits a symplectic adjoint LL^{\star} in the symplectic space (𝕎2,22nx,ω)(\mathbb{W}_{2,2}^{2n_{x}},\omega). This adjoint is given by

L=L.\displaystyle L^{\star}=-L\;. (25)

Proof. For any function Φ𝕎2,22nx\Phi\in\mathbb{W}_{2,2}^{2n_{x}}, we have

(LΦ)=dFdyΦ+2ΦF,\displaystyle\nabla(L\Phi)=\frac{\mathrm{d}F}{\mathrm{d}y}\nabla\Phi+\nabla^{2}\Phi F\;, (26)

which follows by differentiating (24) on both sides. Next, we know from Theorem 9 that UtU_{t} is a symplectic operator. Thus, we can differentiate the equation

ω(Utφ,UtΦ)=ω(φ,Φ)\omega(U_{t}\varphi,U_{t}\Phi)=\omega(\varphi,\Phi)

on both sides with respect to tt, which yields

ddtω(Utφ,UtΦ)=0.\frac{\mathrm{d}}{\mathrm{d}t}\omega(U_{t}\varphi,U_{t}\Phi)=0\;.

Now, if φ,Φ𝕎2,22nx\varphi,\Phi\in\mathbb{W}_{2,2}^{2n_{x}}, we can expand the derivative on the right as

0=ddtω(Utφ,UtΦ)\displaystyle 0=\frac{\mathrm{d}}{\mathrm{d}t}\,\omega(U_{t}\varphi,U_{t}\Phi) =(3.2)ddt2nxφ(Γt(z))ΩΦ(Γt(z))dz\displaystyle\overset{\eqref{eq::aux4}}{=}\frac{\mathrm{d}}{\mathrm{d}t}\,\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(\Gamma_{t}(z))^{\intercal}\Omega\nabla\Phi(\Gamma_{t}(z))\,\mathrm{d}z
=2nxF(Γt(z))2φ(Γt(z))ΩΦ(Γt(z))dz+2nxφ(Γt(z))Ω2Φ(Γt(z))F(Γt(z))dz\displaystyle=\int_{\mathbb{R}^{2n_{x}}}F(\Gamma_{t}(z))^{\intercal}\nabla^{2}\varphi(\Gamma_{t}(z))\Omega\nabla\Phi(\Gamma_{t}(z))\,\mathrm{d}z+\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(\Gamma_{t}(z))^{\intercal}\Omega\nabla^{2}\Phi(\Gamma_{t}(z))F(\Gamma_{t}(z))\,\mathrm{d}z

and, after changing variables, zΓt(z)z\leftarrow\Gamma_{t}(z), recalling that det(zΓt)=1\mathrm{det}\left(\frac{\partial}{\partial z}\Gamma_{t}\right)=1, and resorting terms,

2nxF(z)2φ(z)ΩΦ(z)dz=2nxφ(z)Ω2Φ(z)F(z)dz.\int_{\mathbb{R}^{2n_{x}}}F(z)^{\intercal}\nabla^{2}\varphi(z)\Omega\nabla\Phi(z)\,\mathrm{d}z=-\int_{\mathbb{R}^{2n_{x}}}\nabla\varphi(z)^{\intercal}\Omega\nabla^{2}\Phi(z)F(z)\,\mathrm{d}z\;. (27)

Moreover, we recall that the equation

ΩdFdy=(dFdy)Ω=(dFdy)Ω\displaystyle\Omega\frac{\mathrm{d}F}{\mathrm{d}y}=\left(\frac{\mathrm{d}F}{\mathrm{d}y}\right)^{\intercal}\Omega^{\intercal}=-\left(\frac{\mathrm{d}F}{\mathrm{d}y}\right)^{\intercal}\Omega (28)

also holds (see Equation (18)). Thus, in summary, we have

ω(φ,LΦ)=2nx[φΩ[LΦ]](z)dz\displaystyle\omega(\varphi,L\Phi)=\displaystyle\int_{\mathbb{R}^{2n_{x}}}\left[\nabla\varphi^{\intercal}\Omega\nabla[L\Phi]\right](z)\,\mathrm{d}z =(26)2nx[φΩ(dFdxΦ+2ΦF)](z)dz\displaystyle\quad\overset{\eqref{eq::NablaL}}{=}\qquad\displaystyle\int_{\mathbb{R}^{2n_{x}}}\left[\nabla\varphi^{\intercal}\Omega\left(\frac{\mathrm{d}F}{\mathrm{d}x}\nabla\Phi+\nabla^{2}\Phi F\right)\right](z)\,\mathrm{d}z
=(27),(28)2nx[(dFdxφ+2φF)ΩΦ](z)dz\displaystyle\overset{\eqref{eq::aux5},\eqref{eq::aux6}}{=}-\displaystyle\int_{\mathbb{R}^{2n_{x}}}\left[\left(\frac{\mathrm{d}F}{\mathrm{d}x}\nabla\varphi+\nabla^{2}\varphi F\right)^{\intercal}\Omega\nabla\Phi\right](z)\,\mathrm{d}z
=ω(Lφ,Φ).\displaystyle\quad\,=\qquad\omega(-L\varphi,\Phi)\;.

