Local Number Fluctuations in Hyperuniform and Nonhyperuniform Systems: Higher-Order Moments and Distribution Functions

Salvatore Torquato∗ Department of , Princeton University, Princeton, NJ 08544, USA and Department of , Princeton Institute for the Science and Technology of Materials, and Program in Applied and Computational Mathematics, Princeton University, Princeton, New Jersey 08544, USA

Jaeuk Kim† and Michael A. Klatt‡ Department of Physics, Princeton University, Princeton, NJ 08544, USA (Dated: December 9, 2020) The local number variance σ2(R) associated with a spherical sampling window of radius R enables d a classification of many-particle systems in d-dimensional Euclidean space R according to the degree to which large-scale density fluctuations are suppressed, resulting in a demarcation between hyperuniform and nonhyperuniform phyla. To more completely characterize density fluctuations, we carry out an extensive study of higher-order moments or cumulants, including the skewness γ1(R), excess kurtosis γ2(R) and the corresponding probability distribution function P [N(R)] of a large family of models across the first three space dimensions, including both hyperuniform and nonhyperuniform systems with varying degrees of short- and long-range order. To carry out this comprehensive program, we derive new theoretical results that apply to general point processes and conduct high-precision numerical studies. Specifically, we derive explicit closed-form integral expressions for γ1(R) and γ2(R) that encode structural information up to three-body and four-body correlation functions, respectively. We also derive rigorous bounds on γ1(R), γ2(R) and P [N(R)] for general point processes and corresponding exact results for general packings of identical spheres. High-quality simulation data for γ1(R), γ2(R) and P [N(R)] are generated for each model. We also ascertain the proximity of P [N(R)] to the normal distribution via a novel Gaussian “distance” metric l2(R). Among all models, the convergence to a central limit theorem (CLT) is generally fastest for the −(d+1)/2 −(d+1) disordered hyperuniform processes such that γ1(R) ∼ l2(R) ∼ R and γ2(R) ∼ R for large R. The convergence to a CLT is slower for standard nonhyperuniform models and slowest for the “antihyperuniform” model studied here. We prove that one-dimensional hyperuniform systems of class I or any d-dimensional lattice cannot obey a CLT. Remarkably, we discovered that the gamma distribution provides a good approximation to P [N(R)] for all models that obey a CLT across all dimensions for intermediate to large values of R, enabling us to estimate the large-R scalings of γ1(R), γ2(R) and l2(R). For any d-dimensional model that “decorrelates” or “correlates” with d, we elucidate why P [N(R)] increasingly moves toward or away from Gaussian-like behavior, respectively. Our work elucidates the fundamental importance of higher-order structural information to fully characterize density fluctuations in many-body systems across length scales and dimensions, and thus has broad implications for condensed matter physics, engineering, mathematics and biology.

I. INTRODUCTION radius R (see Fig.1) and volume

d/2 d The quantification of density fluctuations in many- π R v1(R) = . (1) particle systems in d-dimensional Euclidean space Rd is Γ(1 + d/2) of great fundamental and practical importance in many fields across the physical, mathematical and biological The local number variance σ2(R) ≡ hN 2(R)i − hN(R)i2 sciences [1–16]. It is well known that long-wavelength is a useful measure of number fluctuations, where the density fluctuations of disordered as well as ordered sys- first moment hN(R)i = ρv1(R) is the average number of tems contain crucial information about the structure as points within a d-dimensional spherical (sampling) win- well as equilibrium and nonequilibrium physical prop- dow of radius R and angular brackets denote an ensem- arXiv:2012.02358v2 [cond-mat.stat-mech] 7 Dec 2020 erties of the systems [1,5, 13, 16]. Clearly, density ble average. The local number variance is exactly deter- fluctuations that occur on some arbitrary local length mined by pair statistics, and can be given either in terms scale [2,3,8,9, 11–13, 15–18] provide considerably more of the pair correlation function g2(r) in direct space or information about the system than those in the long- the structure factor S(k) in reciprocal space [13]: wavelength limit. ï Z ò 2 Consider a statistically homogeneous (translationally σ (R) = ρv1(R) 1 + ρ h(r)α2(r; R) dr , (2) d invariant) point process in d-dimensional Euclidean space R d Z R at number density ρ and sampling for the number of h 1 i = ρv1(R) d S(k)˜α2(k; R)dk , (3) points N(R) within a d-dimensional spherical window of (2π) d R 2 where h(r) ≡ g2(r) − 1 is the total correlation func- tion, α2(r; R) is the intersection volume of two spher- ical windows of radius R, scaled by v1(R), whose cen- ters are separated by the distance r, andα ˜2(k; R) is its Fourier transform. The large-scale behavior of the number variance σ2(R) is central to the hyperuniformity concept, which is attracting attention across many fields [13, 16, 19–23]. Specifically, a hyperuniform point pro- cess is one in which σ2(R) grows slower than the window volume, i.e., Rd, for large R and hence is characterized by large-scale density fluctuations that are anomalously sup- pressed compared to those of typical disordered systems. The hyperuniformity concept generalizes the traditional notion of long-range order of crystals and quasicrystals to also encompass certain exotic disordered states of matter [13, 16]. Disordered hyperuniform systems are diamet- rically opposite to systems at thermal critical points in which the local variance diverges faster than Rd in the limit R → ∞. Any system with such divergent behavior in the local variance has been called anti-hyperuniform [16] (see also Sec.II for additional details). While the local number variance contains useful infor- FIG. 1. Schematic showing a spherical window Ω of radius R used to sample fluctuations in the number of points in a mation, one would like to more completely characterize two-dimensional point process. the fluctuations by ascertaining higher-order moments, i.e., hN k(R)i, where k ≥ 3, as well as the corresponding discrete probability distribution P [N(R)] associated with In this paper, we investigate the skewness γ1(R) (re- finding exactly N(R) particles within a d-dimensional lated to the first three moments), excess kurtosis γ2(R) spherical window of radius R. The mth moment of the (related to the first four moments), and the number distribution is given by distribution function P [N(R)] for general homogeneous point processes in Rd as well as a wide class of models X hN m(R)i = N m(R)P [N(R)]. (4) across dimensions. Specifically, we derive explicit closed- form integral expressions for γ (R) in terms of the num- N(R)=0 1 ber density ρ, the pair correlation function g2 and the three-body correlation function g3 (defined in Sec.II). Due to the fact that the random variable N(R) is discrete Similarly, we derive corresponding formulas for the ex- and cannot take on negative values, the probability dis- cess kurtosis γ (R), which now depends additionally on tribution P [N(R)], for finite R, can never exactly attain 2 the four-body correlation function g4. These integral re- the normal distribution, which is given by lations also involve geometrical information about the spherical windows via the intersection volumes of up to 1 ï(N(R) − hN(R)i)2 ò three and four spheres in the cases of the skewness and ex- P [N(R)] = √ exp . (5) 2πσ(R) 2σ2(R) cess kurtosis, respectively. Thus, the skewness and excess kurtosis encode up to three-body and four-body informa- For example, for a statistically homogeneous Poisson tion about spatial correlations and window geometries, point process in d at number density ρ, respectively. We also derive some exact elementary re- R sults for the skewness, excess kurtosis and number distri- bution that apply to general packings of identical spheres. N(R) [ρv1(R)] P [N(R)] = exp[−ρv1(R)], (6) Via high-precision computer simulation studies, we ac- N(R)! curately determine the number variance, skewness, ex- cess kurtosis and the number distribution for up to eight which deviates significantly from the normal distribution different models of statistically homogeneous point pro- for sufficiently small R. This is one of the rare cases cesses across the first three space dimensions and wide in which a closed-form analytic formula for P [N(R)] is range of window radii R. These models include both known across dimensions for nontrivial point processes. nonhyperuniform and hyperuniform systems with vary- It is only when R tends to infinity that the Poisson dis- ing degrees of short- and long-range order. Among the tribution becomes a normal distribution, i.e., it follows a nonhyperuniform point processes, we characterize fluc- central limit theorem (CLT); see Ref. [24] and references tuations of Poisson, random sequential addition (RSA) therein. The reader is referred to Refs. [25–28] for proofs packings, equilibrium hard spheres, Poisson cluster and of CLT for other point processes. anti-hyperuniform point processes. Among the hyperuni- 3 form point processes, we characterize hypercubic lattices, sults. randomly perturbed lattices, and stealthy disordered hy- An important fundamental question is what is the ef- peruniform systems. We show that our simulation results fect of increasing the space dimension on number fluctua- for γ1(R), γ2(R) and P [N(R)] for all models are in excel- tions for any particular model? To answer this question, lent agreement with the aforementioned rigorous bounds we recall the so-called decorrelation principle [29], which and exact results for the applicable ranges of R. For all roughly states that for any disordered point process, un- disordered hyperuniform models, our explicit general for- constrained correlations that exist in low dimensions van- mulas of these quantities in terms of n-body information ish as d tends to infinity, and all higher-order correlation enable us to infer the existence of “hidden” order that functions gn for n ≥ 3 may be expressed in terms of the manifests itself for the first time at the three-body level number density ρ and pair correlation function g2. The or higher. decorrelation principle was employed to justify the con- For each model considered in this paper, we are in- jecture that the densest sphere packings in sufficiently terested in ascertaining how large R must be such that high dimensions are disordered (as opposed to ordered P [N(R)] is well-approximated by the Gaussian (normal) in low dimensions) [29, 30]. Importantly, decorrelation distribution. We have found that such “distance” metrics in pair statistics has been shown to manifest itself in low proposed previously are not adequate for assessing the dimensions in the case of disordered sphere packings [31– diverse set of models that we consider here across dimen- 33] as well as other disordered systems with strongly re- sions. We quantify this proximity to the normal distri- pulsively interacting particles [34, 35]. Since the number bution for any model by introducing a certain Gaussian distribution function P [N(R)] generally involves certain integrals over all of the n-body correlation functions, the “distance” metric l2(R), defined in Sec.VII. This dis- tance metric enables us to accurately determine when the decorrelation principle implies that P [N(R)] increasingly distribution function P [N(R)] for a particular model is becomes Gaussian-like as the space dimension increases tending to a CLT. Because the distributions for all mod- for any model that decorrelates with d. Similarly, for els (except the lattices) across dimensions are unimodal, models the correlate with d, P [N(R)] increasingly de- the tendency to a CLT corresponds to the skewness and viates from the normal distribution as d increases. We excess kurtosis simultaneously tending to zero. We have confirm such behaviors for all models that obey a CLT found that almost all of the considered models across di- (Sec.VIID). mensions obey a CLT. The convergence to a CLT is slow- In Sec. II, we provide basic definitions and necessary est for the anti-hyperuniform point process, followed by background material. In Sec. III, we derive explicit the Poisson cluster process, and the Poisson process. The closed-form integral expressions for γ1(R) and γ2(R) as nonhyperuniform RSA and equilibrium packings tend to well as rigorous lower bounds on both of these quantities. a CLT at the same rate as a Poisson point process but We also obtain some general exact results for the first few with smaller coefficients of proportionality. Among all cumulants and distribution functions for sphere packings, models, the convergence to a CLT is generally fastest for whether disordered or not. SectionIV describes the large the disordered hyperuniform processes. The only models variety of nonhyperuniform and hyperuniform models in considered that do not achieve a CLT are the hypercubic one, two and three dimensions that we study in this pa- lattices for any d and 1D hyperuniform systems of class per. In Sec.V, we discuss our proposed Gaussian dis- I. The reader is referred to Sec.VII for details. tance metric. SectionVI describes the simulation proce- dure that we employ to sample the first four cumulants We have examined a variety of well-known closed- and number distributions. In Sec.VII, we present our form probability distributions P [N(R)] to ascertain those results. We make concluding remarks in Sec.VIII. which best approximate the actual distributions for fi- nite R for all of our models. Interestingly, the gamma distribution provides a good approximation to the num- ber distribution P [N(R)] for all models that obey a CLT II. DEFINITIONS AND BACKGROUND across all dimensions for intermediate to large values of R (Sec.VIIC). It is noteworthy that the approxima- A stochastic point process in Rd is defined as a map- tion of P [N(R)] by a gamma distribution enables us ping from a probability space to configurations of points d to estimate the large-R scalings of γ1(R), γ2(R) and r1, r2, r3 ... in d-dimensional Euclidean space R ; see Ref. l2(R) for all models across dimensions that obey a CLT. [36] for mathematical details. Let X denote the set of Among all models, the convergence to a CLT is generally configurations such that each configuration x ∈ X is a fastest for the disordered hyperuniform processes such subset of Rd that satisfies two regularity conditions: (i) −(d+1)/2 −(d+1) that γ1(R) ∼ l2(R) ∼ R and γ2(R) ∼ R there are no multiple points (ri 6= rj if i 6= j) and (ii) each for large R. For standard nonhyperuniform models, con- bounded subset of Rd must contain only a finite num- vergence to a CLT is slower such that γ1(R) ∼ l2(R) ∼ ber of points of x (i.e.,, x is “locally finite”). The point −d/2 −d R and γ2(R) ∼ R . Finally, convergence to a process is statistically is characterized by the generic n- n n CLT is slowest for the antihyperuniform model such that particle probability density function ρn(r ), where r is −1/2 −1 γ1(R) ∼ l2(R) ∼ R and γ2(R) ∼ R . These predic- a shorthand notation for the position vectors of any n n tions are corroborated by corresponding simulation re- points, i.e., r ≡ r1, r2,..., rn [5, 37]. In words, the 4

