On the lower bound for packing densities of superballs in high dimensions
Abstract
Define the superball with radius and center in to be the set
which is a generalization of -balls. We give two new proofs for the celebrated result that for , the translative packing density of superballs in is . This bound was first obtained by Schmidt, with subsequent constant factor improvement by Rogers and Schmidt, respectively. Our first proof is based on the hard superball model, and the second proof is based on the independence number of a graph. We also investigate the entropy of packings, which measures how plentiful such packings are.
Keywords: superball packing, hard superball model, uniform convexity, independence number
AMS subject classifications: 52C17, 05C15, 05C69
1 Introduction
The sphere packing problem asks the densest packing of nonoverlapping equal-sized balls in . This is an old and difficult problem in discrete geometry. The exact answer is only known in dimensions , and . In dimension , the problem is trivial; in dimension , the problem is nontrivial and can be solved by elementary geometry; in dimension , the problem is known as the Kepler’s conjecture and was solved by Hales [8]; in dimension , the problem was solved by Viazovska [31] by finding a function that matches the Cohn-Elkies linear programming bound [3]; in dimension , the problem was solved by Cohn et al. [4].
Let be the maximum translative packing density of equal-sized Euclidean balls in . The best upper bound of in high dimension is due to Kabatjanskiĭ and Levenšteĭn [12]: . See also Cohn and Zhao [5] and Sardari and Zargar [25] for constant factor improvement. A lower bound is trivial since doubling the radii of balls will cover the whole space. Rogers [16] improved the bound by a factor . Ball [1] constructed lattice packings of density . The best known lower bound is , due to Venkatesh [30].
In this paper, we consider the packing density of superballs. Let and be a real number. Let be the norm; that is, for , . Let be such that . Denote the superball with radius and center in by
where and . We simply write if the superball is centered at . Here, the role of is to cut the vectors in ; that is, cuts and in into shorter vectors and , respectively. We first calculate the -distance between and . And then we calculate the norm of the new vector formed by these .
Throughout the paper, we assume the superballs of our packings have volume and denote by the radius of a superball of volume . Let denote the maximum translative packing density of copies of superballs of volume ; that is,
where is the volume of covered by superballs of volume with centers in , and the supremum is over all such that the superballs with centers in are nonoverlapping.
Consider a special case that . In this case,
is the -ball in . Let be the packing density of -balls. The upper bound for was first proved by van der Corput and Schaake [29], giving if and if . For , the upper bound was improved exponentially by Sah et al. [24]. The Minkowski-Hlawka theorem [9] gives a lower bound that , where is the Riemann zeta function. Rush and Sloane [23] improved the Minkowski-Hlawka bound for -balls and integers , for instance, . Rush [19] constructed lattice packings with density for every convex body which is symmetric through each of the coordinate hyperplanes. Moreover, Elkies et al. [6] improved the Minkowski-Hlawka bound exponentially for superballs and reals , and they also obtained lower bound for the packing density of more general bodies. For the lower bound constructed by error correcting codes, see Rush [20, 21, 22] and Liu and Xing [14].
We focus on the lower bound in the case that . In this case, only subexponential improvements of the Minkowski-Hlawka bound are known. Rogers [17] obtained a lower bound of . Schmidt [26] obtained a lower bound of . See also Rogers [18] and Schmidt [27] for constant factor improvements. The best result is due to Schmidt [28], where the constant factor is less than . We give new proofs for the lower bound of the packing density of -balls.
Theorem 1.1.
For every , there exists a constant such that
We also obtain a lower bound for .
Theorem 1.2.
For every , there exists a constant such that the following holds. For every with and , we have
The lower bound in Theorem 1.2 is independent of . Note that different will lead to different bodies in the packing. So this result means that, given and letting be large, for every ( and ), the translative packing density of is approximately larger than . Throughout the paper, and appear in the subscripts of most of our symbols and variables since they depend on and . However, we always assume that is a given number and our bounds are independent of (Theorem 1.2, Theorem 3.2, Theorem 5.1 and Theorem 5.2), so and in the subscripts could be ignored for simplicity.
