The Cuboidal Lattices and their Lattice sums
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, 65B101. 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 with [6], i.e. the expression is the quadratic form associated with the lattice (or simply, the associated quadratic form).
In the general case of a -dimensional lattice (), is a positive definite, real and symmetric matrix called the Gram matrix of the lattice , defined by its basis (or lattice) vectors through . is called the generator matrix ( 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 or generator matrix we introduce a few important definitions required in this work [9]. Two generator matrices and are equivalent if , a non-zero real number, a real orthogonal matrix () with describing rotation, reflection or rotoreflection of the lattice, and a matrix containing integers with describing for example permutations of the basis vectors. Given two equivalent generator matrices and , the corresponding Gram matrices are related by
The minimum distance in a lattice is defined by
where is the Euclidean distance. In terms of the Gram matrix this is equivalent to
The minimal norm is related to the minimum distance by . Dividing by assures that used in most lattice sum applications [6]. For dense sphere packings the radius of a sphere is simply . The packing density and the center density of a three-dimensional lattice are given by
The kissing number for dense sphere packings is defined by
In this work we discuss cuboidal lattices and their lattice sums. We first present the characteristics of cuboidal lattices dependent on a single parameter . 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 , where and are non-zero real numbers. To make it specific, take the basis vectors
(2.1) |
Then the generator matrix is given by
which has determinant . The Gram matrix is
(2.2) |
where and is positive definite for . Conway and Sloane [7] use , so . The most important cases, in decreasing numerical order, are:
-
(1)
: the face-centred cubic (f.c.c.) lattice;
-
(2)
: the mean centred-cuboidal (m.c.c.) lattice;
-
(3)
: the body-centred cubic (b.c.c.) lattice;
-
(4)
: 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
(2.3) |
It is easy to check that
(2.4) |
It follows from (2.3) and (2.4) that the minimum distance is given by
(2.5) |
We rescale to make the minimum distance by defining
The examples we are interested are in (f.c.c., m.c.c., b.c.c., a.c.c.) all satisfy . Because of its practical interest, this is the only case we will study, and from here on (unless otherwise mentioned) we will always assume in which case we have
(2.6) |
corresponding to the rescaled Gram matrix
(2.7) |
The packing density is calculated as
where and
It follows that
On using the formula for in (2.5) we obtain the formula for the packing density, given by
Figure 1 shows a graph of the packing density as a function of the parameter . Further information is recorded in Table 1. In the main region of interest , we have
(2.8) |
and so
(2.9) |
It follows that on the interval , the packing density has a maximum of at corresponding to f.c.c., and a minimum of at corresponding to b.c.c.
It is also interesting to note that as the limiting case of the lattice is the two-dimensional square close packing with minimal distance and kissing number ; while in the other extreme case the limit as gives the one-dimensional lattice with minimal distance and kissing number . These cases are briefly analysed in Appendix B.
Region | integer combinationsa | |||
---|---|---|---|---|
I | 2 | |||
a.c.c. | 10 | |||
II | 8 | |||
f.c.c. | 12 | |||
III | 4 |
a The integer combinations which determine in (2.5) for the different regions are as follows: , , , , , , , , , , , , , .

Given a positive definite quadratic form , the corresponding theta series is defined for by
For the quadratic form in (2.6) the theta series is
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 are given respectively by
Since the quadratic form has been normalised to make the minimum distance , the kissing number occurs in each theta series as the coefficient of . That is, we have , , and .
We introduce the following lattice sum important in solid state theory [5],
(2.10) |
where . 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 , and except for the term which is omitted. This lattice sum smoothly connects four different well known lattices, i.e., when , , or 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
We conclude this section by reconciling the Gram matrix in (2.2) with two matrices given by Conway and Sloane [7]. Let
and consider the equivalent matrices and defined by
(2.11) |
and
(2.12) |
When and , the matrix in (2.11) is the Gram matrix for the m.c.c lattice given by Conway and Sloane [7, (10)]. Moreover, when and , the matrix leads to another well-known quadratic form for the b.c.c. lattice, e.g., see [6, (8b)]. When , , the matrix in (2.12) is the Gram matrix for the a.c.c. lattice given in [7, p. 378]. Since it follows that
The corresponding quadratic forms and are defined by
and | ||||
They are related to the quadratic form in (2.3) by
(2.13) |
and
3. A minimum property of the lattice sum
In the previous section — see (2.8) and (2.9) — it was noted that on the interval , the packing density function has a minimum value when . The next result shows that provided , the corresponding lattice sum also attains a minimum at the same value .
