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

The Cuboidal Lattices and their Lattice sums

Antony Burrows Centre for Theoretical Chemistry and Physics, The New Zealand Institute for Advanced Study (NZIAS), Massey University Albany, Private Bag 102904, Auckland 0745, New Zealand Shaun Cooper School of Natural and Computational Sciences
Massey University–Albany
Private Bag 102904, North Shore Mail Centre
Auckland, New Zealand.
E-mail: [email protected]
 and  Peter Schwerdtfeger Centre for Theoretical Chemistry and Physics, The New Zealand Institute for Advanced Study (NZIAS), Massey University Albany, Private Bag 102904, Auckland 0745, New Zealand
Abstract.

Lattice sums of cuboidal lattices, which connect the face-centered with the mean-centered and the body-centered cubic lattices through parameter dependent lattice vectors, are evaluated by decomposing them into two separate lattice sums related to a scaled cubic lattice and a scaled Madelung constant. Using theta functions we were able to derive fast converging series in terms of Bessel functions. Analytical continuations of these lattice sums are discussed in detail.

Key words and phrases:
axial centred cuboidal lattice, body centred cubic, cuboidal lattice, face centred cubic, hexagonal close packing, lattice sum, Madelung constant, mean centred cuboidal lattice, modified Bessel function, sphere packing, theta function.
2000 Mathematics Subject Classification:
Primary 11E45. Secondary 33E20, 40A25, 65B10

1. Introduction

Lattice sums have a long history in solid-state physics and discrete mathematics [5]. They connect lattices to observables such as the equation of state for a bulk system using interacting potentials between the lattice points (atoms or molecules) in three-dimensional space [12, 13, 18, 20]. Most notable cases for such interactions are the Lennard-Jones [14] and the Coulomb potential, leading in the latter case, for example, to the famous Madelung constant derived as early as in 1918 [17]. For such potentials the corresponding lattice sums become functions of quadratic forms 𝒊G𝒊\boldsymbol{i}^{\top}G\boldsymbol{i} with 𝒊3\boldsymbol{i}\in\mathbb{Z}^{3} [6], i.e. the expression 𝒊G𝒊\boldsymbol{i}^{\top}G\boldsymbol{i} is the quadratic form associated with the lattice \mathcal{L} (or simply, the associated quadratic form).

In the general case of a nn-dimensional lattice (𝒊n\boldsymbol{i}\in\mathbb{Z}^{n}), GG is a positive definite, real and symmetric (n×n)(n\times n) matrix called the Gram matrix of the lattice \mathcal{L}, defined by its basis (or lattice) vectors {𝒃i}\{\boldsymbol{b}_{i}\} through G=BBG=BB^{\top}. B=(𝒃1,,𝒃n)B=(\boldsymbol{b}_{1},...,\boldsymbol{b}_{n})^{\top} is called the generator matrix (BB not necessarily positive definite). Lattice sums represent often conditionally convergent series [4], and the theory of converting them into fast converging series has become an intense research field over the past 50 years [5].

Concerning the Gram matrix GG or generator matrix we introduce a few important definitions required in this work [9]. Two generator matrices B1B_{1} and B2B_{2} are equivalent if B2=cUB1𝒪B_{2}=cUB_{1}\mathcal{O}, cc a non-zero real number, 𝒪\mathcal{O} a real orthogonal matrix (𝒪𝒪=1\mathcal{O}\mathcal{O}^{\top}=1) with det(𝒪)=±1\textrm{det}(\mathcal{O})=\pm 1 describing rotation, reflection or rotoreflection of the lattice, and UU a matrix containing integers with detU=1\textrm{det}U=1 describing for example permutations of the basis vectors. Given two equivalent generator matrices B1B_{1} and B2B_{2}, the corresponding Gram matrices are related by

G2=B2B2=cUB1𝒪(cUB1𝒪)=c2UB1𝒪𝒪B1U=c2UG1U.G_{2}=B_{2}B_{2}^{\top}=cUB_{1}\mathcal{O}\left(cUB_{1}\mathcal{O}\right)^{\top}=c^{2}UB_{1}\mathcal{O}\mathcal{O}^{\top}B_{1}^{\top}U^{\top}=c^{2}UG_{1}U^{\top}.

The minimum distance dmind_{\textrm{min}} in a lattice \mathcal{L} is defined by

dmin()=min{d(𝒗1;𝒗2)|𝒗1,𝒗2;𝒗1𝒗2}d_{\textrm{min}}(\mathcal{L})=\textrm{min}\{d(\boldsymbol{v}_{1};\boldsymbol{v}_{2})|\boldsymbol{v}_{1},\boldsymbol{v}_{2}\in\mathcal{L};\boldsymbol{v}_{1}\neq\boldsymbol{v}_{2}\}

where d(𝒗1;𝒗2)=|𝒗1𝒗2|d(\boldsymbol{v}_{1};\boldsymbol{v}_{2})=|\boldsymbol{v}_{1}-\boldsymbol{v}_{2}| is the Euclidean distance. In terms of the Gram matrix this is equivalent to

dmin=min{+𝒊G𝒊|𝒊3\(0,0,0)}.d_{\textrm{min}}=\textrm{min}\{+\sqrt{\boldsymbol{i}^{\top}G\boldsymbol{i}}\leavevmode\nobreak\ |\leavevmode\nobreak\ \boldsymbol{i}\in\mathbb{Z}^{3}\backslash(0,0,0)^{\top}\}.

The minimal norm is related to the minimum distance by μ=dmin2\mu=d_{\textrm{min}}^{2}. Dividing GG by μ\mu assures that dmin=1d_{\textrm{min}}=1 used in most lattice sum applications [6]. For dense sphere packings the radius of a sphere ρ\rho is simply ρ=dmin/2\rho=d_{\textrm{min}}/2. The packing density ΔL\Delta_{L} and the center density δL\delta_{L} of a three-dimensional lattice are given by

Δ=4π3δ=4π3ρ3vol()=4π3ρ3|det(B)|=4π3ρ3det(G).\Delta_{\mathcal{L}}=\frac{4\pi}{3}\delta_{\mathcal{L}}=\frac{4\pi}{3}\frac{\rho^{3}}{\textrm{vol}(\mathcal{L})}=\frac{4\pi}{3}\frac{\rho^{3}}{|\textrm{det}(B)|}=\frac{4\pi}{3}\frac{\rho^{3}}{\sqrt{\textrm{det}(G)}}.

The kissing number for dense sphere packings is defined by

kiss()=#{𝒗||𝒗|=dmin()}.\textrm{kiss}(\mathcal{L})=\#\{\boldsymbol{v}\in\mathcal{L}\leavevmode\nobreak\ |\leavevmode\nobreak\ |\boldsymbol{v}|=d_{\textrm{min}}(\mathcal{L})\}.

In this work we discuss cuboidal lattices and their lattice sums. We first present the characteristics of cuboidal lattices (A)\mathcal{L}(A) dependent on a single parameter AA. In what follows we decompose the corresponding lattice sum into two lattice sums, where one is related to a scaled cubic lattice and the other to a scaled Madelung constant. We evaluate these lattice sums in two ways using theta functions. We discuss these lattice sums including their analytical continuations and provide a more complete analysis for the lattice sum difference between f.c.c. (face centred cubic) and h.c.p. (hexagonal close packing).

2. The cuboidal lattices

Following Conway and Sloane [7, Sec. 3] we consider the lattice generated by the vectors (±u,±v,0)and(0,±v,±v)(\pm u,\pm v,0)^{\top}\quad\text{and}\quad(0,\pm v,\pm v)^{\top}, where uu and vv are non-zero real numbers. To make it specific, take the basis vectors

𝒃1=(u,v,0),𝒃2=(u,0,v),𝒃3=(0,v,v).\boldsymbol{b}_{1}=(u,v,0)^{\top},\quad\boldsymbol{b}_{2}=(u,0,v)^{\top},\quad\boldsymbol{b}_{3}=(0,v,v)^{\top}. (2.1)

Then the generator matrix BB is given by

B=(uv0u0v0vv)B=\begin{pmatrix}u&v&0\\ u&0&v\\ 0&v&v\end{pmatrix}

which has determinant 2uv2-2uv^{2}. The Gram matrix is

G=BB=(u2+v2u2v2u2u2+v2v2v2v22v2)=v2(1+AA1A1+A1112)G=B\,B^{\top}=\begin{pmatrix}u^{2}+v^{2}&u^{2}&v^{2}\\ u^{2}&u^{2}+v^{2}&v^{2}\\ v^{2}&v^{2}&2v^{2}\end{pmatrix}=v^{2}\begin{pmatrix}1+A&A&1\\ A&1+A&1\\ 1&1&2\end{pmatrix} (2.2)

where A=u2/v2A=u^{2}/v^{2} and GG is positive definite for A>0A>0. Conway and Sloane [7] use σ=u/v\sigma=u/v, so A=σ2A=\sigma^{2}. The most important cases, in decreasing numerical order, are:

  1. (1)

    A=1A=1: the face-centred cubic (f.c.c.) lattice;

  2. (2)

    A=1/2A=1/\sqrt{2}: the mean centred-cuboidal (m.c.c.) lattice;

  3. (3)

    A=1/2A=1/2: the body-centred cubic (b.c.c.) lattice;

  4. (4)

    A=1/3A=1/3: the axial centred-cuboidal (a.c.c.) lattice.

The f.c.c. and b.c.c. lattices are well known. The corresponding Gram matrices for the f.c.c. and b.c.c lattices are identical to the ones shown in our previous paper on lattice sums [6]. The m.c.c. and a.c.c. lattices occur in [7] and [8]. The m.c.c. lattice is the densest isodual lattice in three-dimensional space.

The quadratic form associated with the lattice is

g(i,j,k)\displaystyle g(i,j,k) =(i,j,k)G(i,j,k)\displaystyle=(i,j,k)\,G(i,j,k)^{\top}
=(u2+v2)i2+(u2+v2)j2+2v2k2+2u2ij+2v2ik+2v2jk\displaystyle=(u^{2}+v^{2})i^{2}+(u^{2}+v^{2})j^{2}+2v^{2}k^{2}+2u^{2}ij+2v^{2}ik+2v^{2}jk
=u2(i2+j2)+v2(j+k)2+v2(i+k)2\displaystyle=u^{2}(i^{2}+j^{2})+v^{2}(j+k)^{2}+v^{2}(i+k)^{2}
=v2(A(i+j)2+(j+k)2+(i+k)2).\displaystyle=v^{2}\left(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\right). (2.3)

It is easy to check that

mini,j,k(i,j,k)(0,0,0){A(i+j)2+(j+k)2+(i+k)2}={4Aif 0<A<1/3,A+1if 1/3A1,2if A>1.\min_{i,j,k\in\mathbb{Z}\atop(i,j,k)\neq(0,0,0)}\{A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\}=\begin{cases}4A&\text{if $0<A<1/3$,}\\ A+1&\text{if $1/3\leq A\leq 1$,}\\ 2&\text{if $A>1$.}\end{cases} (2.4)

It follows from (2.3) and (2.4) that the minimum distance is given by

dmin=mini,j,k(i,j,k)(0,0,0)g(i,j,k)={2vAif 0<A<1/3,vA+1if 1/3A1,v2if A>1.d_{\textrm{min}}=\min_{i,j,k\in\mathbb{Z}\atop(i,j,k)\neq(0,0,0)}\,\sqrt{g(i,j,k)}=\begin{cases}2v\sqrt{A}&\text{if $0<A<1/3$,}\\ v\sqrt{A+1}&\text{if $1/3\leq A\leq 1$,}\\ v\sqrt{2}&\text{if $A>1$.}\end{cases} (2.5)

We rescale to make the minimum distance 11 by defining

g(A;i,j,k)\displaystyle g(A;i,j,k) =g(i,j,k)(dmin)2\displaystyle=\frac{g(i,j,k)}{\left(d_{\textrm{min}}\right)^{2}}
={14A(A(i+j)2+(j+k)2+(i+k)2)if 0<A<1/3,1A+1(A(i+j)2+(j+k)2+(i+k)2)if 1/3A1,12(A(i+j)2+(j+k)2+(i+k)2)if A>1.\displaystyle=\begin{cases}\displaystyle{\frac{1}{4A}\left(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\right)}&\text{if $0<A<1/3$,}\\ \\ \displaystyle{\frac{1}{A+1}\left(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\right)}&\text{if $1/3\leq A\leq 1$,}\\ \\ \displaystyle{\frac{1}{2}\left(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\right)}&\text{if $A>1$.}\end{cases}

The examples we are interested are in (f.c.c., m.c.c., b.c.c., a.c.c.) all satisfy 1/3A11/3\leq A\leq 1. Because of its practical interest, this is the only case we will study, and from here on (unless otherwise mentioned) we will always assume 1/3A11/3\leq A\leq 1 in which case we have

g(A;i,j,k)=1A+1(A(i+j)2+(j+k)2+(i+k)2),g(A;i,j,k)=\frac{1}{A+1}\left(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}\right), (2.6)

corresponding to the rescaled Gram matrix

G(A):=1(dmin)2G=1A+1(1+AA1A1+A1112).G(A):=\frac{1}{\left(d_{\textrm{min}}\right)^{2}}\;G=\frac{1}{A+1}\begin{pmatrix}1+A&A&1\\ A&1+A&1\\ 1&1&2\end{pmatrix}. (2.7)

The packing density is calculated as

Δ=4π3ρ3det(G)\Delta_{\mathcal{L}}=\frac{4\pi}{3}\frac{\rho^{3}}{\sqrt{\textrm{det}(G)}}

where ρ=dmin/2\rho=d_{\textrm{min}}/2 and

det(G)=(det(B))2=4u2v4=4Av6.\det(G)=\left(\det(B)\right)^{2}=4u^{2}v^{4}=4Av^{6}.

It follows that

Δ=π12A(dminv)3.\Delta_{\mathcal{L}}=\frac{\pi}{12\,\sqrt{A}}\left(\frac{d_{\textrm{min}}}{v}\right)^{3}.

On using the formula for dmind_{\textrm{min}} in (2.5) we obtain the formula for the packing density, given by

Δ={2πA3if 0<A<1/3,π12(A+1)3Aif 1/3A1,π62Aif A>1.\Delta_{\mathcal{L}}=\begin{cases}\displaystyle{\frac{2\pi A}{3}}&\text{if $0<A<1/3$,}\\ \\ \displaystyle{\frac{\pi}{12}\sqrt{\frac{(A+1)^{3}}{A}}}&\text{if $1/3\leq A\leq 1$,}\\ \\ \displaystyle{\frac{\pi}{6}\sqrt{\frac{2}{A}}}&\text{if $A>1$.}\end{cases}

Figure 1 shows a graph of the packing density as a function of the parameter AA. Further information is recorded in Table 1. In the main region of interest 1/3A11/3\leq A\leq 1, we have

Δ=π12(A+1)3A\Delta_{\mathcal{L}}=\frac{\pi}{12}\sqrt{\frac{(A+1)^{3}}{A}} (2.8)

and so

dΔdA=π12(A12)A+1A3.\frac{\mathrm{d}\Delta_{\mathcal{L}}}{\mathrm{d}A}=\frac{\pi}{12}\left(A-\frac{1}{2}\right)\sqrt{\frac{A+1}{A^{3}}}. (2.9)

It follows that on the interval 1/3A11/3\leq A\leq 1, the packing density has a maximum of π2/60.74048\pi\sqrt{2}/6\approx 0.74048 at A=1A=1 corresponding to f.c.c., and a minimum of π3/80.68017\pi\sqrt{3}/8\approx 0.68017 at A=1/2A=1/2 corresponding to b.c.c.

It is also interesting to note that as AA\rightarrow\infty the limiting case of the lattice is the two-dimensional square close packing with minimal distance 11 and kissing number 44; while in the other extreme case the limit as A0A\rightarrow 0 gives the one-dimensional lattice with minimal distance 11 and kissing number 22. These cases are briefly analysed in Appendix B.

Table 1. Kissing number, packing density Δ\Delta_{\mathcal{L}} and integer combinations 𝒊n\boldsymbol{i}_{n} for the lattice associated with the Gram matrix GG defined in (2.2). The values in the table depend only on AA and are independent of vv.
Region AA kiss()\textrm{kiss}(\mathcal{L}) Δ\Delta_{\mathcal{L}} integer combinationsa
I (0,13)(0,\frac{1}{3}) 2 2πA3\frac{2\pi A}{3} 𝒊1,𝒊2\boldsymbol{i}_{1},\boldsymbol{i}_{2}
a.c.c. 13\frac{1}{3} 10 2π9\frac{2\pi}{9} 𝒊1,,𝒊10\boldsymbol{i}_{1},...,\boldsymbol{i}_{10}
II (13,1)(\frac{1}{3},1) 8 π12(A+1)3A\frac{\pi}{12}\sqrt{\frac{(A+1)^{3}}{A}} 𝒊3,,𝒊10\boldsymbol{i}_{3},...,\boldsymbol{i}_{10}
f.c.c. 11 12 π26\frac{\pi\,\sqrt{2}}{6} 𝒊3,,𝒊14\boldsymbol{i}_{3},...,\boldsymbol{i}_{14}
III (1,)(1,\infty) 4 π62A\frac{\pi}{6}\sqrt{\frac{2}{A}} 𝒊11,,𝒊14\boldsymbol{i}_{11},...,\boldsymbol{i}_{14}

a The integer combinations 𝒊\boldsymbol{i} which determine dmind_{\textrm{min}} in (2.5) for the different regions are as follows: 𝒊1I=(1,1,1)\boldsymbol{i}_{1}^{\textrm{I}}=(-1,-1,1), 𝒊2I=(1,1,1)\boldsymbol{i}_{2}^{\textrm{I}}=(1,1,-1), 𝒊3II=(1,0,0)\boldsymbol{i}_{3}^{\textrm{II}}=(-1,0,0), 𝒊4II=(1,0,1)\boldsymbol{i}_{4}^{\textrm{II}}=(-1,0,1), 𝒊5II=(0,1,0)\boldsymbol{i}_{5}^{\textrm{II}}=(0,-1,0), 𝒊6II=(0,1,1)\boldsymbol{i}_{6}^{\textrm{II}}=(0,-1,1), 𝒊7II=(0,1,1)\boldsymbol{i}_{7}^{\textrm{II}}=(0,1,-1), 𝒊8II=(0,1,0)\boldsymbol{i}_{8}^{\textrm{II}}=(0,1,0), 𝒊9II=(1,0,1)\boldsymbol{i}_{9}^{\textrm{II}}=(1,0,-1), 𝒊10II=(1,0,0)\boldsymbol{i}_{10}^{\textrm{II}}=(1,0,0), 𝒊11III=(1,1,0)\boldsymbol{i}_{11}^{\textrm{III}}=(-1,1,0), 𝒊12III=(0,0,1)\boldsymbol{i}_{12}^{\textrm{III}}=(0,0,-1), 𝒊13III=(0,0,1)\boldsymbol{i}_{13}^{\textrm{III}}=(0,0,1), 𝒊14III=(1,1,0)\boldsymbol{i}_{14}^{\textrm{III}}=(1,-1,0).

