Abstract
We consider the distribution of first passage time events in the presence of non-ergodic modes that drive otherwise ergodic dynamics on a potential landscape. We find that in the limit of slow and large enough fluctuations the distribution of first passage time events, f(t), exhibits heavy tails dominated by a power law with exponent f(t) ~ t−2, and corrections that depend on the strength and the nature of fluctuations. We support our theoretical findings through direct numerical simulations in illustrative examples.
Introduction.—
Complex dynamics are ubiquitous in the natural world. Despite their intrinsic irregularity and unpredictability, they can nonetheless exhibit coherent and universal emergent properties. Of particular importance in the study of complex systems is the understanding of the time it takes for rare events to occur [1–3]. Notable examples include natural disasters [4] or the spreading of a virus [5]. In fact, first passage times are central to many fields within physics and beyond, with important examples stemming from chemistry, biology and finance (see, e.g., [6–12] and references therein). Biology in particular is ripe with examples where time is of the essence [13], such as fertilization [14], intracellular events [15–21], search processes [22–24], neural activity [25, 26] or population dynamics [27].
We here consider the estimation of first passage time distributions (FPTDs) from finite-time observations in an experimental context. In particular, we are interested in systems with intrinsic time scales comparable to the observation time, for which weak ergodicity breaking becomes evident [28, 29]. Such dynamics can be found for instance in glassy systems [30–33], where the time scales of equilibration are so long that one can decompose the dynamics into a stationary component and an “aging” component that breaks time-translation invariance.
Our main inspiration comes from the less traditional branch of the physics of animal behavior [34, 35]. Remarkably, recent advances in machine vision (see, e.g., [36–39]) have resulted in an explosion of high spatiotemporal resolution behavioral data. Analysis of fine-scale posture movements shows that much like the runand-tumble behavior of bacteria [40], more complex organisms also exhibit stereotyped behaviors, albeit with a more intricate structure [41–48]. The notion of stereotypy in behavior inherently stems from the time scale separation between variations on what is defined as a behavioral state, and the transitions between behavioral states, much like a particle hopping between wells in a potential landscape. For example, while foraging for food the nematode worm C. elegans transitions between coarsegrained “runs” and “pirouettes”, which are stereotyped sequences of finer scale movements [48, 49]. However, unlike the particle hopping among potential wells which has a characteristic exponential distribution of transition times, the time spent in a given behavior can be heavy-tailed (see, e.g. Fig.4E of [48] or Fig.3 of [50]). We here hypothesize that such heavy-tailed distributions reflect the slow continuous modulation of behavior on longer time scales, resulting from environmental factors or fluctuating internal states driven by neuromodulation, such as hunger, stress or arousal (see, e.g., [51–53]). Indeed, it has been shown that C. elegans continuously modulates its rate of reorientation events to explore larger and larger arenas in search of food [54]. In order to truly capture the multiscale nature of behavior, we therefore need to account for the fact that it can be modulated on time scales comparable to the observation time.
In this Letter, we introduce a general model of behavior in which the pose dynamics evolves in potential landscapes that fluctuate over time. We then study how these dynamics impact the estimation of the distribution of times spent in a given behavior. In the first section, we introduce our phenomenological description of the behavioral dynamics, decomposing it into ergodic dynamics on a potential landscape and the non-ergodic modulation of the landscape. We then derive a general result for the distribution of first passage times, and illustrate it through direct numerical simulations in three example systems.