n n quantity ρn(r )dr is proportional to the probability of and hence h(r) must exhibit negative pair correlations, finding any n particles with configuration rn in volume i.e., anticorrelations, for some values of r [16]. By con- n element dr ≡ dr1dr2 ··· drn, i.e., it is the probability trast, an anti-hyperuniform point process is one in which measure. For any subvolume Ω ∈ Rd, the following nor- S(k) tends to +∞ in the limit |k| → 0 [16]. malization (average) condition involving the fluctuating A lattice Λ in Rd is a subgroup consisting of the integer number of particles within this subvolume, NΩ, immedi- linear combinations of independent vectors that span Rd ately follows and thus represents a special subset of point processes. d ≠ N ! ∑ Z Z Z In a lattice Λ, the space R can be geometrically divided Ω = ··· ρ (rn)drn, (7) (N − n)! n into identical regions F called fundamental cells, each of Ω Ω Ω Ω Ω which contains just one point specified by the lattice vec- Note that this random setting is quite general; it incor- tor [38, 39]. In the physical sciences, a lattice is equiva- porates cases in which the locations of the points are lent to a Bravais lattice. Unless otherwise stated, we will deterministically known, such as a lattice. use the term lattice. A periodic point process is a more n For statistically homogeneous media, ρn(r ) is transla- general notion than a lattice because it is is obtained by tionally invariant and hence depends only on the relative placing a fixed configuration of N points (where N ≥ 1), displacements, say with respect to r1: called the basis, within one fundamental cell of a lattice n Λ, which is then periodically replicated. Thus, the point ρn(r ) = ρn(r12, r13,..., r1n), (8) process is still periodic under translations by Λ, but the N points can occur anywhere in the chosen fundamental where rij = rj − ri. In particular, the one-particle func- cell. Any lattice or periodic point configuration can be tion ρ1 is just equal to the constant number density of particles ρ. For statistically homogeneous point patterns, made statistically homogeneous by uniform translations it is convenient to define the so-called n-particle correla- of the pattern within the fundamental cell. tion function We call a packing in Rd a collection of nonoverlapping particles [40]. The centroids of the particles constitute a n n ρn(r ) special point process in which no two particles can closer gn(r ) = . (9) ρn than some minimal distance. In this paper, we consider packings of identical spheres of diameter D. The packing In systems without long-range order and in which the fraction φ = ρv (D/2) is the fraction of space covered by particles are mutually far from one another, ρ (rn) → ρn 1 n the spheres, where v (R) is the volume of a sphere of and we have from (9) that g (rn) → 1. Thus, the devi- 1 n radius R given by (1). Any periodic point configuration ation of g from unity provides a measure of the degree n with a finite basis can be regarded to be a packing since of spatial correlation between the particles, with unity there is a minimal pair distance. corresponding to no spatial correlation. The important two-particle quantity g2(r12) is usually referred to as the pair correlation function. The total correlation function h(r ) is defined as 12 III. GENERAL LOCAL MOMENT FORMULAS h(r ) = g (r ) − 1, (10) AND PROBABILITY DISTRIBUTION FOR 12 2 12 HOMOGENEOUS POINT PROCESSES and thus is a function that is zero when there are no spa- tial correlations in the system. Observe that the struc- We consider the determination of local moment for- ture factor S(k) is related to the Fourier transform of mulas using the formalism of Torquato and Stillinger h(r), denoted by h˜(k), via the expression [13] that was used to obtain formulas for the local ˜ number variance. Here we immediately begin with S(k) = 1 + ρh(k), (11) a d-dimensional spherical window of radius R in d- d where dimensional Euclidean space R with window indicator function Z h˜(k) = h(r)e−ik·r dr. (12) d ® R 1, |r − x0| ≤ R, w(r − x0; R) = (15) A hyperuniform point process is one in which single- 0, |r − x0| > R scattering events at infinite wavelength vanishes, i.e., d lim S(k) = 0. (13) where x is some arbitrary position vector in R and x0 |k|→0 is the position vector of the window center. The number of points N(R; x0) within the window at position x0 is This implies that a hyperuniform system obeys the fol- given by lowing sum rule: in direct space Z X ρ h(r) dr = −1, (14) N(R; x0) = w(ri − x0; R), (16) d R i=1 5 which must be a finite number. We will subsequently use Following the same procedure used in Ref. [13] to ob- the fact that tain an explicit formula for the second moment, we obtain from (18) that the mth moment hN m(R)i for a homoge- Z n int n Y neous process is given by integrals involving the finite set vn (r ; R) = w(ri − x0; R)dx0 (17) d of correlation functions g , g , . . . g weighted with the R 2 3 m i=1 int int int set of intersection volumes v2 , v3 , . . . , vm . For exam- is the intersection volume of n spheres of radius R cen- ple, expanding the sums in (18) into one-body, two-body, tered at positions r1, r2, ··· , rn. ... and m-body terms in a manner analogous to the one given in Ref. [13], third and fourth moments are explic- itly given by A. Moments Z 3 2 3 int hN (R)i = ρv1(R) + 3ρ g2(r )v2 (r; R)dr The mth moment associated with the random variable d R N(R) is given by the following ensemble average: Z Z 3 3 int 3 + ρ g3(r )v3 (r ; R)dr2dr3, (19) d d * R R m X X X hN (R)i = ··· w(ri − x0; R)w(ri − x0; R) 1 2 and i1=1 i2=1 im=1 + Z · · · × w(r − x ; R) . (18) 4 2 int im 0 hN (R)i = ρv1(R) + 7ρ g2(r)v2 (r; R)dr d Z Z R 3 3 int 3 Here we have implicitly assumed homogeneity, which ren- +6ρ g3(r )v3 (r ; R)dr2dr3 d d ders the mth moment independent of the position of the Z R Z R Z window x . Under the ergodic assumption, the ensemble 4 4 int 4 0 +ρ g4(r )v4 (r ; R)dr2dr3dr4, (20) d d d average indicated on the left-hand side of relation (18) is R R R equivalent to averaging by uniformly window sampling a int single realization over the infinite space. where vn is given by (17).

We are generally interested in the mth order cumulant Cm(R), which is directly related to the mth central moment h[N(R) − hN(R)i]mi. For example, the first several cumulants are given by

2 C2(R) ≡ σ (R) Z 2 int = ρv1(R) + ρ h(r)v2 (r; R)dr d R Z 2 int = ρv1(R)[1 − ρv1(R)] + ρ g2(r)v2 (r; R)dr, (21) d R

3 C3(R) ≡ h[N(R) − hN(R)i] i = hN 3(R)i − 3hN 2(R)ihN(R)i + 2hN(R)i3 Z Z Z 2 int 3 3 int 3 = ρv1(R) + 3ρ h(r)v2 (r; R)dr + ρ [g3(r ) − 3g2(r12) + 2]v3 (r ; R)dr2dr3, (22) d d d R R R and

4 4 C4(R) ≡ h[N(R) − hN(R)i] i − 3σ (R) = hN 4(R)i − 4hN 3(R)ihN(R)i − 3hN 2(R)i2 + 12hN 2(R)ihN(R)i2 − 6hN(R)i4 Z Z Z 2 int 3 3 int 3 = ρv1(R) + 7ρ h(r)v2 (r; R)dr + 6ρ [g3(r ) − 3g2(r12) + 2]v3 (r ; R)dr2dr3 d d d R R R Z Z Z ï Z ò2 4 4 3 int 4 2 int + ρ [g4(r ) − 4g3(r ) + 12g2(r12) − 6]v4 (r ; R)dr2dr3dr4 − 3 ρ g2(r)v2 (r; R)dr (23), d d d d R R R R 6 where we have used the following identities: Z int int 3 v1(R)v2 (r12; R) = v3 (r ; R)dr3, (24) d ZR Z 3 int 3 v1(R) = v3 (r ; R)dr2dr3, (25) d d ZR R int 3 int 4 v1(R)v3 (r ; R) = v4 (r ; R)dr4, (26) d ZR Z 2 int int 4 v1(R)v2 (r12; R) = v4 (r ; R)dr3dr4, (27) d d ZR ZR Z 4 int 4 v1(R) = v4 (r ; R)dr2dr3dr4. (28) d d d R R R

We now derive a lower bound on C3(R) in terms of the first and second cumulants for general point processes. 3 int 3 This easily follows from the fact that the product g3(r )v3 (r ; R) in the last integral of (22) is nonnegative for all positions and so dropping this integral yields the following lower bound: 2  C3(R) ≥ ρv1(R)[1 − ρv1(R)][1 − 2ρv1(R)] + 3[1 − ρv1(R)] σ (R) − ρv1[1 − ρv1(R)] . (29)

4 int 4 Similarly, dropping the positive integral in (23) involving the product g4(r )v4 (r ; R) gives the following lower bound on C4(R) in terms of the first, second and third cumulants: 2 2 C4(R) ≥ ρv1(R)[1 − ρv1(R)][2 − ρv1(R)][3 − ρv1(R)] − σ (R) 11 − 18ρv1(R) + 6[ρv1(R)] 4 −3σ + 2C3(R)[3 − 2ρv1(R)]. (30) These bounds are relatively tight for sufficiently small values of R and can be exact (sharp) for such R for packings, as discussed in Sec.IIIC.

The third- and fourth-order cumulants are directly re- In the case of a homogeneous Poisson point process, lated to the skewness and excess kurtosis, respectively. gn = 1 for all n. Hence, from the expressions above, The skewness is often defined as it immediately follows that C2(R) = C3(R) = C4(R) = ρv1(R), which are the expected well-known results for C3(R) γ1(R) ≡ , (31) this point process. It follows that the corresponding σ3(R) skewness and excess kurtosis are exactly given by which, qualitatively speaking, is a measure of the asym- 1 metry of the probability distribution. The excess kurto- γ (R) = ∝ R−d/2 (34) 1 1/2 sis, which is a measure of the heaviness of the “tails” of [ρv1(R)] the probability distribution, is defined to be and C4(R) γ2(R) ≡ 4 . (32) 1 −d σ (R) γ2(R) = ∝ R , (35) ρv1(R) For a general random variable, Pearson derived a lower bound on the excess kurtosis in terms of the skewness respectively. Observe that both γ1(R) and γ2(R) tend to [41]: zero in the limit R → ∞, which is consistent with the fact that the Poisson point process obeys a CLT. 2 γ2(R) ≥ γ1 (R) − 2, (33)

