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

Metastability associated with many-body explosion of eigenmode expansion coefficients

Takashi Mori RIKEN Center for Emergent Matter Science (CEMS), Wako 351-0198, Japan
Abstract

Metastable states in stochastic systems are often characterized by the presence of small eigenvalues in the generator of the stochastic dynamics. We here show that metastability in many-body systems is not necessarily associated with small eigenvalues. Instead, many-body explosion of eigenmode expansion coefficients characterizes slow relaxation, which is demonstrated for two models, interacting particles in a double-well potential and the Fredrickson-Andersen model, the latter of which is a prototypical example of kinetically constrained models studied in glass and jamming transitions. Our results provide insights into slow relaxation and metastability in many-body stochastic systems.

I Introduction

Metastability in stochastic systems is ubiquitous in nature. In particular, it is often associated with first-order phase transitions Langer (1969, 1974); Penrose (1995), (spin) glasses Bray and Moore (1980); Biroli and Kurchan (2001); Berthier and Biroli (2011), protein folding Baldwin et al. (2011), and so forth. Despite its importance, the definition of metastability is quite subtle. In mean-field models, there is a static characterization of metastable states, i.e., local minima of the free energy as a function of order parameters Griffiths et al. (1966). For non mean-field models, however, a metastable state has a finite lifetime even in the thermodynamic limit Lanford and Ruelle (1969), and hence we should treat metastability as a timescale-dependent dynamical concept 111See Refs. Langer (1969); Penrose (1995) for attempts to give a static characterization of finite-lifetime metastable states associated with first-order phase transitions via an analytic continuation of the thermodynamic function..

Since the generator GG of the master equation contains full information on the stochastic dynamics, it is natural to expect that metastability is dynamically characterized by some special properties of eigenvalues and/or eigenvectors of GG. Indeed, the real part of an eigenvalue of GG gives the decay rate of the corresponding eigenmode, which suggests that metastability is characterized by the presence of small eigenvalues separated from the other large eigenvalues Gaveau and Schulman (1998); Bovier et al. (2000); Biroli and Kurchan (2001) (see Refs. Macieszczak et al. (2016); Rose et al. (2016, ) for quantum systems). Such a spectral characterization of metastable states is appealing since it is purely dynamical; we do not have to rely on thermodynamic notions such as the (free) energy landscape, which is difficult to define precisely except for mean-field models Frenkel (2013).

In this paper, we revisit this problem. We show that metastability in many-body stochastic systems does not necessarily accompany small eigenvalues. By analyzing two simple models, i.e. interacting particles in a double-well potential and the Fredrickson-Andersen (FA) model Fredrickson and Andersen (1984), we demonstrate that slow relaxation in those models is characterized by many-body explosion of eigenmode expansion coefficients, which was recently studied in the context of open quantum systems Mori and Shirai (2020); Haga et al. (2021) (see also Ref. Bensa and Znidaric ). Our results provide insights into slow relaxation and metastability in many-body stochastic systems.

The remaining part of the paper is organized as follows. In Sec. II, we give a general formulation within a framework of classical Markov processes on discrete states. We introduce eigenmode relaxation times in Sec. II.2. In Sec. III, we numerically show that eigenmode expansion coefficients become huge for concrete models. In Sec. III.1, we consider non-interacting particles in a double-well potential. Metastability in this model is due to large barrier of a single-particle potential, and hence we call it “potential-induced metastability”. In Sec. III.2, we consider the same model with inter-particle interactions. Metastability is induced by strong inter-particle interactions, and hence we call it “interaction-induced metastability”. In Sec. III.3, the FA model is investigated. Metastability in this model is due to kinetic constraints and not of energetic origin. It turns out that potential-induced metastability can be explained by the emergence of vanishingly small eigenvalues, whereas the other two are not associated with small eigenvalues: they are explained by many-body explosion of expansion coefficients. Finally, we discuss our results in Sec. IV.

II General formulation

II.1 Setup

Let us consider a system with discrete states n=1,2,,Dn=1,2,\dots,D. The energy of the state nn is denoted by EnE_{n}. The probability distribution Pn(t)P_{n}(t) evolves following the master equation

dPn(t)dt=m=1DGn,mPm(t),\frac{dP_{n}(t)}{dt}=-\sum_{m=1}^{D}G_{n,m}P_{m}(t), (1)

where the matrix GG is the generator of the stochastic dynamics. For nmn\neq m, Gn,mG_{n,m} represents the transition rate from the state mm to nn, and Gn,n=m(n)Gm,nG_{n,n}=-\sum_{m(\neq n)}G_{m,n}. When the system is coupled to a thermal bath at the inverse temperature β\beta, GG satisfies the detailed-balance condition, Gn,meβEm=Gm,neβEnG_{n,m}e^{-\beta E_{m}}=G_{m,n}e^{-\beta E_{n}}.

The detailed-balance condition ensures that GG is diagonalizable with real (and non-negative) eigenvalues since it is made symmetric via the similarity transformation GG=eβH/2GeβH/2G\to G^{\prime}=e^{\beta H/2}Ge^{-\beta H/2} with H=diag(E1,E2,,ED)H=\mathrm{diag}(E_{1},E_{2},\dots,E_{D}). We here assume that the stationary state is unique and eigenvalues are ordered as 0=Λ0<Λ1Λ2ΛD10=\Lambda_{0}<\Lambda_{1}\leq\Lambda_{2}\dots\leq\Lambda_{D-1}. The right and left eigenvectors for an eigenvalue Λα\Lambda_{\alpha} are respectively denoted by Φ(α)\vec{\Phi}^{(\alpha)} and Π(α)\vec{\Pi}^{(\alpha)}: GΦ(α)=ΛαΦ(α)G\vec{\Phi}^{(\alpha)}=\Lambda_{\alpha}\vec{\Phi}^{(\alpha)} and GTΠ(α)=ΛαΠ(α)G^{\mathrm{T}}\vec{\Pi}^{(\alpha)}=\Lambda_{\alpha}\vec{\Pi}^{(\alpha)}.

