Path integrals for stochastic hybrid reaction-diffusion processes

Paul C. Bressloff Department of Mathematics, University of Utah, Salt Lake City, UT, USA

We construct a functional path integral for a stochastic hybrid reaction-diffusion (RD) equation, in which the reaction term depends on the discrete state of a randomly switching environment. We proceed by spatially discretizing the RD system and using operator methods and coherent spin states to derive a path integral representation of the lattice model. The path integral specifies the distribution of trajectories in a state-space consisting of the set of local concentrations and the environmental state. Taking the continuum limit yields a path integral for the stochastic hybrid RD equation. We show how in the semi-classical limit, the stochastic dynamics reduces to a system of Langevin equations. The latter consists of an RD equation coupled to a stochastic probability function z that evolved according to an Ornstein-Uhlenbeck process with multiplicative noise. On eliminating the variable z, we obtain an RD equation driven by nonlocal spatiotemporal noise. This is used to determine leading order corrections to the mean-field RD equation. We then apply the theory to the example of diffusion with stochastically gated adsorption under dimerization. Com- parison with the classical problem of diffusion-limited pair-annihilation suggests that the adsorption model exhibits anomalous large-time behavior in one and two spatial . This motivates an alternative path integral formulation based on an RD master equation that treats diffusion as a set of hopping reactions and incorporates the effects of a randomly switching environment. A small noise expansion now yields a stochastic RD equation with both real and complex noise sources. Finally, diagrammatic and methods are used to determine the anomalous large-time behavior of the mean concentration at the upper critical dc = 2.

I. INTRODUCTION in the weak noise (adiabatic) limit. In particular, we have shown how to derive a hybrid path integral using two alternative methods: (i) integral representations of the Diffusion processes in randomly switching environ- Dirac delta function [32, 33], which is analogous to the ments are finding increasing applications in cell biology construction of path integrals for stochastic differential and biophysics. Examples include diffusion in domains equations (SDEs) [34–36]; (ii) bra-kets and “quantum- with stochastically-gated boundaries [1–6], diffusion over mechanical” operators [37], similar in spirit to the Doi- fluctuating barriers [7–10], and stochastic gap junctions Peliti formalism for master equations [38–41]. In both [11]. Mathematically speaking, diffusion in a randomly cases the Hamiltonian of the resulting action functional switching environment is an infinite-dimensional version corresponds to the principal eigenvalue of a linear opera- of a so-called stochastic hybrid system. Stochastic hybrid tor, which combines the generator of the discrete Markov systems involve a coupling between a discrete Markov process and the vector fields of the piecewise determin- chain and a continuous stochastic process [12]. If the lat- istic dynamics. This is consistent with more rigorous re- ter evolves deterministically between jumps in the dis- sults obtained using large deviation theory [42–45]. Re- crete state, then the system reduces to a piecewise de- cently, we have derived a different hybrid path integral terministic Markov process (PDMP) [13]. Well known representation that holds outside the weak noise regime examples of finite-dimensional hybrid systems include [46]. Rather than using the “bosonic” annihilation and stochastic gene expression [14–19], voltage fluctuations in creation operators of the Doi-Peliti formalism, we de- neurons [20–27], and motor-driven intracellular transport composed the discrete states in terms of coherent “spin” [28–31]. One method for analyzing the diffusion equation states [47, 48]. Such a decomposition has recently been with switching boundary conditions is to discretize space used to study stochastic gene expression in the presence and construct the Chapman-Kolmogorov (CK) equation of promoter noise and low protein copy numbers [49–51]. for the resulting finite-dimensional stochastic hybrid sys- For example, one can effectively map the stochastic dy- tem [5]. The CK equation can then be used to derive mo- namics of a single genetic switch to a quantum spin-boson ment equations for the stochastic concentration. In the system. continuum limit, this yields a hierarchy of moment equa- arXiv:2103.07759v1 [cond-mat.stat-mech] 13 Mar 2021 tions, with the equation at r-th order taking the form In this paper, we use coherent spin states to construct a of an r-dimensional parabolic partial differential equa- path-integral formulation of a stochastic hybrid reaction- tion (PDE) that couples to lower order moments at the diffusion (RD) equation, under the assumption that the boundaries. Although the diffusing particles are non- switching environment affects the reaction term rather interacting, statistical correlations arise at the popula- than the boundary conditions. A simple example is diffu- tion level due to the fact that they all move in the same sion with stochastically-gated adsorption, where the ad- randomly switching environment. sorption rate depends on the current state of the environ- We have previously developed path integral methods ment [52]. We proceed by first constructing a path in- for studying finite-dimensional stochastic hybrid systems tegral representation of the spatially discretized RD sys- 2 tem and then taking a continuum limit. We show how a Note that U(x, t) is a stochastic concentration field. In semi-classical approximation of the resulting continuum the case of a finite interval, Ω = [0,L], the PDE has to path integral action yields the action for a stochastic RD be supplemented by boundary conditions at x = 0,L. equation with real, spatially colored noise. The latter is In previous work [5], we took the boundary conditions used to determine leading order corrections to the mean- themselves to depend on the environmental state (with field RD equation. We illustrate the theory by consid- Fn = 0). That is, ering the example of diffusion with stochastically gated adsorption. Comparison with the classical problem of bnU(0, t) + cn∂xU(0, t) = dn, (2.3a) diffusion-limited pair-annihilation suggests that the ad- 0 0 0 bnU(L, t) + cn∂xU(L, t) = dn (2.3b) sorption model exhibits anomalous large-time behavior in one and two spatial dimensions [53]. Therefore, we 0 0 for N(t) = n and constant coefficients bn, bn, cn, cn, derive an alternative path-integral representation based 0 dn, dn. We showed that one way to analyze the effects on an RD master equation that includes the effects of a of a random environment is to spatially discretize the switching environment, and treats diffusion as a set of PDE so that it is converted to a stochastic hybrid ODE hopping reactions that supplement the reaction associ- [5]. Introducing the lattice spacing h and setting uj = ated with absorption. A small noise expansion now yields u(jh), j ∈ J = {0, 1,..., N} with N a = L, leads to the a stochastic RD equation with both real and complex piecewise deterministic ODE noise sources. Finally, we use diagrammatic and renor- malization group (RG) methods in statistical field theory dUi X = ∆n U + F (U ), i ∈ J , (2.4) [54–58] to determine the anomalous large-time behavior dt ij j n i of the mean concentration at the upper critical dimension j∈J d = 2. c n where U = (Uj, j ∈ J ) and ∆ij is the discrete Laplacian The structure of the paper is as follows. In Sect. II we n introduce the stochastic hybrid RD equation for a sin- for N(t) = n. Away from the boundaries, ∆ij = ∆ij gle chemical species, and derive moment equations for with the concentration. We introduce a two-state model for D stochastically-gated adsorption, and indicate how a non- ∆ij = [δi,j+1 + δi,j−1 − 2δi,j], 0 < i < N , (2.5) h2 linear adsorption rate can occur under dimerization, lead- ing to a moment closure problem. In Sect. III, we present n whereas ∆ij is modified at the boundaries i = 0, N in a detailed derivation of the path integral for the spatially order to be consistent with the boundary conditions (see discretized hybrid RD model. We perform the small noise [5] for details). The corresponding probability density expansion of the continuum path integral action in Sect.

IV and derive leading order corrections to the determin- Prob{U(t) ∈ (u, u + du),N(t) = n} = Pn(u, t)du (2.6) istic mean-field RD equation. In Sect. V we apply the theory to the example of stochastically-gated adsorption evolves according to a differential Chapman-Kolmogorov under dimerization and construct the alternative path in- (CK) equation for the stochastic hybrid system (2.4): tegral representation based on an RD master equation. Finally, the RG analysis is developed in Sect. VI.    ∂Pn X ∂ X n = −  ∆ijuj + Fn(ui) Pn(u, t) ∂t ∂ui i∈J j∈J II. STOCHASTIC HYBRID PDE X + QnmPm(u, t). (2.7) Consider a one-dimensional RD process in which the m∈I reaction term depends on the current discrete state of It is important to note that the above spatial dis- the environment, N(t). The latter is assumed to evolve cretization is distinct from the scheme used to derive according to a continuous time Markov chain on I ⊂ an RD master equation [39–41]. In the derivation of Z+. Setting P (t) = [N(t) = n], we have the master n P Eq. (2.7) we assumed that h is sufficiently large so that equation the expected number of molecules per lattice site is large enough to write down a local concentration. In contrast, dPn X X = Q P (t), Q = 0, (2.1) dt nm m nm the derivation of an RD master equation allows h to be m∈I n∈I smaller and considers the discrete number of molecules per lattice site. The diffusive hopping between neighbor- where Q is the matrix generator. In between jumps in the ing lattice sites is treated as a set of single step reactions, environmental state, with N(t) = n, the concentration that supplement the local chemical reactions. U(x, t) evolves according to the piecewise RD equation Now define the conditional first moments ∂U ∂2U Z = D + F (U), x ∈ Ω ⊆ . (2.2) ∂t ∂x2 n R Un,k(t) = E[Uk(t)1N(t)=n] = du Pn(u, t)uk(t), (2.8) 3

R Q R ∞ where du = j −∞ duj. Multiplying both sides of such that the interface is adsorbing when N(t) = 0 and the CK Eq. (2.7) by uk(t) and integrating by parts with reflecting when N(t) = 1, see Fig.1. (One application of respect to u yields the moment equation such a model is to diffusive intracellular transport [52].) The matrix generator is N dUn,k X n X   = ∆kjUn,j +E[Fn(Uk)1N(t)=n]+ QnmUm,k. −β α dt Q = , (2.15) j=1 m∈I β −α (2.9)

For the sake of illustration, suppose that Fn(u) = −γnu and there exists a steady-state distribution ρn of the mas- where γn is some environmentally-dependent absorption ter Eq. (2.1) with rate [52]. In this case, E[Fn(Uk)1N(t)=n] = −γnUn,k, and Eq. (2.9) becomes a closed equation for the first β α ρ = , ρ = . (2.16) moments. Retaking the continuum limit h → 0 then 0 α + β 1 α + β yields the coupled system of deterministic PDEs for Un(x, t) = E[U(x, t)1N(t)=n] If the rate of adsorption per unit concentration is γ, then F0(u) = −γu and F1(u) = 0. For the sake of generality, 2 ∂Un ∂ Un(x, t) X we also impose the switching boundary conditions = D − γ U (x, t) + Q U (x, t) ∂t ∂x2 n n nm m m∈I ∂ U(0, t) = 0 = ∂ U(L, t) when N(t) = 0 (2.10) x x U(0, t) = A1,U(L, t) = 0 when N(t) = 1. for x ∈ [0,L], with the boundary conditions Eq. (2.10) reduces to the pair of equations bnUn(0, t) + cn∂xUn(0, t) = dn, (2.11a) 2 0 0 0 ∂U0 ∂ U0 bnUn(L, t) + cn∂xUn(L, t) = dn. (2.11b) = D − γU − βU + αU , (2.17a) ∂t ∂x2 0 0 1 As we have previously highlighted [5], since all the parti- ∂U ∂2U 1 = D 1 + βU − αU , (2.17b) cles diffuse in the same randomly switching environment, ∂t ∂x2 0 1 there are non-trivial statistical correlations between the particles. For example, consider the second-order mo- with ments ∂xU0(0, t) = ∂xU0(L, t) = 0, U1(0, t) = A1, U1(L, t) = 0. Cn(x, y, t) = E[U(x, t)U(y, t)1N(t)=n]. (2.12) (2.18) These evolve according to the moment equation

