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

Overcrowding and Separation Estimates for the Coulomb Gas

Eric Thoma
(Date: February 20, 2023)
Abstract.

We prove several results for the Coulomb gas in any dimension d2d\geq 2 that follow from isotropic averaging, a transport method based on Newton’s theorem. First, we prove a high-density Jancovici-Lebowitz-Manificat law, extending the microscopic density bounds of Armstrong and Serfaty and establishing strictly sub-Gaussian tails for charge excess in dimension 22. The existence of microscopic limiting point processes is proved at the edge of the droplet. Next, we prove optimal upper bounds on the kk-point correlation function for merging points, including a Wegner estimate for the Coulomb gas for k=1k=1. We prove the tightness of the properly rescaled kkth minimal particle gap, identifying the correct order in d=2d=2 and a three term expansion in d3d\geq 3, as well as upper and lower tail estimates. In particular, we extend the two-dimensional “perfect-freezing regime” identified by Ameur and Romero to higher dimensions. Finally, we give positive charge discrepancy bounds which are state of the art near the droplet boundary and prove incompressibility of Laughlin states in the fractional quantum Hall effect, starting at large microscopic scales. Using rigidity for fluctuations of smooth linear statistics, we show how to upgrade positive discrepancy bounds to estimates on the absolute discrepancy in certain regions.

2020 Mathematics Subject Classification:
82B05, 60G55, 60G70, 49S05

1. Introduction

1.1. The setting.

For d2d\geq 2, the dd-dimensional Coulomb gas (or one-component plasma) at inverse temperature β(0,)\beta\in(0,\infty) is a probability measure on point configurations XN=(x1,,xN)(d)NX_{N}=(x_{1},\ldots,x_{N})\in(\mathbb{R}^{d})^{N} given by

N,βW(dXN)=1𝒵exp(βW(XN))dXN\mathbb{P}^{W}_{N,\beta}(dX_{N})=\frac{1}{\mathcal{Z}}\exp\left(-\beta\mathcal{H}^{W}(X_{N})\right)dX_{N} (1.1)

where dXNdX_{N} is Lebesgue measure on (d)N(\mathbb{R}^{d})^{N}, 𝒵\mathcal{Z} is a normalizing constant, and

W(XN)=121ijN𝗀(xixj)+i=1NW(xi)\mathcal{H}^{W}(X_{N})=\frac{1}{2}\sum_{1\leq i\neq j\leq N}\mathsf{g}(x_{i}-x_{j})+\sum_{i=1}^{N}W(x_{i}) (1.2)

is the Coulomb energy of XNX_{N} with confining potential WW. The kernel 𝗀\mathsf{g} is the Coulomb interaction given by

𝗀(x)={log|x|if d=21|x|d2if d3.\mathsf{g}(x)=\begin{cases}-\log|x|\quad&\text{if }d=2\\ \frac{1}{|x|^{d-2}}\quad&\text{if }d\geq 3.\end{cases} (1.3)

While we will rarely require it, we have in mind the scaling W=VN:=N2/dV(N1/d)W=V_{N}:=N^{2/d}V(N^{-1/d}\cdot) for a potential VV satisfying certain conditions. The normalization in VNV_{N} is chosen so that the typical interstitial distance is of size O(1)O(1), i.e. the Coulomb gas N,βVN\mathbb{P}^{V_{N}}_{N,\beta} is on the “blown-up” scale. However, unless otherwise stated, we will work only under the assumption that ΔW\Delta W exists and is bounded from above on d\mathbb{R}^{d} and (1.1) is well-defined, though see Remark 1.9 for comments on how this can be loosened significantly. For some results, we will need additional assumptions on WW.

Up to normalization, the kernel 𝗀\mathsf{g} gives the repulsive interaction between two positive point charges, and so the Coulomb gas exhibits a competition between particle repulsion, given by the first sum in (1.2), and particle confinement, given by the second sum in (1.2). The behavior of XNX_{N} at the macroscopic scale (i.e. in a box of side length O(N1/d)O(N^{1/d})) is largely dictated by the equilibrium measure μeq\mu_{\mathrm{eq}}, a compactly supported probability measure on d\mathbb{R}^{d} solving a variational problem involving VV, in the sense that the empirical measure N1i=1NδN1/dxiN^{-1}\sum_{i=1}^{N}\delta_{N^{-1/d}x_{i}} is well-approximated weakly by μeq\mu_{\mathrm{eq}} with high probability as NN\to\infty. In particular, the rescaled points condense on the droplet, i.e. the support of μeq\mu_{\mathrm{eq}}. Even on mesoscales O(Nα)O(N^{\alpha}), 0<α<1/d0<\alpha<1/d, the equilibrium measure gives a good approximation for particle density. Letting μeqN\mu_{\mathrm{eq}}^{N} be defined by μeqN(A)=Nμeq(N1dA)\mu_{\mathrm{eq}}^{N}(A)=N\mu_{\mathrm{eq}}(N^{-\frac{1}{d}}A) for measurable AdA\subset\mathbb{R}^{d}, one can form the random fluctuation measure

fluct(dx)=i=1Nδxi(dx)μeqN(dx),\mathrm{fluct}(dx)=\sum_{i=1}^{N}\delta_{x_{i}}(dx)-\mu_{\mathrm{eq}}^{N}(dx), (1.4)

which, despite being of size O(N)O(N) in total variation, is typically of size O(1)O(1) when acting on smooth functions (e.g. [Ser20, LS18, AS21, BBNY19]).

Most of the time, we will work with the more general probability measure N,βW,U\mathbb{P}^{W,U}_{N,\beta} defined by

N,βW,U(dXN)exp(βW,U(XN))dXN,W,U(XN)=W(XN)+U(x1,,xN),\mathbb{P}^{W,U}_{N,\beta}(dX_{N})\propto\exp\left(-\beta\mathcal{H}^{W,U}(X_{N})\right)dX_{N},\quad\mathcal{H}^{W,U}(X_{N})=\mathcal{H}^{W}(X_{N})+U(x_{1},\ldots,x_{N}), (1.5)

where U=UN:(d)NU=U_{N}:(\mathbb{R}^{d})^{N}\to\mathbb{R} is symmetric, superharmonic and locally integrable in each variable xix_{i}, and such that the measure N,βW,U\mathbb{P}^{W,U}_{N,\beta} is well-defined. Measures of this form capture behavior of the gas under conditioning. For example, the Coulomb gas (1.1) disintegrates along xnx_{n} to N1,βW,U\mathbb{P}^{W,U}_{N-1,\beta} with U(XN1)=i=1N1𝗀(xixN)U(X_{N-1})=\sum_{i=1}^{N-1}\mathsf{g}(x_{i}-x_{N}). They also play an important role in the study of the fractional quantum Hall effect; see Section 1.6 for further discussion as well the surveys [Rou22b, Rou22a].

We will apply a transport-type argument, called isotropic averaging, to give upper bounds for N,βW,U\mathbb{P}^{W,U}_{N,\beta} on a variety of events, all concerning the overcrowding of particles. This terminology and a similar method was first used in [Leb21], but a technical issue limited its applicability. Our main contribution is to demonstrate that the method has wide-ranging applicability by giving relatively short and intuitive solutions to several open problems. We believe that it will be an important tool in future studies of the Coulomb gas.

1.2. A model computation.

The basic idea behind isotropic averaging will be motivated through the following model computation. We will refer to this computation, in more general forms, throughout the paper.

We start by defining certain isotropic averaging operators. Given an index set {1,2,,N}\mathcal{I}\subset\{1,2,\ldots,N\} and a rotationally symmetric probability measure ν\nu on d\mathbb{R}^{d}, we define

Iso,νF((xi)i)=(d)F((xi+yi)i)iν(dyi)\mathrm{Iso}_{\mathcal{I},\nu}F((x_{i})_{i\in\mathcal{I}})=\int_{(\mathbb{R}^{d})^{\mathcal{I}}}F\left((x_{i}+y_{i})_{i\in\mathcal{I}}\right)\prod_{i\in\mathcal{I}}\nu(dy_{i})

for any nice enough function F:(d)F:(\mathbb{R}^{d})^{\mathcal{I}}\to\mathbb{R}. We also consider the operator Iso,ν\mathrm{Iso}_{\mathcal{I},\nu} acting on functions of XNX_{N}, or more generally any set of labeled coordinates, by convolution with ν\nu on the coordinates with labels in \mathcal{I}. For example, we have

Iso,νF(XN)=(d)F(XN+(yi𝟏(i))i=1N)iν(dyi)\mathrm{Iso}_{\mathcal{I},\nu}F(X_{N})=\int_{(\mathbb{R}^{d})^{\mathcal{I}}}F\left(X_{N}+(y_{i}\mathbf{1}_{\mathcal{I}}(i))_{i=1}^{N}\right)\prod_{i\in\mathcal{I}}\nu(dy_{i})

by convention, and Iso,νF(x1)=F(x1)\mathrm{Iso}_{\mathcal{I},\nu}F(x_{1})=F(x_{1}) if 11\not\in\mathcal{I}, otherwise Iso,νF(x1)=Fν(x1)\mathrm{Iso}_{\mathcal{I},\nu}F(x_{1})=F\ast\nu(x_{1}).

An important observation is that the kernel 𝗀\mathsf{g} is superharmonic everywhere and harmonic away from the origin, and thus we have the mean value inequality

Iso,ν𝗀(xixj)𝗀(xixj)i,j.\mathrm{Iso}_{\mathcal{I},\nu}\mathsf{g}(x_{i}-x_{j})\leq\mathsf{g}(x_{i}-x_{j})\quad\forall i,j. (1.6)

In our physical context, the isotropic averaging operator replaces each point charge xix_{i}, ii\in\mathcal{I}, by a charge distribution shaped like ν\nu centered at xix_{i}. Newton’s theorem implies that the electric interaction between two disjoint, radial, unit charge distributions is the same as the interaction between two point charges located at the respective centers. More generally, if the charge distributions are not disjoint, then the interaction is more mild than that of the point charge system (this is because 𝗀(r)\mathsf{g}(r) is decreasing in rr and the electric field generated by a uniform spherical charge is 0 in the interior).

Consider an event EE which we wish to show to be unlikely. For definiteness, we let EE be the event “Br(z)B_{r}(z) contains at least 22 particles” for some fixed r1r\ll 1 and zdz\in\mathbb{R}^{d}. By a union bound, we have

N,βW,U(E)i<jN,βW,U(E{i,j})=(N2)N,βW,U(E{1,2}),E{i,j}:={xiBr(z)}{xjBr(z)}.\mathbb{P}^{W,U}_{N,\beta}(E)\leq\sum_{i<j}\mathbb{P}^{W,U}_{N,\beta}(E_{\{i,j\}})=\binom{N}{2}\mathbb{P}^{W,U}_{N,\beta}(E_{\{1,2\}}),\quad E_{\{i,j\}}:=\{x_{i}\in B_{r}(z)\}\cap\{x_{j}\in B_{r}(z)\}. (1.7)

We can bound the likelihood of E{1,2}E_{\{1,2\}} by comparing each XNE{1,2}X_{N}\in E_{\{1,2\}} to the weighted family of configurations generated by replacing x1x_{1} and x2x_{2} by unit charged annuli of inner radius 1/21/2 and outer radius 11. Letting ν\nu be the uniform probability measure supported on the centered annulus Ann[1/2,1](0)d\mathrm{Ann}_{[1/2,1]}(0)\subset\mathbb{R}^{d}, we have by Jensen’s inequality

N,βW,U(E{1,2})\displaystyle\mathbb{P}^{W,U}_{N,\beta}(E_{\{1,2\}}) =1𝒵E{1,2}eβW,U(XN)𝑑XNeβΔ𝒵E{1,2}eβIso{1,2},νW,U(XN)𝑑XN\displaystyle=\frac{1}{\mathcal{Z}}\int_{E_{\{1,2\}}}e^{-\beta\mathcal{H}^{W,U}(X_{N})}dX_{N}\leq\frac{e^{-\beta\Delta}}{\mathcal{Z}}\int_{E_{\{1,2\}}}e^{-\beta\mathrm{Iso}_{\{1,2\},\nu}\mathcal{H}^{W,U}(X_{N})}dX_{N} (1.8)
eβΔ𝒵E{1,2}Iso{1,2},νeβW,U(XN)𝑑XN\displaystyle\leq\frac{e^{-\beta\Delta}}{\mathcal{Z}}\int_{E_{\{1,2\}}}\mathrm{Iso}_{\{1,2\},\nu}e^{-\beta\mathcal{H}^{W,U}(X_{N})}dX_{N}

for

Δ=infXNE{1,2}W,U(XN)Iso{1,2},νW,U(XN).\Delta=\inf_{X_{N}\in E_{\{1,2\}}}\mathcal{H}^{W,U}(X_{N})-\mathrm{Iso}_{\{1,2\},\nu}\mathcal{H}^{W,U}(X_{N}).

We can then consider the L2((d){1,2})L^{2}((\mathbb{R}^{d})^{\{1,2\}})-adjoint of Iso{1,2},ν\mathrm{Iso}_{\{1,2\},\nu}, which we call Iso{1,2},ν\mathrm{Iso}^{\ast}_{\{1,2\},\nu}, to bound

N,βW,U(E{1,2})\displaystyle\mathbb{P}^{W,U}_{N,\beta}(E_{\{1,2\}}) eβΔ𝒵(d)NIso{1,2},ν𝟏E{1,2}(XN)eβW,U(XN)𝑑XN\displaystyle\leq\frac{e^{-\beta\Delta}}{\mathcal{Z}}\int_{(\mathbb{R}^{d})^{N}}\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}(X_{N})e^{-\beta\mathcal{H}^{W,U}(X_{N})}dX_{N} (1.9)
=eβΔ𝔼N,βW,U[Iso{1,2},ν𝟏E{1,2}].\displaystyle=e^{-\beta\Delta}\mathbb{E}^{W,U}_{N,\beta}[\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}].

We call the above calculation, namely (1.8) and (1.9), the model computation. There are now two tasks: (1) give a lower bound for Δ\Delta and (2) give an upper bound for the expectation of Iso{1,2},ν𝟏E{1,2}\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}.

Regarding task (1), we expect Δ\Delta will be large: two particles initially clustered in Br(z)B_{r}(z) are replaced by annular charges of microscopic length scale, which interact mildly. It is a simple calculation to see the pairwise interaction between the charged annuli is bounded by 𝗀(1/2)\mathsf{g}(1/2) (with the abuse of notation 𝗀(x)=𝗀(|x|)\mathsf{g}(x)=\mathsf{g}(|x|)). Regarding the potential term i=1NW(xi)\sum_{i=1}^{N}W(x_{i}) within W,U(XN)\mathcal{H}^{W,U}(X_{N}), it will increase by at most a constant after isotropic averaging since ΔWC\Delta W\leq C. The superharmonic term UU does not increase. Therefore, we have Δ𝗀(2r)C.\Delta\geq\mathsf{g}(2r)-C.

Regarding task (2), since Iso{1,2},ν\mathrm{Iso}_{\{1,2\},\nu} is a convolution by ν2\nu^{\otimes 2}, we have

Iso{1,2},ν𝟏E{1,2}(XN)=Iso{1,2},ν𝟏E{1,2}(XN)νL2𝟏E{1,2}(,,x3,,xN)L1(2)Cr2d.\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}(X_{N})=\mathrm{Iso}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}(X_{N})\leq\|\nu\|^{2}_{L^{\infty}}\|\mathbf{1}_{E_{\{1,2\}}}(\cdot,\cdot,x_{3},\ldots,x_{N})\|_{L^{1}(\mathbb{R}^{2})}\leq Cr^{2d}.

Moreover, we have Iso{1,2},ν𝟏E{1,2}(XN)=0\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}(X_{N})=0 if x1x_{1} or x2x_{2} is not in B1+r(z)B2(z)B_{1+r}(z)\subset B_{2}(z). Thus

𝔼N,βW,U[Iso{1,2},ν𝟏E{1,2}]Cr2dN,βW,U({x1,x2B2(z)}).\mathbb{E}^{W,U}_{N,\beta}[\mathrm{Iso}^{\ast}_{\{1,2\},\nu}\mathbf{1}_{E_{\{1,2\}}}]\leq Cr^{2d}\mathbb{P}^{W,U}_{N,\beta}(\{x_{1},x_{2}\in B_{2}(z)\}).

Assembling the above, starting with (1.7), we find

N,βW,U(E)Ceβ𝗀(2r)r2dN2N,βW,U({x1,x2B2(z)}).\mathbb{P}^{W,U}_{N,\beta}(E)\leq Ce^{-\beta\mathsf{g}(2r)}r^{2d}N^{2}\mathbb{P}^{W,U}_{N,\beta}(\{x_{1},x_{2}\in B_{2}(z)\}). (1.10)

The probability appearing in the RHS will be bounded by CN2CN^{-2} by our microscopic local law Theorem 1, which is proved using a separate isotropic averaging argument, and we see that the probability of EE is bounded by Cr2deβ𝗀(2r)Cr^{2d}e^{-\beta\mathsf{g}(2r)}. This is optimal in d=2d=2, but can be improved to Cr3d2eβ𝗀(2r)Cr^{3d-2}e^{-\beta\mathsf{g}(2r)} in d3d\geq 3 (see Theorem 3). The CN2r2dCN^{-2}r^{2d} bound for the probability of E{1,2}E_{\{1,2\}} comes from the decrease in phase space volume available to x1x_{1} and x2x_{2} from the full macroscopic scale of O(N)O(N) volume per particle to a specific sub-microscopic ball of O(rd)O(r^{d}) volume upon restricting to E{1,2}E_{\{1,2\}}. In d3d\geq 3, the polynomial singularity of 𝗀\mathsf{g} generates additional effective constraints on x1x_{1} and x2x_{2} within Br(z)B_{r}(z).

We remark that our technique exhibits perfect localization and gives quantitative estimates with computable constants. In particular, it is robust to certain types of conditioning and randomization of the ball Br(z)B_{r}(z), as well as allowing to prove disparate phenomena on vastly different scales. It can also be generalized to use operators other than Iso,ν\mathrm{Iso}_{\mathcal{I},\nu}, as in the proof of Theorem 4 where we give both upper and lower bounds on the minimal inter-particle difference. For the lower bound, we must apply our model computation with a “mimicry” operator defined in Proposition 4.3. The method, in particular techniques for estimating Δ\Delta, can be made very precise, as in Theorem 5. Our model computation bears resemblance to the Mermin-Wagner argument from statistical physics [MW66]. It is also similar to an argument of Lieb, which applies only to ground states (β=\beta=\infty) and was generously shared and eventually generalized and published in [NS15, RS16, PS17].

Notation.

We identify N,βW,U\mathbb{P}^{W,U}_{N,\beta} with the law of a point process XX, with the translation between XNX_{N} and XX given by X=i=1NδxiX=\sum_{i=1}^{N}\delta_{x_{i}}. All point processes will be assumed to be simple. We also define the “index” process 𝕏\mathbb{X} given by 𝕏(A)={i:xiA}\mathbb{X}(A)=\{i:x_{i}\in A\} for measurable sets AA. For example, we have E={X(Br(z))2}E=\{X(B_{r}(z))\geq 2\} and E{1,2}={{1,2}𝕏(Br(z))}E_{\{1,2\}}=\{\{1,2\}\subset\mathbb{X}(B_{r}(z))\} for the events EE and E{1,2}E_{\{1,2\}} considered in this subsection.

1.3. JLM laws.

Introduced in [JLM93], Jancovici-Lebowitz-Manificat (JLM) laws give the probability of large charge discrepencies in the Coulomb gas. The authors considered an infinite volume jellium and approximated the probability of an absolute net charge of size much larger than R(d1)/2R^{(d-1)/2} in a ball of radius RR as RR\to\infty. The jellium is a Coulomb gas with a uniform negative background charge, making the whole system net neutral in an appropriate sense. Since the typical net charge in BR(0)B_{R}(0) is expected to be of order R(d1)/2R^{(d-1)/2} (see [MY80]), the JLM laws are moderate to large deviation results and exhibit tail probabilities with very strong decay in the charge excess. The arguments of [JLM93] are based on electrostatic principles and consider several different regimes of the charge discrepancy size.

We are interested in a rigorous proof of the high density versions of the JLM laws. These versions apply when X(BR(z))X(B_{R}(z)) exceeds the expected charge μeq(BR(z))\mu_{\mathrm{eq}}(B_{R}(z)) by a large multiplicative factor CC. They predict that

jell({X(BR(z))Q}){exp(β4Q2logQQ0)if d=2,exp(β4RQ2)if d=3,\mathbb{P}_{\mathrm{jell}}(\{X(B_{R}(z))\geq Q\})\sim\begin{cases}\exp\left(-\frac{\beta}{4}Q^{2}\log\frac{Q}{Q_{0}}\right)\quad\text{if }d=2,\\ \exp\left(-\frac{\beta}{4R}Q^{2}\right)\quad\text{if }d=3,\end{cases} (1.11)

for Q0=|BR(z)|Q_{0}=|B_{R}(z)|. The prediction applies for QRdQ\gg R^{d} as RR\to\infty

Our main results prove the high density JLM law upper bounds in all dimensions in the ultra-high positive charge excess regime. We do so for N,βW,U\mathbb{P}^{W,U}_{N,\beta}, a Coulomb gas with potential confinement and superharmonic perturbation, though the result holds also for the jellium mutatis mutandis. We note that our result does not require RR\to\infty. Indeed, we have found it very useful at R=1R=1 as a local law upper bound valid on all of d\mathbb{R}^{d}, an extension of the microscale local law in [AS21] which is only proved for zz sufficiently far into the interior of the droplet and under other more restrictive assumptions. Note that although we do not obtain a sharp coefficient on Q2Q^{2} in the exponent of the d3d\geq 3 case, it could be improved with additional effort in Proposition 2.1.

Theorem 1 (High Density JLM Law).

For any R1R\geq 1, integer λ100\lambda\geq 100, and integer QQ satisfying

Q{Cλ2R2+Cβ1log(14λ)if d=2,CRd+Cβ1Rd2if d3,Q\geq\begin{cases}\frac{C\lambda^{2}R^{2}+C\beta^{-1}}{\log(\frac{1}{4}\lambda)}\quad&\text{if }d=2,\\ CR^{d}+C\beta^{-1}R^{d-2}\quad&\text{if }d\geq 3,\end{cases} (1.12)

we have

N,βW,U({X(BR(z))Q}){e12βlog(14λ)Q2+C(1+βλ2R2)Qif d=2,e2dβRd+2Q(Q1)if d3,\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{R}(z))\geq Q\})\leq\begin{cases}e^{-\frac{1}{2}\beta\log(\frac{1}{4}\lambda)Q^{2}+C(1+\beta\lambda^{2}R^{2})Q}\quad&\text{if }d=2,\\ e^{-2^{-d}\beta R^{-d+2}Q(Q-1)}\quad&\text{if }d\geq 3,\end{cases} (1.13)

and the result remains true if zz is replaced by x1x_{1}. The constant CC depends only on the dimension and the upper bound for ΔW\Delta W. In particular if d=2d=2 and QCβ,WR2Q\geq C_{\beta,W}R^{2}, we may choose λ=QR2\lambda=\sqrt{\frac{Q}{R^{2}}} to see

N,βW({X(BR(z))Q})eβ4log(QR2)Q2+CβQ2+CQ.\mathbb{P}^{W}_{N,\beta}(\{X(B_{R}(z))\geq Q\})\leq e^{-\frac{\beta}{4}\log\left(\frac{Q}{R^{2}}\right)Q^{2}+C\beta Q^{2}+CQ}.
Remark 1.1.

The physical principles leading to the law (1.11) focus on the change of free energy between an unconstrained Coulomb gas and one constrained to have charge QQ in BR(z)B_{R}(z). For the constrained gas, the most likely particle configurations involve a build up of positive charge on an inner boundary layer of BR(z)B_{R}(z) and a near vacuum outside of BR(z)B_{R}(z) which “screens” the excess charge. Since the negative charge density is bounded (in a jellium by definition and in N,βW,U\mathbb{P}^{W,U}_{N,\beta} by ΔWC\Delta W\leq C), the negative screening region must be extremely thick when QRdQ\gg R^{d}. The self-energy of the negative screening region is the dominant contributor to the (1.11) bounds in [JLM93]. In our proof, we apply an isotropic averaging operator that moves the particles within BR(z)B_{R}(z) to the bulk of the vacuum region, extracting a large average energy change per particle, thus providing a different perspective on the JLM law.

Remark 1.2.

Theorem 1 applies to small β>0\beta>0. In particular, one sees that charge excesses of order TRdTR^{d}, T1T\gg 1, become unlikely as soon as RC1β1/2R\geq C^{-1}\beta^{-1/2}. For this particular estimate type, Theorem 1 therefore improves the minimal effective distance given in [AS21, Theorem 1] in dimensions d=2d=2 and d5d\geq 5 (RCβ1/2(logβ1)1/2R\geq C\beta^{-1/2}(\log\beta^{-1})^{1/2} and RCβ1d21R\geq C\beta^{\frac{1}{d-2}-1}, respectively).

Theorem 1 immediately allows us to generalize [AS21, Corollary 1.1], which established the existence of limiting microscopic point processes for (x1,,xN)(x_{1},\ldots,x_{N}) re-centered around a point zz. Previous to the work of Armstrong and Serfaty, the existence of such a process was only known in d=2d=2 and β=2\beta=2, where it is the Ginibre point process with an explicit correlation kernel. In [AS21], the point zz must be in the droplet ΣN\Sigma_{N} and a mesoscopic distance CN1d+2CN^{\frac{1}{d+2}} distance from the edge of the droplet ΣN\partial\Sigma_{N}. We are able to lift that restriction, and in particular we can take W=VNW=V_{N} and z=zNz=z_{N} near or in ΣN\partial\Sigma_{N}, in which case one would expect a genuinely different limit than the bulk case.

Corollary 1.3.

For any sequence of points zNdz_{N}\in\mathbb{R}^{d}, the law under N,βW,U\mathbb{P}^{W,U}_{N,\beta} of the point process i=1NδxizN\sum_{i=1}^{N}\delta_{x_{i}-z_{N}} converges weakly along subsequences as NN\to\infty to a simple point process.

Proof.

Tightness of the law of the finite dimensional distributions (X(A1),,X(An))(X(A_{1}),\ldots,X(A_{n})) for bounded Borel sets A1,,AnA_{1},\ldots,A_{n} (or for shifted versions of XX) follows from Theorem 1. This implies weak convergence of the laws of the point processes (see [DVJ08, Theorem 11.1.VII]). ∎

Remark 1.4.

Any limit from Corollary 1.3 will also enjoy analogs of Theorem 1, Theorem 2, and Theorem 5.

1.4. Clustering and the kk-point function.

We have already seen in Section 1.2 that isotropic averaging can be applied to the description of the gas below the microscale, and we now state our full results. We are interested in pointwise bounds for the kk-point correlation function ρk\rho_{k}, defined by

A1×A2××Akρk(y1,y2,,yk)𝑑y1𝑑yk=N!(Nk)!N,βW,U(i=1k{xiAi})\int_{A_{1}\times A_{2}\times\cdots\times A_{k}}\rho_{k}(y_{1},y_{2},\ldots,y_{k})dy_{1}\cdots dy_{k}=\frac{N!}{(N-k)!}\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{i=1}^{k}\{x_{i}\in A_{i}\}\right) (1.14)

for measurable sets A1,A2,,AkdA_{1},A_{2},\ldots,A_{k}\subset\mathbb{R}^{d}.

The functions ρk\rho_{k}, and their truncated versions, are objects of intense interest. For example, in the physics literature, they are known to capture the charge screening behavior of the gas and satisfy sum rules and BBGKY equations [GLM80, Mar88]. For d=2d=2, spatial oscillations of ρ1\rho_{1} are expected to occur for large enough β\beta [CSA20, Cif06, CW03]. Starting at β>2\beta>2, the oscillations occur near the edge of the droplet, and as β\beta increases the oscillations penetrate the bulk of the droplet (numerically, it is present at β=200\beta=200) [CSA20]. This phenomenon is part of a debated freezing or crystallization transition in the two-dimensional Coulomb gas [KK16].

Many results on ρk\rho_{k} are known when integrated on the microscale or higher, though these results are often stated in terms of integration of the empirical measure N1XN^{-1}X against test functions. We will not comprehensively review previous results, but only mention that [AS21] proves that B1(z)ρ1(y)𝑑y\int_{B_{1}(z)}\rho_{1}(y)dy is uniformly bounded in NN for zz sufficiently far inside the droplet.