Refer to caption
Figure 1. Graph of the packing density Δ\Delta_{\mathcal{L}} versus AA. The regions I, II and III are according to the different kissing numbers. Explicit formulas are given in Table 1. The location of the f.c.c., m.c.c., b.c.c. and a.c.c. lattices are indicated on the graph.

Given a positive definite quadratic form g(i,j,k)g(i,j,k), the corresponding theta series is defined for |q|<1|q|<1 by

θg(q)=i=j=k=qg(i,j,k).\theta_{g}(q)=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{g(i,j,k)}.

For the quadratic form in (2.6) the theta series is

θ(A;q)=i=j=k=q(A(i+j)2+(j+k)2+(i+k)2)/(A+1) where 1/3A1.\theta(A;q)=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2})/(A+1)}\quad\text{ where $1/3\leq A\leq 1$.}

The first few terms in the theta series for f.c.c., m.c.c., b.c.c. and a.c.c. as far as q9q^{9} are given respectively by

θ(1;q)\displaystyle\theta(1;q) =1+12q+6q2+24q3+12q4+24q5+8q6+48q7+6q8+36q9+,\displaystyle=1+12q+6q^{2}+24q^{3}+12q^{4}+24q^{5}+8q^{6}+48q^{7}+6q^{8}+36q^{9}+\cdots,
θ(12;q)\displaystyle\theta(\frac{1}{\sqrt{2}};q) =1+8q+4q422+2q424+4q842+8q22+16q42+9\displaystyle=1+8q+4q^{4-2\sqrt{2}}+2q^{4\sqrt{2}-4}+4q^{8-4\sqrt{2}}+8q^{2\sqrt{2}}+16q^{-4\sqrt{2}+9}
+8q4+8q827+4q1682+8q82+17+8q20102+8q42+12\displaystyle\quad+8q^{4}+8q^{8\sqrt{2}-7}+4q^{16-8\sqrt{2}}+8q^{-8\sqrt{2}+17}+8q^{20-10\sqrt{2}}+8q^{-4\sqrt{2}+12}
+2q16216+16q42+1+16q62+16+8q14212+16q122+25\displaystyle\quad+2q^{16\sqrt{2}-16}+16q^{4\sqrt{2}+1}+16q^{-6\sqrt{2}+16}+8q^{14\sqrt{2}-12}+16q^{-12\sqrt{2}+25}
+8q8+122+8q9+,\displaystyle\quad+8q^{-8+12\sqrt{2}}+8q^{9}+\cdots,
θ(12;q)\displaystyle\theta(\frac{1}{2};q) =1+8q+6q4/3+12q8/3+8q4+24q11/3+6q16/3+24q19/3+24q20/3\displaystyle=1+8q+6q^{4/3}+12q^{8/3}+8q^{4}+24q^{11/3}+6q^{16/3}+24q^{19/3}+24q^{20/3}
+24q8+32q9+,\displaystyle\quad+24q^{8}+32q^{9}+\cdots,
θ(13;q)\displaystyle\theta(\frac{1}{3};q) =1+10q+4q3/2+8q5/2+12q3+26q4+8q11/2+20q6+32q7\displaystyle=1+10q+4q^{3/2}+8q^{5/2}+12q^{3}+26q^{4}+8q^{11/2}+20q^{6}+32q^{7}
+8q15/2+16q17/2+10q9+.\displaystyle\quad+8q^{15/2}+16q^{17/2}+10q^{9}+\cdots.

Since the quadratic form g(A;i,j,k)g(A;i,j,k) has been normalised to make the minimum distance 11, the kissing number occurs in each theta series as the coefficient of qq. That is, we have kiss(f.c.c.)=12\textrm{kiss(f.c.c.)}=12, kiss(m.c.c.)=8\textrm{kiss(m.c.c.)}=8, kiss(b.c.c.)=8\textrm{kiss(b.c.c.)}=8 and kiss(a.c.c.)=10\textrm{kiss(a.c.c.)}=10.

We introduce the following lattice sum important in solid state theory [5],

L(A;s)=i,j,k(1g(A;i,j,k))s=i,j,k(A+1A(i+j)2+(j+k)2+(i+k)2)sL(A;s)={\sum_{i,j,k}}^{\prime}\left(\frac{1}{g(A;i,j,k)}\right)^{s}={\sum_{i,j,k}}^{\prime}\left(\frac{A+1}{A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}}\right)^{s} (2.10)

where 1/3A11/3\leq A\leq 1. Here and throughout this work, a prime on the summation symbol will denote that the sum ranges over all integer values except for the term when all of the summation indices are simultaneously zero. Thus, the sums in (2.10) are over all integer values of ii, jj and kk except for the term (i,j,k)=(0,0,0)(i,j,k)=(0,0,0) which is omitted. This lattice sum smoothly connects four different well known lattices, i.e., when A=1A=1, 1/21/\sqrt{2}, 1/21/2 or 1/31/3 we obtain the expressions for the lattices f.c.c, m.c.c., b.c.c. and a.c.c. respectively (face-centred cubic, mean centred-cuboidal, body-centred cubic, and axial centred cuboidal). In these cases, we also write

L3FCC(s)\displaystyle L_{3}^{\text{FCC}}(s) =L(1;s),\displaystyle=L(1;s),
L3MCC(s)\displaystyle L_{3}^{\text{MCC}}(s) =L(1/2;s),\displaystyle=L(1/\sqrt{2};s),
L3BCC(s)\displaystyle L_{3}^{\text{BCC}}(s) =L(1/2;s),\displaystyle=L(1/2;s),
andL3ACC(s)\displaystyle\text{and}\quad L_{3}^{\text{ACC}}(s) =L(1/3;s).\displaystyle=L(1/3;s).

We conclude this section by reconciling the Gram matrix GG in (2.2) with two matrices given by Conway and Sloane [7]. Let

U1=(100101010)andU2=(111100010)U_{1}=\begin{pmatrix}1&0&0\\ -1&0&1\\ 0&-1&0\end{pmatrix}\quad\text{and}\quad U_{2}=\begin{pmatrix}1&1&-1\\ 1&0&0\\ 0&1&0\end{pmatrix}

and consider the equivalent matrices G1G_{1} and G2G_{2} defined by

G1=U1GU1=(u2+v2u2u2u2u2+v2u2v2u2u2v2u2+v2)G_{1}=U_{1}\,G\,U_{1}^{\top}=\begin{pmatrix}u^{2}+v^{2}&-u^{2}&-u^{2}\\ -u^{2}&u^{2}+v^{2}&u^{2}-v^{2}\\ -u^{2}&u^{2}-v^{2}&u^{2}+v^{2}\end{pmatrix} (2.11)

and

G2=U2GU2=(4u22u22u22u2u2+v2u22u2u2u2+v2).G_{2}=U_{2}\,G\,U_{2}^{\top}=\begin{pmatrix}4u^{2}&2u^{2}&2u^{2}\\ 2u^{2}&u^{2}+v^{2}&u^{2}\\ 2u^{2}&u^{2}&u^{2}+v^{2}\end{pmatrix}. (2.12)

When u=1/2u=1/\sqrt{2} and v=1/24v=1/\sqrt[4]{2}, the matrix G1G_{1} in (2.11) is the Gram matrix for the m.c.c lattice given by Conway and Sloane [7, (10)]. Moreover, when u=1/3u=\sqrt{1/3} and v=2/3v=\sqrt{2/3}, the matrix G1G_{1} leads to another well-known quadratic form for the b.c.c. lattice, e.g., see [6, (8b)]. When u=1u=1, v=3v=\sqrt{3}, the matrix G2G_{2} in (2.12) is the Gram matrix for the a.c.c. lattice given in [7, p. 378]. Since detU1=detU2=1\det U_{1}=\det U_{2}=1 it follows that

detG1=detG2=detG=(detB)2=4u2v4=4v6A.\det G_{1}=\det G_{2}=\det G=(\det B)^{2}=4u^{2}v^{4}=4v^{6}A.

The corresponding quadratic forms g1g_{1} and g2g_{2} are defined by

g1(i,j,k)\displaystyle g_{1}(i,j,k) =(i,j,k)G1(i,j,k)\displaystyle=(i,j,k)\,G_{1}(i,j,k)^{\top}
=(u2+v2)i2+(u2+v2)j2+(u2+v2)k22u2ij2u2ik+2(u2v2)jk\displaystyle=(u^{2}+v^{2})i^{2}+(u^{2}+v^{2})j^{2}+(u^{2}+v^{2})k^{2}-2u^{2}ij-2u^{2}ik+2(u^{2}-v^{2})jk
and
g2(i,j,k)\displaystyle g_{2}(i,j,k) =(i,j,k)G2(i,j,k)\displaystyle=(i,j,k)\,G_{2}(i,j,k)^{\top}
=4u2i2+(u2+v2)j2+(u2+v2)k2+4u2ij+4u2ik+2u2jk.\displaystyle=4u^{2}i^{2}+(u^{2}+v^{2})j^{2}+(u^{2}+v^{2})k^{2}+4u^{2}ij+4u^{2}ik+2u^{2}jk.

They are related to the quadratic form gg in (2.3) by

g1(i,j,k)=g((i,j,k)U1)=g(ij,k,j)g_{1}(i,j,k)=g\left((i,j,k)U_{1}\right)=g(i-j,-k,j) (2.13)

and

g2(i,j,k)=g((i,j,k)U2)=g(i+j,i+k,i).g_{2}(i,j,k)=g\left((i,j,k)U_{2}\right)=g(i+j,i+k,-i).

3. A minimum property of the lattice sum L(A;s)L(A;s)

In the previous section — see (2.8) and (2.9) — it was noted that on the interval 1/3A11/3\leq A\leq 1, the packing density function Δ\Delta_{\mathcal{L}} has a minimum value when A=1/2A=1/2. The next result shows that provided s>3/2s>3/2, the corresponding lattice sum L(A;s)L(A;s) also attains a minimum at the same value A=1/2A=1/2.

Theorem 3.1.

Let L(A;s)L(A;s) be the lattice defined by (2.10), that is,

L(A;s)=i,j,k(1g(A;i,j,k))s=i,j,k(A+1A(i+j)2+(j+k)2+(i+k)2)sL(A;s)={\sum_{i,j,k}}^{\prime}\left(\frac{1}{g(A;i,j,k)}\right)^{s}={\sum_{i,j,k}}^{\prime}\left(\frac{A+1}{A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}}\right)^{s}

where s>3/2s>3/2 and 1/3A11/3\leq A\leq 1. Then

AL(A;s)|A=1/2=0and2A2L(A;s)|A=1/2>0.\left.\frac{\partial}{\partial A}L(A;s)\right|_{A=1/2}=0\quad\text{and}\quad\left.\frac{\partial^{2}}{\partial A^{2}}L(A;s)\right|_{A=1/2}>0.
Proof.

By definition we have

L(A;s)=I,J,K(1g(A;I,J,K))sL(A;s)={\sum_{I,J,K}}^{\prime}\left(\frac{1}{g(A;I,J,K)}\right)^{s}

where

g(A;I,J,K)=1A+1(A(I+J)2+(J+K)2+(I+K)2).g(A;I,J,K)=\frac{1}{A+1}\left(A(I+J)^{2}+(J+K)^{2}+(I+K)^{2}\right).

Now make the change of variables given by (2.13), namely

(I,J,K)=(ij,k,j).(I,J,K)=(i-j,-k,j).

This is a bijection since

(i,j,k)=(I+K,K,J),(i,j,k)=(I+K,K,-J),

and it follows that

L(A;s)\displaystyle L(A;s) =i,j,k(1g(A;ij,k,j))s\displaystyle={\sum_{i,j,k}}^{\prime}\left(\frac{1}{g(A;i-j,-k,j)}\right)^{s}
=i,j,k1(i2+j2+k22(ij+ik)(AA+1)+2jk(A1A+1))s.\displaystyle={\sum_{i,j,k}}^{\prime}\frac{1}{\left(i^{2}+j^{2}+k^{2}-2(ij+ik)\left(\frac{A}{A+1}\right)+2jk\left(\frac{A-1}{A+1}\right)\right)^{s}}.

By direct calculation, the derivative is given by

AL(A;s)=2s(A+1)2i,j,kij+ik2jk(i2+j2+k22(ij+ik)(AA+1)+2jk(A1A+1))s+1.\frac{\partial}{\partial A}L(A;s)=\frac{2s}{(A+1)^{2}}{\sum_{i,j,k}}^{\prime}\frac{ij+ik-2jk}{\left(i^{2}+j^{2}+k^{2}-2(ij+ik)\left(\frac{A}{A+1}\right)+2jk\left(\frac{A-1}{A+1}\right)\right)^{s+1}}. (3.1)

Setting A=1/2A=1/2 gives

AL(A;s)|A=1/2=8s9i,j,kij+ik2jk(i2+j2+k223(ij+ik+jk))s+1.\left.\frac{\partial}{\partial A}L(A;s)\right|_{A=1/2}=\frac{8s}{9}\;{\sum_{i,j,k}}^{\prime}\frac{ij+ik-2jk}{\left(i^{2}+j^{2}+k^{2}-\frac{2}{3}(ij+ik+jk)\right)^{s+1}}. (3.2)

Switching ii and jj gives

AL(A;s)|A=1/2=8s9i,j,kij+jk2ik(i2+j2+k223(ij+ik+jk))s+1,\left.\frac{\partial}{\partial A}L(A;s)\right|_{A=1/2}=\frac{8s}{9}\;{\sum_{i,j,k}}^{\prime}\frac{ij+jk-2ik}{\left(i^{2}+j^{2}+k^{2}-\frac{2}{3}(ij+ik+jk)\right)^{s+1}}, (3.3)

while switching ii and kk in (3.2) gives

AL(A;s)|A=1/2=8s9i,j,kjk+ik2ij(i2+j2+k223(ij+ik+jk))s+1.\left.\frac{\partial}{\partial A}L(A;s)\right|_{A=1/2}=\frac{8s}{9}\;{\sum_{i,j,k}}^{\prime}\frac{jk+ik-2ij}{\left(i^{2}+j^{2}+k^{2}-\frac{2}{3}(ij+ik+jk)\right)^{s+1}}. (3.4)

On adding (3.2), (3.3) and (3.4) and noting that

(ij+ik2jk)+(ij+jk2ik)+(jk+ik2ij)=0(ij+ik-2jk)+(ij+jk-2ik)+(jk+ik-2ij)=0

it follows that

AL(A;s)|A=1/2=0.\left.\frac{\partial}{\partial A}L(A;s)\right|_{A=1/2}=0.

Next, taking the derivative of (3.1) gives

2A2L(A;s)\displaystyle\frac{\partial^{2}}{\partial A^{2}}L(A;s) =4s(A+1)3i,j,kij+ik2jk(i2+j2+k22(ij+ik)(AA+1)+2jk(A1A+1))s+1\displaystyle=\frac{-4s}{(A+1)^{3}}{\sum_{i,j,k}}^{\prime}\frac{ij+ik-2jk}{\left(i^{2}+j^{2}+k^{2}-2(ij+ik)\left(\frac{A}{A+1}\right)+2jk\left(\frac{A-1}{A+1}\right)\right)^{s+1}}
+4s(s+1)(A+1)4i,j,k(ij+ik2jk)2(i2+j2+k22(ij+ik)(AA+1)+2jk(A1A+1))s+2.\displaystyle\quad+\frac{4s(s+1)}{(A+1)^{4}}{\sum_{i,j,k}}^{\prime}\frac{(ij+ik-2jk)^{2}}{\left(i^{2}+j^{2}+k^{2}-2(ij+ik)\left(\frac{A}{A+1}\right)+2jk\left(\frac{A-1}{A+1}\right)\right)^{s+2}}.

When A=1/2A=1/2 the first sum is zero by the calculations in the first part of the proof. Therefore,

2A2L(A;s)|A=1/2\displaystyle\left.\frac{\partial^{2}}{\partial A^{2}}L(A;s)\right|_{A=1/2} =64s(s+1)81i,j,k(jk+ik2ij)2(i2+j2+k223(ij+ik+jk))s+2.\displaystyle=\frac{64s(s+1)}{81}{\sum_{i,j,k}}^{\prime}\frac{(jk+ik-2ij)^{2}}{\left(i^{2}+j^{2}+k^{2}-\frac{2}{3}(ij+ik+jk)\right)^{s+2}}.

The term (jk+ik2ij)2(jk+ik-2ij)^{2} in the numerator is non-negative and not always zero. The denominator is always positive because the quadratic form is positive definite. It follows that

2A2L(A;s)|A=1/2>0\left.\frac{\partial^{2}}{\partial A^{2}}L(A;s)\right|_{A=1/2}>0

as required.

