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

Estimating Phase from Observed Trajectories Using the Temporal 1-Form

Simon Wilshin Royal Vet College, London, UK    Matthew D. Kvalheim University of Pennsylvania, Philadelphia, PA, USA    Clayton Scott University of Michigan, Ann Arbor, MI, USA    Shai Revzen University of Michigan, Ann Arbor, MI, USA
Abstract

Oscillators are ubiquitous in nature, and usually associated with the existence of an “asymptotic phase” that governs the long-term dynamics of the oscillator. We show that asymptotic phase can be estimated using a carefully chosen series expansion which directly computes the phase response curve and provide an algorithm for estimating the co-efficients of this series. Unlike all previously available data driven phase estimation methods, our algorithm can: (i) use observations that are much shorter than a cycle; (ii) recover phase within any forward invariant region for which sufficient data are available; (iii) recover the phase response curves (PRC-s) that govern weak oscillator coupling; (iv) show isochron curvature, and recover nonlinear features of isochron geometry. Our method may find application wherever models of oscillator dynamics need to be constructed from measured or simulated time-series.

I Phase as an organizing principle

Our interest in oscillators arose from their utility as models of animal locomotion (Revzen and Kvalheim, 2015; Seipel et al., 2017), but oscillators appear in virtually every physical science: in biological models (Kvalheim and Bloch, 2021) at all scales, from the coupled neuronal firing Hoppensteadt and Izhikevich (1997) to coupled oscillations of predator and prey populations May (1972); in chemistry Epstein and Pojman (1998); in physics Głazek and Wilson (2002); in electrical Van der Pol (1934), civil Duffing (1918), and mechanical Stoker (1950) engineering; etc. In the typical case that transients decay at an exponential rate, the underlying mathematical structure of oscillators shares many properties across all these cases, in particular the existence of an “asymptotic phase” which encodes the long term outcome of any (recoverable) perturbation. We have discovered that by representing phase as a “Temporal 1-Form”, we could bring to bear interpolation tools from machine learning to allow asymptotic phase to be computed in a data driven way—directly from time series of measurements—and without a need to know the governing equations.

This will allow investigators to perform “phase reduction” directly from experimental measurements, and thereby construct models of oscillatory systems and how they couple to each other for a broad swath of systems. In this paper we present our algorithm applied to simulations with known ground truth, and some classical oscillators (Fitzhugh-Nagumo oscillator FitzHugh (1961); Nagumo et al. (1962), Selkov oscillator Sel’Kov (1968)) from the literature. Additional simulation tests, and examples of use with animal locomotion data and chemical oscillator data are included, with the details of our simulations in the supplemental information (SI). These tests demonstrate the broad applicability of our methods.

II Differential equations

Mathematical models in the physical sciences frequently use differential equations of order 1 or higher. Newton’s laws, and the equations governing electrical circuits are usually written as second order systems x¨=g(x,x˙)\ddot{x}=g(x,\dot{x}), whereas first order systems govern x˙=f(x)\dot{x}=f(x) simple chemical reactions. By re-writing the second order equations in terms of positions and velocities (or momenta), the former can be reduced to a special case of the latter, and we will therefore consider first order systems.

Sometimes the solutions of x˙=f(x)\dot{x}=f(x) are periodic. We consider the case that the curve traced by such a solution attracts nearby solutions. More specifically, we consider only the typical case that this convergence occurs at an exponential rate, which is equivalent to the periodic solution being (normally) “hyperbolic”; for brevity, we will simply refer to such periodic solutions as “limit cycles”. We use the term “oscillator” to describe the dynamics f()f(\cdot) restricted to the set of initial conditions which, after a sufficiently long relaxation period, converge arbitrarily close to the designated limit cycle. That set of initial conditions which evolve toward the limit cycle is termed the “stability basin” of the limit cycle or the domain of the oscillator.

Once the system settles on (or arbitrarily near to) the limit cycle it oscillates with the same pattern repeatedly, allowing points from this closed curve to be invertibly mapped onto the unit circle so that they cycle at a constant angular rate ω\omega. The angle of the image on this circle is what is commonly refered to as the “phase (angle)” of the oscillator at that state.

This phase can be naturally extended to the entire stability basin thanks to the following observation: all initial conditions converge to the limit cycle, but points on the limit cycle forever remain apart. To model the long term behavior of an oscillator we could represent each off-cycle point by a point on the cycle which most closely represents its long term behavior, constructing a phase map. In the typical case that convergence to the limit cycle is at an exponential rate, such long-term representatives actually exist (and are unique), so, by extension, the phase angle of the representative can be taken to be the phase angle of all points it represents.

Since we can assign to every point in our stability basin a phase angle, we can divide up the stability basins into sections (hypersurfaces) with the same phase. In the mathematics literature these are called “isochrons”. The fact that isochrons are manifolds and therefore have no cusps or kinks is a non-trivial result (Guckenheimer, 1975). For a DD dimensional system this will be a set of D1D-1 dimensional hypersurfaces on each of which phase is constant. In Euclidean space the derivatives of the phase are the normals of these surfaces. These derivatives at the intersection points with the limit cycle govern the long term sensitivity of the system to small perturbations, and are called the “phase response curves”. They provide a complete first order description of the long-term responses of the oscillator to small perturbations.

III Estimating phase by series expansion

We note that for our deterministic systems our phase, ϕ\phi satisfies the following condition: ϕ(x(t))=ωt+ϕ(x(0))\phi(x(t))=\omega t+\phi(x(0)) when following a trajectory x()x(\,\cdot\,) for a system with angular frequency ω\omega. This is obviously true on the limit cycle. Off of the limit cycle, x(0)x(0) and x(t)x(t) are tt time units apart on the same solution if and only if x(s)x(s) and x(s+t)x(s+t) are as well for any ss. Considering ss values sufficiently large that x(s)x(s) and x(s+t)x(s+t) are effectively on the limit cycle, the result follows. Rearranging slightly we have:

ϕ(x(s+t))ϕ(x(s))=ωt\phi(x(s+t))-\phi(x(s))=\omega t (1)

Taking the limit t0t\to 0:

ω=limt0ϕ(x(s+t))ϕ(x(s))t=ddtϕ(x(t))|s\omega=\lim_{t\rightarrow 0}\frac{\phi(x(s+t))-\phi(x(s))}{t}=\left.\frac{d}{dt}\phi(x(t))\right|_{s} (2)

Since x(s)x(s) is an arbitrary point, and x˙(s)=f(x(s))\dot{x}(s)=f(x(s)) we obtain from the chain rule applied to eqn. (2):

dϕ(x)f(x)=ωd\phi(x)\cdot f(x)=\omega (3)

Here dd is intended to evoke the exterior derivative 𝐝\mathrm{\mathbf{d}}{}111 This is the exterior derivative and for Riemannian spaces when acting on a scalar is the operator \nabla with its contravariant index lowered by the metric. Since our method is not metric dependent we use 𝐝\mathrm{\mathbf{d}}{} here, however readers unfamiliar with differential geometry can treat 𝐝\mathrm{\mathbf{d}}{} as \nabla. , which represents the set of first order partial derivatives when acting on a scalar. But dϕd\phi is actually not globally the exterior derivative of any smooth real-valued function (ϕ\phi is not continuous on all of 𝐗\mathbf{X}), and we reserve the boldface 𝐝\mathrm{\mathbf{d}}{} for actual exterior derivatives.

This form has a few advantages. First, it permits us to write down a point-wise condition our phase function must observe, rather than a condition about its differences along trajectories. Second, phase angle has an arbitrary gauge in that it is defined only up to choice of a point at phase 0, but the form dϕd\phi does not and is, at least potentially, a physical quantity. This coordinate free form which we dubbed the “Temporal 1-Form” (SI sec. A.3) is unique for each oscillator, is mathematically equivalent to other representations of asymptotic phase (Theorem 1 in SI sec. A.4), and can be approximated using our algorithm with roughly the same statistical efficiency as the law of large numbers for data sampling and system and measurement noise (Theorems 2 and 3 and Remark 3 in SI sec. A.5). This provable convergence property is unique among all data-driven phase estimation algorithms we are aware of.

As mentioned, the function ϕ\phi does not have a conventional derivative everywhere. First, it suffers a jump discontinuity when moving from 2π2\pi to 0. However, a suitable derivative can be calculated at any point by noting that: (1) Under a different gauge ϕ~(x):=ϕ(x)θ0\tilde{\phi}(x):=\phi(x)-\theta_{0} for phase angle, the jump would occur at the isochron ϕ(x)=θ0\phi(x)=\theta_{0} instead of ϕ(x)=0\phi(x)=0. (2) Wherever they are both defined, 𝐝ϕ=𝐝ϕ~\mathrm{\mathbf{d}}{\phi}=\mathrm{\mathbf{d}}{\tilde{\phi}}. Thus we may, in a self-consistent way, define dϕd\phi on all of 𝐗\mathbf{X} to be given by 𝐝ϕ\mathrm{\mathbf{d}}{\phi} or 𝐝ϕ~\mathrm{\mathbf{d}}{\tilde{\phi}} depending on which quantity exists.

Second, phase is undefined at the boundaries of the stability basin (and everywhere else outside the stability basin). If the solution starting at point yy does not return to the limit cycle, ϕ\phi is not defined at yy. If, in addition, y=limkxky=\lim_{k\to\infty}x_{k} of points xkx_{k} that do return to the limit cycle, then ϕ(xk)\phi(x_{k}) need not converge 222While it need not, it could; e.g. θ˙=1+cr;r˙=(r2)(r1)r\dot{\theta}=1+cr;\dot{r}=(r-2)(r-1)r has phase identically θ\theta for c=0c=0, but phase diverges at r=2r=2 when c0c\neq 0. .

IV Algorithm

Refer to caption
Figure 1: Steps of the algorithm, showing phase (as color; isochrons as contours), trajectories (white traces), and a limit cycle estimate (dark line) as applied to a randomly generated simulation system.We projected the data down to its two principal components, and then built a limit cycle model from the polar angle in the circulation plane [A]. Using this we rectified the radius to the unit circle [B], and then rectified the phase to a constant circulation rate [C]. We then add basis terms (up to order 22 [D], order 66 [E]) and convert the resulting series approximation back to the original coordinates [F].

Our algorithm takes as input pairs (xk,x˙k)(x_{k},\dot{x}_{k}) k=1Nk=1\ldots N of observed states and state velocities, and learns both an angle valued function that takes states to phase, and a matching vector valued function giving the Temporal 1-Form.

We perform our algorithm in the following way. First we rectify the data by constructing a smooth change of coordinates that maps an estimated limit cycle to the unit circle at constant angular rate. Then subtract this constant circulation and expand the remainder with vector valued basis functions in the rectified coordinates and select their coefficients by minimizing the cost of a loss function which is zero for the residual produced from the true Temporal 1-Form. The basis functions we use here are derivatives of known scalar valued functions. The sum of the scalar valued functions and the phase angle contribution of the circulation is our resulting angle valued phase estimate.

IV.1 Rectification

The first step of our algorithm is fairly similar to Revzen & Guckenheimer Revzen and Guckenheimer (2008) in that it estimates the limit cycle, and computes a phase estimate that evolves uniformly on that limit cycle estimate. However, here we use this estimate to construct a smooth coordinate change that rectifies the data so that the estimated limit cycle is the unit circle in the first two coordinates.

In particular, we translate the data to zero mean, then use principal component analysis (PCA) on the positions xkx_{k} to rotate the two major covariance axes into the first two coordinates. We assume that in these new coordinates the trajectories wind around the origin, and that data is mostly constrained to an annulus in this 2D “circulation plane” comprising the first two coordinates. Data which does not meet this assumption would require pre-processing before it could be used—some other transformation, more elaborate than PCA, would be needed to bring it to meet the circulation requirement.

Consider the data in cylindrical coordinates: scalar angle θ\theta and radius rr for the circulation plane, and cartesian zz with the remaining coordinates of each data point (see figure 1 [A] showing such data). We fit Fourier series r^(θ)\hat{r}(\theta) and z^(θ)\hat{z}(\theta) to the data, and thereby constructed an approximate model of the limit cycle. This also allowed us to estimate the period TT as the rate of circulation of this limit cycle model. Using the same approach as Phaser Revzen and Guckenheimer (2008), we constructed a map ϕ^(θ)\hat{\phi}(\theta) so that |2πt/T(ϕ^+θ)|2\left|2\pi t/T-(\hat{\phi}+\theta)\right|^{2} is minimised. Using this we constructed the map mapping the input xx coordinates to rectified coordinates

(θ,r,z)(θϕ^(θ),r/r^(θ),zz^(θ))\displaystyle(\theta,r,z)\mapsto(\theta-\hat{\phi}(\theta),r/\hat{r}(\theta),z-\hat{z}(\theta)) (4)

which rectifies the data to motion on the unit circle (see figure 1 [B]) at close to a constant angular rate (see figure 1 [C]). We refer to the state xx in its rectified coordinates by the symbol qq.

IV.2 Approximation by topologically motivated basis functions

We reconstructed the Temporal 1-Form by a series approximation using a basis 333We use the term “basis” informally, as it is used e.g. in the notion of “radial basis functions” in machine learning, to mean a collection of functions whose finite linear spans are dense in the function space of interest. described below. Most of the forms comprising this basis were themselves derivatives of real-valued functions; we will refer to such as “basis forms”.

Unfortunately, it is not possible to reconstruct the Temporal 1-Form from any linear combination of basis forms, for the following reasons. Consider the change in the value of ϕ\phi after a single cycle on the limit cycle. For a single cycle this will be 2π2\pi (or 11, or 360360^{\circ} depending on the arbitrary units used). This difference can be computed using a line integral of the Temporal 1-Form  which amounts to just integrating eqn. (3). If we build dϕd\phi exclusively from sums of derivatives of any real-valued functions, this integral will necessarily be zero, since the sum of the integrals will just be the integral of the sum. This demonstrates that the Temporal 1-Form is not exact, i.e. it is not the derivative of any scalar valued function, and therefore cannot be the sum of basis forms.