Let us normalize right eigenvectors by using the L1L^{1} norm 1\|\cdot\|_{1} as follows:

Φ(α)1:=n=1D|Φn(α)|=1,\|\vec{\Phi}^{(\alpha)}\|_{1}:=\sum_{n=1}^{D}|\Phi^{(\alpha)}_{n}|=1, (2)

where Φn(α)\Phi^{(\alpha)}_{n} are nnth component of Φ(α)\vec{\Phi}^{(\alpha)}. We will later see that this normalization is convenient for our purpose. Since the dual norm of the L1L^{1} norm is the LL^{\infty} norm \|\cdot\|_{\infty} in the sense that |AB|AB1|\vec{A}\cdot\vec{B}|\leq\|\vec{A}\|_{\infty}\|\vec{B}\|_{1} for any set of vectors A\vec{A} and B\vec{B}, we normalize left eigenvectors by using LL^{\infty} norm as follows:

Π(α):=maxn|Πn(α)|=1,\|\vec{\Pi}^{(\alpha)}\|_{\infty}:=\max_{n}|\Pi^{(\alpha)}_{n}|=1, (3)

where Πn(α)\Pi^{(\alpha)}_{n} are nnth component of Π(α)\vec{\Pi}^{(\alpha)}.

The mode α=0\alpha=0 corresponds to the stationary state, and Φn(0)=eβEn/Z\Phi^{(0)}_{n}=e^{-\beta E_{n}}/Z with Z=neβEnZ=\sum_{n}e^{-\beta E_{n}} and Πn(0)=1\Pi^{(0)}_{n}=1 for all nn. Without loss of generality, we choose the signs of Φ(α)\vec{\Phi}^{(\alpha)} and Π(α)\vec{\Pi}^{(\alpha)} so that maxnΠn(α)=1\max_{n}\Pi^{(\alpha)}_{n}=1 and Π(α)Φ(α)>0\vec{\Pi}^{(\alpha)}\cdot\vec{\Phi}^{(\alpha)}>0.

The probability distribution P(t)=(P1(t),,PD(t))T\vec{P}(t)=(P_{1}(t),\dots,P_{D}(t))^{\mathrm{T}} can be expanded in terms of right eigenvectors of GG as P(t)=αCαeΛαtΦ(α)\vec{P}(t)=\sum_{\alpha}C_{\alpha}e^{-\Lambda_{\alpha}t}\vec{\Phi}^{(\alpha)}, where

Cα=Π(α)P(0)Π(α)Φ(α)C_{\alpha}=\frac{\vec{\Pi}^{(\alpha)}\cdot\vec{P}(0)}{\vec{\Pi}^{(\alpha)}\cdot\vec{\Phi}^{(\alpha)}} (4)

is an eigenmode expansion coefficient in the initial state.

II.2 Eigenmode relaxation time

Obviously, Λα\Lambda_{\alpha} gives the decay rate of α\alphath eigenmode, and hence the slowest decay rate is given by the spectral gap Λ1\Lambda_{1} of GG. This observation indicates that slow relaxation should accompany small Λ1\Lambda_{1}. Following this idea, in previous works Gaveau and Schulman (1987, 1998); Biroli and Kurchan (2001); Macieszczak et al. (2016), metastability is characterized by the presence of small eigenvalues. This argument implicitly assumes that CαC_{\alpha} is not too large. As we point out below, however, this assumption is typically not satisfied in many-body systems.

For a given initial state P(0)=αCαΦ(α)\vec{P}(0)=\sum_{\alpha}C_{\alpha}\vec{\Phi}^{(\alpha)}, let us define the relaxation time τα\tau_{\alpha} for each eigenmode. The expectation value of a physical quantity O=(O1,O2,,OD)T\vec{O}=(O_{1},O_{2},\dots,O_{D})^{\mathrm{T}} with O=1\|\vec{O}\|_{\infty}=1 is given by

O=OP(t)=αCαeΛαtOΦ(α).\braket{O}=\vec{O}\cdot\vec{P}(t)=\sum_{\alpha}C_{\alpha}e^{-\Lambda_{\alpha}t}\vec{O}\cdot\vec{\Phi}^{(\alpha)}. (5)

The contribution from α\alphath eigenmode is thus bounded as

|CαeΛαtOΦ(α)||Cα|eΛαtOΦ(α)1=|Cα|eΛαt,|C_{\alpha}e^{-\Lambda_{\alpha}t}\vec{O}\cdot\vec{\Phi}^{(\alpha)}|\leq|C_{\alpha}|e^{-\Lambda_{\alpha}t}\|\vec{O}\|_{\infty}\|\vec{\Phi}^{(\alpha)}\|_{1}=|C_{\alpha}|e^{-\Lambda_{\alpha}t}, (6)

where the normalization (2) is used. Thus the contribution from α\alphath eigenmode is negligible when |Cα|eΛαt1|C_{\alpha}|e^{-\Lambda_{\alpha}t}\ll 1, and hence it is appropriate to define the eigenmode relaxation time τα\tau_{\alpha} as the time at which |Cα|eΛαt=δ|C_{\alpha}|e^{-\Lambda_{\alpha}t}=\delta for a fixed constant δ\delta. Throughout the paper, we fix δ=0.5\delta=0.5.

