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

Response Theory Identifies Reaction Coordinates and Explains Critical Phenomena in Noisy Interacting Systems

N Zagli 1 4, V Lucarini 2 4 and G A Pavliotis 3 1 Nordita, Stockholm University and KTH Royal Institute of Technology - Hannes Alfvéns väg 12, SE-106 91 Stockholm, Sweden 2 Department of Mathematics and Statistics, University of Reading - Reading, RG6 6AX, UK 3 Department of Mathematics, Imperial College London - London, SW7 2AZ, UK 4 Centre for the Mathematics of Planet Earth, University of Reading - Reading, RG6 6AX, UK [email protected]
Abstract

We consider a class of nonequilibrium systems of interacting agents with pairwise interactions and quenched disorder in the dynamics featuring, in the thermodynamic limit, phase transitions. We provide conditions on the microscopic structure of interactions among the agents that lead to a dimension reduction of the system in terms of a finite number of reaction coordinates. Such reaction coordinates prove to be proper nonequilibrium thermodynamic variables as they carry information on correlation, memory and resilience properties of the system. Phase transitions can be identified and quantitatively characterised as singularities of the complex valued susceptibility functions associated to the reaction coordinates. We provide analytical and numerical evidence of how the singularities affect the physical properties of finite size systems.

Keywords: Multiagent models, Critical phenomena, Resilience, Collective variables, Statistical mechanics, Linear Response, Nonequilibrium systems

1 Introduction

Interacting agent models are at the basis of the microscopic description of the rich variety of collective emergent phenomena that high dimensional complex systems often exhibit. Common applications of such models range from synchronisation of nonlinear oscillators  [1, 2, 3, 4, 5, 6, 7, 8], see also the recent special issue [9], phase transitions in complex energy landscapes [10, 11] to opinion dynamics and consensus formation [12, 13], socio-economic siences [14, 15], life sciences [16], the dynamics of the brain [17], formation of swarms [18, 19], dynamical networks [20], self-gravitating systems [21, 22] and algorithms for optimisation and training of neural networks [23, 24, 25]. In the thermodynamic limit, such systems can exhibit phase transitions resulting from the interplay between the interaction among the agents, their internal dynamics and the noise. Such critical phenomena are associated with the so called critical slowing down phenomenon, that is by long lasting, persistent fluctuations exhibiting statistics far away from Gaussianity, see [26] for a detailed description of the critical fluctuations for an equilibrium continuous phase transition.

1.1 From a microscopic to a macroscopic description of the system

From an operational point of view, the investigation of these critical phenomena heavily relies on the identification of a suitable set of observables, commonly referred to as order parameters (OPs), collective variables (CVs) or reaction coordinates (RCs), acting as thermodynamic variables by providing a low dimensional description of the macroscopic feature of the system. Whereas it is true that all the three notions of OP, CV and RC are related to a dimension reduction procedure provided by a projection from a high dimensional system to a reduced state space, they refer to different features of the system. Along the lines of [27], we define a CV as any function of the full phase space. An OP is then a suitable combination of CVs that is able to distinguish between macroscopic, possibly metastable, states of the sytem. As some relevant parameters of the system are varied, one wishes to identify critical points by observing asymptotic properties of the OPs, that is by constructing phase diagrams. On the other hand, the term RC refers to a suitable OP that is able to capture time dependent properties of the system, e.g. relaxation phenomena or transitions between metastable states.
A major challenge in the identification of critical phenomena is that, while order parameters or reaction coordinates can in many cases be deduced for equilibrium systems using, e.g., symmetry arguments, the identification of reaction coordinates in nonequilibrium settings is far less trivial [28, 29, 27] and is tailored to the specific problem at hand. Here we provide a constructive, as opposed to ad hoc, criterion to identify a set of reaction coordinates {Ai}\{A_{i}\} for a large class of interacting agent systems based on the microscopic interaction structure among the agents. Inspired by the success of response theory in explaining critical phenomena in a class of interacting systems [30, 31, 32, 33], we show that, according to this perspective, the {Ai}\{A_{i}\} not only represent OPs for the system but correspond to suitable RCs as they act as proper nonequilibrium thermodynamic variables. On the one hand, their stationary asymptotic properties identify the different macroscopic states of the system. On the other hand, they also carry information on correlation properties and resilience of the system to perturbations. Most importantly, we show that such observables represent the optimal observables to fully characterise the development of a critical phenomenon as the critical, strongly peaked, resonances in their linear response to external perturbations allow for a quantitative description of the critical behaviour.

1.2 Linear Response

Linear Response Theory provides a powerful framework to predict how statistical properties of a system change as a result of weak, yet arbitrary exogenous perturbations. Originally developed for statistical mechanical systems near equilibrium, the theory has much larger validity [34] with applications ranging from driven, strongly dissipative systems such as the climate [35, 36] to neural networks [37, 38, 39] and financial markets [40]. In brief, a linear response perspective corresponds to the following conceptual and mathematical framework. Given a system with a unique, physical, ergodic (not necessarily smooth) measure μ0\mu_{0}, linear response theory aims at evaluating changes in the expectation value

Φ0:=Φ(𝐱)μ0(d𝐱)=limT+1T0TΦ(𝐱(t))dt\langle\Phi\rangle_{0}:=\int\Phi(\mathbf{x})\mu_{0}(\mathrm{d}\mathbf{x})=\lim_{T\rightarrow+\infty}\frac{1}{T}\int_{0}^{T}\Phi(\mathbf{x}(t))\mathrm{d}t (1)

where Φ\Phi is a generic observable of the system, 𝐱(t)\mathbf{x}(t) is a typical trajectory and the last equality hold because of ergodicity. Linear response theory is valid for a vast class of systems ranging from deterministic dynamical systems [34, 41], Markov chain models [42] to continuous stochastic systems [43, 44] where it can be justified on rigorous mathematical grounds [45]. We here consider a finite dimensional system described by a set of Ito stochastic differential equations with deterministic drift :dd\mathbf{\mathcal{F}}:\mathbb{R}^{d}\to\mathbb{R}^{d} and noise law given by the matrix valued function 𝚺:dd×q\boldsymbol{\Sigma}:\mathbb{R}^{d}\to\mathbb{R}^{d\times q}

d𝐲(t)=(𝐲)dt+𝚺(𝐲)d𝐖t\mathrm{d}\mathbf{y}(t)=\mathbf{\mathcal{F}}(\mathbf{y})\mathrm{d}t+\boldsymbol{\Sigma}(\mathbf{y})\mathrm{d}\mathbf{W}_{t} (2)

where d𝐖t\mathrm{d}\mathbf{W}_{t} is a qq-dimensional Wiener process. If we assume that the stochastic process is hypoelliptic [44], the invariant measure μ0\mu_{0} associated to the above stochastic process is absolutely continuous with respect to the Lebesgue measure with smooth probability density ρ0\rho_{0}, that is μ0(d𝐲)=ρ0(𝐲)d𝐲\mu_{0}(\mathrm{d}\mathbf{y})=\rho_{0}(\mathbf{y})\mathrm{d}\mathbf{y}, even when the deterministic drift \mathbf{\mathcal{F}} induces a chaotic dynamics. We perturb the system by letting +ε𝐘(𝐲)T(t)\mathbf{\mathcal{F}}\to\mathbf{\mathcal{F}}+\varepsilon\mathbf{Y}(\mathbf{y})T(t) where the bounded function T(t)T(t) represents the time modulation of the external perturbation and Y(y)Y(y) its phase space state dependence. In this setting, the statistical properties of the system change as

Φ(t)=Φ0+εΦ1(t)+o(ε),withΦ1(t)=(GΦT)(t)\langle\Phi\rangle(t)=\langle\Phi\rangle_{0}+\varepsilon\langle\Phi\rangle_{1}(t)+o(\varepsilon),\quad\text{with}\quad\langle\Phi\rangle_{1}(t)=\left(G_{\Phi}\star T\right)(t) (3)

where GΦG_{\Phi} is the Green’s function associated to the observable Φ\Phi. The Green’s function GΦG_{\Phi} encodes all the positive and negative feedbacks of the system and allows to evaluate the response of the system to general time dependent forcings. Response formulas provide a useful decomposition of any Green’s function as111In expression (4) the contribution of the continuous part of the spectrum is neglected, which is equivalent to assuming that the Markov semigroup describing the dynamics is quasi compact [46].

GΦ(t)=k=1l=1mkgk(l)tl1eλktG_{\Phi}(t)=\sum_{k=1}\sum_{l=1}^{m_{k}}g^{(l)}_{k}t^{l-1}e^{\lambda_{k}t} (4)

where the complex valued exponential rates λk\lambda_{k}\in\mathbb{C} and 𝐑𝐞λk<0\mathbf{Re}\lambda_{k}<0 are the eigenvalues of the generator of the Markov semigroup associated to the unperturbed dynamics (2) and are referred to as stochastic Ruelle Pollicott resonances and gk(l)g_{k}^{(l)} are suitable coefficients [47, 48, 49]. Moreover, if the eigenvalue λk\lambda_{k} is not simple, that is its multiplicity is mk>1m_{k}>1, modulating polynomial terms arise in the expression for the Green’s function. We remark that the exponential rates λk\lambda_{k} represent an intrinsic property of the unperturbed dynamics and correspond to natural relaxation and oscillatory timescales of the system. Green’s functions are thus the relevant quantities to be investigated in order to detect the onset of critical instabilities of general systems. In the context of interacting systems, it has been shown that long lasting and persistent modes with exponential rates of small real part are associated to poorly damped instabilities due to phase transitions in the thermodynamic limit [31].
It is important to stress that the choice of the observable under investigation remains a critical issue in detecting criticalities as the coefficients gk(l)g^{(l)}_{k} are not an intrinsic property of the system but depend both on the observable Φ\Phi and the state dependence 𝐘(𝐲)\mathbf{Y}(\mathbf{y}) of the perturbation [47]. In particular, if the observable Φ\Phi has vanishingly small projection on the most unstable mode, its associated Green’s function GΦG_{\Phi} would not represent a good proxy for the detection of the critical behaviour of the system.

1.3 This paper

The main contributions of this paper can be summarised as follows:

  • We provide a constructive way to define a finite set of RCs {Ai}\{A_{i}\} for the thermodynamic limit of an ensemble of interacting agents. The RCs correspond to an effective dimension reduction of the system in terms of a few relevant observables. We provide mathematical conditions on the interaction structure among the agents for which the construction of RCs is exact. In the context of synchronisation phenomena for ensembles of phase oscillators with purely harmonic coupling, dimension reduction is usually obtained by invoking the Ott-Antonsen ansatz and generalisations thereof [50, 51, 52, 53]. Our results are instead valid for any coupling structure and even in settings where phase reduction techniques of nonlinear oscillators [54, 55] cannot be applied. In this context, this paper represents a generalisation of [6] to settings with (a) possibly multiplicative, white noise sources and (b) an interaction structure that is not a priori assumed to depend on (functions of) the empirical mean of the ensemble.

  • We derive response formulas for the thermodynamic limit of the ensemble of agents. We show that the response of general mean field observables of the system is mediated by these RCs. In particular, we show that the coupling among the agents manifest itself in the thermodynamic limit as a memory contribution, conveyed by the RCs, to the sensitivity properties of the system. We prove that the macroscopic Green’s function describing the response of the system at a macroscopic level encodes feedbacks stemming from both microscopic properties of each agent and the interaction structure among them.

  • We show that the investigation of criticalities arising from phase transitions are associated, in the thermodynamic limit, to the divergent behaviour of Green’s functions relative to the RCs. Phase transitions can be quantitatively identified as the singular part of the complex valued susceptibility, the Fourier Transform of the Green’s Function. These singularities can be characterised by a pole-residue pair (ω,κ)(\omega^{*},\kappa) with ω\omega^{*}\in\mathbb{R} and κ\kappa\in\mathbb{C}. The pole ω\omega^{*} corresponds to frequencies of the driving force that would critically destabilise the system, whereas the residue κ\kappa determines the strength and detailed nature of the singularity.

  • We provide analytical and numerical evidence that the analysis of response properties of the RCs in a finite size ensemble of agents is able to capture the signature of a phase transition as a resonance around the singular thermodynamic behaviour. We show that both the pole ω\omega^{*} and the residue κ\kappa can be inferred in a finite system. In particular, we illustrate the physical role that the residue κ\kappa plays on response properties of the system.

The paper is structured as follows. In section 2 we introduce the class of interacting systems under investigation and we describe when a dimension reduction in terms of RCs is possible. Section 3 is dedicated to the description of linear response properties of the system in terms of the RCs. In section 4 we investigate the critical response properties of a finite ensemble of Kuramoto oscillators. In the conclusions we mention future research directions of this linear response framework to investigate resilience features of general interacting systems.

2 The interacting agent models

We consider an ensemble of NN interacting MM-dimensional systems {𝐱k}k=1N\{\mathbf{x}^{k}\}_{k=1}^{N} whose dynamics is described by the following stochastic differential equations

d𝐱k=𝐅α(𝐱k;𝐡k)dtθNj=1N𝐊(𝐱k,𝐱j)dt+σ𝐬(𝐱𝐤)d𝐖(k)\mathrm{d}\mathbf{x}^{k}=\mathbf{\mathbf{F}}_{\alpha}(\mathbf{x}^{k};\mathbf{h}^{k})\mathrm{d}t-\frac{\theta}{N}\sum_{j=1}^{N}\mathbf{K}(\mathbf{x}^{k},\mathbf{x}^{j})\mathrm{d}t+\sigma\mathbf{s}(\mathbf{x^{k}})\mathrm{d}\mathbf{W}^{(k)} (5)

