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

New upper bounds for spherical codes and packings

Naser Talebizadeh Sardari and Masoud Zargar Penn State department of Mathematics, McAllister Building, Pollock Rd, State College, PA 16802 USA [email protected] Max Planck Institute for Mathematics, Vivatsgasse 7, 53111 Bonn, Germany [email protected]
Abstract.

We improve the previously best known upper bounds on the sizes of θ\theta-spherical codes for every θ<θ62.997\theta<\theta^{*}\approx 62.997^{\circ} at least by a factor of 0.43250.4325, in sufficiently high dimensions. Furthermore, for sphere packing densities in dimensions n2000n\geq 2000 we have an improvement at least by a factor of 0.4325+51n0.4325+\frac{51}{n}. Our method also breaks many non-numerical sphere packing density bounds in smaller dimensions. This is the first such improvement for each dimension since the work of Kabatyanskii and Levenshtein [KL78] and its later improvement by Levenshtein [Lev79]. Novelties of this paper include the analysis of triple correlations, usage of the concentration of mass in high dimensions, and the study of the spacings between the roots of Jacobi polynomials.

1. Introduction

1.1. Spherical codes and packings

Packing densities have been studied extensively, for purely mathematical reasons as well as for their connections to coding theory. The work of Conway and Sloane is a comprehensive reference for this subject [CS99]. We proceed by defining the basics of this subject. Consider n\mathbb{R}^{n} equipped with the Euclidean metric |.||.| and the associated volume vol(.)\textup{vol}(.). For each real r>0r>0 and each xnx\in\mathbb{R}^{n}, we denote by Bn(x,r)B_{n}(x,r) the open ball in n\mathbb{R}^{n} centered at xx and of radius rr. For each discrete set of points SnS\subset\mathbb{R}^{n} such that any two distinct points x,ySx,y\in S satisfy |xy|2|x-y|\geq 2, we can consider

𝒫:=𝓍𝒮𝓃(𝓍,1),\cal{P}:=\cup_{x\in S}B_{n}(x,1),

the union of non-overlapping unit open balls centered at the points of SS. This is called a sphere packing (SS may vary), and we may associate to it the function mapping each real r>0r>0 to

δ𝒫(r):=vol(𝒫𝓃(0,𝓇))vol(Bn(0,r)).\delta_{\cal{P}}(r):=\frac{\textup{vol}(\cal{P}\cap B_{n}(0,r))}{\textup{vol}(B_{n}(0,r))}.

The packing density of 𝒫\cal{P} is defined as

δ𝒫:=lim suprδ𝒫(r).\delta_{\cal{P}}:=\limsup_{r\rightarrow\infty}\delta_{\cal{P}}(r).

Clearly, this is a finite number. The maximal sphere packing density in n\mathbb{R}^{n} is defined as

δn:=sup𝒫𝓃δ𝒫,\delta_{n}:=\sup_{\cal{P}\subset\mathbb{R}^{n}}\delta_{\cal{P}},

a supremum over all sphere packings 𝒫\cal{P} of n\mathbb{R}^{n} by non-overlapping unit balls.

The linear programming method initiated by Delsarte is a powerful tool for giving upper bounds on sphere packing densities [Del72]. That being said, we only know the optimal sphere packing densities in dimensions 1,2,3,8 and 24 [FT43, Hal05, Via17, CKM+17]. Very recently, the first author proved an optimal upper bound on the sphere packing density of all but a tiny fraction of even unimodular lattices in high dimensions; see [Sar19, Theorem 1.1].

The best known linear programming upper bounds on sphere packing densities in low dimensions are based on the linear programming method developed by Cohn–Elkies [CE03] which itself was inspired by Delsarte’s linear programming method. As far as the exponent is concerned, in high dimensions, the best asymptotic upper bound goes back to Kabatyanskii–Levenshtein from 1978 [KL78] stating that δn2(0.599+o(1))n\delta_{n}\leq 2^{-(0.599+o(1))n} as nn\rightarrow\infty. More recently, de Laat–de Oliveira Filho–Vallentin improved upper bounds in very low dimensions using the semi-definite programming method [dLdOFV14], partially based on the semi-definite programming method developed by Bachoc–Vallentin [BV08] for bounding kissing numbers. The work of Bachoc–Vallentin was further improved by Mittelmann–Vallentin [MV10], Machado–de Oliveira Filho [MdOF18], and very recently after the writing of our paper by de Laat–Leijenhorst [Ld22].

Another recent development is the discovery by Hartman–Mazác–Rastelli [HMR19] of a connection between the spinless modular bootstrap for two-dimensional conformal field theories and the linear programming bound for sphere packing densities. After the writing of our paper, Afkhami-Jeddi–Cohn–Hartman–de Laat–Tajdini [AJCH+20] numerically constructed solutions to the Cohn–Elkies linear programming problem and conjectured that the linear programming method is capable of producing an upper bound on sphere packing densities in high dimensions that is exponentially better than that of Kabatyanskii–Levenshtein.

A notion closely related to sphere packings in Euclidean spaces is that of spherical codes. By inequalities relating sphere packing densities to the sizes of spherical codes, Kabatyanskii–Levenshtein [KL78] obtained their bound on sphere packing densities stated above. The sizes of spherical codes are bounded from above using Delsarte’s linear programming method. In what follows, we define spherical codes and this linear programming method.

Given Sn1S^{n-1}, the unit sphere in n\mathbb{R}^{n}, a θ\theta-spherical code is a finite subset ASn1A\subset S^{n-1} such that no two distinct x,yAx,y\in A are at an angular distance less than θ\theta. For each 0<θπ0<\theta\leq\pi, we define M(n,θ)M(n,\theta) to be the largest cardinality of a θ\theta-spherical code ASn1.A\subset S^{n-1}.

The Delsarte linear programming method is applied to spherical codes as follows. Throughout this paper, we work with probability measures μ\mu on [1,1].[-1,1]. μ\mu gives an inner product on the \mathbb{R}-vector space of real polynomials [t]\mathbb{R}[t], and let {pi}i=0\{p_{i}\}_{i=0}^{\infty} be an orthonormal basis with respect to μ\mu such that the degree of pip_{i} is ii and pi(1)>0p_{i}(1)>0 for every ii. Note that p0=1p_{0}=1. Suppose that the basis elements pkp_{k} define positive definite functions on Sn1S^{n-1}, that is,

(1) xi,xjAhihjpk(xi,xj)0\sum_{x_{i},x_{j}\in A}h_{i}h_{j}p_{k}\left(\left<x_{i},x_{j}\right>\right)\geq 0

for any finite subset ASn1A\subset S^{n-1} and any real numbers hi.h_{i}\in\mathbb{R}. An example of a probability measure satisfying inequality (1) is

dμα=(1t2)α11(1t2)α𝑑tdt,d\mu_{\alpha}=\frac{(1-t^{2})^{\alpha}}{\int_{-1}^{1}(1-t^{2})^{\alpha}dt}dt,

where αn32\alpha\geq\frac{n-3}{2} and 2α2\alpha\in\mathbb{Z}. Given s[1,1]s\in[-1,1], consider the space D(μ,s)D(\mu,s) of all functions f(t)=i=0fipi(t)f(t)=\sum_{i=0}^{\infty}f_{i}p_{i}(t), fif_{i}\in\mathbb{R}, such that

  1. (1)

    fi0f_{i}\geq 0 for every ii, and f0>0,f_{0}>0,

  2. (2)

    f(t)0f(t)\leq 0 for 1ts.-1\leq t\leq s.

Suppose 0<θ<π0<\theta<\pi, and A={x1,,xN}A=\{x_{1},\ldots,x_{N}\} is a θ\theta-spherical code in Sn1S^{n-1}. Given a function fD(μ,cosθ)f\in D(\mu,\cos\theta), we consider

i,jf(xi,xj).\sum_{i,j}f(\left<x_{i},x_{j}\right>).

This may be written in two different ways as

Nf(1)+ijf(xi,xj)=f0N2+k=1fki,jpk(xi,xj).Nf(1)+\sum_{i\neq j}f(\left<x_{i},x_{j}\right>)=f_{0}N^{2}+\sum_{k=1}^{\infty}f_{k}\sum_{i,j}p_{k}(\left<x_{i},x_{j}\right>).

Since fD(μ,cosθ)f\in D(\mu,\cos\theta) and xi,xjcosθ\left<x_{i},x_{j}\right>\leq\cos\theta for every iji\neq j, this gives us the inequality

Nf(1)f0.N\leq\frac{f(1)}{f_{0}}.

We define

(2) (𝒻):=𝒻(1)𝒻0.\cal{L}(f):=\frac{f(1)}{f_{0}}.

In particular, this method leads to the linear programming bound

(3) M(n,θ)inffD(dμn32,cosθ)(𝒻).M(n,\theta)\leq\inf_{f\in D(d\mu_{\frac{n-3}{2}},\cos\theta)}\cal{L}(f).

One of the novelties of our work is the construction using triple points of new test functions in Section 3 satisfying conditions (1) and (2) of the Delsarte linear programming method. In fact, our functions are infinite linear combinations of coefficients of the matrices appearing in Theorem 3.2 of Bachoc–Vallentin [BV08]. Bachoc–Vallentin use semi-definite programming to obtain an upper bound on kissing numbers M(n,π3)M(n,\frac{\pi}{3}) by summing over triples of points in spherical codes. On the other hand, we average one of the three points over the sphere, and take the other two points from the spherical code. Semi-definite programming is computationally feasible in very low dimensions, and improves upon linear programming bounds [dDV22]. After the writing of our paper, a semi-definite programming method using triple point correlations for sphere packings was developed by Cohn–de Laat–Salmon [CdS22], improving upon linear programming bounds on sphere packings in special low dimensions. In the semi-definite programming methods, the functions are numerically constructed. Furthermore, in high dimensions, there is no asymptotic bound using semi-definite programming which improves upon the linear programming bound of Kabatyanskii–Levenshtein [KL78], even up to a constant factor. The functions that we non-numerically construct improve upon [KL78] by a constant factor. In the same spirit, we also improve upon sphere packing density upper bounds in high dimensions.

Upper bounds on spherical codes are used to obtain upper bounds on sphere packing densities through inequalities proved using geometric methods. For example, for any 0<θπ/20<\theta\leq\pi/2, Sidelnikov [Sid74] proved using an elementary argument that

(4) δnsinn(θ/2)M(n+1,θ).\delta_{n}\leq\sin^{n}(\theta/2)M(n+1,\theta).

Let 0<θ<θπ0<\theta<\theta^{\prime}\leq\pi. We write λn(θ,θ)\lambda_{n}(\theta,\theta^{\prime}) for the ratio of volume of the spherical cap with radius sin(θ/2)sin(θ/2)\frac{\sin(\theta/2)}{\sin(\theta^{\prime}/2)} on the unit sphere Sn1S^{n-1} to the volume of the whole sphere. Sidelnikov [Sid74] used a similar argument to show that for 0<θ<θπ0<\theta<\theta^{\prime}\leq\pi

(5) M(n,θ)M(n+1,θ)λn(θ,θ).M(n,\theta)\leq\frac{M(n+1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})}.

Kabatyanskii–Levenshstein used the Delsarte linear programming method and Jacobi polynomials to give an upper bound on M(n,θ)M(n,\theta) [KL78]. A year later, Levenshtein [Lev79] found optimal polynomials up to a certain degree and improved the Kabatyanskii–Levenshtein bound by a constant factor. Levenshtein obtained the upper bound

(6) M(n,θ)MLev(n,θ),M(n,\theta)\leq M_{\textup{Lev}}(n,\theta),

where

(7) MLev(n,θ):={2(d+n1n1) if t1,dα+1,α<cos(θ)t1,dα+1,α+1,(d+n1n1)+(d+n2n1) if t1,d1α+1,α+1<cos(θ)t1,dα+1,α.M_{\textup{Lev}}(n,\theta):=\begin{cases}2{\binom{d+n-1}{n-1}}&\text{ if }t^{\alpha+1,\alpha}_{1,d}<\cos(\theta)\leq t^{\alpha+1,\alpha+1}_{1,d},\\ \binom{d+n-1}{n-1}+\binom{d+n-2}{n-1}&\text{ if }t^{\alpha+1,\alpha+1}_{1,d-1}<\cos(\theta)\leq t^{\alpha+1,\alpha}_{1,d}.\end{cases}

Here, α=n32\alpha=\frac{n-3}{2} and t1,dα,βt^{\alpha,\beta}_{1,d} is the largest root of the Jacobi polynomial pdα,βp^{\alpha,\beta}_{d} of degree d=d(n,θ)d=d(n,\theta), a function of nn and θ\theta. We carefully define d=d(n,θ)d=d(n,\theta) and Levenshtein’s optimal polynomials in Subsection 5.1. Also see Subsection 5.1 for the definition and properties of Jacobi polynomials.

Throughout this paper, θ=62.997\theta^{*}=62.997...^{\circ} is the unique root of the equation

cosθlog(1+sinθ1sinθ)(1+cosθ)sinθ=0\cos\theta\log(\frac{1+\sin\theta}{1-\sin\theta})-(1+\cos\theta)\sin\theta=0

in (0,π/2)(0,\pi/2). As demonstrated by Kabatyanskii–Levenshtein in  [KL78], for 0<θ<θ0<\theta<\theta^{*}, the bound M(n,θ)MLev(n,θ)M(n,\theta)\leq M_{\textup{Lev}}(n,\theta) is asymptotically exponentially weaker in nn than

(8) M(n,θ)MLev(n+1,θ)λn(θ,θ)M(n,\theta)\leq\frac{M_{\textup{Lev}}(n+1,\theta^{*})}{\lambda_{n}(\theta,\theta^{*})}

obtained from inequality (5). Barg–Musin [BM07, p.11 (8)], based on the work [AVZ00] of Agrell–Vargy–Zeger, improved inequality (5) and showed that

(9) M(n,θ)M(n1,θ)λn(θ,θ)M(n,\theta)\leq\frac{M(n-1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})}

whenever π>θ>2arcsin(12cos(θ/2))\pi>\theta^{\prime}>2\arcsin\left(\frac{1}{2\cos(\theta/2)}\right). Cohn–Zhao [CZ14] improved sphere packing density upper bounds by combining the upper bound of Kabatyanskii–Levenshtein on M(n,θ)M(n,\theta) with their analogue [CZ14, Proposition 2.1] of (9) stating that for π/3θπ\pi/3\leq\theta\leq\pi,

(10) δnsinn(θ/2)M(n,θ),\delta_{n}\leq\sin^{n}(\theta/2)M(n,\theta),

leading to better bounds than those obtained from (4). Aside from our main results discussed in the next subsection, in Proposition 2.2 of Section 2 we remove the angular restrictions on inequality (9) in the large nn regime by using a concentration of mass phenomenon. An analogous result removes the restriction θπ/3\theta\geq\pi/3 on inequality (10) for large nn.

Inequalities (9) and (10) give, respectively, the bounds

(11) M(n,θ)MLev(n1,θ)λn(θ,θ),M(n,\theta)\leq\frac{M_{\textup{Lev}}(n-1,\theta^{*})}{\lambda_{n}(\theta,\theta^{*})},

for θ\theta and θ\theta^{*} restricted as in the conditions for inequality (9), and

(12) δnsinn(θ/2)MLev(n,θ).\delta_{n}\leq\sin^{n}(\theta^{*}/2)M_{\textup{Lev}}(n,\theta^{*}).

Prior to our work, the above were the best bounds for large enough nn. In Theorems 1.1 and 1.2, we improve both by a constant factor for large nn, the first such improvement since the work of Levenshtein [Lev79] more than forty years ago. We also relax the angular condition in inequality (11) to 0<θ<θ0<\theta<\theta^{*} for large nn. For θθπ\theta^{*}\leq\theta\leq\pi, M(n,θ)MLev(n,θ)M(n,\theta)\leq M_{\textup{Lev}}(n,\theta) is still the best bound. We also prove a number of other results, including the construction of general test functions in Section 3 that are of independent interest.

1.2. Main results and general strategy

We improve inequalities (11) and (12) with an extra factor 0.43250.4325 for each sufficiently large nn. In the case of sphere packings, we obtain an improvement by a factor of 0.4325+51n0.4325+\frac{51}{n} for dimensions n2000n\geq 2000. In low dimensions, our geometric ideas combined with numerics lead to improvements that are better than 0.43250.4325. In Section 6, we provide the results of our extensive numerical calculations. We now state our main theorems.

Theorem 1.1.

Suppose that 0<θ<θ0<\theta<\theta^{*}. Then

M(n,θ)cnMLev(n1,θ)λn(θ,θ),M(n,\theta)\leq c_{n}\frac{M_{\textup{Lev}}(n-1,\theta^{*})}{\lambda_{n}(\theta,\theta^{*})},

where cn0.4325c_{n}\leq 0.4325 for large enough nn independent of θ.\theta.

We also have a uniform version of this theorem for sphere packing densities.

Theorem 1.2.

Suppose that 13cos(θ)12\frac{1}{3}\leq\cos(\theta)\leq\frac{1}{2}. We have

δncn(θ)sinn(θ/2)MLev(n,θ),\delta_{n}\leq c_{n}(\theta)\sin^{n}(\theta/2)M_{\textup{Lev}}(n,\theta),

where cn(θ)<1c_{n}(\theta)<1 for every nn. If, additionally, n2000n\geq 2000 we have cn(θ)0.515+74nc_{n}(\theta)\leq 0.515+\frac{74}{n}. Furthermore, for n2000n\geq 2000 we have cn(θ)0.4325+51nc_{n}(\theta^{*})\leq 0.4325+\frac{51}{n}.

By Kabatyanskii–Levenshtein [KL78], the best bound on sphere packing densities δn\delta_{n} for large nn comes from θ=θ\theta=\theta^{*}; comparisons using other angles are exponentially worse. Consequently, this theorem implies that we have an improvement by 0.43250.4325 for sphere packing density upper bounds in high dimensions. Furthermore, note that the constants of improvement cn(θ)c_{n}(\theta) are bounded from above uniformly in θ\theta. The lower bound in 13cosθ12\frac{1}{3}\leq\cos\theta\leq\frac{1}{2} is not conceptually significant in the sense that a change in 13\frac{1}{3} would lead to a change in the bound cn(θ)0.515+74nc_{n}(\theta)\leq 0.515+\frac{74}{n}.

We prove Theorem 1.2 by constructing a new test function that satisfies the Cohn–Elkies linear programming conditions.

Refer to caption
Figure 1. Schematic diagram for sphere packings.

The geometric idea behind the construction of our new test functions is the following (see Figure 1). In [CZ14], for every πθπ3\pi\geq\theta\geq\frac{\pi}{3}, Cohn and Zhao choose a ball of radius r=12sin(θ/2)r=\frac{1}{2\sin(\theta/2)} around each point of the sphere packing so that for every two points 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} that are centers of balls in the sphere packing, for every point 𝒛\boldsymbol{z} in the shaded region, 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} make an angle φθ\varphi\geq\theta with respect to 𝒛\boldsymbol{z}. By averaging a function satisfying the Delsarte linear programming conditions for the angle θ\theta, a new function satisfying the conditions of the Cohn–Elkies linear programming method is produced. The negativity condition easily follows as φθ\varphi\geq\theta for every point 𝒛\boldsymbol{z} in the shaded region. Our insight is that since we are taking an average, we do not need pointwise negativity for each point 𝒛\boldsymbol{z} to ensure the negativity condition for the new averaged function. In fact, we can enlarge the radii of the balls by a quantity δ=O(1/n)\delta=O(1/n) so that the conditions of the Cohn–Elkies linear programming method continue to hold for this averaged function. Since the condition φθ\varphi\geq\theta is no longer satisfied for every point 𝒛\boldsymbol{z}, determining how large δ\delta may be chosen is delicate. We develop analytic methods for determining such δ\delta. This requires us to estimate triple density functions in Section 4, and estimating the Jacobi polynomials near their extreme roots in Subsection 5.1. It is known that the latter problem is difficult [Kra07, Conjecture 1]. In Subsection 5.1, we treat this difficulty by using the relation between the zeros of Jacobi polynomials; these ideas go back to the work of Stieltjes [Sti87]. More precisely, we use the underlying differential equations satisfied by Jacobi polynomials, and the fact that the roots of the family of Jacobi polynomials are interlacing. For Theorem 1.1, we use a similar idea but consider how much larger we can make certain appropriately chosen strips on a sphere. See Figure 3.

nn Rogers Levenshtein79 K.–L. Cohn–Zhao C.–Z.+L79 Our bound
12 8.759×1028.759\times 10^{-2} 1.065×1011.065\times 10^{-1} 1.038×1001.038\times 10^{0} 9.666×1019.666\times 10^{-1} 3.253×1013.253\times 10^{-1} 1.228×1011.228\times 10^{-1}
24 2.456×1032.456\times 10^{-3} 3.420×1033.420\times 10^{-3} 2.930×1022.930\times 10^{-2} 2.637×1022.637\times 10^{-2} 8.464×1038.464\times 10^{-3} 3.194×1033.194\times 10^{-3}
36 5.527×1055.527\times 10^{-5} 8.109×1058.109\times 10^{-5} 5.547×1045.547\times 10^{-4} 4.951×1044.951\times 10^{-4} 1.610×1041.610\times 10^{-4} 6.035×1056.035\times 10^{-5}
48 1.128×1061.128\times 10^{-6} 1.643×1061.643\times 10^{-6} 8.745×1068.745\times 10^{-6} 7.649×1067.649\times 10^{-6} 2.534×1062.534\times 10^{-6} 9.487×1079.487\times 10^{-7}
60 2.173×1082.173\times 10^{-8} 3.009×1083.009\times 10^{-8} 1.223×1071.223\times 10^{-7} 1.046×1071.046\times 10^{-7} 3.521×1083.521\times 10^{-8} 1.317×1081.317\times 10^{-8}
72 4.039×10104.039\times 10^{-10} 5.135×10105.135\times 10^{-10} 1.550×1091.550\times 10^{-9} 1.322×1091.322\times 10^{-9} 4.496×10104.496\times 10^{-10} 1.678×10101.678\times 10^{-10}
84 7.315×10127.315\times 10^{-12} 8.312×10128.312\times 10^{-12} 1.850×10111.850\times 10^{-11} 1.574×10111.574\times 10^{-11} 5.381×10125.381\times 10^{-12} 2.007×10122.007\times 10^{-12}
96 1.300×10131.300\times 10^{-13} 1.291×10131.291\times 10^{-13} 2.111×10132.111\times 10^{-13} 1.786×10131.786\times 10^{-13} 6.101×10146.101\times 10^{-14} 2.273×10142.273\times 10^{-14}
108 2.277×10152.277\times 10^{-15} 1.937×10151.937\times 10^{-15} 2.320×10152.320\times 10^{-15} 1.942×10151.942\times 10^{-15} 6.662×10166.662\times 10^{-16} 2.480×10162.480\times 10^{-16}
120 3.940×10173.940\times 10^{-17} 2.826×10172.826\times 10^{-17} 2.452×10172.452\times 10^{-17} 2.051×10172.051\times 10^{-17} 7.058×10187.058\times 10^{-18} 2.626×10182.626\times 10^{-18}
Table 1. Upper bounds on maximal sphere packing densities δn\delta_{n} in n\mathbb{R}^{n} using functions that are not constructed using computer assistance. The last column is obtained using the method of this paper.