We then have

τα=ln(|Cα|/δ)Λα,\tau_{\alpha}=\frac{\ln\left(|C_{\alpha}|/\delta\right)}{\Lambda_{\alpha}}, (7)

which implies that large expansion coefficients can cause the delay of the relaxation 222If we choose another normalization condition for right eigenvectors, the criterion of the relaxation |Cα|eΛαt=δ|C_{\alpha}|e^{-\Lambda_{\alpha}t}=\delta should be modified. For example, if we employ the normalization Φ(α)2:=(n=1D|Φn(α)|2)1/2=1\|\vec{\Phi}^{(\alpha)}\|_{2}:=\left(\sum_{n=1}^{D}|\Phi^{(\alpha)}_{n}|^{2}\right)^{1/2}=1, we have |OΦ(α)|D1/2|\vec{O}\cdot\vec{\Phi}^{(\alpha)}|\sim D^{1/2}, and hence the criterion of the relaxation should be |Cα|eΛαt=δD1/2|C_{\alpha}|e^{-\Lambda_{\alpha}t}=\delta D^{-1/2} with δ\delta being a fixed constant. Accordingly, the formula (7) should also be modified. In other words, Eq. (7) is appropriate only for the normalization scheme (2).

Which eigenmodes are responsible for metastable states? To answer this question, let us examine the possible largest value of τα\tau_{\alpha} over all initial distributions. From Eq. (4), we have an upper bound

|Cα|Π(α)P(0)1Π(α)Φ(α)=1Π(α)Φ(α)=:Ψα.|C_{\alpha}|\leq\frac{\|\vec{\Pi}^{(\alpha)}\|_{\infty}\|\vec{P}(0)\|_{1}}{\vec{\Pi}^{(\alpha)}\cdot\vec{\Phi}^{(\alpha)}}=\frac{1}{\vec{\Pi}^{(\alpha)}\cdot\vec{\Phi}^{(\alpha)}}=:\Psi_{\alpha}. (8)

This upper bound is always realizable: |Cα|=Ψα|C_{\alpha}|=\Psi_{\alpha} when Pn(0)P_{n}(0) is nonzero only for nn such that Πn(α)=1\Pi^{(\alpha)}_{n}=1. Thus, we obtain

maxP(0)τα=ln(Ψα/δ)Λα=:τ~α.\max_{\vec{P}(0)}\tau_{\alpha}=\frac{\ln(\Psi_{\alpha}/\delta)}{\Lambda_{\alpha}}=:\tilde{\tau}_{\alpha}. (9)

The eigenmode relaxation times {τ~α}\{\tilde{\tau}_{\alpha}\} tell us about which eigenmodes can compose metastable states. In bra-ket notation, the time evolution operator is written as

eGt=αΨαeΛαt|Φ(α)Π(α)|.e^{-Gt}=\sum_{\alpha}\Psi_{\alpha}e^{-\Lambda_{\alpha}t}\ket{\Phi^{(\alpha)}}\bra{\Pi^{(\alpha)}}. (10)

For a timescale specified by tt, we consider a subset \mathcal{M} of eigenmodes such that τ~αt\tilde{\tau}_{\alpha}\ll t for any α\alpha\notin\mathcal{M}. Since ΨαeΛαt=δeΛα(τ~αt)1\Psi_{\alpha}e^{-\Lambda_{\alpha}t}=\delta e^{\Lambda_{\alpha}(\tilde{\tau}_{\alpha}-t)}\ll 1 for α\alpha\notin\mathcal{M}, we have eGtαΨαeΛαt|Φ(α)Π(α)|e^{-Gt}\approx\sum_{\alpha\in\mathcal{M}}\Psi_{\alpha}e^{-\Lambda_{\alpha}t}\ket{\Phi^{(\alpha)}}\bra{\Pi^{(\alpha)}}, which implies that a metastable state consists of eigenmodes in \mathcal{M}.

Refer to caption
Figure 1: Schematic picture of the model. Here, EBE_{B} denotes the energy barrier and ε\varepsilon the potential energy of the left well. In the interacting model, particles in the same potential well interact with each other (gg denotes the interaction strength).

III Metastability and explosive growth of eigenmode expansion coefficients

In this section, we show that Ψα\Psi_{\alpha} becomes huge and significantly changes the relaxation time in many-body stochastic models, i.e., an NN-particle system in a double-well potential and the FA model. Through the analysis of those models, we consider three kinds of metastability, i.e., the potential-induced metastability, the interaction-induced metastability, and the metastability due to kinetic constraints. It is shown that the potential-induced metastability is associated with small eigenvalues, whereas the others are caused by an explosive growth of expansion coefficients Ψα\Psi_{\alpha}.

III.1 Non-interacting particles in a double-well potential

Let us begin with the simplest model, i.e., NN independent particles in a double-well potential under the two-state approximation (see Fig. 1). Each particle ii (i=1,2,,Ni=1,2,\dots,N) is either in the left or right well, which is expressed by a spin variable σi=±1\sigma_{i}=\pm 1 (σi=1\sigma_{i}=1 corresponds to the left well). Then NN-particle state is specified by a set of spin variables (σ1,σ2,,σN)(\sigma_{1},\sigma_{2},\dots,\sigma_{N}), and there are D=2ND=2^{N} different states. The state σi=+1\sigma_{i}=+1 (left well) has the energy ε>0\varepsilon>0, and the energy barrier undergoing in the transition from σi=+1\sigma_{i}=+1 to 1-1 is given by EB>0E_{B}>0. By choosing the temperature as the unit of energy (i.e., β=1\beta=1), transitions of iith particle from σi=1\sigma_{i}=1 to 1-1 and vice versa happen at rates given by the Arrhenius law τ01eEB\tau_{0}^{-1}e^{-E_{B}} and τ01e(EB+ε)\tau_{0}^{-1}e^{-(E_{B}+\varepsilon)}, respectively. Here, τ0\tau_{0} is a certain microscopic timescale, which is chosen as the unit of time, i.e., τ0=1\tau_{0}=1. This transition rate satisfies the detailed-balance condition for the Hamiltonian H=εN+:=εi=1N(σi+1)/2H=\varepsilon N_{+}:=\varepsilon\sum_{i=1}^{N}(\sigma_{i}+1)/2, where the number of particles in the left well is denoted by N+N_{+}. Similarly, we define N=NN+N_{-}=N-N_{+} as that in the right well.

