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

Population Level Activity in Large Random Neural Networks

James MacLaurin, Moshe Silverstein, Pedro Vilanova
Abstract

We determine limiting equations for large asymmetric ‘spin glass’ networks. The initial conditions are not assumed to be independent of the disordered connectivity: one of the main motivations for this is that allows one to understand how the structure of the limiting equations depends on the energy landscape of the random connectivity. The method is to determine the convergence of the double empirical measure (this yields population density equations for the joint distribution of the spins and fields). An additional advantage to utilizing the double empirical measure is that it yields a means of obtaining accurate finite-dimensional approximations to the dynamics.

1 Introduction

This paper concerns the high-dimensional dynamics of asymmetric random neural networks of the form, for jIN={1,2,,N}j\in I_{N}=\{1,2,\ldots,N\}.

dxtj=(xtj/τ+βN1/2k=1NJjkλ(xtk))dt+σtdWtj,dx^{j}_{t}=\big{(}-x^{j}_{t}/\tau+\beta N^{-1/2}\sum_{k=1}^{N}J^{jk}\lambda(x^{k}_{t})\big{)}dt+\sigma_{t}dW^{j}_{t}, (1)

where λ\lambda is a Lipschitz function, τ\tau is a constant, and {Jjk}j,kIN\{J^{jk}\}_{j,k\in I_{N}} are sampled independently from a centered normal distribution of variance 11, {Wtj}jIN\{W^{j}_{t}\}_{j\in I_{N}} are Brownian Motions. We study the convergence of the double empirical measure

N1jINδ(𝐳[0,T]j,𝐆[0,T]j),\displaystyle N^{-1}\sum_{j\in I_{N}}\delta_{(\mathbf{z}^{j}_{[0,T]},\mathbf{G}^{j}_{[0,T]})}, (2)

where Gtj=N1/2k=1NJjkλ(xtk))G^{j}_{t}=N^{-1/2}\sum_{k=1}^{N}J^{jk}\lambda(x^{k}_{t})\big{)}. The dynamics of high-dimensional recurrent neural networks have many applications. They have been heavily applied to neuroscientific problems: many scholars think that they can be used to explain how the brain balances excitation and inhibition [7, 32, 19, 28, 12]. They have been used to study spatially-extended patterns in the brain [33, 34]. Most recently, it has been recognized that they are of fundamental importance to data science [6, 2, 22, 35]. For more applications, see the mongraph of Helias and Dahmen [26] and the recent survey in [13].

There exist limiting ‘correlation equations’ [14, 26] that describe the effective dynamics of high dimensional random neural networks. These constitute delayed integro-differential equations that have proven very difficult to analyze, particularly over short timescales. A related problem is that the correlation equations have only been determined from initial conditions that are independent of the connectivity. This means that they may not be accurate over longer timescales that diverge with NN. For example, many scholars are interested in understanding the nature of the limiting dynamics after the system attains a particular state (such as, if it enters an ‘energy well’ of specified characteristics, does it escape?). To address this question, one needs to start the dynamics at a particular point in the energy landscape of the connectivity (and therefore the initial condition is disorder-dependent).

The literature concerning large NN limiting equations for random neural networks has a complex history. Sompolinsky, Crisanti and Sommers anticipated that Path Integral methods would yield limiting dynamical equations [36] - the derivation was published in a later work [14]. We refer the reader to the excellent discussion in the monograph of Helias and Dahmen [26].

Path Integral methods (as practiced by physicists) yield population density equations by determining where the probability measure for the NN-dimensional system concentrates. In the probability literature, one of the most powerful means of addressing this question is the theory of Large Deviations [18]. Large Deviations theory was used to determine spin glass dynamics in the pioneering papers of Ben Arous and Guionnet [3, 5, 24]; they obtained the first rigorous results concerning the large NN limit of random neural networks. After this work, Grunwald employed Large Deviations theory to obtain correlation / response equations for random neural networks whose spins flip randomly between discrete states [23]. Moynot and Samuelides studied the non-Gaussian case [31]. Faugeras and MacLaurin extended the work of Ben Arous and Guionnet to include correlations in the connectivity [20]. Touboul and Cabana determined limiting equations for spatially-extended systems [10, 11]. Faugeras, Soret and Tanre [21] determined novel integral equations to describe the state of these systems.

On a related note, correlation-response equations for symmetric random neural networks were first derived by Crisanti, Horner Sommers [15] and Cugliandolo and Kurchan [16]. Ben Arous, Dembo and Guionnet [4] proved the accuracy of these correlation / response equations for symmetric random neural networks, employing concentration inequalities.

Broadly-speaking, this paper follows the approach of Ben Arous and Guionnet [3]. We employ the theory of Large Deviations to determine the large NN limit of the empirical measure. However the main novelties of our approach are:

  • We employ a general class of connectivity-dependent initial conditions. This unsurprisingly yields a different limiting dynamics as NN\to\infty. Connectivity-dependent initial conditions were employed in the papers of Ben Arous and Guionnet [5] (who studied dynamics started at the equilibrium distribution, in the high temperature regime) and Dembo and Subag [17].

  • We study the double empirical measure, that includes information about both the spins and the fields. This has several advantages: it facilitates finite-dimensional approximations to the dynamics that are very accurate, and it facilitates a broader class of disorder-dependent initial condition. For spin-glass dynamics, the Large Deviations of the double empirical measure was determined by Grunwald [23] for jump-Markov systems.

  • We include Replicas (i.e. MM copies of the system with the same connectivity, but independent Brownian Motions). This broadens the class of admissible disorder-dependent initial conditions.

  • The function λ\lambda can be unbounded and the diffusion coefficient σt\sigma_{t} can vary in time. The time-varying nature of σt\sigma_{t} is essential for studying how periodic environmental noise in the brain shapes the dynamics of random neural networks.

Notation: Let IN={1,2,,N}I_{N}=\{1,2,\ldots,N\} be the set of neuron indices. For any Polish space 𝒳\mathcal{X}, let 𝒫(𝒳)\mathcal{P}(\mathcal{X}) denote all probability measures over 𝒳\mathcal{X}. The space 𝒞([0,T],)\mathcal{C}([0,T],\mathbb{R}) is always endowed with the supremum topology (unless indicated otherwise), i.e.

x[0,T]=supt[0,T]|xt|\left\|x_{[0,T]}\right\|=\sup_{t\in[0,T]}|x_{t}|

For 𝐲N\mathbf{y}\in\mathbb{R}^{N}, 𝐲\left\|\mathbf{y}\right\| is the Euclidean norm. For any probability measures μ\mu and ν\nu over a Polish Space, let (μ||ν)\mathcal{R}(\mu||\nu) denotes the relative entropy of measure μ\mu with respect to ν\nu. For any two measures on the same metric space with metric dd, dW(,)d_{W}(\cdot,\cdot) indicates the Wasserstein distance, i.e.

dW(μ,ν)=infζ𝔼ζ[d(x,y)],d_{W}(\mu,\nu)=\inf_{\zeta}\mathbb{E}^{\zeta}\big{[}d(x,y)\big{]}, (3)

where the infimum is taken over all ζ\zeta on the product space such that the marginal of the first variable is equal to μ\mu and the marginal of the second variable is equal to ν\nu. In the particular case that μ,ν𝒞([0,T],M)\mu,\nu\in\mathcal{C}([0,T],\mathbb{R}^{M}), the distance is (unless otherwise indicated) d(x,y)=supt[0,T]suppIM|xtpytp|d(x,y)=\sup_{t\in[0,T]}\sup_{p\in I_{M}}\big{|}x^{p}_{t}-y^{p}_{t}\big{|}.

For any μ𝒫(𝒞([0,T],M)2)\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}, we write μ(1),μ(2)𝒫(𝒞([0,T],M))\mu^{(1)},\mu^{(2)}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)} to be the marginals over (respectively) the first MM variables and last MM variables.

2 Outline of Model and Main Results

We are going to rigorously determine the limiting dynamics of multiple replicas (with identical connections 𝐉\mathbf{J}, but with independent initial conditions and independent Brownian Motions). We let the superscript aa denote replica aIM={1,2,,M}a\in I_{M}=\{1,2,\ldots,M\}, and consider the system

dzta,j=\displaystyle dz^{a,j}_{t}= (zta,j/τ+Gta,j)dt+σtdWta,j where\displaystyle\big{(}-z^{a,j}_{t}/\tau+G^{a,j}_{t}\big{)}dt+\sigma_{t}dW^{a,j}_{t}\text{ where } (4)
Gta,j=\displaystyle G^{a,j}_{t}= N1/2kINJjkλ(zta,k).\displaystyle N^{-1/2}\sum_{k\in I_{N}}J^{jk}\lambda(z^{a,k}_{t}). (5)

We assume that λ𝒞2()\lambda\in\mathcal{C}^{2}(\mathbb{R}): this means in particular that there is a constant CλC_{\lambda} such that |λ(x)λ(y)|Cλ|xy||\lambda(x)-\lambda(y)|\leq C_{\lambda}|x-y|. The noise intensity tσtt\to\sigma_{t} is taken to be continuous and non-random, and such that for constants σ¯\underline{\sigma} and σ¯\bar{\sigma},

0<σ¯σtσ¯.0<\underline{\sigma}\leq\sigma_{t}\leq\bar{\sigma}. (6)

Our major motivation for time-varying diffusivity lies in neuroscience: often synaptic noise exhibits particular rhythms. It has been of major interest how these rhythms shape pattern formation [8].

The connectivities {Jjk}\{J^{jk}\} are taken to be independent centered Gaussian variables, with variance

𝔼[JjkJlm]=δ(j,l)δ(k,m).\mathbb{E}\big{[}J^{jk}J^{lm}\big{]}=\delta(j,l)\delta(k,m).

Let γN𝒫(N2)\gamma^{N}\in\mathcal{P}\big{(}\mathbb{R}^{N^{2}}\big{)} be their joint probability law. There are two cases for the initial conditions {z0a,j}jIN,aIM\{z^{a,j}_{0}\}_{j\in I_{N},a\in I_{M}} that are considered in this paper.

2.1 Assumptions on the Initial Conditions

2.1.1 Case 1: Connectivity-Dependent Initial Conditions

The probability law of the initial conditions is assumed to be such that for any measurable set 𝒜MN\mathcal{A}\subset\mathbb{R}^{MN},

(𝐙0𝒜)=𝒜ρ𝐉N(𝐱)𝑑𝐱,\displaystyle\mathbb{P}\big{(}\mathbf{Z}_{0}\in\mathcal{A}\big{)}=\int_{\mathcal{A}}\rho^{N}_{\mathbf{J}}(\mathbf{x})d\mathbf{x}, (7)

where the probability density ρ𝐉N:MN+\rho^{N}_{\mathbf{J}}:\mathbb{R}^{MN}\to\mathbb{R}^{+} is defined as follows.

Let μN\mu^{N} be the uniform Lebesgue Measure over the set MN\mathbb{R}^{MN}. Conditionally on a realization 𝐉\mathbf{J} of the random connections, let the probability density ρ𝐉N:MN+\rho^{N}_{\mathbf{J}}:\mathbb{R}^{MN}\to\mathbb{R}^{+} be such that for δN>0\delta_{N}>0 (with δ1=1\delta_{1}=1 and δN\delta_{N} decreasing to 0 as NN\to\infty), there exists some κ𝒫(2M)\kappa\in\mathcal{P}\big{(}\mathbb{R}^{2M}\big{)} such that

ρ𝐉N(𝐠)=\displaystyle\rho^{N}_{\mathbf{J}}(\mathbf{g})= χ{dW(μ^N(𝐳0,𝐆0),κ)δN}/Z𝐉N where\displaystyle\chi\big{\{}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0}),\kappa\big{)}\leq\delta_{N}\big{\}}/Z^{N}_{\mathbf{J}}\text{ where } (8)
Z𝐉N=\displaystyle Z^{N}_{\mathbf{J}}= μN(dW(μ^N(𝐳0,𝐆0),κ)δN) and\displaystyle\mu^{N}\big{(}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0}),\kappa\big{)}\leq\delta_{N}\big{)}\text{ and } (9)
μ^N(𝐳0,𝐆0)=\displaystyle\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0})= N1jINδ𝐳0j,𝐆0j𝒫(2M)\displaystyle N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{z}^{j}_{0},\mathbf{G}^{j}_{0}}\in\mathcal{P}\big{(}\mathbb{R}^{2M}\big{)} (10)

Roughly speaking, we need to assume that as NN\to\infty, the law of 𝐳0N\mathbf{z}^{N}_{0} behaves like its annealed average. Its assumed that (i) κ\kappa has a finite second moment in each of its variables, and (ii) we have the bound

lim¯NN1log𝔼[Z𝐉N]>,\displaystyle\underset{N\to\infty}{\underline{\lim}}N^{-1}\log\mathbb{E}[Z^{N}_{\mathbf{J}}]>-\infty, (11)

It is also assumed that for any ϵ>0\epsilon>0,

limN¯N1log(|N1logZ𝐉N𝔼[Z𝐉N]|ϵ)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\big{|}N^{-1}\log Z^{N}_{\mathbf{J}}-\mathbb{E}[Z^{N}_{\mathbf{J}}]\big{|}\geq\epsilon\big{)}<0. (12)

Define 𝔙0,𝔙~0M×M\mathfrak{V}_{0},\tilde{\mathfrak{V}}_{0}\in\mathbb{R}^{M\times M} to be the covariance matrix with entries, p,qIMp,q\in I_{M},

𝔙0pq\displaystyle\mathfrak{V}_{0}^{pq} =𝔼κ[λ(z0p)λ(z0q)]\displaystyle=\mathbb{E}^{\kappa}\big{[}\lambda(z^{p}_{0})\lambda(z^{q}_{0})\big{]} (13)
𝔙~0pq\displaystyle\tilde{\mathfrak{V}}_{0}^{pq} =𝔼κ[z0pz0q].\displaystyle=\mathbb{E}^{\kappa}\big{[}z^{p}_{0}z^{q}_{0}\big{]}. (14)

It is also assumed that both 𝔙0\mathfrak{V}_{0} and 𝔙~0\tilde{\mathfrak{V}}_{0} are invertible.

2.1.2 Case 2: Connectivity-Independent Initial Conditions

One can also assume that the initial conditions (z0j)jIN(z^{j}_{0})_{j\in I_{N}} are (i) independent of the connectivity, and (ii) sampled independently from a M\mathbb{R}^{M}-valued probabilistic distribution of bounded variance. This distribution is written as κ^𝒫(M)\hat{\kappa}\in\mathcal{P}(\mathbb{R}^{M}).

2.2 Main Result

Our main result is that the empirical measure converges to a fixed point of a mapping Φ:𝒰𝒰\Phi:\mathcal{U}\to\mathcal{U}. Here 𝒰𝒫(𝒞([0,T],M)2)\mathcal{U}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} is defined in (54), consisting of (i) a broad class of measures with nice regularity properties, and (ii) such that the empirical measure inhabits 𝒰\mathcal{U} with unit probability.

For any μ𝒰\mu\in\mathcal{U}, in the case of connectivity-dependent initial conditions Φ(μ)\Phi(\mu) is specified as follows. It is defined to be the law of Gaussian random variables (ztp,Gtp)pIMt[0,T]\big{(}z^{p}_{t},G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} such that (i) (z0p,G0p)pIM(z^{p}_{0},G^{p}_{0})_{p\in I_{M}} are distributed according to κ\kappa, and (ii) conditionally on the initial conditions, (Gtp)pIMt[0,T]\big{(}G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} is a Gaussian system such that 𝔼[Gsp]=𝔪sp(μ,𝐆0)\mathbb{E}[G^{p}_{s}]=\mathfrak{m}^{p}_{s}(\mu,\mathbf{G}_{0}) and the conditional variance is

𝔼[(Gsp𝔪sp(μ,𝐆0))(Gtq𝔪tq(μ,𝐆0))|𝐆0,𝐳0]=𝔚stμ,pq.\displaystyle\mathbb{E}\big{[}\big{(}G^{p}_{s}-\mathfrak{m}^{p}_{s}(\mu,\mathbf{G}_{0})\big{)}\big{(}G^{q}_{t}-\mathfrak{m}^{q}_{t}(\mu,\mathbf{G}_{0})\big{)}\;|\;\mathbf{G}_{0},\mathbf{z}_{0}\big{]}=\mathfrak{W}^{\mu,pq}_{st}. (15)

Here

𝔚stμ,pq\displaystyle\mathfrak{W}^{\mu,pq}_{st} =a,bIM𝔼μ[λ(zsp)λ(z0a)]𝔼μ[λ(ztq)λ(z0b](𝔙μ,01)ab\displaystyle=\sum_{a,b\in I_{M}}\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{s})\lambda(z^{a}_{0})\big{]}\mathbb{E}^{\mu}\big{[}\lambda(z^{q}_{t})\lambda(z^{b}_{0}\big{]}\big{(}\mathfrak{V}_{\mu,0}^{-1}\big{)}^{ab} (16)
𝔙μ,0pq\displaystyle\mathfrak{V}_{\mu,0}^{pq} =𝔼μ[λ(z0p)λ(z0q)]\displaystyle=\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{0})\lambda(z^{q}_{0})\big{]} (17)
𝔪sp(μ,𝐠)\displaystyle\mathfrak{m}^{p}_{s}(\mu,\mathbf{g}) =a,bIM𝔼μ[λ(zsp)λ(z0a](𝔙μ,01)abgb\displaystyle=\sum_{a,b\in I_{M}}\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{s})\lambda(z^{a}_{0}\big{]}\big{(}\mathfrak{V}_{\mu,0}^{-1}\big{)}^{ab}g^{b} (18)

Letting (W[0,T]p)pIM\big{(}W^{p}_{[0,T]}\big{)}_{p\in I_{M}} be Brownian Motions that are independent of 𝐆μ\mathbf{G}^{\mu} , we define (ztp)pIMt[0,T](z^{p}_{t})_{p\in I_{M}\fatsemi t\in[0,T]} to be the strong solution to the stochastic differential equation

dztp=(τ1ztp+Gtμ,p)dt+σtdWtp.\displaystyle dz^{p}_{t}=\big{(}-\tau^{-1}z^{p}_{t}+G^{\mu,p}_{t}\big{)}dt+\sigma_{t}dW^{p}_{t}. (19)

In the case of connectivity-independent initial conditions, Φ\Phi is defined as follows. One first defines (Gtp)pIMt[0,T]\big{(}G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} to be a centered Gaussian system such that

𝔼[GtpGsq]=𝔼μ[λ(ztp)λ(zsq)].\mathbb{E}\big{[}G^{p}_{t}G^{q}_{s}\big{]}=\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{t})\lambda(z^{q}_{s})\big{]}.

(z0p)pIM(z^{p}_{0})_{p\in I_{M}} is independent of (Gtp)pIMt[0,T]\big{(}G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} and distributed according to κ^\hat{\kappa}. For Brownian Motions (W[0,T]p)pIM\big{(}W^{p}_{[0,T]}\big{)}_{p\in I_{M}}, that are independent of 𝐆μ\mathbf{G}^{\mu}, ztpz^{p}_{t} is the strong solution of (19).

Theorem 1.

The mapping Φ\Phi is well-defined for all μ𝒰\mu\in\mathcal{U}. Furthermore there exists a unique probability measure ξ𝒫(𝒞([0,T],M)2)\xi\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} such that with unit probability,

limNN1jINδ(𝐳[0,T]j,𝐆[0,T]j)=ξ.\lim_{N\to\infty}N^{-1}\sum_{j\in I_{N}}\delta_{(\mathbf{z}^{j}_{[0,T]},\mathbf{G}^{j}_{[0,T]})}=\xi. (20)

ξ\xi is the unique measure such that Φ(ξ)=ξ\Phi(\xi)=\xi. Furthermore,

ξ=limnξ(n),\displaystyle\xi=\lim_{n\to\infty}\xi^{(n)}, (21)

where ξ(n+1)=Φ(ξ(n))\xi^{(n+1)}=\Phi(\xi^{(n)}) and ξ(1)\xi^{(1)} is any measure in 𝒰\mathcal{U}.

Remark. This theorem is useful because it also implies a means to efficiently determine the large NN limiting equations, through repeated application of the mapping Φ\Phi. Because the limiting system is Gaussian, one only needs to solve for its covariance matrix. See the discussion in Helias and Dahmen [26] for an alternative formulation of the limiting covariance function in terms of a PDE.

2.3 An Example System that Satisfies the Conditions of Section 2.1.1

We now outline a general system that satisfies the conditions of Section 2.1.1. Suppose that λ:\lambda:\mathbb{R}\to\mathbb{R} is an odd function. Define ι:𝒫(2×2)+\iota:\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)}\to\mathbb{R}^{+} to be

ι(μ)=(μ||p).\displaystyle\iota(\mu)=\mathcal{R}(\mu||p). (22)

Here p𝒫(2×2)p\in\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)} is centered and Gaussian, the law of random variables (z01,g01,z02,g02)(z^{1}_{0},g^{1}_{0},z^{2}_{0},g^{2}_{0}) that are such that 𝔼[(z0p)2]=1\mathbb{E}[(z^{p}_{0})^{2}]=1, 𝔼[z0pg0q]=0\mathbb{E}[z^{p}_{0}g^{q}_{0}]=0 and 𝔼[g0pg0q]=𝔼[λ(z0p)λ(z0q)]\mathbb{E}[g^{p}_{0}g^{q}_{0}]=\mathbb{E}[\lambda(z^{p}_{0})\lambda(z^{q}_{0})]. Let Ξ:𝒞(2)\Xi:\mathcal{C}\big{(}\mathbb{R}^{2}\big{)}\to\mathbb{R} be bounded.

Lemma 2.

Suppose that there is a unique η𝒫(2×2)\eta\in\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)} such that

supμ𝒫(2×2){𝔼μ[Ξ(z1,g1)]ι(μ)}=𝔼η[Ξ(z1,g1)]ι(η).\displaystyle\sup_{\mu\in\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)}}\big{\{}\mathbb{E}^{\mu}[\Xi(z^{1},g^{1})]-\iota(\mu)\big{\}}=\mathbb{E}^{\eta}[\Xi(z^{1},g^{1})]-\iota(\eta). (23)

Suppose also that

sup{𝔼μ[Ξ(z1,g1)+Ξ(z2,g2)]ι(μ):μ𝒫(2×2) and μ(1)=η(1) and μ(2)=η(1)}=2{𝔼η[Ξ(z1,g1)]ι(η)}.\sup\big{\{}\mathbb{E}^{\mu}[\Xi(z^{1},g^{1})+\Xi(z^{2},g^{2})]-\iota(\mu)\;:\;\mu\in\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)}\text{ and }\mu^{(1)}=\eta^{(1)}\text{ and }\mu^{(2)}=\eta^{(1)}\big{\}}\\ =2\big{\{}\mathbb{E}^{\eta}[\Xi(z^{1},g^{1})]-\iota(\eta)\big{\}}. (24)

Then the conditions of Section 2.1.1 are satisfied, substituting M=1M=1, κ=η(1)\kappa=\eta^{(1)} and δN=(logN)1\delta_{N}=(\log N)^{-1}.

