Quick viewing(Text Mode)

Arxiv:1911.11996V4 [Math.DS] 31 Mar 2021

Arxiv:1911.11996V4 [Math.DS] 31 Mar 2021

EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS FOR STABLE FIXED POINTS AND PERIODIC ORBITS

MATTHEW D. KVALHEIM AND SHAI REVZEN

Abstract. We consider C1 dynamical systems having an attracting hyperbolic fixed point or periodic orbit k k,α and prove existence and uniqueness results for C (actually Cloc ) linearizing semiconjugacies—of which Koopman eigenfunctions are a special case—defined on the entire basin of attraction. Our main results both generalize and sharpen Sternberg’s Ck linearization theorem for hyperbolic sinks, and in particular our corollaries include uniqueness statements for Sternberg linearizations and Floquet normal forms. Using our main results we also prove new existence and uniqueness statements for Ck Koopman eigenfunctions, including a complete classification of C∞ eigenfunctions assuming a C∞ dynamical system with semisimple and nonresonant linearization. We give an intrinsic definition of “principal Koopman eigenfunctions” which generalizes the definition of Mohr and Mezić for linear systems, and which includes the notions of “isostables” and “isostable coordinates” appearing in work by Ermentrout, Mauroy, Mezić, Moehlis, Wilson, and others. Our main results yield existence and uniqueness theorems for the principal eigenfunctions and isostable coordinates and also show, e.g., that the (a priori non-unique) “pullback algebra” defined in [MM16b] is unique under certain conditions. We also discuss the limit used to define the “faster” isostable coordinates in [WE18, MWMM19] in light of our main results.

Contents 1. Introduction 1 1.1. Nontechnical overview of results2 1.2. Technical overview of results and organization of the paper3 1.3. Notation and terminology5 2. Main results 5 3. Applications 11 3.1. Sternberg linearizations and Floquet normal forms 11 3.2. Principal Koopman eigenfunctions, isostables, and isostable coordinates 13 3.3. Classification of all C∞ Koopman eigenfunctions 19 4. Proofs of the main results 21 4.1. Proof of Theorem1 21 4.2. Proof of Theorem2 32 Acknowledgments 32 References 33 arXiv:1911.11996v4 [math.DS] 31 Mar 2021

1. Introduction This paper fills a significant technical gap between the linearization results known from classical dy- namical systems theory—e.g., the linearization theorems of Poincaré-Siegel [Poi79, Sie42, Sie52], Sternberg [Ste57], Grobman-Hartman [Gro59, Har60a], and Hartman [Har60b]—and the growing interest in applied

Department of Electrical and Systems Engineering, University of Pennsylvania, Philadelphia, PA 19104 Department of Electrical Engineering and Computer Science, Ecology and Evolutionary Biology Depart- ment, Robotics Institute, University of Michigan, Ann Arbor, MI 48109 E-mail addresses: [email protected], [email protected]. 1 2 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

fields such as engineering and fluid dynamics in using linearizations based on Koopman theory.1 Moti- vated largely by data-driven applications, this “applied Koopmanism” literature has experienced a surge of interest initiated by [Mez94, MB04, Mez05] more than 70 years after Koopman’s seminal work [Koo31].2 The practical application of computational Koopman eigenfunction representations of dynamical sys- tems is grounded in (i) eigenfunction existence theorems based on such classical linearization theorems [LM13, MMM13] and (ii) uniqueness theorems applying only to analytic eigenfunctions for certain an- alytic dynamical systems [MMM13]. Existence and uniqueness results are desirable since, in analyzing the theoretical properties of any algorithm for computing some quantity, it is desirable to know whether the computation is well-posed [Had02], and in particular whether the quantity in question exists and is uniquely determined. The results in the present paper yield new precise conditions under which various quantities in the applied Koopmanism literature—including targets of numerical algorithms—exist and are unique, and are especially relevant to work on principal eigenfunctions and isostables for point attractors [MM16b, MMM13] and to work on isostable coordinates for periodic orbit (limit cycle) attractors [WM16a, SKN17, WE18, MWMM19]. Isostables and isostable coordinates are useful tools for nonlinear model reduction, and it has been proposed that they could prove useful in real-world applications such as treatment design for Parkinson’s disease, migraines, cardiac arrhythmias [WM16b], and jet lag [WM14].

1.1. Nontechnical overview of results. This paper was motivated by the following three questions. (A Ck is one which has continuous mixed partial derivatives up to order k.) Eigenfunction uniqueness: When—and in what sense—are Ck (1 ≤ k ≤ +∞) Koopman eigen- functions unique? Linearization uniqueness: When—and in what sense—are full Ck (1 ≤ k ≤ +∞) linearizing coordinate changes unique? Eigenfunction existence: Can the existence of specific Ck (2 ≤ k ≤ +∞) Koopman eigenfunctions be guaranteed under assumptions which are weaker than those needed to invoke a classical result guaranteeing a full set of linearizing coordinates exists? We provide answers to each of these three questions for eigenfunctions and coordinate changes defined on the basin of an attracting hyperbolic fixed point or periodic orbit. To the best of our knowledge, our answers to the eigenfunction uniqueness and existence questions are new. During the review process of this paper we discovered that an answer to the full linearization uniqueness question for the case k = +∞ (equivalent to our answer in that case) can also be obtained from results in the recent book [FH19, Thm 6.8.21, 6.8.22]. Sternberg’s work showed that, for Ck (1 ≤ k ≤ +∞) dynamical systems such as those arising from a Ck ordinary differential equation d x(t) = f(x(t)) x(t) ∈ n dt R n having an attracting fixed point x0 ∈ R , there is an intimate relationship between the eigenvalues of k the derivative Dx0 f of f at x0 and the existence of C linearizing coordinates defined near x0 [Ste57]. Our results establish a similarly intimate relationship between certain eigenvalues and the answers to our three questions. In the case of the eigenfunction uniqueness and existence questions, we show that the corresponding relationship is less restrictive than in Sternberg’s case. In particular, the answer to the eigenfunction existence question is “yes” (especially in the case of the “slowest” principal eigenfunc- tions/isostable coordinates). The answer to the linearization uniqueness question is this: under Sternberg’s hypotheses guaranteeing that Ck linearizing coordinates exist in the vicinity (hence also on the entire basin [LM13, EKR18]) of an attracting hyperbolic fixed point, these coordinates are uniquely determined by their derivatives at the fixed point.

1Here “linearization” refers to a nonlinear change of coordinates in which a nonlinear dynamical system becomes exactly linear, and is distinct from approximate linearization. A recent paper extending and discussing some of the state of the art is [New17]. 2See, e.g., [BMM12, MM12, MMM13, LM13, MM14, GSZ15, Mez15, WKR15, MM16a, MM16b, BBPK16, Sur16, SB16, AM17a, AM17b, KKB17, Mez19, PBK18, KM18, KPM18, KM19, DG19, BRV19, DTK19, AT19]. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 3

To obtain the answers to our three questions, we prove a general result on the existence and uniqueness of Ck linearizing semiconjugacies (partial linearizations), of which Koopman eigenfunctions and linearizing conjugacies are special cases. To provide more refined answers to the same three questions, we further k,α k k k,0 examine the Cloc smoothness classes refining the familiar C smoothness classes, wherein C = Cloc . Our main results concern the existence and uniqueness of linearizing semiconjugacies defined on the basin of an attracting hyperbolic fixed point or periodic orbit. Notable consequences worked out in this paper, for dynamics in the basin of an attracting hyperbolic fixed point or periodic orbit, include: • a fairly tight relationship between properties of the system eigenvalues/Floquet multipliers and k,α 1 k,α the uniqueness of Cloc principal eigenfunctions for C dynamics (Proposition6) or Cloc dynamics (Proposition7); k,α • sufficient conditions for the existence of such Cloc principal eigenfunctions when the dynamics are k,α Cloc (Propositions6 and7); k,α • sufficient conditions for when such Cloc eigenfunctions (or isostable coordinates) can be constructed by limiting procedures such as Laplace averages (Propositions6 and7 and Remark 14); • a full classification of all C∞ eigenfunctions in terms of principal eigenfunctions for C∞ dynamics satisfying a nondegeneracy assumption (Theorems3 and4); and • sufficient conditions under which a full set of linearizing Ck coordinates exist and are uniquely determined by their first-order approximation (Propositions2 and3). To the best of our knowledge, our uniqueness results for principal Koopman eigenfunctions are the first uniqueness results known for non-analytic eigenfunctions. Similarly, our classification of all C∞ Koopman eigenfunctions appears to be the first such classification theorem for non-analytic eigenfunctions. While certain existence results for principal Koopman eigenfunctions defined on the basin of an attracting hyperbolic equilibrium or periodic orbit have been known for some time [LM13, MMM13], we believe our new existence results to be the strongest known for Ck eigenfunctions with 2 ≤ k ≤ +∞. This is because prior existence results (cf. [Mez19, Sec. 5, 7, 8]) construct Ck eigenfunctions by pulling back linear eigenfunctions through the linearizing conjugacies provided by the Sternberg or Poincaré-Siegel linearization theorems mentioned above, and the full hypotheses of one of these linearization theorems must be assumed in order to invoke it; in contrast, our existence result for specific principal eigenfunctions requires much weaker assumptions. These assumptions are extremely mild in the case of the “slowest” principal eigenfunctions/isostable coordinates; the following subsection contains a precise statement (with more details in Remark 15) as part of a more technical overview of our results.

1.2. Technical overview of results and organization of the paper. In this paper, we consider C1 dynamical systems Φ: Q × T → Q for which Q is the basin of attraction of a stable hyperbolic fixed point or periodic orbit. Here Q is a smooth , either T = Z or T = R, and Φ could be the restriction of a n dynamical system defined on a larger space (e.g., R ) to some basin of attraction Q. That Φ is a dynamical 0 t+s t s t system means that Φ = idQ and Φ = Φ ◦ Φ for all t, s ∈ T, where Φ := Φ( · , t); in particular, it t 1 −t follows that Φ : Q → Q is a C diffeomorphism with inverse Φ for any t ∈ T. When T = R, Φ is called t a flow; a common example is that of t 7→ Φ (x0) being the solution to the initial value problem d x(t) = f(x(t)), x(0) = x dt 0 determined by a complete C1 vector field f on Q. Our main contributions are existence and uniqueness k,α m results regarding Cloc linearizing semiconjugacies ψ : Q → C defined on the entire basin of attraction Q, where we do not assume that m has any relationship to the dimension of Q; in particular, ψ need not be a diffeomorphism or a . By definition, such a semiconjugacy makes the diagram

t Q Φ Q

(1) ψ ψ m etA m C C 4 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

m×m k,α commute for some A ∈ C and all t ∈ T. By Cloc with k ∈ N≥1 and 0 ≤ α ≤ 1, we mean that k m ψ ∈ C (Q, C ) and that all k-th partial derivatives of ψ are locally α-Hölder continuous in local coordinates, ∞,α +∞,α ∞ k,0 k and by definition Cloc := Cloc := C . We note that Cloc = C , so the reader uninterested in Hölder k,0 k k+1 k,α continuity can simply keep in mind the case Cloc = C and the fact that every C function is also Cloc for every 0 ≤ α ≤ 1. Our motivation for including local Hölder continuity of derivatives is that, once it is included, our main results become fairly close to optimal, at least in the interesting case m = 1 of Koopman eigenfunctions ψ (see Examples2 and3). Linearizing semiconjugacies are also known as linearizing factors or factor maps in the literature and can be viewed as a further generalization of the generalized Koopman eigenfunctions of [Mez19, KM19]. We note that such semiconjugacies are distinct from those in the diagram

t Q Φ Q

(2) K K m etA m C C obtained from (1) by flipping the vertical arrows (although the diagrams are equivalent if, e.g., ψ and K are diffeomorphisms). In (2) K is a factor of etA, whereas ψ is a factor of Φt in (1). Existence results for semiconjugacies of the type in (2) were obtained by [CFdlL03a, CFdlL03b, CFdlL05] in the context of proving invariant manifold results using the parameterization method. Our main result for the case of an attracting hyperbolic fixed point both generalizes and sharpens Sternberg’s linearization theorem [Ste57, Thms 2,3,4] which provides conditions ensuring the existence of a linearizing local Ck diffeomorphism defined on a neighborhood of the fixed point; the results of [LM13, EKR18] show that this local diffeomorphism can be extended to a Ck diffeomorphism ψ : Q → n n R ⊂ C defined on the entire basin of attraction Q making (1) commute. Under Sternberg’s conditions, a corollary of our main result is that this global linearizing diffeomorphism is in fact uniquely determined by its derivative at the fixed point (cf. [FH19, Thm 6.8.21, 6.8.22] for the case k = +∞). Additionally, k k,α we sharpen Sternberg’s result from C to Cloc linearizations. For the case of an attracting hyperbolic periodic orbit of a flow, our main result also yields a similar existence and uniqueness corollary for the Floquet normal form, a nonlinear change of coordinates in which the dynamics become the product of a linear system with constant-rate rotation on a circle [AR67, BK94, AMR]. We remark that the Floquet normal form is a nonlinear generalization of the comparatively well-known classical Floquet theory of linear time-periodic systems [Hal, Sec. III.7] Using our two main results, we make the following contributions to the theory of Koopman eigenfunc- tions. We give an intrinsic definition of principal eigenfunctions for nonlinear dynamical systems which generalizes the definition for linear systems in [MM16b]. We provide existence and uniqueness results k,α for Cloc principal eigenfunctions, and we also show that the (a priori non-unique) “pullback algebra” de- fined in [MM16b] is unique under certain conditions. For the case of periodic orbit attractors, principal eigenfunctions essentially coincide with the notion of isostable coordinates defined in [WE18, MWMM19], except that the definition in these references involves a limit which might not exist except for the “slowest” isostable coordinate. Our techniques shed light on this issue, and our results imply that this limit does in fact always exist for the “slowest” isostable coordinate if the dynamical system is at least smoother than 1,α Cloc with α > 0. In fact, our results imply—assuming that there is a unique and algebraically simple 1,α “slowest” Floquet multiplier which is real—that a corresponding “slowest” Cloc isostable coordinate with 1,α 2 α > 0 always exists and is unique modulo scalar multiplication for a Cloc dynamical system (e.g., a C dynamical system), without the need for any nonresonance or spectral spread assumptions. Similarly, if instead there is a unique and algebraically simple “slowest” pair of Floquet multipliers which are complex 1,α conjugates, then a corresponding “slowest” complex conjugate pair of Cloc isostable coordinates always 1,α exists and is unique modulo scalar multiplication for a Cloc dynamical system with α > 0. As a final application of our main results, we give a complete classification of C∞ eigenfunctions for a C∞ dynamical system with semisimple (diagonalizable over C) and nonresonant linearization, generalizing known results for analytic dynamics and analytic eigenfunctions [MMM13, Mez19]. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 5

The remainder of the paper is organized as follows. We explain notation and terminology below to be used in the sequel. After some definitions, in §2 we state Theorems1 and2, our two main results, without proof. We also state a proposition on the uniqueness of linearizing factors which does not assume any nonresonance conditions. As applications we derive in §3 several results which are essentially corollaries of this proposition and the two main theorems. §3.1 contains existence and uniqueness theorems for global Sternberg linearizations and Floquet normal forms. In §3.2 we define principal Koopman eigenfunctions and isostable coordinates for nonlinear dynamical systems and discuss how Theorems1 and2 yield cor- responding existence and uniqueness results. We then discuss the relationship between various notions defined in [MM16b] and our definitions, and we also discuss the convergence of the isostable coordinate limits in [WE18, MWMM19]. §3.3 contains our theorem which completely classifies the C∞ eigenfunctions of C∞ dynamical systems on the basin of an attracting hyperbolic fixed point or periodic orbit. Finally, §4 contains the proofs of Theorems1 and2. 1.3. Notation and terminology. In this paper we employ the following mostly standard notation and terminology.

1.3.1. Sets of numbers. We denote the real numbers by R, complex numbers by C, integers by Z, and nonnegative integers by N. Given c ∈ R and S ⊂ R, we define S≥c := {s ∈ S : s ≥ c} and S>c := {s ∈ S : s > c} so that, e.g., Z≥0 = N≥0 = N. m×m 1.3.2. Linear algebra. Given m ∈ N≥1, we denote by GL(m, C) ⊂ C the invertible m × m matrices m×m with entries in C and by GL(m, R) ⊂ GL(m, C) those with entries in R. Given A ∈ C , we denote by m×m spec(A) ⊂ C the set of eigenvalues of A; given A ∈ R , we denote by spec(A) ⊂ C the set of eigenvalues m×m m×m of A ∈ R ⊂ C when viewed as a complex matrix. If A: V → V is a linear self-map with V a complex vector space, spec(A) ⊂ C also denotes the eigenvalues of A. If V is a real vector space, then its complexification VC is the complex vector space given by all formal linear combinations of vectors in V with complex coefficients (cf. [HS74, p. 64]); if A: V → W is an R- between real vector spaces, then the complexification AC : VC → WC of A is the unique C-linear extension of A (cf. [HS74, p. 65]), and if W = V we define spec(A) := spec(AC) ⊂ C. If E1,E2 ⊂ V are linear subspaces of a real or complex vector space V , we say that E1 and E2 are complementary if V = {e1 + e2 : e1 ∈ E1, e2 ∈ E2} and if E1 ∩ E2 = {0}. 1.3.3. Derivatives. Given a differentiable map F : M → N between smooth , we use the notation DxF for the derivative of F at the point x ∈ M. (Recall that the derivative DxF : TxM → TF (x)N is a linear map between tangent spaces [Lee13], which can be identified with the Jacobian of F evaluated at x in local coordinates.) In particular, given a dynamical system Φ: Q × T → Q and fixed t ∈ T, we t t write DxΦ : TxQ → TΦt(x)Q for the derivative of the time-t map Φ : Q → Q at the point x ∈ Q. A map k k F : M → N satisfies F ∈ C (M,N) with k ∈ N≥0 ∪ {+∞}, or briefly F ∈ C , if every x ∈ M is contained in a coordinate chart in which all mixed partial derivatives of F of order less than k + 1 exist and are 0 continuous. For convenience, we define the 0-th derivative DxF := F (x) to coincide with F for all x ∈ M. i Several of our results include conditions such as “Dx0 F = 0 for all integers 0 ≤ i < r.” This is to be interpreted to mean that, in local coordinates, all mixed partial derivatives of F of order less than r vanish j ⊗j at x0. This can be made more formal in the following way. Inductively, if i ≥ 2 and Dx0 F :(Tx0 M) → i ⊗i TF (x0)N is well-defined and zero for all 1 ≤ j ≤ i−1, then the i-th derivative Dx0 F :(Tx0 M) → TF (x0)N is ⊗i a well-defined linear map from the i-th tensor power (Tx0 M) to TF (x0)N represented in local coordinates 3 by the (1 + i)-dimensional array of i-th partial derivatives of F evaluated at x0.

