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

On the lower bound for packing densities of superballs in high dimensions

Chengfei Xie  and Gennian Ge C. Xie is with the School of Mathematical Sciences, Capital Normal University, Beijing 100048, China (email: cfxie@cnu.edu.cn).Corresponding author. G. Ge is with the School of Mathematical Sciences, Capital Normal University, Beijing 100048, China (e-mail: gnge@zju.edu.cn). The research of G. Ge is supported by the National Key Research and Development Program of China under Grant Nos. 2020YFA0712100 and 2018YFA0704703, National Natural Science Foundation of China under Grant No. 11971325, and Beijing Scholars Program.
Abstract

Define the superball with radius rr and center 𝟎\bm{0} in n\mathbb{R}^{n} to be the set

{𝒙n:j=1m(xkj+12+xkj+22++xkj+12)p/2rp},0=k1<k2<<km+1=n,\left\{\bm{x}\in\mathbb{R}^{n}:\sum_{j=1}^{m}\left(x_{k_{j}+1}^{2}+x_{k_{j}+2}^{2}+\cdots+x_{k_{j+1}}^{2}\right)^{p/2}\leq r^{p}\right\},0=k_{1}<k_{2}<\cdots<k_{m+1}=n,

which is a generalization of p\ell_{p}-balls. We give two new proofs for the celebrated result that for 1<p21<p\leq 2, the translative packing density of superballs in n\mathbb{R}^{n} is Ω(n/2n)\Omega(n/2^{n}). This bound was first obtained by Schmidt, with subsequent constant factor improvement by Rogers and Schmidt, respectively. Our first proof is based on the hard superball model, and the second proof is based on the independence number of a graph. We also investigate the entropy of packings, which measures how plentiful such packings are.

Keywords: superball packing, hard superball model, uniform convexity, independence number

AMS subject classifications: 52C17, 05C15, 05C69

1 Introduction

The sphere packing problem asks the densest packing of nonoverlapping equal-sized balls in n\mathbb{R}^{n}. This is an old and difficult problem in discrete geometry. The exact answer is only known in dimensions 1,2,3,81,2,3,8, and 2424. In dimension 11, the problem is trivial; in dimension 22, the problem is nontrivial and can be solved by elementary geometry; in dimension 33, the problem is known as the Kepler’s conjecture and was solved by Hales [8]; in dimension 88, the problem was solved by Viazovska [31] by finding a function that matches the Cohn-Elkies linear programming bound [3]; in dimension 2424, the problem was solved by Cohn et al. [4].

Let Δ2(n)\Delta_{2}(n) be the maximum translative packing density of equal-sized Euclidean balls in n\mathbb{R}^{n}. The best upper bound of Δ2(n)\Delta_{2}(n) in high dimension is due to Kabatjanskiĭ and Levenšteĭn [12]: Δ2(n)2(0.599+o(1))n\Delta_{2}(n)\leq 2^{(-0.599\ldots+o(1))n}. See also Cohn and Zhao [5] and Sardari and Zargar [25] for constant factor improvement. A lower bound Δ2(n)2n\Delta_{2}(n)\geq 2^{-n} is trivial since doubling the radii of balls will cover the whole space. Rogers [16] improved the bound by a factor nn. Ball [1] constructed lattice packings of density 2(n1)2nζ(n)2(n-1)2^{-n}\zeta(n). The best known lower bound is (65963+o(1))n2n(65963+o(1))n2^{-n}, due to Venkatesh [30].

In this paper, we consider the packing density of superballs. Let kk\in\mathbb{N} and p1p\geq 1 be a real number. Let 2\|\cdot\|_{2} be the 2\ell_{2} norm; that is, for (x1,x2,,xk)k(x_{1},x_{2},\ldots,x_{k})\in\mathbb{R}^{k}, (x1,x2,,xk)2=x12+x22++xk2\|(x_{1},x_{2},\ldots,x_{k})\|_{2}=\sqrt{x_{1}^{2}+x_{2}^{2}+\cdots+x_{k}^{2}}. Let 𝒌=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) be such that 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n. Denote the superball with radius rr and center 𝒙\bm{x} in n\mathbb{R}^{n} by

Bp,𝒌n(𝒙,r)={𝒚n:(j=1m(xkj+1ykj+1,xkj+2ykj+2,,xkj+1ykj+1)2p)1/pr},B^{n}_{p,\bm{k}}(\bm{x},r)=\left\{\bm{y}\in\mathbb{R}^{n}:\left(\sum_{j=1}^{m}\left\|\left(x_{k_{j}+1}-y_{k_{j}+1},x_{k_{j}+2}-y_{k_{j}+2},\ldots,x_{k_{j+1}}-y_{k_{j+1}}\right)\right\|_{2}^{p}\right)^{1/p}\leq r\right\},

where 𝒙=(x1,x2,,xn)\bm{x}=(x_{1},x_{2},\ldots,x_{n}) and 𝒚=(y1,y2,,yn)\bm{y}=(y_{1},y_{2},\ldots,y_{n}). We simply write Bp,𝒌n(r)=Bp,𝒌n(𝟎,r)B^{n}_{p,\bm{k}}(r)=B^{n}_{p,\bm{k}}(\bm{0},r) if the superball is centered at 𝟎\bm{0}. Here, the role of 𝒌\bm{k} is to cut the vectors in n\mathbb{R}^{n}; that is, 𝒌\bm{k} cuts 𝒙\bm{x} and 𝒚\bm{y} in n\mathbb{R}^{n} into shorter vectors 𝒙j:=(xkj+1,xkj+2,,xkj+1)\bm{x}_{j}:=\left(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}}\right) and 𝒚j:=(ykj+1,ykj+2,,ykj+1)\bm{y}_{j}:=\left(y_{k_{j}+1},y_{k_{j}+2},\ldots,y_{k_{j+1}}\right), respectively. We first calculate the 2\ell_{2}-distance between 𝒙j\bm{x}_{j} and 𝒚j\bm{y}_{j}. And then we calculate the p\ell_{p} norm of the new vector formed by these 𝒙j𝒚j2\|\bm{x}_{j}-\bm{y}_{j}\|_{2}.

Throughout the paper, we assume the superballs of our packings have volume 11 and denote by rp,𝒌,nr_{p,\bm{k},n} the radius of a superball of volume 11. Let Δp,𝒌(n)\Delta_{p,\bm{k}}(n) denote the maximum translative packing density of copies of superballs of volume 11; that is,

Δp,𝒌(n)=lim supRsup𝒫vol(𝒫Bp,𝒌n(R))vol(Bp,𝒌n(R)),\Delta_{p,\bm{k}}(n)=\limsup_{R\rightarrow\infty}\sup_{\mathcal{P}}\frac{\textrm{vol}(\mathcal{P}\cap B_{p,\bm{k}}^{n}(R))}{\textrm{vol}(B_{p,\bm{k}}^{n}(R))},

where vol(𝒫Bp,𝒌n(R))\textrm{vol}(\mathcal{P}\cap B_{p,\bm{k}}^{n}(R)) is the volume of Bp,𝒌n(R)B_{p,\bm{k}}^{n}(R) covered by superballs of volume 11 with centers in 𝒫\mathcal{P}, and the supremum is over all 𝒫n\mathcal{P}\subseteq\mathbb{R}^{n} such that the superballs with centers in 𝒫\mathcal{P} are nonoverlapping.

Consider a special case that 𝒌=𝒌n:=(0,1,2,,n)\bm{k}=\bm{k}_{n}:=(0,1,2,\ldots,n). In this case,

Bp,𝒌nn(𝒙,r)={𝒚n:(j=1n|xjyj|p)1/pr}B^{n}_{p,\bm{k}_{n}}(\bm{x},r)=\left\{\bm{y}\in\mathbb{R}^{n}:\left(\sum_{j=1}^{n}\left|x_{j}-y_{j}\right|^{p}\right)^{1/p}\leq r\right\}

is the p\ell_{p}-ball in n\mathbb{R}^{n}. Let Δp(n):=Δp,𝒌n(n)\Delta_{p}(n):=\Delta_{p,\bm{k}_{n}}(n) be the packing density of p\ell_{p}-balls. The upper bound for Δp(n)\Delta_{p}(n) was first proved by van der Corput and Schaake [29], giving Δp(n)1+n/p2n/p\Delta_{p}(n)\leq\frac{1+n/p}{2^{n/p}} if p2p\geq 2 and Δp(n)1+(11/p)n2n/p\Delta_{p}(n)\leq\frac{1+(1-1/p)n}{2^{n/p}} if 1p21\leq p\leq 2. For p1.494p\geq 1.494\ldots, the upper bound was improved exponentially by Sah et al. [24]. The Minkowski-Hlawka theorem [9] gives a lower bound that Δp,𝒌(n)ζ(n)2n+1\Delta_{p,\bm{k}}(n)\geq\zeta(n)2^{-n+1}, where ζ(n)\zeta(n) is the Riemann zeta function. Rush and Sloane [23] improved the Minkowski-Hlawka bound for p\ell_{p}-balls and integers p3p\geq 3, for instance, Δ3(n)20.8226n+o(n)\Delta_{3}(n)\geq 2^{-0.8226\ldots n+o(n)}. Rush [19] constructed lattice packings with density 2n+o(n)2^{-n+o(n)} for every convex body which is symmetric through each of the coordinate hyperplanes. Moreover, Elkies et al. [6] improved the Minkowski-Hlawka bound exponentially for superballs and reals p>2p>2, and they also obtained lower bound for the packing density of more general bodies. For the lower bound constructed by error correcting codes, see Rush [20, 21, 22] and Liu and Xing [14].

We focus on the lower bound in the case that 1<p21<p\leq 2. In this case, only subexponential improvements of the Minkowski-Hlawka bound are known. Rogers [17] obtained a lower bound of Ω(n/2n)\Omega(\sqrt{n}/2^{n}). Schmidt [26] obtained a lower bound of Ω(n/2n)\Omega(n/2^{n}). See also Rogers [18] and Schmidt [27] for constant factor improvements. The best result is due to Schmidt [28], where the constant factor is less than log20.346\log\sqrt{2}\approx 0.346. We give new proofs for the lower bound Ω(n/2n)\Omega(n/2^{n}) of the packing density of p\ell_{p}-balls.

Theorem 1.1.

For every 1<p21<p\leq 2, there exists a constant cp(0,2)c_{p}\in(0,2) such that

Δp(n)(1+on(1))log(2/cp)n2n.\Delta_{p}(n)\geq(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}.

We also obtain a lower bound for Δp,𝒌(n)\Delta_{p,\bm{k}}(n).

Theorem 1.2.

For every p(1,2]p\in(1,2], there exists a constant cp(0,2)c_{p}\in(0,2) such that the following holds. For every 𝐤=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) with kj{0}k_{j}\in\mathbb{N}\cup\{0\} and 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n, we have

Δp,𝒌(n)(1+on(1))log(2/cp)n2n.\Delta_{p,\bm{k}}(n)\geq(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}.

Observe that Theorem 1.1 follows directly from Theorem 1.2 by setting 𝒌=(0,1,2,,n)\bm{k}=(0,1,2,\ldots,n).

The lower bound in Theorem 1.2 is independent of 𝒌\bm{k}. Note that different 𝒌\bm{k} will lead to different bodies in the packing. So this result means that, given 1<p21<p\leq 2 and letting nn be large, for every 𝒌=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) (kj{0}k_{j}\in\mathbb{N}\cup\{0\} and 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n), the translative packing density of Bp,𝒌n(𝟎,rp,𝒌,n)B^{n}_{p,\bm{k}}(\bm{0},r_{p,\bm{k},n}) is approximately larger than log(2/cp)n2n\frac{\log(2/c_{p})n}{2^{n}}. Throughout the paper, pp and 𝒌\bm{k} appear in the subscripts of most of our symbols and variables since they depend on pp and 𝒌\bm{k}. However, we always assume that pp is a given number and our bounds are independent of 𝒌\bm{k} (Theorem 1.2, Theorem 3.2, Theorem 5.1 and Theorem 5.2), so pp and 𝒌\bm{k} in the subscripts could be ignored for simplicity.

For p>2p>2, we can obtain a lower bound of Ωp(n/2n)\Omega_{p}(n/2^{n}) as well. But this bound is worse than the bound in [6].

We will give two proofs of Theorem 1.2. In Section 3, we give the first proof. The method we use is the hard superball model, which is developed from statistical physics. It was used by Jenssen et al. to prove the lower bound of kissing number [10] and packing density of Euclidean balls [11], where the methods are called hard cap model and hard sphere model, respectively. Recently, Fernández et al. [7] gave a constant factor improvement of results in [10] and [11]. In Section 4, we give the second proof, which uses the independence number of a graph. We also use the so-called uniform convexity to overcome the difficulty for the non-Euclideanness.

We also investigate the entropy density and pressure of packings, which we will define in Section 5. This measures how plentiful such packings are.

2 Uniformly convex spaces

Throughout the paper, we always assume that 𝒌\bm{k} is a (finite or infinite) sequence of strictly increasing nonnegative integers. Let p1p\geq 1. For 𝒌=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) such that 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n and 𝒙=(x1,x2,,xn)n\bm{x}=(x_{1},x_{2},\ldots,x_{n})\in\mathbb{R}^{n}, define

𝒙p,𝒌,n=(j=1m(xkj+1,xkj+2,,xkj+1)2p)1/p=(j=1m𝒙j2p)1/p,\|\bm{x}\|_{p,\bm{k},n}=\left(\sum_{j=1}^{m}\left\|\left(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}}\right)\right\|_{2}^{p}\right)^{1/p}=\left(\sum_{j=1}^{m}\|\bm{x}_{j}\|_{2}^{p}\right)^{1/p},

where 𝒙j=(xkj+1,xkj+2,,xkj+1)\bm{x}_{j}=(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}}). For 𝒌=(k1,k2,,km+1,)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1},\ldots) with 0=k1<k2<<km+1<0=k_{1}<k_{2}<\cdots<k_{m+1}<\cdots (obviously limjkj=\lim_{j\rightarrow\infty}k_{j}=\infty, since kjk_{j}’s are nonnegative integers) and 𝒙=(x1,x2,x3,)\bm{x}=(x_{1},x_{2},x_{3},\ldots), define

𝒙p,𝒌=(j=1(xkj+1,xkj+2,,xkj+1)2p)1/p=(j=1𝒙j2p)1/p,\|\bm{x}\|_{p,\bm{k}}=\left(\sum_{j=1}^{\infty}\left\|\left(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}}\right)\right\|_{2}^{p}\right)^{1/p}=\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}\|_{2}^{p}\right)^{1/p},

where 𝒙j=(xkj+1,xkj+2,,xkj+1)\bm{x}_{j}=(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}}). Let p,𝒌\ell_{p,\bm{k}} be the set consisting of all real sequences 𝒙\bm{x} satisfying 𝒙p,𝒌<\|\bm{x}\|_{p,\bm{k}}<\infty.

Proposition 2.1.

For every 𝐤=(k1,k2,,km+1,)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1},\ldots) with 0=k1<k2<<km+1<0=k_{1}<k_{2}<\cdots<k_{m+1}<\cdots, and for every p1p\geq 1, p,𝐤\ell_{p,\bm{k}} is a normed linear space equipped with the norm p,𝐤\|\cdot\|_{p,\bm{k}}.

Proof.

For every 𝒙=(x1,x2,),𝒚=(y1,y2,)p,𝒌\bm{x}=(x_{1},x_{2},\ldots),\bm{y}=(y_{1},y_{2},\ldots)\in\ell_{p,\bm{k}} and aa\in\mathbb{R}, let

𝒙j=(xkj+1,xkj+2,,xkj+1)\bm{x}_{j}=(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}})

and

𝒚j=(ykj+1,ykj+2,,ykj+1).\bm{y}_{j}=(y_{k_{j}+1},y_{k_{j}+2},\ldots,y_{k_{j+1}}).

We have

𝒙+𝒚p,𝒌=(j=1𝒙j+𝒚j2p)1/p(j=1(𝒙j2+𝒚j2)p)1/p(j=1𝒙j2p)1/p+(j=1𝒚j2p)1/p=𝒙p,𝒌+𝒚p,𝒌<,\begin{split}\|\bm{x}+\bm{y}\|_{p,\bm{k}}=&\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{p}\right)^{1/p}\\ \leq&\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}\|_{2}+\|\bm{y}_{j}\|_{2}\right)^{p}\right)^{1/p}\\ \leq&\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}\|_{2}^{p}\right)^{1/p}+\left(\sum_{j=1}^{\infty}\|\bm{y}_{j}\|_{2}^{p}\right)^{1/p}\\ =&\|\bm{x}\|_{p,\bm{k}}+\|\bm{y}\|_{p,\bm{k}}<\infty,\end{split} (1)

where the first ‘\leq’ follows from the triangle inequality of 2\|\cdot\|_{2} and the second ‘\leq’ follows from the Minkowski inequality

(j=1(Aj+Bj)s)1/s(j=1Ajs)1/s+(j=1Bjs)1/s\left(\sum_{j=1}^{\infty}\left(A_{j}+B_{j}\right)^{s}\right)^{1/s}\leq\left(\sum_{j=1}^{\infty}A_{j}^{s}\right)^{1/s}+\left(\sum_{j=1}^{\infty}B_{j}^{s}\right)^{1/s} (2)

for every nonnegative real sequences (Aj)j(A_{j})_{j\in\mathbb{N}} and (Bj)j(B_{j})_{j\in\mathbb{N}} and s1s\geq 1 (if 0<s10<s\leq 1, then inequality (2) is in the reverse sense). And

a𝒙p,𝒌=(j=1a𝒙j2p)1/p=(j=1ap𝒙j2p)1/p=|a|(j=1𝒙j2p)1/p<.\begin{split}\|a\bm{x}\|_{p,\bm{k}}=&\left(\sum_{j=1}^{\infty}\|a\bm{x}_{j}\|_{2}^{p}\right)^{1/p}\\ =&\left(\sum_{j=1}^{\infty}a^{p}\|\bm{x}_{j}\|_{2}^{p}\right)^{1/p}\\ =&|a|\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}\|_{2}^{p}\right)^{1/p}<\infty.\end{split} (3)

So p,𝒌\ell_{p,\bm{k}} is a linear space with the usual addition and scalar multiplication.