Slowly driven ergodic dynamics.—
Given a set of observations of animal locomotion (e.g. from video imaging), we consider that the dynamics can be decomposed into ergodic and non-ergodic components. The former are the state-space variables that mix sufficiently well and define the potential wells that correspond to the stereotyped behaviors; the latter non-ergodic components evolve on time scales comparable to the observation time and slowly modulate the potential landscape. The full dynamics is thus given by
| (1) |
where represents the ergodic components, represents the non-ergodic degrees of freedoms, F and G are nonlinear, possibly noisy, functions, and τλ is assumed to be of the order of the measurement time Texp, τλ = 𝒪(Texp), such that the dynamics do not mix. Given the time scale separation between the dynamics of and , we assume that the dynamics of is well approximated by quasi-stationary Fokker-Planck dynamics , where ℒ represents the Fokker-Planck operator. Since we are primarily interested in the long time scale behavior of the system, we consider a projection of the dynamics onto the slowest mode of ℒ, yielding a generalized Langevin equation [55, 56] with history-dependent friction and fluctuations. Assuming that we can sample the system on a time scale longer than the noise correlation time, we obtain an effective overdamped description:
| (2) |
where Tx captures the effective temperature, ηx is Gaussian white noise, and λ is a slow control parameter that modulates the effective potential landscape on slow time scales. Similarly, we consider that λ also obeys an effective overdamped Langevin equation,
| (3) |
where V is assumed to be uncoupled from the dynamics of x for simplicity, Tλ captures the degree of fluctuations in λ and ηλ is Gaussian white noise.
First passage time distributions.—
We are primarily interested in studying the time spent in a given behavioral state. Within the context of the Langevin dynamics of Eq.2, this is given by the first passage time to reach an energy barrier xf from the bottom of the potential x0, defined as,
| (4) |
Despite the general interest in this concept, finding analytical expressions for the density of first passage time events is generally a formidable task [57]. Remarkably few closed-form expressions for the FPTD are known, with most results concerning only the mean first passage time (MFPT) which is more tractable (see, e.g., [1, 6, 9]). However, the MFPT provides only limited information, especially when multiple time scales are involved [15]. Here, we are interested in studying the behavior of the full first passage time distribution, with particular focus on its long time behavior in the presence of weakly non-ergodic dynamics, Eqs.2 and 3.
As previously discussed, the measurement time Texp essentially separates ergodic from non-ergodic dynamics. In addition, it also sets a lower bound on the slowest observed hopping rates , such that when τλ = 𝒪(Texp) we can make an adiabatic approximation and assume that transition events occur within a nearly static potential. For a given hopping rate ω, the first passage time distribution is given by
where is the dominating slow kinetic transition rate which implicitly depends on the dynamics of λ. When we allow λ to fluctuate slowly, the distribution of first passage times f(t) is given by the expectation value of f(t, ω) over the distribution of ω, p(ω), weighted by the effective number of transition observed within Texp, which is proportional to ω. Marginalizing over ω we get
| (5) |
While the barrier height is going to depend on the dynamics of a slow control parameter λ, the tail of the distribution is going to be dominated by instances in which the barrier height is the largest, motivating the use of Kramers approximation (see, e.g., [2]),
| (6) |
where ΔU(λ) = U(xf, λ)−U(x0, λ) and ω0 is a constant. For multiple realization of Eq.3 with different initial conditions, the distribution of λ is given by the Boltzmann weight [58],
| (7) |
Leveraging Eqs.S1,6,7 we can obtain an asymptotic approximation of the FPTD in the large t limit (see Supplemental Material),
| (8) |
where ΔU−1(·) represents the inverse function of the potential difference defined by Eq. 6 and we have kept only the dominant order of the asymptotic approximation (see Supplemental Material). For very general conditions on V (λ) and U(x, λ), we thus get f(t) ~ t−2 for t → ∞ and Tλ ≫ 1. In the following section we will demonstrate the validity of this result in three illustrative examples.
