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

A parametrisation method for high-order phase reduction in coupled oscillator networks

Sören von der Gracht, Eddie Nijholt, Bob Rink Department of Mathematics, Paderborn University, Germany, [email protected]Department of Mathematics, Imperial College London, United Kingdom, [email protected]Department of Mathematics, Vrije Universiteit Amsterdam, The Netherlands, [email protected]
Abstract

We present a novel method for high-order phase reduction in networks of weakly coupled oscillators and, more generally, perturbations of reducible normally hyperbolic (quasi-)periodic tori. Our method works by computing an asymptotic expansion for an embedding of the perturbed invariant torus, as well as for the reduced phase dynamics in local coordinates. Both can be determined to arbitrary degrees of accuracy, and we show that the phase dynamics may directly be obtained in normal form. We apply the method to predict remote synchronisation in a chain of coupled Stuart-Landau oscillators.

1 Introduction

Many systems in science and engineering consist of coupled periodic processes. Examples vary from the motion of the planets, to the synchronous flashing of fireflies [5], and from the activity of neurons in the brain [11], to power grids and electronic circuits. The functioning and malfunctioning of these coupled systems is often determined by a form of collective behaviour of its constituents, perhaps most notably their synchronisation [1, 26]. For example, synchronisation of neurons plays a critical role in cognitive processes [23, 24].

In this paper, we consider the situation where the coupling between the periodic processes is weak, a case that is amenable to rigorous mathematical analysis. Specifically, we assume that the evolution of the processes can be modelled by a system of differential equations of the form

x˙j=Fj(xj)+εGj(x1,,xm)forxjMjandj=1,,m.\displaystyle\dot{x}_{j}=F_{j}(x_{j})+\varepsilon\,G_{j}(x_{1},\ldots,x_{m})\ \mbox{for}\ x_{j}\in\mathbb{R}^{M_{j}}\ \mbox{and}\ j=1,\ldots,m\,. (1.1)

The vector fields Fj:MjMjF_{j}:\mathbb{R}^{M_{j}}\to\mathbb{R}^{M_{j}} in (1.1) determine the dynamics of the uncoupled oscillators: we assume that each FjF_{j} possesses a hyperbolic TjT_{j}-periodic orbit Xj(t)X_{j}(t). In the uncoupled limit—when ε=0\varepsilon=0—equations (1.1) thus admit a normally hyperbolic periodic or quasi-periodic invariant torus 𝕋0M\mathbb{T}_{0}\subset\mathbb{R}^{M} (where M:=M1++MmM:=M_{1}+\ldots+M_{m}), consisting of the product of these periodic orbits. The functions GjG_{j} in (1.1) model the interaction between the oscillators, for example through a (hyper-)network. The interaction strength 0ε10\leq\varepsilon\ll 1 is assumed small, so that the unperturbed torus 𝕋0\mathbb{T}_{0} persists as an invariant manifold 𝕋ε\mathbb{T}_{\varepsilon} for (1.1), depending smoothly on ε\varepsilon, as is guaranteed by Fénichel’s theorem [8, 30].

The process of finding the equations of motion that govern the dynamics on the persisting torus 𝕋ε\mathbb{T}_{\varepsilon} is usually referred to as phase reduction [20, 21, 25]. Phase reduction has proved a powerful tool in the study of the synchronisation of coupled oscillators, especially because it often realises a considerable reduction of the dimension—and hence complexity—of the system. Various methods of phase reduction have been introduced over the past decades, the most well-known appearing perhaps in the work on chemical oscillations of Kuramoto [17]. We refer to [25] for an extensive overview of established phase reduction techniques, and refrain from providing an overview of these methods here.

Most existing phase reduction methods provide a first-order approximation of the dynamics on the persisting invariant torus in terms of the small coupling parameter. However, there are various instances where such a first-order approximation is insufficient, see [4, 16, 18, 27, 22], in particular when the first-order reduced dynamics is structurally unstable. For instance, it was observed in [16] that “remote synchronisation” [3] cannot be analysed with first-order methods. More accurate “high-order phase reduction” techniques (that go beyond the first-order approximation) have only been introduced very recently [4, 10, 18]. They have already been applied successfully, for example to predict remote synchronisation [16]. However, to the best of our knowledge, mathematically rigorous high-order phase reduction methods have only been derived in the special case that the unperturbed oscillators are either Stuart-Landau oscillators [10, 18] or deformations thereof [2, 4]. In that setting, phase reduction can be performed by computing an expansion of the phase-amplitude relation that defines the invariant torus. However, this procedure does not generalise to arbitrary systems of the form (1.1).

This paper presents a novel method for high-order phase reduction, that applies to general coupled oscillator systems of the form (1.1). Our method works by computing an expansion (in the small parameter ε\varepsilon) of an embedding

e:(/2π)mMe:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M}\,

of the persisting invariant torus 𝕋ε\mathbb{T_{\varepsilon}}. In addition, it computes an expansion of the dynamics on 𝕋ε\mathbb{T}_{\varepsilon} in local coordinates, in the form of a so-called “reduced phase vector field”

𝐟:(/2π)mm{\bf f}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m}

on the standard torus (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m}. We find these ee and 𝐟{\bf f} by solving a so-called “conjugacy equation”. Our method is thus inspired by the work of De la Llave et al. [13], who popularised the idea of finding invariant manifolds by solving conjugacy equations. In fact, this idea was used in [6] to design a quadratically convergent iterative scheme for finding normally hyperbolic invariant tori. However, in [6] these tori are required to carry Diophantine quasi-periodic motion, not only before but also after the perturbation.

The phase reduction method presented in this paper is more similar in nature to the parametrisation method developed in [19]. There the idea of parametrisation is used to calculate expansions of slow manifolds and their flows in geometric singular perturbation problems [30]. Just like the method in [19], the phase reduction method presented here yields asymptotic expansions to finite order, but it poses no restrictions on the nature of the dynamics on the invariant torus.

We now sketch the idea behind our method. Let us write 𝐅0{\bf F}_{0} for the vector field on M=M1××Mm\mathbb{R}^{M}=\mathbb{R}^{M_{1}}\times\ldots\times\mathbb{R}^{M_{m}} that governs the dynamics of the uncoupled oscillators in (1.1), that is,

𝐅0(x1,,xm):=(F1(x1),,Fm(xm)).\displaystyle{\bf F}_{0}(x_{1},\ldots,x_{m}):=(F_{1}(x_{1}),\ldots,F_{m}(x_{m}))\,. (1.2)

Our starting point is an embedding of the invariant torus 𝕋0\mathbb{T}_{0} for this 𝐅0{\bf F}_{0}. Recall our assumption that every FjF_{j} possesses a hyperbolic periodic orbit Xj(t)X_{j}(t) of minimal period Tj>0T_{j}>0. We denote the frequency of this orbit by ωj:=2πTj\omega_{j}:=\frac{2\pi}{T_{j}}. An obvious embedding of 𝕋0\mathbb{T}_{0} is the map e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} defined by

e0(ϕ)=e0(ϕ1,,ϕm):=(X1(ω11ϕ1),,Xm(ωm1ϕm)).\displaystyle e_{0}(\phi)=e_{0}(\phi_{1},\ldots,\phi_{m}):=\left(X_{1}\left(\omega_{1}^{-1}\phi_{1}\right),\ldots,X_{m}\left(\omega_{m}^{-1}\phi_{m}\right)\right)\,. (1.3)

In fact, this e0e_{0} sends the periodic or quasi-periodic solutions of the ODEs

ϕ˙=ω:=(ω1,,ωm)\dot{\phi}=\omega:=(\omega_{1},\ldots,\omega_{m})

on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to integral curves of 𝐅0{\bf F}_{0}. In other words—see also Lemma 2.1 below—it satisfies the conjugacy equation

e0ω=𝐅0e0.e_{0}^{\prime}\cdot\omega={\bf F}_{0}\circ e_{0}\,.

The idea is now that we search for an asymptotic approximation of an embedding of the persisting torus 𝕋ε\mathbb{T}_{\varepsilon} by solving a similar conjugacy equation. We do this by making a series expansion ansatz for such an embedding, of the form

e=e0+εe1+ε2e2+:(/2π)mM,e=e_{0}+\varepsilon e_{1}+\varepsilon^{2}e_{2}+\ldots:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M}\,,

as well as for a reduced phase vector field

𝐟=ω+ε𝐟1+ε2𝐟2+:(/2π)mm.{\bf f}=\omega+\varepsilon{\bf f}_{1}+\varepsilon^{2}{\bf f}_{2}+\ldots:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m}\,.

Indeed, writing 𝐅=𝐅𝟎+ε𝐅1:MM{\bf F}={\bf F_{0}}+\varepsilon{\bf F}_{1}:\mathbb{R}^{M}\to\mathbb{R}^{M}, with 𝐅0{\bf F}_{0} as above, and

𝐅1(x):=(G1(x),,Gm(x)){\bf F}_{1}(x):=(G_{1}(x),\ldots,G_{m}(x))\,

denoting the coupled part of (1.1), we have that ee maps integral curves of 𝐟{\bf f} to solutions of (1.1), exactly when the conjugacy equation

e𝐟=𝐅ee^{\prime}\cdot{\bf f}={\bf F}\circ e

holds. If this is the case, then 𝕋ε=e((/2π)m)\mathbb{T}_{\varepsilon}=e((\mathbb{R}/2\pi\mathbb{Z})^{m}) is the persisting invariant torus, whereas the vector field 𝐟{\bf f} on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} represents the dynamic on 𝕋ε\mathbb{T}_{\varepsilon} in local coordinates, that is, it determines the reduced phase dynamics.

We will see that the conjugacy equation for (e,𝐟)(e,{\bf f}) translates into a sequence of iterative equations for (e1,𝐟1),(e2,𝐟2),(e_{1},{\bf f}_{1}),(e_{2},{\bf f}_{2}),\ldots. We will show how to solve these iterative equations, which then allows us to compute the expansions for ee and 𝐟{\bf f} to any desired order in the small parameter. Because the embedding of the torus 𝕋ε\mathbb{T}_{\varepsilon} is not unique, neither are the solutions (ej,𝐟j)(e_{j},{\bf f}_{j}) to these iterative equations. We characterize the extent to which one is free to choose these solutions, and we show how this freedom can be exploited to obtain 𝐟j{\bf f}_{j} that are in normal form. This means that “nonresonant” terms have been removed from the reduced phase equations to high order.

A crucial requirement for the solvability of the iterative equations is that the torus 𝕋0\mathbb{T}_{0} is reducible. Reducibility is a property of the unperturbed dynamics normal to 𝕋0\mathbb{T}_{0}. We shall define it at the hand of an embedding of the so-called fast fibre bundle of 𝕋0\mathbb{T}_{0}. We call such an embedding a fast fibre map. The fast fibre map is an important ingredient of our method. An invariant torus for an uncoupled oscillator system is always reducible. We show in Section 5 how, in this case, the fast fibre map can be obtained from the Floquet decompositions of the fundamental matrix solutions of the periodic orbits Xj(t)X_{j}(t). We remark that by using fast fibre maps, we are able to avoid the use of isochrons [12] to characterise the dynamics normal to 𝕋0\mathbb{T}_{0}. Our parametrisation method is therefore not restricted to the case where the periodic orbits Xj(t)X_{j}(t) are stable limit cycles—it suffices if they are hyperbolic. We also stress that our method is not restricted to weakly coupled oscillator systems: it applies whenever the unperturbed embedded torus 𝕋0\mathbb{T}_{0} is quasi-periodic, normally hyperbolic and reducible.

This paper is organised as follows. In section 2 we discuss the conjugacy problem for (e,𝐟)(e,{\bf f}) in more detail, and derive the iterative equations for (ej,𝐟j)(e_{j},{\bf f}_{j}). In section 3 we introduce fast fibre maps and use them to define when an embedded (quasi-)periodic torus is reducible. In section 4 we explain how the fast fibre map can be used to solve the iterative equations for (ej,𝐟j)(e_{j},{\bf f}_{j}). We give formulas for the solutions, and discuss their properties. Section 5 shows how to compute the fast fibre map for a coupled oscillator system, treating the Stuart-Landau oscillator as an example. We finish with an application/illustration of our method in section 6, in which we prove that remote synchronisation occurs in a chain of weakly coupled Stuart-Landau oscillators.

2 An iterative scheme

We start this section with a proof of our earlier claim about the embedding e0e_{0}. In the formulation of Lemma 2.1 below, we use the notation

ωe0:=e0ω=dds|s=0e0(+sω)\displaystyle\partial_{\omega}e_{0}:=e_{0}^{\prime}\cdot\omega=\left.\frac{d}{ds}\right|_{s=0}\!\!\!\!\!\!\!e_{0}(\,\cdot+s\omega)\, (2.1)

for the (directional) derivative of e0e_{0} in the direction of the vector ωm\omega\in\mathbb{R}^{m}. Like e0e_{0} itself, ωe0\partial_{\omega}e_{0} is a smooth map from (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to M\mathbb{R}^{M}.

Lemma 2.1.

The embedding e0e_{0} defined in (1.3) satisfies the conjugacy equation

ωe0(=e0ω)=𝐅0e0.\partial_{\omega}e_{0}\ (=e_{0}^{\prime}\cdot\omega)={\bf F}_{0}\circ e_{0}\,.
Proof.

Recall from (1.3) that (e0)j(ϕ)=Xj(ωj1ϕj)(e_{0})_{j}(\phi)=X_{j}(\omega_{j}^{-1}\phi_{j}), where XjX_{j} is a hyperbolic periodic orbit of FjF_{j}. It follows that

(ωe0)j(ϕ)\displaystyle(\partial_{\omega}e_{0})_{j}(\phi) =dds|s=0(e0)j(ϕ+sω)=dds|s=0Xj(ωj1(ϕj+sωj))\displaystyle=\left.\frac{d}{ds}\right|_{s=0}\!\!\!(e_{0})_{j}(\phi+s\omega)=\left.\frac{d}{ds}\right|_{s=0}\!\!\!X_{j}(\omega_{j}^{-1}(\phi_{j}+s\omega_{j}))
=X˙j(ωj1ϕj)=Fj(Xj(ωj1ϕj))=(𝐅0)j((e0(ϕ)),\displaystyle=\dot{X}_{j}(\omega_{j}^{-1}\phi_{j})=F_{j}(X_{j}(\omega_{j}^{-1}\phi_{j}))=({\bf F}_{0})_{j}((e_{0}(\phi))\,,

because X˙j(t)=Fj(Xj(t))\dot{X}_{j}(t)=F_{j}(X_{j}(t)) for all tt\in\mathbb{R}. ∎

Lemma 2.1 implies that e0e_{0} sends integral curves of the constant vector field ω\omega on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to integral curves of the vector field 𝐅0{\bf F}_{0} given in (1.2). Because the integral curves of the ODEs ϕ˙=ω\dot{\phi}=\omega on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} are clearly either periodic or quasi-periodic, we call 𝕋0=e0((/2π)m)\mathbb{T}_{0}=e_{0}((\mathbb{R}/2\pi\mathbb{Z})^{m}) an embedded (quasi-)periodic torus.

At this point we temporarily abandon the setting of coupled oscillators and consider a general ODE x˙=𝐅0(x)\dot{x}={\bf F}_{0}(x) defined by a smooth vector field 𝐅0:MM{\bf F}_{0}:\mathbb{R}^{M}\to\mathbb{R}^{M}. That is, we do not assume that this ODE decouples into mutually independent ODEs. However, we will assume throughout this paper that 𝐅0{\bf F}_{0} possesses a normally hyperbolic periodic or quasi-periodic invariant torus 𝕋0\mathbb{T}_{0} which admits an embedding e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} that semi-conjugates the constant vector field ω\omega on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to 𝐅0{\bf F}_{0}. In other words, we assume that e0e_{0} and 𝐅0{\bf F}_{0} satisfy

ωe0=𝐅0e0,\displaystyle\partial_{\omega}e_{0}={\bf F}_{0}\circ e_{0}\,, (2.2)

just as in Lemma 2.1. We return to coupled oscillator systems in section 5.

We now study any smooth perturbation of 𝐅0{\bf F}_{0} of the form

𝐅=𝐅(x)=𝐅0(x)+ε𝐅1(x)+ε2𝐅2(x)+:MM.{\bf F}={\bf F}(x)={\bf F}_{0}(x)+\varepsilon\,{\bf F}_{1}(x)+\varepsilon^{2}\,{\bf F}_{2}(x)+\ldots:\ \mathbb{R}^{M}\to\mathbb{R}^{M}\,.

