We consider a periodically forced 1D Langevin equation that possesses two stable periodic solutions in the absence of noise. We ask the question: is there a most likely noise-induced transition path between these periodic solutions that allows us to identify a preferred phase of the forcing when tipping occurs? The quasistatic regime, where the forcing period is long compared to the adiabatic relaxation time, has been well studied; our work instead explores the case when these time scales are comparable. We compute optimal paths using the path integral method incorporating the Onsager–Machlup functional and validate results with Monte Carlo simulations. Results for the preferred tipping phase are compared with the deterministic aspects of the problem. We identify parameter regimes where nullclines, associated with the deterministic problem in a 2D extended phase space, form passageways through which the optimal paths transit. As the nullclines are independent of the relaxation time and the noise strength, this leads to a robust deterministic predictor of the preferred tipping phase in a regime where forcing is neither too fast nor too slow.
The study of “tipping points” has applications to disease eradication,1–4 cell fate decisions,5 climate dynamics,6–8 and many other problems and has been explored in both time-autonomous and time-dependent models. Mathematical mechanisms for such tipping points in low-dimensional dynamical systems have been classified according to whether they involve, predominantly, a bifurcation, noise, or parameter drift.9,10 This paper focuses on noise-induced tipping, between distinct periodic attractors, in systems with periodic forcing. While many studies predict the time it takes for tipping to occur, we are motivated by applications with intrinsic seasonal forcing (e.g., annual cycling of Arctic sea ice extent6,7 or yearly disease outbreaks11) and instead focus on the likely phase for tipping events. In particular, we are interested in determining whether there is a preferred phase of the forcing when the system is most likely to tip from one attractor to another, and, if so, which deterministic features might predict that phase. A better understanding of these features, and the possible role of intrinsic relaxation vs exogenous forcing time scales, may help identify a phase of the forcing when the system is most vulnerable to a noise-induced abrupt transition.
I. INTRODUCTION
In this paper, we focus on a simple stochastic differential equation for which we can readily control the characteristic time scales of the problem. Specifically, we consider
Here, Xt ∈ ℝ is a stochastic process parameterized by time t ≥ 0, Wt is a standard Wiener process, and σ > 0 denotes the noise strength. We choose the parameters ɛ, α, A > 0 so that the deterministic problem has two stable periodic solutions separated by an unstable one [represented by solid and dashed curves in Fig. 1(a)]. The parameter ɛ represents a ratio of the characteristic relaxation time of the flow to the period of the forcing and arises when we nondimensionalize time by the forcing period. Throughout the paper, the “lower” and “upper” stable periodic solutions of are denoted by and , respectively, and the “middle” unstable one is denoted by , where for t ∈ ℝ. The presence of noise introduces a mechanism for transition from one stable periodic solution to the other. Figure 1(a) shows some example realizations of (1). Because we are interested in the phase of tipping events relative to the forcing, rather than the absolute timing of expected tipping, we often wrap noisy realizations of (1) to a cylinder; see Fig. 1(b).
(a) Schematic (ɛ, σ)-parameter plane. Insets are tipping realizations of Eq. (1). Black solid (dashed) curves represent stable (unstable) solutions to the deterministic equation (1) with σ = 0. Parameters in each inset are taken as follows: (ɛ, σ) = (0.25, 0.1) in regime I; (ɛ, σ) = (0.025, 0.1) in regime II; (ɛ, σ) = (0.25, 0.8) in regime III; (ɛ, σ) = (1, 0.2) in regime IV, with (α, A) = (0.25, 0.5) in all regimes. (b) We are interested in the phase of tipping and, thus, view our stochastic system (1) as lying in a cylindrical phase space. For (b), we used (ɛ, σ, α, A) = (0.01, 0.5, 0.15, 0.2).
(a) Schematic (ɛ, σ)-parameter plane. Insets are tipping realizations of Eq. (1). Black solid (dashed) curves represent stable (unstable) solutions to the deterministic equation (1) with σ = 0. Parameters in each inset are taken as follows: (ɛ, σ) = (0.25, 0.1) in regime I; (ɛ, σ) = (0.025, 0.1) in regime II; (ɛ, σ) = (0.25, 0.8) in regime III; (ɛ, σ) = (1, 0.2) in regime IV, with (α, A) = (0.25, 0.5) in all regimes. (b) We are interested in the phase of tipping and, thus, view our stochastic system (1) as lying in a cylindrical phase space. For (b), we used (ɛ, σ, α, A) = (0.01, 0.5, 0.15, 0.2).
In Fig. 1(a), we present a schematic diagram of different ɛ and σ regimes in which noise-induced transitions differ qualitatively. Comparing insets in Fig. 1(a), we see that the value of ɛ in (1) can strongly impact the geometry and separation of the periodic solutions. This timescale ratio, in turn, may affect when and how the sample paths cross the unstable periodic solution. The adiabatic regime (regime II) in Fig. 1(a), where ɛ ≪ 1, has been well studied.8,10,12–17 In this case, the relaxation time to stable periodic orbits is fast compared to the forcing period. Consequently, noise-induced transitions are “instantaneous” and occur with high probability when the separation between the stable and unstable periodic orbits is minimal and when an associated potential barrier height is lowest.
In contrast, regime I of Fig. 1(a), obtained with and otherwise the same parameters, has been the focus of only a small number of studies.18–20 In this regime, the transition occurs more slowly, especially after crossing the unstable periodic solution, and the separation between the deterministic solutions does not vary significantly over the forcing period. The distribution of tipping events may display more than one narrow peak when crossing the unstable limit cycle.19 Moreover, in this regime, where noise is not too large, the impacts of noise and deterministic drift may act on comparable time scales. These observations suggest challenges with identifying a preferred tipping phase outside of the adiabatic regime. Indeed, while formal matched asymptotic expansions of solutions to the Fokker–Planck equation yield quantitative information about the distribution of tipping events when crossing the unstable orbit,19 little is known about what deterministic quantities govern when the most probable paths depart from the lower stable orbit. Such deterministic indicators are important for determining when the system is most susceptible to noise-induced transitions.