We use our geometrically constructed test functions to obtain the last column of Table 1. Table 1 is a comparison of upper bounds on sphere packing densities. This table does not include the computer-assisted mathematically rigorous bounds obtained using the Cohn–Elkies linear programming method or those of semi-definite programming. We now describe the different columns. The Rogers column corresponds to the bounds on sphere packing densities obtained by Rogers [Rog58]. The Levenshtein79 column corresponds to the bound obtained by Levenshtein in terms of roots of Bessel functions [Lev79]. The K.–L. column corresponds to the bound on M(n,θ)M(n,\theta) proved by Kabatyanskii and Levenshtein [KL78] combined with inequality (5). The Cohn–Zhao column corresponds to the column found in the work of Cohn and Zhao [CZ14]; they combined their inequality (10) with the bound on M(n,θ)M(n,\theta) proved by Kabatyanskii–Levenshtein [KL78]. We also include the column C.–Z.+L79 which corresponds to combining Cohn and Zhao’s inequality with improved bounds on M(n,θ)M(n,\theta) using Levenshtein’s optimal polynomials [Lev79]. The final column corresponds to the bounds on sphere packing densities obtained by our method. In Table 1, the highlighted entries are the best bounds obtained from these methods. Our bounds break most of the other bounds also in low dimensions. Our bounds are obtained from explicit geometrically constructed functions satisfying the Cohn–Elkies linear programming method. Our method only involves explicit integral calculations; in contrast to the numerical method in [CE03], we do not rely on any searching algorithm. Moreover, compared to the Cohn–Elkies linear programming method, in n=120n=120 dimensions, we improve upon the sphere packing density upper bound of 1.164×10171.164\times 10^{-17} obtained by forcing eight double roots.

1.3. Structure of the paper

In Section 2, we setup some of the notation used in this paper and prove Proposition 2.2. Section 3 concerns the general construction of our test functions that are used in conjunction with the Delsarte and Cohn–Elkies linear programming methods. In Section 5, we prove our main Theorems 1.1 and 1.2. In this section, we use our estimates on the triple density functions proved in Section 4. In Subsection 5.1, we describe Jacobi polynomials, Levenshtein’s optimal polynomials, and locally approximate Jacobi polynomials near their largest roots. In the final Section 6, we provide a table of improvement factors.

Acknowledgments. Both authors are thankful to Alexander Barg, Peter Boyvalenkov, Henry Cohn, Matthew de Courcy-Ireland, Mehrdad Khani Shirkoohi, Peter Sarnak, and Hamid Zargar. N.T.Sardari’s work is supported partially by the National Science Foundation under Grant No. DMS-2015305 N.T.S and is grateful to the Institute for Advanced Study and the Max Planck Institute for Mathematics in Bonn for their hospitality and financial support. M.Zargar was supported by the Max Planck Institute for Mathematics in Bonn, SFB1085 at the University of Regensburg, and the University of Southern California.

2. Geometric improvement

In this section, we prove Proposition 2.2, improving inequality (9) by removing the restrictions on the angles for large dimensions nn. This is achieved via a concentration of mass phenomenon, at the expense of an exponentially decaying error term. First, we introduce some notations that we use throughout this paper.

Let 0<θ<θ<π0<\theta<\theta^{\prime}<\pi be given angles, and let s:=cosθs:=\cos\theta and s:=cosθs^{\prime}:=\cos\theta^{\prime}. Throughout, Sn1S^{n-1} is the unit sphere centered at the origin of n\mathbb{R}^{n}. Suppose 𝒛Sn1\boldsymbol{z}\in S^{n-1} is a fixed point.

Consider the hyperplane N𝒛:={𝒘n:𝒘,𝒛=0}N_{\boldsymbol{z}}:=\{\boldsymbol{w}\in\mathbb{R}^{n}:\left<\boldsymbol{w},\boldsymbol{z}\right>=0\}. For each 𝒂,𝒃Sn1\boldsymbol{a},\boldsymbol{b}\in S^{n-1} of radial angle at least θ\theta from each other, we may orthogonally project them onto N𝒛N_{\boldsymbol{z}} via the map Π𝒛:Sn1{±𝒛}N𝒛\Pi_{\boldsymbol{z}}:S^{n-1}\setminus\{\pm\boldsymbol{z}\}\rightarrow N_{\boldsymbol{z}}. For every 𝒂Sn1\boldsymbol{a}\in S^{n-1},

Π𝒛(𝒂)=𝒂𝒂,𝒛𝒛.\Pi_{\boldsymbol{z}}(\boldsymbol{a})=\boldsymbol{a}-\left<\boldsymbol{a},\boldsymbol{z}\right>\boldsymbol{z}.

For brevity, when 𝒂±𝒛\boldsymbol{a}\neq\pm\boldsymbol{z} we denote Π𝒛(𝒂)|Π𝒛(𝒂)|\frac{\Pi_{\boldsymbol{z}}(\boldsymbol{a})}{|\Pi_{\boldsymbol{z}}(\boldsymbol{a})|} by 𝒂~\tilde{\boldsymbol{a}} lying inside the unit sphere Sn2S^{n-2} in N𝒛N_{\boldsymbol{z}} centered at the origin. Given 𝒂,𝒃Sn1{±𝒛}\boldsymbol{a},\boldsymbol{b}\in S^{n-1}\setminus\{\pm\boldsymbol{z}\}, we obtain points 𝒂~,𝒃~Sn2\tilde{\boldsymbol{a}},\tilde{\boldsymbol{b}}\in S^{n-2}. We will use the following notation.

u:=𝒂,𝒛,u:=\left<\boldsymbol{a},\boldsymbol{z}\right>,
v:=𝒃,𝒛,v:=\left<\boldsymbol{b},\boldsymbol{z}\right>,

and

t:=𝒂,𝒃.t:=\left<\boldsymbol{a},\boldsymbol{b}\right>.

It is easy to see that

(13) 𝒂~,𝒃~=tuv(1u2)(1v2).\left<\tilde{\boldsymbol{a}},\tilde{\boldsymbol{b}}\right>=\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}.
Refer to caption
Figure 2. Spherical cap Capθ,θ(𝒛)\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z}) containing points 𝒂,𝒃\boldsymbol{a},\boldsymbol{b}.

Consider the cap Capθ,θ(𝒛)\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z}) on Sn1S^{n-1} centered at 𝒛\boldsymbol{z} and of radius sin(θ/2)sin(θ/2)\frac{\sin(\theta/2)}{\sin(\theta^{\prime}/2)}. Capθ,θ(𝒛)\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z}) is the spherical cap centered at 𝒛\boldsymbol{z} with the defining property that any two points 𝒂\boldsymbol{a}, 𝒃\boldsymbol{b} on its boundary of radial angle θ\theta are sent to points 𝒂~\tilde{\boldsymbol{a}}, 𝒃~Sn2\tilde{\boldsymbol{b}}\in S^{n-2} having radial angle θ\theta^{\prime}.

It will be convenient for us to write the radius of the cap as 1r2=sin(θ/2)sin(θ/2)\sqrt{1-r^{2}}=\frac{\sin(\theta/2)}{\sin(\theta^{\prime}/2)}, from which it follows that

(14) r=ss1s.r=\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}.

This rr is the distance from the center of the cross-section defining the cap Capθ,θ(𝒛)\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z}) to the center of Sn1S^{n-1}. In fact,

Capθ,θ(𝒛)={𝒂Sn1:𝒂,𝒛r}.\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z})=\{\boldsymbol{a}\in S^{n-1}:\left<\boldsymbol{a},\boldsymbol{z}\right>\geq r\}.


For 0<θ<θ<π0<\theta<\theta^{\prime}<\pi, we define

(15) γθ,θ:=2arctans(1s)(ss)+arccos(r)π,\gamma_{\theta,\theta^{\prime}}:=2\arctan\frac{s}{\sqrt{(1-s)(s-s^{\prime})}}+\arccos(r)-\pi,

and

(16) R:=cos(γθ,θ).R:=\cos(\gamma_{\theta,\theta^{\prime}}).

Then R>rR>r, and we define the strip

Strθ,θ(𝒛):={𝒂Sn1:r𝒂,𝒛R}.\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}):=\left\{\boldsymbol{a}\in S^{n-1}:r\leq\left<\boldsymbol{a},\boldsymbol{z}\right>\leq R\right\}.
Refer to caption
Figure 3. Spherical strip Strθ,θ(𝒛)\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}).
Lemma 2.1.

The strip Strθ,θ(𝐳)\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}) is contained in Capθ,θ(𝐳)\textup{Cap}_{\theta,\theta^{\prime}}(\boldsymbol{z}). Furthermore, any two points in Strθ,θ(𝐳){±𝐳}\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z})\setminus\{\pm\boldsymbol{z}\} of radial angle at least θ\theta apart are mapped to points in Sn2S^{n-2} (unit sphere in N𝐳N_{\boldsymbol{z}}) of radial angle at least θ\theta^{\prime}.

Proof.

The first statement follows from 𝒂,𝒛r\left<\boldsymbol{a},\boldsymbol{z}\right>\geq r for any point 𝒂Strθ,θ(𝒛)\boldsymbol{a}\in\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}). The second statement follows from equation (147) and Lemma 42 of [AVZ00]. ∎

With this in mind, we are now ready to prove Proposition 2.2. When discussing the measure of strips Strθ,θ(𝒛)\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}), we drop 𝒛\boldsymbol{z} from the notation and simply write Strθ,θ\textup{Str}_{\theta,\theta^{\prime}}.

Proposition 2.2.

Let 0<θ<θ<π.0<\theta<\theta^{\prime}<\pi. We have

(17) M(n,θ)M(n1,θ)λn(θ,θ)(1+O(nenc)),M(n,\theta)\leq\frac{M(n-1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})}(1+O(ne^{-nc})),

where c:=12log(1r21R2)>0c:=\frac{1}{2}\log\left(\frac{1-r^{2}}{1-R^{2}}\right)>0 is independent of nn and only depends on θ\theta and θ\theta^{\prime}.

Proof.

Suppose {x1,,xN}Sn1\{x_{1},\ldots,x_{N}\}\subset S^{n-1} is a maximal θ\theta-spherical code. Given xSn1x\in S^{n-1}, let m(x)m(x) be the number of such strips Strθ,θ(xi)\textup{Str}_{\theta,\theta^{\prime}}(x_{i}) such that xStrθ,θ(xi)x\in\textup{Str}_{\theta,\theta^{\prime}}(x_{i}). Note that xStrθ,θ(xi)x\in\textup{Str}_{\theta,\theta^{\prime}}(x_{i}) if and only if xiStrθ,θ(x)x_{i}\in\textup{Str}_{\theta,\theta^{\prime}}(x). Therefore, the strip Strθ,θ(x)\textup{Str}_{\theta,\theta^{\prime}}(x) contains m(x)m(x) points of {x1,,xN}\{x_{1},\ldots,x_{N}\}. From the previous lemma, we know that these m(x)m(x) points are mapped to points in Sn2S^{n-2} that have pairwise radial angles at least θ\theta^{\prime}. As a result,

m(x)M(n1,θ),m(x)\leq M(n-1,\theta^{\prime}),

using which we obtain

Nμ(Strθ,θ)=i=1NStrθ,θ(xi)𝑑μ(x)=Sn1m(x)𝑑μ(x)M(n1,θ)Sn1𝑑μ(x)=M(n1,θ),N\cdot\mu(\textup{Str}_{\theta,\theta^{\prime}})=\sum_{i=1}^{N}\int_{\textup{Str}_{\theta,\theta^{\prime}}(x_{i})}d\mu(x)=\int_{S^{n-1}}m(x)d\mu(x)\leq M(n-1,\theta^{\prime})\int_{S^{n-1}}d\mu(x)=M(n-1,\theta^{\prime}),

where μ\mu is the uniform probability measure on Sn1S^{n-1}. Hence,

(18) M(n,θ)M(n1,θ)μ(Strθ,θ).M(n,\theta)\leq\frac{M(n-1,\theta^{\prime})}{\mu(\textup{Str}_{\theta,\theta^{\prime}})}.

Note that the masses of Strθ,θ\textup{Str}_{\theta,\theta^{\prime}} and the cap Capθ,θ\textup{Cap}_{\theta,\theta^{\prime}} have the property that

1μ(Strθ,θ)λn(θ,θ)\displaystyle 1-\frac{\mu(\textup{Str}_{\theta,\theta^{\prime}})}{\lambda_{n}(\theta,\theta^{\prime})} =\displaystyle= 1ωnλn(θ,θ)R1(1t2)n32𝑑t\displaystyle\frac{1}{\omega_{n}\lambda_{n}(\theta,\theta^{\prime})}\int_{R}^{1}(1-t^{2})^{\frac{n-3}{2}}dt
\displaystyle\leq (1R2)n32ωnλn(θ,θ).\displaystyle\frac{(1-R^{2})^{\frac{n-3}{2}}}{\omega_{n}\lambda_{n}(\theta,\theta^{\prime})}.

Here, ωn=11(1t2)n32𝑑t\omega_{n}=\int_{-1}^{1}(1-t^{2})^{\frac{n-3}{2}}dt. On the other hand, we may also give a lower bound on λn(θ,θ)\lambda_{n}(\theta,\theta^{\prime}) by noting that

λn(θ,θ)\displaystyle\lambda_{n}(\theta,\theta^{\prime}) =\displaystyle= 1ωnr1(1t2)n32𝑑t\displaystyle\frac{1}{\omega_{n}}\int_{r}^{1}(1-t^{2})^{\frac{n-3}{2}}dt
\displaystyle\geq (1+r1r)n3ωnr1(1t)n3𝑑t\displaystyle\frac{\left(\sqrt{\frac{1+r}{1-r}}\right)^{n-3}}{\omega_{n}}\int_{r}^{1}(1-t)^{n-3}dt
=\displaystyle= (1+r1r)n3(1r)n2(n2)ωn.\displaystyle\frac{\left(\sqrt{\frac{1+r}{1-r}}\right)^{n-3}(1-r)^{n-2}}{(n-2)\omega_{n}}.

The first inequality follows from

1+t1t1+r1r\frac{1+t}{1-t}\geq\frac{1+r}{1-r}

for t[r,1)t\in[r,1). Combining this inequality for λn(θ,θ)\lambda_{n}(\theta,\theta^{\prime}) with the above, we obtain

1μ(Strθ,θ)λn(θ,θ)1(n2)(1R2)n32(1r)(1r2)n32=1(n2)(1r)en32log(1r21R2).1\geq\frac{\mu(\textup{Str}_{\theta,\theta^{\prime}})}{\lambda_{n}(\theta,\theta^{\prime})}\geq 1-\frac{(n-2)(1-R^{2})^{\frac{n-3}{2}}}{(1-r)(1-r^{2})^{\frac{n-3}{2}}}=1-\frac{(n-2)}{(1-r)}e^{-\frac{n-3}{2}\log\left(\frac{1-r^{2}}{1-R^{2}}\right)}.

The conclusion follows using (18). ∎

Remark 19.

Proposition 2.2 removes the angular condition on inequality (9) of Barg–Musin at the expense of an error term that is exponentially decaying in the dimension nn.

3. New test functions

In this section, we prove general linear programming bounds on the sizes of spherical codes and sphere packing densities by constructing new test functions.

3.1. Spherical codes

Recall the definition of D(μ,s)D(\mu,s) from the discussion of the Delsarte linear programming method in the introduction. In this subsection, we construct a function inside D(dμn32,cosθ)D(d\mu_{\frac{n-3}{2}},\cos\theta) from a given one inside D(dμn42,cosθ),D(d\mu_{\frac{n-4}{2}},\cos\theta^{\prime}), where θ>θ.\theta^{\prime}>\theta.

Suppose that gθD(dμn42,cosθ).g_{\theta^{\prime}}\in D(d\mu_{\frac{n-4}{2}},\cos\theta^{\prime}). Fix 𝒛Sn1\boldsymbol{z}\in S^{n-1}. Given 𝒂,𝒃Sn1\boldsymbol{a},\boldsymbol{b}\in S^{n-1}, we define

(20) h(𝒂,𝒃;𝒛):=F(𝒂,𝒛)F(𝒃,𝒛)gθ(𝒂~,𝒃~),h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}):=F(\left<\boldsymbol{a},\boldsymbol{z}\right>)F(\left<\boldsymbol{b},\boldsymbol{z}\right>)g_{\theta^{\prime}}\left(\left<\tilde{\boldsymbol{a}},\tilde{\boldsymbol{b}}\right>\right),

where FF is an arbitrary integrable real valued function on [1,1][-1,1], and 𝒂~\tilde{\boldsymbol{a}} and 𝒃~\tilde{\boldsymbol{b}} are unit vectors on the tangent space of the sphere at 𝒛\boldsymbol{z} as defined in the previous section. Using the notation u,v,tu,v,t of Section 2,

(21) h(𝒂,𝒃;𝒛):=F(u)F(v)gθ(tuv(1u2)(1v2)).h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}):=F(u)F(v)g_{\theta^{\prime}}\left(\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}\right).

Note that the projection is not defined at ±𝒛\pm\boldsymbol{z}. If either 𝒂\boldsymbol{a} or 𝒃\boldsymbol{b} is ±𝒛\pm\boldsymbol{z}, we let h(𝒂,𝒃;𝒛)=0h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z})=0.

Lemma 3.1.

h(𝒂,𝒃;𝒛)h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}) is a positive semi-definite function in the variables 𝐚,𝐛\boldsymbol{a},\boldsymbol{b} on Sn1,S^{n-1}, namely

xi,xjAaiajh(𝒂i,𝒂j;𝒛)0\sum_{x_{i},x_{j}\in A}a_{i}a_{j}h(\boldsymbol{a}_{i},\boldsymbol{a}_{j};\boldsymbol{z})\geq 0

for every finite subset ASn1A\subset S^{n-1}, and coefficients ai.a_{i}\in\mathbb{R}. Moreover, hh is invariant under the diagonal action of the orthogonal group O(n)O(n), namely

h(𝒂,𝒃;𝒛)=h(k𝒂,k𝒃;k𝒛)h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z})=h(k\boldsymbol{a},k\boldsymbol{b};k\boldsymbol{z})

for every kO(n)k\in O(n).

Proof.

By (20), we have

xi,xjAaiajh(𝒂i,𝒂j;𝒛)=xi,xjAaiF(𝒙𝒊,𝒛)ajF(𝒙𝒋,𝒛)gθ(𝒙𝒊~,𝒙𝒋~).\sum_{x_{i},x_{j}\in A}a_{i}a_{j}h(\boldsymbol{a}_{i},\boldsymbol{a}_{j};\boldsymbol{z})=\sum_{{x_{i},x_{j}\in A}}a_{i}F(\left<\boldsymbol{x_{i}},\boldsymbol{z}\right>)a_{j}F(\left<\boldsymbol{x_{j}},\boldsymbol{z}\right>)g_{\theta^{\prime}}(\left<\tilde{\boldsymbol{x_{i}}},\tilde{\boldsymbol{x_{j}}}\right>).

By (1) and the positivity of the Fourier coefficients of gθg_{\theta^{\prime}}, we have

xi,xjAaiF(𝒙𝒊,𝒛)ajF(𝒙𝒋,𝒛)gθ(𝒙𝒊~,𝒙𝒋~)0.\sum_{{x_{i},x_{j}\in A}}a_{i}F(\left<\boldsymbol{x_{i}},\boldsymbol{z}\right>)a_{j}F(\left<\boldsymbol{x_{j}},\boldsymbol{z}\right>)g_{\theta^{\prime}}(\left<\tilde{\boldsymbol{x_{i}}},\tilde{\boldsymbol{x_{j}}}\right>)\geq 0.

