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

Large Deviations Principles for Coulomb gases at intermediate temperature regime

David Padilla-Garza
Abstract

This paper deals with Coulomb gases at an intermediate temperature regime. We define a local empirical measure and identify a critical temperature scaling. We show that if the scaling of the temperature is supercritical, the local empirical measure satisfies an LDP with an entropy-based rate function. We also show that if the scaling of the temperature is subcritical, the local empirical measure satisfies an LDP with an energy-based rate function. In the critical temperature scaling regime, we derive an LDP-type result in which the "rate function" features the competition of entropy and energy terms.

1 Introduction

Coulomb gases are a system of particles of the same charge that interact via a repulsive kernel, and are confined by an external potential. Let XN=(x1,x2,xN)X_{N}=(x_{1},x_{2},...x_{N}) with xi𝐑dx_{i}\in\mathbf{R}^{d} and let

N(XN)=12ijg(xixj)+Ni=1NV(xi),\mathcal{H}_{N}\left(X_{N}\right)=\frac{1}{2}\sum_{i\neq j}g\left(x_{i}-x_{j}\right)+N\sum_{i=1}^{N}V\left(x_{i}\right), (1)

where

{g(x)=1|x|d2 if d3g(x)=log(|x|) if d=1,2\begin{cases}g(x)=\frac{1}{|x|^{d-2}}\text{ if }d\geq 3\\ g(x)=-\log(|x|)\text{ if }d=1,2\end{cases} (2)

is the Coulomb kernel for d2d\geq 2, i.e. gg satisfies for d2d\geq 2,

Δg=cdδ0,\Delta g=c_{d}\delta_{0}, (3)

where cdc_{d} is a constant that depends only on dd. Often, Coulomb gases at non-zero temperature are considered, these are modeled by a point process whose density is given by the Gibbs measure associated to the Hamiltonian:

d𝐏N,β=1ZN,βexp(βN)dXN,d\mathbf{P}_{N,\beta}=\frac{1}{Z_{{N},\beta}}\exp\left(-\beta\mathcal{H}_{N}\right)dX_{N}, (4)

where

ZN,β=(𝐑d)Nexp(βN)𝑑XN.Z_{N,\beta}=\int_{(\mathbf{R}^{d})^{N}}\exp\left(-\beta\mathcal{H}_{N}\right)d\,X_{N}. (5)

In this notation β\beta is the inverse temperature (which may depend on NN).

As long as 1Nβ,\frac{1}{N}\ll\beta, we have that the empirical measure

empN:=1Ni=1Nδxi{\rm emp}_{N}:=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}} (6)

converges (weakly in the sense of probability measures) almost surely under the Gibbs measure to μV,\mu_{V}, where μV\mu_{V} is the minimizer of the mean-field limit

V(μ)=12𝐑d×𝐑dg(xy)𝑑μ(x)𝑑μ(y)+𝐑dV𝑑μ\mathcal{I}_{V}\left(\mu\right)=\frac{1}{2}\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)d\mu(x)\,d\mu(y)+\int_{\mathbf{R}^{d}}V\,d\mu (7)

among probability measures.

Coulomb gases have a wide range of applications in Statistical Mechanics and Random Matrix Theory, among other areas, see [9, 29] for a more in-depth discussion.

The most fundamental LDP in log gases is found in [3]. Adapting their results to our setting, it was proved that in the regime β=1\beta=1 and d=1,d=1, the push-forward of 𝐏N,β\mathbf{P}_{N,\beta} by empN{\rm emp}_{N} satisfies an LDP at speed N2N^{2} with rate function given by

(μ)=V(μ)V(μV).\mathcal{F}(\mu)=\mathcal{I}_{V}\left(\mu\right)-\mathcal{I}_{V}\left(\mu_{V}\right). (8)

This result was originally motivated by Hermitian Random Matrix Theory.

An analogous statement in dimension 22 was proved in [7]. In [15] the authors deal with a general repulsive interaction gg in dimension d1.d\geq 1. In [26], the authors derive an LDP for the eigenvalues of some non-symmetric random matrices.

As mentioned before, the regime β=CN\beta=\frac{C}{N} is substantially different since the empirical measure does not converge to the equilibrium measure. We call this the high temperature regime. Nevertheless, it is possible to identify the limit of the empirical measures as the thermal equilibrium measure:

μβ=argminμV(μ)+1Cent[μ],\mu_{\beta}=\operatorname{argmin}_{\mu}\mathcal{I}_{V}\left(\mu\right)+\frac{1}{C}\mbox{ent}[\mu], (9)

where ent{\rm ent} is given by Definition 2.2, and the minimum is taken over probablity measures. Moreover, the push-forward of the Gibbs measure 𝐏N,β\mathbf{P}_{N,\beta} by the empirical measure satisfies an LDP at speed CNCN with rate function

(μ)=V(μ)+1Cent[μ](V(μβ)+1Cent[μβ]).\mathcal{F}(\mu)=\mathcal{I}_{V}\left(\mu\right)+\frac{1}{C}\rm{ent}[\mu]-\left(\mathcal{I}_{V}\left(\mu_{\beta}\right)+\frac{1}{C}\rm{ent}[\mu_{\beta}]\right). (10)

This result can be found in [17], which also treats Coulomb gases on compact manifolds.

In our setting, the intensity and sign of the charge of the particles are fixed; in reference, [8], however, the authors also consider the case of the intensity of the charges being a random variable which takes values in {1,1}\{-1,1\}. Having positive and negative charges implies that there are attractive interactions, which are harder to deal with.

A widely studied question in Coulomb gases is that of the fluctuations of the difference between empN{\rm emp}_{N} and μV.\mu_{V}. In order to understand these fluctuations, it is convenient to multiply this difference by a test function φ,\varphi, the resulting object is called the first order statistic:

FluctN(φ)=Nφd(empNμV).\mbox{Fluct}_{N}(\varphi)=N\int\varphi\,d\left({\rm emp}_{N}-\mu_{V}\right). (11)

In [23] it was proved that in two dimensions (under mild technical additional conditions) FluctN(φ)\mbox{Fluct}_{N}(\varphi) converges in law to a Gaussian random variable with mean

mean=12π(1β14)𝐑2Δφ(𝟏Σ+(logΔV)Σ)\mbox{mean}=\frac{1}{2\pi}\left(\frac{1}{\beta}-\frac{1}{4}\right)\int_{\mathbf{R}^{2}}\Delta\varphi\left(\mathbf{1}_{\Sigma}+(\log\Delta V)^{\Sigma}\right) (12)

and variance

Var=12πβ𝐑2|φΣ|2.\mbox{Var}=\frac{1}{2\pi\beta}\int_{\mathbf{R}^{2}}|\nabla\varphi^{\Sigma}|^{2}. (13)

In this notation, Σ\Sigma is the support of the equilibrium measure, and gΣg^{\Sigma} is the harmonic extension of gg outside Σ,\Sigma, i.e. the only continuous function which agrees with gg in Σ\Sigma up to the boundary and is harmonic and bounded in 𝐑2Σ.\mathbf{R}^{2}\setminus\Sigma. A related, very similar result was obtained simultaneously in [5]. Analogous results were obtained in one dimension in [6] and [21], generalizing the work of [20], [32], and [10]. In [4], the authors derive local laws and moderate deviation bounds. In [19], the authors derive a CLT for linear statistics of β\beta-ensembles at high temperature, in this reference, the authors also derive concentration inequalities. In [31], the author deals with linear statistics replacing μV\mu_{V} with the thermal equilibrium measure. All of the references just mentioned, except [31] and [19] deal with β\beta proportional to N12dN^{1-\frac{2}{d}} for d2d\geq 2 (in our notation), or β\beta constant in d=1d=1. This paper wanders into the mainly unexplored territory of Coulomb gases at general temperature regimes and high dimensions.

Coulomb gases are also widely studied due to their connection with Random Matrix Theory, see [13, 14, 12, 11]for recent developments. Most of the problems studied in connection with Random Matrix Theory are in dimensions 11 and 22, therefore the result in this paper (which holds only in dimensions 33 and higher) is not applicable to that setting. However, the main result in this paper is applicable to quaternionic Gaussian ensembles.

Since the equilibrium measure typically has compact support, there are NN particles in a bounded domain in 𝐑d,\mathbf{R}^{d}, and so, typically the particles are at distance N1dN^{-\frac{1}{d}} of each other. After applying a dilation of magnitude N1dN^{\frac{1}{d}} to Euclidean space, one observes individual particles. An LDP at speed NN for Coulomb gases at this scale was obtained in [22], and the rate function combines two terms: one comes from the Hamiltonian and the other one is related to entropy. Similar results were obtained in [18] for hyper-singular Riesz gases, and in [24] for two-component plasmas.

Details of the convergence of empN{\rm emp}_{N} to μV\mu_{V} were obtained in [16]. In this reference, the authors also study the relation between the electric energy and norms on probability measures. One of their results concerning the convergence of empN{\rm emp}_{N} to μV\mu_{V} is the following: If β>0\beta>0, then under mild additional assumptions, there exist constants u,v>0u,v>0 depending on β\beta and VV only such that, for any N2N\geq 2 and

r{vlogNN if d=2vN1d if d>2r\geq\begin{cases}v\sqrt{\frac{\log N}{N}}\mbox{ if }d=2\\ vN^{-\frac{1}{d}}\mbox{ if }d>2\end{cases} (14)

we have

𝐏N,β(W1(μ,ν)r)exp(uN2r2),\mathbf{P}_{N,\beta}(W_{1}(\mu,\nu)\geq r)\leq\exp(-uN^{2}r^{2}), (15)

where W1W_{1} is the Wasserstein distance (see [16]).

The main contribution of this paper is to clarify the relationship between temperature scales and length scales for the mesoscopic behavior of particle systems with Coulomb interactions. The way to do this is to look at rare events at a mesoscale and understand them by means of an LDP. An important idea in this work is to exploit the different scaling relations satisfied by the Coulomb energy and the entropy. This work also exploits the smearing technique, used for example in [16, 22, 27, 28]. A large part of this work is devoted to simplifying expressions for partition functions, derived via a variational characterization.

At a macroscopic scale, it has been well-known that the temperature and the energy compete if the energy is of order β=1N\beta=\frac{1}{N}, in the sense that the empirical field (the macroscopic observable) satisfies an LDP that involves energy and entropy terms. At a microscopic scale, it was recently proved in [22] that the energy and the entropy compete if the temperature is of order β=N2dd\beta=N^{\frac{2-d}{d}} (for d3d\geq 3), in the sense that the tagged empirical field (the microscopic observable) satisfies an LDP that involves energy and entropy terms. This raises the natural question: given a length scale between the macroscopic and microscopic, is there a temperature regime in which the temperature and entropy compete (in the sense that there is an LDP containing energy and entropy terms)? This paper answers this question.

Given a length scale, we will identify a critical temperature regime. Of course, this problem is equivalent to identifying a critical length scale given a temperature regime. This last approach was taken in [1]. However, despite analyzing the interplay between temperature and length scale, this work and [1] are pretty much independent. [1] deals with the tagged empirical field: an observable obtained by averaging the empirical field over a certain region. This observable is fundamentally different from the local empirical measure. Furthermore, the ρβ\rho_{\beta} identified in [1] does not coincide with the "critical length scale" identified in our work. The main results in this work and in [1] are independent: our results do not follow from [1], and the results in [1] do not follow from our results. The techniques used are also fundamentally different. This work is not an attempt to prove any conjecture in [1].

A significant part of this work is devoted to computing, with high precision, certain partition functions. This is similar to obtaining a Laplace principle, as in [17]. However, the techniques in [17] would only allow us to obtain the leading order term in the partition function. This would be greatly insufficient to conclude, and so it is necessary to take a different approach in order to identify the next-order terms.

2 Main definitions and statement of main results

This section defines the most important objects for the rest of the paper and states the main results.

We begin with definitions related to the empirical measure.

Definition 2.1.

Given XN𝐑d×N,X_{N}\in\mathbf{R}^{d\times N}, with

XN=(x1,xN),X_{N}=(x_{1},...x_{N}), (16)

we denote

empN(XN)=1Ni=1Nδxi.{\rm emp}_{N}(X_{N})=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}. (17)

In order to make the notation more clear, we will often write

empN{\rm emp}_{N} (18)

instead of

empN(XN).{\rm emp}_{N}(X_{N}). (19)

Given x𝐑dx\in\mathbf{R}^{d} and R𝐑+R\in\mathbf{R}^{+} we denote by

(x,R)=(R2,R2)d+x.\square(x,R)=\left(-\frac{R}{2},\frac{R}{2}\right)^{d}+x. (20)

We will also use the notation

R=(0,R).\square_{R}=\square(0,R). (21)

Let

xiλ=Nλxi.x_{i}^{\lambda}=N^{\lambda}x_{i}. (22)

We now define the main observable of this paper: the local empirical measure

lempNλ(XN)=1N1λdi=1Nδxiλ|R.{\rm lemp}_{N}^{\lambda}(X_{N})=\frac{1}{N^{1-\lambda d}}\sum_{i=1}^{N}\delta_{x_{i}^{\lambda}}|_{\square_{R}}. (23)

Even though lempNλ(XN){\rm lemp}_{N}^{\lambda}(X_{N}) depends on λ\lambda and XNX_{N}, we will sometimes omit this dependence for ease of notation and simply write lempN{\rm lemp}_{N}. Note that lempN{\rm lemp}_{N} is a measure with support contained in R,\square_{R}, and with mass which we expect remains bounded if XNX_{N} is distributed according to 𝐏N,β\mathbf{P}_{N,\beta}.

This paper deals with the empirical measure at a mesoscopic scale, i.e. at a scale Nλ,N^{-\lambda}, where

λ(0,1d).\lambda\in\left(0,\frac{1}{d}\right). (24)

We choose the name mesoscopic because the scale λ=0\lambda=0 is macroscopic, while the scale λ=1d\lambda=\frac{1}{d} is microscopic, i.e. a scale which is of the same order of magnitude as the distance between particles. Without loss of generality, we assume that we blow up around the origin. For the general case, we may simply consider a modified potential.

The idea is to define a mesoscopic observable. This definition is inspired by interpolating between the empirical measure: 1Ni=1Nδxi\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}, and the empirical field: i=1NδxiN1d\sum_{i=1}^{N}\delta_{x_{i}^{N^{\frac{1}{d}}}}. The factor of NλN^{\lambda} in the dilation corresponds to a mesoscale, while the normalizing factor of N1λdN^{1-\lambda d} is necessary to obtain a bounded nonzero quantity. Note that the local empirical measure is, in general, not a probability measure but only a positive measure. The local empirical measure is more similar to the empirical measure in the sense that it converges to a continuous measure.

We now define the most basic functionals used in the paper: energy and entropy. We also define a few modifications of the functionals which will be used throughout the paper.

Definition 2.2.

Given two measures μ\mu and ρ\rho, the relative entropy ent[μ|ρ]{\rm ent}[\mu|\rho] is defined as

ent[μ|ρ]={log(dμdρ)𝑑μifμρ o.w.{\rm ent}[\mu|\rho]=\begin{cases}\int\log\left(\frac{d\mu}{d\rho}\right)\,d{\mu}\quad if\mu\ll\rho\\ \infty\quad\text{ o.w.}\end{cases} (25)

The entropy of a measure ent[μ]{\rm ent}[\mu] is defined as ent[μ|]{\rm ent}[\mu|\mathcal{L}], where \mathcal{L} denotes the Lebesgue measure on 𝐑d.\mathbf{R}^{d}.

In the remainder of the paper, we commit the abuse of notation of not distinguishing between a measure and its density.

Definition 2.3.

The electric energy of a measure ν\nu is defined as

(ν)=𝐑d×𝐑dg(xy)𝑑ν(x)𝑑ν(y),\mathcal{E}(\nu)=\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)d\nu(x)\,d\nu(y), (26)

and

(ν)=𝐑d×𝐑dΔg(xy)𝑑ν(x)𝑑ν(y),\mathcal{E}^{\neq}(\nu)=\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}\setminus\Delta}g(x-y)d\nu(x)\,d\nu(y), (27)

where

Δ={(x,x)𝐑d×𝐑d}.\Delta=\{(x,x)\in\mathbf{R}^{d}\times\mathbf{R}^{d}\}. (28)

Given a measurable set Ω𝐑d\Omega\subset\mathbf{R}^{d}, we will also use the notation

Ω(ν)=𝐑d×𝐑dΔΩg(xy)𝑑ν(x)𝑑ν(y),\mathcal{E}^{\neq}_{\Omega}(\nu)=\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}\setminus\Delta_{\Omega}}g(x-y)d\nu(x)\,d\nu(y), (29)

where

ΔΩ={(x,x)Ω×Ω}.\Delta_{\Omega}=\{(x,x)\in\Omega\times\Omega\}. (30)
Definition 2.4.

We define the free energy of a measure as

β(μ)=V(μ)+1Nβ𝐑dμlog(μ).\mathcal{E}_{\beta}\left(\mu\right)=\mathcal{I}_{V}\left(\mu\right)+\frac{1}{N\beta}\int_{\mathbf{R}^{d}}\mu\log\left(\mu\right). (31)

We define the thermal equilibrium measure μβ\mu_{\beta} as

μβ=argminμ𝒫(𝐑d)β(μ),\mu_{\beta}=\operatorname{argmin}_{\mu\in\mathcal{P}(\mathbf{R}^{d})}\mathcal{E}_{\beta}(\mu), (32)

where 𝒫(𝐑d)\mathcal{P}(\mathbf{R}^{d}) denotes the set of probability measures on 𝐑d\mathbf{R}^{d}. More generally, we will use the notation 𝒫(Ω)\mathcal{P}(\Omega) for the set of probability measures on Ω𝐑d\Omega\subset\mathbf{R}^{d}.

For existence, uniqueness and basic properties of μβ\mu_{\beta} see [2].

We also define the equilibrium measure μV\mu_{V} by

μV=argminμ𝒫(𝐑d)V(μ),\mu_{V}=\operatorname{argmin}_{\mu\in\mathcal{P}(\mathbf{R}^{d})}\mathcal{I}_{V}\left(\mu\right), (33)

where V\mathcal{I}_{V} is given by equation (7).

For existence, uniqueness and basic properties of μV\mu_{V} see for example [30] and references therein.

We proceed with a few definitions regarding measures.

Definition 2.5.

Given a measurable set Ω𝐑d\Omega\subset\mathbf{R}^{d}, we denote (Ω)\mathcal{M}(\Omega) the space of measures on Ω\Omega which are either of bounded variation or have a definite sign. We also define, for any μ(Ω)\mu\in\mathcal{M}(\Omega)

|μ|=μ(Ω).|\mu|=\mu(\Omega). (34)
Definition 2.6.

Given a measurable set Ω\Omega, and a measure μ\mu on Ω\Omega, we define the bounded Lipschitz norm of μ\mu, denoted μBL\|\mu\|_{BL} as

μBL=supfLip1(Ω)𝐑df𝑑μ,\|\mu\|_{BL}=\sup_{f\in{\rm Lip}_{1}(\Omega)}\int_{\mathbf{R}^{d}}f\,d\mu, (35)

where Lip1(Ω){\rm Lip}_{1}(\Omega) denotes the set of Lipschitz functions on Ω\Omega whose absolute value is bounded by 11, and Lipschitz constant is also smaller than 11.

Unless otherwise specified, any distance between measures will refer to the bounded Lipschitz norm. In particular, given ϵ>0\epsilon>0 we define

B(ν,ϵ)={μ(Ω)|μνBLϵ}.B(\nu,\epsilon)=\{\mu\in\mathcal{M}(\Omega)|\|\mu-\nu\|_{BL}\leq\epsilon\}. (36)

We recall that the bounded Lipschitz norm metricizes the topology of weak convergence.

Definition 2.7.

Let μ,ν+(Ω),\mu,\nu\in\mathcal{M}^{+}(\Omega), we define

𝒩[μ|ν]=ent[μ|ν]+|ν||μ|,\mathcal{N}[\mu|\nu]={\rm ent}[\mu|\nu]+|\nu|-|\mu|, (37)

where +(Ω)\mathcal{M}^{+}(\Omega) denotes the set of positive measures on a set Ω\Omega.

We will now introduce the rate functions for the different LDP’s. These rate functions are based on the entropy functional, the energy functional, or both.

Definition 2.8.

Given a domain Ω𝐑d\Omega\subset\mathbf{R}^{d} and a scalar α𝐑+,\alpha\in\mathbf{R}^{+}, we define the function ΦΩα,\Phi^{\alpha}_{\Omega}, defined for an absolutely continuous measure μ\mu on Ω\Omega as

ΦΩα(μ)=infρ:𝐑dΩ𝐑+𝐑d×𝐑dg(xy)(μ(x)+ρ(x)α)𝑑x(μ(y)+ρ(y)α)𝑑y.\Phi^{\alpha}_{\Omega}(\mu)=\inf_{\rho:\mathbf{R}^{d}\setminus\Omega\to\mathbf{R}^{+}}\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)(\mu(x)+\rho(x)-\alpha)dx(\mu(y)+\rho(y)-\alpha)dy. (38)
Definition 2.9.