2 2 no adsorption ∂Cn ∂ Cn ∂ Cn X = D + D − 2γ C + Q C , ∂t ∂x2 ∂y2 n n nm m N(t) = 1 m∈I (2.13) diffusion together with boundary conditions that couple to the first-order moments [5]. The latter can be derived from the spatially discretized CK Eq. (2.7) after multiply- α β ing both sides by the product uk(t)ul(t), integrating by parts and retaking the continuum limit. Clearly Cn(x, y, t) 6= Un(x, t)Un(y, t), which means that the two- point correlation function is non-zero. A similar com- γ adsorption ment holds for higher-order moments. N(t) = 0 diffusion A. Two-state model with adsorption. x = 0 x = L For the sake of illustration, consider one-dimensional diffusion in a domain with an adsorbing interface. Sup- FIG. 1. Schematic diagram of diffusion in a one-dimensional pose that the environment switches between two states domain with stochastically gated adsorption. The environ- ment switches between two discrete states according to a two- N(t) ∈ {0, 1} according to the two-state Markov chain state Markov chain with transition rates α, β. Adsorption can only occur when N(t) = 0. In addition, the ends are closed β 0 1, (2.14) when N(t) = 0 and open when N(t) = 1 α 4

The steady-state solution can be calculated by setting adsorption of dimers) so that A(t) = 0 whenever N(t) = 2 all time derivatives to zero and performing the change of 0, and hence γ2A  γ1U when N(t) = 0. (The latter variables βUb0 = βU0 − αU1. This gives condition will hold provided that the switching rates α, β are sufficiently fast, which is the weak noise assumption 2 2 d Ub0 α d U1 αγ considered in section IV.) Under these conditions, we can 0 = D − (γ + β)Ub0 + D − U1, (2.19a) 2 dx2 β dx2 β take the effective adsorption rate of U to be −γ1U . 2 d U1 0 = D 2 + βUb0. (2.19b) dx III. CONSTRUCTION OF A SPATIALLY Eqs. (2.19a)and Eq. (2.19b) lead to the fourth-order DISCRETIZED HYBRID PATH INTEGRAL equation We now use the operator method recently introduced d4U d2U 1 = Γ 1 − Γ U , (2.20) in [46] to construct a path integral representation of the dx4 2 dx2 0 1 spatially discretized hybrid system (2.4). Taking the con- tinuum limit then yields a functional path integral for the with stochastic hybrid RD Eq. (2.2), which can be used to de- γ + α + β αγ velop various weak noise approximations, see sections 4 Γ = , Γ = . 2 D 0 D2 and 5. For convenience, we ignore boundary effects by taking x ∈ in Eq. (2.2) so that the lattice in Eq. (2.4) This has a general solution of the form R is J = Z. Alternatively, we could take x ∈ [0,L] and im-

λ+x −λ+x λ−x −λ−x pose periodic boundary conditions so that J is finite. In U0(x) = a+e + b+e + a−e + b−e , both cases, the discrete Laplacian is given by Eq. (2.5) with the coefficients a±, b± determined by the boundary for all i ∈ J and is thus independent of the environ- conditions, and mental state. In order to develop the basic theory we focus on the case of two discrete states, N(t) ∈ {0, 1} as s 1  q  exemplified by the adsorption model of Sect. II.A. λ = Γ ± Γ2 − 4Γ . ± 2 2 2 0

In the special case γ = 0, the general solution becomes A. Coherent spin states

Γ2x −Γ2x U1(x) = a+e + a−e + b+x + b−. Consider the master equation for a two-state Markov chain, written in matrix form The steady-state equations for higher-order moments can be solved in a similar fashion. dP = QP(t), P(t) = (P (t),P (t))>, (3.1) A non-trivial extension of the above model is to as- dt 0 1 sume that when the environment is in the adsorbing state N(t) = 0, the rate of adsorption is a nonlinear function with Q the matrix (2.15). Introduce the Pauli spin ma- 2 trices of u. For example, suppose that F0(u) = −γu . The term −γU0 in Eq. (2.17a) is now replaced by the second 1  0 1  1  0 −i  1  1 0  σ = , σ = , σ = , order moment −γC0(x, x, t) and we have a moment clo- x 2 1 0 y 2 i 0 z 2 0 −1 sure problem. (The latter can be handled using a weak (3.2) noise approximation, see Sect. IV.) One possible mecha- and set σ± = σx ± iσy. It follows that the generator can nism for the quadratic rate of adsorption is dimerization. be rewritten as Let U denote the concentration of monomers and A the concentration of dimers. Consider the generalized RD 1  1  Q = −β 1 + σ −α 1 − σ +ασ +βσ . (3.3) system 2 z 2 z + − ∂U ∂2U = D − γ U 2 + γ A, (2.21a) Next we define the coherent spin-1/2 state [47, 49–51] ∂t ∂x2 1 2 2  eiφ/2 cos2 θ/2  ∂A ∂ A 2 = D + γ U − γ A − κδ A. (2.21b) |si = −iφ/2 2 , 0 ≤ θ ≤ π, 0 ≤ φ < 2π, ∂t ∂x2 1 2 n,0 e sin θ/2 (3.4) for N(t) = n. Here γ1 is the rate of dimerization under together with the adjoint the reaction scheme  −iφ/2 iφ/2 γ1 hs| = e , e . (3.5) U + U U2, γ2 Note that and adsorption only occurs for dimers at a rate κ when 0 0 N(t) = 0. Furthermore, suppose that κ → ∞ (instant hs0|si = ei(φ−φ )/2 cos2 θ/2 + e−i(φ−φ )/2 sin2 θ/2, (3.6) 5 so that hs|si = 1 and It immediately follows that |vi is an eigenvector of the momentum operatorv ˆk, since 1 hs + ∆s|si = 1 − i∆φ cos θ + O(∆φ2). (3.7)  ←  2 Z Z iv·u ∂ iv·u vˆk|vi = du e −i  |ui = du e vk|ui We also have the completeness relation ∂vk 1 Z π Z 2π = v |vi. (3.16) sin θ dθ dφ |sihs| = 1. (3.8) k 2π 0 0 Using the inverse Fourier transform, we also have It can checked that the following identities hold: Z Z Z ∞ Y dvj 1 |ui = dv e−iv·u|vi, dv = , (3.17) hs|σ |si = cos θ, (3.9a) 2π z 2 j∈J −∞ 1 1 hs|σ |si = eiφ sin θ, hs|σ |si = e−iφ sin θ. (3.9b) + 2 − 2 and the completeness relation Hence, Z dv|vihv| = 1. (3.18) 1 + cos θ hs|Q|si = Q(θ, φ) ≡ −β 1 − eiφ 2 1 − cos θ − α 1 − e−iφ . (3.10) C. Operator version of CK equation 2 Introduce the state vectors B. Hilbert space for continuous states Z |ψn(t)i = du Pn(u, t)|ui, n = 0, 1. (3.19) The Hilbert space spanned by the continuous vectors |ui is taken to have the inner product Differentiating both sides with respect to time and using Eq. (2.7) with ∆n = ∆ gives 0 Y 0 ij ij hu |ui = δ(uj − uj), j∈J Z  d X ∂Fj(n, u)Pn(u, t) |ψ (t)i = du − dt n ∂u and is based on a recent construction of path integrals j j for SDEs developed in Ref. [59]. Define the conjugate X  pair of operatorsu ˆj, vˆj, j ∈ J such that + QnmPm(u, t) |ui, m=0,1 ← ∂   uˆj|ui = uj|ui, vˆj|xi = −i |ui. (3.11) X X ∂uj = −iδn,m vˆjFj(n, uˆ) + Qnm |ψm(t)i, m=0,1 j The arrow on the differential operator indicates that it operates to the left. Alternatively, the action ofv ˆj can where we have set be defined in terms of state vectors, X Fj(n, u) = ∆jkuk + Fn(uj). (3.20) ∂ψ Z k∈J hu|vˆj|ψi = −i , |ψi = du ψ(u)|ui. (3.12) ∂uj Let The operators satisfy the commutation relations X Hbn = −i vˆjFj(n, uˆ) n = 0, 1, (3.21) [ˆu , vˆ ] = iδ , [ˆu , uˆ ] = 0, [ˆv , vˆ ] = 0. (3.13) j k j,k j k j k j

The Hilbert space has the completeness relation and consider the diagonal matrix operator Z du |uihu| = 1. (3.14)       Hb0 0 1 1 Kb = = 1 + σz Hb0 + 1 − σz Hb1. 0 Hb1 2 2 It is also convenient to introduce the “momentum” rep- (3.22) resentation (analogous to taking Fourier transforms), We can then rewrite (2.7) ) as the operator equation Z iv·u d |vi = du e |ui. (3.15) |ψ(t)i = Hb |ψ(t)i, (3.23) dt 6

> with |ψ(t)i = (|ψ0(t)i, |ψ1(t)i) and In the limit N → ∞ and ∆t → 0 with N∆t = t fixed, we make the approximation 1  1  Hb = Kb + Q = 1 + σz Hb0 + 1 − σz Hb1 2 2 hs , u |eHb ∆t|s , u i ≈ hs , u |1 + H∆t|s , u i 1  1  `+1 `+1 ` ` `+1 `+1 b ` ` − β 1 + σ − α 1 − σ + ασ + βσ .  2 z 2 z + − = hs`+1|s`i δ(u`+1 − u`) (3.24)  2 Moreover, + hu`+1|H(θ`, φ`, u`, vˆ`)∆t|u`i + O(∆t ). (3.30)

  iφ  1 + cos θ hs|Hb |si = H(θ, φ, uˆ, vˆ) ≡ − β 1 − e − Hb0 2 In addition, Eq. (3.7) implies that   −iφ  1 − cos θ − α 1 − e − Hb1 . (3.25) 2 1 2 hs`+1|s`i = 1 − i(φ`+1 − φ`) cos θ` + O(∆φ ) Formally integrating Eq. (3.23) gives 2 1 dφ` 2 = 1 − i∆t cos θ` + O(∆t ). (3.31) |ψ(t)i = eHb t|ψ(0)i. (3.26) 2 dt Each small-time propagator thus becomes

D. Spatially discrete path integral Hb ∆t hs`+1, u`+1|e |s`, u`i (3.32) In order to derive a path integral representation of the  i dφ   ≈ hu | exp H(θ , φ , u , vˆ ) − ` cos θ ∆t |u i. formal solution (3.26), we divide the time interval [0, t] `+1 ` ` ` ` 2 dt ` ` into N subintervals of size ∆t = t/N and set Substituting the momentum completeness relation (3.18) |ψ(t)i = eHb ∆teHb ∆t ··· eHb ∆t|ψ(0)i. (3.27) into the small-time propagator (3.32) then gives We then insert multiple copies of appropriately chosen completeness relations. Here we use the completeness re- Z Hb ∆t lations (3.8) and (3.14), which are applied to the product hs`+1, u`+1|e |s`, uji ≈ dv`hu`+1|v`ihv`|u`i Hilbert space with |s, ui = |si⊗|ui. Introducing the solid  i dφ   angle integral × exp H(θ , φ , u , v ) − ` cos θ ∆t . (3.33) ` ` ` ` 2 dt ` Z 1 Z π Z 2π ds = sin θ dθ dφ, (3.28) Ω 2π 0 0 Furthermore, we have iv`·(u`+1−u`) Z Z Z Z hu`+1|v`ihv`|u`i = e (3.34)   |ψ(t)i = ds0 ··· dsN du0 ··· dun |sN , uN i duj 2 Ω Ω = exp iv` · ∆t + O(∆t ). "N−1 # dt Y Hb ∆t hs`+1, u`+1|e |s`, u`i hs0, u0|ψ(0)i, (3.29) `=0 Substituting Eqs. (3.33) and (3.34) into (3.29) yields

