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

Boundary-layer analysis of repelling particles pushed to an impenetrable barrier

Patrick van Meurs
Abstract

This paper considers the equilibrium positions of nn particles in one dimension. Two forces act on the particles; a nonlocal repulsive particle-interaction force and an external force which pushes them to an impenetrable barrier. While the continuum limit as nn\to\infty is known for a certain class of potentials, numerical simulations show that a discrete boundary layer appears at the impenetrable barrier, i.e. the positions of o(n)o(n) particles do not fit to the particle density predicted by the continuum limit. In this paper we establish a first-order Γ\Gamma-convergence result which guarantees that these o(n)o(n) particles converge to a specific continuum boundary-layer profile.

keywords: discrete-to-continuum; boundary layers; Γ\Gamma-convergence; Γ\Gamma-development.

MSC: 74Q05, 74G10, 49J45, 82C22.

1 Introduction

This paper contributes to a recent trend in interacting particle systems which aims to find more detailed information on the particle positions at equilibrium than the information which the continuum limit provides. There are roughly two directions which are currently pursued; convergence rates and particle patterns on mesoscopic scales. The studies on convergence rates (see, e.g. [BO20, EOS16, HvMP20, KvM21, PZ20, TS19, vM18]) aim to find a topology in which the distance between the configuration of nn particles and the continuum particle density can be measured and bounded by a small value which vanishes as nn\to\infty. The studies on particle patterns zoom in on a mesoscopic scale, and reveal how the particles are distributed on this scale, either in the bulk (see the paper series started in [PS17, SS15]) or near the end of the support [GvMPS16, HCO10, HHvM18, Hud13]. This paper contributes to the latter, in which case we call the particle pattern a boundary layer. More precisely, this paper fills the important gap that was left open in [GvMPS16] on the characterization of boundary layers.

The gap in [GvMPS16]

To describe the gap in [GvMPS16], we first recall the corresponding setting. Consider n+1n+1 many particles (n1n\geq 1) confined to the half-line

Ω=[0,).\Omega=[0,\infty).

We label their positions as 𝐱:=(x0,x1,,xn)\mathbf{x}:=(x_{0},x_{1},\ldots,x_{n}) and assume that they are ordered, i.e. 𝐱Ωn\mathbf{x}\in\Omega_{n}, where

Ωn:={𝐱n+1:0=x0<x1<<xn}.\Omega_{n}:=\{\mathbf{x}\in\mathbb{R}^{n+1}:0=x_{0}<x_{1}<\ldots<x_{n}\}.

The discrete (i.e. n<n<\infty) particle interaction energy is given by

En:Ωn[0,),En(𝐱):=1n2i=1nj=0i1γnV(γn(xixj))+1ni=0nU(xi),E_{n}:\Omega_{n}\to[0,\infty),\qquad E_{n}(\mathbf{x}):=\frac{1}{n^{2}}\sum_{i=1}^{n}\sum_{j=0}^{i-1}\gamma_{n}V(\gamma_{n}(x_{i}-x_{j}))+\frac{1}{n}\sum_{i=0}^{n}U(x_{i}), (1)

where VV is an interaction potential, UU is a confining potential and γn>0\gamma_{n}>0 is a parameter. The double sum accounts for each pair of two particles. Figure 1 illustrates typical examples for VV and UU. The assumptions and properties of VV and UU are roughly as follows. UC1(Ω)U\in C^{1}(\Omega) with minΩU=0\min_{\Omega}U=0 and U(x)U(x)\to\infty as xx\to\infty. VL1()V\in L^{1}(\mathbb{R}) is normalized to V=1\int_{\mathbb{R}}V=1, nonnegative, even and singular at 0 with the bound

V(x)C{|x|aif a>0log|x|if a=0for all x(0,12)V(x)\leq C\left\{\begin{array}[]{ll}|x|^{-a}&\text{if }a>0\\ -\log|x|&\text{if }a=0\end{array}\right.\qquad\text{for all }x\in(0,\tfrac{1}{2}) (2)

for some C>0C>0, where a[0,1)a\in[0,1) is a parameter which bounds the strength of the singularity. We also assume that V|(0,)V|_{(0,\infty)} is non-increasing and convex. We state the precise list of assumptions and resulting properties in Section 3. With respect to [GvMPS16], we put more assumptions on VV, but allow for a general confining potential UU instead of the specific choice U(x)=xU(x)=x. This generalization of UU does not result in further complications. Instead, it clarifies the dependence of UU on the boundary layer.

xx0V(x)V(x)0xxU(x)U(x)
Figure 1: Typical examples of VV and UU.

One particular choice of VV which we have in mind is

Vwall(x)=xcothxlog|2sinhx|.V_{\operatorname{wall}}(x)=x\coth x-\log|2\sinh x|. (3)

For this potential, EnE_{n} is a model for the pile-up of dislocation walls at a lock. We refer to [GvMPS16] for the discussion of this model in the literature and its physical relevance. We show in Section 3 that it satisfies all the assumptions that we put on VV.

The asymptotic behaviour of the parameter γn\gamma_{n} in (1) as nn\to\infty plays a decisive role for the limiting energy EE of EnE_{n} as nn\to\infty. This can be expected from (1) by noting that the scaled potential

Vγ(x)=γV(γx),V_{\gamma}(x)=\gamma V(\gamma x), (4)

which has unit integral for all γ>0\gamma>0, is squeezed to a delta-peak at 0 as γ\gamma\to\infty. Hence, as γ\gamma increases, the particle interactions become more localized. In [GPPS13] the Γ\Gamma-limit of EnE_{n} is obtained as nn\to\infty. Depending on the asymptotic behavior of γn\gamma_{n}, five different limiting energies are obtained: two of these belong to the critical scaling regimes γnγ>0\gamma_{n}\to\gamma>0 and γn/nΓ>0\gamma_{n}/n\to\Gamma>0 as nn\to\infty, and the other three belong to the three regimes separated by these two critical regimes (the outer two regimes require a rescaling of EnE_{n} and 𝐱\mathbf{x}; we refer to [GPPS13] for the details).

In this paper we focus on the regime in between the two critical ones, i.e.

limnγn=andlimnγnn=0.\lim_{n\to\infty}\gamma_{n}=\infty\quad\text{and}\quad\lim_{n\to\infty}\frac{\gamma_{n}}{n}=0. (5)

In this regime the Γ\Gamma-limit of EnE_{n} (see [GPPS13, Thm. 7]) is given by

E:𝒫(Ω)[0,],E(μ)=12μL2(Ω)2+ΩU(x)𝑑μ(x),E:\mathcal{P}(\Omega)\to[0,\infty],\qquad E(\mu)=\frac{1}{2}\|\mu\|_{L^{2}(\Omega)}^{2}+\int_{\Omega}U(x)\,d\mu(x), (6)

where 𝒫(Ω)\mathcal{P}(\Omega) is the space of probability measures on Ω\Omega. The L2L^{2}-norm is extended to measures (see (27) for details) and may be infinite. It is well-known (see, e.g. [KS80, Thm. 2.1] with minor modications to account for UL2(Ω)U\notin L^{2}(\Omega)) that the minimization problem of EE over 𝒫(Ω)\mathcal{P}(\Omega) has a unique minimizer μL2(Ω)𝒫(Ω)\mu_{*}\in L^{2}(\Omega)\cap\mathcal{P}(\Omega) and that its density ρ\rho_{\ast} is characterized by

{ρ+UCUon Ωρ+U=CUon suppρ,\left\{\begin{aligned} \rho_{\ast}+U&\geq C_{U}\quad\text{on }\Omega\\ \rho_{\ast}+U&=C_{U}\quad\text{on }\operatorname{supp}\rho_{\ast},\\ \end{aligned}\right. (7)

where the constant CU>0C_{U}>0 is such that Ωρ(x)𝑑x=1\int_{\Omega}\rho_{\ast}(x)\,dx=1. Obviously,

ρ=[CUU]+.\rho_{\ast}=[C_{U}-U]^{+}. (8)

Figure 2 illustrates ρ\rho_{\ast}.

0xxCUC_{U}ρ\rho_{*}U(x)U(x)
Figure 2: Geometrical interpretation of ρ\rho_{*} in (8). ρ\rho_{*} is the height function of the region (colored in gray) with unit area below CUC_{U} and above UU.

As pointed out in [GvMPS16], the Γ\Gamma-convergence of EnE_{n} is not completely satisfactory, because it does not detect any particle patterns on mesoscopic scales. For the scaling regime in (5), the numerical computations of the minimizer 𝐱\mathbf{x}_{*} in [GvMPS16] indicate that O(n/γn)O(n/\gamma_{n}) particles are not distributed according to ρ\rho_{*}; see Figure 3. The main result [GvMPS16, Thm. 1.1] captures the continuum boundary-layer profile according to which these O(n/γn)O(n/\gamma_{n}) particles are distributed. This profile is obtained by firstly proving a first-order Γ\Gamma-convergence result for the continuous counterpart of EnE_{n} (see (9) below) and by secondly minimizing the first-order Γ\Gamma-limit. While Figure 3 suggests strongly that the obtained boundary-layer profile accurately describes the discrete boundary layer, any proof for this observation was left open. This is the gap in [GvMPS16] which we aim to fill in this paper.

Refer to caption
Figure 3: This is a copy of [GvMPS16, Fig. 4]; copyright by World Scientific Publishing Co., Inc. The Figure illustrates the minimizers 𝐱\mathbf{x}_{*} and ρ\rho_{*} of respectively EnE_{n} and EE for the potentials VV as in (3) and U(x)=xU(x)=x. (a) is obtained from (b) by zooming in with the scaling operator (γn)(\gamma_{n})_{\to} defined in (10). The crosses ρn\rho_{n} illustrate the discrete density profile of 𝐱\mathbf{x}_{*}. The xx- and yy-coordinates of these crosses are respectively x,ix_{*,i} and 2n(x,i+1x,i1)1\frac{2}{n}(x_{*,i+1}-x_{*,i-1})^{-1}. ργn\rho_{*}^{\gamma_{n}} is the continuum boundary-layer profile; see (13).

The first-order Γ\Gamma-convergence result of [GvMPS16]

In order to describe this paper’s first-order Γ\Gamma-convergence result which will fill the gap in [GvMPS16], we first recall that of [GvMPS16]. By [GPPS13, Thm. 5] the Γ\Gamma-limit of EnE_{n} in the regime γnγ>0\gamma_{n}\to\gamma>0 is given by

Eγ:𝒫(Ω)[0,],Eγ(μ):=12ΩΩVγ(xy)𝑑μ(y)𝑑μ(x)+ΩU(x)𝑑μ(x),E^{\gamma}:\mathcal{P}(\Omega)\to[0,\infty],\qquad E^{\gamma}(\mu):=\frac{1}{2}\int_{\Omega}\int_{\Omega}V_{\gamma}(x-y)\,d\mu(y)d\mu(x)+\int_{\Omega}U(x)\,d\mu(x), (9)

where VγV_{\gamma} is defined in (4). The energy EγE^{\gamma} is the continuous counterpart of EnE_{n} which is considered in [GvMPS16]. It Γ\Gamma-converges to EE as γ\gamma\to\infty (see [GvMPS16, Thm. 2.1]). In particular, this means that there exists a sequence (μγ)γ(\mu^{\gamma})_{\gamma} such that

Eγ(μγ)E(ρ)=o(1)as γ.E^{\gamma}(\mu^{\gamma})-E(\rho_{*})=o(1)\qquad\text{as }\gamma\to\infty.

The idea of the authors of [GvMPS16] to upgrade this to a first-order Γ\Gamma-convergence result was to characterize the o(1)o(1)-term. They predicted from a priori computations that this term is O(1/γ)O(1/\gamma), and that it is easier to replace E(ρ)E(\rho_{*}) by Eγ(ρ)E^{\gamma}(\rho_{*}). This motivated them to consider the functional

Fγ(μ):=γ(Eγ(μ)Eγ(ρ)).F^{\gamma}(\mu):=\gamma\big{(}E^{\gamma}(\mu)-E^{\gamma}(\rho_{*})\big{)}.

They call FγF^{\gamma} the boundary-layer energy. The first-order Γ\Gamma-convergence of EγE^{\gamma} is simply the (zeroth-order) Γ\Gamma-convergence of FγF^{\gamma}.

For the Γ\Gamma-convergence of FγF^{\gamma} the topology needed to be chosen carefully. From the formal asymptotics in [Hal11] and their own numerical simulations the authors guessed that the width of the boundary layer is O(1/γ)O(1/\gamma). This motivated them to use the following spatial rescaling. For a measure μ𝒫(Ω)\mu\in\mathcal{P}(\Omega), let

μ~:=γμ:=γ(γid)#μ.\tilde{\mu}:=\gamma_{\to}\mu:=\gamma\;(\gamma\operatorname{id})_{\#}\mu. (10)

The inverse scaling is given by

μ:=γμ~:=1γ(1γid)#μ~.\mu:=\gamma_{\leftarrow}\tilde{\mu}:=\frac{1}{\gamma}\Big{(}\frac{1}{\gamma}\operatorname{id}\Big{)}_{\#}\tilde{\mu}.

Note that if μ\mu has a density ρ\rho, then the density of μ~\tilde{\mu} satisfies

ρ~(x)=ρ(x/γ)=:γρ(x).\tilde{\rho}(x)=\rho(x/\gamma)=:\gamma_{\to}\rho(x).

Using this scaling, the authors of [GvMPS16] employed the following change of variables:

νγ:=μ~ρ~,μ=γνγ+ρ.\nu^{\gamma}:=\tilde{\mu}-\tilde{\rho}_{*},\qquad\mu=\gamma_{\leftarrow}\nu^{\gamma}+\rho_{\ast}.

By subtracting ρ\rho_{*} the bulk behaviour gets separated from the boundary layer.

For the signed Radon measures νγ\nu^{\gamma} with total variation that growths linearly with γ\gamma, the authors used the vague topology. This topology is defined as follows on the space (Ω)\mathcal{M}(\Omega) of signed Radon measures on Ω\Omega. A sequence (νε)ε>0(Ω)(\nu_{\varepsilon})_{\varepsilon>0}\subset\mathcal{M}(\Omega) converges to ν(Ω)\nu\in\mathcal{M}(\Omega) vaguely (denoted by νεvν\nu_{\varepsilon}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu) as ε0\varepsilon\to 0 if

Ωφ𝑑νεε0Ωφ𝑑νfor all φCc(Ω).\int_{\Omega}\varphi\,d\nu_{\varepsilon}\xrightarrow{\varepsilon\to 0}\int_{\Omega}\varphi\,d\nu\qquad\text{for all }\varphi\in C_{c}(\Omega).

The main result [GvMPS16, Thm. 1.1] states that FγF^{\gamma} Γ\Gamma-converges with respect to the vague topology to a certain limiting boundary-layer energy FF. This functional FF is defined on

𝒜={ν(Ω)ν(dx)ρ(0)dx,supx0ν+([x,x+1])<},\mathcal{A}=\Big{\{}\nu\in\mathcal{M}(\Omega)\mid\nu^{-}(dx)\leq\rho_{\ast}(0)dx,\ \sup_{x\geq 0}\nu^{+}([x,x+1])<\infty\Big{\}}, (11)

where ν+,ν0\nu^{+},\nu^{-}\geq 0 are respectively the positive and negative part of ν\nu such that ν=ν+ν\nu=\nu^{+}-\nu^{-}. While ν𝒜\nu\in\mathcal{A} may have infinite total variation, we have that νL(Ω)\nu^{-}\in L^{\infty}(\Omega) and that the local bound on ν+\nu^{+} is translation invariant. For ν𝒜L2(Ω)\nu\in\mathcal{A}\cap L^{2}(\Omega),

F(ν):=12ΩΩV(xy)ν(y)ν(x)𝑑y𝑑xρ(0)0(Vν)(x)𝑑x.F(\nu):=\frac{1}{2}\int_{\Omega}\int_{\Omega}V(x-y)\nu(y)\nu(x)\,dydx-\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu)(x)\,dx. (12)

It is not obvious to extend this definition to ν𝒜\nu\in\mathcal{A}, because ν\nu need not be of finite total variation. We recall this extension briefly in Section 5.1. Finally, [GvMPS16] noted that FF has a unique minimizer ν\nu_{*} (existence follows from the Γ\Gamma-convergence result in [GvMPS16] and uniqueness follows from the convexity of 𝒜\mathcal{A} and the strict convexity of FF), and that the continuous boundary-layer profile is given by

ρ~γ:=ν+ρ~,ργ:=γν+ρ.\tilde{\rho}_{*}^{\gamma}:=\nu_{*}+\tilde{\rho}_{*},\qquad\rho_{*}^{\gamma}:=\gamma_{\leftarrow}\nu_{*}+\rho_{*}. (13)

Figure 3 and all other numerical simulations performed in [GvMPS16] suggest that ργ\rho_{*}^{\gamma} gives a very good prediction for both the bulk and the boundary layer in the minimizer 𝐱\mathbf{x}_{*}.

However, the match between 𝐱\mathbf{x}_{*} and ργ\rho_{*}^{\gamma} has only been observed and has not been proven. Hence, there is no guarantee that such a match extrapolates to any other choices for the potentials UU and VV and for the parameter γn\gamma_{n}. This motivates our aim to establish a first-order Γ\Gamma-convergence result for the discrete energy EnE_{n} instead of its continuous counterpart EγnE^{\gamma_{n}}.

First-order Γ\Gamma-convergence result of EnE_{n}

To establish a first-order Γ\Gamma-convergence result for EnE_{n}, we follow largely the same setup as the one just described. In fact, from Figure 3 we expect the same limiting boundary-layer energy FF. Also, there is a close connection between EnE_{n} and EγnE^{\gamma_{n}}, which can be seen as follows. Given 𝐱Ωn\mathbf{x}\in\Omega_{n}, consider the corresponding empirical measure

μn:=1ni=0nδxin+1n𝒫(Ω).\mu_{n}:=\frac{1}{n}\sum_{i=0}^{n}\delta_{x_{i}}\in\frac{n+1}{n}\mathcal{P}(\Omega). (14)

Then, we can express En(𝐱)E_{n}(\mathbf{x}) in terms of μn\mu_{n} as

En(μn):=12ΔcVγn(xy)𝑑μn(y)𝑑μn(x)+ΩU(x)𝑑μn(x),E_{n}(\mu_{n}):=\frac{1}{2}\iint_{\Delta^{c}}V_{\gamma_{n}}(x-y)\,d\mu_{n}(y)d\mu_{n}(x)+\int_{\Omega}U(x)\,d\mu_{n}(x), (15)

where the diagonal

Δ={(x,x)T:x}2\Delta=\{(x,x)^{T}:x\in\mathbb{R}\}\subset\mathbb{R}^{2}

is removed from the integration domain to avoid self-interactions. Apart from removing the diagonal, the expressions for EnE_{n} and EγnE^{\gamma_{n}} are the same. Yet, the removal of the diagonal and the difference in the admissible sets on which EnE_{n} and EγnE^{\gamma_{n}} are defined are crucial. Indeed, (x,y)Vγn(xy)(x,y)\mapsto V_{\gamma_{n}}(x-y) concentrates around the diagonal as nn\to\infty and thus careful analysis is required.