Clearly, 𝒙p,𝒌0\|\bm{x}\|_{p,\bm{k}}\geq 0 for every 𝒙p,𝒌\bm{x}\in\ell_{p,\bm{k}}, and if 𝒙p,𝒌=0\|\bm{x}\|_{p,\bm{k}}=0, then 𝒙j=𝟎\bm{x}_{j}=\bm{0} for every jj, i.e. 𝒙=𝟎\bm{x}=\bm{0}. We have already proved positive homogeneity (equation (3)) and triangle inequality (inequality (1)). We are done. ∎

For similar reasons, 𝒙p,𝒌,n\|\bm{x}\|_{p,\bm{k},n} is a norm on n\mathbb{R}^{n} as well. For 𝒙,𝒚n\bm{x},\bm{y}\in\mathbb{R}^{n}, let dp,𝒌,n(𝒙,𝒚):=𝒚𝒙p,𝒌,nd_{p,\bm{k},n}(\bm{x},\bm{y}):=\|\bm{y}-\bm{x}\|_{p,\bm{k},n} be the p,𝒌\ell_{p,\bm{k}}-distance between 𝒙\bm{x} and 𝒚\bm{y} in n\mathbb{R}^{n}. For sequences 𝒙,𝒚p,𝒌\bm{x},\bm{y}\in\ell_{p,\bm{k}}, let dp,𝒌(𝒙,𝒚):=𝒚𝒙p,𝒌d_{p,\bm{k}}(\bm{x},\bm{y}):=\|\bm{y}-\bm{x}\|_{p,\bm{k}} be the p,𝒌\ell_{p,\bm{k}}-distance between 𝒙\bm{x} and 𝒚\bm{y} in p,𝒌\ell_{p,\bm{k}}.

Definition 2.2.

A normed linear space equipped with norm \|\cdot\| is said to be uniformly convex if to each ϵ(0,2]\epsilon\in(0,2], there corresponds a δ(ϵ)>0\delta(\epsilon)>0 such that the conditions x=y=1\|x\|=\|y\|=1 and xyϵ\|x-y\|\geq\epsilon imply x+y21δ(ϵ)\left\|\frac{x+y}{2}\right\|\leq 1-\delta(\epsilon).

If we write 𝒌=(0,1,2,3,)\bm{k}_{\infty}=(0,1,2,3,\ldots), then the space p,𝒌\ell_{p,\bm{k}_{\infty}} is the usual p\ell^{p} space. And we have the following theorem.

Theorem 2.3 ([2]).

For p>1p>1, the p,𝐤\ell_{p,\bm{k}_{\infty}} space is uniformly convex. If 1<p21<p\leq 2 and δp,𝐤(ϵ)\delta_{p,\bm{k}_{\infty}}(\epsilon) is the constant with respect to the uniform convexity of p,𝐤\ell_{p,\bm{k}_{\infty}} space, then we can choose δp,𝐤(ϵ)=1(1(ϵ2)q)1/q\delta_{p,\bm{k}_{\infty}}(\epsilon)=1-\left(1-\left(\frac{\epsilon}{2}\right)^{q}\right)^{1/q}, where q=p/(p1)q=p/(p-1) is the conjugate index.

We have the following generalization of Theorem 2.3.

Proposition 2.4.

For every 𝐤=(k1,k2,,km+1,)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1},\ldots) with 0=k1<k2<<km+1<0=k_{1}<k_{2}<\cdots<k_{m+1}<\cdots, and for every p>1p>1, p,𝐤\ell_{p,\bm{k}} is uniformly convex. If 1<p21<p\leq 2 and δp(ϵ)\delta_{p}(\epsilon) is the constant with respect to the uniform convexity of p,𝐤\ell_{p,\bm{k}} space, then we can choose δp(ϵ)=1(1(ϵ2)q)1/q\delta_{p}(\epsilon)=1-\left(1-\left(\frac{\epsilon}{2}\right)^{q}\right)^{1/q}, where q=p/(p1)q=p/(p-1) is the conjugate index.

Remark 2.5.

We can choose δp(ϵ)\delta_{p}(\epsilon) independent of 𝐤\bm{k}.

The proof of Proposition 2.4 is similar to the proof of uniform convexity of p,𝒌\ell_{p,\bm{k}_{\infty}} in [2], so we first introduce the following lemma, which is a generalization of [2, Theorem 2].

Lemma 2.6.

For the space p,𝐤\ell_{p,\bm{k}} defined above, with p2p\geq 2, the following inequalities between the norms of two arbitrary elements 𝐱\bm{x} and 𝐲\bm{y} of the space are valid (Here qq is the conjugate index, q=p/(p1)q=p/(p-1)).

2(𝒙p,𝒌p+𝒚p,𝒌p)𝒙+𝒚p,𝒌p+𝒙𝒚p,𝒌p2p1(𝒙p,𝒌p+𝒚p,𝒌p);2\left(\|\bm{x}\|_{p,\bm{k}}^{p}+\|\bm{y}\|_{p,\bm{k}}^{p}\right)\leq\|\bm{x}+\bm{y}\|_{p,\bm{k}}^{p}+\|\bm{x}-\bm{y}\|_{p,\bm{k}}^{p}\leq 2^{p-1}\left(\|\bm{x}\|_{p,\bm{k}}^{p}+\|\bm{y}\|_{p,\bm{k}}^{p}\right); (4)
2(𝒙p,𝒌p+𝒚p,𝒌p)q1𝒙+𝒚p,𝒌q+𝒙𝒚p,𝒌q;2\left(\|\bm{x}\|_{p,\bm{k}}^{p}+\|\bm{y}\|_{p,\bm{k}}^{p}\right)^{q-1}\leq\|\bm{x}+\bm{y}\|_{p,\bm{k}}^{q}+\|\bm{x}-\bm{y}\|_{p,\bm{k}}^{q}; (5)
𝒙+𝒚p,𝒌p+𝒙𝒚p,𝒌p2(𝒙p,𝒌q+𝒚p,𝒌q)p1.\|\bm{x}+\bm{y}\|_{p,\bm{k}}^{p}+\|\bm{x}-\bm{y}\|_{p,\bm{k}}^{p}\leq 2\left(\|\bm{x}\|_{p,\bm{k}}^{q}+\|\bm{y}\|_{p,\bm{k}}^{q}\right)^{p-1}. (6)

For 1<p21<p\leq 2 these inequalities hold in the reverse sense.

Proof.

For simplicity, we write =p,𝒌\ell=\ell_{p,\bm{k}} and =p,𝒌\|\cdot\|=\|\cdot\|_{p,\bm{k}} in the proof. First note that for all values of pp the right hand side of inequality (4) is equivalent to the left hand side, while inequality (5) is equivalent to inequality (6); to see this, set 𝒙+𝒚=𝝃,𝒙𝒚=𝜼\bm{x}+\bm{y}=\bm{\xi},\bm{x}-\bm{y}=\bm{\eta} and we have

𝒙+𝒚p+𝒙𝒚p2p1(𝒙p+𝒚p)𝝃p+𝜼p2p1((𝝃+𝜼)/2p+(𝝃𝜼)/2p)𝝃p+𝜼p21(𝝃+𝜼p+𝝃𝜼p)2(𝝃p+𝜼p)𝝃+𝜼p+𝝃𝜼p,\begin{split}&\|\bm{x}+\bm{y}\|^{p}+\|\bm{x}-\bm{y}\|^{p}\leq 2^{p-1}(\|\bm{x}\|^{p}+\|\bm{y}\|^{p})\\ \Leftrightarrow&\|\bm{\xi}\|^{p}+\|\bm{\eta}\|^{p}\leq 2^{p-1}(\|(\bm{\xi}+\bm{\eta})/2\|^{p}+\|(\bm{\xi}-\bm{\eta})/2\|^{p})\\ \Leftrightarrow&\|\bm{\xi}\|^{p}+\|\bm{\eta}\|^{p}\leq 2^{-1}(\|\bm{\xi}+\bm{\eta}\|^{p}+\|\bm{\xi}-\bm{\eta}\|^{p})\\ \Leftrightarrow&2(\|\bm{\xi}\|^{p}+\|\bm{\eta}\|^{p})\leq\|\bm{\xi}+\bm{\eta}\|^{p}+\|\bm{\xi}-\bm{\eta}\|^{p},\end{split}

and

2(𝒙p+𝒚p)q1𝒙+𝒚q+𝒙𝒚q2((𝝃+𝜼)/2p+(𝝃𝜼)/2p)q1𝝃q+𝜼q21p(q1)(𝝃+𝜼p+𝝃𝜼p)q1𝝃q+𝜼q𝝃+𝜼p+𝝃𝜼p2(𝝃q+𝜼q)p1.\begin{split}&2(\|\bm{x}\|^{p}+\|\bm{y}\|^{p})^{q-1}\leq\|\bm{x}+\bm{y}\|^{q}+\|\bm{x}-\bm{y}\|^{q}\\ \Leftrightarrow&2(\|(\bm{\xi}+\bm{\eta})/2\|^{p}+\|(\bm{\xi}-\bm{\eta})/2\|^{p})^{q-1}\leq\|\bm{\xi}\|^{q}+\|\bm{\eta}\|^{q}\\ \Leftrightarrow&2^{1-p(q-1)}(\|\bm{\xi}+\bm{\eta}\|^{p}+\|\bm{\xi}-\bm{\eta}\|^{p})^{q-1}\leq\|\bm{\xi}\|^{q}+\|\bm{\eta}\|^{q}\\ \Leftrightarrow&\|\bm{\xi}+\bm{\eta}\|^{p}+\|\bm{\xi}-\bm{\eta}\|^{p}\leq 2(\|\bm{\xi}\|^{q}+\|\bm{\eta}\|^{q})^{p-1}.\end{split}

We first prove inequality (5) for 1<p21<p\leq 2, i.e.

2(𝒙p+𝒚p)q1𝒙+𝒚q+𝒙𝒚q.2(\|\bm{x}\|^{p}+\|\bm{y}\|^{p})^{q-1}\geq\|\bm{x}+\bm{y}\|^{q}+\|\bm{x}-\bm{y}\|^{q}. (7)

We claim that for 𝜶,𝜷k\bm{\alpha},\bm{\beta}\in\mathbb{R}^{k}, we have

2(𝜶2p+𝜷2p)q1𝜶+𝜷2q+𝜶𝜷2q.2\left(\|\bm{\alpha}\|_{2}^{p}+\|\bm{\beta}\|_{2}^{p}\right)^{q-1}\geq\|\bm{\alpha}+\bm{\beta}\|_{2}^{q}+\|\bm{\alpha}-\bm{\beta}\|_{2}^{q}. (8)

If one of 𝜶\bm{\alpha} and 𝜷\bm{\beta} is 𝟎\bm{0}, the inequality holds trivially. Assume that 𝜶2𝜷2>0\|\bm{\alpha}\|_{2}\geq\|\bm{\beta}\|_{2}>0. Divide by 𝜶2q\|\bm{\alpha}\|_{2}^{q} in the both sides of inequality (8) and use the homogeneity, reducing inequality (8) to

2(𝜶𝜶22p+𝜷𝜶22p)q1𝜶𝜶2+𝜷𝜶22q+𝜶𝜶2𝜷𝜶22q.2\left(\left\|\frac{\bm{\alpha}}{\|\bm{\alpha}\|_{2}}\right\|_{2}^{p}+\left\|\frac{\bm{\beta}}{\|\bm{\alpha}\|_{2}}\right\|_{2}^{p}\right)^{q-1}\geq\left\|\frac{\bm{\alpha}}{\|\bm{\alpha}\|_{2}}+\frac{\bm{\beta}}{\|\bm{\alpha}\|_{2}}\right\|_{2}^{q}+\left\|\frac{\bm{\alpha}}{\|\bm{\alpha}\|_{2}}-\frac{\bm{\beta}}{\|\bm{\alpha}\|_{2}}\right\|_{2}^{q}. (9)

Let 𝒂=𝜶𝜶2\bm{a}=\frac{\bm{\alpha}}{\|\bm{\alpha}\|_{2}} and 𝒃=𝜷𝜶2\bm{b}=\frac{\bm{\beta}}{\|\bm{\alpha}\|_{2}}. Then 𝒂2=1\|\bm{a}\|_{2}=1 and 𝒃21\|\bm{b}\|_{2}\leq 1, and inequality (9) is equivalent to

2(1+𝒃2p)q1𝒂+𝒃2q+𝒂𝒃2q,2\left(1+\left\|\bm{b}\right\|_{2}^{p}\right)^{q-1}\geq\left\|\bm{a}+\bm{b}\right\|_{2}^{q}+\left\|\bm{a}-\bm{b}\right\|_{2}^{q},

i.e.

2(1+𝒃2p)q1(𝒂22+𝒃22+2𝒂𝒃)q/2+(𝒂22+𝒃222𝒂𝒃)q/2,2\left(1+\left\|\bm{b}\right\|_{2}^{p}\right)^{q-1}\geq\left(\left\|\bm{a}\right\|_{2}^{2}+\left\|\bm{b}\right\|_{2}^{2}+2\bm{a}\cdot\bm{b}\right)^{q/2}+\left(\left\|\bm{a}\right\|_{2}^{2}+\left\|\bm{b}\right\|_{2}^{2}-2\bm{a}\cdot\bm{b}\right)^{q/2},

or equivalently,

2(1+𝒃2p)q1(1+𝒃22+2𝒂𝒃)q/2+(1+𝒃222𝒂𝒃)q/2,2\left(1+\left\|\bm{b}\right\|_{2}^{p}\right)^{q-1}\geq\left(1+\left\|\bm{b}\right\|_{2}^{2}+2\bm{a}\cdot\bm{b}\right)^{q/2}+\left(1+\left\|\bm{b}\right\|_{2}^{2}-2\bm{a}\cdot\bm{b}\right)^{q/2},

where 𝒂𝒃\bm{a}\cdot\bm{b} is the usual inner product. Suppose 𝒃2=b[0,1]\left\|\bm{b}\right\|_{2}=b\in[0,1], so

|𝒂𝒃|𝒂2𝒃2=b.|\bm{a}\cdot\bm{b}|\leq\left\|\bm{a}\right\|_{2}\left\|\bm{b}\right\|_{2}=b.

Without loss of generality assume that 𝒂𝒃[0,b]\bm{a}\cdot\bm{b}\in[0,b]. Consider the function

g(x)=(1+b2+2x)q/2+(1+b22x)q/2,x[0,b].g(x)=(1+b^{2}+2x)^{q/2}+(1+b^{2}-2x)^{q/2},x\in[0,b].

We have g(x)=q(1+b2+2x)q/21q(1+b22x)q/210g^{\prime}(x)=q(1+b^{2}+2x)^{q/2-1}-q(1+b^{2}-2x)^{q/2-1}\geq 0 for x[0,b]x\in[0,b], since q2q\geq 2. Thus maxg(x)=g(b)=(1+b2+2b)q/2+(1+b22b)q/2=(b+1)q+(1b)q\max g(x)=g(b)=(1+b^{2}+2b)^{q/2}+(1+b^{2}-2b)^{q/2}=(b+1)^{q}+(1-b)^{q} and it suffices to prove

2(1+bp)q1(1+b)q+(1b)q.2\left(1+b^{p}\right)^{q-1}\geq\left(1+b\right)^{q}+\left(1-b\right)^{q}. (10)

Inequality (10) has been proved in [2] (see the proof of [2, Theorem 2]). So inequality (8) is proved.

Now we are able to prove inequality (7). Let 𝒙=(𝒙1,𝒙2,)=(x1,x2,),𝒚=(𝒚1,𝒚2,)=(y1,y2,)\bm{x}=(\bm{x}_{1},\bm{x}_{2},\ldots)=(x_{1},x_{2},\ldots),\bm{y}=(\bm{y}_{1},\bm{y}_{2},\ldots)=(y_{1},y_{2},\ldots)\in\ell, where

𝒙j=(xkj+1,xkj+2,,xkj+1)\bm{x}_{j}=(x_{k_{j}+1},x_{k_{j}+2},\ldots,x_{k_{j+1}})

and

𝒚j=(ykj+1,ykj+2,,ykj+1).\bm{y}_{j}=(y_{k_{j}+1},y_{k_{j}+2},\ldots,y_{k_{j+1}}).

