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

[2,3]\fnmYuka \surHashimoto

1]\orgdivDepartment of Mathematics, \orgnameDartmouth University, \orgaddress\street27 N. Main Street, \cityHanover, \postcode03755, \stateNew Hampshire, \countryUSA

[2]\orgdivNTT Network Service Systems Laboratories, \orgnameNTT Corporation, \orgaddress\street3-9-11, Midori-cho, \cityMusashino, \postcode180-8585, \stateTokyo, \countryJapan

3]\orgdivCenter for Advanced Intelligence Project, \orgnameRIKEN, \orgaddress\street1-4-1 Nihonbashi, \cityChuo, \postcode103-0027, \stateTokyo, \countryJapan

4]\orgdivFaculty of Science and Technology, \orgnameKeio University, \orgaddress\street3-14-1, Hiyoshi, \cityKohoku, Yokohama, \postcode223-8522, \stateKanagawa, \countryJapan

5]\orgdivCenter for Data Science, \orgnameEhime University \orgaddress\street2-5, Bunkyo-cho, \cityMatsuyama, \postcode790-8577, \stateEhime, \countryJapan

Koopman spectral analysis of skew-product dynamics on Hilbert CC^{*}-modules

\fnmDimitrios \surGiannakis [email protected]    [email protected]    \fnmMasahiro \surIkeda [email protected]    \fnmIsao \surIshikawa [email protected]    \fnmJoanna \surSlawinska [email protected] [ * [ [ [
Abstract

We introduce a linear operator on a Hilbert CC^{*}-module for analyzing skew-product dynamical systems. The operator is defined by composition and multiplication. We show that it admits a decomposition in the Hilbert CC^{*}-module, called eigenoperator decomposition, that generalizes the concept of the eigenvalue decomposition. This decomposition reconstructs the Koopman operator of the system in a manner that represents the continuous spectrum through eigenoperators. In addition, it is related to the notions of cocycle and Oseledets subspaces and it is useful for characterizing coherent structures under skew-product dynamics. We present numerical applications to simple systems on two-dimensional domains.

keywords:
Koopman operator, transfer operator, operator cocycle, Hilbert CC^{*}-module, skew-product dynamical system

1 Introduction

1.1 Background and motivation

Operator-theoretic methods have been used extensively in analysis and computational techniques for dynamical systems. Let f:𝒳𝒳f:\mathcal{X}\to\mathcal{X} be a dynamical system on a state space 𝒳\mathcal{X}. Then, the Koopman operator UfU_{f} associated with ff is defined as a composition operator on an ff-invariant function space \mathcal{F} on 𝒳\mathcal{X},

Ufv=vf,U_{f}v=v\circ f,

for vv\in\mathcal{F} [1, 2]. In many cases, \mathcal{F} is chosen as a Banach space or Hilbert space, such as the Lebesgue spaces Lp(𝒳)L^{p}(\mathcal{X}) for a measure space 𝒳\mathcal{X} and the Hardy space Hp(𝔻)H^{p}(\mathbb{D}) on the unit disk 𝔻\mathbb{D}, where p[1,]p\in[1,\infty]. Meanwhile, the Perron–Frobenius, or transfer, operator associated with ff is defined as the adjoint PfP_{f} of the Koopman operator acting on the continuous dual \mathcal{F}^{\prime} of \mathcal{F}, i.e., Pfν=νUfP_{f}\nu=\nu\circ U_{f} for ν\nu\in\mathcal{F}^{\prime} [3]. In a number of important cases (e.g., =Lp(𝒳)\mathcal{F}=L^{p}(\mathcal{X}) with p[1,)p\in[1,\infty) or =C(𝒳)\mathcal{F}=C(\mathcal{X}) for a compact Hausdorff space 𝒳\mathcal{X}), \mathcal{F}^{\prime} can be identified with a space of measures on 𝒳\mathcal{X}; the transfer operator is then identified with the pushforward map on measures, Pfν=νf1P_{f}\nu=\nu\circ f^{-1}. When \mathcal{F} has a predual, \mathcal{F}_{*}\subseteq\mathcal{F}^{\prime} it is common to define PfP_{f} as the predual of the Koopman operator, i.e., (Ufv)ν=v(Pfν)(U_{f}v)\nu=v(P_{f}\nu); an important such example is =L(𝒳)\mathcal{F}=L^{\infty}(\mathcal{X}) with =L1(𝒳)\mathcal{F}_{*}=L^{1}(\mathcal{X}). A central tenet of modern ergodic theory is to leverage the duality relationships between f:𝒳𝒳f:\mathcal{X}\to\mathcal{X}, Uf:U_{f}:\mathcal{F}\to\mathcal{F}, and Pf:P_{f}:\mathcal{F}^{\prime}\to\mathcal{F}^{\prime} to characterize properties of nonlinear dynamics such ergodicity, mixing, and existence of factor maps, using linear operator-theoretic techniques [4].

Starting from work in the late 1990s and early 2000s [5, 6, 7, 8], operator-theoretic techniques have also proven highly successful in data-driven applications [9, 10, 11, 12]. A primary such application is the modal decomposition (e.g., [13, 14, 15, 16, 17, 18]). This approach applies eigenvalue decomposition to the Koopman operator to identify the long-term behavior of the dynamical system. Assume \mathcal{F} is a Hilbert space equipped with an inner product ,\left\langle\cdot,\cdot\right\rangle, and UfU_{f} is normal, bounded, and diagonalizable, with eigenvalues λ1,λ2,\lambda_{1},\lambda_{2},\ldots\in\mathbb{C} and corresponding basis of orthonormal eigenvectors v1,v2,v_{1},v_{2},\ldots\in\mathcal{F}. Then, for uu\in\mathcal{F} and a.e. x𝒳x\in\mathcal{X} we have u(fi(x))=Ufiu(x)=j=1λjivj(x)vj,uu(f^{i}(x))=U_{f}^{i}u(x)=\sum_{j=1}^{\infty}\lambda_{j}^{i}v_{j}(x)\left\langle v_{j},u\right\rangle, where ii\in\mathbb{N} represents discrete time. Therefore, the time evolution of observables is described by the Koopman eigenvalues and corresponding eigenvectors. By computing the eigenvalues of the Koopman operator, we obtain oscillating elements and decaying elements in the dynamical system.

Several attempts have been made to generalize the above decomposition to the case where the Koopman operator has continuous or residual spectrum. Korda et al. [19] approximate the spectral measure of the Koopman operator on L2(𝒳)L^{2}(\mathcal{X}) for measure-preserving dynamics using Christoffel–Darboux kernels in spectral space. Slipantschuk et al. [20] consider a riddged Hilbert space and extend the Koopman operator to a space of distributions so that it becomes compact. Colbrook and Townsend [21] employ a residual-based approach that consistently approximates the spectral measure by removing spurious eigenvalues from DMD-type spectral computations. Spectrally approximating the Koopman operator in measure-preserving, ergodic flows by compact operators on reproducing kernel Hilbert spaces (RKHSs) has also been investigated [22]. However, dealing with continuous and residual Koopman spectra is still a challenging problem.

On the transfer operator side, popular approximation techniques are based on the Ulam method [23]. The Ulam method has been shown to yield spectrally consistent approximations for particular classes of systems such as expanding maps and Anosov diffeomorphisms on compact manifolds [5]. In some cases, spectral computations from the Ulam method has been shown to recover eigenvalues of transfer operators on anisotropic Banach spaces adapted to the expanding/contracting subspaces of such systems [24]; however, these results depend on carefully chosen state space partitions that may be hard to construct in high dimensions and/or under unknown dynamics. Various modifications of the basic Ulam method have been proposed that are appropriate for high-dimensional applications; e.g., sparse grid techniques [25].

1.2 Skew-product dynamical systems

We focus on measure-preserving skew-product systems in discrete time, T(y,z)=(h(y),g(y,z))T(y,z)=(h(y),g(y,z)), or continuous-time, Φt(y,z)=(ht(y),gt(y,z))\Phi_{t}(y,z)=(h_{t}(y),g_{t}(y,z)), on a product space 𝒳=𝒴×𝒵\mathcal{X}=\mathcal{Y}\times\mathcal{Z}. Here, 𝒴\mathcal{Y} and 𝒵\mathcal{Z} are measure spaces, oftentimes referred to as the “base” and “fiber”, respectively. In such systems, the driving dynamics on 𝒴\mathcal{Y} is autonomous, but the dynamics on 𝒵\mathcal{Z} depends on the configuration y𝒴y\in\mathcal{Y}. In many cases, one is interested in the time-dependent fiber dynamics, rather than the autonomous dynamics on the base. A typical example of skew-product dynamics is Lagrangian tracer advection under a time-dependent fluid flow [26, 27, 28], where 𝒴\mathcal{Y} is the state space of the fluid dynamical equations of motion and 𝒵\mathcal{Z} is the spatial domain where tracer advection takes place.

A well-studied approach for analysis of skew-product systems involves replacing the spectral decomposition of Koopman/transfer operators acting on functions on 𝒳\mathcal{X} by decomposition of associated operator cocycles acting on functions on 𝒴\mathcal{Y} using multiplicative ergodic theorems. In a standard formulation of the multiplicative ergodic theorem, first proved by Oseledets [29], one considers an invertible measure-preserving map h:𝒴𝒴h:\mathcal{Y}\to\mathcal{Y} and the cocycle generated by a matrix-valued map AA on 𝒴\mathcal{Y}. The multiplicative ergodic theorem then shows the existence of subspaces 𝒱1(y),,𝒱k(y)\mathcal{V}_{1}(y),\ldots,\mathcal{V}_{k}(y) such that A(y)𝒱j(y)=𝒱j(h(y))A(y)\mathcal{V}_{j}(y)=\mathcal{V}_{j}(h(y)). The subspace 𝒱j(y)\mathcal{V}_{j}(y) is called an Oseledets subspace (or equivariant subspace). Each Oseledets subspace has an associated Lyapunov exponent and associated covariant vectors, which are the analogs of the eigenvalues and eigenvectors of Koopman/transfer operators, respectively, in the setting of cocycles. Since its inception, the multiplicative ergodic theorem has been extended in many ways to infinite-dimensional operator cocycles [30, 31, 32, 33, 34]. Under appropriate quasi-compactness assumptions, it has been shown that the Lyapunov exponent spectrum is at most countably infinite and the associated Oseledets subspaces are finite-dimensional, e.g., [34, Theorem A].

A primary application of Oseledets decompositions is the detection of coherent sets and coherent structures in natural and engineered systems [35]. A family of sets {𝒮(y)}y𝒴\{\mathcal{S}(y)\}_{y\in\mathcal{Y}} is called coherent if ν(𝒮(y)g(h(y),𝒮(y)))/ν(𝒮(y))\nu(\mathcal{S}(y)\bigcap g(h(y),\mathcal{S}(y)))/\nu(\mathcal{S}(y)) is large for a reference measure ν\nu on 𝒵\mathcal{Z}. If an Oseledets subspace 𝒱(y)\mathcal{V}(y) with respect to the transfer operator cocycle Ug(y,)1U_{g(y,\cdot)}^{-1} is represented as 𝒱(y)=Span{vy}\mathcal{V}(y)=\operatorname{Span}\{v_{y}\} for a covariant vector vyL2(𝒵)v_{y}\in L^{2}(\mathcal{Z}) satisfying Ug(y,)vh(y)=vyU_{g(y,\cdot)}v_{h(y)}=v_{y}, then setting 𝒮(y)\mathcal{S}(y) to a level set of vyv_{y} leads to a family of coherent sets. Finite-time coherent sets and Lagrangian coherent structures as the boundaries of the finite-time coherent sets have also been studied [26, 36, 37].

1.3 Eigenoperator decomposition

In this paper, we investigate a different approach to deal with continuous and residual spectra of Koopman operators on L2L^{2} associated with skew-product dynamical systems. We propose a new decomposition, called eigenoperator decomposition, which reconstructs the Koopman operator from multiplication operators acting on certain subspaces, referred to here as generalized Oseledets spaces. These multiplication operators are obtained by solving an eigenvalue-type equation, but they can individually have continuous spectrum. Intuitively, this decomposition provides a factorization of the (potentially continuous) spectrum of the underlying Koopman operator into the spectra of eigenoperator families.

Our approach is based on the theory of Hilbert CC^{*}-modules [38], which generalizes Hilbert space theory by replacing the complex-valued inner product by a product that takes values in a CC^{*}-algebra. In this work, we employ the CC^{*}-algebra of bounded linear operators on L2(𝒵)L^{2}(\mathcal{Z}), denoted by (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z})). A standard operator-theoretic approach for skew-product dynamics is to define the Koopman or transfer operator on the product Hilbert space =L2(𝒴)L2(𝒵)\mathcal{H}=L^{2}(\mathcal{Y})\otimes L^{2}(\mathcal{Z}) [28, 39]. In contrast, here we consider the Hilbert CC^{*}-module =L2(𝒴)(L2(𝒵))\mathcal{M}=L^{2}(\mathcal{Y})\otimes\mathcal{B}(L^{2}(\mathcal{Z})) over (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z})). By considering (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z})) instead of L2(𝒵)L^{2}(\mathcal{Z}), we aim to push information about the continuous spectrum of the Koopman operator onto the CC^{*}-algebra (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z})).

In more detail, starting from discrete-time systems, we define a (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z}))-linear operator KTK_{T} on \mathcal{M}, which can be thought of as a lift of the standard Koopman operator on \mathcal{H} to the Hilbert CC^{*}-module setting. In addition, KTK_{T} can be used to reconstruct the Koopman operator of the full skew-product system on 𝒳\mathcal{X}. We show that KTK_{T} admits a decomposition

KTw^i,j=w^i,jM^i,j,K_{T}\hat{w}_{i,j}=\hat{w}_{i,j}\cdot\hat{M}_{i,j}, (1)

where M^i,j\hat{M}_{i,j} is a (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z}))-linear multiplication operator (which we call eigenoperator), and w^i,j\hat{w}_{i,j}\in\mathcal{M} are eigenvectors associated with the operator cocycle on L2(𝒵)L^{2}(\mathcal{Z}) induced by the skew-product dynamics. We also derive an analogous version of (1) for continuous-time systems, formulated in terms of the generator of the Koopman group {KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} acting on \mathcal{M}. A schematic overview of our approach for the continuous-time case is displayed in Fig. 1.

The eigenoperator decomposition (1) and its continuous-time variant have associated equivariant subspaces of L2(𝒵)L^{2}(\mathcal{Z}) as in the multiplicative ergodic theorem. In particular, to each eigenoperator M^i,j\hat{M}_{i,j} there is an associated family {𝒱j(y)}y𝒴\{\mathcal{V}_{j}(y)\}_{y\in\mathcal{Y}} of closed subspaces 𝒱j(y)L2(𝒵)\mathcal{V}_{j}(y)\subseteq L^{2}(\mathcal{Z}) such that Ug(y,)U_{g(y,\cdot)} maps vectors in 𝒱j(h(y))\mathcal{V}_{j}(h(y)) to vectors in 𝒱j(y)\mathcal{V}_{j}(y). Since we consider cocycles generated by unitary Koopman/transfer operators, the equivariant subspaces 𝒱j(y)\mathcal{V}_{j}(y) can be infinite-dimensional. Therefore, we call them generalized Oseledets subspaces. Spectral analysis of M^i,j\hat{M}_{i,j} then reveals coherent structures under the skew-product dynamics.

The rest of this paper is organized as follows. In Section 2, we derive our eigenoperator decomposition for discrete-time systems, and establish the correspondence between KTK_{T} and the Koopman operator. We illustrate the decomposition in Section 3 by means of analytical examples with fiber dynamics on abelian and non-abelian groups. In these examples, the generalized Oseledets subspaces can be constructed explicitly, which provides intuition about the behavior of eigenoperator decomposition. In Section 4, we describe the construction of the infinitesimal generator and the associated eigenoperator decomposition for continuous-time systems. Section 5 contains numerical applications of the decomposition for continuous-time systems to simple time-dependent flows in two-dimensional domains. Section 6 contains a conclusory discussion. The paper includes an Appendix collecting auxiliary results.

Φt(y,z)=(ht(y),gt(y,z))\Phi_{t}(y,z)=(h_{t}(y),g_{t}(y,z)): Skew product flowKΦtK_{\Phi_{t}}: Linear operator on KΦtK_{\Phi_{t}}: the Hilbert CC^{*}-moduleKΦtK_{\Phi_{t}}: =L2(𝒴)(L2(𝒵))\mathcal{M}=L^{2}(\mathcal{Y})\otimes\mathcal{B}(L^{2}(\mathcal{Z}))LΦL_{\Phi}: GeneratorLΦL_{\Phi}: of KΦK_{\Phi}UΦtU_{\Phi_{t}}: Koopman operator on UΦtU_{\Phi_{t}}: a Banach space UΦtU_{\Phi_{t}}: 𝒩=C(𝒴)L2(𝒵)\mathcal{N}=C(\mathcal{Y})\otimes L^{2}(\mathcal{Z})VΦV_{\Phi}: GeneratorVΦV_{\Phi}: of UΦtU_{\Phi_{t}}EigenvectorEigenoperatorDecomposition (Theorems 8 and 21)LΦw^s,j=w^s,jN^s,jL_{\Phi}{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\hat{w}_{s,j}}={\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\hat{w}_{s,j}}\cdot{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\hat{N}_{s,j}}w^s,j{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\hat{w}_{s,j}}\in\mathcal{M}, w^s,j=wspj{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\hat{w}_{s,j}}=w_{s}p_{j}(j=1,2,,s)(j=1,2,\ldots,\ s\in\mathbb{R})   wN^s,j(y)=w(y)(VΦpj)(hs(y))w\cdot{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\hat{N}_{s,j}}(y)=w(y)({\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}V_{\Phi}}p_{j})(h_{s}(y))𝒱j\mathcal{V}_{j}: Invariant𝒱j\mathcal{V}_{j}: subspace of VΦV_{\Phi}Fix yy𝒱j(y)\mathcal{V}_{j}(y): Oseledets space on L2(𝒵)L^{2}(\mathcal{Z})pj(y)p_{j}(y): Projection onto 𝒱j(y)\mathcal{V}_{j}(y)
Figure 1: Overview of eigenoperator decomposition for continuous-time systems.

2 Discrete-time systems

2.1 Skew product system and Koopman operator on Hilbert space

Let 𝒴\mathcal{Y} and 𝒵\mathcal{Z} be separable measure spaces equipped with measures μ\mu and ν\nu, respectively and let 𝒳=𝒴×𝒵\mathcal{X}=\mathcal{Y}\times\mathcal{Z}, the direct product measure space of 𝒴\mathcal{Y} and 𝒵\mathcal{Z}. Let h:𝒴𝒴h:\mathcal{Y}\to\mathcal{Y} be a measure preserving and invertible map and let g:𝒳𝒵g:\mathcal{X}\to\mathcal{Z} be a measurable map such that g(y,)g(y,\cdot) is measure preserving and invertible for any y𝒴y\in\mathcal{Y}. Consider the following skew product transformation TT on 𝒳\mathcal{X}:

T(y,z)=(h(y),g(y,z)).T(y,z)=(h(y),g(y,z)).

We consider the Koopman operator UTU_{T} on L2(𝒳)L^{2}(\mathcal{X}). Note that since L2(𝒴)L^{2}(\mathcal{Y}) and L2(𝒵)L^{2}(\mathcal{Z}) are separable, their tensor product L2(𝒴)L2(𝒵)L^{2}(\mathcal{Y})\otimes L^{2}(\mathcal{Z}) satisfies

L2(𝒴)L2(𝒵)L2(𝒳).\displaystyle L^{2}(\mathcal{Y})\otimes L^{2}(\mathcal{Z})\simeq L^{2}(\mathcal{X}).
Definition 1.

The Koopman operator UTU_{T} on L2(𝒳)L^{2}(\mathcal{X}) is defined as

UTf=fTU_{T}f=f\circ T

for fL2(𝒳)f\in L^{2}(\mathcal{X}).

Since TT is measure preserving, the Koopman operator UTU_{T} is an unitary operator, but UTU_{T} does not always have an eigenvalue decomposition since it has continuous spectrum in general.

2.2 Operator on Hilbert CC^{*}-module related to the Koopman operator

We extend the Koopman operator UTU_{T} to an operator on a Hilbert CC^{*}-module. We first introduce Hilbert CC^{*}-module [38, 40].

Definition 2.

For a module \mathcal{M} over a CC^{*}-algebra 𝒜\mathcal{A}, a map ,:×𝒜\left\langle\cdot,\cdot\right\rangle_{\mathcal{M}}:\mathcal{M}\times\mathcal{M}\to\mathcal{A} is referred to as an 𝒜\mathcal{A}-valued inner product if it is \mathbb{C}-linear with respect to the second variable and has the following properties: For w1,w2,w3w_{1},w_{2},w_{3}\in\mathcal{M} and a,b𝒜a,b\in\mathcal{A},

  1. 1.

    w1,w2a+w3b=w1,w2a+w1,w3b\left\langle w_{1},w_{2}a+w_{3}b\right\rangle_{\mathcal{M}}=\left\langle w_{1},w_{2}\right\rangle_{\mathcal{M}}a+\left\langle w_{1},w_{3}\right\rangle_{\mathcal{M}}b,

  2. 2.

    w1,w2=w2,w1\left\langle w_{1},w_{2}\right\rangle_{\mathcal{M}}=\left\langle w_{2},w_{1}\right\rangle_{\mathcal{M}}^{*},

  3. 3.

    w1,w1\left\langle w_{1},w_{1}\right\rangle_{\mathcal{M}} is positive,

  4. 4.

    If w1,w1=0\left\langle w_{1},w_{1}\right\rangle_{\mathcal{M}}=0 then w1=0w_{1}=0.

If ,\left\langle\cdot,\cdot\right\rangle satisfies the conditions 1\sim3, but not 4, then it is called a semi-inner product. Let w=w,w𝒜1/2\|w\|_{\mathcal{M}}=\|\left\langle w,w\right\rangle_{\mathcal{M}}\|_{\mathcal{A}}^{1/2} for ww\in\mathcal{M}. Then \|\cdot\|_{\mathcal{M}} is a norm in \mathcal{M}.

Definition 3.

A Hilbert CC^{*}-module over 𝒜\mathcal{A} or Hilbert 𝒜\mathcal{A}-module is a module over 𝒜\mathcal{A} equipped with an 𝒜\mathcal{A}-valued inner product and complete with respect to the norm induced by the 𝒜\mathcal{A}-valued inner product.

Let 𝒜\mathcal{A} be the CC^{*}-algebra (L2(𝒵))\mathcal{B}(L^{2}(\mathcal{Z})). Let

=L2(𝒴)𝒜,\displaystyle\mathcal{M}=L^{2}(\mathcal{Y})\otimes\mathcal{A},

i.e., the (right) Hilbert 𝒜\mathcal{A}-module defined by the tensor product of the Hilbert \mathbb{C}-module L2(𝒴)L^{2}(\mathcal{Y}) and (right) Hilbert 𝒜\mathcal{A}-module 𝒜\mathcal{A} [38]. We now define an operator on a Hilbert CC^{*}-module.

Definition 4.

We define the a right 𝒜\mathcal{A}-linear operator KTK_{T} on \mathcal{M} (i.e., KTK_{T} is linear and satisfies KT(wa)=(Ktw)aK_{T}(wa)=(K_{t}w)a for all a𝒜a\in\mathcal{A} and ww\in\mathcal{M}) by

KT(va)(y)=v(h(y))Ug(y,)aK_{T}(v\otimes a)(y)=v(h(y))U_{g(y,\cdot)}a

for vL2(𝒴)v\in L^{2}(\mathcal{Y}), a𝒜a\in\mathcal{A}, and y𝒴y\in\mathcal{Y}. Here, for y𝒴y\in\mathcal{Y}, Ug(y,)U_{g(y,\cdot)} is the Koopman operator on L2(𝒵)L^{2}(\mathcal{Z}) with respect to the map g(y,)g(y,\cdot).