Given a measure ν\nu, we define the measure ντ\nu^{\tau} as

ντ(Ω)=τdν(τ1Ω).\nu^{\tau}(\Omega)=\tau^{d}\nu(\tau^{-1}\Omega). (39)

For a measure μ\mu defined on Ω,\Omega, we denote by

𝐓λN(ν)=infρ+𝐑dΩ(12(ν+ρμβNλ)𝐑dlog(μβNλ)𝑑ρ+ent[ρ]),\mathbf{T}_{\lambda}^{N}(\nu)=\inf_{\rho\in\mathcal{M}^{+}\mathbf{R}^{d}\setminus\Omega}\left(\frac{1}{2}\mathcal{E}\left(\nu+\rho-\mu_{\beta}^{N^{\lambda}}\right)-\int_{\mathbf{R}^{d}}\log\left(\mu_{\beta}^{N^{\lambda}}\right)d\rho+{\rm ent}[\rho]\right), (40)

where the infimum is taken over ρ\rho such that

𝐑dν+ρμβNλ=0.\int_{\mathbf{R}^{d}}\nu+\rho-\mu_{\beta}^{N^{\lambda}}=0. (41)

We also define

𝒯λN(ν)=𝐓λN(ν)+ent[ν|μV(0)𝟏Ω].\mathcal{T}_{\lambda}^{N}(\nu)=\mathbf{T}_{\lambda}^{N}(\nu)+{\rm ent}[\nu|\mu_{V}(0)\mathbf{1}_{\Omega}]. (42)

In this paper, we deal with general a general potential VV. We only impose some regularity and growth conditions, which we make precise in the next definition.

Definition 2.10.

We call a potential V:𝐑d𝐑V:\mathbf{R}^{d}\to\mathbf{R}, with d3d\geq 3 admissible if:

  • 1.

    VC2V\in C^{2}.

  • 2.
    limxV(x)=.\lim_{x\to\infty}V(x)=\infty. (43)
  • 3.
    |x|>1exp(αV(x))𝑑x<\int_{|x|>1}\exp\left(-\alpha V(x)\right)\,dx<\infty (44)

    for all α>0\alpha>0.

  • 4.
    ΔVα>0\Delta V\geq\alpha>0 (45)

    for some α\alpha, in a neighborhood of Σ\Sigma, defined as the support of μV\mu_{V}.

  • 5.

    V|ΣC(Σ)V|_{\Sigma}\in C^{\infty}(\Sigma).

  • 6.

    0int(Σ)0\in{\rm int}(\Sigma).

Remark 2.1.

If VV is admissible, the equilibrium measure μV\mu_{V} is bounded and has compact support, see [30].

Finally, before stating the main result, we recall the definition of rate function and Large Deviations Principle (LDP).

Definition 2.11.

(Rate function) Let XX be a metric space (or a topological space). A rate function is a l.s.c. function I:X[0,],I:X\to[0,\infty], it is called a good rate function if its sublevel sets are compact.

Definition 2.12 (LDP).

Let PNP_{N} be a sequence of Borel probability measures on XX and aNa_{N} a sequence of positive reals such that aN.a_{N}\to\infty. Let II be a good rate function on X.X. The sequence PNP_{N} is said to satisfy a Large Deviations Principle (LDP) at speed aNa_{N} with (good) rate function II if for every Borel set EXE\subset X the following inequalities hold:

infEoIlim infN1aNlog(PN(E))lim supN1aNlog(PN(E))infE¯I,-\inf_{E^{\mathrm{o}}}I\leq\liminf_{N\to\infty}\frac{1}{a_{N}}\log\left(P_{N}(E)\right)\leq\limsup_{N\to\infty}\frac{1}{a_{N}}\log\left(P_{N}(E)\right)\leq-\inf_{\overline{E}}I, (46)

where EoE^{\mathrm{o}} and E¯\overline{E} denote respectively the interior and the closure of a set E.E. Formally, this means that PN(E)exp(aNinfEI).P_{N}(E)\simeq\exp(-a_{N}\inf_{{E}}I).

The main result of this paper is the following theorem:

Theorem 2.1.

Assume that d3d\geq 3 and the potential VV are admissible. Let β=Nγ\beta=N^{-\gamma} with γ(d2d,1).\gamma\in(\frac{d-2}{d},1). Assume that μV\mu_{V} is bounded away from 0 inside its support. Let

γ=12λ,\gamma^{*}=1-2\lambda, (47)

and assume that

λ<1d(d+2).\lambda<\frac{1}{d(d+2)}. (48)

Then:

  • \bullet

    If γ<γ\gamma<\gamma^{*} (subcritical regime) then the push-forward of 𝐏N,β\mathbf{P}_{N,\beta} by lempN{\rm lemp}_{N} satisfies an LDP in the topology of weak convergence at speed βN2(d+2)λ\beta N^{2-(d+2)\lambda} and rate function

    12ΦRμV(0)(μ).\frac{1}{2}\Phi^{\mu_{V}(0)}_{\square_{R}}(\mu). (49)
  • \bullet

    If γ>γ\gamma>\gamma^{*} (supercritical regime) then the push-forward of 𝐏N,β\mathbf{P}_{N,\beta} by lempN{\rm lemp}_{N} satisfies an LDP in the topology of weak convergence at speed N1λdN^{1-\lambda d} and rate function

    𝒩[μ|μV(0)𝟏R].\mathcal{N}[\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]. (50)
  • \bullet

    If γ=γ\gamma=\gamma^{*} (critical regime) and νL\nu\in L^{\infty} then

    limϵ0lim supN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))=0.\lim_{\epsilon\to 0}\limsup_{N\to\infty}\left(\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))\right)+\mathcal{T}^{N}_{\lambda}(\nu)\right)=0. (51)

    Similarly,

    limϵ0lim infN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))=0.\lim_{\epsilon\to 0}\liminf_{N\to\infty}\left(\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))\right)+\mathcal{T}^{N}_{\lambda}(\nu)\right)=0. (52)
Remark 2.2.

The rate functions have the same minimizer in all cases: μV(0)𝟏R\mu_{V}(0)\mathbf{1}_{\square_{R}}. This is clear, since in this temperature regime, the empirical measure concentrates on the thermal equilibrium measure for all scales larger than N1dN^{-\frac{1}{d}}, as was proved in [25]. The typical event, therefore, is trivial; and it is a rare event that deserves to be looked at.

Remark 2.3.

Nearly all hypotheses in Theorem 2.1 are essential. The hypothesis that λ<1d(d+2)\lambda<\frac{1}{d(d+2)}, however is not. It is an artifact of the proof, and it is needed to bound a specific error term. Bounding this error term is necessary if one uses the regularization procedure, i.e. it is needed to bound the difference between the energy of a discrete probability measure and a continuous one. This technique is very common in the field. We expect that a similar result will be true for λ1d(d+2)\lambda\geq\frac{1}{d(d+2)}, but proving this would require essentially different techniques.

Remark 2.4.

It is natural to ask if there is an analog of Theorem 2.1 in the extreme cases

β=N1,λ=0andβ=Nd2d,λ=1d.\beta=N^{-1},\lambda=0\quad\text{and}\quad\beta=N^{-\frac{d-2}{d}},\lambda=\frac{1}{d}. (53)

In the case

β=N1,λ=0,\beta=N^{-1},\lambda=0, (54)

Theorem 2.1 has a very natural generalization, as mentioned in the introduction. It was proved in [17] that empN{\rm emp}_{N} satisfies an LDP at speed NN with rate function

(μ)=V(μ)+ent[μ](V(μβ)+ent[μβ]).\mathcal{F}(\mu)=\mathcal{I}_{V}\left(\mu\right)+\rm{ent}[\mu]-\left(\mathcal{I}_{V}\left(\mu_{\beta}\right)+\rm{ent}[\mu_{\beta}]\right). (55)

The case

β=Nd2d,λ=1d\beta=N^{-\frac{d-2}{d}},\lambda=\frac{1}{d} (56)

is substantially different, because at a microscopic scale, we do not observe a continuous distribution but rather individual particles. A similar problem was treated in [22]. Even though the result is substantially different, it has a similar flavor, since the authors prove an LDP at speed NN in which the rate function contains the sum of an entropy term and an electric energy term.

3 Additional definitions

We proceed with a few additional definitions related to the empirical measure.

Definition 3.1.

Let R𝐑+R\in\mathbf{R}^{+} be fixed, and XN𝐑d×N.X_{N}\in\mathbf{R}^{d\times N}. We define yi,zjy_{i},z_{j} such that

XN=(y1,y2yiN,z1,z2,zjN),X_{N}=(y_{1},y_{2}...y_{i_{N}},z_{1},z_{2},...z_{j_{N}}), (57)

where

ykRNλ,zkRNλ,y_{k}\in\square_{\frac{R}{N^{\lambda}}},\quad z_{k}\notin\square_{\frac{R}{N^{\lambda}}}, (58)

and

iN+jN=N.i_{N}+j_{N}=N. (59)

Let

YN=(y1,yiN)Y_{N}=(y_{1},...y_{i_{N}}) (60)

and

empN(YN)=1Nk=1iNδyk.{\rm emp}_{N}^{\prime}(Y_{N})=\frac{1}{N}\sum_{k=1}^{i_{N}}\delta_{y_{k}}. (61)

Similarly, let

ZN=(z1,ziN).Z_{N}=(z_{1},...z_{i_{N}}). (62)
Definition 3.2.

Given an integer M,M, and ϵ>0,\epsilon>0, we denote by 𝒜Mϵ(Ω)\mathcal{A}_{M}^{\epsilon}(\Omega) the set of measures which are purely atomic with weight ϵ,\epsilon, i.e.

𝒜Mϵ(Ω)={μ+(Ω)|μ=ϵi=1Mδxi}.\mathcal{A}_{M}^{\epsilon}(\Omega)=\{\mu\in\mathcal{M}^{+}(\Omega)|\mu=\epsilon\sum_{i=1}^{M}\delta_{x_{i}}\}. (63)

Given a measure μ+(𝐑d),\mu\in\mathcal{M}^{+}(\mathbf{R}^{d}), an integer M,M, a region Ω𝐑d\Omega\subset\mathbf{R}^{d} and ϵ>0,\epsilon>0, we define

𝐖Ω,ϵM,μ(ν)=infρ𝒜Mϵ(𝐑dΩ)(ν+ρμ),\mathbf{W}_{\Omega,\epsilon}^{M,\mu}(\nu)=\inf_{\rho\in\mathcal{A}_{M}^{\epsilon}(\mathbf{R}^{d}\setminus\Omega)}\mathcal{E}^{\neq}(\nu+\rho-\mu), (64)

where ν+(Ω).\nu\in\mathcal{M}^{+}(\Omega).

We also define

𝐓λN,(ν)=infρ+𝐑dΩ(12R(ν+ρμβNλ)𝐑dlog(μβNλ)𝑑ρ+ent[ρ]),\mathbf{T}_{\lambda}^{N,\neq}(\nu)=\\ \inf_{\rho\in\mathcal{M}^{+}\mathbf{R}^{d}\setminus\Omega}\Bigg{(}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{R}}\left(\nu+\rho-\mu_{\beta}^{N^{\lambda}}\right)-\int_{\mathbf{R}^{d}}\log\left(\mu_{\beta}^{N^{\lambda}}\right)d\rho+\mbox{ent}[\rho]\Bigg{)}, (65)

where the infimum is taken over ρ\rho such that

𝐑dν+ρμβNλ=0.\int_{\mathbf{R}^{d}}\nu+\rho-\mu_{\beta}^{N^{\lambda}}=0. (66)

The definition of 𝐓λN,\mathbf{T}_{\lambda}^{N,\neq} is almost the same as 𝐓λN\mathbf{T}_{\lambda}^{N} but omitting the diagonal inside the square R\square_{R} in the computation of the Coulomb energy. This modification allows for the quantity to be finite for atomic measures inside the cube.

Given α𝐑+\alpha\in\mathbf{R}^{+} we also define

𝐅Rα(ν)=infρ+(𝐑dR)(ν+ρα),\mathbf{F}_{\square_{R}}^{\alpha}(\nu)=\inf_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\nu+\rho-\alpha), (67)

where the inf is taken over all ρC\rho\in C^{\infty} such that

𝐑dρ+ναdx=0.\int_{\mathbf{R}^{d}}\rho+\nu-\alpha\,dx=0. (68)

We generalize the definition of ΦΩα(ν)\Phi_{\Omega}^{\alpha}(\nu) to a more general setting in which the background measure is not necessarily constant out of Ω\Omega. Given a set Ω𝐑d\Omega\in\mathbf{R}^{d} and a background measure μ(𝐑d),\mu\in\mathcal{M}(\mathbf{R}^{d}),

ΦΩμ(ν)=infρ+(𝐑dΩ)(ν+ρμ).\Phi_{\Omega}^{\mu}(\nu)=\inf_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\Omega)}\mathcal{E}(\nu+\rho-\mu). (69)

We now define an analog of ΦΩμ\Phi^{\mu}_{\Omega} for measures that are not absolutely continuous. Given a measurable set Ω𝐑d\Omega\subset\mathbf{R}^{d}, a positive measure μ\mu on 𝐑d\mathbf{R}^{d}, ΦΩ,μ(ν)\Phi_{\Omega,\neq}^{\mu}(\nu) is defined for a measure ν\nu on Ω\Omega as

ΦΩ,μ(ν)=infρ+(𝐑dΩ)Ω(ν+ρμ).\Phi_{\Omega,\neq}^{\mu}(\nu)=\inf_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\Omega)}\mathcal{E}^{\neq}_{\Omega}(\nu+\rho-\mu). (70)

We also introduce the notation.

𝒢(μ,ν)=𝐑d×𝐑dg(xy)𝑑μ(x)𝑑ν(y).\mathcal{G}(\mu,\nu)=\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)d\mu({x})d\nu({y}). (71)
Remark 3.1.

Note that

(empN(YN))Nλ=lempN.({\rm emp}_{N}^{\prime}(Y_{N}))^{N^{\lambda}}={\rm lemp}_{N}. (72)

Note also that for any α𝐑+\alpha\in\mathbf{R}^{+}, \mathcal{E} has the scaling relation

(μα)=αd+2(μ),\mathcal{E}(\mu^{\alpha})=\alpha^{d+2}\mathcal{E}(\mu), (73)

and therefore 𝐖Ω,ϵM,μ(ν)\mathbf{W}^{M,\mu}_{\Omega,\epsilon}(\nu) has the scaling relation

𝐖Ω,ϵM,μ(ν)=α(d+2)𝐖αΩ,αdϵM,μα(να).\mathbf{W}^{M,\mu}_{\Omega,\epsilon}(\nu)=\alpha^{-(d+2)}\mathbf{W}^{M,\mu^{\alpha}}_{\alpha\Omega,\alpha^{d}\epsilon}(\nu^{\alpha}). (74)

Lastly, we introduce notation that will be used throughout the work.

Remark 3.2 (Notation).

Given ϵ>0\epsilon>0, we denote by λϵ\lambda_{\epsilon} the uniform probability measure on B(0,ϵ)\partial B(0,\epsilon).

4 Preliminary results

In this section, we will prove some preliminary results needed for the main Theorem.

We begin with a splitting formula around the thermal equilibrium measure, which is an analog of the usual splitting formula (see for example [28]).

Proposition 4.1.

The Hamiltonian N\mathcal{H}_{N} can be split into:

N(XN)=N2β(μβ)+Ni=1Nζβ(xi)+N22(empNμβ),\mathcal{H}_{N}\left(X_{N}\right)=N^{2}\mathcal{E}_{\beta}\left(\mu_{\beta}\right)+N\sum_{i=1}^{N}\zeta_{\beta}\left(x_{i}\right)+\frac{N^{2}}{2}\mathcal{E}^{\neq}\left({\rm emp}_{N}-\mu_{\beta}\right), (75)

where

ζβ=1Nβlog(μβ).\zeta_{\beta}=-\frac{1}{N\beta}\log\left(\mu_{\beta}\right). (76)
Proof.

See [1]. ∎

Definition 4.1.

In analogy with previous work in this field [1, 23, 6, 22], we define a next order partition function KN,β,K_{N,\beta}, as

KN,β=ZN,βexp(N2ββ(μβ)).K_{N,\beta}=Z_{N,\beta}\exp\left(-N^{2}\beta\mathcal{E}_{\beta}\left(\mu_{\beta}\right)\right). (77)

Using (77), we may rewrite the Gibbs measure as

d𝐏N,β(x1xN)=1KN,βexp(12N2β(empNμβ))Πi=1Nμβ(xi)dxi.d\mathbf{P}_{N,\beta}(x_{1}...x_{N})=\frac{1}{K_{N,\beta}}\exp\left(-\frac{1}{2}N^{2}\beta\mathcal{E}^{\neq}({\rm emp}_{N}-\mu_{\beta})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}. (78)

We also need the following piece of information about μβ\mu_{\beta}, which can be deduced from [2], Theorem 1.

Remark 4.1.

Let T>0,λ>0T>0,\lambda>0 and assume that limNNβ=\lim_{N\to\infty}N\beta=\infty, and the potential VV is admissible, then

μβNλμV(0)L(B(0,T))0.\|\mu_{\beta}^{N^{\lambda}}-\mu_{V}(0)\|_{L^{\infty}(B(0,T))}\to 0. (79)

We proceed to prove some elementary properties about the rate functions in Theorem 2.1.

Claim 4.1.

For any α,R>0\alpha,R>0, the function 𝒩(ν|α𝟏R)\mathcal{N}(\nu|\alpha\mathbf{1}_{\square_{R}}) is a convex (in ν\nu) rate function.

Proof.

Since convexity and l.s.c. are immediate from the convexity and l.s.c. of ent, we need only show that 𝒩(ν|α𝟏R)\mathcal{N}(\nu|\alpha\mathbf{1}_{\square_{R}}) is positive for any ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}). Throughout the proof, we will use the notation

ν¯=1RdRν𝑑x.\overline{\nu}=\frac{1}{R^{d}}\int_{\square_{R}}\nu dx. (80)

Using Jensen’s inequality, the convexity of xxlog(x),x\mapsto x\log(x), and doing a first-order Taylor expansion of xlog(x),x\log(x), we have

𝒩(ν|α𝟏R)=Rlog(να)ν𝑑x+Rdα|ν|=Rlog(να)ναα𝑑x+Rdα|ν|Rlog(ν¯α)ν¯αα𝑑x+Rdα|ν|=Rdαlog(ν¯α)ν¯α+Rdα|ν|Rdα(ν¯α1)+Rdα|ν|=0.\begin{split}\mathcal{N}(\nu|\alpha\mathbf{1}_{\square_{R}})&=\int_{\square_{R}}\log\left(\frac{\nu}{\alpha}\right)\nu dx+R^{d}\alpha-|\nu|\\ &=\int_{\square_{R}}\log\left(\frac{\nu}{\alpha}\right)\frac{\nu}{\alpha}\alpha dx+R^{d}\alpha-|\nu|\\ &\geq\int_{\square_{R}}\log\left(\frac{\overline{\nu}}{\alpha}\right)\frac{\overline{\nu}}{\alpha}\alpha dx+R^{d}\alpha-|\nu|\\ &=R^{d}\alpha\log\left(\frac{\overline{\nu}}{\alpha}\right)\frac{\overline{\nu}}{\alpha}+R^{d}\alpha-|\nu|\\ &\geq R^{d}\alpha\left(\frac{\overline{\nu}}{\alpha}-1\right)+R^{d}\alpha-|\nu|\\ &=0.\end{split} (81)

The following claim is standard and can be found, for example, in [25].

Lemma 4.1.

The energy \mathcal{E} is l.s.c. w.r.t. to weak H1H^{-1} convergence.

With the help of Lemma 4.1, we can prove some elementary properties about ΦRα\Phi_{\square_{R}}^{\alpha}.

Lemma 4.2.

For any R,α>0R,\alpha>0 and any measure μ\mu in R\square_{R} such that (μ)<\mathcal{E}(\mu)<\infty, the infimum in the definition of ΦRα(μ)\Phi_{\square_{R}}^{\alpha}(\mu) is achieved.

Proof.

Let ρN\rho_{N} be a minimizing sequence for

infρ+(𝐑dR)(μ+ρα).\inf_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\mu+\rho-\alpha). (82)

Note that

lim supN(μ+ρNα)<.\limsup_{N\to\infty}\mathcal{E}(\mu+\rho_{N}-\alpha)<\infty. (83)

Hence, modulo a subsequence,

ρNρ\rho_{N}\rightharpoonup\rho (84)

weakly in H1H^{-1} for some ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}). By l.s.c. of ,\mathcal{E}, we have

(μ+ρα)lim infN(μ+ρNα)=infρ+(𝐑dR)(μ+ρα).\begin{split}\mathcal{E}(\mu+\rho-\alpha)&\leq\liminf_{N\to\infty}\mathcal{E}(\mu+\rho_{N}-\alpha)\\ &=\inf_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\mu+\rho-\alpha).\end{split} (85)

We now prove that the function ΦRα\Phi_{\square_{R}}^{\alpha} is a convex rate function for any α,R>0\alpha,R>0.

Claim 4.2.

For any α,R>0\alpha,R>0, the function ΦRα\Phi_{\square_{R}}^{\alpha} is a convex rate function.

Proof.