The calculations above are valid provided term-by-term differentiation of the series is allowed. All of the series encountered above converge absolutely and uniformly on compact subsets of the region Re(s)>3/2(s)>3/2. On restricting ss to real values, the conclusion about positivity is valid for s>3/2s>3/2. ∎

A consequence of Theorem 3.1 is that for any fixed value s>3/2s>3/2, the lattice sum L(A;s)L(A;s) attains a minimum when A=1/2A=1/2. Graphs of y=L(A;s)y=L(A;s) to illustrate this minimum property are shown in Fig. 2. In the limiting case ss\rightarrow\infty we have

L(A;)=limsL(A;s)=kiss()={10if A=1/3,8if 1/3<A<1,12if A=1.L(A;\infty)=\lim_{s\rightarrow\infty}L(A;s)=\textrm{kiss}(\mathcal{L})=\begin{cases}10&\text{if $A=1/3$,}\\ 8&\text{if $1/3<A<1$,}\\ 12&\text{if $A=1$.}\end{cases}

This graph is also shown in Fig. 2.

Refer to caption
Figure 2. Graph of L(A;s)L(A;s) versus AA for various values of ss.

4. Evaluation of the sum L(A;s)L(A;s)

We now turn to the evaluation of L(A;s)L(A;s). Our objectives are to find formulas that are both simple and computationally efficient. The formulas we obtain can be used to show that L(A;s)L(A;s) can be analytically continued to complex values of ss, with a simple pole at s=3/2s=3/2 and no other singularities.

One method of evaluating the sum L(A;s)L(A;s) is to use the Terras decomposition. This was done for f.c.c. and b.c.c. in [6] and can in principle also be used for L(A;s)L(A;s). Here we use an easier method that also works the whole parameter range 1/3A11/3\leq A\leq 1 and hence gives the lattice sum for all four lattices f.c.c., m.c.c., b.c.c. and a.c.c. In fact, the advantage here is that we obtain two formulas which not only can be used as checks, but also contain different information about their analytic continuation.

We begin by writing the lattice sum in the form

L(A;s)\displaystyle L(A;s) =i,j,k(A+1A(i+j)2+(j+k)2+(i+k)2)s\displaystyle={\sum_{i,j,k}}^{\prime}\left(\frac{A+1}{A(i+j)^{2}+(j+k)^{2}+(i+k)^{2}}\right)^{s}
=I,J,KI+J+Keven(A+1AI2+J2+K2)s\displaystyle={\sum_{I,J,K\atop I+J+K\,\text{even}}}^{\!\!\!\!\!\!\!\!\prime}\left(\frac{A+1}{AI^{2}+J^{2}+K^{2}}\right)^{s}
=(A+1)s2i,j,k1+(1)i+j+k(Ai2+j2+k2)s.\displaystyle=\frac{(A+1)^{s}}{2}{\sum_{i,j,k}}^{\prime}\;\;\;\frac{1+(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}. (4.1)

Therefore, we evaluate the sums

T1(A;A;s):=i,j,k1(Ai2+j2+k2)sT_{1}(A;A;s):={\sum_{i,j,k}}^{\prime}\;\;\;\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}} (4.2)

and

T2(A;s):=i,j,k(1)i+j+k(Ai2+j2+k2)s.T_{2}(A;s):={\sum_{i,j,k}}^{\prime}\;\;\;\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}. (4.3)

By (4.1), (4.2) and (4.3), the required lattice sum is given by

L(A;s)=(A+1)s2(T1(A;s)+T2(A;s)).L(A;s)=\frac{(A+1)^{s}}{2}\left(T_{1}(A;s)+T_{2}(A;s)\right). (4.4)

4.1. The sum T1(A;s)T_{1}(A;s)

We shall consider two ways for handling the sum in (4.2). The first is to separate the terms according to whether i=0i=0 or i0i\neq 0, which gives rise to

T1(A;s)=f(s)+2F(s)T_{1}(A;s)=f(s)+2F(s) (4.5)

where

f(s)\displaystyle f(s) =j,k1(j2+k2)s\displaystyle={\sum_{j,k}}^{\,\prime}\;\;\;\frac{1}{(j^{2}+k^{2})^{s}}
and
F(s)\displaystyle F(s) =i=1j=k=1(Ai2+j2+k2)s.\displaystyle=\sum_{i=1}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}.

This is the starting point of the approach taken by Selberg and Chowla [19, Section 7]. Another way is to separate the terms according to whether (j,k)=(0,0)(j,k)=(0,0) or (j,k)(0,0)(j,k)\neq(0,0) and write

T1(A;s)=2g(s)+G(s)T_{1}(A;s)=2g(s)+G(s) (4.6)

where

g(s)\displaystyle g(s) =i=11(Ai2)s\displaystyle=\sum_{i=1}^{\infty}\frac{1}{(Ai^{2})^{s}}
and
G(s)\displaystyle G(s) =j,ki=1(Ai2+j2+k2)s.\displaystyle={\sum_{j,k}}^{\,\prime}\;\sum_{i=-\infty}^{\infty}\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}.

The series F(s)F(s), g(s)g(s) and G(s)G(s) also depend on AA. For simplicity we omit the parameter AA from the notation and just write F(s)F(s), g(s)g(s) and G(s)G(s) in place of F(A;s)F(A;s), g(A;s)g(A;s) and G(A;s)G(A;s), respectively.

We will now analyse (4.5); the corresponding analysis for (4.6) will be given in Section 4.2. By the well-known result (A.20) we have

f(s)=j,k1(j2+k2)s=4ζ(s)L4(s)f(s)={\sum_{j,k}}^{\,\prime}\;\;\;\frac{1}{(j^{2}+k^{2})^{s}}=4\zeta(s)L_{-4}(s)

where

ζ(s)=n=11ns\zeta(s)=\sum_{n=1}^{\infty}\frac{1}{n^{s}}

is the Riemann zeta function, and

L4(s)=n=1sinnπ2ns=11s13s+15s17s+.L_{-4}(s)=\sum_{n=1}^{\infty}\frac{\sin\frac{n\pi}{2}}{n^{s}}=\frac{1}{1^{s}}-\frac{1}{3^{s}}+\frac{1}{5^{s}}-\frac{1}{7^{s}}+\cdots.

is the Dirichlet beta series. It remains to analyse F(s)F(s). By the integral formula for the gamma function (A.2) we have

πsΓ(s)F(s)\displaystyle\pi^{-s}\Gamma(s)F(s) =0xs1i=1eπAxi2j=k=eπx(j2+k2)dx\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}e^{-\pi Axi^{2}}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}e^{-\pi x(j^{2}+k^{2})}\;\mathrm{d}x
=0xs1i=1eπAxi2(j=eπxj2)2dx.\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}e^{-\pi Axi^{2}}\left(\sum_{j=-\infty}^{\infty}e^{-\pi xj^{2}}\right)^{2}\;\mathrm{d}x.
Now apply the modular transformation for theta functions (A.12) to obtain
πsΓ(s)F(s)\displaystyle\pi^{-s}\Gamma(s)F(s) =0xs1i=1eπAxi2(1xj=eπj2/x)2dx\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}e^{-\pi Axi^{2}}\left(\frac{1}{\sqrt{x}}\sum_{j=-\infty}^{\infty}e^{-\pi j^{2}/x}\right)^{2}\;\mathrm{d}x
=0xs2i=1eπAxi2N=0r2(N)eπN/xdx\displaystyle=\int_{0}^{\infty}x^{s-2}\sum_{i=1}^{\infty}e^{-\pi Axi^{2}}\sum_{N=0}^{\infty}r_{2}(N)e^{-\pi N/x}\ \;\mathrm{d}x
where r2(N)r_{2}(N) is the number of representations of NN as a sum of two squares, e.g., see (A.13). Now separate out the N=0N=0 term and evaluate the resulting integrals. We find that
πsΓ(s)F(s)\displaystyle\pi^{-s}\Gamma(s)F(s) =i=10xs2eπAxi2dx+i=1N=1r2(N)0xs2eπAxi2πN/xdx\displaystyle=\sum_{i=1}^{\infty}\int_{0}^{\infty}x^{s-2}e^{-\pi Axi^{2}}\;\mathrm{d}x+\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\int_{0}^{\infty}x^{s-2}e^{-\pi Axi^{2}-\pi N/x}\ \;\mathrm{d}x
=Γ(s1)ζ(2s2)As1πs1+2i=1N=1r2(N)(NAi2)(s1)/2Ks1(2πiAN)\displaystyle=\frac{\Gamma(s-1)\zeta(2s-2)}{A^{s-1}\pi^{s-1}}+2\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{N}{Ai^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{AN}\right)

where we have used the formula (A.3) for the KK-Bessel function. On using all of the above back in (4.5) we obtain

i,j,k1(Ai2+j2+k2)s=4ζ(s)L4(s)+2π(s1)ζ(2s2)As1\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}=4\zeta(s)L_{-4}(s)+\frac{2\pi}{(s-1)}\frac{\zeta(2s-2)}{A^{s-1}}
+4πsΓ(s)A(1s)/2i=1N=1r2(N)(Ni2)(s1)/2Ks1(2πiAN).\displaystyle\quad+\frac{4\pi^{s}}{\Gamma(s)}\,A^{(1-s)/2}\,\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{N}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{AN}\right). (4.7)

This is essentially Selberg and Chowla’s formula [19, (45)]. They write it in a slightly different form in terms of a sum over the divisors of NN to minimise the number of Bessel function evaluations. We will leave it as it is for simplicity.

4.2. A second formula for the sum T1(A;s)T_{1}(A;s)

This time we split the terms according as in (4.6) and start with

T1(A;s)=2g(s)+G(s)T_{1}(A;s)=2g(s)+G(s) (4.8)

where

g(s)\displaystyle g(s) =i=11(Ai2)s=Asζ(2s)\displaystyle=\sum_{i=1}^{\infty}\frac{1}{(Ai^{2})^{s}}=A^{-s}\,\zeta(2s)
and
G(s)\displaystyle G(s) =j,ki=1(Ai2+j2+k2)s.\displaystyle={\sum_{j,k}}^{\,\prime}\;\sum_{i=-\infty}^{\infty}\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}.

Now apply the integral formula for the gamma function (A.2) and then the modular transformation for the theta function (A.9) to obtain

πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s) =0xs1j,keπ(j2+k2)xi=eπi2Axdx\displaystyle=\int_{0}^{\infty}x^{s-1}{\sum_{j,k}}^{\;\prime}e^{-\pi(j^{2}+k^{2})x}\sum_{i=-\infty}^{\infty}e^{-\pi i^{2}Ax}\,\mathrm{d}x
=1A0xs3/2j,keπ(j2+k2)xi=eπi2/Axdx.\displaystyle=\frac{1}{\sqrt{A}}\int_{0}^{\infty}x^{s-3/2}\,{\sum_{j,k}}^{\;\prime}e^{-\pi(j^{2}+k^{2})x}\sum_{i=-\infty}^{\infty}e^{-\pi i^{2}/Ax}\,\mathrm{d}x.
Separate the i=0i=0 term, to get
πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s) =1A0xs3/2j,keπ(j2+k2)xdx\displaystyle=\frac{1}{\sqrt{A}}\int_{0}^{\infty}x^{s-3/2}\,{\sum_{j,k}}^{\;\prime}e^{-\pi(j^{2}+k^{2})x}\,\mathrm{d}x
+2A0xs3/2j,keπ(j2+k2)xi=1eπi2/Axdx.\displaystyle\quad+\frac{2}{\sqrt{A}}\int_{0}^{\infty}x^{s-3/2}\,{\sum_{j,k}}^{\;\prime}e^{-\pi(j^{2}+k^{2})x}\sum_{i=1}^{\infty}e^{-\pi i^{2}/Ax}\,\mathrm{d}x.

The first integral can be evaluated in terms of the gamma function by (A.2), while the second integral can be expressed in terms of the modified Bessel function by (A.3). The result is

πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s) =Γ(s12)Aπs12j,k1(j2+k2)s12\displaystyle=\frac{\Gamma\left(s-\frac{1}{2}\right)}{\sqrt{A}\,\pi^{s-\frac{1}{2}}}\;{\sum_{j,k}}^{\;\prime}\;\frac{1}{(j^{2}+k^{2})^{s-\frac{1}{2}}}
+4As2+14j,ki=1(ij2+k2)s12Ks12(2πij2+k2A)\displaystyle\quad+\frac{4}{A^{\frac{s}{2}+\frac{1}{4}}}\,{\sum_{j,k}}^{\;\prime}\;\;\sum_{i=1}^{\infty}\left(\frac{i}{\sqrt{j^{2}+k^{2}}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi i\sqrt{\frac{j^{2}+k^{2}}{A}}\right)
=4Aπ(s12)Γ(s12)ζ(s12)L4(s12)\displaystyle=\frac{4}{\sqrt{A}}\,{\pi^{-(s-\frac{1}{2})}}\,\Gamma\left(s-\frac{1}{2}\right)\,\zeta\left(s-\frac{1}{2}\right)\,L_{-4}\left(s-\frac{1}{2}\right)
+4As2+14N=1i=1r2(N)(iN)s12Ks12(2πiNA).\displaystyle\quad+\frac{4}{A^{\frac{s}{2}+\frac{1}{4}}}\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}r_{2}(N)\left(\frac{i}{\sqrt{N}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi i\sqrt{\frac{N}{A}}\right).

On using all of the above back in (4.8) we obtain

i,j,k1(Ai2+j2+k2)s\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}
=2Asζ(2s)+4πAΓ(s12)Γ(s)ζ(s12)L4(s12)\displaystyle=2A^{-s}\zeta(2s)+4\,\sqrt{\frac{\pi}{A}}\,\frac{\Gamma\left(s-\frac{1}{2}\right)}{\Gamma(s)}\,\zeta\left(s-\frac{1}{2}\right)\,L_{-4}\left(s-\frac{1}{2}\right)
+4As2+14πsΓ(s)N=1i=1r2(N)(iN)s12Ks12(2πiNA).\displaystyle\quad+\frac{4}{A^{\frac{s}{2}+\frac{1}{4}}}\,\frac{\pi^{s}}{\Gamma(s)}\,\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}r_{2}(N)\left(\frac{i}{\sqrt{N}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi i\sqrt{\frac{N}{A}}\right). (4.9)

The terms in (4.7) involve Ks1K_{s-1} Bessel functions whereas Ks1/2K_{s-1/2} Bessel functions occur in (4.9). That is because each application of the theta function transformation formula lowers the subscript in the resulting Bessel function by 1/21/2, due to the creation of a x1/2x^{-1/2} factor in the integral. The theta function transformation formula is used twice (i.e., the formula is squared) in the derivation of (4.7) and only once in the derivation of (4.9).

Each of (4.7) and (4.9) turns out to have its own advantages, as will be seen in Sections 6.4 and 6.3.

4.3. The alternating sum T2(A;s)T_{2}(A;s)

The analysis in the previous sections can be modified to handle the alternating series (4.3) which has the term (1)i+j+k(-1)^{i+j+k} in the numerator, as follows. Separating the terms according to whether i=0i=0 or i0i\neq 0 gives

T2(A;s)\displaystyle T_{2}(A;s) =h(s)+2H(S)\displaystyle=h(s)+2H(S) (4.10)
where
h(s)\displaystyle h(s) =j,k(1)j+k(j2+k2)s\displaystyle={\sum_{j,k}}^{\prime}\;\;\;\frac{(-1)^{j+k}}{(j^{2}+k^{2})^{s}}
and
H(s)\displaystyle H(s) =i=1j=k=(1)i+j+k(Ai2+j2+k2)s.\displaystyle=\sum_{i=1}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}.

By a known result (A.21), we have

h(s)=4(121s)ζ(s)L4(s).h(s)=-4(1-2^{1-s})\zeta(s)L_{-4}(s).

Next, by the integral formula for the gamma function (A.2) we have

πsΓ(s)H(s)\displaystyle\pi^{-s}\Gamma(s)H(s) =0xs1i=1(1)ieπAxi2j=k=(1)j+keπx(j2+k2)dx\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}(-1)^{i}e^{-\pi Axi^{2}}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}(-1)^{j+k}e^{-\pi x(j^{2}+k^{2})}\;\mathrm{d}x
=0xs1i=1(1)ieπAxi2(j=(1)jeπxj2)2dx.\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}(-1)^{i}e^{-\pi Axi^{2}}\left(\sum_{j=-\infty}^{\infty}(-1)^{j}e^{-\pi xj^{2}}\right)^{2}\;\mathrm{d}x.
Now apply the modular transformation for theta functions to obtain
πsΓ(s)H(s)\displaystyle\pi^{-s}\Gamma(s)H(s) =0xs1i=1(1)ieπAxi2(1xj=eπ(j+12)2/x)2dx.\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{i=1}^{\infty}(-1)^{i}e^{-\pi Axi^{2}}\left(\frac{1}{\sqrt{x}}\sum_{j=-\infty}^{\infty}e^{-\pi(j+\frac{1}{2})^{2}/x}\right)^{2}\;\mathrm{d}x.
By formula (A.14) this can be expressed as
πsΓ(s)H(s)\displaystyle\pi^{-s}\Gamma(s)H(s) =0xs2i=1(1)ieπAxi2N=0r2(4N+1)eπ(4N+1)/2xdx\displaystyle=\int_{0}^{\infty}x^{s-2}\sum_{i=1}^{\infty}(-1)^{i}e^{-\pi Axi^{2}}\sum_{N=0}^{\infty}r_{2}(4N+1)e^{-\pi(4N+1)/2x}\ \;\mathrm{d}x
=i=1N=0(1)ir2(4N+1)0xs2eπAxi2π(4N+1)/2xdx.\displaystyle=\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)\int_{0}^{\infty}x^{s-2}e^{-\pi Axi^{2}-\pi(4N+1)/2x}\ \;\mathrm{d}x.
The integral can be expressed in terms of Bessel functions by (A.3) to give
πsΓ(s)H(s)\displaystyle\pi^{-s}\Gamma(s)H(s) =2i=1N=0(1)ir2(4N+1)(2N+12Ai2)(s1)/2Ks1(2πiA(2N+12)).\displaystyle=2\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)\left(\frac{2N+\frac{1}{2}}{Ai^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{A(2N+\frac{1}{2})}\right).

