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

Effect of periodic arrays of defects on lattice energy minimizers

Laurent Bétermin

Faculty of Mathematics, University of Vienna,
Oskar-Morgenstern-Platz 1, 1090 Vienna, Austria
[email protected]. ORCID id: 0000-0003-4070-3344
(date)
Abstract

We consider interaction energies Ef[L]E_{f}[L] between a point OdO\in\mathbb{R}^{d}, d2d\geq 2, and a lattice LL containing OO, where the interaction potential ff is assumed to be radially symmetric and decaying sufficiently fast at infinity. We investigate the conservation of optimality results for EfE_{f} when integer sublattices kLkL are removed (periodic arrays of vacancies) or substituted (periodic arrays of substitutional defects). We consider separately the non-shifted (OkLO\in kL) and shifted (OkLO\not\in kL) cases and we derive several general conditions ensuring the (non-)optimality of a universal optimizer among lattices for the new energy including defects. Furthermore, in the case of inverse power laws and Lennard-Jones type potentials, we give necessary and sufficient conditions on non-shifted periodic vacancies or substitutional defects for the conservation of minimality results at fixed density. Different examples of applications are presented, including optimality results for the Kagome lattice and energy comparisons of certain ionic-like structures.

AMS Classification: Primary 74G65 ; Secondary 82B20.
Keywords: Lattice energy; Universal optimality; Defects; Theta functions; Epstein zeta functions; Ionic crystals; Kagome lattice.

1 Introduction, setting and goal of the paper

1.1 Lattice energy minimization and setting

Mathematical results for identifying the lattice ground states of interacting systems have recently attracted a lot of attention. Even though the ‘Crystallization Conjecture’ [16] – the proof of existence and uniqueness of a periodic minimizer for systems with free particles – is still open in full generality, many interesting results have been derived in various settings for showing the global minimality of certain periodic structures including the uniform chain \mathbb{Z}, the triangular lattice 𝖠2\mathsf{A}_{2}, the square lattice 2\mathbb{Z}^{2}, the face-centred cubic lattice 𝖣3\mathsf{D}_{3} (see Fig. 1), as well as the other best packings 𝖤8\mathsf{E}_{8} and the Leech lattice Λ24\Lambda_{24} (see [12, 23] and references therein). Moreover, the same kind of investigation has been made for multi-component systems (e.g. in [29, 30, 35, 10, 36]) where the presence of charged particles yield to rich energetically optimal structures. These problems of optimal point configurations are known to be at the interface of Mathematical Physics, Chemistry, Cryptography, Geometry, Signal processing, Approximation, Arithmetic, etc. The point of view adopted in this work is the one of Material Science where the points are thought as particles or atoms.

Refer to caption
Refer to caption
Refer to caption
Figure 1: In dimension d=2d=2, representation of the triangular and square lattices respectively defined by 𝖠2=λ1[(1,0)(1/2,3/2)]\mathsf{A}_{2}=\lambda_{1}\left[\mathbb{Z}(1,0)\oplus\mathbb{Z}(1/2,\sqrt{3}/2)\right] and 2\mathbb{Z}^{2}. In dimension d=3d=3, representation of the simple cubic and the face-centred cubic lattices respectively defined by 3\mathbb{Z}^{3} and 𝖣3:=λ2[(1,0,1)(0,1,1)(1,1,0)]\mathsf{D}_{3}:=\lambda_{2}\left[\mathbb{Z}(1,0,1)\oplus\mathbb{Z}(0,1,1)\oplus\mathbb{Z}(1,1,0)\right]. The constants λ1,λ2\lambda_{1},\lambda_{2} are such that the lattices have unit density.

In this paper, our general goal is to show mathematically how the presence of periodic arrays of charges (called here ‘defects’ in contrast with the initial crystal ‘atoms’) in a perfect crystal affects the minimizers of interaction energies when the interaction between species is radially symmetric. Since the structure of crystals are often given by the same kind of lattices, it is an important question to know the conditions on the added periodic distribution of defects and on the interaction energy in order to have conservation of the initial ground state structure. Furthermore, only very few rigorous results are available on minimization of charged structures among lattices (see e.g. our recent works [10, 9]).

We therefore assume the periodicity of our systems, and once we restrict this kind of problem to the class of (simple) lattices and radially symmetric interaction potentials, an interesting non-trivial problem is to find the minimizers of a given energy per point among these simple periodic sets of points, with or without a fixed density. In this paper, we keep the same kind of notations we have used in our previous works (see e.g. [14, 8, 10]). More precisely, for any d2d\geq 2 we called d\mathcal{L}_{d} the class of dd-dimensional lattices, i.e. discrete co-compact subgroups or d\mathbb{R}^{d},

d:={L=i=1dui:{u1,,ud} is a basis of d},\mathcal{L}_{d}:=\left\{L=\bigoplus_{i=1}^{d}\mathbb{Z}u_{i}:\textnormal{$\{u_{1},...,u_{d}\}$ is a basis of $\mathbb{R}^{d}$}\right\},

and, for any V>0V>0, d(V)d\mathcal{L}_{d}(V)\subset\mathcal{L}_{d} denotes the set of lattices with volume |det(u1,,ud)|=V|\det(u_{1},...,u_{d})|=V, i.e. such that its unit cell QLQ_{L} defined by

QL:={x=i=1dλiui:i{1,,d},λi[0,1)},Q_{L}:=\left\{x=\sum_{i=1}^{d}\lambda_{i}u_{i}:\forall i\in\{1,...,d\},\lambda_{i}\in[0,1)\right\}, (1.1)

has volume |QL|=V|Q_{L}|=V. We will also say that Ld(V)L\in\mathcal{L}_{d}(V) has density V1V^{-1}. The class d\mathcal{F}_{d} of radially symmetric functions we consider is, calling d\mathcal{M}_{d} the space of Radon measures on +\mathbb{R}_{+},

d:={f:+:f(r)=0ert𝑑μf(t),μfd,|f(r)|=O(rpf) as rpf>d/2}.\mathcal{F}_{d}:=\left\{f:\mathbb{R}_{+}\to\mathbb{R}:f(r)=\int_{0}^{\infty}e^{-rt}d\mu_{f}(t),\mu_{f}\in\mathcal{M}_{d},|f(r)|=O(r^{-p_{f}})\textnormal{ as $r\to\infty$, $p_{f}>d/2$}\right\}.

When μf\mu_{f} is non-negative, ff is a completely monotone function, which is equivalent by Hausdorff-Bernstein-Widder Theorem [3] with the property that for all r>0r>0 and all kk\in\mathbb{N}, (1)kf(k)(r)0(-1)^{k}f^{(k)}(r)\geq 0. We will write this class of completely monotone functions as

dcm:={fd:μf0}.\mathcal{F}^{cm}_{d}:=\left\{f\in\mathcal{F}_{d}:\mu_{f}\geq 0\right\}.

For any fdf\in\mathcal{F}_{d}, we thus defined the ff-energy Ef[L]E_{f}[L] of a lattice LL, which is actually the interaction energy between the origin OO of d\mathbb{R}^{d} and all the other points of LL, by

Ef[L]:=pL\{0}f(|p|2).E_{f}[L]:=\sum_{p\in L\backslash\{0\}}f(|p|^{2}). (1.2)

Notice that this sum is absolutely convergent as a simple consequence of the definition of d\mathcal{F}_{d}. We could also define EfE_{f} without such decay assumption by renormalizing the sum using, for instance, a uniform background of opposite charges (see e.g. [34]) or an analytic continuation in case of parametrized potential such as rsr^{-s} (see [17]).

One can interpret the problem of minimizing EfE_{f} in d\mathcal{L}_{d} (or in d(V)\mathcal{L}_{d}(V) for fixed V>0V>0) as a geometry optimization problem for solid crystals where the potential energy landscape of a system with an infinite number of particles is studied in the restricted class of lattice structures. Even though the interactions in a solid crystal are very complex (quantum effects, angle-dependent energies, etc.), it is known that the Born-Oppenheimer adiabatic approximation used to describe the interaction between atoms or ions in a solid by a sum of pairwise contributions (see e.g. [40, p. 33 and p. 945] and [46]) is a good model for ‘classical crystals’ (compared to ‘quantum crystals’ [18]), i.e. where the atoms are sufficiently heavy. Moreover, since all the optimality properties we are deriving in this paper are invariant under rotations, all the results will be tacitly considered up to rotations.

Furthermore, studying such interacting systems with this periodicity constraint is a good method to keep or exclude possible candidates for a crystallization problem (i.e. with free particles). We are in particular interested in a type of lattice LdL_{d} that is the unique minimizer of EfE_{f} in d(V)\mathcal{L}_{d}(V) for any fixed V>0V>0 and any completely monotone potential fdcmf\in\mathcal{F}_{d}^{cm}. Following Cohn and Kumar [21] notion (originally defined among all periodic configurations), we call this property the universal optimality among lattices of LdL_{d} (or universal optimality in d(1)\mathcal{L}_{d}(1)).

Only few methods are available to carry out the minimization of EfE_{f}. Historically, the first one consists to parametrize all the lattices of d(1)\mathcal{L}_{d}(1) in an Euclidean fundamental domain 𝒟dd(d+1)21\mathcal{D}_{d}\subset\mathbb{R}^{\frac{d(d+1)}{2}-1} (see e.g. [44, Sec. 1.4]) and to study the variations of the energy in 𝒟d\mathcal{D}_{d}. It has been done in dimension 2 for showing the optimality of the triangular lattice 𝖠2\mathsf{A}_{2} at fixed density for the Epstein zeta function [42, 28, 19, 27] and the lattice theta function [38] respectively defined for s>ds>d and α>0\alpha>0 by

ζL(s):=pL\{0}1|p|s,andθL(α):=pLeπα|p|2.\zeta_{L}(s):=\sum_{p\in L\backslash\{0\}}\frac{1}{|p|^{s}},\qquad\textnormal{and}\qquad\theta_{L}(\alpha):=\sum_{p\in L}e^{-\pi\alpha|p|^{2}}. (1.3)

In particular, a simple consequence of Montgomery’s result [38] for the lattice theta function is the universal optimality among lattices of 𝖠2\mathsf{A}_{2} (see e.g. [4, Prop. 3.1]). Other consequences of the universal optimality of 𝖠2\mathsf{A}_{2} among lattices have been derived for other potentials (including the Lennard-Jones one) [15, 4, 14, 7] as well as masses interactions [11]. Furthermore, new interesting and general consequences of universal optimality will be derived in this paper, including a sufficient condition for the minimality of a universal minimizer at fixed density (see Theorem 2.9).

This variational method is also the one we have recently chosen in [9] for showing the maximality of 𝖠2\mathsf{A}_{2} in 2(1)\mathcal{L}_{2}(1) – and conjectured the same results in dimensions d{8,24}d\in\{8,24\} for the lattices 𝖤8\mathsf{E}_{8} and Λ24\Lambda_{24} – for the alternating and centered lattice theta function respectively defined, for all α>0\alpha>0, by

θL±(α):=pLφ±(p)eπα|p|2,andθLc(α):=pLeπα|p+cL|2,\theta_{L}^{\pm}(\alpha):=\sum_{p\in L}\varphi_{\pm}(p)e^{-\pi\alpha|p|^{2}},\quad\textnormal{and}\quad\theta_{L}^{c}(\alpha):=\sum_{p\in L}e^{-\pi\alpha|p+c_{L}|^{2}}, (1.4)

where L=i=1duiL=\bigoplus_{i=1}^{d}\mathbb{Z}u_{i}, {ui}i\{u_{i}\}_{i} being a Minkowski (reduced) basis of LL (see e.g. [44, Sect. 1.4.2]), φ±(p):=i=1dmi\varphi_{\pm}(p):=\sum_{i=1}^{d}m_{i} for p=i=1dmiuip=\sum_{i=1}^{d}m_{i}u_{i}, mim_{i}\in\mathbb{Z} for all ii, and cL=12iuic_{L}=\frac{1}{2}\sum_{i}u_{i} is the center of its unit cell QLQ_{L}. In particular, the alternate lattice theta function θL±(α)\theta_{L}^{\pm}(\alpha) can be viewed as the Gaussian interaction energy of a lattice LL with an alternating distribution of charges ±1\pm 1, which can be itself seen as the energy once we have removed 22 times the union of sublattices i=1d(L+ui)\cup_{i=1}^{d}(L+u_{i}) from the original lattice LL. This result shows another example of universal optimality – we will call it universal maximality – among lattices, i.e. the maximality of 𝖠2\mathsf{A}_{2} in 2(1)\mathcal{L}_{2}(1) for the energies Ef±E_{f}^{\pm} and EfcE_{f}^{c} defined by

Ef±[L]:=pL\{0}φ±(p)f(|p|2),orEfc[L]:=pLf(|p+cL|2),E_{f}^{\pm}[L]:=\sum_{p\in L\backslash\{0\}}\varphi_{\pm}(p)f(|p|^{2}),\quad\textnormal{or}\quad E_{f}^{c}[L]:=\sum_{p\in L}f(|p+c_{L}|^{2}), (1.5)

where fdcmf\in\mathcal{F}_{d}^{cm}. This kind of problem was actually our first motivation for investigating the effects of periodic arrays of defects on lattice energy minimizers, since removing two times the sublattices 2L+u12L+u_{1} and 2L+u22L+u_{2} totally inverse the type of optimality among lattices. Furthermore, this maximality result will also be used in Theorem 2.4, applied – in the general case of a universal maximizer Ld±L_{d}^{\pm} for Ef±E_{f}^{\pm} in any dimension where this property could be shown – for other potentials d\dcm\mathcal{F}_{d}\backslash\mathcal{F}_{d}^{cm} in Theorem 2.11 and compared with other optimality results in Section 3.2.

