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

Optimal local laws and CLT for the circular
Riesz gas

Jeanne Boursier (JB) École Normale Supérieure de Lyon, UMR 5669, CNRS, UMPA, 69364 Lyon, France
Abstract.

We study the long-range one-dimensional Riesz gas on the circle, a continuous system of particles interacting through a Riesz kernel. We establish near-optimal rigidity estimates on gaps valid at any scale. Leveraging on these local laws together with Stein’s method, we prove a quantitative Central Limit Theorem for linear statistics. The proof is based on a mean-field transport and a fine analysis of the fluctuations of local error terms using various convexity and monotonicity arguments. By using a comparison principle for the Helffer-Sjöstrand equation, the method can handle very singular test-functions, including characteristic functions of intervals.

1. Introduction

1.1. Setting of the problem

In this paper, we study the one-dimensional Riesz gas on the circle. We denote 𝕋:=/\mathbb{T}:=\mathbb{R}/\mathbb{Z}. For a parameter s(0,1)s\in(0,1), let us consider the Riesz ss-kernel on 𝕋\mathbb{T}, defined by

g(x)=limn(k=nn1|k+x|s21sn1s)=ζ(s,x)+ζ(s,1x),g(x)=\lim_{n\to\infty}\Bigr{(}\sum_{k=-n}^{n}\frac{1}{|k+x|^{s}}-\frac{2}{1-s}n^{1-s}\Bigr{)}=\zeta(s,x)+\zeta(s,1-x), (1.1)

where ζ(s,x)\zeta(s,x) stands for the Hurwitz zeta function [Ber72]. Note that gg is the fundamental solution of the fractional Laplace equation on the circle

{(Δ)1s2g=cs(δ01)g=0,\begin{cases}(-\Delta)^{\frac{1-s}{2}}g=c_{s}(\delta_{0}-1)\\ \int g=0,\end{cases} (1.2)

with csc_{s} given by

cs=Γ(1s2)Γ(s2)π21s.c_{s}=\frac{\Gamma(\frac{1-s}{2})}{\Gamma(\frac{s}{2})}\frac{\sqrt{\pi}}{2^{1-s}}. (1.3)

We endow 𝕋\mathbb{T} with the natural order x<yx<y if x=x+kx=x^{\prime}+k, y=y+ky=y^{\prime}+k^{\prime} with k,kk,k^{\prime}\in\mathbb{Z}, x,y[0,1)x^{\prime},y^{\prime}\in[0,1) and x<yx^{\prime}<y^{\prime} and work on the set of ordered configurations

DN={XN=(x1,,xN)𝕋N:x2x1<<xNx1}.D_{N}=\{X_{N}=(x_{1},\ldots,x_{N})\in\mathbb{T}^{N}:x_{2}-x_{1}<\ldots<x_{N}-x_{1}\}.

On DND_{N} let us consider the energy

N:XNDNNsijg(xixj),\mathcal{H}_{N}:X_{N}\in D_{N}\mapsto N^{-s}\sum_{i\neq j}g(x_{i}-x_{j}), (1.4)

where gg is given by (1.2). The circular Riesz gas, at the inverse temperature β>0\beta>0, is defined by the Gibbs measure

dN,β=1ZN,βexp(βN(XN))𝟙DN(XN)dXN,\mathrm{d}\mathbb{P}_{N,\beta}=\frac{1}{Z_{N,\beta}}\exp(-\beta\mathcal{H}_{N}(X_{N}))\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N},

where ZN,βZ_{N,\beta} is the normalizing constant, called the partition function, given by

ZN,β=DNeβN(XN)dXN.Z_{N,\beta}=\int_{D_{N}}e^{-\beta\mathcal{H}_{N}(X_{N})}\mathrm{d}X_{N}.

Throughout the paper, ss is a fixed parameter in (0,1)(0,1).

The choice of the normalization in the definition of the energy (1.4) appears to be a natural choice, making β\beta the effective inverse temperature governing the microscopic scale behavior.

The model described above belongs to a family of interacting particle systems named Riesz gases. On d\mathbb{R}^{\mathrm{d}}, these are associated to a kernel of the form |x|s|x|^{-s} with s>0s>0. The Riesz family also contains in dimensions 11 and 22 the so-called log-gases with kernel log|x|-\log|x|. For d3\mathrm{d}\geq 3 and s=d2s=\mathrm{d-2}, |x|s|x|^{-s} is the fundamental solution of the Laplace equation on d\mathbb{R}^{d} and therefore corresponds to the Coulombian interaction. The parameter ss determines the singularity as well as the range of the interaction. When s>ds>\mathrm{d}, the interaction is short-range and the system, referred to as hypersingular Riesz gas, resembles a nearest-neighbour model. For s(0,d)s\in(0,\mathrm{d}) or s=0s=0 and d=1,2\mathrm{d}=1,2, Riesz gases are long-range particle systems, which have, as such, attracted much attention in both mathematical and physical contexts.

The 1D log-gas, also called the β\beta-ensemble has been extensively studied in the last few decades, partly for its connection to random matrix theory. Indeed, it corresponds, in the cases β{1,2,4}\beta\in\{1,2,4\}, to the distribution of the eigenvalues of N×NN\times N symmetric/Hermitian/symplectic random matrices with independent Gaussian entries (see the original paper of Dyson [Dys62]). The 1D log-gas also appears in many other contexts such as zeros of random polynomials, zeros of the Riemann function and is conjectured to be related to the eigenvalues of random Schrödinger operators [Lew22]. The 2D Coulomb gas is another fundamental model. Among many other examples, it is connected to non-unitary random matrices, Ginzburg-Landau vertices, Fekete points, complex geometry, the XY model and the KT transition [Ser18]. For other values of ss, let us mention that the case s=2s=2 in dimension 11 is an integrable system, called the classical Calogero-Sutherland model. The study of minimizers of Riesz interactions is also a dynamic topic [HS05, CSW21] and is the object of long-standing conjectures related to sphere packing problems [BL15]. From a statistical physics perspective, even in dimension 11, the Riesz gas is not fully elucidated since the classical theory of the 60-70s [Rue74] cannot be applied due the long-range nature of the interaction. The reader may refer to the nice review [Lew22], where an account of the literature and many open problems on Riesz gases are given.

As the number of particles NN tends to infinity, the empirical measure

μN:=1Ni=1Nδxi\mu_{N}:=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}

converges almost surely under N,β\mathbb{P}_{N,\beta} (in a suitable topology) to the uniform measure on the circle. This result can be obtained through standard large deviations techniques (see for instance [Ser14, Ch. 2] for the case of Riesz gases on the real line, which adapts readily to the periodic setting). In the large NN limit, particles tend to spread uniformly on the circle, which suggests that particles spacing (or gaps) N(xi+kxi)N(x_{i+k}-x_{i}) concentrate around the value kk. The first goal of this paper is to quantify the fluctuations of the gaps around their mean. We establish the optimal size of the fluctuations of N(xi+kxi)N(x_{i+k}-x_{i}), which turns out to be in O(β12ks2)O(\beta^{-\frac{1}{2}}k^{\frac{s}{2}}), as conjectured in the recent physics paper [SKA+21]. This type of result is referred to in the literature as a rigidity estimate. It was intensively investigated for β\beta-ensembles (see for instance [BEY12, BEY+14b, BEY14a, BMP22]), but the correct observable in that case is xiγix_{i}-\gamma_{i}, where xix_{i} is the ii-th particle and γi\gamma_{i} the classical location of the ii-th particle, that is the corresponding quantile of the equilibrium measure arising in the mean-field limit.

A complementary way to study the rigidity of the system is to investigate the fluctuations of linear statistics of the form

FluctN[ξ(N1)]:=i=1Nξ(N1xi)NN𝕋ξ,\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]:=\sum_{i=1}^{N}\xi(\ell_{N}^{-1}x_{i})-N\ell_{N}\int_{\mathbb{T}}\xi, (1.5)

where ξ:𝕋\xi:\mathbb{T}\to\mathbb{R} is a given measurable test-function and {N}\{\ell_{N}\} a sequence of numbers in (0,1](0,1]. For smooth test-functions, many central limit theorem (CLT) results are available in the literature on 1D-log gases, including [J+98, Shc13, BG13, BEY12, BEY+14b, BLS+18, BMP22, HL21]. For 1D Riesz gases with s(0,1)s\in(0,1), to our knowledge, no prior results on CLT for linear statistics are known. In this paper we obtain a quantitative CLT for (1.5), which is valid at all scales {N}\{\ell_{N}\} down to microscopic scales N1N\ell_{N}\gg\frac{1}{N}. A major direction in random matrix theory is to establish CLTs for (1.5) allowing test-functions which are as singular as possible. Indeed it is a natural question to capture the fluctuations of the number of points and of the logarithmic potential, which are key observables for the log-gas. The question of the optimal regularity of the test-function has also drawn a lot of interest because it encapsulates non-universality features in the context of Wigner matrices. One of the goals of this paper is to provide a robust method which allows singular test-functions to be treated in a systematic way. The main question we investigate is therefore a regularity issue. Using new concentration inequalities we are able to treat singular test-functions, including characteristic functions of intervals and inverse power functions up to the critical power s2\frac{s}{2}. In particular we obtain a CLT for the number of points, thus extending to a Riesz (periodic and fully convex) setting some of the recent results of [BMP22].

Let us now introduce the main tools and objects used in the proofs. For any reasonable Gibbs measure on DND_{N} (or N\mathbb{R}^{N}), the fluctuations of any (smooth) statistic F:DNF:D_{N}\to\mathbb{R} are related to the properties of a partial differential equation called the Helffer-Sjöstrand (H.-S.) equation, which is sometimes referred to as a Witten Laplacian (on 11-forms). This equation appears in [Sjo93a, Sjo93b, HS94]. It is more substantially studied in [Hel98b, Hel98a, NS97], where it is used to establish correlation decay, uniqueness of the limiting measure and log-Sobolev inequalities for models with convex interactions. The purpose of the present paper is to show how the analysis of Helffer-Sjöstrand equations provides powerful tools to study the fluctuations of linear statistics with singular test-functions.

The proof of the near-optimal rigidity is essentially similar to [BEY12]. It exploits the convexity of the interaction and is thus very specific to 1D systems. The method is mainly based on a concentration inequality for divergence-free functions and on a key convexity result due to Brascamp [BL02].

The method of proof of the CLT for linear statistics starts by performing the mean-field transport argument usually attributed to Johansson [J+98]. When studying the Laplace transform of linear statistics FluctN[ξ]\mathrm{Fluct}_{N}[\xi], this consists in applying a well-chosen change of variables on each point, depending only on its position, to transport the uniform measure on the circle to the perturbed equilibrium measure (perturbed by the effect of adding tFluctN[ξ]t\mathrm{Fluct}_{N}[\xi] to the energy). In this paper, one computes variances instead of Laplace transforms and the implementation of the transport of [J+98] takes the form of an integration by parts. This argument is a variation of the so-called loop equations (see [BEY12] for various comments on this topic). It is the starting point of many CLTs on β\beta-ensembles and Coulomb gases but it received a more systematic analysis in the series of works [LS18, BLS+18, Leb18, LZ20, Ser20]. One can also interpret this transport as a mean-field approximation of the H.-S. equation associated to (the gradient of) linear statistics.

Since the transport is an approximate solution (a mean-field approximation) of the H.-S. equation, it creates an error term, sometimes called loop equation term, which is essentially a local weighted energy and the heart of the problem is to estimate its fluctuations. In contrast, in [LS18, BLS+18, Leb18, Ser20] the typical size in a large deviation sense of this error term is evaluated, rather than the size of its fluctuations. Let us point out however that the latter seems intractable in dimension d2\mathrm{d}\geq 2 because of the lack of convexity. The core of this paper is about the control on the fluctuations of this loop equation term through the analysis of the related H.-S. equation. Our proof is based on the Brascamp-Lieb inequality, the near-optimal rigidity estimates on gaps and nearest-neighbour distances and a comparison principle for the Helffer-Sjöstrand equation.

The use of the comparison principle mentioned above (also known in [Car74, HS94]) is one of the main novelties of the paper. This is the key technical tool to be able to treat linear statistics with singular test-functions. Indeed, after performing loop equations techniques, we will study singular local quantities, for which standard concentration inequalities, such as the Brascamp-Lieb or log-Sobolev inequalities, do not give the right order of fluctuations.

The central limit theorem is then obtained from a rather straightforward application of Stein’s method. We show how the mean-field transport naturally leads to an approximate Gaussian integration by parts formula. As a result, quantifying normality boils down to controlling the variance of the loop equation term. The CLT then follows from the variance bound discussed in the above comments.

1.2. Main results

The following results are valid for s(0,1)s\in(0,1), thus covering the entire long-range regime.

Our first result concerns the fluctuations of gaps and discrepancies. We establish the following near-optimal decay estimate:

Theorem 1 (Near-optimal rigidity).

There exists ε0\varepsilon_{0} such that for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), letting δ:=ε4(s+2)\delta:=\frac{\varepsilon}{4(s+2)}, there exist c>0c>0 and C>0C>0 such that for each i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2},

N,β(|N(xi+kxi)k|ks2+ε)1Ceckδ.\mathbb{P}_{N,\beta}(|N(x_{i+k}-x_{i})-k|\leq k^{\frac{s}{2}+\varepsilon})\geq 1-Ce^{-ck^{\delta}}. (1.6)

In addition for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), letting δ:=ε4(s+2)\delta:=\frac{\varepsilon}{4(s+2)}, there exist c>0c>0 and C>0C>0 such that for all a𝕋a\in\mathbb{T} and N(0,14]\ell_{N}\in(0,\frac{1}{4}],

N,β(|i=1N𝟙(aN,a+N)(xi)2NN|(NN)s2+ε)1Cec(NN)δ.\mathbb{P}_{N,\beta}\Bigr{(}\Bigr{|}\sum_{i=1}^{N}\mathds{1}_{(a-\ell_{N},a+\ell_{N})}(x_{i})-2N\ell_{N}\Bigr{|}\leq(N\ell_{N})^{\frac{s}{2}+\varepsilon}\Bigr{)}\geq 1-Ce^{-c(N\ell_{N})^{\delta}}. (1.7)

Theorem 1 is the natural extension of the rigidity result of [BEY12, Th. 3.1] to the Riesz setting. Since the probability of deviations are exponentially small, the estimates of Theorem 1 allow one to reduce the phase space to an event where all gaps are close to their standard value. Theorem 1 is proved in Section 4.

We next prove that ks2k^{\frac{s}{2}} is the exact fluctuation scale of N(xi+kxi)N(x_{i+k}-x_{i}). This is equivalent to proving that i=1N𝟙(0,N)(xi)\sum_{i=1}^{N}\mathds{1}_{(0,\ell_{N})}(x_{i}) fluctuates at scale (NN)s2(N\ell_{N})^{\frac{s}{2}}. For any map ξ:𝕋\xi:\mathbb{T}\to\mathbb{R}, consider the linear statistic

FluctN[ξ]:=i=1Nξ(xi)N𝕋ξ.\mathrm{Fluct}_{N}[\xi]:=\sum_{i=1}^{N}\xi(x_{i})-N\int_{\mathbb{T}}\xi.

We prove sharp variance estimates valid at any scale for possibly singular test-functions satisfying the following assumptions:

Assumptions 1.1.

Let ξ:𝕋{,+}\xi:\mathbb{T}\to\mathbb{R}\cup\{-\infty,+\infty\}.

  1. (1)

    (Piecewise regularity) ξ\xi is piecewise 𝒞2s\mathcal{C}^{2-s}: there exists p0p\geq 0 and a1<<apa_{1}<\ldots<a_{p} such that, letting ap+1=a1a_{p+1}=a_{1}, for each i=1,,pi=1,\ldots,p, ξ\xi is 𝒞2s\mathcal{C}^{2-s} on (ai,ai+1)(a_{i},a_{i+1}).

  2. (2)

    (Global regularity) ξ𝒞s+ε(𝕋,)\xi\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}) for some ε>0\varepsilon>0.

  3. (3)

    (Singularity) Let ϕ:=(Δ)1s2ξ\phi:=(-\Delta)^{\frac{1-s}{2}}\xi. There exists α(s1,s2)\alpha\in(s-1,\frac{s}{2}) such that for all x𝕋{a1,,ap}x\in\mathbb{T}\setminus\{a_{1},\ldots,a_{p}\},

    |ϕ|(x)C(1+k=1p|xak|(1s+α)).|\phi^{\prime}|(x)\leq C\Bigr{(}1+\sum_{k=1}^{p}{|x-a_{k}|^{-(1-s+\alpha)}}\Bigr{)}. (1.8)
  4. (4)

    (Support) Let {N}\{\ell_{N}\} be a sequence in (0,1](0,1]. Assume that N0\ell_{N}\to 0 or that N=1\ell_{N}=1 for each NN. In the case where N0\ell_{N}\to 0, we will assume that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}).

Let us first comment upon the above assumptions.

Remark 1.2.
  • Assumption (3) compares the singularities of ξ\xi with the singularity of |x|s2+ε|x|^{-\frac{s}{2}+\varepsilon} at 0: the derivative of order 2s2-s of ξ\xi near a singular point a𝕋a\in\mathbb{T} is bounded by the derivative of order 2s2-s of x|xa|s2+εx\mapsto|x-a|^{-\frac{s}{2}+\varepsilon}. Note that the function x|x|s2x\mapsto|x|^{-\frac{s}{2}} is the critical inverse power function which does not lie in H1s2(𝕋,)H^{\frac{1-s}{2}}(\mathbb{T},\mathbb{R}).

  • When the scale N\ell_{N} tends to 0, Assumption (4) ensures that FluctN[ξ(N1)]\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)] is at most of order O(NN)O(N\ell_{N}).

  • The characteristic function ξ:=𝟙(a,a)\xi:=\mathds{1}_{(-a,a)} satisfies Assumptions 1.1. Indeed the map (Δ)s2ξ(-\Delta)^{-\frac{s}{2}}\xi is piecewise 𝒞2\mathcal{C}^{2} with singularities at a-a and aa. Moreover, letting ϕ:=(Δ)1s2ξ\phi:=(-\Delta)^{\frac{1-s}{2}}\xi, we have ϕ(x)xac0|x+a|2s\phi^{\prime}(x)\underset{x\to-a}{\sim}\frac{c_{0}}{|x+a|^{2-s}} and ϕ(x)xac0|xa|2s\phi^{\prime}(x)\underset{x\to a}{\sim}\frac{c_{0}}{|x-a|^{2-s}} for some c00c_{0}\neq 0.

Before stating the theorem, recall the definition of the fractional Sobolev seminorm on the circle ||Hα|\cdot|_{H^{\alpha}}, for α>0\alpha>0. Let h:𝕋h:\mathbb{T}\to\mathbb{R} in L2(𝕋,)L^{2}(\mathbb{T},\mathbb{R}) with Fourier coefficients h^(k)\hat{h}(k), kk\in\mathbb{Z}. Whenever it is finite, we call |h|Hα2|h|_{H^{\alpha}}^{2} the quantity

|h|Hα2:=k|k|2α|h^|2(k).|h|_{H^{\alpha}}^{2}:=\sum_{k\in\mathbb{Z}}|k|^{2\alpha}|\hat{h}|^{2}(k).

Similarly, the fractional Sobolev seminorm of a function hL2(,)h\in L^{2}(\mathbb{R},\mathbb{R}), also denoted |h|Hα|h|_{H^{\alpha}}, is defined by

|h|Hα2:=|ξ|2α|h^|2(ξ)dξ,|h|_{H^{\alpha}}^{2}:=\int|\xi|^{2\alpha}|\hat{h}|^{2}(\xi)\mathrm{d}\xi, (1.9)

where h^\hat{h} stands for the Fourier transform of hh.

Let {N}\{\ell_{N}\} be a sequence taking values in (0,1](0,1]. For any function ξH1s2(𝕋,)\xi\in H^{\frac{1-s}{2}}(\mathbb{T},\mathbb{R}), define

σN2(ξ):=Ns|ξ(N1)|H1s22.\sigma_{\ell_{N}}^{2}(\xi):=\ell_{N}^{-s}|\xi(\ell_{N}^{-1}\cdot)|_{H^{\frac{1-s}{2}}}^{2}. (1.10)

If ξH1s2(𝕋,)\xi\in H^{\frac{1-s}{2}}(\mathbb{T},\mathbb{R}) is supported on (12,12)(-\frac{1}{2},\frac{1}{2}), let ξ0:\xi_{0}:\mathbb{R}\to\mathbb{R} given for all xx\in\mathbb{R} by

ξ0(x)={ξ(x)if |x|120if |x|>12.\xi_{0}(x)=\begin{cases}\xi(x)&\text{if $|x|\leq\frac{1}{2}$}\\ 0&\text{if $|x|>\frac{1}{2}$}.\end{cases} (1.11)

For all ξH1s2(𝕋,)\xi\in H^{\frac{1-s}{2}}(\mathbb{T},\mathbb{R}), we then let

σ2(ξ):=12βcs{|ξ|H1s22if N=1 for each N|ξ0|H1s22if N0, with ξ0 as in (1.11).\sigma_{\infty}^{2}(\xi):=\frac{1}{2\beta c_{s}}\begin{cases}|\xi|_{H^{\frac{1-s}{2}}}^{2}&\text{if $\ell_{N}=1$ for each $N$}\\ |\xi_{0}|^{2}_{H^{\frac{1-s}{2}}}&\text{if $\ell_{N}\to 0$, with $\xi_{0}$ as in (\ref{eq:defxi0})}.\end{cases} (1.12)

The variance of FluctN[ξ(N1)]\mathrm{Fluct}_{N}[\xi(\cdot\ell_{N}^{-1})] under N,β\mathbb{P}_{N,\beta} may be expanded as follows:

Theorem 2 (Variance of singular linear statistics).

Let ξ\xi and {N}\{\ell_{N}\} satisfy Assumptions 1.1. Let α(s1,s2)\alpha\in(s-1,\frac{s}{2}) such that (1.8) is satisfied.

For all ε>0\varepsilon>0, there holds

VarN,β[FluctN[ξ(N1)]]=(NN)sσN2(ξ)+O((NN)max(2s1,2α)+ε)=(NN)s(σ2(ξ)+𝟙N0O(N2s)+O((NN)max(2s1,2α)+ε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=(N\ell_{N})^{s}\sigma_{\ell_{N}}^{2}(\xi)+O((N\ell_{N})^{\max(2s-1,2\alpha)+\varepsilon})\\ =(N\ell_{N})^{s}(\sigma_{\infty}^{2}(\xi)+\mathds{1}_{\ell_{N}\to 0}O(\ell_{N}^{2-s})+O((N\ell_{N})^{\max(2s-1,2\alpha)+\varepsilon}). (1.13)

Since α<s2\alpha<\frac{s}{2}, note that ξH1s2(𝕋,)\xi\in H^{\frac{1-s}{2}}(\mathbb{T},\mathbb{R}) and that the remaining term in (1.13) is o((NN)s)o((N\ell_{N})^{s}).

Remark 1.3 (On the adaptation to β\beta-ensembles).

We expect that our method can also give a (log-correlated) CLT for the test-functions 𝟙(a,a)\mathds{1}_{(-a,a)} and xlog|xa|x\mapsto\log|x-a| for 1D log-gases on the circle or on the real line when the external potential is convex.

By Remark 1.2, the number-variance (i.e variance of the number of points) of the Riesz gas grows in O(Ns)O(N^{s}), like the variance of smooth linear statistics. In comparison, for the 1D log-gas, smooth linear statistics fluctuate in O(1)O(1) with an asymptotic variance proportional to the squared Sobolev norm ||H122|\cdot|_{H^{\frac{1}{2}}}^{2} (see [BLS+18] for instance) whereas the number of points in an interval (a,a)(-a,a) fluctuates in O(logN)O(\log N) since 1(a,a)H12(𝕋,)\mathcal{1}_{(-a,a)}\notin H^{\frac{1}{2}}(\mathbb{T},\mathbb{R}). Theorem 2 shows that the Riesz gas with s(0,1)s\in(0,1) interpolates between the 1D log-gas case s=0s=0 and the Poisson-type case s=1s=1. Moreover Theorem 2 makes the 1D long-range Riesz gas a hyperuniform particle system in the sense of [Tor16] (meaning that the number-variance is much smaller than for i.i.d variables).

As mentioned in the beginning of the introduction, the next-order term in the expansion (1.13) corresponds to the variance of a local energy arising from the mean-field transport of [J+98], sometimes referred in the literature as the loop equation term. One could extract the leading-order of this variance and relate it to the second derivative of the free energy of the infinite Riesz gas.

The next question we address concerns the asymptotic behavior of rescaled linear statistics. We show that if ξ\xi satisfies Assumptions 1.1 and provided N1N\ell_{N}\gg\frac{1}{N}, FluctN[ξ(N1)]\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)] converges after suitable rescaling to a Gaussian random variable. For any probability measures μ\mu and ν\nu on \mathbb{R} let us denote 𝖽(μ,ν)\mathsf{d}(\mu,\nu) the distance

𝖽(μ,ν):=supf𝒞2(,){fd(μν):|f|1,|f|1,|f′′|1}.\mathsf{d}(\mu,\nu):=\sup_{f\in\mathcal{C}^{2}(\mathbb{R},\mathbb{R})}\Bigr{\{}\int f\mathrm{d}(\mu-\nu):|f|_{\infty}\leq 1,|f^{\prime}|_{\infty}\leq 1,|f^{\prime\prime}|_{\infty}\leq 1\Bigr{\}}.

We establish the following result:

Theorem 3 (CLT for singular linear statistics).

Let ξ\xi and {N}\{\ell_{N}\} satisfy Assumptions 1.1. Assume that N1N\ell_{N}\gg\frac{1}{N}.

  • The sequence of random variables (NN)s2FluctN[ξ(N1)](N\ell_{N})^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)] converges in distribution to a centered Gaussian random variable with variance σ2(ξ)\sigma_{\infty}^{2}(\xi) defined in (1.12).

  • Let σN2(ξ)\sigma^{2}_{\ell_{N}}(\xi) be as in (1.10). Let

    ZN𝒩(0,σN2(ξ)),Z𝒩(0,σ2(ξ)).Z_{N}\sim\mathcal{N}(0,\sigma_{\ell_{N}}^{2}(\xi)),\quad Z\sim\mathcal{N}(0,\sigma_{\infty}^{2}(\xi)).

    Then for all ε>0\varepsilon>0, we have

    𝖽(Law((NN)s2FluctN[ξ(N1)]),Law(ZN))=O((NN)ε+max(s12,αs2))\mathsf{d}(\mathrm{Law}((N\ell_{N})^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]),\mathrm{Law}(Z_{N}))=O\Bigr{(}(N\ell_{N})^{\varepsilon+\max(\frac{s-1}{2},\alpha-\frac{s}{2})}\Bigr{)}

    and

    𝖽(Law((NN)s2FluctN[ξ(N1)]),Law(Z))=O((NN)ε+max(s12,αs2)+N2s𝟙N0).\mathsf{d}(\mathrm{Law}((N\ell_{N})^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]),\mathrm{Law}(Z))=O\Bigr{(}(N\ell_{N})^{\varepsilon+\max(\frac{s-1}{2},\alpha-\frac{s}{2})}+\ell_{N}^{2-s}\mathds{1}_{\ell_{N}\to 0}\Bigr{)}.

Theorem 3 can be interpreted as the convergence of the field

i=1Ng(xi)Ng(x)dx\sum_{i=1}^{N}g(x_{i}-\cdot)-N\int g(x-\cdot)\mathrm{d}x

to a fractional Gaussian field. Observe that if ξ\xi and χ\chi are smooth test-functions with disjoint support, then the asymptotic covariance is, in general, not equal to 0, meaning that the corresponding fractional field does not exhibit spatial independence. This reflects the non-local nature of the fractional Laplacian (Δ)1s2(-\Delta)^{\frac{1-s}{2}} for s(0,1)s\in(0,1).

Following Remark 1.2, Theorem 3 gives a CLT for gaps and discrepancies:

Corollary 1.1 (CLT for the number of points).

Let a(0,12)a\in(0,\frac{1}{2}) and {N}\{\ell_{N}\} be a sequence taking values on (0,1](0,1] such that N1N\ell_{N}\gg\frac{1}{N}. The sequence of random variables

Ns2ζ(s,2aN)12(i=1N𝟙(aN,aN)(xi)2NN)N^{-\frac{s}{2}}\zeta(-s,2a\ell_{N})^{-\frac{1}{2}}\Bigr{(}\sum_{i=1}^{N}\mathds{1}_{(-a\ell_{N},a\ell_{N})}(x_{i})-2N\ell_{N}\Bigr{)}

converges in distribution to a centered Gaussian random variables with variance

σ2:=cotan(π2s)βπs4s1.\sigma^{2}:=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta\pi s4^{s-1}}.

Moreover

Nsζ(s,2Na)(NN)sN{ζ(s,2a)if N1(2a)sif N0.\frac{N^{s}\zeta(-s,2\ell_{N}a)}{(N\ell_{N})^{s}}\underset{N\to\infty}{\longrightarrow}\begin{cases}\zeta(-s,2a)&\text{if $\ell_{N}\to 1$}\\ (2a)^{s}&\text{if $\ell_{N}\to 0$}.\end{cases}

Let {kN}\{k_{N}\} be a sequence taking values in {1,,N2}\{1,\ldots,\frac{N}{2}\} such that kNk_{N}\to\infty as NN\to\infty. The sequence of random variables

Ns2ζ(s,kNN)(N(xi+kNxi)kN)N^{-\frac{s}{2}}\zeta(-s,\frac{k_{N}}{N})(N(x_{i+k_{N}}-x_{i})-k_{N})

converges in distribution to a centered Gaussian random variables with variance σ2\sigma^{2}.

Corollary 1.1 is an extension of the results on the fluctuations of single particles in the bulk for β\beta-ensembles, see [Gus05, CFLW21] for the GUE case and [BMP22]. Theorem 3 can also be applied to singular function having singularity around 0 in |x|α|x|^{-\alpha} for α(0,s2)\alpha\in(0,\frac{s}{2}).

Corollary 1.2 (CLT for power-type functions).

Let α(0,s2)\alpha\in(0,\frac{s}{2}). The sequence of random variables

Ns2(i=1Nζ(α,xi)Nζ(α,x)dx)N^{-\frac{s}{2}}\Bigr{(}\sum_{i=1}^{N}\zeta(-\alpha,x_{i})-N\int\zeta(-\alpha,x)\mathrm{d}x\Bigr{)}

converges in distribution to a centered Gaussian random variables with variance

cα2βcsζ(1+s2α),\frac{c_{\alpha}^{2}}{\beta c_{s}}\zeta(1+s-2\alpha),

where cαc_{\alpha}, csc_{s} are as in (1.3).

The test-function x𝕋|x|s2x\in\mathbb{T}\mapsto|x|^{-\frac{s}{2}} is the critical inverse power which does not lie in H1s2H^{\frac{1-s}{2}}. This should be compared in the case s=0s=0 to the test-functions 𝟙(a,a)\mathds{1}_{(-a,a)} and log|x|-\log|x|, for which the associated linear statistics satisfy a log-correlated central limit theorem as shown for instance in [BMP22].

1.3. Context, related results, open questions

1.3.1. Rigidity of β\beta-ensembles

As mentioned in the introduction, Theorems 1 and 3 are the natural extensions to the circular Riesz gas of some known results on the fluctuations of β\beta-ensembles. We refer again to [BEY12, BEY+14b, BEY14a, BMP22] for rigidity estimates, to [J+98, Shc13, BG13, BLS+18, LLW+19, HL20, Pei22] for CLTs for linear statistics with smooth test-functions and to [HL20, Lam21] for the case of the circular β\beta-ensemble. In the case of the GUE, that is for β=2\beta=2 with a quadratic potential, a CLT for test-functions in H12H^{\frac{1}{2}} is obtained in [SW13] using a Littlewood-Paley type decomposition argument. However as observed in [Lam21, Rem. 1.3], the minimal regularity of the test-function depends on β\beta. Indeed for β=4\beta=4, leveraging on variances expansions of [JM15], [Lam21] exhibits a test-function in H12H^{\frac{1}{2}} such that the associated linear statistics does not have a finite limit. Since the characteristic function of a given interval is not is H12H^{\frac{1}{2}}, the asymptotic scaling of discrepancies in intervals is not of order 11. It is proved in [Gus05] that for the GUE, eigenvalues xix_{i} in the bulk fluctuate in O(logi)O(\sqrt{\log i}) and that discrepancies are of order logN\sqrt{\log N}. A general CLT for the characteristic functions of intervals and for the logarithm function is given in the recent paper [BMP22]. Concerning the method of proof, let us point out a very similar variation on Stein’s method developed in [LLW+19], see also [HL20] for a high-temperature regime.

1.3.2. Local laws and fluctuations for the Langevin dynamics

A related and much studied question concerns the rigidity of the Dyson Brownian motion, an evolving gas of particles whose invariant distribution is given by β\beta-ensemble. The time to equilibrium at the microscopic scale of Dyson Brownian motion was studied in many papers including [ESY11, EY12], see also [Bou21] for optimal relaxation estimates. A central limit theorem at mesoscopic scale for linear statistics of the Dyson Brownian Motion is established in [HL16], thus exhibiting a time dependent covariance structure.

1.3.3. Decay of the correlations and Helffer-Sjöstrand representation

The decay of the gaps correlations of β\beta-ensembles have been extensively studied in [EY15], where a power-law decay in the inverse squared distance is established. The starting point of [EY15] is based on a a representation of the correlation function by a random walk in a dynamic random environment or in other words on a dynamic interpretation of the Helffer-Sjöstrand operator. The paper [EY15] then develops a sophisticated homogenization theory for a system of discrete parabolic equations. In a different context, a more direct analysis of the Helffer-Sjöstrand operator has been developed in the groundwork [NS97] to characterize the scaling limit of the gradient interface model in arbitrary dimension d1\mathrm{d}\geq 1. Combining ideas from [NS97] and from quantitative stochastic homogenization, the paper [AW19] then shows that the free energy associated to this model is at least 𝒞2,α\mathcal{C}^{2,\alpha} for some α>0\alpha>0. We also refer to the recent paper [Tho21] which studies in a similar framework the scaling limit of the non-Gaussian membrane model. In non-convex settings, much fewer results are available in the literature. One can mention the work [DW20] which establishes the optimal decay for the two-point correlation function of the Villain rotator model in d\mathbb{Z}^{d}, for d3\mathrm{d}\geq 3 at low temperature. It could be interesting to develop a direct method to analyze the large scale decay of the Helffer-Sjöstrand equation in the context of one-dimensional Riesz gases. We plan to address this question in future work.

1.3.4. Uniqueness of the limiting point process

The question of the decay of the correlations mentioned above is related to a property of uniqueness of the limiting measure. One expects that after rescaling, chosen so that the typical distance between consecutive points is of order 11, the point process converges, in a suitable topology, to a certain point process Rieszβ\mathrm{Riesz}_{\beta}. For s=0s=0, the limiting point process called Sineβ\mathrm{Sine}_{\beta}, is unique and universal as proved in [BEY12, BEY+14b]. The existence of a limit was first established in [VV09] for β\beta-ensembles with quadratic exterior potential, together with a sophisticated description and in [KS09] for the circular β\beta-ensemble. The Sineβ\mathrm{Sine}_{\beta} process has also been characterized as the unique minimizer of the free energy functional governing the microscopic behavior in [EHL18] using a displacement convexity argument. In [Bou22], we prove the existence of a limiting point process Rieszβ\mathrm{Riesz}_{\beta} for the circular Riesz ensemble.

1.3.5. 1D hypersingular Riesz gases

Athough the 1D hypersingular Riesz gas (i.e s>1s>1) is not hyperuniform, its fluctuations are also of interest. In such a system, the macroscopic and microscopic behaviors are coupled, a fact which translates into the linear response associated to linear statistics (in contrast with long-range particle systems, the linear response is a combination of a mean-field change of variables, moving each point according to its position only, and of local perturbations). Simple heuristic computations shows that the limiting variance is then proportional to a L2L^{2} norm (after subtraction of the mean) with a factor depending on the second order derivative of the free energy.

1.3.6. Fluctuations of Riesz gases in higher dimension

For d1\mathrm{d}\geq 1 and ss smaller than d\mathrm{d}, the existence of a thermodynamic limit for the Riesz gas (after extraction of a subsequence) is delicate as the energy is long-range. It was obtained in [LS17] for s(min(d2,0),d)s\in(\min(\mathrm{d}-2,0),\mathrm{d}), leveraging among many other ingredients on an electric formulation of the Riesz energy, see [PS17], and on a screening procedure introduced in [SS12] and then improved in [RS13, PS17]. The first task to study the fluctuations of higher dimensional long-range Riesz gases is to establish local laws, that is to control the number of points and the energy in cubes of small scales. This was done for the Coulomb gas in arbitrary dimension down to the microscopic scale in the paper [AS19] using subadditive and supperadditive approximate energies. Due to the lack of convexity, establishing a CLT or even a Poissonian rigidity estimate for linear statistics of Riesz gases in arbitrary dimension is a very delicate task. In dimension 2, since long-range interactions are overwhelmingly dominant, a CLT for linear statistics with smooth test-functions can be proved, see [LS18, BBNY16, Leb17], without proving any “probabilistic cancellation” on local quantities, but only a “quenched cancellation” on some angle term. Let us finally mention the work [Leb21] where the 2D Coulomb gas is shown to be hyperuniform, meaning that the variance of the number of points in a ball scales much smaller than the volume. The paper [Leb21] establishes an important quantitative translation invariance property based on refinements of Mermin-Wagner type arguments, see also [Tho22]. In higher dimension much fewer results are available. One can mention the result of [Ser20] which treats the 3D Coulomb gas at high temperature “under a no phase transition assumption”. A simpler variation of the 3D Colomb gas, named hierarchical Coulomb gas, has also been investigated in the work [Cha19], followed by [GS20].

1.4. Outline of the main proofs

Rigidity

The proof of Theorem 1 is similar to the proof of [BEY12, Th. 3.1]. Let us explain the main steps.

First, we establish a local law on gaps saying that for all δ>0\delta>0, there exist κ>0\kappa>0, c>0,C>0c>0,C>0 such that for each k=1,,N/2k=1,\ldots,\lfloor N/2\rfloor and each i=1,,Ni=1,\ldots,N,

N,β(N(xi+kxi)k1+δ)Ceckκ.\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq Ce^{-ck^{\kappa}}. (1.14)

This estimate is obtained through a bootstrap procedure and some concentration estimates that we do not detail here.

Let us explain how to prove Theorem 1 from (1.14). For each k=1,,N/2k=1,\ldots,\lfloor N/2\rfloor and i=1,,Ni=1,\ldots,N, define

Ik(i):={j{1,,N}:d(i,j)k}I_{k}(i):=\{j\in\{1,\ldots,N\}:d(i,j)\leq k\}

and the block average

xi[k]:=12k+1jIk(i)xj.x_{i}^{[k]}:=\frac{1}{2k+1}\sum_{j\in I_{k}(i)}x_{j}.

Fix kk, ii and let p1p\geq 1 be a large number. Let α:=1p\alpha:=\frac{1}{p}. The idea of [BEY12] is to decompose xix_{i} into

Nxi=Nxi[k]+m=0p1N(xi[kmα]xi[k(m+1)α]).Nx_{i}=Nx_{i}^{[k]}+\sum_{m=0}^{p-1}N(x_{i}^{[\lfloor k^{m\alpha}\rfloor]}-x_{i}^{[\lfloor k^{(m+1)\alpha}\rfloor]}).

For each k=1,,p1k=1,\ldots,p-1, denote

Gm:=N(xi[kmα]xi[k(m+1)α]).G_{m}:=N(x_{i}^{[\lfloor k^{m\alpha}]}-x_{i}^{[\lfloor k^{(m+1)\alpha\rfloor}]}).

One may note that for small mm, |Gm|2|\nabla G_{m}|^{2} gets larger but GmG_{m} is more local since it depends only on the variables in Jm+1:=I[k(m+1)α]J_{m+1}:=I_{[\lfloor k^{(m+1)\alpha}\rfloor]}. We therefore need to exploit the convexity of interactions in Jm+1J_{m+1}. To gain uniform convexity, we let θ:++\theta:\mathbb{R}^{+}\to\mathbb{R}^{+} be a smooth non-negative convex function such that θ=0\theta=0 on [0,12][0,\frac{1}{2}] and θ(x)=x2\theta(x)=x^{2} for x>1x>1 and define for some fixed ε>0\varepsilon>0, the forcing

F:=2i<jJm+1θ(N(xjxi)k1+ε)\mathrm{F}:=2\sum_{i<j\in J_{m+1}}\theta\Bigr{(}\frac{N(x_{j}-x_{i})}{k^{1+\varepsilon}}\Bigr{)}

and the locally constrained measure

dN,β(XN)eβ(N+F)(XN)𝟙DN(XN)dXN.\mathrm{d}\mathbb{Q}_{N,\beta}(X_{N})\propto e^{-\beta(\mathcal{H}_{N}+\mathrm{F})(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N}.

Let μ\mu be the law of (xi)iJm+1(x_{i})_{i\in J_{m+1}} when XNN,βX_{N}\sim\mathbb{Q}_{N,\beta}. Using the so-called Brascamp-Lieb inequality, one can show that

dμ(x)=eβ(H+F(x)+H~(x))𝟙D|Jm+1|(x)dx,\mathrm{d}\mu(x)=e^{-\beta(H+\mathrm{F}(x)+\tilde{H}(x))}\mathds{1}_{D_{|J_{m+1}|}}(x)\mathrm{d}x,

with 2H~0\nabla^{2}\tilde{H}\geq 0 and HH defined by

H:x=(xi)iJm+1D|Jm+1|NsijJm+1g(xjxi).H:x=(x_{i})_{i\in J_{m+1}}\in D_{|J_{m+1}|}\mapsto N^{-s}\sum_{i\neq j\in J_{m+1}}g(x_{j}-x_{i}).

For all U|Jm+1|U\in\mathbb{R}^{|J_{m+1}|}, by construction, we have

U2(H+F)U=2i<jJm+1(Nsg′′(xjxi)+θ′′(NK(xjxi))(Nk)2)(ujui)22Ns(k(m+1)α(1+ε)N)(s+2)i<jJm+1(ujui)2.U\cdot\nabla^{2}(H+\mathrm{F})U=2\sum_{i<j\in J_{m+1}}\Bigr{(}N^{-s}g^{\prime\prime}(x_{j}-x_{i})+\theta^{\prime\prime}\Bigr{(}\frac{N}{K}(x_{j}-x_{i})\Bigr{)}\Bigr{(}\frac{N}{k}\Bigr{)}^{2}\Bigr{)}(u_{j}-u_{i})^{2}\\ \geq 2N^{-s}\Bigr{(}\frac{k^{(m+1)\alpha(1+\varepsilon)}}{N}\Bigr{)}^{-(s+2)}\sum_{i<j\in J_{m+1}}(u_{j}-u_{i})^{2}.

The point is that when iJm+1ui=0\sum_{i\in J_{m+1}}u_{i}=0,

U2HUc1iJm+1ui2where c1:=(|Jm+1|1)Ns(k(m+1)αN)(s+2)(1+ε).U\cdot\nabla^{2}HU\geq c_{1}\sum_{i\in J_{m+1}}u_{i}^{2}\quad\text{where $c_{1}:=(|J_{m+1}|-1)N^{-s}\Bigr{(}\frac{k^{(m+1)\alpha}}{N}\Bigr{)}^{-(s+2)(1+\varepsilon)}$}.

Besides, one can observe that

|Gm|2=O(N2kmα).|\nabla G_{m}|^{2}=O\Bigr{(}\frac{N^{2}}{k^{m\alpha}}\Bigr{)}.

Since HH, F\mathrm{F}, H~\tilde{H} and GmG_{m} are independent of iJm+1xi\sum_{i\in J_{m+1}}x_{i}, one can argue with the Bakry-Emery criterion that for all t0t\geq 0,

log𝔼N,β[etGm]t𝔼N,β[Gm]+O(t2N2kmαc11)=O(ks+κε+2α).\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{tG_{m}}]\leq t\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G_{m}]+O\Bigr{(}t^{2}\frac{N^{2}}{k^{m\alpha}}c_{1}^{-1}\Bigr{)}=O(k^{s+\kappa\varepsilon+2\alpha}).

Using the log-Sobolev inequality and the local law (1.14), one can easily show by taking α\alpha small enough, that for all ε>0\varepsilon>0, Gm=O(ks2+ε)G_{m}=O(k^{\frac{s}{2}+\varepsilon}) with high probability (under N,β\mathbb{P}_{N,\beta}), which shows that

N(xi+kxi)=N(xi+k[k]xi[k])+O(ks2+ε),N(x_{i+k}-x_{i})=N(x_{i+k}^{[k]}-x_{i}^{[k]})+O(k^{\frac{s}{2}+\varepsilon}),

with high probability. By similar augments, it is easy to prove that N(xi+k[k]xi[k])=k+O(ks2+ε)N(x_{i+k}^{[k]}-x_{i}^{[k]})=k+O(k^{\frac{s}{2}+\varepsilon}) with high probability, which will conclude the proof of Theorem 1.

We now explain the general strategy to obtain the variance expansion formula of Theorem 2 and the CLT of Theorem 3. To simplify assume that N=1\ell_{N}=1. We are interested in the fluctuations of the linear statistic FluctN[ξ]\mathrm{Fluct}_{N}[\xi], where ξ:𝕋\xi:\mathbb{T}\to\mathbb{R} is a piecewise smooth function satisfying Assumptions 1.1. Let α(s1,s2)\alpha\in(s-1,\frac{s}{2}) such that (1.8) holds.

Mollification

Let KK be a smooth kernel supported on (12,12)(-\frac{1}{2},\frac{1}{2}). Let :=N1\ell:=N^{-1}. Define K:=1K(1)K_{\ell}:=\ell^{-1}K(\ell^{-1}\cdot). Let us recall that for all fL2(𝕋,)f\in L^{2}(\mathbb{T},\mathbb{R}),

VarN,β[FluctN[f]]N|f|L22.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[f]]\leq N|f|^{2}_{L^{2}}. (1.15)

Using this, one may replace ξ\xi by ξK\xi*K_{\ell} up to an error

VarN,β[FluctN[ξξK]]N|ξξK|L22C(N1+N2α)=o(Ns).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi-\xi*K_{\ell}]]\leq N|\xi-\xi*K_{\ell}|_{L^{2}}^{2}\leq C(N^{-1}+N^{2\alpha})=o(N^{s}). (1.16)

It is therefore sufficient to study the fluctuations of the function ξreg:=ξK\xi_{\mathrm{reg}}:=\xi*K_{\ell}.

The Helffer-Sjöstrand equation

Let F:DNF:D_{N}\to\mathbb{R} be smooth enough. The fluctuations of FF are related to a partial differential equation through the representation

VarN,β[F]=𝔼N,β[Fϕ],\mathrm{Var}_{\mathbb{P}_{N,\beta}}[F]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\nabla F\cdot\nabla\phi], (1.17)

where ϕ:DN\phi:D_{N}\to\mathbb{R} solves the Poisson equation

{ϕ=F𝔼N,β[F]on DNϕn=0on DN,\begin{cases}\mathcal{L}\phi=F-\mathbb{E}_{\mathbb{P}_{N,\beta}}[F]&\text{on }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{on }\partial D_{N},\end{cases} (1.18)

where \mathcal{L} stands for the generator

:=βNΔ.\mathcal{L}:=\beta\nabla\mathcal{H}_{N}\cdot\nabla-\Delta.

Note that (1.17) directly follows by integration by parts once it is known that (1.18) has a solution. Differentiating (1.18), one obtains the so-called Helffer-Sjöstrand equation which reads

{A1ϕ=Fon DNϕn=0on DN,\begin{cases}A_{1}\nabla\phi=\nabla F&\text{on }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{on }\partial D_{N},\end{cases} (1.19)

with A1A_{1} formally given by

A1:=β2N+IN.A_{1}:=\beta\nabla^{2}\mathcal{H}_{N}+\mathcal{L}\otimes I_{N}.

When FF is a function of the gaps or when gg is replaced by a bounded kernel, the existence and uniqueness of a solution of (1.18) roughly follow from Lax-Milgram’s lemma. Indeed, under such assumptions, a Poincaré inequality holds, which ensures the coerciveness of the appropriate bilinear form. To analyze the solution of (1.19), we will use various tools based on mean-field approximations, convexity and monotonicity.

Since ijN0\partial_{ij}\mathcal{H}_{N}\leq 0 for each iji\neq j, it is standard that N,β\mathbb{P}_{N,\beta} satisfies the FKG inequality, meaning that the covariance between two increasing functions is non-negative. This can be formulated by saying that 1\mathcal{L}^{-1} preserves the cone of increasing functions: if F0\nabla F\geq 0 (coordinate wise), then ϕ:=A11ϕ0\nabla\phi:=A_{1}^{-1}\nabla\phi\geq 0. A nice consequence is the following: if F,G:DNF,G:D_{N}\to\mathbb{R} are such that |F|G|\nabla F|\leq\nabla G, then

VarN,β[F]VarN,β[G].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[F]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[G]. (1.20)

This comparison principle can be extended to non-gradient vector-fields, which will be used as a key argument to handle the fluctuations of some complicated singular functions.

Mean-field transport

It turns out that when FF is a linear statistic, i.e F=FluctN[f]F=\mathrm{Fluct}_{N}[f] for some smooth enough test-function f:𝕋f:\mathbb{T}\to\mathbb{R}, then the solution ϕ\nabla\phi of (1.19) can be approximated by a transport ΨN\Psi_{N} in the form XNDN1N1s(ψ(x1),,ψ(xN))X_{N}\in D_{N}\mapsto\frac{1}{N^{1-s}}(\psi(x_{1}),\ldots,\psi(x_{N})) for some well-chosen map ψ:𝕋\psi:\mathbb{T}\to\mathbb{R}. Letting Δ:={(x,y)𝕋2:x=y}\Delta:=\{(x,y)\in\mathbb{T}^{2}:x=y\}, one may write

NΨN=NΔcg(xy)(ψ(x)ψ(y))dμN(x)dμN(y),\nabla\mathcal{H}_{N}\cdot\Psi_{N}=N\iint_{\Delta^{c}}g^{\prime}(x-y)(\psi(x)-\psi(y))\mathrm{d}\mu_{N}(x)\mathrm{d}\mu_{N}(y),

where μN:=1Ni=1Nδxi\mu_{N}:=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}. Let us expand μN\mu_{N} around the Lebesgue measure dx\mathrm{d}x on 𝕋\mathbb{T} and denote fluctN:=N(μNdx)\mathrm{fluct}_{N}:=N(\mu_{N}-\mathrm{d}x). Noting that the constant term vanishes, one can check that

NΨN=2(gψ)fluctN+1N1sA1[ψ]\nabla\mathcal{H}_{N}\cdot\Psi_{N}=2\int(-g^{\prime}*\psi)\mathrm{fluct}_{N}+\frac{1}{N^{1-s}}\mathrm{A}_{1}[\psi] (1.21)

with

A1[ψ]:=Δc(ψ(x)ψ(y))Nsg(xy)dfluctN(x)dfluctN(y).\mathrm{A}_{1}[\psi]:=\iint_{\Delta^{c}}(\psi(x)-\psi(y))N^{-s}g^{\prime}(x-y)\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y). (1.22)

The leading-order of (1.21) being a linear statistic, one can choose ψ\psi such that

βNΨNdivΨNF\beta\nabla\mathcal{H}_{N}\cdot\Psi_{N}-\mathrm{div}\ \Psi_{N}\simeq F

by letting ψ\psi such that

ψ=12βcs(Δ)1s2f\psi^{\prime}=-\frac{1}{2\beta c_{s}}(-\Delta)^{\frac{1-s}{2}}f

and ψ=0\int\psi=0. We will apply this to f=ξregf=\xi_{\mathrm{reg}}. Denoting ψreg\psi_{\mathrm{reg}} the above transport, the central task of the paper is to show that

VarN,β[A1[ψreg]]CN1+ε|ψreg|L22,\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{1}[\psi_{\mathrm{reg}}]]\leq CN^{1+\varepsilon}|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2}, (1.23)

for all ε>0\varepsilon>0.

Splitting the variance of the next-order term

Denote

ζ:(x,y)𝕋2ψ(x)ψ(y)xy\zeta:(x,y)\in\mathbb{T}^{2}\mapsto\frac{\psi(x)-\psi(y)}{x-y}

so that

A1[ψ]=Δcζ(x,y)Nsg~(xy)dfluctN(x)dfluctN(y),\mathrm{A}_{1}[\psi]=\iint_{\Delta^{c}}\zeta(x,y)N^{-s}\tilde{g}(x-y)\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y), (1.24)

where g~:x𝕋{0}xg(x)\tilde{g}:x\in\mathbb{T}\setminus\{0\}\mapsto xg^{\prime}(x). Note that for each index i=1,,Ni=1,\ldots,N

iA1[ψ]\displaystyle\partial_{i}\mathrm{A}_{1}[\psi]
=2yxi1ζ(xi,y)Nsg~(xy)dfluctN(y)Vi+2yxiζ(xi,y)Nsg~(xy)dfluctN(y)Wi.\displaystyle\quad=\underbrace{2\int_{y\neq x_{i}}\partial_{1}\zeta(x_{i},y)N^{-s}\tilde{g}(x-y)\mathrm{d}\mathrm{fluct}_{N}(y)}_{\simeq\mathrm{V}_{i}}+\underbrace{2\int_{y\neq x_{i}}\zeta(x_{i},y)N^{-s}\tilde{g}^{\prime}(x-y)\mathrm{d}\mathrm{fluct}_{N}(y)}_{\simeq\mathrm{W}_{i}}.