Illustrative examples: Slowly-driven harmonic oscillator.—
Consider that x evolves in a harmonic potential, U(x, s) = (x − sxf)2, that is driven by a slow parameter s that fluctuates within V (s) = s2/2, pushing U(x, s) closer or further from xf in a time scale τs, Fig.1(a). The equations of motion are given by a set of Ito stochastic differential equation, corresponding to coupled Ornstein-Uhlenbeck processes,
| (9) |
where Tx and Ts captures the degree of fluctuations, dWt is a Wiener Gaussian white noise process. We are interested in the density of first passage time events from the minimum of the potential x0 = s to xf = 1, for which it is challenging to find a closed form analytical expression, even when [57]. In the Supplemental Material, we derive the FPTD in Laplace space [59] and leverage it to estimate the FPTD through numerical inversion [60] for varying values of τs (as in Ref. [61]), see Fig.S2. We find that when s fluctuates fast enough, τs → 0, we can average out s and get the simpler dynamics . In this case, the FPTD is well approximated by f(t) ≈ f(t, 〈ω〉) = 〈ω〉e−〈ω〉t, where 〈ω〉 is the average hopping rate which is set by 〈s〉. Even when τs > 0 but short, it is possible to obtain a self-consistent Markovian dynamics for x(t) (see e.g., [1]). In this case, the distribution of first passage times is still dominantly exponential, but with a corrected first passage time which depends on the ratio of temperatures Ts/Tx and the slow time scale τs. However, as we have shown in the previous section τs is large enough, τs ~ Texp, the distribution of first passage times becomes heavy-tailed. In this limit, we can leverage Eq.8 to derive an asymptotic approximation to the distribution of first passage times. The tail of the distribution will be dominated by low ω values, which correspond to |s| ≫ 1. In this limit, the barrier height primarily behaves as ΔU(s) = s2/2 + 𝒪(s). In addition, since V (s) = s2/2, we see that V (ΔU−1(x)) = x and Eq.8 yields (see Supplemental Material),
| (10) |
which matches what we obtain from direct numerical simulations of Eq.9, Figs.1(b),S2,S3(a).
FIG. 1.
Heavy-tailed first passage time distributions for a slowly-driven overdamped harmonic oscillator. (a) We simulate the dynamics of a particle in a harmonic oscillator while slowly driving the potential landscape, and estimate the distribution of times it takes to reach xf. The gray line represents the minimum of potential, x0 = s, and the color scheme different values of s. (b) FPTDs obtained from direct numerical simulations of Eq.9 for different values of the temperature Ts that controls the level of fluctuations for the parameter driving the slow variations of the potential landscape. As predicted, the tail of the distribution behaves as a power law with an exponent f(t) ~ t−2−α, with . The color scheme represents different ratios of temperatures, and the black dashed line the Ts →∞ limit.
Illustrative examples: Slowly-driven double-well potential.—
We now consider a symmetric double-well potential in which the barrier height is slowly modulated according to an Ornstein-Uhlenbeck process, Fig.2(a),
| (11) |
where all the parameters are the same as in Eq.9 with an extra μs that represents the expectation value of s, which we set as μs = 1. In this case, we have a quartic potential for x, U(x, s) = s2(x2 − 1)2, which yields ΔU(s) = s2. Since V (s) = s2/2, we see that V (ΔU−1(x)) = x/2 and Eq.8 yields (see Supplemental Material),
| (12) |
matching what we find through direct numerical simulations of Eq.11, Figs.2(b),S3(b)
FIG. 2.
Heavy-tailed first passage time distribution of a slowly-driven double-well potential. (a) Schematic of the variation in the double-well potential with s (colored from blue to red; the black line represents s = μs). (b) FPTDs from direct numerical simulations of Eq.11 for different values of Ts. As expected, the tail of the distribution behaves as a power law f(t) ~ t−2−α, where (colored line). The black dashed line represents the Ts →∞ limit.