Fénichel’s theorem [8, 30] guarantees that, for 0ε10\leq\varepsilon\ll 1, the perturbed ODE x˙=𝐅(x)\dot{x}={\bf F}(x) admits an invariant torus 𝕋ε\mathbb{T}_{\varepsilon} close to 𝕋0\mathbb{T}_{0}, that depends smoothly on ε\varepsilon. Our strategy to find 𝕋ε\mathbb{T}_{\varepsilon} will be to search for an embedding e:(/2π)mMe:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} close to e0e_{0}, and a reduced vector field 𝐟:(/2π)mm{\bf f}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m} close to ω\omega satisfying the conjugacy equation

(e,𝐟):=e𝐟𝐅e=0.\displaystyle\mathfrak{C}(e,{\bf f}):=e^{\prime}\cdot{\bf f}-{\bf F}\circ e=0\,. (2.3)

Any solution (e,𝐟)(e,{\bf f}) to (2.3) indeed yields an embedded 𝐅{\bf F}-invariant torus 𝕋ε:=e((/2π)m)M\mathbb{T}_{\varepsilon}:=e((\mathbb{R}/2\pi\mathbb{Z})^{m})\subset\mathbb{R}^{M}, as we see from (2.3) that at any point x=e(ϕ)𝕋εx=e(\phi)\in\mathbb{T}_{\varepsilon} the vector 𝐅(x){\bf F}(x) lies in the image of the derivative e(ϕ)e^{\prime}(\phi), and is thus tangent to 𝕋ε\mathbb{T}_{\varepsilon}. Moreover, ee semi-conjugates 𝐟{\bf f} to 𝐅{\bf F}, that is, 𝐟{\bf f} is the restriction of 𝐅{\bf F} to 𝕋ε\mathbb{T}_{\varepsilon} represented in (or “pulled back to”) the local coordinate chart (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m}.

As explained in the introduction, we try to find solutions to (2.3) by making a series expansion ansatz

e=e0+εe1+ε2e2+and𝐟=ω+ε𝐟1+ε2𝐟2+e=e_{0}+\varepsilon e_{1}+\varepsilon^{2}e_{2}+\ldots\ \mbox{and}\ {\bf f}=\omega+\varepsilon{\bf f}_{1}+\varepsilon^{2}{\bf f}_{2}+\ldots

for e1,e2,:(/2π)mMe_{1},e_{2},\ldots:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} and 𝐟1,𝐟2,:(/2π)mm{\bf f}_{1},{\bf f}_{2},\ldots:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m}. Substitution of this ansatz in (2.3), and Taylor expansion to ε\varepsilon, yields the following list of recursive equations for the eje_{j} and 𝐟j{\bf f}_{j}:

(ω𝐅0e0)e1+e0𝐟1=𝐅1e0=:𝐆1(ω𝐅0e0)e2+e0𝐟2=𝐅2e0+(𝐅1e0)e1+12(𝐅0′′e0)(e1,e1)e1𝐟1=:𝐆2(ω𝐅0e0)ej+e0𝐟j==:𝐆j\displaystyle\begin{array}[]{ccc}(\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{1}+e_{0}^{\prime}\cdot{\bf f}_{1}=&{\bf F}_{1}\circ e_{0}&=:{\bf G}_{1}\\ (\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{2}+e_{0}^{\prime}\cdot{\bf f}_{2}=&{\bf F}_{2}\circ e_{0}+({\bf F}_{1}^{\prime}\circ e_{0})\cdot e_{1}\\ &+\frac{1}{2}({\bf F}_{0}^{\prime\prime}\circ e_{0})(e_{1},e_{1})-e_{1}^{\prime}\cdot{\bf f}_{1}&=:{\bf G}_{2}\\ \vdots&\vdots&\vdots\\ (\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{j}+e_{0}^{\prime}\cdot{\bf f}_{j}=&\ldots&=:{\bf G}_{j}\\ \vdots&\vdots&\vdots\end{array} (2.10)

Here, each 𝐆j:(/2π)mM{\bf G}_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} is an “inhomogeneous term” that can iteratively be determined and depends on 𝐅1,,𝐅j,𝐟1,,𝐟j1{\bf F}_{1},\ldots,{\bf F}_{j},{\bf f}_{1},\ldots,{\bf f}_{j-1} and e1,,ej1e_{1},\ldots,e_{j-1}. Concretely, 𝐆j{\bf G}_{j} is given by

𝐆j:=1j!djdεj|ε=0(𝐅0+ε𝐅1++εj𝐅j)(e0+εe1++εj1ej1)(e0+εe1+εj1ej1)(ω+ε𝐟1++εj1𝐟j1).\displaystyle\!\!\!\!{\bf G}_{j}\!:=\!\!\left.\frac{1}{j!}\frac{d^{j}}{d\varepsilon^{j}}\right|_{\varepsilon=0}\!\!\!\!\!\!\begin{array}[]{c}\!\!\!({\bf F}_{0}+\varepsilon{\bf F}_{1}+\ldots+\varepsilon^{j}{\bf F}_{j})(e_{0}+\varepsilon e_{1}+\ldots+\varepsilon^{j-1}e_{j-1})\\ -(e_{0}+\varepsilon e_{1}\ldots+\varepsilon^{j-1}e_{j-1})^{\prime}\cdot(\omega+\varepsilon{\bf f}_{1}+\ldots+\varepsilon^{j-1}{\bf f}_{j-1})\end{array}\!\!\!\!\!\!. (2.13)

Explicit formulas for 𝐆1{\bf G}_{1} and 𝐆2{\bf G}_{2} are given in (2.10). Note that equations (2.10) are all of the form

𝔠(ej,𝐟j)=𝐆jforj=1,2,,\displaystyle\mathfrak{c}(e_{j},{\bf f}_{j})={\bf G}_{j}\ \mbox{for}\ j=1,2,\ldots\,, (2.14)

in which

𝔠(ej,𝐟j):=(ω𝐅0e0)ej+e0𝐟j\displaystyle\mathfrak{c}(e_{j},{\bf f}_{j}):=(\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{j}+e_{0}^{\prime}\cdot{\bf f}_{j}\, (2.15)

is the linearisation of the operator \mathfrak{C} defined in (2.3) at the point (e,𝐟)=(e0,ω)(e,{\bf f})=(e_{0},\omega), where ε=0\varepsilon=0. This linearisation 𝔠\mathfrak{c} is not invertible, but we will see that 𝔠\mathfrak{c} is surjective under the assumption that 𝕋0\mathbb{T}_{0} is reducible. This implies that equations (2.10) can iteratively be solved.

Remark 1.

We think of \mathfrak{C} and 𝔠\mathfrak{c} as operators between function spaces. For example, for 𝐅0Cr+1(M,M),𝐅Cr(M,M){\bf F}_{0}\in C^{r+1}(\mathbb{R}^{M},\mathbb{R}^{M}),{\bf F}\in C^{r}(\mathbb{R}^{M},\mathbb{R}^{M}), and e0Cr+1((/2π)m,M)e_{0}\in C^{r+1}((\mathbb{R}/2\pi\mathbb{Z})^{m},\mathbb{R}^{M}),

,𝔠:Cr+1((/2π)m,M)×Cr((/2π)m,m)Cr((/2π)m,M).\mathfrak{C},\mathfrak{c}:C^{r+1}((\mathbb{R}/2\pi\mathbb{Z})^{m},\mathbb{R}^{M})\times C^{r}((\mathbb{R}/2\pi\mathbb{Z})^{m},\mathbb{R}^{m})\to C^{r}((\mathbb{R}/2\pi\mathbb{Z})^{m},\mathbb{R}^{M})\,.
Remark 2.

The solutions to equation (2.3) are not unique because an invariant torus can be embedded in many different ways. In fact, if e:(/2π)mMe:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} is an embedding of 𝕋ε\mathbb{T}_{\varepsilon} and Ψ:(/2π)m(/2π)m\Psi:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to(\mathbb{R}/2\pi\mathbb{Z})^{m} is any diffeomorphism of the standard torus, then also eΨe\circ\Psi is an embedding of 𝕋ε\mathbb{T}_{\varepsilon}. The operator \mathfrak{C} defined in (2.3) is thus equivariant under the group of diffeomorphisms of (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m}. As a consequence, solutions of (2.14) are not unique either.

Remark 3.

For the interested reader we provide additional details on Remark 2. Let us denote by Ψ𝐟\Psi^{*}{\bf{f}} the pullback of the vector field 𝐟{\bf{f}} by Ψ\Psi defined by the formula (Ψ𝐟)(ϕ):=(Ψ(ϕ))1𝐟(Ψ(ϕ))(\Psi^{*}{\bf{f}})(\phi):=(\Psi^{\prime}(\phi))^{-1}\cdot{\bf f}(\Psi(\phi)) for all ϕ(/2π)m\phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m}. We claim that

(eΨ,Ψ𝐟)=(e,f)Ψ.\displaystyle\mathfrak{C}(e\circ\Psi,\Psi^{*}{\bf{f}})=\mathfrak{C}(e,f)\circ\Psi\,. (2.16)

This follows from a straightforward calculation. Indeed,

(eΨ,Ψ𝐟)(ϕ)\displaystyle\mathfrak{C}(e\circ\Psi,\Psi^{*}{\bf{f}})(\phi) =e(Ψ(ϕ))Ψ(ϕ)(Ψ(ϕ))1𝐟(Ψ(ϕ))𝐅((eΨ)(ϕ))\displaystyle=e^{\prime}(\Psi(\phi))\cdot\Psi^{\prime}(\phi)\cdot(\Psi^{\prime}(\phi))^{-1}\cdot{\bf f}(\Psi(\phi))-{\bf F}((e\circ\Psi)(\phi))
=e(Ψ(ϕ))𝐟(Ψ(ϕ))(𝐅e)(Ψ(ϕ))=(e,f)(Ψ(ϕ)).\displaystyle=e^{\prime}(\Psi(\phi))\cdot{\bf f}(\Psi(\phi))-({\bf F}\circ e)(\Psi(\phi))=\mathfrak{C}(e,f)(\Psi(\phi))\,.

As we may view vector fields as infinitesimal diffeomorphisms, this allows us to find many elements in the kernel of 𝔠\mathfrak{c}. Namely, if XX is any vector field on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} with corresponding flow φt\varphi_{t}, then

ddt|t=0(e0φt,φtω)=(e0X,[X,ω])ker𝔠.\displaystyle\left.\frac{d}{dt}\right|_{t=0}(e_{0}\circ\varphi_{t},\varphi_{t}^{*}\omega)=(e_{0}^{\prime}\cdot X,[X,\omega])\in\ker\mathfrak{c}. (2.17)

Here [X,ω]=Xω=ωX[X,\omega]=-X^{\prime}\cdot\omega=-\partial_{\omega}X denotes the Lie bracket between XX and ω\omega.

Formula (2.17) may also be verified directly. Differentiating the identity

(e0,ω)(ϕ)=e0(ϕ)ω(𝐅0e0)(ϕ)=0\displaystyle\mathfrak{C}(e_{0},\omega)(\phi)=e_{0}^{\prime}(\phi)\cdot\omega-({\bf F}_{0}\circ e_{0})(\phi)=0 (2.18)

at any ϕ\phi, in the direction of any vector uu, we first of all find that

e0′′(ϕ)(ω,u)(𝐅0e0)(ϕ)e0(ϕ)u=0.\displaystyle e_{0}^{\prime\prime}(\phi)(\omega,u)-({\bf F}_{0}^{\prime}\circ e_{0})(\phi)\cdot e_{0}^{\prime}(\phi)\cdot u=0\,. (2.19)

From this we see that indeed

𝔠(e0X,[X,ω])\displaystyle\mathfrak{c}(e_{0}^{\prime}\cdot X,[X,\omega]) =(ω𝐅0e0)e0Xe0ωX\displaystyle=(\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{0}^{\prime}\cdot X-e^{\prime}_{0}\cdot\partial_{\omega}X
=e0′′(ω,X)+e0ωX(𝐅0e0)e0Xe0ωX\displaystyle=e_{0}^{\prime\prime}(\omega,X)+e_{0}^{\prime}\cdot\partial_{\omega}X-({\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{0}^{\prime}\cdot X-e^{\prime}_{0}\cdot\partial_{\omega}X
=e0′′(ω,X)(𝐅0e0)e0X=0,\displaystyle=e_{0}^{\prime\prime}(\omega,X)-({\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{0}^{\prime}\cdot X=0\,,

where the last step follows from equation (2.19).

3 Reducibility and the fast fibre map

As was indicated in Remarks 2 and 3, the solutions to the iterative equations 𝔠(ej,𝐟j)=𝐆j\mathfrak{c}(e_{j},{\bf f}_{j})={\bf G}_{j} are not unique. However, we show in section 4 that solutions can be found if we assume that the unperturbed torus 𝕋0\mathbb{T}_{0} is reducible. We define this concept by means of a parametrisation of the linearised dynamics of 𝐅0{\bf F}_{0} normal to 𝕋0\mathbb{T}_{0}. But we start with the observation that the linearised dynamics tangent to 𝕋0\mathbb{T}_{0} is trivial. Recall that if e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} is an embedding of 𝕋0M\mathbb{T}_{0}\subset\mathbb{R}^{M}, then the tangent map 𝐓e0:(/2π)m×mM×M{\bf T}e_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\times\mathbb{R}^{m}\to\mathbb{R}^{M}\times\mathbb{R}^{M} defined by

𝐓e0(ϕ,u)=(e0(ϕ),e0(ϕ)u)\displaystyle{\bf T}e_{0}(\phi,u)=(e_{0}(\phi),e_{0}^{\prime}(\phi)\cdot u) (3.1)

is an embedding as well. Its image is the tangent bundle 𝐓𝕋0M×M{\bf T}\mathbb{T}_{0}\subset\mathbb{R}^{M}\times\mathbb{R}^{M}.

Lemma 3.1.

Assume that the embedding e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} semi-conjugates the constant vector field ωm\omega\in\mathbb{R}^{m} on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to the vector field 𝐅0{\bf F}_{0} on M\mathbb{R}^{M}. Then 𝐓e0{\bf T}e_{0} sends solution curves of the system of ODEs

ϕ˙=ω,u˙=0on(/2π)m×m\dot{\phi}=\omega\,,\ \dot{u}=0\ \mbox{on}\ (\mathbb{R}/2\pi\mathbb{Z})^{m}\times\mathbb{R}^{m}

to integral curves of the tangent vector field 𝐓𝐅0{\bf T}{\bf F}_{0} on M×M\mathbb{R}^{M}\times\mathbb{R}^{M} defined by

𝐓𝐅0(x,v):=(𝐅0(x),𝐅0(x)v).\displaystyle{\bf T}{\bf F}_{0}(x,v):=({\bf F}_{0}(x),{\bf F}_{0}^{\prime}(x)\cdot v)\,. (3.2)
Proof.

Our assumption simply means that ωe0=𝐅0e0\partial_{\omega}e_{0}={\bf F}_{0}\circ e_{0}. As we already observed in (2.19), differentiation of this identity at a point ϕ(/2π)m\phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m} in the direction of a vector umu\in\mathbb{R}^{m} yields that

e0′′(ϕ)(u,ω)=𝐅0(e0(ϕ))e0(ϕ)u.\displaystyle e_{0}^{\prime\prime}(\phi)(u,\omega)={\bf F}_{0}^{\prime}(e_{0}(\phi))\cdot e_{0}^{\prime}(\phi)\cdot u\,.

From this it follows that

(𝐓e0)(ϕ,u)(ω,0)=\displaystyle({\bf T}e_{0})^{\prime}(\phi,u)\cdot(\omega,0)= dds|s=0𝐓e0(ϕ+sω,u)\displaystyle\left.\frac{d}{ds}\right|_{s=0}\!\!\!\!\!{\bf T}e_{0}(\phi+s\omega,u)
=\displaystyle= dds|s=0(e0(ϕ+sω),e0(ϕ+sω)u)\displaystyle\left.\frac{d}{ds}\right|_{s=0}\!\!\!\!\!\left(e_{0}(\phi+s\omega),e_{0}^{\prime}(\phi+s\omega)\cdot u\right)
=\displaystyle= ((ωe0)(ϕ),e0′′(ϕ)(u,ω))\displaystyle\left((\partial_{\omega}e_{0})(\phi),e_{0}^{\prime\prime}(\phi)(u,\omega)\right)
=\displaystyle= (𝐅0(e0(ϕ)),𝐅0(e0(ϕ))e0(ϕ)u)=𝐓𝐅0(𝐓e0(ϕ,u)).\displaystyle\left({\bf F}_{0}(e_{0}(\phi)),{\bf F}_{0}^{\prime}(e_{0}(\phi))\cdot e_{0}^{\prime}(\phi)\cdot u\right)={\bf T}{\bf F}_{0}({\bf T}e_{0}(\phi,u))\,.

In the last equality we used Definitions (3.1) and (3.2). ∎

Lemma 3.1 shows that 𝐓e0{\bf T}e_{0} trivialises the linearised dynamics of 𝐅0{\bf F}_{0} in the direction tangent to 𝕋0\mathbb{T}_{0}. In what follows, we assume that something similar happens in the direction normal to 𝕋0\mathbb{T}_{0}, that is, we assume that 𝕋0\mathbb{T}_{0} is reducible. We define this concept now.

Definition 3.2.

