Skip to main content
Log in

Optimal \(N\)-Point Configurations on the Sphere: “Magic” Numbers and Smale’s 7th Problem

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

This paper inquires into the concavity of the map \(N\mapsto v_s(N)\) from the integers \(N\ge 2\) into the minimal average standardized Riesz pair-energies \(v_s(N)\) of \(N\)-point configurations on the sphere \(\mathbb {S}^2\) for various \(s\in \mathbb {R}\). The standardized Riesz pair-energy of a pair of points on \(\mathbb {S}^2\) a chordal distance \(r\) apart is \(V_s(r)= s^{-1}\left( r^{-s}-1 \right) \), \(s \ne 0\), which becomes \(V_0(r) = \ln \frac{1}{r}\) in the limit \(s\rightarrow 0\). Averaging it over the \(\left( \begin{array}{c} N\\ 2\end{array}\right) \) distinct pairs in a configuration and minimizing over all possible \(N\)-point configurations defines \(v_s(N)\). It is known that \(N\mapsto v_s(N)\) is strictly increasing for each \(s\in \mathbb {R}\), and for \(s<2\) also bounded above, thus “overall concave.” It is (easily) proved that \(N\mapsto v_{-2}^{}(N)\) is even locally strictly concave, and that so is the map \(2n\mapsto v_s(2n)\) for \(s<-2\). By analyzing computer-experimental data of putatively minimal average Riesz pair-energies \(v_s^x(N)\) for \(s\in \{-1,0,1,2,3\}\) and \(N\in \{2,\ldots ,200\}\), it is found that the map \(N\mapsto {v}_{-1}^x(N)\) is locally strictly concave, while \(N\mapsto {v}_s^x(N)\) is not always locally strictly concave for \(s\in \{0,1,2,3\}\): concavity defects occur whenever \(N\in {\mathcal {C}}^{x}_+(s)\) (an \(s\)-specific empirical set of integers). It is found that the empirical map \(s\mapsto {\mathcal {C}}^{x}_+(s),\ s\in \{-2,-1,0,1,2,3\}\), is set-theoretically increasing; moreover, the percentage of odd numbers in \({\mathcal {C}}^{x}_+(s),\ s\in \{0,1,2,3\}\) is found to increase with \(s\). The integers in \({\mathcal {C}}^{x}_+(0)\) are few and far between, forming a curious sequence of numbers, reminiscent of the “magic numbers” in nuclear physics. It is conjectured that these new “magic numbers” are associated with optimally symmetric optimal-log-energy \(N\)-point configurations on \(\mathbb {S}^2\). A list of interesting open problems is extracted from the empirical findings, and some rigorous first steps toward their solutions are presented. It is emphasized how concavity can assist in the solution to Smale’s \(7\)th Problem, which asks for an efficient algorithm to find near-optimal \(N\)-point configurations on \(\mathbb {S}^2\) and higher-dimensional spheres.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Notes

  1. Traditionally the Riesz pair-energy is defined as \(\widetilde{V}_s(r)=r^{-s}\) for \(s\ne 0\), and \(\widetilde{V}_0(r)=-\ln r\) for \(s=0\). This has the disadvantages that \(\widetilde{V}_0(r)\ne \lim _{s\rightarrow 0}\widetilde{V}_s(r)\), and that one has to seek energy-minimizing configurations for \(s\ge 0\) yet energy-maximizing ones for \(s<0\).

  2. Our \(v_s (N)\) equals \(2\varepsilon _s(N)\), where \(\varepsilon _s(N)\) denotes the so-called “pair-specific ground state energy” in physics (cf. [13]). While \(\varepsilon _s(N)\) is indeed a physically meaningful quantity, its attribute “pair-specific” is a misnomer—it should actually refer to the statistically meaningful \(v_s (N)\), for the number of different pairs is \(N(N-1)/2\).

  3. Originally, Fekete (cf. [4]) studied points from an infinite compact set in the complex plane that maximize the product of all mutual distances, which is equivalent to minimizing the average standardized Riesz pair-energy for \(s\rightarrow 0\).

  4. By the lower semi-continuity of the standardized Riesz pair-energy and the compactness of the sphere, there always exist \(N\) labeled points (not necessarily pairwise different if \(s \le -2\)) whose average pair-energy equals \(v_s (N)\). A minimizing set of \(N\) labeled points is not a proper minimizing \(N\)-point configuration unless all points are pairwise different.

  5. The growth rate should have a significance similar to “the complexity of the energy landscape,” see [15]. Studies of the Riesz \(s\)-energy landscape for \(N\)-point configurations on \(\mathbb {S}^2\) have only begun recently, see [16] and references therein.

  6. For an exponential time algorithm which provides rational points on the sphere whose logarithmic energy differs from the optimal value by at most 1/9, see Proposition 1.11 in [17].

  7. A good collection of existing search algorithms can be found at the website [18].

  8. In [20] it is conjectured that \(c = \ln \big (2(2/3)^{1/4}\pi ^{3/4}/\Gamma (1/3)^{3/2}\big )\). Recently, a rigorous determination of \(c\) for weighted logarithmic Fekete problems in \(\mathbb {R}^2\), to which the logarithmic Fekete problem on \(\mathbb {S}^2\) is related by stereographic projection, was given in [21]; unfortunately, their conditions on the weights barely miss the weight obtained by stereographic projection. (Note added: After submission of the revised version of our paper we were informed by Laurent Bétermin that in [22] the order-\(N\) term in (5) is proved with the Sandier–Serfaty method; see also [23]).

  9. Currently, only numerical evidence is available for the fourth term in the putative asymptotic expansion, and it is also conceivable that this term is actually not truly asymptotic.