The smooth vector field 𝐅α(𝐱;):MM\mathbf{F}_{\alpha}(\mathbf{x};\cdot):\mathbb{R}^{M}\to\mathbb{R}^{M}, possibly depending on a set of parameters α\alpha, defines the local dynamics of each agent corresponding to, in general, nonequilibrium, dissipative conditions.The (pair-wise) interactions among the systems are introduced via the function 𝐊:M×MM\mathbf{K}:\mathbb{R}^{M}\times\mathbb{R}^{M}\to\mathbb{R}^{M} with θ\theta representing the coupling strength. The function 𝐊\mathbf{K} is assumed to be antisymmetric in its arguments, 𝐊(𝐱,𝐲)=𝐊(𝐲,𝐱)\mathbf{K}(\mathbf{x},\mathbf{y})=-\mathbf{K}(\mathbf{y},\mathbf{x}), to model Newton’s third law of motion for the internal forces acting between the systems. The volatility matrix 𝐬:MM×M\mathbf{s}:\mathbb{R}^{M}\to\mathbb{R}^{M\times M} determines the state dependent noise term, with σ\sigma being its amplitude and d𝐖=(d𝐖(1),,d𝐖(N))\mathrm{d}\mathbf{W}=(\mathrm{d}\mathbf{W}^{(1)},\dots,\mathrm{d}\mathbf{W}^{(N)}) representing a M×NM\times N Brownian motion. In the following we will adopt the Itô’s convention. Furthermore, we introduce a source of quenched disorder by letting the function 𝐅α\mathbf{F}_{\alpha} depend on a vector of parameters 𝐡km\mathbf{h}^{k}\in\mathbb{R}^{m} drawn from a fixed known distribution μ\mu, that is 𝐡kμ\mathbf{h}^{k}\sim\mu k=1,,N\forall k=1,\dots,N. The microscopic architecture of the system given by μ\mu can be interpreted either as an intrinsic property of the system, such as the natural frequencies of an ensemble of oscillators, or as a model error feature arising from partial knowledge of the microscopic properties of the agents.
In the thermodynamic limit N+N\to+\infty, the mean field nature of the coupling allows to obtain a macroscopic hydrodynamic description of the ensemble of agents [56, 57, 58]. More specifically, we define the empirical density of the NN particle system as ρN(𝐱,𝐡,t)=1Nj=1Nδ(𝐱𝐱j)δ(𝐡𝐡j)\rho_{N}(\mathbf{x},\mathbf{h},t)=\frac{1}{N}\sum_{j=1}^{N}\delta(\mathbf{x}-\mathbf{x}^{j})\delta(\mathbf{h}-\mathbf{h}^{j}), where δ()\delta(\cdot) represents the Dirac delta distribution. Under general conditions, it is possible to show that the system exhibits propagation of molecular chaos [16, 59], i.e. , fixed T\mathrm{T}\in\mathbb{R} and given a chaotic initial condition 𝐱k(t=0)ρ^(𝐱)\mathbf{x}^{k}(t=0)\sim\hat{\rho}(\mathbf{x}) k=1,,N\forall k=1,\dots,N , the empirical density ρN\rho_{N} converges weakly to ρ(𝐱,t;𝐡)μ(𝐡)\rho(\mathbf{x},t;\mathbf{h})\mu(\mathbf{h}) where the the one-particle distribution ρ(𝐱,t;𝐡)\rho(\mathbf{x},t;\mathbf{h}) satisfies the nonlinear Fokker Planck Equation (NFPE)

tρ=[(𝐅α(𝐱,𝐡)θ(𝐊2ρ)μ(𝐡)d𝐡)ρ]++σ222:(𝐬(𝐱)𝐬T(𝐱)ρ)\begin{split}\partial_{t}\rho&=-\nabla\cdot\left[\left(\mathbf{\mathbf{F}}_{\alpha}(\mathbf{x},\mathbf{h})-\theta\int\left(\mathbf{K}\star_{2}\rho\right)\mu(\mathbf{h})\mathrm{d}\mathbf{h}\right)\rho\right]+\\ &+\frac{\sigma^{2}}{2}\nabla^{2}:\left(\mathbf{s}(\mathbf{x})\mathbf{s}^{T}(\mathbf{x})\rho\right)\end{split} (6)

where ρ(𝐱,0;)=ρ^(𝐱)\rho(\mathbf{x},0;\cdot)=\hat{\rho}(\mathbf{x}) and t[0,T]t\in[0,\mathrm{T}]. In equation (6) the drift term depends on the probability distribution ρ\rho itself through (𝐊2ρ)(𝐱;𝐡)=𝐊(𝐱,𝐲)ρ(𝐲,t;𝐡)d𝐲(\mathbf{K}\star_{2}\rho)(\mathbf{x};\mathbf{h})=\int\mathbf{K}(\mathbf{x},\mathbf{y})\rho(\mathbf{y},t;\mathbf{h})\mathrm{d}\mathbf{y} originating from the coupling among the agents and 2:(𝐬𝐬Tρ)=i,j,k=1Mij2(sik(𝐱)sjk(𝐱)ρ)\nabla^{2}:\left(\mathbf{s}\mathbf{s}^{T}\rho\right)=\sum_{i,j,k=1}^{M}\partial^{2}_{ij}\left(s_{ik}(\mathbf{x})s_{jk}(\mathbf{x})\rho\right) represents the state dependent diffusive part. We remark that the microscopic architecture of the system manifests itself in equation (6) as an expectation value over the distribution μ\mu. In particular, the homogeneous case where no quenched disorder is present is obtained by setting μ(𝐡)=δ(𝐡𝐡0)\mu(\mathbf{h})=\delta(\mathbf{h}-\mathbf{h}_{0}) where 𝐡0\mathbf{h}_{0} is a constant parameter. We observe that equation (9) represent a nonlinear and nonlocal (integro-differential) equation and can thus support multiple coexisting stationary measures, corresponding to different macroscopic state of the system. Furthermore, exchange of stability of such measures, as the parameters of the system are varied, correspond to phase transition for the system [60, 20].
We remark that the phenomenon of propagation of molecular chaos stated above is equivalent to a mean field approximation of the Fokker Planck associated to (5), where the N-particle probability distribution is assumed to be factorised as a product of one-particle distributions ρ(𝐱,t;𝐡)\rho(\mathbf{x},t;\mathbf{h}). This corresponds to the simplest closure scheme, characterised by considering all agents as statistically independent, of the BBGKY hierarchy [61], an infinite set of equations describing the correlations among the N subsystems. It is known that, at a phase transition point, correlations can no longer be neglected at a macroscopic level [26]. Thus, near a phase transition point the NFPE is not expected to capture the behaviour of fluctuations, correlation and relaxation properties of macroscopic observables. We mention that the Linear Response perspective described later in section 3 and 4 can be interpreted as an approach to characterising critical fluctuations.

2.1 Identification of Reaction Coordinates

Equation (6) represents a coarse grained, mesoscopic, description of the system valid for a big, strictly infinite, ensemble of agents. However, even at this coarse grained level, this equation represents an infinite dimensional problem as the drift coefficient depends on the full one-particle probability distribution through the convolution product (𝐊2ρ(\mathbf{K}\star_{2}\rho. It is then fundamental to find conditions on the dynamics of the system for which suitable OPs or RCs can be found. It turns out that the characterisation of RCs for these systems is strictly related to the properties of the interaction kernel 𝐊\mathbf{K}. We consider a separable interaction kernel, that is, admitting a decomposition

Ki(𝐱,𝐲)=l=1dki;l(𝐱)ai;l(𝐲)K_{i}(\mathbf{x},\mathbf{y})=\sum_{l=1}^{d}k_{i;l}(\mathbf{x})a_{i;l}(\mathbf{y}) (7)

where dd\in\mathbb{N}, d<+d<+\infty. We remark that it is always possible to assume that the functions in the above decomposition are linearly independent [62]. In this case, the effect of the nonlinearity in equation (6) can be written in terms of a finite number of suitable observables rather than on the full probability distribution ρ\rho since, using (7), the interaction term results in

(Ki2ρ)(𝐱;𝐡)μ(𝐡)d𝐡=Ki(𝐱,𝐲)ρ(𝐲,t;𝐡)μ(𝐡)d𝐡d𝐲==l=1dki;l(𝐱)ai;l(t):=𝒦i(𝐱;{ai;l}l)\begin{split}\int\left(K_{i}\star_{2}\rho\right)(\mathbf{x};\mathbf{h})\mu(\mathbf{h})\mathrm{d}\mathbf{h}&=\int\int K_{i}(\mathbf{x},\mathbf{y})\rho(\mathbf{y},t;\mathbf{h})\mu(\mathbf{h})\mathrm{d}\mathbf{h}\mathrm{d}\mathbf{y}=\\ &=\sum_{l=1}^{d}k_{i;l}(\mathbf{x})\langle\langle a_{i;l}\rangle\rangle(t):=\mathcal{K}_{i}(\mathbf{x};\{\langle\langle a_{i;l}\rangle\rangle\}_{l})\end{split} (8)

where \langle\langle\cdot\rangle\rangle represents the double expectation value with respect to ρ(𝐱,t;𝐡)\rho(\mathbf{x},t;\mathbf{h}) and μ\mu. Equation (6) can now be written as

tρ=((𝐅α(𝐱,𝐡)θ𝓚(𝐱,𝐀))ρ)++σ222:(𝐬(𝐱)𝐬T(𝐱)ρ):=𝐀ρ\begin{split}\partial_{t}\rho&=-\nabla\cdot\bigg{(}\big{(}\mathbf{\mathbf{F}}_{\alpha}(\mathbf{x},\mathbf{h})-\theta\boldsymbol{\mathcal{K}}\left(\mathbf{x},\langle\langle\mathbf{A}\rangle\rangle\right)\big{)}\rho\bigg{)}+\\ &+\frac{\sigma^{2}}{2}\nabla^{2}:\left(\mathbf{s}(\mathbf{x})\mathbf{s}^{T}(\mathbf{x})\rho\right):=\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle}\rho\end{split} (9)

where 𝐀(𝐱)={ai;l(𝐱)}i=1,l=1M,dp\langle\langle\mathbf{A}(\mathbf{x})\rangle\rangle=\{\langle\langle a_{i;l}(\mathbf{x})\rangle\rangle\}^{M,d}_{i=1,l=1}\in\mathbb{R}^{p} is a finite vector (pp not necessarily equals to dd and pNp\ll N, see later discussion) of observables and Aj(𝐱)A_{j}(\mathbf{x}) j=1,,pj=1,\dots,p are linearly independent functions. We have also defined the nonlinear operator 𝐀\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle} depending on the probability distribution through the vector 𝐀\langle\langle\mathbf{A}\rangle\rangle. Following the terminology of [60], this scenario corresponds to a finite nonlinearity dimension equals to pp for equation (6) and it corresponds to a natural dimension reduction of the ensemble of agents in terms of the variables 𝐀\langle\langle\mathbf{A}\rangle\rangle.
Separable kernels 𝐊(𝐱,𝐲)=𝒰(𝐱𝐲)\mathbf{K}(\mathbf{x},\mathbf{y})=\nabla\mathcal{U}(\mathbf{x}-\mathbf{y}) arising from interaction potentials are routinely used in applications, with radial potentials of the form 𝒰(𝐱)=u(|𝐱|)\mathcal{U}(\mathbf{x})=u(|\mathbf{x}|) being a common choice. A paradigmatic example is represented by quadratic interactions u(x)=x22u(x)=\frac{x^{2}}{2}, yielding a separable kernel 𝐊(𝐱,𝐲)=𝐱𝐲\mathbf{K}(\mathbf{x},\mathbf{y})=\mathbf{x}-\mathbf{y}. This corresponds to linear forces among the particles trying to synchronise them towards their centre of mass x¯=1Nj𝐱j(t)\bar{x}=\frac{1}{N}\sum_{j}{\mathbf{x}}_{j}(t) and has numerous applications, see our previous work [30, 31] on this. In this case, i=1,,M\forall i=1,\dots,M, the expansion (7) of the interaction kernel holds with d=2d=2 and (fi;1,gi;1,fi;2,gi;2)=(x,0,0,y)(f_{i;1},g_{i;1},f_{i;2},g_{i;2})=(x,0,0,-y) resulting in a non linear dimension p=Mp=M for equation (9) given by the observables 𝐀(𝐱)=𝐱\langle\langle\mathbf{A}(\mathbf{x})\rangle\rangle=\langle\langle\mathbf{x}\rangle\rangle. Similarly, higher order polynomials u(x)=k=2nukxku(x)=\sum_{k=2}^{n}u_{k}x^{k} with n2n\geq 2 give rise to separable kernels and a finite nonlinearity dimension. As a further example, we mention separable kernels provided by trigonometric interactions of any order with applications to synchronisation phenomena, see [63, 8] and the section below on the Kuramoto model. We refer the reader to the conclusions section for a discussion regarding non separable kernels.
We remark that equation (9) can be interpreted as a thermodynamic formulation of the system with the 𝐀\langle\langle\mathbf{A}\rangle\rangles being a generalisation of thermodynamic variables to nonequilibrium settings [64]. Firstly, stationary distributions ρ0=ρ0(𝐱;𝐡)\rho_{0}=\rho_{0}(\mathbf{x};\mathbf{h}) of (9) can be parametrised in terms of the stationary values of the thermodynamic variables 𝐀0\langle\langle\mathbf{A}\rangle\rangle_{0} according to the self consistency equation given by the stationary condition 0ρ0=0\mathcal{L}_{0}\rho_{0}=0 where 0=𝐀0\mathcal{L}_{0}=\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle_{0}} is the operator defined in (9) evaluated at stationarity. This is better understood in the case of homogeneous equilibrium systems, characterised by a gradient dynamics 𝐅α(𝐱)=Vα(𝐱)\mathbf{F}_{\alpha}(\mathbf{x})=-\nabla V_{\alpha}(\mathbf{x}), an interaction potential 𝒰(𝐱)\mathcal{U}(\mathbf{x}), thermal noise 𝐬(𝐱)=𝟏M×M\mathbf{s}(\mathbf{x})=\mathbf{1}_{M\times M} and no quenched disorder. In such systems, stationary solutions of (6) satisfy the self consistency equation [20]

ρeq(𝐱)=1Zexp(2σ2(Vα(𝐱)𝒰ρeq))\rho_{eq}(\mathbf{x})=\frac{1}{Z}\exp\left(-\frac{2}{\sigma^{2}}\left(V_{\alpha}(\mathbf{x})-\mathcal{U}\star\rho_{eq}\right)\right) (10)

where ZZ is a normalisation constant. Assuming a separable kernel, equilibrium stationary solutions are parametrised by a finite number of observables 𝐀eq\langle\mathbf{A}\rangle_{eq} since they can be written as ρeq(𝐱;𝐀eq)=1Zexp(2σ2(Vα(𝐱)f(𝐱,𝐀eq)))\rho_{eq}\left(\mathbf{x};\langle\mathbf{A}\rangle_{eq}\rangle\right)=\frac{1}{Z}\exp\left(-\frac{2}{\sigma^{2}}\left(V_{\alpha}(\mathbf{x})-f(\mathbf{x},\langle\mathbf{A}\rangle_{eq})\right)\right) where ff is a suitable function deriving by the convolution product 𝒰ρeq\mathcal{U}\star\rho_{eq}, similarly to equation (8).
Moreover, the 𝐀(𝐱)\langle\langle\mathbf{A}(\mathbf{x})\rangle\rangles represent proper RCs of the system as they govern time dependent properties of the ensemble of agents, such as its response, correlation and memory features, as we elucidate in the next section.