We first prove convexity. Let μ,ν\mu,\nu be measures on R\square_{R} such that ΦRα(μ)+ΦRα(ν)<\Phi_{\square_{R}}^{\alpha}(\mu)+\Phi_{\square_{R}}^{\alpha}(\nu)<\infty. Let

ρμ=argminρ+(𝐑dR)(μ+ρα),\rho_{\mu}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\mu+\rho-\alpha), (86)

and

ρν=argminρ+(𝐑dR)(ν+ρα).\rho_{\nu}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\nu+\rho-\alpha). (87)

Then, using the convexity of \mathcal{E} we have

ΦRα(12(μ+ν))(12(μ+ν)+12(ρμ+ρν)α))12((μ+ρμα)+(μ+ρμα))=12(ΦRα(μ)+ΦRα(ν)).\begin{split}\Phi_{\square_{R}}^{\alpha}\left(\frac{1}{2}\left(\mu+\nu\right)\right)&\leq\mathcal{E}\left(\frac{1}{2}(\mu+\nu)+\frac{1}{2}(\rho_{\mu}+\rho_{\nu})-\alpha)\right)\\ &\leq\frac{1}{2}\bigg{(}\mathcal{E}\left(\mu+\rho_{\mu}-\alpha\right)+\mathcal{E}\left(\mu+\rho_{\mu}-\alpha\right)\bigg{)}\\ &=\frac{1}{2}\left(\Phi_{\square_{R}}^{\alpha}(\mu)+\Phi_{\square_{R}}^{\alpha}(\nu)\right).\end{split} (88)

This proves the convexity of ΦRα\Phi_{\square_{R}}^{\alpha}. We now turn to prove that ΦRα\Phi_{\square_{R}}^{\alpha} is l.s.c. Since it is clearly positive, this will conclude the proof. Let μ\mu be a measure in R\square_{R} such that ΦRα(μ)<\Phi_{\square_{R}}^{\alpha}(\mu)<\infty and let μn\mu_{n} be a sequence of measures in R\square_{R} such that

μnμ\mu_{n}\rightharpoonup\mu (89)

weakly in the sense of measures. Let

ρn=argminρ+(𝐑dR)((μn+ρα)).\rho_{n}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\left(\mathcal{E}(\mu_{n}+\rho-\alpha)\right). (90)

Note that

lim supn(μn+ρnα)<.\limsup_{n\to\infty}\mathcal{E}(\mu_{n}+\rho_{n}-\alpha)<\infty. (91)

Then by precompactness, we have that the sequence μn+ρnα\mu_{n}+\rho_{n}-\alpha is precompact in the weak H1H^{-1} topology (note that we are not claiming precompactness for convergence in the BL metric, which is clearly not true in general). Let σ\sigma be such that

μn+ρnασ.\mu_{n}+\rho_{n}-\alpha\rightharpoonup\sigma. (92)

It is easy to see that σ\sigma and μα𝟏R\mu-\alpha\mathbf{1}_{\square_{R}} agree in the interior of R.\square_{R}. Note also that

ρ:=σ(μα𝟏R)\rho:=\sigma-(\mu-\alpha\mathbf{1}_{\square_{R}}) (93)

is a positive measure, and therefore it can be used as a test function in the definition of ΦRα\Phi_{\square_{R}}^{\alpha}. Then, using l.s.c. of \mathcal{E} we have

ΦRα(μ)(μ+ρα))lim infn(μn+ρnα))=lim infnΦRα(μn).\begin{split}\Phi_{\square_{R}}^{\alpha}(\mu)&\leq\mathcal{E}\left(\mu+\rho-\alpha)\right)\\ &\leq\liminf_{n\to\infty}\mathcal{E}\left(\mu_{n}+\rho_{n}-\alpha)\right)\\ &=\liminf_{n\to\infty}\Phi_{\square_{R}}^{\alpha}(\mu_{n}).\end{split} (94)

We will now prove that the rate functions are good.

Claim 4.3.

For any R,α>0,R,\alpha>0, the function 𝒩(μ|α𝟏R)\mathcal{N}(\mu|\alpha\mathbf{1}_{\square_{R}}) is a good rate function, i.e. sublevel sets are precompact in the topology of weak convergence of measures.

Proof.

Consider the sublevel sets

LM={μ+(R)|𝒩(μ|α𝟏R)<M}.L_{M}=\{\mu\in\mathcal{M}^{+}(\square_{R})|\mathcal{N}(\mu|\alpha\mathbf{1}_{\square_{R}})<M\}. (95)

We will prove that there exists NN such that if

μLM\mu\in L_{M} (96)

then

|μ|N,|\mu|\leq N, (97)

which will imply the desired compactness. Let

μ¯=1RdR𝑑μ.\overline{\mu}=\frac{1}{R^{d}}\int_{\square_{R}}d\mu. (98)

Using Jensen’s inequality, we have

𝒩(μ|α𝟏R)𝒩(μ¯|α𝟏R)=Rdα(μ¯αlog(μ¯α)μ¯α+1).\begin{split}\mathcal{N}(\mu|\alpha\mathbf{1}_{\square_{R}})&\geq\mathcal{N}(\overline{\mu}|\alpha\mathbf{1}_{\square_{R}})\\ &=R^{d}\alpha\left(\frac{\overline{\mu}}{\alpha}\log\left(\frac{\overline{\mu}}{\alpha}\right)-\frac{\overline{\mu}}{\alpha}+1\right).\end{split} (99)

Since xlog(x)xx\log(x)-x\to\infty as x,x\to\infty, we have that there exists NN such that |μ|<N|\mu|<N if μLM.\mu\in L_{M}. Hence, LML_{M} is precompact in the topology of weak convergence. ∎

We now prove that ΦRα\Phi^{\alpha}_{\square_{R}} is a good rate function.

Claim 4.4.

For any R,α>0R,\alpha>0 the function ΦRα\Phi^{\alpha}_{\square_{R}} is a good rate function, i.e. sublevel sets are precompact in the topology of weak convergence of measures..

Proof.

Let μn+(R)\mu_{n}\in\mathcal{M}^{+}(\square_{R}) be such that

lim supnΦRα(μn)<.\limsup_{n\to\infty}\Phi^{\alpha}_{\square_{R}}(\mu_{n})<\infty. (100)

Let

ρn=argminρ+(𝐑dR)(μn+ρα).\rho_{n}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\mu_{n}+\rho-\alpha). (101)

Since we are assuming equation (100), we have that μn+ρn\mu_{n}+\rho_{n} converges, modulo a subsequence (not relabelled) weakly in the H1H^{-1} topology. Hence the restriction to R\square_{R}, μn\mu_{n} converges weakly in the H1H^{-1} topology. In particular,

lim supn(μn)<.\limsup_{n\to\infty}\mathcal{E}(\mu_{n})<\infty. (102)

Since μn\mu_{n} is a positive measure, equation (102) implies that

lim supn|μn|<,\limsup_{n\to\infty}|\mu_{n}|<\infty, (103)

which implies that modulo a subsequence (not relabelled) μn\mu_{n} converges in the topology of weak convergence of probability measures.

5 Proof of upper bound

In this section, we prove the upper bound of Theorem 2.1. Recall that we use the notation

XN=(y1,yiN,z1,zjN),X_{N}=(y_{1},...y_{i_{N}},z_{1},...z_{j_{N}}), (104)

with

ymR,zmR.y_{m}\in\square_{R},\quad z_{m}\notin\square_{R}. (105)
Proof of Theorem 2.1, upper bound.

We begin by using the splitting formula for the thermal equilibrium measure (Proposition (4.1)). Let ϵ,R>0\epsilon,R>0 and ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}), then

𝐏N,β(lempNB(ν,ϵ))=1ZN,βXN:lempNB(ν,ϵ)exp(β(XN))𝑑XN=1KN,βXN:lempNB(ν,ϵ)exp(12βN2(empNμβ))Πi=1Nμβ(xi)𝑑xi1KN,βXN:lempNB(ν,ϵ)exp(12βN2𝐖RNλ,1NNiN,μβ(empN))Πi=1Nμβ(xi)𝑑xi=1KN,βXN:lempNB(ν,ϵ)exp(12βN2(d+2)λ𝐖R,N1+λdNiN,μβNλ(lempN))Πi=1Nμβ(xi)dxi1KN,βsupμB(ν,ϵ)𝒜iNN1+λd(R){exp(12βN2(d+2)λ𝐖R,N1+λdNiN,μβNλ(μ))}XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi.\begin{split}&\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in B(\nu,\epsilon)\right)=\\ &\frac{1}{Z_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\exp\left(-\beta\mathcal{H}(X_{N})\right)dX_{N}=\\ &\frac{1}{K_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\exp\left(-\frac{1}{2}\beta N^{2}\mathcal{E}^{\neq}({\rm emp}_{N}-\mu_{\beta})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\leq\\ &\frac{1}{K_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\exp\left(-\frac{1}{2}\beta N^{2}\mathbf{W}_{\square_{\frac{R}{N^{\lambda}}},\frac{1}{N}}^{N-i_{N},\mu_{\beta}}({\rm emp}^{\prime}_{N})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}=\\ &\frac{1}{K_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\\ &\quad\exp\left(-\frac{1}{2}\beta N^{2-(d+2)\lambda}\mathbf{W}_{\square_{R},N^{-1+\lambda d}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}({\rm lemp}_{N})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\leq\\ &\frac{1}{K_{N,\beta}}\sup_{\mu\in B(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\\ &\quad\left\{\exp\left(-\frac{1}{2}\beta N^{2-(d+2)\lambda}\mathbf{W}_{\square_{R},N^{-1+\lambda d}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu)\right)\right\}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}.\end{split} (106)

In order to pass from the third to the fourth line, we have used that

1KN,βXN:lempNB(ν,ϵ)exp(12βN2(empNμβ))Πi=1Nμβ(xi)𝑑xi1KN,βXN:lempNB(ν,ϵ)exp(12βN2infzi𝐑dRNλ(empNμβ))Πi=1Nμβ(xi)𝑑xi,\begin{split}&\frac{1}{K_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\exp\left(-\frac{1}{2}\beta N^{2}\mathcal{E}^{\neq}({\rm emp}_{N}-\mu_{\beta})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\leq\\ &\frac{1}{K_{N,\beta}}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\exp\left(-\frac{1}{2}\beta N^{2}\inf_{z_{i}\in\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}}}\mathcal{E}^{\neq}({\rm emp}_{N}-\mu_{\beta})\right)\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i},\end{split} (107)

since for any ZN(𝐑dRNλ)jNZ_{N}^{*}\in(\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}})^{j_{N}}

(empN(YN,ZN)μβ)infzi𝐑dRNλ(empN(YN,ZN)μβ).\mathcal{E}^{\neq}({\rm emp}_{N}(Y_{N},Z_{N}^{*})-\mu_{\beta})\geq{\inf_{z_{i}\in\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}}}}\mathcal{E}^{\neq}({\rm emp}_{N}(Y_{N},Z_{N})-\mu_{\beta}). (108)

But given yiRNλ,y_{i}\in\square_{RN^{-\lambda}}, we have

infZN(𝐑dRNλ)jN(empNμβ)𝐖RNλ,1NNiN,μβ(empN),\inf_{Z_{N}\in(\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}})^{j_{N}}}\mathcal{E}^{\neq}({\rm emp}_{N}-\mu_{\beta}){\geq}\mathbf{W}_{\square_{\frac{R}{N^{\lambda}}},\frac{1}{N}}^{N-i_{N},\mu_{\beta}}({\rm emp}^{\prime}_{N}), (109)

see equation (64).

We now treat each of the terms in the last line of equation (106) individually. The second term is the easier, and we will will deal with it at the end of this section. More specifically, we will prove the following lemma:

Lemma 5.1.

Let R,ϵ>0R,\epsilon>0 and ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}). Then

lim supN(1N1λdlog(XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi))infμB(ν,ϵ)(𝒩(μ|μV(0)𝟏R)).\begin{split}&\limsup_{N\to\infty}\Bigg{(}\frac{1}{N^{1-\lambda d}}\log\Bigg{(}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\Bigg{)}\Bigg{)}\leq\\ &-\inf_{\mu\in B(\nu,\epsilon)}(\mathcal{N}(\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}})).\end{split} (110)

The analysis of the first term is more delicate, and we deal with it in section 66. The result we prove is the following:

Lemma 5.2.

Let R,ϵ>0R,\epsilon>0, let ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) and iNi_{N} be an integer smaller than or equal to NN. Then

infμB(ν,δ)ΦRμV(0)(μ)lim infNinfμB(ν,δ)𝒜iNN1+λd(R)𝐖R,N1+λdNiN,μβNλ(μ).\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{V}(0)}(\mu)\leq\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\mathbf{W}_{\square_{R},N^{-1+\lambda d}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu). (111)

Furthermore, for any ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) such that (ν)<\mathcal{E}(\nu)<\infty we have

limN|ΦRμV(0)(ν)ΦRμβNλ(ν)|=0.\lim_{N\to\infty}\left|\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)-\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)\right|=0. (112)

We will now finish the proof of the upper bound in Theorem 2.1 using Lemmas 5.1 and 5.2. We start with the last line of equation (106):

𝐏N,β(lempNB(ν,ϵ))1KN,βsupμB(ν,ϵ)𝒜iNN1+λd(R){exp(12βN2(d+2)λ𝐖R,N1+λdNiN,μβNλ(μ))}XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi.\begin{split}&\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in B(\nu,\epsilon)\right)\leq\\ &\frac{1}{K_{N,\beta}}\sup_{\mu\in B(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\\ &\quad\left\{\exp\left(-\frac{1}{2}\beta N^{2-(d+2)\lambda}\mathbf{W}_{\square_{R},N^{-1+\lambda d}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu)\right)\right\}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}.\end{split} (113)

Using results from [27], or from [1], we know that

|log(KN,β)|CβN22d,|\log(K_{N,\beta})|\leq C\beta N^{2-\frac{2}{d}}, (114)

using the hypothesis that λ<1d(d+2)\lambda<\frac{1}{d(d+2)} we have that

|log(KN,β)|=o(min(βN2λ(d+2),N1λd).|\log(K_{N,\beta})|=o(\min(\beta N^{2-\lambda(d+2)},N^{1-\lambda d}). (115)

Bounding this error term (and bounding a similar error term in the upper bound) is the only step in which we use the hypothesis that λ<1d(d+2)\lambda<\frac{1}{d(d+2)}.

Note that, if γ<γ\gamma<\gamma^{*} then

2(d+2)λγ>1λd,2-(d+2)\lambda-\gamma>1-\lambda d, (116)

and so

lim supN1βN2(d+2)λlog(𝐏N,β(lempNB(ν,ϵ)))12infμB(ν,ϵ)ΦRμV(0)(μ).\limsup_{N\to\infty}\frac{1}{\beta N^{2-(d+2)\lambda}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in B(\nu,\epsilon)\right)\right)\leq-\frac{1}{2}\inf_{\mu\in B(\nu,\epsilon)}\Phi^{\mu_{V}(0)}_{\square_{R}}(\mu). (117)

And finally, if γ>γ\gamma>\gamma^{*} then

2(d+2)λγ<1λd,2-(d+2)\lambda-\gamma<1-\lambda d, (118)

and so

lim supN1N1λdlog(𝐏N,β(lempNB(ν,ϵ)))infμB(ν,ϵ)𝒩[μ|μV(0)𝟏R].\limsup_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in B(\nu,\epsilon)\right)\right)\leq-\inf_{\mu\in B(\nu,\epsilon)}\mathcal{N}[\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]. (119)

This concludes the proof of the upper bound of Theorem 2.1. We now turn to the proof of the auxiliary lemmas (Lemmas 5.1 and 5.2). We start with Lemma 5.1, which we restate here for convenience:

Lemma 5.3.

Let R,ϵ>0R,\epsilon>0 and ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}). Then

lim supN(1N1λdlog(XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi))infμB(ν,ϵ)(𝒩(μ|μV(0)𝟏R)).\begin{split}&\limsup_{N\to\infty}\Bigg{(}\frac{1}{N^{1-\lambda d}}\log\Bigg{(}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\Bigg{)}\Bigg{)}\leq\\ &-\inf_{\mu\in B(\nu,\epsilon)}(\mathcal{N}(\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}})).\end{split} (120)
Proof.

Using Sanov’s theorem and the scaling relation of ent{\rm ent}, we have that

limN1N1λdlog(XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi)infρent[ρ|μβNλ],\lim_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\right)\leq-\inf_{\rho}{\rm ent}[{\rho}|\mu_{\beta}^{N^{\lambda}}], (121)

where the infimum is taken over ρ\rho such that |ρ|=Nλd|\rho|=N^{\lambda d} and ρ|RB(ν,ϵ).\rho|\square_{R}\in B(\nu,\epsilon). Note that we may rewrite equation (121) as

limN1N1λdlog(XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi)infμB(ν,ϵ)(ent[μ|μβNλ|R]+infρent[ρ|μβ]),\begin{split}&\lim_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\right)\leq\\ &-\inf_{\mu\in B(\nu,\epsilon)}\left({\rm ent}[\mu|\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}]+\inf_{\rho}{\rm ent}[\rho|\mu_{\beta}]\right),\end{split} (122)

where the infimum is taken over all ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}) such that |ρ|=Nλd|μ|.|\rho|=N^{\lambda d}-|\mu|.

We first determine the optimal ρ\rho in the minimization problem of equation (122) for a given μ\mu. This can be done by adding a Lagrange multiplier for the constraint of mass and then computing the Euler Lagrange equations. The solution is that the minimizer ρ\rho^{*} is given by

ρ=κμβNλ𝟏𝐑dR,\rho^{*}=\kappa\mu_{\beta}^{N^{\lambda}}\mathbf{1}_{\mathbf{R}^{d}\setminus\square_{R}}, (123)

where κ\kappa is given by

κ=Nλd|μ|𝐑dRμβNλ.\kappa=\frac{N^{\lambda d}-|\mu|}{\int_{\mathbf{R}^{d}\setminus\square_{R}}\mu_{\beta}^{N^{\lambda}}}. (124)

Hence we have that, for each μB(ν,ϵ),\mu\in B(\nu,\epsilon),

limNent[μ+ρ|μβNλ]=limNent[μ|μβNλ|R]+𝐑dRlog(κ)κμβNλ𝑑x=ent[μ|μV(0)𝟏R]+limNκ(κ1)𝐑dRμβNλ𝑑x.\begin{split}\lim_{N\to\infty}{\rm ent}[\mu+\rho^{*}|\mu_{\beta}^{N^{\lambda}}]&=\lim_{N\to\infty}{\rm ent}[\mu|\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}]+\int_{\mathbf{R}^{d}\setminus\square_{R}}\log(\kappa)\kappa\mu_{\beta}^{N^{\lambda}}dx\\ &={\rm ent}[\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]+\lim_{N\to\infty}\kappa(\kappa-1)\int_{\mathbf{R}^{d}\setminus\square_{R}}\mu_{\beta}^{N^{\lambda}}dx.\end{split} (125)

In the last equation, we have used Remark 4.1 and the approximation log(κ)κ1\log(\kappa)\simeq\kappa-1, since κ\kappa tends to 11 as NN tends to \infty. Recalling that

𝐑dμβNλ=Nλd,\int_{\mathbf{R}^{d}}\mu_{\beta}^{N^{\lambda}}=N^{\lambda d}, (126)

and using again Remark 4.1 we have that

limNκ(κ1)𝐑dRμβNλ𝑑x=RdμV(0)|μ|.\lim_{N\to\infty}\kappa(\kappa-1)\int_{\mathbf{R}^{d}\setminus\square_{R}}\mu_{\beta}^{N^{\lambda}}dx=R^{d}\mu_{V}(0)-|\mu|. (127)

Therefore

lim supN(1N1λdlog(XN:lempNB(ν,ϵ)Πi=1Nμβ(xi)𝑑xi))infμB(ν,ϵ)(𝒩(μ|μV(0)𝟏R)).\begin{split}&\limsup_{N\to\infty}\Bigg{(}\frac{1}{N^{1-\lambda d}}\log\Bigg{(}\int_{X_{N}:{\rm lemp}_{N}\in B(\nu,\epsilon)}\Pi_{i=1}^{N}\mu_{\beta}(x_{i})dx_{i}\Bigg{)}\Bigg{)}\leq\\ &-\inf_{\mu\in B(\nu,\epsilon)}(\mathcal{N}(\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}})).\end{split} (128)

6 Proof of Lemma 5.2

In this section, we prove Lemma 5.2, which we restate here for convenience:

Lemma 6.1.

Let R,ϵ>0R,\epsilon>0, let ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) and iNi_{N} be an integer smaller than or equal to NN. Then

infμB(ν,δ)ΦRμV(0)(μ)lim infNinfμB(ν,δ)𝒜iNN1+λd(R)𝐖R,N1+λdNiN,μβNλ(μ).\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{V}(0)}(\mu)\leq\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\mathbf{W}_{\square_{R},N^{-1+\lambda d}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu). (129)

Furthermore, for any ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) such that (ν)<\mathcal{E}(\nu)<\infty we have

limN|ΦRμV(0)(ν)ΦRμβNλ(ν)|=0.\lim_{N\to\infty}\left|\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)-\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)\right|=0. (130)

The idea is that, on the one hand, given our choice of dilation, (empN)Nλ({\rm emp}_{N})^{N^{\lambda}} will converge to a continuous measure on every compact set. This implies that we can replace the infimum over purely atomic measures with the infimum over absolutely continuous measures in +(𝐑dR).\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}). On the other hand, μβNλ\mu_{\beta}^{N^{\lambda}} will converge to μV(0)\mu_{V}(0) on compact sets, so we can replace the background measure μβNλ\mu_{\beta}^{N^{\lambda}} with μV(0).\mu_{V}(0). We will now make this intuition more rigorous.