  10. For a state-of-the-art survey, see [24].

  11. For a link to random polynomials, see [25]; in particular see their Thm.0.2.

  12. When the monotonicity proof was recently rediscovered [2, 3], that author remarked ([3], p. 276) that the “[monotonic increase of \(N\mapsto v_s (N)\)] and its proof are quite elementary and presumably known, yet after a serious search in the pertinent literature I came up empty-handed,\(\ldots \)”. (cf. also [2], p. 1188). M. K. likes to thank Ed Saff for subsequently pointing out to him that the monotonicity result and its proof were already given in [26], indeed. Happily, the applications of the monotonicity presented in [2] and [3] were novel.

  13. Of course, as a test criterion for empirical data concavity or strict concavity are equally fine.

  14. Actually, what is symmetric is the structure of the wave function of the electrons.

  15. Since there are protons and neutrons in the nucleus, some nuclei are “doubly magic.”

  16. Save the exactly computable data for \(N=2\) and \(N=3\).

  17. For \(N\gg 200\) failures of monotonicity were spotted in some data lists at [18]; cf. [3].

  18. We have completed all lists by computing \(v_{s}(2)=(1/s)(2^{-s}-1)\) whenever necessary.

  19. The data lists are available online in our preprint [35].

  20. For \(s=0\) and \(1\) see, respectively, also Figs. 1 and 2 in [3].

  21. In addition, the limited resolution of the plotting programs can yield deceptive plots.

  22. We note that it is futile to look for \(N\)-values for which \(\ddot{v}_{s}^{x}(N)=0\) in the empirical data.

  23. Since the \(s\)-range has not been—and cannot be—covered exhaustively with a computer, our list of 5-point and 7-point optimizers should be seen as preliminary.

  24. Recall that \(\omega _N^s\) depends on \(s\in (-\infty ,-2]\) when \(N\) is odd; recall also that at \(s=-2\) the minimizer is not unique even after factoring out \(SO(3)\), except when \(N=2\).

  25. The reason for the prefix “quasi” is the absence of a rigorous proof that the triangular bi-pyramid is the \(N=5\) optimizer for \(s\in (-2,2+\epsilon )\). For such a small \(N\)-value it is reasonable, though, to take the numerically found optimizers for granted.

  26. Digital Library of Mathematical Functions. http://dlmf.nist.gov/15.4.E20.

  27. Note that the expressions at the r.h.s.s are invariant under the permutation group \(S_N\), which is why the mapping \({\omega _N }\leftrightarrow \{\underline{\Delta } ^{(n)}_{\omega _N }\}_{n=1}^N\) is one-to-one only for unlabeled configurations.

  28. In fact, one can show these hold for all general \(N\)-point sets \(\omega _N\).

  29. Of course, the \(N=2\) configuration is also universally optimal, but “ \(\ddot{v}_s (2)\)” is ill-defined.

  30. The energy functionals \(\langle U_s \rangle ( \omega _N )\) (without the normalization \(1 / (N (N-1))\)) were studied by Wagner [39, 40] who first derived two-sided bounds for optimal \(N\)-point configurations in terms of the correct order of decay of \(N\) for the complete range \(-2 < s < 2\).

  31. It is furthermore well-known that for a sequence \(\{\omega _{N}^{s}\}_{N \ge 2}\) of optimal \(N\)-point configurations with \(-2 < s < 2\) (\(s \ne 0\)) one has the limit relation

    $$\begin{aligned} \lim _{N \rightarrow \infty } \frac{2}{N \left( N - 1 \right) } \mathop {\sum \sum }_{1 \le j < k \le N} \left| {\varvec{{ q}} }_{j,N}^{s} - {\varvec{{ q}} }_{k,N}^{s} \right| ^{-s} = W_s. \end{aligned}$$
    (99)

    The reciprocal of the quantity under the limit symbol is also known as the N-th generalized diameter of \(\mathbb {S}^2\). It converges monotonically to the generalized transfinite diameter of \(\mathbb {S}^2\), introduced by Pólya and Szegő in [42], which equals the so-called s-capacity \(1/W_s\) of \(\mathbb {S}^2\). Incidentally, it should also be noted that \(- | {\varvec{{ p}} }- {\varvec{{ q}} }|^{-s}\) is a conditionally positive definite function of order \(1\) for \(-2 < s < 0\); cf. [43].

  32. It is also well-known that, for a sequence \(\{\omega _{N}^{\mathrm {log}}\}_{N \ge 2}\) of optimal \(N\)-point configurations,

    $$\begin{aligned} \textstyle \lim \limits _{N \rightarrow \infty } \frac{2}{N \left( N - 1 \right) } \mathop {\sum \sum }\limits _{1 \le j < k \le N}\log \frac{1}{\big | {\varvec{{ q}} }_{j,N}^{\mathrm {log}}-{\varvec{{ q}} }_{k,N}^{\mathrm {log}} \big |} = W_{\mathrm {log}}. \end{aligned}$$

    The reciprocal of the exponential function of the quantity under the limit symbol is also known as the N-th diameter of \(\mathbb {S}^2\) in the logarithmic case. It converges monotonically to the transfinite diameter of \(\mathbb {S}^2\) (in the logarithmic case), introduced in [42], which equals the logarithmic capacity \(\exp (-W_{\mathrm {log}})\) of \(\mathbb {S}^2\); see [44] for a recent account.

  33. By this we only mean a vague qualitative reminiscence. Of course, Thomson [45] may have hoped to find the actual quantitative pattern of the periodic table of the chemists; for a most recent inquiry in this spirit, see [46].