Assume that the embedding e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} semi-conjugates the constant vector field ωm\omega\in\mathbb{R}^{m} on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} to the vector field 𝐅0{\bf F}_{0} on M\mathbb{R}^{M}. We say that the (quasi-)periodic invariant torus 𝕋0=e0((/2π)m)\mathbb{T}_{0}=e_{0}((\mathbb{R}/2\pi\mathbb{Z})^{m}) is reducible if there is a map 𝐍e0:(/2π)m×MmM×M{\bf N}e_{0}\!:\!(\mathbb{R}/2\pi\mathbb{Z})^{m}\!\times\!\mathbb{R}^{M-m}\to\mathbb{R}^{M}\!\times\!\mathbb{R}^{M} of the form

𝐍e0(ϕ,u):=(e0(ϕ),N(ϕ)u),\displaystyle{\bf N}e_{0}(\phi,u):=(e_{0}(\phi),N(\phi)\cdot u)\,, (3.3)

for a smooth family of linear maps

N:(/2π)m(Mm,M),N:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathcal{L}(\mathbb{R}^{M-m},\mathbb{R}^{M})\,,

with the following two properties:

  • i)

    𝐍e0{\bf N}e_{0} is transverse to 𝐓e0{\bf T}e_{0}. By this we mean that

    M=ime0(ϕ)imN(ϕ)for everyϕ(/2π)m.\displaystyle\mathbb{R}^{M}={\rm im}\,e_{0}^{\prime}(\phi)\oplus{\rm im}\,N(\phi)\ \mbox{for every}\ \phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m}\,. (3.4)

    In particular, every N(ϕ)N(\phi) is injective.

  • ii)

    There is a linear map L:MmMmL:\mathbb{R}^{M-m}\to\mathbb{R}^{M-m} such that 𝐍e0{\bf N}e_{0} sends solution curves of the system of ODEs

    ϕ˙=ω,u˙=Ludefined on(/2π)m×Mm\dot{\phi}=\omega\,,\,\dot{u}=L\cdot u\ \mbox{defined on}\ (\mathbb{R}/2\pi\mathbb{Z})^{m}\times\mathbb{R}^{M-m}

    to integral curves of the tangent vector field 𝐓𝐅0{\bf T}{\bf F}_{0} on M×M\mathbb{R}^{M}\times\mathbb{R}^{M}.

When 𝕋0\mathbb{T}_{0} is reducible, the matrix LL is called a Floquet matrix for 𝕋0\mathbb{T}_{0}, and its eigenvalues the Floquet exponents of 𝕋0\mathbb{T}_{0}.

If LL is hyperbolic (no Floquet exponents lie on the imaginary axis) then 𝕋0\mathbb{T}_{0} is normally hyperbolic, and we call 𝐍e0{\bf N}e_{0} a fast fibre map for 𝕋0\mathbb{T}_{0}. Its image

𝐍𝕋0:=𝐍e0((/2π)m×Mm)M×M{\bf N}\mathbb{T}_{0}:={\bf N}e_{0}((\mathbb{R}/2\pi\mathbb{Z})^{m}\times\mathbb{R}^{M-m})\subset\mathbb{R}^{M}\times\mathbb{R}^{M}\,

is then called the fast fibre bundle of 𝕋0\mathbb{T}_{0}.

We note that the map 𝐍e0{\bf N}e_{0} appearing in Definition 3.2 is an embedding because e0e_{0} is an embedding and the linear maps N(ϕ)N(\phi) are all injective. Therefore its image 𝐍𝕋0{\bf N}\mathbb{T}_{0} is a smooth MM-dimensional manifold. Condition i) ensures that 𝐍𝕋0{\bf N}\mathbb{T}_{0} is in fact a normal bundle for 𝕋0\mathbb{T}_{0}.

We finish this section with an alternative characterisation of property 𝑖𝑖)\it ii) in Definition 3.2.

Lemma 3.3.

Assume that the embedding e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} semi-conjugates the constant vector field ω\omega to the vector field 𝐅0{\bf F}_{0}. Let L:MmMmL:\mathbb{R}^{M-m}\to\mathbb{R}^{M-m} be a linear map, and let 𝐍e0{\bf N}e_{0} be a map of the form (3.3) for a smooth family of linear maps N:(/2π)m(Mm,M)N:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathcal{L}(\mathbb{R}^{M-m},\mathbb{R}^{M}). The following are equivalent:

  • i)

    𝐍e0{\bf N}e_{0} sends solution curves of the system of ODEs

    ϕ˙=ω,u˙=Ludefined on(/2π)m×Mm\dot{\phi}=\omega\,,\,\dot{u}=L\cdot u\ \mbox{defined on}\ (\mathbb{R}/2\pi\mathbb{Z})^{m}\times\mathbb{R}^{M-m}

    to integral curves of the tangent vector field 𝐓𝐅0{\bf T}{\bf F}_{0} on M×M\mathbb{R}^{M}\times\mathbb{R}^{M};

  • ii)

    N=N(ϕ)N=N(\phi) satisfies the partial differential equation

    ωN+NL=(𝐅0e0)Non(/2π)m.\displaystyle\partial_{\omega}N+N\!\cdot\!L=({\bf F}_{0}^{\prime}\circ e_{0})\!\cdot\!N\ \mbox{on}\ (\mathbb{R}/2\pi\mathbb{Z})^{m}\,. (3.5)
Proof.

It holds that

(𝐍e0)(ϕ,u)(ω,Lu)\displaystyle({\bf N}e_{0})^{\prime}(\phi,u)\cdot(\omega,L\cdot u) =dds|s=0(e0(ϕ+sω),N(ϕ+sω)(u+sLu))\displaystyle=\left.\frac{d}{ds}\right|_{s=0}\!\!\!\!\!\left(e_{0}(\phi+s\omega),N(\phi+s\omega)\cdot(u+sL\cdot u)\right)
=((ωe0)(ϕ),ωN(ϕ)u+N(ϕ)Lu).\displaystyle=((\partial_{\omega}e_{0})(\phi),\partial_{\omega}N(\phi)\cdot u+N(\phi)\cdot L\cdot u)\,.

At the same time,

𝐓𝐅0(𝐍e0(ϕ,u))=(𝐅0(e0(ϕ)),𝐅0(e0(ϕ))N(ϕ)u).{\bf T}{\bf F}_{0}({\bf N}e_{0}(\phi,u))=({\bf F}_{0}(e_{0}(\phi)),{\bf F}_{0}^{\prime}(e_{0}(\phi))\cdot N(\phi)\cdot u)\,.

It holds that ωe0=𝐅0e0\partial_{\omega}e_{0}={\bf F}_{0}\circ e_{0} by assumption, so the first components of these two expressions are equal. The conclusion of the lemma therefore follows from comparing the second components. ∎

Remark 4.

Reducibility of a (quasi-)periodic invariant torus of an arbitrary vector field 𝐅0{\bf F}_{0} can only be quaranteed under strong conditions, e.g., that 𝐅0{\bf F}_{0} is Hamiltonian [7], or that the frequency vector ω\omega satisfies certain Diophantine inequalities [14]. We do not assume such conditions here. Even the question whether reducibility is preserved under perturbation is subtle [15].

However, hyperbolic periodic orbits (which are one-dimensional normally hyperbolic invariant tori) are always reducible (at least if we allow the matrix LL to be complex, see Section 5). This relatively well-known fact is a consequence of Floquet’s theorem [9], as we show in Theorem 5.1. The (quasi-)periodic torus occurring in an uncoupled oscillator system such as (1.1) is a product of hyperbolic periodic orbits, and is therefore reducible as well, see Lemma 5.3.

4 Solving the iterative equations

We now return to solving the iterative equations (2.10), assuming from here on out that 𝕋0\mathbb{T}_{0} is an embedded (quasi-)periodic reducible and normally hyperbolic invariant torus for 𝐅0{\bf F}_{0}. The main result of this section can be summarised (at this point still somewhat imprecisely) as follows.

Theorem 4.1.

Assume that 𝕋0=e0((/2π)m)M\mathbb{T}_{0}=e_{0}((\mathbb{R}/2\pi\mathbb{Z})^{m})\subset\mathbb{R}^{M} is a smooth embedded (quasi-)periodic reducible normally hyperbolic invariant torus for 𝐅0{\bf F}_{0}. Then

  • i)

    there are smooth solutions (ej,𝐟j)(e_{j},{\bf f}_{j}) to the iterative equations 𝔠(ej,𝐟j)=𝐆j\mathfrak{c}(e_{j},{\bf f}_{j})={\bf G}_{j} for every jj\in\mathbb{N}, for which we provide explicit formulas in this section;

  • ii)

    the component of each eje_{j} tangential to 𝕋0\mathbb{T}_{0} can be chosen freely, but every such choice for e1,,ej1e_{1},\ldots,e_{j-1} uniquely determines the component of eje_{j} normal to 𝕋0\mathbb{T}_{0} (see Theorem 4.4);

  • iii)

    the tangential component of eje_{j} can be chosen in such a way that 𝐟j{\bf f}_{j} is in “normal form” to arbitrarily high order in its Fourier expansion. We say that 𝐟j{\bf f}_{j} is in normal form if it is a sum of “resonant terms” only (see Corollary 4.6).

The precise meaning of the statements in this theorem will be made clear below. Theorem 4.1 follows directly from the results presented in this section.

To prove the theorem, recall that (because 𝕋0\mathbb{T}_{0} is reducible) we have at our disposal a fast fibre map 𝐍e0{\bf N}e_{0} for 𝕋0\mathbb{T}_{0}, defined by a family of injective matrices N=N(ϕ)N=N(\phi) that satisfies M=ime0(ϕ)imN(ϕ)\mathbb{R}^{M}={\rm im}\,e_{0}^{\prime}(\phi)\oplus{\rm im}\,N(\phi) for every ϕ(/2π)m\phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m}. This enables us to make the ansatz

ej(ϕ)M=e0(ϕ)(m,M)𝐠j(ϕ)m+N(ϕ)(Mm,M)𝐡j(ϕ)Mm,\displaystyle\underbrace{e_{j}(\phi)}_{\footnotesize\begin{array}[]{c}\in\\ \mathbb{R}^{M}\end{array}}\ \ =\!\!\underbrace{e_{0}^{\prime}(\phi)}_{\footnotesize\begin{array}[]{c}\in\\ \mathcal{L}(\mathbb{R}^{m},\mathbb{R}^{M})\end{array}}\!\!\!\!\!\cdot\ \underbrace{{\bf g}_{j}(\phi)}_{\footnotesize\begin{array}[]{c}\in\\ \mathbb{R}^{m}\end{array}}\ \ +\!\!\!\!\!\underbrace{N(\phi)}_{\footnotesize\begin{array}[]{c}\in\\ \mathcal{L}(\mathbb{R}^{M-m},\mathbb{R}^{M})\end{array}}\!\!\!\!\!\!\!\cdot\ \underbrace{{\bf h}_{j}(\phi)}_{\footnotesize\begin{array}[]{c}\in\\ \mathbb{R}^{M-m}\end{array}}\,, (4.11)

for (unknown) smooth functions 𝐠j:(/2π)mm{\bf g}_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m} and 𝐡j:(/2π)mMm{\bf h}_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M-m}. This ansatz decomposes eje_{j} into components in the direction of the tangent bundle 𝐓e0{\bf T}e_{0} and the fast fibre bundle 𝐍e0{\bf N}e_{0}.

Lemma 4.2.

The ansatz (4.11) transforms equation (2.14) into

𝔠(ej,𝐟j)=e0(ω𝐠j+𝐟j)+N(ωL)(𝐡j)=𝐆j.\displaystyle\mathfrak{c}(e_{j},{\bf f}_{j})=e_{0}^{\prime}\cdot\left(\partial_{\omega}{\bf g}_{j}+{\bf f}_{j}\right)+N\cdot(\partial_{\omega}-L)({\bf h}_{j})={\bf G}_{j}\,. (4.12)
Proof.

We use our definitions, and results derived above, to compute:

𝐆j=𝔠(ej,𝐟j)=\displaystyle{\bf G}_{j}=\mathfrak{c}(e_{j},{\bf f}_{j})= (ω𝐅0e0)ej+e0𝐟j\displaystyle\ (\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot e_{j}+e_{0}^{\prime}\cdot{\bf f}_{j}
=\displaystyle= (ω𝐅0e0)(e0𝐠j+N𝐡j)+e0𝐟j\displaystyle\ (\partial_{\omega}-{\bf F}_{0}^{\prime}\circ e_{0})\cdot\left(e_{0}^{\prime}\cdot{\bf g}_{j}+N\cdot{\bf h}_{j}\right)+e_{0}^{\prime}\cdot{\bf f}_{j}
=\displaystyle= e0′′(𝐠j,ω)+e0ω𝐠j+ωN𝐡j+Nω𝐡j\displaystyle\ e_{0}^{\prime\prime}({\bf g}_{j},\omega)+e_{0}^{\prime}\cdot\partial_{\omega}{\bf g}_{j}+\partial_{\omega}N\cdot{\bf h}_{j}+N\cdot\partial_{\omega}{\bf h}_{j}
(𝐅0.e0)e0𝐠j(𝐅𝟎e0)N𝐡j+e0𝐟j\displaystyle\ -({\bf F}_{0}^{\prime}.\circ e_{0})\cdot e_{0}^{\prime}\cdot{\bf g}_{j}-({\bf F_{0}}^{\prime}\circ e_{0})\cdot N\cdot{\bf h}_{j}+e_{0}^{\prime}\cdot{\bf f}_{j}
=\displaystyle= e0′′(𝐠j,ω)(𝐅0.e0)e0𝐠j=0+e0ω𝐠j+e0𝐟j\displaystyle\ \underbrace{e_{0}^{\prime\prime}({\bf g}_{j},\omega)-({\bf F}_{0}^{\prime}.\circ e_{0})\cdot e_{0}^{\prime}\cdot{\bf g}_{j}}_{=0}+e_{0}^{\prime}\cdot\partial_{\omega}{\bf g}_{j}+e_{0}^{\prime}\cdot{\bf f}_{j}
+Nω𝐡j+ωN𝐡j(𝐅𝟎e0)N𝐡j=NL𝐡j\displaystyle\ +N\cdot\partial_{\omega}{\bf h}_{j}+\underbrace{\partial_{\omega}N\cdot{\bf h}_{j}-({\bf F_{0}}^{\prime}\circ e_{0})\cdot N\cdot{\bf h}_{j}}_{=-N\cdot L\cdot{\bf h}_{j}}
=\displaystyle= e0(ω𝐠j+𝐟j)+N(ωL)𝐡j.\displaystyle\ e_{0}^{\prime}\cdot\left(\partial_{\omega}{\bf g}_{j}+{\bf f}_{j}\right)+N\cdot\left(\partial_{\omega}-L\right)\cdot{\bf h}_{j}\,.

We clarify these equalities below:

  • 1.

    The first equality is (2.14);

  • 2.

    In the second equality, we used (2.15);

  • 3.

    The third equality is our ansatz (4.11);

  • 4.

    The fourth equality follows from the product rule (applied twice);

  • 5.

    In the fifth equality, the terms in the sum were re-ordered;

  • 6.

    The final equality follows from (2.19) and (3.5).

This proves the lemma. ∎

Lemma 4.2 allows us to solve equation (4.12) by splitting it into a component along the tangent bundle 𝐓𝕋0{\bf T}\mathbb{T}_{0} and a component along the fast fibre bundle 𝐍𝕋0{\bf N}\mathbb{T}_{0} of 𝕋0\mathbb{T}_{0}. In what follows we denote by

π:(/2π)m(M,M)\pi:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathcal{L}(\mathbb{R}^{M},\mathbb{R}^{M})

the family of projections onto the tangent bundle along the fast fibre bundle. That is, each π(ϕ):MM\pi(\phi):\mathbb{R}^{M}\to\mathbb{R}^{M} is the unique projection that satisfies

π(ϕ)e0(ϕ)=e0(ϕ)andπ(ϕ)N(ϕ)=0.\pi(\phi)\cdot e_{0}^{\prime}(\phi)=e_{0}^{\prime}(\phi)\ \mbox{and}\ \pi(\phi)\cdot N(\phi)=0\,.

Proposition 4.7 below provides an explicit formula for π(ϕ)\pi(\phi). It is clear from this formula that π\pi depends smoothly on the base point ϕ(/2π)m\phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m}.

Applying π\pi and 1π1-\pi to (4.12) produces, respectively,

e0(ω𝐠j+𝐟j)=π𝐆j,\displaystyle e_{0}^{\prime}\cdot(\partial_{\omega}{\bf g}_{j}+{\bf f}_{j})=\pi\cdot{\bf G}_{j}\,,
N(ωL)(𝐡j)=(1π)𝐆j.\displaystyle N\cdot(\partial_{\omega}-L)({\bf h}_{j})=(1-\pi)\cdot{\bf G}_{j}\,.