We will be interested in pointwise bounds on ρk(y1,,yk)\rho_{k}(y_{1},\ldots,y_{k}), particularly when some of the yiy_{i} within sub-microscopic distances of each other. One should see ρk0\rho_{k}\to 0 as y1y2y_{1}\to y_{2} due to the repulsion between particles. There are no previously existing rigorous results for pointwise values for general β\beta; even boundedness of ρ1\rho_{1} was until now unproved.

Theorem 2.

We have that

ρ1(y)C\rho_{1}(y)\leq C (1.15)

for some constant CC independent of NN and yy. We also have

ρk(y1,y2,,yk)C1i<jk(1|yiyj|β)if d=2\rho_{k}(y_{1},y_{2},\ldots,y_{k})\leq C\prod_{1\leq i<j\leq k}(1\wedge|y_{i}-y_{j}|^{\beta})\quad\text{if }d=2 (1.16)

and

ρk(y1,y2,,yk)Cexp(β0(y1,,yk))if d3.\rho_{k}(y_{1},y_{2},\ldots,y_{k})\leq C\exp\left(-\beta\mathcal{H}^{0}(y_{1},\ldots,y_{k})\right)\quad\text{if }d\geq 3. (1.17)

The following bound on sub-microscopic particle clusters is easily derived by integrating Theorem 2. We point out the enhanced r2d2r^{2d-2} factor in (1.22), which will be crucial for Theorem 4.

Theorem 3.

Let QQ be a positive integer. There exists a constant CC, dependent only on β\beta, QQ, and supΔW\sup\Delta W, such that for all r>0r>0 we have

N,βW,U({X(Br(z))Q})Cr2Q+β(Q2)if d=2,\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(z))\geq Q\})\leq Cr^{2Q+\beta\binom{Q}{2}}\quad\text{if }d=2, (1.18)

and

N,βW,U({X(Br(z))Q})CrdQeβ2d21rd2(Q2)if d3.\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(z))\geq Q\})\leq Cr^{dQ}e^{-\frac{\beta}{2^{d-2}}\cdot\frac{1}{r^{d-2}}\binom{Q}{2}}\quad\text{if }d\geq 3. (1.19)

We also have for Q2Q\geq 2 and d=2d=2 that

N,βW,U({X(Br(x1))Q})Cr2(Q1)+β(Q2)\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq Q\})\leq Cr^{2(Q-1)+\beta\binom{Q}{2}} (1.20)

and d3d\geq 3 that

N,βW,U({X(Br(x1))Q})Crd(Q1)eβ2d21rd2(Q12)eβ1rd2(Q1).\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq Q\})\leq Cr^{d(Q-1)}e^{-\frac{\beta}{2^{d-2}}\cdot\frac{1}{r^{d-2}}\binom{Q-1}{2}}e^{-\beta\frac{1}{r^{d-2}}(Q-1)}. (1.21)

In the case of Q=2Q=2 and d3d\geq 3 this can be improved to

N,βW,U({X(Br(x1))2})Cr2d2eβrd2.\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq 2\})\leq Cr^{2d-2}e^{-\frac{\beta}{r^{d-2}}}. (1.22)

We remark that the k=1k=1 and Q=1Q=1 cases of Theorem 2 and Theorem 3, respectively, are instances of Wegner estimates. In the context of β\beta-ensembles on the line, Wegner estimates were proved in [BMP21] and for Wigner matrices in [ESY10]. The Q=2Q=2 cases of Theorem 3 are particle repulsion estimates. These estimates, as well as eigenvalue minimal gaps, have been considered for random matrices in many articles, e.g. [NTV17, Tao13, TV11, EKYY12].

Remark 1.5.

We claim that our results in Theorem 2 are essentially optimal and that Theorem 3 is optimal if d=2d=2 or Q=2Q=2. For d,Q3d,Q\geq 3, one can improve Theorem 3 by more carefully integrating Theorem 2, though an optimal, explicit solution for all QQ would be difficult. Our claim is evidenced by the tightness results we prove in Theorem 4 and by computations for merging points with fixed NN.

1.5. Minimal particle gaps.

We will also study the law of the kkth smallest particle gap ηk\eta_{k}, i.e.

ηk(XN)=thekth smallest element of{|xixj|:i,j{1,,N},ij}.\eta_{k}(X_{N})=\text{the}\ k\text{th smallest element of}\ \{|x_{i}-x_{j}|\ :\ i,j\in\{1,\ldots,N\},i\neq j\}. (1.23)

Note that the particle gaps |xixj||x_{i}-x_{j}|, iji\neq j, are almost surely unique under N,βW,U\mathbb{P}^{W,U}_{N,\beta}.

Previously, the order of η1\eta_{1} was investigated dimension two in [Ame18] and [AR22]. The latter article proves that η1(NlogN)1β\eta_{1}\geq(N\log N)^{-\frac{1}{\beta}} with high probability as NN\to\infty for all β>1\beta>1. It is also proved that η1>C1\eta_{1}>C^{-1} with high probability if β=βN\beta=\beta_{N} grows like logN\log N. This suggests that the gas “freezes”, even at the level of the extremal statistic η1\eta_{1}, in this temperature regime.

We remark that Theorem 3 already improves this bound.

Corollary 1.6.

Let d=2d=2. We have

N,βW({η1γN12+β})Cγ2+βγ>0\mathbb{P}^{W}_{N,\beta}(\{\eta_{1}\leq\gamma N^{-\frac{1}{2+\beta}}\})\leq C\gamma^{2+\beta}\quad\forall\gamma>0 (1.24)

for a constant CC independent of NN.

Proof.

This follows from Theorem 3, specifically (1.20) with Q=2Q=2 and r=γN12+βr=\gamma N^{-\frac{1}{2+\beta}}, and a union bound. ∎

Furthermore, an examination of the proof’s dependence on β\beta shows we can let β=βN=c0logN\beta=\beta_{N}=c_{0}\log N and take a constant CC equal to eCβ=NCe^{C\beta}=N^{C} on the RHS of (1.24). Letting γ=N1/(2+β)γ\gamma=N^{1/(2+\beta)}\gamma^{\prime} for a small enough γ>0\gamma^{\prime}>0 in Corollary 1.6 shows that η1\eta_{1} is bounded below by a constant independent of NN with high probability, offering an alternative proof of the freezing regime identified in [Ame18, AR22]. An identical idea using (1.21) in place of (1.20) proves that the gas freezes in dimension d3d\geq 3 in the logN\log N inverse temperature regime as well.

It is natural to wonder whether Corollary 1.6 is sharp in some sense and whether there are versions for ηk\eta_{k} and d3d\geq 3. We will give an affirmative answer to these questions.

For the eigenvalues of random matrices, the law of ηk\eta_{k} has been of great interest. While one dimensional (for the most studied cases), these models are particularly relevant to the d=2d=2 case because the interaction between eigenvalues is also given by 𝗀=log\mathsf{g}=-\log for certain ensembles. In [AB13], the authors prove for the CUE and GUE ensembles that N4/3η1N^{4/3}\eta_{1} is tight and has limiting law with density proportional to x3k1ex3x^{3k-1}e^{-x^{3}}. Note that the interstitial distance is order N1N^{-1} for these ensembles, whereas for us it will be order 11. The proof uses determinantal correlation kernel methods. Recently, the extremal statistics of generalized Hermitian Wigner matrices was studied by Bourgade in [Bou22] with dynamical methods, proving that η1\eta_{1} is also of order N4/3N^{-4/3} and the rescaled limiting law is identical to the GUE case. For a symmetric Wigner matrix, [Bou22] proves that the minimal gap is of order N3/2N^{-3/2}.

The articles [FW21, FTW19] prove even more detailed results on the minimal particle gaps for certain random matrix models. Specifically, for the GOE ensemble and the circular β\beta ensemble with positive integer β\beta, the joint limiting law of the minimal particle gaps is determined. Actually, convergence of a point process containing the data of the minimal particle gaps and their location is proved, showing in particular that the gap locations are asymptotically Poissonian. The methods crucially rely on certain exact identities unavailable in our case.

The only optimal d=2d=2 result is in [SJ12]. Here, the authors consider the kkth smallest eigenvalue gap of certain normal random matrix ensembles. The Ginibre ensemble, after rescaling the eigenvalues by N1/2N^{1/2}, corresponds to N,βVN\mathbb{P}^{V_{N}}_{N,\beta} for a certain quadratic VV and β=2\beta=2. Using determinantal correlation kernel methods, [SJ12] proves that, for the Ginibre ensemble, N14ηkN^{\frac{1}{4}}\eta_{k} is tight as NN\to\infty (with interstitial distance O(1)O(1)), and its limiting law on \mathbb{R} has density proportional to x4k1ex4dxx^{4k-1}e^{-x^{4}}dx. Theorem 4 extends the tightness result to all β\beta and general potential VV.

For our main result, we will take W=VNW=V_{N} and require some extra assumptions on VV. These assumptions are only used to prove Lemma 4.2, which is used in proving lower bounds on ηk\eta_{k}. In particular, our upper bounds on ηk\eta_{k} below hold for N,βW,U\mathbb{P}^{W,U}_{N,\beta} in full generality. For Theorem 4, we require

lim|x|V(x)+𝗀(x)=+anddeβ(V(x)log(1+|x|))𝑑x<if d=2,\lim_{|x|\to\infty}V(x)+\mathsf{g}(x)=+\infty\quad\text{and}\quad\int_{\mathbb{R}^{d}}e^{-\beta(V(x)-\log(1+|x|))}dx<\infty\quad\text{if }d=2, (A1)
ε>0lim inf|x|V(x)|x|ε>0if d3.\exists\varepsilon>0\quad\liminf_{|x|\to\infty}\frac{V(x)}{|x|^{\varepsilon}}>0\quad\text{if }d\geq 3. (A2)
Theorem 4.

Let W=VNW=V_{N} with ΔWC\Delta W\leq C, in d=2d=2 condition (A1) satisfied, and in d3d\geq 3 condition (A2) satisfied. Then, in d=2d=2 the law of N12+βηkN^{\frac{1}{2+\beta}}\eta_{k} is tight as NN\to\infty. Moreover, we have

lim supNN,βVN({N12+βηkγ})Cγk(2+β),\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{N^{\frac{1}{2+\beta}}\eta_{k}\leq\gamma\})\leq C\gamma^{k(2+\beta)},
lim supNN,βVN({N12+βηkγ})Cγ4+2β4+β\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{N^{\frac{1}{2+\beta}}\eta_{k}\geq\gamma\})\leq C\gamma^{-\frac{4+2\beta}{4+\beta}}

for all γ>0\gamma>0.

In d3d\geq 3, let ZkZ_{k} be defined by

ηk=(βlogN)1d2(1+2d2(d2)2loglogNlogN+Zk(d2)logN).\eta_{k}=\left(\frac{\beta}{\log N}\right)^{\frac{1}{d-2}}\left(1+\frac{2d-2}{(d-2)^{2}}\frac{\log\log N}{\log N}+\frac{Z_{k}}{(d-2)\log N}\right).

Then the law of ZkZ_{k} is tight as NN\to\infty. We have

lim supNN,βVN({Zkγ})Cekγ,\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{Z_{k}\leq-\gamma\})\leq Ce^{-k\gamma},
lim supNN,βVN({Zkγ})Ce12γ\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{Z_{k}\geq\gamma\})\leq Ce^{-\frac{1}{2}\gamma}

for γ>0\gamma>0.

Proof.

The theorem follows easily from combining Proposition 4.1 and Proposition 4.4 from Section 4. ∎

Remark 1.7.

The estimates for ηk\eta_{k} in Theorem 4 can be understood by the following Poissonian ansatz. Consider xix_{i}, i=1,,Ni=1,\ldots,N, to be an i.i.d. family with x1x_{1} having law with density proportional to eβ𝗀(x)dxe^{-\beta\mathsf{g}(x)}dx on B1(0)dB_{1}(0)\subset\mathbb{R}^{d}. If we let ηk\eta_{k} be the kkth smallest element of {|xi|:i=1,,N}\{|x_{i}|:i=1,\ldots,N\}, then the order of ηk\eta_{k} agrees with Theorem 4.

The proof of the upper bounds on ηk\eta_{k} in Theorem 4 is an application of the ideas from the model computation in Section 1.2, except with some more precision. The proof of the lower bounds can be understood via an idea related to isotropic averaging that we term “mimicry.” See the beginning of Section 4 for a high-level discussion of the proof.

1.6. Discrepancy and incompressibility bounds.

Our previously mentioned results all concerned either sub-microscopic length scales or high particle densities, but we can in fact effectively use isotropic averaging on mesoscopic length scales and under only slight particle density excesses. We are interested in the discrepancy

Disc(Ω)=X(Ω)Nμeq(N1dΩ),Ωd\mathrm{Disc}(\Omega)=X(\Omega)-N\mu_{\mathrm{eq}}(N^{-\frac{1}{d}}\Omega),\quad\Omega\subset\mathbb{R}^{d} (1.25)

which measures the deficit or excess of particles with reference to the equilibrium measure in a measurable set Ω\Omega. It is also useful to define the compression

DiscW(Ω)=X(Ω)1cdΩΔW(x)𝑑x\mathrm{Disc}_{W}(\Omega)=X(\Omega)-\frac{1}{c_{d}}\int_{\Omega}\Delta W(x)dx (1.26)

where cdc_{d} is such that Δ𝗀=cdδ0\Delta\mathsf{g}=-c_{d}\delta_{0}. Note that Disc\mathrm{Disc} and DiscW\mathrm{Disc}_{W} agree when ΩΣN:=N1dsuppμeq\Omega\subset\Sigma_{N}:=N^{\frac{1}{d}}\mathrm{supp}\ \mu_{\mathrm{eq}} and the equilibrium measure exists.

In [AS21], it is proved that the discrepancy in Ω=BR(z)\Omega=B_{R}(z) is typically not more than O(Rd1)O(R^{d-1}) in size whenever zz is sufficiently far in the interior of the droplet and RR is sufficiently large. The proof involves a multiscale argument inspired by stochastic homogenization [AKM19], as well as a technical screening procedure. More generally, [AS21] gives local bounds on the electric energy, which provide a technical basis for central limit theorems [Ser20] among other things. The upper bound of O(Rd1)O(R^{d-1}) represents that the dominant error contribution within the argument comes from surface terms appearing in the multiscale argument. The JLM laws [JLM93] predict that the discrepancy in BR(z)B_{R}(z) is actually typically of size O(R(d1)/2)O(R^{(d-1)/2}); in particular, the Coulomb gas is hyperuniform. This motivates the search for other methods to prove discrepancy bounds that, with enough refinement, may be able to overcome surface error terms.

A motivation for studying the compression DiscW\mathrm{Disc}_{W} comes from fractional quantum Hall effect (FQHE) physics (see [Rou22b, Rou22a] and references therein for a more comprehensive discussion). In FQHE physics, one considers wave functions ΨF:N\Psi_{F}:\mathbb{C}^{N}\to\mathbb{C} of the form (considering xi2x_{i}\in\mathbb{C}\cong\mathbb{R}^{2})

ΨF(XN)=F(XN)ΨLau(XN),ΨLau(XN)1i<jN(xixj)eBi=1N|xi|2/4\Psi_{F}(X_{N})=F(X_{N})\Psi_{\mathrm{Lau}}(X_{N}),\quad\Psi_{\mathrm{Lau}}(X_{N})\propto\prod_{1\leq i<j\leq N}(x_{i}-x_{j})^{\ell}e^{-B\sum_{i=1}^{N}|x_{i}|^{2}/4} (1.27)

for integers 1\ell\geq 1, symmetric, analytic functions F:NF:\mathbb{C}^{N}\to\mathbb{C}, and magnetic field strengths B>0B>0. All wave functions are assumed to be L2(N)L^{2}(\mathbb{C}^{N}) normalized. This means that |ΨF(XN)|2dXN=N,βW,U(dXN)|\Psi_{F}(X_{N})|^{2}dX_{N}=\mathbb{P}^{W,U}_{N,\beta}(dX_{N}) for special values W=WB,W=W_{B,\ell}, β=\beta=\ell, and

U(XN)=β1log|F(XN)|2.U(X_{N})=-\beta^{-1}\log|F(X_{N})|^{2}.

In particular, the function UU is superharmonic in each variable since FF is analytic. The connection between the Laughlin wave function ΨLau\Psi_{\mathrm{Lau}} and the Coulomb gas is termed the plasma analogy.

When studying the robustness of FQHE under e.g. material impurities, a physically relevant variational problem is to find FF minimizing the functional

E𝒪(N,B,)=infF𝔼N,βW,U[i=1N𝒪(xi)]=infF2𝒪(x)ρ1,F(x)E_{\mathcal{O}}(N,B,\ell)=\inf_{F}\mathbb{E}^{W,U}_{N,\beta}\left[\sum_{i=1}^{N}\mathcal{O}(x_{i})\right]=\inf_{F}\int_{\mathbb{R}^{2}}\mathcal{O}(x)\rho_{1,F}(x) (1.28)

for a certain 𝒪:2\mathcal{O}:\mathbb{R}^{2}\to\mathbb{R} giving a one-body energy associated to material impurities and trapping. Here N,βW,U(dXN)=|ΨF(XN)|2dXN\mathbb{P}^{W,U}_{N,\beta}(dX_{N})=|\Psi_{F}(X_{N})|^{2}dX_{N} as in the plasma analogy and ρ1,F\rho_{1,F} is the 11-point function of N,βW,U\mathbb{P}^{W,U}_{N,\beta}. It is expected that the infimum within E𝒪(N,B,)E_{\mathcal{O}}(N,B,\ell) is approximately achieved as NN\to\infty with FF of the form F(XN)=i=1Nf(xi)F(X_{N})=\prod_{i=1}^{N}f(x_{i}) for an analytic function ff. Such a factorization is important physically as it indicates the presence of uncorrelated “quasi-holes” at the zeros of ff, a remarkably simple system when compared to the those with nontrivial couplings between particles and quasi-holes.

Lieb, Rougerie, and Yngvason, in the main result of [LRY19], proved ρ1,Fcd1ΔW(1+o(1))\rho_{1,F}\leq c_{d}^{-1}\Delta W(1+o(1)) as NN\to\infty when integrated on scales of size N1/4+εN^{1/4+\varepsilon} for ε>0\varepsilon>0. In other words, we have

DiscW(BR(z))o(1)as N\mathrm{Disc}_{W}(B_{R}(z))\leq o(1)\quad\text{as }N\to\infty

for RN1/4+εR\geq N^{1/4+\varepsilon} with high probability. Such a result is called an incompressibility estimate. A consequence is that if 𝒪\mathcal{O} varies on a scale larger than N1/4N^{1/4}, then we have that the “bathtub” energy

E𝒪bt(N,B,)=inf{2𝒪(x)ρ(x):2ρ(x)𝑑x=N, 0ρcd1ΔW}E_{\mathcal{O}}^{\mathrm{bt}}(N,B,\ell)=\inf\left\{\int_{\mathbb{R}^{2}}\mathcal{O}(x)\rho(x)\ :\ \int_{\mathbb{R}^{2}}\rho(x)dx=N,\ 0\leq\rho\leq c_{d}^{-1}\Delta W\right\} (1.29)

is an approximate lower bound for E𝒪(N,B,)E_{\mathcal{O}}(N,B,\ell). The infimum in (1.29) is over ρ\rho that do not necessarily come as a 11-point function for some N,βW,U\mathbb{P}^{W,U}_{N,\beta}. The restriction on the length scale of 𝒪\mathcal{O} does not capture all physically realistic scenarios. To prove that F(XN)=i=1Nf(xi)F(X_{N})=\prod_{i=1}^{N}f(x_{i}) approximately saturates the infimum in E𝒪(N,B,)E_{\mathcal{O}}(N,B,\ell), the remaining task is to show that a set of profiles ρ\rho saturating the infimum in the bathtub energy are well approximated by a 11-point function ρ1,F\rho_{1,F} with FF of the factorized form, a task that was considered in [RY18, OR20].

We present a new method to prove incompressibility down to large microscopic scales, i.e. for R1R\gg 1, and also to give quantitative estimates for o(1)o(1) terms. Our result is also interesting because it gives density upper bounds on balls BR(z)B_{R}(z) with weaker restrictions on the location of zz than in [AS21]. For technical reasons appearing in Proposition 5.3, we will need to approximate the density of μ\mu by a constant in B2R(z)B_{2R}(z), and the resulting error begins to dominate for RN3/(5d)R\geq N^{3/(5d)}. We have therefore chosen to restrict to small enough mesoscales RR. For FQHE applications, one has that μ\mu is a multiple of Lebesgue measure and so this restriction is unnecessary.

Theorem 5 (Incompressibility).

Let R1R\geq 1 and zdz\in\mathbb{R}^{d}. Suppose either ΔW\Delta W is constant in B2R(z)B_{2R}(z) or both of the following: firstly that

ΔWL(B2R(z))CN1/d,\|\nabla\Delta W\|_{L^{\infty}(B_{2R}(z))}\leq CN^{-1/d},

and secondly RN3/(5d)R\leq N^{3/(5d)}. Assume also ΔW(x)C1\Delta W(x)\geq C^{-1} for xB2R(z)x\in B_{2R}(z). Then we have

N,βW,U({DiscW(BR(z))TRd2/3(1+𝟏d=2logR)})ecRdT+ecR(d+2)/3T2+eRd+2\mathbb{P}^{W,U}_{N,\beta}(\{\mathrm{Disc}_{W}(B_{R}(z))\geq TR^{d-2/3}(1+\mathbf{1}_{d=2}\log R)\})\leq e^{-cR^{d}T}+e^{-cR^{(d+2)/3}T^{2}}+e^{-R^{d+2}} (1.30)

for some c>0c>0 for all T1T\geq 1 large enough.

Theorem 5 affirmatively answers an important question posed in [LRY19], demonstrating the remarkable property that the Coulomb gas cannot be significantly compressed beyond density cd1ΔWc_{d}^{-1}\Delta W by any choice of superharmonic perturbation UU. The importance of incompressibility of the Laughlin phase was first raised in [RY15a] and first progress was made in [RY15b]. Further progress and the previously best result appears in [LRY19]. There, the authors transfer β=\beta=\infty incompressibility estimates to the positive temperature system, whereas we work directly with the positive temperature Gibbs measure. We also avoid completely the use of potential theoretic subharmonic quadrature domains (also termed “screening regions”). We expect our result will have significant applications toward proving the stability of the Laughlin phase, which appears in the study of 2D electron gases and rotating Bose gases [RSY14].

Remark 1.8.

Theorem 5 may, at first, seem to be in disagreement with numerical results showing oscillations in ρ1\rho_{1} at the edge of the droplet [CSA20, Cif06, CW03] in d=2d=2, with supρ1\sup\rho_{1} significantly larger than (2π)1ΔW(2\pi)^{-1}\Delta W. The oscillation wavelength appears to be of order of the inter-particle distance R=1R=1, whereas our theorem becomes effective for R1R\gg 1, resolving the apparent disagreement.

While Theorem 5 controls only positive discrepancies, when combined with an estimate on the fluctuations of smooth linear statistics, it can also give lower bounds. Indeed, when BR(z)B_{R}(z) has a large discrepancy, the physically realistic scenario is that positive charge excess builds up either on the inside or outside of BR(z)\partial B_{R}(z). We call a thin annulus with the positive charge buildup a “screening region”, the existence of which is implied by rigidity for smooth linear statistics. We use rigidity from [Ser20], which we note does not take “heavy-lifting”, e.g. it is independent of the multi-scale argument from [AS21]. Proposition 5.6 proves a stronger version of Theorem 5 that applies to screening regions.

We will need some additional assumptions to apply results from [Ser20] and [AS19]. We refer to the introduction of [Ser20] for commentary on the conditions. While the conditions hold in significant generality, our main purpose is to demonstrate the usefulness of Theorem 5 rather than optimize the conditions. Assume that W=VNW=V_{N} where VC7(d)V\in C^{7}(\mathbb{R}^{d}), the droplet Σ=suppμeq\Sigma=\mathrm{supp}\ \mu_{\mathrm{eq}} has C1C^{1} boundary, ΔVC1>0\Delta V\geq C^{-1}>0 in a neighborhood of Σ\Sigma, and supΔVC\sup\Delta V\leq C. Assume further

{deβ2N(V(x)log(1+|x|))𝑑x+deβN(V(x)log(1+|x|))(|x|log(1+|x|))2𝑑x<if d=2,deβ2V(x)𝑑x<if d=3,lim|x|V(x)+𝗀(x)=+.\begin{cases}&\int_{\mathbb{R}^{d}}e^{-\frac{\beta}{2}N(V(x)-\log(1+|x|))}dx+\int_{\mathbb{R}^{d}}e^{-\beta N(V(x)-\log(1+|x|))}(|x|\log(1+|x|))^{2}dx<\infty\quad\text{if }d=2,\\ &\int_{\mathbb{R}^{d}}e^{-\frac{\beta}{2}V(x)}dx<\infty\quad\text{if }d=3,\\ &\lim_{|x|\to\infty}V(x)+\mathsf{g}(x)=+\infty.\end{cases}

Finally, assume there exists a constant KK such that

𝗀μeq(x)+V(x)KC1min(dist(x,Σ)2,1)xd.\mathsf{g}\ast\mu_{\mathrm{eq}}(x)+V(x)-K\geq C^{-1}\min(\mathrm{dist}(x,\Sigma)^{2},1)\quad\forall x\in\mathbb{R}^{d}.
Theorem 6.

Let R[1,N57d]R\in[1,N^{\frac{5}{7d}}] and zdz\in\mathbb{R}^{d} be such that B2R(z){xΣN:dist(x,ΣN)C0N1/(d+2)}B_{2R}(z)\subset\{x\in\Sigma_{N}\ :\ \mathrm{dist}(x,\partial\Sigma_{N})\geq C_{0}N^{1/(d+2)}\} for a large enough constant C0C_{0}. Assume that W=VNW=V_{N} with VV satisfying the above conditions. Then we have

N,βVN({|Disc(BR(z))|TRd4/5(1+𝟏d=2logR)})ecRd10/15T+ecR2/5+d/15T2+ecR(d+2)/5.\mathbb{P}^{V_{N}}_{N,\beta}(\{|\mathrm{Disc}(B_{R}(z))|\geq TR^{d-4/5}(1+\mathbf{1}_{d=2}\log R)\})\leq e^{-cR^{d-10/15}T}+e^{-cR^{2/5+d/15}T^{2}}+e^{-cR^{(d+2)/5}}.

for large enough T1T\gg 1 and some c>0c>0.

We note that by applying isotropic averaging to a screening region, not only do we obtain bounds on the absolute discrepancy, we also give a sharper bound on the positive part than Theorem 5, albeit with some extra restrictions on RR and zz.

1.7. Notation.

We now introduce some notation and conventions used throughout the paper. First, we recall the point process XX and “index” process 𝕏\mathbb{X} introduced at the end of Section 1.1. All point processes will be simple.

Implicit constants CC will change from line to line and may depend on supΔW\sup\Delta W and dd without further comment. In all sections except Section 2, we will also allow CC to depend on continuously β\beta and β1\beta^{-1}. A numbered constant like C0,C1C_{0},C_{1} will be fixed, but may be needed to be taken large depending on various parameters. For positive quantities a,ba,b, we write aba\gg b to mean that aCba\geq Cb for a large enough constant C>0C>0, and aba\ll b for aC1ba\leq C^{-1}b for large enough C>0C>0.

For brevity we will sometimes write \mathbb{P} for N,βW,U\mathbb{P}^{W,U}_{N,\beta} and 𝔼\mathbb{E} for 𝔼N,βW,U\mathbb{E}^{W,U}_{N,\beta}. This will only be done in proofs or sections where the probability measure is fixed throughout. We will write 𝗀(s)\mathsf{g}(s) to mean logs-\log s in dimension 22 or |s|d+2|s|^{-d+2} in d3d\geq 3 when s>0s>0. For a measure with a Lebesgue density, we often denote the density with the same symbol as the measure, e.g. ν(x)\nu(x) as the density of ν(dx)\nu(dx).

When it exists, we let μeq\mu_{\mathrm{eq}} be the equilibrium measure associated to VV, and let Σ=suppμeq\Sigma=\mathrm{supp}\ \mu_{\mathrm{eq}} be the droplet. Note that μeq\mu_{\mathrm{eq}} is a probability measure and Σ\Sigma has length scale 11. We let μeqN\mu_{\mathrm{eq}}^{N} be the blown-up equilibrium measure with μeqN(A)=Nμeq(N1/dA)\mu_{\mathrm{eq}}^{N}(A)=N\mu_{\mathrm{eq}}(N^{-1/d}A) for Borel sets AA and ΣN=suppμeqN=N1/dsuppμeq\Sigma_{N}=\mathrm{supp}\ \mu_{\mathrm{eq}}^{N}=N^{1/d}\mathrm{supp}\mu_{\mathrm{eq}} be the blown-up droplet. Finally, we define

μ(dx)=1cdΔW(x)dx\mu(dx)=\frac{1}{c_{d}}\Delta W(x)dx (1.31)

where cdc_{d} is such that Δ𝗀=cdδ0\Delta\mathsf{g}=-c_{d}\delta_{0}. When we take W=VNW=V_{N}, we have that μ=μeqN\mu=\mu_{\mathrm{eq}}^{N} on the blown-up droplet ΣN\Sigma_{N}.