  34. By Theorem 7 of [38], the infimum is not achieved by a proper \(N\)-point configuration.

  35. This was already noted by Rachmanov, Saff, and Zhou [6].

  36. An important special case of Quasi-Monte Carlo schemes are the so-called \(t\)-designs, which can be characterized by polynomial energy functionals [47].

  37. The potential function \({\varvec{{ q}} }\mapsto V_{s}( \left| {\varvec{{ q}} }\right| )\) is strictly superharmonic (i.e. \(\Delta V_{s}( \left| {\varvec{{ q}} }\right| ) < 0\) in \(\dot{\mathbb {R}}^{3}\)) for \(s\in (-2,1)\), harmonic (\(\Delta V_{s}( \left| {\varvec{{ q}} }\right| ) = 0\)) for \(s = 1\), and strictly subharmonic (\(\Delta V_{s}( \left| {\varvec{{ q}} }\right| ) > 0\)) for \(s\in (1,2)\); here, \(\Delta \) is the Laplacian in the ambient space \(\mathbb {R}^3\), and \(\dot{\mathbb {R}}^3\) is \(\mathbb {R}^3\) with its origin removed. One consequence is that \(N\)-point configurations with minimal average standardized Riesz pair-energy in the closed unit ball in \(\mathbb {R}^3\) live on its boundary \(\mathbb {S}^2\) in the superharmonic case, but extend “into the solid” in the subharmonic case as some points need to move into the volume to lower the energy (“charge injection”) (cf. [26, 51]).

  38. From [74] it readily follows that a certain limit of the leading coefficient in the asymptotic expansion of \(v_s (N)\) for large \(N\) (and the \(d\)-sphere) is closely related to the largest sphere packing density in \(\mathbb {R}^{\mathrm {d}}\). Only the densities for \(d = 1\), \(2\) and \(3\) are known, and only quite recently Hales [75] could settle the last case by proving the famous Kepler Conjecture, which states that no packing of congruent balls in Euclidean space has density greater than the density of the face-centered cubic packing (which is identical to the density of the hexagonal close packing).

  39. See [36, 76] for tables of of \(\omega _N^\infty \) and numerical values of \(\rho (N)\).

  40. Very recently, a proof for \(N=13\) has been proposed in [82].

  41. In particular, \(\rho (12)\approx 1.051462225\) implies that the 12 calotte arrangement \(\omega _{hcp}\) obtained from the hexagonal close packing of \(\mathbb {R}^3\), which has \(\varrho (\omega _{hcp})=1\), is not the optimizer of the Tammes problem with \(N=12\), which is \(\omega _{12}^\infty \): the regular icosahedron.

  42. Surprisingly, perhaps, the vertices of the Platonic cube (\(N = 8\)) have a higher average pair-energy than the square-antiprism derived from the cube by twisting (angle of 45 degrees) and squeezing together two opposite faces of the cube. Similarly, the dodecahedron (N=20) is not a minimizing configuration either, for any \(s>-2\).

  43. The numerical study for \(s\le -2\) is our own. For \(1\le s\le 400\), and \(s\rightarrow \infty \), cf. [36].

  44. Curiously, for \(-2\le s\le 0\) the optimal height of the square-pyramidal configuration is constant, equal to \(5/4\). Only for \(s>0\) does the optimized height depend on \(s\).

  45. The five point problem on the sphere can be also studied as (unconstrained) external field problem in the plane (J.B., manuscript in preparation).

  46. Our usage here of both \({\varvec{{ p}} }\) and \({\varvec{{ q}} }\) as points in space (i.e., on \(\mathbb {S}^2\)) should not be confused with the usage in Hamiltonian dynamics of \(({\varvec{{ p}} },{\varvec{{ q}} })\) as a pair of canonical variables.

  47. The investigation of the “energy integral” \(\int \!\!\int _{\mathbb {S}^2\times \mathbb {S}^2} | {\varvec{{ p}} }- {\varvec{{ q}} }|^\lambda \mu (d{\varvec{{ p}} })\mu (d{\varvec{{ q}} })\), \(\lambda \) real, can be traced back to [42].

  48. This also works for \(s<2\); cf. [51], where a “semi-continuous approach” is used to model the large-\(N\) behavior.

  49. Numerical results based on the Sobol’ points implemented in Matlab for \(N\) up to \(1\) Million points support this conjecture (cf. [100]).

References

  1. Eyink, G., Spohn, H.: Negative temperature states and large-scale, long-lived vortices in two-dimensional turbulence. J. Stat. Phys. 70, 833–886 (1993)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  2. Kiessling, M.K.-H.: The Vlasov continuum limit for the classical microcanonical ensemble. Rev. Math. Phys. 21, 1145–1195 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  3. Kiessling, M.K.-H.: A note on classical ground state energies. J. Stat. Phys. 136, 275–284 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  4. Fekete, M.: Über die Verteilung der Wurzeln bei gewissen algebraischen Gleichungen mit ganzzahligen Koeffizienten. Math. Z. 17(1), 228–249 (1923)

    Article  MATH  MathSciNet  Google Scholar 

  5. Erber, T., Hockney, G.M.: Complex systems: equilibrium configurations of \(N\) equal charges on a sphere \((2\le N\le 112)\). In Prigogine, I., Rice, S.A. (eds.) Adv. Chem. Phys. XCVIII , pp. 495–594. Wiley, New York (1997).

  6. Rakhmanov, E.A., Saff, E.B., Zhou, Y.M.: Minimal discrete energy on the sphere. Math. Res. Lett. 1, 647–662 (1994)

    Article  MATH  MathSciNet  Google Scholar 

  7. Rakhmanov, E.A., Saff, E.B., Zhou, Y.M.: Electrons on the sphere. In Ali, R.M., Ruscheweyh, S., Saff, E.B., (eds.) Computational Methods and Function Theory, pp. 111–127. World Scientific, Singapore (1995).