Illustrative examples: Slowly-driven rugged parabolic potential.—
Finally, we consider a rugged parabolic potential as a simple model of the rough energy landscapes found across complex systems, from glasses to proteins (see, e.g., [19, 20, 62]). We construct a rugged landscape by superimposing a sinusoidal perturbation onto a harmonic potential [63], U(x, s) = U0(x, s) + U1(x), where U0(x, s) = (x − s)2/2 and U1(x) = −cos(2πkx)/(kπ). The corresponding dynamics are given by,
| (13) |
where k sets the number of smaller barriers between the global minimum of the potential and xf = 1. We set k = 10 resulting in a rugged potential as illustrated in Fig.3(a). In this case, since U(x, s) is not as simple as before, it is more challenging to derive the correction terms to the power law. However, it has been shown [63] that by spatial averaging of U1(x) = −cos(2πkx)/(kπ) over one period, the resulting hopping rate is simply corrected by a constant prefactor , where I0 is the modified Bessel function and ω0 is the hopping rate in the absence of the sinusoidal perturbation (from U0(x, s) = (x − s)2/2). As such, we expect the asymptotic behavior of f(t) to be the same as for the slowly driven harmonic potential, Eq.10. Indeed, this is what we observe in Figs.3(b),S3(c).
FIG. 3.
Heavy-tailed first passage time distribution in slowly driven rugged parabolic potential. (a) We estimate the first passage time to reach xf from the global minimum of a rugged parabolic potential. (b) FPTDs from direct numerical simulations of Eq.13 for different values of Ts. As expected, the tail of the distribution behaves as a power law f(t) ~ t−2−α (colored lines) with . The black dashed line corresponds to the Ts →∞ limit.
Discussion.—
Inspired by quantitative analysis of animal behavior, we here examined how the existence of slow non-ergodic modes impacts the statistics collected experimentally, focusing on the distribution of first passage time events. Our results show the emergence of heavy-tailed distributions. In particular, we find that the distribution asymptotes to a power law with an exponent f(t) ~ t−2 in the limit of large fluctuations, regardless of the details of the dynamics. As remarked in the Introduction, our results have important implications to a wide variety of fields, and we here discuss some of these in detail.
In the context of animal behavior, heavy-tailed first passage times with an exponent f(t) ≈ t−2 have been found extensively across multiple species, from bacteria [64], termites [65] and rats [66] to marine animals [67, 68], humans [69] and even fossil records [70]. In the context of search behaviors (e.g., when foraging for food), such observations have led researchers to hypothesize that Lévyflights (power law distributed run lengths) are efficient search strategies and thus evolutionarily favorable [71–75]. However, we here show that such fat tails may emerge when the animal is continuously adapting its behavior (slowly modulating the potential landscape), even in the absence of external drives. We therefore predict that disrupting the internal mechanisms for slow modulation of behavior (e.g. neuromodulatory pathways) should result in distribution of first passage times that have exponential tails.
Power laws have been observed in a wide variety of systems, from solar flares [76, 77] to the brain [78] and different hypotheses have been put forward to explain their emergence (for a review see e.g. [79]). Among these, work inspired by phase transitions in statistical mechanics associates power laws to “criticality”, mostly due to the fact that models inferred from the data appear to require fine-tuning of the model parameters to a special regime between two qualitatively different “phases” (see, e.g., [80]). However, as we have shown here, power laws can emerge without fine tuning and far from “criticality”. Indeed, slow modes that evolve on time scales comparable to the observation time are challenging to infer from data, and can give rise to best-fit models that appear “critical”. While some of the arguments we have put forward have also been proposed in other contexts [22, 81–83], we here place them into the framework of out-ofequilibrium statistical mechanics, explicitly connecting the long time scale emergent behavior with the underlying effective fluctuations. In addition, unlike other approaches [82, 84], our framework does not require explicit external drives, but simply collective modes that evolve in a weakly non-ergodic fashion.
