We introduce a new mathematical framework for the qualitative analysis of dynamical stability, designed particularly for finite-time processes subject to slow-timescale external influences. In particular, our approach is to treat finite-time dynamical systems in terms of a slow–fast formalism in which the slow time only exists in a bounded interval, and consider stability in the singular limit. Applying this to one-dimensional phase dynamics, we provide stability definitions somewhat analogous to the classical infinite-time definitions associated with Aleksandr Lyapunov. With this, we mathematically formalize and generalize a phase-stabilization phenomenon previously described in the physics literature for which the classical stability definitions are inapplicable and instead our new framework is required.

An important phenomenon in nonlinear science is the stabilization of processes by external forces. This phenomenon can manifest as the mutual synchronization of an array of processes evolving under the same law but starting at different initial states, where the synchronization is due to the processes being subject to a common external forcing. Such stabilization is traditionally analyzed in terms of Lyapunov exponents and other related long-time-asymptotic conceptions of dynamical stability; these can be applied to both deterministic and stochastic dynamical systems that are well-defined over infinite time. References 1 and 2 present a stabilization phenomenon in which the kind of external forcing considered is a slow-timescale process; our present paper concerns mathematically rigorous formulation, proof, and generalization of this stabilization phenomenon. We will see that the traditional approaches to theoretical stability analysis such as Lyapunov exponents and other infinite-time approaches cannot generally be applied to describe the stabilization phenomenon of Refs. 1 and 2. In fact, the critical forcing strength required to induce stability is generally dependent on the time-interval over which the system is being considered. This means that the phenomenon should be understood in terms of finite-time stability, in contrast to other stabilization phenomena such as the well-known case of stabilization by stationary noise. Our work represents a significant contribution to the important and growing field of finite-time dynamical systems (FTDSs) and finite-time stability, whose necessity in the study of various nonlinear systems, such as climate and biological systems, is increasingly being recognized.

For either an autonomous dynamical system

(1)

with a time-independent vector field F:XTX, or a nonautonomous3–5 dynamical system

with a time-dependent vector field F:X×RTX, dynamics is traditionally analyzed in terms of behavior as t±. For example, attractors, repellers, asymptotic stability, and Lyapunov exponents are typically defined in terms of such long-time-asymptotic behavior. From a practical perspective, this approach is based on the assumption that one is interested in the system’s behavior after a sufficiently long time that “transients have decayed.”

However, in more recent decades, the study of “finite-time dynamical systems” (FTDSs) has grown in popularity;6–14 these are dynamical systems in which time is constrained to a compact interval, i.e.,

for some time-dependent vector field F:X×[0,T]TX. As described in Secs. 1.1 and 6 of Ref. 11, situations where the FTDS framework is appropriate include

  • when one is interested in behavior on a shorter timescale than that required for a given model’s long-time-asymptotic dynamical picture to emerge, or

  • when the temporal variations of the external forcing over finite timescales of interest are “unextendible,” meaning that they do not (or are not know to) follow any particular indefinite-time deterministic or stochastic model, and so there is no such thing as “long-time-asymptotic” behavior.

Now sometimes, the temporal variations of the external forcing may be relatively slow, and this slowness may play an important role in shaping the consequent dynamical behavior.15 In this paper:

  • We will introduce (immediately below, and further in Sec. III) a framework for investigating the role of slow variation of external forcing on the consequent dynamical behavior, specifically designed to be applicable to “unextendible” temporal variations defined over compact time-intervals. (The need for such a framework will be further illustrated in Sec. II B.)

  • We will apply this framework to provide a rigorous mathematical description of a certain one-dimensional phase-stabilization phenomenon described in the physics literature, in Refs. 1 and 2.

The framework we introduce is a slow–fast setup in which the slow time is constrained to a compact interval (without loss of generality, the unit interval). That is, we consider systems of the form

(2)

for some [0,1]-parameterized vector field,

While traditional mathematical definitions of stability and Lyapunov exponents are formulated in terms of the t behavior, we will formulate analogous definitions in terms of the ε0 behavior of (2). While a classical t limit represents “non-transient dynamical phenomena,” the ε0 limit represents “non-rate-induced16 dynamical phenomena.” Here, ε represents the “timescale separation” between the slower timescale of the external forcing and the faster timescale of the internal dynamics; this setup can equivalently take the form

(3)

Now, in investigating the stability properties of (2) and (3), this paper focuses on the natural “starting case” to consider, namely, dynamics on a one-dimensional phase space (specifically, the circle X=S1). This already provides significant non-trivialities to deal with (see Sec. II C), especially the canard phenomenon17 of multiple-timescale dynamics where solutions spend a long time traveling along an unstable portion of the slow manifold. Our results in this paper will both formalize and considerably generalize the core findings of Refs. 1 and 2 regarding stabilization of the one-dimensional Adler equation by slow-timescale additive forcing. The two main steps that will be involved are

  • showing that “generic” time-dependent vector fields F:S1×[0,1]TS1 admit an adiabatically traceable moving sink analogous to the moving sink of the slowly varying Adler equation (Theorem 14); and

  • making mathematically rigorous the heuristic stability arguments used in Refs. 1 and 2 for the Adler equation (Theorem 24), which we mainly achieve by adapting methods that were used in Ref. 17 for investigating periodic orbits of an Adler equation subject to low-frequency sinusoidal forcing.

We will also analyze the generic behavior of the transition to stability from neutral stability induced by slow-timescale forcing, and its contrast to analogous classical saddle-node bifurcations (see Secs. III G, IV D, and IV E 3). Let us also mention that fast-timescale forcing of the Adler equation as a FTDS has been considered in Theorem 6.3 of Ref. 18, as part of a study of phase-synchronization transitions.

From a physical perspective, dynamical systems on S1 can be used to describe the phase of a cyclic process via phase reduction, which has been established for systems subject to slow-timescale forcing in Refs. 19–21. The study of one-dimensional phase dynamics has particular relevance for various biological oscillatory processes or coupled pairs of oscillatory processes;1,2,18,21–23 and indeed, in such contexts, it is especially important that temporal variations in external forcing are “unextendible” in the sense described above, otherwise, the organism cannot properly interact with the outside world and will quickly die.

However, we repeat that the rationale behind this paper is not simply to obtain results about one-dimensional phase dynamics. We hope through our treatment of the one-dimensional case of (2) to help lay the foundations for the development of more general methodology for treating FTDSs involving multiple timescales. The need for theory of higher-dimensional dynamics within the framework of Eqs. (2) and (3) is already evidenced in Sec. III of Ref. 2, where numerics indicate that the phenomenon of stabilization of solutions by slowly varying forcing can also occur in higher-dimensional systems, including chaotic systems. Additionally, Ref. 24 shows that the same mechanism of stabilization described in Refs. 1 and 2 for one-dimensional dynamics can also induce local stability of synchronous solutions in a driven Kuramoto network of arbitrary size.

So, we anticipate that the line of inquiry initiated by this paper may find utility in the investigation of various complex systems such as biological oscillator networks (where our earlier comment about the necessity of “unextendible” temporal variations still applies) and climate systems subject to natural or anthropogenic forcing. Furthermore, we believe that mathematical theory of multiple-timescale FTDSs can provide deeper insight into the nature and application of time-series analysis tools designed for studying time-localized dynamics, such as those described in Refs. 5, 25, and 26.

The paper is organized as follows.

In Sec. II, we describe the stabilization of the Adler equation presented in Refs. 1 and 2, its physical motivations, and its contrast to the traditional cases of stabilization by constant forcing and by noise. We illustrate how the stability and neutral stability observed in Refs. 1 and 2 cannot generally be formulated in terms of the traditional definitions and quantification of stability due to Aleksandr Lyapunov, such as asymptotic stability and asymptotic Lyapunov exponents. We then outline the primary mathematical non-trivialities for our main task of providing a mathematically rigorous formulation and proof of the stabilization phenomenon.

In Sec. III, we first introduce our concepts of stability for systems of the form (2) defined on the circle X=S1. Then, we present other definitions necessary for the formulation of our main results (such as curves of the stable and unstable slow manifolds and tracking of such curves) and present our main results that mathematically formalize and generalize the basic argument of Refs. 1 and 2. We also outline our main methods of proof and their relation to the methods in Ref. 17. Finally, we give a result regarding the generic scenario of a transition between stability and neutral stability within our framework, highlighting, in particular, its signature characteristic of approximately linear growth of the stability exponent on the stable side of the transition (in contrast to the square-root growth in classical saddle-node bifurcations).

In Sec. IV, we carry out a comparison of our new framework of stability analysis with the classical long-time-asymptotic framework, particularly through the example of low-frequency-periodically forced Adler equations.

In Sec. V, we prove the results in Sec. III.

In Sec. VI, we summarize and conclude.

In  Appendix A, for the sake of completeness, we obtain a much more refined version of one of the results (Corollary 45) needed within the proof of Theorem 24(C).

In  Appendix B, we state the numerical procedures used for all the numerics in this paper.

A relatively non-technical presentation of some of the discussions in this paper (in particular, in Secs. II A, II B, IV B 1, IV B 3, IV C, and IV E 1) has been published in Ref. 27 (which is itself the sequel to an exposition of phase synchronization by time-dependent driving given in Ref. 28).

In Sec. II A, we first give preliminaries regarding the general concept of stabilization by external forcing and two basic mechanisms of stabilization of the Adler equation; then, we describe the stabilization by slow-timescale forcing studied in Refs. 1 and 2. In Sec. II B, we will demonstrate an example of this latter mechanism of stabilization, where we use a slow-timescale forcing that is “unextendible” in the sense described in Sec. I. In Sec. II C, we will outline the main non-trivialities in obtaining a mathematically rigorous formulation and proof of this stabilization phenomenon.

It is well-known that trajectories of an autonomous dynamical system (1) can often be “stabilized” by the introduction of external forcing, e.g., an additive forcing ξ(t) giving rise to a nonautonomous dynamical system of the form

(4)

This “stabilization” can be defined in terms of synchronization between solutions starting at different initial conditions:29 namely, it refers to the scenario that for an array of distinct initial conditions x1,,xn, the solutions of (1) starting at these initial conditions remain clearly distinct at all times but the solutions of (4) starting at these same initial conditions become synchronized with each other. Note that this synchronization is not induced by any “direct” coupling between the solutions but rather by the “indirect coupling” of being simultaneously subject to the same external driving ξ(t). A natural question then is what kinds of forcing function ξ(t) will give rise to such synchronization.

As a prototype: on the circle S1=R/(2πZ), all solutions of the autonomous Adler equation

(5)

with constant parameters k>a>0, move strictly periodically round the circle with common period 2πk2a2; and so, in particular, the solutions of any two distinct initial conditions do not synchronize with each other. We now consider what types of external forcing can cause the solutions of different initial conditions to become synchronized with each other.

1. Stabilization by constant forcing

One crude way to stabilize system (5) is to apply a perpetually constant external forcing ξ such that the “effective new k-value” knew=k+ξ lies in (a,a). In this case, all solutions now converge to the point yknew:=arcsin(knewa) apart from one single exceptional solution that remains fixed at the point πyknew; and thus, the phase differences between these solutions trivially tend to 0 as t. This mutual synchronization of solutions takes place at an exponential rate with exponent a2knew2, called the Lyapunov exponent associated with the fixed point yknew. [This exponent is obtained as the derivative of asin()+knew at yknew.]

Note that if one were instead to take the constant ξ sufficiently close to zero, then the Adler equation would not be stabilized: solutions would move strictly periodically round the circle with common period 2πknew2a2. The bifurcation that takes place as ξ crosses the critical threshold ak is a saddle-node bifurcation, namely, the creation of a stable and an unstable equilibrium point.

Now, from a practical point of view, this crude approach to the removal of asynchrony among solutions through simply forcing them perpetually to remain at a fixed point may (depending on context) be problematic due to the inability of solutions to move freely throughout the phase space. It may also (again depending on context) be rather energy-inefficient to keep this forcing perpetually maintained.

2. Stabilization by noise

One kind of forcing that is often known to induce stability in dynamical systems is noise.29–31 Noise-induced stability is very common in general but is especially typical for one-dimensional phase-oscillator dynamics, where, for example, memoryless stationary noise is guaranteed to lead to mutual synchronization of solutions under weak conditions.32,33

As an example, the addition of stationary Gaussian white noise of intensity A>0 to the Adler equation (5) will stabilize the Adler equation, regardless of how small A is.29,34 To be more precise: if we fix the values of k, a and A, almost every realization W:[0,)R of a Wiener process has the property that all solutions of the corresponding nonautonomous dynamical system

(6)

synchronize with each other in the limit as t (i.e., their phase differences tend to 0), apart from one solution starting at a single exceptional initial condition. Moreover, this synchronization of solutions comes with a well-defined Lyapunov exponent (i.e., long-time-asymptotic exponential rate of decay of phase differences), which is exactly the same value across almost all realizations W of the Wiener process.

The mechanism behind this exponential synchronization of solutions is that the infinite time-horizon allows for statistical large-number laws to come into play, whereby random events in the noise which bring the different solutions toward those regions on the circle where they are pushed closer together are guaranteed to occur with sufficient long-time-asymptotic regularity. Note, in particular, that if the noise intensity A is very small, then it will take a very long time for the synchrony among different solutions to begin to emerge.

3. The stabilization phenomenon of Refs. 1 and 2

In contrast to stationary white noise, a very different type of forcing that one can consider is deterministic35 slowly varying forcing. In particular, in the physics literature, Refs. 1 and 2 considered equations of the form

(7)

with G(t) having slow dependence on t. In both of these papers, the variable θ(t) in (7) is regarded as representing the phase difference between a unidirectionally coupled pair of phase oscillators (where the driving oscillator has time-variable frequency), so that synchronization between the solutions of (7) starting at different initial conditions corresponds to stabilization of the driven oscillator by the driving oscillator. Physical motivations in Ref. 1 include the van der Pol oscillator subject to aperiodic external driving [where a phase-reduction approximation brings the system to the form (7)], as well as neuron voltage spiking dynamics, where stabilization by external driving is known to occur.36–38 One physical motivation behind Ref. 2 was to develop upon some of the main findings of Refs. 22 and 23 for cardiorespiratory phase synchronization.

The foundational point of Refs. 1 and 2 is essentially the following argument:

  • If there are times t at which the slowly varying function G(t) lies in the interval (a,a), then during such times, the solutions of (7) will cluster together into an increasingly tight cluster around the “slowly moving sink” yG(t)=arcsin(G(t)a).

  • By contrast, while G(t) is outside the interval [a,a], solutions move unboundedly round the circle without either synchronizing with each other or becoming desynchronized from each other.

  • Thus, overall, we have mutual synchronization of solutions if there are times when G(t) enters the interval (a,a), but not if G(t) stays outside the interval [a,a]. (We will illustrate this further in Sec. II B.)

As in Ref. 1, we can derive an approximation Λ of the average exponential rate of mutual synchronization of solutions, by adiabatically following the Lyapunov exponent of the stable fixed point yG(t) of the time-frozen vector field asin()+G(t) when it exists. More precisely, over a time-interval [0,T], the approximate average exponential rate of synchronization (taken as negative when such synchronization does occur) is given by

(8)

Note that Λ is 0 if G(t) never enters the interval (a,a).

Now, as in Ref. 2, we can consider G(t) of the form G(t)=k+Ag(t) for some k>a, some function g(t) taking both positive and negative values, and some parameter A0; so (7) becomes

(9)

This represents the addition of an external forcing +Ag(t) to the autonomous system (5). Parameter A is somewhat analogous to the noise intensity parameter in (6), except that we are now considering slow-timescale deterministic variation rather than fast-timescale stochastic variation. Fixing the function g, if A is sufficiently large, then there will be times t at which k+Ag(t) enters the interval (a,a), and so the added forcing +Ag(t) will induce stabilization.

References 1 and 2 also provide strong numerical support for the above description of the dynamics of (7) and (9). The numerics in Ref. 2 use low-frequency sinusoidal forcing (see also Secs. IV C and IV E 2 of this paper), and Ref. 1 considers various other forms of forcing. In Sec. II B, we will go on to provide further numerical support using a form of forcing deliberately designed to be “unextendible” in the sense of point (ii) in the Introduction.

To illustrate the need for a FTDS framework in order to be able to formulate mathematically the stabilization phenomenon described in Refs. 1 and 2, we consider Eq. (9) with g(t) being a function that is only meaningfully defined on a compact time-interval.

One way in which such a function g(t) could arise is as an empirically recorded time-series:11 if one defines a nonautonomous dynamical system with reference to an actual empirically measured signal for the external forcing, then the finiteness of the signal duration means that the system is inherently a FTDS, for which traditional infinite-time dynamical systems theory cannot be used as a framework for dynamics analysis. Forcing a fast-timescale Adler equation by a finite-duration empirical signal measured from a relatively slow-timescale physical process will give results in line with the picture described in Sec. II A 3.

But here, we will construct g(t) numerically. Specifically, we will generate a sample realization of a Brownian bridge—this is a finite-time stochastic process, which represents the inhomogeneities of a large sample of random times selected with uniform probability from a pre-specified compact time-interval. Then, as in Ref. 1, Sec. III B, we will pass the result through a low-pass filter to obtain our slowly time-dependent forcing. Our choice of using a sample realization of a Brownian bridge is not motivated by any specific physical application but is simply for illustrative purposes as a good example of “unextendible” temporal variations.