In the noise-dominated regime [regime III in Fig. 1(a)], randomness overpowers the deterministic dynamics and tipping samples are extremely noisy, with deterministic features largely washed out. In the fast forcing regime, marked as regime IV in Fig. 1(a), where the timescale ratio ɛ is large, the relaxation time of the dynamics is slow compared to the forcing period and transitions can take a long time compared to the forcing period, possibly making the notion of the preferred tipping phase moot.
As shown in Fig. 2(a), we let tl, tm, and tu denote the times when a sample path departs from , first crosses , and arrives at , respectively. They are defined as
Because we are interested in the phase of tipping, relative to the forcing, rather than the absolute time, we will typically mod these times by the forcing period T and speak of the “tipping phases” that then derive from tl, tm, and tu. In Figs. 2(b) and 2(c), we present the distributions of tl, tm, and tu modded by T = 1. We take deterministic parameters in regime I and consider noise intensity σ = 0.15 for Fig. 2(b) and σ = 0.3 for Fig. 2(c). We notice that for both values of σ, the histograms of tl have a more prominent peak than the histograms of tm and tu. Moreover, the location of the peak in tl does not shift with the change in σ the way that the peak in tm does. Indeed, we find that the histograms for values of σ ∈ [0.15, 0.4] (not shown here) give consistent results; the peak of tl broadens but does not shift significantly with increasing noise strength in this range. We will identify tl as a robust “tipping phase” in this intermediate regime I. In addition to looking for conditions on ɛ and σ under which we can identify a robust “tipping phase,” we aim to investigate what features of the deterministic dynamics select such a phase when it exists.
(a) Example tipping trajectories for (1) and diagram showing the notation used in this paper. Parameters taken in the simulations are (ɛ, A, α, σ) = (0.25, 0.65, 0.15, 0.2). (b) and (c) Normalized histograms of tl, tm, and tu, which denote the phases when sample paths of (1) cross , , and , respectively. The noise intensity is σ = 0.15 in (b) and σ = 0.3 in (c).
(a) Example tipping trajectories for (1) and diagram showing the notation used in this paper. Parameters taken in the simulations are (ɛ, A, α, σ) = (0.25, 0.65, 0.15, 0.2). (b) and (c) Normalized histograms of tl, tm, and tu, which denote the phases when sample paths of (1) cross , , and , respectively. The noise intensity is σ = 0.15 in (b) and σ = 0.3 in (c).
In this paper, we study noise-induced transitions between stable periodic solutions in the intermediate forcing regime [regime I shown in Fig. 1(a)]. We use the framework of “exit problems,”21 where the most probable path of escape, from one domain of attraction to another, is typically determined by taking the limit σ → 0. In this setting, “most probable paths” (also called “optimal paths”) are often approximated by calculating the critical points of either the Freidlin–Wentzell (FW)18,19,21–23 or the Onsager–Machlup (OM) rate functional, which include an additional Ito correction term to the FW functional.24,25 In the autonomous setting, the use of the FW functional can be rigorously justified using the theory of large deviations,21,26 and for autonomous drift with a single time scale, the FW and OM functionals yield essentially identical most probable paths.27,28 Efficient computational techniques for computing most probable paths range from direct minimization algorithms,29,30 ordered upwind methods applied to the corresponding Hamilton–Jacobi equations,31 finite difference schemes that exploit the Hamiltonian structure of the Euler–Lagrange equations,32 and numerical continuation methods.33 For nonautonomous systems with periodic forcing, the analysis is nuanced since the additional time scale inherent to the periodic forcing may introduce discrepancies between minimizers of the FW and OM functionals. Moreover, for our specific problem, consideration of other small parameters in the problem, such as the timescale ratio ɛ, may be important in resolving discrepancies.
The rest of the paper is organized as follows. In Sec. II, we formulate the problem using a path integral approach. We describe the FW and OM functionals that we consider, and how the parameters ɛ and σ enter the variational problem for computing optimal paths. Section III contains the main contributions of this paper. Focusing on regime I where ɛ = O(1), we perform numerical simulations over a range of parameters in order to identify a robust definition of the preferred tipping phase. We find that it is associated with the time of initial departure from the lower stable periodic solution, as suggested by the histograms of Figs. 2(b) and 2(c). It occurs at a phase of the forcing when the flow near this stable periodic solution changes, momentarily, from negative to positive, i.e., from downwards to upwards in Fig. 2(a). This phase is determined by an intersection of the periodic solution with the nullclines f(x, t) = 0 of the deterministic system. We conclude in Sec. IV with a discussion of our results, limitations of the method, and some avenues for future research. We further numerically investigate the most probable transition phase in a conceptual Arctic sea ice model due to Eisenman and Wettlaufer,7 which has a strong seasonal forcing and possesses bistable parameter regimes. We use insights from our case study to conjecture that this model is in regime IV of Fig. 1(a), the fast forcing case.
To simplify the notation, throughout the paper, we use to represent time derivative dx/dt and subscript to denote partial derivative, e.g., Vx(x, t) : = ∂V/∂x.
II. MATHEMATICAL FORMULATION
A. Deterministic problem
We first discuss the role of the parameters in the deterministic problem
where V(x, t) is a potential associated with the deterministic dynamics f(x, t)/ɛ, which is the drift term in (1)
The parameter A determines the forcing strength and α breaks a reflection symmetry of the potential. The timescale ratio ɛ controls how (un)stable the periodic solutions of (2) are; their Floquet multipliers, , depend exponentially on 1/ɛ
Figure 3 summarizes the dependence of the deterministic dynamics of (2) on the parameters. The phase diagram in Fig. 3(b) shows that (2) admits four qualitatively distinct parameter regions for α > 0; these are characterized by different configurations of the nullclines f(x, t) = 0, in the (t, x)-phase plane. In this paper, we focus on regions R1 − R3 of the (A, α)-parameter plane where, for a given ɛ, there are two stable periodic solutions. (There is only one stable periodic solution in R4.) The location of the saddle-node bifurcations of Fig. 3(a) depends on ɛ as indicated. We describe these three parameter regions and their nullclines as follows:
The nullclines in R1, defined by , consist of three curves that partition extended phase space into four regions with vertical flow directions that alternate in sign. A standard Poincaré analysis proves existence of two stable periodic solutions for all ɛ values in R1.