In the single-particle problem, the generator is a two-by-two matrix, and it has two eigenvalues λ0=0\lambda_{0}=0 and λ1=eEB(1+eε)\lambda_{1}=e^{-E_{B}}(1+e^{-\varepsilon}). The right and left eigenvectors are denoted by ϕ(α)\vec{\phi}^{(\alpha)} and π(α)\vec{\pi}^{(\alpha)} (α=0\alpha=0 or 1), respectively. In the NN-particle case, the eigenvalues {Λα}\{\Lambda_{\alpha}\} are given by the sum of single-particle eigenvalues as Λα=i=1Nαiλ1\Lambda_{\alpha}=\sum_{i=1}^{N}\alpha_{i}\lambda_{1}, where αi=0,1\alpha_{i}=0,1 specifies which eigenmode is occupied by iith particle. The corresponding eigenvectors are given by the product of single-particle eigenvectors: Φ(α)=ϕ(α1)ϕ(α2)ϕ(αN)\vec{\Phi}^{(\alpha)}=\vec{\phi}^{(\alpha_{1})}\otimes\vec{\phi}^{(\alpha_{2})}\dots\otimes\vec{\phi}^{(\alpha_{N})} and Π(α)=π(α1)π(α2)π(αN)\vec{\Pi}^{(\alpha)}=\vec{\pi}^{(\alpha_{1})}\otimes\vec{\pi}^{(\alpha_{2})}\dots\otimes\vec{\pi}^{(\alpha_{N})}.

The maximum expansion coefficient Ψα\Psi_{\alpha} in Eq. (8) for each α\alpha is explicitly calculated as follows:

Ψα=(21+eε)i=1Nαi.\Psi_{\alpha}=\left(\frac{2}{1+e^{-\varepsilon}}\right)^{\sum_{i=1}^{N}\alpha_{i}}. (11)

When i=1Nαi=O(N)\sum_{i=1}^{N}\alpha_{i}=O(N), Ψα=eO(N)\Psi_{\alpha}=e^{O(N)} for ε>0\varepsilon>0. Expansion coefficients diverge exponentially in NN. Although the corresponding eigenvalue is extensively large Λα=O(N)\Lambda_{\alpha}=O(N), it is not correct that this eigenmode decays in a timescale tΛα1=O(N1)t\sim\Lambda_{\alpha}^{-1}=O(N^{-1}). The eigenmode relaxation time τ~α\tilde{\tau}_{\alpha} is evaluated as

τ~α=ln(δ)Λα+eEB1+eεln(21+eε),\tilde{\tau}_{\alpha}=-\frac{\ln(\delta)}{\Lambda_{\alpha}}+\frac{e^{E_{B}}}{1+e^{-\varepsilon}}\ln\left(\frac{2}{1+e^{-\varepsilon}}\right), (12)

whose second term is independent of the system size and Λα\Lambda_{\alpha}.

In this model, the “all-left” state, i.e., σ1=σ2==σN=1\sigma_{1}=\sigma_{2}=\dots=\sigma_{N}=1, is considered to be a many-body metastable state for large EBE_{B}. Expansion coefficients for this state are explicitly calculated, and it is found that Cα=ΨαC_{\alpha}=\Psi_{\alpha} and τα=τ~α\tau_{\alpha}=\tilde{\tau}_{\alpha} for all α\alpha, i.e., this state gives the maximum expansion coefficients for all eigenmodes.

Even in this simplest case, explosive growth of expansion coefficients occurs. However, Ψα\Psi_{\alpha} does not depend on EBE_{B}, and hence large Ψα\Psi_{\alpha} is not related to metastability in this non-interacting model.

Refer to caption
Figure 2: Spectrum gap Λ1\Lambda_{1} of GG against the system size NN for g=2g=2, 3, and 4. The saturated value of the gap at large NN is almost independent of NN.
Refer to caption
Figure 3: Eigemmode relaxation times {τα}\{\tau_{\alpha}\} against their eigenvalues Λα\Lambda_{\alpha} for different NN. We find a peak structure, and the peak shifts towards larger eigenvalues as NN increases. We also plot {τ~α}\{\tilde{\tau}_{\alpha}\}, which is indistinguishable from {τα}\{\tau_{\alpha}\}. (Inset) Logarithmic plot of τmax\tau_{\mathrm{max}} against gg for N=30N=30.
Refer to caption
Figure 4: Dynamics of the expectation value of N+N_{+} for different values of gg at N=30N=30. Vertical dashed lines represent τmax\tau_{\mathrm{max}} for each gg.

III.2 Interacting particles in a double-well potential

Let us introduce interactions. We here consider a simple model in which two particles in the same potential well equally interact with each other. The interaction energy is given by Eint=(g/N)[N+(N+1)/2+N(N1)/2]E_{\mathrm{int}}=-(g/N)[N_{+}(N_{+}-1)/2+N_{-}(N_{-}-1)/2], where g>0g>0 denotes the interaction strength (the scaling 1/N1/N ensures the extensivity of the energy). The interaction changes the effective potential felt by a single particle: the energy is written as E(N+)=ε+N++εNE(N_{+})=\varepsilon_{+}N_{+}+\varepsilon_{-}N_{-} with ε+=εg(N+1)/N\varepsilon_{+}=\varepsilon-g(N_{+}-1)/N and ε=g(N1)/N\varepsilon_{-}=-g(N_{-}-1)/N. The effective energy barrier from σi=±1\sigma_{i}=\pm 1 to 1\mp 1 is thus given by EB(±)=EB+εε±E_{B}^{(\pm)}=E_{B}+\varepsilon-\varepsilon_{\pm}. The model with interactions is obtained by replacing (ε,EB)(\varepsilon,E_{B}) in the non-interacting model by (ε+ε,EB(+))(\varepsilon_{+}-\varepsilon_{-},E_{B}^{(+)}). Obviously, interactions g>0g>0 effectively increases the energy barrier and causes slower relaxation, which we call interaction-induced metastability.