Following the procedure from [GvMPS16], we consider the blown-up energy difference γn[En(μn)Eγn(ρ)]\gamma_{n}[E_{n}(\mu_{n})-E^{\gamma_{n}}(\rho_{*})] and employ the following change of variables. For μn\mu_{n} as in (14), we set

νn:=μ~nρ~,μn=(γn)νn+ρ.\nu_{n}:=\tilde{\mu}_{n}-\tilde{\rho}_{*},\qquad\mu_{n}=(\gamma_{n})_{\leftarrow}\nu_{n}+\rho_{\ast}. (16)

Then, the discrete boundary-layer energy FnF_{n} is defined on the admissible set

𝒜n:={νn(Ω)νn=ρ~,𝐲Ωn:νn+=γnni=0nδyi}\mathcal{A}_{n}:=\bigg{\{}\nu_{n}\in\mathcal{M}(\Omega)\mid\nu_{n}^{-}=\tilde{\rho}_{*},\ \exists\,\mathbf{y}\in\Omega_{n}:\nu_{n}^{+}=\frac{\gamma_{n}}{n}\sum_{i=0}^{n}\delta_{y_{i}}\bigg{\}} (17)

and given by

Fn(νn):=γn[En((γn)νn+ρ)+Eγn(ρ)].F_{n}(\nu_{n}):=\gamma_{n}\big{[}E_{n}\big{(}(\gamma_{n})_{\leftarrow}\nu_{n}+\rho_{\ast}\big{)}+E^{\gamma_{n}}(\rho_{\ast})\big{]}. (18)

Note that if νn\nu_{n} is constructed from μn\mu_{n} by (16), then

Fn(νn)=γn(En(μn)Eγn(ρ)).F_{n}(\nu_{n})=\gamma_{n}\left(E_{n}(\mu_{n})-E^{\gamma_{n}}(\rho_{\ast})\right). (19)

The main result of this paper in the following Γ\Gamma-convergence result of FnF_{n}:

Theorem 1.1.

Let UU and VV satisfy Assumptions 3.1 and 3.2, and let a[0,1)a\in[0,1) be such that (2) holds. If

1γn{n1a2aif 0<a<1nlognif a=0,1\ll\gamma_{n}\ll\left\{\begin{aligned} &n^{\tfrac{1-a}{2-a}}&&\text{if }0<a<1\\ &\sqrt{\frac{n}{\log n}}&&\text{if }a=0,\end{aligned}\right.

then any sequence (νn)n=1(\nu_{n})_{n=1}^{\infty} with νn𝒜n\nu_{n}\in\mathcal{A}_{n} and supn1Fn(νn)<\sup_{n\geq 1}F_{n}(\nu_{n})<\infty is pre-compact in 𝒜\mathcal{A} in the vague topology. Moreover, the functionals FnF_{n} Γ\Gamma-converge to FF with respect to the vague topology, i.e.

ν𝒜νn𝒜n with νnvν:\displaystyle\forall\,\nu\in\mathcal{A}\ \forall\,\nu_{n}\in\mathcal{A}_{n}\text{ with }\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu: lim infnFn(νn)\displaystyle\liminf_{n\to\infty}F_{n}(\nu_{n}) F(ν)\displaystyle\geq F(\nu)
ν𝒜νn𝒜n with νnvν:\displaystyle\forall\,\nu\in\mathcal{A}\ \exists\,\nu_{n}\in\mathcal{A}_{n}\text{ with }\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu: lim supnFn(νn)\displaystyle\limsup_{n\to\infty}F_{n}(\nu_{n}) F(ν).\displaystyle\leq F(\nu).

More precisely, the assumption on γn\gamma_{n} is equivalent to

limnγn=and{limnγnn1a2a=0if 0<a<1limnγnlognn=0if a=0.\lim_{n\to\infty}\gamma_{n}=\infty\quad\text{and}\quad\left\{\begin{aligned} \lim_{n\to\infty}\gamma_{n}n^{-\tfrac{1-a}{2-a}}&=0&&\text{if }0<a<1\\ \lim_{n\to\infty}\gamma_{n}\sqrt{\frac{\log n}{n}}&=0&&\text{if }a=0.\end{aligned}\right.

The proof of Theorem 1.1 is given in Section 6 with preliminaries in Sections 4 and 5. It follows the proof in [GvMPS16] with major modifications to allow for the discreteness. Here, we briefly describe the main features of Theorem 1.1 and focus in particular on these major modifications.

First, we recall from [GvMPS16] that the expression for FF in (12) arises naturally when the right-hand side in (18) is explicitly expressed in terms of νn\nu_{n}. In Section 4 we redo this computation, which in our case deals with the discrete setting and with a general confining potential UU.

The main difficulty with respect to [GvMPS16] is that the diagonal Δ\Delta is removed from the integration domain (see (15)) and that the domain of FnF_{n} is discrete (i.e. the degrees of freedom are empirical measures). To deal with this, we use essentially the particular regularization VβV^{\beta} of VV constructed in [KvM21], which approximates VV from below as β0\beta\to 0. Using this regularization, we add and subtract the contribution of the diagonal. By adding the diagonal, we can apply similar arguments as those in [GvMPS16] to establish the liminf inequality. However, β\beta needs to be chosen carefully. If β\beta is too small, then the contribution of the diagonal is too large and may not vanish in the limit. On the other hand, if β\beta is too large, then we cannot control the error made by the replacement of VV by VβV^{\beta}. Balancing out these two errors results in the asymptotic upper bound on γn\gamma_{n} in Theorem 1.1. This bound is a stronger requirement than in (5), which is sufficient for the (zeroth-order) Γ\Gamma-convergence of EnE_{n} to EE.

Establishing the limsup inequality is also significantly more challenging than in [GvMPS16]. The discreteness of 𝒜n\mathcal{A}_{n} forces us to discretize ν\nu, which was not necessary in the continuous setting in [GvMPS16]. Since we blow up the energy difference by the factor γn\gamma_{n}, we need to show that the discretization error is asymptotically smaller than 1/γn1/\gamma_{n}. This is much more intricate than for the zeroth-order Γ\Gamma-limit of EnE_{n} (see [GPPS13, Thm. 7]), where it was sufficient to show that the discretization error simply vanishes as nn\to\infty.

Discussion

In conclusion, Theorem 1.1 extends its continuous counterpart [GvMPS16, Thm. 1.1] (i.e. the Γ\Gamma-convergence of FγF^{\gamma} to FF) in two manners. First, on a minor note, it allows for a general confining potential UU. This highlights the fact that the dependence of FF on UU is restricted to the single value ρ(0)0\rho_{\ast}(0)\geq 0, which depends nonlocally on UU (see (8) and Figure 2). Second, on a major note, Theorem 1.1 considers the discrete energy FnF_{n}. As a consequence of Theorem 1.1, any sequence of minimizers ν,n\nu_{*,n} of FnF_{n} converges to ν\nu_{*}. Since this convergence happens on the mesoscopic scale of the boundary layer, this proves that ργn\rho_{*}^{\gamma_{n}} (see (13)) indeed describes the boundary layer which appears in 𝐱\mathbf{x}_{*}. This gives the first theoretical motivation for the observations in Figure 3 and any other numerical computation in [GvMPS16] that fits to the regime of γn\gamma_{n} assumed in Theorem 1.1. This fills the main gap that was left open in [GvMPS16].

Yet, the story is not complete; [GvMPS16] contains a number of conjectures sparked by numerical simulations to which Theorem 1.1 does not provide an answer. Here, we focus on the main limitation of Theorem 1.1, which is the upper bound on γn\gamma_{n}. Indeed, the numerical simulations in [GvMPS16] suggest that ργn\rho_{*}^{\gamma_{n}} is the correct boundary-layer profile for the whole regime of γn\gamma_{n} in (5). However, [GvMPS16, Table 1] suggests that the anticipated scaling of the energy difference, i.e. En(𝐱)Eγn(ρ)1/γnE_{n}(\mathbf{x}_{*})-E^{\gamma_{n}}(\rho_{\ast})\sim 1/\gamma_{n}, ceases to hold at the upper bound on γn\gamma_{n} in Theorem 1.1. Hence, this upper bound is not simply an artefact of our proof. Looking deeper into the proof in Section 6, it seems that this upper bound is caused by the contribution to FnF_{n} from a narrow region around the diagonal in the double integral in (15). A more precise treatment of this diagonal region could perhaps reveal a contribution of the right-hand side in (19) which diverges to \infty as nn\to\infty. Specifying this contribution, subtracting it from FnF_{n} and proving Γ\Gamma-convergence of the resulting energy functional (provided that his is possible) would reveal that ργn\rho_{*}^{\gamma_{n}} remains the correct boundary layer profile beyond the upper bound on γn\gamma_{n} in Theorem 1.1. Pursuing this direction is beyond our scope.

Organization of the paper

In Section 2 we set the notation. In Section 3 we state the precise assumptions on the potentials VV and UU, and derive further properties that follow from these assumptions. In Section 4 we rewrite FnF_{n} defined in (18) explicitly in terms of νn\nu_{n}, which will clarify the connection with the expression for FF in (12). In Section 5 we build the functional setting on which our proof of Theorem 1.1 relies. We also provide several a priori estimates. Finally, Section 6 is devoted to the proof of Theorem 1.1.

2 Notation

Here we list some symbols and abbreviations that we use throughout the paper.

aa smallest constant such that V(x)C|x|aV(x)\leq C|x|^{-a} As. 3.2(ii)
𝒜n\mathcal{A}_{n}, 𝒜\mathcal{A} admissible sets for FnF_{n} and FF (17), (11)
β\beta regularization parameter for VV and TT (36)
γn\gamma_{n} modelling parameter Thm. 1.1
γμ\gamma_{\to}\mu, γμ\gamma_{\leftarrow}\mu transforms of μ\mu by scaling space by γ>0\gamma>0 (10)
EnE_{n} discrete energy (1)
EγE^{\gamma} Γ\Gamma-limit of EnE_{n} for γn=γ\gamma_{n}=\gamma (9)
EE Γ\Gamma-limit of EnE_{n} for 1γnn1\ll\gamma_{n}\ll n (6)
f^\widehat{f}, f\mathcal{F}f Fourier transform of ff;
(f)(ω)=f^(ω):=f(x)e2πixω𝑑x(\mathcal{F}f)(\omega)=\widehat{f}(\omega):=\int_{\mathbb{R}}f(x)e^{-2\pi ix\omega}\,dx
1f\mathcal{F}^{-1}f inverse Fourier transform of ff;
FnF_{n} discrete boundary-layer energy (18), (24)
FF continuum boundary-layer energy (12), (30)
\mathcal{L} the Lebesgue measure on Ω\Omega; (Ω)\mathcal{L}\in\mathcal{M}(\Omega)
(Ω)\mathcal{M}(\Omega) signed Radon measures on \mathbb{R} with support in Ω\Omega
ν+,ν\nu^{+},\nu^{-} positive and negative part of a
measure ν(Ω)\nu\in\mathcal{M}(\Omega); ν±0\nu^{\pm}\geq 0
𝒫(Ω)\mathcal{P}(\Omega) 𝒫(Ω)(Ω)\mathcal{P}(\Omega)\subset\mathcal{M}(\Omega) is the set of probability measures
ρ\rho_{\ast} minimizer of EE (8)
ρ~\tilde{\rho}_{\ast} rescaled version; ρ~(x):=(γn)ρ(x)=ρ(x/γn)\tilde{\rho}_{\ast}(x):=(\gamma_{n})_{\to}\rho_{*}(x)=\rho_{*}(x/\gamma_{n})
TT ‘convolutional square root’ of VV; T2f=VfT^{2}f=V\ast f (28), Lem. 5.1
Xk,jX_{k,j} Hilbert space; Xk,jL2()X_{k,j}\subset L^{2}(\mathbb{R}) (32)
p\|\cdot\|_{p} Lp()L^{p}(\mathbb{R})-norm; 1p1\leq p\leq\infty.

We reserve c,C>0c,C>0 for generic constants which do not depend on any of the relevant variables. We use CC in upper bounds (and think of it as possibly large) and cc in lower bounds (and think of it as possibly small). While c,Cc,C may vary from line to line, in the same display they refer to the same value. If different constants appear in the same display, we denote them by C,C,C′′,C,C^{\prime},C^{\prime\prime},\ldots.

To avoid clutter, we often omit the integration variable. For instance, we use

ΩU𝑑ρ\displaystyle\int_{\Omega}U\,d\rho :=ΩU(x)𝑑ρ(x)\displaystyle:=\int_{\Omega}U(x)\,d\rho(x)
V\displaystyle\int_{\mathbb{R}}V :=V(x)𝑑x\displaystyle:=\int_{\mathbb{R}}V(x)\,dx
(Vνn)(x)\displaystyle(V*\nu_{n})(x) :=ΩV(xy)𝑑νn(y),\displaystyle:=\int_{\Omega}V(x-y)\,d\nu_{n}(y),

and extrapolate this notation to other integrands. Other than the framework of measures, we will also work with distributions. To connect the two notions, we often interpret measures on Ω\Omega as distributions on \mathbb{R} supported in Ω\Omega.

3 The potentials UU and VV

To the potential UU we add one more assumption to those mentioned in the introduction. We recall that CUC_{U} is the constant in (7); see also Figure 2.

Assumption 3.1.

UC1(Ω)U\in C^{1}(\Omega) satisfies minΩU=0\min_{\Omega}U=0 and U(x)U(x)\to\infty as xx\to\infty. Moreover, there exist finitely many disjoint closed intervals I1,,ImI_{1},\ldots,I_{m} such that

supp[CUU]+=i=1mIi.\operatorname{supp}[C_{U}-U]^{+}=\bigcup_{i=1}^{m}I_{i}. (20)

The assumption minΩU=0\min_{\Omega}U=0 is not restrictive, as otherwise one can achieve this by adding a constant to EnE_{n}. The assumption (20) is technical; it excludes pathological cases in which the graph of UU crosses the value CUC_{U} infinitely many times. In fact, for the choice U(x)=xU(x)=x in [GvMPS16], (20) holds for m=1m=1. For UU as in Figure 2, (20) holds for m=2m=2.

In view of (8), Assumption 3.1 directly translates to assumptions on ρ\rho_{\ast}, independent of the assumptions on VV. Indeed, from (8) it is clear that Assumption 3.1 implies that ρ\rho_{\ast} is Lipschitz continuous, and that ρ\rho_{*}^{\prime} is uniformly continuous on Ω(suppρ)\Omega\setminus\partial(\operatorname{supp}\rho_{*}). Hence, (20) implies that only at finitely many points ρ\rho_{*} is not of class C1C^{1}.

Next we turn to the potential VV:

Assumption 3.2.

VC({0})V\in C(\mathbb{R}\setminus\{0\}) satisfies

  1. (i)

    (Evenness). V:V:\mathbb{R}\to\mathbb{R} is even;

  2. (ii)

    (Singularity). V(x)V(x)\to\infty as x0x\to 0, and there exist C>0C>0 and a[0,1)a\in[0,1) such that for all x(0,12]x\in(0,\frac{1}{2}]

    V(x)C{|x|aif a>0log|x|if a=0;V(x)\leq C\left\{\begin{array}[]{ll}|x|^{-a}&\text{if }a>0\\ -\log|x|&\text{if }a=0;\end{array}\right.
  3. (iii)

    (Convexity). VV is convex on (0,)(0,\infty) and λ\lambda-convex near x=0x=0, i.e.

    λ,δ>0:xV(x)λ2x2 is convex on (0,δ);\exists\ \lambda,\delta>0:x\mapsto V(x)-\frac{\lambda}{2}x^{2}\text{ is convex on }(0,\delta)\text{;}
  4. (iv)

    (Integrability). VV is normalized to V1=1\|V\|_{1}=1 and has bounded first moment, i.e.

    |x|V(x)𝑑x<;\int_{\mathbb{R}}|x|V(x)\,dx<\infty;
  5. (v)

    (Regularity). VWloc2,1(0,)V\in W_{\operatorname{loc}}^{2,1}(0,\infty) and V^W2,()\sqrt{\widehat{V}}\in W^{2,\infty}(\mathbb{R}).

First, we mention several properties of VV which follow from Assumption 3.2. The evenness, convexity and integrability imply that V0V\geq 0 is non-increasing on (0,)(0,\infty) and that V^\widehat{V} is real-valued, nonnegative and even, which is sufficient for Assumption 3.2(v) to be well-defined. A less obvious consequence is Lemma 3.3.

Lemma 3.3.

There exists a constant c>0c>0 such that for all ω\omega\in\mathbb{R}

v(ω):=V^(ω)cmin{1,|ω|2}.v(\omega):=\widehat{V}(\omega)\geq c\min\{1,|\omega|^{-2}\}.
Proof.

Since vv is even and real-valued, it is enough to focus on ω>0\omega>0. [KvM20, (A.3)] provides the characterization

v(ω)=1πω3k=0012(012V′′(k+x+yω)𝑑y)sin(2πx)𝑑x.v(\omega)=\frac{1}{\pi\omega^{3}}\sum_{k=0}^{\infty}\int_{0}^{\frac{1}{2}}\bigg{(}\int_{0}^{\frac{1}{2}}V^{\prime\prime}\Big{(}\frac{k+x+y}{\omega}\Big{)}dy\bigg{)}\sin(2\pi x)\,dx.

Since the integrand is nonnegative, we may bound it from below by shrinking the integration domain. Then, on 0x140\leq x\leq\frac{1}{4}, we bound sin(2πx)4x\sin(2\pi x)\geq 4x. For V′′V^{\prime\prime} we note from Assumption 3.2(iii) that

V′′(z)λ1(z<δ)for all z>0,V^{\prime\prime}(z)\geq\lambda 1(z<\delta)\qquad\text{for all }z>0,

where 1(P)1(P) equals 11 if the statement PP is true and 0 otherwise. Then,

V^(ω)cω3k=0014(014x1((k+x+y)<δω)𝑑y)x𝑑x.\widehat{V}(\omega)\geq\frac{c}{\omega^{3}}\sum_{k=0}^{\infty}\int_{0}^{\frac{1}{4}}\bigg{(}\int_{0}^{\frac{1}{4}-x}1\big{(}(k+x+y)<\delta\omega\big{)}dy\bigg{)}x\,dx.

We split two cases. If ω14δ\omega\leq\frac{1}{4\delta}, then only the term corresponding to k=0k=0 is nonzero, and the right-hand side equals

cω30δω0δωxx𝑑y𝑑xδ3c.\displaystyle\frac{c}{\omega^{3}}\int_{0}^{\delta\omega}\int_{0}^{\delta\omega-x}xdydx\geq\delta^{3}c^{\prime}.

If ω>14δ\omega>\frac{1}{4\delta}, then we estimate

V^(ω)\displaystyle\widehat{V}(\omega) cω3k=0014(014x1((k+14)<δω)𝑑y)x𝑑x\displaystyle\geq\frac{c}{\omega^{3}}\sum_{k=0}^{\infty}\int_{0}^{\frac{1}{4}}\bigg{(}\int_{0}^{\frac{1}{4}-x}1\big{(}(k+\tfrac{1}{4})<\delta\omega\big{)}dy\bigg{)}x\,dx
=cω3k=01((k+14)<δω)=cω3δω1445δcω2.\displaystyle=\frac{c^{\prime}}{\omega^{3}}\sum_{k=0}^{\infty}1\big{(}(k+\tfrac{1}{4})<\delta\omega\big{)}=\frac{c^{\prime}}{\omega^{3}}\lceil\delta\omega-\tfrac{1}{4}\rceil\geq\frac{4}{5}\delta\frac{c^{\prime}}{\omega^{2}}.