Z Z Z Z |ψ(t)i = ds0 ··· dsN du0dv0 ··· duN dvN |sN , uN i Ω Ω N−1    Y i dφ` duj × exp H(θ , φ , u , v ) − cos θ + iv · ∆t hs , u |ψ(0)i. ` ` ` ` 2 dt ` ` dt 0 0 `=0

The final step is to take the limit N → ∞, ∆t → 0 with that < u0|ψn(0)i = ρnδ(u − u0), and set N∆t = t fixed, u` = u(`∆t) etc. We will also assume

Pn(u, t|u0, 0) = hu, n|ψ(t)i. 7

After Wick ordering, v → −iv, integrating by parts the A. Semi-classical limit term involving dφ/dt, and performing the change of co- ordinates z = (1 + cos θ)/2, we obtain the following func- In general it is not possible to evaluate a stochas- tional path integral: tic path integral without some form of approximation scheme. One of the best known is the so-called semi- Z u(t)=u classical approximation, which involves expanding the P (u, t|u , 0) = N D[z]D[φ]D[u]D[v] n 0 n path integral action to second order in the momentum u(0)=u0 variables, assuming that the system operates in the weak  Z t  du dz   exp − v · − iφ − H dτ , (3.35) noise regime. In the case of the piecewise PDE (2.2), we 0 dτ dτ define the weak noise limit by introducing the scalings α → α/, β → β/ with  → 0 representing the adiabatic where H is the effective “Hamiltonian” limit of fast switching between the discrete states. Intro-   ducing the additional scaling φ → φ, the path integral action for the two-state model becomes  iφ X H = − β 1 − e − [vj∆jkuk + F0(uj)] z Z t  Z ∂u  j∈J S = dτ dx v − D∇2u d ∂τ   0 R  −iφ X dz β α − α 1 − e − vj[∆jkuk + F1(uj)] (1 − z), − iφ + 1 − eiφ z + 1 − e−iφ (1 − z) j∈J dτ   Z Z  (3.36) − z dx vF0(u) − (1 − z) dx vF1(u) (4.3) d d R R with “position” coordinates (u, z) and “conjugate” mo- Under the approximation 1−e±iφ = ∓iφ+2φ2/2+..., menta (v, −iφ). Here N is a normalization constant. n the action is quadratic in φ. Comparison with the path Finally, note that the derivation of the path integral car- integral action of an SDE [34–36] then establishes that ries over to the spatially discretized version of the higher- the resulting path integral represents the probability den- dimensional stochastic hybrid RD equation sity functional of an effective stochastic processes evolv- ing according to the following pair of coupled Langevin ∂U = D∇2U + F (U). (3.37) equations: ∂t n ∂U 2 = D∇ U + F0(U)z + F1(U)(1 − z), (4.4a) The only change is that ∆ij is the discrete Laplacian on ∂t a d-dimensional square lattice. We will assume d ≥ 1 in dz √  = −βz + α(1 − z) + ξ , (4.4b) the following. dt z

where ξz is an independent Gaussian white noise pro- cesses with hξzi = 0 and IV. CONTINUUM PATH INTEGRAL AND THE 0 0 WEAK NOISE LIMIT hξz(t)ξz(t )i = (βz + α(1 − z))δ(t − t ). (4.4c) Note that Eq. (4.4b) is in the form of an Ornstein- Having obtained the path integral of the spatially dis- Uhlenbeck process with multiplicative noise. The latter cretized hybrid system evolving according to the CK Eq. ensures that if z(0) ∈ [0, 1] then z(t) ∈ [0, 1] for all t > 0, (2.7), we now take the continuum limit to determine the since the noise vanishes at z = 0, 1. Therefore, the auxil- corresponding functional path integral for the original iary variable z(t) represents the effective probability that stochastic hybrid diffusion Eq. (2.2): N(t) = 0 at t = 0. If 0 <   1 then the stochastic variable z may be elim- Z P [u] = N D[z]D[φ]D[u]D[v]e−S[u,v,z,φ], inated using a linear noise approximation. First, taking n n ∗ u(x,t)=u(x) the limit  → 0 shows that z(t) → z = α/(α + β) and (4.1) U(x, t) satisfies the deterministic mean-field RD equation

∂U 2 ∗ ∗ with action = D∇ U+F (U), F (U) = F0(U)z +F1(U)(1−z ). ∂t t (4.5) Z  dz ∗ ∗ S = − iφ + β 1 − eiφ z + α 1 − e−iφ (1 − z) Clearly ρn, n = 0, 1 with ρ0 = z and ρ1 = 1 − z is 0 dτ the stationary distribution of the two-state Markov chain Z   Z ∗ ∂u 2 (3.1). Setting z(t) = z + y(t) and substituting into Eq. + dx v − D∇ u − z dx vF0(u) (4.2) (4.4b) implies that to leading order, d ∂τ d R R √ Z   − (1 − z) dx vF1(u) dτ. y(t) = ξ0(t), d R α + β 8 with Let u be the solution to the deterministic RD Eq. (4.5) and introduce the linear noise approximation hξ (t)i = 0, hξ (t)ξ (t0)i = 2βz∗δ(t − t0). 0 0 0 √ U(x, t) = u(x, t) + Ue(x, t). (4.12) Applying the linear noise approximation to Eq. (4.4a) then gives We then have

∂U 2 ∂hUi 2 √ = D∇ U + F (U) + [F0(U) − F1(U)]y(t) = D∇ hUi + hF (u + Ue)i ∂t ∂t 2 √ √ = D∇ U(x, t) + F (U(x, t)) + ξ(x, t), (4.6) = D∇2hUi + hF (hUi + (Ue − hUei)i √ 0 where hξ(x, t)i = 0, and = D∇2hUi + F (hUi) + F (hUi)h(Ue − hUei)i  00 0 0 + F (hUi)h(Ue − hUei)2i + O(2) hξ(x, t)ξ(x , t )i = [F0(U(x, t)) − F1(U(x, t))] 2 0 0 0 0 0 × [F0(U(x , t )) − F1(U(x , t ))]hξ0(t)ξ0(t )i 0 0 That is, we have the following leading order correction = 2B(U(x, t),U(x , t))δ(t − t ), (4.7) to Eq. (4.5): with ∂hUi 2  X 00 = D∇ hUi + F (hUi) + ∆(x, x, t) ρnF (hUi), αβ F (u(x)) − F (u(x)) ∂t 2 n B(u(x), u(y)) ≡ 0 1 n α + β α + β (4.13) F (u(y)) − F (u(y)) where ∆(x, y, t) is the equal-time two-point correlation × 0 1 . (4.8) α + β function A little algebra shows that ∆(x, y, t) = hUe(x, t)Ue(y, t)i. (4.14)

B(u(x), u(y)) (4.9) Since the first-order moment equation couples to the two-point correlation function, we need to obtain an α(F (u(x)) − F (u(x))(F (u(y)) − F (u(y)) = 0 0 equation for the latter. Again this can be obtained using (α + β)2 the linear noise approximation. More specifically, lin- β(F (u(x)) − F (u(x))(F (u(y)) − F (u(y)) earizing Eq. (4.6) about u gives + 1 1 . (α + β)2 ∂Ue 2 0 k = D∇ Ue + F (u)Ue + ξe(x, t), (4.15) In the simple example F0(u) = −γu and F1(u) = 0, we ∂t have with hξ(x, t)i = 0 and 2 e 2 k 2 αβγ B(u(x), u(y)) = σγ (u(x)u(y)) , σγ = . (α + β)3 hξe(x, t)ξe(x0, t)i = 2B(u(x, t), u(x0, t0))δ(t − t0). (4.16) (4.10) We can calculate moments of the stochastic density Ue(x, t) by expressing the solution in terms of the causal B. O() corrections to the mean field equation Green’s function or propagator G, which is defined ac- cording to Eq. (4.4) or Eq. (4.6) can now be used to derive O()  ∂ 0  corrections to the mean field Eq. (4.5). As we have − D∇2 − F (u(x, t)) G(x, t; x0, t0) (4.17) shown elsewhere for finite-dimensional stochastic hybrid ∂t systems [46], the former allows one to explore the effects = δ(x − x0)δ(t − t0). of noise outside the adiabatic regime. In particular, if the initial state of the environment is far from the steady- Assuming the initial condition Ue(x, 0) = 0, it follows that ∗ state distribution ρn, that is z(0) 6= z , then there are Z t Z additional contributions to the moments of U(x, t) that 0 0 0 0 0 0 are not taken into account by Eq. (4.6). However, such Ue(x, t) = dt dx G(x, t; x , t )ξe(x , t ). (4.18) 0 d contributions vary as e−(α+β)t/ so that for times t  R  they are negligible. Therefore, for simplicity, we will It follows from Eq. (4.16) that assume that  is sufficiently small, and use the Ito SPDE (4.6). Taking expectations of both sides of this equation Z t Z Z ∆(x, y, t) = 2 dt0 dx0 dx00 G(x, t; x0, t0) shows that d d 0 R R (4.19) ∂hUi = D∇2hUi + hF (U)i. (4.11) 00 0 0 0 00 0 ∂t × G(y, t; x , t )B(u(x , t ), u(x , t )). 9

Differentiating both sides using the equation for G gives the steady-state solution, but the approach to steady- state. Suppose that we ignore diffusion by taking a uni-   ∂ 2 2 0 0 form initial density and solve the resulting differential − D∇x − D∇y − F (u(x, t)) − F (u(y, t)) 2 ∂t equationu ˙ = −γu under the initial condition u(0) = u0. × ∆(x, y, t) = 2B(u(x, t), u(y, t)). (4.20) This gives 1 Finally noting that u − hUi = O(), we have u(t) = , (5.2) 1/u0 + γt