We let BR(z)dB_{R}(z)\subset\mathbb{R}^{d} be the open Euclidean ball of radius RR centered at zz, and for a nonempty interval I0I\subset\mathbb{R}^{\geq 0}, we let AnnI(z)\mathrm{Ann}_{I}(z) be the annulus containing points xx with |xz|I|x-z|\in I. We use |||\cdot| to denote both Lebesgue measure for subsets of d\mathbb{R}^{d} and cardinality for finite sets. It will be clear via context which is meant.

Remark 1.9.

For most applications, the requirement ΔWC\Delta W\leq C can be loosened to boundedness on the macroscopic scale. This is because the points XNX_{N} are typically confined to a vicinity of the droplet and our arguments are localized to macroscopic neighborhoods of the regions to which they are applied. In d=2d=2, for example, Ameur [Ame21] has proved strong confinement estimates. Furthermore, one can apply our arguments to other Coulomb-type systems like finite volume jelliums provided our arguments do not “run into” domain boundaries. The limiting factor is generally in iterating Proposition 2.2 as in the proof of Theorem 1.

Acknowledgments.

The author would like to thank his advisor, Sylvia Serfaty, for her encouragement and comments. He also thanks Nicolas Rougerie for helpful comments on the incompressibility estimates, especially in connection to the fractional quantum Hall effect. The author was partially supported by NSF grant DMS-2000205.

2. High Density Estimates

In this section, we prove Theorem 1. We will write \mathbb{P} for N,βW,U\mathbb{P}_{N,\beta}^{W,U}, and implicit constants CC will depend only on dd and supΔW\sup\Delta W; they are independent of β\beta.

The idea behind the proof is as follows. We will consider two scales r,Rr,R with 1r=λ1R1\leq r=\lambda^{-1}R for λ10\lambda\geq 10 and two concentric balls Br(z)B_{r}(z) and BR(z)B_{R}(z). In the event that X(Br(z))rdX(B_{r}(z))\gg r^{d}, there is a large pairwise Coulomb energy benefit upon replacing each point charge within Br(z)B_{r}(z) by annuli of scale RR. In particular, this energy benefit dominates any loss from the potential term iW(xi)\sum_{i}W(x_{i}) in W,U\mathcal{H}^{W,U}. We must however consider entropy factors: after applying isotropic averaging, the particles originally confined to Br(z)B_{r}(z) become indistinguishable from the particles within BR(z)B_{R}(z). If X(BR(z))λdX(Br(z))X(B_{R}(z))\leq\lambda^{d}X(B_{r}(z)), i.e. the density of particles in BR(z)B_{R}(z) is not larger than that of Br(z)B_{r}(z), the entropy costs are manageable. If, on the other hand, BR(z)B_{R}(z) has an extremely high particle density, we may iterate our estimate to larger scales RR and λR\lambda R. The iteration terminates once RdNR^{d}\gg N and the considered overcrowding event becomes impossible.

Proposition 2.1 will compute the energy change upon isotropic averaging and estimate the adjoint isotropic averaging operator on the relevant event, which Proposition 2.2 uses as in the model computation to obtain the iteration step. Theorem 1 will then follow shortly. Note that the below computation leading up to (2.3) and (2.4) will be used often in slightly different contexts.

Proposition 2.1.

Consider 0<r<110R0<r<\frac{1}{10}R and νR\nu_{R} the uniform probability measure on the annulus Ann[12R,R2r](0)\mathrm{Ann}_{[\frac{1}{2}R,R-2r]}(0). We have

Iso𝕏(Br(z)),νRW,U(XN)W,U(XN)+CR2X(Br(z))+(X(Br(z))2)(𝗀(R2)𝗀(2r)),\mathrm{Iso}_{\mathbb{X}(B_{r}(z)),\nu_{R}}\mathcal{H}^{W,U}(X_{N})\leq\mathcal{H}^{W,U}(X_{N})+CR^{2}X(B_{r}(z))+\binom{X(B_{r}(z))}{2}\left(\mathsf{g}\left(\frac{R}{2}\right)-\mathsf{g}(2r)\right), (2.1)

and for any index sets 𝒩,{1,,N}\mathcal{N},\mathcal{M}\subset\{1,\ldots,N\} we have

Iso𝒩,νR𝟏{𝕏(Br(z))=𝒩}{𝕏(BR(z))=}eC|𝒩|(rR)d|𝒩|𝟏{𝕏(BR(z))=}.\mathrm{Iso}_{\mathcal{N},\nu_{R}}^{\ast}\mathbf{1}_{\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}}\leq e^{C|\mathcal{N}|}\left(\frac{r}{R}\right)^{d|\mathcal{N}|}\mathbf{1}_{\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}}. (2.2)
Proof.

We first prove (2.1) by considering the effect of the isotropic averaging operator on 𝗀(xixj)\mathsf{g}(x_{i}-x_{j}) and W(xi)W(x_{i}). Let σ\sigma denote (d1)(d-1)-dimensional Hausdorff measure. First, since Δ𝗀=cdδ0-\Delta\mathsf{g}=c_{d}\delta_{0} for cd=σ(B1)(d2+𝟏d=2)c_{d}=\sigma(\partial B_{1})(d-2+\mathbf{1}_{d=2}), we have for any s>0s>0 and ydy\in\mathbb{R}^{d} by Green’s third identity that

1cdBs(y)𝗀(xy)ΔW(x)𝑑x+W(y)\displaystyle\frac{1}{c_{d}}\int_{B_{s}(y)}\mathsf{g}(x-y)\Delta W(x)dx+W(y)
=1cdsBs(y)(𝗀(xy)W(x)(xy)W(x)𝗀(xy)(xy))σ(dx).\displaystyle=\frac{1}{c_{d}s}\int_{\partial B_{s}(y)}\left(\mathsf{g}(x-y)\nabla W(x)\cdot(x-y)-W(x)\nabla\mathsf{g}(x-y)\cdot(x-y)\right)\sigma(dx).

Using the divergence theorem, the RHS can be simplified further to

𝗀(s)cdBs(y)ΔW(x)𝑑x+d2+𝟏d=2cdsd1Bs(y)W(x)σ(dx).\frac{\mathsf{g}(s)}{c_{d}}\int_{B_{s}(y)}\Delta W(x)dx+\frac{d-2+\mathbf{1}_{d=2}}{c_{d}s^{d-1}}\int_{\partial B_{s}(y)}W(x)\sigma(dx).

After a rearrangement, we see

1σ(Bs)Bs(y)W(x)σ(dx)=W(y)+1cdBs(y)(𝗀(xy)𝗀(s))ΔW(x)𝑑x,\frac{1}{\sigma(\partial B_{s})}\int_{\partial B_{s}(y)}W(x)\sigma(dx)=W(y)+\frac{1}{c_{d}}\int_{B_{s}(y)}(\mathsf{g}(x-y)-\mathsf{g}(s))\Delta W(x)dx,

and we can then integrate against σ(Bs)ds\sigma(\partial B_{s})ds with y=x1y=x_{1} to see

Iso{1},νRW(x1)\displaystyle\mathrm{Iso}_{\{1\},\nu_{R}}W(x_{1})
=W(x1)+1cd|Ann[12R,R2r](0)|12RR2rσ(Bs)Bs(x1)(𝗀(xx1)𝗀(s))ΔW(x)𝑑x𝑑s.\displaystyle=W(x_{1})+\frac{1}{c_{d}|\mathrm{Ann}_{[\frac{1}{2}R,R-2r]}(0)|}\int_{\frac{1}{2}R}^{R-2r}\sigma(\partial B_{s})\int_{B_{s}(x_{1})}(\mathsf{g}(x-x_{1})-\mathsf{g}(s))\Delta W(x)dxds.

If we bound ΔWC\Delta W\leq C, one can check by explicit integration that

Iso{1},νRW(x1)W(x1)+CR2.\mathrm{Iso}_{\{1\},\nu_{R}}W(x_{1})\leq W(x_{1})+CR^{2}. (2.3)

A similar computation, this time using Δ𝗀0\Delta\mathsf{g}\leq 0 or superharmonicity of UU in each variable, shows that

Iso𝕏(Br(z)),νR𝗀(xixj)𝗀(xixj),Iso𝕏(Br(z)),νRU(XN)U(XN)i,j{1,,N}.\mathrm{Iso}_{\mathbb{X}(B_{r}(z)),\nu_{R}}\mathsf{g}(x_{i}-x_{j})\leq\mathsf{g}(x_{i}-x_{j}),\quad\mathrm{Iso}_{\mathbb{X}(B_{r}(z)),\nu_{R}}U(X_{N})\leq U(X_{N})\quad\forall i,j\in\{1,\ldots,N\}. (2.4)

Finally, by Newton’s theorem, the Coulomb interaction between a sphere of unit charge and radius ss and a point charge is bounded above by 𝗀(s)\mathsf{g}(s). It follows from superposition that

Iso𝕏(Br(z)),νR𝗀(xixj)𝗀(R2)\mathrm{Iso}_{\mathbb{X}(B_{r}(z)),\nu_{R}}\mathsf{g}(x_{i}-x_{j})\leq\mathsf{g}\left(\frac{R}{2}\right)

whenever i𝕏(Br(z))i\in\mathbb{X}(B_{r}(z)), particularly whenever i,j𝕏(Br(z))i,j\in\mathbb{X}(B_{r}(z)) in which case 𝗀(xixj)𝗀(2r)\mathsf{g}(x_{i}-x_{j})\geq\mathsf{g}(2r). Putting the above results together proves (2.1).

We now consider (2.2). Using that Iso𝒩,νR=Iso𝒩,νR\mathrm{Iso}_{\mathcal{N},\nu_{R}}=\mathrm{Iso}_{\mathcal{N},\nu_{R}}^{\ast} is a convolution on (d)𝒩(\mathbb{R}^{d})^{\mathcal{N}} and Young’s inequality, we have

Iso𝒩,νR𝟏{𝕏(Br(z))=𝒩}{𝕏(BR(z))=}\displaystyle\mathrm{Iso}_{\mathcal{N},\nu_{R}}^{\ast}\mathbf{1}_{\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}} (2.5)
νRL(d)|𝒩|𝟏{𝕏(Br(z))=𝒩}{𝕏(BR(z))=}L1((d)𝒩)eC|𝒩|(rR)d|𝒩|.\displaystyle\leq\|\nu_{R}\|_{L^{\infty}(\mathbb{R}^{d})}^{|\mathcal{N}|}\|\mathbf{1}_{\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}}\|_{L^{1}((\mathbb{R}^{d})^{\mathcal{N}})}\leq e^{C|\mathcal{N}|}\left(\frac{r}{R}\right)^{d|\mathcal{N}|}.

For any configuration XNX_{N} with 𝕏(BR(z))\mathbb{X}(B_{R}(z))\neq\mathcal{M}, we claim that

Iso𝒩,νR𝟏{𝕏(Br(z))=𝒩}{𝕏(BR(z))=}(XN)=0.\mathrm{Iso}_{\mathcal{N},\nu_{R}}^{\ast}\mathbf{1}_{\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}}(X_{N})=0.

First, our claim follows if 𝕏(BR(z))𝒩𝒩\mathbb{X}(B_{R}(z))\setminus\mathcal{N}\neq\mathcal{M}\setminus\mathcal{N} simply because our isotropic averaging operator leaves coordinates with labels in 𝒩c\mathcal{N}^{c} fixed. So we may instead assume there exists i𝒩i\in\mathcal{N} with xiBR(z)x_{i}\not\in B_{R}(z). Then convolution with νR\nu_{R} in the xix_{i} coordinate considers translates xi+yx_{i}+y with |y|R2r|y|\leq R-2r, none of which can be found in Br(z)B_{r}(z). Our claim follows, and together with our pointwise bound (2.5) this establishes (2.2). ∎

We are ready to prove the main iterative estimate that establishes Theorem 1.

Proposition 2.2.

Let 0<r<R0<r<R be such that λ:=Rr10\lambda:=\frac{R}{r}\geq 10. Then we have that

({X(Br(z))Q})({X(BR(z))λdQ})+eC(1+βλ2r2)Qβ(Q2)(𝗀(2r)𝗀(λr/2))\mathbb{P}(\{X(B_{r}(z))\geq Q\})\leq\mathbb{P}(\{X(B_{R}(z))\geq\lambda^{d}Q\})+e^{C(1+\beta\lambda^{2}r^{2})Q-\beta\binom{Q}{2}(\mathsf{g}(2r)-\mathsf{g}(\lambda r/2))}

for all zdz\in\mathbb{R}^{d} and integers QQ with

Q{Cλ2r2+Cβ1log(14λ)if d=2,Cλ2rd+Cβ1rd2if d3.Q\geq\begin{cases}\frac{C\lambda^{2}r^{2}+C\beta^{-1}}{\log(\frac{1}{4}\lambda)}\quad&\text{if }d=2,\\ C\lambda^{2}r^{d}+C\beta^{-1}r^{d-2}\quad&\text{if }d\geq 3.\end{cases} (2.6)
Proof.

For simplicity, we consider λ\lambda an integer. Let 𝒩\mathcal{N}\subset\mathcal{M} be index sets of size nn and mm, respectively. We apply the model computation detailed in Section 1.2 and Proposition 2.1 to see

({𝕏(Br(z))=𝒩}{𝕏(BR(z))=})\displaystyle\mathbb{P}(\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\})
eC(1+βR2)nβ(n2)(𝗀(2r)𝗀(R/2))λdn({𝕏(BR(z))=}).\displaystyle\leq e^{C(1+\beta R^{2})n-\beta\binom{n}{2}(\mathsf{g}(2r)-\mathsf{g}(R/2))}\lambda^{-dn}\mathbb{P}(\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}).

By particle exchangeability, we have

({X(Br(z))=n}{X(BR(z)=m)})=(Nm)(mn)({𝕏(Br(z))=𝒩}{𝕏(BR(z))=}),\mathbb{P}(\{X(B_{r}(z))=n\}\cap\{X(B_{R}(z)=m)\})=\binom{N}{m}\binom{m}{n}\mathbb{P}(\{\mathbb{X}(B_{r}(z))=\mathcal{N}\}\cap\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}),
({X(BR(z)=m)})=(Nm)({𝕏(BR(z))=}),\mathbb{P}(\{X(B_{R}(z)=m)\})=\binom{N}{m}\mathbb{P}(\{\mathbb{X}(B_{R}(z))=\mathcal{M}\}),

whence

({X(Br(z))=n}{X(BR(z)=m)})\displaystyle\mathbb{P}(\{X(B_{r}(z))=n\}\cap\{X(B_{R}(z)=m)\})
eC(1+βR2)nβ(n2)(𝗀(2r)𝗀(R/2))(mn)λdn({X(BR(z))=m}).\displaystyle\leq e^{C(1+\beta R^{2})n-\beta\binom{n}{2}(\mathsf{g}(2r)-\mathsf{g}(R/2))}\binom{m}{n}\lambda^{-dn}\mathbb{P}(\{X(B_{R}(z))=m\}).

By Stirling’s approximation, we can estimate for 1nm1\leq n\leq m that

(mn)2mn(mn)enlogmn+(mn)logmmneCnenlogmn.\binom{m}{n}\leq 2\sqrt{\frac{m}{n(m-n)}}e^{n\log\frac{m}{n}+(m-n)\log\frac{m}{m-n}}\leq e^{Cn}e^{n\log\frac{m}{n}}.

The RHS is bounded by eCnλdne^{Cn}\lambda^{dn} in the case that mλdnm\leq\lambda^{d}n, which we consider. We thus have

({X(Br(z))Q}{X(BR(z))Qλd})\displaystyle\mathbb{P}(\{X(B_{r}(z))\geq Q\}\cap\{X(B_{R}(z))\leq Q\lambda^{d}\})
n=QQλdeC(1+βR2)nβ(n2)(𝗀(2r)𝗀(R/2))m=nQλd({X(BR(z))=m})\displaystyle\leq\sum_{n=Q}^{Q\lambda^{d}}e^{C(1+\beta R^{2})n-\beta\binom{n}{2}(\mathsf{g}(2r)-\mathsf{g}(R/2))}\sum_{m=n}^{Q\lambda^{d}}\mathbb{P}(\{X(B_{R}(z))=m\})
n=QQλdeC(1+βR2)nβ(n2)(𝗀(2r)𝗀(R/2)).\displaystyle\leq\sum_{n=Q}^{Q\lambda^{d}}e^{C(1+\beta R^{2})n-\beta\binom{n}{2}(\mathsf{g}(2r)-\mathsf{g}(R/2))}.

We can bound the ratio between successive terms in the last sum above as

eC(1+βR2)βn(𝗀(2r)𝗀(R/2))12e^{C(1+\beta R^{2})-\beta n(\mathsf{g}(2r)-\mathsf{g}(R/2))}\leq\frac{1}{2}

if Qlog(λ/4)Cβ1+Cr2λ2Q\log(\lambda/4)\geq C\beta^{-1}+Cr^{2}\lambda^{2} in d=2d=2 and if QCβ1rd2+Cλ2rdQ\geq C\beta^{-1}r^{d-2}+C\lambda^{2}r^{d} in d3d\geq 3. We conclude

({X(Br(z))Q}{X(BR(z))Qλd})eC(1+βλ2r2)Qβ(Q2)(𝗀(2r)𝗀(λr/2)),\mathbb{P}(\{X(B_{r}(z))\geq Q\}\cap\{X(B_{R}(z))\leq Q\lambda^{d}\})\leq e^{C(1+\beta\lambda^{2}r^{2})Q-\beta\binom{Q}{2}(\mathsf{g}(2r)-\mathsf{g}(\lambda r/2))},

and the proposition follows. ∎

We conclude the section with a proof of the high density JLM laws.

Proof of Theorem 1.

In d=2d=2, we let λ10\lambda\geq 10 be a free parameter, and in d3d\geq 3 we fix λ\lambda large enough dependent on dd. We apply Proposition 2.2 iteratively to a series of radii rk=λrk1r_{k}=\lambda r_{k-1}, k1k\geq 1, with r0=Rr_{0}=R to achieve

({X(BR(z))Q})lim supk({X(Brk(z))λdkQ})+k=0eak\mathbb{P}(\{X(B_{R}(z))\geq Q\})\leq\limsup_{k\to\infty}\mathbb{P}(\{X(B_{r_{k}}(z))\geq\lambda^{dk}Q\})+\sum_{k=0}^{\infty}e^{a_{k}}

for

ak=C(1+βλ2k+2R2)λdkQβ(λdkQ2)(𝗀(2λkR)𝗀(λk+1R/2)).a_{k}={C(1+\beta\lambda^{2k+2}R^{2})\lambda^{dk}Q-\beta{\binom{\lambda^{dk}Q}{2}}(\mathsf{g}(2\lambda^{k}R)-\mathsf{g}(\lambda^{k+1}R/2))}.

If d=2d=2, we compute

ak+1akβλ4(k+1)log(λ/4)Q24+C(1+βλ2k+4R2)λ2(k+1)Q.a_{k+1}-a_{k}\leq-\beta\frac{\lambda^{4(k+1)}\log(\lambda/4)Q^{2}}{4}+C(1+\beta\lambda^{2k+4}R^{2})\lambda^{2(k+1)}Q.

We can use Qlog(λ/4)Cλ2R2+Cβ1Q\log(\lambda/4)\geq C\lambda^{2}R^{2}+C\beta^{-1} for a large enough CC to see

ak+1akβλ4(k+1)log(λ/4)Q28log2.a_{k+1}-a_{k}\leq-\beta\frac{\lambda^{4(k+1)}\log(\lambda/4)Q^{2}}{8}\leq-\log 2.

In d3d\geq 3, we have

ak+1\displaystyle a_{k+1} C(1+βλ2k+2R2)λdkQβ(λdkQ2)12d2Rd2λ(k+1)(d2)(122d4λd2)\displaystyle\leq C(1+\beta\lambda^{2k+2}R^{2})\lambda^{dk}Q-\beta\binom{\lambda^{dk}Q}{2}\frac{1}{2^{d-2}R^{d-2}\lambda^{(k+1)(d-2)}}\left(1-\frac{2^{2d-4}}{\lambda^{d-2}}\right)
C(1+βλ2k+2R2)λdkQβλdk+2k+2dQ22dRd2.\displaystyle\leq C(1+\beta\lambda^{2k+2}R^{2})\lambda^{dk}Q-\frac{\beta\lambda^{dk+2k+2-d}Q^{2}}{2^{d}R^{d-2}}.

By using QCRd+Cβ1Rd2Q\geq CR^{d}+C\beta^{-1}R^{d-2} (and λ\lambda fixed), we find

ak+1βλ(d+2)kd+2Q22d+1Rd2.a_{k+1}\leq-\beta\frac{\lambda^{(d+2)k-d+2}Q^{2}}{2^{d+1}R^{d-2}}.

One can also compute akβRd+2λ2dk(d+2)kQ2a_{k}\geq-\beta R^{-d+2}\lambda^{2dk-(d+2)k}Q^{2}, which is dominated by ak+1a_{k+1} if λ\lambda is large enough. It follows that ak+1aklog2a_{k+1}-a_{k}\leq-\log 2.

We obtain

({X(BR(z))Q})2ea0eβ(Q2)(𝗀(2R)𝗀(λR/2))+C(1+βλ2R2)Q.\mathbb{P}(\{X(B_{R}(z))\geq Q\})\leq 2e^{a_{0}}\leq e^{-\beta{\binom{Q}{2}}(\mathsf{g}(2R)-\mathsf{g}(\lambda R/2))+C(1+\beta\lambda^{2}R^{2})Q}.

The desired result for balls centered at a fixed zz follows from some routine simplifications.

Finally, it will be useful to have versions of the overcrowding estimates for z=x1z=x_{1}. For this, note that conditioning N,βW,U\mathbb{P}^{W,U}_{N,\beta} on x1x_{1} gives a new measure N1,βW,U+i=2N𝗀(x1xi)\mathbb{P}^{W,U+\sum_{i=2}^{N}\mathsf{g}(x_{1}-x_{i})}_{N-1,\beta} on (x2,,xN)(x_{2},\ldots,x_{N}). We can then apply our results to this (N1)(N-1)-particle Coulomb gas with modified potential. Actually, we could even extract an extra beneficial term in (2.1) from strict superharmonicity of 𝗀(x1)\mathsf{g}(x_{1}-\cdot), but it is mostly inconsequential for the large QQ results. We omit the details. ∎

3. Clustering Estimates

The goal of this section is to prove Theorem 2 and Theorem 3. Our idea is similar to that of the previous section, except we will work with submicroscopic scales and transport particles distances of order 11. We will precisely compute energy and volume gains associated to the transport and control entropy costs using Theorem 1 with R=1R=1. Here, it will be important that we work with measures N,βW,U\mathbb{P}^{W,U}_{N,\beta}, since changing UU will effectively allow us to condition the gas without deteriorating the estimates. In this section, we will allow implicit constants CC to depend continuously on β\beta, β1\beta^{-1}, and supΔW\sup\Delta W.

3.1. kk-point function bounds

We will first bound the kk-point function ρk(y1,,yk)\rho_{k}(y_{1},\ldots,y_{k}), y1,,ykdy_{1},\ldots,y_{k}\in\mathbb{R}^{d}, which was defined in (1.14). In particular, we will prove Theorem 2.

Note that

ρk(y1,,yk)=ρk1(y1,,yk1)ρ1(yk|y1,,yk1)\rho_{k}(y_{1},\ldots,y_{k})=\rho_{k-1}(y_{1},\ldots,y_{k-1})\rho_{1}(y_{k}|y_{1},\ldots,y_{k-1}) (3.1)

where ρ1(|y1,,yk1)\rho_{1}(\cdot|y_{1},\ldots,y_{k-1}) is the 11-point function of the gas Nk+1,βW,Uy1,,yk1\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta} with

Uy1,,yk1(xk,xk+1,,xN)=U(y1,,yk1,xk,,xN)+i=1k1j=kN𝗀(yixj).U_{y_{1},\ldots,y_{k-1}}(x_{k},x_{k+1},\ldots,x_{N})=U(y_{1},\ldots,y_{k-1},x_{k},\ldots,x_{N})+\sum_{i=1}^{k-1}\sum_{j=k}^{N}\mathsf{g}(y_{i}-x_{j}).

We let XN,k=(xk,,xN)X_{N,k}=(x_{k},\ldots,x_{N}) and the representation

ρ1(yk|y1,,yk1)=limr0+Nk+1|B1(0)|rdNk+1,βW,Uy1,,yk1({xkBr(yk)})\rho_{1}(y_{k}|y_{1},\ldots,y_{k-1})=\lim_{r\to 0^{+}}\frac{N-k+1}{|B_{1}(0)|r^{d}}\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}(\{x_{k}\in B_{r}(y_{k})\}) (3.2)

and bound the probability in the RHS using isotropic averaging. Theorem 2 will follow easily from iterating our one-point function bound. Let ν\nu be the uniform probability measure on the annulus Ann[12+2r,12r](0)\mathrm{Ann}_{[\frac{1}{2}+2r,1-2r]}(0). We will replace xkx_{k} by a charge shaped like ν\nu, and the below lemma gives estimates on the energy change and on the isotropic averaging operator.

Lemma 3.1.

With the definitions above and 0<r<11000<r<\frac{1}{100}, we have

infXN,kd(Nk+1)x1Br(yk)W,Uy1,,yk1(XN,k)Iso{k},νW,Uy1,,yk1(XN,k)\displaystyle\inf_{\begin{subarray}{c}X_{N,k}\in\mathbb{R}^{d(N-k+1)}\\ x_{1}\in B_{r}(y_{k})\end{subarray}}\mathcal{H}^{W,U_{y_{1},\ldots,y_{k-1}}}(X_{N,k})-\mathrm{Iso}_{\{k\},\nu}\mathcal{H}^{W,U_{y_{1},\ldots,y_{k-1}}}(X_{N,k}) (3.3)
Ck+i=1k1max(𝗀(|ykyi|+r),0).\displaystyle\quad\geq-Ck+\sum_{i=1}^{k-1}\max(\mathsf{g}(|y_{k}-y_{i}|+r),0).

We also have

Iso{k},ν𝟏Br(yk)(xk)Crd𝟏B1(yk)(xk).\mathrm{Iso}^{\ast}_{\{k\},\nu}\mathbf{1}_{B_{r}(y_{k})}(x_{k})\leq Cr^{d}\mathbf{1}_{B_{1}(y_{k})}(x_{k}). (3.4)
Proof.

We begin by noting that (see Proposition 2.1 for a similar computation)

Iso{k},ν𝗀(xixj)𝗀(xixj),Iso{k},νW(xk)W(xk)+C.\mathrm{Iso}_{\{k\},\nu}\mathsf{g}(x_{i}-x_{j})\leq\mathsf{g}(x_{i}-x_{j}),\quad\mathrm{Iso}_{\{k\},\nu}W(x_{k})\leq W(x_{k})+C.

It follows that

Iso{k},νW,0(XN,k)W,0(XN,k)+C,\mathrm{Iso}_{\{k\},\nu}\mathcal{H}^{W,0}(X_{N,k})\leq\mathcal{H}^{W,0}(X_{N,k})+C,

and it remains to consider the Uy1,,yk1U_{y_{1},\ldots,y_{k-1}} term. For this term, we compute

Iso{k},ν𝗀(yixk)C,Iso{k},ν𝗀(yixk)=𝗀(yixk) if |yiyk|1\mathrm{Iso}_{\{k\},\nu}\mathsf{g}(y_{i}-x_{k})\leq C,\quad\mathrm{Iso}_{\{k\},\nu}\mathsf{g}(y_{i}-x_{k})=\mathsf{g}(y_{i}-x_{k})\text{ if }|y_{i}-y_{k}|\geq 1

using Newton’s theorem. We also have 𝗀(yixk)𝗀(|ykyi|+r)\mathsf{g}(y_{i}-x_{k})\geq\mathsf{g}(|y_{k}-y_{i}|+r) since xkBr(yk)x_{k}\in B_{r}(y_{k}). Thus we have

Iso{k},νUy1,,yk1(XN,k)\displaystyle\mathrm{Iso}_{\{k\},\nu}U_{y_{1},\ldots,y_{k-1}}(X_{N,k}) Uy1,,yk1(XN,k)+i=1k1𝟏|yiyk|1(C𝗀(yixk))\displaystyle\leq U_{y_{1},\ldots,y_{k-1}}(X_{N,k})+\sum_{i=1}^{k-1}\mathbf{1}_{|y_{i}-y_{k}|\leq 1}(C-\mathsf{g}(y_{i}-x_{k}))
Uy1,,yk1(XN,k)+Cki=1k1max(𝗀(|ykyi|+r,0)).\displaystyle\leq U_{y_{1},\ldots,y_{k-1}}(X_{N,k})+Ck-\sum_{i=1}^{k-1}\max(\mathsf{g}(|y_{k}-y_{i}|+r,0)).