Next we compare Assumption 3.2 to the assumptions on VV made in [GvMPS16], which are weaker. Indeed, in [GvMPS16] Assumption 3.2(ii) and the regularity on VV are not required, and Assumption 3.2(iii) is relaxed to the requirement that V|(0,)V|_{(0,\infty)} is non-increasing. While [GvMPS16] has a further assumption that VV can be approximated from below by a certain class of functions, we show that this holds under Assumption 3.2 by constructing such an approximation explicitly; see (36).

Next we motivate the assumptions which are new with respect to [GvMPS16]. We believe that these additional assumptions are minor, and still allow for most of the potentials in practice which satisfy the assumptions in [GvMPS16]. Regarding Assumption 3.2(ii), it is obvious from the bound on γn\gamma_{n} in Theorem 1.1 that a bound on the singularity of VV is needed. The requirement V(x)V(x)\to\infty as x0x\to 0 might not be necessary. However, including this case in the proof would require a further case splitting. Since we are not aware of any application for this case, we omit it. While the convexity in Assumption 3.2(iii) is new, it captures the following three assumptions in [GvMPS16]: V^>0\widehat{V}>0 (see the proof of Lemma 3.3), V|(0,)V|_{(0,\infty)} is non-increasing, and VV can be approximated from below by a special class of functions. The local λ\lambda-convexity is a technical addition which simplifies several steps in the proof of Theorem 1.1. Finally, in Assumption 3.2(v), the regularity on VV is only a small upgrade of VWloc1,(0,)V\in W_{\operatorname{loc}}^{1,\infty}(0,\infty), which follows from convexity. The regularity of v\sqrt{v} is required in [GvMPS16] to extend FF from L2(Ω)L^{2}(\Omega) to 𝒜\mathcal{A}. We further exploit this assumption when proving properties of the regularization VβV^{\beta}. This is the single assumption which can be hard to check in practice.

Finally, we show that VwallV_{\operatorname{wall}} defined in (3) satisfies Assumption 3.2. We recall from [GPPS13] that VwallV_{\operatorname{wall}} is strictly convex, has a logarithmic singularity (in particular, a=0a=0) and exponential tails. Then, the only non-trivial property left to check is the regularity of v\sqrt{v} with v:=Vwallv:=\mathcal{F}V_{\operatorname{wall}}. By the strict convexity and the exponential tails of VwallV_{\operatorname{wall}}, it follows that vC()v\in C^{\infty}(\mathbb{R}) is positive, and thus vWloc2,()\sqrt{v}\in W_{\operatorname{loc}}^{2,\infty}(\mathbb{R}). To extend this to large ω\omega, we recall from [GPPS13, App. A.1] that

v(ω)=12ω(cothidsinh2)(π2ω)=:φ(ω)ω.v(\omega)=\frac{1}{2\omega}\Big{(}\coth-\frac{\operatorname{id}}{\sinh^{2}}\Big{)}(\pi^{2}\omega)=:\frac{\varphi(\omega)}{\omega}.

Note that φ(ω)1\varphi(\omega)\to 1 and φ(ω),φ′′(ω)0\varphi^{\prime}(\omega),\varphi^{\prime\prime}(\omega)\to 0 as ω\omega\to\infty. Then, we obtain vW2,()\sqrt{v}\in W^{2,\infty}(\mathbb{R}) from

(v)(ω)\displaystyle\big{(}\sqrt{v}\big{)}^{\prime}(\omega) =(φ)(ω)ω1/2φ(ω)2ω3/2ω0\displaystyle=\frac{\big{(}\sqrt{\varphi}\big{)}^{\prime}(\omega)}{\omega^{1/2}}-\frac{\sqrt{\varphi}(\omega)}{2\omega^{3/2}}\xrightarrow{\omega\to\infty}0
(v)′′(ω)\displaystyle\big{(}\sqrt{v}\big{)}^{\prime\prime}(\omega) =(φ)′′(ω)ω1/2(φ)(ω)ω3/2+3φ(ω)4ω5/2ω0.\displaystyle=\frac{\big{(}\sqrt{\varphi}\big{)}^{\prime\prime}(\omega)}{\omega^{1/2}}-\frac{\big{(}\sqrt{\varphi}\big{)}^{\prime}(\omega)}{\omega^{3/2}}+\frac{3\sqrt{\varphi}(\omega)}{4\omega^{5/2}}\xrightarrow{\omega\to\infty}0.

4 Explicit expression of Fn(νn)F_{n}(\nu_{n})

Here we derive an explicit expression for Fn(νn)F_{n}(\nu_{n}) in terms of νn\nu_{n} and motivate the prefactor γn\gamma_{n} in (18). By scaling back, note from (16) that νn𝒜n\nu_{n}\in\mathcal{A}_{n} can be written as

σn:=(γn)νn=μnρ\sigma_{n}:=(\gamma_{n})_{\leftarrow}\nu_{n}=\mu_{n}-\rho_{\ast}

for some empirical measure μn\mu_{n} of the form (14). Then, in view of the right-hand side in (19), we set Vγ(x):=γV(γx)V_{\gamma}(x):=\gamma V(\gamma x) and compute

En(μn)Eγn(ρ)\displaystyle E_{n}(\mu_{n})-E^{\gamma_{n}}(\rho_{\ast})
=12ΔcVγn(xy)𝑑μn(y)𝑑μn(x)12ΔcVγn(xy)𝑑ρ(y)𝑑ρ(x)+ΩU𝑑σn\displaystyle=\frac{1}{2}\iint_{\Delta^{c}}V_{\gamma_{n}}(x-y)\,d\mu_{n}(y)d\mu_{n}(x)-\frac{1}{2}\iint_{\Delta^{c}}V_{\gamma_{n}}(x-y)\,d\rho_{\ast}(y)d\rho_{\ast}(x)+\int_{\Omega}U\,d\sigma_{n}
=12ΔcVγn(xy)𝑑σn(y)𝑑σn(x)+Ω[(Vγnρ)+U]𝑑σn,\displaystyle=\frac{1}{2}\iint_{\Delta^{c}}V_{\gamma_{n}}(x-y)\,d\sigma_{n}(y)d\sigma_{n}(x)+\int_{\Omega}\big{[}(V_{\gamma_{n}}*\rho_{\ast})+U\big{]}\,d\sigma_{n},

where we recall that ρ\rho_{*} and μn\mu_{n} are extended from Ω\Omega to \mathbb{R} by 0.

Next we rewrite the second term. With this aim, we set

ρ¯(x):={ρ(0)if x<0ρ(x)if x0\overline{\rho}_{*}(x):=\left\{\begin{aligned} &\rho_{\ast}(0)&&\text{if }x<0\\ &\rho_{\ast}(x)&&\text{if }x\geq 0\end{aligned}\right. (21)

and expand

Ω[(Vγnρ)+U]𝑑σn=Ω(Vγn(ρρ¯))𝑑σn+Ω((Vγnρ¯)ρ¯)𝑑σn+Ω(ρ+U)𝑑σn.\int_{\Omega}\big{[}(V_{\gamma_{n}}*\rho_{\ast})+U\big{]}\,d\sigma_{n}\\ =\int_{\Omega}\big{(}V_{\gamma_{n}}*(\rho_{\ast}-\overline{\rho}_{*})\big{)}\,d\sigma_{n}+\int_{\Omega}\big{(}(V_{\gamma_{n}}*\overline{\rho}_{*})-\overline{\rho}_{*}\big{)}\,d\sigma_{n}+\int_{\Omega}\big{(}\rho_{\ast}+U\big{)}\,d\sigma_{n}. (22)

The first term equals

ρ(0)Ω0Vγn(xy)𝑑y𝑑σn(x)=ρ(0)0(Vγnσn)(y)𝑑y.-\rho_{\ast}(0)\int_{\Omega}\int_{-\infty}^{0}V_{\gamma_{n}}(x-y)\,dy\,d\sigma_{n}(x)=-\rho_{\ast}(0)\int_{-\infty}^{0}(V_{\gamma_{n}}*\sigma_{n})(y)\,dy. (23)

For the integrand of the third term in (22), we note from (8) that

Ω(ρ+U)𝑑σn=ΩCU𝑑σn+(suppρ)c(UCU)𝑑σn=CUn+(suppρ)c(UCU)𝑑μn.\int_{\Omega}\big{(}\rho_{\ast}+U\big{)}\,d\sigma_{n}=\int_{\Omega}C_{U}\,d\sigma_{n}+\int_{(\operatorname{supp}\rho_{\ast})^{c}}(U-C_{U})\,d\sigma_{n}=\frac{C_{U}}{n}+\int_{(\operatorname{supp}\rho_{\ast})^{c}}(U-C_{U})\,d\mu_{n}.

Collecting our computations, we obtain

En(μn)Eγn(ρ)=12ΔcVγn(xy)𝑑σn(y)𝑑σn(x)ρ(0)0(Vγnσn)(x)𝑑x+Ω(Vγnδ0)ρ¯𝑑σn+(suppρ)c(UCU)𝑑μn+CUn.E_{n}(\mu_{n})-E^{\gamma_{n}}(\rho_{\ast})=\frac{1}{2}\iint_{\Delta^{c}}V_{\gamma_{n}}(x-y)\,d\sigma_{n}(y)d\sigma_{n}(x)-\rho_{\ast}(0)\int_{-\infty}^{0}(V_{\gamma_{n}}*\sigma_{n})(x)\,dx\\ +\int_{\Omega}(V_{\gamma_{n}}-\delta_{0})*\overline{\rho}_{*}\,d\sigma_{n}+\int_{(\operatorname{supp}\rho_{\ast})^{c}}(U-C_{U})\,d\mu_{n}+\frac{C_{U}}{n}.

Multiplying by γn\gamma_{n} and changing variables (recall νn=(γn)σn\nu_{n}=(\gamma_{n})_{\to}\sigma_{n}), we get

Fn(νn)=12ΔcV(xy)𝑑νn(y)𝑑νn(x)ρ(0)0(Vνn)(x)𝑑x+Ωγn(Vγnδ0)ρ¯𝑑σn+(suppρ~)c(U(x/γn)CU)𝑑νn+(x)+CUγnn.F_{n}(\nu_{n})=\frac{1}{2}\iint_{\Delta^{c}}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)-\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu_{n})(x)\,dx\\ +\int_{\Omega}\gamma_{n}(V_{\gamma_{n}}-\delta_{0})*\overline{\rho}_{*}\,d\sigma_{n}+\int_{(\operatorname{supp}\tilde{\rho}_{*})^{c}}\big{(}U(x/\gamma_{n})-C_{U}\big{)}\,d\nu_{n}^{+}(x)+C_{U}\frac{\gamma_{n}}{n}. (24)

For later use we note that the integral in the second term can be rewritten as (recall (23))

0(Vνn)(y)𝑑y=ΩxV(y)𝑑y𝑑νn(x).\int_{-\infty}^{0}(V*\nu_{n})(y)\,dy=\int_{\Omega}\int_{x}^{\infty}V(y)\,dy\,d\nu_{n}(x). (25)

Note that the first two terms in (24) resemble the expression of FF in (30). This motivates the scaling of the energy difference in (18) by γn\gamma_{n}. We treat the latter three terms in (24) as error terms when proving Theorem 1.1. While the third term obviously vanishes as nn\to\infty, the other two terms may not for certain sequences (νn)n(\nu_{n})_{n}. We rely on the fact that the second term is nonnegative, and that the integrand in the first term is expected to be small because Vγvδ0V_{\gamma}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\delta_{0} as γ\gamma\to\infty. We give a precise bound later in Lemma 5.6.

5 Functional setting and preliminaries

In Section 5.1 we recall from [GvMPS16] the necessary functional framework to extend the definition of FF in (12) to 𝒜\mathcal{A}. Since this functional framework also facilitates the statements and proofs of several preliminary estimates, we treat them in the subsequent Section 5.2. In this functional framework we identify measures on Ω\Omega as tempered distributions on \mathbb{R} supported in Ω\Omega.

5.1 Proper definition of FF

Since FF is the same as in [GvMPS16], we briefly recall the extension of the definition in (12) on L2()L^{2}(\mathbb{R}) to 𝒜\mathcal{A}. Ideally, if there exists a function uu such that V=uuV=u*u, then for fL2()f\in L^{2}(\mathbb{R}) we have for the interaction term that

(Vf)f=(uf)2=uf22.\int_{\mathbb{R}}(V*f)f=\int_{\mathbb{R}}(u*f)^{2}=\|u*f\|_{2}^{2}. (26)

This expression can be extended to distributions by noting that

ξ22:=supφCc()(2ξ,φφ22)[0,].\|\xi\|_{2}^{2}:=\sup_{\varphi\in C_{c}^{\infty}(\mathbb{R})}\big{(}2\langle\xi,\varphi\rangle-\|\varphi\|_{2}^{2}\big{)}\in[0,\infty]. (27)

However, from Assumption 3.2 it is not clear whether such a function uu exists.

One way to avoid characterizing uu is to work in Fourier space. Since convolution transforms into multiplication by the Fourier transform, the linear operation of convolving by uu turns into multiplication by V\sqrt{\mathcal{F}V}, which is a function due to Assumption 3.2(v). Precisely, we set

T:L2()L2(),Tf:=1(vf^),T:L^{2}(\mathbb{R})\to L^{2}(\mathbb{R}),\qquad Tf:=\mathcal{F}^{-1}(\sqrt{v}\widehat{f}), (28)

where v=Vv=\mathcal{F}V. Then, by translation to Fourier space, we observe that (26) turns into

(Vf)f=Tf22.\displaystyle\int_{\mathbb{R}}(V*f)f=\|Tf\|_{2}^{2}. (29)

Together with the observation in (25) this yields

F(ν)=12Tν22ρ(0)Ωg𝑑ν\displaystyle F(\nu)=\frac{1}{2}\|T\nu\|_{2}^{2}-\rho_{\ast}(0)\int_{\Omega}g\,d\nu (30)

for all ν𝒜L2()\nu\in\mathcal{A}\cap L^{2}(\mathbb{R}), where

g(x):=xV(y)𝑑yfor all x0.\displaystyle g(x):=\int_{x}^{\infty}V(y)\,dy\qquad\text{for all }x\geq 0. (31)

To extend FF to 𝒜\mathcal{A}, we show that the linear term in (30) is bounded and that the operator TT can be extended to 𝒜\mathcal{A}. We do this in Lemmas 5.1 and 5.2. For later use, we state these lemmas in a general form. With this aim, we introduce the Hilbert spaces

Xk,j()\displaystyle X_{k,j}(\mathbb{C}) :={fHk(;):xjf(x)L2(;)}\displaystyle:=\big{\{}f\in H^{k}(\mathbb{R};\mathbb{C}):x^{j}f(x)\in L^{2}(\mathbb{R};\mathbb{C})\big{\}}
(f,ϕ)Xk,j\displaystyle(f,\phi)_{X_{k,j}} :==0kf()ϕ()¯+x2jf(x)ϕ(x)¯𝑑x\displaystyle:=\sum_{\ell=0}^{k}\int_{\mathbb{R}}f^{(\ell)}\overline{\phi^{(\ell)}}+\int_{\mathbb{R}}x^{2j}f(x)\overline{\phi(x)}\,dx

for all k,jk,j\in\mathbb{N}. In case the functions are real-valued, we set

Xk,j:={fHk():xjf(x)L2()}.X_{k,j}:=\Big{\{}f\in H^{k}(\mathbb{R}):x^{j}f(x)\in L^{2}(\mathbb{R})\Big{\}}. (32)

Note that L2()=X0,0Xk,j𝒮()L^{2}(\mathbb{R})=X_{0,0}\supset X_{k,j}\supset\mathcal{S}(\mathbb{R}) for all k,jk,j\in\mathbb{N}, where 𝒮()\mathcal{S}(\mathbb{R}) is the space of Schwarz functions. Note from the Fourier transform property

(xjf(k))=ik+j(2π)kj(ωkf^)(j)\mathcal{F}(x^{j}f^{(k)})=i^{k+j}(2\pi)^{k-j}\big{(}\omega^{k}\widehat{f}\big{)}^{(j)}

that \mathcal{F} is an invertible bounded linear operator from Xk,j()X_{k,j}(\mathbb{C}) to Xj,k()X_{j,k}(\mathbb{C}). We further set Xk,jX_{k,j}^{\prime} as the dual of Xk,jX_{k,j} with respect to the L2L^{2}-topology, and (Xk,j)\mathcal{L}(X_{k,j}) as the space of all bounded linear operators from Xk,jX_{k,j} to Xk,jX_{k,j}.

Lemma 5.1 ([GvMPS16, Lem. A.2]).

The operator TT and the set 𝒜\mathcal{A} satisfy

  1. (i)

    𝒜X1,2\mathcal{A}\subset X_{1,2}^{\prime};

  2. (ii)

    For each kk\in\mathbb{N} and each {0,1,2}\ell\in\{0,1,2\}, T(Xk,j)T\in\mathcal{L}(X_{k,j}) is symmetric, and can be extended to T(Xk,j)T\in\mathcal{L}(X_{k,j}^{\prime}) by

    Tξ,f:=ξ,Tffor all ξXk,j,fXk,j.\langle T\xi,f\rangle:=\langle\xi,Tf\rangle\qquad\text{for all }\xi\in X_{k,j}^{\prime},\ f\in X_{k,j}.
Lemma 5.2 ([GvMPS16, Lem. 3.4]).

There exist constants C,εM>0C,\varepsilon_{M}>0 with

supM0εM<andεMM0\sup_{M\geq 0}\varepsilon_{M}<\infty\quad\text{and}\quad\varepsilon_{M}\xrightarrow{M\to\infty}0

such that for all ν𝒜\nu\in\mathcal{A} and all M0M\geq 0 it holds that

Mgd|ν|\displaystyle\int_{M}^{\infty}g\,d|\nu| εMsupx0|ν|([x,x+1]),where g is as in (31), and\displaystyle\leq\varepsilon_{M}\sup_{x\geq 0}|\nu|([x,x+1]),\quad\text{where $g$ is as in \eqref{g}, and} (33)
|Ωfν|\displaystyle\bigg{|}\int_{\Omega}f\nu\,\bigg{|} CfX1,2supx0|ν|([x,x+1])for all fX1,2.\displaystyle\leq C\|f\|_{X_{1,2}}\sup_{x\geq 0}|\nu|([x,x+1])\quad\text{for all }f\in X_{1,2}. (34)

We remark that while Lemma 5.2 is a stronger statement than [GvMPS16, Lem. 3.4], the proof in [GvMPS16] instantly implies Lemma 5.2.