This proves the first part, and the second part follows from (20) and the O(n)O(n)-invariance of the inner product. ∎

Let

(22) h(𝒂,𝒃):=O(n)h(𝒂,𝒃;k𝒛)𝑑μ(k),h(\boldsymbol{a},\boldsymbol{b}):=\int_{O(n)}h(\boldsymbol{a},\boldsymbol{b};k\boldsymbol{z})d\mu(k),

where dμ(k)d\mu(k) is the probability Haar measure on O(n).O(n).

Lemma 3.2.

h(𝒂,𝒃)h(\boldsymbol{a},\boldsymbol{b}) is a positive semi-definite point pair invariant function on Sn1.S^{n-1}.

Proof.

This follows from the previous lemma. ∎

Since h(𝒂,𝒃)h(\boldsymbol{a},\boldsymbol{b}) is a point pair invariant function, it only depends on t=𝒂,𝒃.t=\langle\boldsymbol{a},\boldsymbol{b}\rangle. For the rest of this paper, we abuse notation and consider hh as a real valued function of tt on [1,1],[-1,1], and write h(𝒂,𝒃)=h(t).h(\boldsymbol{a},\boldsymbol{b})=h(t).

3.1.1. Computing (𝒽)\cal{L}(h)

See (2) for the definition of \cal{L}. We proceed to computing the value of (𝒽)\cal{L}(h) in terms of FF and gθ.g_{\theta^{\prime}}. First, we compute the value of h(1).h(1). Let F22:=11F(u)2𝑑μn32(u).\|F\|^{2}_{2}:=\int_{-1}^{1}F(u)^{2}d\mu_{\frac{n-3}{2}}(u).

Lemma 3.3.

We have

h(1)=gθ(1)F22.h(1)=g_{\theta^{\prime}}(1)\|F\|^{2}_{2}.
Proof.

Indeed, by definition, h(1)h(1) corresponds to taking 𝒂=𝒃\boldsymbol{a}=\boldsymbol{b}, from which it follows that t=1t=1, u=vu=v and tuv(1u2)(1v2)=1\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}=1. Therefore, we obtain

h(1)=11F(u)2gθ(1)𝑑μn32(u)=gθ(1)F22.h(1)=\int_{-1}^{1}F(u)^{2}g_{\theta^{\prime}}(1)d\mu_{\frac{n-3}{2}}(u)=g_{\theta^{\prime}}(1)\|F\|^{2}_{2}.

Next, we compute the zero Fourier coefficient of h,h, that is, h0:=11h(t)𝑑μn32(t)h_{0}:=\int_{-1}^{1}h(t)d\mu_{\frac{n-3}{2}}(t). Let F0=11F(u)𝑑μn32(u)F_{0}=\int_{-1}^{1}F(u)d\mu_{\frac{n-3}{2}}(u) and gθ,0=11gθ𝑑μn42(u)g_{\theta^{\prime},0}=\int_{-1}^{1}g_{\theta^{\prime}}d\mu_{\frac{n-4}{2}}(u).

Lemma 3.4.

We have

h0=gθ,0F02.h_{0}=g_{\theta^{\prime},0}F_{0}^{2}.
Proof.

Let O(n1)O(n)O(n-1)\subset O(n) be the stabilizer of 𝒛.\boldsymbol{z}. We identify O(n)/O(n1)O(n)/O(n-1) with Sn1S^{n-1} and write [k1]:=k1𝒛Sn1[k_{1}]:=k_{1}\boldsymbol{z}\in S^{n-1} for any k1O(n).k_{1}\in O(n). Then we write the probability Haar measure of O(n)O(n) as the product of the probability Haar measure of O(n1)O(n-1) and the uniform probability measure dσd\sigma of Sn1:S^{n-1}:

dμ(k1)=dμ(k1)dσ([k1]),d\mu(k_{1})=d\mu(k_{1}^{\prime})d\sigma([k_{1}]),

where k1O(n1).k_{1}^{\prime}\in O(n-1). By equations (20),  (22) and the above, we obtain

h0=kiO(n1)[ki]Sn1F(k1[k1],𝒛)F(k2[k2],𝒛)gθ(k1[k1]~,k2[k2]~)𝑑μ(k1)𝑑σ([k1])𝑑μ(k2)𝑑σ([k2])=[ki]Sn1F([k1],𝒛)F([k2],𝒛)𝑑σ([k1])𝑑σ([k2])kiO(n1)gθ(k1[k1]~,k2[k2]~)𝑑μ(k1)𝑑μ(k2).\begin{split}h_{0}&=\iint_{k_{i}^{\prime}\in O(n-1)}\iint_{[k_{i}]\in S^{n-1}}F(\left<k_{1}^{\prime}[k_{1}],\boldsymbol{z}\right>)F(\left<k_{2}^{\prime}[k_{2}],\boldsymbol{z}\right>)g_{\theta^{\prime}}\left(\left<\widetilde{k_{1}^{\prime}[k_{1}]},\widetilde{k_{2}^{\prime}[k_{2}]}\right>\right)d\mu(k_{1}^{\prime})d\sigma([k_{1}])d\mu(k_{2}^{\prime})d\sigma([k_{2}])\\ &=\iint_{[k_{i}]\in S^{n-1}}F(\left<[k_{1}],\boldsymbol{z}\right>)F(\left<[k_{2}],\boldsymbol{z}\right>)d\sigma([k_{1}])d\sigma([k_{2}])\iint_{k_{i}^{\prime}\in O(n-1)}g_{\theta^{\prime}}\left(\left<k_{1}^{\prime}\widetilde{[k_{1}]},k_{2}^{\prime}\widetilde{[k_{2}]}\right>\right)d\mu(k_{1}^{\prime})d\mu(k_{2}^{\prime}).\end{split}

We note that

kiO(n1)gθ(k1[k1]~,k2[k2]~)𝑑μ(k1)𝑑μ(k2)=gθ,0,\iint_{k_{i}^{\prime}\in O(n-1)}g_{\theta^{\prime}}\left(\left<k_{1}^{\prime}\widetilde{[k_{1}]},k_{2}^{\prime}\widetilde{[k_{2}]}\right>\right)d\mu(k_{1}^{\prime})d\mu(k_{2}^{\prime})=g_{\theta^{\prime},0},

and

[ki]Sn1F([k1],𝒛)F([k2],𝒛)𝑑σ([k1])𝑑σ([k2])=1111F(u)F(v)𝑑μn32(u)μn32(v)=F02.\iint_{[k_{i}]\in S^{n-1}}F(\left<[k_{1}],\boldsymbol{z}\right>)F(\left<[k_{2}],\boldsymbol{z}\right>)d\sigma([k_{1}])d\sigma([k_{2}])=\int_{-1}^{1}\int_{-1}^{1}F(u)F(v)d\mu_{\frac{n-3}{2}}(u)\mu_{\frac{n-3}{2}}(v)=F_{0}^{2}.

Therefore,

h0=gθ,0F02,h_{0}=g_{\theta^{\prime},0}F_{0}^{2},

as required. ∎

Proposition 3.5.

We have

(𝒽)=(θ)2202.\cal{L}(h)=\cal{L}(g_{\theta^{\prime}})\frac{\|F\|^{2}_{2}}{F_{0}^{2}}.
Proof.

This follows immediately from Lemmas 3.3 and 3.4. ∎

3.1.2. Criterion for hD(dμn32,cosθ)h\in D(d\mu_{\frac{n-3}{2}},\cos\theta)

Finally, we give a criterion which implies hD(dμn32,cosθ).h\in D(d\mu_{\frac{n-3}{2}},\cos\theta). Recall that 0<θ<θ,0<\theta<\theta^{\prime}, and 0<r<R<10<r<R<1 are as defined in Section 2. Let s=cos(θ)s^{\prime}=\cos(\theta^{\prime}) and s=cos(θ).s=\cos(\theta). Note that s<ss^{\prime}<s, r=ss1sr=\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}, and R=cosγθ,θR=\cos\gamma_{\theta,\theta^{\prime}} as in equation (15). We define

χ(y):={1 for ryR,0otherwise.\chi(y):=\begin{cases}1&\text{ for }r\leq y\leq R,\\ 0&\text{otherwise}.\end{cases}
Proposition 3.6.

Suppose that gθD(dμn42,cosθ)g_{\theta^{\prime}}\in D(d\mu_{\frac{n-4}{2}},\cos\theta^{\prime}) is given and hh is defined as in (22) for some FF. Suppose that F(x)F(x) is a positive integrable function giving rise to an hh such that h(t)0h(t)\leq 0 for every 1tcosθ-1\leq t\leq\cos\theta. Then

hD(dμn32,cosθ),h\in D(d\mu_{\frac{n-3}{2}},\cos\theta),

and

M(n,θ)(𝒽)=(θ)2202.M(n,\theta)\leq\cal{L}(h)=\cal{L}(g_{\theta^{\prime}})\frac{\|F\|^{2}_{2}}{F_{0}^{2}}.

Among all positive integrable functions FF with compact support inside [r,R][r,R], χ\chi minimize the value of (𝒽),\cal{L}(h), and for F=χF=\chi we have

M(n,θ)(𝒽)(θ)λ𝓃(θ,θ)(1+𝒪(𝓃𝓃𝒸)),M(n,\theta)\leq\cal{L}(h)\leq\frac{\cal{L}(g_{\theta^{\prime}})}{\lambda_{n}(\theta,\theta^{\prime})}(1+O(ne^{-nc})),

where c=12log(1r21R2)>0.c=\frac{1}{2}\log\left(\frac{1-r^{2}}{1-R^{2}}\right)>0.

Proof.

The first part follows from the previous lemmas and propositions. Let us specialize to the situation where FF is merely assumed to be a positive integrable function with compact support inside [r,R][r,R]. Let us first show that h(t)0h(t)\leq 0 for tst\leq s. We have

h(t):=O(n)h(𝒂,𝒃;k𝒛)𝑑μ(k),h(t):=\int_{O(n)}h(\boldsymbol{a},\boldsymbol{b};k\boldsymbol{z})d\mu(k),

where h(𝒂,𝒃;𝒛):=F(u)F(v)gθ(tuv(1u2)(1v2)).h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}):=F(u)F(v)g_{\theta^{\prime}}\left(\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}\right). First, note that F(u)F(v)0F(u)F(v)\neq 0 implies that 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} belong to Strθ,θ(𝒛).\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}). By Lemma 2.1, the radial angle between 𝒂~\tilde{\boldsymbol{a}} and 𝒃~\tilde{\boldsymbol{b}} is at least θ\theta^{\prime}, and so

tuv(1u2)(1v2)=𝒂~,𝒃~[1,cosθ].\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}=\left<\tilde{\boldsymbol{a}},\tilde{\boldsymbol{b}}\right>\in[-1,\cos\theta^{\prime}].

Therefore

gθ(tuv(1u2)(1v2))0g_{\theta^{\prime}}\left(\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}\right)\leq 0

when F(u)F(v)0.F(u)F(v)\neq 0. Hence, the integrand h(𝒂,𝒃,𝒛)h(\boldsymbol{a},\boldsymbol{b},\boldsymbol{z}) is non-positive when t[1,cosθ]t\in[-1,\cos\theta], and so h(t)0h(t)\leq 0 for tst\leq s.

It is easy to see that when F=χF=\chi,

F22=μ(Strθ,θ(𝒛))\|F\|_{2}^{2}=\mu(\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}))

and

F0=μ(Strθ,θ(𝒛)).F_{0}=\mu(\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z})).

Therefore, by our estimate in the proof of Proposition 2.2 we have

M(n,θ)(𝒽)(θ)λ𝓃(θ,θ)(1+𝒪(𝓃𝓃2log(1𝓇212))).M(n,\theta)\leq\cal{L}(h)\leq\frac{\cal{L}(g_{\theta^{\prime}})}{\lambda_{n}(\theta,\theta^{\prime})}\left(1+O\left(ne^{-\frac{n}{2}\log\left(\frac{1-r^{2}}{1-R^{2}}\right)}\right)\right).

Finally, the optimality follows from the Cauchy–Schwarz inequality. More precisely, since F(x)F(x) has compact support inside [r,R][r,R], we have

μ(Strθ,θ(𝒛))F22F02.\mu(\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}))\|F\|_{2}^{2}\geq F_{0}^{2}.

Therefore, (𝒽)=θ(1)θ,02202(θ)μ(Strθ,θ(𝔃))\cal{L}(h)=\frac{g_{\theta^{\prime}}(1)}{g_{\theta^{\prime},0}}\frac{\|F\|^{2}_{2}}{F_{0}^{2}}\geq\frac{\cal{L}(g_{\theta^{\prime}})}{\mu(\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}))} with equality only when F=χ.F=\chi.

3.2. Sphere packings

Suppose 0<θπ0<\theta\leq\pi is a given angle, and suppose gθD(dμn32,cosθ)g_{\theta}\in D(d\mu_{\frac{n-3}{2}},\cos\theta). Fixing 𝒛n\boldsymbol{z}\in\mathbb{R}^{n}, for each pair of points 𝒂,𝒃n\{𝒛}\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{n}\backslash\{\boldsymbol{z}\} consider

H(𝒂,𝒃;𝒛):=F(|𝒂𝒛|)F(|𝒃𝒛|)gθ(𝒂𝒛|𝒂𝒛|,𝒃𝒛|𝒃𝒛|),H(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}):=F(|\boldsymbol{a}-\boldsymbol{z}|)F(|\boldsymbol{b}-\boldsymbol{z}|)g_{\theta}\left(\left<\frac{\boldsymbol{a}-\boldsymbol{z}}{|\boldsymbol{a}-\boldsymbol{z}|},\frac{\boldsymbol{b}-\boldsymbol{z}}{|\boldsymbol{b}-\boldsymbol{z}|}\right>\right),

where FF is an even positive function on \mathbb{R} such that it is in L1(n)L2(n)L^{1}(\mathbb{R}^{n})\cap L^{2}(\mathbb{R}^{n}). We may then define H(𝒂,𝒃)H(\boldsymbol{a},\boldsymbol{b}) by averaging over all 𝒛n\boldsymbol{z}\in\mathbb{R}^{n}:

(23) H(𝒂,𝒃):=nH(𝒂,𝒃;𝒛)𝑑𝒛.H(\boldsymbol{a},\boldsymbol{b}):=\int_{\mathbb{R}^{n}}H(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z})d\boldsymbol{z}.
Lemma 3.7.

H(𝒂,𝒃)H(\boldsymbol{a},\boldsymbol{b}) is a positive semi-definite kernel on n\mathbb{R}^{n} and depends only on T=|𝐚𝐛|T=|\boldsymbol{a}-\boldsymbol{b}|.

Proof.

The proof is similar to that of Lemma 3.1. ∎

As before, we abuse notation and write H(T)H(T) instead of H(𝒂,𝒃)H(\boldsymbol{a},\boldsymbol{b}) when T=|𝒂𝒃|T=|\boldsymbol{a}-\boldsymbol{b}|. The analogue of Proposition 3.6 is then the following.

Proposition 3.8.

Let 0<θπ0<\theta\leq\pi and suppose gθD(dμn32,cosθ)g_{\theta}\in D(d\mu_{\frac{n-3}{2}},\cos\theta). Suppose FF is as above such that H(T)0H(T)\leq 0 for every T1T\geq 1. Then

(24) δnvol(B1n)FL2(n)22nFL1(n)2(θ),\delta_{n}\leq\frac{\textup{vol}(B_{1}^{n})\|F\|_{L^{2}(\mathbb{R}^{n})}^{2}}{2^{n}\|F\|_{L^{1}(\mathbb{R}^{n})}^{2}}\cal{L}(g_{\theta}),

where vol(B1n)\textup{vol}(B_{1}^{n}) is the volume of the nn-dimensional unit ball. In particular, if F=χ[0,r]F=\chi_{[0,r]}, where 0r10\leq r\leq 1, is such that it gives rise to an HH satisfying H(T)0H(T)\leq 0 for every T1T\geq 1, then

(25) δn(θ)(2r)n.\delta_{n}\leq\frac{\cal{L}(g_{\theta})}{(2r)^{n}}.
Proof.

The proof of this proposition is similar to that of Theorem 3.4 of Cohn–Zhao [CZ14]. We focus our attention on proving inequality (24). Suppose we have a packing of n\mathbb{R}^{n} of density Δ\Delta by non-overlapping balls of radius 12\frac{1}{2}. By Theorem 3.1 of Cohn–Elkies [CE03], we have

Δvol(B1n)H(0)2nH^(0).\Delta\leq\frac{\textup{vol}(B_{1}^{n})H(0)}{2^{n}\widehat{H}(0)}.

Note that H(0)=gθ(1)FL2(n)2H(0)=g_{\theta}(1)\|F\|_{L^{2}(\mathbb{R}^{n})}^{2}, and that

H^(0)=nH(|𝒛|)𝑑𝒛=gθ,0FL1(n)2.\widehat{H}(0)=\int_{\mathbb{R}^{n}}H(|\boldsymbol{z}|)d\boldsymbol{z}=g_{\theta,0}\|F\|_{L^{1}(\mathbb{R}^{n})}^{2}.

As a result, we obtain inequality (24). The rest follows from a simple computation. ∎

Note that the situation r=12sin(θ/2)r=\frac{1}{2\sin(\theta/2)} with π/3θπ\pi/3\leq\theta\leq\pi corresponds to Theorem 3.4 of Cohn–Zhao [CZ14], as checking the negativity condition H(T)0H(T)\leq 0 for T1T\geq 1 follows from Lemma 2.2 therein. The factor 2n2^{n} comes from considering functions where the negativity condition is for T1T\geq 1 instead of T2T\geq 2.

Remark 26.

In this paper, we consider characteristic functions; however, it is an interesting open question to determine the optimal such FF in order to obtain the best bounds on sphere packing densities through this method.

3.3. Incorporating geometric improvement into linear programming

In proving upper bounds on M(n,θ)M(n,\theta), Levenshtein [Lev79, Lev98], building on Kabatyanskii–Levenshtein [KL78], constructed feasible test functions gθD(dμn32,cosθ)g_{\theta}\in D(d\mu_{\frac{n-3}{2}},\cos\theta) for the Delsarte linear programming problem with (θ)=Lev(𝓃,θ)\cal{L}(g_{\theta})=M_{\textup{Lev}}(n,\theta) defined in (7). This gave the bound

(27) M(n,θ)MLev(n,θ).M(n,\theta)\leq M_{\textup{Lev}}(n,\theta).

Sidelnikov’s geometric inequality (5) gives

M(n,θ)M(n+1,θ)λn(θ,θ)M(n,\theta)\leq\frac{M(n+1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})}

for angles 0<θ<θ<π20<\theta<\theta^{\prime}<\frac{\pi}{2}. Applying this for 0<θ<θ0<\theta<\theta^{*} and combining with inequality (27), Kabatyanskii and Levenshtein obtained

(28) M(n,θ)MLev(n+1,θ)λn(θ,θ).M(n,\theta)\leq\frac{M_{\textup{Lev}}(n+1,\theta^{*})}{\lambda_{n}(\theta,\theta^{*})}.

From [KL78], it is known that this is exponentially better than inequality (27) for 0<θ<θ0<\theta<\theta^{*}. Finding functions hD(dμn32,cosθ)h\in D(d\mu_{\frac{n-3}{2}},\cos\theta) with (𝒽)<(θ)\cal{L}(h)<\cal{L}(g_{\theta}) was suggested by Levenshtein in [Lev98, page 117]. In fact, Boyvalenkov–Danev–Bumova [BDB96] gives necessary and sufficient conditions for constructing extremal polynomials that improve Levenshtein’s bound. However, their construction does not exponentially improve inequality (27) for 0<θ<θ0<\theta<\theta^{*}. In contrast to their construction, Proposition 3.6 gives the following corollary stating that our construction of the function hh gives an exponential improvement in the linear programming problem comparing to Levenshtein’s optimal polynomials for 0<θ<θ0<\theta<\theta^{*}. This is not to say that we exponentially improve the bounds given for spherical codes and sphere packings; we provide a single function using which the Delsarte linear programming method gives a better version of the exponentially better inequality (28).

Corollary 3.9.

Fix 0<θ<θ.0<\theta<\theta^{*}. Let hθD(dμn32,cosθ)h_{\theta^{*}}\in D(d\mu_{\frac{n-3}{2}},\cos\theta) be the function associated to Levenshtein’s gn1,θg_{n-1,\theta^{*}} constructed in Proposition 3.6. Then

M(n,θ)(𝒽θ)𝓃(δθ+(1))Lev(𝓃,θ),M(n,\theta)\leq\cal{L}(h_{\theta^{*}})\leq e^{-n(\delta_{\theta}+o(1))}M_{\textup{Lev}}(n,\theta),

where o(1)0o(1)\rightarrow 0 as nn\rightarrow\infty and δθ:=Δ(θ)Δ(θ)>0\delta_{\theta}:=\Delta(\theta)-\Delta(\theta^{*})>0, with

Δ(θ):=1+sinθ2sinθlog1+sinθ2sinθ1sinθ2sinθlog1sinθ2sinθ+12log(1cosθ).\Delta(\theta):=\frac{1+\sin\theta}{2\sin\theta}\log\frac{1+\sin\theta}{2\sin\theta}-\frac{1-\sin\theta}{2\sin\theta}\log\frac{1-\sin\theta}{2\sin\theta}+\frac{1}{2}\log(1-\cos\theta).
Proof.