  8. Altschuler, E.L., Williams, T.J., Ratner, E.R., Tipton, R., Stong, R., Dowla, F., Wooten, F.: Possible global minimum lattice configurations for Thomson’s problem of charges on the sphere. Phys. Rev. Lett. 78, 2681–2685 (1997)

    Article  ADS  Google Scholar 

  9. Pérez-Garrido, A., Dodgson, M.J.W., Moore, M.A., Ortuño, M., Díaz-Sánchez, A.: Comment on: ‘possible global minimum lattice configurations for Thomson’s problem of charges on a sphere. Phys. Rev. Lett. 79, 1417 (1997)

    Article  ADS  Google Scholar 

  10. Bowick, M.J., Cacciuto, C., Nelson, D.R., Travesset, A.: Crystalline order on a sphere and the generalized Thomson problem. Phys. Rev. Lett. 89, 185502ff (2002).

  11. Bowick, M.J., Cacciuto, C., Nelson, D.R., Travesset, A.: Crystalline particle packings on a sphere with long range power law potentials. Phys. Rev. B 73, 024115ff (2006).

  12. Bendito, E., Carmona, A., Encinas, A.M., Gesto, J.M.: Estimation of Fekete points. J. Comp. Phys. 225, 2354–2376 (2007)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  13. Wales, D.J., Ulker, S.: Structure and dynamics of spherical crystals characterised for the Thomson problem. Phys. Rev. B 74, 212101 (2006)

    Article  ADS  Google Scholar 

  14. Wales, D.J., Mackay, H., Altshuler, E.L.: Defect motifs for spherical topologies. Phys. Rev. B 79, 224115 (2009)

  15. Wales, D.J.: Energy Landscapes: Applications to Clusters, Biomolecules and Glasses. Cambridge University Press, Cambridge (2004)

    Book  Google Scholar 

  16. Calef, M.T., Fichtl, Ch.A., Goulart, W.C., Hardin, D.P., Schulz, A.E.: Asymptotic differences in energies of stable and minimal point configurations on \({\mathbb{S}}^2\) and the role of defects. J. Math. Phys. 54:101901, 20 (2013).

  17. Beltran, C.: Harmonic properties of the logarithmic potential and the computability of elliptic Fekete points. Constr. Approx. 37, 135–165 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  18. Bowick, M.J., Cecka, C., Giomi, L., Middleton, A., Zielnicki, K.: Thomson Problem @ S.U. http://thomson.phy.syr.edu/

  19. Smale, S.: Mathematical problems for the next century. Math. Intell. 20, 7–15 (1998); see also version 2 on Steve Smale’s home page. http://www6.cityu.edu.hk/ma/doc/people/smales/pap104

  20. Brauchart, J.S., Hardin, D.P., Saff, E.B.: The next-order term for optimal Riesz and logarithmic energy asymptotics on the sphere. In Recent advances in orthogonal polynomials, special functions, and their applications Arvesú, J., López Lagomasino, G. (eds.), Contemporary Mathematics, vol. 578, pp. 31–61. AMS, Providence, RI (2012)

  21. Sandier, E., Serfaty, S.: 2D Coulomb gases and the renormalized energy. arXiv:1201.3503v2 (2013); to appear in Annals Prob.

  22. Bétermin, L.: Renormalized energy and asymptotic expansion of optimal logarithmic energy on the sphere. Submitted to LAMA (2014). arXiv:1404.4485v2 (2014).

  23. Bétermin, L., Zhang, P.: Minimization of energy per particle among Bravais lattices in \({\mathbb{R}}^2\): Lennard–Jones and Thomas-Fermi cases. arXiv:1402.2751v2 (2014).

  24. Beltran, C.: The state of the art in Smale’s 7th problem. In: Foundations of computational mathematics, Budapest 2011. London Mathematical Society Lecture Note Series 403, 1–15 (2013).

  25. Armentano, D., Beltrán, C., Shub, M.: Minimizing the discrete logarithmic energy on the sphere: the role of random polynomials. Trans. Am. Math. Soc. 363, 2955–2965 (2011)

    Article  MATH  Google Scholar 

  26. Landkof, N.S.: Foundations of Modern Potential Theory. Springer, Berlin (1972)

    Book  MATH  Google Scholar 

  27. Kuijlaars, A.B.J., Saff, E.B.: Asymptotics for minimal discrete energy on the sphere. Trans. Am. Math. Soc. 350, 523–538 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  28. Hardin, D.P., Saff, E.B.: Minimal Riesz energy point configurations for rectifiable d-dimensional manifolds. Adv. Math. 193, 174–204 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  29. Brauchart, J.S., Hardin, D.P., Saff, E.B.: The Riesz energy of the Nth roots of unity: an asymptotic expansion for large N. Bull. London Math. Soc. 41, 621–633 (2009)

  30. Brauchart, J.S.: A remark on exact formulas for the Riesz energy of the \(N\)th roots of unity. arXiv:1105.5530v1 (2011).

  31. Calef, M.T.: Theoretical and computational investigations of minimal energy problems, Thesis (Ph.D.), Vanderbilt University, 2009.