The master equation in this model is generated by a 2N2^{N}-dimensional generator. Instead of considering the full generator, let us employ the simplified description by focusing on the dynamics of a collective variable N+{0,1,,N}N_{+}\in\{0,1,\dots,N\} only. Since the energy only depends on N+N_{+} in our model, we can exactly obtain the dynamics of N+N_{+}. Let the probability distribution of N+N_{+} at time tt be denoted by PN+(t)P_{N_{+}}(t). Then the master equation for PN+(t)P_{N_{+}}(t) is written as

dPN+(t)dt=GN+,N++1PN++1(t)GN+,N+1PN+1(t)+(GN++1,N++GN+1,N+)PN+(t),\frac{dP_{N_{+}}(t)}{dt}=-G_{N_{+},N_{+}+1}P_{N_{+}+1}(t)-G_{N_{+},N_{+}-1}P_{N_{+}-1}(t)\\ +(G_{N_{+}+1,N_{+}}+G_{N_{+}-1,N_{+}})P_{N_{+}}(t), (13)

where GN+1,N+=N±eEB(±)G_{N_{+}\mp 1,N_{+}}=-N_{\pm}e^{-E_{B}^{(\pm)}}. This generator satisfies the detailed-balance condition of the form

GN+1,N+eF(N+)=GN+,N+1eF(N+1),G_{N_{+}-1,N_{+}}e^{-F(N_{+})}=G_{N_{+},N_{+}-1}e^{-F(N_{+}-1)}, (14)

where F(N+)=E(N+)S(N+)F(N_{+})=E(N_{+})-S(N_{+}) denotes the free energy of a state specified by N+N_{+}. The entropy S(N+)S(N_{+}) is given by S(N+)=ln(N!/(N+!N!)S(N_{+})=\ln(N!/(N_{+}!N_{-}!).

The (N+1)(N+1)-dimensional generator in Eq. (13) is a diagonal block of the original 2N2^{N}-dimensional full generator. Thus, the full generator has other 2N(N+1)2^{N}-(N+1) eigenmodes, which are dropped in this description. However, it does not affect the conclusion (see Appendix A).

We now fix EB=1E_{B}=1 and ε=5\varepsilon=5. In our model, there is the non-ergodic phase for g8.8g\gtrsim 8.8, in which we have an extensive free energy barrier ΔFN\Delta F\propto N and the relaxation time τrel\tau_{\mathrm{rel}} diverges in the thermodynamic limit as τrel=eΔF=eO(N)\tau_{\mathrm{rel}}=e^{\Delta F}=e^{O(N)}. We discuss the non-ergodic phase in Appendix B, and here we focus on the ergodic phase at which the free energy F(N+)F(N_{+}) is a monotonic function of N+N_{+}.

First let us investigate the gg-dependence of the spectral gap of GG. Numerical results for different gg and NN are presented in Fig. 2. It clearly shows that the spectral gap of GG is almost independent of the value of gg for large NN, although the finite-size effect is stronger for larger gg. It means that the spectrum gap Λ1\Lambda_{1} does not reflect the increase of the relaxation time due to interactions.

Next, we compute {τ~α}\{\tilde{\tau}_{\alpha}\} at g=2g=2 for different NN and show the result in Fig. 3. Although the typical value of Λα\Lambda_{\alpha} linearly grows with NN, that of τ~α\tilde{\tau}_{\alpha} looks convergent for large NN. According to Eq. (9), it means that expansion coefficients typically grows exponentially in NN, which is already seen in the non-interacting model. In Fig. 3, τα\tau_{\alpha} for the all-left initial state is also plotted by squares with a dashed line in Fig. 3, which completely agrees with τ~α\tilde{\tau}_{\alpha} as in non-interacting case. The all-left state has the maximum expansion coefficients for all eigenmodes, |Cα|=Ψα|C_{\alpha}|=\Psi_{\alpha}.

It turns out that the eigenmode α\alpha^{*} that gives the largest relaxation time τmax=maxατ~α=τ~α\tau_{\mathrm{max}}=\max_{\alpha}\tilde{\tau}_{\alpha}=\tilde{\tau}_{\alpha^{*}} appears in the middle of the spectrum, ΛαN\Lambda_{\alpha^{*}}\propto N, not at the spectrum edge. Since the relaxation particularly slows down far from equilibrium with large N+N_{+}, an extensive number of particles simultaneously undergo dissipation in a many-body metastable state, which would be the reason why eigenmodes with extensively large eigenvalues are important here.

We plot τmax\tau_{\mathrm{max}} against gg in the inset of Fig. 3. We find that τmax\tau_{\mathrm{max}} increases exponentially in gg, which is consistent with the intuitive picture that interactions effectively increase the energy barrier. This increase of the relaxation time stems from rapid growth of expansion coefficients with gg.

In order to compare the physical relaxation time with τmax\tau_{\mathrm{max}}, we numerically compute the dynamics of N+=N+N+PN+(t)\braket{N_{+}}=\sum_{N_{+}}N_{+}P_{N_{+}}(t) starting from the all-left initial state. The numerical result for different gg at N=30N=30 is shown in Fig. 4. The quantity N+\braket{N_{+}} reaches the stationary value roughly at t=τmaxt=\tau_{\mathrm{max}}, which is shown by vertical dashed lines in Fig. 4. It is thus confirmed that τmax\tau_{\mathrm{max}} coincides with the physical relaxation time of the system.