We also apply this lower bound to validate our numerical results for all of our model point processes. Such details B. Probability Distribution Function are reported in the SM. Both γ1 and γ2 are identically zero for the normal It is straightforward to show that for an arbitrarily- distribution and hence such lower-order information can shaped window region Ω, the probability distribution herald at what value of R a general point configuration function is given by [2,3] can be approximated by a normal distribution. Note that the higher-order cumulants become increasingly more ∞ X m−M Am(Ω) complicated but can be written as a determinant involv- P [NΩ = M] = (−1) , (36) m M!(m − M)! ing the moments hN (R)i [42]. m=M 7 where packings. The difficulty in ascertaining P [N(R)] exactly ≠ ∑ for nontrivial models can be appreciated by appealing to NΩ! Am(Ω) ≡ the ghost random sequential addition (RSA) packing pro- (NΩ − m)! Ω cess [29], for which the gm are known exactly for any m. Z Z Z m m While the evaluation of the integral (39) for ghost RSA =ρ ··· gm(r )dr1dr2 . . . drm, (37) can be carried out exactly for very small m, its exact de- Ω Ω Ω termination becomes impossible for general values of m. is exactly the same as the average given in (7) under the This points to the importance of devising accurate nu- assumption of homogeneity. Thus, we see that the aver- merical methods to determine the number distributions age Am(Ω) is the nontrivial and common contribution to of packings. For example, P [N(R)] has be determined in P [NΩ] for any specific value of NΩ. The only differences simulations for equilibrium hard spheres and MRJ sphere in P [NΩ] for different values of NΩ are the combinatoric packings in Ref. [15]. factors multiplying the coefficients Am(Ω) in the series It has already been established rigorously that trun- (36). Because these coefficients are intrinsically positive, cations of the alternating series (38) for the special case relation (36) is an alternating series. P [N(R) = 0] = EV (R) at an even and odd number of When Ω is a spherical window of radius R, it simply terms yield successive upper and lower bounds on the follows that probability distribution P [N(R) = 0], respectively. Such ∞ bounds are consequences of an inclusion-exclusion princi- X Am(R) P [N(R) = M] = (−1)m−M , (38) ple associated with this alternating series. Here we make M!(m − M)! the simple observation that the same inclusion-exclusion m=M principle applies for any value of N(R). The first several where of such bounds are given by ≠ N(R)! ∑ A (R) ≡ AN (R) m (N(R) − m)! P [N(R)] ≤ (41) v1(R) N(R)! Z Z Z m m AN (R) AN+1(R) =ρ ··· gm(r ) P [N(R)] ≥ − (42) d d d R R R N(R)! N(R)! × vint(rm; R)dr dr . . . dr . (39) A (R) A (R) A (R) m 2 3 m P [N(R)] ≤ N − N+1 + N+2 (43) N(R)! N(R)! 2!N(R)! For a fixed value of M ∈ , P [N(R) < M] as a func- N0 A (R) A (R) A (R) tion of R is a complementary cumulative distribution P [N(R)] ≥ N − N+1 + N+2 function, which is associated with the “void” probabil- N(R)! N(R)! 2!N(R)! ity density function H (R; M)[9], where H (R; M)dR A (R) V V − N+3 , (44) is the probability that the distance to the Mth near- 3!N(R)! est neighbor from an arbitrary point in space is between

R and R + dR. For example, in the case M = 1, where AN (R) is a shorthand for AN(R)(R). These P [N(R) < 1] = P [N(R) = 0] is the well-known void bounds become increasingly sharper as more terms are exclusion probability function EV (R), which is associ- included. Moreover, these bounds can be sharp (exact) ated with the void nearest-neighbor probability density for sufficiently small R for sphere packings, as discussed function HV (R)[37, 39, 43]. In the case of a homoge- in Sec.IIIC. We utilize these bounds to validate our neous Poisson point process, we can immediately recover simulations results for all models across dimensions in the exact result (6) for the number distribution from re- the SM. lations (38) and (39) using the fact that gn = 1 for all n and the identity Z Z Z C. Elementary Results for Packings m int m v1(R) = ··· vm (r ; R)dr2dr3 . . . drm . d d d R R R (40) Here we obtain some general exact results for the first Exact results for P [N(R)] for non-Poissonian point few cumulants and distribution functions for packings of processes are rare. One exception is the one-dimensional identical spheres of diameter D, whether disordered or model of equilibrium hard rods for which P [N(R)] is not. Results that apply to lattice packings are also de- known exactly [9]. In the case of disordered equilib- rived. rium hard-disk (d = 2) and hard-sphere (d = 3) pack- Because no two spheres can overlap when their cen- ings, accurate but approximate expressions for the ex- ters are separated by a distance less than or equal to D, clusion probability EV (R) ≡ P [N(R) = 0] are available the cumulants can be written explicitly for d ≥ 1 and [37, 43]. Thus, in principle, one can extract from such R ≤ D/2, since such a window region can accommodate formulas approximations for Am(R) to get corresponding at most a single sphere. For example, this means that approximations for P [N(R)] for any N(R) ≥ 1 for such integral involving g2 in the last line of Eq. (21) is identi- 8 cally zero, and hence for R ≤ D/2, This means that for R ≤ D/2,

1 − 2ρv (R) γ (R) = 1 (48) 1 [ρv (R)(1 − ρv (R))]1/2 C2(R) = ρv1(R)[1 − ρv1(R)], (45) 1 1 and

2 2 3 3 which was noted by Torquato and Stillinger [13]. More 1 − 7ρv1(R) + 12ρ v1(R) − 6ρ v1(R) m γ2(R) = 2 . (49) generally, noting that hN(R) i = ρv1(R) for all m be- ρv1(R)[1 − ρv1(R)] cause all terms in (19) and (20) involving either g2, g3 and g4, we have from (22) and (23) for d ≥ 1 that for As in the Poisson case, both the skewness γ1(R) and ex- R ≤ D/2, cess kurtosis γ2(R) diverge to +∞ in the limit R → 0. For 0 ≤ R ≤ D/2, while γ1(R) is generally a monotoni- cally decreasing function of R that can be nonnegative, γ (R) is generally a nonmonotonic function but also can C (R) = ρv (R)[1 − ρv (R)][1 − 2ρv (R)], (46) 2 3 1 1 1 be nonnegative. The corresponding results for the dis- tribution function P [N(R)] for R ≤ D/2 follow immedi- ately from (36) and (39), namely,

P [N(R) = 0] = 1 − ρv1(R), (50)

P [N(R) = 1] = ρv1(R), (51) 2 2 3 3 C4(R) = ρv1(R)[1 − 7ρv1(R) + 12ρ v1(R) − 6ρ v1(R)], P [N(R) ≥ 2] = 0. (52) (47)

√ When D/2 < R ≤ D/ 3 and d ≥ 2, the window can accommodate at most two but not three hard spheres. Therefore, following√ the same reasoning as above for R ≤ D/2, we find from (22) and (23) that for d ≥ 2 and 0 ≤ R ≤ D/ 3 Z Z 2 int 3 int 3 3 C3(R) = ρv1(R) + 3ρ h(r)v2 (r; R)dr − 3ρ v1(R) g2(r12)v2 (r; R)dr + 2ρ v1(R) d d R R Z 2 int = ρv1(R)[1 − ρv1(R)][1 − 2ρv1(R)] + 3[1 − ρv1(R)]ρ g2(r)v2 (r; R)dr d R  2  = ρv1(R)[1 − ρv1(R)][1 − 2ρv1(R)] + 3[1 − ρv1(R)] σ (R) − ρv1[1 − ρv1(R)] (53) and Z Z Z 2 int 3 int 3 C4(R) = ρv1(R) + 7ρ h(r)v2 (r; R)dr + 6ρ [−3g2(r12) + 2]v3 (r ; R)dr2dr3 d d d R R R Z Z Z ï Z ò2 4 int 4 2 int + ρ [12g2(r12) − 6]v4 (r ; R)dr2dr3dr4 − 3 ρ g2(r)v2 (r; R)dr d d d d R R R Z R Z 2 2 3 3 2 int 3 int = ρv1(R)[1 − 7ρv1(R) + 12ρ v1(R) − 6ρ v1(R)] + 7ρ g2(r)v2 (r; R)dr − 18ρ v1(R) g2(r)v2 (r; R)dr d d R R Z ï Z ò2 4 2 int 2 int + 12ρ v1(R) 12g2(r)v2 (r; R)dr − 3 ρ g2(r)v2 (r; R)dr d d R R 2 2 3 3 = ρv1(R)[1 − 7ρv1(R) + 12ρ v1(R) − 6ρ v1(R)] Z ï Z ò2 2 2 2 int 2 int + [7 − 18ρv1(R) + 12ρ v1(R)] ρ g2(r)v2 (r; R)dr − 3 ρ g2(r)v2 (r; R)dr d d R R 2 2 3 3 2 2 2 = ρv1(R)[1 − 7ρv1(R) + 12ρ v1(R) − 6ρ v1(R)] + [7 − 18ρv1(R) + 12ρ v1(R)][σ (R) − ρv1(R)(1 − ρv1(R))] 2 2 − 3 [σ (R) − ρv1(R)(1 − ρv1(R))] . (54)

We see that both C3(R) and C4(R) are given purely in terms of the mean hN(R)i = ρv1(R) and number variance σ2(R) for such R. Observe also that formula (53) is identical the lower bound (29) for general point processes. It is seen that for R ≤ D/2, relations (46) and (47) are recovered, as expected. Finally, for d = 1, these formulas actually apply for R ≤ D. When the window can accommodate at most three spheres, the fourth cumulant C4(R) for 0 ≤ R ≤ R∗(d) is exactly 9 given in terms of the first three cumulants, i.e., 2  2 C4(R) = ρv1(R)[1 − ρv1(R)][2 − ρv1(R)][3 − ρv1(R)] − σ (R) 11 − 18ρv1(R) + 6[ρv1(R)] 4 −3σ + 2C3(R)[3 − 2ρv1(R)], (55) √ where R∗(d) > D/ 3 is a threshold that depends on the space dimension d. For example, R∗(1) = 3D/2, R∗(2) = √ p 2D/2, and R∗(3) = 3/8D. Note that formula (55) is identical to the lower bound (30) for√ general point processes. When the spherical window can accommodate at most two spheres, implying that R ≤ D/ 3 for d ≥ 2, three-body and higher-order terms in the series (39) for the probability distribution vanish identically, yielding the following exact result for P [N(R)]: 1 σ2(R) P [N(R) = 0] = [1 − ρv (R)][1 − ρv (R)] + , (56) 2 1 1 2 2 2 P [N(R) = 1] = 2ρv1(R) − [ρv1(R)] − σ (R), (57) σ2(R) ρ P [N(R) = 2] = − v (R)[1 − ρv (R)], (58) 2 2 1 1 P [N(R) ≥ 3] = 0. (59) We see that for such windows, the entire distribution function is completely determined by the first and second cumulants. A nontrivial upper bound on the variance σ2(R) of a packing follows immediately from (57) and the fact that P [N(R) = 1] must be nonnegative, i.e.,

2 σ (R) ≤ ρv1(R)[2 − ρv1(R)]. (60)

For the same reasons, relation (58) yields the following general lower bound on the variance

2 σ (R) ≥ ρv1(R)[1 − ρv1(R)]. (61)

As before, these formulas actually apply for R ≤ D for d = 1. Similarly, when the spherical window can accommodate at most three spheres, implying that 0 ≤ R ≤ R∗(d), four-body and higher-order terms in the series (39) for P [N(R)] vanish identically, yielding the following exact result for P [N(R)]: ρv (R) 1 C (R) P [N(R) = 0] = 1 − ρv (R) − 1 [1 − ρv (R)][5 − ρv (R)] + [2 − ρv (R)]σ2(R) − 3 , (62) 1 6 1 1 2 1 6 1 1 1 P [N(R) = 1] = ρv (R)[2 − ρv (R)][3 − ρv (R)] + [−5 + 3ρv (R)]σ2(R) + C (R), (63) 2 1 1 1 2 1 2 3 ρv (R) 1 1 P [N(R) = 2] = − 1 [1 − ρv (R)][3 − ρv (R)] + [4 − 3ρv (R)]σ2(R) − C (R), (64) 2 1 1 2 1 2 3 ρv (R) σ2(R) C (R) P [N(R) = 3] = 1 [1 − ρv (R)][2 − ρv (R)] − [1 − ρv (R)] + 3 , (65) 6 1 1 2 1 6 P [N(R) ≥ 4] = 0. (66) Thus, for such R, the entire probability distribution function is completely determined by the first three cumulants. The nonnegativities of the probabilities P [N(R) = 0] [Eq. (62)] and P [N(R) = 2] [Eq. (64)] yield upper bounds on C3(R) in terms of the first two cumulants and so the minimum of these two upper bounds are to be chosen. Similarly, the P [N(R) = 1] [Eq. (63)] and P [N(R) = 3] [Eq. (65)] yield lower bounds on C3(R) in terms of the first two cumulants and so the maximum of these two lower bounds are to be chosen. In the SM, we demonstrate that the aforementioned exact results for C3(R), C4(R) and certain P [N(R)] are in excellent agreement with our corresponding simulations data for sphere packings examined in the paper across dimensions.