1. The forcing and the system parameters

Let us first briefly recall the definition of a Brownian bridge, and its connection to Brownian motion. A Brownian bridge of parameter σ>0 on a time-interval [0,T] is a finite-time stochastic process (Bt)t[0,T] approximated by taking a very large simple random sample of times S={t1,,tN} from the uniform distribution on [0,T] and then setting

where E() is the “signed deviation from homogeneity” of the random sample S, that is,

This definition is made rigorous by Donsker’s theorem.39 Now, although a Brownian motion(Mt)t0 is an indefinite-time process, one can construct up to time T a zero-drift Brownian motion Mt of diffusion parameter σ by taking a random variable YN(0,σ2T) that is independent of the Brownian bridge (Bt)t[0,T] and then letting

(10)

For each t[0,T], the variance of Bt is given by

This quadratic expression with respect to t is non-negative on [0,T] but becomes negative when t is outside of [0,T]. Thus, there is no natural way to extend a Brownian bridge on [0,T] beyond time T. [And indeed, there is no way to extend the construction (10) beyond time T, as the variance of tY then exceeds the variance of Mt.]

Now, in our case, we define g(t) on a compact interval [0,T] with T=2π×105s; we constructed g(t) by first generating a sample realization of a Brownian bridge on [0,T] of parameter σ=1T1.3×103s12 and then passing the result through a fifth order Butterworth low-pass filter with cut-off frequency 1/(2π×103)1.6×104 Hz. The resulting function is shown is Fig. 1. Further details of this construction and of all numerics in this paper are given in  Appendix B. For other parameters in Eq. (9), we take a=13rad/s and k=1rad/s, and we consider how the dynamics depends on the parameter A0. Due to the low-pass filter, g(t) has very slow dependence on t compared to the timescale of the “internal dynamics” of the system as represented by Eq. (5), whose solutions have a frequency of k2a22π0.15 Hz.

FIG. 1.

Graph of g(t) obtained by passing a sample realization of a Brownian bridge on [0s,2π×105s] through a low-pass filter of cut-off frequency 1/(2π×103) Hz. Reproduced with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

FIG. 1.

Graph of g(t) obtained by passing a sample realization of a Brownian bridge on [0s,2π×105s] through a low-pass filter of cut-off frequency 1/(2π×103) Hz. Reproduced with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

Close modal

2. The stabilization phenomenon

Considering (9) over the whole time-interval [0,T], the argument described in Sec. II A 3 tells us that if A is less than A:=akmint[0,T]g(t) then the solutions of (9) do not synchronize with each other, but if A is greater than A then the solutions of (9) do synchronize with each other. This is exactly what we observe in numerics: Fig. 2(b) shows a “numerical bifurcation diagram” of (9) where for each A-value, the trajectories at time T of 50 evenly spaced initial conditions are shown; and in dashed black is marked the value of A.

FIG. 2.

Dynamics of (9) with g as in Fig. 1, with varying A, over the time-interval [0s,2π×105s] of duration T=2π×105s in (a)–(c), and over the shorter time-interval [0s,π×104s] of duration T=π×104s in (d)–(f). Other parameters are a=13 rad/s and k=1rad/s. In (a), (b), (d), and (e), for each A-value, results for the evolution θ(t) of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49, are shown: (a) and (d) show the finite-time Lyapunov exponents λT, as defined by (11), for these trajectories, with T=T in (a) and with T=T in (d); (b) and (e) show the positions θ(T) of these trajectories at time T=T in (b) and T=T in (e). In (c) and (f), for each A-value, the positions of θ(0) for the 50 trajectories ending at the points θ(T)=2πi50, i=0,,49, are shown, with T=T in (c) and T=T in (f). In (a)–(c), the value A:=akmint[0,T]g(t) is marked in dashed black. In (d)–(f), the value A:=akmint[0,T]g(t) is marked in dashed black. Reproduced (with minor modification) with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

FIG. 2.

Dynamics of (9) with g as in Fig. 1, with varying A, over the time-interval [0s,2π×105s] of duration T=2π×105s in (a)–(c), and over the shorter time-interval [0s,π×104s] of duration T=π×104s in (d)–(f). Other parameters are a=13 rad/s and k=1rad/s. In (a), (b), (d), and (e), for each A-value, results for the evolution θ(t) of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49, are shown: (a) and (d) show the finite-time Lyapunov exponents λT, as defined by (11), for these trajectories, with T=T in (a) and with T=T in (d); (b) and (e) show the positions θ(T) of these trajectories at time T=T in (b) and T=T in (e). In (c) and (f), for each A-value, the positions of θ(0) for the 50 trajectories ending at the points θ(T)=2πi50, i=0,,49, are shown, with T=T in (c) and T=T in (f). In (a)–(c), the value A:=akmint[0,T]g(t) is marked in dashed black. In (d)–(f), the value A:=akmint[0,T]g(t) is marked in dashed black. Reproduced (with minor modification) with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

Close modal
  • For A<A, we see the trajectories distributed fairly evenly distributed throughout the circle, just as we would have seen for the unperturbed system (5). We refer to this lack of clear mutual attraction or mutual repulsion between the trajectories of different initial conditions as “neutral stability.” Now, in the setting of infinite-time dynamical systems, one can give a rigorous definition of “neutral stability” in terms of the traditional framework of long-time-asymptotic dynamics, such as we do in Sec. IV A. (The definition there implies, in particular, that all solutions are Lyapunov stable but not asymptotically stable, and have a Lyapunov exponent of zero.)

  • For A>A, we see that the trajectories are “stabilized,” i.e., they have become mutually synchronized with each other, such that they appear like a single point in Fig. 2(b).

Thus, a clear transition from “neutral stability” to “stability” takes place at A=A. Figure 2(c) shows the “numerical reverse-time bifurcation diagram,” where 50 evenly spaced initial conditions are run in backward time under (9) from time T to time 0; we see from this that when A rises above A, the forward-time solutions of (9) are being mutually repelled from a very small region [appearing like a single point in Fig. 2(c)]. So, overall, the picture indicated by plots (b) and (c) of Fig. 2 is rather resemblant of a saddle-node bifurcation in classical autonomous dynamical systems theory, where a lack of equilibria bifurcates into a repelling (unstable) and an attracting (stable) equilibrium. However, the system is a finite-time nonautonomous dynamical system: mathematically rigorous formulation of the “bifurcation” seen in plots (b) and (c) cannot be in terms of traditional “t” considerations by which concepts such as “stable” and “unstable” equilibria are defined.

3. Analysis via finite-time Lyapunov exponents

Local stability of solutions of a dynamical system can be quantified in terms of Lyapunov exponents. Traditional infinite-time dynamical systems theory tends to focus on long-time-asymptotic Lyapunov exponents, simply referred to as “Lyapunov exponents” without a qualifier. These Lyapunov exponents are defined as the long-time-asymptotic limit of finite-time Lyapunov exponents (FTLEs). In some contexts, one can also study other long-time-asymptotic features of the set of the FTLEs, such as given by large deviation theory29 or the dichotomy spectrum.40 Now, in our finite-time system, we cannot apply any long-time-asymptotic concepts; but we can still apply the concept of FTLEs within the bounded time-interval [0,T].

The finite-time Lyapunov exponent for a solution θ() of (9) over a time-window [0,T] is calculated by averaging over time t the derivative of the time-frozen vector field θasin(θ)+k+Ag(t) at θ(t), i.e.,

(11)

Working over the whole time-interval [0,T], Fig. 2(a) shows the values of λT for the trajectories of 50 initial conditions. For each A-value, we see that the 50 trajectories share indistinguishably the same FTLE value as each other (appearing like a single point), being indistinguishable from 0 for A<A and clearly negative for A>A. Again, this suggests a transition from neutral stability to stability as A crosses A.

4. Restricting to a time-subinterval

Now, for further illustration, let us consider the dynamics of (9) not over the whole time-interval, but rather over a subinterval [0,T] with T=π×104s. Figure 2(e) shows the numerical bifurcation diagram for simulation only up to time T, and likewise Fig. 2(f) shows the numerical reverse-time bifurcation when running initial conditions backward in time from time T to time 0. Here, we see the saddle-node-like “bifurcation” occurring at A:=akmint[0,T]g(t), again exactly in accordance with the reasoning of Sec. II A 3. In Fig. 2(d) are shown FTLEs of trajectories over the time-window [0,T], and again we see a transition from 0 to negative at A.

Note that A is larger than A—i.e., the critical parameter value for the saddle-node-like bifurcation changes when we consider the system over [0,T] vs when we consider the system over the whole time-interval [0,T]. This further shows how long-time-asymptotic approaches to considering stability (such as traditional Lyapunov exponents as opposed to finite-time Lyapunov exponents) are completely unsuitable for attempting to analyze the behavior of systems like the one we have just examined.

We seek to obtain a rigorous mathematical description, justification and generalization for the above-described stabilization phenomenon exhibited by Eqs. (7) and (9) and the quantification of stability given by Eq. (8).

In seeking to achieve this, the following three non-trivialities arise:

  • As we have discussed in Sec. II B, the traditional long-time-asymptotic framework of stability analysis and stability quantification via Lyapunov exponents—as used, for example, to describe the stabilization phenomena in Secs. II A 1 and II A 2—cannot generally be used to describe the stabilization phenomenon in Eq. (7) or (9). So, we need to formulate new mathematical definitions of stability within a suitable finite-time-dynamics framework.

  • The reasoning of Refs. 1 and 2 requires that while G(t)[a,a], there is no significant net stabilizing or destabilizing of trajectories. It is clear that this is true over timescales comparable with the “fast” timescale of the time-frozen vector field: all trajectories move approximately periodically round the circle with common approximate period 2πG(t)2a2. However, it is somewhat less obvious that no significant net stabilizing or destabilizing of trajectories can eventually build up over the “slow” timescale of the forcing.

  • If G(t) passes in and out of the interval (a,a) multiple times, then each new time that G(t) enters the interval (a,a), the previously synchronized solutions may experience some level of desynchronization due to a canard phenomenon17 in which the synchronized solutions spend a long time following the slow motion of moving sourcezG(t):=πyG(t). This phenomenon requires extremely fine tuning of parameters but we shall see that it still makes it impossible for us to take a “true” limit as ε0 in our slow–fast analysis. Rather, as we will prove, we will need that our “limit as ε0” is allowed to pass over an exponentially vanishing set of “bad” ε-intervals.

The way in which we address these issues is essentially by adapting the methods used in Ref. 17 to investigate periodic orbits of (9) when g(t) is sinusoidal. However, we will develop a much more elementary and more explicitly constructive approach to obtaining the necessary results about behavior while G(t)[a,a] than that used in Ref. 17.

In addition to the above issues, our general results will require us to take into account another dynamical phenomenon that cannot occur in the Adler equation, namely, bifurcation-induced tipping16,41,42 between stable equilibria of the time-frozen vector fields θF(θ,εt). This is not necessary when considering a forced Adler equation (7) because the time-frozen vector fields θasin(θ)+G(t) have at most one stable equilibrium. However, the presence of such tipping will not add much complication to the proofs.

Throughout this paper, we identify the circle S1 with R/(2πZ). (Hence, each tangent space TxS1 is naturally identified with R.)

An arc will always mean a connected proper subset of S1 with non-empty interior. For any two distinct points a,bS1, we define the “open arc extending anticlockwise from a to b”—written as (a,b)—to be the arc obtained by projecting an interval (a,b)R onto S1, where aR is any lift of a and b is the unique lift of b in the interval (a,a+2π). For such an open arc (a,b), we define length((a,b)):=ba; and for a non-open arc J, we set length(J):=length(J°). Just as (a,b)S1 denotes an open arc, so likewise we use the analogous notations [a,b), (a,b], and [a,b] for half-open arcs and closed arcs. For any aS1 and xR, we write a+xS1 for the projection of a+xR, where aR is any lift of a.

For any points a,bS1, we define the unsigned distance

For pS1 and δ>0, write Bδ(p)={θS1:d(θ,p)<δ}. So, if δ(0,π), then Bδ(p)=(pδ,p+δ).

Generally speaking, a variable θ will denote position on the circle S1, and a variable τ will denote “slow time,” which in this paper belongs to the unit interval [0,1]. Given a function F:S1×[0,1]R, we write m+nFθmτn (m,n0) to denote the nth partial derivative with respect to the second input of the mth partial derivative with respect to the first input of F.

Most results in this section will be proved in Sec. V.

We denote by F the set of functions F:S1×[0,1]R for which the partial derivatives Fθ, Fτ, 2Fθ2, and 2Fθτ all exist and are continuous on the whole of S1×[0,1]. We equip F with its natural norm,

where denotes the supremum norm for continuous functions on S1×[0,1].

Throughout this paper, we consider the differential equation

(12)

where FF.

Given an arc J0S1, for any ε>0 and t[0,1ε], we define the arc Jtε,F to be the image of J0 under the time-0-to-t mapping of (12), that is,

As in traditional dynamical systems theory, one can formulate local notions of stability around trajectories starting at individual initial conditions, as well as global-scale notions of stability that concern the whole ensemble of trajectories of all the initial conditions. Throughout this paper, we only consider such global-scale concepts of stability (and, therefore, we do not include the word “global” in our definitions).

Definition 1
We will say that F is neutrally stable if there exists c1 such that for every ε>0 and s,t[0,1ε], for every arc J0,
(13)
Definition 2
We will say that F is robustly neutrally stable if there exists c1 and a neighborhood U of F in F such that for every F~U, for every ε>0, s,t[0,1ε] and arc J0,
Definition 3
We will say that F is exponentially stable with rate Λ(,0) if given any η,Δ>0, for sufficiently small ε, there is an arc Pε with length(Pε)<Δ such that for every arc J0 not intersecting Pε,
(14)

Due to canard phenomena, the exponential stability formulated in Definition 3 will often be too strong, and so we will need to exclude a small set of ε-values. In the following, we write Leb() for the Lebesgue measure on (0,).

Definition 4
We will say that a set E(0,) is exponentially vanishing if there exists κ(0,1) such that for all sufficiently small h>0,

As an equivalent formulation, E is exponentially vanishing if and only if there exists κ(0,1) such that for all sufficiently large ξ>0,

Definition 5

We will say that F is almost exponentially stable with rateΛ(,0) if given any η,Δ>0, there is an exponentially vanishing set E(0,) such that for every ε(0,)E, there is an arc Pε with length(Pε)<Δ such that for every arc J0 not intersecting Pε, Eq. (14) holds.

All four stability definitions above can equivalently be formulated in terms of FTLEs. For an initial condition θ0S1, letting θ() be the solution of (12) starting at θ(0)=θ0, we define the corresponding FTLE over a time-window [s,t] by

(15)

Note that stFθ(θ(u),εu)du is precisely the logarithm of the derivative at θ(s) of the time-s-to-t mapping of (12). Hence, by the mean value theorem, we immediately have the following:

Proposition 6

For any r>0, ε>0, s,t[0,1ε] with s<t, and any connected subset S of S1 with non-empty interior, the following two statements are equivalent:

  • for every arc J0S,
  • for every θ0S,

Likewise, the corresponding two statements with in place of are equivalent to each other.

Thus, in Definition 1, the statement that Eq. (13) holds for every arc J0 can be reformulated as saying that for every θ0S1,

and similarly, in Definition 3 or Definition 5, the statement that Eq. (14) holds for every arc J0 not intersecting Pε can be reformulated as saying that for every θ0S1Pε,

Throughout the rest of Sec. III, we assume that FF.

Since we consider small ε, the dynamics of Eq. (12) will be essentially determined by the zeros of F.

Definition 7

A zero ofF is a point (θ,τ)S1×[0,1] such that F(θ,τ)=0. A zero (θ,τ) of F is called

  • hyperbolic stable if Fθ(θ,τ)<0;

  • hyperbolic unstable if Fθ(θ,τ)>0;and

  • non-hyperbolic if Fθ(θ,τ)=0.

A non-hyperbolic zero (θ,τ) is called non-degenerate if 2Fθ2(θ,τ) and Fτ(θ,τ) are both non-zero.

Definition 8

A curve of the stable (respectively, unstable) slow manifold is a pair (U,y) consisting of a set U[0,1] and a continuous function y:US1 such that for all τU, (y(τ),τ) is a hyperbolic stable (respectively, unstable) zero of F. We will also refer to the function y itself as being a curve of the stable (respectively, unstable) slow manifold over U.

We emphasize that in Definition 8, U is allowed to be the empty set (in which case y is the “empty function” S1), or a disconnected set, and y need not admit a continuous extension to U¯. In particular, if τ is a shared boundary point of two connected components of U and the limits limστy(σ) and limστy(σ) both exist, these limits do not have to be equal to each other.

Remark 9

Let us now comment on the behavior of F locally around its zeros.

  • Suppose (θ0,τ0) is a hyperbolic stable (respectively, unstable) zero of F. Then, by the implicit function theorem, there is a curve y of the stable (respectively, unstable) slow manifold over a neighborhood U of τ0 such that every zero of F close to (θ0,τ0) lies on graphy.

  • Suppose (θ0,τ0) is a non-degenerate non-hyperbolic zero of F with τ0(0,1). Then, the family of vector fields {F(,τ)}τ[0,1] undergoes a saddle-node bifurcation near θ0 as τ passes through τ0 (see also the description at the start of Sec. III G). In particular, over an interval U[0,1] with τ0U, there is both a curve y of the stable slow manifold and a curve z of the unstable slow manifold such that every zero of F close to (θ0,τ0) lies on the curve {(θ0,τ0)}graphygraphz.

Definition 10

Suppose we have τ[0,1] and θ1,θ2S1 such that (θ1,τ) and (θ2,τ) are zeros of F. We say that there is a fast connection from(θ1,τ) to (θ2,τ) if at least one of the following two statements holds:

  • F is strictly positive on (θ1,θ2)×{τ};

  • F is strictly negative on (θ2,θ1)×{τ}.

For a classical infinite-time autonomous differential equation θ˙=f(θ) with a stable equilibrium point y, one can consider the set of solutions that approach y arbitrarily closely as t, and moreover, one can quantify the long-time-asymptotic stability of such solutions using the Lyapunov exponent associated with y, defined as dfdθ(y). [Likewise, for an infinite-time stationary-noise-driven system such as our example (6) of noise-induced stabilization, one can define analogous concepts,43 with the analog of equilibrium points being invariant random Dirac measures.]