Because the latter equation holds for all φ,Φ𝕎2,22nx\varphi,\Phi\in\mathbb{W}_{2,2}^{2n_{x}}, we find that L=LL^{\star}=-L is the symplectic adjoint of LL, as claimed by the statement of this lemma. ∎

Remark 12.

In many articles on Koopman operators, a so-called transport differential equation of the form

tϕ=div(Fϕ)\frac{\partial}{\partial t}\phi=-\mathrm{div}(F\phi)

is considered. If one introduces suitable boundary conditions, this advection PDE can be interpreted as the derivative of a Perron-Frobenius operator that is dual to the Koopman operator UtU_{t} with respect to the standard 𝕃2\mathbb{L}_{2}-scalar product [10]. Notice that in our case, the symplectic adjoint operator

L:ϕFϕ=div(Fϕ)+div(F)=0ϕL^{\star}:\;\phi\,\to\,-F^{\intercal}\nabla\phi=-\mathrm{div}(F\phi)+\underbrace{\mathrm{div}(F)}_{=0}\phi

happens to coincide with the right-hand operator of the above advection PDE. Thus, physically, one could interpret the operator LL^{\star} as an advection operator of the incompressible flow field FF recalling that div(F)=0\mathrm{div}(F)=0.

4 Spectral Analysis

In this section, we show that the symplectic structures of the Koopman operator, as analyzed in the previous section, have important consequences on its set of eigenvalues. Moreover, Section 5 uses these spectral properties of the Koopman operator to characterize global optimal control laws that are associated with the infinite horizon optimal control problem (1).

4.1 Eigenfunctions and Eigenvalues

As already mentioned in the introduction, the spectrum of general Koopman operators has been analyzed by many authors; for example in [1, 6, 26]. In the context of this paper, we call a weakly differentiable function Ψ\Psi an eigenfunction of the differential Pontryagin-Koopman operator LL, if the Lebesgue measure of the set supp(Ψ)\,\mathrm{supp}(\Psi)\, is either unbounded, or, if it exists, is not equal to 0, and

LΨ=κΨΨF=κΨL\Psi=\kappa\Psi\qquad\Longleftrightarrow\qquad\nabla\Psi^{\intercal}F=\kappa\Psi

for a potentially complex eigenvalue κ\kappa\in\mathbb{C}. The fact that the symplectic adjoint of the operator LL is given by L=LL^{\star}=-L has important consequences on its spectrum, which can be summarized as follows.

Theorem 13.

Let Assumptions 12, and 4 hold and let Ψ1,Ψ2𝕎2,22nx\Psi_{1},\Psi_{2}\in\mathbb{W}_{2,2}^{2n_{x}} be two eigenfunctions of LL with eigenvalues κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{C}. If ω(Ψ1,Ψ2)0\omega(\Psi_{1},\Psi_{2})\neq 0, then we must have κ1=κ2\kappa_{1}=-\kappa_{2}.

Proof. Because the functions Ψ1,Ψ2𝕎2,22nx\Psi_{1},\Psi_{2}\in\mathbb{W}_{2,2}^{2n_{x}} are eigenfunctions of LL, they satisfy

LΨ1=κ1Ψ1andLΨ2=κ2Ψ2.\displaystyle L\Psi_{1}=\kappa_{1}\Psi_{1}\quad\text{and}\quad L\Psi_{2}=\kappa_{2}\Psi_{2}\;. (29)

Thus, since ω\omega is a bilinear form, we have

κ2ω(Ψ1,Ψ2)=(29)ω(Ψ1,LΨ2)=(25)ω(LΨ1,Ψ2)=(29)κ1ω(Ψ1,Ψ2)\kappa_{2}\omega(\Psi_{1},\Psi_{2})\overset{\eqref{eq::aux7}}{=}\omega(\Psi_{1},L\Psi_{2})\overset{\eqref{eq::LL}}{=}-\omega(L\Psi_{1},\Psi_{2})\overset{\eqref{eq::aux7}}{=}-\kappa_{1}\omega(\Psi_{1},\Psi_{2})

and, after resorting terms,

(κ1+κ2)ω(Ψ1,Ψ2)=0.(\kappa_{1}+\kappa_{2})\omega(\Psi_{1},\Psi_{2})=0\;.

Thus, if ω(Ψ1,Ψ2)0\omega(\Psi_{1},\Psi_{2})\neq 0, we must have κ1+κ2=0\kappa_{1}+\kappa_{2}=0. This is equivalent to the statement of the theorem. ∎

Remark 14.