Proof.

Step 1

We claim that

lim infNinfμB(ν,δ)ΦRμβNλ(μ)lim infNinfμB(ν,δ)𝒜iNN1+λd(R)𝐖R,Nλd1NiN,μβNλ(μ).\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{\beta}^{N^{\lambda}}}(\mu)\leq\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\mathbf{W}_{\square_{R},N^{\lambda d-1}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu). (131)

To see this, let

μ=1Ni=1iNδyi,\mu=\frac{1}{N}\sum_{i=1}^{i_{N}}\delta_{y_{i}}, (132)

and

ρ=1Ni=1jNδzi,\rho=\frac{1}{N}\sum_{i=1}^{j_{N}}\delta_{z_{i}}, (133)

with yiRy_{i}\in\square_{R}, zi𝐑dRz_{i}\in\mathbf{R}^{d}\setminus\square_{R} and iN+jN=Ni_{N}+j_{N}=N.

Now we define

μ~=μλN1d\widetilde{\mu}=\mu\ast\lambda_{N^{-\frac{1}{d}}} (134)

and

ρ~=ρλN1d,\widetilde{\rho}=\rho\ast\lambda_{N^{-\frac{1}{d}}}, (135)

(see Remark 3.2 for notation).

Then

μμ~BLN1d\|\mu-\widetilde{\mu}\|_{BL}\leq N^{-\frac{1}{d}} (136)

and we also have, because of Lemmas 9.1, 9.2, 9.3, 9.4 that

(μ~+ρ~μβ)(μ+ρμβ)+CN2d.\mathcal{E}(\widetilde{\mu}+\widetilde{\rho}-\mu_{\beta})\leq\mathcal{E}^{\neq}(\mu+{\rho}-\mu_{\beta})+CN^{-\frac{2}{d}}. (137)

Note that CC depends only on VV and dd, since μβ\mu_{\beta} is uniformly bounded in NN for NN large enough, with a bound that depends only on VV and dd.

Using the hypothesis that

λ<1d(d+2),\lambda<\frac{1}{d(d+2)}, (138)

we have that

N2d<<Nλ(d+2),N^{-\frac{2}{d}}<<N^{-\lambda(d+2)}, (139)

which implies, using the scaling relations of ΦRμβNλ\Phi_{\square_{R}}^{\mu_{\beta}^{N^{\lambda}}} and 𝐖R,Nλd1NiN,μβNλ\mathbf{W}_{\square_{R},N^{\lambda d-1}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}, that

lim infNinfμB(ν,δ)ΦRμβNλ(μ)lim infNinfμB(ν,δ)𝒜iNN1+λd(R)𝐖R,Nλd1NiN,μβNλ(μ).\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{\beta}^{N^{\lambda}}}(\mu)\leq\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)\cap\mathcal{A}_{i_{N}}^{N^{-1+\lambda d}}(\square_{R})}\mathbf{W}_{\square_{R},N^{\lambda d-1}}^{N-i_{N},\mu_{\beta}^{N^{\lambda}}}(\mu). (140)

Step 2 We now prove the second part of the claim: that for any ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) such that (ν)<\mathcal{E}(\nu)<\infty we have

limN|ΦRμV(0)(ν)ΦRμβNλ(ν)|=0.\lim_{N\to\infty}\left|\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)-\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)\right|=0. (141)

We will first prove that

lim supNΦRμβNλ(ν)ΦRμV(0)(ν).\limsup_{N\to\infty}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)\leq\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu). (142)

To this end, let

ρ¯=argminρ+(𝐑dR)(ρ+νμV(0)).\overline{\rho}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\rho+\nu-\mu_{V}(0)). (143)

For any ϵ>0\epsilon>0 let TT be such that

|((ρ¯+νμV(0))𝟏B(0,T))(ρ¯+νμV(0))|ϵ.\begin{split}&\left|\mathcal{E}\left(\left(\overline{\rho}+\nu-\mu_{V}(0)\right)\mathbf{1}_{B(0,T)}\right)-\mathcal{E}\left(\overline{\rho}+\nu-\mu_{V}(0)\right)\right|\leq\epsilon.\end{split} (144)

Taking ρ¯𝟏B(0,T)+μβNλ𝟏𝐑dB(0,T)\overline{\rho}\mathbf{1}_{B(0,T)}+\mu_{\beta}^{N^{\lambda}}\mathbf{1}_{\mathbf{R}^{d}\setminus B(0,T)} as a test function in the definition of ΦRμβNλ(ν)\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu) and using Remark 4.1 we have

lim supNΦRμβNλ(ν)lim supN((ρ¯+νμβNλ)𝟏B(0,T))=((ρ¯+νμV(0))𝟏B(0,T))(ρ¯+νμV(0))+ϵ=ΦRμV(0)(ν)+ϵ.\begin{split}\limsup_{N\to\infty}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)&\leq\limsup_{N\to\infty}\mathcal{E}\left(\left(\overline{\rho}+\nu-\mu_{\beta}^{N^{\lambda}}\right)\mathbf{1}_{B(0,T)}\right)\\ &=\mathcal{E}\left(\left(\overline{\rho}+\nu-\mu_{V}(0)\right)\mathbf{1}_{B(0,T)}\right)\\ &\leq\mathcal{E}(\overline{\rho}+{\nu}-\mu_{V}(0))+\epsilon\\ &=\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)+\epsilon.\end{split} (145)

Since ϵ>0\epsilon>0 is arbitrary, we can conclude that

lim supNΦRμβNλ(ν)ΦRμV(0)(ν).\limsup_{N\to\infty}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)\leq\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu). (146)

We now turn to prove

ΦRμV(0)(ν)lim infNΦRμβNλ(ν).\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)\leq\liminf_{N\to\infty}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu). (147)

To this end, let

ρN=argminρ(ρ+ν𝟏RμβNλ),{\rho}_{N}=\operatorname{argmin}_{\rho}\mathcal{E}(\rho+\nu-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}}), (148)

where ρ\rho is minimized over measures satisfying ρμβNλ\rho\geq-\mu_{\beta}^{N^{\lambda}} and which are supported in 𝐑dR.\mathbf{R}^{d}\setminus\square_{R}.

Note that

ΦRμβNλ(ν)=(ρN+ν𝟏RμβNλ).\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu)=\mathcal{E}(\rho_{N}+\nu-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}}). (149)

Then, since

lim supN(ρN+ν𝟏RμβNλ)<,\limsup_{N\to\infty}\mathcal{E}(\rho_{N}+\nu-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}})<\infty, (150)

we have that

ρNρ^,\rho_{N}\rightharpoonup\widehat{\rho}, (151)

weakly in H1,H^{-1}, for some ρ^\widehat{\rho}. It is easy to check that ρ^μV(0)\widehat{\rho}\geq-\mu_{V}(0) a.e. Using l.s.c. of ,\mathcal{E}, we then have that

ΦRμV(0)(ν)(νμV(0)𝟏R+ρ^)lim infN(νμβNλ𝟏R+ρN)=lim infNΦRμβNλ(ν).\begin{split}\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)&\leq\mathcal{E}(\nu-\mu_{V}(0)\mathbf{1}_{\square_{R}}+\widehat{\rho})\\ &\leq\liminf_{N\to\infty}\mathcal{E}(\nu-\mu_{\beta}^{N^{\lambda}}\mathbf{1}_{\square_{R}}+{\rho}_{N})\\ &=\liminf_{N\to\infty}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\nu).\end{split} (152)

Step 3

We now prove the first part of the statement of Lemma 5.2. In view of equation (140), we will prove that for any δ>0\delta>0 and any measure ν\nu on R\square_{R} such that (ν)<\mathcal{E}(\nu)<\infty,

infμB(ν,δ)ΦRμV(0)(μ)lim infNinfμB(ν,δ)ΦRμβNλ(μ).\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{V}(0)}(\mu)\leq\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\mu). (153)

To this end, let μNB(ν,δ)\mu_{N}\in B(\nu,\delta) be such that

infμB(ν,δ)ΦRμβNλ(μ)=ΦRμβNλ(μN),\inf_{\mu\in B(\nu,\delta)}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\mu)=\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\mu_{N}), (154)

we assume that the infimum is achieved for clarity of exposition, otherwise, we could prove the claim up to an arbitrary error by taking a minimizing sequence.

Since μNB(ν,δ),\mu_{N}\in B(\nu,\delta), we have that as NN tends to \infty

μNμ¯,\mu_{N}\rightharpoonup\overline{\mu}, (155)

weakly in the sense of measures, for some μ¯B(ν,δ).\overline{\mu}\in B(\nu,\delta). Let

ρN=argminρ(μ+μN𝟏RμβNλ),{\rho}_{N}=\operatorname{argmin}_{\rho}\mathcal{E}(\mu+\mu_{N}-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}}), (156)

where ρ\rho is minimized over ρμβNλ\rho\geq-\mu_{\beta}^{N^{\lambda}} supported in 𝐑dR.\mathbf{R}^{d}\setminus\square_{R}.

Note that

lim supN(μN+ρN𝟏RμβNλ)<,\limsup_{N\to\infty}\mathcal{E}(\mu_{N}+\rho_{N}-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}})<\infty, (157)

therefore, for a subsequence

ρN+μNρ¯+μ¯,\rho_{N}+\mu_{N}\rightharpoonup\overline{\rho}+\overline{\mu}, (158)

weakly in H1,H^{-1}, for some ρ¯μV(0)\overline{\rho}\geq-\mu_{V}(0). Therefore we can use ρ¯\overline{\rho} as a test function in the definition of ΦRμV(0)\Phi_{\square_{R}}^{\mu_{V}(0)} and get

infμB(ν,δ)ΦRμV(0)(μ)ΦRμV(0)(μ¯)(ρ¯+μ¯𝟏RμV(0))lim infN(μN+ρN𝟏RμβNλ)=lim infNinfμB(ν,δ)ΦRμβNλ(μ).\begin{split}\inf_{\mu\in B(\nu,\delta)}\Phi_{\square_{R}}^{\mu_{V}(0)}(\mu)&\leq\Phi_{\square_{R}}^{\mu_{V}(0)}(\overline{\mu})\\ &\leq\mathcal{E}(\overline{\rho}+\overline{\mu}-\mathbf{1}_{\square_{R}}\mu_{V}(0))\\ &\leq\liminf_{N\to\infty}\mathcal{E}({\mu}_{N}+\rho_{N}-\mathbf{1}_{\square_{R}}\mu_{\beta}^{N^{\lambda}})\\ &=\liminf_{N\to\infty}\inf_{\mu\in B(\nu,\delta)}\Phi^{\mu_{\beta}^{N^{\lambda}}}_{\square_{R}}(\mu).\end{split} (159)

7 Proof of lower bound

This section is devoted to proving the lower bound of the LDP’s of Theorem 2.1. The approach will be to construct a family of point configurations that has correct energy and sufficient volume.

We start with a lemma, which builds upon a construction found in unpublished class notes by Sylvia Serfaty.

Lemma 7.1.

Let μn,ν¯\mu_{n},\overline{\nu} be probability measures on a compact set Ω\Omega such that

lim supnent[μn]<,ent[ν¯]<,\limsup_{n\to\infty}{\rm ent}[\mu_{n}]<\infty,\quad{\rm ent}[\overline{\nu}]<\infty, (160)

ν¯L(Ω)\overline{\nu}\in L^{\infty}(\Omega), and

(ν¯)<.\mathcal{E}(\overline{\nu})<\infty. (161)

Assume that μn(x)\mu_{n}(x) is uniformly equi-continuous and bounded away from 0 uniformly in xx and nn. Then for every ϵ,δ,η,\epsilon,\delta,\eta, there exists a family of configurations

Λδη,ϵ𝐑d×n\Lambda_{\delta}^{\eta,\epsilon}\subset\mathbf{R}^{d\times n} (162)

such that

  • \bullet
    empn(Xn)B(ν¯,ϵ){\rm emp}_{n}(X_{n})\in B(\overline{\nu},\epsilon) (163)

    for any XnΛδη,ϵ.X_{n}\in\Lambda_{\delta}^{\eta,\epsilon}.

  • \bullet
    lim infn1nlog(XnΛδη,ϵΠi=1nμn(xi)𝑑Xn)lim infnent[ν¯|μn]δ.\liminf_{n\to\infty}\frac{1}{n}\log\left(\int_{X_{n}\in\Lambda_{\delta}^{\eta,\epsilon}}\Pi_{i=1}^{n}\mu_{n}(x_{i})dX_{n}\right)\geq-\liminf_{n\to\infty}{\rm ent}[\overline{\nu}|\mu_{n}]-\delta. (164)
  • \bullet
    lim supn|(empn(Xn)ν¯)|η2.\limsup_{n\to\infty}\left|\mathcal{E}^{\neq}({\rm emp}_{n}(X_{n})-\overline{\nu})\right|\leq\eta^{2}. (165)
  • There exists r>0r>0 such that

    d(xi,Ω)>rn1dandd(xi,xj)>rn1d,d(x_{i},\partial\Omega)>rn^{-\frac{1}{d}}\quad{\rm and}\quad d(x_{i},x_{j})>rn^{-\frac{1}{d}}, (166)

    for iji\neq j.

The proof of Lemma 7.1 is found in Section 10.

We will also require the following lemma, which deals with approximating certain partition functions.

Lemma 7.2.

Let YN=(y1,y2,yiN)Y_{N}=(y_{1},y_{2},...y_{i_{N}}) with yjRNλy_{j}\in\square_{RN^{-\lambda}} for each jj. Let ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) such that

(ν)<,νL(R).\mathcal{E}(\nu)<\infty,\quad\nu\in L^{\infty}(\square_{R}). (167)

Assume that

lim supN|(lempN(YN)ν)|η2,\limsup_{N\to\infty}\left|\mathcal{E}^{\neq}({\rm lemp}_{N}(Y_{N})-\nu)\right|\leq\eta^{2}, (168)

also that

limN|ν|N1+λdiN0,\lim_{N\to\infty}|\nu|-N^{-1+\lambda d}i_{N}\to 0, (169)

and that there exists r>0r>0 such that

d(yi,RNλ)rN1dd(y_{i},\partial\square_{RN^{-\lambda}})\geq rN^{-\frac{1}{d}} (170)

and

d(yi,yj)rN1d.d(y_{i},y_{j})\geq rN^{-\frac{1}{d}}. (171)

Let

ZN,βYN=ZN(𝐑dRNλ)jkexp(βN(YN,ZN))𝑑ZN.Z_{N,\beta}^{Y_{N}}=\iint_{Z_{N}\in(\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}})^{j_{k}}}\exp\left(-\beta\mathcal{H}_{N}(Y_{N},Z_{N})\right)dZ_{N}. (172)

Then for γ<γ\gamma<\gamma^{*} we have

1βN2λ(d+2)(log(ZN,βYN)N2ββ(μβ))12ΦRμV(0)(ν)+Cη+oN(1).\frac{1}{\beta N^{2-\lambda(d+2)}}\left(-\log(Z_{N,\beta}^{Y_{N}})-N^{2}\beta\mathcal{E}_{\beta}(\mu_{\beta})\right)\leq\frac{1}{2}\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)+C\eta+o_{N}(1). (173)

where CC and oN(1)o_{N}(1) are independent of YN.Y_{N}.

For γ>γ\gamma>\gamma^{*} we have

N1+λdβ(1N2βlog(ZN,βYN)β(μβ))𝐑dlog(μβNλ)d(lempN(YN))|ν|+RdμV(0)+oN(1),\begin{split}&N^{1+\lambda d}\beta\left(-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\right)\leq\\ &\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))-|\nu|+R^{d}\mu_{V}(0)+o_{N}(1),\end{split} (174)

where oN(1)o_{N}(1) is independent of YN.Y_{N}.

Proof.

We will divide the proof in 3 steps. The idea of the proof is that using the variational formulation of the partition function, as well as the splitting formula for the equilibrium measure (Proposition (4.1)), we can reduce the integral in equation (172) to

log(ZN,βYN)N2β(β(μβ)+infρ12RNλ(empN(YN)+ρμβ)1Nβ𝐑dlog(μβ)d(ρ+empN(YN))+1Nβent[ρ]).\begin{split}&-\log(Z_{N,\beta}^{Y_{N}})\simeq N^{2}\beta\bigg{(}\mathcal{E}_{\beta}(\mu_{\beta})+\inf_{\rho}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{RN^{-\lambda}}}({\rm emp}_{N}^{\prime}(Y_{N})+\rho-\mu_{\beta})-\\ &\frac{1}{N\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta})d(\rho+{\rm emp}_{N}^{\prime}(Y_{N}))+\frac{1}{N\beta}{\rm ent}[\rho]\bigg{)}.\end{split} (175)

This is done in step 1. Steps 2 and 3 simplify this expression, and show that either the electric energy or the entropy dominates, depending on whether γ>γ\gamma>\gamma^{*} or γ<γ.\gamma<\gamma^{*}.

Step 1

We start with the characterization

log(ZN,βYN)β=minμ𝒫([𝐑dRN1d]N)(μ),-\frac{\log(Z_{N,\beta}^{Y_{N}})}{\beta}=\min_{\mu\in\mathcal{P}([\mathbf{R}^{d}\setminus\square_{RN^{-\frac{1}{d}}}]^{N})}\mathcal{F}(\mu), (176)

in this equation,

(μ)=𝐑Ndμ(ZN)NYN(ZN)dZN+1β𝐑dNμ(ZN)log(μ(ZN))dZN\mathcal{F}(\mu)=\int_{\mathbf{R}^{Nd}}\mu(Z_{N})\mathcal{H}^{Y_{N}}_{N}(Z_{N})\,\mathrm{d}Z_{N}+\frac{1}{\beta}\int_{\mathbf{R}^{dN}}\mu(Z_{N})\log(\mu(Z_{N}))\ \mathrm{d}Z_{N} (177)

where

NYN(ZN)=12yi,yjRN1dg(yiyj)+12zi,zjRN1dg(zizj)+NiV~(xi),\mathcal{H}^{Y_{N}}_{N}\left(Z_{N}\right)=\frac{1}{2}\sum_{y_{i},y_{j}\in\square_{RN^{-\frac{1}{d}}}}g\left(y_{i}-y_{j}\right)+\frac{1}{2}\sum_{z_{i},z_{j}\notin\square_{RN^{-\frac{1}{d}}}}g\left(z_{i}-z_{j}\right)+N\sum_{i}\widetilde{V}\left(x_{i}\right), (178)

and

V~=V+12empN(YN)g.\widetilde{V}=V+\frac{1}{2}{\rm emp}_{N}^{\prime}(Y_{N})\ast g. (179)

This is a particular case of a characterization of the partition function which is valid in general, see for example [27].

We now define

μβYN=argminρ𝒫(𝐑dRN1d)¯VYN(ρ)+1(NiN)βent[ρ],\mu_{\beta}^{Y_{N}}=\text{argmin}_{\rho\in\mathcal{P}(\mathbf{R}^{d}\setminus\square_{RN^{-\frac{1}{d}}})}\overline{\mathcal{I}}_{V}^{Y_{N}}(\rho)+\frac{1}{\left(N-i_{N}\right)\beta}\text{ent}[\rho], (180)

where

¯VYN(ρ)=12(empN(YN))+12(1iNN)2𝐑d×𝐑dg(xy)𝑑ρx𝑑ρy+(1iNN)V~𝑑ρ.\begin{split}\overline{\mathcal{I}}_{V}^{Y_{N}}(\rho)=\frac{1}{2}\mathcal{E}_{\neq}({\rm emp}_{N}^{\prime}(Y_{N}))+\frac{1}{2}\left(1-\frac{i_{N}}{N}\right)^{2}\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)d\rho_{x}d\rho_{y}+\\ \left(1-\frac{i_{N}}{N}\right)\int\widetilde{V}d\rho.\end{split} (181)

The reason for introducing the extra factors of 1iNN1-\frac{i_{N}}{N} is that we need to normalize the total charge outside the cube to be 11. We then have that probability measures satisfy a splitting formula around μβYN\mu_{\beta}^{Y_{N}}, analogous to equation (75).