For our system (12), the analogous concept to a stable equilibrium point is what we will call a complete curve of the stable slow manifold.44 Analogous to the arbitrarily close approach of an equilibrium point as t is the arbitrarily close tracking of such a curve as ε0 (see Sec. III D); and in analogy to the Lyapunov exponent of an equilibrium, we will define the adiabatic Lyapunov exponent associated with a complete curve of the stable slow manifold, which (by Lemma 46) quantifies the stability of solutions that track the curve when ε is small.

First, we need to introduce a “genericity” assumption on F that will be required for us to be able to carry out the above-described analysis.

Definition 11

We will say that F is generic if

  • every non-hyperbolic zero of F is non-degenerate;

  • S1×{0,1} contains no non-hyperbolic zeros of F; and

  • for each τ(0,1), S1×{τ} contains at most one non-hyperbolic zero of F.

The first two points imply that every non-hyperbolic zero of F is as described in Remark 9(B).

If F is generic, then the number of zeros in S1×{0} is even, with half being hyperbolic stable and half being hyperbolic unstable.

Definition 12

We will say that F is initially multistable if S1×{0} contains at least two hyperbolic stable zeros.

Note, in particular, that if F has no zeros, then F is generic and not initially multistable.

The main results of this paper concern the stability properties of (12) when F is generic and not initially multistable. (If Fis initially multistable, then F will generally not exhibit the “global-scale” stability properties formulated in Sec. III A.)

Definition 13

Suppose F is generic. A curve (U,y) of the stable slow manifold will be called complete if the following statements hold:

  • U is open relative to [0,1] and has finitely many connected components;

  • there are no zeros of F in S1×([0,1]U¯); and

  • for any τ(0,1)U that is a shared boundary point of two connected components of U,

    • the limit y(τ):=limστy(σ) exists and (y(τ),τ) is a non-hyperbolic zero of F;

    • the limit y(τ+):=limστy(σ) exists and (y(τ+),τ) is a hyperbolic stable zero of F; and

    • there is a fast connection from (y(τ),τ) to (y(τ+),τ).

Theorem 14

Suppose F is generic.

  • If S1×{0} contains no zeros of F, then there exists a unique complete curve (U,y) of the stable slow manifold. Furthermore, U is empty if and only if F has no zeros.

  • For any θS1 with (θ,0) being a hyperbolic stable zero of F, there exists a unique complete curve (Uθ,yθ) of the stable slow manifold for which yθ(0)=θ.

We have the following immediate corollary:

Corollary 15

If F is generic and not initially multistable, then there exists a unique complete curve of the stable slow manifold.

The key method in the proof of Theorem 14 is continuation of a curve of zeros by virtue of Remark 9. The proof is given in Sec. V B.

Remark 16

As we will see from the proof of Theorem 14, if F is generic and initially multistable, then U¯θ is the same across all hyperbolic stable zeros (θ,0) in S1×{0}; and letting C1 be the connected component of U¯θ containing 0, we have that UθC1 and yθ|UθC1 are the same across all hyperbolic stable zeros (θ,0) in S1×{0}.

We now define adiabatic Lyapunov exponents, which generalize Eq. (8).

Definition 17
Suppose F is generic. Then, for each complete curve (U,y) of the stable slow manifold, we define the corresponding adiabatic Lyapunov exponent by
(16)
In the case that F is not initially multistable, we simply write Λad to refer to the adiabatic Lyapunov exponent associated with the unique complete curve of the stable slow manifold.

Note that Λad(U,y) must be finite and non-positive and that Λad(U,y)=0 if and only if U= (i.e., if and only if F has no zeros).

So far, we have formulated various definitions connected with the zeros of F, but not yet connected with how this relates to the dynamics of (12). We now introduce notions of tracking, similar to that of Ref. 41.

In the following definitions, we assume that we have a set V[0,1] and a function q:VS1. Where we will apply these definitions, (V,q) will always be either a curve of the stable slow manifold or a curve of the unstable slow manifold.

Definition 18

Fix ε>0. Given δ>0, we say that a function θ:[0,1ε]S1δ-tracks (V,q) if for every τV, d(θ(τε),q(τ))<δ.

Definition 19

Fix ε>0. Given Δ,δ>0, we say that (12)exhibits (Δ,δ)-tracking of (V,q) if there is an arc P with length(P)<Δ such that every solution θ() of (12) with θ(0)Pδ-tracks (V,q).

Definition 20

We say that

  • F exhibits strict tracking of (V,q) if given any Δ,δ>0, we have that for sufficiently small ε, (12) exhibits (Δ,δ)-tracking of (V,q);

  • F almost exhibits strict tracking of (V,q) if given any Δ,δ>0, there is an exponentially vanishing set E(0,) such that for every ε(0,)E, (12) exhibits (Δ,δ)-tracking of (V,q); and

  • F potentially exhibits strict tracking of (V,q) if given any Δ,δ>0, there is an open set A(0,) with infA=0 such that for every εA, (12) exhibits (Δ,δ)-tracking of (V,q).

Definition 21

We say that F admits a global canard phenomenon if there exists an open set U(0,1) and a curve z of the unstable slow manifold over U such that F potentially exhibits strict tracking of (U,z).

We now define what it means for F to exhibit “not necessarily strict” tracking of (V,q).

Definition 22

Assume that V is σ-compact. We say that Fexhibits (respectively, almost exhibits, potentially exhibits) tracking of (V,q) if for every compact subset S of V{0}, F exhibits (respectively, almost exhibits, potentially exhibits) strict tracking of (S,q|S).

Remark 23

Suppose V(0,1], and we have functions q1,q2:VS1 that disagree on at least one point in the interior of V; then, we cannot have both that F exhibits tracking of q1 and that F potentially exhibits tracking of q2. However (as in Sec. III E), we may have that F almost exhibits tracking of q1 and that F potentially exhibits tracking of q2.

Standing assumption. Throughout this section, if F is generic and not initially multistable, then we take (U,y) to be the unique complete curve of the stable slow manifold, and we take Λad to be the corresponding adiabatic Lyapunov exponent.

In our main result (Theorem 24), we consider functions F that are generic and not initially multistable and split into the following three cases: U¯ has zero, one, and more than one connected component.

Theorem 24

  • Suppose F has no zeros. Then, F is robustly neutrally stable.

  • Suppose F is generic and not initially multistable, and U¯ is non-empty and connected. Then, F exhibits tracking of (U,y) and is exponentially stable with rate Λad.

  • Suppose F is generic and not initially multistable, and U¯ is not connected. Then, F almost exhibits tracking of (U,y) and is almost exponentially stable with rate Λad.

Now, in case (C), a natural question is whether we have that F exhibits tracking of (U,y). In view of Remark 23, a negative answer is provided by the following result.

Proposition 25

Suppose F is generic and not initially multistable and U¯ is not connected. Then, F admits a global canard phenomenon.

Specifically, letting τ1 be the lower end point of the second connected component of U¯, we will show that there exists τ>τ1 and a curve z of the unstable slow manifold over (τ1,τ] such that F potentially exhibits tracking of ((τ1,τ],z).

Now, let us consider case (A) further. Define

(17)

and

The neutral stability in case (A) obviously implies that cF(ε), and hence rF(ε), remains bounded as ε0. This fact also follows immediately from Ref. 17, Theorem 2, but the proof of that result is highly involved and not explicitly constructive, meaning that one cannot easily write down a bound for these quantities as ε0. However, our analysis of the behavior of (12) when F has no zeros will be more elementary and explicitly constructive, allowing us to obtain the following explicit bound.

Proposition 26
Define the function e3:[0,)[16,) by
Suppose F has no zeros. For each τ[0,1], define
and let
Then,

Our results in Sec. III E are, in large part, obtained by adapting and generalizing methods used in Ref. 17 to study periodic orbits and their exhibition of the canard phenomenon in the periodic differential equation

(18)

In particular, Theorem 24(B) of this paper is crudely analogous to Proposition 4(1) of Ref. 17, Theorem 24(C) of this paper is crudely analogous to Theorem 1(3) of Ref. 17, and Proposition 25 of this paper is crudely analogous to Theorem 1(4) of Ref. 17.

Let us now outline the basic structure of the set of proofs of results in Sec. III E.

First, foundational to the proofs are the basic formulas [Eqs. (33) and (34)] for the differentiable dependence of solutions on their initial condition and on ε, which we establish [along with our general notations for the solution mapping of (12)] in Sec. V A.

1. When F has no zeros

When F has no zeros, a kind of neutral stability is obtained in Ref. 17: namely, as we have already said, Theorem 2 of Ref. 17 immediately implies that cF(ε) [as in Eq. (17)] remains bounded as ε0. However, it would be difficult or impossible to obtain from this proof an explicit bound on cF(ε), or to deduce from this proof that F is robustly neutrally stable in the sense of Definition 2.

So, in Sec. V C, we obtain Proposition 26 and Theorem 24(A) by a different approach. Specifically, our strategy is to approximate the nonautonomous differential equation (12) by a differential equation that is piecewise-autonomous on almost the whole time-interval, where on each “piece” all the solutions move strictly periodically round the circle with common period. The pieces are chosen to be of duration exactly equal to that common period, so that solutions do not mutually approach or move away from each other. Since F(,εt) has slow dependence on t, the piecewise-autonomous approximation is of high accuracy, and Grönwall’s Lemma is used to obtain that even over the whole time-interval [0,1ε], the inaccuracy of the piecewise-autonomous approximation does not allow mutual synchronization of trajectories to build up.

Now, in addition to the neutral stability, a further issue that needs addressing is the dependence of θ(1ε) on ε given the initial condition θ(0), since this will ultimately play a central role in the proof of Theorem 24(C). In the proof of Lemma 2 of Ref. 17, this is addressed by building further upon the abovementioned Theorem 2 of Ref. 17. By contrast, in Corollary 45, we obtain the necessary result [namely, the existence of a strictly positive lower bound on the magnitude of the derivative of θ(1ε) with respect to 1ε] as an immediate consequence of the fact that F is neutrally stable in the sense of Definition 1.

2. Tracking and stability

In cases (B) and (C) of Theorem 24, we have both a statement about tracking and a statement about stability. In fact, the statement about stability follows from the statement about tracking. This is fairly obvious when U¯ is the whole interval [0,1]; and when U¯ is a proper subset of [0,1], we apply Theorem 24(A) to the subintervals of [0,1] over which F has no zeros. Specifically, the statement linking tracking and stability is provided by Lemma 46 in Sec. V D.

3. Theorem 24(B)

In case (B) of Theorem 24, i.e., when U¯ is connected, if 0U, then it is not so obvious that F exhibits tracking of (U,y); specifically, the question arises of whether, as εt first enters into U, the trajectories of a significant proportion of initial conditions may exhibit the canard phenomenon of spending a long time tracking the curve of hyperbolic unstable zeros of F. So, we first study the case that 0U, in Sec. V E; and then [along somewhat similar lines to the proof of Proposition 4(1) of Ref. 17] for the case that 0U, we apply the results of this to the time-reversal of F starting at a time slightly after the first entry into U and going back to time 0.

It is not very hard to show that F exhibits tracking of (U,y) when U¯ is connected and 0U. In fact, letting (z0,0) be the hyperbolic unstable zero of F in S1×{0}, we have (Proposition 50) that F exhibits tracking of (U,y) “away from z0”; this is defined in Sec. V D and essentially means that the arc P in Definition 19 contains z0 regardless of the value of ε. [By contrast, in the general definition of F exhibiting tracking of (U,y), the arc P is free to move around as ε0.] As a result of this, by Lemma 46, we have (Proposition 51) that F is exponentially stable “away from z0” with rate Λad. More precisely, this notion is defined as follows:

Definition 27

Fix pS1. We say that F is exponentially stable away from p with rate Λ(,0) if given any η,Δ>0, for sufficiently small ε, for every arc J0 not intersecting BΔ2(p), Eq. (14) holds.

Remark 28

By Proposition 6, Definition 27 is equivalent to saying that λ0,1εε,F() converges to Λ uniformly on compact subsets of S1{p} as ε0.

Section V F then deals with case (B) in the scenario that 0U. To show that F exhibits tracking of (U,y), we apply Proposition 50 to deal with what happens after εt has entered U, and we apply the exponential stability result of Proposition 51 to the time-reversal of F as described above, in order to show that problems do not arise as εt enters into U. Again, Lemma 46 then gives that F is exponentially stable with rate Λad.

4. Proposition 25

Having dealt with case (B), we next prove Proposition 25 in Sec. V G. We do this before proving Theorem 24(C), in order to illustrate the canard phenomenon that will need to be avoided outside an exponentially vanishing set of ε-values in Theorem 24(C). Reference 17 specifically looks for the situation that (18) has an almost-everywhere-attracting stable periodic orbit exhibiting the canard phenomenon; the fact that this occurs for intervals of ε-values arbitrarily close to 0 is proved by analyzing the time-2πε mapping of (18). In a sense, our task of proving Proposition 25 is easier, in that we do not require any of the trajectories exhibiting the canard phenomenon to have any special additional property like “periodicity.” [For example, the system (18) exhibits the canard phenomenon also for ε-values where there are no periodic orbits.] But, on the other hand, we are considering a free form of temporal dependence rather than a periodic dynamical system, and so the approach of simply analyzing the shape of the graph of a time-t mapping of the system cannot be applied.

The idea of the proof of Proposition 25 is as follows. Take a value σ that lies strictly in between the first and the second connected component of U¯. The stability results that we have obtained so far for when U¯ is connected can be applied over [0,σ] (since U¯[0,σ] is connected) and can also be applied in reverse time over [σ,τ]. This gives us two high-density clusters of trajectories at time σε, one of which moves unboundedly anticlockwise round the circle as ε decreases, while the other moves unboundedly clockwise round the circle as ε decreases (Proposition 52), such that the clusters cross each other infinitely many times as ε0. [A similar idea is illustrated in Fig. 5(c) in Sec. IV E 2.] However, for the desired result, we need there to be arbitrarily small ε-values for which not only do these clusters meet each other, but actually the cluster obtained by evolving (12) forward in time over εt[0,σ] is entirely contained in the cluster obtained by evolving (12) backward in time over εt[σ,τ]. We achieve this by choosing τ sufficiently close to the lower end point of the second connected component of U¯ that the latter cluster shrinks at a slower rate than the former cluster as ε0.

FIG. 3.

Dynamics of (23), with varying A in (a)–(c) and varying k in (d)–(f). In all plots, ω=103rad/s, T=2π×105 s, and a=13rad/s. For (a)–(c), we have k=1rad/s, and the value ka=23rad/s is marked by the black dashed line. For (d)–(f), we have A=13rad/s, and the value a+A=23rad/s is marked by the black dashed line. In (a), (b), (d), and (e), results for the evolution θ(t) of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49, are shown: (a) and (d) show the finite-time Lyapunov exponents λT, as defined by (11), for these trajectories, and also show Λa,k,A [defined in Eq. (31)] in gray; (b) and (e) show the positions θ(T) of these trajectories at time T. In (c) and (f), the positions of θ(0) for the 50 trajectories of (23) with θ(T)=2πi50, i=0,,49, are shown. Plots (a)–(c) reproduced with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

FIG. 3.

Dynamics of (23), with varying A in (a)–(c) and varying k in (d)–(f). In all plots, ω=103rad/s, T=2π×105 s, and a=13rad/s. For (a)–(c), we have k=1rad/s, and the value ka=23rad/s is marked by the black dashed line. For (d)–(f), we have A=13rad/s, and the value a+A=23rad/s is marked by the black dashed line. In (a), (b), (d), and (e), results for the evolution θ(t) of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49, are shown: (a) and (d) show the finite-time Lyapunov exponents λT, as defined by (11), for these trajectories, and also show Λa,k,A [defined in Eq. (31)] in gray; (b) and (e) show the positions θ(T) of these trajectories at time T. In (c) and (f), the positions of θ(0) for the 50 trajectories of (23) with θ(T)=2πi50, i=0,,49, are shown. Plots (a)–(c) reproduced with permission from Newman et al., Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129. Copyright 2021 Springer Nature Switzerland AG.

Close modal

5. Theorem 24(C)

We prove Theorem 24(C) in Sec. V H by a similar approach to the proof in Ref. 17 that outside small intervals of ε-values system (18) has an almost-everywhere-attracting stable periodic orbit (Theorem 1, parts 1–3). In particular, Proposition 53 at the start of Sec. V H (which is an almost immediate consequence of Corollary 45 in Sec. V C) plays an analogous role to Lemma 2(1) of Ref. 17.

As before, the main task is to show that F almost exhibits tracking of (U,y), since Lemma 46 would then yield that F is almost exponentially stable with rate Λad. Essentially, the idea is to show that in constructions like the one used to prove Proposition 25, the crossings between the clusters of trajectories only take place during an exponentially vanishing set of ε-values. Since the clusters are themselves shrinking at an exponential rate as ε0, it is sufficient to find both a lower bound on the velocity (with respect to 1ε) at which the clusters move round the circle and an upper bound on the linear growth (again with respect to 1ε) of the cumulative distance of the circle swept out by the clusters (i.e., an upper bound on the “average velocity” of the clusters). The latter is fairly trivial [being essentially given by Eq. (35) in Sec. V A], while the former is essentially given by Proposition 53.

We now consider the typical behavior of a transition between case (A) and case (B) of Theorem 24 as a parameter of F is varied. First, let us recall the analogous scenario for autonomous dynamical systems.