The statement of Theorem 13 is formally not directly applicable for eigenfunctions Ψ1,Ψ2\Psi_{1},\Psi_{2} of LL that are locally weakly twice differentiable but whose derivatives are not square integrable, such that Ψ1,Ψ2\Psi_{1},\Psi_{2} are not elements of the Sobolev space 𝕎2,22nx\mathbb{W}_{2,2}^{2n_{x}}. However, in practice, one is usually interested in constructing eigenfunctions of LL on a compact domain C2nxC\subset\mathbb{R}^{2n_{x}}, in the following called the region of interest, such that

xC,[LΨ](x)=κΨ(x).\displaystyle\forall x\in C,\quad[L\Psi](x)=\kappa\Psi(x)\;. (30)

This region of interest CC can, for example, model a-priori bounds on the optimal primal and dual trajectories of (1). Consequently, because we are simply not interested in how the flow Γt\Gamma_{t} and the associated eigenfunctions Ψ\Psi are defined outside of the set CC, these functions can simply be redefined arbitrarily for xCx\notin C, for example, such that the desired eigenfunctions of LL satisfy Ψ𝕎2,22nx\Psi\in\mathbb{W}_{2,2}^{2n_{x}} by construction. Thus, for the purpose of this paper, it is not restrictive at all to assume that the derivatives of the eigenfunctions of LL are square integrable. Notice that this technique is also illustrated by our tutorial case study in Section 6, where we explain how to choose CC and how to discretize (30) on CC.

Let σ(L)\sigma(L)\subseteq\mathbb{C} denote the spectrum of LL; that is, the set of eigenvalues of the linear operator LL. In the following, we use the notation Ψκ𝕎2,22nx\Psi_{\kappa}\in\mathbb{W}_{2,2}^{2n_{x}} to denote an eigenfunction that is associated with an eigenvalue κσ(L)\kappa\in\sigma(L). Moreover, for a function q:2nx2nxq:\mathbb{R}^{2n_{x}}\to\mathbb{R}^{2n_{x}}, we use the shorthand notation

ω(q,q)=(ω(q1,q1)ω(q1,q2)ω(q1,q2nx)ω(q2,q1)ω(q2,q2)ω(q2,q2nx)ω(q2nx,q1)ω(q2nx,q2)ω(q2nx,q2nx))\omega(q,q)=\left(\begin{array}[]{cccc}\omega(q_{1},q_{1})&\omega(q_{1},q_{2})&\ldots&\omega(q_{1},q_{2n_{x}})\\ \omega(q_{2},q_{1})&\omega(q_{2},q_{2})&\ldots&\omega(q_{2},q_{2n_{x}})\\ \vdots&\vdots&\ddots&\vdots\\ \omega(q_{2n_{x}},q_{1})&\omega(q_{2n_{x}},q_{2})&\ldots&\omega(q_{2n_{x}},q_{2n_{x}})\\ \end{array}\right)

to denote the matrix that is obtained by evaluating the skew symmetric bilinear form ω\omega for all possible combinations of the components of qq. We call q\,q\, a skew-orthogonal function in the symplectic space (𝕎1,22nx,ω)(\mathbb{W}_{1,2}^{2n_{x}},\omega), if it satisfies ω(q,q)=I\omega(q,q)=I. Notice that the construction of such skew orthogonal functions is straightforward by using the standard skew-symmetric variant of the Gram-Schmidt algorithm [2].

Definition 15.

The operator LL is said to admit a spectral decomposition with respect to a given function qq, if there exists a generalized function a:σ(L)¯2nxa:\sigma(L)\to\overline{\mathbb{C}}^{2n_{x}} such that

q=σ(L)Ψκa(κ)dκ.\displaystyle q=\int_{\sigma(L)}\Psi_{\kappa}a(\kappa)\,\mathrm{d}\kappa\;. (31)

Notice that the above definition uses the “control engineering notation” for generalized functions, which means that we use aa as if it was a standard function, although this notation suppresses the distributional nature of aa. Thus, in mathematical terms, this notation has to be translated as “aa represents a distribution of order 0; that is, a linear operator on σ(L)\sigma(L) that is Lipschitz continuous with respect to the LL_{\infty}-norm”.222Notice that, for example, aa could be a Dirac distribution, which is not a function in the traditional sense but, by construction, a Lipschitz continuous linear operator. The following statement is a consequence of Theorem 13.

Corollary 16.

Let Assumptions 12, and 4 hold. Let the operator LL admit a spectral decomposition with respect to a skew-orthogonal function qq. Then, there exist at least 2nx2n_{x} eigenvalues

κ1+,κ2+,,κnx+,κ1,κ2,,κnxσ(L),\kappa_{1}^{+},\kappa_{2}^{+},\ldots,\kappa_{n_{x}}^{+},\kappa_{1}^{-},\kappa_{2}^{-},\ldots,\kappa_{n_{x}}^{-}\in\sigma(L)\;,

such that κi+=κi\kappa_{i}^{+}=-\kappa_{i}^{-}, where κi+\kappa_{i}^{+} has a non-negative real part for all i{1,2,,nx}i\in\{1,2,\ldots,n_{x}\}.