5.2 A priori bounds on νn\nu_{n}

First we state the counterpart of Lemma 5.1(i) in the discrete setting. Since Dirac-delta measures are included in H1()H^{-1}(\mathbb{R}), we obtain

𝒜nH1()=X1,0.\mathcal{A}_{n}\subset H^{-1}(\mathbb{R})=X_{1,0}^{\prime}. (35)

While 𝒜n𝒜\mathcal{A}_{n}\subset\mathcal{A}, it is pointless to consider FF on 𝒜n\mathcal{A}_{n}, because F(νn)=F(\nu_{n})=\infty for each νn𝒜n\nu_{n}\in\mathcal{A}_{n} due to the discreteness and the singularity of VV at 0. This is the crucial difference with the continuous setting in [GvMPS16] in which FnF_{n} is a small perturbation of F|𝒜nF|_{\mathcal{A}_{n}}.

To deal with the discreteness, we construct a careful regularization of VV. With this aim, let λ,δ>0\lambda,\delta>0 be as in Assumption 3.2(iii). For β(0,δ)\beta\in(0,\delta) we define the regularization

Vβ(x)={V(β)+(xβ)V(β)+λ2(xβ)2if 0x<βV(x)if xβ,V^{\beta}(x)=\left\{\begin{aligned} &V(\beta)+(x-\beta)V^{\prime}(\beta)+\frac{\lambda}{2}(x-\beta)^{2}&&\text{if }0\leq x<\beta\\ &V(x)&&\text{if }x\geq\beta,\end{aligned}\right. (36)

and consider the even extension of VβV^{\beta} to \mathbb{R}. Figure 4 illustrates VβV^{\beta}. Note that on (0,β)(0,\beta), VβV^{\beta} is a parabola which is tangent to VV at x=βx=\beta. We also set

Wβ:=VVβ.W^{\beta}:=V-V^{\beta}.

Lemma 5.3 lists several properties of Vβ,WβV^{\beta},W^{\beta}.

xx0β\betaVβ(x)V^{\beta}(x)Wβ(x)W^{\beta}(x)V(x)V(x)
Figure 4: Sketch of VβV^{\beta}. The first two terms in (36) for x<βx<\beta describe the tangent line (red) of VV at x=βx=\beta.
Lemma 5.3 (Properties of Vβ,WβV^{\beta},W^{\beta}).

There exist constants C,c>0C,c>0 such that for all β\beta small enough

  1. (i)

    (Pointwise bounds). 0VβV0\leq V^{\beta}\leq V and 0WβV0\leq W^{\beta}\leq V;

  2. (ii)

    (Convexity). VβV^{\beta} and WβW^{\beta} are convex and non-increasing on (0,)(0,\infty). Moreover, VβV^{\beta} satisfies Assumption 3.2(iii) with the same constants λ,δ\lambda,\delta;

  3. (iii)

    (Narrow support). suppWβ[β,β]\operatorname{supp}W^{\beta}\subset[-\beta,\beta];

  4. (iv)

    (Fourier transform). vβ:=Vβv_{\beta}:=\mathcal{F}V^{\beta} is real-valued and even, and satisfies

    vβvandvβ(ω)cmin{1,|ω|2}>0for all ω.\|v_{\beta}\|_{\infty}\leq\|v\|_{\infty}\quad\text{and}\quad v_{\beta}(\omega)\geq c\min\{1,|\omega|^{-2}\}>0\quad\text{for all }\omega\in\mathbb{R}.
  5. (v)

    (L1L^{1}, LL^{\infty} bounds). Vβ=Vβ(0)Cβa\|V^{\beta}\|_{\infty}=V^{\beta}(0)\leq C\beta^{-a} and Wβ1=WβCβ1a\|W^{\beta}\|_{1}=\int_{\mathbb{R}}W^{\beta}\leq C\beta^{1-a}.

Proof.

Properties (i), (ii), (iii) and the fact that vβv_{\beta} is real-valued and even hold by construction. From these properties, we observe that the proof of Lemma 3.3 also applies to vβv_{\beta}; this proves the lower bound in (iv). The upper bound follows simply from vβ=Vβ1\|v_{\beta}\|_{\infty}=\|V^{\beta}\|_{1} and (i). The bound on WβW^{\beta} in (v) follows from (iii) through

Wβ=ββWβββVCββ|x|a𝑑x=Cβ1a.\int_{\mathbb{R}}W^{\beta}=\int_{-\beta}^{\beta}W^{\beta}\leq\int_{-\beta}^{\beta}V\leq C\int_{-\beta}^{\beta}|x|^{-a}\,dx=C^{\prime}\beta^{1-a}.

To prove the bound on Vβ(0)V^{\beta}(0) in (v), we observe from (36) that

Vβ(0)=V(β)βV(β)+λ2β2Cβa+β|V(β)|.V^{\beta}(0)=V(\beta)-\beta V^{\prime}(\beta)+\frac{\lambda}{2}\beta^{2}\leq C\beta^{-a}+\beta|V^{\prime}(\beta)|.

By the Mean Value Theorem and the fact that |V||V^{\prime}| is non-increasing, we get

|V(β)|V(β/2)V(β)β/22Cβa0β=2Cβ1a.|V^{\prime}(\beta)|\leq\frac{V(\beta/2)-V(\beta)}{\beta/2}\leq 2\frac{C\beta^{-a}-0}{\beta}=2C\beta^{-1-a}.

This completes the proof of Lemma 5.3. ∎

Lemma 5.3(iv) allows us to define, similarly to VV, the convolutional square root operator

Tβ:L2()L2(),Tβf:=1(vβf^),T^{\beta}:L^{2}(\mathbb{R})\to L^{2}(\mathbb{R}),\qquad T^{\beta}f:=\mathcal{F}^{-1}(\sqrt{v_{\beta}}\widehat{f}),

where vβ=Vβv_{\beta}=\mathcal{F}V^{\beta}. As in (29), it is easy to see (in Fourier space) that Vβf=TβTβfV^{\beta}*f=T^{\beta}T^{\beta}f and

(Vβf)f=Tβf22.\int_{\mathbb{R}}(V^{\beta}*f)f=\|T^{\beta}f\|_{2}^{2}. (37)

Lemma 5.4 states further properties of TβT^{\beta}.

Lemma 5.4.

The operator TβT^{\beta} defined above satisfies

  1. (i)

    For each kk\in\mathbb{N} and all β\beta small enough, Tβ(Xk,0)T^{\beta}\in\mathcal{L}(X_{k,0}) is symmetric, and can be extended to Tβ(Xk,0)T^{\beta}\in\mathcal{L}(X_{k,0}^{\prime}) as in Lemma 5.1;

  2. (ii)

    lim supβ0Tβ(Xk,0)<\displaystyle\limsup_{\beta\to 0}\|T^{\beta}\|_{\mathcal{L}(X_{k,0})}<\infty for each kk\in\mathbb{N};

  3. (iii)

    For all β\beta small enough, each n1n\geq 1 and all νn𝒜n\nu_{n}\in\mathcal{A}_{n}, there holds TβνnL2()T^{\beta}\nu_{n}\in L^{2}(\mathbb{R}),

    Vβνn=TβTβνnandΩ(Vβνn)𝑑νn=Tβνn22;V^{\beta}\ast\nu_{n}=T^{\beta}T^{\beta}\nu_{n}\quad\text{and}\quad\int_{\Omega}(V^{\beta}\ast\nu_{n})\,d\nu_{n}=\|T^{\beta}\nu_{n}\|_{2}^{2}\text{;}
  4. (iv)

    There exists β0>0\beta_{0}>0 such that for all Schwarz functions f𝒮()f\in\mathcal{S}(\mathbb{R}) there exists Cf>0C_{f}>0 such that for all β(0,β0)\beta\in(0,\beta_{0})

    (TTβ)fX1,2Cfβ1a.\|(T-T^{\beta})f\|_{X_{1,2}}\leq C_{f}\beta^{1-a}.
Proof.

Since vβ=(Vβ)1/2\sqrt{v_{\beta}}=(\mathcal{F}V^{\beta})^{1/2} is even and bounded, (i) and (ii) follow from the same argument used in the proof of [GvMPS16, Lem. A.2(ii)]. In particular,

Tβ(Xk,0)Cvβ,\|T^{\beta}\|_{\mathcal{L}(X_{k,0})}\leq C\|\sqrt{v_{\beta}}\|_{\infty},

for which Lemma 5.3(iv) provides a sufficient bound.

Next we proof (iii). It is clear that Vβνn=TβTβνnV^{\beta}\ast\nu_{n}=T^{\beta}T^{\beta}\nu_{n} holds in distributional sense. Note that (37) defines a seminorm, which can be extended to distributions similarly to (27). Using this and applying (i), we write

>Ω(Vβνn)𝑑νn\displaystyle\infty>\int_{\Omega}(V^{\beta}\ast\nu_{n})\,d\nu_{n} =supφX1,0(2Vβφ,νnΩ(Vβφ)𝑑φ)\displaystyle=\sup_{\varphi\in X_{1,0}}\bigg{(}2\langle V^{\beta}*\varphi,\nu_{n}\rangle-\int_{\Omega}(V^{\beta}\ast\varphi)\,d\varphi\bigg{)}
=supφX1,0(2TβTβφ,νnTβφ22)\displaystyle=\sup_{\varphi\in X_{1,0}}\big{(}2\langle T^{\beta}T^{\beta}\varphi,\nu_{n}\rangle-\|T^{\beta}\varphi\|_{2}^{2}\big{)}
=supψTβX1,0(2ψ,Tβνnψ22),\displaystyle=\sup_{\psi\in T^{\beta}X_{1,0}}\big{(}2\langle\psi,T^{\beta}\nu_{n}\rangle-\|\psi\|_{2}^{2}\big{)},

where TβX1,0:={TβffX1,0}T^{\beta}X_{1,0}:=\{T^{\beta}f\mid f\in X_{1,0}\}. We claim that TβX1,0T^{\beta}X_{1,0} is dense in X1,0X_{1,0}. From this claim, (iii) follows by applying (27). To prove the claim, we note that, by translating it to Fourier space, it is equivalent to the claim that {vβffX0,1}\{\sqrt{v_{\beta}}f\mid f\in X_{0,1}\} is dense in X0,1X_{0,1}. This is easily seen to be true; for fX0,1f\in X_{0,1}, set fk:=f|(k,k)f_{k}:=f|_{(-k,k)} and note from the lower bound in Lemma 5.3(iv) that fk/vβX0,1f_{k}/\sqrt{v_{\beta}}\in X_{0,1} for all kk\in\mathbb{N}.

Finally we prove (iv). Note that

(TTβ)fX1,2C((TTβ)f)X2,1.\|(T-T^{\beta})f\|_{X_{1,2}}\leq C\big{\|}\mathcal{F}\big{(}(T-T^{\beta})f\big{)}\big{\|}_{X_{2,1}}. (38)

Recalling v=Vv=\mathcal{F}V, we set wβ:=Wβ=vvβw_{\beta}:=\mathcal{F}W^{\beta}=v-v_{\beta} and compute

((TTβ)f)=(vvβ)f^=wβv+vβf^=:wβuβf^,\mathcal{F}\big{(}(T-T^{\beta})f\big{)}=(\sqrt{v}-\sqrt{v_{\beta}})\widehat{f}=\frac{w_{\beta}}{\sqrt{v}+\sqrt{v_{\beta}}}\widehat{f}=:w_{\beta}u_{\beta}\widehat{f},

where uβ=(v+vβ)1/2u_{\beta}=(\sqrt{v}+\sqrt{v_{\beta}})^{-1/2}. We observe from Lemma 5.3(i),(iii) that for any jj\in\mathbb{N}

wβ(j)=(2π)j(xjWβ)=(2π)j|x|jWβ(x)𝑑xCj0βxja𝑑x=Cjβj+1a.\|w_{\beta}^{(j)}\|_{\infty}=(2\pi)^{j}\|\mathcal{F}(x^{j}W^{\beta})\|_{\infty}=(2\pi)^{j}\int_{\mathbb{R}}|x|^{j}W^{\beta}(x)\,dx\leq C_{j}\int_{0}^{\beta}x^{j-a}\,dx=C_{j}^{\prime}\beta^{j+1-a}.

This will eventually result in the prefactor in (iv).

Next we bound uβu_{\beta}. We recall from Assumption 3.2(v) that

v,(v),(v)′′L()\sqrt{v},(\sqrt{v})^{\prime},(\sqrt{v})^{\prime\prime}\in L^{\infty}(\mathbb{R})

and from Lemmas 3.3 and 5.3(iv) that

min{v,vβ}cmin{1,1|ω|}.\min\big{\{}\sqrt{v},\sqrt{v_{\beta}}\big{\}}\geq c\min\Big{\{}1,\frac{1}{|\omega|}\Big{\}}.

Here and henceforth, we often abuse notation by removing the variable ω\omega from the notation for functions in Fourier space. Then,

uβ=1v+vβC(1+|ω|)=:Cω¯1,u_{\beta}=\frac{1}{\sqrt{v}+\sqrt{v_{\beta}}}\leq C(1+|\omega|)=:C\overline{\omega}^{1},

where CC is independent of β\beta and

ω¯k:=(1+|ω|k)for all k.\overline{\omega}^{k}:=(1+|\omega|^{k})\qquad\text{for all }k\in\mathbb{N}.

Using this, we obtain for the X0,1X_{0,1}-part of the norm in (38) that

((TTβ)f)X0,1=wβuβf^X0,1ω¯1wβuβf^2Cβ1aω¯2f^2=Cfβ1a.\big{\|}\mathcal{F}\big{(}(T-T^{\beta})f\big{)}\big{\|}_{X_{0,1}}=\big{\|}w_{\beta}u_{\beta}\widehat{f}\big{\|}_{X_{0,1}}\leq\big{\|}\overline{\omega}^{1}w_{\beta}u_{\beta}\widehat{f}\big{\|}_{2}\leq C\beta^{1-a}\|\overline{\omega}^{2}\widehat{f}\|_{2}=C_{f}\beta^{1-a}.

For the H1H^{1}-part of the norm, we compute

uβ=1(v+vβ)2((v)+vβ2vβ)=uβ2((v)+vwβ2vβ).u_{\beta}^{\prime}=\frac{-1}{(\sqrt{v}+\sqrt{v_{\beta}})^{2}}\Big{(}(\sqrt{v})^{\prime}+\frac{v_{\beta}^{\prime}}{2\sqrt{v_{\beta}}}\Big{)}=-u_{\beta}^{2}\bigg{(}(\sqrt{v})^{\prime}+\frac{v^{\prime}-w_{\beta}^{\prime}}{2\sqrt{v_{\beta}}}\bigg{)}. (39)

Writing v=2v(v)L()v^{\prime}=2\sqrt{v}(\sqrt{v})^{\prime}\in L^{\infty}(\mathbb{R}), all terms in the expression for uβu_{\beta}^{\prime} can be bounded from the estimates above. This yields

|uβ|Cω¯2(C+(C′′+β2a)ω¯1)C′′′ω¯3.|u_{\beta}^{\prime}|\leq C\overline{\omega}^{2}\Big{(}C^{\prime}+(C^{\prime\prime}+\beta^{2-a})\overline{\omega}^{1}\Big{)}\leq C^{\prime\prime\prime}\overline{\omega}^{3}.

Hence,

(((TTβ)f))2\displaystyle\big{\|}\big{(}\mathcal{F}\big{(}(T-T^{\beta})f\big{)}\big{)}^{\prime}\big{\|}_{2} wβuβf^2+wβuβf^2+wβuβ(f^)2\displaystyle\leq\big{\|}w_{\beta}^{\prime}u_{\beta}\widehat{f}\big{\|}_{2}+\big{\|}w_{\beta}u_{\beta}^{\prime}\widehat{f}\big{\|}_{2}+\big{\|}w_{\beta}u_{\beta}(\widehat{f})^{\prime}\big{\|}_{2}
Cβ1a(βω¯1f^2+ω¯3f^2+ω¯1(f^)2)Cfβ1a.\displaystyle\leq C\beta^{1-a}\Big{(}\beta\big{\|}\overline{\omega}^{1}\widehat{f}\big{\|}_{2}+\big{\|}\overline{\omega}^{3}\widehat{f}\big{\|}_{2}+\big{\|}\overline{\omega}^{1}(\widehat{f})^{\prime}\big{\|}_{2}\Big{)}\leq C_{f}\beta^{1-a}.

Finally we bound the H2H^{2}-part of the norm. By the estimates obtained so far,

(((TTβ)f))′′2\displaystyle\big{\|}\big{(}\mathcal{F}\big{(}(T-T^{\beta})f\big{)}\big{)}^{\prime\prime}\big{\|}_{2} wβ(uβf^)′′2+2wβ(uβf^)2+wβ′′uβf^2\displaystyle\leq\big{\|}w_{\beta}(u_{\beta}\widehat{f})^{\prime\prime}\big{\|}_{2}+2\big{\|}w_{\beta}^{\prime}(u_{\beta}\widehat{f})^{\prime}\big{\|}_{2}+\big{\|}w_{\beta}^{\prime\prime}u_{\beta}\widehat{f}\big{\|}_{2}
Cβ1a(uβf^)′′2+Cfβ2a.\displaystyle\leq C\beta^{1-a}\big{\|}(u_{\beta}\widehat{f})^{\prime\prime}\big{\|}_{2}+C_{f}\beta^{2-a}. (40)

In preparation for estimating the second derivative, we compute (using the expression in (39))

uβ′′=2uβuβ((v)+vwβ2vβ)uβ2((v)′′+v′′wβ′′2vβ(vwβ)24vβ3/2).\displaystyle u_{\beta}^{\prime\prime}=-2u_{\beta}u_{\beta}^{\prime}\bigg{(}(\sqrt{v})^{\prime}+\frac{v^{\prime}-w_{\beta}^{\prime}}{2\sqrt{v_{\beta}}}\bigg{)}-u_{\beta}^{2}\bigg{(}(\sqrt{v})^{\prime\prime}+\frac{v^{\prime\prime}-w_{\beta}^{\prime\prime}}{2\sqrt{v_{\beta}}}-\frac{(v^{\prime}-w_{\beta}^{\prime})^{2}}{4v_{\beta}^{3/2}}\bigg{)}.

Then, rewriting

v′′=((v)2)′′=2v(v)′′+2((v))2L(),v^{\prime\prime}=((\sqrt{v})^{2})^{\prime\prime}=2\sqrt{v}(\sqrt{v})^{\prime\prime}+2((\sqrt{v})^{\prime})^{2}\in L^{\infty}(\mathbb{R}),

we estimate all terms in the expression for uβ′′u_{\beta}^{\prime\prime} by the bounds obtained above. This yields

|uβ′′|Cω¯1ω¯3(1+ω¯1)+Cω¯2(1+ω¯1+ω¯3)C′′ω¯5.\displaystyle|u_{\beta}^{\prime\prime}|\leq C\overline{\omega}^{1}\overline{\omega}^{3}(1+\overline{\omega}^{1})+C^{\prime}\overline{\omega}^{2}(1+\overline{\omega}^{1}+\overline{\omega}^{3})\leq C^{\prime\prime}\overline{\omega}^{5}.