Generically, for a (sufficiently smoothly) parameter-dependent autonomous dynamical system θ˙=fγ(θ), a transition between the absence and presence of equilibrium points will take place via a “non-degenerate saddle-node bifurcation,” where at the critical bifurcation parameter γ0, letting θ0 be the unique zero of fγ0, the following non-degeneracies hold:

  • d1:=fγ0(θ0)0 and

  • d2:=dfγ(θ0)dγ|γ=γ00.

The behavior of fγ(θ) locally around (γ=γ0,θ=θ0) is then similar to that of the normal form

around (γ=0,x=0). The zeros of this quadratic expression are precisely x=±2d2γd1 wherever this is well-defined, and the derivative of fγnormal at each such zero is given by ±sgn(d1)2d1d2γ. Accordingly, for the original parameter-dependent system θ˙=fγ(θ), sufficiently close to the critical parameter value γ0, we have the following:

  • For γ on one side of γ0, there are no equilibria and so all solutions move periodically round the circle with the same period.

  • For γ on the other side of γ0, there is one stable equilibrium yγ and one unstable equilibrium zγ; thus, almost all solutions converge to yγ at an exponential rate with exponent λ(γ):=fγ(yγ). Furthermore,
    (19)
    as γγ0.

Now, for the slow–fast framework (12) with a parameter-dependent F=Fγ, the analogous scenario is a transition between the absence and presence of zeros of Fγ. [This is precisely what we see illustrated numerically in Figs. 2(a)2(c) for the example

and in Figs. 2(d)2(f) for the example

with g() as in Sec. II B.] Assuming sufficient smoothness, generically a transition between the absence and presence of zeros of Fγ will be as follows:

At the critical parameter γ0, Fγ0 has a unique zero (θ0,τ0), and τ0(0,1) and the following non-degeneracies hold:

  • the Hessian matrix
    is invertible;
  • d2:=dFγ(θ0,τ0)dγ|γ=γ00.

As we will prove, the behavior locally around (γ=γ0,θ=θ0,τ=τ0) is then similar to that of the “normal form”

(20)

around (γ=0,x=(0,0)). We will shortly consider the zeros of this quadratic expression and then state a corresponding result for the original system (12), but first we can automatically say more about the matrix D1.

Lemma 29

Suppose we have FC2(S1×[0,1],R) with a zero (θ0,τ0)S1×(0,1) such that F(θ0,τ0)=(0,0) and det(HessF(θ0,τ0))<0. Then, F takes positive, negative, and zero values in an arbitrarily small punctured neighborhood of (θ0,τ0).

Hence, in the above setup, we must have that det(D1)>0. Note that for a 2×2 real symmetric matrix A with positive determinant, the diagonal entries are non-zero and have the same sign; so, we will write diag-sgn(A){±1} for the sign of the diagonal entries of A.

Now, returning to consideration of the normal form (20), let us write

The discriminant of the quadratic function fγ,x2() is given by

We have that

  • if D(fγ,x2)>0, then fγ,x2 has two zeros, and the derivative of fγ,x2 at these zeros is ±D(fγ,x2);

  • if D(fγ,x2)=0, then fγ,x2 has one zero, and the derivative of fγ,x2 at this zero is 0; and

  • if D(fγ,x2)<0, then fγ,x2 has no zeros.

Suppose without loss of generality that diag-sgn(D1)=1 and d2>0. For γ>0, Fγnormal has no zeros. For γ<0, we have

and so Fγnormal has zeros as described above. The value of the integral U~γD(fγ,x2)dx2 is the negative of the area of a semi-ellipse with principal radii 2dθθd2γdet(D1) and 2dθθd2γ, namely,

Now, let us make a rigorous statement about the original system (12) with dependence on a parameter γ.

We equip C2(S1×[0,1],R) with the standard C2-topology. Fix an open interval ΓR, and let F2 be the set of all continuous paths (Fγ)γΓ in the space C2(S1×[0,1],R) for which the functions

are well-defined and continuous on Γ×S1×[0,1]. (For each individual γ, we apply the same partial derivative notations to the function Fγ as were introduced at the start of Sec. III.)

Given (Fγ)γΓF2, a point (γ0,θ0,τ0)Γ×S1×[0,1] is called a critical zero if Fγ0(θ0,τ0)=0 and Fγ0(θ0,τ0)=(0,0). [This implies, in particular, that (θ0,τ0) is a degenerate non-hyperbolic zero of Fγ0.] We say that a critical zero (γ0,θ0,τ0) is non-degenerate if HessFγ0(θ0,τ0) is invertible and dFγdγ(γ0,θ0,τ0) is non-zero.

Theorem 30
Suppose we have (Fγ)γΓF2 with a non-degenerate critical zero (γ0,θ0,τ0) with τ0(0,1), such that (θ0,τ0) is the only zero of Fγ0. Let D1 and d2 be as above. Then, there exists δ>0 such that, defining the intervals Γ and Γ+ by
we have the following:
  • For all γΓ, Fγ has no zeros and thus is neutrally stable.

  • For all γΓ+, Fγ is generic with a unique complete curve (Uγ,yγ) of the stable slow manifold, and Uγ is non-empty and connected. Thus, Fγ is exponentially stable with rate equal to the corresponding adiabatic Lyapunov exponent Λ(γ). Furthermore,
    (21)
    as γγ0.
Remark 31
Note that if dθτ=0, then Eq. (21) simplifies to
(22)

Let us emphasize a key point of contrast between Eq. (21) and Eq. (19): namely, the latter has a square-root dependence on γγ0 while the former has a linear dependence on γγ0. So, although we said in Sec. II B that the bifurcation portraits in Fig. 2 bear resemblance of a classical saddle-node bifurcation, one significant way in which plots (a) and (d) differ from a generic saddle-node bifurcation is the following: as A rises above A and A, respectively, the numerical FTLEs appear to grow in magnitude linearly (as opposed to the faster-than-linear growth that one would see in a generic saddle-node bifurcation).

Lemma 29 and Theorem 30 will be proved in Sec. V I.

In this section, we compare the stability formalisms and results in Sec. III to the traditional long-time-asymptotic framework:

  • We will present definitions of neutral stability and exponential stability for systems on S1 that are closely analogous to Definitions 1 and 27, except now within the traditional t framework rather than our ε0 framework.

  • Whereas Secs. II A 1 and II A 2 considered stabilization of the Adler equation by constant forcing and by noise, and described the stabilization in terms of traditional t concepts, now we will consider stabilization of the Adler equation by low-frequency periodic forcing (especially sinusoidal17,45–47 forcing) so that both the traditional t framework and (by restricting to finite time) the framework of Sec. III can be applied.

Remark 32

In any real-world situation, one will be working with a finite-duration process (as opposed to taking t) subject to “finite-timescale” forcing (as opposed to taking ε0). So, for a model in which both frameworks make sense, such as a low-frequency-periodically forced system, the question of which of the two frameworks (if either) provides a suitable approximation would need to be determined by context.

First, let us note that the discussion of Eq. (9) in Sec. II A 3 can be applied to the case that g(t)=cos(ωt) for small ω: If k>a, and if we consider the system

(23)

over a time-interval of duration greater than πω, then the reasoning of Refs. 1 and 2 (as formalized by Theorem 24) gives that

  • if 0A<ka, then system (23) exhibits neutrally stable dynamics (i.e., no significant mutual synchronization or repulsion of trajectories), and

  • if A>ka, then the Adler equation is stabilized: the trajectories of different initial conditions mutually synchronize into a tight cluster.

In Sec. IV A, we will define some general dynamical properties, especially stability properties, for infinite-time dynamical systems on the circle, within the classical long-time-asymptotic framework.

In Sec. IV B, we will apply these long-time-asymptotic concepts given in Sec. IV A to provide a basic stability analysis of periodically forced Adler equations. In particular, we will consider saddle-node bifurcations of periodic orbits and stabilization of the Adler equation with k>a.

In Sec. IV C, we will use numerical bifurcation diagrams of the same type as in Sec. II B to look at the parameter-dependence of the stability properties of low-frequency-sinusoidally forced Adler equations.

In Sec. IV D, we analyze Eq. (23) over finitely many periods of the forcing in terms of our new framework of Sec. III (in particular, applying Theorem 24). We then consider the numerics of Sec. IV C in the light of this theoretical analysis.

In Sec. IV E, we tie together all the above theoretical and numerical results to discuss how analysis of dynamics under the traditional framework and analysis of dynamics under our new framework relate and compare to each other for periodically forced Adler equations.

We consider systems of the form

(24)

for some continuous function F:S1×[0,)R for which the partial derivatives Fθ and 2Fθ2 exist and are continuous on the whole of S1×[0,). The most fundamental examples are the cases that

  • (24) is autonomous, meaning that F(θ,t) is independent of t, i.e., F(θ,t)=f(θ) for some f:S1R, and

  • (24) is 2πω-periodic for some ω>0, meaning that F(θ,) is 2πω-periodic for all θS1.

More advanced examples include quasiperiodic systems, almost-periodic48 systems, systems subject to forcing from a stationary stochastic process or measure-preserving dynamical system,43 and asymptotically autonomous41,49 systems. In the periodic case, we will write Φ:S1S1 for the time-0-to-2πω mapping of (24), i.e., Φ(θ(2πnω))=θ(2π(n+1)ω) for any solution θ() of (24) and any nN.

Now, given an arc J0S1, for any t[0,), we define the arc JtF to be the image of J0 under the time-0-to-t mapping of (24), that is,

For any θ0S1 and t>s0, we define the finite-time Lyapunov exponent

(25)

where θ() is the solution of (24) with θ(0)=θ0. We define the (asymptotic) Lyapunov exponent by

(26)

if this limit exists in R¯; the existence and the value of the limit are clearly not dependent on s. We also define the (asymptotic) average angular velocity by

where θ^:[0,)R may be any lift of any solution θ:[0,)S1 of (24), if this limit exists. Note that the existence and the value of this limit do not depend on the initial condition θ^(0): if θ^1 and θ^2 are lifts of solutions of (24), then supt0|θ^2(t)θ^1(t)| cannot be larger than the smallest multiple of 2π above |θ^2(0)θ^1(0)|. In the autonomous case F(θ,t)=f(θ), we have

It is also well-known that in the periodic case, ΩF exists and has continuous dependence (in the C0-topology) on Φ.

Now, as in Sec. III, although one can formulate notions of stability and neutral stability locally around individual trajectories, here, we will just present global-scale notions of stability (and, therefore, we do not include the word “global” in our definitions).

Definition 33
We say that system (24) is neutrally stable if there exists c1 such that for every s,t[0,), for every arc J0,
(27)

Note that if system (24) is neutrally stable, then the Lyapunov exponent λF(θ0) associated with every initial condition θ0S1 is 0.

Remark 34

Suppose (24) is 2πω-periodic for some ω>0, and ΩF is an irrational multiple of ω. Then, by Denjoy’s theorem applied to Φ, (24) is neutrally stable.

Definition 35
Fix pS1. We say that the system (24) is exponentially stable away frompwith rateΛ[,0) if given any η,Δ>0, for sufficiently large t>0, for every arc J0 not intersecting BΔ2(p), we have

This is equivalent to saying that λ0,tF()Λ uniformly on compact subsets of S1{p} as t. Hence, in particular, the Lyapunov exponent λF(θ0) associated with each θ0S1{p} is Λ.

Remark 36

Suppose (24) is 2πω-periodic for some ω>0 and that (24) is exponentially stable away from a point p with rate Λ. Then, considering the graph of Φ yields the following: the solution θ() starting at θ(0)=p is 2πω-periodic; there is exactly one other 2πω-periodic solution; letting q be the initial condition of this latter periodic solution, we have Λ=λ0,2πωF(q); and ΩF is an integer multiple of ω.

Typically, if (24) is 2πω-periodic and ΩF is a non-integer rational multiple of ω, then (24) will admit multiple locally exponentially stable periodic solutions. (However, as we will soon see, this does not have to be the case.)

We have formulated the above definitions for differential equations but the same definitions also apply to infinite-time dynamical systems with non-differentiable solutions. For example, for system (6) presented in Sec. II A, there exists pS1 (dependent on W) such that (6) is exponentially stable away from p with some W-independent rate λa,k,A<0.

We now apply the concepts of Sec. IV A to Adler equations with periodic additive forcing.

1. The possible behaviors

The following known result is essentially the main result of Ref. 45 (Theorems 1 and 4).

Proposition 37
Take F in (24) of the form
where aR is constant and G is 2πω-periodic for some ω>0. Exactly one of the following three statements holds:
  • System (24) is neutrally stable.

  • For some pS1 and Λ(,0), system (24) is exponentially stable away from p with rate Λ.

  • System (24) has exactly one 2πω-periodic solution p(t); every solution θ() of (24) has d(θ(t),p(t))0 as t; and λF(θ0)=0 for all θ0S1.

Cases (ii) and (iii) can only occur if ΩF is an integer multiple of ω; i.e., if ΩF is not an integer multiple of ω, then system (24) is neutrally stable.

Case (iii) is a kind of “boundary case” between neutral stability and exponential stability; we expect that for “typical” a and G(), if ΩF is an integer multiple of ω then the system will be in case (ii).

2. Periodic saddle-node bifurcations

A transition between case (i) and case (ii) via case (iii) as some parameter is varied corresponds precisely to a saddle-node bifurcation of the time-0-to-2πω mapping Φ. Let us now describe one such scenario.

Proposition 38

In Proposition 37, consider G(t) of the form G(t)=k+H(t), where we fix the function H() while considering varying values of kR. Suppose we have k0R such that for k=k0, system (24) is in case (iii) of Proposition 37. Then, sufficiently close to k0, we have the following:

  • for k on one side of k0, system (24) is in case (i) of Proposition 37, i.e., is neutrally stable; and

  • for k on the other side of k0, system (24) is in case (ii) of Proposition 37, with
    (28)
    as kk0 for some C>0.

Note that the autonomous system (5) is precisely the case where H0, and that the system (23) (for fixed A and ω) is precisely the case where H(t)=Acos(ωt).

Recall that the autonomous case with a>0 is as in Sec. II A 1: if |k|>a, then the system is neutrally stable, but if |k|<a, then the system is exponentially stable with exponent

which approximates to 2a|ka| for k near ±a. This is precisely in accordance with the general description of saddle-node bifurcations of autonomous systems given at the start of Sec. III G. So, what Proposition 38 essentially says is that as we now go from the autonomous case to the periodic case, all k-parameterized transitions between neutral stability and exponential stability must still likewise take place via a “saddle-node bifurcation of 2πω-periodic orbits” exhibiting the same square-root dependence (28).

Proof of Proposition 38
Proof
Proof of Proposition 38
Let p(t) be as in Proposition 37 for k=k0. By Ref. 45, Theorem 4, stereographic projection conjugates Φ(k=k0) to an orientation-preserving linear fractional transformation Φ~ of R^ with a unique fixed point p~; without loss of generality, we can take p~R, since otherwise we can first rotate the circle by angle π to get p~=0. The second derivative of any linear fractional transformation other than the identity function is non-zero throughout R; and since smooth conjugation simply divides the second derivative at a fixed point by the derivative of the conjugacy at that point, it follows that d1:=Φ(k=k0)(p(0)) is non-zero. Also, the derivative d2 of the map kΦ(p(0)) at k=k0 is given by d2=02πωet2πωFθ(p(u),u)dudt, which is clearly strictly positive. Hence (similarly to the description at the start of Sec. III G), the map Φ undergoes a non-degenerate saddle-node bifurcation as k crosses k0; and by Proposition 37, the two sides of the bifurcation correspond to cases (i) and (ii) of Proposition 37. Furthermore, in case (ii), we have that Λ=logD, where D is the derivative of Φ at its unique stable fixed point, which [similarly to Eq. (19)] takes the form
So,

3. Stabilization of the neutrally stable Adler equation

Let us now consider the question of stabilization of the Adler equation. Specifically, we consider the addition of zero-mean periodic forcing to the autonomous system θ˙=asin(θ)+k, where k>a>0. Obviously, without the addition of any forcing [i.e., G(t)=k in Proposition 37], the system belongs to case (i) of Proposition 37. The question is then whether, with the added zero-mean periodic forcing, the forced system now belongs to case (ii) of Proposition 37.

Let us first note that for the forced system, the average angular velocity must lie within the compact subinterval [ka,k+a] of (0,). This follows immediately from the fact that the term asin(θ(t)) always lies within the range [a,a]. Since the average angular velocity has continuous dependence on the time-0-to-2πω mapping of the system, we immediately obtain from Proposition 37 the following corollary about low-frequency forcing: there is no way to send ω to 0 without passing through the scenario that the system is neutrally stable. Let us phrase this more precisely.

Corollary 39
Fix k>a>0. Consider the ε-parameterized periodic differential equation
(29)
where g~:(0,)×RR is any continuous function with the following properties:
  • g~(ε,) is 2πω(ε)-periodic for some ω(ε)>0 with ω(ε)0 as ε0, and

  • 02πω(ε)g~(ε,t)dt=0.

System (29) is neutrally stable for an open set of ε-values arbitrarily close to 0.

Let us exemplify this with the prototypical case of sinusoidal forcing. Taking

gives the following.

Corollary 40

Fix k>a>0. In the (A,ω)-parameter space of Eq. (23), it is not possible to plot a continuous path that tends toward the ω=0 axis without traveling through a region of the (A,ω)-parameter space where the system (23) is neutrally stable.

In the most basic case, let us simply fix any value AR: then, there is an open set of ω-values arbitrarily close to 0 for which system (23) is neutrally stable, i.e., for which the forcing +Acos(ωt) “fails to stabilize the Adler equation.”

Having investigated periodically forced Adler equations such as (23) theoretically, we now plot numerical bifurcation diagrams for (23). We cannot “plot the long-time-asymptotic dynamics” simply through numerical simulation of trajectories since the simulations are themselves only of finite duration. However, as is typically done in numerical bifurcation diagrams, we can still plot “long-but-finite-time” results obtained from the finite-time simulations. (So, for example, we cannot calculate asymptotic Lyapunov exponents simply from numerically simulated trajectories, but we can calculate finite-time Lyapunov exponents taken over the entire duration of the simulation.)