More generally, for any packing of identical spheres, sions. Therefore, the number distribution P [N(R)] for a a spherical region of radius R can accommodate a max- packing is generally far from a normal distribution for a imum number of spheres, denoted by Nmax(R). This finite-sized window, since it must have compact support maximal number can be determined from tabulations of such that it is zero for N(R) > Nmax(R), i.e., the so-called densest local packings for a finite range of P [N(R)] = 0 for N(R) > N (R). (67) particle numbers in both two [44] and three [45] dimen- max In such instances, the distribution function is determined 10 by a finite set of moments, i.e., the first, second, ..., plots of the variance, skewness and excess kurtosis as Nmaxth moments. Moreover, for any dense packing or a function of R for Z. The highly discrete of the point process in which the nearest neighbor from a parti- number distribution for the integer lattice extends to that cle is narrowly distributed (e.g., “strongly” stealthy sys- for the hypercubic lattice Zd for d ≥ 2, as will see in Sec. tems described below and in Sec. IV B 3), P [N(R)] will VII. be zero for N(R) < Nmin(R), where the cut-off value Nmin(R) grows with R. This situation prevents a strict CLT from applying for finite-sized windows. IV. NONHYPERUNIFORM AND Another important observation is that for point pro- HYPERUNIFORM MODELS cesses in which the “hole” radius R is bounded from above by Rmax, the probability of finding a spheri- We consider eight different models of statistically ho- cal window with radius R > Rmax must be zero, i.e., mogeneous point processes in two and three dimen- P [N(R) = 0] = 0 for R > Rmax, which of course is sions: five nonhyperuniform models, one of which is anti- non-Gaussian behavior. The cut-off value Rmax for a hyperuniform (hyperplanes intersection process or HIP), point process in Rd is its covering radius [39]. Processes and three hyperuniform models. Analogous models are with this bounded-hole property include periodic pack- also examined in one dimension, except for HIP, which is ings with a finite basis [46], quasicrystals [47], as well as not defined in this dimension. The reader is referred to the saturated random sequential addition packing pro- Figs.3 and4 for representative images of configurations cess (see Sec. IV A 3). Disordered stealthy point pro- for each of the models in two dimensions. cesses have bounded holes [46, 48], as discussed in Sec. It is useful to recall scaling relations for hyperuni- IV B 3. form and nonhyperuniform point processes. Consider d We note that for the hypercubic lattice Zd scaled by D any homogeneous point process in R for which the struc- (see Sec. IV B 1 for precise definition), all of the relations ture factor has the following power-law behavior as the derived above for the skewness, excess kurtosis√ and dis- wavenumber tends to zero: tribution function for the situation R ≤ D/ 3 actually √ S(k) ∼ |k|α (|k| → 0). (75) apply as well for the larger range R ≤ D/ 2 when d ≥ 2, where D is the lattice spacing. In fact, in the case of the This scaling implies that the total correlation function scaled integer lattice ρ−1Z at number density ρ, we can h(r) has the corresponding power-law behavior 1/|r|d+α obtain an exact formula for P [N(R)] by invoking the key for large |r| [16]. For hyperuniform systems, the exponent idea of Ref. [13] to yield the exact result for the local α is a positive constant, which implies that there are number variance, namely, the number of points inside a three different scaling regimes (classes) that describe the window of radius R can only take two values, either NR associated large-R of the number variance [13, 16, 49]: or NR + 1, where NR ≡ bρ2Rc and bxc is the floor func-  Rd−1, α > 1 (Class I) tion of a real number x. The probability distribution for  all N(R) is given by σ2(R) ∼ Rd−1 ln R, α = 1 (Class II), (76)  Rd−α, 0 < α < 1 (Class III). P [N(R) < NR] = 0, (68)

P [N(R) = NR] = 1 − {ρ2R}, (69) By contrast, for any nonhyperuniform system, it follows from the asymptotic analysis given in Ref. [16] that P [N(R) = NR + 1] = {ρ2R}, (70) P [N(R) > N + 1] = 0, (71) ® Rd, α = 0 (typical nonhyperuniform) R σ2(R) ∼ Rd−α, α < 0 (anti-hyperuniform). where {x} ≡ x − bxc is the fractional part of a positive number x. Thus, this skewed distribution is highly non- (77) Gaussian with nonexistent left or right tails for almost The scaling for the anti-hyperuniform instance can be ob- all values of N(R), implying values of the skewness and tained using an asymptotic analysis of either the direct- excess kurtosis that are generally far from zero for almost space representation (2) or the Fourier-space representa- R. From the distribution function (71) and relation (4), tion (3) of the number variance, as derived in Ref. [50]. we can immediately obtain the first several cumulants: The typical nonhyperuniform scaling in (77) results from the fact that S(0) is bounded and, indeed, the implied σ2(R) = {ρ2R}(1 − {ρ2R}), (72) constant multiplying Rd is proportional to S(0). 1 − 2{ρ2R} Any nonhyperuniform point process for which S(0) > 1 γ1(R) = , (73) [{ρ2R}(1 − {ρ2R})]1/2 has a large-R asymptotic number variance σ2(R) that is 1 − 6{ρ2R}(1 − {ρ2R}) larger than that for a Poisson point process [S(0) = 1] γ (R) = , (74) 2 {ρ2R}(1 − {ρ2R}) with the same mean hN(R)i. We call such a nonhyper- uniform point process super-Poissonian. Two examples which are all periodic functions with period ρ2R. Rela- of super-Poissonian point processes studied in this work tion (72) for the variance was given in Ref. [13]. The are the Poisson cluster and HIP point processes described reader is referred to the top panel of Fig.2, which shows below. 11

A. Nonhyperuniform Processes 4. Poisson Cluster Process

1. Poisson Point Process The Poisson cluster process is an example of a strongly clustering point process with large density fluctuations A homogeneous Poisson point process in Rd has a on large length scales, i.e., with a large but finite value structure factor S(k) = 1 for all k and hence is non- of S(0), and hence is a nonhyperuniform system that is hyperuniform. At unit mean density (ρ = 1) this process far from being hyperuniform. The construction of the is generated within a hypercubic simulation box of fixed cluster process starts from a homogeneous Poisson point volume V under periodic boundary conditions by a two- process of intensity ρp [24]. Each point of the Poisson step procedure. First, we choose a random number N point process is the center of a cluster of points. The from the Poisson distribution (6) with intensity or mean number of points in each cluster is independent and fol- ρV = V and then place N points in the simulation box lows a Poisson distribution with mean value c. In our uniformly. specific model, the positions of the points relative to the center of the cluster follows an isotropic Gaussian distri- bution with standard deviation r0, which can be regarded 2. Equilibrium Packings to be the characteristic length scale of a single cluster. This model is also known as a (modified) Thomas point We also consider equilibrium packings of identical process, which is an example of a Neyman-Scott pro- sphere (Gibbs hard-sphere processes) across the first cess [36, 57]. In the infinite-volume limit, the pair corre- three space dimensions. For d = 2 and d = 3, we ex- lation function in Rd is exactly given by [57]: amine disordered states that lie along the stable liquid r2 branch [5, 37] as well as disordered states in one dimen- c − 2 g (r) = 1 + e 4r0 . 2 2 d/2 sion, all of which are nonhyperuniform. We generate such ρ(4πr0) equilibrium packings using the well-established Metropo- lis numerical scheme [5, 37]. All configurations that we Thus, the corresponding structure factor for any d is generate are well away from jamming points and hence given by all are nonhyperuniform with bounded S(0) (see Table −k2r2 I). S(k) = 1 + ce 0 . (78)

and hence such processes are nonhyperuniform and 3. Random Sequential Addition Packings super-Poissonian with S(0) = 1 +c. To simulate the pro- cess, which is straightforward, we use periodic boundary The random sequential addition (RSA) process is a conditions and the following parameters across the first time-dependent (nonequilibrium) procedure that gener- three space dimensions: r0 = 1 and unit number density d ates disordered sphere packings in R [32, 51–55]. Start- ρ = ρpc = 1 such that ρp = 0.1 and c = 10. For such ing with an empty but large volume in Rd, the RSA parameters, S(0) = 11 across dimensions (see TableI). process is produced by randomly, irreversibly, and se- quentially placing nonoverlapping spheres into the vol- ume. This procedure is repeated for ever increasing vol- 5. Hyperplanes Intersection Process umes and then an appropriate infinite-volume limit is obtained. In practice, hard spheres are randomly and The hyperplanes intersection process (HIP) is hy- sequentially placed into a large fundamental cell under perfluctuating [16], i.e., its number variances scales periodic boundary conditions and subject to a nonover- faster than the volume of the observation window and lap constraint: If a new sphere does not overlap with any limk→0 S(k) = ∞ [26, 58]. This antihyperuniform and existing spheres, it will be added to the configuration; super-Poissonian point process is defined as the vertices otherwise, the attempt is discarded. One can stop the (i.e., intersections) of a Poisson hyperplane process, that addition process at any time t, obtaining RSA configu- is, of randomly and independently distributed hyper- rations with a range of packing fractions φ(t) up to the planes [36, 59]. In the infinite-volume limit, the pair maximal saturation value φ ≡ φ(∞), which imposes a d s correlation function in R for any d ≥ 2 is exactly given bounded-hole property [46]. For identical spheres, which by [26]: we consider here, φs ≈ 0.74, 0.55, and 0.38 for d = 1, 2, and 3, respectively [33, 51, 53–55]. The pair correlation d−1 Ç å Å ã2 Å ãk X d − 1 ωd−k dωd 1 function g2(r) is known exactly only in one dimension g2(r) = 1 + k , k ωd ωd−1 (sr) [56]. The structure factors at the saturation states across k=1 dimensions have been determined numerically [33, 55]. These results reveal that saturated RSA packings are where s is the specific surface of the hyperplane and ωd nonhyperuniform, even if the values of S(0) are relatively denotes the volume of a d-dimensional sphere of unit ra- small (see TableI). dius. The number density ρ is determined by the specific 12 surface area s of the Poisson hyperplane process (which [62]. The underlying long-range order can be “cloaked”, is the only parameter of the isotropic HIP): under certain conditions, in the sense that it cannot be reconstructed from the pair-correlation function alone. Å ãd ωd−1 d Here we generate the URL model using the hypercubic ρ = ωd s . (79) lattice d and displace each lattice point by a random dωd Z vector that is uniformly distributed in a rescaled funda- According to (77), because α = 1 for any d, the number mental cell of the lattice with aF ≡ [−a/2, a/2)d. The variance has the large-R scaling σ2(R) ∼ R2d−1. Clearly, constant a controls the strength of perturbations. Coun- this process does not exist for d = 1. To simulate this terintuitively, the long-range order suddenly disappears process, we cannot employ periodic boundary conditions; at certain discrete values of a and reemerges for stronger rather, we circumscribe the cubic simulation box by a perturbations, as we will show. Here we cloak the Bragg hypersphere and then generate intersecting hyperplanes peaks of Zd using the special value a = 1. Such cloaked that are Poisson distributed [58]. The orientation of the URLs are hyperuniform such that S(k) ∼ k2 in the limit hyperplanes is uniformly distributed on the unit sphere k → 0 and hence are of class I [see Eq. (76)]. and the distance of the hyperplanes to the center of the simulation box is uniformly distributed between zero and the radius of the circumsphere. The point process at 3. Stealthy Hyperuniform Process unit number density is then simulated by computing all intersections of hyperplanes (within the circumsphere). Stealthy hyperuniform processes are defined by a struc- ture factor that vanishes in a spherical region around the origin, i.e., S(k) = 0 for 0 < |k| ≤ K. Such point pro- B. Hyperuniform Processes cesses are hyperuniform of class I; see Eq. (76). A pow- erful procedure that enables one to generate high-fidelity 1. Hypercubic Lattice stealthy hyperuniform point patterns is the collective- coordinate optimization technique [63–68]. This opti- Interestingly, the problem of determining number fluc- mization methodology involves finding the highly degen- tuations in lattices has deep connections to number the- erate ground states of a class of bounded pair poten- ory, including Gauss’s circle problem [60] and its gener- tials with compact support in Fourier space, which are alizations [16] as well as the Epstein zeta function [61], stealthy and hyperuniform by construction. The control which is directly related to the minimization of the num- parameter χ is a dimensionless measure of the ratio of ber variance [13, 16, 49]. All periodic point patterns in constrained degrees of freedom (i.e., wave vectors con- Rd, including Bravais lattices, are hyperuniform of class I tained within the cut-off wavenumber K) to the total [13, 16, 49]. The hyperuniformity concept enables one to degrees of freedom (approximately dN) in such an opti- rank order lattices and other periodic point patterns ac- mization procedure. A point configuration with a small cording to the degree to which they suppress large-scale value of χ (relatively unconstrained) is disordered, and as density fluctuations as defined by the number variance χ increases, the short-range order increases within a dis- [13, 16, 49]. ordered regime (χ < 1/2 for d = 2 and d = 3) [67]. For For the purposes of this investigation, it is sufficient to d = 1, stealthy hyperuniform states can be disordered consider the higher-order fluctuations of the hypercubic for χ < 1/3 [68]. Here we use the collective-coordinate lattice Zd is defined by procedure to generate “entropically-favored” disordered stealthy point processes by first performing molecular dy- d Z = {(x1, . . . , xd): xi ∈ Z} for d ≥ 1 (80) namics simulations at sufficiently low temperatures and then minimizing the energy to obtain ground states with where Z is the set of integers (...−3, −2, −1, 0, 1, 2, 3 ...) exquisite accuracy [68]. Importantly, stealthy states pos- and x1, . . . , xd denote the components of a lattice vector. sess the bounded-hole property [46, 48] and hence, as dis- cussed in Sec.IIIC, P [N(R) = 0] = 0 for R > Rmax(χ), where Rmax(χ) is the radius of the largest hole in space 2. Uniformly Randomized Lattice dimension d, which depends on the control parameter χ.