On using all of the above back in (4.10) we obtain

i,j,k(1)i+j+k(Ai2+j2+k2)s\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}
=4(121s)ζ(s)L4(s)\displaystyle=-4(1-2^{1-s})\zeta(s)L_{-4}(s)
+4πsΓ(s)A(1s)/2i=1N=0(1)ir2(4N+1)(2N+12i2)(s1)/2Ks1(2πiA(2N+12)).\displaystyle+\frac{4\pi^{s}}{\Gamma(s)}\,A^{(1-s)/2}\,\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)\left(\frac{2N+\frac{1}{2}}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{A(2N+\frac{1}{2})}\right). (4.11)

4.4. A second formula for the alternating sum T2(A;s)T_{2}(A;s)

This time we separate the terms according to whether (j,k)=(0,0)(j,k)=(0,0) or (j,k)(0,0)(j,k)\neq(0,0) and write

T2(A;s)=2i=1(1)i(Ai2)s+J(s)T_{2}(A;s)=2\sum_{i=1}^{\infty}\frac{(-1)^{i}}{(Ai^{2})^{s}}+J(s) (4.12)

where

J(s)=j,ki=(1)i+j+k(Ai2+j2+k2)s.J(s)={\sum_{j,k}}^{\,\prime}\;\sum_{i=-\infty}^{\infty}\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}.

By (A.30) we have

2i=1(1)i(Ai2)s=2As(1212s)ζ(2s).2\sum_{i=1}^{\infty}\frac{(-1)^{i}}{(Ai^{2})^{s}}=-2A^{-s}(1-2^{1-2s})\zeta(2s).

It remains to analyse the sum for J(s)J(s). By the integral formula for the gamma function (A.2) we have

πsΓ(s)J(s)\displaystyle\pi^{-s}\Gamma(s)J(s) =0xs1j,k(1)j+jeπ(j2+k2)xi=(1)ieπi2Axdx.\displaystyle=\int_{0}^{\infty}x^{s-1}{\sum_{j,k}}^{\;\prime}(-1)^{j+j}e^{-\pi(j^{2}+k^{2})x}\sum_{i=-\infty}^{\infty}(-1)^{i}e^{-\pi i^{2}Ax}\,\mathrm{d}x.
Now apply the modular transformation (A.11) to obtain
πsΓ(s)J(s)\displaystyle\pi^{-s}\Gamma(s)J(s) =1A0xs3/2j,k(1)j+keπ(j2+k2)xi=eπ(i+12)2/Axdx.\displaystyle=\frac{1}{\sqrt{A}}\int_{0}^{\infty}x^{s-3/2}\,{\sum_{j,k}}^{\;\prime}(-1)^{j+k}e^{-\pi(j^{2}+k^{2})x}\sum_{i=-\infty}^{\infty}e^{-\pi(i+\frac{1}{2})^{2}/Ax}\,\mathrm{d}x.

Now put N=j2+k2N=j^{2}+k^{2} and use

i=eπ(i+12)2/Ax=2i=0eπ(i+12)2/Ax=2i=1eπ(i12)2/Ax\sum_{i=-\infty}^{\infty}e^{-\pi(i+\frac{1}{2})^{2}/Ax}=2\sum_{i=0}^{\infty}e^{-\pi(i+\frac{1}{2})^{2}/Ax}=2\sum_{i=1}^{\infty}e^{-\pi(i-\frac{1}{2})^{2}/Ax}

to deduce

πsΓ(s)J(s)\displaystyle\pi^{-s}\Gamma(s)J(s) =2AN=1i=1(1)Nr2(N)0xs3/2eπNxπ(i12)2/Axdx.\displaystyle=\frac{2}{\sqrt{A}}\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}(-1)^{N}r_{2}(N)\int_{0}^{\infty}x^{s-3/2}\,e^{-\pi Nx-\pi(i-\frac{1}{2})^{2}/Ax}\,\mathrm{d}x.
The integral can be evaluated in terms of the modified Bessel function by (A.3) to give
πsΓ(s)J(s)\displaystyle\pi^{-s}\Gamma(s)J(s) =4As2+14N=1i=1(1)Nr2(N)(i12N)s12Ks12(2π(i12)NA).\displaystyle=\frac{4}{A^{\frac{s}{2}+\frac{1}{4}}}\,\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}(-1)^{N}\,r_{2}(N)\left(\frac{i-\frac{1}{2}}{\sqrt{N}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi(i-\frac{1}{2})\sqrt{\frac{N}{A}}\right).

It follows that

i,j,k(1)i+j+k(Ai2+j2+k2)s\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}
=2As(1212s)ζ(2s)\displaystyle=-2A^{-s}(1-2^{1-{2s}})\zeta(2s)
+4As2+14πsΓ(s)N=1i=1(1)Nr2(N)(i12N)s12Ks12(2π(i12)NA).\displaystyle\quad+\frac{4}{A^{\frac{s}{2}+\frac{1}{4}}}\,\frac{\pi^{s}}{\Gamma(s)}\,\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}(-1)^{N}\,r_{2}(N)\left(\frac{i-\frac{1}{2}}{\sqrt{N}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi(i-\frac{1}{2})\sqrt{\frac{N}{A}}\right). (4.13)

4.5. Two formulas for L(A;s)L(A;s)

On substituting the results of (4.7) and (4.11) back into (4.4) we obtain a formula for L(A;s)L(A;s) in terms of Ks1K_{s-1} Bessel functions:

L(A;s)\displaystyle L(A;s) =4(A+12)sζ(s)L4(s)+πAs1(1+1A)sζ(2s2)\displaystyle=4\left(\frac{A+1}{2}\right)^{s}\zeta(s)L_{-4}(s)+\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2)
+2πsAΓ(s)(A+1A)si=1N=1r2(N)(Ni2)(s1)/2Ks1(2πiAN)\displaystyle\quad+\frac{2\pi^{s}\sqrt{A}}{\Gamma(s)}\left(\sqrt{A}+\frac{1}{\sqrt{A}}\right)^{s}\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{N}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{AN}\right)
+2πsAΓ(s)(A+1A)si=1N=0(1)ir2(4N+1)\displaystyle\quad+\frac{2\pi^{s}\sqrt{A}}{\Gamma(s)}\left(\sqrt{A}+\frac{1}{\sqrt{A}}\right)^{s}\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)
×(2N+12i2)(s1)/2Ks1(2πiA(2N+12)).\displaystyle\qquad\qquad\qquad\times\left(\frac{2N+\frac{1}{2}}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{A(2N+\frac{1}{2})}\right). (4.14)

On the other hand, if the results of (4.9) and (4.13) are used in (4.4), the resulting formula for L(A;s)L(A;s) involves Ks1/2K_{s-1/2} Bessel functions:

L(A;s)\displaystyle L(A;s) =2(A+14A)sζ(2s)+2πA(A+1)sΓ(s12)Γ(s)ζ(s12)L4(s12)\displaystyle=2\left(\frac{A+1}{4A}\right)^{s}\zeta(2s)+2\,\sqrt{\frac{\pi}{A}}\,(A+1)^{s}\,\frac{\Gamma\left(s-\frac{1}{2}\right)}{\Gamma(s)}\,\zeta\left(s-\frac{1}{2}\right)\,L_{-4}\left(s-\frac{1}{2}\right)
+2A1/4(A+1A)sπsΓ(s)N=1i=1N(12s)/4r2(N)\displaystyle\quad+\frac{2}{A^{1/4}}\left(\sqrt{A}+\frac{1}{\sqrt{A}}\right)^{s}\,\frac{\pi^{s}}{\Gamma(s)}\,\,\sum_{N=1}^{\infty}\sum_{i=1}^{\infty}N^{(1-2s)/4}\,r_{2}(N)
×{is12Ks12(2πiNA)+(1)N(i12)s12Ks12(2π(i12)NA)}.\displaystyle\qquad\times\left\{i^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi i\sqrt{\frac{N}{A}}\right)+(-1)^{N}\left(i-\frac{1}{2}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi(i-\frac{1}{2})\sqrt{\frac{N}{A}}\right)\right\}. (4.15)

The formulas (4.14) and (4.15) can be used as checks against each other. Moreover, the formulas offer different information about special values of the lattice sum, as will be seen in Section 6.

5. Hexagonal close packing

Because of its importance in solid state chemistry and physics, we give a similar analysis of the lattice sum for the hexagonal close packed structure given by [15]

L3HCP(s)=S1(s)+S2(s)L_{3}^{\text{HCP}}(s)=S_{1}(s)+S_{2}(s)

where

S1(s)=i,j,k1(i2+ij+j2+83k2)sS_{1}(s)={\sum_{i,j,k}}^{\prime}\frac{1}{(i^{2}+ij+j^{2}+\frac{8}{3}k^{2})^{s}}

and

S2(s)=i,j,k1((i+13)2+(i+13)(j+13)+(j+13)2+83(k+12)2)s.S_{2}(s)=\sum_{i,j,k}\frac{1}{((i+\frac{1}{3})^{2}+(i+\frac{1}{3})(j+\frac{1}{3})+(j+\frac{1}{3})^{2}+\frac{8}{3}(k+\frac{1}{2})^{2})^{s}}.

As before, the prime on the sum for S1(s)S_{1}(s) indicates that the summation is over all integers except for the term corresponding to i=j=k=0i=j=k=0 which is omitted. The sum for S2(s)S_{2}(s) is over all integer values of ii, jj and kk. We shall analyse S1(s)S_{1}(s) and S2(s)S_{2}(s) one at a time.

5.1. The sum S1(s)S_{1}(s)

Break the sum for S1(s)S_{1}(s) into two, according to whether k=0k=0 or k0k\neq 0. This gives

S1(s)=f(s)+2F(s)S_{1}(s)=f(s)+2F(s) (5.1)

where

f(s)=i,j1(i2+ij+j2)sf(s)={\sum_{i,j}}^{\;\prime}\frac{1}{(i^{2}+ij+j^{2})^{s}}

and

F(s)=k=1i,j1(i2+ij+j2+83k2)s.F(s)=\sum_{k=1}^{\infty}\sum_{i,j}\frac{1}{(i^{2}+ij+j^{2}+\frac{8}{3}k^{2})^{s}}.

By (A.33) we have

f(s)=6ζ(s)L3(s)f(s)=6\zeta(s)\,L_{-3}(s)

where

L3(s)=k=1(sin(2kπ/3)sin(2π/3))1ks=11s12s+14s15s+.L_{-3}(s)=\sum_{k=1}^{\infty}\left(\frac{\sin(2k\pi/3)}{\sin(2\pi/3)}\right)\frac{1}{k^{s}}=\frac{1}{1^{s}}-\frac{1}{2^{s}}+\frac{1}{4^{s}}-\frac{1}{5^{s}}+\cdots.

It remains to calculate F(s)F(s). Applying the gamma function integral (A.2) followed by the theta function transformation formula (A.15), we obtain

(2π)sΓ(s)F(s)\displaystyle(2\pi)^{-s}\Gamma(s)F(s) =0xs1k=1e16πk2x/3i,je2π(i2+ij+j2)xdx\displaystyle=\int_{0}^{\infty}x^{s-1}\sum_{k=1}^{\infty}e^{-16\pi k^{2}x/3}\sum_{i,j}e^{-2\pi(i^{2}+ij+j^{2})x}\,\mathrm{d}x
=130xs2k=1e16πk2x/3i,je2π(i2+ij+j2)/3xdx.\displaystyle=\frac{1}{\sqrt{3}}\,\int_{0}^{\infty}x^{s-2}\sum_{k=1}^{\infty}e^{-16\pi k^{2}x/3}\sum_{i,j}e^{-2\pi(i^{2}+ij+j^{2})/3x}\,\mathrm{d}x.

Now separate out the i=j=0i=j=0 term and evaluate the resulting integrals. The result is

(2π)sΓ(s)F(s)\displaystyle(2\pi)^{-s}\Gamma(s)F(s) =13k=0xs2e16πk2x/3dx\displaystyle=\frac{1}{\sqrt{3}}\,\sum_{k=-\infty}^{\infty}\int_{0}^{\infty}x^{s-2}e^{-16\pi k^{2}x/3}\,\mathrm{d}x
+13k=i,j0xs2e16πk2x/32π(i2+ij+j2)/3xdx\displaystyle\quad+\frac{1}{\sqrt{3}}\,\sum_{k=-\infty}^{\infty}\;{\sum_{i,j}}^{\;\prime}\int_{0}^{\infty}x^{s-2}e^{-16\pi k^{2}x/3-2\pi(i^{2}+ij+j^{2})/3x}\,\mathrm{d}x
=13(316π)s1Γ(s1)ζ(2s2)\displaystyle=\frac{1}{\sqrt{3}}\left(\frac{3}{16\pi}\right)^{s-1}\Gamma(s-1)\zeta(2s-2)
+23k=1i,j(i2+ij+j28k2)(s1)/2Ks1(8πk32(i2+ij+j2)).\displaystyle\quad+\frac{2}{\sqrt{3}}\sum_{k=1}^{\infty}{\sum_{i,j}}^{\;\prime}\left(\frac{i^{2}+ij+j^{2}}{8k^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi k}{3}\sqrt{2(i^{2}+ij+j^{2})}\right).

It follows that

S1(s)\displaystyle S_{1}(s) =6ζ(s)L3(s)+4π3(38)s1(1s1)ζ(2s2)\displaystyle=6\zeta(s)\,L_{-3}(s)+\frac{4\pi}{\sqrt{3}}\left(\frac{3}{8}\right)^{s-1}\left(\frac{1}{s-1}\right)\zeta(2s-2)
+83πsΓ(s)k=1i,j(i2+ij+j22k2)(s1)/2Ks1(8πk32(i2+ij+j2))\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=1}^{\infty}{\sum_{i,j}}^{\;\prime}\left(\frac{i^{2}+ij+j^{2}}{2k^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi k}{3}\sqrt{2(i^{2}+ij+j^{2})}\right)
=6ζ(s)L3(s)+4π3(38)s1(1s1)ζ(2s2)\displaystyle=6\zeta(s)\,L_{-3}(s)+\frac{4\pi}{\sqrt{3}}\left(\frac{3}{8}\right)^{s-1}\left(\frac{1}{s-1}\right)\zeta(2s-2)
+83πsΓ(s)k=1N=1u2(N)(N2k2)(s1)/2Ks1(8πk32N)\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=1}^{\infty}\sum_{N=1}^{\infty}u_{2}(N)\left(\frac{N}{2k^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi k}{3}\sqrt{2N}\right) (5.2)

where u2(N)u_{2}(N) is the number of representations of NN by the form i2+ij+j2i^{2}+ij+j^{2}.

5.2. A second formula for the sum S1(s)S_{1}(s)

A different formula for S1(s)S_{1}(s) can be obtained by separating the terms in the series according to whether i=j=0i=j=0 or ii and jj are not both zero. This gives

S1(s)=2(38)sk=11k2s+G(s)S_{1}(s)=2\left(\frac{3}{8}\right)^{s}\sum_{k=1}^{\infty}\frac{1}{k^{2s}}+G(s)

where

G(s)=i,jk=1(i2+ij+j2+83k2)s.G(s)={\sum_{i,j}}^{\;\prime}\sum_{k=-\infty}^{\infty}\frac{1}{(i^{2}+ij+j^{2}+\frac{8}{3}k^{2})^{s}}.

Applying the gamma function integral (A.2) followed by the theta function transformation formula (A.9), we obtain

πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s) =0xs1i,jeπ(i2+ij+j2)xk=e8πk2x/3dx\displaystyle=\int_{0}^{\infty}x^{s-1}{\sum_{i,j}}^{\;\prime}e^{-\pi(i^{2}+ij+j^{2})x}\sum_{k=-\infty}^{\infty}e^{-8\pi k^{2}x/3}\,\mathrm{d}x
=380xs32i,jeπ(i2+ij+j2)xk=e3πk2/8xdx.\displaystyle=\sqrt{\frac{3}{8}}\int_{0}^{\infty}x^{s-\frac{3}{2}}{\sum_{i,j}}^{\;\prime}e^{-\pi(i^{2}+ij+j^{2})x}\sum_{k=-\infty}^{\infty}e^{-3\pi k^{2}/8x}\,\mathrm{d}x.

Now separate out the k=0k=0 term and evaluate the resulting integrals. The result is

πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s)
=380xs32i,jeπ(i2+ij+j2)xdx\displaystyle=\sqrt{\frac{3}{8}}\int_{0}^{\infty}x^{s-\frac{3}{2}}{\sum_{i,j}}^{\;\prime}e^{-\pi(i^{2}+ij+j^{2})x}\,\mathrm{d}x
+2380xs32i,jeπ(i2+ij+j2)xk=1e3πk2/8xdx\displaystyle\quad+2\sqrt{\frac{3}{8}}\int_{0}^{\infty}x^{s-\frac{3}{2}}{\sum_{i,j}}^{\;\prime}e^{-\pi(i^{2}+ij+j^{2})x}\sum_{k=1}^{\infty}e^{-3\pi k^{2}/8x}\,\mathrm{d}x
=38π(s12)Γ(s12)i,j1(i2+ij+j2)s1/2\displaystyle=\sqrt{\frac{3}{8}}\,\pi^{-(s-\frac{1}{2})}\,\Gamma(s-\frac{1}{2}){\sum_{i,j}}^{\;\prime}\frac{1}{(i^{2}+ij+j^{2})^{s-1/2}}
+4(38)(2s+1)/4i,jk=1(k2i2+ij+j2)(2s1)/4Ks12(32πki2+ij+j2).\displaystyle\quad+4\left(\frac{3}{8}\right)^{(2s+1)/4}{\sum_{i,j}}^{\;\prime}\sum_{k=1}^{\infty}\left(\frac{k^{2}}{i^{2}+ij+j^{2}}\right)^{(2s-1)/4}K_{s-\frac{1}{2}}\left(\sqrt{\frac{3}{2}}\,\pi k\sqrt{i^{2}+ij+j^{2}}\right).