Returning to (40), we obtain

(uβf^)′′2\displaystyle\big{\|}(u_{\beta}\widehat{f})^{\prime\prime}\big{\|}_{2} uβ′′f^2+2uβ(f^)2+uβ(f^)′′2\displaystyle\leq\big{\|}u_{\beta}^{\prime\prime}\widehat{f}\big{\|}_{2}+2\big{\|}u_{\beta}^{\prime}(\widehat{f})^{\prime}\big{\|}_{2}+\big{\|}u_{\beta}(\widehat{f})^{\prime\prime}\big{\|}_{2}
C(ω¯5f^2+ω¯3(f^)2+ω¯1(f^)′′2)Cf.\displaystyle\leq C\big{(}\big{\|}\overline{\omega}^{5}\widehat{f}\big{\|}_{2}+\big{\|}\overline{\omega}^{3}(\widehat{f})^{\prime}\big{\|}_{2}+\big{\|}\overline{\omega}^{1}(\widehat{f})^{\prime\prime}\big{\|}_{2}\big{)}\leq C_{f}.

Plugging this estimate into (40) completes the proof of (iv). ∎

This completes the preliminaries on the regularization VβV^{\beta}. Next we apply them to construct tools for the proof of Theorem 1.1. The first of these tools is a crucial estimate in the proof of compactness. It is the discrete counterpart of [GvMPS16, Lem. 3.3].

Lemma 5.5.

There exists a constant C>0C>0 such that for all β>0\beta>0 small enough and all n1n\geq 1

supx0νn+([x,x+1])+0VνnC(TβνnL2()+1).\displaystyle\sup_{x\geq 0}\nu_{n}^{+}([x,x+1])+\int_{-\infty}^{0}V*\nu_{n}\leq C(\|T^{\beta}\nu_{n}\|_{L^{2}(\mathbb{R})}+1).
Proof.

The proof is a modification of the proof of [GvMPS16, Lem. 3.3]. Let x0x\geq 0 be arbitrary. Using that Vβ0V^{\beta}\geq 0 is non-increasing on [0,)[0,\infty), we obtain

xx+1Vβνn+=Ωxx+1Vβ(yz)𝑑y𝑑νn+(z)[x,x+1]xx+1Vβ(yz)𝑑y𝑑νn+(z)[x,x+1](01Vβ(y)𝑑y)𝑑νn+(z)=(01Vβ)νn+([x,x+1]).\int_{x}^{x+1}V^{\beta}*\nu_{n}^{+}=\int_{\Omega}\int_{x}^{x+1}V^{\beta}(y-z)\,dyd\nu_{n}^{+}(z)\geq\int_{[x,x+1]}\int_{x}^{x+1}V^{\beta}(y-z)\,dyd\nu_{n}^{+}(z)\\ \geq\int_{[x,x+1]}\bigg{(}\int_{0}^{1}V^{\beta}(y)\,dy\bigg{)}d\nu_{n}^{+}(z)=\bigg{(}\int_{0}^{1}V^{\beta}\bigg{)}\nu_{n}^{+}([x,x+1]).

Taking β\beta small enough such that 01Vβ1201V>0\int_{0}^{1}V^{\beta}\geq\frac{1}{2}\int_{0}^{1}V>0, Lemma 5.4(iii) implies

νn+([x,x+1])201Vxx+1Vβνn+=201Vxx+1(TβTβνn)+201Vxx+1Vβνn.\displaystyle\nu_{n}^{+}([x,x+1])\leq\frac{2}{\int_{0}^{1}V}\int_{x}^{x+1}V^{\beta}*\nu_{n}^{+}=\frac{2}{\int_{0}^{1}V}\int_{x}^{x+1}(T^{\beta}T^{\beta}\nu_{n})+\frac{2}{\int_{0}^{1}V}\int_{x}^{x+1}V^{\beta}*\nu_{n}^{-}.

Then, applying the Cauchy-Schwarz Inequality and Lemma 5.4(ii)

νn+([x,x+1])\displaystyle\nu_{n}^{+}([x,x+1]) CTβTβνnL2()+Cxx+10γnVβ(yz)ρ~(z)𝑑z𝑑y\displaystyle\leq C\|T^{\beta}T^{\beta}\nu_{n}\|_{L^{2}(\mathbb{R})}+C\int_{x}^{x+1}\int_{0}^{\gamma_{n}}V^{\beta}(y-z)\tilde{\rho}_{*}(z)\,dzdy
CTβνnL2()+C(V)ρ,\displaystyle\leq C^{\prime}\|T^{\beta}\nu_{n}\|_{L^{2}(\mathbb{R})}+C\bigg{(}\int_{\mathbb{R}}V\bigg{)}\|\rho_{*}\|_{\infty}, (41)

which shows the desired estimate for the first term in the display in Lemma 5.5.

For the second term, we use that V0V\geq 0 is non-increasing on (0,)(0,\infty) together with (41) to estimate

0Vνn\displaystyle\int_{-\infty}^{0}V*\nu_{n} 0ΩV(yz)𝑑νn+(z)𝑑y\displaystyle\leq\int_{-\infty}^{0}\int_{\Omega}V(y-z)d\nu_{n}^{+}(z)dy
0k=0V(yk)νn+([k,k+1])dy\displaystyle\leq\int_{-\infty}^{0}\sum_{k=0}^{\infty}V(y-k)\,\nu_{n}^{+}([k,k+1])dy
C(TβνnL2()+1)k=0kV,\displaystyle\leq C(\|T^{\beta}\nu_{n}\|_{L^{2}(\mathbb{R})}+1)\sum_{k=0}^{\infty}\int_{k}^{\infty}V,

where the sum is finite due to Assumption 3.2(iv). ∎

The local bound on νn+\nu_{n}^{+} in Lemma 5.5 turns out useful when passing to the limit nn\to\infty in the second and third term in the expression for FnF_{n} in (24). This is made precise in the following two lemmas.

Lemma 5.6.

There exists C>0C>0 such that for all n1n\geq 1 and all νn𝒜n\nu_{n}\in\mathcal{A}_{n}

|Ωγn(Vγnδ0)ρ¯𝑑σn|C,\bigg{|}\int_{\Omega}\gamma_{n}(V_{\gamma_{n}}-\delta_{0})*\overline{\rho}_{*}\,d\sigma_{n}\bigg{|}\leq C,

where σn=(γn)νn\sigma_{n}=(\gamma_{n})_{\leftarrow}\nu_{n} and ρ¯\overline{\rho}_{*} is defined in (21). If

supn1supx0νn+([x,x+1])<,\sup_{n\geq 1}\sup_{x\geq 0}\nu_{n}^{+}([x,x+1])<\infty,

then

Ωγn(Vγnδ0)ρ¯𝑑σn\displaystyle\int_{\Omega}\gamma_{n}(V_{\gamma_{n}}-\delta_{0})*\overline{\rho}_{*}\,d\sigma_{n} n0.\displaystyle\xrightarrow{n\to\infty}0.
Proof.

Let un:=γn(Vγnδ0)ρ¯u_{n}:=\gamma_{n}(V_{\gamma_{n}}-\delta_{0})*\overline{\rho}_{*}. Since ρ¯\overline{\rho}_{*} is Lipschitz continuous, we obtain

unL()=supx|γn(ρ¯(xy)ρ¯(x))V(γny)γn𝑑y|L|γny|V(γny)γn𝑑y=L|z|V(z)𝑑z,\|u_{n}\|_{L^{\infty}(\mathbb{R})}=\sup_{x\in\mathbb{R}}\bigg{|}\int_{\mathbb{R}}\gamma_{n}\big{(}\overline{\rho}_{*}(x-y)-\overline{\rho}_{*}(x)\big{)}V(\gamma_{n}y)\gamma_{n}\,dy\bigg{|}\\ \leq L\int_{\mathbb{R}}|\gamma_{n}y|V(\gamma_{n}y)\gamma_{n}\,dy=L\int_{\mathbb{R}}|z|V(z)\,dz, (42)

which is bounded due to Assumption 3.2(iv). Hence,

|Ωun𝑑σn|un|σn|(Ω)C(2+n1).\bigg{|}\int_{\Omega}u_{n}\,d\sigma_{n}\bigg{|}\leq\|u_{n}\|_{\infty}|\sigma_{n}|(\Omega)\leq C(2+n^{-1}). (43)

To prove the convergence statement in Lemma 5.6, we improve the bound on unu_{n} in (42). Let ε>0\varepsilon>0 be arbitrary. By (8) and (20) there exist finitely many points y1,,yy_{1},\ldots,y_{\ell} such that ρ¯\overline{\rho}_{*}^{\prime} is uniformly continuous on {y1,,y}\mathbb{R}\setminus\{y_{1},\ldots,y_{\ell}\}. Let Ii:=(yiε0,yi+ε0)I_{i}:=(y_{i}-\varepsilon_{0},y_{i}+\varepsilon_{0}) where ε0=ε2\varepsilon_{0}=\frac{\varepsilon}{2\ell}. Then, for all x𝒪c:=𝒪x\in\mathcal{O}^{c}:=\mathbb{R}\setminus\mathcal{O},

ρ¯(xy)ρ¯(x)=ρ(x)y+rx(y)for all y,\overline{\rho}_{*}(x-y)-\overline{\rho}_{*}(x)=-\rho_{*}^{\prime}(x)y+r_{x}(y)\qquad\text{for all }y\in\mathbb{R},

where the bounded function yrx(y)/yy\mapsto r_{x}(y)/y vanishes as y0y\to 0 uniformly in x𝒪cx\in\mathcal{O}^{c}. Using this, we estimate, similarly to (42),

unL(𝒪c)\displaystyle\|u_{n}\|_{L^{\infty}(\mathcal{O}^{c})} =supx𝒪c|γn(ρ(x)y+rx(y))V(γny)γn𝑑y|\displaystyle=\sup_{x\in\mathcal{O}^{c}}\bigg{|}\int_{\mathbb{R}}\gamma_{n}\big{(}\rho_{*}^{\prime}(x)y+r_{x}(y)\big{)}V(\gamma_{n}y)\gamma_{n}\,dy\bigg{|}
γnρ|yVγn(y)𝑑y|+supx𝒪c{|z|<γn}|rx(z/γn)z/γn||z|V(z)𝑑z\displaystyle\leq\gamma_{n}\|\rho_{*}^{\prime}\|_{\infty}\bigg{|}\int_{\mathbb{R}}yV_{\gamma_{n}}(y)\,dy\bigg{|}+\sup_{x\in\mathcal{O}^{c}}\int_{\{|z|<\sqrt{\gamma_{n}}\}}\Big{|}\frac{r_{x}(z/\gamma_{n})}{z/\gamma_{n}}\Big{|}|z|V(z)\,dz
+supx𝒪c{|z|>γn}|rx(z/γn)z/γn||z|V(z)𝑑z\displaystyle\qquad+\sup_{x\in\mathcal{O}^{c}}\int_{\{|z|>\sqrt{\gamma_{n}}\}}\Big{|}\frac{r_{x}(z/\gamma_{n})}{z/\gamma_{n}}\Big{|}|z|V(z)\,dz
0+(supx𝒪csup|z|<γn|rx(z/γn)z/γn|)|z|V(z)𝑑z+C{|z|>γn}|z|V(z)𝑑z\displaystyle\leq 0+\bigg{(}\sup_{x\in\mathcal{O}^{c}}\sup_{|z|<\sqrt{\gamma_{n}}}\Big{|}\frac{r_{x}(z/\gamma_{n})}{z/\gamma_{n}}\Big{|}\bigg{)}\int_{\mathbb{R}}|z|V(z)\,dz+C\int_{\{|z|>\sqrt{\gamma_{n}}\}}|z|V(z)\,dz
n0.\displaystyle\xrightarrow{n\to\infty}0.

Using this, we sharpen the bound in (43) by

|Ωun𝑑σn|unL()|σn|(𝒪)+unL(𝒪c)|σn|(Ω).\bigg{|}\int_{\Omega}u_{n}\,d\sigma_{n}\bigg{|}\leq\|u_{n}\|_{L^{\infty}(\mathbb{R})}|\sigma_{n}|(\mathcal{O})+\|u_{n}\|_{L^{\infty}(\mathcal{O}^{c})}|\sigma_{n}|(\Omega).

The second term vanishes as nn\to\infty. For the first term, we note that

|σn|(𝒪)1γni=1|νn|(γnIi).|\sigma_{n}|(\mathcal{O})\leq\frac{1}{\gamma_{n}}\sum_{i=1}^{\ell}|\nu_{n}|(\gamma_{n}I_{i}).

Covering each interval γnIi\gamma_{n}I_{i} with εγn/\lceil\varepsilon\gamma_{n}/\ell\rceil many intervals of length 11, we use the given bound on νn+\nu_{n}^{+} to continue this estimate by

1γni=1|νn|(γnIi)1γni=1C(εγn+1)=Cε.\frac{1}{\gamma_{n}}\sum_{i=1}^{\ell}|\nu_{n}|(\gamma_{n}I_{i})\leq\frac{1}{\gamma_{n}}\sum_{i=1}^{\ell}C\Big{(}\frac{\varepsilon\gamma_{n}}{\ell}+1\Big{)}=C^{\prime}\varepsilon.

In conclusion,

lim supn|Ωun𝑑σn|Cε.\limsup_{n\to\infty}\bigg{|}\int_{\Omega}u_{n}\,d\sigma_{n}\bigg{|}\leq C\varepsilon.

Since ε>0\varepsilon>0 is arbitrary, Lemma 5.6 follows. ∎

Lemma 5.7.

Let νn𝒜n\nu_{n}\in\mathcal{A}_{n} be such that νnvν𝒜\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu\in\mathcal{A} as nn\to\infty. If

supn1supx0νn+([x,x+1])<,\sup_{n\geq 1}\sup_{x\geq 0}\nu_{n}^{+}([x,x+1])<\infty,

then (recall g(x)=xVg(x)=\int_{x}^{\infty}V)

0(Vνn)(x)𝑑xnΩg𝑑ν.\displaystyle\int_{-\infty}^{0}(V*\nu_{n})(x)\,dx\xrightarrow{n\to\infty}\int_{\Omega}g\,d\nu.
Proof.

We follow the proof in [GvMPS16] for the continuum setting and present it here in more detail. Take a continuous cut-off function

ψM:[0,)[0,1]withψM(x)={1if xM0if xM+1.\psi_{M}:[0,\infty)\to[0,1]\quad\text{with}\quad\psi_{M}(x)=\left\{\begin{array}[]{ll}1&\text{if }x\leq M\\ 0&\text{if }x\geq M+1.\end{array}\right. (44)

We recall from (25) that

0(Vνn)(x)𝑑x=Ωg𝑑νn=ΩgψM𝑑νn+Mg(1ψM)𝑑νn,\int_{-\infty}^{0}(V*\nu_{n})(x)\,dx=\int_{\Omega}g\,d\nu_{n}=\int_{\Omega}g\psi_{M}\,d\nu_{n}+\int_{M}^{\infty}g(1-\psi_{M})\,d\nu_{n}, (45)

where M>0M>0 is an arbitrary constant. Since gψMCc([0,))g\psi_{M}\in C_{c}([0,\infty)), we obtain from νnvν\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu that

ΩgψM𝑑νnnΩgψM𝑑ν.\int_{\Omega}g\psi_{M}\,d\nu_{n}\xrightarrow{n\to\infty}\int_{\Omega}g\psi_{M}\,d\nu.

For the second term in (45), we use (33) to estimate

|Mg(1ψM)𝑑νn|Mgd|νn|εMsupx0|νn|([x,x+1])CεM,\bigg{|}\int_{M}^{\infty}g(1-\psi_{M})\,d\nu_{n}\bigg{|}\leq\int_{M}^{\infty}g\,d|\nu_{n}|\leq\varepsilon_{M}\sup_{x\geq 0}|\nu_{n}|([x,x+1])\leq C\varepsilon_{M},

where εM0\varepsilon_{M}\to 0 as MM\to\infty. By a similar argument, it follows from ν𝒜\nu\in\mathcal{A} that

|Mg(1ψM)𝑑ν|CεM.\bigg{|}\int_{M}^{\infty}g(1-\psi_{M})\,d\nu\bigg{|}\leq C\varepsilon_{M}.

Tracing these observations back to (45), we conclude

0(Vνn)(x)𝑑xnΩgψM𝑑ν+Mg(1ψM)𝑑ν+CεM=Ωg𝑑ν+CεM.\int_{-\infty}^{0}(V*\nu_{n})(x)\,dx\xrightarrow{n\to\infty}\int_{\Omega}g\psi_{M}\,d\nu+\int_{M}^{\infty}g(1-\psi_{M})\,d\nu+C\varepsilon_{M}=\int_{\Omega}g\,d\nu+C\varepsilon_{M}.

Since MM is arbitrary, Lemma 5.7 follows. ∎

6 Proof of Theorem 1.1

This section is devoted to the proof of Theorem 1.1. Theorem 1.1 consists of three statements: compactness, the liminf inequality and the limsup inequality. We prove these three statements respectively in Sections 6.1, 6.2 and 6.3 for the power-law case a>0a>0 (see Assumption 3.2(ii)). In Section 6.4 we show that with minor modifications the proof for the logarithmic case a=0a=0 follows.

6.1 Compactness

Let νn𝒜n\nu_{n}\in\mathcal{A}_{n} be such that supn1Fn(νn)<\sup_{n\geq 1}F_{n}(\nu_{n})<\infty. We start from the expression for Fn(νn)F_{n}(\nu_{n}) in (24). Since the fourth and fifth term in (24) are nonnegative (recall (7)), we may neglect them. By Lemma 5.6 the third term is uniformly bounded. Hence, it is sufficient to focus on the first two terms in (24), which we label 𝖥n(νn)\mathsf{F}_{n}(\nu_{n}).

Recalling the regularization VβV^{\beta} defined in (36), we expand

𝖥n(νn)\displaystyle\mathsf{F}_{n}(\nu_{n}) =12ΔcV(xy)𝑑νn(y)𝑑νn(x)ρ(0)0(Vνn)\displaystyle=\frac{1}{2}\iint_{\Delta^{c}}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)-\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu_{n})
=12Δc(VVβ)(xy)𝑑νn(y)𝑑νn(x)+122Vβ(xy)𝑑νn(y)𝑑νn(x)\displaystyle=\frac{1}{2}\iint_{\Delta^{c}}(V-V^{\beta})(x-y)\,d\nu_{n}(y)d\nu_{n}(x)+\frac{1}{2}\iint_{\mathbb{R}^{2}}V^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)
n+12n2γn2Vβ(0)ρ(0)0(Vνn).\displaystyle\qquad-\frac{n+1}{2n^{2}}\gamma_{n}^{2}V^{\beta}(0)-\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu_{n}). (46)

For the second term, note from Lemma 5.4(iii) that

2Vβ(xy)𝑑νn(y)𝑑νn(x)=Tβνn22.\displaystyle\iint_{\mathbb{R}^{2}}V^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)=\|T^{\beta}\nu_{n}\|_{2}^{2}.

The first and third terms in (46) can be bounded from below by small constants. Indeed, using Vβ(0)C/βaV^{\beta}(0)\leq C/\beta^{a}, we bound the third term by

n+12n2γn2Vβ(0)Cγn2nβa.-\frac{n+1}{2n^{2}}\gamma_{n}^{2}V^{\beta}(0)\geq-C\frac{\gamma_{n}^{2}}{n}\beta^{-a}.