We have thus split A[ψreg]\nabla\mathrm{A}[\psi_{\mathrm{reg}}] into a macroscopic force V\mathrm{V} and a microscopic force W\mathrm{W} (the splitting is in fact slightly different). By subadditivity, it follows that

VarN,β[A[ψreg]]2𝔼N,β[VA11V](I)+2𝔼N,β[WA11W](II).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}[\psi_{\mathrm{reg}}]]\leq 2\underbrace{\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{V}\cdot A_{1}^{-1}\mathrm{V}]}_{({I})}+2\underbrace{\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{W}\cdot A_{1}^{-1}\mathrm{W}]}_{({II})}.
Control on (II)({II}) with Poincaré inequality in gap coordinates

In gap coordinates the microscopic force W\mathrm{W} behaves well: there exists W~\tilde{\mathrm{W}} such that for all UNNU_{N}\in\mathbb{R}^{N},

WUN=i=1NW~iN(ui+1ui)\mathrm{W}\cdot U_{N}=-\sum_{i=1}^{N}\tilde{\mathrm{W}}_{i}N(u_{i+1}-u_{i})

satisfying typically (i.e with overwhelming probability) the estimate

|W~|2CN1+ε|ψreg|L22|\tilde{\mathrm{W}}|^{2}\leq CN^{1+\varepsilon}|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2} (1.25)

for all ε>0\varepsilon>0. By penalizing configurations with large nearest-neighbor distances, one can modify the Gibbs measure into a new one being uniformly log-concave with respect to the variables N(xi+1xi),i=1,,NN(x_{i+1}-x_{i}),i=1,\ldots,N. By Brascamp-Lieb inequality, we get using (1.25),

(II)CN1+ε|ψreg|L22,(II)\leq CN^{1+\varepsilon}|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2}, (1.26)

for all ε>0\varepsilon>0.

Control on (I)({I}) with the comparison principle

In substance, one should think of V\mathrm{V} as satisfying for each i=1,,Ni=1,\ldots,N

|Vi|C|ψreg′′(xi)|+“Lower order terms”.|\mathrm{V}_{i}|\leq C|\psi_{\mathrm{reg}}^{\prime\prime}(x_{i})|+\text{``Lower order terms''}. (1.27)

Note that for instance if ξ=𝟙(a,b)\xi=\mathds{1}_{(a,b)}, ψreg′′\psi_{\mathrm{reg}}^{\prime\prime} blows like (|x|η)(2s)(|x|\vee\eta)^{-(2-s)} near aa and bb. Therefore for such singular ψreg\psi_{\mathrm{reg}}, the Poincaré inequality does not provide satisfactory estimates for (I)({I}). If Vi\mathrm{V}_{i} was exactly given by ψreg′′(xi)\psi_{\mathrm{reg}}^{\prime\prime}(x_{i}) for each i=1,,Ni=1,\ldots,N, one could bound 𝔼N,β[VA11V]\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{V}\cdot A_{1}^{-1}\mathrm{V}] by N|ψreg|L22N|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2} by (1.15).

The idea is to use the comparison principle (1.20) to compare 𝔼N,β[VA11V]\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{V}\cdot A_{1}^{-1}\mathrm{V}] to the variance of a linear statistic, which are easier to handle using for instance (1.15). Let ζN:𝕋\zeta_{N}:\mathbb{T}\to\mathbb{R} such that ζN=C|ψreg′′|\zeta_{N}^{\prime}=C|\psi_{\mathrm{reg}}^{\prime\prime}|. Equation (1.27) can be put in the form

|V|FluctN[ζN].|\mathrm{V}|\leq\nabla\mathrm{Fluct}_{N}[\zeta_{N}].

It then follows from (1.20) that

𝔼N,β[VA11V]VarN,β[FluctN[ζN]]+“Lower order terms”\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{V}\cdot A_{1}^{-1}\mathrm{V}]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\zeta_{N}]]+\text{``Lower order terms''}

and the variance of FluctN[ζN]\mathrm{Fluct}_{N}[\zeta_{N}] is then roughly bounded by

VarN,β[FluctN[ζN]]N|ζN|L22CNmax(1,2(1s+α)),\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\zeta_{N}]]\leq N|\zeta_{N}|_{L^{2}}^{2}\leq CN^{\max(1,2(1-s+\alpha))},

which yields

(I)CNmax(1,2(1s+α))CN|ψreg|L22.({I})\leq CN^{\max(1,2(1-s+\alpha))}\leq CN|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2}. (1.28)

Let us emphasize that ζN\zeta_{N} is in fact slightly more complicated staring a term in |x|s2|x|^{-\frac{s}{2}}. We thus need to bound the fluctuations of the critical inverse-power |x|s2|x|^{-\frac{s}{2}}, which is done using the rigidity estimates of Theorem 1 as well as a bootstrap argument, by rerunning the previous steps for a singular test-function.

Combining (1.26) and (1.28) gives (1.23), easily concluding the proof of Theorem 2.

Central limit theorem

The starting point of the proof of the CLT of Theorem 3 is very similar to [LLW+19] and proceeds by Stein’s method. Let GN:=Ns2FluctN[ξreg]G_{N}:=N^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}]. We shall prove that for all η𝒞1(,)\eta\in\mathcal{C}^{1}(\mathbb{R},\mathbb{R}) such that |η|1|\eta^{\prime}|_{\infty}\leq 1, up to a small error term,

𝔼N,β[η(GN)GN]=σ2(ξreg)𝔼N,β[η(GN)]+ErrorN,\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})G_{N}]=\sigma_{\infty}^{2}(\xi_{\mathrm{reg}})\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta^{\prime}(G_{N})]+\mathrm{Error}_{N}, (1.29)

with σ2(ξreg)\sigma_{\infty}^{2}(\xi_{\mathrm{reg}}) as in (1.12). The fundamental observation of Stein is that this approximate integration by parts formula quantifies a distance to normality. Indeed letting ZZ be a centered random variable with variance σ2(ξreg)\sigma_{\infty}^{2}(\xi_{\mathrm{reg}}) and h:h:\mathbb{R}\to\mathbb{R} smooth, one can solve the ODE

σ2(ξreg)xη(x)η(x)=h(x)𝔼[h(Z)]\sigma_{\infty}^{2}(\xi_{\mathrm{reg}})x\eta(x)-\eta^{\prime}(x)=h(x)-\mathbb{E}[h(Z)] (1.30)

and (1.29) can be written in the form

𝔼N,β[h(GN)]𝔼[h(Z)]=ErrorN,\mathbb{E}_{\mathbb{P}_{N,\beta}}[h(G_{N})]-\mathbb{E}[h(Z)]=\mathrm{Error}_{N},

showing that GNG_{N} is approximately Gaussian. Let us explain how to obtain (1.29). Let ϕ=A11GN\nabla\phi=A_{1}^{-1}\nabla G_{N}. By integration by parts we have

𝔼N,β[η(GN)GN]=𝔼N,β[σ2(ξreg)η(GN)GNϕ].\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})G_{N}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\sigma_{\infty}^{2}(\xi_{\mathrm{reg}})\eta^{\prime}(G_{N})\nabla G_{N}\cdot\nabla\phi]. (1.31)

The goal is then to prove that GNϕ\nabla G_{N}\cdot\nabla\phi concentrates around σ2(ξreg)\sigma_{\infty}^{2}(\xi_{\mathrm{reg}}). As explained in the second paragraph, ϕ\nabla\phi may be approximated by the transport N1+s2Ψ{N^{-1+\frac{s}{2}}}\Psi with Ψ:XNDN(ψreg(x1),,ψreg(xN))\Psi:X_{N}\in D_{N}\mapsto(\psi_{\mathrm{reg}}(x_{1}),\ldots,\psi_{\mathrm{reg}}(x_{N})). Performing this approximate transport allows one to replace (1.31) by

𝔼N,β[η(GN)GN]σ2(ξreg)𝔼N,β[η(GN)]=𝔼N,β[η(GN)(1Ni=1Nξreg(xi)ψ(xi)σ2(ξreg))](I)+1N1s2CovN,β[η(GN),βA1[ψreg]FluctN[ψreg]](II).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})G_{N}]-\sigma_{\infty}^{2}(\xi_{\mathrm{reg}})\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta^{\prime}(G_{N})]=\underbrace{\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\eta^{\prime}(G_{N})\Bigr{(}\frac{1}{N}\sum_{i=1}^{N}\xi_{\mathrm{reg}}^{\prime}(x_{i})\psi(x_{i})-\sigma_{\infty}^{2}(\xi_{\mathrm{reg}})\Bigr{)}\Bigr{]}}_{(I)}\\ +\underbrace{\frac{1}{N^{1-\frac{s}{2}}}\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[\eta(G_{N}),\beta\mathrm{A}_{1}[\psi_{\mathrm{reg}}]-\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}]]}_{(II)}. (1.32)

The error term (I)(I) is handled with the local laws, the error term (II)(II) by inserting (1.23) which concludes the proof of the CLT together with (1.16).

1.5. Structure of the paper

  • Section 2 records some useful preliminaries on the fractional Laplacian on the circle.

  • Section 3 shows the well-posedness of the Helffer-Sjöstrand equation and gives various consequences of convexity and monotonicity.

  • Section 4 completes the proof of the near-optimal rigidity result of Theorem 1.

  • Section 5 gives a proof of the variance expansion of Theorem 2.

  • Section 6 contains the proof of the CLT of Theorem 3 and its corollaries.

1.6. Notation

  • We denote d:{1,,N}2d:\{1,\ldots,N\}^{2}\to\mathbb{N} the symmetric distance defined for each 1i,jN1\leq i,j\leq N by

    d(i,j)=min(|ji|,N|ji|).d(i,j)=\min(|j-i|,N-|j-i|).

    We let

    Δ:={(x,y)𝕋2:x=y}.\Delta:=\{(x,y)\in\mathbb{T}^{2}:x=y\}.
  • For any vector-field ψ:Ωm\psi:\Omega\to\mathbb{R}^{m} where Ω\Omega is an open set of n\mathbb{R}^{n}, we let DψD\psi be the matrix of partial derivatives of ψ\psi. We also write 2f\nabla^{2}f for the Hessian of a real-valued function ff.

  • For all α(0,1)\alpha\in(0,1) we let 𝒞α(𝕋,)\mathcal{C}^{\alpha}(\mathbb{T},\mathbb{R}) be the space of α\alpha-Hölder continuous functions from 𝕋\mathbb{T} to \mathbb{R} and 𝒞α(𝕋,)\mathcal{C}^{-\alpha}(\mathbb{T},\mathbb{R}) the dual of 𝒞α(𝕋,)\mathcal{C}^{\alpha}(\mathbb{T},\mathbb{R}).

Throughout the paper, CC and κ\kappa will denote positive constants depending on β\beta and ss that may change from line to line.

1.7. Acknowledgments

I would like to thank Sylvia Serfaty, Thomas Leblé, Djalil Chafaï, Gaultier Lambert and David Garcia-Zelada for many helpful comments. I would also like to thank the anonymous referee for their invaluable feedback on an earlier version of this manuscript.

J.B. was supported by a grant from the “Fondation CFM pour la Recherche” and by the ERC Project LDRAM, ERC-2019-ADG Project 884584.

2. Preliminaries

2.1. The fundamental solution of the fractional Laplacian on the circle

We begin by justifying that the fundamental solution of the fractional Laplace equation on the circle is given by (1.1). Roughly, (1.1) corresponds to the periodic summation of x|x|sx\mapsto|x|^{-s}, which is the fundamental solution of (Δ)1s2(-\Delta)^{\frac{1-s}{2}} on the real line.

For all complex variables ss and aa such that Re(s)>1\mathrm{Re}(s)>1 and a0,1,2,a\neq 0,-1,-2,\ldots, set

ζ(s,a)=n=01(n+a)s.\zeta(s,a)=\sum_{n=0}^{\infty}\frac{1}{(n+a)^{s}}. (2.1)

Given a0,1,2,a\neq 0,-1,-2,\ldots, one can uniquely extend ζ(,a)\zeta(\cdot,a) into a meromorphic function on the full complex plane with a unique pole at s=1s=1, which is simple with a residue equal to 11. This function is called the Hurwitz zeta function [Ber72].

Lemma 2.1 (Fundamental solution).

Let gg be the solution of the fractional Laplace equation on the circle,

{(Δ)1s2g=cs(δ01),g=0,\begin{cases}(-\Delta)^{\frac{1-s}{2}}g=c_{s}(\delta_{0}-1)\,,\\ \int g=0,\end{cases} (2.2)

with csc_{s} as in (1.3). Let ζ(s,x)\zeta(s,x) be the Hurwitz zeta function. Then for all x𝕋{0}x\in\mathbb{T}\setminus\{0\},

g(x)=ζ(s,x)+ζ(s,1x)=limn(k=nn1|k+x|s21sn1s).g(x)=\zeta(s,x)+\zeta(s,1-x)=\lim_{n\to\infty}\Bigr{(}\sum_{k=-n}^{n}\frac{1}{|k+x|^{s}}-\frac{2}{1-s}n^{1-s}\Bigr{)}\,. (2.3)

The above lemma is standard but for completeness we add a proof following roughly [RS16].

Proof.

Let gg be the unique solution of (2.2). We first derive the semi-group representation for (Δ)1s2(-\Delta)^{-\frac{1-s}{2}}. Let α(0,1)\alpha\in(0,1). Recall

λα=1Γ(α)0eλtdtt1α,for all λ>0.\lambda^{-\alpha}=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}e^{-\lambda t}\frac{\mathrm{d}t}{t^{1-\alpha}},\quad\text{for all $\lambda>0$}\,. (2.4)

For an integrable function on the torus on the torus and kk\in\mathbb{Z}, we let f^(k)\hat{f}(k) be kk-th component of the Fourier series of ff, namely

f^(k)=12π𝕋f(y)eikydy.\hat{f}(k)=\frac{1}{2\pi}\int_{\mathbb{T}}f(y)e^{-iky}\mathrm{d}y\,.

The fractional Laplacian (Δ)α(-\Delta)^{-\alpha} is defined by the following Fourier multiplier: for all f𝒞(𝕋,)f\in\mathcal{C}^{\infty}(\mathbb{T},\mathbb{R}) such that f=0\int f=0, letting g=(Δ)αfg=(-\Delta)^{-\alpha}f, we have

g^(k)=|k|2αf^(k),for all k.\hat{g}(k)=|k|^{-2\alpha}\hat{f}(k),\quad\text{for all $k\in\mathbb{Z}$}\,.

Let kk\in\mathbb{Z}. Let f𝒞(𝕋,)f\in\mathcal{C}^{\infty}(\mathbb{T},\mathbb{R}) such that f=0\int f=0 and let g=(Δ)αfg=(-\Delta)^{-\alpha}f. Applying (2.4) to λ=|k|2\lambda=|k|^{2} gives

g^(k)=1Γ(α)0e|k|2tf^(k)dtt1α.\hat{g}(k)=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}e^{-|k|^{2}t}\widehat{f}(k)\frac{\mathrm{d}t}{t^{1-\alpha}}\,. (2.5)

Let (Wt)t0(W_{t})_{t\geq 0} be the heat kernel of the Laplacian on 𝕋\mathbb{T}. Recall that the Fourier coefficients of WtW_{t} are given by

Wt^(k)=e|k|t2,for all kt0.\widehat{W_{t}}(k)=e^{-|k|t^{2}},\quad\text{for all $k\in\mathbb{Z}$, $t\geq 0$}\,.

The heat kernel WtW_{t} can be expressed as

Wt(x)=12πket|k|2eikx=14πtke|xk|24t,for all x𝕋.W_{t}(x)=\frac{1}{2\pi}\sum_{k\in\mathbb{Z}}e^{-t|k|^{2}}e^{ikx}=\frac{1}{\sqrt{4\pi t}}\sum_{k\in\mathbb{Z}}e^{-\frac{|x-k|^{2}}{4t}},\quad\text{for all $x\in\mathbb{T}$}\,. (2.6)

One may rewrite (2.5) as

g^(k)=1Γ(α)0fWt^(k)dtt1α.\hat{g}(k)=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}\widehat{f*W_{t}}(k)\frac{\mathrm{d}t}{t^{1-\alpha}}.

It follows that

(Δ)αf=1Γ(α)0fWtdtt1α.(-\Delta)^{-\alpha}f=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}f*W_{t}\frac{\mathrm{d}t}{t^{1-\alpha}}. (2.7)

Taking α=1s2(0,1)\alpha=\frac{1-s}{2}\in(0,1), we deduce with a regularization argument that

cs(Δ)1s2(δ01)=csΓ(1s2)0(Wt(x)1)dtt1+s2=csΓ(1s2)14π0k(e|xk|24t𝕋e|xk|24tdx)dtt1+s2,c_{s}(-\Delta)^{-\frac{1-s}{2}}(\delta_{0}-1)=\frac{c_{s}}{\Gamma(\frac{1-s}{2})}\int_{0}^{\infty}(W_{t}(x)-1)\frac{\mathrm{d}t}{t^{\frac{1+s}{2}}}\\ =\frac{c_{s}}{\Gamma(\frac{1-s}{2})}\frac{1}{\sqrt{4\pi}}\int_{0}^{\infty}\sum_{k\in\mathbb{Z}}\Bigr{(}e^{-\frac{|x-k|^{2}}{4t}}-\int_{\mathbb{T}}e^{-\frac{|x-k|^{2}}{4t}}\mathrm{d}x\Bigr{)}\frac{\mathrm{d}t}{t^{1+\frac{s}{2}}}, (2.8)

where we have used the second equality in (2.6) to compute the average of WtW_{t} on 𝕋\mathbb{T}.

Define the sequence of functions

uk:t+1t1+s2(e|xk|24t𝕋e|yk|24tdy),k.u_{k}:t\in\mathbb{R}^{+*}\mapsto\frac{1}{t^{1+\frac{s}{2}}}\Bigr{(}e^{-\frac{|x-k|^{2}}{4t}}-\int_{\mathbb{T}}e^{-\frac{|y-k|^{2}}{4t}}\mathrm{d}y\Bigr{)},\quad k\in\mathbb{Z}^{*}.

First observe that when t1t\geq 1,

ke|k|2tet2ke12|k|2tCet2.\sum_{k\in\mathbb{Z}^{*}}e^{-|k|^{2}t}\leq e^{-\frac{t}{2}}\sum_{k\in\mathbb{Z}^{*}}e^{-\frac{1}{2}|k|^{2}t}\leq Ce^{-\frac{t}{2}}.

It follows that

k01|uk|(t)dtCk11t1s2e|k|2tdtC11t1s2et2<.\sum_{k\in\mathbb{Z}^{*}}\int_{0}^{1}|u_{k}|(t)\mathrm{d}t\leq C\sum_{k\in\mathbb{Z}^{*}}\int_{1}^{\infty}\frac{1}{t^{1-\frac{s}{2}}}e^{-|k|^{2}t}\mathrm{d}t\leq C\int_{1}^{\infty}\frac{1}{t^{1-\frac{s}{2}}}e^{-\frac{t}{2}}<\infty. (2.9)

To treat the other part of the integral, we can write

uk(t)=1t1+s2𝕋(e|xk|24te|yk|24t)dy=1t1+s2e|kx|24t|ky|>|kx|(1e|kx|2|ky|24t)dy+1t1+s2|ky|<|kx|e|ky|24t(1e|ky|2|kx|24t)dy.u_{k}(t)=\frac{1}{t^{1+\frac{s}{2}}}\int_{\mathbb{T}}\Bigr{(}e^{-\frac{|x-k|^{2}}{4t}}-e^{-\frac{|y-k|^{2}}{4t}}\Bigr{)}\mathrm{d}y=\frac{1}{t^{1+\frac{s}{2}}}e^{-\frac{|k-x|^{2}}{4t}}\int_{|k-y|>|k-x|}\Bigr{(}1-e^{\frac{|k-x|^{2}-|k-y|^{2}}{4t}}\Bigr{)}\mathrm{d}y\\ +\frac{1}{t^{1+\frac{s}{2}}}\int_{|k-y|<|k-x|}e^{-\frac{|k-y|^{2}}{4t}}\Bigr{(}1-e^{\frac{|k-y|^{2}-|k-x|^{2}}{4t}}\Bigr{)}\mathrm{d}y.

As a consequence, there exists a constant C>0C>0 such that for each kk\in\mathbb{Z} and all tt\in\mathbb{R},

|uk(t)|Ckt2+s2e(k1)24t.|u_{k}(t)|\leq\frac{Ck}{t^{2+\frac{s}{2}}}e^{-\frac{(k-1)^{2}}{4t}}.

When u1u\geq 1, by comparison to a Gaussian integral, one may check that

kke|k|24u2Cu2,\sum_{k\in\mathbb{Z}^{*}}ke^{-\frac{|k|^{2}}{4u^{2}}}\leq Cu^{2},

which leads to

1k|uk(t)|dtC11t1+s2dt<.\int_{1}^{\infty}\sum_{k\in\mathbb{Z}}|u_{k}(t)|\mathrm{d}t\leq C\int_{1}^{\infty}\frac{1}{t^{1+\frac{s}{2}}}\mathrm{d}t<\infty. (2.10)

Combining (2.9) and (2.10), we deduce by Fubini’s theorem that the order of integration and summation in (2.8) can be inverted and we find

g(x)=csΓ(1s2)4πk0uk(t)dt=csΓ(1s2)4πk(0e|xk|24tdtt1+s20𝕋e|yk|24tdtt1+s2dy)=Γ(s2)csΓ(1s2)4πk(1|xk2|s𝕋dy|yk2|s)=limn(k=nn1|x+k|s21sn1s),\begin{split}g(x)&=\frac{c_{s}}{\Gamma(\frac{1-s}{2})\sqrt{4\pi}}\sum_{k\in\mathbb{Z}}\int_{0}^{\infty}u_{k}(t)\mathrm{d}t\\ &=\frac{c_{s}}{\Gamma(\frac{1-s}{2})\sqrt{4\pi}}\sum_{k\in\mathbb{Z}}\Bigr{(}\int_{0}^{\infty}e^{-\frac{|x-k|^{2}}{4}t}\frac{\mathrm{d}t}{t^{1+\frac{s}{2}}}-\int_{0}^{\infty}\int_{\mathbb{T}}e^{-\frac{|y-k|^{2}}{4}t}\frac{\mathrm{d}t}{t^{1+\frac{s}{2}}}\mathrm{d}y\Bigr{)}\\ &=\frac{\Gamma(\frac{s}{2})c_{s}}{\Gamma(\frac{1-s}{2})\sqrt{4\pi}}\sum_{k\in\mathbb{Z}}\Bigr{(}\frac{1}{|\frac{x-k}{2}|^{s}}-\int_{\mathbb{T}}\frac{\mathrm{d}y}{|\frac{y-k}{2}|^{s}}\Bigr{)}\\ &=\lim_{n\to\infty}\Bigr{(}\sum_{k=-n}^{n}\frac{1}{|x+k|^{s}}-\frac{2}{1-s}n^{1-s}\Bigr{)},\end{split}

where we have used the change of variables t+1tt\in\mathbb{R}^{+*}\mapsto\frac{1}{t} and (2.4) in the third equality.

Let us provide an alternative expression of gg. Using the Euler-Maclaurin formula, one gets that for all a𝕋{0}a\in\mathbb{T}\setminus\{0\},

limn(k=0n1(k+a)s11sn1s)=11sa1s+1ass0tt12(t+a)1+sdt.\lim_{n\to\infty}\Bigr{(}\sum_{k=0}^{n}\frac{1}{(k+a)^{s}}-\frac{1}{1-s}n^{1-s}\Bigr{)}=-\frac{1}{1-s}a^{1-s}+\frac{1}{a^{s}}-s\int_{0}^{\infty}\frac{t-\lfloor t\rfloor-\frac{1}{2}}{(t+a)^{1+s}}\mathrm{d}t. (2.11)

If ss is complex-valued with Re(s)>1\mathrm{Re}(s)>1, then the left-hand side of (2.11) is given ζ(s,a)\zeta(s,a) (2.1). It is argued in [Ber72, Eq. (5.2)] via an analytic continuation argument that, that ζ(s,a)\zeta(s,a) coincides with the right-hand side of (2.11) when (s)>1\Re(s)>-1 and a(0,1)a\in(0,1). As a consequence we find that for all x𝕋x\in\mathbb{T} and s(0,1)s\in(0,1)

g(x)=ζ(s,1x)+ζ(s,x).g(x)=\zeta(s,1-x)+\zeta(s,x).

2.2. Inverse Riesz transform

In Section 5, we will need to compute the fractional Laplacian of order 1s2\frac{1-s}{2} of test-functions ξ:𝕋\xi:\mathbb{T}\to\mathbb{R} with poor regularity. In the next lemma, we provide some useful identities.

Lemma 2.2 (Inversion of the Riesz transform).

Let ε>0\varepsilon>0 and ξ𝒞s+ε(𝕋,)\xi\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}). Let ψ𝒞ε(𝕋,)\psi\in\mathcal{C}^{\varepsilon}(\mathbb{T},\mathbb{R}) be given by

ψ=12cs(Δ)1s2ξandψ=0.\psi^{\prime}=-\frac{1}{2c_{s}}(-\Delta)^{\frac{1-s}{2}}\xi\quad\text{and}\quad\int\psi=0. (2.12)
  1. (1)

    We have

    ψ(x)=4sπtan(π2s)kξ(y)ξ|xyk|1ssgn(xyk)dy,for all x𝕋,\psi(x)=-\frac{4^{-s}}{\pi\tan(\frac{\pi}{2}s)}\int\sum_{k\in\mathbb{Z}}\frac{\xi(y)-\textstyle\int\xi}{|x-y-k|^{1-s}}\mathrm{sgn}(x-y-k)\mathrm{d}y,\quad\text{for all $x\in\mathbb{T}$}, (2.13)

    where (2.13) should be understood as a principle value sum.

  2. (2)

    Assume that ξ\xi satisfies Assumptions 1.1. We have

    12cs|ξ|H1s22=g′′(xy)(ψ(x)ψ(y))2dxdy2ξ(x)ψ(x)dx=ξ(x)ψ(x)dx.\frac{1}{2c_{s}}|\xi|_{H^{\frac{1-s}{2}}}^{2}=\iint g^{\prime\prime}(x-y)(\psi(x)-\psi(y))^{2}\mathrm{d}x\mathrm{d}y-2\int\xi^{\prime}(x)\psi(x)\mathrm{d}x=-\int\xi^{\prime}(x)\psi(x)\mathrm{d}x. (2.14)
  3. (3)

    Assume that ξ\xi satisfies Assumptions 1.1 and that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}). Let N(0,1]\ell_{N}\in(0,1]. Let ξ0:\xi_{0}:\mathbb{R}\to\mathbb{R} such that

    ξ0(x)={ξ(x)if |x|120if |x|>12.\xi_{0}(x)=\begin{cases}\xi(x)&\text{if $|x|\leq\frac{1}{2}$}\\ 0&\text{if $|x|>\frac{1}{2}$}.\end{cases} (2.15)

    Then

    |ξ(N1)|H1s2=Ns|ξ0|H1s22+O(N2|ξ0|L22).|\xi(\ell_{N}^{-1}\cdot)|_{H^{\frac{1-s}{2}}}=\ell_{N}^{s}|\xi_{0}|_{H^{\frac{1-s}{2}}}^{2}+O(\ell_{N}^{2}|\xi_{0}|_{L^{2}}^{2}). (2.16)
Proof.

Let ξ𝒞s+ε(𝕋,)\xi\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}) for some ε>0\varepsilon>0. Let ψ:𝕋\psi:\mathbb{T}\to\mathbb{R} solving (2.12). If ξ𝒞1s+ε(𝕋,)\xi\in\mathcal{C}^{1-s+\varepsilon}(\mathbb{T},\mathbb{R}), it is well-known, see for instance [Sti19, Th. 2], that for all x𝕋x\in\mathbb{T},

ψ(x)=cs2cskξ(x)ξ(y)|xyk|2sdy,\psi^{\prime}(x)=-\frac{c_{s}^{\prime}}{2c_{s}}\int\sum_{k\in\mathbb{Z}}\frac{\xi(x)-\xi(y)}{|x-y-k|^{2-s}}\mathrm{d}y, (2.17)

where

cs=21sΓ(1s2)|Γ(1s2)|π12=2s(1s)Γ(1s2)Γ(1+s2)π12.c_{s}^{\prime}=\frac{2^{1-s}\Gamma(1-\frac{s}{2})}{|\Gamma(-\frac{1-s}{2})|\pi^{\frac{1}{2}}}=\frac{2^{-s}(1-s)\Gamma(1-\frac{s}{2})}{\Gamma(\frac{1+s}{2})\pi^{\frac{1}{2}}}.

The above formula amounts to computing the fundamental solution of (Δ)1s2(-\Delta)^{-\frac{1-s}{2}}. This can be done by proceeding as in the proof of Lemma 2.1, starting from the identity

λα=1Γ(α)0(etλ1)dtt1+α,for all λ>0,\lambda^{\alpha}=\frac{1}{\Gamma(-\alpha)}\int_{0}^{\infty}(e^{-t\lambda}-1)\frac{\mathrm{d}t}{t^{1+\alpha}},\quad\text{for all $\lambda>0$},

valid for α>0\alpha>0.

Using Euler’s reflection lemma, we next compute

cscs=212s(1s)πΓ(1s2)Γ(s2)Γ(1s2)Γ(1+s2)=212s(1s)πsin(π(1s)2)sin(πs2)=(1s)212sπtan(π2s).\frac{c_{s}^{\prime}}{c_{s}}=\frac{2^{1-2s}(1-s)}{\pi}\frac{\Gamma(1-\frac{s}{2})\Gamma(\frac{s}{2})}{\Gamma(\frac{1-s}{2})\Gamma(\frac{1+s}{2})}=\frac{2^{1-2s}(1-s)}{\pi}\frac{\sin(\frac{\pi(1-s)}{2})}{\sin(\frac{\pi s}{2})}=\frac{(1-s)2^{1-2s}}{\pi\tan(\frac{\pi}{2}s)}.

Thus for any x𝕋x\in\mathbb{T}

ψ(x)=1s4sπtan(π2s)01kξ(x)ξ(y)|xyk|2sdy.\psi^{\prime}(x)=-\frac{1-s}{4^{s}\pi\tan(\frac{\pi}{2}s)}\int_{0}^{1}\sum_{k\in\mathbb{Z}}\frac{\xi(x)-\xi(y)}{|x-y-k|^{2-s}}\mathrm{d}y.

Integrating the last display with the condition that ψ=0\int\psi=0 yields

ψ(x)=4sπtan(π2s)01kξ(x)ξ(y)|xyk|1ssgn(xyk)dy=4sπtan(π2s)01kξ(y)|xyk|1ssgn(xyk)dy,\begin{split}\psi(x)&=\frac{4^{-s}}{\pi\tan(\frac{\pi}{2}s)}\int_{0}^{1}\sum_{k\in\mathbb{Z}}\frac{\xi(x)-\xi(y)}{|x-y-k|^{1-s}}\mathrm{sgn}(x-y-k)\mathrm{d}y\\ &=-\frac{4^{-s}}{\pi\tan(\frac{\pi}{2}s)}\int_{0}^{1}\sum_{k\in\mathbb{Z}}\frac{\xi(y)}{|x-y-k|^{1-s}}\mathrm{sgn}(x-y-k)\mathrm{d}y,\end{split} (2.18)

showing (2.13) when ξ𝒞1s+ε(𝕋,)\xi\in\mathcal{C}^{1-s+\varepsilon}(\mathbb{T},\mathbb{R}) for some ε>0\varepsilon>0. Note that (2.18) is a principle value sum, i.e should be understood as the limit

ψ(x)=limn=4sπtan(π2s)01k=nnξ(y)ξ|xyk|1ssgn(xyk)dy\psi(x)=\lim_{n\to\infty}=-\frac{4^{-s}}{\pi\tan(\frac{\pi}{2}s)}\int_{0}^{1}\sum_{k=-n}^{n}\frac{\xi(y)-\textstyle\int\xi}{|x-y-k|^{1-s}}\mathrm{sgn}(x-y-k)\mathrm{d}y

By density, we conclude that (2.13) in fact holds as soon as ξ𝒞s+ε(𝕋,)\xi\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}) for some ε>0\varepsilon>0.

Let ξ\xi satisfy Assumptions 1.1. Recall that ξ\xi is piecewise 𝒞2s\mathcal{C}^{2-s} and around a singularity aa of ξ\xi, ψ\psi^{\prime} grows at most in |xa|(1s2)+ε)|x-a|^{-(1-\frac{s}{2})+\varepsilon)} and ξ\xi in |xa|s2+ε|x-a|^{-\frac{s}{2}+\varepsilon}, proving that ξψ\xi\psi^{\prime} is in L1(𝕋,)L^{1}(\mathbb{T},\mathbb{R}) and that ξψ\int\xi^{\prime}\psi is well defined.

Equation (2.14) follows by integration by parts: we have

g′′(xy)(ψ(x)ψ(y))2dxdy=g(xy)ψ(x)2(ψ(x)ψ(y))dxdy=2g(xy)ψ(x)ψ(y)=ψ(2gψ)=ψξ=ψξ,\int g^{\prime\prime}(x-y)(\psi(x)-\psi(y))^{2}\mathrm{d}x\mathrm{d}y=-\int g^{\prime}(x-y)\psi^{\prime}(x)2(\psi(x)-\psi(y))\mathrm{d}x\mathrm{d}y\\ =2\int g(x-y)\psi^{\prime}(x)\psi^{\prime}(y)=\int\psi^{\prime}(2g*\psi^{\prime})=-\int\psi^{\prime}\xi=\int\psi\xi^{\prime},

proving (2.14).

Assume that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}). Let ξ0:\xi_{0}:\mathbb{R}\to\mathbb{R} such that

ξ0(x)={ξ(x)if |x|120if |x|>12.\xi_{0}(x)=\begin{cases}\xi(x)&\text{if $|x|\leq\frac{1}{2}$}\\ 0&\text{if $|x|>\frac{1}{2}$}.\end{cases}

We have

|ξ(N1)|H1s22=𝕋𝟙|x|<12Nξ(N1x)k𝕋ξ(N1x)ξ(N1y)|xy+k|2s𝟙|y|<12Ndydx=ξ0(N1x)kξ0(N1x)ξ0(N1y)|xy+k|2sdydx=Ns|ξ0|H1s22+kξ0(N1x)ξ0(N1x)ξ0(N1y)|xy+k|2sdydx=Ns|ξ0|H1s22+O(N2|ξ0(x)||ξ0(x)ξ0(y)|dxdy)=Ns|ξ0|H1s22+O(N2|ξ0|L22).\begin{split}|\xi(\ell_{N}^{-1}\cdot)|_{H^{\frac{1-s}{2}}}^{2}&=\int_{\mathbb{T}}\mathds{1}_{|x|<\frac{1}{2}\ell_{N}}\xi(\ell_{N}^{-1}x)\sum_{k\in\mathbb{Z}}\int_{\mathbb{T}}\frac{\xi(\ell_{N}^{-1}x)-\xi(\ell_{N}^{-1}y)}{|x-y+k|^{2-s}}\mathds{1}_{|y|<\frac{1}{2}\ell_{N}}\mathrm{d}y\mathrm{d}x\\ &=\int_{\mathbb{R}}\xi_{0}(\ell_{N}^{-1}x)\sum_{k\in\mathbb{Z}}\int_{\mathbb{R}}\frac{\xi_{0}(\ell_{N}^{-1}x)-\xi_{0}(\ell_{N}^{-1}y)}{|x-y+k|^{2-s}}\mathrm{d}y\mathrm{d}x\\ &=\ell_{N}^{s}|\xi_{0}|_{H^{\frac{1-s}{2}}}^{2}+\sum_{k\in\mathbb{Z}^{*}}\int_{\mathbb{R}}\xi_{0}(\ell_{N}^{-1}x)\frac{\xi_{0}(\ell_{N}^{-1}x)-\xi_{0}(\ell_{N}^{-1}y)}{|x-y+k|^{2-s}}\mathrm{d}y\mathrm{d}x\\ &=\ell_{N}^{s}|\xi_{0}|_{H^{\frac{1-s}{2}}}^{2}+O\Bigr{(}\ell_{N}^{2}\int|\xi_{0}(x)||\xi_{0}(x)-\xi_{0}(y)|\mathrm{d}x\mathrm{d}y\Bigr{)}\\ &=\ell_{N}^{s}|\xi_{0}|_{H^{\frac{1-s}{2}}}^{2}+O(\ell_{N}^{2}|\xi_{0}|_{L^{2}}^{2}).\end{split}

Next, we apply the pointwise formula (2.13) to indicator and inverse power functions.

Lemma 2.3 (Explicit formulas).

Let ζ(s,a)\zeta(s,a) be the Hurwitz zeta function. Let ξ=𝟙(a,a)2a\xi=\mathds{1}_{(-a,a)}-2a for some 0<a<120<a<\frac{1}{2} and ψ\psi be given by (2.13). We have

ψ(x)=14sπstan(π2s)(ζ(s,x+a)ζ(s,xa)),for all x𝕋.\psi(x)=\frac{1}{4^{s}\pi s\tan(\frac{\pi}{2}s)}(\zeta(-s,x+a)-\zeta(-s,x-a)),\quad\text{for all $x\in\mathbb{T}$}. (2.19)

Moreover

12cs|ξ|H1s22=cotan(π2s)β4s1πsζ(s,2a).\frac{1}{2c_{s}}{|\xi|_{H^{\frac{1-s}{2}}}^{2}}=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta 4^{s-1}\pi s}\zeta(-s,2a). (2.20)

Let α(0,s2)\alpha\in(0,\frac{s}{2}) and ξ:=ζ(α,)\xi:=\zeta(-\alpha,\cdot). We have

12cs|ξ|H1s22=cα2csζ(1+s2α).\frac{1}{2c_{s}}|\xi|_{H^{\frac{1-s}{2}}}^{2}=\frac{c_{\alpha}^{2}}{c_{s}}\zeta(1+s-2\alpha).
Proof.

Let ξ:=𝟙(a,a)\xi:=\mathds{1}_{(-a,a)} for a(0,12)a\in(0,\frac{1}{2}). Formula (2.19) follows from (2.13). Moreover in view of (2.14) we have

12cs|ξ|H1s22=1βξψ=1β(ψ(a)ψ(a))=cotan(π2s)β4sπs2ζ(s,2a),\frac{1}{2c_{s}}|\xi|_{H^{\frac{1-s}{2}}}^{2}=\frac{1}{\beta}\int\xi\psi^{\prime}=\frac{1}{\beta}(\psi(a)-\psi(-a))=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta 4^{s}\pi s}2\zeta(-s,2a),

since ζ(s,0)=0\zeta(-s,0)=0.

Let α(0,s2)\alpha\in(0,\frac{s}{2}) and ξ:=ζ(α,)\xi:=\zeta(-\alpha,\cdot). Recall that for each kk\in\mathbb{Z},

ξ^(k)=cα𝟙k0|k|1α\hat{\xi}(k)=c_{\alpha}\frac{\mathds{1}_{k\neq 0}}{|k|^{1-\alpha}}

Thus

12cs|ξ|H1s22=cα2k{0}1k1+s2α=2cα2ζ(1+s2α).\frac{1}{2c_{s}}|\xi|_{H^{\frac{1-s}{2}}}^{2}=c_{\alpha}^{2}\sum_{k\in\mathbb{Z}\setminus\{0\}}\frac{1}{k^{1+s-2\alpha}}=2c_{\alpha}^{2}\zeta(1+s-2\alpha).

Lemma 2.4 (Decay).

Let ξ𝒞s+ε(𝕋,)\xi\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}). Assume that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}). Then ψ:=(Δ)1s2ξ\psi:=(-\Delta)^{\frac{1-s}{2}}\xi is 𝒞\mathcal{C}^{\infty} on 𝕋[a,a]\mathbb{T}\setminus[-a,a] and for each k1k\geq 1, there exists a constant Ck>0C_{k}>0 such that for all x𝕋[2a,2a]x\in\mathbb{T}\setminus[-2a,2a],

|ψ(k)|(x)Ck|ξ|L1|x|1s+k.|\psi^{(k)}|(x)\leq\frac{C_{k}|\xi|_{L^{1}}}{|x|^{1-s+k}}. (2.21)
Proof.

Since ξL1(𝕋,)\xi\in L^{1}(\mathbb{T},\mathbb{R}), one may differentiate (2.13) under the integral sign. Using the fact that (ξξ)=0\textstyle\int(\xi-\int\xi)=0, we obtain (2.21). ∎

2.3. Regularization estimates

Let K:𝕋K:\mathbb{T}\to\mathbb{R} be a 𝒞3\mathcal{C}^{3} non-negative function supported on (12,12)(-\frac{1}{2},\frac{1}{2}). For all (0,1)\ell\in(0,1) define

K:=1K(1).K_{\ell}:=\ell^{-1}K(\ell^{-1}\cdot). (2.22)
Lemma 2.5 (Regularization).

Let ξ\xi satisfy Assumption 1.1. Let ψ\psi be the integral of mean 0 of (Δ)1s2ξ(-\Delta)^{\frac{1-s}{2}}\xi. Let a1,,apa_{1},\ldots,a_{p} be the singularities of ξ\xi. Assume that there exists a(0,12]a\in(0,\frac{1}{2}] such that a1,,ap(a,a)a_{1},\ldots,a_{p}\in(-a,a) and α(s1,s2)\alpha\in(s-1,\frac{s}{2}) such that for all x𝕋x\in\mathbb{T},

|ψ′′(x)|k=1pC|xak|α+2s𝟙|x|<2a+1(a+|x|)3s.|\psi^{\prime\prime}(x)|\leq\sum_{k=1}^{p}\frac{C}{|x-a_{k}|^{\alpha+2-s}}\mathds{1}_{|x|<2a}+\frac{1}{(a+|x|)^{3-s}}.

Let (0,12)\ell\in(0,\frac{1}{2}) and KK_{\ell} be as in (2.22).

We have

|ξξK|L22\displaystyle|\xi-\xi*K_{\ell}|^{2}_{L^{2}} C(2+12α)\displaystyle\leq C(\ell^{2}+\ell^{1-2\alpha}) (2.23)
|ψK|L22\displaystyle|\psi^{\prime}*K_{\ell}|_{L^{2}}^{2} C(1+(1+2α2s))\displaystyle\leq C(1+\ell^{-(1+2\alpha-2s)}) (2.24)
|ψ′′K|(x)\displaystyle|\psi^{\prime\prime}*K_{\ell}|(x) C(k=1p1(|xak|)2s+α𝟙|xak|<2a+1(a+|x|)3s)\displaystyle\leq C\Bigr{(}\sum_{k=1}^{p}\frac{1}{(|x-a_{k}|\vee\ell)^{2-s+\alpha}}\mathds{1}_{|x-a_{k}|<2a}+\frac{1}{(a+|x|)^{3-s}}\Bigr{)} (2.25)
|ξK|H1s22\displaystyle|\xi*K_{\ell}|_{H^{\frac{1-s}{2}}}^{2} =|ξ|H1s22+O(+s2α).\displaystyle=|\xi|_{H^{\frac{1-s}{2}}}^{2}+O(\ell+\ell^{s-2\alpha}). (2.26)
Proof.

We decompose ξ\xi into smooth and non-smooth parts as

ξ=k=1pξk+f\xi=\sum_{k=1}^{p}\xi_{k}+f

where the function f𝒞2s(𝕋,)f\in\mathcal{C}^{2-s}(\mathbb{T},\mathbb{R}) is smooth and for each k=1,,pk=1,\ldots,p, the non-smooth function ξk𝒞s+ε(𝕋,)\xi_{k}\in\mathcal{C}^{-s+\varepsilon}(\mathbb{T},\mathbb{R}) satisfies

|ψk′′|(x)C(1|xak|α+2s𝟙|x|<2a+1(a+|x|)3s),|\psi_{k}^{\prime\prime}|(x)\leq C\Bigr{(}\frac{1}{|x-a_{k}|^{\alpha+2-s}}\mathds{1}_{|x|<2a}+\frac{1}{(a+|x|)^{3-s}}\Bigr{)},

where ψk\psi_{k} is the integral of mean 0 of (Δ)1s2ξk(-\Delta)^{\frac{1-s}{2}}\xi_{k}. Also denote ϕ\phi the integral of mean 0 if (Δ)1s2f(-\Delta)^{\frac{1-s}{2}}f.

Let k{1,,p}k\in\{1,\ldots,p\}. For all x𝕋x\in\mathbb{T} such that |xak|>2|x-a_{k}|>2\ell,

|ξkξkK|(x)C|xak|α+1K(y)|y|dyC|xak|α+1.|\xi_{k}-\xi_{k}*K_{\ell}|(x)\leq\frac{C}{|x-a_{k}|^{\alpha+1}}\int K_{\ell}(y)|y|\mathrm{d}y\leq\frac{C\ell}{|x-a_{k}|^{\alpha+1}}. (2.27)

If |xak|2|x-a_{k}|\leq 2\ell then we write

|ξkξkK|(x)|ξkK|(x)+|ξk|(x)C|xak|α.|\xi_{k}-\xi_{k}*K_{\ell}|(x)\leq|\xi_{k}*K_{\ell}|(x)+|\xi_{k}|(x)\leq\frac{C}{|x-a_{k}|^{\alpha}}. (2.28)

Since 2α<12\alpha<1, we deduce from (2.27) and (2.28) that

|ξkξkK|L22C12α.|\xi_{k}-\xi_{k}*K_{\ell}|^{2}_{L^{2}}\leq C\ell^{1-2\alpha}.

Moreover since s<1s<1, f𝒞1(𝕋,)f\in\mathcal{C}^{1}(\mathbb{T},\mathbb{R}) and

|ffK|L22C2|f|2,|f-f*K_{\ell}|^{2}_{L^{2}}\leq C\ell^{2}|f^{\prime}|_{\infty}^{2},

as claimed in (2.23).

By assumption, for each k=1,,pk=1,\ldots,p

|ψk|(x)C|xak|α+1sfor all x𝕋{ak}|\psi_{k}^{\prime}|(x)\leq\frac{C}{|x-a_{k}|^{\alpha+1-s}}\quad\text{for all $x\in\mathbb{T}\setminus\{a_{k}\}$}

and

|ϕ|(x)Cfor all x𝕋.|\phi^{\prime}|(x)\leq C\quad\text{for all $x\in\mathbb{T}$}.

Fix k{1,,p}k\in\{1,\ldots,p\}. For all x(ak2,ak+2)cx\in(a_{k}-2\ell,a_{k}+2\ell)^{c},

|ψkK|(x)C|xak|1+αs.|\psi_{k}^{\prime}*K_{\ell}|(x)\leq\frac{C}{|x-a_{k}|^{1+\alpha-s}}.

Now if x(ak2,ak+2)x\in(a_{k}-2\ell,a_{k}+2\ell) we get by integration by parts that

|ψkK(x)|=|ψk(xy)K(y)dy|C2|xaky|sα|K(1y)|dyC21+sα.|\psi_{k}^{\prime}*K_{\ell}(x)|=\Bigr{|}\int\psi_{k}(x-y)K_{\ell}^{\prime}(y)\mathrm{d}y\Bigr{|}\leq C\ell^{-2}\int_{-\ell}^{\ell}|x-a_{k}-y|^{s-\alpha}|K^{\prime}(\ell^{-1}y)|\mathrm{d}y\\ \leq C^{\prime}\ell^{-2}\ell^{1+s-\alpha}.

Combining the two last displays shows that

|ψkK|(x)C(|xak|)1+sα,|\psi_{k}^{\prime}*K_{\ell}|(x)\leq\frac{C}{(|x-a_{k}|\vee\ell)^{1+s-\alpha}},

which gives

|ψkK|L22C(1+(1+2α2s)),|\psi_{k}^{\prime}*K_{\ell}|_{L^{2}}^{2}\leq C(1+\ell^{-(1+2\alpha-2s)}),

which proves (2.24). Arguing as above for the map ψ′′\psi^{\prime\prime} gives the bound (2.25).

Recall that by (2.14),

|ξ|H1s22=1csξψ.|\xi|_{H^{\frac{1-s}{2}}}^{2}=\frac{1}{c_{s}}\int\xi\psi^{\prime}.

One can write

cs(|ξ|H1s22|ξK|H1s22)=ξ(ψψK)+(ξKξ)ψK.c_{s}(|\xi|_{H^{\frac{1-s}{2}}}^{2}-|\xi*K_{\ell}|_{H^{\frac{1-s}{2}}}^{2})=\int\xi(\psi^{\prime}-\psi^{\prime}*K_{\ell})+\int(\xi-K_{\ell}*\xi)\psi^{\prime}*K_{\ell}. (2.29)

For each k=1,,pk=1,\ldots,p, using α<s2\alpha<\frac{s}{2} and

|ψkψkK|(x)C(1|xak|1+αs𝟙|xak|<2+|xak|2+αs𝟙|xak|2),|\psi_{k}^{\prime}-\psi_{k}^{\prime}*K_{\ell}|(x)\leq C\Bigr{(}\frac{1}{|x-a_{k}|^{1+\alpha-s}}\mathds{1}_{|x-a_{k}|<2\ell}+\frac{\ell}{|x-a_{k}|^{2+\alpha-s}}\mathds{1}_{|x-a_{k}|\geq 2\ell}\Bigr{)},

we get

|ξ(ψkψkK)|Cs2α.\Bigr{|}\int\xi(\psi_{k}^{\prime}-\psi_{k}^{\prime}*K_{\ell})\Bigr{|}\leq C\ell^{s-2\alpha}.

Furthermore

|ξ(ϕϕK)|C.\Bigr{|}\int\xi(\phi^{\prime}-\phi^{\prime}*K_{\ell})\Bigr{|}\leq C\ell.

By similar arguments one can check that for each k=1,,pk=1,\ldots,p,

|(ξKξ)ψkK|Cs2α\Bigr{|}\int(\xi-K_{\ell}*\xi)\psi_{k}^{\prime}*K_{\ell}\Bigr{|}\leq C\ell^{s-2\alpha}

and

|(ξKξ)ϕK|C.\Bigr{|}\int(\xi-K_{\ell}*\xi)\phi^{\prime}*K_{\ell}\Bigr{|}\leq C\ell.

Inserting the four last displays in (2.29) yields (2.26). ∎

3. The Helffer-Sjöstrand equation

In this section we review some basic properties of the Helffer-Sjöstrand equation. We first state some existence and uniqueness results, then derive a known comparison principle that we adapt to the circular setting and finally record various concentration estimate.

3.1. Basic properties

In this subsection we introduce the H.-S. equation and state some standard existence and uniqueness results following partly [AW19]. Let μ\mu be a probability measure on DND_{N} in the form

dμ(XN)=eH(XN)𝟙DN(XN)dXN,\mathrm{d}\mu(X_{N})=e^{-H(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N},

with H:DNH:D_{N}\to\mathbb{R} measurable. We make the following assumptions on HH:

Assumptions 3.1.

Assume that H:DNH:D_{N}\to\mathbb{R} is in the form

H:XNijχij(|xixj|),H:X_{N}\mapsto\sum_{i\neq j}\chi_{ij}(|x_{i}-x_{j}|), (3.1)

for a family of functions χij:+\chi_{ij}:\mathbb{R}^{+*}\to\mathbb{R} satisfying

χij𝒞(+,),χij′′c>0.\chi_{ij}\in\mathcal{C}^{\infty}(\mathbb{R}^{+*},\mathbb{R}),\quad\chi_{ij}^{\prime\prime}\geq c>0.

Let F:DNF:D_{N}\to\mathbb{R} be a smooth enough function. We seek to rewrite the variance of FF under μ\mu in a tractable way. Let us recall the integration by parts formula for μ\mu. Define on 𝒞c(DN)\mathcal{C}_{c}^{\infty}(D_{N}) the Langevin operator

μ=HΔ,\mathcal{L}^{\mu}=\nabla H\cdot\nabla-\Delta,

with \nabla and Δ\Delta the gradient and Laplace operators of 𝕋N\mathbb{T}^{N}. By integration by parts, for any functions ϕ,ψ𝒞(DN)\phi,\psi\in\mathcal{C}^{\infty}(D_{N}) such that ϕn=0\nabla\phi\cdot\vec{n}=0 a.e on DN\partial D_{N}, we have

𝔼μ[ψμϕ]=𝔼μ[ψϕ].\mathbb{E}_{\mu}[\psi\mathcal{L}^{\mu}\phi]=\mathbb{E}_{\mu}[\nabla\psi\cdot\nabla\phi]. (3.2)

Note that when limx0χ(x)=\lim_{x\to 0}\chi(x)=\infty, the condition ϕn=0\nabla\phi\cdot\vec{n}=0 is not necessary. Let us now assume that the Poisson equation

{μϕ=F𝔼μ[F]on DNϕn=0a.e on DN\begin{cases}\mathcal{L}^{\mu}\phi=F-\mathbb{E}_{\mu}[F]&\text{on }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{a.e on }\partial D_{N}\end{cases} (3.3)

admits a solution in an appropriate sense. Then, by (3.2), one may rewrite the variance of FF as

Varμ[F]=𝔼μ[Fϕ].\mathrm{Var}_{\mu}[F]=\mathbb{E}_{\mu}[\nabla F\cdot\nabla\phi].

The above formula is called the Helffer-Sjöstrand representation formula. Formally differentiating (3.3) yields

μϕ=A1μϕ,\nabla\mathcal{L}^{\mu}\phi=A_{1}^{\mu}\nabla\phi,

where A1A_{1} is the so-called Helffer-Sjöstrand operator defined by

A1μ=2H+μIN.A_{1}^{\mu}=\nabla^{2}H+\mathcal{L}^{\mu}\otimes I_{N}.

The solution of (3.3) therefore formally satisfies

{A1μϕ=Fon DNϕn=0a.e on DN.\begin{cases}A_{1}^{\mu}\nabla\phi=\nabla F&\text{on }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{a.e on }\partial D_{N}.\\ \end{cases} (3.4)

The above is called the Helffer-Sjöstrand equation [AW19].