Region R2, defined by and , has two nullcline curves that divide the phase space into three regions.
In R3, defined by , the phase space is partitioned by the nullcline into two regions.
(a) Hysteretic bifurcation diagrams (in α) associated with (2) for two ɛ values and A = 0.25. The inset figure shows the solutions for α = 0.1. (b) Regimes R1 − R4, in the (A, α)-parameter plane, with qualitatively different flow geometry; inset figures illustrate the flow geometry for each, with the nullclines [curves where f(x, t) = 0] overlaid in red. The dashed curves track the location of the right-most saddle-node bifurcation in α, corresponding to various ɛ values (the region with three periodic solutions is below these curves). (c) and (d) Phase portraits of (2) with black solid (dashed) curves denoting the stable (unstable) periodic orbits and blue curves tracking the nullclines; as shown, the nullclines no longer approximate the orbits well for larger ɛ. (A, α) = (0.25, 0.1) in these two examples.
(a) Hysteretic bifurcation diagrams (in α) associated with (2) for two ɛ values and A = 0.25. The inset figure shows the solutions for α = 0.1. (b) Regimes R1 − R4, in the (A, α)-parameter plane, with qualitatively different flow geometry; inset figures illustrate the flow geometry for each, with the nullclines [curves where f(x, t) = 0] overlaid in red. The dashed curves track the location of the right-most saddle-node bifurcation in α, corresponding to various ɛ values (the region with three periodic solutions is below these curves). (c) and (d) Phase portraits of (2) with black solid (dashed) curves denoting the stable (unstable) periodic orbits and blue curves tracking the nullclines; as shown, the nullclines no longer approximate the orbits well for larger ɛ. (A, α) = (0.25, 0.1) in these two examples.
Figure 3 highlights another ɛ-dependent feature of the dynamics. When ɛ ≪ 1 in (2), the periodic solutions are well-approximated by the nullclines f(x, t) = 0, as seen in Fig. 3(c) obtained for ɛ = 0.01. In contrast, for the moderate case ɛ = 0.5 of Fig. 3(d), there are significant deviations between the periodic solutions and the nullclines, leading to a pronounced intertwining of the (blue and black) curves in that phase portrait. This indicates intervals of time when the flow is upwards, say, along the lower stable solution, i.e., toward the upper unstable solution, and so on. These intervals become more pronounced in R2 and R3, where, provided ɛ is large enough, three periodic solutions may exist even though there are fewer than three curves comprising the nullcline; in these regions, the double well structure of V(x, t) is temporarily lost during certain phases of the forcing.
B. Path integral approach
We use the path integral formulation to identify a possible preferred phase for tipping. For tf > t0 ≥ 0, let denote the set of paths connecting a point (t0, x0) on and a point (tf, xf) on . If is differentiable, then the probability of the tipping trajectories lying in an infinitesimally small tubular neighborhood of satisfies
where dW [x] is the Wiener measure,25,34 and the proportionality reflects the missing normalization prefactor. Heuristically, this formulation can be derived by considering an equally spaced partition t0 = t1 < t2 < ⋅ ⋅ ⋅ < tn = tf of width Δt, calculating the probability that a realization passes through a sequence of intervals [ai, bi] at time ti, and taking the limit |bi − ai| → 0 and Δt → 0. Derivations following this approach can be found in the physics literature,25,34 and we include our version of the formal derivation in Appendix A.
The probability of a tipping event lying in a small tubular neighborhood of a tipping path is then determined by integrating over this neighborhood. Similar to Laplace’s method for the asymptotic expansion of exponential integrals, we expect as σ → 0 that the dominant contribution to this integral is concentrated about local minimizers of the argument of the exponential in the definition of (5). With this as motivation, we define the “most probable path” or “optimal path” connecting points (t0, x0) and (tf, xf) to be minimizers of the Onsager–Machlup (OM) functional given by
which can be expressed as the sum of the Freidlin–Wentzell (FW) rate functional IFW21 and an additional integral IOM2. Here, the admissible set is and denotes the Lagrangian. We note that for our model problem, the integrand of IOM2 is simply fx = 1 − 3x2, so that this is typically a positive contribution to when the paths are near the unstable solution and a negative contribution when the paths are near the stable solutions and .
Since is convex in , to prove the existence of a minimum in , it is sufficient to show that Iσ is coercive with respect to the H1 norm. The next theorem establishes this fact (as ); the existence of a minimizer is then a corollary following standard techniques from the direct method of the calculus of variations.35
For all σ ∈ [0, 1], there exists such that for all .
Local minimizers of (6) satisfy the Euler–Lagrange equations:
These equations constitute a second order nonlinear boundary value problem in x on the domain [t0, tf] and can be rewritten in the Hamiltonian form, for time-varying Hamiltonian function
as
Here, the “momentum” measures the deviation from the deterministic flow. The FW functional in (6) can then be expressed in terms of Ψ as .
The path integral formulation assumes that the noise strength σ is small. If σ is sufficiently small compared to the timescale ratio ɛ, such that σ2/ɛ ≪ 1, the naive regular perturbation in σ of (9), at leading order in σ, is
In this system, Ψ = 0 forms an invariant submanifold foliated by solutions of the deterministic dynamics . Thus, for σ → 0, optimal paths are well-approximated by the heteroclinic orbits of (9) connecting deterministic solutions.36,37 Moreover, (10) is the Hamiltonian form of the Euler–Lagrange equations for the FW rate functional IFW. Since IOM2 is independent of , minimizers of Iσ will converge uniformly as σ → 0 to a minimizer of IFW; see Appendix B.
III. NONADIABATIC FORCING REGIME:
The remainder of the paper addresses the intermediate regime I in Fig. 1(a), where ɛ is . In this case, there is no small parameter to exploit to simplify the analysis. We investigate the contribution IOM2 to the functional (6) in determining optimal paths, which are then validated by Monte Carlo simulations. This section concludes with a parameter study of (1), based on the path integral method using the Onsager–Machlup functional (6).