Refer to caption
Figure 5: Eigemmode relaxation times {τ~α}\{\tilde{\tau}_{\alpha}\} against their eigenvalues Λα\Lambda_{\alpha} for different LL in the FA model with ρ=0.1\rho=0.1. (Inset) τmax\tau_{\mathrm{max}} against LL.

III.3 Fredrickson-Andersen model

In the model considered so far, interaction-induced metastability is explained by the many-body explosion of expansion coefficients. Here we examine the FA model Fredrickson and Andersen (1984) as a prototypical example of kinetically constrained models Ritort and Sollich (2003) studied in glass and jamming transitions Berthier and Biroli (2011). Metastability in a kinetically constrained model is not due to energy barriers but due to dynamical rules. As we see below, the many-body explosion of expansion coefficients also explains this kind of metastability.

Suppose that each site i{1,2,,L}i\in\{1,2,\dots,L\} is occupied (ni=1n_{i}=1) or empty (ni=0n_{i}=0), and the transition rates from ni=0n_{i}=0 to 1 and that from ni=1n_{i}=1 to 0 are given by CiρC_{i}\rho and Ci(1ρ)C_{i}(1-\rho), respectively, where ρ\rho corresponds to the average density at equilibrium and CiC_{i} expresses dynamical constraints. In the FA model Fredrickson and Andersen (1984), Ci=1ni1ni+1C_{i}=1-n_{i-1}n_{i+1}, which means that the site ii can change its state only if at least one of the two neighbors i1i-1 and i+1i+1 is empty. Since this transition rate satisfies the detailed balance condition with respect to the non-interacting Hamiltonian H=μi=1LniH=-\mu\sum_{i=1}^{L}n_{i}, where μ=ln[ρ/(1ρ)]\mu=\ln[\rho/(1-\rho)] denotes the chemical potential, the equilibrium state is trivial but its dynamics exhibits interesting properties, which is a common characteristic of kinetically constrained models.

In the FA model, it is rigorously proved that the spectral gap of GG remains finite in the thermodynamic limit for any 0ρ<10\leq\rho<1 Cancrini et al. (2008). However, the maximum relaxation time exhibited by this model diverges in the thermodynamic limit for any ρ\rho. Because of the kinetic constraint, a cluster of particles can decay or grow only from its edges. Therefore, if the initial state has a macroscopically large cluster of size O(L)O(L), the relaxation time is at least of O(L)O(L). A state with macroscopically large clusters is therefore regarded as a metastable state. This increase of the maximum relaxation time should be due to an exponential growth of Ψα\Psi_{\alpha} 333At ρ=1/2\rho=1/2, the Hamiltonian vanishes and GG is a symmetric matrix. Even in this case, the many-body explosion of expansion coefficients occurs..

Under the periodic boundary condition, {τ~α}\{\tilde{\tau}_{\alpha}\} for ρ=0.1\rho=0.1 are computed and plotted in Fig. 5. We see that in the FA model Ψα=eO(L)\Psi_{\alpha}=e^{O(L)} for α\alpha with Λα=O(1)\Lambda_{\alpha}=O(1), which results in τmaxL\tau_{\mathrm{max}}\propto L even though the spectral gap remains finite.

IV Discussion

We have shown that the many-body explosion of eigenmode expansion coefficients is responsible for slow relaxation in many-body stochastic systems. We emphasize that this is a generic phenomenon, and it would also occur in nonequilibrium dynamics without detailed balance. Concepts related to slow relaxation, such as metastable states, must be reconsidered.

Since the maximum value Ψα\Psi_{\alpha} of the expansion coefficient for an eigenmode α\alpha is given by the inverse of the overlap between the left eigenvector Π(α)\vec{\Pi}^{(\alpha)} and the right eigenvector Φ(α)\vec{\Phi}^{(\alpha)}, the many-body explosion of Ψα\Psi_{\alpha} indicates a mismatch between the supports of these two vectors. Under the detailed balance condition, left and right eigenvectors are given by Φ(α)eβH/2W(α)\vec{\Phi}^{(\alpha)}\propto e^{-\beta H/2}\vec{W}^{(\alpha)} and Π(α)eβH/2W(α)\vec{\Pi}^{(\alpha)}\propto e^{\beta H/2}\vec{W}^{(\alpha)}, where W(α)\vec{W}^{(\alpha)} is the eigenvector of the symmetric matrix G=eβH/2GeβH/2G^{\prime}=e^{\beta H/2}Ge^{-\beta H/2} with the eigenvalue Λα\Lambda_{\alpha}. Therefore, the support of Φα\vec{\Phi}_{\alpha} (Π(α)\vec{\Pi}^{(\alpha)}) tends to be localized at low (high) energies, which may result in an extremely small overlap Π(α)Φ(α)=eO(N)\vec{\Pi}^{(\alpha)}\cdot\vec{\Phi}^{(\alpha)}=e^{-O(N)}. We however point out that explosive growth of Ψα\Psi_{\alpha} occurs even when GG is a symmetric matrix, i.e., H=0H=0, which is realized in the FA model with ρ=1/2\rho=1/2. We also emphasize that it would also occur in nonequilibrium stochastic dynamics without detailed balance.

In the two models examined in this work, {τ~α}\{\tilde{\tau}_{\alpha}\} are broadly distributed, and no clear separation in {τ~α}\{\tilde{\tau}_{\alpha}\} is observed. In such a case, it would be a difficult problem to identify metastable states Tǎnase-Nicola and Kurchan (2004); Kurchan and Levine (2011); Jack (2013). Existing mathematical theory on spectral characterization of metastable states Gaveau and Schulman (1998) provides a recipe to construct a metastable state as a linear combination of relevant eigenmodes of GG, but it is not applicable to our case, in which slow relaxation is caused by the many-body explosion of expansion coefficients. Although we can identify which eigenmodes are important for metastable states by looking at {τ~α}\{\tilde{\tau}_{\alpha}\}, it is not straightforward to find a systematic method to construct metastable states by using them. We leave it as a future problem.