Specifically, we will construct the same kind of plots as in Sec. II B. Fix a=13rad/s (as in Sec. II B) and ω=103rad/s (which is the same as the cut-off angular frequency of the low-pass filter in Sec. II B). We will consider both

  • varying A with fixed k=1rad/s>a (as in Sec. II B) and

  • varying k with fixed A=13rad/s.

Note that ω is very small, i.e., we are considering low-frequency forcing; however, we still make sure that the duration of the numerical simulations is long enough to include many complete cycles of the low-frequency forcing. Specifically, we will work with 100 periods of the forcing [which is the same duration as the duration T of the simulations represented in Figs. 2(a)2(c) in Sec. II B]. In other words, we can think of the results as being for the 100th iterate of the circle map Φ. Results are shown in Fig. 3.

FIG. 4.

Graph of Λa,k,A [as defined in Eq. (31)] against k for different values of A, with a=13rad/s. The blue curve Λa,k,0=max(0,a2k2) for A=0 has infinite gradient on the left at k=a where the curve transitions from negative to zero; the orange, gray, and green curves, respectively, have finite gradient 12, 12, and 122 on the left, where they transition from negative to zero (see Proposition 43).

FIG. 4.

Graph of Λa,k,A [as defined in Eq. (31)] against k for different values of A, with a=13rad/s. The blue curve Λa,k,0=max(0,a2k2) for A=0 has infinite gradient on the left at k=a where the curve transitions from negative to zero; the orange, gray, and green curves, respectively, have finite gradient 12, 12, and 122 on the left, where they transition from negative to zero (see Proposition 43).

Close modal

In Figs. 3(b) and 3(e), for the various combinations of A and k, the trajectories at time T=200πω of 50 evenly spaced initial conditions are shown. Figures 3(c) and 3(f) show the analogous results for the reverse-time system; due to the symmetries of (23), plots (c) and (f) turn out to be the reflection of plots (b) and (e) about π2. Figures 3(a) and 3(d) show FTLEs for the trajectories shown in Figs. 3(b) and 3(e), respectively. These FTLEs are computed according to Eq. (11), which is precisely the same as Eq. (25) with F(θ,t)=asin(θ)+k+Acos(ωt).

Plots (a)–(c) show a picture somewhat resemblant of a saddle-node bifurcation: recalling that the system is neutrally stable for A=0, we see the persistence of this exhibition of neutrally stable behavior for A going up to about A=23rad/s, after which the system appears to be clearly stabilized.

Plots (d)–(f) likewise show a picture somewhat resemblant of a saddle-node bifurcation: we see stability for k going up to about k=23rad/s, after which we see neutrally stable behavior. This is similar to the autonomous case A=0, except with a higher critical k-threshold for the transition from stability to neutral stability.

However, plot (d) also reveals a feature distinctly different from a saddle-node bifurcation: as k approaches the critical threshold from below, the dependence of the finite-time Lyapunov exponents on k appears to be linear (with a gradient of about 12) rather than of square-root form. [Plot (a) likewise appears potentially to indicate a linear dependence on A as the critical A-threshold is approached from above, but this is less clear.]

Standing assumption. In all subsequent discussion of Eq. (23), we take a>0 and k,A0.

We now consider Eq. (23) on finite time-intervals, within the slow–fast framework (12). Specifically, we will consider (23) over time-intervals [0,nπω], i.e., over n2 cycles of the periodic forcing, for positive integers n. This can be expressed in the form of (12) by taking

(30)

with ε=ωnπ. Obviously, F is not initially multistable. We have that F is generic if and only if A{k+a,ka,ak}; in this case, we will denote the adiabatic Lyapunov exponent by Λa,k,A0, which clearly does not depend on n. We extend Λa,k,A continuously to cover those (a,k,A)-values for which F is not generic. Explicitly,

(31)

Applying Theorem 24 gives the following.

Proposition 41

For F as in Eq. (30):

  • If k>a and A<ka, then F is neutrally stable.

  • If

    • k<a and A<ak, or

    • |ak|<A<a+k and n{1,2}, or

    • A>a+k and n=1,

    then F is exponentially stable with rate Λa,k,A.

  • If

    • |ak|<A<a+k and n3, or

    • A>a+k and n2,

    then F is almost exponentially stable with rate Λa,k,A.

Let us mention, in particular, the following cases:

Corollary 42

Suppose n3.

  • Fixing k>a and letting A:=ka, we have

    • if A<A then F is neutrally stable, but

    • if A(A,){k+a} then F is almost exponentially stable with rate Λa,k,A.

  • Fixing a>0 and A0 and letting k:=a+A, we have

    • if k>k, then F is neutrally stable, but

    • if k[0,k){|aA|}, then F is almost exponentially stable with rate Λa,k,A (with the “almost” not being necessary in the case that A<a and k<aA).

We illustrate case (B) of Corollary 42 with a graph of Λa,k,A against k for different values of A, in Fig. 4. One immediate observation from this figure is the following:

  • for A=0, i.e., the autonomous case (where a saddle-node bifurcation takes place as k crosses a), we see an infinite gradient in the graph as ka, corresponding to the square-root dependence Λa,k,02a|ka| exactly as described just after the statement of Proposition 38; but

  • when we now go to the nonautonomous periodic case A>0, the graph has a finite gradient as ka+A.

FIG. 5.

Dynamics of (23) with varying ω. Other parameters are a=13 rad/s and A=k=1rad/s. (a) For each ω-value, taking T=200πω (i.e., 100 periods), the finite-time Lyapunov exponents λT as defined by (11) are shown for the trajectories of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49. The horizontal gray line marks the value of Λa,k,A defined in (31). (b) Zoomed-in version of (a); the red points indicate the location of the small intervals of ω-values for which all initial conditions have zero asymptotic Lyapunov exponent. (c) Forward- and backward-time dynamics over the time-interval [0,2πω]; for each ω-value, hollow circles show the positions of θ(2πω) for the 50 trajectories of (23) with θ(0)=2πi50, i=0,,49, while solid circles show the positions of θ(0) for the 50 trajectories of (23) with θ(2πω)=2πi50, i=0,,49. In both cases, the 50 points are clustered together; the values of ω where the two curves of clustered points cross correspond to where the red points are marked in (b); see  Appendix B for further explanation.

FIG. 5.

Dynamics of (23) with varying ω. Other parameters are a=13 rad/s and A=k=1rad/s. (a) For each ω-value, taking T=200πω (i.e., 100 periods), the finite-time Lyapunov exponents λT as defined by (11) are shown for the trajectories of 50 equally spaced initial conditions θ(0)=2πi50, i=0,,49. The horizontal gray line marks the value of Λa,k,A defined in (31). (b) Zoomed-in version of (a); the red points indicate the location of the small intervals of ω-values for which all initial conditions have zero asymptotic Lyapunov exponent. (c) Forward- and backward-time dynamics over the time-interval [0,2πω]; for each ω-value, hollow circles show the positions of θ(2πω) for the 50 trajectories of (23) with θ(0)=2πi50, i=0,,49, while solid circles show the positions of θ(0) for the 50 trajectories of (23) with θ(2πω)=2πi50, i=0,,49. In both cases, the 50 points are clustered together; the values of ω where the two curves of clustered points cross correspond to where the red points are marked in (b); see  Appendix B for further explanation.

Close modal

These finite gradients can be calculated using Theorem 30, as follows.

Proposition 43

  • Fixing k>a and letting A:=ka, the map AΛa,k,A is right-sided-differentiable at A with derivative 12aA.

  • Fixing a,A>0 and letting k:=a+A, the map kΛa,k,A is left-sided-differentiable at k with derivative 12aA.

Proof.
In case (A), take Fγ to be the function F in Eq. (30) with n=2 and A=γ. Then, taking (γ0,θ0,τ0)=(A,π4,12), we have that (θ0,τ0) is the only zero of Fγ0, and
Thus, the conditions of Theorem 30 are satisfied, and applying Eq. (22) gives the result. Case (B) is proved the same way, with k=γ and γ0=k.

Now, just as one can seek numerical help in investigating the long-time-asymptotic behavior of an infinite-time dynamical system through long-but-finite-time simulations, so likewise one can seek numerical help in investigating the ε0 behavior of a dynamical system of form (2) through small-but-positive-ε simulations. Now, combining the two: suppose we have a finite-time numerical simulation of an infinite-time system subject to external forcing, where the timescale of the forcing in the simulation is much slower than the timescale of the system’s internal dynamics, and the duration of the simulation is very large compared to the timescale of the forcing. Then, exactly the same comment that we made in Remark 32 regarding real-world processes also applies to the simulation: one can interpret it as either an approximation of the t behavior or an approximation of the ε0 behavior. Note that Fig. 3 comes under exactly this category: ω is very small, but at the same time, n is very large.

Accordingly, let us now reconsider Fig. 3 in the light of Corollary 42. In all six plots, we have indicated the value of A=ka or k=a+A as appropriate by a dashed black line. Additionally, in plots (a) and (d), we have shown in gray the value of Λa,k,A as given in Eq. (31); so, the gray curve in plot (d) is the same as the gray curve in Fig. 4. The numerical results in Fig. 3 show perfect agreement with the theoretical predictions of Corollary 42 regarding stability, neutral stability, and quantification of stability, just as the numerics in Sec. II B for “unextendible” finite-time forcing likewise showed perfect agreement with the theoretical predictions of Theorem 24 regarding stability and neutral stability. (Recall that all these theoretical predictions are simply the mathematically precise description of what one would heuristically expect to see by applying the reasoning of Refs. 1 and 2.)

So far, we have presented theoretical results within the long-time-asymptotic framework, numerical results, and theoretical results within the finite-time slow–fast framework. Let us now tie these together and draw conclusions.

1. Stable vs neutrally stable dynamics for A > k − a

At the end of Sec. IV B 3, we saw that when k>a, for any fixed value of A there are robust ω-values arbitrarily close to 0 for which the system (23) is neutrally stable in the sense of Sec. IV A. (Recall that this implies, in particular, that all trajectories have an asymptotic Lyapunov exponent of zero.) The proof of this result made no recognition whatsoever of a distinction between A<ka and A>ka. Yet, Figs. 3(a)3(c) show a very clear distinction between neutrally stable behavior for A<ka and stable behavior A>ka exactly as in Corollary 42. The fact that these ω-values of neutral stability exist arbitrarily close to zero even when A>ka indicates that, at such ω-values, the heuristic reasoning described in Sec. II A 3 must break down.

The cause of this breakdown is precisely the canard phenomenon where solutions spend a long time tracking the motion of the slowly moving source and thus experience mutual desynchronization. It is well-known that the canard phenomenon is an extremely fine-tuned phenomenon; in other words, the open set of ω-values described at the end of Sec. IV B 3 must consist of extremely narrow ω-intervals if A>ka. Indeed, for the range of parameter-values considered in Ref. 17, it was proved that this set of ω-values is exponentially vanishing; and similarly in this present paper, we have shown [Theorem 24(B)–(C)] that for any fixed finite-length shape of slow-timescale variation, the set of timescale-separation values for which the canard phenomenon causes the reasoning in Sec. II A 3 to break down is likewise exponentially vanishing. However, the fact that the A>ka case requires this fine-tuned canard mechanism of desynchronization in order for the neutral stability (in the sense of Sec. IV A) to be achieved is altogether missed by the bifurcation and stability analysis given in Sec. IV B.

2. Illustration through further numerics

Let us now consider through further numerics how the dynamics of Eq. (23) depends on ω when A>ka. Figures 5(a)5(b) are FTLE plots similar to those in plots (a) and (d) of Figs. 2 and 3, except that the varied parameter is now ω rather than A or k. We take a=13rad/s and A=k=1rad/s, so ka<A<k+a, implying that F is exponentially stable for n{1,2} and almost exponentially stable for n3; here, as in Fig. 3, we take 100 cycles of the forcing, i.e., the results are shown over a duration of 200πω. In gray is shown the value of Λa,k,A. For ω sufficiently small, we see a close match of the FTLEs to Λa,k,A. In plot (b) are also indicated the approximate locations of the narrow ω-intervals for which system (23) is neutrally stable in the sense of Sec. IV A. (The procedure by which this is achieved is described in  Appendix B.) Despite the high density of ω-values for which FTLEs are plotted in plot (b), all the results are close to Λa,k,A and none are close to 0.

For further illustration, we now zoom in around one of the points marked in Fig. 5(b) as being the approximate location of an ω-interval of neutral stability. In Fig. 6, FTLE results are shown for both 100 cycles and 1000 cycles of the periodic forcing. Despite the extremely high density of ω-values for which results are shown, all the shown FTLEs approximate the value Λa,k,A<0 to an accuracy of at least about 97%.

3. Dependence of Lyapunov exponents on k

The apparently linear dependence of the FTLEs in Fig. 3(d) to the left of k=23rad/s is perfectly explained by Corollary 42 and Proposition 43: the left-sided derivative of kΛa,k,A at k=a+A is equal to 12. This is in contrast to Proposition 38, which gives us the following:

  • for any transition between stability and neutral stability (understood within the framework of Sec. IV A) that arises from keeping the parameters a, A, and ω fixed while varying k past a critical value k0, the Lyapunov exponent on the stable side of the transition must have square-root dependence on |kk0| as k approaches k0.

One question that then naturally arises is the following: if we fix ω, A, and a as in Figs. 3(d)3(f), then over an interval of k-values such as, say, [0.65,1]rad/s, which k-values correspond to neutral stability and which k-values correspond to exponential stability with small Lyapunov exponent? Nonetheless, the fundamental mechanism of stabilization (as described in Sec. II A 3 and formalized by Theorem 24) that gives rise to what we see in Fig. 3 is essentially blind to the distinction between these two scenarios; indeed, it is exactly this same mechanism of stabilization that also gave rise to the equally clear pictures seen in Fig. 2, where questions of long-time-asymptotic dynamics do not make sense.

We begin with some further notations.

We will always assume that FF. Define

Given σ1,σ2[0,1], define Fσ1,σ2F by

So, for fixed ε, if θ() is a solution of Eq. (12), then

is a solution of Eq. (12) with Fσ1,σ2 in place of F: denoting this function by θσ1,σ2, we have

Note that if F is generic and σ1σ2, then Fσ1,σ2 is also generic.

Recall the finite-time Lyapunov exponents λs,tε,F(θ0) defined in Eq. (15); for convenience, also define the notation

where the first equality is used to define λ¯s,tε,F(θ0) even for st. Note that

Define the function

such that, given any ε>0, for any solution θ() of (12) and any s,t[0,1ε], we have

Equivalently, given any ε>0, for any solution ϕ:[0,1]S1 of the differential equation

(32)

and any s,t[0,1],

We denote the partial derivative of Φ with respect to its first input and its second input, respectively, by

We now give formulas for these partial derivatives.

Proposition 44
Take any ε>0 and let θ() be a solution of (12) starting at θ(0)=θ0. For all s,t[0,1ε],
(33)
(34)
Proof.
Equation (33) is a standard result about the differentiable dependence of solutions on their initial condition. Now, let ϕ(τ)=θ(τε) for each τ[0,1]; so, ϕ satisfies Eq. (32). The derivative of the map rrF(ϕ(τ),τ) at r=1ε is obviously just F(ϕ(τ),τ), and so we have

It will also be useful to define the function

such that, given any ε, for any lift θ^:[0,1ε]R of a solution θ:[0,1ε]S1 of (12), we have

Since solutions θ() of (12) have |θ˙(t)|M, we have that

(35)

for all ε>0 and θ0S1.

Assume that F is generic.

If F has no zeros, then it is clear that (,S1) is a complete curve of the stable slow manifold (and obviously there is no other).

Now, suppose that F has zeros, and let C={τ[0,1]:θS1 s.t.F(θ,τ)=0}, i.e., C is the image of the set of zeros of F under (θ,τ)τ. Since F is continuous, we obviously have that C is closed. Note that if we have a complete curve (U,y) of the stable slow manifold, then C=U¯.

By Remark 9(A), if we have τ[0,1] such that S1×{τ} includes a hyperbolic stable or unstable zero of F, then τ is in the interior of C relative to [0,1]. Hence, for every boundary point τ of C relative to [0,1], S1×{τ} only contains a non-hyperbolic zero of F.

Now, if there are infinitely many τ[0,1] for which S1×{τ} contains a non-hyperbolic zero, then, in particular, we can find a convergent sequence of distinct non-hyperbolic zeros (θn,τn) converging to a zero (θ,τ); but by Remark 9, no zero of F can have non-hyperbolic zeros of F arbitrarily close to it. Hence, there are only finitely many τ[0,1] for which S1×{τ} contains a non-hyperbolic zero. Hence, in particular, C has only finitely many boundary points, and, therefore, only finitely many connected components. Furthermore, by Remark 9, no connected component of C can be a singleton.

So, let C1,,Cn be the connected components of C arranged in increasing order. For each connected component Ci=[ai,bi] of C,

  • if ai0, then since S1×{ai} contains exactly one zero of F and this zero is non-hyperbolic, on the basis of Remark 9(B) there must exist δ>0 and a curve y of the stable slow manifold over (ai,ai+δ) such that for each τ(ai,ai+δ), (yτ,τ) is the only hyperbolic stable zero of F in S1×{τ}; but

  • if ai=0 (obviously implying, in particular, that i=1), then for each hyperbolic stable zero (θ0,0) of F in S1×{0}, on the basis of Remark 9(A), we have that for some δ>0, there is a unique curve y of the stable slow manifold over [0,δ) for which y(0)=θ0.