The key insight that enabled our algorithm comes from algebraic topology 444 Although the closed 1-form we seek is not exact, algebraic topology teaches us that, on the basin of attraction associated to an oscillator, any two closed 1-forms α,b\alpha,b can be related as a=ωb+ca=\omega b+c for some ω\omega real and cc exact. Exact forms are exterior derivatives (“gradients”) of scalar functions. The authors are deeply indebted to Dan Guralnik for pointing out this insight on the de Rham cohomology of oscillators. That this insight extends from 𝒞\mathcal{C}^{\infty} closed forms (used in de Rham cohomology) to 𝒞0\mathcal{C}^{0} closed forms follows from general results on smoothing currents (de Rham, 1984, pp. 61-70); alternatively, it is not difficult to give a direct proof of this result for oscillators. and is that we can always write our expression for the Temporal 1-Form of one system, aa, in terms of another Temporal 1-Form for another system, bb, via a=ωb+𝐝ca=\omega b+\mathrm{\mathbf{d}}{c} with ω\omega real and cc real-valued. We chose to use a trivial oscillator for bb by using dθd\theta with θ\theta the angle in the plane. This dθd\theta is the Temporal 1-Form associated with any oscillator obeying equations of the form θ˙=ω\dot{\theta}=\omega r˙=g(r)\dot{r}=g(r).

dϕ(θ,r,z):=dθ+\displaystyle d\phi(\theta,r,z):=d\theta+ μmμ𝐝vμ(θ,r,z)\displaystyle\sum_{\mu}m_{\mu}\mathrm{\mathbf{d}}{v_{\mu}(\theta,r,z)} (5)
μ:=(i,j,k)0in2,j,k;vμ(θ,r,z):=ξi(z)ρj(r)uk(θ)\displaystyle\mu:=(i,j,k)~{}~{}0\leq i\leq n-2,~{}j,k\in\mathbb{N};~{}~{}v_{\mu}(\theta,r,z):=\xi_{i}(z)\rho_{j}(r)u_{k}(\theta)
Name Integral Differential 1-form
angle θarctan2(q0,q1)(no global integral exists)\displaystyle\theta\approx\text{arctan2}(q_{0},q_{1})~{}~{}\text{{(no global integral exists)}} dθ\displaystyle d\theta (6)
Fourier uk(θ):=akcos(kθ)+bksin(kθ)\displaystyle u_{k}(\theta):=a_{k}\cos(k\theta)+b_{k}\sin(k\theta) 𝐝uk(θ)=kaksin(kθ)dθkbkcos(kθ)dθ\displaystyle\mathrm{\mathbf{d}}{u_{k}}(\theta)=ka_{k}\sin(k\theta)d\theta-kb_{k}\cos(k\theta)d\theta (7)
polynomial ρj(r):=(r1)j\displaystyle\rho_{j}(r):=(r-1)^{j} 𝐝ρj(r)=j(r1)j1𝐝r\displaystyle\mathrm{\mathbf{d}}{\rho_{j}}(r)=j(r-1)^{j-1}\mathrm{\mathbf{d}}{r} (8)
out of plane ξi(z):=zifor i>0, or1, for i = 0\displaystyle\xi_{i}(z):=z_{i}~{}\text{for i>0, or}~{}1\text{, for i = 0} 𝐝ξi(z)=𝐝zifor i>0, or0, for i = 0\displaystyle\mathrm{\mathbf{d}}{\xi_{i}}(z)=\mathrm{\mathbf{d}}{z_{i}}~{}\text{for i>0, or}~{}0\text{, for i = 0} (9)
Table 1: Basis forms and their integrals in rectified coordinates θ,r,z\theta,r,z. The general dϕd\phi structure is in [5]. After the rectification step, dθd\theta is a trivial expression [6], and provides the circulation needed to ensure that remaining terms can be exact—the sum of derivatives 𝐝vμ\mathrm{\mathbf{d}}{v_{\mu}}. The vμv_{\mu} basis functions are a first order expansion in zz, via [9], a Fourier expansion in θ\theta via [7], and a polynomial [8] expansion in rr.

Our approximation algorithm proceeds by optimizing the parameters of eqn. (5) (in table 1) as indicated by the following expression, where the constant CC is selected to be consistent with the previously computed period:

argmindϕi(dϕx˙iC)2.\arg\min_{d\phi}\sum_{i}\left({d\phi\cdot\dot{x}_{i}-C}\right)^{2}. (10)

Here we have multiple observations of state and its derivative, (x,x˙)(x,\dot{x}) and hence xx and x˙\dot{x} have an index, ii. First, working in the rectified co-ordinates qq, we selected a scale for dθ(q)d\theta(q) such that Γ𝑑θ(q)\int_{\Gamma}d\theta(q) is the period of oscillation, where Γ𝐗\Gamma\subset\mathbf{X} is the image of the limit cycle.

Refer to caption
Figure 2: A plot of uk(θ)ρj(r)u_{k}(\theta)\rho_{j}(r) terms for orders up to five. From left to right the order of the Fourier term increases and the cosine and sine terms alternate. From top to bottom we have increasing order of the radial polynomial term. In the top right corner is the dθd\theta term.

To find the remaining coefficients mμm_{\mu} we solved

argminmμ,Cμ,i(mμ𝐝vi(qi)q˙i(Cdθ(qi)q˙i))2.\arg\min_{m_{\mu},C}\sum_{\mu,i}\left({m_{\mu}\mathrm{\mathbf{d}}{v_{i}}(q_{i})\cdot{\dot{q}}_{i}-\left({C-d\theta(q_{i})\cdot{\dot{q}}_{i}}\right)}\right)^{2}. (11)

We note that eqn. (11) is a conventional ordinary least squares regression problem, and thus mμm_{\mu} can be solved for using standard tools. The contribution of individual Fourier-polynomial terms can be seen in figure 2.

V Performance of the new algorithm

V.1 Performance on 2D, 3D, and 8D simulations

We used the approach described in Appendix sec. A.1 to produce three simulation systems of dimension two, three and eight. The first is easy to visualize; the second is the minimal dimension in which general eigenvalues can appear in the Floquet structure; the third is more typical of the dimensionality of biological systems for which we developed this estimation method. We examined the performance of three different estimates of the phase of these systems and compared the estimated phase to the ground truth provided by θ\theta of eqn. (A.1.1).

The first phase estimator we used is comparable to “event-based” phase estimates commonly used in biological research. Researchers often use distinguished events, such as voltage levels in the nervous system Danner et al. (2015), footfall Ting et al. (1994); Jindrich and Full (2002), or anterior extreme position of a limb Cruse and Schwarze (1988) to identify the beginning of a cycle and presume that phase evolves uniformly in time between these events. As our event, we used the zero crossing of the first principal component of the data. The second phase estimator we used was our previously published Phaser algorithm Revzen and Guckenheimer (2008). The third phase estimator was derived from the Temporal 1-Form as described in this paper, and is referred to as “form phase”. A visualization of the execution of the form-phase estimation algorithm is depicted in figure 1.

Even though the phase we wish to reconstruct is that of a deterministic dynamical system, experimental data more closely resembles the sample paths of a stochastic differential equation. We trained our phase estimation methods on one set of simulated sample paths and tested them on another, generated from the same stochastic dynamical system. We calculated a model residual by subtracting the ground truth phase from the estimated phases obtained with each method. We removed the trial-to-trial indeterminate phase offset by taking the circular mean of the model residual to be zero.

The simulations along with the code used to generate them are available at https://purl.archive.org/purl/formphase. The results show that the form phase method has lower mean-square residuals than the Phaser method, which in turn has lower mean-square residual than the event-based method; see residuals in figure 3.

Table 2: Variance of residual phase for different phase estimation techniques with different initial condition noise and system noise levels.
noise res. var.
D initial system phase event Phaser form
2 0.1 0.01 0.1 0.128 0.0473 0.0185
2 0.2 0.01 0.1 0.127 0.0456 0.0205
2 0.1 0.02 0.1 0.135 0.0480 0.0199
2 0.1 0.01 0.2 0.222 0.0885 0.0389
3 0.066 0.0066 0.066 0.102 0.156 0.0164
3 0.133 0.0066 0.066 0.102 0.0920 0.0189
3 0.066 0.0133 0.066 0.108 0.0863 0.0214
3 0.066 0.0066 0.133 0.159 0.0897 0.0252
8 0.025 0.0025 0.025 0.0789 0.0232 0.0340
8 0.05 0.0025 0.025 0.0776 0.0246 0.0455
8 0.025 0.005 0.025 0.0816 0.0352 0.0671
8 0.025 0.0025 0.05 0.0889 0.0273 0.0460
Refer to caption
Figure 3: Comparison of phase estimators on 2D, 3D, and 8D systems. Comparison of event phase (green) Phaser (black) and form phase (turquoise) phase noise distributions on simulated trajectories. Plots show four different conditions for each dimension: one baseline condition (top left), an increase in the magnitude of the stochastic diffusion term (top right), an increase in the levels of variability in the initial conditions (bottom left) and an increase in noise on the coordinate corresponding to phase in the equivalent deterministic system (bottom right; see table 2).

V.2 Selkov and Fitzhugh Nagumo oscillators

Refer to caption
Figure 4: Comparison of stochastically driven Selkov oscillator isochrons obtained from the form-phase, Phaser and event-based phase estimates. We plotted trajectories (blue), the limit cycle (true cycle—black, data driven estimate—teal) and ground-truth isochrons from forward integration (orange lines). On top of those we plotted scatter plots of trajectory points falling in 20 equally spaced intervals of width 0.02 period; ([A], green, alternating shades), Phaser ([B], black), and form-based phase estimates ([C], gray). Event phase estimates become noticeably poorer away from event; Phaser estimates remain equally tight, but their angle to the limit cycle—the PRC—is incorrect; Form phase isochrons closely match the forward integration “ground truth”, that is to say the angle between the thin lines in [C] are more acute than the angle between the black dots and orange lines in [B])
Refer to caption
Figure 5: Temporal 1-Form based estimation of the FitzHugh-Nagumo oscillator isochrons (colored bands and solid black curves transverse to cycle). We uniformly sampled points within an annulus around the limit cycle in the rectified coordinates (60006000 points, black dots). We then estimated the Temporal 1-Form with polynomial terms up to order six, and Fourier terms of order up to ten. For comparison, we chose the same parameters as fig. 4 from Langfield et al. (2014) and overlaid their isochrons (pastel lines) on ours (black lines between colored regions). The results are shown in the original coordinates (left; limit cycle in black), and in rectified coordinates (middle). The phase response curves (right) for this system as estimated by form phase (blue) and by forward numerical integration (in the sense used in Wilshin et al. (2021); red) for both xx (white line interior) and yy (black line interior) perturbations. Adapted from Langfield et al. (2014), with the permission of AIP Publishing.

One of the particular strengths of our algorithm is that it inherently estimates the phase response curves (PRC (Ermentrout, 1996; Wilshin et al., 2021)) of the system. We tested this capability with the Selkov oscillator (figure 4), and a classical neuronal oscillator model—the FitzHugh-Nagumo (FitzHugh (1961); Nagumo et al. (1962), figure 5) system.

For Selkov, we took the two-dimensional system of the previous section and estimated the isochrons from three methods of phase estimation: event-based, taking phase to be the fraction of time between events; the Phaser algorithm Revzen and Guckenheimer (2008); and the form-phase based algorithm. We focused on the case of low system noise and highest initial condition noise, which corresponds closely with deterministic dynamics whose state space has been fully sampled. We found that event-based estimates of the ground-truth isochrons are poor, with a large spread in position and poor agreement with the ground truth. Phaser performed well on the limit cycle, but the angle of the limit cycle to the estimated isochrons was incorrect, indicating an erroneous PRC estimate. The form-phase estimate matched the deterministic isochrons far more closely. We conclude that form-phase based phase estimates provided a superior estimate of the PRC.

Form-phase based estimates can also easily be used for obtaining isochrons of systems for which the equations are known, by using an importance sampling scheme. Figure 5 demonstrates this approach applied to the FitzHugh-Nagumo oscillator, and compares our results with those of Langfield et al. (2015), which is a state of the art method for systems with known equations of motion. FitzHugh-Nagumo is particularly challenging as it is a “relaxation oscillator” which produces rapidly changing spikes with gaps of inactivity between them in a neuron-like spike train. Relaxation oscillators inherently have large changes in the magnitude of dϕd\phi on the limit cycle, exacerbating the effects of numerical errors on the estimation of isochrons. The results demonstrate that we can reproduce the isochrons next to the limit cycle, and thus correctly reproduce the PRC for the FitzHugh-Nagumo oscillator when the complete state is available for measurement, or when the conditions of Wilshin et al. (2021) are met. This suggests we could reasonably expect our new algorithm to recover the PRC of neuronal oscillators when provided with sufficient data.

V.3 Phase estimation from partial data in guineafowl

Perhaps the most significant feature of our algorithm for its intended users is that our algorithm requires only state and state velocity pairs to calculate a phase estimate. It can be trained using short, disjointed training examples—even if no example contains more than a small portion of a cycle (Theorem 3 in SI sec. A.5). This is not true of any event-based phase, since the phase of all segments that do not contain the event cannot be determined at all. Similarly, Phaser cannot determine the phase of any data coming from a segment that does not cross the phase zero Poincaré section. Furthermore, because it uses the Hilbert transform to produce a protophase, Phaser requires each time series in its training data to be multiple cycles long—otherwise the Hilbert transform exhibits ringing artifacts.

We demonstrate the ability of our Temporal 1-Form algorithm to recover phase from short segments of data in figure 6. Not only does it recover phase from input that only contains trajectory segments from partial cycles, the new algorithm also produces a similar result even when some of the data is omitted systematically from a large part of space. This dataset with systematic omissions does not meet the requirements of our estimator convergence result (Theorem 3 in SI sec. A.5)—roughly speaking, it does not have a uniquely defined phase because a portion of the limit cycle is not observed—but it still seems to be usable, for reasons which remain to be understood (cf. Remark 1 in SI sec. A.5).

Refer to caption
Figure 6: Estimating phase from partial data. Results for guineafowl foot data, post-processed adversely. We randomly split Numida meleagris foot motions (292 cycles; 250sps250\text{sps}; running at 3.Hz3.\text{Hz}) into 2020 sample segments with gaps >40>40 samples in between. Segments were far shorter than cycles ([A] blue dots all data; example segment wide solid green). We used the Temporal 1-Form phase estimator (order six Fourier, and order six polynomial terms), plotted the limit cycle estimate (order 10 Fourier model; [B], solid black) and isochrons ([B] light teal, every π/20\pi/20 radians). We removed all data with x<0.5x<-0.5 (arbitrary units; data is z-scored), about one third of every cycle ([A,B,C] vertical line; omitted data faded on left), and recomputed phase (order six Fourier terms; order 6 polynomial terms), and plotted the limit cycle ([C] solid green; same method) and the same isochrons ([C] dark teal; full data isochrons faded light teal). We indicated amount of training data using a kernel smoothed density plot (50%, 70% and 90% of max density; strong orange to pale orange). Even with fractional cycles and systematic inability to observe a large fraction of the cycle, most structure is recovered in both treatments wherever >70%>70\% of max data density is available. We remark that, while accuracy of the isochron estimates in [B] is to be expected from Theorem 3 (SI sec. A.5), it remains to be understood why reasonable accuracy is obtained in [C] (cf. Remark 1 in SI sec. A.5).