It is well known that if the sites of a lattice are stochas- tically displaced by certain finite distances, the scatter- V. GAUSSIAN DISTANCE METRIC ing intensity (structure factor) inherits the Bragg peaks (long-range order) of the original lattice, in addition to a As noted in the Introduction, we are interested in as- diffuse contribution. It has recently been demonstrated certaining how large R must be such that P [N(R)] is that these Bragg peaks can be hidden in the scatter- well approximated by the normal distribution. There are ing pattern for certain independent and identically dis- several candidate “distance” metrics that we have con- tributed perturbations. We have referred to this pro- sidered that could be used to quantify proximity to the tocol as the uniformly randomized lattice (URL) model Gaussian distribution (beyond the skewness and excess 13 kurtosis). One possible distance metric that we consid- G(R), i.e., ered is the Kolmogorov-Smirnov test statistic [69]. While n it is statistically robust, we found it be too insensitive X for our purposes. The Kullback-Leibler divergence (also FG(n) = PG(R)[G(R) = m], (82) called the relative entropy) is a well-known measure of the m=0 difference between probability distributions [70]. How- and FN (n) is the cumulative probability distribution of ever, it is not well-defined for comparing a discrete to a N(R), i.e., continuous distribution; see the SM for details. n A recent study that considered distance metrics be- X tween a pair of general functions that depend on d- FN (n) = P [N(R) = m]. (83) dimensional vectors was based on the integrated squared m=0 d difference in R , i.e., an L2 distance metric [71]. These Note that the series of the squared differences in (81) is authors also found that the Kullback-Leibler distance was scaled by 1/σ(R) because for a Gaussian distribution the not useful for their purposes. These findings motivated range of values for which FG ∈ [ε, 1 − ε] is proportional us to consider distance metrics based on the squared dif- to σ(R). If a particular number distribution converges ference between the Gaussian and number distributions. (sufficiently fast) to a Gaussian distribution, l2(R) will We identify here two different contributions to the “dis- tend to zero. In summary, the Gaussian distance metric tance” between a number distribution P [N(R)] and a (81) is designed to be sensitive to small deviations in Gaussian distribution: (1) deviations in the functional the distribution functions and at the same time robust form of P [N(R)] and (2) the discreteness of N(R) that against statistical fluctuations. can only approximate the continuous Gaussian distribu- tion. We found that the second contribution is essentially 2 determined by the value of the number variance σ (R), VI. SAMPLING OF MOMENTS AND NUMBER i.e., the contribution is smaller for larger values of σ2(R) DISTRIBUTION for the following reason. Consider the standardized ran- dom variable [N(R)−hN(R)i]/σ(R) whose discrete prob- We sample number fluctuations within a spherical win- ability mass function is to approximate the continuous dow of radius R for all models using a two-step procedure. normal density. Then, the bin width of the probability First, we randomly place the observation window in the mass function is given by 1/σ(R) and converges to zero sample (using a uniform distribution for its center). Sec- for σ(R) → ∞. ond, we determine the number of points N(R) within the Moreover, we found that the weight of the two con- observation window using periodic boundary conditions, tributions (relative to each other) strongly depends on except for the hyperplanes intersection process (HIP), as the representation of the number distribution, e.g., via described in Sec. IV A 5. To reduce computational re- the characteristic function (Fourier representation) [69] sources, we use the same centers for all radii that we or via direct space representations either in discrete or consider. Relevant simulation parameters and properties continuous forms. In fact, the choice of representation for each of the models described above across dimensions can virtually reverse the order of the distance metrics are listed in TableI. for our point processes. For example, a strong contri- Thus, we empirically determine the probability mass bution (2) may result in a lower distance metric for the function P [N(R)] and compute the mean value, vari- highly skew distribution of the HIP than for the Poisson ance, skewness, excess kurtosis. We determine the first process. Because contribution (2) of the discreteness of four central moments using the unbiased estimators from 2 P [N(R)] is already essentially given by σ (R), we here Ref. [77]. Finally, we compute the l2 distance metric, as choose a representation that focuses on deviations in the described above. A source of systematic errors, which functional form of P [N(R)] from that of a Gaussian ran- must be avoided for both the moments and distance mea- dom variable. sures, can arise when the number of observation windows Therefore, we define an integer-valued random variable Nwindow per sample is too large. This effect can cause G(R), whose probability mass function PG(R) is propor- the distance metric l2(R) to artificially increase again tional to a Gaussian distribution with the same first and for large radii. Therefore, we have used between 1 and second moment as our number distribution at radius R. 104 observation windows per sample depending on the We introduce a type of L2 distance metric, denoted by system size and computational cost, so that the the sys- l2(R), which is a Gaussian distance metric that employs tematic error is either avoided completely or smaller than the cumulative distribution function: the effects caused by statistical fluctuations. As a rule of thumb, we chose Nwindow such that the volume fraction " ∞ #1/2 of the union of all observation windows of the largest win- 1 X l (R) ≡ |F (n) − F (n)|2 (81) dow radius R in one sample should not exceed 50%, 2 σ(R) G N max n=0 i.e., 1 − exp [−Nwindowv1(Rmax)/V ] < 0.5, where V is the volume of a single sample. Our estimators for l2(R) where FG(n) is the cumulative distribution function of are statistically robust for values that are larger than the 14

TABLE I. Simulation parameters all of the model point processes across the first three space dimensions considered in the work. Here, N is the average of point number inside a fundamental cell, Nc is the number of point patterns considered, Nwindow is the number of observation windows per point pattern, and Rmax is the largest radius of an observation window. We have also indicated the values of the structure factors at the origin, S(0). In the cases of the equilibrium packings, they are obtained from the exact result for hard rods (d = 1) [72] and highly accurate approximations for the conditional nearest-neighbor function GP (r) in the limit r → ∞ [73, 74]. Note that our computer simulation results are in excellent agreement with these analytical estimates of S(0) for equilibrium packings.

1/d Models N Nc Nwindow Rmaxρ S(0) 2D HIP ∞ 106 1 50 ∞ Antihyperuniform 3D HIP ∞ 106 1 30.1 ∞ 4D HIP ∞ 104 1 5.4 ∞ 1D Poisson cluster 105 106 1 50 10 2D Poisson cluster 2052 107 1 100 10 3D Poisson cluster 453 107 1 20 10 1D Poisson [75] ∞ 1 ∞ 50 1 2D Poisson [75] ∞ 1 ∞ 501 Nonhyperuniform 3D Poisson [75] ∞ 1 ∞ 501 1D RSA (φ = 0.74) 107 9987 102 50 0.051 [33] 2D RSA (φ = 0.55) 104 104 103 25 0.05869(4) [55] 3D RSA (φ = 0.38) 106 250 104 25 0.05581(5) [55] 1D Equil. hard rods (φ = 0.75) 5 × 103 103 10 100 0.0629(2) 2D Equil. hard disks (φ = 0.65) 104 103 10 15 0.0260(4) 3D Equil. hard spheres (φ = 0.48) 104 100 103 20 0.022(1) 1D Cloaked URL 104 106 1 50 0 2D Cloaked URL 104 107 1 500 3D Cloaked URL 443 107 1 20 0 3 3 Hyperuniform 1D Stealthy (χ = 0.30) 10 900 10 50 0 2D Stealthy (χ = 0.49) 104 700 102 500 3D Stealthy (χ = 0.49) 103 5.3 × 103 103 5 0 Integer lattice [76] ∞ 1 ∞ 500 Square lattice 104 1 105 50 0 SC lattice 5.12 × 105 1 104 40 0 inverse of the total number n of observation windows, VII. RESULTS where n = Nwindow × Nc with Nc being the number of configurations. For a finite number of samples, we em- We describe results that we have obtained for the pirically find that l2(R) typically cannot be smaller than second, third and fourth cumulants, σ (R), γ (R) and approximately O(1/n). Therefore, we apply a datacut 2 1 γ2(R), as a function of the window radius R for all mod- and only consider radii up to els across the first three space dimensions at unit number density (ρ = 1), as well as the corresponding full prob- √ ability distributions P [N(R)] and the Gaussian distance Rcut ≡ min{R : l2(R) < 1/ n}. R>3 metric l2(R). A testament to the high-precision of the data is the excellent agreement with rigorous bounds for these quantities for general cases as well as with exact We apply the same datacut to the skewness and excess results for the cases of packings for certain R reported in kurtosis [78]. We show all data without the datacut in Sec.III (see the SM for details). the SM. All simulated data are available at a Zenodo repository [79]. 2 Another obvious source of systematic errors can arise A. Cumulants σ (R), γ1(R) and γ2(R) for the when the size of the sampling window is not much smaller Models Across Dimensions than the size of the simulation box with a fixed number of particles, i.e., when canonical ensembles are employed. Figure2 shows the second, third and fourth cumu- It is well-known that such finite-size effects lead to an lants as a function of R versus the window radius R for 2 underestimation of the local number variance σfinite(R) all models for d = 1, d = 2 and d = 3. The large-R when compared to its value in the thermodynamic limit. asymptotic behavior of the variance is determined by the An empirical formula to estimate the first-order correc- structure factor at the origin, S(0) [13]. As expected, 2 tion to the thermodynamic limit [17] shows that the error the scaled number variance σ (R)/v1(R) grows with R term is proportional S(0) (because there are larger fluc- for the HIP process for two-dimensional (2D) and three- tuations in the number of points per simulation box). dimensional (3D) cases. For all other nonhyperuniform 2 Importantly, this implies that hyperuniform models, de- models, σ (R)/v1(R) asymptotes to a constant for large fined by limk→0 S(k) = 0, are more robust against such R in all dimensions. Of course, this scaled variance de- finite-size effects. creases with R for the three hyperuniform models (hy- 15 percubic lattice, URL and stealthy systems). for this model is a monotonically decreasing function of For d ≥ 2, we have found via analyses given in the R, γ2(R) exhibits oscillations. This is entirely consistent SM and immediately below that the skewness γ1(R) and with the fact that the periodicity of the underlying lat- excess kurtosis γ2(R) for our disordered hyperuniform tice is completely hidden at the level of the three-point systems vanish faster with increasing R than those for correlation function but manifests itself for the first time nonhyperuniform systems. Among all models studied, in g4 [62]. the quantities γ1(R) and γ2(R) vanish slowest for the antihyperuniform HIP models for d ≥ 2. The specific decay rates for all models are described below. B. Number Distributions for the Models Across It is noteworthy that one-dimensional systems can Dimensions present fluctuation anomalies not present in higher di- mensions. For example, in the case of the integer lat- Figures3 and4 show representative configurations of tice, the random variable N(R) can only take at most all of the 2D models and their corresponding standard- two values for any R, which of course is abnormally non- ized number distributions. (We provide corresponding Gaussian. While the hypercubic lattice for d ≥ 2 never figures for all 1D and 3D models in the SM). Except achieves a CLT (as discussed below), the variance is con- for hypercubic lattices for any d and 1D hyperuniform siderably broader than that for d = 1. Another anoma- systems of class I, all of the considered models across lous category is class I hyperuniform systems, which have dimensions obey a CLT. It is noteworthy that for all bounded variance for d = 1 [see Eq. (76)] and thus any models, except the hypercubic lattices, the number dis- such hyperuniform point process cannot obey a CLT be- tribution functions P [N(R)] are unimodal, i.e., one with cause the standardized distribution cannot converge to a a single peak. (For small radii R / 0.5, the number dis- continuous distribution. This is clearly borne out by the tributions are monotonically decreasing, which are still distance metric plots, shown in Fig.5 and Fig. S8 of the considered to be unimodal.) Recall that for all models, SM, for both the 1D disordered stealthy point process except the hypercubic lattice and the class I hyperuni- and integer lattice. form models in one dimension, the skewness and excess For the Poisson and super-Poissonian models (cluster kurtosis tend to zero for large R for all dimensions. For and HIP) across dimensions, both the skewness and ex- well-behaved unimodal distributions, such a vanishing of cess kurtosis decay monotonically; see Fig.2. By con- both γ1(R) and γ2(R) indicates a tendency to a CLT. trast, the skewness and excess kurtosis oscillate about Indeed, this is confirmed by visual inspection of our cor- zero for the lattice, stealthy, equilibrium and RSA sys- responding evaluations of the number distributions for all tems across all dimensions because all of them exhibit dimensions; see Figs.3 and4. Our conclusions about the at least short-range order. It should not go unnoticed tendencies to CLTs for sufficiently large R are further how the oscillations in γ1(R) and γ2(R) are related to confirmed by our evaluations of the Gaussian distance short- or long-range order at the level of three- and four- metric l2(R) for all models across dimensions, which are body correlation functions, g3 and g4, as can be seen plotted in Fig.5. For d = 2 and d = 3, one clearly sees from the explicit formulas (22) and (23) for the skewness that disordered hyperuniform point processes are better and excess kurtosis. In the instances of RSA and equilib- approximated by the normal distribution than their non- rium packings, oscillations in γ1(R) and γ2(R) arise from hyperuniform counterparts at a given large value of R. strong short-range order exhibited by g3 and g4. This is consistent with the visual inspections of the plots The disordered hyperuniform systems that we con- presented Figs.3 and4. From Fig.5 and Fig. S8 of sider, stealthy and URL point processes, have extraor- the SM, one can ascertain, for each model, a radius R0 dinary number fluctuation behaviors. For 2D and 3D above which the distribution can be deemed to be approx- stealthy systems, both γ1(R) and γ2(R) strongly oscil- imately Gaussian, i.e., l2(R) is below some threshold for late about zero; see Fig.2. This suggests that at the R > R0, implying that it is approximately determined level of g3, stealthy systems, counterintuitively, exhibit by only the mean and the variance. Typically, we find significant ordering on much larger length scales than that R0 is orders of magnitude smaller for our hyperuni- the short-range order seen in the pair correlation func- form and standard nonhyperuniform models than for our tion [67]. For 1D stealthy systems, γ1(R) and γ2(R) show super-Poissonian models. even stronger oscillations than their higher-dimensional Our results clearly show that the hypercubic lattices counterparts, indeed reflecting possible long-range order across dimensions do not obey a CLT; see Figs.2,4 and present in g3 and g4. It is remarkable that the skewness 5. This is due to the fact that lattice points in general and excess kurtosis can detect such anomalous long-range are “rigid” in the sense that fluctuations in N(R) are order that would not be expected based solely on the be- always stringently bounded from below and above for havior of the pair correlation function. Another reason any value of R (as discussed in Sec.IIIC) due to their that supports the capacity of γ1(R) and γ2(R) to de- inherent long-range order. Hence, P [N(R)] is highly sen- tect unusual long-range order is the cloaked URL model, sitive to the value of R, so that both γ1(R) and γ2(R) which at the level of the pair correlation function would rapidly oscillate around zero, but the amplitudes of the be considered to be highly disordered. Whereas γ1(R) oscillations do not vanish as R increases. It is already 16