The second method for showing such optimality result is based on the Cohn-Elkies linear programming bound that was successfully used for showing the best packing results in dimensions 8 and 24 for 𝖤8\mathsf{E}_{8} and Λ24\Lambda_{24} in [47, 22], as well as their universal optimality among all periodic configurations in [23]. As in the two-dimensional case, many consequences of these optimality results have been shown for other potentials [14, 39] and masses interactions [8].

1.2 Problem studied in this paper and connection to Material Science

The goal of this work is to investigate the effect on the minimizers of EfE_{f} when we change, given a lattice LdL\subset\mathcal{L}_{d} and K\{1}K\subset\mathbb{N}\backslash\{1\}, a certain real number ak0a_{k}\neq 0 of integer sublattices kLkL, kKk\in K, in the original lattice, and where the lattices kLkL might be shifted by a finite number of lattice points Lk:={pi,k}iIkLL_{k}:=\{p_{i,k}\}_{i\in I_{k}}\subset L for some finite set IkI_{k}. Writing

κ:={K,AK,PK},K\{1},AK={ak}kK,PK=kKLk,Lk={pi,k}iIkL,\kappa:=\{K,A_{K},P_{K}\},\quad K\subset\mathbb{N}\backslash\{1\},\quad A_{K}=\{a_{k}\}_{k\in K}\subset\mathbb{R}^{*},\quad P_{K}=\bigcup_{k\in K}L_{k},\quad L_{k}=\{p_{i,k}\}_{i\in I_{k}}\subset L, (1.6)

the new energy EfκE^{\kappa}_{f} we consider, defined for fdf\in\mathcal{F}_{d} and κ\kappa as in (1.6) and such that the following sum on KK is absolutely convergent, is given by

Efκ[L]:=Ef[L]kKiIkakEf[pi,k+kL].E^{\kappa}_{f}[L]:=E_{f}[L]-\sum_{k\in K}\sum_{i\in I_{k}}a_{k}E_{f}[p_{i,k}+kL]. (1.7)

In particular, in the non-shifted case, i.e. PK=P_{K}=\emptyset, then

Efκ[L]=Efκ[L],wherefκ(r):=f(r)kKakf(k2r).E^{\kappa}_{f}[L]=E_{f_{\kappa}}[L],\quad\textnormal{where}\quad f_{\kappa}(r):=f(r)-\sum_{k\in K}a_{k}f(k^{2}r). (1.8)

Since we are interested in the effects of defects on lattice energy ground states, we therefore want to derive conditions on κ\kappa and ff such that EfE_{f} and EfκE_{f}^{\kappa} have the same minimizers in d\mathcal{L}_{d} or d(V)\mathcal{L}_{d}(V) for fixed V>0V>0. In particular, we also want to know if the universal minimality among lattices of a lattice LdL_{d} is conserved while removing or substituting integer sublattices. This a natural step for investigating the robustness of the optimality results stated in the previous section of this paper when the interaction potential is completely monotone or, for instance, of Lennard-Jones type. Furthermore, it is also the opportunity to derive new applications and generalizations of the methods recently developed in [4, 14, 9] for more ‘exotic’ ionic-like structures.

Replacing integer sublattices as described above can be interpreted and classified in two relevant cases in Material Science:

  1. 1.

    If ak=1a_{k}=1, then removing only once the sublattice kLkL from LL creates a periodic array of vacancies (also called periodic Schottky defects [45, Sect. 3.4.3]);

  2. 2.

    If ak1a_{k}\neq 1, then ‘removing’ aka_{k} times the sublattice kLkL from LL creates a periodic array of substitutional defects (also called impurities), where the original lattice points (initially with charges +1+1) are replaced by points with ‘charges’ (or ‘weights’) 1ak01-a_{k}\neq 0.

In Figure 2, we have constructed three examples of two-dimensional lattices with periodic arrays of defects which certainly do not exist in the real world. In contrast, Figure 3 shows two important examples of crystal structures arising in nature: the Kagome lattice and the rock-salt structure. These examples are discussed further in Section 3.

Refer to caption
Refer to caption
Refer to caption
Figure 2: Mathematical examples of periodic array of defects performed on a patch of the square lattice 2\mathbb{Z}^{2} (left and right) and the triangular lattice 𝖠2\mathsf{A}_{2} (middle). The cross ×{\color[rgb]{0,0,1}\times} represents the origin OO of 2\mathbb{R}^{2}. The points marked by {\color[rgb]{0,0,1}\bullet} are the original points of the lattice whereas the points marked by +{\color[rgb]{1,.5,0}+} and {\color[rgb]{1,0,0}\circ} are substitutional defects of charge 1ak1-a_{k} for some ak\{1}a_{k}\in\mathbb{R}^{*}\backslash\{1\} and some kK:={2,3,4,5}k\in K_{:}=\{2,3,4,5\}. The missing lattice points are the vacancy defects. The patch on the right contains two shifted periodic arrays of defects.
Refer to caption
Refer to caption
Figure 3: Two examples of 2d lattices patches with a periodic array of defect arising in nature. The left-hand structure is the Kagome lattice obtained by removing from the triangular lattice 𝖠2\mathsf{A}_{2} the sublattice 2𝖠2+(1,0)+(1/2,3/2)2\mathsf{A}_{2}+(1,0)+(1/2,\sqrt{3}/2). It appears to be a layer of the jarosite. The right-hand structure is the 2d rock-salt structure obtained by removing from the square lattice 2\mathbb{Z}^{2} two times the sublattices 22+(1,0)2\mathbb{Z}^{2}+(1,0) and 22+(0,1)2\mathbb{Z}^{2}+(0,1) in such a way that particles of opposites signs ±1\pm 1 alternate ({\color[rgb]{0,0,1}\bullet} and {\color[rgb]{1,0,0}\circ} correspond respectively to charges of signs 11 and 1-1). It is itself a layer of the three-dimensional rock-salt structure NaCl.

While the substitutional defects case has different interpretations and applications in terms of optimal multi-component (ionic) crystals (see e.g. Section 3.2), the vacancy case is also of interest when we look for accelerating the computational time for checking numerically the minimality of a structure. Indeed, if the minimizer does not change once several periodic arrays of points are removed from all lattices, then a computer will be faster to check this minimality. This is of practical relevance in particular in low dimensions since the computational time of such lattice energies, which grows exponentially with the dimension, are extremely long in dimension d8d\geq 8 – even with the presence of periodic arrays of vacancies – and shows how important are rigorous minimality results in these cases.

Furthermore, from a Physics point of view, it is well-known (see e.g. [45]) that point defects play an important role in crystal properties. As explained in [1]: ‘Crystals are like people, it is the defects in them which tend to make them interesting’. For instance, they reduce the electric and thermal conductivity in metals and modify the colors of solids and their mechanical strength. We also notice that substitutional defects control the electronic conductivity in semi-conductors, whereas the vacancies control the diffusion and the ionic conductivity in a solid. In particular, there is no perfect crystal in nature and it is then interesting and physically relevant to study optimality results for periodic systems with defects, in particular for models at positive temperature where the number of vacancies per unit volume increases exponentially with the temperature (see e.g. [45, Sec. 3.4.3]). Notice that the raise of temperature also creates another kind of defects called self-interstitial – i.e. the presence of extra atoms out of lattice sites – but they are known to be negligible (at least if they are of the same type than the solid’s atoms) compared to the vacancies when disorder appears, excepted for Silicon.

Plan of the paper. Our main results are presented in Section 2 whereas their proofs are postponed to Section 4. Many applications of our results are discussed in Section 3, including explicit examples of minimality results for the Kagome lattice and other ionic structures.

2 Statement of the main results

2.1 On the minimality of a universal optimizer

We start by recalling the notion of universal optimality among lattices as defined by Cohn and Kumar in [21].

Definition 2.1 (Universal optimality among lattices).

Let d2d\geq 2. We say that LdL_{d} is universally optimal in d(1)\mathcal{L}_{d}(1) if LdL_{d} is a minimizer of EfE_{f} defined by (1.2) in d(1)\mathcal{L}_{d}(1) for any fdcmf\in\mathcal{F}_{d}^{cm}.

Remark 2.1 (Universally optimal lattices).

We recall again that the only known universally optimal lattices in dimension d2d\geq 2 are 𝖠2\mathsf{A}_{2} (see [38]), 𝖤8\mathsf{E}_{8} and the Leech lattice Λ24\Lambda_{24} (see [23]) in dimensions d{2,8,24}d\in\{2,8,24\}. It is also shown in [43, p. 117] that there is no such universally optimal lattice in dimension d=3d=3. There are also clear indications (see [14, Sect. 6.1]) that the space of functions for which the minimality at all the scales of LdL_{d} holds is much larger than dcm\mathcal{F}_{d}^{cm}.

Before stating our results, notice that all of them are stated in terms of global optimality but could be rephrased for showing local optimality properties. This is important, in particular in dimensions d=3d=3 where only local minimality results are available for EfE_{f} (see e.g. [6]) and can be generalized for energies of type EfκE_{f}^{\kappa}, ensuring the local stability of certain crystal structures.

We now show that the universal optimalities among lattices in dimension d{2,8,24}d\in\{2,8,24\} proved in [38, 23] are not conserved in the non-shifted case once we only removed a single integer sublattice a positive number ak>0a_{k}>0 of times, whereas they are conserved when ak<0a_{k}<0.

Theorem 2.2 (Conservation of universal optimalities - Non-shifted case).

Let ff be defined by f(r)=eπαrf(r)=e^{-\pi\alpha r}, α>0\alpha>0. For all d{2,8,24}d\in\{2,8,24\}, all k\{1}k\in\mathbb{N}\backslash\{1\}, all ak>0a_{k}>0 and κ={k,ak,}\kappa=\{k,a_{k},\emptyset\}, there exists αd\alpha_{d} such that for all α(0,αd)\alpha\in(0,\alpha_{d}), 𝖠2\mathsf{A}_{2}, 𝖤8\mathsf{E}_{8} and the Leech lattice Λ24\Lambda_{24} are not minimizers of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1).
Furthermore, for any d{2,8,24}d\in\{2,8,24\}, for any K\{1}K\subset\mathbb{N}\backslash\{1\}, any AK={ak}kKA_{K}=\{a_{k}\}_{k\in K}\subset\mathbb{R}_{-} and κ={K,AK,}\kappa=\{K,A_{K},\emptyset\}, 𝖠2\mathsf{A}_{2}, 𝖤8\mathsf{E}_{8} and the Leech lattice Λ24\Lambda_{24} are the unique minimizers of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1) for all α>0\alpha>0.

Remark 2.3 (Generalization to 44-designs).

The non-optimality result in Theorem 2.2 is obtained by using the Taylor expansion of the theta function found by Coulangeon and Schürmann in [25, Eq. (21)]. Therefore, the result is actually generalizable to any universal optimal lattice LdL_{d} such that all its layers (or shells) are 44-designs, i.e. such that for all r>0r>0 with {BrLd}\{\partial B_{r}\cap L_{d}\}\neq\emptyset, BrB_{r} being the ball centred at the origin and with radius rr, and all polynomial PP of degree up to 44 we have

1|Br|BrP(x)𝑑x=1{BrLd}xBrLdP(x).\frac{1}{|\partial B_{r}|}\int_{\partial B_{r}}P(x)dx=\frac{1}{\sharp\{\partial B_{r}\cap L_{d}\}}\sum_{x\in\partial B_{r}\cap L_{d}}P(x).

We now present a sufficient condition on PKP_{K} such that the triangular lattice is universally optimal in 2(1)\mathcal{L}_{2}(1) for EfκE_{f}^{\kappa}. This result is based on our recent work [9] where we have proven the maximality of 𝖠2\mathsf{A}_{2} in 2(1)\mathcal{L}_{2}(1) for the centred lattice theta functions, i.e. LθL+cL(α)L\mapsto\theta_{L+c_{L}}(\alpha), where cLc_{L} is the center of the unit cell QLQ_{L} (see also Remark 2.12).

Theorem 2.4 (Conservation of universal optimality - 2d shifted case).