The well-definedness of KTK_{T} is not trivial. The following proposition shows the well-definedness of KTK_{T} as an operator from \mathcal{M} to \mathcal{M}.

Proposition 1.

The operator KTK_{T} is a right 𝒜\mathcal{A}-linear unitary operator from \mathcal{M} to \mathcal{M}.

The proof of Proposition 1 is documented in Appendix.

The next proposition shows the relationship between UTU_{T} and KTK_{T}, which enables us to connect existing studies of Koopman operators with our framework.

Proposition 2.

Let {γi}i=1\{\gamma_{i}\}_{i=1}^{\infty} be an orthonormal basis of L2(𝒵)L^{2}(\mathcal{Z}). Let ιi:L2(𝒳)\iota_{i}:L^{2}(\mathcal{X})\to\mathcal{M} and Pi:L2(𝒳)P_{i}:\mathcal{M}\to L^{2}(\mathcal{X}) be linear operators defined as vuvuγiv\otimes u\mapsto v\otimes u\gamma_{i}^{\prime} and vavaγiv\otimes a\mapsto v\otimes a\gamma_{i}, respectively. Then, we have UT=PiKTιiU_{T}=P_{i}K_{T}\iota_{i} for any i=1,2,i=1,2,\ldots. Moreover, we have KT=i=1ιiUTPiK_{T}=\sum_{i=1}^{\infty}\iota_{i}U_{T}P_{i}, where the sum converges strongly to KTK_{T} in \mathcal{M}.

Proof.

For vL2(𝒴)v\in L^{2}(\mathcal{Y}), uL2(𝒵)u\in L^{2}(\mathcal{Z}), y𝒴y\in\mathcal{Y}, and i=1,2,i=1,2,\ldots, we have

KTιi(vu)(y)\displaystyle K_{T}\iota_{i}(v\otimes u)(y) =KT(vuγi)(y)=v(h(y))Ug(y,)uγi\displaystyle=K_{T}(v\otimes u\gamma_{i}^{\prime})(y)=v(h(y))U_{g(y,\cdot)}u\gamma_{i}^{\prime}
=v(h(y))u(g(y,))γi=ιiUT(vu).\displaystyle=v(h(y))u(g(y,\cdot))\gamma_{i}^{\prime}=\iota_{i}U_{T}(v\otimes u). (2)

By acting PiP_{i} on the both sides of Eq. (2), we have PiKTιi(vu)=UT(vu)P_{i}K_{T}\iota_{i}(v\otimes u)=U_{T}(v\otimes u). Since UTU_{T} and KTK_{T} are bounded, PiKTιi=UTP_{i}K_{T}\iota_{i}=U_{T} holds on \mathcal{M}. Moreover, for vL2(𝒴)v\in L^{2}(\mathcal{Y}) and a𝒜a\in\mathcal{A}, we have

i=1ιiPi(va)=i=1vaγiγi=va,\displaystyle\sum_{i=1}^{\infty}\iota_{i}P_{i}(v\otimes a)=\sum_{i=1}^{\infty}v\otimes a\gamma_{i}\gamma_{i}^{\prime}=v\otimes a, (3)

where the convergence is the strong convergence. Since KTK_{T} is bounded, by Eqs. (2) and (3), KT(va)=i=1ιiUTPi(va)K_{T}(v\otimes a)=\sum_{i=1}^{\infty}\iota_{i}U_{T}P_{i}(v\otimes a) holds for vL2(𝒴)v\in L^{2}(\mathcal{Y}) and a𝒜a\in\mathcal{A}. Since UTU_{T} and KTK_{T} are bounded, KT=i=1ιiUTPiK_{T}=\sum_{i=1}^{\infty}\iota_{i}U_{T}P_{i} holds on \mathcal{M}. ∎

2.3 Decomposition of KTK_{T}

We derive a decomposition of KTK_{T} called the eigenoperator decomposition. We first derive a fundamental decomposition using a cocycle on 𝒴\mathcal{Y}. Then, we refine the decomposition using generalized Oseledets subspaces.

2.3.1 Fundamental decomposition using cocycle

We first define vectors to decompose the operator KTK_{T} using a cocycle on 𝒴\mathcal{Y}.

Definition 5.

For ii\in\mathbb{Z}, we define a linear operator wi:L2(𝒵)L2(𝒳)w_{i}:L^{2}(\mathcal{Z})\to L^{2}(\mathcal{X}) as

(wiu)(y,z)={(Ug(y,)Ug(h(y),)Ug(hi1(y),)u)(z)(i>0)I(i=0)(Ug(h1(y),)Ug(h2(y),)Ug(hi(y),)u)(z)(i<0).(w_{i}u)(y,z)=\left\{\begin{array}[]{ll}(U_{g(y,\cdot)}U_{g(h(y),\cdot)}\cdots U_{g(h^{i-1}(y),\cdot)}u)(z)&\quad(i>0)\\ I&\quad(i=0)\\ (U_{g(h^{-1}(y),\cdot)}^{*}U_{g(h^{-2}(y),\cdot)}^{*}\cdots U_{g(h^{i}(y),\cdot)}^{*}u)(z)&\quad(i<0).\end{array}\right.

We can see that \mathcal{M} can be also regarded as a left 𝒜\mathcal{A}-module. Thus, we can also consider left 𝒜\mathcal{A}-linear operators on \mathcal{M}. In the following, we denote the action of a left 𝒜\mathcal{A}-linear operator AA on a vector ww\in\mathcal{M} by wMw\cdot M.

Proposition 3.

For ii\in\mathbb{Z}, we have wiw_{i}\in\mathcal{M}. Moreover, KTwi=wi+1=wiMiK_{T}w_{i}=w_{i+1}=w_{i}\cdot M_{i}, where MiM_{i} is a left 𝒜\mathcal{A}-linear multiplication operator on \mathcal{M} defined as (wMi)(y)=w(y)Ug(hi(y),)(w\cdot M_{i})(y)=w(y)U_{g(h^{i}(y),\cdot)}.

Proof.

We obtain wiw_{i}\in\mathcal{M} in the same manner as the proof of Proposition 1. The identities KTwi=wi+1=wiMiK_{T}w_{i}=w_{i+1}=w_{i}\cdot M_{i} follow by the definition of wiw_{i}. ∎

The vectors wiw_{i} characterize the dynamics within 𝒵\mathcal{Z}, which are specific to skew product dynamical systems and of particular interest to us.

Proposition 4.

The action of the Koopman operator UTU_{T} is decomposed into two parts as

UTi(vu)(y,z)=Uhiv(z)UTi1ug(y,z)U_{T}^{i}(v\otimes u)(y,z)=U_{h}^{i}v(z)\cdot U_{T}^{i-1}u\circ g(y,z) (4)

for vL2(𝒴)v\in L^{2}(\mathcal{Y}), uL2(𝒵)u\in L^{2}(\mathcal{Z}), and ii\in\mathbb{Z}.

Proof.

For i>0i>0, it follows by the definition of UTU_{T}. Regarding the case of i0i\leq 0, UT1U_{T}^{-1} is calculated as follows:

v1u1,UT(v2u2)\displaystyle\left\langle v_{1}\otimes u_{1},U_{T}(v_{2}\otimes u_{2})\right\rangle =z𝒵y𝒴v1(y)u1(z)¯v2(h(y))u2(g(y,z))dμ(y)dν(z)\displaystyle=\int_{z\in\mathcal{Z}}\int_{y\in\mathcal{Y}}\overline{v_{1}(y)u_{1}(z)}v_{2}(h(y))u_{2}(g(y,z))\mathrm{d}\mu(y)\mathrm{d}\nu(z)
=z𝒵y𝒴v1(h1(y))u1(gh1(y)(z))¯v2(y)u2(z)dμ(y)dν(z),\displaystyle=\int_{z\in\mathcal{Z}}\int_{y\in\mathcal{Y}}\overline{v_{1}(h^{-1}(y))u_{1}(g_{h^{-1}(y)}(z))}v_{2}(y)u_{2}(z)\mathrm{d}\mu(y)\mathrm{d}\nu(z),

where for y𝒴y\in\mathcal{Y}, the map gy:𝒵𝒵g_{y}:\mathcal{Z}\to\mathcal{Z} is defined as gy(z)=g(y,z)g_{y}(z)=g(y,z). Thus, we have UT1(uv)(y,z)=v(h1(y))u(gh1(y)(z))U_{T}^{-1}(u\otimes v)(y,z)=v(h^{-1}(y))u(g_{h^{-1}(y)}(z)). As a result, the equation (4) is derived also for i0i\leq 0. ∎

We define a submodule 𝒲\mathcal{W} of \mathcal{M}, which is composed of the vectors wiw_{i} (ii\in\mathbb{Z}). Let

𝒲0={iFwiciF:finite set,ci𝒜}\mathcal{W}_{0}=\bigg{\{}\sum_{i\in F}w_{i}c_{i}\,\mid\,F\subseteq\mathbb{Z}:\ \mbox{finite set},\ c_{i}\in\mathcal{A}\bigg{\}}

and let 𝒲\mathcal{W} be the completion of 𝒲0\mathcal{W}_{0} with respect to the norm in \mathcal{M}. Note that 𝒲\mathcal{W} is a submodule of \mathcal{M} and Hilbert 𝒜\mathcal{A}-module. Moreover, for uL2(𝒵)u\in L^{2}(\mathcal{Z}), let w~u,iL2(𝒳)\tilde{w}_{u,i}\in L^{2}(\mathcal{X}) be defined as w~u,i(y,z)=u(g(hi1(y),,g(h(y),g(y,z))))\tilde{w}_{u,i}(y,z)=u(g(h^{i-1}(y),\ldots,g(h(y),g(y,z))\ldots)) for i>0i>0, w~u,0(y,z)=u(z)\tilde{w}_{u,0}(y,z)=u(z), and w~u,i(y,z)=u(g(hi(y),,g(h2(y),g(h1(y),z))))\tilde{w}_{u,i}(y,z)=u(g(h^{i}(y),\ldots,g(h^{-2}(y),g(h^{-1}(y),z))\ldots)) for i<0i<0. Let

𝒲~0={j=1niFciw~uj,in,F:finite set,ci,ujL2(𝒵)}\tilde{\mathcal{W}}_{0}=\bigg{\{}\sum_{j=1}^{n}\sum_{i\in F}c_{i}\tilde{w}_{u_{j},i}\,\mid\,n\in\mathbb{N},\ F\subseteq\mathbb{Z}:\ \mbox{finite set},\ c_{i}\in\mathbb{C},\ u_{j}\in L^{2}(\mathcal{Z})\bigg{\}}

and 𝒲~\tilde{\mathcal{W}} be the completion of 𝒲~0\tilde{\mathcal{W}}_{0} with respect to the norm in L2(𝒳)L^{2}(\mathcal{X}).

We show the connection of the operator KTK_{T} restricted on 𝒲\mathcal{W} with the Koopman operator UTU_{T}.

Proposition 5.

With the notation defined in Proposition 2, we have KT|𝒲ιi|𝒲~=ιiUT|𝒲~K_{T}|_{\mathcal{W}}\,\iota_{i}|_{\tilde{\mathcal{W}}}=\iota_{i}U_{T}|_{\tilde{\mathcal{W}}} for i=1,2,….

Proof.

For uL2(𝒵)u\in L^{2}(\mathcal{Z}), jj\in\mathbb{Z}, and i>0i>0, we have

(ιiw~u,j)(y)\displaystyle(\iota_{i}\tilde{w}_{u,j})(y) =u(g(hi1(y),,g(h(y),g(y,))))γi\displaystyle=u(g(h^{i-1}(y),\ldots,g(h(y),g(y,\cdot))\ldots))\gamma_{i}^{\prime}
=Ug(hi1(y),)Ug(h(y),)Ug(h(y),)uγi=wi(y)(uγi).\displaystyle=U_{g(h^{i-1}(y),\cdot)}\cdots U_{g(h(y),\cdot)}U_{g(h(y),\cdot)}u\gamma_{i}^{\prime}=w_{i}(y)(u\gamma_{i}^{\prime}).

Thus, we obtain ιiw~u,j𝒲\iota_{i}\tilde{w}_{u,j}\in\mathcal{W}. We obtain ιiw~u,j𝒲\iota_{i}\tilde{w}_{u,j}\in\mathcal{W} for i0i\leq 0 in the same manner as the case of i>0i>0. Therefore, the range of ιi|W~\iota_{i}|_{\tilde{W}} is contained in 𝒲\mathcal{W}. The equality is deduced by the definitions of KTK_{T} and UTU_{T}. ∎

We can describe the decomposition proposed in Proposition 3 using operators on Hilbert CC^{*}-modules. Let

𝒞0={(,c1,c0,c1,)ci𝒜,ci=0 for all but finite i},\displaystyle\mathcal{C}_{0}=\{(\ldots,c_{-1},c_{0},c_{1},\ldots)\,\mid\,c_{i}\in\mathcal{A},\ c_{i}=0\mbox{ for all but finite }i\in\mathbb{Z}\},
𝒞0={(,A1,A0,A1,)Ai:left 𝒜-linear operator on 𝒲,\displaystyle\mathcal{C}^{\prime}_{0}=\{(\ldots,A_{-1},A_{0},A_{1},\ldots)\,\mid\,A_{i}:\ \mbox{left }\mathcal{A}\mbox{-linear operator on }\mathcal{W},
Ai=0 for all but finite i}.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad A_{i}=0\mbox{ for all but finite }i\in\mathbb{Z}\}.

We can see 𝒞0\mathcal{C}_{0} and 𝒞0\mathcal{C}^{\prime}_{0} are right 𝒜\mathcal{A}-modules. We define 𝒜\mathcal{A}-valued semi-inner products in 𝒞0\mathcal{C}_{0} and 𝒞0\mathcal{C}^{\prime}_{0} as

(,c1,c0,c1,),(,d1,d0,d1,)𝒞0=i,jciwi,wjdj,\displaystyle\left\langle(\ldots,c_{-1},c_{0},c_{1},\ldots),(\ldots,d_{-1},d_{0},d_{1},\ldots)\right\rangle_{\mathcal{C}_{0}}=\sum_{i,j\in\mathbb{Z}}c_{i}^{*}\left\langle w_{i},w_{j}\right\rangle_{\mathcal{M}}d_{j},
(,A1,A0,A1,),(,B1,B0,B1,)𝒞0=i,jwiAi,wjBj,\displaystyle\left\langle(\ldots,A_{-1},A_{0},A_{1},\ldots),(\ldots,B_{-1},B_{0},B_{1},\ldots)\right\rangle_{\mathcal{C}^{\prime}_{0}}=\sum_{i,j\in\mathbb{Z}}\left\langle w_{i}\cdot A_{i},w_{j}\cdot B_{j}\right\rangle_{\mathcal{M}},

respectively.

We define an equivalent relation cdc\sim d by cd𝒩c-d\in\mathcal{N} for c,d𝒞0c,d\in\mathcal{C}_{0}, where 𝒩={c𝒞0c,c𝒞0=0}\mathcal{N}=\{c\in\mathcal{C}_{0}\,\mid\,\left\langle c,c\right\rangle_{\mathcal{C}_{0}}=0\}. There is an 𝒜\mathcal{A}-valued inner product on 𝒞0/\mathcal{C}_{0}/\sim given by c,d=c+𝒩,d+𝒩\left\langle c,d\right\rangle=\left\langle c+\mathcal{N},d+\mathcal{N}\right\rangle. We denote by 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime} the completions of 𝒞0/\mathcal{C}_{0}/\sim and 𝒞0/\mathcal{C}^{\prime}_{0}/\sim with respect to the norms induced by the above inner products. We abuse the notation and denote by (,c1,c0,c1,)(\ldots,c_{-1},c_{0},c_{1},\ldots) the equivalent class of (,c1,c0,c1,)(\ldots,c_{-1},c_{0},c_{1},\ldots) with respect to Let WW be a right 𝒜\mathcal{A}-linear operator from 𝒞\mathcal{C}^{\prime} to 𝒲\mathcal{W} defined as

W(,A1,A0,A1,)=iwiAiW(\ldots,A_{-1},A_{0},A_{1},\ldots)=\sum_{i\in\mathbb{Z}}w_{i}\cdot A_{i}

for (,A1,A0,A1,)𝒞0(\ldots,A_{-1},A_{0},A_{1},\ldots)\in\mathcal{C}^{\prime}_{0} and let XX be a right 𝒜\mathcal{A}-linear operator from 𝒲\mathcal{W} to 𝒞\mathcal{C} defined as

XiFwici=(,c1,c0,c1,)X\sum_{i\in F}w_{i}c_{i}=(\ldots,c_{-1},c_{0},c_{1},\ldots)

for a finite set FF\subseteq\mathbb{Z}. In addition, let MM be a right 𝒜\mathcal{A}-linear operator from 𝒞\mathcal{C} to 𝒞\mathcal{C}^{\prime} defined as

M(,c1,c0,c1,)=(,M1c1,M0c0,M1c1,)M(\ldots,c_{-1},c_{0},c_{1},\ldots)=(\ldots,M_{-1}c_{-1},M_{0}c_{0},M_{1}c_{1},\ldots)

for (,c1,c0,c1,)𝒞0(\ldots,c_{-1},c_{0},c_{1},\ldots)\in\mathcal{C}_{0}, which is formally denoted by diag{Mi}i\operatorname{diag}\{M_{i}\}_{i\in\mathbb{Z}}. Here, MiM_{i} is the multiplication operator defined in Proposition 3.

Proposition 6.

The operators WW and XX are unitary operators. Therefore, XWXW is an unitary operator from 𝒞\mathcal{C}^{\prime} to 𝒞\mathcal{C}.

Proof.

Let W~:𝒲𝒞\tilde{W}:\mathcal{W}\to\mathcal{C}^{\prime} be the right 𝒜\mathcal{A}-linear operator defined as W~(iFwici)=(,C1,C0,C1,)\tilde{W}(\sum_{i\in F}w_{i}c_{i})=(\ldots,C_{-1},C_{0},C_{1},\ldots), where CiC_{i} is the left 𝒜\mathcal{A}-linear multiplication operator on 𝒲\mathcal{W} with respect to the constant function cic_{i}. In addition, let X~:𝒞𝒲\tilde{X}:\mathcal{C}\to\mathcal{W} be the right 𝒜\mathcal{A}-linear operator defined as X~(,c1,c0,c1,)=iFwici\tilde{X}(\ldots,c_{-1},c_{0},c_{1},\ldots)=\sum_{i\in F}w_{i}c_{i}. Then, W~\tilde{W} and X~\tilde{X} are the inverses of WW and XX, respectively. Moreover, for (,A1,A0,A1,),(,B1,B0,B1,)𝒞0(\ldots,A_{-1},A_{0},A_{1},\ldots),(\ldots,B_{-1},B_{0},B_{1},\ldots)\in\mathcal{C}^{\prime}_{0}, we have

W(,A1,A0,A1,),W(,B1,B0,B1,)=iwiAi,iwiBi\displaystyle\left\langle W(\ldots,A_{-1},A_{0},A_{1},\ldots),W(\ldots,B_{-1},B_{0},B_{1},\ldots)\right\rangle_{\mathcal{M}}=\bigg{\langle}\sum_{i\in\mathbb{Z}}w_{i}\cdot A_{i},\sum_{i\in\mathbb{Z}}w_{i}\cdot B_{i}\bigg{\rangle}_{\mathcal{M}}
=(,A1,A0,A1,),(,B1,B0,B1,)𝒞\displaystyle\qquad=\left\langle(\ldots,A_{-1},A_{0},A_{1},\ldots),(\ldots,B_{-1},B_{0},B_{1},\ldots)\right\rangle_{\mathcal{C}^{\prime}}

and for ci,di𝒜c_{i},d_{i}\in\mathcal{A}, finite subsets FF and GG of \mathbb{Z}, we have

XiFwici,XiGwidi𝒞\displaystyle\bigg{\langle}X\sum_{i\in F}w_{i}c_{i},X\sum_{i\in G}w_{i}d_{i}\bigg{\rangle}_{\mathcal{C}} =(,c1,c0,c1,),(,d1,d0,d1,)𝒞\displaystyle=\left\langle(\ldots,c_{-1},c_{0},c_{1},\ldots),(\ldots,d_{-1},d_{0},d_{1},\ldots)\right\rangle_{\mathcal{C}}
=iFwici,iGwidi.\displaystyle=\bigg{\langle}\sum_{i\in F}w_{i}c_{i},\sum_{i\in G}w_{i}d_{i}\bigg{\rangle}_{\mathcal{M}}.

Proposition 7.

The operator MM is well-defined and KT|𝒲=WMXK_{T}|_{\mathcal{W}}=WMX.

Proof.

Since KTK_{T} is unitary, we have

wi+1,wj+1=KTwi,KTwj=wi,wj\left\langle w_{i+1},w_{j+1}\right\rangle=\left\langle K_{T}w_{i},K_{T}w_{j}\right\rangle=\left\langle w_{i},w_{j}\right\rangle (5)

for i,ji,j\in\mathbb{Z}. Assume (,c1,c0,c1,)=0(\ldots,c_{-1},c_{0},c_{1},\ldots)=0. Then, by the equation (5), we obtain

M(,c1,c0,c1,),M(,c1,c0,c1,)𝒞\displaystyle\left\langle M(\ldots,c_{-1},c_{0},c_{1},\ldots),M(\ldots,c_{-1},c_{0},c_{1},\ldots)\right\rangle_{\mathcal{C}^{\prime}}
=(,M1c1,M0c0,M1c1,),(,M1c1,M0c0,M1c1,)𝒞\displaystyle\qquad=\left\langle(\ldots,M_{-1}c_{-1},M_{0}c_{0},M_{1}c_{1},\ldots),(\ldots,M_{-1}c_{-1},M_{0}c_{0},M_{1}c_{1},\ldots)\right\rangle_{\mathcal{C}^{\prime}}
=i,jFwiMici,wjMjcj=i,jFwi+1ci,wj+1cj\displaystyle\qquad=\sum_{i,j\in F}\left\langle w_{i}\cdot M_{i}c_{i},w_{j}\cdot M_{j}c_{j}\right\rangle_{\mathcal{M}}=\sum_{i,j\in F}\left\langle w_{i+1}c_{i},w_{j+1}c_{j}\right\rangle_{\mathcal{M}}
=i,jFwici,wjcj=0,\displaystyle\qquad=\sum_{i,j\in F}\left\langle w_{i}c_{i},w_{j}c_{j}\right\rangle_{\mathcal{M}}=0,

which shows the well-definiteness of MM. The decomposition KT|𝒲=WMXK_{T}|_{\mathcal{W}}=WMX is derived by Proposition 3. ∎

In summary, we obtain the following commutative diagram:

𝒲\textstyle{\mathcal{W}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KT\scriptstyle{K_{T}}X\scriptstyle{X}𝒲\textstyle{\mathcal{W}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}W\scriptstyle{W^{*}}𝒞\textstyle{\mathcal{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\scriptstyle{M}X\scriptstyle{X^{*}}𝒞\textstyle{\mathcal{C}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}W\scriptstyle{W}

2.3.2 Further decomposition

We further decompose wiw_{i} and MiM_{i} and obtain a more detailed decomposition of KT|𝒲K_{T}|_{\mathcal{W}}. Let 𝒱1,𝒱2,\mathcal{V}_{1},\mathcal{V}_{2},\ldots be a sequence of maps from 𝒴\mathcal{Y} to the set of all closed subspaces of L2(𝒵)L^{2}(\mathcal{Z}) satisfying L2(𝒵)=Span{j=1𝒱j(y)}¯L^{2}(\mathcal{Z})=\overline{\operatorname{Span}\{\bigcup_{j=1}^{\infty}\mathcal{V}_{j}(y)\}} for a.s. y𝒴y\in\mathcal{Y}. Let pj(y)𝒜p_{j}(y)\in\mathcal{A} be the projection onto 𝒱j(y)\mathcal{V}_{j}(y), i.e., it satisfies pj(y)2=pj(y)p_{j}(y)^{2}=p_{j}(y) and pj(y)=pj(y)p_{j}(y)^{*}=p_{j}(y). For ii\in\mathbb{Z} and j=1,2,j=1,2,\ldots, we define a linear map w^i,j\hat{w}_{i,j} from L2(𝒵)L^{2}(\mathcal{Z}) to L2(𝒳)L^{2}(\mathcal{X}) as (w^i,ju)(y,z)=(wi(y)pj(hi(y))u)(z)(\hat{w}_{i,j}u)(y,z)=(w_{i}(y)p_{j}(h^{i}(y))u)(z). We decompose KT|𝒲K_{T}|_{\mathcal{W}} using w^i,j\hat{w}_{i,j}. For each j=1,2,j=1,2,\ldots, the following theorem holds:

Theorem 8 (Eigenoperator decomposition for discrete-time systems).