3 Response Formulas

We remark that a stationary state ρ0\rho_{0} of equation (9) is characterised by the constant set of observables 𝐀0\langle\langle\mathbf{A}\rangle\rangle_{0}, where the subscript denotes that the expectation value is taken with respect to ρ0\rho_{0} (and μ\mu). We perturb such state with a time and state dependent perturbation by letting 𝐅α(𝐱;𝐡)𝐅α(𝐱;𝐡)+εT(t)𝐗(𝐱)\mathbf{F}_{\alpha}(\mathbf{x};\mathbf{h})\rightarrow\mathbf{F}_{\alpha}(\mathbf{x};\mathbf{h})+\varepsilon T(t)\mathbf{X}(\mathbf{x}) and we write ρ(𝐱,t;𝐡)ρ0(𝐱;𝐡)+ερ1(𝐱,t;𝐡)\rho(\mathbf{x},t;\mathbf{h})\approx\rho_{0}(\mathbf{x};\mathbf{h})+\varepsilon\rho_{1}(\mathbf{x},t;\mathbf{h}). We observe that, since 𝐀𝐀0+ε𝐀1(t)\langle\langle\mathbf{A}\rangle\rangle\approx\langle\langle\mathbf{A}\rangle\rangle_{0}+\varepsilon\langle\langle\mathbf{A}\rangle\rangle_{1}(t), where the last subscript denotes the expectation value with respect to ρ1\rho_{1}, the perturbation to the nonlinear part of the drift term can be written as

𝓚(𝐱,𝐀)𝓚(𝐱,𝐀0)+ε𝐉(𝐱)𝐀1(t)\boldsymbol{\mathcal{K}}\left(\mathbf{x},\langle\langle\mathbf{A}\rangle\rangle\right)\approx\boldsymbol{\mathcal{K}}\left(\mathbf{x},\langle\langle\mathbf{A}\rangle\rangle_{0}\right)+\varepsilon\mathbf{J}\left(\mathbf{x}\right)\langle\langle\mathbf{A}\rangle\rangle_{1}(t) (11)

where the matrix 𝐉M×p\mathbf{J}\in\mathbb{R}^{M\times p} provides the information on how the drift term changes, in a linear approximation, with respect to the observables 𝐀\langle\langle\mathbf{A}\rangle\rangle and is defined as

Jij(𝐱)=Jij(𝐱,𝐀0)=𝒦iAj(𝐱,𝐀0)J_{ij}\left(\mathbf{x}\right)=J_{ij}(\mathbf{x},\langle\langle\mathbf{A}\rangle\rangle_{0})=\frac{\partial\mathcal{K}_{i}}{\partial\langle\langle A_{j}\rangle\rangle}\left(\mathbf{x},\langle\langle\mathbf{A}\rangle\rangle_{0}\right) (12)

From (9) we obtain an equation for the perturbation to the stationary measure

ρ1(𝐱,t;𝐡)=0tT(s)e(ts)0pρ0ds++θj=1p0tAj1(s)k=1Me(ts)0xk(Jkj(𝐱)ρ0)ds\begin{split}\rho_{1}(&\mathbf{x},t;\mathbf{h})=\int_{0}^{t}T(s)e^{\left(t-s\right)\mathcal{L}_{0}}\mathcal{L}_{p}\rho_{0}\mathrm{d}s+\\ &+\theta\sum_{j=1}^{p}\int_{0}^{t}\langle\langle A_{j}\rangle\rangle_{1}(s)\sum_{k=1}^{M}e^{\left(t-s\right)\mathcal{L}_{0}}\partial_{x_{k}}\left(J_{kj}\left(\mathbf{x}\right)\rho_{0}\right)\mathrm{d}s\end{split} (13)

where p:=(𝐗(𝐱))\mathcal{L}_{p}\cdot:=-\nabla\cdot\left(\mathbf{X}(\mathbf{x})\quad\!\!\!\!\cdot\quad\!\!\!\!\right) is the perturbation operator. We now evaluate the perturbation to the observable Ai(𝐱)A_{i}(\mathbf{x}) by taking the expected value with respect to ρ1\rho_{1} and μ\mu to obtain

Ai1(t)=(GiT)(t)+θj=1p(YijAj1)(t)\begin{split}\langle\langle A_{i}\rangle\rangle_{1}(t)=\left(G_{i}\star T\right)(t)+\theta\sum_{j=1}^{p}\left(Y_{ij}\star\langle\langle A_{j}\rangle\rangle_{1}\right)(t)\end{split} (14)

where \star denotes the convolution product and we have defined the microscopic Green’s functions

Gi(t)\displaystyle G_{i}(t) =Θ(t)d𝐡μ(𝐡)d𝐱Ai(𝐱)et0pρ0,\displaystyle=\Theta(t)\int\mathrm{d}\mathbf{h}\mu(\mathbf{h})\int\mathrm{d}\mathbf{x}A_{i}(\mathbf{x})e^{t\mathcal{L}_{0}}\mathcal{L}_{p}\rho_{0}, (15a)
Yij(t)\displaystyle Y_{ij}(t) =Θ(t)d𝐡μ(𝐡)d𝐱Ai(𝐱)et0k=1Mxk(Jkj(𝐱)ρ0).\displaystyle=\Theta(t)\int\mathrm{d}\mathbf{h}\mu(\mathbf{h})\int\mathrm{d}\mathbf{x}A_{i}(\mathbf{x})e^{t\mathcal{L}_{{}_{0}}}\sum_{k=1}^{M}\partial_{x_{k}}\left(J_{kj}\left(\mathbf{x}\right)\rho_{0}\right). (15b)

We observe that GiG_{i} depends on the perturbation whereas YijY_{ij} does not, representing thus an intrinsic property of the system describing its internal feedbacks, see [30]. The coupling among the agents manifests itself as a memory term in (14) conveyed by the variables Aj\langle\langle A_{j}\rangle\rangle through the microscopic Green’s Function YijY_{ij}. Taking the Fourier transform of (14) we obtain

j=1pPij(ω)Aj1(ω)=T(ω)χi(ω)\sum_{j=1}^{p}P_{ij}(\omega)\langle\langle A_{j}\rangle\rangle_{1}(\omega)=T(\omega)\chi_{i}(\omega) (16)

where we have defined the p×pp\times p matrix

Pij(ω)=δijθYij(ω)P_{ij}(\omega)=\delta_{ij}-\theta Y_{ij}(\omega) (17)

and χi(ω)\chi_{i}(\omega) (Yij(ω)Y_{ij}(\omega)) are the Fourier transform of the microscopic Green functions Gi(t)G_{i}(t) (Yij(t)Y_{ij}(t)) respectively. Equations (16) and (17) completely characterise the response of statistical properties of the system to perturbations in terms of the observables 𝐀\langle\langle\mathbf{A}\rangle\rangle. When χi(ω)\chi_{i}(\omega) is an analytic function in the upper complex ω\omega plane and Pij(ω)P_{ij}(\omega) is invertible the response of the system is smooth and given by

𝐀(ω)=𝝌~(ω)T(ω),𝝌~(ω)=𝐏1(ω)𝝌(ω)\langle\langle\mathbf{A}\rangle\rangle(\omega)=\tilde{\boldsymbol{\chi}}(\omega)T(\omega),\quad\tilde{\boldsymbol{\chi}}(\omega)=\mathbf{P}^{-1}(\omega)\boldsymbol{\chi}(\omega) (18)

where 𝝌~\tilde{\boldsymbol{\chi}} is the complex valued macroscopic susceptibility associated to the RCs. In this case, G~i(t)\tilde{G}_{i}(t), the inverse Fourier transform of χ~i(ω)\tilde{\chi}_{i}(\omega), defines a causal macroscopic Green’s function describing the response properties of the system as

Ai(t)1(t)=(G~iT)(t).\langle\langle A_{i}(t)\rangle\rangle_{1}(t)=(\tilde{G}_{i}\star T)(t). (19)

From an operational point of view, the macroscopic Green’s function G~i(t)\tilde{G}_{i}(t) is the quantity that is experimentally accessible at a macroscopic level. This function encodes all positive and negative feedbacks of the system, including feedbacks stemming from microscopic single agent features and from the interaction among them. We remark that the macroscopic Green’s function G~i(t)\tilde{G}_{i}(t), being the inverse Fourier transform of the macroscopic susceptibility defined in (18), has a non trivial expression in terms of the microscopic Green’s function Gi(t)G_{i}(t) and Yij(t)Y_{ij}(t).
Criticalities in the response of the system arise when the negative feedbacks are no longer successful in damping the external perturbations and are signalled by the fact that some modes of the macroscopic Green’s function G~i(t)\tilde{G}_{i}(t) do not decay in time. More specifically, the response of the system breaks down as a result of a pole in 𝝌~(ω)\tilde{\boldsymbol{\chi}}(\omega) when either χi(ω)\chi_{i}(\omega) develops a singularity, corresponding to a destabilisation of each agent in the system [30], or when Pij(ω)P_{ij}(\omega) becomes noninvertible. Equation (17) shows that the latter case arises due to endogenous instabilities associated with positive feedbacks resulting from the coupling among the agents. This identifies the occurrence of a phase transition in the system as it is intimately related to taking the thermodynamic limit [30]. We remark that, generally, a linear stability analysis of nonlinear Fokker Planck equations such as (9), e.g. for the Kuramoto model described below, involves the investigation of spectral properties of (infinite dimensional) linear integral operators [60, 65]. However, the identification of RCs allows to simplify the stability problem into the analysis of the finite dimensional matrix Pijp×pP_{ij}\in\mathbb{C}^{p\times p} encoding their intrinsic response properties.

4 Critical phenomena for the Kuramoto model

We elucidate the above results by investigating critical phenomena in the Kuramoto model [65, 66], the paradigmatic example of synchronisation phenomena of phase oscillators. We remark that, even if the Kuramoto model has been thoroughly investigated, a stability theory of the stationary solutions of the Kuramoto model is still lacking, see [67] for partial results in this direction in deterministic settings. The goal of this section is to shed light on the physical mechanisms underlying the onset of synchronisation encoded in the properties of the Green’s function of the system. In particular, we are interested in showing how the critical response of a finite dimensional system relates to the properties of the complex valued susceptibility 𝝌~(ω)\tilde{\boldsymbol{\chi}}(\omega).
The Kuramoto model is defined by the following SDEs

dxk=[hkθNj=1Nsin(xkxj)]dt+σdWk\mathrm{d}x_{k}=\big{[}h_{k}-\frac{\theta}{N}\sum_{j=1}^{N}\sin\left(x_{k}-x_{j}\right)\big{]}\mathrm{d}t+\sigma\mathrm{d}W_{k} (20)

where the quenched disorder in this case corresponds to the presence of a non-trivial distribution of intrinsic frequencies across the oscillators, so that hkμh_{k}\sim\mu. The interaction kernel K(x,y)=sin(xy)=sin(x)cos(y)cos(x)sin(y)K(x,y)=\sin(x-y)=\sin(x)\cos(y)-\cos(x)\sin(y) is separable and leads, in the thermodynamic limit, to the following NFPE of nonlinearity dimension p=2p=2

tρ=x(h𝒦(x,𝐀))ρ+σ22xxρ=𝐀ρ\partial_{t}\rho=-\partial_{x}\big{(}h-\mathcal{K}(x,\langle\langle\mathbf{A}\rangle\rangle)\big{)}\rho+\frac{\sigma^{2}}{2}\partial_{xx}\rho=\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle}\rho (21)

where 𝒦(x,𝐀)=sin(x)A1(x)cos(x)A2(x)\mathcal{K}(x,\langle\langle\mathbf{A}\rangle\rangle)=\sin(x)\langle\langle A_{1}(x)\rangle\rangle-\cos(x)\langle\langle A_{2}(x)\rangle\rangle and 𝐀(x)=(cos(x),sin(x))\mathbf{A}(x)=\left(\cos(x),\sin(x)\right) represents the set of reaction coordinates for the system. The uniform distribution ρ0=12π\rho_{0}=\frac{1}{2\pi} is always a solution of equation (21) corresponding to a incoherent state characterised by a set of reaction coordinates 𝐀0=(0,0)\langle\langle\mathbf{A}\rangle\rangle_{0}=(0,0). We here focus on the task of assessing the response properties of the incoherent state ρ0\rho_{0}. From (12), we evaluate 𝐉=(sin(x),cos(x))\mathbf{J}=\left(\sin(x),-\cos(x)\right) yielding222We refer the reader to the Appendix for the details regarding the calculations regarding this section. the microscopic Green’s Function Yij(t)Y_{ij}(t):

Yij(t)=Θ(t)dhμ(h)Dh2+D2ddtCij(t;h),Y_{ij}(t)=-\Theta(t)\int\mathrm{d}h\mu(h)\frac{D}{h^{2}+D^{2}}\frac{\mathrm{d}}{\mathrm{d}t}C_{ij}(t;h), (22)

where D=σ22D=\frac{\sigma^{2}}{2} and Cij(t;h)C_{ij}(t;h) is the correlation function in the stationary state between observable Ai(x)A_{i}(x) and ψj(x;h)=Aj(x)hDxAj(x)\psi_{j}(x;h)=A_{j}(x)-\frac{h}{D}\partial_{x}A_{j}(x). If we consider an even distribution of intrinsic frequencies μ(h)=μ(h)\mu(-h)=\mu(h) the matrix (17) is diagonal Pij(ω)=P(ω)δijP_{ij}(\omega)=P(\omega)\delta_{ij} and

P(ω)=1θD2μ(h)D2+h2(1+2iωC(ω;h))dhP(\omega)=1-\frac{\theta D}{2}\int\frac{\mu(h)}{D^{2}+h^{2}}\left(1+2i\omega C\left(\omega;h\right)\right)\mathrm{d}h (23)

where C(ω;h)=CR(ω;h)+iCI(ω;h)C(\omega;h)=C_{R}(\omega;h)+iC_{I}(\omega;h) is the Fourier Transform of the (diagonal part of the) correlation function Cij(t;h)C_{ij}(t;h) and the explicit expressions for CRC_{R} and CIC_{I} are in the Appendix. Singularities in the response of the reaction coordinate Ai(x)A_{i}(x), due to a phase transition, are characterised by a value ω\omega^{*}\in\mathbb{R} such that the matrix Pij(ω)P_{ij}(\omega^{*}) is not invertible, or, equivalently, such that P(ω)=0P(\omega^{*})=0. Separating real and imaginary part of P(ω)=0P(\omega^{*})=0 results respectively in

Dθ2μ(h)h2+D2(12ωCI(ω;h))dh=1,\displaystyle\frac{D\theta}{2}\int\frac{\mu(h)}{h^{2}+D^{2}}\left(1-2\omega^{*}C_{I}(\omega^{*};h)\right)\mathrm{d}h=1, (24a)
ωμ(h)h2+D2CR(ω;h)dh=0.\displaystyle\omega^{*}\int\frac{\mu(h)}{h^{2}+D^{2}}C_{R}(\omega^{*};h)\mathrm{d}h=0. (24b)

Equation (24b) determines the critical value ω\omega^{*}\in\mathbb{R} whereas (24a) identifies the corresponding transition point in the parameter space (θ,D)(\theta,D). An immediate solution of (24b) is ω=ω0=0\omega^{*}=\omega_{0}=0, which takes place, according to (24a), at the critical coupling strength θcstat=2[Dh2+D2μ(h)dh]1\theta^{stat}_{c}=2[\int\frac{D}{h^{2}+D^{2}}\mu(h)\mathrm{d}h]^{-1} in agreement with [68].

Refer to caption
Figure 1: Real part of the susceptibility χ~1(ω)\tilde{\chi}_{1}(\omega) associated to the reaction coordinate A1(x)A_{1}(x). Panel (a) refers to the homogeneous case μ(h)=δ(0)\mu(h)=\delta(0) and panel (b) to a bimodal distribution μh0(h)=12(δ(h0)+δ(h0))\mu_{h_{0}}(h)=\frac{1}{2}(\delta(h_{0})+\delta(-h_{0})). As the thermodynamic limit is approached, 𝐑𝐞χ~1(ω)\mathbf{Re}\tilde{\chi}_{1}(\omega) develops a diverging behaviour for ω=ω0=0\omega=\omega_{0}=0 (panel (a)) and ω=ω=h02D2\omega=\omega^{\star}=\sqrt{h_{0}^{2}-D^{2}} (panel (b)). The investigation of 𝐈𝐦χ~1\mathbf{Im}\tilde{\chi}_{1} (bottom insets) provides information on the residue associated to the pole ω0=0\omega_{0}=0 of the susceptibility for the homogeneous case. The top insets show the macroscopic Green function G~1(t)\tilde{G}_{1}(t). Red (black) lines refer to settings at (away from) the phase transition. The numerical response experiments have been performed with N=6,12,24,48×103N=6,12,24,48\times 10^{3}, see more details in the Appendix.