For , we can obtain a lower bound of as well. But this bound is worse than the bound in [6].
We will give two proofs of Theorem 1.2. In Section 3, we give the first proof. The method we use is the hard superball model, which is developed from statistical physics. It was used by Jenssen et al. to prove the lower bound of kissing number [10] and packing density of Euclidean balls [11], where the methods are called hard cap model and hard sphere model, respectively. Recently, Fernández et al. [7] gave a constant factor improvement of results in [10] and [11]. In Section 4, we give the second proof, which uses the independence number of a graph. We also use the so-called uniform convexity to overcome the difficulty for the non-Euclideanness.
We also investigate the entropy density and pressure of packings, which we will define in Section 5. This measures how plentiful such packings are.
2 Uniformly convex spaces
Throughout the paper, we always assume that is a (finite or infinite) sequence of strictly increasing nonnegative integers. Let . For such that and , define
where . For with (obviously , since ’s are nonnegative integers) and , define
where . Let be the set consisting of all real sequences satisfying .
Proposition 2.1.
For every with , and for every , is a normed linear space equipped with the norm .
Proof.
For every and , let
and
We have
(1) |
where the first ‘’ follows from the triangle inequality of and the second ‘’ follows from the Minkowski inequality
(2) |
for every nonnegative real sequences and and (if , then inequality (2) is in the reverse sense). And
(3) |
So is a linear space with the usual addition and scalar multiplication.
For similar reasons, is a norm on as well. For , let be the -distance between and in . For sequences , let be the -distance between and in .
Definition 2.2.
A normed linear space equipped with norm is said to be uniformly convex if to each , there corresponds a such that the conditions and imply .
If we write , then the space is the usual space. And we have the following theorem.
Theorem 2.3 ([2]).
For , the space is uniformly convex. If and is the constant with respect to the uniform convexity of space, then we can choose , where is the conjugate index.
We have the following generalization of Theorem 2.3.
Proposition 2.4.
For every with , and for every , is uniformly convex. If and is the constant with respect to the uniform convexity of space, then we can choose , where is the conjugate index.
Remark 2.5.
We can choose independent of .
The proof of Proposition 2.4 is similar to the proof of uniform convexity of in [2], so we first introduce the following lemma, which is a generalization of [2, Theorem 2].
Lemma 2.6.
For the space defined above, with , the following inequalities between the norms of two arbitrary elements and of the space are valid (Here is the conjugate index, ).
(4) |
(5) |
(6) |
For these inequalities hold in the reverse sense.
Proof.
For simplicity, we write and in the proof. First note that for all values of the right hand side of inequality (4) is equivalent to the left hand side, while inequality (5) is equivalent to inequality (6); to see this, set and we have
and
We first prove inequality (5) for , i.e.
(7) |
We claim that for , we have
(8) |
If one of and is , the inequality holds trivially. Assume that . Divide by in the both sides of inequality (8) and use the homogeneity, reducing inequality (8) to
(9) |
Let and . Then and , and inequality (9) is equivalent to
i.e.
or equivalently,
where is the usual inner product. Suppose , so
Without loss of generality assume that . Consider the function
We have for , since . Thus and it suffices to prove
(10) |
Inequality (10) has been proved in [2] (see the proof of [2, Theorem 2]). So inequality (8) is proved.
Now we are able to prove inequality (7). Let , where
and
Inequality (7) states that
(11) |
Using Minkowski inequality (2) in the reverse sense with and , we infer that the right hand side of inequality (11) is
which by inequality (8) is
Since , this is our result. It completes the proof of inequality (5) for .
Now we prove inequality (5) for , and we need to prove inequality (11) in the reverse sense, i.e.