V.4 Application to chemical oscillators

To illustrate how our algorithm could be applied to data from diverse sources, we obtained a sample of chemical oscillator data 555Our thanks for these data go to Seth Fraden and Michael Norton. . The raw measurements are seen in figure 7. In these relaxation oscillators the observable state is nearly constant away from their rapid “spike”, and so it is difficult to extract useful state information far from these spikes. To provide a better numerically conditioned state, after renormalizing the oscillations to account for baseline drift and reagent depletion (figure 8 top), we passed the time series through a linear filter bank. We passed each signal through three Butterworth lowpass filters of order 2, with cutoffs computed from the median inter-spike interval (ISI) as half ISI, ISI, and twice ISI. We then used the difference of the first two as one set of “lagged” signals (figure 8 middle), and the difference of the last two as a second set of “lagged” signals (figure 8 bottom). These lagged signals have the property of being fairly sinusoidal, possessing phase responses close to π/2\pi/2 apart, and having very little response at “DC” (frequency 0). In short, they appear to provide a good augmented state-space for the phase estimator to operate on (and it seems likely that this intuitive statement can be made mathematically rigorous using methods similar to those of Sauer et al. (1991, Sec. 3)). We then took the resulting phase estimates as real numbers, unwrapped them, and subtracted their time-dependent mean to obtain the relative phases (Revzen et al., 2008), showing clearly the convergence into two pairs of anti-phase oscillators (figure 9).

Refer to caption
Figure 7: Raw data of optical transmission through four coupled cells running a Belousov-Zhabotinsky reaction. Physical units are not provided as phase estimation does not require it.
Refer to caption
Figure 8: Data preparation of chemical oscillator data. [top] detrended and rescaled time series. We then computed the median inter-spike interval and filtered at 2 ISI (hereon s1), 1 ISI (s2), and 0.5 ISI (s3). [middle] the difference s1-s2; [bottom] the difference s2-s3
Refer to caption
Figure 9: Relative phases of the oscillators in figure 7, taken by training a phase estimator on the data, and subtracting the mean phase of all the oscillators.

VI Conclusion

The definition of the Temporal 1-Form we provided shows that for smooth exponentially stable oscillators the phase obtained from the Temporal 1-Form agrees with the classical definition of asymptotic phase. Unlike the classical definition, the Temporal 1-Form makes evident (SI sec. A.4) that asymptotic phase is in fact uniquely determined by a pointwise condition on forward invariant sets.

This theoretical property allowed us to develop the presented algorithm and its performance guarantees (SI sec. A.5) for approximating the Temporal 1-Form of an oscillator from which a sample of noisy trajectory segments is available. We have shown this algorithm performs well on both simulated and experimental data. On simulated data of the Fitzhugh-Nagumo oscillator, a challenging relaxation oscillator system with known equations of motion, it performs comparably to the state-of-the-art algorithms which use the equations, even though our algorithm is entirely data driven, and does not require the dynamics themselves to be integrated.

Our algorithm offers significant improvements over previously known methods for phase estimation.

  1. 1.

    Most importantly, the ability to estimate phase from state and state velocity pairs (x,x˙)(x,\dot{x}). This capability in particular would make it attractive for modeling systems for which we are unable to obtain records of full periods of oscillation.

  2. 2.

    The ability to directly obtain phase response curves from experimental time series data.

  3. 3.

    More generally, the ability to estimate phase far from the limit cycle, in any forward invariant region containing enough data (Theorem 3 in SI sec. A.5).

  4. 4.

    All these abilities come from an estimation procedure with provable convergence and convergence rates under certain assumptions (SI sec. A.5).

No other phase estimation method we are aware of possesses all of these important properties.

Because the Temporal 1-Form can be estimated from short time series segments, it offers the opportunity to study a broad range of oscillatory phenomena that were previously hard to analyze—either because of prohibitively long oscillation periods, or because of technical difficulties in obtaining multi-period measurements. We have shown that a Temporal 1-Form estimator can be successfully trained on short fragments of cycles, and will still estimate the isochrons accurately and consistently (e.g. the oscillator associated with the guineafowl feet). Event-based methods such as anterior extreme position detection (Cruse and Schwarze, 1988) or heel-strike (Ting et al., 1994; Jindrich and Full, 2002) detection cannot use such fragmented data as the decisive event need not appear in every fragment of data. The Phaser algorithm cannot solve this problem as it performs a Hilbert transform to calculate its initial proto-phases, introducing transients on the order of 1 to 3 cycles.

The Temporal 1-Form allows one to detect the phase response to perturbations of the underlying deterministic dynamics. The phase response describes only the effect of perturbation that persists after a long time, and even the infinitesimal phase response approximation—the phase response curve (PRC)—allows the observed system to be modeled in the weakly coupled oscillator approximation. This PRC is available directly from the Temporal 1-Form phase estimation procedure, paving the way to empirically generated oscillator coupling models in many domains.

For scientists studying oscillatory systems, a phase-based model allows the experimentalist to compare the outcome of experimental treatment (e.g. a mechanical perturbation to locomotion) to that of the counter-factual unperturbed system modeled as a deterministic function of phase (Revzen et al., 2008). The results in sec. V.1 indicate that for systems of different dimensions, levels of diffusion, and initial condition noise, approximating phase using a Temporal 1-Form produced by the algorithm presented here is superior to event-based phase estimates commonly used in the experimental literature, and to the Phaser algorithm. In the examples we studied, the mean-square error of the phase estimate from the form-based technique is smaller. The inferred isochrons are closer to the ground truth, and the phase response curves are more accurate.

Future work may extend our algorithm by application of domain-specific knowledge, in particular by incorporating new basis forms and coordinate change methods. Such extensions would certainly speed up the algorithm and improve its accuracy. Another future direction could include more general large-scale numerical solvers which could be applied to approximating the Temporal 1-Form without resorting to explicit basis function expansions.

The Temporal 1-Form offers a new perspective on the notion of phase in nonlinear oscillators, as well as a new way to obtain oscillator and PRC models from experimental data.

Acknowledgements

We are deeply indebted to John M. Guckenheimer for conceiving of the Temporal 1-Form and providing guidance and comments. We also thank Dan Guralnik and G. Bard Ermentrout for valuable insights. This work was supported by supported by EPSRC grant EP/H04924X/1, BBSRC grant BB/J021504/1, and U. S. Army Research Office under contract/grant number 64929EG supporting S. Wilshin, ARO W911NF-18-1-0327 and ONR N000141612817 (held by D. E. Koditschek) supporting M. D. Kvalheim, NSF Grants 1422157 and 1838179 to C. Scott, and ARO Young Investigator Award #61770 and W911NF-14-1-0573 to S. Revzen. The guineafowl kinematics were gathered under the supervision of M. Daley with the support of a Human Frontier Science Program grant (RGY0062/2010) and ethical approval was obtained from Royal Veterinary College Ethics and Welfare Committee under the protocol title, “Kinematics and kinetics in birds running over uneven terrain”. The optical transmission data for the Belousov-Zhabotinsky reaction were obtained from S. Fraden and M. Norton of Brandeis University, under awards from the National Science Foundation NSF DMREF-1534890 and U. S. Army Research Laboratory and U. S. Army Research Office W911NF-16-1-0094.