Acknowledgements.
The author thanks Tatsuhiko Shirai for carefully reading the manuscript. This work was supported by JSPS KAKENHI Grant Numbers JP19K14622, JP21H05185.

Appendix A Diagonalization of full generator

In the main text, we consider the model of NN interacting particles in a double-well potential within the two-state approximation. In this model, each microscopic state is specified by the set of “spin” variables {σi}={σ1,σ2,,σN}\{\sigma_{i}\}=\{\sigma_{1},\sigma_{2},\dots,\sigma_{N}\}, where σi=+1\sigma_{i}=+1 (1-1) means that iith particle is in the left (right) potential well. In the main text, the master equation for the probability distribution PN+(t)P_{N_{+}}(t) of the collective variable N+=i=1N(σi+1)/2N_{+}=\sum_{i=1}^{N}(\sigma_{i}+1)/2 is examined.

Here we diagonalize the generator of the master equation for the probability distribution P{σi}(t)P_{\{\sigma_{i}\}}(t) over all microscopic states, which we call “the full generator”, and compare with the description using the collective variable N+N_{+}. The result for N=15N=15, g=2g=2, EB=1E_{B}=1, and ε=5\varepsilon=5 is shown in Fig. 6. In the calculation, in order to avoid eigenvalue degeneracies, we introduced very weak inhomogeneity: ε\varepsilon is replaced by εi=ε+δi\varepsilon_{i}=\varepsilon+\delta_{i}, where {δi}\{\delta_{i}\} are independent random variables uniformly drawn from [0,104][0,10^{-4}].

Figure 6 shows that the full generator contains all the eigenmodes of the generator of the master equation for the collective variable (shown by open circles). This is because the collective variable is completely decoupled from the other degrees of freedom in our model. In general, of course, coarse-grained description using collective variables is not exact.

Refer to caption
Figure 6: Eigenmode relaxation times {τ~α}\{\tilde{\tau}_{\alpha}\} obtained by the diagonalization of the full generator of the master equation for N=15N=15, g=2g=2, EB=1E_{B}=1, and ε=5\varepsilon=5. Those obtained by using the generator of the master equation for PN+(t)P_{N_{+}}(t) are shown by open circles.

Appendix B Non-ergodic phase

In the two-state model of interacting particles in the double-well potential, we have the non-ergodic phase for g>gcg>g_{c}, where gc8.8g_{c}\approx 8.8 for ε=5\varepsilon=5. When g<gcg<g_{c}, F(N+)F(N_{+}) is a monotonic function of N+N_{+}, whereas when g>gcg>g_{c}, it becomes non-monotonic and has a local minimum, which is due to the mean-field character of the model. The non-ergodicity in the thermodynamic limit is a result of an extensive free energy barrier ΔFN\Delta F\propto N. The local minimum is interpreted as a metastable state in this case, and its lifetime grows with NN as eΔF=eO(N)e^{\Delta F}=e^{O(N)}.

In the non-ergodic phase, metastability is fully explained by the gap closing of GG. We find that α=1\alpha^{*}=1 and Λ1eΔF\Lambda_{1}\propto e^{-\Delta F}. The corresponding maximum expansion coefficient Ψ1\Psi_{1} does not grow with NN. This is explained by the fact that relaxation of a metastable state across the free energy barrier is regarded as a single-particle problem with the cooridinate x=N+/Nx=N_{+}/N. There is no “many-body” explosion in the relevant expansion coefficient. We show numerical results for ε=5\varepsilon=5, EB=1E_{B}=1, and g=10>gcg=10>g_{c} in Fig. 7. The maximum eigenmode relaxation time τmax=maxατ~α\tau_{\mathrm{max}}=\max_{\alpha}\tilde{\tau}_{\alpha} agrees with const.×eΔF\text{const.}\times e^{\Delta F}, where ΔF0.113\Delta F\approx 0.113 (see the inset of Fig. 7.

Refer to caption
Figure 7: Eigenmode relaxation times {τ~α}\{\tilde{\tau}_{\alpha}\} against {Λα}\{\Lambda_{\alpha}\} in the non-erogdic phase. (Inset) τmax=maxατ~α\tau_{\mathrm{max}}=\max_{\alpha}\tilde{\tau}_{\alpha} against NN. It increases exponentially in NN, in agreement with const.×eΔF\text{const.}\times e^{\Delta F} that is shown by the dashed line. The parameters are chosen as ε=5\varepsilon=5, EB=1E_{B}=1, and g=10g=10.