∂∆ 2 2 0 0 −1 = D∇x∆ + D∇y∆ + F (hU(x, t)i) + F (hU(y, t)i) which exhibits the asymptotic behavior u(t) ∼ t that ∂t is independent of the initial density. On the other hand, + 2B(hU(x, t)i, hU(x, t)i) + O(). (4.21) Monte-Carlo simulations of a spatially discretized version of the model, in which particles randomly hop between Higher-order corrections to the moment equations can neighboring sites on a lattice, reveal more complex be- be obtained in a similar fashion, and can be represented havior that is dependent on the spatial dimension d [53]: graphically using Feynman diagrams, see Sect. VI. Fi- nally, note that an alternative way of deriving the O() u ∼ t−1/2 (d = 1), u ∼ t−1 ln t (d = 2), u ∼ t−1 (d > 2). correction to the mean field equation is to perform a semi- (5.3) classical expansion of the corresponding moment generat- Hence the RD equation only captures the asymptotic dy- ing functional [61, 62], see also recent studies of a neural namics in dimensions higher than two. This is a conse- master equation [63, 64] and the Kuramoto model [65]. quence of the fact that (5.1) is a macroscopic equation for the mean particle density u(x, t), which ignores any spatial fluctuations and statistical correlations between V. RD MASTER EQUATION FOR particles due to low copy numbers. STOCHASTICALLY-GATED ADSORPTION One way to understand the breakdown of the macro- scopic RD equation at d = 2 is to recall that random As we highlighted in Sect. II, the derivation of the walks display certain universal properties. In particular, continuum path integral (4.1) is based on a spatial dis- an unbiased random walk on a d-dimensional lattice is cretization in which the number of molecules per lattice recurrent if d ≤ 2 and transient if d > 2. In light of this, site is large so we can define a local concentration. Al- consider a decay process such as A + A → ∅ where at though the theory applies to any reasonable choice of large times the surviving particles are separated by large reaction terms Fn(u), the particular example of stochas- distances. This means that the probability of a pair of tically gated adsorption introduced in Sect. II.A raises an diffusing particles coming into close proximity to annihi- additional issue. That is, eventually adsorption will lead late each other is strongly dependent on whether diffusion to a small number of well-separated molecules, for which is recurrent or transient. For d ≤ 2 (recurrent diffusion) a the existence of a local concentration breaks down. An pair of particles find each other with probability 1, even analogous issue holds in the classical problem of diffusion- if they are represented by points in a continuum limit. limited pair annihilation [54, 55], which combines diffu- Hence, the effective diffusion-limited reaction rate will sion and single-species pair annihilation be governed by universal features of diffusion. On the other hand, if d > 2 (transient diffusion) then the prob- γ A + A → ∅ + diffusion. ability of point particles meeting vanishes. That is, for any reaction to occur, the particles must have finite size The corresponding mean-field RD equation is or reaction radius, or be placed on a lattice. The effective reaction rate will then depend on the microscopic details ∂u = D∇2u − γu2, (5.1) of the short-distance spatial regularization, meaning that ∂t a degree of universality is lost. The occurrence of univer- sal behavior at and below some dimension d , known as 2 c with the reaction term −γu representing pair annihila- the upper critical dimension, is typically indicative of a tion under mass action kinetics. In our adsorption model, break down of macroscopic mean field equations due to this term represents instantaneous adsorption following the effects of statistical fluctuations. In the case of binary dimerization, which is equivalent to pair-annihilation in reactions such as pair annihilation, dc = 2. the diffusing medium. Let us now return to our example of stochastically- gated adsorption. In the case of a spatially uniform so- lution, Eqs. (4.13) and (4.20) become (to O()) A. Large-t behavior dhUi  00 = F (hUi) + ∆(t)F (hUi) (5.4a) It is straightforward to show that Eq. (5.1) has the dt 2 d∆ 0 trivial solution u(x, t) → 0 pointwise as t → ∞. The in- = 2F (hUi)∆(t) + 2B(hUi, hUi). (5.4b) terest in the above system is not, therefore, with regards dt 10

In particular, if F (u) = −2γu2, see also Sect. VI, then applies to one- and two-dimensional interfaces. There- B is given by Eq. (4.10) so that fore, the spatial discretization scheme developed in Sect. II breaks down when the number of molecules per lat- dhUi tice site becomes sufficiently small at very large times, = −2γhUi2 − 2γ∆(t) (5.5a) dt since fluctuations in the local concentrations blow up. d∆ In classical field theoretic treatments of pair annihila- = −8γhUi∆ + 8σ2hUi4. (5.5b) dt γ tion with diffusion [54–58], the effects of small molecular numbers at each lattice site is dealt with by consider- In order to extract the large time behavior, we use reg- ing an RD master equation, in which diffusive hopping ular perturbation theory with hUi = u0 + u1 + ... and between neighboring lattice sites is treated as an addi- ∆ = ∆0 +∆1 +.... Substituting into equations (5.5) and tional set of single step reactions that supplement the collecting equal powers of  yields a hierarchy of equa- pair-annihilation reaction. The Doi-Peliti formalism is tions, the first few of which are as follows: then used to construct a corresponding path integral, which yields a statistical field theoretic description of the du0 2 = −2γu0(t), (5.6a) RD process in the continuum limit. We show how coher- dt ent spin states can be used to extend the classical field du1 theoretic treatment of RD master equations in order to = −4γu0(t)u1(t) − 2γ∆0(t), (5.6b) dt take into account stochastically gated reaction rates. d∆ 0 = −8γu (t)∆ (t) + 8σ2u4(t). (5.6c) dt 0 0 γ 0 B. RD master equation with switching reaction The first equation has the solution (for large times) rates 1 1 u0(t) = ∼ . Returning to the spatial discretization scheme of Sect. 1/u0(0) + 2γt 2γt II, we now set uj = zj/Ω at the j-th lattice site, where d Plugging into the third equation implies that zj is the number of molecules and Ω = h is the hyper- volume of a single discrete cell. If zj is sufficiently large, 2 then we can treat zj as a continuous variable and write d∆0 −1 8σγ ∼ −4t ∆0(t) + , down the analog of Eq. (2.4): dt (2γt)4 dzj X which can be rewritten as = ∆ z − γ z (z − 1), i ∈ J (5.8) dt ij j n j j j∈J d(t4∆ ) σ2 σ2 0 ∼ γ =⇒ ∆ (t) ∼ γ . dt 2γ4 0 2γ(γt)3 when N(t) = n and γn is an n-dependent adsorption rate. We have rescaled the adsorption rate in Eq. (2.4) d Finally, substituting for u0, ∆0 into Eq. (5.6b) gives according to γn → γnh . If one views equation (5.8) as the rate equation of a Markov process with discrete 2 P du σ variables z(t) = j zj(t)ej, then the associated master 1 = −2t−1u (t) − γ , dt 1 (γt)3 equation for the probability distribution P (n, z, t) where n is the discrete state of the environment is which can be rewritten as dPn(z, t) 2 2 2 d(t u1) σγ σγ dt ∼ − 3 =⇒ u1(t) ∼ − 2 ln t. D X dt γ t γ(γt) = [(z + 1)P (z + e − e 0 , t) − z P (z, t)] h2 j n j j j n hj,j0i Combining the various results, we obtain the large-t be- D X havior + [(z 0 + 1)P (z − e + e 0 , t) − z 0 P (z, t)] h2 j n j j j n 0 1 σ2 σ2 hj,j i hU(t)i ∼ − γ ln t, ∆(t) ∼ γ . (5.7) X 2γt γ(γt)2 2γ(γt)3 + γn [(zj + 2)(zj + 1)Pn(z + 2ej, t) j Hence, environmental switching contributes to the sub- X − z (z − 1)P (z, t)] + Q P (z, t), (5.9) leading large-t behavior. j j n nm m m As in the classical pair-annihilation system, we expect 0 the above scaling behavior only to hold for d > dc, where where hj, j i indicates that we are summing over nearest dc is the upper critical dimension. Since dimerization re- neighbors on the lattice without double counting. Note quires two molecules to be in close proximity in the same that z + ej − ej0 = (. . . , zj + 1, zj0 − 1,...) etc. The way as pair annihilation, it follows that dc = 2. This is first and second lines represent single-step hopping tran- particularly significant for adsorption, since it typically sitions j → j0 and j0 → j, respectively, while the third 11

P Q P Q † zj line represents instantaneous adsorption following dimer- where = d and |zi = (a ) |0i. In z j∈Z zj ≥0 j j ization, A + A → ∅ at each lattice site, and changes in addition, environmental state. The next step is to convert the RD master equation to operator form by introducing Doi-Peliti annihilation † and creation operators aj, aj at each lattice site. The † aj|zi = zj| . . . , zj−1, zj − 1, zj+1,...i, corresponding commutation relations for ai, ai are † aj|zi = | . . . , zj−1, zj + 1, zj+1,...i † † † [ai, aj] = δi,j, [ai, aj] = [ai , aj]. (5.10) Define the state vectors X |ψn(t)i = Pn(z, t)|zi, (5.11) Differentiating with respect to time and plugging in the z master equation gives

d D X X Y † zj |ψ (t)i = [(z + 1)P (z + e − e 0 , t) − z P (z, t)] (a ) |0i dt n h2 j n j j j n j z hj,j0i j D X X Y † zj + [(z 0 + 1)P (z − e + e 0 , t) − z P (z, t)] (a ) |0i h2 j n j j j n j z hj,j0i j

X X Y † zj X + γn [(zj + 2)(zj + 1)Pn(z + 2ej, t) − zj(zj − 1)Pn(z, t)] (aj) |0i + Qnm|ψm(t)i. z j j m=0,1

Following the same steps as in the analysis of pair- In order to determine the action of Hbn on |ϕi it is first annihilation [54, 55], we thus obtain the following op- necessary to normal-order Hbn by moving all creation op- erator version of the master equation: † erators aj to the left of all annihilation operators aj using the commutation relations. The normal-ordered operator d X |ψn(t)i = Hbn|ψn(t)i + Qnm|ψm(t)i, (5.12) is dt m=0,1 D X † † X † 2 2 Hbn = − (a 0 − a )(aj0 − aj) + γn [1 − (a ) ]a . h2 j j j j where hj,j0i j (5.16) D X † † Hbn = − (a 0 − a )(aj0 − aj) (5.13) h2 j j such that hj,j0i  X h 2 † † i D X ∗ ∗ + γn aj − a aj(a aj − 1) . Hbn|ϕi = − (ϕ 0 − ϕ )(ϕj0 − ϕj) (5.17) j j h2 j j j hj,j0i  X ∗ 2 2 An important step in the construction of a Doi-Peliti + γn [1 − (ϕj ) ]ϕj |ϕi. path integral is the choice of basis vectors. Following j along similar lines to the case of pair-annihilation, we introduce the coherent-state representation Following the same procedure as Sect. III, we intro- duce the diagonal matrix   ! 1       X 2 X † Hb0 0 1 1 |ϕi = exp − |ϕi| exp ϕia |0i, (5.14) K = = 1 + σ H + 1 − σ H .  2  i b z b0 z b1 j i 0 Hb1 2 2 (5.18) where ϕi is the complex-valued eigenvalue of the anni- and obtain the operator equation ∗ hilation operator ai, with complex conjugate ϕi . The d coherent states satisfy the completeness relation |ψ(t)i = Hb |ψ(t)i, (5.19) dt Z ∗ Y dϕidϕ > i |ϕihϕ| = 1. (5.15) with |ψ(t)i = (|ψ0(t)i, |ψ1(t)i) , Hb given by Eq. (3.24) π i and hs|Hb |si satisfying Eq. (3.25). The only difference 12 is the definition of the operators Hbn. Formally speak- C. Doi-Peliti path integral ing, the solution to the operator version of the master equation can be written as Following Sect. III, we divide the time interval [0, t] into N subintervals of size ∆t = t/N and write |ψ(t)i = eHb t|ψ(0)i, (5.20) H∆t H∆t H∆t P a |ψ(t)i = e b e b ··· e b e j j |ψ(0)i, (5.26) and the expectation value of the number of molecules We then insert multiple copies of the completeness re- zj(t) expressed as lations (3.8) and (5.15) for the basis vectors |s, ϕi =   |si ⊗ |ϕi so that   Z Z Z ∗ Z ∗ Y † Hb t dϕ0dϕ0 dϕN dϕN hzj(t)i = (1, 1) ⊗ h0| exp (aj) ajaje |ψ(0)i, |ψ(t)i = ds0 ··· dsN ···  j  Ω Ω π π (5.21) N−1 Y Hb ∆t where × |sN , ϕN i hs`+1, ϕ`+1|e |s`, ϕ`i `=0 Y P aj h∅| = h0| exp (aj) (5.22) × hs0, ϕ0|e j |ψ(0)i. (5.27) j In the limit N → ∞ and ∆t → 0 with N∆t = t fixed, we is the projection state and can make the approximation Hb ∆t X hs`+1, ϕ`+1|e |s`, ϕ`i (1, 1) ⊗ h∅|ψ(t)i = h∅|ψn(t)i.   n=0,1 ∗ ≈ hs`+1|s`ihϕ`+1|ϕ`i 1 + H(θ`, φ`, ϕ`, ϕ` )∆t ,