We also have that μβYN\mu_{\beta}^{Y_{N}} satisfies the EL equation

(1iNN)2hμβYN+(1iNN)V~+1(NiN)βlog(μβYN)=k,\left(1-\frac{i_{N}}{N}\right)^{2}h^{\mu_{\beta}^{Y_{N}}}+\left(1-\frac{i_{N}}{N}\right)\widetilde{V}+\frac{1}{\left(N-i_{N}\right)\beta}\log(\mu_{\beta}^{Y_{N}})=k, (182)

for some constant k.k. In order to find k,k, we can multiply equation (182) by μβYN,\mu_{\beta}^{Y_{N}}, integrate, and use that μβYN\mu_{\beta}^{Y_{N}} has integral 11. The result is

k=¯VYN(μβYN)+1Nβent[μβYN]+12(1iNN)(μβYN).k=\overline{\mathcal{I}}_{V}^{Y_{N}}(\mu_{\beta}^{Y_{N}})+\frac{1}{N\beta}\text{ent}[\mu_{\beta}^{Y_{N}}]+\frac{1}{2}\left(1-\frac{i_{N}}{N}\right)\mathcal{E}(\mu_{\beta}^{Y_{N}}). (183)

Plugging in μ=(μβYN)NiN\mu=(\mu_{\beta}^{Y_{N}})^{\otimes N-i_{N}} as a test function in equation (176) we get

log(ZN,βYN)β((μβYN)NiN)=N2(¯VYN(μβYN)+1(NiN)βent[μβYN])(NiN2)(μβYN)N2(minρ¯VYN(ρ)+1(NiN)βent[ρ])CN.\begin{split}-\frac{\log(Z_{N,\beta}^{Y_{N}})}{\beta}&\leq\mathcal{F}((\mu_{\beta}^{Y_{N}})^{\otimes N-i_{N}})\\ &=N^{2}\left(\overline{\mathcal{I}}_{V}^{Y_{N}}(\mu_{\beta}^{Y_{N}})+\frac{1}{\left(N-i_{N}\right)\beta}\text{ent}[\mu_{\beta}^{Y_{N}}]\right)-\left(\frac{N-i_{N}}{2}\right)\mathcal{E}(\mu_{\beta}^{Y_{N}})\\ &\leq N^{2}\left(\min_{\rho}\overline{\mathcal{I}}_{V}^{Y_{N}}(\rho)+\frac{1}{\left(N-i_{N}\right)\beta}\text{ent}[\rho]\right)-CN.\end{split} (184)

The negative term of order (NiN)\left(N-i_{N}\right) is due to the fact that there are (NiN)(NiN1)2\frac{(N-i_{N})(N-i_{N}-1)}{2} pair of particles, and not (NiN)22\frac{(N-i_{N})^{2}}{2} pairs.

Finally, note that the last term in equation (184) is (up to a negligible error) equivalent to

minρ+(𝐑dRN1d)~VYN(ρ)+1Nβent[ρ],\min_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{RN^{-\frac{1}{d}}})}\widetilde{\mathcal{I}}_{V}^{Y_{N}}(\rho)+\frac{1}{N\beta}\text{ent}[\rho], (185)

where

~VYN(ρ)=12(empN(YN))+12𝐑d×𝐑dg(xy)𝑑ρx𝑑ρy+V~𝑑ρ,\widetilde{\mathcal{I}}_{V}^{Y_{N}}(\rho)=\frac{1}{2}\mathcal{E}_{\neq}({\rm emp}_{N}^{\prime}(Y_{N}))+\frac{1}{2}\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x-y)d\rho_{x}d\rho_{y}+\int\widetilde{V}d\rho, (186)

and the minimum is taken over ρ\rho satisfying |ρ|=1iNN|\rho|=1-\frac{i_{N}}{N}.

Using the splitting formula for the thermal equilibrium measure (Proposition 4.1) we have that

log(ZN,βYN)N2β(β(μβ)+infρ12RNλ(empN(YN)+ρμβ)1Nβ𝐑dlog(μβ)d(ρ+empN(YN))+1Nβent[ρ]),\begin{split}&-\log(Z_{N,\beta}^{Y_{N}})\leq N^{2}\beta\bigg{(}\mathcal{E}_{\beta}(\mu_{\beta})+\inf_{\rho}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{RN^{-\lambda}}}({\rm emp}_{N}^{\prime}(Y_{N})+\rho-\mu_{\beta})-\\ &\frac{1}{N\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta})d(\rho+{\rm emp}_{N}^{\prime}(Y_{N}))+\frac{1}{N\beta}{\rm ent}[\rho]\bigg{)},\end{split} (187)

where the infimum is taken over all measures ρ\rho on 𝐑dRNλ\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}} which are positive and such that

|ρ|=1iNN.|\rho|=1-\frac{i_{N}}{N}. (188)

Step 2

This step is divided into two cases. The case γ<γ\gamma<\gamma^{*} and the case γ>γ\gamma>\gamma^{*}. First we deal with the case γ<γ\gamma<\gamma^{*}.

Substep 2.1: Regime γ<γ\gamma<\gamma^{*}.

In this case, we claim that

Nλ(d+2)(1N2βlog(ZN,βYN)β(μβ))12𝐅RμV(0)(ν)+Cη+oN(1),N^{\lambda(d+2)}\left(-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\right)\leq\frac{1}{2}\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)+C\eta+o_{N}(1), (189)

where 𝐅RμV(0)\mathbf{F}_{\square_{R}}^{\mu_{V}(0)} is given by (67). The proof will be divided into 33 further subsubsteps.

Subsubstep 2.1.1

We now begin the proof of the claim. Using the scaling relations

ent[ρ]=Nλdent[ρNλ]{\rm ent}[\rho]=N^{-\lambda d}{\rm ent}[\rho^{N^{\lambda}}] (190)

and

(ρ)=Nλ(d+2)(ρNλ),\mathcal{E}(\rho)=N^{-\lambda(d+2)}\mathcal{E}(\rho^{N^{\lambda}}), (191)

we can rewrite equation (187) as

1N2βlog(ZN,βYN)β(μβ)infρ(Nλ(d+2)2R(lempN(YN)+ρμβNλ)1N1+λdβ𝐑dlog(μβNλ)d(ρ+lempN(YN))+1βN1+λdent[ρ]+CN2d),\begin{split}&-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\leq\\ &\inf_{\rho}\Bigg{(}\frac{N^{-\lambda(d+2)}}{2}\mathcal{E}^{\neq}_{\square_{R}}({\rm lemp}_{N}(Y_{N})+\rho-\mu_{\beta}^{N^{\lambda}})-\\ &\frac{1}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+{\rm lemp}_{N}(Y_{N}))+\frac{1}{\beta N^{1+\lambda d}}{\rm ent}[\rho]+CN^{-\frac{2}{d}}\Bigg{)},\end{split} (192)

where the infimum is taken over all ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}) such that

|ρ|=Nλd(1iNN).|\rho|=N^{\lambda d}\left(1-\frac{i_{N}}{N}\right). (193)

Next we will argue that we can deal with ν\nu instead of lempN(YN){\rm lemp}_{N}(Y_{N}) since we make a small error when approximating ν\nu by lempN(YN){\rm lemp}_{N}(Y_{N}). This will be the subject of the next subsubstep.

Substep 2.1.2

First of all, note that the constraint |ρ|=Nλd(1iNN)|\rho|=N^{\lambda d}\left(1-\frac{i_{N}}{N}\right) can be replaced by the constraint |ρ|=Nλd|ν||\rho|=N^{\lambda d}-|\nu| while making a negligible error because of equation (169). We now introduce the functionals 𝐋\mathbf{L}^{*} and 𝐋\mathbf{L}^{*}_{\neq}, defined for measures ν\nu and lempN{\rm lemp}_{N} on R\square_{R} as

𝐋(ν)=infρ12(ν+ρμβNλ)Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(ρ+ν)+Nλ(d+2)βN1+λdent[ρ],\begin{split}&\mathbf{L}^{*}(\nu)=\\ &\inf_{\rho}\frac{1}{2}\mathcal{E}(\nu+\rho-\mu_{\beta}^{N^{\lambda}})-\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+\nu)+\frac{N^{\lambda(d+2)}}{\beta N^{1+\lambda d}}{\rm ent}[\rho],\end{split} (194)

and

𝐋(lempN)=infρ12R(lempN(YN))+ρμβNλ)Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(ρ+lempN(YN)))+Nλ(d+2)βN1+λdent[ρ],\begin{split}&\mathbf{L}^{*}_{\neq}({\rm lemp}_{N})=\inf_{\rho}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{R}}({\rm lemp}_{N}(Y_{N}))+\rho-\mu_{\beta}^{N^{\lambda}})-\\ &\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+{\rm lemp}_{N}(Y_{N})))+\frac{N^{\lambda(d+2)}}{\beta N^{1+\lambda d}}{\rm ent}[\rho],\end{split} (195)

where the infimum in equations (194) and (195) is taken over all ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}) such that

|ρ|=Nλd|ν|.|\rho|=N^{\lambda d}-|\nu|. (196)

Given ν+(R)\nu\in\mathcal{M}^{+}(\square_{R}) and YNRiNY_{N}\in\square_{R}^{i_{N}}, let

ρν=argminρ12(ν+ρμβNλ)Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(ρ+ν)+Nλ(d+2)βN1+λdent[ρ],\begin{split}&\rho^{*}_{\nu}=\\ &\operatorname{argmin}_{\rho}\frac{1}{2}\mathcal{E}(\nu+\rho-\mu_{\beta}^{N^{\lambda}})-\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+\nu)+\frac{N^{\lambda(d+2)}}{\beta N^{1+\lambda d}}{\rm ent}[\rho],\end{split} (197)

and similarly, let ρYN\rho^{*}_{Y_{N}} be defined as

ρYN=infρ12R(lempN(YN))+ρμβNλ)Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(ρ+lempN(YN)))+Nλ(d+2)βN1+λdent[ρ],\begin{split}&\rho^{*}_{Y_{N}}=\inf_{\rho}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{R}}({\rm lemp}_{N}(Y_{N}))+\rho-\mu_{\beta}^{N^{\lambda}})-\\ &\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+{\rm lemp}_{N}(Y_{N})))+\frac{N^{\lambda(d+2)}}{\beta N^{1+\lambda d}}{\rm ent}[\rho],\end{split} (198)

where the infimum is taken over all ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}) such that

|ρ|=Nλd|ν|.|\rho|=N^{\lambda d}-|\nu|. (199)

We assume that the infimum is achieved for clarity of exposition. Otherwise, we could repeat the argument up to an arbitrarily small error. Then we can use ρν\rho^{*}_{\nu} as a test function in equation (194) and get

𝐋(ν)𝐋(lempN(YN))12(ν)12(lempN(YN))+G(ρνμβNλ,νlempN(YN))+Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(lempNν).\begin{split}&\mathbf{L}^{*}(\nu)-\mathbf{L}^{*}_{\neq}({\rm lemp}_{N}(Y_{N}))\leq\\ &\frac{1}{2}\mathcal{E}(\nu)-\frac{1}{2}\mathcal{E}^{\neq}({\rm lemp}_{N}(Y_{N}))+G(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})})+\\ &\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}-\nu).\end{split} (200)

Similarly, we can use ρYN\rho^{*}_{Y_{N}} as test function in equation (194) and get

𝐋(lempN(YN))𝐋(ν)12(lempN(YN))12(ν)+G(ρYNμβNλ,νlempN(YN))Nλ(d+2)N1+λdβ𝐑dlog(μβNλ)d(lempNν).\begin{split}&\mathbf{L}^{*}_{\neq}({\rm lemp}_{N}(Y_{N}))-\mathbf{L}^{*}(\nu)\leq\\ &\frac{1}{2}\mathcal{E}^{\neq}({\rm lemp}_{N}(Y_{N}))-\frac{1}{2}\mathcal{E}(\nu)+G(\rho^{*}_{Y_{N}}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})})-\\ &\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}-\nu).\end{split} (201)

Using equation (168) we have that

|(lempN(YN))(ν)|Cη,|\mathcal{E}^{\neq}({\rm lemp}_{N}(Y_{N}))-\mathcal{E}(\nu)|\leq C\eta, (202)

where CC depends on ν\nu.

Then, since the points are at distance at least rN1drN^{-\frac{1}{d}} from RNλ\partial\square_{RN^{-\lambda}}, we have by Lemma 9.1, that, for xRx\notin\square_{R}

glemp(x)=glempλrNλ1d(x),g\ast{\rm lemp}(x)=g\ast{\rm lemp}\ast\lambda_{rN^{\lambda-\frac{1}{d}}}(x), (203)

(see Remark 3.2 for notation).

Using Cauchy-Schwartz we get

𝒢(ρνμβNλ,νlempN(YN))=𝒢(ρνμβNλ,νlempN(YN)λrNλ1d)(ρνμβNλ)(νlempN(YN)λrNλ1d).\begin{split}&\mathcal{G}(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})})=\\ &\mathcal{G}(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})}\ast\lambda_{rN^{\lambda-\frac{1}{d}}})\leq\\ &\sqrt{\mathcal{E}(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}})\mathcal{E}({\nu}-{{\rm lemp}_{N}(Y_{N})}\ast\lambda_{rN^{\lambda-\frac{1}{d}}})}.\end{split} (204)

Using now the hypothesis that ν\nu has LL^{\infty} regularity, along with Lemmas 9.1, 9.2, 9.3, 9.4 we have that

(νlempN(YN)λrNλ1d)Cη2,\mathcal{E}({\nu}-{{\rm lemp}_{N}(Y_{N})}\ast\lambda_{rN^{\lambda-\frac{1}{d}}})\leq C\eta^{2}, (205)

where CC depends on ν\nu and rr.

On the other hand, it is easy to see that

(ρνμβNλ)C,\mathcal{E}(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}})\leq C, (206)

where CC depends on ν\nu. Therefore

𝒢(ρνμβNλ,νlempN(YN))Cη,\mathcal{G}(\rho^{*}_{\nu}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})})\leq C\eta, (207)

where CC depends on ν\nu and rr. Similarly,

𝒢(ρYNμβNλ,νlempN(YN))Cη,\mathcal{G}(\rho^{*}_{Y_{N}}-\mu_{\beta}^{N^{\lambda}},{\nu}-{{\rm lemp}_{N}(Y_{N})})\leq C\eta, (208)

where CC depends on ν\nu and rr.

This implies that

|𝐋(ν)𝐋(lempN(YN))|Cη,\left|\mathbf{L}^{*}(\nu)-\mathbf{L}^{*}_{\neq}({\rm lemp}_{N}(Y_{N}))\right|\leq C\eta, (209)

where CC depends on ν\nu.

We have proved that we can deal with ν\nu instead of lempN(YN){\rm lemp}_{N}(Y_{N}) since we make a small error when approximating ν\nu by lempN(YN){\rm lemp}_{N}(Y_{N}). The last subsubstep will consist in proving that

lim supN𝐋(ν)12𝐅RμV(0)(ν).\limsup_{N\to\infty}\mathbf{L}^{*}(\nu)\leq\frac{1}{2}\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu). (210)

Substep 2.1.3

The proof of equation (210) will consist in taking a minimizing sequence of the problem in the RHS, and modifying it so that it is a valid test function to the problem in the LHS.

Let ρϵ0\rho_{\epsilon}\geq 0 be such that

𝐑dρϵ+νμV(0)dx=0\int_{\mathbf{R}^{d}}\rho_{\epsilon}+\nu-\mu_{V}(0)dx=0 (211)

and

(ν+ρϵμV(0))𝐅RμV(0)(ν)+ϵ.\mathcal{E}(\nu+\rho_{\epsilon}-\mu_{V}(0))\leq\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)+\epsilon. (212)

Then for every δ>0\delta>0 there exists T>0T>0 such that

|B(0,T)ρϵ+νμV(0)dx|δ\left|\int_{B(0,T)}\rho_{\epsilon}+\nu-\mu_{V}(0)dx\right|\leq\delta (213)

and

|(ν+ρϵμV(0))((ν+ρϵμV(0))𝟏B(0,T))|δ.\begin{split}&\left|\mathcal{E}\left(\nu+\rho_{\epsilon}-\mu_{V}(0)\right)-\mathcal{E}\left((\nu+\rho_{\epsilon}-\mu_{V}(0)\right)\mathbf{1}_{B(0,T)})\right|\leq\delta.\end{split} (214)

Now take a truncated ρϵη¯\rho_{\epsilon}^{\overline{\eta}} such that (LABEL:errorinenergy) and (213) hold with an error δ+η¯\delta+\overline{\eta} in the right hand side, and in addition

ρϵη¯L.\rho_{\epsilon}^{\overline{\eta}}\in L^{\infty}. (215)

Note that ρϵη¯\rho_{\epsilon}^{\overline{\eta}} exists because the sequence

ρϵ𝟏|ρϵ|<M\rho_{\epsilon}\mathbf{1}_{|\rho_{\epsilon}|<M} (216)

is bounded, and by Dominated Convergence Theorem, its integral converges to the integral of ρϵ\rho_{\epsilon} as MM\to\infty.

Now define ρϵ,Tη¯\rho_{\epsilon,T}^{\overline{\eta}} as

ρϵ,Tη¯=ρϵη¯𝟏B(0,T)+μβNλ.\rho_{\epsilon,T}^{\overline{\eta}}=\rho_{\epsilon}^{\overline{\eta}}\mathbf{1}_{B(0,T)}+\mu_{\beta}^{N^{\lambda}}. (217)

Note that

|𝐑dlog(μβNλ)d(ρϵ,Tη¯+ν)+ent[ρϵ,Tη¯]|C,\left|\int_{\mathbf{R}^{d}}-\log(\mu_{\beta}^{N^{\lambda}})d(\rho_{\epsilon,T}^{\overline{\eta}}+\nu)+{\rm ent}[\rho_{\epsilon,T}^{\overline{\eta}}]\right|\leq\\ C, (218)

where CC depends on TT and η¯\overline{\eta} but does not depend on N.N. Since we are in the regime γ<γ,\gamma<\gamma^{*}, we have that

1+λdγ>λ(d+2),1+\lambda d-\gamma>\lambda(d+2), (219)

and therefore

limNNλ(d+2)N1+λdβ|𝐑dlog(μβNλ)d(ρϵ,Tη¯+ν)+ent[ρϵ,Tη¯]|=0.\lim_{N\to\infty}\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\left|\int_{\mathbf{R}^{d}}-\log(\mu_{\beta}^{N^{\lambda}})d(\rho_{\epsilon,T}^{\overline{\eta}}+\nu)+{\rm ent}[\rho_{\epsilon,T}^{\overline{\eta}}]\right|=0. (220)

Using ρϵ,Tη¯\rho_{\epsilon,T}^{\overline{\eta}} as a test function in the definition of 𝐋(ν)\mathbf{L}^{*}(\nu), and appealing once again to Remark 4.1 we have that

lim supN𝐋(ν)lim supN12(ν+ρϵ,Tη¯μβNλ)+Nλ(d+2)N1+λdβ|𝐑dlog(μβNλ)d(ρϵ,Tη¯+ν)+ent[ρϵ,Tη¯]|lim supN12((ν+ρϵ,Rη¯μβNλ)𝟏B(0,T))=12((ν+ρϵ,Rη¯μV(0)))𝟏B(0,T))𝐅RμV(0)(ν)2+ϵ+δ+η¯.\begin{split}&\limsup_{N\to\infty}\mathbf{L}^{*}(\nu)\\ \leq&\limsup_{N\to\infty}\frac{1}{2}\mathcal{E}(\nu+\rho_{\epsilon,T}^{\overline{\eta}}-\mu_{\beta}^{N^{\lambda}})+\\ &\quad\frac{N^{\lambda(d+2)}}{N^{1+\lambda d}\beta}\left|\int_{\mathbf{R}^{d}}-\log(\mu_{\beta}^{N^{\lambda}})d(\rho_{\epsilon,T}^{\overline{\eta}}+\nu)+{\rm ent}[\rho_{\epsilon,T}^{\overline{\eta}}]\right|\\ \leq&\limsup_{N\to\infty}\frac{1}{2}\mathcal{E}((\nu+\rho_{\epsilon,R}^{\overline{\eta}}-\mu_{\beta}^{N^{\lambda}})\mathbf{1}_{B(0,T)})\\ =&\frac{1}{2}\mathcal{E}\left((\nu+\rho_{\epsilon,R}^{\overline{\eta}}-\mu_{V}(0))\right)\mathbf{1}_{B(0,T)})\\ \leq&\frac{\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)}{2}+\epsilon+\delta+\overline{\eta}.\end{split} (221)

Since ϵ,δ,η¯\epsilon,\delta,\overline{\eta} are arbitrary, we conclude

lim supN𝐋(ν)12𝐅RμV(0)(ν).\limsup_{N\to\infty}\mathbf{L}^{*}(\nu)\leq\frac{1}{2}\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu). (222)

The proof of substep 2.1 is now complete.

Substep 2.2 Now we deal with the case γ>γ.\gamma>\gamma^{*}. In this case we go back to working in unreescaled coordinates.