References

  • Langer (1969) J. S. Langer, Statistical theory of the decay of metastable states, Ann. Phys. (N. Y). 54, 258–275 (1969).
  • Langer (1974) J. S. Langer, Metastable states, Physica 73, 61 (1974).
  • Penrose (1995) O. Penrose, Metastable decay rates, asymptotic expansions, and analytic continuation of thermodynamic functions, J. Stat. Phys. 78, 267–283 (1995).
  • Bray and Moore (1980) A. J. Bray and M. A. Moore, Metastable states in spin glasses, J. Phys. C Solid State Phys. 13, L469 (1980).
  • Biroli and Kurchan (2001) G. Biroli and J. Kurchan, Metastable states in glassy systems, Phys. Rev. E 64, 016101 (2001).
  • Berthier and Biroli (2011) L. Berthier and G. Biroli, Theoretical perspective on the glass transition and amorphous materials, Rev. Mod. Phys. 83, 587–645 (2011)arXiv:1011.2578 .
  • Baldwin et al. (2011) A. J. Baldwin, T. P. Knowles, G. G. Tartaglia, A. W. Fitzpatrick, G. L. Devlin, S. L. Shammas, C. A. Waudby, M. F. Mossuto, S. Meehan, S. L. Gras, J. Christodoulou, S. J. Anthony-Cahill, P. D. Barker, M. Vendruscolo, and C. M. Dobson, Metastability of native proteins and the phenomenon of amyloid formation, J. Am. Chem. Soc. 133, 14160–14163 (2011).
  • Griffiths et al. (1966) R. B. Griffiths, C. Y. Weng, and J. S. Langer, Relaxation times for metastable states in the mean-field model of a ferromagnet, Phys. Rev. 149, 301–305 (1966).
  • Lanford and Ruelle (1969) O. E. Lanford and D. Ruelle, Observables at infinity and states with short range correlations in statistical mechanics, Commun. Math. Phys. 13, 194–215 (1969).
  • Note (1) See Refs. Langer (1969); Penrose (1995) for attempts to give a static characterization of finite-lifetime metastable states associated with first-order phase transitions via an analytic continuation of the thermodynamic function.
  • Gaveau and Schulman (1998) B. Gaveau and L. S. Schulman, Theory of nonequilibrium first-order phase transitions for stochastic dynamics, J. Math. Phys. 39, 1517–1533 (1998).
  • Bovier et al. (2000) A. Bovier, M. Eckhoff, V. Gayrard, and M. Klein, Metastability and small eigenvalues in Markov chains, J. Phys. A. Math. Gen. 33, L447 (2000).
  • Macieszczak et al. (2016) K. Macieszczak, M. Guţă, I. Lesanovsky, and J. P. Garrahan, Towards a Theory of Metastability in Open Quantum Dynamics, Phys. Rev. Lett. 116, 240404 (2016)arXiv:1512.05801 .
  • Rose et al. (2016) D. C. Rose, K. Macieszczak, I. Lesanovsky, and J. P. Garrahan, Metastability in an open quantum Ising model, Phys. Rev. E 94, 052132 (2016)arXiv:1607.06780 .
  • (15) D. C. Rose, K. Macieszczak, I. Lesanovsky, and J. P. Garrahan, Hierarchical classical metastability in an open quantum East model, arXiv:2010.15304 .
  • Frenkel (2013) D. Frenkel, Simulations: The dark side, Eur. Phys. J. Plus 128, 10 (2013).
  • Fredrickson and Andersen (1984) G. H. Fredrickson and H. C. Andersen, Kinetic Ising model of the glass transition, Phys. Rev. Lett. 53, 1244–1247 (1984).
  • Mori and Shirai (2020) T. Mori and T. Shirai, Resolving a Discrepancy between Liouvillian Gap and Relaxation Time in Boundary-Dissipated Quantum Many-Body Systems, Phys. Rev. Lett. 125, 230604 (2020)arXiv:2006.10953 .
  • Haga et al. (2021) T. Haga, M. Nakagawa, R. Hamazaki, and M. Ueda, Liouvillian Skin Effect: Slowing down of Relaxation Processes without Gap Closing, Phys. Rev. Lett. 127, 070402 (2021)arXiv:2005.00824 .
  • (20) J. Bensa and M. Znidaric, Fastest local entanglement scrambler, multistage thermalization, and a non-Hermitian phantom, arXiv:2101.05579 .
  • Gaveau and Schulman (1987) B. Gaveau and L. S. Schulman, Dynamical metastability, J. Phys. A. Math. Gen. 20, 2865–2873 (1987).
  • Note (2) If we choose another normalization condition for right eigenvectors, the criterion of the relaxation |Cα|eΛαt=δ|C_{\alpha}|e^{-\Lambda_{\alpha}t}=\delta should be modified. For example, if we employ the normalization Φ(α)2:=(\sumop\slimits@n=1D|Φn(α)|2)1/2=1\|\vec{\Phi}^{(\alpha)}\|_{2}:=\left(\sumop\slimits@_{n=1}^{D}|\Phi^{(\alpha)}_{n}|^{2}\right)^{1/2}=1, we have |OΦ(α)|D1/2|\vec{O}\cdot\vec{\Phi}^{(\alpha)}|\sim D^{1/2}, and hence the criterion of the relaxation should be |Cα|eΛαt=δD1/2|C_{\alpha}|e^{-\Lambda_{\alpha}t}=\delta D^{-1/2} with δ\delta being a fixed constant. Accordingly, the formula (7) should also be modified. In other words, Eq. (7) is appropriate only for the normalization scheme (2).
  • Ritort and Sollich (2003) F. Ritort and P. Sollich, Glassy dynamics of kinetically constrained models, Adv. Phys. 52, 219–342 (2003)arXiv:0210382 [cond-mat] .
  • Cancrini et al. (2008) N. Cancrini, F. Martinelli, C. Roberto, and C. Toninelli, Kinetically constrained spin models, Probab. Theory Relat. Fields 140, 459–504 (2008).
  • Note (3) At ρ=1/2\rho=1/2, the Hamiltonian vanishes and GG is a symmetric matrix. Even in this case, the many-body explosion of expansion coefficients occurs.
  • Tǎnase-Nicola and Kurchan (2004) S. Tǎnase-Nicola and J. Kurchan, Metastable states, transitions, basins and borders at finite temperatures, J. Stat. Phys. 116, 1201–1245 (2004)arXiv:0311273 [cond-mat] .
  • Kurchan and Levine (2011) J. Kurchan and D. Levine, Order in glassy systems, J. Phys. A Math. Theor. 44, 035001 (2011)arXiv:1008.4068 .
  • Jack (2013) R. L. Jack, Counting metastable states in a kinetically constrained model using a patch repetition analysis, Phys. Rev. E 88, 062113 (2013).