Proof.

We show that

limN¯N1{log𝔼γ[(Z𝐉N)2]2log𝔼γ[Z𝐉N]}=0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\big{\{}\log\mathbb{E}^{\gamma}\big{[}(Z^{N}_{\mathbf{J}})^{2}\big{]}-2\log\mathbb{E}^{\gamma}\big{[}Z^{N}_{\mathbf{J}}\big{]}\big{\}}=0. (25)

It is immediate from (25) that (12) must be satisfied. Let {y0p,j}jIN1p2\{y^{p,j}_{0}\}_{j\in I_{N}\fatsemi 1\leq p\leq 2} be iid 𝒩(0,1)\mathcal{N}(0,1) random variables. Form the empirical measure μ^0N:=N1jINδ𝐲j,𝐆0j𝒫(4)\hat{\mu}^{N}_{0}:=N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{y}^{j},\mathbf{G}^{j}_{0}}\in\mathcal{P}\big{(}\mathbb{R}^{4}\big{)}.

Define

𝒜N=χ{dW(η(1),N1jINδy1,j,g1,j)(logN)1,dW(η(2),N1jINδy2,j,g2,j)(logN)1}.\displaystyle\mathcal{A}_{N}=\chi\bigg{\{}d_{W}\big{(}\eta^{(1)},N^{-1}\sum_{j\in I_{N}}\delta_{y^{1,j},g^{1,j}}\big{)}\leq(\log N)^{-1}\;,\;d_{W}\big{(}\eta^{(2)},N^{-1}\sum_{j\in I_{N}}\delta_{y^{2,j},g^{2,j}}\big{)}\leq(\log N)^{-1}\bigg{\}}.

We will first demonstrate that

limNN1log𝔼[𝒜Nexp(N𝔼μ^N[Ξ(z1,g1)+Ξ(z2,g2)])]=sup{𝔼μ[Ξ(z1,g1)+Ξ(z2,g2)]ι(μ):μ𝒫(2×2) and μ(1)=η(1) and μ(2)=η(1)}.\lim_{N\to\infty}N^{-1}\log\mathbb{E}\big{[}\mathcal{A}_{N}\exp\big{(}N\mathbb{E}^{\hat{\mu}^{N}}[\Xi(z^{1},g^{1})+\Xi(z^{2},g^{2})]\big{)}\big{]}=\\ \sup\big{\{}\mathbb{E}^{\mu}[\Xi(z^{1},g^{1})+\Xi(z^{2},g^{2})]-\iota(\mu)\;:\;\mu\in\mathcal{P}\big{(}\mathbb{R}^{2}\times\mathbb{R}^{2}\big{)}\text{ and }\mu^{(1)}=\eta^{(1)}\text{ and }\mu^{(2)}=\eta^{(1)}\big{\}}. (26)

It is straightforward to prove that for any a>0a>0 there exists a compact subset 𝔘a𝒫(4)\mathfrak{U}_{a}\subset\mathcal{P}\big{(}\mathbb{R}^{4}\big{)} such that

limN¯N1log(μ^0N𝔘a)a.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}_{0}\notin\mathfrak{U}_{a}\big{)}\leq-a. (27)

A simple Large Deviations estimate yields that, for any μ𝒫(4)\mu\in\mathcal{P}\big{(}\mathbb{R}^{4}\big{)}, and ϵ1\epsilon\ll 1,

N1log(μ^0NBϵ(μ))=ι(μ)+O(ϵ).\displaystyle N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}_{0}\in B_{\epsilon}(\mu)\big{)}=-\iota(\mu)+O(\sqrt{\epsilon}). (28)

(In fact this could also be proved using Corollary 11). Our choice that δN\delta_{N} goes to zero sufficiently slowly (i.e. δN=(logN)1\delta_{N}=(\log N)^{-1}) ensures that the rate function in (28) dominates the asymptotic estimate of the probability as NN\to\infty. Thus discretizing 𝔘a\mathfrak{U}_{a} into ϵ\epsilon balls, and then taking ϵ0+\epsilon\to 0^{+}, one obtains (26).

Now (26) also implies that

limNN1log𝔼[𝒜N]=2ι(η).\displaystyle\lim_{N\to\infty}N^{-1}\log\mathbb{E}\big{[}\mathcal{A}_{N}\big{]}=-2\iota(\eta). (29)

This holds because since μ𝔼μ[Ξ(y1,g1+Ξ(y2,g2)]\mu\to\mathbb{E}^{\mu}[\Xi(y^{1},g^{1}+\Xi(y^{2},g^{2})] is continuous, it becomes almost constant as NN\to\infty, as the radius of 𝒜N\mathcal{A}_{N} shrinks. (29) in turn implies (25). ∎

3 Proof Outline

The main goal of this paper is to prove Theorem 1 employing the theory of Large Deviations [18]. The method - similarly to the original work by Ben Arous and Guionnet [3] - is to (i) prove a Large Deviations Principle for the uncoupled system, and then (ii) perform an exponential change-of-measure using Girsanov’s Theorem to obtain the Large Deviations Principle for the coupled system, before (iii) proving that the rate function has a unique zero.

The three main differences between this paper and the early papers of Ben Arous and Guionnet is that we (i) study the convergence of the double empirical measure (2) (whereas Ben Arous and Guionnet study the convergence of the annealed empirical measure in their earlier papers [3]) (in the later works [25, 4] quenched asymptotics are determined) (ii) we employ disorder-dependent initial conditions and (iii) we employ replicas.

Our main focus is on proving Case 1 (i.e. the connectivity-dependent initial conditions). The proofs are broadly similar, however Case 1 is more difficult because it requires a uniform Large Deviations Principle for the conditioned probability laws.

3.1 Large Deviations of the Uncoupled System

We start by noting a Large Deviation Principle for the uncoupled system. Because we are employing general disorder-dependent initial conditions, we must determine a Large Deviations Principle for the conditioned probability law. To this end, we must first define the set 𝒴N\mathcal{Y}^{N} of all ‘valid initial conditions’ (basically the set of all initial points such that the empirical measure at time 0 is close to its limit). More precisely, we define 𝒴NNM×NM\mathcal{Y}^{N}\subset\mathbb{R}^{NM}\times\mathbb{R}^{NM} to be such that

𝒴N={(𝐳0,𝐠0):dW(μ^N(𝐳0,𝐠0),κ)δ~N},\displaystyle\mathcal{Y}^{N}=\big{\{}(\mathbf{z}_{0},\mathbf{g}_{0})\;:\;d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{g}_{0}),\kappa\big{)}\leq\tilde{\delta}_{N}\big{\}}, (30)

where δ~N=max{δN,(logN)1}\tilde{\delta}_{N}=\max\big{\{}\delta_{N},(\log N)^{-1}\big{\}} and μ^N(𝐳0,𝐠0)=N1jINδ𝐳0j,𝐠0j𝒫(2M)\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{g}_{0})=N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{z}^{j}_{0},\mathbf{g}^{j}_{0}}\in\mathcal{P}(\mathbb{R}^{2M}). Define the uncoupled dynamics,

ytp,j=z0p,j+0tσs𝑑Wsp,j,\displaystyle y^{p,j}_{t}=z^{p,j}_{0}+\int_{0}^{t}\sigma_{s}dW_{s}^{p,j}, (31)

and let P𝐳0N𝒫(𝒞([0,T],M)N)P^{N}_{\mathbf{z}_{0}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} be the law of {y[0,T]j}jIN\{y^{j}_{[0,T]}\}_{j\in I_{N}}, conditioned on the values at time 0. Write

G~tp,j=N1/2kINJjkλ(ytp,k).\displaystyle\tilde{G}^{p,j}_{t}=N^{-1/2}\sum_{k\in I_{N}}J^{jk}\lambda(y^{p,k}_{t}). (32)

Define γ𝐠0N𝒫(N2)\gamma^{N}_{\mathbf{g}_{0}}\in\mathcal{P}\big{(}\mathbb{R}^{N^{2}}\big{)} to be the law of the connections {Jjk}j,kIN\{J^{jk}\}_{j,k\in I_{N}}, conditioned on the event

{G~0p,j=g0p,j for each jIN,pIM}.\displaystyle\big{\{}\tilde{G}^{p,j}_{0}=g^{p,j}_{0}\text{ for each }j\in I_{N},\;p\in I_{M}\big{\}}. (33)

We note that γ𝐠0N\gamma^{N}_{\mathbf{g}_{0}} is Gaussian, but no longer of zero mean (in general). The mean of G~tp,j\tilde{G}^{p,j}_{t} is a function of the empirical measure and 𝐠0j\mathbf{g}^{j}_{0} (explicit formulae are outlined further below). Let Q𝐳0,𝐠0N𝒫(N2×𝒞([0,T],M)N)Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\in\mathcal{P}\big{(}\mathbb{R}^{N^{2}}\times\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} be the joint law of the uncoupled system, i.e.

Q𝐳0,𝐠0N=γ𝐠0NP𝐳0N.Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}=\gamma^{N}_{\mathbf{g}_{0}}\otimes P^{N}_{\mathbf{z}_{0}}. (34)

The first main result is a uniform Large Deviations Principle for the conditioned system.

Theorem 3.

Let 𝒜,𝒪(𝒫(𝒞([0,T],M)2))\mathcal{A},\mathcal{O}\in\mathcal{B}\big{(}\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\big{)}, such that 𝒪\mathcal{O} is open and 𝒜\mathcal{A} closed. Then

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N(𝐲[0,T],𝐆~[0,T])𝒜)\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}_{[0,T]})\in\mathcal{A}\big{)} infμ𝒜(μ)\displaystyle\leq-\inf_{\mu\in\mathcal{A}}\mathcal{I}(\mu) (35)
lim¯Ninf(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T])𝒪)\displaystyle\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}_{[0,T]})\in\mathcal{O}\big{)} infμ𝒪(μ).\displaystyle\geq-\inf_{\mu\in\mathcal{O}}\mathcal{I}(\mu). (36)

Here the rate function :𝒫(𝒞([0,T],M)2)+\mathcal{I}:\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\to\mathbb{R}^{+} is such that (μ)=\mathcal{I}(\mu)=\infty if μ𝒰\mu\notin\mathcal{U}, else otherwise

(μ)=I~μ(μ),\displaystyle\mathcal{I}(\mu)=\tilde{I}_{\mu}(\mu), (37)

where, I~μ(μ)\tilde{I}_{\mu}(\mu) is defined in (117). Furthermore \mathcal{I} is lower-semi-continuous and has compact level sets.

We will prove Theorem 3 by locally freezing the dependence of the fields {G~tp,j}\{\tilde{G}^{p,j}_{t}\} on the empirical measure. In order that we may do this, we must first define a regular subset 𝒬𝔞\mathcal{Q}_{\mathfrak{a}} (for a positive integer 𝔞1\mathfrak{a}\gg 1) which is such that (i) the empirical measure μ^N(𝐲)=N1jINδ𝐲[0,T]j𝒫(𝒞([0,T],M))\hat{\mu}^{N}(\mathbf{y})=N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{y}^{j}_{[0,T]}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)} inhabits with high probability, and (ii) there exist uniform bounds on the fluctuations in time. To this end, writing 𝒦𝔞\mathcal{K}_{\mathfrak{a}} to be the compact set specified in Lemma 24, define the set

𝒬𝔞={μ𝒫(𝒞([0,T],M)):μ𝒦𝔞 and suppIM𝔼μ[supt[0,T](ytp)2]𝔞 and  For all integers m𝔞 it holds that sup0im𝔼μ[suppIM(wti+1(m)pwti(m)p)2]Δm1/4}\mathcal{Q}_{\mathfrak{a}}=\bigg{\{}\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)}\;:\;\mu\in\mathcal{K}_{\mathfrak{a}}\text{ and }\sup_{p\in I_{M}}\mathbb{E}^{\mu}[\sup_{t\in[0,T]}(y^{p}_{t})^{2}\big{]}\leq\mathfrak{a}\text{ and }\\ \text{ For all integers }m\geq\mathfrak{a}\text{ it holds that }\sup_{0\leq i\leq m}\mathbb{E}^{\mu}\big{[}\sup_{p\in I_{M}}(w^{p}_{t^{(m)}_{i+1}}-w^{p}_{t^{(m)}_{i}})^{2}\big{]}\leq\Delta_{m}^{1/4}\bigg{\}} (38)

where Δm=T/m\Delta_{m}=T/m and ti(m)=iT/mt^{(m)}_{i}=iT/m. Write

𝔔=𝔞1𝔔𝔞.\displaystyle\mathfrak{Q}=\bigcup_{\mathfrak{a}\geq 1}\mathfrak{Q}_{\mathfrak{a}}. (39)
Lemma 4.

For any L>0L>0, there exists a>0a>0 such that

limN¯N1log(μ^N(𝐲)𝒬a)L\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{Q}_{a}\big{)}\leq-L (40)

The above lemma is proved in the Appendix. Next, for any ν𝒬\nu\in\mathcal{Q}, we define a centered Gaussian law βν𝒫(𝒞([0,T],)M)\beta_{\nu}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R})^{M}\big{)} as follows. We stipulate that βν\beta_{\nu} is the law of Gaussian random variables {Gtν,p}t[0,T],pIM\{G^{\nu,p}_{t}\}_{t\in[0,T],p\in I_{M}} with covariance structure

𝔼βν[Gsν,pGtν,q]=𝔼ν[λ(xsp)λ(xtq)]\displaystyle\mathbb{E}^{\beta_{\nu}}\big{[}G^{\nu,p}_{s}G^{\nu,q}_{t}\big{]}=\mathbb{E}^{\nu}\big{[}\lambda(x^{p}_{s})\lambda(x^{q}_{t})\big{]} (41)

This definition will be useful because for any jINj\in I_{N}, the law of G~[0,T]j\tilde{G}^{j}_{[0,T]} under γN\gamma^{N} is βμ^N(𝐲)\beta_{\hat{\mu}^{N}(\mathbf{y})}. In the following Lemma we collect some regularity estimates for the Gaussian Law βν\beta_{\nu}.

Lemma 5.

(i) βν\beta_{\nu} is a well-defined Gaussian probability law. (ii) Furthermore, the map tGtν,pt\to G^{\nu,p}_{t} is ‘uniformly continuous’ for all measures in 𝒰a\mathcal{U}_{a}, in the following sense. For any a>0a>0, and any ϵ>0\epsilon>0, there exists δ(a,ϵ)\delta(a,\epsilon) such that for all ν𝒰a\nu\in\mathcal{U}_{a},

supν𝒰asuppIM𝔼βν[sups,t[0,T]|st|δ(a,ϵ)|Gsν,pGtν,p|]ϵ\displaystyle\sup_{\nu\in\mathcal{U}_{a}}\sup_{p\in I_{M}}\mathbb{E}^{\beta_{\nu}}\bigg{[}\sup_{s,t\in[0,T]\fatsemi|s-t|\leq\delta(a,\epsilon)}\big{|}G^{\nu,p}_{s}-G^{\nu,p}_{t}\big{|}\bigg{]}\leq\epsilon (42)

In order that we may make sense of the disorder-dependent initial condition, we also require an understanding of the distribution of βν\beta_{\nu}, conditioned on the value of G~0\tilde{G}_{0}. To this end, for any ν𝒰\nu\in\mathcal{U} and any 𝐠M\mathbf{g}\in\mathbb{R}^{M}, let βν,𝐠𝒫(𝒞([0,T],M))\beta_{\nu,\mathbf{g}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)} be the probability law βν\beta_{\nu} conditioned on the event 𝐆0ν=𝐠\mathbf{G}^{\nu}_{0}=\mathbf{g}. Standard Gaussian identities [29] imply that βν,𝐠\beta_{\nu,\mathbf{g}} is also Gaussian, with the following mean and variance,

𝔼βν,𝐠[Gsν,p]=\displaystyle\mathbb{E}^{\beta_{\nu,\mathbf{g}}}\big{[}G^{\nu,p}_{s}\big{]}= 𝔪sp(ν,𝐠)\displaystyle\mathfrak{m}^{p}_{s}(\nu,\mathbf{g}) (43)
𝔼βν,𝐠[(Gsν,p𝔪sp(ν,𝐠))(Gtν,q𝔪tq(ν,𝐠)]=\displaystyle\mathbb{E}^{\beta_{\nu,\mathbf{g}}}\big{[}\big{(}G^{\nu,p}_{s}-\mathfrak{m}^{p}_{s}(\nu,\mathbf{g})\big{)}\big{(}G^{\nu,q}_{t}-\mathfrak{m}^{q}_{t}(\nu,\mathbf{g}\big{)}\big{]}= 𝔚stν,pq,\displaystyle\mathfrak{W}^{\nu,pq}_{st}, (44)

where

𝔚stμ,pq\displaystyle\mathfrak{W}^{\mu,pq}_{st} =a,bIM𝔼μ[λ(zsp)λ(z0a)]𝔼μ[λ(ztq)λ(z0b](𝔙μ,01)ab\displaystyle=\sum_{a,b\in I_{M}}\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{s})\lambda(z^{a}_{0})\big{]}\mathbb{E}^{\mu}\big{[}\lambda(z^{q}_{t})\lambda(z^{b}_{0}\big{]}\big{(}\mathfrak{V}_{\mu,0}^{-1}\big{)}^{ab} (45)
𝔪sp(μ,𝐠)\displaystyle\mathfrak{m}^{p}_{s}(\mu,\mathbf{g}) =a,bIM𝔼μ[λ(zsp)λ(z0a](𝔙μ,01)abgb\displaystyle=\sum_{a,b\in I_{M}}\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{s})\lambda(z^{a}_{0}\big{]}\big{(}\mathfrak{V}_{\mu,0}^{-1}\big{)}^{ab}g^{b} (46)
𝔙μ,0pq\displaystyle\mathfrak{V}_{\mu,0}^{pq} =𝔼μ[λ(z0p)λ(z0q)].\displaystyle=\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{0})\lambda(z^{q}_{0})\big{]}. (47)
Corollary 6.

(i) For any a>0a>0, and any ϵ>0\epsilon>0, there exists δ~(a,ϵ)\tilde{\delta}(a,\epsilon) such that for all ν𝒰a\nu\in\mathcal{U}_{a}, and all 𝐠M\mathbf{g}\in\mathbb{R}^{M},

supν𝒰asuppIM𝔼βν,𝐠[sups,t[0,T]|st|δ~(a,ϵ)|Gsν,p𝔪sp(ν,𝐠)Gtν,p+𝔪tp(ν,𝐠)|]ϵ\displaystyle\sup_{\nu\in\mathcal{U}_{a}}\sup_{p\in I_{M}}\mathbb{E}^{\beta_{\nu,\mathbf{g}}}\bigg{[}\sup_{s,t\in[0,T]\fatsemi|s-t|\leq\tilde{\delta}(a,\epsilon)}\big{|}G^{\nu,p}_{s}-\mathfrak{m}^{p}_{s}(\nu,\mathbf{g})-G^{\nu,p}_{t}+\mathfrak{m}^{p}_{t}(\nu,\mathbf{g})\big{|}\bigg{]}\leq\epsilon (48)

(ii) For any ϵ,a>0\epsilon,a>0, there exists a compact set 𝒞ϵ,a𝒞([0,T],M)\mathcal{C}_{\epsilon,a}\subset\mathcal{C}([0,T],\mathbb{R}^{M}) such that for all ν𝒬a\nu\in\mathcal{Q}_{a},

βν(𝒞ϵ,a)1ϵ,\displaystyle\beta_{\nu}(\mathcal{C}_{\epsilon,a})\geq 1-\epsilon, (49)

and for all 𝐠0M\mathbf{g}_{0}\in\mathbb{R}^{M} such that 𝐠0a\left\|\mathbf{g}_{0}\right\|\leq a,

βν,𝐠0(𝒞ϵ,a)1ϵ.\displaystyle\beta_{\nu,\mathbf{g}_{0}}(\mathcal{C}_{\epsilon,a})\geq 1-\epsilon. (50)

(iii) For 1jN1\leq j\leq N and any 𝐠0M\mathbf{g}_{0}\in\mathbb{R}^{M}, the law of G~[0,T]j\tilde{G}^{j}_{[0,T]} under γ𝐠0N\gamma^{N}_{\mathbf{g}_{0}} is identical to βμ^N(𝐲),𝐠0\beta_{\hat{\mu}^{N}(\mathbf{y}),\mathbf{g}_{0}}.

3.2 Exponential Tightness

To prove a Large Deviation Principle, one requires that the empirical measure inhabits a compact set with arbitrarily high probability. For any 𝐲𝒞([0,T],M)N\mathbf{y}\in\mathcal{C}([0,T],\mathbb{R}^{M})^{N}, write γ~𝐲N𝒫(𝒞([0,T],M)N)\tilde{\gamma}^{N}_{\mathbf{y}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} to be the law of the random variables (G~tp,j)jINpIMt[0,T](\tilde{G}^{p,j}_{t})_{j\in I_{N}\fatsemi p\in I_{M}\fatsemi t\in[0,T]}, and write γ~𝐲,𝐠0N\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}} to be γ~𝐲N\tilde{\gamma}^{N}_{\mathbf{y}} conditioned on the event in (33).

The following lemmas are needed for this proof.

Lemma 7.