Inequality (7) states that

2(j=1(𝒙j2p+𝒚j2p))q1(j=1𝒙j+𝒚j2p)q/p+(j=1𝒙j𝒚j2p)q/p.2\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)\right)^{q-1}\geq\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{p}\right)^{q/p}+\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}-\bm{y}_{j}\|_{2}^{p}\right)^{q/p}. (11)

Using Minkowski inequality (2) in the reverse sense with Aj=𝒙j+𝒚j2q,Bj=𝒙j𝒚j2qA_{j}=\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{q},B_{j}=\|\bm{x}_{j}-\bm{y}_{j}\|_{2}^{q} and s=p/q1s=p/q\leq 1, we infer that the right hand side of inequality (11) is

(j=1Ajs)1/s+(j=1Bjs)1/s(j=1(Aj+Bj)s)1/s=(j=1(𝒙j+𝒚j2q+𝒙j𝒚j2q)p/q)q/p,\left(\sum_{j=1}^{\infty}A_{j}^{s}\right)^{1/s}+\left(\sum_{j=1}^{\infty}B_{j}^{s}\right)^{1/s}\leq\left(\sum_{j=1}^{\infty}\left(A_{j}+B_{j}\right)^{s}\right)^{1/s}=\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{q}+\|\bm{x}_{j}-\bm{y}_{j}\|_{2}^{q}\right)^{p/q}\right)^{q/p},

which by inequality (8) is

(j=1(2(𝒙j2p+𝒚j2p)q1)p/q)q/p=2(j=1(𝒙j2p+𝒚j2p))q/p.\leq\left(\sum_{j=1}^{\infty}\left(2\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)^{q-1}\right)^{p/q}\right)^{q/p}=2\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)\right)^{q/p}.

Since q/p=q1q/p=q-1, this is our result. It completes the proof of inequality (5) for 1<p21<p\leq 2.

Now we prove inequality (5) for p2p\geq 2, and we need to prove inequality (11) in the reverse sense, i.e.

2(j=1(𝒙j2p+𝒚j2p))q1(j=1𝒙j+𝒚j2p)q/p+(j=1𝒙j𝒚j2p)q/p.2\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)\right)^{q-1}\leq\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{p}\right)^{q/p}+\left(\sum_{j=1}^{\infty}\|\bm{x}_{j}-\bm{y}_{j}\|_{2}^{p}\right)^{q/p}. (12)

Letting Aj,BjA_{j},B_{j} and ss have the same values as above, and again applying Minkowski inequality (2), we conclude that the right hand side of inequality (12) is

(j=1Ajs)1/s+(j=1Bjs)1/s(j=1(Aj+Bj)s)1/s=(j=1(𝒙j+𝒚j2q+𝒙j𝒚j2q)p/q)q/p.\left(\sum_{j=1}^{\infty}A_{j}^{s}\right)^{1/s}+\left(\sum_{j=1}^{\infty}B_{j}^{s}\right)^{1/s}\geq\left(\sum_{j=1}^{\infty}\left(A_{j}+B_{j}\right)^{s}\right)^{1/s}=\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}+\bm{y}_{j}\|_{2}^{q}+\|\bm{x}_{j}-\bm{y}_{j}\|_{2}^{q}\right)^{p/q}\right)^{q/p}. (13)

Recall that inequality (8) holds for 1<p21<p\leq 2, and it is equivalent to

𝜶+𝜷2p+𝜶𝜷2p2(𝜶2q+𝜷2q)p1.\|\bm{\alpha}+\bm{\beta}\|_{2}^{p}+\|\bm{\alpha}-\bm{\beta}\|_{2}^{p}\geq 2\left(\|\bm{\alpha}\|_{2}^{q}+\|\bm{\beta}\|_{2}^{q}\right)^{p-1}.

After exchanging the role of pp and qq, we have

𝜶+𝜷2q+𝜶𝜷2q2(𝜶2p+𝜷2p)q1,\|\bm{\alpha}+\bm{\beta}\|_{2}^{q}+\|\bm{\alpha}-\bm{\beta}\|_{2}^{q}\geq 2\left(\|\bm{\alpha}\|_{2}^{p}+\|\bm{\beta}\|_{2}^{p}\right)^{q-1},

for p2p\geq 2. Thus inequality (13) is

(j=1(2(𝒙j2p+𝒚j2p)q1)p/q)q/p=2(j=1(𝒙j2p+𝒚j2p))q1.\geq\left(\sum_{j=1}^{\infty}\left(2\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)^{q-1}\right)^{p/q}\right)^{q/p}=2\left(\sum_{j=1}^{\infty}\left(\|\bm{x}_{j}\|_{2}^{p}+\|\bm{y}_{j}\|_{2}^{p}\right)\right)^{q-1}.

Therefore, we complete the proof of inequality (5) for p2p\geq 2.

Finally we prove inequality (4). Let p2p\geq 2 and consider the right hand inequality. This is simply implied by inequality (6): For x,y0x,y\geq 0, we have

2(xq+yq)p12p1(xp+yp).2(x^{q}+y^{q})^{p-1}\leq 2^{p-1}(x^{p}+y^{p}). (14)

Inequality (14) has been proved in [2] (see the proof of [2, Theorem 2]). For 1<p21<p\leq 2, it follows from inequality (14) that

2(xq+yq)p12p1(xp+yp).2(x^{q}+y^{q})^{p-1}\geq 2^{p-1}(x^{p}+y^{p}). (15)

The right hand side of inequality (4) (in the reverse sense) follows from inequality (6) (in the reverse sense) and inequality (15). ∎

Now we are able to prove Proposition 2.4.

Proof of Proposition 2.4.

For p2p\geq 2, letting 𝒙p,𝒌=𝒚p,𝒌=1\|\bm{x}\|_{p,\bm{k}}=\|\bm{y}\|_{p,\bm{k}}=1 in inequality (4), we have

𝒙+𝒚p,𝒌p+𝒙𝒚p,𝒌p2p.\|\bm{x}+\bm{y}\|_{p,\bm{k}}^{p}+\|\bm{x}-\bm{y}\|_{p,\bm{k}}^{p}\leq 2^{p}.

If 𝒙𝒚p,𝒌ϵ\|\bm{x}-\bm{y}\|_{p,\bm{k}}\geq\epsilon, then

𝒙+𝒚2p,𝒌(1(ϵ/2)p)1/p.\left\|\frac{\bm{x}+\bm{y}}{2}\right\|_{p,\bm{k}}\leq(1-(\epsilon/2)^{p})^{1/p}.

So we can choose δ(ϵ)=1(1(ϵ/2)p)1/p\delta(\epsilon)=1-(1-(\epsilon/2)^{p})^{1/p}. Similarly, for 1<p21<p\leq 2, we can choose δ(ϵ)=1(1(ϵ/2)q)1/q\delta(\epsilon)=1-(1-(\epsilon/2)^{q})^{1/q} by inequality (5). ∎

3 The hard superball model

Given p>1,np>1,n\in\mathbb{N} and 𝒌=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) with 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n, consider the translative packing of Bp,𝒌n(𝟎,rp,𝒌,n)B_{p,\bm{k}}^{n}(\bm{0},r_{p,\bm{k},n}). For a bounded, measurable subset SnS\subseteq\mathbb{R}^{n}, let

Pt,p,𝒌(S)={{𝒙1,𝒙2,,𝒙t}S:dp,𝒌,n(𝒙i,𝒙j)2rp,𝒌,n for every ij}.P_{t,p,\bm{k}}(S)=\{\{\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}\}\subseteq S:d_{p,\bm{k},n}(\bm{x}_{i},\bm{x}_{j})\geq 2r_{p,\bm{k},n}\text{ for every }i\neq j\}.

be the family consisting of unordered tt-tuples of points from SS that can form a packing.

The canonical hard superball model on SS with tt centers is simply a uniformly random tt-tuple Xt,p,𝒌Pt,p,𝒌(S)X_{t,p,\bm{k}}\in P_{t,p,\bm{k}}(S). The partition function of the canonical hard superball model on SS is the function

Z^S,p,𝒌(t)=1t!St1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙t,\hat{Z}_{S,p,\bm{k}}(t)=\frac{1}{t!}\int_{S^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}, (16)

where for 𝒙1,𝒙2,,𝒙tn\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}\in\mathbb{R}^{n}, the expression 𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}) denotes the event that dp,𝒌,n(𝒙i,𝒙j)2rp,𝒌,nd_{p,\bm{k},n}(\bm{x}_{i},\bm{x}_{j})\geq 2r_{p,\bm{k},n} for every iji\neq j.

The grand canonical hard superball model on SS at fugacity λ\lambda is a random set XX of unordered points, with XX distributed according to a Poisson point process of intensity λ\lambda conditioned on the event that dp,𝒌,n(𝒙,𝒚)2rp,𝒌,nd_{p,\bm{k},n}(\bm{x},\bm{y})\geq 2r_{p,\bm{k},n} for all distinct 𝒙,𝒚X\bm{x},\bm{y}\in X. In other words, we first choose tt from {0,1,2,}\{0,1,2,\ldots\} with probability \propto λtZ^S,p,𝒌(t)\lambda^{t}\hat{Z}_{S,p,\bm{k}}(t), then we independently, uniformly choose XX from Pt,p,𝒌(S)P_{t,p,\bm{k}}(S). The partition function of the grand canonical hard superball model on SS is

ZS,p,𝒌(λ)=t=0λtZ^S,p,𝒌(t),Z_{S,p,\bm{k}}(\lambda)=\sum_{t=0}^{\infty}\lambda^{t}\hat{Z}_{S,p,\bm{k}}(t), (17)

where we take Z^S,p,𝒌(0)=1\hat{Z}_{S,p,\bm{k}}(0)=1. Note that if SS is bounded then ZS,p,𝒌(λ)Z_{S,p,\bm{k}}(\lambda) is a polynomial in λ\lambda.

The expected packing density, αS,p,𝒌(λ)\alpha_{S,p,\bm{k}}(\lambda), of the hard superball model is simply the expected number of centers in SS normalized by the volume of SS; that is,

αS,p,𝒌(λ)=𝔼S,p,𝒌,λ|X|vol(S).\alpha_{S,p,\bm{k}}(\lambda)=\frac{\mathbb{E}_{S,p,\bm{k},\lambda}|X|}{\text{vol}(S)}.

Here and in what follows the notation S,p,𝒌,λ\mathbb{P}_{S,p,\bm{k},\lambda} and 𝔼S,p,𝒌,λ\mathbb{E}_{S,p,\bm{k},\lambda} indicates probabilities and expectations with respect to the grand canonical hard superball model on a region SS at fugacity λ\lambda.

The expected packing density can be expressed as the derivative of the normalized log partition function.

αS,p,𝒌(λ)=1vol(S)t=1tS,p,𝒌,λ(|X|=t)=1vol(S)t=1tλtZ^S,p,𝒌(t)ZS,p,𝒌(λ)=1vol(S)λ(ZS,p,𝒌(λ))ZS,p,𝒌(λ)=λvol(S)(logZS,p,𝒌(λ)).\begin{split}\alpha_{S,p,\bm{k}}(\lambda)&=\frac{1}{\text{vol}(S)}\sum_{t=1}^{\infty}t\cdot\mathbb{P}_{S,p,\bm{k},\lambda}(|X|=t)\\ &=\frac{1}{\text{vol}(S)}\sum_{t=1}^{\infty}\frac{t\cdot\lambda^{t}\hat{Z}_{S,p,\bm{k}}(t)}{Z_{S,p,\bm{k}}(\lambda)}\\ &=\frac{1}{\text{vol}(S)}\frac{\lambda\cdot(Z_{S,p,\bm{k}}(\lambda))^{\prime}}{Z_{S,p,\bm{k}}(\lambda)}\\ &=\frac{\lambda}{\text{vol}(S)}(\log Z_{S,p,\bm{k}}(\lambda))^{\prime}.\end{split} (18)

Here and in what follows logx\log x always denotes the natural logarithm of xx.

Lemma 3.1.

The asymptotic expected packing density of Bp,𝐤n(R)nB^{n}_{p,\bm{k}}(R)\subseteq\mathbb{R}^{n} is a lower bound on the maximum superball packing density. That is, for any λ>0\lambda>0,

Δp,𝒌(n)lim supRαBp,𝒌n(R),p,𝒌(λ).\Delta_{p,\bm{k}}(n)\geq\limsup_{R\rightarrow\infty}\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda).
Proof.

By the definition of Δp,𝒌(n)\Delta_{p,\bm{k}}(n),

Δp,𝒌(n)=lim supRsupX𝒫|X|(R/rp,𝒌,n)n,\Delta_{p,\bm{k}}(n)=\limsup_{R\rightarrow\infty}\sup_{X\in\mathcal{P}}\frac{|X|}{(R/r_{p,\bm{k},n})^{n}},

where 𝒫\mathcal{P} is the family consisting of all packings of Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R) by superballs of radius r=rp,𝒌,nr=r_{p,\bm{k},n}, and (R/rp,𝒌,n)n(R/r_{p,\bm{k},n})^{n} is the volume of Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R). Note that in our model, some centers may be near the boundary of Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R), which should be deleted from the packing. However, we can enlarge the radius of the superball to make it be a packing. In other words, if XX is chosen from the model on Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R), then XX is a packing of Bp,𝒌n(R+100)B^{n}_{p,\bm{k}}(R+100). So

Δp,𝒌(n)lim supR𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)((R+100)/rp,𝒌,n)n=lim supR𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)((R+100)/rp,𝒌,n)n((R+100)/rp,𝒌,n)n(R/rp,𝒌,n)n=lim supR𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)(R/rp,𝒌,n)n=lim supRαBp,𝒌n(R),p,𝒌(λ).\begin{split}\Delta_{p,\bm{k}}(n)&\geq\limsup_{R\rightarrow\infty}\frac{\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)}{((R+100)/r_{p,\bm{k},n})^{n}}\\ &=\limsup_{R\rightarrow\infty}\frac{\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)}{((R+100)/r_{p,\bm{k},n})^{n}}\cdot\frac{((R+100)/r_{p,\bm{k},n})^{n}}{(R/r_{p,\bm{k},n})^{n}}\\ &=\limsup_{R\rightarrow\infty}\frac{\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)}{(R/r_{p,\bm{k},n})^{n}}\\ &=\limsup_{R\rightarrow\infty}\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda).\end{split}

Theorem 3.2.

For every p(1,2]p\in(1,2], there exists a constant cp(0,2)c_{p}\in(0,2) such that the following holds. Let SnS\subseteq\mathbb{R}^{n} be bounded, measurable, and of positive volume. Then for every 𝐤=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) such that 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n and every λn1cpn\lambda\geq n^{-1}c_{p}^{-n}, we have

αS,p,𝒌(λ)(1+on(1))log(2/cp)n2n.\alpha_{S,p,\bm{k}}(\lambda)\geq(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}.
Remark 3.3.

The constant cpc_{p} in Theorem 3.2 is independent of nn and 𝐤\bm{k}.

Theorem 1.2 follows immediately from Lemma 3.1 and Theorem 3.2.

Lemma 3.4.

Let SnS\subseteq\mathbb{R}^{n} be bounded, measurable, and of positive volume. Then the expected packing density αS,p,𝐤(λ)\alpha_{S,p,\bm{k}}(\lambda) is strictly increasing in λ\lambda.

Proof.