In either case, let τi(1) be the supremum of the set of all τ>ai for which there is a unique curve of the stable slow manifold over (ai,τ) agreeing with y on (ai,ai+δ); note that if ai=0, then τi(1) will depend on θ0. It is clear that this “supremum” is indeed a maximum, i.e., that there is a unique curve of the stable slow manifold over (ai,τi(1)) agreeing with y on (ai,ai+δ). So, let us extend the domain of y to include (ai,τi(1)). Now, let σn be a strictly increasing sequence converging to this maximum τi(1) such that y(σn) is convergent to some limit l. We have that (l,τi(1)) is a zero of F; and so by Remark 9 together with the fact that y is continuous on (ai,τi(1)), we have that, in fact, y(τ)l as ττi(1). If τi(1)=1, then (l,τi(1)) cannot be non-hyperbolic and so is a hyperbolic stable zero, and, therefore, the domain of y can be extended to include 1 itself.

Thus, if τi(1)=bi, then we have already defined y on the whole of the interior of Ci relative to [0,1]. So, now suppose that τi(1)bi. Note that (l,τi(1)) must be non-hyperbolic, otherwise by Remark 9(A), τi(1) would not be the supremum of the set of times up to which we can uniquely extend y. Since (l,τi(1)) is non-hyperbolic, Remark 9(B) implies that there are no zeros (θ,τ) close to (l,τi(1)) with τ>τi(1), and, therefore, S1×{τi(1)} must contain zeros other than just (l,τi(1)). So, let (θ1,τi(1)) be the zero for which there is a fast connection from (l,τi(1)) to (θ1,τi(1)). Since l is a local extremum of F(,τi(1)), it is clear that θ1 is unique. Since (l,τi(1)) is non-hyperbolic, (θ1,τi(1)) cannot also be non-hyperbolic and, therefore, (θ1,τi(1)) must be hyperbolic stable. Hence, Remark 9(A) gives that for some δ1>0, there is a unique curve of the stable slow manifold over (τi(1),τi(1)+δ1) whose right-sided limit at τi(1) coincides with θ1. So, extend the domain of y to include the open interval (τi(1),τi(1)+δ1), with y|(τi(1),τi(1)+δ1) being the unique curve of the stable slow manifold over (τi(1),τi(1)+δ1) for which y(τ)θ1 as ττi(1). Now, define τi(2) as the maximum of the set of all τ>τi(1) for which there is a unique curve of the stable slow manifold over (τi(1),τ) agreeing with y on (τi(1),τi(1)+δ1), and extend y to include (τi(1),τi(2)). We can then treat τi(2) the same way we treated τi(1) earlier; and continuing the procedure as necessary, we obtain an increasing list of terms τi(1),τi(2),τi(3), that terminates at the term τi(mi) for which τi(mi)=bi. Note that this termination must take place due to the fact that there are only finitely many τ-values for which S1×{τ} contains a non-hyperbolic zero.

Thus, overall, we have defined y on the set

where τi(0)=ai. By construction, if 0C, then (U,y) is the unique complete curve of the stable slow manifold, and if 0C, then for our given θ0, (U,y) is the unique complete curve of the stable slow manifold fulfilling y(0)=θ0.

In this section, we will prove Proposition 26 and Theorem 24(A) as well as obtaining a corollary of Theorem 24(A) that will be needed later for the proof of Theorem 24(C).

For each n1, let en:RR be the continuous function for which en(x)xn is the remainder in the order-(n1) Taylor expansion of ex; in other words, en(x)=i=0xi(i+n)!. Note, in particular, that en is strictly increasing on [0,). It is easy to check that for any c,tR,

and

If F has no zeros, then we will work with all the notations introduced in Proposition 26, and also let

Proof of Proposition 26

proof
Proof of Proposition 26
Let τ be a global minimizer of τm1(τ)k(τ). Given ε>0, define an integer n1 and times
such that
  • ti+1=ti+k(εti) for each 1i<n, and

  • τε[tn,tn+k(εtn));

and then define an integer n~1 and times
such that
  • t~i+1=t~ik(εt~i) for each 1i<n~, and

  • tn(t~n~k(εt~n~),t~n~].

We will approximate 01r(τ)dτ by Riemann-like sums over the partition of [0,1] whose set of end points is given by {εti}1in{εt~i}1in~. Note that the quantities ti+1ti, t~it~i+1, and t~n~tn are all bounded by the ε-independent constant K:=maxτ[0,1]k(τ), and hence the mesh size of this partition tends to 0 as ε0.

For any continuous function g:[0,1]R, define
Let
(36)
(37)
We have that r(εti)riri and r(εt~i)r~ir~i; hence,
and, therefore,
(38)
as ε0. We also have that
and hence
(39)
as ε0. So, to prove the desired result, by Eqs. (38) and (39), it is sufficient to show that
(40)
for all θ0S1.
Fix θ0S1. Let θ:[0,1ε]S1 be the solution of (12) starting at θ(0)=θ0, and let θ^:[0,1ε]R be a lift of θ. For each 1i<n, let ψi:RS1 be the solution of the autonomous differential equation
(41)
for which ψi(ti)=θ(ti), and let ψ^i:RR be a lift of ψi with ψ^i(ti)=θ^(ti). For all t[ti,ti+1], we have
and so a suitable version of Grönwall’s inequality (Ref. 50, Corollary 2) gives that
Now, all solutions of (41) are k(εti)-periodic, and so Eq. (33) applied to the τ-independent function F(θ,εti) gives
Hence,
By the same reasoning applied to the time-reversal of (12) (i.e., applied to F1,0 in place of F), we likewise have that for all 1i<n~,
Finally,
Combining our bounds on |λ¯ti,ti+1ε,F(θ0)|, |λ¯t~i+1,t~iε,F(θ0)|, and |λ¯tn,t~n~ε,F(θ0)| yields the required Eq. (40).

Proof of Theorem 24(A)

proof
Proof of Theorem 24(A)
Continuing with the notations in the proof of Proposition 26, and letting
we have that riR for all 1i<n and r~iR for all 1i<n~, and we also have that
Hence,
and hence, by Eq. (40),
for all ε>0 and θ0S1. Now, it is easy to see that C[Fσ1,σ2]C[F], for any σ1,σ2[0,1]. Hence,
(42)
for all s,t[0,1ε]. So, by Proposition 6, F is neutrally stable with constant c=eC[F]. It is clear that C[] is locally bounded (in fact, continuous) with respect to the norm F, and hence F is robustly neutrally stable.

The following corollary will play a similar role in the proof of Theorem 24(C) to the role played by Lemma 2 of Ref. 17 in the main results of Ref. 17.

Corollary 45
Suppose F has no zeros, and let
Then, for all ε>0 and θ0S1,
where c=eC[F] is as in the proof of Theorem 24(A).
Proof.

Follows immediately from Eqs. (42) and (34).

For the sake of completeness, in  Appendix A, we give bounds on the limiting range of Φ(ε1)(1ε,θ0,0,1) as ε0.

Now, as an immediate consequence of Corollary 45, we obtain that if F has no zeros, then

(43)

for all ε>0 and θ0S1. [In fact, as one would intuitively expect, it is not hard to show by elementary means that

uniformly across θ0S1 as ε0. Further results on this convergence are given in Ref. 17 between Eqs. (4.20) and (4.22).]

We first give the result that provides the connection between tracking and stability in Theorem 24(B) and (C); it is a consequence of Theorem 24(A) or Proposition 26. In the following, Leb() denotes the Lebesgue measure on [0,1].

Lemma 46
Suppose F is generic, and let (U,y) be a complete curve of the stable manifold, with Λad being the corresponding adiabatic Lyapunov exponent. Let SU be a Borel set. For all η>0, there exist δ0,ε0>0 such that for any ε(0,ε0) and δ(0,δ0), if θ() is a solution of (12) that δ-tracks (S,y|S) over [0,1ε], then [writing θ(0)=:θ0],
Proof.
Essentially by definition,
with
Now, first observe that for any solution θ() of (12) (with any ε>0), we have
(44)
Now, fix η>0. Taking δ0=η3M11, we have that for any ε>0 and δ(0,δ0), if θ() is a solution of (12) that δ-tracks (S,y|S) over [0,1ε], then
(45)
Now, let W[0,1]U¯ be a disjoint union of finitely many intervals [wi,wi+], i=1,,n, such that
So, for any solution θ() of (12) (with any ε>0), we have
(46)
For each 1in, Fwi,wi+ has no zeros; and so on the basis of Theorem 24(A) or Proposition 26, let ε0>0 be such that for all ε(0,ε0) and 1in, every solution θ() of (12) has
(47)
Combining the four Eqs. (44)–(47) yields the desired result.

Let us now introduce a few further tracking-related definitions. In the following, we assume that we have a set V[0,1], a function q:VS1, and a point pS1.

Definition 47

Fix ε>0. Given Δ,δ>0, we say that (12)exhibits (Δ,δ)-tracking of (V,q) away from p if every solution θ() of (12) with d(θ(0),p)Δ2δ-tracks (V,q).

Definition 48

We say that Fexhibits strict tracking of(V,q)away fromp if given any Δ,δ>0, for sufficiently small ε, (12) exhibits (Δ,δ)-tracking of (V,q) away from p.

Definition 49

Assume that V is σ-compact. We say that F exhibits tracking of (V,q) away from p if for every compact subset S of V{0}, F exhibits strict tracking of (S,q|S) away from p.

Proposition 50

Suppose F is generic and not initially multistable, and that U¯ is connected and 0U. Let (z0,0) be the unique hyperbolic unstable zero of F in S1×{0}. Then, F exhibits tracking of (U,y) away from z0.

In other words, the conclusion of Proposition 50 states that by taking ε sufficiently small, we can guarantee that solutions θ(t) that do not start too close to z0 will be close to y(εt) at all times t1εS, where SU is bounded away from {0}U.

This is fairly obvious given the definition of (U,y). Unsurprisingly, to write out a rigorous proof would involve no conceptual difficulties or subtleties but would, nonetheless, be technically tedious. Therefore, we will omit a proof.

Combining Proposition 50 with Lemma 46 gives the following immediate consequence:

Proposition 51

In the setting of Proposition 50, F is exponentially stable away from z0 with rate Λad.

Proof.

Fix η,Δ>0, let SU{0} be a compact set with 2M1Leb(US)<η2, let δ0,ε0 be the values given by Lemma 46 applied to S with η2 in place of η, and take any δ(0,δ0). On the basis of Proposition 50, suppose ε<ε0 is sufficiently small that (12) exhibits (Δ,δ)-tracking of (S,y|S) away from z0. Then, every solution θ() of (12) with θ(0)BΔ2(z0) has |λ0,1εε,F(θ(0))Λad|η.

Assume that F is generic, S1×{0} contains no zeros, and U¯ is connected.

To show that F exhibits tracking of (U,y), we need to show that for any compact SU and any Δ,δ>0, if ε is sufficiently small, then there is an arc Pε of length less of Δ such that every solution of (12) starting outside of Pεδ-tracks (S,y|S). Fix S, Δ and δ. Let τ(0)=infU and let σ0=minS. We have that S1×{τ(0)} contains exactly one zero of F, namely, a non-hyperbolic zero; so, on the basis of Remark 9(B), choose a value τ~(τ(0),σ0) such that the only zeros in S1×{τ~} are (y(τ~),τ~) and a hyperbolic unstable zero (z(τ~),τ~). Note that Fτ~,0 is generic and not initially multistable, and the corresponding complete curve (Uτ~,0,yτ~,0) of the stable slow manifold has

Take any Δ~<2d(y(τ~),z(τ~)) and let P~=BΔ~2(z(τ~)). We have that

  • by Proposition 50 applied to Fτ~,1, if ε is sufficiently small then every solution θ() of (12) with θ(τ~ε)P~δ-tracks (S,y|S), and

  • by Proposition 51 applied to Fτ~,0, since y(τ~) has a neighborhood that does not intersect P~, the length of the arc
    (48)
    tends to 0 exponentially as ε0; and so length(Pε)<Δ for sufficiently small ε.

Combining (a) and (b) gives the desired result.

Having shown that F exhibits tracking of (U,y), combining this with Lemma 46 gives that F is exponentially stable with rate Λad, by exactly the same reasoning as in the proof of Proposition 51.

We start with the following general result.

Proposition 52

Suppose FF is such that for some τ(0,1),

  • F has no zeros in S1×(τ,1];

  • for all τ[0,τ), S1×{τ} contains exactly two zeros of F, namely, a hyperbolic stable zero (y(τ),τ) and a hyperbolic unstable zero (z(τ),τ); and

  • letting
    we have sFFτ(y(τ),τ)>0 and sFFτ(z(τ),τ)>0 for all τ[0,τ).

Then,
Proof.
First, take θ0 with sFF(θ0,0)>0. Let θ() be the solution of (12) with θ(0)=θ0, and let θ^() be a lift of θ(). By the implicit function theorem, on the whole of [0,τ), the derivative of τy(τ) has sign sF and the derivative of τz(τ) has sign sF. So, it is clear that θ(t) cannot cross past either y(εt) or z(εt) during t[0,τε). Therefore, sFθ˙(t)=sFF(θ(t),εt)>0 for all t[0,τε), and so
It follows that if we now let θ0 be an arbitrary point in S1, then
Furthermore, obviously sFF(θ(t),εt)>0 for all t(τε,1ε]. And, taking an arbitrary τ(τ,1), we have that μ:=min(θ,τ)S1×[τ,1]sFF(θ,τ)>0. Hence,

Now, suppose F is generic and not initially multistable, and U¯ is not connected. Let [τ0,τ0+] be the first connected component of U¯ and let τ1 be the lower end point of the second connected component of U¯. Fix an arbitrary σ(τ0+,τ1). On the basis of Remark 9(B) applied to the non-hyperbolic zero in S1×{τ0+} and in S1×{τ1}, we can find τ~0<τ0+ and τ~1>τ1 such that Fτ~0,σ and Fτ~1,σ each fulfill the conditions of Proposition 52. For each τ(τ1,τ~1], let (z(τ),τ) be the hyperbolic unstable zero in S1×{τ}. Let

and fix τ(τ1,τ~1] sufficiently close to τ1 that

Fix an arbitrary η>0 sufficiently small that Λ~+η<Λ0η.

Now, we will show that for every compact S(τ1,τ] and every Δ,δ>0, there are intervals of ε-values reaching arbitrarily close to 0 for which one can find an arc Pε of length less than Δ such that every solution θ() of (12) starting outside Pεδ-tracks (S,z|S). Without loss of generality, take S of the form S=[τ,τ] for some τ(τ1,τ). Without loss of generality, take δ<d(y(τ),z(τ)), and let Q=Bδ(z(τ)). Since Fτ~1,σ fulfills the conditions of Proposition 52, we have, in particular, that δ<d(y(τ),z(τ)). Fix any Δ>0. Also define the positive quantities

where z denotes the derivative of z. Using Theorem 24(B) applied to F0,σ and Proposition 51 applied to Fτ,σ [where in the latter case, we note that y(τ) has a neighborhood not intersecting the arc Q], we can find ε0>0 such that for all ε<ε0, the following statements hold:

  1. There is an arc Pε of length less than Δ such that, writing
    we have
  2. Writing
    we have
  3. We have εν<υ; and therefore, for any solution θ() of (12), if θ(τε)Q, then d(θ(t),z(εt))<δ for all t with εtS.

Note that J~~(ε) depends continuously on ε. One can also take Pε and hence J~(ε) to depend continuously on ε: if 0U, then by Proposition 51 Pε can be taken independent of ε, and if 0U then one can see from the proof of Theorem 24(B) [see, in particular, Eq. (48)] that Pε can be taken to depend continuously on ε. By Proposition 52 applied to Fτ,σ, J~~(ε) moves unboundedly round the circle as ε0. Also, by the tracking statement in Theorem 24(B) applied to F0,τ~0, it is clear that the set

remains within an ε-independent proper subset of S1 for all sufficiently small ε, and, therefore, applying Proposition 52 to Fτ~0,σ gives that J~(ε) moves unboundedly round the circle as ε0 in the opposite direction to J~~(ε). However, by points (1) and (2), we also have that length(J~(ε))<length(J~~(ε)) for all sufficiently small ε. Hence, there are intervals of ε-values arbitrarily close to 0 for which J~(ε)J~~(ε). For each such ε, every solution θ() of (12) starting outside Pε has θ(τε)Q and hence, by point (3), d(θ(t),z(εt))<δ for all t with εtS.

We start with the following general result, which is a consequence of Corollary 45.

Proposition 53
In the setting of Proposition 52, every θ0S1 with sFF(θ0,0)>0 has
Proof.
Take an arbitrary τ(τ,1). By Eq. (34), for any ε>0 and θ0S1, letting θ() be the solution of (12) with θ(0)=θ0, we have
By Corollary 45 applied to Fτ,1, we have that
and as in the proof of Proposition 52, we have that if sFF(θ0,0)>0, then sFF(θ(u),εu)0 for all u. Hence, the result.

Now, suppose F is generic and not initially multistable, and U¯ has n2 connected components. Let

be the end points of the connected components of U¯. For each 1i<n, take an arbitrary value σ~i(τi+,τi+1)[0,1]U¯.

To show that F exhibits tracking of (U,y), we need to show that for any compact SU{0} and any Δ,δ>0, if ε is sufficiently small, then there is an arc Pε of length less of Δ such that every solution of (12) starting outside of Pεδ-tracks (S,y|S). Fix S, and for each 1in, let

Without loss of generality, assume that for all 1i<n, σi+ is sufficiently close to τi+ that Fσi+,σ~i fulfills the conditions of Proposition 52. For each 1i<n, let (z(σi+),σi+) be the hyperbolic unstable zero in S1×{σi+}. Without loss of generality, let δ>0 be such that for all 1i<n, δ<d(y(σi+),z(σi+)). Let Qi=Bδ(y(σi+)) for each 1i<n. Fix Δ>0, and on the basis of Theorem 24(B) applied to F0,σ1+, for all sufficiently small ε, let Pε be an arc of length less than Δ such that every solution θ() of (12) starting outside Pεδ-tracks (S[τ1,τ1+],y|S[τ1,τ1+]).