Let us now study the well-posedness of (3.4). Define the norm

FH1(μ)=𝔼μ[F2]12+𝔼μ[|F|2]12.\|F\|_{H^{1}(\mu)}=\mathbb{E}_{\mu}[F^{2}]^{\frac{1}{2}}+\mathbb{E}_{\mu}[|\nabla F|^{2}]^{\frac{1}{2}}.

Let H1(μ)H^{1}(\mu) be the completion of 𝒞c(DN)\mathcal{C}^{\infty}_{c}(D_{N}) with respect to the norm H1(μ)\|\cdot\|_{H^{1}(\mu)}. Also define the norm

FH1(μ)=sup{|𝔼μ[FG]|:GH1(μ),GH1(μ)1}\|F\|_{H^{-1}(\mu)}=\sup\{|\mathbb{E}_{\mu}[FG]|:G\in H^{1}(\mu),\|G\|_{H^{1}(\mu)}\leq 1\}

and let H1(μ)H^{-1}(\mu) be the dual of H1(μ)H^{1}(\mu), defined as the completion of 𝒞c(DN)\mathcal{C}_{c}^{\infty}(D_{N}) with respect to the norm H1(μ)\|\cdot\|_{H^{-1}(\mu)}.

Since the density of μ\mu with respect to the Lebesgue measure on DND_{N} is not bounded from below, the existence of a solution of (3.4) is not straightforward. However one can easily prove existence and uniqueness when FF is a function of the gaps. Define the map

GapN:XNDNN(x2x1,x3x1,,xNx1)N\mathrm{Gap}_{N}:X_{N}\in D_{N}\mapsto N(x_{2}-x_{1},x_{3}-x_{1},\ldots,x_{N}-x_{1})\in\mathbb{R}^{N} (3.5)

and the push-forward of μ\mu by the map GapN\mathrm{Gap}_{N}

μ=GapN#μ.\mu^{\prime}=\mathrm{Gap}_{N}\#\mu. (3.6)
Proposition 3.1 (Existence and representation).

Let μ\mu satisfy Assumptions 3.1. Assume that FF is in the form F=GGapNF=G\circ\mathrm{Gap}_{N}, GH1(μ)G\in H^{1}(\mu^{\prime}) or that χ\chi is bounded. Then there exists a unique ϕL2({1,,N},H1(μ))\nabla\phi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)) such that

{A1μϕ=F in DNϕn=0a.e on DN,\begin{cases}A_{1}^{\mu}\nabla\phi=\nabla F&\text{ in }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{a.e on }\partial D_{N},\end{cases} (3.7)

with the first equality being, for each coordinate, an equality of elements of H1(μ)H^{-1}(\mu).

Moreover the solution of (3.7) is the unique minimizer of the functional

ϕ𝔼μ[ϕ2Hϕ+|2ϕ|22Fϕ],\nabla\phi\mapsto\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla^{2}H\nabla\phi+|\nabla^{2}\phi|^{2}-2\nabla F\cdot\nabla\phi], (3.8)

on maps ϕL2({1,,N},H1(μ))\nabla\phi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)) such that ϕn=0\nabla\phi\cdot\vec{n}=0 on DN\partial D_{N}. The variance of FF may be represented as

Varμ[F]=𝔼μ[ϕF]\mathrm{Var}_{\mu}[F]=\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla F] (3.9)

and the covariance between FF and any function GH1(μ)G\in H^{1}(\mu) as

Covμ[F,G]=𝔼μ[ϕG].\mathrm{Cov}_{\mu}[F,G]=\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla G].

The identity (3.9) is called the Helffer-Sjöstrand formula. The proof of Proposition 3.1 is postponed to Appendix A.

Remark 3.2 (Remark on the boundary condition).

In the case where

limx0χij(x)=+\lim_{x\to 0}\chi_{ij}(x)=+\infty

for each i,ji,j, the boundary terms in the integration by parts vanish, making the analysis much simpler. In the rest of the paper, we will only consider the case where the χijs\chi_{ij}^{\prime}s are bounded from above in the proof of Lemma 3.9.

Note that for a map ϕH1(μ)\phi\in H^{1}(\mu), ϕn=0\nabla\phi\cdot\vec{n}=0 a.e on DN\partial D_{N} if and only if iϕ=jϕ\partial_{i}\phi=\partial_{j}\phi a.e whenever i=ji=j, for each i,j{1,,N}i,j\in\{1,\ldots,N\}.

When F\nabla F is replaced by a non-gradient vector-field vv, the solution is in general non unique. In order to have uniqueness, we need to also assume that i=1Nvi=0\sum_{i=1}^{N}v_{i}=0 and that each coordinate viv_{i} is a function of the gaps.

Proposition 3.2 (Well-posedness for non-gradient vector-fields).

Let μ\mu satisfy Assumption 3.1. Assume also that for each i,j=1,,Ni,j=1,\ldots,N,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let vL2({1,,N},H1(μ))v\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu)) such that v(e1++eN)=0v\cdot(e_{1}+\ldots+e_{N})=0 and for each i{1,,N}i\in\{1,\ldots,N\}, vi=wiGapNv_{i}=w_{i}\circ\mathrm{Gap}_{N} for some wiH1(μ)w_{i}\in H^{-1}(\mu^{\prime}). There exists a unique ψL2({1,,N},H1(μ))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)) such that

{A1μψ=von DNψ(e1++eN)=0on DN.\begin{cases}A_{1}^{\mu}\psi=v&\text{on }D_{N}\\ \psi\cdot(e_{1}+\ldots+e_{N})=0&\text{on }D_{N}.\end{cases} (3.10)

Moreover the solution of (3.10) is also the unique minimizer of

ψ𝔼μ[ψ2Hψ+|Dψ|22vψ],\psi\mapsto\mathbb{E}_{\mu}[\psi\cdot\nabla^{2}H\psi+|D\psi|^{2}-2v\cdot\psi],

over maps ψL2({1,,N},H1(μ))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)).

The proof of Proposition 3.2 is postponed to Appendix A. When vv satisfies the assumptions of Proposition 3.2, we unambiguously denote ψ=(A1μ)1v\psi=(A_{1}^{\mu})^{-1}v as the solution of (3.10).

Lemma 3.3.

Let μ\mu satisfy Assumptions 3.1. Assume also that for each i,j=1,,Ni,j=1,\ldots,N,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let v,wL2({1,,N},H1(μ))v,w\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu)) satisfy the assumptions of Proposition 3.2. We have

𝔼μ[(v+w)(A1μ)1(w+w)]2(𝔼μ[v(A1μ)1v]+𝔼μ[w(A1μ)1w]).\mathbb{E}_{\mu}[(v+w)\cdot(A_{1}^{\mu})^{-1}(w+w)]\leq 2\Bigr{(}\mathbb{E}_{\mu}[v\cdot(A_{1}^{\mu})^{-1}v]+\mathbb{E}_{\mu}[w\cdot(A_{1}^{\mu})^{-1}w]\Bigr{)}. (3.11)

Let F=GGapNF=G\circ\mathrm{Gap}_{N} with GL2({1,,N},H1(μ))\nabla G\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{\prime})). Then the solution of (3.10) is the solution ϕ\nabla\phi of (3.7).

Proof.

Since vwv-w satisfies (e1++eN)(vw)=0(e_{1}+\ldots+e_{N})\cdot(v-w)=0, one can define (A1μ)1(vw)(A_{1}^{\mu})^{-1}(v-w). Moreover note that by integration by parts

𝔼μ[(vw)(A1μ)1(vw)]0.\mathbb{E}_{\mu}[(v-w)\cdot(A_{1}^{\mu})^{-1}(v-w)]\geq 0.

By linearity this implies (3.11). The second part of the statement is straightforward. Indeed when F=GGapNF=G\circ\mathrm{Gap}_{N} with GL2({1,,N},H1(μ))\nabla G\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{\prime})), then F\nabla F satisfies the assumptions of Proposition 3.2 and we conclude by uniqueness. ∎

3.2. Monotonicity and consequences

We now state some monotonicity results related to the FKG inequality. Recall that a measure ρ\rho on N\mathbb{R}^{N} is said to satisfy the FKG inequality if for any functions F:NF:\mathbb{R}^{N}\to\mathbb{R}, G:NG:\mathbb{R}^{N}\to\mathbb{R} in L2L^{2} which are increasing in each variable, we have

Covρ[F,G]0.\mathrm{Cov}_{\rho}[F,G]\geq 0.

A standard condition for ρ\rho to satisfy the FKG inequality [BÉ85] is that ρ\rho can be written as

dρ(XN)=eH(XN)dXN\mathrm{d}\rho(X_{N})=e^{-H(X_{N})}\mathrm{d}X_{N}

with H:NH:\mathbb{R}^{N}\to\mathbb{R} smooth enough satisfying

ijH0for each 1ijN.\partial_{ij}H\leq 0\quad\text{for each $1\leq i\neq j\leq N$}. (3.12)

In fact, the FKG property can be reformulated as a maximum principle for (3.7).

We prove a stronger maximum principle allowing to compare the solutions of (3.10) for general source vector-fields. This requires, in addition to (3.12), some strict convexity property on HH. Note that if μ\mu satisfies Assumptions 3.1, then μ\mu is not strictly log-concave. To make our argument work, we thus need to fix a point.

Fix xDNx\in D_{N} and let

DNx:=DN{x1=x},μx:=μ(x1=x).D_{N}^{x}:=D_{N}\cap\{x_{1}=x\},\quad\mu^{x}:=\mu(\cdot\mid x_{1}=x). (3.13)

On L2({1,,N},H1(μx))L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})) define

A1μx:=2H+μxIN.A_{1}^{\mu_{x}}:=\nabla^{2}H+\mathcal{L}^{\mu_{x}}\otimes I_{N}.
Proposition 3.4 (Existence with a fixed point).

Let μ\mu satisfy Assumptions 3.1. Assume also that for each i,j=1,,Ni,j=1,\ldots,N,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let x𝕋x\in\mathbb{T} and DNxD_{N}^{x}, μx\mu^{x} as in (3.13). Let vL2({1,,N},H1(μx))v\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{x})).

There exists a unique ψL2({1,,N},H1(μx))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})) solution of

{A1μxψ=v+λe1on DNxψ1=0on DNxψ=0on DNx.\left\{\begin{array}[]{ll}A_{1}^{\mu_{x}}\psi=v+\lambda e_{1}&\text{on }D_{N}^{x}\\ \psi_{1}=0&\text{on }D_{N}^{x}\\ {\psi}=0&\text{on }\partial D_{N}^{x}.\end{array}\right. (3.14)

If FH1(μx)F\in H^{1}(\mu^{x}), then the solution of (3.14) is in the form ψ=ϕL2({1,,N},H1(μx))\psi=\nabla\phi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})). Moreover the variance of FF under μx\mu^{x} may be represented as

Varμx[F]=𝔼μx[ϕF].\mathrm{Var}_{\mu^{x}}[F]=\mathbb{E}_{\mu^{x}}[\nabla\phi\cdot\nabla F].

In the sequel given vL2({1,,N},H1(μx))v\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{x})), we denote (A1μx)1v({A}_{1}^{\mu_{x}})^{-1}v the solution of (3.14).

Remark 3.3.

The coefficient λ\lambda in (3.14) is a Lagrange multiplier associated to the constraint x1=0x_{1}=0.

The proof of Proposition 3.4 is entirely similar to the proof of Proposition 3.2.

We can now state the maximum principle for (3.14), derived for instance in [Car74, HS94].

Lemma 3.5 (Monotonicity).

Let μ\mu satisfy Assumptions 3.1. Assume also that for each i,j=1,,Ni,j=1,\ldots,N,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let x𝕋x\in\mathbb{T} and vL2({1,,N},H1(μx))v\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{x})).

Let ψL2({1,,N},H1(μx))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})) be the solution of (3.14). Assume that

vi0a.e on DNx,for each i{1,,N}.v_{i}\geq 0\quad\text{a.e on }D_{N}^{x},\quad\text{for each $i\in\{1,\ldots,N\}$}.

Then

ψi0a.e on DNx,for each i{1,,N}.\psi_{i}\geq 0\quad\text{a.e on }D_{N}^{x},\quad\text{for each $i\in\{1,\ldots,N\}$}.
Proof.

Let ψL2({1,,N},H1(μx))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})) be the solution of (3.14). Let us prove that for each i{1,,N}i\in\{1,\ldots,N\}, ψi0\psi_{i}\geq 0 a.e on DNxD_{N}^{x}. Let ψ+\psi^{+} and ψ\psi^{-} be the positive and negative parts of ψ\psi. Taking the scalar product of the equation A1xψ=vA_{1}^{x}\psi=v with ψ\psi^{-} gives

ψ2Hψ+ψxψ=ψv0.\psi^{-}\cdot\nabla^{2}H\psi+\psi^{-}\cdot\mathcal{L}^{x}\psi=\psi^{-}\cdot v\geq 0.

By integration by parts under μx\mu^{x}, since limx0χij(x)=+\lim_{x\to 0}\chi_{ij}(x)=+\infty for each i,ji,j, one can observe that

𝔼μx[ψ2Hψ+i=1Nψiψi]0\mathbb{E}_{\mu^{x}}\Bigr{[}\psi^{-}\cdot\nabla^{2}H\psi+\sum_{i=1}^{N}\nabla\psi_{i}^{-}\cdot\nabla\psi_{i}\Bigr{]}\geq 0

since the boundary term in the above integration by parts vanishes. Note that ψiψi=|ψi|2\nabla\psi_{i}^{-}\cdot\nabla\psi_{i}=-|\nabla\psi_{i}^{-}|^{2} and

ψ2Hψ+=2i<jχ′′(xjxi)(ψiψi+ψiψj+)=2i<jχ′′(xjxi)ψiψj+0,\psi^{-}\cdot\nabla^{2}H\psi^{+}=2\sum_{i<j}\chi^{\prime\prime}(x_{j}-x_{i})(\psi_{i}^{-}\psi_{i}^{+}-\psi_{i}^{-}\psi_{j}^{+})=-2\sum_{i<j}\chi^{\prime\prime}(x_{j}-x_{i})\psi_{i}^{-}\psi_{j}^{+}\leq 0, (3.15)

since χ′′0\chi^{\prime\prime}\geq 0. One deduces that

𝔼μx[ψ2Hxψ+|Dψ|2]0.\mathbb{E}_{\mu^{x}}[\psi^{-}\cdot\nabla^{2}H^{x}\psi^{-}+|D\psi^{-}|^{2}]\leq 0.

Therefore

𝔼μx[ψ2Hψ]0.\mathbb{E}_{\mu^{x}}[\psi^{-}\cdot\nabla^{2}H\psi^{-}]\leq 0.

By Assumptions 3.1, there exists c>0c>0 such that

𝔼μx[ψ2Hψ]c𝔼μx[i=1N(ψi+1ψi)2].\mathbb{E}_{\mu^{x}}[\psi^{-}\cdot\nabla^{2}H\psi^{-}]\geq c\mathbb{E}_{\mu^{x}}\Bigr{[}\sum_{i=1}^{N}(\psi_{i+1}^{-}-\psi_{i}^{-})^{2}\Bigr{]}.

This shows that ψi=ψ1\psi_{i}^{-}=\psi_{1}^{-} for each i=1,,Ni=1,\ldots,N. Since ψ1=0\psi_{1}^{-}=0 we get ψ=0\psi^{-}=0 which concludes the proof. ∎

In view of Lemma 3.5, we compare the variances of two functions by comparison of their gradients. We derive the following new observation:

Lemma 3.6 (Energy comparison).

Let μ\mu satisfy Assumptions 3.1. Assume also that for each i,j=1,,Ni,j=1,\ldots,N,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let x𝕋x\in\mathbb{T}, DNxD_{N}^{x}, μx\mu^{x} be as in (3.13). Let v,wL2({1,,N},H1(μx))v,w\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{x})). Assume that for each i{1,,N}i\in\{1,\ldots,N\},

|vi|wi,a.e on DNx.|v_{i}|\leq w_{i},\quad\text{a.e on }D_{N}^{x}. (3.16)

Then

𝔼μx[v(A1μx)1v]𝔼μx[w(A1μx)1w].\mathbb{E}_{\mu^{x}}[v\cdot(A_{1}^{\mu^{x}})^{-1}v]\leq\mathbb{E}_{\mu^{x}}[w\cdot(A_{1}^{\mu^{x}})^{-1}w].

In particular if F,GH1(μ)F,G\in H^{1}(\mu) satisfy for each i{1,,N}i\in\{1,\ldots,N\},

|iF|iG,a.e on DNx,|\partial_{i}F|\leq\partial_{i}G,\quad\text{a.e on }D_{N}^{x},

then

Varμx[F]Varμx[G].\mathrm{Var}_{\mu^{x}}[F]\leq\mathrm{Var}_{\mu^{x}}[G].
Proof.

For x=(x1,,xN)Nx=(x_{1},\ldots,x_{N})\in\mathbb{R}^{N}, we use the notation x0x\geq 0 when for each i{1,,N}i\in\{1,\ldots,N\}, xi0x_{i}\geq 0.

Let v,wL2({1,,N},H1(μx))v,w\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{x})) as in the statement of Lemma 3.6. Let v+v^{+} and vv^{-} be the positive and negative parts of vv. Using the fact that A1μxA_{1}^{\mu_{x}} is self adjoint on L2({1,,N},H1(μx))L^{2}(\{1,\ldots,N\},H^{1}(\mu^{x})), one finds that

𝔼μx[w(A1μx)1w]𝔼μx[v(A1μx)1v]=𝔼μx[(v+w)(A1μx)1(wv)].\mathbb{E}_{\mu^{x}}[w\cdot(A_{1}^{\mu^{x}})^{-1}w]-\mathbb{E}_{\mu^{x}}[v\cdot(A_{1}^{\mu^{x}})^{-1}v]=\mathbb{E}_{\mu^{x}}[(v+w)\cdot(A_{1}^{\mu^{x}})^{-1}(w-v)].

Note that since wv0w-v\geq 0, by Lemma 3.5, (A1μx)1(wv)0(A_{1}^{\mu^{x}})^{-1}(w-v)\geq 0 and that w+v0w+v\geq 0, one gets

𝔼μx[(v+w)(A1μx)1(wv)]0,\mathbb{E}_{\mu^{x}}[(v+w)\cdot(A_{1}^{\mu^{x}})^{-1}(w-v)]\geq 0,

which gives the desired result. The second part of statement is straightforward. ∎

Now, if vv and ww are gradients, then much less is required on the measure μ\mu, as shown in the following lemma:

Lemma 3.7.

Let μ\mu be a probability measure on DND_{N} in the form dμ=eHdXN\mathrm{d}\mu=e^{-H}\mathrm{d}X_{N} with H:DNH:D_{N}\to\mathbb{R} in 𝒞2\mathcal{C}^{2} such that

limd(x,DN)0H(x)=+.\lim_{d(x,\partial D_{N})\to 0}H(x)=+\infty.

Assume that

ijH0for each ij.\partial_{ij}H\leq 0\quad\text{for each $i\neq j$}.

Let F,GH1(μ)F,G\in H^{1}(\mu) such that for each i{1,,N}i\in\{1,\ldots,N\},

|iF|iG.|\partial_{i}F|\leq\partial_{i}G. (3.17)

Then

Varμ[F]Varμ[G].\mathrm{Var}_{\mu}[F]\leq\mathrm{Var}_{\mu}[G]. (3.18)
Proof.

It is standard that μ\mu satisfies the FKG inequality meaning that for all measurable non-decreasing functions ff and gg, the covariance between ff and gg under μ\mu is non-negative. We refer to [BM92, Th. 1.3] in the N\mathbb{R}^{N} case.

Let F,GH1(μ)F,G\in H^{1}(\mu) be as in (3.17). One may write

Varμ[G]=Varμ[F]+Covμ[G+F,GF]\mathrm{Var}_{\mu}[G]=\mathrm{Var}_{\mu}[F]+\mathrm{Cov}_{\mu}[G+F,G-F]

Since GFG-F and F+GF+G are non-decreasing, their covariance is non-negative, concluding the proof of (3.18). ∎

3.3. Variances upper bounds

We review some well-known variance upper bounds.

Lemma 3.8 (Brascamp-Lieb inequality [BL02]).

Let μ\mu satisfy Assumptions 3.1. Let 𝒜DN\mathcal{A}\subset D_{N} be a convex domain with a piecewise smooth boundary. Let F:DNF:D_{N}\to\mathbb{R} in the form F=GGapNF=G\circ\mathrm{Gap}_{N} with GL2({1,,N},H1(μ))\nabla G\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu^{\prime})). There holds

Varμ[F𝒜]𝔼μ[minUNN(UN2HUN2FUN)𝒜]=𝔼μ[F(2H)1F𝒜].\mathrm{Var}_{\mu}[F\mid\mathcal{A}]\leq-\mathbb{E}_{\mu}\Bigr{[}\min_{U_{N}\in\mathbb{R}^{N}}\Bigr{(}U_{N}\cdot\nabla^{2}HU_{N}-2\nabla F\cdot U_{N}\Bigr{)}\mid\mathcal{A}\Bigr{]}=\mathbb{E}_{\mu}[\nabla F\cdot(\nabla^{2}H)^{-1}\nabla F\mid\mathcal{A}]. (3.19)
Elements of proof.

Let us illustrate the proof in the case 𝒜=DN\mathcal{A}=D_{N}. By Proposition 3.1, the variance of FF may be expressed as

Varμ[F]=minϕH1(μ)𝔼μ[ϕ2Hϕ+|2ϕ|22Fϕ].\mathrm{Var}_{\mu}[F]=-\min_{\phi\in H^{1}(\mu)}\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla^{2}H\nabla\phi+|\nabla^{2}\phi|^{2}-2\nabla F\cdot\nabla\phi].

Since 𝔼μ[|2ϕ|2]0\mathbb{E}_{\mu}[|\nabla^{2}\phi|^{2}]\geq 0, one gets

Varμ[F]minϕH1(μ)𝔼μ[ϕ2Hϕ2Fϕ]𝔼μ[minUNNUN2HUN2FUN]=𝔼μ[F(2H)1F].\begin{split}\mathrm{Var}_{\mu}[F]&\leq-\min_{\phi\in H^{1}(\mu)}\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla^{2}H\nabla\phi-2\nabla F\cdot\nabla\phi]\\ &\leq-\mathbb{E}_{\mu}\Bigr{[}\min_{U_{N}\in\mathbb{R}^{N}}U_{N}\cdot\nabla^{2}HU_{N}-2\nabla F\cdot U_{N}\Bigr{]}\\ &=\mathbb{E}_{\mu}[\nabla F\cdot(\nabla^{2}H)^{-1}\nabla F].\end{split}

The Brascamp-Lieb inequality requires some regularity on FF. We now give a simple upper bound on Var[F]\mathrm{Var}[F] when FF is a linear statistic, which depends only on the L2L^{2} norm of the test-function.

Lemma 3.9 (Poissonian variance estimate).

Let μ\mu satisfy Assumptions 3.1. Let ξL2(𝕋,)\xi\in L^{2}(\mathbb{T},\mathbb{R}). We have

Varμ[i=1Nξ(N1xi)]NN(𝕋ξ2(𝕋ξ)2).\mathrm{Var}_{\mu}\Bigr{[}\sum_{i=1}^{N}\xi(\ell_{N}^{-1}x_{i})\Bigr{]}\leq N{\ell_{N}}\Bigr{(}\int_{\mathbb{T}}\xi^{2}-\Bigr{(}\int_{\mathbb{T}}\xi\Bigr{)}^{2}\Bigr{)}. (3.20)
Proof.

Let μ\mu satisfy Assumptions 3.1. Let ξL2(𝕋,)\xi\in L^{2}(\mathbb{T},\mathbb{R}). Let (ξk)(\xi_{k}) be a sequence of elements of 𝒞2(𝕋,)\mathcal{C}^{2}(\mathbb{T},\mathbb{R}) such that (ξk)(\xi_{k}) converge to ξ\xi in L2(𝕋,)L^{2}(\mathbb{T},\mathbb{R}). Let us prove that (3.20) holds for ξk\xi_{k}. For η>0\eta>0 and iji\neq j, let χijη\chi_{ij}^{\eta} be such that χijη\chi_{ij}^{\eta} is bounded by η1\eta^{-1}, (χijη)′′0(\chi_{ij}^{\eta})^{\prime\prime}\geq 0 and χijηχ\chi_{ij}^{\eta}\leq\chi. Define

dμη=1ZηeHηdXN,whereHη=ijχijη(|xixj|).\mathrm{d}\mu_{\eta}=\frac{1}{Z_{\eta}}e^{-H_{\eta}}\mathrm{d}X_{N},\quad\text{where}\quad H_{\eta}=\sum_{i\neq j}\chi_{ij}^{\eta}(|x_{i}-x_{j}|).

Denote μη\mathcal{L}^{\mu_{\eta}} the operator acting on H1(DN,)H^{-1}(D_{N},\mathbb{R}),

μη:=HηΔ.\mathcal{L}^{\mu_{\eta}}:=\nabla H_{\eta}\cdot\nabla-\Delta.

Since the density of μη\mu_{\eta} is bounded from below and from above with respect to the Lebesgue measure on DND_{N}, one may apply Proposition 3.1, which allows to express the variance of ξk\xi_{k} under μη\mu_{\eta} as

Varμη[FluctN[ξk]]=min𝔼μη[ϕ2Hηϕ+|2ϕ|22ϕFluctN[ξk]],\mathrm{Var}_{\mu_{\eta}}[\mathrm{Fluct}_{N}[\xi_{k}]]=-\min\mathbb{E}_{\mu_{\eta}}[\nabla\phi\cdot\nabla^{2}H_{\eta}\nabla\phi+|\nabla^{2}\phi|^{2}-2\nabla\phi\cdot\nabla\mathrm{Fluct}_{N}[\xi_{k}]],

where the minimum is taken over maps ϕ:DN\phi:D_{N}\to\mathbb{R} such that ϕL2({1,,N},H1(μη))\nabla\phi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu_{\eta})) and ϕn=0\nabla\phi\cdot\vec{n}=0. Since 2Hη\nabla^{2}H_{\eta} is non-negative, one may bound this by

Varμη[FluctN[ξk]]min𝔼μη[|2ϕ|22ϕFluctN[ξk]],\mathrm{Var}_{\mu_{\eta}}[\mathrm{Fluct}_{N}[\xi_{k}]]\leq-\min\mathbb{E}_{\mu_{\eta}}[|\nabla^{2}\phi|^{2}-2\nabla\phi\cdot\nabla\mathrm{Fluct}_{N}[\xi_{k}]], (3.21)

where the minimum is taken over maps ϕH1(μη)\phi\in H^{1}(\mu_{\eta}). By integration by parts under μη\mu_{\eta}, the minimizer ϕ\nabla\phi of (3.21) is the unique solution of

{μηϕ=FluctN[ξk]ϕn=0on DN.\begin{cases}\mathcal{L}^{\mu_{\eta}}\nabla\phi=\nabla\mathrm{Fluct}_{N}[\xi_{k}]\\ \nabla\phi\cdot\vec{n}=0&\text{on $\partial D_{N}$}.\end{cases} (3.22)

Let θk:𝕋\theta_{k}:\mathbb{T}\to\mathbb{R} such that θk′′=(ξkξk)\theta_{k}^{\prime\prime}=-(\xi_{k}-\int\xi_{k}) and let

ϕk:XNDNθk(x1)++θk(xN),\phi_{k}:X_{N}\in D_{N}\mapsto\theta_{k}(x_{1})+\ldots+\theta_{k}(x_{N}),

which satisfies ϕkn=0\nabla\phi_{k}\cdot\vec{n}=0 a.e on DN\partial D_{N} by Remark 3.2.

One can observe that for each i=1,,Ni=1,\ldots,N

μη(iϕk)=Δ(iϕk)+2j:j<i(χijη)(xixj)(iϕkjϕk)=θk(3)(xi)+2j:j<i(χijη)(xixj)θk′′(xi)=θk(3)(xi)=ξk(xi).\mathcal{L}^{\mu_{\eta}}(\partial_{i}\phi_{k})=-\Delta(\partial_{i}\phi_{k})+2\sum_{j:j<i}(\chi_{ij}^{\eta})^{\prime}(x_{i}-x_{j})(\partial_{i}\phi_{k}-\partial_{j}\phi_{k})=-\theta_{k}^{(3)}(x_{i})\\ +2\sum_{j:j<i}(\chi_{ij}^{\eta})^{\prime}(x_{i}-x_{j})\theta_{k}^{\prime\prime}(x_{i})=-\theta_{k}^{(3)}(x_{i})=\xi^{\prime}_{k}(x_{i}).

We thus see that ϕk\nabla\phi_{k} solves (3.22). It follows that

Varμη[FluctN[ξk]]𝔼μη[i=1Nθk(3)(xi)22i=1Nξk(xi)θk′′(xi)]=𝔼μη[i=1Nξk2(xi)]=N(𝕋ξk2(𝕋ξk)2),\mathrm{Var}_{\mu_{\eta}}[\mathrm{Fluct}_{N}[\xi_{k}]]\leq-\mathbb{E}_{\mu_{\eta}}\Bigr{[}\sum_{i=1}^{N}\theta_{k}^{(3)}(x_{i})^{2}-2\sum_{i=1}^{N}\xi_{k}^{\prime}(x_{i})\theta_{k}^{\prime\prime}(x_{i})\Bigr{]}=\mathbb{E}_{\mu_{\eta}}\Bigr{[}\sum_{i=1}^{N}\xi_{k}^{2}(x_{i})\Bigr{]}\\ =N\Bigr{(}\int_{\mathbb{T}}\xi_{k}^{2}-\Bigr{(}\int_{\mathbb{T}}\xi_{k}\Bigr{)}^{2}\Bigr{)},

where we use the fact that the first marginal of μη\mu_{\eta} is the Lebesgue measure on 𝕋\mathbb{T}. Letting kk go to ++\infty yields

Varμη[FluctN[ξ]]N(𝕋ξ2(𝕋ξ)2).\mathrm{Var}_{\mu_{\eta}}[\mathrm{Fluct}_{N}[\xi]]\leq N\Bigr{(}\int_{\mathbb{T}}\xi^{2}-\Bigr{(}\int_{\mathbb{T}}\xi\Bigr{)}^{2}\Bigr{)}.

Then, letting η\eta tend to 0, we deduce by monotone convergence that (3.20) holds. ∎

3.4. Log-Sobolev inequalities and Gaussian concentration

We now review some standard results on log-Sobolev inequalities and Gaussian concentration for log-concave measures on N\mathbb{R}^{N} and derive some stronger estimates valid for measures satisfying Assumptions 3.1 following [BEY12].

Let us first recall a crucial convexity result proved in [BL02].

Lemma 3.10.

Let μ\mu satisfy Assumptions 3.1. Assume that for each i,ji,j,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Let I{1,,N}I\subset\{1,\ldots,N\} of cardinal mm. Denote πI\pi_{I} and πIc\pi_{I^{c}} the projections on the coordinates (xi)iI(x_{i})_{i\in I} and (xi)iIc(x_{i})_{i\in I^{c}}. Split HH into H=H1πI+H2H=H_{1}\circ\pi_{I}+H_{2} with

H1:(xi)iIDmijIχ(|xixj|),H2:XNDNiIcjiχ(|xixj|).H_{1}:(x_{i})_{i\in I}\in D_{m}\mapsto\sum_{i\neq j\in I}\chi(|x_{i}-x_{j}|),\quad H_{2}:X_{N}\in D_{N}\mapsto\sum_{i\in I^{c}}\sum_{j\neq i}\chi(|x_{i}-x_{j}|).

Let ν=πI#μ\nu=\pi_{I}\#\mu. The density of ν\nu may be written

dν(x)e(H1+H~)𝟙Dm(x)dx,\mathrm{d}\nu(x)\propto e^{-(H_{1}+\tilde{H})}\mathds{1}_{D_{m}}(x)\mathrm{d}x, (3.23)

with

H~:xDmlogeH2(x,y)𝟙DN(x,y)dy.\tilde{H}:x\in D_{m}\mapsto-\log\int e^{-H_{2}(x,y)}\mathds{1}_{D_{N}}(x,y)\mathrm{d}y. (3.24)

Moreover

2H~0.\nabla^{2}\tilde{H}\geq 0.
Proof.

Let I{1,,N}I\subset\{1,\ldots,N\} of cardinal mm. The fact that ν\nu is as in (3.23) is straightforward. Let us prove that the Hessian of (3.24) is non-negative. On DND_{N}, introduce the coordinates

(z,y)=((xi)iI,(xi)iIc).(z,y)=((x_{i})_{i\in I},(x_{i})_{i\in I^{c}}).

For smooth functions f:DNf:D_{N}\to\mathbb{R}, denote 1f=(if)iI,\nabla_{1}f=(\partial_{i}f)_{i\in I}, 2f=(if)iIc\nabla_{2}f=(\partial_{i}f)_{i\in I^{c}} and 112f,122f,212f,222f\nabla^{2}_{11}f,\nabla_{12}^{2}f,\nabla^{2}_{21}f,\nabla^{2}_{22}f the matrices of second partial derivatives. Fix zDmz\in D_{m}, vmv\in\mathbb{R}^{m} and let

h:tH~(z+tv).h:t\mapsto\tilde{H}(z+tv).

Since for each i,ji,j,

limx0χij(x)=+,\lim_{x\to 0}\chi_{ij}(x)=+\infty,

one can check that for all tt\in\mathbb{R},

h′′(t)=𝔼(x+tv)[v112H2v]Var(x+tv)[v1H2],h^{\prime\prime}(t)=\mathbb{E}_{\mathbb{P}(x+tv)}[v\cdot\nabla^{2}_{11}H_{2}v]-\mathrm{Var}_{\mathbb{P}(x+tv)}[v\cdot\nabla_{1}H_{2}], (3.25)

where for any zDmz\in D_{m}, (z)\mathbb{P}(z) is the probability measure

d(z)=1Z(z)eH2(z,y)𝟙(z,y)DNdy.\mathrm{d}\mathbb{P}(z)=\frac{1}{Z(z)}e^{-H_{2}(z,y)}\mathds{1}_{(z,y)\in D_{N}}\mathrm{d}y.

Since yH2(x,y)y\mapsto H_{2}(x,y) is convex, the Brascamp-Lieb inequality implies

Var(x+tv)[v1H2]𝔼(x+tv)[v(122H2)(222H2)1(122H2)v].\mathrm{Var}_{\mathbb{P}(x+tv)}[v\cdot\partial_{1}H_{2}]\leq\mathbb{E}_{\mathbb{P}(x+tv)}[v\cdot(\nabla^{2}_{12}H_{2})(\nabla^{2}_{22}H_{2})^{-1}(\nabla^{2}_{12}H_{2})v].

Furthermore since 2H2\nabla^{2}H_{2} is non-negative, its Schur complement is non-negative, which gives

112H2122H2(222H2)1122H20.\nabla^{2}_{11}H_{2}-\nabla^{2}_{12}H_{2}(\nabla^{2}_{22}H_{2})^{-1}\nabla^{2}_{12}H_{2}\geq 0.

Inserting this into (3.25)(\ref{eq:h''}), this shows that 2H~0\nabla^{2}\tilde{H}\geq 0. ∎

Let us recall the standard log-Sobolev inequality for uniformly log-concave measures on N\mathbb{R}^{N}, which is a special case of the Bakry-Emery criterion [BÉ85]. Let ν\nu and μ\mu be two probability measures on DND_{N}. Recall that the relative entropy of μ\mu with respect to ν\nu is defined by

Ent(μν)=logdμdνdμ[0,+],\mathrm{Ent}(\mu\mid\nu)=\int\log\frac{\mathrm{d}\mu}{\mathrm{d}\nu}\mathrm{d}\mu\in[0,+\infty],

if ν\nu is absolutely continuous with respect to μ\mu and Ent(νμ)=+\mathrm{Ent}(\nu\mid\mu)=+\infty otherwise. Let also recall the Fisher information of μ\mu with respect to ν\nu,

Fisher(μν)=|logdνdμ|2dν,\mathrm{Fisher}(\mu\mid\nu)=\Bigr{|}\nabla\log\frac{\mathrm{d}\nu}{\mathrm{d}\mu}\Bigr{|}^{2}\mathrm{d}\nu,

if ν\nu is absolutely continuous with respect to μ\mu and Fisher(νμ)=+\mathrm{Fisher}(\nu\mid\mu)=+\infty otherwise.

Lemma 3.11 (Bakry-Emery [BÉ85]).

Let K{K} be a convex domain of N\mathbb{R}^{N}. Let w>0w>0 and γw\gamma^{w} be a centered Gaussian distribution on N\mathbb{R}^{N} with covariance matrix 1wIN\frac{1}{w}I_{N}. Let γKw\gamma_{K}^{w} defined by conditioning γw\gamma^{w} into KK. Assume that μ\mu is a measure on KK in the form dμ=fdγKw\mathrm{d}\mu=f\mathrm{d}\gamma_{K}^{w} with f:Kf:K\to\mathbb{R} Borel and log-concave. Then ν\nu satisfies a log-Sobolev inequality with constant 2w2w, meaning for all probability measure μ𝒫(K)\mu\in\mathcal{P}(K),

Ent(μν)2wFisher(μν).\mathrm{Ent}(\mu\mid\nu)\leq 2w\mathrm{Fisher}(\mu\mid\nu).

Moreover μ\mu satisfies Gaussian concentration: for all FH1F\in H^{1}, we have

log𝔼μ[etF]t𝔼μ[F]+w2t2supK|F|2,for all t.\log\mathbb{E}_{\mu}[e^{tF}]\leq t\mathbb{E}_{\mu}[F]+\frac{w}{2}t^{2}\sup_{K}|\nabla F|^{2},\quad\text{for all $t\in\mathbb{R}$}. (3.26)

We now state a key concentration result due to [BEY12]. Recall that if μ\mu satisfies Assumptions 1.1, then tere exists c>0c>0 such that

UN2HUNcij(uiuj)2,for all UNN.U_{N}\cdot\nabla^{2}HU_{N}\geq c\sum_{i\neq j}(u_{i}-u_{j})^{2},\quad\text{for all }U_{N}\in\mathbb{R}^{N}.

The crucial observation is that when i=1Nui=0\sum_{i=1}^{N}u_{i}=0, the Hessian of the energy controls N1N-1 times the Euclidean norm of uu:

UN2HUN(N1)ci=1Nui2.U_{N}\cdot\nabla^{2}HU_{N}\geq(N-1)c\sum_{i=1}^{N}u_{i}^{2}. (3.27)

Furthermore one can observe that ϕ:=(A1μ)1F\nabla\phi:=(A_{1}^{\mu})^{-1}\nabla F satisfies 1ϕ++Nϕ=0\partial_{1}\phi+\ldots+\partial_{N}\phi=0 when 1F++NF=0\partial_{1}F+\ldots+\partial_{N}F=0 since ϕ\phi is a function of the gaps. Combining this with (3.27) gives the following Gaussian estimate:

Lemma 3.12.

Let I{1,,N}I\subset\{1,\ldots,N\} and πI\pi_{I} the projection on the coordinates (xi)iI(x_{i})_{i\in I}. Let μ\mu satisfy Assumptions 3.1. Assume that for each i,ji,j,

limx0χij(x)=+.\lim_{x\to 0}\chi_{ij}(x)=+\infty.

Assume that for each i,jIi,j\in I, iji\neq j and XNDNX_{N}\in D_{N},

χij′′(xjxi)c12>0.\chi_{ij}^{\prime\prime}(x_{j}-x_{i})\geq\frac{c_{1}}{2}>0. (3.28)

Let F=GπIH1(μ)F=G\circ\pi_{I}\in H^{1}(\mu). Assume that FF is independent of iIxi\sum_{i\in I}x_{i}, i.e iIiF=0\sum_{i\in I}\partial_{i}F=0. For all tt\in\mathbb{R} we have

log𝔼μ[etF]t𝔼μ[F]+t22c1(|I|1)sup|F|2.\log\mathbb{E}_{\mu}[e^{tF}]\leq t\mathbb{E}_{\mu}[F]+\frac{t^{2}}{2c_{1}(|I|-1)}\sup|\nabla F|^{2}.

The proof of Lemma 3.12 can be found in [BEY12, Le. 3.9]. It can be adapted readily to our circular setting. For completeness we sketch the main arguments below and follow line by line the proof of [BEY12].

Proof.

Let I{1,,N}I\in\{1,\ldots,N\} of cardinal mm. To simplify the notation assume that I={1,,m}I=\{1,\ldots,m\}.

On DND_{N} introduce the coordinates (x,y)(x,y) with x=(xi)iIDmx=(x_{i})_{i\in I}\in D_{m} and y=(xi)iIcDNmy=(x_{i})_{i\in I^{c}}\in D_{N-m}. The energy HH can be split into H(x,y)=H1(x)+H2(x,y)H(x,y)=H_{1}(x)+H_{2}(x,y) with H1H_{1} uniformly convex, H2H_{2} convex and H1H_{1} independent of iIxi\sum_{i\in I}x_{i}, i.e iIiH1=0\sum_{i\in I}\partial_{i}H_{1}=0. Now we introduce on DmD_{m} the coordinates x=(z,q)x=(z,q) with z=(x1,,xm1)Dm1z=(x_{1},\ldots,x_{m-1})\in D_{m-1} and q=m12i=1mxiq=m^{-\frac{1}{2}}\sum_{i=1}^{m}x_{i}. Observe that this change of variables can be written (z,w)=M(x1,,xm)(z,w)=M(x_{1},\ldots,x_{m}), with M𝒪m()M\in\mathcal{O}_{m}(\mathbb{R}). Since H1H_{1} is independent of qq, one can write it in the form H1=H~1(z)H_{1}=\tilde{H}_{1}(z). Similarly FF can be written F=F~(z)F=\tilde{F}(z).

Let ν\nu be the push-forward of μ\mu by (z,q,y)z(z,q,y)\mapsto z. The density of ν\nu is given by dν(z)eH~(z)dz\mathrm{d}\nu(z)\propto e^{-\tilde{H}(z)}\mathrm{d}z where

H~(z)=logeH(z,q,y)dqdy.\tilde{H}(z)=-\log\int e^{-H(z,q,y)}\mathrm{d}q\mathrm{d}y. (3.29)

Fix zDm1z\in D_{m-1} and vm1v\in\mathbb{R}^{m-1}. Consider h:tH~(z+tv)h:t\mapsto\tilde{H}(z+tv). As in the proof of Lemma 3.10, we have

h′′(t)=𝔼(z+tv)[vzz2HvVar(z+tv)[vzH],h^{\prime\prime}(t)=\mathbb{E}_{\mathbb{P}(z+tv)}[v\cdot\nabla^{2}_{zz}Hv-\mathrm{Var}_{\mathbb{P}(z+tv)}[v\cdot\nabla_{z}H],

where for any zDm1z\in D_{m-1}, (z)\mathbb{P}(z) stands for the probability measure

d(z)=1Z(z)eH(z,q,y)dqdy.\mathrm{d}\mathbb{P}(z)=\frac{1}{Z(z)}e^{-H(z,q,y)}\mathrm{d}q\mathrm{d}y.

Using the Brascamp-Lieb inequality this entails

h′′(t)𝔼(z+tv)[vzz2Hvvzq2H(qq2H)1zq2Hv].h^{\prime\prime}(t)\geq\mathbb{E}_{\mathbb{P}(z+tv)}[v\cdot\nabla^{2}_{zz}Hv-v\cdot\nabla_{zq}^{2}H(\nabla_{qq}^{2}H)^{-1}\nabla_{zq}^{2}Hv].

Thus

zz2H~𝔼(x+tv)[zz2Hzq2H(qq2H)1zq2H)].\nabla^{2}_{zz}\tilde{H}\geq\mathbb{E}_{\mathbb{P}(x+tv)}[\nabla_{zz}^{2}H-\nabla_{zq}^{2}H(\nabla_{qq}^{2}H)^{-1}\nabla_{zq}^{2}H)]. (3.30)

Since H1H_{1} is independent of qq, one has

zq2H(qq2H)1zq2H=zq2H2(qq2H2)1zq2H2.\nabla^{2}_{zq}H(\nabla^{2}_{qq}H)^{-1}\nabla^{2}_{zq}H=\nabla^{2}_{zq}H_{2}(\nabla^{2}_{qq}H_{2})^{-1}\nabla^{2}_{zq}H_{2}.

Hence, by positivity of zz2H2\nabla^{2}_{zz}H_{2}, its Schur complement is positive and

zz2Hzq2H(qq2H)1zq2H=zz2H1+zz2H2zq2H2(qq2H2)1zq2H2zz2H1.\nabla_{zz}^{2}H-\nabla_{zq}^{2}H(\nabla_{qq}^{2}H)^{-1}\nabla^{2}_{zq}H=\nabla^{2}_{zz}H_{1}+\nabla^{2}_{zz}H_{2}-\nabla^{2}_{zq}H_{2}(\nabla^{2}_{qq}H_{2})^{-1}\nabla^{2}_{zq}H_{2}\geq\nabla^{2}_{zz}H_{1}.

Inserting this into (3.30) we deduce that for all um1u\in\mathbb{R}^{m-1},

uzz2H~uuzz2H1u=M~uxx2H1M~ucij((M~u)i(M~u)j)2,u\cdot\nabla^{2}_{zz}\tilde{H}u\geq u\cdot\nabla^{2}_{zz}H_{1}u={\tilde{M}u}\cdot\nabla_{xx}^{2}H_{1}{\tilde{M}u}\geq c\sum_{i\neq j}((\tilde{M}u)_{i}-(\tilde{M}u)_{j})^{2},

where M~\tilde{M} denotes the first m1m-1 columns of MM. Moreover we can observe that

ij((M~u)i(M~u)j)2=(m1)i=1m1ui2.\sum_{i\neq j}((\tilde{M}u)_{i}-(\tilde{M}u)_{j})^{2}=(m-1)\sum_{i=1}^{m-1}u_{i}^{2}.

Since ν\nu is uniformly log-concave for the constant (m1)c(m-1)c, one can apply the Bakry-Emery criterion stated in Lemma 3.11, which gives that for all tt\in\mathbb{R},

𝔼μ[etF]=𝔼ν[etF~]et𝔼ν[F~]+t22c(m1)sup|zF~|2.\mathbb{E}_{\mu}[e^{tF}]=\mathbb{E}_{\nu}[e^{t\tilde{F}}]\leq e^{t\mathbb{E}_{\nu}[\tilde{F}]+\frac{t^{2}}{2c(m-1)}\sup|\nabla_{z}\tilde{F}|^{2}}.

We can now observe that, since MM is orthogonal, |zF~|2=|F|2|\nabla_{z}\tilde{F}|^{2}=|\nabla F|^{2} and this concludes the proof. ∎

4. Near-optimal rigidity

This section is devoted to the proof of the rigidity result of Theorem 1. The method uses various techniques introduced in the seminal paper [BEY12, Th. 3.1]. Since we are working on the circle, it is straightforward to compute the expectation of gaps and one does not need to estimate the accuracy of standard positions, which was one of the main issues of [BEY12]. The first task is to obtain a local law on gaps saying that for each i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2}, N(xi+kxi)N(x_{i+k}-x_{i}) is typically of order kk. To this end we perform a mutliscale analysis similar to [BEY12] allowing one to bootstrap the local law down to microscale. The argument is based on a convexifying procedure that we first detail.

4.1. Comparison to a constrained Gibbs measures

Since the Hessian of the energy degenerates when particles are far away from each other, one cannot directly derive Gaussian concentration estimates for N,β\mathbb{P}_{N,\beta}. Following [BEY12], one may add to the Hamiltonian a convexifying term, which penalizes configurations with large gaps. Let θ:++\theta:\mathbb{R}^{+}\to\mathbb{R}^{+} be a smooth function such that θ(x)=x2\theta(x)=x^{2} for x>1x>1, θ=0\theta=0 on [0,12][0,\frac{1}{2}] and θ′′0\theta^{\prime\prime}\geq 0 on +\mathbb{R}^{+}. Let i0{1,,N}i_{0}\in\{1,\ldots,N\}, L{1,,N2}L\in\{1,\ldots,\frac{N}{2}\} and K>0K>0. Let

I:={j:d(i0,j)L}.I:=\{j:d(i_{0},j)\leq L\}.

Define

F:=2i<jIθ(NK(xjxi))\mathrm{F}:=2\sum_{i<j\in I}\theta\Bigr{(}\frac{N}{K}(x_{j}-x_{i})\Bigr{)} (4.1)

and N,β\mathbb{Q}_{N,\beta} the locally constrained Gibbs measure

dN,β(XN)=1KN,βeβ(N+F)(XN)𝟙DN(XN)dXN.\mathrm{d}\mathbb{Q}_{N,\beta}(X_{N})=\frac{1}{K_{N,\beta}}e^{-\beta(\mathcal{H}_{N}+\mathrm{F})(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N}. (4.2)

In the sequel we will often take K=L1+εK=L^{1+\varepsilon} for some ε>0\varepsilon>0. Recall the total variation distance between two measures μ\mu and ν\nu on DND_{N}:

TV(μ,ν)=sup𝒜(DN)|μ(𝒜)ν(𝒜)|.\mathrm{TV}(\mu,\nu)=\sup_{\mathcal{A}\in\mathcal{B}(D_{N})}|\mu(\mathcal{A})-\nu(\mathcal{A})|.

The Pinsker inequality, see [ABC+00, Ch. 5] for a proof, asserts that

TV(μ,ν)22Ent(μν),\mathrm{TV}(\mu,\nu)^{2}\leq 2\mathrm{Ent}(\mu\mid\nu), (4.3)

where Ent(ν)\mathrm{Ent}(\cdot\mid\nu) is the relative entropy with respect to ν\nu. Using (4.3) and the log-concavity of the constrained measure (4.2), one may derive the following control:

Lemma 4.1.

Let N,β\mathbb{Q}_{N,\beta} be the measure (4.2). Denote πI\pi_{I} the projection πI:XNDN(xi)iID2L+1\pi_{I}:X_{N}\in D_{N}\mapsto(x_{i})_{i\in I}\in D_{2L+1}. There exists a constant C>0C>0 depending only on β\beta and ss such that

TV(πI#N,β,πI#N,β)2CL5Ks𝔼N,β[(NK(xi0+Lxi0L))2𝟙xi0+Lxi0LK2N].\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})^{2}\leq CL^{5}K^{s}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\Bigr{(}\frac{N}{K}(x_{i_{0}+L}-x_{i_{0}-L})\Bigr{)}^{2}\mathds{1}_{x_{i_{0}+L}-x_{i_{0}-L}\geq\frac{K}{2N}}\Bigr{]}.
Proof.

For each k=1,,N/2,k=1,\ldots,\lfloor N/2\rfloor, let

Gapk:XkDk(N(x2x1),,N(xkxk1))k1.\mathrm{Gap}_{k}:X_{k}\in D_{k}\mapsto(N(x_{2}-x_{1}),\ldots,N(x_{k}-x_{k-1}))\in\mathbb{R}^{k-1}.

Applying Pinsker’s inequality (4.3) to μ:=πI#N,β\mu:=\pi_{I}\#\mathbb{P}_{N,\beta} and ν:=πI#N,β\nu:=\pi_{I}\#\mathbb{Q}_{N,\beta} gives

TV(πI#N,β,πI#N,β)22Ent(πI#N,βπI#N,β).\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})^{2}\leq 2\mathrm{Ent}(\pi_{I}\#\mathbb{P}_{N,\beta}\mid\pi_{I}\#\mathbb{Q}_{N,\beta}). (4.4)

Note that

Ent(πI#N,βπI#N,β)=Ent((Gap2L+1πI)#N,β(Gap2L+1πI)#N,β).\mathrm{Ent}(\pi_{I}\#\mathbb{P}_{N,\beta}\mid\pi_{I}\#\mathbb{Q}_{N,\beta})=\mathrm{Ent}((\mathrm{Gap}_{2L+1}\circ\pi_{I})\#\mathbb{P}_{N,\beta}\mid(\mathrm{Gap}_{2L+1}\circ\pi_{I})\#\mathbb{Q}_{N,\beta}).

The law of xix_{i} under N,β\mathbb{P}_{N,\beta} and N,β\mathbb{Q}_{N,\beta} is uniform on the circle and independent of Gap2L+1πI(XN)\mathrm{Gap}_{2L+1}\circ\pi_{I}(X_{N}). By Lemma 3.10, the measure Gap2L+1πI(XN)\mathrm{Gap}_{2L+1}\circ\pi_{I}(X_{N}) has density exp(β(H+H~))\exp(-\beta(H+\tilde{H})) with 2H~0\nabla^{2}\tilde{H}\geq 0 and HH defined by

H:(xi)iD2L+1D2L+1β(NsijD2L+1g(xixj)+FπI).H:(x_{i})_{i\in D_{2L+1}}\in D_{2L+1}\mapsto\beta(N^{-s}\sum_{i\neq j\in D_{2L+1}}g(x_{i}-x_{j})+\mathrm{F}\circ\pi_{I}).