Let 0<θ<θπ/20<\theta<\theta^{\prime}\leq\pi/2. Recall the notation of Proposition 3.6. Associated to a function gθg_{\theta^{\prime}} satisfying the Delsarte linear programming conditions in dimension n1n-1 and for angle θ\theta^{\prime}, in Proposition 3.6 we constructed an explicit hθD(dμn32,cosθ)h_{\theta^{\prime}}\in D(d\mu_{\frac{n-3}{2}},\cos\theta) such that

(𝒽θ)(θ)λ𝓃(θ,θ)(1+𝒪(𝓃𝓃𝒸)),\cal{L}(h_{\theta^{\prime}})\leq\frac{\cal{L}(g_{\theta^{\prime}})}{\lambda_{n}(\theta,\theta^{\prime})}(1+O(ne^{-nc})),

where c>0c>0 is a specific constant depending only on θ\theta and θ\theta^{\prime}. Therefore,

1nlog(𝒽θ)1𝓃log(θ)logλ𝓃(θ,θ)𝓃+𝒪(𝓃𝒸).\frac{1}{n}\log\cal{L}(h_{\theta^{\prime}})\leq\frac{1}{n}\log\cal{L}(g_{\theta^{\prime}})-\frac{\log\lambda_{n}(\theta,\theta^{\prime})}{n}+O(e^{-nc}).

By [KL78, Theorem 4]

limn1nlog(θ)=1+sinθ2sinθlog1+sinθ2sinθ1sinθ2sinθlog1sinθ2sinθ.\lim_{n\to\infty}\frac{1}{n}\log\cal{L}(g_{\theta^{\prime}})=\frac{1+\sin\theta^{\prime}}{2\sin\theta^{\prime}}\log\frac{1+\sin\theta^{\prime}}{2\sin\theta^{\prime}}-\frac{1-\sin\theta^{\prime}}{2\sin\theta^{\prime}}\log\frac{1-\sin\theta^{\prime}}{2\sin\theta^{\prime}}.

It is easy to show that [KL78, Proof of Theorem 4]

limnlogλn(θ,θ)n=12log1cosθ1cosθ.\lim_{n\to\infty}\frac{\log\lambda_{n}(\theta,\theta^{\prime})}{n}=\frac{1}{2}\log\frac{1-\cos\theta}{1-\cos\theta^{\prime}}.

Hence,

1nlog(𝒽θ)Δ(θ)12log(1cosθ)+(1),\frac{1}{n}\log\cal{L}(h_{\theta^{\prime}})\leq\Delta(\theta^{\prime})-\frac{1}{2}\log(1-\cos\theta)+o(1),

where

Δ(θ):=1+sinθ2sinθlog1+sinθ2sinθ1sinθ2sinθlog1sinθ2sinθ+12log(1cosθ).\Delta(\theta^{\prime}):=\frac{1+\sin\theta^{\prime}}{2\sin\theta^{\prime}}\log\frac{1+\sin\theta^{\prime}}{2\sin\theta^{\prime}}-\frac{1-\sin\theta^{\prime}}{2\sin\theta^{\prime}}\log\frac{1-\sin\theta^{\prime}}{2\sin\theta^{\prime}}+\frac{1}{2}\log(1-\cos\theta^{\prime}).

Note that

ddθΔ(θ)=csc2(θ)2(cosθlog(1+sinθ1sinθ)(1+cosθ)sinθ).\frac{d}{d\theta^{\prime}}\Delta(\theta^{\prime})=-\frac{\csc^{2}(\theta^{\prime})}{2}\left(\cos\theta^{\prime}\log(\frac{1+\sin\theta^{\prime}}{1-\sin\theta^{\prime}})-(1+\cos\theta^{\prime})\sin\theta^{\prime}\right).

As we mentioned before, θ:=62.997\theta^{*}:=62.997...^{\circ} is the unique root of the equation ddθΔ(θ)=0\frac{d}{d\theta^{\prime}}\Delta(\theta^{\prime})=0 [KL78, Theorem 4] in the interval 0<θ<π/20<\theta^{\prime}<\pi/2, which is the unique minimum of Δ(θ)\Delta(\theta^{\prime}) for 0<θ<π/2.0<\theta^{\prime}<\pi/2. Hence, for 0<θ<θ,0<\theta<\theta^{*}, taking hθh_{\theta^{*}} the function constructed in Proposition 3.6 associated to Levenshtein’s optimal polynomial gn1,θg_{n-1,\theta^{*}} for angle θ\theta^{*} in dimension n1n-1, and gn,θg_{n,\theta} Levenshtein’s optimal polynomial for angle θ\theta in dimension nn, we obtain

1n(log(𝒽θ)log(𝓃,θ))Δ(θ)Δ(θ)+o(1)<0\frac{1}{n}\left(\log\cal{L}(h_{\theta^{*}})-\log\cal{L}(g_{n,\theta})\right)\leq\Delta(\theta^{*})-\Delta(\theta)+o(1)<0

for sufficiently large nn. This concludes the proof of this proposition. ∎

3.4. Constant improvement to Barg–Musin and Cohn–Zhao

In this subsection, we sketch the ideas that go into proving Theorems 1.1 that improves inequality (9) of Barg–Musin by a constant factor of at least 0.43250.4325 for every angle θ\theta with 0<θ<θ0<\theta<\theta^{*}. The improvement to the Cohn–Zhao inequality (10) for sphere packings is similar. We will complete the technical details in the rest of the paper.

Refer to caption
Figure 4. rr and RR defined in (14) and (16).

Recall that given any test function gθD(dμn42,cosθ)g_{\theta^{\prime}}\in D(d\mu_{\frac{n-4}{2}},\cos\theta^{\prime}) and FF an arbitrary integrable real valued function on [1,1][-1,1], we defined

h(𝒂,𝒃;𝒛):=F(u)F(v)gθ(tuv(1u2)(1v2))h(\boldsymbol{a},\boldsymbol{b};\boldsymbol{z}):=F(u)F(v)g_{\theta^{\prime}}\left(\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}\right)

in equation (21), using which we defined the function

(29) h(𝒂,𝒃):=O(n)h(𝒂,𝒃;k𝒛)𝑑μ(k).h(\boldsymbol{a},\boldsymbol{b}):=\int_{O(n)}h(\boldsymbol{a},\boldsymbol{b};k\boldsymbol{z})d\mu(k).

h(𝒂,𝒃)h(\boldsymbol{a},\boldsymbol{b}) is a point-pair invariant function, and so we viewed it as a function h(t)h(t) of t:=𝒂,𝒃t:=\left<\boldsymbol{a},\boldsymbol{b}\right>. As we saw in the proof of Proposition 3.6, for FF positive integrable and compactly supported on [r,R][r,R], F(u)F(v)0F(u)F(v)\neq 0 implies that 𝒂,𝒃Strθ,θ(𝒛)\boldsymbol{a},\boldsymbol{b}\in\text{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}), where

Strθ,θ(𝒛):={𝒂Sn1:r𝒂,𝒛R}\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}):=\{\boldsymbol{a}\in S^{n-1}:r\leq\left<\boldsymbol{a},\boldsymbol{z}\right>\leq R\}

is the grey strip illustrated above in Figure 4. From this, we obtained that

gθ(tuv(1u2)(1v2))0,g_{\theta^{\prime}}\left(\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}\right)\leq 0,

using which we obtained that the integrand in (29) is non-positive, giving us

(30) h(t)0 for t[1,cosθ].h(t)\leq 0\textup{ for }t\in[-1,\cos\theta].

Our insight is that the integrand in (29) need not be non-positive everywhere in order to have (30). In fact, we will show that when FF is the characteristic function with support [rδ,R][r-\delta,R] for some δ>cn\delta>\frac{c}{n}, where c>0c>0 is independent of nn, and gθg_{\theta^{\prime}} are Levenshtein’s optimal polynomials, we continue to have (30). This corresponds to allowing 𝒂,𝒃\boldsymbol{a},\boldsymbol{b} be contained in a slight enlargement of the strip Strθ,θ(𝒛)\textup{Str}_{\theta,\theta^{\prime}}(\boldsymbol{z}) including also the yellow region in Figure 4.

There are two main ingredients that go into determining an explicit lower bound for δ\delta. The first is related to understanding the behavior of Levenshtein’s optimal polynomials near their largest roots. This reduces to understanding the behavior of Jacobi polynomials near their largest roots. This is done in Subsection 5.1. The other idea is estimating the density function of the inner product matrix of triple uniformly distributed points on high-dimensional spheres. This allows us to rewrite the integral (29) and its sphere packing analogue using different coordinates. This is done in Section 4.

4. Conditional density functions

In this section, we rewrite the averaging integral (29) when FF is a characteristic function in different coordinates. We also do the same for sphere packings. See equations (31) and (33). In order to prove Theorems 1.1 and 1.2 in Section 5, we will also need to estimate certain conditional densities, which is the main purpose of this section.

4.1. Conditional density for spherical codes

Let 𝒮\mathcal{S} be the space of positive semi-definite symmetric 3×33\times 3 matrices with 11 on diagonal. Let

π:(Sn1)3𝒮\pi:(S^{n-1})^{3}\rightarrow\mathcal{S}

be the map that sends triple points to their pairwise inner products via

(𝒂,𝒃,𝒛)[1tut1vuv1],(\boldsymbol{a},\boldsymbol{b},\boldsymbol{z})\mapsto\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix},

where (t,u,v):=(𝒂,𝒃,𝒂,𝒛,𝒃,𝒛)(t,u,v):=(\left<\boldsymbol{a},\boldsymbol{b}\right>,\left<\boldsymbol{a},\boldsymbol{z}\right>,\left<\boldsymbol{b},\boldsymbol{z}\right>). Let μ(u,v,t)dudvdt\mu(u,v,t)dudvdt be the density function of the pushforward of the product of uniform probability measures on (Sn1)3(S^{n-1})^{3} to the coordinates (u,v,t)(u,v,t).

Proposition 4.1.

We have

μ(u,v,t):=Cdet[1tut1vuv1]n42.\mu(u,v,t):=C\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{\frac{n-4}{2}}.

where CC is a normalization constant such that 𝒮μ(u,v,t)𝑑u𝑑v𝑑t=1\int_{\mathcal{S}}\mu(u,v,t)dudvdt=1

Proof.

The density of the conditional measure is given by

μ(u,v,t)=μ(u,v;t)μ(t)\mu(u,v,t)=\mu(u,v;t)\mu(t)

where μ(t)=(1t2)n32\mu(t)=(1-t^{2})^{\frac{n-3}{2}} and μ(u,v;t)\mu(u,v;t) is the conditional density of u,vu,v given tt. We note that this conditional density function is proportional to

limε0vol(Str(u,v,ε;𝒂,𝒃))ε2.\lim_{\varepsilon\to 0}\frac{\text{vol}(Str(u,v,\varepsilon;\boldsymbol{a},\boldsymbol{b}))}{\varepsilon^{2}}.

where

Str(u,v,ε;𝒂,𝒃):={𝒛Sn1:0𝒂,𝒛u,𝒃,𝒛vε}.Str(u,v,\varepsilon;\boldsymbol{a},\boldsymbol{b}):=\{\boldsymbol{z}\in S^{n-1}:0\leq\left<\boldsymbol{a},\boldsymbol{z}\right>-u,\left<\boldsymbol{b},\boldsymbol{z}\right>-v\leq\varepsilon\}.

We write 𝒛=𝒛+𝒛\boldsymbol{z}=\boldsymbol{z}^{\perp}+\boldsymbol{z}^{\|}, where 𝒛\boldsymbol{z}^{\|} is the projection of 𝒛\boldsymbol{z} onto the two dimensional plane spanned by 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} and 𝒛\boldsymbol{z}^{\perp} is orthogonal to 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b}. Fix 𝒂,𝒃\boldsymbol{a},\boldsymbol{b} and consider the following map from Sn1S^{n-1}

𝒛[𝒛,𝒛].\boldsymbol{z}\to[\boldsymbol{z}^{\|},\boldsymbol{z}^{\perp}].

The above map has Jacobian det[1tut1vuv1]1/21t2\frac{\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{1/2}}{\sqrt{1-t^{2}}} with respect to the Euclidean metric. We note that the geometric locus of 𝒛\boldsymbol{z}^{\|} is a rhombus with area ε21t2\frac{\varepsilon^{2}}{\sqrt{1-t^{2}}}. Moreover, given 𝒛||\boldsymbol{z}^{||}, the geometric locus of 𝒛\boldsymbol{z}^{\perp} is a sphere of dimension n3n-3 and radius

|𝒛|=det[1tut1vuv1]1/21t2+O(ε).|\boldsymbol{z}^{\perp}|=\frac{\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{1/2}}{\sqrt{1-t^{2}}}+O(\varepsilon).

Therefore,

vol(Str(u,v,ε;x,y))=ε21t21t2det[1tut1vuv1]1/2(det[1tut1vuv1]1/21t2)n3(1+O(ε)),\text{vol}(Str(u,v,\varepsilon;x,y))=\frac{\varepsilon^{2}}{\sqrt{1-t^{2}}}\frac{\sqrt{1-t^{2}}}{\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{1/2}}\left(\frac{\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{1/2}}{\sqrt{1-t^{2}}}\right)^{n-3}(1+O(\varepsilon)),

from which it follows that

μ(u,v;t)=Cdet[1tut1vuv1]n42(1t2)n32\mu(u,v;t)=C\frac{\det\begin{bmatrix}1&t&u\\ t&1&v\\ u&v&1\end{bmatrix}^{\frac{n-4}{2}}}{(1-t^{2})^{\frac{n-3}{2}}}

for some constant C>0C>0. This implies our proposition.

Recall that θ<θ\theta<\theta^{\prime}. Let s=cos(θ)s^{\prime}=\cos(\theta^{\prime}) and s=cos(θ).s=\cos(\theta). Note that s<ss^{\prime}<s and for r=ss1sr=\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}, we have s=sr21r2s^{\prime}=\frac{s-r^{2}}{1-r^{2}} and 0<r<10<r<1. Let 0<δ=o(1n)0<\delta=o(\frac{1}{\sqrt{n}}) that we specify later, and define

χ(y):={1 for rδyR,0otherwise.\chi(y):=\begin{cases}1&\text{ for }r-\delta\leq y\leq R,\\ 0&\text{otherwise}.\end{cases}

See Figure 4. 𝒂,𝒃\boldsymbol{a},\boldsymbol{b} are in the shaded areas (both grey and yellow) correspond to u,vu,v being in the support of χ\chi. Recall (13), and denote

x:=tuv(1u2)(1v2).x:=\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}.

Let μ(x;t,χ)\mu(x;t,\chi) be the induced density function on xx subjected to the conditions of fixed t,t, and rδu,vRr-\delta\leq u,v\leq R. Precisely, up to a positive constant multiple that depends on n,t,R,rδn,t,R,r-\delta, we have

μ(x;t,χ)=Cx,tχ(u)χ(v)μ(x,l,t)𝑑l,\mu(x;t,\chi)=\int_{C_{x,t}}\chi(u)\chi(v)\mu^{*}(x,l,t)dl,

where the integral is over the curve Cx,t2C_{x,t}\subset\mathbb{R}^{2} that is given by tuv(1u2)(1v2)=x\frac{t-uv}{\sqrt{(1-u^{2})(1-v^{2})}}=x, dldl is the induced Euclidean metric on Cx,t,C_{x,t}, and

μ(x,l,t)dxdldt=μ(u,v,t)dudvdt.\mu^{*}(x,l,t)dxdldt=\mu(u,v,t)dudvdt.
Refer to caption
Figure 5. 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} are two points with fixed t=𝒂,𝒃t=\left<\boldsymbol{a},\boldsymbol{b}\right>. The red curve is the 22-dimensional representation of the locus of points 𝒛\boldsymbol{z} having (u,v)(u,v) on the curve Cx,tC_{x,t}. Shaded region corresponds to the support of χ(u)χ(v)\chi(u)\chi(v).

We explicitly compute μ(x,l,t).\mu^{*}(x,l,t). We have

dudvdt=1(xu)2+(xv)2dxdldt,dudvdt=\frac{1}{\sqrt{(\frac{\partial x}{\partial u})^{2}+(\frac{\partial x}{\partial v})^{2}}}dxdldt,

from which it follows that

μ(x,l,t)=μ(u,v,t)(xu)2+(xv)2.\mu^{*}(x,l,t)=\frac{\mu(u,v,t)}{\sqrt{(\frac{\partial x}{\partial u})^{2}+(\frac{\partial x}{\partial v})^{2}}}.

Hence, up to a positive constant multiple that depends on n,t,R,rδn,t,R,r-\delta,

μ(x;t,χ)=Cx,tχ(u)χ(v)μ(u,v,t)(xu)2+(xv)2𝑑l.\mu(x;t,\chi)=\int_{C_{x,t}}\frac{\chi(u)\chi(v)\mu(u,v,t)}{\sqrt{(\frac{\partial x}{\partial u})^{2}+(\frac{\partial x}{\partial v})^{2}}}dl.

We may write the test function constructed in equation (22) with F=χF=\chi and any given gg as

(31) h(t)=11g(x)μ(x;t,χ)𝑑x,h(t)=\int_{-1}^{1}g(x)\mu(x;t,\chi)dx,

where t=𝒂,𝒃t=\left<\boldsymbol{a},\boldsymbol{b}\right>; see Subsection 3.4. We define for complex xx

x+:={x for x0,0otherwise.x^{+}:=\begin{cases}x&\text{ for }x\geq 0,\\ 0&\text{otherwise.}\end{cases}
Proposition 4.2.

Suppose that |xs|=o(1n).|x-s^{\prime}|=o(\frac{1}{\sqrt{n}}). We have, up to a positive constant multiple depending on n,s,R,rδn,s,R,r-\delta,

μ(x;s,χ)=(2(1r2)2r(1s)+o(1))(δ+sx1xr)+((1x2x2)(sr2)2)n42e(2nr(sx1xr)sr2).\mu(x;s,\chi)=\left(\frac{2(1-r^{2})^{2}}{r(1-s)}+o(1)\right)\left(\delta+\sqrt{\frac{s-x}{1-x}}-r\right)^{+}\left(\left(\frac{1-x^{2}}{x^{2}}\right)\left(s-r^{2}\right)^{2}\right)^{\frac{n-4}{2}}e^{\left(-\frac{2nr\left(\sqrt{\frac{s-x}{1-x}}-r\right)}{s-r^{2}}\right)}.
Proof.

Let Cx:=Cx,sC_{x}:=C_{x,s}. Note that

x=suv(1u2)(1v2)x=\frac{s-uv}{\sqrt{(1-u^{2})(1-v^{2})}}

and

μ(u,v,s)=((1x2)(1u2)(1v2))n42.\mu(u,v,s)=\left((1-x^{2})(1-u^{2})(1-v^{2})\right)^{\frac{n-4}{2}}.

Furthermore,

(xu)2+(xv)2=(usv)2(1v2)(1u2)3+(vsu)2(1u2)(1v2)3=2r2(1s)2(1r2)4+o(1).\left(\frac{\partial x}{\partial u}\right)^{2}+\left(\frac{\partial x}{\partial v}\right)^{2}=\frac{(us-v)^{2}}{(1-v^{2})(1-u^{2})^{3}}+\frac{(vs-u)^{2}}{(1-u^{2})(1-v^{2})^{3}}=2\frac{r^{2}(1-s)^{2}}{(1-r^{2})^{4}}+o(1).