P a It is also convenient to shift the term e j j to the right (5.28) by using the identity where ∗ aj † † aj H(θ, φ, ϕ, ϕ ) ≡ hs|Hb |si (5.29) e f(aj) = f(aj + 1)e , D X ∗ ∗ = − (ϕ 0 − ϕ )(ϕj0 − ϕj) for an arbitrary function f (the so-called Doi shift). In h2 j j hj,j0i particular, since the annihilation and creation operators 1 + cos θ commute with the matrix Q, we have − β 1 − eiφ + γ Γ(ϕ, ϕ∗) 0 2 P 1 − cos θ Hb t j aj  −iφ ∗  hzj(t)i = (1, 1) ⊗ h0|aje e |ψ(0)i, (5.23) − α 1 − e + γ Γ(ϕ, ϕ ) . (5.30) 1 2 with the modified operators and ∗ X ∗ 2 ∗ 2 Γ(ϕ, ϕ ) = [(ϕj ) + 2ϕj ]ϕj (5.31) D X † † Hbn = − (a 0 − a )(aj0 − aj) j h2 j j hj,j0i The inner product hs`+1|s`i satisfies Eq. (3.31). In addi- X † 2 † 2 − γn [(aj) + 2aj]aj . (5.24) tion, using standard properties of coherent states, j   1 X 1 X hϕ |ϕ i = exp |ϕ |2 − |ϕ |2 We also take the initial state to be distributed according `+1 ` 2 `+1,i 2 `,i  to a product of Poisson distributions, j j ! z X ∗ j −r0 × exp − ϕ (ϕ`+1,i − ϕ`,i) (5.32) Y r0 e `+1,i Pn(z, 0) = ρn , i zj! j Substituting the expression for the small-time propa- gator back into Eq. (5.27) yields where hzj(0)i = r0. Then, for each j, Z Z Z ∗ Z ∗ dϕ0dϕ0 dϕN dϕN |ψ(t)i = ds0 ··· dsN ··· zj † zj −r0 −r0 † π π X r0 e X [r0aj] e −r r a Ω Ω |z i = |0i = e 0 e 0 j |0i, z ! j z ! N−1  j j Y i dφ` zj zj × |s , ϕ i exp H(θ , φ , ϕ , ϕ∗) − cos θ N N ` ` ` ` 2 dt ` `=0 which is a coherent state. It follows that   dϕ` P ∗ j aj P P † + iϕ` · ∆t hs0, ϕ0|e |ψ(0)i. j aj r0 j aj dt e |ψn(0)i = ρne |0i. (5.25) 13

We now take the limit N → ∞, ∆t → 0 with N∆t = t [34], we obtain the following Ito SPDE with two indepen- ∗ ∗ fixed, ϕ` = ϕ(`∆t) and ϕ` = ϕ (`∆t). After Wick dent multiplicative noise sources: ordering, ϕ∗ → −iϕ∗, integrating by parts the term in- √ volving dφ/dt, and performing the change of coordinates ∂Φ 2 2 = D∇ Φ − 2[γ0z + γ1(1 − z)]Φ + ζ, (5.38a) z = (1+cos θ)/2, we obtain the following functional path ∂t integral: dz √  = −βz + α(1 − z) + ξ , (5.38b) dt z Z ∗ hzj(t)i ∼ N D[z]D[φ]D[ϕ]D[ϕ ]ϕj(t) where ζ and ξzare independent Gaussian white noise pro- cesses with hξ i = 0 = hζi and  Z t    z ∗ dϕ dz exp − ϕ · − iφ − H dτ , (5.33) 0 0 0 dτ dτ hξz(t)ξz(t )i = (βz + α(1 − z))δ(t − t ) (5.38c) hζ(x, t)ζ(x0, t)i (5.38d) where Nn is a constant and H is the effective Hamiltonian 2 0 0 = −2[γ0z + γ1(1 − z)]Φ(x, t) δ(x − x )δ(t − t ). D X ∗ ∗ H = − (ϕ 0 − ϕ )(ϕ 0 − ϕ ) ∗ h2 j j j j Taking the limit  → 0 with z(t) → z = α/(α+β) yields hj,j0i the deterministic mean-field RD equation  iφ ∗  − β 1 − e + γ0Γ(ϕ, ϕ ) z ∂U 2 2 αγ0 βγ1  −iφ ∗  = D∇ U − 2γU , γ = + . (5.39) − α 1 − e + γ1Γ(ϕ, ϕ ) (1 − z). (5.34) ∂t α + β α + β

Following Sect. IV, we now take the continuum limit For small but non-zero  we can use the linear-noise ap- P d proximation of Sect. IV.B to obtain an SPDE for Φ of by letting the lattice spacing h → 0 such that j h → R the form dx. This requires using the unscaled rate γn and re- defining the fields as follows: ∂Φ √ √ = D∇2Φ − 2γΦ2 + ξ(x, t) + ζ(x, t), (5.40) −d ∗ ∂t h ϕj(t) → ϕ(x, t), ϕj (t) → ϕe(x, t). (5.35) with hξ(x, t)i = 0 = hζ(x, t)i and We thus arrive at the field theoretic path integral 0 2 2 0 0 2 0 hξ(x, t)ξ(x , t)i = 8σγ Φ(x, t) Φ(x , t ) δ(t − t ), (5.41a) Z 0 2 0 0 −S[z,φ,ϕ,ϕe] hζ(x, t)ζ(x , t)i = −2¯γΦ(x, t) δ(x − x )δ(t − t ). hz(x, t)i ∼ D[z]D[φ]D[ϕ] D[ϕe] φ(x, t)e (5.41b) (5.36) Eq. (5.41b) represents the effects of demographic noise with the action functional based on the Doi-Peliti construction. However, the neg- ative sign in the noise correlations means that ζ is Z t  dz S = − iφ + β 1 − eiφ z + α 1 − e−iφ (1 − z) complex-valued, which limits the phenomenological in- 0 dτ terpretation of the Langevin equation. On the other- Z     ∂ϕ 2 hand, the real-valued noise satisfying (5.41a) arises from + dx ϕ − D∇ ϕ − ρ0φδe (τ) (5.37) d e ∂τ environmental switching. In the limit ζ → ∞, Eq. (5.40) R Z  reduces to the SPDE (4.6), whereas in the limit ξ → 0 2 2 − [zγ0 + (1 − z)γ1] dx [ϕ + 2ϕ]ϕ dτ. we obtain the equivalent Langevin equation for pair an- d e e R nihilation [54, 55].

D. Semi-classical limit VI. DIAGRAMMATIC EXPANSION AND RENORMALIZATION In contrast to the path-integral (4.1), there are now two distinct sources of fluctuations, one given by the A. Feynman diagrams switching environment and the other from demographic noise, which is represented by the quartic terms φe2φ2 Following along the lines of [55], we develop a pertur- in the action functional. As it stands the path integral bative solution of the SPDE (5.40) by iterating the formal is analytically intractable. Therefore, we derive a weak solution noise approximation along analogous lines to Sect. IV.B. Z t Z 0 0 0 0 0 That is, we set φe = φe and consider the approximation Φ(x, t) = dt dx G0(x − x , t − t ) (6.1) ±iφ 2 2 d 1−e ≈ ∓iφ+ φ /2, so that the action is quadratic 0 R in φ and ϕ. Again comparing with the stochastic path  √ √  e × − 2γΦ2(x0, t0) + ξ(x0, t0) + ζ(x0, t0) + ρ δ(t0) . integral for an SPDE with Gaussian spatiotemporal noise 0 14 where G0 is the free propagator Now suppose that we switch on the noise terms by taking  ∂   > 0. Following Sect. V.D, let us shift the stochastic − D∇2 G (x, t) = δ(x)δ(t). (6.2) field according to ∂t 0 and the final term arises from the Doi-shifted initial co- Φ(x, t) = φ(x, t) + Φ(e x, t). (6.6) herent state. Eq. (6.2) has the solution Substituting into Eq. (5.40) and taking expectations of 1 0 2 0 G (x, t; x0, t0) = e−|x−x | /4D(t−t )H(t − t0), both sides gives 0 (4πDt)d/2 (6.3) ∂hΦi = D∇2hΦi − 2γh(φ + Φ)e 2i where H(t) is the left continuous Heaviside step function: ∂t H(0) = 0 for t ≤ 0 and H(1) = 1 for t > 0. (The choice = D∇2hΦi − 2γhΦi2 − 2γh(Φ − hΦi)2i H(0) = 0 is consistent with the Ito condition for the e e SDE, and ensures that Φ(x, t) is uncorrelated with the = D∇2hΦi − 2γhΦi2 − 2γ∆(x, x, t), noise terms. Other forms of stochastic calculus would be obtained using different choices of H(0). For example, where ∆(x, y, t) is the equal-time two-point correlation H(0) = 1/2 would correspond to Stratonovich calculus.) function The iterative solution for Φ can be represented graph- ically in terms of Feynman diagrams. First consider the ∆(x, y, t) = hΦ(e x, t)Φ(e y, t)i. (6.7) noise-free case ( → 0). The formal solution can be ex- The first moment equation thus has the formal solution pressed compactly in terms of convolutions: 2 Z t Z φ = ρ0G0 ◦ δ − 2γG0 ◦ φ 0 0 0 0 hΦ(x, t)i = dt dx G0(x − x , t − t ) (6.8) 2 2 d = ρ G ◦ δ − 2ρ γG ◦ (G ◦ δ) + ... (6.4) 0 R 0 0 0 0 0   0 0 2 0 0 0 0 The diagrammatic representation of this iterated solution × − 2γhΦ(x , t )i − 2γ∆(x , x , t ) + ρ0δ(t ) . is shown in Fig.2, which establishes the well known result that the sum of tree level diagrams generates the Iterating Eq. (6.8) generates both tree-level diagrams solution to the deterministic RD equation (5.1). Using as in the deterministic case, and loop diagrams associated the fact that with the correlation function ∆. Here we focus on the 1- Z loop diagrams, which can be obtained by linearizing Eq. (G0 ◦ δ)(x, t) = G0(x, t)dx = H(t), d R (5.40): it follows that ∂Φe √ √ 2 2 4 2 = D∇2Φe − 4γφΦe + ξe(x, t) + ζe(x, t), (6.9) φ(x, t) = ρ0 − 2γρ0t + 4γ ρ0t − ... ∂t ρ = 0 , (6.5) 1 + 2ρ0γt with hξe(x, t)i = 0, hζe(x, t)i = 0 and