A. Gradient flow
To solve the second order boundary value problem (7), we introduce a gradient flow associated with . Specifically, introducing the artificial “time” s ≥ 0, we consider solutions to the following partial differential equation:
where xg(t) is a sufficiently smooth initial guess of the most probable path. To numerically solve (11), we use MATLAB’s built-in parabolic initial-boundary value solver pdepe.38
Stationary solutions of (11) correspond to solutions of the Euler–Lagrange equations (7). If we let u(s, t) denote a solution to (11), we find [after integrating by parts and applying (11)] that
where we have suppressed the arguments in f and u in order to make the notation compact. Hence, Iσ[u(s, ⋅ )] is a monotonically decreasing function in s. Furthermore, by Theorem 1, Iσ[u(s, ⋅ )] is bounded from below, and, thus, the pointwise limit lim s→∞Iσ[u(s, ⋅ )] exists and . Iσ is a Lyapunov function on , and, by Corollary 2, lim s→∞u(s, t) = x(t), where x(t) is a local minimizer of the OM functional and, hence, a solution to the Euler–Lagrange equations.39 Convergence to a steady state is assessed by the value of , i.e., when the Euler–Lagrange equations are approximately satisfied.
For the gradient flow, we impose a starting position at and a final position at . Figure 4 presents results for six forcing periods between t0 and tf. The initial condition for the gradient flow is chosen as a piecewise constant function
where tj is varied on a uniform grid between t0 and tf. A few examples of such initial conditions u(0, t) are shown as blue curves in Fig. 4(a), where tj takes values uniformly in [3, 4], and all of the initial solutions converge to the same stationary solution represented by the red curve. In this example, we find six different stationary solutions {Pi: i = 1, …, 6}, each of which represents a different local minimizer of given by (6); see Fig. 4(b). The paths {P2, …, P6} are the same up to translation by a multiple of the forcing period. P1 is also a local minimizer of , but its form depends on t0, whereas the remaining optimal paths are insensitive to the choice of t0 and tf; see Appendix C for details.
Results for (ɛ, A, α, σ) = (0.25, 0.5, 0.2, 0.4). (a) Examples of 16 initial guesses (blue) evolving under the gradient flow to the same transition path (red). (b) Stationary solutions of (11) for 64 initial guesses of the form (12) with uniformly spaced tj ∈ [0, 6]. Note that P3, …, P6 are translates, by successive forcing periods, of P2.
Results for (ɛ, A, α, σ) = (0.25, 0.5, 0.2, 0.4). (a) Examples of 16 initial guesses (blue) evolving under the gradient flow to the same transition path (red). (b) Stationary solutions of (11) for 64 initial guesses of the form (12) with uniformly spaced tj ∈ [0, 6]. Note that P3, …, P6 are translates, by successive forcing periods, of P2.
Figure 5 compares optimal paths, obtained from the gradient flow (11), for various values of σ. This includes setting σ = 0, which corresponds to using the FW functional. This example enforces boundary conditions so that all optimal transition paths start at the same initial point on and terminate four periods later on . We make the following observations:
Before crossing , the optimal paths are concentrated along a similar trajectory, and they all leave around the same phase of the forcing, marked by the green square. However, after crossing , the optimal paths spread out in a way that is strongly dependent on the noise intensity.
The minimizer of IFW stays near for a full cycle before it exits to . This is in contrast to the minimizers of , where the path may transition quickly to the upper stable solution .
Optimal paths for different σ cross at different times and quickly relax to follow the deterministic flow after crossing.
For larger σ, a small gap between the optimal path and the stable periodic solutions before and after the transition is more visible. This points to a possible failure of the minimizers of the OM functional to adequately approximate the most probable path for larger σ since we expect the system to track the deterministic dynamics there. [We note that the contribution of IOM2 to the functional (6) is more negative if the path is shifted from to below since fx = 1 − 3x2.]
Stationary solutions to (11) for (ɛ, A, α) = (0.25, 0.5, 0.2) and various σ, where “using IFW” corresponds to setting σ = 0 in the functional (6). The green square marks where paths cross .
The above observations can be understood by considering the relative scales of IFW and IOM2 both before and after crossing the unstable periodic solution for different regimes in Fig. 1, as described below.
Before crossing the unstable solution, the optimal path must go against the flow to reach that barrier, and, consequently, IFW is necessarily large and bounded away from 0. For small σ, the contribution from the IOM2 term is negligible in comparison with the IFW term along this part of the optimal path. After crossing the unstable orbit, the optimal path can track the deterministic drift, and, hence, the contribution from IFW can vanish for that portion of the path.19 As noted in Sec. II B, for the adiabatic regime, the IOM2 term remains negligible throughout time and IFW alone provides a good method for computing optimal paths.
In the nonadiabatic forcing regime, we consider, however, IOM2 has the opportunity to play a role in shaping the portion of the optimal path after reaching the unstable orbit since contributions to IFW may vanish there. This quality distinguishes optimal path dynamics in the adiabatic and nonadiabatic regimes. In particular, for the intermediate forcing regime, the IOM2 term, which is related to the local expansion/contraction rate of the flow, selects when the optimal path departs the unstable orbit. Without this term, optimal paths computed using only IFW (the blue curve in Fig. 5) track the unstable orbit for a full cycle under intermediate forcing. This is because the path can follow the deterministic dynamics without contributing to the value of IFW. The inclusion of the IOM2 term introduces a cost for noisy trajectories to remain near the unstable orbit, and this is important for computing optimal paths in our intermediate forcing regime. [Note that the contribution from IOM2 is positive for portions of the path in the phase space band with .]
B. Comparison with Monte Carlo simulations
We numerically validate the path integral method by comparing optimal paths with the results of Monte Carlo simulations. The reported mean squared error associated with halving the number of samples quantifies our convergence test of the Monte Carlo results to the distribution. It is defined as
where is the fraction of samples in each bin m; see Appendix D for more details about the convergence test and six comparisons for different values of (ɛ, σ).