References

  • Revzen and Kvalheim (2015) S. Revzen and M. Kvalheim, in Micro-and Nanotechnology Sensors, Systems, and Applications VII, Vol. 9467 (International Society for Optics and Photonics, 2015) p. 94671V.
  • Seipel et al. (2017) J. Seipel, M. Kvalheim, S. Revzen, M. A. Sharbafi,  and A. Seyfarth, in Bioinspired Legged Locomotion (Elsevier, 2017) pp. 55–131.
  • Kvalheim and Bloch (2021) M. D. Kvalheim and A. M. Bloch, J. Differential Equations 285, 211 (2021).
  • Hoppensteadt and Izhikevich (1997) F. C. Hoppensteadt and E. M. Izhikevich, Weakly connected neural networks, Vol. 126 (Springer-Verlag, 1997).
  • May (1972) R. M. May, Science 177, 900 (1972).
  • Epstein and Pojman (1998) I. R. Epstein and J. A. Pojman, An introduction to nonlinear chemical dynamics: oscillations, waves, patterns, and chaos (Oxford University Press, 1998).
  • Głazek and Wilson (2002) S. D. Głazek and K. G. Wilson, Physical review letters 89, 230401 (2002).
  • Van der Pol (1934) B. Van der Pol, Proceedings of the Institute of Radio Engineers 22, 1051 (1934).
  • Duffing (1918) G. Duffing, Erzwungene Schwingungen bei veränderlicher Eigenfrequenz und ihre technische Bedeutung, 41-42 (R, Vieweg & Sohn, 1918).
  • Stoker (1950) J. J. Stoker, Nonlinear vibrations in mechanical and electrical systems, Vol. 2 (Interscience Publishers New York, 1950).
  • FitzHugh (1961) R. FitzHugh, Biophysical journal 1, 445 (1961).
  • Nagumo et al. (1962) J. Nagumo, S. Arimoto,  and S. Yoshizawa, Proceedings of the IRE 50, 2061 (1962).
  • Sel’Kov (1968) E. E. Sel’Kov, The FEBS Journal 4, 79 (1968).
  • Guckenheimer (1975) J. Guckenheimer, J Math Biol 1, 259 (1975).
  • Note (1) This is the exterior derivative and for Riemannian spaces when acting on a scalar is the operator \nabla with its contravariant index lowered by the metric. Since our method is not metric dependent we use 𝐝\mathrm{\mathbf{d}}{} here, however readers unfamiliar with differential geometry can treat 𝐝\mathrm{\mathbf{d}}{} as \nabla.
  • Note (2) While it need not, it could; e.g. θ˙=1+cr;r˙=(r2)(r1)r\dot{\theta}=1+cr;\dot{r}=(r-2)(r-1)r has phase identically θ\theta for c=0c=0, but phase diverges at r=2r=2 when c0c\neq 0.
  • Revzen and Guckenheimer (2008) S. Revzen and J. M. Guckenheimer, Phys Rev E 78, 051907 (2008).
  • Note (3) We use the term “basis” informally, as it is used e.g. in the notion of “radial basis functions” in machine learning, to mean a collection of functions whose finite linear spans are dense in the function space of interest.
  • Note (4) Although the closed 1-form we seek is not exact, algebraic topology teaches us that, on the basin of attraction associated to an oscillator, any two closed 1-forms α,b\alpha,b can be related as a=ωb+ca=\omega b+c for some ω\omega real and cc exact. Exact forms are exterior derivatives (“gradients”) of scalar functions. The authors are deeply indebted to Dan Guralnik for pointing out this insight on the de Rham cohomology of oscillators. That this insight extends from 𝒞\mathcal{C}^{\infty} closed forms (used in de Rham cohomology) to 𝒞0\mathcal{C}^{0} closed forms follows from general results on smoothing currents (de Rham, 1984, pp. 61-70); alternatively, it is not difficult to give a direct proof of this result for oscillators.
  • Danner et al. (2015) S. M. Danner, U. S. Hofstoetter, B. Freundl, H. Binder, W. Mayr, F. Rattay,  and K. Minassian, Brain 138, 577 (2015).
  • Ting et al. (1994) L. H. Ting, R. Blickhan,  and R. J. Full, J Exp Biol 197, 251 (1994).
  • Jindrich and Full (2002) D. L. Jindrich and R. J. Full, J Exp Biol 205, 2803 (2002).
  • Cruse and Schwarze (1988) H. Cruse and W. Schwarze, J Exp Biol 138, 455 (1988).
  • Langfield et al. (2014) P. Langfield, B. Krauskopf,  and H. M. Osinga, Chaos: An Interdisciplinary Journal of Nonlinear Science 24, 013131 (2014).
  • Wilshin et al. (2021) S. Wilshin, M. D. Kvalheim,  and S. Revzen, “Phase response curves and the role of coordinates,”  (2021), arXiv:2111.06511 [q-bio.QM] .
  • Ermentrout (1996) B. Ermentrout, Neural computation 8, 979 (1996).
  • Langfield et al. (2015) P. Langfield, B. Krauskopf,  and H. M. Osinga, SIAM Journal on Applied Dynamical Systems 14, 1418 (2015).
  • Note (5) Our thanks for these data go to Seth Fraden and Michael Norton.
  • Sauer et al. (1991) T. Sauer, J. A. Yorke,  and M. Casdagli, J. Statist. Phys. 65, 579 (1991).
  • Revzen et al. (2008) S. Revzen, D. E. Koditschek,  and R. J. Full, “Progress in motor control - a multidisciplinary perspective,”  (Springer Science+Business Media, LLC - NY, 2008) Chap. Towards testable neuromechanical control architectures for running, pp. 25–56.
  • de Rham (1984) G. de Rham, Differentiable manifolds (Springer-Verlag, 1984).
  • Eldering et al. (2018) J. Eldering, M. Kvalheim,  and S. Revzen, Nonlinearity 31, 4202 (2018).
  • Lee (2013) J. M. Lee, Introduction to Smooth Manifolds, 2nd ed. (Springer, 2013).
  • Guillemin and Pollack (2010) V. Guillemin and A. Pollack, Differential topology (AMS Chelsea Publishing, Providence, RI, 2010) pp. xviii+224, reprint of the 1974 original.
  • Hirsch (1976) M. W. Hirsch, Differential topology (Springer-Verlag, 1976).
  • Spivak (1971) M. Spivak, Calculus on Manifolds (Perseus Books Publishing, 1971).
  • Arnold (1974) L. Arnold, Stochastic differential equations: theory and applications, 1st ed. (John Wiley and Sons, 1974).
  • Gardiner (2004) C. W. Gardiner, Handbook of Stochastic Methods, 3rd ed. (Springer-Verlag, 2004).
  • Evans (2013) L. C. Evans, An Introduction to Stochastic Differential Equations, 1st ed. (American Mathematical Society, 2013).
  • Øksendal (2013) B. Øksendal, Stochastic differential equations, 6th ed. (Springer-Verlag, 2013).
  • Hénon (1976) M. Hénon, Communications in Mathematical Physics 50, 69 (1976).
  • Bastani and Hosseini (2007) A. F. Bastani and S. M. Hosseini, J Comp Appl Math 206, 631 (2007).
  • Burrage (1999) P. M. Burrage, Runge-Kutta methods for stochastic differential equations, Ph.D. thesis, U Queensland, Australia (1999).
  • Schwabedal and Pikovsky (2013) J. T. C. Schwabedal and A. Pikovsky, Physical review letters 110, 204102 (2013).
  • Thomas and Lindner (2014) P. J. Thomas and B. Lindner, Physical review letters 113, 254101 (2014).
  • Cao et al. (2020) A. Cao, B. Lindner,  and P. J. Thomas, SIAM J. Appl. Math. 80, 422 (2020).
  • Engel and Kuehn (2021) M. Engel and C. Kuehn, Comm. Math. Phys. 386, 1603 (2021).
  • Briers et al. (2009) M. Briers, A. Doucet,  and S. Maskell, Annals of the Institute of Statistical Mathematics 62, 61 (2009).
  • Hirsch et al. (1977) M. W. Hirsch, C. C. Pugh,  and M. Shub, Invariant manifolds, Lecture Notes in Mathematics (Springer-Verlag, Berlin-New York, 1977) pp. ii+149.
  • Winfree (1980) A. T. Winfree, The Geometry of Biological Time (Springer-Verlag, New York, 1980).
  • Kostelich and Yorke (1990) E. J. Kostelich and J. A. Yorke, Phys D: Nonlinear Phenom 41, 183 (1990).
  • Takens (1981) F. Takens, in Dynamical systems and turbulence, Warwick 1980 (Coventry, 1979/1980), Lecture Notes in Math., Vol. 898 (Springer, Berlin-New York, 1981) pp. 366–381.
  • Schelter et al. (2006) B. Schelter, M. Winterhalder,  and J. Timmer, eds., Handbook of time series analysis: recent theoretical developments and applications (Wiley, 2006) iSBN: 9783527406234.
  • Gibson et al. (1992) J. F. Gibson, J. D. Farmer, M. Casdagli,  and S. Eubank, Physica D: Nonlinear Phenomena 57, 1 (1992).
  • Huguet and de-la Llave (2012) G. Huguet and R. de-la Llave, Computation of limit cycles and their isochrones: fast algorithms and their convergence, Tech. Rep. 2415 (Institute for mathematics and its applications, University of Minnesota, 2012).
  • Osinga and Moehlis (2010) H. M. Osinga and J. Moehlis, SIAM Journal on Applied Dynamical Systems 9, 1201 (2010).
  • Revzen and Guckenheimer (2012) S. Revzen and J. M. Guckenheimer, J R Soc Lond Interface 9, 957 (2012).
  • Peters and Schaal (2008) J. Peters and S. Schaal, Neurocomputing 71, 1180 (2008).
  • Schaal and Atkeson (2010) S. Schaal and C. Atkeson, Robotics Automation Magazine, IEEE 17, 20 (2010).
  • Schaal et al. (2005) S. Schaal, J. Peters, J. Nakanishi,  and A. Ijspeert, in Robotics Research (Springer, 2005) pp. 561–572.
  • Pikovsky (2015) A. Pikovsky, Physical review letters 115, 069401 (2015).
  • Thomas and Lindner (2015) P. J. Thomas and B. Lindner, Physical review letters 115, 069402 (2015).
  • Mauroy and Mezić (2012) A. Mauroy and I. Mezić, Chaos: An Interdisciplinary Journal of Nonlinear Science 22, 033112 (2012).
  • Kvalheim and Revzen (2021) M. D. Kvalheim and S. Revzen, Physica D 425, Paper No. 132959, 20 (2021).
  • Kvalheim et al. (2021a) M. D. Kvalheim, D. Hong,  and S. Revzen, IFAC-PapersOnLine 54, 267 (2021a).
  • Mauroy et al. (2013) A. Mauroy, I. Mezić,  and J. Moehlis, Physica D: Nonlinear Phenomena 261, 19 (2013).
  • Brunton et al. (2021) S. L. Brunton, M. Budišić, E. Kaiser,  and J. N. Kutz, arXiv preprint arXiv:2102.12086  (2021).
  • Kutz et al. (2016) J. N. Kutz, S. L. Brunton, B. W. Brunton,  and J. L. Proctor, Dynamic mode decomposition: data-driven modeling of complex systems (SIAM, 2016).
  • Wu et al. (2021) Z. Wu, S. L. Brunton,  and S. Revzen, Journal of the Royal Society Interface 18, 20210686 (2021).
  • Revzen et al. (2011) S. Revzen, J. M. Guckenheimer,  and R. J. Full, in Yearly meeting of the Society for Integrative and Comparative Biology (Society for Integrative and Comparative Biology, 2011).
  • Revzen (2009) S. Revzen, Neuromechanical Control Architectures in Arthropod Locomotion, Ph.D. thesis, University of California, Berkeley (2009), department of Integrative Biology.
  • Tytell (2013) E. D. Tytell, in Soc Integ Compar Biol (2013) p. 40.1.
  • Wang and Srinivasan (2012) Y. Wang and M. Srinivasan, in ASME 5th Annual Dynamic Systems and Control Conference, Vol. 2 (ASME, 2012) pp. 19–23.
  • Ankarali and Cowan (2014) M. M. Ankarali and N. J. Cowan, in 53rd IEEE Conference on Decision and Control (IEEE, 2014) pp. 1017–1022.
  • Back et al. (1993) A. Back, J. Guckenheimer,  and M. Myers, Hybrid Systems , 255–267 (1993).
  • Simić et al. (2000) S. N. Simić, K. H. Johansson, S. Sastry,  and J. Lygeros, in International Workshop on Hybrid Systems: Computation and Control (Springer, 2000) pp. 421–436.
  • Westervelt et al. (2003) E. R. Westervelt, J. W. Grizzle,  and D. E. Koditschek, IEEE transactions on automatic control 48, 42 (2003).
  • Ames and Sastry (2005) A. D. Ames and S. Sastry, in International Workshop on Hybrid Systems: Computation and Control (Springer, 2005) pp. 86–102.
  • Goebel et al. (2009) R. Goebel, R. G. Sanfelice,  and A. Teel, Control Systems, IEEE 29, 28–93 (2009).
  • Lerman (2016) E. Lerman, arXiv preprint arXiv:1612.01950  (2016).
  • Lerman and Schmidt (2020) E. Lerman and J. Schmidt, J. Geom. Phys. 149, 103582, 30 (2020).
  • Clark and Bloch (2021) W. Clark and A. Bloch, arXiv preprint arXiv:2101.11128  (2021).
  • Kvalheim et al. (2021b) M. D. Kvalheim, P. Gustafson,  and D. E. Koditschek, SIAM J. Appl. Dyn. Syst. 20, 784 (2021b).
  • Note (6) In more detail: fix x𝐗x\in\mathbf{X}, let UΓU\subsetneq\Gamma be a contractible open neighborhood of P(x)\mathrm{P}(x), and define τx[0,)\tau_{x}\in[0,\infty) to be the smallest nonnegative number such that PΦτx(x)=o(0)\mathrm{P}\circ\Phi_{-\tau_{x}}({x})=o(0). Since the isochron P1(o(0))\mathrm{P}^{-1}(o(0)) is a 𝒞k1\mathcal{C}^{k\geq 1} manifold transverse to the vector field Guckenheimer (1975); Eldering et al. (2018), the implicit function theorem implies that there is a 𝒞k\mathcal{C}^{k} function τ\nonscript:P1(U)\tau\nobreak\mskip 2.0mu\mathpunct{}\nonscript\mkern-3.0mu{:}\mskip 6.0mu plus 1.0mu\mathrm{P}^{-1}(U)\to\mathbb{R} satisfying τ(x)=τx\tau(x)=\tau_{x} and PΦτ(y)(y)=o(0)\mathrm{P}\circ\Phi_{-\tau(y)}({y})=o(0) for all yP1(U)y\in\mathrm{P}^{-1}(U). Thus, P|P1(U)=oτ\mathrm{P}|_{\mathrm{P}^{-1}(U)}=o\circ\tau is 𝒞k\mathcal{C}^{k}. Since x𝐗x\in\mathbf{X} was arbitrary, it follows that P𝒞k\mathrm{P}\in\mathcal{C}^{k}.
  • Note (7) The geometry of isochrons can be surprisingly complicated, especially when the vector field ff has multiple time scales (Langfield et al., 2015).
  • Note (8) In a Euclidean space, α\alpha can be expressed as α:=(f𝖳f)1ωf𝖳\alpha:=(f^{\mathsf{T}}f)^{-1}\omega f^{\mathsf{T}}.
  • Farber et al. (2004) M. Farber, T. Kappeler, J. Latschev,  and E. Zehnder, Ergodic theory and dynamical systems 24, 1451 (2004).
  • Note (9) In fact, a strong deformation retract. As an embedded submanifold of UU, the limit cycle is a strong deformation retract of a “tubular neighborhood” (Lee (2013) Chapter 6) containing it. Using the flow, one may construct a strong deformation retract of UU onto a tubular neighborhood. Following this homotopy with the strong deformation retraction of the tubular neighborhood onto the limit cycle then gives the desired strong deformation retraction.
  • Pajitnov (2006) A. V. Pajitnov, Circle-valued Morse theory, De Gruyter Studies in Mathematics, Vol. 32 (Walter de Gruyter & Co., Berlin, 2006) pp. x+454.
  • Kvalheim (2018) M. Kvalheim, Aspects of invariant manifold theory and applications, Ph.D. thesis, University of Michigan (2018), department of Electrical Engineering and Computer Science.
  • Doob (1990) J. L. Doob, Stochastic processes, Wiley Classics Library (John Wiley & Sons, Inc., New York, 1990) pp. viii+654, reprint of the 1953 original, A Wiley-Interscience Publication.
  • Andrews (1987) D. W. K. Andrews, Econometrica 55, 1465 (1987).
  • Pötscher and Prucha (1989) B. M. Pötscher and I. R. Prucha, Econometrica 57, 675 (1989).
  • Note (10) Note that, by the Arzelà-Ascoli theorem, the compact subsets of Ω01(K)\Omega^{1}_{0}(K) are precisely the closed and bounded subsets of 11-forms which are equicontinuous.
  • Aliprantis and Border (2006) C. D. Aliprantis and K. C. Border, Infinite dimensional analysis, 3rd ed. (Springer, Berlin, 2006) pp. xxii+703, a hitchhiker’s guide.

Supplemental Information (SI)

Appendix A Algorithm notes

The Supplemental Information is written with the mathematically inclined reader in mind. As such, we presume familiarity with fundamental concepts of differential topology (Guillemin and Pollack, 2010; Hirsch, 1976; Lee, 2013), differential forms (Spivak, 1971; Guillemin and Pollack, 2010; Lee, 2013), and stochastic differential equations Arnold (1974); Gardiner (2004); Evans (2013); Øksendal (2013).

A.1 Generating ground truth test data—details

This section provides details of the method we used to generate randomized test systems with adjustible complexity and noise characteristics. We provide these details in the hope that other investigators may use them to evaluate the performance of algorithms which aim to analyze the structure of nonlinear oscillators.

Very few oscillators admit a closed-form representation of their phase as a function of state, making validation of our algorithm a significant challenge all on its own. To produce a well specified class of tests for our algorithm, we produced sample paths arising from a Stratonovich stochastic differential equation of the form

dx(t)=f(x(t))dt+G(x(t))dW(t),\displaystyle\mathrm{d}{x}(t)=f(x(t))\mathrm{d}{t}+G(x(t))\circ\mathrm{d}{W}(t), (12)

and requiring that the deterministic dynamics of dx(t)=f(x(t))dt\mathrm{d}{x}(t)=f(x(t))\mathrm{d}{t} are known to be a (hyperbolic) limit cycle oscillator, with a known phase coordinate.

We obtained such a deterministic oscillator by starting with a randomly chosen linearized (Floquet) structure converging on the trajectory x(t)=[t,0,]x(t)=[t,0,\ldots], then winding this affine system around the unit circle with x1x_{1} as the angle, and finally pushing forward the dynamics through a composition of randomly generated diffeomorphisms to generate f()f(\cdot) of eqn. (12). We chose diffeomorphisms such that their tangent maps were computable and invertible in closed form. We used a second composition of random diffeomorphisms to generate a function g()g(\cdot), and the inverse of the tangent map of g()g(\cdot) was used for G()G(\cdot). By using the inverse diffeomorphisms, we directly computed the phase of each trajectory point with respect to the deterministic dynamics, to use as a baseline for comparing with other forms of phase estimation.

A.1.1 Implementation of the Floquet Structure

While our exposition so far focused on differential forms, allowing us to emphasize the coordinate-invariant properties of the Temporal 1-Form, we provide the equations in this section in a computationally friendly coordinate dependent matrix notation.

Given an nn-dimensional state space (n>1n>1), we defined a vector field by first taking the polar decomposition Ω:nS1×+×n2\Omega:\,\mathbb{R}^{n}\rightarrow S^{1}\times\mathbb{R}^{+}\times\mathbb{R}^{n-2} of the first two coordinates: p:=[x3,x4,]p:=[x_{3},x_{4},\ldots], θ:=atan2(x1,x2)\theta:=\mathrm{atan2}(x_{1},x_{2}), r:=x12+x22r:=\sqrt{x_{1}^{2}+x_{2}^{2}} and thus Ω(x)=[θ,r]p\Omega(x)=[\theta,r]\oplus p.

In these polar coordinates, the equation of motion we implemented is affine:

θ˙\displaystyle\dot{\theta} =1\displaystyle=1
r˙\displaystyle\dot{r} =β(r1)+cr𝖳p\displaystyle=\beta(r-1)+c_{r}^{\mathsf{T}}p (13)
p˙\displaystyle\dot{p} =γ(r1)+Mp\displaystyle=\gamma(r-1)+Mp

This means that θ˙\dot{\theta} is, for a deterministic system, uniformly advancing in time, and the latter two equations comprise an autonomous linear subsystem which evolves independently of θ\theta. These equations can be combined to form a single equation in homogeneous coordinates:

[θ˙,r˙,p˙,0]𝖳=M~[θ,r,p,1]𝖳\displaystyle[\dot{\theta},\dot{r},\dot{p},0]^{\mathsf{T}}=\tilde{M}[\theta,r,p,1]^{\mathsf{T}} (14)

where the matrix M~\tilde{M} can be chosen to obtain any eigenvalue structure desired for the dynamics.