The first sum can be evaluated in terms of the Riemann zeta function and the L3L_{-3} function by (A.33). In the second sum, we again use the notation u2(N)u_{2}(N) for the number of representations of NN by the form i2+ij+j2i^{2}+ij+j^{2}. The result is

πsΓ(s)G(s)\displaystyle\pi^{-s}\Gamma(s)G(s) =272π(s12)Γ(s12)ζ(s12)L3(s12)\displaystyle=\sqrt{\frac{27}{2}}\,\pi^{-(s-\frac{1}{2})}\,\Gamma\left(s-\frac{1}{2}\right)\zeta\left(s-\frac{1}{2}\right)L_{-3}\left(s-\frac{1}{2}\right)
+4(38)(2s+1)/4N=1k=1u2(N)(k2N)(2s1)/4Ks12(πk3N2).\displaystyle\quad+4\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{N=1}^{\infty}\sum_{k=1}^{\infty}u_{2}(N)\,\left(\frac{k^{2}}{N}\right)^{(2s-1)/4}K_{s-\frac{1}{2}}\left(\pi k\sqrt{\frac{3N}{2}}\right).

It follows that

S1(s)\displaystyle S_{1}(s) =2(38)sζ(2s)+27π2Γ(s12)Γ(s)ζ(s12)L3(s12)\displaystyle=2\left(\frac{3}{8}\right)^{s}\zeta(2s)+\sqrt{\frac{27\pi}{2}}\;\frac{\Gamma(s-\frac{1}{2})}{\Gamma(s)}\;\zeta\left(s-\frac{1}{2}\right)L_{-3}\left(s-\frac{1}{2}\right)
+4πsΓ(s)(38)(2s+1)/4N=1k=1u2(N)(k2N)(2s1)/4Ks12(πk3N2).\displaystyle\quad+\frac{4\pi^{s}}{\Gamma(s)}\,\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{N=1}^{\infty}\sum_{k=1}^{\infty}u_{2}(N)\,\left(\frac{k^{2}}{N}\right)^{(2s-1)/4}K_{s-\frac{1}{2}}\left(\pi k\sqrt{\frac{3N}{2}}\right). (5.3)

5.3. The sum S2(s)S_{2}(s)

The analysis in this case is a little simpler because the summation is over all integers ii, jj and kk. We apply the gamma function integral (A.2) to write

S2(s)\displaystyle S_{2}(s) =i,j,k1((i+13)2+(i+13)(j+13)+(j+13)2+83(k+12)2)s\displaystyle=\sum_{i,j,k}\frac{1}{((i+\frac{1}{3})^{2}+(i+\frac{1}{3})(j+\frac{1}{3})+(j+\frac{1}{3})^{2}+\frac{8}{3}(k+\frac{1}{2})^{2})^{s}}
=(2π)sΓ(s)0xs1k=e16π(k+12)2x/3i,j=e2π((i+13)2+(i+13)(j+13)+(j+13)2)xdx.\displaystyle=\frac{(2\pi)^{s}}{\Gamma(s)}\int_{0}^{\infty}x^{s-1}\sum_{k=-\infty}^{\infty}e^{-16\pi(k+\frac{1}{2})^{2}x/3}\sum_{i,j=-\infty}^{\infty}e^{-2\pi((i+\frac{1}{3})^{2}+(i+\frac{1}{3})(j+\frac{1}{3})+(j+\frac{1}{3})^{2})x}\,\mathrm{d}x. (5.4)

Now make use of the transformation formula  (A.16) to deduce

S2(s)\displaystyle S_{2}(s) =(2π)sΓ(s)0xs1(2k=0e16π(k+12)2x/3)(1x3i,j=ωije2π(i2+ij+j2)/3x)dx\displaystyle=\frac{(2\pi)^{s}}{\Gamma(s)}\int_{0}^{\infty}x^{s-1}\left(2\sum_{k=0}^{\infty}e^{-16\pi(k+\frac{1}{2})^{2}x/3}\right)\left(\frac{1}{x\sqrt{3}}\sum_{i,j=-\infty}^{\infty}\omega^{i-j}e^{-2\pi(i^{2}+ij+j^{2})/3x}\right)\,\mathrm{d}x
where ω=e2πi/3\omega=e^{2\pi i/3} is a primitive cube root of 1. Now separate the term i=j=0i=j=0 to deduce
S2(s)\displaystyle S_{2}(s) =(2π)sΓ(s)230xs2k=0e16π(k+12)2x/3dx\displaystyle=\frac{(2\pi)^{s}}{\Gamma(s)}\frac{2}{\sqrt{3}}\int_{0}^{\infty}x^{s-2}\sum_{k=0}^{\infty}e^{-16\pi(k+\frac{1}{2})^{2}x/3}\,\mathrm{d}x
+(2π)sΓ(s)230xs2k=0e16π(k+12)2x/3N=1ωNu2(N)e2πN/3xdx\displaystyle\quad+\frac{(2\pi)^{s}}{\Gamma(s)}\frac{2}{\sqrt{3}}\int_{0}^{\infty}x^{s-2}\sum_{k=0}^{\infty}e^{-16\pi(k+\frac{1}{2})^{2}x/3}\;\;\sum_{N=1}^{\infty}\omega^{N}\,u_{2}(N)\,e^{-2\pi N/3x}\,\mathrm{d}x

where u2(N)u_{2}(N) is the number of representations of NN by the form i2+ij+j2i^{2}+ij+j^{2}, as before. On evaluating the integrals using (A.2) and (A.3) we obtain

S2(s)\displaystyle S_{2}(s) =4π3(38)s1(1s1)k=01(k+12)2s2\displaystyle=\frac{4\pi}{\sqrt{3}}\left(\frac{3}{8}\right)^{s-1}\left(\frac{1}{s-1}\right)\,\sum_{k=0}^{\infty}\frac{1}{(k+\frac{1}{2})^{2s-2}}
+83πsΓ(s)k=0N=1ωNu2(N)(N2(k+12)2)(s1)/2Ks1(8π(k+12)32N).\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=0}^{\infty}\sum_{N=1}^{\infty}\omega^{N}\,u_{2}(N)\,\left(\frac{N}{2(k+\frac{1}{2})^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi\left(k+\frac{1}{2}\right)}{3}\sqrt{2N}\right).

The first sum can be evaluated in terms of the Riemann zeta function by using (A.29). The ωN\omega^{N} term can be handled by using

ωN=cos2πN3+isin2πN3\omega^{N}=\cos\frac{2\pi N}{3}+i\sin\frac{2\pi N}{3}

along with the fact that S2(s)S_{2}(s) is real valued when ss is real. It follows that

S2(s)\displaystyle S_{2}(s) =4π3(38)s1(22s21)(1s1)ζ(2s2)\displaystyle=\frac{4\pi}{\sqrt{3}}\left(\frac{3}{8}\right)^{s-1}(2^{2s-2}-1)\left(\frac{1}{s-1}\right)\zeta(2s-2)
+83πsΓ(s)k=0N=1cos2πN3u2(N)(N2(k+12)2)(s1)/2Ks1(8π(k+12)32N).\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=0}^{\infty}\sum_{N=1}^{\infty}\cos\frac{2\pi N}{3}\,u_{2}(N)\,\left(\frac{N}{2(k+\frac{1}{2})^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi\left(k+\frac{1}{2}\right)}{3}\sqrt{2N}\right). (5.5)

5.4. A second formula for the sum S2(s)S_{2}(s)

We introduce the abbreviation

Yi,j=(i+13)2+(i+13)(j+13)+(j+13)2Y_{i,j}=\left(i+\frac{1}{3}\right)^{2}+\left(i+\frac{1}{3}\right)\left(j+\frac{1}{3}\right)+\left(j+\frac{1}{3}\right)^{2}

to write (5.4) in the form

S2(s)\displaystyle S_{2}(s) =(2π)sΓ(s)0xs1i,j=e2πYi,jxk=e16π(k+12)2x/3dx.\displaystyle=\frac{(2\pi)^{s}}{\Gamma(s)}\int_{0}^{\infty}x^{s-1}\sum_{i,j=-\infty}^{\infty}e^{-2\pi Y_{i,j}x}\sum_{k=-\infty}^{\infty}e^{-16\pi(k+\frac{1}{2})^{2}x/3}\,\mathrm{d}x.
This time we apply the transformation formula (A.11) to the sum over kk to obtain
S2(s)\displaystyle S_{2}(s) =34(2π)sΓ(s)0xs3/2i,j=e2πYi,jxk=(1)ke3πk2/16xdx.\displaystyle=\frac{\sqrt{3}}{4}\,\frac{(2\pi)^{s}}{\Gamma(s)}\int_{0}^{\infty}x^{s-3/2}\sum_{i,j=-\infty}^{\infty}e^{-2\pi Y_{i,j}x}\sum_{k=-\infty}^{\infty}(-1)^{k}\,e^{-3\pi k^{2}/16x}\,\mathrm{d}x.

Now separate the terms according to whether k=0k=0 or k0k\neq 0 and evaluate the resulting integrals by (A.2) and (A.3). The result is

S2(s)\displaystyle S_{2}(s) =3π8Γ(s12)Γ(s)i,j=1Yijs1/2\displaystyle=\sqrt{\frac{3\pi}{8}}\,\frac{\Gamma(s-\frac{1}{2})}{\Gamma(s)}\,\sum_{i,j=-\infty}^{\infty}\frac{1}{Y_{ij}^{s-1/2}}
+4πsΓ(s)(38)(2s+1)/4k=1(1)ki,j=(kYij)s12Ks12(πk3Yi,j/2).\displaystyle\quad+\frac{4\pi^{s}}{\Gamma(s)}\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{k=1}^{\infty}(-1)^{k}\sum_{i,j=-\infty}^{\infty}\left(\frac{k}{\sqrt{Y_{ij}}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(\pi k\sqrt{3Y_{i,j}/2}\right).

The first sum can be handled by (A.34) to give

i,j=1Yijs1/2=3(3s1/21)ζ(s12)L3(s12).\sum_{i,j=-\infty}^{\infty}\frac{1}{Y_{ij}^{s-1/2}}=3(3^{s-1/2}-1)\zeta\left(s-\frac{1}{2}\right)L_{-3}\left(s-\frac{1}{2}\right).

For the other sum, observe that

3Yi,j=3i2+3ij+3j2+3i+3j+1,3Y_{i,j}=3i^{2}+3ij+3j^{2}+3i+3j+1,

that is to say 3Yi,j3Y_{i,j} is a positive integer and 3Yi,j1(mod3)3Y_{i,j}\equiv 1\pmod{3}. Therefore we set 3Yi,j=3N+13Y_{i,j}=3N+1 and use (A.19) to deduce that the number of solutions of

3i2+3ij+3j2+3i+3j+1=3N+13i^{2}+3ij+3j^{2}+3i+3j+1=3N+1

is equal to 12u2(3N+1)\frac{1}{2}u_{2}(3N+1).

Taking all of the above into account, we finally obtain

S2(s)\displaystyle S_{2}(s) =27π8Γ(s12)Γ(s)(3s1/21)ζ(s12)L3(s12)\displaystyle=\sqrt{\frac{27\pi}{8}}\;\frac{\Gamma(s-\frac{1}{2})}{\Gamma(s)}\;(3^{s-1/2}-1)\;\zeta\left(s-\frac{1}{2}\right)L_{-3}\left(s-\frac{1}{2}\right)
+2πsΓ(s)(38)(2s+1)/4k=1N=0(1)ku2(3N+1)\displaystyle\quad+\frac{2\pi^{s}}{\Gamma(s)}\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{k=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{k}\,u_{2}(3N+1)
×(kN+13)s12Ks12(πk3N+12).\displaystyle\qquad\qquad\qquad\times\left(\frac{k}{\sqrt{N+\frac{1}{3}}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(\pi k\sqrt{\frac{3N+1}{2}}\right). (5.6)

5.5. The lattice sum for hexagonal close packing

On adding the results for S1(s)S_{1}(s) and S2(s)S_{2}(s) in (5.2) and (5.5) we obtain

L3HCP(s)\displaystyle L_{3}^{\text{HCP}}(s) =6ζ(s)L3(s)+4π3(32)s1(1s1)ζ(2s2)\displaystyle=6\zeta(s)\,L_{-3}(s)+\frac{4\pi}{\sqrt{3}}\left(\frac{3}{2}\right)^{s-1}\left(\frac{1}{s-1}\right)\zeta(2s-2)
+83πsΓ(s)k=1N=1u2(N)(N2k2)(s1)/2Ks1(8πk32N)\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=1}^{\infty}\sum_{N=1}^{\infty}u_{2}(N)\left(\frac{N}{2k^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi k}{3}\sqrt{2N}\right)
+83πsΓ(s)k=0N=1cos2πN3u2(N)\displaystyle\quad+\frac{8}{\sqrt{3}}\,\frac{\pi^{s}}{\Gamma(s)}\sum_{k=0}^{\infty}\sum_{N=1}^{\infty}\cos\frac{2\pi N}{3}\,u_{2}(N)\,
×(N2(k+12)2)(s1)/2Ks1(8π(k+12)32N).\displaystyle\qquad\qquad\qquad\times\left(\frac{N}{2(k+\frac{1}{2})^{2}}\right)^{(s-1)/2}K_{s-1}\left(\frac{8\pi\left(k+\frac{1}{2}\right)}{3}\sqrt{2N}\right). (5.7)

On the other hand, if we add the results of (5.3) and (5.6) we obtain

L3HCP(s)\displaystyle L_{3}^{\text{HCP}}(s) =2(38)sζ(2s)+27π8Γ(s12)Γ(s)(3s1/2+1)ζ(s12)L3(s12)\displaystyle=2\left(\frac{3}{8}\right)^{s}\zeta(2s)+\sqrt{\frac{27\pi}{8}}\;\frac{\Gamma(s-\frac{1}{2})}{\Gamma(s)}\;(3^{s-1/2}+1)\;\zeta\left(s-\frac{1}{2}\right)L_{-3}\left(s-\frac{1}{2}\right)
+4πsΓ(s)(38)(2s+1)/4N=1k=1u2(N)(kN)s1/2Ks12(πk3N2)\displaystyle\quad+\frac{4\pi^{s}}{\Gamma(s)}\,\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{N=1}^{\infty}\sum_{k=1}^{\infty}u_{2}(N)\,\left(\frac{k}{\sqrt{N}}\right)^{s-1/2}K_{s-\frac{1}{2}}\left(\pi k\sqrt{\frac{3N}{2}}\right)
+2πsΓ(s)(38)(2s+1)/4k=1N=0(1)ku2(3N+1)\displaystyle\quad+\frac{2\pi^{s}}{\Gamma(s)}\left(\frac{3}{8}\right)^{(2s+1)/4}\sum_{k=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{k}\,u_{2}(3N+1)
×(kN+13)s12Ks12(πk3N+12)\displaystyle\qquad\qquad\qquad\times\left(\frac{k}{\sqrt{N+\frac{1}{3}}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(\pi k\sqrt{\frac{3N+1}{2}}\right) (5.8)

6. Analytic continuations of the lattice sums L(A;s)L(A;s) and L3HCP(s)L_{3}^{\text{HCP}}(s)

We will now show that the lattice sums L(A;s)L(A;s) and L3HCP(s)L_{3}^{\text{HCP}}(s) can be continued analytically to the whole ss-plane, and that the resulting functions have a single simple pole at s=3/2s=3/2 and no other singularities. We do this in steps. First, in Section 6.1 we show that the lattice sums each have a simple pole at s=3/2s=3/2 and determine the residue. Then, in Section 6.2 we show that the analytic continuations obtained are valid for the whole ss-plane and there are no other singularities. Finally, in Sections 6.36.5, values of the analytic continuations at the points s=1/2s=1/2 and s=1, 0,1,2,s=1,\,0,\,-1,\,-2,\ldots are computed. In particular, the evaluation of T2(A;s)T_{2}(A;s) at s=1/2s=1/2 in the case A=1A=1 gives the Madelung constant, e.g., see [4],  [5, pp. xiii, 39–51][17].

6.1. Behaviour of the lattice sums at s=3/2s=3/2

We start by showing that L(A;s)L(A;s) has a simple pole at s=3/2s=3/2 and determine the residue. In the formula (4.14), all of the terms are analytic at s=3/2s=3/2 except for the term involving ζ(2s2)\zeta(2s-2). It follows that

lims3/2(s3/2)L(A;s)\displaystyle\lim_{s\rightarrow 3/2}(s-3/2)L(A;s) =lims3/2(s3/2)πAs1(1+1A)sζ(2s2)\displaystyle=\lim_{s\rightarrow 3/2}(s-3/2)\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2)
=2πA(1+1A)3/2lims3/2(s3/2)ζ(2s2)\displaystyle=2\pi A\left(1+\frac{1}{A}\right)^{3/2}\lim_{s\rightarrow 3/2}(s-3/2)\zeta(2s-2)
=2πA(A+1)3/2×12limu1(u1)ζ(u)\displaystyle=\frac{2\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\times\frac{1}{2}\;\lim_{u\rightarrow 1}(u-1)\zeta(u)
=πA(A+1)3/2\displaystyle=\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}

where (A.23) was used in the last step of the calculation. It follows further that L(A;s)L(A;s) has a simple pole at s=3/2s=3/2 and the residue is given by

Res(L(A;s),3/2)=πA(A+1)3/2.\text{Res}(L(A;s),3/2)=\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}.