103

101 1D 2 2D 3D 10 Zd 102 Stealthy 0 1 URL

) 10 10 1 Equil. R 10 ( RSA 1 −1 0 Poisson /v 10 10 2 Cluster ) 100 HIP R ( −1 σ 10−2 10 10−1 −2 10−3 10 10−2 0 5 10 15 20 25 0 5 10 15 20 25 0 2 4 6 8 10 12 10 10 15 1D 2D 3D Zd 2 2 4 Stealthy 10 2 URL 0 0 Equil. 0

) 5 5 RSA − − − Poisson R 2 2 2 ( Cluster 1 0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2

γ 5 HIP

0 0 0

0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 20 20 1D 2D 3D 60 Zd 2 2 2 Stealthy URL 0 0 0 Equil.

) RSA 10 − 10 − − Poisson R 2 2 2 ( Cluster 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2 30 0 0.5 1 1.5 2

γ HIP

0 0 0 0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10 R R R

FIG. 2. Graphs of the number variance, skewness and excess kurtosis versus the window radius R for all considered models across the first three space dimensions. well-established that the local variance σ2(R) of lattices values of R. (Recall that the hypercubic lattices and 1D exhibits such rapid oscillations; see Ref. [13] and refer- hyperuniform systems do not obey a CLT.) The gamma ences therein. Because of the rapid oscillations in γ1(R), distribution is defined by γ2(R), and l2(R) for hypercubic lattices, we represent the 1 data by points instead of curves, which would require in- P [N(R) = m] = mk−1e−m/θ, (84) terpolation between the data points. Note that the corre- Γ(k)θk sponding distance metrics l2(R) are bounded from below for any value of R. These observations are consistent with where k and θ are shape and scale parameters, respec- the visual inspections of the number distribution shown tively, which are related to the mean and variance of in Fig.4 for d = 2 and those for d = 1 and d = 3 given P [N(R)] as follows: in the SM. hN(R)i =kθ, (85) σ2(R) =kθ2. (86) C. Gamma-Distribution Approximation for All Models Obeying a CLT It immediately follows that the associated skewness and excess kurtosis can be expressed simply in terms of the We have studied a variety of well-known closed-form mean and number variance, yielding probability distributions to determine those which best ñ 2 ô1/2 approximate the actual distributions for finite R for all of 2 σ (R) γ1(R) =√ = 2 , (87) our models. Remarkably, we have ascertained that the k hN(R)i2 gamma distribution provides a good approximation to 6 ñ σ2(R) ô the number distribution P [N(R)] for all models that obey γ2(R) = = 6 2 . (88) a CLT across all dimensions for intermediate to large k hN(R)i 17

. 102 . 102 2D HIP R 2D Poisson cluster R

− − 10 1 101 10 1 101 )] )] R R ( ( N N [ [

P 0 P 0 −2 10 −2 10

× 10 × 10 ) ) R R ( ( σ σ 10−1 10−1 10−3 10−3

−6 −4 −20 2 4 6 810 −6 −4 −20 2 4 6 810 . [N(R) −hN(R)i]/σ(R) . [N(R) −hN(R)i]/σ(R) . . 102 102 2D Poisson R 2D Equil. disks R

− − 10 1 101 10 1 1 )] )] 10 R R ( ( N N [ [

P 0 P 0 −2 10 −2 10 × 10 × 10 ) ) R R ( ( σ σ 10−1 10−1 10−3 10−3

−6 −4 −20 2 4 6 810 −6 −4 −20 2 4 6 810 . [N(R) −hN(R)i]/σ(R) . [N(R) −hN(R)i]/σ(R) . 102 2D RSA R

− 10 1 101 )] R ( N [

P 0 −2 10

× 10 ) R ( σ 10−1 10−3

−6 −4 −20 2 4 6 810 . [N(R) −hN(R)i]/σ(R)

FIG. 3. Representative images of configurations of each of the five 2D nonhyperuniform models, beginning with the most nonhyperuniform one (HIP) down to one that has the smallest structure factor at the origin (saturated RSA). The corresponding standardized probability distributions for different window radii are also included; deepest blue color (darkest shade) corresponds to zero radius and the deep orange color corresponds to the largest considered radius.

2 TABLE II. Large-R asymptotic scalings of σ (R), γ1(R), γ2(R) and l2(R) for all of our models across dimensions that obey a CLT, as obtained by the approximation of P [N(R)] by the gamma distribution function for each model for any d. Descriptor Stealthy & URL Cluster, Poisson, RSA, & Equilibrium HIP σ2(R) R(d−1) Rd R2d−1 −(d+1)/2 −d/2 −1/2 γ1(R) R R R −(d+1) −d −1 γ2(R) R R R −(d+1)/2 −d/2 −1/2 l2(R) R R R

−(d+1)/2 Figure6 compares the gammma-distribution approxi- runiform models, and γ1(R) ∼ R and γ2(R) ∼ mations to simulation data for a representative disor- R−(d+1) for the hyperuniform models. TableII summa- dered hyperuniform model (stealthy), a standard nonhy- rizes these scaling behaviors. The fact that the excess peruniform model (RSA) and an antihyperuniform model kurtosis decays to zero faster than the skewness for any (HIP) in two dimensions for R = 5. Visual inspection model that obeys a CLT, whether hyperuniform or not, reveals the gamma distribution provides a good approxi- implies that the dominant asymptotic correction of the mation in each case. Similar good agreement between the gamma distribution to a CLT for large R is determined gamma-distribution approximation and simulation data by the skewness; see AppendixA for a proof. It is note- was found for other models (not shown in Fig.6) obeying worthy that these predictions based on the gamma dis- a CLT across dimensions, as discussed in the SM. tribution are consistent with numerical findings for all nonhyperuniform systems and the antihyperuniform HIP Importantly, the approximation of P [N(R)] by a using an independent method that employs certain run- gamma distribution enables us to estimate the large-R ning averages of γ1(R) and γ2(R); see the SM for de- scalings of γ1(R) and γ2(R) for all models across dimen- tails. Moreover, for 2D and 3D disordered hyperuniform sions that obey a CLT. Specifically, we find γ1(R) ∼ (stealthy and URL) models, the predictions from the −1/2 −1 R and γ2(R) ∼ R for the antihyperuniform HIP, gamma-distribution approximations are also consistent −d/2 −d γ1(R) ∼ R and γ2(R) ∼ R for standard nonhype- 18

. . 102 102 2D Cloaked URL R 2D Stealthy R

− − 10 1 101 10 1 1 )] )] 10 R R ( ( N N [ [

P 0 P 0 −2 10 −2 10 × 10 × 10 ) ) R R ( ( σ σ 10−1 10−1 10−3 10−3

−6 −4 −20 2 4 6 810 −6 −4 −20 2 4 6 810 . [N(R) −hN(R)i]/σ(R) . [N(R) −hN(R)i]/σ(R) . 102 2D Lattice R

− 10 1 101 )] R ( N [

P 0 −2 10

× 10 ) R ( σ 10−1 10−3

−6 −4 −20 2 4 6 810 . [N(R) −hN(R)i]/σ(R)

FIG. 4. Representative images of configurations of each of the three 2D hyperuniform models and their corresponding stan- dardized probability distributions, where the deepest blue color (darkest shade) corresponds to zero radius and the deep orange color corresponds to the largest considered radius.

100 100 100 1D 2D Zd 3D Stealthy −1 −1 URL 10 10 Equil. 10−1 RSA Poisson

) Cluster −2 −2 R 10 10 HIP ( 2 l 10−2 10−3 10−3

10−3 10−4 10−4 10−2 10−1 100 101 10−2 10−1 100 101 10−2 10−1 100 101 R R R

FIG. 5. The Gaussian distance metric l2(R) versus the window radius R for all considered models across the first three space dimensions. with the observed scalings of the skewness. Due to the described in AppendixA. strong oscillations in the excess kurtosis for 2D and 3D According to TableII, convergence to a CLT is slowest stealthy and URL processes described above, it was nu- for the HIP (proportional to R−1/2 for d ≥ 2), followed merically difficult to definitively determine their scalings by the Poisson cluster process, and the Poisson process. from the running-average method. It should not go unno- RSA and equilibrium packings have the same scaling be- ticed that the exact formulas for the skewness and excess haviors as the Poisson cluster and Poisson processes, but kurtosis, Eqs. (34) and (35), respectively, of the Pois- with smaller coefficients of proportionality. The conver- son distribution are consistent with the scalings predicted gence to a CLT is fastest for the disordered hyperuniform by the gamma-distribution approximation, lending addi- processes for d = 2 and d = 3 [80] tional validity to the latter. Furthermore, we have found that for all models across dimensions that obey a CLT, the running-average procedure yields scalings for l2(R) that agree with the corresponding gamma-distribution D. Effect of Dimensionality on the Approach to a CLT approximations (see SM), which are identical to the scal- ings for the skewness γ1(R). The proof that l2(R) has exactly the same scaling as γ1(R) for the gamma dis- For any particular d-dimensional model that eventu- tribution is given in AppendixA. Furthermore, our ap- ally obeys a CLT, does the number distribution func- proximations of P [N(R)] by gamma distributions are also tion P [N(R)] tend to Gaussian-like behavior faster as consistent with CLTs, since they converge to the Gaus- the space dimension d increases? As noted in the In- sian distribution in the limit of k → ∞ (i.e, R → ∞), as troduction, this question can be answered by appealing to the decorrelation principle [29], which states uncon- 19

100 Simulation Stealthy R =5.0 Simulation RSA R =5.0 Simulation HIP R =5.0 −1 Gaussian Gaussian Gaussian 10 Gamma Gamma Gamma

10−2

)] −3

R 10 ( −4 N

[ 10 P 10−5

10−6

10−7 65 70 75 80 85 90 95 65 70 75 80 85 90 95 0 70 140 210 280 350 N(R) N(R) N(R)

FIG. 6. Comparison of simulation data for the probability distributions P [N(R)] with R = 5 for three 2D models (stealthy hyperuniform, RSA, and HIP) to corresponding gamma-distribution approximations with the same mean value and variance. The gamma distribution provides good approximations for all models that obey a CLT across the first three dimensions considered, including the ones not shown here.