Theorem 3.1.
Proof.
By definition we have
where
Now make the change of variables given by (2.13), namely
This is a bijection since
and it follows that
By direct calculation, the derivative is given by
(3.1) |
Setting gives
(3.2) |
Switching and gives
(3.3) |
while switching and in (3.2) gives
(3.4) |
On adding (3.2), (3.3) and (3.4) and noting that
it follows that
Next, taking the derivative of (3.1) gives
When the first sum is zero by the calculations in the first part of the proof. Therefore,
The term 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
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. On restricting to real values, the conclusion about positivity is valid for . ∎
A consequence of Theorem 3.1 is that for any fixed value , the lattice sum attains a minimum when . Graphs of to illustrate this minimum property are shown in Fig. 2. In the limiting case we have
This graph is also shown in Fig. 2.

4. Evaluation of the sum
We now turn to the evaluation of . Our objectives are to find formulas that are both simple and computationally efficient. The formulas we obtain can be used to show that can be analytically continued to complex values of , with a simple pole at and no other singularities.
One method of evaluating the sum 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 . Here we use an easier method that also works the whole parameter range 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
(4.1) |
Therefore, we evaluate the sums
(4.2) |
and
(4.3) |
By (4.1), (4.2) and (4.3), the required lattice sum is given by
(4.4) |
4.1. The sum
We shall consider two ways for handling the sum in (4.2). The first is to separate the terms according to whether or , which gives rise to
(4.5) |
where
and | ||||
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 or and write
(4.6) |
where
and | ||||
The series , and also depend on . For simplicity we omit the parameter from the notation and just write , and in place of , and , 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
where
is the Riemann zeta function, and
is the Dirichlet beta series. It remains to analyse . By the integral formula for the gamma function (A.2) we have
Now apply the modular transformation for theta functions (A.12) to obtain | ||||
where is the number of representations of as a sum of two squares, e.g., see (A.13). Now separate out the term and evaluate the resulting integrals. We find that | ||||
where we have used the formula (A.3) for the -Bessel function. On using all of the above back in (4.5) we obtain
(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 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
This time we split the terms according as in (4.6) and start with
(4.8) |
where
and | ||||
Now apply the integral formula for the gamma function (A.2) and then the modular transformation for the theta function (A.9) to obtain
Separate the term, to get | ||||
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
On using all of the above back in (4.8) we obtain
(4.9) |
The terms in (4.7) involve Bessel functions whereas 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 , due to the creation of a 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).
4.3. The alternating sum
The analysis in the previous sections can be modified to handle the alternating series (4.3) which has the term in the numerator, as follows. Separating the terms according to whether or gives
(4.10) | ||||
where | ||||
and | ||||
By a known result (A.21), we have
Next, by the integral formula for the gamma function (A.2) we have
Now apply the modular transformation for theta functions to obtain | ||||
By formula (A.14) this can be expressed as | ||||
The integral can be expressed in terms of Bessel functions by (A.3) to give | ||||
On using all of the above back in (4.10) we obtain
(4.11) |
4.4. A second formula for the alternating sum
This time we separate the terms according to whether or and write
(4.12) |
where
By (A.30) we have
It remains to analyse the sum for . By the integral formula for the gamma function (A.2) we have
Now apply the modular transformation (A.11) to obtain | ||||
Now put and use
to deduce
The integral can be evaluated in terms of the modified Bessel function by (A.3) to give | ||||
It follows that
(4.13) |
4.5. Two formulas for
On substituting the results of (4.7) and (4.11) back into (4.4) we obtain a formula for in terms of Bessel functions:
(4.14) |
On the other hand, if the results of (4.9) and (4.13) are used in (4.4), the resulting formula for involves Bessel functions:
(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]
where
and
As before, the prime on the sum for indicates that the summation is over all integers except for the term corresponding to which is omitted. The sum for is over all integer values of , and . We shall analyse and one at a time.
5.1. The sum
Break the sum for into two, according to whether or . This gives
(5.1) |
where
and
By (A.33) we have
where
It remains to calculate . Applying the gamma function integral (A.2) followed by the theta function transformation formula (A.15), we obtain
Now separate out the term and evaluate the resulting integrals. The result is
It follows that
(5.2) |
where is the number of representations of by the form .