We start with formula (187). Since in the regime γ>γ,\gamma>\gamma^{*}, we expect the term

R(empN(YN)+ρμβ)\mathcal{E}^{\neq}_{\square_{R}}({\rm emp}_{N}^{\prime}(Y_{N})+\rho-\mu_{\beta}) (223)

to be negligible, we focus on the remaining part of the functional, i.e.

infρ𝐑dlog(μβ)d(ρ+empN(YN))+ent[ρ],\inf_{\rho}-\int_{\mathbf{R}^{d}}\log(\mu_{\beta})d(\rho+{\rm emp}_{N}^{\prime}(Y_{N}))+{\rm ent}[\rho], (224)

where the infimum is taken over all measures ρ\rho on 𝐑dRNλ\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}} which are positive and such that

|ρ|=1iNN.|\rho|=1-\frac{i_{N}}{N}. (225)

The minimizer in equation (224) can be easily found by adding a Lagrange multiplier for the mass constraint. It can be easily checked that the unique minimizer of (224) in the corresponding space is given by ρ\rho^{*}, where

ρ=αμβ𝟏𝐑dRNλ,\rho^{*}=\alpha\mu_{\beta}\mathbf{1}_{\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}}}, (226)

where

α=1iNN𝐑dRNλμβ𝑑x.\alpha=\frac{1-\frac{i_{N}}{N}}{\int_{\mathbf{R}^{d}\setminus\square_{RN^{-\lambda}}}\mu_{\beta}dx}. (227)

Using the identity

ent[Aμ]=Aent[μ]+Alog(A)μ𝑑x,{\rm ent}[A\mu]=A{\rm ent}[\mu]+A\log(A)\int\mu dx, (228)

valid for any A𝐑+A\in\mathbf{R}^{+}, we have that

𝐑dlog(μβ)d(ρ+empN(YN))+ent[ρ]=𝐑dlog(μβ)d(empN(YN))+(1iNN)log(α).\begin{split}&\int_{\mathbf{R}^{d}}\log(\mu_{\beta})d(\rho^{*}+{\rm emp}_{N}^{\prime}(Y_{N}))+{\rm ent}[\rho^{*}]=\\ &\int_{\mathbf{R}^{d}}\log(\mu_{\beta})d({\rm emp}_{N}^{\prime}(Y_{N}))+\left(1-\frac{i_{N}}{N}\right)\log(\alpha).\end{split} (229)

It can be checked that, as a consequence of equation (169), limNα=1\lim_{N\to\infty}\alpha=1, and therefore we may use the approximation logαα1\log\alpha\simeq\alpha-1. Proceeding as in the proof of Lemma 5.1 and using equation (169), we have that

limNNλd(1iNN)log(α)=RdμV(0)|ν|.\lim_{N\to\infty}N^{\lambda d}\left(1-\frac{i_{N}}{N}\right)\log(\alpha)=R^{d}\mu_{V}(0)-|\nu|. (230)

Note that

limN0N1+λdβNλ(d+2)(ρμβ)=0\lim_{N\to 0}\frac{N^{1+\lambda d}\beta}{N^{\lambda(d+2)}}\mathcal{E}(\rho^{*}-\mu_{\beta})=0 (231)

since we are in the regime γ>γ\gamma>\gamma^{*}. Using again formula (187) and switching to rescaled coordinates, we have

N1+λdβ(1N2βlog(ZN,βYN)β(μβ))𝐑dlog(μβNλ)d(lempN(YN))|ν|+RdμV(0)+oN(1),\begin{split}&N^{1+\lambda d}\beta\left(-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\right)\leq\\ &\quad\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))-|\nu|+R^{d}\mu_{V}(0)+o_{N}(1),\end{split} (232)

where o(1)o(1) is independent of YN.Y_{N}.

Lemma 7.2 is proved for γ>γ\gamma>\gamma^{*}.

Step 3

This step only deals with the case γ<γ.\gamma<\gamma^{*}. Once again we work with rescaled coordinates.

We now claim that for any measure ν\nu on R\square_{R} such that (ν)<\mathcal{E}(\nu)<\infty we have

𝐅RμV(0)(ν)=ΦRμV(0)(ν).\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)=\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu). (233)

In other words, we claim that we can drop the mass constraint. We now prove the claim. Since clearly

𝐅RμV(0)(ν)ΦRμV(0)(ν),\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)\geq\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu), (234)

we will prove that

𝐅RμV(0)(ν)ΦRμV(0)(ν).\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)\leq\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu). (235)

In order to prove this claim, we reformulate the definition of ΦRμV(0)(ν)\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu) as as

ΦRμV(0)(ν)=infρ(νμV(0)𝟏R+ρ)\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)=\inf_{\rho}\mathcal{E}(\nu-\mu_{V}(0)\mathbf{1}_{\square_{R}}+\rho) (236)

where the infimum is taken over all ρ\rho such that ρ\rho is supported in 𝐑dR\mathbf{R}^{d}\setminus\square_{R} and ρμV(0).\rho\geq-\mu_{V}(0).

Let ϵ>0\epsilon>0 and let ρϵC0\rho_{\epsilon}\in C^{\infty}_{0} be such that ρϵ\rho_{\epsilon} is supported in 𝐑dR\mathbf{R}^{d}\setminus\square_{R}, ρϵμV(0)\rho_{\epsilon}\geq-\mu_{V}(0) and

(νμV(0)𝟏R+ρϵ)ΦRμV(0)(ν)+ϵ.\mathcal{E}(\nu-\mu_{V}(0)\mathbf{1}_{\square_{R}}+\rho_{\epsilon})\leq\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)+\epsilon. (237)

Let

K=supp(ρϵ),K=\text{supp}(\rho_{\epsilon}), (238)

and let

E=|ρϵ|(|ν|RdμV(0)).E=|\rho_{\epsilon}|-(|\nu|-R^{d}\mu_{V}(0)). (239)

Let RnR_{n} be a sequence such that RnR_{n} tends to \infty monotonically, and

KB(0,R1).K\subset B(0,R_{1}). (240)

Define

ρϵn=ρϵ+E(B(0,2Rn)B(0,Rn))𝟏B(0,2Rn)B(0,Rn),\rho_{\epsilon}^{n}=\rho_{\epsilon}+\frac{E}{\mathcal{L}(B(0,2R_{n})\setminus B(0,R_{n}))}\mathbf{1}_{B(0,2R_{n})\setminus B(0,R_{n})}, (241)

where \mathcal{L} denotes the Lebesgue measure. Then it’s easy to see that

|ρϵn|=(|ν|RdμV(0)),|\rho_{\epsilon}^{n}|=-(|\nu|-R^{d}\mu_{V}(0)), (242)

and

limn(νμV(0)𝟏R+ρϵn)=(νμV(0)𝟏R+ρϵ).\lim_{n\to\infty}\mathcal{E}(\nu-\mu_{V}(0)\mathbf{1}_{\square_{R}}+\rho_{\epsilon}^{n})=\mathcal{E}(\nu-\mu_{V}(0)\mathbf{1}_{\square_{R}}+\rho_{\epsilon}). (243)

Therefore

𝐅RμV(0)(ν)ΦRμV(0)(ν)+ϵ.\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)\leq\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)+\epsilon. (244)

Since ϵ\epsilon is arbitrary, we conclude that

𝐅RμV(0)(ν)=ΦRμV(0)(ν).\mathbf{F}_{\square_{R}}^{\mu_{V}(0)}(\nu)=\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu). (245)

We can now complete the proof of Theorem 2.1 by giving the lower bound.

Proof of Theorem 2.1, lower bound.

We start with the case γ<γ.\gamma<\gamma^{*}. Let ν\nu be a measure on R.\square_{R}. Using a density argument, we may assume that νL\nu\in L^{\infty} and ent[ν]<.{\rm ent}[\nu]<\infty. Let ϵ,η,δ>0\epsilon,\eta,\delta>0 and let Λδη,ϵ\Lambda_{\delta}^{\eta,\epsilon} be as in Lemma 7.1 with Ω=R\Omega=\square_{R} and μN=μβNλ|R|μβNλ|R|\mu_{N}=\frac{\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}}{|\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}|}, n=|ν|N1λdn=|\nu|N^{1-\lambda d} (rounded to an integer) and ν¯=ν|ν|\overline{\nu}=\frac{\nu}{|\nu|}. Note that equation (168) is satisfied with this choice of nn. We claim that equation (165) implies that there exists CC which depends on ν\nu such that for any XNΛδη,ϵX_{N}\in\Lambda_{\delta}^{\eta,\epsilon},

ΦR,μV(0)(empN(XN))ΦRμV(0)(ν)+Cη.\Phi^{\mu_{V}(0)}_{\square_{R},\neq}({\rm emp}_{N}(X_{N}))\leq\Phi^{\mu_{V}(0)}_{\square_{R}}(\nu)+C\eta. (246)

This is because for

ρ=argminρ+(𝐑dR)(ν+ρμV(0))\rho^{*}=\operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\mathcal{E}(\nu+\rho-\mu_{V}(0)) (247)

we have

ΦR,μV(0)(empN(XN))(empN(XN)+ρμV(0))=(empN(XN)+νν+ρμV(0))=(ν+ρμV(0))+2𝒢(ν+ρμV(0),empN(XN)ν)+(empN(XN)ν)ΦRμV(0)(ν)+Cη+η2,\begin{split}\Phi^{\mu_{V}(0)}_{\square_{R},\neq}({\rm emp}_{N}(X_{N}))&\leq\mathcal{E}^{\neq}({\rm emp}_{N}(X_{N})+\rho^{*}-\mu_{V}(0))\\ &=\mathcal{E}^{\neq}({\rm emp}_{N}(X_{N})+\nu-\nu+\rho^{*}-\mu_{V}(0))\\ &=\mathcal{E}(\nu+\rho^{*}-\mu_{V}(0))+\\ &\quad 2\mathcal{G}(\nu+\rho^{*}-\mu_{V}(0),{\rm emp}_{N}(X_{N})-\nu)+\\ &\quad\mathcal{E}^{\neq}({\rm emp}_{N}(X_{N})-\nu)\\ &\leq\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)+C\eta+\eta^{2},\end{split} (248)

where CC depends on ν\nu. Using Lemma 7.2 and equation (114) we then have that

𝐏N,β(lempN(YN)B(ν,ϵ))𝐏N,β(YNλΛδη,ϵ)=1ZN,βYNΛδη,ϵZN,βYN𝑑YNYNΛδη,ϵexp(βN2λ(d+2)[12ΦRμV(0)(ν)+Cη+oN(1)])𝑑YN=exp(βN2λ(d+2)[12ΦRμV(0)(ν)+Cη+oN(1)])YNΛδη,ϵ𝑑YN.\begin{split}&\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))\geq\\ &\mathbf{P}_{N,\beta}(Y_{N}^{\lambda}\in\Lambda_{\delta}^{\eta,\epsilon})=\\ &\frac{1}{Z_{N,\beta}}\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}Z_{N,\beta}^{Y_{N}}dY_{N}\geq\\ &\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\exp\left(-\beta N^{2-\lambda(d+2)}\left[\frac{1}{2}\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)+C\eta+o_{N}(1)\right]\right)dY_{N}=\\ &\exp\left(-\beta N^{2-\lambda(d+2)}\left[\frac{1}{2}\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)+C\eta+o_{N}(1)\right]\right)\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}dY_{N}.\end{split} (249)

Since we are in the regime γ<γ\gamma<\gamma^{*} we have

|log(YNΛδη,ϵ𝑑YN)|CN1λdent[ν]=o(βN2λ(d+2)).\begin{split}\left|\log\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}dY_{N}\right)\right|&\leq CN^{1-\lambda d}{\rm ent}[\nu]\\ &=o(\beta N^{2-\lambda(d+2)}).\end{split} (250)

Therefore

lim infN1βN2λ(d+2)log(𝐏N,β(lempN(YN)B(ν,ϵ)))12ΦRμV(0)(ν)Cη.\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))\right)\geq-\frac{1}{2}\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu)-C\eta. (251)

Since η\eta is arbitrary, we can conclude that

lim infN1βN2λ(d+2)log(𝐏N,β(lempN(YN)B(ν,ϵ)))12ΦRμV(0)(ν).\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))\right)\geq-\frac{1}{2}\Phi_{\square_{R}}^{\mu_{V}(0)}(\nu). (252)

Now we proceed with the case γ>γ.\gamma>\gamma^{*}. Let ν\nu be a positive measure in R\square_{R}, let ϵ,η,δ>0\epsilon,\eta,\delta>0 and let Λδη,ϵ\Lambda_{\delta}^{\eta,\epsilon} be as in Lemma 7.1 with Ω=R\Omega=\square_{R} and μN=μβNλ|R|μβNλ|R|\mu_{N}=\frac{\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}}{|\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}|}, n=|ν|N1λdn=|\nu|N^{1-\lambda d} (rounded to an integer) and ν¯=ν|ν|\overline{\nu}=\frac{\nu}{|\nu|}. Then, starting as in the previous case, we have

𝐏N,β(lempN(YN)B(ν,ϵ))𝐏N,β(YNλΛδη,ϵ)=1ZN,βYNΛδη,ϵZN,βYN𝑑YN.\begin{split}\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))&\geq\mathbf{P}_{N,\beta}(Y_{N}^{\lambda}\in\Lambda_{\delta}^{\eta,\epsilon})\\ &=\frac{1}{Z_{N,\beta}}\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}Z_{N,\beta}^{Y_{N}}dY_{N}.\end{split} (253)

We then have that

lim infN1N1λdlog(𝐏N,β(lempN(YN)B(ν,ϵ)))lim infN1ZN,β1N1λdlog(YNΛδη,ϵZN,βYN𝑑YN)=lim infN1N1λdlog(YNΛδη,ϵexp(N1λd[Rlog(μβNλ)dlempN)dYN|ν|+RdμV(0)]).\begin{split}&\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))\right)\geq\\ &\liminf_{N\to\infty}\frac{1}{Z_{N,\beta}}\frac{1}{N^{1-\lambda d}}\log\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}Z_{N,\beta}^{Y_{N}}dY_{N}\right)=\\ &\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\\ &\quad\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\exp\left(-N^{1-\lambda d}\left[\int_{\square_{R}}\log(\mu_{\beta}^{N^{\lambda}})d\,{\rm lemp}_{N}\right)dY_{N}-|\nu|+R^{d}\mu_{V}(0)\right]\right).\end{split} (254)

Recalling that

lempN(YN)=1N1λdi=1iNδyiλ,{\rm lemp}_{N}(Y_{N})=\frac{1}{N^{1-\lambda d}}\sum_{i=1}^{i_{N}}\delta_{y_{i}^{\lambda}}, (255)

and iN=N1λd|ν|i_{N}=N^{1-\lambda d}|\nu| (rounded to an integer), we have that

lim infN1N1λdlog(𝐏N,β(lempN(YN)B(ν,ϵ)))lim infN1N1λdlog(YNΛδη,ϵexp(N1λd[Rlog(μβNλ)dlempN)dYN|ν|+RdμV(0)])=|ν|RdμV(0)+lim infN1N1λdlog(YNΛδη,ϵΠi=1iNμβNλ(yi)𝑑YN).\begin{split}&\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon))\right)\geq\\ &\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\\ &\quad\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\exp\left(-N^{1-\lambda d}\left[\int_{\square_{R}}\log(\mu_{\beta}^{N^{\lambda}})d\,{\rm lemp}_{N}\right)dY_{N}-|\nu|+R^{d}\mu_{V}(0)\right]\right)=\\ &|\nu|-R^{d}\mu_{V}(0)+\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dY_{N}\right).\end{split} (256)

By construction we have that

lim infN1N1λdlog(YNΛδη,ϵΠi=1iNμβNλ(yi)𝑑yi)lim infNent[ν|μβNλ]δ.\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dy_{{i}}\right)\geq-\liminf_{N\to\infty}{\rm ent}[\nu|\mu_{\beta}^{N^{\lambda}}]-\delta. (257)

Combining the last equation with Remark 4.1 we have

lim infN1N1λdlog(YNΛδη,ϵΠi=1iNμβNλ(yi)𝑑yi)ent[ν|μV(0)𝟏R]δ,\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\int_{Y_{N}\in\Lambda_{\delta}^{\eta,\epsilon}}\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dy_{{i}}\right)\geq-{\rm ent}[\nu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]-\delta, (258)

and therefore

lim infN1N1λdlog(𝐏N,β(lempN(YN)B(ν,ϵ)))ent[ν|μV(0)𝟏R]+|ν|RdμV(0)δ.\begin{split}&\liminf_{N\to\infty}\frac{1}{N^{1-\lambda d}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}(Y_{N})\in B(\nu,\epsilon)\right)\right)\geq\\ &-{\rm ent}[\nu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]+|\nu|-R^{d}\mu_{V}(0)-\delta.\end{split} (259)

Since δ\delta is arbitrary, we can conclude.

8 Proof of statement about regime γ=γ\gamma=\gamma^{*}

In this section, we prove the third part of Theorem 2.1, which we repeat here for convenience: If γ=γ\gamma=\gamma^{*} (critical regime) and νL\nu\in L^{\infty} then

limϵ0lim supN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))=0.\lim_{\epsilon\to 0}\limsup_{N\to\infty}\left(\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))\right)+\mathcal{T}^{N}_{\lambda}(\nu)\right)=0. (260)

Similarly,

limϵ0lim infN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))=0.\lim_{\epsilon\to 0}\liminf_{N\to\infty}\left(\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))\right)+\mathcal{T}^{N}_{\lambda}(\nu)\right)=0. (261)

Before starting the proof, we note that since we are in the critical regime, we have βN2λ(d+2)=N1+λd\beta N^{2-\lambda(d+2)}=N^{1+\lambda d}.

Proof of lim inf\liminf inequality.

Let ν\nu be a positive measure in R\square_{R}, let ϵ,η,δ>0\epsilon,\eta,\delta>0 and let Λδη,ϵ\Lambda_{\delta}^{\eta,\epsilon} be as in Lemma 7.1 with Ω=R\Omega=\square_{R} and μN=μβNλ|R|μβNλ|R|\mu_{N}=\frac{\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}}{|\mu_{\beta}^{N^{\lambda}}|_{\square_{R}}|}, n=|ν|N1λdn=|\nu|N^{1-\lambda d} (rounded to an integer) and ν¯=ν|ν|\overline{\nu}=\frac{\nu}{|\nu|}. Then we have that

1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))1βN2λ(d+2)log(Λδη,ϵZN,βYN𝑑YN).\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in{B}(\nu,\epsilon)\right)\right)\geq\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\int_{\Lambda_{\delta}^{\eta,\epsilon}}Z_{N,\beta}^{Y_{N}}dY_{N}\right). (262)

Using equation (192), and the hypothesis that γ=γ,\gamma=\gamma^{*}, we have

1N2βlog(ZN,βYN)β(μβ)infρNλ(d+2)(12(lempN(YN)+ρμβNλ)𝐑dlog(μβNλ)d(ρ+lempN(YN))+ent[ρ])Nλ(d+2)(𝐓λN(ν)+𝐑dlog(μβNλ)d(lempN(YN))Cη),\begin{split}&-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\leq\\ &\inf_{\rho}N^{-\lambda(d+2)}\Bigg{(}\frac{1}{2}\mathcal{E}^{\neq}({\rm lemp}_{N}(Y_{N})+\rho-\mu_{\beta}^{N^{\lambda}})-\\ &\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+{\rm lemp}_{N}(Y_{N}))+{\rm ent}[\rho]\Bigg{)}\leq\\ &N^{-\lambda(d+2)}\Bigg{(}\mathbf{T}^{N}_{\lambda}(\nu)+\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))-C\eta\Bigg{)},\end{split} (263)

where the infimum is taken over ρ\rho such that

𝐑dlempN+ρμβNλ=0,\int_{\mathbf{R}^{d}}{\rm lemp}_{N}+\rho-\mu_{\beta}^{N^{\lambda}}=0, (264)

and 𝐓λN,𝐓λN,\mathbf{T}^{N}_{\lambda},\mathbf{T}^{N,\neq}_{\lambda} are given by equations (40) and (65) respectively. We have used that, if νL\nu\in L^{\infty} then

|𝐓λN(ν)𝐓λN,(lempN(YN))|Cη,\left|\mathbf{T}^{N}_{\lambda}(\nu)-\mathbf{T}^{N,\neq}_{\lambda}({\rm lemp}_{N}(Y_{N}))\right|\leq C\eta, (265)

where CC depends on ν.\nu. The proof of this statement is the same as the proof that

|𝐋(ν)𝐋(lempN(YN))|Cη,\left|\mathbf{L}^{*}(\nu)-\mathbf{L}^{*}_{\neq}({\rm lemp}_{N}(Y_{N}))\right|\leq C\eta, (266)

where CC depends on ν\nu, see the proof of Lemma 7.2, step 2.1 (in fact, in the critical regime, we have that 𝐓λN=𝐋\mathbf{T}^{N}_{\lambda}=\mathbf{L}^{*} and 𝐓λN,=𝐋\mathbf{T}^{N,\neq}_{\lambda}=\mathbf{L}^{*}_{\neq}).

Therefore we can rewrite equation (263) as

lim infN1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))lim infN1βN2λ(d+2)log(Λδη,ϵZN,βYN𝑑YN)lim infN1βN2λ(d+2)log(Λδη,ϵexp(βN2λ(d+2)(𝐓λN(ν)+Cη+𝐑dlog(μβNλ)d(lempN(YN))))dYN)=lim infN1βN2λ(d+2)log(Λδη,ϵexp(βN2λ(d+2)(𝐓λN(ν)+Cη))Πi=1iNμβNλ(yi)𝑑yi)=lim infN1βN2λ(d+2)log(exp(βN2λ(d+2)𝐓λN(ν)+Cη))ent[ν|μV(0)𝟏R]δ=lim infN𝒯λN(ν)C(δ+η),\begin{split}&\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in{B}(\nu,\epsilon)\right)\right)\\ &\geq\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\int_{\Lambda_{\delta}^{\eta,\epsilon}}Z_{N,\beta}^{Y_{N}}dY_{N}\right)\\ &\geq\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\Bigg{(}\int_{\Lambda_{\delta}^{\eta,\epsilon}}\exp\Bigg{(}-\beta N^{2-\lambda(d+2)}\Bigg{(}\mathbf{T}^{N}_{\lambda}(\nu)+C\eta+\\ &\quad\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))\Bigg{)}\Bigg{)}dY_{N}\Bigg{)}\\ &=\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\\ &\quad\left(\int_{\Lambda_{\delta}^{\eta,\epsilon}}\exp\left(-\beta N^{2-\lambda(d+2)}\Bigg{(}\mathbf{T}^{N}_{\lambda}(\nu)+C\eta\Bigg{)}\right)\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dy_{i}\right)\\ &=\liminf_{N\to\infty}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\exp\left(-\beta N^{2-\lambda(d+2)}\mathbf{T}^{N}_{\lambda}(\nu)+C\eta\right)\right)-\\ &\quad\mbox{ent}[\nu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]-\delta\\ &=\liminf_{N\to\infty}-\mathcal{T}^{N}_{\lambda}(\nu)-C(\delta+\eta),\end{split} (267)

where CC depends on ν\nu.