  32. Hardin, R.H., Sloane, N.J.A., Smith, W.D.: Minimal energy arrangements of points on a sphere. \(\copyright \) (1994) HSS; http://neilsloane.com/electrons

  33. Wales, D.J., Ulker, S.: Global Minima for the Thomson Problem. http://www-wales.ch.cam.ac.uk/wales/CCD/Thomson/table.html

  34. Morris, J.R., Deaven, D.M., Ho, K.M.: Genetic-algorithm energy minimization for point charges on a sphere. Phys. Rev. B 53(14), R1740–R1743 (1996)

    Article  ADS  Google Scholar 

  35. Nerattini, R., Brauchart, J.S., Kiessling, M.K.-H.: “Magic” numbers in Smale’s 7th problem. arXiv:1307.2834 (2013).

  36. Melnyk, T.W., Knop, O., Smith, W.R.: Extremal arrangements of points and unit charges on a sphere: equilibrium configurations revisited. Can. J. Chem. 55(10), 1745–1761 (1977)

    Article  MathSciNet  Google Scholar 

  37. Berman, J., Hanes, K.: Optimizing the arrangement of points on the unit sphere. Math. Comp. 31, 1006–1008 (1977)

    Article  MATH  MathSciNet  Google Scholar 

  38. Björck, G.: Distributions of positive mass, which maximize a certain generalized energy integral. Ark. Mat. 3, 255–269 (1956)

    Article  MATH  MathSciNet  Google Scholar 

  39. Wagner, G.: On means of distances on the surface of a sphere (lower bounds). Pac. J. Math. 144, 389–398 (1990)

    Article  ADS  MATH  Google Scholar 

  40. Wagner, G.: On means of distances on the surface of a sphere II. Upper bounds. Pac. J. Math. 154, 381–396 (1992)

    Article  MATH  Google Scholar 

  41. Hardin, D.P., Saff, E.B.: Discretizing manifolds via minimum energy points. Notices AMS 51, 1186–1194 (2004)

    MATH  MathSciNet  Google Scholar 

  42. Pólya, G., Szegö, G.: Über den transfiniten Durchmesser (Kapazitätskonstante) von ebenen und räumlichen Punktmengen. J. Reine Angew. Math. 165, 4–49 (1931)

    Google Scholar 

  43. Schoenberg, I.J.: Metric spaces and positive definite functions. Trans. Am. Math. Soc. 44(3), 522–536 (1938)

    Article  MathSciNet  Google Scholar 

  44. Pritsker, I.E.: Equidistribution of points via energy. Ark. Mat. 49, 149–173 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  45. Thomson, J.J.: On the structure of the atom: an investigation of the stability and periods of oscillation of a number of corpuscles arranged at equal intervals around the circumference of a circle; with application of the results to the theory of atomic structure. Philos. Mag. 7, 237–265 (1904)

    Article  MATH  MathSciNet  Google Scholar 

  46. LaFave, T.: Correspondence between the classical electrostatic Thomson problem and atomic electronic structure. J. Electrostat. 71, 1029–1035 (2013)

    Article  Google Scholar 

  47. Sloan, I.H.,Womersley, R.S.: A variational characterisation of spherical designs. J. Approx. Theory 159, 308–318 (2009)

  48. Brauchart, J.S., Dick, J.: A characterization of Sobolev spaces on the sphere and an extension of Stolarsky’s invariance principle to arbitrary smoothness. Constr. Approx. 38, 397–445 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  49. Brauchart, J.S., Dick, J.: A simple proof of Stolarsky’s invariance principle. Proc. Am. Math. Soc. 141, 2085–2096 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  50. Brauchart, J.S., Saff, E.B., Sloan, I.H., Womersley, R.S.: QMC designs: optimal order Quasi Monte Carlo integration schemes on the sphere. Math. Comp. 2014. DOI: 10.1090/S0025-5718-2014-02839-1

  51. Berezin, A.A.: Asymptotics of the maximum number of repulsive particles on a spherical surface. J. Math. Phys. 27, 1533–1536 (1986)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  52. Beck, J.: Sums of distances between points on a sphere—an application of the theory of irregularities of distribution to discrete geometry. Mathematica 31, 33–41 (1984)

    MATH  Google Scholar 

  53. Fejes Tóth, L.: On the sum of distances determined by a point set. Acta. Math. Acad. Sci. Hungar. 7, 397–401 (1956).

  54. Stolarsky, K.B.: Spherical distributions of \(N\) points with maximal distance sums are well spaced. Proc. Am. Math. Soc. 48, 203–206 (1975)

    MATH  MathSciNet  Google Scholar 

  55. Kirchhoff, G.: Vorlesungen uber mathematische Physik. Mechanik, Chap. 20, 3rd ed., Teubner, Leipzig (1883).

  56. Chanillo, S., Kiessling, M.K.-H.: Rotational symmetry of solutions of some nonlinear problems in statistical mechanics and in geometry. Commun. Math. Phys. 160, 217–238 (1994)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  57. Forrester, P.J.: Log-gases and random matrices. Princeton University Press, Princeton (2010)

    MATH  Google Scholar 

  58. Forrester, P.J., Jancovici, B., Madore, J.: The two-dimensional Coulomb gas on a sphere: exact results. J. Stat. Phys. 69, 179–192 (1992)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  59. Kiessling, M.K.-H., Wang, Y.: Onsager’s ensemble for point vortices with random circulations on the sphere. J. Stat. Phys. 148, 896–932 (2012)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  60. Berkenbusch, M.K., Claus, I., Dunn, C., Kadanoff, L.P., Nicewicz, M., Venkataramani, S.C.: Discrete charges on a two dimensional conductor. J. Stat. Phys. 116, 1301–1358 (2004)

  61. Newton, P.K., Chamoun, G.: Vortex lattice theory: a particle interaction perspective. SIAM Rev. 51(3), 501–542 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  62. Saff, E.B., Totik, V.: Logarithmic potentials with external fields. Grundl. d. Math. Wiss. vol. 316. Springer, Berlin (1997).