We use equation (18) and calculate

λvol(S)αS,p,𝒌(λ)=λ(λ(ZS,p,𝒌(λ))ZS,p,𝒌(λ))=λ((ZS,p,𝒌(λ))+λ(ZS,p,𝒌(λ))′′ZS,p,𝒌(λ)λ[(ZS,p,𝒌(λ))]2(ZS,p,𝒌(λ))2).\begin{split}\lambda\cdot{\rm vol}(S)\cdot\alpha_{S,p,\bm{k}}^{\prime}(\lambda)&=\lambda\cdot\left(\frac{\lambda\cdot(Z_{S,p,\bm{k}}(\lambda))^{\prime}}{Z_{S,p,\bm{k}}(\lambda)}\right)^{\prime}\\ &=\lambda\cdot\left(\frac{(Z_{S,p,\bm{k}}(\lambda))^{\prime}+\lambda\cdot(Z_{S,p,\bm{k}}(\lambda))^{\prime\prime}}{Z_{S,p,\bm{k}}(\lambda)}-\frac{\lambda\cdot\left[(Z_{S,p,\bm{k}}(\lambda))^{\prime}\right]^{2}}{\left(Z_{S,p,\bm{k}}(\lambda)\right)^{2}}\right).\end{split}

Since

λ(ZS,p,𝒌(λ))ZS,p,𝒌(λ)=vol(S)αS,p,𝒌=𝔼S,p,𝒌,λ|X|\lambda\cdot\frac{(Z_{S,p,\bm{k}}(\lambda))^{\prime}}{Z_{S,p,\bm{k}}(\lambda)}={\rm vol}(S)\cdot\alpha_{S,p,\bm{k}}=\mathbb{E}_{S,p,\bm{k},\lambda}|X|

and

λλ(ZS,p,𝒌(λ))′′ZS,p,𝒌(λ)=λ2t=2t(t1)λt2Z^S,p,𝒌(t)ZS,p,𝒌(λ)=t=2t(t1)S,p,𝒌,λ(|X|=t)=𝔼S,p,𝒌,λ[|X|(|X|1)],\lambda\cdot\frac{\lambda\cdot(Z_{S,p,\bm{k}}(\lambda))^{\prime\prime}}{Z_{S,p,\bm{k}}(\lambda)}=\lambda^{2}\cdot\frac{\sum_{t=2}^{\infty}t(t-1)\lambda^{t-2}\hat{Z}_{S,p,\bm{k}}(t)}{Z_{S,p,\bm{k}}(\lambda)}=\sum_{t=2}^{\infty}t(t-1)\mathbb{P}_{S,p,\bm{k},\lambda}(|X|=t)=\mathbb{E}_{S,p,\bm{k},\lambda}[|X|(|X|-1)],

it follows that

λvol(S)αS,p,𝒌(λ)=𝔼S,p,𝒌,λ|X|+𝔼S,p,𝒌,λ[|X|(|X|1)](𝔼S,p,𝒌,λ|X|)2=Var(|X|)>0.\lambda\cdot{\rm vol}(S)\cdot\alpha_{S,p,\bm{k}}^{\prime}(\lambda)=\mathbb{E}_{S,p,\bm{k},\lambda}|X|+\mathbb{E}_{S,p,\bm{k},\lambda}[|X|(|X|-1)]-\left(\mathbb{E}_{S,p,\bm{k},\lambda}|X|\right)^{2}={\rm Var}(|X|)>0. (19)

Thus αS,p,𝒌(λ)\alpha_{S,p,\bm{k}}(\lambda) is strictly increasing. ∎

Let FVS,p,𝒌(λ)\text{FV}_{S,p,\bm{k}}(\lambda) denote the expected free volume of the hard superball model; that is, the expected fraction of the volume of SS containing points that are at distance at least 2rp,𝒌,n2r_{p,\bm{k},n} from the nearest center. That is,

FVS,p,𝒌(λ)=𝔼S,p,𝒌,λ[vol({𝒚S:dp,𝒌,n(𝒚,𝒙)2rp,𝒌,n for every 𝒙X})]vol(S).\text{FV}_{S,p,\bm{k}}(\lambda)=\frac{\mathbb{E}_{S,p,\bm{k},\lambda}\left[\text{vol}(\{\bm{y}\in S:d_{p,\bm{k},n}(\bm{y},\bm{x})\geq 2r_{p,\bm{k},n}\text{ for every }\bm{x}\in X\})\right]}{\text{vol}(S)}.
Lemma 3.5.

Let SnS\subseteq\mathbb{R}^{n} be bounded, measurable, and of positive volume. Then

αS,p,𝒌(λ)=λFVS,p,𝒌(λ).\alpha_{S,p,\bm{k}}(\lambda)=\lambda\cdot{\rm FV}_{S,p,\bm{k}}(\lambda).
Proof.

We use the definitions of αS,p,𝒌(λ)\alpha_{S,p,\bm{k}}(\lambda) and FVS,p,𝒌(λ){\rm FV}_{S,p,\bm{k}}(\lambda).

αS,p,𝒌(λ)=𝔼S,p,𝒌,λ|X|vol(S)=1vol(S)t=0(t+1)S,p,𝒌,λ(|X|=t+1)=1vol(S)ZS,p,𝒌(λ)t=0(t+1)St+1λt+1(t+1)!1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0𝑑𝒙1𝑑𝒙t=λvol(S)ZS,p,𝒌(λ)t=0St+1λtt!1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0𝑑𝒙1𝑑𝒙t=λvol(S)ZS,p,𝒌(λ)t=0Stλtt!(S1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0)𝑑𝒙1𝑑𝒙t.\begin{split}\alpha_{S,p,\bm{k}}(\lambda)&=\frac{\mathbb{E}_{S,p,\bm{k},\lambda}|X|}{\text{vol}(S)}\\ &=\frac{1}{\text{vol}(S)}\sum_{t=0}^{\infty}(t+1)\cdot\mathbb{P}_{S,p,\bm{k},\lambda}(|X|=t+1)\\ &=\frac{1}{\text{vol}(S)Z_{S,p,\bm{k}}(\lambda)}\sum_{t=0}^{\infty}(t+1)\cdot\int_{S^{t+1}}\frac{\lambda^{t+1}}{(t+1)!}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}d\bm{x}_{1}\cdots d\bm{x}_{t}\\ &=\frac{\lambda}{\text{vol}(S)Z_{S,p,\bm{k}}(\lambda)}\sum_{t=0}^{\infty}\int_{S^{t+1}}\frac{\lambda^{t}}{t!}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}d\bm{x}_{1}\cdots d\bm{x}_{t}\\ &=\frac{\lambda}{\text{vol}(S)Z_{S,p,\bm{k}}(\lambda)}\sum_{t=0}^{\infty}\int_{S^{t}}\frac{\lambda^{t}}{t!}\left(\int_{S}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}\right)d\bm{x}_{1}\cdots d\bm{x}_{t}.\end{split}

Let Y=vol({𝒚S:dp,𝒌,n(𝒚,𝒙)2rp,𝒌,n for every 𝒙X})Y=\text{vol}(\{\bm{y}\in S:d_{p,\bm{k},n}(\bm{y},\bm{x})\geq 2r_{p,\bm{k},n}\text{ for every }\bm{x}\in X\}). Then

𝔼S,p,𝒌,λ(Y|X={𝒙1,𝒙2,,𝒙t})=S1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0.\mathbb{E}_{S,p,\bm{k},\lambda}\left(Y|X=\{\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}\}\right)=\int_{S}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}.

So

λFVS,p,𝒌(λ)=λvol(S)𝔼S,p,𝒌,λ(Y)=λvol(S)𝔼S,p,𝒌,λ[𝔼S,p,𝒌,λ(Y|X={𝒙1,𝒙2,,𝒙t})]=λvol(S)t=0λtZS,p,𝒌(λ)1t!St(S1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0)1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙t=λvol(S)ZS,p,𝒌(λ)t=0Stλtt!(S1𝒟p,𝒌(𝒙0,𝒙1,𝒙2,,𝒙t)𝑑𝒙0)𝑑𝒙1𝑑𝒙t=αS,p,𝒌(λ),\begin{split}&\lambda\cdot\text{FV}_{S,p,\bm{k}}(\lambda)\\ =&\frac{\lambda}{\text{vol}(S)}\cdot\mathbb{E}_{S,p,\bm{k},\lambda}\left(Y\right)\\ =&\frac{\lambda}{\text{vol}(S)}\cdot\mathbb{E}_{S,p,\bm{k},\lambda}\left[\mathbb{E}_{S,p,\bm{k},\lambda}\left(Y|X=\{\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}\}\right)\right]\\ =&\frac{\lambda}{\text{vol}(S)}\sum_{t=0}^{\infty}\frac{\lambda^{t}}{Z_{S,p,\bm{k}}(\lambda)}\frac{1}{t!}\int_{S^{t}}\left(\int_{S}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}\right)\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}\cdots d\bm{x}_{t}\\ =&\frac{\lambda}{\text{vol}(S)Z_{S,p,\bm{k}}(\lambda)}\sum_{t=0}^{\infty}\int_{S^{t}}\frac{\lambda^{t}}{t!}\left(\int_{S}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{0},\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{0}\right)d\bm{x}_{1}\cdots d\bm{x}_{t}\\ =&\alpha_{S,p,\bm{k}}(\lambda),\end{split}

where for t=0t=0, 1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)1\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}\equiv 1 (since Z^S,p,𝒌(0)=1\hat{Z}_{S,p,\bm{k}}(0)=1). ∎

Now consider the following two-part experiment: sample a configuration of centers XX from the hard superball model on SS at fugacity λ\lambda and independently choose a point 𝒗\bm{v} from SS. We define the random set

T={𝒙Bp,𝒌n(𝒗,2rp,𝒌,n)S:dp,𝒌,n(𝒙,𝒚)2rp,𝒌,n for every 𝒚XBp,𝒌n(𝒗,2rp,𝒌,n)c}.\text{T}=\{\bm{x}\in B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n})\cap S:d_{p,\bm{k},n}(\bm{x},\bm{y})\geq 2r_{p,\bm{k},n}\text{ for every }\bm{y}\in X\cap B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n})^{c}\}.

That is, T is the set of all points of SS in Bp,𝒌n(𝒗,2rp,𝒌,n)B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n}) that are not blocked from being a center by a center outside of Bp,𝒌n(𝒗,2rp,𝒌,n)B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n}).

Lemma 3.6.

Let SnS\subseteq\mathbb{R}^{n} be bounded, measurable, and of positive volume. Then

αS,p,𝒌(λ)=λ𝔼[1ZT,p,𝒌(λ)]\alpha_{S,p,\bm{k}}(\lambda)=\lambda\cdot\mathbb{E}\left[\frac{1}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right] (20)

and

αS,p,𝒌(λ)2n𝔼[λ(ZT,p,𝒌(λ))ZT,p,𝒌(λ)],\alpha_{S,p,\bm{k}}(\lambda)\geq 2^{-n}\cdot\mathbb{E}\left[\frac{\lambda\cdot(Z_{{\rm T},p,\bm{k}}(\lambda))^{\prime}}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right], (21)

where both expectations are with respect to the random set T generated by the two-part experiment defined above.

Proof.

Using Lemma 3.5, we have

αS,p,𝒌(λ)=λFVS,p,𝒌(λ)=λvol(S)S[dp,𝒌,n(𝒙,𝒚)2rp,𝒌,n,𝒙X]𝑑𝒚=λ𝔼(𝟏TX=)=λ𝔼[1ZT,p,𝒌(λ)],\begin{split}\alpha_{S,p,\bm{k}}(\lambda)&=\lambda\cdot{\rm FV}_{S,p,\bm{k}}(\lambda)\\ &=\frac{\lambda}{{\rm vol}(S)}\cdot\int_{S}\mathbb{P}[d_{p,\bm{k},n}(\bm{x},\bm{y})\geq 2r_{p,\bm{k},n},\forall\bm{x}\in X]d\bm{y}\\ &=\lambda\cdot\mathbb{E}(\mathbf{1}_{{\rm T}\cap X=\emptyset})\\ &=\lambda\cdot\mathbb{E}\left[\frac{1}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right],\end{split}

where the last equation uses the spatial Markov property of the hard superball model: conditioned on XBp,𝒌n(𝒗,2rp,𝒌,n)cX\cap B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n})^{c}, the distribution of XBp,𝒌n(𝒗,2rp,𝒌,n)X\cap B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n}) is exactly that of the hard superball model on the set T{\rm T}.

For every 𝒙S\bm{x}\in S, (𝒙Bp,𝒌n(𝒗,2rp,𝒌,n))=vol(Bp,𝒌n(𝒗,2rp,𝒌,n)S)/vol(S)2n/vol(S)\mathbb{P}(\bm{x}\in B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n}))={\rm vol}(B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n})\cap S)/{\rm vol}(S)\leq 2^{n}/{\rm vol}(S). So

αS,p,𝒌(λ)=𝔼S,p,𝒌,λ|X|vol(S)2n𝔼(|XBp,𝒌n(𝒗,2rp,𝒌,n)|)=2n𝔼(αT,p,𝒌(λ)vol(T))=2n𝔼[λ(ZT,p,𝒌(λ))ZT,p,𝒌(λ)].\begin{split}\alpha_{S,p,\bm{k}}(\lambda)&=\frac{\mathbb{E}_{S,p,\bm{k},\lambda}|X|}{\text{vol}(S)}\\ &\geq 2^{-n}\mathbb{E}(|X\cap B^{n}_{p,\bm{k}}(\bm{v},2r_{p,\bm{k},n})|)\\ &=2^{-n}\mathbb{E}(\alpha_{{\rm T},p,\bm{k}}(\lambda)\cdot{\rm vol(T)})\\ &=2^{-n}\cdot\mathbb{E}\left[\frac{\lambda\cdot(Z_{{\rm T},p,\bm{k}}(\lambda))^{\prime}}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right].\end{split}

Lemma 3.7.

Let SnS\subseteq\mathbb{R}^{n} be bounded and measurable. Then

logZS,p,𝒌(λ)λvol(S)\log Z_{S,p,\bm{k}}(\lambda)\leq\lambda\cdot{\rm vol}(S) (22)

and if in addition SS is of positive volume, then

αS,p,𝒌(λ)λeλ𝔼[vol(T)].\alpha_{S,p,\bm{k}}(\lambda)\geq\lambda\cdot e^{-\lambda\cdot\mathbb{E}\left[\rm vol(T)\right]}. (23)
Proof.

By the definition of ZS,p,𝒌(λ)Z_{S,p,\bm{k}}(\lambda), we have

ZS,p,𝒌(λ)=t=0λtZ^S,p,𝒌(t)=t=0λtt!St1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙tt=0λtt!St𝑑𝒙1𝑑𝒙2𝑑𝒙t=t=0λtt!(vol(S))t=eλvol(S).\begin{split}Z_{S,p,\bm{k}}(\lambda)&=\sum_{t=0}^{\infty}\lambda^{t}\hat{Z}_{S,p,\bm{k}}(t)\\ &=\sum_{t=0}^{\infty}\frac{\lambda^{t}}{t!}\int_{S^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}\\ &\leq\sum_{t=0}^{\infty}\frac{\lambda^{t}}{t!}\int_{S^{t}}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}\\ &=\sum_{t=0}^{\infty}\frac{\lambda^{t}}{t!}({\rm vol}(S))^{t}\\ &=e^{\lambda\cdot{\rm vol}(S)}.\end{split}

Take logarithm and we obtain inequality (22).

For inequality (23), we use equation (20) and inequality (22). So

αS,p,𝒌(λ)=λ𝔼[1ZT,p,𝒌(λ)]λ𝔼[eλvol(T)]λeλ𝔼[vol(T)],\begin{split}\alpha_{S,p,\bm{k}}(\lambda)&=\lambda\cdot\mathbb{E}\left[\frac{1}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right]\\ &\geq\lambda\cdot\mathbb{E}\left[e^{-\lambda\cdot{\rm vol(T)}}\right]\\ &\geq\lambda\cdot e^{-\lambda\cdot\mathbb{E}\left[{\rm vol(T)}\right]},\end{split}

where the last inequality follows from Jensen’s inequality. ∎

Remark 3.8.

In [11], Lemmas 3.4-3.7 are with respect to the packing of Euclidean balls. Indeed, these lemmas are valid for the packing of superballs defined above as well.

Consider the function

h(x)=(x4+121x)q+(x+24)q1.h(x)=\left(\frac{x}{4}+\frac{1}{2}-\frac{1}{x}\right)^{q}+\left(\frac{x+2}{4}\right)^{q}-1.

Note that h(x)h(x) is continuous for x1.5x\geq 1.5 and h(2)=1/2q>0h(2)=1/2^{q}>0, so there exists xp(1.5,2)x_{p}\in(1.5,2) such that h(x)>1/3qh(x)>1/3^{q} for every x[xp,2]x\in[x_{p},2]. We will use xpx_{p} in the proof of the following lemma.

Lemma 3.9.

For every p(1,2]p\in(1,2], there exists a constant cp(0,2)c_{p}\in(0,2) such that the following holds. For every nn\in\mathbb{N} and 𝐤=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) with 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n, let SBp,𝐤n(2rp,𝐤,n)S\subseteq B^{n}_{p,\bm{k}}(2r_{p,\bm{k},n}) be measurable. Then

𝔼[vol(Bp,𝒌n(𝒖,2rp,𝒌,n)S)]2cpn,\mathbb{E}\left[{\rm vol}(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap S)\right]\leq 2\cdot c_{p}^{n}, (24)

where 𝐮\bm{u} is a uniformly chosen point in SS. In particular,

αS,p,𝒌λeλ2cpn.\alpha_{S,p,\bm{k}}\geq\lambda\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}. (25)
Remark 3.10.

The constant cpc_{p} in Theorem 3.2 will be chosen as the same as the constant cpc_{p} here. So we use the same symbol.

Proof.

Clearly, we may assume that SS has positive volume. We have

𝔼[vol(Bp,𝒌n(𝒖,2rp,𝒌,n)S)]=1vol(S)SS𝟏dp,𝒌,n(𝒖,𝒗)2rp,𝒌,n𝑑𝒗𝑑𝒖=2vol(S)SS𝟏dp,𝒌,n(𝒖,𝒗)2rp,𝒌,n𝟏vp,𝒌,nup,𝒌,n𝑑𝒗𝑑𝒖2max𝒖Bp,𝒌n(2rp,𝒌,n)S𝟏dp,𝒌,n(𝒖,𝒗)2rp,𝒌,n𝟏vp,𝒌,nup,𝒌,n𝑑𝒗2max𝒖Bp,𝒌n(2rp,𝒌,n)vol(Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n)).\begin{split}\mathbb{E}\left[{\rm vol}(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap S)\right]&=\frac{1}{{\rm vol}(S)}\int_{S}\int_{S}\mathbf{1}_{d_{p,\bm{k},n}(\bm{u},\bm{v})\leq 2r_{p,\bm{k},n}}d\bm{v}d\bm{u}\\ &=\frac{2}{{\rm vol}(S)}\int_{S}\int_{S}\mathbf{1}_{d_{p,\bm{k},n}(\bm{u},\bm{v})\leq 2r_{p,\bm{k},n}}\cdot\mathbf{1}_{\|v\|_{p,\bm{k},n}\leq\|u\|_{p,\bm{k},n}}d\bm{v}d\bm{u}\\ &\leq 2\max_{\bm{u}\in B^{n}_{p,\bm{k}}(2r_{p,\bm{k},n})}\int_{S}\mathbf{1}_{d_{p,\bm{k},n}(\bm{u},\bm{v})\leq 2r_{p,\bm{k},n}}\cdot\mathbf{1}_{\|v\|_{p,\bm{k},n}\leq\|u\|_{p,\bm{k},n}}d\bm{v}\\ &\leq 2\max_{\bm{u}\in B^{n}_{p,\bm{k}}(2r_{p,\bm{k},n})}{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\right).\end{split}

In order to prove inequality (24), it suffices to prove that there exists a constant cp(0,2)c_{p}\in(0,2) such that for every n,𝒌n,\bm{k} and for every 𝒖Bp,𝒌n(𝟎,2rp,𝒌,n)\bm{u}\in B^{n}_{p,\bm{k}}(\bm{0},2r_{p,\bm{k},n}), we have

vol(Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n))cpn.{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\right)\leq c_{p}^{n}.