100 1D URL 1D RSA 2D HIP 2D 2D 3D 3D 3D 4D

10−1 ) R (

2 −2

l 10

10−3

10−2 10−1 100 101 10−2 10−1 100 101 10−2 10−1 100 101 ρ1/dR ρ1/dR ρ1/dR

FIG. 7. The Gaussian distance metric l2(R) versus R for a representative disordered hyperuniform model (URL), sub-Poisson nonhyperuniform model (RSA) and anti-hyperuniform model (HIP) across the first three space dimensions, respectively. strained correlations in disordered packings that exist the URL and RSA models for a fixed value of R (for in low dimensions vanish as d tends to infinity, and all R > 1), l2(R) decreases with increasing d (increasingly higher-order correlation functions gn for n ≥ 3 may be tends to Gaussian-like behavior) because of decorrela- expressed in terms of the number density ρ and pair cor- tion, while for the HIP, l2(R) increases with increasing relation function g2. The decorrelation principle begins d (moves away from Gaussian-like behavior) because it to manifest itself in low dimensions for disordered pack- increasingly correlates with d. Importantly, when com- ings [31–33] but other disordered systems with strongly paring fluctuations across dimensions, one must choose repulsively interacting particles, including fermionic [34] a meaningful length scale to make the window radius R and Gaussian-core point processes [35]. We know that dimensionless. A simple and good choice is ρ1/d, which the number distribution function P [N(R)] generally in- is proportional to the mean nearest-neighbor distance in volves certain integrals over all of the n-body correla- space dimension d, and explains why the horizontal axes tion functions. Therefore, the decorrelation principle im- in each subfigure of Fig.7 is ρ1/dR. plies that for any d-dimensional model that decorrelates with d, P [N(R)] increasingly becomes Gaussian-like as d increases, since the first and second moments, deter- VIII. CONCLUSIONS AND DISCUSSION mined by ρ and g2, dominate the distribution. By the same token, for any model that correlates with increas- ing d, P [N(R)] increasingly deviates from the normal Via theoretical methods and high-precise simulation distribution as d increases. We have verified these broad studies, we accurately quantified the skewness γ1(R), ex- conclusions for the models studied in this article. In cess kurtosis γ2(R) and the number distribution P [N(R)] for eight different models of statistically homogeneous Fig.7, we plot the Gaussian distance metric l2(R) for a representative disordered hyperuniform model (URL), a point processes in two and three dimensions: five non- sub-Poisson nonhyperuniform model (RSA) and an anti- hyperuniform models, one of which is anti-hyperuniform hyperuniform model (HIP) across the first three space (HIP), and three hyperuniform models. Analogous mod- dimensions, respectively. (Recall that the URL obeys els were also examined in one dimension, except for HIP, a CLT for d ≥ 2.) As expected, we see that for both which is not defined in this dimension. We validated our simulation results for γ1(R), γ2(R) and P [N(R)] for all 20 models by showing that they are in excellent agreement system [16]. These distributions are distinguished from with rigorous bounds and exact results that we derived those of our models in that they have heavy left tails for the applicable ranges of R. For all disordered hype- (γ1(R) < 0). This implies that they cannot be approxi- runiform models, our explicit general formulas for γ1(R) mated by gamma distributions, as in the majority of the and γ2(R) in terms of n-body information [Eqs. (22) and models studied in the present paper across dimensions. (23)] enable us to infer the existence of “hidden” type of A related but distinctly different study from the one long-range order that manifests itself for the first time considered in the present work concerns fluctuations at the three-body level or higher. Thus, the skewness when the window is centered on a point of the point and excess kurtosis have the capacity to detect anoma- process [9, 43]. It will be interesting in future work to lous long-range order that would not be expected based carry out the analogous investigation of the correspond- on the behavior of the pair correlation function alone. ing skewness, excess kurtosis and number distribution for We have introduced a novel Gaussian “distance” met- such particle-based quantities for hyperuniform and non- ric l2(R) to ascertain the proximity of P [N(R)] to the hyperuniform models. normal distribution for each model as a function of R. We have verified that l2(R) is a sensitive metric via numerical and theoretical methods. Since the distributions for all Appendix A: Asymptotic Behavior of l2(R) for the models (except the lattices) across dimensions are uni- Gamma Distribution modal, the tendency to a CLT corresponds to the skew- ness and excess kurtosis simultaneously tending to zero. We prove that the large-R scalings of the distance Almost all of the considered models across dimensions metric l2(R) and the skewness γ1(R) are the same for obey a CLT. We found that disordered hyperuniform the gamma distribution. It is important to note that point processes are better approximated by the normal the asymptotic analysis is facilitated by transforming the distribution than their nonhyperuniform counterparts at random variable N(R) to the standardized random vari- a given large value of R. We proved that any 1D hyper- able X ≡ (N(R) − hN(R)i)/σ(R). In the large-R limit, uniform system of class I as well as the hypercubic lattice the discrete random variable tends to a continuous one for any d cannot obey a CLT. Similarly, a general lattice and hence the continuous-variable analog of the distance in any dimension cannot obey a CLT. metric l2(R), defined Eq. (81), is given by It is noteworthy that we discovered that the gamma ïZ ∞ ò1/2 distribution provides a good approximation to the num- 2 ber distribution P [N(R)] for all models that obey a CLT l2(R) ≡ |FG(X) − FΓ(X)| dX (A1) −∞ across all dimensions for intermediate to large values of R, enabling us to estimate the large-R scalings of γ1(R), where FG(X) and FΓ(X) are the cumulative distribution γ2(R) and l2(R). These predictions were corroborated by functions for the Gaussian and gamma distribution func- corresponding simulation results in almost all instances, tions, respectively. Specifically, as detailed in the SM. It is only in the special cases of 1 î √ ó the excess kurtosis for 2D and 3D stealthy and URL pro- F (X) = 1 + erf(X/ 2) (A2) cesses where the running-average method was not reli- G 2 able enough to definitively determine their scalings due to strong oscillations, and thus this represents a simu- and lation challenge for future research. Among all mod- √ Γ(k, X k + k) els, the convergence to a CLT is generally fastest for FΓ(X) = 1 − (A3) the disordered hyperuniform processes in two and three Γ(k) −(d+1)/2 dimensions such that γ1(R) ∼ l2(R) ∼ R and R ∞ s−1 −(d+1) where Γ(s, y) = t exp(−t)dt is the upper incom- γ2(R) ∼ R for large R. The convergence to a CLT y is slower for standard nonhyperuniform models such that plete gamma function. −d/2 −d Since the (n+1)th cumulant of the gamma distribution γ1(R) ∼ l2(R) ∼ R and γ2(R) ∼ R . Not surpris- ingly, the convergence to a CLT is slowest for the an- tends to zero faster than its nth cumulant for n ≥ 3, the greatest deviation of the gamma distribution from the tihyperuniform HIP model such that γ1(R) ∼ l2(R) ∼ −1/2 −1 normal one occurs in the vicinity of the origin, i.e., X = R and γ2(R) ∼ R . Using the decorrelation prin- ciple, we elucidated why any d-dimensional model that 0, in the large k-limit (see Fig.8) and is thus dominated “decorrelates” or “correlates” with d corresponds to a by the asymptotic behavior of the skewness. Thus, we P [N(R)] that increasingly moves toward or away from a require the Taylor series expansions of the distributions CLT, respectively. about X = 0: It has recently been reported that in a 2D absorb- … … … 1 1 2 1 2 3 1 2 5 ing phase of a lattice gas model, the density distribu- FG(X) = + X − X + X tion at a hyperuniform length scale increasingly deviates 2 2 π 12 π 80 π 1 … 2 from a Gaussian distribution on approach to the criti- − X7 + O(X9) (A4) cal point [18], corresponding to a class III hyperuniform 672 π 21

0.0002 We see that for odd m, the leading-order terms of am k=100 are exactly the same as the coefficients multiplying Xm k=1000 0.00015 k in the series expansion (A4) for the standard normal dis- 2 =10000

)| tribution. Thus, combining (A4), (A5) and the afore- X ( mentioned asymptotic expansions yield Γ

F 0.0001 )- X (

G f(X) 1

F F (X) − F (X) = √ + O( ), (A15) | Γ G 5e-05 k k3/2

0 where f(X) is a function that is localized about the origin -4 -2 0 2 4 X with the following corresponding Taylor series expansion:

FIG. 8. Integrand of (A1) versus X for select values of the shape parameter k. … 2 ï1 1 5 7 1 ò f(X) = − X2 − X4 − X6 − X8 + O(X10) . π 6 4 48 288 256 2 3 FΓ(X) =a0(k) + a1(k)X + a2(k)X + a3(k)X (A16) 4 5 6 7 Substitution of (A15) into (A1) gives the large-k asymp- + a4(k)X + a5(k)X + a6(k)X + a7(k)X totic expansion of the Gaussian distance metric, i.e., 8 9 + a8(k)X + O(X ), (A5) c 1 where the coefficients am(k)(m = 1, 2, 3,...) depend on l2(R) = √ + O( ), (A17) the shape parameter k. We know these coefficients ex- k k3/2 plicitly but do not indicate them for reasons of brevity. The corresponding large-k asymptotic expansions of the where the square of the constant c is given by first nine coefficients are … 1 1 2 Z ∞ a (k) = + + O(1/k) (A6) 2 2 0 2 6 kπ c = f (X)dX. (A18) −∞ 1… 2 a (k) = + O(1/k) (A7) 1 2 π √ 1… 2 Finally, since the skewness γ1(R) scales like 1/ k, we a (k) = − + O(1/k3/2) (A8) conclude that l (R) ∼ γ (R) for large values of R. Note 2 4 kπ 2 1 that the vanishing of l2(R) for R → ∞ implies a CLT for 1 … 2 a (k) = − + O(1/k) (A9) the gamma distribution as described in the text. 3 12 π 5 … 2 a (k) = + O(1/k3/2) (A10) 4 48 kπ 1 … 2 a (k) = + O(1/k) (A11) 5 80 π ACKNOWLEDGMENTS … 7 2 3/2 a6(k) = − + O(1/k ) (A12) 288 kπ We thank Steven Atkinson for his configurations of 3D 1 … 2 equilibrium hard spheres. This work was supported in a (k) = − + O(1/k) (A13) 7 672 π part by the National Science Foundation under Award 1 … 2 No. DGE-2039656 and by the Princeton University In- a (k) = + O(1/k3/2) (A14) novation Fund for New Ideas in the Natural Sciences. 8 256 kπ

∗ Email: [email protected] 66123 Saarbr¨ucken, Germany † Email: [email protected] [1] P. Schofield, “Wavelength-dependent fluctuations in clas- ‡ Email: [email protected]; Present address: sical fluids: I. The long wavelength limit,” Proc. Phys. Friedrich-Alexander-Universit¨at Erlangen-N¨urnberg Soc. 88, 149–170 (1966). (FAU), Institut f¨urTheoretische Physik, Staudtstr. 7, [2] D. J. Vezzetti, “A new derivation of some fluctuation 91058 Erlangen, Germany, and Department of Exper- theorems in ,” J. Math. Phys. 16, imental Physics, Saarland University, Campus E2 9, 31–33 (1975). 22