For any L>0L>0, there exists a compact set 𝒞~L𝒫(𝒞([0,T],M))\tilde{\mathcal{C}}_{L}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)} such that the following holds. For any N1N\geq 1, and any {𝐲[0,T]j}jIN\{\mathbf{y}^{j}_{[0,T]}\}_{j\in I_{N}} such that μ^N(𝐲)𝒬L\hat{\mu}^{N}(\mathbf{y})\in\mathcal{Q}_{L},

N1logγ~𝐲N(μ^N(𝐆~)𝒞~L)L.\displaystyle N^{-1}\log\tilde{\gamma}^{N}_{\mathbf{y}}\big{(}\hat{\mu}^{N}(\tilde{\mathbf{G}})\notin\tilde{\mathcal{C}}_{L}\big{)}\leq-L. (51)

Also, as long as (𝐲0,𝐠0)𝒴N(\mathbf{y}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N},

limN¯sup(𝐲0,𝐠0)𝒴NN1logγ~𝐲,𝐠0N(μ^N(𝐆~)𝒞~L)L.\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{y}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}(\tilde{\mathbf{G}})\notin\tilde{\mathcal{C}}_{L}\big{)}\leq-L. (52)

For μ𝒫(𝒞([0,T],)M×𝒞([0,T],)M)\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R})^{M}\times\mathcal{C}([0,T],\mathbb{R})^{M}\big{)}, write μ(1)𝒫(𝒞([0,T],)M)\mu^{(1)}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R})^{M}\big{)} to be the marginal of μ\mu over its first MM variables, and μ(2)\mu^{(2)} to be the marginal of μ\mu over its last MM variables. Next, define the set

𝒰a={μ𝒫(𝒞([0,T],)M×𝒞([0,T],)M):μ(1)𝒬a,μ(2)𝒞~a and supt[0,T]suppIM𝔼μ[(Gtp)2]Cλ2a,for all 0s,tT,suppIM𝔼μ[(GtpGsp)2]aCλ2|ts|1/2},\mathcal{U}_{a}=\bigg{\{}\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R})^{M}\times\mathcal{C}([0,T],\mathbb{R})^{M}\big{)}\;:\;\mu^{(1)}\in\mathcal{Q}_{a},\;\mu^{(2)}\in\tilde{\mathcal{C}}_{a}\text{ and }\\ \sup_{t\in[0,T]}\sup_{p\in I_{M}}\mathbb{E}^{\mu}[(G^{p}_{t})^{2}]\leq C_{\lambda}^{2}a,\;\text{for all }0\leq s,t\leq T,\;\sup_{p\in I_{M}}\mathbb{E}^{\mu}[(G^{p}_{t}-G^{p}_{s})^{2}]\leq aC_{\lambda}^{2}|t-s|^{1/2}\bigg{\}}, (53)

and let

𝒰=a0𝒰a.\displaystyle\mathcal{U}=\bigcup_{a\geq 0}\mathcal{U}_{a}. (54)

It follows immediately from the above definition that dW(μ,ν)<d_{W}(\mu,\nu)<\infty for any μ,ν𝒰\mu,\nu\in\mathcal{U}. We can now prove an ‘exponential tightness’ result.

Lemma 8.

For any a0a\geq 0, 𝒰a\mathcal{U}_{a} is compact. For any L>0L>0, there exists a>0a>0 such that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒰a)L.\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\notin\mathcal{U}_{a}\big{)}\leq-L. (55)
Proof.

Since the sets 𝒬a\mathcal{Q}_{a} and 𝒞~a\tilde{\mathcal{C}}_{a} are compact, this follows almost immediately from Lemmas 4 and 7. ∎

3.3 The Coupled System (with connectivity-dependent initial conditions)

Define 𝒥:𝒫(𝒞([0,T],M)2)\mathcal{J}:\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\to\mathbb{R} to be such that 𝒥(μ)=\mathcal{J}(\mu)=\infty if μ𝒰\mu\notin\mathcal{U}, or if the marginal of μ\mu at time 0 is not κ\kappa. Else otherwise, for any μ𝒫(𝒞([0,T],M)2)\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} and 𝐳0,𝐠0M\mathbf{z}_{0},\mathbf{g}_{0}\in\mathbb{R}^{M}, writing μ𝐳0,𝐠0\mu_{\mathbf{z}_{0},\mathbf{g}_{0}} for μ\mu conditioned on the values of its variables at time 0, define

𝒥(μ)=𝔼κ[(μ𝐳0,𝐠0||Sμ,𝐳0,𝐠0)]\displaystyle\mathcal{J}(\mu)=\mathbb{E}^{\kappa}\bigg{[}\mathcal{R}\big{(}\mu_{\mathbf{z}_{0},\mathbf{g}_{0}}||S_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{)}\bigg{]} (56)

Here Sμ,𝐳0,𝐠0𝒫(𝒞([0,T],M)2)S_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} is defined to be the probability law of (𝐳,𝐆μ)(\mathbf{z},\mathbf{G}^{\mu}), where 𝐆μ\mathbf{G}^{\mu} is distributed according to βμ,𝐠0\beta_{\mu,\mathbf{g}_{0}}, and for Brownian Motions (W[0,T]p)pIM\big{(}W^{p}_{[0,T]}\big{)}_{p\in I_{M}} that are independent of 𝐆ν\mathbf{G}^{\nu}

dztp=(τ1ztp+Gtμ,p)dt+σtdWtp\displaystyle dz^{p}_{t}=\big{(}-\tau^{-1}z^{p}_{t}+G^{\mu,p}_{t}\big{)}dt+\sigma_{t}dW^{p}_{t} (57)

Define Φ:𝒰𝒰\Phi:\mathcal{U}\to\mathcal{U} to be the following map. For some μ𝒰\mu\in\mathcal{U}, write Φ(μ)\Phi(\mu) to be the law of the following random variables (𝐱,𝐡)(\mathbf{x},\mathbf{h}). First, it is stipulated that (𝐱0,𝐡0)(\mathbf{x}_{0},\mathbf{h}_{0}) have probability law κ\kappa. Second, conditionally on (𝐱0,𝐡0)(\mathbf{x}_{0},\mathbf{h}_{0}), the distribution of (𝐱[0,T],𝐡[0,T])(\mathbf{x}_{[0,T]},\mathbf{h}_{[0,T]}) is given by Sμ,𝐱0,𝐠0S_{\mu,\mathbf{x}_{0},\mathbf{g}_{0}}.

Lemma 9.

The probability law Sμ,𝐳0,𝐠0S_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}} is well-defined. Furthermore, there exists a unique zero ξ\xi of the rate function 𝒥\mathcal{J}. ξ\xi is the unique measure in 𝒰\mathcal{U} such that Φ(ξ)=ξ\Phi(\xi)=\xi.

Proof.

We have already proved in Lemma 5 that (G[0,T]μ,p)pIM\big{(}G^{\mu,p}_{[0,T]}\big{)}_{p\in I_{M}} is well-defined. It is straightforward to check that for any path (G[0,T]μ,p)pIM\big{(}G^{\mu,p}_{[0,T]}\big{)}_{p\in I_{M}}, there exists a unique strong solution to the stochastic differential equation (57). Thus the probability law is well-defined.

It is well-known that the Relative Entropy is zero if and only if its two arguments are identical [9]. Thus, from the form of 𝒥\mathcal{J} in (56), any zero must be a fixed point of Φ\Phi. Furthermore, there must exist at least one zero of the rate function (if not, the total probability mass could not be one as NN\to\infty). The uniqueness of the zero is proved in Lemma 22. ∎

Theorem 10.

For any ϵ>0\epsilon>0,

limN¯N1log(dW(μ^N(𝐳,𝐆),ξ)ϵ)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}d_{W}(\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\xi)\geq\epsilon\big{)}<0. (58)

Thus with unit probability,

limNμ^N(𝐳,𝐆)=ξ.\lim_{N\to\infty}\hat{\mu}^{N}(\mathbf{z},\mathbf{G})=\xi. (59)

Furthermore,

ξ=limnξ(n),\displaystyle\xi=\lim_{n\to\infty}\xi^{(n)}, (60)

where ξ(n+1)=Φ(ξ(n))\xi^{(n+1)}=\Phi(\xi^{(n)}) and ξ(1)\xi^{(1)} is any measure in 𝒰\mathcal{U}.

3.4 Connectivity-Independent Initial Conditions

The above reasoning can be adapted to prove a Large Deviation Principle for the unconditioned system. This is needed for proving the main theorem for Case 2 (connectivity-independent initial conditions). Write QN=γNPNQ^{N}=\gamma^{N}\otimes P^{N} to be the law of the random variables (𝐲,𝐆)(\mathbf{y},\mathbf{G}) (with no conditioning), and for any ν𝔔\nu\in\mathfrak{Q}, define Sν𝒫(𝒞([0,T],M)2)S_{\nu}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} to be Sν=PβνS_{\nu}=P\otimes\beta_{\nu}. In the following corollary to Theorem 3, we prove the Large Deviation Principle for the unconditioned and uncoupled system.

Corollary 11.

Let 𝒜,𝒪(𝒫(𝒞([0,T],M)2))\mathcal{A},\mathcal{O}\in\mathcal{B}\big{(}\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\big{)}, such that 𝒪\mathcal{O} is open and 𝒜\mathcal{A} closed. Then

limN¯N1logQN(μ^N(𝐲[0,T],𝐆~[0,T])𝒜)\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log Q^{N}\big{(}\hat{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}_{[0,T]})\in\mathcal{A}\big{)} infμ𝒜(μ||Sμ(1))\displaystyle\leq-\inf_{\mu\in\mathcal{A}}\mathcal{R}\big{(}\mu||S_{\mu^{(1)}}\big{)} (61)
lim¯NN1logQN(μ^N(𝐲[0,T],𝐆~[0,T])𝒪)\displaystyle\underset{N\to\infty}{\underline{\lim}}N^{-1}\log Q^{N}\big{(}\hat{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}_{[0,T]})\in\mathcal{O}\big{)} infμ𝒪(μ||Sμ(1)).\displaystyle\geq-\inf_{\mu\in\mathcal{O}}\mathcal{R}\big{(}\mu||S_{\mu^{(1)}}\big{)}. (62)

Here the rate function μ(μ||Sμ(1))\mu\to\mathcal{R}\big{(}\mu||S_{\mu^{(1)}}\big{)} is lower semi-continuous and has compact level sets.

We now specify the operator Φ~:𝒰𝒰\tilde{\Phi}:\mathcal{U}\to\mathcal{U}. Fix μ𝒰\mu\in\mathcal{U} and defined Φ~(μ)\tilde{\Phi}(\mu) to be the law of processes (z[0,T]p,G[0,T]p)pIMt[0,T]\big{(}z^{p}_{[0,T]},G^{p}_{[0,T]}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]}. One first defines (Gtp)pIMt[0,T]\big{(}G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} to be centered Gaussian system such that

𝔼[GtpGsq]=𝔼μ[λ(ztp)λ(zsq)].\mathbb{E}\big{[}G^{p}_{t}G^{q}_{s}\big{]}=\mathbb{E}^{\mu}\big{[}\lambda(z^{p}_{t})\lambda(z^{q}_{s})\big{]}.

(z0p)pIM(z^{p}_{0})_{p\in I_{M}} is independent of (Gtp)pIMt[0,T]\big{(}G^{p}_{t}\big{)}_{p\in I_{M}\fatsemi t\in[0,T]} and distributed according to κ^\hat{\kappa}. Letting (W[0,T]p)pIM\big{(}W^{p}_{[0,T]}\big{)}_{p\in I_{M}} be Brownian Motions that are independent of 𝐆μ\mathbf{G}^{\mu} , we define (ztp)pIMt[0,T](z^{p}_{t})_{p\in I_{M}\fatsemi t\in[0,T]} to be the strong solution to the stochastic differential equation

dztp=(τ1ztp+Gtμ,p)dt+σtdWtp.\displaystyle dz^{p}_{t}=\big{(}-\tau^{-1}z^{p}_{t}+G^{\mu,p}_{t}\big{)}dt+\sigma_{t}dW^{p}_{t}. (63)
Theorem 12.

Assume the connectivity-independent initial conditions (Case 2). For any ϵ>0\epsilon>0,

limN¯N1log(dW(μ^N(𝐳,𝐆),ξ)ϵ)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}d_{W}(\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\xi)\geq\epsilon\big{)}<0. (64)

Thus with unit probability,

limNμ^N(𝐳,𝐆)=ξ.\lim_{N\to\infty}\hat{\mu}^{N}(\mathbf{z},\mathbf{G})=\xi. (65)

Furthermore,

ξ=limnξ(n),\displaystyle\xi=\lim_{n\to\infty}\xi^{(n)}, (66)

where ξ(n+1)=Φ~(ξ(n))\xi^{(n+1)}=\tilde{\Phi}(\xi^{(n)}) and ξ(1)\xi^{(1)} is any measure in 𝒰\mathcal{U}.

4 Proofs

We have divided the proofs into three main sections. In Section 4.1, we prove general regularity properties of the stochastic processes. In Section 4.2, we prove the LDP for the uncoupled system. In Section 4.3, we determine the limiting dynamics of the coupled system.

4.1 Regularity Estimates and Compactness

We first prove Lemma 5.

Proof.

We first check that the covariance function is positive definite (when restricted to a finite set of times). Let {ti}1im[0,T]\{t_{i}\}_{1\leq i\leq m}\subset[0,T] be a finite set of times. Then evidently for any constants {αip}pIM,1im\{\alpha^{p}_{i}\}_{p\in I_{M},1\leq i\leq m}, it must be that

p,qIM1i,jmαipαjq𝔼ν[λ(xtip)λ(xtjq)]=𝔼ν[(pIM1imαipλ(xtip))2]0.\displaystyle\sum_{p,q\in I_{M}}\sum_{1\leq i,j\leq m}\alpha^{p}_{i}\alpha^{q}_{j}\mathbb{E}^{\nu}\big{[}\lambda\big{(}x^{p}_{t_{i}}\big{)}\lambda\big{(}x^{q}_{t_{j}}\big{)}\big{]}=\mathbb{E}^{\nu}\big{[}\big{(}\sum_{p\in I_{M}}\sum_{1\leq i\leq m}\alpha^{p}_{i}\lambda\big{(}x^{p}_{t_{i}}\big{)}\big{)}^{2}\big{]}\geq 0. (67)

This means that there exists a finite set of centered Gaussian variables {Gti(m)ν,p}pIM1im\{G^{\nu,p}_{t^{(m)}_{i}}\}_{p\in I_{M}\fatsemi 1\leq i\leq m} such that (41) holds. It then follows from the Komolgorov Extension Theorem that βν\beta_{\nu} is well-defined on any countably dense subset of times of [0,T][0,T]. It remains for us to demonstrate continuity, i.e. that there exists a Gaussian probability law such that (41) holds for all time. We do this using standard theory for the continuity of Gaussian Processes (following Chapter 2 of [1]).

First, we notice that

suppIMsupt[0,T]𝔼[(Gtν,p)2]<.\displaystyle\sup_{p\in I_{M}}\sup_{t\in[0,T]}\mathbb{E}\big{[}(G^{\nu,p}_{t})^{2}\big{]}<\infty. (68)

Now define the canonical metric,

d¯p(s,t)=\displaystyle\bar{d}_{p}(s,t)= 𝔼[(Gsν,pGtν,p)2]12=𝔼ν[(λ(xsp)λ(xtp))2]12\displaystyle\mathbb{E}\big{[}\big{(}G^{\nu,p}_{s}-G^{\nu,p}_{t}\big{)}^{2}\big{]}^{\frac{1}{2}}=\mathbb{E}^{\nu}\big{[}\big{(}\lambda(x^{p}_{s})-\lambda(x^{p}_{t})\big{)}^{2}\big{]}^{\frac{1}{2}} (69)
\displaystyle\leq ConstsuppIM𝔼ν[|xspxtp|2]12a(ts)14\displaystyle\rm{Const}\sup_{p\in I_{M}}\mathbb{E}^{\nu}\big{[}\big{|}x^{p}_{s}-x^{p}_{t}\big{|}^{2}\big{]}^{\frac{1}{2}}\leq a\;(t-s)^{\frac{1}{4}} (70)

thanks to properties of the set 𝒬a\mathcal{Q}_{a}, for all s,ts,t such that |st||s-t| is smaller than some constant depending on aa. It follows from Theorem 1.4.1 of [1] that the Gaussian Process is almost-surely continuous.

Write Bt(ϵ)={s[0,T]:d¯(s,t)ϵ}B_{t}(\epsilon)=\big{\{}s\in[0,T]:\bar{d}(s,t)\leq\epsilon\big{\}} to be the ϵ\epsilon-ball about tt, and let 𝒩(ϵ)\mathcal{N}(\epsilon) denote the smallest number of such balls that cover TT. We see that there exists a constant 𝔠a>0\mathfrak{c}_{a}>0 such that

𝒩(ϵ)𝔠aϵ4.\displaystyle\mathcal{N}(\epsilon)\leq\mathfrak{c}_{a}\epsilon^{-4}. (71)

Writing H(ϵ)=log𝒩(ϵ)H(\epsilon)=\log\mathcal{N}(\epsilon), it follows from Theorem 1.3.5 in [1] that there exist MM Gaussian Processes (Gtν,p)t[0,T](G^{\nu,p}_{t})_{t\in[0,T]} such that tGtν,pt\to G^{\nu,p}_{t} is almost-surely continuous, and there exists a universal constant 𝔎>0\mathfrak{K}>0 and a random η>0\eta>0 such that for all δ<η\delta<\eta,

suppIMs,tTd¯(s,t)δ|Gsν,pGtν,p|\displaystyle\sup_{p\in I_{M}\fatsemi s,t\leq T\fatsemi\bar{d}(s,t)\leq\delta}\big{|}G^{\nu,p}_{s}-G^{\nu,p}_{t}\big{|} 𝔎0δH1/2(ϵ)𝑑ϵ\displaystyle\leq\mathfrak{K}\int_{0}^{\delta}H^{1/2}(\epsilon)d\epsilon (72)
𝔎0δ(4log(ϵ1)+log𝔠a)12𝑑ϵ,\displaystyle\leq\mathfrak{K}\int_{0}^{\delta}\big{(}4\log\big{(}\epsilon^{-1}\big{)}+\log\mathfrak{c}_{a}\big{)}^{\frac{1}{2}}d\epsilon, (73)

and we note that the above goes to 0 as δ0+\delta\to 0^{+}. This also implies (42). ∎

We next prove Corollary 6.

Proof.

The proof of (48) is analogous to the proof of (42). Notice that 𝔪tp(μ,𝐠)\mathfrak{m}^{p}_{t}(\mu,\mathbf{g}) depends continuously on 𝐠\mathbf{g}.

(42) and (48) imply (respectively) (49) and (50) . ∎

We can now prove Lemma 7.

Proof.

We prove (52) only. The other proof is very similar.

It follows from Lemma 5 that for any ϵ>0\epsilon>0, there exists a compact set 𝒞ϵ𝒞([0,T],M)\mathcal{C}_{\epsilon}\subset\mathcal{C}([0,T],\mathbb{R}^{M}) such that for any μ𝒬L\mu\in\mathcal{Q}_{L}, and all 𝐠0M\mathbf{g}_{0}\in\mathbb{R}^{M} such that 𝐠0ϵ1\left\|\mathbf{g}_{0}\right\|\leq\epsilon^{-1},

βμ,𝐠0(G[0,T]μ𝒞ϵ)ϵ.\displaystyle\beta_{\mu,\mathbf{g}_{0}}\big{(}G^{\mu}_{[0,T]}\notin\mathcal{C}_{\epsilon}\big{)}\leq\epsilon. (74)

It has already been noted above that for any {𝐲[0,T]j}jIN𝒞([0,T],M)\{\mathbf{y}_{[0,T]}^{j}\}_{j\in I_{N}}\subset\mathcal{C}([0,T],\mathbb{R}^{M}), {G~[0,T]j}jIN\{\tilde{G}^{j}_{[0,T]}\}_{j\in I_{N}} are independent, and the probability law of G~[0,T]j\tilde{G}^{j}_{[0,T]} is βμ^N(𝐲),𝐠0\beta_{\hat{\mu}^{N}(\mathbf{y}),\mathbf{g}_{0}}. Thus as long as μ^N(𝐲)𝒬L\hat{\mu}^{N}(\mathbf{y})\in\mathcal{Q}_{L}, the estimate in (74) holds for any G~[0,T]j\tilde{G}^{j}_{[0,T]} such that 𝐠0jϵ1\|\mathbf{g}^{j}_{0}\|\leq\epsilon^{-1}.

Define

θN,ϵ(𝐠0)=N1jINχ{𝐠0j>ϵ1}.\displaystyle\theta_{N,\epsilon}(\mathbf{g}_{0})=N^{-1}\sum_{j\in I_{N}}\chi\{\|\mathbf{g}^{j}_{0}\|>\epsilon^{-1}\}. (75)

Our construction of 𝒴N\mathcal{Y}^{N} implies that

limϵ0+limNsup(𝐳0,𝐠0)𝒴NθN,ϵ(𝐠0)=0.\displaystyle\lim_{\epsilon\to 0^{+}}\lim_{N\to\infty}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}\theta_{N,\epsilon}(\mathbf{g}_{0})=0. (76)

Write IN,ϵ={j:𝐠0jϵ1}I_{N,\epsilon}=\big{\{}j\;:\|\mathbf{g}^{j}_{0}\|\leq\epsilon^{-1}\big{\}}. Next, for some δ>ϵ/2\delta>\epsilon/2, and b>0b>0, by Chernoff’s Inequality, for any N1N\geq 1 and any (𝐲0,𝐠0)𝒴N(\mathbf{y}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N},

γ~𝐲,𝐠0N(N1jIN{G[0,T]j𝒞ϵ}δ)\displaystyle\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\bigg{(}N^{-1}\sum_{j\in I_{N}}\{G^{j}_{[0,T]}\notin\mathcal{C}_{\epsilon}\}\geq\delta\bigg{)} γ~𝐲,𝐠0N(N1jI~N,ϵ{G[0,T]j𝒞ϵ}δ/2)\displaystyle\leq\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\bigg{(}N^{-1}\sum_{j\in\tilde{I}_{N,\epsilon}}\{G^{j}_{[0,T]}\notin\mathcal{C}_{\epsilon}\}\geq\delta/2\bigg{)}
𝔼γ~𝐲,𝐠0N[exp(bjI~N,ϵ{G[0,T]j𝒞ϵ}bNδ/2)]\displaystyle\leq\mathbb{E}^{\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}}\bigg{[}\exp\bigg{(}b\sum_{j\in\tilde{I}_{N,\epsilon}}\{G^{j}_{[0,T]}\notin\mathcal{C}_{\epsilon}\}-bN\delta/2\bigg{)}\bigg{]}
{ϵexp(b)+1ϵ}Nexp(Nbδ/2)\displaystyle\leq\big{\{}\epsilon\exp(b)+1-\epsilon\big{\}}^{N}\exp\big{(}-Nb\delta/2\big{)}

We thus find that (by taking small enough ϵ\epsilon, and 1blogϵ1\ll b\ll-\log\epsilon), for any integer nn, that there must exist a compact set 𝒞(n)\mathcal{C}^{(n)} such that for all N1N\geq 1, all (𝐲0,𝐠0)𝒴N(\mathbf{y}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N} and all 𝐲\mathbf{y} such that μ^N(𝐲)𝒬L\hat{\mu}^{N}(\mathbf{y})\in\mathcal{Q}_{L}

N1logγ~𝐲,𝐠0N(N1jIN{G[0,T]j𝒞(n)}n1)n2.\displaystyle N^{-1}\log\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\big{(}N^{-1}\sum_{j\in I_{N}}\{G^{j}_{[0,T]}\notin\mathcal{C}^{(n)}\}\geq n^{-1}\big{)}\leq-n^{2}. (77)

This motivates us to define the compact set 𝒞~L𝒫(𝒞([0,T],M)2)\tilde{\mathcal{C}}_{L}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} to consist of all measures μ\mu such that for all nmn\geq m,

μ(𝒞(n))1n1.\displaystyle\mu\big{(}\mathcal{C}^{(n)}\big{)}\geq 1-n^{-1}. (78)

Thus using a union-of-events bound,

γ~𝐲,𝐠0N(μ^N(𝐆)𝒞~L)\displaystyle\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{G})\notin\tilde{\mathcal{C}}_{L}\big{)} nmγ~𝐲,𝐠0N(N1jIN{G[0,T]j𝒦(n)}n1)\displaystyle\leq\sum_{n\geq m}\tilde{\gamma}^{N}_{\mathbf{y},\mathbf{g}_{0}}\bigg{(}N^{-1}\sum_{j\in I_{N}}\{G^{j}_{[0,T]}\notin\mathcal{K}^{(n)}\}\geq n^{-1}\bigg{)} (79)
nmexp(Nn2)\displaystyle\leq\sum_{n\geq m}\exp\big{(}-Nn^{2}\big{)} (80)
exp(NL),\displaystyle\leq\exp(-NL), (81)

for all N1N\geq 1, as long as mm is large enough.