This finishes the proof of (3.3). The proof of (3.4) is straightforward using that Iso{k},ν\mathrm{Iso}^{\ast}_{\{k\},\nu} is convolution by ν\nu. ∎

Proposition 3.2.

We have

ρ1(yk|y1,,yk1)eCkβi=1k1max(𝗀(yiyk),0)\rho_{1}(y_{k}|y_{1},\ldots,y_{k-1})\leq e^{Ck-\beta\sum_{i=1}^{k-1}\max(\mathsf{g}(y_{i}-y_{k}),0)} (3.5)

for the one-point function of Nk+1,βW,Uy1,,yk1\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}.

Proof.

An application of the model computation with the operator Iso{k},ν\mathrm{Iso}_{\{k\},\nu} and Lemma 3.1 shows

Nk+1,βW,Uy1,,yk1({xkBr(yk)})rdeCkβi=1k1max(𝗀(yiyk+r),0)Nk+1,βW,Uy1,,yk1({xkB1(yk)})\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}(\{x_{k}\in B_{r}(y_{k})\})\leq r^{d}e^{Ck-\beta\sum_{i=1}^{k-1}\max(\mathsf{g}(y_{i}-y_{k}+r),0)}\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}(\{x_{k}\in B_{1}(y_{k})\}) (3.6)

for all rr small enough. Let Y=i=kNδxiY=\sum_{i=k}^{N}\delta_{x_{i}}. Since

i=kN𝟏{xiB1(yk)}n=1Nk+1n𝟏{Y(B1(yk))=n}\sum_{i=k}^{N}\mathbf{1}_{\{x_{i}\in B_{1}(y_{k})\}}\leq\sum_{n=1}^{N-k+1}n\mathbf{1}_{\{Y(B_{1}(y_{k}))=n\}}

and by Theorem 1 we have

PNk+1,βW,Uy1,,yk1({Y(B1(yk))=n})Cecn2P^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}(\{Y(B_{1}(y_{k}))=n\})\leq Ce^{-cn^{2}}

for some c>0c>0, we find by exchangeability that

Nk+1,βW,Uy1,,yk1({xkB1(yk)})CNk+1n=1necn2CNk+1.\mathbb{P}^{W,U_{y_{1},\ldots,y_{k-1}}}_{N-k+1,\beta}(\{x_{k}\in B_{1}(y_{k})\})\leq\frac{C}{N-k+1}\sum_{n=1}^{\infty}ne^{-cn^{2}}\leq\frac{C}{N-k+1}.

Combining this with (3.6) and taking r0+r\to 0^{+} in (3.2) proves the proposition. ∎

Proof of Theorem 2..

The theorem follows easily by applying Proposition 3.2 iteratively to the representation (3.1) ∎

3.2. Bounds on particle clusters.

We apply our kk-point function estimates to prove the clustering estimates of Theorem 3.

Proof of Theorem 3..

We will need slightly different arguments based on d=2d=2 or d3d\geq 3. We integrate the results of Theorem 2. We have

N,βW,U({X(Br(z))Q})\displaystyle\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(z))\geq Q\}) 1Q!i1,,iQdistinctN,βW,U(=1Q{xiBr(z)})\displaystyle\leq\frac{1}{Q!}\sum_{i_{1},\ldots,i_{Q}\ \text{distinct}}\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{Q}\{x_{i_{\ell}}\in B_{r}(z)\}\right)
=1Q!(Br(z))QρQ(y1,,yQ)𝑑y1𝑑yQ.\displaystyle=\frac{1}{Q!}\int_{(B_{r}(z))^{Q}}\rho_{Q}(y_{1},\ldots,y_{Q})dy_{1}\cdots dy_{Q}.

Thus, for a QQ dependent constant CC, we have

N,βW,U({X(Br(z))Q})CrQdρQL((Br(z))Q).\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(z))\geq Q\})\leq Cr^{Qd}\|\rho_{Q}\|_{L^{\infty}((B_{r}(z))^{Q})}.

In d=2d=2 one estimates

ρQL((Br(z))Q)Crβ(Q2)\|\rho_{Q}\|_{L^{\infty}((B_{r}(z))^{Q})}\leq Cr^{\beta{\binom{Q}{2}}}

which establishes (1.18). In d3d\geq 3, we can estimate

ρQL((Br(z))Q)Ceβ(Q2)infx,yBr(z)𝗀(xy)Ceβ2d2rd2(Q2),\|\rho_{Q}\|_{L^{\infty}((B_{r}(z))^{Q})}\leq Ce^{-\beta{\binom{Q}{2}}\inf_{x,y\in B_{r}(z)}\mathsf{g}(x-y)}\leq Ce^{-\frac{\beta}{2^{d-2}r^{d-2}}{\binom{Q}{2}}},

establishing (1.19). To prove the estimates with z=x1z=x_{1}, we consider (x2,,xN)(x_{2},\ldots,x_{N}) drawn from the Coulomb gas N1,βW,Ux1\mathbb{P}^{W,U_{x_{1}}}_{N-1,\beta} where Ux1(x2,,xN)=U(XN)+i=2N𝗀(xix1)U_{x_{1}}(x_{2},\ldots,x_{N})=U(X_{N})+\sum_{i=2}^{N}\mathsf{g}(x_{i}-x_{1}). The argument of Theorem 2 applies to this gas, except we have an extra term when applying our isotropic averaging operator to Ux1U_{x_{1}} coming from particle repulsion generated by x1x_{1}. It is a straightforward modification to include this term in Lemma 3.1 and the proof of Theorem 2. We can prove

ρQ1(y2,,yQ|y1){C1i<jQmin(1,|yiyj|β)ifd=2,Ceβ0(y1,,yQ)ifd3,\rho_{Q-1}(y_{2},\ldots,y_{Q}|y_{1})\leq\begin{cases}C\prod_{1\leq i<j\leq Q}\min(1,|y_{i}-y_{j}|^{\beta})\quad&\text{if}\ d=2,\\ Ce^{-\beta\mathcal{H}^{0}(y_{1},\ldots,y_{Q})}\quad&\text{if}\ d\geq 3,\end{cases} (3.7)

where ρQ1(y2,,yQ|y1)\rho_{Q-1}(y_{2},\ldots,y_{Q}|y_{1}) is the (Q1)(Q-1)-point function of N1,βW,Ux1\mathbb{P}^{W,U_{x_{1}}}_{N-1,\beta} with x1=y1x_{1}=y_{1}. Then since

N,βW,U({X(Br(x1))Q})Cr(Q1)dsupy1dρQ1(|y1)L((Br(y1))Q1),\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq Q\})\leq Cr^{(Q-1)d}\sup_{y_{1}\in\mathbb{R}^{d}}\|\rho_{Q-1}(\cdot|y_{1})\|_{L^{\infty}((B_{r}(y_{1}))^{Q-1})},

we obtain

N,βW,U({X(Br(x1))Q})Cr(Q1)d+β(Q2)ifd=2,\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq Q\})\leq Cr^{(Q-1)d+\beta{\binom{Q}{2}}}\quad\text{if}\ d=2,

and

N,βW,U({X(Br(x1))Q})Cr(Q1)deβQ1rd2β12d2rd2(Q12)ifd3.\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq Q\})\leq Cr^{(Q-1)d}e^{-\beta\frac{Q-1}{r^{d-2}}-\beta\frac{1}{2^{d-2}r^{d-2}}{\binom{Q-1}{2}}}\quad\text{if}\ d\geq 3. (3.8)

In the latter estimate, we used that |yiy1|r|y_{i}-y_{1}|\leq r in our L(Br(y1)Q1)L^{\infty}(B_{r}(y_{1})^{Q-1}) bound on ρQ1(|y1)\rho_{Q-1}(\cdot|y_{1}).

We expect one can obtain significant improvements to the d3d\geq 3 bound by more accurately estimating the minimum value of 0(y1,,yQ)\mathcal{H}^{0}(y_{1},\ldots,y_{Q}) and the relative volume within (Br(z))Q(B_{r}(z))^{Q} of near-minimizers. We will now do so for the case of Q=2Q=2 and z=x1z=x_{1} since it is relevant to the minimal separation problem.

Let d3d\geq 3. We find

N,βW,U({X(Br(x1))2})Csupy1dBr(y1)eβ|y1y2|d2𝑑y2\displaystyle\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq 2\})\leq C\sup_{y_{1}\in\mathbb{R}^{d}}\int_{B_{r}(y_{1})}e^{-\frac{\beta}{|y_{1}-y_{2}|^{d-2}}}dy_{2} CrdB1(0)eβrd2|y|d2𝑑y\displaystyle\leq Cr^{d}\int_{B_{1}(0)}e^{-\frac{\beta}{r^{d-2}|y|^{d-2}}}dy (3.9)
Crd01sd1eβrd21sd2𝑑s.\displaystyle\leq Cr^{d}\int_{0}^{1}s^{d-1}e^{-\frac{\beta}{r^{d-2}}\frac{1}{s^{d-2}}}ds.

For any α>0\alpha>0, we can compute

01sd1eαsd2𝑑s011sd1eαsd2𝑑s=eαα.\int_{0}^{1}s^{d-1}e^{-\frac{\alpha}{s^{d-2}}}ds\leq\int_{0}^{1}\frac{1}{s^{d-1}}e^{-\frac{\alpha}{s^{d-2}}}ds=\frac{e^{-\alpha}}{\alpha}.

Applying this at α=rd+2β\alpha=r^{-d+2}\beta allows us to conclude

N,βW,U({X(Br(x1))2})Cr2d2eβrd2.\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq 2\})\leq Cr^{2d-2}e^{-\frac{\beta}{r^{d-2}}}.

Notice that this is significantly stronger than (3.8) with Q=2Q=2. The “volume factor” has been reduced from rdr^{d} to r2d2r^{2d-2}. ∎

4. Minimal Separation

In this section, we prove the minimal separation theorem Theorem 4 via Proposition 4.1 and Proposition 4.4. The proof that ηk\eta_{k} is not smaller than expected is a relatively straightforward application of the clustering bounds of Theorem 3. The proof the ηk\eta_{k} is not too large takes a new idea that we term “mimicry”, for which we now give some intuition.

Consider a configuration XNX_{N} with η1>r\eta_{1}>r, let R=O(1)R=O(1) be a large microscopic scale, and let r<Rr<R. Consider two particles, say x1x_{1} and x2x_{2}, with |x1x2|(r,R)|x_{1}-x_{2}|\in(r,R). We can move the particles closer together by “replacing” x2x_{2} with a charged annulus of inner radius r/2r/2 and outer radius rr centered x1x_{1}, and applying our model computation to associate a family of new point configurations to XNX_{N}. A key difference from isotropic averaging is that the annulus is centered at x1x_{1} instead of x2x_{2}.

By Newton’s theorem, the interaction of the annulus with the particles x3,,xNx_{3},\ldots,x_{N} is the same or more mild than that of x1x_{1} and the potential term is less than W(x1)+Cr2W(x_{1})+Cr^{2}. The interaction between x1x_{1} and x2x_{2} has increased, in the worst case, from 𝗀(R)\mathsf{g}(R) to 𝗀(r/2)\mathsf{g}(r/2). The particle x2x_{2} is “mimicking” x1x_{1}. We can also have x1x_{1} mimic x2x_{2}; one of the two mimicking operations is favorable energy-wise up to O(r2)+𝗀(r/2)𝗀(R)O(r^{2})+\mathsf{g}(r/2)-\mathsf{g}(R). There is also an entropy cost associated to mimicry. The particle x2x_{2} originally occupies a O(Rd)O(R^{d}) volume region around x1x_{1} and afterwards is restricted to O(rd)O(r^{d}) volume. We can apply this argument to any of the O(N)O(N) particle pairs within distance RR, creating new configurations from XNX_{N} with either (1) all particles separated by rr except a single pair or (2) with a cluster of three or more particles within a ball of radius rr. Situation (2) can be proved unlikely using our clustering estimates from Theorem 3. Assuming situation (1), we can never create the same configuration by applying a mimicry argument to two distinct index pairs, and so we achieve a volume benefit factor of O(N)O(N). We find that η1>r\eta_{1}>r is unlikely as soon as

NrdRdeβ(𝗀(r/2)𝗀(R))1.\frac{Nr^{d}}{R^{d}}e^{-\beta(\mathsf{g}(r/2)-\mathsf{g}(R))}\gg 1.

In d=2d=2, this happens as soon as rN1/(2+β)r\ll N^{1/(2+\beta)}, matching our desired result Theorem 4. In d3d\geq 3, we must be more careful and consider thinner annuli for mimicry, but a similar intuition holds. This argument is carried out in Proposition 4.4, but before then we must take some care to provide an appropriate parameter RR, which is done in Lemma 4.2 using results from [CHM18]. The main idea is that most of the particles are contained within some volume of size O(N)O(N) with high probability.

Proposition 4.1.

There is an absolute constant C0>0C_{0}>0 such that in d=2d=2 we have

N,βW,U({ηkγN12+β})Cγ(2+β)k+CN2+2β2+βγ4+3β+CN7kγC0(1+β)kγ>0,\mathbb{P}^{W,U}_{N,\beta}(\{\eta_{k}\leq\gamma N^{-\frac{1}{2+\beta}}\})\leq C\gamma^{(2+\beta)k}+CN^{-\frac{2+2\beta}{2+\beta}}\gamma^{4+3\beta}+CN^{-7k}\gamma^{C_{0}(1+\beta)k}\quad\forall\gamma>0, (4.1)

and in d=3d=3 we have

N,βW,U({ηk(βlogN2d2d2loglogN+γ)1d2})CN+Cekγγ>0.\mathbb{P}^{W,U}_{N,\beta}\left(\left\{\eta_{k}\leq\left(\frac{\beta}{\log N-\frac{2d-2}{d-2}\log\log N+\gamma}\right)^{\frac{1}{d-2}}\right\}\right)\leq\frac{C}{\sqrt{N}}+Ce^{-k\gamma}\quad\forall\gamma>0. (4.2)

The constant CC depends on β\beta, supΔW\sup\Delta W, and kk.

Proof.

Let r(0,1)r\in(0,1). Let SS be the event that there exists i{1,,N}i\in\{1,\ldots,N\} such that Br(xi)B_{r}(x_{i}) contains 33 or more particles xjx_{j}. We have

N,βW,U(S)CNN,βW,U({X(Br(x1))3}){CNr4+3βifd=2,CNr2de2βrd2ifd3,\mathbb{P}^{W,U}_{N,\beta}(S)\leq CN\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{r}(x_{1}))\geq 3\})\leq\begin{cases}CNr^{4+3\beta}\quad&\text{if}\ d=2,\\ CNr^{2d}e^{-\frac{2\beta}{r^{d-2}}}\quad&\text{if}\ d\geq 3,\end{cases} (4.3)

by a union bound and Theorem 3.

Let Ek,rE_{k,r} be the event that ηkr\eta_{k}\leq r, and let aj<bja_{j}<b_{j} be random indices such that ηj=|xajxbj|\eta_{j}=|x_{a_{j}}-x_{b_{j}}| for j=1,,kj=1,\ldots,k, which are well-defined almost surely. On the event ScEk,rS^{c}\cap E_{k,r}, we have that {aj,bj}{a,b}=\{a_{j},b_{j}\}\cap\{a_{\ell},b_{\ell}\}=\varnothing for j\ell\neq j almost surely. Let Fj==1j{a=21}{b=2}F_{j}=\bigcap_{\ell=1}^{j}\{a_{\ell}=2\ell-1\}\cap\{b_{\ell}=2\ell\}. By exchangeability, we have

N,βW,U(Ek,rSc)(N2)kN,βW,U(Ek,rFk)(N2)kN,βW,U(=1k{|x2x21|r}).\mathbb{P}^{W,U}_{N,\beta}(E_{k,r}\cap S^{c})\leq{\binom{N}{2}}^{k}\mathbb{P}^{W,U}_{N,\beta}(E_{k,r}\cap F_{k})\leq{\binom{N}{2}}^{k}\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right).

Let Y=Xi=12k2δxiY=X-\sum_{i=1}^{2k-2}\delta_{x_{i}}, i.e. the point process with the first 2k22k-2 points deleted. We estimate

N,βW,U(=1k{|x2x21|r})\displaystyle\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right) (4.4)
=N,βW,U(=1k1{|x2x21|r})N,βW,U({|x2kx2k1|r}|=1k1{|x2x21|r}),\displaystyle=\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k-1}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\mathbb{P}^{W,U}_{N,\beta}\left(\{|x_{2k}-x_{2k-1}|\leq r\}\ \bigg{|}\ \bigcap_{\ell=1}^{k-1}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right),

and furthermore

N,βW,U({|x2kx2k1|r}|=1k1{|x2x21|r})=𝔼[N2k+2,βW,Uk1({|x2kx2k1|r})].\mathbb{P}^{W,U}_{N,\beta}\left(\{|x_{2k}-x_{2k-1}|\leq r\}\ \bigg{|}\ \bigcap_{\ell=1}^{k-1}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)=\mathbb{E}\left[\mathbb{P}^{W,U_{k-1}}_{N-2k+2,\beta}(\{|x_{2k}-x_{2k-1}|\leq r\})\right]. (4.5)

for the random potential Uk1(x2k1,x2k,,xN)=U(XN)+j=2k1N=12k2𝗀(xxj).U_{k-1}(x_{2k-1},x_{2k},\ldots,x_{N})=U(X_{N})+\sum_{j=2k-1}^{N}\sum_{\ell=1}^{2k-2}\mathsf{g}(x_{\ell}-x_{j}). The expectation 𝔼\mathbb{E} is over the law of the points xx_{\ell}, {1,,2k2}\ell\in\{1,\ldots,2k-2\}, conditioned to be pairwise close as above, and N2k+2,βW,Uk1\mathbb{P}^{W,U_{k-1}}_{N-2k+2,\beta} is a measure on the particles (x2k1,x2k,,xN)(x_{2k-1},x_{2k},\ldots,x_{N}).

For any positive integer nn, we have by exchangeability that

N2k+2,βW,Uk1({|x2kx2k1|r})\displaystyle\mathbb{P}^{W,U_{k-1}}_{N-2k+2,\beta}(\{|x_{2k}-x_{2k-1}|\leq r\})
CnNN2k+2,βW,Uk1({Y(Br(x2k))2})+N2k+2,βW,Uk1({Y(Br(x2k))n}).\displaystyle\leq\frac{C_{n}}{N}\mathbb{P}^{W,U_{k-1}}_{N-2k+2,\beta}(\{Y(B_{r}(x_{2k}))\geq 2\})+\mathbb{P}^{W,U_{k-1}}_{N-2k+2,\beta}(\{Y(B_{r}(x_{2k}))\geq n\}).

We apply Theorem 3 to bound each piece above. Collecting the estimates, (4.5), and (4.4), we have proved

N,βW,U(=1k{|x2x21|r})\displaystyle\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)
{C(1Nr2+β+r2(n1)+β(n2))N,βW,U(=1k1{|x2x21|r})if d=2,C(1Nr2d2eβrd2+r(n1)decβn2rd2)N,βW,U(=1k1{|x2x21|r})if d3.\displaystyle\leq\begin{cases}C\left(\frac{1}{N}r^{2+\beta}+r^{2(n-1)+\beta{\binom{n}{2}}}\right)\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k-1}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\quad\text{if }d=2,\\ C\left(\frac{1}{N}r^{2d-2}e^{-\frac{\beta}{r^{d-2}}}+r^{(n-1)d}e^{-\frac{c\beta n^{2}}{r^{d-2}}}\right)\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k-1}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\quad\text{if }d\geq 3.\end{cases}

for a dimensional constant c>0c>0. We can iterate this estimate to see

N,βW,U(=1k{|x2x21|r}){C(1Nr2+β+r2(n1)+β(n2))kifd=2,C(1Nr2d2eβrd2+r(n1)decβn2rd2)kifd3.\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\leq\begin{cases}C\left(\frac{1}{N}r^{2+\beta}+r^{2(n-1)+\beta{\binom{n}{2}}}\right)^{k}\quad&\text{if}\ d=2,\\ C\left(\frac{1}{N}r^{2d-2}e^{-\frac{\beta}{r^{d-2}}}+r^{(n-1)d}e^{-\frac{c\beta n^{2}}{r^{d-2}}}\right)^{k}\quad&\text{if}\ d\geq 3.\end{cases} (4.6)

We conclude the general argument by writing

N,βW,U(Ek,r)N,βW,U(S)+N,βW,U(Ek,rSc)N,βW,U(S)+(N2)kN,βW,U(=1k{|x2x21|r}).\mathbb{P}^{W,U}_{N,\beta}(E_{k,r})\leq\mathbb{P}^{W,U}_{N,\beta}(S)+\mathbb{P}^{W,U}_{N,\beta}(E_{k,r}\cap S^{c})\leq\mathbb{P}^{W,U}_{N,\beta}(S)+\binom{N}{2}^{k}\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right). (4.7)

The probability of SS is bounded in (4.3).

Next, we choose specific rr and nn to conclude the proposition. We must consider d=2d=2 and d3d\geq 3 separately. For d=2d=2, we choose r=γN12+βr=\gamma N^{-\frac{1}{2+\beta}} and n=10n=10 to see

N,βW,U(S)CN2+2β2+βγ4+3β,\displaystyle\mathbb{P}^{W,U}_{N,\beta}(S)\leq CN^{-\frac{2+2\beta}{2+\beta}}\gamma^{4+3\beta},
N,βW,U(=1k{|x2x21|r})CN2kγ(2+β)k+CN9kγC0(1+β)k\displaystyle\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\leq CN^{-2k}\gamma^{(2+\beta)k}+CN^{-9k}\gamma^{C_{0}(1+\beta)k}

for some C0>0C_{0}>0, and (4.1) follows.

For d3d\geq 3, we will assume rd24β3logNr^{d-2}\leq\frac{4\beta}{3\log N}, which with our choice below will happen if NCN\geq C, so that the first summand in the RHS of (4.6) dominates for nn large enough. It follows that

N,βW,U(=1k{|x2x21|r})Cr(2d2)kNkeβkrd2,\mathbb{P}^{W,U}_{N,\beta}\left(\bigcap_{\ell=1}^{k}\{|x_{2\ell}-x_{2\ell-1}|\leq r\}\right)\leq C\frac{r^{(2d-2)k}}{N^{k}}e^{-\frac{\beta k}{r^{d-2}}},

if nn is chosen large enough, and

N,βW,U(S)CNexp(2β(3logN4β))CN12.\mathbb{P}^{W,U}_{N,\beta}(S)\leq CN\exp\left(-2\beta\left(\frac{3\log N}{4\beta}\right)\right)\leq CN^{-\frac{1}{2}}.

We can then let r=(βlogN2d2d2loglogN+γ)1d2r=\left(\frac{\beta}{\log N-\frac{2d-2}{d-2}\log\log N+\gamma}\right)^{\frac{1}{d-2}}, and after a short computation using (4.7), we find (4.2) holds. ∎

Our next goal is to prove that ηk\eta_{k}, properly normalized, is tight as NN\to\infty. To do this, we must create close particle pairs using the model computation, except with a new class of operators, and precisely estimate energy and entropy costs.

For i{1,,N}i\in\{1,\ldots,N\}, let Li,RL_{i,R} be the event that minji|xjxi|R\min_{j\neq i}|x_{j}-x_{i}|\geq R, i.e. that the point xix_{i} is “lonely”. Clearly, if RR is much larger than the interstitial distance, the event Li,RL_{i,R} is rare. We use [CHM18] to quantify this fact. The below result is their Theorem 1.5, written in our blown-up coordinates and corresponding to their inverse temperature chosen as N1+2/dβN^{-1+2/d}\beta in terms of our β\beta.

Theorem 7 ([CHM18], Theorem 1.5).

Let W=VNW=V_{N} with VV satisfying (A1) in d=2d=2 and (A2) in d3d\geq 3. Recall the blown-up equilibrium measure μeqN\mu_{\mathrm{eq}}^{N} from Section 1.7. We have

N,βVN({dBL,N(X,μeqN)NlogNr})ecNlogNr2\mathbb{P}^{V_{N}}_{N,\beta}(\{d_{\mathrm{BL},N}(X,\mu_{\mathrm{eq}}^{N})\geq N\sqrt{\log N}r\})\leq e^{-cN\log Nr^{2}} (4.8)

for any rc1r\geq c^{-1} for some β\beta-dependent c>0c>0, where, for nonnegative measures ν1\nu_{1} and ν2\nu_{2} of mass NN, we define

dBL,N(ν1,ν2):=supfCN0,1(d)fCN0,1(d)1df(x)(ν1ν2)(dx),d_{\mathrm{BL},N}(\nu_{1},\nu_{2}):=\sup_{\begin{subarray}{c}f\in C_{N}^{0,1}(\mathbb{R}^{d})\\ \|f\|_{C_{N}^{0,1}(\mathbb{R}^{d})}\leq 1\end{subarray}}\int_{\mathbb{R}^{d}}f(x)(\nu_{1}-\nu_{2})(dx),

and CN0,1(d)C_{N}^{0,1}(\mathbb{R}^{d}) is the space of bounded Lipschitz functions with norm

fC0,1(d)=supxdN1/d|f(x)|+supxyd|f(x)f(y)||xy|.\|f\|_{C^{0,1}(\mathbb{R}^{d})}=\sup_{x\in\mathbb{R}^{d}}N^{-1/d}|f(x)|+\sup_{x\neq y\in\mathbb{R}^{d}}\frac{|f(x)-f(y)|}{|x-y|}.
Lemma 4.2.

Let W=VNW=V_{N} with VV satisfying (A1) in d=2d=2 and (A2) in d3d\geq 3. We have

N,βVN(L1,R)CRd+CN1dlogN\mathbb{P}^{V_{N}}_{N,\beta}(L_{1,R})\leq CR^{-d}+CN^{-\frac{1}{d}}\sqrt{\log N}

for all R1R\geq 1.

Proof.

We will mostly omit VN,N,βV_{N},N,\beta from the notation.

Defining φ(x)=max(0,1N1/ddist(x,suppμeqN))\varphi(x)=\max(0,1-N^{-1/d}\mathrm{dist}(x,\mathrm{supp}\ \mu_{\mathrm{eq}}^{N})), one can easily check

φ(x)CN0,1(d)CN1/d.\|\varphi(x)\|_{C^{0,1}_{N}(\mathbb{R}^{d})}\leq CN^{-1/d}.

Therefore

|dφ(x)(XμeqN)(dx)|min(CN1ddBL,N(X,μeqN),N),\left|\int_{\mathbb{R}^{d}}\varphi(x)(X-\mu_{\mathrm{eq}}^{N})(dx)\right|\leq\min\left(CN^{-\frac{1}{d}}d_{\mathrm{BL},N}(X,\mu_{\mathrm{eq}}^{N}),N\right),

and so for r=c1r=c^{-1} large enough we have by Theorem 4.8

𝔼[|dφ(x)(XμeqN)(dx)|]\displaystyle\mathbb{E}\left[\left|\int_{\mathbb{R}^{d}}\varphi(x)(X-\mu_{\mathrm{eq}}^{N})(dx)\right|\right] N({dBL,N(X,μeqN)NlogNr})+CN11dlogNr\displaystyle\leq N\mathbb{P}(\{d_{\mathrm{BL},N}(X,\mu_{\mathrm{eq}}^{N})\geq N\sqrt{\log N}r\})+CN^{1-\frac{1}{d}}\sqrt{\log N}r (4.9)
NecNlogNr2+CN11dlogNrCN11dlogN.\displaystyle\leq Ne^{-cN\log Nr^{2}}+CN^{1-\frac{1}{d}}\sqrt{\log N}r\leq CN^{1-\frac{1}{d}}\sqrt{\log N}.

We will use (4.9) to show that most points are typically within suppφ\mathrm{supp}\ \varphi. Indeed, we have

𝔼[dφ(x)(XμeqN)(dx)]𝔼[X((suppφ)c)]=i=1N({xisuppφ}),\mathbb{E}\left[\int_{\mathbb{R}^{d}}\varphi(x)(X-\mu_{\mathrm{eq}}^{N})(dx)\right]\leq\mathbb{E}\left[-X((\mathrm{supp}\ \varphi)^{c})\right]=-\sum_{i=1}^{N}\mathbb{P}(\{x_{i}\not\in\mathrm{supp}\ \varphi\}),

and so ({x1suppφ})CN1dlogN\mathbb{P}(\{x_{1}\not\in\mathrm{supp}\ \varphi\})\leq CN^{-\frac{1}{d}}\sqrt{\log N}.