Our starting point is an effective description of the long time scale dynamics, and further work will be required to fully bridge between microscopic dynamics, and the emergent long time behavior of the first passage time distribution that we uncovered. For example, we find that for intermediate values of 1 ≪ Tλ ≪ Texp the FPTD behaves as a truncated power law with an effective exponent that is slightly smaller that −2 (see Supplemental Material), which goes beyond arguments presented here. What are the minimum τλ and Tλ for power laws to be measurable, and how do simple exponentials (Tλ ≪ Texp) transition to power law behavior? These are important questions if one hopes to test our predictions in an experimental context (using for example a set-up akin to the ones used to test stochastic resonance [85, 86]). Additionally, we note that when Tλ ≫ Texp, the distribution of initial conditions determines the emergent behavior, see Fig.S4. Inspired by experiments in animal behavior, which are typically done with multiple animals, we here assume that the initial condition for the slow mode is sampled according to its Boltzmann distribution , reflecting the variability across individuals. In this case, the emergent behavior we have derived holds true from τλ = Texp to τλ → ∞. However, if the variability across experiments is smaller than that of the Boltzmann distribution, the τλ → ∞ limit will differ from the behavior at τλ ~ Texp. Indeed, if the variance of the initial distribution of λ is smaller than that of the Boltzmann distribution, the temperature Tλ in our derivation should be changed to a new effective temperature reflecting the lower variance of the initial conditions. Making this transformation we still get a power law distribution of first passage times, but with a modified exponent that reflects the lower variance (see Supplemental Material).
To conclude, we have considered the effect of slow non-ergodic modulations and theoretically captured their effects on the distribution of first passage times, a result that we believe is widely relevant to a range of natural systems.
Supplementary Material
Acknowledgments
We thank Adrian van Kan, Stéphan Fauve, Federica Ferretti, Tosif Ahamed and Arghyadip Mukherjee for comments. This work was partially supported by the LabEx ENS-ICFP: ANR-10-LABX-0010/ANR-10-IDEX-0001-02 PSL* and by the NIH Grant 1RF1NS128865-01. AC also acknowledges useful discussions at the Aspen Center for Physics, which is supported by National Science Foundation Grant PHY-1607611.
References
- [1].Hänggi P., Chemical Physics 180, 157 (1994). [Google Scholar]
- [2].Van Kampen N. G., Stochastic processes in physics and chemistry (North-Holland, Amsterdam, 1981). [Google Scholar]
- [3].Redner S., A Guide to First-Passage Processes (Cambridge University Press, 2001). [Google Scholar]
- [4].Wijesundera I., Halgamuge M. N., Nanayakkara T., and Nirmalathas T., Natural Disasters, When Will They Reach Me?, Springer Natural Hazards (Springer Singapore, Singapore, 2016). [Google Scholar]
- [5].Verma A. K., Bhatnagar A., Mitra D., and Pandit R., Phys. Rev. Research 2, 033239 (2020). [Google Scholar]
- [6].Szabo A., Schulten K., and Schulten Z., The Journal of Chemical Physics 72, 4350 (1980). [Google Scholar]
- [7].Condamin S., Tejedor V., Voituriez R., Bénichou O., and Klafter J., Proceedings of the National Academy of Sciences 105, 10.1073/pnas.0712158105 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Bénichou O. and Voituriez R., Physics Reports 539, 225 (2014). [Google Scholar]
- [9].Bénichou O., Guérin T., and Voituriez R., Journal of Physics A: Mathematical and Theoretical 48, 163001 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Ghusinga K. R., Dennehy J. J., and Singh A., Proceedings of the National Academy of Sciences 114, 10.1073/pnas.1609012114 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [11].Chicheportiche R. and Bouchaud J.-P., Some applications of first-passage ideas to finance, in First-Passage Phenomena and Their Applications (World Scientific, 2014) Chap. 1, pp. 447–476. [Google Scholar]
- [12].de Wit X. M., van Kan A., and Alexakis A., Journal of Fluid Mechanics 939, R2 (2022). [Google Scholar]
- [13].Polizzi N. F., Therien M. J., and Beratan D. N., Israel Journal of Chemistry 56, 816 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Meerson B. and Redner S., Phys. Rev. Lett. 114, 198101 (2015). [DOI] [PubMed] [Google Scholar]