Let d=2d=2 and κ={K,AK,PK}\kappa=\{K,A_{K},P_{K}\} be as in (1.6) where AK+A_{K}\subset\mathbb{R}_{+}, and be such that

kK,iIk,pi,kk=cL modulo QL,L=u1u2,cL:=u1+u22,\forall k\in K,\forall i\in I_{k},\quad\frac{p_{i,k}}{k}=c_{L}\textnormal{ modulo $Q_{L}$,}\quad L=\mathbb{Z}u_{1}\oplus\mathbb{Z}u_{2},\quad c_{L}:=\frac{u_{1}+u_{2}}{2}, (2.1)

where QLQ_{L} is the unit cell of LL defined by (1.1) with a Minkowski basis {u1,u2}\{u_{1},u_{2}\} and its center cLc_{L}. Then, for all f2cmf\in\mathcal{F}_{2}^{cm}, 𝖠2\mathsf{A}_{2} is the unique minimizer of EfκE^{\kappa}_{f} in 2(1)\mathcal{L}_{2}(1).

Example 2.5.

Theorem 2.4 holds in a particularly simple case, when k=2k=2 and pi,2=u1+u2Lp_{i,2}=u_{1}+u_{2}\in L.

Remark 2.6 (Conjecture in dimensions d{8,24}d\in\{8,24\}).

Theorem 2.4 is based on the fact that 𝖠2\mathsf{A}_{2} has been shown to be the unique maximizer of EfcE_{f}^{c} defined in (1.5) in d(1)\mathcal{L}_{d}(1) for any fdcmf\in\mathcal{F}_{d}^{cm} (see also Remark 2.12). As discussed in [9], we believe that this result still holds in dimensions 8 and 24 for 𝖤8\mathsf{E}_{8} and the Leech lattice Λ24\Lambda_{24}, as well as our Theorem 2.4.

Remark 2.7 (Phase transition for the minimizer in the Gaussian case - Numerical observation).

In the non-universally optimal case of Theorem 2.2 and the shifted case satisfying (2.1), numerical investigations suggest that the minimizer of EfκE_{f}^{\kappa} exhibits a phase transition as the density decreases.
Non-shifted case. Let us consider the example f(r)=eπαrf(r)=e^{-\pi\alpha r} given in Theorem 2.2 (i.e. f(r2)f(r^{2}) is a Gaussian function) and fκ(r)=eπαr0.1e2παrf_{\kappa}(r)=e^{-\pi\alpha r}-0.1e^{-2\pi\alpha r} (defined by (1.8)), κ:={2,0.1,}\kappa:=\{2,0.1,\emptyset\}, corresponding to removing a2=0.1a_{2}=0.1 times the sublattice 2L2L (k=2k=2) from the original lattice LL. In dimension d=2d=2, we numerically observe an interesting phase transition of type ‘triangular-rhombic-square-rectangular’ for the minimizer of EfκE_{f}^{\kappa} in 2(1)\mathcal{L}_{2}(1) as α\alpha (which plays the role of the inverse density here) increases.
Shifted case with ak<0a_{k}<0. Let us assume that K={2}K=\{2\}, AK:={a2<0}A_{K}:=\{a_{2}<0\}, I2={1}I_{2}=\{1\} and p1,2=u1+u2p_{1,2}=u_{1}+u_{2} in such a way that (2.1) is satisfied. If we consider f(r)=eπαrf(r)=e^{-\pi\alpha r}, then for all the negative parameters a2a_{2} we have chosen, the minimizer of Efκ[L]:=θL(α)+|a2|θL+cL(α)E_{f}^{\kappa}[L]:=\theta_{L}(\alpha)+|a_{2}|\theta_{L+c_{L}}(\alpha) in 2(1)\mathcal{L}_{2}(1) numerically shows the same phase transition of type ‘triangular-rhombic-square-rectangular’ as α\alpha increases.
This type of phase transition seems to have a certain universality in dimension 2 since it was also observed for Lennard-Jones energy [5], Morse energy [7], Madelung-like energies [10] and proved for 3-blocks copolymers [35] and two-component Bose-Einstein condensates [36] by Wei et al..

Remark 2.8 (Optimality of d\mathbb{Z}^{d} in the orthorhombic case).

Another type of universal optimality is known in the set of orthorhombic lattices, i.e. the lattice LL which can be represented by an orthogonal basis. As proved by Montgomery in [38, Thm. 2], the cubic lattice d\mathbb{Z}^{d} is universally minimal among orthorhombic lattices of unit density in any dimension (see also [10, Rmk. 3.1]). The proof of Theorem 2.2 can be easily adapted to show the same optimality result for d\mathbb{Z}^{d} among orthorhombic lattices of unit density. Furthermore, it has also been shown (see e.g. [13, Prop. 1.4]) that d\mathbb{Z}^{d} is the unique maximum of LEf[L+cL]L\mapsto E_{f}[L+c_{L}] among orthorhombic lattices of fixed density for any fdcmf\in\mathcal{F}_{d}^{cm}. Therefore, the proof of Theorem 2.4 can be also easily adapted in this orthorhombic case in order to show the universal optimality of d\mathbb{Z}^{d} in this particular shifted case. Moreover, all the next results involving any universally optimal lattice can be rephrased for d\mathbb{Z}^{d} in the space of orthorhombic lattices. Examples of applications of such result will be discussed in Section 3.2.

We now give a general criterion that ensures the conservation of an universal optimizer’s minimality for EfκE_{f}^{\kappa}.

Theorem 2.9 (General criterion for minimality conservation - Non-shifted case).

Let d2d\geq 2, κ={K,AK,}\kappa=\{K,A_{K},\emptyset\} be as in (1.6) (possibly empty) where AK+A_{K}\subset\mathbb{R}_{+}, and LdL_{d} be universally optimal in d(1)\mathcal{L}_{d}(1). Furthermore, let fdf\in\mathcal{F}_{d} be such that dμf(t)=ρf(t)dtd\mu_{f}(t)=\rho_{f}(t)dt and fκf_{\kappa} be defined by (1.8). Then:

  1. 1.

    For any κ\kappa, we have fκ(r)=0ert𝑑μfκ(t)f_{\kappa}(r)=\displaystyle\int_{0}^{\infty}e^{-rt}d\mu_{f_{\kappa}}(t) where

    dμfκ(t)=ρfκ(t)dt,ρfκ(t)=ρf(t)kKakk2ρf(tk2).d\mu_{f_{\kappa}}(t)=\rho_{f_{\kappa}}(t)dt,\quad\rho_{f_{\kappa}}(t)=\rho_{f}(t)-\sum_{k\in K}\frac{a_{k}}{k^{2}}\rho_{f}\left(\frac{t}{k^{2}}\right).
  2. 2.

    The following equivalence holds: fκdcmf_{\kappa}\in\mathcal{F}_{d}^{cm} if and only if

    t>0,ρf(t)kKakk2ρf(tk2);\forall t>0,\quad\rho_{f}(t)\geq\sum_{k\in K}\frac{a_{k}}{k^{2}}\rho_{f}\left(\frac{t}{k^{2}}\right); (2.2)
  3. 3.

    If (2.2) holds, then LdL_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1).

  4. 4.

    If there exists V>0V>0 such that for a.e. y1y\geq 1 there holds

    gV(y):=ρfκ(πyV2d)+yd22ρfκ(πV2dy)0,g_{V}(y):=\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)+y^{\frac{d}{2}-2}\rho_{f_{\kappa}}\left(\frac{\pi}{V^{\frac{2}{d}}y}\right)\geq 0, (2.3)

    then V1dLdV^{\frac{1}{d}}L_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(V)\mathcal{L}_{d}(V).

The fourth point on Theorem 2.9 generalizes our two-dimensional result [4, Thm. 1.1] to any dimension and with possible periodic arrays of defects. It is an important result since only few minimality results for EfE_{f} are available for non-completely monotone potentials fd\dcmf\in\mathcal{F}_{d}\backslash\mathcal{F}_{d}^{cm}, and this also the first result of this kind for charged lattices (i.e. when the particles are not of the same kind). Condition (2.3) has been used in dimension d=2d=2 in [4, 7] for proving the optimality of a triangular lattice at fixed density for non-convex sums of inverse power laws, differences of Yukawa potentials, Lennard-Jones potentials and Morse potentials and we expect the same property to hold in higher dimension. In Theorem 2.17, we will give an example of such application in any dimension dd by applying the fourth point of Theorem 2.9 to Lennard-Jones type potentials. We now add a very important remark concerning the adaptation of the fourth point of Theorem 2.9 in the general periodic case, i.e. for crystallographic point packings (see [2, Def. 2.5]).

Remark 2.10 (Crystallization at fixed density as a consequence of Cohn-Kumar Conjecture).

When κ=\kappa=\emptyset, i.e. all the particles are present and of the same kind, the proof of point 4. of Theorem 2.9 admits a straightforward adaptation in the periodic case, i.e among all configurations 𝒞=i=1N(Λ+vk)𝒮\mathcal{C}=\bigcup_{i=1}^{N}\left(\Lambda+v_{k}\right)\in\mathcal{S} being Λ\Lambda-periodic of unit density, where Λd\Lambda\in\mathcal{L}_{d}, i.e. such that |Λ|=N|\Lambda|=N, and with a ff-energy defined for V>0V>0 by

Ef[V1d𝒞]:=1Nj,k=1NxΛ\{vkvj}f(V2d|x+vkvj|2).E_{f}[V^{\frac{1}{d}}\mathcal{C}]:=\frac{1}{N}\sum_{j,k=1}^{N}\sum_{x\in\Lambda\backslash\{v_{k}-v_{j}\}}f\left(V^{\frac{2}{d}}|x+v_{k}-v_{j}|^{2}\right).

Using again the representation of ff as a superposition of Gaussians combined with the Jacobi transformation formula (see the proof of Theorem 2.9), the same condition (2.3) ensures the crystallization on LdL_{d} at fixed density once we know its universal optimality in the set of all periodic configurations with fixed density V1V^{-1}. This result is in the same spirit as the one derived by Petrache and Serfaty in [39] for Coulomb and Riesz interactions. In dimensions d{8,24}d\in\{8,24\}, (2.3) implies the crystallization on 𝖤8\mathsf{E}_{8} and Λ24\Lambda_{24} at fixed density V1V^{-1} as a consequence of [23] whereas in dimension d=2d=2 it is conjectured by Cohn and Kumar in [21] that the same holds on the triangular lattice. It is in particular true for the Lennard-Jones potential at high density as a simple application of our Theorem 2.17.

Using exactly the same arguments as the fourth point of Theorem 2.9, we show the following result which gives a sufficient condition on an interaction potential ff for a universal maximizer Ld±L_{d}^{\pm} of θL±(α)\theta_{L}^{\pm}(\alpha) to be optimal for Ef±E_{f}^{\pm}, where

θL±(α):=pLφ±(p)eπα|p|2,andEf±[L]:=pL\{0}φ±(p)f(|p|2),\theta_{L}^{\pm}(\alpha):=\sum_{p\in L}\varphi_{\pm}(p)e^{-\pi\alpha|p|^{2}},\quad\textnormal{and}\quad E_{f}^{\pm}[L]:=\sum_{p\in L\backslash\{0\}}\varphi_{\pm}(p)f(|p|^{2}), (2.4)

with L=i=1duiL=\bigoplus_{i=1}^{d}\mathbb{Z}u_{i}, {u1,,ud}\{u_{1},...,u_{d}\} being its Minkowski basis, and φ±(p)=i=1dmi\varphi_{\pm}(p)=\sum_{i=1}^{d}m_{i} for p=i=1dmiuip=\sum_{i=1}^{d}m_{i}u_{i}, mim_{i}\in\mathbb{Z} for all ii. Remark that Ef±=EfκE_{f}^{\pm}=E_{f}^{\kappa} when κ={2,{2,.2},{u1,,ud}}\kappa=\{2,\{2,....2\},\{u_{1},...,u_{d}\}\}, L=i=1duiL=\bigoplus_{i=1}^{d}\mathbb{Z}u_{i}. In particular, it holds for the triangular lattice 𝖠2\mathsf{A}_{2} as a simple application of our main result in [9].

Theorem 2.11 (Maximality of a universal maximizer for θL±\theta_{L}^{\pm} - Shifted case).

Let d2d\geq 2, V>0V>0, κ={2,{2,.2},{u1,,ud}}\kappa=\{2,\{2,....2\},\{u_{1},...,u_{d}\}\}, where a generic lattice is written L=i=1duiL=\bigoplus_{i=1}^{d}\mathbb{Z}u_{i}, {u1,,ud}\{u_{1},...,u_{d}\} being its Minkowski basis, and Ld±L_{d}^{\pm} be the unique maximizer of θL±(α)\theta_{L}^{\pm}(\alpha), defined by (2.4), in d(1)\mathcal{L}_{d}(1) and for all α>0\alpha>0. If fdf\in\mathcal{F}_{d} satisfies (2.3), then V1dLd±V^{\frac{1}{d}}L_{d}^{\pm} is the unique maximizer of EfκE_{f}^{\kappa} (equivalently of Ef±E_{f}^{\pm} defined by (2.4)) in d(V)\mathcal{L}_{d}(V).