By definition of θ\theta,

U2L+12HU2L+1Ci<jI(N(ujui))2Ks+2CiImaxI(N(ui+1ui))2Ks+2,for all U2L+12L+1,U_{2L+1}\cdot\nabla^{2}HU_{2L+1}\geq C\sum_{i<j\in I}\frac{(N(u_{j}-u_{i}))^{2}}{K^{s+2}}\\ \geq C\sum_{i\in I\setminus\max I}\frac{(N(u_{i+1}-u_{i}))^{2}}{K^{s+2}},\quad\text{for all $U_{2L+1}\in\mathbb{R}^{2L+1}$},

for some constant C>0C>0 independent of NN, LL and LL. Therefore by Lemma 3.11, (Gap2L+1πI)#N,β(\mathrm{Gap}_{2L+1}\circ\pi_{I})\#\mathbb{Q}_{N,\beta} satisfies a log-Sobolev inequality with constant 2c12c^{-1} where c:=CβKs+2c:=\frac{C\beta}{K^{s+2}}. Writing F=GGap2L+1πI\mathrm{F}=\mathrm{G}\circ\mathrm{Gap}_{2L+1}\circ\pi_{I} for some G:2L\mathrm{G}:\mathbb{R}^{2L}\to\mathbb{R}, this gives

Ent((Gap2L+1πI)#N,βν)CKs+2𝔼N,β[|(GGap2L+1πI)|2].\mathrm{Ent}((\mathrm{Gap}_{2L+1}\circ\pi_{I})\#\mathbb{P}_{N,\beta}\mid\nu)\leq CK^{s+2}\mathbb{E}_{\mathbb{P}_{N,\beta}}[|\nabla(\mathrm{G}\circ\mathrm{Gap}_{2L+1}\circ\pi_{I})|^{2}]. (4.5)

We can next bound the Fisher information by

𝔼N,β[|G|2Gap2L+1πI]CL5K2𝔼N,β[(θ(NK(xi0+Lxi0)))2]CL5K2𝔼N,β[(NK(xi0+Lxi0L)))2𝟙xi0+Lxi0LK2N].\begin{split}\mathbb{E}_{\mathbb{P}_{N,\beta}}[|\nabla\mathrm{G}|^{2}\circ\mathrm{Gap}_{2L+1}\circ\pi_{I}]&\leq CL^{5}K^{-2}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\Bigr{(}\theta^{\prime}\Bigr{(}\frac{N}{K}(x_{i_{0}+L}-x_{i_{0}})\Bigr{)}\Bigr{)}^{2}\Bigr{]}\\ &\leq CL^{5}K^{-2}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\Bigr{(}\frac{N}{K}(x_{i_{0}+L}-x_{i_{0}-L})\Bigr{)}\Bigr{)}^{2}\mathds{1}_{x_{i_{0}+L}-x_{i_{0}-L}\geq\frac{K}{2N}}\Bigr{]}.\end{split}

Inserting this into (4.5) and using (4.4) concludes the proof of Lemma 4.1. ∎

4.2. First local law

We now prove that each gap N(xi+kxi)N(x_{i+k}-x_{i}) is typically of order kk with an exponentially small probability of deviation.

Lemma 4.2.

Let δ>0\delta>0. There exist c>0c>0 and C>0C>0 such that for each i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2},

N,β(N(xi+1xi)k1+δ)Cexp(ck2min(δ,1s2(2+s))).\mathbb{P}_{N,\beta}(N(x_{i+1}-x_{i})\geq k^{1+\delta})\leq C\exp\Bigr{(}{-ck^{2\min(\delta,\frac{1-s}{2(2+s)})}}\Bigr{)}. (4.6)

The proof of Lemma 4.2 is inspired by the multiscale analysis of [BEY12]. We proceed by a bootstrap on scales: if the local law (4.6) is assumed to hold for some k{1,,N2}k\in\{1,\ldots,\frac{N}{2}\}, then in view of Lemma 4.1, one may convexify the measure in a window of size kk without changing much the measure. Moreover, the convexified measure satisfies better concentration estimates, allowing one to prove through Lemma 3.12 that (4.6) holds at a slightly smaller scale.

Proof of Lemma 4.2.
Step 1: setting the bootstrap

Define

δ0:=1s2(s+2).\delta_{0}:=\frac{1-s}{2(s+2)}.

We wish to prove that there exist c0>0c_{0}>0 and C0>0C_{0}>0 independent of NN such that for each i{1,,N}i\in\{1,\ldots,N\}, 1kN21\leq k\leq\frac{N}{2} and all δ>0\delta>0,

N,β(N(xi+kxi)k1+δ){C0ec0k2δif δ(0,δ0]C0ec0k2δ0(1+δ1+δ0)2if δ>δ0..\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq\begin{cases}C_{0}e^{-c_{0}k^{2\delta}}&\text{if $\delta\in(0,\delta_{0}]$}\\ C_{0}e^{-c_{0}k^{2\delta_{0}(\frac{1+\delta}{1+\delta_{0}})^{2}}}&\text{if $\delta>\delta_{0}$}.\end{cases}. (4.7)

Observe that (4.7) trivially holds for K=N/2K=\lfloor N/2\rfloor since 0N(xi+N/2xi)N0\leq N(x_{i+\lfloor N/2\rfloor}-x_{i})\leq N.

Let K{1,,N/2}K\in\{1,\ldots,N/2\}. Assume that (4.7) holds for each kKk\geq K. Fix δ(0,δ0]\delta\in(0,\delta_{0}] and let

α0(0,11+s2δ(s+2)).\alpha_{0}\in\Bigr{(}0,1-\frac{1+s}{2-\delta(s+2)}\Bigr{)}. (4.8)

Note that since δδ0<1ss+2\delta\leq\delta_{0}<\frac{1-s}{s+2}, we have 11+s2δ(s+2)>01-\frac{1+s}{2-\delta(s+2)}>0. Let us prove that (4.7) holds for each kK1α0k\geq K^{1-\alpha_{0}}.

Let i{1,,N}i\in\{1,\ldots,N\}, kK1α0k\geq K^{1-\alpha_{0}} and

γ(δ(1α0),δ0).\gamma\in(\delta(1-\alpha_{0}),\delta_{0}). (4.9)

Define

I:={i,,i+K}.I:=\{i,\cdots,i+K\}.

Let θ:++\theta:\mathbb{R}^{+}\to\mathbb{R}^{+} be a smooth cutoff function such that θ(x)=x2\theta(x)=x^{2} for x>1x>1, θ=0\theta=0 on [0,12][0,\frac{1}{2}] and θ′′0\theta^{\prime\prime}\geq 0 on +\mathbb{R}^{+}. Define

F:=l<jIθ(NK1+γ(xlxj)).\mathrm{F}:=\sum_{l<j\in I}\theta\Bigr{(}\frac{N}{K^{1+\gamma}}(x_{l}-x_{j})\Bigr{)}. (4.10)

Let N,β\mathbb{Q}_{N,\beta} be the constrained Gibbs measure

dN,β(XN)=1KN,βeβ(N+F)(XN)𝟙DN(XN)dXN.\mathrm{d}\mathbb{Q}_{N,\beta}(X_{N})=\frac{1}{K_{N,\beta}}e^{-\beta(\mathcal{H}_{N}+\mathrm{F})(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N}.

Since xi+kxix_{i+k}-x_{i} is a function of (xi)iI(x_{i})_{i\in I}, one can write

N,β(N(xi+kxi)k1+δ)N,β(N(xi+kxi)k1+δ)+TV(πI#N,β,πI#N,β).\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq\mathbb{Q}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})+\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta}). (4.11)
Step 2: upper bound on the total variation distance

By Lemma 4.1, we have

TV(πI#N,β,πI#N,β)CK7𝔼N,β[(1K1+γN(xi+Kxi))2𝟙N(xi+Kxi)12K1+γ]CK7(N,β(12K1+γN(xi+Kxi)K1+γ)+jK1+γj2K2(1+γ)N,β(N(xi+Kxi)=j)).\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})\leq CK^{7}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\Bigr{(}\frac{1}{K^{1+\gamma}}N(x_{i+K}-x_{i})\Bigr{)}^{2}\mathds{1}_{N(x_{i+K}-x_{i})\geq\frac{1}{2}K^{1+\gamma}}\Bigr{]}\\ \leq CK^{7}\Bigr{(}\mathbb{P}_{N,\beta}\Bigr{(}\frac{1}{2}K^{1+\gamma}\leq N(x_{i+K}-x_{i})\leq K^{1+\gamma}\Bigr{)}+\sum_{j\geq K^{1+\gamma}}\frac{j^{2}}{K^{2(1+\gamma)}}\mathbb{P}_{N,\beta}(N(x_{i+K}-x_{i})=j)\Bigr{)}. (4.12)

By (4.7), for each j{1,,N/2},j\in\{1,\ldots,\lfloor N/2\rfloor\},

N,β(N(xi+kxi)=j)C0exp(c0j2min(δ1+δ,δ01+δ0(1+δ1+δ0))).\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})=j)\leq C_{0}\exp(-c_{0}j^{2\min(\frac{\delta}{1+\delta},\frac{\delta_{0}}{1+\delta_{0}}(\frac{1+\delta}{1+\delta_{0}}))}). (4.13)

Inserting this into (4.12) and using the fact that kK1α0k\geq K^{1-\alpha_{0}}, one gets that for all κ<γ1α0\kappa<\frac{\gamma}{1-\alpha_{0}}, there exists C>0C>0 independent of NN and KK such that

TV(πI#N,β,πI#N,β)Cec0k2κ.\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})\leq Ce^{-c_{0}k^{2\kappa}}.

Since γ>δ(1α0)\gamma>\delta(1-\alpha_{0}), we deduce that there exists k0k_{0} independent of KK such that for each kmax(K,k0)k\geq\max(K,k_{0}),

TV(πI#N,β,πI#N,β)14C0ec0k2δ.\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})\leq\frac{1}{4}C_{0}e^{-c_{0}k^{2\delta}}. (4.14)
Step 3: accuracy under N,β\mathbb{Q}_{N,\beta}

Since N(xi+kxi)N(x_{i+k}-x_{i}) is not bounded from above independently of NN, one cannot directly apply (4.14) to approximate the expectation of N(xi+kxi)N(x_{i+k}-x_{i}) under N,β\mathbb{Q}_{N,\beta} and one needs to first prove a tightness result. Let ε>0\varepsilon^{\prime}>0. We have

log𝔼N,β[e(N(xi+kxi))ε]=log𝔼N,β[e(N(xi+kxi))εβF]log𝔼N,β[eβF]log𝔼N,β[e(N(xi+kxi))ε]+β𝔼N,β[F],\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{(N(x_{i+k}-x_{i}))^{\varepsilon^{\prime}}}]=\log\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{(N(x_{i+k}-x_{i}))^{\varepsilon^{\prime}}-\beta\mathrm{F}}]-\log\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{-\beta\mathrm{F}}]\\ \leq\log\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{(N(x_{i+k}-x_{i}))^{\varepsilon^{\prime}}}]+\beta\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{F}], (4.15)

by Jensen’s inequality. In view of (4.13), we have

𝔼N,β[F]CK,\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{F}]\leq CK, (4.16)

for some constant C>0C>0 independent of NN and KK. Moreover

𝔼N,β[e(N(xi+kxi))ε]𝔼N,β[e(N(xi+Kxi))ε]eKε+jK+1ejεN,β(N(xi+Kxi)=j).\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{(N(x_{i+k}-x_{i}))^{\varepsilon^{\prime}}}]\leq\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{(N(x_{i+K}-x_{i}))^{\varepsilon^{\prime}}}]\leq e^{K^{\varepsilon^{\prime}}}+\sum_{j\geq K+1}e^{j^{\varepsilon^{\prime}}}\mathbb{P}_{N,\beta}(N(x_{i+K}-x_{i})=j).

Using (4.13), one finds that for ε>0\varepsilon^{\prime}>0 small enough,

log𝔼N,β[e(N(xi+kxi))ε]CK,\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{(N(x_{i+k}-x_{i}))^{\varepsilon^{\prime}}}]\leq CK, (4.17)

for some constant C>0C>0 independent of NN and KK. Using Markov’s inequality one gets

𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)K2/ε]eCKjK2/εkekε=O(1),\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})\geq K^{2/\varepsilon^{\prime}}}]\leq e^{CK}\sum_{j\geq K^{2/\varepsilon^{\prime}}}ke^{-k^{\varepsilon^{\prime}}}=O(1), (4.18)

for some O(1)O(1) uniform in NN and KK. Besides, using (4.13), we get that

𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)>K2/ε]=O(1),\mathbb{E}_{\mathbb{P}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})>K^{2/\varepsilon^{\prime}}}]=O(1), (4.19)

for some O(1)O(1) uniform in NN and KK.

For all κ<γ1α0\kappa<\frac{\gamma}{1-\alpha_{0}}, we have

|𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)K2/ε]𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)K2/ε]|K2/εTV(πI#N,β,πI#N,β)14K2/εC0ec0k2κ,|\mathbb{E}_{\mathbb{P}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})\leq K^{2/\varepsilon^{\prime}}}]-\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})\leq K^{2/\varepsilon^{\prime}}}]|\\ \leq K^{2/\varepsilon^{\prime}}\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})\leq\frac{1}{4}K^{2/\varepsilon^{\prime}}C_{0}e^{-c_{0}k^{2\kappa}},

for each kmax(K,k0)k\geq\max(K,k_{0}). Furthermore, combining (4.18) and (4.19), one gets

𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)>K2/ε]𝔼N,β[N(xi+kxi)𝟙N(xi+kxi)>K2/ε]=O(1),\mathbb{E}_{\mathbb{P}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})>K^{2/\varepsilon^{\prime}}}]-\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})\mathds{1}_{N(x_{i+k}-x_{i})>K^{2/\varepsilon^{\prime}}}]=O(1),

for some O(1)O(1) uniform in NN and KK. We deduce that

𝔼N,β[N(xi+kxi)]=𝔼N,β[N(xi+kxi)]+O(1)=k+O(1),\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[N(x_{i+k}-x_{i})]+O(1)=k+O(1), (4.20)

for some O(1)O(1) uniform in NN and KK.

Step 4: fluctuations under N,β\mathbb{Q}_{N,\beta}

We now study the fluctuations of N(xi+kxi)N(x_{i+k}-x_{i}) under N,β\mathbb{Q}_{N,\beta}. Denote

G:XNDNN(xi+kxi).G:X_{N}\in D_{N}\mapsto N(x_{i+k}-x_{i}).

Observe that jIiG=0\sum_{j\in I}\partial_{i}G=0, jG=0\partial_{j}G=0 for each iIci\in I^{c} and sup|G|2=2N2\sup|\nabla G|^{2}=2N^{2}. Moreover N,β\mathbb{Q}_{N,\beta} satisfies Assumptions 3.1 with

χij:x𝕋β(Nsg(x)+θ(NK1+γx)𝟙i,jI).\chi_{ij}:x\in\mathbb{T}\mapsto\beta\Bigr{(}N^{-s}g(x)+\theta\Bigr{(}\frac{N}{K^{1+\gamma}}x\Bigr{)}\mathds{1}_{i,j\in I}\Bigr{)}.

Moreover, for each i,jIi,j\in I and x𝕋x\in\mathbb{T},

χij′′(x)=β(Nsg′′(x)+θ′′(NK1+γx)(NK1+γ)2).\chi_{ij}^{\prime\prime}(x)=\beta\Bigr{(}N^{-s}g^{\prime\prime}(x)+\theta^{\prime\prime}\Bigr{(}\frac{N}{K^{1+\gamma}}x\Bigr{)}\Bigr{(}\frac{N}{K^{1+\gamma}}\Bigr{)}^{2}\Bigr{)}.

If |x|K1+γ|x|\leq K^{1+\gamma}, then one can write

Nsg′′(x)+θ′′(NK1+γx)(NK1+γ)2Nsg′′(x)N2K(s+2)(1+γ).N^{-s}g^{\prime\prime}(x)+\theta^{\prime\prime}\Bigr{(}\frac{N}{K^{1+\gamma}}x\Bigr{)}\Bigr{(}\frac{N}{K^{1+\gamma}}\Bigr{)}^{2}\geq N^{-s}g^{\prime\prime}(x)\geq\frac{N^{2}}{K^{(s+2)(1+\gamma)}}.

If |x|K1+γ|x|\geq K^{1+\gamma}, then

Nsg′′(x)+θ′′(NK1+γx)(NK1+γ)2(NK1+γ)2.N^{-s}g^{\prime\prime}(x)+\theta^{\prime\prime}\Bigr{(}\frac{N}{K^{1+\gamma}}x\Bigr{)}\Bigr{(}\frac{N}{K^{1+\gamma}}\Bigr{)}^{2}\geq\Bigr{(}\frac{N}{K^{1+\gamma}}\Bigr{)}^{2}.

Combining the two last displays we deduce that for each i,jIi,j\in I,

χij′′(x)c1where c1:=βN2K(1+γ)(s+2).\chi_{ij}^{\prime\prime}(x)\geq c_{1}\quad\text{where $c_{1}:=\frac{\beta N^{2}}{K^{(1+\gamma)(s+2)}}$}.

Therefore, by applying Lemma 3.12, we obtain that for all tt\in\mathbb{R},

|log𝔼N,β[etG]t𝔼N,β[G]|t22c1Ksup|G|2Ct2K(1+γ)(s+2)1,|\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{tG}]-t\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G]|\leq\frac{t^{2}}{2c_{1}K}\sup|\nabla G|^{2}\leq Ct^{2}K^{(1+\gamma)(s+2)-1},

for some constant C>0C>0 independent of NN and KK. By Markov’s inequality, recalling that kK1α0k\geq K^{1-\alpha_{0}}, it implies that

N,β(|N(xi+kxi)𝔼N,β[N(xi+kxi)]|k1+δ)Cexp(ck2(1+δ)K(1+γ)(s+2)1)Cexp(ck2(1+δ)(1α0)1((1+γ)(s+2)1)),\mathbb{Q}_{N,\beta}(|N(x_{i+k}-x_{i})-\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})]|\geq k^{1+\delta})\leq C\exp\Bigr{(}-c\frac{k^{2(1+\delta)}}{K^{(1+\gamma)(s+2)-1}}\Bigr{)}\\ \leq C\exp\Bigr{(}-ck^{2(1+\delta)-(1-\alpha_{0})^{-1}((1+\gamma)(s+2)-1)}\Bigr{)},

for some constants c>0,C>0c>0,C>0 independent of NN and KK. Using (4.20), this gives

N,β(|N(xi+kxi)k|k1+δ)Cexp(ck2(1+δ)(1α0)1((1+γ)(s+2)1)),\mathbb{Q}_{N,\beta}(|N(x_{i+k}-x_{i})-k|\geq k^{1+\delta})\leq C\exp\Bigr{(}-ck^{2(1+\delta)-(1-\alpha_{0})^{-1}((1+\gamma)(s+2)-1)}\Bigr{)}, (4.21)

for some constants c>0,C>0c>0,C>0 independent of NN and KK.

Step 5: conclusion

We have

2(1+δ)11α0((1+γ)(s+2)1)>2δγ<2(1α0)(1+s)2+s.2(1+\delta)-\frac{1}{1-\alpha_{0}}((1+\gamma)(s+2)-1)>2\delta\Longleftrightarrow\gamma<\frac{2(1-\alpha_{0})-(1+s)}{2+s}. (4.22)

Observe that the conditions (4.9) and (4.22) can be satisfied if and only if

(1α0)δ<2(1α0)(1+s)2+s1α0>1+s2δ(s+2).(1-\alpha_{0})\delta<\frac{2(1-\alpha_{0})-(1+s)}{2+s}\Longleftrightarrow 1-\alpha_{0}>\frac{1+s}{2-\delta(s+2)}. (4.23)

Therefore by (4.8) we can choose γ(δ(1α0),δ0)\gamma\in(\delta(1-\alpha_{0}),\delta_{0}) satisfying (4.22). Therefore by (4.20) and (4.21), there exists δ>δ\delta^{\prime}>\delta and C>0C>0, c>0c>0 independent of NN and KK such that

N,β(N(xi+kxi)k1+δ)Ceck2δ.\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq Ce^{-ck^{2\delta^{\prime}}}.

In combination with (4.14), we deduce that there exists k0k_{0} independent of NN and KK such that for each kmax(k0,K)k\geq\max(k_{0},K),

N,β(N(xi+kxi)k1+δ)14C0ec0k2δ.\mathbb{Q}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq\frac{1}{4}C_{0}e^{-c_{0}k^{2\delta}}. (4.24)

In combination with (4.14) this implies the existence of k0k_{0} independent of KK such that for each kmax(K,k0)k\geq\max(K,k_{0}),

N,β(N(xi+kxi)k1+δ)12C0ec0k2δ.\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq\frac{1}{2}C_{0}e^{-c_{0}k^{2\delta}}.

Let δ>δ0\delta>\delta_{0} and L:=k1+δ1+δ0KL:=\lfloor k^{\frac{1+\delta}{1+\delta_{0}}}\rfloor\geq K. We have

N,β(N(xi+kxi)k1+δ)N,β(N(xi+Lxi)k1+δ)N,β(N(xi+Lxi)L1+δ0)12C0ec0L2δ0C0ec0k2δ0(1+δ1+δ0)2,\mathbb{P}_{N,\beta}(N(x_{i+k}-x_{i})\geq k^{1+\delta})\leq\mathbb{P}_{N,\beta}(N(x_{i+L}-x_{i})\geq k^{1+\delta})\\ \leq\mathbb{P}_{N,\beta}(N(x_{i+L}-x_{i})\geq L^{1+\delta_{0}})\leq\frac{1}{2}C_{0}e^{-c_{0}L^{2\delta_{0}}}\leq C_{0}e^{-c_{0}k^{2\delta_{0}(\frac{1+\delta}{1+\delta_{0}})^{2}}},

if kmax(k0,K)k\geq\max(k_{0},K) for some k0k_{0} independent of KK. We conclude that there exists k0k_{0} independent of KK such that (4.7) holds for each kk0k\geq k_{0}. At the cost of changing C0C_{0}, we find that (4.7) also holds for each k1k\geq 1. ∎

From the proof of Lemma 4.2 we deduce the following estimate on the expectation of gaps under the locally constrained measure:

Lemma 4.3.

Let L{1,,N/2}L\in\{1,\ldots,N/2\} and i{1,,N}i\in\{1,\ldots,N\}. Let

I:={iL,,i+L}.I:=\{i-L,\ldots,i+L\}.

Let jIj\in I and k{1,,N2}k\in\{1,\ldots,\frac{N}{2}\} such that j+kIj+k\in I. Let γ>0\gamma>0. Let N,β\mathbb{Q}_{N,\beta} be the locally constrained measure (4.2) with K:=L1+γK:=L^{1+\gamma}. There exists C>0C>0 depending on γ\gamma and independent of NN, LL and kk such that

|𝔼N,β[N(xi+kxi)]k|C.|\mathbb{E}_{\mathbb{Q}_{N,\beta}}[N(x_{i+k}-x_{i})]-k|\leq C.
Proof.

We have proved in Lemma 4.2 that N,β\mathbb{P}_{N,\beta} satisfies (4.7). We thus conclude by (4.20). ∎

4.3. Reduction to a block average

In this subsection we implement a method developed in [BEY12] to study the fluctuations of particles positions. The strategy consists in replacing a single point xix_{i} by the average of the xjsx_{j}^{\prime}s over a certain block centered around xix_{i}.

For each i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2}, let Ik(i)I_{k}(i) stand for the interval of indices

Ik(i):={j{1,,N}:d(i,j)k}.I_{k}(i):=\{j\in\{1,\ldots,N\}:d(i,j)\leq k\}.

Define the block average

xi[k]:=12k+1jIk(i)xj.x_{i}^{[k]}:=\frac{1}{2k+1}\sum_{j\in I_{k}(i)}x_{j}. (4.25)
Lemma 4.4 (Comparison to a block average).

Let ε>0\varepsilon>0 be small enough. There exist C>0C>0 and c>0c>0 independent of NN such that for each i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2},

N,β(|N(xixi[k])|ks2+ε)Ceckε4(s+2).\mathbb{P}_{N,\beta}(|N(x_{i}-x_{i}^{[k]})|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\frac{\varepsilon}{4(s+2)}}}.
Proof.

Let i{1,,N}i\in\{1,\ldots,N\} and 1kN21\leq k\leq\frac{N}{2}. Fix ε>0\varepsilon>0 such that 2(1+s)ε\frac{2(1+s)}{\varepsilon}\in\mathbb{N}^{*}. Let p1p\geq 1 be a large number and α:=1p\alpha:=\frac{1}{p}. Since xi[0]=xix_{i}^{[0]}=x_{i}, one can break N(xixi[k])N(x_{i}-x_{i}^{[k]}) into

N(xixi[k])=m=0p1N(xi[kmα]xi[k(m+1)α]).N(x_{i}-x_{i}^{[k]})=\sum_{m=0}^{p-1}N(x_{i}^{[\lfloor k^{m\alpha}\rfloor]}-x_{i}^{[\lfloor k^{(m+1)\alpha}\rfloor]}).

For each m{0,,p1}m\in\{0,\ldots,p-1\}, denote

Gm:=N(xi[kmα]xi[k(m+1)α])andIm:=Ik(m+1)α(i).G_{m}:=N(x_{i}^{[\lfloor k^{m\alpha}\rfloor]}-x_{i}^{[\lfloor k^{(m+1)\alpha}\rfloor]})\quad\text{and}\quad I_{m}:=I_{\lfloor k^{(m+1)\alpha}\rfloor}(i).

The function GmG_{m} only depends on the variables (xj)jIm(x_{j})_{j\in I_{m}} and jImjGm=0\sum_{j\in I_{m}}\partial_{j}G_{m}=0. Let ε>0\varepsilon^{\prime}>0. Let N,β\mathbb{Q}_{N,\beta} be the constrained Gibbs measure (4.2) with I=ImI=I_{m} and K=k(m+1)α+εK=k^{(m+1)\alpha+\varepsilon^{\prime}}. The measure N,β\mathbb{Q}_{N,\beta} satisfies the assumptions of Lemma 3.12 with c1:=βN2k((m+1)α+ε)(s+2)c_{1}:=\beta N^{2}k^{-((m+1)\alpha+\varepsilon^{\prime})(s+2)}. Moreover sup|Gm|2=O(N2kαm)\sup|\nabla G_{m}|^{2}=O(N^{2}k^{-\alpha m}). Thus, by Lemma 3.12, for all tt\in\mathbb{R}, we have

|log𝔼N,β[etGm]t𝔼N,β[Gm]|Ct2kαms+α(1+s)+ε(s+2)Ct2ks+α(1+s)+ε(s+2),|\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{tG_{m}}]-t\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G_{m}]|\leq Ct^{2}k^{\alpha ms+\alpha(1+s)+\varepsilon^{\prime}(s+2)}\leq Ct^{2}k^{s+\alpha(1+s)+\varepsilon^{\prime}(s+2)},

since αmp1p1\alpha m\leq\frac{p-1}{p}\leq 1. Let us now choose p=2(1+s)εp=\frac{2(1+s)}{\varepsilon}\in\mathbb{N}^{*} and ε=ε4(s+2)\varepsilon^{\prime}=\frac{\varepsilon}{4(s+2)} so that

α(1+s)+ε(s+2)=ε2.\alpha(1+s)+\varepsilon^{\prime}(s+2)=\frac{\varepsilon}{2}.

We get

N,β(|Gm𝔼N,β[Gm]|ks2+ε)Ceckε.\mathbb{Q}_{N,\beta}(|G_{m}-\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G_{m}]|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\varepsilon}}. (4.26)

Inserting the result of Lemma 4.3 and (4.13), we get

𝔼N,β[Gm]=O(1),\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G_{m}]=O(1), (4.27)

for some O(1)O(1) depending on ε\varepsilon and α\alpha. Consequently by (4.26) and (4.27), one has

N,β(|Gm|ks2+ε)Ceckε.\mathbb{Q}_{N,\beta}(|G_{m}|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\varepsilon}}.

Meanwhile by Lemma 4.1 and Lemma 4.2, one obtains that for ε>0\varepsilon^{\prime}>0 small enough,

TV(πIm#N,β,πIm#N,β)Ceckε.\mathrm{TV}(\pi_{I_{m}}\#\mathbb{P}_{N,\beta},\pi_{I_{m}}\#\mathbb{Q}_{N,\beta})\leq Ce^{-ck^{\varepsilon^{\prime}}}.

Since εε\varepsilon^{\prime}\leq\varepsilon, one finds

N,β(|Gm|ks2+ε)Ceckε4(s+2).\mathbb{P}_{N,\beta}(|G_{m}|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\frac{\varepsilon}{4(s+2)}}}.

We conclude that there exist constants C>0,c>0C>0,c>0 depending on ε\varepsilon such that

N,β(m=0p1|Gm|α1kε)Ceckε4(s+2).\mathbb{P}_{N,\beta}\Bigr{(}\sum_{m=0}^{p-1}|G_{m}|\geq\alpha^{-1}k^{\varepsilon}\Bigr{)}\leq Ce^{-ck^{\frac{\varepsilon}{4(s+2)}}}.

This concludes the proof of Lemma 4.4. ∎

4.4. Proof of Theorem 1

Theorem 1 quickly follows from Lemma 4.4. We study the fluctuations of N(xi+kxi)N(x_{i+k}-x_{i}). By Lemma 4.4, one can replace xix_{i} and xi+kx_{i+k} by their block average at scale kk, up to a well-controlled error. Moreover, the difference of these block averages can be easily bounded using Lemma 3.12.

Proof of Theorem 1.

Let i{1,,N}i\in\{1,\ldots,N\}, 1kN21\leq k\leq\frac{N}{2} and ε>0\varepsilon>0. Let us split the gap N(xi+kxi)N(x_{i+k}-x_{i}) into

N(xi+kxi)=N(xi+kxi+k[k])N(xixi[k])+N(xi+k[k]xi[k]).N(x_{i+k}-x_{i})=N(x_{i+k}-x_{i+k}^{[k]})-N(x_{i}-x_{i}^{[k]})+N(x_{i+k}^{[k]}-x_{i}^{[k]}). (4.28)

By Lemma 4.4, letting δ=ε4(s+2)\delta=\frac{\varepsilon}{4(s+2)}, we have

N,β(|N(xi[k]xi)|ks2+ε)Ceckδ,\mathbb{P}_{N,\beta}(|N(x_{i}^{[k]}-x_{i})|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\delta}}, (4.29)
N,β(|N(xi+k[k]xi+k)|ks2+ε)Ceckδ,\mathbb{P}_{N,\beta}(|N(x_{i+k}^{[k]}-x_{i+k})|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\delta}}, (4.30)

Let us define

G:XNDNN(xi+k[k]xi[k]).G:X_{N}\in D_{N}\mapsto N(x_{i+k}^{[k]}-x_{i}^{[k]}).

Let N,β\mathbb{Q}_{N,\beta} be the constrained Gibbs measure (4.1) with I:={j:d(i,j)k}I:=\{j:d(i,j)\leq k\} and K:=k(1+γ)K:=k^{(1+\gamma)} for some γ>0\gamma>0 to fix later. We have sup|G|2=O(N2k)\sup|\nabla G|^{2}=O(\frac{N^{2}}{k}). Moreover N,β\mathbb{Q}_{N,\beta} satisfies the assumptions of Lemma 3.12 with c1:=βN2k(1+γ)(s+2)c_{1}:=\beta N^{2}k^{-(1+\gamma)(s+2)}. Consequently one gets from 3.12 that for all tt\in\mathbb{R},

log𝔼N,β[etG]=t𝔼N,β[G]+O(t2k(1+γ)(s+2)2).\log\mathbb{E}_{\mathbb{Q}_{N,\beta}}[e^{tG}]=t\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G]+O(t^{2}k^{(1+\gamma)(s+2)-2}). (4.31)

Inserting the accuracy estimate of Lemma 4.3 we find

𝔼N,β[G]=O(1).\mathbb{E}_{\mathbb{Q}_{N,\beta}}[G]=O(1). (4.32)

Fix γ=εs+2\gamma=\frac{\varepsilon}{s+2}. By (4.31) and (4.32) one finds

N,β(|G|ks2+ε)Ceckε.\mathbb{Q}_{N,\beta}(|G|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\varepsilon}}. (4.33)

Again, by Lemmas 4.1, 4.2 and (4.13), one has

TV(πI#N,β,πI#N,β)kκ𝔼N,β[(N(xi+kxi))2𝟙N(xi+kxi)k1+γ]kκj=k1+γj2ecj2γ1+γCeckεs+2.\mathrm{TV}(\pi_{I}\#\mathbb{P}_{N,\beta},\pi_{I}\#\mathbb{Q}_{N,\beta})\leq k^{\kappa}\mathbb{E}_{\mathbb{P}_{N,\beta}}[(N(x_{i+k}-x_{i}))^{2}\mathds{1}_{N(x_{i+k}-x_{i})\geq k^{1+\gamma}}]\\ \leq k^{\kappa}\sum_{j=k^{1+\gamma}}j^{2}e^{-cj^{\frac{2\gamma}{1+\gamma}}}\leq Ce^{-ck^{\frac{\varepsilon}{s+2}}}. (4.34)

Combining (4.33) and (4.34) one deduces that

N,β(|G|ks2+ε)Ceckεs+2.\mathbb{Q}_{N,\beta}(|G|\geq k^{\frac{s}{2}+\varepsilon})\leq Ce^{-ck^{\frac{\varepsilon}{s+2}}}.

Together with (4.29) and (4.30), this proves (1.6). Since for each i{1,,N}i\in\{1,\ldots,N\}, xix_{i} is uniformly distributed on 𝕋\mathbb{T}, one easily concludes the proof of (1.7). ∎

4.5. Control of the probability of near collisions

Let us control the probability of having two particles very close to each other.

Lemma 4.5.

Let α(0,s2)\alpha\in(0,\frac{s}{2}). There exist C>0,ε0>0C>0,\varepsilon_{0}>0 independent of NN such that for each i{1,,N}i\in\{1,\ldots,N\} and ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}),

N,β(N(xi+1xi)ε)Ceεα.\mathbb{P}_{N,\beta}(N(x_{i+1}-x_{i})\leq\varepsilon)\leq Ce^{-\varepsilon^{-\alpha}}.
Proof.

Let η:(0,N)+\eta:(0,N)\to\mathbb{R}^{+} be a smooth cutoff function independent of NN such that

η(x)={1if x[0,1]0if x[2,N].\eta(x)=\begin{cases}1&\text{if $x\in[0,1]$}\\ 0&\text{if $x\in[2,N]$}.\end{cases}

Let α(0,s2)\alpha\in(0,\frac{s}{2}). Let ξ:+\xi:\mathbb{R}^{+}\to\mathbb{R} smooth such that ξ(x)=η(x)|x|α\xi(x)=\eta(x)|x|^{-\alpha} for x(0,2)x\in(0,2), ξ\xi supported on [0,4][0,4] and such that 𝔼N,β[ξ(N(x2x1))]=0\mathbb{E}_{\mathbb{P}_{N,\beta}}[\xi(N(x_{2}-x_{1}))]=0. Define

F:XNDNξ(N(x2x1)).F:X_{N}\in D_{N}\mapsto\xi(N(x_{2}-x_{1})).
Step 1: starting point

Let μ\mu be the law of (x1,x2)(x_{1},x_{2}) under N,β\mathbb{P}_{N,\beta}, which can be written

dμ(x1,x2)eSN(x1,x2)𝟙D2(x1,x2)dx1dx2,\mathrm{d}\mu(x_{1},x_{2})\propto e^{-S_{N}(x_{1},x_{2})}\mathds{1}_{D_{2}}(x_{1},x_{2})\mathrm{d}x_{1}\mathrm{d}x_{2},

where

SN:(x1,x2)D2eβN(x1,x2,,xN)dx3dxN.S_{N}:(x_{1},x_{2})\in D_{2}\mapsto-\int e^{-\beta\mathcal{H}_{N}(x_{1},x_{2},\ldots,x_{N})}\mathrm{d}x_{3}\ldots\mathrm{d}x_{N}.

Let us define

GN:x(0,N)logeβN(0,x/N,x2,,xN)𝟙x/N<x3<xN<1dx3dxN.G_{N}:x\in(0,N)\mapsto-\log\int e^{-\beta\mathcal{H}_{N}(0,x/N,x_{2},\ldots,x_{N})}\mathds{1}_{x/N<x_{3}<\ldots x_{N}<1}\mathrm{d}x_{3}\ldots\mathrm{d}x_{N}.

By rotational invariance, SN(x1,x2)=GN(N(x2x1))S_{N}(x_{1},x_{2})=G_{N}(N(x_{2}-x_{1})). Consider the solution ΦL2({1,2},H1(μ))\nabla\Phi\in L^{2}(\{1,2\},H^{1}(\mu)) of

ΔΦ+SNΦ=F.-\Delta\Phi+\nabla S_{N}\cdot\nabla\Phi=F. (4.35)

Since SNS_{N} and FF depend only on x2x1x_{2}-x_{1}, Φ\Phi is also a function of x2x1x_{2}-x_{1} and one may write it as Φ(x1,x2)=1N2ϕ(N(x2x1))\Phi(x_{1},x_{2})=\frac{1}{N^{2}}\phi(N(x_{2}-x_{1})) for some ϕ:(0,N)\phi:(0,N)\to\mathbb{R}. Let ψ:=ϕ\psi:=\phi^{\prime}, which solves

ψ+GNψ=ξ2.-\psi^{\prime}+G_{N}^{\prime}\psi=\frac{\xi}{2}. (4.36)
Step 2: estimates on GNG_{N}

One may split GNG_{N} into

GN=Nsβg+βJNG_{N}=N^{-s}\beta g+\beta J_{N}

where

JN:x(0,N)1βlogeβN(0,x/N,x2,,xN)𝟙x/N<x3<<xN<1dx3dxN,J_{N}:x\in(0,N)\mapsto-\frac{1}{\beta}\log\int e^{-\beta\mathcal{H}_{N}^{*}(0,x/N,x_{2},\ldots,x_{N})}\mathds{1}_{x/N<x_{3}<\ldots<x_{N}<1}\mathrm{d}x_{3}\ldots\mathrm{d}x_{N},

where

N:XNDNNsij:(i,j)(1,2)g(xixj).\mathcal{H}_{N}^{*}:X_{N}\in D_{N}\mapsto N^{-s}\sum_{i\neq j:(i,j)\neq(1,2)}g(x_{i}-x_{j}).

Since limx0g(x)=+\lim_{x\to 0}g^{\prime}(x)=+\infty, for all x(0,N)x\in(0,N),

JN(x)=1N𝔼(x)[2N(0,x/N,x3,,xN)],J_{N}^{\prime}(x)=\frac{1}{N}\mathbb{E}_{\mathbb{P}(x)}[\partial_{2}\mathcal{H}_{N}^{*}(0,x/N,x_{3},\ldots,x_{N})],

where (x)\mathbb{P}(x) is the probability measure with density

d(x)(x3,,xN)eβN(0,x/N,x3,,xN)𝟙x/N<x3<xN<1dx3dxN.\mathrm{d}\mathbb{P}(x)(x_{3},\ldots,x_{N})\propto e^{-\beta\mathcal{H}_{N}^{*}(0,x/N,x_{3},\ldots,x_{N})}\mathds{1}_{x/N<x_{3}<\ldots x_{N}<1}\mathrm{d}x_{3}\ldots\mathrm{d}x_{N}.

We then expand

JN(x)=N(1+s)βk{1,2}𝔼N,β[g(xkx/N)x1=0,x2=x/N].J_{N}^{\prime}(x)=N^{-(1+s)}\beta\sum_{k\neq\{1,2\}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[g^{\prime}(x_{k}-x/N)\mid x_{1}=0,x_{2}=x/N]. (4.37)

Using that for each k=1,,N2k=1,\ldots,\lfloor\frac{N}{2}\rfloor,

𝔼N,β[g(x2+kx/N)x1=0,x2=x/N]𝔼N,β[g(x1kx/N)x1=0,x2=x/N]0,\mathbb{E}_{\mathbb{P}_{N,\beta}}[g^{\prime}(x_{2+k}-x/N)\mid x_{1}=0,x_{2}=x/N]-\mathbb{E}_{\mathbb{P}_{N,\beta}}[g^{\prime}(x_{1-k}-x/N)\mid x_{1}=0,x_{2}=x/N]\leq 0,

we deduce that

JN0.J_{N}^{\prime}\leq 0. (4.38)

Recall that g<0g^{\prime}<0 on (0,12)(0,\frac{1}{2}) and g>0g^{\prime}>0 on (12,1)(\frac{1}{2},1). It follows that GN<0G_{N}^{\prime}<0 on (0,12)(0,\frac{1}{2}).

Following the proof of the convexity result of Lemma 3.10, we have JN′′0J_{N}^{\prime\prime}\geq 0. Indeed, for all x(0,N)x\in(0,N),

JN′′(x)=1N2𝔼(x)[22N(0,x/N,)]βN2Var(x)[2N(0,x/N,)].J_{N}^{\prime\prime}(x)=\frac{1}{N^{2}}\mathbb{E}_{\mathbb{P}(x)}[\partial_{22}\mathcal{H}_{N}^{*}(0,x/N,\cdot)]-\frac{\beta}{N^{2}}\mathrm{Var}_{\mathbb{P}(x)}[\partial_{2}\mathcal{H}_{N}^{*}(0,x/N,\cdot)].

Let us denote 3:N2N=(ijN)3i,jN\nabla_{3:N}^{2}\mathcal{H}_{N}^{*}=(\partial_{ij}\mathcal{H}_{N}^{*})_{3\leq i,j\leq N}. Using the Brascamp-Lieb inequality,

βVar(x)[2N(0,x/N,)]𝔼(x)[3:N,2N(3:N2N)13:N,2N(0,x/N,)].\beta\mathrm{Var}_{\mathbb{P}(x)}[\partial_{2}\mathcal{H}_{N}^{*}(0,x/N,\cdot)]\leq\mathbb{E}_{\mathbb{P}(x)}[\nabla_{3:N,2}\mathcal{H}_{N}^{*}(\nabla_{3:N}^{2}\mathcal{H}_{N}^{*})^{-1}\nabla_{3:N,2}\mathcal{H}_{N}^{*}(0,x/N,\cdot)].

Therefore for all x(0,N)x\in(0,N),

JN′′(x)βN2𝔼(x)[(22N3:N,2N(3:N2N)13:N,2N)(0,x/N,x3,,xN)]0.J_{N}^{\prime\prime}(x)\geq\frac{\beta}{N^{2}}\mathbb{E}_{\mathbb{P}}(x)[(\partial_{22}\mathcal{H}_{N}^{*}-\nabla_{3:N,2}\mathcal{H}_{N}^{*}(\nabla_{3:N}^{2}\mathcal{H}_{N}^{*})^{-1}\nabla_{3:N,2}\mathcal{H}_{N}^{*})(0,x/N,x_{3},\ldots,x_{N})]\geq 0.

Since 2:N2N(0,)\nabla_{2:N}^{2}\mathcal{H}_{N}^{*}(0,\cdot) is non-negative the Schur-complement

(22N3:N,2N(3:N2N)13:N,2N)(0,x/N,)(\partial_{22}\mathcal{H}_{N}^{*}-\nabla_{3:N,2}\mathcal{H}_{N}^{*}(\nabla_{3:N}^{2}\mathcal{H}_{N}^{*})^{-1}\nabla_{3:N,2}\mathcal{H}_{N}^{*})(0,x/N,\cdot)

is also non-negative and therefore JN′′0J_{N}^{\prime\prime}\geq 0.

Step 3: study of the one-point transport

One can solve (4.36) explicitly: for all x(0,N)x\in(0,N), we have

ψ(x)=12eGN(x)0xeGN(y)ξ(y)dy.\psi(x)=-\frac{1}{2}e^{G_{N}(x)}\int_{0}^{x}e^{-G_{N}(y)}\xi(y)\mathrm{d}y.

In view of (4.38), for all y[0,x]y\in[0,x], JN(x)JN(y)J_{N}(x)\leq J_{N}(y). Therefore for all x(0,N)x\in(0,N),

|ψ|(x)12eβ(g(x)g(y))|ξ(y)|dy.|\psi|(x)\leq\frac{1}{2}\int e^{\beta(g(x)-g(y))}|\xi(y)|\mathrm{d}y.

From this we get that there exists C0>0C_{0}>0 independent of NN such that for all x(0,N)x\in(0,N)

|ψ(x)|C0|x|1+sα.|\psi(x)|\leq C_{0}|x|^{1+s-\alpha}. (4.39)

Let us now derive an estimate for ψ\psi^{\prime}. Recall that GN|(0,4)G_{N}|_{(0,4)} is strictly decreasing. Consider its inverse GN1:(GN(4),+)(0,4)G_{N}^{-1}:(G_{N}(4),+\infty)\to(0,4). Using the change of variables y=GN1(x)y=G_{N}^{-1}(x), we obtain that for all x(0,4)x\in(0,4),

ψ(x)=12eGN(x)GN(x)+ey(ξGN)GN1(y)dy.\psi(x)=\frac{1}{2}e^{-G_{N}(x)}\int_{G_{N}(x)}^{+\infty}e^{-y}\Bigr{(}\frac{\xi}{G_{N}^{\prime}}\Bigr{)}\circ G_{N}^{-1}(y)\mathrm{d}y.

We then get by integration by parts that for all x(0,4)x\in(0,4),

ψ(x)=12ξ(x)GN(x)+12eGN(x)GN(x)+ey(ξGN)GN11GNGN1(y)dy.\psi(x)=\frac{1}{2}\frac{\xi(x)}{G_{N}^{\prime}(x)}+\frac{1}{2}e^{-G_{N}(x)}\int_{G_{N}(x)}^{+\infty}e^{-y}\Bigr{(}\frac{\xi}{G_{N}^{\prime}}\Bigr{)}^{\prime}\circ G_{N}^{-1}\frac{1}{G_{N}^{\prime}\circ G_{N}^{-1}}(y)\mathrm{d}y.

Inserting this into (4.36) we deduce that for all x(0,4),x\in(0,4),

ψ(x)=12eGN(x)GN(x)GN(x)+ey(ξGN)GN11GNGN1(y)dy.\psi^{\prime}(x)=\frac{1}{2}e^{-G_{N}(x)}G_{N}^{\prime}(x)\int_{G_{N}(x)}^{+\infty}e^{-y}\Bigr{(}\frac{\xi}{G_{N}^{\prime}}\Bigr{)}^{\prime}\circ G_{N}^{-1}\frac{1}{G_{N}^{\prime}\circ G_{N}^{-1}}(y)\mathrm{d}y.

We have

(ξGN)=ξGNξGN′′(GN)2.\Bigr{(}\frac{\xi}{G_{N}^{\prime}}\Bigr{)}^{\prime}=\frac{\xi^{\prime}}{G_{N}^{\prime}}-\frac{\xi G_{N}^{\prime\prime}}{(G_{N}^{\prime})^{2}}.

Note that since JN′′0J_{N}^{\prime\prime}\geq 0 and JN0J_{N}^{\prime}\leq 0, we have |GN(z)|β|g(z)||G_{N}^{\prime}(z)|\geq\beta|g^{\prime}(z)|, GN′′(z)βg′′(z)0G_{N}^{\prime\prime}(z)\geq\beta g^{\prime\prime}(z)\geq 0, |GN1(z)||(βg)1(z)||G_{N}^{-1}(z)|\leq|(\beta g)^{-1}(z)| on (0,N)(0,N). Thus there exists K>0K>0 independent of NN such that for all y(0,4)y\in(0,4),

|(ξGN)|(y)K|y|sα.\Bigr{|}\Bigr{(}\frac{\xi}{G_{N}}\Bigr{)}^{\prime}\Bigr{|}(y)\leq K|y|^{s-\alpha}.

and for all y>GN1(4)y>G_{N}^{-1}(4),

|1GNGN1|(y)K|y|11s.\Bigr{|}\frac{1}{G_{N}^{\prime}\circ G_{N}^{-1}}\Bigr{|}(y)\leq K|y|^{-1-\frac{1}{s}}.

Thus for all x(0,4)x\in(0,4),

|ψ(x)|KeGN(x)|GN(x)|GN(x)+ey|y|αs1s2K|GN(x)||x|2s+1αK|x|sα.|\psi^{\prime}(x)|\leq Ke^{-G_{N}(x)}|G_{N}^{\prime}(x)|\int_{G_{N}(x)}^{+\infty}e^{-y}|y|^{\frac{\alpha}{s}-\frac{1}{s}-2}\leq K|G_{N}^{\prime}(x)||x|^{2s+1-\alpha}\leq K|x|^{s-\alpha}. (4.40)

For x4x\geq 4, since ξ\xi is supported on (0,4)(0,4), we have

ψ(x)=12eGN(x)04eGN(y)ξ(y)dy\psi(x)=-\frac{1}{2}e^{G_{N}(x)}\int_{0}^{4}e^{-G_{N}(y)}\xi(y)\mathrm{d}y

It follows that ψ(N)=0\psi(N)=0. Moreover

ψ(x)=12GN(x)eGN(x)04eGN(y)ξ(y)dy.\psi^{\prime}(x)=-\frac{1}{2}G_{N}^{\prime}(x)e^{-G_{N}(x)}\int_{0}^{4}e^{-G_{N}(y)}\xi(y)\mathrm{d}y.

Since GN0G_{N}^{\prime}\leq 0, we get combining the last display and (4.40) that there exists K0>0K_{0}>0 independent of NN such that

sup(0,N)min(ψ,0)K0.\sup_{(0,N)}\min(\psi^{\prime},0)\leq K_{0}. (4.41)
Step 5: Gaussian estimate

Let us prove a Gaussian concentration estimate for FF. Let t(0,12K0)t\in(0,\frac{1}{2K_{0}}). Let Φ\Phi be as in (4.35). Consider the map

(x1,x2)Id+tΦ(x1,x2).(x_{1},x_{2})\mapsto\mathrm{Id}+t\nabla\Phi(x_{1},x_{2}).

By (4.39), we have

det(D(Id+tΦ(x1,x2)))=1+2tψ(x2x1)>0.\det(D(\mathrm{Id}+t\nabla\Phi(x_{1},x_{2})))=1+2t\psi^{\prime}(x_{2}-x_{1})>0.

Moreover since 1+tψ(x2x1)>01+t\psi^{\prime}(x_{2}-x_{1})>0, we deduce that Id+tD2Φ(x1,x2)>0\mathrm{Id}+tD^{2}\Phi(x_{1},x_{2})>0. Using the fact that ψ(0)=ψ(N)=0\psi(0)=\psi(N)=0, one gets that Id+tΦ(x1,x2)\mathrm{Id}+t\nabla\Phi(x_{1},x_{2}) defines a valid change of variables on D2D_{2}.

Therefore one may rewrite the Laplace transform of ξ\xi under μ\mu as

𝔼μ[exp(tξ)]=𝔼μ[exp(tF(Id+tΦ)GN(Id+tΦ)+GN+logdet(Id+tD2Φ))].\mathbb{E}_{\mu}[\exp({t\xi})]=\mathbb{E}_{\mu}\Bigr{[}\exp\Bigr{(}tF\circ(\mathrm{Id}+t\nabla\Phi)-G_{N}\circ(\mathrm{Id}+t\nabla\Phi)+G_{N}+\log\det(\mathrm{Id}+tD^{2}\Phi)\Bigr{)}\Bigr{]}.

Since GN′′0G_{N}^{\prime\prime}\geq 0, we have

GN(Id+tΦ)GNtGNΦ.G_{N}\circ(\mathrm{Id}+t\nabla\Phi)-G_{N}\leq t\nabla G_{N}\cdot\nabla\Phi.

Define

ϕ:t(0,12K0)logdet(Id+tD2ϕ).\phi:t\in\Bigr{(}0,\frac{1}{2K_{0}}\Bigr{)}\mapsto\log\det(\mathrm{Id}+tD^{2}\phi).

One can check that for all t(0,12K0)t\in\Bigr{(}0,\frac{1}{2K_{0}}\Bigr{)},

ϕ′′(t)=tr((Id+tD2ϕ)1(D2ϕ)(Id+tD2ϕ)1D2ϕ)0,\phi^{\prime\prime}(t)=-\mathrm{tr}((\mathrm{Id}+tD^{2}\phi)^{-1}(D^{2}\phi)(\mathrm{Id}+tD^{2}\phi)^{-1}D^{2}\phi)\leq 0,

since Id+tD2ϕ>0\mathrm{Id}+tD^{2}\phi>0. It follows that

log(Id+tD2ϕ)tΔΦ.\log(\mathrm{Id}+tD^{2}\phi)\leq t\Delta\Phi.

Then, since ξ\xi is supported on [0,4][0,4], one can notice that there exists C>0C>0 independent of NN such that for all x𝕋x\in\mathbb{T},

|F(Id+tΦ)F|(x)Ct2supy(0,4)|ψ(y)||y|1+sα.|F\circ(\mathrm{Id}+t\nabla\Phi)-F|(x)\leq Ct^{2}\sup_{y\in(0,4)}\frac{|\psi^{\prime}(y)|}{|y|^{1+s-\alpha}}.

Now, inserting the estimate (4.39), there exists C0>0C_{0}>0 independent of NN such that

|F(Id+tΦ)tξ|C0t2.|F\circ(\mathrm{Id}+t\nabla\Phi)-t\xi|_{\infty}\leq C_{0}t^{2}.

Combining the above gives

tF(Id+tΦ)GN(Id+tΦ)+GN+log(Id+tD2Φ)C0t2+tFdμ+t(FFdμGNΦ+ΔΦ)=C0t2+tFdμ,tF\circ(\mathrm{Id}+t\nabla\Phi)-G_{N}\circ(\mathrm{Id}+t\nabla\Phi)+G_{N}+\log(\mathrm{Id}+tD^{2}\Phi)\\ \leq C_{0}t^{2}+t\int F\mathrm{d}\mu+t\Bigr{(}F-\int F\mathrm{d}\mu-\nabla G_{N}\cdot\nabla\Phi+\Delta\Phi\Bigr{)}=C_{0}t^{2}+t\int F\mathrm{d}\mu,

since Φ\nabla\Phi solves (4.35). Thus there exists C0>0C_{0}>0 such that for all t(0,12K0)t\in(0,\frac{1}{2K_{0}}),

log𝔼μ[etF]t𝔼μ[F]C0t2.\log\mathbb{E}_{\mu}[e^{tF}]-t\mathbb{E}_{\mu}[F]\leq C_{0}t^{2}. (4.42)
Step 5: conclusion

To control the expectation, one may write by symmetry

𝔼μ[F]=1N𝔼N,β[i=1Nξ(N(xi+1xi))].\mathbb{E}_{\mu}[F]=\frac{1}{N}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\xi(N(x_{i+1}-x_{i}))\Bigr{]}.