2. Main results Before stating our main results, we give two definitions which are essentially asymmetric versions of some appearing in [Ste57, Sel85]. When discussing eigenvalues and eigenvectors of a matrix or linear self-map

3 i To define Dx0 F , use local coordinates. The inductive assumption that the first i − 1 derivatives are well-defined and zero at x0 ensures that the result is independent of the choice of local coordinates, hence well-defined. 6 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

(endomorphism) in the remainder of the paper, we are always discussing eigenvalues and eigenvectors of its complexification, although we do not always make this explicit.

d×d n×n Definition 1 ((X,Y ) k-nonresonance). Let X ∈ C and Y ∈ C be matrices with eigenvalues µ1, . . . , µd and λ1, . . . , λn, respectively, repeated with multiplicities. For any k ∈ N≥1 ∪ {+∞}, we say n that (X,Y ) is k-nonresonant if, for any i ∈ {1, . . . , d} and any m = (m1, . . . , mn) ∈ N≥0 satisfying 2 ≤ m1 + ··· + mn < k + 1,

m1 mn (3) µi 6= λ1 ··· λn . (Note this condition vacuously holds if k = 1; i.e., any two matrices are 1-nonresonant.) We also extend the definition of k-nonresonance to general linear self-maps X,Y of finite-dimensional complex vector spaces by identifying X,Y with their matrix representations with respect to any choice of bases. We say that linear self-maps (X,Y ) of finite-dimensional real vector spaces are k-nonresonant if the complexifications (XC,YC) are k-nonresonant. For the definition below, recall that the spectral radius ρ(X) of a matrix is defined to be the largest modulus (absolute value) of the eigenvalues of (the complexification of) X.

Definition 2 ((X,Y ) spectral spread). Let X ∈ GL(m, C) and Y ∈ GL(n, C) be invertible matrices with the spectral radius ρ(Y ) satisfying ρ(Y ) < 1. We define the spectral spread ν(X,Y ) to be ln(|µ|) ν(X,Y ) := max µ∈spec(X) ln(|λ|) λ∈spec(Y ) ( ! !) (4) r = min r ∈ R: min |µ| ≥ max |λ| µ∈spec(X) λ∈spec(Y )

n −1 r o = min r ∈ R: ρ(X )(ρ(Y )) ≤ 1 . We also extend the definition of ν(X,Y ) to general linear automorphisms X,Y of finite-dimensional complex vector spaces by identifying X,Y with their matrix representations with respect to any choice of bases. We extend the definition to linear automorphisms (X,Y ) of real vector spaces by defining ν(X,Y ) := ν(XC,YC) to be the spectral spread of the complexifications. The second line in (4) follows since ln(|µ|)/ ln(|λ|) ≤ r if and only if ln(|µ|) ≥ r ln(|λ|), with the inequality flipping because ln(|λ|) < 0 since |λ| ≤ ρ(Y ) < 1, and this in turn holds if and only if |µ| ≥ |λ|r. The third line follows from the definition of the spectral radius ρ(Y ). Figure1 illustrates the condition A 1 k,α ν(e , Dx0 Φ ) < k + α in Theorem1 below. Finally, we now recall the definition of Cloc functions. k,α k Definition 3 (Cloc functions). Let M,N be smooth manifolds of dimensions m and n, let ψ ∈ C (M,N) k k,α be a C map ψ : M → N with k ∈ N≥0, and let 0 ≤ α ≤ 1. We will say that ψ ∈ Cloc (M,N) if for every x ∈ M there exist charts (U1, ϕ1) and (U2, ϕ2) containing x and ψ(x) such that all k-th partial −1 derivatives of ϕ2 ◦ ψ ◦ ϕ1 are Hölder continuous with exponent α. If k = +∞, we use the convention +∞,α ∞ Cloc (M,N) := C (M,N) for any 0 ≤ α ≤ 1. If the domain and M and N are clear from k k,α k k,α k context, we will sometimes write C and Cloc instead of C (M,N) and Cloc (M,N) and write, e.g., ψ ∈ C k,α k,β k,α k,0 or ψ ∈ Cloc . We note that Cloc ⊂ Cloc for any k ∈ N ∪ {+∞} and 0 ≤ α ≤ β ≤ 1, and that ψ ∈ Cloc if and only if ψ ∈ Ck. Remark 1. Using the chain rule and the fact that compositions and products of locally α-Hölder contin- k,α uous functions are again locally α-Hölder, it follows that the property of being Cloc on a manifold does not depend on the choice of charts in Definition3. We now state our main results, Theorems1 and2, as well as Proposition1. An example and several clarifying remarks follow the statement of Theorem1; in particular, see Remark2 for intuitive remarks and Example1 for concreteness. Simple analytic (counter-)examples demonstrating Theorems1 and2 in EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 7 Im

1   k+α minλ∈spec(eA) |λ|

Re

1 spec(Dx0 Φ )

A 1 Figure 1. An illustration of the condition ν(e , Dx0 Φ ) < k + α of Theorem1. This 1 condition is equivalent to every eigenvalue of Dx0 Φ (represented by an “×” above) belonging to the open disk with radius given by raising the smallest modulus of the eigenvalues of eA 1 to the power k+α . the case m = 1 of Koopman eigenfunctions—demonstrating in particular the sharpness of the uniqueness statement—are Examples2 and3 of §3.2. We emphasize that Theorems1 and2 do not assume that the linear map B is invertible, and do not m claim anything about the semiconjugacy ψ : Q → C being a diffeomorphism (but see Propositions2 and 3); moreover, nothing is assumed about the relationship of m to dim(Q).

k,α Theorem 1 (Existence and uniqueness of Cloc global linearizing factors for a point attractor). Let Φ: Q× 1 T → Q be a C dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q A is a smooth manifold with dim(Q) ≥ 1 and either T = Z or T = R. Let m ∈ N≥1 and e ∈ GL(m, C) have A m spectral radius ρ(e ) < 1, and let the linear map B : Tx0 Q → C satisfy t tA (5) ∀t ∈ T: BDx0 Φ = e B. A 1 Fix k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1, assume that (e , Dx0 Φ ) is k-nonresonant, and assume that A 1 ν(e , Dx0 Φ ) < k + α. k,α m Uniqueness. Any ψ ∈ Cloc (Q, C ) satisfying 1 A ψ ◦ Φ = e ψ, Dx0 ψ = B m m A m m is unique, and if B : Tx0 Q → R ⊂ C and e ∈ GL(m, R) ⊂ GL(m, C) are real, then ψ : Q → R ⊂ C is real. k,α k,α m Existence. If furthermore Φ ∈ Cloc , then such a unique ψ ∈ Cloc (Q, C ) exists and additionally satisfies t tA (6) ∀t ∈ T: ψ ◦ Φ = e ψ. k,α m In fact, if P ∈ Cloc (Q, C ) is any “approximate linearizing factor” satisfying Dx0 P = B and (7) P ◦ Φ1 = eAP + R i with Dx0 R = 0 for all integers 0 ≤ i < k + α, then (8) ψ = lim e−tAP ◦ Φt t→∞ in the of Ck,α-uniform convergence on compact subsets of Q if k < +∞, and in the topology of k0 0 C -uniform convergence on compact subsets of Q for any k ∈ N≥1 if k = +∞. 8 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

A 1 Remark 2. The spectral spread condition ν(e , Dx0 Φ ) < k + α means that, if λ is any eigenvalue 1 k+α A of Dx0 Φ , then |λ| is smaller than |µ| for every eigenvalue µ of e . The k-nonresonance condition means that no eigenvalue µ of eA can be written as a product (with repetitions allowed) of ` ∈ {2, . . . , k} 1 eigenvalues of Dx0 Φ . The uniqueness statement of Theorem1 then says that, under these two conditions, k,α m any linearizing semiconjugacy ψ ∈ Cloc (Q, C ) is uniquely determined by its derivative at the fixed k,α 1 point x0. Under the additional assumption that Φ ∈ Cloc rather than merely Φ ∈ C , the existence statement of Theorem1 gives sufficient conditions ensuring that, given a linear linearizing semiconjugacy m t B : Tx0 Q → C for the linear dynamical system (v, t) 7→ Dx0 Φ · v, there exists a unique nonlinear k,α m linearizing semiconjugacy ψ ∈ Cloc (Q, C ) for the nonlinear dynamical system Φ satisfying Dx0 ψ = B. Thus, the existence statement can be thought of as supplying sufficient conditions under which a linearizing semiconjugacy can be constructed from an “infinitesimal” one.

Remark 3 (Weaker nonresonance assumption in the case T = R). Assume T = R in the setting of t+s t s Theorem1. Differentiating the identity Φ = Φ ◦ Φ at x0 yields t+s t s (9) ∀t, s ∈ R: Dx0 Φ = Dx0 Φ ◦ Dx0 Φ . 1 t If Φ ∈ C , then the map t 7→ Dx0 Φ is, a priori, merely continuous. However, continuity together with (9) actually implies the existence of a linear map J : Tx0 Q → Tx0 Q such that t tJ (10) ∀t ∈ R: Dx0 Φ = e .

See [EN00, Thm 2.9] and [EN00, p. 9, para. 1]. Let λ1, . . . , λn be the eigenvalues of (the complexification of) J, repeated with multiplicities. Taking the natural logarithm of (3), we see that k-nonresonance of A 1 n (e , Dx0 Φ ) means that, for any (possibly complex) eigenvalue µ of A, any n-tuple (m1, . . . , mn) ∈ N≥0 satisfying 2 ≤ m1 + ··· + mn < k + 1, and any ` ∈ Z,

(11) µ 6= m1λ1 + ··· + mnλn + i2π`, √ where i = −1. If c > 0 and we define the rescaled linear map Ac := cA and the time-rescaled flow ct t ct ctA tAc t ct ctA tAc Φc := Φ , then we see that BDx0 Φc = BDx0 Φ = e B = e B and ψ ◦ Φc = ψ ◦ Φ = e ψ = e ψ for all t ∈ R. Thus, B and ψ satisfy (5) and (6) with T = R if and only if B and ψ satisfy (5) and (6) with Ac 1 A and Φ replaced by Ac and Φc for every c > 0. Now, the k-nonresonance condition for (e , Dx0 Φc ) is obtained from (11) by multiplying the eigenvalues µ and λ1, . . . , λn by c; dividing by c then yields 2π (12) µ 6= m λ + ··· + m λ + i `. 1 1 n n c Because there are only finitely many eigenvalues of A and J,(12) can be violated for all c > 0 if and only if it is violated with ` = 0. It follows that, if T = R, the k-nonresonance assumption in Theorem1 (as well as in Theorem3 and Propositions2 and6) can be replaced with the less restrictive condition

(13) µ 6= m1λ1 + ··· + mnλn n for all (m1, . . . , mn) ∈ N≥0 satisfying 2 ≤ m1 + ··· mn < k + 1 and all (possibly complex) eigenvalues µ of A, where λ1, . . . , λn are the eigenvalues of (the complexification of) J, repeated with multiplicity. n Example 1. Consider the setting of Theorem1 in the special case that Q ⊂ R , T = R, and with Φ the flow of the ordinary differential equation dx = f(x) dt 1 m×n with f ∈ C a complete vector field, so that f(x0) = 0. Let B ∈ C be any matrix such that m×m BDx0 f = AB for some A ∈ C . (For example, in the case m = 1, B is a left eigenvector of Dx0 f if B 6= 0.) It follows that Bh(tDx0 f) = h(tA)B for any analytic function h and t ∈ R, so in particular tDx0 f tA t tDx0 f t tA Be = e B. Since Dx0 Φ = e for all t ∈ R (because f(x0) = 0), it follows that BDx0 Φ = e B for all t ∈ R. Thus, Theorem1 can be applied in this situation as long as the spectral spread and nonresonance 1 Dx0 f conditions are satisfied. Since Dx0 Φ = e in the present setting, satisfaction of the spectral spread A 1 condition ν(e , Dx0 Φ ) < k + α according to Definition2 (after taking logarithms) means that, if λ is any EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 9

(possibly complex) eigenvalue of Dx0 f, and if µ is any (possibly complex) eigenvalue of A, then the real part Re(λ) < 0 satisfies (k + α)Re(λ) < Re(µ). Satisfaction of the k-nonresonance condition means that, for any (possibly complex) eigenvalue µ of A, any n n-tuple (m1, . . . , mn) ∈ N≥0 satisfying 2 ≤ m1 + ··· + mn < k + 1, and any ` ∈ Z,

(14) µ 6= m1λ1 + ··· + mnλn + i2π`, √ where λ1, . . . , λn are the (possibly complex) eigenvalues of Dx0 f, repeated with multiplicity, and i = −1. By Remark3, the k-nonresonance condition of Theorem1 can actually be replaced with the less restrictive condition

(15) µ 6= m1λ1 + ··· + mnλn in the setting of the present example. To make easier the application of the existence portion of Theorem k,α k,α 1, we note that if the vector field f ∈ Cloc , then also the flow Φ ∈ Cloc [Eld13, Thm A.6]. Remark 4. Definitions1 and2 are not independent. In particular, if (X,Y ) is (` − 1)-nonresonant and ν(X,Y ) < ` for ` ∈ N≥2, then it follows that (X,Y ) is ∞-nonresonant. Hence an equivalent statement of Theorem1 could be obtained by replacing k-nonresonance with ∞-resonance everywhere (alternatively, for the existence statement only (k − 1)-nonresonance need be assumed in the case α = 0). We prefer to use the stronger-sounding statement of the theorem above since it makes it clear that the set of matrix A 1 pairs (e , Dx0 Φ ) satisfying its hypotheses are open in the space of all matrix pairs. Openness for k < +∞ A 1 is immediate, and openness for k = +∞ follows the fact that ν(e , Dx0 Φ ) is always finite. Remark 5. The statement in Theorem1 regarding the limit in (8) actually holds without any nonresonance k,α m assumptions if an approximate linearizing factor P ∈ Cloc (Q, C ) satisfying (7) exists; see Lemma5 in §4.1.2. ∞ A 1 Remark 6 (the C case). In the case that k = +∞, the hypothesis ν(e , Dx0 Φ ) < k + α becomes A 1 A 1 ν(e , Dx0 Φ ) < +∞ which is automatically satisfied since ν(e , Dx0 Φ ) is always finite. Hence for the case k = +∞, no assumption is needed on the spectral spread in Theorem1; we need only assume that A 1 (e , Dx0 Φ ) is ∞-nonresonant. Similar remarks hold for all of the following results in this paper which include a condition of the form ν( · , · ) < k + α. Remark 7 (sketch of the proof of the existence portion of Theorem1) . Here we sketch the proof of the existence statement of Theorem1, which is somewhat more involved than the uniqueness proof. (The k,α existence proof also yields uniqueness, but under the additional assumption Φ ∈ Cloc not needed for the n uniqueness statement in Theorem1.) Since the basin of attraction Q of x0 is always diffeomorphic to R n [Wil67, Lem 2.1], we may assume that Q = R and x0 = 0. For now we consider the case k < +∞. First, the k-nonresonance assumption implies that we can uniquely solve (7) order by order (in the sense of Taylor polynomials) for P up to order k. Once we obtain a polynomial P of sufficiently high order, we derive a fixed point equation for the high-order remainder term ϕ, where ψ = P + ϕ is the desired linearizing factor. Given a sufficiently small, positively invariant, closed ball N centered at the fixed point, A 1 the proof of Lemma5 shows that the spectral spread condition ν(e , Dx0 Φ ) < k + α implies that the k,α m k,α m restriction ϕ|N of the desired high-order term is the fixed point of a map S : C (N, C ) → C (N, C ) k,α k,α m which is a contraction, with respect to the standard C norm k · kk,α making C (N, C ) a , k,α m when restricted to the closed linear S-invariant subspace F ⊂ C (N, C ) of functions with vanishing i-th derivatives at the fixed point for all integers 0 ≤ i < k + α.4 In fact, S is the affine map defined by −A 1 (16) S(ϕ|N ) := −P |N + e (P |N + ϕ|N ) ◦ Φ .

Hence we can obtain ϕ|N by the standard contraction mapping theorem, thereby obtaining the function k,α m 1 A ψ|N = ϕ|N + P |N ∈ C (N, C ) satisfying ψ|N ◦ Φ |N = e ψ|N . (The preceding techniques are an extension of Sternberg’s [Ste57] and owe much to Sternberg’s work.) We then extend the domain of ψ|N

4 n Note, however, that k · kk,α must be induced by an appropriate underlying adapted norm [CFdlL03a, Sec. A.1] on R to ensure that S is a contraction. 10 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

k,α m using the globalization techniques of [LM13, EKR18] to obtain a function ψ ∈ Cloc (Q, C ) defined on the 1 A entire basin Q and satisfying ψ ◦Φ = e ψ. To show that the function ψ satisfies (6) when T = R, i.e., that t ψ is actually a linearizing factor of Φ for all t ∈ R, we use an argument of Sternberg [Ste57, Lem. 4] in combination with the uniqueness statement of Theorem1. We extend the result to the case that k = +∞ using a bootstrapping argument. Remark 8 (a numerical consideration). Our proof of the existence portion of Theorem1, outlined above, was inspired by Sternberg’s proof of his linearization theorem [Ste57, Thms 2, 3, 4] and also has strong similarities with the techniques used to prove the existence of semiconjugacies of the type (2) using the parameterization method [CFdlL03a, CFdlL03b, CFdlL05]. We repeat here an observation of [CFdlL03a, Sec. 3] and [CFdlL05, Rem. 5.5] which is also relevant for numerical computations of linearizing semicon- jugacies of the type (1) (such as Koopman eigenfunctions) based on our proof of Theorem1. Consider k,α n m P ∈ Cloc (R , C ) satisfying (7) as in Remark7; N, S, and F as in the same remark; and an initial guess ψ0|N = P |N + ϕ0|N for a local linearizing factor with ϕ0|N ∈ F. If

Lip(S) ≤ κ < 1 kS(ϕ0|N ) − ϕ0|N k ≤ δ where Lip(S) is the Lipschitz constant of S, then the standard proof of the contraction mapping theorem implies the estimate

(17) kϕ|N − ϕ0|N k ≤ δ/(1 − κ), where ϕ|N ∈ F is such that ψ|N = P |N + ϕ|N is the unique actual local linearizing factor. Thus equation (17) furnishes an upper bound on the distance between the initial guess ϕ0|N and the true solution ϕ|N , and can be used for a posteriori estimates in numerical analysis. Theorem1 gave conditions ensuring existence and uniqueness of linearizing factors under spectral spread and nonresonance conditions. Before stating Theorem2, we state a result on the uniqueness of linearizing factors which does not assume any nonresonance conditions. Proposition1 follows immediately from Lemma3 (used to prove the uniqueness statement of Theorem1) and the fact that Q is diffeomorphic to dim(Q) R as mentioned above. 1 Proposition 1. Fix k ∈ N≥1 ∪{+∞} and 0 ≤ α ≤ 1, and let Φ: Q×T → Q be a C dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q is a smooth manifold with dim(Q) ≥ 1 A A and either T = Z or T = R. Let m ∈ N≥1 and e ∈ GL(m, C) have spectral radius ρ(e ) < 1 and satisfy A k,α m i 5 ν(e , D0F ) < k + α. Let ϕ ∈ Cloc (Q, C ) satisfy Dx0 ϕ = 0 for all integers 0 ≤ i < k + α and ϕ ◦ Φ1 = eAϕ.