(12) |
Letting and have the same values as above, and again applying Minkowski inequality (2), we conclude that the right hand side of inequality (12) is
(13) |
Recall that inequality (8) holds for , and it is equivalent to
After exchanging the role of and , we have
for . Thus inequality (13) is
Therefore, we complete the proof of inequality (5) for .
Finally we prove inequality (4). Let and consider the right hand inequality. This is simply implied by inequality (6): For , we have
(14) |
Inequality (14) has been proved in [2] (see the proof of [2, Theorem 2]). For , it follows from inequality (14) that
(15) |
The right hand side of inequality (4) (in the reverse sense) follows from inequality (6) (in the reverse sense) and inequality (15). ∎
Now we are able to prove Proposition 2.4.
3 The hard superball model
Given and with , consider the translative packing of . For a bounded, measurable subset , let
be the family consisting of unordered -tuples of points from that can form a packing.
The canonical hard superball model on with centers is simply a uniformly random -tuple . The partition function of the canonical hard superball model on is the function
(16) |
where for , the expression denotes the event that for every .
The grand canonical hard superball model on at fugacity is a random set of unordered points, with distributed according to a Poisson point process of intensity conditioned on the event that for all distinct . In other words, we first choose from with probability , then we independently, uniformly choose from . The partition function of the grand canonical hard superball model on is
(17) |
where we take . Note that if is bounded then is a polynomial in .
The expected packing density, , of the hard superball model is simply the expected number of centers in normalized by the volume of ; that is,
Here and in what follows the notation and indicates probabilities and expectations with respect to the grand canonical hard superball model on a region at fugacity .
The expected packing density can be expressed as the derivative of the normalized log partition function.
(18) |
Here and in what follows always denotes the natural logarithm of .
Lemma 3.1.
The asymptotic expected packing density of is a lower bound on the maximum superball packing density. That is, for any ,
Proof.
By the definition of ,
where is the family consisting of all packings of by superballs of radius , and is the volume of . Note that in our model, some centers may be near the boundary of , which should be deleted from the packing. However, we can enlarge the radius of the superball to make it be a packing. In other words, if is chosen from the model on , then is a packing of . So
∎
Theorem 3.2.
For every , there exists a constant such that the following holds. Let be bounded, measurable, and of positive volume. Then for every such that and every , we have
Remark 3.3.
The constant in Theorem 3.2 is independent of and .
Lemma 3.4.
Let be bounded, measurable, and of positive volume. Then the expected packing density is strictly increasing in .
Proof.
Let denote the expected free volume of the hard superball model; that is, the expected fraction of the volume of containing points that are at distance at least from the nearest center. That is,
Lemma 3.5.
Let be bounded, measurable, and of positive volume. Then
Proof.
We use the definitions of and .
Let . Then
So
where for , (since ). ∎
Now consider the following two-part experiment: sample a configuration of centers from the hard superball model on at fugacity and independently choose a point from . We define the random set
That is, T is the set of all points of in that are not blocked from being a center by a center outside of .
Lemma 3.6.
Let be bounded, measurable, and of positive volume. Then
(20) |
and
(21) |
where both expectations are with respect to the random set T generated by the two-part experiment defined above.
Proof.
Using Lemma 3.5, we have
where the last equation uses the spatial Markov property of the hard superball model: conditioned on , the distribution of is exactly that of the hard superball model on the set .
For every , . So
∎
Lemma 3.7.
Let be bounded and measurable. Then
(22) |
and if in addition is of positive volume, then
(23) |
Proof.
Remark 3.8.
Consider the function
Note that is continuous for and , so there exists such that for every . We will use in the proof of the following lemma.
Lemma 3.9.
For every , there exists a constant such that the following holds. For every and with , let be measurable. Then
(24) |
where is a uniformly chosen point in . In particular,
(25) |
Remark 3.10.
The constant in Theorem 3.2 will be chosen as the same as the constant here. So we use the same symbol.
Proof.