Figure 6 illustrates results obtained with the path integral method compared to those from Monte Carlo simulations for parameters (ɛ, A, α, σ) = (0.1, 0.3, 0.15, 0.2). This figure shows the distribution of the sample paths on the (unwrapped) cylinder and two examples of distributions in x at t = −0.2 and 0, compared with the values of optimal paths (marked by red lines) at those times. These distributions are fairly concentrated as shown in the middle row of Fig. 6. We zoom in near the peaks of the distributions, in the bottom row, to better reveal the agreement between the optimal path and the Monte Carlo simulation results. Moreover, we notice that distributions broaden after tm, the time when paths cross the unstable periodic solutions , after which paths will simply follow the deterministic flow to .
Comparison of optimal paths with the time-evolved distribution obtained via Monte Carlo simulations at (ɛ, A, α, σ) = (0.1, 0.3, 0.15, 0.2). The mean absolute percentage error between the mode of the time-evolved distribution and the optimal path for t ∈ [ − 0.5, tm] is 4%, where tm ≈ 0.1. Two examples of normalized distributions of x are shown in yellow at t = −0.2 and t = 0 and compared to the value of the optimal path position, indicated by a red vertical line. The bottom two figures are zoomed-in distributions in x at t = −0.2 and t = 0. E = 8 × 10−6 defined as (13) in the convergence test. There are N = 12 772 tipping samples; M = 50 bins used for estimating the distribution of x ∈ [ − 1.5, 1.5], and the number of time steps in each sample is K = 104.
Comparison of optimal paths with the time-evolved distribution obtained via Monte Carlo simulations at (ɛ, A, α, σ) = (0.1, 0.3, 0.15, 0.2). The mean absolute percentage error between the mode of the time-evolved distribution and the optimal path for t ∈ [ − 0.5, tm] is 4%, where tm ≈ 0.1. Two examples of normalized distributions of x are shown in yellow at t = −0.2 and t = 0 and compared to the value of the optimal path position, indicated by a red vertical line. The bottom two figures are zoomed-in distributions in x at t = −0.2 and t = 0. E = 8 × 10−6 defined as (13) in the convergence test. There are N = 12 772 tipping samples; M = 50 bins used for estimating the distribution of x ∈ [ − 1.5, 1.5], and the number of time steps in each sample is K = 104.
To quantify the discrepancy between the optimal path and Monte Carlo simulation results, we compute the mean absolute percentage error (MAPE) between the mode of time-evolved distributions and optimal paths for t ∈ [− 0.5, tm], i.e., before crossing . The MAPE is calculated as
where mi denotes the mode of the distribution at each time step ti, oi is the value of the optimal path in x at ti, is the average of the mode of the distribution over t ∈ [− 0.5, tm], and k represents the total number of time steps in t ∈ [− 0.5, tm]. More comparison results for different values of ɛ and σ are summarized in Table I, and plots are presented in Appendix D. Results from the Monte Carlo comparison suggest that the optimal path obtained from the OM functional gives a good estimate of the mode of the distribution prior to crossing in this intermediate forcing regime. We also find that it does better than the FW functional alone; see comparison in Table I.
Results of a comparison of optimal paths, for both OM and FW functionals, with the mode of distributions from Monte Carlo simulations for different (ɛ, σ)-pairs and (α, A) = (0.15, 0.7). Here, N is the number of tipping samples used to generate the time-evolved distribution; MAPE [defined by (14)] is the mean absolute percentage error between the mode of the distribution and the optimal path for t ∈ [− 0.5, tm], where tm ∈ [0, 0.3] for these parameter sets. The phase space plots showing the optimal paths overlaid on the distributions from Monte Carlo simulations can be found in Fig. 11 in Appendix D.
. | σ . | N . | MAPE (OM) . | MAPE (FW) . |
---|---|---|---|---|
ɛ = 0.25 | 0.2 | 49 002 | 5.0% | 7.0% |
0.25 | 46 010 | 5.6% | 8.9% | |
0.3 | 21 183 | 6.3% | 12.6% | |
ɛ = 0.4 | 0.2 | 10 662 | 6.0% | 7.3% |
0.25 | 24 882 | 6.8% | 8.5% | |
0.3 | 48 160 | 8.3% | 12.4% |
. | σ . | N . | MAPE (OM) . | MAPE (FW) . |
---|---|---|---|---|
ɛ = 0.25 | 0.2 | 49 002 | 5.0% | 7.0% |
0.25 | 46 010 | 5.6% | 8.9% | |
0.3 | 21 183 | 6.3% | 12.6% | |
ɛ = 0.4 | 0.2 | 10 662 | 6.0% | 7.3% |
0.25 | 24 882 | 6.8% | 8.5% | |
0.3 | 48 160 | 8.3% | 12.4% |
C. Parameter study
In this section, we use the optimal path method to investigate the intermediate forcing regime I of Fig. 1(a) and explore features of the deterministic problem that influence the tipping phase. Figure 5 demonstrated that transitions from the lower stable solution, , to the unstable solution, , may be insensitive to changes in σ, leading to a well-defined transition path associated with this part of the transition. Here, we explore the transitions between and more systematically for parameters in the nullcline regions R2 and R3, which were introduced in Fig. 3(b) for which the number of nullcline curves corresponds to 2 and 1, respectively.
Figure 7 summarizes results of a parameter study for α fixed at 0.1, σ ∈ [0.1, 0.4], and ɛ = 0.25 and 0.4. A is 0.4 in the left column and 0.7 in the right column, each associated with the different nullcline configurations that are in R2 and R3 of Fig. 3(b). We make the following observations:
For all four deterministic parameter sets, the optimal paths leave near when a nullcline crosses this lower stable orbit, indicating a time when the flow changes direction from downwards to upwards near .
The transition between and is concentrated around the same trajectory and varies little with the change of σ. Moreover, independent of noise intensity, the transitions from to all happen in the gap between the nullclines, which is where the deterministic flow is pointing upwards from .
For ɛ = 0.25, the optimal paths arrive at around the same time for different values of σ. However, for ɛ = 0.4, the times tm differ by quite a bit. This is likely a result of the weaker deterministic flow associated with the larger value of ɛ.
Upon crossing the unstable periodic solution , the optimal paths spread out and follow the deterministic flow to reach the upper stable periodic solution .