Then it follows that ϕ ≡ 0. In particular, if ϕ = ψ1 − ψ2, then

ψ1 = ψ2.

k,α Theorem 2 (Existence and uniqueness of Cloc global linearizing factors for a limit cycle attractor). Fix k,α k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1. Let Φ: Q × R → Q be a Cloc flow with Q the basin of an attracting hyperbolic τ-periodic orbit with Γ ⊂ Q, where Q is a smooth manifold with dim(Q) ≥ 2. Fix s τ x0 ∈ Γ and let Ex0 denote the unique Dx0 Φ -invariant subspace complementary to Tx0 Γ. Let m ∈ N≥1 and τA τA s m e ∈ GL(m, C) have spectral radius ρ(e ) < 1, and let the linear map B : Ex0 → C satisfy τ τA (18) BDx Φ |Es = e B. 0 x0 τA τ τA τ Assume that (e , Dx Φ |Es ) is k-nonresonant, and assume that ν(e , Dx Φ |Es ) < k + α. 0 x0 0 x0 k,α m Then there exists a unique ψ ∈ Cloc (Q, C ) satisfying t tA (19) ∀t ∈ : ψ ◦ Φ = e ψ, Dx ψ|Es = B, R 0 x0 s m m m×m m×m m m and if B : Ex0 → R ⊂ C and A ∈ R ⊂ C are real, then ψ : Q → R ⊂ C is real. 5Note the case α = 0 which does not require vanishing of the k-th derivative. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 11

Remark 9. For the uniqueness statement of Theorem1 (and Proposition6) it is only assumed that 1 k,α Φ ∈ C , whereas Φ ∈ Cloc is assumed for the uniqueness statement of Theorem2 (and Proposition7). k,α This is because our proof of the uniqueness statement of Theorem2 relies on Cloc smoothness of the isochrons [Guc75] or (equivalently) strong stable manifolds [Fen74, Fen77, HPS77] of the periodic orbit, a k,α property which is ensured by the assumption that Φ ∈ Cloc . We leave open the question as to whether the need for this additional assumption is merely an artifact of our proof of Theorem2. Remark 10. By considering the Poincaré (first-return) map with Poincaré section an isochron [Guc75] and using Theorem1, it is readily seen that the linearizing factor ψ of Theorem2 can also be represented as a limit analogous to (8) which converges in the pointwise sense. While we believe that this limit also converges in the topology of Ck,α-uniform convergence on compact subsets of Q as in Theorem1, we did not attempt to prove this stronger convergence statement.

3. Applications In this section, we give some applications of Theorems1 and2 and Proposition1. §3.1 contains results on Sternberg linearizations and Floquet normal forms. §3.2 gives applications to principal Koopman eigenfunctions and isostable coordinates. §3.3 contains our classification theorems for C∞ eigenfunctions of C∞ dynamical systems. 3.1. Sternberg linearizations and Floquet normal forms. The following result is an improved state- ment of Sternberg’s linearization theorem for hyperbolic sinks [Ste57, Thms 2,3,4]. Our improvements k,α k include: uniqueness of the linearizing conjugacy, refined Cloc regularity rather than just C , and global definition of the linearization on the entire basin of attraction Q rather than just on some small neighbor- hood of x0. Our techniques for globalizing the domain of the linearization are essentially the same as those used in [LM13, EKR18]. For the case T = R, the k-nonresonance assumption in the following result can be replaced by the slightly less restrictive nonresonance condition in Remark3. k,α Proposition 2 (Existence and uniqueness of global Cloc Sternberg linearizations). Fix k ∈ N≥1 ∪ {+∞} k,α and 0 ≤ α ≤ 1. Let Φ: Q × T → Q be a Cloc dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q is a smooth manifold with dim(Q) ≥ 1 and either T = Z or T = R. Assume 1 1 1 1 that ν(Dx0 Φ , Dx0 Φ ) < k + α, and assume that (Dx0 Φ , Dx0 Φ ) is k-nonresonant. k,α Then there exists a unique diffeomorphism ψ ∈ Cloc (Q, Tx0 Q) satisfying (20) ∀t ∈ : ψ ◦ Φt = D Φtψ, D ψ = id . T x0 x0 Tx0 Q (In writing D ψ = id , we are making the standard and canonical identification T (T Q) =∼ T Q.) x0 Tx0 Q 0 x0 x0 Remark 11 (Uniqueness of general linearizing conjugacies modulo a linear coordinate transformation).

Under the hypotheses of Proposition2, let L1,L2 : Tx0 Q × T → Tx0 Q be any linear dynamical systems t k,α (i.e., such that each time-t map Li is linear) and ψ1, ψ2 ∈ Cloc (Q, Tx0 Q) be any diffeomorphisms satisfying t t ψi ◦ Φ = Liψi for all t ∈ T and i ∈ {1, 2}. Differentiation at the fixed point x0 using the chain rule t t t t −1 yields Dx0 ψiDx0 Φ = LiDx0 ψi for all t ∈ T, so Li = BiDx0 Φ Bi where Bi := Dx0 ψi. It follows that −1 t t −1 −1 Bi ψi ◦ Φ = Dx0 Φ Bi ψi for all t ∈ T, so Bi ψi satisfies (20) for i ∈ {1, 2}. The uniqueness statement −1 of Proposition2 then implies that ψi = Biψ, from which it follows that ψ1 = B1B2 ψ2. n Proof. Identifying Tx0 Q with R by choosing a basis and letting A ∈ GL(n, C) be any matrix logarithm of D Φ1, we apply Theorem1 with etA = D Φt and B = id to obtain a unique ψ ∈ Ck,α(Q, T Q) x0 x0 Tx0 Q loc x0 satisfying (20) and D ψ = id . It remains only to show that ψ is a diffeomorphism. To do this, we x0 Tx0 Q separately show that ψ is injective, surjective, and a local diffeomorphism. By continuity, D ψ = id implies that D ψ is invertible for all x in some neighborhood U 3 x . x0 Tx0 Q x 0 S −t Since Q = t≥0 Φ (U) by asymptotic stability of x0,(20) and the chain rule imply that Dxψ is invertible for all x ∈ Q. Hence ψ is a local diffeomorphism. To see that ψ is injective, let U be a neighborhood of x0 such that ψ|U : U → ψ(U) is a diffeomorphism, and let x, y ∈ Q be such that ψ(x) = ψ(y). By asymptotic stability of x0, there is T > 0 such that 12 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

T T T T Φ (x), Φ (y) ∈ U, and (20) implies that ψ ◦ Φ (x) = ψ ◦ Φ (y). Injectivity of ψ|U then implies that ΦT (x) = ΦT (y), and injectivity of ΦT then implies that x = y. Hence ψ is injective.

To see that ψ is surjective, fix any y ∈ Tx0 Q and let the neighborhood U be as in the last paragraph. T Asymptotic stability of 0 for Dx0 Φ: Tx0 Q × T → Tx0 Q implies that there is T > 0 such that Dx0 Φ · y ∈ T −T −T ψ(U), so there exists x ∈ U with Dx0 Φ · y = ψ(x). Hence y = Dx0 Φ · ψ(x) = ψ ◦ Φ (x), where we have used (20). It follows that ψ is surjective. This completes the proof.  k,α The following is an existence and uniqueness result for the Cloc Floquet normal form of an attracting hyperbolic periodic orbit of a flow, a nonlinear change of coordinates in which the dynamics become the product of a linear system with constant-rate rotation on a circle. References discussing the Floquet normal form include [AR67, Sec. 26], [BK94, Sec. I.3], and [AMR, Sec. 4.3]; the Floquet normal form is a nonlinear generalization of the classical Floquet theory of linear time-periodic systems [Hal, Sec. III.7]. The following result is proved using a combination of Proposition2 and stable manifold theory [Fen74, Fen77, HPS77, Rue89] specialized to the theory of isochrons [Guc75]. For the statement, recall that a map between manifolds is a C1 if it is a homeomorphism onto its image (equipped with the k,α k,α ) and if its derivative is everywhere one-to-one [Hir94, p. 21]; a Cloc embedding is a Cloc map which is also a C1 embedding. A map between topological spaces is proper if the preimage of every compact subset is compact [Lee13, p. 610].

k,α Proposition 3 (Existence and uniqueness of Cloc global Floquet normal forms). Fix k ∈ N≥1 ∪{+∞} and k,α 0 ≤ α ≤ 1. Let Φ: Q × R → Q be a Cloc flow with Q the basin of an attracting hyperbolic τ-periodic orbit s with image Γ ⊂ Q, where Q is a smooth manifold with dim(Q) ≥ 2. Fix x0 ∈ Γ and let Ex0 ⊂ Tx0 Q denote τ τ τ the unique Dx Φ -invariant subspace complementary to Tx Γ. Assume that ν(Dx Φ |Es , Dx Φ |Es ) < 0 0 0 x0 0 x0 τ τ k + α, and assume that (Dx Φ |Es , Dx Φ |Es ) is k-nonresonant. 0 x0 0 x0 τ τA s s Then if we write Dx Φ |Es = e for some complex linear A:(E ) → (E ) , there exists a unique, 0 x0 x0 C x0 C k,α 1 s s proper, C embedding ψ = (ψθ, ψz): Q → S × (E ) such that ψθ(x0) = 1, (Dx ψz)|Es = (E ,→ loc x0 C 0 x0 x0 s (Ex0 )C), and t 2πi t t tA (21) ∀t ∈ R: ψθ ◦ Φ (x) = e τ ψθ(x), ψz ◦ Φ (x) = e ψz(x), 1 √ s s s k,α s where S ⊂ is the unit circle and i = −1. If A|Es : E → E ⊂ (E ) is real, then ψz ∈ C (Q, E ) C x0 x0 x0 x0 C loc x0 1 s 1 s is real, and the codomain-restricted map ψ : Q → S × Ex0 ⊂ S × (Ex0 )C is a diffeomorphism. k,α s Proof. Theorem2 implies that a map ψz ∈ Cloc (Q, (Ex0 )C) satisfying all applicable conclusions above s s s exists. Letting Wx0 denote the global strong stable manifold (isochron) through x0, we have Tx0 Wx0 = Ex0 τ s s s k,α τ and Φ (Wx0 ) = Wx0 . Since Wx0 is the stable manifold for the fixed point x0 of the Cloc diffeomorphism Φ , s k,α it follows that Wx0 is a Cloc [Rue89, pp. 2, 27; Thm 6.1] which is properly embedded (rather than merely immersed) in Q because Γ is stable [EKR18, pp. 4208–4209]. Proposition2 then implies that s s s s 6 s s ψz|W s : W → E ⊂ (E ) is a diffeomorphism onto its image E . Since E is closed in (E ) , it x0 x0 x0 x0 C x0 x0 x0 C s s k,α follows that ψz|W s : W → (E ) is a proper C embedding [Lee13, Prop. A.53(c)]. x0 x0 x0 C loc s s : tA s Define expA :(Ex0 )C × R → (Ex0 )C via expA(z, t) = e z, let K ⊂ (Ex0 )C be any subset, define −1 Kτ := exp (K × [−τ, 0]), and define Jτ := (ψz|W s ) (Kτ ). From the second equation of (21), we have A x0 that −1 [ t  −1 −tA  ψ (K) = Φ (ψz|W s ) (e K) ⊂ Φ(Jτ × [0, τ]). z x0 t∈[0,τ] s If K is compact, then so are Kτ and Jτ by continuity of exp , properness of ψz|W s , and the fact that W A x0 x0 is a closed subset of Q since it is properly embedded [Lee13, Prop. 5.5]. Thus, if K is compact, continuity

6Strictly speaking, Proposition2 was—for simplicity—stated for smooth manifolds. Hence in order to apply Proposition2 s ∞ here (and also in the proofs of Theorems2 and4) we must first give Wx0 a compatible C structure, but this can always be done [Hir94, Thm 2.2.9], so we will not mention this anymore. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 13

−1 implies that ψz (K) is a closed subset of the compact set Φ(Jτ × [0, τ]) and is therefore compact. This s establishes that ψz : Q → (Ex0 )C is proper. s Since the vector field generating Φ intersects Wx0 transversely, a standard argument [HS74, p. 243] and k,α k,α s the Cloc implicit function theorem [Eld13, Cor. A.4] imply that a real-valued Cloc “time-to-impact Wx0 ” function can be defined on a neighborhood of any point. Using these facts, one can show that the function t k,α 1 s t s 2πi τ ψθ : Q → S defined via ψθ(Wx0 ) ≡ 1 and ψθ(Φ (Wx0 )) ≡ e is a Cloc function. By construction, this function ψθ satisfies ψθ(x0) = 1 and (21). ψθ is unique among all continuous functions satisfying these ˜ equalities, since if ψθ is any other such function, then asymptotic stability of Γ implies that the quotient ˜ ˜ ˜ (ψθ/ψθ) is constant on Q, and since (ψθ/ψθ)(x0) = 1 it follows that ψθ ≡ ψθ. s t s Note that, for all t ∈ and x ∈ W t = Φ (W ), R Φ (x0) x0 s s ker (D ψ ) = T W t and ker (D ψ ) ∩ T W t = {0}, x θ x Φ (x0) x z x Φ (x0) tA −t k,α with the second equality following since ψ | s = e ◦ ψ | s ◦ Φ | s is a C embedding. It z W t z Wx W t loc Φ (x0) 0 Φ (x0) : 1 s follows that ψ = (ψθ, ψz): Q → S × (Ex0 )C is an immersion, i.e., that Dxψ is injective for all x ∈ Q. s Furthermore, ψ is injective since the restriction of ψ to any level set W t of ψ is the composition of z Φ (x0) θ tA −t 1 s s injective maps e ◦ ψ | s ◦ Φ | s . Let π : S × (E ) → (E ) be onto the second z Wx W t z x0 C x0 C 0 Φ (x0) −1 −1 factor. Since ψ (K) ⊂ ψz (πz(K)) for any subset K, properness of ψz and continuity of πz and ψ imply that ψ is also proper. Since (i) proper maps between manifolds are closed maps [Lee13, Thm A.57], (ii) closed injective continuous maps are onto their images [Lee13, Lem. A.52.C], and (iii) ψ k,α k,α is a Cloc injective immersion, it follows that ψ is a proper Cloc embedding. 1 s 1 s If A is real, then the image of the proper embedding ψ is contained in S × Ex0 ⊂ S × (Ex0 )C. Since 1 s 1 dim(S × Ex0 ) = dim(Q), and since C proper between manifolds of the same dimension are both open and closed maps [Lee13, Prop. 4.28, Thm A.57], it follows that the image of ψ is both open and 1 s 1 s 1 s closed in S × Ex0 . Since S × Ex0 is connected, it follows that the image of ψ is all of S × Ex0 [Lee13, k,α 1 s Prop. A.39(e)]. Thus, ψ is a Cloc diffeomorphism onto S × Ex0 if A is real. This completes the proof.  3.2. Principal Koopman eigenfunctions, isostables, and isostable coordinates. Given a C1 dy- namical system Φ: Q × T → Q, where Q is a smooth manifold and either T = Z or T = R, we say that ψ : Q → C is a Koopman eigenfunction if ψ is not identically zero and satisfies t µt (22) ∀t ∈ T: ψ ◦ Φ = e ψ for some µ ∈ C. The following are intrinsic definitions of principal eigenfunctions and the principal algebra which extend the definitions for linear systems given in [MM16b, Def. 2.2–2.3]; a more detailed comparison is given later in Remark 16. The condition ψ|Γ ≡ 0 was motivated in part by the definition of a certain space FAc of functions in [MM16a, p. 3358].

Definition 4. If Q is the basin of an asymptotically stable fixed point x0 ∈ Q for Φ, we say that an 1 eigenfunction ψ ∈ C (Q) is a principal eigenfunction if ψ(x0) = 0 and Dx0 ψ 6= 0. If instead Q is the basin of an asymptotically stable periodic orbit with image Γ ⊂ Q for Φ, we say that an eigenfunction 1 7 ψ ∈ C (Q) is a principal eigenfunction if ψ|Γ ≡ 0 and Dx0 ψ 6= 0 for all x0 ∈ Γ. In either case, we define k,α k,α k,α k,α the Cloc principal algebra AΦ to be the complex subalgebra of Cloc (Q, C) generated by all Cloc principal eigenfunctions.

Given a (real or complex) linear self-map Y : V → V , we say that a linear map w : V → C is a left eigenvector of Y with eigenvalue λ ∈ C if wY = λw. (If a basis is chosen for V , then Y can be identified with a matrix and w can be identified with a row vector so that wY = λw in the usual sense of matrix multiplication.) Differentiating (22) and using the chain rule immediately yields Propositions4 and5, which have previously appeared in the literature (see, e.g., the proof of [MM16a, Prop. 2]; our stability assumptions are for convenience of exposition and are not necessary).