  63. Shub, M., Smale, S.: Complexity of Bezout’s theorem, III: condition number and packing. J. Complex. 9, 4–14 (1993)

    Article  MATH  MathSciNet  Google Scholar 

  64. Whyte, L.L.: Unique arrangements of points on a sphere. Am. Math. Monthly 59, 606–611 (1952)

    Article  MATH  MathSciNet  Google Scholar 

  65. Calef, M.T., Hardin, D.P.: Riesz \(s\)-equilibrium measures on \(d\)-rectifiable sets as \(s\) approaches \(d\). Potential Anal. 30(4), 385–401 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  66. Manev, G.: Die gravitation und das Prinzip von Wirkung und Gegenwirkung. Z. Phys. 31, 786–802 (1925)

    Article  ADS  Google Scholar 

  67. Bobylev, A.V., Ibragimov, NKh: Interconnectivity of symmetry properties for equations of dynamics, kinetic theory of gases, and hydrodynamics. Matem. Mod. 1, 100–109 (1989). (in Russian)

    MATH  MathSciNet  Google Scholar 

  68. Calogero, F., Leyvraz, F.: A macroscopic system with undamped periodic compressional oscillations. J. Stat. Phys. 151, 922–937 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  69. Lynden-Bell, D., Lynden-Bell, R.M.: Exact general solutions to extraordinary \(N\)-body problems. Proc. R. Soc. Lond. A 445, 475–489 (1999).

  70. Calogero, F.: Solution of the one-dimensional N-body problem with quadratic and/or inversely quadratic pair potentials. J. Math. Phys. 12, 419–436 (1971). Erratum, ibid. 37:3646 (1996).

  71. Moser, J.: Three integrable Hamiltonian systems connected with isospectral deformations. Adv. Math. 16, 197–220 (1975)

    Article  ADS  MATH  Google Scholar 

  72. Tammes, R.M.L.: On the origin number and arrangement of the places of exits on the surface of pollengrains. Rec. Trv. Bot. Neerl. 27, 1–84 (1930)

    Google Scholar 

  73. Conway, J.H., Sloane, N.J.A.: Sphere packings, lattices and groups. With additional contributions by Bannai, E., Borcherds, J. Leech, R.E., Norton, S.P., Odlyzko, A.M., Parker, R.A., Queen, L., Venkov, B.B. (3rd ed.). Grundlehren der Math. Wiss. vol. 290. Springer, New York (1999).

  74. Borodachov, S.V., Hardin, D.P., Saff, E.B.: Asymptotics of best-packing on rectifiable sets. Proc. Am. Math. Soc. 135(8), 2369–2380 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  75. Hales, T.C.: A proof of the Kepler conjecture. Ann. Math. (2) 162(3), 1065–1185 (2005).

  76. Clare, B.W., Kepert, D.L.: The closest packing of equal circles on a sphere. Proc. Roy. Soc. London Ser. A 405(1829), 329–344 (1986)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  77. Böröczky, K.: The problem of Tammes for \(n=11\). Studia Sci. Math. Hungar. 18(2–4), 165–171 (1983)

    MATH  MathSciNet  Google Scholar 

  78. Danzer, L.: Finite point-sets on \(S^2\) with minimum distance as large as possible. Discret. Math. 60, 3–66 (1986)

    Article  MATH  MathSciNet  Google Scholar 

  79. Fejes Tóth, L.: Über eine Abschätzung des kürzesten Abstandes zweier Punkte eines auf einer Kugelfläche liegenden Punktsystems. Jber. Deutsch. Math. Verein. 53, 66–68 (1943).

  80. Robinson, R.M.: Arrangements of 24 points on a sphere. Math. Ann. 144, 17–48 (1961)

    Article  MATH  MathSciNet  Google Scholar 

  81. Schütte, K., van der Waerden, B.L.: Auf welcher Kugel haben 5, 6, 7, 8 oder 9 Punkte mit Mindestabstand Eins Platz? Math. Ann. 123, 96–124 (1951)

    Article  MATH  MathSciNet  Google Scholar 

  82. Musin, O., Tarasov, A.: The strong thirteen spheres problem. Discret. Comput. Geom. 48, 128–141 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  83. Bachoc, C., Vallentin, F.: New upper bounds for kissing numbers from semidefinite programming. J. Am. Math. Soc. 21(3), 909–924 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  84. Cohn, H., Kumar, A.: Universally optimal distribution of points on spheres. J. Am. Math. Soc. 20(1), 99–148 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  85. Leech, J.: Equilibrium of sets of particles on a sphere. Math. Gaz. 41, 81–90 (1957)

    Article  MATH  MathSciNet  Google Scholar 

  86. Dragnev, P.D., Legg, D.A., Townsend, D.W.: Discrete logarithmic energy on the sphere. Pac. J. Math. 207(2), 345–358 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  87. Hou, X., Shao, J.: Spherical distribution of 5 points with maximal distance sum. Discret. Comput. Geom. 46(1), 156–174 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  88. Schwartz, R.E.: The 5 electron case of Thomson’s problem. Exper. Math. 22, 157–186 (2013)

    Article  MATH  Google Scholar 

  89. Kiessling, M.K.-H., Spohn, H.: A note on the eigenvalue density of random matrices. Commun. Math. Phys. 199, 683–695 (1999)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  90. Szegő, G.: Bemerkungen zu einer Arbeit von Herrn M. Fekete: Über die Verteilung der Wurzeln bei gewissen algebraischen Gleichungen mit ganzzahligen Koeffizienten. Math. Z. 21, 203–208 (1924)