A.1.2 A natural class of invertible diffeomorphisms

To produce randomized invertible diffeomorphisms, we used a structure inspired by the Hénon map (Hénon, 1976) (and therefore refer to these as H-maps) and recommended to us by J.M. Guckenheimer. Given: A (split) vector space 𝐐:=𝐗𝐘\mathbf{Q}:=\mathbf{X}\oplus\mathbf{Y}; Invertible maps gX:𝐗𝐗g_{X}:\,\mathbf{X}\rightarrow\mathbf{X} and gY:𝐘𝐘g_{Y}:\,\mathbf{Y}\rightarrow\mathbf{Y}; and a (possibly non-invertible) map f:𝐘𝐗f:\,\mathbf{Y}\rightarrow\mathbf{X}, we constructed the following mapping:

x~:=gX(x)+(fgY)(y)\displaystyle\tilde{x}:=g_{X}(x)+(f\circ g_{Y})(y) y~:=gY(y),\displaystyle\tilde{y}:=g_{Y}(y), (15)

which admits the inverse (by construction):

x:=gX1(x~f(y~))\displaystyle x:=g^{-1}_{X}\left(\tilde{x}-f(\tilde{y})\right) y:=gY1(y~).\displaystyle y:=g^{-1}_{Y}(\tilde{y}). (16)

In our implementation we have chosen to make gXg_{X} and gYg_{Y} affine maps. For our nonlinear map f()f(\cdot), we used inversions of the form

f(x)\displaystyle f(x) :=m(β+2)ϱ(x)P(x)(ϱ(x)2+βϱ(x)+1)\displaystyle:=m\left({\beta+2}\right)\frac{\varrho\left({x}\right)P\left({x}\right)}{\left({\varrho\left({x}\right)^{2}+\beta\varrho\left({x}\right)+1}\right)} (17)
ϱ(x)\displaystyle\varrho(x) :=(x𝖳Ax),\displaystyle:=\left({x^{\mathsf{T}}Ax}\right), (18)

for some positive symmetric AA with eigenvalues close to one (uniformly randomly sampled between 0.950.95 and 1.051.05 for our simulations). ff is a function which is approximately zero when ϱ(x)\varrho\left({x}\right) is small and large, and is stationary with value mm at ϱ(x)=1\varrho\left({x}\right)=1. It has one stationary point for positive values of ϱ(x)\varrho\left({x}\right) with β\beta controlling how “peaked” the function is (the more negative beta is, the greater the “peaking”). PP maps from 𝐗\mathbf{X} of dimension DXD_{X} to 𝐘\mathbf{Y} of dimension DYD_{Y} and the components of PP are given by.

Pij={1if ij[modDX]0if ij[modDX]P_{ij}=\left\{\begin{array}[]{ll}1&\mbox{if $i\equiv j\,[{\mathrm{mod}~{}}D_{X}]$}\\ 0&\mbox{if $i\not\equiv j\,[{\mathrm{mod}~{}}D_{X}]$}\end{array}\right. (19)

with 1iDY1\leq i\leq D_{Y} and 1jDX1\leq j\leq D_{X}.

We composed together a sequence of these maps to produce a highly nonlinear, but invertible, distortion of the Floquet system after it has been transformed to polar coordinates. Between each pair of consecutive H-maps, we inserted a randomly chosen full-rank affine transformation.

A.1.3 Stochastic integration of the combined system

We implemented individual diffeomorphisms as objects of a Python class using the SciPy scientific programming environment. Each mapping object was capable of forward and inverse mappings, pushforwards and pullbacks. We composed these mapping objects with each other to construct the fully formed random diffeomorphisms for both the deterministic and noise terms of eqn. (12). The noise term uses the Jacobian of an H-map chain to transform the Wiener process when integrating the SDE. The final integrated Floquet system is then transformed by a different H-map chain. We integrated the SDE of eqn. (12) using an implementation of the R3 stochastic integration scheme (Bastani and Hosseini, 2007; Burrage, 1999), and have made this integrator available publicly at http://github/BIRDSlab/BIRDSode. We also note that while the “ground truth” isochrons of the deterministic system are known precisely, they may not be compatible with various “ground truth” definitions of stochastic isochrons (Schwabedal and Pikovsky, 2013; Thomas and Lindner, 2014; Cao et al., 2020; Engel and Kuehn, 2021) for the SDE. For example, the presence of noise can change the average frequency of a limit cycle oscillator, and by extension, may deform the isochrons. We expect that while at the low noise limit our recovered phase will closely match the deterministic system (this expectation is mathematically justified in sec. A.5.2), as noise is increased the discrepancy between the two may justifiably increase. We therefore caution the reader that in the presence of consistent systematic errors and higher noise conditions, errors in estimates of asymptotic phase may represent an error in the “ground truth” rather than an estimation error on the part of the algorithm.

A.1.4 Preparation of data for phase estimation

In each simulation, we treated these DD dimensional data similarly to how experimental data would be treated. We filtered the data using a Kalman smoother Briers et al. (2009) with the system states the position and its derivative. The state transition matrix assumed no acceleration and that the last DD coordinates were derivatives of the first DD coordinates. The observation matrix was a D×2DD\times 2D matrix with an identity in the first D×DD\times D sub-matrix. We assumed the system noise matrix to be diagonal with the first DD system variables having the same level of variance, and the last DD variables having a (typically) different shared level variance. We estimated these two system covariance parameters by maximizing the likelihood of the training observations.

After filtering, we performed a principal component analysis and rotated the system into the corresponding orthogonal coordinate system. We then z-scored the first two principal components.

A.2 Review of differential forms

Differential forms Ω\Omega over 𝐗\mathbf{X} are an exterior algebra generated using two operators: the “exterior product” \wedge—the universal skew symmetric product, and the “exterior derivative” 𝐝{\mathbf{d}}. Elements of rank 0 correspond to (sufficiently smooth) scalar valued functions 𝐗\mathbf{X}\to\mathbb{C}. When αΩr\alpha\in\Omega_{r} and βΩs\beta\in\Omega_{s}, 𝐝αΩr+1{\mathbf{d}}\alpha\in\Omega_{r+1} and αβΩr+s\alpha\wedge\beta\in\Omega_{r+s}.

These operations are familiar to many in their 3-dimensional special cases, where ranks Ω0Ω3\Omega_{0}\ldots\Omega_{3} correspond to scalar functions, vectors, directed areas, and directed volumes. Here the exterior derivative 𝐝{\mathbf{d}} functions as the gradient, curl, and divergence. The exterior product \wedge functions as a scalar-vector product when one argument is rank 0, as a cross product (resulting in a directed area) when the arguments are rank 1, and as a box-product when applied to rank 2 and rank 1 arguments.

Formally defined, 𝒞k\mathcal{C}^{k} differential 1-forms α\alpha on 𝐗\mathbf{X} are 𝒞k\mathcal{C}^{k} sections of its cotangent bundle 𝖳𝐗\mathsf{T}^{*}\mathbf{X}: α(x)𝖳x𝐗\alpha(x)\in\mathsf{T}^{*}_{x}\mathbf{X}. If vv is a 𝒞k\mathcal{C}^{k} vector field, then α,v\langle\alpha,v\rangle is a 𝒞k\mathcal{C}^{k} scalar function. If α\alpha is a 𝒞1\mathcal{C}^{1} curve in 𝐗\mathbf{X}, the line integral αα\int_{\alpha}\alpha is well defined and independent of coordinate system.

A 1-form α\alpha is “closed” if its exterior derivative vanishes. It is “exact” if it is the exterior derivative of a scalar function. The identity 𝐝𝐝=0{\mathbf{d}}\circ{\mathbf{d}}=0 implies that all exact forms are closed. On contractible regions of 𝐗\mathbf{X} the converse is true, i.e. all closed forms are exact (a.k.a. the Poincaré Lemma). The quotient space of closed to exact kk-forms is known as the kk-th de-Rham cohomology group. We use some properties of this cohomology to construct our algorithm.

If a closed curve γ1\gamma_{1} can be smoothly deformed into a curve γ2\gamma_{2}, γ1α\int_{\gamma_{1}}\alpha of a closed 1-form α\alpha equals γ2α\int_{\gamma_{2}}\alpha. In other words, the integral of a closed 1-form depends only on the homotopy class of the underlying closed curve.

A more complete introduction to the theory of differential forms can be found in Spivak (1971); Guillemin and Pollack (2010); Lee (2013).

A.3 Temporal 1-Form

This self-contained section serves as an alternative introduction to our paper which contains more details and is aimed at mathematically inclined readers. It includes a description of the Temporal 1-Form assuming some familiarity with differential forms, which are reviewed in sec. A.2.

We consider continuous-time dynamical systems, or flows Φ\Phi, on a smooth nn-dimensional manifold 𝐗\mathbf{X}. That Φ\Phi is a flow means that Φ:𝐗×𝐗\Phi:\mathbf{X}\times\mathbb{R}\to\mathbf{X}, Φ0=id𝐗\Phi_{0}=\textnormal{id}_{\mathbf{X}}, and ΦtΦs=Φt+s\Phi_{t}\circ\Phi_{s}=\Phi_{t+s}. We assume that the vector field f:=tΦt|t=0f:=\frac{\partial}{\partial t}\Phi_{t}|_{t=0} is well-defined and 𝒞k1\mathcal{C}^{k\geq 1}, so that Φ𝒞k\Phi\in\mathcal{C}^{k}. The trajectories tΦt(x0)x(t)t\mapsto\Phi_{t}({x_{0}})\eqqcolon x(t) satisfy the ordinary differential equation (ODE) x˙=f(x)\dot{x}=f(x). A periodic orbit oo of period T>0T>0 is a trajectory that satisfies o(t+T)=o(t)o(t+T)=o(t) with o(t)o(0)o(t)\neq o(0) for all 0<t<T0<t<T. We will always assume that T>0T>0 is the minimal period of a nonstationary periodic orbit. We focus on dynamical systems possessing only one nonstationary periodic trajectory oo and denote by Γ𝐗\Gamma\subset\mathbf{X} the image of this oo.

We further assume that this periodic trajectory is exponentially stable, a property that holds in many practical cases. Exponentially stable limit cycles are always “normally hyperbolic” (Hirsch et al., 1977; Eldering et al., 2018); for brevity we will simply refer to these as “limit cycles”. We refer to the stability basin of a limit cycle and the dynamics within it as an “oscillator”. The asymptotic behavior of any oscillator is described fully by its “asymptotic phase”: the stability basin of an oscillator is partitioned into codimension-11 𝒞k\mathcal{C}^{k} embedded submanifolds (Eldering et al., 2018, pp. 4208–4209) called “isochrons” which are asymptotically in phase with one another Guckenheimer (1975); Winfree (1980).

For systems in which the dimension is low, the noise is small or the data is plentiful, data pairs (xk,x˙k),k=1N(x_{k},{\dot{x}}_{k}),~{}k=1\ldots N may be used to estimate the vector field ff. Numerous investigators proposed such approaches, primarily in the 1990s. Classical papers include Kostelich and Yorke Kostelich and Yorke (1990), who fit arbitrary continuous dynamics to trajectories, and the rich literature on time-delay embeddings, derivatives and principal components Takens (1981); Sauer et al. (1991); Schelter et al. (2006); Gibson et al. (1992). When dynamical equations are known precisely, automatic differentiation and continuation methods can be used to compute the isochrons, a problem equivalent to that of phase estimation Huguet and de-la Llave (2012); Osinga and Moehlis (2010). However, care must be taken if it is desired to extract phase response curves from this, as these quantities are coordinate dependent (Wilshin et al., 2021).

With fewer data points and higher levels of noise, one expects that the full off-cycle dynamics are less amenable to accurate estimation, and that only a few of the slower-decaying “modes” might be observable off of the limit cycle Revzen and Guckenheimer (2012). At the extreme, only phase itself might be detectable, through methods such as we proposed in Revzen & Guckenheimer Revzen and Guckenheimer (2008) using various phase-locking approaches Peters and Schaal (2008); Schaal and Atkeson (2010); Schaal et al. (2005). This motivates the main goal of the present paper: estimate asymptotic phase from empirically observed dynamics, e.g., from an ensemble of noisy measurements of (possibly short and noisy) system trajectory segments.

We note that, while there has been work to define generalized notions of asymptotic phase for stochastic oscillators (Schwabedal and Pikovsky, 2013; Thomas and Lindner, 2014; Cao et al., 2020; Engel and Kuehn, 2021) (see Pikovsky (2015); Thomas and Lindner (2015) for a spirited discussion), in this paper we restrict ourselves to estimation of the classical asymptotic phase of a deterministic oscillator using data from a perturbed version of the underlying deterministic system. We also note that there are operator-theoretic methods of recent interest revealing phase as the generator of a family of eigenfunctions of the Koopman operator with purely imaginary eigenvalues Mauroy and Mezić (2012); Kvalheim and Revzen (2021); Kvalheim et al. (2021a). These eigenfunctions and others can be estimated using Fourier/Laplace averages Mauroy and Mezić (2012); Mauroy et al. (2013); Kvalheim and Revzen (2021); Kvalheim et al. (2021a); Brunton et al. (2021) when dynamical equations are known, and from data using Dynamic Mode Decomposition Kutz et al. (2016), at least in certain cases. However, the results of these emerging spectral methods seem to be sensitive to the choice of observables and their nonlinearities (Wu et al., 2021).

Several investigators have pursued the construction of linearized (Floquet) models for dynamics near the limit cycle. We proposed an approach termed “Data Driven Floquet Analysis (DDFA)” Revzen et al. (2011); Revzen (2009) consisting of estimating phase, then constructing affine models conditioned on phase. Tytell Tytell (2013) reported a DDFA approach based on harmonic balance. Wang and Srinivasan Wang and Srinivasan (2012) proposed to construct a Floquet model using “factored Poincaré maps” and derive a phase estimate from this model. Ankarali et. al Ankarali and Cowan (2014) applied “harmonic transfer functions” and achieved good results on a “hybrid” Back et al. (1993); Simić et al. (2000); Westervelt et al. (2003); Ames and Sastry (2005); Goebel et al. (2009); Lerman (2016); Lerman and Schmidt (2020); Clark and Bloch (2021); Kvalheim et al. (2021b) spring-mass hopper model. Regardless of how they are obtained, linear models of the dynamics around the limit cycle can at best only represent the hyperplanes tangent to the isochrons at their intersection with the limit cycle. By contrast, the method we propose here can construct nonlinear isochron approximations on suitable neighborhoods of the limit cycle, as we explain in sec. A.5.