7 By (22) and the chain rule, it suffices to assume there exists one point x0 ∈ Γ such that ψ(x0) 6= 0 and Dx0 ψ 6= 0. 14 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

1 Proposition 4. Let Q be the basin of an asymptotically stable fixed point x0 for the C dynamical system Φ: Q × T → Q. If ψ is a principal Koopman eigenfunction for Φ satisfying (22) with exponent µ ∈ C, then t µt for any t ∈ T, it follows that Dx0 ψ is a left eigenvector of Dx0 Φ with eigenvalue e . Proposition 5. Let Q be the basin of an asymptotically stable τ-periodic orbit with image Γ ⊂ Q for the 1 C dynamical system Φ: Q × R → Q. If ψ is a principal Koopman eigenfunction for Φ satisfying (22) with τ exponent µ ∈ C, then for any x0 ∈ Γ, it follows that Dx0 ψ is a left eigenvector of Dx0 Φ with eigenvalue eµτ ; in particular, eµτ is a Floquet multiplier for Γ. Remark 12. For a dynamical system with Q the basin of an attracting compact invariant set M, any µ continuous eigenfunction defined on Q satisfying (22) with exponent µ ∈ C must have |e | ≤ 1. If this attracting set M is furthermore a hyperbolic fixed point, then there is the stronger observations that either eµ = 1 or |eµ| < 1. These observations are straightforward consequences of continuity and (22). Remark 13. Let µ ∈ C and consider the case that Q is the basin of an attracting hyperbolic fixed point 1 1 x0 ∈ Q for the C dynamical system Φ: Q × T → Q. Recall that the spectral radius ρ(Dx0 Φ ) ∈ (0, 1) of 1 Dx0 Φ is defined to be the largest modulus (absolute value) of the eigenvalues of (the complexification of) 1 µ 1 Dx0 Φ . From Equation (4) of Definition2, the spectral spread ν(e , Dx0 Φ ) satisfies r µ 1 n µ   1 o (23) ν(e , Dx0 Φ ) = min r ∈ R: |e | ≥ ρ Dx0 Φ . It follows that, for any r ∈ R ∪ {+∞}, r µ 1 µ   1 (24) ν(e , Dx0 Φ ) < r ⇐⇒ |e | > ρ Dx0 Φ

1+∞ where ρ Dx0 Φ := 0 in the special case that r = +∞. In light of Remarks 12 and 13 (taking r = k + α in (24)), Proposition6 below is now nearly immediate from Theorem1 and Proposition4. We prove the non-immediate portion following the statement of Proposition6. We emphasize that, when k = +∞ in Proposition6, the spectral spread condition k+α +∞ µ   1   1 |e | > ρ Dx0 Φ = ρ Dx0 Φ := 0 µ is automatically satisfied since |e | > 0 for all µ ∈ C (cf. Remark6). For the case T = R, the k-nonresonance assumptions in the following result can be replaced by the slightly less restrictive nonresonance condition in Remark3. Applications-oriented readers may find it helpful to consult Example1 with m = 1 for remarks relevant to the following proposition. k,α Proposition 6 (Existence and uniqueness of Cloc Koopman eigenvalues and principal eigenfunctions for a 1 point attractor). Let Φ: Q × T → Q be a C dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q is a smooth manifold with dim(Q) ≥ 1 and either T = Z or T = R. Fix 1 k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1, and assume that the spectral radius ρ Dx0 Φ ∈ (0, 1) satisfies k+α µ   1 |e | > ρ Dx0 Φ in all of the following statements. k,α Uniqueness of Koopman eigenvalues and principal eigenfunctions. Let ψ1 ∈ Cloc (Q, C) be any Koopman eigenfunction satisfying (22) with exponent µ ∈ C. n (1) Then there exists m = (m1, . . . , mn) ∈ N≥0 such that eµ = em·λ,

λ1 λn 1 where e , . . . , e are the eigenvalues of Dx0 Φ repeated with multiplicities and λ := (λ1, . . . , λn). µ 1 µ 1 (2) Assume that ψ1 is a principal eigenfunction so that e ∈ spec(Dx0 Φ ), and assume that (e , Dx0 Φ ) µ is k-nonresonant. Then ψ1 is uniquely determined by Dx0 ψ1, and if e and Dx0 ψ1 are real, then µ ψ1 : Q → R ⊂ C is real. In particular, if e is an algebraically simple eigenvalue of (the complex- 1 ification of) Dx0 Φ and if ψ2 is any other principal eigenfunction satisfying (22) with the same exponent µ, then there exists c ∈ C \{0} such that

ψ1 = cψ2. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 15

k,α µ 1 Existence of principal eigenfunctions. Assume that Φ ∈ Cloc and that (e , Dx0 Φ ) is k-nonresonant. 1 µ Let w : Tx0 Q → C be any left eigenvector of Dx0 Φ with eigenvalue e such that t µt ∀t ∈ T: wDx0 Φ = e w. k,α (1) Then there exists a unique principal eigenfunction ψ ∈ Cloc (Q, C) satisfying (22) with exponent µ

and satisfying Dx0 ψ = w. k,α (2) In fact, if P ∈ Cloc (Q, C) is any “approximate eigenfunction” satisfying Dx0 P = w and (25) P ◦ Φ1 = eµP + R i with Dx0 R = 0 for all integers 0 ≤ i < k + α, then (26) ψ = lim e−µtP ◦ Φt t→∞ in the topology of Ck,α-uniform convergence on compact subsets of Q if k < +∞, and in the topology k0 0 of C -uniform convergence on compact subsets of Q for any k ∈ N≥1 if k = +∞. Proof. All of the claims are immediate from Theorem1 and Proposition4 except for the first uniqueness claim, which we now justify. Suppose (to obtain a contradiction) that the first uniqueness claim does not µ 1 µ 1 hold. Then (i) e is not an eigenvalue of Dx0 Φ and (ii) (e , Dx0 Φ ) is ∞-nonresonant. The first observation together with Proposition4 implies ψ is not a principal eigenfunction, i.e., Dx0 ψ = 0. From Remark 13 and the uniqueness portion of Theorem1 (with A = µ and B = 0), it follows that ψ ≡ 0 is identically zero. However, Koopman eigenfunctions are (by definition) not identically zero, so we have obtained a contradiction.  Remark 14 (Laplace averages). Given P : Q → C and T = R, in the Koopman literature the Laplace average 1 Z T ψ := lim e−µtP ◦ Φt dt T →∞ T 0 is used to produce a Koopman eigenfunction satisfying (22) with exponent µ as long as the limit exists [MMM13, MM14]. (When T = Z, a similar definition can be given with a sum replacing the integral.) Since convergence of the limit (26) easily implies convergence of the Laplace average to the same limiting function, the existence portion of Proposition6 gives sufficient conditions under which the Laplace average k,α of P exists and is equal to a unique Cloc principal eigenfunction ψ satisfying Dx0 ψ = Dx0 P . Remark 15 (Isostables and isostable coordinates). It follows from the discussion after [MMM13, Def. 2] that the definition of isostables given in that paper—for Φ having an attracting hyperbolic fixed point x0 1 µ1 with basin of attraction Q and with Dx0 Φ having a unique eigenvalue e (or complex conjugate pair of eigenvalues) of largest modulus—is equivalent to the following. Isostables as defined in [MMM13] are the level sets of the modulus |ψ1| of a principal eigenfunction ψ1 defined on Q and satisfying (22) with exponent µ1 1 µ = µ1. Because e is the “slowest” eigenvalue of Dx0 Φ , Proposition6 implies that, for any α > 0, any such 1,α 1 ψ1 ∈ Cloc (Q, C) satisfying (22) with exponent µ1 is unique modulo scalar multiplication for a C dynamical µ1 µ1 1+α 11+α system Φ without any further assumptions (since |e | > |e | = ρ Dx0 Φ ). Furthermore, such 1,α t µt a unique eigenfunction always exists if Φ ∈ Cloc and if wDx0 Φ = e w for all t ∈ T, where w is a left 1 µ 8 ¯ eigenvector of Dx0 Φ with eigenvalue e . Since the complex conjugate ψ1 is a principal eigenfunction satisfying (22) with exponent µ =µ ¯1, it follows that the isostables as defined in [MMM13] are unique even 1 1 if µ1 ∈ C\R. A uniqueness proof for analytic isostables under the additional assumptions of (Dx0 Φ , Dx0 Φ ) ∞-nonresonance and of dynamics generated by an analytic vector field was given in [MMM13, App. A]. For the special case that the eigenvalue of largest modulus is real, unique, and algebraically simple, in [MMM13, p. 23] these authors do point out that uniqueness of C1 isostables (if they exist) follows from the fact that they coincide with the unique C1 global (strong) stable manifolds [EKR18, pp. 4208, 4211]

8 1,α By Example1, if T = R and Φ is the flow of a Cloc vector field f, these assumptions are automatically satisfied if w is a left eigenvector of Dx0 f with eigenvalue µ. 16 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS over9 an attracting, normally hyperbolic [EKR18, p. 4207], 1-dimensional, inflowing invariant manifold [EKR18, p. 4211]; this argument works even if the dynamical system is only C1 (see [EKR18] for detailed information on the global stable foliation of an inflowing invariant manifold). The 1-dimensional invariant “slow” manifold is itself generally non-unique without further assumptions, but this does not affect the isostable uniqueness argument. However, as pointed out in [MMM13, p. 23], this argument does not work when the eigenvalue of largest modulus is not real, because in this case the isostables can no longer be interpreted as strong stable manifolds (e.g., the relevant slow manifold is now 2-dimensional, so the dimension of the codimension-1 isostables is too large by 1). For the case that T = R and Φ has an attracting hyperbolic periodic orbit, several authors have in- vestigated various versions of isostable coordinates without restricting attention to the “slowest” isostable coordinate. The authors in [WM16a, Eq. 5] defined a “finite-time” approximate version of isostable co- ordinates which provide an approximation of our principal eigenfunctions. Subsequently, [SKN17, Sec. 2] defined a version of “exact” isostable coordinates (termed amplitudes and phases) directly in terms of Koopman eigenfunctions, and in particular our Proposition7 and Theorem4 can be used to directly infer existence, uniqueness and regularity properties of these coordinates under relatively weak assumptions. It appears that [WE18, MWMM19] intended to define a different version of “exact” isostable coordinates close in spirit to the approximate version in [WM16a]. However, these definitions [WE18, MWMM19, Eq. 24, Eq. 58] are given in terms of a limit which might not exist for principal eigenfunctions other than the “slowest”, as we show in Example3 below. In any case, it appears that principal Koopman eigenfunc- tions provide a means for defining all of the isostable coordinates for a periodic orbit attractor which does not require such limits. Remark 16 (Relationship to the principal eigenfunctions, principal algebras, and pullback algebras of [MM16b]). Given a nonlinear dynamical system Φ: Q×T → Q with Q the basin of an attracting hyperbolic fixed point x0, Mohr and Mezić defined (in our notation) the principal eigenfunctions for the associated linearization Dx0 Φ: Tx0 Q × T → Tx0 Q to be those of the form v 7→ w(v), where w : Tx0 Q → C is a left 1 eigenvector of Dx Φ [MM16b, Def. 2.2], and they defined the principal algebra A 1 to be the subalgebra 0 Dx0 Φ 0 of C (Tx0 Q, C) generated by the principal eigenfunctions [MM16b, Def. 2.3]. Mohr and Mezić do not define principal eigenfunctions or the principal algebra for the nonlinear system itself but, given a topological conjugacy τ : Tx0 Q → Q between Φ and Dx0 Φ, they define the pullback algebra   −1 −1 (27) A 1 ◦ τ := {ϕ ◦ τ : ϕ ∈ A 1 }. Dx0 Φ Dx0 Φ k,α Assuming that Φ ∈ Cloc , Proposition6 implies that the relationship between the concepts in our Definition 1 1 1 1 4 and those of [MM16b] is as follows. If (Dx0 Φ , Dx0 Φ ) is k-nonresonant and ν(Dx0 Φ , Dx0 Φ ) < k +α (see 1 Definition2), our principal eigenfunctions for Dx0 Φ coincide precisely with their principal eigenfunctions k,α 1 w : Tx0 Q → C. This implies that our principal algebra A 1 coincides with their ADx Φ . Next, notice Dx0 Φ 0 −1 that the pullback algebra (27) is generated by the functions w ◦ τ where w : Tx0 Q → C is a principal k,α eigenfunction of the linearization. If we further assume that the conjugacy τ is a Cloc diffeomorphism, then −1 k,α   the chain rule implies that each w ◦ τ is a C principal eigenfunction for Φ, and therefore A 1 ◦ loc Dx0 Φ −1 k,α   −1 τ = A by Proposition6. In particular, under the above hypotheses it follows that A 1 ◦ τ Φ Dx0 Φ k,α is independent of τ and generated by at most n Cloc principal eigenfunctions for Φ. This is perhaps k,α surprising since (27) depends on the a priori non-unique conjugacy τ; here the assumption that τ is a Cloc diffeomorphism is essential. k,α For an attracting hyperbolic τ-periodic orbit of a Cloc flow with image Γ and basin Q ⊃ Γ, stable manifold theory [Fen74, Fen77, HPS77, Guc75, Rue89] can be used to show that the global strong stable s k,α τ s s manifold (isochron) Wx0 through x0 ∈ Γ is a Cloc submanifold of Q, and Φ (Wx0 ) = Wx0 . Furthermore,

9In general, the (strong) stable manifolds are the leaves of the unique global (strong) stable foliation [EKR18, p. 4208] of the global (center-)stable manifold [EKR18, p. 4208] of an (inflowing) normally hyperbolic invariant manifold [Fen71, HPS77, Eld13]; these leaves generalize the isochrons [Guc75] of an attracting hyperbolic limit cycle. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 17

k,α s τ τ µτ any eigenfunction ϕ ∈ C (W , ) of Φ |W s satisfying ϕ ◦ Φ |W s = e ϕ admits the unique extension loc x0 C x0 x0 k,α to an eigenfunction ψ ∈ Cloc (Q, C) with exponent µ given by µt −t ψ| s := e ϕ ◦ Φ | s W t W t Φ (x0) Φ (x0) k,α k,α s for all t ∈ R. That ψ ∈ Cloc follows from considering locally-defined Cloc “time-to-impact Wx0 ” functions as in the proof of Proposition3. This observation combined with Propositions5 and6 yields the following result. (Alternatively, the statement concerning existence and uniqueness of principal eigenfunctions follows from Theorem2 and Proposition5.) k,α Proposition 7 (Existence and uniqueness of Cloc Koopman eigenvalues and principal eigenfunctions for k,α a limit cycle attractor). Fix k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1. Let Φ: Q × R → Q be a Cloc flow with Q the basin of an attracting hyperbolic τ-periodic orbit with image Γ ⊂ Q, where Q is a smooth manifold with s s τ dim(Q) ≥ 2. Fix x0 ∈ Γ and let Ex0 = Tx0 Wx0 denote the unique Dx0 Φ -invariant subspace complementary  τ  to Tx Γ. Assume that the spectral radius ρ Dx Φ |Es ∈ (0, 1) satisfies 0 0 x0 k+α µτ   τ  |e | > ρ Dx Φ |Es 0 x0 in all of the following statements. k,α Uniqueness of Koopman eigenvalues. Let ψ1 ∈ Cloc (Q, C) be any Koopman eigenfunction satisfying n (22) with exponent µ ∈ C and T = R. Then there exists m = (m1, . . . , mn) ∈ N≥0 such that eµτ = e(m·λ)τ ,

λ τ λnτ τ where e 1 , . . . , e are the eigenvalues of Dx Φ |Es repeated with multiplicities and λ := (λ1, . . . , λn). 0 x0 µτ τ Existence and uniqueness of principal eigenfunctions. Assume that (e , Dx Φ |Es ) is k-nonresonant. 0 x0 s τ µτ Let w : E → be any left eigenvector of Dx Φ |Es with eigenvalue e . Then there exists a unique x0 C 0 x0 k,α principal eigenfunction ψ ∈ Cloc (Q, C) for Φ satisfying (22) with exponent µ and T = R and satisfying Dx ψ|Es = w. Additionally, if µ and w are real, then ψ : Q → ⊂ is real. 0 x0 R C A well-known example of Sternberg shows that, even for an analytic diffeomorphism Φ1 of the plane having the globally attracting fixed point 0, there need not exist a C2 principal eigenfunction corresponding µ 1 µ 1 to e ∈ spec(D0Φ ) if (e , D0Φ ) is not 2-nonresonant [Ste57, p. 812]. Concentrating now on the issue of uniqueness of principal eigenfunctions, the following example shows that our nonresonance and spectral spread conditions are both necessary for the uniqueness statements of Propositions6 and7 (hence also for the uniqueness statements of Theorems1 and2). 2 2 Example 2 (Uniqueness of principal eigenfunctions). Consider Φ = (Φ1, Φ2): R × T → R defined by t −t Φ1(x, y) = e x (28) t −(k+α)t Φ2(x, y) = e y where k ∈ N≥1, 0 ≤ α ≤ 1, and either T = Z or T = R. Φ is a diagonal linear dynamical system with 1 1 −1 −(k+α) x0 = 0 a globally exponentially stable fixed point, and the eigenvalues of D0Φ = Φ are e and e . −(k+α) 1 Furthermore, for any irrational α ∈ [0, 1], (e , D0Φ ) is ∞-nonresonant; ∞-nonresonance also holds if, e.g., k = 1 and α = 0 (see Definition1). However, if we define σα(x) := |x| for α > 0 and σα(x) := x for α = 0, then for any k ∈ N≥1 and α ∈ [0, 1] both