The following bound on the operator norm of the connectivity matrix is well-known (and the proof is omitted).

Lemma 13.

For any L>0L>0, there exists \ell such that

limN¯N1log(𝒥N)L,\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\geq\ell\big{)}\leq-L, (82)

where 𝒥NN×N\mathcal{J}_{N}\in\mathbb{R}^{N\times N} has (j,k)(j,k) entry

𝒥N,jk=N1/2Jjk\mathcal{J}_{N,jk}=N^{-1/2}J^{jk}
Lemma 14.

For any >0\ell>0, there exists L>0L>0 such that for all pIMp\in I_{M} and all N1N\geq 1,

N1log(𝒜c,supt[0,T]jIN(ztp,j)2N)L\displaystyle N^{-1}\log\mathbb{P}\big{(}\mathcal{A}_{c}\;,\;\sup_{t\in[0,T]}\sum_{j\in I_{N}}(z^{p,j}_{t})^{2}\geq N\ell\big{)}\leq-L (83)

where

𝒜c={𝒥Nc,suppIMjIN(z0p,j)2N𝔼κ[(z0p)2]+N}.\mathcal{A}_{c}=\bigg{\{}\left\|\mathcal{J}_{N}\right\|\leq c\;,\;\sup_{p\in I_{M}}\sum_{j\in I_{N}}(z^{p,j}_{0})^{2}\leq N\mathbb{E}^{\kappa}[(z^{p}_{0})^{2}]+N\bigg{\}}.
Proof.

Write

ut=N1jIN(ztp,j)2.u_{t}=N^{-1}\sum_{j\in I_{N}}(z^{p,j}_{t})^{2}.

If the event 𝒜c\mathcal{A}_{c} holds, then thanks to Ito’s Lemma it must be that

dut=\displaystyle du_{t}= {2τ1ut+1+N1jINztp,jGtp,j}dt+N1jINztp,jdWtp,j\displaystyle\big{\{}-2\tau^{-1}u_{t}+1+N^{-1}\sum_{j\in I_{N}}z^{p,j}_{t}G^{p,j}_{t}\big{\}}dt+N^{-1}\sum_{j\in I_{N}}z^{p,j}_{t}dW^{p,j}_{t} (84)
\displaystyle\leq {2τ1ut+1+cCλut}dt+N1jINztp,jdWtp,j,\displaystyle\big{\{}-2\tau^{-1}u_{t}+1+cC_{\lambda}u_{t}\big{\}}dt+N^{-1}\sum_{j\in I_{N}}z^{p,j}_{t}dW^{p,j}_{t}, (85)

since N1jINλ(ztp,j)2Cλ2utN^{-1}\sum_{j\in I_{N}}\lambda(z^{p,j}_{t})^{2}\leq C_{\lambda}^{2}u_{t}. Write

vt=sups[0,t]N1|jIN0tzsp,j𝑑Wsp,j|,\displaystyle v_{t}=\sup_{s\in[0,t]}N^{-1}\bigg{|}\sum_{j\in I_{N}}\int_{0}^{t}z^{p,j}_{s}dW^{p,j}_{s}\bigg{|}, (86)

and define the stopping time, for a constant A>0A>0,

τA=inf{t0:vtexp(At)+A}.\displaystyle\tau_{A}=\inf\big{\{}t\geq 0\;:\;v_{t}\geq\exp(At)+A\big{\}}. (87)

Gronwall’s Inequality implies that for all tτAt\leq\tau_{A},

ut(A+u0+t)exp(c~t)\displaystyle u_{t}\leq\big{(}A+u_{0}+t\big{)}\exp\big{(}\tilde{c}t\big{)}

where c~=A+cCλ2τ1\tilde{c}=A+cC_{\lambda}-2\tau^{-1}. The quadratic variation of x(t):=N1jIN0tzsp,j𝑑Wsp,jx(t):=N^{-1}\sum_{j\in I_{N}}\int_{0}^{t}z^{p,j}_{s}dW^{p,j}_{s} is

(QV)tN=N2jIN0t(zsp,j)2𝑑s.\displaystyle(QV)_{t}^{N}=N^{-2}\sum_{j\in I_{N}}\int_{0}^{t}(z^{p,j}_{s})^{2}ds. (88)

For all tτAt\leq\tau_{A},

(QV)tN\displaystyle(QV)_{t}^{N} N1c~1(A+u0+t)exp(c~t):=N1ht,\displaystyle\leq N^{-1}\tilde{c}^{-1}\big{(}A+u_{0}+t\big{)}\exp\big{(}\tilde{c}t\big{)}:=N^{-1}h_{t}, (89)

and notice that hth_{t} is independent of the Brownian Motions. Now define the stochastic process w(t)w(t) to be such that

w(t)=\displaystyle w(t)= x(αtN) where\displaystyle x\big{(}\alpha^{N}_{t}\big{)}\text{ where } (90)
αtN=\displaystyle\alpha^{N}_{t}= inf{s0:(QV)sN=t}\displaystyle\inf\big{\{}s\geq 0\;:\;(QV)_{s}^{N}=t\big{\}} (91)

Thanks to the time-rescaled representation of a stochastic integral, w(t)w(t) is a Brownian Motion [27]. Writing f(t)=exp(At)+Af(t)=\exp(At)+A, it follows that

( There exists\displaystyle\mathbb{P}\bigg{(}\text{ There exists } sT such that |x(s)|f(s))\displaystyle s\leq T\text{ such that }\big{|}x(s)\big{|}\geq f(s)\bigg{)}
\displaystyle\leq ( There exists sT such that |w(N1hs)|f(s))\displaystyle\mathbb{P}\bigg{(}\text{ There exists }s\leq T\text{ such that }\big{|}w(N^{-1}h_{s})\big{|}\geq f(s)\bigg{)}
\displaystyle\leq ( There exists sT such that |w(N1hs(m))|f(s(m)))\displaystyle\mathbb{P}\bigg{(}\text{ There exists }s\leq T\text{ such that }\big{|}w(N^{-1}h_{s^{(m)}})\big{|}\geq f(s_{(m)})\bigg{)}

and we have written

s(m)\displaystyle s^{(m)} =inf{ta(m):ta(m)s}\displaystyle=\inf\big{\{}t^{(m)}_{a}\;:t^{(m)}_{a}\geq s\big{\}} (92)
s(m)\displaystyle s_{(m)} =sup{ta(m):ta(m)s}.\displaystyle=\sup\big{\{}t^{(m)}_{a}\;:t^{(m)}_{a}\leq s\big{\}}. (93)

and we recall that ta(m)=Ta/mt^{(m)}_{a}=Ta/m. Employing a union-of-events bound,

( There exists sT such that |w(hs(m))|f(s(m)))a=0m1{(w(N1hta+1(m))f(ta(m)))+(w(N1hta+1(m))f(ta(m)))}\mathbb{P}\bigg{(}\text{ There exists }s\leq T\text{ such that }\big{|}w(h_{s^{(m)}})\big{|}\geq f(s_{(m)})\bigg{)}\leq\\ \sum_{a=0}^{m-1}\bigg{\{}\mathbb{P}\bigg{(}w\big{(}N^{-1}h_{t_{a+1}^{(m)}}\big{)}\geq f\big{(}t_{a}^{(m)}\big{)}\bigg{)}+\mathbb{P}\bigg{(}w\big{(}N^{-1}h_{t_{a+1}^{(m)}}\big{)}\leq-f\big{(}t_{a}^{(m)}\big{)}\bigg{)}\bigg{\}} (94)

Now since w(t)w(t) is centered and Gaussian, with variance of tt,

N1log(w(N1hta+1(m))f(ta(m)))=\displaystyle N^{-1}\log\mathbb{P}\bigg{(}w\big{(}N^{-1}h_{t_{a+1}^{(m)}}\big{)}\geq f\big{(}t_{a}^{(m)}\big{)}\bigg{)}= N2f(ta(m))2(hta+1(m))1+O(logN)\displaystyle-\frac{N}{2}f\big{(}t_{a}^{(m)}\big{)}^{2}\big{(}h_{t_{a+1}^{(m)}}\big{)}^{-1}+O\big{(}\log N\big{)} (95)
N1log(w(N1hta+1(m))f(ta(m)))=\displaystyle N^{-1}\log\mathbb{P}\bigg{(}w\big{(}N^{-1}h_{t_{a+1}^{(m)}}\big{)}\leq-f\big{(}t_{a}^{(m)}\big{)}\bigg{)}= N2f(ta(m))2(hta+1(m))1+O(logN).\displaystyle-\frac{N}{2}f\big{(}t_{a}^{(m)}\big{)}^{2}\big{(}h_{t_{a+1}^{(m)}}\big{)}^{-1}+O\big{(}\log N\big{)}. (96)

We fix m=Am=A, and take AA to be arbitrarily large. Then

limAinf0am1f(ta(m))2(hta+1(m))1=.\lim_{A\to\infty}\inf_{0\leq a\leq m-1}f\big{(}t_{a}^{(m)}\big{)}^{2}\big{(}h_{t_{a+1}^{(m)}}\big{)}^{-1}=\infty.

We thus find that, for large enough AA,

limN¯N1log(𝒜c, There exists sT such that |x(s)|f(s))L.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\bigg{(}\mathcal{A}_{c},\text{ There exists }s\leq T\text{ such that }\big{|}x(s)\big{|}\geq f(s)\bigg{)}\leq-L. (97)

We have already demonstrated in the course of the proof that if the event 𝒜c\mathcal{A}_{c} holds, and sups[0,T]|x(s)|f(s)\sup_{s\in[0,T]}|x(s)|\leq f(s), then there exists a constant such that supt[0,T]utConst\sup_{t\in[0,T]}u_{t}\leq\rm{Const}. We have thus established the Lemma. ∎

The following L2L^{2}-Wasserstein distance provides a very useful way of controlling the dependence of the fields (Gtν)(G^{\nu}_{t}) on the measure ν\nu. Define dt(2)(,)d^{(2)}_{t}(\cdot,\cdot) to be such that for any μ,ν𝒰\mu,\nu\in\mathcal{U},

dt(2)(μ,ν)=infζ𝔼ζ[pIM0t{(yspy~sp)2+(GspG~sp)2}𝑑s]1/2,\displaystyle d^{(2)}_{t}(\mu,\nu)=\inf_{\zeta}\mathbb{E}^{\zeta}\bigg{[}\sum_{p\in I_{M}}\int_{0}^{t}\big{\{}(y^{p}_{s}-\tilde{y}^{p}_{s})^{2}+(G^{p}_{s}-\tilde{G}^{p}_{s})^{2}\big{\}}ds\bigg{]}^{1/2}, (98)

where the infimum is over all ζ𝒫(𝒞([0,T],2M)×𝒞([0,T],2M)\zeta\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{2M})\times\mathcal{C}([0,T],\mathbb{R}^{2M}\big{)}, such that the law of the first 2M2M processes is given by μ\mu, and the law of the last 2M2M processes is given by ν\nu. Let d(2)(μ,ν):=dT(2)(μ,ν)d^{(2)}(\mu,\nu):=d^{(2)}_{T}(\mu,\nu).

Lemma 15.

For any a>0a>0, d(2)(,)d^{(2)}(\cdot,\cdot) metrizes weak convergence in 𝒰a\mathcal{U}_{a}. Furthermore,

limϵ0+sup{dW(μ,ν):μ,ν𝒰a and d(2)(μ,ν)ϵ}=0.\displaystyle\lim_{\epsilon\to 0^{+}}\sup\big{\{}d_{W}(\mu,\nu)\;:\mu,\nu\in\mathcal{U}_{a}\text{ and }d^{(2)}(\mu,\nu)\leq\epsilon\big{\}}=0. (99)
Proof.

Since 𝒰a\mathcal{U}_{a} is compact, Prokhorov’s Theorem implies that for any ϵ~>0\tilde{\epsilon}>0, there exists a compact set 𝒟ϵ𝒞([0,T],M)2\mathcal{D}_{\epsilon}\subset\mathcal{C}([0,T],\mathbb{R}^{M})^{2} such that for all μ𝒰a\mu\in\mathcal{U}_{a},

μ(𝒟ϵ)1ϵ~.\displaystyle\mu\big{(}\mathcal{D}_{\epsilon}\big{)}\geq 1-\tilde{\epsilon}. (100)

Since 𝒟ϵ\mathcal{D}_{\epsilon} is compact, it follows from the Arzela-Ascoli Theorem that for any δ>0\delta>0, there exists υ(ϵ,δ)\upsilon(\epsilon,\delta) such that for all f,g𝒟ϵf,g\in\mathcal{D}_{\epsilon} such that for all pI2Mp\in I_{2M},

0T(fp(t)gp(t))2𝑑tυ(ϵ,δ),\displaystyle\int_{0}^{T}(f^{p}(t)-g^{p}(t))^{2}dt\leq\upsilon(\epsilon,\delta), (101)

it necessarily holds that

suppIMsupt[0,T]|fp(t)gp(t)|δ.\displaystyle\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}f^{p}(t)-g^{p}(t)\big{|}\leq\delta. (102)

Let ζ\zeta be any measure that is within η1\eta\ll 1 of realizing the infimum in (98). Then, writing

𝒜ϵ=χ{ For each pIM,yp,y~p,gp,g~p𝒟ϵ},\mathcal{A}_{\epsilon}=\chi\big{\{}\text{ For each }p\in I_{M},\;\;y^{p},\tilde{y}^{p},g^{p},\tilde{g}^{p}\in\mathcal{D}_{\epsilon}\big{\}},

we have the bound

𝔼ζ[suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|]𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)𝒜ϵ]+𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)(1𝒜ϵ)]\mathbb{E}^{\zeta}\bigg{[}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{]}\\ \leq\mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\mathcal{A}_{\epsilon}\bigg{]}+\\ \mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\big{(}1-\mathcal{A}_{\epsilon}\big{)}\bigg{]} (103)

Now we take d(2)(μ,ν)0+d^{(2)}(\mu,\nu)\to 0^{+}, and η0+\eta\to 0^{+} too. Since 𝒜ϵ\mathcal{A}_{\epsilon} is closed, thanks to the Portmanteau Theorem, we thus find that for any ϵ>0\epsilon>0,

𝔼ζ[𝒜ϵpIM0T{(yspy~sp)2+(GspG~sp)2}𝑑s]0.\displaystyle\mathbb{E}^{\zeta}\bigg{[}\mathcal{A}_{\epsilon}\sum_{p\in I_{M}}\int_{0}^{T}\big{\{}(y^{p}_{s}-\tilde{y}^{p}_{s})^{2}+(G^{p}_{s}-\tilde{G}^{p}_{s})^{2}\big{\}}ds\bigg{]}\to 0. (104)

which in turn implies that (making use of the uniform convergence over 𝒟ϵ\mathcal{D}_{\epsilon} in (102))

𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)𝒜ϵ]0.\mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\mathcal{A}_{\epsilon}\bigg{]}\to 0. (105)

For the other term on the RHS of (103), for b>0b>0, write

b=χ{ For each pIM,supt[0,T]|ytp|b,supt[0,T]|y~tp|b,supt[0,T]|gtp|b,supt[0,T]|g~tp|b}\displaystyle\mathcal{B}_{b}=\chi\bigg{\{}\text{ For each }p\in I_{M},\;\sup_{t\in[0,T]}\big{|}y^{p}_{t}\big{|}\leq b\;,\;\sup_{t\in[0,T]}\big{|}\tilde{y}^{p}_{t}\big{|}\leq b\;,\;\sup_{t\in[0,T]}\big{|}g^{p}_{t}\big{|}\leq b\;,\;\sup_{t\in[0,T]}\big{|}\tilde{g}^{p}_{t}\big{|}\leq b\bigg{\}}

Then,

𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)(1𝒜ϵ)]𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)(1𝒜ϵ)b]+𝔼ζ[(suppIMsupt[0,T]|yp(t)y~p(t)|+suppIMsupt[0,T]|gp(t)g~p(t)|)(1𝒜ϵ)(1b)].\mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\big{(}1-\mathcal{A}_{\epsilon}\big{)}\bigg{]}\\ \leq\mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\big{(}1-\mathcal{A}_{\epsilon}\big{)}\mathcal{B}_{b}\bigg{]}\\ +\mathbb{E}^{\zeta}\bigg{[}\bigg{(}\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}y^{p}(t)-\tilde{y}^{p}(t)\big{|}+\sup_{p\in I_{M}}\sup_{t\in[0,T]}\big{|}g^{p}(t)-\tilde{g}^{p}(t)\big{|}\bigg{)}\big{(}1-\mathcal{A}_{\epsilon}\big{)}\big{(}1-\mathcal{B}_{b}\big{)}\bigg{]}. (106)

Thanks to the fact that, for all μ𝒰a\mu\in\mathcal{U}_{a},

suppIM𝔼μ[supt[0,T](ytp)2]a,\sup_{p\in I_{M}}\mathbb{E}^{\mu}\big{[}\sup_{t\in[0,T]}(y^{p}_{t})^{2}\big{]}\leq a,

one finds that the second term on the RHS of (106) goes to 0 as bb\to\infty, uniformly for all ϵ>0\epsilon>0 and all μ,ν𝒰a\mu,\nu\in\mathcal{U}_{a} . For any fixed b1b\gg 1, the first term on the RHS of (106) must go to zero as ϵ0+\epsilon\to 0^{+}, thanks to (100). We have thus proved the Lemma. ∎

For μ,𝔔\mu,\in\mathfrak{Q}, we define dt(2)(μ,ν)d^{(2)}_{t}(\mu,\nu) analogously to (98).

Lemma 16.

There exists a constant >0\mathfrak{C}>0 such that for all μ,ν𝔔\mu,\nu\in\mathfrak{Q} and all t[0,T]t\in[0,T],

dt(2)(βν,βμ)\displaystyle d^{(2)}_{t}(\beta_{\nu},\beta_{\mu}) dt(2)(ν,μ).\displaystyle\leq\mathfrak{C}d^{(2)}_{t}(\nu,\mu). (107)

Also for all μ,ν𝔔\mu,\nu\in\mathfrak{Q} such that for some b>0b>0, det(𝔙μ,0),det(𝔙ν,0)b>0\det(\mathfrak{V}_{\mu,0}),\det(\mathfrak{V}_{\nu,0})\geq b>0, there exists a constant b\mathfrak{C}_{b} such that

dt(2)(βν,𝐠,βμ,𝐠)\displaystyle d^{(2)}_{t}\big{(}\beta_{\nu,\mathbf{g}},\beta_{\mu,\mathbf{g}}\big{)} ~b(1+𝐠)dt(2)(ν,μ),\displaystyle\leq\tilde{\mathfrak{C}}_{b}(1+\left\|\mathbf{g}\right\|)d^{(2)}_{t}(\nu,\mu), (108)

and \left\|\cdot\right\| is the Euclidean norm on M\mathbb{R}^{M}.

Proof.