5.2. A second formula for the sum
A different formula for can be obtained by separating the terms in the series according to whether or and are not both zero. This gives
where
Applying the gamma function integral (A.2) followed by the theta function transformation formula (A.9), we obtain
Now separate out the term and evaluate the resulting integrals. The result is
The first sum can be evaluated in terms of the Riemann zeta function and the function by (A.33). In the second sum, we again use the notation for the number of representations of by the form . The result is
It follows that
(5.3) |
5.3. The sum
The analysis in this case is a little simpler because the summation is over all integers , and . We apply the gamma function integral (A.2) to write
(5.4) |
Now make use of the transformation formula (A.16) to deduce
where is a primitive cube root of 1. Now separate the term to deduce | ||||
where is the number of representations of by the form , as before. On evaluating the integrals using (A.2) and (A.3) we obtain
The first sum can be evaluated in terms of the Riemann zeta function by using (A.29). The term can be handled by using
along with the fact that is real valued when is real. It follows that
(5.5) |
5.4. A second formula for the sum
We introduce the abbreviation
to write (5.4) in the form
This time we apply the transformation formula (A.11) to the sum over to obtain | ||||
Now separate the terms according to whether or and evaluate the resulting integrals by (A.2) and (A.3). The result is
The first sum can be handled by (A.34) to give
For the other sum, observe that
that is to say is a positive integer and . Therefore we set and use (A.19) to deduce that the number of solutions of
is equal to .
Taking all of the above into account, we finally obtain
(5.6) |
5.5. The lattice sum for hexagonal close packing
6. Analytic continuations of the lattice sums and
We will now show that the lattice sums and can be continued analytically to the whole -plane, and that the resulting functions have a single simple pole at 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 and determine the residue. Then, in Section 6.2 we show that the analytic continuations obtained are valid for the whole -plane and there are no other singularities. Finally, in Sections 6.3–6.5, values of the analytic continuations at the points and are computed. In particular, the evaluation of at in the case gives the Madelung constant, e.g., see [4], [5, pp. xiii, 39–51], [17].
6.1. Behaviour of the lattice sums at
We start by showing that has a simple pole at and determine the residue. In the formula (4.14), all of the terms are analytic at except for the term involving . It follows that
where (A.23) was used in the last step of the calculation. It follows further that has a simple pole at and the residue is given by
By (2.8) this is just 12 times the packing density, i.e.,
For example, taking gives
(6.1) |
while taking gives
Laurent’s theorem implies there is an expansion of the form
(6.2) |
where
and the coefficients , , depend on but not on . To calculate , start with the fact that
where is Euler’s constant. Then use (4.14) and (A.5) to deduce
Interchanging the order of summation and evaluating the sum over gives | ||||
Numerical evaluation in the case gives
(6.3) |
A similar analysis can be given for 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
(6.4) |
where
(6.5) |
and
(6.6) |
In particular, the pole of at is simple. By (6.1) and (6.5) we have
It follows that the difference has a removable singularity at and from the Laurent expansions we deduce that
Using the numerical values from (6.3) and (6.6) we obtain
This gives the value at the left hand end of the graph in [6, Fig. 3]. The value used here corresponds to taking 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
By (A.6), the double series of Bessel functions in (4.14) converges absolutely and uniformly on compact subsets of the -plane and therefore represents an entire function of . It follows that has an analytic continuation to a meromorphic function which is analytic except possibly at the singularities of the terms
(6.7) |
and
(6.8) |
The function in (6.7) is analytic except at due to the pole of , as the function and the exponential function are both entire. The function in (6.8) is analytic except at and . We studied the singularity at in the previous section, so this leaves only the point . Using (A.23) and the values of and in (A.26) and (A.27) we find that
and
It follows that the sum of the functions in (6.7) and (6.8) has a removable singularity at and so is also analytic at . The analyticity at can also be seen directly from the alternative formula for in (4.15).
In conclusion, it has been shown that has an analytic continuation to a meromorphic function of which has a simple pole at and no other singularities. Because has only one singularity, namely , the Laurent expansion (6.2) is valid in the annulus , i.e., for all .
In a similar way, (5.7) and (5.8) can be used to show that also has an analytic continuation to a meromorphic function of which has a simple pole at and no other singularities. The Laurent expansion (6.4) converges for all .
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
(6.9) |
This is interpreted as being the value of the analytic continuation of the series at , because
obviously diverges if . From now on, we shall use the expression “the value of a series at a point ” to mean “the value of the analytic continuation of the series at the point ”.
6.3. Values at and the Madelung constant
On putting in (4.11) we obtain an analytic expression for the value of
which specialises to the Madelung constant in the case . We have | ||||
Now use (A.4) and (A.5) to express the Bessel functions in terms of exponential functions. The result simplifies to
On interchanging the order of summation and summing the geometric series, we obtain | ||||
When this gives the Madelung constant defined by (6.9). Numerical evaluation gives
(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 in [5, pp 39–51].