(29) h1(x, y) := y and k+α (30) h2(x, y) := y + σα(x) are Ck,α principal eigenfunctions satisfying (22) with the same exponent µ = −(k + α). 18 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

k,α In particular, this shows that Cloc principal Koopman eigenfunctions are not necessarily unique (modulo k,α 2 scalar multiplication) even if the ∞-nonresonance condition is satisfied. Since here h2 ∈ Cloc (R , C) −(k+α) 1r −1 r and r = (k + α) is the smallest r ∈ R satisfying e ≥ ρ D0Φ = (e ) , this shows that the µ 1k+α spectral spread condition |e | > ρ Dx0 Φ is both necessary and sharp for the principal eigenfunction uniqueness statement of Proposition6 to hold, at least in the case that α > 0.10 (Note that Proposition6 k0,α0 0 0 does imply that Cloc principal eigenfunctions are unique for any k + α > k + α.) If instead k = 1 and α = 0, then h1 and h2 are both analytic eigenfunctions satisfying (22) with the same exponent µ = −1, −1 1 while (e , D0Φ ) is ∞-nonresonant, but now these eigenfunctions are distinguished by their derivatives at the origin; this is consistent with the uniqueness statement of Proposition6. On the other hand, if k = 2 −2 1 and α = 0 so that (e , D0Φ ) is not 2-nonresonant, (29) and (30) show that analytic eigenfunctions are −2 1+∞ not unique despite the fact that the spectral spread condition |e | > ρ D0Φ = 0 certainly holds. Hence the nonresonance condition is also necessary for the principal eigenfunction uniqueness statement 2 2 1 of Proposition6 to hold. Finally, by taking T = R, changing the state space R above to R × S , and 1 t it √ prescribing the unit circle S ⊂ C with the decoupled dynamics Φ3(x, y, θ) := e θ (where i = −1) yields an example showing that the spectral spread and nonresonance conditions are both necessary for the uniqueness statement in Proposition7 to hold as well. Example 3 (Existence of the limit (26) and isostable coordinates). Existence of the limit in (26) is not automatic if the “approximate eigenfunction” P is not an approximation to sufficiently high order. To demonstrate this, fix k ∈ N≥1, α ∈ [0, 1], r ∈ R≥1, and  ∈ R>0. Define σα(x) := |x| for α > 0 and k,α 2 2 σα(x) := x for α = 0, and consider the Cloc dynamical system Φ: R × T → R defined by t −t Φ1(x, y) = e x (31) t −rt k+α −(k+α)t k+α Φ2(x, y) = e (y − σα(x) ) + e σα(x) , where either T = Z or T = R. To see that Φ is indeed a dynamical system (i.e., that Φ satisfies the group property Φt+s = Φt ◦ Φs), define the diagonal linear system Φe t(x, y) = (e−tx, e−rty) and the Ck,α 2 2 k+α t t −1 diffeomorphism H : R → R via H(x, y) := (x, y + σα(x) ), and note that Φ = H ◦ Φe ◦ H . In other words, Φ is obtained from a diagonal linear dynamical system via a Ck,α change of coordinates; note also that this change of coordinates can be made arbitrarily close to the identity by taking  arbitrarily small. Since x0 = 0 is a globally exponentially stable fixed point for Φe, it is also so for Φ. We note that r0 = r ≥ 1 1 −1 −r 1r0 11 is the smallest r0 ∈ R such that the spectral radius ρ D0Φ = e ∈ (0, 1) satisfies |e | ≥ ρ D0Φ . We further note that, for any choice of , the analytic function P (x, y) := y satisfies P ◦ Φ1 = e−rP + R j where D(0,0)R = 0 for all integers 0 ≤ j < k + α. However, rt t k+α k+α (r−(k+α))t lim e P ◦ Φ (x, y) = y − σα(x) + σα(x) lim e t→∞ t→∞  k+α y − σα(x) 1 ≤ r < k + α (32)  = y r = k + α  +∞ r > k + α

−r 1k+α for any x 6= 0 and  > 0. We see that the limit (32) diverges when r > k+α (so that |e | < ρ D0Φ ), −r 1k+α but the limit converges when r ≤ k + α (so that |e | ≥ ρ D0Φ ). For the case that r < k + α and k,α r 6∈ N≥2, this is consistent with Proposition6 which guarantees that the limit converges if Φ ∈ Cloc , if −r 1k+α −r 1 |e | > ρ D0Φ , and if (e , D0Φ ) is k-nonresonant. When r = k + α and r 6∈ N≥2, convergence is

10 µ 1k+α In the terminology and notation of Theorem1 with m = 1, the condition |e | > ρ D0Φ is equivalent to the µ 1 condition ν(e , D0Φ ) < k + α, where the spectral spread ν( · , · ) is defined in Definition2. See Remark 13. 11 −r 1 That is, the spectral spread satisfies ν(e , D0Φ ) = r in the terminology of Definition2 and Theorem1. See Remark 13. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 19

−r 1 also guaranteed by Proposition6 for this specific example, because then (i) (e , D0Φ ) is ∞-nonresonant, (ii) Φ is linear (the nonlinear terms cancel when r = k + α) and hence C∞, and (iii) Proposition6 ∞ −r 1 guarantees that this limit always exists if Φ ∈ C and (e , D0Φ ) is ∞-nonresonant because the spectral −r 1+∞ spread condition |e | > ρ D0Φ = 0 always holds. As alluded to in Remark5, the preceding reasoning can actually be applied even without the assumption that r is not an integer if Lemma5 is used instead of Proposition6 as the tool of inference (i.e., nothing about nonresonance actually needs to be assumed for this example). We emphasize that the divergence in (32) is associated purely with the spectral −r 1 spread condition since, e.g., we can choose r ≥ 1 so that (e , D0Φ ) is ∞-nonresonant and take α = 0 so that Φ is analytic. 2 2 1 1 Note that by taking T = R, changing the state space R to R ×S , and prescribing the unit circle S ⊂ C t it √ with the decoupled dynamics Φ3(x, y, θ) := e θ (where i = −1) yields a corresponding example with a globally attracting hyperbolic periodic orbit {(0, 0)}×S1. In this case, for this example [WE18, MWMM19, Eq. 24, Eq. 58] would attempt to define the “faster” isostable coordinate (principal eigenfunction in our terminology) ψ2 satisfying (22) with exponent µ2 := −r via the limit (32), but (32) shows that this limit does not exist if r > k + α. This phenomenon should be compared with the explanation in the preceding paragraph based on our general results. 3.3. Classification of all C∞ Koopman eigenfunctions. Notation. To improve the readability of Theorems3 and4 below, we introduce the following multi-index n notation. We define an n-dimensional multi-index to be an n-tuple i = (i1, . . . , in) ∈ N≥0 of nonnegative n n integers, and define its sum to be |i| := i1 +···+in. For a multi-index i ∈ N≥0 and z = (z1, . . . , zn) ∈ C , we [i] i1 in n n [i] define z := z1 ··· zn . Given a C -valued function ψ = (ψ1, . . . , ψn): Q → C , we define ψ : Q → C via [i] [i] ψ (x) := (ψ(x)) for all x ∈ Q. We also define the complex conjugate of ψ = (ψ1, . . . , ψn) element-wise: ψ¯ := (ψ¯1,..., ψ¯n). For the case T = R, the ∞-nonresonance assumption in the following result can be replaced by the slightly less restrictive nonresonance condition in Remark3. Applications-oriented readers may find it helpful to consult Example1 with m = 1 for relevant remarks. ∞ ∞ Theorem 3 (Classification of all C eigenfunctions for a point attractor). Let Φ: Q × T → Q be a C dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q is a smooth 1 manifold with dim(Q) ≥ 1 and either T = Z or T = R. Assume that Dx0 Φ is semisimple (diagonalizable 1 1 over C) and that (Dx0 Φ , Dx0 Φ ) is ∞-nonresonant. Letting n = dim(Q), it follows that there exists an n-tuple

ψ = (ψ1, . . . , ψn) of C∞ principal eigenfunctions such that every C∞ Koopman eigenfunction ϕ is a finite sum of scalar multiples of products of the ψj and their complex conjugates ψ¯j: X [`] ¯[m] (33) ϕ = c`,mψ ψ |`|+|m|≤k for some k ∈ N≥0 and some coefficients c`,m ∈ C. ∞ n Proof. By Proposition2 and linear algebra, there exists a C embedding Q,→ C which maps Q diffeo- n morphically onto an R-linear subspace of C , maps x0 to 0, and semiconjugates Φ to the diagonal C-linear t λ t λnt 12 dynamical system Θ (z1, . . . , zn) = (e 1 z1, . . . , e zn). Thus, for simplicity, we may (and do) view Q n ∞ ∞ as a Θ-invariant R-linear subspace of C with Φ = Θ|Q×T. Let ϕ ∈ C (Q, C) be any C Koopman eigenfunction satisfying (22) with exponent µ ∈ C. n ∞ Write z = (z1, . . . , zn) ∈ C . For any k ∈ N≥0, Taylor’s theorem implies the existence of Rk ∈ C (Q, C) k satisfying 0 = Rk(0) = D0Rk = ··· = D0Rk and coefficients c`,m such that X [`] [m] (34) ϕ(z) = c`,mz z¯ + Rk(z) |`|+|m|≤k

12The standard definition of C∞ embedding is recalled preceding Proposition3 (take α = 0 and k = +∞). 20 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

1 µ for all z ∈ Q. Defining λ := (λ1, . . . , λn) and writing the eigenfunction equation ϕ ◦ Θ = e ϕ in terms of the expansion (34) yields, for all z ∈ Q, X `·λ+m·λ¯ [`] [m] 1 X µ  [`] [m]  e c`,mz z¯ + Rk ◦ Φ (z) = e c`,mz z¯ + Rk(z) . |`|+|m|≤k |`|+|m|≤k 1 µ Upon subtracting terms, we see that Sk := Rk ◦ Φ − e Rk is the restriction of a k-th order polynomial to k Q. On the other hand, we infer that 0 = Sk(0) = D0Sk = ··· = D0Sk from the corresponding property of 1 µ Rk. Thus, Sk is the zero function, so Rk ◦ Φ = e Rk. Recall Definition2 of the (always finite) spectral µ 1 spread ν( · , · ). If we choose k sufficiently large so that k > ν(e , D0Φ ), Proposition1 implies that Rk ≡ 0. From (34), we then see that ϕ is equal to a sum of products of the principal eigenfunctions ψj(z) := zj, ¯ ψj(z) =z ¯j as desired.  k,α s For a an attracting hyperbolic τ-periodic orbit of a Cloc flow with image Γ and basin Q ⊃ Γ, let Wx0 be the global strong stable manifold (isochron) through the point x0 ∈ Γ. As discussed in the proof of Proposition3, there is a unique (modulo scalar multiplication) continuous eigenfunction satisfying (22) 2π √ ∞ ∞ with exponent µ = i τ and T = R, where i = −1, and this eigenfunction is in fact C for a C flow. s Moreover, such an eigenfunction is constant on Wx0 . In the theorem below, let ψθ be the unique such eigenfunction satisfying ψθ|W s ≡ 1, where the point x0 is as in the theorem statement. Explicitly, ψθ is x0 given by i 2π t ψ | s = e τ θ W t Φ (x0) S s for all t ∈ . This defines ψ on all of Q since Q = W t , and the definition makes sense since R θ t∈R Φ (x0) s s W jτ = W for all j ∈ . Φ (x0) x0 Z ∞ Theorem 4 (Classification of all C eigenfunctions for a limit cycle attractor). Let Φ: Q × R → Q be a C∞ dynamical system with Q the basin of an attracting hyperbolic τ-periodic orbit with image Γ ⊂ Q, s where Q is a smooth manifold with dim(Q) ≥ 2. Fix x0 ∈ Γ and denote by Ex0 the unique τ-invariant τ τ τ subspace complementary to Tx Γ. Assume that Dx Φ |Es is semisimple and that (Dx Φ |Es , Dx Φ |Es ) 0 0 x0 0 x0 0 x0 is ∞-nonresonant. Letting n + 1 = dim(Q), it follows that there exists an n-tuple

ψ = (ψ1, . . . , ψn) of C∞ principal eigenfunctions such that every C∞ Koopman eigenfunction ϕ is a finite sum of scalar ¯ multiples of products of integer powers of ψθ with products of the ψj and their complex conjugates ψj: X [`] ¯[m] j`,m (35) ϕ = c`,mψ ψ ψθ |`|+|m|≤k for some k ∈ N≥0, some coefficients c`,m ∈ C, and j`,m ∈ Z. s ∞ Proof. Let Wx0 be the C global strong stable manifold through x0. We remind the reader of the facts S s s t s Q = W t and W t = Φ (W ) which are implicitly used in the remainder of the proof. t∈R Φ (x0) Φ (x0) x0 ∞ s j jτ First, we note that every eigenfunction χ ∈ C (W , ) of F (x) := Φ |W s (x) satisfying (22) with x0 C x0 ∞ exponent µ ∈ C and T = Z admits a unique extension to an eigenfunction χ˜ ∈ C (Q, C) of Φ satisfying (22) with exponent µ and T = R; this unique extension χ˜ is defined via −µt t (36) χ˜| s = e χ ◦ Φ | s W −t W −t Φ (x0) Φ (x0) ∞ ∞ s for all t ∈ R. That χ˜ ∈ C follows from considering locally-defined C “time-to-impact Wx0 ” functions as in the proof of Proposition3, and χ is a principal eigenfunction if and only if its extension χ˜ is. ∞ Next, let ϕ ∈ C (Q, C) be a eigenfunction satisfying (22) with exponent µ and T = R. Theorem3 τ implies that ϕ|W s is equal to a sum of products of principal eigenfunctions χ1, . . . , χn, χ¯1,..., χ¯n of Φ |W s x0 x0 of the form: X [`] [m] (37) ϕ|W s = c`,mχ χ¯ x0 |`|+|m|≤k EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 21

n for some k ∈ N≥0, where χ = (χ1, . . . , χn). Let λ = (λ1, . . . λn) ∈ C be such that each χj satisfies τ λ τ χj ◦ Φ |W s = e j χj. Since ϕ satisfies (22) with exponent µ, it follows that x0 ¯ (38) eµτ = e(`·λ+m·λ)τ n for all `, m ∈ N≥0 such that c`,m 6= 0, so for such `, m we have 2π (39) µ = ` · λ + m · λ¯ + i j τ `,m √ for some j`,m ∈ , where i = −1. By the previous paragraph, we may uniquely write χ = ψ|W s = Z x0 (ψ1|W s , . . . , ψn|W s ) for principal eigenfunctions ψj of Φ satisfying (22) with exponent λj and = . x0 x0 T R Using (37),(39), and the extension formula (36), we obtain X −µt  [`] [m] t ϕ| s = c e · χ χ¯ ◦ Φ | s W −t `,m W −t Φ (x0) Φ (x0) |`|+|m|≤k X −µt  [`] [m] t = c e · ψ ψ¯ | s ◦ Φ | s `,m Wx W −t 0 Φ (x0) |`|+|m|≤k X −(i 2π j )t  [`] [m] = c e τ `,m · ψ ψ¯ | s `,m W −t Φ (x0) |`|+|m|≤k

X  [`] [m] j`,m  = c ψ ψ¯ ψ | s `,m θ W −t Φ (x0) |`|+|m|≤k for all t ∈ as desired. To obtain the last equality we used the fact that ψθ|W s ≡ 1, so the extension for- R x0 −i 2π t  j`,m  −(i 2π j )t mula (36) implies that ψ | s ≡ e τ and hence also ψ | s ≡ e τ `,m . This completes θ W −t θ W −t Φ (x0) Φ (x0) the proof. 

4. Proofs of the main results 4.1. Proof of Theorem1. In this section we prove Theorem1, which we repeat here for convenience. k,α Theorem 1 (Existence and uniqueness of Cloc global linearizing factors for a point attractor). Let Φ: Q× 1 T → Q be a C dynamical system with Q the basin of an attracting hyperbolic fixed point x0 ∈ Q, where Q A is a smooth manifold with dim(Q) ≥ 1 and either T = Z or T = R. Let m ∈ N≥1 and e ∈ GL(m, C) have A m spectral radius ρ(e ) < 1, and let the linear map B : Tx0 Q → C satisfy t tA (5) ∀t ∈ T: BDx0 Φ = e B. A 1 Fix k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1, assume that (e , Dx0 Φ ) is k-nonresonant, and assume that A 1 ν(e , Dx0 Φ ) < k + α. k,α m Uniqueness. Any ψ ∈ Cloc (Q, C ) satisfying 1 A ψ ◦ Φ = e ψ, Dx0 ψ = B m m A m m is unique, and if B : Tx0 Q → R ⊂ C and e ∈ GL(m, R) ⊂ GL(m, C) are real, then ψ : Q → R ⊂ C is real. k,α k,α m Existence. If furthermore Φ ∈ Cloc , then such a unique ψ ∈ Cloc (Q, C ) exists and additionally satisfies t tA (6) ∀t ∈ T: ψ ◦ Φ = e ψ. k,α m In fact, if P ∈ Cloc (Q, C ) is any “approximate linearizing factor” satisfying Dx0 P = B and (7) P ◦ Φ1 = eAP + R i with Dx0 R = 0 for all integers 0 ≤ i < k + α, then (8) ψ = lim e−tAP ◦ Φt t→∞ 22 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS in the topology of Ck,α-uniform convergence on compact subsets of Q if k < +∞, and in the topology of k0 0 C -uniform convergence on compact subsets of Q for any k ∈ N≥1 if k = +∞. We prove the uniqueness and existence portions of Theorem1 in the following §4.1.1 and §4.1.2, respec- tively. Some of our statements and proofs make use of higher-order derivatives of maps between Euclidean spaces [BK94, Sec. A.5] and the fact that a multilinear map is equivalent to a linear map out of a tensor product (the “universal property of the tensor product”). 4.1.1. Proof of uniqueness. In this section, we prove the uniqueness portion of Theorem1. The proof of uniqueness consists of an algebraic part and an analytic part. The algebraic portion is carried out in Lemmas1 and2, and the analytic portion is carried out in Lemma3. m×m n×n Lemma 1. Let k ∈ N≥1 ∪ {+∞}, m, n ∈ N≥1, X ∈ C , and Y ∈ R be such that (X,Y ) is k- n ⊗i m nonresonant. For all 1 < i < k + 1, let L((R ) , C ) denote the space of linear maps from the i-fold n ⊗i m tensor product (R ) to C , and define the linear operator n ⊗i m n ⊗i m ⊗i (40) Ti : L((R ) , C ) → L((R ) , C ),Ti(P ) := PY − XP. n ⊗i ⊗i (By this formula we mean that Ti(P ) acts on tensors τ ∈ (R ) via τ 7→ P (Y (τ)) − XP (τ).) Then for all 1 < i < k + 1, Ti is a linear isomorphism. (The conclusion holds vacuously if k = 1.)