Remark 2.12 (Adaptation to shifted ff-energy).

We believe that Theorem 2.11 also holds for 𝖤8\mathsf{E}_{8} and Λ24\Lambda_{24} (see [9, Conj. 1.3] and Remark 2.6). Furthermore, the same kind of optimality result could be easily derived for any energy shifted energy of type LEf[L+c]L\mapsto E_{f}[L+c] where cQLc\in Q_{L} is fixed as a function of the vectors in the Minkowski basis {ui}\{u_{i}\} and when one knows a universal minimizer or maximizer for LEf[L+c]L\mapsto E_{f}[L+c], fdcmf\in\mathcal{F}_{d}^{cm}. However, no other result concerning any optimality of a lattice for such kind of energy is currently available when c{L,cL}c\not\in\{L,c_{L}\}.

The rest of our results are all given in the non-shifted case PK=P_{K}=\emptyset. It is indeed a rather difficult task to minimize the sum of shifted and/or non-shifted energies of type EfE_{f}. Very few results are available and the recent work by Luo and Wei [36] has shown the extreme difficulty to obtain any general result for completely monotone function ff. Shifting the lattices by a non-lattice point which is not the center cLc_{L} appears to be deeply more tricky in terms of energy optimization.

We remark that, since dcm\mathcal{F}_{d}^{cm} is not stable by difference, it is not totally surprising that Theorem 2.2 holds. Furthermore, identifying the largest space of all functions ff such that EfE_{f} is uniquely minimized by LdL_{d} in d(1)\mathcal{L}_{d}(1) seems to be very challenging (see [14]). Therefore a natural question in order to identify a large class of potentials ff such that the minimality of an universal optimizer LdL_{d} holds for EfκE_{f}^{\kappa} is the following: what are the completely monotone potentials fdcmf\in\mathcal{F}_{d}^{cm} satisfying (2.2), i.e. such that fκdcmf_{\kappa}\in\mathcal{F}_{d}^{cm}? The following corollary of Theorem 2.9 gives an example of such potentials, where we define, for s>0s>0 and any AK={ak}kKA_{K}=\{a_{k}\}_{k\in K}, K\{1}K\subset\mathbb{N}\backslash\{1\},

𝖫(AK,s):=kKakks.\mathsf{L}(A_{K},s):=\sum_{k\in K}\frac{a_{k}}{k^{s}}. (2.5)

Notice that the notation of (2.5) is inspired by the one of Dirichlet L-series that are generalizing the Riemann zeta function (see e.g. [20, Chap. 10]). For us, the arithmetic function appearing in a Dirichlet series is simply replaced by AKA_{K} and can be finite.

Corollary 2.13 (Minimality conservation for special ff - Non-shifted case).

Let d2d\geq 2 and fdcmf\in\mathcal{F}_{d}^{cm} be such that dμf(t)=ρf(t)dtd\mu_{f}(t)=\rho_{f}(t)dt and ρf\rho_{f} be an increasing function on +\mathbb{R}_{+}. Let κ={K,AK,}\kappa=\{K,A_{K},\emptyset\} be as in (1.6) where AK={ak}kK+A_{K}=\{a_{k}\}_{k\in K}\subset\mathbb{R}_{+} and be such that 𝖫(AK,s)\mathsf{L}(A_{K},s) defined by (2.5) satisfies 𝖫(AK,2)1\mathsf{L}(A_{K},2)\leq 1. If LdL_{d} is universally optimal in d(1)\mathcal{L}_{d}(1), then LdL_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1).

Example 2.14 (Potentials satisfying the assumptions of Corollary 2.13).

There are many examples of potentials ff such that Corollary 2.13 holds. For instance, this is the case for the parametrized potential f=fσ,sf=f_{\sigma,s} defined for all r>0r>0 by fσ,s(r)=eσrrsf_{\sigma,s}(r)=\frac{e^{-\sigma r}}{r^{s}}, σ>0\sigma>0, s>1s>1, since dμfσ,s(t)=(tσ)s1Γ(s)𝟙[σ,)(t)dtd\mu_{f_{\sigma,s}}(t)=\frac{(t-\sigma)^{s-1}}{\Gamma(s)}\mathds{1}_{[\sigma,\infty)}(t)dt and t(tσ)s1Γ(s)𝟙[σ,)(t)t\mapsto\frac{(t-\sigma)^{s-1}}{\Gamma(s)}\mathds{1}_{[\sigma,\infty)}(t) are increasing functions on +\mathbb{R}_{+}. Notice that the inverse power law f(r)=rsf(r)=r^{-s} with exponent s>d/21s>d/2\geq 1 (if σ=0\sigma=0) and the Yukawa potential f(r)=eσrr1f(r)=e^{-\sigma r}r^{-1} with parameter σ>0\sigma>0 (if s=1s=1) are special cases of fσ,sf_{\sigma,s}.

2.2 The inverse power law and Lennard-Jones cases

In this subsection, we restrict our study to combinations of inverse power laws, since they are the building blocks of many interaction potentials used in molecular simulations (see e.g. [33]). Their homogeneity simplifies a lot the energy computations and allows us to give a complete picture of the periodic arrays of defects effects with respect to the values of 𝖫\mathsf{L} defined by (2.5).

In the following result, we show that the values of 𝖫(AK,2s)\mathsf{L}(A_{K},2s) plays a fundamental role in the minimization of EfκE_{f}^{\kappa} when ff is an inverse power law.

Theorem 2.15 (The inverse power law case - Non-shifted case).

Let d2d\geq 2 and f(r)=rsf(r)=r^{-s} where s>d/2s>d/2. Let κ={K,AK,}\kappa=\{K,A_{K},\emptyset\} be as in (1.6) and be such that 𝖫(AK,2s)\mathsf{L}(A_{K},2s) defined by (2.5) is absolutely convergent. We have:

  1. 1.

    If 𝖫(AK,2s)<1\mathsf{L}(A_{K},2s)<1, then L0L_{0} is a minimizer of LζL(2s)L\mapsto\zeta_{L}(2s) in d(1)\mathcal{L}_{d}(1) if and only if L0L_{0} is a minimizer of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1).

  2. 2.

    If 𝖫(AK,2s)>1\mathsf{L}(A_{K},2s)>1, then L0L_{0} is a minimizer of LζL(2s)L\mapsto\zeta_{L}(2s) in d(1)\mathcal{L}_{d}(1) if and only if L0L_{0} is a maximizer of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1).

In particular, for any K\{1}K\subset\mathbb{N}\backslash\{1\}, if ak=1a_{k}=1 for all kKk\in K, then LζL(2s)L\mapsto\zeta_{L}(2s) and EfκE_{f}^{\kappa} have the same minimizers in d(1)\mathcal{L}_{d}(1).

Examples 2.16 (Minimizers of the Epstein zeta function).

In dimensions d{2,8,24}d\in\{2,8,24\}, the minimizer L0L_{0} of LζL(2s)L\mapsto\zeta_{L}(2s) in d(1)\mathcal{L}_{d}(1) is, respectively, 𝖠2\mathsf{A}_{2}, 𝖤8\mathsf{E}_{8} and Λ24\Lambda_{24} as consequences of [38, 23]. In dimension d=3d=3, Sarnak and Strömbergsson have conjectured in [43, Eq. (44)] that the face-centred cubic lattice 𝖣3\mathsf{D}_{3} (see Fig. 1) is the unique minimizer of LζL(2s)L\mapsto\zeta_{L}(2s) in 3(1)\mathcal{L}_{3}(1) if s>3/2s>3/2.

Many applications of point 4. of Theorem 2.9 can then be shown for non-convex sums of inverse power laws, differences of Yukawa potentials or Morse potentials by following the lines of [4]. In this paper, we have chosen to focus on Lennard-Jones type potentials since it is possible to have a complete description of the effect of non-shifted periodic arrays of vacancies using the homogeneity of the Epstein zeta functions. It is also known that Lennard-Jones type potentials play an important role in molecular simulation (see e.g. [4, Sect. 6.3] and [33, Sect. 5.1.2]).

In our last results, we define the Lennard-Jones type potential by

f(r)=c2rx2c1rx1where(c1,c2)(0,),x2>x1>d/2,f(r)=\frac{c_{2}}{r^{x_{2}}}-\frac{c_{1}}{r^{x_{1}}}\quad\textnormal{where}\quad(c_{1},c_{2})\in(0,\infty),\quad x_{2}>x_{1}>d/2, (2.6)

which is a prototypical example of function where μf\mu_{f} is not nonnegative everywhere, and a difference of completely monotone functions. We discuss the optimality of a universally optimal lattice LdL_{d} for EfκE_{f}^{\kappa} with respect to the values of 𝖫(AK,2xi)\mathsf{L}(A_{K},2x_{i}), i{1,2}i\in\{1,2\} as well as the shape of the global minimizer of EfκE_{f}^{\kappa}, i.e. its equivalence class in d\mathcal{L}_{d} modulo rotation and dilation (as previously defined in [14]).

Theorem 2.17 (The Lennard-Jones case - Non-shifted case).

Let d2d\geq 2, ff be defined by (2.6) and κ={K,AK,}\kappa=\{K,A_{K},\emptyset\} be as in (1.6) (possibly empty) and be such that 𝖫(AK,2xi)\mathsf{L}(A_{K},2x_{i}), i{1,2}i\in\{1,2\} defined by (2.5) are absolutely convergent. Let LdL_{d} be universally optimal in d(1)\mathcal{L}_{d}(1). Then:

  1. 1.

    If 𝖫(AK,2x2)<𝖫(AK,2x1)<1\mathsf{L}(A_{K},2x_{2})<\mathsf{L}(A_{K},2x_{1})<1, then for all V>0V>0 such that

    VVκ:=πd2(c2(1𝖫(AK,2x2))Γ(x1)c1(1𝖫(AK,2x1))Γ(x2))d2(x2x1),V\leq V_{\kappa}:=\pi^{\frac{d}{2}}\left(\frac{c_{2}(1-\mathsf{L}(A_{K},2x_{2}))\Gamma(x_{1})}{c_{1}(1-\mathsf{L}(A_{K},2x_{1}))\Gamma(x_{2})}\right)^{\frac{d}{2(x_{2}-x_{1})}},

    the lattice V1dLdV^{\frac{1}{d}}L_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(V)\mathcal{L}_{d}(V) and there exists V1>0V_{1}>0 such that it is not a minimizer of EfκE_{f}^{\kappa} for V>V1V>V_{1}. Furthermore, the shape of the minimizer of EfE_{f} and EfκE_{f}^{\kappa} are the same in d\mathcal{L}_{d}.

  2. 2.

    If 𝖫(AK,2x1)>𝖫(AK,2x2)>1\mathsf{L}(A_{K},2x_{1})>\mathsf{L}(A_{K},2x_{2})>1, then EfκE_{f}^{\kappa} does not have any minimizer in d\mathcal{L}_{d} and for all V<VκV<V_{\kappa}, V1dLdV^{\frac{1}{d}}L_{d} is the unique maximizer of EfκE_{f}^{\kappa} in d(V)\mathcal{L}_{d}(V).

  3. 3.

    If 𝖫(AK,2x1)>1>𝖫(AK,2x2)\mathsf{L}(A_{K},2x_{1})>1>\mathsf{L}(A_{K},2x_{2}), then EfκE_{f}^{\kappa} does not have any minimizer in d\mathcal{L}_{d} but V1dLdV^{\frac{1}{d}}L_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(V)\mathcal{L}_{d}(V) for all V>0V>0.

Remark 2.18 (Increasing of the threshold value VκV_{\kappa}).

The fact that 1𝖫(AK,2x2)>1𝖫(AK,2x1)1-\mathsf{L}(A_{K},2x_{2})>1-\mathsf{L}(A_{K},2x_{1}) implies that the threshold value VκV_{\kappa} is larger in the κ\kappa\neq\emptyset case than in the case without defect κ=\kappa=\emptyset. The same is expected to be true for any non-convex sum of inverse power law with a positive main term as r0r\to 0 (see [4, Prop. 6.4] for a twp-dimensional example in the no-defect case κ=\kappa=\emptyset). It is also totally straightforward to show that VκVV_{\kappa}\to V_{\emptyset} as minK\min K tend to ++\infty.

Remark 2.19 (Global minimality of 𝖠2\mathsf{A}_{2} among lattices for Lennard-Jones type potentials).