- [15].Godec A. and Metzler R., Phys. Rev. X 6, 041037 (2016). [Google Scholar]
- [16].Eliazar I., Koren T., and Klafter J., Journal of Physics: Condensed Matter 19, 065140 (2007). [Google Scholar]
- [17].Szymanski J. and Weiss M., Phys. Rev. Lett. 103, 038102 (2009). [DOI] [PubMed] [Google Scholar]
- [18].Zhang Y. and Dudko O. K., Annual Review of Biophysics 45, 117 (2016). [DOI] [PubMed] [Google Scholar]
- [19].Hyeon C. and Thirumalai D., Proceedings of the National Academy of Sciences 100, 10.1073/pnas.1833310100 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Nevo R., Brumfeld V., Kapon R., Hinterdorfer P., and Reich Z., EMBO reports 6, 482 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [21].Biswas K., Shreshtha M., Surendran A., and Ghosh A., The European Physical Journal E 42, 24 (2019). [DOI] [PubMed] [Google Scholar]
- [22].Tu Y. and Grinstein G., Phys. Rev. Lett. 94, 208101 (2005). [DOI] [PubMed] [Google Scholar]
- [23].Dev S. and Chatterjee S., Phys. Rev. E 91, 042714 (2015). [DOI] [PubMed] [Google Scholar]
- [24].Benichou O., Coppey M., Moreau M., Suet P.-H., and Voituriez R., Physical Review Letters 94, 198101 (2005). [DOI] [PubMed] [Google Scholar]
- [25].Taillefumier T. and Magnasco M. O., Proceedings of the National Academy of Sciences 110, 10.1073/pnas.1212479110 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [26].Pakdaman K., Tanabe S., and Shimokawa T., Neural Networks 14, 895 (2001). [DOI] [PubMed] [Google Scholar]
- [27].Spanio T., Hidalgo J., and Muñoz M. A., Phys. Rev. E 96, 042301 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Bouchaud J. P., Phys J.. I France 2, 10.1051/jp1:1992238 (1992). [DOI] [Google Scholar]
- [29].Cugliandolo L. F. and Kurchan J., Phys. Rev. Lett. 71, 173 (1993). [DOI] [PubMed] [Google Scholar]
- [30].Binder K. and Young A. P., Reviews of Modern Physics 58, 801 (1986). [Google Scholar]
- [31].Jackle J., Reports on Progress in Physics 49, 171 (1986). [Google Scholar]
- [32].Debenedetti P. G. and Stillinger F. H., Nature 410, 259 (2001). [DOI] [PubMed] [Google Scholar]
- [33].Young A. P., Spin Glasses and Random Fields, Directions in condensed matter physics (World Scientific, 1997). [Google Scholar]
- [34].Brown A. X. and de Bivort B., Nature Physics 14, 653 (2018). [Google Scholar]
- [35].Berman G. J., BMC Biol. 16, 10.1186/s12915-018-0494-7 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [36].Pereira T. D., Shaevitz J. W., and Murthy M., Nature Neuroscience 23, 1537 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Mathis M. W. and Mathis A., Current Opinion in Neurobiology 60, 1 (2020). [DOI] [PubMed] [Google Scholar]
- [38].Yemini E., Jucikas T., Grundy L. J., Brown A. E., and Schafer W. R., Nat Methods 10, 10.1038/nmeth.2560 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Hebert L., Ahamed T., Costa A. C., O’Shaughnessy L., and Stephens G. J., PLoS Computational Biology 17, e1008914 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Berg H. C., coli in motion E., Biological and medical physics series (Springer, New York, 2004). [Google Scholar]
- [41].Berman G. J., Choi D. M., Bialek W., and Shaevitz J. W., J. Royal Soc. Interface 11, 1 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Stephens G. J., Johnson-Kerner B., Bialek W., and Ryu W. S., PLoS Comput. Biol. 4, e1000028 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Costa A. C., Ahamed T., and Stephens G. J., Proceedings of the National Academy of Sciences of the United States of America 116, 10.1073/pnas.1813476116 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [44].Ahamed T., Costa A. C., and Stephens G. J., Nature Physics 17, 275 (2021). [Google Scholar]
- [45].Wiltschko A. B., Johnson M. J., Iurilli G., Peterson R. E., Katon J. M., Pashkovski S. L., Abraira V. E., Adams R. P., and Datta S. R., Neuron 88, 1121 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [46].Johnson R. E., Linderman S., Panier T., Wee C. L., Song E., Herrera K. J., Miller A., and Engert F., Current Biology 30, 70 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [47].Alba V., Berman G. J., Bialek W., and Shaevitz J. W., Exploring a strongly non-markovian animal behavior (2020), arXiv:2012.15681 [q-bio.NC].