0 2 2 0 0 2 0 ρ0 hξe(x, t)ξe(x , t)i = 8σγ φ(x, t) φ(x , t ) δ(t − t ), (6.10a) hζe(x, t)ζe(x0, t)i = −2¯γφ(x, t)2δ(x − x0)δ(t − t0). φ( x,t) _ = ρ0 + −2γ (6.10b) Introduce the propagator G (cf. Eq. (4.17)), ρ0 ρ0   ∂ 2 0 0 _ ρ0 − D∇ − 4γφ(x, t) G(x, t; x , t ) (6.11) −2γ ∂t 0 0 _ _ = δ(x − x )δ(t − t ). + + −2γ +... −2γ ρ0 ρ0 _ Note that G is related to the “bare” propagator G0 by −2γ dressing it with all tree-level diagrams. Hence, in the ρ0 weak coupling, G = G0 + O(γ). Assuming the initial ρ0 condition Φ(e x, 0) = 0, it follows that

FIG. 2. Tree level diagrams contributing to the determinis- Φ(e x, t) (6.12) tic solution, which is indicated by the thick line. Filled dots √ Z t Z   0 0 0 0 0 0 0 0 0 0 correspond to factors of ρ0 and thin lines from t to t, t < t =  dt dx G(x, t; x , t ) ξe(x , t ) + ζe(x , t ) . 0 0 d represent the free propagator G0(t − t ). Each 3-vertex intro- R duces a factor −2γ and a time integral. and thus 15

Z t Z Z 0 0 0 0 0 0 0 2 2 0 0 2 0 0 2 0 0 0 0  ∆(x, y, t) =  dt dx dy G(x, t; x , t )G(y, t; y , t ) 8σγ φ (x , t )φ (y , t ) − 2γφ (x , t )δ(y − x ) . d d 0 R R (6.13)

In terms of a weak coupling expansion, the lowest order contribution to the two-point correlation function is

Z t Z Z 0 0 0 0 0 0 0 2 4 2 0 0  ∆(x, y, t) =  dt dx dy G0(x − x , t − t )G0(y − y , t − t ) 8σγ ρ0 − 2ρ0γδ(x − y ) d d 0 R R  Z t Z  2 4 2 0 0 0 0 0 0 =  µ ρ0t − λρ0 dt dx G0(x − x , t − t )G0(y − x , t − t ) . (6.14) d 0 R

(a) Reversing the order of integration and using ρ0 −λ -λε Z 0 dx e−ix·(k−k ) = (2π)dδ(k − k0), x,t x’,t’ x’‘ ,t’ ρ0

we have

(b) x’‘ ,t’’ ρ0 0 Z dk 02 _ −2Dk t −d/2 Γ(t) = d e = 2(8πDt) (6.16) −2γ μ√ε ρ0 (2π) ρ x,t x’,t’ μ√ε 0 It follows that the 1-loop contribution is given by ρ y’‘ ,t’’ 0 Z t Z τ1 2 2 Iloop = λ ρ0 dτ1 Γ(τ1 − τ2) FIG. 3. One-loop diagrams contributing to√ hΦi(x, t). The 0 0 coupling constants are λ = 2γ and µ = 2 2σ (a) Demo- λ2ρ2 Z t Z τ1 dτ γ = 0 dτ 2 graphic noise ζ. (b) Environmental noise ξ. d/2 1 d/2 (8πD) 0 0 (τ1 − τ2) 4λ2ρ2 t2−d/2 = 0 , d 6= 2, 4. (8πD)d/2 (2 − d)(4 − d) We have introduced two coupling constants 2 √ Comparison with the result Itree = −2λρ0t for the corre- λ = 2γ, µ = 2 2σγ . (6.15) sponding tree diagram (second graph in Fig.2), shows The two terms on the right-hand side of Eq. (6.14) correspond to the 1-loop diagrams for hΦi shown in Fig. 3. The linear term in t reflects the short-time solution of Eq. (5.6c) with ∆0(0) = 0 and u0(0) = ρ0. On the other hand, it is well known from studies of pair annihilation that the 1-loop diagram associated with demographic -λ -λε noise has ultraviolet (UV) divergences for d ≥ dc = 2 and infrared (IV) divergences for d ≤ 2 [54, 55] – these do not arise in the case of environmental noise (at least at the given level of approximation) due to the non-local 3-vertex 4-vertex form of the correlation function (6.10a). A direct method for identifying the singular behavior is to set x = y and consider the Fourier transform μ√ε Z dk Z Γ(t) = dxdx0eik·xG (x − x0, t)2 μ√ε (2π)d 0  0 Z dk Z Z dk 0 0 02 = dxdx0eik·x e−ik ·(x−x )e−Dk t (2π)d (2π)d 6-vertex 00  Z dk 00 0 002 × e−ik ·(x−x )e−Dk t . (2π)d FIG. 4. Vertices used to construct Feynman diagrams. 16

-λε -λε -λε μ√ε -λ -λ μ√ε

- λε μ√ε μ√ε -λε -λ μ√ε -λ μ√ε

μ√ε μ√ε -λ μ√ε -λ μ√ε

FIG. 5. One-particle irreducible 2-loop (or O(2)) diagrams contributing to hΦi. that the loop diagram has an effective dimensionless cou- Laplace-Fourier transforms. Summing the resulting ge- pling constant ometric series yields an effective 3-vertex coupling Γe3(s) given by [54, 55] 2λ λ ∼ t1−d/2. (6.17) eff Dd/2 λ Γe3(s) = , (6.18) 1 + λIe(s) This indicates that for d < dc = 2, the perturbation expansion is finite at small times but blows up at large where s is the Laplace frequency, times, whereas the opposite holds when d > dc. In the critical two-dimensional case one finds that the effective 2 Ie(s) = Γ(1 − d/2)s−(1−d/2), (6.19) coupling diverges as (λ/D) ln(Dt) for both t → 0 and (8πD)d/2 t → ∞. Higher-order contributions can be delineated diagram- and Γ(z) is the Gamma function matically by combining propagators G (or the dressed 0 Z ∞ version G) with the vertices shown in Fig.4. The 6- Γ(z) = xz−1e−xdx. (6.20) vertex corresponds to the right-hand vertex in Fig.3(b). 0 Examples of one-particle irreducible (1PI) 2-loop dia- grams are given in Fig.5. Here 1PI means that a diagram The latter has poles at z = 0, −1, −2, ... with Γ(z) = cannot be disconnected by cutting a single internal line. 1/z + γ + ... for z → 0 and Euler’s constant γ. Hence, the IR and UV divergences are neatly separated. The

B. Renormalization -λε = + The divergent contributions are the same as standard pair annihilation [54, 55], namely the set of loop dia- grams shown in Fig.6. A powerful method for han- -λε -λε dling these divergences is to use renormalization theory + [54–56, 58, 61]. This combines dimensional analysis with scaling theory to extend perturbation results that hold -λε -λε -λε at intermediate time and length scales to regimes where + + . . . perturbation theory breaks down. In the case of the cur- rent model the only model parameter that needs to be renormalized is the coupling parameter γ. The first step is to note that the divergent loop dia- FIG. 6. Loop diagrams contributing to singularities in the grams involve convolutions that can be analyzed using effective 3-vertex coupling. 17 former occurs in the limit s → 0 (large time limit) when with β given by Eq, (6.23). Above the critical dimension d < dc, whereas the latter correspond to poles in the (d > 2), the renormalized coupling constant flows to zero −(d/2−1) Gamma function. We now fix s such that t0 = 1/s rep- algebraically as t , which means that the effective resents a large but intermediate time where perturbation reaction rate λ(t) → constant, as assumed in mean field theory still holds, and define a renormalized dimension- theory. On the other hand, below the critical dimension, less coupling parameter ∗ 1 − d/2 −(1−d/2) −d/2 gR(t) → gR = as t → ∞, gR(s) = Γe3(s)s D . (6.21) e B Now suppose that we rewrite the perturbation expansion and of hΦi in terms of the renormalized parameter gR, the d/2−1 ∗ λ(t) ∼ D(Dt) gR for t  t0. (6.30) initial density ρ0, the bare coupling parameter µ and the reference frequency s. Then Finally, at the critical dimension,

hΦi = Φ(t, ρ0, µ, gR(s), s). dgeR 2 t = −BgR, (6.31) Since s does not appear in the unrenormalized the- dt e ory, any statistical quantity should be independent of s. which has the solution Therefore, we require 4π dΦ  ∂ ∂  geR(t) = . (6.32) s = s + β(g ) φ = 0. (6.22) ln t ds ∂s R ∂g R Hence, were we have introduced the renormalization group β 4πD function λ(t) ∼ , t  t0. (6.33) ln(t/t0) dgR β(gR) = s = gR [−(1 − d/2) + BgR] , (6.23) ds Since the running coupling µe(t) → ∞ for d = 1, whereas µe(t) = µ for all t at the critical dimension d = 2, where perturbation theory can only be used in the latter case. 2 − d The large-t behavior for the mean hΦi is obtained by B = Γ(1 − d/2). (6.24) (8π)d/2 making the replacement λ → λ(t) = D/ ln t in Eqs. (5.5): (Here  represent the noise strength rather than the pa- dhUi rameter dc −d commonly used in dimensional regulariza- = −λ(t)hUi2 − λ(t)∆(t) (6.34a) tion.) From dimensional analysis, dt d∆ = −4λ(t)hUi∆ + µ2hUi4. (6.34b) Φ(t, ρ0, µ, gR, s) (6.25) dt −d/2 d/2 d/2 −1/2 ∼ t F (st, ρ0(D/s) , µ(s/D) s , gR), In order to extract the large time behavior, we again use which eliminates one of the independent variables and regular perturbation theory by setting hUi = u0+u1+... yields the Callan-Symanzik equation for the mean den- and ∆ = ∆0 + ∆1 + .... Substituting into equations sity: (6.34) and collecting equal powers of  yields

  du0 ∂ dρ0 ∂ d − 1 ∂ ∂ d = −λ(t)u2(t), (6.35a) t − + µ + βg(gR) + Φ = 0. dt 0 ∂t 2 ∂ρ0 2 ∂µ ∂gR 2 du1 (6.26) = −2λ(t)u0(t)u1(t) − λ(t)∆0(t), (6.35b) Equation (6.26) can be solved using the method of dt characteristics. This allows us to link the asymptotic d∆0 2 4 −1 = −4λ(t)u0(t)∆0(t) + µ u0(t). (6.35c) regime t  t0 = s with the intermediate regime at t0: dt −d/2 The large-time asymptotics for u (t) for d = 2 is the same Φ(t, ρ0, µ, gR, s) = (t/t0) Φ(t0, ρe0(t), µe(t), geR(t)), 0 (6.27) as that found for standard pair-annihilation, see also Eq. (5.3): where ln(Dt) d/2 1−d/2 u0(t) ∼ . (6.36) ρe(t) = (t/t0) ρ0, µe(t) = (t/t0) µ, (6.28) 8πDt and the running coupling geR(t) is the solution to the Eq. (6.35c) becomes equation d∆ 2 ln t4 µ2 dg 0 ∼ − ∆ + b , b = . t eR = −β(g ). (6.29) dt t 0 t (8πD)4 dt eR 18