This setting corresponds to a static phase transition, characterised by a vanishing pole ω0=0\omega_{0}=0 of the susceptibility and a lack of oscillations in the system, see panel (a) of Figure 1 where we have investigated this scenario for μ0(h)=δ(h)\mu_{0}(h)=\delta(h). The introduction of a distribution of natural frequencies μ\mu brings the system to a nonequilibrium regime and it is interesting to investigate whether a dynamical phase transition, characterised by a pair of poles ω1=ω2=ω>0\omega_{1}=-\omega_{2}=\omega^{*}>0, can arise in the system. This involves the study of the dynamical properties encoded in C(ω;h)C(\omega;h). Further insight can be gained by observing that, if (24b) is satisfied for ω>0\omega^{*}>0, the following equation holds, see details in the Appendix,

ω±hD2+(ω±h)2μ(h)dh=0,\int\frac{\omega^{*}\pm h}{D^{2}+(\omega^{*}\pm h)^{2}}\mu(h)\mathrm{d}h=0, (25)

which is a known result for the Kuramoto model [66]. In particular, no dynamical phase transition can take place if μ(h)\mu(h) is unimodal as no ω0\omega^{*}\neq 0 satisfying (25) exists. We here consider, as in [69], a bimodal distribution μh0(h)=12(δ(hh0)+δ(h+h0))\mu_{h_{0}}(h)=\frac{1}{2}(\delta(h-h_{0})+\delta(h+h_{0})). In this setting, equation (24b) admits, further to the static transition scenario, a new solution ω=h02D2\omega^{*}=\sqrt{h_{0}^{2}-D^{2}} when h0>Dh_{0}>D. According to (24a), this dynamical phase transition scenario takes place at the transition point θdyn=4D\theta_{dyn}=4D. Panel (b) of Figure 1 confirms that, at the phase transition, response properties break down as the susceptibility develops a singular resonant behaviour for ω=ω\omega=\omega^{*} as NN\to\infty, resulting in greatly amplified and long lasting oscillations of the macroscopic Green’s function G~i(t)\tilde{G}_{i}(t).

4.1 Response properties of the finite system

Following [43], we evaluate response properties of the finite system by performing nn simulations of equations (20) where the initial conditions are sampled according to ρ0=12π\rho_{0}=\frac{1}{2\pi} and where at time t=0t=0 we apply a macroscopic perturbation F(x;h)=hF(x;h)=h+εX(x)δ(t)F(x;h)=h\to F(x;h)=h+\varepsilon X(x)\delta(t). The macroscopic Green’s function G~1(t)\tilde{G}_{1}(t) is obtained as the average over the nn ensemble members of A11(t)\langle\langle A_{1}\rangle\rangle_{1}(t) which we estimate as 1Ni=1NA1(xi(t))\frac{1}{N}\sum_{i=1}^{N}A_{1}(x_{i}(t)). We remark that, since the unperturbed measure is uniform, a constant perturbation X(x)=1X(x)=1 would result in a vanishing response as G1(t)0G_{1}(t)\equiv 0, see (15a). We have here chosen a non trivial state dependent perturbation X(x)sin(x)X(x)\propto\sin(x). We note that such term has been used to model excitability features in Kuramoto-like models for brain dynamics [70]. The numerical experiments indicate that the finite system susceptibility can be written, at the phase transition, as

χ~1(ω)={κω+iγ(N)+ψ(ω)if ω=0κωω+iγ(N)κω+ω+iγ(N)+ψ(ω)if ω>0\tilde{\chi}_{1}(\omega)=\begin{cases}\frac{\kappa}{\omega+i\gamma(N)}+\psi(\omega)&\text{if $\omega^{*}=0$}\\ \frac{\kappa}{\omega-\omega^{*}+i\gamma(N)}-\frac{\kappa^{*}}{\omega+\omega^{*}+i\gamma(N)}+\psi(\omega)&\text{if $\omega^{*}>0$}\end{cases} (26)

where ψ(ω)\psi(\omega) is an analytic function in the upper complex ω\omega plane and the residue κ\kappa characterises the strength of the divergence. The residue is related to the properties of P(ω)P(\omega) and can be found as κ=(θP(ω))1\kappa=\left(\theta P^{\prime}(\omega^{*})\right)^{-1}, see Appendix for more details.

Refer to caption
Figure 2: Top panels: Response of the reaction coordinate A1A_{1} to a sinusoidal forcing T(t)=sin(ωt)T(t)=\sin(\omega t), with ω=ω\omega=\omega^{*} (left) and ω=2ω\omega=2\omega^{*} (right) . The dashed line corresponds to T(t)T(t), the continuous one to the analytical result and the red dots to the numerical response evaluated as G~1T\tilde{G}_{1}\star T with G~1(t)\tilde{G}_{1}(t) obtained from the numerical simulations for N=48×103N=48\times 10^{3}. The amplitude of the numerical response has been re-scaled by its maximum value. Bottom panel: Correlation function (CF) between T(t)T(t) and the numerical response as a function of the angle ωt\omega t. Its maximum value yields an estimate of the phase shift ϕ\phi. The vertical dashed line corresponds to the analytical value of ϕ\phi. Red dots correspond to multiples of π2\frac{\pi}{2} and are just a visual aid.

We remark that the finiteness of the system results in a mollifying effect of the singularity into a resonant behaviour, where the poles are slightly shifted as ±ω±ω+iγ(N)\pm\omega^{*}\to\pm\omega^{*}+i\gamma(N), with γ(N)0\gamma(N)\to 0 as N+N\to+\infty. In the homogeneous case the residue turns out to be completely imaginary and equal to κ=i2\kappa=\frac{i}{2}, whereas for the bimodal case it results in κ=κR+iκI=D4ω+i4\kappa=\kappa_{R}+i\kappa_{I}=-\frac{D}{4\omega^{*}}+\frac{i}{4}. In the former case, one expects that, as N+N\to+\infty, the function ω𝐈𝐦χ~1(ω)\omega\mathbf{Im}\tilde{\chi}_{1}(\omega) will be constant and equal to |κ|=12|\kappa|=\frac{1}{2} in the proximity of the pole and to quickly drop to zero for ω=0\omega=0. A visual inspection of the bottom inset of panel (a) of Figure 1 provides a clear evidence of such theoretical prediction.
A fundamentally different situation occurs in the latter case as the residue has a non-vanishing real part leading to a distortion of the resonance at ω\omega^{*} as 𝐈𝐦χ~(ω)0\mathbf{Im}\tilde{\chi}(\omega^{*})\neq 0 and, thus, affecting the way the amplitude of the oscillations of the response track the forcing signal. This is more easily understood when considering a sinusoidal driving forcing T(t)=sin(ω^t)T(t)=\sin(\hat{\omega}t). The response (in the time domain) can be obtained by performing an inverse Fourier transform of (18) and results in

A11(t;ω^)=|χ~(ω^)|sin(ω^t+arctan(𝐑𝐞χ~(ω^)𝐈𝐦χ~(ω^))π2):=|χ~(ω^)|sin(ω^t+ϕ)\langle\langle A_{1}\rangle\rangle_{1}(t;\hat{\omega})=|\tilde{\chi}(\hat{\omega})|\sin\left(\hat{\omega}t+\arctan\left(\frac{\mathbf{Re}\tilde{\chi}(\hat{\omega})}{\mathbf{Im}\tilde{\chi}(\hat{\omega})}\right)-\frac{\pi}{2}\right):=|\tilde{\chi}(\hat{\omega})|\sin\left(\hat{\omega}t+\phi\right) (27)

The amplitude of the response is given by the amplitude of the macroscopic susceptibility. We expect that driving frequencies around the pole ω\omega^{*} would destabilise the system resulting in greatly amplified oscillations, whereas the system would be much more resilient for higher driving frequencies. Moreover, if the imaginary part of the susceptibility were to vanish, the response would feature either no phase shift as ϕ=0\phi=0 or phase reversal as ϕ=π\phi=-\pi. At the critical frequency ω^=ω\hat{\omega}=\omega^{*} we have, from (26), that the response is given by A11(t;ω)=|κ|γ(N)sin(ωt+ϕ)\langle\langle A_{1}\rangle\rangle_{1}(t;\omega^{*})=\frac{|\kappa|}{\gamma(N)}\sin\left(\omega^{*}t+\phi\right) with ϕ=arctan(κIκR)π2=arctan(ωD)π2\phi=\arctan(-\frac{\kappa_{I}}{\kappa_{R}})-\frac{\pi}{2}=\arctan(\frac{\omega^{*}}{D})-\frac{\pi}{2}. The non vanishing real part of the residue yields a spontaneous phase shift of the response to perturbations as confirmed by Figure 2 (left panels). We remark that this provides a quantitative estimate, at the phase transition, of the phase shift phenomenon observed (in deterministic settings) in [71]. On the other hand, if the system is driven at higher frequency, e.g. ω^=2ω\hat{\omega}=2\omega^{*}, we expect the real part of the susceptibility to decay faster than the imaginary part [72] and the response to be in quadrature with the forcing, i.e. A11(t;2ω)cos(2ωt)\langle\langle A_{1}\rangle\rangle_{1}(t;2\omega^{*})\propto\cos(2\omega^{*}t), as confirmed by Figure 2 (right panels).

4.2 Conclusions

In this paper we have considered the relevant issue of the identification of a set of reaction coordinates {Ai}\{A_{i}\} for the thermodynamic limit of ensembles of interacting agents in the general case of pairwise interaction and quenched disorder in the dynamics. Given a general class of interaction structure among the agents, we have provided a framework to define the {Ai}s\{A_{i}\}s and construct a nonequilibrium thermodynamical formulation of the macroscopic description of the system. These reaction coordinates provide a parametrization of the stationary measure and fully characterise the linear response of the system to perturbations. Such a model reduction strategy is particularly useful as it allows to study not only smooth regimes of the response but also critical behaviours, characterised as an emergence of singular poles of the complex valued susceptibility of the system. The analysis in terms of the reaction coordinates of the paradigmatic Kuramoto model allows us to show (a) that our framework encompasses known results on the stability properties of the incoherent state and (b) that the time dependent properties of the onset of synchronisation can be quantitatively investigated through the susceptibility function describing the response of the system to general driving forces.
Our results are exact in the case of separable interactions. However, non separable kernels are also found in applications. If 𝐊(𝐱,𝐲)\mathbf{K}(\mathbf{x},\mathbf{y}) is a smooth function, an expansion as in (8) still holds by using a Schauder decomposition [73] but it contains in general an infinite number of terms, resulting in an infinite number of observables {Ai}\{A_{i}\} with a clear loss of dimension reduction of the system. However, there exist numerous rigorous techniques to approximate in a controlled manner a general kernel 𝐊(𝐱,𝐲)\mathbf{K}(\mathbf{x},\mathbf{y}) in terms of suitable separable kernels, see [62, Ch. 11][74]. Such approximating methods would be the basis for the development of phenomenological theories of interacting systems with suitable separable interactions for which the identification of a finite number of observables can be performed exactly.
On the one hand, the investigation of mean field models is very relevant as it provides a way of gaining a mathematical and analytical understanding of the smooth and critical features of ensembles of interacting agents. On the other hand, the lack of an underlying geometrical structure might pose a too restrictive limitation to the applications of these models. Future research in this direction will involve the generalisation of this framework to systems with a microscopic network structure [75] and in particular to all systems amenable to a description in terms of graphon [76, 77] and graphop theory [78, 79]. We envision that this framework can be extended to include the physically relevant cases where the dynamics is not assumed to be overdamped, in the direction of understanding nonequilibrium ensembles in statistical mechanics [80] and to non-Markovian interacting particles [81], where appropriate Markovian closures have to be considered [44, 82].

VL acknowledges the support received by the European Union’s Horizon 2020 research and innovation program through the project TiPES (Grant Agreement No. 820970) and Marie Curie ITN CriticalEarth (Grant Agreement No. 956170). The work of GP was partially funded by the EPSRC, grant number EP/P031587/1, by J.P. Morgan Chase & Co through a Faculty Research Award 2019 and 2021 and by Advanced ERC grant Machine-Aided General Framework for Fluctuating Dynamic Density Functional Theory. NZ has been supported by the Wallenberg Initiative on Networks and Quantum Information (WINQ).

Appendix A Dynamical Properties of the Kuramoto model

We provide here the detailed calculations leading to the results reported in the main text regarding the Kuramoto model. This model describes an ensemble of one dimensional (M=1M=1) phase oscillators on the torus 𝕋1=[0,2π]\mathbb{T}^{1}=[0,2\pi] whose dynamics is determined by the following Stochastic Differential Equations

dxk=[hkθNj=1Nsin(xkxj)]dt+σdWk,k=1,,N\mathrm{d}x_{k}=\big{[}h_{k}-\frac{\theta}{N}\sum_{j=1}^{N}\sin\left(x_{k}-x_{j}\right)\big{]}\mathrm{d}t+\sigma\mathrm{d}W_{k},\quad k=1,\dots,N (28)

where hkμh_{k}\sim\mu, with μ(h)=μ(h)\mu(-h)=\mu(h), represents the intrinsic frequencies of each oscillator and the symbols are defined as in the main text. The interaction kernel K(x,y)=sin(xy)=sin(x)cos(y)cos(x)sin(y)K(x,y)=\sin(x-y)=\sin(x)\cos(y)-\cos(x)\sin(y) is separable and leads, in the thermodynamic limit N+N\to+\infty [16, 59], to the following nonlinear nonlocal Fokker Planck equation of nonlinearity dimension p=2p=2

tρ=x(h𝒦(x,𝐀))ρ+σ22xxρ=𝐀ρ\partial_{t}\rho=-\partial_{x}\big{(}h-\mathcal{K}(x,\langle\langle\mathbf{A}\rangle\rangle)\big{)}\rho+\frac{\sigma^{2}}{2}\partial_{xx}\rho=\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle}\rho (29)

where 𝒦(x,𝐀)=sin(x)A1(x)cos(x)A2(x)\mathcal{K}(x,\langle\langle\mathbf{A}\rangle\rangle)=\sin(x)\langle\langle A_{1}(x)\rangle\rangle-\cos(x)\langle\langle A_{2}(x)\rangle\rangle where 𝐀(x)=(cos(x),sin(x))\mathbf{A}(x)=\left(\cos(x),\sin(x)\right) represents the set of reaction coordinates for the Kuramoto model. It is easy to verify that the uniform solution ρ0=12π\rho_{0}=\frac{1}{2\pi} is always a stationary solution of equation (29) corresponding to a disordered state characterised by a set of observables 𝐀0=(0,0)\langle\langle\mathbf{A}\rangle\rangle_{0}=(0,0). Moreover, this stationary solution ρ0\rho_{0} satisfies 0ρ0=0\mathcal{L}_{0}\rho_{0}=0 where

0:=𝐀0=hx+σ22xx\mathcal{L}_{0}:=\mathcal{L}_{\langle\langle\mathbf{A}\rangle\rangle_{0}}=-h\partial_{x}+\frac{\sigma^{2}}{2}\partial_{xx} (30)

Below we investigate the response properties of the disordered state and detect phase transitions of the system by looking at the singularities of the susceptibility of the system originating from non invertibility properties of the matrix Pij(ω)P_{ij}(\omega), see equation (12)(12) in the main text. We evaluate the matrix 𝐉2\mathbf{J}\in\mathbb{R}^{2} (here a vector since M=1M=1)