Hence, up to a positive constant multiple depending on n,s,R,rδn,s,R,r-\delta,

μ(x;s,χ)\displaystyle\mu(x;s,\chi) =\displaystyle= ((1r2)22r(1s)+o(1))Cxχ(u)χ(v)μ(u,v,s)𝑑l\displaystyle\left(\frac{(1-r^{2})^{2}}{\sqrt{2}r(1-s)}+o(1)\right)\int_{C_{x}}\chi(u)\chi(v)\mu(u,v,s)dl
=\displaystyle= ((1r2)22r(1s)+o(1))(1x2)n42Cxχ(u)χ(v)((1u2)(1v2))n42𝑑l\displaystyle\left(\frac{(1-r^{2})^{2}}{\sqrt{2}r(1-s)}+o(1)\right)(1-x^{2})^{\frac{n-4}{2}}\int_{C_{x}}\chi(u)\chi(v)\left((1-u^{2})(1-v^{2})\right)^{\frac{n-4}{2}}dl
=\displaystyle= ((1r2)22r(1s)+o(1))(1x2)n42Cxχ(u)χ(v)(suvx)n4𝑑l\displaystyle\left(\frac{(1-r^{2})^{2}}{\sqrt{2}r(1-s)}+o(1)\right)(1-x^{2})^{\frac{n-4}{2}}\int_{C_{x}}\chi(u)\chi(v)\left(\frac{s-uv}{x}\right)^{{n-4}}dl
=\displaystyle= ((1r2)22r(1s)+o(1))(1x2x2)n42Cxχ(u)χ(v)(suv)n4𝑑l\displaystyle\left(\frac{(1-r^{2})^{2}}{\sqrt{2}r(1-s)}+o(1)\right)\left(\frac{1-x^{2}}{x^{2}}\right)^{\frac{n-4}{2}}\int_{C_{x}}\chi(u)\chi(v)\left(s-uv\right)^{{n-4}}dl

Suppose that ru,vRr\leq u,v\leq R. Then, by the definition of RR, the equality

s=suv(1u2)(1v2)s^{\prime}=\frac{s-uv}{\sqrt{(1-u^{2})(1-v^{2})}}

occurs when (u,v){(r,R),(R,r),(r,r)}(u,v)\in\{(r,R),(R,r),(r,r)\}. Furthermore, when |xs|=o(1n)|x-s^{\prime}|=o\left(\frac{1}{\sqrt{n}}\right), we must have (u,v){(r,R),(R,r),(r,r)}(u,v)\in\{(r,R),(R,r),(r,r)\} up to o(1n)o\left(\frac{1}{\sqrt{n}}\right). Since the integrand χ(u)χ(v)(suv)n4\chi(u)\chi(v)(s-uv)^{n-4} of the last integral above is exponentially larger when u,v=r+o(1n)u,v=r+o\left(\frac{1}{\sqrt{n}}\right), the main contribution of the integral comes when uu and vv are near rr up to o(1n)o\left(\frac{1}{\sqrt{n}}\right). Writing u=r+u~u=r+\tilde{u} and v=r+v~,v=r+\tilde{v}, where u~,v~=o(1n),\tilde{u},\tilde{v}=o(\frac{1}{\sqrt{n}}), we have

suv=sr2r(u~+v~)u~v~=(sr2)(1r(u~+v~)+u~v~sr2).s-uv=s-r^{2}-r(\tilde{u}+\tilde{v})-\tilde{u}\tilde{v}=(s-r^{2})\left(1-\frac{r(\tilde{u}+\tilde{v})+\tilde{u}\tilde{v}}{s-r^{2}}\right).

Hence, up to a positive constant multiple depending on n,s,R,rδn,s,R,r-\delta,

μ(x;s,χ)=((1r2)22r(1s)+o(1))((1x2x2)(sr2)2)n42Cxχ(u)χ(v)(1r(u~+v~)+u~v~sr2)n4𝑑l.\mu(x;s,\chi)=\left(\frac{(1-r^{2})^{2}}{\sqrt{2}r(1-s)}+o(1)\right)\left(\left(\frac{1-x^{2}}{x^{2}}\right)\left(s-r^{2}\right)^{2}\right)^{\frac{n-4}{2}}\int_{C_{x}}\chi(u)\chi(v)\left(1-\frac{r(\tilde{u}+\tilde{v})+\tilde{u}\tilde{v}}{s-r^{2}}\right)^{{n-4}}dl.

Recall the following inequalities, which follow easily from the Taylor expansion of log(1+x)\log(1+x)

eaa2n(1+an)nea\begin{split}e^{a-\frac{a^{2}}{n}}\leq(1+\frac{a}{n})^{n}\leq e^{a}\end{split}

for |a|n/2.|a|\leq n/2. We apply the above inequalities to estimate the integral, and obtain

Cxχ(u)χ(v)(1r(u~+v~)+u~v~sr2)n4𝑑l=(1+o(1))Cxe(nr(u~+v~)sr2)χ(u)χ(v)𝑑l.\int_{C_{x}}\chi(u)\chi(v)\left(1-\frac{r(\tilde{u}+\tilde{v})+\tilde{u}\tilde{v}}{s-r^{2}}\right)^{{n-4}}dl=\left(1+o(1)\right)\int_{C_{x}}e^{\left(-\frac{nr(\tilde{u}+\tilde{v})}{s-r^{2}}\right)}\chi(u)\chi(v)dl.

We approximate the curve CxC_{x} with the following line

u~+v~=2(sx1xr)+o(1n).\tilde{u}+\tilde{v}=2\left(\sqrt{\frac{s-x}{1-x}}-r\right)+o\left(\frac{1}{n}\right).

It follows that

Cxe(nr(u~+v~)sr2)χ(u)χ(v)𝑑l=(1+o(1))22(δ+sx1xr)+e(2nr(sx1xr)sr2),\int_{C_{x}}e^{\left(-\frac{nr(\tilde{u}+\tilde{v})}{s-r^{2}}\right)}\chi(u)\chi(v)dl=(1+o(1))2\sqrt{2}\left(\delta+\sqrt{\frac{s-x}{1-x}}-r\right)^{+}e^{\left(-\frac{2nr\left(\sqrt{\frac{s-x}{1-x}}-r\right)}{s-r^{2}}\right)},

from which the conclusion follows.

4.2. Conditional density for sphere packings

Let s=cos(θ)s=\cos(\theta) and r=12(1s)r=\frac{1}{\sqrt{2(1-s)}}, where 13s12\frac{1}{3}\leq s\leq\frac{1}{2}. Let 0<δ=c1n0<\delta=\frac{c_{1}}{n} for some fixed c1>0c_{1}>0 that we specify later, and define

χ(y):={1 for 0yr+δ,0otherwise.\chi(y):=\begin{cases}1&\text{ for }0\leq y\leq r+\delta,\\ 0&\text{otherwise}.\end{cases}

Let 𝒂,𝒃\boldsymbol{a},\boldsymbol{b} be two randomly independently chosen points on n\mathbb{R}^{n} with respect to the Euclidean measure such that |𝒂|,|𝒃|r+δ,|\boldsymbol{a}|,|\boldsymbol{b}|\leq r+\delta, where |.||.| is the Euclidean norm. Let

U:=|𝒂|,U:=|\boldsymbol{a}|,
V:=|𝒃|,V:=|\boldsymbol{b}|,
T:=|𝒂𝒃|,T:=|\boldsymbol{a}-\boldsymbol{b}|,

and α\alpha be the angle between 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b}. The pushforward of the product measure on n×n\mathbb{R}^{n}\times\mathbb{R}^{n} onto the coordinates (U,V,α)(U,V,\alpha) is, up to a positive scalar depending only on nn, the measure

μ(U,V,α)dUdVdα=Un1Vn1sin(α)n3dUdVdα.\mu(U,V,\alpha)dUdVd\alpha=U^{n-1}V^{n-1}\sin(\alpha)^{n-3}dUdVd\alpha.

Let

(32) x:=cosα=U2+V2T22UV,x:=\cos\alpha=\frac{U^{2}+V^{2}-T^{2}}{2UV},

which follows from the cosine law. We have

sinαdα=TUVdTU2+V2T22UVUdUU2+V2T22UVVdV.\sin\alpha d\alpha=\frac{T}{UV}dT-\frac{\partial\frac{U^{2}+V^{2}-T^{2}}{2UV}}{\partial U}dU-\frac{\partial\frac{U^{2}+V^{2}-T^{2}}{2UV}}{\partial V}dV.

Hence, the pushforward of the product measure on n×n\mathbb{R}^{n}\times\mathbb{R}^{n} onto the coordinates (U,V,T)3(U,V,T)\in\mathbb{R}^{3} has the following density function up to a positive scalar depending only on nn:

μ(U,V,T)=(U2V2T)Δ(U,V,T)n4,\mu(U,V,T)=(U^{2}V^{2}T)\Delta(U,V,T)^{n-4},

where Δ(U,V,T)\Delta(U,V,T) is the Euclidean area of the triangle with sides U,V,TU,V,T. If no such triangle exists, then Δ(U,V,T)=0\Delta(U,V,T)=0. Let μ(x;T,χ)\mu(x;T,\chi) be the induced density function on xx subjected to the conditions of fixed T,T, and U,Vr+δU,V\leq r+\delta. Precisely, we have, up to a positive constant multiple depending on n,r+δn,r+\delta and TT, that

μ(x;T,χ):=Cx,Tχ(U)χ(V)μ(x,l,T)𝑑l,\mu(x;T,\chi):=\int_{C_{x,T}}\chi(U)\chi(V)\mu(x,l,T)dl,

where the integral is over the curve Cx,T2C_{x,T}\subset\mathbb{R}^{2} that is given by U2+V2T22UV=x\frac{U^{2}+V^{2}-T^{2}}{2UV}=x and dldl is the induced Euclidean metric on Cx,T,C_{x,T}, and

μ(x,l,T)dxdldT=μ(U,V,T)dUdVdT.\mu(x,l,T)dxdldT=\mu(U,V,T)dUdVdT.
Refer to caption
Figure 6. 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} are two points with T=|𝒂𝒃|T=|\boldsymbol{a}-\boldsymbol{b}|. The red curve is the locus of points 𝒛\boldsymbol{z} having (U,V)(U,V) on the curve Cx,TC_{x,T}, that is, 𝒛\boldsymbol{z} with respect to which 𝒂\boldsymbol{a} and 𝒃\boldsymbol{b} have angle whose cosine is xx. Shaded region corresponds to the support of χ(U)χ(V)\chi(U)\chi(V).

We have

dUdVdT=1(xU)2+(xV)2dxdldT,dUdVdT=\frac{1}{\sqrt{(\frac{\partial x}{\partial U})^{2}+(\frac{\partial x}{\partial V})^{2}}}dxdldT,

and

(xU)2+(xV)2=(TUV)2(1+x2)2x(1x2)UV.\left(\frac{\partial x}{\partial U}\right)^{2}+\left(\frac{\partial x}{\partial V}\right)^{2}=\left(\frac{T}{UV}\right)^{2}(1+x^{2})-\frac{2x(1-x^{2})}{UV}.

Hence,

μ(x;T,χ)=Cx,Tχ(U)χ(V)μ(U,V,T)(TUV)2(1+x2)2x(1x2)UV𝑑l\mu(x;T,\chi)=\int_{C_{x,T}}\frac{\chi(U)\chi(V)\mu(U,V,T)}{\sqrt{\left(\frac{T}{UV}\right)^{2}(1+x^{2})-\frac{2x(1-x^{2})}{UV}}}dl

up to a positive constant multiple depending on n,r+δn,r+\delta and TT.

From Lemma 3.7, equation (23) with F=χF=\chi and any given gg may be written as

(33) H(T)=11g(x)μ(x;T,χ)𝑑x.H(T)=\int_{-1}^{1}g(x)\mu(x;T,\chi)dx.

We now study the scaling property of μ(x;T,χ).\mu(x;T,\chi). Let χT(x):=χ(xT).\chi_{T}(x):=\chi(\frac{x}{T}).

Lemma 4.3.

We have

μ(x;T,χ)=T2n1μ(x;1,χT).\mu(x;T,\chi)=T^{2n-1}\mu(x;1,\chi_{T}).
Proof.

Note that

μ(U,V,T)=T2n3μ(UT,VT,1),\mu(U,V,T)=T^{2n-3}\mu\left(\frac{U}{T},\frac{V}{T},1\right),

and xx is invariant by scaling U,V,T.U,V,T. The conclusion follows. ∎

Let μ(x;χ):=μ(x;1,χ),\mu(x;\chi):=\mu(x;1,\chi), and

r1:=1(1x2)(r+δ)2+x(r+δ).r_{1}:=\sqrt{1-(1-x^{2})(r+\delta)^{2}}+x(r+\delta).
Proposition 4.4.

Suppose that |xs|c2n|x-s|\leq\frac{c_{2}}{n}, x<1/2x<1/2 and δ=c1n\delta=\frac{c_{1}}{n}. We have

μ(x;χ)=(1+E)(1x2)n42(12(1x))n111x1+(x(1x2)r1(1x2)r2)2(r+δr1)+\mu(x;\chi)=(1+E)(1-x^{2})^{\frac{n-4}{2}}\left(\frac{1}{2(1-x)}\right)^{{n-1}}\frac{1}{\sqrt{1-x}}\sqrt{1+\left(x-\frac{(1-x^{2})r}{\sqrt{1-(1-x^{2})r^{2}}}\right)^{2}}\left(r+\delta-r_{1}\right)^{+}

up to a positive scalar multiple making this a probability measure on [1,1][-1,1], and where EE is a function of xx with the uniform bound |E|<(4c2+2c1+2)2n|E|<\frac{(4c_{2}+2c_{1}+2)^{2}}{n} for n2000n\geq 2000.

Proof.

We have

μ(x;χ)=Cxχ(U)χ(V)μ(U,V,1)1+x2(UV)22x(1x2)UV𝑑l,\mu(x;\chi)=\int_{C_{x}}\chi(U)\chi(V)\frac{\mu(U,V,1)}{{\sqrt{\frac{1+x^{2}}{\left(UV\right)^{2}}-\frac{2x(1-x^{2})}{UV}}}}dl,

where Cx:=Cx,1.C_{x}:=C_{x,1}. We note that up to a positive scalar depending only on nn,

μ(U,V,1)=U2V2((1x2)U2V2)n42.\mu(U,V,1)=U^{2}V^{2}\left((1-x^{2})U^{2}V^{2}\right)^{\frac{n-4}{2}}.

Hence,

μ(x;χ)=(1x2)n42Cxχ(U)χ(V)(UV)n11+x22x(1x2)UV𝑑l.\begin{split}\mu(x;\chi)=(1-x^{2})^{\frac{n-4}{2}}\int_{C_{x}}\chi(U)\chi(V)\frac{\left(UV\right)^{n-1}}{{\sqrt{1+x^{2}-2x(1-x^{2})UV}}}dl.\end{split}

Suppose that UU and VV are in the support of χ.\chi. By concentration of mass, we may assume that U=r+U~U=r+\tilde{U} and V=r+V~,V=r+\tilde{V}, with U~,V~c1n.\tilde{U},\tilde{V}\leq\frac{c_{1}}{n}. From (32), we have

U~+V~=12r(1x)r+E1,\tilde{U}+\tilde{V}=\frac{1}{2r(1-x)}-r+E_{1},

where

E1:=12r(1x)(2xU~V~U~2V~2).E_{1}:=\frac{1}{2r(1-x)}\left(2x\tilde{U}\tilde{V}-\tilde{U}^{2}-\tilde{V}^{2}\right).

Since |xs|c2n,|x-s|\leq\frac{c_{2}}{n}, and 12<2r(1x)2,\frac{1}{2}<2r(1-x)^{2}, it follows that c1+2c2n<U~,V~c1n-\frac{c_{1}+2c_{2}}{n}<\tilde{U},\tilde{V}\leq\frac{c_{1}}{n} for n2000n\geq 2000. Hence,

|E1|3(c1+2c2)2n2.|E_{1}|\leq 3\frac{(c_{1}+2c_{2})^{2}}{n^{2}}.

We have

UV=r2+r(U~+V~)+U~V~=r2(1+U~+V~r)+U~V~=12(1x)+E2,UV=r^{2}+r(\tilde{U}+\tilde{V})+\tilde{U}\tilde{V}=r^{2}\left(1+\frac{\tilde{U}+\tilde{V}}{r}\right)+\tilde{U}\tilde{V}=\frac{1}{2(1-x)}+E_{2},

where |E2|<4(c1+2c2)2n2.|E_{2}|<4\frac{(c_{1}+2c_{2})^{2}}{n^{2}}. Furthermore,

1+x22x(1x2)UV=1x2x(1x2)E2=1x+E2,\sqrt{1+x^{2}-2x(1-x^{2})UV}=\sqrt{1-x-2x(1-x^{2})E_{2}}=\sqrt{1-x}+E_{2}^{\prime},

where |E2|<|E2|.|E_{2}^{\prime}|<|E_{2}|. Hence,

μ(x;χ)=(1x2)n42(12(1x))n111x+E2(1+E22(1x))n1Cxχ(U)χ(V)𝑑l.\begin{split}\mu(x;\chi)&=(1-x^{2})^{\frac{n-4}{2}}\left(\frac{1}{2(1-x)}\right)^{{n-1}}\frac{1}{\sqrt{1-x}+E_{2}^{\prime}}\left(1+\frac{E_{2}}{2(1-x)}\right)^{n-1}\int_{C_{x}}\chi(U)\chi(V)dl.\end{split}

Note that

11x+E2(1+E22(1x))n1=11x(1+E2′′),\frac{1}{\sqrt{1-x}+E_{2}^{\prime}}\left(1+\frac{E_{2}}{2(1-x)}\right)^{n-1}=\frac{1}{\sqrt{1-x}}(1+E_{2}^{\prime\prime}),

where E2′′4(c1+2c2)2nE_{2}^{\prime\prime}\leq 4\frac{(c_{1}+2c_{2})^{2}}{n} for n2000n\geq 2000. We parametrize the curve CxC_{x} with VV to obtain

U(V)=1(1x2)V2+xV.U(V)=\sqrt{1-(1-x^{2})V^{2}}+xV.

We have

dUdV=x(1x2)V1(1x2)V2=x(1x2)r1(1x2)r2+E3,\frac{dU}{dV}=x-\frac{(1-x^{2})V}{\sqrt{1-(1-x^{2})V^{2}}}=x-\frac{(1-x^{2})r}{\sqrt{1-(1-x^{2})r^{2}}}+E_{3},

where

|E3|=|(Vr)(1x2)(1(1x2)V12)3/2||E_{3}|=\left|(V-r)\frac{(1-x^{2})}{(1-(1-x^{2})V_{1}^{2})^{3/2}}\right|

for some V1(rc1+2c2n,r+c1n),V_{1}\in\left(r-\frac{c_{1}+2c_{2}}{n},r+\frac{c_{1}}{n}\right), which implies |(1x2)(1(1x2)V12)3/2|<6.\left|\frac{(1-x^{2})}{(1-(1-x^{2})V_{1}^{2})^{3/2}}\right|<6. Hence,

|E3|6(c1+2c2n).|E_{3}|\leq 6\left(\frac{c_{1}+2c_{2}}{n}\right).

Hence,

Cxχ(U)χ(V)𝑑l=+r1r+δ1+(dUdV)2𝑑V=1+(x(1x2)r1(1x2)r2)2(r+δr1)+(1+E3¯)\begin{split}\int_{C_{x}}\chi(U)\chi(V)dl&=^{+}\int_{r_{1}}^{r+\delta}\sqrt{1+\left(\frac{dU}{dV}\right)^{2}}dV\\ &=\sqrt{1+\left(x-\frac{(1-x^{2})r}{\sqrt{1-(1-x^{2})r^{2}}}\right)^{2}}\left(r+\delta-r_{1}\right)^{+}(1+\bar{E_{3}})\end{split}

where

r1:=1(1x2)(r+δ)2+x(r+δ),r_{1}:=\sqrt{1-(1-x^{2})(r+\delta)^{2}}+x(r+\delta),

and

|E3|6(c1+2c2n).|E_{3}|\leq 6\left(\frac{c_{1}+2c_{2}}{n}\right).

The first equality =+=^{+} above means equal to 0 if r+δ<r1r+\delta<r_{1}. Therefore,

μ(x;χ)=(1+E)(1x2)n42(12(1x))n111x1+(x(1x2)r1(1x2)r2)2(r+δr1)+\mu(x;\chi)=(1+E)(1-x^{2})^{\frac{n-4}{2}}\left(\frac{1}{2(1-x)}\right)^{{n-1}}\frac{1}{\sqrt{1-x}}\sqrt{1+\left(x-\frac{(1-x^{2})r}{\sqrt{1-(1-x^{2})r^{2}}}\right)^{2}}\left(r+\delta-r_{1}\right)^{+}

up to a positive scalar multiple, where |E|<(4c2+2c1+2)2n|E|<\frac{(4c_{2}+2c_{1}+2)^{2}}{n} for n2000n\geq 2000. This completes the proof of our Proposition. ∎

5. Comparison with previous bounds

We define Jacobi polynomials, state some of their properties, and prove a local approximation result for Jacobi polynomials in Subsection 5.1. In Subsection 5.2, we improve bounds on θ\theta-spherical codes. In Subsection 5.3, we improve upper bounds on sphere packing densities. The general constructions of our test functions were provided in Section 3. For each of our main theorems, we determine the largest values of δ=O(1/n)\delta=O(1/n) measuring the extent to which the supports of the characteristic functions χ\chi in Propositions 3.6 and 3.8 could be enlarged. See the end of Section 3 for an outline of the general strategy.

5.1. Jacobi polynomials and their local approximation

Recall definition (7) of MLev(n,θ)M_{\textup{Lev}}(n,\theta). Levenshtein proved inequality (6) by applying Delsarte’s linear programming bound to a family of even and odd degree polynomials inside D(dμn32,cosθ),D(d\mu_{\frac{n-3}{2}},\cos\theta), which we now discuss. In order to define these Levenshtein polynomials, we record some well-known properties of Jacobi polynomials (see [Sze39, Chapter IV]) that we will also use in the rest of the paper.

We denote by pdα,β(t)p^{\alpha,\beta}_{d}(t) Jacobi polynomials of degree dd with parameters α\alpha and β\beta. These are orthogonal polynomials with respect to the probability measure

dμα,β:=(1t)α(1+t)βdt11(1t)α(1+t)β𝑑td\mu_{\alpha,\beta}:=\frac{(1-t)^{\alpha}(1+t)^{\beta}dt}{\int_{-1}^{1}(1-t)^{\alpha}(1+t)^{\beta}dt}

on the interval [1,1][-1,1] with the normalization that gives

pdα,β(1)=(d+αd).p^{\alpha,\beta}_{d}(1)=\binom{d+\alpha}{d}.

pdα,β(t)p^{\alpha,\beta}_{d}(t) has nn simple real roots t1,dα,β>t2,dα,β>>td,dα,β.t_{1,d}^{\alpha,\beta}>t_{2,d}^{\alpha,\beta}>\dots>t_{d,d}^{\alpha,\beta}. When α=β\alpha=\beta, we denote the measure dμα,αd\mu_{\alpha,\alpha} simply as dμαd\mu_{\alpha}.