We now begin our theoretical description of the object we call the “Temporal 1-Form”. By redefining 𝐗\mathbf{X} to be the basin of attraction of Γ\Gamma, we henceforth assume that Γ\Gamma is globally asymptotically stable. Asymptotic phase can be viewed as a map P:𝐗Γ\mathrm{P}:\mathbf{X}\to\Gamma which is a retraction (P|Γ=id|Γ\mathrm{P}|_{\Gamma}=\textnormal{id}|_{\Gamma}) and a semiconjugacy (t:PΦt=ΦtP\forall t\in\mathbb{R}:\mathrm{P}\circ\Phi_{t}=\Phi_{t}\circ\mathrm{P}). From this it follows that P𝒞k\mathrm{P}\in\mathcal{C}^{k} when f𝒞kf\in\mathcal{C}^{k}666 In more detail: fix x𝐗x\in\mathbf{X}, let UΓU\subsetneq\Gamma be a contractible open neighborhood of P(x)\mathrm{P}(x), and define τx[0,)\tau_{x}\in[0,\infty) to be the smallest nonnegative number such that PΦτx(x)=o(0)\mathrm{P}\circ\Phi_{-\tau_{x}}({x})=o(0). Since the isochron P1(o(0))\mathrm{P}^{-1}(o(0)) is a 𝒞k1\mathcal{C}^{k\geq 1} manifold transverse to the vector field Guckenheimer (1975); Eldering et al. (2018), the implicit function theorem implies that there is a 𝒞k\mathcal{C}^{k} function τ:P1(U)\tau\colon\mathrm{P}^{-1}(U)\to\mathbb{R} satisfying τ(x)=τx\tau(x)=\tau_{x} and PΦτ(y)(y)=o(0)\mathrm{P}\circ\Phi_{-\tau(y)}({y})=o(0) for all yP1(U)y\in\mathrm{P}^{-1}(U). Thus, P|P1(U)=oτ\mathrm{P}|_{\mathrm{P}^{-1}(U)}=o\circ\tau is 𝒞k\mathcal{C}^{k}. Since x𝐗x\in\mathbf{X} was arbitrary, it follows that P𝒞k\mathrm{P}\in\mathcal{C}^{k}. . Each pΓp\in\Gamma represents the “isochron” P1(p)\mathrm{P}^{-1}(p), which is the set of initial conditions from which trajectories converge with the one starting at pp777 The geometry of isochrons can be surprisingly complicated, especially when the vector field ff has multiple time scales (Langfield et al., 2015) . In the present context of oscillators, asymptotic phase is more commonly represented as a circle-valued (“phasor”-valued) map φ:𝐗S1\varphi\colon\mathbf{X}\to\mathrm{S}^{1}; we explain the relationships between different representations of phase in sec. A.4.

The Temporal 1-Form appears naturally as a consequence of the existence of P\mathrm{P}. Noting that ff is nowhere zero on the 11-dimensional 𝒞k\mathcal{C}^{k} manifold Γ\Gamma and defining the angular frequency ω2π/T\omega\coloneqq 2\pi/T, it follows that there exists a unique 𝒞k1\mathcal{C}^{k-1} differential 1-form α\alpha on Γ\Gamma satisfying α,f=ω\langle\alpha,f\rangle=\omega identically on Γ\Gamma888In a Euclidean space, α\alpha can be expressed as α:=(f𝖳f)1ωf𝖳\alpha:=(f^{\mathsf{T}}f)^{-1}\omega f^{\mathsf{T}}. The “Temporal 1-Form ” dϕ:𝐗𝖳𝐗d\phi:\,\mathbf{X}\to\mathsf{T}^{*}\mathbf{X} is defined to be the pullback Pα\mathrm{P}^{*}\alpha of α\alpha, which means that dϕ(x),v:=α(P(x)),𝖣P(x)v\langle d\phi(x),v\rangle:=\langle\alpha(\mathrm{P}(x)),\mathsf{D}\mathrm{P}(x)\cdot v\rangle for any v𝖳x𝐗v\in\mathsf{T}_{x}\mathbf{X}. Since P𝒞k\mathrm{P}\in\mathcal{C}^{k}, it follows that dϕ𝒞k1d\phi\in\mathcal{C}^{k-1}. It is easy to see that the Temporal 1-Form satisfies two properties (further discussed in sec. A.4): (1) dϕ,f=ω\langle d\phi,f\rangle=\omega everywhere, and (2) dϕd\phi is closed.

By “closed” we mean that each x𝐗x\in\mathbf{X} possesses an open neighborhood UxU_{x} on which dϕd\phi is the exterior derivative of a 𝒞k\mathcal{C}^{k} function. This definition permits us to discuss continuous closed forms, needed for the case f𝒞1f\in\mathcal{C}^{1}; it reduces to the usual definition for 𝒞1\mathcal{C}^{1} forms (cf. Farber et al. (2004)).

Somewhat less obvious is the fact that the Temporal 1-Form is the unique continuous closed 11-form on 𝐗\mathbf{X} satisfying the preceding two properties; we show that this is the case in sec. A.4. In that section we also relate P\mathrm{P}, dϕd\phi, and circle-valued phases φ:𝐗S1\varphi\colon\mathbf{X}\to\mathrm{S}^{1}. By a “circle-valued (asymptotic) phase” φ:𝐗S1\varphi\colon\mathbf{X}\to\mathrm{S}^{1}\subset\mathbb{C} we mean a continuous map satisfying φΦt=eiωtφ\varphi\circ\Phi_{t}=e^{i\omega t}\varphi for all tt\in\mathbb{R}. To summarize Theorem 1 of sec. A.4, some of these relationships are as follows: φP=φ\varphi\circ\mathrm{P}=\varphi, φ\varphi is unique modulo rotations of S1\mathrm{S}^{1}, and dϕ=φ(dθ)d\phi=\varphi^{*}(d\theta) is the pullback of the standard angular form on S1\mathrm{S}^{1} via any circle-valued phase. Moreover, any choice of basepoint x0𝐗x_{0}\in\mathbf{X} uniquely determines a circle-valued phase by integrating dϕd\phi along continuous curves from x0x_{0}.

While we are interested in asymptotic phase as defined for deterministic oscillators, our primary goal is to compute asymptotic phase from real-world data which we assume may be subject to system and measurement noise. In sec. A.5 we derive fairly general performance guarantees applicable to our algorithm and potentially others. In that section, we define “unobserved” and “observed” empirical cost functions JNJ_{N} and JN^\widehat{J_{N}} mapping 11-forms to nonnegative numbers, where NN is the number of observed state-velocity data pairs. Intuitively, we would like to minimize the former cost (for which JN(dϕ)=0J_{N}(d\phi)=0), but only the latter cost is observable, so our algorithms attempts to minimize the latter.

Assuming that our algorithm outputs estimates dϕ^N\widehat{d\phi}_{N} satisfying dϕ^Ndϕ+ϵ\widehat{d\phi}_{N}\leq d\phi+\epsilon for some ϵ0\epsilon\geq 0, that 𝐗\mathbf{X} is an open subset of n\mathbb{R}^{n}, and that state measurements belong to some fixed compact set K𝐗K\subset\mathbf{X}, we establish in Theorem 2 the following inequality holding with probability 11:

lim supN\displaystyle\limsup_{N\to\infty} JN(dϕ^N)2ϵdϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K)a.s.2κ.\displaystyle\frac{\sqrt{J_{N}(\widehat{d\phi}_{N})}-2\sqrt{\epsilon}}{\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}}\stackrel{{\scriptstyle\textnormal{a.s.}}}{{\leq}}2\sqrt{\kappa}.

Here 𝒞0(K)\|\,\cdot\,\|_{\mathcal{C}^{0}(K)} denotes the supremum norm. Assuming ϵ=0\epsilon=0 for simplicity, it follows in particular that, with ff fixed, the performance of the estimator with respect to JNJ_{N} is within 𝒪(κ)\mathcal{O}(\kappa) of optimality in the limit of large data. Moreover, the full statement of Theorem 2 also includes explicit convergence rates (see Remark 3). See the remarks following Theorem 2 for further discussion.

However, as discussed in sec. A.5, bounding the performance with respect to JNJ_{N} as in Theorem 2 does not say anything about closeness of the estimates dϕ^N\widehat{d\phi}_{N} themselves to dϕd\phi. Thus, even if JN(dϕ^N)J_{N}(\widehat{d\phi}_{N}) is small, the estimated isochrons corresponding to dϕ^N\widehat{d\phi}_{N} need not resemble the true isochrons. This motivates Theorem 3 which shows, roughly speaking, that under some additional assumptions the estimated isochrons converge 𝒞1\mathcal{C}^{1}-uniformly to the true isochrons on forward invariant open subsets of KK. Moreover, the full statement of Theorem 1 also includes explicit convergence rates.

A.4 Relationship of the Temporal 1-Form to other representations of asymptotic phase

In this section we explain the relationship (mentioned in sec. A.3) between the asymptotic phase map P:𝐗Γ\mathrm{P}\colon\mathbf{X}\to\Gamma, circle-valued phase maps φ:𝐗S1\varphi\colon\mathbf{X}\to\mathrm{S}^{1}, and the Temporal 1-Form dϕd\phi. We continue under the assumptions and notation of sec. A.3; in particular, the state space 𝐗\mathbf{X} is also the basin of Γ\Gamma.

Lemma 1.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. The Temporal 1-Form dϕ:𝐗𝖳𝐗d\phi:\,\mathbf{X}\to\mathsf{T}^{*}\mathbf{X} is closed and satisfies dϕ,fω\langle d\phi,f\rangle\equiv\omega.

Proof.

Since P:𝐗Γ\mathrm{P}\colon\mathbf{X}\to\Gamma is a semiconjugacy, it follows that 𝖣Pf=fP\mathsf{D}\mathrm{P}\circ f=f\circ\mathrm{P}. Since dϕPαd\phi\coloneqq\mathrm{P}^{*}\alpha (sec. A.3), it follows that dϕ,fPα,fα,𝖣Pfα,fPω\langle d\phi,f\rangle\equiv\langle P^{*}\alpha,f\rangle\equiv\langle\alpha,\mathsf{D}P\circ f\rangle\equiv\langle\alpha,f\circ P\rangle\equiv\omega. Since α\alpha is closed and pullbacks commute with 𝐝\mathrm{\mathbf{d}}{}, dϕ=Pαd\phi=\mathrm{P}^{*}\alpha is closed. ∎

Given an open neighborhood U𝐗U\subset\mathbf{X} of Γ\Gamma which is forward invariant for the flow of ff, we extend the definition of circle-valued phase (sec. A.3) to include continuous maps φ:US1\varphi\colon U\to\mathrm{S}^{1}\subset\mathbb{C} satisfying φΦt=eiωtφ\varphi\circ\Phi_{t}=e^{i\omega t}\varphi for all t0t\geq 0. (We use this generality in the proof of Theorem 3.) For the following lemmas and theorem, dθd\theta denotes the standard angular form on the circle S1\mathrm{S}^{1} and i=1i=\sqrt{-1}.

Lemma 2.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. Let UU be a forward invariant open neighborhood of Γ\Gamma. Given any x0Ux_{0}\in U, there exists a unique circle-valued phase φ:US1\varphi\colon U\to\mathrm{S}^{1} satisfying φ(x0)=1\varphi(x_{0})=1. Moreover, φ𝒞k1\varphi\in\mathcal{C}^{k\geq 1} if f𝒞kf\in\mathcal{C}^{k}.

Proof.

Existence follows by first defining φ|Γ:ΓS1\varphi|_{\Gamma}\colon\Gamma\to\mathrm{S}^{1} by φ|ΓP(x0)=1\varphi|_{\Gamma}\circ\mathrm{P}(x_{0})=1 and φ|ΓΦt=eiωtφ|Γ\varphi|_{\Gamma}\circ\Phi_{t}=e^{i\omega t}\varphi|_{\Gamma}, then defining φφ|ΓP\varphi\coloneqq\varphi|_{\Gamma}\circ\mathrm{P}. (As explained in sec. A.3, P𝒞k\mathrm{P}\in\mathcal{C}^{k} when f𝒞kf\in\mathcal{C}^{k}.) To show uniqueness, let φ1\varphi_{1} and φ2\varphi_{2} be two such circle-valued phases. The ratio φ1/φ2\varphi_{1}/\varphi_{2} is constant along trajectories of ff, so φ1/φ2\varphi_{1}/\varphi_{2} is constant on Γ\Gamma. Since all trajectories converge to Γ\Gamma, it follows from continuity that φ1/φ2φ1(x0)/φ2(x0)=1\varphi_{1}/\varphi_{2}\equiv\varphi_{1}(x_{0})/\varphi_{2}(x_{0})=1 on UU, so φ1=φ2\varphi_{1}=\varphi_{2} as desired. ∎

Lemma 3.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. Let UU be a forward invariant open neighborhood of Γ\Gamma, and let β\beta be any 𝒞k0\mathcal{C}^{k\geq 0} closed 11-form on UU satisfying β,fω\langle\beta,f\rangle\equiv\omega. Then for any x0Ux_{0}\in U, there is a unique 𝒞k+1\mathcal{C}^{k+1} circle-valued phase φ:US1\varphi\colon U\to\mathrm{S}^{1} satisfying φ(x0)=1\varphi(x_{0})=1 and φ(dθ)=β\varphi^{*}(d\theta)=\beta.

Proof.

Since UU is forward invariant, it follows that Γ\Gamma is a deformation retraction of UU999In fact, a strong deformation retract. As an embedded submanifold of UU, the limit cycle is a strong deformation retract of a “tubular neighborhood” (Lee (2013) Chapter 6) containing it. Using the flow, one may construct a strong deformation retract of UU onto a tubular neighborhood. Following this homotopy with the strong deformation retraction of the tubular neighborhood onto the limit cycle then gives the desired strong deformation retraction. . Since β\beta is closed and satisfies β,fω\langle\beta,f\rangle\equiv\omega, it follows that the line integral of β\beta along any continuous closed curve in UU is an integer multiple of 2π2\pi. Thus, exponentiating ii times the line integrals of β\beta along arbitrary continuous paths from x0x_{0} defines a 𝒞k+1\mathcal{C}^{k+1} map φ:US1\varphi\colon U\to\mathrm{S}^{1} satisfying β=φ(dθ)\beta=\varphi^{*}(d\theta) (cf. the proof of (Pajitnov, 2006, Lem. 1.17)). Moreover, the condition β,fω\langle\beta,f\rangle\equiv\omega implies that φ\varphi is a circle-valued phase. Uniqueness of φ\varphi follows from Lemma 2. ∎

The following result summarizes some relationships between different representations of asymptotic phase.