Proof. By substituting (31) in the equation ω(q,q)=I\omega(q,q)=I, we find that the equation

I=σ(L)σ(L)ω(Ψκ,Ψκ)a(κ)a(κ)rank 1dκdκ,I=\int_{\sigma(L)}\int_{\sigma(L)}\omega(\Psi_{\kappa^{\prime}},\Psi_{\kappa})\underbrace{a(\kappa^{\prime})a(\kappa)^{\intercal}}_{\mathrm{rank}\;1}\,\mathrm{d}\kappa\,\mathrm{d}\kappa^{\prime}\;, (32)

holds. Let us have a closer look at the terms in (32). Clearly, the unit matrix on the left has full rank. But, on the other side, we have an integral over the rank 11 matrices a(κ)a(κ)a(\kappa^{\prime})a(\kappa)^{\intercal}. This integral term can only have full rank, if there are at least 2nx2n_{x} pairs of eigenvalues (κ,κ)σ(L)×σ(L)(\kappa^{\prime},\kappa)\in\sigma(L)\times\sigma(L) for which

ω(Ψκ,Ψκ)0.\omega(\Psi_{\kappa^{\prime}},\Psi_{\kappa})\neq 0\;.

But now Theorem 13 implies that all these pairs must be such that κ=κ\kappa^{\prime}=-\kappa and, after sorting all eigenvalues with respect to their real-part, we find that there must be at least nxn_{x} eigenvalues with non-negative real part and nxn_{x} associated mirrored eigenvalues with non-positive real-part, as claimed by the statement of this corollary. ∎

Remark 17.

Notice that Corollary 16 makes a statement about the spectrum of LL under the assumption that this operator admits a spectral decomposition with respect to at least one skew orthogonal function. Thus, one should further ask the question under which assumptions the existence of such a spectral decomposition can be guaranteed for at least one such skew orthogonal function. In full generality, this question is difficult to answer, but sufficient conditions for the existence of (much more general and, in our context, sufficient) spectral decompositions can be found in [1, 6, 25, 26], which use ideas from the field of ergodic theory [33] as well as Yoshida’s theorem [34]. For example, if the monodromy matrix, zΓT(xp(0))\frac{\partial}{\partial z}\Gamma_{T}(x_{\mathrm{p}}(0)), of a periodic optimal orbit with period length T>0T>0 is diagonalizable, one can ensure that a spectral decomposition is possible. This sufficient condition follows by applying Proposition 3 in [25] to the Pontryagin differential equation. A more complete review of such results from the field of functional analysis would, however, go beyond the scope of the present paper.

5 Optimal Feedback Control Laws

The theoretical results from the previous sections can be used to derive optimality conditions, which, in turn, can be used to develop practical numerical algorithms for solving (1). Let us introduce the set

σ+(L)\displaystyle\sigma^{+}(L) ={κσ(L)Re(κ)>0}\displaystyle=\{\kappa\in\sigma(L)\mid\mathrm{Re}(\kappa)>0\}

of unstable eigenvalues and its associated invariant manifold

+\displaystyle\mathcal{M}^{+} ={[x,λ]2nx|κσ+(L),Ψκ(x,λ)=0}.\displaystyle=\left\{[x^{\intercal},\lambda^{\intercal}]^{\intercal}\in\mathbb{R}^{2n_{x}}\;\middle|\;\begin{array}[]{l}\forall\kappa\in\sigma^{+}(L),\\ \Psi_{\kappa}(x,\lambda)=0\end{array}\right\}\;. (35)

Because we assume that the eigenfunctions Ψκ\Psi_{\kappa} are weakly differentiable, we may assume without loss of generality that the functions Ψκ\Psi_{\kappa} are also continuous—otherwise there exists a continuous function Ψ~κ(x)=Ψκ(x)\tilde{\Psi}_{\kappa}(x)=\Psi_{\kappa}(x) for almost every x2nxx\in\mathbb{R}^{2n_{x}} and we can use Ψ~κ\tilde{\Psi}_{\kappa} instead of Ψκ\Psi_{\kappa} recalling that we work with Sobolev spaces, in which such arguments are indeed possible.

In the following, we say that μ\mu^{\star} is a regular optimal control law of (1), if the closed-loop trajectories,

x˙(t)=f(x(t),μ(x(t)))withx(0)=x0,\dot{x}^{\star}(t)=f(x^{\star}(t),\mu^{\star}(x^{\star}(t)))\quad\text{with}\quad x^{\star}(0)=x_{0}\;,

are minimizers of (1) at which Pontryagin’s necessary optimality conditions are satisfied such that xx^{\star} converges to an optimal periodic orbit at which the conditions of Proposition 5 are satisfied.

Theorem 18.