Clearly, we may assume that has positive volume. We have
In order to prove inequality (24), it suffices to prove that there exists a constant such that for every and for every , we have
For every and for every , if , then
Next we consider the case that . For every , we have and . If or , then by properties of a norm (the homogeneity condition and the triangle inequality), we have
(26) |
Now suppose that and . Fix . Let . By Proposition 2.4, for every and for every , if and , then . For every and with , we view a point as a point in by adding zeros after the -th coordinate of and , so . In the rest of proof, we use norm instead of .
Let , , and . We have known that and . Since and
it follows from the uniform convexity of that
(27) |
where . So . Multiplying by in the both sides of the inequality (27) and rearranging it, we have
Therefore,
(28) |
Since
and by the definition of
it follows that
(29) |
Now we are able to prove Theorem 3.2.
Proof of Theorem 3.2.
Let be bounded, measurable, and of positive volume. Let be the constant in Lemma 3.9 and . Then by Jensen’s inequality, we have
where the above expectation is with respect to the two-part experiment in forming the random set and the first equality follows from the equation (20).
On the other hand, we have
Combining these two lower bounds and letting , we have
Since is decreasing in and is increasing in , the infimum over of the maximum of the two expressions occurs when they are equal, i.e., , where is the solution to
In other words,
(30) |
where is the Lambert-W function. For , is defined to be the unique solution to the equation . So
As ,
We take . So as . The equation (30) becomes
(31) |
So
Since is increasing in , this bound holds for every , completing the proof. ∎
In the above proof, if we take for some , then we have
i.e.,
(32) |
for .
4 An alternate proof of Theorem 1.2
In this section, we give an alternate proof of Theorem 1.2 without using the hard superball model. The idea of the proof comes from [13, Section 6.1]. We fix and , and let be a sufficiently large number. For convenience, in this section, we simplify most of our symbols and variables: , and .
Let be the superball centered at with radius . Consider the packings in using . Let be a small positive real number (we will determine later), be a basic cube and be a lattice. tiles the whole space, and it also gives a partition of . Let . We have the following lemma.
Lemma 4.1.
covers .
Proof.
Suppose on the contrary that there is some such that for every . Let such that for every , in other words, . By the choice of , we have that . On the other hand, for every , we have
Note that since the number of is at most , and since . So
This holds for every . Thus , a contradiction. ∎
Let . We arbitrarily choose from each , and construct an auxiliary graph with vertex set , where and are adjacent if and only if . So the maximum number of superballs with radius that can be packed in is larger than or equal to , where is the independence number of . In other words,
(33) |
Here, using the same trick in the proof of Lemma 3.1, the points near the boundary of do not affect the result.
Lemma 4.2.
Every vertex in has degree at most .
Proof.
We write for the closed neighborhood of . We claim that for every ,
(34) |
For the first inclusion, let . By Lemma 4.1, there exists an index such that . So . As , it follows that , i.e. . For the second inclusion, let for some index with . We have and . So , as desired.
Since ’s are nonoverlapping, it follows from the claim that
i.e.,
where we choose satisfying . ∎
Let and , where is the constant in Lemma 3.9. Let denote the induced subgraph of whose vertex set is . We have the following lemma.
Lemma 4.3.
Suppose and are sufficiently large. For every , the average degree of is at most .
Proof.
We calculate . Let and where . So . By the definition, the average degree of is at most
(35) |
By the definition of integration,
in other words, for every , there exists such that
(36) |
whenever .
Recall that in the proof of Lemma 4.2 we have . Let . So and
(37) |
where we choose small enough so that . For instance, if we choose satisfying , then .
Recall that in the proof of Lemma 3.9, we have
(38) |
Combining equation (35) and inequalities (36)-(38), the average degree of is at most
(39) |
Next we give a lower bound of . Recalling the inclusion relation (34), we have
Denote . If , then
If , then we have
where . To see this, letting , we have . So
And
Since . Thus
In conclusion, we have
Therefore, if we choose and , then by inequality (39), we know that the average degree of is at most
for large . Note that this is independent of the choice of , so we are done. ∎
By Lemmas 4.2 and 4.3, we know that the -vertex graph has maximum degree and every subgraph induced by a neighborhood in has average degree at most . We say such a graph is locally sparse. By a result of Hurley and Pirot [15, Corollary 1], the chromatic number of is at most . So the independence number of is at least
(40) |
and
This completes the proof of Theorem 1.2.