By (2.8) this is just 12 times the packing density, i.e.,

Res(L(A;s),3/2)=12Δ.\text{Res}(L(A;s),3/2)=12\Delta_{\mathcal{L}}.

For example, taking A=1A=1 gives

Res(L3FCC(s),3/2)=22π\text{Res}(L_{3}^{\text{FCC}}(s),3/2)=2\sqrt{2}\,\pi (6.1)

while taking A=1/2A=1/2 gives

Res(L3BCC(s),3/2)=33π/2.\text{Res}(L_{3}^{\text{BCC}}(s),3/2)=3\sqrt{3}\,\pi/2.

Laurent’s theorem implies there is an expansion of the form

L(A;s)=c1s3/2+c0+n=1cn(s3/2)nL(A;s)=\frac{c_{-1}}{s-3/2}+c_{0}+\sum_{n=1}^{\infty}c_{n}(s-3/2)^{n} (6.2)

where

c1=Res(L(A;s),3/2)=πA(A+1)3/2c_{-1}=\text{Res}(L(A;s),3/2)=\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}

and the coefficients c0c_{0}, c1c_{1}, c2,c_{2},\ldots depend on AA but not on ss. To calculate c0c_{0}, start with the fact that

lims3/2(πAs1(1+1A)sζ(2s2)c1s3/2)\lim_{s\rightarrow 3/2}\left(\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2)-\frac{c_{-1}}{s-3/2}\right)
=πA(A+1)3/2(2γ2+log(1+1A))=\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\left(2\gamma-2+\log\left(1+\frac{1}{A}\right)\right)

where γ=0.57721 56649 01532 86060\gamma=0.57721\,56649\,01532\,86060\,\cdots is Euler’s constant. Then use (4.14) and (A.5) to deduce

c0\displaystyle c_{0} =lims3/2(L(A;s)c1s3/2)\displaystyle=\lim_{s\rightarrow 3/2}\left(L(A;s)-\frac{c_{-1}}{s-3/2}\right)
=2(A+1)3/2ζ(32)L4(32)\displaystyle=\sqrt{2}\left(A+1\right)^{3/2}\zeta\left(\frac{3}{2}\right)L_{-4}\left(\frac{3}{2}\right)
+πA(A+1)3/2(2γ2+log(1+1A))\displaystyle\quad+\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\left(2\gamma-2+\log\left(1+\frac{1}{A}\right)\right)
+2πA(A+1)3/2k=1N=11kr2(N)exp(2πkAN)\displaystyle\quad+\frac{2\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\sum_{k=1}^{\infty}\sum_{N=1}^{\infty}\frac{1}{k}\,r_{2}(N)\exp\left(-2\pi k\sqrt{AN}\right)
+2πA(A+1)3/2k=1N=0(1)kkr2(4N+1)exp(2πkA(2N+12)).\displaystyle\quad+\frac{2\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\sum_{k=1}^{\infty}\sum_{N=0}^{\infty}\frac{(-1)^{k}}{k}\,r_{2}(4N+1)\,\exp\left(-2\pi k\sqrt{A\left(2N+\frac{1}{2}\right)}\right).
Interchanging the order of summation and evaluating the sum over kk gives
c0\displaystyle c_{0} =2(A+1)3/2ζ(32)L4(32)\displaystyle=\sqrt{2}\left(A+1\right)^{3/2}\zeta\left(\frac{3}{2}\right)L_{-4}\left(\frac{3}{2}\right)
+πA(A+1)3/2(2γ2+log(1+1A))\displaystyle\quad+\frac{\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\left(2\gamma-2+\log\left(1+\frac{1}{A}\right)\right)
2πA(A+1)3/2N=1r2(N)log(1e2πAN)\displaystyle\quad-\frac{2\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\sum_{N=1}^{\infty}r_{2}(N)\log\left(1-e^{-2\pi\sqrt{AN}}\right)
2πA(A+1)3/2N=0r2(4N+1)log(1+eπ2A(4N+1)).\displaystyle\quad-\frac{2\pi}{\sqrt{A}}\left(A+1\right)^{3/2}\sum_{N=0}^{\infty}r_{2}(4N+1)\,\log\left(1+e^{-\pi\sqrt{2A\left(4N+1\right)}}\right).

Numerical evaluation in the case A=1A=1 gives

c0|A=1=6.98405 25503 22247 93406.\left.c_{0}\right|_{A=1}=6.98405\,25503\,22247\,93406\cdots. (6.3)

A similar analysis can be given for L3HCP(s)L_{3}^{\text{HCP}}(s) using (5.7). We omit the details of the calculations as they are similar to the above. The end result is a Laurent expansion of the form

L3HCP(s)=d1s3/2+d0+n=1dn(s3/2)nL_{3}^{\text{HCP}}(s)=\frac{d_{-1}}{s-3/2}+d_{0}+\sum_{n=1}^{\infty}d_{n}(s-3/2)^{n} (6.4)

where

d1=Res(L3HCP(s),3/2)=22πd_{-1}=\text{Res}(L_{3}^{\text{HCP}}(s),3/2)=2\sqrt{2}\pi (6.5)

and

d0\displaystyle d_{0} =6ζ(32)L3(32)+ 22π(2γ2+log32)\displaystyle=6\zeta\left(\frac{3}{2}\right)L_{-3}\left(\frac{3}{2}\right)\;+\;2\,\sqrt{2}\,\pi\left(2\gamma-2+\log\frac{3}{2}\right)
+22πk=1N=11ku2(N)exp(83πk2N)\displaystyle\quad+2\sqrt{2}\pi\sum_{k=1}^{\infty}\sum_{N=1}^{\infty}\frac{1}{k}\,u_{2}(N)\exp\left(-\frac{8}{3}\pi k\sqrt{2N}\right)
+22πk=0N=1cos(2πN3)(k+12)u2(N)exp(83π(k+12)2N)\displaystyle\quad+2\sqrt{2}\pi\sum_{k=0}^{\infty}\sum_{N=1}^{\infty}\frac{\cos\left(\frac{2\pi N}{3}\right)}{(k+\frac{1}{2})}\,u_{2}(N)\,\exp\left(-\frac{8}{3}\pi\left(k+\frac{1}{2}\right)\,\sqrt{2N}\right)
=6ζ(32)L3(32)+ 22π(2γ2+log32)\displaystyle=6\zeta\left(\frac{3}{2}\right)L_{-3}\left(\frac{3}{2}\right)\;+\;2\,\sqrt{2}\,\pi\left(2\gamma-2+\log\frac{3}{2}\right)
22πN=1u2(N)log(1e83π2N)\displaystyle\quad-2\sqrt{2}\pi\sum_{N=1}^{\infty}u_{2}(N)\log\left(1-e^{-\frac{8}{3}\pi\sqrt{2N}}\right)
+22πN=1cos(2πN3)u2(N)log(1+e43π2N1e43π2N)\displaystyle\quad+2\sqrt{2}\pi\sum_{N=1}^{\infty}\cos\left(\frac{2\pi N}{3}\right)\,u_{2}(N)\,\log\left(\frac{1+e^{-\frac{4}{3}\pi\sqrt{2N}}}{1-e^{-\frac{4}{3}\pi\sqrt{2N}}}\right)
=6.98462 37414 38416 61307.\displaystyle=6.98462\;37414\;38416\;61307\;\cdots. (6.6)

In particular, the pole of L3HCP(s)L_{3}^{\text{HCP}}(s) at s=3/2s=3/2 is simple. By (6.1) and (6.5) we have

Res(L3HCP(s),3/2)=Res(L3FCC(s),3/2).\text{Res}(L_{3}^{\text{HCP}}(s),3/2)=\text{Res}(L_{3}^{\text{FCC}}(s),3/2).

It follows that the difference L3FCC(s)L3HCP(s)L_{3}^{\text{FCC}}(s)-L_{3}^{\text{HCP}}(s) has a removable singularity at s=3/2s=3/2 and from the Laurent expansions we deduce that

lims32(L3FCC(s)L3HCP(s))=c0|A=1d0.\lim_{s\rightarrow\frac{3}{2}}\left(L_{3}^{\text{FCC}}(s)-L_{3}^{\text{HCP}}(s)\right)=\left.c_{0}\right|_{A=1}-d_{0}.

Using the numerical values from (6.3) and (6.6) we obtain

lims32(L3FCC(s)L3HCP(s))=0.00057 11911 16168 67901.\lim_{s\rightarrow\frac{3}{2}}\left(L_{3}^{\text{FCC}}(s)-L_{3}^{\text{HCP}}(s)\right)=-0.00057\;11911\;16168\;67901\cdots.

This gives the value at the left hand end of the graph in [6, Fig. 3]. The value s=3/2s=3/2 used here corresponds to taking s=3s=3 in [6] because of the different way the exponents are used in the definitions.

6.2. Analyticity of the lattice sums at other values of ss

By (A.6), the double series of Bessel functions in (4.14) converges absolutely and uniformly on compact subsets of the ss-plane and therefore represents an entire function of ss. It follows that L(A;s)L(A;s) has an analytic continuation to a meromorphic function which is analytic except possibly at the singularities of the terms

4(A+12)sζ(s)L4(s)4\left(\frac{A+1}{2}\right)^{s}\zeta(s)L_{-4}(s) (6.7)

and

πAs1(1+1A)sζ(2s2).\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2). (6.8)

The function in (6.7) is analytic except at s=1s=1 due to the pole of ζ(s)\zeta(s), as the function L4(s)L_{-4}(s) and the exponential function are both entire. The function in (6.8) is analytic except at s=1s=1 and s=3/2s=3/2. We studied the singularity at s=3/2s=3/2 in the previous section, so this leaves only the point s=1s=1. Using (A.23) and the values of ζ(0)\zeta(0) and L4(1)L_{-4}(1) in (A.26) and (A.27) we find that

4(A+12)sζ(s)L4(s)=(A+1)π2(s1)+O(1)as s14\left(\frac{A+1}{2}\right)^{s}\zeta(s)L_{-4}(s)=\frac{(A+1)\pi}{2(s-1)}+O(1)\quad\text{as $s\rightarrow 1$}

and

πAs1(1+1A)sζ(2s2)=(A+1)π2(s1)+O(1)as s1.\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2)=-\frac{(A+1)\pi}{2(s-1)}+O(1)\quad\text{as $s\rightarrow 1$}.

It follows that the sum of the functions in (6.7) and (6.8) has a removable singularity at s=1s=1 and so L(A;s)L(A;s) is also analytic at s=1s=1. The analyticity at s=1s=1 can also be seen directly from the alternative formula for L(A;s)L(A;s) in (4.15).

In conclusion, it has been shown that L(A;s)L(A;s) has an analytic continuation to a meromorphic function of ss which has a simple pole at s=3/2s=3/2 and no other singularities. Because L(A;s)L(A;s) has only one singularity, namely s=3/2s=3/2, the Laurent expansion (6.2) is valid in the annulus 0<|s3/2|<0<|s-3/2|<\infty, i.e., for all s3/2s\neq 3/2.

In a similar way, (5.7) and (5.8) can be used to show that L3HCP(s)L_{3}^{\text{HCP}}(s) also has an analytic continuation to a meromorphic function of ss which has a simple pole at s=3/2s=3/2 and no other singularities. The Laurent expansion (6.4) converges for all s3/2s\neq 3/2.

By the theory of complex variables, the analytic continuation, if one exists, is unique, e.g., see [16, p. 147, Th. 1]. Therefore analytic continuation formulas can be used to assign values to divergent series. For example, the Madelung constant is defined by

M=i,j,k(1)i+j+k(i2+j2+k2)s|s=1/2.M={\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{(-1)^{i+j+k}}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}. (6.9)

This is interpreted as being the value of the analytic continuation of the series at s=1/2s=1/2, because

i,j,k(1)i+j+k(i2+j2+k2)s{\sum_{i,j,k}}^{\prime}\;\;\;\frac{(-1)^{i+j+k}}{(i^{2}+j^{2}+k^{2})^{s}}

obviously diverges if s=1/2s=1/2. From now on, we shall use the expression “the value of a series at a point ss” to mean “the value of the analytic continuation of the series at the point ss”.

6.3. Values at s=1/2s=1/2 and the Madelung constant

On putting s=1/2s=1/2 in (4.11) we obtain an analytic expression for the value of

M(A)=i,j,k(1)i+j+k(Ai2+j2+k2)s|s=1/2\displaystyle M(A)={\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}
which specialises to the Madelung constant in the case A=1A=1. We have
M(A)\displaystyle M(A) =4(121s)ζ(s)L4(s)|s=1/2\displaystyle=-4(1-2^{1-s})\zeta(s)L_{-4}(s)\Bigg{|}_{s=1/2}
+4πsΓ(s)A(1s)/2i=1N=0(1)ir2(4N+1)(2N+12i2)(s1)/2\displaystyle+\frac{4\pi^{s}}{\Gamma(s)}\,A^{(1-s)/2}\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)\left(\frac{2N+\frac{1}{2}}{i^{2}}\right)^{(s-1)/2}
×Ks1(2πiA(2N+12))|s=1/2.\displaystyle\qquad\qquad\qquad\qquad\qquad\times K_{s-1}\left(2\pi i\sqrt{A(2N+\frac{1}{2})}\,\right)\Bigg{|}_{s=1/2}.

Now use (A.4) and (A.5) to express the Bessel functions in terms of exponential functions. The result simplifies to

M(A)\displaystyle M(A) =4(21)ζ(12)L4(12)+2i=1N=0(1)ir2(4N+1)2N+12e2πiA(2N+1/2).\displaystyle=4(\sqrt{2}-1)\zeta\left(\frac{1}{2}\right)L_{-4}\left(\frac{1}{2}\right)+2\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}\,\frac{r_{2}(4N+1)}{\sqrt{2N+\frac{1}{2}}}\,e^{-2\pi i\sqrt{A(2N+1/2)}}.
On interchanging the order of summation and summing the geometric series, we obtain
M(A)\displaystyle M(A) =4(21)ζ(12)L4(12)22N=0r2(4N+1)4N+1(1eπ2A(4N+1)+1).\displaystyle=4(\sqrt{2}-1)\zeta\left(\frac{1}{2}\right)L_{-4}\left(\frac{1}{2}\right)-2\sqrt{2}\sum_{N=0}^{\infty}\frac{r_{2}(4N+1)}{\sqrt{4N+1}}\left(\frac{1}{e^{\pi\sqrt{2A(4N+1)}}+1}\right).

When A=1A=1 this gives the Madelung constant defined by (6.9). Numerical evaluation gives

M=M(1)=1.74756 45946 33182 19063M=M(1)=-1.74756\;45946\;33182\;19063\cdots (6.10)

which is in agreement with [5, p. xiii] (apart from the minus sign which we have corrected here) and matches the value of d(1)d(1) in [5, pp 39–51].


In a similar way, starting from (4.7) and using (A.5) and (A.26) we obtain

i,j,k1(Ai2+j2+k2)s|s=1/2\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}
=4ζ(12)L4(12)+πA3+2i=1N=1r2(N)Ne2πiAN\displaystyle=4\zeta\left(\frac{1}{2}\right)L_{-4}\left(\frac{1}{2}\right)+\frac{\pi\sqrt{A}}{3}+2\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}\frac{r_{2}(N)}{\sqrt{N}}\,e^{-2\pi i\sqrt{AN}}
=4ζ(12)L4(12)+πA3+2N=1r2(N)N(1e2πAN1).\displaystyle=4\zeta\left(\frac{1}{2}\right)L_{-4}\left(\frac{1}{2}\right)+\frac{\pi\sqrt{A}}{3}+2\sum_{N=1}^{\infty}\frac{r_{2}(N)}{\sqrt{N}}\left(\frac{1}{e^{2\pi\sqrt{AN}}-1}\right). (6.11)

Numerical evaluation in the case A=1A=1 gives

i,j,k1(i2+j2+k2)s|s=1/2=2.83729 74794 80619 47666.{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}=-2.83729\;74794\;80619\;47666\cdots. (6.12)

Now, from (4.4) we have

L3FCC(1/2)\displaystyle L_{3}^{\text{FCC}}(1/2) =12i,j,k1(i2+j2+k2)s|s=1/2+12i,j,k(1)i+j+k(i2+j2+k2)s|s=1/2.\displaystyle=\frac{1}{\sqrt{2}}{\sum_{i,j,k}}^{\prime}\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}+\frac{1}{\sqrt{2}}{\sum_{i,j,k}}^{\prime}\left.\frac{(-1)^{i+j+k}}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}.
Hence, using the values from (6.10) and (6.12) we obtain
L3FCC(1/2)\displaystyle L_{3}^{\text{FCC}}(1/2) =3.24198 70634 10888 39428.\displaystyle=-3.24198\;70634\;10888\;39428\cdots.

We also record the result

L3HCP(1/2)\displaystyle L_{3}^{\text{HCP}}(1/2) =6ζ(12)L3(12)+22π9+2N=1u2(N)N(1e8π2N/31)\displaystyle=6\zeta\left(\frac{1}{2}\right)L_{-3}\left(\frac{1}{2}\right)+\frac{2\sqrt{2}\pi}{9}+2\sum_{N=1}^{\infty}\frac{u_{2}(N)}{\sqrt{N}}\left(\frac{1}{e^{8\pi\sqrt{2N}/3}-1}\right)
+2N=1cos(2πN3)u2(N)N(1e4π2N/3e4π2N/3)\displaystyle\qquad+2\sum_{N=1}^{\infty}\cos\left(\frac{2\pi N}{3}\right)\frac{u_{2}(N)}{\sqrt{N}}\left(\frac{1}{e^{4\pi\sqrt{2N}/3}-e^{-4\pi\sqrt{2N}/3}}\right)
=3.24185 86150 75732 86473\displaystyle=-3.24185\;86150\;75732\;86473\cdots

which is obtained in the same way, starting from (5.7).