Let Assumptions 12, and 4 hold. We further assume (without loss of generality) that the eigenfunctions ΨκW2,22nx\Psi_{\kappa}\in W_{2,2}^{2n_{x}} are continuous. Let μ:nxnx\mu^{\star}:\mathbb{R}^{n_{x}}\to\mathbb{R}^{n_{x}} be a regular optimal control law of (1). Then, there exists a function Λ:nxnx\Lambda:\mathbb{R}^{n_{x}}\to\mathbb{R}^{n_{x}} such that

[x,Λ(x)]+andμ(x)=u(x,Λ(x)).[x^{\intercal},\Lambda(x)^{\intercal}]^{\intercal}\in\mathcal{M}^{+}\quad\text{and}\quad\mu^{\star}(x)=u^{\star}(x,\Lambda(x))\;.

Proof. Since LL is a time-autonomous infinitesimal generator of UtU_{t}, the eigenfunctions of LL also satisfy [26]

UtΨκ=eκtΨκ\displaystyle U_{t}\Psi_{\kappa}=e^{\kappa t}\Psi_{\kappa} (36)

for all κσ(L)\kappa\in\sigma(L). The remainder of the proof is divided into two parts. In the first part, we show that (36) implies that the optimal periodic limit orbit is a subset of the manifold +\mathcal{M}^{+} and, in the second part, we use this property of the limit orbit to construct a globally optimal feedback law.

Part I: Let (xp,λp)(x_{\mathrm{p}},\lambda_{\mathrm{p}}) denote the optimal periodic limit orbit of (1) with period time TT. If we would have (xp(t),λp(t))+(x_{\mathrm{p}}(t),\lambda_{\mathrm{p}}(t))\notin\mathcal{M}^{+} for a time t[0,T]t\in[0,T], we must have

Ψκ(xp(t),λp(t))0\Psi_{\kappa}(x_{\mathrm{p}}(t),\lambda_{\mathrm{p}}(t))\neq 0

for at least one κσ+(L)\kappa\in\sigma^{+}(L). Because the optimal periodic orbit satisfies Pontryagin’s differential equation, this implies in particular that

|Ψκ(xp(t+T),λp(t+T))|=(36)|eκTΨκ(xp(t),λp(t))|>|Ψκ(xp(t),λp(t))|\left|\Psi_{\kappa}(x_{\mathrm{p}}(t+T),\lambda_{\mathrm{p}}(t+T))\right|\overset{\eqref{eq::Ut}}{=}\left|e^{\kappa T}\Psi_{\kappa}(x_{\mathrm{p}}(t),\lambda_{\mathrm{p}}(t))\right|>\left|\Psi_{\kappa}(x_{\mathrm{p}}(t),\lambda_{\mathrm{p}}(t))\right|\,

as κ\kappa has a strictly positive real part. But this is a contradiction, since (xp,λp)(x_{\mathrm{p}},\lambda_{\mathrm{p}}) is periodic. Thus, in summary, we must have

t[0,T],(xp(t),λp(t))+.\displaystyle\forall t\in[0,T],\quad(x_{\mathrm{p}}(t),\lambda_{\mathrm{p}}(t))\in\mathcal{M}^{+}\;. (37)

Part II: Let (x,λ)(x,\lambda) denote an optimal solution of (1) with x(0)=x0x(0)=x_{0}. Now, if we would have (x(0),λ(0))+(x(0),\lambda(0))\notin\mathcal{M}^{+}, then there would have to exist at least one κσ+(L)\kappa\in\sigma^{+}(L) for which

Ψκ(x(0),λ(0))0.\Psi_{\kappa}(x(0),\lambda(0))\neq 0\;.

But if this would be the case, then we would also have

limt|Ψκ(x(t),λ(t))|=,\displaystyle\lim_{t\to\infty}|\Psi_{\kappa}(x(t),\lambda(t))|=\infty\;, (38)

as Ψκ\Psi_{\kappa} is strictly unstable. But this limit statement is in conflict with (37), since Ψκ\Psi_{\kappa} is continuous and we assume that (x(t),λ(t))(x(t),\lambda(t)) converges to the optimal periodic limit orbit. Thus, in summary, there exists for every x0x_{0} a λ0=λ(0)\lambda_{0}=\lambda(0) with (x0,λ0)+(x_{0},\lambda_{0})\in\mathcal{M}^{+} and the corresponding map from x0x_{0} to λ0\lambda_{0} can be denoted by Λ\Lambda. The associated optimal control input is given by

μ(x0)=u(x0,Λ(x0)).\mu^{\star}(x_{0})=u^{\star}(x_{0},\Lambda(x_{0}))\;.