Let ξi=min(R,minji|xixj|)\xi_{i}=\min(R,\min_{j\neq i}|x_{i}-x_{j}|), so the event Li,RL_{i,R} is equivalent to ξr=R\xi_{r}=R. Since the balls Bξi/2(xi)B_{\xi_{i}/2}(x_{i}) are disjoint, we must have

|{i:xisuppφ,ξi=R}|RdC|{x:dist(x,suppφ)R}|CN,|\{i:x_{i}\in\mathrm{supp}\ \varphi,\xi_{i}=R\}|\cdot R^{d}\leq C|\{x:\mathrm{dist}(x,\mathrm{supp}\ \varphi)\leq R\}|\leq CN,

where we assumed WLOG RN1dR\leq N^{\frac{1}{d}} and used suppμeq\mathrm{supp}\ \mu_{\mathrm{eq}} compact in the last inequality. Thus

(L1,R{x1suppφ})CRd,\mathbb{P}(L_{1,R}\cap\{x_{1}\in\mathrm{supp}\ \varphi\})\leq CR^{-d},

and the lemma follows. ∎

For a configuration XNX_{N}, let ϕ(i)=ϕXN(i){1,,N}\phi(i)=\phi_{X_{N}}(i)\in\{1,\ldots,N\} be the index of the closest particle to ii, i.e. |xixϕ(i)|=minij|xixj||x_{i}-x_{\phi(i)}|=\min_{i\neq j}|x_{i}-x_{j}|. This is almost surely well-defined. Also, let Tk,rT_{k,r} be the event that the cardinality of {{i,j}:|xixj|<r,ij}\{\{i,j\}\ :\ |x_{i}-x_{j}|<r,i\neq j\} is at most kk. Define the random indices a<ba_{\ell}<b_{\ell} such that η=|xaxb|\eta_{\ell}=|x_{a_{\ell}}-x_{b_{\ell}}|. The next proposition uses a mimicry operator to move xix_{i} and xϕ(i)x_{\phi(i)} closer together.

Proposition 4.3.

Suppose U(XN)=i=1NU1(xi)U(X_{N})=\sum_{i=1}^{N}U_{1}(x_{i}) for a superharmonic function U1:dU_{1}:\mathbb{R}^{d}\to\mathbb{R}. Let r(0,1)r\in(0,1) and let ν\nu be a rotationally symmetric probability measure supported in Br(0)¯\overline{B_{r}(0)} with a Lebesgue density, also denoted ν\nu. Abbreviate =N,βW,U\mathbb{P}=\mathbb{P}^{W,U}_{N,\beta}. For any R1R\geq 1, integer n3n\geq 3, and k{1,,N}k\in\{1,\ldots,N\}, we have that ({ηkr})\mathbb{P}(\{\eta_{k}\geq r\}) is bounded by

CeCβr2+βΔννL(d)Mr,R(k+nN+({X(Br(x1))3})+N({X(Br(x1))n}))\displaystyle{Ce^{C\beta r^{2}+\beta\Delta_{\nu}}}\|\nu\|_{L^{\infty}(\mathbb{R}^{d})}M_{r,R}\left(\frac{k+n}{N}+\mathbb{P}(\{X(B_{r}(x_{1}))\geq 3\})+N\mathbb{P}(\{X(B_{r}(x_{1}))\geq n\})\right)
+2k2N+(L1,R),\displaystyle\quad+\frac{2k-2}{N}+\mathbb{P}(L_{1,R}),

where Δν=d𝗀(y)ν(dy)\Delta_{\nu}=\int_{\mathbb{R}^{d}}\mathsf{g}(y)\nu(dy) and Mr,R=Ann[r,R](0)eβ𝗀(y)𝑑yM_{r,R}=\int_{\mathrm{Ann}_{[r,R]}(0)}e^{-\beta\mathsf{g}(y)}dy. The constant CC depends on supΔW\sup\Delta W and nn.

Proof.

We will abbreviate =N,βW,U\mathbb{P}=\mathbb{P}^{W,U}_{N,\beta} and 𝔼=𝔼N,βW,U\mathbb{E}=\mathbb{E}^{W,U}_{N,\beta} throughout the proof.

First note that Tk1,r={ηkr}T_{k-1,r}=\{\eta_{k}\geq r\}. Furthermore, on Tk1,rT_{k-1,r} we have |x1xϕ(1)|r|x_{1}-x_{\phi(1)}|\geq r unless a=1a_{\ell}=1 or b=1b_{\ell}=1 for some {1,,k1}\ell\in\{1,\ldots,k-1\}. We will furthermore want to fix the label ϕ(1)\phi(1) and ensure x1x_{1} is not RR-lonely, which inspires the bound

(Tk1,r)\displaystyle\mathbb{P}(T_{k-1,r}) (N1)({|x1x2|r}{ϕ(1)=2}L1,RcTk1,r)+(L1,R)\displaystyle\leq(N-1)\mathbb{P}(\{|x_{1}-x_{2}|\geq r\}\cap\{\phi(1)=2\}\cap L_{1,R}^{c}\cap T_{k-1,r})+\mathbb{P}(L_{1,R}) (4.10)
+({{1,,k1}a=1orb=1}).\displaystyle\quad+\mathbb{P}(\{\exists\ell\in\{1,\ldots,k-1\}\ a_{\ell}=1\ \text{or}\ b_{\ell}=1\}).

The N1N-1 factor comes from the fact that ϕ(1)\phi(1) is equally likely to be each of {2,,N}\{2,\ldots,N\}. The first probability on the RHS of (4.10) is suitable to apply a transport procedure to move x2x_{2} closer to x1x_{1}, but first we bound the last term. By exchangeability, we have

({{1,,k1}a=1orb=1})2k2N\mathbb{P}(\{\exists\ell\in\{1,\ldots,k-1\}\ a_{\ell}=1\ \text{or}\ b_{\ell}=1\})\leq\frac{2k-2}{N} (4.11)

since {a:=1,,k1}{a:=1,,k1}\{a_{\ell}:\ell=1,\ldots,k-1\}\cup\{a_{\ell}:\ell=1,\ldots,k-1\} is random subset of {1,,N}\{1,\ldots,N\} size at most 2k22k-2.

We condition on XN,3=(x3,x4,,xN)X_{N,3}=(x_{3},x_{4},\ldots,x_{N}) to rewrite

({|x1x2|r}{ϕ(1)=2}L1,RcTk1,r)\displaystyle\mathbb{P}(\{|x_{1}-x_{2}|\geq r\}\cap\{\phi(1)=2\}\cap L_{1,R}^{c}\cap T_{k-1,r}) (4.12)
(Tk1,r{|x1x2|[r,R]}{ϕ(1)=2})\displaystyle\leq\mathbb{P}(T^{\prime}_{k-1,r}\cap\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\})
=𝔼[𝟏Tk1,r({|x1x2|[r,R]}{ϕ(1)=2}|XN,3)],\displaystyle=\mathbb{E}\left[\mathbf{1}_{T^{\prime}_{k-1,r}}\mathbb{P}(\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}\ |\ X_{N,3})\right],

where Tk1,rT^{\prime}_{k-1,r} is the event that the cardinality of {{i,j}{3,,N}:|xixj|<r,ij}\{\{i,j\}\subset\{3,\ldots,N\}\ :\ |x_{i}-x_{j}|<r,i\neq j\} is at most k1k-1.

We next define a new type of isotropic averaging operator to apply to the conditional probability above. For a rotationally symmetric probability measure ν\nu on d\mathbb{R}^{d}, define the mimicry operators

Mim1,2,νF(x1,x2)\displaystyle\mathrm{Mim}_{1,2,\nu}F(x_{1},x_{2}) =dF(x1,x1+y)ν(dy),\displaystyle=\int_{\mathbb{R}^{d}}F(x_{1},x_{1}+y)\nu(dy),
Mim2,1,νF(x1,x2)\displaystyle\mathrm{Mim}_{2,1,\nu}F(x_{1},x_{2}) =dF(x2+y,x2)ν(dy).\displaystyle=\int_{\mathbb{R}^{d}}F(x_{2}+y,x_{2})\nu(dy).

Note that

Mim1,2,νj=3N𝗀(x2xj)j=3N𝗀(x1xj)\mathrm{Mim}_{1,2,\nu}\sum_{j=3}^{N}\mathsf{g}(x_{2}-x_{j})\leq\sum_{j=3}^{N}\mathsf{g}(x_{1}-x_{j})

and

Mim1,2,νU1(x2)U1(x1),Mim1,2,νW(x2)W(x1)+Cr2\mathrm{Mim}_{1,2,\nu}U_{1}(x_{2})\leq U_{1}(x_{1}),\quad\mathrm{Mim}_{1,2,\nu}W(x_{2})\leq W(x_{1})+Cr^{2}

since U1U_{1} is superharmonic and suppνBr(0)¯\mathrm{supp}\ \nu\subset\overline{B_{r}(0)}. Define

Δν=Mim1,2,ν𝗀(x1x2)=Mim2,1,ν𝗀(x1x2)\Delta_{\nu}=\mathrm{Mim}_{1,2,\nu}\mathsf{g}(x_{1}-x_{2})=\mathrm{Mim}_{2,1,\nu}\mathsf{g}(x_{1}-x_{2})

which depends only on ν\nu. Analogous results as above hold for Mim2,1,ν\mathrm{Mim}_{2,1,\nu}. It follows that

eβMim1,2,νW,U(XN)+eβMim2,1,νW,U(XN)eCβr2eβ(𝗀(x1x2)Δν)eβW,U(XN).e^{-\beta\mathrm{Mim}_{1,2,\nu}\mathcal{H}^{W,U}(X_{N})}+e^{-\beta\mathrm{Mim}_{2,1,\nu}\mathcal{H}^{W,U}(X_{N})}\geq e^{-C\beta r^{2}}e^{\beta(\mathsf{g}(x_{1}-x_{2})-\Delta_{\nu})}e^{-\beta\mathcal{H}^{W,U}(X_{N})}.

Indeed, the first summand on the LHS dominates the RHS in the event that

j=3N𝗀(x1xj)+U1(x1)+W(x1)j=3N𝗀(x2xj)+U1(x2)+W(x2)\sum_{j=3}^{N}\mathsf{g}(x_{1}-x_{j})+U_{1}(x_{1})+W(x_{1})\leq\sum_{j=3}^{N}\mathsf{g}(x_{2}-x_{j})+U_{1}(x_{2})+W(x_{2})

since x2x_{2} interacts with x3,,xNx_{3},\ldots,x_{N} similar to how x1x_{1} does after applying the operator Mim1,2,ν\mathrm{Mim}_{1,2,\nu}. When the reverse inequality is true, the second summand on the LHS is dominating. The adjoints are easily computed as

Mim1,2,νF(x1,x2)\displaystyle\mathrm{Mim}_{1,2,\nu}^{\ast}F(x_{1},x_{2}) =ν(x2x1)dF(x1,y)𝑑y,\displaystyle=\nu(x_{2}-x_{1})\int_{\mathbb{R}^{d}}F(x_{1},y)dy,
Mim1,2,νF(x1,x2)\displaystyle\mathrm{Mim}_{1,2,\nu}^{\ast}F(x_{1},x_{2}) =ν(x1x2)dF(y,x2)𝑑y.\displaystyle=\nu(x_{1}-x_{2})\int_{\mathbb{R}^{d}}F(y,x_{2})dy.

We will now apply the model computation to (|XN,3)\mathbb{P}(\cdot\ |\ X_{N,3}). For a normalizing factor 𝒵(XN,3)\mathcal{Z}(X_{N,3}), we have a.s.

({|x1x2|[r,R]}{ϕ(1)=2}|XN,3)\displaystyle\mathbb{P}(\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}\ |\ X_{N,3})
=1𝒵(XN,3)d×d𝟏{|x1x2|[r,R]}{ϕ(1)=2}(x1,x2)eβW,U(XN)𝑑x1𝑑x2\displaystyle=\frac{1}{\mathcal{Z}(X_{N,3})}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathbf{1}_{\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}}(x_{1},x_{2})e^{-\beta\mathcal{H}^{W,U}(X_{N})}dx_{1}dx_{2}
eCβr2+βΔν𝒵(XN,3)(A1+A2),\displaystyle\leq\frac{e^{C\beta r^{2}+\beta\Delta_{\nu}}}{\mathcal{Z}(X_{N,3})}(A_{1}+A_{2}),

where (with a slight abuse of notation)

A1=d×dMim1,2,ν(eβ𝗀(x1x2)𝟏{|x1x2|[r,R]}{ϕ(1)=2})(x1,x2)eβW,U(XN)𝑑x1𝑑x2A_{1}=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathrm{Mim}_{1,2,\nu}^{\ast}\left(e^{-\beta\mathsf{g}(x_{1}-x_{2})}\mathbf{1}_{\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}}\right)(x_{1},x_{2})e^{-\beta\mathcal{H}^{W,U}(X_{N})}dx_{1}dx_{2}

and A2A_{2} identical but with Mim2,1,ν\mathrm{Mim}_{2,1,\nu}^{\ast} in the place of Mim1,2,ν\mathrm{Mim}_{1,2,\nu}^{\ast}. Note that Mim1,2,ν\mathrm{Mim}_{1,2,\nu}^{\ast} is monotonic and

𝟏{|x1x2|[r,R]}{ϕ(1)=2}𝟏{|x1x2|[r,R]},\mathbf{1}_{\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}}\leq\mathbf{1}_{\{|x_{1}-x_{2}|\in[r,R]\}},

whence

Mim1,2,ν(eβ𝗀(x1x2)𝟏{|x1x2|[r,R]}{ϕ(1)=2})(x1,x2)ν(x2x1)Ann[r,R](x1)eβ𝗀(x1y)𝑑y.\mathrm{Mim}_{1,2,\nu}^{\ast}\left(e^{-\beta\mathsf{g}(x_{1}-x_{2})}\mathbf{1}_{\{|x_{1}-x_{2}|\in[r,R]\}\cap\{\phi(1)=2\}}\right)(x_{1},x_{2})\leq\nu(x_{2}-x_{1})\int_{\mathrm{Ann}_{[r,R]}(x_{1})}e^{-\beta\mathsf{g}(x_{1}-y)}dy.

Define Mr,R=Ann[r,R](0)eβ𝗀(y)𝑑yM_{r,R}=\int_{\mathrm{Ann}_{[r,R]}(0)}e^{-\beta\mathsf{g}(y)}dy. Since ν\nu is supported in Br(0)¯\overline{B_{r}(0)}, the RHS above is a.s. bounded by

νL(d)Mr,R𝟏Br(0)(x1x2),\|\nu\|_{L^{\infty}(\mathbb{R}^{d})}M_{r,R}\mathbf{1}_{B_{r}(0)}(x_{1}-x_{2}),

and we find

A1𝒵(XN,3)Mr,RνL(d)({|x1x2|<r}|XN,3).\frac{A_{1}}{\mathcal{Z}(X_{N,3})}\leq M_{r,R}\|\nu\|_{L^{\infty}(\mathbb{R}^{d})}\mathbb{P}(\{|x_{1}-x_{2}|<r\}\ |\ X_{N,3}).

We can prove an identical bound for A2A_{2} with the same argument, and using the bounds in (4.12) shows

({|x1x2|r}{ϕ(1)=2}L1,RcTk1,r)κ({|x1x2|<r}Tk1,r).\mathbb{P}(\{|x_{1}-x_{2}|\geq r\}\cap\{\phi(1)=2\}\cap L_{1,R}^{c}\cap T_{k-1,r})\leq\kappa\mathbb{P}(\{|x_{1}-x_{2}|<r\}\cap{T^{\prime}_{k-1,r}}). (4.13)

where

κ:=2eCβr2+βΔνMr,RνL(d).\kappa:=2e^{C\beta r^{2}+\beta\Delta_{\nu}}M_{r,R}\|\nu\|_{L^{\infty}(\mathbb{R}^{d})}.

We wish to partially recover the information ϕ(1)=2\phi(1)=2 after the transport, i.e. on the RHS of (4.13). To do so, we use the bound

({|x1x2|<r}{ϕ(1)2})CnN({X(Br(x1))3})+({X(Br(x1))n}),\mathbb{P}(\{|x_{1}-x_{2}|<r\}\cap\{\phi(1)\neq 2\})\leq\frac{C_{n}}{N}\mathbb{P}(\{X(B_{r}(x_{1}))\geq 3\})+\mathbb{P}(\{X(B_{r}(x_{1}))\geq n\}),

finding (starting from (4.13))

({|x1x2|r}{ϕ(1)=2}L1,RcTk1,r)\displaystyle\mathbb{P}(\{|x_{1}-x_{2}|\geq r\}\cap\{\phi(1)=2\}\cap L_{1,R}^{c}\cap T_{k-1,r}) (4.14)
κ(({|x1x2|<r}{ϕ(1)=2}Tk1,r)\displaystyle\leq\kappa\bigg{(}\mathbb{P}(\{|x_{1}-x_{2}|<r\}\cap\{\phi(1)=2\}\cap{T^{\prime}_{k-1,r}})
+CnN({X(Br(x1))3})+({X(Br(x1))n})).\displaystyle\quad+\frac{C_{n}}{N}\mathbb{P}(\{X(B_{r}(x_{1}))\geq 3\})+\mathbb{P}(\{X(B_{r}(x_{1}))\geq n\})\bigg{)}.

Let Tk1,r(j)T^{\prime}_{k-1,r}(j) be the event that {xi:i{2,,N},ij}\{x_{i}:i\in\{2,\ldots,N\},i\neq j\} contains at most k1k-1 pairs of points within distance rr. For example, Tk1,r(2)=Tk1,rT^{\prime}_{k-1,r}(2)=T^{\prime}_{k-1,r}. Since the events {ϕ(i)=j}\{\phi(i)=j\}, j=2,,Nj=2,\ldots,N are disjoint up to probability 0, we see

({|x1x2|<r}{ϕ(1)=2}Tk1,r)\displaystyle\mathbb{P}(\{|x_{1}-x_{2}|<r\}\cap\{\phi(1)=2\}\cap{T^{\prime}_{k-1,r}}) (4.15)
=1N1j=2N({|x1xj|<r}{ϕ(1)=j}Tk1,r(j))\displaystyle=\frac{1}{N-1}\sum_{j=2}^{N}\mathbb{P}(\{|x_{1}-x_{j}|<r\}\cap\{\phi(1)=j\}\cap{T^{\prime}_{k-1,r}(j)})
1N1({|x1xϕ(1)|<r}Tk1,r(ϕ(1))}).\displaystyle\leq\frac{1}{N-1}\mathbb{P}(\{|x_{1}-x_{\phi(1)}|<r\}\cap T^{\prime}_{k-1,r}(\phi(1))\}).

On the event {|x1xϕ(1)|<r}Tk1,r(ϕ(1))}\{|x_{1}-x_{\phi(1)}|<r\}\cap T^{\prime}_{k-1,r}(\phi(1))\}, the index 11 is an exceptional index since there likely are very few particles within distance rr of each other, so we expect the probability of the event is of order O(1/N)O(1/N). To be precise, if X(Br(xi))nX(B_{r}(x_{i}))\leq n for all i{1,,N}i\in\{1,\ldots,N\}, then Tk1,j(ϕ(1))T^{\prime}_{k-1,j}(\phi(1)) occurring implies Tk1+2n,rT_{k-1+2n,r}. Thus

({|x1xϕ(1)|<r}Tk1,j(ϕ(1)))\displaystyle\mathbb{P}({\{|x_{1}-x_{\phi(1)}|<r\}}\cap{T^{\prime}_{k-1,j}(\phi(1))}) (4.16)
({iX(Br(xi))n})+({|x1xϕ(1)|<r}Tk1+2n,r)\displaystyle\leq\mathbb{P}(\{\exists i\ X(B_{r}(x_{i}))\geq n\})+\mathbb{P}({\{|x_{1}-x_{\phi(1)}|<r\}}\cap T_{k-1+2n,r})
N({X(Br(x1))n})+({|x1xϕ(1)|<r}Tk1+2n,r).\displaystyle\leq N\mathbb{P}(\{X(B_{r}(x_{1}))\geq n\})+\mathbb{P}({\{|x_{1}-x_{\phi(1)}|<r\}}\cap T_{k-1+2n,r}).

Since, by definition of Tk1+2n,rT_{k-1+2n,r}, we have pointwise a.s.

i=1N𝟏{|xixϕ(i)|<r}𝟏Tk1+2n,r2k2+4n,\sum_{i=1}^{N}\mathbf{1}_{\{|x_{i}-x_{\phi(i)}|<r\}}\mathbf{1}_{T_{k-1+2n,r}}\leq 2k-2+4n,

we can apply exhchangeability to see

({|x1xϕ(1)|<r}Tk1+2n,r)2k2+4nN.\mathbb{P}({\{|x_{1}-x_{\phi(1)}|<r\}}\cap T_{k-1+2n,r})\leq\frac{2k-2+4n}{N}.

Collecting this estimate, (4.16), (4.15), and (4.14), we have

({|x1x2|r}{ϕ(1)=2}L1,RcTk1,r)\displaystyle\mathbb{P}(\{|x_{1}-x_{2}|\geq r\}\cap\{\phi(1)=2\}\cap L_{1,R}^{c}\cap T_{k-1,r})
κ(2k2+4nN(N1)+CnN({X(Br(x1))3})+2({X(Br(x1))n})).\displaystyle\leq\kappa\left(\frac{2k-2+4n}{N(N-1)}+\frac{C_{n}}{N}\mathbb{P}(\{X(B_{r}(x_{1}))\geq 3\})+2\mathbb{P}(\{X(B_{r}(x_{1}))\geq n\})\right).

Finally, plugging this bound into (4.10) along with (4.11) proves the proposition. ∎

Proposition 4.4.

Consider W=VNW=V_{N} with VV satisfying (A1) in d=2d=2 and (A2) in d3d\geq 3. In d=2d=2, the law of N12+βηkN^{\frac{1}{2+\beta}}\eta_{k} is tight as NN\to\infty and lim supNN,βVN({N12+βηkγ})Cγ4+2β4+β\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{N^{\frac{1}{2+\beta}}\eta_{k}\geq\gamma\})\leq C\gamma^{-\frac{4+2\beta}{4+\beta}} for γ>0\gamma>0. For d3d\geq 3, let ZkZ_{k} be defined by

ηk=(βlogN)1d2(1+2d2(d2)2loglogNlogN+Zk(d2)logN).\eta_{k}=\left(\frac{\beta}{\log N}\right)^{\frac{1}{d-2}}\left(1+\frac{2d-2}{(d-2)^{2}}\frac{\log\log N}{\log N}+\frac{Z_{k}}{(d-2)\log N}\right).

Then we have lim supNN,βVN({Zkγ})Ceγ/2\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{Z_{k}\geq\gamma\})\leq Ce^{-\gamma/2}.

Proof.

Both results are consequences of Proposition 4.3, Lemma 4.2, and our clustering result Theorem 3. We adopt the notation from Proposition 4.3. For the d=2d=2 result, choose r=γN12+βr=\gamma N^{-\frac{1}{2+\beta}} for γ1\gamma\geq 1, n=5n=5 (say), and let ν\nu be the uniform probability measure on Ann[r/2,r](0)\mathrm{Ann}_{[r/2,r]}(0). Without loss of generality, we assume r1r\leq 1. We compute

Mr,RCR2+β,\displaystyle M_{r,R}\leq CR^{2+\beta},
Δνlogr+C,\displaystyle\Delta_{\nu}\leq-\log r+C,
νL(2)Cr2.\displaystyle\|\nu\|_{L^{\infty}(\mathbb{R}^{2})}\leq Cr^{-2}.

Applying Proposition 4.3, Lemma 4.2, and Theorem 3, we see

N,βVN({ηkr})CN1dlogN+CR2+CR2+βγ2βeCβN22+βγ2(1+N14+3β2+2βγcβ)\mathbb{P}^{V_{N}}_{N,\beta}(\{\eta_{k}\geq r\})\leq CN^{-\frac{1}{d}}\sqrt{\log N}+CR^{-2}+CR^{2+\beta}\gamma^{-2-\beta}e^{C\beta N^{-\frac{2}{2+\beta}}\gamma^{2}}\left(1+N^{1-\frac{4+3\beta}{2+2\beta}}\gamma^{c_{\beta}}\right) (4.17)

for a constant CC depending on kk and some constant cβ>0c_{\beta}>0. Taking lim supN\limsup_{N\to\infty} of both sides and optimizing in RR proves that

lim supNN,βVN({N12+βηkγ})CRd+CR2+βγ2βCγ4+2β4+β.\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{N^{\frac{1}{2+\beta}}\eta_{k}\geq\gamma\})\leq CR^{-d}+CR^{2+\beta}\gamma^{-2-\beta}\leq C\gamma^{-\frac{4+2\beta}{4+\beta}}.

For the d3d\geq 3 result, choose r=(βlogN2d2d2loglogNγ)1d2r=\left(\frac{\beta}{\log N-\frac{2d-2}{d-2}\log\log N-\gamma}\right)^{\frac{1}{d-2}} for γ>0\gamma>0 and WLOG assume r1r\leq 1. Let ν\nu be the uniform measure on the annulus Ann[(1(logN)1)r,r](0)\mathrm{Ann}_{[(1-(\log N)^{-1})r,r]}(0), and compute

Mr,RCRd,\displaystyle M_{r,R}\leq CR^{d},
Δν1(1(logN)1)d2rd21rd2(1+d2logN+C(logN)2),\displaystyle\Delta_{\nu}\leq\frac{1}{(1-(\log N)^{-1})^{d-2}r^{d-2}}\leq\frac{1}{r^{d-2}}\left(1+\frac{d-2}{\log N}+C(\log N)^{-2}\right),
νL(d)ClogNrd.\displaystyle\|\nu\|_{L^{\infty}(\mathbb{R}^{d})}\leq\frac{C\log N}{r^{d}}.

We can then estimate

eCβr2+βΔνNMr,RνL(d)\displaystyle\frac{e^{C\beta r^{2}+\beta\Delta_{\nu}}}{N}M_{r,R}\|\nu\|_{L^{\infty}(\mathbb{R}^{d})} CRdexp(βrd2(1+d2logN)+loglogNdlogrlogN)\displaystyle\leq CR^{d}\exp\left(\frac{\beta}{r^{d-2}}\left(1+\frac{d-2}{\log N}\right)+\log\log N-d\log r-\log N\right)
CRdexp(γ+d22d2d2loglogNlogN)CRdeγ.\displaystyle\leq CR^{d}\exp\left(-\gamma+d-2-\frac{2d-2}{d-2}\frac{\log\log N}{\log N}\right)\leq CR^{d}e^{-\gamma}.

Thus we have

N,βVN({ηkr})\displaystyle\mathbb{P}^{V_{N}}_{N,\beta}(\{\eta_{k}\geq r\}) CN1dlogN+CRd+CRdeγ(1+NN,βVN({X(Br(x1))3})\displaystyle\leq CN^{-\frac{1}{d}}\sqrt{\log N}+CR^{-d}+CR^{d}e^{-\gamma}(1+N\mathbb{P}^{V_{N}}_{N,\beta}(\{X(B_{r}(x_{1}))\geq 3\})
+N2N,βVN({X(Br(x1))n}))\displaystyle\quad+N^{2}\mathbb{P}^{V_{N}}_{N,\beta}(\{X(B_{r}(x_{1}))\geq n\}))
CN1dlogN+CRd+CRdeγ(1+Ne2βrd2+N2e(n1)βrd2)\displaystyle\leq CN^{-\frac{1}{d}}\sqrt{\log N}+CR^{-d}+CR^{d}e^{-\gamma}(1+Ne^{-\frac{2\beta}{r^{d-2}}}+N^{2}e^{-\frac{(n-1)\beta}{r^{d-2}}})
CN1dlogN+CRd+CRdeγ(1+eCγlogN+CloglogN),\displaystyle\leq CN^{-\frac{1}{d}}\sqrt{\log N}+CR^{-d}+CR^{d}e^{-\gamma}(1+e^{C\gamma-\log N+C\log\log N}),

where we chose n=4n=4. We conclude

lim supNN,βVN({ηkr})CRd+CRdeγCeγ/2.\limsup_{N\to\infty}\mathbb{P}^{V_{N}}_{N,\beta}(\{\eta_{k}\geq r\})\leq CR^{-d}+CR^{d}e^{-\gamma}\leq Ce^{-\gamma/2}.