For the first term in (46), we recall Wβ=VVβW^{\beta}=V-V^{\beta} and expand νn=μ~nρ~\nu_{n}=\tilde{\mu}_{n}-\tilde{\rho}_{*}:

12ΔcWβ(xy)𝑑νn(y)𝑑νn(x)=12ΔcWβ(xy)𝑑μ~n(y)𝑑μ~n(x)Ω(Wβρ~)𝑑μ~n+12Ω(Wβρ~)ρ~.\frac{1}{2}\iint_{\Delta^{c}}W^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)\\ =\frac{1}{2}\iint_{\Delta^{c}}W^{\beta}(x-y)\,d\tilde{\mu}_{n}(y)d\tilde{\mu}_{n}(x)-\int_{\Omega}(W^{\beta}*\tilde{\rho}_{*})\,d\tilde{\mu}_{n}+\frac{1}{2}\int_{\Omega}(W^{\beta}*\tilde{\rho}_{*})\tilde{\rho}_{*}.

Since Wβ0W^{\beta}\geq 0, the first and third term are nonnegative. Using WβCβ1a\int_{\mathbb{R}}W^{\beta}\leq C\beta^{1-a}, we estimate the second term by

Ω(Wβρ~)𝑑μ~nΩ(Wβ)ρ~𝑑μ~nCγnβ1a.-\int_{\Omega}(W^{\beta}*\tilde{\rho}_{*})\,d\tilde{\mu}_{n}\geq-\int_{\Omega}\bigg{(}\int_{\mathbb{R}}W^{\beta}\bigg{)}\|\tilde{\rho}_{*}\|_{\infty}\,d\tilde{\mu}_{n}\geq-C\gamma_{n}\beta^{1-a}.

Hence, the first and third term in (46) are bounded from below by

Cβaγn(β+γnn).-C\beta^{-a}\gamma_{n}\Big{(}\beta+\frac{\gamma_{n}}{n}\Big{)}.

We choose β\beta such that this quantity is maximal. Up to a constant, this yields

β=βn:=γnn.\beta=\beta_{n}:=\frac{\gamma_{n}}{n}.

Then, by the assumption on γn\gamma_{n} in Theorem 1.1, we obtain

Cβnaγn(βn+γnn)=2Cγn(γnn)1a=:εnn0.\displaystyle C\beta_{n}^{-a}\gamma_{n}\Big{(}\beta_{n}+\frac{\gamma_{n}}{n}\Big{)}=\underbrace{2C\gamma_{n}\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{1-a}}_{=:\varepsilon_{n}}\xrightarrow{n\to\infty}0.

Collecting these estimates, we obtain from (46) that

𝖥n(νn)12Tβnνn22εnρ(0)0(Vνn).\displaystyle\mathsf{F}_{n}(\nu_{n})\geq\frac{1}{2}\|T^{\beta_{n}}\nu_{n}\|_{2}^{2}-\varepsilon_{n}-\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu_{n}). (47)

For the fourth term in (47), we use a rougher estimate. By Lemma 5.5,

ρ(0)0(Vνn)C(Tβnνn2+1)\rho_{\ast}(0)\int_{-\infty}^{0}(V*\nu_{n})\leq C(\|T^{\beta_{n}}\nu_{n}\|_{2}+1)

for all nn large enough. Then, together with the first two terms in (47), we obtain

𝖥n(νn)12Tβnνn22C.\displaystyle\mathsf{F}_{n}(\nu_{n})\geq\frac{1}{2}\|T^{\beta_{n}}\nu_{n}\|_{2}^{2}-C^{\prime}. (48)

By (48), Tβnνn2\|T^{\beta_{n}}\nu_{n}\|_{2} is bounded. This implies two useful properties. First,

Tβnνnϕ\displaystyle T^{\beta_{n}}\nu_{n}\rightharpoonup\phi

in L2()L^{2}(\mathbb{R}) as nn\to\infty along a subsequence (not relabelled) for some ϕL2()\phi\in L^{2}(\mathbb{R}). Then, by Lemma 5.5, supn1|νn|([0,M])\sup_{n\geq 1}|\nu_{n}|([0,M]) is bounded for any M>0M>0. Hence,

νnvν(Ω)\displaystyle\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu\in\mathcal{M}(\Omega)

along a further subsequence as nn\to\infty.

It is left to show that ν𝒜\nu\in\mathcal{A}, i.e.

νρ(0)andsupx0ν+([x,x+1])<.\displaystyle\nu^{-}\leq\rho_{\ast}(0)\mathcal{L}\quad\text{and}\quad\sup_{x\geq 0}\nu^{+}([x,x+1])<\infty. (49)

Since ρ~(x)=ρ(x/γn)\tilde{\rho}_{*}(x)=\rho_{\ast}(x/\gamma_{n}) and ρ\rho_{*} is continuous, it follows that

νn=ρ~vρ(0)\nu_{n}^{-}=\tilde{\rho}_{*}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\rho_{\ast}(0)\mathcal{L}

as nn\to\infty. Then, together with νnvν\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu, we obtain

νn+=νn+νnvν+ρ(0)\nu_{n}^{+}=\nu_{n}+\nu_{n}^{-}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu+\rho_{\ast}(0)\mathcal{L}

as nn\to\infty. In particular, ν+ρ(0)0\nu+\rho_{\ast}(0)\mathcal{L}\geq 0, which shows the first statement in (49). To prove the second statement, we take x0x\geq 0 arbitrary, and take a test function φCc(Ω)\varphi\in C_{c}(\Omega) which satisfies 0φ10\leq\varphi\leq 1 and

φ(y)={1if xyx+10if yx1 or yx+2.\varphi(y)=\left\{\begin{array}[]{ll}1&\text{if }x\leq y\leq x+1\\ 0&\text{if }y\leq x-1\text{ or }y\geq x+2.\end{array}\right.

Then,

ν+([x,x+1])\displaystyle\nu^{+}([x,x+1]) x1x+2φ𝑑ν+=x1x+2φ𝑑ν+x1x+2φ𝑑ν\displaystyle\leq\int_{x-1}^{x+2}\varphi\,d\nu^{+}=\int_{x-1}^{x+2}\varphi\,d\nu+\int_{x-1}^{x+2}\varphi\,d\nu^{-}
limnx1x+2φ𝑑νn+ρ(0)x1x+2φ(x)𝑑x\displaystyle\leq\lim_{n\to\infty}\int_{x-1}^{x+2}\varphi\,d\nu_{n}+\rho_{\ast}(0)\int_{x-1}^{x+2}\varphi(x)\,dx
lim supn(k=02|νn|([x+k1,x+k]))+3ρ(0),\displaystyle\leq\limsup_{n\to\infty}\bigg{(}\sum_{k=0}^{2}|\nu_{n}|([x+k-1,x+k])\bigg{)}+3\rho_{\ast}(0),

which by Lemma 5.5 and (48) is bounded uniformly in xx. This implies the second statement in (49).

6.2 Liminf inequality

To prove the liminf inequality in Theorem 1.1, let νnvν\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu be given. We may assume that Fn(νn)F_{n}(\nu_{n}) is bounded along a subsequence in nn (not relabelled), as otherwise the liminf inequality is trivial. Then, as in the compactness proof, we obtain (48), which shows that (Tβnνn)n(T^{\beta_{n}}\nu_{n})_{n} is bounded in L2()L^{2}(\mathbb{R}). Then, Lemma 5.5 implies

supn1supx0νn+([x,x+1])C.\sup_{n\geq 1}\sup_{x\geq 0}\nu_{n}^{+}([x,x+1])\leq C. (50)

Let 𝖥n\mathsf{F}_{n} be as in (46). First, we observe that

lim infnFn(νn)lim infn𝖥n(νn).\liminf_{n\to\infty}F_{n}(\nu_{n})\geq\liminf_{n\to\infty}\mathsf{F}_{n}(\nu_{n}). (51)

Indeed, in the compactness proof we already showed that the fourth and fifth term in (24) are nonnegative. By (50) and Lemma 5.6, the third term vanishes as nn\to\infty.

Next we bound the right-hand side in (51) from below. As in the proof for the compactness, we obtain (47). Together with Lemma 5.7 this yields

lim infn𝖥n(νn)12lim infnTβnνn22ρ(0)Ωg𝑑ν.\displaystyle\liminf_{n\to\infty}\mathsf{F}_{n}(\nu_{n})\geq\frac{1}{2}\liminf_{n\to\infty}\|T^{\beta_{n}}\nu_{n}\|_{2}^{2}-\rho_{\ast}(0)\int_{\Omega}g\,d\nu.

In particular, TβnνnϕT^{\beta_{n}}\nu_{n}\rightharpoonup\phi in L2()L^{2}(\mathbb{R}) as nn\to\infty along a subsequence to some ϕL2()\phi\in L^{2}(\mathbb{R}). It is left to prove that ϕ=Tν\phi=T\nu. With this aim, let φCc()\varphi\in C_{c}^{\infty}(\mathbb{R}) be a test function. Using (35) and Lemma 5.4(i) we obtain

φ(Tβnνn)=Ω(Tβnφ)𝑑νn=0M+1(Tβnφ)ψM𝑑νn+M(Tβnφ)(1ψM)𝑑νn,\displaystyle\int_{\mathbb{R}}\varphi(T^{\beta_{n}}\nu_{n})=\int_{\Omega}(T^{\beta_{n}}\varphi)\,d\nu_{n}=\int_{0}^{M+1}(T^{\beta_{n}}\varphi)\psi_{M}\,d\nu_{n}+\int_{M}^{\infty}(T^{\beta_{n}}\varphi)(1-\psi_{M})\,d\nu_{n}, (52)

where ψM\psi_{M} is the cut-off function introduced in (44), and M>0M>0 is an arbitrary constant. We pass to the limit nn\to\infty in both terms separately.

For the first term in (52), we note from Lemma 5.4(iv) that (Tβnφ)ψM(Tφ)ψM(T^{\beta_{n}}\varphi)\psi_{M}\to(T\varphi)\psi_{M} in C([0,M+1])C([0,M+1]) as nn\to\infty. Together with νnvν\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu this yields

0M+1(Tβnφ)ψM𝑑νnn0M+1(Tφ)ψM𝑑ν.\int_{0}^{M+1}(T^{\beta_{n}}\varphi)\psi_{M}\,d\nu_{n}\xrightarrow{n\to\infty}\int_{0}^{M+1}(T\varphi)\psi_{M}\,d\nu.

For the second term in (52), we set ψ¯M:=1ψM\overline{\psi}_{M}:=1-\psi_{M} and obtain from (34), Lemma 5.5, Tβnνn2C\|T^{\beta_{n}}\nu_{n}\|_{2}\leq C and the triangle inequality that

|M(Tβnφ)ψ¯M𝑑νn|\displaystyle\bigg{|}\int_{M}^{\infty}(T^{\beta_{n}}\varphi)\overline{\psi}_{M}\,d\nu_{n}\bigg{|} Cψ¯MTβnφX1,2supx0|νn|([x,x+1])\displaystyle\leq C\big{\|}\overline{\psi}_{M}T^{\beta_{n}}\varphi\big{\|}_{X_{1,2}}\sup_{x\geq 0}|\nu_{n}|([x,x+1])
C(ψ¯M(TTβn)φX1,2+ψ¯MTφX1,2).\displaystyle\leq C^{\prime}\big{(}\|\overline{\psi}_{M}(T-T^{\beta_{n}})\varphi\|_{X_{1,2}}+\|\overline{\psi}_{M}T\varphi\|_{X_{1,2}}\big{)}. (53)

For the second term, we set f:=Tφf:=T\varphi and compute

fψ¯MX1,22\displaystyle\|f\overline{\psi}_{M}\|_{X_{1,2}}^{2} =M(x4+1)f(x)2ψ¯M(x)2𝑑x+M(fψ¯M)(x)2𝑑x\displaystyle=\int_{M}^{\infty}(x^{4}+1)f(x)^{2}\overline{\psi}_{M}(x)^{2}\,dx+\int_{M}^{\infty}\big{(}f\overline{\psi}_{M}\big{)}^{\prime}(x)^{2}\,dx
M(x4+1)f(x)2𝑑x+Cψ¯MW1,()2M(f(x)2+f(x)2)𝑑x.\displaystyle\leq\int_{M}^{\infty}(x^{4}+1)f(x)^{2}\,dx+C\|\overline{\psi}_{M}\|_{W^{1,\infty}(\mathbb{R})}^{2}\int_{M}^{\infty}\big{(}f(x)^{2}+f^{\prime}(x)^{2}\big{)}\,dx.

Since fX1,2f\in X_{1,2}, this value vanishes as MM\to\infty. For the first term in (53), we obtain from Lemma 5.4(iv) that

ψ¯M(TTβn)φX1,2Cψ¯MW1,()(TTβn)φX1,2Cβn1a,\displaystyle\|\overline{\psi}_{M}(T-T^{\beta_{n}})\varphi\|_{X_{1,2}}\leq C\|\overline{\psi}_{M}\|_{W^{1,\infty}(\mathbb{R})}\|(T-T^{\beta_{n}})\varphi\|_{X_{1,2}}\leq C^{\prime}\beta_{n}^{1-a},

which vanishes as nn\to\infty uniformly in MM. In conclusion, by tracing these observations back to (52), we obtain

φ(Tβnνn)n0M+1(Tφ)ψM𝑑ν+cM,\displaystyle\int_{\mathbb{R}}\varphi(T^{\beta_{n}}\nu_{n})\xrightarrow{n\to\infty}\int_{0}^{M+1}(T\varphi)\psi_{M}\,d\nu+c_{M}, (54)

where cM0c_{M}\to 0 as MM\to\infty.

Next we pass to the limit MM\to\infty. Since also supx0|ν|([x,x+1])\sup_{x\geq 0}|\nu|([x,x+1]) is bounded, the computation in (53) shows that

|M(Tφ)ψ¯M𝑑ν|C(Tφ)ψ¯MX1,2M0.\bigg{|}\int_{M}^{\infty}(T\varphi)\overline{\psi}_{M}\,d\nu\bigg{|}\leq C\|(T\varphi)\overline{\psi}_{M}\|_{X_{1,2}}\xrightarrow{M\to\infty}0.

Hence, the right-hand side in (54) can be cast into

0M+1(Tφ)ψM𝑑ν+M(Tφ)(1ψM)𝑑ν+o(1)=(Tφ)𝑑ν+o(1)\int_{0}^{M+1}(T\varphi)\psi_{M}\,d\nu+\int_{M}^{\infty}(T\varphi)(1-\psi_{M})\,d\nu+o(1)=\int_{\mathbb{R}}(T\varphi)\,d\nu+o(1)

as MM\to\infty. Finally, by Lemma 5.1 we conclude that ϕ=Tν.\phi=T\nu.

6.3 Limsup inequality

We first assume ρ(0)>0\rho_{*}(0)>0 and treat the special case ρ(0)=0\rho_{*}(0)=0 afterwards. Since FF is the same as in [GvMPS16] and only depends on UU through the constant ρ(0)\rho_{*}(0), we may use the density result in [GvMPS16]. This result states that it is sufficient to prove the limsup inequality only for those ν𝒜\nu\in\mathcal{A} for which νL2()\nu\in L^{2}(\mathbb{R}), suppν[0,M]\operatorname{supp}\nu\subset[0,M] for some M>0M>0 and νρ(0)δ\nu^{-}\leq\rho_{*}(0)-\delta on [0,M][0,M] for some δ>0\delta>0. In particular, we treat ν\nu as a density. Since FF is continuous as a functional on L2()L^{2}(\mathbb{R}), we may further assume that νC1(Ω)\nu\in C^{1}(\Omega).

We first treat the case Ων0\int_{\Omega}\nu\geq 0, and comment on the case Ων<0\int_{\Omega}\nu<0 afterwards. To choose νn𝒜n\nu_{n}\in\mathcal{A}_{n}, we note from (17) that 𝒜n\mathcal{A}_{n} can be parametrized by 𝐱Ωn\mathbf{x}\in\Omega_{n} through

νn=γnni=0nδxiρ~,\nu_{n}=\frac{\gamma_{n}}{n}\sum_{i=0}^{n}\delta_{x_{i}}-\tilde{\rho}_{*}, (55)

where we recall that ρ~(x)=ρ(x/γn)\tilde{\rho}_{*}(x)=\rho_{*}(x/\gamma_{n}) depends on nn. We choose xix_{i} such that

xi1xiν(x)+ρ~(x)dx=γnnfor all i=1,,n.\int_{x_{i-1}}^{x_{i}}\nu(x)+\tilde{\rho}_{*}(x)\,dx=\frac{\gamma_{n}}{n}\qquad\text{for all }i=1,\ldots,n. (56)

Note from Ω(ν+ρ~)=γn+Ωνγn\int_{\Omega}(\nu+\tilde{\rho}_{*})=\gamma_{n}+\int_{\Omega}\nu\geq\gamma_{n} that such an 𝐱Ωn\mathbf{x}\in\Omega_{n} exists. By taking nn large enough, we may further assume that ν+ρ~0\nu+\tilde{\rho}_{*}\geq 0 and that

xisuppρ~for all i=0,,n.\displaystyle x_{i}\in\operatorname{supp}\tilde{\rho}_{*}\qquad\text{for all }i=0,\ldots,n. (57)

Next we prove several properties of νn\nu_{n} and 𝐱\mathbf{x}. We observe from (56) that

xixi1γnn(ν+ρ)for all i=1,,n.\displaystyle x_{i}-x_{i-1}\geq\frac{\gamma_{n}}{n\big{(}\|\nu\|_{\infty}+\|\rho_{*}\|_{\infty}\big{)}}\qquad\text{for all }i=1,\ldots,n. (58)

Hence,

supx0νn+([x,x+1])=supx0(#{xi[x,x+1]})γnnC.\sup_{x\geq 0}\nu_{n}^{+}([x,x+1])=\sup_{x\geq 0}\big{(}\#\{x_{i}\in[x,x+1]\}\big{)}\frac{\gamma_{n}}{n}\leq C. (59)

To prove

νnvνas n,\nu_{n}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu\qquad\text{as }n\to\infty, (60)

we take a test function φCc(Ω)\varphi\in C_{c}(\Omega) and set N:=maxsuppφN:=\max\operatorname{supp}\varphi. Taking nn large enough such that ρ~δ2\tilde{\rho}_{*}\geq\frac{\delta}{2} on [0,N+1][0,N+1] and ν+ρ~δ2\nu+\tilde{\rho}_{*}\geq\frac{\delta}{2} on [0,M][0,M], we note from (56) that

xixi12δγnnfor all i such that xi1<N,\displaystyle x_{i}-x_{i-1}\leq\frac{2}{\delta}\frac{\gamma_{n}}{n}\qquad\text{for all $i$ such that }x_{i-1}<N, (61)

which vanishes as nn\to\infty. Then, from

Ωφd(νnν)\displaystyle\int_{\Omega}\varphi\,d(\nu_{n}-\nu) =0Nφ𝑑μ~n0Nφ(ν+ρ~)\displaystyle=\int_{0}^{N}\varphi\,d\tilde{\mu}_{n}-\int_{0}^{N}\varphi(\nu+\tilde{\rho}_{*})
=i:xi1<Nxi1xi(φ(xi1)φ(x))(ν+ρ~)(x)𝑑x\displaystyle=\sum_{i\,:\,x_{i-1}<N}\int_{x_{i-1}}^{x_{i}}\big{(}\varphi(x_{i-1})-\varphi(x)\big{)}(\nu+\tilde{\rho}_{*})(x)\,dx
maxi:xi1<N(maxx[xi1,xi]|φ(xi1)φ(x)|)0N(ν+ρ~)\displaystyle\leq\max_{i\,:\,x_{i-1}<N}\Big{(}\max_{x\in[x_{i-1},x_{i}]}\big{|}\varphi(x_{i-1})-\varphi(x)\big{|}\Big{)}\int_{0}^{N}(\nu+\tilde{\rho}_{*})

we obtain by the continuity of φ\varphi that the right-hand side vanishes as nn\to\infty. This proves (60).