Because e0(ϕ)e_{0}^{\prime}(\phi) and N(ϕ)N(\phi) are injective, these equations are equivalent to

ω𝐠j+𝐟j=(e0)+π𝐆j=:Uj,(ωL)(𝐡j)=N+(1π)𝐆j=:Vj.\displaystyle\begin{array}[]{rll}\partial_{\omega}{\bf g}_{j}+{\bf f}_{j}=&(e_{0}^{\prime})^{+}\cdot\pi\cdot{\bf G}_{j}&=:U_{j}\,,\\ (\partial_{\omega}-L)({\bf h}_{j})=&N^{+}\cdot(1-\pi)\cdot{\bf G}_{j}&=:V_{j}\,.\end{array} (4.15)

Here, A+:=(ATA)1ATA^{+}:=(A^{T}A)^{-1}A^{T} denotes the Moore-Penrose pseudo-inverse, which is well-defined for an injective linear map AA. Clearly, (e0)+(e_{0}^{\prime})^{+} and N+N^{+} depend smoothly on ϕ(/2π)m\phi\in(\mathbb{R}/2\pi\mathbb{Z})^{m}. We give these equations a special name.

Definition 4.3.

We call the first equation in (4.15),

ω𝐠j+𝐟j=Uj,\displaystyle\partial_{\omega}{\bf g}_{j}+{\bf f}_{j}=U_{j}\,, (4.16)

the jj-th tangential homological equation. The second equation in (4.15),

(ωL)(𝐡j)=Vj,\displaystyle(\partial_{\omega}-L)({\bf h}_{j})=V_{j}\,, (4.17)

is called the jj-th normal homological equation.

Remark 5.

To recap, we note that (4.16) and (4.17) are inhomogeneous linear equations for the three unknown smooth functions 𝐟j,𝐠j,𝐡j{\bf f}_{j},{\bf g}_{j},{\bf h}_{j} and with the inhomogeneous right hand sides Uj,VjU_{j},V_{j}. The domains and co-domains of these functions are given by

𝐟j,𝐠j,Uj:(/2π)mmand𝐡j,Vj:(/2π)mMm.\displaystyle{\bf f}_{j},{\bf g}_{j},U_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m}\ \mbox{and}\ {\bf h}_{j},V_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M-m}\,.

The following theorem shows how the homological equations can be solved. Explicit expressions for the Fourier series of the solutions are given in formulas (4.18) and (4.21), that appear in the proof of the theorem.

Theorem 4.4.

For any smooth functions 𝐠j,Uj:(/2π)mm{\bf g}_{j},U_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m} and Vj:(/2π)mMmV_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M-m}, there are unique smooth functions 𝐟j:(/2π)mm{\bf f}_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m} and 𝐡j:(/2π)mMm{\bf h}_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M-m} that solve (4.16) and (4.17).

Proof.

The tangential homological equation (4.16) can be rewritten as

𝐟j=Ujω𝐠j.{\bf f}_{j}=U_{j}-\partial_{\omega}{\bf g}_{j}\,.

This shows that for any smooth 𝐠j{\bf g}_{j} and UjU_{j} there exists a unique solution 𝐟j{\bf f}_{j}. However, in view of Corollary 4.6 below, we would also like a formula for the solution of the tangential homological equation in the form of a Fourier series. To this end, we expand UjU_{j} and 𝐠j{\bf g}_{j} in Fourier series as

Uj(ϕ)=kmUj,keik,ϕand𝐠j(ϕ)=kmgj,keik,ϕ.U_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}U_{j,k}e^{i\langle k,\phi\rangle}\ \mbox{and}\ {\bf g}_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}g_{j,k}e^{i\langle k,\phi\rangle}\,.

We use the notation

k,ϕ:=k1ϕ1++kmϕm\langle k,\phi\rangle:=k_{1}\phi_{1}+\ldots+k_{m}\phi_{m}\,

for what is often called the kk-th combination angle. Note that the Fourier coefficients Uj,k,gj,kmU_{j,k},g_{j,k}\in\mathbb{C}^{m} are complex vectors satisfying Uj,k=U¯j,kU_{j,-k}=\overline{U}_{j,k} and gj,k=g¯j,kg_{j,-k}=\overline{g}_{j,k}, because UjU_{j} and 𝐠j{\bf g}_{j} are real-valued. We similarly expand 𝐟j{\bf f}_{j} in a Fourier series by making the solution ansatz

𝐟j(ϕ)=kmfj,keik,ϕ,{\bf f}_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}f_{j,k}e^{i\langle k,\phi\rangle},

with fj,kmf_{j,k}\in\mathbb{C}^{m}. In terms of these Fourier series, equation (4.16) becomes

km(iω,kgj,k+fj,k)eik,ϕ=kmUj,keik,ϕ,\sum_{k\in\mathbb{Z}^{m}}(i\langle\omega,k\rangle g_{j,k}+f_{j,k})e^{i\langle k,\phi\rangle}=\sum_{k\in\mathbb{Z}^{m}}U_{j,k}e^{i\langle k,\phi\rangle}\,,

or, equivalently,

iω,kgj,k+fj,k=Uj,kfor allkm.i\langle\omega,k\rangle g_{j,k}+f_{j,k}=U_{j,k}\ \mbox{for all}\ k\in\mathbb{Z}^{m}\,.

This shows that for any choice of Fourier coefficients Uj,kU_{j,k} for UjU_{j} and gj,kg_{j,k} for 𝐠j{\bf g}_{j} there are unique Fourier coefficients fj,kf_{j,k} for the solution 𝐟j{\bf f}_{j} to the tangential homological equation. These coefficients are given by

fj,k=Uj,kiω,kgj,kfor allkm.\displaystyle f_{j,k}=U_{j,k}-i\langle\omega,k\rangle g_{j,k}\ \mbox{for all}\ k\in\mathbb{Z}^{m}\,. (4.18)

It is clear from this equation that fj,k=f¯j,kf_{j,-k}=\overline{f}_{j,k} so that 𝐟j{\bf f}_{j} is real-valued.

We proceed to solve the normal homological equation (4.17). We again use Fourier series, and thus we expand 𝐡j{\bf h}_{j} and VjV_{j} as

𝐡j(ϕ)=kmhj,keik,ϕandVj(ϕ)=kmVj,keik,ϕ,\displaystyle{\bf h}_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}h_{j,k}e^{i\langle k,\phi\rangle}\ \mbox{and}\ V_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}V_{j,k}e^{i\langle k,\phi\rangle}\,, (4.19)

for hj,k,Vj,kMmh_{j,k},V_{j,k}\in\mathbb{C}^{M-m} satisfying Vj,k=V¯j,kV_{j,-k}=\overline{V}_{j,k}. Substitution of (4.19) into (4.17) produces

km(iω,kL)hj,keik,ϕ=kmVj,keik,ϕ,\sum_{k\in\mathbb{Z}^{m}}(i\langle\omega,k\rangle-L){h}_{j,k}e^{i\langle k,\phi\rangle}=\sum_{k\in\mathbb{Z}^{m}}V_{j,k}e^{i\langle k,\phi\rangle}\,,

so that we obtain the equations

(iω,kL)hj,k=Vj,kfor allkm.\displaystyle(i\langle\omega,k\rangle-L)\,{h}_{j,k}=V_{j,k}\ \mbox{for all}\ k\in\mathbb{Z}^{m}\,. (4.20)

Because LL has no eigenvalues on the imaginary axis, the matrix iω,kLi\langle\omega,k\rangle-L is invertible. Each of the equations in (4.20) therefore possesses a unique solution, which is given by

hj,k=(iω,kL)1Vj,k.{h}_{j,k}=(i\langle\omega,k\rangle-L)^{-1}V_{j,k}\,. (4.21)

Because the matrix LL is real, it follows that hj,k=h¯j,kh_{j,-k}=\overline{h}_{j,k}. This proves the theorem. ∎

Remark 6.

Formulas (4.18) and (4.21) allow us to estimate the smoothness of the solutions 𝐟j{\bf f}_{j} and 𝐡j{\bf h}_{j} to equations (4.16), (4.17) in terms of the smoothness of 𝐠j,Uj{\bf g}_{j},U_{j} and VjV_{j}. To see this, let 𝐀:(/2π)mp{\bf A}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{C}^{p} be a function with Fourier series

𝐀(ϕ)=kmAkeik,ϕ.{\bf A}(\phi)=\sum_{k\in\mathbb{Z}^{m}}A_{k}e^{i\langle k,\phi\rangle}\,.

For kmk\in\mathbb{Z}^{m}, define |k|:=(|k1|2++|km|2)12|k|:=\left(|k_{1}|^{2}+\ldots+|k_{m}|^{2}\right)^{\frac{1}{2}}, and let W|k|>0W_{|k|}\in\mathbb{R}_{>0} be weights satisfying W|k|W_{|k|}\to\infty as |k||k|\to\infty. When ||||||\cdot|| is any norm on p\mathbb{C}^{p}, then

𝐀W:=(kmAk2W|k|2)12||{\bf A}||_{W}:=\left(\sum_{k\in\mathbb{Z}^{m}}||A_{k}||^{2}W_{|k|}^{2}\right)^{\frac{1}{2}}

defines a norm of 𝐀{\bf A} that measures the growth of its Fourier coefficients. For example, when W|k|=(1+|k|2)s/2W_{|k|}=(1+|k|^{2})^{s/2} for some s>0s>0, then it is a Sobolev norm. It follows directly from (4.18) that 𝐟jWUjW+ω𝐠jW||{\bf f}_{j}||_{W}\leq||U_{j}||_{W}+||\partial_{\omega}{\bf g}_{j}||_{W}, which shows that 𝐟j{\bf f}_{j} is at least as smooth as UjU_{j} and ω𝐠j\partial_{\omega}{\bf g}_{j}.

To find a similar bound for 𝐡jW||{\bf h}_{j}||_{W}, note that the hyperbolicity of LL implies that the function λ(iλL)1op\lambda\mapsto||(i\lambda-L)^{-1}||_{\rm op} on \mathbb{R}, that assigns to λ\lambda the operator norm of (iλL)1(i\lambda-L)^{-1}, is well-defined, and therefore also continuous. It converges to 0 as λ±\lambda\to\pm\infty. Hence it is uniformly bounded in λ\lambda. In particular,

(ik,ωL)1opCL:=maxλ(iλL)1op.||(i\langle k,\omega\rangle-L)^{-1}||_{\rm op}\leq C_{L}:=\max_{\lambda\in\mathbb{R}}||(i\lambda-L)^{-1}||_{\rm op}\,.

It thus follows from (4.21) that

𝐡jWCLVjW.||{\bf h}_{j}||_{W}\leq C_{L}||V_{j}||_{W}\,.

This means that 𝐡j{\bf h}_{j} is at least as smooth as VjV_{j}.

Theorem 4.4 shows that one can choose 𝐠j{\bf g}_{j} (and thus the component of eje_{j} tangent to 𝕋0\mathbb{T}_{0}) freely when solving the homological equations (4.16) and (4.17). This reflects the fact that the embedding of 𝕋ε\mathbb{T}_{\varepsilon} is not unique. Corollary 4.6 below states that it is possible to choose 𝐠j{\bf g}_{j} in such a way that 𝐟j{\bf f}_{j} is in “normal form”. We first define this concept.

Definition 4.5.

Let

𝐟=ω+ε𝐟1+ε2𝐟2+:(/2π)mm{\bf f}=\omega+\varepsilon{\bf f}_{1}+\varepsilon^{2}{\bf f}_{2}+\ldots:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{m}

be an asymptotic expansion of a vector field on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m}. Assume that the Fourier series of 𝐟j{\bf f}_{j} is given by

𝐟j(ϕ)=kmfj,keik,ϕfor certainfj,km.{\bf f}_{j}(\phi)=\sum_{k\in\mathbb{Z}^{m}}f_{j,k}e^{i\langle k,\phi\rangle}\ \mbox{for certain}\ f_{j,k}\in\mathbb{C}^{m}\,.

For kmk\in\mathbb{Z}^{m}, denote |k|=(|k1|2++|km|2)12|k|=\left(|k_{1}|^{2}+\ldots+|k_{m}|^{2}\right)^{\frac{1}{2}} as before. We say that 𝐟j{\bf f}_{j} is in normal form to order K{}K\in\mathbb{N}\cup\{\infty\} in its Fourier expansion if

fj,k=0for allkmwithω,k0and|k|K.f_{j,k}=0\ \mbox{for all}\ k\in\mathbb{Z}^{m}\ \mbox{with}\ \langle\omega,k\rangle\neq 0\ \mbox{and}\ |k|\leq K\,.
Remark 7.

We remark that 𝐟j{\bf f}_{j} is in normal form to order KK in its Fourier expansion, if and only if its truncated Fourier series

𝐟jK(ϕ):=|k|Kfj,keik,ϕ{\bf f}_{j}^{K}(\phi):=\sum_{|k|\leq K}f_{j,k}e^{i\langle k,\phi\rangle}

depends only on so-called resonant combination angles. A combination angle k,ϕ\langle k,\phi\rangle is called resonant when k,ω=0\langle k,\omega\rangle=0.

The following result shows that we can arrange for the reduced phase vector field to be in normal form to arbitrarily high-order in its Fourier expansion.

Corollary 4.6.

For any (finite) KK\in\mathbb{N} the function 𝐠j{\bf g}_{j} can be chosen in such a way that the solution 𝐟j{\bf f}_{j} to the tangential homological equation

ω𝐠j+𝐟j=Uj\partial_{\omega}{\bf g}_{j}+{\bf f}_{j}=U_{j}

is in normal form to order KK in its Fourier expansion.

Proof.

Recall that the tangential homological equation reduces to the equations

iω,kgj,k+fj,k=Uj,k\displaystyle i\langle\omega,k\rangle g_{j,k}+f_{j,k}=U_{j,k} (4.22)

for the Fourier coefficients of 𝐟j{\bf f}_{j}, 𝐠j{\bf g}_{j} and UjU_{j}—see (4.18). Given KK\in\mathbb{N}, choose

gj,k=Uj,kik,ωwhenk,ω0and|k|K,gj,k=0whenk,ω=0or|k|>K.\displaystyle\begin{array}[]{ll}g_{j,k}=\frac{U_{j,k}}{i\langle k,\omega\rangle}&\mbox{when}\ \langle k,\omega\rangle\neq 0\ \mbox{and}\ |k|\leq K,\\ g_{j,k}=0&\mbox{when}\ \langle k,\omega\rangle=0\ \mbox{or}\ |k|>K.\end{array} (4.25)

The (unique) solutions to (4.22) are then given by

fj,k=0whenk,ω0and|k|K,fj,k=Uj,kwhenk,ω=0or|k|>K.\displaystyle\begin{array}[]{ll}f_{j,k}=0&\mbox{when}\ \langle k,\omega\rangle\neq 0\ \mbox{and}\ |k|\leq K,\\ f_{j,k}=U_{j,k}&\mbox{when}\ \langle k,\omega\rangle=0\ \mbox{or}\ |k|>K.\end{array} (4.28)

With these choices, 𝐠j{\bf g}_{j} is a smooth function, as its Fourier expansion is finite. It is also clear that 𝐟j{\bf f}_{j} is in normal form to order KK in its Fourier expansion. ∎

Remark 8.

Recall that the flow of the ODE ϕ˙=ω\dot{\phi}=\omega on (/2π)m(\mathbb{R}/2\pi\mathbb{Z})^{m} is periodic or quasi-periodic and given by the formula ϕϕ+ωtmod(2π)m\phi\mapsto\phi+\omega t\ \mbox{mod}\,(2\pi\mathbb{Z})^{m}. It follows that the time-average over this (quasi-)periodic flow, of a complex exponential vector field fkeik,ϕf_{k}e^{i\langle k,\phi\rangle} (with fkmf_{k}\in\mathbb{C}^{m}) is given by

limT1T0T\displaystyle\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T} fkeik,ϕ+ωtdt={fkeik,ϕwhenk,ω=0,0whenk,ω0.\displaystyle f_{k}e^{i\langle k,\phi+\omega t\rangle}\,dt=\left\{\begin{array}[]{ll}f_{k}e^{i\langle k,\phi\rangle}&\mbox{when}\ \langle k,\omega\rangle=0\,,\\ 0&\mbox{when}\ \langle k,\omega\rangle\neq 0\,.\end{array}\right. (4.31)

This shows that fkeik,ϕf_{k}e^{i\langle k,\phi\rangle} is resonant (that is: it depends on a resonant combination angle) precisely when it is equal to its average over the (quasi-)periodic flow, whereas fkeik,ϕf_{k}e^{i\langle k,\phi\rangle} is nonresonant precisely when this average is zero. For an arbitrary (and sufficiently regular) Fourier series it follows that

limT1T0T(kmfkeik,ϕ+ωt)𝑑t=kmω,k=0fkeik,ϕ.\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\left(\sum_{k\in\mathbb{Z}^{m}}f_{k}\,e^{i\langle k,\phi+\omega t\rangle}\right)dt=\sum_{\footnotesize\begin{array}[]{c}k\in\mathbb{Z}^{m}\\ \langle\omega,k\rangle=0\end{array}}\!\!\!\!f_{k}\,e^{i\langle k,\phi\rangle}\,.