Assume 𝒱j\mathcal{V}_{j} satisfies Ug(y,)𝒱j(h(y))𝒱j(y)U_{g(y,\cdot)}\mathcal{V}_{j}(h(y))\subseteq\mathcal{V}_{j}(y) for a.s. y𝒴y\in\mathcal{Y}. Assume in addition, for any uL2(𝒵)u\in L^{2}(\mathcal{Z}) and ii\in\mathbb{Z}, the map (y,z)(pj(hi(y))u)(z)(y,z)\mapsto(p_{j}(h^{i}(y))u)(z) is measurable. Then, w^i,j\hat{w}_{i,j} is contained in \mathcal{M} and we have KTw^i,j=w^i+1,j=w^i,jM^i,jK_{T}\hat{w}_{i,j}=\hat{w}_{i+1,j}=\hat{w}_{i,j}\cdot\hat{M}_{i,j}. Here, M^i,j\hat{M}_{i,j} is a left 𝒜\mathcal{A}-linear multiplication operator on \mathcal{M} defined as (wM^i,j)(y)=w(y)Ug(hi(y),)pj(hi(y))(w\cdot\hat{M}_{i,j})(y)=w(y)U_{g(h^{i}(y),\cdot)}p_{j}(h^{i}(y)).

Proof.

We have

KTw^i,j(y)=Ug(y,)w^i,j(h(y))=wi+1(y)pj(hi+1(y))=w^i+1,j(y).K_{T}\hat{w}_{i,j}(y)=U_{g(y,\cdot)}\hat{w}_{i,j}(h(y))=w_{i+1}(y)p_{j}(h^{i+1}(y))=\hat{w}_{i+1,j}(y).

In addition, since by the assumption, the range of Ug(hi(y),)pj(hi+1(y))U_{g(h^{i}(y),\cdot)}p_{j}(h^{i+1}(y)) is contained in 𝒱(hi(y))\mathcal{V}(h^{i}(y)), we have

KTw^i,j(y)\displaystyle K_{T}\hat{w}_{i,j}(y) =wi(y)Ug(hi(y),)pj(hi+1(y))\displaystyle=w_{i}(y)U_{g(h^{i}(y),\cdot)}p_{j}(h^{i+1}(y))
=wi(y)pj(hi(y))Ug(hi(y),)pj(hi+1(y))=w^i,j(y)M^i,j.\displaystyle=w_{i}(y)p_{j}(h^{i}(y))U_{g(h^{i}(y),\cdot)}p_{j}(h^{i+1}(y))=\hat{w}_{i,j}(y)\cdot\hat{M}_{i,j}.

Corollary 9.

By replacing {wi}i\{w_{i}\}_{i\in\mathbb{Z}} and {Mi}i\{M_{i}\}_{i\in\mathbb{Z}} by {w^i,j}i,j\{\hat{w}_{i,j}\}_{i\in\mathbb{Z},j\in\mathbb{N}} and {M^i,j}i,j\{\hat{M}_{i,j}\}_{i\in\mathbb{Z},j\in\mathbb{N}}, respectively, we define 𝒲^\hat{\mathcal{W}}, 𝒞^\hat{\mathcal{C}}, 𝒞^\hat{\mathcal{C}^{\prime}}, W^\hat{W}, X^\hat{X}, and M^\hat{M} in the same manner as 𝒲\mathcal{W}, 𝒞\mathcal{C}, 𝒞\mathcal{C}^{\prime}, WW, XX, and MM, respectively. Then, under the assumptions of Theorem 8, we obtain KT|𝒲=W^M^X^K_{T}|_{\mathcal{W}}=\hat{W}\hat{M}\hat{X}.

We call M^i,j\hat{M}_{i,j} an eigenoperator and w^i,j\hat{w}_{i,j} an eigenvector. In addition, we call the subspace 𝒱j(y)\mathcal{V}_{j}(y) satisfying the assumption in Theorem 8 generalized Oseledets space.

If 𝒱j(y)\mathcal{V}_{j}(y) is a finite-dimensional space, then we can explicitly calculate the spectrum of M^i,j\hat{M}_{i,j} as follows.

Proposition 10.

Assume dim(𝒱j(y))\operatorname{dim}(\mathcal{V}_{j}(y)) is finite and constant with respect to y𝒴y\in\mathcal{Y}. Then, σ(M^i,j)={λϵ>0,μ({y𝒴λσϵ(fi,j(y))})>0}\sigma(\hat{M}_{i,j})=\{\lambda\in\mathbb{C}\,\mid\,^{\forall}\epsilon>0,\ \mu(\{y\in\mathcal{Y}\,\mid\,\lambda\in\sigma_{\epsilon}(f_{i,j}(y))\})>0\}. Here, fi,j(y)=Ug(hi(y),)pj(hi(y))f_{i,j}(y)=U_{g(h^{i}(y),\cdot)}p_{j}(h^{i}(y)). In addition, σ(a)\sigma(a) and σϵ(a)\sigma_{\epsilon}(a) for a𝒜a\in\mathcal{A} are the spectrum and the essential spectrum of aa, respectively.

Proof.

For λ\lambda\in\mathbb{C}, λIM^i,j\lambda I-\hat{M}_{i,j} is not invertible if and only if there exists ww\in\mathcal{M} with w0w\neq 0 such that w(y)(λIfi,j(y))=0w(y)(\lambda I-f_{i,j}(y))=0 for a.s. y𝒴y\in\mathcal{Y}, which is equivalent to

μ({y𝒴λIfi,j(y) is not invertible})>0.\displaystyle\mu(\{y\in\mathcal{Y}\,\mid\,\lambda I-f_{i,j}(y)\mbox{ is not invertible}\})>0.

Assume λIfi,j(y)\lambda I-f_{i,j}(y) is invertible for a.s. y𝒴y\in\mathcal{Y}. Then, we have (λIM^i,j)1w(y)=w(y)(λIfi,j(y))1(\lambda I-\hat{M}_{i,j})^{-1}w(y)=w(y)(\lambda I-f_{i,j}(y))^{-1}. For ww\in\mathcal{M}, we have

(λIM^i,j)1w2\displaystyle\|(\lambda I-\hat{M}_{i,j})^{-1}w\|_{\mathcal{M}}^{2} =𝒴(λIfi,j(y))w(y)w(y)(λIfi,j(y))1dμ(y)𝒜\displaystyle=\bigg{\|}\int_{\mathcal{Y}}(\lambda I-f_{i,j}(y))^{-*}w(y)^{*}w(y)(\lambda I-f_{i,j}(y))^{-1}\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}
tr(𝒴(λIfi,j(y))w(y)w(y)(λIfi,j(y))1dμ(y))\displaystyle\leq\operatorname{tr}\bigg{(}\int_{\mathcal{Y}}(\lambda I-f_{i,j}(y))^{-*}w(y)^{*}w(y)(\lambda I-f_{i,j}(y))^{-1}\mathrm{d}\mu(y)\bigg{)}
=tr(𝒴w(y)(λIfi,j(y))(λIfi,j(y))1w(y)dμ(y)).\displaystyle=\operatorname{tr}\bigg{(}\int_{\mathcal{Y}}w(y)(\lambda I-f_{i,j}(y))^{-*}(\lambda I-f_{i,j}(y))^{-1}w(y)^{*}\mathrm{d}\mu(y)\bigg{)}.

Thus, we have

1dj𝒴w(y)(λIfi,j(y))(λIfi,j(y))1w(y)dμ(y)𝒜(λIM^i,j)1w2\displaystyle\frac{1}{d_{j}}\bigg{\|}\int_{\mathcal{Y}}w(y)(\lambda I-f_{i,j}(y))^{-*}(\lambda I-f_{i,j}(y))^{-1}w(y)^{*}\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}\leq\|(\lambda I-\hat{M}_{i,j})^{-1}w\|_{\mathcal{M}}^{2}
dj𝒴w(y)(λIfi,j(y))(λIfi,j(y))1w(y)dμ(y)𝒜,\displaystyle\qquad\leq d_{j}\bigg{\|}\int_{\mathcal{Y}}w(y)(\lambda I-f_{i,j}(y))^{-*}(\lambda I-f_{i,j}(y))^{-1}w(y)^{*}\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}},

where dj=dim(𝒱j(y))d_{j}=\operatorname{dim}(\mathcal{V}_{j}(y)). Assume for any ϵ>0\epsilon>0, μ({y𝒴(λIfi,j(y))1𝒜>1/ϵ})>0\mu(\{y\in\mathcal{Y}\,\mid\,\|(\lambda I-f_{i,j}(y))^{-1}\|_{\mathcal{A}}>1/\epsilon\})>0. We set ai,j(y)=vi,j(y)vi,j(y)a_{i,j}(y)=v_{i,j}(y)v_{i,j}(y)^{*}, where vi,j(y)v_{i,j}(y) is the orthonormal eigenvector corresponding to the largest eigenvalue of (λIfi,j(y))λIfi,j(y))1(\lambda I-f_{i,j}(y))^{-*}\lambda I-f_{i,j}(y))^{-1}. Then, we have

(λIM^i,j)1w21dj𝒴(λIfi,j(y))1𝒜2w(y)ai,j(y)w(y)dμ(y)𝒜\displaystyle\|(\lambda I-\hat{M}_{i,j})^{-1}w\|_{\mathcal{M}}^{2}\geq\frac{1}{d_{j}}\bigg{\|}\int_{\mathcal{Y}}\|(\lambda I-f_{i,j}(y))^{-1}\|_{\mathcal{A}}^{2}w(y)a_{i,j}(y)w(y)^{*}\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}
μ({y𝒴(λIfi,j(y))1𝒜>1/ϵ})1ϵ2dj𝒴w(y)a(y)w(y)dμ(y)𝒜.\displaystyle\qquad\geq\mu(\{y\in\mathcal{Y}\,\mid\,\|(\lambda I-f_{i,j}(y))^{-1}\|_{\mathcal{A}}>1/\epsilon\})\frac{1}{\epsilon^{2}d_{j}}\bigg{\|}\int_{\mathcal{Y}}w(y)a(y)w(y)^{*}\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}.

Setting w(y)=ai,j(y)w(y)=a_{i,j}(y), we derive that (λIM^i,j)1(\lambda I-\hat{M}_{i,j})^{-1} is unbounded. Conversely, assume (λIM^i,j)1(\lambda I-\hat{M}_{i,j})^{-1} is unbounded. Then, we obtain

(λIM^i,j)1w2djesssupy𝒴(λIfi,j(y))12𝒴w(y)w(y)dμ(y)𝒜\displaystyle\|(\lambda I-\hat{M}_{i,j})^{-1}w\|_{\mathcal{M}}^{2}\leq d_{j}\operatorname*{ess\,sup}_{y\in\mathcal{Y}}\|(\lambda I-f_{i,j}(y))^{-1}\|^{2}\bigg{\|}\int_{\mathcal{Y}}w(y)^{*}w(y)\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}
djesssupy𝒴(λIfi,j(y))12w2.\displaystyle\qquad\leq d_{j}\operatorname*{ess\,sup}_{y\in\mathcal{Y}}\|(\lambda I-f_{i,j}(y))^{-1}\|^{2}\|w\|_{\mathcal{M}}^{2}.

Thus, for any ϵ>0\epsilon>0, μ({y𝒴(λIfi,j(y))1𝒜>1/ϵ})>0\mu(\{y\in\mathcal{Y}\,\mid\,\|(\lambda I-f_{i,j}(y))^{-1}\|_{\mathcal{A}}>1/\epsilon\})>0, which completes the proof. ∎

2.3.3 Construction of the generalized Oseledets space 𝒱j(y)\mathcal{V}_{j}(y)

For cocycles generated by matrices, the existence of the Oseledets space is guaranteed by the multiplicative ergodic theorem. This theorem has been generalized for cocycles generated by compact operators or operators that have similar properties to the compactness [31, 32]. In our case, since the cocycle is generated by a unitary operator, we can construct 𝒱j\mathcal{V}_{j} explicitly if hh is periodic.

Proposition 11.

Assume hn(y)=yh^{n}(y)=y for any y𝒴y\in\mathcal{Y}. Let U:𝒴(k=1nL2(𝒵))U:\mathcal{Y}\to\mathcal{B}(\oplus_{k=1}^{n}L^{2}(\mathcal{Z})) be defined as

U(y)=(Ug(h(y),)Ug(h2(y),)Ug(hn1(y),)Ug(y,))S,U(y)=(U_{g(h(y),\cdot)}\oplus U_{g(h^{2}(y),\cdot)}\oplus\cdots\oplus U_{g(h^{n-1}(y),\cdot)}\oplus U_{g(y,\cdot)})S,

where SS is the permutation operator defined as Sk=1nuk=k=1n1uk+1u1S\oplus_{k=1}^{n}u_{k}=\oplus_{k=1}^{n-1}u_{k+1}\oplus u_{1}. Let E(y)E(y) be the spectral measure with respect to U(y)U(y) and 𝒯[0,2π)\mathcal{T}\subset[0,2\pi) be a subset of [0,2π)[0,2\pi). Let 𝒱~k(y)\tilde{\mathcal{V}}_{k}(y) be the range of PkE(y)(T)P_{k}E(y)(T) and let 𝒱(y)=𝒱~n(y)\mathcal{V}(y)=\tilde{\mathcal{V}}_{n}(y), where Pk:l=1nL2(𝒵)L2(𝒵)P_{k}:\oplus_{l=1}^{n}L^{2}(\mathcal{Z})\to L^{2}(\mathcal{Z}) is the projection defined as l=1nuluk\oplus_{l=1}^{n}u_{l}\mapsto u_{k}. Then, we have Ug(y,)𝒱(h(y))𝒱(y)U_{g(y,\cdot)}\mathcal{V}(h(y))\subseteq\mathcal{V}(y).

Proof.

We first show k=1n𝒱~k(y)\oplus_{k=1}^{n}\tilde{\mathcal{V}}_{k}(y) is an invariant subspace of U(y)U(y). Let uE(y)(𝒯)u\in E(y)(\mathcal{T}). Then, we have

U(y)P~ku=Ug(hk1(y),)uk=P~kU(y)uU(y)\tilde{P}_{k}u=U_{g(h^{k-1}(y),\cdot)}u_{k}=\tilde{P}_{k}U(y)u

for k=1,,nk=1,\ldots,n. Here, P~k:l=1nL2(𝒵)l=1nL2(𝒵)\tilde{P}_{k}:\oplus_{l=1}^{n}L^{2}(\mathcal{Z})\to\oplus_{l=1}^{n}L^{2}(\mathcal{Z}) is the linear operator defined as l=1null=1nu~l\oplus_{l=1}^{n}u_{l}\mapsto\oplus_{l=1}^{n}\tilde{u}_{l}, where u~l=0\tilde{u}_{l}=0 for lkl\neq k and u~l=uk\tilde{u}_{l}=u_{k} for l=kl=k. Since E(y)(𝒯)E(y)(\mathcal{T}) is an invariant subspace of U(y)U(y), we have U(y)uE(y)(𝒯)U(y)u\in E(y)(\mathcal{T}). Thus, we have U(y)k=1n𝒱~k(y)k=1n𝒱~k(y)U(y)\oplus_{k=1}^{n}\tilde{\mathcal{V}}_{k}(y)\subseteq\oplus_{k=1}^{n}\tilde{\mathcal{V}}_{k}(y). Therefore, we have Ug(y,)𝒱~n(h(y))𝒱~n1(h(y))U_{g(y,\cdot)}\tilde{\mathcal{V}}_{n}(h(y))\subseteq\tilde{\mathcal{V}}_{n-1}(h(y)). In addition, for k=1nukk=1nL2(𝒵)\oplus_{k=1}^{n}u_{k}\in\oplus_{k=1}^{n}L^{2}(\mathcal{Z}), we have

S1U(h(y))Sk=1nuk\displaystyle S^{-1}U(h(y))S\oplus_{k=1}^{n}u_{k} =S1U(h(y))k=1nuk+1=S1k=1nUg(hk+1(y),)(y)uk+2\displaystyle=S^{-1}U(h(y))\oplus_{k=1}^{n}u_{k+1}=S^{-1}\oplus_{k=1}^{n}U_{g(h^{k+1}(y),\cdot)}(y)u_{k+2}
=k=1nUg(hk(y),)(y)uk+1=U(y)k=1nuk,\displaystyle=\oplus_{k=1}^{n}U_{g(h^{k}(y),\cdot)}(y)u_{k+1}=U(y)\oplus_{k=1}^{n}u_{k},

where uk+n=uku_{k+n}=u_{k} for k=1,2k=1,2. Thus, we have S1U(h(y))S=U(y)S^{-1}U(h(y))S=U(y), and the spectral measure E(h(y))E(h(y)) of U(h(y))U(h(y)) is represented as E(h(y))=SE(y)S1E(h(y))=SE(y)S^{-1}. Therefore, for k=1,2,k=1,2,\ldots, we obtain

PkE(h(y))(𝒯)=PkSE(y)(𝒯)S1=Pk+1E(y)(𝒯)S1,P_{k}E(h(y))(\mathcal{T})=P_{k}SE(y)(\mathcal{T})S^{-1}=P_{k+1}E(y)(\mathcal{T})S^{-1},

where Pk+1=P1P_{k+1}=P_{1}, which implies 𝒱~k(h(y))=𝒱~k+1(y)\tilde{\mathcal{V}}_{k}(h(y))=\tilde{\mathcal{V}}_{k+1}(y). Combining this identity with the inclusion Ug(y,)𝒱n(h(y))𝒱n1(h(y))U_{g(y,\cdot)}\mathcal{V}_{n}(h(y))\subseteq\mathcal{V}_{n-1}(h(y)), we have Ug(y,)𝒱~n(h(y))𝒱~n(y)U_{g(y,\cdot)}\tilde{\mathcal{V}}_{n}(h(y))\subseteq\tilde{\mathcal{V}}_{n}(y). ∎

Corollary 12.

Let 𝒯1,𝒯2,[0,2π)\mathcal{T}_{1},\mathcal{T}_{2},\ldots\subset[0,2\pi) be a sequence of countable disjoint subsets of [0,2π)[0,2\pi) such that [0,2π)=j=1𝒯j[0,2\pi)=\bigcup_{j=1}^{\infty}\mathcal{T}_{j}. If we set 𝒱j(y)\mathcal{V}_{j}(y) as 𝒱(y)\mathcal{V}(y) in Proposition 11 by replacing 𝒯\mathcal{T} with 𝒯j\mathcal{T}_{j}, it satisfies the assumption in Theorem 8.

2.3.4 Connection with Koopman operator on Hilbert space

Assume pj(y)=p^jp_{j}(y)=\hat{p}_{j} for any y𝒴y\in\mathcal{Y}, where p^j𝒜\hat{p}_{j}\in\mathcal{A} is a projection. By Proposition 8, we obtain the following commutative diagram:

\textstyle{\mathcal{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KT\scriptstyle{K_{T}}Pj\scriptstyle{P_{j}}\textstyle{\mathcal{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Pj\scriptstyle{P_{j}}L2(𝒴)𝒜^j\textstyle{L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KT\scriptstyle{K_{T}}L2(𝒴)𝒜^j\textstyle{L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}}

where 𝒜^j={p^jap^ja𝒜}\hat{\mathcal{A}}_{j}=\{\hat{p}_{j}a\hat{p}_{j}\,\mid\,a\in\mathcal{A}\} is a CC^{*}-subalgebra of 𝒜\mathcal{A} and Pj:L2(𝒴)𝒜^jP_{j}:\mathcal{M}\to L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j} is defined as wPj=wp^jw\cdot P_{j}=w\hat{p}_{j}. If 𝒱j\mathcal{V}_{j} is a finite dimensional space, then L2(𝒴)𝒜^jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j} is isomorphic to a Hilbert space and the action of KTK_{T} on L2(𝒴)𝒜^jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j} is reduced to that of UTU_{T}.

Proposition 13.

Assume 𝒱j\mathcal{V}_{j} is an nn-dimensional space. Let {γ1,,γn}\{\gamma_{1},\ldots,\gamma_{n}\} be an orthonormal basis of 𝒱j\mathcal{V}_{j} and let λj:L2(𝒴)𝒜^ji=1nL2(𝒳)\lambda_{j}:L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}\to\oplus_{i=1}^{n}L^{2}(\mathcal{X}) be a linear operator defined as λj(va)=(vaγ1,,vaγn)\lambda_{j}(v\otimes a)=(v\otimes a\gamma_{1},\ldots,v\otimes a\gamma_{n}). Then, λj\lambda_{j} is an isomorphism and we have the following commutative diagram:

L2(𝒴)𝒜^j\textstyle{L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KT\scriptstyle{K_{T}}λj\scriptstyle{\lambda_{j}}L2(𝒴)𝒜^j\textstyle{L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λj\scriptstyle{\lambda_{j}}i=1nL2(𝒳)\textstyle{\oplus_{i=1}^{n}L^{2}(\mathcal{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i=1nUT\scriptstyle{\oplus_{i=1}^{n}U_{T}}i=1nL2(𝒳)\textstyle{\oplus_{i=1}^{n}L^{2}(\mathcal{X})}
Proof.

Let λ~j:i=1nL2(𝒳)L2(𝒴)𝒜^j\tilde{\lambda}_{j}:\oplus_{i=1}^{n}L^{2}(\mathcal{X})\to L^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j} be a linear operator defined as λ~j(v1u1,,vnun)=i=1nviuiγi\tilde{\lambda}_{j}(v_{1}\otimes u_{1},\ldots,v_{n}\otimes u_{n})=\sum_{i=1}^{n}v_{i}\otimes u_{i}\gamma_{i}^{\prime}. Then, λ~j\tilde{\lambda}_{j} is the inverse of λj\lambda_{j}. In addition, we have

λj(va),λj(va)i=1nL2(𝒳)=i=1nvaγi,vaγiL2(𝒳)\displaystyle\left\langle\lambda_{j}(v\otimes a),\lambda_{j}(v\otimes a)\right\rangle_{\oplus_{i=1}^{n}L^{2}(\mathcal{X})}=\sum_{i=1}^{n}\left\langle v\otimes a\gamma_{i},v\otimes a\gamma_{i}\right\rangle_{L^{2}(\mathcal{X})}
=v,vL2(𝒴)i=1nγi,aaγiL2(𝒵)nv,vL2(𝒴)aa𝒜=nva2.\displaystyle\qquad=\left\langle v,v\right\rangle_{L^{2}(\mathcal{Y})}\sum_{i=1}^{n}\left\langle\gamma_{i},a^{*}a\gamma_{i}\right\rangle_{L^{2}(\mathcal{Z})}\leq n\|\left\langle v,v\right\rangle_{L^{2}(\mathcal{Y})}a^{*}a\|_{\mathcal{A}}=n\|v\otimes a\|_{\mathcal{M}}^{2}.