Next we prove the limsup inequality in Theorem 1.1 for νn\nu_{n} as constructed above. With this aim, we treat all five terms of FnF_{n} in (24) separately. The latter four terms all converge as nn\to\infty. Indeed, the fifth term in (24) vanishes as nn\to\infty. By (57) the fourth term equals 0. Due to (59), Lemma 5.6 implies that the third term vanishes as nn\to\infty. Due to (59) and (60), Lemma 5.7 guarantees the convergence of the second term.

Therefore, it is sufficient to prove the limsup inequality only for the first term in (24). We first show that for all β>0\beta>0 small enough

lim supnΔcV(xy)𝑑νn(y)𝑑νn(x)Ω(Vβν)𝑑ν+Cβ1a.\limsup_{n\to\infty}\iint_{\Delta^{c}}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)\leq\int_{\Omega}(V^{\beta}*\nu)d\nu+C\beta^{1-a}. (62)

Let II be the smallest integer for which xIMx_{I}\geq M. Using that VV is even, we expand

ΔcV(xy)𝑑νn(y)𝑑νn(x)\displaystyle\iint_{\Delta^{c}}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)
=[0,xI]2Vβ(xy)𝑑νn(y)𝑑νn(x)(I+1)(γnn)2Vβ(0)\displaystyle=\iint_{[0,x_{I}]^{2}}V^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)-(I+1)\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}V^{\beta}(0) (63a)
+[0,xI]2ΔWβ(xy)𝑑νn(y)𝑑νn(x)\displaystyle\quad+\iint_{[0,x_{I}]^{2}\setminus\Delta}W^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x) (63b)
+2(xI,xn][0,xI]V(xy)𝑑νn(y)𝑑νn(x)\displaystyle\quad+2\int_{(x_{I},x_{n}]}\int_{[0,x_{I}]}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x) (63c)
+(xI,xn]2ΔV(xy)𝑑νn(y)𝑑νn(x)\displaystyle\quad+\iint_{(x_{I},x_{n}]^{2}\setminus\Delta}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x) (63d)
+2[0,xn](xn,)V(xy)𝑑νn(y)𝑑νn(x)\displaystyle\quad+2\int_{[0,x_{n}]}\int_{(x_{n},\infty)}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x) (63e)
+(xn,)2ΔV(xy)𝑑νn(y)𝑑νn(x).\displaystyle\quad+\iint_{(x_{n},\infty)^{2}\setminus\Delta}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x). (63f)

The second term in (63a) is negative; we simply bound it from above by 0. By (60) and ν,ρL()\nu,\rho_{*}\in L^{\infty}(\mathbb{R}), we obtain that νn|[0,xI]vν|[0,M]\nu_{n}|_{[0,x_{I}]}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}\nu|_{[0,M]} as nn\to\infty. By [AFP00, Thm. 1.59] we then also have that (νnνn)|[0,xI]2v(νν)|[0,M]2(\nu_{n}\otimes\nu_{n})|_{[0,x_{I}]^{2}}\stackrel{{\scriptstyle v}}{{\rightharpoonup}}(\nu\otimes\nu)|_{[0,M]^{2}} as nn\to\infty. Since VβV^{\beta} is continuous, this implies for the first term in (63a) that

[0,xI]2Vβ(xy)𝑑νn(y)𝑑νn(x)n[0,M]2Vβ(xy)𝑑ν(y)𝑑ν(x),\iint_{[0,x_{I}]^{2}}V^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)\xrightarrow{n\to\infty}\iint_{[0,M]^{2}}V^{\beta}(x-y)\,d\nu(y)d\nu(x),

which equals the integral in the right-hand side of (62).

Next we bound (63b). Neglecting the negative cross terms, we estimate

[0,xI]2ΔWβ(xy)𝑑νn(y)𝑑νn(x)[0,xI]2ΔWβ(xy)𝑑νn+(y)𝑑νn+(x)+[0,xI]2Wβ(xy)𝑑ρ~(y)𝑑ρ~(x).\iint_{[0,x_{I}]^{2}\setminus\Delta}W^{\beta}(x-y)\,d\nu_{n}(y)d\nu_{n}(x)\\ \leq\iint_{[0,x_{I}]^{2}\setminus\Delta}W^{\beta}(x-y)\,d\nu_{n}^{+}(y)d\nu_{n}^{+}(x)+\iint_{[0,x_{I}]^{2}}W^{\beta}(x-y)\,d\tilde{\rho}_{*}(y)d\tilde{\rho}_{*}(x). (64)

For the second term, we recall from Lemma 5.3(v) that WCβ1a\int_{\mathbb{R}}W\leq C\beta^{1-a}. Since xI<M+1x_{I}<M+1 for nn large enough, we obtain

[0,xI]2Wβ(xy)𝑑ρ~(y)𝑑ρ~(x)0M+1(Wβ)ρ𝑑ρ~(x)C(M+1)ρ2β1a.\displaystyle\iint_{[0,x_{I}]^{2}}W^{\beta}(x-y)\,d\tilde{\rho}_{*}(y)d\tilde{\rho}_{*}(x)\leq\int_{0}^{M+1}\bigg{(}\int_{\mathbb{R}}W^{\beta}\bigg{)}\|\rho_{*}\|_{\infty}\,d\tilde{\rho}_{*}(x)\leq C(M+1)\|\rho_{*}\|_{\infty}^{2}\beta^{1-a}.

The first term in (64) is the discrete counterpart of the second term, and can be treated similarly. Relying on (58) and I=O(n/γn)I=O(n/\gamma_{n}), we bound it by

[0,xI]2ΔWβ(xy)𝑑νn+(y)𝑑νn+(x)=2(γnn)2i=0I1k=1IiWβ(xi+kxi)Cγnni=0I1k=1cγnnWβ(cγnnk)C0Wβ=C′′β1a.\iint_{[0,x_{I}]^{2}\setminus\Delta}W^{\beta}(x-y)\,d\nu_{n}^{+}(y)d\nu_{n}^{+}(x)=2\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}\sum_{i=0}^{I-1}\sum_{k=1}^{I-i}W^{\beta}(x_{i+k}-x_{i})\\ \leq C\frac{\gamma_{n}}{n}\sum_{i=0}^{I-1}\sum_{k=1}^{\infty}c\frac{\gamma_{n}}{n}W^{\beta}\Big{(}c\frac{\gamma_{n}}{n}k\Big{)}\leq C^{\prime}\int_{0}^{\infty}W^{\beta}=C^{\prime\prime}\beta^{1-a}.

Hence, (63b) is bounded by Cβ1aC\beta^{1-a} uniformly in nn. In view of (62), it is therefore left to show that the limsup of the remaining terms in (63) are nonpositive.

We start with (63d). Expanding νn=νn+ρ~\nu_{n}=\nu_{n}^{+}-\tilde{\rho}_{*} and using that VV is even, we rewrite

(xI,xn]2ΔV(xy)𝑑νn(y)𝑑νn(x)\displaystyle\iint_{(x_{I},x_{n}]^{2}\setminus\Delta}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)
=2γnni=I+1n(γnnj=I+1i1V(xixj)j=I+1ixj1xjV(xix)ρ~(x)𝑑x)\displaystyle=2\frac{\gamma_{n}}{n}\sum_{i=I+1}^{n}\bigg{(}\frac{\gamma_{n}}{n}\sum_{j=I+1}^{i-1}V(x_{i}-x_{j})-\sum_{j=I+1}^{i}\int_{x_{j-1}}^{x_{j}}V(x_{i}-x)\tilde{\rho}_{*}(x)\,dx\bigg{)}
+2i=I+1nxi1xi(j=I+1i1xj1xjV(xy)ρ~(y)𝑑yγnnj=I+1i1V(xxj))ρ~(x)𝑑x\displaystyle\quad+2\sum_{i=I+1}^{n}\int_{x_{i-1}}^{x_{i}}\bigg{(}\sum_{j=I+1}^{i-1}\int_{x_{j-1}}^{x_{j}}V(x-y)\tilde{\rho}_{*}(y)\,dy-\frac{\gamma_{n}}{n}\sum_{j=I+1}^{i-1}V(x-x_{j})\bigg{)}\tilde{\rho}_{*}(x)\,dx
+i=I+1nxi1xixi1xiV(xy)ρ~(y)ρ~(x)𝑑x\displaystyle\quad+\sum_{i=I+1}^{n}\int_{x_{i-1}}^{x_{i}}\int_{x_{i-1}}^{x_{i}}V(x-y)\tilde{\rho}_{*}(y)\tilde{\rho}_{*}(x)\,dx
=:T1+T2+T3.\displaystyle=:T_{1}+T_{2}+T_{3}. (65)

Using (56) and the fact that VV is decreasing, we obtain for the integrals inside the parentheses that

xj1xjV(xix)ρ~(x)𝑑xV(xixj1)xj1xjρ~(x)𝑑x=γnnV(xixj1)\int_{x_{j-1}}^{x_{j}}V(x_{i}-x)\tilde{\rho}_{*}(x)\,dx\geq V(x_{i}-x_{j-1})\int_{x_{j-1}}^{x_{j}}\tilde{\rho}_{*}(x)\,dx=\frac{\gamma_{n}}{n}V(x_{i}-x_{j-1}) (66)

and, similarly,

xj1xjV(xy)ρ~(y)𝑑yγnnV(xxj)\int_{x_{j-1}}^{x_{j}}V(x-y)\tilde{\rho}_{*}(y)\,dy\leq\frac{\gamma_{n}}{n}V(x-x_{j})

for all x(xi1,xi)x\in(x_{i-1},x_{i}). Hence, T1,T20T_{1},T_{2}\leq 0. For T3T_{3}, we observe from V(x)C|x|aV(x)\leq C|x|^{-a} and (56) that

xi1xiV(xy)ρ~(y)𝑑y20γn/(2nρ)V(z)ρ𝑑zC0γn/(2nρ)1za𝑑z=C(γnn)1a.\int_{x_{i-1}}^{x_{i}}V(x-y)\tilde{\rho}_{*}(y)dy\leq 2\int_{0}^{\gamma_{n}/(2n\|\rho_{*}\|_{\infty})}V(z)\|\rho_{*}\|_{\infty}\,dz\\ \leq C\int_{0}^{\gamma_{n}/(2n\|\rho_{*}\|_{\infty})}\frac{1}{z^{a}}\,dz=C^{\prime}\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{1-a}. (67)

Hence, T3Cγn2a/n1aT_{3}\leq C^{\prime}\gamma_{n}^{2-a}/n^{1-a}, which by the assumption on γn\gamma_{n} in Theorem 1.1 vanishes as nn\to\infty. In conclusion, the limsup of (63d) is nonpositive.

For (63f), we note that (xn,)(x_{n},\infty) is disjoint with suppνn+\operatorname{supp}\nu_{n}^{+}. Hence,

(xn,)2ΔV(xy)𝑑νn(y)𝑑νn(x)=(xn,)2V(xy)ρ~(y)ρ~(x)𝑑y𝑑x.\iint_{(x_{n},\infty)^{2}\setminus\Delta}V(x-y)\,d\nu_{n}(y)d\nu_{n}(x)=\iint_{(x_{n},\infty)^{2}}V(x-y)\tilde{\rho}_{*}(y)\tilde{\rho}_{*}(x)\,dydx. (68)

Then, since

xnρ~=ρ~0xn(ρ~+ν)+0xnν=Ων=C0,\int_{x_{n}}^{\infty}\tilde{\rho}_{*}=\int_{\mathbb{R}}\tilde{\rho}_{*}-\int_{0}^{x_{n}}(\tilde{\rho}_{*}+\nu)+\int_{0}^{x_{n}}\nu=\int_{\Omega}\nu=C\geq 0, (69)

we get xn/γnρ=C/γn\int_{x_{n}/\gamma_{n}}^{\infty}\rho_{*}=C/\gamma_{n}. Since ρ\rho_{*} is Lipschitz continuous, this implies that

ρ~L(xn,)=ρL(xn/γn,)C/γn.\|\tilde{\rho}_{*}\|_{L^{\infty}(x_{n},\infty)}=\|\rho_{*}\|_{L^{\infty}(x_{n}/\gamma_{n},\infty)}\leq C^{\prime}/\sqrt{\gamma_{n}}.

Then, similarly to (67), we estimate the right-hand side of (68) as

(xn,)2V(xy)ρ~(y)ρ~(x)𝑑y𝑑xxn(20CγnV(z)Cγn𝑑z)ρ~(x)𝑑xC′′γn3/2V1,\displaystyle\iint_{(x_{n},\infty)^{2}}V(x-y)\tilde{\rho}_{*}(y)\tilde{\rho}_{*}(x)\,dydx\leq\int_{x_{n}}^{\infty}\bigg{(}2\int_{0}^{C\sqrt{\gamma_{n}}}V(z)\frac{C^{\prime}}{\sqrt{\gamma_{n}}}\,dz\bigg{)}\tilde{\rho}_{*}(x)\,dx\leq\frac{C^{\prime\prime}}{\gamma_{n}^{3/2}}\|V\|_{1},

which vanishes as nn\to\infty.

Next we treat (63c). We split νn=(νnν)+ν\nu_{n}=(\nu_{n}-\nu)+\nu in the inner integral. For the first part, we expand as in (65),

(xI,xn][0,xI]V(xy)d(νnν)(y)𝑑νn(x)\displaystyle\int_{(x_{I},x_{n}]}\int_{[0,x_{I}]}V(x-y)\,d(\nu_{n}-\nu)(y)d\nu_{n}(x)
=γnni=I+1n(γnnj=0IV(xixj)j=1Ixj1xjV(xix)(ν+ρ~)(x)𝑑x)\displaystyle=\frac{\gamma_{n}}{n}\sum_{i=I+1}^{n}\bigg{(}\frac{\gamma_{n}}{n}\sum_{j=0}^{I}V(x_{i}-x_{j})-\sum_{j=1}^{I}\int_{x_{j-1}}^{x_{j}}V(x_{i}-x)(\nu+\tilde{\rho}_{*})(x)\,dx\bigg{)}
+i=I+1nxi1xi(j=1Ixj1xjV(xy)(ν+ρ~)(y)𝑑yγnnj=0IV(xxj))ρ~(x)𝑑x\displaystyle\quad+\sum_{i=I+1}^{n}\int_{x_{i-1}}^{x_{i}}\bigg{(}\sum_{j=1}^{I}\int_{x_{j-1}}^{x_{j}}V(x-y)(\nu+\tilde{\rho}_{*})(y)\,dy-\frac{\gamma_{n}}{n}\sum_{j=0}^{I}V(x-x_{j})\bigg{)}\tilde{\rho}_{*}(x)\,dx
=:T4+T5.\displaystyle=:T_{4}+T_{5}.

A similar argument as that in (66) yields T50T_{5}\leq 0 and

T4γnni=I+1nγnnV(xixI).T_{4}\leq\frac{\gamma_{n}}{n}\sum_{i=I+1}^{n}\frac{\gamma_{n}}{n}V(x_{i}-x_{I}).

Noting from (58) that

i=I+1nV(xixI)k=1V(ckγn/n)1cnγn0V=Cnγn,\sum_{i=I+1}^{n}V(x_{i}-x_{I})\leq\sum_{k=1}^{\infty}V(ck\gamma_{n}/n)\leq\frac{1}{c}\frac{n}{\gamma_{n}}\int_{0}^{\infty}V=C\frac{n}{\gamma_{n}},

we obtain lim supnT40\limsup_{n\to\infty}T_{4}\leq 0.

The second part of (63c) equals (setting f:=VνLip(Ω)f:=V*\nu\in\operatorname{Lip}(\Omega))

(xI,xn][0,xI]V(xy)ν(y)𝑑νn(x)=(xI,xn]f𝑑νn=i=I+1nxi1xi(f(xi)f(x))ρ~(x)𝑑x.\int_{(x_{I},x_{n}]}\int_{[0,x_{I}]}V(x-y)\nu(y)\,d\nu_{n}(x)=\int_{(x_{I},x_{n}]}f\,d\nu_{n}=\sum_{i=I+1}^{n}\int_{x_{i-1}}^{x_{i}}\big{(}f(x_{i})-f(x)\big{)}\tilde{\rho}_{*}(x)\,dx. (70)

To estimate this in absolute value, we split the sum in two parts. Let J=n/γnJ=\lceil n/\sqrt{\gamma_{n}}\rceil and take nn large enough such that JI1cn/γnJ-I-1\geq cn/\sqrt{\gamma_{n}} (recall that I=O(n/γn)I=O(n/\gamma_{n})) and such that (61) holds for all iJi\leq J. Then,

i=I+1Jxi1xi|f(xi)f(x)|ρ~(x)𝑑xCi=I+1Jxi1xi|xixi1|ρ~(x)𝑑xCi=I+1J(γnn)2Cγn3/2nn0\sum_{i=I+1}^{J}\int_{x_{i-1}}^{x_{i}}\big{|}f(x_{i})-f(x)\big{|}\tilde{\rho}_{*}(x)\,dx\leq C\sum_{i=I+1}^{J}\int_{x_{i-1}}^{x_{i}}|x_{i}-x_{i-1}|\tilde{\rho}_{*}(x)\,dx\\ \leq C\sum_{i=I+1}^{J}\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}\leq C^{\prime}\frac{\gamma_{n}^{3/2}}{n}\xrightarrow{n\to\infty}0

and, recalling (58),

i=J+1nxi1xi|f(xi)f(x)|ρ~(x)𝑑xi=J+1nxi1xi2fL(xi1,xi)ρ~(x)𝑑x\displaystyle\sum_{i=J+1}^{n}\int_{x_{i-1}}^{x_{i}}\big{|}f(x_{i})-f(x)\big{|}\tilde{\rho}_{*}(x)\,dx\leq\sum_{i=J+1}^{n}\int_{x_{i-1}}^{x_{i}}2\|f\|_{L^{\infty}(x_{i-1},x_{i})}\tilde{\rho}_{*}(x)\,dx
i=J+1n2γnnV(xi1xI)Ω|ν|Ci=J+1ncγnnV(cγnn(i1I))\displaystyle\leq\sum_{i=J+1}^{n}2\frac{\gamma_{n}}{n}V(x_{i-1}-x_{I})\int_{\Omega}|\nu|\leq C\sum_{i=J+1}^{n}c\frac{\gamma_{n}}{n}V\Big{(}c\frac{\gamma_{n}}{n}(i-1-I)\Big{)}
Cc(JI1)γn/nVCcγnVn0.\displaystyle\leq C\int_{c(J-I-1)\gamma_{n}/n}^{\infty}V\leq C\int_{c^{\prime}\sqrt{\gamma_{n}}}^{\infty}V\xrightarrow{n\to\infty}0.