For each 2in, pick a value τi(τi,σi) sufficiently close to τi that Fτi,σ~i1 fulfills the conditions of Proposition 52. For each 2in, let (z(τi),τi) be the hyperbolic unstable zero in S1×{τi}, and let P(i) be a neighborhood of z(τi) such y(τi)P¯(i). For each 1i<n and each ε>0, let

Using Proposition 50 applied to Fτi,σi+ for all 2in, we obtain that for every sufficiently small ε>0, if

(49)

then all solutions of (12) starting outside Pεδ-track S. So, it remains to show that the set of ε-values for which (49) fails is exponentially vanishing. For each 1i<n, let

It is clear that a finite union of exponentially vanishing sets is exponentially vanishing, so it is sufficient to show that Ei is exponentially vanishing for each 1i<n.

Fix i. By Proposition 51 applied to Fσi+,σ~i and Fτi+1,σ~i, we can find λ<0 such that for all sufficiently small ε, J(i)(ε) and J(i)+(ε) are both of length less than eλε. Let

Let θ~iQi and θ~i+P(i+1) be such that siF(θ~i,σi+)>0 and siF(θ~i+,τi+1)>0. Let x1:(0,)R be a lift of the function TΦ(T,θ~i,σi+,σ~i), and let x2:(0,)R be a lift of the function TΦ(T,θ~i+,τi+1,σ~i). Define x:(0,)R by x=si(x1x2), and let dxd(ε1):(0,)R denote the derivative of x. For sufficiently small ε, we have

(50)

By Eq. (35) applied to Fσi+,σ~i and Fτi+1,σ~i, we have that

(51)

for all ε>0. By Proposition 53 applied to Fσi+,σ~i and Fτi+1,σ~i, there exists μ>0 such that

(52)

Applying elementary analysis to Eqs. (50)–(52) yields that E¯i is exponentially vanishing. Specifically, this analysis is as follows:

Given jN and sufficiently small ε>0, if

then

and so,

where α>0 and κ(0,1). So, letting

we have that

for all sufficiently small ε. So, it is sufficient to show that the set {ε>0:x(1ε)Xi} is exponentially vanishing; this is justified as follows: for each ξ>0,

Thus, we have shown that F almost exhibits tracking of (U,y). Combining this with Lemma 46 gives that F is almost exponentially stable with rate Λad, by exactly the same reasoning as in the proof of Proposition 51.

Recall from the discussion of the normal form (20) in Sec. III G that the graph of the map x2D(fγ,x2) was a semi-ellipse. Since the quadratic function Fγnormal is only an approximation of the behavior of (Fγ)γΓ around the critical zero, we will need to consider perturbation of ellipses. So, before addressing the specific setting of Sec. III G, let us first give some general notations and a basic fact regarding ellipses and perturbation of ellipses.

For any τ0R and r1,r2>0, define the ellipse

i.e., Sτ0,r1,r2 has center (τ0,0) and its principal axes are a horizontal axis of radius r1 and a vertical axis of radius r2. We will use the following notation for perturbation of this ellipse: given hR5 with r1+h1>0 and r2+h2>0, define

Now, for Δ(0,1), define the solid elliptical annulus

Also, define Sτ0,r1,r2 and Aτ0,r1,r2(Δ) to be the intersection, respectively, of Sτ0,r1,r2 and Aτ0,r1,r2(Δ) with the lower half-plane R×(,0).

We now address the question of how strongly one can perturb the ellipse Sτ0,r1,r2 in the manner described above while remaining within Aτ0,r1,r2(Δ).

Lemma 54
Given Δ(0,1) and hR5 with r1+h1>0 and r2+h2>0, if
then S~τ0,r1,r2(h)Aτ0,r1,r2(Δ).
Proof.
First, note that the transformation
sends Sτ0,r1,r2 and Aτ0,r1,r2(Δ) onto S0,1,1 and A0,1,1(Δ), respectively, and also sends S~τ0,r1,r2(h) onto the set S~0,1,1(h) with
So, without loss of generality take τ0=0 and r1=r2=1.
Note that the nearest point on A0,1,1(Δ) to any given point on S0,1,1 is of distance Δ away. Now, S0,1,1 is mapped onto S~0,1,1(h) by first applying the transformation
and then the transformation
The first transformation moves points horizontally through a maximum distance of |h1|+|h3| and vertically through a maximum distance of |h2|+|h4|. Since |h1|+|h3|12Δ, the horizontal coordinate of the resulting points must remain within distance 1+12Δ of 0. Therefore, the application of the second transformation moves points through a maximum vertical distance of (1+12Δ)|h5| while leaving the horizontal coordinate the same. Thus, overall, in going from S0,1,1 to S~0,1,1(h), the horizontal and vertical distances moved by each point are a maximum of 12Δ, and, therefore, the overall distance is a maximum of Δ.

Now, given a 2×2 real symmetric matrix A and a vector xR2, write

It is easy to verify the following:

  • If det(A)<0, then for all x2R{0}, the quadratic function x1A[x1,x2] takes both positive and negative values.

  • If det(A)>0, then sgn(a11)=sgn(a22)0, i.e., diag-sgn(A) is well-defined, and furthermore,
    for all (x1,x2)R2{(0,0)}.

Now, for FC2(S1×[0,1],R), if we fix (θ,τ)S1×[0,1] and vR2 and define a function w() by

then the second derivative of w is given by

(53)

It is now straightforward to prove Lemma 29.

Proof of Lemma 29

proof
Proof of Lemma 29
Let us fix vectors v,v~R2 such that HessF(θ0,τ0)[v]>0 and HessF(θ0,τ0)[v~]<0 and define
for t in a neighborhood of 0. Since F(θ0,τ0)=0, we have that w(0)=w~(0)=0, and, therefore, by Eq. (53), w()>0 and w~()<0 on a punctured neighborhood of 0. It follows, in particular, that v and v~ are linearly independent. Now, for all t in such a punctured neighborhood, the intermediate value theorem gives a point (θt,τt) on the line segment joining w(t) and w~(t) such that F(θt,τt)=0; and since v and v~ are linearly independent, we have that (θt,τt)(θ0,τ0).

We now prove Theorem 30.

Standing assumption. Without loss of generality, take diag-sgn(D1)=1 and d2>0.

For convenience, for all small δ>0, define

Note that Fγ0 is strictly positive on (S1×[0,1]){(θ0,τ0)}. Therefore, by continuity, for all small δ>0, we can find r(δ)(0,δ) such that Fγ(θ,τ)>0 for all

Now, take δ1>0 sufficiently small that the following statements hold:

  • det(HessFγ(θ,τ))>0 for all (γ,θ,τ)Cδ1 and

  • sgn(Fγ(θ0,τ0))=sgn(γγ0) for all γΓδ1.

Define positive finite constants m22,M12, and m0 such that the following statements hold:

  • for all (γ,θ,τ)Cδ1 and all |v|=1,
  • for all γΓδ1,

Let δ2=r(δ1). We will prove Theorem 30 in the following steps:

  • Step 1. There does not exist a sequence (γn,θn,τn) converging to (γ0,θn,τ0) such that F(γn,θn,τn)=0 and γn>γ0 for all n.

  • So, we can define δ3>0 sufficiently small that for all γ(γ0,γ0+δ3), Fγ has no zeros in Sδ3; and hence, for all γ(γ0,γ0+r(δ3)), Fγ has no zeros at all. We then take δ in the statement of Theorem 30 to be min(δ2,r(δ3)).

  • Step 2. For all γ(γ0δ2,γ0), Fγ is generic with a unique complete curve (Uγ,yγ) of the stable slow manifold, and Uγ is non-empty and connected. Write
    so that
  • Step 3. We will propose an approximation (U~γ,Y~γ) of (Uγ,Yγ), based on the normal form (20); this gives
  • Step 4. For any η>0, taking γ>γ0 sufficiently close to γ0 guarantees that

1. Step 1

Suppose for a contradiction that such a sequence (γn,θn,τn) exists. For all sufficiently large n, we have that γnΓδ1, and so (θn,τn)(θ0,τn). So, let tn>0 and v(n) with |v(n)|=1 be such that

By Taylor’s theorem, there exists t~n(0,tn) such that, letting

we have

and, therefore,

For sufficiently large n, we have (γn,θ~n,τ~n)Cδ1 and so it follows that

and, therefore,

However, this contradicts the fact that γnγ0 as n.

2. Step 2

Fix γ(γ0δ2,γ0). Recall that Fγ is strictly positive outside Sδ1 and so all zeros of Fγ lie inside Sδ1. For each τTδ1, we have that HessFγ()[1,0] is strictly positive on the whole of Θδ1×{τ} and so Θδ1×{τ} contains at most two zeros, and if Θδ1×{τ} contains two zeros, then one is hyperbolic stable [which we denote (yγ(τ),τ)], and the other is hyperbolic unstable [which we denote (zγ(τ),τ)]. Let

By Remark 9(A), Uγ is open and yγ() and zγ() are continuous on Uγ. Now, Fγ(τ0,θ0)<0, and so by the intermediate value theorem, Θδ1×{τ0} has at least (and hence exactly) two zeros. In other words, τ0Uγ. So, let (τγ,τγ+) be the connected component of Uγ containing τ0. Since the set of zeros of F is closed, it is clear that Θδ1×{τγ} and Θδ1×{τγ+} each contain a zero of F, which we denote (θγ,τγ) and (θγ+,τγ+), and it is clear that both yγ(τ) and zγ(τ) converge to θγ± as ττγ±. Hence, in particular, by Remark 9(A), (θγ,τγ) and (θγ+,τγ+) are non-hyperbolic zeros.

We next show that (θγ,τγ) and (θγ+,τγ+) are non-degenerate. By definition of δ1, since (γ,θγ±,τγ±)Cδ1, we have 2Fγθ2(θγ±,τγ±)>0. Now, take v±R2 such that

Since Fγ(θ0,τ0)<0 and HessFγ()[v±] is strictly positive throughout the line segment

we have that

Since Fθ(θγ±,τγ±)=0, it follows that Fτ(θγ±,τγ±)0.

Finally, we show that there are no zeros of Fγ other than those which we have already mentioned, namely, those in the set

By definition, all zeros in Θδ1×(τγ,τγ+) belong to this set. So, it remains to show that given any

we have Fγ(θ,τ)>0. Assume that τ(τ0δ1,τγ); the case that τ(τγ+,τ0+δ1) proceeds in exactly the same way. Take t>1 and vR2 such that

Once again, we have that Fγ(θ0,τ0)<0 and HessFγ()[v] is strictly positive throughout the line segment

Therefore, splitting into the cases that

  • θ0+v1=θγ,

  • θ0+v1(θγ,θ0+δ1),

  • θ0+v1(θ0δ1,θγ),

we have the following: In case (i), Fγ(s(1))=0 and so Fγ(θ,τ)=Fγ(s(t))>0. In case (ii), by the intermediate value theorem, there exists t(0,1) such that, letting τ=τ0+tv2, we have

so Fγ(s(t))=0 and hence, again, Fγ(θ,τ)>0. Case (iii) is the same, with yγ in place of zγ.

3. Step 3

For convenience, let ρ1=2dθθd2det(D1) and ρ2=2dθθd2. For each γ<γ0, define the set U~γ and the negative-valued function Y~γ() on U~γ such that

For each τU~γ, we can write Y~γ(τ) more explicitly as

4. Step 4

Fix η>0, and let Δ(0,1) be sufficiently small that the area of the elliptical semi-annulus A0,ρ1,ρ2(Δ) is at most η. For each γ(γ0δ2,γ0), we have that Yγ<0 on Uγ and Yγ tends to 0 at the boundary points of Uγ. Therefore, to show the desired result, it is sufficient to show that for all γ<γ0 sufficiently close to γ we have

For each γ(γ0δ2,γ0), let δ(γ)>0 be the smallest value for which all zeros of Fγ are contained in S¯δ(γ). Note that δ(γ)0 as γγ0: for any small η~>0, taking γ(γ0r(η~),γ0) gives δ(γ)<η~.

Now, for each γ(γ0δ2,γ0), for each τUγ, we can apply Taylor’s theorem to the functions Fγ(θ0,), Fγ(,τ), and Fγθ(θ0,), respectively, to obtain expressions

where d~ττγ,τ, d~θθγ,τ, and d~θτγ,τ are, respectively, the values of 2Fγτ2, 2Fγθ2, and 2Fγθτ at certain points in Sδ(γ). Since δ(γ)0 as γγ0, letting

we have that D~1γ,τ converges to D1 uniformly across τUγ as γγ0. Also, recall that |Fγ(θ0,τ0)| is O(γ0γ) as γγ0. Therefore, in the above three equations, if we substitute the expressions for Fγ(θ0,τ) and Fγθ(θ0,τ) given by the first and third equation into the second equation and then rearrange the second equation to make yγ(τ) the subject, we obtain an expression of the form

(54)

where υγ,τ=Fγθ(θ0,τ) is as given by the third equation, and Eγ,τ takes the form

where

  • aγ,τ0,

  • bγ,τ is O(γ0γ), and

  • cγ,τ is o(γ0γ),

uniformly across τUγ as γγ0. Now, by applying Taylor’s theorem to Fγθ(,τ), using Eq. (54), we obtain an expression

where d~~θθγ,τ is the value of 2Fγθ2 at some point in Sδ(γ). Again, d~~θθγ,τdθθ uniformly across τUγ as γγ0. Therefore, one obtains that, after suitable modification of the coefficients aγ,τ,bγ,τ,cγ,τ (while keeping the same convergence properties as above), we have

where

Now, for convenience, let us write a:=det(D1) and cγ:=2dθθd2(γ0γ), so that

The point (τ,Yγ(τ)) lies on the curve

with h4γ,τ and h5γ,τ as above and

The general inequalities |AB||AB| and A+B|A|+|B| (understood as valid for all A,BR for which they are well-defined) give

So,

  • h1γ,τ and h2γ,τ are o(γ0γ),

  • h3γ,τ and h4γ,τ are, respectively, O(γ0γ) and o(γ0γ), and hence are both o(γ0γ), and

  • h5γ,τ0,

uniformly across τUγ as γγ0. Hence, in particular, taking γ sufficiently close to γ0 will mean that hγ,τ fulfills the condition in Lemma 54 (with ri=ρiγ0γ) for all τUγ, and hence

This completes the proof.

The modern theory of dynamical systems was introduced by Henri Poincaré and Aleksandr Lyapunov to provide a rigorous mathematical basis for defining and investigating qualitative dynamical properties of real-world physical systems—properties such as synchronization, stability, and neutral stability. Using a numerically simulated sample realization of a Brownian bridge as a conceptual representation of inherently finite-time processes, we have illustrated in Fig. 2 how a finite-time system can exhibit precisely such properties in a way that cannot be modeled so as to make the framework of Poincaré and Lyapunov applicable. This is because this framework is defined in terms of the infinite-time behavior of dynamical systems.

Therefore, we have introduced an alternative, finite-time framework for stability analysis by essentially “replacing t with ε0” in the traditional formalisms of qualitative stability analysis, where ε represents the timescale separation between a system’s “internal timescales” and the slower timescale of a finite-time external forcing process. This non-traditional framework has enabled us to provide a rigorous mathematical statement (Theorem 24) and proof of the stabilization phenomenon of Refs. 1 and 2 exemplified in Fig. 2.

We have also explored how our new framework of stability analysis compares with the traditional framework through the example of low-frequency-periodically forced systems, where both frameworks can be applied. In particular, we saw in Sec. IV E how analysis based on Theorem 24 can more readily yield the basic description of dynamics than seeking to analyze the system within the traditional apparatus of stability analysis (such as asymptotic Lyapunov exponents).

This paper has worked entirely with one-dimensional dynamics but we hope that the work here will also prompt further research into non-traditional qualitative stability analysis of higher-dimensional finite-time systems. Whereas the work of Poincaré was originally motivated by celestial mechanics, we anticipate that finite-time stability theory will increasingly play a fundamental and necessary role in the understanding of more complex “terrestrial” systems such as biological processes and climate systems.

This is TiPES contribution #127. This project received funding from the European Union’s Horizon 2020 Research and Innovation Programme under Grant Agreement No. 820970 (TiPES), the European Union’s Horizon 2020 Research and Innovation programme under the Marie Skłodowska-Curie Grant Agreement No. 642563, an EPSRC Doctoral Prize Fellowship, the DFG (Grant No. CRC 701), and the EPSRC (Grant No. EP/M006298/1). The authors would like to express their gratitude to Peter Ashwin, Boštjan Dolenc, Edgar Knobloch, Colin Lambert, Peter McClintock, Woosok Moon, Arkady Pikovsky, Antonio Politi, Martin Rasmussen, Michael Rosenblum, Janne Ruostekoski, David Sloan, and Yevhen Suprunenko for interesting and insightful discussions.

The authors have no conflicts to disclose.

Codes for the numerics carried out in this paper (as described in  Appendix B) are available from the corresponding author upon reasonable request and are openly available in Lancaster Publications and Research system Pure at https://doi.org/10.17635/lancaster/researchdata/292, Ref. 51.