Thus, λj\lambda_{j} is an isomorphism. The commutativity of the diagram is derived by Proposition 2. ∎

3 Examples

3.1 The case of 𝒵\mathcal{Z} is a compact Hausdorff group

Let 𝒵\mathcal{Z} be a compact Hausdorff group equipped with the (normalized) Haar measure ν\nu. Let 𝒵^\hat{\mathcal{Z}} be the set of equivalent classes of irreducible unitary representations. For an irreducible representation ρ\rho, let ρ\mathcal{E}_{\rho} be the representation space of ρ\rho and let nρn_{\rho} be the dimension of ρ\mathcal{E}_{\rho}. Note that since 𝒵\mathcal{Z} is a compact group, nρn_{\rho} is finite. Let {eρ,1,,eρ,nρ}\{e_{\rho,1},\ldots,e_{\rho,n_{\rho}}\} be an orthonormal basis of ρ\mathcal{E}_{\rho} and let γρ,i,j:𝒵\gamma_{\rho,i,j}:\mathcal{Z}\to\mathbb{C} be the matrix coefficient defined as γρ,i,j(z)=eρ,i,ρ(z)eρ,j\gamma_{\rho,i,j}(z)=\left\langle e_{\rho,i},\rho(z)e_{\rho,j}\right\rangle. By the Peter–Weyl theorem, [ρ]𝒵^{γρ,i,ji,j=1,,nρ}\bigcup_{[\rho]\in\hat{\mathcal{Z}}}\{\gamma_{\rho,i,j}\;\mid\;i,j=1,\ldots,n_{\rho}\} is an orthonormal basis of L2(𝒵)L^{2}(\mathcal{Z}), where [ρ][\rho] is the equivalent class of an irreducible representation ρ\rho. We set the map g:𝒴×𝒵𝒵g:\mathcal{Y}\times\mathcal{Z}\to\mathcal{Z} as g(y,z)=zg~(y)g(y,z)=z\tilde{g}(y), where g~:𝒴𝒵\tilde{g}:\mathcal{Y}\to\mathcal{Z} is a measurable map. Let Γρ,i:ρL2(𝒵)\Gamma_{\rho,i}:\mathcal{E}_{\rho}\to L^{2}(\mathcal{Z}) be the linear operator defined as eρ,jγρ,i,je_{\rho,j}\mapsto\gamma_{\rho,i,j} for i,j=1,,nρi,j=1,\ldots,n_{\rho}. Note that the adjoint Γρ,i:L2(𝒵)ρ\Gamma_{\rho,i}^{*}:L^{2}(\mathcal{Z})\to\mathcal{E}_{\rho} is written as uj=1nργρ,i,j,ueρ,ju\mapsto\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle e_{\rho,j}. Then, regarding the Koopman operator Ug(y,)U_{g(y,\cdot)} on L2(𝒵)L^{2}(\mathcal{Z}), we have

Ug(y,)Γρ,iΓρ,iu(z)=Ug(y,)j=1nργρ,i,j,uγρ,i,j(z)=j=1nργρ,i,j,uγρ,i,j(g~(y)z)\displaystyle U_{g(y,\cdot)}\Gamma_{\rho,i}\Gamma_{\rho,i}^{*}u(z)=U_{g(y,\cdot)}\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\gamma_{\rho,i,j}(z)=\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\gamma_{\rho,i,j}(\tilde{g}(y)z)
=j=1nργρ,i,j,ueρ,i,ρ(zg~(y))eρ,j=j=1nργρ,i,j,uρ(z)eρ,i,ρ(g~(y))eρ,j\displaystyle\qquad=\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\left\langle e_{\rho,i},\rho(z\tilde{g}(y))e_{\rho,j}\right\rangle=\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\left\langle\rho(z)^{*}e_{\rho,i},\rho(\tilde{g}(y))e_{\rho,j}\right\rangle
=j=1nργρ,i,j,uρ(z)eρ,i,k=1nρeρ,k,ρ(g~(y))eρ,jeρ,k\displaystyle\qquad=\sum_{j=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\left\langle\rho(z)^{*}e_{\rho,i},\sum_{k=1}^{n_{\rho}}\left\langle e_{\rho,k},\rho(\tilde{g}(y))e_{\rho,j}\right\rangle e_{\rho,k}\right\rangle
=j,k=1nργρ,i,j,uρ(g~(y))eρ,k,eρ,jγρ,i,k(z)\displaystyle\qquad=\sum_{j,k=1}^{n_{\rho}}\left\langle\gamma_{\rho,i,j},u\right\rangle\left\langle\rho(\tilde{g}(y))^{*}e_{\rho,k},e_{\rho,j}\right\rangle\gamma_{\rho,i,k}(z)
=k=1nρρ(g~(y))eρ,k,Γρ,iuγρ,i,k(z)=(Γρ,ik=1nρρ(g~(y))eρ,k,Γρ,iueρ,k)(z)\displaystyle\qquad=\sum_{k=1}^{n_{\rho}}\left\langle\rho(\tilde{g}(y))^{*}e_{\rho,k},\Gamma_{\rho,i}^{*}u\right\rangle\gamma_{\rho,i,k}(z)=\bigg{(}\Gamma_{\rho,i}\sum_{k=1}^{n_{\rho}}\left\langle\rho(\tilde{g}(y))^{*}e_{\rho,k},\Gamma_{\rho,i}^{*}u\right\rangle e_{\rho,k}\bigg{)}(z)
=Γρ,iρ(g~(y))Γρ,iu(z)\displaystyle\qquad=\Gamma_{\rho,i}\rho(\tilde{g}(y))\Gamma_{\rho,i}^{*}u(z)

for uL2(𝒵)u\in L^{2}(\mathcal{Z}), z𝒵z\in\mathcal{Z}, and i=1,,nρi=1,\ldots,n_{\rho}. Thus, we have Ug(y,)=[ρ]𝒵^i=1nρΓρ,iρ(g~(y))Γρ,iU_{g(y,\cdot)}=\sum_{[\rho]\in\hat{\mathcal{Z}}}\sum_{i=1}^{n_{\rho}}\Gamma_{\rho,i}\rho(\tilde{g}(y))\Gamma_{\rho,i}^{*}. Therefore, the range of Γρ,i\Gamma_{\rho,i} is an invariant subspace of Ug(y,)U_{g(y,\cdot)} for any y𝒴y\in\mathcal{Y}. Thus, we set 𝒱[ρ],j\mathcal{V}_{[\rho],j} as the constant map which takes its value the range of Γρ,j\Gamma_{\rho,j}, and apply Proposition 8. In this case, the multiplication operator M^i,[ρ],j\hat{M}_{i,[\rho],j} is calculated as (wM^i,[ρ],j)(y)=w(y)Γρ,jρ(g~(hi(y)))Γρ,j(w\cdot\hat{M}_{i,[\rho],j})(y)=w(y)\Gamma_{\rho,j}\rho(\tilde{g}(h^{i}(y)))\Gamma_{\rho,j}^{*}, and by Proposition 10, its spectrum is calculated as

σ(M^i,[ρ],j)\displaystyle\sigma(\hat{M}_{i,[\rho],j}) ={λϵ>0,μ({y𝒴λσϵ(Γρ,jρ(g~(hi(y)))Γρ,j)})>0}\displaystyle=\{\lambda\in\mathbb{C}\;\mid\;^{\forall}\epsilon>0,\ \mu(\{y\in\mathcal{Y}\;\mid\;\lambda\in\sigma_{\epsilon}(\Gamma_{\rho,j}\rho(\tilde{g}(h^{i}(y)))\Gamma_{\rho,j}^{*})\})>0\}

Note that since ρ(g~(hi(y)))\rho(\tilde{g}(h^{i}(y))) is a linear operator on a finite dimensional space, it has only point spectra. By Corollary 9, we obtain a discrete decomposition of KT|𝒲K_{T}|_{\mathcal{W}} with the multiplication operators M^i,[ρ],j\hat{M}_{i,[\rho],j}. Let p^[ρ],j=Γρ,jΓρ,j\hat{p}_{[\rho],j}=\Gamma_{\rho,j}\Gamma_{\rho,j}^{*}. Then, KTK_{T} maps L2(𝒴)𝒜^[ρ],jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{[\rho],j} to L2(𝒴)𝒜^[ρ],jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{[\rho],j}, where 𝒜^[ρ],j={p^[ρ],jap^[ρ],ja𝒜}\hat{\mathcal{A}}_{[\rho],j}=\{\hat{p}_{[\rho],j}a\hat{p}_{[\rho],j}\;\mid\;a\in\mathcal{A}\}. Since 𝒱[ρ],j\mathcal{V}_{[\rho],j} is a finite dimensional space, by Proposition 13, the action of KTK_{T} restricted to L2(𝒴)𝒜^[ρ],jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{[\rho],j} is reduced to that of i=1nρUT\otimes_{i=1}^{n_{\rho}}U_{T} on i=1nρ(L2(𝒳))\otimes_{i=1}^{n_{\rho}}(L^{2}(\mathcal{X})) as KT=λj,[ρ](i=1nρUT)λ[ρ],j1K_{T}=\lambda_{j,[\rho]}(\otimes_{i=1}^{n_{\rho}}U_{T})\lambda_{[\rho],j}^{-1}, where λ[ρ],j(va)=(vaγρ,j,1,,vaγρ,j,nρ)\lambda_{[\rho],j}(v\otimes a)=(v\otimes a\gamma_{\rho,j,1},\ldots,v\otimes a\gamma_{\rho,j,n_{\rho}}).

3.2 The case of 𝒵=\mathcal{Z}=\mathbb{Z}

Let 𝒵=\mathcal{Z}=\mathbb{Z} equipped with the counting measure. We set the map g:𝒴×𝒵𝒵g:\mathcal{Y}\times\mathcal{Z}\to\mathcal{Z} as g(y,z)=z+g~(y)g(y,z)=z+\tilde{g}(y), where g~:𝒴𝒵\tilde{g}:\mathcal{Y}\to\mathcal{Z} is a measurable map. For ii\in\mathbb{Z}, let ei:𝕋e_{i}:\mathbb{T}\to\mathbb{C} be defined as ei(ω)=e1iωe_{i}(\omega)=\mathrm{e}^{\sqrt{-1}i\omega}, where 𝕋=/2π\mathbb{T}=\mathbb{R}/2\pi\mathbb{Z}. Note that {eii}\{e_{i}\;\mid\;i\in\mathbb{Z}\} is an orthonormal basis of L2(𝕋)L^{2}(\mathbb{T}). In addition, for ii\in\mathbb{Z}, let γi:\gamma_{i}:\mathbb{Z}\to\mathbb{C} be defined as γi(i)=1\gamma_{i}(i)=1, γi(z)=0(zi)\gamma_{i}(z)=0\ (z\neq i). Note also that {γii}\{\gamma_{i}\;\mid\;i\in\mathbb{Z}\} is an orthonormal basis of L2()L^{2}(\mathbb{Z}). Let Γ:L2(𝕋)L2()\Gamma:L^{2}(\mathbb{T})\to L^{2}(\mathbb{Z}) be the linear operator defined as eiγie_{i}\mapsto\gamma_{i} for any ii\in\mathbb{Z} and let ϕy(ω)=e1g~(y)ω\phi_{y}(\omega)=\mathrm{e}^{\sqrt{-1}\tilde{g}(y)\omega}. Then, we have

ΓMϕyΓγi=ΓMϕyei=Γeiϕ=Γei+g~(y)=γi+g~(y)=Ug(y,)γi,\displaystyle\Gamma M_{\phi_{y}}\Gamma^{*}\gamma_{i}=\Gamma M_{\phi_{y}}e_{i}=\Gamma e_{i}\phi=\Gamma e_{i+\tilde{g}(y)}=\gamma_{i+\tilde{g}(y)}=U_{g(y,\cdot)}\gamma_{i},

where MϕyM_{\phi_{y}} is the multiplication operator on L2(𝕋)L^{2}(\mathbb{T}) defined as Mϕyu(ω)=u(ω)ϕy(ω)=u(ω)e1g~(y)ωM_{\phi_{y}}u(\omega)=u(\omega)\phi_{y}(\omega)=u(\omega)\mathrm{e}^{\sqrt{-1}\tilde{g}(y)\omega}. Thus, we have the spectral decomposition Ug(y,)=ω𝕋e1g~(y)ωdE(ω)U_{g(y,\cdot)}=\int_{\omega\in\mathbb{T}}\mathrm{e}^{\sqrt{-1}\tilde{g}(y)\omega}\mathrm{d}E(\omega), where EE is the spectral measure defined as E(Ω)=ΓMχΩΓE(\Omega)=\Gamma M_{\chi_{\Omega}}\Gamma^{*} for a Borel set Ω\Omega and χΩ\chi_{\Omega} is the characteristic function of Ω\Omega. Let T1,T2,T_{1},T_{2},\ldots be a sequence of countable disjoint subsets of 𝕋\mathbb{T} such that 𝕋=j=1Tj\mathbb{T}=\bigcup_{j=1}^{\infty}T_{j}. Then, the range of E(Tj)E(T_{j}) is an invariant subspace of Ug(y,)U_{g(y,\cdot)} for any y𝒴y\in\mathcal{Y}. Thus, we set 𝒱j\mathcal{V}_{j} as the constant map which takes its value the range of E(Tj)E(T_{j}), and apply Proposition 8. In this case, M^i,j\hat{M}_{i,j} is calculated as (wM^i,j)(y)=w(y)ΓMϕhi(y)MχTjΓ(w\cdot\hat{M}_{i,j})(y)=w(y)\Gamma M_{\phi_{h^{i}(y)}}M_{\chi_{T_{j}}}\Gamma^{*}. Let p^j=E(Tj)\hat{p}_{j}=E(T_{j}). Then, KTK_{T} maps L2(𝒴)𝒜^jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j} to L2(𝒴)𝒜^jL^{2}(\mathcal{Y})\otimes\hat{\mathcal{A}}_{j}. Since 𝒱j\mathcal{V}_{j} is an infinite dimensional space, we cannot reduce the action of KTK_{T} restricted to L2(𝒴)A^jL^{2}(\mathcal{Y})\otimes\hat{A}_{j} to that of UTU_{T} on a Hilbert space. However, by Corollary 9, we obtain a discrete decomposition of KT|𝒲K_{T}|_{\mathcal{W}} in the Hilbert CC^{*}-module even in this case of the spectral decomposition of Ug(y,)U_{g(y,\cdot)} is continuous.

4 Continuous-time systems

4.1 Skew product system and Koopman operator on Hilbert space

As in the Section 2, let 𝒴\mathcal{Y} and 𝒵\mathcal{Z} be separable measure spaces equipped with measures μ\mu and ν\nu, respectively and let 𝒳=𝒴×𝒵\mathcal{X}=\mathcal{Y}\times\mathcal{Z}, the direct product measure space of 𝒴\mathcal{Y} and 𝒵\mathcal{Z}. Let h:×𝒴𝒴h:\mathbb{R}\times\mathcal{Y}\to\mathcal{Y} be a map such that for any tt\in\mathbb{R}, h(t,)h(t,\cdot) is a measure preserving and invertible map on 𝒴\mathcal{Y}. Moreover, let g:×𝒳𝒵g:\mathbb{R}\times\mathcal{X}\to\mathcal{Z} be a map such that for any tt\in\mathbb{R}, g(t,,)g(t,\cdot,\cdot) is a measurable map from 𝒳\mathcal{X} to 𝒵\mathcal{Z} and for any y𝒴y\in\mathcal{Y}, g(t,y,)g(t,y,\cdot) is measure preserving and invertible on 𝒵\mathcal{Z}. Consider the following skew product flow on 𝒳\mathcal{X}:

Φ(t,y,z)=(h(t,y),g(t,y,z))\Phi(t,y,z)=(h(t,y),g(t,y,z))

that satisfies Φ(0,y,z)=(y,z)\Phi(0,y,z)=(y,z) and Φ(s,Φ(t,y,z))=Φ(s+t,y,z)\Phi(s,\Phi(t,y,z))=\Phi(s+t,y,z) for any s,ts,t\in\mathbb{R}, y𝒴y\in\mathcal{Y}, and z𝒵z\in\mathcal{Z}. We denote Φ(t,,)=Φt\Phi(t,\cdot,\cdot)=\Phi_{t}, h(t,)=hth(t,\cdot)=h_{t}, and g(t,,)=gtg(t,\cdot,\cdot)=g_{t}, respectively. For tt\in\mathbb{R}, we consider the Koopman operator UΦtU_{\Phi_{t}} on L2(𝒳)L^{2}(\mathcal{X}). Instead of UTU_{T} for discrete systems, we consider a family of Koopman operators {UΦt}t\{U_{\Phi_{t}}\}_{t\in\mathbb{R}} for continuous systems.

4.2 Operator on Hilbert CC^{*}-module related to the Koopman operator

Analogous to the case of discrete systems, we extend the Koopman operator UΦtU_{\Phi_{t}} to an operator on the Hilbert CC^{*}-module \mathcal{M}.

Definition 6.

For tt\in\mathbb{R}, we define a right 𝒜\mathcal{A}-linear operator KΦtK_{\Phi_{t}} on \mathcal{M} by

KΦt(va)(y)=v(ht(y))Ugt(y,)aK_{\Phi_{t}}(v\otimes a)(y)=v(h_{t}(y))U_{g_{t}(y,\cdot)}a

for vL2(𝒴)v\in L^{2}(\mathcal{Y}), a𝒜a\in\mathcal{A}, and y𝒴y\in\mathcal{Y}. Here, for x𝒴x\in\mathcal{Y}, Ugt(y,)U_{g_{t}(y,\cdot)} is the Koopman operator on L2(𝒵)L^{2}(\mathcal{Z}) with respect to the map gt(y,)g_{t}(y,\cdot).

Remark 1.

The operator family {KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} satisfies KΦsKΦt=KΦs+tK_{\Phi_{s}}K_{\Phi_{t}}=K_{\Phi_{s+t}} for any s,ts,t\in\mathbb{R} and KΦ0=IK_{\Phi_{0}}=I. However, it is not strongly continuous even for a simple case. Let 𝒵=/2π\mathcal{Z}=\mathbb{R}/2\pi\mathbb{Z} equipped with the normalized Haar measure on 𝒵\mathcal{Z}. Let gt(y,z)=z+tαg_{t}(y,z)=z+t\alpha for α0\alpha\neq 0. For v1v\equiv 1 and a=Ia=I, we have

KΦtvava2\displaystyle\|K_{\Phi_{t}}v\otimes a-v\otimes a\|_{\mathcal{M}}^{2} =y𝒴(Ugt(y,)I)(Ugt(y,)I)dμ(y)𝒜\displaystyle=\bigg{\|}\int_{y\in\mathcal{Y}}(U_{g_{t}(y,\cdot)}-I)^{*}(U_{g_{t}(y,\cdot)}-I)\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}
=y𝒴(2IUgt(y,)Ugt(y,))dμ(y)𝒜\displaystyle=\bigg{\|}\int_{y\in\mathcal{Y}}(2I-U_{g_{t}(y,\cdot)}-U_{g_{t}(y,\cdot)}^{*})\mathrm{d}\mu(y)\bigg{\|}_{\mathcal{A}}
=2IU~MtU~U~MtU~𝒜\displaystyle=\|2I-\tilde{U}^{*}M_{t}\tilde{U}-\tilde{U}^{*}M_{t}^{*}\tilde{U}\|_{\mathcal{A}}
=2IMtMt𝒜\displaystyle=\|2I-M_{t}-M_{t}^{*}\|_{\mathcal{A}}
=supn|2e1ntαe1ntα|,\displaystyle=\sup_{n\in\mathbb{Z}}|2-\mathrm{e}^{\sqrt{-1}nt\alpha}-\mathrm{e}^{-\sqrt{-1}nt\alpha}|,

where U~:L2(𝒵)L2()\tilde{U}:L^{2}(\mathcal{Z})\to L^{2}(\mathbb{Z}) the unitary operator defined as γiei\gamma_{i}\mapsto e_{i}, γi(z)=e1iz\gamma_{i}(z)=\mathrm{e}^{\sqrt{-1}iz}, and eie_{i} is the map on \mathbb{Z} defined as ei(i)=1e_{i}(i)=1 and ei(n)=0e_{i}(n)=0 for nin\neq i. Moreover, Mt:L2()L2()M_{t}:L^{2}(\mathbb{Z})\to L^{2}(\mathbb{Z}) is the multiplication operator with respect to the map ne1αtnn\mapsto\mathrm{e}^{\sqrt{-1}\alpha tn}. The third equality holds since

U~MtU~γi=U~e1αtiei=e1αtiγi=Ugt(y,)γi.\displaystyle\tilde{U}^{*}M_{t}\tilde{U}\gamma_{i}=\tilde{U}^{*}\mathrm{e}^{\sqrt{-1}\alpha ti}e_{i}=\mathrm{e}^{\sqrt{-1}\alpha ti}\gamma_{i}=U_{g_{t}(y,\cdot)}\gamma_{i}.

Let ϵ=|2e1αe1α|\epsilon=|2-\mathrm{e}^{\sqrt{-1}\alpha}-\mathrm{e}^{-\sqrt{-1}\alpha}|. For any δ>0\delta>0, let n0n_{0}\in\mathbb{Z} such that n01/δn_{0}\geq 1/\delta and let t=1/n0t=1/n_{0}. Then, we have

KΦtvava2\displaystyle\|K_{\Phi_{t}}v\otimes a-v\otimes a\|_{\mathcal{M}}^{2} |2e1n0tαe1n0tα|=ϵ.\displaystyle\geq|2-\mathrm{e}^{\sqrt{-1}n_{0}t\alpha}-\mathrm{e}^{-\sqrt{-1}n_{0}t\alpha}|=\epsilon.

We adopt the generator defined using a weaker topology than the topology of the Hilbert CC^{*}-module.

Definition 7 (Equicontinuous C0C_{0}-group [41]).

Let MM be a sequentially complete locally convex space and for any tt\in\mathbb{R}, let κt:MM\kappa_{t}:M\to M be a linear operator on MM which satisfies

  1. 1.

    κ0=I\kappa_{0}=I,

  2. 2.

    κsκt=κs+t\kappa_{s}\kappa_{t}=\kappa_{s+t} for any s,ts,t\in\mathbb{R},

  3. 3.

    limt0κtw=w\lim_{t\to 0}\kappa_{t}w=w for any wMw\in M,

  4. 4.

    For any continuous seminorm pp on MM, there exists a continuous seminorm qq such that p(κtw)q(w)p(\kappa_{t}w)\leq q(w) for any wMw\in M and tt\in\mathbb{R}.

The family {κt}t\{\kappa_{t}\}_{t\in\mathbb{R}} is called an equicontinuous C0C_{0}-group.

Proposition 14.

The space (L2(𝒵),L2(𝒳))\mathcal{M}\subseteq\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})) equipped with the strong operator topology is a sequentially complete locally convex space. In addition, assume 𝒴\mathcal{Y} and 𝒵\mathcal{Z} are locally compact Hausdorff spaces, μ\mu and ν\nu are regular probability measures, and hh and gg are continuous. Then, {KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} is an equicontinuous C0C_{0}-group.

To prove Proposition 14, we use the following lemma:

Lemma 15.

Let Ω\Omega and 𝒳\mathcal{X} be topological spaces. If a map Ψ:Ω×𝒳\Psi:\Omega\times\mathcal{X}\to\mathbb{C} is continuous and compactly supported, then the map ΩtΨ(t,)Cc(𝒳)\Omega\ni t\mapsto\Psi(t,\cdot)\in C_{c}(\mathcal{X}) is continuous. Here, Cc(𝒳)C_{c}(\mathcal{X}) is the space of compactly supported continuous functions on 𝒳\mathcal{X}.

Proof.