Since η\eta and δ\delta are arbitrary, we have

lim infN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))0.\liminf_{N\to\infty}\left(\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))\right)+\mathcal{T}^{N}_{\lambda}(\nu)\right)\geq 0. (268)

In particular, this implies

limϵ0lim infN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+(𝒯λN(ν)))0.\lim_{\epsilon\to 0}\liminf_{N\to\infty}\Bigg{(}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in{B}(\nu,\epsilon)\right)\right)+\\ \left(\mathcal{T}^{N}_{\lambda}(\nu)\right)\Bigg{)}\geq 0. (269)

We now turn to the proof of the lim sup\limsup inequality:

Proof of lim sup\limsup inequality.

We start with equation (192), which in the critical regime γ=γ\gamma=\gamma^{*} reads

1N2βlog(ZN,βYN)β(μβ)=Nλ(d+2)infρ(12R(lempN(YN)+ρμβNλ)𝐑dlog(μβNλ)d(ρ+lempN(YN))+ent[ρ]+oN(1))=Nλ(d+2)(𝐓λN(lempN(YN))+𝐑dlog(μβNλ)d(lempN(YN))+oN(1)),\begin{split}&-\frac{1}{N^{2}\beta}\log(Z_{N,\beta}^{Y_{N}})-\mathcal{E}_{\beta}(\mu_{\beta})\\ =&N^{-\lambda(d+2)}\inf_{\rho}\Bigg{(}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{R}}({\rm lemp}_{N}(Y_{N})+\rho-\mu_{\beta}^{N^{\lambda}})-\\ &\quad\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d(\rho+{\rm lemp}_{N}(Y_{N}))+{\rm ent}[\rho]+o_{N}(1)\Bigg{)}=\\ &N^{-\lambda(d+2)}\left(\mathbf{T}^{N\neq}_{\lambda}({\rm lemp}_{N}(Y_{N}))+\int_{\mathbf{R}^{d}}\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))+o_{N}(1)\right),\end{split} (270)

where the infimum in line 22 of the last equation is taken over all ρ+(𝐑dR)\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R}) such that

|ρ|=Nλd(1iNN).|\rho|=N^{\lambda d}\left(1-\frac{i_{N}}{N}\right). (271)

We proceed by writing

𝐏N,β(lempNB(ν,ϵ))=lempNB(ν,ϵ)ZN,βYN𝑑YN=lempNB(ν,ϵ)exp(βN2λ(d+2)[𝐓λN(lempN(YN))+log(μβNλ)d(lempN(YN))])dYN=lempNB(ν,ϵ)exp(βN2λ(d+2)𝐓λN(lempN(YN)))Πi=1iNμβNλ(yi)𝑑yiexp(βN2λ(d+2)infμB(ν,ϵ)𝒜iNNλd1(R)𝐓λN,(μ))lempNB(ν,ϵ)Πi=1iNμβNλ(yi)𝑑yiexp(βN2λ(d+2)[infμB(ν,ϵ)𝒜iNNλd1(R)𝐓λN(μ)+infμB(ν,ϵ)ent[μ|μβNλ𝟏R]]).\begin{split}&\mathbf{P}_{N,\beta}({\rm lemp}_{N}\in{B}(\nu,\epsilon))=\\ &\int_{{\rm lemp}_{N}\in{B}(\nu,\epsilon)}Z_{N,\beta}^{Y_{N}}dY_{N}=\\ &\int_{{\rm lemp}_{N}\in{B}(\nu,\epsilon)}\exp\Bigg{(}-\beta N^{2-\lambda(d+2)}[\mathbf{T}^{N\neq}_{\lambda}({\rm lemp}_{N}(Y_{N}))+\\ &\quad\int\log(\mu_{\beta}^{N^{\lambda}})d({\rm lemp}_{N}(Y_{N}))]\Bigg{)}dY_{N}=\\ &\int_{{\rm lemp}_{N}\in{B}(\nu,\epsilon)}\exp\left(-\beta N^{2-\lambda(d+2)}\mathbf{T}^{N\neq}_{\lambda}({\rm lemp}_{N}(Y_{N}))\right)\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dy_{i}\leq\\ &\exp\left(-\beta N^{2-\lambda(d+2)}\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\mathbf{T}^{N,\neq}_{\lambda}(\mu)\right)\\ &\quad\int_{{\rm lemp}_{N}\in{B}(\nu,\epsilon)}\Pi_{i=1}^{i_{N}}\mu_{\beta}^{N^{\lambda}}(y_{i})dy_{i}\leq\\ &\exp\left(-\beta N^{2-\lambda(d+2)}\left[\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\mathbf{T}^{N\neq}_{\lambda}(\mu)+\inf_{\mu\in{B}(\nu,\epsilon)}\mbox{ent}[\mu|\mu_{\beta}^{N^{\lambda}}\mathbf{1}_{\square_{R}}]\right]\right).\end{split} (272)

Letting NN tend to \infty and using Remark 4.1 we have that

lim supN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+(infμB(ν,ϵ)𝒜iNNλd1(R)𝐓λN,(μ)+infμB(ν,ϵ)ent[μ|μV(0)𝟏R]))0.\begin{split}\limsup_{N\to\infty}\Bigg{(}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in{B}(\nu,\epsilon)\right)\right)+\\ \left(\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\mathbf{T}^{N,\neq}_{\lambda}(\mu)+\inf_{\mu\in{B}(\nu,\epsilon)}\mbox{ent}[\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]\right)\Bigg{)}\leq 0.\end{split} (273)

It’s well known that ent[ν|μ][\nu|\mu] is l.s.c. in ν\nu for fixed μ.\mu. Therefore

limϵ0infμB(ν,ϵ)ent[μ|μV(0)𝟏R]=ent[ν|μV(0)𝟏R].\lim_{\epsilon\to 0}\inf_{\mu\in{B}(\nu,\epsilon)}\mbox{ent}[\mu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]=\mbox{ent}[\nu|\mu_{V}(0)\mathbf{1}_{\square_{R}}]. (274)

We will also use a property of 𝐓λN,,\mathbf{T}^{N,\neq}_{\lambda}, which we prove at the end of this section: we will show that

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)𝒜iNNλd1(R)𝐓λN,(μ))=0.\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\mathbf{T}^{N,\neq}_{\lambda}(\mu)\right)=0. (275)

Using equation (275), we have

limϵ0lim supN(1βN2λ(d+2)log(𝐏N,β(lempNB(ν,ϵ)))+𝒯λN(ν))0.\lim_{\epsilon\to 0}\limsup_{N\to\infty}\Bigg{(}\frac{1}{\beta N^{2-\lambda(d+2)}}\log\left(\mathbf{P}_{N,\beta}\left({\rm lemp}_{N}\in{B}(\nu,\epsilon)\right)\right)+\mathcal{T}^{N}_{\lambda}(\nu)\Bigg{)}\leq 0. (276)

This concludes the proof. ∎

We now prove equation (275), used in the proof and restated here for convenience.

Lemma 8.1.

Let ν\nu be a measure on R\square_{R} such that (ν)<\mathcal{E}(\nu)<\infty and νL\nu\in L^{\infty}, and iNi_{N} be such that

limNiNN1λd=|ν|.\lim_{N\to\infty}\frac{i_{N}}{N^{1-\lambda d}}=|\nu|. (277)

Then

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)𝒜iNNλd1(R)𝐓λN,(μ))=0.\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\mathbf{T}^{N,\neq}_{\lambda}(\mu)\right)=0. (278)
Proof.

Let μB(ν,ϵ)𝒜iNNλd1(R)\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R}), and let

μ=μλNλ1d,\mu^{*}=\mu\ast\lambda_{N^{\lambda-\frac{1}{d}}}, (279)

(see Remark 3.2 for notation). We claim that

𝐓λN(μ)𝐓λN,(μ)+CN2(λ1d),\mathbf{T}^{N}_{\lambda}(\mu^{*})\leq\mathbf{T}^{N,\neq}_{\lambda}(\mu)+CN^{2\left(\lambda-\frac{1}{d}\right)}, (280)

where CC depends on ν\nu. To see this, let

ρ=argminρ+(𝐑dR)(12R(μ+ρμβNλ)𝐑dlog(μβNλ)𝑑ρ+ent[ρ]),\rho^{*}=\\ \operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\Bigg{(}\frac{1}{2}\mathcal{E}^{\neq}_{\square_{R}}\left(\mu+\rho-\mu_{\beta}^{N^{\lambda}}\right)-\int_{\mathbf{R}^{d}}\log\left(\mu_{\beta}^{N^{\lambda}}\right)d\rho+\mbox{ent}[\rho]\Bigg{)}, (281)

where the minimum is taken over ρ\rho such that

𝐑dμ+ρμβNλ=0.\int_{\mathbf{R}^{d}}\mu+\rho-\mu_{\beta}^{N^{\lambda}}=0. (282)

Then we can use ρ\rho^{*} as a test function in the definition of 𝐓λN(μ)\mathbf{T}^{N}_{\lambda}(\mu^{*}) and get

𝐓λN(μ)12(μ+ρμβNλ)𝐑dlog(μβNλ)𝑑ρ+ent[ρ].\mathbf{T}^{N}_{\lambda}(\mu^{*})\leq\frac{1}{2}\mathcal{E}\left(\mu^{*}+\rho^{*}-\mu_{\beta}^{N^{\lambda}}\right)-\int_{\mathbf{R}^{d}}\log\left(\mu_{\beta}^{N^{\lambda}}\right)d\rho^{*}+\mbox{ent}[\rho^{*}]. (283)

Using Lemmas 9.1, 9.2, 9.3, 9.4, we have that

(μ+ρμβNλ)R(μ+ρμβNλ)+CN2(λ1d),\mathcal{E}\left(\mu^{*}+\rho^{*}-\mu_{\beta}^{N^{\lambda}}\right)\leq\mathcal{E}^{\neq}_{\square_{R}}\left(\mu+\rho^{*}-\mu_{\beta}^{N^{\lambda}}\right)+CN^{2\left(\lambda-\frac{1}{d}\right)}, (284)

where CC depends on ν\nu. We, therefore, get that

𝐓λN(μ)𝐓λN,(μ)+CN2(λ1d),\mathbf{T}^{N}_{\lambda}(\mu^{*})\leq\mathbf{T}^{N,\neq}_{\lambda}(\mu)+CN^{2\left(\lambda-\frac{1}{d}\right)}, (285)

where CC depends on ν\nu. Note that

limNμμBL=0,\lim_{N\to\infty}\|\mu-\mu^{*}\|_{BL}=0, (286)

therefore we are left with proving that

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)(𝐓λN(μ)))=0.\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)}\left(\mathbf{T}^{N}_{\lambda}(\mu)\right)\right)=0. (287)

To see this, let

μNϵ=argminμB(ν,ϵ)𝐓λN(μ).\mu_{N}^{\epsilon}=\operatorname{argmin}_{\mu\in{B}(\nu,\epsilon)}\mathbf{T}^{N}_{\lambda}(\mu). (288)

We assume that the infimum is achieved for clarity of exposition. Otherwise, we would repeat the argument up to an arbitrarily small error. Let

ρϵN=argminρ+(𝐑dR)(12(μNϵ+ρμβNλ)𝐑dlog(μβNλ)𝑑ρ+ent[ρ]),\rho^{N}_{\epsilon}=\\ \operatorname{argmin}_{\rho\in\mathcal{M}^{+}(\mathbf{R}^{d}\setminus\square_{R})}\Bigg{(}\frac{1}{2}\mathcal{E}\left(\mu_{N}^{\epsilon}+\rho-\mu_{\beta}^{N^{\lambda}}\right)-\int_{\mathbf{R}^{d}}\log\left(\mu_{\beta}^{N^{\lambda}}\right)d\rho+\mbox{ent}[\rho]\Bigg{)}, (289)

where the minimum is taken over ρ\rho such that

𝐑dμNϵ+ρμβNλ=0.\int_{\mathbf{R}^{d}}\mu_{N}^{\epsilon}+\rho-\mu_{\beta}^{N^{\lambda}}=0. (290)

Then we can use ρϵN\rho^{N}_{\epsilon} as a test function in the definition of 𝐓λN(ν)\mathbf{T}^{N}_{\lambda}(\nu) and get

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)(𝐓λN(μ)))limϵ0limN12(ν+ρϵNμβNλ)12(μNϵ+ρϵNμβNλ)=limϵ0limN12(ν)12(μNϵ)+𝒢(νμNϵ,ρϵNμβNλ).\begin{split}&\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)}\left(\mathbf{T}^{N}_{\lambda}(\mu)\right)\right)\leq\\ &\lim_{\epsilon\to 0}\lim_{N\to\infty}\frac{1}{2}\mathcal{E}\left(\nu+\rho^{N}_{\epsilon}-\mu_{\beta}^{N^{\lambda}}\right)-\frac{1}{2}\mathcal{E}\left(\mu_{N}^{\epsilon}+\rho^{N}_{\epsilon}-\mu_{\beta}^{N^{\lambda}}\right)=\\ &\lim_{\epsilon\to 0}\lim_{N\to\infty}\frac{1}{2}\mathcal{E}\left(\nu\right)-\frac{1}{2}\mathcal{E}\left(\mu_{N}^{\epsilon}\right)+\mathcal{G}\left(\nu-\mu_{N}^{\epsilon},\rho^{N}_{\epsilon}-\mu_{\beta}^{N^{\lambda}}\right).\end{split} (291)

Note that as ϵ\epsilon tends to 0 and NN tends to \infty, μNϵ\mu_{N}^{\epsilon} converges weakly to ν\nu, therefore

limϵ0limN(ν)(μNϵ)0,\lim_{\epsilon\to 0}\lim_{N\to\infty}\mathcal{E}\left(\nu\right)-\mathcal{E}\left(\mu_{N}^{\epsilon}\right)\leq 0, (292)

and

limϵ0limN𝒢(νμNϵ,ρϵNμβNλ)=0.\lim_{\epsilon\to 0}\lim_{N\to\infty}\mathcal{G}\left(\nu-\mu_{N}^{\epsilon},\rho^{N}_{\epsilon}-\mu_{\beta}^{N^{\lambda}}\right)=0. (293)

This implies that

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)(𝐓λN(μ)))0,\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)}\left(\mathbf{T}^{N}_{\lambda}(\mu)\right)\right)\leq 0, (294)

and since clearly

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)(𝐓λN(μ)))0,\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)}\left(\mathbf{T}^{N}_{\lambda}(\mu)\right)\right)\geq 0, (295)

we conclude that

limϵ0limN(𝐓λN(ν)infμB(ν,ϵ)𝒜iNNλd1(R)(𝐓λN,(μ)))=0.\lim_{\epsilon\to 0}\lim_{N\to\infty}\left(\mathbf{T}^{N}_{\lambda}(\nu)-\inf_{\mu\in{B}(\nu,\epsilon)\cap\mathcal{A}_{i_{N}}^{N^{\lambda d-1}}(\square_{R})}\left(\mathbf{T}^{N,\neq}_{\lambda}(\mu)\right)\right)=0. (296)

9 Appendix A

In this appendix, we prove some fundamental properties of the smearing technique and energy minimizers. Loosely speaking, the smearing technique consists in studying properties about empN{\rm emp}_{N} by analyzing instead the more regular measure empNλϵ,{\rm emp}_{N}\ast\lambda_{\epsilon}, where λϵ\lambda_{\epsilon} is a measure that approximates a Dirac delta on a scale ϵ\epsilon.

We start by recalling a few facts about smearing and electric energy. These are standard and can be found, for example, in [16], [22], or [27]. The proof uses that gg is superharmonic in its domain, and harmonic away from 0.0.

Lemma 9.1.

For every x𝐑dx\in\mathbf{R}^{d} and ϵ>0\epsilon>0 we have that

𝐑dg(x+u)𝑑λϵ(u)g(x)\int_{\mathbf{R}^{d}}g(x+u)\,d\lambda_{\epsilon}(u)\leq g(x) (297)

and also that

𝐑d×𝐑dg(x+uv)𝑑λϵ(u)𝑑λϵ(v)g(x),\iint_{\mathbf{R}^{d}\times\mathbf{R}^{d}}g(x+u-v)\,d\lambda_{\epsilon}(u)\,d\lambda_{\epsilon}(v)\leq g(x), (298)

(see Remark 3.2 for notation). Furthermore, eqs (297) and (298) become an equality if |x|>ϵ.|x|>\epsilon.

The next lemma can also be found in [16] (or verified by direct computation).

Lemma 9.2.

Let ϵ>0\epsilon>0, then for d3d\geq 3,

(λϵ)=g(ϵ)(λ1).\mathcal{E}(\lambda_{\epsilon})=g(\epsilon)\mathcal{E}(\lambda_{1}). (299)

For d=2d=2,

(λϵ)=g(ϵ)+(λ1).\mathcal{E}(\lambda_{\epsilon})=g(\epsilon)+\mathcal{E}(\lambda_{1}). (300)
Lemma 9.3.

Let {xi}i=1N𝐑d,\left\{x_{i}\right\}_{i=1}^{N}\in\mathbf{R}^{d}, let ϕ=1Ni=1Nδxi\phi=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}} and ϕϵ=ϕλϵ.\phi_{\epsilon}=\phi\ast\lambda_{\epsilon}. Then

1N2ijg(xixj)(ϕϵ)1Ng(ϵ)(λ1).\frac{1}{N^{2}}\sum_{i\neq j}g(x_{i}-x_{j})\geq\mathcal{E}\left(\phi_{\epsilon}\right)-\frac{1}{N}g(\epsilon)\mathcal{E}(\lambda_{1}). (301)

Furthermore, eq. (301) is an equality if ϵmin{|xixj|}.\epsilon\leq\min\left\{|x_{i}-x_{j}|\right\}.

Lemma 9.4.

Let ϕ=1Ni=1Nδxi\phi=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}} for {xi}i=1N𝐑d.\left\{x_{i}\right\}_{i=1}^{N}\in\mathbf{R}^{d}. Let ϕϵ=ϕλϵ\phi_{\epsilon}=\phi\ast\lambda_{\epsilon} for ϵ>0.\epsilon>0. Let μ\mu be a measure with an LL^{\infty} density. Then there exists C>0,C>0, which depends only on μL\|\mu\|_{L^{\infty}} such that

|𝒢(Pϵ,μ)𝒢(P,μ)|Cϵ2.\left|\mathcal{G}(P_{\epsilon},\mu)-\mathcal{G}(P,\mu)\right|\leq C\epsilon^{2}. (302)

10 Appendix B

We will now prove Lemma 7.1, which we restate here for convenience.

Lemma 10.1.

Let μn,ν¯\mu_{n},\overline{\nu} be probability measures on a compact set Ω\Omega such that

lim supnent[μn]<,ent[ν¯]<,\limsup_{n\to\infty}{\rm ent}[\mu_{n}]<\infty,\quad{\rm ent}[\overline{\nu}]<\infty, (303)

ν¯L(Ω)\overline{\nu}\in L^{\infty}(\Omega), and

(ν¯)<.\mathcal{E}(\overline{\nu})<\infty. (304)

Assume that μn(x)\mu_{n}(x) is uniformly equi-continuous and bounded away from 0 uniformly in xx and nn. Then for every ϵ,δ,η,\epsilon,\delta,\eta, there exists a family of configurations

Λδη,ϵ𝐑d×n\Lambda_{\delta}^{\eta,\epsilon}\subset\mathbf{R}^{d\times n} (305)

such that

  • \bullet
    empn(Xn)B(ν¯,ϵ){\rm emp}_{n}(X_{n})\in B(\overline{\nu},\epsilon) (306)

    for any XnΛδη,ϵ.X_{n}\in\Lambda_{\delta}^{\eta,\epsilon}.

  • \bullet
    lim infn1nlog(XnΛδη,ϵΠi=1nμn(xi)𝑑Xn)lim infnent[ν¯|μn]δ.\liminf_{n\to\infty}\frac{1}{n}\log\left(\int_{X_{n}\in\Lambda_{\delta}^{\eta,\epsilon}}\Pi_{i=1}^{n}\mu_{n}(x_{i})dX_{n}\right)\geq-\liminf_{n\to\infty}{\rm ent}[\overline{\nu}|\mu_{n}]-\delta. (307)
  • \bullet
    lim supn|(empn(Xn)ν¯)|η2.\limsup_{n\to\infty}\left|\mathcal{E}^{\neq}({\rm emp}_{n}(X_{n})-\overline{\nu})\right|\leq\eta^{2}. (308)
  • There exists r>0r>0 such that

    d(xi,Ω)>rn1dandd(xi,xj)>rn1d,d(x_{i},\partial\Omega)>rn^{-\frac{1}{d}}\quad{\rm and}\quad d(x_{i},x_{j})>rn^{-\frac{1}{d}}, (309)

    for iji\neq j.