In dimension d=2d=2, the triangular lattice L2=𝖠2L_{2}=\mathsf{A}_{2} has been shown in [4, Thm. 1.2.2] to be the shape of the global minimizer of EfE_{f} in 2\mathcal{L}_{2} when πx2Γ(x2)x2<πx1Γ(x1)x1\pi^{-x_{2}}\Gamma(x_{2})x_{2}<\pi^{-x_{1}}\Gamma(x_{1})x_{1}. Point 1. of Theorem 2.17 implies that the same holds when 𝖫(AK,2x2)<𝖫(AK,2x1)<1\mathsf{L}(A_{K},2x_{2})<\mathsf{L}(A_{K},2x_{1})<1.

2.3 Conclusion

From all our results, we conclude that is possible to remove or substitute several infinite periodic sets of points from all the lattices (i.e. an integer sublattices) and to conserve the already existing minimality properties, but only in a certain class of potentials or sublattices. Physically, it means that adding point defects to a crystal can be without any effect on its ground state if we assume the interaction between atoms to be well-approximate by a pairwise potential (Born model [46]) and the sublattices to satisfy some simple properties. We give several examples in Section 3 and our result are the first known general results giving global optimality of ionic crystals. In particular, the Kagome lattice (see Figure 3) is shown to be the global minimizer for the interaction energies discussed in this paper in the class of (potentially shifted) lattices L\2LL\backslash 2L where L2(1)L\in\mathcal{L}_{2}(1). This is, as far as we know, the first results of this kind for the Kagome lattice. We also believe that the results and techniques derived in this paper can be applied to other ionic crystals and other general periodic systems.

Furthermore, this paper also shows the possibility to check the optimality of a structure while ’forgetting’ many points which, in a certain sense, do not play any role (vacancy case). This allow to simplify both numerical investigations – leading to a shorter computational time – and mathematical estimates for these energies. We voluntarily did not explore further this fact since it is only relevant in low dimensions because the computational time of such lattice sums is exponentially growing and gives unreachable durations in dimension d4d\geq 4 for computing many values of the energies, especially in dimensions d{8,24}d\in\{8,24\} where our global optimality results are applicable.

In dimension d=3d=3, i.e. where the everyday life real crystals exist, our results only apply – combined with the one from [6] – to the conservation of local minimality in the cubic lattices cases (3\mathbb{Z}^{3}, 𝖣3\mathsf{D}_{3} and 𝖣3\mathsf{D}_{3}^{*}) for the Epstein zeta function, the lattice theta function and the Lennard-Jones type energies. We believe that our result will find other very interesting applications in dimension 3 once global optimality properties will be shown for the lattice theta functions and the Epstein zeta functions (Sarnak-Strömbergsson conjectures [43]).

Even though the inverse power laws and Lennard-Jones cases have been completely solved here, we still ignore what is the optimal result that holds for ensuring the robustness of the universal optimality among lattices. An interesting problem would be to find a necessary condition for this robustness. Furthermore, we can also ask the following question: is it enough to study this kind of minimization problem in a (small) ball centred at the origin? In other words: can we remove all the points that are far enough from OO and conserving the minimality results? Numerical investigations and Figure 5 tend to confirm this fact, and a rigorous proof of such property would deeply simplify the analysis of such lattice energies.

3 Applications: The Kagome lattice and other ionic structures

We now give several examples of applications of our results. In particular, we identify interesting structures that are minimizers of EfE_{f} in classes of sparse and charged lattices.

3.1 The Kagome lattice

Being the vertices of a trihexagonal tiling, this structure – wich is actually not a lattice as we defined it in this paper – that we will write 𝖪:=𝖠2\2𝖠2\mathsf{K}:=\mathsf{A}_{2}\backslash 2\mathsf{A}_{2} is the difference of two triangular lattices of scale ratio 22 (see Fig. 4). Some minerals – which display novel physical properties connected with geometrically frustrated magnetism – like jarosites and herbertsmithite contain layers having this structure (see [37] and references therein). We can therefore apply our results of Section 2 with κ={2,1,}\kappa=\{2,1,\emptyset\} or κ={2,1,u1+u2}\kappa=\{2,1,u_{1}+u_{2}\}. The following optimality results for EfE_{f} in the class of lattices of the form L\2LL\backslash 2L (or L\(2L+u1+u2)L\backslash(2L+u_{1}+u_{2}) in the shifted case) are simple consequences of our results combined with the universal optimality of 𝖠2\mathsf{A}_{2} among lattices proved by Montgomery in [38]:

  1. 1.

    Universal optimality of 𝖪\mathsf{K}. Applying Theorem 2.4 to κ={2,1,u1+u2}\kappa=\{2,1,u_{1}+u_{2}\}, it follows that for all f2cmf\in\mathcal{F}_{2}^{cm}, the shifted Kagome lattice 𝖪+(1/2,3/2)\mathsf{K}+(1/2,-\sqrt{3}/2) (see Fig. 4) is the unique minimizer of EfE_{f} among lattices of the form L\(2L+u1+u2)L\backslash(2L+u_{1}+u_{2}), where L=u1u22(1)L=\mathbb{Z}u_{1}\oplus\mathbb{Z}u_{2}\in\mathcal{L}_{2}(1).

  2. 2.

    Minimality of 𝖪\mathsf{K} at all densities for certain completely monotone potentials. A direct consequence of Theorem 2.9 is the following. For any completely monotone function f2cmf\in\mathcal{F}_{2}^{cm} such that dμf(t)=ρf(t)dtd\mu_{f}(t)=\rho_{f}(t)dt and ρf\rho_{f} is an increasing function, the Kagome lattice 𝖪\mathsf{K} is the unique minimizer of EfE_{f} among all the two-dimensional sparse lattices L\2LL\backslash 2L where L2(1)L\in\mathcal{L}_{2}(1). This is the case for instance for f=fσ,sf=f_{\sigma,s} defined in Example 2.14, including the inverse power laws and the Yukawa potential.

  3. 3.

    Optimality at high density for Lennard-Jones interactions. Applying Theorem 2.17, we obtain its optimality at high density: if f(r)=c2rx2c1rx1f(r)=c_{2}r^{-x_{2}}-c_{1}r^{-x_{1}}, x2>x1>1x_{2}>x_{1}>1 is a Lennard-Jones potential, then the unique minimizer of EfE_{f} at high density among all the two-dimensional sparse lattices L\2LL\backslash 2L, where LL has fixed density, has the shape of 𝖪\mathsf{K}.

  4. 4.

    Global optimality for Lennard-Jones interactions with small exponents. Furthermore, using Theorem 2.17 and [4, Thm. 1.2.2] (see also Remark 2.19), we obtain the following interesting result in the Lennard-Jones potential case: if πx2Γ(x2)x2<πx1Γ(x1)x1\pi^{-x_{2}}\Gamma(x_{2})x_{2}<\pi^{-x_{1}}\Gamma(x_{1})x_{1}, then the unique global minimizer of EfE_{f} among all the possible sparse lattices L\2LL\backslash 2L has the shape of 𝖪\mathsf{K}.

Refer to caption
Refer to caption
Figure 4: Two patches of the Kagome lattice. On the left, the origin OO does not belong to 𝖪\mathsf{K} and is the center of one of the hexagons. On the right, OO belongs to a shifted version 𝖪+(1/2,3/2)\mathsf{K}+(1/2,-\sqrt{3}/2).

These are the first minimality results for 𝖪\mathsf{K} in a class of periodic configurations. We recall that a non-optimality result has also been derived by Grivopoulos in [31] for Lennard-Jones potential in the case of free particles, and different attempts have been made for obtaining numerically or experimentally a Kagome structure as an energy ground state (see e.g. [32, 41, 26]).

Remark 3.1 (The honeycomb lattice).

We notice that the honeycomb lattice 𝖧:=𝖠2\3𝖠2\mathsf{H}:=\mathsf{A}_{2}\backslash\sqrt{3}\mathsf{A}_{2}, also constructed from the triangular lattice, does not belong to the set of sparse lattices L\kLL\backslash kL, kk\in\mathbb{N}. That is why no optimality result for 𝖧\mathsf{H} is included in this paper.

3.2 Rock-salt vs. other ionic structures

We recall that, in [9], we have shown with Faulhuber the universal optimality of the triangular lattice among lattices with alternating charges, i.e. the fact that 𝖠2\mathsf{A}_{2} uniquely maximizes

LθL±(α):=pLφ±(p)eπα|p|2andζL±(s):=pL\{0}φ±(p)|p|s,L=u1u2,L\mapsto\theta_{L}^{\pm}(\alpha):=\sum_{p\in L}\varphi_{\pm}(p)e^{-\pi\alpha|p|^{2}}\quad\textnormal{and}\quad\zeta_{L}^{\pm}(s):=\sum_{p\in L\backslash\{0\}}\frac{\varphi_{\pm}(p)}{|p|^{s}},\quad L=\mathbb{Z}u_{1}\oplus\mathbb{Z}u_{2}, (3.1)

in 2(1)\mathcal{L}_{2}(1), where, for all p=mu1+nu2p=mu_{1}+nu_{2}, φ±(p):=m+n\varphi_{\pm}(p):=m+n. Notice that the maximality result at all scales for the alternating lattice theta function is equivalent with the fact that 𝖠2\mathsf{A}_{2} maximizes

LEfκ[L]:=Ef[L]2Ef[2L+u1]2Ef[2L+u2],whereκ:={2,{2,2},{u1,u2}}L\mapsto E_{f}^{\kappa}[L]:=E_{f}[L]-2E_{f}[2L+u_{1}]-2E_{f}[2L+u_{2}],\quad\textnormal{where}\quad\kappa:=\{2,\{2,2\},\{u_{1},u_{2}\}\}

in 2(1)\mathcal{L}_{2}(1) for any f2cmf\in\mathcal{F}_{2}^{cm}. It has been also proven in [13, Thm. 1.4] that d\mathbb{Z}^{d} is the unique maximizer of the dd-dimensional generalization of the two lattice energies θL±(α)\theta_{L}^{\pm}(\alpha) and ζL±(s)\zeta_{L}^{\pm}(s) among dd-dimensional orthorhombic (rectangular) lattices of fixed unit density, whereas it is a minimizer of the lattice theta functions and the Epstein zeta functions defined in (1.3). Furthermore, applying Theorem 2.9 in dimension d=2d=2 (resp. any dd), we see that 𝖠2\mathsf{A}_{2} (resp. d\mathbb{Z}^{d}) minimizes in 2(1)\mathcal{L}_{2}(1) (resp. among the orthorhombic lattices of unit density) the energy

Efκ[L]:=ζL(s)2ζkL(s),f(r)=rs,K={k},ak=2,E_{f}^{\kappa}[L]:=\zeta_{L}(s)-2\zeta_{kL}(s),\quad f(r)=r^{-s},\quad K=\{k\},\quad a_{k}=2, (3.2)

for all s>d/2s>d/2. We remark that d\mathbb{Z}^{d}, d{2,3}d\in\{2,3\} is also a saddle point (see [38, 6]) of EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1). It is then interesting to see how the array of substitutional defects with charges 1-1 plays a totally different role for this energy (see also Fig. 5 and Fig. 6). This seems to confirm that the role of the nearest-neighbors of the origin is fundamental, since they are actually the main terms of the energy when the potential is decreasing fast at infinity.

Refer to caption
Refer to caption
Refer to caption
Figure 5: Three periodic arrays of defects on 2\mathbb{Z}^{2}. Blue points {\color[rgb]{0,0,1}\bullet} are points with charges +1+1 and red points {\color[rgb]{1,0,0}\circ} are with charges 1-1. For the inverse power laws energies, the left-hand configuration is the unique maximizer among rectangular lattices of fixed density with alternation of charges whereas the centred configuration is its unique minimizer with this distribution of charges among rectangular lattices. However, the configuration on the right is a saddle point of any energy on the form EfE_{f}, f2cmf\in\mathcal{F}_{2}^{cm} in this class of charged configurations. For the two structures on the left, the same is true in higher dimension while generalizing the ionic-like distribution on orthorhombic lattices.
Refer to caption
Refer to caption
Refer to caption
Figure 6: Three periodic arrays of defects on a patch of 𝖠2\mathsf{A}_{2}. Blue points {\color[rgb]{0,0,1}\bullet} are points with charges +1+1 and red points {\color[rgb]{1,0,0}\circ} are with charges 1-1. On the left, the triangular alternate configuration maximizes ζL±(s)\zeta_{L}^{\pm}(s) in 2(1)\mathcal{L}_{2}(1) with this alternation of charges, while the configuration in the middle minimizes the inverse power law energy in this class of charged lattices. The configuration on the right minimizes any energy on the form EfE_{f}, f2cmf\in\mathcal{F}_{2}^{cm} in this class of charged configurations.

4 Proofs of the main results

We first show Theorem 2.2, i.e. the non-robustness of universal optimality results under non-shifted periodic arrays of defects.

Proof of Theorem 2.2.