Proof. Let λ1, . . . , λn and µ1, . . . , µm respectively be the eigenvalues of Y and X repeated with multiplicity. First assume that Y and X are both semisimple, i.e., diagonalizable over C. Identifying Y with its n 1 n n ∗ complexification, let e1, . . . , en ∈ C be a basis of eigenvectors for Y and let e , . . . , e ∈ (C ) be the m associated dual basis. Let f1, . . . , fm ∈ C be a basis of eigenvectors for X. Fix any integer i with i ⊗[`] `1 `i 1 < i < k + 1, any p ∈ {1, . . . , m}, and any multi-indices `, j ∈ N≥1; defining e := e ⊗ · · · ⊗ e and similarly for e⊗[j], we compute  ⊗[`] ⊗[`] ⊗[`] Ti fp ⊗ e · e⊗[j] = λj1 ··· λjn · (e · e⊗[j])fp − µp · (e · e⊗[j])fp ` = δj · (λj1 ··· λji − µp) fp ` ` (no summation implied), where the multi-index Kronecker delta is defined by δ` = 1 and δj = 0 if ` 6= j. ⊗[`] Hence the fp ⊗e are eigenvectors of Ti with eigenvalues (λ`1 ··· λ`i − µp), and dimension counting implies that these are all of the eigenvector/eigenvalue pairs. The k-nonresonance assumption implies that none of these eigenvalues are zero, so Ti is invertible if Y and X are both semisimple. Since the operator Ti depends continuously on the matrices X and Y , since eigenvalues of a matrix depend continuously on the matrix, and since semisimple matrices are dense, it follows by continuity that the eigenvalues of Ti are all of the form (λ`1 ··· λ`i − µp) even if one or both of X and Y are not semisimple (cf. [Nel70, p. 37]), and these eigenvalues are all nonzero by the asssumption that (X,Y ) is k-nonresonant. Hence Ti is still invertible in the case of general X and Y .  1 n n Lemma 2. Let F ∈ C (R , R ) have the origin as a fixed point, where n ≥ 1. Let k ∈ N≥1 ∪ {+∞}, m×m k n m m ∈ N≥1, and X ∈ C be such that (X, D0F ) is k-nonresonant. Assume that ψ ∈ C (R , C ) satisfies D0ψ = 0 and (41) ψ ◦ F = Xψ. Then it follows that ψ(0) = Xψ(0) and i D0ψ = 0 for all 1 < i < k + 1. (The conclusion holds vacuously if k = 1.) Remark 17. We can restate the conclusion of Lemma2 in the language of jets [Hir94, GS85, BK94]. If ψ 1 k is a linearizing factor such that the 1-jet j0 (ψ − ψ(0)) = 0, then automatically the k-jet j0 (ψ − ψ(0)) = 0. Proof. That ψ(0) = Xψ(0) follows from setting x = 0 in (41) and using the assumption that F (0) = 0. i We will prove the remaining claim that D0ψ = 0 for 1 ≤ i < k + 1 by induction on i. The base case of the 1 induction, D0ψ = D0ψ = 0, is one of the hypotheses of the lemma. For the inductive step, assume that i i+1 D0ψ = ··· = D0ψ = 0 for an integer i satisfying 1 ≤ i < k. If it were the case that F ∈ C , one way to EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 23

i+1 proceed would be to differentiate (41) (i + 1) times and somehow deduce that D0 ψ = 0. However, we are assuming only that F ∈ C1, so that approach is problematic. We instead proceed as follows. By Taylor’s theorem and the inductive hypothesis, we have (here x⊗` denotes the tensor product of x with itself ` times) i+1 ⊗(i+1) (42) F (x) = D0F · x + RF (x), ψ(x) = ψ(0) + D0ψ · x + Rψ(x)

RF (x) Rψ(x) for all x, where the remainders satisfy limx→0 kxk = 0 and limx→0 kxki+1 = 0. It follows that i+1 ⊗(i+1) (43) ψ(F (x)) = ψ(0) + D0ψ · (D0F · x) + R(x), where i i+1 X ⊗` ⊗(i+1−`) (44) R(x) := Rψ(F (x)) + D0ψ · C`[(D0F · x) ⊗ (RF (x)) ] `=0 for suitable combinatorially determined constants C` > 0. Rewriting (41) using (42) and (43), we obtain i+1 ⊗(i+1) i+1 ⊗(i+1) (45) ψ(0) + D0ψ · (D0F · x) + R(x) = Xψ(0) + XD0ψ · x + XRψ(x). R(x) In order to deduce the information we need from (45), we now show that limx→0 kxki+1 = 0. Using the tensor product property ap+qv⊗p ⊗ w⊗q = (av)⊗p ⊗ (aw)⊗q for a scalar a and vectors v and w, this follows from (44) and the computation

i+1 i+1 " ⊗(i+1−`)  ⊗`# R(x) Rψ(F (x)) kF (x)k i+1 X D0F · x RF (x) lim = lim + D0ψ · lim C` ⊗ x→0 kxki+1 x→0 kF (x)ki+1 kxki+1 x→0 kxk kxk (46) `=1 = 0. The first limit on the right is zero since F (x) → 0 as x → 0 and since F ∈ C1, so kF (x)k/kxk ≤ maxkyk≤1kDyF k < +∞ when kxk ≤ 1 by the mean value theorem; the second limit on the right is zero since kD0F · xk/kxk ≤ kD0F k < +∞. Let r > 0 and xˆ be a unit vector. Set x = rxˆ in (45). By the first sentence of the proof, ψ(0) = Xψ(0). Canceling these equal terms from (45), dividing both sides of the resulting equation by ri+1, and taking i+1 ⊗(i+1) ⊗(i+1) i+1 ⊗(i+1) the limit as r → 0 using (46) yields D0ψ (D0F ) · xˆ = XD0ψ · xˆ for all unit vectors xˆ. Since derivatives are symmetric tensors, the latter equation has the form S · xˆ⊗(i+1) = T · xˆ⊗(i+1) with i+1 ⊗(i+1) i+1 symmetric tensors S := D0ψ (D0F ) and T := XD0ψ . Since symmetric tensors are completely determined by their action on tensors of the form xˆ⊗(i+1) [Tho14, Thm 1], it follows that 0 = S − T , or

i+1 ⊗(i+1) i+1 i+1 (47) 0 = D0 ψ (D0F ) − XD0 ψ = T(i+1)(D0 ψ), n ⊗(i+1) m n ⊗(i+1) m where the linear operator T(i+1) : L((R ) , C ) → L((R ) , C ) is as defined in Lemma1 (taking Y := D0F ). Lemma1 implies that T(i+1) is invertible since (X, D0F ) is k-nonresonant, so (47) implies that i+1 D0 ψ = 0. This completes the inductive step and the proof.  1 n n Lemma 3. Let F ∈ C (R , R ) be a diffeomorphism such that the origin is a globally attracting hyperbolic fixed point for the dynamical system defined by iterating F , where n ≥ 1. Fix k ∈ N≥1 ∪ {+∞} and A A A 0 ≤ α ≤ 1. Let e ∈ GL(m, C) have spectral radius ρ(e ) < 1 and satisfy ν(e , D0F ) < k + α. Assume k,α n m ψ ∈ Cloc (R , R ) satisfies (48) ψ ◦ F = eAψ and i (49) D0ψ = 0 for all integers 0 ≤ i < k + α (note the case α = 0 which does not require vanishing of the k-th derivative). Then ψ ≡ 0. 24 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

Proof. We first observe that since (i) 0 is asymptotically stable for the iterated dynamical system defined by F , (ii) ψ is continuous, and (iii) ρ(eA) < 1, it follows that ψ(0) = 0 since nA n (50) 0 = lim e ψ(x0) = lim ψ(F (x0)) = ψ(0) n→∞ n→∞ n for any x0 ∈ R . The second equality follows from (48). A n For the remainder of the proof, fix any r > 0 with ν(e , D0F ) < r < k + α, fix x0 ∈ R \{0}, and define j 0 k,α xj := F (x0) for j ∈ N. Let 0 ≤ k < r be the largest integer smaller than r. Taylor’s theorem for Cloc functions [dlLO99, p. 162] says that

k0 X i ⊗i ψ(x) = D0ψ · x + R(x), i=0 R(x) where limx→0 kxkr = 0. Equations (49) and (50) imply that all of the terms in the sum above vanish, so jA j j j we obtain ψ = R. Using (48) it follows that e ψ = ψ ◦ F = R ◦ F , and since xj = F (x0) we obtain

jA R(x) (51) e ψ(x0) = R(xj), lim = 0. x→0 kxkr A Denote by M := ρ(D0F ) < 1 the spectral radius of D0F . Since ν(e , D0F ) < r,(4) implies that all eigenvalues µ of eA satisfy |µ| > M r. Since this inequality is strict, by continuity there is  > 0 such that 0 < (M + ) < 1 and (52) ∀µ ∈ spec(eA): |µ| > (M + )r. 13 By replacing k · k with an adapted norm, we may assume that kD0F k < (M + /2). Since kF (x) − D0F · xk/kxk → 0 as kxk → 0, there exists b > 0 such that kF (x)k < (M + )kxk if kxk < b (cf. [HS74, p. 281]). Since the origin is globally asymptotically stable and since {kxk < b} is positively invariant by the preceding sentence (recall that (M + ) < 1), there exists j0 ∈ N≥1 such that kxjk < b for all j ≥ j0. Hence for all j ≥ j0:

j−j0 j (53) kxjk < (M + ) kxj0 k = C · (M + ) kx0k,

−j0 −1 where C := (M + ) kxj0 kkx0k . r (M+)jr Dividing both sides of (51) by kxjk , multiplying by 1 = (M+)jr and taking the limit as j → ∞ yields

j !r A !j jA ψ(x0) (M + ) e (54) 0 = lim e r = lim r ψ(x0). j→∞ kxjk j→∞ kxjk (M + )

eA Since (52) implies that all eigenvalues of (M+)r have modulus strictly larger than 1, the moduli of all eA j nonzero entries in the (upper triangular, complex) Jordan normal form of ( (M+)r ) approach ∞ as 14 eA j j → ∞. If ψ(x0) 6= 0, it follows that the absolute value of at least one component of ( (M+)r ) ψ(x0)  j r with respect to the Jordan basis approaches ∞ as j → ∞. Moreover, (53) implies that (M+) > kxj k −r −r eA j C kx0k > 0 for all j, so the product of this quantity with the diverging quantity ( (M+)r ) ψ(x0) also n diverges as j → ∞. It follows that (54) holds if and only if ψ(x0) = 0. Since x0 ∈ R \{0} was arbitrary, n and since we already obtained ψ(0) = 0 in (50), it follows that ψ ≡ 0 on R . This completes the proof.  Using Lemmas2 and3, we now prove the uniqueness portion of Theorem1.

13Such an adapted norm always exists. It can be constructed as the Euclidean norm with respect to a choice of basis placing D0F in “-Jordan form” [HS74, pp. 279–280], wherein the off-diagonal unity entries of the usual Jordan normal form are replaced by . An alternative construction of an adapted norm proceeds by suitably averaging a given norm along the dynamics linearized at the fixed point [CFdlL03a, Sec. A.1]; an analogous technique also works in more general situations. 14The desire for this conclusion was part of what motivated our definition of the spectral spread ν( · , · ). EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 25

Proof of the uniqueness portion of Theorem1. Since x0 is globally asymptotically stable, the Brown-Stallings n theorem [Wil67, Lem 2.1] implies that there is a diffeomorphism Q ≈ R sending x0 to 0, where n = dim(Q), n 15 1 so we may assume that Q = R and x0 = 0. Define the diffeomorphism F := Φ to be the time-1 map. A Let ψ1 and ψ2 be two functions satisfying D0ψi = B and ψi ◦ F = e ψi for i = 1, 2. Then ψ := ψ1 − ψ2 A i satisfies D0ψ = 0 and ψ ◦ F = e ψ. Lemma2 implies that D0ψ = 0 for all 1 ≤ i < k + 1, and Lemma3 A then implies that ψ1 − ψ2 = ψ ≡ 0. If e and B are real, then we can define ψ2 := ψ¯1 to be the complex A conjugate of ψ1, and so the preceding implies that ψ1 = ψ¯1; hence ψ1 is real if e and B are real. This completes the proof of the uniqueness statement of Theorem1. 

4.1.2. Proof of existence. In this section, we prove the existence portion of Theorem1. As with the proof of the uniqueness portion, the proof consists of an algebraic part and an analytic part. The techniques we use in the existence proof are similar to those used in [Ste57, CFdlL03a]. The algebraic portion of our proof is carried out in Lemma4, and the analytic portion is carried out in Lemma5.

Lemma 4 (Existence and uniqueness of approximate polynomial linearizing factors for diffeomorphisms). k n n Fix k ∈ N≥1 and let F ∈ C (R , R ) have the origin as a fixed point, where n ≥ 1. Let m ∈ N≥1 and m×m m×n X ∈ C be such that (X, D0F ) is k-nonresonant, and assume B ∈ C satisfies

BD0F = XB.

n m Then there exists a unique degree-k symmetric polynomial P : R → C vanishing at 0 such that D0P = B and such that

(55) P ◦ F = XP + R,

i m×m m×n where R satisfies D0R = 0 for all 0 ≤ i ≤ k. Furthermore, if X ∈ R and B ∈ R are real, then this n m m unique polynomial P : R → R ⊂ C is real. Remark 18. We state and prove Lemma4 for the case of finite k only and rely on a bootstrapping method to prove the existence portion of Theorem1 for the case k = +∞ at the end of this section. We believe it is possible to prove a C∞ version of the existence portion of Lemma4 (loosely speaking) using the fact that for every formal power series there exists a C∞ function with matching derivatives [Nel70, p. 34], but we did not attempt to do this.

Proof. By Lemma1, the linear operator

n ⊗i m n ⊗i m ⊗i (56) Ti : L((R ) , C ) → L((R ) , C ),Ti(Pi) := Pi(D0F ) − XPi

i n ⊗i m n ⊗i m is invertible for all 1 < i ≤ k. Denoting by Sym ((R ) , C ) ⊂ L((R ) , C ) the linear subspace n i m corresponding to symmetric multilinear maps (R ) → C via the universal property of the tensor product, i n ⊗i m i n ⊗i m we see from (56) that T (Sym ((R ) , C )) ⊂ Sym ((R ) , C ). Invertibility of Ti and dimension counting i n ⊗i m imply the opposite inclusion, so Ti restricts to a well-defined linear automorphism of Sym ((R ) , C ). By Taylor’s theorem we may uniquely write F as a degree-k symmetric polynomial plus remainder, Pk ⊗i i F (x) = i=1 Fi ·x +R1, where F1 = D0F and D0R1 = 0 for all 0 ≤ i ≤ k. Defining F⊗[j] := Fj1 ⊗· · ·⊗Fj` for any multi-index j ∈ N`≥1 and using the notation |j| := j1 + ··· + j`,(55) is equivalent to

k k X X ⊗|j| X ⊗` (57) P` · F⊗[j] · x = X P` · x , `=1 ` `=1 j∈N≥1 |j|≤k

15For example, Wilson states in [Wil67, Thm 2.2] this result for the special case of a flow generated by a C1 vector field, but his argument based on the Brown-Stallings theorem [Wil67, Lem 2.1] works equally well for any C1 flow or diffeomorphism having a globally asymptotically stable fixed point. 26 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

P` Pk ⊗` where j = |(j1, . . . , j`)| = i=1 ji, P (x) = `=1 P` ·x , and P1 = B. It follows from an inductive argument (equating coefficients of x⊗i) that (55) is equivalent to    i  X X  ⊗i ⊗i (58) ∀i ∈ {1, . . . , k}:  P F  · x = XPi · x .  ` ⊗[j] `=1 `  j∈N≥1  |j|=i ⊗i If we require that all tensors P` are symmetric then, for each fixed i, the two tensors acting on x in (58) are symmetric. Since symmetric tensors are completely determined by their action on all tensors of the form x⊗i [Tho14, Thm 1], it follows that (55) is equivalent to i ∀i ∈ {1, . . . , k}: X P X F = XP (59) ` ⊗[j] i `=1 ` j∈N≥1 |j|=i or, after rearranging terms, i−1 ∀i ∈ {1, . . . , k}: X P X F = XP − P (D F )⊗i (60) ` ⊗[j] i i 0 `=1 j∈ ` | {z } N≥1 −Ti(Pi) |j|=i since D0F = F1. By our assumptions, XB −BD0F = 0 and P1 = D0P = B. Moreover, as discussed above, i n ⊗i m T | i is a well-defined linear automorphism of the subspace Sym (( ) , ). Thus, (60) can i Sym ((Rn)⊗i,Cm) R C be inductively solved for the tensors P1,...,Pk, and the preceding sentence implies that these solutions are unique and symmetric. Thus, so is P . m×m m×n Finally, assume that X ∈ R and B ∈ R are real, and assume by induction that B = P1,P2,...,Pi−1 are real. Taking the complex conjugate of (60), we see that Pi solves (60) if and only if its complex con- jugate P¯i solves (60). Invertibility of Ti thus implies that Pi = P¯i, so Pi is real. By induction, P1,...,Pk are real. Thus, so is P . This completes the proof. 