Proposition 26 can be expressed as a statement about the limiting behavior of Φθ(1ε,θ0,0,1) as ε0. We now give an analogous statement about the limiting behavior of Φ(ε1)(1ε,θ0,0,1) as ε0. Let us first make a couple of general remarks regarding Φ(ε1):

  • Given t1,,tn[0,1ε], using Eq. (34), we have
    (A1)
  • Suppose F(,τ)=:f() is independent of τ [meaning that (12) is just a time-restricted autonomous dynamical system]. Then, Φ(ε1)(1ε,θ(s),εs,εt) is, by definition, the derivative of Tθ(s+ε(ts)T) at T=1ε, which is ε(ts)θ˙(t). In other words, we have
    (A2)

We use the notations present in Proposition 26, as well as the following:

Proposition 55
Suppose F has no zeros. Then,
and
Proof.
Given ε>0, define an integer N1 and times
such that
  • t~i+1=t~ik(εt~i) for each 1i<N and

  • 0[t~Nk(εt~N),t~N).

For each 1iN, for any continuous function g:[0,1]R, let
and let r~i be as in Eq. (37). Similarly to the proof of Proposition 26, we have that
and hence,
as ε0. Hence,
(A3)
and
(A4)
as ε0. Now, letting θ() be a solution of (12) with θ(0)=θ0, as in the proof of Proposition 26, we have
and so the left-hand sides of (A3) and (A4) are, respectively, a lower and an upper bound of the expression
Therefore, to prove the desired result, we will show that there are constants ε0>0 and E0 (dependent only on F) such that for all ε(0,ε0) and θ0S1,
(A5)
Given ε>0 and θ0S1, let θ() be the solution of (12) with θ(0)=θ0, and let
for each 1iN. By Eq. (A1),
and it is clear from Eq. (34) that sgn(F)di0, and so we have
where R is as in the proof of Theorem 24(A). Letting K=maxτ[0,1]k(τ), Eq. (34) implies
and so it remains to find an ε-independent bound on 1εi=1N1|diεk(εt~i)F(θ(t~i),εt~i)|. For each 1i<N, let ψi() be the solution of
for which ψi(t~i)=θ(t~i). Writing Fi for the τ-independent function Fi(θ,τ)=F(θ,εt~i), Eqs. (34) and (A2) applied to Fi give
Hence, applying Eq. (34) also to di, we have
By the same reasoning as in the proof of Proposition 26, for each u[t~i+1,t~i], we have
So, if ε<ε0, then
We also have, again by the same reasoning as in Proposition 26, that for each u[t~i+1,t~i],
Hence, if ε<ε0, then for all u[t~i+1,t~i],
So,
and hence,
Thus, overall, Eq. (A5) is satisfied with

Let us note that the bound E obtained in the above proof can probably be considerably improved through more detailed calculation and less estimation; but we have obtained a cruder bound for the sake of a faster proof.

Throughout this paper, additively forced Adler equations θ˙(t)=asin(θ(t))+G(t) were simulated by numerical integration using a fourth order Runge–Kutta scheme, with a time step of 0.01 s. The FTLEs associated with trajectories θ(t) were calculated simply by taking the time-averaged value of acos(θ(t)) over the time-interval of interest, as in (11). To obtain the initial condition θ(0) for a given final state θ(T), the value of θ(T) was used as the initial condition of a forward-time simulation of the equation θ˙(t)=asin(θ(t))G(Tt).

In Sec. II B, the function g(t) was constructed as follows. First, a sample path Wt of a zero-drift Brownian motion on [0,T] of diffusion parameter σ=1T was constructed by cumulative addition of Gaussian random increments over a time step of 0.01 s. Then, the Brownian bridge was constructed as WttTWT. Then, the result was passed through a fifth order Butterworth low-pass filter with cut-off frequency 1/(2π×103) Hz, performed via cascaded second-order sections (in Python, with the function scipy.signal.sosfilt), and we linearly interpolated the output of the filter.

In Figs. 5 and 6, the ω-values indicated in red were numerically obtained as follows: For the unwrapped phase x(t) as governed by the differential equation

on the real line, setting x(0)=0, it was observed by simulation that x(2πω) increased approximately linearly with 1/ω, with increments across consecutive values in the (1/ω)-discretization being strictly positive and very small compared to 2π. Hence, it was possible to carry out linear interpolation of the simulated wrapped phase θ(2πω) as a function of 1/ω [with θ(0)=0]. Where this linearly interpolated function of 1/ω crossed π2 and 3π2 is where the ω-values were marked; as in Fig. 5(c), the locations of θ(2πω) are indistinguishably the same for the other 49 initial conditions θ(0)=2πi50 as for θ(0)=0.

FIG. 6.

Dynamics of (23) with varying ω near a red-marked point in Fig. 5(b). Parameters a, k, and A are as in Fig. 5. For each ω-value, in hollow blue are shown the finite-time Lyapunov exponents λT associated with the trajectories of 10 initial conditions θ(0)=2πi10 with T=200πω (i.e., 100 periods), and in solid orange are shown the finite-time Lyapunov exponents λT associated with the trajectories of three initial conditions θ(0)=0,2π3,4π3 with T=2000πω (i.e., 1000 periods). The red dashed line indicates the approximate location of a small interval of ω-values for which all trajectories have zero asymptotic Lyapunov exponent. The horizontal gray line marks the value of Λa,k,A defined in (31).

FIG. 6.

Dynamics of (23) with varying ω near a red-marked point in Fig. 5(b). Parameters a, k, and A are as in Fig. 5. For each ω-value, in hollow blue are shown the finite-time Lyapunov exponents λT associated with the trajectories of 10 initial conditions θ(0)=2πi10 with T=200πω (i.e., 100 periods), and in solid orange are shown the finite-time Lyapunov exponents λT associated with the trajectories of three initial conditions θ(0)=0,2π3,4π3 with T=2000πω (i.e., 1000 periods). The red dashed line indicates the approximate location of a small interval of ω-values for which all trajectories have zero asymptotic Lyapunov exponent. The horizontal gray line marks the value of Λa,k,A defined in (31).

Close modal

The basis for this procedure is as follows: since k>a and A(ka,k+a), Proposition 41(B) can be applied to the time-0-to-2πω mapping Φ of (23) to give that for all sufficiently small ω, Φ maps all points starting outside some small arc P into a small arc Q. By the symmetries of (23), we have that Q is close to P if and only if Q is close to either π2 or 3π2. Now, on the one hand, when P and Q are not close to each other, it is clear by considering the graph of Φ that Φ has a stable fixed point near Q and an unstable fixed point near P, and thus system (23) is exponentially stable in the sense of Definition 35. However, on the other hand, if Q and P cross past each other (which, again, is equivalent to Q crossing past either π2 or 3π2) as ω is varied then, again by considering the graph of Φ, during such crossing there must occur an interval of ω-values for which Φ has no fixed points and, therefore, by Proposition 37, the system (23) is neutrally stable in the sense of Definition 33 (see also Ref. 17).

1.
R. V.
Jensen
, “
Synchronization of driven nonlinear oscillators
,”
Am. J. Phys.
70
,
607
619
(
2002
).
2.
M.
Lucas
,
J.
Newman
, and
A.
Stefanovska
, “
Stabilization of dynamics of oscillatory systems by nonautonomous perturbation
,”
Phys. Rev. E
97
,
042209
(
2018
).
3.
V.
Anishchenko
,
T.
Vadivasova
, and
G.
Strelkova
, “
Stochastic self-sustained oscillations of non-autonomous systems
,”
Eur. Phys. J. Spec. Top.
187
,
109
125
(
2010
).
4.
P. E.
Kloeden
and
M.
Rasmussen
,
Nonautonomous Dynamical Systems
(
American Mathematical Society
,
Providence, RI
,
2011
).
5.
P. T.
Clemson
and
A.
Stefanovska
, “
Discerning non-autonomous dynamics
,”
Phys. Rep.
542
,
297
368
(
2014
).
6.
G.
Haller
, “
Finding finite-time invariant manifolds in two-dimensional velocity fields
,”
Chaos
10
,
99
108
(
2000
).
7.
F.
Lekien
,
S. C.
Shadden
, and
J. E.
Marsden
, “
Lagrangian coherent structures in n-dimensional systems
,”
J. Math. Phys.
48
,
065404
(
2007
).
8.
A.
Berger
,
T. S.
Doan
, and
S.
Siegmund
, “
Nonautonomous finite-time dynamics
,”
Discrete Contin. Dyn. Syst. B
9
,
463
492
(
2008
).
9.
M.
Rasmussen
, “
Finite-time attractivity and bifurcation for nonautonomous differential equations
,”
Differ. Equ. Dyn. Syst.
18
,
57
78
(
2010
).
10.
T. S.
Doan
,
D.
Karrasch
,
T. Y.
Nguyen
, and
S.
Siegmund
, “
A unified approach to finite-time hyperbolicity which extends finite-time Lyapunov exponents
,”
J. Differ. Equ.
252
,
5535
5554
(
2012
).
11.
D.
Karrasch
, “
Linearization of hyperbolic finite-time processes
,”
J. Differ. Equ.
254
,
256
282
(
2013
).
12.
L. H.
Duc
,
J. P.
Chávez
,
T. S.
Doan
, and
S.
Siegmund
, “
Finite-time Lyapunov exponents and metabolic control coefficients for threshold detection of stimulus–response curves
,”
J. Biol. Dyn.
10
,
379
394
(
2016
).
13.
P.
Giesl
and
J.
McMichen
, “
Determination of the area of exponential attraction in one-dimensional finite-time systems using meshless collocation
,”
Discrete Contin. Dyn. Syst. B
23
,
1835
1850
(
2018
).
14.
B.
Kaszás
,
U.
Feudel
, and
T.
Tél
, “
Leaking in history space: A way to analyze systems subjected to arbitrary driving
,”
Chaos
28
,
033612
(
2018
).
15.
C.
Kuehn
, Multiple Time Scale Dynamics, Applied Mathematical Sciences, Vol. 191 (Springer, Cham, 2015).
16.
P.
Ashwin
,
S.
Wieczorek
,
R.
Vitolo
, and
P.
Cox
, “
Tipping points in open systems: Bifurcation, noise-induced and rate-dependent examples in the climate system
,”
Philos. Trans. R. Soc. London, Ser. A
370
,
1166
1184
(
2012
).
17.
J.
Guckenheimer
and
Y. S.
Ilyashenko
, “
The duck and the devil: Canards on the staircase
,”
Moscow Math. J.
1
,
27
47
(
2001
).
18.
Z.
Hagos
,
T.
Stankovski
,
J.
Newman
,
T.
Pereira
,
P. V. E.
McClintock
, and
A.
Stefanovska
, “
Synchronization transitions caused by time-varying coupling functions
,”
Philos. Trans. R. Soc. A
377
,
20190275
(
2019
).
19.
W.
Kurebayashi
,
S.
Shirasaka
, and
H.
Nakao
, “
Phase reduction method for strongly perturbed limit cycle oscillators
,”
Phys. Rev. Lett.
111
,
214101
(
2013
).
20.
W.
Kurebayashi
,
S.
Shirasaka
, and
H.
Nakao
, “
A criterion for timescale decomposition of external inputs for generalized phase reduction of limit-cycle oscillators
,”
Nonlinear Theory Appl.
6
,
171
180
(
2015
).
21.
Y.
Park
and
G. B.
Ermentrout
, “
Weakly coupled oscillators in a slowly varying world
,”
J. Comput. Neurosci.
40
,
269
281
(
2016
).
22.
A.
Stefanovska
,
D. G.
Luchinsky
, and
P. V. E.
McClintock
, “
Modelling couplings among the oscillators of the cardiovascular system
,”
Physiol. Meas.
22
,
551
564
(
2001
).
23.
Y. F.
Suprunenko
,
P. T.
Clemson
, and
A.
Stefanovska
, “
Chronotaxic systems: A new class of self-sustained nonautonomous oscillators
,”
Phys. Rev. Lett.
111
,
024101
(
2013
).
24.
M.
Lucas
,
D.
Fanelli
, and
A.
Stefanovska
, “
Nonautonomous driving induces stability in network of identical oscillators
,”
Phys. Rev. E
99
,
012309
(
2019
).
25.
D.
Lukarski
,
M.
Ginovska
,
H.
Spasevska
, and
T.
Stankovski
, “
Time window determination for inference of time-varying dynamics: Application to cardiorespiratory interaction
,”
Front. Physiol.
11
,
702
(
2020
).
26.
J.
Newman
,
A.
Pidde
, and
A.
Stefanovska
, “
Defining the wavelet bispectrum
,”
Appl. Comput. Harmon. Anal.
51
,
171
224
(
2021
).
27.
J.
Newman
,
M.
Lucas
, and
A.
Stefanovska
, “Non-asymptotic-time dynamics,” in Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 7, pp. 111–129.
28.
M.
Lucas
,
J.
Newman
, and
A.
Stefanovska
, “Synchronization and non-autonomicity,” in Physics of Biological Oscillators, edited by A. Stefanovska and P. V. E. McClintock (Springer, 2021), Chap. 6, pp. 85–110.
29.
A.
Pikovsky
,
M.
Rosenblum
, and
J.
Kurths
,
Synchronization: A Universal Concept in Nonlinear Sciences
(
Cambridge University Press
,
Cambridge
,
2003
), Vol.
12
.
30.
A. S.
Pikovskii
, “
Synchronization and stochastization of array of self-excited oscillators by external noise
,”
Radiophys. Quantum Electron.
27
,
390
395
(
1984
).
31.
F.
Flandoli
,
B.
Gess
, and
M.
Scheutzow
, “
Synchronization by noise
,”
Probab. Theory Relat. Fields
168
,
511
556
(
2017
).
32.
V. A.
Antonov
, “
Modeling of processes of cyclic evolution type. Synchronization by a random signal
,”
Vestnik Leningrad. Univ. Mat. Mekh. Astronom.
vyp. 2
,
67
76
(
1984
).
33.
D.
Malicet
, “
Random walks on Homeo(S1)
,”
Commun. Math. Phys.
356
,
1083
1116
(
2017
).
34.
A. E.
Hramov
,
A. A.
Koronovskii
,
M. K.
Kurovskaya
, and
O. I.
Moskalenko
, “
Analytical expression for zero Lyapunov exponent of chaotic noised oscillators
,”
Chaos, Solitons Fractals
78
,
118
123
(
2015
).
35.
As in Sec. III B of Ref. 1, this is allowed to include a forcing function that happens to have arisen as a sample realization of a random process. The point is that—in contrast to the phenomenon that we explained in Sec. II A 2—the actual mechanism of stabilization here makes no reference to probability-theoretic properties of the underlying random process.
36.
Z. F.
Mainen
and
T. J.
Sejnowski
, “
Reliability of spike timing in neocortical neurons
,”
Science
268
,
1503
1506
(
1995
).
37.
J. D.
Hunter
,
J. G.
Milton
,
P. J.
Thomas
, and
J. D.
Cowan
, “
Resonance effect for neural spike time reliability
,”
J. Neurophysiol.
80
,
1427
1438
(
1998
).
38.
R. V.
Jensen
,
L.
Jones
, and
D. H.
Gartner
, “Synchronization of randomly driven nonlinear oscillators and the reliable firing of cortical neurons,” in Computational Neuroscience: Trends in Research, 1998, edited by J. M. Bower (Springer US, Boston, MA, 1998), pp. 403–408.
39.
R. M.
Dudley
, Uniform Central Limit Theorems, Cambridge Studies in Advanced Mathematics (Cambridge University Press, 1999).
40.
M.
Callaway
,
T. S.
Doan
,
J. S. W.
Lamb
, and
M.
Rasmussen
, “
The dichotomy spectrum for random dynamical systems and pitchfork bifurcations with additive noise
,”
Ann. Henri Poincaré B
53
,
1548
1574
(
2017
).
41.
P.
Ashwin
,
C.
Perryman
, and
S.
Wieczorek
, “
Parameter shifts for nonautonomous systems in low dimension: Bifurcation- and rate-induced tipping
,”
Nonlinearity
30
,
2185
2210
(
2017
).
42.
U.
Feudel
,
A. N.
Pisarchik
, and
K.
Showalter
, “
Multistability and tipping: From mathematics and physics to climate and brain—Minireview and preface to the focus issue
,”
Chaos
28
,
033501
(
2018
).
43.
L.
Arnold
, Random Dynamical Systems, Springer Monographs in Mathematics (Springer-Verlag, Berlin, 1995), pp. xvi + 586.
44.
One potentially confusing imperfection in the analogy: the analogous situation to “f having no equilibria” is not that “there is no complete curve of the stable slow manifold,” but is rather that “the unique complete curve of the stable slow manifold is the empty curve.”
45.
Y. S.
Ilyashenko
,
D. A.
Ryzhov
, and
D. A.
Filimonov
, “
Phase-lock effect for equations modeling resistively shunted Josephson junctions and for their perturbations
,”
Funct. Anal. Appl.
45
,
192
(
2011
).
46.
P.
Gandhi
,
E.
Knobloch
, and
C.
Beaume
, “
Dynamics of phase slips in systems with time-periodic modulation
,”
Phys. Rev. E
92
,
062914
(
2015
).
47.
V. M.
Buchstaber
and
A. A.
Glutsyuk
, “
On monodromy eigenfunctions of Heun equations and boundaries of phase-lock areas in a model of overdamped Josephson effect
,”
Proc. Steklov Inst. Math.
297
,
50
89
(
2017
).
48.
P.
Giesl
and
M.
Rasmussen
, “
Borg’s criterion for almost periodic differential equations
,”
Nonlinear Anal.: Theory Methods Appl.
69
,
3722
3733
(
2008
).
49.
S.
Wieczorek
,
C.
Xie
, and
C. K. R. T.
Jones
, “
Compactification for asymptotically autonomous dynamical systems: Theory, applications and invariant manifolds
,”
Nonlinearity
34
,
2970
3000
(
2021
).
50.
S. S.
Dragomir
,
Some Gronwall Type Inequalities and Applications
(
Nova Science Publishers
,
2003
).
51.
J.
Newman
,
M.
Lucas
, and
A.
Stefanovska
, “
Codes and data for paper “Stabilisation of cyclic processes by slowly varying forcing”
,” Lancaster University (
2019
).