𝐉=𝒦𝐀=(sin(x),cos(x))\mathbf{J}=\frac{\partial\mathcal{K}}{\partial\langle\langle\mathbf{A}\rangle\rangle}=\left(\sin(x),-\cos(x)\right) (31)

Considering that 𝐉(x)=𝐀(x)\partial\mathbf{J}(x)=\mathbf{A}(x), the microscopic Green’s function Yij(t)Y_{ij}(t) can be written as Yij(t)=dhμ(h)Yij(t;h)Y_{ij}(t)=\int\mathrm{d}h\mu(h)Y_{ij}(t;h) where

Yij(t;h)=Θ(t)02πdx2πAi(x)et0Aj(x)Y_{ij}(t;h)=\Theta(t)\int_{0}^{2\pi}\frac{\mathrm{d}x}{2\pi}A_{i}(x)e^{t\mathcal{L}_{0}}A_{j}(x) (32)

Now, it is easy to verify that

𝐀=Dh2+D20𝝍\mathbf{A}=-\frac{D}{h^{2}+D^{2}}\mathcal{L}_{0}\boldsymbol{\psi} (33)

where D=σ22D=\frac{\sigma^{2}}{2} and

𝝍(x;h)=(cos(x)+hDsin(x)sin(x)hDcos(x))=𝐀(x)hDx𝐀(x)\boldsymbol{\psi}(x;h)=\begin{pmatrix}\cos(x)+\frac{h}{D}\sin(x)\\ \sin(x)-\frac{h}{D}\cos(x)\end{pmatrix}=\mathbf{A}(x)-\frac{h}{D}\partial_{x}\mathbf{A}(x) (34)

The microscopic Green Function can then be written as

Yij(t;h)=Dh2+D2Θ(t)02πdx2πAi(x)et00ψj(x;h)==Dh2+D2Θ(t)ddt02πdx2πAi(x)et0ψj(x;h)=:=Dh2+D2Θ(t)ddtCij(t;h)\begin{split}Y_{ij}(t;h)&=-\frac{D}{h^{2}+D^{2}}\Theta(t)\int_{0}^{2\pi}\frac{\mathrm{d}x}{2\pi}A_{i}(x)e^{t\mathcal{L}_{0}}\mathcal{L}_{0}\psi_{j}(x;h)=\\ &=-\frac{D}{h^{2}+D^{2}}\Theta(t)\frac{\mathrm{d}}{\mathrm{d}t}\int_{0}^{2\pi}\frac{\mathrm{d}x}{2\pi}A_{i}(x)e^{t\mathcal{L}_{0}}\psi_{j}(x;h)=\\ &:=-\frac{D}{h^{2}+D^{2}}\Theta(t)\frac{\mathrm{d}}{\mathrm{d}t}C_{ij}(t;h)\end{split} (35)

where Cij(t;h)C_{ij}(t;h) is the correlation function between observable Ai(x)A_{i}(x) and ψj(x;h)\psi_{j}(x;h). The corresponding microscopic susceptibility is then given by

Yij(ω;h)=+eiωtYij(t)dt=Dh2+D2(Cij(t=0;h)+iωCij(ω;h))Y_{ij}(\omega;h)=\int_{-\infty}^{+\infty}e^{i\omega t}Y_{ij}(t)\mathrm{d}t=\frac{D}{h^{2}+D^{2}}\left(C_{ij}(t=0;h)+i\omega C_{ij}(\omega;h)\right) (36)

where Cij(ω;h)=0+eiωtCij(t;h)dtC_{ij}(\omega;h)=\int_{0}^{+\infty}e^{i\omega t}C_{ij}(t;h)\mathrm{d}t is the (one sided) Fourier Transform of Cij(t;h)C_{ij}(t;h). Since the correlation function at time t=0t=0 is Cij(t=0;h)=12δijh2D(1δij)C_{ij}(t=0;h)=\frac{1}{2}\delta_{ij}-\frac{h}{2D}(1-\delta_{ij}) the microscopic susceptibility associated to Yij(t)Y_{ij}(t) is

Yij(ω)=Yij(ω;h)μ(h)dh=D2δijμ(h)D2+h2dh+iωDCij(ω;h)h2+D2μ(h)dhY_{ij}(\omega)=\int Y_{ij}(\omega;h)\mu(h)\mathrm{d}h=\frac{D}{2}\delta_{ij}\int\frac{\mu(h)}{D^{2}+h^{2}}\mathrm{d}h+i\omega D\int\frac{C_{ij}(\omega;h)}{h^{2}+D^{2}}\mu(h)\mathrm{d}h (37)

where we have used the fact that the distribution of frequencies is even. The matrix Pij(ω)P_{ij}(\omega) describing the response to the observable Ai(x)\langle\langle A_{i}(x)\rangle\rangle is given by

Pij(ω)=δijθYij(ω)=(1θD2μ(h)D2+h2dh)δijiθωDCij(ω;h)h2+D2μ(h)dhP_{ij}(\omega)=\delta_{ij}-\theta Y_{ij}(\omega)=\left(1-\frac{\theta D}{2}\int\frac{\mu(h)}{D^{2}+h^{2}}\mathrm{d}h\right)\delta_{ij}-i\theta\omega D\int\frac{{C}_{ij}(\omega;h)}{h^{2}+D^{2}}\mu(h)\mathrm{d}h (38)

A.0.1 Evaluation of the Correlation Function

The correlation function Cij(t;h)C_{ij}(t;h) defined in (35) is completely determined by the spectral properties of the operator 0\mathcal{L}_{0} which is a linear differential operator with constant coefficients, see equation (30). As such, its spectral features on the space of functions that are square integrable with respect to the invariant measure L2(𝕋;ρ0)L^{2}(\mathbb{T};\rho_{0}) is easily found to be given by the eigenfunctions ϕk=eikx\phi_{k}=e^{ikx} with relative eigenvalue λk=ihkk2D\lambda_{k}=-ihk-k^{2}D where k=0,±1,±2,k=0,\pm 1,\pm 2,\dots . We remark that the eigenfunctions ϕk\phi_{k} are orthonormal in L2(𝕋;ρ0)L^{2}(\mathbb{T};\rho_{0}) since ϕk|ϕk0=02πϕk(x)ϕk(x)ρ0=12π02πei(kk)x=δkk\langle\phi_{k}|\phi_{k^{\prime}}\rangle_{0}=\int_{0}^{2\pi}\phi_{k}^{*}(x)\phi_{k^{\prime}}(x)\rho_{0}=\frac{1}{2\pi}\int_{0}^{2\pi}e^{i(k^{\prime}-k)x}=\delta_{kk^{\prime}} where |0\langle\cdot|\cdot\rangle_{0} represents the L2(𝕋,ρ)L^{2}(\mathbb{T},\rho) scalar product. We can then obtain a spectral decomposition [46, 83] of the operator et0e^{t\mathcal{L}_{0}} as

et0=k=+etλkΠke^{t\mathcal{L}_{0}}=\sum_{k=-\infty}^{+\infty}e^{t\lambda_{k}}\Pi_{k} (39)

where Πk=|ϕkϕk|\Pi_{k}=|\phi_{k}\rangle\langle\phi_{k}| is the projector onto the eigenspace in L2(𝕋;ρ0)L^{2}(\mathbb{T};\rho_{0}) relative to the eigenvalue λk\lambda_{k}. We then obtain

Cij(t;h)=k=+etλkdxAi(x)Πkψj(x;h)=k=+etλkAiρ0|ϕk0ϕk|ψjρ00:=k=+etλkαij(k)(h)\begin{split}C_{ij}(t;h)&=\sum_{k=-\infty}^{+\infty}e^{t\lambda_{k}}\int\mathrm{d}xA_{i}(x)\Pi_{k}\psi_{j}(x;h)=\sum_{k=-\infty}^{+\infty}e^{t\lambda_{k}}\langle\frac{A_{i}}{\rho_{0}}|\phi_{k}\rangle_{0}\langle\phi_{k}|\psi_{j}\rho_{0}\rangle_{0}:=\sum_{k=-\infty}^{+\infty}e^{t\lambda_{k}}\alpha_{ij}^{(k)}(h)\end{split} (40)

Since λk=λk\lambda_{-k}=\lambda_{k}^{*} and ϕk=ϕk\phi_{-k}=\phi_{k}^{*}, the coefficients αij(k)\alpha_{ij}^{(k)} satisfy αij(k)=(αij(k))\alpha_{ij}^{(-k)}=\left(\alpha_{ij}^{(k)}\right)^{*} and the correlation function can be written as

Cij(t;h)=2𝐑𝐞k=1+eλktαij(k)C_{ij}(t;h)=2\mathbf{Re}\sum_{k=1}^{+\infty}e^{\lambda_{k}t}\alpha_{ij}^{(k)} (41)

where the mode k=0k=0 does not contribute to the correlation function, as Π0\Pi_{0} projects onto the invariant measure ρ0=12π\rho_{0}=\frac{1}{2\pi} yielding aij(0)=0a_{ij}^{(0)}=0. Given the particular structure of the coupling kernel K(x,y)=sin(xy)K(x,y)=sin(x-y) selecting the reaction coordinates 𝐀(x)\mathbf{A}(x), it is possible to show that only the first k=1k=1 contributes to the correlation functions, since αij(k)=0\alpha_{ij}^{(k)}=0 for k=2,3,k=2,3,\dots . Indeed a simple but lengthy calculation shows that

αij(k)=δk14(δijhD(1δij)Sij+i(((1δij)AijhDδij)))\alpha_{ij}^{(k)}=\frac{\delta_{k1}}{4}\left(\delta_{ij}-\frac{h}{D}\left(1-\delta_{ij}\right)S_{ij}+i\left(\left(\left(1-\delta_{ij}\right)A_{ij}-\frac{h}{D}\delta_{ij}\right)\right)\right) (42)

where 𝐒=(0110)\mathbf{S}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix} and 𝐀=(0110)\mathbf{A}=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix} are two antidiagonal matrices describing the interaction between the observables AiA_{i} and ψj\psi_{j} with iji\neq j. The correlation function can then be written as

Cij(t;h)=2𝐑𝐞(eλ1tαij(1))=δij2eDt(cos(ht)hDsin(ht))+cij(t;h)C_{ij}(t;h)=2\mathbf{Re}\left(e^{\lambda_{1}t}\alpha_{ij}^{(1)}\right)=\frac{\delta_{ij}}{2}e^{-Dt}\left(\cos(ht)-\frac{h}{D}\sin(ht)\right)+c_{ij}(t;h) (43)

where cij(t;h)c_{ij}(t;h) is an off diagonal contribution and is an odd function of the intrinsic frequency, cij(t;h)=cij(t;h)c_{ij}(t;-h)=-c_{ij}(t;h). Since we have assumed that the intrinsic frequency distribution is even, we observe that equation (38) shows that cij(t;h)c_{ij}(t;h) provides a vanishing contribution to the matrix Pij(ω)P_{ij}(\omega) and in the following calculations it will be neglected. By taking the (one sided) Fourier transform we obtain

Cij(ω;h)=δij2D+h2/D+iωω2(D2+h2)+2iωD:=C(ω;h)δij,C_{ij}(\omega;h)=\frac{\delta_{ij}}{2}\frac{-D+h^{2}/D+i\omega}{\omega^{2}-(D^{2}+h^{2})+2i\omega D}:=C(\omega;h)\delta_{ij}, (44)

where we have defined

C(ω;h)\displaystyle C(\omega;h) =CR(ω;h)+iCI(ω;h)\displaystyle=C_{R}(\omega;h)+iC_{I}(\omega;h) (45a)
CR(ω;h)\displaystyle C_{R}(\omega;h) =D2+h22Dω2+D2h2(ω2(D2+h2))2+4ω2D2,\displaystyle=\frac{D^{2}+h^{2}}{2D}\frac{\omega^{2}+D^{2}-h^{2}}{\left(\omega^{2}-\left(D^{2}+h^{2}\right)\right)^{2}+4\omega^{2}D^{2}}, (45b)
CI(ω;h)\displaystyle C_{I}(\omega;h) =ω2ω2+D23h2(ω2(D2+h2))2+4ω2D2.\displaystyle=\frac{\omega}{2}\frac{\omega^{2}+D^{2}-3h^{2}}{\left(\omega^{2}-\left(D^{2}+h^{2}\right)\right)^{2}+4\omega^{2}D^{2}}. (45c)

From equation (38) and (44) we note that the renormalisation matrix is diagonal Pij(ω)=P(ω)δijP_{ij}(\omega)=P(\omega)\delta_{ij}, with

P(ω)=1θD2μ(h)D2+h2(1+2iωC(ω;h))dh.P(\omega)=1-\frac{\theta D}{2}\int\frac{\mu(h)}{D^{2}+h^{2}}\left(1+2i\omega C\left(\omega;h\right)\right)\mathrm{d}h. (46)

A.0.2 Phase transitions for the Kuramoto model

As explained in the main text, phase transitions are associated to settings where the matrix Pij(ω)P_{ij}(\omega^{*}) is not invertible for ω\omega^{*}\in\mathbb{R}. Given equation (46), this is equivalent to the condition P(ω)=0P(\omega^{*})=0. Now we separate real and imaginary part of the above equation and, considering (45a) and (46) we get, respectively,

Dθ2\displaystyle\frac{D\theta}{2} μ(h)h2+D2(12ωCI(ω;h))dh=1,\displaystyle\int\frac{\mu(h)}{h^{2}+D^{2}}\left(1-2\omega^{*}C_{I}(\omega^{*};h)\right)\mathrm{d}h=1, (47a)
ω\displaystyle\omega^{*} μ(h)h2+D2CR(ω;h)dh=0.\displaystyle\int\frac{\mu(h)}{h^{2}+D^{2}}C_{R}(\omega^{*};h)\mathrm{d}h=0. (47b)

Equation (47b) provides the value of ω\omega^{*} in terms of the parameters of the system and the properties of the quenched disorder distribution μ\mu whereas (47a) gives the associated transition point, that is a relationship among the parameters of the system (θ,D)(\theta,D) and the properties of μ\mu. In general, we can identify two typical scenarios of phase transitions: static phase transitions associated to a single simple pole ω=0\omega^{*}=0 and dynamic phase transitions associated to an opposite pair of pole ω1=ω2=ω>0\omega_{1}=-\omega_{2}=\omega^{*}>0.

A.0.3 Static phase transitions

It is immediate to see that ω=ω0=0\omega^{*}=\omega_{0}=0 is a solution of (47b). Evaluating (47a) for ω=ω0=0\omega^{*}=\omega_{0}=0 yields

1θD2μ(h)h2+D2dh=01-\frac{\theta D}{2}\int\frac{\mu(h)}{h^{2}+D^{2}}\mathrm{d}h=0 (48)

from which we obtain the critical value of the coupling strength

θcstat=2[Dh2+D2μ(h)dh]1\theta^{stat}_{c}=2\bigg{[}\int\frac{D}{h^{2}+D^{2}}\mu(h)\mathrm{d}h\bigg{]}^{-1} (49)

We remark that this phase transition is characterised by a vanishing simple pole ω0=0\omega_{0}=0, resulting in a static transition where the static susceptibility 𝝌~(0)=(𝐏1𝝌)(0)\tilde{\boldsymbol{\chi}}(0)=\left(\mathbf{P}^{-1}\boldsymbol{\chi}\right)\left(0\right) diverges.