Lemma 5 (Making approximate linearizing factors exact). Fix k ∈ N≥1 ∪ {+∞}, 0 ≤ α ≤ 1, and let n n k,α F : R → R be a Cloc diffeomorphism such that the origin is a globally attracting hyperbolic fixed point A for the dynamical system defined by iterating F , where n ≥ 1. Let m ∈ N≥1 and e ∈ GL(m, C) satisfy A k,α n m ν(e , D0F ) < k + α, and assume that there exists P ∈ Cloc (R , C ) such that P ◦ F = eAP + R, k,α n m i where R ∈ Cloc (R , C ) satisfies D0R = 0 for all integers 0 ≤ i < k + α (note the case α = 0 which does not require vanishing of the k-th derivative). k,α n m i Then there exists a unique ϕ ∈ Cloc (R , C ) such that D0ϕ = 0 for all integers 0 ≤ i < k + α and such that ψ := P + ϕ satisfies ψ ◦ F = eAψ. In fact, ψ = lim e−jAP ◦ F j j→∞ k,α n in the topology of C -uniform convergence on compact subsets of R if k < +∞, and in the topology k0 0 of C -uniform convergence on compact subsets of Q for any k ∈ N≥1 if k = +∞. Furthermore, if A k,α n m n m m e ∈ GL(m, R) is real and P ∈ Cloc (R , R ) is real, then ϕ, ψ : R → R ⊂ C are real. Proof. We first assume that k < +∞ and delay consideration of the case k = +∞ until the end of the proof. Adapted norms. Later in the proof we will require that the following bound on operator norms holds (needed following (72)): −A k+α (61) ke kkD0F k < 1. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 27

A Due to our assumption that ν(e , D0F ) < k + α, this bound can always be made to hold by using an appropriate choice of “adapted” norms (which induce the operator norms) on the underlying vector spaces n m R and C , and so we may (and do) assume that (61) holds in the remainder of the proof. A But first we argue that such norms can indeed be chosen. Let λ ∈ spec(D0F ) and µ ∈ spec(e ) be the A eigenvalues of D0F and e with largest and smallest modulus, respectively. For any κ > 0, there exist n m adapted norms (both denoted by k·k) on R and C having the property that the induced operator norms −A ke k and kD0F k satisfy [HS74, pp. 279–280], [CFdlL03a, Sec. A.1]:

−A −1 (62) |kD0F k − |λ|| ≤ κ, ke k − |µ| ≤ κ. A −1 k+α Now since ν(e , D0F ) < k + α and since |λ| < 1, it follows from (4) that |µ| |λ| < 1. The inequalities −A k+α −1 k+α (62) imply that ke kkD0F k can be made arbitrarily close to |µ| |λ| by making κ small, so choosing κ sufficiently small yields (61) as claimed. For later use we also note (62) implies that D0F is a strict contraction if κ is small enough since |λ| < 1. This in turn implies that (63) F (B) ⊂ B n if B ⊂ R is a sufficiently small (adapted norm) ball centered at the origin [HS74, p. 281]. We henceforth assume this is the case. n Definition of function spaces. Let U ⊂ R be a precompact and denote by U¯ = cl(U) its k ¯ compact . Given any Banach space X and k ∈ N≥0, let C (U,X) be the space of continuous functions G: U¯ → X having partial derivatives of order less than or equal to k which are uniformly continuous on U, in which case they extend to continuous functions on U¯. We equip Ck(U,X¯ ) with the standard norm k X i kGkk := supkDxGk i=0 x∈U making Ck(U,X¯ ) into a Banach space [dlLO99]. For a Banach space Y and 0 < α ≤ 1, we define the α-Hölder constant [H]α of a map H : U¯ → Y via kH(x) − H(y)k : [H]α = sup α . x,y∈U kx − yk x6=y k,α k k For 0 < α ≤ 1 we let C (U,X¯ ) be the subset of functions G ∈ C (U,X¯ ) for which [D G]α < +∞, and we equip Ck,α(U,X¯ ) with the standard norm k (64) kGkk,α := kGkk + [D G]α making Ck,α(U,X¯ ) into a Banach space [dlLO99].16 For α = 0, we identify Ck,0(U,X¯ ) with Ck(U,X¯ ) and make the special definition k · kk,0 := k · kk. n k,α m In what follows, let B ⊂ R be a closed ball centered at the origin and let F ⊂ C (B, C ) denote the i subspace of functions ϕ such that D0ϕ = 0 for all integers 0 ≤ i < k + α; F is a closed linear subspace of k,α m C (B, C ), hence also a Banach space. i−1 Preliminary estimates. For any ϕ ∈ F, x ∈ B, and all integers 1 ≤ i < k + α, we have that kDx ϕk ≤ R 1 i kxk · 0 kDtxϕk dt by the fundamental theorem of calculus, the chain rule, and the definition of F. The derivatives Diϕ vanish at x = 0 for all integers 0 ≤ i < k + α by the definition of F, so the preceding sentence and an induction argument imply that, for any  > 0, if the radius of B is sufficiently small then for any ϕ ∈ F: k kϕkk−1 ≤ kD ϕk0 (65) k kϕkk ≤ (1 + )kD ϕk0.

16 k k,α i Different C and C norms are actually used in [dlLO99, Def. 2.1, Def. 2.5], namely, max0≤i≤kkD Gk0 and i k max(max0≤i≤kkD Gk0, [D G]α), but these two norms are equivalent (in the sense of norms) to the corresponding norms we have chosen. 28 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

k k If α > 0, the additional fact that kDxϕk ≤ kxk[D ϕ]α further implies that k kϕkk ≤ [D ϕ]α (66) k kϕkk,α ≤ (1 + )[D ϕ]α if the radius of B is sufficiently small. n n Defining a linear contraction mapping on F. Recall that F : R → R is the diffeomorphism from the n statement of the lemma. By (63), all sufficiently small closed (adapted norm) balls B ⊂ R centered at k,α n n k,α n the origin satisfy F (B) ⊂ B. Additionally, since F ∈ Cloc (R , R ) and B is compact, F |B ∈ C (B, R ). k,α m k,α m 17 It follows that there is a well-defined linear operator T : C (B, C ) → C (B, C ) given by −A (67) T (ϕ) := e ϕ ◦ F |B. Note that T (F) ⊂ F, so that F is an invariant subspace for T . We claim that there is a choice of B so that T |F : F → F is a (strict) contraction with constant β < 1:

(68) ∀ϕ ∈ F : kT (ϕ)kk,α ≤ βkϕkk,α. To show this we give an argument essentially due to Sternberg, but which generalizes the proof of [Ste57, Thm 2] to the case of linearizing semiconjugacies and to the Ck,α setting (allowing α > 0). Using the ⊗[j] j1 ji i notation Dx F := Dx F ⊗ · · · ⊗ Dx F for a multi-index j ∈ N≥1, we compute k−1 k −A k ⊗k −A X X i ⊗[j] (69) Dx(T (ϕ)) = e DF (x)ϕ · (DxF ) + e Ci,jDF (x)ϕ · Dx F, i=1 i j∈N≥1 |j|=k where the integer coefficients Ci,j ∈ N≥1 are combinatorially determined by Faà di Bruno’s formula for the “higher-order chain rule” and are therefore independent of B.18 We choose B sufficiently small that its diameter is less than 1 and note that, by continuity and compactness of {kxk ≤ 1} and the fact that k,α 19 F ∈ Cloc , there exists a constant N0 > 0 such that   k−1 ! ! ⊗[j] ⊗[j] X X  ⊗[j] α kDx F − Dy F k (70) Ci,j ·  sup kDx F k 1 + sup kDxF k + sup α  < N0.  kxk≤1 kxk≤1 kxk,kyk≤1 kx − yk  i=1 j∈ ` N≥1 x6=y |j|=k Using (65) and (70) to bound the sum in (69), it follows that k −A k k (71) kD T (ϕ)k0 ≤ ke k(kDF k0 + N0)kD ϕk0. k For the case that α > 0, we will now use (66) to obtain a bound on [DxT (ϕ)]α analogous to (71). k k α In order to do this, we use the estimate [x 7→ DF (x)ϕ]α ≤ [D ϕ]αkDF k0 and the product rule [fg]α ≤ kfk0[g]α + [f]αkgk0 for Hölder constants (see, e.g., [Eld13, Lem 1.19]) to bound the first term of (69) by −A  k k+α k ⊗k  −A  k+α ⊗k  k ke k [D ϕ]αkDF k0 + kD ϕk0[(DF ) ]α ≤ ke k kDF k0 + [(DF ) ]α [D ϕ]α, where we have used (66) to bound the second term in parentheses on the left side. Next, we use (66), i (70), the product rule for Hölder constants again, and for 1 ≤ i ≤ k − 1 the estimates [x 7→ DF (x)ϕ]α ≤ i α k α −A k [D ϕ]αkDF k0 ≤ [D ϕ]αkDF k0 to bound the second term of (69) by N0ke k[D ϕ]α. This last estimate

17That DkT (ϕ) is α-Hölder follows from the chain rule, the fact that the first k − 1 derivatives of F and of ϕ are C1 and hence Lipschitz (hence also α-Hölder), the fact that the composition of an α-Hölder function with a Lipschitz function is again α-Hölder, and the fact that the product of bounded α-Hölder functions is again α-Hölder (see, e.g., [Eld13, Lem 1.19]). 18The “higher-order chain rule,” also known as Faà di Bruno’s formula, gives a general expression for higher-order derivatives of the composition of two functions (see [Jac14] for an exposition). 19Since we have not yet chosen B (other than stipulating that its diameter be smaller than 1), to avoid circular reasoning we are using {kxk ≤ 1} in place of B in (70) to make clear that the estimate holds for all closed balls B ⊂ {kxk ≤ 1} centered at 0. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 29 we used follows from (66) and the fact that we are requiring B to have diameter less than 1, so that i i+1 k [D ϕ]α ≤ kD ϕk0 ≤ [D ϕ]α. We finally obtain h k i −A  k+α ⊗k  k (72) DxT (ϕ) ≤ ke k kDF k + [(DF ) ]α + N0 [D ϕ]α. α 0 −A k+α The estimate (61) and continuity imply that ke kkDF k0 < 1 if B is sufficiently small. Hence if  k k is sufficiently small, the quantities respectively multiplying kD ϕk0 and [D ϕ]α in (71) and (72) will be bounded above by some positive constant β0 < 1. The discussion preceding (65) and (66) implies that we k 0 k can indeed take  this small after possibly further shrinking B, so it follows that kD T (ϕ)k0 < β kD ϕk0 k 0 k and, if α > 0, also [D T (ϕ)]α < β [D ϕ]α. We therefore obtain a contraction estimate on the highest derivative and its Hölder constant (if α > 0) only. However, we can combine this observation with the second inequalities from each of the two displays (65) and (66), together with the fact that T (F) ⊂ F, to obtain in both cases (α = 0 and α > 0) the following estimate involving all of the derivatives: 0 (73) kT (ϕ)kk,α ≤ (1 + )β kϕkk,α. (This technique for the case α = 0 was also used in the proof of [Ste57, Thm 2].) Define β := (1 + )β0. Since β0 < 1, if necessary we may shrink B further to ensure that  may be taken sufficiently small that β < 1. It then follows that T |F is a (strict) contraction; this completes the proof of (68). Existence and uniqueness of a linearizing factor defined on B. We will now find a locally-defined (i.e., n ˜ k,α m ˜ defined on B rather than on all of R ) linearizing factor ψ ∈ C (B, C ) of the form ψ = P |B +ϕ ˜, where n m ϕ˜ ∈ F and P : R → C is as in the statement of the lemma. By definition, ψ˜ is linearizing if and only ˜ −A ˜ ˜ if ψ = e ψ ◦ F |B = T (ψ), so we need to solve the equation P |B +ϕ ˜ = T (P |B +ϕ ˜) for ϕ˜ ∈ F. (We are n writing P |B rather than P because ϕ˜ is a function with domain B rather than R , and also because T is a linear operator defined on functions with domain B.) Since T is linear, after rearranging we see that this amounts to solving   (74) idCk,α(B,Cm) − T ϕ˜ = T (P |B) − P |B.  A  One of the assumptions of the lemma directly implies that P ◦ F − e P |B ∈ F, and this implies that the right hand side of (74) belongs to F since e−A ·F ⊂ F. Since T (F) ⊂ F, it follows that we may rewrite (74) as

(75) (idF − T |F )ϕ ˜ = T (P |B) − P |B.

We showed earlier that T |F is a strict contraction, i.e., its operator norm satisfies kT |F kk,α < 1. It follows that (idF − T |F ): F → F has a bounded inverse given by the corresponding Neumann series, so that (75) has a unique solution ϕ˜ given by ∞ −1 X n (76) ϕ˜ = (idF − T |F ) · (T (P |B) − P |B) = (T |F ) · (T (P |B) − P |B) , n=0 ˜ ˜ −A ˜ ˜ and ψ := P |B +ϕ ˜ satisfies ψ = e ψ ◦ F |B = T (ψ) as discussed above. Extension to a unique global linearizing factor. Since x0 is globally asymptotically stable and since B is j positively invariant, for every x ∈ B there exists j(x) ∈ N≥0 such that, for all j > j(x), F (x) ∈ int(B). If j is large enough that F j(x) ∈ int(B) and ` > j, then −`A ˜ ` −jA  (j−`)A ˜ (`−j)  j −jA  (`−j) ˜  j −jA ˜ j e ψ(F (x)) = e e ψ ◦ F |B (F (x)) = e T (ψ) (F (x)) = e ψ(F (x)), n m so there is a well-defined map ψ : R → C given by (77) ψ(x) := e−jAψ˜(F j(x)), j ˜ A ˜ where j ∈ N≥0 is any nonnegative integer sufficiently large that F (x) ∈ int(B). Since ψ ◦ F |B = e ψ, A n j it follows from (77) that ψ ◦ F = e ψ. If x ∈ R and F (x) ∈ int(B), then x has a neighborhood U j with F (U) ⊂ int(B) by continuity, so ψ|U is given by (77) with j constant on U. By the chain rule and k,α n m standard properties of locally α-Hölder functions (see Footnote 17), this shows that ψ ∈ Cloc (R , C ). From (77) we see that ψ and hence also ϕ := ψ−P are uniquely determined by ψ|B = P |B +ϕ|B = P |B +ϕ ˜, 30 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS which is in turn uniquely determined by ϕ˜, and since ϕ˜ is unique it follows that ϕ and ψ are also unique. A k,α n m ¯ If e ∈ GL(m, R) and P ∈ Cloc (R , R ) are real, then the complex conjugate ψ = P +ϕ ¯ also satisfies A n m m ψ¯ ◦ F = e ψ¯ with ϕ¯ ∈ F, so uniqueness implies that ψ¯ = ψ and hence ψ, ϕ: R → R ⊂ C are real. Convergence to the global linearizing factor. We now complete the proof of the lemma by proving the sole remaining claim that e−jAP ◦F j → ψ in the topology of Ck,α-uniform convergence on compact subsets n of R . To do this, we first inspect the finite truncations of the infinite series in (76). We see that, since j j X n X n+1 n j+1 (T |F ) · (T (P |B) − P |B) = T (P |B) − T (P |B) = T (P |B) − P |B n=0 n=0 j for each j ∈ N≥1, taking the limit j → ∞ shows that the series in (76) is equal to −P |B +limj→∞ T (P |B), k,α m with convergence in the Banach space C (B, C ). In other words, ˜ −jA j (78) ψ = lim e P ◦ F |B j→∞ k,α m with convergence in C (B, C ). n k,α m Next, let K ⊂ R be the closure of any precompact open set (so that the Banach space C (K, C ) is defined as in the text containing (64)). Since 0 is globally asymptotically stable and since B contains j a neighborhood of 0, there exists j0 > 0 such that F (K) ⊂ B for all j ≥ j0. We justify the following computation below:   −jA j −j0A −jA j j0 lim e P ◦ F |K = lim e e P ◦ F |B ◦ F |K j→∞ j→∞   −j0A −jA j j0 = e lim e P ◦ F |B ◦ F |K j→∞

−j0A j0 = e ψ|B ◦ F |K

= ψ|K , k,α m k,α m with convergence in C (K, C ). Since we are considering convergence in the space C (K, C )—rather than mere pointwise convergence—it is not obvious that we can move the limit inside the parentheses to k,α m k,α m obtain the second equality. The reason this is valid is that composition maps C (K, C ) → C (K, C ) ∞ k,α m k,α of the form g 7→ f ◦ g ◦ h, where f ∈ C and h ∈ C (K, C ), are continuous with respect to the C normed [dlLO99, Prop. 6.1, Prop. 6.2 (iii)].20 This completes the proof for the case k < +∞. Consideration of the case k = +∞. For the case k = +∞, repeating the proof above for any 1 ≤ k0 < +∞ A 0 k0 n m i such that ν(e , D0F ) < k yields unique C functions ϕ: R → C and ψ := P + ϕ such that D0ϕ = 0 for all 0 ≤ i ≤ k0 − 1. By the uniqueness statement already proved for the case k < +∞, these functions ϕ, ψ 0 A 0 ∞ n m are independent of k > ν(e , D0F ), and since k is arbitrary it follows that ϕ, ψ ∈ C (R , C ). Finally, −jA j for the closure K of any precompact open set, we have already shown that e ◦ P ◦ F |K → ψ|K in the k0 m 0 Banach space C (K, C ) for every k ∈ N≥1, as desired. This completes the proof.  Using Lemmas4 and5, we now complete the proof of Theorem1 by proving the existence portion of its statement. Proof of the existence portion of Theorem1. As in the proof of the uniqueness portion of Theorem1 at n the end of §4.1.1, we may assume that Q = R and x0 = 0. We first consider the case that T = Z, and define the time-1 map F := Φ1. First suppose that k < +∞. Lemma4 implies that there exists a polynomial P such that D0P = B and A k,α n m i 21 P ◦ F = e P + R, where R ∈ Cloc (R , C ) satisfies D0R = 0 for all integers 0 ≤ i < k + α. Furthermore, A k,α n m P and R are real if e and B are real. Lemma5 then implies that there exists ϕ ∈ Cloc (R , C ) such that k,α n m A −jA ˜ j k,α ψ = P + ϕ ∈ Cloc (R , C ) satisfies D0ψ = D0P = B, ψ ◦ F = e ψ, e P ◦ Φ → ψ C -uniformly on

20We remark that, for maps between finite-dimensional spaces, there are somewhat weaker assumptions ensuring continuity of such composition maps [dlLO99, Rem. 6.5], but our present situation does not require this. 21 i Actually, Lemma4 implies that we can find P such that D0R = 0 for all integers 0 ≤ i ≤ k, with the only difference arising when α = 0. However, we do not need this in the following. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 31

n compact subsets of R for any “approximate linearizing factor” P˜ satisfying the hypotheses of Theorem1 (such as P ), and that ψ is real if eA and B are real. This completes the proof for the case k < +∞. 0 A k0 n m Now suppose that k = +∞. Repeating the proof above for finite k > ν(e , D0) yields ψ ∈ C (R , C ) A satisfying D0ψ = B and ψ ◦ F = e ψ. The uniqueness statement of Theorem1 proved in §4.1.1 implies 0 A 0 ∞ that ψ is independent of k > ν(e , D0F ), so since k is arbitrary it follows that ψ ∈ C . Additionally, we −jA ˜ j k0 n 0 have already argued above that e P ◦ Φ → ψ C -uniformly on compact subsets of R for any k ∈ N≥1 and any “approximate linearizing factor” P˜ satisfying the hypotheses of Theorem1. This completes the proof for the case that T = Z. It remains only to consider the case that T = R, i.e., the case that Φ is a flow. By the proof of the case ˜ k,α n m ˜ ˜ j jA ˜ T = Z, there exists ψ ∈ Cloc (R , C ) satisfying D0ψ = B and ψ ◦ Φ = e ψ for all j ∈ Z. By adapting ˜ k,α n m a technique of Sternberg [Ste57, Lem 4], from ψ we will construct a map ψ ∈ Cloc (R , C ) satisfying t tA D0ψ = B and ψ ◦ Φ = e ψ for all t ∈ R. In fact, we will show that Z 1 (79) ψ := e−sAψ˜ ◦ Φs ds 0 has these properties. By Leibniz’s rule for differentiating under the integral sign and standard properties k,α n m of locally α-Hölder functions (see Footnote 17), ψ ∈ Cloc (R , C ), and using the assumption (5) we have that Z 1 Z 1 −sA s D0ψ = e BD0Φ ds = B ds = B. 0 0 t tA To prove that ψ ◦ Φ = e ψ for all t ∈ R, we compute Z 1 Z 1+t ψ ◦ Φt = e−sAψ˜ ◦ Φs+t ds = e(t−s)Aψ˜ ◦ Φs ds 0 t Z 1 Z 1+t = etA e−sAψ˜ ◦ Φs ds + etA e−sAψ˜ ◦ Φs ds t 1 Z 1 Z 1+t   = etA e−sAψ˜ ◦ Φs ds + etA e−sA eAψ˜ ◦ Φ−1 ◦ Φs ds t 1 Z 1 Z t = etA e−sAψ˜ ◦ Φs ds + etA e−sAψ˜ ◦ Φs ds t 0 = etAψ