To conclude the result on ZkZ_{k}, note that

r=(βlogN)1d2(1+2d2(d2)2loglogNlogN+γ(d2)logN)+O((logN)1d22+ε)r=\left(\frac{\beta}{\log N}\right)^{\frac{1}{d-2}}\left(1+\frac{2d-2}{(d-2)^{2}}\frac{\log\log N}{\log N}+\frac{\gamma}{(d-2)\log N}\right)+O((\log N)^{-\frac{1}{d-2}-2+\varepsilon})

for any ε>0\varepsilon>0. ∎

5. Discrepancy Bounds

In this section, we prove Theorem 5 and Theorem 6. All implicit constants CC may depend on β\beta and on various characteristics of WW or VV. We let μ(dx)=1cdΔW(x)dx\mu(dx)=\frac{1}{c_{d}}\Delta W(x)dx throughout.

It will be necessary to consider more general domains Ω\Omega for our discrepancy bounds. We will work with a domain Ωd\Omega\subset\mathbb{R}^{d} that is an α\alpha-thin annulus for fixed α(0,1]\alpha\in(0,1]. To be precise, we will take either Ω\Omega to be a ball of radius RR if α=1\alpha=1 or Ω=Ann[RαR,R](z)\Omega=\mathrm{Ann}_{[R-\alpha R,R]}(z) for α(0,1)\alpha\in(0,1). This means there exists some C>0C>0 such that

C1αRd|Ω|CαRdwhere 2R=diam(Ω).C^{-1}\alpha R^{d}\leq|\Omega|\leq C\alpha R^{d}\quad\text{where}\ 2R=\mathrm{diam}(\Omega). (5.1)

We also define thickened and thinned versions of Ω\Omega. For s0s\geq 0, define Ωs=Ω{xd:dist(x,Ω)s}\Omega_{s}=\Omega\cup\{x\in\mathbb{R}^{d}:\mathrm{dist}(x,\partial\Omega)\leq s\}, and for s<0s<0, define Ωs={xΩ:dist(x,Ω)|s|}\Omega_{s}=\{x\in\Omega:\mathrm{dist}(x,\partial\Omega)\geq|s|\}. We will use 2R2R for diam(Ω)\mathrm{diam}(\Omega) throughout the first two subsections. We assume αRC\alpha R\geq C dependent on β\beta, supΔW\sup\Delta W, and infΩRΔW\inf_{\Omega_{R}}\Delta W.

We will consider the event Eρ,r,ME_{\rho,r,M} for parameters ρ,r,M>0\rho,r,M>0. It is defined by

Eρ,r,M={X(Ω)μ(Ω)+ρ|Ω|}{X(Ω5rΩ5r)MRd1r}.E_{\rho,r,M}=\{X(\Omega)\geq\mu(\Omega)+\rho|\Omega|\}\cap\{X(\Omega_{5r}\setminus\Omega_{-5r})\leq MR^{d-1}r\}. (5.2)

Let ϕ:d\phi:\mathbb{R}^{d}\to\mathbb{R} be a smooth, nonnegative, radial function with dϕ(x)𝑑x=1\int_{\mathbb{R}^{d}}\phi(x)dx=1 and support within B1(0)B_{1}(0). For s>0s>0, define ϕs=sdϕ(x/s)\phi_{s}=s^{-d}\phi(x/s), and ψs=ϕsϕ1\psi_{s}=\phi_{s}\ast\phi_{1}. For a measure (or function) λ\lambda, we will write Φsλ\Phi_{s}\lambda for ϕsλ\phi_{s}\ast\lambda when it makes sense. We will apply the isotropic averaging argument with operator Iso𝕏(Ω),ψr\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}} (identifying ψr\psi_{r} with the measure ψrdx\psi_{r}dx). The analysis mainly hinges on a precise lower bound for

Δρ,r,M:=infXNEρ,r,MW,U(XN)Iso𝕏(Ω),ψrW,U(XN).\Delta_{\rho,r,M}:=\inf_{X_{N}\in E_{\rho,r,M}}\mathcal{H}^{W,U}(X_{N})-\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\mathcal{H}^{W,U}(X_{N}). (5.3)

We want to be able to take ρ\rho to be very small and still achieve Δρ,r,M1\Delta_{\rho,r,M}\gg 1 for appropriate choices of rr and MM.

5.1. Proof idea.

The main idea of the proof is to use a continuum approximation for the energy change upon isotropic averaging to find a lower bound for Δρ,r,M\Delta_{\rho,r,M}. To a continuous charge density, we associate an isotropic averaging procedure adapted to ϕr\phi_{r} in which every “infinitesimal point charge” constituting the continuum is replaced by a infinitesimal charge shaped like ϕr\phi_{r}.

Crucially, the Coulomb energy of a continuous charge distribution contains the self-energy or “the diagonal”. That is, to a bounded charge distribution ν\nu, we associate the Coulomb energy 12𝗀(xy)ν(dx)ν(dy)\frac{1}{2}\iint\mathsf{g}(x-y)\nu(dx)\nu(dy). After isotropic averaging, i.e. replacing ν\nu by Φrν\Phi_{r}\nu, the Coulomb energy is

12d×d𝗀(xy)Φrν(dx)Φrν(dy)=12d×d(Φr2𝗀)(xy)ν(dx)ν(dy).\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathsf{g}(x-y)\Phi_{r}\nu(dx)\Phi_{r}\nu(dy)=\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\Phi_{r}^{2}\mathsf{g})(x-y)\nu(dx)\nu(dy).

The continuum Coulomb energy change is therefore

r(ν)wherer(ν):=12d×d(𝗀Φr2𝗀)(xy)ν(dx)ν(dy).-\mathcal{E}_{r}(\nu)\quad\text{where}\quad\mathcal{E}_{r}(\nu):=\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)\nu(dx)\nu(dy). (5.4)

Two facts form the crux of the method: (1) r(ν)\mathcal{E}_{r}(\nu) is convex and (2) 𝗀Φr2𝗀\mathsf{g}-\Phi_{r}^{2}\mathsf{g} is compactly supported (on scale rr).

Considering the energy term W(x)ν(dx)\int W(x)\nu(dx) associated to the potential WW, which one can think of as the Coulomb interaction between ν\nu and a signed background charge density of μ-\mu, we can precisely compare (5.4) to the change in this term upon smearing ν\nu by ϕr\phi_{r}. The energy change is d×d(𝗀Φr𝗀)(xy)ν(dx)μ(dy)\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)\nu(dx)\mu(dy). Note that the kernel is different from that of (5.4), but also the 12\frac{1}{2} factor is absent in comparison. In the case that ν(Ω)μ(Ω)+ρ|Ω|\nu(\Omega)\geq\mu(\Omega)+\rho|\Omega| and ν\nu is supported on Ω\Omega, the net energy change is favorable for the choice r=(αR)1/3r=(\alpha R)^{1/3} under some conditions on ρ>0\rho>0, provided we apply some appropriate modifications near the boundary of Ω\Omega to overcome any boundary layer effects and approximate μ\mu by a constant measure near Ω\Omega. The modifications create only boundary errors since 𝗀Φr2𝗀\mathsf{g}-\Phi_{r}^{2}\mathsf{g} is supported on a scale rr much smaller than RR. Our argument is capturing the fact that overcrowding is unfavorable for the interaction on length scale rr and below.

The remainder of the proof involves relating our continuum approximation above to the true change in energy of a point configuration upon isotropic averaging. The first step, and the reason for isotropic averaging using ψr\psi_{r} instead of ϕr\phi_{r}, is that we must renormalize by replacing our point charges by microscopic continuous charges shaped like ϕ1\phi_{1}. This allows us to make sense of the self energy of the charges (whereas the self-energy of a point charge is infinite) and so directly relate to the continuum problem. As alluded to above, we must also deal with boundary effects, which are well controlled by the parameter MM in the event Eρ,r,ME_{\rho,r,M} and our local law Theorem 1. Finally, there are some lower order entropy factors to consider, after which our model computation applies to conclude Theorem 5.

Regarding Theorem 6, we use rigidity for the fluctuation of smooth linear statistics from [Ser20] to find a screening region whenever the absolute discrepancy in a ball BR(z)B_{R}(z) is large. A screening region takes the form of an α\alpha-thin annulus Ω\Omega just inside or outside BR(z)\partial B_{R}(z) that has an excess of positive charge. We can then apply our incompressibility estimate to conclude.

5.2. Discrepancy upper bound.

In this subsection, we prove Theorem 5. We will actually generalize the result to α\alpha-thin annuli Ω\Omega under some conditions.

We will first control Δρ,r,M\Delta_{\rho,r,M} from (5.3) in terms of an “energy” functional r(ν)\mathcal{E}_{r}(\nu) defined in (5.4). We will only need to consider measures ν\nu with bounded densities, so for simplicity we restrict to this case from the start.

We let qq\in\mathbb{R} denote a positive constant to be chosen later and introduce Leb(dx)\mathrm{Leb}(dx) to denote Lebesgue measure on d\mathbb{R}^{d}. We let ν|A\nu_{|A} denote the restriction of ν\nu to a set AA for any measure ν\nu. Define

r(ν;q)=r(ν+qLeb|Ω2rΩ)qd×d(𝗀Φr𝗀)(xy)𝑑x(ν(dy)+qLeb|ΩrΩ(dy)).\mathcal{E}_{r}(\nu;q)=\mathcal{E}_{r}(\nu+q\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})-q\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx(\nu(dy)+q\mathrm{Leb}_{|\Omega_{r}\setminus\Omega}(dy)). (5.5)

The above definition models the full energy change for the continuum approximation discussed in Section 5.1. Note that we extend the measure ν\nu slightly past the boundary Ω\Omega by qdxqdx to eliminate boundary layer effects. In Proposition 5.3, we will see that ν=qLeb|Ω\nu=q\mathrm{Leb}_{|\Omega} is a minimizer of r(;q)\mathcal{E}_{r}(\cdot;q) among bounded measures supported on Ω\Omega with mass q|Ω|q|\Omega|.

Lemma 5.1.

We have

Iso{1,2},ψr𝗀(x1x2)\displaystyle\mathrm{Iso}_{\{1,2\},\psi_{r}}\mathsf{g}(x_{1}-x_{2}) =(ϕ1ϕ1ϕrϕr𝗀)(x1x2),\displaystyle=(\phi_{1}\ast\phi_{1}\ast\phi_{r}\ast\phi_{r}\ast\mathsf{g})(x_{1}-x_{2}),
Iso{1,2},ψr𝗀(x1x3)\displaystyle\mathrm{Iso}_{\{1,2\},\psi_{r}}\mathsf{g}(x_{1}-x_{3}) 𝗀(x1x3).\displaystyle\leq\mathsf{g}(x_{1}-x_{3}).

Furthermore,

Iso{1},ψrW(x1)=W(x1)\displaystyle\mathrm{Iso}_{\{1\},\psi_{r}}W(x_{1})=W(x_{1}) +d(𝗀Φ1𝗀)(yx1)μ(dy)\displaystyle+\int_{\mathbb{R}^{d}}(\mathsf{g}-\Phi_{1}\mathsf{g})(y-x_{1})\mu(dy) (5.6)
+d×d(𝗀Φr𝗀)(yx)ϕ1(xx1)𝑑xμ(dy).\displaystyle+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(y-x)\phi_{1}(x-x_{1})dx\mu(dy).
Proof.

The first two equations follow immediately from the definition of Iso{1,2},ψr\mathrm{Iso}_{\{1,2\},\psi_{r}} and superharmonicity of 𝗀\mathsf{g}; see the proof of Proposition 2.1 for a similar calculation.

Considering now isotropic averaging of the potential term and letting σ\sigma be surface measure on the unit sphere, we have

1σ(B1(0))B1(0)W(z+rθ)σ(dθ)=W(z)+Br(z)(𝗀(xz)𝗀(r))μ(dx).\frac{1}{\sigma(B_{1}(0))}\int_{\partial B_{1}(0)}W(z+r\theta)\sigma(d\theta)=W(z)+\int_{B_{r}(z)}(\mathsf{g}(x-z)-\mathsf{g}(r))\mu(dx). (5.7)

Defining 𝗀s=(𝗀(x)𝗀(s))+\mathsf{g}_{s}=(\mathsf{g}(x)-\mathsf{g}(s))_{+}, we compute using radial symmetry of ϕr\phi_{r}:

ϕrW(z)=dϕr(y)W(z+y)𝑑y\displaystyle\phi_{r}\ast W(z)=\int_{\mathbb{R}^{d}}\phi_{r}(y)W(z+y)dy =dϕr(y)1σ(B1(0))B1(0)W(z+|y|θ)𝑑θ𝑑y\displaystyle=\int_{\mathbb{R}^{d}}\phi_{r}(y)\frac{1}{\sigma(B_{1}(0))}\int_{\partial B_{1}(0)}W(z+|y|{\theta})d\theta dy
=W(z)+(dϕr(y)𝗀|y|𝑑y)μ(x).\displaystyle=W(z)+\left(\int_{\mathbb{R}^{d}}\phi_{r}(y)\mathsf{g}_{|y|}dy\right)\ast\mu(x).

Note that 𝗀s=𝗀δ(s)𝗀\mathsf{g}_{s}=\mathsf{g}-\delta^{(s)}\ast\mathsf{g}, where δ(s)\delta^{(s)} is Dirac delta “smeared” evenly on the sphere Bs(0)\partial B_{s}(0), and so

dϕr(y)𝗀|y|(x)𝑑y=𝗀(x)dd𝗀(xz)ϕr(y)δ(|y|)(dz)𝑑y=𝗀(x)d𝗀(xz)ϕr(z)𝑑z.\int_{\mathbb{R}^{d}}\phi_{r}(y)\mathsf{g}_{|y|}(x)dy=\mathsf{g}(x)-\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}\mathsf{g}(x-z)\phi_{r}(y)\delta^{(|y|)}(dz)dy=\mathsf{g}(x)-\int_{\mathbb{R}^{d}}\mathsf{g}(x-z)\phi_{r}(z)dz.

This is (𝗀Φr𝗀)(x)(\mathsf{g}-\Phi_{r}\mathsf{g})(x). In the last equality, we used that (δ(|y|)(dz)ϕr(y))𝑑y=ϕr(z)dz\int(\delta^{(|y|)}(dz)\phi_{r}(y))dy=\phi_{r}(z)dz, formally, since ϕ\phi is radial. By applying the above computation twice, we have

ψrW(xi)\displaystyle\psi_{r}\ast W(x_{i}) =ϕr(ϕ1W)(xi)=(ϕ1W)(xi)+((𝗀Φr𝗀)ϕ1μ)(xi)\displaystyle=\phi_{r}\ast(\phi_{1}\ast W)(x_{i})=(\phi_{1}\ast W)(x_{i})+((\mathsf{g}-\Phi_{r}\mathsf{g})\ast\phi_{1}\ast\mu)(x_{i})
=W(xi)+(𝗀Φ1𝗀)μ(xi)+((𝗀Φr𝗀)ϕ1μ)(xi).\displaystyle=W(x_{i})+(\mathsf{g}-\Phi_{1}\mathsf{g})\ast\mu(x_{i})+((\mathsf{g}-\Phi_{r}\mathsf{g})\ast\phi_{1}\ast\mu)(x_{i}).

The last line with i=1i=1 gives (5.6). ∎

The next lemma directly relates the energy change of a point charge system after isotropic averaging to the functional r\mathcal{E}_{r}. There are error terms corresponding to self-energy terms (Errorvol\mathrm{Error}_{\mathrm{vol}}), boundary layer effects (Errorbl\mathrm{Error}_{\mathrm{bl}}), and approximation of μ\mu by qq.

Lemma 5.2.

Let r>1r>1 and qq\in\mathbb{R}. We have

W,U(XN)Iso𝕏(Ω),ψrW,U(XN)\displaystyle\mathcal{H}^{W,U}(X_{N})-\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\mathcal{H}^{W,U}(X_{N}) (5.8)
r((Φ1X|Ω)|Ω;q)+d×Ω(𝗀Φr𝗀)(yx)(qLebμ)(dx)(Φ1X|Ω)(dy)\displaystyle\geq\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q)+\iint_{\mathbb{R}^{d}\times\Omega}(\mathsf{g}-\Phi_{r}\mathsf{g})(y-x)(q\mathrm{Leb}-\mu)(dx)(\Phi_{1}X_{|\Omega})(dy)
ErrorblErrorvol,\displaystyle\quad-\mathrm{Error}_{\mathrm{bl}}-\mathrm{Error}_{\mathrm{vol}},

where

ErrorblCr2|q|X(ΩΩ3r)+Cr3q2Rd1+Cr2(1+|q|)X(ΩΩ1),\mathrm{Error}_{\mathrm{bl}}\leq Cr^{2}|q|X(\Omega\setminus\Omega_{-3r})+Cr^{3}q^{2}R^{d-1}+Cr^{2}(1+|q|)X(\Omega\setminus\Omega_{-1}), (5.9)

and

ErrorvolX(Ω)𝗀(4r)+CX(Ω).\mathrm{Error}_{\mathrm{vol}}\leq-X(\Omega)\mathsf{g}(4r)+CX(\Omega). (5.10)
Proof.

By Lemma 5.1, we have

Iso𝕏(Ω),ψr0(XN)12ij{i,j}𝕏(Ω)𝗀(xixj)+12i,j𝕏(Ω)(Φ12Φr2𝗀)(xixj)12X(Ω)Φ12Φr2𝗀(0).\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\mathcal{H}^{0}(X_{N})\leq\frac{1}{2}\sum_{\begin{subarray}{c}i\neq j\\ \{i,j\}\not\subset\mathbb{X}(\Omega)\end{subarray}}\mathsf{g}(x_{i}-x_{j})+\frac{1}{2}\sum_{i,j\in\mathbb{X}(\Omega)}(\Phi_{1}^{2}\Phi_{r}^{2}\mathsf{g})(x_{i}-x_{j})-\frac{1}{2}X(\Omega)\Phi_{1}^{2}\Phi_{r}^{2}\mathsf{g}(0).

Note that we have added and subtracted the diagonal terms in the second sum on the RHS. We use that Φ1\Phi_{1} is self-dual to write

12i,j𝕏(Ω)(Φ12Φr2𝗀)(xixj)=12d×dΦr2𝗀(xy)(Φ1X|Ω)(dx)(Φ1X|Ω)(dy).\frac{1}{2}\sum_{i,j\in\mathbb{X}(\Omega)}(\Phi_{1}^{2}\Phi_{r}^{2}\mathsf{g})(x_{i}-x_{j})=\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\Phi_{r}^{2}\mathsf{g}(x-y)(\Phi_{1}X_{|\Omega})(dx)(\Phi_{1}X_{|\Omega})(dy).

Considering the interaction within 0(XN)\mathcal{H}^{0}(X_{N}), we estimate

12i,j𝕏(Ω)ij𝗀(xixj)12d×d𝗀(xy)(Φ1X|Ω)(dx)(Φ1X|Ω)(dy)12X(Ω)Φ12𝗀(0)\frac{1}{2}\sum_{\begin{subarray}{c}i,j\in\mathbb{X}(\Omega)\\ i\neq j\end{subarray}}\mathsf{g}(x_{i}-x_{j})\geq\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathsf{g}(x-y)(\Phi_{1}X_{|\Omega})(dx)(\Phi_{1}X_{|\Omega})(dy)-\frac{1}{2}X(\Omega)\Phi_{1}^{2}\mathsf{g}(0)

using Φ12𝗀𝗀\Phi_{1}^{2}\mathsf{g}\leq\mathsf{g} pointwise and duality like above. It follows that

0(XN)Iso𝕏(Ω),ψr0(XN)\displaystyle\mathcal{H}^{0}(X_{N})-\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\mathcal{H}^{0}(X_{N}) 12d×d(𝗀Φr2𝗀)(xy)(Φ1X|Ω)(dx)(Φ1X|Ω)(dy)\displaystyle\geq\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})(dx)(\Phi_{1}X_{|\Omega})(dy) (5.11)
+12X(Ω)(Φ12Φr2𝗀(0)Φ12𝗀(0)).\displaystyle\quad+\frac{1}{2}X(\Omega)(\Phi_{1}^{2}\Phi_{r}^{2}\mathsf{g}(0)-\Phi_{1}^{2}\mathsf{g}(0)).

We can compare the double integral above to r((Φ1X|Ω)|Ω+qLeb|Ω2rΩ)\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega}+q\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega}). They will only differ by boundary layer terms. We start by simply restricting the integral to Ω×Ω\Omega\times\Omega using 𝗀Φr2𝗀0\mathsf{g}-\Phi_{r}^{2}\mathsf{g}\geq 0. After doing so, we find

12d×d(𝗀Φr2𝗀)(xy)(Φ1X|Ω)2(dx,dy)r((Φ1X|Ω)|Ω+qLeb|Ω2rΩ)T1T2\frac{1}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})^{\otimes 2}(dx,dy)\geq\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega}+q\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})-T_{1}-T_{2}

where

T1\displaystyle T_{1} =qΩ×d(𝗀Φr2𝗀)(xy)Φ1X|Ω(dx)Leb|Ω2rΩ(dy)Cr2|q|X(ΩΩ2r1),\displaystyle=q\iint_{\Omega\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)\Phi_{1}X_{|\Omega}(dx)\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega}(dy)\leq Cr^{2}|q|X(\Omega\setminus\Omega_{-2r-1}),
T2\displaystyle T_{2} =q2d×d(𝗀Φr2𝗀)(xy)Leb|Ω2rΩ(dy)Leb|Ω2rΩ(dy)Cr3q2Rd1.\displaystyle=q^{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega}(dy)\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega}(dy)\leq Cr^{3}q^{2}R^{d-1}.

We also bound

12X(Ω)(Φ12Φr2𝗀(0)Φ12𝗀(0))X(Ω)(𝗀(4r)C).\frac{1}{2}X(\Omega)(\Phi_{1}^{2}\Phi_{r}^{2}\mathsf{g}(0)-\Phi_{1}^{2}\mathsf{g}(0))\geq X(\Omega)(\mathsf{g}(4r)-C).

This term, as well as T1T_{1} and T2T_{2}, are absorbed into Errorvol\mathrm{Error}_{\mathrm{vol}} and Errorbl\mathrm{Error}_{\mathrm{bl}}.

We now handle the potential terms. Like usual, we can handle the UU term using superharmonicity. For the WW term, we use Lemma 5.1 to see

Iso𝕏(Ω),ψri𝕏(Ω)W(xi)=i𝕏(Ω)W(xi)\displaystyle\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\sum_{i\in\mathbb{X}(\Omega)}W(x_{i})=\sum_{i\in\mathbb{X}(\Omega)}W(x_{i}) (5.12)
+d×d(𝗀Φ1𝗀)(xy)X|Ω(dx)μ(dy)+d×d(𝗀Φr𝗀)(xy)(Φ1X|Ω)(dx)μ(dy).\displaystyle\quad+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{1}\mathsf{g})(x-y)X_{|\Omega}(dx)\mu(dy)+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})(dx)\mu(dy).

The middle term on the RHS in (5.12) contributes to the volume error Errorvol\mathrm{Error}_{\mathrm{vol}}. It is bounded by

d×d(𝗀Φ1𝗀)(xy)X|Ω(dx)μ(dy)CX(Ω).\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{1}\mathsf{g})(x-y)X_{|\Omega}(dx)\mu(dy)\leq CX(\Omega).

For the last term in (5.12), we replace μ\mu by qq to generate the term

d×d(𝗀Φr𝗀)(xy)(Φ1X|Ω)(dx)(qLebμ)(dy)\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})(dx)(q\mathrm{Leb}-\mu)(dy)

in (5.8). We then estimate

qd×d(𝗀Φr𝗀)(xy)(Φ1X|Ω)(dx)𝑑y\displaystyle q\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})(dx)dy
qΩ×d(𝗀Φr𝗀)(xy)(Φ1X|Ω)(dx)𝑑y+Cr2|q|X(ΩΩ1)\displaystyle\leq q\iint_{\Omega\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)(\Phi_{1}X_{|\Omega})(dx)dy+Cr^{2}|q|X(\Omega\setminus\Omega_{-1})
qd×d(𝗀Φr𝗀)(xy)((Φ1X|Ω)|Ω+qLeb|ΩrΩ)(dx)𝑑y+Cr2|q|X(ΩΩ1),\displaystyle\leq q\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)((\Phi_{1}X_{|\Omega})_{|\Omega}+q\mathrm{Leb}_{|\Omega_{r}\setminus\Omega})(dx)dy+Cr^{2}|q|X(\Omega\setminus\Omega_{-1}),

where we used 𝗀Φr𝗀L1(d)Cr2\|\mathsf{g}-\Phi_{r}\mathsf{g}\|_{L^{1}(\mathbb{R}^{d})}\leq Cr^{2} and the fact that the mass of Φ1X|Ω\Phi_{1}X_{|\Omega} lying outside of Ω\Omega is bounded by X(ΩΩ1)X(\Omega\setminus\Omega_{-1}). Assembling the above estimates proves the lemma. ∎

We now turn to studying minimizers of r(;q)\mathcal{E}_{r}(\cdot;q) conditioned on the weight given to Ω\Omega. The following proposition is the key technical result of Section 5, and it is the reason for considering the precise form of the energy r(;q)\mathcal{E}_{r}(\cdot;q).

Proposition 5.3.

Let Ω\Omega an α\alpha-thin annulus and qq\in\mathbb{R}. We have

infν:ν(Ω)=q|Ω|r(ν;q)0\inf_{\nu:\nu(\Omega)=q|\Omega|}\mathcal{E}_{r}(\nu;q)\geq 0 (5.13)

where the infimum is over measures ν\nu supported on Ω¯\overline{\Omega} with a bounded Lebesgue density.

Proof.

First, note that r()\mathcal{E}_{r}(\cdot) as defined in (5.4) is non-negative when applied to measures with bounded density and compact support. This is because the Fourier transform of δ0ϕrϕr\delta_{0}-\phi_{r}\ast\phi_{r} is non-negative. Indeed, the Fourier transform on d\mathbb{R}^{d} of ϕrϕr\phi_{r}\ast\phi_{r} is real since ϕr\phi_{r} is radial, and it is bounded above by 11 (in the normalization for which δ^0=1\hat{\delta}_{0}=1) since it is a probability density. Since 𝗀\mathsf{g} has positive Fourier transform, the Fourier transform of 𝗀Φr2𝗀=(δ0ϕrϕr)𝗀\mathsf{g}-\Phi_{r}^{2}\mathsf{g}=(\delta_{0}-\phi_{r}\ast\phi_{r})\ast\mathsf{g} is non-negative, and non-negativity of r()\mathcal{E}_{r}(\cdot) follows from Plancherel’s theorem.

Let ν\nu be a measure supported on Ω¯\overline{\Omega} with a bounded Lebesgue density and ν(Ω)=q|Ω|\nu(\Omega)=q|\Omega|. We use that r(;q)\mathcal{E}_{r}(\cdot;q) is quadratic to expand

r(ν;q)\displaystyle\mathcal{E}_{r}(\nu;q) =r(qLeb|Ω;q)+r(qLeb|Ων)\displaystyle=\mathcal{E}_{r}(q\mathrm{Leb}_{|\Omega};q)+\mathcal{E}_{r}(q\mathrm{Leb}_{|\Omega}-\nu) (5.14)
+d×d(𝗀Φr2𝗀)(xy)(νqLeb|Ω)(dx)(qLeb|Ω2r)(dy)\displaystyle\quad+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)(\nu-q\mathrm{Leb}_{|\Omega})(dx)(q\mathrm{Leb}_{|\Omega_{2r}})(dy)
qd×d(𝗀Φr𝗀)(xy)𝑑x(νqLeb|Ω)(dy).\displaystyle\quad-q\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx(\nu-q\mathrm{Leb}_{|\Omega})(dy).

We claim that both terms on the last line are 0. Indeed, we have

d(𝗀Φr𝗀)(xy)𝑑y=c1,ϕ,r,d(𝗀Φr2𝗀)(xy)𝑑y=c2,ϕ,r\int_{\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dy=c_{1,\phi,r},\quad\int_{\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)dy=c_{2,\phi,r} (5.15)

for constants c1,ϕ,r,c2,ϕ,rc_{1,\phi,r},c_{2,\phi,r} independent of xx, and the same holds for Leb|Ω2r(dy)\mathrm{Leb}_{|\Omega_{2r}}(dy) in place of dydy as long as xΩx\in\Omega. The claim follows from ν(Ω)=q\nu(\Omega)=q. Since r(qLeb|Ων)0\mathcal{E}_{r}(q\mathrm{Leb}_{|\Omega}-\nu)\geq 0, we have proved that the infimum in (5.13) is attained at ν=qLeb|Ω\nu=q\mathrm{Leb}_{|\Omega}.