This is already sufficient to establish the statement of this theorem, as the optimal feedback law must be time-autonomous.∎

Notice that Theorem 18 can be used to systematically search for globally optimal solutions of (1). Here, one of the key observations is that if, in addition to the assumptions of Theorem 18, the assumptions of Corollary 16 are also satisfied and if none of the non-negative eigenvalues κ1+,κ2+,,κnx+\kappa_{1}^{+},\kappa_{2}^{+},\ldots,\kappa_{n_{x}}^{+} happens to be on the imaginary axis, then the parametric equation system

κσ+(L),Ψκ(x0,λ0)=0.\displaystyle\forall\kappa\in\sigma^{+}(L),\quad\Psi_{\kappa}(x_{0},\lambda_{0})=0\;. (39)

consists of at least nxn_{x} independent equations while the number of variables, dim(λ)=nx\mathrm{dim}(\lambda)=n_{x}, is equal to nxn_{x}, too. Thus, this equation system can, in many practical instances, be expected to admit a finite number of parametric solutions λ0=Λ(x0)\lambda_{0}=\Lambda(x_{0}) only. This means that, if all the above regularity assumptions are satisfied, Theorem 18 singles out a finite number of candidate control laws μ(x)=u(x,Λ(x))\mu^{\star}(x)=u^{\star}(x,\Lambda(x))—at least one of which must be globally optimal.

6 Numerical example

This section illustrates how Theorem 18 can be used to construct accurate approximations of globally optimal control laws of (1) for a Van der Pol oscillator system with

f(x,u)\displaystyle f(x,u) =(x2x112(1x12)x2+x1u)\displaystyle=\left(\begin{array}[]{l}x_{2}\\ x_{1}-\frac{1}{2}\left(1-x_{1}^{2}\right)x_{2}+x_{1}u\end{array}\right)\quad (42)
andl(x,u)\displaystyle\text{and}\quad l(x,u) =12(x22+u2).\displaystyle=\frac{1}{2}\left(x_{2}^{2}+u^{2}\right)\;. (43)

The associated system, x˙(t)=f(x(t),u(t))\dot{x}(t)=f(x(t),u(t)), has a linearly uncontrollable equilibrium at x=(0,0)x=(0,0)^{\intercal}. Notice that for this particular example an explicit solution of the Hamilton-Jacobi-Bellman equation is known [30]: it turns out that the optimal value function VV and the globally optimal feedback law μ\mu^{\star} are given, respectively, by

V(x)=12(x12+x22)andμ(x)=x1x2.\displaystyle V(x)=\frac{1}{2}(x_{1}^{2}+x_{2}^{2})\quad\text{and}\quad\mu^{\star}(x)=-x_{1}x_{2}\;. (44)

In the following, this explicit solution is, however, only used to assess the accuracy of the proposed method; that is, the numerical procedure below neither knows the above expression for VV nor for μ\mu^{\star}.

6.1 Galerkin discretization

In order to illustrate the practical applicability of the developments in Section 5, we introduce a simple Galerkin discretization of the operator LL. Let φ1,,φN𝕎1,22nx\varphi_{1},\ldots,\varphi_{N}\in\mathbb{W}_{1,2}^{2n_{x}} be orthogonal functions with respect to the standard 𝕃2\mathbb{L}_{2}-scalar product ,\langle\cdot,\cdot\rangle on 𝕎1,22nx\mathbb{W}_{1,2}^{2n_{x}}. Next, if we compute the coefficients

Mi,j=Lφi,φj\displaystyle M_{i,j}=\langle L\varphi_{i},\varphi_{j}\rangle\;

for all i,j{1,,N}i,j\in\{1,\ldots,N\}, the matrix MM^{\intercal} can be interpreted as a Galerkin approximation of the operator LL over the subspace spanned by φ1,,φN\varphi_{1},\ldots,\varphi_{N}. In our implementation, we set φi\varphi_{i} to the iith multivariate Legendre polynomial on the 44-dimensional compact interval box C=[12,12]4C=[-\frac{1}{2},\frac{1}{2}]^{4} and we set φi(x,λ)=0\varphi_{i}(x,\lambda)=0 outside of this domain; that is, for (x,λ)C(x,\lambda)\notin C. Consequently, our discretization can only be expected to be accurate inside our region of interest CC, but other choices for CC and for the basis functions would be possible, too.

Remark 19.

Standard Galerkin methods are, in general, numerically unstable when applied to advection operators and, consequently, although this method happens to yield reasonable approximations for our particular example, such naive discretization schemes cannot be recommended in general. More advanced discretization schemes for linear advection operators can be found in the modern PDE literature; see, for example [31]. A more complete discussion of such numerical discretization methods for Pontryagin-Koopman operators, is, however, beyond the scope of this paper.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Left: Comparison between the optimal feedback law μ(x)=x1x2\mu^{\star}(x)=-x_{1}x_{2}, and the approximate feedback law μ~(x)\widetilde{\mu}(x) in (46). Middle: trajectories of the closed-loop system under the approximate feedback law. Right: Eigenvalues of the projection MN×NM\in\mathbb{R}^{N\times N} for N=4N=4 (blue circles), N=15N=15 (red squares), and N=35N=35 (black circles with a cross).