5 A lower bound on entropy density and pressure
In this section, we investigate the entropy density and pressure of packings. Let be the region of packings and . Define
to be the entropy density of the packings. Consider a random packing in with density . The superballs we use in the packings have volume , so there are centers. measures all the choices of unordered points chosen in , and measures the choices of unordered points that can form a packing. Dividing by ensures that is independent of the choice of the size of superballs in our packings. So measures how plentiful the packings of density are. We also define
to be the pressure of the packings. measures how plentiful the packings with fugacity are. We have the following theorems.
Theorem 5.1.
Let be the constant in Theorem 3.2. For , we have
(41) |
Note that in Theorem 5.1, .
Theorem 5.2.
Let be the constant in Theorem 3.2. There exists such that
The proofs of Theorems 5.1 and 5.2 are similar to those of [11, Theorem 4] and [11, Theorem 5], respectively. For the sake of completeness, we include them here.
Proof of Theorem 5.1.
Before proving Theorem 5.2, we give a simple fact of : the larger the density is, the less such packings exist.
Lemma 5.3.
is decreasing in .
Proof.
If , then for large . So when .
Suppose , and let and satisfy (so ). We claim that
(42) |
where and (so ). In order to prove inequality (42), it suffices to prove
Using equation (16), the definition of , it reduces to
(43) |
where . Consider
the right hand side of inequality (43). Let . Then
where means the event that for every .
The left hand side of inequality (43) is the probability of the event that uniformly chosen points in with pairwise distance larger than or equal to , while the right hand side of inequality (43) is the probability of the event that uniformly chosen points in with pairwise distance larger than or equal to . The latter event is contained in the former event, i.e.
Thus inequality (43) holds and the claim is proved.
Finally, we give a proof of Theorem 5.2.
Proof of Theorem 5.2.
Consider packings in and . Assume that for every . Using equation (19) in the proof of Lemma 3.4, we have
a contradiction. So there exists such that .
Choose such that . Using Chebyshev’s inequality, we have
So
Note that is an integer. By averaging, there exists an integer
such that
for large . Recall that
So
On the other hand, by Theorem 3.2 and the definition of , we have
Let be a small positive number (we will choose dependent on but independent of ) and . So for large ,
Applying Lemma 5.3, it follows that
We calculate
Note that when . Hence Stirling’s formula implies that
Therefore,
Choose so that , completing the proof. ∎
Acknowledgements
The authors would like to thank Professor Yufei Zhao and Professor Chuanming Zong for many helpful and valuable comments on the manuscript. The authors also thank Professor Hong Liu for sharing his nice idea about the alternate proof.
References
- [1] K. Ball. A lower bound for the optimal density of lattice packings. Internat. Math. Res. Notices, (10):217–221, 1992.
- [2] J. A. Clarkson. Uniformly convex spaces. Trans. Amer. Math. Soc., 40(3):396–414, 1936.
- [3] H. Cohn and N. Elkies. New upper bounds on sphere packings. I. Ann. of Math. (2), 157(2):689–714, 2003.
- [4] H. Cohn, A. Kumar, S. D. Miller, D. Radchenko, and M. Viazovska. The sphere packing problem in dimension 24. Ann. of Math. (2), 185(3):1017–1033, 2017.
- [5] H. Cohn and Y. Zhao. Sphere packing bounds via spherical codes. Duke Math. J., 163(10):1965–2002, 2014.
- [6] N. D. Elkies, A. M. Odlyzko, and J. A. Rush. On the packing densities of superballs and other bodies. Invent. Math., 105(3):613–639, 1991.