We conclude that averaging a Fourier series removes its nonresonant terms, while keeping its resonant terms untouched. Corollary 4.6 shows that it can be arranged that the term 𝐟j{\bf f}_{j} in the reduced vector field 𝐟{\bf f} is a sum of resonant terms only (to arbitrarily high order). We may thus loosely interpret Corollary 4.6 as a high-order averaging theorem, see [28].

Remark 9.

We include the following result for completeness. Applied to A=e0(ϕ)A=e_{0}^{\prime}(\phi) and B=N(ϕ)B=N(\phi) it gives a formula for the projection π(ϕ)\pi(\phi) onto the tangent space to 𝕋0\mathbb{T}_{0} at e0(ϕ)e_{0}(\phi) along the fast fibre at that point. This formula is not only useful for practical computations, but also shows explicitly that π(ϕ)\pi(\phi) depends smoothly on ϕ\phi. A proof of Proposition 4.7 is given in [19].

Proposition 4.7.

Let 1mM1\leq m\leq M and assume that A(m,M)A\in\mathcal{L}(\mathbb{R}^{m},\mathbb{R}^{M}) and B(Mm,M)B\in\mathcal{L}(\mathbb{R}^{M-m},\mathbb{R}^{M}) are linear maps satisfying M=imAimB\mathbb{R}^{M}={\rm im}\,A\oplus{\rm im}\,B. We denote by π(M,M)\pi\in\mathcal{L}(\mathbb{R}^{M},\mathbb{R}^{M}) the “oblique projection” onto the image of AA along the image of BB, i.e., π\pi is the unique linear map satisfying πA=A\pi A=A and πB=0\pi B=0. Then π\pi is given by the formula

π=A(ATπ(B)A)1ATπ(B)in whichπ(B):=(1B(BTB)1BT).\pi=A(A^{T}\pi(B)^{\perp}A)^{-1}A^{T}\pi(B)^{\perp}\ \mbox{in which}\ \pi(B)^{\perp}:=(1-B(B^{T}B)^{-1}B^{T})\,.

The TT denotes matrix transpose. All the inverses in this formula exist. Note that π(B)\pi(B)^{\perp} is the orthogonal projection onto kerBT\ker B^{T} along imB{\rm im}\,B.

5 Reducibility for oscillator systems

In this section we show that the invariant torus of a system of uncoupled oscillators (see the introduction) is reducible. We also give a formula for the fast fibre map for such a torus. The results in this section are a consequence of Floquet’s theorem, which implies that the invariant circle defined by a single hyperbolic periodic solution of an ODE is reducible. The results in this section should thus be considered well-known, but for completeness we include them in detail. We start with the result for single hyperbolic periodic orbits.

Theorem 5.1.

Let X:MX:\mathbb{R}\to\mathbb{R}^{M} be a hyperbolic TT-periodic orbit of a smooth vector field 𝐅:MM{\bf F}:\mathbb{R}^{M}\to\mathbb{R}^{M}. Then the invariant circle 𝕋0=X()M\mathbb{T}_{0}=X(\mathbb{R})\subset\mathbb{R}^{M} is reducible and normally hyperbolic. Its fast fibre map is given by formula (5.1).

Proof.

Assume that the ODE x˙=𝐅(x)onM\dot{x}={\bf F}(x)\ \mbox{on}\ \mathbb{R}^{M} possesses a hyperbolic periodic orbit X=X(t)X=X(t) with minimal period T>0T>0. We think of it as an invariant circle 𝕋0\mathbb{T}_{0} embedded by the map e0:/2πMe_{0}:\mathbb{R}/2\pi\mathbb{Z}\to\mathbb{R}^{M} defined by e0(ϕ):=X(ω1ϕ)e_{0}(\phi):=X(\omega^{-1}\phi), where ω:=2πT\omega:=\frac{2\pi}{T}. Let Φ=Φ(t)GL(M)\Phi=\Phi(t)\in{\rm GL}(\mathbb{R}^{M}) be the principal fundamental matrix solution of the linearisation around this periodic orbit. This means that

Φ˙(t)=𝐅(X(t))Φ(t)andΦ(0)=IdM.\dot{\Phi}(t)={\bf F}^{\prime}(X(t))\cdot\Phi(t)\ \mbox{and}\ \Phi(0)={\rm Id}_{\mathbb{R}^{M}}\,.

Floquet’s theorem [9, 29] states that Φ(t)\Phi(t) admits a factorisation

Φ(t)=P(t)eBtwithP(t+T)=P(t)andP(0)=IdM.\Phi(t)=P(t)e^{Bt}\ \mbox{with}\ P(t+T)=P(t)\ \mbox{and}\ P(0)={\rm Id}_{\mathbb{R}^{M}}\,.

The constant (and perhaps complex) Floquet matrix BB satisfies eBT=Φ(T)e^{BT}=\Phi(T), for example B=1TlogΦ(T)B=\frac{1}{T}\log\Phi(T) for a choice of matrix logarithm. Note that a matrix logarithm of Φ(T)\Phi(T) exists because Φ(T)\Phi(T) is invertible. We shall assume here that BB is a real matrix. This can always be arranged by replacing TT by 2T2T and considering a double cover of 𝕋0\mathbb{T}_{0} if necessary, but we ignore this (somewhat annoying) subtlety here.

Substituting the Floquet decomposition in the definition of the fundamental matrix solution, we obtain that P˙(t)eBt+P(t)BeBt=𝐅(X(t))P(t)eBt\dot{P}(t)e^{Bt}+P(t)Be^{Bt}={\bf F}^{\prime}(X(t))P(t)e^{Bt}. Thus,

P˙(t)+P(t)B=𝐅(X(t))P(t).\dot{P}(t)+P(t)B={\bf F}^{\prime}(X(t))P(t)\,.

This implies that we found a solution to Equation (3.5) in Lemma 3.3. Indeed, if we define

L~=BandN~(ϕ)=P(ω1ϕ),\tilde{L}=B\ \mbox{and}\ \tilde{N}(\phi)=P(\omega^{-1}\phi)\,,

then we have, recalling that e0(ϕ)=X(ω1ϕ)e_{0}(\phi)=X(\omega^{-1}\phi)),

ωN~(ϕ)+N~(ϕ)L~\displaystyle\partial_{\omega}\tilde{N}(\phi)+\tilde{N}(\phi)\cdot\tilde{L} =N~(ϕ)ω+N~(ϕ)L~\displaystyle=\tilde{N}^{\prime}(\phi)\cdot\omega+\tilde{N}(\phi)\cdot\tilde{L}
=P˙(ω1ϕ)+P(ω1ϕ)B\displaystyle=\dot{P}(\omega^{-1}\phi)+P(\omega^{-1}\phi)\cdot B
=𝐅(X(ω1ϕ))P(ω1ϕ)=𝐅(e0(ϕ))N~(ϕ).\displaystyle={\bf F}^{\prime}(X(\omega^{-1}\phi))\cdot P(\omega^{-1}\phi)={\bf F}^{\prime}(e_{0}(\phi))\cdot\tilde{N}(\phi)\,.

However, this does not yet prove that the periodic orbit is reducible, because N~=N~(ϕ)\tilde{N}=\tilde{N}(\phi) defines a family of M×MM\times M-matrices, and hence the image of N~(ϕ)\tilde{N}(\phi) is not normal to the tangent vector ωe0(ϕ)=X˙(ω1ϕ)\omega e_{0}^{\prime}(\phi)=\dot{X}(\omega^{-1}\phi) to the periodic orbit. To resolve this issue, recall that Φ(T)\Phi(T) always has a unit eigenvalue. This follows from differentiating the identity X˙(t)=𝐅(X(t))\dot{X}(t)={\bf F}(X(t)) to tt, which gives that ddtX˙(t)=𝐅(X(t))X˙(t)\frac{d}{dt}\dot{X}(t)={\bf F}^{\prime}(X(t))\cdot\dot{X}(t), so that

X˙(0)=X˙(T)=Φ(T)X˙(0).\dot{X}(0)=\dot{X}(T)=\Phi(T)\cdot\dot{X}(0)\,.

Because Φ(T)=eBT\Phi(T)=e^{BT}, we conclude that BB has a purely imaginary eigenvalue in 2πiT\frac{2\pi i}{T}\mathbb{Z}. Our assumption that XX is hyperbolic implies that none of the other eigenvalues of BB lie on the imaginary axis. Because BB is real and its eigenvalues must thus come in complex conjugate pairs, we conclude that the purely imaginary eigenvalue of BB must in fact be zero.

We now choose an injective linear map A:M1MA:\mathbb{R}^{M-1}\to\mathbb{R}^{M} whose image coincides with the (M1)(M-1)-dimensional image of BB. For any such choice of AA there is a unique map L:M1M1L:\mathbb{R}^{M-1}\to\mathbb{R}^{M-1} for which

AL=BA.A\cdot L=B\cdot A\,.

Clearly, the eigenvalues of LL are the nonzero eigenvalues of BB, showing that LL is hyperbolic. We also define N:/2π(M1,M)N:\mathbb{R}/2\pi\mathbb{Z}\to\mathcal{L}(\mathbb{R}^{M-1},\mathbb{R}^{M}) by

N(ϕ):=P(ω1ϕ)A.\displaystyle N(\phi):=P(\omega^{-1}\phi)A\,. (5.1)

By definition, imN(0)=imA=imB{\rm im}\,N(0)={\rm im}\,A={\rm im}\,B is transverse to the tangent vector X˙(0)kerB\dot{X}(0)\in\ker B to the periodic orbit. Because each P(t)P(t) is invertible, this transversality persists along the entire orbit. Indeed, writing t=ω1ϕt=\omega^{-1}\phi, note that

X˙(t)=Φ(t)X˙(0)=P(t)eBtX˙(0)=P(t)X˙(0)P(t)(kerB)\dot{X}(t)=\Phi(t)\dot{X}(0)=P(t)e^{Bt}\dot{X}(0)=P(t)\dot{X}(0)\in P(t)({\rm ker}\,B)

is transversal to imN(ϕ)=im(P(t)A)=P(t)(imB){\rm im}\,N(\phi)={\rm im}(P(t)A)=P(t)({\rm im}\,B). Finally, we compute

ωN(ϕ)+N(ϕ)L\displaystyle\partial_{\omega}N(\phi)+N(\phi)L =N(ϕ)ω+N(ϕ)L\displaystyle=N^{\prime}(\phi)\,\omega+N(\phi)L
=P˙(ω1ϕ)A+P(ω1ϕ)AL\displaystyle=\dot{P}(\omega^{-1}\phi)A+P(\omega^{-1}\phi)AL
=P˙(ω1ϕ)A+P(ω1ϕ)BA\displaystyle=\dot{P}(\omega^{-1}\phi)A+P(\omega^{-1}\phi)BA
=𝐅(X(ω1ϕ))P(ω1ϕ)A=𝐅(e0(ϕ))N(ϕ).\displaystyle={\bf F}^{\prime}(X(\omega^{-1}\phi))P(\omega^{-1}\phi)A={\bf F}^{\prime}(e_{0}(\phi))N(\phi)\,.

This proves that the invariant circle 𝕋0\mathbb{T}_{0} defined by X(t)X(t) is reducible. ∎

Example 5.2.

As an example consider a single Stuart-Landau oscillator

z˙=(α+iβ)z+(γ+iδ)|z|2zforz2.\displaystyle\dot{z}=\left(\alpha+i\beta\right)z+\left(\gamma+i\delta\right)|z|^{2}z\quad\mbox{for}\ z\in\mathbb{C}\cong\mathbb{R}^{2}\,. (5.2)

Here α,β,γ,δ\alpha,\beta,\gamma,\delta\in\mathbb{R} are parameters. We assume that αγ<0\alpha\gamma<0 and αδβγ0\alpha\delta-\beta\gamma\neq 0, so that (5.2) possesses a unique (up to rotation) circular periodic orbit

X(t)=ReiωtwhereR:=α/γandω:=βαδ/γ0.X(t)=Re^{i\omega t}\ \mbox{where}\ R:=\sqrt{-\alpha/\gamma}\ \mbox{and}\ \omega:=\beta-\alpha\delta/\gamma\neq 0\,.

Thus, the embedding

e0:/2πϕz:=Reiϕe_{0}:\mathbb{R}/2\pi\mathbb{Z}\ni\phi\mapsto z:=R\,e^{i\phi}\in\mathbb{C}

sends solutions of ϕ˙=ω\dot{\phi}=\omega on /2π\mathbb{R}/2\pi\mathbb{Z} to solutions of (5.2). The Floquet decomposition of the fundamental matrix solution around this periodic orbit can be found by anticipating that P(t)=eiωtP(t)=e^{i\omega t} and thus making the ansatz

Φ(t)=eiωteBt\Phi(t)=e^{i\omega t}e^{Bt}\,

for an unknown linear map B:B:\mathbb{C}\to\mathbb{C}. With this in mind we expand solutions to (5.2) nearby the periodic orbit as

z(t)=Reiωt+εeiωtv(t).z(t)=R\,e^{i\omega t}+\varepsilon\,e^{i\omega t}v(t)\,.

To first order in ε\varepsilon this gives the linear differential equations

v˙=v˙1+iv˙2=2R2(γ+iδ)v1,\dot{v}=\dot{v}_{1}+i\dot{v}_{2}=2R^{2}(\gamma+i\delta)v_{1}\,,

which shows that the Floquet map B:B:\mathbb{C}\to\mathbb{C} must be given by

B(v1+iv2)=2R2(γ+iδ)v1.B(v_{1}+iv_{2})=2R^{2}(\gamma+i\delta)v_{1}\,.

This BB has an eigenvalue 0 (with eigenvector ii corresponding to the tangent space to the invariant circle) and an eigenvalue 2γR2=2α02\gamma R^{2}=-2\alpha\neq 0 (with eigenvector γ+iδ\gamma+i\delta). We conclude that the map

𝐍e0:(ϕ,u)(Reiϕ,eiϕ(γ+iδ)u)from/2π×to×{\bf N}e_{0}:(\phi,u)\mapsto(Re^{i\phi},e^{i\phi}(\gamma+i\delta)u)\ \mbox{from}\ \mathbb{R}/2\pi\mathbb{Z}\times\mathbb{R}\ \mbox{to}\ \mathbb{C}\times\mathbb{C}

sends solutions of

ϕ˙=ω,u˙=2αuforϕ/2πandu\dot{\phi}=\omega\,,\,\dot{u}=-2\alpha\,u\ \mbox{for}\ \phi\in\mathbb{R}/2\pi\mathbb{Z}\ \mbox{and}\ u\in\mathbb{R}

to solutions of the linearised dynamics of (5.2) on ×\mathbb{C}\times\mathbb{C} around the invariant circle. In particular, we have L=2αL=-2\alpha and N(ϕ)=eiϕ(γ+iδ)N(\phi)=e^{i\phi}(\gamma+i\delta). The projection onto the tangent bundle of the invariant circle along its fast fibre bundle is given by the formulas

π(0)(x+iy)=i(y(δ/γ)x)andπ(ϕ)=eiϕπ(0)eiϕ.\pi(0)\cdot(x+iy)=i(y-(\delta/\gamma)x)\ \mbox{and}\ \pi(\phi)=e^{i\phi}\cdot\pi(0)\cdot e^{-i\phi}\,.

Indeed, it is easy to check that π(ϕ)ieiϕ=ieiϕ\pi(\phi)\cdot ie^{i\phi}=ie^{i\phi} and π(ϕ)eiϕ(γ+iδ)=0\pi(\phi)\cdot e^{i\phi}(\gamma+i\delta)=0.

We now extend the result of Theorem 5.1 to systems of multiple uncoupled oscillators, that is, systems of the form

x˙1=F1(x1),,x˙m=Fm(xm)withxjMj,\dot{x}_{1}=F_{1}(x_{1})\,,\,\ldots\,,\,\dot{x}_{m}=F_{m}(x_{m})\,\ \mbox{with}\ x_{j}\in\mathbb{R}^{M_{j}}\,,

that each have a hyperbolic TjT_{j}-periodic orbit Xj(t)X_{j}(t). Recall that the product of these periodic orbits forms an invariant torus. The fact that this torus is reducible follows from the following lemma. Its proof is straightforward, but included here for completeness.

Lemma 5.3.