1 A as desired. We remark that, since ψ satisfies ψ ◦ Φ = e ψ and D0ψ = B, the uniqueness result for the case T = Z actually implies the (non-obvious) fact that ψ = ψ˜. n Suppose now that k < +∞. Letting K ⊂ R be the closure of any precompact open set (so that the k,α m Banach space C (K, C ) is defined as in the text containing (64)) which is also positively invariant, k,α m k,α m −rA r the map G: [0, 1] × C (K, C ) → C (K, C ) given by G(r, f) := e f ◦ Φ |K is continuous [dlLO99, Thm 6.10] and satisfies G(r, ψ|K ) = ψ|K for all r ∈ [0, 1]. Thus, compactness of [0, 1] implies that, for k,α m every neighborhood V ⊂ C (K, C ) of ψ|K , there is a smaller neighborhood U ⊂ V of ψ|K such that −rA r G([0, 1] × U) ⊂ V , i.e., e ϕ ◦ Φ |K ⊂ V for every ϕ ∈ U and r ∈ [0, 1]. Fix any such neighborhoods V,U k,α n m and fix any “approximate linearizing factor” P ∈ Cloc (R , C ) satisfying the hypotheses of Theorem1. −jA j By the proof for the case T = Z there exists N ∈ N≥0 such that, for all j > N, e P ◦ Φ |K ∈ U. By the −tA t definition of U it follows that e P ◦ Φ |K ⊂ V for all t > N + 1. Since the neighborhood V 3 ψ|K was arbitrary, this implies that −tA t (80) ψ|K = lim e P ◦ Φ |K t→∞ k,α m with convergence in the Banach space C (K, C ). If instead k = +∞, the same argument shows that k0 m 0 n (80) converges in the Banach space C (K, C ) for every k ∈ N≥1. Since every compact subset of R is contained in the closure K of some positively invariant precompact open set (e.g., a sublevel set of a smooth Lyapunov function [Wil69, FP19]), this proves that e−tAP ◦Φt → ψ in the topology of Ck,α-uniform 32 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS convergence on compact sets if k < +∞, and in the topology of Ck0 -uniform convergence on compact sets 0 for any k ∈ N≥1 if k = +∞. This completes the proof of Theorem1.  4.2. Proof of Theorem2. In this section we prove Theorem2, which we repeat here for convenience. This proof invokes Theorem1 and is much shorter because of this. k,α Theorem 2 (Existence and uniqueness of Cloc global linearizing factors for a limit cycle attractor). Fix k,α k ∈ N≥1 ∪ {+∞} and 0 ≤ α ≤ 1. Let Φ: Q × R → Q be a Cloc flow with Q the basin of an attracting hyperbolic τ-periodic orbit with image Γ ⊂ Q, where Q is a smooth manifold with dim(Q) ≥ 2. Fix s τ x0 ∈ Γ and let Ex0 denote the unique Dx0 Φ -invariant subspace complementary to Tx0 Γ. Let m ∈ N≥1 and τA τA s m e ∈ GL(m, C) have spectral radius ρ(e ) < 1, and let the linear map B : Ex0 → C satisfy τ τA (18) BDx Φ |Es = e B. 0 x0 τA τ τA τ Assume that (e , Dx Φ |Es ) is k-nonresonant, and assume that ν(e , Dx Φ |Es ) < k + α. 0 x0 0 x0 k,α m Then there exists a unique ψ ∈ Cloc (Q, C ) satisfying t tA (19) ∀t ∈ : ψ ◦ Φ = e ψ, Dx ψ|Es = B, R 0 x0 s m m m×m m×m m m and if B : Ex0 → R ⊂ C and A ∈ R ⊂ C are real, then ψ : Q → R ⊂ C is real. s s Proof. Let Wx0 be the global strong stable manifold (isochron) through x0 [EKR18, Sec. 2.1]. Since Wx0 k,α τ s k,α is the stable manifold for the fixed point x0 of the Cloc diffeomorphism Φ , it follows that Wx0 is a Cloc submanifold [Rue89, pp. 2, 27; Thm 6.1] which is properly embedded (rather than merely immersed) in Q because Γ is stable [EKR18, pp. 4208–4209]. s n After identifying E with , the uniqueness portion of Theorem1 applied to ψ|W s implies that ψ|W s x0 R x0 x0 is unique for any ψ satisfying the uniqueness hypotheses, and furthermore ψ|W s is real if A and B are x0 S t s real. Since ψ is uniquely determined by ψ|W s and (19) (which is true because Q = Φ (W )), this x0 t∈R x0 implies that ψ is unique and that ψ is real if A and B are real. This completes the proof of the uniqueness statement of Theorem2. Under the existence hypotheses, the existence portion of Theorem1 similarly implies that there exists k,α s m a unique ϕ ∈ Cloc (Wx0 , C ) satisfying Dx0 ϕ = B and jτ jτA (81) ∀j ∈ : ϕ ◦ Φ |W s = e ϕ. Z x0 k,α m The unique extension of ϕ to a Cloc function ψ : Q → C satisfying (19) is given by −tA t (82) ∀t ∈ : ψ| s := e ϕ ◦ Φ | s . R W −t W −t Φ (x0) Φ (x0) τ s s τA −τ k,α ψ is well-defined because Φ (W ) = W and e ϕ ◦ Φ |W s = ϕ by (81). That ψ ∈ C follows from x0 x0 x0 loc k,α s considering locally-defined Cloc “time-to-impact Wx0 ” functions as in the proof of Proposition3. This completes the proof. 

Acknowledgments The majority of this work was performed while Kvalheim was a postoctoral researcher at the University of Michigan. Both authors were supported by ARO award W911NF-14-1-0573 to Revzen and by the ARO under the Multidisciplinary University Research Initiatives (MURI) Program, award W911NF-17-1-0306 to Revzen. Kvalheim was also supported by the ARO under the SLICE MURI Program, award W911NF- 18-1-0327. We thank George Haller, David Hong, Igor Mezić, Jeff Moehlis, Corbinian Schlosser, and Dan Wilson for helpful discussions and comments. We owe special gratitude to Alexandre Mauroy and Ryan Mohr for their careful reading of the manuscript and valuable feedback; in particular, one of Mauroy’s comments led to a sharpening of the uniqueness statements of Theorems1 and2 and Propositions6 and 7, and Mohr found an error in Definition1. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 33

References [AM17a] H Arbabi and I Mezić, Ergodic theory, dynamic mode decomposition, and computation of spectral properties of the Koopman operator, SIAM Journal on Applied Dynamical Systems 16 (2017), no. 4, 2096–2126. [AM17b] , Study of dynamics in post-transient flows using Koopman mode decomposition, Physical Review Fluids 2 (2017), no. 12, 124402. [AMR] R Abraham, J E Marsden, and T Ratiu, Manifolds, tensor analysis, and applications, 3 ed. [AR67] R Abraham and J Robbin, Transversal mappings and flows, W A Benjamin New York, 1967. [AT19] H Arbabi and S Themistoklis, Data-driven modeling of strongly nonlinear chaotic systems with non-Gaussian statistics, arXiv preprint arXiv:1908.08941 (2019). [BBPK16] S L Brunton, Bingni W Brunton, J L Proctor, and J N Kutz, Koopman invariant subspaces and finite linear representations of nonlinear dynamical systems for control, PloS one 11 (2016), no. 2, e0150171. [BK94] I U Bronstein and A Y Kopanski˘ı, Smooth invariant manifolds and normal forms, World Scientific Series on Nonlinear Science. Series A: Monographs and Treatises, vol. 7, World Scientific Publishing Co., Inc., River Edge, NJ, 1994. MR 1337026 [BMM12] M Budišić, R Mohr, and I Mezić, Applied Koopmanism, Chaos: An Interdisciplinary Journal of Nonlinear Science 22 (2012), no. 4, 047510. [BRV19] D Bruder, C D Remy, and R Vasudevan, Nonlinear system identification of soft robot dynamics using Koopman operator theory, 2019 International Conference on Robotics and Automation (ICRA), IEEE, 2019, pp. 6244–6250. [CFdlL03a] X Cabré, E Fontich, and R de la Llave, The parameterization method for invariant manifolds I: manifolds associated to non-resonant subspaces, Indiana University mathematics journal (2003), 283–328. [CFdlL03b] , The parameterization method for invariant manifolds II: regularity with respect to parameters, Indiana University mathematics journal (2003), 329–360. [CFdlL05] , The parameterization method for invariant manifolds III: overview and applications, Journal of Differ- ential Equations 218 (2005), no. 2, 444–515. [DG19] S Das and D Giannakis, Delay-coordinate maps and the spectra of Koopman operators, Journal of Statistical Physics (2019), 1–39. [dlLO99] R de la Llave and R Obaya, Regularity of the composition operator in spaces of Hölder functions, Discrete and Continuous Dynamical Systems 5 (1999), 157–184. [DTK19] F Dietrich, T N Thiem, and I G Kevrekidis, On the Koopman operator of algorithms, arXiv preprint arXiv:1907.10807 (2019). [EKR18] J Eldering, M Kvalheim, and S Revzen, Global linearization and fiber bundle structure of invariant manifolds, Nonlinearity 31 (2018), no. 9, 4202–4245. [Eld13] J Eldering, Normally hyperbolic invariant manifolds: the noncompact case, Atlantis Press, 2013. [EN00] K-J Engel and R Nagel, One-parameter semigroups for linear evolution equations, Springer, 2000. [Fen71] N Fenichel, Persistence and smoothness of invariant manifolds for flows, Indiana Univ. Math. J. 21 (1971), 193–226. MR 0287106 [Fen74] , Asymptotic stability with rate conditions, Indiana University Mathematics Journal 23 (1974), no. 12, 1109–1137. [Fen77] , Asymptotic stability with rate conditions II, Indiana University Mathematics Journal 26 (1977), no. 1, 81–93. [FH19] T Fisher and B Hasselblatt, Hyperbolic flows, European Mathematical Society, 2019. [FP19] A Fathi and P Pageault, Smoothing Lyapunov functions, Transactions of the American Mathematical Society 371 (2019), no. 3, 1677–1700. [Gro59] D M Grobman, Homeomorphism of systems of differential equations, Doklady Akademii Nauk SSSR 128 (1959), no. 5, 880–881. [GS85] M Golubitsky and D Schaeffer, Singularities and groups in bifurcation theory, Springer-Verlag, New York, 1985. [GSZ15] D Giannakis, J Slawinska, and Z Zhao, Spatiotemporal feature extraction with data-driven Koopman operators, Feature Extraction: Modern Questions and Challenges, 2015, pp. 103–115. [Guc75] J M Guckenheimer, Isochrons and phaseless sets, Journal of Mathematical Biology 1 (1975), 259–273. [Had02] J Hadamard, Sur les problèmes aux dérivées partielles et leur signification physique, Princeton University Bulletin (1902), 49–52. [Hal] J K Hale, Ordinary differential equations, 2 ed. [Har60a] P Hartman, A lemma in the theory of structural stability of differential equations, Proceedings of the American Mathematical Society 11 (1960), no. 4, 610–620. [Har60b] , On local homeomorphisms of Euclidean spaces, Bol. Soc. Mat. Mexicana 5 (1960), no. 2, 220–241. [Hir94] M W Hirsch, Differential topology, Graduate Texts in Mathematics, vol. 33, Springer-Verlag, New York, 1994, Corrected reprint of the 1976 original. MR 1336822 [HPS77] M W Hirsch, C C Pugh, and M Shub, Invariant manifolds, Lecture Notes in Mathematics, Vol. 583, Springer- Verlag, Berlin-New York, 1977. MR 0501173 [HS74] M W Hirsch and S Smale, Differential equations, dynamical systems, and linear algebra, Academic press, 1974. 34 EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS

[Jac14] H O Jacobs, How to stare at the higher-order n-dimensional chain rule without losing your marbles, arXiv preprint arXiv:1410.3493 (2014). [KKB17] E Kaiser, J N Kutz, and S L Brunton, Data-driven discovery of Koopman eigenfunctions for control, arXiv preprint arXiv:1707.01146 (2017). [KM18] M Korda and I Mezić, On convergence of extended dynamic mode decomposition to the Koopman operator, Journal of Nonlinear Science 28 (2018), no. 2, 687–710. [KM19] , Optimal construction of Koopman eigenfunctions for prediction and control, arXiv preprint arXiv:1810.08733 (2019). [Koo31] B O Koopman, Hamiltonian systems and transformation in , Proceedings of the National Academy of Sciences of the United States of America 17 (1931), no. 5, 315. [KPM18] M Korda, M Putinar, and I Mezić, Data-driven spectral analysis of the Koopman operator, Applied and Compu- tational Harmonic Analysis (2018). [Lee13] J M Lee, Introduction to smooth manifolds, 2 ed., Springer, 2013. [LM13] Y Lan and I Mezić, Linearization in the large of nonlinear systems and Koopman operator spectrum, Phys. D 242 (2013), 42–53. MR 3001394 [MB04] I Mezić and A Banaszuk, Comparison of systems with complex behavior, Physica D: Nonlinear Phenomena 197 (2004), no. 1-2, 101–133. [Mez94] I Mezic, On the geometrical and statistical properties of dynamical systems: theory and applications, Ph.D. thesis, California Institute of Technology, 1994. [Mez05] I Mezić, Spectral properties of dynamical systems, model reduction and decompositions, Nonlinear Dynamics 41 (2005), no. 1-3, 309–325. [Mez15] , On applications of the spectral theory of the Koopman operator in dynamical systems and control theory, 2015 54th IEEE Conference on Decision and Control (CDC), IEEE, 2015, pp. 7034–7041. [Mez19] , Spectrum of the Koopman operator, spectral expansions in functional spaces, and state-space geometry, Journal of Nonlinear Science (2019), 1–55. [MM12] A Mauroy and I Mezić, On the use of Fourier averages to compute the global isochrons of (quasi) periodic dynamics, Chaos: An Interdisciplinary Journal of Nonlinear Science 22 (2012), no. 3, 033112. [MM14] R Mohr and I Mezić, Construction of eigenfunctions for scalar-type operators via Laplace averages with connec- tions to the Koopman operator, arXiv preprint arXiv:1403.6559 (2014). [MM16a] A Mauroy and I Mezić, Global stability analysis using the eigenfunctions of the Koopman operator, IEEE Trans- actions on Automatic Control 61 (2016), no. 11, 3356–3369. [MM16b] R Mohr and I Mezić, Koopman principle eigenfunctions and linearization of diffeomorphisms, arXiv preprint arXiv:1611.01209 (2016). [MMM13] A Mauroy, I Mezić, and J Moehlis, Isostables, isochrons, and Koopman spectrum for the action–angle represen- tation of stable fixed point dynamics, Physica D: Nonlinear Phenomena 261 (2013), 19–30. [MWMM19] B Monga, D Wilson, T Matchen, and J Moehlis, Phase reduction and phase-based optimal control for biological systems: a tutorial, Biological cybernetics 113 (2019), no. 1-2, 11–46. [Nel70] E Nelson, Topics in dynamics I: Flows, Princeton University Press, 1970. [New17] S E Newhouse, On a differentiable linearization theorem of Philip Hartman, Contemporary Mathematics 692 (2017), 209–262. [PBK18] J L Proctor, S L Brunton, and J N Kutz, Generalizing Koopman theory to allow for inputs and control, SIAM Journal on Applied Dynamical Systems 17 (2018), no. 1, 909–930. [Poi79] H Poincaré, Sur les propriétés des fonctions définies par les équations aux différences partielles, 1879. [Rue89] D Ruelle, Elements of differentiable dynamics and bifurcation theory, Elsevier, 1989. [SB16] A Surana and A Banaszuk, Linear observer synthesis for nonlinear systems using Koopman operator framework, IFAC-PapersOnLine 49 (2016), no. 18, 716–723. [Sel85] G R Sell, Smooth linearization near a fixed point, American Journal of Mathematics (1985), 1035–1091. [Sie42] C L Siegel, Iteration of analytic functions, Annals of Mathematics (1942), 607–612. [Sie52] , Über die normalform analytischer differentialgleichungen in der nähe einer gleichgewichtslösung, Nachr. Akad. Wiss. Göttingen. Math.-Phys. Kl. Math.-Phys.-Chem. Abt. 1952 (1952), 21–30. [SKN17] S Shirasaka, W Kurebayashi, and H Nakao, Phase-amplitude reduction of transient dynamics far from attractors for limit-cycling systems, Chaos: An Interdisciplinary Journal of Nonlinear Science 27 (2017), no. 2, 023119. [Ste57] S Sternberg, Local contractions and a theorem of Poincaré, American Journal of Mathematics 79 (1957), no. 4, 809–824. [Sur16] A Surana, Koopman operator based observer synthesis for control-affine nonlinear systems, 2016 IEEE 55th Conference on Decision and Control (CDC), IEEE, 2016, pp. 6492–6499. [Tho14] E G F Thomas, A polarization identity for multilinear maps, Indagationes Mathematicae 25 (2014), no. 3, 468–474. [WE18] D Wilson and B Ermentrout, Greater accuracy and broadened applicability of phase reduction using isostable coordinates, Journal of mathematical biology 76 (2018), no. 1-2, 37–66. EXISTENCE AND UNIQUENESS OF GLOBAL KOOPMAN EIGENFUNCTIONS 35

[Wil67] F W Wilson, Jr., The structure of the level surfaces of a Lyapunov function, J. Differential Equations 3 (1967), 323–329. MR 0231409 [Wil69] , Smoothing derivatives of functions and applications, Trans. Amer. Math. Soc. 139 (1969), 413–428. MR 0251747 [WKR15] M O Williams, I G Kevrekidis, and C W Rowley, A data–driven approximation of the Koopman operator: Extending dynamic mode decomposition, Journal of Nonlinear Science 25 (2015), no. 6, 1307–1346. [WM14] D Wilson and J Moehlis, An energy-optimal approach for entrainment of uncertain circadian oscillators, Bio- physical Journal 107 (2014), no. 7, 1744–1755. [WM16a] , Isostable reduction of periodic orbits, Physical Review E 94 (2016), no. 5, 052213. [WM16b] , Isostable reduction with applications to time-dependent partial differential equations, Physical Review E 94 (2016), no. 1, 012211.