We start by proving (107). Recalling the definition of the distance dt(2)d^{(2)}_{t} in (98), let ζϵ\zeta_{\epsilon} be such that

dt(2)(μ,ν)2𝔼ζϵ[pIM0t(zspysp)2𝑑s]+ϵ.\displaystyle d^{(2)}_{t}(\mu,\nu)^{2}\geq\mathbb{E}^{\zeta_{\epsilon}}\bigg{[}\sum_{p\in I_{M}}\int_{0}^{t}(z^{p}_{s}-y^{p}_{s})^{2}ds\bigg{]}+\epsilon. (109)

Furthermore define centered Gaussian processes Gs(ϵ),μ,p,Gs(ϵ),ν,pG^{(\epsilon),\mu,p}_{s},G^{(\epsilon),\nu,p}_{s} to be such that for any p,qIMp,q\in I_{M} and any s,t[0,T]s,t\in[0,T],

𝔼[Gs(ϵ),μ,pGt(ϵ),ν,q]\displaystyle\mathbb{E}\big{[}G^{(\epsilon),\mu,p}_{s}G^{(\epsilon),\nu,q}_{t}\big{]} =𝔼ζϵ[λ(zsp)λ(ytq)]\displaystyle=\mathbb{E}^{\zeta_{\epsilon}}\big{[}\lambda(z^{p}_{s})\lambda(y^{q}_{t})\big{]} (110)
𝔼[Gs(ϵ),μ,pGt(ϵ),μ,q]\displaystyle\mathbb{E}\big{[}G^{(\epsilon),\mu,p}_{s}G^{(\epsilon),\mu,q}_{t}\big{]} =𝔼ζϵ[λ(zsp)λ(ztq)]\displaystyle=\mathbb{E}^{\zeta_{\epsilon}}\big{[}\lambda(z^{p}_{s})\lambda(z^{q}_{t})\big{]} (111)
𝔼[Gs(ϵ),ν,pGt(ϵ),ν,q]\displaystyle\mathbb{E}\big{[}G^{(\epsilon),\nu,p}_{s}G^{(\epsilon),\nu,q}_{t}\big{]} =𝔼ζϵ[λ(ysp)λ(ytq)].\displaystyle=\mathbb{E}^{\zeta_{\epsilon}}\big{[}\lambda(y^{p}_{s})\lambda(y^{q}_{t})\big{]}. (112)

This definition is possible thanks to a trivial modification of Lemma 5 (switching M2MM\to 2M). We thus find that

limϵ0+𝔼[pIM0r(Gt(ϵ),μ,pGt(ϵ),ν,p)2𝑑t]\displaystyle\lim_{\epsilon\to 0^{+}}\mathbb{E}\bigg{[}\sum_{p\in I_{M}}\int_{0}^{r}(G^{(\epsilon),\mu,p}_{t}-G^{(\epsilon),\nu,p}_{t})^{2}dt\bigg{]}\leq limϵ0+𝔼ζϵ[pIM0r{λ(ztp)λ(ytp)}2𝑑t]\displaystyle\lim_{\epsilon\to 0^{+}}\mathbb{E}^{\zeta_{\epsilon}}\bigg{[}\sum_{p\in I_{M}}\int_{0}^{r}\big{\{}\lambda(z^{p}_{t})-\lambda(y^{p}_{t})\big{\}}^{2}dt\bigg{]} (113)
\displaystyle\leq limϵ0+Cλ2𝔼ζϵ[pIM0r{ztpytp}2𝑑t]\displaystyle\lim_{\epsilon\to 0^{+}}C_{\lambda}^{2}\mathbb{E}^{\zeta_{\epsilon}}\bigg{[}\sum_{p\in I_{M}}\int_{0}^{r}\big{\{}z^{p}_{t}-y^{p}_{t}\big{\}}^{2}dt\bigg{]} (114)
=\displaystyle= Cλ2dr(2)(μ,ν).\displaystyle C_{\lambda}^{2}d_{r}^{(2)}(\mu,\nu). (115)

Now as ϵ0+\epsilon\to 0^{+}, the LHS of (113) must decrease to d(2)(βμ,βν)d^{(2)}(\beta_{\mu},\beta_{\nu}). (108) follows analogously.

The proof of (108) is analogous, since the mean and variance functions in (45) and (46) are such that for all μ,ν𝒰\mu,\nu\in\mathcal{U}, there is a constant Cb>0C_{b}>0 such that

supp,qIMs.t[0,T]|𝔚stμ,pq𝔚stν,pq|\displaystyle\sup_{p,q\in I_{M}\fatsemi s.t\in[0,T]}\big{|}\mathfrak{W}^{\mu,pq}_{st}-\mathfrak{W}^{\nu,pq}_{st}\big{|} Cblimϵ0+suprIMs[0,T]𝔼ζϵ[(λ(zsr)λ(ysr))2]\displaystyle\leq C_{b}\lim_{\epsilon\to 0^{+}}\sup_{r\in I_{M}\fatsemi s\in[0,T]}\mathbb{E}^{\zeta_{\epsilon}}\big{[}\big{(}\lambda(z^{r}_{s})-\lambda(y^{r}_{s})\big{)}^{2}\big{]}
supp,qIMs.t[0,T]|𝔪sp(μ,𝐠)𝔪sp(ν,𝐠)|\displaystyle\sup_{p,q\in I_{M}\fatsemi s.t\in[0,T]}\big{|}\mathfrak{m}^{p}_{s}(\mu,\mathbf{g})-\mathfrak{m}^{p}_{s}(\nu,\mathbf{g})\big{|} Cb(1+𝐠)limϵ0+suprIMs[0,T]𝔼ζϵ[(λ(zsr)λ(ysr))2]\displaystyle\leq C_{b}\big{(}1+\|\mathbf{g}\|\big{)}\lim_{\epsilon\to 0^{+}}\sup_{r\in I_{M}\fatsemi s\in[0,T]}\mathbb{E}^{\zeta_{\epsilon}}\big{[}\big{(}\lambda(z^{r}_{s})-\lambda(y^{r}_{s})\big{)}^{2}\big{]}

We have also employed the fact that det(𝔙μ,0)\det(\mathfrak{V}_{\mu,0}) is uniformly lower-bounded by a positive constant bb (as noted in the statement of the Lemma). ∎

4.2 Large Deviations of the Uncoupled System

Our first aim is to prove a Large Deviation Principle in the case of fields with a frozen interaction structure (in Lemma 17 below). This would ordinarily be a trivial application of Sanov’s Theorem. However the proof is slightly complicated by the need for the LDP to be uniform with respect to the variables (𝐳0,𝐠0)MN×MN(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathbb{R}^{MN}\times\mathbb{R}^{MN} that the probability laws are conditioned on.

For any 𝐠0MN\mathbf{g}_{0}\in\mathbb{R}^{MN} and ν𝒰\nu\in\mathcal{U}, define γ~ν,𝐠0N:=j=1Nβν,𝐠0j𝒫(𝒞([0,T],M)N)\tilde{\gamma}^{N}_{\nu,\mathbf{g}_{0}}:=\otimes_{j=1}^{N}\beta_{\nu,\mathbf{g}_{0}^{j}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)}. In other words, γ~ν,𝐠0N\tilde{\gamma}^{N}_{\nu,\mathbf{g}_{0}} is the law of NN independent 𝒞([0,T],M)\mathcal{C}([0,T],\mathbb{R}^{M})-valued Gaussian variables {G~ν,j}jIN\{\tilde{G}^{\nu,j}\}_{j\in I_{N}}. The mean and variance of these variables is specified in (43) and (44).

Let Qν,𝐳0,𝐠0N𝒫(𝒞([0,T],M)N×𝒞([0,T],M)N)Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\times\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} be the joint law of the uncoupled system, i.e.

Qν,𝐳0,𝐠0N=γν,𝐳0,𝐠0NP𝐳0N.Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}=\gamma^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\otimes P^{N}_{\mathbf{z}_{0}}. (116)

For ν𝒰\nu\in\mathcal{U}, define I~ν:𝒫(𝒞([0,T],M)2)\tilde{I}_{\nu}:\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\to\mathbb{R} as follows. We specify that I~ν(μ)=\tilde{I}_{\nu}(\mu)=\infty if either the marginal of μ\mu at time 0 is not equal to κ\kappa, and / or μ𝒰\mu\notin\mathcal{U}. Otherwise, for any ζ𝒰\zeta\in\mathcal{U}, writing ζ𝐳0,𝐠0\zeta_{\mathbf{z}_{0},\mathbf{g}_{0}} to be the law of ζ\zeta, conditioned on the values of its variables at time 0, define

I~ν(ζ)=𝔼κ[(ζ𝐳0,𝐠0||P𝐳0γ~ν,𝐠0)].\displaystyle\tilde{I}_{\nu}(\zeta)=\mathbb{E}^{\kappa}\big{[}\mathcal{R}\big{(}\zeta_{\mathbf{z}_{0},\mathbf{g}_{0}}||P_{\mathbf{z}_{0}}\otimes\tilde{\gamma}_{\nu,\mathbf{g}_{0}}\big{)}\big{]}. (117)

Define the empirical measure μ~N𝒫(𝒞([0,T],)M×𝒞([0,T],)M)\tilde{\mu}^{N}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R})^{M}\times\mathcal{C}([0,T],\mathbb{R})^{M}\big{)} to be

μ~N=N1jINδ𝐲[0,T]j,G~[0,T]ν,j,\displaystyle\tilde{\mu}^{N}=N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{y}^{j}_{[0,T]},\tilde{G}^{\nu,j}_{[0,T]}}, (118)

where we recall that

ytp=z0p+0tσs𝑑Wsp.\displaystyle y^{p}_{t}=z^{p}_{0}+\int_{0}^{t}\sigma_{s}dW_{s}^{p}. (119)
Lemma 17.

Fix some ν𝒰\nu\in\mathcal{U}. Let 𝒜,𝒪𝒫(𝒞([0,T],M)2))\mathcal{A},\mathcal{O}\subseteq\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\big{)}, such that 𝒪\mathcal{O} is open and 𝒜\mathcal{A} closed. Then

limN¯sup(𝐳0,𝐠0)𝒴NN1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜)\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)} infμ𝒜I~ν(μ)\displaystyle\leq-\inf_{\mu\in\mathcal{A}}\tilde{I}_{\nu}(\mu) (120)
lim¯Ninf(𝐳0,𝐠0)𝒴NN1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒪)\displaystyle\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{O}\big{)} infμ𝒪I~ν(μ).\displaystyle\geq-\inf_{\mu\in\mathcal{O}}\tilde{I}_{\nu}(\mu). (121)

Furthermore I~ν()\tilde{I}_{\nu}(\cdot) is lower semi-continuous, and has compact level sets.

Proof.

First, fix any sequence (𝐳0(N),𝐠0(N))N1\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}\big{)}_{N\geq 1}, such that (𝐳0(N),𝐠0(N))𝒴N\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}\big{)}\in\mathcal{Y}^{N}. Necessarily, thanks to the definition of 𝒴N\mathcal{Y}^{N}, it must be that

μ^N(𝐳0(N),𝐠0(N))κ𝒫(2M).\displaystyle\hat{\mu}^{N}\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}\big{)}\to\kappa\in\mathcal{P}\big{(}\mathbb{R}^{2M}\big{)}. (122)

It follows from (122) that

limN¯N1logQν,𝐳0(N),𝐠0(N)N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜)\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)} infμ𝒜I~ν(μ)\displaystyle\leq-\inf_{\mu\in\mathcal{A}}\tilde{I}_{\nu}(\mu) (123)
lim¯NN1logQν,𝐳0(N),𝐠0(N)N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒪)\displaystyle\underset{N\to\infty}{\underline{\lim}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{O}\big{)} infμ𝒪I~ν(μ).\displaystyle\geq-\inf_{\mu\in\mathcal{O}}\tilde{I}_{\nu}(\mu). (124)

See for instance [30] for a proof of this fact. Furthermore I~ν:𝒫(𝒞([0,T],M)2)+\tilde{I}_{\nu}:\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\to\mathbb{R}^{+} is lower-semi-continuous and has compact level sets.

We next have to show that the convergence is uniform over 𝒴N\mathcal{Y}^{N} (as in the statement of the Theorem). To do this, we first wish to show that for any measurable set 𝒫(𝒞([0,T],M)2)\mathcal{E}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)} and any ϵ>0\epsilon>0, for all large enough NN,

sup(𝐳0,𝐠0)𝒴NQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν))inf(𝐳0,𝐠0)𝒴NQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)(ϵ))\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{E}\big{)}\leq\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{E}^{(\epsilon)}\big{)} (125)

and (ϵ)\mathcal{E}^{(\epsilon)} is the closed ϵ\epsilon-blowup of \mathcal{E} with respect to dWd_{W}. To do this, we are going to compare the conditioned probability to the conditioned probability induced by any other sequence in 𝒴N\mathcal{Y}^{N}. This comparison is facilitated by using the following permutation-averaged probability law.

Define the set

𝔖N={(𝐲,𝐠)𝒞([0,T],M)N×𝒞([0,T],M)N:N1jIN{suppIM|y0p,jz0(N),p,j|+suppIM|g0p,jg0(N),p,j|}2δ~N},\mathfrak{S}^{N}=\bigg{\{}(\mathbf{y},\mathbf{g})\in\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N}\times\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N}:\\ N^{-1}\sum_{j\in I_{N}}\big{\{}\sup_{p\in I_{M}}\big{|}y^{p,j}_{0}-z^{(N),p,j}_{0}\big{|}+\sup_{p\in I_{M}}\big{|}g^{p,j}_{0}-g^{(N),p,j}_{0}\big{|}\big{\}}\leq 2\tilde{\delta}_{N}\bigg{\}}, (126)

and we recall that (δ~N)N1(\tilde{\delta}_{N})_{N\geq 1} is a sequence that decreases to 0, as defined in (30). We endow 𝔖N\mathfrak{S}^{N} with the topology that it inherits from 𝒞([0,T],M)N×𝒞([0,T],M)N\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N}\times\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N}. Write 𝔓N\mathfrak{P}^{N} to be the set of all permutations on INI_{N}, and define the measure Q¯ν,𝐳0,𝐠0N𝒫(𝔖N)\bar{Q}^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\in\mathcal{P}\big{(}\mathfrak{S}^{N}\big{)} to be the average over all permutations, i.e. for any measurable 𝒜𝔖N\mathcal{A}\subseteq\mathfrak{S}^{N},

Q¯ν,𝐳0,𝐠0N(𝒜)=|𝔓N|1π𝔓NQν,𝐳0,𝐠0N(π𝒜)\displaystyle\bar{Q}^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}(\mathcal{A})=\big{|}\mathfrak{P}^{N}\big{|}^{-1}\sum_{\pi\in\mathfrak{P}^{N}}Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\pi\cdot\mathcal{A}\big{)} (127)

and here we denote π:𝒞([0,T],M)N×𝒞([0,T],M)N\pi:\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N}\times\mathcal{C}\big{(}[0,T],\mathbb{R}^{M}\big{)}^{N} to be the permutation,

(π(𝐲,𝐠))j=(𝐲π(j),𝐠π(j)).\displaystyle\big{(}\pi\cdot(\mathbf{y},\mathbf{g})\big{)}^{j}=(\mathbf{y}^{\pi(j)},\mathbf{g}^{\pi(j)}). (128)

Since the empirical measure is invariant under any permutation of its arguments, for any measurable 𝒜𝒫(𝒞([0,T],M)2)\mathcal{A}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}

Qν,𝐳0(N),𝐠0(N)N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜)=Q¯ν,𝐳0(N),𝐠0(N)N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜).\displaystyle Q^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)}=\bar{Q}^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)}. (129)

We can without loss of generality take (𝐳0(N),𝐠0(N))𝔖N\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}\big{)}\in\mathfrak{S}^{N}. Now consider any other sequence (𝐳`0(N),𝐠`0(N))𝔖N𝒴N\big{(}\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0}\big{)}\in\mathfrak{S}^{N}\cap\mathcal{Y}^{N}. Let dWd_{W} be the 11-Wasserstein Distance on 𝒫(𝒞([0,T],M)N×𝒞([0,T],M)N)\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\times\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} induced by the norm

(𝐲,𝐆)(𝐲¯,𝐆¯)=N1jIN{suppIMt[0,T]|ytp,jy¯tp,j|+suppIMt[0,T]|Gtp,jG¯tp,j|}.\displaystyle\left\|(\mathbf{y},\mathbf{G})-(\bar{\mathbf{y}},\bar{\mathbf{G}})\right\|=N^{-1}\sum_{j\in I_{N}}\big{\{}\sup_{p\in I_{M}\fatsemi t\in[0,T]}\big{|}y^{p,j}_{t}-\bar{y}^{p,j}_{t}\big{|}+\sup_{p\in I_{M}\fatsemi t\in[0,T]}\big{|}G^{p,j}_{t}-\bar{G}^{p,j}_{t}\big{|}\big{\}}. (130)

We claim that

dW(Q¯ν,𝐳0(N),𝐠0(N)N,Q¯ν,𝐳`0(N),𝐠`0(N)N)\displaystyle d_{W}\big{(}\bar{Q}^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}},\bar{Q}^{N}_{\nu,\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0}}\big{)}\leq N1jIN{suppIM|z0(N),p,jz`0(N),p,j|\displaystyle N^{-1}\sum_{j\in I_{N}}\big{\{}\sup_{p\in I_{M}}\big{|}z^{(N),p,j}_{0}-\grave{z}^{(N),p,j}_{0}\big{|}
+suppIMt[0,T]|𝔪tp(ν,𝐠0(N),j)𝔪tp(ν,𝐠`0(N),j)|}\displaystyle+\sup_{p\in I_{M}\fatsemi t\in[0,T]}\big{|}\mathfrak{m}^{p}_{t}(\nu,\mathbf{g}^{(N),j}_{0})-\mathfrak{m}^{p}_{t}(\nu,\grave{\mathbf{g}}^{(N),j}_{0})\big{|}\big{\}} (131)
:=\displaystyle:= fN(𝐳0(N),𝐠0(N),𝐳`0(N),𝐠`0(N))\displaystyle f^{N}\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0},\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0}\big{)} (132)

This identity follows from the fact that z~tp,j:=ztp,jz0p,j\tilde{z}^{p,j}_{t}:=z^{p,j}_{t}-z^{p,j}_{0} and g~tp,j:=Gtp,j𝔪tp(ν,𝐆0j)\tilde{g}^{p,j}_{t}:=G^{p,j}_{t}-\mathfrak{m}^{p}_{t}(\nu,\mathbf{G}^{j}_{0}) are identically distributed, both (i) for all jINj\in I_{N}, and (ii) with respect to both probability laws Q¯ν,𝐳0(N),𝐠0(N)N\bar{Q}^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}} and Q¯ν,𝐳`0(N),𝐠`0(N)N\bar{Q}^{N}_{\nu,\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0}}.

It follows from the definition of 𝒴N\mathcal{Y}^{N} that

limNsup(𝐳0(N),𝐠0(N)),(𝐳`0(N),𝐠`0(N))𝔖N𝒴NfN(𝐳0(N),𝐠0(N),𝐳`0(N),𝐠`0(N))=0.\displaystyle\lim_{N\to\infty}\sup_{(\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}),(\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0})\in\mathfrak{S}^{N}\cap\mathcal{Y}^{N}}f^{N}\big{(}\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0},\grave{\mathbf{z}}^{(N)}_{0},\grave{\mathbf{g}}^{(N)}_{0}\big{)}=0. (133)

We have thus proved (125).

To now must prove the Large Deviations bounds in the statement of the theorem. We start with the upper-bound (120). It follows from (123) and (125) that for any ϵ>0\epsilon>0,

limN¯sup(𝐳0,𝐠0)𝒴NN1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜)infμ𝒜(ϵ)I~ν(μ).\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)}\leq-\inf_{\mu\in\mathcal{A}^{(\epsilon)}}\tilde{I}_{\nu}(\mu). (134)

The lower-semi-continuity of I~ν(μ)\tilde{I}_{\nu}(\mu) dictates that

limϵ0+infμ𝒜(ϵ)I~ν(μ)=infμ𝒜I~ν(μ),\lim_{\epsilon\to 0^{+}}\inf_{\mu\in\mathcal{A}^{(\epsilon)}}\tilde{I}_{\nu}(\mu)=\inf_{\mu\in\mathcal{A}}\tilde{I}_{\nu}(\mu),

and we have proved the upperbound. For the lower-bound, let 𝒪\mathcal{O} be open, and for any μ𝒪\mu\in\mathcal{O}, take ϵ>0\epsilon>0 to be such that B2ϵ(μ)𝒪B_{2\epsilon}(\mu)\subset\mathcal{O}. Then

lim¯Ninf(𝐳0,𝐠0)𝒴N\displaystyle\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}} N1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒪)\displaystyle N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{O}\big{)} (135)
lim¯Ninf(𝐳0,𝐠0)𝒴NN1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)B2ϵ(μ))\displaystyle\geq\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in B_{2\epsilon}(\mu)\big{)} (136)
lim¯NN1logQν,𝐳0(N),𝐠0(N)N(μ~N(𝐲[0,T],𝐆~[0,T]ν)Bϵ(μ))\displaystyle\geq\underset{N\to\infty}{\underline{\lim}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}^{(N)}_{0},\mathbf{g}^{(N)}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in B_{\epsilon}(\mu)\big{)} (137)
infζBϵ(μ)I~ν(ζ),\displaystyle\geq-\inf_{\zeta\in B_{\epsilon}(\mu)}\tilde{I}_{\nu}(\zeta), (138)

using the Large Deviations estimate (124). Taking ϵ0+\epsilon\to 0^{+}, it must be that for any μ𝒪\mu\in\mathcal{O},

lim¯Ninf(𝐳0,𝐠0)𝒴NN1logQν,𝐳0,𝐠0N(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒪)I~ν(μ).\displaystyle\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{O}\big{)}\geq-\tilde{I}_{\nu}(\mu). (139)

Since μ𝒪\mu\in\mathcal{O} is arbitrary, (121) follows immediately. ∎

We can now state the proof of Theorem 3.

Proof.