It remains to compute r(qLeb|Ω;q)\mathcal{E}_{r}(q\mathrm{Leb}_{|\Omega};q). We write r(qLeb|Ω;q)=T1+T2T3\mathcal{E}_{r}(q\mathrm{Leb}_{|\Omega};q)=T_{1}+T_{2}-T_{3} for

T1\displaystyle T_{1} =q22d×d(𝗀Φr𝗀)(xy)Leb|Ω2r(dx)Leb|Ω2r(dy),\displaystyle=\frac{q^{2}}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)\mathrm{Leb}_{|\Omega_{2r}}(dx)\mathrm{Leb}_{|\Omega_{2r}}(dy),
T2\displaystyle T_{2} =q22d×dΦr(𝗀Φr𝗀)(xy)Leb|Ω2r(dx)Leb|Ω2r(dy),\displaystyle=\frac{q^{2}}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\Phi_{r}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)\mathrm{Leb}_{|\Omega_{2r}}(dx)\mathrm{Leb}_{|\Omega_{2r}}(dy),
T3\displaystyle T_{3} =q2d×d(𝗀Φr𝗀)(xy)𝑑xLeb|Ωr(dy).\displaystyle=q^{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx\mathrm{Leb}_{|\Omega_{r}}(dy).

Note that 𝗀Φr𝗀0\mathsf{g}-\Phi_{r}\mathsf{g}\geq 0, and so

T1q22|Ωr|infyΩrΩ2r(𝗀Φr𝗀)(xy)=12q2c1,ϕ,r|Ωr|.T_{1}\geq\frac{q^{2}}{2}|\Omega_{r}|\inf_{y\in\Omega_{r}}\int_{\Omega_{2r}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)=\frac{1}{2}q^{2}c_{1,\phi,r}|\Omega_{r}|.

We use ΦrLeb|Ω2rLeb|Ωr\Phi_{r}\mathrm{Leb}_{|\Omega_{2r}}\geq\mathrm{Leb}_{|\Omega_{r}} to see

T2\displaystyle T_{2} =q22d×d(𝗀Φr𝗀)(xy)ΦrLeb|Ω2r(dx)Leb|Ω2r(dy)\displaystyle=\frac{q^{2}}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)\Phi_{r}\mathrm{Leb}_{|\Omega_{2r}}(dx)\mathrm{Leb}_{|\Omega_{2r}}(dy)
q22Ωr×Ω2r(𝗀Φr𝗀)(xy)𝑑x𝑑yc1,ϕ,rq22|Ωr|.\displaystyle\geq\frac{q^{2}}{2}\iint_{\Omega_{r}\times\Omega_{2r}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dxdy\geq\frac{c_{1,\phi,r}q^{2}}{2}|\Omega_{r}|.

Similarly, we have T3q2c1,ϕ,r|Ωr|T_{3}\leq q^{2}c_{1,\phi,r}|\Omega_{r}|, and combining the bounds on T1,T2,T3T_{1},T_{2},T_{3} finishes the proof. ∎

Proposition 5.4.

Let Ω\Omega be an α\alpha-thin annulus with diameter 2R2R. Assume that the parameters ρ,α,q>0\rho,\alpha,q>0 and M1M\geq 1 satisfy the following for fixed constants CiC_{i}, i=1,2,3,4,5i=1,2,3,4,5.

  1. (1)

    There is bounded excess:

    ρC1.\rho\leq C_{1}. (5.16)
  2. (2)

    The constant qq approximates μ\mu:

    μqL(ΩR)C21ρ.\|\mu-q\|_{L^{\infty}(\Omega_{R})}\leq C_{2}^{-1}\rho. (5.17)
  3. (3)

    The annulus is not too thin:

    αRC3.\alpha R\geq C_{3}. (5.18)
  4. (4)

    There is significant charge excess:

    (αR)2/3ρ1+𝟏d=2log(αR)C4.\frac{(\alpha R)^{2/3}\rho}{1+\mathbf{1}_{d=2}\log(\alpha R)}\geq C_{4}. (5.19)
  5. (5)

    The boundary layer density is not too high:

    MC51(αR)2/3ρ.M\leq C_{5}^{-1}(\alpha R)^{2/3}\rho. (5.20)

Assume that, dependent on (infΩμ)1(\inf_{\Omega}\mu)^{-1}, supΩRμ\sup_{\Omega_{R}}\mu, β\beta, we have Ci1C_{i}\gg 1 for i=2,3,4,5i=2,3,4,5 and C2C1C_{2}\geq C_{1}. Then we have

Δρ,r,MC1(αR)2/3ρ(μ(Ω)+ρ|Ω|)\Delta_{\rho,r,M}\geq C^{-1}(\alpha R)^{2/3}\rho(\mu(\Omega)+\rho|\Omega|) (5.21)

for r=(αR)1/3r=(\alpha R)^{1/3} and the quantity Δρ,r,M\Delta_{\rho,r,M} defined in (5.3).

Proof.

Let r=(αR)1/3r=(\alpha R)^{1/3} and let XNEρ,r,MX_{N}\in E_{\rho,r,M} be arbitrary.

Step 1: We begin by estimating the energy change for the continuum approximation discussed in Section 5.1. First, note that C2C1C_{2}\geq C_{1} means that

qμqL(ΩR)+μL(ΩR)C21ρ+CC21C1+CC.q\leq\|\mu-q\|_{L^{\infty}(\Omega_{R})}+\|\mu\|_{L^{\infty}(\Omega_{R})}\leq C_{2}^{-1}\rho+C\leq C_{2}^{-1}C_{1}+C\leq C.

Define qX=q+mXq_{X}=q+m_{X} for

mX:=1|Ω|(Φ1X|Ω(Ω)q|Ω|).m_{X}:=\frac{1}{|\Omega|}(\Phi_{1}X_{|\Omega}(\Omega)-q|\Omega|).

The parameter mXm_{X} acts as an excess charge density beyond qq accounting for boundary layer effects. We estimate it by

Φ1X|Ω(Ω)X(Ω)X(ΩΩ1)\displaystyle\Phi_{1}X_{|\Omega}(\Omega)\geq X(\Omega)-X(\Omega\setminus\Omega_{-1}) μ(Ω)+ρ|Ω|MRd1r\displaystyle\geq\mu(\Omega)+\rho|\Omega|-MR^{d-1}r
(q+ρ)|Ω|MRd1rμqL(Ω)|Ω|.\displaystyle\geq(q+\rho)|\Omega|-MR^{d-1}r-\|\mu-q\|_{L^{\infty}(\Omega)}|\Omega|.

Applying our assumptions, we see

mXρMRd1r|Ω|μqL(Ω)\displaystyle m_{X}\geq\rho-\frac{MR^{d-1}r}{|\Omega|}-\|\mu-q\|_{L^{\infty}(\Omega)} ρCM(αR)2/3μqL(Ω)\displaystyle\geq\rho-CM(\alpha R)^{-2/3}-\|\mu-q\|_{L^{\infty}(\Omega)} (5.22)
ρCC51ρC21ρ12ρ.\displaystyle\geq\rho-CC_{5}^{-1}\rho-C_{2}^{-1}\rho\geq\frac{1}{2}\rho.

Since qX|Ω|=Φ1X|Ω(Ω)q_{X}|\Omega|=\Phi_{1}X_{|\Omega}(\Omega), by Proposition 5.3 we have

r((Φ1X|Ω)|Ω;qX)0.\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q_{X})\geq 0.

We now lower bound r((Φ1X|Ω)|Ω;q)\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q) by comparison to r((Φ1X|Ω)|Ω;qX)\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q_{X}). First, for any measure ν\nu with bounded density, we use that r()\mathcal{E}_{r}(\cdot) is quadratic to compute

r(ν+qLeb|Ω2rΩ)r(ν+qXLeb|Ω2rΩ)\displaystyle\mathcal{E}_{r}(\nu+q\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})-\mathcal{E}_{r}(\nu+q_{X}\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})
=qqX2d×(Ω2rΩ)(𝗀Φr2𝗀)(xy)(2ν+(qX+q)Leb|Ω2rΩ)(dx)𝑑y\displaystyle=\frac{q-q_{X}}{2}\iint_{\mathbb{R}^{d}\times(\Omega_{2r}\setminus\Omega)}(\mathsf{g}-\Phi_{r}^{2}\mathsf{g})(x-y)(2\nu+(q_{X}+q)\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})(dx)dy
|qqX|2𝗀Φr2𝗀L1(d)(2|ν|+(qX+q)Leb|Ω2rΩ)(Ω5rΩ3r)\displaystyle\geq-\frac{|q-q_{X}|}{2}\|\mathsf{g}-\Phi_{r}^{2}\mathsf{g}\|_{L^{1}(\mathbb{R}^{d})}(2|\nu|+(q_{X}+q)\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})(\Omega_{5r}\setminus\Omega_{-3r})
C|qqX|r2(|ν|(Ω5rΩ2r)+CRd1r).\displaystyle\geq-C|q-q_{X}|r^{2}(|\nu|(\Omega_{5r}\setminus\Omega_{-2r})+CR^{d-1}r).

The bound in the second to last line follows from the fact that 𝗀Φr2𝗀\mathsf{g}-\Phi_{r}^{2}\mathsf{g} has support of diameter at most 4r4r and so we only need to integrate over xΩ5rΩ3rx\in\Omega_{5r}\setminus\Omega_{-3r}. For ν=(Φ1X|Ω)|Ω\nu=(\Phi_{1}X_{|\Omega})_{|\Omega}, we see

r((Φ1X|Ω)|Ω+qLeb|Ω2rΩ)r((Φ1X|Ω)|Ω+qXLeb|Ω2rΩ)CmX(M+C)Rd1r3.\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega}+q\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})-\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega}+q_{X}\mathrm{Leb}_{|\Omega_{2r}\setminus\Omega})\geq-Cm_{X}(M+C)R^{d-1}r^{3}.

Plugging this computation into (5.5), we find

r((Φ1X|Ω)|Ω;q)r((Φ1X|Ω)|Ω;qX)\displaystyle\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q)-\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q_{X})
CmX(M+C)Rd1r3+qXd×d(𝗀Φr𝗀)(xy)𝑑x(Φ1X|Ω+qXLeb|ΩrΩ)(dy)\displaystyle\geq-Cm_{X}(M+C)R^{d-1}r^{3}+q_{X}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx(\Phi_{1}X_{|\Omega}+q_{X}\mathrm{Leb}_{|\Omega_{r}\setminus\Omega})(dy)
qd×d(𝗀Φr𝗀)(xy)𝑑x(Φ1X|Ω+qLeb|ΩrΩ)(dy)\displaystyle\quad-q\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx(\Phi_{1}X_{|\Omega}+q\mathrm{Leb}_{|\Omega_{r}\setminus\Omega})(dy)
CmX(M+C)Rd1r3+(qXq)d×d(𝗀Φr𝗀)(xy)𝑑x(Φ1X|Ω+qLeb|ΩrΩ)(dy)\displaystyle\geq-Cm_{X}(M+C)R^{d-1}r^{3}+(q_{X}-q)\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(\mathsf{g}-\Phi_{r}\mathsf{g})(x-y)dx(\Phi_{1}X_{|\Omega}+q\mathrm{Leb}_{|\Omega_{r}\setminus\Omega})(dy)
CmX(M+C)αRd+mX𝗀Φr𝗀L1(d)Φ1X|Ω(Ω).\displaystyle\geq-Cm_{X}(M+C)\alpha R^{d}+m_{X}\|\mathsf{g}-\Phi_{r}\mathsf{g}\|_{L^{1}(\mathbb{R}^{d})}\Phi_{1}X_{|\Omega}(\Omega).

We can compute 𝗀Φr𝗀L1(d)=cϕr2\|\mathsf{g}-\Phi_{r}\mathsf{g}\|_{L^{1}(\mathbb{R}^{d})}=c_{\phi}r^{2} for some constant cϕ>0c_{\phi}>0. Furthermore by (5.16), (5.18), and (5.20), we have (M+C)αRdr2αRdCcϕr2Φ1X|Ω(Ω)(M+C)\alpha R^{d}\ll r^{2}\alpha R^{d}\leq Cc_{\phi}r^{2}\Phi_{1}X_{|\Omega}(\Omega). It follows

r((Φ1X|Ω)|Ω;q)r((Φ1X|Ω)|Ω;qX)12cϕmXr2Φ1X|Ω(Ω).\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q)-\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q_{X})\geq\frac{1}{2}c_{\phi}m_{X}r^{2}\Phi_{1}X_{|\Omega}(\Omega).

We apply our estimate (5.22) on mXm_{X} and the similar estimate Φ1X|Ω(Ω)12X(Ω)\Phi_{1}X_{|\Omega}(\Omega)\geq\frac{1}{2}X(\Omega) to see

r((Φ1X|Ω)|Ω;q)14cϕr2ρX(Ω).\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q)\geq\frac{1}{4}c_{\phi}r^{2}\rho X(\Omega). (5.23)

This term will be the main energy benefit of isotropic averaging. Using infΩμC1\inf_{\Omega}\mu\geq C^{-1}, we can estimate

r((Φ1X|Ω)|Ω;q)C1ρr2αRdC1C4αRd.\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q)\geq C^{-1}\rho r^{2}\alpha R^{d}\geq C^{-1}C_{4}\alpha R^{d}. (5.24)

Step 2: We now relate the isotropic averaging energy change associated to XX to the quantity r((Φ1X|Ω)|Ω;q)\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q) using Lemma 5.1. We claim that

W,U(XN)Iso𝕏(Ω),ψrW,U(XN)12r((Φ1X|Ω)|Ω;q).\mathcal{H}^{W,U}(X_{N})-\mathrm{Iso}_{\mathbb{X}(\Omega),\psi_{r}}\mathcal{H}^{W,U}(X_{N})\geq\frac{1}{2}\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q). (5.25)

This claim, combined with (5.23), will finish the proof. By step 1, it is sufficient to bound the error terms within Lemma 5.2 by 18cϕr2ρX(Ω)\frac{1}{8}c_{\phi}r^{2}\rho X(\Omega). Using the notation from that lemma, we compute

Errorbl\displaystyle\mathrm{Error}_{\mathrm{bl}} Cr3(1+q)MRd1+Cr3q2Rd1CαRd(M+1)\displaystyle\leq Cr^{3}(1+q)MR^{d-1}+Cr^{3}q^{2}R^{d-1}\leq C\alpha R^{d}(M+1)
CαRd(C51(αR)2/3ρ+1)CαRd+CC51ρr2X(Ω).\displaystyle\leq C\alpha R^{d}(C_{5}^{-1}(\alpha R)^{2/3}\rho+1)\leq C\alpha R^{d}+CC_{5}^{-1}\rho r^{2}X(\Omega).

Applying (5.23) and (5.24), we see that

Errorbl1100r((Φ1X|Ω)|Ω;q).\mathrm{Error}_{\mathrm{bl}}\leq\frac{1}{100}\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q).

Next, we estimate the volume error term. In d=2d=2, we can compute using (5.19)

ErrorvolX(Ω)(log(αR)+C)C41X(Ω)(αR)2/3ρ+CX(Ω)1400cϕr2ρX(Ω).\mathrm{Error}_{\mathrm{vol}}\leq X(\Omega)(\log(\alpha R)+C)\leq C_{4}^{-1}X(\Omega)(\alpha R)^{2/3}\rho+CX(\Omega)\leq\frac{1}{400}c_{\phi}r^{2}\rho X(\Omega). (5.26)

In the last inequality, we used that r2ρC41r^{2}\rho\geq C_{4}\gg 1. In d3d\geq 3, we can delete log(αR)\log(\alpha R) above and have the same final result. By (5.23), we have dominated the volume error term by 1100r((Φ1X|Ω)|Ω;q)\frac{1}{100}\mathcal{E}_{r}((\Phi_{1}X_{|\Omega})_{|\Omega};q). The last remaining error term is related to the approximation of μ\mu by qq:

d×Ω(𝗀Φr𝗀)(qμ)(dx)(Φ1X|Ω)(dy)cϕr2qμL(Ωr)X(Ω)C21cϕr2ρX(Ω).\iint_{\mathbb{R}^{d}\times\Omega}(\mathsf{g}-\Phi_{r}\mathsf{g})(q-\mu)(dx)(\Phi_{1}X_{|\Omega})(dy)\geq-c_{\phi}r^{2}\|q-\mu\|_{L^{\infty}(\Omega_{r})}X(\Omega)\geq-C_{2}^{-1}c_{\phi}r^{2}\rho X(\Omega). (5.27)

Using (5.23) and C21C_{2}\gg 1 allows us to dominate this term as well. Assembling the above allows us to conclude the claim (5.25) and the proof. ∎

We will apply Proposition 5.4 to prove Theorem 5, but first we take care of the case in which ρ\rho is very large using our high density law Theorem 1.

Proposition 5.5.

Let Ω\Omega be an α\alpha-thin annulus of diameter 2R2R such that αR1logα\alpha R\geq 1-\log\alpha. Let ρC1\rho\geq C_{1} for a fixed constant C1C_{1} taken large enough. Then we have

N,βW,U({X(Ω)μ(Ω)+ρ|Ω|})eαd+2Rd+2.\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega)\geq\mu(\Omega)+\rho|\Omega|\})\leq e^{-\alpha^{d+2}R^{d+2}}.
Proof.

Let (Aλ)λΛ(A_{\lambda})_{\lambda\in\Lambda} be a covering of Ω\Omega by balls AλA_{\lambda} of radius αR\alpha R of cardinality at most Cαd+1C\alpha^{-d+1}. We have

N,βW,U({X(Ω)μ(Ω)+ρ|Ω|})\displaystyle\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega)\geq\mu(\Omega)+\rho|\Omega|\}) N,βW,U({X(Ω)ρ|Ω|})\displaystyle\leq\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega)\geq\rho|\Omega|\})
|Λ|supλΛN,βW,U({X(Aλ)C1αd1ρ|Ω|}).\displaystyle\leq|\Lambda|\sup_{\lambda\in\Lambda}\mathbb{P}^{W,U}_{N,\beta}(\{X(A_{\lambda})\geq C^{-1}\alpha^{d-1}\rho|\Omega|\}).

If C1C_{1} is large enough, we have that αd1ρ|Ω||Aλ|\alpha^{d-1}\rho|\Omega|\gg|A_{\lambda}|, so we may apply Theorem 1 to see

N,βW,U({X(Ω)μ(Ω)+ρ|Ω|})\displaystyle\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega)\geq\mu(\Omega)+\rho|\Omega|\}) Cαd+1eC1(αρ|Ω|)d+2\displaystyle\leq C\alpha^{-d+1}e^{-C^{-1}(\alpha\rho|\Omega|)^{d+2}}
Cα1eC1C12αd+2Rd+2eαd+2Rd+2.\displaystyle\leq C\alpha^{-1}e^{-C^{-1}C_{1}^{2}\alpha^{d+2}R^{d+2}}\leq e^{-\alpha^{d+2}R^{d+2}}.

We now consider the more general case.

Proposition 5.6.

Let Ω\Omega be an α\alpha-thin annulus of diameter 2R2R. Let C1C_{1} be the constant fixed in Proposition 5.5. Suppose we have parameters q>0q>0 and ρ>0\rho>0 such that conditions (5.17), (5.18), (5.19) hold for large enough constants C2,C3,C4C_{2},C_{3},C_{4} with CiC1C_{i}\gg C_{1} for i=2,3,4i=2,3,4. Then we have

N,βW,U({X(Ω)μ(Ω)+ρ|Ω|})ec(αR)2/3ρ(μ(Ω)+ρ|Ω|)+ec(αR)d/3+2ρ2+eαd+2Rd+2,\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega)\geq\mu(\Omega)+\rho|\Omega|\})\leq e^{-c(\alpha R)^{2/3}\rho(\mu(\Omega)+\rho|\Omega|)}+e^{-c(\alpha R)^{d/3+2}\rho^{2}}+e^{-\alpha^{d+2}R^{d+2}}, (5.28)

for some c>0c>0.

Proof.

Let r=(αR)1/3r=(\alpha R)^{1/3} and M=C51ρr2M=C_{5}^{-1}\rho r^{2} and for a large enough constant C5C_{5} (C5=C4C_{5}=\sqrt{C_{4}} will work if C4C_{4} is chosen large enough).

Step 1: We first bound the probability of Eρ,r,ME_{\rho,r,M} by the isotropic averaging argument. Let nρ=ρ|Ω|+μ(Ω)n_{\rho}=\lceil\rho|\Omega|+\mu(\Omega)\rceil and Mρ=MRd1rM_{\rho}=\lceil MR^{d-1}r\rceil and N1=C1|Ω|+μ(Ω)N_{1}=\lfloor C_{1}|\Omega|+\mu(\Omega)\rfloor. We write

Eρ,r,M{X(Ω)μ(Ω)+C1|Ω|}m=0Mρn=nρN1{1,,N}||=m𝒩{1,,N}|𝒩|=nE,𝒩,r,ME_{\rho,r,M}\subset\{X(\Omega)\geq\mu(\Omega)+C_{1}|\Omega|\}\cup\bigcup_{m=0}^{M_{\rho}}\bigcup_{n=n_{\rho}}^{N_{1}}\bigcup_{\begin{subarray}{c}\mathcal{M}\subset\{1,\ldots,N\}\\ |\mathcal{M}|=m\end{subarray}}\bigcup_{\begin{subarray}{c}\mathcal{N}\subset\{1,\ldots,N\}\\ |\mathcal{N}|=n\end{subarray}}E_{\mathcal{M},\mathcal{N},r,M} (5.29)

for

E,𝒩,r,M={𝕏(Ω)=𝒩}{𝕏(Ω5rΩ)=}{X(Ω5rΩ5r)MRd1r}.E_{\mathcal{M},\mathcal{N},r,M}=\{\mathbb{X}(\Omega)=\mathcal{N}\}\cap\{\mathbb{X}(\Omega_{5r}\setminus\Omega)=\mathcal{M}\}\cap\{X(\Omega_{5r}\setminus\Omega_{-5r})\leq MR^{d-1}r\}.

Note that

Iso𝒩,ψr𝟏E,𝒩,r,M𝟏{𝕏(Ω5r)=𝒩}.\mathrm{Iso}_{\mathcal{N},\psi_{r}}^{\ast}\mathbf{1}_{E_{\mathcal{M},\mathcal{N},r,M}}\leq\mathbf{1}_{\{\mathbb{X}(\Omega_{5r})=\mathcal{M}\cup\mathcal{N}\}}.

Indeed, since Iso𝒩,ψr\mathrm{Iso}^{\ast}_{\mathcal{N},\psi_{r}} is convolution by a probability measure, it is bounded by 11 as an operator LLL^{\infty}\to L^{\infty}. By a similar argument as in the proof of Proposition 2.1, we have Iso𝒩,ψr𝟏E,𝒩,r,M(XN)=0\mathrm{Iso}_{\mathcal{N},\psi_{r}}^{\ast}\mathbf{1}_{E_{\mathcal{M},\mathcal{N},r,M}}(X_{N})=0 if 𝕏(Ω5r)𝒩\mathbb{X}(\Omega_{5r})\neq\mathcal{M}\cup\mathcal{N}.

By isotropic averaging, we conclude, for |𝒩|=nN1|\mathcal{N}|=n\leq N_{1}, that

N,βW,U(E,𝒩,r,M)eβΔρn,r,MN,βW,U({𝕏(Ω5r)=𝒩})\mathbb{P}^{W,U}_{N,\beta}(E_{\mathcal{M},\mathcal{N},r,M})\leq e^{-\beta\Delta_{\rho_{n},r,M}}\mathbb{P}^{W,U}_{N,\beta}(\{\mathbb{X}(\Omega_{5r})=\mathcal{M}\cup\mathcal{N}\})

where ρn:=nμ(Ω)|Ω|\rho_{n}:=\frac{n-\mu(\Omega)}{|\Omega|} and Δρn,r,M\Delta_{\rho_{n},r,M} is as in (5.3). By Proposition 5.5, a union bound, and exchangeability, we have

N,βW,U(Eρ,r,M)\displaystyle\mathbb{P}^{W,U}_{N,\beta}(E_{\rho,r,M}) eαd+2Rd+2\displaystyle\leq e^{-\alpha^{d+2}R^{d+2}}
+m=0Mρn=nρN1(Nm)(Nmn)eβΔρn,r,MN,βW,U({𝕏(Ω5r)={1,,m+n}}).\displaystyle\quad+\sum_{m=0}^{M_{\rho}}\sum_{n=n_{\rho}}^{N_{1}}{\binom{N}{m}}{\binom{N-m}{n}}e^{-\beta\Delta_{\rho_{n},r,M}}\mathbb{P}^{W,U}_{N,\beta}(\{\mathbb{X}(\Omega_{5r})=\{1,\ldots,m+n\}\}).

Above, we only summed over \mathcal{M} and 𝒩\mathcal{N} disjoint. We have

({𝕏(Ω5r)={1,,m+n}})=1(Nm+n)({X(Ω5r)=m+n}})\mathbb{P}(\{\mathbb{X}(\Omega_{5r})=\{1,\ldots,m+n\}\})=\frac{1}{{\binom{N}{m+n}}}\mathbb{P}(\{X(\Omega_{5r})=m+n\}\})

and

(Nm)(Nmn)(Nm+n)=(m+n)!m!n!(n+m)mm!2nm\frac{{\binom{N}{m}}{\binom{N-m}{n}}}{{\binom{N}{m+n}}}=\frac{(m+n)!}{m!n!}\leq\frac{(n+m)^{m}}{m!}\leq 2n^{m}

whenever mnm\leq n, which for us is always the case if C5C_{5} is large enough (independently of CiC_{i}, i=2,3,4i=2,3,4). Letting j=m+nj=m+n, we have

N,βW,U(Eρ,r,M)\displaystyle\mathbb{P}^{W,U}_{N,\beta}(E_{\rho,r,M}) eαd+2Rd+2+n=nρN1j=nn+Mρ2njneβΔρn,r,MN,βW,U({X(Ω5r)=j})\displaystyle\leq e^{-\alpha^{d+2}R^{d+2}}+\sum_{n=n_{\rho}}^{N_{1}}\sum_{j=n}^{n+M_{\rho}}2n^{j-n}e^{-\beta\Delta_{\rho_{n},r,M}}\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega_{5r})=j\}) (5.30)
eαd+2Rd+2+n=nρN12nMρeβΔρn,r,M.\displaystyle\leq e^{-\alpha^{d+2}R^{d+2}}+\sum_{n=n_{\rho}}^{N_{1}}2n^{M_{\rho}}e^{-\beta\Delta_{\rho_{n},r,M}}.

Note that Mρ2MRd1r=2C51Rd1r3ρC51αRdr2ρ/log(N1)M_{\rho}\leq 2MR^{d-1}r=2C_{5}^{-1}R^{d-1}r^{3}\rho\leq C_{5}^{-1}\alpha R^{d}r^{2}\rho/\log(N_{1}). Thus for nN1n\leq N_{1} we have

nMρeMρlogN1eCC51r2ρ|Ω|.n^{M_{\rho}}\leq e^{M_{\rho}\log N_{1}}\leq e^{CC_{5}^{-1}r^{2}\rho|\Omega|}.

Since βΔρn,r,MC1r2ρ|Ω|\beta\Delta_{\rho_{n},r,M}\geq C^{-1}r^{2}\rho|\Omega| by Proposition 5.4, we can bound

nMρeβΔρn,r,Meβ2cr2ρn(μ(Ω)+ρn|Ω|)n^{M_{\rho}}e^{-\beta\Delta_{\rho_{n},r,M}}\leq e^{-\frac{\beta}{2}cr^{2}\rho_{n}(\mu(\Omega)+\rho_{n}|\Omega|)}

for some c>0c>0 for C5C_{5} large enough. The series in the RHS of (5.30) can therefore be dominated by a geometric series with rate 1/21/2 with the same first term, and so

N,βW,U(Eρ,r,M)eαd+2Rd+2+4eβ2cr2ρ(μ(Ω)+ρ|Ω|).\mathbb{P}^{W,U}_{N,\beta}(E_{\rho,r,M})\leq e^{-\alpha^{d+2}R^{d+2}}+4e^{-\frac{\beta}{2}cr^{2}\rho(\mu(\Omega)+\rho|\Omega|)}. (5.31)

Step 2: We now consider the event that X(Ω)μ(Ω)+ρ|Ω|X(\Omega)\geq\mu(\Omega)+\rho|\Omega| on the complement of Eρ,r,ME_{\rho,r,M}. It suffices to bound the probability that X(Ω5rΩ5r)MRd1r.X(\Omega_{5r}\setminus\Omega_{-5r})\geq MR^{d-1}r. Let {Aλ}λΛ\{A_{\lambda}\}_{\lambda\in\Lambda} be a covering of Ω5rΩ5r\Omega_{5r}\setminus\Omega_{-5r} by balls of radius rr of cardinality at most C(R/r)d1C(R/r)^{d-1}. We have

N,βW,U({X(Ω5rΩ5r)MRd1r})(CRr)d1supλΛN,βW,U({X(Aλ)C1Mrd}).\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega_{5r}\setminus\Omega_{-5r})\geq MR^{d-1}r\})\leq\left(\frac{CR}{r}\right)^{d-1}\sup_{\lambda\in\Lambda}\mathbb{P}^{W,U}_{N,\beta}(\{X(A_{\lambda})\geq C^{-1}Mr^{d}\}).