A.0.4 Dynamic phase transitions: properties of μ\mu

The investigation of the onset of a dynamical phase transition is more complicated as it requires the study of frequency dependent properties of C(ω;h)C(\omega;h). Since we are considering a dynamic phase transition characterised by ω0\omega^{*}\neq 0, equation (47b) can be written as

μ(h)h2+D2CR(ω;h)dh=12Dω2+D2h2(ω2(D2+h2))2+4ω2D2μ(h)=0.\int\frac{\mu(h)}{h^{2}+D^{2}}C_{R}(\omega^{*};h)\mathrm{d}h=\frac{1}{2D}\int\frac{\omega^{2}+D^{2}-h^{2}}{\left(\omega^{2}-\left(D^{2}+h^{2}\right)\right)^{2}+4\omega^{2}D^{2}}\mu(h)=0. (50)

where in the last equality we have used (45b). We can prove that (50) implies a known result for the Kuramoto model by noticing that such equation contains the notion that, if ω0\omega^{*}\neq 0 is a solution, then ω-\omega^{*} is also a solution. Indeed it is easy to show that for ω0\omega\neq 0 the following identity holds

ω2+D2h2(ω2(D2+h2))2+4ω2D2=12ω(ω+hD2+(ω+h)2+ωhD2+(ωh)2)\frac{\omega^{2}+D^{2}-h^{2}}{\left(\omega^{2}-\left(D^{2}+h^{2}\right)\right)^{2}+4\omega^{2}D^{2}}=\frac{1}{2\omega}\left(\frac{\omega+h}{D^{2}+(\omega+h)^{2}}+\frac{\omega-h}{D^{2}+(\omega-h)^{2}}\right) (51)

so that (50) implies that

(ω+hD2+(ω+h)2+ωhD2+(ωh)2)μ(h)dh=0.\int\left(\frac{\omega+h}{D^{2}+(\omega+h)^{2}}+\frac{\omega-h}{D^{2}+(\omega-h)^{2}}\right)\mu(h)\mathrm{d}h=0. (52)

Considering that μ(h)\mu(h) is an even function the previous equation implies that

ω±hD2+(ω±h)2μ(h)dh=0,\int\frac{\omega^{*}\pm h}{D^{2}+(\omega^{*}\pm h)^{2}}\mu(h)\mathrm{d}h=0, (53)

which is a known formula for the Kuramoto model [66]. The above equation shows that the existence of a nonvanishing singularity ω\omega^{*} is closely related to the properties of the intrinsic frequency distribution μ(h)\mu(h). In particular, if μ(h)\mu(h) is a unimodal function then there does not exist a ω0\omega^{*}\neq 0 satisfying (53). In other words, an even unimodal distribution cannot support the development of a dynamic phase transition.

A.0.5 Dynamic phase transition for a bimodal frequency distribution

The simplest setting that can support a dynamic phase transition scenario is represented by a bimodal distribution

μh0(h)=δ(h+h0)+δ(hh0)2.\mu_{h_{0}}(h)=\frac{\delta(h+h_{0})+\delta(h-h_{0})}{2}. (54)

In this case, (50) becomes simply

CR(ω;h0)=0C_{R}(\omega^{*};h_{0})=0 (55)

that, given (45b), yields a solution ω=h02D2\omega^{*}=\sqrt{h_{0}^{2}-D^{2}}. This solution ω\omega^{*} is real only if h0>Dh_{0}>D, that is, a sufficiently large separation of the peaks of the bimodal distribution is needed, as observed in [69, 65]. The associated phase transition point is given by (47a), which reads for a bimodal distribution

θD2(12ωCI(ω;h0))=h02+D2\frac{\theta D}{2}\left(1-2\omega^{*}C_{I}(\omega^{*};h_{0})\right)=h_{0}^{2}+D^{2} (56)

We first evaluate the denominator of CI(ω;h0)C_{I}(\omega^{*};h_{0})

(ω2(D2+h02))2+4ω2D2=(2D2)2+4D2ω2=4D2(D2+ω2)\left(\omega_{*}^{2}-\left(D^{2}+h_{0}^{2}\right)\right)^{2}+4\omega_{*}^{2}D^{2}=(2D^{2})^{2}+4D^{2}\omega_{*}^{2}=4D^{2}(D^{2}+\omega_{*}^{2}) (57)

where for the sake of notation clarity we have defined ω2:=(ω)2=h02D2\omega_{*}^{2}:=(\omega^{*})^{2}=h_{0}^{2}-D^{2}. We can then write from (45c)

CI(ω;h0)=ω2ω2+D23h02(ω2(D2+h02))2+4ω2D2=ωh024D2(D2+ω2)=ω4D2C_{I}(\omega^{*};h_{0})=\frac{\omega^{*}}{2}\frac{\omega_{*}^{2}+D^{2}-3h_{0}^{2}}{\left(\omega_{*}^{2}-\left(D^{2}+h_{0}^{2}\right)\right)^{2}+4\omega_{*}^{2}D^{2}}=-\omega^{*}\frac{h_{0}^{2}}{4D^{2}(D^{2}+\omega_{*}^{2})}=-\frac{\omega^{*}}{4D^{2}} (58)

where in the last equality we have used the fact that by definition of ω\omega^{*} we have that D2+ω2=h02D^{2}+\omega_{*}^{2}=h_{0}^{2}. Finally, equation (56) results in

θD2(1+ω22D2)=h02+D2,θD22D2+ω22D2=h02+D2,θD2h02+D22D2=h02+D2,θD212D2=1,\begin{split}\frac{\theta D}{2}\left(1+\frac{\omega^{2}}{2D^{2}}\right)&=h_{0}^{2}+D^{2},\\ \frac{\theta D}{2}\frac{2D^{2}+\omega_{*}^{2}}{2D^{2}}&=h_{0}^{2}+D^{2},\\ \frac{\theta D}{2}\frac{h_{0}^{2}+D^{2}}{2D^{2}}&=h_{0}^{2}+D^{2},\\ \frac{\theta D}{2}\frac{1}{2D^{2}}&=1,\end{split} (59)

yielding eventually the critical phase transition point θcdyn=4D\theta_{c}^{dyn}=4D.

A.1 Characterisation of phase tranisitions: residues

The response properties of the system to external perturbations are characterised by the total susceptibility

χ~i(ω)=(𝐏1𝝌)i(ω)=P1(ω)χi(ω),\tilde{\chi}_{i}(\omega)=\left(\mathbf{P}^{-1}\boldsymbol{\chi}\right)_{i}\left(\omega\right)=P^{-1}(\omega)\chi_{i}(\omega), (60)

where in the last equality we have used the fact that Pij(ω)=P(ω)δijP_{ij}(\omega)=P(\omega)\delta_{ij}, with P(ω)P(\omega) given by (46), when the distribution of frequencies μ(h)\mu(h) is an even function. Since the microscopic Green’s Function Gi(t)G_{i}(t) is given by, see equation (10a)(10a) in the main text,

Gi(t)=dhμ(h)dxAi(x)et0pρ0,G_{i}(t)=\int\mathrm{d}h\mu(h)\int\mathrm{d}xA_{i}(x)e^{t\mathcal{L}_{0}}\mathcal{L}_{p}\rho_{0}, (61)

where pρ0=x(X(x)ρ0)\mathcal{L}_{p}\rho_{0}=-\partial_{x}\left(X(x)\rho_{0}\right), we see that a homogeneous perturbation X(x)=constX(x)=const would result in a null response Gi(t)=0G_{i}(t)=0, and consequently χi(ω)=0\chi_{i}(\omega)=0, given that the stationary measure is the uniform distribution ρ=12π\rho=\frac{1}{2\pi}. We have chosen to perform the numerical response simulations with X(x)=sin(x)=A2(x)X(x)=-\sin(x)=A_{2}(x), so that the microscopic Green’s Function is Gi(t)=Yi1(t)G_{i}(t)=Y_{i1}(t) and the corresponding susceptibility is χi(ω)=Yi1(ω)\chi_{i}(\omega)=Y_{i1}(\omega).

A.1.1 Residue for static phase transitions

A static phase transition is characterised by a simple pole at ω=ω=0\omega=\omega^{*}=0, so that the critical response of the reaction coordinate A1(x)A_{1}(x) can be written as

χ~1(ω)=Y11(ω)P(ω)=Y11(0)P(0)ω+ψ(ω)=1/θP(0)ω+ψ(ω):=κω+ψ(ω),\tilde{\chi}_{1}(\omega)=\frac{Y_{11}(\omega)}{P(\omega)}=\frac{Y_{11}(0)}{P^{\prime}(0)\omega}+\psi(\omega)=\frac{1/\theta}{P^{\prime}(0)\omega}+\psi(\omega):=\frac{\kappa}{\omega}+\psi(\omega), (62)

where ψ(ω)\psi(\omega) is an analytic function in the upper complex ω\omega^{*} plane and we have used the fact that, at a static phase transition, P(0)=0P(0)=0 and Y11(0)=(1P(0))/θ=1/θY_{11}(0)=(1-P(0))/\theta=1/\theta. The quantity κ=1/θP(0)\kappa=\frac{1/\theta}{P^{\prime}(0)} represents the residue of the pole ω=0\omega^{*}=0 and determines the properties of the singularity. From (46) we evaluate the derivative of P(ω)P(\omega)

P(ω)=iθDμ(h)D2+h2(C(ω;h)+ωC(ω;h))dhP^{\prime}(\omega)=-i\theta D\int\frac{\mu(h)}{D^{2}+h^{2}}\left(C(\omega;h)+\omega C^{\prime}(\omega;h)\right)\mathrm{d}h (63)

For a static phase transition we are interested in

P(0)=iθDμ(h)D2+h2C(0;h)dh=iθ2μ(h)D2h2(D2+h2)2dhP^{\prime}(0)=-i\theta D\int\frac{\mu(h)}{D^{2}+h^{2}}C(0;h)\mathrm{d}h=-\frac{i\theta}{2}\int\mu(h)\frac{D^{2}-h^{2}}{\left(D^{2}+h^{2}\right)^{2}}\mathrm{d}h (64)

where in the last equality we have used that C(0;h)=CR(0;h)+iCI(0;h)=CR(0;h)=12DD2h2D2+h2C(0;h)=C_{R}(0;h)+iC_{I}(0;h)=C_{R}(0;h)=\frac{1}{2D}\frac{D^{2}-h^{2}}{D^{2}+h^{2}}. The residue is thus

κ=2iθ2μ(h)D2h2(D2+h2)2dh=i(Dh2+D2μ(h)dh)22D2h2(D2+h2)2μ(h)dh.\kappa=\frac{2i}{\theta^{2}\int\mu(h)\frac{D^{2}-h^{2}}{\left(D^{2}+h^{2}\right)^{2}}\mathrm{d}h}=i\frac{\left(\int\frac{D}{h^{2}+D^{2}}\mu(h)\mathrm{d}h\right)^{2}}{2\int\frac{D^{2}-h^{2}}{\left(D^{2}+h^{2}\right)^{2}}\mu(h)\mathrm{d}h}. (65)

where we have taken into account that, at the phase transition, θ=θcstat\theta=\theta_{c}^{stat} where θcstat\theta_{c}^{stat} is given by (49). The above expression shows that the residue for a static phase transition is completely imaginary.
In the main text, we have investigated the static phase transition arising in the homogeneous Kuramoto model characterised by a delta-distribution of frequency μ(h)=δ(h)\mu(h)=\delta(h). It is easy to see from (65) that, in this case, the residue is independent of the parameter of the system and equals to κ=i2\kappa=\frac{i}{2}.

A.1.2 Residue for dynamic phase transitions

We here investigate the residue for dynamic phase transitions characterised by the development of singularities of the susceptibility for a pair of opposite real frequencies ω1=ω2=ω>0\omega_{1}=-\omega_{2}=\omega^{*}>0.
A similar argument of the previous section holds but it has to be modified to take into account both poles. We have that

χ~1(ω)=Y11(ω)P(ω)=Y11(ω)P(ω)(ωω)+Y11(ω)P(ω)(ω+ω)+ψ(ω)\tilde{\chi}_{1}(\omega)=\frac{Y_{11}(\omega)}{P(\omega)}=\frac{Y_{11}(\omega^{*})}{P^{\prime}(\omega^{*})\left(\omega-\omega^{*}\right)}+\frac{Y_{11}(-\omega^{*})}{P^{\prime}(-\omega^{*})\left(\omega+\omega^{*}\right)}+\psi(\omega) (66)

where ψ(ω)\psi(\omega) is an analytic function in the upper complex ω\omega plane. As before we define the residue κ=Y11(ω)P(ω)\kappa=\frac{Y_{11}(\omega)}{P^{\prime}(\omega^{*})} and we observe that, since the susceptibility, being the Fourier transform of a real function, has to satisfy the condition χ~i(ω)=χ~i(ω)\tilde{\chi}_{i}(-\omega)=\tilde{\chi}_{i}^{*}(\omega). This implies that Y11(ω)P(ω)=κ\frac{Y_{11}(-\omega^{*})}{P^{\prime}(-\omega^{*})}=-\kappa^{*} so that the previous equation can be written as

χ~1(ω)=κωωκω+ω+ψ(ω),\tilde{\chi}_{1}(\omega)=\frac{\kappa}{\omega-\omega^{*}}-\frac{\kappa^{*}}{\omega+\omega^{*}}+\psi(\omega), (67)

Similarly to the previous section Y11(ω)=1/θY_{11}(\omega^{*})=1/\theta and the residue is

κ=1θP(ω)\kappa=\frac{1}{\theta P^{\prime}(\omega^{*})} (68)

As opposed to the static phase transition, the evaluation of the residue now requires a more careful study of P(ω)P^{\prime}(\omega^{*}) where ω>0\omega^{*}>0.
We are here interested in the evaluation of the residue for the dynamic phase transition arising from the bimodal distribution μ0\mu_{0} discussed in the previous sections. We have seen that such transition is associated to a pole ω=h02D2\omega^{*}=\sqrt{h_{0}^{2}-D^{2}} arising at the transition point θcdyn=4D\theta_{c}^{dyn}=4D. Evaluating (63) for a bimodal distribution μ0\mu_{0} and separating real and imaginary part we obtain the following equations

P(ω)\displaystyle P^{\prime}(\omega) =PR(ω)+iPI(ω)\displaystyle=P_{R}^{\prime}(\omega)+iP_{I}^{\prime}(\omega) (69a)
PR(ω)\displaystyle P_{R}^{\prime}(\omega^{*}) =θD(D2+h02)(CI(ω;h0)+ωCI(ω;h0))\displaystyle=\frac{\theta D}{\left(D^{2}+h_{0}^{2}\right)}\left(C_{I}(\omega^{*};h_{0})+\omega^{*}C_{I}^{\prime}(\omega^{*};h_{0})\right) (69b)
PI(ω)\displaystyle P_{I}^{\prime}(\omega^{*}) =θD(D2+h02)(CR(ω;h0)+ωCR(ω;h0))=θD(D2+h02)ωCR(ω;h0)\displaystyle=-\frac{\theta D}{\left(D^{2}+h_{0}^{2}\right)}\left(C_{R}(\omega^{*};h_{0})+\omega^{*}C_{R}^{\prime}(\omega^{*};h_{0})\right)=-\frac{\theta D}{\left(D^{2}+h_{0}^{2}\right)}\omega^{*}C_{R}^{\prime}(\omega^{*};h_{0}) (69c)

where we have taken into account (45a) and in the last equality we have imposed condition (55). We can evaluate the derivative of C(ω;h0)C(\omega;h_{0}) from equations (45a),(45b) and (45c) with similar arguments. We omit the details, as it is just algebra, but we mention that we make extensive use of the fact that ω2:=(ω)2=h02+D2\omega^{2}_{*}:=(\omega^{*})^{2}=h_{0}^{2}+D^{2}. The result is