6.2 Approximations of the optimal feedback law

The above Galerkin approximation of the operator LL can be used to construct approximate eigenfunctions. In detail, if aNa\in\mathbb{R}^{N} is a left eigenvector of MM with eigenvalue κ\kappa\in\mathbb{C}, then

Ψ=i=1NaiφiLΨκΨ\Psi=\sum_{i=1}^{N}a_{i}\varphi_{i}\quad\Longrightarrow\quad L\Psi\approx\kappa\Psi

is an approximation of an eigenfunction of LL. The right plot in Figure 1 shows the spectrum of the matrix MM for different choices of NN (blue circles: N=4N=4, red squares: N=15N=15, and black circles with a cross: N=35N=35). In order to understand the structure of this spectrum it is helpful to recall Corollary 16, which predicts that there exits at least 22 eigenvalues κ1,κ2\kappa_{1},\kappa_{2} of LL such that κ1-\kappa_{1} and κ2-\kappa_{2} are also eigenvalues of LL. Notice that such symmetric eigenvalue pairs are indeed present in the spectrum of MM, although MM is only a Galerkin approximation of LL.

In order to further illustrate how the above spectral analysis of MM can be used to construct approximations of the globally optimal control law μ\mu^{\star}, we can compute an approximation of the manifold +\mathcal{M}^{+} by using the approximate eigenfunctions instead of the exact ones (see Theorem 18). For example, for N=4N=4, the matrix MM has the eigenvalues

κ1,2±=148(βm±βp1)withβp=+983+96143andβm=983+96143\kappa^{\pm}_{1,2}=\frac{1}{48}(\beta_{m}\pm\beta_{p}\sqrt{-1})\quad\text{with}\quad\beta_{p}=\sqrt{+983+96\sqrt{143}}\quad\text{and}\quad\beta_{m}=\sqrt{-983+96\sqrt{143}} (45)

and the associated Galerkin approximation of the globally optimal control law is given by

μ~(x)=(12143)x12+(11βm)2x1x2.\widetilde{\mu}(x)=(12-\sqrt{143})x_{1}^{2}+\frac{(11-\beta_{m})}{2}x_{1}x_{2}\;. (46)

The left plot in Figure 1 shows μ~\widetilde{\mu} and compares it to the optimal feedback law μ\mu^{\star}. In fact, the squared integral error over CC is approximately 6×1056\times 10^{-5}. Quite remarkably, this approximate feedback law can even be used to control the system for initial values outside of CC. The plot in the middle of Figure 1 shows the corresponding trajectories of the closed-loop system that are obtained by using the approximately optimal feedback law μ~\widetilde{\mu}.

7 Conclusions

This paper has presented an analysis of infinite horizon nonlinear optimal control problems, whose minimizers satisfy Pontryagin’s necessary conditions of optimality. The proposed formalism is based on Pontryagin-Koopman operators, which have been shown to possess a symplectic structure, as revealed by Theorem 9. Moreover, Theorem 13 and Corollary 16 have established conditions under which the spectrum of the differential Pontryagin-Koopman operator contains at least 2nx2n_{x} mirrored eigenvalues. This spectral structure is used in Theorem 18 to characterize optimal control laws.

The theoretical findings of this paper have been applied to construct accurate approximations of a globally optimal control law for a Van der Pol oscillator, which illustrates the potential of the proposed Pontryagin-Koopman operator based framework for the design of global optimal control algorithms. Here, it needs to be highlighted that, in contrast to dynamic programming methods, which rely on the discretization of nonlinear Hamilton-Jacobi-Bellman PDEs, the proposed framework for global optimal control relies on the computation of eigenfunctions of a linear differential operator. This opens the door to an application of linear algebra methods and tailored discretization schemes for linear PDEs, which have never been considered for the computation of optimal control laws. Therefore, the development of tailored, structure-exploiting projection and linear algebra methods for symplectic Pontryagin-Koopman operators and their application to global optimal control can be regarded as a promising direction for future research.