- [7] I. G. Fernández, J. Kim, H. Liu, and O. Pikhurko. New lower bounds on kissing numbers and spherical codes in high dimensions. arXiv e-prints, page arXiv:2111.01255, Nov. 2021.
- [8] T. C. Hales. A proof of the Kepler conjecture. Ann. of Math. (2), 162(3):1065–1185, 2005.
- [9] E. Hlawka. Zur Geometrie der Zahlen. Math. Z., 49:285–312, 1943.
- [10] M. Jenssen, F. Joos, and W. Perkins. On kissing numbers and spherical codes in high dimensions. Adv. Math., 335:307–321, 2018.
- [11] M. Jenssen, F. Joos, and W. Perkins. On the hard sphere model and sphere packings in high dimensions. Forum Math. Sigma, 7:Paper No. e1, 19, 2019.
- [12] G. A. Kabatjanskiĭ and V. I. Levenšteĭn. Bounds for packings on the sphere and in space. Problemy Peredači Informacii, 14(1):3–25, 1978.
- [13] J. Kim, H. Liu, and T. Tran. Exponential decay of intersection volume with applications on list-decodability and Gilbert-Varshamov type bound. arXiv e-prints, page arXiv:2112.11274, Dec. 2021.
- [14] L. Liu and C. P. Xing. Packing superballs from codes and algebraic curves. Acta Math. Sin. (Engl. Ser.), 24(1):1–6, 2008.
- [15] F. Pirot and E. Hurley. Colouring locally sparse graphs with the first moment method. arXiv e-prints, page arXiv:2109.15215, Sept. 2021.
- [16] C. A. Rogers. Existence theorems in the geometry of numbers. Ann. of Math. (2), 48:994–1002, 1947.
- [17] C. A. Rogers. The number of lattice points in a set. Proc. London Math. Soc. (3), 6:305–320, 1956.
- [18] C. A. Rogers. Lattice covering of space: The Minkowski-Hlawka theorem. Proc. London Math. Soc. (3), 8:447–465, 1958.
- [19] J. A. Rush. A lower bound on packing density. Invent. Math., 98(3):499–509, 1989.
- [20] J. A. Rush. Constructive packings of cross polytopes. Mathematika, 38(2):376–380 (1992), 1991.
- [21] J. A. Rush. An indexed set of density bounds on lattice packings. Geom. Dedicata, 53(2):217–221, 1994.
- [22] J. A. Rush. Lattice packing of nearly-Euclidean balls in spaces of even dimension. Proc. Edinburgh Math. Soc. (2), 39(1):163–169, 1996.
- [23] J. A. Rush and N. J. A. Sloane. An improvement to the Minkowski-Hlawka bound for packing superballs. Mathematika, 34(1):8–18, 1987.
- [24] A. Sah, M. Sawhney, D. Stoner, and Y. Zhao. Exponential improvements for superball packing upper bounds. Adv. Math., 365:107056, 9, 2020.
- [25] N. T. Sardari and M. Zargar. New upper bounds for spherical codes and packings. arXiv e-prints, page arXiv:2001.00185, Jan. 2020.
- [26] W. M. Schmidt. The measure of the set of admissible lattices. Proc. Amer. Math. Soc., 9:390–403, 1958.
- [27] W. M. Schmidt. Masstheorie in der Geometrie der Zahlen. Acta Math., 102:159–224, 1959.
- [28] W. M. Schmidt. On the Minkowski-Hlawka theorem. Illinois J. Math., 7:18–23, 1963.
- [29] J. van der Corput and G. Schaake. Anwendung einer blichfeldtschen beweismethode in der geometrie der zahlen. Acta Arithmetica, 2:152–160, 1936.
- [30] A. Venkatesh. A note on sphere packings in high dimension. Int. Math. Res. Not. IMRN, (7):1628–1642, 2013.
- [31] M. S. Viazovska. The sphere packing problem in dimension 8. Ann. of Math. (2), 185(3):991–1015, 2017.