    Article  MathSciNet  Google Scholar 

  91. Saff, E.B., Kuijlaars, A.B.J.: Distributing many points on a sphere. Math. Intell. 19, 5–11 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  92. Sadoc, J.-F., Mosseri, R.: Geometrical Frustration. Cambridge University Press, Cambridge (1999)

    Book  Google Scholar 

  93. Bowick, M.J., Giomi, L.: Two-dimensional matter: order, curvature and defects. Adv. Phys. 58(5), 449–563 (2009)

    Article  ADS  Google Scholar 

  94. Womersley, R.S.: Robert Womersley’s home page. http://web.maths.unsw.edu.au/~rsw/.

  95. Atiyah, M., Sutcliffe, P.: Polyhedra in physics, chemistry, and geometry. Milan J. Math. 71, 33–58 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  96. Brauchart, J.S., Womersley, R.S.: Weighted QMC designs: numerical integration over the unit sphere, \(\mathbb{L}_2\)-discrepancy and sum of distances. Manuscript in preparation.

  97. Niederreiter, H.: Point sets and sequences with small discrepancy. Monatsh. Math. 104(4), 273–337 (1987)

    Article  MATH  MathSciNet  Google Scholar 

  98. Dick, J., Pillichshammer, F.: Digital nets and sequences. Discrepancy Theory and Quasi-Monte Carlo Integration. Cambridge University Press, Cambridge (2010)

  99. Joe, S., Kuo, F.Y.: Remark on algorithm 659: implementing Sobol’s quasirandom sequence generator. ACM Trans. Math. Softw. 29, 49–57 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  100. Brauchart, J.S., Dick, J.: Quasi-Monte Carlo rules for numerical integration over the unit sphere \({\mathbb{S}}^2\). Numer. Math. 121(3), 473–502 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  101. Aistleitner, Ch., Brauchart, J.S., Dick, J.: Point sets on the sphere \(\mathbb{S}^2\) with small spherical cap discrepancy. Discrete Comput. Geom. 48(4), 990–1024 (2012)

    MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

J.B. and M.K thank Ed Saff for his role as “matchmaker” that started our collaboration at the conference OPTIMAL 2010 at Vanderbilt University, Nashville, Tennessee, and for interesting comments. R.N. and M.K. thank Michael Kastner for providing the opportunity to get our collaboration started at the 2011 STIAS workshop on “Equilibrium and Equilibration” in Stellenbosch, South Africa, and Lapo Casetti for his support of our collaboration and his interesting comments on energy landscapes. We would like to thank Rob Womersley (UNSW) for numerical data and helpful discussions regarding the numerics. J.B. greatfully acknowledges partial support by an APART-Fellowship of the Austrian Academy of Sciences and the hospitality of the School of Mathematics and Statistics at UNSW and the support of the Australian Research Council; R.N. and M.K. greatfully acknowledge partial financial support by the NSF through grant DMS 0807705, and by the INFN, Sezione di Firenze. Lastly, we thank the referee for the careful reading of the original manuscript and for constructive criticism.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. K.-H. Kiessling.

Additional information

In celebration of Doron Zeilberger’s \((1+\surd \epsilon )\cdot 60\)-th birthday.

Appendices

Appendix 1: Optimal Riesz Energy Configurations on \(\mathbb {S}^2\): A Brief Survey

The problem to determine \(v_s (N)\) together with the minimizing configuration(s) \(\omega _N^s\) has been solved completely for all \(N\ge 2\) only at the distinguished value \(s=-2\), by explicit calculation. Namely, \(s=-2\) yields the energy law for the completely integrable Newtonian \(N\)-body problem with repulsive harmonic forces. Any \(N\)-point configuration satisfying \(\sum _{i=1}^N {\varvec{{ q}} }_i = \mathbf {0}\) is a minimizing configuration of \(\langle V_s\rangle (\omega _N)\), and only such are. The minimal energy reads

$$\begin{aligned} v_{-2}(N) = -\frac{1}{2}\frac{N+1}{N-1}. \end{aligned}$$
(109)

The (presumably) next-simplest parameter regime is \(s<-2\). Here one is confronted with the possibly startling observation that for large \(N\) the \(N\)-tuple Fekete points accumulate around two opposite points, and the localization sharpens as \(N\) is getting larger; this is a consequence of Theorem 7 in [38]. In particular, it follows right away from Theorem 7 in [38] that for even \(N\) the infimum \(v_s (N),\, s<-2\), is achievedFootnote 34 if and only if half of the particles each are placed at two antipodal points,Footnote 35 yielding

$$\begin{aligned} v_s (N) = -\frac{1}{|s|}\,\frac{\left( 2^{|s|-1}-1 \right) N+1}{N-1},\qquad s< -2,\qquad N=2n, \end{aligned}$$
(110)

which converges to \(v_{-2}(N)\) when taking the limit \(s\uparrow -2\) of (110). When \(N\) is odd the situation is already more tricky, and more interesting! For instance, for the smallest allowed odd \(N=3\) it is suggestive to conjecture that the minimizing configuration consists of the corners of an equilateral triangle in an arbitrary equatorial plane; yet comparison with an antipodal “configuration” (arrangement) with two labeled points in the North and one in the South Pole reveals that the equilateral configuration yields a lower average standardized Riesz pair-energy only for \(s_3 < s < -2\), where \(s_3 \equiv \ln (4/9)/\ln (4/3)\), while for \(s < s_3\) the antipodal arrangement yields the lower average standardized Riesz pair-energy; in this case one can easily show rigorously that the antipodal arrangement is in fact optimal: namely, the equilateral triangle and the antipodal arrangement are the only equilibrium arrangements of 3 labeled points. When comparing the average standardized Riesz pair-energy for antipodal and equilateral arrangements for other odd \(N\), this changeover happens only if \(N\) is a multiple of \(3\). The critical \(s_{3(2n-1)}\) tends monotonically to \(-2\) as \(N =3(2n-1)\rightarrow \infty \). Of course, this does not prove that either arrangement is optimal in the respective range of \(s\). To the best of our knowledge, the optimal arrangement of odd-\(N\) points as a function of \(s<-2\) is far from being settled.