Let 𝕋1M1\mathbb{T}_{1}\subset\mathbb{R}^{M_{1}} and 𝕋2M2\mathbb{T}_{2}\subset\mathbb{R}^{M_{2}} be embedded reducible normally hyperbolic (quasi-)periodic invariant tori for the vector fields 𝐅1{\bf F}_{1} and 𝐅2{\bf F}_{2} respectively. Then the product torus 𝕋0:=𝕋1×𝕋2M\mathbb{T}_{0}:=\mathbb{T}_{1}\times\mathbb{T}_{2}\subset\mathbb{R}^{M} (with M:=M1+M2M:=M_{1}+M_{2}) is an embedded reducible normally hyperbolic quasi-periodic invariant torus for the product vector field 𝐅0{\bf F}_{0} on M\mathbb{R}^{M} defined by 𝐅0(x1,x2):=(𝐅1(x1),𝐅2(x2)){\bf F}_{0}(x_{1},x_{2}):=({\bf F}_{1}(x_{1}),{\bf F}_{2}(x_{2})).

Proof.

Assume that ej:(/2π)mjMje_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m_{j}}\to\mathbb{R}^{M_{j}} (for j=1,2j=1,2) is an embedding of a reducible normally hyperbolic (quasi-)periodic invariant torus for the vector field 𝐅j{\bf F}_{j}. This means that there are frequency vectors ωjmj\omega_{j}\in\mathbb{R}^{m_{j}} such that ωjej=𝐅jej\partial_{\omega_{j}}e_{j}={\bf F}_{j}\circ e_{j} and fast fibre maps 𝐍ej:(/2π)mj×MjmjMj×Mj{\bf N}e_{j}:(\mathbb{R}/2\pi\mathbb{Z})^{m_{j}}\times\mathbb{R}^{M_{j}-m_{j}}\to\mathbb{R}^{M_{j}}\times\mathbb{R}^{M_{j}} of the form 𝐍ej(ϕj,uj)=(ej(ϕj),Nj(ϕj)uj){\bf N}e_{j}(\phi_{j},u_{j})=(e_{j}(\phi_{j}),N_{j}(\phi_{j})\cdot u_{j}) satisfying ωjNj+NjLj=(𝐅jej)Nj\partial_{\omega_{j}}N_{j}+N_{j}\cdot L_{j}=({\bf F}_{j}^{\prime}\circ e_{j})\cdot N_{j} for certain hyperbolic Floquet matrices LjL_{j}.

If we now define m:=m1+m2m:=m_{1}+m_{2}, ω:=(ω1,ω2)m\omega:=(\omega_{1},\omega_{2})\in\mathbb{R}^{m} and e0:(/2π)mMe_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M} by e0(ϕ)=e0(ϕ1,ϕ2):=(e1(ϕ1),e2(ϕ2))e_{0}(\phi)=e_{0}(\phi_{1},\phi_{2}):=(e_{1}(\phi_{1}),e_{2}(\phi_{2})), then e0e_{0} is clearly an embedding of 𝕋0\mathbb{T}_{0} and the equality ωe0=𝐅0e0\partial_{\omega}e_{0}={\bf F}_{0}\circ e_{0} holds. In other words, the product torus 𝕋0\mathbb{T}_{0} is an embedded quasi-periodic invariant torus for 𝐅0{\bf F}_{0}.

If we also define N(ϕ)u=N(ϕ1,ϕ2)(u1,u2):=(N1(ϕ1)u1,N2(ϕ2)u2)N(\phi)\cdot u=N(\phi_{1},\phi_{2})\cdot(u_{1},u_{2}):=(N_{1}(\phi_{1})\cdot u_{1},N_{2}(\phi_{2})\cdot u_{2}), then clearly N(ϕ)N(\phi) is injective, and therefore 𝐍e0:(/2π)mM×M{\bf N}e_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{m}\to\mathbb{R}^{M}\times\mathbb{R}^{M} defined by

𝐍e0((ϕ1,ϕ2),(u1,u2))=(e0(ϕ1,ϕ2),N(ϕ1,ϕ2)(u1,u2)){\bf N}e_{0}((\phi_{1},\phi_{2}),(u_{1},u_{2}))=(e_{0}(\phi_{1},\phi_{2}),N(\phi_{1},\phi_{2})\cdot(u_{1},u_{2}))

is a fast fibre map for 𝕋0\mathbb{T}_{0} that satisfies ωN+NL=(𝐅0e0)N\partial_{\omega}N+N\cdot L=({\bf F}_{0}^{\prime}\circ e_{0})\cdot N. Here L:MmMmL:\mathbb{R}^{M-m}\to\mathbb{R}^{M-m} is defined by L(u1,u2):=(L1u1,L2u2)L(u_{1},u_{2}):=(L_{1}u_{1},L_{2}u_{2}). This LL is hyperbolic, its eigenvalues being those of L1L_{1} and L2L_{2}. This proves that 𝕋0\mathbb{T}_{0} is reducible and normally hyperbolic and concludes the proof of the lemma. ∎

6 Application to remote synchronisation

In this final section we apply and illustrate our phase reduction method in a small network of three weakly linearly coupled Stuart-Landau oscillators

z˙1=(α+iβ)z1+(γ+iδ)|z1|2z1+εz2,z˙2=(a+ib)z2+(c+id)|z2|2z2+εz1,z˙3=(α+iβ)z3+(γ+iδ)|z3|2z3+εz2,\displaystyle\begin{array}[]{llll}\dot{z}_{1}=&(\alpha+i\beta)z_{1}&+\hskip 2.84526pt(\gamma+i\delta)|z_{1}|^{2}z_{1}&+\hskip 2.84526pt\varepsilon z_{2}\,,\\ \dot{z}_{2}=&(a\,+\,\!ib)z_{2}&+\hskip 2.84526pt(c+id)|z_{2}|^{2}z_{2}&+\hskip 2.84526pt\varepsilon z_{1}\,,\\ \dot{z}_{3}=&(\alpha+i\beta)z_{3}&+\hskip 2.84526pt(\gamma+i\delta)|z_{3}|^{2}z_{3}&+\hskip 2.84526pt\varepsilon z_{2}\,,\end{array} (6.4)

with z1,z2,z3z_{1},z_{2},z_{3}\in\mathbb{C}. Figure 1 depicts the coupling architecture of this network. Note that the first and third oscillator in equations (6.4) are identical. We choose parameters so that each uncoupled oscillator has a nonzero hyperbolic periodic orbit, with frequencies ω1=ω3ω2\omega_{1}=\omega_{3}\neq\omega_{2}. These periodic orbits form a 33-dimensional invariant torus 𝕋0\mathbb{T}_{0} for the uncoupled system, which persists as a perturbed torus 𝕋ε\mathbb{T}_{\varepsilon} for small nonzero coupling.

Despite that fact that the first and third oscillators in (6.4) are not coupled directly, a numerical study of equations (6.4) reveals that these oscillators synchronise when appropriate parameter values are chosen, see Figure 2. This “remote synchronisation” appears to be mediated by the second oscillator, which allows the two other oscillators to communicate. Figure 3 demonstrates, again numerically, that the timescale of remote synchronisation is of the order tε2t\sim\varepsilon^{-2}. This suggests that proving the synchronisation rigorously would require second-order phase reduction.

In [3], remote synchronisation of Stuart-Landau oscillators was observed numerically for the first time. A first rigorous proof of the phenomenon, for a chain of three Stuart-Landau oscillators, occurs in [16]. The proof in that paper employs the high-order phase reduction method developed in [10]. However, the method in [10] does not yield the reduced phase equations in normal form. As a result, the timescale tε2t\sim\varepsilon^{-2} is not observed in [16].

Here we apply the parametrisation method developed in this paper, to prove that the first and third oscillator in (6.4) synchronise over a timescale tε2t\sim\varepsilon^{-2}. We are also able to determine how the parameters in (6.4) influence this synchronisation. To this end, we will compute an asymptotic expansion of an embedding e:(/2π)33e:(\mathbb{R}/2\pi\mathbb{Z})^{3}\to\mathbb{C}^{3} and a reduced phase vector field 𝐟:(/2π)33{\bf f}:(\mathbb{R}/2\pi\mathbb{Z})^{3}\to\mathbb{R}^{3} to second order in the small parameter. As we are primarily interested in the synchronisation of the first and third oscillator, we do not calculate the full reduced phase vector field. Instead, we only explicitly compute an evolution equation for the resonant combination angle Φ:=ϕ1ϕ3\Phi:=\phi_{1}-\phi_{3}. We will show that

Φ˙=ε2(AsinΦ+B(1cosΦ))+𝒪(ε3),\displaystyle\dot{\Phi}=\varepsilon^{2}\left(-A\sin\Phi+B(1-\cos\Phi)\right)+\mathcal{O}(\varepsilon^{3})\,, (6.5)

in which the constants AA and BB are given by the formulas

A=14a2+(ω1ω2)2(δγ(ω2ω1)+a(1+dδcγ)+2a2(dc+δγ)1ω2ω1),B=14a2+(ω1ω2)2((ω2ω1)+a(dcδγ)+2a2(1dδcγ)1ω2ω1).\displaystyle\begin{array}[]{ll}A&\!\!\!=\frac{1}{4a^{2}+(\omega_{1}-\omega_{2})^{2}}\left(\frac{\delta}{\gamma}(\omega_{2}-\omega_{1})+a\left(1+\frac{d\delta}{c\gamma}\right)+2a^{2}\left(\frac{d}{c}+\frac{\delta}{\gamma}\right)\frac{1}{\omega_{2}-\omega_{1}}\right)\,,\\ B&\!\!\!=\frac{1}{4a^{2}+(\omega_{1}-\omega_{2})^{2}}\left((\omega_{2}-\omega_{1})+a\left(\frac{d}{c}-\frac{\delta}{\gamma}\right)+2a^{2}\left(1-\frac{d\delta}{c\gamma}\right)\frac{1}{\omega_{2}-\omega_{1}}\right)\,.\end{array} (6.8)
1122332211
Figure 1: Representation of the network of Stuart-Landau oscillators (6.4).

Before we prove formulas (6.5) and (6.8), let us investigate their dynamical implications. After rescaling time tτ:=ε2tt\mapsto\tau:=\varepsilon^{2}t, equation (6.5) becomes

dΦdτ=(AsinΦ+B(1cosΦ))+𝒪(ε).\frac{d\Phi}{d\tau}=\left(-A\sin\Phi+B(1-\cos\Phi)\right)+\mathcal{O}(\varepsilon)\,.

For ε=0\varepsilon=0 the time-rescaled reduced flow on (/2π)3(\mathbb{R}/2\pi\mathbb{Z})^{3} therefore admits a 22-dimensional invariant torus

S={ϕ1=ϕ3}(/2π)3S=\{\phi_{1}=\phi_{3}\}\subset(\mathbb{R}/2\pi\mathbb{Z})^{3}

on which the phases of the first and third oscillator are synchronised. This torus is stable when A>0A>0 and unstable when A<0A<0. For A0A\neq 0, there also exists exactly one 22-dimensional invariant torus of the form

P={ϕ1=ϕ3+c}(/2π)3for somec0P=\{\phi_{1}=\phi_{3}+c\}\subset(\mathbb{R}/2\pi\mathbb{Z})^{3}\ \mbox{for some}\ c\neq 0\,

with the opposite stability type. The phases of the first and third oscillator are phase-locked but not synchronised on PP. Fénichel’s theorem guarantees that both SS and PP persist as invariant submanifolds of (/2π)3(\mathbb{R}/2\pi\mathbb{Z})^{3} for small ε0\varepsilon\neq 0. Hence, so do their images e(S),e(P)𝕋ε3e(S),e(P)\subset\mathbb{T}_{\varepsilon}\subset\mathbb{C}^{3} as invariant manifolds for (6.4).

For small ε0\varepsilon\neq 0, a typical solution of (6.4) will therefore first converge to the 33-dimensional invariant torus 𝕋ε\mathbb{T}_{\varepsilon} on a timescale of the order t1t\sim 1. It will subsequently converge to either e(S)e(S) or e(P)e(P) on the much longer timescale tε2t\sim\varepsilon^{-2}, and it is this slow dynamics that governs the synchronisation of the first and third oscillator. This multiple timescale dynamical process is illustrated in Figure 2. Figure 3 confirms numerically that the timescale of synchronisation of z1z_{1} and z3z_{3} is indeed of the order ε2\varepsilon^{-2}.

Remark 10.

We point out that the parameters in (6.4) can be tuned so that either of the two low-dimensional tori SS or PP is the stable one. Assume for instance that α,a>0\alpha,a>0 and γ,c<0\gamma,c<0, so that 𝕋0\mathbb{T}_{0} (and hence 𝕋ε\mathbb{T}_{\varepsilon}) is stable. If in addition we choose the parameters so that cδ+dγ=0c\delta+d\gamma=0, then the expression for AA simplifies to a+(bβ)(δ/γ)+α(δ/γ)24a2+(ω1ω2)2\frac{a+(b-\beta)(\delta/\gamma)+\alpha(\delta/\gamma)^{2}}{4a^{2}+(\omega_{1}-\omega_{2})^{2}}. If δ0\delta\neq 0, then it is clear that we can make this both positive and negative, for instance by varying the parameter bb. Interestingly, this shows that properties of the second oscillator may determine whether the first and third oscillator converge to the synchronised state SS or the phase-locked state PP.

Numerics

Refer to caption
(a) Slow convergence of Φ^\hat{\Phi} to zero.
Refer to caption
(b) Convergence of Φ^\hat{\Phi} to a non-zero value.
Figure 2: Numerically obtained plots of the phase-difference Φ^=Arg(z1z3¯)ϕ1ϕ3\hat{\Phi}=\operatorname{Arg}(z_{1}\overline{z_{3}})\approx\phi_{1}-\phi_{3} against time, for two different realisations of system (6.4).

Before proving (6.5) and (6.8), we present some numerical results on system (6.4). Figure 2 shows numerically obtained plots of Φ^=Arg(z1z3¯)\hat{\Phi}=\operatorname{Arg}(z_{1}\overline{z_{3}}) against time, for two different realisations of system (6.4). We use Φ^\hat{\Phi} as a proxy for Φ=ϕ1ϕ3\Phi=\phi_{1}-\phi_{3}. As this approximation does not take into account the distortion of the perturbed invariant torus, we observe small amplitude, rapid oscillations in Φ^\hat{\Phi}, causing the lines in Figure 2 to be thick. In Figure 2(a), we have chosen the parameter values

α=1β=1γ=1δ=1;a=1b=2c=1d=1,\begin{array}[]{llll}\alpha=1&\beta=1&\gamma=-1&\delta=1\,;\\ a=1&b=2&c=-1&d=-1\,,\end{array} (6.9)

together with ε=0.1\varepsilon=0.1. It follows that cδ+dγ=0c\delta+d\gamma=0, and so A=15>0A=\frac{1}{5}>0, see Remark 10. The above analysis therefore predicts that Φ^\hat{\Phi} should converge to zero, which the figure indeed shows. The convergence is very slow, as only around t=2000t=2000 do we find that Φ^\hat{\Phi} is indistinguishably close to zero. We will comment more on the rate of convergence below. Figure 2(a) was generated using Euler’s method with time steps of 0.050.05, starting from the point in phase space (z1,z2,z3)=(1,1+0.4i,1+0.3i)3(z_{1},z_{2},z_{3})=(-1,1+0.4i,-1+0.3i)\in\mathbb{C}^{3}.

For Figure 2(b) we have likewise set ε=0.1\varepsilon=0.1, but have instead chosen

α=1β=0.1γ=1δ=1;a=1b=6c=1d=1,\begin{array}[]{llll}\alpha=1&\beta=0.1&\gamma=-1&\delta=1\,;\\ a=1&b=6&c=-1&d=-1\,,\end{array} (6.10)

which yields A=3.94+(3.9)2=0.203<0A=\frac{-3.9}{4+(3.9)^{2}}=-0.203\ldots<0. Hence, our theory predicts Φ^\hat{\Phi} to converge to a non-zero constant value, which is indeed seen to be the case. Again the thickness of the line is due to rapid oscillations. Figure 2(b) is generated in the same way as Figure 2(a), except that the starting point for Euler’s method is now (z1,z2,z3)=(1+0.3i,1+0.4i,0.2+0.9i)(z_{1},z_{2},z_{3})=(1+0.3i,1+0.4i,-0.2+0.9i).

Refer to caption
Figure 3: Log-log plot of the time T0.1T_{0.1} it takes for Φ^\hat{\Phi} to decrease by a factor of 1010, against the coupling parameter ε\varepsilon.

Finally, Figure 3 displays the rate of convergence to synchrony as a function of ε\varepsilon. The figure was made using Euler’s method with time-steps of 0.050.05, all starting from the same point (z1,z2,z3)=(1+0.3i,1+0.4i,1+0.5i)(z_{1},z_{2},z_{3})=(-1+0.3i,1+0.4i,-1+0.5i). We have again chosen the parameters as in (6.9), so that we may expect Φ^\hat{\Phi} to converge to zero. However, the rate at which this occurs depends on ε\varepsilon. We measure this rate by recording T0.1T_{0.1}, which is the smallest time tt for which |Φ^(t)|0.1|Φ^(0)||\hat{\Phi}(t)|\leq 0.1|\hat{\Phi}(0)|.