6.4. The value at
It was noted above that (4.14), which involves Bessel functions, contains terms with singularities at and therefore is not suitable for calculations at that value of . Instead we can use (4.15), which involves Bessel functions. As in the previous section, two steps are involved. First, the the 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.
For example, when the above formulas give
(6.14) | ||||
and | ||||
Then taking and in (4.4) gives
For HCP, the formula (5.7) cannot be used to evaluate because two of the terms have cancelling singularities at . Therefore we take in (5.8) instead to obtain
We end this section by noting a connection between two of the values in the above analysis. By setting in each of (6.11) and (6.13) we obtain the remarkable result
(6.15) |
This is consistent with [5, p. 46 (1.3.44)] and is the special case of the functional equation
(6.16) |
This functional equation can be deduced from the two formulas for in (4.7) and (4.9), as follows. Replace with in (4.7), then multiply by and set to get
where we have used the functional equation for the gamma function in the form
to obtain the second term on the right hand side. Now apply the functional equations (A.4), (A.24) and (A.25) to deduce
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
7. Graphs
The formulas (4.14), (4.15), (5.7) and (5.8) have been used to produce the following graphs of on the intervals and in Figure 3. The graph of has a similar appearance, and so to allow a comparison the difference
is plotted using a finer vertical scale in Figure 4.



The graphs appear to suggest the following:
Conjecture:
and
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 by
(A.1) |
By the change of variable this can be rewritten in the useful form
(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:
(A.3) |
By the change of variable it can be shown that
(A.4) |
When the modified Bessel function reduces to an elementary function:
(A.5) |
The asymptotic formula holds:
(A.6) |
For all of these properties, see [1, pp. 223, 237] or [21, pp. 233–248].
A.3. Characters
For an integer , let and be defined by
(A.7) |
and
(A.8) |
A.4. Theta functions
The transformation formula for theta functions is [1, p. 119], [3, (2.2.5)]:
(A.9) |
We will need the special cases and , which are
(A.10) |
and
(A.11) |
respectively. The sum of two squares formula is [11, (3.111)]
(A.12) |
where
(A.13) |
the sum being is over the positive divisors of . For example,
By [11, (3.15) and (3.111)] we also have
(A.14) |
A.5. The cubic theta function
The cubic analogues of the transformation formula are [3, (2.2)], [10, Cor. 5.19]
(A.15) |
and
(A.16) |
where is a primitive cube root of unity. The analogue of the sum of two squares result is [11, (3.124)]
(A.17) |
where
(A.18) |
where the sum is again over the positive divisors of . By [11, (3.18) and (3.124)] we also have
(A.19) |
which is the analogue of (A.14).
A.6. The Riemann zeta function and functions
The definitions are:
(A.20) | ||||
(A.21) | ||||
(A.22) |
The function is the Riemann zeta function. It has a pole of order at , and in fact
(A.23) |
This is a consequence of [1, (1.3.2)]. See also [21, p. 58].
We will require the functional equations
(A.24) |
and
(A.25) |
and the special values
(A.26) |
(A.27) |
and
(A.28) |
See [2, Ch. 12] or [22]. Other results used are
(A.29) |
(A.30) |
(A.31) |
(A.32) |
(A.33) |
(A.34) |
The identities (A.29) and (A.30) follow from the definition of 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 and
We briefly consider the behaviour of the lattices in the limiting cases and . Some of the basis vectors become infinite in the limit, leaving a sublattice of lower dimension. We discuss each case and both in terms of theta functions and then in terms of the basis vectors.
First, consider the limit . In the interval the theta function is
As we have
unless and , respectively. Hence,
This corresponds to the one-dimensional lattice with minimal distance . The kissing number is 2, which is in agreement with the other lattices in the range , as indicated in Table 1. In terms of the basis vectors, from (2.1) we have
The only linear combinations (for ) that remain finite in the limit occur when , in which case we obtain
That is, the limiting lattice is just the one-dimensional lattice consisting of integer multiples of .
Now consider the limit . For the theta function is
Since as unless , it follows that
This is the theta series for the two-dimensional square close packing lattice with minimal distance . The kissing number is , in agreement with other values in the range given by Table 1. In terms of the basis vectors, from (2.1) we have
The only linear combinations (for ) that remain finite in the limit occur when , in which case we obtain
This is isomorphic to the two-dimensional square close packing lattice with minimal distance , 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).