Let Λ{𝖠2,𝖤8,Λ24}\Lambda\in\{\mathsf{A}_{2},\mathsf{E}_{8},\Lambda_{24}\}. We consider the potential f(r):=eπαrf(r):=e^{-\pi\alpha r} where α>0\alpha>0. For all k\{1}k\in\mathbb{N}\backslash\{1\}, all ak>0a_{k}>0 and all Ld(1)L\in\mathcal{L}_{d}(1), we have, using the fact that θkL(α)=θL(k2α)\theta_{kL}(\alpha)=\theta_{L}(k^{2}\alpha),

Efκ[L]=θL(α)akθ(k2α).E_{f}^{\kappa}[L]=\theta_{L}(\alpha)-a_{k}\theta(k^{2}\alpha).

Let us show that there exists αd\alpha_{d} such that for all 0<α<αd0<\alpha<\alpha_{d}, Λ\Lambda does not minimize EfκE_{f}^{\kappa} in d(1)\mathcal{L}_{d}(1). Indeed, we have the following equivalence: for all Ld(1)\{Λ}L\in\mathcal{L}_{d}(1)\backslash\{\Lambda\}, Efκ[L]>Efκ[Λ]E_{f}^{\kappa}[L]>E_{f}^{\kappa}[\Lambda] if and only if

infLd(1)LΛθL(α)θΛ(α)θL(k2α)θΛ(k2α)>ak.\inf_{L\in\mathcal{L}_{d}(1)\atop L\neq\Lambda}\frac{\theta_{L}(\alpha)-\theta_{\Lambda}(\alpha)}{\theta_{L}(k^{2}\alpha)-\theta_{\Lambda}(k^{2}\alpha)}>a_{k}. (4.1)

Let us show that (4.1) does not hold for small α\alpha, and in particular that the left term tends to 0 as α0\alpha\to 0. We use Coulangeon and Schürmann’s work [25, Eq. (21)], in the lattice case, who derived the Taylor expansion of the theta function as LΛL\mapsto\Lambda in d(1)\mathcal{L}_{d}(1). We then obtain

limLΛLΛθL(α)θΛ(α)θL(k2α)θΛ(k2α)\displaystyle\lim_{L\mapsto\Lambda\atop L\neq\Lambda}\frac{\theta_{L}(\alpha)-\theta_{\Lambda}(\alpha)}{\theta_{L}(k^{2}\alpha)-\theta_{\Lambda}(k^{2}\alpha)} =pΛ\{0}πα|p|2(πα|p|22)eπα|p|2pΛ\{0}παk2|p|2(παk2|p|22)eπαk2|p|2\displaystyle=\frac{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}\pi\alpha|p|^{2}\left(\pi\alpha|p|^{2}-2\right)e^{-\pi\alpha|p|^{2}}}{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}\pi\alpha k^{2}|p|^{2}\left(\pi\alpha k^{2}|p|^{2}-2\right)e^{-\pi\alpha k^{2}|p|^{2}}}
=k2pΛ\{0}πα|p|4eπα|p|22pΛ\{0}|p|2eπα|p|2pΛ\{0}παk2|p|4eπαk2|p|22pΛ\{0}|p|2eπαk2|p|2.\displaystyle=k^{-2}\frac{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}\pi\alpha|p|^{4}e^{-\pi\alpha|p|^{2}}-2\sum_{p\in\Lambda\backslash\{0\}}|p|^{2}e^{-\pi\alpha|p|^{2}}}{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}\pi\alpha k^{2}|p|^{4}e^{-\pi\alpha k^{2}|p|^{2}}-2\sum_{p\in\Lambda\backslash\{0\}}|p|^{2}e^{-\pi\alpha k^{2}|p|^{2}}}.

By absolute convergence, the first term of both numerator and denominator are vanishing as α0\alpha\to 0. We therefore obtain that

limα0limLΛLΛθL(α)θΛ(α)θL(k2α)θΛ(k2α)=limα0k2pΛ\{0}|p|2eπα|p|2pΛ\{0}|p|2eπαk2|p|2=0,\lim_{\alpha\to 0}\lim_{L\mapsto\Lambda\atop L\neq\Lambda}\frac{\theta_{L}(\alpha)-\theta_{\Lambda}(\alpha)}{\theta_{L}(k^{2}\alpha)-\theta_{\Lambda}(k^{2}\alpha)}=\lim_{\alpha\to 0}k^{-2}\frac{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}|p|^{2}e^{-\pi\alpha|p|^{2}}}{\displaystyle\sum_{p\in\Lambda\backslash\{0\}}|p|^{2}e^{-\pi\alpha k^{2}|p|^{2}}}=0,

by comparing the convergence rate of these two exponential sums that are going to ++\infty as α0\alpha\to 0. It follows that (4.1) does not hold for α<αd\alpha<\alpha_{d} where αd\alpha_{d} depends on dd, kk and aka_{k}, and the proof of the first part of the theorem is completed.

The second part of the theorem is a simple consequence of the fact that fκf_{\kappa} defined by (1.8) belongs to dcm\mathcal{F}_{d}^{cm} if fdcmf\in\mathcal{F}_{d}^{cm} and ak<0a_{k}<0 for all kKk\in K. ∎

The proof of our second result, namely Theorem 2.4, is a direct and simple consequence of our work [9].

Proof of Theorem 2.4.

If pi,k/k=cLp_{i,k}/k=c_{L} modulo QLQ_{L} for all kKk\in K and all iIki\in I_{k}, we obtain

Efκ[L]=Ef[L]kKakiIkpL\{0}f(k2|pi,kk+p|2)=Ef[L]kKakLkEf(k2)[L+cL].E_{f}^{\kappa}[L]=E_{f}[L]-\sum_{k\in K}a_{k}\sum_{i\in I_{k}}\sum_{p\in L\backslash\{0\}}f\left(k^{2}\left|\frac{p_{i,k}}{k}+p\right|^{2}\right)=E_{f}[L]-\sum_{k\in K}a_{k}\sharp L_{k}E_{f(k^{2}\cdot)}[L+c_{L}].

As proved in [9], for any f2cmf\in\mathcal{F}_{2}^{cm}, 𝖠2\mathsf{A}_{2} is the unique maximizer of LEf[L+cL]L\mapsto E_{f}[L+c_{L}] in 2(1)\mathcal{L}_{2}(1). It follows that 𝖠2\mathsf{A}_{2}, which uniquely minimizes EfE_{f} in 2(1)\mathcal{L}_{2}(1) is the unique minimizer of EfκE_{f}^{\kappa} in 2(1)\mathcal{L}_{2}(1) since ak>0a_{k}>0 for all kKk\in K. ∎

We now show Theorem 2.9 which gives a simple criterion for the conservation of the minimality of a universal optimizer.

Proof of Theorem 2.9.

In order to show the three first points, it is sufficient to show the first point of our theorem, i.e. the fact that dμfκ(t)=(ρf(t)kKakk2ρf(tk2))dtd\mu_{f_{\kappa}}(t)=\left(\rho_{f}(t)-\sum_{k\in K}a_{k}k^{-2}\rho_{f}\left(\frac{t}{k^{2}}\right)\right)dt. We remark that ρf\rho_{f} is the inverse Laplace transform of ff, i.e. ρf(t)=1[f](t)\rho_{f}(t)=\mathcal{L}^{-1}[f](t). By linearity, it follows that

dμfκ(t)=ρfκ(t)dt,whereρfκ(t)=ρf(t)kKak1[f(k2)](t).d\mu_{f_{\kappa}}(t)=\rho_{f_{\kappa}}(t)dt,\quad\textnormal{where}\quad\rho_{f_{\kappa}}(t)=\rho_{f}(t)-\sum_{k\in K}a_{k}\mathcal{L}^{-1}[f(k^{2}\cdot)](t).

By the basic properties of the inverse Laplace transform, we obtain that, for all t>0t>0,

1[f(k2)](t)=k21[f](k2t)=k2ρf(k2t),\mathcal{L}^{-1}[f(k^{2}\cdot)](t)=k^{-2}\mathcal{L}^{-1}[f](k^{-2}t)=k^{-2}\rho_{f}(k^{-2}t),

and our result follows by the universal optimality of LdL_{d} in d(1)\mathcal{L}_{d}(1) and the definition of completely monotone function.

To show the last point of our theorem, we adapt [4, Thm 1.1]. Let Ld(1)L\in\mathcal{L}_{d}(1) and V>0V>0, then we have

Efκ[V1dL]\displaystyle E_{f}^{\kappa}[V^{\frac{1}{d}}L] =pL\{0}fκ(V2d|p|2)=0[θL(V2dtπ)1]ρfκ(t)𝑑t\displaystyle=\sum_{p\in L\backslash\{0\}}f_{\kappa}\left(V^{\frac{2}{d}}|p|^{2}\right)=\int_{0}^{\infty}\left[\theta_{L}\left(\frac{V^{\frac{2}{d}}t}{\pi}\right)-1\right]\rho_{f_{\kappa}}(t)dt
=πV2d0[θL(y)1]ρfκ(πyV2d)𝑑y\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{0}^{\infty}\left[\theta_{L}(y)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy
=πV2d01[θL(y)1]ρfκ(πyV2d)𝑑y+πV2d1[θL(y)1]ρfκ(πyV2d)𝑑y\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{0}^{1}\left[\theta_{L}(y)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}(y)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy
=πV2d1[θL(1y)1]ρfκ(πyV2d)y2𝑑y+πV2d1[θL(y)1]ρfκ(πyV2d)𝑑y.\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}\left(\frac{1}{y}\right)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi}{yV^{\frac{2}{d}}}\right)y^{-2}dy+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}(y)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy. (4.2)

A simple consequence of the Poisson summation formula is the well-known identity (see e.g. [24, Eq. (43)])

y>0,θL(1y)=yd2θL(y).\forall y>0,\quad\theta_{L}\left(\frac{1}{y}\right)=y^{\frac{d}{2}}\theta_{L^{*}}(y). (4.3)

From (4.3), we see that if LdL_{d} is the unique minimizer of LθL(α)L\mapsto\theta_{L}(\alpha) for all α>0,Ld(1)\alpha>0,L\in\mathcal{L}_{d}(1) then Ld=LdL_{d}^{*}=L_{d}. From (4.2) and (4.3), for all V>0,Ld(1)V>0,L\in\mathcal{L}_{d}(1), we have

Efκ[V1dL]\displaystyle E_{f}^{\kappa}[V^{\frac{1}{d}}L] =πV2d1[yd2θL(y)1]ρfκ(πyV2d)y2𝑑y+πV2d1[θL(y)1]ρfκ(πyV2d)𝑑y.\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[y^{\frac{d}{2}}\theta_{L^{*}}\left(y\right)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi}{yV^{\frac{2}{d}}}\right)y^{-2}dy+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}(y)-1\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy. (4.4)

and

Efκ[V1dL]Efκ[V1dLd]\displaystyle E_{f}^{\kappa}[V^{\frac{1}{d}}L]-E_{f}^{\kappa}[V^{\frac{1}{d}}L_{d}] =πV2d1[θL(y)θLd(y)]ρfκ(πyV2d)yd22𝑑y\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L^{*}}\left(y\right)-\theta_{L_{d}}(y)\right]\rho_{f_{\kappa}}\left(\frac{\pi}{yV^{\frac{2}{d}}}\right)y^{\frac{d}{2}-2}dy
+πV2d1[θL(y)θLd(y)]ρfκ(πyV2d)𝑑y.\displaystyle\quad\quad+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}(y)-\theta_{L_{d}}(y)\right]\rho_{f_{\kappa}}\left(\frac{\pi y}{V^{\frac{2}{d}}}\right)dy. (4.5)

By (4) and the definition of gVg_{V}, if VV is such that gV(y)0g_{V}(y)\geq 0 for a.e. y1y\geq 1 then

Efκ[V1dL]Efκ[V1dLd]+Efκ[V1dL]Efκ[V1dLd]\displaystyle E_{f}^{\kappa}[V^{\frac{1}{d}}L]-E_{f}^{\kappa}[V^{\frac{1}{d}}L_{d}]+E_{f}^{\kappa}[V^{\frac{1}{d}}L^{*}]-E_{f}^{\kappa}[V^{\frac{1}{d}}L_{d}]
=πV2d1[θL(y)θLd(y)]gV(y)𝑑y+πV2d1[θL(y)θLd(y)]gV(y)𝑑y\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L^{*}}\left(y\right)-\theta_{L_{d}}(y)\right]g_{V}(y)dy+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}(y)-\theta_{L_{d}}(y)\right]g_{V}(y)dy
πV2d1mL(y)gV(y)𝑑y,\displaystyle\geq\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}m_{L}(y)g_{V}(y)dy, (4.6)

where

mL(y):=min{θL(y)θLd(y),θL(y)θLd(y)}.m_{L}(y):=\min\{\theta_{L^{*}}\left(y\right)-\theta_{L_{d}}(y),\theta_{L}(y)-\theta_{L_{d}}(y)\}.