Figure 3 shows a log-log plot of T0.1T_{0.1} against ε\varepsilon. The crosses in the figure represent numerical results for 2020 different values of ε\varepsilon. Shown in green is the line with slope 2-2 through the leftmost cross. We see that ln(T0.1)=2ln(ε)+C\ln(T_{0.1})=-2\ln(\varepsilon)+C for some CC\in\mathbb{R} to very good approximation. Hence we find T0.1ε2T_{0.1}\sim\varepsilon^{-2}, which is fully in agreement with our predictions.

Setup: the unperturbed problem

We now start our proof of formulas (6.5) and (6.8). We first recall some observations from Example 5.2, and make assumptions on the parameters that appear in (6.4). Specifically, we assume that these parameters are chosen so that

αγ<0,ac<0,βγαδ0,bcad0andω1=ω3ω2.\alpha\gamma<0,\ ac<0,\ \beta\gamma-\alpha\delta\neq 0,\ bc-ad\neq 0\,\ \mbox{and}\ \omega_{1}=\omega_{3}\neq\omega_{2}\,.

Recall from Example 5.2 that this ensures that all three uncoupled oscillators possess a unique hyperbolic periodic orbit, with nonzero frequencies ω1=ω3=βαδ/γ\omega_{1}=\omega_{3}=\beta-\alpha\delta/\gamma and ω2=bad/cω1\omega_{2}=b-ad/c\neq\omega_{1}. The product of these periodic orbits forms a 33-dimensional reducible normally hyperbolic (quasi-)periodic invariant torus 𝕋03\mathbb{T}_{0}\subset\mathbb{C}^{3}. An embedding of 𝕋0\mathbb{T}_{0} is given by

e0:(/2π)33defined bye0(ϕ1,ϕ2,ϕ3)=(R1eiϕ1,R2eiϕ2,R3eiϕ3)e_{0}:(\mathbb{R}/2\pi\mathbb{Z})^{3}\to\mathbb{C}^{3}\ \mbox{defined by}\ e_{0}(\phi_{1},\phi_{2},\phi_{3})=(R_{1}\,e^{i\phi_{1}},R_{2}\,e^{i\phi_{2}},R_{3}\,e^{i\phi_{3}})\,

where

R1=R3=α/γ>0andR2=a/c>0.R_{1}=R_{3}=\sqrt{-\alpha/\gamma}>0\ \mbox{and}\ R_{2}=\sqrt{-a/c}>0\ .

This embedding sends integral curves of the constant vector field ω=(ω1,ω2,ω3)\omega=(\omega_{1},\omega_{2},\omega_{3}) on (/2π)3(\mathbb{R}/2\pi\mathbb{Z})^{3} to solutions of (6.4) (with ε=0\varepsilon=0) on 3\mathbb{C}^{3}.

It follows from Example 5.2 and Lemma 5.3 that a Floquet matrix for 𝕋0\mathbb{T}_{0} is

L=diag(2α,2a,2α),L={\rm diag}(-2\alpha,-2a,-2\alpha)\,,

with corresponding fast fibre map given by the family of injective linear maps N:(/2π)3(3,3)N:(\mathbb{R}/2\pi\mathbb{Z})^{3}\to\mathcal{L}(\mathbb{R}^{3},\mathbb{C}^{3}) defined by

N(ϕ1,ϕ2,ϕ3)=diag(eiϕ1(γ+iδ),eiϕ2(c+id),eiϕ3(γ+iδ)).N(\phi_{1},\phi_{2},\phi_{3})={\rm diag}(e^{i\phi_{1}}(\gamma+i\delta),e^{i\phi_{2}}(c+id),e^{i\phi_{3}}(\gamma+i\delta))\,.

The projection onto the tangent bundle along the fast fibre bundle is given by

π(ϕ1,ϕ2,ϕ3)=diag(eiϕ1π1(0)eiϕ1,eiϕ2π2(0)eiϕ2,eiϕ3π3(0)eiϕ3).\pi(\phi_{1},\phi_{2},\phi_{3})={\rm diag}(e^{i\phi_{1}}\pi_{1}(0)e^{-i\phi_{1}},e^{i\phi_{2}}\pi_{2}(0)e^{-i\phi_{2}},e^{i\phi_{3}}\pi_{3}(0)e^{-i\phi_{3}})\,.

Here,

π1(0)(x1+iy1)=i(y1(δ/γ)x1),π2(0)(x2+iy2)=i(y2(d/c)x2),\pi_{1}(0)(x_{1}+iy_{1})=i(y_{1}-(\delta/\gamma)x_{1})\,,\ \pi_{2}(0)(x_{2}+iy_{2})=i(y_{2}-(d/c)x_{2})\,,

and π3(0)=π1(0)\pi_{3}(0)=\pi_{1}(0).

The first tangential homological equation

We now compute 𝐟1{\bf f}_{1} and 𝐠1{\bf g}_{1} from the first tangential homological equation, see (4.16), with U1U_{1} as given in (4.15). A short calculation shows that the projection of the inhomogeneous term 𝐆1(ϕ)=𝐅1(e0(ϕ))=(R2eiϕ2,R1eiϕ1,R2eiϕ2){\bf G}_{1}(\phi)={\bf F}_{1}(e_{0}(\phi))=(R_{2}e^{i\phi_{2}},R_{1}e^{i\phi_{1}},R_{2}e^{i\phi_{2}}) is

(π𝐆1)(ϕ)=(iR2eiϕ1(sin(ϕ2ϕ1)(δ/γ)cos(ϕ2ϕ1))iR1eiϕ2(sin(ϕ1ϕ2)(d/c)cos(ϕ1ϕ2))iR2eiϕ3(sin(ϕ2ϕ3)(δ/γ)cos(ϕ2ϕ3))).(\pi\cdot{\bf G}_{1})(\phi)=\left(\begin{array}[]{c}iR_{2}e^{i\phi_{1}}\left(\sin(\phi_{2}-\phi_{1})-(\delta/\gamma)\cos(\phi_{2}-\phi_{1})\right)\\ iR_{1}e^{i\phi_{2}}\left(\sin(\phi_{1}-\phi_{2})-(d/c)\cos(\phi_{1}-\phi_{2})\right)\\ iR_{2}e^{i\phi_{3}}\left(\sin(\phi_{2}-\phi_{3})-(\delta/\gamma)\cos(\phi_{2}-\phi_{3})\right)\end{array}\right)\,.

This is clearly in the range of e0(ϕ)=diag(iR1eiϕ1,iR2eiϕ2,iR3eiϕ3)e_{0}^{\prime}(\phi)={\rm diag}(iR_{1}e^{i\phi_{1}},iR_{2}e^{i\phi_{2}},iR_{3}e^{i\phi_{3}}). Thus the first tangential homological equation becomes

ω𝐠1(ϕ)+𝐟1(ϕ)=U1(ϕ)=((R2/R1)(sin(ϕ2ϕ1)(δ/γ)cos(ϕ2ϕ1))(R1/R2)(sin(ϕ1ϕ2)(d/c)cos(ϕ1ϕ2))(R2/R3)(sin(ϕ2ϕ3)(δ/γ)cos(ϕ2ϕ3))).\partial_{\omega}{\bf g}_{1}(\phi)+{\bf f}_{1}(\phi)=U_{1}(\phi)=\left(\begin{array}[]{c}(R_{2}/R_{1})\left(\sin(\phi_{2}-\phi_{1})-(\delta/\gamma)\cos(\phi_{2}-\phi_{1})\right)\\ (R_{1}/R_{2})\left(\sin(\phi_{1}-\phi_{2})-(d/c)\cos(\phi_{1}-\phi_{2})\right)\\ (R_{2}/R_{3})\left(\sin(\phi_{2}-\phi_{3})-(\delta/\gamma)\cos(\phi_{2}-\phi_{3})\right)\end{array}\right)\,.

Because ω1ω2\omega_{1}\neq\omega_{2} we are able to choose the solutions 𝐟1(ϕ)=(0,0,0){\bf f}_{1}(\phi)=(0,0,0) and

𝐠1(ϕ)=1ω1ω2((R2/R1)(cos(ϕ2ϕ1)+(δ/γ)sin(ϕ2ϕ1))(R1/R2)(cos(ϕ1ϕ2)+(d/c)sin(ϕ1ϕ2))(R2/R3)(cos(ϕ2ϕ3)+(δ/γ)sin(ϕ2ϕ3))).{\bf g}_{1}(\phi)=\frac{1}{\omega_{1}-\omega_{2}}\left(\begin{array}[]{r}(R_{2}/R_{1})\left(\cos(\phi_{2}-\phi_{1})+(\delta/\gamma)\sin(\phi_{2}-\phi_{1})\right)\\ -(R_{1}/R_{2})\left(\cos(\phi_{1}-\phi_{2})+(d/c)\sin(\phi_{1}-\phi_{2})\right)\\ (R_{2}/R_{3})\left(\cos(\phi_{2}-\phi_{3})+(\delta/\gamma)\sin(\phi_{2}-\phi_{3})\right)\end{array}\right)\,.

The first normal homological equation

Another short computation allows us to express the projection (1π)𝐆1(1-\pi)\cdot{\bf G}_{1} as

((1π)𝐆1)(ϕ)=(eiϕ1(γ+iδ)(R2/γ)cos(ϕ2ϕ1)eiϕ2(c+id)(R1/c)cos(ϕ1ϕ2)eiϕ3(γ+iδ)(R2/γ)cos(ϕ2ϕ3)).((1-\pi)\cdot{\bf G}_{1})(\phi)=\left(\begin{array}[]{l}e^{i\phi_{1}}(\gamma+i\delta)(R_{2}/\gamma)\cos(\phi_{2}-\phi_{1})\\ e^{i\phi_{2}}(c+id)(R_{1}/c)\cos(\phi_{1}-\phi_{2})\\ e^{i\phi_{3}}(\gamma+i\delta)(R_{2}/\gamma)\cos(\phi_{2}-\phi_{3})\end{array}\right)\,.

This is clearly in the range of N(ϕ)=diag(eiϕ1(γ+iδ),eiϕ2(c+id),eiϕ3(γ+iδ))N(\phi)={\rm diag}(e^{i\phi_{1}}(\gamma+i\delta),e^{i\phi_{2}}(c+id),e^{i\phi_{3}}(\gamma+i\delta)). Thus the first normal homological equation, see (4.17) and (4.15), reads

ω𝐡1(ϕ)+diag(2α,2a,2α)𝐡1(ϕ)=V1(ϕ)=((R2/γ)cos(ϕ2ϕ1)(R1/c)cos(ϕ1ϕ2)(R2/γ)cos(ϕ2ϕ3)).\partial_{\omega}{\bf h}_{1}(\phi)+{\rm diag}(2\alpha,2a,2\alpha){\bf h}_{1}(\phi)=V_{1}(\phi)=\left(\begin{array}[]{l}(R_{2}/\gamma)\cos(\phi_{2}-\phi_{1})\\ (R_{1}/c)\cos(\phi_{1}-\phi_{2})\\ (R_{2}/\gamma)\cos(\phi_{2}-\phi_{3})\end{array}\right)\,.

The solution reads

𝐡1(ϕ)=(R2γ(4α2+(ω1ω2)2)(2αcos(ϕ2ϕ1)+(ω2ω1)sin(ϕ2ϕ1))R1c(4a2+(ω1ω2)2)(2acos(ϕ1ϕ2)+(ω1ω2)sin(ϕ1ϕ2))R2γ(4α2+(ω1ω2)2)(2αcos(ϕ2ϕ3)+(ω2ω1)sin(ϕ2ϕ3))).{\bf h}_{1}(\phi)=\left(\begin{array}[]{l}\frac{R_{2}}{\gamma(4\alpha^{2}+(\omega_{1}-\omega_{2})^{2})}\left(2\alpha\cos(\phi_{2}-\phi_{1})+(\omega_{2}-\omega_{1})\sin(\phi_{2}-\phi_{1})\right)\\ \frac{R_{1}}{c(4a^{2}+(\omega_{1}-\omega_{2})^{2})}\left(2a\cos(\phi_{1}-\phi_{2})+(\omega_{1}-\omega_{2})\sin(\phi_{1}-\phi_{2})\right)\\ \frac{R_{2}}{\gamma(4\alpha^{2}+(\omega_{1}-\omega_{2})^{2})}\left(2\alpha\cos(\phi_{2}-\phi_{3})+(\omega_{2}-\omega_{1})\sin(\phi_{2}-\phi_{3})\right)\end{array}\right)\,.

Second order terms

Let us clarify that we will not solve the second order homological equations completely. Instead, the only second order terms that we compute explicitly are the first and third components 𝐟2(1){\bf f}_{2}^{(1)} and 𝐟2(3){\bf f}_{2}^{(3)} of the second order part 𝐟2{\bf f}_{2} of the reduced phase vector field. As was explained above, this suffices to obtain the desired asymptotic expression for ddt(ϕ1ϕ3)=ε2(𝐟2(1)(ϕ)𝐟2(3)(ϕ))+ε3\frac{d}{dt}(\phi_{1}-\phi_{3})=\varepsilon^{2}\left({\bf f}_{2}^{(1)}(\phi)-{\bf f}_{2}^{(3)}(\phi)\right)+\varepsilon^{3}\ldots.

We first compute the inhomogeneous term 𝐆2{\bf G}_{2} as given in (2.10). Because 𝐅2=0{\bf F}_{2}=0 and 𝐟1=0{\bf f}_{1}=0, we see that

𝐆2=12(𝐅0′′e0)(e1,e1)+(𝐅1e0)e1{\bf G}_{2}=\frac{1}{2}({\bf F}_{0}^{\prime\prime}\circ e_{0})(e_{1},e_{1})+({\bf F}_{1}^{\prime}\circ e_{0})\cdot e_{1}

consists only of two terms. It also turns out that the first of these terms contributes in a rather trivial manner to the phase dynamics at order ε2\varepsilon^{2}. This term can be computed by making use of the expansion

|Rjeiϕj+ε\displaystyle|R_{j}e^{i\phi_{j}}+\varepsilon e1(j)(ϕ)|2(Rjeiϕj+εe1(j)(ϕ))=Rj3eiϕj+εRj2(2e1(j)(ϕ)+e2iϕje1(j)(ϕ)¯)\displaystyle e_{1}^{(j)}(\phi)|^{2}(R_{j}e^{i\phi_{j}}+\varepsilon e_{1}^{(j)}(\phi))=R^{3}_{j}e^{i\phi_{j}}+\varepsilon R_{j}^{2}\left(2e_{1}^{(j)}(\phi)+e^{2i\phi_{j}}\overline{e_{1}^{(j)}(\phi)}\right)
+ε2Rj(2eiϕj|e1(j)(ϕ)|2+eiϕj(e1(j)(ϕ))2)+𝒪(ε3).\displaystyle+\varepsilon^{2}R_{j}\left(2e^{i\phi_{j}}|e_{1}^{(j)}(\phi)|^{2}+e^{-i\phi_{j}}(e_{1}^{(j)}(\phi))^{2}\right)+\mathcal{O}(\varepsilon^{3})\,.

This leads to the formula

12\displaystyle\frac{1}{2} (𝐅0′′(e0(ϕ))(e1(ϕ),e1(ϕ))=\displaystyle({\bf F}_{0}^{\prime\prime}(e_{0}(\phi))(e_{1}(\phi),e_{1}(\phi))\!=
(R1(γ+iδ)(2eiϕ1|e1(1)(ϕ)|2R2(c+id)(2eiϕ2|e1(2)(ϕ)|2R3(γ+iδ)(2eiϕ3|e1(3)(ϕ)|2)=:T1(ϕ)imN(ϕ)+(R1(γ+iδ)eiϕ1(e1(1)(ϕ))2R2(c+id)eiϕ2(e1(2)(ϕ))2R3(γ+iδ)eiϕ3(e1(3)(ϕ))2)=:T2(ϕ).\displaystyle\!\underbrace{\left(\!\!\begin{array}[]{l}R_{1}(\gamma+i\delta)(2e^{i\phi_{1}}|e_{1}^{(1)}(\phi)|^{2}\\ R_{2}(c+id)(2e^{i\phi_{2}}|e_{1}^{(2)}(\phi)|^{2}\\ R_{3}(\gamma+i\delta)(2e^{i\phi_{3}}|e_{1}^{(3)}(\phi)|^{2}\end{array}\!\!\right)}_{=:T_{1}(\phi)\in\,{\rm im}\,N(\phi)}+\underbrace{\left(\!\!\begin{array}[]{l}R_{1}(\gamma+i\delta)e^{-i\phi_{1}}(e_{1}^{(1)}(\phi))^{2}\\ R_{2}(c+id)e^{-i\phi_{2}}(e_{1}^{(2)}(\phi))^{2}\\ R_{3}(\gamma+i\delta)e^{-i\phi_{3}}(e_{1}^{(3)}(\phi))^{2}\end{array}\!\!\right)}_{=:T_{2}(\phi)}\,. (6.17)

It is clear that the first term on the right hand side of (6)—which we called T1(ϕ)T_{1}(\phi)—lies in the range of N(ϕ)N(\phi) because 2Rj|e1(j)(ϕ)|22R_{j}|e_{1}^{(j)}(\phi)|^{2}\in\mathbb{R} for j=1,2,3j=1,2,3. So this first term vanishes when we apply the projection π(ϕ)\pi(\phi).