6.4. The value at s=1s=1

It was noted above that (4.14), which involves Ks1K_{s-1} Bessel functions, contains terms with singularities at s=1s=1 and therefore is not suitable for calculations at that value of ss. Instead we can use (4.15), which involves Ks1/2K_{s-1/2} Bessel functions. As in the previous section, two steps are involved. First, the the K1/2K_{1/2} Bessel functions can be expressed in terms of the exponential function by (A.5). Then, the double sum can be reduced to a single sum by geometric series. We omit the details and just record the final results and corresponding numerical values.

From (4.9) we have

i,j,k1(Ai2+j2+k2)s|s=1\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(Ai^{2}+j^{2}+k^{2})^{s}}\right|_{s=1}
=π23A+4πAζ(12)L4(12)+2πAN=1r2(N)N(1e2πN/A1)\displaystyle=\frac{\pi^{2}}{3A}+\frac{4\pi}{\sqrt{A}}\zeta\left(\frac{1}{2}\right)L_{-4}\left(\frac{1}{2}\right)+\frac{2\pi}{\sqrt{A}}\sum_{N=1}^{\infty}\frac{r_{2}(N)}{\sqrt{N}}\left(\frac{1}{e^{2\pi\sqrt{N/A}}-1}\right) (6.13)
while (4.13) gives
i,j,k(1)i+j+k(Ai2+j2+k2)s|s=1\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{(-1)^{i+j+k}}{(Ai^{2}+j^{2}+k^{2})^{s}}\right|_{s=1}
=π26A+2πAN=1(1)Nr2(N)N(1eπN/AeπN/A).\displaystyle=\frac{-\pi^{2}}{6A}+\frac{2\pi}{\sqrt{A}}\sum_{N=1}^{\infty}(-1)^{N}\,\frac{r_{2}(N)}{\sqrt{N}}\left(\frac{1}{e^{\pi\sqrt{N/A}}-e^{-\pi\sqrt{N/A}}}\right).

Then (4.4) can be used to write down the value of L(A;s)L(A;s).

For example, when A=1A=1 the above formulas give

i,j,k1(i2+j2+k2)s|s=1\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1} =8.91363 29175 85151 27268\displaystyle=-8.91363\;29175\;85151\;27268\cdots (6.14)
and
i,j,k(1)i+j+k(i2+j2+k2)s|s=1\displaystyle{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{(-1)^{i+j+k}}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1} =2.51935 61520 89445 31334.\displaystyle=-2.51935\;61520\;89445\;31334\cdots.

Then taking A=1A=1 and s=1s=1 in (4.4) gives

L3FCC(1)\displaystyle L_{3}^{\text{FCC}}(1) =i,j,k1(i2+j2+k2)s|s=1+i,j,k(1)i+j+k(i2+j2+k2)s|s=1\displaystyle={\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1}+{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{(-1)^{i+j+k}}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1}
=11.43298 90696 74596 58602.\displaystyle=-11.43298\;90696\;74596\;58602\cdots.

For HCP, the formula (5.7) cannot be used to evaluate L3HCP(1)L_{3}^{\text{HCP}}(1) because two of the terms have cancelling singularities at s=1s=1. Therefore we take s=1s=1 in (5.8) instead to obtain

L3HCP(1)\displaystyle L_{3}^{\text{HCP}}(1) =π28+π278(3+1)ζ(12)L3(12)\displaystyle=\frac{\pi^{2}}{8}+\pi\;\sqrt{\frac{27}{8}}\;(\sqrt{3}+1)\;\zeta\left(\frac{1}{2}\right)\,L_{-3}\left(\frac{1}{2}\right)
+π32N=1u2(N)N(1eπ3N/21)\displaystyle\quad+\pi\,\sqrt{\frac{3}{2}}\,\sum_{N=1}^{\infty}\frac{u_{2}(N)}{\sqrt{N}}\,\left(\frac{1}{e^{\pi\sqrt{3N/2}}-1}\right)
3π8N=0u2(3N+1)3N+1(1eπ(3N+1)/2+1)\displaystyle\quad-\frac{3\pi}{\sqrt{8}}\,\sum_{N=0}^{\infty}\frac{u_{2}(3N+1)}{\sqrt{3N+1}}\left(\frac{1}{e^{\pi\sqrt{(3N+1)/2}}+1}\right)
=11.43265 30014 95285 63572.\displaystyle=-11.43265\;30014\;95285\;63572\cdots.

We end this section by noting a connection between two of the values in the above analysis. By setting A=1A=1 in each of (6.11) and (6.13) we obtain the remarkable result

i,j,k1(i2+j2+k2)s|s=1=πi,j,k1(i2+j2+k2)s|s=1/2.{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1}=\pi\;{\sum_{i,j,k}}^{\prime}\;\;\;\left.\frac{1}{(i^{2}+j^{2}+k^{2})^{s}}\right|_{s=1/2}. (6.15)

This is consistent with [5, p. 46 (1.3.44)] and is the special case s=1s=1 of the functional equation

πsΓ(s)T1(1;s)=π(32s)Γ(32s)T1(1;32s).\pi^{-s}\Gamma(s)T_{1}(1;s)=\pi^{-(\frac{3}{2}-s)}\Gamma\left(\frac{3}{2}-s\right)T_{1}\left(1;\frac{3}{2}-s\right). (6.16)

This functional equation can be deduced from the two formulas for T1(A;s)T_{1}(A;s) in (4.7) and (4.9), as follows. Replace ss with 32s\frac{3}{2}-s in (4.7), then multiply by πs32Γ(32s)\pi^{s-\frac{3}{2}}\Gamma(\frac{3}{2}-s) and set A=1A=1 to get

πs32Γ(32s)T1(1;32s)\displaystyle\pi^{s-\frac{3}{2}}\Gamma\left(\frac{3}{2}-s\right)T_{1}\left(1;\frac{3}{2}-s\right)
=4πs32Γ(32s)ζ(32s)L4(32s)+2πs12Γ(12s)ζ(12s)\displaystyle=4\pi^{s-\frac{3}{2}}\Gamma\left(\frac{3}{2}-s\right)\zeta\left(\frac{3}{2}-s\right)L_{-4}\left(\frac{3}{2}-s\right)+2\pi^{s-\frac{1}{2}}\Gamma\left(\frac{1}{2}-s\right)\zeta(1-2s)
+4i=1N=1r2(N)(Ni2)(12s)/2K12s(2πiN),\displaystyle\quad+4\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{N}{i^{2}}\right)^{(\frac{1}{2}-s)/2}K_{\frac{1}{2}-s}\left(2\pi i\sqrt{N}\right),

where we have used the functional equation for the gamma function in the form

Γ(3/2s)=(1/2s)Γ(1/2s)\Gamma(3/2-s)=(1/2-s)\Gamma(1/2-s)

to obtain the second term on the right hand side. Now apply the functional equations (A.4), (A.24) and (A.25) to deduce

π(32s)Γ(32s)T1(1;32s)\displaystyle\pi^{-(\frac{3}{2}-s)}\Gamma\left(\frac{3}{2}-s\right)T_{1}\left(1;\frac{3}{2}-s\right)
=4π12sΓ(s12)ζ(s12)L4(s12)+2πsΓ(s)ζ(2s)\displaystyle=4\pi^{\frac{1}{2}-s}\,\Gamma\left(s-\frac{1}{2}\right)\zeta\left(s-\frac{1}{2}\right)L_{-4}\left(s-\frac{1}{2}\right)+2\pi^{-s}\,\Gamma(s)\zeta(2s)
+4i=1N=1r2(N)(iN)s12Ks12(2πiN).\displaystyle\quad+4\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{i}{\sqrt{N}}\right)^{s-\frac{1}{2}}K_{s-\frac{1}{2}}\left(2\pi i\sqrt{N}\right).

The functional equation (6.16) follows from this by using (4.9). In addition to providing another proof of the functional equation, the calculation above also demonstrates the interconnection between the formulas (4.7) and (4.9). Further functional equations of this type are considered in [5, p. 46].

6.5. Values at s=0,1,2,3,s=0,\,-1,\,-2,\,-3,\ldots

Recall from (4.14) that

L(A;s)\displaystyle L(A;s) =4(A+12)sζ(s)L4(s)+πAs1(1+1A)sζ(2s2)\displaystyle=4\left(\frac{A+1}{2}\right)^{s}\zeta(s)L_{-4}(s)+\frac{\pi A}{s-1}\left(1+\frac{1}{A}\right)^{s}\zeta(2s-2)
+2πsAΓ(s)(A+1A)si=1N=1r2(N)(Ni2)(s1)/2Ks1(2πiAN)\displaystyle\quad+\frac{2\pi^{s}\sqrt{A}}{\Gamma(s)}\left(\sqrt{A}+\frac{1}{\sqrt{A}}\right)^{s}\sum_{i=1}^{\infty}\sum_{N=1}^{\infty}r_{2}(N)\left(\frac{N}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{AN}\right)
+2πsAΓ(s)(A+1A)si=1N=0(1)ir2(4N+1)\displaystyle\quad+\frac{2\pi^{s}\sqrt{A}}{\Gamma(s)}\left(\sqrt{A}+\frac{1}{\sqrt{A}}\right)^{s}\sum_{i=1}^{\infty}\sum_{N=0}^{\infty}(-1)^{i}r_{2}(4N+1)
×(2N+12i2)(s1)/2Ks1(2πiA(2N+12)).\displaystyle\qquad\qquad\times\left(\frac{2N+\frac{1}{2}}{i^{2}}\right)^{(s-1)/2}K_{s-1}\left(2\pi i\sqrt{A(2N+\frac{1}{2})}\right).

On using the values ζ(0)=12\zeta(0)=-\frac{1}{2}, ζ(2)=0\zeta(-2)=0, L4(0)=12L_{-4}(0)=\frac{1}{2} and the limiting value

lims01Γ(s)=0\lim_{s\rightarrow 0}\frac{1}{\Gamma(s)}=0

we readily obtain the result

L(A;0)=1.L(A;0)=-1.

Moreover, since

ζ(2)=ζ(4)=ζ(6)==0,\zeta(-2)=\zeta(-4)=\zeta(-6)=\cdots=0,
L4(1)=L4(3)=ζ(5)==0,L_{-4}(-1)=L_{-4}(-3)=\zeta(-5)=\cdots=0,

and

limsN1Γ(s)=0if N=0,1,2,\lim_{s\rightarrow N}\frac{1}{\Gamma(s)}=0\quad\text{if $N=0,\,-1,\,-2,\,\cdots$}

it follows that

L(A;1)=L(A;2)=L(A;3)==0.L(A;-1)=L(A;-2)=L(A;-3)=\cdots=0.

In a similar way, it can be shown using (5.7) that

L3HCP(0)=1L_{3}^{\text{HCP}}(0)=-1

and

L3HCP(1)=L3HCP(2)=L3HCP(3)==0.L_{3}^{\text{HCP}}(-1)=L_{3}^{\text{HCP}}(-2)=L_{3}^{\text{HCP}}(-3)=\cdots=0.

7. Graphs

The formulas  (4.14), (4.15), (5.7) and (5.8) have been used to produce the following graphs of y=L3FCC(s)y=L_{3}^{\text{FCC}}(s) on the intervals 10<s<10-10<s<10 and 7<s<0-7<s<0 in Figure 3. The graph of y=L3HCP(s)y=L_{3}^{\text{HCP}}(s) has a similar appearance, and so to allow a comparison the difference

y=L3HCP(s)L3FCC(s)y=L_{3}^{\text{HCP}}(s)-L_{3}^{\text{FCC}}(s)

is plotted using a finer vertical scale in Figure 4.

Refer to caption
Figure 3. Graph of y=L3FCC(s)y=L_{3}^{\text{FCC}}(s) for 10<s<10-10<s<10.
Refer to caption
Figure 4. Graph of y=L3FCC(s)y=L_{3}^{\text{FCC}}(s) for 7<s<0-7<s<0.
Refer to caption
Figure 5. Graph of y=L3HCP(s)L3FCC(s)y=L_{3}^{\text{HCP}}(s)-L_{3}^{\text{FCC}}(s).

The graphs appear to suggest the following:

Conjecture:

L3HCP(s)>L3FCC(s)>0\displaystyle L_{3}^{\text{HCP}}(s)>L_{3}^{\text{FCC}}(s)>0 fors(6,5)(4,3)(2,1)(3/2,)\displaystyle\quad\text{for}\quad s\in\cdots\cup(-6,-5)\cup(-4,-3)\cup(-2,-1)\cup(3/2,\infty)
L3HCP(s)<L3FCC(s)<0\displaystyle L_{3}^{\text{HCP}}(s)<L_{3}^{\text{FCC}}(s)<0 fors(5,4)(3,2)(1,0)\displaystyle\quad\text{for}\quad s\in\cdots\cup(-5,-4)\cup(-3,-2)\cup(-1,0)

and

1>L3HCP(s)>L3FCC(s)fors(0,3/2).-1>L_{3}^{\text{HCP}}(s)>L_{3}^{\text{FCC}}(s)\quad\text{for}\quad s\in(0,3/2).

Appendix A Formulas for special functions

Many results for special functions and analytic number theory have been used in this work. For clarity and ease of use, they are stated here along with references.

A.1. The gamma function

The gamma function may be defined for s>0s>0 by

Γ(s)=0ts1etdt.\Gamma(s)=\int_{0}^{\infty}t^{s-1}\,e^{-t}\,\mathrm{d}t. (A.1)

By the change of variable t=wxt=wx this can be rewritten in the useful form

1ws=1Γ(s)0xs1ewxdx.\frac{1}{w^{s}}=\frac{1}{\Gamma(s)}\int_{0}^{\infty}x^{s-1}\,e^{-wx}\,\mathrm{d}x. (A.2)

See [1, (1.1.18)].

A.2. The modified Bessel function

The following integral may be evaluated in terms of the modified Bessel function:

0xs1eaxb/xdx=2(ba)s/2Ks(2ab).\int_{0}^{\infty}x^{s-1}e^{-ax-b/x}\mathrm{d}x=2\left(\frac{b}{a}\right)^{s/2}K_{s}(2\sqrt{ab}). (A.3)

By the change of variable x=u1x=u^{-1} it can be shown that

Ks(z)=Ks(z).K_{s}(z)=K_{-s}(z). (A.4)

When s=1/2s=1/2 the modified Bessel function reduces to an elementary function:

K1/2(z)=π2zez.K_{1/2}(z)=\sqrt{\frac{\pi}{2z}}\,e^{-z}. (A.5)

The asymptotic formula holds:

Ks(z)π2zezasz,(|argz|<3π/2).K_{s}(z)\sim\sqrt{\frac{\pi}{2z}}\,e^{-z}\quad\text{as}\quad z\rightarrow\infty,\quad(\,|\text{arg}\,z|<3\pi/2). (A.6)

For all of these properties, see [1, pp. 223, 237] or [21, pp. 233–248].

A.3. Characters

For an integer nn, let χ4(n)\chi_{-4}(n) and χ3(n)\chi_{-3}(n) be defined by

χ4(n)=sin(πn/2)={1if n1(mod4),1if n3(mod4),0otherwise\chi_{-4}(n)=\sin(\pi n/2)=\begin{cases}1&\text{if $n\equiv 1\pmod{4}$,}\\ -1&\text{if $n\equiv 3\pmod{4}$,}\\ 0&\text{otherwise}\end{cases} (A.7)

and

χ3(n)=sin(2πn/3)sin(2π/3)={1if n1(mod3),1if n2(mod3),0otherwise.\chi_{-3}(n)=\frac{\sin(2\pi n/3)}{\sin(2\pi/3)}=\begin{cases}1&\text{if $n\equiv 1\pmod{3}$,}\\ -1&\text{if $n\equiv 2\pmod{3}$,}\\ 0&\text{otherwise.}\end{cases} (A.8)

A.4. Theta functions

The transformation formula for theta functions is [1, p. 119], [3, (2.2.5)]:

n=eπn2t+2πina=1tn=eπ(n+a)2/t,assuming Re(t)>0.\sum_{n=-\infty}^{\infty}e^{-\pi n^{2}t+2\pi ina}=\frac{1}{\sqrt{t}}\sum_{n=-\infty}^{\infty}e^{-\pi(n+a)^{2}/t},\quad\text{assuming Re$(t)>0$.} (A.9)

We will need the special cases a=0a=0 and a=1/2a=1/2, which are

n=eπn2t=1tn=eπn2/t\sum_{n=-\infty}^{\infty}e^{-\pi n^{2}t}=\frac{1}{\sqrt{t}}\sum_{n=-\infty}^{\infty}e^{-\pi n^{2}/t} (A.10)

and

n=(1)neπn2t=1tn=eπ(n+12)2/t\sum_{n=-\infty}^{\infty}(-1)^{n}e^{-\pi n^{2}t}=\frac{1}{\sqrt{t}}\sum_{n=-\infty}^{\infty}e^{-\pi(n+\frac{1}{2})^{2}/t} (A.11)

respectively. The sum of two squares formula is [11, (3.111)]

(j=qj2)2=j=k=qj2+k2=N=0r2(N)qN\left(\sum_{j=-\infty}^{\infty}q^{j^{2}}\right)^{2}=\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{j^{2}+k^{2}}=\sum_{N=0}^{\infty}r_{2}(N)q^{N} (A.12)

where

r2(N)=#{j2+k2=N}={1if N=0,4d|Nχ4(d)if N1,r_{2}(N)=\#\left\{j^{2}+k^{2}=N\right\}=\begin{cases}1&\text{if $N=0$},\\ \\ \displaystyle{4\sum_{d|N}\chi_{-4}(d)}&\text{if $N\geq 1$,}\end{cases} (A.13)

the sum being is over the positive divisors dd of NN. For example,

r2(18)\displaystyle r_{2}(18) =4(χ4(1)+χ4(2)+χ4(3)+χ4(6)+χ4(9)+χ4(18))\displaystyle=4\left(\chi_{-4}(1)+\chi_{-4}(2)+\chi_{-4}(3)+\chi_{-4}(6)+\chi_{-4}(9)+\chi_{-4}(18)\right)
=4(1+01+0+1+0)=4.\displaystyle=4\left(1+0-1+0+1+0\right)=4.