The statement follows from Lemma 4.16 by Eisner et al. [42]. ∎

Proof of Proposition 14.

aaa
(\mathcal{M} is a sequentially complete locally convex space) For pL2(𝒵)p\in L^{2}(\mathcal{Z}), let p:+\|\cdot\|_{p}:\mathcal{M}\to\mathbb{R}_{+} be defined as wp=wpL2(𝒳)\|w\|_{p}=\|wp\|_{L^{2}(\mathcal{X})} for w(L2(𝒵),L2(𝒳))w\in\mathcal{M}\subseteq\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})). Then, p\|\cdot\|_{p} is a seminorm in \mathcal{M}. Moreover, let {wi}i\{w_{i}\}_{i\in\mathbb{N}} be a countable Cauchy sequence in \mathcal{M}. Then, for any vL2(𝒵)v\in L^{2}(\mathcal{Z}), {wiv}i\{w_{i}v\}_{i\in\mathbb{N}} is a Cauchy sequence in the Hilbert space L2(𝒳)L^{2}(\mathcal{X}). Thus, there exists w~L2(𝒳)\tilde{w}\in L^{2}(\mathcal{X}) such that limiwiv=w~\lim_{i\to\infty}w_{i}v=\tilde{w}. Let w:L2(𝒵)L2(𝒳)w:L^{2}(\mathcal{Z})\to L^{2}(\mathcal{X}) be the map defined as w:vw~w:v\mapsto\tilde{w}. Then, ww is linear and

wvL2(𝒳)=limiwivL2(𝒳)supiwivL2(𝒳)supiwivL2(𝒵)\displaystyle\|wv\|_{L^{2}(\mathcal{X})}=\|\lim_{i\to\infty}w_{i}v\|_{L^{2}(\mathcal{X})}\leq\sup_{i\in\mathbb{N}}\|w_{i}v\|_{L^{2}(\mathcal{X})}\leq\sup_{i\in\mathbb{N}}\|w_{i}\|_{\mathcal{M}}\,\|v\|_{L^{2}(\mathcal{Z})}

for vL2(𝒵)v\in L^{2}(\mathcal{Z}). By the uniform boundedness principle, supiwi<\sup_{i\in\mathbb{N}}\|w_{i}\|<\infty. Thus, w(L2(𝒵),L2(𝒳))w\in\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})). Since (L2(𝒵),L2(𝒳))\mathcal{M}\subseteq\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})) is closed with respect to the strong operator topology, we obtain ww\in\mathcal{M}. Therefore, {wi}i\{w_{i}\}_{i\in\mathbb{N}} converges to ww in \mathcal{M}.
({KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} is an equicontinuous C0C_{0}-group) For any vL2(𝒴)v\in L^{2}(\mathcal{Y}), a𝒜a\in\mathcal{A}, and uL2(𝒵)u\in L^{2}(\mathcal{Z}), we have

(KΦtva)uL2(𝒳)2\displaystyle\|(K_{\Phi_{t}}v\otimes a)u\|_{L^{2}(\mathcal{X})}^{2} =y𝒴z𝒵|v(ht(y))(au)(gt(y,z))|2dν(z)dμ(y)\displaystyle=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}|v(h_{t}(y))(au)(g_{t}(y,z))|^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|v(y)(au)(z)|2dν(z)dμ(y)=(va)uL2(𝒳),\displaystyle=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}|v(y)(au)(z)|^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)=\|(v\otimes a)u\|_{L^{2}(\mathcal{X})},

which shows that the condition 4 of Definition 7 is satisfied.

Regarding the condition 3, let ϵ>0\epsilon>0, let {γi}i=1\{\gamma_{i}\}_{i=1}^{\infty} be an orthonormal basis of L2(𝒵)L^{2}(\mathcal{Z}), and let 𝒟={iFciγiF: finite, ci}\mathcal{D}=\{\sum_{i\in F}c_{i}\gamma_{i}\,\mid\,F\subset\mathbb{Z}:\mbox{ finite, }c_{i}\in\mathbb{C}\}. Since Cc(𝒵)C_{c}(\mathcal{Z}), Cc(𝒴)C_{c}(\mathcal{Y}), and 𝒟\mathcal{D} are dense in L2(𝒵)L^{2}(\mathcal{Z}), L2(𝒴)L^{2}(\mathcal{Y}), and L2(𝒵)L^{2}(\mathcal{Z}), respectively, for any ii\in\mathbb{N} and any vL2(𝒴)v\in L^{2}(\mathcal{Y}), a𝒜a\in\mathcal{A}, and uL2(𝒵)u\in L^{2}(\mathcal{Z}), there exist v~Cc(𝒴)\tilde{v}\in C_{c}(\mathcal{Y}), γ~iCc(𝒵)\tilde{\gamma}_{i}\in C_{c}(\mathcal{Z}), and u~𝒟\tilde{u}\in\mathcal{D} such that v~vL2(𝒴)ϵ\|\tilde{v}-v\|_{L^{2}(\mathcal{Y})}\leq\epsilon, γ~iaγiL2(𝒵)ϵ/(2i)\|\tilde{\gamma}_{i}-a\gamma_{i}\|_{L^{2}(\mathcal{Z})}\leq\epsilon/(\sqrt{2}^{i}), and u~uL2(𝒵)ϵ\|\tilde{u}-u\|_{L^{2}(\mathcal{Z})}\leq\epsilon. Let a~=i=1γ~iγi\tilde{a}=\sum_{i=1}^{\infty}\tilde{\gamma}_{i}\gamma_{i}^{\prime}, where the limit is taken with respect to the strong operator topology. The operator a~\tilde{a} is bounded since we have

a~uL2(𝒵)auL2(𝒵)+a~uauL2(𝒵)a𝒜uL2(𝒵)+i=1(aγiγiuγ~iγiu)L2(𝒵)\displaystyle\|\tilde{a}u\|_{L^{2}(\mathcal{Z})}\leq\|au\|_{L^{2}(\mathcal{Z})}+\|\tilde{a}u-au\|_{L^{2}(\mathcal{Z})}\leq\|a\|_{\mathcal{A}}\|u\|_{L^{2}(\mathcal{Z})}+\bigg{\|}\sum_{i=1}^{\infty}(a\gamma_{i}\gamma_{i}^{\prime}u-\tilde{\gamma}_{i}\gamma_{i}^{\prime}u)\bigg{\|}_{L^{2}(\mathcal{Z})}
=a𝒜uL2(𝒵)+i=1(γiu)(aγiγ~i)L2(𝒵)\displaystyle\qquad=\|a\|_{\mathcal{A}}\|u\|_{L^{2}(\mathcal{Z})}+\bigg{\|}\sum_{i=1}^{\infty}(\gamma_{i}^{\prime}u)(a\gamma_{i}-\tilde{\gamma}_{i})\bigg{\|}_{L^{2}(\mathcal{Z})}
a𝒜uL2(𝒵)+(i=1|γiu|2)1/2(i=1ϵ22i)1/2=uL2(𝒵)(a𝒜+ϵ).\displaystyle\qquad\leq\|a\|_{\mathcal{A}}\|u\|_{L^{2}(\mathcal{Z})}+\bigg{(}\sum_{i=1}^{\infty}|\gamma_{i}^{\prime}u|^{2}\bigg{)}^{1/2}\bigg{(}\sum_{i=1}^{\infty}\frac{\epsilon^{2}}{2^{i}}\bigg{)}^{1/2}=\|u\|_{L^{2}(\mathcal{Z})}(\|a\|_{\mathcal{A}}+\epsilon).

Thus, we have a~𝒜\tilde{a}\in\mathcal{A}. In addition, we have

(KΦtv~a~)u~(v~a~)u~L2(𝒳)2\displaystyle\|(K_{\Phi_{t}}\tilde{v}\otimes\tilde{a})\tilde{u}-(\tilde{v}\otimes\tilde{a})\tilde{u}\|_{L^{2}(\mathcal{X})}^{2}
=y𝒴z𝒵|v~(ht(y))(Ugt(y,)a~u~)(z)v~(y)(a~u~)(z)|2dν(z)dμ(y)\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}|\tilde{v}(h_{t}(y))(U_{g_{t}(y,\cdot)}\tilde{a}\tilde{u})(z)-\tilde{v}(y)(\tilde{a}\tilde{u})(z)|^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|v~(ht(y))(a~u~)(gt(y,z))v~(y)(a~u~)(z)|2dν(z)dμ(y).\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}|\tilde{v}(h_{t}(y))(\tilde{a}\tilde{u})(g_{t}(y,z))-\tilde{v}(y)(\tilde{a}\tilde{u})(z)|^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y).

Let Ψ:×𝒴×𝒵\Psi:\mathbb{R}\times\mathcal{Y}\times\mathcal{Z}\to\mathbb{C} be defined as (t,y,z)v~(ht(y))(a~u~)(gt(y,z))v~(y)(a~u~)(z)(t,y,z)\mapsto\tilde{v}(h_{t}(y))(\tilde{a}\tilde{u})(g_{t}(y,z))-\tilde{v}(y)(\tilde{a}\tilde{u})(z). Since Ψ\Psi is continuous, by Lemma 15, the map tΨ(t,,)Cc(𝒳)t\mapsto\Psi(t,\cdot,\cdot)\in C_{c}(\mathcal{X}) is also continuous. Thus, we have

limt0(KΦtv~a~)u~(v~a~)u~L2(𝒳)\displaystyle\lim_{t\to 0}\|(K_{\Phi_{t}}\tilde{v}\otimes\tilde{a})\tilde{u}-(\tilde{v}\otimes\tilde{a})\tilde{u}\|_{L^{2}(\mathcal{X})} limt0(KΦtv~a~)u~(v~a~)u~L(𝒳)\displaystyle\leq\lim_{t\to 0}\|(K_{\Phi_{t}}\tilde{v}\otimes\tilde{a})\tilde{u}-(\tilde{v}\otimes\tilde{a})\tilde{u}\|_{L^{\infty}(\mathcal{X})}
=limt0Ψt=0,\displaystyle=\lim_{t\to 0}\|\Psi_{t}\|_{\infty}=0,

where \|\cdot\|_{\infty} is the sup norm in Cc(𝒳)C_{c}(\mathcal{X}). Therefore, limt0(KΦtva)u(va)uL2(𝒳)=0\lim_{t\to 0}\|(K_{\Phi_{t}}{v}\otimes{a}){u}-({v}\otimes{a}){u}\|_{L^{2}(\mathcal{X})}=0. Indeed, we have

(KΦtva)u(KΦtv~a~)u~L2(𝒳)=(va)u(v~a~)u~L2(𝒳)\displaystyle\|(K_{\Phi_{t}}v\otimes a)u-(K_{\Phi_{t}}\tilde{v}\otimes\tilde{a})\tilde{u}\|_{L^{2}(\mathcal{X})}=\|(v\otimes a)u-(\tilde{v}\otimes\tilde{a})\tilde{u}\|_{L^{2}(\mathcal{X})}
(v~(aa~))u~L2(𝒳)+((vv~)a)u~L2(𝒳)+(va)(uu~)L2(𝒳)\displaystyle\qquad\leq\|(\tilde{v}\otimes(a-\tilde{a}))\tilde{u}\|_{L^{2}(\mathcal{X})}+\|((v-\tilde{v})\otimes a)\tilde{u}\|_{L^{2}(\mathcal{X})}+\|(v\otimes a)(u-\tilde{u})\|_{L^{2}(\mathcal{X})}
v~L2(𝒴)(aa~)u~L2(𝒵)+vv~L2(𝒴)au~L2(𝒵)+vL2(𝒴)a𝒜uu~L2(𝒵)\displaystyle\qquad\leq\|\tilde{v}\|_{L^{2}(\mathcal{Y})}\|(a-\tilde{a})\tilde{u}\|_{L^{2}(\mathcal{Z})}+\|v-\tilde{v}\|_{L^{2}(\mathcal{Y})}\|{a}\tilde{u}\|_{L^{2}(\mathcal{Z})}+\|v\|_{L^{2}(\mathcal{Y})}\|a\|_{\mathcal{A}}\|u-\tilde{u}\|_{L^{2}(\mathcal{Z})}
(vL2(𝒴)+ϵ)uL2(𝒵)ϵ+ϵauL2(𝒵)+vL2(𝒴)a𝒜ϵ.\displaystyle\qquad\leq(\|v\|_{L^{2}(\mathcal{Y})}+\epsilon)\|u\|_{L^{2}(\mathcal{Z})}\epsilon+\epsilon\|au\|_{L^{2}(\mathcal{Z})}+\|v\|_{L^{2}(\mathcal{Y})}\|a\|_{\mathcal{A}}\epsilon.

As a result, {KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} satisfies the condition 3 of Definition 7. ∎

Definition 8.

The generator LΦL_{\Phi} of {KΦt}t\{K_{\Phi_{t}}\}_{t\in\mathbb{R}} is defined as

LΦw=limt0KΦtwwt,L_{\Phi}w=\lim_{t\to 0}\frac{K_{\Phi_{t}}w-w}{t},

where the limit is with respect to the strong operator topology in \mathcal{M}.

Proposition 16 (Choe, 1985 [41]).

The generator LΦL_{\Phi} is a densely defined linear operator in \mathcal{M} with respect to the strong operator topology.

4.3 Decomposition of KΦtK_{\Phi_{t}} and LΦL_{\Phi}

We derive the eigenoperator decomposition for continuous systems. In the following, we assume 𝒴\mathcal{Y} and 𝒵\mathcal{Z} are differentiable manifolds, μ\mu and ν\nu are regular probability measures, and hh and gg are differentiable.

4.3.1 Fundamental decomposition

We first define vectors to decompose the operator KΦtK_{\Phi_{t}} using the cocycle.

Definition 9.

For ss\in\mathbb{R}, we define a linear operator ws:L2(𝒵)L2(𝒳)w_{s}:L^{2}(\mathcal{Z})\to L^{2}(\mathcal{X}) as

(wsu)(y,z)=(Ugs(y,)u)(z).(w_{s}u)(y,z)=(U_{g_{s}(y,\cdot)}u)(z).
Proposition 17.

For ss\in\mathbb{R}, we have wsw_{s}\in\mathcal{M}. Moreover, KΦtws=ws+t=wsMs,tK_{\Phi_{t}}w_{s}=w_{s+t}=w_{s}\cdot M_{s,t}, where Ms,tM_{s,t} is a left 𝒜\mathcal{A}-linear multiplication operator on \mathcal{M} defined as (wMs,t)(y)=w(y)Ugt(hs(y),)(w\cdot M_{s,t})(y)=w(y)U_{g_{t}(h_{s}(y),\cdot)}.

Proof.

We obtain wsw_{s}\in\mathcal{M} by Lemma 30. Moreover, we have

KΦtws=Ugt(y,)Ugs(ht(y),)=Ugs(ht(y),gt(y,))=Ugs+t(y,)=Ugs(y,)Ugt(hs(y),).K_{\Phi_{t}}w_{s}=U_{g_{t}(y,\cdot)}U_{g_{s}(h_{t}(y),\cdot)}=U_{g_{s}(h_{t}(y),g_{t}(y,\cdot))}=U_{g_{s+t}(y,\cdot)}=U_{g_{s}(y,\cdot)}U_{g_{t}(h_{s}(y),\cdot)}.

Proposition 18.

For ss\in\mathbb{R} and uCc1(𝒵)u\in C_{c}^{1}(\mathcal{Z}), let w~s,u(y,z)=ugt(s,y,z)\tilde{w}_{s,u}(y,z)=\frac{\partial u\circ g}{\partial t}(s,y,z). Then, (LΦws)u=w~s,u(L_{\Phi}w_{s})u=\tilde{w}_{s,u} and LΦws=wsNsL_{\Phi}w_{s}=w_{s}\cdot N_{s}, where (wNs)(y)=w(y)(Mgt(0,hs(y),)z)(w\cdot N_{s})(y)=w(y)(M_{\frac{\partial g}{\partial t}(0,h_{s}(y),\cdot)}\frac{\partial}{\partial z}). Here, Cc1(𝒵)C_{c}^{1}(\mathcal{Z}) is the space of compactly supported and continuously differentiable functions on 𝒵\mathcal{Z}.

Proof.

For uCc1(𝒵)u\in C_{c}^{1}(\mathcal{Z}), we have

1t(KΦtwsws)uw~s,uL2(𝒳)2\displaystyle\bigg{\|}\frac{1}{t}(K_{\Phi_{t}}w_{s}-w_{s})u-\tilde{w}_{s,u}\bigg{\|}^{2}_{L^{2}(\mathcal{X})}
=y𝒴z𝒵|1t(Ugt(y,)Ugs(ht(y),)Ugs(y,))u(z)w~s,u(y,z)|2dν(z)dμ(y)\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}(U_{g_{t}(y,\cdot)}U_{g_{s}(h_{t}(y),\cdot)}-U_{g_{s}(y,\cdot)})u(z)-\tilde{w}_{s,u}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|1t(Ugs+t(y,)Ugs(y,))u(z)w~s,u(y,z)|2dν(z)dμ(y)\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}(U_{g_{s+t}(y,\cdot)}-U_{g_{s}(y,\cdot)})u(z)-\tilde{w}_{s,u}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|1t(u(gs+t(y,z))u(gs(y,z)))w~s,u(y,z)|2dν(z)dμ(y).\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}(u(g_{s+t}(y,z))-u(g_{s}(y,z)))-\tilde{w}_{s,u}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y).

Since gg is continuous, there exists D>0D>0 such that for any ss\in\mathbb{R}, y𝒴y\in\mathcal{Y}, and z𝒵z\in\mathcal{Z}, |ugt(s,y,z)|<D|\frac{\partial u\circ g}{\partial t}(s,y,z)|<D. By the mean-value theorem, for any y𝒴y\in\mathcal{Y} and z𝒵z\in\mathcal{Z}, there exists c(s,s+t)c\in(s,s+t) for t>0t>0 or c(s+t,s)c\in(s+t,s) for t<0t<0 such that

|1t(u(gs+t(y,z))u(gs(y,z)))|=|ugt(c,y,z)|D.\bigg{|}\frac{1}{t}(u(g_{s+t}(y,z))-u(g_{s}(y,z)))\bigg{|}=\bigg{|}\frac{\partial u\circ g}{\partial t}(c,y,z)\bigg{|}\leq D.

Thus, by the Lebesgue’s dominated convergence theorem, we obtain

limt01t(KΦtwsws)uw~s,uL2(𝒳)2\displaystyle\lim_{t\to 0}\bigg{\|}\frac{1}{t}(K_{\Phi_{t}}w_{s}-w_{s})u-\tilde{w}_{s,u}\bigg{\|}^{2}_{L^{2}(\mathcal{X})}
=y𝒴z𝒵limt0|1t(u(gs+t(y,z))u(gs(y,z)))w~s,u(y,z)|2dν(z)dμ(y)=0.\displaystyle\qquad=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\lim_{t\to 0}\bigg{|}\frac{1}{t}(u(g_{s+t}(y,z))-u(g_{s}(y,z)))-\tilde{w}_{s,u}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)=0.

Thus, we have (LΦws)u=w~s,u(L_{\Phi}w_{s})u=\tilde{w}_{s,u}. Moreover, w~s,u\tilde{w}_{s,u} is represented as

w~s,u(y,z)\displaystyle\tilde{w}_{s,u}(y,z) =ugt(s,y,z)=uz(g(s,y,z))gt(s,y,z)\displaystyle=\frac{\partial u\circ g}{\partial t}(s,y,z)=\frac{\partial u}{\partial z}(g(s,y,z))\frac{\partial g}{\partial t}(s,y,z)
=(Ugs(y,)uz)(z)gt(s,y,z)=Ugs(y,)Mgt(s,y,gs1(y,))zu(z).\displaystyle=\bigg{(}U_{g_{s}(y,\cdot)}\frac{\partial u}{\partial z}\bigg{)}(z)\frac{\partial g}{\partial t}(s,y,z)=U_{g_{s}(y,\cdot)}M_{\frac{\partial g}{\partial t}(s,y,g_{s}^{-1}(y,\cdot))}\frac{\partial}{\partial z}u(z).

Since gs(y,gs(hs(y),z))=gs(hs(hs(y)),gs(hs(y),z))=g0(hs(y),z)=zg_{s}(y,g_{-s}(h_{s}(y),z))=g_{s}(h_{-s}(h_{s}(y)),g_{-s}(h_{s}(y),z))=g_{0}(h_{s}(y),z)=z, gs(y,)1=gs(hs(y),)g_{s}(y,\cdot)^{-1}=g_{-s}(h_{s}(y),\cdot). Thus, we have

gt(s,y,gs1(y,z))=gt(s,hs(hs(y)),gs(hs(y),z))=gt(0,hs(y),z).\displaystyle\frac{\partial g}{\partial t}(s,y,g_{s}^{-1}(y,z))=\frac{\partial g}{\partial t}(s,h_{-s}(h_{s}(y)),g_{-s}(h_{s}(y),z))=\frac{\partial g}{\partial t}(0,h_{s}(y),z).

The vectors wsw_{s} describe the dynamics on 𝒵\mathcal{Z}, which is specific for the skew product dynamical systems and we are interested in.

Proposition 19.

The action of the Koopman operator UΦtU_{\Phi_{t}} is decomposed into two parts as

UΦt(vu)=UhtvUΦsugtsU_{\Phi_{t}}(v\otimes u)=U_{h_{t}}v\otimes U_{\Phi_{s}}u\circ g_{t-s}

for vL2(𝒴)v\in L^{2}(\mathcal{Y}), uL2(𝒵)u\in L^{2}(\mathcal{Z}), and s,ts,t\in\mathbb{R}.

Proof.

By the definition of UΦtU_{\Phi_{t}}, we have

UΦt(vu)(y,z)\displaystyle U_{\Phi_{t}}(v\otimes u)(y,z) =v(ht(y))u(gt(y,z))=v(ht(y))u(gts(hs(y),gs(y,z)))\displaystyle=v(h_{t}(y))u(g_{t}(y,z))=v(h_{t}(y))u(g_{t-s}(h_{s}(y),g_{s}(y,z)))
=Uhtv(y)UΦsugts(y,z).\displaystyle=U_{h_{t}}v(y)U_{\Phi_{s}}u\circ g_{t-s}(y,z).

Let

𝒲0={sFwscsF:finite set,cs𝒜}\mathcal{W}_{0}=\bigg{\{}\sum_{s\in F}w_{s}c_{s}\,\mid\,F\subseteq\mathbb{R}:\ \mbox{finite set},\ c_{s}\in\mathcal{A}\bigg{\}}

and 𝒲\mathcal{W} be the completion of 𝒲0\mathcal{W}_{0} with respect to the norm in \mathcal{M} (𝒲\mathcal{W} is a submodule of \mathcal{M} and Hilbert 𝒜\mathcal{A}-module). Moreover, for ss\in\mathbb{R} and uL2(𝒵)u\in L^{2}(\mathcal{Z}), let w~u,sL2(𝒳)\tilde{w}_{u,s}\in L^{2}(\mathcal{X}) be defined as w~u,s(y,z)=u(gs(y,z))\tilde{w}_{u,s}(y,z)=u(g_{s}(y,z)). Let

𝒲~0={j=1nsFcsw~uj,sn,F:finite set,cs,ujL2(𝒵)}\tilde{\mathcal{W}}_{0}=\bigg{\{}\sum_{j=1}^{n}\sum_{s\in F}c_{s}\tilde{w}_{u_{j},s}\,\mid\,n\in\mathbb{N},\ F\subseteq\mathbb{R}:\ \mbox{finite set},\ c_{s}\in\mathbb{C},\ u_{j}\in L^{2}(\mathcal{Z})\bigg{\}}

and 𝒲~\tilde{\mathcal{W}} be the completion of 𝒲~0\tilde{\mathcal{W}}_{0} with respect to the norm in L2(𝒳)L^{2}(\mathcal{X}).

Proposition 20.