This can be rewritten as randomly switching environment. A small noise expan- sion now yielded a stochastic RD equation with both real d(t2∆ ) 0 ∼ bt−2(ln t)4, and complex noise sources. Finally, diagrammatic and dt renormalization group methods were used to determine the anomalous large-time behavior of the mean concen- Using the formula tration at the upper critical dimension dc = 2. Z (ln t)n (ln t)n n Z (ln t)n−1 The statistical field theory of stochastic hybrid RD pro- dt = − + dt cesses developed in this paper has a variety of possible tm (m − 1)tm−1 m − 1 tm extensions. Here we describe two of them. (6.37) for m 6= 1, we have

b(ln t)4 A. N > 2 discrete states ∆ (t) ∼ . 0 t3 First, for pedagogical reasons we restricted the analysis Finally, substituting for u0, ∆0 into Eq. (5.6b) gives to the case of two environmental states. As we have shown elsewhere [46], it is possible to apply coherent S- du 1 4πDb(ln t)3 spin states to represent the master equation (2.1) with 1 = − u − , dt t 1 t3 |I| = (2S + 1). Here we briefly sketch the extension to a 3-state model with N(t) ∈ {0, 1, 2} and a matrix which can be rewritten as generator of the form 3 d(tu1) 4πDb(ln t)   −β+ α+ 0 ∼ − 2 . dt t Q =  β+ −α+ − α− β−  . (7.1) 0 α −β We thus find that − − 4πDb(ln t)3 In terms of an adsorption model, this could represent u ∼ − , two different adsorbing states (n = 0, 2) with different 1 t2 adsorption rates, and a non-adsorbing state (n = 1). and hence Moreover, transitions can only occur either into or out of the non-adsorbing state. We develop the correspond- ln(t) 4πDb(ln t)3 b(ln t)4 hU(t)i ∼ − , ∆(t) ∼ . ing coherent spin-1 state construction along the lines of 8πDt t2 t3 [51]. First, rewrite the generator (7.1) in the form (6.38) Q = α+T+ + α−T− + β+S+ + β−S−, (7.2)

VII. DISCUSSION with  0 1 0   0 0 0  In this paper we used operator methods and coher- T+ =  0 −1 0  , T− =  0 −1 0  , (7.3) ent spin states to derive a path integral formulation of 0 0 0 0 1 0 a stochastic RD equation, in which the reaction term depends on the discrete state of a randomly switch- and ing environment. For simplicity, we assumed that the  −1 0 0   0 0 0  environment switched between two states according to S+ =  1 0 0  , S− =  0 0 1  , (7.4) a continuous-time Markov chain. We showed how in 0 0 0 0 0 −1 the semi-classical limit, the stochastic dynamics could be approximated by a system of Langevin equations. Next we introduce the coherent spin-1 state The latter consisted of an RD equation coupled to a  iφ 4  stochastic probability function z that evolved accord- e cos θ/2 2 ing to an Ornstein-Uhlenbeck process with multiplica- |si =  2 cos2 θ/2 sin θ/2  , 0 ≤ θ ≤ π, 0 ≤ φ < 2π, tive noise. On eliminating the variable z, we obtained e−iφ sin4 θ/2 an RD equation driven by nonlocal spatiotemporal noise. (7.5) This was then used to determine leading order corrections together with the adjoint to the mean-field RD equation. We applied the theory −iφ iφ to the example of diffusion with stochastically gated ad- hs| = e , 1, e . (7.6) sorption under dimerization. Since the adsorption model Note that exhibits anomalous large-time behavior in one and two 0 spatial dimensions, analogous to classical diffusion with hs0|si = ei(φ−φ ) cos4 θ + 2 cos2 θ/2 sin2 θ/2 pair-annihilation, we constructed an alternative path in- −i(φ−φ0) 4 tegral formulation based on an RD master equation in a + e sin θ/2, (7.7) 19 so that hs|si = 1 and Hence,

hs|Q|si = Q(θ, φ) (7.11) 4 4 2 hs + ∆s|si = 1 − i∆φ(cos θ/2 − sin θ/2) + O(∆φ ) iφ 2 2 4  ≡ − 1 − e 2α− cos θ/2 sin θ/2 + β+ cos θ/2 ≈ 1 − i∆φ cos θ. (7.8) −iφ 2 2 4  − 1 − e 2α+ cos θ/2 sin θ/2 + β− sin θ/2 .

Finally, the operator equation (3.23) becomes We also have the completeness relation d |ψ(t)i = Hb |ψ(t)i, (7.12) dt 3 Z π Z 2π sin θ dθ dφ |sihs| = 1. (7.9) > with |ψ(t)i = (|ψ0(t)i, |ψ1(t)i, |ψ2(t)i) , 4π 0 0 4 2 2 hs|Hb |si = Q(θ, φ) + Hb0 cos θ/2 + 2Hb1 cos θ/2 sin θ/2 4 It can checked that the following identities hold: + Hb2 sin θ/2, (7.13) and hs|T |si = 2(e−iφ − 1) cos2 θ/2 sin2 θ/2, (7.10a) ! + X X iφ 2 2 H = −i vˆ ∆ u + F (u ) (7.14) hs|T−|si = 2(e − 1) cos θ/2 sin θ/2, (7.10b) bn j jk k n j iφ 4 j k∈J hs|S+|si = (e − 1) cos θ/2, (7.10c) −iφ 4 hs|S−|si = (e − 1) sin θ/2. (7.10d) for n = 0, 1, 2.

The derivations of the path integrals (4.1) and (5.36) proceed along similar lines to the 2-state model. The action (4.2) becomes

Z t  dz  iφ  −iφ 2  iφ  −iφ S = − iφ + β+ 1 − e z + β− 1 − e (1 − z) + 2 α− 1 − e + α+ 1 − e z(1 − z) 0 dτ Z   Z Z Z  ∂u 2 2 2 2 + dx v − vD∇ u − z dx vF0(u) − 2z(1 − z) dx vF1(u) − (1 − z) dx vF2(u) dτ, d ∂t d d d R R R R 2 Similarly, the action (5.37) for the RD master equation with Fn(u) = −γnu , n = 0, 1, 2, becomes Z t  dz  iφ  −iφ 2  iφ  −iφ S = − iφ + β+ 1 − e z + β− 1 − e (1 − z) + 2 α− 1 − e + α+ 1 − e z(1 − z) 0 dτ Z     Z  ∂ϕ 2 2 2 2 2 + dx ϕ − D∇ ϕ − ρ0φδe (τ) − [z γ0 + 2z(1 − z)γ1 + (1 − z) γ2] dx [ϕ + 2ϕ]ϕ dτ. d e ∂t d e e R R

Note that the effective probability distribution ρn(t) across the three discrete states has the representation

2 2 ρ0 = z , ρ1 = 2z(1 − z), ρ2 = (1 − z) , P with m=0,1,2 ρm = 1.

B. Least action paths more rigorous results obtained using large deviation the- ory [42–45]. Recently we have obtained the same least action principle from the path integral based on coherent One of the original applications of path integrals in spin states. Here we sketch how such a derivation can be the case of finite-dimensional stochastic hybrid systems extended to the RD path integral (4.1). was calculating least-action paths in noise-induced escape As a first step, consider the scalings α → α/, β → β/ problems [32, 33]. In the weak-noise limit we showed and v → v/, and rewrite the path integral (4.1) as how the Hamiltonian of the path integral corresponded to the principal eigenvalue of a linear operator; the latter Z −S[u,v,z,φ]/ incorporated both the generator of the discrete Markov Pn[u] = Nn D[z]D[φ]D[u]D[v]e , process and the vector fields of the piecewise determin- u(x,t)=u(x) istic dynamics. The formal analysis was consistent with (7.15) 20 with action It then follows from equation (7.21) that Z t  dz  iφ  −iφ Z Z S = − iφ + β 1 − e z + α 1 − e (1 − z) q = dx vF0(u) − dx vF1(u). (7.26) 0 dτ d d Z  R R + dx h(u, v; z) dτ. (7.16) Finally, defining d R Z Here Λ = λ + dx vF1(u) d   Z R ∂u 2 h(u, v, z) = v − D∇ u − z dx vF0(u) ∂τ d and substituting for q and λ in equations (6.14) we obtain R Z the following path-integral for small : − (1 − z) dx vF1(u). d Z R P [u] ∼ D[v]D[u]e−S[u,v]/, (7.27a) (In contrast to equation (4.3), we have not rescaled φ.) Next, define the functions u(x,t)=u(x)  iφ  −iφ q = β+ 1 − e + λ, λ = −α 1 − e , (7.17) where S is the action functional and rewrite the action as Z t  Z  ∂  S[u, v] = dτ dx v(x, τ) u(x, τ) − D∇2u(x, τ) t d ∂τ Z  Z  0 R S = − iφz˙ + q(φ)z − λ(φ) + dxh(u, v; z) dτ  d 0 R − Λ[u, v] , (7.27b) (7.18) Note that λ and q are related according to and Λ is the principal eigenvalue of the functional equa- tion αβ q − β + − λ = 0. (7.19) Z  λ + α X dx v(x)Fn(u(x))δm,n + Qnm Rm[u, v] d which is the characteristic equation for the eigenvalue m≥0 R equation = Λ[u, v]Rn[u, v], (7.28) (q − β)r + αr = λr , βr − αr = λr . (7.20) 0 1 0 0 1 1 with Q given by the matrix (2.15). In the limit  → 0 the path integral will be dominated It is important to note that although the path integral by paths that satisfy (7.27) was derived for small , the resulting approxima- tion goes beyond the semi-classical limit. Indeed, the dz dq dλ  = z − . (7.21) Langevin Eq. (4.6) corresponds to expanding the func- dt dφ dφ tional Λ[u, v] to second order in v. A general feature of This allows us to set dz/dt = 0 in the action (7.18) and stochastic path integrals is that in the weak noise limit eliminate the independent variable φ by requiring they are dominated by least action paths. That is, we can interpret S[u, v] as the action functional of a Hamiltonian dq dλ α(1 − z) system with conjugate fields u, v and Hamiltonian z − = 0 =⇒ e2iφ = . (7.22) dφ dφ βz Z H[u, v] = D v(x)∇2u(x) + Λ[u, v], (7.29) We thus obtain the reduced path integral d R Z −S/b Pn[u] = Nn D[z]D[u]D[v]e , (7.23) with principal eigenvalue Λ[u, v]. Least action paths then u(x,t)=u(x) correspond to solutions of Hamilton’s equations with ∂u(x, t) δΛ[u, v] Z t  Z  = D∇2u(x, t) + , (7.30) Sb ≈ qz − λ + dx h(u, v; z) dτ. (7.24) ∂t δv(x, t) d 0 R ∂v(x, t) δΛ[u, v] = −D∇2v(x, t) − . (7.31) We can now eliminate z by functionally minimizing the ∂t δu(x, t) action Sb with respect to z, noting that q and λ are func- tions of z via their dependence on φ: Using Hamilton-Jacobi theory, it follows that evaluating the action along a “zero energy” least action path gives Z δSb dq dφ dλ dφ Sopt = Φ[u], where Φ is a solution of the Hamilton-Jacobi 0 = = z + q − − dx vF0(u) δz(t) dφ dz dφ dz d equation R Z δΦ + dx vF1(u). (7.25) d H[u, δuΦ] = 0, δuΦ = . (7.32) R δu(x) 21

P One solution is v = δuΦ = 0 for which Λ[u, 0] = 0, with n Qnm = 0 then gives δΛ[u, 0]/δu(x) = 0 and R [u, 0] = ρ with ρ the sta- n n tionary distribution of the Markov chain. Functionally X δΛ[u, v] Fn(u(x)ρn = . (7.34) differentiating the eigenvalue Eq. (7.28) with respect to δv(x) n v=0 v(x) shows that Hence, along the least action path, u(x, t) evolves accord- ing to the deterministic mean field equation

∂u X = D∇2u + F (u), F (u(x)) = ρ F (u(x)). Z  ∂t n n X δRm n dx v(x)Fn(u(x))δm,n + Qnm d δv(x) (7.35) m≥0 R δR δΛ + F (u(x))R = Λ n + R . (7.33) In future work one could explore how least action paths n n δv(x) δv(x) n with v 6= 0 contribute to the Kramer’s escape rate from a metastable state.