We start with the upper bound (35). We write μ^N:=μ^N(𝐲[0,T],𝐆~[0,T])\hat{\mu}^{N}:=\hat{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}_{[0,T]}). Using a union-of-events bound, for any a>0a>0,

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜)max{limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜𝒰a),limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒰a)}max{limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜𝒰a),L},\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\big{)}\leq\\ \max\bigg{\{}\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\cap\mathcal{U}_{a}\big{)},\\ \underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\notin\mathcal{U}_{a}\big{)}\bigg{\}}\\ \leq\max\bigg{\{}\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\cap\mathcal{U}_{a}\big{)},-L\bigg{\}}, (140)

for any L>0L>0, as long as aa is sufficiently large, thanks to the exponential tightness proved in Lemma 8. By taking aa\to\infty, it thus suffices that we prove that for arbitrary 𝒰a\mathcal{U}_{a} such that 𝒜𝒰a\mathcal{A}\cap\mathcal{U}_{a}\neq\emptyset,

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜𝒰a)=infμ𝒜𝒰aI~μ(μ).\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\cap\mathcal{U}_{a}\big{)}=-\inf_{\mu\in\mathcal{A}\cap\mathcal{U}_{a}}\tilde{I}_{\mu}(\mu). (141)

Since 𝒜𝒰a\mathcal{A}\cap\mathcal{U}_{a} is compact, for any ϵ>0\epsilon>0 we can always find an open covering of the form, for some positive integer 𝒩ϵ\mathcal{N}_{\epsilon}, {ζi}1i𝒩ϵ𝒜𝒰a\{\zeta_{i}\}_{1\leq i\leq\mathcal{N}_{\epsilon}}\subseteq\mathcal{A}\cap\mathcal{U}_{a},

𝒜𝒰ai=1𝒩ϵBϵ(ζi).\displaystyle\mathcal{A}\cap\mathcal{U}_{a}\subseteq\bigcup_{i=1}^{\mathcal{N}_{\epsilon}}B_{\epsilon}(\zeta_{i}). (142)

We thus find that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜𝒰a)sup1i𝒩ϵ{limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^NBϵ(ζi))}.\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\cap\mathcal{U}_{a}\big{)}\\ \leq\sup_{1\leq i\leq\mathcal{N}_{\epsilon}}\bigg{\{}\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in B_{\epsilon}(\zeta_{i})\big{)}\bigg{\}}. (143)

Thus, employing Lemma 5 in the third line below,

limN¯sup(𝐳0,𝐠0)𝒴N\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}} N1logQ𝐳0,𝐠0N(μ^NBϵ(ζi))\displaystyle N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in B_{\epsilon}(\zeta_{i})\big{)} (144)
=limN¯sup(𝐳0,𝐠0)𝒴NN1log𝔼P𝐳0N[γ𝐲,𝐠0N(μ^NBϵ(ζi))]\displaystyle=\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\gamma^{N}_{\mathbf{y},\mathbf{g}_{0}}\bigg{(}\hat{\mu}^{N}\in B_{\epsilon}(\zeta_{i})\bigg{)}\bigg{]} (145)
=limN¯sup(𝐳0,𝐠0)𝒴NN1log𝔼P𝐳0N[γ~μ^N(𝐲),𝐠0N(μ^NBϵ(ζi))]\displaystyle=\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\tilde{\gamma}^{N}_{\hat{\mu}^{N}(\mathbf{y}),\mathbf{g}_{0}}\bigg{(}\hat{\mu}^{N}\in B_{\epsilon}(\zeta_{i})\bigg{)}\bigg{]} (146)
limN¯sup(𝐳0,𝐠0)𝒴NN1log𝔼P𝐳0N[supνBϵ(ζi)γν(1),𝐠0N(μ^NBϵ(ζi))]\displaystyle\leq\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\sup_{\nu\in B_{\epsilon}(\zeta_{i})}\gamma^{N}_{\nu^{(1)},\mathbf{g}_{0}}\bigg{(}\hat{\mu}^{N}\in B_{\epsilon}(\zeta_{i})\bigg{)}\bigg{]} (147)
=infμ,νBϵ(ζi)I~ν(μ),\displaystyle=-\inf_{\mu,\nu\in B_{\epsilon}(\zeta_{i})}\tilde{I}_{\nu}(\mu), (148)

thanks to Theorem 17. We thus find that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒜𝒰a)\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{A}\cap\mathcal{U}_{a}\big{)} inf1i𝒩ϵinfν,μBϵ(ζi)I~ν(μ).\displaystyle\leq-\inf_{1\leq i\leq\mathcal{N}_{\epsilon}}\inf_{\nu,\mu\in B_{\epsilon}(\zeta_{i})}\tilde{I}_{\nu}(\mu). (149)

Now, it is proved in Lemma 16 that νQν,𝐳0,𝐠0\nu\to Q_{\nu,\mathbf{z}_{0},\mathbf{g}_{0}} is continuous. Since the Relative Entropy is lower-semi-continuous in both of its arguments, we thus find that the following map is lower-semi-continuous,

(ν,μ)I~ν(μ).(\nu,\mu)\to\tilde{I}_{\nu}(\mu).

Thus taking ϵ0+\epsilon\to 0^{+}, we obtain that

limϵ0+inf1i𝒩ϵinfν,μBϵ(ζi)I~ν(μ)=infμ𝒜𝒰aI~μ(μ),\displaystyle\lim_{\epsilon\to 0^{+}}\inf_{1\leq i\leq\mathcal{N}_{\epsilon}}\inf_{\nu,\mu\in B_{\epsilon}(\zeta_{i})}\tilde{I}_{\nu}(\mu)=\inf_{\mu\in\mathcal{A}\cap\mathcal{U}_{a}}\tilde{I}_{\mu}(\mu), (150)

and we have proved (141).

Turning to the lower bound (36), consider an arbitrary open set 𝒪\mathcal{O}. If 𝒪𝒰=\mathcal{O}\cap\mathcal{U}=\emptyset, then

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒪)==infμ𝒪(μ),\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{O}\big{)}=-\infty=-\inf_{\mu\in\mathcal{O}}\mathcal{I}(\mu),

since \mathcal{I} is identically \infty outside of 𝒰\mathcal{U}. In this case, its clear that (121) holds.

We can thus assume that 𝒪𝒰\mathcal{O}\cap\mathcal{U}\neq\emptyset. Let μ𝒪\mu\in\mathcal{O} be such that μ\mu is in the interior of 𝒰a\mathcal{U}_{a}, for some a>0a>0. We can thus find a sequence of neighborhoods {𝒩i}i1\{\mathcal{N}_{i}\}_{i\geq 1} of μ\mu such that 𝒩j𝒪𝒰aBj1(μ)\mathcal{N}_{j}\subseteq\mathcal{O}\cap\mathcal{U}_{a}\cap B_{j^{-1}}(\mu). We thus find that for any j1j\geq 1,

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒪)limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒩j).\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{O}\big{)}\geq\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{N}_{j}\big{)}. (151)

Similarly to the bound for the closed sets, we obtain that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒩j)supν𝒩jinfμ𝒩jI~ν(μ).\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{N}_{j}\big{)}\geq-\sup_{\nu\in\mathcal{N}_{j}}\inf_{\mu\in\mathcal{N}_{j}}\tilde{I}_{\nu}(\mu). (152)

Taking jj\to\infty, since (ν,μ)I~ν(μ)(\nu,\mu)\to\tilde{I}_{\nu}(\mu) is lower semicontinuous, it must be that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒩j)I~μ(μ).\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{N}_{j}\big{)}\geq-\tilde{I}_{\mu}(\mu). (153)

Since μ𝒪\mu\in\mathcal{O} is arbitrary, it must be that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N𝒪)infμ𝒪I~μ(μ).\displaystyle\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}\in\mathcal{O}\big{)}\geq-\inf_{\mu\in\mathcal{O}}\tilde{I}_{\mu}(\mu). (154)

4.2.1 Uncoupled System (with no conditioning)

In this subsection we prove Corollary 11.

For some ν𝔔\nu\in\mathfrak{Q}, let QνN𝒫(𝒞([0,T],M)N×𝒞([0,T],M)N)Q^{N}_{\nu}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\times\mathcal{C}([0,T],\mathbb{R}^{M})^{N}\big{)} be the joint law of the uncoupled system (with no conditioning), i.e.

QνN=(βνP𝐳)N.Q^{N}_{\nu}=\big{(}\beta_{\nu}\otimes P_{\mathbf{z}}\big{)}^{\otimes N}. (155)

We reach a corollary to Lemma 17.

Corollary 18.

Fix some ν𝒰\nu\in\mathcal{U}. Let 𝒜,𝒪𝒫(𝒞([0,T],M)2))\mathcal{A},\mathcal{O}\subseteq\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})^{2}\big{)}\big{)}, such that 𝒪\mathcal{O} is open and 𝒜\mathcal{A} closed. Then

limN¯N1logQνN(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒜)\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log Q^{N}_{\nu}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{A}\big{)} infμ𝒜(μ||Sν)\displaystyle\leq-\inf_{\mu\in\mathcal{A}}\mathcal{R}(\mu||S_{\nu}) (156)
lim¯NN1logQνN(μ~N(𝐲[0,T],𝐆~[0,T]ν)𝒪)\displaystyle\underset{N\to\infty}{\underline{\lim}}N^{-1}\log Q^{N}_{\nu}\big{(}\tilde{\mu}^{N}(\mathbf{y}_{[0,T]},\tilde{\mathbf{G}}^{\nu}_{[0,T]})\in\mathcal{O}\big{)} infμ𝒪(μ||Sν).\displaystyle\geq-\inf_{\mu\in\mathcal{O}}\mathcal{R}(\mu||S_{\nu}). (157)

Furthermore μ(μ||Sν)\mu\to\mathcal{R}(\mu||S_{\nu}) is lower semi-continuous, and has compact level sets.

Proof.

This is a consequence of Sanov’s Theorem. ∎

The proof of Corollary 11 now follows analogously to the proof of 3.

4.3 Coupled System

Girsanov’s Theorem implies that

dP𝐉,𝐳0NdP𝐳0N|T(𝐲)=exp(NΓ𝐉,TN(𝐲))\displaystyle\frac{dP^{N}_{\mathbf{J},\mathbf{z}_{0}}}{dP^{N}_{\mathbf{z}_{0}}}\bigg{|}_{\mathcal{F}_{T}}(\mathbf{y})=\exp\big{(}N\Gamma^{N}_{\mathbf{J},T}(\mathbf{y})\big{)} (158)

where Γ𝐉,TN:MN\Gamma^{N}_{\mathbf{J},T}:\mathbb{R}^{MN}\to\mathbb{R} is

Γ𝐉,TN(𝐲)=N1jINpIM0Tσs2(G~sp,jτ1ysp,j)𝑑ysp,j12σs2(G~sp,jτ1ysp,j)2ds.\Gamma^{N}_{\mathbf{J},T}(\mathbf{y})=N^{-1}\sum_{j\in I_{N}\fatsemi p\in I_{M}}\int_{0}^{T}\sigma_{s}^{-2}\big{(}\tilde{G}^{p,j}_{s}-\tau^{-1}y^{p,j}_{s}\big{)}dy^{p,j}_{s}-\frac{1}{2}\sigma_{s}^{-2}\big{(}\tilde{G}^{p,j}_{s}-\tau^{-1}y^{p,j}_{s}\big{)}^{2}ds. (159)

We wish to specify a map Γ:𝒰\Gamma:\mathcal{U}\to\mathbb{R} with (i) as nice regularity properties as possible, and (ii) such that with unit probability

Γ𝐉,TN(𝐲)=Γ(μ^N(𝐲,𝐆~)).\displaystyle\Gamma^{N}_{\mathbf{J},T}(\mathbf{y})=\Gamma\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}. (160)

It is well-known that the stochastic integral is not a continuous function of the driving Brownian motion. Thus we define the map Γ\Gamma to be a limit of time-discretized approximations, and we will show that this limit must always converge for any measure in 𝒰\mathcal{U}.

Our precise definition of Γ:𝒰\Gamma:\mathcal{U}\to\mathbb{R} is as follows. We first define a time-discretized approximation to Γ\Gamma. Γ(m):𝒰+\Gamma^{(m)}:\mathcal{U}\to\mathbb{R}^{+},

Γ(m)(μ)=pIMa=0m1𝔼μ[σta(m)2(Gta(m)pτ1zta(m)p)(zta+1(m)pzta(m)p+Δmτ1zta(m)p)12σta(m)2Δm(Gta(m)pτ1zta(m)p)2].\Gamma^{(m)}(\mu)=\sum_{p\in I_{M}}\sum_{a=0}^{m-1}\mathbb{E}^{\mu}\bigg{[}\sigma_{t^{(m)}_{a}}^{-2}\big{(}G^{p}_{t_{a}^{(m)}}-\tau^{-1}z^{p}_{t_{a}^{(m)}}\big{)}\big{(}z^{p}_{t^{(m)}_{a+1}}-z^{p}_{t^{(m)}_{a}}+\Delta_{m}\tau^{-1}z^{p}_{t^{(m)}_{a}}\big{)}\\ -\frac{1}{2}\sigma_{t^{(m)}_{a}}^{-2}\Delta_{m}\big{(}G^{p}_{t_{a}^{(m)}}-\tau^{-1}z^{p}_{t_{a}^{(m)}}\big{)}^{2}\bigg{]}. (161)

We now define Γ:𝒰\Gamma:\mathcal{U}\to\mathbb{R} to be such that (in the case that the following limit exists)

Γ(μ)=limjΓ(mj,j)(μ),\displaystyle\Gamma(\mu)=\lim_{j\to\infty}\Gamma^{(m_{j,j})}(\mu), (162)

where mj,jm_{j,j} is a positive integer defined further below in Lemma 20. If the above limit does not exist, then we define Γ(μ)=0\Gamma(\mu)=0 (in fact we will see that the limit always exists if μ𝒰\mu\in\mathcal{U}). It may be observed that Γ\Gamma is a well-defined measurable function.

Lemma 19.

For every N1N\geq 1, every (𝐳0,𝐠0)𝒴N(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}, and for Q𝐳0,𝐠0NQ^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}} almost every (𝐲,𝐆~)(\mathbf{y},\tilde{\mathbf{G}}), the following limit exists

limjΓ(mj,j)(μ^N(𝐲,𝐆~))\lim_{j\to\infty}\Gamma^{(m_{j,j})}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)} (163)

With unit probability, the Radon-Nikodym Derivative in (158) is such that

dP𝐉,𝐳0NdP𝐳0N|T=exp(Γ(μ^N(𝐲,𝐆~)))\displaystyle\frac{dP^{N}_{\mathbf{J},\mathbf{z}_{0}}}{dP^{N}_{\mathbf{z}_{0}}}\bigg{|}_{\mathcal{F}_{T}}=\exp\big{(}\Gamma\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}\big{)} (164)

Also for any ϵ,L>0\epsilon,L>0, there exists k+k\in\mathbb{Z}^{+} such that for all N1N\geq 1,

supjksup𝐳0,𝐠0𝒴NN1logQ𝐳0,𝐠0N(|Γ(mj,j)(μ^N(𝐲,𝐆~))Γ(μ^N(𝐲,𝐆~))|ϵ)L\displaystyle\sup_{j\geq k}\sup_{\mathbf{z}_{0},\mathbf{g}_{0}\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\bigg{(}\big{|}\Gamma^{(m_{j,j})}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}-\Gamma\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}\big{|}\geq\epsilon\bigg{)}\leq-L (165)
Proof.

Define the set

𝒜j={μ𝒰:|Γ(mj,j)(μ)Γ(mj+1,j+1)(μ)|21j}\displaystyle\mathcal{A}_{j}=\big{\{}\mu\in\mathcal{U}:\big{|}\Gamma^{(m_{j,j})}(\mu)-\Gamma^{(m_{j+1,j+1})}(\mu)\big{|}\geq 2^{1-j}\big{\}} (166)

Thanks to a union-of-events bound, for any N1N\geq 1, and using the bound in Lemma 20,

sup𝐳0,𝐠0𝒴NQ𝐳0,𝐠0N(μ^Njk𝒜j)j=kexp(N2j).\displaystyle\sup_{\mathbf{z}_{0},\mathbf{g}_{0}\in\mathcal{Y}^{N}}Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\bigg{(}\hat{\mu}^{N}\in\bigcup_{j\geq k}\mathcal{A}_{j}\bigg{)}\leq\sum_{j=k}^{\infty}\exp\big{(}-N2^{j}\big{)}. (167)

It thus follows from the Borel-Cantelli Lemma that there exists a random kk such that μ^N𝒜j\hat{\mu}^{N}\notin\mathcal{A}_{j} for all jkj\geq k, and so the limit in (163) exists (almost surely). (165) follows analogously.

Lemma 20.

(i) Γ(m):𝒰\Gamma^{(m)}:\mathcal{U}\to\mathbb{R} is continuous. (ii) Moreover, for any a,j+a,j\in\mathbb{Z}^{+}, there exists ma,jm_{a,j} such that for all mma,jm\geq m_{a,j} and all nmn\geq m,

sup𝐳0,𝐠0𝒴NN1logQ𝐳0,𝐠0N(|Γ(m)(μ^N(𝐲,𝐆~))Γ(n)(μ^N(𝐲,𝐆~))2j)2a.\displaystyle\sup_{\mathbf{z}_{0},\mathbf{g}_{0}\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\big{|}\Gamma^{(m)}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}-\Gamma^{(n)}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}\geq 2^{-j}\big{)}\leq-2^{a}. (168)
Proof.

(i) The continuity of Γ(m)\Gamma^{(m)} is almost immediate from the definition.

(ii) For any t[0,T]t\in[0,T], write t(m)=sup{tb(m):tb(m)t}t^{(m)}=\sup\{t^{(m)}_{b}\;:t^{(m)}_{b}\leq t\}. Starting with the discrete approximation to the stochastic integral, we can thus write

b=0m1σta(m)2(Gtb(m)pτ1ztb(m)p)(ztb+1(m)pztb(m)p)=0Tσt(m)2(Gt(m)pτ1zt(m)p)𝑑ztp.\displaystyle\sum_{b=0}^{m-1}\sigma_{t^{(m)}_{a}}^{-2}\big{(}G^{p}_{t_{b}^{(m)}}-\tau^{-1}z^{p}_{t_{b}^{(m)}}\big{)}\big{(}z^{p}_{t^{(m)}_{b+1}}-z^{p}_{t^{(m)}_{b}}\big{)}=\int_{0}^{T}\sigma^{-2}_{t^{(m)}}\big{(}G^{p}_{t^{(m)}}-\tau^{-1}z^{p}_{t^{(m)}}\big{)}dz^{p}_{t}. (169)

Hence,

b=0m1𝔼μ[σtb(m)2(Gtb(m)pτ1ztb(m)p)(ztb+1(m)pzta(m)p)]ab=0n1𝔼μ[σtb(n)2(Gtb(n)pτ1ztb(n)p)(ztb+1(n)pztb(n)p)]=𝔼μ[0T{σt(m)2(Gt(m)pτ1zt(m)p)σt(n)2(Gt(n)pτ1zt(n)p)}𝑑ztp]=𝔼μ[0TpIM(ft(m)pft(n)p)dztp]\sum_{b=0}^{m-1}\mathbb{E}^{\mu}\bigg{[}\sigma_{t^{(m)}_{b}}^{-2}\big{(}G^{p}_{t_{b}^{(m)}}-\tau^{-1}z^{p}_{t_{b}^{(m)}}\big{)}\big{(}z^{p}_{t^{(m)}_{b+1}}-z^{p}_{t^{(m)}_{a}}\big{)}\bigg{]}-\\ \sum_{ab=0}^{n-1}\mathbb{E}^{\mu}\bigg{[}\sigma_{t^{(n)}_{b}}^{-2}\big{(}G^{p}_{t_{b}^{(n)}}-\tau^{-1}z^{p}_{t_{b}^{(n)}}\big{)}\big{(}z^{p}_{t^{(n)}_{b+1}}-z^{p}_{t^{(n)}_{b}}\big{)}\bigg{]}\\ =\mathbb{E}^{\mu}\bigg{[}\int_{0}^{T}\bigg{\{}\sigma^{-2}_{t^{(m)}}\big{(}G^{p}_{t^{(m)}}-\tau^{-1}z^{p}_{t^{(m)}}\big{)}-\sigma^{-2}_{t^{(n)}}\big{(}G^{p}_{t^{(n)}}-\tau^{-1}z^{p}_{t^{(n)}}\big{)}\bigg{\}}dz^{p}_{t}\bigg{]}\\ =\mathbb{E}^{\mu}\bigg{[}\int_{0}^{T}\sum_{p\in I_{M}}(f^{p}_{t^{(m)}}-f^{p}_{t^{(n)}})dz_{t}^{p}\bigg{]} (170)

where ftp=σt2(Gtpτ1ztp)f^{p}_{t}=\sigma_{t}^{-2}\big{(}G^{p}_{t}-\tau^{-1}z^{p}_{t}\big{)}. Writing

ftp,j=σt2(Gtp,jτ1ztp,j),\displaystyle f^{p,j}_{t}=\sigma_{t}^{-2}\big{(}G^{p,j}_{t}-\tau^{-1}z^{p,j}_{t}\big{)}, (171)

we obtain that

𝔼μ^N[0TpIM(ft(m)pft(n)p)dztp]=N1jINpIM0T(ft(m)p,jft(n)p,j)𝑑ytp,j.\displaystyle\mathbb{E}^{\hat{\mu}^{N}}\bigg{[}\int_{0}^{T}\sum_{p\in I_{M}}(f^{p}_{t^{(m)}}-f^{p}_{t^{(n)}})dz_{t}^{p}\bigg{]}=N^{-1}\sum_{j\in I_{N}\fatsemi p\in I_{M}}\int_{0}^{T}\big{(}f^{p,j}_{t^{(m)}}-f^{p,j}_{t^{(n)}}\big{)}dy^{p,j}_{t}. (172)

The quadratic variation of this stochastic integral is

(QV)t(m,n),N=N2jINpIM0t(fs(m)p,jfs(n)p,j)2σs2𝑑s\displaystyle(QV)^{(m,n),N}_{t}=N^{-2}\sum_{j\in I_{N}\fatsemi p\in I_{M}}\int_{0}^{t}\big{(}f^{p,j}_{s^{(m)}}-f^{p,j}_{s^{(n)}}\big{)}^{2}\sigma_{s}^{2}ds (173)

By definition of the set 𝒰a\mathcal{U}_{a}, if μ^N𝒰a\hat{\mu}^{N}\in\mathcal{U}_{a}, then for any δ>0\delta>0, one can find mδm_{\delta} such that as long as m,nmδm,n\geq m_{\delta}, necessarily

(QV)T(m,n),NN1δ.(QV)^{(m,n),N}_{T}\leq N^{-1}\delta.

Then writing w()w(\cdot) to be a standard Brownian Motion, using the Dambin-Dubins-Schwarz [27] time-rescaled representation of the stochastic integral, as long as (m,n)mδ(m,n)\geq m_{\delta},

(μ^N𝒰a,|0TpIM(ft(m)pft(n)p)dztp|ϵ2)\displaystyle\mathbb{P}\bigg{(}\hat{\mu}^{N}\in\mathcal{U}_{a}\;,\;\bigg{|}\int_{0}^{T}\sum_{p\in I_{M}}(f^{p}_{t^{(m)}}-f^{p}_{t^{(n)}})dz_{t}^{p}\bigg{|}\geq\frac{\epsilon}{2}\bigg{)}\leq (|w(N1δ)|ϵ)\displaystyle\mathbb{P}\big{(}\big{|}w\big{(}N^{-1}\delta\big{)}\big{|}\geq\epsilon\big{)} (174)
=\displaystyle= exp(Nϵ2/(8δ))exp(NL),\displaystyle\exp\big{(}-N\epsilon^{2}/(8\delta)\big{)}\leq\exp(-NL), (175)

as long as we choose δ\delta sufficiently small.

The other terms in

Γ(m)(μ^N(𝐲,𝐆~))Γ(n)(μ^N(𝐲,𝐆~))\Gamma^{(m)}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}-\Gamma^{(n)}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\big{)}

are treated similarly (observe that they are just Riemann Sums, so its straightforward to control their difference from the limiting integral).

We now prove Theorems 10 and 12.

Proof.