With the notation defined in Proposition 2, we have KΦt|𝒲ιi|𝒲~=ιiUΦt|𝒲~K_{\Phi_{t}}|_{\mathcal{W}}\,\iota_{i}|_{\tilde{\mathcal{W}}}=\iota_{i}U_{\Phi_{t}}|_{\tilde{\mathcal{W}}} for tt\in\mathbb{R} and i=1,2,i=1,2,\ldots.

Proof.

For uL2(𝒵)u\in L^{2}(\mathcal{Z}), ss\in\mathbb{R}, and i>0i>0, we have

(ιiw~u,s)(y)=u(gs(y,))γi=Ugs(y,)uγi=ws(y)(uγi).\displaystyle(\iota_{i}\tilde{w}_{u,s})(y)=u(g_{s}(y,\cdot))\gamma_{i}^{\prime}=U_{g_{s}(y,\cdot)}u\gamma_{i}^{\prime}=w_{s}(y)(u\gamma_{i}^{\prime}).

Thus, we obtain ιiw~u,s𝒲\iota_{i}\tilde{w}_{u,s}\in\mathcal{W}. Therefore, the range of ιi|𝒲~\iota_{i}|_{\tilde{\mathcal{W}}} is contained in 𝒲\mathcal{W}. The equality is deduced by the definitions of KΦtK_{\Phi_{t}} and UΦtU_{\Phi_{t}}. ∎

4.3.2 Further decomposition

We further decompose wsw_{s} and NsN_{s} and obtain a more detailed decomposition of LΦ|𝒲L_{\Phi}|_{\mathcal{W}}. For y𝒴y\in\mathcal{Y}, let 𝒱1(y),𝒱2(y),\mathcal{V}_{1}(y),\mathcal{V}_{2}(y),\ldots be a sequence of closed subspaces of L2(𝒵)L^{2}(\mathcal{Z}) which satisfies L2(𝒵)=Span{j=1𝒱j(y)}¯L^{2}(\mathcal{Z})=\overline{\operatorname{Span}\{\bigcup_{j=1}^{\infty}\mathcal{V}_{j}(y)\}} for a.s. y𝒴y\in\mathcal{Y}. For ss\in\mathbb{R} and j=1,2,j=1,2,\ldots, we define a linear map w^s,j\hat{w}_{s,j} from L2(𝒵)L^{2}(\mathcal{Z}) to L2(𝒳)L^{2}(\mathcal{X}) as (w^s,ju)(y,z)=(ws(y)pj(hs(y))u)(y,z)(\hat{w}_{s,j}u)(y,z)=(w_{s}(y)p_{j}(h_{s}(y))u)(y,z), where pj(y):L2(𝒵)𝒱j(y)p_{j}(y):L^{2}(\mathcal{Z})\to\mathcal{V}_{j}(y) is the projection onto 𝒱j(y)\mathcal{V}_{j}(y). Assume pj(y)p_{j}(y) satisfies (y,z)(pj(y)u)(z)L2(𝒳)(y,z)\mapsto(p_{j}(y)u)(z)\in L^{2}(\mathcal{X}). We denote by pjp_{j} the linear operator from L2(𝒵)L2(𝒳)L^{2}(\mathcal{Z})\to L^{2}(\mathcal{X}) defined as pju(y,z)=(pj(y)u)(z)p_{j}u(y,z)=(p_{j}(y)u)(z). For each j=1,2,j=1,2,\ldots, the following proposition holds. Here, we define a differential operator VΦV_{\Phi} by

VΦv(y,z)=vy(y,z)ht(0,y)+vz(y,z)gt(0,y,z)\displaystyle V_{\Phi}v(y,z)=\frac{\partial v}{\partial y}(y,z)\frac{\partial h}{\partial t}(0,y)+\frac{\partial v}{\partial z}(y,z)\frac{\partial g}{\partial t}(0,y,z) (6)

for vCc1(𝒳)v\in C^{1}_{c}(\mathcal{X}).

Theorem 21 (Eigenoperator decomposition for continuous-time systems).

For ss\in\mathbb{R} and uL2(𝒵)u\in L^{2}(\mathcal{Z}), let

w~s,u,j(y,z)=(pj(ht(y))ug(t,y,z))t|t=s.\displaystyle\tilde{w}_{s,u,j}(y,z)=\frac{\partial(p_{j}(h_{t}(y))u\circ g(t,y,z))}{\partial t}\bigg{|}_{t=s}.

Assume for any upj1(Cc1(𝒳))u\in p_{j}^{-1}(C_{c}^{1}(\mathcal{X})) and any y𝒴y\in\mathcal{Y}, (pj(ht(y))u)(gt(y,))t|t=0=(VΦpju)(y,)𝒱j(y)\frac{\partial(p_{j}(h_{t}(y))u)(g_{t}(y,\cdot))}{\partial t}\big{|}_{t=0}=(V_{\Phi}p_{j}u)(y,\cdot)\in\mathcal{V}_{j}(y). Then, (LΦw^s,j)u=w~s,u,j=(w^s,jN^s,j)u(L_{\Phi}\hat{w}_{s,j})u=\tilde{w}_{s,u,j}=(\hat{w}_{s,j}\cdot\hat{N}_{s,j})u, where N^s,j\hat{N}_{s,j} is defined as (wN^s,j)u(y,z)=w(y)(VΦpju)(hs(y),z)(w\cdot\hat{N}_{s,j})u(y,z)=w(y)(V_{\Phi}p_{j}u)(h_{s}(y),z).

We call N^s,j\hat{N}_{s,j} an eigenoperator and w^s,j\hat{w}_{s,j} an eigenvector.

Proof.

For upj1(Cc1(𝒳))u\in p_{j}^{-1}(C_{c}^{1}(\mathcal{X})), we have

1t(KΦtw^s,jw^s,j)uw~s,u,jL2(𝒳)2\displaystyle\bigg{\|}\frac{1}{t}(K_{\Phi_{t}}\hat{w}_{s,j}-\hat{w}_{s,j})u-\tilde{w}_{s,u,j}\bigg{\|}^{2}_{L^{2}(\mathcal{X})}
=y𝒴z𝒵|1t(Ugt(y,)Ugs(ht(y),)pj(hs+t(y))Ugs(y,)pj(hs(y)))u(z)w~s,u,j(y,z)|2dν(z)dμ(y)\displaystyle=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}(U_{g_{t}(y,\cdot)}U_{g_{s}(h_{t}(y),\cdot)}p_{j}(h_{s+t}(y))-U_{g_{s}(y,\cdot)}p_{j}(h_{s}(y)))u(z)-\tilde{w}_{s,u,j}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|1t(Ugs+t(y,)pj(hs+t(y))Ugs(y,)pj(hs(y)))u(z)w~s,u,j(y,z)|2dν(z)dμ(y)\displaystyle=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}(U_{g_{s+t}(y,\cdot)}p_{j}(h_{s+t}(y))-U_{g_{s}(y,\cdot)}p_{j}(h_{s}(y)))u(z)-\tilde{w}_{s,u,j}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
=y𝒴z𝒵|1t((pj(hs+t(y))u)(gs+t(y,z))(pj(hs(y))u)(gs(y,z)))w~s,u,j(y,z)|2dν(z)dμ(y)\displaystyle=\int_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}\bigg{|}\frac{1}{t}((p_{j}(h_{s+t}(y))u)(g_{s+t}(y,z))-(p_{j}(h_{s}(y))u)(g_{s}(y,z)))-\tilde{w}_{s,u,j}(y,z)\bigg{|}^{2}\mathrm{d}\nu(z)\mathrm{d}\mu(y)
0(t0).\displaystyle\to 0\ (t\to 0).

Thus, we have (LΦw^s,j)u=w~s,u,j(L_{\Phi}\hat{w}_{s,j})u=\tilde{w}_{s,u,j}. Moreover, w~s,u,j\tilde{w}_{s,u,j} is represented as

w~s,u,j(y,z)\displaystyle\tilde{w}_{s,u,j}(y,z) =t(Ugt(y,)pj(ht(y))u)(z)|t=s=Ugs(y,)t(Ugt(hs(y),)pj(hs+t(y))u)(z)|t=0\displaystyle=\frac{\partial}{\partial t}(U_{g_{t}(y,\cdot)}p_{j}(h_{t}(y))u)(z)\bigg{|}_{t=s}=U_{g_{s}(y,\cdot)}\frac{\partial}{\partial t}(U_{g_{t}(h_{s}(y),\cdot)}p_{j}(h_{s+t}(y))u)(z)\bigg{|}_{t=0}
=Ugs(y,)(pj(ht(hs(y)))u)(g(t,hs(y),z))t|t=0\displaystyle=U_{g_{s}(y,\cdot)}\frac{\partial(p_{j}(h_{t}(h_{s}(y)))u)(g(t,h_{s}(y),z))}{\partial t}\bigg{|}_{t=0}
=Ugs(y,)pj(hs(y))(ps,j(ht(y))u)(g(t,hs(y),z))t|t=0\displaystyle=U_{g_{s}(y,\cdot)}p_{j}(h_{s}(y))\frac{\partial(p_{s,j}(h_{t}(y))u)(g(t,h_{s}(y),z))}{\partial t}\bigg{|}_{t=0}
=w^s,j(y)((ps,j(y)u)(z)y|y=h(0,y),z=g(0,hs(y),z)h(t,y)t|t=0\displaystyle=\hat{w}_{s,j}(y)\bigg{(}\frac{\partial(p_{s,j}(y)u)(z)}{\partial y}\bigg{|}_{\begin{subarray}{c}y=h(0,y),\\ z=g(0,h_{s}(y),z)\end{subarray}}\frac{\partial h(t,y)}{\partial t}\bigg{|}_{t=0}
+(ps,j(y)u)(z)z|y=h(0,y),z=g(0,hs(y),z)g(t,hs(y),z)t|t=0)\displaystyle\qquad\qquad\qquad+\frac{\partial(p_{s,j}(y)u)(z)}{\partial z}\bigg{|}_{\begin{subarray}{c}y=h(0,y),\\ z=g(0,h_{s}(y),z)\end{subarray}}\frac{\partial g(t,h_{s}(y),z)}{\partial t}\bigg{|}_{t=0}\bigg{)}
=w^s,j(y)(yht(0,y)+zgt(0,hs(y),z))ps,ju(y,z),\displaystyle=\hat{w}_{s,j}(y)\bigg{(}\frac{\partial}{\partial y}\cdot\frac{\partial h}{\partial t}(0,y)+\frac{\partial}{\partial z}\cdot\frac{\partial g}{\partial t}(0,h_{s}(y),z)\bigg{)}p_{s,j}u(y,z),

where ps,j(y)=pj(hs(y))p_{s,j}(y)=p_{j}(h_{s}(y)). Furthermore, we have

(yht(0,y)+zgt(0,hs(y),z))ps,ju(y,z)\displaystyle\bigg{(}\frac{\partial}{\partial y}\cdot\frac{\partial h}{\partial t}(0,y)+\frac{\partial}{\partial z}\cdot\frac{\partial g}{\partial t}(0,h_{s}(y),z)\bigg{)}p_{s,j}u(y,z)
=ps,juy(y,z)ht(0,y)+ps,juz(y,z)gt(0,hs(y),z)\displaystyle\qquad=\frac{\partial p_{s,j}u}{\partial y}(y,z)\frac{\partial h}{\partial t}(0,y)+\frac{\partial p_{s,j}u}{\partial z}(y,z)\frac{\partial g}{\partial t}(0,h_{s}(y),z)
=pjuy(hs(y),z)hsy(y)ht(0,y)+pjuz(hs(y),z)gt(0,hs(y),z)\displaystyle\qquad=\frac{\partial p_{j}u}{\partial y}(h_{s}(y),z)\frac{\partial h_{s}}{\partial y}(y)\frac{\partial h}{\partial t}(0,y)+\frac{\partial p_{j}u}{\partial z}(h_{s}(y),z)\frac{\partial g}{\partial t}(0,h_{s}(y),z)
=pjuy(hs(y),z)hs(h(t,y))t|t=0+pjuz(hs(y),z)gt(0,hs(y),z)\displaystyle\qquad=\frac{\partial p_{j}u}{\partial y}(h_{s}(y),z)\frac{\partial h_{s}(h(t,y))}{\partial t}\bigg{|}_{t=0}+\frac{\partial p_{j}u}{\partial z}(h_{s}(y),z)\frac{\partial g}{\partial t}(0,h_{s}(y),z)
=pjuy(hs(y),z)ht(0,hs(y))+pjuz(hs(y),z)gt(0,hs(y),z)=VΦpju(hs(y),z).\displaystyle\qquad=\frac{\partial p_{j}u}{\partial y}(h_{s}(y),z)\frac{\partial h}{\partial t}(0,h_{s}(y))+\frac{\partial p_{j}u}{\partial z}(h_{s}(y),z)\frac{\partial g}{\partial t}(0,h_{s}(y),z)=V_{\Phi}p_{j}u(h_{s}(y),z).

By Proposition 10, we have the following proposition regarding the spectrum of N^s,j\hat{N}_{s,j}.

Proposition 22.

Assume dim(𝒱j(y))\operatorname{dim}(\mathcal{V}_{j}(y)) is finite and constant with respect to y𝒴y\in\mathcal{Y}. Then, we have σ(N^0,j)=σ(N^s,j)\sigma(\hat{N}_{0,j})=\sigma(\hat{N}_{s,j}) for any ss\in\mathbb{R}.

Proof.

Since hsh_{s} is measure preserving for any ss\in\mathbb{R}, by Proposition 10, we have

σ(N^0,j)\displaystyle\sigma(\hat{N}_{0,j}) ={λϵ>0,μ({y𝒴λσϵ(VΦpj(y))})>0}\displaystyle=\{\lambda\in\mathbb{C}\,\mid\,^{\forall}\epsilon>0,\ \mu(\{y\in\mathcal{Y}\,\mid\,\lambda\in\sigma_{\epsilon}(V_{\Phi}p_{j}(y))\})>0\}
={λϵ>0,μ({y𝒴λσϵ(VΦpj(hs(y)))})>0}=σ(N^s,j).\displaystyle=\{\lambda\in\mathbb{C}\,\mid\,^{\forall}\epsilon>0,\ \mu(\{y\in\mathcal{Y}\,\mid\,\lambda\in\sigma_{\epsilon}(V_{\Phi}p_{j}(h_{s}(y)))\})>0\}=\sigma(\hat{N}_{s,j}).

Remark 2.

If Ugt(y,)𝒱j(ht(y))𝒱(y)U_{g_{t}(y,\cdot)}\mathcal{V}_{j}(h_{t}(y))\subseteq\mathcal{V}(y) for any tt in a neighborhood of 0 in \mathbb{R}, then we have

t(Ugt(y,)pj(ht(y))u)(z)|t=0=limt01t(Ugt(y,)pj(ht(y))pj(y))u(z)\displaystyle\frac{\partial}{\partial t}(U_{g_{t}(y,\cdot)}p_{j}(h_{t}(y))u)(z)\bigg{|}_{t=0}=\lim_{t\to 0}\frac{1}{t}(U_{g_{t}(y,\cdot)}p_{j}(h_{t}(y))-p_{j}(y))u(z)
=limt01tpj(y)(Ugt(y,)pj(ht(y))pj(y))u(z)=pj(y)t(Ugt(y,)pj(ht(y))u)(z)|t=0.\displaystyle\quad=\lim_{t\to 0}\frac{1}{t}p_{j}(y)(U_{g_{t}(y,\cdot)}p_{j}(h_{t}(y))-p_{j}(y))u(z)=p_{j}(y)\frac{\partial}{\partial t}(U_{g_{t}(y,\cdot)}p_{j}(h_{t}(y))u)(z)\bigg{|}_{t=0}.

Thus, the assumption (pj(ht(y))u)(gt(y,))t|t=0𝒱j(y)\frac{\partial(p_{j}(h_{t}(y))u)(g_{t}(y,\cdot))}{\partial t}\big{|}_{t=0}\in\mathcal{V}_{j}(y) in Theorem 21 is satisfied.

Example 1.

Let 𝒴=𝒵=/2π=:𝕋\mathcal{Y}=\mathcal{Z}=\mathbb{R}/2\pi\mathbb{Z}=:\mathbb{T}. For α,β>0\alpha,\beta>0, consider the following continuous dynamical system:

(dy(t)dt,dz(t)dt)=(1,α(1+βcos(y(t)))).\bigg{(}\frac{\mathrm{d}y(t)}{\mathrm{d}t},\frac{\mathrm{d}z(t)}{\mathrm{d}t}\bigg{)}=(1,\alpha(1+\beta\cos(y(t)))). (7)

In this case, we have gt(0,y,z)=α(1+βcos(y(0)))=α(1+βcos(y))\frac{\partial g}{\partial t}(0,y,z)=\alpha(1+\beta\cos(y(0)))=\alpha(1+\beta\cos(y)) and hs(y)=y+sh_{s}(y)=y+s. Let γk,j(y,z)=e1(ky+jz)\gamma_{k,j}(y,z)=\mathrm{e}^{\sqrt{-1}(ky+jz)}. Then, we have

VΦγk,j(y,z)=(1k+1jα(1+βcos(y)))γk,j(y,z).\displaystyle V_{\Phi}\gamma_{k,j}(y,z)=\big{(}\sqrt{-1}k+\sqrt{-1}j\alpha(1+\beta\cos(y))\big{)}\gamma_{k,j}(y,z).

Let 𝒱j=Span{γk,jk}¯\mathcal{V}_{j}=\overline{\operatorname{Span}\{\gamma_{k,j}\,\mid\,k\in\mathbb{Z}\}}. We can see 𝒱j\mathcal{V}_{j} is an invariant subspace of VΦV_{\Phi}. In addition, let 𝒱j(y)=Span{γk,j(y,)k}¯\mathcal{V}_{j}(y)=\overline{\operatorname{Span}\{\gamma_{k,j}(y,\cdot)\,\mid\,k\in\mathbb{Z}\}}, and let pj(y)p_{j}(y) be the projection onto 𝒱j(y)\mathcal{V}_{j}(y). Then, since we have

(VΦpj)(y)γk,j(y,)=(VΦγk,j)(y,),\displaystyle(V_{\Phi}p_{j})(y)\gamma_{k,j}(y,\cdot)=(V_{\Phi}\gamma_{k,j})(y,\cdot),

the spectrum of (VΦpj)(y)(V_{\Phi}p_{j})(y) is calculated as

σ((VΦpj)(y))={1k+1jα(1+βcos(y))k}.\displaystyle\sigma((V_{\Phi}p_{j})(y))=\{\sqrt{-1}k+\sqrt{-1}j\alpha(1+\beta\cos(y))\,\mid\,k\in\mathbb{Z}\}.

Therefore, we have

y𝒴σ((VΦpj)(y))=y𝒴σ((VΦpj)(hs(y)))=y𝒴{1k+1jα(1+βcos(y))k}.\displaystyle\bigcup_{y\in\mathcal{Y}}\sigma((V_{\Phi}p_{j})(y))=\bigcup_{y\in\mathcal{Y}}\sigma((V_{\Phi}p_{j})(h_{s}(y)))=\bigcup_{y\in\mathcal{Y}}\{\sqrt{-1}k+\sqrt{-1}j\alpha(1+\beta\cos(y))\,\mid\,k\in\mathbb{Z}\}.

Regarding w^s,j\hat{w}_{s,j}, we have

Ugs(y,)γk,j(y,)\displaystyle U_{g_{s}(y,\cdot)}\gamma_{k,j}(y,\cdot) =γk,j(y,+α(s+β(sin(y+s)sin(y))))\displaystyle=\gamma_{k,j}(y,\cdot+\alpha(s+\beta(\sin(y+s)-\sin(y))))
=γk,j(y,)e1jα(s+β(sin(y+s)sin(y))).\displaystyle=\gamma_{k,j}(y,\cdot)\mathrm{e}^{\sqrt{-1}j\alpha(s+\beta(\sin(y+s)-\sin(y)))}.

Thus, the spectrum of the family of operators {w^s,j(y)}s\{\hat{w}_{s,j}(y)\}_{s} on L2(𝒵)L^{2}(\mathcal{Z}) is γj(αs+αβ(sin(y+s)sin(y)))\gamma_{j}\big{(}\alpha s+\alpha\beta(\sin(y+s)-\sin(y))\big{)}.

In the following subsections, we will generalize the arguments in Example 1.

4.3.3 Construction of the generalized Oseledets space 𝒱j(y)\mathcal{V}_{j}(y) using a function space on 𝒳\mathcal{X}

We show how we can construct the generalized Oseledets space 𝒱j(y)\mathcal{V}_{j}(y) required for obtaining pjp_{j} appearing in Theorem 21. In this subsection, we assume 𝒴\mathcal{Y} is compact. Let

𝒩=C(𝒴)L2(𝒵)\displaystyle\mathcal{N}=C(\mathcal{Y})\otimes L^{2}(\mathcal{Z}) (8)

be the Hilbert C(𝒴)C(\mathcal{Y})-module. Note that a Hilbert CC^{*}-module is also a Banach space. Here, we just regard 𝒩\mathcal{N} as a Banach space equipped with the norm au𝒩2=supy𝒴z𝒵|a(y)u(z)|2dν(z)\|a\otimes u\|_{\mathcal{N}}^{2}=\sup_{y\in\mathcal{Y}}\int_{z\in\mathcal{Z}}|a(y)u(z)|^{2}\mathrm{d}\nu(z). For tt\in\mathbb{R}, let UΦtU_{\Phi_{t}} be the Koopman operator on 𝒩\mathcal{N} with respect to Φt\Phi_{t}.

Proposition 23.

The family of operators {UΦt}t\{U_{\Phi_{t}}\}_{t\in\mathbb{R}} is a strongly continuous one-parameter group.

Proof.

Let vC(𝒴)algCc(𝒵)v\in C(\mathcal{Y})\otimes_{\operatorname{alg}}C_{c}(\mathcal{Z}) and let ϵ>0\epsilon>0. Then, there exists δ>0\delta>0 such that for any |t|δ|t|\leq\delta, y𝒴y\in\mathcal{Y}, and z𝒵z\in\mathcal{Z}, |v(ht(y),gt(y,z))v(y,z)|ϵ|v(h_{t}(y),g_{t}(y,z))-v(y,z)|\leq\epsilon. Thus, we have

UΦtvv𝒩=supy𝒴(z𝒵|v(ht(y),gt(y,z))v(y,z)|2dν(z))1/2ϵ.\|U_{\Phi_{t}}v-v\|_{\mathcal{N}}=\sup_{y\in\mathcal{Y}}\bigg{(}\int_{z\in\mathcal{Z}}|v(h_{t}(y),g_{t}(y,z))-v(y,z)|^{2}\mathrm{d}\nu(z)\bigg{)}^{1/2}\leq\epsilon. (9)

In addition, for any vC(𝒴)algL2(𝒵)v\in C(\mathcal{Y})\otimes_{\operatorname{alg}}L^{2}(\mathcal{Z}), we have

UΦtv𝒩\displaystyle\|U_{\Phi_{t}}v\|_{\mathcal{N}} =supy𝒴(z𝒵|v(ht(y),gt(y,z))|2dν(z))1/2\displaystyle=\sup_{y\in\mathcal{Y}}\bigg{(}\int_{z\in\mathcal{Z}}|v(h_{t}(y),g_{t}(y,z))|^{2}\mathrm{d}\nu(z)\bigg{)}^{1/2}
=supy𝒴(z𝒵|v(y,z)|2dν(z))1/2=v𝒩.\displaystyle=\sup_{y\in\mathcal{Y}}\bigg{(}\int_{z\in\mathcal{Z}}|v(y,z)|^{2}\mathrm{d}\nu(z)\bigg{)}^{1/2}=\|v\|_{\mathcal{N}}.