By [11, (3.15) and (3.111)] we also have

(j=q(j+12)2)2=N=0r2(4N+1)q(4N+1)/2.\left(\sum_{j=-\infty}^{\infty}q^{(j+\frac{1}{2})^{2}}\right)^{2}=\sum_{N=0}^{\infty}r_{2}(4N+1)q^{(4N+1)/2}. (A.14)

A.5. The cubic theta function

The cubic analogues of the transformation formula are [3, (2.2)], [10, Cor. 5.19]

j=k=e2π(j2+jk+k2)t=13j=k=e2π(j2+jk+k2)/3t\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}e^{-2\pi(j^{2}+jk+k^{2})t}=\frac{1}{\sqrt{3}}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}e^{-2\pi(j^{2}+jk+k^{2})/3t} (A.15)

and

j=k=e2π((j+13)2+(j+13)(k+13)+(k+13)2)t=13j=k=ωjke2π(j2+jk+k2)/3t\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}e^{-2\pi((j+\frac{1}{3})^{2}+(j+\frac{1}{3})(k+\frac{1}{3})+(k+\frac{1}{3})^{2})t}=\frac{1}{\sqrt{3}}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}\omega^{j-k}e^{-2\pi(j^{2}+jk+k^{2})/3t} (A.16)

where ω=exp(2πi/3)\omega=\exp(2\pi i/3) is a primitive cube root of unity. The analogue of the sum of two squares result is [11, (3.124)]

j=k=qj2+jk+k2=N=0u2(N)qN\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{j^{2}+jk+k^{2}}=\sum_{N=0}^{\infty}u_{2}(N)q^{N} (A.17)

where

u2(N)=#{j2+jk+k2=N}={1if N=0,6d|Nχ3(d)if N1,u_{2}(N)=\#\left\{j^{2}+jk+k^{2}=N\right\}=\begin{cases}1&\text{if $N=0$},\\ \\ \displaystyle{6\sum_{d|N}\chi_{-3}(d)}&\text{if $N\geq 1$,}\end{cases} (A.18)

where the sum is again over the positive divisors dd of NN. By [11, (3.18) and (3.124)] we also have

j=k=q(j+13)2+(j+13)(k+13)+(k+13)2=12N=0u2(3N+1)qN+13\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{(j+\frac{1}{3})^{2}+(j+\frac{1}{3})(k+\frac{1}{3})+(k+\frac{1}{3})^{2}}=\frac{1}{2}\sum_{N=0}^{\infty}u_{2}(3N+1)q^{N+\frac{1}{3}} (A.19)

which is the analogue of (A.14).

A.6. The Riemann zeta function and LL functions

The definitions are:

ζ(s)\displaystyle\zeta(s) =j=11js\displaystyle=\sum_{j=1}^{\infty}\frac{1}{j^{s}} (A.20)
L4(s)\displaystyle L_{-4}(s) =j=1χ4(j)js=113s+15s17s+.\displaystyle=\sum_{j=1}^{\infty}\frac{\chi_{-4}(j)}{j^{s}}=1-\frac{1}{3^{s}}+\frac{1}{5^{s}}-\frac{1}{7^{s}}+\cdots. (A.21)
L3(s)\displaystyle L_{-3}(s) =j=1χ3(j)js=112s+14s15s+17s18s+.\displaystyle=\sum_{j=1}^{\infty}\frac{\chi_{-3}(j)}{j^{s}}=1-\frac{1}{2^{s}}+\frac{1}{4^{s}}-\frac{1}{5^{s}}+\frac{1}{7^{s}}-\frac{1}{8^{s}}+\cdots. (A.22)

The function ζ(s)\zeta(s) is the Riemann zeta function. It has a pole of order 11 at s=1s=1, and in fact

lims1(s1)ζ(s)=1.\lim_{s\rightarrow 1}(s-1)\zeta(s)=1. (A.23)

This is a consequence of [1, (1.3.2)]. See also [21, p. 58].
We will require the functional equations

πs/2Γ(s/2)ζ(s)=π(1s)/2Γ((1s)/2)ζ(1s)\pi^{-s/2}\Gamma(s/2)\zeta(s)=\pi^{-(1-s)/2}\Gamma((1-s)/2)\zeta(1-s) (A.24)

and

πsΓ(s)ζ(s)L4(s)=π(1s)Γ(1s)ζ(1s)L4(1s)\pi^{-s}\Gamma\left(s\right)\zeta(s)L_{-4}(s)=\pi^{-(1-s)}\Gamma\left(1-s\right)\zeta(1-s)L_{-4}(1-s) (A.25)

and the special values

ζ(2)=π26,ζ(0)=12,ζ(1)=112,ζ(2)=ζ(4)=ζ(6)==0,\zeta(2)=\frac{\pi^{2}}{6},\quad\zeta(0)=-\frac{1}{2},\quad\zeta(-1)=-\frac{1}{12},\quad\zeta(-2)=\zeta(-4)=\zeta(-6)=\cdots=0, (A.26)
L4(1)=π4,L4(0)=12,L4(1)=L4(3)=L4(5)==0,L_{-4}(1)=\frac{\pi}{4},\quad L_{-4}(0)=\frac{1}{2},\quad L_{-4}(-1)=L_{-4}(-3)=L_{-4}(-5)=\cdots=0, (A.27)

and

L3(1)=π39,L3(0)=13,L3(1)=L3(3)=L3(5)==0.L_{-3}(1)=\frac{\pi\sqrt{3}}{9},\quad L_{-3}(0)=\frac{1}{3},\quad L_{-3}(-1)=L_{-3}(-3)=L_{-3}(-5)=\cdots=0. (A.28)

See [2, Ch. 12] or [22]. Other results used are

j=01(j+12)s=(2s1)ζ(s)\sum_{j=0}^{\infty}\frac{1}{(j+\frac{1}{2})^{s}}=(2^{s}-1)\zeta(s) (A.29)
j=1(1)jjs=(121s)ζ(s)\sum_{j=1}^{\infty}\frac{(-1)^{j}}{j^{s}}=-(1-2^{1-s})\zeta(s) (A.30)
j,k1(j2+k2)s=4ζ(s)L4(s){\sum_{j,k}}^{\prime}\;\;\;\frac{1}{(j^{2}+k^{2})^{s}}=4\zeta(s)L_{-4}(s) (A.31)
j,k(1)j+k(j2+k2)s=4(121s)ζ(s)L4(s).{\sum_{j,k}}^{\prime}\;\;\;\frac{(-1)^{j+k}}{(j^{2}+k^{2})^{s}}=-4(1-2^{1-s})\zeta(s)L_{-4}(s). (A.32)
i,j1(i2+ij+j2)s=6ζ(s)L3(s){\sum_{i,j}}^{\;\prime}\frac{1}{(i^{2}+ij+j^{2})^{s}}=6\zeta(s)L_{-3}(s) (A.33)
i,j1((i+13)2+(i+13)(j+13)+(j+13)2)s=3(3s1)ζ(s)L3(s).{\sum_{i,j}}\frac{1}{((i+\frac{1}{3})^{2}+(i+\frac{1}{3})(j+\frac{1}{3})+(j+\frac{1}{3})^{2})^{s}}=3(3^{s}-1)\zeta(s)L_{-3}(s). (A.34)

The identities (A.29) and (A.30) follow from the definition of ζ(s)\zeta(s) by series rearrangements. For (A.31), (A.32) and (A.33), see (1.4.14), (1.7.5) and (1.4.16), respectively, of [5]. The identity (A.34) can be obtained by the method of Mellin transforms (e.g., see Appendix A of [6]) starting with [11, (3.36)].

Appendix B Behaviour as A0+A\rightarrow 0^{+} and A+A\rightarrow+\infty

We briefly consider the behaviour of the lattices in the limiting cases A0+A\rightarrow 0^{+} and A+A\rightarrow+\infty. Some of the basis vectors become infinite in the limit, leaving a sublattice of lower dimension. We discuss each case A0+A\rightarrow 0^{+} and A+A\rightarrow+\infty both in terms of theta functions and then in terms of the basis vectors.

First, consider the limit A0+A\rightarrow 0^{+}. In the interval 0<A<1/30<A<1/3 the theta function is

θ(A;q)\displaystyle\theta(A;q) =i=j=k=qg(A;i,j,k)\displaystyle=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{g(A;i,j,k)}
=i=j=k=q(A(i+j)2+(j+k)2+(i+k)2)/4A.\displaystyle=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2})/4A}.

As A0+A\rightarrow 0^{+} we have

q(j+k)2/4A0andq(i+k)2/4A0q^{(j+k)^{2}/4A}\rightarrow 0\quad\text{and}\quad q^{(i+k)^{2}/4A}\rightarrow 0

unless j=kj=-k and i=ki=-k, respectively. Hence,

limA0+θ(A;q)\displaystyle\lim_{A\rightarrow 0^{+}}\theta(A;q) =limA0+k=(i=kj=kq(A(i+j)2+(j+k)2+(i+k)2)/4A)\displaystyle=\lim_{A\rightarrow 0^{+}}\sum_{k=-\infty}^{\infty}\left(\sum_{i=-k}\sum_{j=-k}q^{(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2})/4A}\right)
=limA0+k=qA(kk)2/4A\displaystyle=\lim_{A\rightarrow 0^{+}}\sum_{k=-\infty}^{\infty}q^{A(-k-k)^{2}/4A}
=k=qk2.\displaystyle=\sum_{k=-\infty}^{\infty}q^{k^{2}}.

This corresponds to the one-dimensional lattice with minimal distance 11. The kissing number is 2, which is in agreement with the other lattices in the range 0<A<1/30<A<1/3, as indicated in Table 1. In terms of the basis vectors, from (2.1) we have

𝒃1=(12,12A,0),𝒃2=(12,0,12A),𝒃3=(0,12A,12A).\boldsymbol{b}_{1}=\left(\frac{1}{2},\frac{1}{2\sqrt{A}},0\right)^{\top},\quad\boldsymbol{b}_{2}=\left(\frac{1}{2},0,\frac{1}{2\sqrt{A}}\right)^{\top},\quad\boldsymbol{b}_{3}=\left(0,\frac{1}{2\sqrt{A}},\frac{1}{2\sqrt{A}}\right)^{\top}.

The only linear combinations 𝒗=i𝒃1+j𝒃2+k𝒃3\boldsymbol{v}=i\boldsymbol{b}_{1}+j\boldsymbol{b}_{2}+k\boldsymbol{b}_{3} (for i,j,k𝐙i,j,k\in\mathbf{Z}) that remain finite in the limit A0+A\rightarrow 0^{+} occur when i=ki=-k, j=kj=-k in which case we obtain

𝒗=k𝒃1k𝒃2+k𝒃3=k(1,0,0).\boldsymbol{v}=-k\boldsymbol{b}_{1}-k\boldsymbol{b}_{2}+k\boldsymbol{b}_{3}=-k(1,0,0)^{\top}.

That is, the limiting lattice is just the one-dimensional lattice consisting of integer multiples of (1,0,0)(1,0,0)^{\top}.

Now consider the limit A+A\rightarrow+\infty. For A>1A>1 the theta function is

θ(A;q)\displaystyle\theta(A;q) =i=j=k=qg(A;i,j,k)\displaystyle=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{g(A;i,j,k)}
=i=j=k=q(A(i+j)2+(j+k)2+(i+k)2)/2.\displaystyle=\sum_{i=-\infty}^{\infty}\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2})/2}.

Since qA(i+j)2/20q^{A(i+j)^{2}/2}\rightarrow 0 as A+A\rightarrow+\infty unless i=ji=-j, it follows that

limA+θ(A;q)\displaystyle\lim_{A\rightarrow+\infty}\theta(A;q) =j=k=(i=jq(A(i+j)2+(j+k)2+(i+k)2)/2)\displaystyle=\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}\left(\sum_{i=-j}q^{(A(i+j)^{2}+(j+k)^{2}+(i+k)^{2})/2}\right)
=j=k=q((j+k)2+(j+k)2)/2\displaystyle=\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{((j+k)^{2}+(-j+k)^{2})/2}
=j=k=qj2+k2.\displaystyle=\sum_{j=-\infty}^{\infty}\sum_{k=-\infty}^{\infty}q^{j^{2}+k^{2}}.

This is the theta series for the two-dimensional square close packing lattice with minimal distance 11. The kissing number is 44, in agreement with other values in the range A>1A>1 given by Table 1. In terms of the basis vectors, from (2.1) we have

𝒃1=12(A,1,0),𝒃2=12(A,0,1),𝒃3=12(0,1,1).\boldsymbol{b}_{1}=\frac{1}{\sqrt{2}}(\sqrt{A},1,0)^{\top},\quad\boldsymbol{b}_{2}=\frac{1}{\sqrt{2}}(\sqrt{A},0,1)^{\top},\quad\boldsymbol{b}_{3}=\frac{1}{\sqrt{2}}(0,1,1)^{\top}.

The only linear combinations 𝒗=i𝒃1+j𝒃2+k𝒃3\boldsymbol{v}=i\boldsymbol{b}_{1}+j\boldsymbol{b}_{2}+k\boldsymbol{b}_{3} (for i,j,k𝐙i,j,k\in\mathbf{Z}) that remain finite in the limit A+A\rightarrow+\infty occur when i=ji=-j, in which case we obtain

𝒗=j𝒃1+j𝒃2+k𝒃3=12[j(0,1,1)+k(0,1,1)].\boldsymbol{v}=-j\boldsymbol{b}_{1}+j\boldsymbol{b}_{2}+k\boldsymbol{b}_{3}=\frac{1}{\sqrt{2}}\left[j(0,-1,1)^{\top}+k(0,1,1)^{\top}\right].

This is isomorphic to the two-dimensional square close packing lattice with minimal distance 11, rotated from the coordinate axes by 45 degrees.

References

  • [1] G. Andrews, R. Askey and R. Roy, Special Functions, Cambridge University Press, Cambridge, 1999.
  • [2] T. M. Apostol, Introduction to Analytic Number Theory, Springer, New York, 1976.
  • [3] J. M. Borwein and P. B. Borwein, A cubic counterpart of Jacobi’s identity and the AGM, Trans. Amer. Math. Soc. 323 (2) 691–701 (1991).
  • [4] D. Borwein, J. M. Borwein, and C. Pinner, Convergence of Madelung-like lattice sums, Trans. Am. Math. Soc. 350, 3131–3167 (1998).
  • [5] J. M. Borwein, M. L. Glasser, R. C. McPhedran, J. G. Wan, and I. J. Zucker, Lattice Sums Then and Now, Cambridge University Press, Cambridge, 2013.
  • [6] A. Burrows, S. Cooper, E. Pahl and P. Schwerdtfeger, Analytical methods for fast converging lattice sums for cubic and hexagonal close-packed structures, J. Math. Phys. 61, 123503 (2020), https://doi.org/10.1063/5.0021159
  • [7] J. H. Conway and N. J. A. Sloane, On lattices equivalent to their duals, J. Number Theory 48, 373–382 (1994).
  • [8] J. H. Conway and N. J. A. Sloane, The optimal isodual lattice quantizer in three dimensions, Adv. Math. Commun. 1 (2007), 257–260. arXiv preprint math/0701080 (2007).
  • [9] J. H. Conway and N. J. A. Sloane, Sphere packings, lattices and groups, Vol. 290. Springer Science & Business Media, 2013.
  • [10] S. Cooper, Cubic theta functions, J. Comput. Appl. Math. 160, 77–94 (2003).
  • [11] S. Cooper, Ramanujan’s Theta Functions, Springer, New York, 2017.
  • [12] R. Fürth, On the equation of state for solids, Proc. Royal Soc. Lond. A. Math Phys. Sci. 183, 87–110 (1944).
  • [13] E. Grüneisen, Theorie des festen Zustandes einatomiger Elemente, Ann. Phys. 344, 257–306 (1912).
  • [14] J. E. Jones and A. E. Ingham, On the Calculation of Certain Crystal Potential Constants, and on the Cubic Crystal of Least Potential Energy, Proc. Royal Soc. Lond. A. Math Phys. Sci. 107, 636–653 (1925).
  • [15] G. Kane, and M. Goeppert-Mayer, Lattice Summations for Hexagonal Close-Packed Crystals, J. Chem. Phys. 8, 642 (1940).
  • [16] N. Levinson and R. M. Redheffer, Complex Variables, Holden-Day, San Francisco, 1970.
  • [17] E. Madelung, Das elektrische Feld in Systemen von regelmäßig angeordneten Punktladungen, Phys. Z. 19, 32 (1918).
  • [18] P. Schwerdtfeger, A. Burrows and O. R. Smits, The Lennard Jones Potential Revisited–Analytical Expressions for Vibrational Effects in Cubic and Hexagonal Close-Packed Lattices, J. Phys. Chem. A 125, 3037–3057 (2021).
  • [19] A. Selberg and S. Chowla, On Epstein’s zeta-function, J. Reine Angew. Math. 227, 86–110 (1967).
  • [20] F. H. Stillinger, Lattice sums and their phase diagram implications for the classical Lennard-Jones model, J. Chem. Phys. 115, 5208–5212 (2001).
  • [21] N. Temme, Special Functions. An Introduction to the Classical Functions of Mathematical Physics, Wiley, New York, 1996.
  • [22] I. J. Zucker and M. M. Robertson, Exact values for some two-dimensional lattice sums, J. Phys. A 8, 874–881, (1975).