Notice that

|𝔼[i=1Nξ(N(xi+1xi))]|C(N+𝔼[i=1NNsg(xi+1xi)]).\Bigr{|}\mathbb{E}\Bigr{[}\sum_{i=1}^{N}\xi(N(x_{i+1}-x_{i}))\Bigr{]}\Bigr{|}\leq C\Bigr{(}N+\mathbb{E}\Bigr{[}\sum_{i=1}^{N}N^{-s}g(x_{i+1}-x_{i})\Bigr{]}\Bigr{)}.

We prove in Lemma B.7 that

𝔼[i=1NNsg(xi+1xi)]=O(N).\mathbb{E}\Bigr{[}\sum_{i=1}^{N}N^{-s}g(x_{i+1}-x_{i})\Bigr{]}=O(N). (4.43)

The proof of Lemma 4.5 then follows from (4.42), (4.43) and Markov’s inequality. ∎

5. Optimal rigidity for singular linear statistics

In this section, we prove the optimal variance estimate stated in Theorem 2. This will give the optimal scaling of the fluctuations of gaps and discrepancies.

5.1. Mean-field transport

We first introduce the transportation argument of [J+98, Shc13] which is the starting point of many CLTs on β\beta-ensembles and Coulomb gases including [LS18, BLS+18, Leb18, Ser20]. The method consists in identifying a mean-field approximation of the solution Φ\nabla\Phi of the equation

ΔΦ+βNΦ=F𝔼N,β[F]-\Delta\Phi+\beta\nabla\mathcal{H}_{N}\cdot\nabla\Phi=F-\mathbb{E}_{\mathbb{P}_{N,\beta}}[F]

when FF is a linear statistic and then using this to give an expansion for the variance of FF. This mean-field transport creates a non-local error term, sometimes called the “loop equation term”, defined for all measurable maps ψ:N1𝕋\psi:\ell_{N}^{-1}\mathbb{T}\to\mathbb{R} by

AN[ψ]=ΔcNN(ψ(N1x)ψ(N1y))N(1+s)g(xy)dfluctN(x)dfluctN(y),\mathrm{A}_{\ell_{N}}[\psi]=\iint_{\Delta^{c}}N\ell_{N}(\psi(\ell_{N}^{-1}x)-\psi(\ell_{N}^{-1}y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y), (5.1)

where

fluctN:=i=1NδxiNdx\mathrm{fluct}_{N}:=\sum_{i=1}^{N}\delta_{x_{i}}-N\mathrm{d}x

and dx\mathrm{d}x is the Lebesgue measure on 𝕋\mathbb{T}.

Remark 5.1.

For 2D Coulomb gases, the loop equation term is replaced by an angle term, as seen in [LS18] and [BBNY16]. In the case of β\beta-ensembles, (5.1) is smooth and so can be controlled directly by bounding the measure fluctN\mathrm{fluct}_{N} using local laws. In our case, (5.1) is as singular as the energy, which makes the treatment more delicate.

Proposition 5.1.

Let ξ𝒞2s(𝕋,)\xi\in\mathcal{C}^{2-s}(\mathbb{T},\mathbb{R}) and N(0,1]\ell_{N}\in(0,1]. Assume either that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}) or that N=1\ell_{N}=1. Let ψ𝒞2(N1𝕋,)\psi\in\mathcal{C}^{2}(\ell_{N}^{-1}\mathbb{T},\mathbb{R}) given by

ψ=12βcsN1s(Δ)1s2(ξ(N1))(N)andψ=0\psi^{\prime}=-\frac{1}{2\beta c_{s}}\ell_{N}^{1-s}(-\Delta)^{\frac{1-s}{2}}(\xi(\ell_{N}^{-1}\cdot))(\ell_{N}\cdot)\quad\text{and}\quad\int\psi=0 (5.2)

and Ψ𝒞2(DN,N)\Psi\in\mathcal{C}^{2}(D_{N},\mathbb{R}^{N}) given by

Ψ:XNDNN(ψ(N1x1),,ψ(N1xN)).\Psi:X_{N}\in D_{N}\mapsto\ell_{N}(\psi(\ell_{N}^{-1}x_{1}),\ldots,\psi(\ell_{N}^{-1}x_{N})).

We have

VarN,β[FluctN[ξ(N1)]]=1(NN)2(1s)𝔼N,β[βΨ2NΨ+|DΨ|2]+2(NN)1s𝔼N,β[i=1NN1ξ(N1xi)ψ(N1xi)]+1(NN)2(1s)VarN,β[βAN[ψ]i=1Nψ(N1xi)].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=-\frac{1}{(N\ell_{N})^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi\cdot\nabla^{2}\mathcal{H}_{N}\Psi+|D\Psi|^{2}]\\ +\frac{2}{(N\ell_{N})^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\ell_{N}^{-1}\xi^{\prime}(\ell_{N}^{-1}x_{i})\psi(\ell_{N}^{-1}x_{i})\Bigr{]}+\frac{1}{(N\ell_{N})^{2(1-s)}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\beta\mathrm{A}_{\ell_{N}}[\psi]-\sum_{i=1}^{N}\psi^{\prime}(\ell_{N}^{-1}x_{i})\Bigr{]}. (5.3)
Remark 5.2 (Mean-field approximation).

Proposition 5.1 can be interpreted as a mean-field approximation of the solution ϕ\nabla\phi of

{ϕ=FluctN[ξ]on DNϕn=0on DN,\begin{cases}\mathcal{L}\phi=\mathrm{Fluct}_{N}[\xi]&\text{on $D_{N}$}\\ \nabla\phi\cdot\vec{n}=0&\text{on $\partial D_{N}$},\end{cases} (5.4)

where :=N,β\mathcal{L}:=\mathcal{L}^{\mathbb{P}_{N,\beta}}. Due to the scaling of the energy and to the long-range nature of the interaction, one can find an approximate solution in the class of “diagonal transports”, Ψ:XNDN(ψ(x1),,ψ(xN))\Psi:X_{N}\in D_{N}\mapsto(\psi(x_{1}),\ldots,\psi(x_{N})) with ψ:𝕋\psi:\mathbb{T}\to\mathbb{R}.

For the hypersingular Riesz gas, i.e the Riesz gas with g(x)|x|sg(x)\sim|x|^{-s} for s>1s>1, one cannot approximate the solution of (5.4) by a diagonal transport. Let indeed ψ:𝕋\psi:\mathbb{T}\to\mathbb{R} smooth enough and Ψ:XNDN(ψ(x1),,ψ(xN))\Psi:X_{N}\in D_{N}\mapsto(\psi(x_{1}),\ldots,\psi(x_{N})). Since the energy is dominated by local interactions, the term div(Ψ)+βNΨ-\mathrm{div}(\Psi)+\beta\nabla\mathcal{H}_{N}\cdot\Psi is not at first order a linear statistic.

Remark 5.3 (Scaling relation).

The Riesz kernel on the circle does not satisfy any scaling relation, hence the intricate formula (5.2). However observe that if ξ0:\xi_{0}:\mathbb{R}\to\mathbb{R}, then

N1s(Δ)1s2(ξ0(N1))=((Δ)1s2ξ0)(N1).\ell_{N}^{1-s}(-\Delta)^{\frac{1-s}{2}}(\xi_{0}(\ell_{N}^{-1}\cdot))=((-\Delta)^{\frac{1-s}{2}}\xi_{0})(\ell_{N}^{-1}\cdot).

Therefore if ξ:𝕋\xi:\mathbb{T}\to\mathbb{R} is supported on (12,12)(-\frac{1}{2},\frac{1}{2}), letting ξ0:\xi_{0}:\mathbb{R}\to\mathbb{R} defined by

ξ0(x)={ξ(x)if |x|120if |x|>12,\xi_{0}(x)=\begin{cases}\xi(x)&\text{if $|x|\leq\frac{1}{2}$}\\ 0&\text{if $|x|>\frac{1}{2}$},\end{cases}

the solution of (5.2) approaches as N0\ell_{N}\to 0 the solution ψ0𝒞2(,)\psi_{0}\in\mathcal{C}^{2}(\mathbb{R},\mathbb{R}) of

ψ0=12βcs(Δ)1s2ξ0withψ0=0.\psi_{0}^{\prime}=-\frac{1}{2\beta c_{s}}(-\Delta)^{\frac{1-s}{2}}\xi_{0}\quad\text{with}\quad\int\psi_{0}=0.
Proof of Proposition 5.1.

Let :=N,β\mathcal{L}:=\mathcal{L}^{\mathbb{P}_{N,\beta}} be the operator

=Δ+βN,\mathcal{L}=-\Delta+\beta\nabla\mathcal{H}_{N}\cdot\nabla,

acting on H1(N,β)H^{1}(\mathbb{P}_{N,\beta}). Let ξ𝒞2s(𝕋,)\xi\in\mathcal{C}^{2-s}(\mathbb{T},\mathbb{R}) and

F=FluctN[ξ].F=\mathrm{Fluct}_{N}[\xi].

Let ϕ𝒞3(𝕋,)\phi\in\mathcal{C}^{3}(\mathbb{T},\mathbb{R}) and ψ=ϕ\psi=\phi^{\prime}. Define

Φ:XNDNϕ(x1)++ϕ(xN).\Phi:X_{N}\in D_{N}\mapsto\phi(x_{1})+\ldots+\phi(x_{N}).

We wish to find ϕ\phi such that ΦF\mathcal{L}\Phi\simeq F. Let us expand NΦ\nabla\mathcal{H}_{N}\cdot\nabla\Phi. Letting μN:=1Ni=1Nδxi\mu_{N}:=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}, we have

NΦ=ΔcN(1+s)g(xy)N(ψ(x)ψ(y))d(NμN)(x)d(NμN)(y),\nabla\mathcal{H}_{N}\cdot\nabla\Phi=\iint_{\Delta^{c}}N^{-(1+s)}g^{\prime}(x-y)N(\psi(x)-\psi(y))\mathrm{d}(N\mu_{N})(x)\mathrm{d}(N\mu_{N})(y),

a.e, where Δ:={(x,y)𝕋2:x=y}\Delta:=\{(x,y)\in\mathbb{T}^{2}:x=y\}. By decomposing μN\mu_{N} into μN=dx+1NfluctN\mu_{N}=\mathrm{d}x+\frac{1}{N}\mathrm{fluct}_{N}, one can break NΦ\nabla\mathcal{H}_{N}\cdot\nabla\Phi into

NΦ=N2N(ψ(x)ψ(y))N(1+s)g(xy)dxdy+2N(N(ψ(x)ψ(y))N(1+s)g(xy)dy)dfluctN(x)+A1[ψ],\nabla\mathcal{H}_{N}\cdot\nabla\Phi=N^{2}\iint N(\psi(x)-\psi(y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}x\mathrm{d}y\\ +2N\int\Bigr{(}\int N(\psi(x)-\psi(y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}y\Bigr{)}\mathrm{d}\mathrm{fluct}_{N}(x)+\mathrm{A}_{1}[\psi], (5.5)

with A1[ψ]\mathrm{A}_{1}[\psi] as in (5.1) for N=1\ell_{N}=1. For the crossed term we can write

NN(ψ(x)ψ(y))N(1+s)g(xy)dy=N1sgψ,N\int N(\psi(x)-\psi(y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}y=-N^{1-s}g^{\prime}*\psi,

where gψ=gψg^{\prime}*\psi=g*\psi^{\prime} is well-defined since ψ𝒞2(𝕋,)\psi\in\mathcal{C}^{2}(\mathbb{T},\mathbb{R}). Thus

ΦF=CN+(2Nβgψ+ξξ)dfluctNFluctN[ψ]+βA1[ψ],\mathcal{L}\Phi-F=C_{N}+\int\Bigr{(}-2N\beta g^{\prime}*\psi+\xi-\int\xi\Bigr{)}\mathrm{d}\mathrm{fluct}_{N}-\mathrm{Fluct}_{N}[\psi^{\prime}]+\beta A_{1}[\psi],

where CNC_{N} is the constant term in (5.5).

Let ψ0𝒞2(𝕋,)\psi_{0}\in\mathcal{C}^{2}(\mathbb{T},\mathbb{R}) be the solution of the convolution equation

2βgψ0=ξξwithψ0=0.-2\beta g^{\prime}*\psi_{0}=\xi-\int\xi\quad\text{with}\quad\int\psi_{0}=0.

Since gg is the fundamental solution of the fractional Laplace equation (1.2), ψ0\psi_{0} is the unique solution of

ψ0=12βcs(Δ)1s2ξwithψ0=0.\psi_{0}^{\prime}=-\frac{1}{2\beta c_{s}}(-\Delta)^{\frac{1-s}{2}}\xi\quad\text{with}\quad\int\psi_{0}=0. (5.6)

For this map, one can observe that the constant term in the splitting (5.5) vanishes:

NN(ψ0(x)ψ0(y))N(1+s)g(xy)dxdy=N1sgψ0=0.N\iint N(\psi_{0}(x)-\psi_{0}(y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}x\mathrm{d}y=-N^{1-s}\int g^{\prime}*\psi_{0}=0.

Let us choose ψ=1N1sψ0\psi=\frac{1}{N^{1-s}}\psi_{0} so that by (5.7),

ΦF=1N1sFluctN[ψ0]+βN1sA1[ψ0].\mathcal{L}\Phi-F=-\frac{1}{N^{1-s}}\mathrm{Fluct}_{N}[\psi_{0}^{\prime}]+\frac{\beta}{N^{1-s}}\mathrm{A}_{1}[\psi_{0}]. (5.7)

One has

VarN,β[F]=VarN,β[FΦ]VarN,β[Φ]+2CovN,β[F,Φ].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[F]=\mathrm{Var}_{\mathbb{P}_{N,\beta}}[F-\mathcal{L}\Phi]-\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathcal{L}\Phi]+2\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[F,\mathcal{L}\Phi].

Since Φn=0\nabla\Phi\cdot\vec{n}=0 a.e on DN\partial D_{N} (see Remark 3.2) we get by integration by parts under N,β\mathbb{P}_{N,\beta} that

CovN,β[F,Φ]=𝔼N,β[FΦ]\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[F,\mathcal{L}\Phi]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\nabla F\cdot\nabla\Phi]

and

VarN,β[Φ]=𝔼N,β[βΦ2NΦ+|2Φ|2].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathcal{L}\Phi]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\nabla\Phi\cdot\nabla^{2}\mathcal{H}_{N}\nabla\Phi+|\nabla^{2}\Phi|^{2}].

Assembling the above gives

VarN,β[F]=1N2(1s)𝔼N,β[βΨ02NΨ0+|DΨ0|2]+2N1s𝔼N,β[i=1Nξ(xi)ψ0(xi)]+1N2(1s)VarN,β[βA1[ψ0]i=1Nψ0(xi)],\mathrm{Var}_{\mathbb{P}_{N,\beta}}[F]=-\frac{1}{N^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi_{0}\cdot\nabla^{2}\mathcal{H}_{N}\Psi_{0}+|D\Psi_{0}|^{2}]+\frac{2}{N^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\xi^{\prime}(x_{i})\psi_{0}(x_{i})\Bigr{]}\\ +\frac{1}{N^{2(1-s)}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\beta\mathrm{A}_{1}[\psi_{0}]-\sum_{i=1}^{N}\psi_{0}^{\prime}(x_{i})\Bigr{]}, (5.8)

where

Ψ0:XNDN(ψ0(x1),,ψ0(xN)).\Psi_{0}:X_{N}\in D_{N}\mapsto(\psi_{0}(x_{1}),\ldots,\psi_{0}(x_{N})).

Let ξ\xi supported on (12,12)(-\frac{1}{2},\frac{1}{2}) and N(0,1]\ell_{N}\in(0,1]. Let ψ0𝒞2(N1𝕋,)\psi_{0}\in\mathcal{C}^{2}{(\ell_{N}^{-1}\mathbb{T},\mathbb{R})} defined by

ψ0=12βcsN1s(Δ)1s2(ξ(N1))(N)withψ0=0\psi_{0}^{\prime}=-\frac{1}{2\beta c_{s}}\ell_{N}^{1-s}(-\Delta)^{\frac{1-s}{2}}(\xi(\ell_{N}^{-1}\cdot))(\ell_{N}\cdot)\quad\text{with}\quad\int\psi_{0}=0 (5.9)

By construction, N(1s)ψ0\ell_{N}^{-(1-s)}\psi_{0} solves (5.6) when ξ\xi is replaced by ξ(N1)\xi(\ell_{N}^{-1}\cdot). Therefore replacing ξ\xi by ξ(N1)\xi(\ell_{N}^{-1}\cdot) and ψ0\psi_{0} by N(1s)ψ0\ell_{N}^{-(1-s)}\psi_{0} in (5.8) gives

VarN,β[FluctN[ξ(N1)]]=1(NN)2(1s)𝔼N,β[βΨ02NΨ0+|DΨ0|2]+2(NN)1s𝔼N,β[i=1NN1ξ(N1xi)ψ0(N1xi)]+1(NN)2(1s)VarN,β[βAN[ψ0]i=1Nψ0(N1xi)],\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=-\frac{1}{(N\ell_{N})^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi_{0}\cdot\nabla^{2}\mathcal{H}_{N}\Psi_{0}+|D\Psi_{0}|^{2}]\\ +\frac{2}{(N\ell_{N})^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\ell_{N}^{-1}\xi^{\prime}(\ell_{N}^{-1}x_{i})\psi_{0}(\ell_{N}^{-1}x_{i})\Bigr{]}+\frac{1}{(N\ell_{N})^{2(1-s)}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\beta\mathrm{A}_{\ell_{N}}[\psi_{0}]-\sum_{i=1}^{N}\psi_{0}^{\prime}(\ell_{N}^{-1}x_{i})\Bigr{]},

where

Ψ0:XNDNN(ψ0(N1x1),,ψ0(N1xN)).\Psi_{0}:X_{N}\in D_{N}\mapsto\ell_{N}(\psi_{0}(\ell_{N}^{-1}x_{1}),\ldots,\psi_{0}(\ell_{N}^{-1}x_{N})).

5.2. Splitting of the loop equation term

In view of Proposition 5.1, expanding the variance of a linear statistic reduces to controlling the loop equation term (5.1). Let us first discard a strategy based on local laws only. Recall that for all ψ:N1𝕋\psi:\ell_{N}^{-1}\mathbb{T}\to\mathbb{R} smooth enough,

AN[ψ]=ΔcNN(ψ(N1x)ψ(N1y))N(1+s)g(xy)dfluctN(x)dfluctN(y).\mathrm{A}_{\ell_{N}}[\psi]=\iint_{\Delta^{c}}N\ell_{N}(\psi(\ell_{N}^{-1}x)-\psi(\ell_{N}^{-1}y))N^{-(1+s)}g^{\prime}(x-y)\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y).

By using local laws on gaps, one may control the above integral away from the diagonal. Nevertheless AN[ψ]\mathrm{A}_{\ell_{N}}[\psi] contains local terms such as

i=1NN(ψ(N1xi+1)ψ(N1xi))N(1+s)g(xi+1xi),\sum_{i=1}^{N}N(\psi(\ell_{N}^{-1}x_{i+1})-\psi(\ell_{N}^{-1}x_{i}))N^{-(1+s)}g^{\prime}(x_{i+1}-x_{i}),

which is O(NN|ψ|)O(N\ell_{N}|\psi^{\prime}|_{\infty}) with overwhelming probability. Therefore applying a local law estimate will give in the best case, the bound

VarN,β[AN[ψ]]=O((NN)2|ψ|2).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi]]=O((N\ell_{N})^{2}|\psi^{\prime}|_{\infty}^{2}).

Inserting this into Proposition 5.1 gives an error term of order O((NN)2s)O((N\ell_{N})^{2s}), which is larger than the expected order of fluctuations of linear statistics. One should instead exploit the convexity of the interaction and bound the fluctuations of AN[ψ]\mathrm{A}_{\ell_{N}}[\psi] using various concentration inequalities. As emphasized in Section 3, the variance of a smooth function under a log-concave probability measure is related to the norm of its gradient and one should therefore first differentiate (5.1).

Before entering into the main computations, we first define a localized version of AN[ψ]\mathrm{A}_{\ell_{N}}[\psi]. We will assume that 0 is in the support of ξ\xi and let i0i_{0} be the index (defined almost surely) such that xix_{i} is the closest point to 0:

i0:=argmin1iN|xi|.i_{0}:=\underset{1\leq i\leq N}{\mathrm{argmin}}|x_{i}|. (5.10)

Fix γ>1\gamma>1 and let

η:={(NN)γNif N012if N=1 for each N.\eta:=\begin{cases}\frac{(N\ell_{N})^{\gamma}}{N}&\text{if $\ell_{N}\to 0$}\\ \frac{1}{2}&\text{if $\ell_{N}=1$ for each $N$}.\end{cases} (5.11)

Then define

IN={i{1,,N}:d(i,i0)Nη}.I_{N}=\{i\in\{1,\ldots,N\}:d(i,i_{0})\leq N\eta\}. (5.12)

For ψ:N1𝕋\psi:\ell_{N}^{-1}\mathbb{T}\to\mathbb{R} smooth enough, we define a localized version of AN[ψ]\mathrm{A}_{\ell_{N}}[\psi] by letting

A~N[ψ]:=ijINNN(ψ(N1xi)ψ(N1xj))N(1+s)g(xixj)2NiIN|y|ηNN(ψ(N1x)ψ(N1y))N(1+s)g(xiy)dy.\tilde{\mathrm{A}}_{\ell_{N}}[\psi]:=\sum_{i\neq j\in I_{N}}N\ell_{N}(\psi(\ell_{N}^{-1}x_{i})-\psi(\ell_{N}^{-1}x_{j}))N^{-(1+s)}g^{\prime}(x_{i}-x_{j})\\ -2N\sum_{i\in I_{N}}\int_{|y|\leq\eta}N\ell_{N}(\psi(\ell_{N}^{-1}x)-\psi(\ell_{N}^{-1}y))N^{-(1+s)}g^{\prime}(x_{i}-y)\mathrm{d}y. (5.13)

For ε>0\varepsilon>0, define the good event

𝒜={XNDN:iIN,k{N/2,,N/2}:i+kIN,N|xi+kxik/N|(NN)εks2,iIN,(NN)εN|xi+1xi|(NN)ε}.\mathcal{A}=\{X_{N}\in D_{N}:\forall i\in I_{N},k\in\{-N/2,\ldots,N/2\}:i+k\in I_{N},\\ N|x_{i+k}-x_{i}-k/N|\leq(N\ell_{N})^{\varepsilon}k^{\frac{s}{2}},\forall i\in I_{N},(N\ell_{N})^{-\varepsilon}\leq N|x_{i+1}-x_{i}|\leq(N\ell_{N})^{\varepsilon}\}. (5.14)

Let K0>1K_{0}>1. Consider h1,h2:𝕋{+}h_{1},h_{2}:\mathbb{T}\to\mathbb{R}\cup\{+\infty\} defined for all x𝕋x\in\mathbb{T} by

h1(x):=(N1+l=1pNα+1s1(|xNal|1N)2s+α)𝟙|x|<2N+1(NN)γsN1(N1|x|+1)2s.h_{1}(x):=\Bigr{(}\ell_{N}^{-1}+\sum_{l=1}^{p}\ell_{N}^{\alpha+1-s}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\Bigr{)}\mathds{1}_{|x|<2\ell_{N}}+\frac{1}{(N\ell_{N})^{\gamma s}}\frac{\ell_{N}^{-1}}{(\ell_{N}^{-1}|x|+1)^{2-s}}. (5.15)

and

h2(x)=(l=1p(NN)α+1sNs21(|xNal|1N)1+s2+1(NN)(2s)(1γ)Ns2(d(x,I)1N)1+s2+Ns2(N+|x|)1+s2)𝟙|x|K0η.h_{2}(x)=\Bigr{(}\sum_{l=1}^{p}(N\ell_{N})^{\alpha+1-s}N^{-\frac{s}{2}}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{1+\frac{s}{2}}}+\frac{1}{(N\ell_{N})^{(2-s)(1-\gamma)}}\frac{N^{-\frac{s}{2}}}{(d(x,\partial I)\vee\frac{1}{N})^{1+\frac{s}{2}}}\\ +\frac{N^{-\frac{s}{2}}}{(\ell_{N}+|x|)^{1+\frac{s}{2}}}\Bigr{)}\mathds{1}_{|x|\leq K_{0}\eta}. (5.16)
Lemma 5.2.

Let ψ𝒞2(N1𝕋,)\psi\in\mathcal{C}^{2}(\ell_{N}^{-1}\mathbb{T},\mathbb{R}). Assume that there exists α(s1,s2)\alpha\in(s-1,\frac{s}{2}) and a1,,ap𝕋a_{1},\ldots,a_{p}\in\mathbb{T} such that

|ψ′′|(x)C(1+l=1p1(|xal|1N)2s+α𝟙|x|<2N+1(|x|+N)3s).|\psi^{\prime\prime}|(x)\leq C\Bigr{(}1+\sum_{l=1}^{p}\frac{1}{(|x-a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\mathds{1}_{|x|<2\ell_{N}}+\frac{1}{(|x|+\ell_{N})^{3-s}}\Bigr{)}. (5.17)

Let γ>1\gamma>1, INI_{N} be as in (5.12) and I:=[η,η]𝕋I:=[-\eta,\eta]\subset\mathbb{T}. Let A~N[ψ]\tilde{\mathrm{A}}_{\ell_{N}}[\psi] be as in (5.13). One can break A~N[ψ]\nabla\tilde{\mathrm{A}}_{\ell_{N}}[\psi] into

A~N[ψ]=V+W,\nabla\tilde{\mathrm{A}}_{\ell_{N}}[\psi]=\mathrm{V}+\mathrm{W},

with V,WL2({1,,N},H1(N,β))\mathrm{V},\mathrm{W}\in L^{2}(\{1,\ldots,N\},H^{1}(\mathbb{P}_{N,\beta})) satisfying

  1. (1)

    For each iINci\in I_{N}^{c}, Vi=Wi=0\mathrm{V}_{i}=\mathrm{W}_{i}=0,

  2. (2)

    There exist C>0C>0 and K0>1K_{0}>1 such that, letting h1h_{1} and h2h_{2} be as in (5.15) and (5.16), for each iINi\in I_{N} and XN𝒜X_{N}\in\mathcal{A},

    |Vi|C(NN)κε(h1+h2)(xi).|\mathrm{V}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}(h_{1}+h_{2})(x_{i}). (5.18)
  3. (3)

    There exists W~L2({1,,N},H1(N,β))\tilde{\mathrm{W}}\in L^{2}(\{1,\ldots,N\},H^{1}(\mathbb{P}_{N,\beta})) such that for all UNNU_{N}\in\mathbb{R}^{N},

    WUN=i=1NW~N(ui+1ui)\mathrm{W}\cdot U_{N}=-\sum_{i=1}^{N}\tilde{\mathrm{W}}N(u_{i+1}-u_{i})

    with

    sup𝒜|W~|2C(NN)κε(NN+(NN)2(α+1s)).\sup_{\mathcal{A}}|\tilde{\mathrm{W}}|^{2}\leq C(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N}+(N\ell_{N})^{2(\alpha+1-s)}). (5.19)

The proof of Lemma 5.2 is deferred to Appendix B.

5.3. Quantitative variance expansion

We proceed to the proof of Theorem 2.

Starting from Proposition 5.1, we first prove that Theorem 2 holds up to the control of the variance of the loop equation term (5.1).

Proposition 5.3.

Let ξ\xi satisfy Assumptions 1.1. Let α(s1,s2)\alpha\in(s-1,\frac{s}{2}) be such that (1.8) is satisfied. Let {N}\{\ell_{N}\} be a sequence of positive numbers in (0,1](0,1]. Assume that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}) or that N=1\ell_{N}=1 for each NN. Let :=(NN)1\ell:=(N\ell_{N})^{-1} and ξreg:=ξK\xi_{\mathrm{reg}}:=\xi*K_{\ell} where KK_{\ell} is the kernel (2.22). Consider the map ψreg𝒞2(𝕋,)\psi_{\mathrm{reg}}\in\mathcal{C}^{2}(\mathbb{T},\mathbb{R}) defined by

ψreg=12csN1s(Δ)1s2(ξreg(N1))(N)andψreg=0.\psi_{\mathrm{reg}}^{\prime}=-\frac{1}{2c_{s}}\ell_{N}^{1-s}(-\Delta)^{\frac{1-s}{2}}(\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot))(\ell_{N}\cdot)\quad\text{and}\quad\int\psi_{\mathrm{reg}}=0. (5.20)

Let σN2(ξ)\sigma_{\ell_{N}}^{2}(\xi) be as in (1.10). For all ε>0\varepsilon>0,

VarN,β[FluctN[ξ(N1)]]=(NN)sσN2(ξ)+O((NN)2(s1)VarN,β[AN[ψreg]]+(NN)max(2s1,2α)+κε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=(N\ell_{N})^{s}\sigma_{\ell_{N}}^{2}(\xi)\\ +O\Bigr{(}(N\ell_{N})^{2(s-1)}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]+(N\ell_{N})^{\max(2s-1,2\alpha)+\kappa\varepsilon}\Bigr{)}. (5.21)

Moreover

σN2(ξ)=σ2(ξ)+O(N2s).\sigma_{\ell_{N}}^{2}(\xi)=\sigma_{\infty}^{2}(\xi)+O(\ell_{N}^{2-s}). (5.22)
Proof.

From the Poissonian estimate of Lemma 3.9 we get the crude bound

VarN,β[FluctN[ξreg(N1)ξ(N1)]]NN|ξregξ|L22.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)-\xi(\ell_{N}^{-1}\cdot)]]\leq N\ell_{N}|\xi_{\mathrm{reg}}-\xi|_{L^{2}}^{2}.

Inserting (2.23) into the above display yields

VarN,β[FluctN[ξreg(N1)ξ(N1)]]CNN((NN)1+(NN)1(12α))C(1+(NN)2α).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)-\xi(\ell_{N}^{-1}\cdot)]]\leq CN\ell_{N}((N\ell_{N})^{-1}+(N\ell_{N})^{-1(1-2\alpha)})\\ \leq C(1+(N\ell_{N})^{2\alpha}). (5.23)

Note that since α<s2\alpha<\frac{s}{2}, the above term is o((NN)s)o((N\ell_{N})^{s}).

Let us now study the fluctuations of FluctN[ξreg(N1)]\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)]. Let ψreg\psi_{\mathrm{reg}} be as in (5.20). Define

Ψ:XNDNN(ψreg(N1x1),,ψreg(N1xN)).\Psi:X_{N}\in D_{N}\mapsto\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{1}),\ldots,\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{N})).

Applying Proposition 5.1 one gets

VarN,β[FluctN[ξreg(N1)]]=1(NN)2(1s)𝔼N,β[βΨ2NΨ+|DΨ|2]+2(NN)1s𝔼N,β[i=1N(N1ξregψreg)(N1xi)]+1(NN)2(1s)VarN,β[βAN[ψreg]i=1Nψreg(N1xi)].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)]]=-\frac{1}{(N\ell_{N})^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi\cdot\nabla^{2}\mathcal{H}_{N}\Psi+|D\Psi|^{2}]\\ +\frac{2}{(N\ell_{N})^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}(\ell_{N}^{-1}\xi_{\mathrm{reg}}^{\prime}\psi_{\mathrm{reg}})(\ell_{N}^{-1}x_{i})\Bigr{]}+\frac{1}{(N\ell_{N})^{2(1-s)}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\beta\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\sum_{i=1}^{N}\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})\Bigr{]}. (5.24)

Using the Poissonian estimate of Lemma 3.9 again and the estimate (2.24), we obtain

VarN,β[FluctN[ψreg(N1)]]NN|ψreg|L22C(NN+(NN)2+2(αs)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]]\leq N\ell_{N}|\psi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2}\leq C(N\ell_{N}+(N\ell_{N})^{2+2(\alpha-s)}). (5.25)

Splitting μN\mu_{N} into dx+1NfluctN\mathrm{d}x+\frac{1}{N}\mathrm{fluct}_{N}, one can write

𝔼N,β[Ψ2NΨ]=𝔼N,β[ΔcN(s+2)g′′(xy)(NN(ψreg(N1x)ψreg(N1y))2d(NμN)(x)d(NμN)(y)]=N(s+2)g′′(xy)(NN(ψreg(N1x)ψreg(N1y))2(Ndx)(Ndy)+BN[ψreg]\mathbb{E}_{\mathbb{P}_{N,\beta}}[\Psi\cdot\nabla^{2}\mathcal{H}_{N}\Psi]\\ =\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\iint_{\Delta^{c}}N^{-(s+2)}g^{\prime\prime}(x-y)(N\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}\mathrm{d}(N\mu_{N})(x)\mathrm{d}(N\mu_{N})(y)\Bigr{]}\\ =\iint N^{-(s+2)}g^{\prime\prime}(x-y)(N\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}(N\mathrm{d}x)(N\mathrm{d}y)+\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]

where

BN[ψreg]:=ΔcN(s+2)g′′(xy)(NN)2(ψreg(N1x)ψreg(N1y))2(Ndx)dfluctN(y).\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]:=\iint_{\Delta^{c}}N^{-(s+2)}g^{\prime\prime}(x-y)(N\ell_{N})^{2}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}(N\mathrm{d}x)\mathrm{d}\mathrm{fluct}_{N}(y). (5.26)

Therefore

1(NN)2(1s)𝔼N,β[βΨ2NΨ+|DΨ|2]+2(NN)1s𝔼N,β[i=1N(N1ξregψreg)(N1xi)]=NsCN(NN)2s1(ψreg)2β(NN)2(1s)𝔼N,β[BN[ψreg]],-\frac{1}{(N\ell_{N})^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi\cdot\nabla^{2}\mathcal{H}_{N}\Psi+|D\Psi|^{2}]+\frac{2}{(N\ell_{N})^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}(\ell_{N}^{-1}\xi_{\mathrm{reg}}^{\prime}\psi_{\mathrm{reg}})(\ell_{N}^{-1}x_{i})\Bigr{]}\\ =N^{s}C_{N}-(N\ell_{N})^{2s-1}\int(\psi_{\mathrm{reg}}^{\prime})^{2}-{\beta}{(N\ell_{N})^{-2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]], (5.27)

where

CN:=βg′′(xy)(N(1s)ψreg(N1x)N(1s)ψreg(N1y))2dxdy+2ξreg(N1)N(1s)ψreg(N1)=12βcs|ξreg(N1)|H1s22=NsσN(ξreg),C_{N}:=-\beta\int g^{\prime\prime}(x-y)(\ell_{N}^{-(1-s)}\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\ell_{N}^{-(1-s)}\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}\mathrm{d}x\mathrm{d}y\\ +2\int\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)\ell_{N}^{-(1-s)}\psi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)=\frac{1}{2\beta c_{s}}|\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)|_{H^{\frac{1-s}{2}}}^{2}=\ell_{N}^{s}\sigma_{\ell_{N}}(\xi_{\mathrm{reg}}),

where we have applied (2.14) in the second equality. We will prove in Lemma B.3 that for all ε>0\varepsilon>0, there exists C>0C>0 such that

𝔼N,β[BN[ψreg]]C(NN)ε((NN+(NN)2(1s+α)).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{\varepsilon}((N\ell_{N}+(N\ell_{N})^{2(1-s+\alpha)}). (5.28)

By the estimate (2.26) of Lemma 2.5, one has

|σN2(ξ)σN2(ξreg)|C((NN)1+(NN)(s2α)).|\sigma_{\ell_{N}}^{2}(\xi)-\sigma_{\ell_{N}}^{2}(\xi_{\mathrm{reg}})|\leq C((N\ell_{N})^{-1}+(N\ell_{N})^{-(s-2\alpha)}). (5.29)

Therefore combining (5.27), (5.29) and inserting (2.24) to control |ψreg|L22|\psi^{\prime}_{\mathrm{reg}}|_{L^{2}}^{2}, we get

1(NN)2(1s)𝔼N,β[βΨ2NΨ+|DΨ|2]+2(NN)1s𝔼N,β[i=1N(N1ξregψreg)(N1xi)]=(NN)sσN2(ξ)+O((NN)max(2s1,2α)).-\frac{1}{(N\ell_{N})^{2(1-s)}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\beta\Psi\cdot\nabla^{2}\mathcal{H}_{N}\Psi+|D\Psi|^{2}]+\frac{2}{(N\ell_{N})^{1-s}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}(\ell_{N}^{-1}\xi_{\mathrm{reg}}^{\prime}\psi_{\mathrm{reg}})(\ell_{N}^{-1}x_{i})\Bigr{]}\\ =(N\ell_{N})^{s}\sigma_{\ell_{N}}^{2}(\xi)+O((N\ell_{N})^{\max(2s-1,2\alpha)}).

Using this and (5.23), (5.24) and (5.25), we conclude that

VarN,β[FluctN[ξ(N1)]]=(NN)sσN2(ξ)+O((NN)2(s1)VarN,β[AN[ψreg]]+(NN)max(2s1,2α)+ε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=(N\ell_{N})^{s}\sigma^{2}_{N}(\xi)\\ +O\Bigr{(}(N\ell_{N})^{2(s-1)}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]+(N\ell_{N})^{\max(2s-1,2\alpha)+\varepsilon}\Bigr{)}.

In the case where N0\ell_{N}\to 0, as proved in (2.16),

|ξ(N1)|H1s22=Nsσ2(ξ)+O(N2|ξ|L22),|\xi(\ell_{N}^{-1}\cdot)|_{H^{\frac{1-s}{2}}}^{2}=\ell_{N}^{s}\sigma_{\infty}^{2}(\xi)+O(\ell_{N}^{2}|\xi|_{L^{2}}^{2}),

which gives (5.22). ∎

We turn to the control on the variance of (5.1). When N0\ell_{N}\to 0, we first localize the term AN[ψreg]\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}] into a smaller window by discarding long-range interactions using the concentration estimate of Theorem 1. We then split the gradient of the main term into two vector-fields V\mathrm{V} and W\mathrm{W} as in Lemma 5.2. The vector-field W\mathrm{W} is easy to deal with, using the bound (5.19) and the log-concavity of GapN#N,β\mathrm{Gap}_{N}\#\mathbb{P}_{N,\beta}. Using the comparison principle of Lemma 3.6 we are able control the Dirichlet energy of V\mathrm{V} by the variance of an auxiliary (singular) linear statistic.

Let η\eta be as in (5.11). Let K0>1K_{0}>1 and t:=min(K0η,14)t:=\min(K_{0}\eta,\frac{1}{4}). Define

χ:x𝕋2ssgn(|x|1N)(|x|1N)s2𝟙|x|K0η.\chi:x\in\mathbb{T}\mapsto-\frac{2}{s}\frac{\mathrm{sgn}(|x|\vee\frac{1}{N})}{(|x|\vee\frac{1}{N})^{\frac{s}{2}}}\mathds{1}_{|x|\leq K_{0}\eta}. (5.30)

Let us consider χ1\chi_{1} of mean 0 such that for all x𝕋x\in\mathbb{T},

χ1(x)={1(|x|1N)1+s2if |x|<t1t2+s2(xt)+1t1+s2if x[t,2t]1t2+s2(x+t)+1t1+s2if x[2t,t]0otherwise.\chi_{1}^{\prime}(x)=\begin{cases}\frac{1}{(|x|\vee\frac{1}{N})^{1+\frac{s}{2}}}&\text{if $|x|<t$}\\ -\frac{1}{t^{2+\frac{s}{2}}}(x-t)+\frac{1}{t^{1+\frac{s}{2}}}&\text{if $x\in[t,2t]$}\\ \frac{1}{t^{2+\frac{s}{2}}}(x+t)+\frac{1}{t^{1+\frac{s}{2}}}&\text{if $x\in[-2t,-t]$}\\ 0&\text{otherwise}.\end{cases} (5.31)

Define

χ2:=χχ1.\chi_{2}:=\chi-\chi_{1}. (5.32)

Note that in the case where N=1\ell_{N}=1 for each NN, t=14t=\frac{1}{4} and χ2:𝕋\chi_{2}:\mathbb{T}\to\mathbb{R} is 𝒞2\mathcal{C}^{2} on 𝕋{12}\mathbb{T}\setminus\{\frac{1}{2}\} and discontinuous at 12\frac{1}{2}. The function χ2\chi_{2} satisfies Assumptions 1.1 with α:=0<s2\alpha:=0<\frac{s}{2}.

Proposition 5.4.

Let ξ\xi satisfy Assumptions 1.1 and α(s1,s2)\alpha\in(s-1,\frac{s}{2}) as in (1.8). Let {N}\{\ell_{N}\} be a sequence of positive numbers in (0,1](0,1]. Assume that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}) or that N=1\ell_{N}=1 for each NN. Let :=1/(NN)\ell:=1/(N\ell_{N}) and ξreg:=ξK\xi_{\mathrm{reg}}:=\xi*K_{\ell} with KK_{\ell} as in (2.22). Let ψreg\psi_{\mathrm{reg}} be as in (5.20). Let γ>3s2s1s(1+11s)\gamma>\frac{3-s}{2-s}\vee\frac{1}{s}\vee(1+\frac{1}{1-s}). Let χ1\chi_{1} be as in (5.31).

For all ε>0\varepsilon>0, there exists C>0C>0 depending on ε>0\varepsilon>0 and γ\gamma such that

VarN,β[AN[ψreg]]C(NN)ε+max(1,2(α+1s))+C(NN)κε+2(1s+α)Ns(VarN,β[FluctN[χ1]]+VarN,β[FluctN[χ2]]𝟙N=1).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{\varepsilon+\max(1,2(\alpha+1-s))}\\ +C(N\ell_{N})^{\kappa\varepsilon+2(1-s+\alpha)}N^{-s}\Bigr{(}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\mathds{1}_{\ell_{N}=1}\Bigr{)}. (5.33)
Proof.
Step 1: reduction to a finite-range quantity

Let γ>3s2s1s(1+11s)\gamma>\frac{3-s}{2-s}\vee\frac{1}{s}\vee(1+\frac{1}{1-s}). Let i0i_{0} and INI_{N} be as in (5.10) and (5.12). Let us split AN[ψreg]\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}] into AN[ψreg]=A~N[ψreg]+ANext[ψreg]\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]=\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]+\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}] with A~N[ψreg]\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}] as defined in (5.13):

A~N[ψreg]=iIN,d(i,j)NηNN(ψreg(N1xi)ψreg(N1xj))N(1+s)g(xixj)2NiIN|y|ηNN(ψreg(N1x)ψreg(N1y))N(1+s)g(xiy)dy,\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]=\sum_{i\in I_{N},d(i,j)\leq N\eta}N\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})-\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{j}))N^{-(1+s)}g^{\prime}(x_{i}-x_{j})\\ -2N\sum_{i\in I_{N}}\int_{|y|\leq\eta}N\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))N^{-(1+s)}g^{\prime}(x_{i}-y)\mathrm{d}y,

where η\eta is as in (5.11). We will prove in Appendix B (see Lemma B.4) that there exist C>0C>0 and δ>0\delta>0 depending on γ\gamma such that

VarN,β[ANext[ψreg]]C(NN)1δ.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{1-\delta}. (5.34)

In the rest of the proof, we denote for shortcut A~:=A~N[ψreg]\tilde{\mathrm{A}}:=\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}].

Step 2: fixing a point

As explained in Subsection 3.2, applying our FKG-type inequalities requires to fix a point. Recall

VarN,β[A~]=𝔼N,β[VarN,β[A~x1]]+VarN,β[𝔼N,β[A~x1]].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}\mid x_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathbb{E}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}\mid x_{1}]].

We claim that

VarN,β[𝔼N,β[A~x1]]C((NN)2(1s)+(NN)2(α+1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathbb{E}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}\mid x_{1}]]\leq C((N\ell_{N})^{2(1-s)}+(N\ell_{N})^{2(\alpha+1-s)}). (5.35)

The proof of (5.35) uses the fact that the law of x2,,xNx_{2},\ldots,x_{N} under N,β(x1=x)\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x) is the law of x2x1+x,,xNx1+xx_{2}-x_{1}+x,\ldots,x_{N}-x_{1}+x under N,β\mathbb{P}_{N,\beta} as well as the rigidity estimate of Theorem 1. We postpone the details to Appendix B, see Lemma B.5.

It therefore remains to study the variance of A~\tilde{\mathrm{A}} conditionally on x1=xx_{1}=x.

Step 3: convexification and conditioning

Fix x𝕋x\in\mathbb{T}. We first convexify the measure N,β(x1=x)\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x) by penalizing large nearest-neighbor gaps in the window INI_{N} as in Section 4.

Let θ:++\theta:\mathbb{R}^{+}\to\mathbb{R}^{+} be a smooth cutoff function such that θ(x)=|x|2\theta(x)=|x|^{2} for x>1x>1, θ=0\theta=0 on [0,12][0,\frac{1}{2}] and θ′′0\theta^{\prime\prime}\geq 0 on +\mathbb{R}^{+}. Fix ε>0\varepsilon>0 and let

F:=i:i,i+1INθ(N(xi+1xi)(NN)ε).\mathrm{F}:=\sum_{i:i,i+1\in I_{N}}\theta\Bigr{(}\frac{N(x_{i+1}-x_{i})}{(N\ell_{N})^{\varepsilon}}\Bigr{)}.

Let N,β\mathbb{Q}_{N,\beta} be the locally constrained measure

dN,β(XN)=1KN,βeβ(N+F)(XN)𝟙DN(XN)dXN.\mathrm{d}\mathbb{Q}_{N,\beta}(X_{N})=\frac{1}{K_{N,\beta}}e^{-\beta(\mathcal{H}_{N}+\mathrm{F})(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N}. (5.36)

In view of Lemma 4.1 and Theorem 1, the total variation distance between the measures N,β(x1=x)\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x) and N,β(x1=x)\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x) satisfies

TV(N,β(x1=x),N,β(x1=x))e(NN)δ,\mathrm{TV}(\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x),\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x))\leq e^{-(N\ell_{N})^{\delta}}, (5.37)

for some constant δ>0\delta>0 depending on ε\varepsilon. Using the rigidity estimates of Theorem 1 and (5.37), one can easily show that there exists κ>1\kappa>1, δ>0\delta>0 such that

𝔼N,β[A~2𝟙|A~|(NN)κx1=x]e(NN)δ,𝔼N,β[A~2𝟙|A~|(NN)κx1=x]e(NN)δ.\mathbb{E}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}^{2}\mathds{1}_{|\tilde{\mathrm{A}}|\geq(N\ell_{N})^{\kappa}}\mid x_{1}=x]\leq e^{-(N\ell_{N})^{\delta}},\quad\mathbb{E}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}^{2}\mathds{1}_{|\tilde{\mathrm{A}}|\geq(N\ell_{N})^{\kappa}}\mid x_{1}=x]\leq e^{-(N\ell_{N})^{\delta}}.

In view of (5.37), this implies that there exists δ>0\delta>0 depending on ε\varepsilon such that

|VarN,β[A~x1=x]VarN,β[A~x1=x]|e(NN)δ.|\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}\mid x_{1}=x]-\mathrm{Var}_{\mathbb{Q}_{N,\beta}}[\tilde{\mathrm{A}}\mid x_{1}=x]|\leq e^{-(N\ell_{N})^{\delta}}.

Together with (5.35) we get

VarN,β[A~]=𝔼N,β[VarN,β(x1=x)[A~]]+O((NN)2(α+1s)+(NN)2(1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\tilde{\mathrm{A}}]]+O((N\ell_{N})^{2(\alpha+1-s)}+(N\ell_{N})^{2(1-s)}). (5.38)
Step 4: reduction to a good event

Recall the good event (5.14),

𝒜={XNDN:iIN,k{N/2,,N/2}:i+kIN,N|xi+kxik/N|(NN)εks2,iIN,(NN)εN|xi+1xi|(NN)ε}.\mathcal{A}=\{X_{N}\in D_{N}:\forall i\in I_{N},k\in\{-N/2,\ldots,N/2\}:i+k\in I_{N},\\ N|x_{i+k}-x_{i}-k/N|\leq(N\ell_{N})^{\varepsilon}k^{\frac{s}{2}},\forall i\in I_{N},(N\ell_{N})^{-\varepsilon}\leq N|x_{i+1}-x_{i}|\leq(N\ell_{N})^{\varepsilon}\}.

Let ϕ:\phi:\mathbb{R}\to\mathbb{R} piecewise linear such that ϕ(x)=0\phi(x)=0 for |x|>1|x|>1, ϕ(x)=0\phi(x)=0 for |x|<12|x|<\frac{1}{2} and ϕ(x)=22|x|\phi(x)=2-2|x| for 12|x|1\frac{1}{2}\leq|x|\leq 1. Define the cutoff function

η:=iINN2kN2:i+kINϕ(N(xi+kxi)k|k|s2(NN)ε)iIN:i+1INϕ(N(xi+1xi)(NN)ε).\eta:=\prod_{i\in I_{N}}\prod_{-\frac{N}{2}\leq k\leq\frac{N}{2}:i+k\in I_{N}}\phi\Bigr{(}\frac{N(x_{i+k}-x_{i})-k}{|k|^{\frac{s}{2}}(N\ell_{N})^{\varepsilon}}\Bigr{)}\prod_{i\in I_{N}:i+1\in I_{N}}\phi(N(x_{i+1}-x_{i})(N\ell_{N})^{\varepsilon}). (5.39)

Using the rigidity estimate of Theorem 1 and (5.37), one can show that there exists δ>0\delta>0 depending on ε\varepsilon and γ\gamma such that

VarN,β[(1η)A~x1=x]e(NN)δ.\mathrm{Var}_{\mathbb{Q}_{N,\beta}}[(1-\eta)\widetilde{\mathrm{A}}\mid x_{1}=x]\leq e^{-(N\ell_{N})^{\delta}}.

Denote for shortcut

A1x:=A1N,β(|x1=x).A_{1}^{x}:=A_{1}^{{\mathbb{Q}_{N,\beta}}(\cdot|x_{1}=x)}.

Since A1xA_{1}^{x} positive on L2({1,,N},H1(N,β(x1=x)))L^{2}(\{1,\ldots,N\},H^{1}(\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x))), we get by subadditivity that

VarN,β(x1=x)[ηA~]=𝔼N,β(x1=x)[(ηA~)(A1x)1(ηA~)]2𝔼N,β(x1=x)[(ηA~)(A1x)1(ηA~)]+2𝔼N,β(x1=x)[(ηA~)(A1x)1(ηA~)].\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta\tilde{\mathrm{A}}]=\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\nabla(\eta\tilde{\mathrm{A}})\cdot(A_{1}^{x})^{-1}\nabla(\eta\tilde{\mathrm{A}})]\\ \leq 2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\nabla\tilde{\mathrm{A}})(A_{1}^{x})^{-1}(\eta\nabla\tilde{\mathrm{A}})]+2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\nabla\eta\tilde{\mathrm{A}})(A_{1}^{x})^{-1}(\nabla\eta\tilde{\mathrm{A}})].

Let us split A~\nabla\tilde{\mathrm{A}} into

A~=V+W\nabla\tilde{\mathrm{A}}=\mathrm{V}+\mathrm{W}

with V\mathrm{V}, W\mathrm{W} as in Lemma 5.2. By subadditivity again

𝔼N,β(x1=x)[(ηA~)(A1x)1(ηA~)]2𝔼N,β(x1=x)[(ηV)(A1x)1(ηV)]+2𝔼N,β(x1=x)[(ηW)(A1x)1(ηW)].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\nabla\tilde{\mathrm{A}})(A_{1}^{x})^{-1}(\eta\nabla\tilde{\mathrm{A}})]\leq 2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\mathrm{V})\cdot(A_{1}^{x})^{-1}(\eta\mathrm{V})]\\ +2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\mathrm{W})\cdot(A_{1}^{x})^{-1}(\eta\mathrm{W})]. (5.40)
Step 5: using Poincaré in gap coordinates for ηW\eta\mathrm{W} and A~η\tilde{\mathrm{A}}\nabla\eta

To estimate the second term of the right-hand side of (5.40), one can take advantage of the fact that N,β(x1=x)\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x) is uniformly log-concave in gap coordinates. Indeed, using the Brascamp-Lieb inequality, one can write

𝔼N,β(x1=x)[ηW(A1x)1(ηW)]β1𝔼N,β(x1=x)[η2W(2(N+F))1W]=β1𝔼N,β(x1=x)[minUNNUN2(N+F)UN2(ηW)UN].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta\mathrm{W}\cdot(A_{1}^{x})^{-1}(\eta\mathrm{W})]\leq\beta^{-1}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta^{2}\mathrm{W}\cdot(\nabla^{2}(\mathcal{H}_{N}+\mathrm{F}))^{-1}\mathrm{W}]\\ =-\beta^{-1}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\min_{U_{N}\in\mathbb{R}^{N}}U_{N}\cdot\nabla^{2}(\mathcal{H}_{N}+\mathrm{F})U_{N}-2(\eta\mathrm{W})\cdot U_{N}\Bigr{]}. (5.41)

By definition of F\mathrm{F}, for all UNNU_{N}\in\mathbb{R}^{N},

UN2(N+F)UN(NN)(s+2)εi=1N(N(ui+1ui))2U_{N}\cdot\nabla^{2}(\mathcal{H}_{N}+\mathrm{F})U_{N}\geq(N\ell_{N})^{-(s+2)\varepsilon}\sum_{i=1}^{N}(N(u_{i+1}-u_{i}))^{2}

and besides

ηWUN=i=1NηW~iN(ui+1ui)\eta\mathrm{W}\cdot U_{N}=-\sum_{i=1}^{N}\eta\tilde{\mathrm{W}}_{i}N(u_{i+1}-u_{i})

with |ηW~||\eta\tilde{\mathrm{W}}| satisfying (5.19). Inserting these into (5.41) we deduce that there exist constants C>0C>0 and κ>0\kappa>0 such that

𝔼N,β(x1=x)[ηW(A1x)1(ηW)]C(NN)κε𝔼N,β(x1=x)[η2|W~|2].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta\mathrm{W}\cdot(A_{1}^{x})^{-1}(\eta\mathrm{W})]\leq C(N\ell_{N})^{\kappa\varepsilon}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta^{2}|\tilde{\mathrm{W}}|^{2}].