[3] R. M. Ziff, “On the bulk distribution functions and fluc- [23] Z. Ma, E. Lomba, and S. Torquato, “Optimized large tuation theorems,” J. Math. Phys. 18, 1825–1831 (1977). hyperuniform binary colloidal suspensions in two dimen- [4] F. Carmona and P. Delhaes, “Effect of density fluctua- sions,” Phys. Rev. Lett. 125, 068002 (2020). tions on the physical properties of a disordered carbon,” [24] G. Last and M. Penrose, Lectures on the Poisson Process, J. Appl. Phys. 49, 618–628 (1978). Institute of Mathematical Statistics Textbooks (Cam- [5] J. P. Hansen and I. R. McDonald, Theory of Simple Liq- bridge University Press, Cambridge, United Kingdom, uids (Academic Press, New York, 1986). 2017). [6] K. Jørgensen, J. H. Ipsen, O. G. Mouritsen, D. Bennett, [25] Mathew D. Penrose and J. E. Yukich, “Limit Theory for and M. J. Zuckermann, “The effects of density fluctua- Random Sequential Packing and Deposition,” Ann. Appl. tions on the partitioning of foreign molecules into lipid Probab. 12, 272–301 (2002). bilayers: application to anaesthetics and insecticides,” [26] L. Heinrich, H. Schmidt, and V. Schmidt, “Central limit Biochimica et Biophysica Acta (BBA)-Biomembranes theorems for Poisson hyperplane tessellations,” Annals 1067, 241–253 (1991). Applied Prob. 16, 919–950 (2006). [7] P. J. E. Peebles, Principles of Physical Cosmology [27] T. Schreiber and J. E. Yukich, “Limit theorems for geo- (Princeton University Press, Princeton, 1993). metric functionals of Gibbs point processes,” Ann. Inst. [8] P. M. Bleher, F. J. Dyson, and J. L. Lebowitz, “Non- H. Poincar´eProbab. Statist. 49, 1158–1182 (2013). Gaussian energy level statistics for some integrable sys- [28] B. Blaszczyszyn, D. Yogeshwaran, and J. E. Yukich, tems,” Phys. Rev. Lett. 71, 3047–3050 (1993). “Limit theory for geometric statistics of point processes [9] T. M. Truskett, S. Torquato, and P. G. Debenedetti, having fast decay of correlations,” Ann. Probab. 47, 835– “Density fluctuations in many-body systems,” Phys. Rev. 895 (2019). E 58, 7639–7380 (1998). [29] S. Torquato and F. H. Stillinger, “New conjectural lower [10] S. Torquato, “Modeling of physical properties of com- bounds on the optimal density of sphere packings,” Ex- posite materials,” Int. J. Solids Structures 37, 411–422 perimental Math. 15, 307–331 (2006). (2000). [30] A. Scardicchio, F. H. Stillinger, and S. Torquato, “Esti- [11] A. Gabrielli, M. Joyce, and F. Sylos Labini, “Glass-like mates of the optimal density of sphere packings in high universe: Real-space correlation properties of standard dimensions,,” J. Math. Phys. 49, 043301 (2008). cosmological models,” Phys. Rev. D 65, 083523 (2002). [31] M. Skoge, A. Donev, F. H. Stillinger, and S. Torquato, [12] A. Wax, C. Yang, V. Backman, K. Badizadegan, C. W. “Packing hyperspheres in high-dimensional Euclidean Boone, R. R. Dasari, and M. S. Feld, “Cellular organiza- spaces,” Phys. Rev. E 74, 041127 (2006). tion and substructure measured using angle-resolved low- [32] S. Torquato and F. H. Stillinger, “Exactly solvable dis- coherence interferometry,” Biophys. J. 82, 2256 – 2264 ordered sphere-packing model in arbitrary-dimensional (2002). Euclidean spaces,” Phys. Rev. E 73, 031106 (2006). [13] S. Torquato and F. H. Stillinger, “Local density fluctua- [33] S. Torquato, O. U. Uche, and F. H. Stillinger, “Random tions, hyperuniform systems, and order metrics,” Phys. sequential addition of hard spheres in high Euclidean di- Rev. E 68, 041113 (2003). mensions,” Phys. Rev. E 74, 061308 (2006). [14] A. C. Lavery, R. W. Schmitt, and T. K. Stanton, “High- [34] S. Torquato, A. Scardicchio, and C. E. Zachary, “Point frequency acoustic scattering from turbulent oceanic mi- processes in arbitrary dimension from Fermionic gases, crostructure: the importance of density fluctuations,” J. random matrix theory, and number theory,” J. Stat. Acoustical Soc. America 114, 2685–2697 (2003). Mech.: Theory Exp. 2008, P11019 (2008). [15] M. A. Klatt and S. Torquato, “Characterization of maxi- [35] C. E. Zachary, F. H. Stillinger, and S. Torquato, mally random jammed sphere packings. II. Correlation “Gaussian-core model phase diagram and pair correla- functions and density fluctuations,” Phys. Rev. E 94, tions in high Euclidean dimensions,” J. Chem. Phys. 128, 022152 (2016). 224505 (2008). [16] S. Torquato, “Hyperuniform states of matter,” Physics [36] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke, Reports 745, 1–95 (2018). Stochastic Geometry and Its Applications, 3rd ed. (Wiley, [17] F. L. Rom´an,J. A. White, A. Gonzalez, and S. Velasco, Chichester, 2013). “Fluctuations in a small hard-disk system: Implicit finite [37] S. Torquato, Random Heterogeneous Materials: Mi- size effects,” J. Chem. Phys. 110, 9821–9824 (1999). crostructure and Macroscopic Properties (Springer- [18] Y. Zheng and M. P. Ciamarra, “Spatio-temporal Hetero- Verlag, New York, 2002). geneity and Hyperuniformity in 2D Conserved Lattice [38] J. H. Conway and N. J. A. Sloane, Sphere Packings, Lat- Gas,” arXiv:2009.07187 (2020). tices and Groups (Springer-Verlag, New York, 1998). [19] A. Chremos and J. F Douglas, “Hidden hyperunifor- [39] S. Torquato, “Reformulation of the covering and quan- mity in soft polymeric materials,” Phys. Rev. Lett. 121, tizer problems as ground states of interacting particles,” 258002 (2018). Phys. Rev. E 82, 056109 (2010). [20] Y. et al. Zheng, “Disordered hyperuniformity in two- [40] S. Torquato, “Perspective: Basic understanding of con- dimensional amorphous silica,” Science Adv. , eaba0826 densed phases of matter via packing models,” J. Chem. (2020). Phys. 149, 020901 (2018). [21] A. Sheremet, R. Pierrat, and R. Carminati, “Absorption [41] Karl Pearson, “IX. mathematical contributions to the of scalar waves in correlated disordered media and its theory of evolution.—XIX. second supplement to a mem- maximization using stealth hyperuniformity,” Phys. Rev. oir on skew variation,” Phil. Trans. R. Soc. Lond. A 216, A 101, 053829 (2020). 429–457 (1916). [22] S. Wilken, R. E. Guerra, D. J. Pine, and P. M. Chaikin, [42] D. S. Broca, “Cumulant-moment relations through de- “Hyperuniform structures formed by shearing colloidal terminants,” Int. J. Math. Ed. Sci. Tech. 35, 917–921 suspensions,” Phys. Rev. Lett. 125, 148001 (2020). (2004). 23

[43] S. Torquato, B. Lu, and J. Rubinstein, “Nearest- 165, 115–151 (2006). neighbor distribution functions in many-body sytems,” [62] M. A. Klatt, J. Kim, and S. Torquato, “Cloaking the Phys. Rev. A 41, 2059–2075 (1990). underlying long-range order of randomly perturbed lat- [44] A. B. Hopkins, F. H. Stillinger, and S. Torquato, “Dens- tices,” Phys. Rev. E 101, 032118 (2020). est local sphere-packing diversity: General concepts and [63] Y. Fan, J. K. Percus, D. K. Stillinger, and F. H. Still- application to two dimensions,” Phys. Rev. E 81, 041305 inger, “Constraints on collective density variables: One (2010). dimension,” Phys. Rev. A 44, 2394–2402 (1991). [45] A. B. Hopkins, F. H. Stillinger, and S. Torquato, “Dens- [64] O. U. Uche, F. H. Stillinger, and S. Torquato, “Con- est local sphere-packing diversity: Application to three straints on collective density variables: Two dimensions,” dimensions,” Phys. Rev. E 83, 011304 (2011). Phys. Rev. E 70, 046122 (2004). [46] G. Zhang, F. H. Stillinger, and S. Torquato, “Can exotic [65] O. U. Uche, S. Torquato, and F. H. Stillinger, “Collective disordered “stealthy” particle configurations tolerate ar- coordinates control of density distributions,” Phys. Rev. bitrarily large holes?” Soft Matter 13, 6197–6207 (2017). E 74, 031104 (2006). [47] D. Levine and P. J. Steinhardt, “Quasicrystals: A new [66] R. D. Batten, F. H. Stillinger, and S. Torquato, “Clas- class of ordered structures,” Phys. Rev. Lett. 53, 2477– sical disordered ground states: Super-ideal gases, and 2480 (1984). stealth and equi-luminous materials,” J. Appl. Phys. [48] S. Ghosh and J. L. Lebowitz, “Generalized stealthy hy- 104, 033504 (2008). peruniform processes: Maximal rigidity and the bounded [67] S. Torquato, G. Zhang, and F. H. Stillinger, “Ensem- holes conjecture,” Comm. Math. Phys. 363, 97–110 ble theory for stealthy hyperuniform disordered ground (2018). states,” Phys. Rev. X 5, 021020 (2015). [49] C. E. Zachary and S. Torquato, “Hyperuniformity in [68] G. Zhang, F. Stillinger, and S. Torquato, “Ground states point patterns and two-phase heterogeneous media,” J. of stealthy hyperuniform potentials: I. Entropically fa- Stat. Mech.: Theory & Exp. 2009, P12015 (2009). vored configurations,” Phys. Rev. E 92, 022119 (2015). [50] (2020). [69] Norbert Henze, “Invariant tests for multivariate normal- [51] A. Re´nyi,“On a one-dimensional problem concerning ity: A critical review,” Statistical Papers 43, 467–506 random space filling,” Sel. Trans. Math. Stat. Prob. 4, (2002). 203–218 (1963). [70] S. Kullback and R. A. Leibler, “On Information and Suf- [52] B. Widom, “Random sequential addition of hard spheres ficiency,” Ann. Math. Statist. 22, 79–86 (1951). to a volume,” J. Chem. Phys. 44, 3888–3894 (1966). [71] H. Wang, F. H. Stillinger, and S. Torquato, “Sensitivity [53] J. Feder, “Random sequential adsorption,” J. Theor. of pair statistics on pair potentials in many-body sys- Biol. 87, 237–254 (1980). tems,” J. Chem. Phys. (2020). [54] D. W. Cooper, “Random-sequential-packing simulations [72] F. Zernike and J. A. Prins, “Die Beugung von in three dimensions for spheres,” Phys. Rev. A 38, 522– R¨ontgenstrahlen in Fl¨ussigkeiten als Effekt der 524 (1988). Molek¨ulanordnung,” Z. Phys. 41, 184–194 (1927). [55] G. Zhang and S. Torquato, “Precise algorithm to gener- [73] S. Torquato, “Mean nearest-neighbor distance in random ate random sequential addition of hard hyperspheres at packings of hard d-dimensional spheres,” Phys. Rev. Lett. saturation,” Phys. Rev. E 88, 053312 (2013). 74, 2156–2159 (1995). d [56] B. Bonnier, D. Boyer, and P. Viot, “Pair correlation [74] For equilibrium packings in R , S(0) is estimated from function in random sequential adsorption processes,” J. the fluctuation-compressibility relation S(0) = ρkB T κT , Phys. A: Math. Gen. 27, 3671–3682 (1994). where kB is the Boltzmann constant, T is temperature, −1 ∂ρ [57] J. Illian, A. Penttinen, H. Stoyan, and D. Stoyan, Sta- κT ≡ ρ ( ∂p )T is the isothermal compressibility, and tistical Analysis and Modelling of Spatial Point Patterns p/(ρkB T ) = GP (∞) is the reduced pressure. (John Wiley & Sons, Chichester, 2008). [75] Analytic formula (6) is employed. [58] M. A. Klatt, J. Lovri´c,D. Chen, S. C. Kapfer, F. M. [76] Analytic formulas (70)-(73) are employed. Schaller, P. W. A. Sch¨onh¨ofer, B. S. Gardiner, A-S. [77] B. Klemens, Modeling with Data: Tools and Techniques Smith, G. E. Schr¨oder-Turk, and S. Torquato, “Uni- for Scientific Computing (Princeton University Press, versal hidden order in amorphous cellular geometries,” Princeton, N.J, 2009) Online appendix M. Nature Comm. 10, 811 (2019). [78] We expect this datacut to be conservative, because the [59] R. Schneider and W. Weil, Stochastic and Integral Geom- l2 distance metric contains information of all moments etry (Probability and Its Applications) (Springer, Berlin, and higher moments less numerically robustly. By visual Germany, 2008). inspection, we found similar range of radii for which we [60] C. F. Gauss, “Besprechung des Buchs von L. A. Seeber: have reliable estimates for γ1, γ2, and l2. Untersuchungen ¨uber die Eigenschaften der positiven [79] The data will be published together with the paper. tern¨arenquadratischen formen,” G¨ottingsche Gelehrte [80] In a less detailed fluctuation study of the ideal gas, equi- Anzeigen (1831), see also J. reine angew. Math., vol. librium hard-spheres, and MRJ sphere packings, Klatt 20, 1840, 312-320. and Torquato [15] consistently found that the conver- [61] P. Sarnak and A. Str¨ombergsson, “Minima of Epstein’s gence of the number distributions to a Gaussian distribu- zeta function and heights of flat tori,” Inventiones Math. tion was slowest for the Poisson distribution and fastest for hyperuniform MRJ sphere packings.