Since C(𝒴)algCc(𝒵)C(\mathcal{Y})\otimes_{\operatorname{alg}}C_{c}(\mathcal{Z}) is dense in 𝒩\mathcal{N}, Eq. (9) is satisfied for any v𝒩v\in\mathcal{N}. ∎

We note that the generator of {UΦt}t\{U_{\Phi_{t}}\}_{t\in\mathbb{R}} is VΦV_{\Phi} defined in Eq. (6). If we set 𝒱j\mathcal{V}_{j} as 𝒱\mathcal{V} in the following proposition, it satisfies the assumption of Theorem 21 (see also Remark 2).

Proposition 24.

Let 𝒱\mathcal{V} be an invariant subspace of UΦtU_{\Phi_{t}} and let 𝒱(y)=Ry𝒱¯\mathcal{V}(y)=\overline{R_{y}\mathcal{V}}. Then, we have Ugt(y,)𝒱(ht(y))𝒱(y)U_{g_{t}(y,\cdot)}\mathcal{V}(h_{t}(y))\subseteq\mathcal{V}(y). Here, Ry:𝒩L2(𝒵)R_{y}:\mathcal{N}\to L^{2}(\mathcal{Z}) be a linear map defined as Ry(au)(z)=a(y)u(z)R_{y}(a\otimes u)(z)=a(y)u(z) for y𝒴y\in\mathcal{Y}.

Proof.

For tt\in\mathbb{R}, y𝒴y\in\mathcal{Y}, and vC(𝒴)algL2(𝒵)v\in C(\mathcal{Y})\otimes_{\operatorname{alg}}L^{2}(\mathcal{Z}), we have

Ugt(y,)Rht(y)v(z)=v(ht(y),gt(y,z))=RyUΦtv(z).\displaystyle U_{g_{t}(y,\cdot)}R_{h_{t}(y)}v(z)=v(h_{t}(y),g_{t}(y,z))=R_{y}U_{\Phi_{t}}v(z).

Thus, we have Ugt(y,)Rht(y)=RyUΦtU_{g_{t}(y,\cdot)}R_{h_{t}(y)}=R_{y}U_{\Phi_{t}}. Since 𝒱\mathcal{V} is an invariant subspace of UΦtU_{\Phi_{t}}, we have Ugt(y,)𝒱(ht(y))𝒱(y)U_{g_{t}(y,\cdot)}\mathcal{V}(h_{t}(y))\subseteq\mathcal{V}(y). ∎

The following proposition shows an example of 𝒱\mathcal{V} constructed in Proposition 24. It is for a simple case where VΦV_{\Phi} has an eigenvalue, but provides us with an intuition of what the eigenoperators describe.

Proposition 25.

Assume there exists y𝒴y\in\mathcal{Y} such that {h(t,y)t}\{h(t,y)\,\mid\,t\in\mathbb{R}\} is dense in 𝒴\mathcal{Y}. Assume VΦV_{\Phi} has an eigenvalue λ\lambda and the corresponding eigenvector v~Cc1(𝒳)\tilde{v}\in C_{c}^{1}(\mathcal{X}). Then, there exists C0C\geq 0 such that for a.s. y𝒴y\in\mathcal{Y}, v~(y,)L2(𝒵)=C\|\tilde{v}(y,\cdot)\|_{L^{2}(\mathcal{Z})}=C. Assume C>0C>0 and let p(y)u=v(y,)v(y,)up(y)u={v(y,\cdot)v(y,\cdot)^{*}u} for uL2(𝒵)u\in L^{2}(\mathcal{Z}), where v=v~/Cv=\tilde{v}/C. Then, p(y)(VΦpu)(y,)=(VΦpu)(y,)p(y)(V_{\Phi}pu)(y,\cdot)=(V_{\Phi}pu)(y,\cdot) and

σ((VΦp)(y))=λ𝒵vy(y,z)v¯(y,z)dν(z)ht(0,y)=𝒵vz(y,z)gt(0,y,z)v¯(y,z)dν(z).\displaystyle\sigma((V_{\Phi}p)(y))=\lambda-\int_{\mathcal{Z}}\frac{\partial{v}}{\partial y}(y,z)\overline{v}(y,z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y)=\int_{\mathcal{Z}}\frac{\partial{v}}{\partial z}(y,z)\frac{\partial{g}}{\partial t}(0,y,z)\overline{v}(y,z)\mathrm{d}\nu(z).

Moreover, σ((VΦp)(y))1\sigma((V_{\Phi}p)(y))\subseteq\sqrt{-1}\mathbb{R}.

Proof.

The vector v~\tilde{v} is an eigenvector of UΦtU_{\Phi_{t}} for any tt\in\mathbb{R}, and its corresponding eigenvalue is eλt\mathrm{e}^{\lambda t} (λ1\lambda\in\sqrt{-1}\mathbb{R}). Thus, we have

𝒵v~(y,z)v~¯(y,z)dν(z)\displaystyle\int_{\mathcal{Z}}\tilde{v}(y,z)\overline{\tilde{v}}(y,z)\mathrm{d}\nu(z) =𝒵eλtUΦtv~(y,z)eλtUΦtv~(y,z)¯dν(z)\displaystyle=\int_{\mathcal{Z}}\mathrm{e}^{-\lambda t}U_{\Phi_{t}}\tilde{v}(y,z)\overline{\mathrm{e}^{-\lambda t}U_{\Phi_{t}}\tilde{v}(y,z)}\mathrm{d}\nu(z)
=𝒵v~(ht(y),gt(y,z))v~¯(ht(y),gt(y,z))dν(z)\displaystyle=\int_{\mathcal{Z}}\tilde{v}(h_{t}(y),g_{t}(y,z))\overline{\tilde{v}}(h_{t}(y),g_{t}(y,z))\mathrm{d}\nu(z)
=𝒵v~(ht(y),z)v~¯(ht(y),z)dν(z).\displaystyle=\int_{\mathcal{Z}}\tilde{v}(h_{t}(y),z)\overline{\tilde{v}}(h_{t}(y),z)\mathrm{d}\nu(z).

For uL2(𝒵)u\in L^{2}(\mathcal{Z}), we have

(VΦpu)(y,z)\displaystyle(V_{\Phi}pu)(y,z)
=(vy(y,z)𝒵v¯(y,z)u(z)dν(z)+v(y,z)𝒵v¯y(y,z)u(z)dν(z))ht(0,y)\displaystyle=\bigg{(}\frac{\partial v}{\partial y}(y,z)\int_{\mathcal{Z}}\overline{v}(y,z)u(z)\mathrm{d}\nu(z)+v(y,z)\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)u(z)\mathrm{d}\nu(z)\bigg{)}\frac{\partial h}{\partial t}(0,y)
+vz(y,z)𝒵v¯(y,z)u(z)dν(z)gt(0,y,z)\displaystyle\qquad\qquad+\frac{\partial v}{\partial z}(y,z)\int_{\mathcal{Z}}\overline{v}(y,z)u(z)\mathrm{d}\nu(z)\frac{\partial g}{\partial t}(0,y,z)
=(VΦv)(y,z)𝒵v¯(y,z)u(z)dν(z)+v(y,z)𝒵v¯y(y,z)u(z)dν(z)ht(0,y)\displaystyle=(V_{\Phi}v)(y,z)\int_{\mathcal{Z}}\overline{v}(y,z)u(z)\mathrm{d}\nu(z)+v(y,z)\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)u(z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y)
=λv(y,z)𝒵v¯(y,z)u(z)dν(z)+v(y,z)𝒵v¯y(y,z)u(z)dν(z)ht(0,y).\displaystyle=\lambda v(y,z)\int_{\mathcal{Z}}\overline{v}(y,z)u(z)\mathrm{d}\nu(z)+v(y,z)\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)u(z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y).

Thus, p(y)(VΦpu)(y)=(VΦpu)(y)p(y)(V_{\Phi}pu)(y)=(V_{\Phi}pu)(y). In addition, we have

(VΦp)(y)v(y,)\displaystyle(V_{\Phi}p)(y)v(y,\cdot) =λv(y,)𝒵v¯(y,z)v(y,z)dν(z)+v(y,)𝒵v¯y(y,z)v(y,z)dν(z)ht(0,y)\displaystyle=\lambda v(y,\cdot)\int_{\mathcal{Z}}\overline{v}(y,z)v(y,z)\mathrm{d}\nu(z)+v(y,\cdot)\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)v(y,z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y)
=(λ+𝒵v¯y(y,z)v(y,z)dν(z)ht(0,y))v(y,).\displaystyle=\bigg{(}\lambda+\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)v(y,z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y)\bigg{)}v(y,\cdot).

Moreover, we have

0\displaystyle 0 =𝒵v(y,z)v¯(y,z)ydν(z)=𝒵v¯y(y,z)v(y,z)dν(z)+𝒵vy(y,z)v¯(y,z)dν(z)\displaystyle=\int_{\mathcal{Z}}\frac{\partial v(y,z)\overline{v}(y,z)}{\partial y}\mathrm{d}\nu(z)=\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)v(y,z)\mathrm{d}\nu(z)+\int_{\mathcal{Z}}\frac{\partial{v}}{\partial y}(y,z)\overline{v}(y,z)\mathrm{d}\nu(z)
=2(𝒵v¯y(y,z)v(y,z)dν(z)),\displaystyle=2\Re\bigg{(}\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)v(y,z)\mathrm{d}\nu(z)\bigg{)},

and

𝒵v¯y(y,z)v(y,z)dν(z)ht(0,y)\displaystyle\int_{\mathcal{Z}}\frac{\partial\overline{v}}{\partial y}(y,z)v(y,z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y) =𝒵vy(y,z)v¯(y,z)dν(z)ht(0,y)\displaystyle=-\int_{\mathcal{Z}}\frac{\partial{v}}{\partial y}(y,z)\overline{v}(y,z)\mathrm{d}\nu(z)\frac{\partial h}{\partial t}(0,y)
=𝒵(VΦv(y,z)vz(y,z)gt(0,y,z))v¯(y,z)dν(z)\displaystyle=-\int_{\mathcal{Z}}\bigg{(}V_{\Phi}v(y,z)-\frac{\partial{v}}{\partial z}(y,z)\frac{\partial g}{\partial t}(0,y,z)\bigg{)}\overline{v}(y,z)\mathrm{d}\nu(z)
=𝒵(λv(y,z)+vz(y,z)gt(0,y,z))v¯(y,z)dν(z).\displaystyle=\int_{\mathcal{Z}}\bigg{(}-\lambda v(y,z)+\frac{\partial{v}}{\partial z}(y,z)\frac{\partial g}{\partial t}(0,y,z)\bigg{)}\overline{v}(y,z)\mathrm{d}\nu(z).

Remark 3.

Proposition 25 implies that the eigenoperators have the information of the dynamics on 𝒵\mathcal{Z} for each yy, which cannot be extracted by eigenvalues of VΦV_{\Phi}. Indeed, if an eigenvector vjv_{j} of VΦV_{\Phi} depends only on yy, then σ(N^s,j)=σ(VΦp(y))=0\sigma(\hat{N}_{s,j})=\sigma(V_{\Phi}p(y))=0. On the other hand, the corresponding eigenvalue of VΦV_{\Phi} can be nonzero.

4.3.4 Approximation of VΦV_{\Phi} in RKHS

For numerical computations, to construct the subspace 𝒱j(y)\mathcal{V}_{j}(y), we need to approximate VΦV_{\Phi} using RKHSs. Approximating the generator of the Koopman operator in RKHSs was proposed by Das et al. [22]. Here, we apply a similar technique to approximating VΦV_{\Phi}. In this subsection, we assume 𝒴=𝕋\mathcal{Y}=\mathbb{T}. We also assume 𝒵\mathcal{Z} is compact and ν\nu is a Borel probability measure satisfying supp(ν)=𝒵\operatorname{supp}(\nu)=\mathcal{Z}.

Let ϕi(l)=e1il\phi_{i}(l)=\mathrm{e}^{\sqrt{-1}il} and λi=e|i|\lambda_{i}=\mathrm{e}^{-|i|} for ii\in\mathbb{Z} and l𝕋l\in\mathbb{T}. Let p1:𝕋×𝕋p_{1}:\mathbb{T}\times\mathbb{T}\to\mathbb{C} be the positive definite kernel defined as p1(l1,l2)=iλiϕi(l1)ϕi(l2)¯p_{1}(l_{1},l_{2})=\sum_{i\in\mathbb{Z}}\lambda_{i}\phi_{i}(l_{1})\overline{\phi_{i}(l_{2})}. In addition, let p2:𝒵×𝒵p_{2}:\mathcal{Z}\times\mathcal{Z}\to\mathbb{C} be a positive definite kernel, and let P~:L2(𝒵)L2(𝒵)\tilde{P}:L^{2}(\mathcal{Z})\to L^{2}(\mathcal{Z}) be the integral operator with respect to p2p_{2}. Let λ~1λ~2>0\tilde{\lambda}_{1}\geq\tilde{\lambda}_{2}\geq\ldots>0 and ϕ~1,ϕ~2,\tilde{\phi}_{1},\tilde{\phi}_{2},\ldots be eigenvalues and the corresponding orthonormal eigenvectors of P~\tilde{P}, respectively. By Mercer’s theorem, p2(z1,z2)=i=1λ~iϕ~i(z1)ϕ~i(z2)¯p_{2}(z_{1},z_{2})=\sum_{i=1}^{\infty}\tilde{\lambda}_{i}\tilde{\phi}_{i}(z_{1})\overline{\tilde{\phi}_{i}(z_{2})}, where the sum converges uniformly on 𝒵×𝒵\mathcal{Z}\times\mathcal{Z}. Let τ>0\tau>0 and let λτ,i,j=eτ(1λi1λ~j1)\lambda_{\tau,i,j}=\mathrm{e}^{\tau(1-\lambda_{i}^{-1}\tilde{\lambda}_{j}^{-1})} for ii\in\mathbb{Z} and j=1,2,j=1,2,\ldots. Let

pτ((l1,z1),(l2,z2))=ij=1λτ,i,jϕi(l1)ϕ~j(z1)ϕ~j(z2)¯ϕi(l2)¯\displaystyle p_{\tau}((l_{1},z_{1}),(l_{2},z_{2}))=\sum_{i\in\mathbb{Z}}\sum_{j=1}^{\infty}\lambda_{\tau,i,j}\phi_{i}(l_{1})\tilde{\phi}_{j}(z_{1})\overline{\tilde{\phi}_{j}(z_{2})}\overline{\phi_{i}(l_{2})}

and let τ\mathcal{H}_{\tau} be the RKHS associated with pτp_{\tau}. In addition, let Pτ{P}_{\tau} be the integral operator with respect to pτp_{\tau}.

Proposition 26.

Let ιτ:τ𝒩\iota_{\tau}:\mathcal{H}_{\tau}\to\mathcal{N} be the inclusion map, where 𝒩\mathcal{N} is defined as Eq. (8). Then, for any v𝒩v\in\mathcal{N}, ιτPτvv𝒩\|\iota_{\tau}P_{\tau}v-v\|_{\mathcal{N}} converges to 0 as τ0\tau\to 0.

Proof.

Let ψi,j=ϕiϕ~j\psi_{i,j}=\phi_{i}\otimes\tilde{\phi}_{j}. Since {ϕ~i}i=1\{\tilde{\phi}_{i}\}_{i=1}^{\infty} is an orthonormal basis in L2(𝒵)L^{2}(\mathcal{Z}), the subspace {i=nnj=1mai,jψi,jn,m,ai,j}\{\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}\,\mid\,n,m\in\mathbb{N},a_{i,j}\in\mathbb{C}\} is dense in 𝒩\mathcal{N} with ψi,j𝒩=1\|\psi_{i,j}\|_{\mathcal{N}}=1. In addition, since we have ιτPτψi,j=λτ,i,jψi,j\iota_{\tau}P_{\tau}\psi_{i,j}=\lambda_{\tau,i,j}\psi_{i,j} and 0λτ,i,j10\leq\lambda_{\tau,i,j}\leq 1, we obtain ιτPτ𝒩1\|\iota_{\tau}P_{\tau}\|_{\mathcal{N}}\leq 1. For any ϵ>0\epsilon>0 and v𝒩v\in\mathcal{N}, there exist n,mn,m\in\mathbb{N}, ai,ja_{i,j}\in\mathbb{C}, and τ0>0\tau_{0}>0 such that i=nnj=1mai,jψi,jv𝒩ϵ\|\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}-v\|_{\mathcal{N}}\leq\epsilon and (1λτ,i,j)(i=nnj=1m|ai,j|)ϵ(1-\lambda_{\tau,i,j})(\sum_{i=-n}^{n}\sum_{j=1}^{m}|a_{i,j}|)\leq\epsilon for i=n,,ni=-n,\ldots,n, j=1,,mj=1,\ldots,m, and ττ0\tau\leq\tau_{0}. Thus, for ττ0\tau\leq\tau_{0}, we have

ιτPτvv𝒩\displaystyle\|\iota_{\tau}P_{\tau}v-v\|_{\mathcal{N}}
ιτPτvιτPτi=nnj=1mai,jψi,j𝒩+ιτPτi=nnj=1mai,jψi,ji=nnj=1mai,jψi,j𝒩\displaystyle\leq\bigg{\|}\iota_{\tau}P_{\tau}v-\iota_{\tau}P_{\tau}\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}\bigg{\|}_{\mathcal{N}}+\bigg{\|}\iota_{\tau}P_{\tau}\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}-\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}\bigg{\|}_{\mathcal{N}}
+i=nnj=1mai,jψi,jv𝒩\displaystyle\qquad\qquad+\bigg{\|}\sum_{i=-n}^{n}\sum_{j=1}^{m}a_{i,j}\psi_{i,j}-v\bigg{\|}_{\mathcal{N}}
ϵ+i=nnj=1m(1λτ,i,j)|ai,j|+ϵ=3ϵ.\displaystyle\leq\epsilon+\sum_{i=-n}^{n}\sum_{j=1}^{m}(1-\lambda_{\tau,i,j})|a_{i,j}|+\epsilon=3\epsilon.

By Proposition 26, we can see that VΦV_{\Phi} can be approximated by PτVΦιτP_{\tau}V_{\Phi}\iota_{\tau} in the following sense.

Corollary 27.

Let τ0>0\tau_{0}>0. For any vτ0v\in\mathcal{H}_{\tau_{0}}, ιτPτVΦιτvVΦιτv𝒩\|\iota_{\tau}P_{\tau}V_{\Phi}\iota_{\tau}v-V_{\Phi}\iota_{\tau}v\|_{\mathcal{N}} converges to 0 as τ0\tau\to 0.

5 Numerical examples

We numerically investigate the eigenoperator decomposition.

5.1 Moving Gaussian vortex

We first visualize w^s,j\hat{w}_{s,j}, the eigenvector in Theorem 21. Let 𝒴=𝕋\mathcal{Y}=\mathbb{T} and 𝒵=𝕋2\mathcal{Z}=\mathbb{T}^{2}. Consider the dynamical system

(dy(t)dt,(dz1(t)dt,dz2(t)dt))=(1,(ζz2(y(t),z(t)),ζz1(y(t),z(t)))),\bigg{(}\frac{\mathrm{d}y(t)}{\mathrm{d}t},\bigg{(}\frac{\mathrm{d}z_{1}(t)}{\mathrm{d}t},\frac{\mathrm{d}z_{2}(t)}{\mathrm{d}t}\bigg{)}\bigg{)}=\bigg{(}1,\bigg{(}-\frac{\partial\zeta}{\partial z_{2}}(y(t),z(t)),\frac{\partial\zeta}{\partial z_{1}}(y(t),z(t))\bigg{)}\bigg{)}, (10)

where ζ(y,z)=eκ(cos(z1y)+cosz2)\zeta(y,z)=\mathrm{e}^{\kappa(\cos(z_{1}-y)+\cos z_{2})}. This problem is also studied by Giannakis and Das [28]. In this case, gt(0,y,z)=(ζz2(y,z),ζz1(y,z))\frac{\partial g}{\partial t}(0,y,z)=(-\frac{\partial\zeta}{\partial z_{2}}(y,z),\frac{\partial\zeta}{\partial z_{1}}(y,z)) and VΦ=yζz2(y,z)z1+ζz1(y,z)z2V_{\Phi}=\frac{\partial}{\partial y}-\frac{\partial\zeta}{\partial z_{2}}(y,z)\frac{\partial}{\partial z_{1}}+\frac{\partial\zeta}{\partial z_{1}}(y,z)\frac{\partial}{\partial z_{2}}. To construct the subspace 𝒱j(y)\mathcal{V}_{j}(y) approximately, we set κ=0.5\kappa=0.5 and approximated VΦV_{\Phi} in the RKHS τττ\mathcal{H}_{\tau}\otimes\mathcal{H}_{\tau}\otimes\mathcal{H}_{\tau} with τ=0.1\tau=0.1. Here, τ\mathcal{H}_{\tau} is the RKHS associated with the positive definite kernel pτ:𝕋×𝕋p_{\tau}:\mathbb{T}\times\mathbb{T}\to\mathbb{C} defined as pτ(y1,y2)=iλτ,iϕi(y1)ϕi(y2)¯p_{\tau}(y_{1},y_{2})=\sum_{i\in\mathbb{Z}}\lambda_{\tau,i}\phi_{i}(y_{1})\overline{\phi_{i}(y_{2})}. In addition, ϕi(y)=e1iy\phi_{i}(y)=\mathrm{e}^{\sqrt{-1}iy} and λτ,i=eτ|i|p\lambda_{\tau,i}=\mathrm{e}^{-\tau|i|^{p}} for ii\in\mathbb{Z}. We set p=0.1p=0.1. We computed eigenvectors v~1,v~2,\tilde{v}_{1},\tilde{v}_{2},\ldots of the approximated operator in the RKHS. Here, the index is ordered from the eigenvector corresponding to the closest eigenvalue to 101010^{-10}. We set 𝒱1(y)=Span{v~1(y,),,v~d(y,)}\mathcal{V}_{1}(y)=\operatorname{Span}\{\tilde{v}_{1}(y,\cdot),\ldots,\tilde{v}_{d}(y,\cdot)\}. Since the eigenvector ws,j(y)w_{s,j}(y) is an operator, its visualization is not easy. Thus, we set any test vector qy,d𝒱1(y)q_{y,d}\in\mathcal{V}_{1}(y) and visualize ws,j(y)qy,dw_{s,j}(y)q_{y,d} instead of ws,jw_{s,j} itself. As the test vector, we set qy,d=1/di=1dv~i(y,)q_{y,d}=1/d\sum_{i=1}^{d}\tilde{v}_{i}(y,\cdot) as an example. Figure 2 shows the eigenvector w^s,1\hat{w}_{s,1} acting on the vector qy,dq_{y,d}. We can see that the pattern becomes more clear as dd becomes large, which implies that considering higher dimensional subspaces 𝒱1(y)\mathcal{V}_{1}(y) than a one dimensional space catches the feature of the dynamical system in this case.

Refer to caption
d=1d=1
Refer to caption
d=2d=2
Refer to caption
d=10d=10
Refer to caption
d=20d=20
Refer to caption
d=30d=30
Figure 2: Real part of w^s,1qy,d\hat{w}_{s,1}q_{y,d} for y=0y=0 and s=0.1s=0.1.