Inserting the estimate (5.19) of Lemma 5.2 we obtain

𝔼N,β(x1=x)[ηW(A1x)1(ηW)]C(NN)κε+max(1,2(α+1s)).\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\eta\mathrm{W}\cdot(A_{1}^{x})^{-1}(\eta\mathrm{W})]\leq C(N\ell_{N})^{\kappa\varepsilon+\max(1,2(\alpha+1-s))}. (5.42)

Besides, recall that η\eta is a function of the gaps: η=η~GapN\eta=\tilde{\eta}\circ\mathrm{Gap}_{N}, with η~:N\tilde{\eta}:\mathbb{R}^{N}\to\mathbb{R}. Thus reasoning as for ηW\eta\mathrm{W} we get

𝔼N,β(x1=x)[(ηA~)(A1x)1(ηA~)]C(NN)κε𝔼N,β(x1=x)[|A~(η~)GapN|2].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\nabla\eta\tilde{\mathrm{A}})(A_{1}^{x})^{-1}(\nabla\eta\tilde{\mathrm{A}})]\leq C(N\ell_{N})^{\kappa\varepsilon}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[|\tilde{\mathrm{A}}(\nabla\tilde{\eta})\circ\mathrm{Gap}_{N}|^{2}].

Note that (η~)GapN=0(\nabla\tilde{\eta})\circ\mathrm{Gap}_{N}=0 on 𝒜c\mathcal{A}^{c} and that A~\tilde{\mathrm{A}} is uniformly bounded on 𝒜\mathcal{A} by (NN)κ(N\ell_{N})^{\kappa} for some κ>0\kappa>0. Thus there exist C,κ,δ>0C,\kappa,\delta>0 such that

𝔼N,β(x1=x)[|A~(η~)GapN|2]C(NN)κ𝔼N,β(x1=x)[|(η~)GapN|2]Cec(NN)δ,\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[|\tilde{\mathrm{A}}(\nabla\tilde{\eta})\circ\mathrm{Gap}_{N}|^{2}]\leq C(N\ell_{N})^{\kappa}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[|(\nabla\tilde{\eta})\circ\mathrm{Gap}_{N}|^{2}]\leq Ce^{-c(N\ell_{N})^{\delta}}, (5.43)

where we have used Theorem 1 and Lemma 4.5 in the last inequality.

Step 6: using the comparison principle for ηV\eta\mathrm{V}

Let h~1,h~2\tilde{h}_{1},\tilde{h}_{2} be integrals of h1,h2h_{1},h_{2} defined in (5.15), (5.16) of mean 0. Since η\eta is supported on 𝒜\mathcal{A}, one gets by Lemma 5.2 that

η|V|C(NN)κεFluctN[h~1+h~2],\eta|\mathrm{V}|\leq C(N\ell_{N})^{\kappa\varepsilon}\nabla\mathrm{Fluct}_{N}[\tilde{h}_{1}+\tilde{h}_{2}],

Therefore, applying Lemma 3.6, one gets

𝔼N,β(x1=x)[(ηV)(A1x)1(ηV)]C(NN)κεVarN,β(x1=x)[[FluctN[h~1+h~2]]C(NN)κε(2VarN,β(x1=x)[FluctN[h~1]]+2VarN,β(x1=x)[FluctN[h~2]]).\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\mathrm{V})\cdot(A_{1}^{x})^{-1}(\eta\mathrm{V})]\leq C(N\ell_{N})^{\kappa\varepsilon}\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[[\mathrm{Fluct}_{N}[\tilde{h}_{1}+\tilde{h}_{2}]]\\ \leq C(N\ell_{N})^{\kappa\varepsilon}(2\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]+2\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]). (5.44)

Since |FluctN[h~1]||IN||h~1|(NN)κ|\mathrm{Fluct}_{N}[\tilde{h}_{1}]|\leq|I_{N}||\tilde{h}_{1}|_{\infty}\leq(N\ell_{N})^{\kappa} for some κ>0\kappa>0 we deduce from (5.37) that there exists δ>0\delta>0 such that

VarN,β(x1)[FluctN[h~1]]=VarN,β(x1=x)[FluctN[h~1]]+O(e(NN)δ).\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1})}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]=\mathrm{Var}_{\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]+O(e^{-(N\ell_{N})^{\delta}}).

It follows that

𝔼N,β[VarN,β(x1=x)[FluctN[h~1]]]VarN,β[FluctN[h~1]]+O(e(NN)δ).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]+O(e^{-(N\ell_{N})^{\delta}}). (5.45)

Using the Poissonian control of Lemma 3.9 one can write

VarN,β[FluctN[h~1]]N|h~1|L22CN(N20Nx2dx+N2(α+1s)0N1(|x|1N)2(2s+α)dx)C(NN)max(1,2(α+1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\tilde{h}_{1}]]\leq N|\tilde{h}_{1}|_{L^{2}}^{2}\\ \leq CN\Bigr{(}\ell_{N}^{-2}\int_{0}^{\ell_{N}}x^{2}\mathrm{d}x+\ell_{N}^{2(\alpha+1-s)}\int_{0}^{\ell_{N}}\frac{1}{(|x|\vee\frac{1}{N})^{2(2-s+\alpha)}}\mathrm{d}x\Bigr{)}\leq C(N\ell_{N})^{\max(1,2(\alpha+1-s))}. (5.46)

A more involved argument is needed to deal with h~2\tilde{h}_{2}. Using (5.37) and the fact that |FluctN[h~2]||IN||h~2||\mathrm{Fluct}_{N}[\tilde{h}_{2}]|\leq|I_{N}||\tilde{h}_{2}|_{\infty}, one gets

VarN,β(x1=x)[FluctN[h~2]]=VarN,β(x1=x)[FluctN[h~2]]+O(e(NN)δ).\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]=\mathrm{Var}_{\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]+O(e^{-(N\ell_{N})^{\delta}}).

Thus, integrating the last display with respect to N,β\mathbb{P}_{N,\beta}, we obtain

𝔼N,β[VarN,β(x1)[FluctN[h~2]]]VarN,β[FluctN[h~2]]+O(e(NN)δ).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1})}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]+O(e^{-(N\ell_{N})^{\delta}}). (5.47)

By rotational invariance,

VarN,β[FluctN[h~2]]C(NN)2(α+1s)NsVarN,β[FluctN[χ]],\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]\leq C(N\ell_{N})^{2(\alpha+1-s)}N^{-s}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]], (5.48)

where χ\chi is as in (5.30).

Step 7: bound on the fluctuations of χ\chi

Let χ1\chi_{1} be as in (5.31) and let χ2:=χχ1\chi_{2}:=\chi-\chi_{1}.

Assume first that N0\ell_{N}\to 0. We have

VarN,β[FluctN[χ]]=𝔼N,β[VarN,β[FluctN[χ]x1]]+O(Ns(NN)ε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]\mid x_{1}]]+O(N^{s}(N\ell_{N})^{\varepsilon}).

Then using the fact that

0χ(x)χ1(x)for all x𝕋,0\leq\chi^{\prime}(x)\leq\chi_{1}^{\prime}(x)\quad\text{for all $x\in\mathbb{T}$},

one gets by Lemma 3.6 that

VarN,β[FluctN[χ]x1]VarN,β[FluctN[χ1]x1],\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]\mid x_{1}]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid x_{1}],

which yields

VarN,β[FluctN[χ]]VarN,β[FluctN[χ1]]+O(Ns(NN)ε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]]\leq\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]+O(N^{s}(N\ell_{N})^{\varepsilon}). (5.49)

Let us now assume that N=1\ell_{N}=1 for each NN. In this case, the support of χ\chi is 𝕋\mathbb{T}. One may write

VarN,β[FluctN[χ]x1=x]2(VarN,β[FluctN[χ1]]+VarN,β[FluctN[χ2]]).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi]\mid x_{1}=x]\leq 2(\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]). (5.50)

One can summarize (5.48), (5.49) and (5.50) into

VarN,β[FluctN[h~2]]CNs(NN)2(α+1s)(VarN,β[FluctN[χ1]]+VarN,β[FluctN[χ2]]𝟙N=1).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\tilde{h}_{2}]]\\ \leq CN^{-s}(N\ell_{N})^{2(\alpha+1-s)}(\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\mathds{1}_{\ell_{N}=1}).

Combining this with (5.44), (5.45), (5.46) and (5.47) we find

𝔼N,β(x1=x)[(ηV)(A1x)1(ηV)]C(NN)κε(NN)max(1,2(α+1s))+C(NN)κε(NN)2(α+1s)Ns(VarN,β[FluctN[χ1]]+VarN,β[FluctN[χ2]]𝟙N=1),\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\mathrm{V})\cdot(A_{1}^{x})^{-1}(\eta\mathrm{V})]\leq C(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N})^{\max(1,2(\alpha+1-s))}\\ +C(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N})^{2(\alpha+1-s)}N^{-s}\Bigr{(}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\mathds{1}_{\ell_{N}=1}\Bigr{)}, (5.51)

which concludes the proof of the proposition together with (5.34), (5.35), (5.40), (5.42), (5.43). ∎

In view of Propositions 5.3 and 5.4, there remains to bound the variance of FluctN[χ1]\mathrm{Fluct}_{N}[\chi_{1}] and FluctN[χ2]\mathrm{Fluct}_{N}[\chi_{2}] for the singular linear statistics χ1,χ2\chi_{1},\chi_{2} defined in (5.31) and (5.32). Let us emphasize that x|x|s2x\mapsto|x|^{-\frac{s}{2}} is not in H1s2H^{\frac{1-s}{2}}.

Proposition 5.5.

Let χ1\chi_{1} be as in (5.31). For all ε>0\varepsilon>0, there exists C>0C>0 depending on ε\varepsilon and γ\gamma such that

VarN,β[FluctN[χ1]]CNs(NN)ε.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]\leq CN^{s}(N\ell_{N})^{\varepsilon}. (5.52)
Proof.
Step 1: conditioning on a point near the singularity

Since χ1\chi_{1} is very singular at 0, we need to condition on a point near 0. Let ε>0\varepsilon>0. Let us define

J:={i{1,,N},|xi|(NN)εN}.J:=\Bigr{\{}i\in\{1,\ldots,N\},|x_{i}|\leq\frac{(N\ell_{N})^{\varepsilon}}{N}\Bigr{\}}.

By Theorem 1, there exists δ>0\delta>0 depending on ε\varepsilon such that

𝔼N,β[FluctN[χ1]2]=𝔼N,β[FluctN[χ1]2𝟙J]+O(e(NN)δ).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}\mathds{1}_{J\neq\emptyset}]+O(e^{-(N\ell_{N})^{\delta}}).

Moreover

𝟙Jj=1N𝟙jJ.\mathds{1}_{J\neq\emptyset}\leq\sum_{j=1}^{N}\mathds{1}_{j\in J}.

Thus

𝔼N,β[FluctN[χ1]2]j=1N𝔼N,β[FluctN[χ1]2jJ]N,β(jJ)+O(e(NN)δ).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}]\leq\sum_{j=1}^{N}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}\mid j\in J]\mathbb{P}_{N,\beta}(j\in J)+O(e^{-(N\ell_{N})^{\delta}}).

Since FluctN[χ1]\mathrm{Fluct}_{N}[\chi_{1}] is invariant by permutation of indices, one can observe that for each j=1,,Nj=1,\ldots,N,

𝔼N,β[FluctN[χ1]2jJ]=𝔼N,β[FluctN[χ1]21J].\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}\mid j\in J]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}\mid 1\in J].

Moreover since xjx_{j} is uniformly distributed on the circle, one has

N,β(jJ)=2(NN)εN.\mathbb{P}_{N,\beta}(j\in J)=\frac{2(N\ell_{N})^{\varepsilon}}{N}.

Since χ1=0\int\chi_{1}=0, it follows that

VarN,β[FluctN[χ1]]=𝔼N,β[FluctN[χ1]2]2(NN)ε𝔼N,β[FluctN[χ1]21J]+O(e(NN)δ)=2(NN)ε(VarN,β[FluctN[χ1]1J]+𝔼N,β[FluctN[χ1]1J]2)+O(e(NN)δ).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}]\leq 2(N\ell_{N})^{\varepsilon}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}\mid 1\in J]\\ +O(e^{-(N\ell_{N})^{\delta}})=2(N\ell_{N})^{\varepsilon}(\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid 1\in J]+\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid 1\in J]^{2})\\ +O(e^{-(N\ell_{N})^{\delta}}).

As proved in Lemma B.6,

𝔼N,β[FluctN[χ1]1J]=O(Ns2(NN)ε).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid 1\in J]=O(N^{\frac{s}{2}}(N\ell_{N})^{\varepsilon}).

Inserting this into the last display, we get that for all ε>0\varepsilon>0,

VarN,β[FluctN[χ1]]2(NN)εVarN,β[FluctN[χ1]x1J]+O((NN)εNs).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]]\leq 2(N\ell_{N})^{\varepsilon}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid x_{1}\in J]+O((N\ell_{N})^{\varepsilon}N^{s}). (5.53)

Fix x𝕋,|x|(NN)εNx\in\mathbb{T},|x|\leq\frac{(N\ell_{N})^{\varepsilon}}{N}. In the sequel, we will control the variance of FluctN[χ1]\mathrm{Fluct}_{N}[\chi_{1}] conditionally on x1=xx_{1}=x.

Step 2: convexification and reduction to a good event

Let ε>0\varepsilon>0 and N,β\mathbb{Q}_{N,\beta} be as in (5.36). There exists δ>0\delta>0 depending on ε\varepsilon such that

VarN,β(x1=x)[FluctN[χ1]]=VarN,β(x1=x)[FluctN[χ1]]+O(e(NN)δ).\mathrm{Var}_{\mathbb{P}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\chi_{1}]]=\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\chi_{1}]]+O(e^{-(N\ell_{N})^{\delta}}). (5.54)

Let η\eta be the cutoff function defined in (5.39) and 𝒜\mathcal{A} be the event (5.14). Let us recall

A1x:=A1N,β(x1=x).A_{1}^{x}:=A_{1}^{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}.

By subadditivity, using that A1xA_{1}^{x} is non-negative on L2({1,,N},N,β(x1=x))L^{2}(\{1,\ldots,N\},\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)), we get

VarN,β[ηFluctN[χ1]x1=x]=𝔼N,β(x1=x)[(ηFluctN[χ1])(A1x)1(ηFluctN[χ1])]2𝔼N,β(x1=x)[(ηFluctN[χ1])(A1x)1(ηFluctN[χ1])]+2𝔼N,β(x1=x)[(FluctN[χ1]η)(A1x)1(FluctN[χ1]η)]\mathrm{Var}_{\mathbb{Q}_{N,\beta}}[\eta\mathrm{Fluct}_{N}[\chi_{1}]\mid x_{1}=x]=\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\nabla(\eta\mathrm{Fluct}_{N}[\chi_{1}])\cdot(A_{1}^{x})^{-1}\nabla(\eta\mathrm{Fluct}_{N}[\chi_{1}])]\\ \leq 2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\nabla\mathrm{Fluct}_{N}[\chi_{1}])\cdot(A_{1}^{x})^{-1}(\eta\nabla\mathrm{Fluct}_{N}[\chi_{1}])]\\ +2\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)\cdot(A_{1}^{x})^{-1}(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)] (5.55)

We proceed as in Step 4 of the proof of Proposition 5.4. Let us write η\eta as η~GapN\tilde{\eta}\circ\mathrm{Gap}_{N} with η~:N\tilde{\eta}:\mathbb{R}^{N}\to\mathbb{R}. Using the log-concavity of GapN#N,β\mathrm{Gap}_{N}\#\mathbb{Q}_{N,\beta}, one can write

𝔼N,β(x1=x)[(FluctN[χ1]η)(A1x)1(FluctN[χ1]η)](NN)κε𝔼N,β(x1=x)[FluctN[χ1]2|η~GapN|2]\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)\cdot(A_{1}^{x})^{-1}(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)]\\ \leq(N\ell_{N})^{\kappa\varepsilon}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{Fluct}_{N}[\chi_{1}]^{2}|\nabla\tilde{\eta}\circ\mathrm{Gap}_{N}|^{2}]

Recall that η~\nabla\tilde{\eta} is supported on 𝒜\mathcal{A}. Moreover, there exists C,κ>0C,\kappa>0 such that on the event 𝒜\mathcal{A},

FluctN[χ1]2C(NN)κ.\mathrm{Fluct}_{N}[\chi_{1}]^{2}\leq C(N\ell_{N})^{\kappa}.

It follows that

𝔼N,β(x1=x)[(FluctN[χ1]η)(A1x)1(FluctN[χ1]η)]C(NN)κε𝔼N,β(x1=x)[|η~GapN|2].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)\cdot(A_{1}^{x})^{-1}(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)]\leq C(N\ell_{N})^{\kappa\varepsilon}\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[|\nabla\tilde{\eta}\circ\mathrm{Gap}_{N}|^{2}].

Using Theorem 1 and (5.37), we find that there exists C>0C>0 and δ>0\delta>0 such that

𝔼N,β(x1=x)[(FluctN[χ1]η)(A1x)1(FluctN[χ1]η)]Ce(NN)δ.\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)\cdot(A_{1}^{x})^{-1}(\mathrm{Fluct}_{N}[\chi_{1}]\nabla\eta)]\leq Ce^{-(N\ell_{N})^{\delta}}. (5.56)
Step 3: control on the main term

Define

V:=ηNs2FluctN[χ1].\mathrm{V}:=\eta N^{-\frac{s}{2}}\nabla\mathrm{Fluct}_{N}[\chi_{1}].

There exists K0>0,C>0,C>0,κ>0K_{0}>0,C>0,C^{\prime}>0,\kappa>0 such that for each iINi\in I_{N}, we have

|Vi|CηNs21(|xi|1N)1+s2𝟙d(i,1)K0(NN)γ=CηN1(N|xi|1)1+s2𝟙d(i,1)K0(NN)γ.|\mathrm{V}_{i}|\leq C\eta N^{-\frac{s}{2}}\frac{1}{(|x_{i}|\vee\frac{1}{N})^{1+\frac{s}{2}}}\mathds{1}_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}=C\eta\frac{N^{-1}}{(N|x_{i}|\vee 1)^{1+\frac{s}{2}}}\mathds{1}_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}.

Recall that on the event 𝒜\mathcal{A}, there exists K>0K>0 and κ>0\kappa>0 such that for each ii such that d(i,1)K(NN)γd(i,1)\leq K(N\ell_{N})^{\gamma},

N|xix1|1(NN)κεd(i,1)+1.N|x_{i}-x_{1}|\vee 1\geq(N\ell_{N})^{-\kappa\varepsilon}d(i,1)+1.

Therefore if |xi|>(NN)εN|x_{i}|>\frac{(N\ell_{N})^{\varepsilon}}{N}, then on 𝒜\mathcal{A},

1(N|xi|1)1+s2(NN)κεd(i,1)1+s2+1.\frac{1}{(N|x_{i}|\vee 1)^{1+\frac{s}{2}}}\leq\frac{(N\ell_{N})^{\kappa\varepsilon}}{d(i,1)^{1+\frac{s}{2}}+1}.

Moreover, on the event 𝒜\mathcal{A}, there exists κ>0\kappa>0 such that d(i,1)(NN)κεd(i,1)\geq(N\ell_{N})^{\kappa\varepsilon} implies |xi|(NN)εN|x_{i}|\geq\frac{(N\ell_{N})^{\varepsilon}}{N}. Moreover, since there are at most O((NN)2ε)O((N\ell_{N})^{2\varepsilon}) points in ((NN)εN,(NN)εN)(-\frac{(N\ell_{N})^{\varepsilon}}{N},\frac{(N\ell_{N})^{\varepsilon}}{N}). We conclude that there exists C,κ,K0>0C,\kappa,K_{0}>0 such that

|Vi|C(NN)κεN1(d(i,1)+1)1+s2𝟙d(i,1)K0(NN)γ.|\mathrm{V}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}\frac{N^{-1}}{(d(i,1)+1)^{1+\frac{s}{2}}}\mathds{1}_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}.

Applying the comparison principle of Lemma 3.6, we thus obtain

𝔼N,β(x1=x)[V(A1x)1V]C(NN)κεVarN,β(x1=x)[i:d(i,1)K0(NN)γNxi(d(i,1)+1)1+s2].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[\mathrm{V}\cdot(A_{1}^{x})^{-1}\mathrm{V}]\leq C(N\ell_{N})^{\kappa\varepsilon}\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\sum_{i:d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{Nx_{i}}{(d(i,1)+1)^{1+\frac{s}{2}}}\Bigr{]}. (5.57)

One can write

VarN,β(x1=x)[d(i,1)K0(NN)γNxi(1+d(i,1))1+s2]=𝔼N,β(x1=x)[(d(i,1)K0(NN)γN(xi𝔼N,β(x1=x)[xi])(d(i,1)+1)1+s2)2].\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{Nx_{i}}{(1+d(i,1))^{1+\frac{s}{2}}}\Bigr{]}\\ =\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\Bigr{(}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{N(x_{i}-\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[x_{i}])}{(d(i,1)+1)^{1+\frac{s}{2}}}\Bigr{)}^{2}\Bigr{]}. (5.58)

Then, using Cauchy-Schwarz inequality, one can bound the last display by

𝔼N,β(x1=x)[(d(i,1)K0(NN)γN(xi𝔼N,β(x1=x)[xi])(d(i,1)+1)1+s2)2]Clog|IN|𝔼N,β(x1=x)[d(i,1)K0(NN)γVarN,β(x1=x)[Nxi](d(i,1)+1)1+s].\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\Bigr{(}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{N(x_{i}-\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[x_{i}])}{(d(i,1)+1)^{1+\frac{s}{2}}}\Bigr{)}^{2}\Bigr{]}\\ \leq C\log|I_{N}|\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[Nx_{i}]}{(d(i,1)+1)^{1+s}}\Bigr{]}.

Moreover,

VarN,β(x1=x)[Nxi]=VarN,β(x1=x)[N(xix1)]+O((NN)κε).\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[Nx_{i}]=\mathrm{Var}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[N(x_{i}-x_{1})]+O((N\ell_{N})^{\kappa\varepsilon}).

Inserting the concentration estimate of Theorem 1 in the last display we get

𝔼N,β(x1=x)[(d(i,1)K0(NN)γN(xi𝔼N,β(x1=x)[xi])(d(i,1)+1)1+s2)2]C(NN)κεd(i,1)K0(NN)γ1(d(i,1)+1)C(NN)κε.\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}\Bigr{[}\Bigr{(}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{N(x_{i}-\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[x_{i}])}{(d(i,1)+1)^{1+\frac{s}{2}}}\Bigr{)}^{2}\Bigr{]}\\ \leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{d(i,1)\leq K_{0}(N\ell_{N})^{\gamma}}\frac{1}{(d(i,1)+1)}\leq C^{\prime}(N\ell_{N})^{\kappa^{\prime}\varepsilon}.

Combined with (5.57) and (5.58) this yields

𝔼N,β(x1=x)[(ηFluctN[χ1])(A1x)1(ηFluctN[χ1])]C(NN)κεNs.\mathbb{E}_{\mathbb{Q}_{N,\beta}(\cdot\mid x_{1}=x)}[(\eta\nabla\mathrm{Fluct}_{N}[\chi_{1}])\cdot(A_{1}^{x})^{-1}(\eta\nabla\mathrm{Fluct}_{N}[\chi_{1}])]\leq C(N\ell_{N})^{\kappa\varepsilon}N^{s}. (5.59)
Step 4: conclusion

Putting (5.59) and (5.56) into (5.55) and using (5.54), we get that there exists C>0C>0 and κ>0\kappa>0 such that for all |x|<(NN)εN|x|<\frac{(N\ell_{N})^{\varepsilon}}{N},

VarN,β[i=1Nχ1(xi)x1=x]CNs(NN)κε.\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\chi_{1}(x_{i})\mid x_{1}=x\Bigr{]}\leq CN^{s}(N\ell_{N})^{\kappa\varepsilon}.

Therefore using (5.53), one concludes that

VarN,β[i=1Nχ1(xi)]CNs(NN)κε.\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\chi_{1}(x_{i})\Bigr{]}\leq CN^{s}(N\ell_{N})^{\kappa\varepsilon}. (5.60)

In the case where N=1\ell_{N}=1 for each NN, one should bound the fluctuations of the discontinuous function χ2\chi_{2} defined in (5.32). To this end, we use a bootstrap argument and apply Propositions 5.3, 5.4, 5.5. After a finite number of steps, we obtain a sharp variance bound, allowing one to conclude the proof of the theorem.

Proof of Theorem 2.

Let χ2\chi_{2} be as in (5.32). Recall that χ2\chi_{2} satisfies Assumptions 1.1 with α:=0<s2\alpha:=0<\frac{s}{2}. Therefore by applying Propositions 5.3 and 5.4, we find that

VarN,β[FluctN[χ2]]CNs+Ns+ε(VarN,β[FluctN[θ1]]+VarN,β[FluctN[θ2]]),\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\leq CN^{s}+N^{-s+\varepsilon}(\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\theta_{1}]]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\theta_{2}]]),

with θ1\theta_{1} as in (5.31) for t=1/4t=1/4 and θ2:𝕋\theta_{2}:\mathbb{T}\to\mathbb{R} satisfying Assumptions 1.1 with α:=0<s2\alpha:=0<\frac{s}{2}. Inserting (5.52) we deduce that for all ε>0\varepsilon>0 there exists C>0C>0 such that

VarN,β[FluctN[χ2]]CNs+Ns+εVarN,β[FluctN[θ2]].\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\leq CN^{s}+N^{-s+\varepsilon}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\theta_{2}]]. (5.61)

One may bootstrap through (5.61) the optimal fluctuation estimate on χ2\chi_{2}, since χ2\chi_{2} and θ2\theta_{2} satisfy the same assumptions. Inserting a crude Poissonian estimate, we get

VarN,β[FluctN[θ2]]N|θ2|L22CN.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\theta_{2}]]\leq N|\theta_{2}|_{L^{2}}^{2}\leq CN.

Substituting this into (5.61), one obtains that there exists C>0C>0 such that

VarN,β[FluctN[χ2]]C(Ns+N12s).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\leq C(N^{s}+N^{1-2s}). (5.62)

Since χ2\chi_{2} and θ2\theta_{2} satisfy the same assumptions, one can insert the estimate (5.62) into (5.61) and we find that for all ε>0\varepsilon>0 there exists C>0C>0 such that

VarN,β[FluctN[χ2]]C(Ns+N13s+ε).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\leq C(N^{s}+N^{1-3s+\varepsilon}).

We conclude after a finite number of steps that

VarN,β[FluctN[χ2]]CNs.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{2}]]\leq CN^{s}.

Combining this with the estimate (5.33) of Proposition 5.4 and the estimate (5.52) of Proposition 5.5, one gets that for all ε>0\varepsilon>0, there exists C>0C>0 such that

VarN,β[AN[ψreg]]C(NN)ε+max(1,2(α+1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{\varepsilon+\max(1,2(\alpha+1-s))}. (5.63)

Finally using (5.21) we conclude that

VarN,β[FluctN[ξ(N1)]]=(NN)sσN2(ξ)+O((NN)max(2s1,2α)+ε),\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]]=(N\ell_{N})^{s}\sigma_{\ell_{N}}^{2}(\xi)\\ +O\Bigr{(}(N\ell_{N})^{\max(2s-1,2\alpha)+\varepsilon}\Bigr{)},

which finishes the proof of Theorem 2. ∎

6. Central Limit Theorem

6.1. Proof of the CLT

We prove the CLT stated in Theorem 3 using Stein’s method, leveraging on the variance estimates of Section 5.

Proof of Theorem 3.

Let ξ\xi satisfy Assumptions 1.1. Let {N}\{\ell_{N}\} such that N1N\ell_{N}\gg\frac{1}{N}. Assume either that ξ\xi is supported on (12,12)(-\frac{1}{2},\frac{1}{2}) or that N=1\ell_{N}=1 for each NN. Let

FN:=(NN)s2FluctN[ξ(N1)].F_{N}:=(N\ell_{N})^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)].

The principle of Stein method is to prove that for all η:\eta:\mathbb{R}\to\mathbb{R} smooth enough,

𝔼N,β[η(FN)FN]σN2(ξ)𝔼N,β[η(FN)],\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(F_{N})F_{N}]\simeq\sigma_{\ell_{N}}^{2}(\xi)\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta^{\prime}(F_{N})], (6.1)

where σN2(ξ)\sigma_{\ell_{N}}^{2}(\xi) is as in (1.10). Let indeed Z𝒩(0,σN2(ξ))Z\sim\mathcal{N}(0,\sigma_{\ell_{N}}^{2}(\xi)) and h:h:\mathbb{R}\to\mathbb{R} smooth enough. Denote σ2:=σN2(ξ)\sigma^{2}:=\sigma_{\ell_{N}}^{2}(\xi). Consider the unique bounded solution η:\eta:\mathbb{R}\to\mathbb{R} of the ODE

σ2η(x)η(x)x=h(x)𝔼[h(Z)],\sigma^{2}\eta^{\prime}(x)-\eta(x)x=h(x)-\mathbb{E}[h(Z)],

which is explicitly given by

η:xex22σ20xey22σ2(h(y)𝔼[h(Z)])dy).\eta:x\in\mathbb{R}\mapsto e^{\frac{x^{2}}{2\sigma^{2}}}\int_{0}^{x}e^{-\frac{y^{2}}{2\sigma^{2}}}(h(y)-\mathbb{E}[h(Z)])\mathrm{d}y).

One can observe that there exists a constant C>0C>0 depending only on σ2\sigma^{2} such that

|η|+|η|+|η′′|C(|h|+|h|+|h′′|).|\eta|_{\infty}+|\eta^{\prime}|_{\infty}+|\eta^{\prime\prime}|_{\infty}\leq C(|h|_{\infty}+|h^{\prime}|_{\infty}+|h^{\prime\prime}|_{\infty}).

Therefore, since (σN2(ξ))(\sigma_{\ell_{N}}^{2}(\xi)) is bounded, there exists a constant C>0C>0 independent of NN such that

supη𝒟|𝔼N,β[η(FN)]𝔼[η(Z)]|C(supη𝒟|𝔼N,β[σN2(ξ)η(FN)FNη(FN)]|)12,\sup_{\eta\in\mathcal{D}}|\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(F_{N})]-\mathbb{E}[\eta(Z)]|\leq C\Bigr{(}\sup_{\eta\in\mathcal{D}}|\mathbb{E}_{\mathbb{P}_{N,\beta}}[\sigma_{\ell_{N}}^{2}(\xi)\eta^{\prime}(F_{N})-F_{N}\eta(F_{N})]|\Bigr{)}^{\frac{1}{2}}, (6.2)

where 𝒟\mathcal{D} is the set of differentiable functions η:\eta:\mathbb{R}\to\mathbb{R} such that |η|1|\eta|_{\infty}\leq 1, |η|1|\eta^{\prime}|_{\infty}\leq 1, |η′′|1|\eta^{\prime\prime}|_{\infty}\leq 1. Let us now prove (6.1).

Step 1: regularization

Let :=(NN)1\ell:=(N\ell_{N})^{-1} and ξreg=ξK\xi_{\mathrm{reg}}=\xi*K_{\ell} with KK_{\ell} as in (2.22). Let

GN:=(NN)s2FluctN[ξreg(N1)]G_{N}:=(N\ell_{N})^{-\frac{s}{2}}\mathrm{Fluct}_{N}[\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)]

Let η𝒟\eta\in\mathcal{D}. We have

|𝔼N,β[η(FN)FNσN2(ξ)η(FN)]𝔼N,β[η(GN)GNσN2(ξreg)η(GN)]|C(1(NN)s/2𝔼N,β[|FluctN[ξξreg]|]+|σN2(ξ)σN2(ξreg)|)C(1(NN)s/2VarN,β[FluctN[ξξreg]]12+|σN2(ξ)σN2(ξreg)|)C((NN)12s2|ξξreg|L2+|σN2(ξ)σN2(ξreg)|)=O((NN)max(αs2,s12)),|\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(F_{N})F_{N}-\sigma_{\ell_{N}}^{2}(\xi)\eta^{\prime}(F_{N})]-\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})G_{N}-\sigma_{\ell_{N}}^{2}(\xi_{\mathrm{reg}})\eta^{\prime}(G_{N})]|\\ \leq C\Bigr{(}\frac{1}{(N\ell_{N})^{s/2}}\mathbb{E}_{\mathbb{P}_{N,\beta}}[|\mathrm{Fluct}_{N}[\xi-\xi_{\mathrm{reg}}]|]+|\sigma_{N}^{2}(\xi)-\sigma_{N}^{2}(\xi_{\mathrm{reg}})|\Bigr{)}\\ \leq C\Bigr{(}\frac{1}{(N\ell_{N})^{s/2}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\xi-\xi_{\mathrm{reg}}]]^{\frac{1}{2}}+|\sigma_{N}^{2}(\xi)-\sigma_{N}^{2}(\xi_{\mathrm{reg}})|\Bigr{)}\\ \leq C((N\ell_{N})^{\frac{1}{2}-\frac{s}{2}}|\xi-\xi_{\mathrm{reg}}|_{L^{2}}+|\sigma_{N}^{2}(\xi)-\sigma_{N}^{2}(\xi_{\mathrm{reg}})|)=O((N\ell_{N})^{\max(\alpha-\frac{s}{2},\frac{s-1}{2})}), (6.3)

where we have used the regularization estimates of Lemma 2.5.

Step 2: main computation

We wish to prove that GNG_{N} approximately solves (6.1). Let ψreg𝒞2(N1𝕋,)\psi_{\mathrm{reg}}\in\mathcal{C}^{2}(\ell_{N}^{-1}\mathbb{T},\mathbb{R}) satisfy

ψreg=12βcsN1s(Δ)1s2(ξreg(N1))(N)andψreg=0.\psi_{\mathrm{reg}}^{\prime}=-\frac{1}{2\beta c_{s}}\ell_{N}^{1-s}(-\Delta)^{\frac{1-s}{2}}(\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot))(\ell_{N}\cdot)\quad\text{and}\quad\textstyle\int\psi_{\mathrm{reg}}=0.

Define

Ψ:XNDN1(NN)1s2N(ψreg(N1x1),,ψreg(N1xN)),\Psi:X_{N}\in D_{N}\mapsto\frac{1}{(N\ell_{N})^{1-\frac{s}{2}}}\ell_{N}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{1}),\ldots,\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{N})),

which can be written Ψ=Φ\Psi=\nabla\Phi for some 𝒞3\mathcal{C}^{3} function Φ:DN\Phi:D_{N}\to\mathbb{R}. Let :=N,β\mathcal{L}:=\mathcal{L}^{\mathbb{P}_{N,\beta}}. Recall that by (5.7),

ΦGN=1(NN)1s2(βAN[ψreg]FluctN[ψreg(N1)].\mathcal{L}\Phi-G_{N}=\frac{1}{(N\ell_{N})^{1-\frac{s}{2}}}(\beta\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)].

Therefore

𝔼N,β[η(GN)GN]=𝔼N,β[η(GN)Φ]1(NN)1s2CovN,β[η(GN),βAN[ψreg]FluctN[ψreg(N1)]]=𝔼N,β[η(GN)GNΦ]1(NN)1s2CovN,β[η(GN),βAN[ψreg]FluctN[ψreg(N1)]]=σN2(ξreg)𝔼N,β[η(GN)]+ErrorN1+ErrorN2,\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})G_{N}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(G_{N})\mathcal{L}\Phi]\\ -\frac{1}{(N\ell_{N})^{1-\frac{s}{2}}}\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[\eta(G_{N}),\beta\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta^{\prime}(G_{N})\nabla G_{N}\cdot\nabla\Phi]\\ -\frac{1}{(N\ell_{N})^{1-\frac{s}{2}}}\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[\eta(G_{N}),\beta\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]]\\ =\sigma_{\ell_{N}}^{2}(\xi_{\mathrm{reg}})\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta^{\prime}(G_{N})]+\mathrm{Error}_{N}^{1}+\mathrm{Error}_{N}^{2}, (6.4)

where

ErrorN1:=𝔼N,β[η(GN)(i=1N1NNψreg(N1xi)ξreg(N1xi)σN2(ξreg))],\mathrm{Error}_{N}^{1}:=\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\eta^{\prime}(G_{N})\Bigr{(}\sum_{i=1}^{N}\frac{1}{N\ell_{N}}\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})-\sigma_{\ell_{N}}^{2}(\xi_{\mathrm{reg}})\Bigr{)}\Bigr{]},
ErrorN2:=1(NN)1s2CovN,β[η(GN),(βAN[ψreg]FluctN[ψreg(N1)]].\mathrm{Error}_{N}^{2}:=-\frac{1}{(N\ell_{N})^{1-\frac{s}{2}}}\mathrm{Cov}_{\mathbb{P}_{N,\beta}}[\eta(G_{N}),(\beta\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]].
Step 3: the error term ErrorN1\mathrm{Error}_{N}^{1}

One can bound ErrorN1\mathrm{Error}_{N}^{1} by

|ErrorN1||η|1NN𝔼N,β[|i=1Nξreg(N1xi)ψreg(N1xi)NNσN2(ξreg)|].|\mathrm{Error}_{N}^{1}|\leq|\eta^{\prime}|_{\infty}\frac{1}{N\ell_{N}}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\Bigr{|}\sum_{i=1}^{N}\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})-N\ell_{N}\sigma_{\ell_{N}}^{2}(\xi_{\mathrm{reg}})\Bigr{|}\Bigr{]}.

Since

1N1sψreg(N1)=12βcs(Δ)1s2(ξreg(N1)),\frac{1}{\ell_{N}^{1-s}}\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)=-\frac{1}{2\beta c_{s}}(-\Delta)^{\frac{1-s}{2}}(\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)),

we get using (2.14) that

𝔼N,β[i=1Nξreg(N1xi)ψreg(N1xi)]=NN1s|ξreg(N1)|2H1s2=NNσ2N(ξreg).\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})\Bigr{]}=N\ell_{N}^{1-s}|\xi_{\mathrm{reg}}(\ell_{N}^{-1}\cdot)|^{2}_{H^{\frac{1-s}{2}}}=N\ell_{N}\sigma^{2}_{\ell_{N}}(\xi_{\mathrm{reg}}).

We therefore obtain

|ErrorN1|C|η|1NNVarN,β[i=1Nξreg(N1xi)ψreg(N1xi)]12.|\mathrm{Error}_{N}^{1}|\leq C|\eta^{\prime}|_{\infty}\frac{1}{N\ell_{N}}\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})\Bigr{]}^{\frac{1}{2}}.

By the Poissonian estimate of Lemma 3.9,

VarN,β[i=1Nξreg(N1xi)ψreg(N1xi)]12(NN)12|ψregξreg|L2.\mathrm{Var}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\xi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}x_{i})\psi_{\mathrm{reg}}(\ell_{N}^{-1}x_{i})\Bigr{]}^{\frac{1}{2}}\leq(N\ell_{N})^{\frac{1}{2}}|\psi_{\mathrm{reg}}\ \xi_{\mathrm{reg}}^{\prime}|_{L^{2}}.

By Assumptions 1.1, ψreg\psi_{\mathrm{reg}} is O(1)O(1) around a singularity aa of ξreg\xi_{\mathrm{reg}} while ξreg\xi_{\mathrm{reg}}^{\prime} grows in O(|xa|(1+α))O(|x-a|^{-(1+\alpha)}). It follows that

|ψregξreg|L22C(1+1NN11|x|2(1+α)dx)C(1+(NN)1+2α).|\psi_{\mathrm{reg}}\ \xi_{\mathrm{reg}}^{\prime}|_{L^{2}}^{2}\leq C\Bigr{(}1+\int_{\frac{1}{N\ell_{N}}}^{1}\frac{1}{|x|^{2(1+\alpha)}}\mathrm{d}x\Bigr{)}\leq C(1+(N\ell_{N})^{1+2\alpha}).

Hence

|ErrorN1|C(NN)κε((NN)1+(NN)α12)C(NN)s12,|\mathrm{Error}_{N}^{1}|\leq C(N\ell_{N})^{\kappa\varepsilon}((N\ell_{N})^{-1}+(N\ell_{N})^{\alpha-\frac{1}{2}})\leq C(N\ell_{N})^{\frac{s-1}{2}}, (6.5)

since α<s2\alpha<\frac{s}{2}.

Step 4: the error term ErrorN2\mathrm{Error}_{N}^{2}

To bound ErrorN2\mathrm{Error}_{N}^{2}, one can write

|ErrorN2|2β(NN)1s2|η|VarN,β[AN[ψreg]]12+2(NN)1s2|η|VarN,β[FluctN[ψreg(N1)]]12.|\mathrm{Error}_{N}^{2}|\leq\frac{2\beta}{(N\ell_{N})^{1-\frac{s}{2}}}|\eta|_{\infty}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]^{\frac{1}{2}}+\frac{2}{(N\ell_{N})^{1-\frac{s}{2}}}|\eta|_{\infty}\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]]^{\frac{1}{2}}. (6.6)

Inserting (5.63), one gets

VarN,β[AN[ψreg]]C(NN)κε+max(1,2(α+1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{\kappa\varepsilon+\max(1,2(\alpha+1-s))}.

Moreover from the Poissonian of Lemma 3.9,

VarN,β[FluctN[ψreg(N1)]]NN(ψreg)2C(NN)max(1,2(α+1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\psi_{\mathrm{reg}}^{\prime}(\ell_{N}^{-1}\cdot)]]\leq N\ell_{N}\int(\psi_{\mathrm{reg}}^{\prime})^{2}\leq C(N\ell_{N})^{\max(1,2(\alpha+1-s))}.

It follows that

|ErrorN2|C|η|(NN)κε+max(s12,αs2).|\mathrm{Error}_{N}^{2}|\leq C|\eta|_{\infty}(N\ell_{N})^{\kappa\varepsilon+\max(\frac{s-1}{2},\alpha-\frac{s}{2})}. (6.7)
Step 5: conclusion

Inserting (6.5), (6.7) into (6.4) and using (6.3) one obtains

|𝔼N,β[η(FN)FNσN2(ξ)η(FN)]|C(NN)κε+max(s12,αs2).|\mathbb{E}_{\mathbb{P}_{N,\beta}}[\eta(F_{N})F_{N}-\sigma_{\ell_{N}}^{2}(\xi)\eta^{\prime}(F_{N})]|\leq C(N\ell_{N})^{\kappa\varepsilon+\max(\frac{s-1}{2},\alpha-\frac{s}{2})}.

Since α<s2\alpha<\frac{s}{2}, the above term is oNN(1)o_{N\ell_{N}}(1). Combined with (6.2) and (5.22) this concludes the proof of Theorem 3. ∎

6.2. Proof of Corollary 1.1

Proof of Corollary 1.1.

By Lemma 2.3, the function ξ:=𝟙(a,a)\xi:=\mathds{1}_{(-a,a)} satisfies Assumptions 1.1 and one may apply Theorem 3. Let {N}\{\ell_{N}\} be a sequence in (0,1](0,1]. Let ξ0:,ξ0=𝟙(a,a)\xi_{0}:\mathbb{R}\to\mathbb{R},\xi_{0}=\mathds{1}_{(-a,a)}. Define

σ2(ξ):=12βcs{|ξ|H1s22if N=1 for each N|ξ0|H1s22if N0, with ξ0 as in (1.11).\sigma_{\infty}^{2}(\xi):=\frac{1}{2\beta c_{s}}\begin{cases}|\xi|_{H^{\frac{1-s}{2}}}^{2}&\text{if $\ell_{N}=1$ for each $N$}\\ |\xi_{0}|^{2}_{H^{\frac{1-s}{2}}}&\text{if $\ell_{N}\to 0$, with $\xi_{0}$ as in (\ref{eq:defxi0})}.\end{cases} (6.8)

By formula (2.20), if N=1\ell_{N}=1 for each NN,

σ2(ξ)=cotan(π2s)4s1βπsζ(s,2a).\sigma_{\infty}^{2}(\xi)=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{4^{s-1}\beta\pi s}\zeta(-s,2a). (6.9)

Now in the case where N0\ell_{N}\to 0, by expanding (6.9) as aa tends to 0, we find

σ2(ξ)=cotan(π2s)βπs4s1(2a)s=cotan(π2s)βπ2sas.\sigma_{\infty}^{2}(\xi)=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta\pi s4^{s-1}}(2a)^{s}=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta\frac{\pi}{2}s}a^{s}.

One can synthesize the convergence result as

Ns2ζ(s,2aN)12FluctN[𝟙(a,a)(N^{-\frac{s}{2}}\zeta(-s,2a\ell_{N})^{-\frac{1}{2}}\mathrm{Fluct}_{N}[\mathds{1}_{(-a,a)}(\ell_{N}^{-1}\cdot)]\underset{\mathrm{Law}}{\Longrightarrow}\mathcal{N}(0,\sigma^{2}), (6.10)

where

σ2:=cotan(π2s)βπs4s1.\sigma^{2}:=\frac{\mathrm{cotan}(\frac{\pi}{2}s)}{\beta\pi s4^{s-1}}. (6.11)

Let {kN}\{k_{N}\} be a sequence in {1,,N2}\{1,\ldots,\frac{N}{2}\} such that kNk_{N}\to\infty. Let

i0=argmin1iN|xi|.i_{0}=\underset{1\leq i\leq N}{\mathrm{argmin}}|x_{i}|.

Let us prove that kNs2(N(xi0+kNxi0)kN)k_{N}^{-\frac{s}{2}}(N(x_{i_{0}+k_{N}}-x_{i_{0}})-k_{N}) converges in distribution. Let ε>0\varepsilon>0. Denote uN\mathrm{u}^{N} the regular grid on 𝕋\mathbb{T}

ukN=kN,for each k=1,,N.\mathrm{u}_{k}^{N}=\frac{k}{N},\quad\text{for each $k=1,\ldots,N$}.

Define the event

𝒜={XNDN:|Nxi0|kNε,i,j{k:d(k,i0)2kN},|N(xjxi)NujiN|d(j,i)s2+ε}.\mathcal{A}=\{X_{N}\in D_{N}:|Nx_{i_{0}}|\leq k_{N}^{\varepsilon},\forall i,j\in\{k:d(k,i_{0})\leq 2k_{N}\},|N(x_{j}-x_{i})-N\mathrm{u}_{j-i}^{N}|\leq d(j,i)^{\frac{s}{2}+\varepsilon}\}.

By Theorem 1, there exists δ>0\delta>0 depending on ε\varepsilon such that

N,β(𝒜c)CeckNδ.\mathbb{P}_{N,\beta}(\mathcal{A}^{c})\leq Ce^{-ck_{N}^{\delta}}.

Moreover, on the event 𝒜\mathcal{A}, we have

N(xkN+i0xi0)kN=(i=1N𝟙(0,N)(xi)NN)+O(kNε+kN(s2+ε)2),N(x_{k_{N}+i_{0}}-x_{i_{0}})-k_{N}=-\Bigr{(}\sum_{i=1}^{N}\mathds{1}_{(0,\ell_{N})}(x_{i})-N\ell_{N}\Bigr{)}+O(k_{N}^{\varepsilon}+k_{N}^{(\frac{s}{2}+\varepsilon)^{2}}), (6.12)

where N:=kNN\ell_{N}:=\frac{k_{N}}{N}. Let indeed

ΔN:=NxkNkN.\Delta_{N}:=\lfloor Nx_{k_{N}}-k_{N}\rfloor.

We have

N(xi0+kNxi0+kNΔN)=ΔN+O(ΔNs2+ε).N(x_{i_{0}+k_{N}}-x_{i_{0}+k_{N}-\Delta_{N}})=\Delta_{N}+O(\Delta_{N}^{\frac{s}{2}+\varepsilon}).

Thus

Nxi0+kNΔN=NN+O(ΔNs2+ε).Nx_{i_{0}+k_{N}-\Delta_{N}}=N\ell_{N}+O(\Delta_{N}^{\frac{s}{2}+\varepsilon}).

Hence

|{i:xi(0,N)}|=|{i:xi(xi0,xi0+kNΔN}|+O({i:|xi|(xi0,0)(xi0+kNΔN,N))=kNΔN+O(kNε+kN(s2+ε)2),|\{i:x_{i}\in(0,\ell_{N})\}|=|\{i:x_{i}\in(x_{i_{0}},x_{i_{0}+k_{N}-\Delta_{N}}\}|\\ +O(\{i:|x_{i}|\in(x_{i_{0}},0)\cup(x_{i_{0}+k_{N}-\Delta_{N}},\ell_{N}))=k_{N}-\Delta_{N}+O(k_{N}^{\varepsilon}+k_{N}^{(\frac{s}{2}+\varepsilon)^{2}}),

thus proving (6.12).

Therefore choosing ε>0\varepsilon>0 small enough, we deduce from (6.10) that

Ns2ζ(s,kNN)12(N(xkN+i0xi0)kN)Law𝒩(0,σ2),N^{-\frac{s}{2}}\zeta(-s,\frac{k_{N}}{N})^{-\frac{1}{2}}(N(x_{k_{N}+i_{0}}-x_{i_{0}})-k_{N})\underset{\mathrm{Law}}{\Longrightarrow}\mathcal{N}(0,\sigma^{2}),

where σ2\sigma^{2} is a in (6.11). By rotational invariance, we deduce that for each i=1,,Ni=1,\ldots,N,

Ns2ζ(s,kNN)12(N(xkN+ixi)kN)Law𝒩(0,σ2).N^{-\frac{s}{2}}\zeta(-s,\frac{k_{N}}{N})^{-\frac{1}{2}}(N(x_{k_{N}+i}-x_{i})-k_{N})\underset{\mathrm{Law}}{\Longrightarrow}\mathcal{N}(0,\sigma^{2}).

Appendix A Well-posedness of the H.-S. equation

A.1. Well-posedness for gradients

Let μ\mu satisfy Assumptions 3.1. The formal adjoint with respect to μ\mu of the derivation i\partial_{i}, i{1,,N}i\in\{1,\ldots,N\} is given by

iw=iw(iH)w,\partial_{i}^{*}w=\partial_{i}w-(\partial_{i}H)w,

meaning that for all v,w𝒞(DN)v,w\in\mathcal{C}^{\infty}(D_{N}) such that vn=0v\cdot\vec{n}=0 on DN\partial D_{N},

𝔼μ[(iv)w]=𝔼μ[viw].\mathbb{E}_{\mu}[(\partial_{i}v)w]=\mathbb{E}_{\mu}[v\partial_{i}^{*}w]. (A.1)

The above can be shown by integration by parts under the Lebesgue measure on DND_{N}. Recall the map

GapN:XNDNN(x2x1,,xNx1)N\mathrm{Gap}_{N}:X_{N}\in D_{N}\mapsto N(x_{2}-x_{1},\ldots,x_{N}-x_{1})\in\mathbb{R}^{N}

and μ:=GapN#μ.\mu^{\prime}:=\mathrm{Gap}_{N}\#\mu.

Lemma A.1.

Let μ\mu satisfy Assumptions 3.1. Let FH1(μ)F\in H^{-1}(\mu). Assume either that FF is in the form F=GGapNF=G\circ\mathrm{Gap}_{N} with GH1(μ)G\in H^{-1}(\mu^{\prime}) or that χ\chi is bounded. Then there exists a unique ϕH1(μ)\phi\in H^{1}(\mu) such that that

{μϕ=F𝔼μ[F]on DNϕn=0a.e on DN\begin{cases}\mathcal{L}^{\mu}\phi=F-\mathbb{E}_{\mu}[F]&\text{on }D_{N}\\ \nabla\phi\cdot\vec{n}=0&\text{a.e on }\partial D_{N}\end{cases} (A.2)

Moreover the solution ϕ\phi of (A.2) is the unique minimizer of

ϕ𝔼μ[|ϕ|22ϕF],\phi\mapsto\mathbb{E}_{\mu}[|\nabla\phi|^{2}-2\phi F],

over functions ϕH1(μ)\phi\in H^{1}(\mu) such that 𝔼μ[ϕ]=0\mathbb{E}_{\mu}[\phi]=0.

Lemma A.1 is a variation on Lax-Milgram’s lemma. When the interaction kernel χ\chi is bounded, a uniform Poincaré inequality holds. If χ\chi is not assumed to be bounded, then the Poincaré inequality holds for all functions of the gaps.

Lemma A.2 (Poincaré inéquality).

Let μ\mu satisfy Assumptions 3.1. There exists a constant C>0C>0 such that for all ϕ=ψGapN\phi=\psi\circ\mathrm{Gap}_{N} with ψH1(μ)\psi\in H^{1}(\mu^{\prime}),

Varμ[ϕ]C𝔼μ[|ϕ|2].\mathrm{Var}_{\mu}[\phi]\leq C\mathbb{E}_{\mu}[|\nabla\phi|^{2}].

Assume in addition that χ\chi (3.1) is bounded. Then, there exists a constant C>0C>0 such that for all ϕH1(μ)\phi\in H^{1}(\mu),

Varμ[ϕ]C𝔼μ[|ϕ|2].\mathrm{Var}_{\mu}[\phi]\leq C\mathbb{E}_{\mu}[|\nabla\phi|^{2}]. (A.3)
Proof.