Since mL(y)0m_{L}(y)\geq 0 for all Ld(1),y>0L\in\mathcal{L}_{d}(1),y>0 with equality if and only if L=LdL=L_{d}, and gV(y)0g_{V}(y)\geq 0 for a.e. y[1,)y\in[1,\infty), we get from (4) that

Efκ[V1dL]+Efκ[V1dL]2Efκ[V1dLd],with equality if and only if L=Ld.E_{f}^{\kappa}[V^{\frac{1}{d}}L]+E_{f}^{\kappa}[V^{\frac{1}{d}}L^{*}]\geq 2E_{f}^{\kappa}[V^{\frac{1}{d}}L_{d}],\quad\mbox{with equality if and only if }L=L_{d}.

It follows that LdL_{d} is the unique minimizer of LEfκ[V1dL]L\mapsto E_{f}^{\kappa}[V^{\frac{1}{d}}L] on d(1)\mathcal{L}_{d}(1), or equivalently that V1dLdV^{\frac{1}{d}}L_{d} is the unique minimizer of EfκE_{f}^{\kappa} in d(V)\mathcal{L}_{d}(V), and the result is proved. ∎

The previous proof contains the main ingredients for showing Theorem 2.11.

Proof of Theorem 2.11.

Following exactly the same sequence of arguments as in the proof of the fourth point of Theorem 2.9, we obtain the maximality result of V1dLd±V^{\frac{1}{d}}L_{d}^{\pm} at fixed density for Ef±E_{f}^{\pm}. Indeed, (4.3) is replaced by

θL±(α)=yd2θL+cL(α),\theta_{L}^{\pm}(\alpha)=y^{\frac{d}{2}}\theta_{L^{*}+c_{L^{*}}}(\alpha),

and, by using the maximality of Ld±L_{d}^{\pm} for LθL±(α)L\mapsto\theta_{L}^{\pm}(\alpha) and LθL+cL(α)L\mapsto\theta_{L+c_{L}}(\alpha) for all α>0\alpha>0, we obtain

Ef±[V1dL]Ef±[V1dLd]+Ef±[V1dL]Ef±[V1dLd]\displaystyle E_{f}^{\pm}[V^{\frac{1}{d}}L]-E_{f}^{\pm}[V^{\frac{1}{d}}L_{d}]+E_{f}^{\pm}[V^{\frac{1}{d}}L^{*}]-E_{f}^{\pm}[V^{\frac{1}{d}}L_{d}]
=πV2d1[θL+cL(y)θLd±+cLd±(y)]gV(y)𝑑y+πV2d1[θL±(y)θLd±±(y)]gV(y)𝑑y\displaystyle=\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L^{*}+c_{L^{*}}}\left(y\right)-\theta_{L_{d}^{\pm}+c_{L_{d}^{\pm}}}(y)\right]g_{V}(y)dy+\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}\left[\theta_{L}^{\pm}(y)-\theta_{L_{d}^{\pm}}^{\pm}(y)\right]g_{V}(y)dy
πV2d1mL±(y)gV(y)𝑑y,\displaystyle\leq\frac{\pi}{V^{\frac{2}{d}}}\int_{1}^{\infty}m_{L}^{\pm}(y)g_{V}(y)dy, (4.7)

where

mL±(y):=max{θL+cL(y)θLd±+cLd±(y),θL±(y)θLd±±(y)}.m_{L}^{\pm}(y):=\max\{\theta_{L^{*}+c_{L^{*}}}\left(y\right)-\theta_{L_{d}^{\pm}+c_{L_{d}^{\pm}}}(y),\theta_{L}^{\pm}(y)-\theta_{L_{d}^{\pm}}^{\pm}(y)\}.

We again remark that mL±(y)0m_{L}^{\pm}(y)\leq 0 for all Ld(1)L\in\mathcal{L}_{d}(1), y>0y>0 with equality if and only if L=Ld±L=L_{d}^{\pm}. Therefore, the positivity of gVg_{V} as well as the universal maximality of Ld±L_{d}^{\pm} implies in the same way that V1dLd±V^{\frac{1}{d}}L_{d}^{\pm} is the unique maximizer of Ef±E_{f}^{\pm} in d(V)\mathcal{L}_{d}(V). ∎

The proof of Corollary 2.13 is a straightforward consequence of Theorem 2.9.

Proof of Corollary 2.13.

Let AK:={ak}kK+A_{K}:=\{a_{k}\}_{k\in K}\subset\mathbb{R}_{+} be such that 𝖫(AK,2)1\mathsf{L}(A_{K},2)\leq 1. Since μf0\mu_{f}\geq 0, it follows that ρf\rho_{f} is positive, and furthermore ρf\rho_{f} is increasing by assumption. Therefore, we have, for all t>0t>0,

kKakk2ρf(tα2)kKakk2ρf(t)=𝖫(AK,2)ρf(t)ρf(t),\sum_{k\in K}\frac{a_{k}}{k^{2}}\rho_{f}\left(\frac{t}{\alpha^{2}}\right)\leq\sum_{k\in K}\frac{a_{k}}{k^{2}}\rho_{f}(t)=\mathsf{L}(A_{K},2)\rho_{f}(t)\leq\rho_{f}(t),

where the first inequality is obtained from the monotonicity of ρf\rho_{f} and the last one from its positivity and the fact that 𝖫(AK,2)1\mathsf{L}(A_{K},2)\leq 1. The proof is completed by applying Theorem 2.9. ∎

We now show Theorem 2.15 which is a simple consequence of the homogeneity of the Epstein zeta function and a property of the Riemann zeta function.

Proof of Theorem 2.15.

Using the homogeneity of the Epstein zeta function, we obtain

Efκ[L]=pL\{0}1|p|2skKpL\{0}akk2s|p|2s=(1𝖫(AK,2s))ζL(2s),E_{f}^{\kappa}[L]=\sum_{p\in L\backslash\{0\}}\frac{1}{|p|^{2s}}-\sum_{k\in K}\sum_{p\in L\backslash\{0\}}\frac{a_{k}}{k^{2s}|p|^{2s}}=\left(1-\mathsf{L}(A_{K},2s)\right)\zeta_{L}(2s),

the exchange of sums being ensured by their absolute summability. If 𝖫(AK,2s)<1\mathsf{L}(A_{K},2s)<1, then LζL(2s)L\mapsto\zeta_{L}(2s) and EfκE_{f}^{\kappa} have exactly the same minimizer. If 𝖫(AK,2s)>1\mathsf{L}(A_{K},2s)>1, then the optimality are reversed and the proof is complete.

Furthermore, if ak=1a_{k}=1 for all kKk\in K, then we have

𝖫(AK,2s)=kK1k2sζ(2s)1,\mathsf{L}(A_{K},2s)=\sum_{k\in K}\frac{1}{k^{2s}}\leq\zeta(2s)-1,

where ζ(s):=nns\zeta(s):=\sum_{n\in\mathbb{N}}n^{-s} is the Riemann zeta function. Since ζ(x)<2\zeta(x)<2 on (0,)(0,\infty) if and only if x>x01.73x>x_{0}\approx 1.73, it follows that ζ(2s)1<1\zeta(2s)-1<1 if and only if s>x0/20.865s>x_{0}/2\approx 0.865 which is true for all s>d/2s>d/2 whenever d2d\geq 2. We thus have 𝖫(AK,2s)<1\mathsf{L}(A_{K},2s)<1 and the proof is completed by application of point 1. of the theorem. ∎

Before proving Theorem 2.17, we derive the following result, a generalization of our two-dimensional theorem [4, Prop. 6.11]. Its proof follows the same main arguments as the two-dimensional version and it is a consequence of point 4. of Theorem 2.9.

Proposition 4.1 (Optimality at high density for Lennard-Jones type potentials).

Let f(r)=b2rx2b1rx1f(r)=\frac{b_{2}}{r^{x_{2}}}-\frac{b_{1}}{r^{x_{1}}} where b1,b2(0,)b_{1},b_{2}\in(0,\infty) and x2>x1>d/2x_{2}>x_{1}>d/2, and let LdL_{d} be universally optimal in d(1)\mathcal{L}_{d}(1). If

Vπd2(b2Γ(x1)b1Γ(x2))d2(x2x1),V\leq\pi^{\frac{d}{2}}\left(\frac{b_{2}\Gamma(x_{1})}{b_{1}\Gamma(x_{2})}\right)^{\frac{d}{2(x_{2}-x_{1})}},

then V1dLdV^{\frac{1}{d}}L_{d} is the unique minimizer of EfE_{f} in d(V)\mathcal{L}_{d}(V).

Proof of Proposition 4.1.

We follow the lines of [4, Prop. 6.10] and we apply point 4. of Theorem 2.9. For i{1,2}i\in\{1,2\}, let βi:=biπxi1Γ(xi)\beta_{i}:=b_{i}\frac{\pi^{x_{i}-1}}{\Gamma(x_{i})} and α:=V2d\alpha:=V^{\frac{2}{d}}, then gV(y)=yd2x21αx11g~V(y)g_{V}(y)=\frac{y^{\frac{d}{2}-x_{2}-1}}{\alpha^{x_{1}-1}}\tilde{g}_{V}(y) where gVg_{V} is given by (2.3) and

g~V(y):=β2αx2x1y2x2d2β1yx2+x1d2β1yx2x1+β2αx2x1.\tilde{g}_{V}(y):=\frac{\beta_{2}}{\alpha^{x_{2}-x_{1}}}y^{2x_{2}-\frac{d}{2}}-\beta_{1}y^{x_{2}+x_{1}-\frac{d}{2}}-\beta_{1}y^{x_{2}-x_{1}}+\frac{\beta_{2}}{\alpha^{x_{2}-x_{1}}}.

We therefore compute g~V(y)=yx2x11uV(y)\tilde{g}_{V}^{\prime}(y)=y^{x_{2}-x_{1}-1}u_{V}(y) where

uV(y):=β2(2x2d2)yx2+x1d2αx2x1β1(x2+x1d2)y2x1d2β1(x2x1).u_{V}(y):=\beta_{2}\left(2x_{2}-\frac{d}{2}\right)\frac{y^{x_{2}+x_{1}-\frac{d}{2}}}{\alpha^{x_{2}-x_{1}}}-\beta_{1}\left(x_{2}+x_{1}-\frac{d}{2}\right)y^{2x_{1}-\frac{d}{2}}-\beta_{1}(x_{2}-x_{1}).

Differentiating again, we obtain

uV(y)=(x2+x1d2)y2x1d21(β2(2x2d2)yx2x1αx2x1β1(2x1d2)),u_{V}^{\prime}(y)=\left(x_{2}+x_{1}-\frac{d}{2}\right)y^{2x_{1}-\frac{d}{2}-1}\left(\beta_{2}\left(2x_{2}-\frac{d}{2}\right)\frac{y^{x_{2}-x_{1}}}{\alpha^{x_{2}-x_{1}}}-\beta_{1}\left(2x_{1}-\frac{d}{2}\right)\right),

and we have that uV(y)0u_{V}^{\prime}(y)\geq 0 if and only if y(β1(2x1d2)β2(2x2d2))1x2x1αy\geq\left(\frac{\beta_{1}(2x_{1}-\frac{d}{2})}{\beta_{2}(2x_{2}-\frac{d}{2})}\right)^{\frac{1}{x_{2}-x_{1}}}\alpha. By assumption, we know that

απ(a2Γ(x1)a1Γ(x2))1x2x1=(β2β1)1x2x1<(β2(2x2d2)β1(2x1d2))1x2x1,\alpha\leq\pi\left(\frac{a_{2}\Gamma(x_{1})}{a_{1}\Gamma(x_{2})}\right)^{\frac{1}{x_{2}-x_{1}}}=\left(\frac{\beta_{2}}{\beta_{1}}\right)^{\frac{1}{x_{2}-x_{1}}}<\left(\frac{\beta_{2}(2x_{2}-\frac{d}{2})}{\beta_{1}(2x_{1}-\frac{d}{2})}\right)^{\frac{1}{x_{2}-x_{1}}},

which implies that uV(y)0u_{V}^{\prime}(y)\geq 0 for all y1y\geq 1. We now remark that

uV(1)=(2x2d2)(β2αx2x1β1)0,u_{V}(1)=\left(2x_{2}-\frac{d}{2}\right)\left(\frac{\beta_{2}}{\alpha^{x_{2}-x_{1}}}-\beta_{1}\right)\geq 0,

by assumption, since p>d/2>d/4p>d/2>d/4 and

απ(b2Γ(x1)b1Γ(x2))1x2x1β2αx2x1β10.\displaystyle\alpha\leq\pi\left(\frac{b_{2}\Gamma(x_{1})}{b_{1}\Gamma(x_{2})}\right)^{\frac{1}{x_{2}-x_{1}}}\iff\frac{\beta_{2}}{\alpha^{x_{2}-x_{1}}}-\beta_{1}\geq 0. (4.8)

It follows that gV(y)0g_{V}^{\prime}(y)\geq 0 for all y1y\geq 1. Since

gV(1)=2(β2αx21β1αx11)0g_{V}(1)=2\left(\frac{\beta_{2}}{\alpha^{x_{2}-1}}-\frac{\beta_{1}}{\alpha^{x_{1}-1}}\right)\geq 0

again by (4.8), gV(y)0g_{V}(y)\geq 0 for all y1y\geq 1 and the proof is complete. ∎

Proof of Theorem 2.17.