Summary of optimal paths obtained using the path integral method for various noise intensities σ ∈ [0.1, 0.4]. Nullclines of the deterministic system (2) are represented by blue curves. Black curves represent the periodic solutions of (2). α is fixed at 0.1. ɛ = 0.25 in the first row and 0.4 in the second row. A = 0.4 in the first column and 0.7 in the second column.
Summary of optimal paths obtained using the path integral method for various noise intensities σ ∈ [0.1, 0.4]. Nullclines of the deterministic system (2) are represented by blue curves. Black curves represent the periodic solutions of (2). α is fixed at 0.1. ɛ = 0.25 in the first row and 0.4 in the second row. A = 0.4 in the first column and 0.7 in the second column.
We propose, based on the explorations summarized by Fig. 7, that a good deterministic predictor of the tipping phase tl is the point where the flow near changes direction from negative to positive. This point is associated with an appropriate nullcline crossing the lower stable orbit . Moreover, for the flow geometries of regions R2 and R3, there is a gap between the nullclines; we conjecture that this gap serves as a passageway through which the most likely path squeezes. To further explore this claim, we focus on parameter sets in R3 where there is a single nullcline, i.e., as in the cases of the right column in Fig. 7. We compute the fraction of the transition time from to , which lies above the nullcline (“outside the gap”), as a function of α, which controls the width of the gap; see Fig. 8. For α small (i.e., α near −0.135, the boundary between R2 and R3), the gap becomes very narrow, yet more than 80% of the transition from to occurs “inside the gap” below the nullcline, indicating the flow is upward. For α > −0.095 and below the saddle-node bifurcation, the entire transition from to is in this gap.
Δt/(tm − tl) as a function of α, where Δt is the time spent in regions of downward flow, denoted by green portions of optimal paths. The insets show these optimal paths plotted, with the nullclines and the periodic solutions, for α = −0.135, − 0.1, and 0.06. The rest of the parameters is fixed with (A, ɛ, σ) = (0.5, 0.25, 0.4).
Δt/(tm − tl) as a function of α, where Δt is the time spent in regions of downward flow, denoted by green portions of optimal paths. The insets show these optimal paths plotted, with the nullclines and the periodic solutions, for α = −0.135, − 0.1, and 0.06. The rest of the parameters is fixed with (A, ɛ, σ) = (0.5, 0.25, 0.4).
IV. DISCUSSION
In this paper, we explored a notion of a “tipping phase” for a periodically forced Langevin equation (1) and extended the deterministic conditions that set it from the adiabatic to the nonadiabatic regime. Previous work in the adiabatic regime shows that the most likely transition occurs when the potential barrier height achieves its minimum,8 which for (2) is also the time when the separation between stable and unstable orbits is a minimum. Outside of the adiabatic regime, the two orbits exhibit roughly constant separation over time and the potential may lose its double well structure over an interval of time.
As a tool in our study, we used the path integral formulation to compute the optimal paths between two stable solutions. Specifically, we used the Onsager–Machlup functional whose minimizers we defined as optimal transition paths. The OM functional is the sum of the Freidlin–Wentzell rate functional and a second term that scales with σ2/ɛ, the latter of which controls the ratio of the noise to drift strength. We argued, and showed, that this second term in the OM functional can be significant in regime I in Fig. 1(a) when and σ ∈ (0, 1), and in particular, when the timescale ratio ɛ and noise intensity σ are comparable.
Our parameter study of Sec. III C suggests that the time when the flow direction changes sign, from negative to positive, near the starting lower stable solution is a robust estimator for the tipping phase, which we defined to be tl, the time when the path leaves . In some sense, this might be seen as an extension of the adiabatic regime results in which the transition happens at minimal barrier height. Specifically, in regions R2 and R3 where there are less than three nullclines, tipping starts when the double well structure first begins to break down, and optimal paths squeeze through regions where there is no barrier. More importantly, though, while in the adiabatic regime, tipping is nearly instantaneous (on the time scale of the forcing period) at the time point when the minimal potential barrier height is reached, in our case of an intermediate forcing regime, tipping takes much longer, making it necessary to define a tipping time more carefully. Our work suggests this estimator, tl, and shows that long noisy paths attempt to make the transition by squeezing and maneuvering between portions of the nullclines.
A. Future directions
We conclude with some future avenues of research that come out of this work and a discussion of some of the limitations of our approach that suggests open problems.
1. Bifurcations in the Hamiltonian system
The Hamiltonian system (9) admits several periodic solutions (xH(t), ΨH(t)) and has a rich bifurcation structure in σ and ɛ. For example, when σ = 0 and for typical parameters in our studies, there exist five periodic solutions of (9); three of these correspond to the solutions of (2) and are given by , while the other two are periodic solutions for which ΨH ≠ 0 (e.g., at A = 0, they satisfy ). We conducted a preliminary bifurcation analysis of (9) using AUTO-07P40 and found that qualitative differences in the tipping trajectories in Figs. 7(d)–7(f) roughly align with saddle-node bifurcations of the Hamiltonian dynamics under changes of σ and ɛ. It would be interesting to explore whether the bifurcations in the Hamiltonian system can serve as a method for partitioning a (σ, ɛ) phase diagram into qualitatively different tipping behaviors. Moreover, near a deterministic bifurcation point, i.e., when periodic orbits collide, the deterministic dynamics enter a new scaling regime in which both noise and deterministic effects are comparable. In fact, near such a bifurcation point, the value of the integrand of IOM2 may change sign. Consequently, we do not expect from our analysis that the intersection points of the nullclines with will serve as a strong deterministic predictor for when tipping events will depart from .
2. Higher dimensions
Our conclusions regarding the way the tipping phase aligns with a nullcline crossing are inherently tied to the low dimensionality of the problem we considered. Specifically, we considered a one degree of freedom, periodically forced system, for which an (extended) 1 + 1 dimensional phase plane analysis allowed us to gain insights. While our approach may not be limited to 1D systems, since we could base an investigation on the OM functional in higher dimensions, the natural analog of the nullcline is unclear. Is there a deterministic tipping phase indicator in some higher dimensional situations?