(34) ddtpdα,β(t)=d+α+β+12pd1α+1,β+1(t).\frac{d}{dt}p_{d}^{\alpha,\beta}(t)=\frac{d+\alpha+\beta+1}{2}p_{d-1}^{\alpha+1,\beta+1}(t).

When proving our local approximation result on Levenshtein’s optimal polynomials in the rest of this subsection, we use the fact that the Jacobi polynomial pdα,β(t)p_{d}^{\alpha,\beta}(t) satisfies the differential equation

(35) (1t2)x′′(t)+(βα(α+β+2)t)x(t)+d(d+α+β+1)x(t)=0.(1-t^{2})x^{\prime\prime}(t)+(\beta-\alpha-(\alpha+\beta+2)t)x^{\prime}(t)+d(d+\alpha+\beta+1)x(t)=0.

By [Lev98, Lemma 5.89],

t1,d1α+1,α+1<t1,dα+1,α<t1,dα+1,α+1.t^{\alpha+1,\alpha+1}_{1,d-1}<t^{\alpha+1,\alpha}_{1,d}<t^{\alpha+1,\alpha+1}_{1,d}.

It is also well-known that for fixed α\alpha, t1,dα+1,α+11t^{\alpha+1,\alpha+1}_{1,d}\rightarrow 1 as dd\rightarrow\infty. Henceforth, let d:=d(n,θ)d:=d(n,\theta) be uniquely determined by t1,d1α+1,α+1<cos(θ)t1,dα+1,α+1t^{\alpha+1,\alpha+1}_{1,d-1}<\cos(\theta)\leq t^{\alpha+1,\alpha+1}_{1,d}. Let [Lev98, Lemma 5.38]

(36) gn,θ(x)={(x+1)2(xt1,dα+1,α+1)(pdα+1,α+1(x))2 if t1,dα+1,α<cos(θ)t1,dα+1,α+1,(x+1)(xt1,dα+1,α)(pdα+1,α(x))2 if t1,d1α+1,α+1<cos(θ)t1,dα+1,α,g_{n,\theta}(x)=\begin{cases}\frac{(x+1)^{2}}{(x-t^{\alpha+1,\alpha+1}_{1,d})}\left(p^{\alpha+1,\alpha+1}_{d}(x)\right)^{2}&\text{ if }t^{\alpha+1,\alpha}_{1,d}<\cos(\theta)\leq t^{\alpha+1,\alpha+1}_{1,d},\\ \frac{(x+1)}{(x-t^{\alpha+1,\alpha}_{1,d})}\left(p^{\alpha+1,\alpha}_{d}(x)\right)^{2}&\text{ if }t^{\alpha+1,\alpha+1}_{1,d-1}<\cos(\theta)\leq t^{\alpha+1,\alpha}_{1,d},\end{cases}

where α:=n32\alpha:=\frac{n-3}{2} in our case. Levenshtein proved that gn,θD(dμn32,cosθ),g_{n,\theta}\in D(d\mu_{\frac{n-3}{2}},\cos\theta), and

(𝓃,θ)=Lev(𝓃,θ).\cal{L}(g_{n,\theta})=M_{\textup{Lev}}(n,\theta).

By (3), this gives

M(n,θ)MLev(n,θ).M(n,\theta)\leq M_{\textup{Lev}}(n,\theta).

As part of our proofs of our main theorems in the next subsections, we need to determine local approximations to Jacobi polynomials pdα,βp_{d}^{\alpha,\beta} in the neighbourhood of points s(1,1)s\in(-1,1) such that st1,dα,βs\geq t_{1,d}^{\alpha,\beta}. This is obtained using the behaviour of the zeros of Jacobi polynomials. Using this, we obtain suitable local approximations of Levenshtein’s optimal functions near ss.

Proposition 5.1.

Suppose that αβ0,\alpha\geq\beta\geq 0, |αβ|1|\alpha-\beta|\leq 1, d0d\geq 0 and s[t1,dα,β,1)s\in[t_{1,d}^{\alpha,\beta},1). Then, we have

pdα,β(t)=pdα,β(s)+(ts)dpdα,βdt(s)(1+A(t)),p_{d}^{\alpha,\beta}(t)=p_{d}^{\alpha,\beta}(s)+(t-s)\frac{dp_{d}^{\alpha,\beta}}{dt}(s)(1+A(t)),

where,

|A(t)|eσ(t)1σ(t)1|A(t)|\leq\frac{e^{\sigma(t)}-1}{\sigma(t)}-1

with

σ(t):=|ts|(2αs+2s+1)1s2.\sigma(t):=\frac{|t-s|\left(2\alpha s+2s+1\right)}{1-s^{2}}.
Proof.

Consider the Taylor expansion

pdα,β(t)=k=0(ts)kk!dkpdα,βdtk(s)p_{d}^{\alpha,\beta}(t)=\sum_{k=0}^{\infty}\frac{(t-s)^{k}}{k!}\frac{d^{k}p_{d}^{\alpha,\beta}}{dt^{k}}(s)

of pdα,βp_{d}^{\alpha,\beta} centered at ss. We prove the proposition by showing that for s[t1,dα,β,1)s\in[t_{1,d}^{\alpha,\beta},1), the higher degree terms in the Taylor expansion are small in comparison to the linear term. Indeed, suppose k1k\geq 1. Then, using equation (34),

(dk+1/dtk+1)pdα,β(s)(dk/dtk)pdα,β(s)=(d/dt)pdkα+k,β+k(s)pdkα+k,β+k(s)=i=1dk1sti,dkα+k,β+ki=1d11sti,d1α+1,β+1,\frac{(d^{k+1}/dt^{k+1})p_{d}^{\alpha,\beta}(s)}{(d^{k}/dt^{k})p_{d}^{\alpha,\beta}(s)}=\frac{(d/dt)p_{d-k}^{\alpha+k,\beta+k}(s)}{p_{d-k}^{\alpha+k,\beta+k}(s)}=\sum_{i=1}^{d-k}\frac{1}{s-t_{i,d-k}^{\alpha+k,\beta+k}}\leq\sum_{i=1}^{d-1}\frac{1}{s-t_{i,d-1}^{\alpha+1,\beta+1}},

where the last inequality follows from the fact that the roots of a Jacobi polynomial interlace with those of its derivative. However, the last quantity is equal to (d2/dt2)pdα,β(s)(d/dt)pdα,β(s)\frac{(d^{2}/dt^{2})p_{d}^{\alpha,\beta}(s)}{(d/dt)p_{d}^{\alpha,\beta}(s)}. We proceed to show that

(37) (d2/dt2)pdα,β(s)(d/dt)pdα,β(s)2αs+2s+11s2.\frac{(d^{2}/dt^{2})p_{d}^{\alpha,\beta}(s)}{(d/dt)p_{d}^{\alpha,\beta}(s)}\leq\frac{2\alpha s+2s+1}{1-s^{2}}.

Indeed, we know from the differential equation (35) that

(38) (1s2)(d2/dt2)pdα,β(s)+(βα(α+β+2)s)(d/dt)pdα,β(s)+d(d+α+β+1)pdα,β(s)=0.(1-s^{2})(d^{2}/dt^{2})p_{d}^{\alpha,\beta}(s)+(\beta-\alpha-(\alpha+\beta+2)s)(d/dt)p_{d}^{\alpha,\beta}(s)+d(d+\alpha+\beta+1)p_{d}^{\alpha,\beta}(s)=0.

However, since ss is to the right of the largest root of pdα,βp_{d}^{\alpha,\beta}, pdα,β(s)0p_{d}^{\alpha,\beta}(s)\geq 0. Therefore,

(1s2)(d2/dt2)pdα,β(s)+(βα(α+β+2)s)(d/dt)pdα,β(s)0,(1-s^{2})(d^{2}/dt^{2})p_{d}^{\alpha,\beta}(s)+(\beta-\alpha-(\alpha+\beta+2)s)(d/dt)p_{d}^{\alpha,\beta}(s)\leq 0,

from which inequality (37) follows. As a result, the degree k+1k+1 term compares to the linear term as

(dk+1/dtk+1)pdα,β(s)(d/dt)pdα,β(s)|ts|k(k+1)!\displaystyle\frac{(d^{k+1}/dt^{k+1})p_{d}^{\alpha,\beta}(s)}{(d/dt)p_{d}^{\alpha,\beta}(s)}\frac{|t-s|^{k}}{(k+1)!} \displaystyle\leq |ts|k(k+1)!(2(α+1)s+11s2)k\displaystyle\frac{|t-s|^{k}}{(k+1)!}\left(\frac{2(\alpha+1)s+1}{1-s^{2}}\right)^{k}

Consequently, for every k1k\geq 1,

|ts|k+1(k+1)!(dk+1/dtk+1)pdα,β(s)|ts|(d/dt)pdα,β(s)(|ts|(2α+2s+1)1s2)k(k+1)!\frac{|t-s|^{k+1}}{(k+1)!}(d^{k+1}/dt^{k+1})p_{d}^{\alpha,\beta}(s)\leq|t-s|(d/dt)p_{d}^{\alpha,\beta}(s)\frac{\left(\frac{|t-s|\left(2\alpha+2s+1\right)}{1-s^{2}}\right)^{k}}{(k+1)!}

As a result, we obtain that

pdα,β(t)=pdα,β(s)+(ts)dpdα,βdt(s)(1+A(t)),p_{d}^{\alpha,\beta}(t)=p_{d}^{\alpha,\beta}(s)+(t-s)\frac{dp_{d}^{\alpha,\beta}}{dt}(s)(1+A(t)),

where

|A(t)|eσ(t)1σ(t)1|A(t)|\leq\frac{e^{\sigma(t)}-1}{\sigma(t)}-1

with

σ(t)=|ts|(2αs+2s+1)1s2.\sigma(t)=\frac{|t-s|\left(2\alpha s+2s+1\right)}{1-s^{2}}.

5.2. Improving spherical codes bound

We prove a stronger version of Theorem 1.1 that we now state.

Theorem 5.2.

Fix θ<θ\theta<\theta^{*} and suppose 0<θ<θπ/20<\theta<\theta^{\prime}\leq\pi/2. Then there is a function hD(dμn32,cosθ)h\in D(d\mu_{\frac{n-3}{2}},\cos\theta) such that

(𝒽)𝒸𝓃Lev(𝓃1,θ)λ𝓃(θ,θ),\cal{L}(h)\leq c_{n}\frac{M_{\textup{Lev}}(n-1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})},

where cn0.4325c_{n}\leq 0.4325 for large enough nn independent of θ\theta and θ.\theta^{\prime}.

Proof.

Without loss of generality, we may assume that cos(θ)=t1,dα+1,α+ε\cos(\theta^{\prime})=t^{\alpha+1,\alpha+\varepsilon}_{1,d} for some ε{0,1}\varepsilon\in\{0,1\} and α:=n42.\alpha:=\frac{n-4}{2}. Indeed, recall from Subsection 5.1 that dd is uniquely determined by t1,d1α+1,α+1<cos(θ)t1,dα+1,α+1t^{\alpha+1,\alpha+1}_{1,d-1}<\cos(\theta^{\prime})\leq t^{\alpha+1,\alpha+1}_{1,d},

g(x)={(x+1)2(xt1,dα+1,α+1)(pdα+1,α+1(x))2 if t1,dα+1,α<cos(θ)t1,dα+1,α+1,(x+1)(xt1,dα+1,α)(pdα+1,α(x))2 if t1,d1α+1,α+1<cos(θ)t1,dα+1,α,g(x)=\begin{cases}\frac{(x+1)^{2}}{(x-t^{\alpha+1,\alpha+1}_{1,d})}\left(p^{\alpha+1,\alpha+1}_{d}(x)\right)^{2}&\text{ if }t^{\alpha+1,\alpha}_{1,d}<\cos(\theta^{\prime})\leq t^{\alpha+1,\alpha+1}_{1,d},\\ \frac{(x+1)}{(x-t^{\alpha+1,\alpha}_{1,d})}\left(p^{\alpha+1,\alpha}_{d}(x)\right)^{2}&\text{ if }t^{\alpha+1,\alpha+1}_{1,d-1}<\cos(\theta^{\prime})\leq t^{\alpha+1,\alpha}_{1,d},\end{cases}

and

MLev(n1,θ)=()=Lev(𝓃1,arccos(𝓉1,𝒹α+1,α+ε)).M_{\textup{Lev}}(n-1,\theta^{\prime})=\cal{L}(g)=M_{\textup{Lev}}(n-1,\arccos(t^{\alpha+1,\alpha+\varepsilon}_{1,d})).

We also have

λn(θ,θ)λn(θ,arccos(t1,dα+1,α+ε)),\lambda_{n}(\theta,\theta^{\prime})\leq\lambda_{n}(\theta,\arccos(t^{\alpha+1,\alpha+\varepsilon}_{1,d})),

where the above follows from cos(θ)t1,dα+1,α+ε\cos(\theta^{\prime})\leq t^{\alpha+1,\alpha+\varepsilon}_{1,d} and λn(θ,θ)\lambda_{n}(\theta,\theta^{\prime}) is the ratio of volume of the spherical cap with radius sin(θ/2)sin(θ/2)\frac{\sin(\theta/2)}{\sin(\theta^{\prime}/2)} on the unit sphere Sn1S^{n-1} to the volume of the whole sphere. Hence,

MLev(n1,arccos(t1,dα+1,α+ε))λn(θ,arccos(t1,dα+1,α+ε))MLev(n1,θ)λn(θ,θ).\frac{M_{\textup{Lev}}(n-1,\arccos(t^{\alpha+1,\alpha+\varepsilon}_{1,d}))}{\lambda_{n}(\theta,\arccos(t^{\alpha+1,\alpha+\varepsilon}_{1,d}))}\leq\frac{M_{\textup{Lev}}(n-1,\theta^{\prime})}{\lambda_{n}(\theta,\theta^{\prime})}.

As before, s=cos(θ),s=\cos(\theta), and s=cos(θ)=t1,dα+1,α+ε.s^{\prime}=\cos(\theta^{\prime})=t^{\alpha+1,\alpha+\varepsilon}_{1,d}. Note that s<ss^{\prime}<s and for r=ss1sr=\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}, we have s=sr21r2s^{\prime}=\frac{s-r^{2}}{1-r^{2}} and 0<r<10<r<1. Let 0<δ=O(1n)0<\delta=O(\frac{1}{n}) that we specify later, and define the function FF for the application of Proposition 3.6 to be

F(y)=χ(y):={1 for rδ<yR,0otherwise.F(y)=\chi(y):=\begin{cases}1&\text{ for }r-\delta<y\leq R,\\ 0&\text{otherwise}.\end{cases}

Recall from (31) that

h(t):=11g(x)μ(x,t;χ)𝑑x.h(t):=\int_{-1}^{1}g(x)\mu(x,t;\chi)dx.

Applying Proposition 3.6 with the function FF as above and the function gg, we obtain the inequality in the statement of the theorem with cn1+o(1)c_{n}\leq 1+o(1). We now prove the desired bound on cnc_{n} for sufficiently large nn.

By Proposition 3.9, for any δ1>0\delta_{1}>0, if |scos(θ)|δ1|s^{\prime}-\cos(\theta^{*})|\geq\delta_{1}, then for large nn, MLev(n1,θ)/λn(θ,θ)M_{\textup{Lev}}(n-1,\theta^{\prime})/\lambda_{n}(\theta,\theta^{\prime}) is exponentially worse than MLev(n1,θ)/λn(θ,θ)M_{\textup{Lev}}(n-1,\theta^{*})/\lambda_{n}(\theta,\theta^{*}). Therefore, we assume that s=cos(θ)=t1,dα+1,α+εs^{\prime}=\cos(\theta^{\prime})=t_{1,d}^{\alpha+1,\alpha+\varepsilon} for some ε{0,1}\varepsilon\in\{0,1\}, and that ss^{\prime} is sufficiently close to cosθ\cos\theta^{*}. We specify the precision of the difference later. Suppose ε=0\varepsilon=0, and so s=t1,dα+1,αs^{\prime}=t^{\alpha+1,\alpha}_{1,d}. By Proposition 5.1, we have

g(x)=(x+1)(xs)dpdα+1,αdt(s)2(1+A(x))2,g(x)=(x+1)(x-s^{\prime})\frac{dp_{d}^{\alpha+1,\alpha}}{dt}(s^{\prime})^{2}(1+A(x))^{2},

where

(39) |A(x)|eσ(x)1σ(x)1|A(x)|\leq\frac{e^{\sigma(x)}-1}{\sigma(x)}-1

with

σ(x):=|xs|(ns+s+1)1s2.\sigma(x):=\frac{|x-s^{\prime}|(ns^{\prime}+s^{\prime}+1)}{1-s^{\prime 2}}.

By Proposition 4.2, we have for |xs|=o(1n)|x-s^{\prime}|=o\left(\frac{1}{\sqrt{n}}\right) the estimate

μ(x;s,χ)=(2(1r2)2r(1s)+o(1))(δ+sx1xr)+((1x2x2)(sr2)2)n42e(2nr(sx1xr)sr2).\mu(x;s,\chi)=\left(\frac{2(1-r^{2})^{2}}{r(1-s)}+o(1)\right)\left(\delta+\sqrt{\frac{s-x}{1-x}}-r\right)^{+}\left(\left(\frac{1-x^{2}}{x^{2}}\right)\left(s-r^{2}\right)^{2}\right)^{\frac{n-4}{2}}e^{\left(-\frac{2nr\left(\sqrt{\frac{s-x}{1-x}}-r\right)}{s-r^{2}}\right)}.

We need to find the maximal δ>0\delta>0 such that

h(t)=11g(x)μ(x;t,χ)𝑑x0h(t)=\int_{-1}^{1}g(x)\mu(x;t,\chi)dx\leq 0

for every 1ts-1\leq t\leq s. We first address the above inequality for t=st=s. Note that the integrand is negative for x<sx<s^{\prime} and positive for x>s.x>s^{\prime}. Hence,

11g(x)μ(x;s,χ)𝑑x=|s1g(x)μ(x;s,χ)𝑑x||1sg(x)μ(x;s,χ)𝑑x|.\displaystyle\int_{-1}^{1}g(x)\mu(x;s,\chi)dx=\left|\int_{s^{\prime}}^{1}g(x)\mu(x;s,\chi)dx\right|-\left|\int_{-1}^{s^{\prime}}g(x)\mu(x;s,\chi)dx\right|.

Its non-positivity is equivalent to

|1sg(x)μ(x;s,χ)𝑑x||s1g(x)μ(x;s,χ)𝑑x|.\left|\int_{-1}^{s^{\prime}}g(x)\mu(x;s,\chi)dx\right|\geq\left|\int_{s^{\prime}}^{1}g(x)\mu(x;s,\chi)dx\right|.

We proceed to give a lower bound on the absolute value of the integral over 1xs.-1\leq x\leq s^{\prime}. Later, we give an upper bound on the right hand side. By (39), we have

|1+A(x)|(2eσ(x)1σ(x))+.|1+A(x)|\geq\left(2-\frac{e^{\sigma(x)}-1}{\sigma(x)}\right)^{+}.

We note that 2eσ1σ2-\frac{e^{\sigma}-1}{\sigma} is a concave function with value 1 as σ0\sigma\rightarrow 0 and a root at σ=1.256431\sigma=1.256431.... Hence, for σ<1.25643\sigma<1.25643, we have

|1+A(x)|(2eσ(x)1σ(x)).|1+A(x)|\geq\left(2-\frac{e^{\sigma(x)}-1}{\sigma(x)}\right).

Note that σ(x)<1.25643\sigma(x)<1.25643 implies

|xs|(1.25643)(1s2)ns.|x-s^{\prime}|\leq\frac{(1.25643)(1-s^{\prime 2})}{ns^{\prime}}.

Let

λ:=(1.25643)(1s2)ns.\lambda:=\frac{(1.25643)(1-s^{\prime 2})}{ns^{\prime}}.

Therefore, as nn\rightarrow\infty

(40) r(1s)2(1r2)2(sr2)(n4)dpd(α+1,α)dt(s)2|1sg(x)μ(x;s,χ)𝑑x|\displaystyle\frac{r(1-s)}{2(1-r^{2})^{2}}\left(s-r^{2}\right)^{-(n-4)}\frac{dp_{d}^{(\alpha+1,\alpha)}}{dt}(s^{\prime})^{-2}\left|\int_{-1}^{s^{\prime}}g(x)\mu(x;s,\chi)dx\right|
sλs(1+x)(sx)(2eσ(x)1σ(x))2(δ+sx1xr)(1x2x2)n42e(2nr(sx1xr)sr2)𝑑x.\displaystyle\gtrsim\int_{s^{\prime}-\lambda}^{s^{\prime}}(1+x)(s^{\prime}-x)\left(2-\frac{e^{\sigma(x)}-1}{\sigma(x)}\right)^{2}\left(\delta+\sqrt{\frac{s-x}{1-x}}-r\right)\left(\frac{1-x^{2}}{x^{2}}\right)^{\frac{n-4}{2}}e^{\left(-\frac{2nr\left(\sqrt{\frac{s-x}{1-x}}-r\right)}{s-r^{2}}\right)}dx.