For every n,𝒌n,\bm{k} and for every 𝒖Bp,𝒌n(𝟎,2rp,𝒌,n)\bm{u}\in B^{n}_{p,\bm{k}}(\bm{0},2r_{p,\bm{k},n}), if 𝒖p,𝒌,nxprp,𝒌,n\|\bm{u}\|_{p,\bm{k},n}\leq x_{p}\cdot r_{p,\bm{k},n}, then

vol(Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n))vol(Bp,𝒌n(𝟎,𝒖p,𝒌,n))xpn.{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\right)\leq{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\right)\leq x_{p}^{n}.

Next we consider the case that 𝒖p,𝒌,nxprp,𝒌,n\|\bm{u}\|_{p,\bm{k},n}\geq x_{p}\cdot r_{p,\bm{k},n}. For every 𝒙Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n)\bm{x}\in B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n}), we have 𝒙𝒖p,𝒌,n2rp,𝒌,n\|\bm{x}-\bm{u}\|_{p,\bm{k},n}\leq 2r_{p,\bm{k},n} and 𝒙p,𝒌,n𝒖p,𝒌,n2rp,𝒌,n\|\bm{x}\|_{p,\bm{k},n}\leq\|\bm{u}\|_{p,\bm{k},n}\leq 2r_{p,\bm{k},n}. If 𝒙𝒖p,𝒌,nxprp,𝒌,n\|\bm{x}-\bm{u}\|_{p,\bm{k},n}\leq x_{p}\cdot r_{p,\bm{k},n} or 𝒙p,𝒌,nxprp,𝒌,n\|\bm{x}\|_{p,\bm{k},n}\leq x_{p}\cdot r_{p,\bm{k},n}, then by properties of a norm (the homogeneity condition and the triangle inequality), we have

𝒙12𝒖p,𝒌,n=122𝒙𝒖p,𝒌,n12(𝒙𝒖p,𝒌,n+𝒙p,𝒌,n)xp+22rp,𝒌,n.\left\|\bm{x}-\frac{1}{2}\bm{u}\right\|_{p,\bm{k},n}=\frac{1}{2}\left\|2\bm{x}-\bm{u}\right\|_{p,\bm{k},n}\leq\frac{1}{2}\left(\left\|\bm{x}-\bm{u}\right\|_{p,\bm{k},n}+\left\|\bm{x}\right\|_{p,\bm{k},n}\right)\leq\frac{x_{p}+2}{2}r_{p,\bm{k},n}. (26)

Now suppose that 𝒙𝒖p,𝒌,nxprp,𝒌,n\|\bm{x}-\bm{u}\|_{p,\bm{k},n}\geq x_{p}\cdot r_{p,\bm{k},n} and 𝒙p,𝒌,nxprp,𝒌,n\|\bm{x}\|_{p,\bm{k},n}\geq x_{p}\cdot r_{p,\bm{k},n}. Fix 1<p21<p\leq 2. Let δp(ϵ)=1(1(ϵ2)q)1/q\delta_{p}(\epsilon)=1-\left(1-\left(\frac{\epsilon}{2}\right)^{q}\right)^{1/q}. By Proposition 2.4, for every 𝒌=(k1,k2,)\bm{k}=(k_{1},k_{2},\ldots) and for every 𝒙,𝒚p,𝒌\bm{x},\bm{y}\in\ell_{p,\bm{k}}, if 𝒙p,𝒌=𝒚p,𝒌=1\|\bm{x}\|_{p,\bm{k}}=\|\bm{y}\|_{p,\bm{k}}=1 and 𝒙𝒚p,𝒌ϵ\|\bm{x}-\bm{y}\|_{p,\bm{k}}\geq\epsilon, then 𝒙+𝒚2p,𝒌1δp(ϵ)\left\|\frac{\bm{x}+\bm{y}}{2}\right\|_{p,\bm{k}}\leq 1-\delta_{p}(\epsilon). For every nn\in\mathbb{N} and 𝒌=(k1,k2,,km+1)\bm{k}=(k_{1},k_{2},\ldots,k_{m+1}) with 0=k1<k2<<km+1=n0=k_{1}<k_{2}<\cdots<k_{m+1}=n, we view a point 𝒙n\bm{x}\in\mathbb{R}^{n} as a point in p,𝒌~\ell_{p,\tilde{\bm{k}}} by adding zeros after the nn-th coordinate of 𝒙\bm{x} and 𝒌~=(k1,k2,,km+1,n+1,n+2,)\tilde{\bm{k}}=(k_{1},k_{2},\ldots,k_{m+1},n+1,n+2,\ldots), so 𝒙p,𝒌,n=𝒙p,𝒌~\|\bm{x}\|_{p,\bm{k},n}=\|\bm{x}\|_{p,\tilde{\bm{k}}}. In the rest of proof, we use norm p,𝒌\|\cdot\|_{p,\bm{k}} instead of p,𝒌,n\|\cdot\|_{p,\bm{k},n}.

Let 𝒙𝒖p,𝒌=a\|\bm{x}-\bm{u}\|_{p,\bm{k}}=a, 𝒙p,𝒌=b\|\bm{x}\|_{p,\bm{k}}=b, and 𝒖p,𝒌=c\|\bm{u}\|_{p,\bm{k}}=c. We have known that xprp,𝒌,na2rp,𝒌,nx_{p}\cdot r_{p,\bm{k},n}\leq a\leq 2r_{p,\bm{k},n} and xprp,𝒌,nbc2rp,𝒌,nx_{p}\cdot r_{p,\bm{k},n}\leq b\leq c\leq 2r_{p,\bm{k},n}. Since 𝒙𝒖ap,𝒌=𝒙bp,𝒌=1\left\|\frac{\bm{x}-\bm{u}}{a}\right\|_{p,\bm{k}}=\left\|\frac{\bm{x}}{b}\right\|_{p,\bm{k}}=1 and

𝒙𝒖a𝒙bp,𝒌=𝒖a+(1a1b)𝒙p,𝒌𝒖ap,𝒌(1a1b)𝒙p,𝒌=ca|baa|xp2|ba|a,\left\|\frac{\bm{x}-\bm{u}}{a}-\frac{\bm{x}}{b}\right\|_{p,\bm{k}}=\left\|\frac{-\bm{u}}{a}+\left(\frac{1}{a}-\frac{1}{b}\right)\bm{x}\right\|_{p,\bm{k}}\geq\left\|\frac{-\bm{u}}{a}\right\|_{p,\bm{k}}-\left\|\left(\frac{1}{a}-\frac{1}{b}\right)\bm{x}\right\|_{p,\bm{k}}=\frac{c}{a}-\left|\frac{b-a}{a}\right|\geq\frac{x_{p}}{2}-\frac{|b-a|}{a},

it follows from the uniform convexity of p,𝒌\ell_{p,\bm{k}} that

𝒙𝒖a+𝒙b2p,𝒌1δp(ϵp),\left\|\frac{\frac{\bm{x}-\bm{u}}{a}+\frac{\bm{x}}{b}}{2}\right\|_{p,\bm{k}}\leq 1-\delta_{p}(\epsilon_{p}), (27)

where ϵp:=xp2|ba|axp22xpxp=1+xp22xp\epsilon_{p}:=\frac{x_{p}}{2}-\frac{|b-a|}{a}\geq\frac{x_{p}}{2}-\frac{2-x_{p}}{x_{p}}=1+\frac{x_{p}}{2}-\frac{2}{x_{p}}. So δp(ϵp)δp(1+xp22xp)\delta_{p}(\epsilon_{p})\geq\delta_{p}\left(1+\frac{x_{p}}{2}-\frac{2}{x_{p}}\right). Multiplying by aa in the both sides of the inequality (27) and rearranging it, we have

1+ab2𝒙12𝒖p,𝒌a(1δp(ϵp)).\left\|\frac{1+\frac{a}{b}}{2}\bm{x}-\frac{1}{2}\bm{u}\right\|_{p,\bm{k}}\leq a(1-\delta_{p}(\epsilon_{p})).

Therefore,

𝒙12𝒖p,𝒌=1+ab2𝒙12𝒖+1ab2𝒙p,𝒌1+ab2𝒙12𝒖p,𝒌+1ab2𝒙p,𝒌a(1δp(ϵp))+|ba|2(2(1δp(ϵp))+2xp2)rp,𝒌,n=[2(2δp(ϵp)2xp2)]rp,𝒌,n.\begin{split}\left\|\bm{x}-\frac{1}{2}\bm{u}\right\|_{p,\bm{k}}&=\left\|\frac{1+\frac{a}{b}}{2}\bm{x}-\frac{1}{2}\bm{u}+\frac{1-\frac{a}{b}}{2}\bm{x}\right\|_{p,\bm{k}}\\ &\leq\left\|\frac{1+\frac{a}{b}}{2}\bm{x}-\frac{1}{2}\bm{u}\right\|_{p,\bm{k}}+\left\|\frac{1-\frac{a}{b}}{2}\bm{x}\right\|_{p,\bm{k}}\\ &\leq a(1-\delta_{p}(\epsilon_{p}))+\frac{|b-a|}{2}\\ &\leq\left(2(1-\delta_{p}(\epsilon_{p}))+\frac{2-x_{p}}{2}\right)r_{p,\bm{k},n}\\ &=\left[2-\left(2\delta_{p}(\epsilon_{p})-\frac{2-x_{p}}{2}\right)\right]r_{p,\bm{k},n}.\end{split} (28)

Since

2δp(ϵp)2xp2=2[1(1(ϵp2)q)1/q]1+xp2=12(1(ϵp2)q)1/q+xp212(1(1+xp22xp2)q)1/q+xp2=12(1(12+xp41xp)q)1/q+xp2\begin{split}2\delta_{p}\left(\epsilon_{p}\right)-\frac{2-x_{p}}{2}&=2\left[1-\left(1-\left(\frac{\epsilon_{p}}{2}\right)^{q}\right)^{1/q}\right]-1+\frac{x_{p}}{2}\\ &=1-2\left(1-\left(\frac{\epsilon_{p}}{2}\right)^{q}\right)^{1/q}+\frac{x_{p}}{2}\\ &\geq 1-2\left(1-\left(\frac{1+\frac{x_{p}}{2}-\frac{2}{x_{p}}}{2}\right)^{q}\right)^{1/q}+\frac{x_{p}}{2}\\ &=1-2\left(1-\left(\frac{1}{2}+\frac{x_{p}}{4}-\frac{1}{x_{p}}\right)^{q}\right)^{1/q}+\frac{x_{p}}{2}\end{split}

and by the definition of xpx_{p}

1(12+xp41xp)q(xp+24)q13q,1-\left(\frac{1}{2}+\frac{x_{p}}{4}-\frac{1}{x_{p}}\right)^{q}\leq\left(\frac{x_{p}+2}{4}\right)^{q}-\frac{1}{3^{q}},

it follows that

2δp(ϵp)2xp212(1(12+xp41xp)q)1/q+xp21+xp2((1+xp2)q2q3q)1/q>0.\begin{split}2\delta_{p}\left(\epsilon_{p}\right)-\frac{2-x_{p}}{2}&\geq 1-2\left(1-\left(\frac{1}{2}+\frac{x_{p}}{4}-\frac{1}{x_{p}}\right)^{q}\right)^{1/q}+\frac{x_{p}}{2}\\ &\geq 1+\frac{x_{p}}{2}-\left(\left(1+\frac{x_{p}}{2}\right)^{q}-\frac{2^{q}}{3^{q}}\right)^{1/q}\\ &>0.\end{split} (29)

Let cp=max{xp+22,2(2δp(ϵp)2xp2)}c^{\prime}_{p}=\max\left\{\frac{x_{p}+2}{2},2-\left(2\delta_{p}(\epsilon_{p})-\frac{2-x_{p}}{2}\right)\right\}. By the inequality (29) we see that cp<2c^{\prime}_{p}<2. By inequalities (26) and (28) we see that

Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n)Bp,𝒌n(𝒖/2,cprp,𝒌,n).B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\subseteq B^{n}_{p,\bm{k}}(\bm{u}/2,c^{\prime}_{p}r_{p,\bm{k},n}).

So

vol(Bp,𝒌n(𝒖,2rp,𝒌,n)Bp,𝒌n(𝟎,𝒖p,𝒌,n))vol(Bp,𝒌n(𝒖/2,cprp,𝒌,n))=(cp)n.{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{u},2r_{p,\bm{k},n})\cap B^{n}_{p,\bm{k}}(\bm{0},\|\bm{u}\|_{p,\bm{k},n})\right)\leq{\rm vol}\left(B^{n}_{p,\bm{k}}(\bm{u}/2,c^{\prime}_{p}r_{p,\bm{k},n})\right)=(c^{\prime}_{p})^{n}.

Let cp=max{xp,cp}c_{p}=\max\{x_{p},c^{\prime}_{p}\} and we are done.

The equation (25) follows from inequalities (23) and (24). ∎

Now we are able to prove Theorem 3.2.

Proof of Theorem 3.2.

Let SnS\subseteq\mathbb{R}^{n} be bounded, measurable, and of positive volume. Let cpc_{p} be the constant in Lemma 3.9 and α=αS,p,𝒌(λ)\alpha=\alpha_{S,p,\bm{k}}(\lambda). Then by Jensen’s inequality, we have

α=λ𝔼[1ZT,p,𝒌(λ)]λe𝔼logZT,p,𝒌(λ),\alpha=\lambda\cdot\mathbb{E}\left[\frac{1}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right]\geq\lambda\cdot e^{-\mathbb{E}\log Z_{{\rm T},p,\bm{k}}(\lambda)},

where the above expectation is with respect to the two-part experiment in forming the random set T{\rm T} and the first equality follows from the equation (20).

On the other hand, we have

α2n𝔼[λZT,p,𝒌(λ)ZT,p,𝒌(λ)]=2n𝔼[vol(T)αT,p,𝒌(λ)]2n𝔼[λvol(T)eλ2cpn]2n𝔼[logZT,p,𝒌(λ)eλ2cpn]=2neλ2cpn𝔼[logZT,p,𝒌(λ)].\begin{split}\alpha&\geq 2^{-n}\cdot\mathbb{E}\left[\frac{\lambda\cdot Z^{\prime}_{{\rm T},p,\bm{k}}(\lambda)}{Z_{{\rm T},p,\bm{k}}(\lambda)}\right]\\ &=2^{-n}\cdot\mathbb{E}\left[{\rm vol(T)}\cdot\alpha_{{\rm T},p,\bm{k}}(\lambda)\right]\\ &\geq 2^{-n}\cdot\mathbb{E}\left[\lambda\cdot{\rm vol(T)}\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\right]\\ &\geq 2^{-n}\cdot\mathbb{E}\left[\log Z_{{\rm T},p,\bm{k}}(\lambda)\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\right]\\ &=2^{-n}\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\cdot\mathbb{E}\left[\log Z_{{\rm T},p,\bm{k}}(\lambda)\right].\end{split}

Combining these two lower bounds and letting z=𝔼[logZT,p,𝒌(λ)]z=\mathbb{E}\left[\log Z_{{\rm T},p,\bm{k}}(\lambda)\right], we have

αinfzmax{λez,2neλ2cpnz}.\alpha\geq\inf_{z}\max\left\{\lambda\cdot e^{-z},2^{-n}\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\cdot z\right\}.

Since λez\lambda\cdot e^{-z} is decreasing in zz and 2neλ2cpnz2^{-n}\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\cdot z is increasing in zz, the infimum over zz of the maximum of the two expressions occurs when they are equal, i.e., αλez\alpha\geq\lambda e^{-z^{*}}, where zz^{*} is the solution to

λez=2neλ2cpnz.\lambda\cdot e^{-z}=2^{-n}\cdot e^{-\lambda\cdot 2\cdot c_{p}^{n}}\cdot z.

In other words,

z=W(λ2ne2λcpn),z^{*}=W(\lambda 2^{n}e^{2\lambda c_{p}^{n}}), (30)

where W(x)W(x) is the Lambert-W function. For x>0x>0, w=W(x)w=W(x) is defined to be the unique solution to the equation wew=xwe^{w}=x. So

w=logxlog(logxlogw)=logxloglogxlog(1logwlogx).w=\log x-\log(\log x-\log w)=\log x-\log\log x-\log\left(1-\frac{\log w}{\log x}\right).