Proof.

Step 1: Definition

First, we subdivide Ω\Omega into cubes KjK_{j} of size η¯>0\overline{\eta}>0 and center xj,x_{j}, for η¯>0\overline{\eta}>0 to be determined later.

Let either

nj=nν¯(Kj)n_{j}=\left\lceil n\overline{\nu}(K_{j})\right\rceil (310)

or

nj=nν¯(Kj),n_{j}=\left\lfloor n\overline{\nu}(K_{j})\right\rfloor, (311)

chosen so that

jnj=n.\sum_{j}n_{j}=n. (312)

The procedure for determining the point configuration of njn_{j} points is: y1y_{1} is chosen at random from Kjτ,K_{j}^{\tau}, where KjτK_{j}^{\tau} is the cube KjK_{j} minus a boundary layer of width τ\tau, y2y_{2} is chosen at random from

KjτB(y1,τ).K_{j}^{\tau}\setminus B(y_{1},\tau). (313)

Then, for i=1nj,i=1...n_{j}, the point yiy_{i} is chosen at random from

Kjτl=1i1B(yl,τ).K_{j}^{\tau}\setminus\bigcup_{l=1}^{i-1}B(y_{l},\tau). (314)

In other words,

Λδη,ϵ=σsym[1:n]ji=1nj(Kjτl=1i1B(yσ(l),τ)).\Lambda_{\delta}^{\eta,\epsilon}=\bigcup_{\sigma\in\text{sym}[1:n]}\bigotimes_{j}\bigotimes_{i=1}^{n_{j}}\left(K_{j}^{\tau}\setminus\bigcup_{l=1}^{i-1}B(y_{\sigma(l)},\tau)\right). (315)

We set τ=αη¯nj1d,\tau=\alpha\overline{\eta}n_{j}^{-\frac{1}{d}}, for some α(0,1)\alpha\in(0,1) to be determined later. For α\alpha small enough, the procedure is well defined, in the sense that it is possible to choose njn_{j} points in this way.

For η¯\overline{\eta} small enough, any XnΛδη,ϵX_{n}\in\Lambda_{\delta}^{\eta,\epsilon} satisfies

empn(Xn)B(ν¯,ϵ).{\rm emp}_{n}(X_{n})\in B(\overline{\nu},\epsilon). (316)

We immediately get that d(xi,Ω)>rn1d,d(xi,xj)>rn1dd(x_{i},\partial\Omega)>rn^{-\frac{1}{d}},d(x_{i},x_{j})>rn^{-\frac{1}{d}} for some r>0r>0. We now prove that these configurations have the right volume and energy.

Step 2: Volume Estimate

To give intuition, we first treat the case μn\mu_{n} as the uniform measure on Ω\Omega. In this case, we have

μnn(Λδη,ϵ)=n!Πini!ΠjΠp=1nj1(η¯dkdη¯d1τcdpτd)=n!Πini!Πjη¯dnjΠp=1nj1(1τkdη¯cdpαdnj),\begin{split}\mu_{n}^{\otimes n}(\Lambda_{\delta}^{\eta,\epsilon})&=\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}\Pi_{p=1}^{n_{j}-1}(\overline{\eta}^{d}-k_{d}\overline{\eta}^{d-1}\tau-c_{d}p\tau^{d})\\ &=\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}\overline{\eta}^{dn_{j}}\Pi_{p=1}^{n_{j}-1}(1-\frac{\tau k_{d}}{\overline{\eta}}-\frac{c_{d}p\alpha^{d}}{n_{j}}),\end{split} (317)

where cd,kdc_{d},k_{d} are constants which depend only on dd. On the other hand, the volume of all configurations with exactly njn_{j} points in cube KjK_{j} is given by

n!Πini!Πjη¯dnj.\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}\overline{\eta}^{dn_{j}}. (318)

By Sanov’s theorem, we have that

n!Πini!Πjη¯dnj=exp(n[ent[ν¯|μn]+on(1)]).\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}\overline{\eta}^{dn_{j}}=\exp(-n[{\rm ent}[\overline{\nu}|\mu_{n}]+o_{n}(1)]). (319)

For a general μn,\mu_{n}, we have that the volume of all configurations with exactly njn_{j} points in cube KjK_{j} is given by

n!Πini!Πj[μn(Kj)]nj,\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}[\mu_{n}(K_{j})]^{n_{j}}, (320)

and that by Sanov’s theorem

n!Πini!Πj[μn(Kj)]nj=exp(n[ent[ν¯|μn]+on(1)]).\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}[\mu_{n}(K_{j})]^{n_{j}}=\exp(-n[{\rm ent}[\overline{\nu}|\mu_{n}]+o_{n}(1)]). (321)

On the other hand, we can estimate

log(ΠjΠp=1nj1(1kdτη¯cdpαdnj))=jp=1nj1log(1kdτη¯cdpαdnj)αkdjnj11d+cdαdjnjCαn,\begin{split}\log\left(\Pi_{j}\Pi_{p=1}^{n_{j}-1}(1-\frac{k_{d}\tau}{\overline{\eta}}-\frac{c_{d}p\alpha^{d}}{n_{j}})\right)&=\sum_{j}\sum_{p=1}^{n_{j}-1}\log\left(1-\frac{k_{d}\tau}{\overline{\eta}}-\frac{c_{d}p\alpha^{d}}{n_{j}}\right)\\ &\leq\alpha k_{d}\sum_{j}n_{j}^{1-\frac{1}{d}}+c_{d}\alpha^{d}\sum_{j}n_{j}\\ &\leq C\alpha n,\end{split} (322)

where CC depends on ν¯\overline{\nu}. Using the hypothesis that μn\mu_{n} is uniformly equi-continuous, we have that for any any δ¯>0\overline{\delta}>0 there exists η¯\overline{\eta}^{*} such that if η¯<η¯\overline{\eta}<\overline{\eta}^{*} we have

μn(x)μn(y)(1δ¯,1+δ¯)\frac{\mu_{n}(x)}{\mu_{n}(y)}\in(1-\overline{\delta},1+\overline{\delta}) (323)

for any x,yΩx,y\in\Omega.

Hence, we have

log(μnn(Λδη,ϵ))log(n!Πini!Πj[μn(Kj)]nj)log(ΠjΠp=1nj1(1τη¯cdpαdnj))on(n)=n(ent[ν¯|μn]Cαon(1)oη¯(1)).\begin{split}&\log\left(\mu_{n}^{\otimes n}(\Lambda_{\delta}^{\eta,\epsilon})\right)\geq\\ &\log\left(\frac{n!}{\Pi_{i}n_{i}!}\Pi_{j}[\mu_{n}(K_{j})]^{n_{j}}\right)-\log\left(\Pi_{j}\Pi_{p=1}^{n_{j}-1}(1-\frac{\tau}{\overline{\eta}}-\frac{c_{d}p\alpha^{d}}{n_{j}})\right)-o_{n}(n)=\\ &-n\left({\rm ent}[\overline{\nu}|\mu_{n}]-C\alpha-o_{n}(1)-o_{\overline{\eta}}(1)\right).\end{split} (324)

Step 3: Energy Estimate

The idea for the energy estimate will be to prove that

hempnν¯h^{{\rm emp}_{n}-\overline{\nu}} (325)

is typically pointwise small. Then the smallness of the energy will be a consequence of the finite mass of the measures ν¯\overline{\nu} and empn.{\rm emp}_{n}.

Let xKi.x\in K_{i}. Then we can write

hempnν¯(x)=Kig(xy)d(empnν¯)(y)+jiKjg(xy)d(empnν¯)(y).h^{{\rm emp}_{n}-\overline{\nu}}(x)=\int_{K_{i}}g(x-y)d({\rm emp}_{n}-\overline{\nu})(y)+\sum_{j\neq i}\int_{K_{j}}g(x-y)d({\rm emp}_{n}-\overline{\nu})(y). (326)

For any ji,j\neq i, note that the minimum distance from xx to KjK_{j} is given by |xxi|cη¯|x-x_{i}|-c\overline{\eta} and the maximum distance from xx to KiK_{i} is given by |xxi|+cη¯|x-x_{i}|+c\overline{\eta}, for some cc which depends on dd and xx. For the rest of the proof, we assume w.l.o.g. that

empn(Kj)ν¯(Kj),{\rm emp}_{n}(K_{j})\geq\overline{\nu}(K_{j}), (327)

then

|Kjg(xy)d(empnν¯)(y)||empn(Ki)(|xxi|cη¯)d2ν¯(Ki)(|xxi|+cη¯)d2|=|empn(Ki)(|xxi|cη¯)d2ν¯(Ki)(|xxi|cη¯)d2+ν¯(Ki)(|xxi|cη¯)d2ν¯(Ki)(|xxi|+cη¯)d2||(empnν¯)(Ki)(|xxi|cη¯)d2|+|Cη¯ν¯(Ki)|xxi|d1|,\begin{split}&\left|\int_{K_{j}}g(x-y)d({\rm emp}_{n}-\overline{\nu})(y)\right|\leq\\ &\left|\frac{{\rm emp}_{n}(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}-\frac{\overline{\nu}(K_{i})}{(|x-x_{i}|+c\overline{\eta})^{d-2}}\right|=\\ &\left|\frac{{\rm emp}_{n}(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}-\frac{\overline{\nu}(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}+\frac{\overline{\nu}(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}-\frac{\overline{\nu}(K_{i})}{(|x-x_{i}|+c\overline{\eta})^{d-2}}\right|\leq\\ &\left|\frac{({\rm emp}_{n}-\overline{\nu})(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}\right|+\left|C\overline{\eta}\frac{\overline{\nu}(K_{i})}{|x-x_{i}|^{d-1}}\right|,\end{split} (328)

for some absolute constant CC.

Using the hypothesis that ν¯L(Ω)\overline{\nu}\in L^{\infty}(\Omega) we get

|(empnν¯)(Kj)|Cn,|({\rm emp}_{n}-\overline{\nu})(K_{j})|\leq\frac{C}{n}, (329)

where CC depends on ν¯L\|\overline{\nu}\|_{L^{\infty}}. Since 1|x|d2\frac{1}{|x|^{d-2}} is integrable at the origin and Ω\Omega is compact, we have

ji|(empnν¯)(Ki)(|xxi|cη¯)d2|Cnη¯d,\sum_{j\neq i}\left|\frac{({\rm emp}_{n}-\overline{\nu})(K_{i})}{(|x-x_{i}|-c\overline{\eta})^{d-2}}\right|\leq\frac{C}{n\overline{\eta}^{d}}, (330)

where CC depends on ν¯L\|\overline{\nu}\|_{L^{\infty}} and Ω\Omega.

Using again the hypothesis that ν¯L\overline{\nu}\in L^{\infty} we have

ji|η¯ν¯(Ki)|xxi|d1|Cη¯Ω1|x|d1d,Cη¯,\begin{split}\sum_{j\neq i}\left|\overline{\eta}\frac{\overline{\nu}(K_{i})}{|x-x_{i}|^{d-1}}\right|&\leq C\overline{\eta}\int_{\Omega}\frac{1}{|x|^{d-1}}d,\\ &\leq C\overline{\eta},\end{split} (331)

where CC depends on ν¯L\|\overline{\nu}\|_{L^{\infty}} and Ω\Omega.

For the second term in equation (326) term, we will instead work with

empn=empnλτ2,{\rm emp}_{n}^{*}={\rm emp}_{n}\ast\lambda_{\frac{\tau}{2}}, (332)

where τ=αη¯nj1d,\tau=\alpha\overline{\eta}n_{j}^{-\frac{1}{d}}, for some α(0,1)\alpha\in(0,1) to be determined later (see Remark 3.2 for notation). Note that by Lemma 9.1, and because d(xi,xj)rn1dd(x_{i},x_{j})\geq rn^{-\frac{1}{d}} we have

jiKjg(xy)d(empnν¯)(y)=jiKjg(xy)d(empnν¯)(y).\sum_{j\neq i}\int_{K_{j}}g(x-y)d({\rm emp}_{n}-\overline{\nu})(y)=\sum_{j\neq i}\int_{K_{j}}g(x-y)d({\rm emp}_{n}^{*}-\overline{\nu})(y). (333)

Note also that

empnLcα,ν¯,\|{\rm emp}_{n}^{*}\|_{L^{\infty}}\leq c_{\alpha,\overline{\nu}}, (334)

where cα,ν¯c_{\alpha,\overline{\nu}} is a constant that depends on α\alpha and ν¯L\|\overline{\nu}\|_{L^{\infty}}. Hence

|Kig(xy)d(empnν¯)(y)|cα,ν¯Ki1|x|d2𝑑xcα,ν¯η¯2,\begin{split}\left|\int_{K_{i}}g(x-y)d({\rm emp}_{n}^{*}-\overline{\nu})(y)\right|&\leq c_{\alpha,\overline{\nu}}\int_{K_{i}}\frac{1}{|x|^{d-2}}dx\\ &\leq c_{\alpha,\overline{\nu}}\overline{\eta}^{2},\end{split} (335)

where cα,ν¯c_{\alpha,\overline{\nu}} is a (new) constant that depends on α\alpha and ν¯L\|\overline{\nu}\|_{L^{\infty}}.

Putting everything together, we get

|hempnν¯|Cnη¯d+Cη¯+cα,ν¯η¯2,|h^{{\rm emp}^{*}_{n}-\overline{\nu}}|\leq\frac{C}{n\overline{\eta}^{d}}+C\overline{\eta}+c_{\alpha,\overline{\nu}}\overline{\eta}^{2}, (336)

where CC depends on ν¯\overline{\nu} and Ω\Omega and cα,ν¯c_{\alpha,\overline{\nu}} depends, in addition, on α\alpha.

Hence

Ω×Ωg(xy)d(empnν¯)(x)d(empnν¯)(y)hempnν¯Lempnν¯TVCnη¯d+Cη¯+cα,ν¯η¯2.\begin{split}&\iint_{\Omega\times\Omega}g(x-y)d({\rm emp}_{n}^{*}-\overline{\nu})(x)d({\rm emp}_{n}^{*}-\overline{\nu})(y)\leq\\ &\|h^{{\rm emp}^{*}_{n}-\overline{\nu}}\|_{L^{\infty}}\|{\rm emp}_{n}^{*}-\overline{\nu}\|_{TV}\leq\\ &\frac{C}{n\overline{\eta}^{d}}+C\overline{\eta}+c_{\alpha,\overline{\nu}}\overline{\eta}^{2}.\end{split} (337)

Making η¯\overline{\eta} small enough after having chosen α,\alpha, while keeping η¯>>n1d,\overline{\eta}>>n^{-\frac{1}{d}}, we have that for any η>0\eta>0 we can find parameters such that

lim supn|(empn(Xn)ν¯)|η2,\limsup_{n\to\infty}\left|\mathcal{E}({\rm emp}_{n}^{*}(X_{n})-\overline{\nu})\right|\leq\eta^{2}, (338)

which implies that

lim supn|(empn(Xn)ν¯)|η2.\limsup_{n\to\infty}\left|\mathcal{E}^{\neq}({\rm emp}_{n}(X_{n})-\overline{\nu})\right|\leq\eta^{2}. (339)

11 Acknowledgements

I thank Sylvia Serfaty for her guidance during this project. I thank Ofer Zeitouni and Thomas Leblé for useful conversations.

References

  • [1] S. Armstrong and S. Serfaty, Local laws and rigidity for Coulomb gases at any temperature, Annals of Probability, in press-arXiv preprint arXiv:1906.09848, (2019).
  • [2]  , Thermal approximation of the equilibrium measure and obstacle problem, arXiv preprint arXiv:1912.13018, (2019).
  • [3] G. B. Arous and A. Guionnet, Large deviations for wigner’s law and voiculescu’s non-commutative entropy, Probability theory and related fields, 108 (1997), pp. 517–542.
  • [4] R. Bauerschmidt, P. Bourgade, M. Nikula, and H.-T. Yau, Local density for two-dimensional one-component plasma, Communications in Mathematical Physics, 356 (2017), pp. 189–230.
  • [5]  , The two-dimensional coulomb plasma: quasi-free approximation and central limit theorem, Advances in Theoretical and Mathematical Physics, 23 (2019), pp. 841–1002.
  • [6] F. Bekerman, T. Leblé, S. Serfaty, et al., Clt for fluctuations of β\beta-ensembles with general potential, Electronic Journal of Probability, 23 (2018).
  • [7] G. Ben Arous and O. Zeitouni, Large deviations from the circular law, ESAIM: Probability and Statistics, 2 (1998), pp. 123–134.
  • [8] T. Bodineau and A. Guionnet, About the stationary states of vortex systems, Annales de l’Institut Henri Poincare (B) Probability and Statistics, 35 (1999), pp. 205–237.
  • [9] A. Borodin, I. Corwin, and A. Guionnet, Random matrices, in IAS/Park City Mathematics Series, vol. 26, American Mathematical Soc., 2019.
  • [10] G. Borot and A. Guionnet, Asymptotic expansion of β\beta matrix models in the one-cut regime, Communications in Mathematical Physics, 317 (2013), pp. 447–483.
  • [11] P. Bourgade, L. Erdős, and H.-T. Yau, Bulk universality of general β\beta-ensembles with non-convex potential, Journal of mathematical physics, 53 (2012), p. 095221.
  • [12] P. Bourgade, L. Erdös, and H.-T. Yau, Edge universality of beta ensembles, Communications in Mathematical Physics, 332 (2014), pp. 261–353.
  • [13] P. Bourgade, L. Erdős, H.-T. Yau, et al., Universality of general β\beta-ensembles, Duke Mathematical Journal, 163 (2014), pp. 1127–1190.
  • [14] P. Bourgade, H.-T. Yau, and J. Yin, Local circular law for random matrices, Probability Theory and Related Fields, 159 (2014), pp. 545–595.
  • [15] D. Chafaï, N. Gozlan, P.-A. Zitt, et al., First-order global asymptotics for confined particles with singular pair repulsion, The Annals of Applied Probability, 24 (2014), pp. 2371–2413.
  • [16] D. Chafai, A. Hardy, and M. Maïda, Concentration for coulomb gases and coulomb transport inequalities, Journal of Functional Analysis, 275 (2018), pp. 1447–1483.
  • [17] D. García-Zelada, A large deviation principle for empirical measures on polish spaces: Application to singular gibbs measures on manifolds, in Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, vol. 55, Institut Henri Poincaré, 2019, pp. 1377–1401.
  • [18] D. P. Hardin, T. Leblé, E. B. Saff, and S. Serfaty, Large deviation principles for hypersingular riesz gases, Constructive Approximation, 48 (2018), pp. 61–100.
  • [19] A. Hardy and G. Lambert, Clt for circular beta-ensembles at high temperature, Journal of Functional Analysis, 280 (2021), p. 108869.
  • [20] K. Johansson et al., On fluctuations of eigenvalues of random hermitian matrices, Duke mathematical journal, 91 (1998), pp. 151–204.
  • [21] G. Lambert, M. Ledoux, C. Webb, et al., Quantitative normal approximation of linear statistics of β\beta-ensembles, The Annals of Probability, 47 (2019), pp. 2619–2685.
  • [22] T. Leblé and S. Serfaty, Large deviation principle for empirical fields of log and riesz gases, Inventiones mathematicae, 210 (2017), pp. 645–757.
  • [23]  , Fluctuations of two dimensional coulomb gases, Geometric and Functional Analysis, 28 (2018), pp. 443–508.
  • [24] T. Leblé, S. Serfaty, and O. Zeitouni, Large deviations for the two-dimensional two-component plasma, Communications in Mathematical Physics, 350 (2017), pp. 301–360.
  • [25] D. Padilla-Garza, Concentration inequality around the thermal equilibrium measure of coulomb gases, Journal of Functional Analysis, 284 (2023), p. 109733.
  • [26] D. Petz and F. Hiai, Logarithmic energy as an entropy functional, Contemporary Mathematics, 217 (1998), pp. 205–221.
  • [27] N. Rougerie and S. Serfaty, Higher-dimensional coulomb gases and renormalized energy functionals, Communications on Pure and Applied Mathematics, 69 (2016), pp. 519–605.
  • [28] E. Sandier, S. Serfaty, et al., 2d coulomb gases and the renormalized energy, The Annals of Probability, 43 (2015), pp. 2026–2083.
  • [29] S. Serfaty, Microscopic description of log and coulomb gases, arXiv preprint arXiv:1709.04089, (2017).
  • [30]  , Systems of points with coulomb interactions, in Proceedings of the International Congress of Mathematicians: Rio de Janeiro 2018, World Scientific, 2018, pp. 935–977.
  • [31]  , Gaussian fluctuations and free energy expansion for 2d and 3d coulomb gases at any temperature, arXiv preprint arXiv:2003.11704, (2020).
  • [32] M. Shcherbina, Fluctuations of linear eigenvalue statistics of β\beta matrix models in the multi-cut regime, Journal of Statistical Physics, 151 (2013), pp. 1004–1034.