Let AK={ak}kKA_{K}=\{a_{k}\}_{k\in K} for some K\{1}K\subset\mathbb{N}\backslash\{1\} and f(r)=c2rx2c1rx1f(r)=c_{2}r^{-x_{2}}-c_{1}r^{-x_{1}}, then we have, using the homogeneity of the Epstein zeta function,

Efκ[L]\displaystyle E_{f}^{\kappa}[L] =c2ζL(2x2)c1ζL(2x1)kKak(c2ζkL(2x2)c1ζkL(2x1))\displaystyle=c_{2}\zeta_{L}(2x_{2})-c_{1}\zeta_{L}(2x_{1})-\sum_{k\in K}a_{k}\left(c_{2}\zeta_{kL}(2x_{2})-c_{1}\zeta_{kL}(2x_{1})\right)
=c2(1𝖫(AK,2x2))ζL(2x2)c1(1𝖫(AK,2x1)))ζL(2x1).\displaystyle=c_{2}\left(1-\mathsf{L}(A_{K},2x_{2})\right)\zeta_{L}(2x_{2})-c_{1}\left(1-\mathsf{L}(A_{K},2x_{1}))\right)\zeta_{L}(2x_{1}).

We now assume that 𝖫(AK,2x2)<𝖫(AK,2x1)<1\mathsf{L}(A_{K},2x_{2})<\mathsf{L}(A_{K},2x_{1})<1. Therefore, the first part of point 1. is a simple consequence of Prop.4.1 applied for the coefficients bi=ci(1𝖫(AK,2xi))>0b_{i}=c_{i}\left(1-\mathsf{L}(A_{K},2x_{i})\right)>0 where i{1,2}i\in\{1,2\}. The fact that EfκE_{f}^{\kappa} is not minimized by LdL_{d} for VV large enough is a direct application of [14, Thm. 1.5(1)] since μf\mu_{f} is negative on (0,r0)(0,r_{0}) for some r0r_{0} depending on the parameters c1,c2,x1,x2,AKc_{1},c_{2},x_{1},x_{2},A_{K}. Furthermore, the fact that the shape of the minimizers are the same follows from [14, Thm. 1.11] where it is shown that the minimizer of the Lennard-Jones type lattice energies does not depend on the coefficients b1,b2b_{1},b_{2} but only on the exponents x1,x2x_{1},x_{2}, which are the same for ff and fκf_{\kappa}.

If 𝖫(AK,2x1)>𝖫(AK,2x2)>1\mathsf{L}(A_{K},2x_{1})>\mathsf{L}(A_{K},2x_{2})>1, then fκ(r)=b2rx2+b1rx1f_{\kappa}(r)=-b_{2}r^{-x_{2}}+b_{1}r^{-x_{1}} where bi:=ci(𝖫(AK,2xi)1)>0b_{i}:=c_{i}\left(\mathsf{L}(A_{K},2x_{i})-1\right)>0, i{1,2}i\in\{1,2\}. If follows that fκ(r)f_{\kappa}(r) tends to -\infty as r0r\to 0, which implies the same for Efκ[L]E_{f}^{\kappa}[L] as LL has its lengths going to 0 and ++\infty, i.e. when LL collapses. This means that EfκE_{f}^{\kappa} does not have a minimizer in d(V)\mathcal{L}_{d}(V) and in d\mathcal{L}_{d}. Furthermore, combining point 1. with the fact that the signs of the coefficients are switched, we obtain the maximality of V1/dLdV^{1/d}L_{d} at high density (i.e. low volume V<VκV<V_{\kappa}).

If 𝖫(AK,2x1)>1>𝖫(AK,2x2)\mathsf{L}(A_{K},2x_{1})>1>\mathsf{L}(A_{K},2x_{2}), then fκ(r)=b2rx2+b1rx1f_{\kappa}(r)=b_{2}r^{-x_{2}}+b_{1}r^{-x_{1}} where b1:=c1(𝖫(AK,2x1)1)>0b_{1}:=c_{1}\left(\mathsf{L}(A_{K},2x_{1})-1\right)>0 and b2:=c2(1𝖫(AK,2x2))>0b_{2}:=c_{2}\left(1-\mathsf{L}(A_{K},2x_{2})\right)>0. Therefore fκdcmf_{\kappa}\in\mathcal{F}_{d}^{cm}, which implies the optimality of V1/dLdV^{1/d}L_{d} in d(V)\mathcal{L}_{d}(V) for all fixed V>0V>0 and the fact that Efκ[L]E_{f}^{\kappa}[L] tends to 0 as all the points are sent to infinity, i.e. EfκE_{f}^{\kappa} does not have a minimizer in d\mathcal{L}_{d}.

Acknowledgement: I am grateful for the support of the WWTF research project ”Variational Modeling of Carbon Nanostructures” (no. MA14-009) and the (partial) financial support from the Austrian Science Fund (FWF) project F65. I also thank Mircea Petrache for our discussions about the crystallization at fixed density as a consequence of Cohn-Kumar conjecture stated in Remark 2.10.

References

  • [1] G. Allaire. Shape optimization by the homogenization method, volume 146 of Applied Mathematical Sciences. Springer New York, 2002.
  • [2] M. Baake and U. Grimm. Aperiodic Order, Volume 1. A Mathematical invitation. Cambridge University Press, 2013.
  • [3] S. Bernstein. Sur les fonctions absolument monotones. Acta Math., 52:1–66, 1929.
  • [4] L. Bétermin. Two-dimensional Theta Functions and Crystallization among Bravais Lattices. SIAM J. Math. Anal., 48(5):3236–3269, 2016.
  • [5] L. Bétermin. Local variational study of 2d lattice energies and application to Lennard-Jones type interactions. Nonlinearity, 31(9):3973–4005, 2018.
  • [6] L. Bétermin. Local optimality of cubic lattices for interaction energies. Anal. Math. Phys., 9(1):403–426, 2019.
  • [7] L. Bétermin. Minimizing lattice structures for Morse potential energy in two and three dimensions. J. Math. Phys., 60(10):102901, 2019.
  • [8] L. Bétermin. Minimal Soft Lattice Theta Functions. Constr. Approx., 52(1):115–138, 2020.
  • [9] L. Bétermin and M. Faulhuber. Maximal theta functions - Universal optimality of the hexagonal lattice for Madelung-like lattice energies. Preprint. arXiv:2007.15977, 2020.
  • [10] L. Bétermin, M. Faulhuber, and H. Knüpfer. On the optimality of the rock-salt structure among lattices and change distributions. Preprint. arXiv:2004.04553, 2020.
  • [11] L. Bétermin and H. Knüpfer. Optimal lattice configurations for interacting spatially extended particles. Lett. Math. Phys., 108(10):2213–2228, 2018.
  • [12] L. Bétermin, L. De Luca, and M. Petrache. Crystallization to the square lattice for a two-body potential. Preprint. arXiv:1907:06105, 2019.
  • [13] L. Bétermin and M. Petrache. Dimension reduction techniques for the minimization of theta functions on lattices. J. Math. Phys., 58:071902, 2017.
  • [14] L. Bétermin and M. Petrache. Optimal and non-optimal lattices for non-completely monotone interaction potentials. Anal. Math. Phys., 9(4):2033–2073, 2019.
  • [15] L. Bétermin and P. Zhang. Minimization of energy per particle among Bravais lattices in 2\mathbb{R}^{2}: Lennard-Jones and Thomas-Fermi cases. Commun. Contemp. Math., 17(6):1450049, 2015.
  • [16] X. Blanc and M. Lewin. The Crystallization Conjecture: A Review. EMS Surv. in Math. Sci., 2:255–306, 2015.
  • [17] J. M. Borwein, M. L. McPhedran, R. C. Wan, and I. J. Zucker. Lattice sums: then and now. volume 150 of Encyclopedia of Mathematics, 2013.
  • [18] M. Buchanan. Quantum crystals. Nature Physics, 13:925, 2017.
  • [19] J.W.S. Cassels. On a Problem of Rankin about the Epstein Zeta-Function. Proceedings of the Glasgow Mathematical Association, 4:73–80, 7 1959.
  • [20] H. Cohen. Number Theory II: Analytic and Modern Methods. Springer, 2007.
  • [21] H. Cohn and A. Kumar. Universally optimal distribution of points on spheres. J. Amer. Math. Soc., 20(1):99–148, 2007.
  • [22] H. Cohn, A. Kumar, S. D. Miller, D. Radchenko, and M. Viazovska. The sphere packing problem in dimension 24. Ann. of Math., 185(3):1017–1033, 2017.
  • [23] H. Cohn, A. Kumar, S. D. Miller, D. Radchenko, and M. Viazovska. Universal optimality of the E8E_{8} and Leech lattices and interpolation formulas. Preprint. arXiv:1902:05438, 2019.
  • [24] J. H. Conway and N. J. A. Sloane. Sphere Packings, Lattices and Groups, volume 290. Springer, 1999.
  • [25] R. Coulangeon and A. Schürmann. Energy Minimization, Periodic Sets and Spherical Designs. Int. Math. Res. Not. IMRN, pages 829–848, 2012.
  • [26] G. Zhang D. Chen and S. Torquato. Inverse Design of Colloidal Crystals via Optimized Patchy Interactions. The journal of Physical Chemistry B, 122:8462–8468, 2018.
  • [27] P. H. Diananda. Notes on Two Lemmas concerning the Epstein Zeta-Function. Proceedings of the Glasgow Mathematical Association, 6:202–204, 7 1964.
  • [28] V. Ennola. A Lemma about the Epstein Zeta-Function. Proceedings of The Glasgow Mathematical Association, 6:198–201, 1964.
  • [29] M. Friedrich and L. Kreutz. Crystallization in the hexagonal lattice for ionic dimers. Math. Models Methods Appl. Sci., 29(10):1853–1900, 2019.
  • [30] M. Friedrich and L. Kreutz. Finite crystallization and Wulff shape emergence for ionic compounds in the square lattice. Preprint. arXiv:arXiv:1903:00331, 2019.
  • [31] S. Grivopoulos. No crystallization to honeycomb or Kagomé in free space. J. Phys. A: Math. Theor., 42(11):1–10, 2009.
  • [32] S. Hyun and S. Torquato. Optimal and Manufacturable Two-dimensional, Kagomé-like Cellular Solids. Journal of Materials Research, 17(1):137–144, 2002.
  • [33] I. G. Kaplan. Intermolecular Interactions : Physical Picture, Computational Methods, Model Potentials. John Wiley and Sons Ltd, 2006.
  • [34] M. Lewin, E. H. Lieb, and R. Seiringer. Floating Wigner crystal with no boundary charge fluctuations. Phys. Rev. B, 100(3):035127, 2019.
  • [35] S. Luo, X. Ren, and J. Wei. Non-hexagonal lattices from a two species interacting system. SIAM J. Math. Anal., 52(2):1903–1942, 2020. Preprint. arXiv:1902.09611.
  • [36] S. Luo and J. Wei. On minima of sum of theta functions and Mueller-Ho Conjecture. Preprint. arXiv:2004.13882, 2020.
  • [37] M. Mekata. Kagome: The story of the basketweave lattice. Physics Today, 56(2):12–13, 2003.
  • [38] H. L. Montgomery. Minimal Theta Functions. Glasg. Math. J., 30(1):75–85, 1988.
  • [39] M. Petrache and S. Serfaty. Crystallization for Coulomb and Riesz Interactions as a Consequence of the Cohn-Kumar Conjecture. Proceedings of the American Mathematical Society, 148:3047–3057, 2020.
  • [40] C. Poole. Encyclopedic Dictionary of Condensed Matter Physics. Elsevier, 1st edition edition, 2004.
  • [41] S. Bae Q. Chen and S. Granick. Directed self-assembly of a colloidal kagome lattice. Nature, 469:381–384, 2011.
  • [42] R. A. Rankin. A Minimum Problem for the Epstein Zeta-Function. Proceedings of The Glasgow Mathematical Association, 1:149–158, 1953.
  • [43] P. Sarnak and A. Strömbergsson. Minima of Epstein’s Zeta Function and Heights of Flat Tori. Invent. Math., 165:115–151, 2006.
  • [44] A. Terras. Harmonic Analysis on Symmetric Spaces and Applications II. Springer New York, 1988.
  • [45] R. J .D. Tilley. Understanding Solids: The Science of Materials. John Wiley &\& Sons, 2004.
  • [46] M.P. Tosi. Cohesion if ionic solids in the Born model. Solid State Physics, 16:1–120, 1964.
  • [47] M. Viazovska. The sphere packing problem in dimension 8. Ann. of Math., 185(3):991–1015, 2017.