3. Limitations of the path integral approach
The benefit of the path integral approach is that it efficiently calculates optimal paths without performing Monte Carlo simulations. However, with this approach, we lose higher order information that is encoded in the distribution of tipping events. In contrast, in the Freidlin–Wentzell theory of large deviations, these quantities can be recovered from the so-called quasipotential in the limit of vanishing noise strength21,23 or by asymptotics of the Fokker–Planck equation.19 A natural and open question is how to use the OM functional to obtain similar quantities in an asymptotic regime in which and noise is small (intermediate forcing rate) or ɛ ≪ 1 (adiabatic). This may be important in the context of the question we pose about a tipping phase for the following reason. If the distribution of tipping events is too broad, then there is no well-defined phase. It only gains meaning when the distribution is narrow compared to the period of the forcing. We expect that the distribution will narrow with decreases in σ, as illustrated in the contrasting histograms in Fig. 2, which were computed with different noise intensities.
4. Application
Noise-induced transition has applications in a range of areas, such as eradication of infectious disease,1–4,11 large fluctuations in chemical reactions,41 and climate systems,8 to name but a few. Our investigation arose, originally, in the setting of tipping points in a bistable conceptual model of Arctic sea ice, which has strong seasonal forcing. We wondered whether there is a season of the year when this important component of the climate system is most vulnerable to tipping from its current state of perennial ice cover to an alternative state where the Arctic is, year-round, ice-free.
Our preliminary investigations were based on an Arctic sea ice model due to Eisenman and Wettlaufer.7 Their model, an energy-balance column model of the Arctic, captures the strong seasonal variation in incoming solar and outgoing long-wave radiation over the course of the year and the key feedbacks associated with the ice-albedo and the sea ice thermodynamics. It is a 1D nonautonomous dynamical system, which shows bistability between the current Arctic state and the one where the Arctic is ice-free. We illustrate the noise-induced tipping phenomenon in this setting, using a version of the model due to Eisenman,6 with additive white noise. A stochastic version of this model has also been investigated by Moon and Wettlauffer,42 who consider both additive and multiplicative noise.
The Eisenman–Wettlaufer (EW) model7 has a prominent hysteresis loop under changes in greenhouse gases, akin to Fig. 3(a). Here, we consider the parameters in this model corresponding to the current state of the Arctic, which has perennial ice-cover and is bistable, in this model, with a perennially ice-free state; see the deterministic orbits in Figs. 9(a) and 9(b). Figure 9(c) presents distributions obtained from Monte Carlo simulations of the model with additive white noise. In contrast with our findings of the histograms for (1), summarized by Fig. 2, which showed the most prominent peak in the distribution for tl, we find that the distribution of tm is most peaked for the EW model, with a peak in April, which is, interestingly, well before the summer melting season. In Fig. 9(c), we shift all the tipping samples to cross the seasonal state in the third year and plot their mean (magenta curve). We note that, on average, the tipping events take many years to complete the transition, even with relatively strong noise intensity. Moreover, the noisy trajectories often stay in a neighborhood of the unstable seasonally ice-free state for almost a year or more before leaving for the perennially ice-free state, making it challenging to identify “a tipping phase,” despite the peaked distributions. These simulations appear to be in a fast forcing parameter regime where the intrinsic relaxation dynamics are slow compared to the yearly exogenous forcing time scale. We also note that the model has a natural nonsmooth limit, leading to a discontinuity boundary at E = 0, due to the transition from ice-covered to ice-free dynamics.43 To further investigate its tipping events, considering the model switches that take place, with the seasonal effects, along the unstable seasonally ice-free state may be important.
Summary of noise-induced tipping events for the Arctic sea ice model given by Eisenman.6 (a) Vertical scale is surface enthalpy density, proportional to ocean mixed-layer temperature when positive (red) and ice thickness (blue) when negative. Red (blue) curves are the stable perennially ice-free (ice-covered) state; green curve is the unstable seasonally ice-free state. Magenta curve is the mean of 832 transition samples. (b) and (c) Percentage histograms of the departure times from neighborhoods of the perennially ice-covered state (tl) in blue and the seasonal state (tm) in green and the arrival times to a neighborhood of the perennially ice-free state (tu) in red; neighborhood size is chosen so that the distribution is invariant when the size is halved. Model parameters are taken from Table 1 in the paper by Eisenman6 with and σ = 0.17. Each simulation runs for 1000 years. (E = 0.4% in the Monte Carlo convergence test.)
Summary of noise-induced tipping events for the Arctic sea ice model given by Eisenman.6 (a) Vertical scale is surface enthalpy density, proportional to ocean mixed-layer temperature when positive (red) and ice thickness (blue) when negative. Red (blue) curves are the stable perennially ice-free (ice-covered) state; green curve is the unstable seasonally ice-free state. Magenta curve is the mean of 832 transition samples. (b) and (c) Percentage histograms of the departure times from neighborhoods of the perennially ice-covered state (tl) in blue and the seasonal state (tm) in green and the arrival times to a neighborhood of the perennially ice-free state (tu) in red; neighborhood size is chosen so that the distribution is invariant when the size is halved. Model parameters are taken from Table 1 in the paper by Eisenman6 with and σ = 0.17. Each simulation runs for 1000 years. (E = 0.4% in the Monte Carlo convergence test.)
ACKNOWLEDGMENTS
We thank Julie Leifeld for some early discussions. A.V. thanks Björn Sandstede for his guidance in AUTO. We thank Paul Ritchie and Jan Sieber for introducing us to the path integral approach. M.S., Y.C., and A.V. are partially funded by the AMS Mathematics Research Communities [National Science Foundation (NSF) Grant No. 1321794]. The work of M.S. and Y.C. is funded in part by NSF Grant No. DMS-1517416. The work of A.V. and J.A.G. was supported in part by NSF-RTG Grant No. DMS-1148284, and A.V. is currently supported by the Mathematical Biosciences Institute and the NSF under Grant No. DMS-1440386.