As xx\rightarrow\infty,

W(x)=logxloglogx+O(loglogxlogx).W(x)=\log x-\log\log x+O\left(\frac{\log\log x}{\log x}\right).

We take λ=n1cpn\lambda=n^{-1}c_{p}^{-n}. So λ2ne2λcpn=1n(2cp)ne2/n\lambda 2^{n}e^{2\lambda c_{p}^{n}}=\frac{1}{n}\left(\frac{2}{c_{p}}\right)^{n}e^{2/n}\rightarrow\infty as nn\rightarrow\infty. The equation (30) becomes

z=W(λ2ne2/n)=logλ+nlog2lognloglog(2/cp)+O(log(logn/n)).\begin{split}z^{*}&=W(\lambda 2^{n}e^{2/n})\\ &=\log\lambda+n\log 2-\log n-\log\log(2/c_{p})+O(\log(\log n/n)).\end{split} (31)

So

αλez=(1+O(logn/n))log(2/cp)n2n.\alpha\geq\lambda e^{-z^{*}}=\left(1+O(\log n/n)\right)\frac{\log(2/c_{p})\cdot n}{2^{n}}.

Since α\alpha is increasing in λ\lambda, this bound holds for every λn1cpn\lambda\geq n^{-1}c_{p}^{-n}, completing the proof. ∎

In the above proof, if we take λ=n1cn\lambda=n^{-1}c^{-n} for some c[cp,2)c\in[c_{p},2), then we have

αS,p,𝒌(enlogc)=αS,p,𝒌(cn)αS,p,𝒌(n1cn)(1+on(1))(log2logc)n2n,\alpha_{S,p,\bm{k}}(e^{-n\log c})=\alpha_{S,p,\bm{k}}(c^{-n})\geq\alpha_{S,p,\bm{k}}(n^{-1}c^{-n})\geq\left(1+o_{n}(1)\right)\frac{(\log 2-\log c)\cdot n}{2^{n}},

i.e.,

αS,p,𝒌(ent)(1+on(1))(log2t)n2n\alpha_{S,p,\bm{k}}(e^{-nt})\geq\left(1+o_{n}(1)\right)\frac{(\log 2-t)\cdot n}{2^{n}} (32)

for t[logcp,log2)t\in[\log c_{p},\log 2).

4 An alternate proof of Theorem 1.2

In this section, we give an alternate proof of Theorem 1.2 without using the hard superball model. The idea of the proof comes from [13, Section 6.1]. We fix p(1,2]p\in(1,2] and 𝒌\bm{k}, and let nn be a sufficiently large number. For convenience, in this section, we simplify most of our symbols and variables: Bn(R)=Bn(𝟎,R):=Bp,𝒌n(R)=Bp,𝒌n(𝟎,R),rn:=rp,𝒌,n,Bn(𝒙,rn):=Bp,𝒌n(𝒙,rp,𝒌,n),:=p,𝒌,nB^{n}(R)=B^{n}(\bm{0},R):=B^{n}_{p,\bm{k}}(R)=B^{n}_{p,\bm{k}}(\bm{0},R),r_{n}:=r_{p,\bm{k},n},B^{n}(\bm{x},r_{n}):=B^{n}_{p,\bm{k}}(\bm{x},r_{p,\bm{k},n}),\|\cdot\|:=\|\cdot\|_{p,\bm{k},n}, and d(,):=dp,𝒌,n(,)d(\cdot,\cdot):=d_{p,\bm{k},n}(\cdot,\cdot).

Let Bn(R)B^{n}(R) be the superball centered at 𝟎\bm{0} with radius RR. Consider the packings in Bn(R)B^{n}(R) using Bn(𝒙,rn)B^{n}(\bm{x},r_{n}). Let ϵ\epsilon be a small positive real number (we will determine ϵ\epsilon later), Cϵ:=[0,ϵ]n{\rm C}_{\epsilon}:=[0,\epsilon]^{n} be a basic cube and L~ϵ:=(ϵ)n\tilde{L}_{\epsilon}:=(\epsilon\mathbb{Z})^{n} be a lattice. Cϵ+L~ϵ{\rm C}_{\epsilon}+\tilde{L}_{\epsilon} tiles the whole space, and it also gives a partition of Bn(R)B^{n}(R). Let Lϵ={𝒙L~ϵ:Cϵ+𝒙Bn(R)}L_{\epsilon}=\{\bm{x}\in\tilde{L}_{\epsilon}:{\rm C}_{\epsilon}+\bm{x}\subseteq B^{n}(R)\}. We have the following lemma.

Lemma 4.1.

{Cϵ+𝒙:𝒙Lϵ}\{{\rm C}_{\epsilon}+\bm{x}:\bm{x}\in L_{\epsilon}\} covers Bn(R2np+22pϵ)B^{n}(R-2n^{\frac{p+2}{2p}}\epsilon).

Proof.

Suppose on the contrary that there is some 𝒚~=(y1,y2,,yn)Bn(R2np+22pϵ)\tilde{\bm{y}}=(y_{1},y_{2},\ldots,y_{n})\in B^{n}(R-2n^{\frac{p+2}{2p}}\epsilon) such that 𝒚Cϵ+𝒙\bm{y}\notin{\rm C}_{\epsilon}+\bm{x} for every 𝒙Lϵ\bm{x}\in L_{\epsilon}. Let 𝒙~=(x1,x2,,xn)L~ϵ\tilde{\bm{x}}=(x_{1},x_{2},\ldots,x_{n})\in\tilde{L}_{\epsilon} such that 0yixiϵ0\leq y_{i}-x_{i}\leq\epsilon for every 1in1\leq i\leq n, in other words, 𝒚~Cϵ+𝒙~\tilde{\bm{y}}\in{\rm C}_{\epsilon}+\tilde{\bm{x}}. By the choice of 𝒚~\tilde{\bm{y}}, we have that Cϵ+𝒙~Bn(R){\rm C}_{\epsilon}+\tilde{\bm{x}}\nsubseteq B^{n}(R). On the other hand, for every 𝒛Cϵ+𝒙~\bm{z}\in{\rm C}_{\epsilon}+\tilde{\bm{x}}, we have

𝒛𝒛𝒚+𝒚(j=1m(ϵ,ϵ,,ϵ)2p)1/p+R2np+22pϵ.\|\bm{z}\|\leq\|\bm{z}-\bm{y}\|+\|\bm{y}\|\leq\left(\sum_{j=1}^{m}\left\|\left(\epsilon,\epsilon,\ldots,\epsilon\right)\right\|_{2}^{p}\right)^{1/p}+R-2n^{\frac{p+2}{2p}}\epsilon.

Note that (ϵ,ϵ,,ϵ)2p=(ϵ2+ϵ2++ϵ2)p/2np/2ϵp\left\|\left(\epsilon,\epsilon,\ldots,\epsilon\right)\right\|_{2}^{p}=(\epsilon^{2}+\epsilon^{2}+\cdots+\epsilon^{2})^{p/2}\leq n^{p/2}\epsilon^{p} since the number of ϵ\epsilon is at most nn, and j=1m(ϵ,ϵ,,ϵ)2pn(p+2)/2ϵp\sum_{j=1}^{m}\left\|\left(\epsilon,\epsilon,\ldots,\epsilon\right)\right\|_{2}^{p}\leq n^{(p+2)/2}\epsilon^{p} since mnm\leq n. So

𝒛(j=1m(ϵ,ϵ,,ϵ)2p)1/p+R2np+22pϵR2np+22pϵ+np+22pϵ<R.\|\bm{z}\|\leq\left(\sum_{j=1}^{m}\left\|\left(\epsilon,\epsilon,\ldots,\epsilon\right)\right\|_{2}^{p}\right)^{1/p}+R-2n^{\frac{p+2}{2p}}\epsilon\leq R-2n^{\frac{p+2}{2p}}\epsilon+n^{\frac{p+2}{2p}}\epsilon<R.

This holds for every 𝒛Cϵ+𝒙~\bm{z}\in{\rm C}_{\epsilon}+\tilde{\bm{x}}. Thus Cϵ+𝒙~Bn(R){\rm C}_{\epsilon}+\tilde{\bm{x}}\subseteq B^{n}(R), a contradiction. ∎

If we write N:=|Lϵ|N:=\left|L_{\epsilon}\right|, we can estimate NN by Lemma 4.1; that is,

vol(Bn(R2np+22pϵ))Nvol(Cϵ)vol(Bn(R)),{\rm vol}\left(B^{n}(R-2n^{\frac{p+2}{2p}}\epsilon)\right)\leq N\cdot{\rm vol}({\rm C}_{\epsilon})\leq{\rm vol}\left(B^{n}(R)\right),

i.e.,

(R2np+22pϵϵrn)nN(Rϵrn)n.\left(\frac{R-2n^{\frac{p+2}{2p}}\epsilon}{\epsilon r_{n}}\right)^{n}\leq N\leq\left(\frac{R}{\epsilon r_{n}}\right)^{n}.

So N=(1oR(1))(Rϵrn)nN=(1-o_{R}(1))\left(\frac{R}{\epsilon r_{n}}\right)^{n}.

Let {Cϵ+𝒙:𝒙Lϵ}={C1,C2,,CN}\{{\rm C}_{\epsilon}+\bm{x}:\bm{x}\in L_{\epsilon}\}=\{C_{1},C_{2},\ldots,C_{N}\}. We arbitrarily choose 𝒗i\bm{v}_{i} from each CiC_{i}, and construct an auxiliary graph GG with vertex set V(G)={𝒗1,𝒗2,,𝒗N}V(G)=\{\bm{v}_{1},\bm{v}_{2},\ldots,\bm{v}_{N}\}, where 𝒗i\bm{v}_{i} and 𝒗j\bm{v}_{j} are adjacent if and only if d(𝒗i,𝒗j)<2rnd(\bm{v}_{i},\bm{v}_{j})<2r_{n}. So the maximum number of superballs with radius rnr_{n} that can be packed in Bn(R)B^{n}(R) is larger than or equal to α(G)\alpha(G), where α(G)\alpha(G) is the independence number of GG. In other words,

Δp,𝒌(n)limRα(G)(R/rn)n.\Delta_{p,\bm{k}}(n)\geq\lim_{R\rightarrow\infty}\frac{\alpha(G)}{\left(R/r_{n}\right)^{n}}. (33)

Here, using the same trick in the proof of Lemma 3.1, the points near the boundary of Bn(R)B^{n}(R) do not affect the result.

Lemma 4.2.

Every vertex in GG has degree at most 1+on(1)1oR(1)(2rnR)nN\frac{1+o_{n}(1)}{1-o_{R}(1)}\left(\frac{2r_{n}}{R}\right)^{n}N.

Proof.

We write N[𝒙]=N(𝒙){𝒙}N[\bm{x}]=N(\bm{x})\cup\{\bm{x}\} for the closed neighborhood of 𝒙\bm{x}. We claim that for every 𝒙V(G)\bm{x}\in V(G),

Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ)𝒗iN[𝒙]CiBn(𝒙,2rn+2np+22pϵ).B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\cap B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon)\subseteq\bigcup_{\bm{v}_{i}\in N[\bm{x}]}C_{i}\subseteq B^{n}(\bm{x},2r_{n}+2n^{\frac{p+2}{2p}}\epsilon). (34)

For the first inclusion, let 𝒚Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ)\bm{y}\in B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\cap B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon). By Lemma 4.1, there exists an index ii such that 𝒚Ci\bm{y}\in C_{i}. So d(𝒚,𝒗i)np+22pϵd(\bm{y},\bm{v}_{i})\leq n^{\frac{p+2}{2p}}\epsilon. As d(𝒙,𝒚)2rn2np+22pϵd(\bm{x},\bm{y})\leq 2r_{n}-2n^{\frac{p+2}{2p}}\epsilon, it follows that d(𝒙,𝒗i)2rn2np+22pϵ+np+22pϵ<2rnd(\bm{x},\bm{v}_{i})\leq 2r_{n}-2n^{\frac{p+2}{2p}}\epsilon+n^{\frac{p+2}{2p}}\epsilon<2r_{n}, i.e. 𝒗iN[𝒙]\bm{v}_{i}\in N[\bm{x}]. For the second inclusion, let 𝒚Ci\bm{y}\in C_{i} for some index ii with 𝒗iN[𝒙]\bm{v}_{i}\in N[\bm{x}]. We have d(𝒚,𝒗i)np+22pϵd(\bm{y},\bm{v}_{i})\leq n^{\frac{p+2}{2p}}\epsilon and d(𝒙,𝒗i)<2rp,𝒌,nd(\bm{x},\bm{v}_{i})<2r_{p,\bm{k},n}. So d(𝒚,𝒙)2rp,𝒌,n+2np+22pϵd(\bm{y},\bm{x})\leq 2r_{p,\bm{k},n}+2n^{\frac{p+2}{2p}}\epsilon, as desired.

Since CiC_{i}’s are nonoverlapping, it follows from the claim that

vol(𝒗iN[𝒙]Ci)=|N[𝒙]|ϵnvol(Bn(𝒙,2rn+2np+22pϵ))=(2rn+2np+22pϵrn)n,{\rm vol}\left(\bigcup_{\bm{v}_{i}\in N[\bm{x}]}C_{i}\right)=|N[\bm{x}]|\epsilon^{n}\leq{\rm vol}(B^{n}(\bm{x},2r_{n}+2n^{\frac{p+2}{2p}}\epsilon))=\left(\frac{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}{r_{n}}\right)^{n},

i.e.,

|N[𝒙]|(1ϵrn)n(2rn+2np+22pϵ)n=1+on(1)1oR(1)(2rnR)nN,|N[\bm{x}]|\leq\left(\frac{1}{\epsilon r_{n}}\right)^{n}(2r_{n}+2n^{\frac{p+2}{2p}}\epsilon)^{n}=\frac{1+o_{n}(1)}{1-o_{R}(1)}\left(\frac{2r_{n}}{R}\right)^{n}N,

where we choose ϵ\epsilon satisfying np+22pϵ/rn<n2n^{\frac{p+2}{2p}}\epsilon/r_{n}<n^{-2}. ∎

Let D:=1+on(1)1oR(1)(2rnR)nND:=\frac{1+o_{n}(1)}{1-o_{R}(1)}\left(\frac{2r_{n}}{R}\right)^{n}N and K:=110(2cp)nK:=\frac{1}{10}\left(\frac{2}{c_{p}}\right)^{n}, where cpc_{p} is the constant in Lemma 3.9. Let G[N(𝒙)]G[N(\bm{x})] denote the induced subgraph of GG whose vertex set is N(𝒙)N(\bm{x}). We have the following lemma.

Lemma 4.3.

Suppose RR and nn are sufficiently large. For every 𝐱V(G)\bm{x}\in V(G), the average degree of G[N(𝐱)]G[N(\bm{x})] is at most DK\frac{D}{K}.

Proof.

We calculate DK=10(1+on(1))(cpϵ)n\frac{D}{K}=10(1+o_{n}(1))\left(\frac{c_{p}}{\epsilon}\right)^{n}. Let S=𝒗iN(𝒙)CiS^{\prime}=\bigcup_{\bm{v}_{i}\in N(\bm{x})}C_{i} and N(𝒙)={𝒙1,𝒙2,,𝒙t}N(\bm{x})=\{\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t}\} where t=|N(𝒙)|t=|N(\bm{x})|. So vol(S)=tϵn{\rm vol}(S^{\prime})=t\epsilon^{n}. By the definition, the average degree of G[N(𝒙)]G[N(\bm{x})] is at most

1ti,j=1t𝟏d(𝒙i,𝒙j)2rn=ϵnvol(S)i,j=1t𝟏d(𝒙i,𝒙j)2rn.\frac{1}{t}\sum_{i,j=1}^{t}\mathbf{1}_{d(\bm{x}_{i},\bm{x}_{j})\leq 2r_{n}}=\frac{\epsilon^{n}}{{\rm vol}(S^{\prime})}\sum_{i,j=1}^{t}\mathbf{1}_{d(\bm{x}_{i},\bm{x}_{j})\leq 2r_{n}}. (35)

By the definition of integration,

limϵ0ϵ2ni,j=1t𝟏d(𝒙i,𝒙j)2rn=SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗,\lim_{\epsilon\rightarrow 0}\epsilon^{2n}\sum_{i,j=1}^{t}\mathbf{1}_{d(\bm{x}_{i},\bm{x}_{j})\leq 2r_{n}}=\int_{S^{\prime}}\int_{S^{\prime}}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v},

in other words, for every δ>0\delta>0, there exists ϵ0(δ)\epsilon_{0}(\delta) such that

ϵ2ni,j=1t𝟏d(𝒙i,𝒙j)2rnSS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗+δ.\epsilon^{2n}\sum_{i,j=1}^{t}\mathbf{1}_{d(\bm{x}_{i},\bm{x}_{j})\leq 2r_{n}}\leq\int_{S^{\prime}}\int_{S^{\prime}}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}+\delta. (36)

whenever ϵ<ϵ0(δ)\epsilon<\epsilon_{0}(\delta).