- [48].Costa A. C., Ahamed T., Jordan D., and Stephens G., Maximally predictive ensemble dynamics from data (2021), arXiv:2105.12811 [physics.bio-ph].
- [49].Pierce-Shimomura J. T., Morse T. M., and Lockery S. R., J. Neurosci. 19, 9557 (1999). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Srivastava N., Clark D. a., and Samuel a. D. T., Journal of Neurophysiology 102, 1172 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [51].Bargmann C. I., BioEssays 34, 458 (2012). [DOI] [PubMed] [Google Scholar]
- [52].Marder E., Neuron 76, 1 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Flavell S. W., Gogolla N., Lovett-Barron M., and Zelikowsky M., Neuron 110, 2545 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [54].Gray J. M., Hill J. J., and Bargmann C. I., Proceedings of the National Academy of Sciences 102, 10.1073/pnas.0409009101 (2005). [DOI] [Google Scholar]
- [55].Mori H., Progress of Theoretical Physics 33 (1965). [Google Scholar]
- [56].Zwanzig R., Journal of Statistical Physics 9, 215 (1973). [Google Scholar]
- [57].Grebenkov D. S., Journal of Physics A: Mathematical and Theoretical 48, 013001 (2014). [Google Scholar]
- [58]. We note that this is only true if τλ is finite. In particular, if Tλ ≫ Texp, then p(ω) is primarily defined by the distribution of initial conditions, a point that we return to in the Discussion.
- [59].Siegert A. J. F., Phys. Rev. 81, 617 (1951). [Google Scholar]
- [60].de Hoog F. R., Knight J. H., and Stokes A. N., SIAM J. Sci. Stat. Comput. 3, 357–366 (1982). [Google Scholar]
- [61].Mullowney P. and Iyengar S., Journal of Computational Neuroscience 24, 179 (2008). [DOI] [PubMed] [Google Scholar]
- [62].Charbonneau P., Kurchan J., Parisi G., Urbani P., and Zamponi F., Nature Communications 5, 3725 (2014). [DOI] [PubMed] [Google Scholar]
- [63].Zwanzig R., Proceedings of the National Academy of Sciences 85, 10.1073/pnas.85.7.2029 (1988). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Korobkova E., Emonet T., Vilar J. M. G., Shimizu T. S., and Cluzel P., Nature 428, 574 (2004). [DOI] [PubMed] [Google Scholar]
- [65].Miramontes O., DeSouza O., Paiva L. R., Marins A., and Orozco S., PLOS ONE 9, 1 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Jung K., Jang H., Kralik J. D., and Jeong J., PLOS Computational Biology 10, 1 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Humphries N. E., Queiroz N., Dyer J. R. M., Pade N. G., Musyl M. K., Schaefer K. M., Fuller D. W., Brunnschweiler J. M., Doyle T. K., Houghton J. D. R., Hays G. C., Jones C. S., Noble L. R., Wearmouth V. J., Southall E. J., and Sims D. W., Nature 465, 1066 (2010). [DOI] [PubMed] [Google Scholar]
- [68].Sims D. W., Southall E. J., Humphries N. E., Hays G. C., Bradshaw C. J. A., Pitchford J. W., James A., Ahmed M. Z., Brierley A. S., Hindell M. A., Morritt D., Musyl M. K., Righton D., Shepard E. L. C., Wearmouth V. J., Wilson R. P., Witt M. J., and Metcalfe J. D., Nature 451, 1098 (2008). [DOI] [PubMed] [Google Scholar]