Let ϕ=ψGapN\phi=\psi\circ\mathrm{Gap}_{N} with ψH1(μ)\psi\in H^{1}(\mu^{\prime}). The measure μ\mu^{\prime} is uniformly log-concave: it can be written

dμ(x)=eH~(x)𝟙x1++xN=NdXN,\mathrm{d}\mu^{\prime}(x)=e^{-\tilde{H}(x)}\mathds{1}_{x_{1}+\ldots+x_{N}=N}\mathrm{d}X_{N},

where dXN\mathrm{d}X_{N} is the Lebesgue measure on (+)N(\mathbb{R}^{+*})^{N} and with H~\tilde{H} satisfying 2H~cId\nabla^{2}\tilde{H}\geq c\mathrm{Id} for some constant c>0c>0. Therefore by the Brascamp-Lieb inequality of Lemma 3.8,

Varμ[ϕ]=Varμ[ψ]c1𝔼μ[|ψ|2]=c1𝔼μ[i=1N(N(i+1ϕiϕ))2]=2c1N2𝔼μ[|ϕ|2].\mathrm{Var}_{\mu}[\phi]=\mathrm{Var}_{\mu^{\prime}}[\psi]\leq c^{-1}\mathbb{E}_{\mu^{\prime}}[|\nabla\psi|^{2}]=c^{-1}\mathbb{E}_{\mu}\Bigr{[}\sum_{i=1}^{N}(N(\partial_{i+1}\phi-\partial_{i}\phi))^{2}\Bigr{]}=2c^{-1}N^{2}\mathbb{E}_{\mu}[|\nabla\phi|^{2}].

If χ\chi is bounded, there exist c1>0,c2>0c_{1}>0,c_{2}>0 such that

c1dμ(x)dxc2.c_{1}\leq\frac{\mathrm{d}\mu(x)}{\mathrm{d}x}\leq c_{2}.

Since the Lebesgue measure on DND_{N} satisfies a Poincaré inequality, so does μ\mu. ∎

Proof of Lemma A.1.

Assume that F=GGapNF=G\circ\mathrm{Gap}_{N} with GH1(μ)G\in H^{-1}(\mu^{\prime}). Let

E={ϕH1(μ):ϕ=ψGapN,ψH1(μ),𝔼μ[ϕ]=0}E=\{\phi\in H^{1}(\mu):\phi=\psi\circ\mathrm{Gap}_{N},\psi\in H^{1}(\mu^{\prime}),\mathbb{E}_{\mu}[\phi]=0\}

and

J:ϕE𝔼μ[|ϕ|2]2𝔼μ[Fϕ].J:\phi\in E\mapsto\mathbb{E}_{\mu}[|\nabla\phi|^{2}]-2\mathbb{E}_{\mu}[F\phi]. (A.4)

Let us prove that JJ admits a unique minimizer. First for all ϕE\phi\in E,

|𝔼μ[Fϕ]|FH1(μ)ϕH1(μ).|\mathbb{E}_{\mu}[F\phi]|\leq\|F\|_{H^{-1}(\mu)}\|\phi\|_{H^{1}(\mu)}.

By Lemma A.2, there exists a constant C>0C>0 such that for all ϕE\phi\in E,

𝔼μ[ϕ2]C𝔼μ[|ϕ|2].\mathbb{E}_{\mu}[\phi^{2}]\leq C\mathbb{E}_{\mu}[|\nabla\phi|^{2}].

Therefore there exists C>0C>0 such that for all ϕE\phi\in E,

𝔼μ[Fϕ]CFH1(μ)𝔼μ[|ϕ|2]12.\mathbb{E}_{\mu}[F\phi]\leq C\|F\|_{H^{-1}(\mu)}\mathbb{E}_{\mu}[|\nabla\phi|^{2}]^{\frac{1}{2}}. (A.5)

It follows that JJ is coercive with respect to the H1(μ)H^{1}(\mu) norm and that JJ is bounded from below. Let (ϕk)(\phi_{k}) be a sequence of elements of EE such that (J(ϕk))(J(\phi_{k})) converges to infJ\inf J. Since (ϕk)(\phi_{k}) is bounded in H1(μ)H^{1}(\mu), there exists a sub-sequence converging weakly to a certain ϕE\phi\in E. It follows from (A.5) that JJ is l.s.c on H1(μ)H^{1}(\mu). Since JJ is convex, JJ is l.s.c for the weak topology on H1(μ)H^{1}(\mu). Therefore ϕ\phi is a minimizer of JJ on EE. The first-order minimality condition for ϕ\phi reads

𝔼μ[ϕh]=𝔼μ[Fh],\mathbb{E}_{\mu}[\nabla\phi\cdot\nabla h]=\mathbb{E}_{\mu}[Fh],

for all hEh\in E. By integration by parts, for all hEh\in E, we have

𝔼μ[(μϕF)h]+DN(ϕn)heH=0.\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F)h]+\int_{\partial D_{N}}(\nabla\phi\cdot\vec{n})he^{-H}=0.

Thus if the χij\chi_{ij}’s are bounded from below, we deduce that ϕn=0\nabla\phi\cdot\vec{n}=0 a.e on DN\partial D_{N}. Moreover for all hEh\in E,

𝔼μ[(μϕF)h]=0.\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F)h]=0.

Since

𝔼μ[(μϕF+𝔼μ[F])]=0,\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F])]=0,

we deduce that for all h=ψGapNh=\psi\circ\mathrm{Gap}_{N} with ψH1(μ)\psi\in H^{1}(\mu^{\prime}),

𝔼μ[(μϕF+𝔼μ[F])h]=0.\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F])h]=0.

Since (μϕF+𝔼μ[F])(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F]) is a function of the gaps, we deduce that for all hH1(μ)h\in H^{1}(\mu),

𝔼μ[(μϕF+𝔼μ[F])h]=𝔼μ[𝔼μ[(μϕF+𝔼μ[F])hx1]]=𝔼μ[(μϕF+𝔼μ[F])𝔼μ[hx1]]=0,\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F])h]=\mathbb{E}_{\mu}[\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F])h\mid x_{1}]]=\mathbb{E}_{\mu}[(\mathcal{L}^{\mu}\phi-F+\mathbb{E}_{\mu}[F])\mathbb{E}_{\mu}[h\mid x_{1}]]=0,

since 𝔼μ[hx1]\mathbb{E}_{\mu}[h\mid x_{1}] is a function of the gaps. We deduce that

μϕ=F𝔼μ[F],\mathcal{L}^{\mu}\phi=F-\mathbb{E}_{\mu}[F],

as elements of H1(μ)H^{-1}(\mu) and that ϕn=0\nabla\phi\cdot\vec{n}=0 a.e on DN\partial D_{N}. The uniqueness is straightforward.

When the density of μ\mu is bounded from below by a positive constant, then by Lemma A.2, μ\mu satisfies a Poincaré inequality. We deduce existence and uniqueness of a solution to (A.2) by the same arguments. ∎

One can now complete the proof of Proposition 3.1.

Proof of Proposition 3.1.

Let F=GGapNF=G\circ\mathrm{Gap}_{N} with GH1(μ)G\in H^{1}(\mu^{\prime}). Recall that if FH1(μ)F\in H^{1}(\mu), FL2({1,,N},H1(μ))\nabla F\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu)). Indeed

i=1NiFH1(μ)2i=1NiFL22FH1(μ)2.\sum_{i=1}^{N}\|\partial_{i}F\|_{H^{-1}(\mu)}^{2}\leq\sum_{i=1}^{N}\|\partial_{i}F\|_{L^{2}}^{2}\leq\|F\|_{H^{1}(\mu)}^{2}.

By Lemma A.1, there exists ϕH1(μ)\phi\in H^{1}(\mu) such that ϕ=F𝔼μ[F]\mathcal{L}\phi=F-\mathbb{E}_{\mu}[F] as elements of H1(μ)H^{-1}(\mu) and such that ϕn=0\nabla\phi\cdot\vec{n}=0 a.e on DN\partial D_{N}. Let w𝒞(DN)w\in\mathcal{C}^{\infty}(D_{N}) such that wn=0\nabla w\cdot\vec{n}=0 a.e on DN\partial D_{N}. For each i{1,,N}i\in\{1,\ldots,N\}, we have

𝔼μ[wiF]=𝔼μ[iw(F𝔼μ[F])]=𝔼μ[iwμϕ]=𝔼μ[iwϕ]=j=1N𝔼μ[(ijw)jϕ]+j=1N𝔼μ[([j,i]w)jϕ].\mathbb{E}_{\mu}[w\partial_{i}F]=\mathbb{E}_{\mu}[\partial_{i}^{*}w(F-\mathbb{E}_{\mu}[F])]=\mathbb{E}_{\mu}[\partial_{i}^{*}w\mathcal{L}^{\mu}\phi]=\mathbb{E}_{\mu}[\nabla\partial_{i}^{*}w\cdot\nabla\phi]\\ =\sum_{j=1}^{N}\mathbb{E}_{\mu}[(\partial_{i}^{*}\partial_{j}w)\partial_{j}\phi]+\sum_{j=1}^{N}\mathbb{E}_{\mu}[([\partial_{j},\partial_{i}^{*}]w)\partial_{j}\phi].

For the first term in the sum above, we have

j=1N𝔼μ[(ijw)jϕ]=j=1N𝔼μ[(jw)ijϕ]=𝔼μ[w(iϕ)]=𝔼μ[wμ(iϕ)].\sum_{j=1}^{N}\mathbb{E}_{\mu}[(\partial_{i}^{*}\partial_{j}w)\partial_{j}\phi]=\sum_{j=1}^{N}\mathbb{E}_{\mu}[(\partial_{j}w)\partial_{i}\partial_{j}\phi]=\mathbb{E}_{\mu}[\nabla w\cdot\nabla(\partial_{i}\phi)]=\mathbb{E}_{\mu}[w\mathcal{L}^{\mu}(\partial_{i}\phi)].

For the second term, using the identity [j,i]=(2H)i,j,[\partial_{j},\partial_{i}^{*}]=(\nabla^{2}H)_{i,j}, one can write

j=1N𝔼μ[([j,i]w)jϕ]=𝔼μ[wei2Hϕ].\sum_{j=1}^{N}\mathbb{E}_{\mu}[([\partial_{j},\partial_{i}^{*}]w)\partial_{j}\phi]=\mathbb{E}_{\mu}[we_{i}\cdot\nabla^{2}H\nabla\phi].

We conclude by density that, in the sense of H1(μ)H^{-1}(\mu), for each i{1,,N}i\in\{1,\ldots,N\},

μ(iϕ)+(2Hϕ)i=iF.\mathcal{L}^{\mu}(\partial_{i}\phi)+(\nabla^{2}H\nabla\phi)_{i}=\partial_{i}F.

This concludes the proof of existence of a solution to (3.7). Uniqueness is straightforward.

The formula (3.9) then easily follows from an integration by parts: letting ϕ\nabla\phi be the solution of (3.7), we can write

Varμ[F]=𝔼μ[(F𝔼μ[F])μϕ]=𝔼μ[Fϕ].\mathrm{Var}_{\mu}[F]=\mathbb{E}_{\mu}[(F-\mathbb{E}_{\mu}[F])\mathcal{L}^{\mu}\phi]=\mathbb{E}_{\mu}[\nabla F\cdot\nabla\phi].

The variational representation (3.8) is straightforward using that ϕ\nabla\phi is the unique minimizer of the function JJ (A.4).

In the case where FF is not a function of the gaps and where μ\mu is assumed to be bounded from below, we conclude likewise. ∎

Proof of Proposition 3.2.

Let vL2({1,,N},H1(μ))v\in L^{2}(\{1,\ldots,N\},H^{-1}(\mu)). Let

E={ψL2({1,,N},H1(μ)):ψ=vGapN,vL2({1,,N},H1(μ)),ψ(e1++eN)=0}E=\{\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)):\psi=v\circ\mathrm{Gap}_{N},v\in L^{2}(\{1,\ldots,N\},H^{1}(\mu^{\prime})),\\ \psi\cdot(e_{1}+\ldots+e_{N})=0\} (A.6)

and

J:ψE𝔼μ[|Dψ|2+ψ2Hψ2vψ].J:\psi\in E\mapsto\mathbb{E}_{\mu}[|D\psi|^{2}+\psi\cdot\nabla^{2}H\psi-2v\cdot\psi].

Let us show that JJ admits a unique minimizer. First

|𝔼μ[vψ]|i=1N𝔼μ[vi2]12𝔼μ[ψi2]12.|\mathbb{E}_{\mu}[v\cdot\psi]|\leq\sum_{i=1}^{N}\mathbb{E}_{\mu}[v_{i}^{2}]^{\frac{1}{2}}\mathbb{E}_{\mu}[\psi_{i}^{2}]^{\frac{1}{2}}.

Now we use the fact that there exists c>0c>0 such that

ψ2ψci=1N(ψi+1ψi)2=2ci=1Nψi2,\psi\cdot\nabla^{2}\psi\geq c\sum_{i=1}^{N}(\psi_{i+1}-\psi_{i})^{2}=2c\sum_{i=1}^{N}\psi_{i}^{2},

since ψ(e1++eN)=0\psi\cdot(e_{1}+\ldots+e_{N})=0. It follows that JJ is coercive with respect to
L2({1,,N},H1(μ))L^{2}(\{1,\ldots,N\},H^{1}(\mu)). Arguing as in the proof of Lemma A.1, one can show that JJ admits a unique minimizer ψE\psi\in E which satisfies

𝔼μ[i=1Nψihi+ψ2Hh]=𝔼μ[vh],\mathbb{E}_{\mu}\Bigr{[}\sum_{i=1}^{N}\nabla\psi_{i}\cdot\nabla h_{i}+\psi\cdot\nabla^{2}Hh\Bigr{]}=\mathbb{E}_{\mu}[v\cdot h],

for all hEh\in E. By integration by parts for all hEh\in E and each i=1,,Ni=1,\ldots,N

𝔼μ[ψihi]=𝔼μ[μψihi]+DN(ψin)hieH=𝔼μ[μψihi],\mathbb{E}_{\mu}[\nabla\psi_{i}\cdot\nabla h_{i}]=\mathbb{E}_{\mu}[\mathcal{L}^{\mu}\psi_{i}h_{i}]+\int_{\partial D_{N}}(\nabla\psi_{i}\cdot\vec{n})h_{i}e^{-H}=\mathbb{E}_{\mu}[\mathcal{L}^{\mu}\psi_{i}h_{i}],

since for each i,ji,j, limx0χij(x)=+\lim_{x\to 0}\chi_{ij}(x)=+\infty. Therefore for all hEh\in E

𝔼μ[(A1μψv)h]=0.\mathbb{E}_{\mu}[(A_{1}^{\mu}\psi-v)h]=0. (A.7)

Since (A1μψv)(e1++eN)=0(A_{1}^{\mu}\psi-v)\cdot(e_{1}+\ldots+e_{N})=0, (A.7) holds for all hL2({1,,N},H1(μ))h\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)) such that for each ii, hih_{i} is a function of the gaps. Finally, since A1μψvA_{1}^{\mu}\psi-v is a function of the gaps, (A.7) holds for any hL2({1,,N},H1(μ))h\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)). We conclude that ψ\psi satisfies the Euler-Lagrange equation

{A1μψ=von DNψ(e1++eN)=0on DN.\begin{cases}A_{1}^{\mu}\psi=v&\text{on }D_{N}\\ \psi\cdot(e_{1}+\ldots+e_{N})=0&\text{on $D_{N}$}.\end{cases} (A.8)

where the first equality is an equality between elements of L2({1,,N},H1(μ))L^{2}(\{1,\ldots,N\},H^{-1}(\mu)).

Let us show uniqueness of the solution. Let ψL2({1,,N},H1(μ))\psi\in L^{2}(\{1,\ldots,N\},H^{1}(\mu)) satisfy

{A1μψ=0on DNψ(e1++eN)=0on DN.\begin{cases}A_{1}^{\mu}\psi=0&\text{on }D_{N}\\ \psi\cdot(e_{1}+\ldots+e_{N})=0&\text{on $D_{N}$}.\end{cases} (A.9)

Taking the scalar product of the first equation with ψ\psi, integrating by parts gives

𝔼μ[i=1N|ψi|2]+𝔼μ[ψ2Hψ]=0.\mathbb{E}_{\mu}\Bigr{[}\sum_{i=1}^{N}|\nabla\psi_{i}|^{2}\Bigr{]}+\mathbb{E}_{\mu}[\psi\cdot\nabla^{2}H\psi]=0.

Since i=1Nψi=0\sum_{i=1}^{N}\psi_{i}=0, we have ψ=0\psi=0. This proves uniqueness. ∎

Appendix B Auxiliary estimates

B.1. Discrete convolution products

Lemma B.1.

Let α\alpha, β\beta be such that α+β>1\alpha+\beta>1. Let k0k_{0}\in\mathbb{N}.

  1. (i)

    If α(0,1)\alpha\in(0,1) and β(0,1)\beta\in(0,1),

    k,kk01kα1|k0k|βCk0α+β1.\sum_{k\in\mathbb{N},k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k_{0}-k|^{\beta}}\leq\frac{C}{k_{0}^{\alpha+\beta-1}}.
  2. (ii)

    If α=1\alpha=1 and β(0,1)\beta\in(0,1),

    k,kk01kα1|k0k|βClogk0k0β,\sum_{k\in\mathbb{N},k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k_{0}-k|^{\beta}}\leq\frac{C\log k_{0}}{k_{0}^{\beta}},
  3. (iii)

    If α>1\alpha>1 and β>0\beta>0,

    k,kk01kα1|k0k|βCk0min(α,β).\sum_{k\in\mathbb{N},k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k_{0}-k|^{\beta}}\leq\frac{C}{k_{0}^{\min(\alpha,\beta)}}.

The above estimates follow from straightforward computations, see for instance [MO+16, DW20]. Let us now adapt Lemma B.1 to truncated convolution products.

Lemma B.2.

Let α,β\alpha,\beta be such that α+β>1\alpha+\beta>1. Let k0k_{0}\in\mathbb{N}.

  1. (i)

    If α>1\alpha>1 and β(0,1)\beta\in(0,1),

    kK,kk01kα1|kk0|β{1Kα+β1 if k0K21k0β1Kα1 if k0K2.\sum_{k\geq K,k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\left\{\begin{array}[]{ll}\frac{1}{K^{\alpha+\beta-1}}&\text{ if }k_{0}\leq\frac{K}{2}\\ \frac{1}{k_{0}^{\beta}}\frac{1}{K^{\alpha-1}}&\text{ if }k_{0}\geq\frac{K}{2}.\end{array}\right. (B.1)
  2. (ii)

    If α>1\alpha>1 and β=1\beta=1,

    kK,kk01kα1|k0k|β{1Kα+β1 if k0K21k0α+β1+1k0βlogk0 if k0K2.\sum_{k\geq K,k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k_{0}-k|^{\beta}}\lesssim\left\{\begin{array}[]{ll}\frac{1}{K^{\alpha+\beta-1}}&\text{ if }k_{0}\leq\frac{K}{2}\\ \frac{1}{k_{0}^{\alpha+\beta-1}}+\frac{1}{k_{0}^{\beta}}\log k_{0}&\text{ if }k_{0}\geq\frac{K}{2}.\end{array}\right.
  3. (iii)

    If α>1\alpha>1 and β>1\beta>1,

    kK,kk01kα1|kk0|β{1Kα+β1 if k0K21k0α1(Kk0)β1 if K2k0<K1k0α+1k0β1Kα1 if k0K.\sum_{k\geq K,k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\left\{\begin{array}[]{ll}\frac{1}{K^{\alpha+\beta-1}}&\text{ if }k_{0}\leq\frac{K}{2}\\ \frac{1}{k_{0}^{\alpha}}\frac{1}{(K-k_{0})^{\beta-1}}&\text{ if }\frac{K}{2}\leq k_{0}<K\\ \frac{1}{k_{0}^{\alpha}}+\frac{1}{k_{0}^{\beta}}\frac{1}{K^{\alpha-1}}&\text{ if }k_{0}\geq K.\end{array}\right.
Proof of Lemma B.2.

Let us prove the three statements together. Let α>1\alpha>1. We split the sum over kk along the condition k2k0k\geq 2k_{0}. If k2k0k\geq 2k_{0}, |kk0|k2|k-k_{0}|\geq\frac{k}{2} and therefore

k2k0,kK1kα1|kk0|βk2k0,kK1kα+β1max(K,k0)α+β1.\sum_{k\geq 2k_{0},k\geq K}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\sum_{k\geq 2k_{0},k\geq K}\frac{1}{k^{\alpha+\beta}}\lesssim\frac{1}{\max(K,k_{0})^{\alpha+\beta-1}}.

If k0K2k_{0}\leq\frac{K}{2}, the remaining part is empty and therefore

kK,kk01kα1|kk0|β1Kα+β1.\sum_{k\geq K,k\neq k_{0}}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\frac{1}{K^{\alpha+\beta-1}}.

Assume now that k0K2k_{0}\geq\frac{K}{2}. We split the remaining term into two parts according to whether |kk0|k02|k-k_{0}|\geq\frac{k_{0}}{2} or not. For the first contribution one has

Kk2k0,|kk0|k021kα1|kk0|β1k0βkK1kα1k0β1Kα1.\sum_{K\leq k\leq 2k_{0},|k-k_{0}|\geq\frac{k_{0}}{2}}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\frac{1}{k_{0}^{\beta}}\sum_{k\geq K}\frac{1}{k^{\alpha}}\lesssim\frac{1}{k_{0}^{\beta}}\frac{1}{K^{\alpha-1}}.

For the second contribution we may write

Kk2k0,|kk0|k02kk01kα1|kk0|β1k0α|kk0|k02,kKkk01|kk0|β.\sum_{\begin{subarray}{c}K\leq k\leq 2k_{0},|k-k_{0}|\leq\frac{k_{0}}{2}\\ k\neq k_{0}\end{subarray}}\frac{1}{k^{\alpha}}\frac{1}{|k-k_{0}|^{\beta}}\lesssim\frac{1}{k_{0}^{\alpha}}\sum_{\begin{subarray}{c}|k-k_{0}|\leq\frac{k_{0}}{2},k\geq K\\ k\neq k_{0}\end{subarray}}\frac{1}{|k-k_{0}|^{\beta}}.

One can bound the sum in the right-hand side by

kK,|kk0|k02kk01|kk0|β{k01β if β(0,1)logk0 if β=11 if β>1 and k0K1(Kk0)β1 if β>1 and K2k0<K.\sum_{\begin{subarray}{c}k\geq K,|k-k_{0}|\leq\frac{k_{0}}{2}\\ k\neq k_{0}\end{subarray}}\frac{1}{|k-k_{0}|^{\beta}}\leq\left\{\begin{array}[]{ll}k_{0}^{1-\beta}&\text{ if }\beta\in(0,1)\\ \log k_{0}&\text{ if }\beta=1\\ 1&\text{ if }\beta>1\text{ and }k_{0}\geq K\\ \frac{1}{(K-k_{0})^{\beta-1}}&\text{ if }\beta>1\text{ and }\frac{K}{2}\leq k_{0}<K.\end{array}\right.

Combining the above estimates concludes the proof of Lemma B.2. ∎

B.2. Proof of Lemma 5.2

Proof of Lemma 5.2.

We assume that N0\ell_{N}\to 0. The case N=1\ell_{N}=1 is similar and easier. Let i0=argmin1iN|xi|i_{0}=\underset{1\leq i\leq N}{\mathrm{argmin}}|x_{i}|, γ>1\gamma>1 and

IN:={i{1,,N}:d(i,i0)(NN)γ}.I_{N}:=\{i\in\{1,\ldots,N\}:d(i,i_{0})\leq(N\ell_{N})^{\gamma}\}.

Let ψ𝒞2(N1𝕋,)\psi\in\mathcal{C}^{2}(\ell_{N}^{-1}\mathbb{T},\mathbb{R}) satisfy (5.17). Define

ηN:xN1𝕋(1+l=1p1(|xal|1N)2s+α)𝟙|x|<2+1(1+|x|)3s,\eta_{N}:x\in\ell_{N}^{-1}\mathbb{T}\mapsto\Bigr{(}1+\sum_{l=1}^{p}\frac{1}{(|x-a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\Bigr{)}\mathds{1}_{|x|<2}+\frac{1}{(1+|x|)^{3-s}}, (B.2)

so that

|ψ|CηN.|\psi^{\prime\prime}|\leq C\eta_{N}.

Also define

γN:xN1𝕋(1+l=1p1|xal|1s+α)𝟙|x|<2+11+|x|2s,\gamma_{N}:x\in\ell_{N}^{-1}\mathbb{T}\mapsto\Bigr{(}1+\sum_{l=1}^{p}\frac{1}{|x-a_{l}|^{1-s+\alpha}}\Bigr{)}\mathds{1}_{|x|<2}+\frac{1}{1+|x|^{2-s}},

so that

|ψ|CγN.|\psi^{\prime}|\leq C\gamma_{N}.

Denote uN\mathrm{u}^{N} the regular grid of spacing 1N\frac{1}{N} on 𝕋\mathbb{T}:

uiN=iN𝕋,i=1,,N.\mathrm{u}_{i}^{N}=\frac{i}{N}\in\mathbb{T},\quad i=1,\ldots,N.

Recall

𝒜={XNDN:iIN,k{N/2,,N/2}:i+kIN,N|xi+kxiukN|(NN)εks2,iIN,(NN)εN|xi+1xi|(NN)ε}.\mathcal{A}=\Bigr{\{}X_{N}\in D_{N}:\forall i\in I_{N},k\in\{-N/2,\ldots,N/2\}:i+k\in I_{N},\\ N|x_{i+k}-x_{i}-\mathrm{u}_{k}^{N}|\leq(N\ell_{N})^{\varepsilon}k^{\frac{s}{2}},\forall i\in I_{N},(N\ell_{N})^{-\varepsilon}\leq N|x_{i+1}-x_{i}|\leq(N\ell_{N})^{\varepsilon}\}. (B.3)

Below we restrict to the event (B.3).

Step 1: splitting

Let

g~:x𝕋{0}g(x)x\tilde{g}:x\in\mathbb{T}\setminus\{0\}\mapsto g^{\prime}(x)x

and

ζ:(x,y)(N1𝕋)2ΔN(ψ(N1y)ψ(N1x))yx,\zeta:(x,y)\in(\ell_{N}^{-1}\mathbb{T})^{2}\setminus\Delta\mapsto\frac{\ell_{N}(\psi(\ell_{N}^{-1}y)-\psi(\ell_{N}^{-1}x))}{y-x},

so that

NN(ψ(N1xi)ψ(N1xj))N(s+1)g(xixj)=ζ(xi,xj)Nsg~(xixj).N\ell_{N}(\psi(\ell_{N}^{-1}x_{i})-\psi(\ell_{N}^{-1}x_{j}))N^{-(s+1)}g^{\prime}(x_{i}-x_{j})=\zeta(x_{i},x_{j})N^{-s}\tilde{g}(x_{i}-x_{j}).

One can express A~\tilde{\mathrm{A}} as

A~=ijINζ(xi,xj)Nsg~(xixj)2NiIN|y|(NN)γNζ(xi,y)Nsg~(xiy)dy.\tilde{\mathrm{A}}=\sum_{i\neq j\in I_{N}}\zeta(x_{i},x_{j})N^{-s}\tilde{g}(x_{i}-x_{j})-2N\sum_{i\in I_{N}}\int_{|y|\leq\frac{(N\ell_{N})^{\gamma}}{N}}\zeta(x_{i},y)N^{-s}\tilde{g}(x_{i}-y)\mathrm{d}y.

Therefore A~=V+W\nabla\tilde{\mathrm{A}}=\mathrm{V}+\mathrm{W} with Vi=Wi=0\mathrm{V}_{i}=\mathrm{W}_{i}=0 for each iINci\in I_{N}^{c} and for each iINi\in I_{N},

Vi=2k:i+kIN1ζ(xi,xi+k)Nsg~(xi+kxi)2N|y|(NN)γN1ζ(xi,y)Nsg~(yxi)dy+2k:i+kINζ(xi,xi+k)Nsg~(ukN)2N|y|(NN)γNζ(xi,y)Nsg~(yxi)dy\mathrm{V}_{i}=2\sum_{k:i+k\in I_{N}}\partial_{1}\zeta(x_{i},x_{i+k})N^{-s}\tilde{g}(x_{i+k}-x_{i})-2N\int_{|y|\leq\frac{(N\ell_{N})^{\gamma}}{N}}\partial_{1}\zeta(x_{i},y)N^{-s}\tilde{g}(y-x_{i})\mathrm{d}y\\ +2\sum_{k:i+k\in I_{N}}\zeta(x_{i},x_{i+k})N^{-s}\tilde{g}^{\prime}(\mathrm{u}_{k}^{N})-2N\int_{|y|\leq\frac{(N\ell_{N})^{\gamma}}{N}}\zeta(x_{i},y)N^{-s}\tilde{g}^{\prime}(y-x_{i})\mathrm{d}y

and

Wi=2k:i+kINζ(xi,xi+k)Ns(g~(xi+kxi)g~(ukN)).\mathrm{W}_{i}=2\sum_{k:i+k\in I_{N}}\zeta(x_{i},x_{i+k})N^{-s}(\tilde{g}^{\prime}(x_{i+k}-x_{i})-\tilde{g}^{\prime}(\mathrm{u}_{k}^{N})).

Let us then isolate from V\mathrm{V} the fluctuations from the discretization error. For each iINi\in I_{N}, write Vi=Vi1+Vi2\mathrm{V}_{i}=\mathrm{V}_{i}^{1}+\mathrm{V}_{i}^{2} with

Vi1=2k:i+kIN1ζ(xi,xi+k)Nsg~(xi+kxi)2k:i+kIN1ζ(xi,xi+ukN)Nsg~(ukN)+2k:i+kIN(ζ(xi,xi+k)ζ(xi,xi+ukN))Nsg~(ukN)\mathrm{V}_{i}^{1}=2\sum_{k:i+k\in I_{N}}\partial_{1}\zeta(x_{i},x_{i+k})N^{-s}\tilde{g}(x_{i+k}-x_{i})-2\sum_{k:i+k\in I_{N}}\partial_{1}\zeta(x_{i},x_{i}+\mathrm{u}_{k}^{N})N^{-s}\tilde{g}(\mathrm{u}_{k}^{N})\\ +2\sum_{k:i+k\in I_{N}}(\zeta(x_{i},x_{i+k})-\zeta(x_{i},x_{i}+\mathrm{u}_{k}^{N}))N^{-s}\tilde{g}^{\prime}(\mathrm{u}_{k}^{N})

and

Vi2=2k:i+kIN(1ζ(xi,ukN)g~(ukN)+ζ(xi,ukN)g~(ukN))2N|y|(NN)γN(1ζ(xi,y)g~(y)+ζ(xi,y)g~(y))dy.\mathrm{V}_{i}^{2}=2\sum_{k:i+k\in I_{N}}(\partial_{1}\zeta(x_{i},\mathrm{u}_{k}^{N})\tilde{g}(\mathrm{u}_{k}^{N})+\zeta(x_{i},\mathrm{u}_{k}^{N})\tilde{g}^{\prime}(\mathrm{u}_{k}^{N}))\\ -2N\int_{|y|\leq\frac{(N\ell_{N})^{\gamma}}{N}}(\partial_{1}\zeta(x_{i},y)\tilde{g}(y)+\zeta(x_{i},y)\tilde{g}^{\prime}(y))\mathrm{d}y.
Step 2: control on V1\mathrm{V}^{1}

For each iINi\in I_{N}, one may write

Vi1=2k:i+kIN1ζ(xi,xi+k)Ns(g~(xi+kxi)g~(ukN))+2k:i+kIN(1ζ(xi,xi+k)1ζ(xi,xi+ukN))Nsg~(ukN)+2k:i+kIN(ζ(xi,xi+k)ζ(xi,xi+ukN))Nsg~(ukN).\mathrm{V}_{i}^{1}=2\sum_{k:i+k\in I_{N}}\partial_{1}\zeta(x_{i},x_{i+k})N^{-s}(\tilde{g}(x_{i+k}-x_{i})-\tilde{g}(\mathrm{u}_{k}^{N}))\\ +2\sum_{k:i+k\in I_{N}}(\partial_{1}\zeta(x_{i},x_{i+k})-\partial_{1}\zeta(x_{i},x_{i}+\mathrm{u}_{k}^{N}))N^{-s}\tilde{g}(\mathrm{u}_{k}^{N})\\ +2\sum_{k:i+k\in I_{N}}(\zeta(x_{i},x_{i+k})-\zeta(x_{i},x_{i}+\mathrm{u}_{k}^{N}))N^{-s}\tilde{g}^{\prime}(\mathrm{u}_{k}^{N}). (B.4)

Denote Vi1,1\mathrm{V}_{i}^{1,1}, Vi1,2\mathrm{V}_{i}^{1,2} and Vi1,3\mathrm{V}_{i}^{1,3} the three terms in (B.4). There exists C>0C>0 such that for all x,y𝕋x,y\in\mathbb{T},

|1ζ(x,y)|CN1(ηN(N1x)+ηN(N1y)).|\partial_{1}\zeta(x,y)|\leq C\ell_{N}^{-1}(\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}y)). (B.5)

Therefore, on the good event 𝒜\mathcal{A} defined in (B.3), one can bound V1,1\mathrm{V}^{1,1} uniformly in iINi\in I_{N} by

|Vi1,1|CN1(NN)κεjIN:ji(ηN(N1xj)+ηN(N1(xi+ujiN)))1d(i,j)1+s2+1.|\mathrm{V}_{i}^{1,1}|\leq C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}\sum_{j\in I_{N}:j\neq i}(\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\frac{1}{d(i,j)^{1+\frac{s}{2}}+1}. (B.6)

We turn to the second term of (B.4). There exists C>0C>0 such that for all a,x,y𝕋a,x,y\in\mathbb{T},

|1ζ(x,y+a)1ζ(x,y)|CN1(ηN(N1y)+ηN(N1(y+a))+ηN(N1x))|a||xy|.|\partial_{1}\zeta(x,y+a)-\partial_{1}\zeta(x,y)|\leq C\ell_{N}^{-1}(\eta_{N}(\ell_{N}^{-1}y)+\eta_{N}(\ell_{N}^{-1}(y+a))+\eta_{N}(\ell_{N}^{-1}x))\frac{|a|}{|x-y|}.

Applying this to x=xix=x_{i}, y=xjy=x_{j}, a=xjxiujiNa=x_{j}-x_{i}-\mathrm{u}_{j-i}^{N}, we find that there exist constants C>0C>0 and κ>0\kappa>0 such that uniformly on 𝒜\mathcal{A} and for each i,jINi,j\in I_{N},

|1ζ(xi,xj)1ζ(xi,xi+ujiN)|CN1(NN)κε(ηN(N1xi)+ηN(N1(xi+ujiN))+ηN(N1xj))1d(i,j)1+s2+1.|\partial_{1}{\zeta}(x_{i},x_{j})-\partial_{1}{\zeta}(x_{i},x_{i}+\mathrm{u}_{j-i}^{N})|\\ \leq C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}(\eta_{N}(\ell_{N}^{-1}x_{i})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N}))+\eta_{N}(\ell_{N}^{-1}x_{j}))\frac{1}{d(i,j)^{1+\frac{s}{2}}+1}.

Thus, uniformly on 𝒜\mathcal{A},

|Vi1,2|CN1(NN)κεjIN:ji(ηN(N1xj)+ηN(N1(xi+ujiN)))1d(j,i)1+s2+1.|\mathrm{V}_{i}^{1,2}|\leq C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}\sum_{j\in I_{N}:j\neq i}(\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\frac{1}{d(j,i)^{1+\frac{s}{2}}+1}. (B.7)

Similarly,

|Vi1,3|CN1(NN)κεjIN:ji(ηN(N1xj)+ηN(N1(xi+ujiN)))1d(i,j)1+s2+1.|\mathrm{V}_{i}^{1,3}|\leq C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}\sum_{j\in I_{N}:j\neq i}(\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\frac{1}{d(i,j)^{1+\frac{s}{2}}+1}. (B.8)

Let us bound the right-hand side of the last display. For each l=1,,pl=1,\ldots,p let

kl:=argmin1iN|xiNal|.k^{l}:=\underset{1\leq i\leq N}{\mathrm{argmin}}|x_{i}-\ell_{N}a_{l}|. (B.9)

There exists C0>1C_{0}>1 such on the event 𝒜\mathcal{A}, for each jj such that d(i0,j)C0NNd(i_{0},j)\geq C_{0}N\ell_{N},

|xj|>2N.|x_{j}|>2\ell_{N}.

Moreover we can write

N1(N1|xal|(NN)1)2s+α=N1(NN)2s+α((N|xal|)1)2s+α.\frac{\ell_{N}^{-1}}{(\ell_{N}^{-1}|x-a_{l}|\vee(N\ell_{N})^{-1})^{2-s+\alpha}}=\frac{\ell_{N}^{-1}(N\ell_{N})^{2-s+\alpha}}{((N|x-a_{l}|)\vee 1)^{2-s+\alpha}}.

It follows that

ηN(N1xj)+ηN(N1(xi+ujiN))C(NN)κε𝟙d(j,i0)C0NN×(1+l=1p(NN)2s+α1(1+d(j,kl))2s+α)+C(NN)κε(NN)3s(d(j,i0)+NN)3s\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N}))\leq C(N\ell_{N})^{\kappa\varepsilon}\mathds{1}_{d(j,i_{0})\leq C_{0}N\ell_{N}}\\ \times\Bigr{(}1+\sum_{l=1}^{p}(N\ell_{N})^{2-s+\alpha}\frac{1}{(1+d(j,k^{l}))^{2-s+\alpha}}\Bigr{)}+C(N\ell_{N})^{\kappa\varepsilon}\frac{(N\ell_{N})^{3-s}}{(d(j,i_{0})+N\ell_{N})^{3-s}}

and

jIN:ji(ηN(N1xj)+ηN(N1(xi+ujiN)))1d(j,i)1+s2+1CN1(NN)κε×d(j,i0)C0NN(1+l=1p(NN)2s+α11+d(j,kl)2s+α)11+d(j,i)1+s2+CN1(NN)κε(NN)3sj=1N1(NN+d(j,i0))3s11+d(j,i)1+s2.\sum_{j\in I_{N}:j\neq i}(\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\frac{1}{d(j,i)^{1+\frac{s}{2}}+1}\leq C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}\\ \times\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\Bigr{(}1+\sum_{l=1}^{p}(N\ell_{N})^{2-s+\alpha}\frac{1}{1+d(j,k^{l})^{2-s+\alpha}}\Bigr{)}\frac{1}{1+d(j,i)^{1+\frac{s}{2}}}\\ +C\ell_{N}^{-1}(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N})^{3-s}\sum_{j=1}^{N}\frac{1}{(N\ell_{N}+d(j,i_{0}))^{3-s}}\frac{1}{1+d(j,i)^{1+\frac{s}{2}}}. (B.10)

One can first check that

d(j,i0)C0NN1d(j,i)1+s2{1if d(i,i0)2C0NNNNd(i,i0)1+s2if d(i,i0)2C0NN.\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\leq\begin{cases}1&\text{if $d(i,i_{0})\leq 2C_{0}N\ell_{N}$}\\ \frac{N\ell_{N}}{d(i,i_{0})^{1+\frac{s}{2}}}&\text{if $d(i,i_{0})\geq 2C_{0}N\ell_{N}$}.\end{cases}

Then, since α>s1\alpha>s-1 and s2>0\frac{s}{2}>0 we may apply Lemma B.2 to get

d(j,i0)C0NN11+d(j,kl)2s+α11+d(j,i)1+s2C{1d(i,kl)2s+α+1+1d(i,kl)1+s2+1if d(i,i0)2C0NN and α>3s211d(i,kl)2s+α+1if d(i,i0)2C0NN and α3s211d(i,i0)2s2+α+1if d(i,i0)2C0NN.\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\frac{1}{1+d(j,k^{l})^{2-s+\alpha}}\frac{1}{1+d(j,i)^{1+\frac{s}{2}}}\\ \leq C\begin{cases}\frac{1}{d(i,k^{l})^{2-s+\alpha}+1}+\frac{1}{d(i,k^{l})^{1+\frac{s}{2}}+1}&\text{if $d(i,i_{0})\leq 2C_{0}N\ell_{N}$ and $\alpha>\frac{3s}{2}-1$}\\ \frac{1}{d(i,k^{l})^{2-s+\alpha}+1}&\text{if $d(i,i_{0})\leq 2C_{0}N\ell_{N}$ and $\alpha\leq\frac{3s}{2}-1$}\\ \frac{1}{d(i,i_{0})^{2-\frac{s}{2}+\alpha}+1}&\text{if $d(i,i_{0})\geq 2C_{0}N\ell_{N}$}.\end{cases} (B.11)

There exist C>0C>0 and κ>0\kappa>0 such that on the event 𝒜\mathcal{A},

d(i,kl)C(NN)κε|NxiNNal|,for each iIN1lp.d(i,k^{l})\geq C(N\ell_{N})^{-\kappa\varepsilon}|Nx_{i}-N\ell_{N}a_{l}|,\quad\text{for each $i\in I_{N}$, $1\leq l\leq p$}.

Besides we can check that

j=1N1(NN+d(j,i0))3s1d(j,i)1+s2C{1(NN)3s2if d(i,i0)2NN1d(i,i0)3s2if d(i,i0)2NN.\sum_{j=1}^{N}\frac{1}{(N\ell_{N}+d(j,i_{0}))^{3-s}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\leq C\begin{cases}\frac{1}{(N\ell_{N})^{3-\frac{s}{2}}}&\text{if $d(i,i_{0})\leq 2N\ell_{N}$}\\ \frac{1}{d(i,i_{0})^{3-\frac{s}{2}}}&\text{if $d(i,i_{0})\geq 2N\ell_{N}$}.\end{cases}

Combining these we obtain

jIN:ji(ηN(N1xj)+ηN(N1(xi+ujiN)))1d(j,i)1+s2C(NN)κε𝟙α>3s21l=1p(NN)1s+αNs21(|xiNal|1N)1+s2𝟙|xi|<2N+C(NN)κεl=1pN1s+α1(|xiNal|1N)1s+α𝟙|xi|<2N+C(NN)κεNs21(|xi|+N)1+s2.\sum_{j\in I_{N}:j\neq i}(\eta_{N}(\ell_{N}^{-1}x_{j})+\eta_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\frac{1}{d(j,i)^{1+\frac{s}{2}}}\\ \leq C(N\ell_{N})^{\kappa\varepsilon}\mathds{1}_{\alpha>\frac{3s}{2}-1}\sum_{l=1}^{p}(N\ell_{N})^{1-s+\alpha}N^{-\frac{s}{2}}\frac{1}{(|x_{i}-\ell_{N}a_{l}|\vee\frac{1}{N})^{1+\frac{s}{2}}}\mathds{1}_{|x_{i}|<2\ell_{N}}\\ +C(N\ell_{N})^{\kappa\varepsilon}\sum_{l=1}^{p}\ell_{N}^{1-s+\alpha}\frac{1}{(|x_{i}-\ell_{N}a_{l}|\vee\frac{1}{N})^{1-s+\alpha}}\mathds{1}_{|x_{i}|<2\ell_{N}}+C(N\ell_{N})^{\kappa\varepsilon}N^{-\frac{s}{2}}\frac{1}{(|x_{i}|+\ell_{N})^{1+\frac{s}{2}}}. (B.12)

Using this and (B.6), (B.7), (B.8), we conclude that

|Vi1|C(NN)κε𝟙α>3s21l=1p(NN)1s+αNs21(|xiNal|1N)1+s2𝟙|xi|<2N+C(NN)κεl=1pN1s+α1(|xiNNal|1N)2s+α𝟙|xi|<2N+C(NN)κεNs21(|xi|+N)1+s2.|\mathrm{V}_{i}^{1}|\leq C(N\ell_{N})^{\kappa\varepsilon}\mathds{1}_{\alpha>\frac{3s}{2}-1}\sum_{l=1}^{p}(N\ell_{N})^{1-s+\alpha}N^{-\frac{s}{2}}\frac{1}{(|x_{i}-\ell_{N}a_{l}|\vee\frac{1}{N})^{1+\frac{s}{2}}}\mathds{1}_{|x_{i}|<2\ell_{N}}\\ +C(N\ell_{N})^{\kappa\varepsilon}\sum_{l=1}^{p}\ell_{N}^{1-s+\alpha}\frac{1}{(|x_{i}-N\ell_{N}a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\mathds{1}_{|x_{i}|<2\ell_{N}}+C(N\ell_{N})^{\kappa\varepsilon}N^{-\frac{s}{2}}\frac{1}{(|x_{i}|+\ell_{N})^{1+\frac{s}{2}}}. (B.13)
Step 3: control on V2\mathrm{V}^{2}

Fix x𝕋x\in\mathbb{T} and define

ϕ:yN𝕋ζ(x,x+yN)Nsg~(yN)+1ζ(x,x+yN)Nsg~(yN).\phi:y\in N\mathbb{T}\mapsto\zeta(x,x+\frac{y}{N})N^{-s}\tilde{g}^{\prime}(\frac{y}{N})+\partial_{1}\zeta(x,x+\frac{y}{N})N^{-s}\tilde{g}(\frac{y}{N}).

For each K1,K2{1,,N2}K_{1},K_{2}\in\{1,\ldots,\frac{N}{2}\}, one may write using the Euler-Maclaurin formula

K1kK2,k0ϕ(k)=[K1,K2][1,1]ϕ(y)dy+ϕ(K2)+ϕ(K1)ϕ(1)ϕ(1)2+O(N21|ϕ(y)|dy+1N2|ϕ(y)|dy).\sum_{-K_{1}\leq k\leq K_{2},k\neq 0}\phi(k)=\int_{[-K_{1},K_{2}]\setminus[-1,1]}\phi(y)\mathrm{d}y+\frac{\phi(K_{2})+\phi(-K_{1})-\phi(1)-\phi(-1)}{2}\\ +O\Bigr{(}\int_{-\frac{N}{2}}^{-1}|\phi^{\prime}(y)|\mathrm{d}y+\int_{1}^{\frac{N}{2}}|\phi^{\prime}(y)|\mathrm{d}y\Bigr{)}.

We compute

ϕ(1)+ϕ(1)=(ζ(x,x+1N)ζ(x,x1N))Nsg~(1N)+(1ζ(x,x+1N)+1ζ(x,x1N))Nsg~(1N).\phi(1)+\phi(-1)=\Bigr{(}\zeta(x,x+\frac{1}{N})-\zeta(x,x-\frac{1}{N})\Bigr{)}N^{-s}\tilde{g}^{\prime}(\frac{1}{N})+\Bigr{(}\partial_{1}\zeta(x,x+\frac{1}{N})+\partial_{1}\zeta(x,x-\frac{1}{N})\Bigr{)}N^{-s}\tilde{g}(\frac{1}{N}).

By Taylor expansion,

|ζ(x,x+1N)ζ(x,x1N)|CN1ηN(N1x)1N,|\zeta(x,x+\frac{1}{N})-\zeta(x,x-\frac{1}{N})|\leq C\ell_{N}^{-1}\eta_{N}(\ell_{N}^{-1}x)\frac{1}{N},

and therefore

|ϕ(1)+ϕ(1)|CN1ηN(N1).|\phi(1)+\phi(-1)|\leq C\ell_{N}^{-1}\eta_{N}(\ell_{N}^{-1}\cdot). (B.14)

Let us control |ϕ|\int|\phi^{\prime}|. First, as in (B.5), for all yN𝕋y\in N\mathbb{T},

|2ζ(x,x+yN)|CN1(ηN(N1x)+ηN(N1(x+yN)).\Bigr{|}\partial_{2}\zeta(x,x+\frac{y}{N})\Bigr{|}\leq C\ell_{N}^{-1}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N})\Bigr{)}.

It follows that

1N2|2ζ(x,x+yN)N(1+s)g~(yN)|dyCN11N2(ηN(N1x)+ηN(N1(x+yN)))1y1+sdy.\int_{1}^{\frac{N}{2}}\Bigr{|}\partial_{2}\zeta(x,x+\frac{y}{N})N^{-(1+s)}\tilde{g}^{\prime}(\frac{y}{N})\Bigr{|}\mathrm{d}y\leq C\ell_{N}^{-1}\int_{1}^{\frac{N}{2}}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{1}{y^{1+s}}\mathrm{d}y.

We recognize an expression similar to the right-hand side of (B.6). After some computations one gets

N11N2(ηN(N1x)+ηN(N1(x+yN)))1y1+sdyC(NN)κεl=1pN1s+α1(|xNal|1N)2s+α𝟙|x|<2N+C(NN)κε𝟙α>2s1l=1p(NN)1s+αNs1(|xNal|1N)1+s+C(NN)κεNs1(|x|+N)1+s.\ell_{N}^{-1}\int_{1}^{\frac{N}{2}}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{1}{y^{1+s}}\mathrm{d}y\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{l=1}^{p}\ell_{N}^{1-s+\alpha}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\mathds{1}_{|x|<2\ell_{N}}\\ +C(N\ell_{N})^{\kappa\varepsilon}\mathds{1}_{\alpha>2s-1}\sum_{l=1}^{p}(N\ell_{N})^{1-s+\alpha}N^{-s}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{1+s}}+C(N\ell_{N})^{\kappa\varepsilon}N^{-s}\frac{1}{(|x|+\ell_{N})^{1+s}}. (B.15)

Using that for all x𝕋x\in\mathbb{T}, y(N2,N2)y\in(-\frac{N}{2},\frac{N}{2}),

|ζ(x,x+yN)|CN1(ηN(N1x)+ηN(N1(x+yN)))|y|N,|\zeta(x,x+\frac{y}{N})|\leq C\ell_{N}^{-1}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{|y|}{N},

one gets

1N2|ζ(x,x+yN)|N(s+1)g~(yN)dyCN11N2(ηN(N1x)+ηN(N1(x+yN)))1y1+sdy,\int_{1}^{\frac{N}{2}}|\zeta(x,x+\frac{y}{N})|N^{-(s+1)}\tilde{g}^{\prime\prime}(\frac{y}{N})\mathrm{d}y\leq C\ell_{N}^{-1}\int_{1}^{\frac{N}{2}}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{1}{y^{1+s}}\mathrm{d}y,

which is bounded by the right-hand side of (B.15). Finally, noting that for all x𝕋x\in\mathbb{T}, y(N2,N2)y\in(-\frac{N}{2},\frac{N}{2}),

|21ζ(x,x+yN)|CN1(ηN(N1x)+ηN(N1(x+yN)))N|y|,|\partial_{21}\zeta(x,x+\frac{y}{N})|\leq C\ell_{N}^{-1}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{N}{|y|},

we have

1N21N|21ζ(x,x+yN)|Nsg~(yN)dyCN11N2(ηN(N1x)+ηN(N1(x+yN)))1y1+sdy.\int_{1}^{\frac{N}{2}}\frac{1}{N}|\partial_{21}\zeta(x,x+\frac{y}{N})|N^{-s}\tilde{g}(\frac{y}{N})\mathrm{d}y\leq C\ell_{N}^{-1}\int_{1}^{\frac{N}{2}}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x)+\eta_{N}(\ell_{N}^{-1}(x+\frac{y}{N}))\Bigr{)}\frac{1}{y^{1+s}}\mathrm{d}y.

Combining these estimates with (B.15) and (B.14) we deduce that

|Vi2|C(NN)κεl=1pN1s+α1(|xNal|1N)2s+α𝟙|x|<2N+C(NN)κε𝟙α>2s1l=1p(NN)1s+αNs1(|xNal|1N)1+s+C(NN)κεNs1(|x|+N)1+s+Cϕ(d(i,IN)).|\mathrm{V}_{i}^{2}|\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{l=1}^{p}\ell_{N}^{1-s+\alpha}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{2-s+\alpha}}\mathds{1}_{|x|<2\ell_{N}}\\ +C(N\ell_{N})^{\kappa\varepsilon}\mathds{1}_{\alpha>2s-1}\sum_{l=1}^{p}(N\ell_{N})^{1-s+\alpha}N^{-s}\frac{1}{(|x-\ell_{N}a_{l}|\vee\frac{1}{N})^{1+s}}\\ +C(N\ell_{N})^{\kappa\varepsilon}N^{-s}\frac{1}{(|x|+\ell_{N})^{1+s}}+C\phi(d(i,\partial I_{N})). (B.16)

Let K{minIN,maxIN}K\in\{\min I_{N},\max I_{N}\}. One has

ϕ(Ki)=ζ(xi,xi+uNKi)Nsg~(uNKi)+1ζ(xi,xi+uNKi)Nsg~(uNKi).\phi(K-i)=\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}^{\prime}(\mathrm{u}^{N}_{K-i})+\partial_{1}\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}(\mathrm{u}^{N}_{K-i}).

On the event 𝒜\mathcal{A} we have

|1ζ(xi,xi+uNKi)Nsg~(uNKi)|C(NN)κεN1ηN(N1xi).|\partial_{1}\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}(\mathrm{u}^{N}_{K-i})|\leq C(N\ell_{N})^{\kappa\varepsilon}\ell_{N}^{-1}\eta_{N}(\ell_{N}^{-1}x_{i}).

Let iINi\in I_{N} such that d(i,i0)12d(i,K)d(i,i_{0})\leq\frac{1}{2}d(i,K). One has

|ζ(xi,xi+uNKi)Nsg~(uNKi)|CNK1+sγN(N1xi)C(NN)γsN1γN(N1xi),|\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}^{\prime}(\mathrm{u}^{N}_{K-i})|\leq\frac{CN}{K^{1+s}}\gamma_{N}(\ell_{N}^{-1}x_{i})\leq\frac{C}{(N\ell_{N})^{\gamma s}}\ell_{N}^{-1}\gamma_{N}(\ell_{N}^{-1}x_{i}),

since γ>1\gamma>1 and NN1N\ell_{N}\geq 1. Let iINi\in I_{N} such that d(i,i0)12d(i,K)d(i,i_{0})\geq\frac{1}{2}d(i,K). One has

|ζ(xi,xi+uNKi)Nsg~(uNKi)|CN|Ki|1+s1(N1|xi|+1)2s.|\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}^{\prime}(\mathrm{u}^{N}_{K-i})|\leq\frac{CN}{|K-i|^{1+s}}\frac{1}{(\ell_{N}^{-1}|x_{i}|+1)^{2-s}}.