5.2 Idealized stratospheric flow

Next, we observe the eigenoperator N^s,j\hat{N}_{s,j}. We study what information the eigenoperators capture. Let 𝒴=𝕋\mathcal{Y}=\mathbb{T} and 𝒵=𝕋×[π,π]\mathcal{Z}=\mathbb{T}\times[-\pi,\pi]. Consider the same dynamical system as Eq. (10) with ζ(y,z)=c3z2U0Ltanh(z2/L)+i=13AiU0Lsech2(z2/L)cos(kiz1σiy)\zeta(y,z)=c_{3}z_{2}-U_{0}L\tanh(z_{2}/L)+\sum_{i=1}^{3}A_{i}U_{0}L\operatorname{sech}^{2}(z_{2}/L)\cos(k_{i}z_{1}-\sigma_{i}y), where L=0.1L=0.1, A1=0.075A_{1}=0.075, A2=0.4A_{2}=0.4, A3=0.2A_{3}=0.2, k1=1k_{1}=1, k2=2k1k_{2}=2k_{1}, k3=3k1k_{3}=3k_{1}, U0=62.66U_{0}=62.66, c3=0.7U0c_{3}=0.7U_{0}, σ2=1\sigma_{2}=-1, and σ1=2σ1\sigma_{1}=2\sigma_{1}. A similar problem is also studied by Froyland et al. [26]. We approximated VΦV_{\Phi} in the RKHS τττ\mathcal{H}_{\tau}\otimes\mathcal{H}_{\tau}\otimes\mathcal{H}_{\tau} with τ=0.1\tau=0.1. We computed eigenvectors v~1,v~2,\tilde{v}_{1},\tilde{v}_{2},\ldots of the approximated operator in the RKHS and set 𝒱j(y)=Span{v~j(y)}\mathcal{V}_{j}(y)=\operatorname{Span}\{\tilde{v}_{j}(y)\} for j=1,2,j=1,2,\ldots. We study N^s,j\hat{N}_{s,j} for different jj and what information we can extract according to N^s,j\hat{N}_{s,j}. Figure 3 shows the heatmap of the function w^0,jv~j(0,)\hat{w}_{0,j}\tilde{v}_{j}(0,\cdot) for different values of jj. Since we have w^0,jv~j(y,)=v~j(y,)\hat{w}_{0,j}\tilde{v}_{j}(y,\cdot)=\tilde{v}_{j}(y,\cdot), it provides us coherent patterns. We computed the spectrum of N^s,j\hat{N}_{s,j} for the corresponding jj. The eigenoperator N^s,j\hat{N}_{s,j} is different from the spectrum of VΦV_{\Phi}, the generator of the Koopman operator. We also computed the spectrum of VΦV_{\Phi}. We can see that the pattern becomes complicated as the magnitude of the spectrum σ(N^0,j)\sigma(\hat{N}_{0,j}) of the eigenoperator becomes large. On the other hand, the spectrum of VΦV_{\Phi} does not provide such an observation.

Refer to caption
σ(N^s,j)=11.532i\sigma(\hat{N}_{s,j})=-11.532\mathrm{i},
σ(VΦ)=0.040001i\sigma(V_{\Phi})=0.040001\mathrm{i}
Refer to caption
σ(N^s,j)=12.4976i\sigma(\hat{N}_{s,j})=12.4976\mathrm{i},
σ(VΦ)=0.22200i\sigma(V_{\Phi})=0.22200\mathrm{i}
Refer to caption
σ(N^s,j)=13.028i\sigma(\hat{N}_{s,j})=13.028\mathrm{i},
σ(VΦ)=0.17188i\sigma(V_{\Phi})=0.17188\mathrm{i}
Figure 3: Real part of w^s,jv~j(y,)\hat{w}_{s,j}\tilde{v}_{j}(y,\cdot) for s=y=0s=y=0

6 Conclusion and discussion

In this paper, we considered a skew product dynamical system on 𝒴×𝒵\mathcal{Y}\times\mathcal{Z} and defined a linear operator on a Hilbert CC^{*}-module related to the Koopman operator. We proposed the eigenoperator decomposition as a generalization of the eigenvalue decomposition. The eigenvectors are constructed using a cocycle. The eigenoperators reconstruct the Koopman operator projected on generalized Oseledets subspaces. Thus, if the Oseledets subspaces are infinite-dimensional spaces, the eigenoperators can have continuous spectra related to the Koopman operator. Our approach is different from existing approach to dealing with continuous and residual spectra of Koopman operators, such as focusing on the spectral measure [19] and approximating Koopman operators using compact operators on different space from the space where the Koopman operators are defined [20, 28]. In addition, the proposed decomposition gives us information of the behavior of coherent patterns on 𝒵\mathcal{Z}. Extracting coherent structure of skew product dynamical systems has been investigated [35, 26, 22]. The proposed decomposition will allow us to classify these coherent patterns.

For future work, studying data-driven approaches to obtaining the decomposition is an important direction of researches. Investigating practical and computationally efficient ways to approximate operators on Hilbert CC^{*}-modules would be essential in that direction of researches. Another interesting direction is applying the proposed decomposition to quantum computation. Decomposing a Koopman operator for quantum computation was proposed [43]. It would be interesting to generalize the decomposition using the proposed decomposition.

Acknowledgments

We thank Suddhasattwa Das for many constructive discussions on this work. DG acknowledges support from the US National Science Foundation under grant DMS-1854383, the US Office of Naval Research under MURI grant N00014-19-1-242, and the US Department of Defense, Basic Research Office under Vannevar Bush Faculty Fellowship grant N00014-21-1-2946. MI and II acknowledge support from the Japan Science and Technlogy Agency under CREST grant JPMJCR1913. II acknowledge support from the Japan Science and Technlogy Agency under ACT-X grant JPMJAX2004.

Appendix A Proof of Proposition 1

To show Proposition 1, we use the following lemmas [44].

Lemma 28.

We have (L2(𝒵),L2(𝒳))\mathcal{M}\subseteq\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})).

Proof.

Let ι:L2(𝒴)alg𝒜(L2(𝒵),L2(𝒳))\iota:L^{2}(\mathcal{Y})\otimes_{\operatorname{alg}}\mathcal{A}\to\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})) be defined as (ι(va)u)(y,z)=v(y)(au)(z)(\iota(v\otimes a)u)(y,z)=v(y)(au)(z). Then, ι\iota is an injection. In addition, we have

ι(va)(L2(𝒵),L2(𝒳))\displaystyle\|\iota(v\otimes a)\|_{\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X}))} =supuL2(𝒵)=1ι(va)uL2(𝒳)\displaystyle=\sup_{\|u\|_{L^{2}(\mathcal{Z})}=1}\|\iota(v\otimes a)u\|_{L^{2}(\mathcal{X})}
=vL2(𝒴)supuL2(𝒵)=1auL2(𝒵)=va.\displaystyle=\|v\|_{L^{2}(\mathcal{Y})}\sup_{\|u\|_{L^{2}(\mathcal{Z})}=1}\|au\|_{L^{2}(\mathcal{Z})}=\|v\otimes a\|_{\mathcal{M}}.

Therefore, we have (L2(𝒵),L2(𝒳))\mathcal{M}\subseteq\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})). ∎

We recall that a Hilbert CC^{*}-module \mathcal{M} is referred to as self-dual if for any bounded 𝒜\mathcal{A}-linear map b:𝒜b:\mathcal{M}\to\mathcal{A}, there exists a unique b^\hat{b}\in\mathcal{M} such that b(w)=b^,wb(w)=\langle\hat{b},w\rangle_{\mathcal{M}}.

Lemma 29.

Let {wi}i=1\{w_{i}\}_{i=1}^{\infty} be a sequence in \mathcal{M}. Assume there exists w(L2(𝒵),L2(𝒳))w\in\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})) such that for any uL2(𝒵)u\in L^{2}(\mathcal{Z}), ι(wi)uwu\iota(w_{i})u\to wu in L2(𝒳)L^{2}(\mathcal{X}), where ι\iota is defined in the proof of Lemma 28. Then, ww\in\mathcal{M}.

Proof.

For w~\tilde{w}\in\mathcal{M} and u1,u2L2(𝒵)u_{1},u_{2}\in L^{2}(\mathcal{Z}), we have

ι(w~)u1,wu2=ι(w~)u1,limiι(wi)u2\displaystyle\left\langle\iota(\tilde{w})u_{1},wu_{2}\right\rangle=\left\langle\iota(\tilde{w})u_{1},\lim_{i\to\infty}\iota(w_{i})u_{2}\right\rangle =limiz𝒵y𝒴u1(z)¯w~(y)wi(y)u2(z)dμ(y)dν(z)\displaystyle=\lim_{i\to\infty}\int_{z\in\mathcal{Z}}\int_{y\in\mathcal{Y}}\overline{u_{1}(z)}\tilde{w}(y)^{*}w_{i}(y)u_{2}(z)\mathrm{d}\mu(y)\mathrm{d}\nu(z)
=limiu1,w~,wiu2L2(𝒵).\displaystyle=\lim_{i\to\infty}\left\langle u_{1},\left\langle\tilde{w},w_{i}\right\rangle_{\mathcal{M}}u_{2}\right\rangle_{L^{2}(\mathcal{Z})}.

Thus, by the Riesz representation theorem, there exists a𝒜a\in\mathcal{A} such that ι(w~)u1,wu2=u1,au2\left\langle\iota(\tilde{w})u_{1},wu_{2}\right\rangle=\left\langle u_{1},au_{2}\right\rangle. The map w~a\tilde{w}\mapsto a^{*} is a bounded 𝒜\mathcal{A}-linear map from \mathcal{M} to 𝒜\mathcal{A}. Since 𝒜\mathcal{A} is self-dual, \mathcal{M} is also self-dual. As a result, there exists w^\hat{w}\in\mathcal{M} such that w^,w~=a\langle\hat{w},\tilde{w}\rangle_{\mathcal{M}}=a^{*} and w=ι(w^)w=\iota(\hat{w}). ∎

Lemma 30.

Let w:𝒴𝒜w:\mathcal{Y}\to\mathcal{A}. Assume for any uL2(𝒵)u\in L^{2}(\mathcal{Z}), the map (y,z)(w(y)u)(z)(y,z)\mapsto(w(y)u)(z) is contained in L2(𝒳)L^{2}(\mathcal{X}). Then, ww\in\mathcal{M}.

Proof.

Let {γi}i=1\{\gamma_{i}\}_{i=1}^{\infty} be an orthonormal basis of L2(𝒵)L^{2}(\mathcal{Z}). For uL2(𝒴)u\in L^{2}(\mathcal{Y}), we have

w(y)u=w(y)i=1γiγiu=i=1w(y)γiγiu.w(y)u=w(y)\sum_{i=1}^{\infty}\gamma_{i}\gamma_{i}^{\prime}u=\sum_{i=1}^{\infty}w(y)\gamma_{i}\gamma_{i}^{\prime}u.

Here, γi\gamma_{i}^{\prime} denotes the dual of γiL2(𝒵)\gamma_{i}\in L^{2}(\mathcal{Z}). Since the map (y,z)(w(y)γi)(z)(y,z)\mapsto(w(y)\gamma_{i})(z) is in L2(𝒳)L^{2}(\mathcal{X}), we obtain w(y)γiγiw(y)\gamma_{i}\gamma_{i}^{\prime}\in\mathcal{M}. Thus, by regarding ww as an element in (L2(𝒵),L2(𝒳))\mathcal{B}(L^{2}(\mathcal{Z}),L^{2}(\mathcal{X})) defined as u((y,z)(w(y)u)(z))u\mapsto((y,z)\mapsto(w(y)u)(z)), by Lemma 29, ww\in\mathcal{M} holds. ∎

Proof of Proposition 1.

Since (y,z)(KT(va)(y)u)(z)=v(h(y))(au)(g(y,z))(y,z)\mapsto(K_{T}(v\otimes a)(y)u)(z)=v(h(y))(au)(g(y,z)) is in L2(𝒳)L^{2}(\mathcal{X}), by Lemma 30, KT(va)K_{T}(v\otimes a)\in\mathcal{M} holds.

Regarding the unitarity of KTK_{T}, let LT:L_{T}:\mathcal{M}\to\mathcal{M} be a right 𝒜\mathcal{A}-linear operator defined as LT(va)=v(h1(y))Ug(h1(y),)aL_{T}(v\otimes a)=v(h^{-1}(y))U_{g(h^{-1}(y),\cdot)}^{*}a. Then, LTL_{T} is the inverse of KTK_{T} and for v1,v2L2(𝒴)v_{1},v_{2}\in L^{2}(\mathcal{Y}) and a1,a2𝒜a_{1},a_{2}\in\mathcal{A}, we have

KT(v1a1),KT(v2a2)=y𝒴v1(h(y))¯a1Ug(y,)Ug(y,)a2v2(h(y))dμ(y)\displaystyle\left\langle K_{T}(v_{1}\otimes a_{1}),K_{T}(v_{2}\otimes a_{2})\right\rangle_{\mathcal{M}}=\int_{y\in\mathcal{Y}}\overline{v_{1}(h(y))}a_{1}^{*}U_{g(y,\cdot)}^{*}U_{g(y,\cdot)}a_{2}v_{2}(h(y))\mathrm{d}\mu(y)
=v1,v2L2(𝒴)a1a2=v1a1,v2a2.\displaystyle\qquad\qquad=\left\langle v_{1},v_{2}\right\rangle_{L^{2}(\mathcal{Y})}a_{1}^{*}a_{2}=\left\langle v_{1}\otimes a_{1},v_{2}\otimes a_{2}\right\rangle_{\mathcal{M}}.

References

  • \bibcommenthead
  • Koopman [1931] Koopman, B.: Hamiltonian systems and transformation in Hilbert space. Proc. Natl. Acad. Sci. 17(5), 315–318 (1931) https://doi.org/10.1073/pnas.17.5.315
  • Koopman and von Neumann [1932] Koopman, B.O., Neumann, J.: Dynamical systems of continuous spectra. Proc. Natl. Acad. Sci. 18(3), 255–263 (1932) https://doi.org/10.1073/pnas.18.3.255
  • Baladi [2000] Baladi, V.: Positive Transfer Operators and Decay of Correlations. Advanced Series in Nonlinear Dynamics, vol. 16. World Scientific, Singapore (2000)
  • Eisner et al. [2015] Eisner, T., Farkas, B., Haase, M., Nagel, R.: Operator Theoretic Aspects of Ergodic Theory. Graduate Texts in Mathematics, vol. 272. Springer, Cham (2015)
  • Froyland [1997] Froyland, G.: Computer-assisted bounds for the rate of decay of correlations. Commun. Math. Phys. 189(NN), 237–257 (1997) https://doi.org/10.1007/s002200050198
  • Dellnitz and Junge [1999] Dellnitz, M., Junge, O.: On the approximation of complicated dynamical behavior. SIAM J. Numer. Anal. 36, 491 (1999) https://doi.org/10.1137/S0036142996313002
  • Dellnitz et al. [2000] Dellnitz, M., Froyland, G., Sertl, S.: On the isolated spectrum of the Perron–Frobenius operator. Nonlinearity 13, 1171–1188 (2000) https://doi.org/10.1088/0951-7715/13/4/310
  • Mezić [2005] Mezić, I.: Spectral properties of dynamical systems, model reduction and decompositions. Nonlinear Dyn. 41, 309–325 (2005) https://doi.org/10.1007/s11071-005-2824-x
  • Giannakis et al. [2015] Giannakis, D., Slawinska, J., Zhao, Z.: Spatiotemporal feature extraction with data-driven Koopman operators. In: Storcheus, D., Rostamizadeh, A., Kumar, S. (eds.) Proceedings of the 1st International Workshop on Feature Extraction: Modern Questions and Challenges at NIPS 2015. Proceedings of Machine Learning Research, vol. 44, pp. 103–115. PMLR, Montreal, Canada (2015). https://proceedings.mlr.press/v44/giannakis15.html
  • Klus et al. [2020] Klus, S., Schuster, I., Muandet, K.: Eigendecompositions of transfer operators in reproducing kernel Hilbert spaces. J. Nonlinear Sci. 30, 283–315 (2020) https://doi.org/10.1007/s00332-019-09574-z
  • Ishikawa et al. [2018] Ishikawa, I., Fujii, K., Ikeda, M., Hashimoto, Y., Kawahara, Y.: Metric on nonlinear dynamical systems with Perron-Frobenius operators. In: Proceedings of Advances in Neural Information Processing Systems 32 (NeurIPS) (2018)
  • Hashimoto et al. [2020] Hashimoto, Y., Ishikawa, I., Ikeda, M., Matsuo, Y., Kawahara, Y.: Krylov subspace method for nonlinear dynamical systems with random noise. J. Mach. Learn. Res. 21(172), 1–29 (2020)
  • Schmid [2010] Schmid, P.J.: Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech. 656, 5–28 (2010) https://doi.org/10.1017/S0022112010001217
  • Rowley et al. [2009] Rowley, C.W., Mezić, I., Bagheri, S., Schlatter, P., Henningson, D.S.: Spectral analysis of nonlinear flows. J. Fluid Mech. 641, 115–127 (2009) https://doi.org/10.1017/s0022112009992059
  • Williams et al. [2015] Williams, M.O., Kevrekidis, I.G., Rowley, C.W.: A data–driven approximation of the Koopman operator: extending dynamic mode decomposition. J. Nonlinear Sci. 25, 1307–1346 (2015) https://doi.org/10.1007/s00332-015-9258-5
  • Kawahara [2016] Kawahara, Y.: Dynamic mode decomposition with reproducing kernels for Koopman spectral analysis. In: Proceedings of Advances in Neural Information Processing Systems 30 (NIPS) (2016)
  • Arbabi and Mezić [2017] Arbabi, H., Mezić, I.: Ergodic theory, dynamic mode decomposition, and computation of spectral properties of the Koopman operator. SIAM J. Appl. Dyn. Syst. 16(4), 2096–2126 (2017) https://doi.org/%****␣main_sn.bbl␣Line␣300␣****10.1137/17M1125
  • Rosenfeld et al. [2022] Rosenfeld, J.A., Kamalapurkar, R., Gruss, L.F., Johnson, T.T.: Dynamic mode decomposition for continuous time systems with the Liouville operator. J. Nonlinear Sci. 32 (2022) https://doi.org/10.1007/s00332-021-09746-w
  • Korda et al. [2020] Korda, M., Putinar, M., Mezić, I.: Data-driven spectral analysis of the Koopman operator. Appl. Comput. Harmon. Anal. 48(2), 599–629 (2020) https://doi.org/10.1016/j.acha.2018.08.002
  • Slipantschuk et al. [2020] Slipantschuk, J., Bandtlow, O.F., Just, W.: Dynamic mode decomposition for analytic maps. Commun. Nonlinear Sci. Numer. Simul. 84, 105179 (2020) https://doi.org/10.1016/j.cnsns.2020.105179
  • Colbrook and Townsend [2021] Colbrook, M.J., Townsend, A.: Rigorous Data-Driven Computation of Spectral Properties of Koopman Operators for Dynamical Systems (2021). https://arxiv.org/abs/2111.14889
  • Das et al. [2021] Das, S., Giannakis, D., Slawinska, J.: Reproducing kernel Hilbert space compactification of unitary evolution groups. Appl. Comput. Harmon. Anal. 54, 75–136 (2021) https://doi.org/%****␣main_sn.bbl␣Line␣375␣****10.1016/j.acha.2021.02.004
  • Ulam [1964] Ulam, S.M.: Problems in Modern Mathematics. Dover Publications, Mineola (1964)
  • Blank et al. [2002] Blank, M., Keller, G., Liverani, C.: Ruelle–Perron–Frobenius spectrum for Anosov maps. Nonlinearity 15(6), 1905–1973 (2002) https://doi.org/10.1088/0951-7715/15/6/309
  • Junge and Koltai [2009] Junge, O., Koltai, P.: Discretization of the Frobenius–Perron operator using a sparse Haar tensor basis: The sparse Ulam method. SIAM J. Numer. Anal. 47, 3464–2485 (2009) https://doi.org/10.1137/080716864
  • Froyland et al. [2010] Froyland, G., Santitissadeekorn, N., Monahan, A.: Transport in time-dependent dynamical systems: Finite-time coherent sets. Chaos: An Interdisciplinary Journal of Nonlinear Science 20(4), 043116 (2010) https://doi.org/10.1063/1.3502450
  • Froyland and Koltai [2017] Froyland, G., Koltai, P.: Estimating long-term behavior of periodically driven flows without trajectory integration. Nonlinearity 30(5), 1948 (2017) https://doi.org/10.1088/1361-6544/aa6693
  • Giannakis and Das [2020] Giannakis, D., Das, S.: Extraction and prediction of coherent patterns in incompressible flows through space-time Koopman analysis. Physica D: Nonlinear Phenomena 402, 132211 (2020) https://doi.org/10.1016/j.physd.2019.132211
  • Oseledets [1968] Oseledets, V.I.: A multiplicative ergodic theorem. Trans. Moscow Math. Soc. 19, 197–231 (1968)
  • Ruelle [1968] Ruelle, D.: Statistical mechanics of a one-dimensional lattice gas. Comm. Math. Phys. 9(4), 267–278 (1968)
  • Thieullen [1987] Thieullen, P.: Fibres dynamiques asymptotiquement compacts exposants de lyapounov. entropie. dimension. Annales de l’Institut Henri Poincaré C, Analyse non linéaire 4(1), 49–97 (1987) https://doi.org/%****␣main_sn.bbl␣Line␣500␣****10.1016/S0294-1449(16)30373-0
  • Schaumlöffel [1991] Schaumlöffel, K.: Multiplicative ergodic theorems in infinite dimensions. In: Lyapunov Exponents. Lecture Notes in Mathematics, vol. 1486. Springer, Heidelberg (1991)
  • Froyland et al. [2010] Froyland, G., Lloyd, S., Quas, A.: Coherent structures and isolated spectrum for Perron–Frobenius cocycles. Ergod. Theory Dyn. Syst. 30(3), 729–756 (2010) https://doi.org/10.1017/S0143385709000339
  • González-Tokman and Quas [2014] González-Tokman, C., Quas, A.: A semi-invertible operator oseledets theorem. Ergod. Theory Dyn. Syst. 34(4), 1230–1272 (2014) https://doi.org/10.1017/etds.2012.189
  • Froyland et al. [2010] Froyland, G., Lloyd, S., Santitissadeekorn, N.: Coherent sets for nonautonomous dynamical systems. Physica D: Nonlinear Phenomena 239(16), 1527–1541 (2010) https://doi.org/10.1016/j.physd.2010.03.009
  • Froyland and Junge [2018] Froyland, G., Junge, O.: Robust FEM-based extraction of finite-time coherent sets using scattered, sparse, and incomplete trajectories. SIAM J. Appl. Dyn. Syst. 17(2), 1891–1924 (2018) https://doi.org/10.1137/17M1129738
  • Froyland [2015] Froyland, G.: Dynamic isoperimetry and the geometry of Lagrangian coherent structures. Nonlinearity 28(10), 3587 (2015) https://doi.org/10.1088/0951-7715/28/10/3587
  • Lance [1995] Lance, E.C.: Hilbert CC^{*}-modules – a Toolkit for Operator Algebraists. London Mathematical Society Lecture Note Series, vol. 210. Cambridge University Press, New York (1995)
  • Froyland and Koltai [2023] Froyland, G., Koltai, P.: Detecting the birth and death of finite-time coherent sets. Commun. Pure Appl. Math (2023) https://doi.org/10.1002/cpa.22115
  • Hashimoto et al. [2023] Hashimoto, Y., Komura, F., Ikeda, M.: Hilbert CC^{*}-module for analyzing structured data. In: Matrix and Operator Equations. Springer, Cham (2023). https://doi.org/10.1007/16618_2023_58
  • Choe [1985] Choe, Y.H.: C0C_{0}-semigroups on a locally convex space. J. Math. Anal. Appl. 106(2), 293–320 (1985) https://doi.org/10.1016/0022-247X(85)90115-5
  • Eisner et al. [2016] Eisner, T., Farkas, B., Haase, M., Nagel, R.: Operator Theoretic Aspects of Ergodic Theory. Springer, Cham (2016)
  • Giannakis et al. [2022] Giannakis, D., Ourmazd, A., Pfeffer, P., Schumacher, J., Slawinska, J.: Embedding classical dynamics in a quantum computer. Phys. Rev. A 105, 052404 (2022) https://doi.org/10.1103/PhysRevA.105.052404
  • Skeide [2000] Skeide, M.: Generalised matrix CC^{\ast}-algebras and representations of Hilbert modules. Math. Proc. Roy. Irish Acad. 100A(1), 11–38 (2000)