Theorem 1.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}, with asymptotic phase map denoted by P:𝐗Γ\mathrm{P}\colon\mathbf{X}\to\Gamma.

  1. 1.

    The Temporal 1-Form dϕd\phi is the unique continuous closed 11-form on 𝐗\mathbf{X} satisfying

    dϕ,fω.\langle d\phi,f\rangle\equiv\omega.
  2. 2.

    If φ\varphi is any circle-valued phase, φP=φ\varphi\circ\mathrm{P}=\varphi and

    dϕ=φ(dθ)=i(𝐝φ)/φ.d\phi=\varphi^{*}(d\theta)=-i(\mathrm{\mathbf{d}}{\varphi})/\varphi.
  3. 3.

    Conversely, a choice of x0𝐗x_{0}\in\mathbf{X} uniquely determines a circle-valued phase φ\varphi satisfying φ(x0)=1\varphi(x_{0})=1 through the formula

    φ(x)=exp(iγ𝑑ϕ),\varphi(x)=\exp\left(i\int_{\gamma}d\phi\right), (20)

    where γ\gamma is any 𝒞0\mathcal{C}^{0} path joining x0x_{0} to xx.

Proof.

Claim 3 is immediate from Lemma 3 and its proof (taking U=𝐗U=\mathbf{X}). The first statement of Claim 2 follows since φP\varphi\circ\mathrm{P} and φ\varphi are both circle-valued phases agreeing on Γ\Gamma (since P|Γ=id|Γ\mathrm{P}|_{\Gamma}=\textnormal{id}|_{\Gamma}), so they coincide by Lemma 2.

Since any two circle-valued phases differ by a rotation of S1\mathrm{S}^{1} (Lemma 2), and since dθd\theta is rotation invariant, Lemma 3 implies Claim 1 and the equality dϕ=φ(dθ)d\phi=\varphi^{*}(d\theta) in Claim 2. The remaining equality in Claim 2 follows from the straightforward verification that φ(dθ)=i(𝐝φ)/φ\varphi^{*}(d\theta)=-i(\mathrm{\mathbf{d}}{\varphi})/\varphi for any 𝒞1\mathcal{C}^{1} circle-valued map φ\varphi, where 𝐝φ\mathrm{\mathbf{d}}{\varphi} is viewed as a complex-valued differential 11-form (Guillemin and Pollack, 2010, p. 192).

A.5 The Temporal 1-Form estimated from uncertain systems

In this section we prove two theorems relating the performance of the true Temporal 1-Form to that of estimates computed in the presence of noise, to which data from real-world systems and measurements are subjected. These are extensions of corresponding results in Kvalheim (2018, Sec. 3.4).

Throughout the remainder of this section, we assume for simplicity that the state space 𝐗\mathbf{X} (also assumed to be the basin of attraction of the limit cycle) is an open subset of n\mathbb{R}^{n}. We also assume for the remainder of this section that data takes values in a fixed compact set K𝐗K\subset\mathbf{X}.

In what follows we let \|\,\cdot\,\| be the Euclidean norm, and we denote induced operator norms by the same symbol. Given a continuous function β\beta on a set A𝐗A\subset\mathbf{X}, we define

β𝒞0(A)supxAβ(x).\|\beta\|_{\mathcal{C}^{0}(A)}\coloneqq\sup_{x\in A}\|\beta(x)\|.

We denote by Ω01(A)\Omega^{1}_{0}(A) the set of continuous 11-forms over AA, i.e., continuous sections of 𝖳𝐗|A\mathsf{T}^{*}\mathbf{X}|_{A}.

A.5.1 Assumptions about Data and Noise

We assume that we have a finite collection of pairs (xi,x˙i)i=1NK×n(x_{i},\dot{x}_{i})_{i=1}^{N}\subset K\times\mathbb{R}^{n}, with x˙i\dot{x}_{i} of the form x˙i=f(xi)+ηi\dot{x}_{i}=f(x_{i})+\eta_{i}. We consider the xiKx_{i}\in K and ηin\eta_{i}\in\mathbb{R}^{n} to be random variables, and we assume that there is a constant κ0\kappa\geq 0 such that

i:𝔼(ηi2)κ.\forall i\colon\mathbb{E}(\|\eta_{i}\|^{2})\leq\kappa. (21)

We also assume that the strong law of large numbers applies to yield the following equality with probability 11; quite general conditions ensuring this can be found in Doob (1990); Andrews (1987); Pötscher and Prucha (1989).

limN1Ni=1Nηi2𝔼(ηi2)\displaystyle\lim_{N\to\infty}\frac{1}{N}\sum_{i=1}^{N}\|\eta_{i}\|^{2}-\mathbb{E}(\|\eta_{i}\|^{2}) =a.s.0\displaystyle\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}0 (22)

We will impose additional conditions involving the xix_{i} for Theorem 3, but not for Theorem 2.

Remark 1.

This setup is sufficiently general that the ηi\eta_{i} could arise from measurement noise, or system noise, or both. In (Kvalheim, 2018, Sec. 3.4.2) it is argued in detail that this formulation applies to data from Itô SDEs. It also applies to data coming from Stratonovich SDEs, but there is a slight twist: to get sharper results, one should replace ff by a modified vector field related to Itô’s formula (Kvalheim, 2018, Sec. 3.4.2.2).

A.5.2 Performance of the estimated Temporal 1-Form

Denote by dϕd\phi the unique true Temporal 1-Form on 𝐗\mathbf{X} corresponding to the vector field f𝒞1f\in\mathcal{C}^{1}. We define the “unobserved” cost function JN:Ω01(K)[0,)J_{N}:\Omega_{0}^{1}(K)\to[0,\infty) which has dϕ𝒞0d\phi\in\mathcal{C}^{0} as its (nonunique) global minimizer, with minimum JN(dϕ)=0J_{N}(d\phi)=0, by

JN(β)\displaystyle J_{N}(\beta) :=1Ni=1N(β(xi),f(xi)ω)2.\displaystyle:=\frac{1}{N}\sum_{i=1}^{N}\left(\left\langle\beta(x_{i}),f(x_{i})\right\rangle-\omega\right)^{2}. (23)

We also define the following “observed” cost function JN^:Ω01(K)[0,)\widehat{J_{N}}:\Omega_{0}^{1}(K)\to[0,\infty) by

JN^(β)\displaystyle\widehat{J_{N}}(\beta) :=1Ni=1N(β(xi),f(xi)+ηiω)2.\displaystyle:=\frac{1}{N}\sum_{i=1}^{N}\left(\left\langle\beta(x_{i}),f(x_{i})+\eta_{i}\right\rangle-\omega\right)^{2}. (24)

Fix ϵ0\epsilon\geq 0. For each NN we fix any dϕ^NΩ01(K)\widehat{d\phi}_{N}\in\Omega_{0}^{1}(K) with the property that

JN^(dϕ^N)JN^(dϕ)+ϵ.\widehat{J_{N}}(\widehat{d\phi}_{N})\leq\widehat{J_{N}}(d\phi)+\epsilon. (25)

We think of dϕ^N\widehat{d\phi}_{N} as being the output of our algorithm to compute the Temporal 1-Form; our algorithm effectively attempts to minimize JN^\widehat{J_{N}}. We have the following performance bound on our algorithm (assuming its output satisfies (25)) with respect to minimizing JNJ_{N}.

Theorem 2.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. Then

14(JN(dϕ^N)2ϵdϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K))21Ni=1Nηi2.\displaystyle\frac{1}{4}\left(\frac{\sqrt{J_{N}(\widehat{d\phi}_{N})}-2\sqrt{\epsilon}}{\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}}\right)^{2}\leq\frac{1}{N}\sum_{i=1}^{N}\|\eta_{i}\|^{2}.

Thus, under the assumptions of the present section,

lim supN\displaystyle\limsup_{N\to\infty} JN(dϕ^N)2ϵdϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K)a.s.2κ.\displaystyle\frac{\sqrt{J_{N}(\widehat{d\phi}_{N})}-2\sqrt{\epsilon}}{\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}}\stackrel{{\scriptstyle\textnormal{a.s.}}}{{\leq}}2\sqrt{\kappa}.

In particular, if dϕ𝒞0(K),dϕ^N𝒞0(K)M0\|d\phi\|_{\mathcal{C}^{0}(K)},\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}\leq M_{0}, then

lim supNJN(dϕ^N)a.s.(4+22)M0κ+2ϵ.\displaystyle\limsup_{N\to\infty}\sqrt{J_{N}(\widehat{d\phi}_{N})}\stackrel{{\scriptstyle\textnormal{a.s.}}}{{\leq}}(4+2\sqrt{2})M_{0}\sqrt{\kappa}+2\sqrt{\epsilon}.

We make several remarks before giving the proof.

Remark 2.

The assumptions (21) and (22) are not needed for the first inequality of the theorem.

Remark 3 (Convergence rates).

The first inequality in the theorem shows that the left side converges to [0,κ][0,\kappa] at least as fast as the rate of convergence in the law of large numbers (22). Similarly, with respect to the weak law of large numbers, the same inequality implies that the rate of convergence in probability of the left side are at least as fast as the corresponding rates of convergence in probability for the empirical second moments of the ηi\eta_{i}. Similar remarks hold for (36) in Theorem 3.

Remark 4.

The appearance of dϕ𝒞0(K)\|d\phi\|_{\mathcal{C}^{0}(K)} in the first and second bounds suggests that a (“wilder”) Temporal 1-Form with large supremum norm is harder to estimate. See also (36) in Theorem 3.

Remark 5.

Since dϕ𝒞0d\phi\in\mathcal{C}^{0}, there exists M0>0M_{0}>0 such that the assumption dϕ𝒞0(K)M0\|d\phi\|_{\mathcal{C}^{0}(K)}\leq M_{0} holds. Since in practice the estimates dϕ^N\widehat{d\phi}_{N} produced by our algorithm are uniformly bounded, for practical purposes it is also the case that dϕ^𝒞0(K)M0\|\widehat{d\phi}\|_{\mathcal{C}^{0}(K)}\leq M_{0} for some M0M_{0} and all NN. Thus, the extra assumptions in the second portion of the theorem are quite mild.

Remark 6.

Since JN(dϕ)=0J_{N}(d\phi)=0, the estimate provided by Theorem 2 bounds how well our algorithm does at minimizing the “unobserved” cost function. Taking ϵ=0\epsilon=0 for simplicity, the final inequality of the theorem asserts that, with ff fixed, the performance of the estimator—as measured by JNJ_{N}—is asymptotically within 𝒪(κ)\mathcal{O}(\kappa) of optimality. However, Theorem 2 does not say anything about how close the estimate dϕ^N\widehat{d\phi}_{N} itself is to the true Temporal 1-Form dϕd\phi. This will be addressed in Theorem 3.

Proof of Theorem 2.

Expanding the summand in the definition of JN^\widehat{J_{N}} and rearranging, we see that, for any βΩ01(K)\beta\in\Omega^{1}_{0}(K):

JN(β)\displaystyle J_{N}(\beta) =JN^(β)1Ni=1Nβ(xi),ηi2\displaystyle=\widehat{J_{N}}(\beta)-\frac{1}{N}\sum_{i=1}^{N}\left\langle\beta(x_{i}),\eta_{i}\right\rangle^{2}
2Ni=1N(β(xi),f(xi)1)β(xi),ηi\displaystyle-\frac{2}{N}\sum_{i=1}^{N}\left(\left\langle\beta(x_{i}),f(x_{i})\right\rangle-1\right)\left\langle\beta(x_{i}),\eta_{i}\right\rangle (26)
JN^(β)+2[JN(β)]12[1Ni=1Nβ(xi),ηi2]12\displaystyle\leq\widehat{J_{N}}(\beta)+2\left[{J_{N}(\beta)}\right]^{{}_{{1}\over{2}}}\left[{\frac{1}{N}\sum_{i=1}^{N}\left\langle\beta(x_{i}),\eta_{i}\right\rangle^{2}}\right]^{{}_{{1}\over{2}}}
+1Ni=1Nβ(xi),ηi2,\displaystyle+\frac{1}{N}\sum_{i=1}^{N}\left\langle\beta(x_{i}),\eta_{i}\right\rangle^{2}, (27)

where the inequality follows from the Cauchy-Schwarz inequality. Since the true Temporal 1-Form dϕd\phi satisfies (dϕ,f1)0(\langle d\phi,f\rangle-1)\equiv 0, it follows immediately from (A.5.2) that

JN^(dϕ)\displaystyle\widehat{J_{N}}(d\phi) =1Ni=1Ndϕ(xi),ηi2.\displaystyle=\frac{1}{N}\sum_{i=1}^{N}\left\langle d\phi(x_{i}),\eta_{i}\right\rangle^{2}. (28)

Next, eqn. (25) and (A.5.2) imply that

JN(dϕ^N)ϵ+JN^(dϕ)+1Ni=1Ndϕ^N(xi),ηi2+2[JN(dϕ^N)]12[1Ni=1Ndϕ^N(xi),ηi2]12.\begin{split}&J_{N}(\widehat{d\phi}_{N})\leq\epsilon+\widehat{J_{N}}(d\phi)+\frac{1}{N}\sum_{i=1}^{N}\left\langle\widehat{d\phi}_{N}(x_{i}),\eta_{i}\right\rangle^{2}\\ &+2\left[{J_{N}(\widehat{d\phi}_{N})}\right]^{{}_{{1}\over{2}}}\left[{\frac{1}{N}\sum_{i=1}^{N}\left\langle\widehat{d\phi}_{N}(x_{i}),\eta_{i}\right\rangle^{2}}\right]^{{}_{{1}\over{2}}}.\end{split} (29)

Defining SN,TN0S_{N},T_{N}\geq 0 by

SN21Ni=1Nηi2,TN2JN(dϕ^N),S_{N}^{2}\coloneqq\frac{1}{N}\sum_{i=1}^{N}\|\eta_{i}\|^{2},\quad T_{N}^{2}\coloneqq J_{N}(\widehat{d\phi}_{N}),

substituting (28) into (29), and using Cauchy-Schwarz yields an inequality quadratic in TNT_{N}:

TN22SNdϕ^N𝒞0(K)TN(dϕ𝒞0(K)2+dϕ^N𝒞0(K)2)SN2ϵ.\begin{split}T_{N}^{2}&-2S_{N}\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}T_{N}-(\|d\phi\|_{\mathcal{C}^{0}(K)}^{2}+\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}^{2})S_{N}^{2}\\ &\leq\epsilon.\end{split}

This and the quadratic formula imply that

TN2SNdϕ^N𝒞0(K)+2[(dϕ𝒞0(K)2+2dϕ^N𝒞0(K)2)SN2+ϵ]12.\begin{split}T_{N}&\leq 2S_{N}\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}\\ &+2\left[{(\|d\phi\|_{\mathcal{C}^{0}(K)}^{2}+2\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}^{2})S_{N}^{2}+\epsilon}\right]^{{}_{{1}\over{2}}}.\end{split}

Using subadditivity of \sqrt{\,\cdot\,}, squaring, and rearranging yields

TN2ϵdϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K)2SN.\frac{T_{N}-2\sqrt{\epsilon}}{\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}}\leq 2S_{N}.