APPENDIX A: DERIVATION OF THE ONSAGER–MACHLUP FUNCTIONAL
Here, we provide a formal derivation of the Onsager–Machlup functional following the presentation in Chaichian and Demichev.25 Let t0, tf ∈ ℝ with tf > t0 ≥ 0 and x0 ∈ ℝ. For t ∈ [t0, tf] and g(x, t) a smooth function, consider the two continuous stochastic processes y(t), x(t) ∈ ℝ that satisfy the following stochastic differential equations interpreted in the Ito sense:
where σ > 0, y(t0) = 0, x(t0) = x0, and W(t) is a Wiener process.
For a uniform discretization {t0, t1, …, tN} of [t0, tf] of width Δt and realizations x(t) and y(t) of Eqs. (A1) and (A2), we define vector valued random variables x = {x0, x1, …xN}, y = {y0, y1, …, yN} by xi = x(ti) and yi = y(ti). It follows from the Kolmogorov extension theorem44 that the probability density ζy for y is given by
In particular, if we let Δyi > 0 be a sequence of spatial widths and a continuous function on [t0, tf] satisfying , then the probability realizations of Eq. (A1) pass through a sequence of “gates” is
The Wiener measure μy is defined as the unique measure on continuous functions satisfying (A4) for all temporal and spatial discretizations, i.e., all cylindrical sets.44 The space of continuous functions is infinite dimensional, and, thus, a Lebesgue measure cannot be defined on it. However, as is customary in the physics literature, it is useful to think of the Wiener measure as heuristically having the following density with respect to Lebesgue measure:
where Z is the appropriate normalization factor.25 Note that this heuristic notation results from considering the formal limiting density of Eq. (A3) as Δt → 0.
We can formally use Eqs. (A1) and (A2) to make a change of variables to construct the density ζx for x. In the integral form, it follows that
Using the composite trapezoidal rule, it follows that
Consequently, to the lowest order, the Jacobian matrix for this transformation is a lower triangular matrix with diagonal entries and thus to the lowest order
Since , it follows, again to the lowest order, that
Finally, we obtain the following formal expression for the change of measure as Δt → 0:
Again, the representation of dμx as a Lebesgue measure is heuristic. However, using Girsanov’s theorem to rigorously compute the Radon–Nikodym derivative , this result can be made precise for autonomous and nonautonomous drift.45,46
APPENDIX B: CONVERGENCE OF OM FUNCTIONAL TO FW FUNCTIONAL IN THE LIMIT σ → 0
APPENDIX C: INDEPENDENCE OF INITIAL AND FINAL TIMES T0 AND TF
To show that the optimal path is not sensitive to the choice of t0 and tf, we vary t0 (tf) by taking 16 uniformly spaced values in [0, 1] ([4, 5]) (Fig. 10). Initial guesses are in the form of (12) with the same intermediate time tj = 2. In blow up box A (B), curves with different colors indicate that the paths start (end) at different positions on (). In the second period, all paths converge to the same curve from to . After arriving at , the paths start to separate shortly before the fourth period since we impose different final positions. Thus, given a large enough interval of time between tl and tu, optimal paths that start and end at different positions transition from to at the same forcing phase.
Illustration of the numerical experiment to test that the optimal path is independent of t0 and tf. We vary t0 and tf in the first and third periods. (ɛ, α, A, σ) = (0.25, 0.1, 0.4, 0.5).
Illustration of the numerical experiment to test that the optimal path is independent of t0 and tf. We vary t0 and tf in the first and third periods. (ɛ, α, A, σ) = (0.25, 0.1, 0.4, 0.5).
APPENDIX D: CONVERGENCE TEST OF MONTE CARLO SIMULATIONS AND MORE COMPARISON RESULTS
Here, we describe the procedure of Monte Carlo simulations. We use the Euler–Maruyama method48 on the interval [t0, tf] with initial data to generate a sequence of samples Xj(t) for (1), each of which runs for K time steps (Fig. 11). We collect N samples that meet the following criteria for transitioning from to . Let denote the first time a sample path crosses . We define the “tipping events” to be {Xi} that satisfy {Xi} = {Xj:τj < tf}. We then construct the distribution of the tipping events using M bins. The convergence test of the Monte Carlo simulations used to obtain distributions of the tipping events shown in Sec. III involves the following steps:
At each time step, t(k) (k = 1, …, K), we divide the samples {Xi(t(k)): i = 1, …, N} into M equal width bins in the x direction. We then record the fraction of the samples that fall into the mth bin. We pick M in the histogram based on the Freedman Diaconis rule,49 and K so that Euler–Maruyama method converges.
We double the number of samples to 2N and repeat Step 1 to get .
We then compute the mean squared error between and (m = 1, …, M, k = 1, …, K), given by .
If E < 10−5, we say that the Monte Carlo simulation has converged. If it has not converged, we double the number of sample paths and repeat the procedure.
Comparison between the optimal path results, represented by the red curves, and the Monte Carlo results. Parameters are taken as follows: ɛ = 0.25 in the left column and ɛ = 0.4 in the right one; σ = 0.2, 0.25, and 0.3 from the top to the bottom rows. α and A are fixed at 0.15 and 0.7, respectively. These are illustrations of the result summarized in Table I. Our criterion for the convergence test of the Monte Carlo simulation is E < 10−5, where E is defined in (13). From top to bottom, E = 1.2 × 10−6, 1.2 × 10−6, and 7.5 × 10−6 in the left column and E = 8.7 × 10−6, 1.4 × 10−6, and 1.7 × 10−6 in the right column.
Comparison between the optimal path results, represented by the red curves, and the Monte Carlo results. Parameters are taken as follows: ɛ = 0.25 in the left column and ɛ = 0.4 in the right one; σ = 0.2, 0.25, and 0.3 from the top to the bottom rows. α and A are fixed at 0.15 and 0.7, respectively. These are illustrations of the result summarized in Table I. Our criterion for the convergence test of the Monte Carlo simulation is E < 10−5, where E is defined in (13). From top to bottom, E = 1.2 × 10−6, 1.2 × 10−6, and 7.5 × 10−6 in the left column and E = 8.7 × 10−6, 1.4 × 10−6, and 1.7 × 10−6 in the right column.