REFERENCES Summing both sides with respect to n and setting v = 0

[1] A. M. Berezhkovskii, M. A. Pusovoit and S. M. Bezrukov, (2014). Channel-facilitated membrane transport: Transit proba- [13] M. H. A. Davis, Piecewise-deterministic Markov pro- bility and interaction with the channel J. Chem. Phys. cesses: A general class of non-diffusion stochastic models. 116 9952-9956 (2002). Journal of the Royal Society, Series B (Methodological), [2] A. M. Berezhkovskii, M. A. Pusovoit and S. M. Bezrukov, 46, 353-388 (1984). Channel-facilitated membrane transport: Average life- [14] T. B. Kepler and T. C. Elston, Stochasticity in transcrip- times in the channel J. Chem. Phys. 119 3943-3951 tional regulation: Origins, consequences, and mathemat- (2003). ical representations. Biophys. J. 81 3116-3136 (2001). [3] J. Reingruber and D. Holcman. Narrow escape for a [15] R. Karmakar and I. Bose, Graded and binary responses stochastically gated Brownian ligand. J. Phys. Cond. in stochastic gene expression Phys. Biol. 1197-204 (2004) Matt. 22 065103 (2010). [16] J. M. Newby, Isolating intrinsic noise sources in a [4] S. D. Lawley, J. C. Mattingly and M. C. Reed, Stochastic stochastic genetic switch. Phys. Biol. 9 026002 (2012) switching in infinite dimensions with applications to ran- [17] J. M. Newby, Bistable switching asymptotics for the self dom parabolic PDEs. SIAM J. Math. Anal. 47 3035-3063 regulating gene. J. Phys. A 48 185001 (2015) (2015). [18] P. G. Hufton, Y. T. Lin, T. Galla and A. J. McKane, [5] P. C. Bressloff and S. D. Lawley, Moment equations for Intrinsic noise in systems with switching environments a piecewise deterministic PDE. J. Phys. A 48 105001 Phys. Rev. E 93 052119 (2016). (2015). [19] P. C. Bressloff, Stochastic switching in biology: from [6] P. C. Bressloff and S. D. Lawley Escape from subcellular genotype to phenotype (Topical Review) J. Phys. A 50 domains with randomly switching boundaries. Multiscale 133001 (2017) Model. Simul. 13 1420-1455 (2015). [20] R. F. Fox and Y. N. Lu, Emergent collective behavior in [7] N. Agmon and J. J. Hopfield, Transient kinetics of chemi- large numbers of globally coupled independent stochastic cal reactions with bounded diffusion perpendicular to the ion channels. Phys. Rev. E 49 3421-3431 (1994) reaction coordinate: Intramolecular processes with slow [21] C. C. Chow and J. A. White, Spontaneous action poten- conformational changes J. Chem. Phys. 80 592 (1984). tials due to channel fluctuations. Biophys. J. 71 3013– [8] C. R. Doering and J. C. Gadoua, Resonant activation 3021 (1996) over a fluctuating barrier Phys. Rev. Lett. 69 2318-2321 [22] J. H. Goldwyn and E. Shea-Brown, The what and where (1992) of adding channel noise to the Hodgkin-Huxley equations. [9] P. Reimann, R. Bartussek and P. Hanggi, Reaction rates PLoS Comp. Biol. 7) e1002247 (2011) when barriers fluctuate: A singular perturbation ap- [23] E. Buckwar and M. G. Riedler, An exact stochastic proach. Chem Phys. 235 11-26 (1998). hybrid model of excitable membranes including spatio- [10] J. Ankerhold and P. Pechukas, Mathematical aspects of temporal evolution J. Math. Biol. 63 1051-1093 (2011) the fluctuating barrier problem. Explicit equilibrium and [24] J. P. Keener and J. M. Newby, Perturbation analysis of relaxation solutions Physica A 261 458-470 (1998). spontaneous action potential initiation by stochastic ion [11] P. C. Bressloff, Diffusion in cells with stochastically-gated channels. Phy. Rev. E 84 011918 (2011) gap junctions. SIAM J. Appl. Math. 76 1658-1682 (2016). [25] J. M. Newby, P. C. Bressloff and J. P. Keeener, Break- [12] P. C. Bressloff, Stochastic Processes in Cell Biology Inter- down of fast-slow analysis in an excitable system with disciplinary Applied Mathematics 41 Springer New York channel noise. Phys. Rev. Lett. 111 128101 (2013) 22

[26] P. C. Bressloff and J. M. Newby, Stochastic hybrid model Stat. Mech. In press (2021). of spontaneous dendritic NMDA spikes. Phys. Biol. Phys. [47] R. M. Radcliffe, Some properties of coherent spin states Biol. 11 016006 (2014) J. Phys. A 4 313-323 (1971) [27] J. M. Newby, Spontaneous excitability in the Morris– [48] E. Fradkin, Field Theories of Condensed Matter Lecar model with ion channel noise. SIAM J. Appl. Dyn. Cambridge University Press (2013) Syst. 13 1756-1791 (2014) [49] M. Sasai and P. G. Wolynes, Stochastic gene expression [28] M. C. Reed, S. Venakides, J. J. Blum, Approximate as a many-body problem. Proc. Natl. Acad. Sci. USA traveling waves in linear reaction-hyperbolic equations. 100 2374-2379 (2003) SIAM J. Appl. Math. 50 167-180 (1990) [50] K. Zhang, M. Sasai and J. Wang, Eddy currents and [29] A. Friedman and G. Craciun, A model of intracellular coupled landscapes for nonadiabatic and nonequilibrium transport of particles in an axon. J. Math. Biol. 51 217- complex system dynamics Proc. Nat. Acad. Sci. USA 110 246 (2005) 14930-14935 (2013) [30] J. M. Newby and P. C. Bressloff, Quasi-steady state re- [51] B. Bhattacharyya, J. Wang and M. Sasai, Stochastic epi- duction of molecular-based models of directed intermit- genetic dynamics of gene switching Phys. Rev. E 102 tent search. Bull. Math. Biol. 72 1840-1866 (2010) 042408 (2020) [31] P. C. Bressloff and J. M. Newby, Stochastic models of in- [52] P. C. Bressloff, B. M. Karamched, E. Levien and S. D. tracellular transport Rev. Mod. Phys. 85 135-196 (2013) Lawley, Diffusive transport in the presence of stochasti- [32] P. C. Bressloff and J. M. Newby, Path integrals and large cally gated absorption Phys. Rev. E 96 022102 (2017). deviations in stochastic hybrid systems. Phys. Rev. E 89 [53] B. A. Reid, U. C. Tauber, and J. C. Brunson, Reaction- 042701 (2014). controlled diffusion: Monte Carlo simulations. Phys. Rev. [33] P. C. Bressloff, Path-integral methods for analyzing the E 64 046121 (2003). effects of fluctuations in stochastic hybrid neural net- [54] U. C. T¨auber, M. Howard M and B. P. Vollmayr-Lee, Ap- works J. Math. Neurosci. 5 4 (2015) plications of field-theoretic renormalization group meth- [34] P. C. Martin, E. D. Siggia and H. A. Rose, Statistical ods to reaction-diffusion problems J. Phys. A: Math. Gen. dynamics of classical systems Phys. Rev. A 8 423-437 38 R79-R131 (2005). (1973). [55] J. Cardy, Reaction-diffusion processes. In S. Nazarenko [35] C. de Dominicis, Techniques de renormalisation de la and O. V. Zaboronski, editors, Non-equilibrium Statis- th´eoriedes champs et dynamique des ph´enom`enescri- tical Mechanics and Turbulence, London Mathematical tiques J. Phys. (Paris) 37 247-253 (1976). Society Lecture Note Series 108-131. Cambridge Univer- [36] H.-K. Janssen, On a Lagrangian for classical field dynam- sity Press, Cambridge (2008). ics and renormalization group calculations of dynamical [56] A. Altland and B. D. Simons, Condensed Matter Field critical properties Z. Phys. B 23 377-380 (1976). Theory 2nd ed Cambridge University Press, Cambridge [37] P. C. Bressloff, Construction of stochastic hybrid (2010). path integrals using “quantum-mechanical” operators. [57] A. Kamenev, Field Theory of Non-Equilibrium Systems arXiv:2012.07770 [cond-mat.stat-mech] (2021). Cambridge University Press, Cambridge (2014). [38] M. Doi, Second quantization representation for classical [58] U. C. T¨auber, Critical Dynamics - A Field Theory Ap- many-particle systems. J. Phys. A 9 1465-1477 (1976). proach to Equilibrium and Non-Equilibrium Scaling Be- [39] M. Doi, Stochastic theory of diffusion controlled reac- havior. Cambridge University Press, Cambridge (2014). tions. J. Phys. A 9 1479-1495 (1976). [59] J. J. Vastola and W. R. Holmes, Stochastic path integrals [40] L. Peliti, Path integral approach to birth-death processes can be derived like quantum mechanical path integrals on a lattice. Journal de Physique 46 1469-1483 (1985). arXiv:1909.12990 (2020). [41] M. F. Weber and E. Frey, Master equations and the [60] F. Cooper and J. Dawson. Auxiliary field loop expan- theory of stochastic path integrals Rep. Prog. Phys. 80 sion of the effective action for a class of stochastic partial 046601 (2017). differential equations. Annal. Phys. 365 118-154 (2016). [42] Y. Kifer, Large deviations and adiabatic transitions for [61] J. Zinn-Justin, and Critical Phe- dynamical systems and Markov processes in fully coupled nomena Oxford Science Publications, Oxford (2002). averaging Memoirs of the AMS 201 issue 944 [62] H. Kleinert, Path Integrals in Quantum Mechanics, [43] A. Faggionato, D. Gabrielli and M. R. Crivellari, Non- Statistics, Polymer Physics, and Financial Markets 5th equilibrium thermodynamics of piecewise deterministic ed World Scientific Publishing Company, New Jersey Markov Processes. J Stat Phys 137 259-304 (2009) (2009). [44] A. Faggionato, D. Gabrielli and M. R. Crivellari, Averag- [63] M. Buice and J. D. Cowan, Field-theoretic approach to ing and large deviation principles for fully-coupled piece- fluctuation effects in neural networks. Phys. Rev. E 75 wise deterministic Markov processes and applications to 051919 (2007). molecular motors. Markov Processes and Related Fields [64] M. Buice, J. D. Cowan and C. C. Chow, Systematic fluc- 16 497-548 (2010) tuation expansion for neural network activity equations. [45] P. C. Bressloff and O. Faugeras, On the Hamiltonian Neural Comput., 22 377 (2010). structure of large deviations in stochastic hybrid systems. [65] M. Buice and C. C. Chow, Correlations, fluctuations, J. Stat. Mech. 033206 (2017) and stability of a finite-size network of coupled oscilla- [46] P. C. Bressloff, Spin coherent states and stochastic hybrid tors Phys. Rev. E, 76 031118 (2007). path integrals. arXiv:2102.03878 [cond-mat.stat-mech] J.