The projection of the second term on the right hand side of (6)—which we called T2(ϕ)T_{2}(\phi)—can be computed as follows. Recall from (4.11) that e1(ϕ)=e0(ϕ)𝐠1(ϕ)+N(ϕ)𝐡1(ϕ)e_{1}(\phi)=e_{0}^{\prime}(\phi)\cdot{\bf g}_{1}(\phi)+N(\phi)\cdot{\bf h}_{1}(\phi), where e0e_{0}, 𝐠1{\bf g}_{1}, NN and 𝐡1{\bf h}_{1} are given in the formulas above. This can be used to expand, first the (e1(j)(ϕ))2(e_{1}^{(j)}(\phi))^{2}, and then T2(ϕ)T_{2}(\phi) in trigonometric polynomials. It is not very hard to see that this must yield a formula of the form

π(ϕ)T2(ϕ)=(R1ieiϕ1(C+Dsin(2ϕ22ϕ1)+Ecos(2ϕ22ϕ1))R2ieiϕ2(C~+D~sin(2ϕ22ϕ1)+E~cos(2ϕ22ϕ1))R3ieiϕ3(C+Dsin(2ϕ22ϕ3)+Ecos(2ϕ22ϕ3)))\displaystyle\pi(\phi)T_{2}(\phi)=\left(\begin{array}[]{l}R_{1}ie^{i\phi_{1}}\left(C+D\sin(2\phi_{2}-2\phi_{1})+E\cos(2\phi_{2}-2\phi_{1})\right)\\ R_{2}ie^{i\phi_{2}}\left(\tilde{C}+\tilde{D}\sin(2\phi_{2}-2\phi_{1})+\tilde{E}\cos(2\phi_{2}-2\phi_{1})\right)\\ R_{3}ie^{i\phi_{3}}\left(C+D\sin(2\phi_{2}-2\phi_{3})+E\cos(2\phi_{2}-2\phi_{3})\right)\end{array}\right) (6.21)

for certain real numbers C,D,E,C~,D~,E~C,D,E,\tilde{C},\tilde{D},\tilde{E} that we shall not explicitly compute here. Note that this clearly lies in the range of e0(ϕ)e_{0}^{\prime}(\phi). It follows that

U21st(ϕ)=(C+Dsin(2ϕ22ϕ1)+Ecos(2ϕ22ϕ1)C~+D~sin(2ϕ22ϕ1)+E~cos(2ϕ22ϕ1)C+Dsin(2ϕ22ϕ3)+Ecos(2ϕ22ϕ3))\displaystyle U_{2}^{1{\rm st}}(\phi)=\left(\begin{array}[]{l}C+D\sin(2\phi_{2}-2\phi_{1})+E\cos(2\phi_{2}-2\phi_{1})\\ \tilde{C}+\tilde{D}\sin(2\phi_{2}-2\phi_{1})+\tilde{E}\cos(2\phi_{2}-2\phi_{1})\\ C+D\sin(2\phi_{2}-2\phi_{3})+E\cos(2\phi_{2}-2\phi_{3})\end{array}\right) (6.25)

is the first part of the inhomogeneous right hand side of the second tangential homogeneous equation ω𝐠𝟐+𝐟2=U2\partial_{\omega}{\bf g_{2}}+{\bf f}_{2}=U_{2}. Because 2ω12ω22\omega_{1}\neq 2\omega_{2}, only the constant part (C,C~,C)(C,\tilde{C},C) of this U21st(ϕ)U_{2}^{1{\rm st}}(\phi) is resonant; all other terms can be absorbed in 𝐠2{\bf g}_{2}. Thus the resonant normal form of this part of 𝐟2{\bf f}_{2} is (C,C~,C)T(C,\tilde{C},C)^{T}. As this constant vector field does not contribute to ddt(ϕ1ϕ3)\frac{d}{dt}\left(\phi_{1}-\phi_{3}\right), we compute neither CC nor C~\tilde{C} explicitly.

We proceed by considering the other term in 𝐆2{\bf G}_{2}, namely (𝐅1e0)e1({\bf F}^{\prime}_{1}\circ e_{0})\cdot e_{1}. Recalling that 𝐅1(z)=(z2,z1,z2){\bf F}_{1}(z)=(z_{2},z_{1},z_{2}), we see that this term equals

𝐅1(e0(ϕ))e1(ϕ)=(e1(2)(ϕ)e1(1)(ϕ)e1(2)(ϕ))=(eiϕ2(iR2𝐠1(2)(ϕ)+(c+id)𝐡1(2)(ϕ))eiϕ1(iR1𝐠1(1)(ϕ)+(γ+iδ)𝐡1(1)(ϕ))eiϕ2(iR2𝐠1(2)(ϕ)+(c+id)𝐡1(2)(ϕ))).{\bf F}_{1}^{\prime}(e_{0}(\phi))\cdot e_{1}(\phi)=\left(\begin{array}[]{c}e_{1}^{(2)}(\phi)\\ e_{1}^{(1)}(\phi)\\ e_{1}^{(2)}(\phi)\end{array}\right)=\left(\begin{array}[]{l}e^{i\phi_{2}}(iR_{2}{\bf g}_{1}^{(2)}(\phi)+(c+id){\bf h}_{1}^{(2)}(\phi))\\ e^{i\phi_{1}}(iR_{1}{\bf g}_{1}^{(1)}(\phi)+(\gamma+i\delta){\bf h}_{1}^{(1)}(\phi))\\ e^{i\phi_{2}}(iR_{2}{\bf g}_{1}^{(2)}(\phi)+(c+id){\bf h}_{1}^{(2)}(\phi))\end{array}\right)\,.

Using the expressions for π(ϕ)\pi(\phi), 𝐠1(ϕ){\bf g}_{1}(\phi) and 𝐡1(ϕ){\bf h}_{1}(\phi) provided above, one can compute that the projection of this term has the form

π(ϕ)𝐅1(e0(ϕ))e1(ϕ)=(iR1eiϕ1000iR2eiϕ2000iR3eiϕ3)U22nd(ϕ),\displaystyle\pi(\phi)\cdot{\bf F}_{1}^{\prime}(e_{0}(\phi))\cdot e_{1}(\phi)=\left(\begin{array}[]{ccc}iR_{1}e^{i\phi_{1}}&0&0\\ 0&iR_{2}e^{i\phi_{2}}&0\\ 0&0&iR_{3}e^{i\phi_{3}}\end{array}\right)\cdot U_{2}^{2{\rm nd}}(\phi)\,, (6.29)

in which now

U22nd(ϕ)=(B+Fsin(2ϕ12ϕ2)+Gcos(2ϕ12ϕ2)B~+F~sin(2ϕ12ϕ2)+G~cos(2ϕ12ϕ2){Asin(ϕ1ϕ3)+Bcos(ϕ1ϕ3)+Fsin(ϕ1+ϕ32ϕ2)+Gcos(ϕ1+ϕ32ϕ2)}).\displaystyle U_{2}^{2{\rm nd}}(\phi)=\left(\begin{array}[]{l}B+F\sin(2\phi_{1}-2\phi_{2})+G\cos(2\phi_{1}-2\phi_{2})\\ \tilde{B}+\tilde{F}\sin(2\phi_{1}-2\phi_{2})+\tilde{G}\cos(2\phi_{1}-2\phi_{2})\\ \left\{\begin{array}[]{l}A\sin(\phi_{1}-\phi_{3})+B\cos(\phi_{1}-\phi_{3})\\ +F\sin(\phi_{1}+\phi_{3}-2\phi_{2})+G\cos(\phi_{1}+\phi_{3}-2\phi_{2})\end{array}\right\}\end{array}\right)\,. (6.35)

With some effort the constants AA and BB can be computed by hand, yielding

A=14a2+(ω1ω2)2(δγ(ω2ω1)+a(1+dδcγ)+2a2(dc+δγ)1ω2ω1),B=14a2+(ω1ω2)2((ω2ω1)+a(dcδγ)+2a2(1dδcγ)1ω2ω1).\displaystyle\begin{array}[]{ll}A&\!\!\!=\frac{1}{4a^{2}+(\omega_{1}-\omega_{2})^{2}}\left(\frac{\delta}{\gamma}(\omega_{2}-\omega_{1})+a\left(1+\frac{d\delta}{c\gamma}\right)+2a^{2}\left(\frac{d}{c}+\frac{\delta}{\gamma}\right)\frac{1}{\omega_{2}-\omega_{1}}\right)\,,\\ B&\!\!\!=\frac{1}{4a^{2}+(\omega_{1}-\omega_{2})^{2}}\left((\omega_{2}-\omega_{1})+a\left(\frac{d}{c}-\frac{\delta}{\gamma}\right)+2a^{2}\left(1-\frac{d\delta}{c\gamma}\right)\frac{1}{\omega_{2}-\omega_{1}}\right)\,.\end{array} (6.38)

We did not compute any of the other constants. As ω1=ω3ω2\omega_{1}=\omega_{3}\neq\omega_{2}, the resonant part of U22nd(ϕ)U_{2}^{2{\rm nd}}(\phi) is given by 𝐟2(ϕ)=(B,B~,Asin(ϕ1ϕ3)+Bcos(ϕ1ϕ3))T{\bf f}_{2}(\phi)=(B,\tilde{B},A\sin(\phi_{1}-\phi_{3})+B\cos(\phi_{1}-\phi_{3}))^{T}. The other terms in U22nd(ϕ)U_{2}^{2{\rm nd}}(\phi) can be absorbed into 𝐠2{\bf g}_{2} when solving the tangential homological equation ω𝐠𝟐+𝐟2=U2\partial_{\omega}{\bf g_{2}}+{\bf f}_{2}=U_{2}.

Conclusion

To summarise, we computed that 𝐟1(ϕ)=(0,0,0)T{\bf f}_{1}(\phi)=(0,0,0)^{T} and

𝐟2(ϕ)=(B+CB~+C~Asin(ϕ1ϕ3)+Bcos(ϕ1ϕ3)+C).\displaystyle{\bf f}_{2}(\phi)=\left(\begin{array}[]{c}B+C\\ \tilde{B}+\tilde{C}\\ A\sin(\phi_{1}-\phi_{3})+B\cos(\phi_{1}-\phi_{3})+C\end{array}\right)\,. (6.42)

The constants AA and BB are given in (6.38), but we did not compute B~,C\tilde{B},C or C~\tilde{C}. Because ω1=ω3\omega_{1}=\omega_{3} and ϕ˙=ω+ε𝐟1(ϕ)+ε2𝐟2(ϕ)+𝒪(ε3)\dot{\phi}=\omega+\varepsilon{\bf f}_{1}(\phi)+\varepsilon^{2}{\bf f}_{2}(\phi)+\mathcal{O}(\varepsilon^{3}), we conclude that

ddt(ϕ1ϕ3)=ε2(Asin(ϕ1ϕ3)Bcos(ϕ1ϕ3)+B)+𝒪(ε3).\displaystyle\frac{d}{dt}(\phi_{1}-\phi_{3})=\varepsilon^{2}\left(-A\sin(\phi_{1}-\phi_{3})-B\cos(\phi_{1}-\phi_{3})+B\right)+\mathcal{O}(\varepsilon^{3})\,. (6.43)

This is exactly equation (6.5).

7 Acknowledgements

We would like to thank Edmilson Roque and Deniz Eroglu for useful tips regarding numerics. S.v.d.G. was partially funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)––453112019. E.N. was partially supported by the Serrapilheira Institute (Grant No. Serra-1709-16124). B.R. acknowledges funding and hospitality of the Sydney Mathematical Research Institute.

References

  • [1] A. Arenas, A. Díaz-Guilera, J. Kurths, Y. Moreno, and C. Zhou, Synchronization in complex networks, Phys. Rep. 469 (2008), no. 3, 93–153.
  • [2] P. Ashwin and A. Rodrigues, Hopf normal form with SN symmetry and reduction to systems of nonlinearly coupled phase oscillators, Physica D 325 (2016), 14–24.
  • [3] A. Bergner, M. Frasca, G. Sciuto, A. Buscarino, E. Ngamga, L. Fortuna, and J. Kurths, Remote synchronization in star networks, Phys. Rev. E 85 (2012), 026208.
  • [4] C. Bick, T. Böhle, and C. Kühn, Higher-order interactions in phase oscillator networks through phase reductions of oscillators with phase dependent amplitude, arXiv (2023), 2305.04277.
  • [5] J. Buck and E. Buck, Mechanism of rhythmic synchronous flashing of fireflies, Science 159 (1968), no. 3821, 1319–1327.
  • [6] M. Canadell and A. Haro, Computation of quasiperiodic normally hyperbolic invariant tori: Rigorous results, J. Nonlin. Sci. 27 (2017), 1–36.
  • [7] R. de la Llave, A. González, A. Jorba, and J. Villanueva, KAM theory without action-angle variables, Nonlinearity 18 (2005), no. 2, 855–895.
  • [8] N. Fénichel, Geometric singular perturbation theory for ordinary differential equations, J. Differ. Eq. 31 (1979), 53–98.
  • [9] G. Floquet, Sur les équations différentielles linéaires à coefficients périodiques, Ann. Sci. Éc. Norm. Supér. 12 (1883), 47–88.
  • [10] E. Gengel, E. Teichmann, M. Rosenblum, and A. Pikovsky, High-order phase reduction for coupled oscillators, Journal of Physics: Complexity 2 (2021), no. 1, 015005.
  • [11] D. Golomb, D. Hansel, and G. Mato, Mechanisms of synchrony of neural activity in large networks, Handbook of Biological Physics, vol. 4, Elsevier, 2001, p. 887–968.
  • [12] J. Guckenheimer, Isochrons and phaseless sets, Journal of Mathematical Biology 1 (1975), no. 3, 259–273.
  • [13] A. Haro, M. Canadell, J. Figueras, A. Luque, and J. Mondelo, The parameterization method for invariant manifolds, Springer, 2016.
  • [14] R. Johnson and G. Sell, Smoothness of spectral subbundles and reducibility of quasi-periodic linear differential systems, J. Differ. Eq. 41 (1981), no. 2, 262–288.
  • [15] A. Jorba and C. Simó, On the reducibility of linear differential equations with quasiperiodic coefficients, J. Differ. Eq. 98 (1992), no. 1, 111–124.
  • [16] M. Kumar and M. Rosenblum, Two mechanisms of remote synchronization in a chain of Stuart-Landau oscillators, Phys. Rev. E 104 (2021), 054202.
  • [17] Y. Kuramoto, Chemical oscillations, waves, and turbulence, Springer, Berlin, Heidelberg, 1984.
  • [18] I. León and D. Pazó, Phase reduction beyond the first order: The case of the mean-field complex Ginzburg-Landau equation, Phys. Rev. E 100 (2019), no. 1, 012211.
  • [19] I. Lizarraga, B. Rink, and M. Wechselberger, Multiple timescales and the parametrisation method in geometric singular perturbation theory, Nonlinearity 34 (2021), no. 6, 4163–4201.
  • [20] 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.
  • [21] H. Nakao, Phase reduction approach to synchronisation of nonlinear oscillators, Contemporary Physics 57 (2016), no. 2, 188–214.
  • [22] E. Nijholt, J. Ocampo-Espindola, D. Eroglu, I. Kiss, and T. Pereira, Emergent hypernetworks in weakly coupled oscillators, Nature Comm. 13 (2022), no. 1, 4849.
  • [23] N. Ognjanovski, S. Schaeffer, and J. Wu, Parvalbumin-expressing interneurons coordinate hippocampal network dynamics required for memory consolidation, Nature Comm. 8 (2017), 15039.
  • [24] J. Palva, S. Palva, and K. Kaila, Phase synchrony among neuronal oscillations in the human cortex, J. Neurosci. 25 (2005), no. 15, 3962–3972.
  • [25] B. Pietras and A. Daffertshofer, Network dynamics of coupled oscillators and phase reduction techniques, Phys. Rep. 819 (2019), 1–105.
  • [26] A. Pikovsky, M. Rosenblum, and J. Kurths, Synchronization; a universal concept in nonlinear sciences, Cambridge University Press, 2001.
  • [27] M. Rosenblum and A. Pikovsky, Numerical phase reduction beyond the first order approximation, Chaos 29 (2019), 011105.
  • [28] J.A. Sanders, F. Verhulst, and J. Murdock, Averaging methods in nonlinear dynamical systems, second ed., Applied Mathematical Sciences, vol. 59, Springer, New York, 2007.
  • [29] G. Teschl, Ordinary differential equations and dynamical systems, American Mathematical Society, 2012.
  • [30] M. Wechselberger, Geometric singular perturbation theory beyond the standard form, Springer, 2020.