C(ω;h0)\displaystyle C^{\prime}(\omega^{*};h_{0}) =CR(ω;h0)+iCI(ω;h0)\displaystyle=C_{R}^{\prime}(\omega^{*};h_{0})+iC_{I}^{\prime}(\omega^{*};h_{0}) (70a)
CR(ω;h0)\displaystyle C_{R}^{\prime}(\omega^{*};h_{0}) =ω4D3D2+h02D2+ω2\displaystyle=\frac{\omega^{*}}{4D^{3}}\frac{D^{2}+h_{0}^{2}}{D^{2}+\omega_{*}^{2}} (70b)
CI(ω;h0)\displaystyle C_{I}^{\prime}(\omega^{*};h_{0}) =14(D2+ω2)\displaystyle=-\frac{1}{4\left(D^{2}+\omega_{*}^{2}\right)} (70c)

From the above equations it is immediate to evaluate

PI(ω)=ω2D2+ω2θ4D2=1Dω2D2+ω2P_{I}^{\prime}(\omega^{*})=-\frac{\omega_{*}^{2}}{D^{2}+\omega_{*}^{2}}\frac{\theta}{4D^{2}}=-\frac{1}{D}\frac{\omega_{*}^{2}}{D^{2}+\omega_{*}^{2}} (71)

where we have used that θ=θcdyn=4D\theta=\theta_{c}^{dyn}=4D. We can then proceed to calculate PR(ω)P^{\prime}_{R}(\omega^{*}). Taking into account (58) and (70c) we get

PR(ω)=θD(D2+h02)(ω4D2ω4(D2+ω2))=θω(2D2+ω2)4D(D2+h02)(D2+ω2)=ωD2+ω2P^{\prime}_{R}(\omega^{*})=\frac{\theta D}{\left(D^{2}+h_{0}^{2}\right)}\left(-\frac{\omega^{*}}{4D^{2}}-\frac{\omega^{*}}{4\left(D^{2}+\omega_{*}^{2}\right)}\right)=-\frac{\theta\omega^{*}\left(2D^{2}+\omega_{*}^{2}\right)}{4D\left(D^{2}+h_{0}^{2}\right)\left(D^{2}+\omega_{*}^{2}\right)}=-\frac{\omega_{*}}{D^{2}+\omega_{*}^{2}} (72)

where in the last equality we have used that, by the definition of ω\omega^{*}, 2D2+ω2=D2+h022D^{2}+\omega_{*}^{2}=D^{2}+h_{0}^{2} and that θ=θcdyn=4D\theta=\theta_{c}^{dyn}=4D. As a result, we can write that

P(ω)=PR(ω)+PI(ω)=ωD2+ω2(1+iωD)P^{\prime}(\omega^{*})=P_{R}^{\prime}(\omega^{*})+P_{I}^{\prime}(\omega^{*})=-\frac{\omega_{*}}{D^{2}+\omega_{*}^{2}}\left(1+i\frac{\omega_{*}}{D}\right) (73)

Since θ=θcdyn=4D\theta=\theta_{c}^{dyn}=4D, the residue (68) turns out to be for the dynamic phase transition associated to the bimodal frequency distribution

κ=κR+iκI=D4ω+i4\kappa=\kappa_{R}+i\kappa_{I}=-\frac{D}{4\omega^{*}}+\frac{i}{4} (74)

where DD satisfies the property θ=4D\theta=4D identifying the phase transition in the parameter space (θ,D)(\theta,D). As opposed to the static transition, the residue for the dynamic phase transition is not completely imaginary but has also a real part that depends on both the strength of the noise and the critical frequency. Consequently, such singular behaviour is fundamentally different from the static transition as will be highlighted in the next section.

A.2 Response for finite NN: singularities and resonances

In order to numerically evaluate the response properties of the finite system we perform, following [43], nn simulations of equations (29) where the initial conditions are sampled according to the unperturbed invariant measure ρ0=12π\rho_{0}=\frac{1}{2\pi} and where at time t=0t=0 we apply a perturbation proportional to a Dirac δ(t)\delta(t) function. We remark that this is a macroscopic perturbation where each agent is perturbed according to F(x;h)F(x;h)+εX(x)δ(t)F(x;h)\to F(x;h)+\varepsilon X(x)\delta(t), where F(x;h)=hF(x;h)=h for the Kuramoto model. The average of the response Ai=Ai0+εAi1=εAi1\langle\langle A_{i}\rangle\rangle=\langle\langle A_{i}\rangle\rangle_{0}+\varepsilon\langle\langle A_{i}\rangle\rangle_{1}=\varepsilon\langle\langle A_{i}\rangle\rangle_{1} over the nn simulations gives an estimate G~i\tilde{G}_{i} of the macroscopic response function defined since

Ai1(t)=(G~iT)(t)=(G~iδ)(t)=G~i(t)\langle\langle A_{i}\rangle\rangle_{1}(t)=\left(\tilde{G}_{i}\star T\right)(t)=\left(\tilde{G}_{i}\star\delta\right)(t)=\tilde{G}_{i}(t) (75)

where we remark that Ai0=0\langle\langle A_{i}\rangle\rangle_{0}=0 for the Kuramoto model. Convergence to the linear response regime has been assessed by performing simulations with different values of ε\varepsilon. The total susceptibility χ~i(ω)\tilde{\chi}_{i}(\omega) of the system is obtained as

χ~i(ω)=+G~i(t)eiωtdt\tilde{\chi}_{i}(\omega)=\int_{-\infty}^{+\infty}\tilde{G}_{i}(t)e^{i\omega t}\mathrm{d}t (76)

For finite NN we observe a mollifying effect of the singular behaviour of the susceptibility [31] where the poles ω\omega^{*} gets slightly shifted towards the lower half of the complex plane, namely ±ω±ωiγ(N)\pm\omega^{*}\to\pm\omega^{*}-i\gamma(N) where γ(N)0\gamma(N)\to 0 as N+N\to+\infty. Singularities of the susceptibility in the thermodynamic limit correspond to resonances for finite NN characterised by

χ~1(ω)=κωω+iγ(N)+ψ(ω)=κωω+iγ(N)κω+ω+iγ(N)+ψ(ω)\tilde{\chi}_{1}(\omega)=\frac{\kappa}{\omega-\omega^{*}+i\gamma(N)}+\psi(\omega)=\frac{\kappa}{\omega-\omega^{*}+i\gamma(N)}-\frac{\kappa^{*}}{\omega+\omega^{*}+i\gamma(N)}+\psi^{\prime}(\omega) (77)

The resonant behaviour of the susceptibility becomes singular in the thermodynamic limit as, for N+N\to+\infty,

limNχ~N(ω)=iπκδ(ωω)+κ𝒫(1ωω)+ψ(ω).\lim_{N\to\infty}\tilde{\chi}_{N}(\omega)=-i\pi\kappa\delta(\omega-\omega^{*})+\kappa\mathcal{P}\left(\frac{1}{\omega-\omega^{*}}\right)+\psi(\omega). (78)

where 𝒫(1ω)\mathcal{P}(\frac{1}{\omega}) represents the Principal Value distribution. For the static transition, the residue is completely imaginary so that 𝐑𝐞χ~i(ω)\mathbf{Re}\tilde{\chi}_{i}(\omega) provides the δ\delta-diverging behaviour whereas the imaginary part yields the Principal Value behaviour. In particular, the function ω𝐈𝐦χ~i(ω)\omega\mathbf{Im}\tilde{\chi}_{i}(\omega) is constant and equals to |κ||\kappa| in a neighborhood of ω=0\omega^{*}=0. A visual inspection of Panel (a)(a) (bottom inset) of Figure 11 in the main text clearly shows that numerical experiments agree with the theoretical prediction |κ|=12|\kappa|=\frac{1}{2} for the homogeneous case μ(h)=δ(h)\mu(h)=\delta(h).
As for the dynamic transition, the residue has a real part too and the above identification of the two different behaviours of the susceptibility is not as straightforward. More importantly, the non vanishing real part of the residue has strong consequences on the response properties of the system, with the development of a spontaneous phase shift of the response with respect to the forcing. In order to clarify this feature, we evaluate (77) at the critical frequency ω=ω\omega=\omega^{*} and neglect non diverging terms in γ(N)\gamma(N)

χ~1(ω)=κiγ(N)=κIγ(N)iκRγ(N).\tilde{\chi}_{1}(\omega^{*})=\frac{\kappa}{i\gamma(N)}=\frac{\kappa_{I}}{\gamma(N)}-i\frac{\kappa_{R}}{\gamma(N)}. (79)

We now consider a pure sinusoidal time dependent forcing T(t;ω^)=sin(ω^t)T(t;\hat{\omega})=\sin(\hat{\omega}t) and evaluate the response A11\langle\langle A_{1}\rangle\rangle_{1} through the response formula

A11(ω)=χ~1(ω)T(ω)\langle\langle A_{1}\rangle\rangle_{1}(\omega)=\tilde{\chi}_{1}(\omega)T(\omega) (80)

Given that T(ω;ω^)=iπ(δ(ω+ω^)δ(ωω^))T(\omega;\hat{\omega})=-i\pi\left(\delta(\omega+\hat{\omega})-\delta(\omega-\hat{\omega})\right), by taking the anti-Fourier Transform of (80) one obtains the response in the time domain

A11(t;ω^)\displaystyle\langle\langle A_{1}\rangle\rangle_{1}(t;\hat{\omega}) =𝐈𝐦(eiω^tχ~(ω^))=𝐈𝐦χ~(ω^)cos(ω^t)+𝐑𝐞χ~(ω^)sin(ω^t)=\displaystyle=-\mathbf{Im}\left(e^{-i\hat{\omega}t}\tilde{\chi}(\hat{\omega})\right)=-\mathbf{Im}\tilde{\chi}(\hat{\omega})\cos(\hat{\omega}t)+\mathbf{Re}\tilde{\chi}(\hat{\omega})\sin(\hat{\omega}t)= (81)
=|χ~(ω^)|sin(ω^t+arctan(𝐑𝐞χ~(ω^)𝐈𝐦χ~(ω^))π2)\displaystyle=|\tilde{\chi}(\hat{\omega})|\sin\left(\hat{\omega}t+\arctan\left(\frac{\mathbf{Re}\tilde{\chi}(\hat{\omega})}{\mathbf{Im}\tilde{\chi}(\hat{\omega})}\right)-\frac{\pi}{2}\right) (82)

The response to perturbations at the critical frequency ω\omega^{*} is then

A11(t;ω)=1γ(N)(κRcos(ωt)+κIsin(ωt))=|κ|γ(N)sin(ωt+φ)\langle\langle A_{1}\rangle\rangle_{1}(t;\omega^{*})=\frac{1}{\gamma(N)}\left(\kappa_{R}\cos(\omega^{*}t)+\kappa_{I}\sin(\omega^{*}t)\right)=\frac{|\kappa|}{\gamma(N)}\sin\left(\omega^{*}t+\varphi^{\prime}\right) (83)

where φ=φπ2\varphi^{\prime}=\varphi-\frac{\pi}{2} and φ\varphi is related to the ratio of the imaginary and real part of the residue as

φ=arctan(κIκR)=arctan(ωD)\varphi=\arctan\left(-\frac{\kappa_{I}}{\kappa_{R}}\right)=\arctan\left(\frac{\omega^{*}}{D}\right) (84)

Firstly, as expected, the amplitude of the oscillation diverges in the thermodynamic limit as γ(N)0\gamma(N)\to 0. Secondly, given that κR0\kappa_{R}\neq 0, the response spontaneously develop a phase shift φ0\varphi^{\prime}\neq 0 with respect to the external sinusoidal forcing. We have investigated the development of the phase shift by comparing the numerical response A11(t)\langle\langle A_{1}\rangle\rangle_{1}(t) obtained as the convolution product (75) and the exact theoretical result (83). A quantitative estimation of the phase shift is obtained by looking at the correlation function between the forcing and the response. The values at which the correlation function has its maxima correspond to the phase shift. A visual inspection of figure 2 clearly indicates a very good agreement between numerical results and theory. It is also interesting to investigate the response of the system to a very fast driving signal. This regime can be investigated by looking at the high frequency behaviour of the critical susceptibility of the system. We expect that 𝐑𝐞χ~i(ω)0\mathbf{Re}\tilde{\chi}_{i}(\omega)\approx 0 for ω\omega\to\infty so that, from (81), the response will be in quadrature with the forcing A1(t)cos(ωt)\langle\langle A_{1}\rangle\rangle(t)\propto\cos(\omega t), as clearly shown in the right column of figure 2 (here we have chosen a frequency ω^=2ω\hat{\omega}=2\omega^{*}).

A.2.1 Parameters for the numerical experiments

We here report the critical and off-critical parameters for the numerical experiments.

  • Static Transition μ(h)=δ(h)\mu(h)=\delta(h): the coupling strength is fixedθ=1\theta=1. The value of the noise strength at the phase transition is σc=1\sigma_{c}=1, whereas away from the transition is σ=1.1\sigma=1.1.

  • Dynamic Transition μh0(h)=12(δ(hh0)+δ(h+h0))\mu_{h_{0}}(h)=\frac{1}{2}\left(\delta(h-h_{0})+\delta(h+h_{0})\right): h0=2h_{0}=2, θ=2\theta=2 are fixed. The value of the noise strength at the transition is σc=1\sigma_{c}=1, away from the transition is σ=1.4\sigma=1.4. As a result, the critical frequency is ω=h02D21.93\omega^{*}=\sqrt{h_{0}^{2}-D^{2}}\approx 1.93 and the phase shift φ=φπ20.25\varphi^{\prime}=\varphi-\frac{\pi}{2}\approx-0.25, see equation (84).