We change the variable sxs^{\prime}-x to zz and note that

(41) sx1xr=z(1s)2(ss)1/2(1s)3/2+O(λ2)\sqrt{\frac{s-x}{1-x}}-r=\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}+O(\lambda^{2})
(42) 1x2x2=1s2s2(1+2zs(1s2)+O(λ2))\frac{1-x^{2}}{x^{2}}=\frac{1-s^{\prime 2}}{s^{\prime 2}}\left(1+\frac{2z}{s^{\prime}(1-s^{\prime 2})}+O(\lambda^{2})\right)

for |xs|<λ.|x-s^{\prime}|<\lambda. Hence, we obtain that as nn\rightarrow\infty and |xs|<λ|x-s^{\prime}|<\lambda,

e(2nr(sx1xr)sr2)\displaystyle e^{\left(-\frac{2nr\left(\sqrt{\frac{s-x}{1-x}}-r\right)}{s-r^{2}}\right)} =\displaystyle= e(2nr(z(1s)2(ss)1/2(1s)3/2)sr2)+O(λ)=e(nzs(1s))+O(λ)\displaystyle e^{\left(-\frac{2nr\left(\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\right)}{s-r^{2}}\right)+O(\lambda)}=e^{\left(\frac{-nz}{s^{\prime}(1-s^{\prime})}\right)+O(\lambda)}
(1x2x2)n42\displaystyle\left(\frac{1-x^{2}}{x^{2}}\right)^{\frac{n-4}{2}} =\displaystyle= (1s2s2)n42e(nzs(1s2))+O(λ)\displaystyle\left(\frac{1-s^{\prime 2}}{s^{\prime 2}}\right)^{\frac{n-4}{2}}e^{\left(\frac{nz}{s^{\prime}(1-s^{\prime 2})}\right)+O(\lambda)}
2eσ(x)1σ(x)\displaystyle 2-\frac{e^{\sigma(x)}-1}{\sigma(x)} =\displaystyle= (2enzs(1s2)1nzs(1s2))(1+O(λ)).\displaystyle\left(2-\frac{e^{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}-1}{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}\right)(1+O(\lambda)).

for |xs|<λ.|x-s^{\prime}|<\lambda. We replace the above asymptotic formulas and obtain that as nn\rightarrow\infty, the right hand side of (40) is at least

(1+s)(1s2s2)n420λz(2enzs(1s2)1nzs(1s2))2(δ+z(1s)2(ss)1/2(1s)3/2)e(nz1s2)𝑑z.(1+s^{\prime})\left(\frac{1-s^{\prime 2}}{s^{\prime 2}}\right)^{\frac{n-4}{2}}\int_{0}^{\lambda}z\left(2-\frac{e^{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}-1}{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}\right)^{2}\left(\delta+\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\right)e^{\left(-\frac{nz}{1-s^{\prime 2}}\right)}dz.

We now give an upper bound on the absolute value of the integral over sx1.s^{\prime}\leq x\leq 1. We note that

sx1xr=z(1s)2(ss)1/2(1s)3/2+O(λ2).\sqrt{\frac{s-x}{1-x}}-r=\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}+O(\lambda^{2}).

Let Λ:=2(ss)1/2(1s)3/2δ(1s)\Lambda:=\frac{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}\delta}{(1-s)}. We have

(δ+sx1xr)+=0\left(\delta+\sqrt{\frac{s-x}{1-x}}-r\right)^{+}=0

for xs>Λ.x-s^{\prime}>\Lambda. We have

|1+A(x)|eσ(x)1σ(x),|1+A(x)|\leq\frac{e^{\sigma(x)}-1}{\sigma(x)},

where

σ(x)=n|xs|s1s2+o(1).\sigma(x)=\frac{n|x-s^{\prime}|s^{\prime}}{1-s^{{}^{\prime}2}}+o(1).

Therefore,

r(1s)2(1r2)2(sr2)(n4)dpd(α+1,α)dt(s)2|s1g(x)μ(x;s,χ)𝑑x|\displaystyle\frac{r(1-s)}{2(1-r^{2})^{2}}\left(s-r^{2}\right)^{-(n-4)}\frac{dp_{d}^{(\alpha+1,\alpha)}}{dt}(s^{\prime})^{-2}\left|\int_{s^{\prime}}^{1}g(x)\mu(x;s,\chi)dx\right|
\displaystyle\lesssim (1+s)(1s2s2)n420Λz(enzs(1s2)1nzs(1s2))2(δz(1s)2(ss)1/2(1s)3/2)e(nz1s2)𝑑z.\displaystyle(1+s^{\prime})\left(\frac{1-s^{\prime 2}}{s^{\prime 2}}\right)^{\frac{n-4}{2}}\int_{0}^{\Lambda}z\left(\frac{e^{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}-1}{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}\right)^{2}\left(\delta-\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\right)e^{\left(\frac{nz}{1-s^{\prime 2}}\right)}dz.

We choose δ\delta such that

0λz(2enzs(1s2)1nzs(1s2))2(δ+z(1s)2(ss)1/2(1s)3/2)e(nz1s2)𝑑z0Λz(enzs(1s2)1nzs(1s2))2(δz(1s)2(ss)1/2(1s)3/2)e(nz1s2)𝑑z.\begin{split}\int_{0}^{\lambda}z\left(2-\frac{e^{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}-1}{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}\right)^{2}\left(\delta+\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\right)e^{\left(-\frac{nz}{1-s^{\prime 2}}\right)}dz\\ \geq\int_{0}^{\Lambda}z\left(\frac{e^{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}-1}{\frac{nzs^{\prime}}{(1-s^{\prime 2})}}\right)^{2}\left(\delta-\frac{z(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\right)e^{\left(\frac{nz}{1-s^{\prime 2}}\right)}dz.\end{split}

For large enough nn we may replace the numerical value cos(1.0995124)\cos(1.0995124) for ss^{\prime} as we may assume that ss^{\prime} is close to cos(θ)\cos(\theta^{*}). Furthermore, write v:=nzv:=nz, and divide by δ\delta to obtain

02.196823v(2e0.571931v10.571931v)2(1+vnΛ)e(1.259674v)𝑑v0nΛv(e0.571931v10.571931v)2(1vnΛ)e(1.259674v)𝑑v.\begin{split}\int_{0}^{2.196823}v\left(2-\frac{e^{0.571931v}-1}{0.571931v}\right)^{2}\left(1+\frac{v}{n\Lambda}\right)e^{\left(-1.259674v\right)}dv\\ \geq\int_{0}^{n\Lambda}v\left(\frac{e^{0.571931v}-1}{0.571931v}\right)^{2}\left(1-\frac{v}{n\Lambda}\right)e^{\left(1.259674v\right)}dv.\end{split}

Here, we have also used that Λ=2δ(ss)1/2(1s)3/2(1s)\Lambda=\frac{2\delta(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}{(1-s)}. Also, note that nλ=2.196823n\lambda=2.196823... when ss^{\prime} is near cos(1.0995124)\cos(1.0995124). We have the Taylor expansion around v=0v=0

v(e0.571931v10.571931v)2e(1.259674v)\displaystyle v\left(\frac{e^{0.571931v}-1}{0.571931v}\right)^{2}e^{\left(1.259674v\right)}
=\displaystyle= v+1.83161v2+1.70465v3+1.07403v4+0.514959v5+0.200242v6+0.0657225v7\displaystyle v+1.83161v^{2}+1.70465v^{3}+1.07403v^{4}+0.514959v^{5}+0.200242v^{6}+0.0657225v^{7}
+\displaystyle+ 0.0187113v8+0.00471321v9+0.0010662v10+0.000219143v11+0.0000413083v12+Er\displaystyle 0.0187113v^{8}+0.00471321v^{9}+0.0010662v^{10}+0.000219143v^{11}+0.0000413083v^{12}+Er

with error |Er|<2×105|Er|<2\times 10^{-5} if nΛ<0.92n\Lambda<0.92, which we assume to be the case. Simplifying, we want to find the maximal nΛn\Lambda such that

0.195783+0.140655nΛ0nΛv(e0.571556v10.571556v)2(1vnΛ)e(1.259392v)𝑑v\displaystyle 0.195783+\frac{0.140655}{n\Lambda}\geq\int_{0}^{n\Lambda}v\left(\frac{e^{0.571556v}-1}{0.571556v}\right)^{2}\left(1-\frac{v}{n\Lambda}\right)e^{\left(1.259392v\right)}dv
=\displaystyle= 3.17756×106(nΛ)2((nΛ)11+5.74715(nΛ)10+30.5037(nΛ)9+148.328(nΛ)8+654.286(nΛ)7\displaystyle 3.17756\times 10^{-6}(n\Lambda)^{2}\Bigg{(}(n\Lambda)^{11}+5.74715(n\Lambda)^{10}+30.5037(n\Lambda)^{9}+148.328(n\Lambda)^{8}+654.286(n\Lambda)^{7}
+\displaystyle+ 2585.41(nΛ)6+9002.5(nΛ)5+27010.2(nΛ)4+67600.9(nΛ)3+134116(nΛ)2+192140(nΛ)+157353)\displaystyle 2585.41(n\Lambda)^{6}+9002.5(n\Lambda)^{5}+27010.2(n\Lambda)^{4}+67600.9(n\Lambda)^{3}+134116(n\Lambda)^{2}+192140(n\Lambda)+157353\Bigg{)}
\displaystyle- 1nΛ(0.360653+2.42691e1.25967(nΛ)3.33819e1.83161(nΛ)+1.27193e2.40354(nΛ))+Er,\displaystyle\frac{1}{n\Lambda}\left(-0.360653+2.42691e^{1.25967(n\Lambda)}-3.33819e^{1.83161(n\Lambda)}+1.27193e^{2.40354(n\Lambda)}\right)+Er,

where the error ErEr again satisfies |Er|2×105|Er|\leq 2\times 10^{-5}. A numerical computation gives us that the inequality is satisfied when nΛ0.915451n\Lambda\leq 0.915451.... Consequently, if we choose δ=/n\delta=\ell/n, then we must have

0.915451(1s)2(ss)1/2(1s)3/2.\ell\leq 0.915451...\frac{(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}.

In this case, the cap of radius 1r2\sqrt{1-r^{2}} becomes 1(rδ)2=1r2(1+rn(1r2))+O(1/n2)\sqrt{1-(r-\delta)^{2}}=\sqrt{1-r^{2}}\left(1+\frac{\ell r}{n(1-r^{2})}\right)+O(1/n^{2}). Note that r=ss1sr=\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}, and so

r1r2=(ss)1/2(1s)1/21s.\displaystyle\frac{r}{1-r^{2}}=\frac{(s-s^{\prime})^{1/2}(1-s^{\prime})^{1/2}}{1-s}.

We deduce that,

r(1r2)0.915451(1s)2(ss)1/2(1s)3/2(ss)1/2(1s)1/21s=0.4578968621s=0.83837237\frac{\ell r}{(1-r^{2})}\leq 0.915451...\frac{(1-s)}{2(s-s^{\prime})^{1/2}(1-s^{\prime})^{3/2}}\cdot\frac{(s-s^{\prime})^{1/2}(1-s^{\prime})^{1/2}}{1-s}=\frac{0.457896862...}{1-s^{\prime}}=0.83837237...

This computation shows that for sufficiently large nn, for any 0δ0.83837237(1r2)nr0\leq\delta\leq\frac{0.83837237(1-r^{2})}{nr}, h(s)0h(s)\leq 0. We now show that h(t)0h(t)\leq 0 for 1t<s-1\leq t<s. Note that

r(t):=(ts)+1s<ss1s=r.r(t):=\sqrt{\frac{(t-s^{\prime})^{+}}{1-s^{\prime}}}<\sqrt{\frac{s-s^{\prime}}{1-s^{\prime}}}=r.

If sts-t is of order greater than O(1/n)O(1/n), then r(t)<rδr(t)<r-\delta for large nn, and so h(t)0h(t)\leq 0 as the integrand is negative in this case. Therefore, we may assume that 1t<s-1\leq t<s and st=O(1/n)s-t=O(1/n). For such tt, replacing ss with tt in the above calculations shows that the negativity h(t)0h(t)\leq 0 is true for all 0δ0.83837237(1r(t)2)nr(t).0\leq\delta\leq\frac{0.83837237(1-r(t)^{2})}{nr(t)}. Since r(t)<rr(t)<r, 0.83837237(1r(t)2)nr(t)>0.83837237(1r2)nr\frac{0.83837237(1-r(t)^{2})}{nr(t)}>\frac{0.83837237(1-r^{2})}{nr}. Consequently, h(t)0h(t)\leq 0 for every 1ts-1\leq t\leq s whenever 0δ0.83837237(1r2)nr0\leq\delta\leq\frac{0.83837237(1-r^{2})}{nr}.

Similarly, one obtains the same conclusion when ε=1\varepsilon=1, that is s=t1,dα+1,α+1s^{\prime}=t^{\alpha+1,\alpha+1}_{1,d}. Therefore, our improvement to Levenshtein’s bound on M(n,θ)M(n,\theta) for large nn is by a factor of 1/e0.83837237=0.4324131/e^{0.83837237...}=0.432413... for any choice of angle 0<θ<θ0<\theta<\theta^{*}. As the error in our computations is less than 2×1052\times 10^{-5}, we deduce that we have an improvement by a factor of 0.43250.4325 for sufficiently large nn. ∎

5.3. Improving bound on sphere packing densities

In this subsection, we give our improvement to Cohn and Zhao’s [CZ14] bound on sphere packings. Recall that in the case of sphere packings, we let s=cos(θ)s=\cos(\theta), r=12(1s)r=\frac{1}{\sqrt{2(1-s)}}, and α=n32\alpha=\frac{n-3}{2}. By assumption, 13s12\frac{1}{3}\leq s\leq\frac{1}{2}. In this case, we define for each 0<δ=c1n0<\delta=\frac{c_{1}}{n} the function FF to be used in Proposition 3.8 to be

F(y):=χ(y):={1 for 0yr+δ,0otherwise.F(y):=\chi(y):=\begin{cases}1&\text{ for }0\leq y\leq r+\delta,\\ 0&\text{otherwise}.\end{cases}
Proof of Theorem 1.2.

As in the proof of Theorem 5.2, we consider g=gn,θD(dμn32,θ)g=g_{n,\theta}\in D(d\mu_{\frac{n-3}{2}},\theta). Note that by Proposition 3.8 applied to the function FF above and gg, we have for the function

H(T)=11g(x)μ(x;T,χ)𝑑xH(T)=\int_{-1}^{1}g(x)\mu(x;T,\chi)dx

of equation (33), the inequality

δn()(2(r+δ))n=MLev(n,θ)(2(r+δ))n\delta_{n}\leq\frac{\cal{L}(g)}{(2(r+\delta))^{n}}=\frac{M_{\textup{Lev}}(n,\theta)}{(2(r+\delta))^{n}}

if δ>0\delta>0 is chosen such that H(T)0H(T)\leq 0 for T1T\geq 1. We wish to maximize δ\delta under this negativity condition. Let n2000n\geq 2000 and ss be a root of the Jacobi polynomial as before. As in the proof of Theorem 1.1, we begin by considering that case where ε=0\varepsilon=0, that is, s=t1,dα+1,αs=t^{\alpha+1,\alpha}_{1,d}, and take gg for this ss. We show that H(1)0H(1)\leq 0. For the other TT, showing H(T)0H(T)\leq 0 follows similarly by using Lemma 4.3. Note that

11g(x)μ(x;χ)𝑑x=|s1g(x)μ(x;χ)𝑑x||1sg(x)μ(x;χ)𝑑x|.\displaystyle\int_{-1}^{1}g(x)\mu(x;\chi)dx=\left|\int_{s}^{1}g(x)\mu(x;\chi)dx\right|-\left|\int_{-1}^{s}g(x)\mu(x;\chi)dx\right|.

To show H(1)0H(1)\leq 0, it suffices to show that

(43) |1sg(x)μ(x;χ)𝑑x||s1g(x)μ(x;χ)𝑑x|.\left|\int_{-1}^{s}g(x)\mu(x;\chi)dx\right|\geq\left|\int_{s}^{1}g(x)\mu(x;\chi)dx\right|.

First, we give a lower bound for the left hand side of equation (43). As before, Proposition 5.1 allows us to write

g(x)=(x+1)(xs)dpdα+1,αdt(s)2(1+A(x))2,g(x)=(x+1)(x-s^{\prime})\frac{dp_{d}^{\alpha+1,\alpha}}{dt}(s^{\prime})^{2}(1+A(x))^{2},

where

|A(x)|eσ(x)1σ(x)1|A(x)|\leq\frac{e^{\sigma(x)}-1}{\sigma(x)}-1

with

σ(x):=|xs|(nss+1)1s2.\sigma(x):=\frac{|x-s|(ns-s+1)}{1-s^{2}}.

As before,

|1+A(x)|(2eσ(x)1σ(x))+.|1+A(x)|\geq\left(2-\frac{e^{\sigma(x)}-1}{\sigma(x)}\right)^{+}.

Note that if σ(x)<1.25643\sigma(x)<1.25643, ensuring that the right hand side of the bound on 1+A(x)1+A(x) is non-negative, then

|xs|1.25643(1s2)ns+1.|x-s|\leq\frac{1.25643(1-s^{2})}{ns+1}.

Let

λ:=1.25643(1s2)ns+1.\lambda:=\frac{1.25643(1-s^{2})}{ns+1}.

Let

r1:=1(1x2)(r+δ)2+x(r+δ).r_{1}:=\sqrt{1-(1-x^{2})(r+\delta)^{2}}+x(r+\delta).

Suppose that |xs|c2n|x-s|\leq\frac{c_{2}}{n} and δ=c1n\delta=\frac{c_{1}}{n} for some 0<c1<0.810<c_{1}<0.81, 0<c2<3.360<c_{2}<3.36. Note that for such a c1c_{1} and n2000n\geq 2000, the assumption s1/2s\leq 1/2 implies that r+δ1r+\delta\leq 1. By Proposition 4.4, we have μ(x;χ)=0\mu(x;\chi)=0 for r1r+δ.r_{1}\geq r+\delta. Otherwise,

μ(x;χ)=Cn(1+E)(1x2)n42(12(1x))n111x1+(x(1x2)r1(1x2)r2)2(r+δr1)\mu(x;\chi)=C_{n}(1+E)(1-x^{2})^{\frac{n-4}{2}}\left(\frac{1}{2(1-x)}\right)^{{n-1}}\frac{1}{\sqrt{1-x}}\sqrt{1+\left(x-\frac{(1-x^{2})r}{\sqrt{1-(1-x^{2})r^{2}}}\right)^{2}}\left(r+\delta-r_{1}\right)

for some constant Cn>0C_{n}>0 making μ(x;χ)\mu(x;\chi) a probability measure on [1,1][-1,1], and where |E|<(4c2+2c1+2)2n|E|<\frac{(4c_{2}+2c_{1}+2)^{2}}{n} for n2000n\geq 2000. In particular, for such c1,c2c_{1},c_{2}, we have |E|<292n|E|<\frac{292}{n}. By letting z:=sxz:=s-x and v:=(ns+1)z(1s2)v:=\frac{(ns+1)z}{(1-s^{2})}, Taylor expansion approximations imply that inequality (43) is satisfied for 13s12\frac{1}{3}\leq s\leq\frac{1}{2} and large nn if

(44) 01.25643v(2ev1v)2e3v(1+3v2c1)𝑑v02c13v(ev1v)2e3v(13v2c1)𝑑v.\int_{0}^{1.25643}v\left(2-\frac{e^{v}-1}{v}\right)^{2}e^{-3v}\left(1+\frac{3v}{2c_{1}}\right)dv\geq\int_{0}^{\frac{2c_{1}}{3}}v\left(\frac{e^{v}-1}{v}\right)^{2}e^{3v}\left(1-\frac{3v}{2c_{1}}\right)dv.

By a numerics similar to that done for spherical codes, one finds that the maximal such c1<1c_{1}<1 is 0.664134700.66413470.... Therefore, for every 13s12\frac{1}{3}\leq s\leq\frac{1}{2} and nn\rightarrow\infty, we have an improvement at least as good as

e0.6641347/r=exp(0.66413472(1s)).e^{-0.6641347/r}=\exp(-0.6641347\sqrt{2(1-s)}).

Note that for such ss, as nn\rightarrow\infty, this gives us an improvement of at most 0.51480.5148, a universal such improvement factor.

Returning to the case of general n2000n\geq 2000 and 13s12\frac{1}{3}\leq s\leq\frac{1}{2}, given such an nn we need to maximize c1<0.81c_{1}<0.81 such that

(1312n)01.25643v(2ev1v)2e3v(1+1.499vc117.31c1n)𝑑v\displaystyle\left(1-\frac{312}{n}\right)\int_{0}^{1.25643}v\left(2-\frac{e^{v}-1}{v}\right)^{2}e^{-3v}\left(1+\frac{1.499v}{c_{1}}-\frac{17.31}{c_{1}n}\right)dv
\displaystyle\geq (1+312n)00.667c1v(ev1v)2e3v(11.499vc1+17.31c1n)𝑑v.\displaystyle\left(1+\frac{312}{n}\right)\int_{0}^{0.667c_{1}}v\left(\frac{e^{v}-1}{v}\right)^{2}e^{3v}\left(1-\frac{1.499v}{c_{1}}+\frac{17.31}{c_{1}n}\right)dv.