Note that (αR)2/3ρC4(\alpha R)^{2/3}\rho\geq C_{4} by (5.19), so we have MC51C4.M\geq C_{5}^{-1}C_{4}. If we take C4C5C_{4}\gg C_{5}, we have M1M\gg 1 and we may apply Theorem 1 to see

N,βW,U({X(Ω5rΩ5r)MRr})(CRr)d1eC1rd+6ρ2ecrd+6ρ2\mathbb{P}^{W,U}_{N,\beta}(\{X(\Omega_{5r}\setminus\Omega_{-5r})\geq MRr\})\leq\left(\frac{CR}{r}\right)^{d-1}e^{-C^{-1}r^{d+6}\rho^{2}}\leq e^{-cr^{d+6}\rho^{2}}

for some c>0c>0. Assembling (5.31) and the above finishes the proof. ∎

It is now a simple matter to prove Theorem 5.

Proof of Theorem 5.

We will apply Proposition 5.6 to the 11-thin annulus Ω=BR(z)\Omega=B_{R}(z) with a certain parameter ρ\rho. For the constant qq approximation to μ=1cdΔW\mu=\frac{1}{c_{d}}\Delta W, we take q=μ(z)q=\mu(z). If WW is quadratic near BR(z)B_{R}(z), this approximation is exact, but we focus on the general case. One finds that

μqL(B2R(z))CN1/dR,\|\mu-q\|_{L^{\infty}(B_{2R}(z))}\leq CN^{-1/d}R,

whence we have the restriction ρN1/dR\rho\gg N^{-1/d}R in applying Proposition 5.6. We also have the restriction ρR2/3(1+𝟏d=2logR)\rho\gg R^{-2/3}(1+\mathbf{1}_{d=2}\log R) from (5.19). If RN35d,R\leq N^{\frac{3}{5d}}, then the latter restriction is the only relevant one, and we achieve

N,βW,U({X(BR(z))μ(BR(z))+ρ|BR(z)|})ecRdT+ecR(d+2)/3T2+eRd+2,\mathbb{P}^{W,U}_{N,\beta}(\{X(B_{R}(z))\geq\mu(B_{R}(z))+\rho|B_{R}(z)|\})\leq e^{-cR^{d}T}+e^{-cR^{(d+2)/3}T^{2}}+e^{-R^{d+2}},

as desired, where we set ρ=TR2/3(1+𝟏d=2logR)\rho=TR^{-2/3}(1+\mathbf{1}_{d=2}\log R) for a large T>0T>0. ∎

5.3. Upgrading the discrepancy bound.

In this subsection, we upgrade Theorem 5 using rigidity results for smooth linear statistics. We will assume conditions stated in Theorem 6 throughout. We do not spend undue effort trying to optimize our bounds in β\beta or VV since our results are generally weaker than those of [AS21]. Instead, our purpose is to show how our overcrowding estimates can be upgraded using known rigidity bounds for smooth linear statistics. In particular, we demonstrate that overcrowding bounds are sufficient to bound the absolute discrepancy, rather than just the positive part of the discrepancy. The mechanism for this is consists of finding a screening region of excess positive charge near the boundary of a ball with large absolute discrepancy. We note that the idea of obtaining a screening region is already present and features prominently in [Leb21], and we do not add anything fundamentally new to this procedure, but rather adapt it for our overcrowding estimate.

For α(0,1]\alpha\in(0,1] and R(0,)R\in(0,\infty), let ξR,α:\xi_{R,\alpha}:\mathbb{R}\to\mathbb{R} be a function satisfying

  • 0ξR,α10\leq\xi_{R,\alpha}\leq 1

  • ξR,α(x)=1x[R,R]\xi_{R,\alpha}(x)=1\quad\forall x\in[-R,R] and ξR,α(x)=0x(RαR,R+αR)\xi_{R,\alpha}(x)=0\quad\forall x\not\in(-R-\alpha R,R+\alpha R)

  • ξR,α(x)0\xi^{\prime}_{R,\alpha}(x)\leq 0 for x0x\geq 0.

  • supx|ξR,α(k)(x)|Ck(αR)k\sup_{x\in\mathbb{R}}|\xi^{(k)}_{R,\alpha}(x)|\leq C_{k}(\alpha R)^{-k} for k=1,2,3,4k=1,2,3,4.

In what follows, we will consider the map xξR,α(|x|)x\mapsto\xi_{R,\alpha}(|x|) from d\mathbb{R}^{d}\to\mathbb{R}, and by abuse of notation we will write this map as ξR,α\xi_{R,\alpha}. We will also write Fluct(ϕ):=ϕ(x)fluct(dx)\mathrm{Fluct}(\phi):=\int\phi(x)\mathrm{fluct}(dx) for the (NN-dependent) fluctuation measure fluct\mathrm{fluct} defined in (1.4).

Theorem 8 (Corollary of [Ser20], Theorem 1).

Under the conditions on W=VNW=V_{N} stated above Theorem 6, for a large enough constant C>0C>0 and αRC\alpha R\geq C, we have

|log𝔼N,βVNexp(tFluct(ξR,α))|C|t|Rd2α3+Ct2(Rd4α4+Rd2α)+Ct4Rd8α8+C|t|RdN2/d.|\log\mathbb{E}^{V_{N}}_{N,\beta}\exp(t\mathrm{Fluct}(\xi_{R,\alpha}))|\leq\frac{C|t|R^{d-2}}{\alpha^{3}}+Ct^{2}\left(\frac{R^{d-4}}{\alpha^{4}}+\frac{R^{d-2}}{\alpha}\right)+\frac{Ct^{4}R^{d-8}}{\alpha^{8}}+\frac{C|t|R^{d}}{N^{2/d}}. (5.32)

for all |t|C1α2R2|t|\leq C^{-1}\alpha^{2}R^{2}. In particular for t=1t=1 and α3R2N2/d\alpha^{3}R^{2}\leq N^{2/d}, we have

|log𝔼N,βVNexp(Fluct(ξR,α))|C|t|Rd2α3.|\log\mathbb{E}^{V_{N}}_{N,\beta}\exp(\mathrm{Fluct}(\xi_{R,\alpha}))|\leq C\frac{|t|R^{d-2}}{\alpha^{3}}. (5.33)
Proof.

A more general estimate on a non-blown up scale, i.e. the scale with interstitial distance of order N1/dN^{-1/d}, and with the thermal equilibrium measure μθ\mu_{\theta} in place of the equilibrium measure μeq\mu_{\mathrm{eq}}, is [Ser20, Theorem 1]. We slightly transform the estimate by using βC1\beta\geq C^{-1}, changing the interstitial length scale to O(1)O(1), and plugging in the specifics of ξR,α\xi_{R,\alpha}. We also use (see [AS19, Theorem 1])

μθμeqL(d)CN2/d\|\mu_{\theta}-\mu_{\mathrm{eq}}\|_{L^{\infty}(\mathbb{R}^{d})}\leq\frac{C}{N^{2/d}}

to replace the thermal equilibrium measure by μeq\mu_{\mathrm{eq}}, generating the C|t|RdN2/dC|t|R^{d}N^{-2/d} term in (5.32). ∎

The following proposition shows that one can find a screening region of excess positive charge whenever the absolute discrepancy is large in a ball. It is a natural consequence of rigidity for fluctuations of smooth linear statistics.

Proposition 5.7.

Fix δ(0,d)\delta\in(0,d) and RCR\geq C for a large enough C>0C>0 and α(0,1]\alpha\in(0,1]. For any T1T\geq 1, if Fluct(ξRαR,α)T10Rδ\mathrm{Fluct}(\xi_{R-\alpha R,\alpha})\leq\frac{T}{10}R^{\delta}, we can find k0k\in\mathbb{Z}^{\geq 0} and a constant C6>0C_{6}>0 such that

Disc(Ann[RαkR,R](z))Disc(BR(z))T2Rδ\mathrm{Disc}(\mathrm{Ann}_{[R-\alpha_{k}R,R]}(z))\geq\mathrm{Disc}(B_{R}(z))-\frac{T}{2}R^{\delta}

for αk=αC61Rd+δk\alpha_{k}=\alpha-C_{6}^{-1}R^{-d+\delta}k with C61Rd+δkαkαC_{6}^{-1}R^{-d+\delta}k\leq\alpha_{k}\leq\alpha. If instead Fluct(ξR,α)T10Rδ\mathrm{Fluct}(\xi_{R,\alpha})\geq-\frac{T}{10}R^{\delta}, we can also find αk\alpha_{k} as above with

Disc(Ann[R,R+αkR](z))Disc(BR(z))T2Rδ.\mathrm{Disc}(\mathrm{Ann}_{[R,R+\alpha_{k}R]}(z))\geq-\mathrm{Disc}(B_{R}(z))-\frac{T}{2}R^{\delta}.
Proof.

First, note that

Disc(Bs+ε(z))Disc(Bs(z)))=Ann[s+ε,s)(z)fluct(dx)\mathrm{Disc}(B_{s+\varepsilon}(z))-\mathrm{Disc}(B_{s}(z)))=\int_{\mathrm{Ann}_{[s+\varepsilon,s)}(z)}\mathrm{fluct}(dx)

whenever fluct\mathrm{fluct} has no atoms on Bs(z)\partial B_{s}(z) or Bs+ε(z)\partial B_{s+\varepsilon}(z). Thus, we have by integration in spherical coordinates that

Fluct(ξRαR,α)\displaystyle\mathrm{Fluct}(\xi_{R-\alpha R,\alpha}) =0RξRαR,α(s)d(Disc(Bs(z)))=0RddsξRαR,α(s)Disc(Bs(z))𝑑s\displaystyle=\int_{0}^{R}\xi_{R-\alpha R,\alpha}(s)d(\mathrm{Disc}(B_{s}(z)))=-\int_{0}^{R}\frac{d}{ds}\xi_{R-\alpha R,\alpha}(s)\mathrm{Disc}(B_{s}(z))ds
=Disc(BR(z))RαRRddsξRαR,α(s)(Disc(Bs(z))Disc(BR(z)))𝑑s.\displaystyle=\mathrm{Disc}(B_{R}(z))-\int_{R-\alpha R}^{R}\frac{d}{ds}\xi_{R-\alpha R,\alpha}(s)\left(\mathrm{Disc}(B_{s}(z))-\mathrm{Disc}(B_{R}(z))\right)ds.

Assuming now Fluct(ξRαR,α)T10Rδ\mathrm{Fluct}(\xi_{R-\alpha R,\alpha})\leq\frac{T}{10}R^{\delta}, by the mean value theorem, there must exist s(RαR,R)s\in(R-\alpha R,R) such that Disc(Bs(z))Disc(BR(z))T10RδDisc(BR(z))\mathrm{Disc}(B_{s}(z))-\mathrm{Disc}(B_{R}(z))\leq\frac{T}{10}R^{\delta}-\mathrm{Disc}(B_{R}(z)). This means that

Disc(Ann[s,R)(z))Disc(BR(z))T10Rδ.\mathrm{Disc}(\mathrm{Ann}_{[s,R)}(z))\geq\mathrm{Disc}(B_{R}(z))-\frac{T}{10}R^{\delta}.

For any kk such that RαRRαkR<sR-\alpha R\leq R-\alpha_{k}R<s, we have

Disc(Ann(RαkR,R)(z))\displaystyle\mathrm{Disc}(\mathrm{Ann}_{(R-\alpha_{k}R,R)}(z)) Disc(Ann[s,R)(z))μeqN(Ann(Rαk,s](z))\displaystyle\geq\mathrm{Disc}(\mathrm{Ann}_{[s,R)}(z))-\mu_{\mathrm{eq}}^{N}(\mathrm{Ann}_{(R-\alpha_{k},s]}(z))
Disc(Ann[s,R)(z))CRd1|sαkR|.\displaystyle\geq\mathrm{Disc}(\mathrm{Ann}_{[s,R)}(z))-CR^{d-1}|s-\alpha_{k}R|.

We choose kk such that |s(RαkR)|C1Rd+1+δ|s-(R-\alpha_{k}R)|\leq C^{-1}R^{-d+1+\delta} for a large enough constant CC to conclude.

If we instead assume Fluct(ξR,α)T10Rδ\mathrm{Fluct}(\xi_{R,\alpha})\geq-\frac{T}{10}R^{\delta}, then since

Fluct(ξR,α)=Disc(BR(z))RR+αRddsξR,α(s)(Disc(Bs(z))Disc(BR(z)))𝑑s,\mathrm{Fluct}(\xi_{R,\alpha})=\mathrm{Disc}(B_{R}(z))-\int_{R}^{R+\alpha R}\frac{d}{ds}\xi_{R,\alpha}(s)\left(\mathrm{Disc}(B_{s}(z))-\mathrm{Disc}(B_{R}(z))\right)ds,

we can find s(R,R+αR)s\in(R,R+\alpha R) with

Disc(Ann[R,s)(z))T10RδDisc(BR(z)).\mathrm{Disc}(\mathrm{Ann}_{[R,s)}(z))\geq-\frac{T}{10}R^{\delta}-\mathrm{Disc}(B_{R}(z)).

Choosing αk\alpha_{k} such that sR+αkRR+αRs\leq R+\alpha_{k}R\leq R+\alpha R and |s(R+αkR)|C1Rd+1+δ|s-(R+\alpha_{k}R)|\leq C^{-1}R^{-d+1+\delta} is sufficient to conclude. ∎

We are now ready to prove Theorem 6.

Proof of Theorem 6.

Let =N,βVN\mathbb{P}=\mathbb{P}^{V_{N}}_{N,\beta}. We choose α=Rλ\alpha=R^{-\lambda} for λ=2/5\lambda=2/5 and consider discrepancies in BR(z)B_{R}(z) of size TRδ(1+𝟏d=2logR)TR^{\delta}(1+\mathbf{1}_{d=2}\log R) for δ=d4/5\delta=d-4/5 and TT sufficiently large. We will however write a mostly generic argument in terms of λ\lambda and δ\delta and insert the specific values later. We will consider the case of positive discrepancy first.

If Fluct(ξRαR,R)T10Rδ\mathrm{Fluct}(\xi_{R-\alpha R,R})\leq\frac{T}{10}R^{\delta}, we can use Proposition 5.7 to find αk(0,α]\alpha_{k}\in(0,\alpha] such that Disc(Ann[RαkR,R](z))T2Rδ(1+𝟏d=2logR)\mathrm{Disc}(\mathrm{Ann}_{[R-\alpha_{k}R,R]}(z))\geq\frac{T}{2}R^{\delta}(1+\mathbf{1}_{d=2}\log R). By a union bound, we have

({Disc(BR(z))TRδlogR})\displaystyle\mathbb{P}(\{\mathrm{Disc}(B_{R}(z))\geq TR^{\delta}\log R\}) (5.34)
({Fluct(ξRαR,α)>T10Rδ})+CαRdδsupαk({Disc(Ann[RαkR,R](z))T2RδlogR}),\displaystyle\leq\mathbb{P}(\{\mathrm{Fluct}(\xi_{R-\alpha R,\alpha})>\frac{T}{10}R^{\delta}\})+C\alpha R^{d-\delta}\sup_{\alpha_{k}}\mathbb{P}(\{\mathrm{Disc}(\mathrm{Ann}_{[R-\alpha_{k}R,R]}(z))\geq\frac{T}{2}R^{\delta}\log R\}),

where the supremum is over all αk=αC1Rd+δk\alpha_{k}=\alpha-C^{-1}R^{-d+\delta}k, kk\in\mathbb{Z} and αk[0,α)\alpha_{k}\in[0,\alpha). Note that we have αkC1Rd+δR1\alpha_{k}\geq C^{-1}R^{-d+\delta}\gg R^{-1} always. Next, we bound the supremum in (5.34).

Let α[C1Rd+δ,α]\alpha^{\prime}\in[C^{-1}R^{-d+\delta},\alpha] and let Ω\Omega be the α\alpha^{\prime}-thin annulus Ann[RαR,R](z)\mathrm{Ann}_{[R-\alpha^{\prime}R,R]}(z). We bound

({Disc(Ω)T2Rδ(1+𝟏d=2logR)})(X(Ω)μ(Ω)+ρ|Ω|)\mathbb{P}(\{\mathrm{Disc}(\Omega)\geq\frac{T}{2}R^{\delta}(1+\mathbf{1}_{d=2}\log R)\})\leq\mathbb{P}(X(\Omega)\geq\mu(\Omega)+\rho|\Omega|)

for ρ=C1T(α)1Rd+δ(1+𝟏d=2logR)\rho=C^{-1}T(\alpha^{\prime})^{-1}R^{-d+\delta}(1+\mathbf{1}_{d=2}\log R). Applying Proposition 5.6 shows

({Disc(Ω)T2Rδ(1+𝟏d=2logR)})ec(α)2/3R2/3+δT+ec(αR)d/3R2+2δ2dT2+e(αR)d+2\mathbb{P}(\{\mathrm{Disc}(\Omega)\geq\frac{T}{2}R^{\delta}(1+\mathbf{1}_{d=2}\log R)\})\leq e^{-c(\alpha^{\prime})^{2/3}R^{2/3+\delta}T}+e^{-c(\alpha^{\prime}R)^{d/3}R^{2+2\delta-2d}T^{2}}+e^{-(\alpha^{\prime}R)^{d+2}} (5.35)

whenever T(α)1/3Rδ+2/3dTRδ+2/3d+λ/3T(\alpha^{\prime})^{-1/3}R^{\delta+2/3-d}\geq TR^{\delta+2/3-d+\lambda/3} is large and we can estimate

μeqNqL(B2R(z))C21TRδd+λ\|\mu_{\mathrm{eq}}^{N}-q\|_{L^{\infty}(B_{2R}(z))}\leq C_{2}^{-1}TR^{\delta-d+\lambda}

for a constant qq. The latter condition happens when R1+dδλN1/dR^{1+d-\delta-\lambda}\ll N^{1/d}.

Considering the other term in (5.34), we apply Theorem 8 with |t|=C1|t|=C^{-1} and Chebyshev’s inequality to bound

({Fluct(ξRαR,α)>T10Rδ})eC1(T10Rδ+CRd2+3λ).\mathbb{P}(\{\mathrm{Fluct}(\xi_{R-\alpha R,\alpha})>\frac{T}{10}R^{\delta}\})\leq e^{C^{-1}(-\frac{T}{10}R^{\delta}+CR^{d-2+3\lambda})}. (5.36)

Note that α3R2=R4/5N2/d\alpha^{3}R^{2}=R^{4/5}\leq N^{2/d} so that (5.33) applies. We thus apply our argument to parameters λ\lambda and δ\delta such that

δ+23d+λ30,δd2+3λ.\delta+\frac{2}{3}-d+\frac{\lambda}{3}\geq 0,\quad\delta\geq d-2+3\lambda. (5.37)

One can check that the smallest choice of δ\delta is δ=d4/5\delta=d-4/5 with λ=2/5\lambda=2/5. With these choices, choosing RN57dR\leq N^{\frac{5}{7d}} guarantees that we can approximate μeqN\mu_{\mathrm{eq}}^{N} by a constant sufficiently well.

Finally, we estimate the RHS of (5.35) and (5.36) and plug them into (5.34). Note that αC1R4/5\alpha^{\prime}\geq C^{-1}R^{-4/5}, so the RHS of (5.35) is bounded by

ecRd10/15T+ecR2/5+d/15T2+eR(d+2)/5.e^{-cR^{d-10/15}T}+e^{-cR^{2/5+d/15}T^{2}}+e^{-R^{(d+2)/5}}.

The factor αRdδ\alpha R^{d-\delta} within (5.34) can be absorbed into the above at the cost of a constant factor within the exponent.

This finishes the one half of the proof of Theorem 6. The proof of the lower bound on Disc(BR(z))\mathrm{Disc}(B_{R}(z)) is nearly identical, except we use fluctuation bounds on ξR,α\xi_{R,\alpha} to find a screening region outside of BR(z)\partial B_{R}(z) with positive discrepancy. ∎

References

  • [AB13] Gérard Ben Arous and Paul Bourgade. Extreme gaps between eigenvalues of random matrices. The Annals of Probability, 41(4):2648 – 2681, 2013.
  • [AKM19] Scott Armstrong, Tuomo Kuusi, and Jean-Christophe Mourrat. Quantitative stochastic homogenization and large-scale regularity, volume 352. Springer, 2019.
  • [Ame18] Yacin Ameur. Repulsion in Low Temperature β\beta-Ensembles. Communications in Mathematical Physics, 359(3):1079–1089, May 2018.
  • [Ame21] Yacin Ameur. A localization theorem for the planar Coulomb gas in an external field. Electronic Journal of Probability, 26(none):1 – 21, 2021.
  • [AR22] Yacin Ameur and José Luis Romero. The planar low temperature Coulomb gas: separation and equidistribution, 2022.
  • [AS19] Scott Armstrong and Sylvia Serfaty. Thermal approximation of the equilibrium measure and obstacle problem, 2019.
  • [AS21] Scott Armstrong and Sylvia Serfaty. Local laws and rigidity for Coulomb gases at any temperature. The Annals of Probability, 49(1):46 – 121, 2021.
  • [BBNY19] Roland Bauerschmidt, Paul Bourgade, Miika Nikula, and Horng-Tzer Yau. The two-dimensional coulomb plasma: quasi-free approximation and central limit theorem. Advances in Theoretical and Mathematical Physics, 23(4):841–1002, 2019.
  • [BMP21] Paul Bourgade, Krishnan Mody, and Michel Pain. Optimal local law and central limit theorem for β\beta-ensembles. arXiv:1806.01831, 2021.
  • [Bou22] Paul Bourgade. Extreme gaps between eigenvalues of Wigner matrices. Journal of the European Mathematical Society, 24(8):2823 – 2873, 2022.
  • [CHM18] Djalil Chafaï, Adrien Hardy, and Mylène Maïda. Concentration for Coulomb gases and Coulomb transport inequalities. Journal of Functional Analysis, 275(6):1447–1483, 2018.
  • [Cif06] Orion Ciftja. Monte Carlo study of Bose Laughlin wave function for filling factors 1/2, 1/4 and 1/6. Europhysics Letters, 74(3):486, 2006.
  • [CSA20] Gabriel Cardoso, Jean-Marie Stéphan, and Alexander G Abanov. The boundary density profile of a Coulomb droplet. freezing at the edge. Journal of Physics A: Mathematical and Theoretical, 54(1):015002, Dec 2020.
  • [CW03] Orion Ciftja and Carlos Wexler. Monte Carlo simulation method for Laughlin-like states in a disk geometry. Phys. Rev. B, 67:075304, Feb 2003.
  • [DVJ08] D. J. Daley and D. Vere-Jones. An Introduction to the Theory of Point Processes. Probability and its Applications. Springer New York, 2nd edition, 2008.
  • [EKYY12] László Erdős, Antti Knowles, Horng-Tzer Yau, and Jun Yin. Spectral statistics of erdős-rényi graphs ii: Eigenvalue spacing and the extreme eigenvalues. Communications in Mathematical Physics, 314(3):587–640, Sep 2012.
  • [ESY10] László Erdős, Benjamin Schlein, and Horng-Tzer Yau. Wegner estimate and level repulsion for Wigner random matrices. International Mathematics Research Notices, 2010(3):436–479, 2010.
  • [FTW19] Renjie Feng, Gang Tian, and Dongyi Wei. Small gaps of GOE. Geometric and Functional Analysis, 29(6):1794–1827, Dec 2019.
  • [FW21] Renjie Feng and Dongyi Wei. Small gaps of circular β\beta-ensemble. The Annals of Probability, 49(2):997 – 1032, 2021.
  • [GLM80] Ch. Gruber, Ch. Lugrin, and Ph. A. Martin. Equilibrium properties of classical systems with long-range forces. BBGKY equation, neutrality, screening, and sum rules. Journal of Statistical Physics, 22(2):193–236, Feb 1980.
  • [JLM93] B Jancovici, Joel L Lebowitz, and G Manificat. Large charge fluctuations in classical Coulomb systems. Journal of statistical physics, 72(3):773–787, 1993.
  • [KK16] S. A. Khrapak and A. G. Khrapak. Internal energy of the classical two- and three-dimensional one-component-plasma. Contributions to Plasma Physics, 56(3-4):270–280, 2016.
  • [Leb21] Thomas Leblé. The two-dimensional one-component plasma is hyperuniform, 2021.
  • [LRY19] Elliott H. Lieb, Nicolas Rougerie, and Jakob Yngvason. Local incompressibility estimates for the Laughlin phase. Communications in Mathematical Physics, 365(2):431–470, Jan 2019.
  • [LS18] Thomas Leblé and Sylvia Serfaty. Fluctuations of two dimensional coulomb gases. Geometric and Functional Analysis, 28:443–508, 2018.
  • [Mar88] Ph. A. Martin. Sum rules in charged fluids. Rev. Mod. Phys., 60:1075–1127, Oct 1988.
  • [MW66] N. D. Mermin and H. Wagner. Absence of ferromagnetism or antiferromagnetism in one- or two-dimensional isotropic Heisenberg models. Phys. Rev. Lett., 17:1133–1136, Nov 1966.
  • [MY80] Ph. A. Martin and T. Yalcin. The charge fluctuations in classical Coulomb systems. Journal of Statistical Physics, 22(4):435–463, Apr 1980.
  • [NS15] Simona Rota Nodari and Sylvia Serfaty. Renormalized energy equidistribution and local charge balance in 2D Coulomb systems. International Mathematics Research Notices, 2015(11):3035–3093, 2015.
  • [NTV17] Hoi Nguyen, Terence Tao, and Van Vu. Random matrices: tail bounds for gaps between eigenvalues. Probability Theory and Related Fields, 167(3):777–816, Apr 2017.
  • [OR20] Alessandro Olgiati and Nicolas Rougerie. Stability of the Laughlin phase against long-range interactions. Archive for Rational Mechanics and Analysis, 237(3):1475–1515, 2020.
  • [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.
  • [Rou22a] Nicolas Rougerie. The classical jellium and the Laughlin phase. arXiv preprint arXiv:2203.06952, 2022.
  • [Rou22b] Nicolas Rougerie. On the stability of Laughlin’s fractional quantum Hall phase. arXiv preprint arXiv:2208.02619, 2022.
  • [RS16] Nicolas Rougerie and Sylvia Serfaty. Higher-dimensional Coulomb gases and renormalized energy functionals. Communications on Pure and Applied Mathematics, 69(3):519–605, 2016.
  • [RSY14] N. Rougerie, S. Serfaty, and J. Yngvason. Quantum Hall phases and plasma analogy in rotating trapped Bose gases. Journal of Statistical Physics, 154(1):2–50, Jan 2014.
  • [RY15a] Nicolas Rougerie and Jakob Yngvason. Incompressibility estimates for the Laughlin phase. Communications in Mathematical Physics, 336(3):1109–1140, Jun 2015.
  • [RY15b] Nicolas Rougerie and Jakob Yngvason. Incompressibility estimates for the Laughlin phase, part ii. Communications in Mathematical Physics, 339(1):263–277, Oct 2015.
  • [RY18] Nicolas Rougerie and Jakob Yngvason. The Laughlin liquid in an external potential. Letters in Mathematical Physics, 108:1007–1029, 2018.
  • [Ser20] Sylvia Serfaty. Gaussian fluctuations and free energy expansion for Coulomb gases at any temperature, 2020.
  • [SJ12] Dai Shi and Yunjiang Jiang. Smallest gaps between eigenvalues of random matrices with complex Ginibre, Wishart and universal unitary ensembles, 2012.
  • [Tao13] Terence Tao. The asymptotic distribution of a single eigenvalue gap of a Wigner matrix. Probability Theory and Related Fields, 157(1):81–106, Oct 2013.
  • [TV11] Terence Tao and Van Vu. Random matrices: Universality of local eigenvalue statistics. Acta Mathematica, 206(1):127 – 204, 2011.

Eric Thoma
Courant Institute, New York University.
Email: eric.thoma@cims.nyu.edu.