Recall that in the proof of Lemma 4.2 we have SBn(𝒙,2rn+2np+22pϵ)S^{\prime}\subseteq B^{n}(\bm{x},2r_{n}+2n^{\frac{p+2}{2p}}\epsilon). Let S=2rn2rn+2np+22pϵSS=\frac{2r_{n}}{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}S^{\prime}. So SBn(𝒙,2rn)S\subseteq B^{n}(\bm{x},2r_{n}) and

SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗=(2rn+2np+22pϵ2rn)2nSS𝟏d(𝒖,𝒗)2rn2rn/(2rn+2np+22pϵ)𝑑𝒖𝑑𝒗(2rn+2np+22pϵ2rn)nvol(S)vol(S)SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗=(1+on(1))vol(S)vol(S)SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗,\begin{split}\int_{S^{\prime}}\int_{S^{\prime}}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}=&\left(\frac{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}{2r_{n}}\right)^{2n}\int_{S}\int_{S}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}\cdot 2r_{n}/(2r_{n}+2n^{\frac{p+2}{2p}}\epsilon)}d\bm{u}d\bm{v}\\ \leq&\left(\frac{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}{2r_{n}}\right)^{n}\frac{{\rm vol}(S^{\prime})}{{\rm vol}(S)}\int_{S}\int_{S}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}\\ =&(1+o_{n}(1))\frac{{\rm vol}(S^{\prime})}{{\rm vol}(S)}\int_{S}\int_{S}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v},\end{split} (37)

where we choose ϵ\epsilon small enough so that (2rn+2np+22pϵ2rn)n=1+on(1)\left(\frac{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}{2r_{n}}\right)^{n}=1+o_{n}(1). For instance, if we choose ϵ\epsilon satisfying np+22pϵ/rn<n2n^{\frac{p+2}{2p}}\epsilon/r_{n}<n^{-2}, then (2rn+2np+22pϵ2rn)n<(1+1n2)n<e1/n=1+on(1)\left(\frac{2r_{n}+2n^{\frac{p+2}{2p}}\epsilon}{2r_{n}}\right)^{n}<\left(1+\frac{1}{n^{2}}\right)^{n}<e^{1/n}=1+o_{n}(1).

Recall that in the proof of Lemma 3.9, we have

SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗2vol(S)cpn.\int_{S}\int_{S}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}\leq 2{\rm vol}(S)c_{p}^{n}. (38)

Combining equation (35) and inequalities (36)-(38), the average degree of G[N(𝒙)]G[N(\bm{x})] is at most

ϵnvol(S)i,j=1t𝟏d(𝒙i,𝒙j)2rn1vol(S)ϵn(SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗+δ)1vol(S)ϵn((1+on(1))vol(S)vol(S)SS𝟏d(𝒖,𝒗)2rn𝑑𝒖𝑑𝒗+δ)1vol(S)ϵn((1+on(1))2cpnvol(S)+δ)=(1+on(1))2(cpϵ)n+δvol(S)ϵn.\begin{split}\frac{\epsilon^{n}}{{\rm vol}(S^{\prime})}\sum_{i,j=1}^{t}\mathbf{1}_{d(\bm{x}_{i},\bm{x}_{j})\leq 2r_{n}}\leq&\frac{1}{{\rm vol}(S^{\prime})\epsilon^{n}}\left(\int_{S^{\prime}}\int_{S^{\prime}}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}+\delta\right)\\ \leq&\frac{1}{{\rm vol}(S^{\prime})\epsilon^{n}}\left((1+o_{n}(1))\frac{{\rm vol}(S^{\prime})}{{\rm vol}(S)}\int_{S}\int_{S}\mathbf{1}_{d(\bm{u},\bm{v})\leq 2r_{n}}d\bm{u}d\bm{v}+\delta\right)\\ \leq&\frac{1}{{\rm vol}(S^{\prime})\epsilon^{n}}\left((1+o_{n}(1))\cdot 2c_{p}^{n}{\rm vol}(S^{\prime})+\delta\right)\\ =&(1+o_{n}(1))\cdot 2\left(\frac{c_{p}}{\epsilon}\right)^{n}+\frac{\delta}{{\rm vol}(S^{\prime})\epsilon^{n}}.\end{split} (39)

Next we give a lower bound of vol(S){\rm vol}(S^{\prime}). Recalling the inclusion relation (34), we have

𝒗iN[𝒙]CiBn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ).\bigcup_{\bm{v}_{i}\in N[\bm{x}]}C_{i}\supseteq B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\cap B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon).

Denote Vlower:=vol(Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ))V_{{\rm lower}}:={\rm vol}\left(B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\cap B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon)\right). If Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ)B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\subseteq B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon), then

Vlowervol(Bn(𝒙,2rn2np+22pϵ))=(2rn2np+22pϵrn)n=(1on(1))2n.V_{{\rm lower}}\geq{\rm vol}\left(B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\right)=\left(\frac{2r_{n}-2n^{\frac{p+2}{2p}}\epsilon}{r_{n}}\right)^{n}=(1-o_{n}(1))2^{n}.

If Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ)B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\nsubseteq B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon), then we have

Bn(𝒙,2rn2np+22pϵ)Bn(𝟎,R2np+22pϵ)Bn(𝒚,rn2np+22pϵ),B^{n}(\bm{x},2r_{n}-2n^{\frac{p+2}{2p}}\epsilon)\cap B^{n}(\bm{0},R-2n^{\frac{p+2}{2p}}\epsilon)\supseteq B^{n}\left(\bm{y},r_{n}-2n^{\frac{p+2}{2p}}\epsilon\right),

where 𝒚:=(1rn𝒙)𝒙\bm{y}:=\left(1-\frac{r_{n}}{\|\bm{x}\|}\right)\bm{x}. To see this, letting 𝒛Bn(𝒚,rn2np+22pϵ)\bm{z}\in B^{n}\left(\bm{y},r_{n}-2n^{\frac{p+2}{2p}}\epsilon\right), we have d(𝒛,𝒚)rn2np+22pϵd(\bm{z},\bm{y})\leq r_{n}-2n^{\frac{p+2}{2p}}\epsilon. So

d(𝒙,𝒛)d(𝒙,𝒚)+d(𝒚,𝒛)rn+rn2np+22pϵ=2rn2np+22pϵ.d(\bm{x},\bm{z})\leq d(\bm{x},\bm{y})+d(\bm{y},\bm{z})\leq r_{n}+r_{n}-2n^{\frac{p+2}{2p}}\epsilon=2r_{n}-2n^{\frac{p+2}{2p}}\epsilon.

And

d(𝟎,𝒛)d(𝟎,𝒚)+d(𝒚,𝒛)𝒙rn+rn2np+22pϵR2np+22pϵ,d(\bm{0},\bm{z})\leq d(\bm{0},\bm{y})+d(\bm{y},\bm{z})\leq\|\bm{x}\|-r_{n}+r_{n}-2n^{\frac{p+2}{2p}}\epsilon\leq R-2n^{\frac{p+2}{2p}}\epsilon,

Since 𝒙Bn(𝟎,R)\bm{x}\in B^{n}(\bm{0},R). Thus

Vlowervol(Bn(𝒚,rn2np+22pϵ))=(rn2np+22pϵrn)n=1on(1).V_{{\rm lower}}\geq{\rm vol}\left(B^{n}\left(\bm{y},r_{n}-2n^{\frac{p+2}{2p}}\epsilon\right)\right)=\left(\frac{r_{n}-2n^{\frac{p+2}{2p}}\epsilon}{r_{n}}\right)^{n}=1-o_{n}(1).

In conclusion, we have

vol(S)=vol(𝒗iN[𝒙]Ci)ϵnVlowerϵn1on(1).{\rm vol}(S^{\prime})={\rm vol}\left(\bigcup_{\bm{v}_{i}\in N[\bm{x}]}C_{i}\right)-\epsilon^{n}\geq V_{{\rm lower}}-\epsilon^{n}\geq 1-o_{n}(1).

Therefore, if we choose δ=cpn\delta=c_{p}^{n} and ϵmin{ϵ0(cpn),rn/n2+p+22p}\epsilon\leq\min\{\epsilon_{0}(c_{p}^{n}),r_{n}/n^{2+\frac{p+2}{2p}}\}, then by inequality (39), we know that the average degree of G[N(𝒙)]G[N(\bm{x})] is at most

(1+on(1))2(cpϵ)n+δvol(S)ϵn(1+on(1))2(cpϵ)n+cpn(1on(1))ϵn<5(cpϵ)n<DK,(1+o_{n}(1))\cdot 2\left(\frac{c_{p}}{\epsilon}\right)^{n}+\frac{\delta}{{\rm vol}(S^{\prime})\epsilon^{n}}\leq(1+o_{n}(1))\cdot 2\left(\frac{c_{p}}{\epsilon}\right)^{n}+\frac{c_{p}^{n}}{(1-o_{n}(1))\epsilon^{n}}<5\left(\frac{c_{p}}{\epsilon}\right)^{n}<\frac{D}{K},

for large nn. Note that this is independent of the choice of 𝒙\bm{x}, so we are done. ∎

By Lemmas 4.2 and 4.3, we know that the NN-vertex graph GG has maximum degree DD and every subgraph induced by a neighborhood in GG has average degree at most D/KD/K. We say such a graph is locally sparse. By a result of Hurley and Pirot [15, Corollary 1], the chromatic number of GG is at most (1+oK(1))DlogK(1+o_{K}(1))\frac{D}{\log K}. So the independence number α(G)\alpha(G) of GG is at least

(1oK(1))NDlogK=(1on(1))1oR(1)1+on(1)(R2rn)nlog(110(2cp)n)=(1on(1))(1oR(1))(Rrn)nlog(2/cp)n2n,\begin{split}(1-o_{K}(1))\frac{N}{D}\log K=&(1-o_{n}(1))\frac{1-o_{R}(1)}{1+o_{n}(1)}\left(\frac{R}{2r_{n}}\right)^{n}\log\left(\frac{1}{10}\left(\frac{2}{c_{p}}\right)^{n}\right)\\ =&(1-o_{n}(1))(1-o_{R}(1))\left(\frac{R}{r_{n}}\right)^{n}\frac{\log(2/c_{p})\cdot n}{2^{n}},\end{split} (40)

and

Δp,𝒌(n)limRα(G)(R/rn)n=(1on(1))log(2/cp)n2n.\Delta_{p,\bm{k}}(n)\geq\lim_{R\rightarrow\infty}\frac{\alpha(G)}{\left(R/r_{n}\right)^{n}}=(1-o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}.

This completes the proof of Theorem 1.2.

5 A lower bound on entropy density and pressure

In this section, we investigate the entropy density and pressure of packings. Let Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R) be the region of packings and V=vol(Bp,𝒌n(R))=(R/rp,𝒌,n)nV={\rm vol}(B^{n}_{p,\bm{k}}(R))=(R/r_{p,\bm{k},n})^{n}. Define

fp,𝒌,n(α)=limV1αVlogZ^Bp,𝒌n(R),p,𝒌(αV)VαV/αV!f_{p,\bm{k},n}(\alpha)=\lim_{V\rightarrow\infty}\frac{1}{\alpha V}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lfloor\alpha V\rfloor)}{V^{\lfloor\alpha V\rfloor}/\lfloor\alpha V\rfloor!}

to be the entropy density of the packings. Consider a random packing in Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R) with density α\alpha. The superballs we use in the packings have volume 11, so there are αV\lfloor\alpha V\rfloor centers. VαV/αV!V^{\lfloor\alpha V\rfloor}/\lfloor\alpha V\rfloor! measures all the choices of αV\lfloor\alpha V\rfloor unordered points chosen in Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R), and Z^Bp,𝒌n(R),p,𝒌(αV)\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lfloor\alpha V\rfloor) measures the choices of αV\lfloor\alpha V\rfloor unordered points that can form a packing. Dividing by αV\alpha V ensures that fp,𝒌,n(α)f_{p,\bm{k},n}(\alpha) is independent of the choice of the size of superballs in our packings. So fp,𝒌,n(α)f_{p,\bm{k},n}(\alpha) measures how plentiful the packings of density α\alpha are. We also define

gp,𝒌,n(λ)=limV1VlogZBp,𝒌n(R),p,𝒌(λ)g_{p,\bm{k},n}(\lambda)=\lim_{V\rightarrow\infty}\frac{1}{V}\log Z_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda)

to be the pressure of the packings. gp,𝒌,n(λ)g_{p,\bm{k},n}(\lambda) measures how plentiful the packings with fugacity λ\lambda are. We have the following theorems.

Theorem 5.1.

Let cpc_{p} be the constant in Theorem 3.2. For λ(2n,cpn]\lambda\in(2^{-n},c_{p}^{-n}], we have

gp,𝒌,n(λ)((log2+1nlogλ)22+on(1))n22n.g_{p,\bm{k},n}(\lambda)\geq\left(\frac{\left(\log 2+\frac{1}{n}\log\lambda\right)^{2}}{2}+o_{n}(1)\right)\frac{n^{2}}{2^{n}}. (41)

Note that in Theorem 5.1, log2+1nlogλ(0,log2logcp]\log 2+\frac{1}{n}\log\lambda\in(0,\log 2-\log c_{p}].

Theorem 5.2.

Let cpc_{p} be the constant in Theorem 3.2. There exists α=(1+on(1))log(2/cp)n2n\alpha=(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}} such that

fp,𝒌,n(α)(1+on(1))log(2/cp)nf_{p,\bm{k},n}(\alpha)\geq-(1+o_{n}(1))\log(2/c_{p})\cdot n

The proofs of Theorems 5.1 and 5.2 are similar to those of [11, Theorem 4] and [11, Theorem 5], respectively. For the sake of completeness, we include them here.

Proof of Theorem 5.1.

We calculate

1VlogZBp,𝒌n(R),p,𝒌(λ)=limϵ0ϵλ1V(logZBp,𝒌n(R),p,𝒌(x))𝑑x=limϵ0ϵλαBp,𝒌n(R),p,𝒌(x)x𝑑x2nλαBp,𝒌n(R),p,𝒌(x)x𝑑x,\begin{split}\frac{1}{V}\log Z_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda)=&\lim_{\epsilon\rightarrow 0}\int_{\epsilon}^{\lambda}\frac{1}{V}\left(\log Z_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(x)\right)^{\prime}dx\\ =&\lim_{\epsilon\rightarrow 0}\int_{\epsilon}^{\lambda}\frac{\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(x)}{x}dx\\ \geq&\int_{2^{-n}}^{\lambda}\frac{\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(x)}{x}dx,\end{split}

where the second equation uses equation (18) and the last inequality follows from the positivity of the integrand. Put x=entx=e^{-nt}. When x=λx=\lambda, t=1nlogλt=-\frac{1}{n}\log\lambda; when x=2nx=2^{-n}, t=log2t=\log 2. Since λ(2n,cpn]\lambda\in(2^{-n},c_{p}^{-n}], it follows that [1nlogλ,log2][logcp,log2][-\frac{1}{n}\log\lambda,\log 2]\subseteq[\log c_{p},\log 2] and we can use inequality (32) to estimate the lower bound of αBp,𝒌n(R),p,𝒌(ent)\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(e^{-nt}). We have

2nλαBp,𝒌n(R),p,𝒌(x)x𝑑x=nlog21nlogλαBp,𝒌n(𝟎,R),p,𝒌(ent)𝑑t=n1nlogλlog2αBp,𝒌n(R),p,𝒌(ent)𝑑tn(1+o(1))1nlogλlog2(log2t)n2n𝑑t=((log2+1nlogλ)22+on(1))n22n.\begin{split}\int_{2^{-n}}^{\lambda}\frac{\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(x)}{x}dx=&-n\int_{\log 2}^{-\frac{1}{n}\log\lambda}\alpha_{B^{n}_{p,\bm{k}}(\bm{0},R),p,\bm{k}}(e^{-nt})dt\\ =&n\int_{-\frac{1}{n}\log\lambda}^{\log 2}\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(e^{-nt})dt\\ \geq&n\left(1+o(1)\right)\int_{-\frac{1}{n}\log\lambda}^{\log 2}\frac{(\log 2-t)\cdot n}{2^{n}}dt\\ =&\left(\frac{\left(\log 2+\frac{1}{n}\log\lambda\right)^{2}}{2}+o_{n}(1)\right)\frac{n^{2}}{2^{n}}.\end{split}

Let VV go to infinity and we obtain equation (41). ∎

Before proving Theorem 5.2, we give a simple fact of fp,𝒌,n(α)f_{p,\bm{k},n}(\alpha): the larger the density is, the less such packings exist.

Lemma 5.3.

fp,𝒌,n(α)f_{p,\bm{k},n}(\alpha) is decreasing in α\alpha.

Proof.

If α>Δp,𝒌(n)\alpha>\Delta_{p,\bm{k}}(n), then Z^Bp,𝒌n(R),p,𝒌(αV)=0\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lfloor\alpha V\rfloor)=0 for large VV. So fp,𝒌,n(α)=0f_{p,\bm{k},n}(\alpha)=0 when α>Δp,𝒌(n)\alpha>\Delta_{p,\bm{k}}(n).