References

References

  • [1] Hidetsugu Sakaguchi. Phase transition in globally coupled Rössler oscillators. Phys. Rev. E, 61:7212–7214, Jun 2000.
  • [2] A. Pikovsky, J. Kurths, M. Rosenblum, and J. Kurths. Synchronization: A Universal Concept in Nonlinear Sciences. Cambridge Nonlinear Science Series. Cambridge University Press, 2003.
  • [3] L. M. Pecora. Synchronization conditions and desynchronizing patterns in coupled limit-cycle and chaotic systems. Phys. Rev. E, 58:347, 1998.
  • [4] Louis M. Pecora and Thomas L. Carroll. Synchronization of chaotic systems. Chaos: An Interdisciplinary Journal of Nonlinear Science, 25(9):097611, 2015.
  • [5] Deniz Eroglu, Jeroen S. W. Lamb, and Tiago Pereira. Synchronisation of chaos and its applications. Contemporary Physics, 58(3):207–243, 2017.
  • [6] Edward Ott, Paul So, Ernest Barreto, and Thomas Antonsen. The onset of synchronization in systems of globally coupled chaotic and periodic oscillators. Physica D: Nonlinear Phenomena, 173(1):29–51, 2002.
  • [7] Juan A. Acebrón, L. L. Bonilla, Conrad J. Pérez Vicente, Félix Ritort, and Renato Spigler. The kuramoto model: A simple paradigm for synchronization phenomena. Rev. Mod. Phys., 77:137–185, Apr 2005.
  • [8] Pau Clusella and Antonio Politi. Noise-induced stabilization of collective dynamics. Phys. Rev. E, 95:062221, Jun 2017.
  • [9] Arkady Pikovsky and Michael Rosenblum. Introduction to focus issue: Dynamics of oscillator populations. Chaos: An Interdisciplinary Journal of Nonlinear Science, 33(1):010401, 2023.
  • [10] Susana N. Gomes, Serafim Kalliadasis, Grigorios A. Pavliotis, and Petr Yatsyshin. Dynamics of the desai-zwanzig model in multiwell and random energy landscapes. Phys. Rev. E, 99:032109, Mar 2019.
  • [11] S.N. Gomes and G.A. Pavliotis. Mean field limits for interacting diffusions in a two-scale potential. J. Nonlin. Sci., 28(3):905–941, 2018.
  • [12] C. Wang, Q. Li, W. E, and B. Chazelle. Noisy hegselmann-krause systems: Phase transition and the 2r-conjecture. Journal of Statistical Physics, 166(5):1209–1225, 2017.
  • [13] B. D. Goddard, B. Gooding, H. Short, and G. A. Pavliotis. Noisy bounded confidence models for opinion dynamics: the effect of boundary conditions on phase transitions. IMA J. Appl. Math., 87(1):80–110, 2022.
  • [14] G. Naldi, L. Pareschi, and G. Toscani. Mathematical Modeling of Collective Behavior in Socio-Economic and Life Sciences. Birkhäuser Basel, 2010.
  • [15] L. Pareschi and G. Toscani. Interacting multiagent systems: kinetic equations and Monte Carlo methods. OUP Oxford, 2013.
  • [16] Paolo Dai Pra. Stochastic mean-field dynamics and applications to life sciences. In Giambattista Giacomin, Stefano Olla, Ellen Saada, Herbert Spohn, and Gabriel Stoltz, editors, Stochastic Dynamics Out of Equilibrium, pages 3–27, Cham, 2019. Springer International Publishing.
  • [17] Ernest Montbrió, Diego Pazó, and Alex Roxin. Macroscopic description for networks of spiking neurons. Phys. Rev. X, 5:021028, Jun 2015.
  • [18] José Antonio Carrillo, Young-Pil Choi, and Maxime Hauray. The derivation of swarming models: Mean-field limit and Wasserstein distances, pages 1–46. Springer Vienna, Vienna, 2014.
  • [19] José A. Carrillo, Massimo Fornasier, Giuseppe Toscani, and Francesco Vecil. Particle, kinetic, and hydrodynamic models of swarming, pages 297–336. Birkhäuser Boston, Boston, 2010.
  • [20] J. A. Carrillo, R. S. Gvalani, G. A. Pavliotis, and A. Schlichting. Long-time behaviour and phase transitions for the mckean–vlasov equation on the torus. Archive for Rational Mechanics and Analysis, 235(1):635–690, 2020.
  • [21] Pierre-Henri Chavanis. The Brownian mean field model. The European Physical Journal B, 87(5):120, 2014.
  • [22] Takayuki Tatekawa, Freddy Bouchet, Thierry Dauxois, and Stefano Ruffo. Thermodynamics of the self-gravitating ring model. Phys. Rev. E, 71:056111, May 2005.
  • [23] A. Borovykh, N. Kantas, P. Parpas, and G. A. Pavliotis. On stochastic mirror descent with interacting particles: convergence properties and variance reduction. Phys. D, 418:Paper No. 132844, 21, 2021.
  • [24] A. Garbuno-Inigo, N. Nüsken, and S. Reich. Affine invariant interacting Langevin dynamics for Bayesian inference. SIAM J. Appl. Dyn. Syst., 19(3):1633–1658, 2020.
  • [25] Grant Rotskoff and Eric Vanden-Eijnden. Trainability and accuracy of artificial neural networks: An interacting particle system approach. Communications on Pure and Applied Mathematics, 75(9):1889–1935, 2022.
  • [26] Donald A. Dawson. Critical dynamics and fluctuations for a mean-field model of cooperative behavior. Journal of Statistical Physics, 31(1):29–85, 1983.
  • [27] Jutta Rogal. Reaction coordinates in complex systems-a perspective. The European Physical Journal B, 94(11):223, 2021.
  • [28] Ao Ma and Aaron R. Dinner. Automatic method for identifying reaction coordinates in complex systems. The Journal of Physical Chemistry B, 109(14):6769–6779, 04 2005.
  • [29] Giovanni Bussi, Alessandro Laio, and Michele Parrinello. Equilibrium free energies from nonequilibrium metadynamics. Phys. Rev. Lett., 96:090601, Mar 2006.
  • [30] V. Lucarini, G. A. Pavliotis, and N. Zagli. Response theory and phase transitions for the thermodynamic limit of interacting identical systems. Proc. R. Soc. A., 476, 2020.
  • [31] Niccolò Zagli, Valerio Lucarini, and Grigorios A. Pavliotis. Spectroscopy of phase transitions for multiagent systems. Chaos: An Interdisciplinary Journal of Nonlinear Science, 31(6):061103, 2021.
  • [32] Niccolò Zagli, Grigorios A. Pavliotis, Valerio Lucarini, and Alexander Alecio. Dimension reduction of noisy interacting systems. Phys. Rev. Res., 5:013078, Feb 2023.
  • [33] Dmitri Topaj, Won-Ho Kye, and Arkady Pikovsky. Transition to coherence in populations of coupled chaotic oscillators: A linear response approach. Phys. Rev. Lett., 87:074101, Jul 2001.
  • [34] D. Ruelle. A review of linear response theory for general differentiable dynamical systems. Nonlinearity, 22(4):855–870, April 2009.
  • [35] Michael Ghil and Valerio Lucarini. The physics of climate variability and climate change. Rev. Mod. Phys., 92:035002, Jul 2020.
  • [36] Valerio Lucarini and Mickaël D. Chekroun. Theoretical tools for understanding the climate crisis from hasselmann’s programme and beyond. Nature Reviews Physics, 2023.
  • [37] Bruno Cessac. Linear response in neuronal networks: from neurons dynamics to collective response. Chaos, 29(10):103105, 24, 2019.
  • [38] Bruno Cessac, Ignacio Ampuero, and Rodrigo Cofré. Linear response of general observables in spiking neuronal network models. Entropy, 23(2):Paper No. 155, 30, 2021.
  • [39] Soon Hoe Lim. Understanding recurrent neural networks using nonequilibrium response theory. J. Mach. Learn. Res., 22:Paper No. 47, 48, 2021.
  • [40] Antonio M. Puertas, Juan E. Trinidad-Segovia, Miguel A. Sánchez-Granero, Joaquim Clara-Rahora, and F. Javier de las Nieves. Linear response theory in stock markets. Scientific Reports, 11(1):23076, 2021.
  • [41] C. Liverani and S. Gouëzel. Banach spaces adapted to Anosov systems. Ergodic Theory and Dynamical Systems, 26:189–217, 2006.
  • [42] Manuel Santos Gutiérrez and Valerio Lucarini. Response and sensitivity using markov chains. Journal of Statistical Physics, 2020.
  • [43] U. Marini Bettolo Marconi, A. Puglisi, L. Rondoni, and A. Vulpiani. Fluctuation-dissipation: Response theory in statistical physics. Phys. Rep., 461:111, 2008.
  • [44] Grigorios A. Pavliotis. Stochastic Processes and Applications, volume 60. Springer, New York, 2014.
  • [45] M. Hairer and A. J. Majda. A simple framework to justify linear response theory. Nonlinearity, 23(4):909–922, 2010.
  • [46] Klaus-Jochen Engel and Rainer Nagel. One-parameter semigroups for linear evolution equations, volume 194 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000. With contributions by S. Brendle, M. Campiti, T. Hahn, G. Metafune, G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R. Schnaubelt.
  • [47] Manuel Santos Gutiérrez and Valerio Lucarini. On some aspects of the response to stochastic and deterministic forcings. Journal of Physics A: Mathematical and Theoretical, 55(42):425002, oct 2022.
  • [48] Mickaël D. Chekroun, Alexis Tantet, Henk A. Dijkstra, and J. David Neelin. Ruelle–pollicott resonances of stochastic systems in reduced state space. part i: Theory. Journal of Statistical Physics, 2020.
  • [49] Andrzej Lasota and Michael C. Mackey. Chaos, fractals and noise. Springer, New York, 1994.
  • [50] Edward Ott and Thomas M. Antonsen. Low dimensional behavior of large systems of globally coupled oscillators. Chaos: An Interdisciplinary Journal of Nonlinear Science, 18(3):037113, 2008.
  • [51] Irina V. Tyulkina, Denis S. Goldobin, Lyudmila S. Klimenko, and Arkady Pikovsky. Dynamics of noisy oscillator populations beyond the ott-antonsen ansatz. Phys. Rev. Lett., 120:264101, Jun 2018.
  • [52] Denis S. Goldobin, Matteo di Volo, and Alessandro Torcini. Reduction methodology for fluctuation driven population dynamics. Phys. Rev. Lett., 127:038301, Jul 2021.
  • [53] Rok Cestnik and Arkady Pikovsky. Exact finite-dimensional reduction for a population of noisy oscillators and its link to ott–antonsen and watanabe–strogatz theories. Chaos: An Interdisciplinary Journal of Nonlinear Science, 32(11):113126, 2022.
  • [54] Dan Wilson and Bard Ermentrout. Phase models beyond weak coupling. Phys. Rev. Lett., 123:164101, Oct 2019.
  • [55] Hiroya Nakao. Phase reduction approach to synchronisation of nonlinear oscillators. Contemporary Physics, 57(2):188–214, 2016.
  • [56] H.P McKean Jr. Propagation of chaos for a class of non-linear parabolic equations. Stochastic Differential Equations Lecture Series in Differential Equations, Session 7, Catholic Univ., 1967.
  • [57] A.S. Sznitman. Topics in propagation of chaos., volume 1464 of Hennequin PL. (eds) Ecole d’Eté de Probabilités de Saint-Flour XIX — 1989. Lecture Notes in Mathematics. Springer, Berlin, Heidelberg, 1989.
  • [58] Louis-Pierre Chaintron and Antoine Diez. Propagation of chaos: a review of models, methods and applications. i. models and methods, 2022.
  • [59] Paolo Dai Pra and Frank den Hollander. McKean-Vlasov limit for interacting random processes in random media. Journal of Statistical Physics, 84(3):735–772, 1996.
  • [60] T. D. Frank. Nonlinear Fokker-Planck Equations Fundamentals and Applications. Springer, Berlin, Heidelberg, 2005.
  • [61] N. Martzel and C. Aslangul. Mean-field treatment of the many-body Fokker-Planck equation. J. Phys. A, 34(50):11225–11240, 2001.
  • [62] Rainer. Kress. Linear Integral Equations. Springer New York, New York, NY, 3rd ed. 2014. edition, 2014.
  • [63] Guy A. Battle. Phase transitions for a continuous system of classical particles in a box. Comm. Math. Phys., 55(3):299–315, 1977.
  • [64] Dmitrii Zubarev, Vladimir Morozov, and Gerd Röpke. Statistical mechanics of nonequilibrium processes. Vol. 1. Akademie Verlag, Berlin, 1996. Basic concepts, kinetic theory.
  • [65] Juan A. Acebrón, L. L. Bonilla, Conrad J. Pérez Vicente, Félix Ritort, and Renato Spigler. The kuramoto model: A simple paradigm for synchronization phenomena. Rev. Mod. Phys., 77:137–185, Apr 2005.
  • [66] Shamik Gupta, Alessandro Campa, and Stefano Ruffo. Kuramoto model of synchronization: equilibrium and nonequilibrium aspects. J. Stat. Mech. Theory Exp., (8):R08001, 61, 2014.
  • [67] D. Iatsenko, S. Petkoski, P. V. E. McClintock, and A. Stefanovska. Stationary and traveling wave states of the kuramoto model with an arbitrary distribution of frequencies and coupling strengths. Phys. Rev. Lett., 110:064101, Feb 2013.
  • [68] Steven H. Strogatz and Renato E. Mirollo. Stability of incoherence in a population of coupled oscillators. J. Statist. Phys., 63(3-4):613–635, 1991.
  • [69] Luis L. Bonilla, John C. Neu, and Renato Spigler. Nonlinear stability of incoherence and collective synchronization in a population of coupled oscillators. J. Statist. Phys., 67(1-2):313–330, 1992.
  • [70] Victor Buendía, Pablo Villegas, Raffaella Burioni, and Miguel A. Muñoz. The broad edge of synchronization: Griffiths effects and collective phenomena in brain networks. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 380(2227):20200424, 2022.
  • [71] Spase Petkoski and Aneta Stefanovska. Kuramoto model with time-varying parameters. Phys. Rev. E, 86:046212, Oct 2012.
  • [72] V. Lucarini, J. J. Saarinen, K.-E. Peiponen, and E. M. Vartiainen. Kramers-Kronig relations in Optical Materials Research. Springer, New York, 2005.
  • [73] Joram Lindenstrauss and Lior Tzafriri. Classical Banach spaces. I. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 92. Springer-Verlag, Berlin-New York, 1977.
  • [74] David R. Dellwo. Accelerated degenerate-kernel methods for linear integral equations. Journal of Computational and Applied Mathematics, 58(2):135–149, 1995.
  • [75] Cristobal Quiñinao and Jonathan Touboul. Limits and dynamics of randomly connected neuronal networks. Acta Applicandae Mathematicae, 136(1):167–192, 2015.
  • [76] László Lovász. Large Networks and Graph Limits., volume 60 of Colloquium Publications. American Mathematical Society, 2012.
  • [77] Fabio Coppini. Long time dynamics for interacting oscillators on graphs. The Annals of Applied Probability, 32(1):360 – 391, 2022.
  • [78] Christian Kuehn. Network dynamics on graphops. New Journal of Physics, 22(5):053030, may 2020.
  • [79] Marios Antonios Gkogkas, Benjamin Jüttner, Christian Kuehn, and Erik Andreas Martens. Graphop mean-field limits and synchronization for the stochastic Kuramoto model. Chaos: An Interdisciplinary Journal of Nonlinear Science, 32(11):113120, 11 2022.
  • [80] G. Gallavotti. Nonequilibrium and irreversibility. Springer, New York, 2014.
  • [81] MH Duong and GA Pavliotis. Mean field limits for non-markovian interacting particles: Convergence to equilibrium, generic formalism, asymptotic limits and phase transitions. Communications in Mathematical Sciences, 16:2199–2230, 2018.
  • [82] Manuel Santos Gutiérrez, Valerio Lucarini, Mickaël D. Chekroun, and Michael Ghil. Reduced-order models for coupled dynamical systems: Data-driven methods and the koopman operator. Chaos: An Interdisciplinary Journal of Nonlinear Science, 31(5):053116, 2021.
  • [83] Mickaël D. Chekroun, Alexis Tantet, Henk A. Dijkstra, and J. David Neelin. Ruelle–pollicott resonances of stochastic systems in reduced state space. part i: Theory. Journal of Statistical Physics, 2020.