By a numerical calculation with Sage, we obtain that the improvement factor for any 13s12\frac{1}{3}\leq s\leq\frac{1}{2} and any n2000n\geq 2000 is at least as good as

0.515+74/n.0.515+74/n.

On the other hand, if we fix ss such that ss is sufficiently close to s=cos(θ)s^{*}=\cos(\theta^{*}), then the same kind of calculations as above give us an asymptotic improvement constant of 0.43250.4325, the same as in the case of spherical codes. In fact, for n2000n\geq 2000, we have an improvement factor at least as good as

0.4325+51/n0.4325+51/n

over the combination of Cohn–Zhao [CZ14] and Levenshtein’s optimal polynomials [Lev79].

The case s=t1,dα+1,α+1s=t^{\alpha+1,\alpha+1}_{1,d} follows in exactly the same way. This completes the proof of our theorem. ∎

Remark 45.

We end this section by saying that our improvements above are based on a local understanding of Levenshtein’s optimal polynomials, and that there is a loss in our computations. By doing numerics, we may do computations without having to rely on such local approximations.

6. Numerics

In Theorem 1.2, we improved the sphere packing densities by a factor of 0.43250.4325 for sufficiently high dimensions. There is a loss in our estimates due to neglecting the contribution of Levenshtein’s polynomials away from their largest roots, as giving rigorous estimates is difficult. In this section, we numerically investigate the behavior of our constant improvement factors for sphere packing densities by considering those neglected terms in dimensions up to 130130. As we noted before in the introduction, in low dimensions, there are better bounds on sphere packing densities using semi-definite programming, and so the objective of this section is guessing the improvement over 0.43250.4325 in high dimensions.

Before our work, the best known upper bound on sphere packing densities in high dimensions was obtained using inequalities

δnsinn(θ/2)Lev(n,θ),\delta_{n}\leq\sin^{n}(\theta/2)\textup{Lev}(n,\theta),

where Lev(n,θ)\text{Lev}(n,\theta) is the linear programming bound using Levenshtein’s optimal polynomials [Lev79, eq.(3),(4)]. Note that Lev(n,θ)MLev(n,θ)\text{Lev}(n,\theta)\leq M_{\textup{Lev}}(n,\theta) and equality occurs when cos(θ)=t1,dα+1,α+ε.\cos(\theta)=t^{\alpha+1,\alpha+\varepsilon}_{1,d}. In this section, we apply Proposition 3.8 to Levenshtein’s optimal polynomials for various angles θ\theta (columns of Table 2) and obtain

δnαn(θ)sinn(θ/2)Lev(n,θ)\delta_{n}\leq\alpha_{n}(\theta)\sin^{n}(\theta/2)\text{Lev}(n,\theta)

with maximal rr (in Proposition 3.8), where αn(θ)\alpha_{n}(\theta) are the entries of the table. Note that in Table 2, the improvement factors appear to gradually become independent of θ\theta as nn enlarges. We conjecture that they tend to 1e\frac{1}{e} as nn\rightarrow\infty.

nn θ\theta 6161^{\circ} 6262^{\circ} 6363^{\circ} 6464^{\circ} 6565^{\circ} 6666^{\circ} 6767^{\circ} 6868^{\circ} 6969^{\circ} 7070^{\circ} 7171^{\circ} 7272^{\circ} 7373^{\circ} 7474^{\circ} 7575^{\circ} 7676^{\circ} 7777^{\circ} 7878^{\circ} 7979^{\circ} 8080^{\circ} 8181^{\circ} 8282^{\circ} 8383^{\circ} 8484^{\circ} 8585^{\circ} 8686^{\circ} 8787^{\circ} 8888^{\circ} 8989^{\circ} 9090^{\circ}
4 .942 .889 .839 .793 .750 .711 .674 .640 .608 .578 .550 .524 .500 .477 .456 .436 .417 .399 .392 .396 .400 .405 .409 .413 .416 .418 .420 .420 .420 .419
5 .928 .863 .803 .748 .698 .653 .611 .572 .537 .504 .474 .446 .420 .400 .374 .371 .377 .382 .387 .392 .396 .401 .404 .408 .410 .412 .412 .412 .411 .408
6 .915 .838 .768 .706 .650 .599 .553 .512 .474 .439 .408 .385 .389 .393 .397 .368 .374 .379 .384 .389 .393 .398 .401 .404 .406 .407 .407 .406 .404 .401
7 .901 .813 .735 .666 .605 .550 .501 .457 .418 .395 .373 .378 .383 .388 .391 .395 .371 .377 .382 .387 .391 .395 .398 .401 .403 .404 .403 .402 .400 .396
8 .888 .789 .704 .629 .563 .505 .454 .409 .394 .394 .393 .373 .378 .383 .387 .391 .369 .374 .380 .385 .389 .393 .396 .399 .400 .401 .401 .399 .396 .392
9 .874 .766 .673 .593 .524 .464 .411 .389 .391 .392 .392 .391 .373 .378 .383 .387 .391 .372 .378 .383 .387 .391 .394 .397 .398 .399 .398 .397 .394 .389
10 .862 .744 .644 .560 .488 .426 .382 .386 .389 .390 .391 .391 .389 .374 .379 .384 .388 .371 .376 .381 .385 .389 .393 .395 .397 .397 .397 .395 .391 .387
11 .849 .722 .617 .528 .454 .391 .378 .382 .386 .388 .390 .390 .389 .370 .376 .380 .385 .369 .374 .379 .384 .388 .391 .394 .395 .396 .395 .393 .390 .385
12 .836 .701 .590 .498 .422 .384 .387 .379 .383 .386 .388 .389 .389 .387 .372 .377 .382 .386 .373 .378 .383 .387 .390 .393 .394 .395 .394 .392 .388 .384
13 .824 .681 .565 .470 .393 .380 .384 .375 .380 .383 .386 .388 .388 .387 .385 .375 .380 .384 .371 .376 .381 .385 .389 .392 .393 .394 .393 .391 .387 .382
14 .811 .661 .540 .444 .387 .376 .380 .384 .377 .381 .384 .387 .388 .387 .386 .372 .377 .382 .370 .375 .380 .384 .388 .391 .392 .393 .392 .390 .386 .381
15 .799 .642 .517 .419 .386 .386 .377 .381 .384 .378 .382 .385 .387 .387 .386 .370 .375 .380 .384 .374 .379 .383 .387 .390 .391 .392 .391 .389 .385 .380
16 .788 .623 .495 .395 .385 .386 .384 .378 .382 .376 .380 .383 .386 .386 .386 .384 .373 .378 .382 .373 .378 .382 .386 .389 .391 .391 .390 .388 .385 .380
17 .776 .605 .474 .381 .384 .385 .385 .375 .380 .383 .378 .382 .384 .386 .386 .384 .371 .376 .381 .371 .377 .381 .385 .388 .390 .391 .390 .388 .384 .379
18 .764 .587 .453 .378 .382 .384 .385 .383 .377 .381 .376 .380 .383 .385 .385 .384 .382 .374 .379 .370 .376 .380 .384 .387 .389 .390 .389 .387 .383 .378
19 .753 .570 .434 .383 .380 .383 .384 .384 .375 .379 .373 .378 .382 .384 .385 .384 .382 .373 .378 .369 .375 .379 .383 .387 .389 .389 .389 .387 .383 .378
20 .742 .553 .415 .381 .377 .381 .383 .384 .382 .377 .381 .376 .380 .383 .385 .384 .383 .371 .376 .381 .374 .379 .383 .386 .388 .389 .388 .386 .382 .377
21 .731 .537 .397 .379 .382 .379 .382 .383 .383 .375 .379 .374 .379 .382 .384 .384 .383 .369 .375 .379 .373 .378 .382 .385 .388 .388 .388 .386 .382 .377
22 .720 .522 .383 .376 .380 .377 .381 .383 .383 .381 .377 .381 .377 .381 .383 .384 .383 .380 .373 .378 .372 .377 .381 .385 .387 .388 .388 .385 .382 .376
23 .709 .506 .383 .382 .375 .381 .379 .382 .383 .382 .375 .379 .376 .380 .382 .384 .383 .381 .372 .377 .371 .376 .381 .384 .387 .388 .387 .385 .381 .376
24 .699 .492 .382 .382 .376 .380 .377 .381 .382 .382 .373 .378 .374 .378 .381 .383 .383 .381 .371 .376 .370 .375 .380 .384 .386 .387 .387 .385 .381 .375
25 .688 .477 .381 .382 .381 .378 .375 .379 .382 .382 .380 .376 .372 .377 .381 .383 .383 .381 .370 .375 .369 .374 .379 .383 .386 .387 .387 .384 .380 .375
26 .678 .463 .379 .381 .381 .376 .380 .378 .381 .382 .381 .375 .379 .376 .380 .382 .383 .382 .379 .374 .369 .374 .379 .383 .385 .387 .386 .384 .380 .375
27 .668 .450 .377 .380 .381 .380 .378 .376 .380 .381 .381 .373 .377 .375 .379 .381 .382 .382 .379 .373 .378 .373 .378 .382 .385 .386 .386 .384 .380 .375
28 .658 .437 .380 .379 .381 .381 .376 .375 .378 .381 .381 .379 .376 .373 .378 .381 .382 .382 .379 .372 .377 .373 .372 .382 .384 .386 .386 .384 .380 .375
29 .648 .424 .379 .378 .380 .381 .375 .379 .377 .380 .381 .380 .375 .372 .377 .380 .382 .382 .380 .371 .376 .372 .377 .381 .384 .386 .385 .384 .380 .374
30 .639 .411 .377 .376 .379 .381 .379 .377 .376 .379 .381 .380 .373 .378 .376 .379 .381 .382 .380 .370 .375 .372 .377 .381 .384 .385 .385 .383 .379 .374
31 .629 .400 .379 .379 .378 .380 .380 .376 .374 .378 .380 .380 .372 .377 .374 .378 .381 .382 .380 .369 .374 .371 .376 .380 .383 .385 .385 .383 .379 .374
32 .620 .388 .380 .377 .377 .380 .380 .374 .378 .377 .380 .380 .378 .375 .373 .378 .380 .381 .380 .377 .374 .370 .376 .380 .383 .385 .385 .383 .379 .374
33 .611 .379 .380 .376 .375 .379 .380 .378 .377 .376 .379 .380 .379 .374 .372 .377 .380 .381 .380 .371 .373 .370 .375 .379 .383 .384 .385 .383 .379 .374
34 .602 .378 .380 .379 .378 .378 .380 .379 .376 .375 .378 .380 .379 .373 .371 .376 .379 .381 .380 .378 .372 .369 .375 .379 .382 .384 .384 .383 .379 .374
35 .593 .377 .379 .379 .377 .377 .379 .379 .375 .374 .378 .380 .379 .372 .376 .375 .379 .381 .380 .378 .371 .369 .374 .379 .382 .384 .384 .382 .379 .373
36 .584 .375 .379 .379 .376 .375 .378 .379 .373 .377 .377 .379 .379 .377 .376 .374 .378 .380 .380 .378 .371 .376 .374 .378 .382 .384 .384 .382 .379 .373
37 .575 .378 .378 .379 .374 .374 .378 .379 .378 .376 .376 .379 .379 .378 .375 .373 .377 .380 .380 .379 .370 .375 .373 .378 .381 .384 .384 .382 .379 .373
38 .567 .376 .376 .379 .378 .377 .377 .379 .378 .375 .375 .378 .379 .378 .374 .373 .377 .380 .380 .379 .369 .375 .373 .378 .381 .383 .384 .382 .378 .373
39 .559 .375 .375 .378 .379 .376 .376 .378 .378 .374 .374 .377 .379 .378 .373 .372 .376 .379 .380 .379 .376 .374 .372 .377 .381 .383 .384 .382 .378 .373
40 .550 .378 .377 .377 .379 .375 .375 .378 .379 .373 .373 .377 .379 .378 .372 .376 .376 .379 .380 .379 .376 .374 .372 .377 .381 .383 .383 .382 .378 .373
41 .542 .379 .376 .376 .378 .377 .377 .377 .378 .377 .376 .376 .378 .379 .376 .376 .375 .378 .380 .379 .376 .373 .372 .376 .380 .383 .383 .382 .378 .373
42 .534 .378 .375 .375 .378 .378 .376 .376 .378 .377 .375 .375 .378 .379 .377 .375 .374 .378 .380 .379 .376 .372 .371 .376 .380 .382 .383 .382 .378 .373
43 .526 .378 .378 .374 .377 .378 .375 .375 .378 .378 .374 .374 .378 .379 .377 .374 .374 .377 .379 .379 .377 .372 .371 .376 .380 .382 .383 .381 .378 .373
44 .518 .377 .378 .376 .377 .378 .374 .374 .377 .378 .373 .373 .377 .378 .377 .373 .373 .377 .379 .379 .377 .371 .370 .375 .379 .382 .383 .381 .378 .372
45 .510 .377 .378 .375 .376 .378 .377 .373 .377 .378 .372 .372 .376 .378 .378 .372 .372 .376 .379 .379 .377 .371 .370 .375 .379 .382 .383 .381 .378 .372
46 .503 .376 .378 .374 .375 .378 .377 .376 .376 .378 .376 .376 .376 .378 .378 .372 .371 .376 .379 .379 .377 .370 .370 .375 .379 .382 .382 .381 .378 .372
47 .495 .374 .377 .377 .374 .377 .377 .375 .376 .378 .377 .375 .375 .378 .378 .371 .371 .375 .378 .379 .377 .370 .369 .375 .379 .381 .382 .381 .378 .372
48 .488 .376 .377 .377 .376 .376 .377 .374 .375 .377 .377 .374 .374 .377 .378 .376 .375 .375 .378 .379 .378 .369 .369 .374 .378 .381 .382 .381 .378 .372
49 .481 .375 .376 .377 .375 .376 .377 .373 .374 .377 .377 .373 .374 .377 .378 .376 .374 .374 .378 .379 .378 .369 .374 .374 .378 .381 .382 .381 .377 .372
50 .474 .374 .375 .377 .374 .375 .377 .376 .373 .377 .377 .372 .373 .377 .378 .376 .374 .374 .377 .379 .378 .374 .374 .374 .378 .381 .382 .381 .377 .372
60 .408 .375 .376 .376 .374 .376 .374 .375 .376 .373 .374 .376 .376 .374 .375 .377 .376 .373 .374 .377 .378 .376 .370 .371 .376 .379 .381 .380 .377 .371
70 .376 .374 .376 .375 .375 .373 .375 .375 .372 .375 .375 .372 .375 .375 .373 .374 .376 .375 .373 .374 .377 .377 .374 .369 .374 .378 .380 .379 .376 .371
80 .375 .374 .374 .374 .374 .374 .374 .373 .375 .373 .375 .374 .372 .375 .375 .373 .375 .376 .373 .371 .375 .377 .375 .370 .372 .377 .379 .379 .376 .371
90 .373 .374 .374 .374 .373 .374 .373 .374 .372 .374 .373 .374 .374 .372 .375 .374 .372 .375 .375 .371 .373 .376 .376 .372 .371 .375 .378 .379 .376 .370
100 .373 .374 .373 .373 .374 .372 .374 .372 .374 .372 .374 .373 .374 .373 .373 .374 .371 .373 .375 .373 .371 .375 .376 .373 .369 .374 .378 .378 .376 .370
110 .372 .372 .374 .373 .372 .374 .372 .374 .372 .373 .372 .374 .372 .374 .372 .374 .374 .370 .374 .374 .371 .373 .375 .374 .371 .373 .377 .378 .375 .370
120 .373 .373 .373 .373 .373 .372 .373 .372 .373 .373 .373 .373 .373 .373 .373 .371 .374 .372 .373 .374 .372 .372 .375 .374 .370 .373 .376 .377 .375 .370
130 .373 .373 .372 .372 .373 .373 .372 .373 .373 .371 .373 .373 .373 .372 .373 .370 .373 .373 .371 .374 .373 .370 .374 .375 .371 .372 .376 .377 .375 .370
Table 2. Table of improvement factors αn(θ)\alpha_{n}(\theta)

References

  • [AJCH+20] Nima Afkhami-Jeddi, Henry Cohn, Thomas Hartman, David de Laat, and Amirhossein Tajdini. High-dimensional sphere packing and the modular bootstrap. J. High Energy Phys., (12):Paper No. 066, 44, 2020.
  • [AVZ00] E. Agrell, A. Vardy, and K. Zeger. Upper bounds for constant-weight codes. IEEE Trans. Inform. Theory, 46(7):2373–2395, 2000.
  • [BDB96] P. G. Boyvalenkov, D. P. Danev, and S. P. Bumova. Upper bounds on the minimum distance of spherical codes. IEEE Trans. Inform. Theory, 42(5):1576–1581, 1996.
  • [BM07] A. Barg and O. R. Musin. Codes in spherical caps. Adv. Math. Commun., 1(1):131–149, 2007.
  • [BV08] Christine Bachoc and Frank Vallentin. New upper bounds for kissing numbers from semidefinite programming. J. Amer. Math. Soc., 21(3):909–924, 2008.
  • [CdS22] Henry Cohn, David de Laat, and Andrew Salmon. Three-point bounds for sphere packing. arXiv e-prints, page arXiv:2206.15373, June 2022.
  • [CE03] H. Cohn and N. Elkies. New upper bounds on sphere packings. I. Ann. of Math. (2), 157(2):689–714, 2003.
  • [CKM+17] 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.
  • [CS99] J. H. Conway and N. J. A. Sloane. Sphere packings, lattices and groups, volume 290 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, New York, third edition, 1999. With additional contributions by E. Bannai, R. E. Borcherds, J. Leech, S. P. Norton, A. M. Odlyzko, R. A. Parker, L. Queen and B. B. Venkov.
  • [CZ14] H. Cohn and Y. Zhao. Sphere packing bounds via spherical codes. Duke Math. J., 163(10):1965–2002, 2014.
  • [dDV22] Matthew de Courcy-Ireland, Maria Dostert, and Maryna Viazovska. Six-dimensional sphere packing and linear programming. arXiv e-prints, page arXiv:2211.09044, November 2022.
  • [Del72] Ph. Delsarte. Bounds for unrestricted codes, by linear programming. Philips Res. Rep., 27:272–289, 1972.
  • [dLdOFV14] D. de Laat, F.M. de Oliveira Filho, and F. Vallentin. Upper bounds for packings of spheres of several radii. Forum Math. Sigma, 2:e23, 42, 2014.
  • [FT43] L. Fejes Tóth. Über die dichteste Kugellagerung. Math. Z., 48:676–684, 1943.
  • [Hal05] T. C. Hales. A proof of the Kepler conjecture. Ann. of Math. (2), 162(3):1065–1185, 2005.
  • [HMR19] Thomas Hartman, Dalimil Mazáč, and Leonardo Rastelli. Sphere packing and quantum gravity. J. High Energy Phys., (12):048, 66, 2019.
  • [KL78] G. A. Kabatjanskiĭ and V. I. Levenšteĭn. Bounds for packings on the sphere and in space. Problemy Peredači Informacii, 14(1):3–25, 1978.
  • [Kra07] I. Krasikov. An upper bound on Jacobi polynomials. J. Approx. Theory, 149(2):116–130, 2007.
  • [Ld22] Nando Leijenhorst and David de Laat. Solving clustered low-rank semidefinite programs arising from polynomial optimization. arXiv e-prints, page arXiv:2202.12077, February 2022.
  • [Lev79] V. I. Levenšteĭn. Boundaries for packings in nn-dimensional Euclidean space. Dokl. Akad. Nauk SSSR, 245(6):1299–1303, 1979.
  • [Lev98] V. I. Levenšteĭn. Universal bounds for codes and designs. In Handbook of coding theory, Vol. I, II, pages 499–648. North-Holland, Amsterdam, 1998.
  • [MdOF18] Fabrício Caluza Machado and Fernando Mário de Oliveira Filho. Improving the semidefinite programming bound for the kissing number by exploiting polynomial symmetry. Exp. Math., 27(3):362–369, 2018.
  • [MV10] Hans D. Mittelmann and Frank Vallentin. High-accuracy semidefinite programming bounds for kissing numbers. Experiment. Math., 19(2):175–179, 2010.
  • [Rog58] C. A. Rogers. The packing of equal spheres. Proc. London Math. Soc. (3), 8:609–620, 1958.
  • [Sar19] N. T. Sardari. The Siegel variance formula for quadratic forms. arXiv e-prints, page arXiv:1904.08041, Apr 2019.
  • [Sid74] V. M. Sidel’nikov. New estimates for the closest packing of spheres in nn-dimensional Euclidean space. Mat. Sb. (N.S.), 95(137):148–158, 160, 1974.
  • [Sti87] T. J. Stieltjes. Sur les racines de l’équation Xn=0X_{n}=0. Acta Math., 9:385–400, 1887.
  • [Sze39] G. Szegö. Orthogonal Polynomials. American Mathematical Society, New York, 1939. American Mathematical Society Colloquium Publications, v. 23.
  • [Via17] M. S. Viazovska. The sphere packing problem in dimension 8. Ann. of Math. (2), 185(3):991–1015, 2017.