Suppose 0<α1<α2Δp,𝒌(n)0<\alpha_{1}<\alpha_{2}\leq\Delta_{p,\bm{k}}(n), and let V1V_{1} and V2V_{2} satisfy α1V1=α2V2\alpha_{1}V_{1}=\alpha_{2}V_{2} (so V1>V2V_{1}>V_{2}). We claim that

1α1V1logZ^Bp,𝒌n(R1),p,𝒌(α1V1)V1α1V1/α1V1!>1α2V2logZ^Bp,𝒌n(R2),p,𝒌(α2V2)V2α2V2/α2V2!,\frac{1}{\alpha_{1}V_{1}}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R_{1}),p,\bm{k}}(\lfloor\alpha_{1}V_{1}\rfloor)}{V_{1}^{\lfloor\alpha_{1}V_{1}\rfloor}/\lfloor\alpha_{1}V_{1}\rfloor!}>\frac{1}{\alpha_{2}V_{2}}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R_{2}),p,\bm{k}}(\lfloor\alpha_{2}V_{2}\rfloor)}{V_{2}^{\lfloor\alpha_{2}V_{2}\rfloor}/\lfloor\alpha_{2}V_{2}\rfloor!}, (42)

where V1=(R1/rp,𝒌,n)nV_{1}=(R_{1}/r_{p,\bm{k},n})^{n} and V2=(R2/rp,𝒌,n)nV_{2}=(R_{2}/r_{p,\bm{k},n})^{n} (so R1>R2R_{1}>R_{2}). In order to prove inequality (42), it suffices to prove

Z^Bp,𝒌n(R1),p,𝒌(α1V1)V1α1V1>Z^Bp,𝒌n(R2),p,𝒌(α2V2)V2α2V2.\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R_{1}),p,\bm{k}}(\lfloor\alpha_{1}V_{1}\rfloor)}{V_{1}^{\lfloor\alpha_{1}V_{1}\rfloor}}>\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R_{2}),p,\bm{k}}(\lfloor\alpha_{2}V_{2}\rfloor)}{V_{2}^{\lfloor\alpha_{2}V_{2}\rfloor}}.

Using equation (16), the definition of Z^S,p,𝒌(t)\hat{Z}_{S,p,\bm{k}}(t), it reduces to

Bp,𝒌n(R1)t1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙tV1t>Bp,𝒌n(R2)t1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙tV2t,\frac{\int_{B^{n}_{p,\bm{k}}(R_{1})^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}}{V_{1}^{t}}>\frac{\int_{B^{n}_{p,\bm{k}}(R_{2})^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}}{V_{2}^{t}}, (43)

where t=α1V1=α2V2t=\lfloor\alpha_{1}V_{1}\rfloor=\lfloor\alpha_{2}V_{2}\rfloor. Consider

Bp,𝒌n(R2)t1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙tV2t,\frac{\int_{B^{n}_{p,\bm{k}}(R_{2})^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}}{V_{2}^{t}},

the right hand side of inequality (43). Let 𝒚i=(R1/R2)𝒙i,i=1,,t\bm{y}_{i}=(R_{1}/R_{2})\bm{x}_{i},i=1,\ldots,t. Then

Bp,𝒌n(R2)t1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)𝑑𝒙1𝑑𝒙2𝑑𝒙tV2t=Bp,𝒌n(R1)t1𝒟p,𝒌(𝒚1,𝒚2,,𝒚t)𝑑𝒚1𝑑𝒚2𝑑𝒚tV1t,\frac{\int_{B^{n}_{p,\bm{k}}(R_{2})^{t}}\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}d\bm{x}_{1}d\bm{x}_{2}\cdots d\bm{x}_{t}}{V_{2}^{t}}=\frac{\int_{B^{n}_{p,\bm{k}}(R_{1})^{t}}\textbf{1}_{\mathcal{D}^{\prime}_{p,\bm{k}}(\bm{y}_{1},\bm{y}_{2},\ldots,\bm{y}_{t})}d\bm{y}_{1}d\bm{y}_{2}\cdots d\bm{y}_{t}}{V_{1}^{t}},

where 𝒟p,𝒌(𝒚1,𝒚2,,𝒚t)\mathcal{D}^{\prime}_{p,\bm{k}}(\bm{y}_{1},\bm{y}_{2},\ldots,\bm{y}_{t}) means the event that dp,𝒌,n(𝒚i,𝒚j)R1R22rp,𝒌,nd_{p,\bm{k},n}(\bm{y}_{i},\bm{y}_{j})\geq\frac{R_{1}}{R_{2}}\cdot 2r_{p,\bm{k},n} for every iji\neq j.

The left hand side of inequality (43) is the probability of the event that tt uniformly chosen points in Bp,𝒌n(R1)B^{n}_{p,\bm{k}}(R_{1}) with pairwise distance larger than or equal to 2rp,𝒌,n2r_{p,\bm{k},n}, while the right hand side of inequality (43) is the probability of the event that tt uniformly chosen points in Bp,𝒌n(R1)B^{n}_{p,\bm{k}}(R_{1}) with pairwise distance larger than or equal to R1R22rp,𝒌,n\frac{R_{1}}{R_{2}}\cdot 2r_{p,\bm{k},n}. The latter event is contained in the former event, i.e.

1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t)1𝒟p,𝒌(𝒙1,𝒙2,,𝒙t).\textbf{1}_{\mathcal{D}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}\geq\textbf{1}_{\mathcal{D}^{\prime}_{p,\bm{k}}(\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{t})}.

Thus inequality (43) holds and the claim is proved.

Letting V1V_{1} go to infinity in both sides of inequality (42), we obtain that

fp,𝒌,n(α1)>fp,𝒌,n(α2).f_{p,\bm{k},n}(\alpha_{1})>f_{p,\bm{k},n}(\alpha_{2}).

This completes the proof. ∎

Finally, we give a proof of Theorem 5.2.

Proof of Theorem 5.2.

Consider packings in Bp,𝒌n(R)B^{n}_{p,\bm{k}}(R) and V=vol(Bp,𝒌n(R))=(R/rp,𝒌,n)nV={\rm vol}(B^{n}_{p,\bm{k}}(R))=(R/r_{p,\bm{k},n})^{n}. Assume that Var(|X|)V3/2{\rm Var}(|X|)\geq V^{3/2} for every λ[cpn,2cpn]\lambda\in[c_{p}^{-n},2c_{p}^{-n}]. Using equation (19) in the proof of Lemma 3.4, we have

αBp,𝒌n(R),p,𝒌(2cpn)=02cpnαBp,𝒌n(R),p,𝒌(x)𝑑x=02cpnVar(|X|)xV𝑑xcpn2cpnVar(|X|)xV𝑑xV1/2(log21)>1,\begin{split}\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(2c_{p}^{-n})=&\int_{0}^{2c_{p}^{-n}}\alpha^{\prime}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(x)dx\\ =&\int_{0}^{2c_{p}^{-n}}\frac{{\rm Var}(|X|)}{xV}dx\\ \geq&\int_{c_{p}^{-n}}^{2c_{p}^{-n}}\frac{{\rm Var}(|X|)}{xV}dx\\ \geq&V^{1/2}(\log 2-1)>1,\end{split}

a contradiction. So there exists λ[cpn,2cpn]\lambda\in[c_{p}^{-n},2c_{p}^{-n}] such that Var(|X|)<V3/2{\rm Var}(|X|)<V^{3/2}.

Choose λ[cpn,2cpn]\lambda\in[c_{p}^{-n},2c_{p}^{-n}] such that Var(|X|)<V3/2{\rm Var}(|X|)<V^{3/2}. Using Chebyshev’s inequality, we have

Bp,𝒌n(R),p,𝒌,λ(||X|𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)|V4/5)Var(|X|)V8/5<V1/10.\mathbb{P}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}\left(\left||X|-\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)\right|\geq V^{4/5}\right)\leq\frac{{\rm Var}(|X|)}{V^{8/5}}<V^{-1/10}.

So

Bp,𝒌n(R),p,𝒌,λ(||X|𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)|V4/5)1V1/10.\mathbb{P}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}\left(\left||X|-\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)\right|\leq V^{4/5}\right)\geq 1-V^{-1/10}.

Note that |X||X| is an integer. By averaging, there exists an integer

t[𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)V4/5,𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)+V4/5]t\in\left[\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)-V^{4/5},\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)+V^{4/5}\right]

such that

Bp,𝒌n(R),p,𝒌,λ(|X|=t)1V1/102V4/51V,\mathbb{P}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}\left(|X|=t\right)\geq\frac{1-V^{-1/10}}{\lfloor 2V^{4/5}\rfloor}\geq\frac{1}{V},

for large VV. Recall that

Bp,𝒌n(R),p,𝒌,λ(|X|=t)=λtZ^Bp,𝒌n(R),p,𝒌(t)ZBp,𝒌n(R),p,𝒌(λ).\mathbb{P}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}\left(|X|=t\right)=\frac{\lambda^{t}\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(t)}{Z_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda)}.

So

Z^Bp,𝒌n(R),p,𝒌(t)1VλtZBp,𝒌n(R),p,𝒌(λ)1Vλt.\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(t)\geq\frac{1}{V\lambda^{t}}Z_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda)\geq\frac{1}{V\lambda^{t}}.

On the other hand, by Theorem 3.2 and the definition of αBp,𝒌n(R),p,𝒌(λ)\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda), we have

𝔼Bp,𝒌n(R),p,𝒌,λ(|X|)=αBp,𝒌n(R),p,𝒌(λ)V(1+on(1))log(2/cp)n2nV.\mathbb{E}_{B^{n}_{p,\bm{k}}(R),p,\bm{k},\lambda}(|X|)=\alpha_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lambda)\cdot V\geq(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}\cdot V.

Let ϵ\epsilon be a small positive number (we will choose ϵ\epsilon dependent on nn but independent of VV) and α=(1+on(1))log(2/cp)n2nϵ\alpha=(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}-\epsilon. So for large VV,

α(1+on(1))log(2/cp)n2nV1/5tV.\alpha\leq(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}-V^{-1/5}\leq\frac{t}{V}.

Applying Lemma 5.3, it follows that

fp,𝒌,n(α)=limV1αVlogZ^Bp,𝒌n(R),p,𝒌(αV)VαV/αV!limV1tlogZ^Bp,𝒌n(R),p,𝒌(t)Vt/t!.f_{p,\bm{k},n}(\alpha)=\lim_{V\rightarrow\infty}\frac{1}{\alpha V}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(\lfloor\alpha V\rfloor)}{V^{\lfloor\alpha V\rfloor}/\lfloor\alpha V\rfloor!}\geq\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(t)}{V^{t}/t!}.

We calculate

limV1tlogZ^Bp,𝒌n(R),p,𝒌(t)Vt/t!limV1tlogt!Vt+1λt=limV1tlogt!Vt+1logλlimV1tlogt!Vt+1+nlogcplog2.\begin{split}\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{\hat{Z}_{B^{n}_{p,\bm{k}}(R),p,\bm{k}}(t)}{V^{t}/t!}\geq&\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{t!}{V^{t+1}\lambda^{t}}\\ =&\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{t!}{V^{t+1}}-\log\lambda\\ \geq&\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{t!}{V^{t+1}}+n\log c_{p}-\log 2.\end{split}

Note that tt\rightarrow\infty when VV\rightarrow\infty. Hence Stirling’s formula implies that

limV1tlogt!Vt+1=limV1tlog2πt(t/e)tVt+1=limV(1tlog2πt+log(t/V)11tlog(1/V))=limVlog(t/V)1log((1+on(1))log(2/cp)n2n)1=(log2+on(1))n.\begin{split}\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{t!}{V^{t+1}}=&\lim_{V\rightarrow\infty}\frac{1}{t}\log\frac{\sqrt{2\pi t}(t/e)^{t}}{V^{t+1}}\\ =&\lim_{V\rightarrow\infty}\left(\frac{1}{t}\log\sqrt{2\pi t}+\log(t/V)-1-\frac{1}{t}\log(1/V)\right)\\ =&\lim_{V\rightarrow\infty}\log(t/V)-1\\ \geq&\log\left((1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}\right)-1\\ =&(-\log 2+o_{n}(1))n.\end{split}

Therefore,

fp,𝒌,n(α)(log2+on(1))n+nlogcplog2=(1+on(1))log(2/cp)n.f_{p,\bm{k},n}(\alpha)\geq(-\log 2+o_{n}(1))n+n\log c_{p}-\log 2=-(1+o_{n}(1))\log(2/c_{p})\cdot n.

Choose ϵ=on(log(2/cp)n2n)\epsilon=o_{n}\left(\frac{\log(2/c_{p})\cdot n}{2^{n}}\right) so that α=(1+on(1))log(2/cp)n2nϵ=(1+on(1))log(2/cp)n2n\alpha=(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}-\epsilon=(1+o_{n}(1))\frac{\log(2/c_{p})\cdot n}{2^{n}}, completing the proof. ∎

Acknowledgements

The authors would like to thank Professor Yufei Zhao and Professor Chuanming Zong for many helpful and valuable comments on the manuscript. The authors also thank Professor Hong Liu for sharing his nice idea about the alternate proof.

References

  • [1] K. Ball. A lower bound for the optimal density of lattice packings. Internat. Math. Res. Notices, (10):217–221, 1992.
  • [2] J. A. Clarkson. Uniformly convex spaces. Trans. Amer. Math. Soc., 40(3):396–414, 1936.
  • [3] H. Cohn and N. Elkies. New upper bounds on sphere packings. I. Ann. of Math. (2), 157(2):689–714, 2003.
  • [4] H. Cohn, A. Kumar, S. D. Miller, D. Radchenko, and M. Viazovska. The sphere packing problem in dimension 24. Ann. of Math. (2), 185(3):1017–1033, 2017.
  • [5] H. Cohn and Y. Zhao. Sphere packing bounds via spherical codes. Duke Math. J., 163(10):1965–2002, 2014.
  • [6] N. D. Elkies, A. M. Odlyzko, and J. A. Rush. On the packing densities of superballs and other bodies. Invent. Math., 105(3):613–639, 1991.
  • [7] I. G. Fernández, J. Kim, H. Liu, and O. Pikhurko. New lower bounds on kissing numbers and spherical codes in high dimensions. arXiv e-prints, page arXiv:2111.01255, Nov. 2021.
  • [8] T. C. Hales. A proof of the Kepler conjecture. Ann. of Math. (2), 162(3):1065–1185, 2005.
  • [9] E. Hlawka. Zur Geometrie der Zahlen. Math. Z., 49:285–312, 1943.
  • [10] M. Jenssen, F. Joos, and W. Perkins. On kissing numbers and spherical codes in high dimensions. Adv. Math., 335:307–321, 2018.
  • [11] M. Jenssen, F. Joos, and W. Perkins. On the hard sphere model and sphere packings in high dimensions. Forum Math. Sigma, 7:Paper No. e1, 19, 2019.
  • [12] G. A. Kabatjanskiĭ and V. I. Levenšteĭn. Bounds for packings on the sphere and in space. Problemy Peredači Informacii, 14(1):3–25, 1978.
  • [13] J. Kim, H. Liu, and T. Tran. Exponential decay of intersection volume with applications on list-decodability and Gilbert-Varshamov type bound. arXiv e-prints, page arXiv:2112.11274, Dec. 2021.
  • [14] L. Liu and C. P. Xing. Packing superballs from codes and algebraic curves. Acta Math. Sin. (Engl. Ser.), 24(1):1–6, 2008.
  • [15] F. Pirot and E. Hurley. Colouring locally sparse graphs with the first moment method. arXiv e-prints, page arXiv:2109.15215, Sept. 2021.
  • [16] C. A. Rogers. Existence theorems in the geometry of numbers. Ann. of Math. (2), 48:994–1002, 1947.
  • [17] C. A. Rogers. The number of lattice points in a set. Proc. London Math. Soc. (3), 6:305–320, 1956.
  • [18] C. A. Rogers. Lattice covering of space: The Minkowski-Hlawka theorem. Proc. London Math. Soc. (3), 8:447–465, 1958.
  • [19] J. A. Rush. A lower bound on packing density. Invent. Math., 98(3):499–509, 1989.
  • [20] J. A. Rush. Constructive packings of cross polytopes. Mathematika, 38(2):376–380 (1992), 1991.
  • [21] J. A. Rush. An indexed set of density bounds on lattice packings. Geom. Dedicata, 53(2):217–221, 1994.
  • [22] J. A. Rush. Lattice packing of nearly-Euclidean balls in spaces of even dimension. Proc. Edinburgh Math. Soc. (2), 39(1):163–169, 1996.
  • [23] J. A. Rush and N. J. A. Sloane. An improvement to the Minkowski-Hlawka bound for packing superballs. Mathematika, 34(1):8–18, 1987.
  • [24] A. Sah, M. Sawhney, D. Stoner, and Y. Zhao. Exponential improvements for superball packing upper bounds. Adv. Math., 365:107056, 9, 2020.
  • [25] N. T. Sardari and M. Zargar. New upper bounds for spherical codes and packings. arXiv e-prints, page arXiv:2001.00185, Jan. 2020.
  • [26] W. M. Schmidt. The measure of the set of admissible lattices. Proc. Amer. Math. Soc., 9:390–403, 1958.
  • [27] W. M. Schmidt. Masstheorie in der Geometrie der Zahlen. Acta Math., 102:159–224, 1959.
  • [28] W. M. Schmidt. On the Minkowski-Hlawka theorem. Illinois J. Math., 7:18–23, 1963.
  • [29] J. van der Corput and G. Schaake. Anwendung einer blichfeldtschen beweismethode in der geometrie der zahlen. Acta Arithmetica, 2:152–160, 1936.
  • [30] A. Venkatesh. A note on sphere packings in high dimension. Int. Math. Res. Not. IMRN, (7):1628–1642, 2013.
  • [31] M. S. Viazovska. The sphere packing problem in dimension 8. Ann. of Math. (2), 185(3):991–1015, 2017.