Define I:=((NN)γN,(NN)γN)I:=(-\frac{(N\ell_{N})^{\gamma}}{N},\frac{(N\ell_{N})^{\gamma}}{N}). One can notice that on 𝒜\mathcal{A},

d(K,i)(NN)εN(d(xi,supI)1N)d(K,i)\geq(N\ell_{N})^{-\varepsilon}N\Bigr{(}d(x_{i},\sup I)\vee\frac{1}{N}\Bigr{)}

and therefore

|ζ(xi,xi+uNKi)Nsg~(uNKi)|C(NN)κεNs1d(xi,supI)1+s1(NN)(γ1)(2s).|\zeta(x_{i},x_{i}+\mathrm{u}^{N}_{K-i})N^{-s}\tilde{g}^{\prime}(\mathrm{u}^{N}_{K-i})|\leq C(N\ell_{N})^{\kappa\varepsilon}N^{-s}\frac{1}{d(x_{i},\sup I)^{1+s}}\frac{1}{(N\ell_{N})^{(\gamma-1)(2-s)}}.

It follows that on the event 𝒜\mathcal{A},

|ϕ(Ki)|C(NN)κε(ηN(N1xi)+Ns1(d(xi,supI)1N)1+s(NN)(γ1)(2s)).|\phi(K-i)|\leq C(N\ell_{N})^{\kappa\varepsilon}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x_{i})+N^{-s}\frac{1}{(d(x_{i},\sup I)\vee\frac{1}{N})^{1+s}}(N\ell_{N})^{-(\gamma-1)(2-s)}\Bigr{)}. (B.17)

Using that for all t1Nt\geq\frac{1}{N},

Nst1+sNs2t1+s2,\frac{N^{-s}}{t^{1+s}}\leq\frac{N^{-\frac{s}{2}}}{t^{1+\frac{s}{2}}},

one may bound (B.17) by

|ϕ(Ki)|C(NN)κε(ηN(N1xi)+Ns21(d(xi,supI)1N)1+s2(NN)(γ1)(2s)+N1γN(N1xi))|\phi(K-i)|\leq C(N\ell_{N})^{\kappa\varepsilon}\Bigr{(}\eta_{N}(\ell_{N}^{-1}x_{i})+N^{-\frac{s}{2}}\frac{1}{(d(x_{i},\sup I)\vee\frac{1}{N})^{1+\frac{s}{2}}}(N\ell_{N})^{-(\gamma-1)(2-s)}\\ +\ell_{N}^{-1}\gamma_{N}(\ell_{N}^{-1}x_{i})\Bigr{)}

Inserting this into (B.16) and using (B.13), we have that

|Vi|C(NN)κε(ζN(1)(xi)+ζN(2)(xi)),|\mathrm{V}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}(\zeta_{N}^{(1)}(x_{i})+\zeta_{N}^{(2)}(x_{i})),

where ζN(1)\zeta_{N}^{(1)} and ζN(2)\zeta_{N}^{(2)} are as in (5.15), (5.16).

Step 4: control on W\mathrm{W} in gap coordinates

Let W~L2({1,,N},H1(N,β))\tilde{\mathrm{W}}\in L^{2}(\{1,\ldots,N\},H^{1}(\mathbb{P}_{N,\beta})) be the vector-field given for each iINi\in I_{N} by

W~i=k=1NlIN:l+kIN,ik<liζ(xl+k,xl)N(1+s)(g~(xl+kxl)g~(ukN))(δkN/2+12δk=N/2)\tilde{\mathrm{W}}_{i}=\sum_{k=1}^{N}\sum_{\begin{subarray}{c}l\in I_{N}:\\ l+k\in I_{N},i-k<l\leq i\end{subarray}}\zeta(x_{l+k},x_{l})N^{-(1+s)}(\tilde{g}^{\prime}(x_{l+k}-x_{l})-\tilde{g}^{\prime}(\mathrm{u}_{k}^{N}))(\delta_{k\neq N/2}+\frac{1}{2}\delta_{k=N/2})

and W~i=0\tilde{\mathrm{W}}_{i}=0 for iINi\notin I_{N}. For all UNNU_{N}\in\mathbb{R}^{N}, we have

WUN=i=1NW~iN(ui+1ui).\mathrm{W}\cdot U_{N}=-\sum_{i=1}^{N}\tilde{\mathrm{W}}_{i}N(u_{i+1}-u_{i}).

There exist constants C,κ>0C,\kappa>0 such that uniformly on 𝒜\mathcal{A}, for each l,l+kINl,l+k\in I_{N}

N(1+s)|g~(xl+kxk)g~(ukN)|C(NN)κεk2+s2.N^{-(1+s)}|\tilde{g}^{\prime}(x_{l+k}-x_{k})-\tilde{g}^{\prime}(\mathrm{u}_{k}^{N})|\leq\frac{C(N\ell_{N})^{\kappa\varepsilon}}{k^{2+\frac{s}{2}}}. (B.18)

One may thus bound W~i\tilde{\mathrm{W}}_{i} uniformly on 𝒜\mathcal{A} by

|W~i|C(NN)κεkINj=i+1K(|γN(N1xk)|+|γN(N1xj)|)1d(j,k)2+s2.|\tilde{\mathrm{W}}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{k\in I_{N}}\sum_{j=i+1}^{K}(|\gamma_{N}(\ell_{N}^{-1}x_{k})|+|\gamma_{N}(\ell_{N}^{-1}x_{j})|)\frac{1}{d(j,k)^{2+\frac{s}{2}}}.

Reindexing this sum gives

|W~i|C(NN)κεj=1K|γN(N1xj)|1d(j,i)1+s2.|\tilde{\mathrm{W}}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{j=1}^{K}|\gamma_{N}(\ell_{N}^{-1}x_{j})|\frac{1}{d(j,i)^{1+\frac{s}{2}}}. (B.19)

Recalling (B.9), for each l=1,,pl=1,\ldots,p and jINj\in I_{N} we have that uniformly on 𝒜\mathcal{A},

N|xjNal|(NN)εd(j,kl).N|x_{j}-\ell_{N}a_{l}|\geq(N\ell_{N})^{-\varepsilon}d(j,k^{l}).

Furthermore there exists a constant C0>0C_{0}>0 such that for each i,jINi,j\in I_{N} and uniformly on 𝒜\mathcal{A}

d(j,i)C0NN|xj|>2N.d(j,i)\geq C_{0}N\ell_{N}\Longrightarrow|x_{j}|>2\ell_{N}.

Inserting this into (B.19) we find that

|W~i|C(NN)κεd(j,i0)C0NNl=1p(NN)1s+α1d(j,kl)1s+α1d(j,i)1+s2+C(NN)2sj=1N1(NN+d(j,i0))2s1d(j,i)1+s2.|\tilde{\mathrm{W}}_{i}|\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\sum_{l=1}^{p}(N\ell_{N})^{1-s+\alpha}\frac{1}{d(j,k^{l})^{1-s+\alpha}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\\ +C(N\ell_{N})^{2-s}\sum_{j=1}^{N}\frac{1}{(N\ell_{N}+d(j,i_{0}))^{2-s}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}. (B.20)

Let 1lp1\leq l\leq p. Since αl(0,1)\alpha_{l}\in(0,1), one can observe that

d(j,i0)C0NN1d(j,kl)1s+α1d(j,i)1+s2C{1d(i,kl)1s+αif d(i,i0)2C0NNNNd(i,i0)2s2+αif d(i,i0)2C0NN.\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\frac{1}{d(j,k^{l})^{1-s+\alpha}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\leq C\begin{cases}\frac{1}{d(i,k^{l})^{1-s+\alpha}}&\text{if $d(i,i_{0})\leq 2C_{0}N\ell_{N}$}\\ \frac{N\ell_{N}}{d(i,i_{0})^{2-\frac{s}{2}+\alpha}}&\text{if $d(i,i_{0})\geq 2C_{0}N\ell_{N}$}.\end{cases}

Summing the squares of these over 1iK11\leq i\leq K-1 therefore gives

(NN)2(1s+α)1iK(d(j,i0)C0NN1d(j,kl)1s+α1d(j,i)1+s2)2C{NNif 1s+α(0,12)(NN)log(NN)if 1s+α=12(NN)2(1s+α)if 1s+α(12,1).(N\ell_{N})^{2(1-s+\alpha)}\sum_{1\leq i\leq K}\Bigr{(}\sum_{d(j,i_{0})\leq C_{0}N\ell_{N}}\frac{1}{d(j,k^{l})^{1-s+\alpha}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\Bigr{)}^{2}\\ \leq C\begin{cases}N\ell_{N}&\text{if $1-s+\alpha\in(0,\frac{1}{2})$}\\ (N\ell_{N})\log(N\ell_{N})&\text{if $1-s+\alpha=\frac{1}{2}$}\\ (N\ell_{N})^{2(1-s+\alpha)}&\text{if $1-s+\alpha\in(\frac{1}{2},1)$}.\end{cases} (B.21)

Besides we can check that

j=1N1(NN+d(j,i0))2s1d(j,i)1+s2C{1(NN)2sif d(i,i0)2NNNNd(i,i0)3sif d(i,i0)2NN.\sum_{j=1}^{N}\frac{1}{(N\ell_{N}+d(j,i_{0}))^{2-s}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\leq C\begin{cases}\frac{1}{(N\ell_{N})^{2-s}}&\text{if $d(i,i_{0})\leq 2N\ell_{N}$}\\ \frac{N\ell_{N}}{d(i,i_{0})^{3-s}}&\text{if $d(i,i_{0})\geq 2N\ell_{N}$}.\end{cases}

Summing the squares gives

(NN)2(2s)iIN(j=1N1(NN+d(j,i0))2s1d(j,i)1+s2)2CNN.(N\ell_{N})^{2(2-s)}\sum_{i\in I_{N}}\Bigr{(}\sum_{j=1}^{N}\frac{1}{(N\ell_{N}+d(j,i_{0}))^{2-s}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\Bigr{)}^{2}\leq CN\ell_{N}. (B.22)

Inserting (B.21) and (B.22) into (B.20) we concludes that there exist constants C>0C>0 and κ>0\kappa>0 such

sup𝒜|W~|2C(NN)κε(NN)max(1,2(1s+α)).\sup_{\mathcal{A}}|\tilde{\mathrm{W}}|^{2}\leq C(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N})^{\max(1,2(1-s+\alpha))}. (B.23)

B.3. Additional useful estimates

Lemma B.3.

Let ψreg\psi_{\mathrm{reg}} be as in (5.20) and BN[ψreg]\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}] be as in (5.28). We have

𝔼N,β[BN[ψreg]]C(NN)κε(NN+(NN)2(1s+α)).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{\kappa\varepsilon}(N\ell_{N}+(N\ell_{N})^{2(1-s+\alpha)}). (B.24)
Proof.

Define

γN:xN1𝕋(1+l=1p1|xal|1s+α)𝟙|x|<2+1(1+|x|)2s,\gamma_{N}:x\in\ell_{N}^{-1}\mathbb{T}\mapsto\Bigr{(}1+\sum_{l=1}^{p}\frac{1}{|x-a_{l}|^{1-s+\alpha}}\Bigr{)}\mathds{1}_{|x|<2}+\frac{1}{(1+|x|)^{2-s}},

so that |ψreg|CγN.|\psi^{\prime}_{\mathrm{reg}}|\leq C\gamma_{N}. Let us recall that

BN[ψreg]=ΔcN(s+2)g(xy)(NN)2(ψreg(N1x)ψreg(N1y))2d(Ndx)dfluctN(y).\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]=\iint_{\Delta^{c}}N^{-(s+2)}g^{\prime\prime}(x-y)(N\ell_{N})^{2}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}\mathrm{d}(N\mathrm{d}x)\mathrm{d}\mathrm{fluct}_{N}(y).

Denote

h:(x,y)(N1𝕋)2ΔN(s+2)g(xy)(NN)2(ψreg(N1x)ψreg(N1y))2,h:(x,y)\in(\ell_{N}^{-1}\mathbb{T})^{2}\setminus\Delta\mapsto N^{-(s+2)}g^{\prime\prime}(x-y)(N\ell_{N})^{2}(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2},

so that

𝔼N,β[BN[ψreg]]=i=1N𝔼N,β[j:jih(xj,xi)Nh(y,xi)dy].\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{B}_{\ell_{N}}[\psi_{\mathrm{reg}}]]=\sum_{i=1}^{N}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{j:j\neq i}h(x_{j},x_{i})-N\int h(y,x_{i})\mathrm{d}y\Bigr{]}.

Let uN\mathrm{u}^{N} be the regular grid on 𝕋\mathbb{T},

uiN=iN𝕋,i=1,,N.\mathrm{u}_{i}^{N}=\frac{i}{N}\in\mathbb{T},\quad i=1,\ldots,N.

One may write

j:jih(xj,xi)Nh(y,xi)dy=j:ji(h(xj,xi)h(xi+ujiN,xi))+j:jih(xi+ujiN,xi)Nh(xi+y,xi)dy:=Ei1+Ei2.\sum_{j:j\neq i}h(x_{j},x_{i})-N\int h(y,x_{i})\mathrm{d}y=\sum_{j:j\neq i}(h(x_{j},x_{i})-h(x_{i}+\mathrm{u}_{j-i}^{N},x_{i}))\\ +\sum_{j:j\neq i}h(x_{i}+\mathrm{u}_{j-i}^{N},x_{i})-N\int h(x_{i}+y,x_{i})\mathrm{d}y:=\mathrm{E}_{i}^{1}+\mathrm{E}_{i}^{2}. (B.25)

Note that for all x,y,ax,y,a,

|(ψreg(N1x)ψreg(N1y))2(ψreg(N1x)ψreg(N1(y+a)))2|C(γN(N1x)+γN(N1y)+γN(N1a))N1(|xy|+|xya|)N1|a|(γN(N1y)+γN(N1(y+a))).|(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}y))^{2}-(\psi_{\mathrm{reg}}(\ell_{N}^{-1}x)-\psi_{\mathrm{reg}}(\ell_{N}^{-1}(y+a)))^{2}|\\ \leq C(\gamma_{N}(\ell_{N}^{-1}x)+\gamma_{N}(\ell_{N}^{-1}y)+\gamma_{N}(\ell_{N}^{-1}a))\ell_{N}^{-1}(|x-y|+|x-y-a|)\ell_{N}^{-1}|a|(\gamma_{N}(\ell_{N}^{-1}y)+\gamma_{N}(\ell_{N}^{-1}(y+a))).

It follows that for each i,ji,j

|h(xi,xj)h(xi,xi+ujiN)|C(γN(N1xi)+γN(N1xj)+γN(N1(xi+ujiN)))2×(N|xjxi|+N|ujiN|)|N(xjxiujiN)||N(xjxi)|s+2,|h(x_{i},x_{j})-h(x_{i},x_{i}+\mathrm{u}_{j-i}^{N})|\leq C\Bigr{(}\gamma_{N}(\ell_{N}^{-1}x_{i})+\gamma_{N}(\ell_{N}^{-1}x_{j})+\gamma_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N}))\Bigr{)}^{2}\\ \times\frac{(N|x_{j}-x_{i}|+N|\mathrm{u}_{j-i}^{N}|)|N(x_{j}-x_{i}-\mathrm{u}_{j-i}^{N})|}{|N(x_{j}-x_{i})|^{s+2}},

which gives

𝔼N,β[|Ei1|]C𝔼N,β[j:ji(γN(N1xi)+γN(N1xj))(γN(N1xj)+γN(N1(xi+ujiN)))×|N(xjxi)NujiN||N(xjxi)|1+s].\mathbb{E}_{\mathbb{P}_{N,\beta}}[|\mathrm{E}_{i}^{1}|]\leq C\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{j:j\neq i}(\gamma_{N}(\ell_{N}^{-1}x_{i})+\gamma_{N}(\ell_{N}^{-1}x_{j}))(\gamma_{N}(\ell_{N}^{-1}x_{j})+\gamma_{N}(\ell_{N}^{-1}(x_{i}+\mathrm{u}_{j-i}^{N})))\\ \times\frac{|N(x_{j}-x_{i})-N\mathrm{u}_{j-i}^{N}|}{|N(x_{j}-x_{i})|^{1+s}}\Bigr{]}.

By applying Theorem 1 to control the fluctuations of gaps, we get

|Ei1|C(NN)κεji:|j|C0NNl=1p(NN)2(α+1s)1|jNNal|2(α+1s)1d(j,i)1+s2+C(NN)κεj:ji(NN)2(2s)1(j+NN)2s1d(j,i)1+s2,|\mathrm{E}_{i}^{1}|\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{\begin{subarray}{c}j\neq i:|j|\leq C_{0}N\ell_{N}\end{subarray}}\sum_{l=1}^{p}(N\ell_{N})^{2(\alpha+1-s)}\frac{1}{|j-N\ell_{N}a_{l}|^{2(\alpha+1-s)}}\frac{1}{d(j,i)^{1+\frac{s}{2}}}\\ +C(N\ell_{N})^{\kappa\varepsilon}\sum_{j:j\neq i}(N\ell_{N})^{2(2-s)}\frac{1}{(j+N\ell_{N})^{2-s}}\frac{1}{d(j,i)^{1+\frac{s}{2}}},

for some constant C0>0C_{0}>0. After some computations similar to the proof of Lemma 5.2, we find

𝔼N,β[i=1N|Ei1|]C(NN)κε+max(2(1s+α),1).\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}|\mathrm{E}_{i}^{1}|\Bigr{]}\leq C(N\ell_{N})^{\kappa\varepsilon+\max(2(1-s+\alpha),1)}. (B.26)

For the discretization error, proceeding as in the proof of Lemma 5.2, one can write

|j:jih(xi+ujiN,xi)Nh(xi+y,xi)dy|1N1N2|1h(xi+yN,xi)|dyC1N21y1+s(γN(N1xi)2+γN(N1(xi+yN))2)CγN(N1xi)2.\Bigr{|}\sum_{j:j\neq i}h(x_{i}+\mathrm{u}_{j-i}^{N},x_{i})-N\int h(x_{i}+y,x_{i})\mathrm{d}y|\leq\frac{1}{N}\int_{1}^{\frac{N}{2}}|\partial_{1}h(x_{i}+\frac{y}{N},x_{i})|\mathrm{d}y\\ \leq C\int_{1}^{\frac{N}{2}}\frac{1}{y^{1+s}}\Bigr{(}\gamma_{N}(\ell_{N}^{-1}x_{i})^{2}+\gamma_{N}(\ell_{N}^{-1}(x_{i}+\frac{y}{N}))^{2}\Bigr{)}\leq C\gamma_{N}(\ell_{N}^{-1}x_{i})^{2}.

Summing the above estimate yields

𝔼N,β[i=1N|Ei2|]C(NN)κε+max(2(1s+α),1).\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}|\mathrm{E}_{i}^{2}|\Bigr{]}\leq C(N\ell_{N})^{\kappa\varepsilon+\max(2(1-s+\alpha),1)}. (B.27)

In combination with (B.25) and (B.26), this gives (B.24). ∎

Lemma B.4.

Let ψreg\psi_{\mathrm{reg}} be as in (5.20). Let ANext[ψreg]:=AN[ψreg]A~N[ψreg]\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]:=\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}] with AN[ψreg]\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}] and A~N[ψreg]\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}] as in (5.1), (5.13). Assume γ>3s2s1s\gamma>\frac{3-s}{2-s}\vee\frac{1}{s}. There exist constants C>0,δ>0C>0,\delta>0 depending on γ\gamma such that

VarN,β[ANext[ψreg]]C(NN)1δ.\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]]\leq C(N\ell_{N})^{1-\delta}. (B.28)
Proof.

Let ε>0\varepsilon>0 be a small number. Define the good event

={XNDN:1iN,1kN/2,|N(xi+kxi)k|ks2+ε(NN)s2+ε}{XNDN:1iN,(NN)εN(xi+1xi)(NN)ε}.\mathcal{B}=\{X_{N}\in D_{N}:\forall 1\leq i\leq N,1\leq k\leq N/2,|N(x_{i+k}-x_{i})-k|\leq k^{\frac{s}{2}+\varepsilon}\vee(N\ell_{N})^{\frac{s}{2}+\varepsilon}\}\\ \cap\{X_{N}\in D_{N}:\forall 1\leq i\leq N,(N\ell_{N})^{-\varepsilon}\leq N(x_{i+1}-x_{i})\leq(N\ell_{N})^{\varepsilon}\}.

Let us bound ANext[ψreg]\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}] on \mathcal{B}. By assumption, |ψreg|CγN|\psi_{\mathrm{reg}}^{\prime}|\leq C\gamma_{N} where

γN:xN1𝕋(1+l=1p1|xal|1s+α)𝟙|x|<2+11+|x|2s.\gamma_{N}:x\in\ell_{N}^{-1}\mathbb{T}\mapsto\Bigr{(}1+\sum_{l=1}^{p}\frac{1}{|x-a_{l}|^{1-s+\alpha}}\Bigr{)}\mathds{1}_{|x|<2}+\frac{1}{1+|x|^{2-s}}.

It follows that ANext[ψreg]\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}] can be bounded on \mathcal{B} by

|ANext[ψreg]|C(NN)κεi=1NjINc(γN(N1xi)+γN(N1xj))11+d(j,i)1+s2εC(NN)κεi=1NγN(N1xi)11+d(i,INc)s2ε+C(NN)κεjINcγN(N1xj)C(NN)κεiINγN(N1xi)1d(i,INc)s2ε+C(NN)κεjINcγN(N1xj).\begin{split}|\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]|&\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{i=1}^{N}\sum_{j\in I_{N}^{c}}(\gamma_{N}(\ell_{N}^{-1}x_{i})+\gamma_{N}(\ell_{N}^{-1}x_{j}))\frac{1}{1+d(j,i)^{1+\frac{s}{2}-\varepsilon}}\\ &\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{i=1}^{N}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{1+d(i,I_{N}^{c})^{\frac{s}{2}-\varepsilon}}+C(N\ell_{N})^{\kappa\varepsilon}\sum_{j\in I_{N}^{c}}\gamma_{N}(\ell_{N}^{-1}x_{j})\\ &\leq C(N\ell_{N})^{\kappa\varepsilon}\sum_{i\in I_{N}}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{d(i,\partial I_{N}^{c})^{\frac{s}{2}-\varepsilon}}+C(N\ell_{N})^{\kappa\varepsilon}\sum_{j\in I_{N}^{c}}\gamma_{N}(\ell_{N}^{-1}x_{j}).\end{split} (B.29)

The second sum of the last display is bounded by

|jINcγN(N1xj)|C(NN)κε+(2s)k(NN)γ1k2sεC(NN)κε+(2s)γ(1s).|\sum_{j\in I_{N}^{c}}\gamma_{N}(\ell_{N}^{-1}x_{j})|\leq C(N\ell_{N})^{\kappa\varepsilon+(2-s)}\sum_{k\geq(N\ell_{N})^{\gamma}}\frac{1}{k^{2-s-\varepsilon}}\leq C^{\prime}(N\ell_{N})^{\kappa^{\prime}\varepsilon+(2-s)-\gamma(1-s)}. (B.30)

Recall |IN|=1+2η|I_{N}|=1+2\lfloor\eta\rfloor with η\eta as in (5.11). One may split the first sum into

iINγN(N1xi)1d(i,INc)s2ε=iIN,d(i,IN)>12(NN)γγN(N1xi)1d(i,INc)γ(s2ε)+iIN,d(i,IN)12(NN)γγN(N1xi)1d(i,INc)s2+εCiINγN(N1xi)1(NN)γ(s2ε)+CiIN,d(i,INc)12(NN)γγ(N1xi).\sum_{i\in I_{N}}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{d(i,\partial I_{N}^{c})^{\frac{s}{2}-\varepsilon}}=\sum_{i\in I_{N},d(i,\partial I_{N})>\frac{1}{2}(N\ell_{N})^{\gamma}}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{d(i,\partial I_{N}^{c})^{\gamma(\frac{s}{2}-\varepsilon)}}\\ +\sum_{i\in I_{N},d(i,\partial I_{N})\leq\frac{1}{2}(N\ell_{N})^{\gamma}}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{d(i,\partial I_{N}^{c})^{\frac{s}{2}+\varepsilon}}\leq C\sum_{i\in I_{N}}\gamma_{N}(\ell_{N}^{-1}x_{i})\frac{1}{(N\ell_{N})^{\gamma(\frac{s}{2}-\varepsilon)}}\\ +C\sum_{i\in I_{N},d(i,I_{N}^{c})\leq\frac{1}{2}(N\ell_{N})^{\gamma}}\gamma(\ell_{N}^{-1}x_{i}).

Since the singularities of ψreg\psi_{\mathrm{reg}}^{\prime} are in L1L^{1}, one can check that uniformly on \mathcal{B},

iINγN(N1xi)C(NN)κε+1.\sum_{i\in I_{N}}\gamma_{N}(\ell_{N}^{-1}x_{i})\leq C(N\ell_{N})^{\kappa\varepsilon+1}. (B.31)

Moreover arguing as in (B.30), we find that on \mathcal{B},

iIN,d(i,INc)12(NN)γγN(N1xi)C(NN)κε+(2s)γ(1s).\sum_{i\in I_{N},d(i,I_{N}^{c})\leq\frac{1}{2}(N\ell_{N})^{\gamma}}\gamma_{N}(\ell_{N}^{-1}x_{i})\leq C(N\ell_{N})^{\kappa\varepsilon+(2-s)-\gamma(1-s)}.

Combining this with (B.29), (B.30) and (B.31), one finally gets that on \mathcal{B},

|ANext[ψreg]|C(NN)κε((NN)2sγ(1sε)+(NN)1γ(s2ε)).|\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]|\leq C(N\ell_{N})^{\kappa\varepsilon}((N\ell_{N})^{2-s-\gamma(1-s-\varepsilon)}+(N\ell_{N})^{1-\gamma(\frac{s}{2}-\varepsilon)}).

Choosing γ>3s2s1s\gamma>\frac{3-s}{2-s}\vee\frac{1}{s}, we thus get that for ε>0\varepsilon>0 small enough, ANext[ψreg]\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}] satisfies

sup|ANext[ψreg]|C(NN)12δ,\sup_{\mathcal{B}}|\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]|\leq C(N\ell_{N})^{\frac{1}{2}-\delta},

for some δ>0\delta>0. Moreover, one can easily prove using Theorem 1 and Lemma 4.5 that there exists δ>0\delta>0 and C>0C>0 depending on ε\varepsilon and γ\gamma such that

𝔼N,β[(ANext[ψreg])2𝟙c]Ce(NN)δ,\mathbb{E}_{\mathbb{P}_{N,\beta}}[(\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}])^{2}\mathds{1}_{\mathcal{B}^{c}}]\leq Ce^{-(N\ell_{N})^{\delta}},

which concludes the proof. ∎

We finish with two conditioning lemmas.

Lemma B.5.

Let ψreg\psi_{\mathrm{reg}} be as in (5.20) and A~N[ψreg]\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}] be as in (5.13). We have

VarN,β[𝔼N,β[A~N[ψreg]x1]]C((NN)2(1s)+NN).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathbb{E}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}]]\leq C((N\ell_{N})^{2(1-s)}+N\ell_{N}).
Proof.

Let :=N,β\mathcal{L}:=\mathcal{L}^{\mathbb{P}_{N,\beta}}. First recall that for any ψ:𝕋\psi:\mathbb{T}\to\mathbb{R} smooth enough such that ψ=0\int\psi=0,

𝔼N,β[AN[ψ]]=0.\mathbb{E}_{\mathbb{P}_{N,\beta}}[{\mathrm{A}}_{\ell_{N}}[\psi]]=0. (B.32)

Let indeed ξ\xi such that ξξ:=2βgψ\xi-\int\xi:=-2\beta g*\psi and

Φ:XN1(NN)1sN(ψ(N1x1),,ψ(N1xN)).\nabla\Phi:X_{N}\mapsto\frac{1}{(N\ell_{N})^{1-s}}\ell_{N}(\psi(\ell_{N}^{-1}x_{1}),\ldots,\psi(\ell_{N}^{-1}x_{N})).

As shown in the proof of Proposition 5.1,

0=𝔼N,β[Φ]=𝔼N,β[FluctN[ξ(N1)]1(NN)1sFluctN[ψ(N1)]+β(NN)1sAN[ψ]].0=\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathcal{L}\Phi]=\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\mathrm{Fluct}_{N}[\xi(\ell_{N}^{-1}\cdot)]-\frac{1}{(N\ell_{N})^{1-s}}\mathrm{Fluct}_{N}[\psi^{\prime}(\ell_{N}^{-1}\cdot)]+\frac{\beta}{(N\ell_{N})^{1-s}}\mathrm{A}_{\ell_{N}}[\psi]\Bigr{]}.

Using the fact that the first marginal of N,β\mathbb{P}_{N,\beta} is the Lebesgue measure on the circle, we obtain (B.32). Let i0:=argmin1iN|xi|i_{0}:=\underset{1\leq i\leq N}{\mathrm{argmin}}|x_{i}|. The point is that

LawN,β(x1,,xNx1=x)=LawN,β(x1+x,,xN+xx1=0)=LawN,β(x1+x,,xN+xxi0=0)=LawN,β(x1xi0+x,,xNxi0+x).\begin{split}\mathrm{Law}_{\mathbb{P}_{N,\beta}}(x_{1},\ldots,x_{N}\mid x_{1}=x)&=\mathrm{Law}_{\mathbb{P}_{N,\beta}}(x_{1}+x,\ldots,x_{N}+x\mid x_{1}=0)\\ &=\mathrm{Law}_{\mathbb{P}_{N,\beta}}(x_{1}+x,\ldots,x_{N}+x\mid x_{i_{0}}=0)\\ &=\mathrm{Law}_{\mathbb{P}_{N,\beta}}(x_{1}-x_{i_{0}}+x,\ldots,x_{N}-x_{i_{0}}+x).\end{split} (B.33)

Fix x0𝕋x_{0}\in\mathbb{T} and let us denote ϕx0:=ψreg(x0+)\phi_{x_{0}}:=\psi_{\mathrm{reg}}(x_{0}+\cdot). In view of (B.33), we have

𝔼N,β[AN[ψreg]x1=x0]=𝔼N,β[ΔcN(1+s)g(xy)×NN(ϕx0(N1(xxi0))ϕx0(N1(yxi0))dfluctN(x)dfluctN(y)].\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\iint_{\Delta^{c}}N^{-(1+s)}g^{\prime}(x-y)\times\\ N\ell_{N}(\phi_{x_{0}}(\ell_{N}^{-1}(x-x_{i_{0}}))-\phi_{x_{0}}(\ell_{N}^{-1}(y-x_{i_{0}}))\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y)\Bigr{]}.

Recall that typically |xi0|=O(1N)|x_{i_{0}}|=O(\frac{1}{N}). Using Theorem 1 to control N,β(|xi0|>M)\mathbb{P}_{N,\beta}(|x_{i_{0}}|>M) and the fact that ψreg\psi_{\mathrm{reg}}^{\prime} is in L1L^{1}, we obtain by Taylor expansion that

𝔼N,β[AN[ψreg]x1=x0]=𝔼N,β[AN[ϕx0]]+O((NN)1s)=O((NN)1s),\mathbb{E}_{\mathbb{P}_{N,\beta}}[{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]=\mathbb{E}_{\mathbb{P}_{N,\beta}}[{\mathrm{A}}_{\ell_{N}}[\phi_{x_{0}}]]+O((N\ell_{N})^{1-s})=O((N\ell_{N})^{1-s}),

where we have used (B.32) for ψ=ϕx0\psi=\phi_{x_{0}} in the last equality. Thus

VarN,β[𝔼N,β[AN[ψreg]x1=x0]]=O((NN)2(1s)).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]]=O((N\ell_{N})^{2(1-s)}). (B.34)

Let us extend this control to the localized quantity A~N[ψreg]\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]. Recall

ANext[ψreg]=AN[ψreg]A~N[ψreg].\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]=\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]-\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}].

One can write

VarN,β[A~N[ψreg]x1=x0]2(VarN,β[AN[ψreg]x1=x0]+VarN,β[ANext[ψreg]x1=x0]).\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]\leq 2(\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]+\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]).

Therefore inserting (B.34) and (B.28), we get

𝔼N,β[VarN,β[A~N[ψreg]x1=x0]]2𝔼N,β[VarN,β[AN[ψreg]x1=x0]]+2VarN,β[ANext[ψreg]]C((NN)2(1s)+NN)).\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\tilde{\mathrm{A}}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]]\leq 2\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}[\psi_{\mathrm{reg}}]\mid x_{1}=x_{0}]]\\ +2\mathrm{Var}_{\mathbb{P}_{N,\beta}}[\mathrm{A}_{\ell_{N}}^{\mathrm{ext}}[\psi_{\mathrm{reg}}]]\leq C((N\ell_{N})^{2(1-s)}+N\ell_{N})).

Lemma B.6.

Let γ>1\gamma>1 and χ1\chi_{1} be as in (5.31). For all ε>0\varepsilon>0, there exists C>0C>0 depending on γ\gamma such that for all x𝕋x\in\mathbb{T},

|𝔼N,β[FluctN[χ1]x1=x]Nχ1|C(NN)εNs2.\Bigr{|}\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]\mid x_{1}=x]-N\int\chi_{1}\Bigr{|}\leq C(N\ell_{N})^{\varepsilon}N^{\frac{s}{2}}.
Proof.

We have

𝔼N,β[FluctN[χ1]|x1=x]=𝔼N,β[i=1Nχ1(xix1+x)].\mathbb{E}_{\mathbb{P}_{N,\beta}}[\mathrm{Fluct}_{N}[\chi_{1}]|x_{1}=x]=\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\chi_{1}(x_{i}-x_{1}+x)\Bigr{]}.

Let uN\mathrm{u}^{N} be the regular grid on 𝕋\mathbb{T}, i.e

ukN=kNfor each k=1,,N.\mathrm{u}_{k}^{N}=\frac{k}{N}\quad\text{for each $k=1,\ldots,N$}.

Using the fact that 𝔼[xix1]=ui1N\mathbb{E}[x_{i}-x_{1}]=\mathrm{u}_{i-1}^{N} for each ii, we get by Taylor expansion that

|𝔼N,β[i=1Nχ1(xix1+x)]i=1Nχ1(ui1N+x)|12𝔼N,β[supy(min(ui1N+x,xix1+x),max(ui1N+x,xix1+x))|χ1|(y)((xix1)ui1N)2].\Bigr{|}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\chi_{1}(x_{i}-x_{1}+x)\Bigr{]}-\sum_{i=1}^{N}\chi_{1}(\mathrm{u}_{i-1}^{N}+x)\Bigr{|}\\ \leq\frac{1}{2}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sup_{y\in(\min(\mathrm{u}_{i-1}^{N}+x,x_{i}-x_{1}+x),\max(\mathrm{u}_{i-1}^{N}+x,x_{i}-x_{1}+x))}|\chi_{1}^{\prime\prime}|(y)((x_{i}-x_{1})-\mathrm{u}_{i-1}^{N})^{2}\Bigr{]}.

Using Theorem 1 and Lemma 4.5, we deduce that for all ε>0\varepsilon>0,

|𝔼N,β[i=1Nχ1(xix1+x)]i=1Nχ1(ui1N+x)|CNs1(NN)ε|χ1|=O((NN)εNs2).\Bigr{|}\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\sum_{i=1}^{N}\chi_{1}(x_{i}-x_{1}+x)\Bigr{]}-\sum_{i=1}^{N}\chi_{1}(\mathrm{u}_{i-1}^{N}+x)\Bigr{|}\leq CN^{s-1}(N\ell_{N})^{\varepsilon}\int|\chi_{1}^{\prime\prime}|=O((N\ell_{N})^{\varepsilon}N^{\frac{s}{2}}).

We conclude by using

|i=1Nχ1(ui1N+x)Nχ1|CN|χ1|=O(Ns2).\Bigr{|}\sum_{i=1}^{N}\chi_{1}(\mathrm{u}_{i-1}^{N}+x)-N\int\chi_{1}\Bigr{|}\leq CN\int|\chi_{1}^{\prime}|=O(N^{\frac{s}{2}}).

B.4. Energy estimate

Lemma B.7.

We have

𝔼N,β[Nsi=1Ng(xi+1xi)]=O(N).\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}N^{-s}\sum_{i=1}^{N}g(x_{i+1}-x_{i})\Bigr{]}=O(N).
Proof.

Let us define

FN:XNDNNsΔcg(xy)dfluctN(x)dfluctN(y).F_{N}:X_{N}\in D_{N}\mapsto N^{-s}\iint_{\Delta^{c}}g(x-y)\mathrm{d}\mathrm{fluct}_{N}(x)\mathrm{d}\mathrm{fluct}_{N}(y).

One can observe that

N=N2sg(xy)dxdy+FN.\mathcal{H}_{N}=N^{2-s}\iint g(x-y)\mathrm{d}x\mathrm{d}y+F_{N}.

Define

KN,β:=eβFN(XN)𝟙DN(XN)dXN=ZNβeβN2g(xy)dxdy.K_{N,\beta}:=\int e^{-\beta F_{N}(X_{N})}\mathds{1}_{D_{N}}(X_{N})\mathrm{d}X_{N}=Z_{N\beta}e^{\beta N^{2}\iint g(x-y)\mathrm{d}x\mathrm{d}y}.

Let us prove following the lecture notes [Ser24, Ch. 5] that logKN,β=O(N)\log K_{N,\beta}=O(N). By [Ser24, Cor. 4.2.4], we have

FNCN,F_{N}\geq-CN,

which gives logKN,βCN\log K_{N,\beta}\leq CN. For the lower bound, one can write by Jensen’s inequality

logKN,ββFN(XN)dx1dxN=βNs(ijg(xixj)2Ni=1Ng(xiy)dy+N2g(xy)dxdy)dx1dxN=βNs(N(N1)2N2+N2)g(xy)dxdy=βN1sg(xy)dxdy0.\log K_{N,\beta}\geq-\beta\int F_{N}(X_{N})\mathrm{d}x_{1}\ldots\mathrm{d}x_{N}\\ =-\beta N^{-s}\int\Bigr{(}\sum_{i\neq j}g(x_{i}-x_{j})-2N\sum_{i=1}^{N}\int g(x_{i}-y)\mathrm{d}y+N^{2}\iint g(x-y)\mathrm{d}x\mathrm{d}y\Bigr{)}\mathrm{d}x_{1}\ldots\mathrm{d}x_{N}\\ =-\beta N^{-s}(N(N-1)-2N^{2}+N^{2})\iint g(x-y)\mathrm{d}x\mathrm{d}y=\beta N^{1-s}\iint g(x-y)\mathrm{d}x\mathrm{d}y\geq 0.

We deduce that

logKN,β=O(N).\log K_{N,\beta}=O(N). (B.35)

It follows that

log𝔼N,β[eβ2FN]=logKN,β2logKN,β=O(N),\log\mathbb{E}_{\mathbb{P}_{N,\beta}}[e^{\frac{\beta}{2}F_{N}}]=\log K_{N,\frac{\beta}{2}}-\log K_{N,\beta}=O(N),

which yields by Hölder’s inequality

𝔼N,β[FN]CN.\mathbb{E}_{\mathbb{P}_{N,\beta}}[F_{N}]\leq CN. (B.36)

Besides note that

NNsi=1Ng(xi+1xi)+i=1Nk2(g(k)𝟙|N(xi+kxi|k+g(k)𝟙N|(xi+kxi|kO(|N(xi+kxi)k|s2+εg(k))).\mathcal{H}_{N}\geq N^{-s}\sum_{i=1}^{N}g(x_{i+1}-x_{i})+\sum_{i=1}^{N}\sum_{k\geq 2}\Bigr{(}g(k)\mathds{1}_{|N(x_{i+k}-x_{i}|\leq k}+g(k)\mathds{1}_{N|(x_{i+k}-x_{i}|\geq k}\\ -O\Bigr{(}\frac{|N(x_{i+k}-x_{i})-k|^{\frac{s}{2}+\varepsilon}}{g^{\prime}(k)}\Bigr{)}\Bigr{)}.

Therefore

𝔼N,β[NN2sg(xy)dxdy]𝔼N,β[Nsi=1Ng(xi+1xi)]O(N),\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}\mathcal{H}_{N}-N^{2-s}\iint g(x-y)\mathrm{d}x\mathrm{d}y\Bigr{]}\geq\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}N^{-s}\sum_{i=1}^{N}g(x_{i+1}-x_{i})\Bigr{]}-O(N),

where we have applied the rigidity estimate of Theorem 1. Using (B.36), we deduce that

𝔼N,β[Nsi=1Ng(xi+1xi)]=O(N).\mathbb{E}_{\mathbb{P}_{N,\beta}}\Bigr{[}N^{-s}\sum_{i=1}^{N}g(x_{i+1}-x_{i})\Bigr{]}=O(N).

References

  • [ABC+00] Cécile Ané, Sébastien Blachère, Djalil Chafaï, Pierre Fougères, Ivan Gentil, Florent Malrieu, Cyril Roberto, and Grégory Scheffer. Sur les inégalités de Sobolev logarithmiques, volume 10. Paris: Société Mathématique de France, 2000.
  • [AS19] Scott Armstrong and Sylvia Serfaty. Local laws and rigidity for coulomb gases at any temperature. arXiv: Mathematical Physics, 2019.
  • [AW19] Scott Armstrong and Wei Wu. 𝒞2\mathcal{C}^{2} regularity of the surface tension for the ϕ\nabla\phi interface model. arXiv preprint arXiv:1909.13325, 2019.
  • [BBNY16] Roland Bauerschmidt, Paul Bourgade, Miika Nikula, and Horng-Tzer Yau. The two-dimensional coulomb plasma: quasi-free approximation and central limit theorem. arXiv preprint arXiv:1609.08582, 2016.
  • [BÉ85] Dominique Bakry and Michel Émery. Diffusions hypercontractives. In Jacques Azéma and Marc Yor, editors, Séminaire de Probabilités XIX 1983/84, pages 177–206, Berlin, Heidelberg, 1985. Springer Berlin Heidelberg.
  • [Ber72] Bruce C. Berndt. On the Hurwitz zeta-function. Rocky Mountain Journal of Mathematics, 2(1):151 – 158, 1972.
  • [BEY12] Paul Bourgade, László Erdős, and Horng-Tzer Yau. Bulk universality of general β\beta-ensembles with non-convex potential. Journal of mathematical physics, 53(9):095221, 2012.
  • [BEY14a] Paul Bourgade, László Erdös, and Horng-Tzer Yau. Edge universality of β\beta ensembles. Communications in Mathematical Physics, 332(1):261–353, 2014.
  • [BEY+14b] Paul Bourgade, László Erdős, Horng-Tzer Yau, et al. Universality of general β\beta-ensembles. Duke Mathematical Journal, 163(6):1127–1190, 2014.
  • [BG13] Gaëtan Borot and Alice Guionnet. Asymptotic expansion of β\beta matrix models in the one-cut regime. Communications in Mathematical Physics, 317(2):447–483, 2013.
  • [BL02] Herm Jan Brascamp and Elliott H. Lieb. On Extensions of the Brunn-Minkowski and Prékopa-Leindler Theorems, Including Inequalities for Log Concave Functions, and with an Application to the Diffusion Equation, pages 441–464. Springer Berlin Heidelberg, Berlin, Heidelberg, 2002.
  • [BL15] Xavier Blanc and Mathieu Lewin. The Crystallization Conjecture: A Review. EMS Surveys in Mathematical Sciences, 2(2):255–306, 2015. Final version to appear in EMS Surv. Math. Sci.
  • [BLS+18] Florent Bekerman, Thomas Leblé, Sylvia Serfaty, et al. CLT for fluctuations of β\beta-ensembles with general potential. Electronic Journal of Probability, 23, 2018.
  • [BM92] Dominique Bakry and Dominique Michel. Sur les inégalités FKG. Séminaire de probabilités de Strasbourg, 26:170–188, 1992.
  • [BMP22] Paul Bourgade, Krishnan Mody, and Michel Pain. Optimal local law and central limit theorem for β\beta-ensembles. Communications in Mathematical Physics, 390(3):1017–1079, 2022.
  • [Bou21] Paul Bourgade. Extreme gaps between eigenvalues of wigner matrices. Journal of the European Mathematical Society, 2021.
  • [Bou22] Jeanne Boursier. Decay of correlations and thermodynamic limit for the circular Riesz gas. arXiv preprint arXiv:2209.00396, 2022.
  • [Car74] Pierre Cartier. Inégalités de corrélation en mécanique statistique. In Séminaire Bourbaki vol. 1972/73 Exposés 418–435, pages 242–264. Springer, 1974.
  • [CFLW21] Tom Claeys, Benjamin Fahs, Gaultier Lambert, and Christian Webb. How much can the eigenvalues of a random hermitian matrix fluctuate? Duke Mathematical Journal, 170(9):2085–2235, 2021.
  • [Cha19] Sourav Chatterjee. Rigidity of the three-dimensional hierarchical coulomb gas. Probability Theory and Related Fields, 175(3):1123–1176, 2019.
  • [CSW21] Djalil Chafaï, Edward B. Saff, and Robert S. Womersley. On the solution of a riesz equilibrium problem and integral identities for special functions, 2021.
  • [DW20] Paul Dario and Wei Wu. Massless phases for the villain model in d\geq3. arXiv preprint arXiv:2002.02946, 2020.
  • [Dys62] Freeman J. Dyson. A Brownian-motion model for the eigenvalues of a random matrix. J. Mathematical Phys., 3:1191–1198, 1962.
  • [EHL18] Matthias Erbar, Martin Huesmann, and Thomas Leblé. The one-dimensional log-gas free energy has a unique minimiser. arXiv preprint arXiv:1812.06929, 2018.
  • [ESY11] László Erdős, Benjamin Schlein, and Horng-Tzer Yau. Universality of random matrices and local relaxation flow. Inventiones mathematicae, 185(1):75–119, 2011.
  • [EY12] László Erdős and Horng-Tzer Yau. Gap universality of generalized wigner and β\beta-ensembles. preprint. arXiv preprint arXiv:1211.3786, 2012.
  • [EY15] László Erdős and Horng-Tzer Yau. Gap universality of generalized wigner and betabeta-ensembles. Journal of the European Mathematical Society, 17(8):1927–2036, 2015.
  • [GS20] Shirshendu Ganguly and Sourav Sarkar. Ground states and hyperuniformity of the hierarchical coulomb gas in all dimensions. Probability Theory and Related Fields, 177(3):621–675, 2020.
  • [Gus05] Jonas Gustavsson. Gaussian fluctuations of eigenvalues in the gue. In Annales de l’IHP Probabilités et statistiques, volume 41, pages 151–178, 2005.
  • [Hel98a] Bernard Helffer. Remarks on decay of correlations and witten laplacians brascamp–lieb inequalities and semiclassical limit. journal of functional analysis, 155(2):571–586, 1998.
  • [Hel98b] Bernard Helffer. Remarks on decay of correlations and witten laplacians brascamp–lieb inequalities and semiclassical limit. Journal of Functional Analysis, 155(2):571–586, 1998.
  • [HL16] Jiaoyang Huang and Benjamin Landon. Local law and mesoscopic fluctuations of dyson brownian motion for general β\beta and potential. arXiv preprint arXiv:1612.06306, 2016.
  • [HL20] Adrien Hardy and Gaultier Lambert. CLT for circular β\beta-ensembles at high temperature. Journal of Functional Analysis, page 108869, 2020.
  • [HL21] Adrien Hardy and Gaultier Lambert. CLT for circular β\beta-ensembles at high temperature. Journal of Functional Analysis, 280(7):108869, 2021.
  • [HS94] Bernard Helffer and Johannes Sjöstrand. On the correlation for kac-like models in the convex case. Journal of statistical physics, 74(1-2):349–409, 1994.
  • [HS05] DP Hardin and EB Saff. Minimal riesz energy point configurations for rectifiable d-dimensional manifolds. Advances in Mathematics, 193(1):174–204, 2005.
  • [J+98] Kurt Johansson et al. On fluctuations of eigenvalues of random hermitian matrices. Duke mathematical journal, 91(1):151–204, 1998.
  • [JM15] Tiefeng Jiang and Sho Matsumoto. Moments of traces of circular beta-ensembles. The Annals of Probability, 43(6):3279–3336, 2015.
  • [KS09] Rowan Killip and Mihai Stoiciu. Eigenvalue statistics for cmv matrices: from poisson to clock via random matrix ensembles. Duke Mathematical Journal, 146(3):361–399, 2009.
  • [Lam21] Gaultier Lambert. Mesoscopic central limit theorem for the circular β\beta-ensembles and applications. Electronic Journal of Probability, 26:1–33, 2021.
  • [Leb17] Thomas Leblé. Local microscopic behavior for 2d coulomb gases. Probability Theory and Related Fields, 169(3):931–976, 2017.
  • [Leb18] Thomas Leblé. CLT for fluctuations of linear statistics in the sine-β\beta process. International Mathematics Research Notices, 2018.
  • [Leb21] Thomas Leblé. The two-dimensional one-component plasma is hyperuniform. arXiv preprint arXiv:2104.05109, 2021.
  • [Lew22] Mathieu Lewin. Coulomb and riesz gases: The known and the unknown. Journal of Mathematical Physics, 63(6):061101, 2022.
  • [LLW+19] Gaultier Lambert, Michel Ledoux, Christian Webb, et al. Quantitative normal approximation of linear statistics of β\beta-ensembles. Annals of Probability, 47(5):2619–2685, 2019.
  • [LS17] Thomas Leblé and Sylvia Serfaty. Large deviation principle for empirical fields of log and riesz gases. Inventiones mathematicae, 210(3):645–757, 2017.
  • [LS18] Thomas Leblé and Sylvia Serfaty. Fluctuations of two dimensional coulomb gases. Geometric and Functional Analysis, 28(2):443–508, 2018.
  • [LZ20] Thomas Leblé and Ofer Zeitouni. A local CLT for linear statistics of 2d coulomb gases. arXiv preprint arXiv:2005.12163, 2020.
  • [MO+16] Jean-Christophe Mourrat, Felix Otto, et al. Correlation structure of the corrector in stochastic homogenization. Annals of Probability, 44(5):3207–3233, 2016.
  • [NS97] Ali Naddaf and Thomas Spencer. On homogenization and scaling limit of some gradient perturbations of a massless free field. Communications in mathematical physics, 183(1):55–84, 1997.
  • [Pei22] Luke Peilen. Local laws and a mesoscopic clt for betabeta-ensembles. arXiv preprint arXiv:2208.14940, 2022.
  • [PS17] Mircea Petrache and Sylvia Serfaty. Next order asymptotics and renormalized energy for riesz interactions. Journal of the Institute of Mathematics of Jussieu, 16(3):501–569, 2017.
  • [RS13] Nicolas Rougerie and Sylvia Serfaty. Higher dimensional coulomb gases and renormalized energy functionals. arXiv: Mathematical Physics, 2013.
  • [RS16] Luz Roncal and Pablo Raúl Stinga. Fractional laplacian on the torus. Communications in Contemporary Mathematics, 18(03):1550033, 2016.
  • [Rue74] David Ruelle. Statistical Mechanics: Rigorous Results. Mathematical physics monograph series. W. A. Benjamin, 1974.
  • [Ser14] Sylvia Serfaty. Coulomb Gases and Ginzburg-Landau Vortices. arXiv e-prints, page arXiv:1403.6860, March 2014.
  • [Ser18] Sylvia Serfaty. Systems of points with coulomb interactions. In Proceedings of the International Congress of Mathematicians: Rio de Janeiro 2018, pages 935–977. World Scientific, 2018.
  • [Ser20] Sylvia Serfaty. Gaussian fluctuations and free energy expansion for 2d and 3d coulomb gases at any temperature. arXiv preprint arXiv:2003.11704, 2020.
  • [Ser24] Sylvia Serfaty. Lectures on coulomb and riesz gases. 2024.
  • [Shc13] Mariya Shcherbina. Fluctuations of linear eigenvalue statistics of β\beta matrix models in the multi-cut regime. Journal of Statistical Physics, 151(6):1004–1034, 2013.
  • [Sjo93a] Johannes Sjostrand. Potential wells in high dimensions i. Annales de l’I.H.P. Physique théorique, 58(1):1–41, 1993.
  • [Sjo93b] Johannes Sjostrand. Potentials wells in high dimensions ii, more about the one well case. Annales de l’I.H.P. Physique théorique, 58, 1993.
  • [SKA+21] Saikat Santra, Jitendra Kethepalli, Sanaa Agarwal, Abhishek Dhar, Manas Kulkarni, and Anupam Kundu. Gap statistics for confined particles with power-law interactions. arXiv preprint arXiv:2109.15026, 2021.
  • [SS12] Etienne Sandier and Sylvia Serfaty. From the Ginzburg-Landau model to vortex lattice problems. Communications in Mathematical Physics, 313(3):635–743, 2012.
  • [Sti19] Pablo Raúl Stinga. User’s guide to the fractional laplacian and the method of semigroups. In Fractional Differential Equations, pages 235–266. De Gruyter, 2019.
  • [SW13] Philippe Sosoe and Percy Wong. Regularity conditions in the CLT for linear eigenvalue statistics of wigner matrices. Advances in Mathematics, 249:37–87, 2013.
  • [Tho21] Eric Thoma. Thermodynamic and scaling limits of the non-gaussian membrane model. arXiv preprint arXiv:2112.07584, 2021.
  • [Tho22] Eric Thoma. Overcrowding and separation estimates for the coulomb gas. arXiv preprint arXiv:2210.05902, 2022.
  • [Tor16] Salvatore Torquato. Hyperuniformity and its generalizations. Physical Review E, 94(2), Aug 2016.
  • [VV09] Benedek Valkó and Bálint Virág. Continuum limits of random matrices and the brownian carousel. Inventiones mathematicae, 177(3):463–508, 2009.