- [69].Raichlen D. A., Wood B. M., Gordon A. D., Mabulla A. Z. P., Marlowe F. W., and Pontzer H., Proceedings of the National Academy of Sciences 111, 10.1073/pnas.1318616111 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Sims D. W., Reynolds A. M., Humphries N. E., Southall E. J., Wearmouth V. J., Metcalfe B., and Twitchett R. J., Proceedings of the National Academy of Sciences 111, 10.1073/pnas.1405966111 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Viswanathan G. M., Buldyrev S. V., Havlin S., da Luz M. G., Raposo E. P., and Stanley H. E., Nature 401, 911 (1999). [DOI] [PubMed] [Google Scholar]
- [72].Wosniack M. E., Santos M. C., Raposo E. P., Viswanathan G. M., and da Luz M. G. E., Phys. Rev. E 91, 052119 (2015). [DOI] [PubMed] [Google Scholar]
- [73].Wosniack M. E., Santos M. C., Raposo E. P., Viswanathan G. M., and da Luz M. G. E., PLOS Computational Biology 13, 1 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Guinard B. and Korman A., Science Advances 7, eabe8211 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [75].Clementi A., d’Amore F., Giakkoupis G., and Natale E., in Proceedings of the 2021 ACM Symposium on Principles of Distributed Computing, PODC’21 (Association for Computing Machinery, New York, NY, USA, 2021) p. 81–91. [Google Scholar]
- [76].Wheatland M. S., Sturrock P. A., and McTiernan J. M., The Astrophysical Journal 509, 448 (1998). [Google Scholar]
- [77].Boffetta G., Carbone V., Giuliani P., Veltri P., and Vulpiani A., Phys. Rev. Lett. 83, 4662 (1999). [Google Scholar]
- [78].Beggs J. M. and Plenz D., Journal of Neuroscience 23, 10.1523/JNEUROSCI.23-35-11167.2003 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [79].Newman M., Contemporary Physics 46, 323 (2005), 10.1080/00107510500052444. [DOI] [Google Scholar]
- [80].Mora T. and Bialek W., Journal of Statistical Physics 144, 268 (2011). [Google Scholar]
- [81].Touboul J. and Destexhe A., Phys. Rev. E 95, 012413 (2017). [DOI] [PubMed] [Google Scholar]
- [82].Priesemann V. and Shriki O., PLOS Computational Biology 14, 1 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [83].Proekt A., Banavar J. R., Maritan A., and Pfaff D. W., Proceedings of the National Academy of Sciences 109, 10.1073/pnas.1206894109 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [84].Schwab D. J., Nemenman I., and Mehta P., Phys. Rev. Lett. 113, 068102 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Fauve S. and Heslot F., Physics Letters A 97, 5 (1983). [Google Scholar]
- [86].Lanzara E., Mantegna R. N., Spagnolo B., and Zangara R., American Journal of Physics 65, 341 (1997), 10.1119/1.18520. [DOI] [Google Scholar]
- [87]. We note that for the general dynamics of Eqs.(2,3), ω0 may have a λ dependency. However, without loss of generality, we consider that ω0 and ΔU(λ) can be redefined to move the λ dependency to the exponential as a subdominant contribution.
- [88].Vergassola M., Deneke V. E., and Talia S. D., Proceedings of the National Academy of Sciences 115, 10.1073/pnas.1714873115 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [89].Duffy D. G., ACM Transactions on Mathematical Software 19, 333 (1993). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