References

  • [1] H. Arbabi and I. Mezić. Ergodic theory, dynamic mode decomposition, and computation of spectral properties of the Koopman operator. SIAM Journal on Applied Dynamical Systems, 16(4), 2096–2126.
  • [2] V.I. Arnold and S.P. Novikov. Symplectic geometry. Dynamical systems IV. Springer, 2001.
  • [3] R. E. Bellman. Dynamic Programming. Princeton University Press, Princeton, 1957.
  • [4] L.T. Biegler. An overview of simultaneous strategies for dynamic optimization. Chemical Engineering and Processing, 46:1043–1053, 2007.
  • [5] S. Bittanti, P. Colaneri, and G. De Nicolao. The periodic Riccati equation. Springer, 1991.
  • [6] M. Budišić, R. Mohr, and I. Mezić. Applied Koopmanism. Chaos, 22(4), 047510, 2012.
  • [7] B. Chachuat, A.B. Singer, and P.I. Barton. Global methods for dynamic optimization and mixed-integer dynamic optimization. Ind. Eng. Chem. Res., 45(25):8373–8392, 2006.
  • [8] H. Diedam and S. Sager. Global optimal control with the direct multiple shooting method. Optimal Control Applications and Methods, DOI. 10.1002/oca.2324, 2017.
  • [9] M. Diehl, H.G. Bock, J.P. Schlöder, R. Findeisen, Z. Nagy, and F. Allgöwer. Real-time optimization and nonlinear model predictive control of processes governed by differential-algebraic equations. Journal of Process Control, 12(4):577–585, 2002.
  • [10] J. Ding. The point spectrum of Frobenius-Perron and Koopman operators. Proceedings of the American Mathematical Society, 126(5), 1998.
  • [11] W. R. Esposito and C. A. Floudas. Deterministic global optimization in nonlinear optimal control problems. J. Global Optim., 17:96–126, 2000.
  • [12] X. Feng, M.E. Villanueva, B. Chachuat, and B. Houska. Branch-and-lift algorithm for obstacle avoidance control. In In Proceedings of the 56th IEEE Conference on Decision and Control, page 745–750, 2017.
  • [13] H. Frankowska. Lower semicontinuous solutions of Hamilton-Jacobi-Bellman equations. SIAM Journal on Control and Optimization, 31(1):257–272.
  • [14] N. Govindarajan, R. Mohr, S. Chandrasekaran, and I. Mezić. On the approximation of Koopman spectra for measure preserving transformations. SIAM Journal on Applied Dynamical Systems, 18(3):1454–1497, 2019.
  • [15] L. Grüne and W. Semmler. Using dynamic programming with adaptive grid scheme to solve nonlinear dynamic models in economics. Computing in Economics and Finance 2002, (99), 2002.
  • [16] B. Houska and B. Chachuat. Branch-and-lift algorithm for deterministic global optimization in nonlinear optimal control. Journal of Optimization Theory and Applications, 162(1):208–248, 2014.
  • [17] B. Houska and B. Chachuat. Global optimization in Hilbert space. Mathematical Programming, Series A, 173:221–249, 2019.
  • [18] B. Houska, H.J. Ferreau, and M. Diehl. An auto-generated real-time iteration algorithm for nonlinear MPC in the microsecond range. Automatica, 47(10):2279–2285, 2011.
  • [19] B.O. Koopman. Hamiltonian systems and transformations in Hilbert space. Proceedings of the National Academy of Sciences of the United States of America, 17(5):315–318, 1931.
  • [20] M. Korda and I. Mezić. Linear predictors for nonlinear dynamical systems: Koopman operator meets model predictive control. Automatica, 93:149–160, 2018.
  • [21] D. Liberzon. Calculus of Variations and Optimal Control Theory: A Concise Introduction. Princeton University Press, 2012.
  • [22] Y. Lin and M. A. Stadtherr. Deterministic global optimization of nonlinear dynamic systems. AIChE J., 53(4):866–875, 2007.
  • [23] R. Luss. Optimal control by dynamic programming using systematic reduction in grid size. International Journal of Control, 51(5):995–1013, 1990.
  • [24] J. Mattingley and S. Boyd. Automatic Code Generation for Real-Time Convex Optimization. Cambridge University Press, 2009.
  • [25] I. Mauroy, A.and Mezić. Global stability analysis using the eigenfunctions of the Koopman operator. IEEE Transactions on Automatic Control, 61(11):3356–3369, 2016.
  • [26] I. Mezić. Spectral properties of dynamical systems, model reduction and decompositions. Nonlinear Dynamics, 41(1–3):309–325, 2005.
  • [27] I. Papamichail and C.S. Adjiman. A rigorous global optimization algorithm for problems with ordinary differential equations. J. Global Optim., 24:1–33, 2002.
  • [28] L.S. Pontryagin, V.G. Boltyanskii, R.V. Gamkrelidze, and E.F. Mishchenko. The mathematical theory of optimal processes. John Wiley & Sons, New York, 1962.
  • [29] S.J. Qin and T.A. Badgwell. A survey of industrial model predictive control technology. Control engineering practice, 11(7):733–764, 2003.
  • [30] Luis Rodrigues, Didier Henrion, and Brian J Cantwell. Symmetries and analytical solutions of the hamilton–jacobi–bellman equation for a class of optimal control problems. Optimal Control Applications and Methods, 37(4):749–764, 2016.
  • [31] Hans-Görg Roos, Martin Stynes, and Lutz Tobiska. Robust numerical methods for singularly perturbed differential equations: convection-diffusion-reaction and flow problems, volume 24. Springer Science & Business Media, 2008.
  • [32] A. B. Singer and P. I. Barton. Global optimization with nonlinear ordinary differential equations. J. Global Optim., 34(2):159–190, 2006.
  • [33] N. Wiener and A. Wintner. Harmonic analysis and ergodic theory. American Journal of Mathematics, 63(2):415–426, 1941.
  • [34] K. Yosida. Functional analysis. Springer, Berlin, 1978.
  • [35] V.M. Zavala and L.T. Biegler. The advanced step NMPC controller: optimality, stability and robustness. Automatica, 45:86–93, 2009.