In the case of connectivity-independent initial conditions (Case 2 of the Assumptions), the theorem follows from Corollary 11. Since the relative entropy is only zero when its two arguments are identical, any zero must be a fixed point of the operator Φ\Phi. It is proved in the following Lemma that there is a unique zero.

For the rest of this proof, we prove the theorem in the case of connectivity-dependent initial conditions. We start by proving that for any ϵ>0\epsilon>0, there must exist a measure μ𝒰\mu\in\mathcal{U} such that

limN¯N1log(dW(μ^N(𝐳,𝐆),μ)ϵ)=0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}d_{W}(\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu)\leq\epsilon\big{)}=0. (176)

Write 𝔘=𝒰a\mathfrak{U}=\mathcal{U}_{a}, where aa is large enough that

limN¯sup(𝐳0,𝐠0)𝒴NN1logQ𝐳0,𝐠0N(μ^N(𝐲,𝐆~)𝒰a)<C.\underset{N\to\infty}{\overline{\lim}}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log Q^{N}_{\mathbf{z}_{0},\mathbf{g}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y},\tilde{\mathbf{G}})\in\mathcal{U}_{a}\big{)}<-C.

where CC is the upperbound for Γ\Gamma in Lemma 21. This is possible thanks to the Exponential Tightness. Thanks to the Radon-Nikodym derivative identity in (160), we thus find that

limN¯N1log(μ^N(𝐳,𝐆~)𝔘)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{z},\tilde{\mathbf{G}})\notin\mathfrak{U}\big{)}<0. (177)

Thus for (176) to hold, it suffices that we prove that there exists μ𝔘\mu\in\mathfrak{U} such that

limN¯N1log(μ^N(𝐳,𝐆)𝔘,dW(μ^N(𝐳,𝐆),μ)ϵ)=0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{z},\mathbf{G})\in\mathfrak{U}\;,\;d_{W}(\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu)\leq\epsilon\big{)}=0. (178)

Since 𝔘\mathfrak{U} is compact, for any ϵ>0\epsilon>0, we can obtain a finite covering of 𝔘\mathfrak{U} of the form

𝔘i=1𝒩ϵBϵ(μi),\displaystyle\mathfrak{U}\subseteq\bigcup_{i=1}^{\mathcal{N}_{\epsilon}}B_{\epsilon}(\mu_{i}), (179)

where μi𝔘\mu_{i}\in\mathfrak{U}. By a union of events bound,

0=\displaystyle 0= limNN1log(μ^N(𝐳,𝐆~)𝔘)\displaystyle\lim_{N\to\infty}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{z},\tilde{\mathbf{G}})\in\mathfrak{U}\big{)} (180)
\displaystyle\leq max1i𝒩ϵ{limN¯N1log(μ^N(𝐳,𝐆~)Bϵ(μi))}\displaystyle\max_{1\leq i\leq\mathcal{N}_{\epsilon}}\bigg{\{}\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{z},\tilde{\mathbf{G}})\in B_{\epsilon}(\mu_{i})\big{)}\bigg{\}} (181)

If our proposition in (178) were to be false, then (181) would be strictly negative, which would be a contradiction.

Write μ(k)𝔘\mu_{(k)}\in\mathfrak{U} to be such that

limN¯N1log(dW(μ^N(𝐳,𝐆),μ(k))k1)=0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}d_{W}(\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu_{(k)})\geq k^{-1}\big{)}=0. (182)

Let μ𝔘\mu\in\mathfrak{U} be any measure such that for some subsequence (pk)k1(p_{k})_{k\geq 1}, limkμ(pk)=μ\lim_{k\to\infty}\mu_{(p_{k})}=\mu (this must be possible because 𝔘\mathfrak{U} is compact).

We next claim that

limϵ0+lim¯NN1log(dW(μ^N(𝐳,𝐆),μ)ϵ)=(μ)+Γ(μ)\displaystyle\lim_{\epsilon\to 0^{+}}\underset{N\to\infty}{\underline{\lim}}N^{-1}\log\mathbb{P}\big{(}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu\big{)}\leq\epsilon\big{)}=-\mathcal{I}(\mu)+\Gamma(\mu) (183)

Indeed writing 𝒜ϵ={dW(μ^N(𝐳,𝐆),μ)ϵ}\mathcal{A}_{\epsilon}=\big{\{}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu\big{)}\leq\epsilon\big{\}},

(dW(μ^N(𝐳,𝐆),μ)ϵ)=\displaystyle\mathbb{P}\big{(}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu\big{)}\leq\epsilon\big{)}= 𝔼γ[MNP𝐉,𝐱N(𝒜ϵ)ρ𝐉N(𝐱)𝑑𝐱]\displaystyle\mathbb{E}^{\gamma}\bigg{[}\int_{\mathbb{R}^{MN}}P^{N}_{\mathbf{J},\mathbf{x}}(\mathcal{A}_{\epsilon})\rho^{N}_{\mathbf{J}}(\mathbf{x})d\mathbf{x}\bigg{]} (184)
=\displaystyle= 𝔼γ[MN𝔼P𝐱N[exp(NΓ(μ^N))χ{𝒜ϵ}]ρ𝐉N(𝐱)𝑑𝐱]\displaystyle\mathbb{E}^{\gamma}\bigg{[}\int_{\mathbb{R}^{MN}}\mathbb{E}^{P^{N}_{\mathbf{x}}}\big{[}\exp\big{(}N\Gamma(\hat{\mu}^{N})\big{)}\chi\{\mathcal{A}_{\epsilon}\}\big{]}\rho^{N}_{\mathbf{J}}(\mathbf{x})d\mathbf{x}\bigg{]} (185)
=\displaystyle= MN𝔼γ[𝔼P𝐱N[exp(NΓ(μ^N))χ{𝒜ϵ}]ρ𝐉N(𝐱)]𝑑𝐱\displaystyle\int_{\mathbb{R}^{MN}}\mathbb{E}^{\gamma}\bigg{[}\mathbb{E}^{P^{N}_{\mathbf{x}}}\big{[}\exp\big{(}N\Gamma(\hat{\mu}^{N})\big{)}\chi\{\mathcal{A}_{\epsilon}\}\big{]}\rho^{N}_{\mathbf{J}}(\mathbf{x})\bigg{]}d\mathbf{x} (186)
=\displaystyle= MN𝔼γ[𝔼γ[𝔼P𝐱N[exp(NΓ(μ^N))χ{𝒜ϵ}]ρ𝐉N(𝐱)|𝐆0]]𝑑𝐱\displaystyle\int_{\mathbb{R}^{MN}}\mathbb{E}^{\gamma}\bigg{[}\mathbb{E}^{\gamma}\bigg{[}\mathbb{E}^{P^{N}_{\mathbf{x}}}\big{[}\exp\big{(}N\Gamma(\hat{\mu}^{N})\big{)}\chi\{\mathcal{A}_{\epsilon}\}\big{]}\rho^{N}_{\mathbf{J}}(\mathbf{x})\;\bigg{|}\;\mathbf{G}_{0}\bigg{]}\bigg{]}d\mathbf{x} (187)

and in this last step we first perform the conditional expectation, for γ\gamma conditioned on the values of {G0p,j}jINpIM\{G^{p,j}_{0}\}_{j\in I_{N}\fatsemi p\in I_{M}}.

Now, recall that

ρ𝐉N(𝐳0)=(Z𝐉N)1χ{μ^N(𝐳0,𝐆0)BδN(κ)}.\rho^{N}_{\mathbf{J}}(\mathbf{z}_{0})=(Z^{N}_{\mathbf{J}})^{-1}\chi\big{\{}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0})\in B_{\delta_{N}}(\kappa)\big{\}}.

Furthermore, writing

uN=N1log𝔼[Z𝐉N],u_{N}=N^{-1}\log\mathbb{E}[Z^{N}_{\mathbf{J}}],

our assumption on the initial condition dictates that for any δ>0\delta>0,

limN¯N1log(|N1logZ𝐉NuN|δ)<0.\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\big{|}N^{-1}\log Z^{N}_{\mathbf{J}}-u_{N}\big{|}\geq\delta\big{)}<0. (188)

Next, we claim that

limϵ0+infν𝔘𝒜ϵΓ(ν)=Γ(μ).\displaystyle\lim_{\epsilon\to 0^{+}}\inf_{\nu\in\mathfrak{U}\cap\mathcal{A}_{\epsilon}}\Gamma(\nu)=\Gamma(\mu). (189)

Indeed (189) is a consequence of Lemma 19: this Lemma implies that Γ\Gamma can be approximated arbitrarily well by continuous functions over 𝔘\mathfrak{U}.

We thus obtain that

limϵ0+lim¯Ninf(𝐳0,𝐆0)N1log(dW(μ^N(𝐳,𝐆),μ)ϵ)\displaystyle\lim_{\epsilon\to 0^{+}}\underset{N\to\infty}{\underline{\lim}}\inf_{(\mathbf{z}_{0},\mathbf{G}_{0})}N^{-1}\log\big{(}d_{W}\big{(}\hat{\mu}^{N}(\mathbf{z},\mathbf{G}),\mu\big{)}\leq\epsilon\big{)} (190)
=\displaystyle= Γ(μ)+limϵ0+lim¯N{uN+N1logMN𝔼γ[Q𝐱,𝐆0N(𝒜ϵ)]χ{μ^N(𝐳0,𝐆0)BδN(κ)}𝑑𝐱}\displaystyle\Gamma(\mu)+\lim_{\epsilon\to 0^{+}}\underset{N\to\infty}{\underline{\lim}}\big{\{}-u_{N}+N^{-1}\log\int_{\mathbb{R}^{MN}}\mathbb{E}^{\gamma}\big{[}Q^{N}_{\mathbf{x},\mathbf{G}_{0}}\big{(}\mathcal{A}_{\epsilon}\big{)}\big{]}\chi\big{\{}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0})\in B_{\delta_{N}}(\kappa)\big{\}}d\mathbf{x}\big{\}} (191)
=\displaystyle= Γ(μ)limϵ0+infν𝒜(ϵ)(ν),\displaystyle\Gamma(\mu)-\lim_{\epsilon\to 0^{+}}\inf_{\nu\in\mathcal{A}_{(\epsilon)}}\mathcal{I}(\nu), (192)

since (by definition)

N1logMN𝔼γ[χ{μ^N(𝐳0,𝐆0)BδN(κ)]d𝐳0=uN,N^{-1}\log\int_{\mathbb{R}^{MN}}\mathbb{E}^{\gamma}\big{[}\chi\big{\{}\hat{\mu}^{N}(\mathbf{z}_{0},\mathbf{G}_{0})\in B_{\delta_{N}}(\kappa)\big{]}d\mathbf{z}_{0}=u_{N},

and we have employed the uniform lower bound in (36). The lower semi-continuity of \mathcal{I} implies that

limϵ0+infν𝒜(ϵ)(ν)=(μ).\lim_{\epsilon\to 0^{+}}\inf_{\nu\in\mathcal{A}_{(\epsilon)}}\mathcal{I}(\nu)=\mathcal{I}(\mu).

We thus obtain (183), as required.

Next, we must show that 𝒥(μ)=(μ)Γ(μ)\mathcal{J}(\mu)=\mathcal{I}(\mu)-\Gamma(\mu) (recall the definition of 𝒥(μ)\mathcal{J}(\mu) in (56). Now

dSμ,𝐳0,𝐠0dQμ,𝐳0,𝐠0|T=exp(pIM0Tσs2(gspτ1zsp)𝑑zsp12σs2(gspτ1zsp)2ds)\displaystyle\frac{dS_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}}}{dQ_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}}}\bigg{|}_{\mathcal{F}_{T}}=\exp\bigg{(}\sum_{p\in I_{M}}\int_{0}^{T}\sigma_{s}^{-2}\big{(}g^{p}_{s}-\tau^{-1}z^{p}_{s}\big{)}dz^{p}_{s}-\frac{1}{2}\sigma_{s}^{-2}\big{(}g^{p}_{s}-\tau^{-1}z^{p}_{s}\big{)}^{2}ds\bigg{)} (193)

Substituting this identity into the proposed rate function definition in (56),

𝔼κ[(μ𝐳0,𝐠0||Sμ,𝐳0,𝐠0)]=(μ)Γ(μ),\displaystyle\mathbb{E}^{\kappa}\bigg{[}\mathcal{R}\big{(}\mu_{\mathbf{z}_{0},\mathbf{g}_{0}}||S_{\mu,\mathbf{z}_{0},\mathbf{g}_{0}}\big{)}\bigg{]}=\mathcal{I}(\mu)-\Gamma(\mu), (194)

as required.

The above reasoning dictates that there must be at least one ξ\xi such that 𝒥(ξ)=0\mathcal{J}(\xi)=0. In fact, it is proved in Lemma 22 that there can only be one measure ξ\xi such that 𝒥(ξ)=0\mathcal{J}(\xi)=0. Furthermore it follows from Lemma 22, over small enough time increments, the mapping Φt\Phi_{t} must be a contraction. This implies (66). ∎

Lemma 21.

There exists a constant C>0C>0 such that

limN¯N1log(Γ𝐉,TN(𝐳)C)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C\big{)}<0. (195)
Proof.

For any >0\ell>0,

limN¯N1log(Γ𝐉,TN(𝐳)C)max{limN¯N1log(𝒥N>),limN¯N1log(𝒥N,Γ𝐉,TN(𝐳)C)}\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\bigg{(}\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C\bigg{)}\leq\max\bigg{\{}\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|>\ell\big{)},\\ \underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell,\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C\big{)}\bigg{\}} (196)

Thanks to Lemma 13, limN¯N1log(𝒥N>)\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|>\ell\big{)} converges to -\infty as \ell\to\infty. It thus suffices that we prove that, for abitrary >0\ell>0, there exists C>0C_{\ell}>0 such that

limN¯N1log(𝒥N,Γ𝐉,TN(𝐳)C)<0.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell,\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C_{\ell}\big{)}<0. (197)

Now, leaving out the negative-semi-definite terms, we find that

Γ𝐉,TN(𝐳)N1jINpIM0Tσs2(G~sp,jτ1ysp,j)𝑑ysp,j\displaystyle\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\leq N^{-1}\sum_{j\in I_{N}\fatsemi p\in I_{M}}\int_{0}^{T}\sigma_{s}^{-2}\big{(}\tilde{G}^{p,j}_{s}-\tau^{-1}y^{p,j}_{s}\big{)}dy^{p,j}_{s} (198)

Furthermore, writing hsp=σs2(G~sp,jτ1ysp,j)h^{p}_{s}=\sigma_{s}^{-2}\big{(}\tilde{G}^{p,j}_{s}-\tau^{-1}y^{p,j}_{s}\big{)}, and assuming that 𝒥N\left\|\mathcal{J}_{N}\right\|\leq\ell, one finds that

jIN(hsp,j)2\displaystyle\sum_{j\in I_{N}}(h^{p,j}_{s})^{2}\leq 2σs4jIN{(G~sp,j)2+τ2(ysp,j)2}\displaystyle 2\sigma_{s}^{-4}\sum_{j\in I_{N}}\big{\{}(\tilde{G}^{p,j}_{s})^{2}+\tau^{-2}(y^{p,j}_{s})^{2}\big{\}} (199)
\displaystyle\leq 2σs4jIN{λ(ysp,j)2+τ2(ysp,j)2}\displaystyle 2\sigma_{s}^{-4}\sum_{j\in I_{N}}\big{\{}\ell\lambda(y^{p,j}_{s})^{2}+\tau^{-2}(y^{p,j}_{s})^{2}\big{\}} (200)
\displaystyle\leq 2σs4jIN{Cλ2(ysp,j)2+τ2(ysp,j)2}.\displaystyle 2\sigma_{s}^{-4}\sum_{j\in I_{N}}\big{\{}\ell C_{\lambda}^{2}(y^{p,j}_{s})^{2}+\tau^{-2}(y^{p,j}_{s})^{2}\big{\}}. (201)

We thus find that, thanks to Lemma , for any L>0L>0 there exists a constant C¯L>0\bar{C}_{L}>0 such that

limN¯N1log(suppIMjIN(hsp,j)2NC¯L)L.\displaystyle\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\sup_{p\in I_{M}}\sum_{j\in I_{N}}(h^{p,j}_{s})^{2}\geq N\bar{C}_{L}\big{)}\leq-L. (202)

Write

N={suppIMjIN(hsp,j)2NC¯L}.\mathcal{H}_{N}=\bigg{\{}\sup_{p\in I_{M}}\sum_{j\in I_{N}}(h^{p,j}_{s})^{2}\leq N\bar{C}_{L}\bigg{\}}.

We thus find that,

limN¯N1log(𝒥N,N,Γ𝐉,TN(𝐳)C)max{limN¯N1log(Nc),limN¯N1log(𝒥N,N,Γ𝐉,TN(𝐳)C)}max{L,limN¯N1log(𝒥N,N,Γ𝐉,TN(𝐳)C)}\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell\;,\;\mathcal{H}_{N}\;,\;\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C_{\ell}\big{)}\\ \leq\max\bigg{\{}\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\mathcal{H}_{N}^{c}\big{)},\\ \underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell\;,\;\mathcal{H}_{N}\;,\;\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C_{\ell}\big{)}\bigg{\}}\\ \leq\max\bigg{\{}-L,\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell\;,\;\mathcal{H}_{N}\;,\;\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C_{\ell}\big{)}\bigg{\}} (203)

Furthermore, using the Dambins-Dubins Schwarz Theorem [27], and writing w(t)w(t) to be 1D Brownian Motion,

limN¯N1log(𝒥N,N,Γ𝐉,TN(𝐳)C)limN¯N1log(sups[0,T]|w(C¯N1s)|C)L,\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\left\|\mathcal{J}_{N}\right\|\leq\ell\;,\;\mathcal{H}_{N}\;,\;\Gamma^{N}_{\mathbf{J},T}(\mathbf{z})\geq C_{\ell}\big{)}\\ \leq\underset{N\to\infty}{\overline{\lim}}N^{-1}\log\mathbb{P}\big{(}\sup_{s\in[0,T]}\big{|}w\big{(}\bar{C}N^{-1}s\big{)}\big{|}\geq C_{\ell}\big{)}\leq-L, (204)

as long as CC_{\ell} is sufficiently large, using standard properties of Brownian Motion. ∎

We now prove that the rate function 𝒥\mathcal{J} has a unique minimizer (i.e. we prove Lemma 9).

Lemma 22.

There exists a unique fixed point ξ\xi of Φ\Phi in 𝒰\mathcal{U}. Furthermore ξ\xi is such that for any μ𝒰\mu\in\mathcal{U}, writing ξ(1)=μ\xi_{(1)}=\mu and ξ(n+1)=Φ(ξ(n))\xi_{(n+1)}=\Phi(\xi_{(n)}), it holds that

ξ=limnξ(n)\displaystyle\xi=\lim_{n\to\infty}\xi_{(n)} (205)
Proof.

We start by considering the following restricted map Φ~:𝔔𝔔\tilde{\Phi}:\mathfrak{Q}\to\mathfrak{Q}. This is such that Φ~(μ)=ν(1)\tilde{\Phi}(\mu)=\nu^{(1)}, where ν=Φ(α)\nu=\Phi(\alpha) for any α𝒰\alpha\in\mathcal{U} such that α(1)=μ\alpha^{(1)}=\mu. Define dt(2):𝔔×𝔔+d^{(2)}_{t}:\mathfrak{Q}\times\mathfrak{Q}\to\mathbb{R}^{+} analogously.

We are going to demonstrate that there is a constant c>0c>0 such that for all μ,ν𝔔\mu,\nu\in\mathfrak{Q},

dt(2)(Φ~t(μ),Φ~t(ν))ctdt(2)(μ,ν).\displaystyle d^{(2)}_{t}\big{(}\tilde{\Phi}_{t}(\mu),\tilde{\Phi}_{t}(\nu)\big{)}\leq c\sqrt{t}d^{(2)}_{t}(\mu,\nu). (206)

For any μ,ν𝔔\mu,\nu\in\mathfrak{Q}, we construct a particular ζ\zeta that is within η1\eta\ll 1 of realizing the infimum in the definition of the Wasserstein distance in (98). To do this, we employ the construction of Lemma 16. Let 𝐆μ,𝐆ν\mathbf{G}^{\mu},\mathbf{G}^{\nu} be 𝒞([0,T],M)\mathcal{C}([0,T],\mathbb{R}^{M})-valued random variables (in the same probability space), with joint probability law βμ,ν\beta_{\mu,\nu}. Then for Brownian motions (Wtp)pIM\big{(}W^{p}_{t}\big{)}_{p\in I_{M}}, define

dztν,p\displaystyle dz^{\nu,p}_{t} =(τ1ztν,p+Gtν,p)dt+σtdWtp\displaystyle=\big{(}-\tau^{-1}z^{\nu,p}_{t}+G^{\nu,p}_{t}\big{)}dt+\sigma_{t}dW^{p}_{t} (207)
dztμ,p\displaystyle dz^{\mu,p}_{t} =(τ1ztμ,p+Gtμ,p)dt+σtdWtp.\displaystyle=\big{(}-\tau^{-1}z^{\mu,p}_{t}+G^{\mu,p}_{t}\big{)}dt+\sigma_{t}dW^{p}_{t}. (208)

The initial conditions are identical: z0ν,p=z0μ,pz^{\nu,p}_{0}=z^{\mu,p}_{0}. We immediately see that

ddt(ztν,pztμ,p)=τ1(ztν,pztμ,p)+Gtν,pGtμ,p,\displaystyle\frac{d}{dt}\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}=-\tau^{-1}\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}+G^{\nu,p}_{t}-G^{\mu,p}_{t}, (209)

and hence

ddt(ztν,pztμ,p)2\displaystyle\frac{d}{dt}\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}^{2} =2τ1(ztν,pztμ,p)2+2(ztν,pztμ,p)(Gtν,pGtμ,p) and thus\displaystyle=-2\tau^{-1}\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}^{2}+2\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}\big{(}G^{\nu,p}_{t}-G^{\mu,p}_{t}\big{)}\text{ and thus } (210)
(ztν,pztμ,p)2\displaystyle\big{(}z^{\nu,p}_{t}-z^{\mu,p}_{t}\big{)}^{2} =0texp(2(st)/τ)2(zsν,pzsμ,p)(Gsν,pGsμ,p)𝑑s.\displaystyle=\int_{0}^{t}\exp\big{(}2(s-t)/\tau\big{)}2\big{(}z^{\nu,p}_{s}-z^{\mu,p}_{s}\big{)}\big{(}G^{\nu,p}_{s}-G^{\mu,p}_{s}\big{)}ds. (211)

It follows from this that there exists a constant c>0c>0 such that for all t[0,T]t\in[0,T],

dt(2)(Φ~t(μ),Φ~t(ν))\displaystyle d^{(2)}_{t}\big{(}\tilde{\Phi}_{t}(\mu),\tilde{\Phi}_{t}(\nu)\big{)} ctdt(2)(βμ,βν)\displaystyle\leq ctd^{(2)}_{t}\big{(}\beta_{\mu},\beta_{\nu}\big{)} (212)
cCλtdt(2)(μ,ν),\displaystyle\leq cC_{\lambda}td^{(2)}_{t}(\mu,\nu), (213)

using Lemma 16. Thus for small enough tt, there is a unique fixed point of Φ~t\tilde{\Phi}_{t} (the mapping upto time tt). Iterating this argument, we find a unique fixed point for Φ~\tilde{\Phi}. The uniqueness for Φ~\tilde{\Phi} in turn implies uniqueness for Φ\Phi, thanks to the identity in Lemma 16.