Hence, the limsup of the second part of (63c) is nonpositive. We conclude that the limsup of (63c) is nonpositive.

Finally, for (63e), we observe that the inner integral

(xn,)V(xy)𝑑νn(y)=(Vρ~|(xn,))(x)\int_{(x_{n},\infty)}V(x-y)\,d\nu_{n}(y)=-(V*\tilde{\rho}_{*}|_{(x_{n},\infty)})(x)

is nonpositive and non-increasing on [0,xn][0,x_{n}]. Then, using a similar argument as for (63c) (simplifications are possible), we conclude that the limsup of (63e) is nonpositive. This completes the proof of (62).

In conclusion, we have proved that

lim supnFn(νn)12Ω(Vβν)𝑑ν+Cβ1aρ(0)Ωg𝑑ν\limsup_{n\to\infty}F_{n}(\nu_{n})\leq\frac{1}{2}\int_{\Omega}(V^{\beta}*\nu)d\nu+C\beta^{1-a}-\rho_{\ast}(0)\int_{\Omega}g\,d\nu

for all β>0\beta>0 small enough. Since νL2()\nu\in L^{2}(\mathbb{R}) and VβVV^{\beta}\to V in L1()L^{1}(\mathbb{R}) as β0\beta\to 0, we conclude the limsup inequality in Theorem 1.1 by taking β0\beta\to 0 and applying (29).

It is left to treat the cases ρ(0)=0\rho_{*}(0)=0 and Ων<0\int_{\Omega}\nu<0. Since ρ(0)=0\rho_{*}(0)=0 implies ν0\nu\geq 0, these two cases are mutually exclusive. We start with the case ρ(0)=0\rho_{*}(0)=0. We follow the proof for the case ρ(0)>0\rho_{*}(0)>0 with minor modifications. By density, instead of assuming νρ(0)δ\nu^{-}\leq\rho_{*}(0)-\delta, we may assume that νψδ\nu\geq\psi^{\delta} on Ω\Omega, where ψδ\psi^{\delta} is a monotone cut-off functions which equals δ\delta on [0,Mδ][0,M-\delta] and 0 on [M,)[M,\infty). Under this assumption, (61) does not hold for large ii, and thus we need to construct a different argument for (60) and for showing that the limsup of (63c) is nonpositive.

To prove (60), we take a test function φCc(Ω)\varphi\in C_{c}(\Omega) and set N:=maxsuppφN:=\max\operatorname{supp}\varphi. From the proof of (60) we observe that νψδ\nu\geq\psi^{\delta} on Ω\Omega implies that [0,M]φd(νnν)0\int_{[0,M]}\varphi\,d(\nu_{n}-\nu)\to 0 as nn\to\infty. For the integral over [M,)[M,\infty), we note from the Lipschitz continuity of ρ\rho_{*} and ρ(0)=0\rho_{\ast}(0)=0 that

MNρ~0Nρ(xγn)𝑑xC0Nxγn𝑑xCγnn0.\int_{M}^{N}\tilde{\rho}_{*}\leq\int_{0}^{N}\rho_{*}\Big{(}\frac{x}{\gamma_{n}}\Big{)}\,dx\leq C\int_{0}^{N}\frac{x}{\gamma_{n}}\,dx\leq\frac{C^{\prime}}{\gamma_{n}}\xrightarrow{n\to\infty}0.

Then, by (56),

(M,N]𝑑νn+γnn+MNρ~n0.\int_{(M,N]}d\nu_{n}^{+}\leq\frac{\gamma_{n}}{n}+\int_{M}^{N}\tilde{\rho}_{*}\xrightarrow{n\to\infty}0.

Hence,

(M,)φd(νnν)=(M,N]φd(νn+ρ~)φ(M,N]d(νn++ρ~)n0.\int_{(M,\infty)}\varphi\,d(\nu_{n}-\nu)=\int_{(M,N]}\varphi\,d(\nu_{n}^{+}-\tilde{\rho}_{*})\leq\|\varphi\|_{\infty}\int_{(M,N]}d(\nu_{n}^{+}+\tilde{\rho}_{*})\xrightarrow{n\to\infty}0.

We conclude (60).

Next we show that the limsup of (63c) is nonpositive. We can follow the proof for the case ρ(0)>0\rho_{*}(0)>0 up to the term (xI,xn]f𝑑νn\int_{(x_{I},x_{n}]}f\,d\nu_{n} in (70). Since ρ(0)=0\rho_{*}(0)=0 implies ν0\nu\geq 0, we observe that f=Vνf=V*\nu is nonnegative and non-increasing on (xI,)(x_{I},\infty). Hence, by a similar argument as that in (66), it follows that (xI,xn]f𝑑νn0\int_{(x_{I},x_{n}]}f\,d\nu_{n}\leq 0. This concludes the proof of the limsup inequality in Theorem 1.1 in the case ρ(0)=0\rho_{*}(0)=0.

It is left to prove the limsup inequality for the case Ων<0\int_{\Omega}\nu<0. We largely follow the proof for the case Ων0\int_{\Omega}\nu\geq 0, and focus on the modifications. For the choice of νn\nu_{n}, we define x0,,xJx_{0},\ldots,x_{J} as in (56), where JJ is the smallest integer at which

xJρ~γnn.\int_{x_{J}}^{\infty}\tilde{\rho}_{*}\leq\frac{\gamma_{n}}{n}.

For i>Ji>J we set

xi=xJ+(iJ)γn3/2n,x_{i}=x_{J}+(i-J)\frac{\gamma_{n}^{3/2}}{n}, (71)

and take νn\nu_{n} as in (55). Clearly, properties (58), (59) and (60) are still satisfied. To obtain the discrete equivalent of (69), we compute

γnnJ=νn+([0,xJ))=[0,xJ)(ν+ρ~)=Ων+Ωρ~xJρ~Ων+γnγnn.\displaystyle\frac{\gamma_{n}}{n}J=\nu_{n}^{+}([0,x_{J}))=\int_{[0,x_{J})}(\nu+\tilde{\rho}_{*})=\int_{\Omega}\nu+\int_{\Omega}\tilde{\rho}_{*}-\int_{x_{J}}^{\infty}\tilde{\rho}_{*}\geq\int_{\Omega}\nu+\gamma_{n}-\frac{\gamma_{n}}{n}.

Hence,

nJ1+nγn|Ων|.n-J\leq 1+\frac{n}{\gamma_{n}}\bigg{|}\int_{\Omega}\nu\bigg{|}. (72)

Finally, instead of (57), we now have

xisuppρ~ifor all i=0,,J.x_{i}\in\operatorname{supp}\tilde{\rho}_{i}\qquad\text{for all }i=0,\ldots,J. (73)

Similarly to the previous case, we pass to the limit nn\to\infty in all five terms in (24) separately. By Lemmas 5.6 and 5.7 we obtain the same limit for the second, third and fifth term. Using (73), the fourth term equals

(suppρ~)c(U(x/γn)CU)𝑑νn+(x)γnn(CU+i=J+1n[U(xi/γn)CU]+).\displaystyle\int_{(\operatorname{supp}\tilde{\rho}_{*})^{c}}\big{(}U(x/\gamma_{n})-C_{U}\big{)}\,d\nu_{n}^{+}(x)\leq\frac{\gamma_{n}}{n}\bigg{(}C_{U}+\sum_{i=J+1}^{n}\big{[}U(x_{i}/\gamma_{n})-C_{U}\big{]}^{+}\bigg{)}.

To bound the sum in the right-hand side, note from xJsuppρ~x_{J}\in\operatorname{supp}\tilde{\rho}_{*} that

i=J+1n[U(xi/γn)CU]+i=J+1n(U(xi/γn)U(xJ/γn)).\sum_{i=J+1}^{n}\big{[}U(x_{i}/\gamma_{n})-C_{U}\big{]}^{+}\leq\sum_{i=J+1}^{n}\big{(}U(x_{i}/\gamma_{n})-U(x_{J}/\gamma_{n})\big{)}.

Since by (72)

xiγnxJγn=(iJ)γnn1γnn0,\frac{x_{i}}{\gamma_{n}}-\frac{x_{J}}{\gamma_{n}}=(i-J)\frac{\sqrt{\gamma_{n}}}{n}\leq\frac{1}{\sqrt{\gamma_{n}}}\xrightarrow{n\to\infty}0,

it follows from UC1()U\in C^{1}(\mathbb{R}) that the summand vanishes as nn\to\infty. Hence, the fourth term in (24) also vanishes as nn\to\infty.

We treat the first term in (24) similarly as in (62) and (63). The modifications to (63) are as follows. First, we replace xnx_{n} by xJx_{J} in all integration domains. Then, we replace in (63e) and (63f) νn\nu_{n} by ρ~-\tilde{\rho}_{*} in the integrals over (xJ,)(x_{J},\infty). Finally, we add two new terms (see (74)) to account for xJ+1,,xnx_{J+1},\ldots,x_{n}. With these modifications, (63a)–(63f) can be treated analogously. The two new terms that need to be added are

2(γnn)2i=J+1nj=J+1i1V(xixj)+2γnni=J+1n[0,xJ]V(xiy)𝑑νn(y).\displaystyle 2\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}\sum_{i=J+1}^{n}\sum_{j=J+1}^{i-1}V(x_{i}-x_{j})+2\frac{\gamma_{n}}{n}\sum_{i=J+1}^{n}\int_{[0,x_{J}]}V(x_{i}-y)\,d\nu_{n}(y). (74)

Using (71) and (72) we estimate the first term by

(γnn)2i=J+1nj=J+1i1V(xixj)(γnn)2i=J+1nk=1nJV(kγn3/2/n)Cγnnk=1nJV(kγn3/2/n).\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}\sum_{i=J+1}^{n}\sum_{j=J+1}^{i-1}V(x_{i}-x_{j})\leq\Big{(}\frac{\gamma_{n}}{n}\Big{)}^{2}\sum_{i=J+1}^{n}\sum_{k=1}^{n-J}V(k\gamma_{n}^{3/2}/n)\leq C\frac{\gamma_{n}}{n}\sum_{k=1}^{n-J}V(k\gamma_{n}^{3/2}/n).

For the second term in (74), we apply a similar argument as for (63c). This yields

γnni=J+1n[0,xJ]V(xiy)d(νnν)(y)γnni=J+1nV(xixJ)=γnnk=1nJV(kγn3/2/n)\frac{\gamma_{n}}{n}\sum_{i=J+1}^{n}\int_{[0,x_{J}]}V(x_{i}-y)\,d(\nu_{n}-\nu)(y)\leq\frac{\gamma_{n}}{n}\sum_{i=J+1}^{n}V(x_{i}-x_{J})=\frac{\gamma_{n}}{n}\sum_{k=1}^{n-J}V(k\gamma_{n}^{3/2}/n)

and

γnn|i=J+1n[0,xJ]V(xiy)ν(y)𝑑y|CV(xJ+1xI)Ω|ν|CV(cγn)n0.\frac{\gamma_{n}}{n}\bigg{|}\sum_{i=J+1}^{n}\int_{[0,x_{J}]}V(x_{i}-y)\nu(y)\,dy\bigg{|}\leq CV(x_{J+1}-x_{I})\int_{\Omega}|\nu|\leq CV(c\gamma_{n})\xrightarrow{n\to\infty}0.

Hence, (74) is bounded from above by

Cγnnk=1nJV(kγn3/2/n)Cγnk=1γn3/2nV(kγn3/2n)Cγn0Vn0.C\frac{\gamma_{n}}{n}\sum_{k=1}^{n-J}V(k\gamma_{n}^{3/2}/n)\leq\frac{C}{\sqrt{\gamma_{n}}}\sum_{k=1}^{\infty}\frac{\gamma_{n}^{3/2}}{n}V\Big{(}k\frac{\gamma_{n}^{3/2}}{n}\Big{)}\leq\frac{C}{\sqrt{\gamma_{n}}}\int_{0}^{\infty}V\xrightarrow{n\to\infty}0.

This completes the proof of the limsup inequality in Theorem 1.1.

6.4 The case a=0a=0

Here we prove Theorem 1.1 for the case a=0a=0. The proof is the same as for the case 0<a<10<a<1, except for minor computational modifications. All these modifications are ramifications from the difference in the bound on VV in Assumption 3.2(ii), which in the current case a=0a=0 produces logarithms. The ramifications in the preliminary estimates are the statement of Lemma 5.3(v), which changes into

Vβ(0)C|logβ|andWβCβ|logβ|,\displaystyle V^{\beta}(0)\leq C|\log\beta|\quad\text{and}\quad\int_{\mathbb{R}}W^{\beta}\leq C\beta|\log\beta|, (75)

and the statement of Lemma 5.4(iv). By observing that 0βxj|logx|𝑑xCβj+1|logβ|\int_{0}^{\beta}x^{j}|\log x|\,dx\leq C\beta^{j+1}|\log\beta|, it follows from the proof of Lemma 5.4(iv) that the corresponding statement becomes

(TTβ)fX1,2Cfβ|logβ|.\displaystyle\|(T-T^{\beta})f\|_{X_{1,2}}\leq C_{f}\beta|\log\beta|. (76)

In the compactness proof, by taking again βn=γn/n\beta_{n}=\gamma_{n}/n, we observe from (75) that the term εn\varepsilon_{n} becomes

εn=Cγn2n|logγnn|.\varepsilon_{n}=C\frac{\gamma_{n}^{2}}{n}\Big{|}\log\frac{\gamma_{n}}{n}\Big{|}.

Applying the asymptotic bound on γn\gamma_{n} in Theorem 1.1, we obtain

εnC1logn(logn+loglogn)C,\varepsilon_{n}\ll C\frac{1}{\log n}(\log\sqrt{n}+\log\sqrt{\log n})\leq C,

which is sufficient for continuing the argument in the proofs for the compactness and the liminf inequality.

For the liminf inequality, we only need that (76) vanishes as β0\beta\to 0, which is obvious. For the limsup inequality, the bounds in (75) yield that (63b) is bounded by Cβ|logβ|C\beta|\log\beta|. Hence, we may replace the term Cβ1aC\beta^{1-a} in (62) by Cβ|logβ|C\beta|\log\beta|. Also, the bound in (67) changes by (75) into

xi1xiV(xy)ρ~(y)𝑑yC0γn/(2nρ)|logz|𝑑z=Cγnn|logγnn|.\int_{x_{i-1}}^{x_{i}}V(x-y)\tilde{\rho}_{*}(y)dy\leq C\int_{0}^{\gamma_{n}/(2n\|\rho_{*}\|_{\infty})}|\log z|\,dz=C^{\prime}\frac{\gamma_{n}}{n}\Big{|}\log\frac{\gamma_{n}}{n}\Big{|}.

From this estimate, we obtain for the term T3T_{3} in (65) that

T3Cγn2n|logγnn|=Cεnn0,T_{3}\leq C\frac{\gamma_{n}^{2}}{n}\Big{|}\log\frac{\gamma_{n}}{n}\Big{|}=C^{\prime}\varepsilon_{n}\xrightarrow{n\to\infty}0,

which is sufficient for the proof of the limsup inequality.

Acknowledgements

The author gratefully acknowledges support from JSPS KAKENHI Grant Number JP20K14358.

References

  • [AFP00] L. Ambrosio, N. Fusco, and D. Pallara. Functions of Bounded Variation and Free Discontinuity Problems. Clarendon Press, Oxford, 2000.
  • [BO20] J. Braun and C. Ortner. Sharp uniform convergence rate of the supercell approximation of a crystalline defect. SIAM Journal on Numerical Analysis, 58, 2020.
  • [EOS16] V. Ehrlacher, C. Ortner, and A. V. Shapeev. Analysis of boundary conditions for crystal defect atomistic simulations. Archive for Rational Mechanics and Analysis, 222(3):1217–1268, 2016.
  • [GPPS13] M. G. D. Geers, R. H. J. Peerlings, M. A. Peletier, and L. Scardia. Asymptotic behaviour of a pile-up of infinite walls of edge dislocations. Archive for Rational Mechanics and Analysis, 209:495–539, 2013.
  • [GvMPS16] A. Garroni, P. van Meurs, M. A. Peletier, and L. Scardia. Boundary-layer analysis of a pile-up of walls of edge dislocations at a lock. Mathematical Models and Methods in Applied Sciences, 26(14):2735–2768, 2016.
  • [Hal11] C. L. Hall. Asymptotic analysis of a pile-up of regular edge dislocation walls. Materials Science and Engineering: A, 530:144–148, 2011.
  • [HCO10] C. L. Hall, S. J. Chapman, and J. R. Ockendon. Asymptotic analysis of a system of algebraic equations arising in dislocation theory. SIAM Journal on Applied Mathematics, 70(7):2729–2749, 2010.
  • [HHvM18] C. L. Hall, T. Hudson, and P. van Meurs. Asymptotic analysis of boundary layers in a repulsive particle system. Acta Applicandae Mathematicae, 153(1):1–54, 2018.
  • [Hud13] T. Hudson. Gamma-expansion for a 1D confined Lennard-Jones model with point defect. Networks & Heterogeneous Media, 8(2):501–527, 2013.
  • [HvMP20] T. Hudson, P. van Meurs, and M. A. Peletier. Atomistic origins of continuum dislocation dynamics. ArXiv: 2001.06120, 2020.
  • [KS80] D. Kinderlehrer and G. Stampacchia. An introduction to variational inequalities and their applications. Academic Press. NY, London, 1980.
  • [KvM20] M. Kimura and P. van Meurs. Regularity of the minimiser of one-dimensional interaction energies. ESAIM: Control, Optimisation and Calculus of Variations, 26:27, 2020.
  • [KvM21] M. Kimura and P. van Meurs. Quantitative estimate of the continuum approximations of interacting particle systems in one dimension. SIAM Journal on Mathematical Analysis, 53(1):681–709, 2021.
  • [PS17] M. Petrache and S. Serfaty. Next order asymptotics and renormalized energy for Riesz interactions. Journal of the Institute of Mathematics of Jussieu, 16(3):501–569, 2017.
  • [PZ20] L. Pronzato and A. Zhigljavsky. Bayesian quadrature, energy minimization, and space-filling design. SIAM/ASA Journal on Uncertainty Quantification, 8(3):959–1011, 2020.
  • [SS15] E. Sandier and S. Serfaty. 1D log gases and the renormalized energy: crystallization at vanishing temperature. Probability Theory and Related Fields, 162(3-4):795–846, 2015.
  • [TS19] K. Tanaka and M. Sugihara. Design of accurate formulas for approximating functions in weighted Hardy spaces by discrete energy minimization. IMA Journal of Numerical Analysis, 39(4):1957–1984, 2019.
  • [vM18] P. van Meurs. Convergence rates for discrete-to-continuum limits in 1D particle systems. In Mathematical Analysis of Continuum Mechanics and Industrial Applications II: Proceedings of the International Conference CoMFoS16, pages 181–193. Springer, 2018.