This yields the first inequality of the theorem statement; taking the lim sup\limsup of both sides as NN\to\infty and using (21) and (22) completes the proof. ∎

We now begin preparations for the statement of Theorem 3. Fix a compact subset Ω01(K)\mathcal{E}\subset\Omega^{1}_{0}(K) 101010Note that, by the Arzelà-Ascoli theorem, the compact subsets of Ω01(K)\Omega^{1}_{0}(K) are precisely the closed and bounded subsets of 11-forms which are equicontinuous. and define

:={β|β is closed}.\mathcal{F}:=\{\beta\in\mathcal{E}|\beta\textnormal{ is closed}\}.

It follows that \mathcal{F} is compact since it is closed in Ω01(K)\Omega^{1}_{0}(K). This follows since the limit in Ω01(K)\Omega^{1}_{0}(K) of a sequence (βi)(\beta_{i})\subset\mathcal{F} must have integral zero over any nullhomotopic continuous loop in KK, since the same is true of each βi\beta_{i}, so the limit is a closed 11-form.

For Theorem 3 we will assume that the strong “uniform law of large numbers” applies to yield

limNsupβ|JN(β)K(β,fω)2ρ𝑑μ|=a.s.0,\begin{split}\lim_{N\to\infty}\sup_{\beta\in\mathcal{F}}&\Bigg{|}J_{N}(\beta)-\int_{K}\left(\langle\beta,f\rangle-\omega\right)^{2}\rho d\mu\Bigg{|}\\ &\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}0,\end{split} (30)

where μ\mu is some Borel probability measure on KK. Quite general conditions ensuring this can be found in Andrews (1987); Pötscher and Prucha (1989). Recall that the “support” of a Borel measure μ\mu is the set of all points xx such that every neighborhood of xx has positive measure.

We first need to prove the following lemma, in which aff()\textnormal{aff}(\,\cdot\,) denotes the “affine hull” of a subset of Ω01(K)\Omega^{1}_{0}(K).

Lemma 4.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. Assume that UKU\subset K is an open neighborhood of Γ\Gamma in 𝐗\mathbf{X} which is forward invariant and contained in the support of μ\mu. Also assume that dϕd\phi\in\mathcal{F}, that \mathcal{F} is star-shaped with respect to dϕd\phi, and that there is δ>0\delta>0 such that

βaff():βdϕ𝒞0(U)δβ.\forall\beta\in\textnormal{aff}(\mathcal{F})\colon\|\beta-d\phi\|_{\mathcal{C}^{0}(U)}\leq\delta\implies\beta\in\mathcal{F}. (31)

Then there is c>0c>0 such that, for all βdϕ+\beta\in-d\phi+\mathcal{F},

cβ𝒞0(U)2Uβ,f2𝑑μ.c\|\beta\|_{\mathcal{C}^{0}(U)}^{2}\leq\int_{U}\langle\beta,f\rangle^{2}d\mu. (32)
Remark 7.

A stronger assumption implying the star-shaped hypothesis is that \mathcal{F} is convex, since then \mathcal{F} is star-shaped with respect to all points in \mathcal{F}.

Proof.

Define Q:Ω01(K)[0,)Q\colon\Omega^{1}_{0}(K)\to[0,\infty) to send β\beta to the right side of (32), ~dϕ+\tilde{\mathcal{F}}\coloneqq-d\phi+\mathcal{F}, and

B{βaff(~):β𝒞0(U)=δ}.B\coloneqq\{\beta\in\textnormal{aff}(\tilde{\mathcal{F}})\colon\|\beta\|_{\mathcal{C}^{0}(U)}=\delta\}. (33)

Since \mathcal{F} is compact and star-shaped with respect to dϕd\phi, it follows that ~\tilde{\mathcal{F}} is compact and star-shaped with respect to 0. Since QQ is continuous and BB is compact (by (31)), Q|BQ|_{B} attains a minimum c00c_{0}\geq 0.

We claim that c0>0c_{0}>0. Indeed, Q(β)=0Q(\beta)=0 implies that β|U,f|U0\langle\beta|_{U},f|_{U}\rangle\equiv 0, which implies that β|U=dh\beta|_{U}=dh for some 𝒞1\mathcal{C}^{1} function h:Uh\colon U\to\mathbb{R} constant along the flow of ff. But hh must be constant everywhere since all trajectories converge to Γ\Gamma, so β𝒞0(U)=0\|\beta\|_{\mathcal{C}^{0}(U)}=0. This and (33) imply that βB\beta\not\in B, so c0>0c_{0}>0.

Since ~\tilde{\mathcal{F}} is star-shaped with respect to 0, each nonzero β~\beta\in\tilde{\mathcal{F}} is of the form tγt\gamma for some γB\gamma\in B and tβ𝒞0(U)/δt\coloneqq\|\beta\|_{\mathcal{C}^{0}(U)}/\delta. Since Q(tγ)=t2Q(γ)t2c0Q(t\gamma)=t^{2}Q(\gamma)\geq t^{2}c_{0}, it follows that Q(β)β𝒞0(U)2c0/δ2Q(\beta)\geq\|\beta\|_{\mathcal{C}^{0}(U)}^{2}c_{0}/\delta^{2} for all β~\beta\in\tilde{\mathcal{F}}. Defining cc0/δ2c\coloneqq c_{0}/\delta^{2} completes the proof. ∎

Theorem 3.

Consider an oscillator generated by the 𝒞1\mathcal{C}^{1} vector field f:𝐗𝖳𝐗f\colon\mathbf{X}\to\mathsf{T}\mathbf{X}. Assume that UKU\subset K is an open neighborhood of Γ\Gamma in 𝐗\mathbf{X} which is forward invariant and contained in the support of μ\mu. Also assume that dϕ^N,dϕ\widehat{d\phi}_{N},d\phi\in\mathcal{F}. Then, under the assumptions of the present section,

limκ+ϵ0lim supNdϕ^Ndϕ𝒞0(U)=a.s.0.\lim_{\kappa+\epsilon\to 0}\limsup_{N\to\infty}\|\widehat{d\phi}_{N}-d\phi\|_{\mathcal{C}^{0}(U)}\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}0. (34)

If also \mathcal{F} is star-shaped with respect to dϕd\phi, (31) holds for some δ>0\delta>0, and ζ,ξ:[0,)\zeta,\xi\colon\mathbb{N}\to[0,\infty) satisfy

1Ni=1Nηi2ζ(N)|JN(dϕ^N)K(dϕ^N,fω)2ρ𝑑μ|ξ(N),\begin{split}&\frac{1}{N}\sum_{i=1}^{N}\|\eta_{i}\|^{2}\leq\zeta(N)\\ &\Bigg{|}J_{N}(\widehat{d\phi}_{N})-\int_{K}\left(\langle\widehat{d\phi}_{N},f\rangle-\omega\right)^{2}\rho d\mu\Bigg{|}\leq\xi(N),\end{split} (35)

then the following explicit convergence rates hold:

cdϕ^Ndϕ𝒞0(U)24[(dϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K))ζ(N)+ϵ]2+ξ(N)4[(2+2)M0ζ(N)+ϵ]2+ξ(N),\begin{split}&c\|\widehat{d\phi}_{N}-d\phi\|_{\mathcal{C}^{0}(U)}^{2}\\ &\leq 4\left[(\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)})\zeta(N)+\sqrt{\epsilon}\right]^{2}\\ &+\xi(N)\\ &\leq 4\left[(2+\sqrt{2})M_{0}\zeta(N)+\sqrt{\epsilon}\right]^{2}+\xi(N),\end{split} (36)

where c>0c>0 is the constant in Lemma 4, and the final inequality holds if dϕ𝒞0(K),dϕ^N𝒞0(K)M0\|d\phi\|_{\mathcal{C}^{0}(K)},\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}\leq M_{0}.

We make several remarks before giving the proof.

Remark 8.

As noted in Remark 7, the star-shaped hypothesis is automatic if \mathcal{F} is convex.

Remark 9.

The assumption (30) is not needed for the second portion of the theorem, since its role is replaced by (35).

Remark 10.

Unlike Theorem 2, Theorem 3 gives conditions under which the estimates dϕ^N\widehat{d\phi}_{N} converge to the true Temporal 1-Form dϕd\phi uniformly on suitable neighborhoods UKU\subset K of Γ\Gamma. This convergence implies convergence of the corresponding isochron estimates to the true isochrons 𝒞1\mathcal{C}^{1}-uniformly on such neighborhoods. This implies, roughly speaking, that the estimated isochrons corresponding to dϕ^N\widehat{d\phi}_{N} will resemble the true isochrons near Γ\Gamma up to one order of smoothness.

Remark 11.

For non-forward invariant UU the conclusions of the theorem do not generally hold, because it is possible to construct examples of such a UU on which there are many closed 11-forms β\beta satisfying β,fω\langle\beta,f\rangle\equiv\omega, and this non-uniqueness makes it possible to construct examples violating the conclusions of the theorem. However, given some curious apparent experimental successes with non-forward invariant UU (cf. the guineafowl example of figure 6), it would be interesting to know whether something useful can be said about a subclass of open sets UU strictly larger than those considered in the theorem.

Remark 12.

When f𝒞k+1f\in\mathcal{C}^{k+1}, it is possible to extend Theorem 3 to obtain convergence in 𝒞k\mathcal{C}^{k} norm rather than merely 𝒞0\mathcal{C}^{0} norm. The proof is nearly the same, but requires additional care in defining function spaces (since KK is an arbitrary compact subset of 𝐗\mathbf{X}). We defer a careful treatment to future work.

Proof of Theorem 3.

We first establish (34). By eqn. (30), with probability 11 the function JN|:[0,)J_{N}|_{\mathcal{F}}\colon\mathcal{F}\to[0,\infty) converges uniformly to the continuous function J:[0,)J\colon\mathcal{F}\to[0,\infty) defined by

J(β)K(β,fω)2𝑑μ.J(\beta)\coloneqq\int_{K}\left(\langle\beta,f\rangle-\omega\right)^{2}d\mu.

Hence

lim supNJ(dϕ^N)=a.s.lim supNJN(dϕ^N),\limsup_{N\to\infty}J(\widehat{d\phi}_{N})\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}\limsup_{N\to\infty}J_{N}(\widehat{d\phi}_{N}),

so Theorem 2 implies that

limκ+ϵ0lim supNJ(dϕ^N)=a.s.0.\lim_{\kappa+\epsilon\to 0}\limsup_{N\to\infty}J(\widehat{d\phi}_{N})\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}0.

Defining \mathcal{M} to be the set of global minimizers of J|J|_{\mathcal{F}} (attaining the minimum value of 0) and using compactness of \mathcal{F}, it follows that

limκ+ϵ0lim supNd(dϕ^N,)𝒞0(K)=a.s.0,\lim_{\kappa+\epsilon\to 0}\limsup_{N\to\infty}\,d\left({\widehat{d\phi}_{N}},{\mathcal{M}}\right)_{\mathcal{C}^{0}(K)}\stackrel{{\scriptstyle\textnormal{a.s.}}}{{=}}0,

where d(α,)𝒞0(K)infβαβ𝒞0(K)d\left({\alpha},{\mathcal{M}}\right)_{\mathcal{C}^{0}(K)}\coloneqq\inf_{\beta\in\mathcal{M}}\|\alpha-\beta\|_{\mathcal{C}^{0}(K)}. (Cf. the Berge maximum theorem (Aliprantis and Border, 2006, Thm 17.31).)

Thus, to complete the proof of (34) it suffices to show that the restriction to UU of every closed 11-form in \mathcal{M} coincides with dϕ|Ud\phi|_{U}. Since JJ has minimum 0 and UU is contained in the support of μ\mu, it follows that any β\beta\in\mathcal{M} must satisfy β,fω\langle\beta,f\rangle\equiv\omega. Since UU is forward invariant, the same argument as in the final paragraph of the proof of Theorem 1 (using Lemmas 2 and 3) then implies that β=dϕ|U\beta=d\phi|_{U}, as desired.

Next, assume that \mathcal{F} is star-shaped with respect to dϕd\phi, (31) holds for some δ>0\delta>0, and (35) holds. Defining βNdϕ^Ndϕdϕ+\beta_{N}\coloneqq\widehat{d\phi}_{N}-d\phi\in-d\phi+\mathcal{F} for each NN, Lemma 4 implies the first inequality below:

cβN𝒞0(U)2UβN,f2𝑑μKβN,f2𝑑μ=J(dϕ^N).\begin{split}c\|\beta_{N}\|_{\mathcal{C}^{0}(U)}^{2}&\leq\int_{U}\langle\beta_{N},f\rangle^{2}d\mu\\ &\leq\int_{K}\langle\beta_{N},f\rangle^{2}d\mu=J(\widehat{d\phi}_{N}).\end{split} (37)

The second inequality follows from UKU\subset K, and the equality follows from expanding the integrand defining JJ. On the other hand, Theorem 2 implies that

14(JN(dϕ^N)2ϵdϕ𝒞0(K)+(1+2)dϕ^N𝒞0(K))21Ni=1Nηi2.\displaystyle\frac{1}{4}\left(\frac{\sqrt{J_{N}(\widehat{d\phi}_{N})}-2\sqrt{\epsilon}}{\|d\phi\|_{\mathcal{C}^{0}(K)}+(1+\sqrt{2})\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}}\right)^{2}\leq\frac{1}{N}\sum_{i=1}^{N}\|\eta_{i}\|^{2}.

Rearranging and using (35) and (37) yields the first inequality in (36). The second inequality in (36) follows directly from this and dϕ𝒞0(K),dϕ^N𝒞0(K)M0\|d\phi\|_{\mathcal{C}^{0}(K)},\|\widehat{d\phi}_{N}\|_{\mathcal{C}^{0}(K)}\leq M_{0}. ∎