To see why (205) holds. First consider arbitrary ν(1)𝔔\nu_{(1)}\in\mathfrak{Q}, and define ν(n+1)=Φ~(ν(n))\nu_{(n+1)}=\tilde{\Phi}(\nu_{(n)}). The above bound in (213) implies that necessarily (ν(n))n1\big{(}\nu_{(n)}\big{)}_{n\geq 1} is Cauchy. It then immediate follows that for any ξ(1)𝒰\xi_{(1)}\in\mathcal{U} with first marginal equal to ν(1)\nu_{(1)}, and writing ξ(n+1)=Φ(ξ(n))\xi_{(n+1)}=\Phi(\xi_{(n)}), it must be that (ξ(n))n1\big{(}\xi_{(n)}\big{)}_{n\geq 1} is Cauchy.

Finally we note that d(2)d^{(2)} metrizes weak convergence, thanks to Lemma 15.

Appendix A Bounding Fluctuations of the Noise

For the processes (𝐲[0,T]j)jIN(\mathbf{y}^{j}_{[0,T]})_{j\in I_{N}} that are defined in (31), define the empirical measure

μ^N(𝐲)=N1jINδ𝐲[0,T]j𝒫(𝒞([0,T],M)).\displaystyle\hat{\mu}^{N}(\mathbf{y})=N^{-1}\sum_{j\in I_{N}}\delta_{\mathbf{y}^{j}_{[0,T]}}\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)}. (214)

Next, we bound the probability of the empirical being in the set 𝔔a\mathfrak{Q}_{a}, defined in (38), which we recall

𝒬𝔞={μ𝒫(𝒞([0,T],M)):supm𝔞sup0im𝔼μ[supMIM(wti+1(m)pwti(m)p)2]>Δm1/4 and μ𝒦𝔞 and suppIM𝔼μ[supt[0,T](ytp)2]𝔞}\mathcal{Q}_{\mathfrak{a}}=\bigg{\{}\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)}\;:\;\sup_{m\geq\mathfrak{a}}\sup_{0\leq i\leq m}\mathbb{E}^{\mu}\big{[}\sup_{M\in I_{M}}(w^{p}_{t^{(m)}_{i+1}}-w^{p}_{t^{(m)}_{i}})^{2}\big{]}>\Delta_{m}^{1/4}\text{ and }\\ \mu\in\mathcal{K}_{\mathfrak{a}}\text{ and }\sup_{p\in I_{M}}\mathbb{E}^{\mu}[\sup_{t\in[0,T]}(y^{p}_{t})^{2}\big{]}\leq\mathfrak{a}\bigg{\}} (215)

where Δm=T/m\Delta_{m}=T/m and ti(m)=iT/mt^{(m)}_{i}=iT/m. The main result of this section is the following.

Lemma 23.

For any L>0L>0, there exists 𝔞+\mathfrak{a}\in\mathbb{Z}^{+} such that for all N1N\geq 1,

sup(𝐳0,𝐠0)N1logP𝐳0N(μ^N(𝐲)𝒬𝔞)L.\displaystyle\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})}N^{-1}\log P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{Q}_{\mathfrak{a}}\big{)}\leq-L. (216)
Proof.

Employing a union-of-events bound, or any (𝐳0,𝐠0)𝒴N(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N},

N1logP𝐳0N(μ^N(𝐲)𝒬𝔞)N1log{P𝐳0N(suppIMN1jINsupt[0,T](ytp,j)2>𝔞)+P𝐳0N(sup0tΔmjINsup0im1suppIM|yt+ti(m)p,jyti(m)p,j|2NaΔm)+P𝐳0N(μ^N(𝐲)𝒦𝔞)}.N^{-1}\log P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{Q}_{\mathfrak{a}}\big{)}\leq N^{-1}\log\bigg{\{}P^{N}_{\mathbf{z}_{0}}\bigg{(}\sup_{p\in I_{M}}N^{-1}\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t})^{2}>\mathfrak{a}\bigg{)}\\ +P^{N}_{\mathbf{z}_{0}}\bigg{(}\sup_{0\leq t\leq\Delta_{m}}\sum_{j\in I_{N}}\sup_{0\leq i\leq m-1}\sup_{p\in I_{M}}\big{|}y^{p,j}_{t+t^{(m)}_{i}}-y^{p,j}_{t^{(m)}_{i}}\big{|}^{2}\geq Na\Delta_{m}\bigg{)}+P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{K}_{\mathfrak{a}}\big{)}\bigg{\}}. (217)

With a view to bounding the first term on the RHS, since y0p,j=z0p,jy^{p,j}_{0}=z^{p,j}_{0},

(ytp,j)22(ytp,jy0p,j)2+2(z0p,j)2.(y^{p,j}_{t})^{2}\leq 2\big{(}y^{p,j}_{t}-y^{p,j}_{0}\big{)}^{2}+2(z^{p,j}_{0})^{2}.

Thus for a positive constant b>0b>0,

𝔼P𝐳0N[exp(bsuppIMjINsupt[0,T](ytp,j)2)]𝔼P𝐳0N[exp(2bsuppIMjIN(z0p,j)2+2bjINsupt[0,T](ytp,jz0p,j)2)].\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\exp\bigg{(}b\sup_{p\in I_{M}}\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t})^{2}\bigg{)}\bigg{]}\\ \leq\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\exp\bigg{(}2b\sup_{p\in I_{M}}\sum_{j\in I_{N}}(z^{p,j}_{0})^{2}+2b\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t}-z^{p,j}_{0})^{2}\bigg{)}\bigg{]}. (218)

Thus, thanks to Chernoff’s Inequality,

N1logP𝐳0N(suppIMN1jINsupt[0,T](ytp,j)2>𝔞)2bNsuppIMjIN(z0p,j)2\displaystyle N^{-1}\log P^{N}_{\mathbf{z}_{0}}\bigg{(}\sup_{p\in I_{M}}N^{-1}\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t})^{2}>\mathfrak{a}\bigg{)}\leq\frac{2b}{N}\sup_{p\in I_{M}}\sum_{j\in I_{N}}(z^{p,j}_{0})^{2} (219)
+N1log𝔼P𝐳0N[exp(2bjINsupt[0,T](ytp,jz0p,j)2)]b𝔞.\displaystyle+N^{-1}\log\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\exp\bigg{(}2b\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t}-z^{p,j}_{0})^{2}\bigg{)}\bigg{]}-b\mathfrak{a}. (220)

The first term on the RHS is bounded for all NN and all (𝐳0,𝐠0)𝒴N(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}. For the second term on the RHS, standard theory on stochastic processes implies that the exponential moment exists, as long as bb is small enough. Thus, taking 𝔞\mathfrak{a}\to\infty, the RHS can be made arbitrarily small. We thus find that

lim𝔞supN1sup(𝐳0,𝐠0)𝒴NN1logP𝐳0N(suppIMN1jINsupt[0,T](ytp,j)2>𝔞)=.\displaystyle\lim_{\mathfrak{a}\to\infty}\sup_{N\geq 1}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log P^{N}_{\mathbf{z}_{0}}\bigg{(}\sup_{p\in I_{M}}N^{-1}\sum_{j\in I_{N}}\sup_{t\in[0,T]}(y^{p,j}_{t})^{2}>\mathfrak{a}\bigg{)}=-\infty. (221)

The Lemma now follows from applying (221), Lemma 25 and Lemma 24 to (217). ∎

The following result is well-known. Nevertheless we sketch a quick proof for clarity.

Lemma 24.

For any L>0L>0, there exists a compact set 𝒦L\mathcal{K}_{L} such that for all N1N\geq 1,

sup(𝐳0,𝐠0)𝒴NN1logP𝐳0N(μ^N(𝐲)𝒦L)L\displaystyle\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{K}_{L}\big{)}\leq-L (222)
Proof.

The following property follows straightforwardly from properties of the stochastic integral (noting that the diffusion coefficient is uniformly bounded): for any ϵ>0\epsilon>0, there exists a compact set 𝒞ϵ𝒞([0,T],M)\mathcal{C}_{\epsilon}\subset\mathcal{C}([0,T],\mathbb{R}^{M}) such that for all jINj\in I_{N} such that z0jϵ1\|z_{0}^{j}\|\leq\epsilon^{-1},

supjINP𝐳0N(y[0,T]j𝒞ϵ)ϵ.\displaystyle\sup_{j\in I_{N}}P^{N}_{\mathbf{z}_{0}}\big{(}y^{j}_{[0,T]}\notin\mathcal{C}_{\epsilon}\big{)}\leq\epsilon. (223)

Write

uϵN=sup(𝐳0,𝐠0)𝒴NN1jINχ{𝐲0jϵ1},\displaystyle u^{N}_{\epsilon}=\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\sum_{j\in I_{N}}\chi\{\|\mathbf{y}^{j}_{0}\|\geq\epsilon^{-1}\}, (224)

and note that our assumptions on 𝒴N\mathcal{Y}^{N} dictates that

limϵ0+limNuϵN=0.\displaystyle\lim_{\epsilon\to 0^{+}}\lim_{N\to\infty}u_{\epsilon}^{N}=0. (225)

For any m+m\in\mathbb{Z}^{+}, define the set m,δ𝒫(𝒞([0,T],M))\mathcal{L}_{m,\delta}\subset\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)} to be such that

m,δ={μ𝒫(𝒞([0,T],M)):μ(𝒞m1)δ)}\displaystyle\mathcal{L}_{m,\delta}=\big{\{}\mu\in\mathcal{P}\big{(}\mathcal{C}([0,T],\mathbb{R}^{M})\big{)}\;:\;\mu(\mathcal{C}_{m^{-1}})\geq\delta\big{)}\big{\}} (226)

We claim that for any m1m\geq 1, there exists δm>0\delta_{m}>0 such that

supN1sup(𝐳0,𝐠0)𝒴NN1logP𝐳0N(μ^N(𝐲)m,δm)m\displaystyle\sup_{N\geq 1}\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}\big{(}\mathbf{y}\big{)}\notin\mathcal{L}_{m,\delta_{m}}\big{)}\leq-m (227)

To see this, employing a Chernoff Inequality, for a constant b>0b>0, for any (𝐳0,𝐠0)𝒴N(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N},

N1logP𝐳0N(μ^N(𝐲)m,δ)\displaystyle N^{-1}\log P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}\big{(}\mathbf{y}\big{)}\notin\mathcal{L}_{m,\delta}\big{)}\leq 𝔼P𝐳0N[exp(bjINχ{𝐲[0,T]j𝒞m1}Nbδ)]\displaystyle\mathbb{E}^{P^{N}_{\mathbf{z}_{0}}}\bigg{[}\exp\bigg{(}b\sum_{j\in I_{N}}\chi\{\mathbf{y}^{j}_{[0,T]}\notin\mathcal{C}_{m^{-1}}\}-Nb\delta\bigg{)}\bigg{]} (228)
\displaystyle\leq bδ+N1log{(ϵ+uϵN)(exp(b)1)+1}N\displaystyle-b\delta+N^{-1}\log\big{\{}(\epsilon+u^{N}_{\epsilon})\big{(}\exp(b)-1\big{)}+1\big{\}}^{N} (229)
=\displaystyle= bδ+log{(ϵ+uϵN)(exp(b)1)+1}\displaystyle-b\delta+\log\big{\{}(\epsilon+u^{N}_{\epsilon})\big{(}\exp(b)-1\big{)}+1\big{\}} (230)

Taking ϵ\epsilon to be sufficiently small, and bb sufficiently large, we obtain (227).

Now, for an integer mLm_{L} to be specified further below, define 𝒦L=mmLm,δm\mathcal{K}_{L}=\bigcap_{m\geq m_{L}}\mathcal{L}_{m,\delta_{m}}. Prokhorov’s Theorem implies that 𝒦L\mathcal{K}_{L} is compact. Employing a union-of-events bound, we obtain that

P𝐳0N(μ^N(𝐲)𝒦L)\displaystyle P^{N}_{\mathbf{z}_{0}}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{K}_{L}\big{)} mmLexp(mN)\displaystyle\leq\sum_{m\geq m_{L}}\exp(-mN) (231)
exp(mLN)supn1j=0exp(jN).\displaystyle\leq\exp(-m_{L}N)\sup_{n\geq 1}\sum_{j=0}^{\infty}\exp(-jN). (232)

We thus find that, for large enough mLm_{L},

supN1N1log(μ^N(𝐲)𝒦L)L,\sup_{N\geq 1}N^{-1}\log\mathbb{P}\big{(}\hat{\mu}^{N}(\mathbf{y})\notin\mathcal{K}_{L}\big{)}\leq-L,

as required. ∎

Lemma 25.

There exists a constant \mathfrak{C} such that for any positive integer mm and any a>0a>0, writing Δm=Tm1\Delta_{m}=Tm^{-1} and ti(m)=Ti/mt^{(m)}_{i}=Ti/m, for any N1N\geq 1,

sup(𝐳0,𝐠0)𝒴NN1logP𝐳0N(sup0tΔmjINsup0im1suppIM|yt+ti(m)p,jyti(m)p,j|2NaΔm)+logma4\displaystyle\sup_{(\mathbf{z}_{0},\mathbf{g}_{0})\in\mathcal{Y}^{N}}N^{-1}\log P^{N}_{\mathbf{z}_{0}}\bigg{(}\sup_{0\leq t\leq\Delta_{m}}\sum_{j\in I_{N}}\sup_{0\leq i\leq m-1}\sup_{p\in I_{M}}\big{|}y^{p,j}_{t+t^{(m)}_{i}}-y^{p,j}_{t^{(m)}_{i}}\big{|}^{2}\geq Na\Delta_{m}\bigg{)}\leq\mathfrak{C}+\log m-\frac{a}{4} (233)
Proof.

Define, for t[0,Δm)t\in[0,\Delta_{m}),

ftN=jINsup0im1suppIM(yt+ti(m)p,jyti(m)p,j)2f^{N}_{t}=\sum_{j\in I_{N}}\sup_{0\leq i\leq m-1}\sup_{p\in I_{M}}\big{(}y^{p,j}_{t+t^{(m)}_{i}}-y^{p,j}_{t^{(m)}_{i}}\big{)}^{2}

Notice that tftNt\to f^{N}_{t} is a submartingale. Thus, writing a=(4Δm)1a=(4\Delta_{m})^{-1}, exp(aftN)\exp\big{(}af^{N}_{t}\big{)} is a submartingale. Therefore, thanks to Doob’s Submartingale Inequality,

(ftNNx)\displaystyle\mathbb{P}\big{(}f^{N}_{t}\geq Nx\big{)} 𝔼[exp(afTNaNx)]\displaystyle\leq\mathbb{E}\bigg{[}\exp\big{(}af^{N}_{T}-aNx\big{)}\bigg{]} (234)
{mM(12Δmσ¯a)1/2}Nexp(aNx)\displaystyle\leq\big{\{}mM\big{(}1-2\Delta_{m}\bar{\sigma}a\big{)}^{-1/2}\big{\}}^{N}\exp(-aNx) (235)
={mM21/2}Nexp(Nx/(4Δm))\displaystyle=\big{\{}mM2^{1/2}\big{\}}^{N}\exp\big{(}-Nx/(4\Delta_{m})\big{)} (236)

References

  • [1] Robert Adler and Jonathan Taylor. Random Fields and Geometry, volume 53. Springer, 2019.
  • [2] Ahmed El Alaoui, Andrea Montanari, and Mark Sellke. Optimization of mean-field spin glasses. Arxiv Preprint, 2020.
  • [3] G Ben Arous and A Guionnet. Large deviations for langevin spin glass dynamics. Probability Theory and Related Fields, 102, 1995.
  • [4] Gerard Ben Arous, Amir Dembo, and Alice Guionnet. Cugliandolo-kurchan equations for dynamics of spin-glasses. Probability Theory and Related Fields, 136:619–660, 2006.
  • [5] Gerard Ben Arous and Alice Guionnet. Symmetric langevin spin glass dynamics. The Annals of Probability, 25:1367–1422, 1997.
  • [6] Gerard Ben Arous, Song Mei, Andrea Montanari, and Mihai Nica. The landscape of the spiked tensor model. Communications on Pure and Applied Mathematics, 72:2282–2330, 11 2019.
  • [7] Nicolas Brunel. Dynamics of sparsely connected networks of excitatory and inhibitory spiking neurons. Journal of Computational Neuroscience, 8:183–208, 2000.
  • [8] Nicolas Brunel and Xiao-Jing Wang. What determines the frequency of fast network oscillations with irregular neural discharges? i. synaptic dynamics and excitation-inhibition balance. Journal of Neurophysiology, 90:415–430, 2003.
  • [9] Amarjit Budhiraja and Paul Dupuis. Analysis and Approximation of Rare Events, volume 94. Springer, 2019.
  • [10] Tanguy Cabana and Jonathan D. Touboul. Large deviations for randomly connected neural networks: I. spatially extended systems. Advances in Applied Probability, 50:983–1004, 2018.
  • [11] Tanguy Cabana and Jonathan D. Touboul. Large deviations for randomly connected neural networks: Ii. state-dependent interactions. Advances in Applied Probability, 50:983–1004, 2018.
  • [12] B. Cessac. Linear response in neuronal networks: from neurons dynamics to collective response. 5 2019.
  • [13] Patrick Charbonneau, Enzo Marinari, Mark Mezard, Giorgio Parisi, Federico Ricci-Tersenghi, Gabriella Sicuro, and Francesco Zamponi, editors. Spin Glass Theory and Far Beyond. World Scientific, 2023.
  • [14] A. Crisanti and H. Sompolinsky. Path integral approach to random neural networks. Physical Review E, 98:1–20, 2018.
  • [15] Andrea Crisanti, Heinz Horner, and H.J. Sommers. The spherical p-spin interaction spin-glass model. Zeitschrift fur Physik B Condensed Matter, 92:257–271, 1993.
  • [16] L. F. Cugliandolo and J. Kurchan. Analytical solution of the off-equilibrium dynamics of a long-range spin-glass model. Physical Review Letters, 71:173–176, 1993.
  • [17] Amir Dembo and Eliran Subag. Dynamics for spherical spin glasses: Disorder dependent initial conditions. Journal of Statistical Physics, 181:465–514, 2020.
  • [18] Amir Dembo and Ofer Zeitouni. Large Deviations Techniques and Applications 2nd Edition. Springer, 1998.
  • [19] Diego Fasoli and Stefano Panzeri. Stationary-state statistics of a binary neural network model with quenched disorder. Entropy, pages 1–30, 2019.
  • [20] Olivier Faugeras and James MacLaurin. Asymptotic description of neural networks with correlated synaptic weights. Entropy, 17:4701–4743, 2015.
  • [21] Olivier Faugeras, Emilie Soret, and Etienne Tanré. Asymptotic behaviour of a network of neurons with random linear interactions. Preprint HAL Id : hal-01986927, 2019.
  • [22] David Gamarnik. The overlap gap property: A topological barrier to optimizing over random structures. Proceedings of the National Academy of Scientists, 2021.
  • [23] M Grunwald. Sanov results for glauber spin-glass dynamics. Probability Theory and Related Fields, 106:187–232, 1996.
  • [24] A Guionnet. Averaged and quenched propagation of chaos for spin glass dynamics. Probability Theory and Related Fields, 109:183–215, 1997.
  • [25] Alice Guionnet and Boguslaw Zegarlinski. Decay to equilibrium in random spin systems on a lattice. Journal of Statistical Physics, 86:899–904, 1997.
  • [26] Moritz Helias and David Dahmen. Statistical Field Theory for Neural Networks. Springer, 2020.
  • [27] Ioannis Karatzas and Steven Shreve. Brownian motion and stochastic calculus 2nd edition, 1991.
  • [28] Itamar Daniel Landau and Haim Sompolinsky. Coherent chaos in a recurrent neural network with structured connectivity. PLoS Computational Biology, 14:1–27, 2018.
  • [29] George Lindgren, Holger Rootzen, and Maria Sandsten. Stationary Stochastic Processes for Scientists and Engineers. Chapman Hall, 2013.
  • [30] Eric Lucon. Quenched large deviations for interacting diffusions in random media. Journal of Statistical Physics, 166:1405–1440, 2017.
  • [31] Olivier Moynot and Manuel Samuelides. Large deviations and mean-field theory for asymmetric random recurrent neural networks. Probability Theory and Related Fields, 123:41–75, 5 2002.
  • [32] Gabriel Koch Ocker, Yu Hu, Michael A. Buice, Brent Doiron, Kresimir Josic, Robert Rosenbaum, and Eric Shea-Brown. From the statistics of connectivity to the statistics of spike times in neuronal networks. Current Opinion in Neurobiology, 46:109–119, 2017.
  • [33] Robert Rosenbaum and Brent Doiron. Balanced networks of spiking neurons with spatially dependent recurrent connections. Physical Review X, pages 1–9, 2014.
  • [34] Robert Rosenbaum, Matthew A. Smith, Adam Kohn, Jonathan E. Rubin, and Brent Doiron. The spatial structure of correlated neuronal variability. Nature Neuroscience, 20:107–114, 2017.
  • [35] Kai Segadlo, Bastian Epping, Alexander Van Meegen, David Dahmen, Michael Krämer, and Moritz Helias. Unified field theoretical approach to deep and recurrent neuronal networks. Journal of Statistical Mechanics: Theory and Experiment, 2022, 10 2022.
  • [36] H. Sompolinsky, A. Crisanti, and H. J. Sommers. Chaos in random neural networks. Physical Review Letters, 61:259–262, 1988.