The concentration of the minimizing “\(N\)-point configuration” for \(s< -2\) at a few distinct points indicates that the optimization problem is incorrectly posed in the set of proper \(N\)-point configurations. (The deeper reason is that the Riesz pair-energy ceases to be positive definite in the sense of Schoenberg [43] for \(s < -2\).) Interestingly, the sum of distance problem for \(s < -2\) plays a central role in the theory of Quasi-Monte Carlo integration schemesFootnote 36 for functions in smooth enough function spaces over \(\mathbb {S}^2\); we refer the interested reader to [4850] and papers cited therein.

When \(s> -2\) the problem becomes drastically more complicated. One needs to distinguish the cases \(-2<s<2\), \(s=2\), \(s>2\), and the limit \({s\rightarrow \infty }\).

The interval \(-2<s<2\) is known as the potential-theoretical regime, since concepts and methods of potential theory can be applied to study both the discrete and the continuous (i.e. \(N\rightarrow \infty \)) optimization problems.Footnote 37 Within this regime the integer values \(s=-1\), \(s=0\), and \(s=1\) are of particular interest. When \(s=-1\) the minimal average standardized Riesz pair-energy problem is equivalent to the maximal average pairwise chordal distance problem; see [5254]. In [49] it is shown that maximum-sum-of-distance configurations are ideal integration nodes for a certain optimal-order Quasi-Monte Carlo integration scheme on \(\mathbb {S}^2\); we will come back to this in Appendix . The case “\(s=0\),” i.e. the limit \(s\rightarrow 0\), which yields the logarithmic pair-energy 2 (also known as the Coulomb energy for a pair of “two-dimensional unit point charges” on \(\mathbb {S}^2\), respectively the Kirchhoff energy [55] of a pair of unit point vortices on \(\mathbb {S}^2\)), occurs in a stunning variety of problems (on \(\mathbb {S}^2\) and other manifolds) in the sciences and mathematics; see, e.g. [10, 5663]. Originally Smale’s 7th problem for the 21st century [19] was formulated for the logarithmic energy. Lastly, the value \(s=1\) yields the Coulomb pair-energy of “three-dimensional unit point charges” associated with the so-called Thomson problem (see [5, 8, 9, 18, 45, 46, 64]).

Amongst the values \(s\ge 2\), the borderline value \(s=2\) is special in the sense that the finite-\(N\) behavior is qualitatively different from both, the regime \(-2<s<2\), and the regime \(s>2\). Yet it can be understood by considering a certain limit process \(s \rightarrow 2\); cf. [65] for the limit process \(s\rightarrow d\) in analogous optimization problems formulated on \(d\)-dimensional manifolds. The Riesz pair interaction for \(s=2\), in physics considered as correction term to Newton’s gravity [66], is also special in the sense that it yields a Newtonian \(N\)-body problem in \(\mathbb {R}^3\) with additional isolating integrals of motion [6769], besides those associated with Galilei invariance. Restricted to \(\mathbb {R}\) the motion is even completely integrable for all \(N\) [70, 71].

The large-\(s\) behavior of \(\langle V_s\rangle (\omega _N)\) (\(N\) fixed) is intimately connected with the classical Tammes’s problem ([72]) or hard sphere (best-packing) problem (cf. [73]); that is, to find a configuration of \(N\) points on the sphere with the minimal pairwise (chordal) distance between the points being as large as possible.Footnote 38 It is not too difficult to see (cf. Appendix ) that for any \(N\)-point configuration \(\omega _N = \{ {\varvec{{ q}} }_1, \dots , {\varvec{{ q}} }_N \} \subset \mathbb {S}^2\) the following limit relation holds:

$$\begin{aligned} \lim _{s \rightarrow \infty } \left[ \langle V_s\rangle (\omega _N) +{\textstyle \frac{1}{s}}\right] ^{-1/s} = \min _{1 \le i < j \le N} \left| {\varvec{{ q}} }_i - {\varvec{{ q}} }_j \right| \equiv \varrho ( \omega _N ). \end{aligned}$$
(111)

Moreover, whenever a family of minimizing configurations \(\omega _N^s\) converges to a limit configuration \(\omega _N^\infty \) one has the relation (cf. Appendix )

$$\begin{aligned} \lim _{s \rightarrow \infty } \left[ v_s (N) +{\textstyle \frac{1}{s}}\right] ^{-1/s} = \varrho ( \omega _N^\infty ) \equiv \rho (N), \end{aligned}$$
(112)

where \(\rho (N)\) is the best-packing (chordal) distance, which maximizes the least distance \(\varrho (\omega _N)\) among all \(N\)-point configurations on \(\mathbb {S}^2\), and \(\omega _N^\infty \) is the best-packing configuration.Footnote 39 The best-packing distance \(\rho (N)\) is only known for \(N = 2,\ 3,\ \dots ,\ 12\) and \(24\) (cf. [7781]).Footnote 40

Another way of obtaining a nontrivial limit problem is to let \(s\rightarrow \infty \) in \(V_s(r)\), which gives

$$\begin{aligned} V_{\infty }(r) \equiv \lim _{s\rightarrow \infty } V_{s}(r) = {\left\{ \begin{array}{ll} \infty &{} \text {if } r < 1, \\ 0 &{} \text {if } r \ge 1. \end{array}\right. } \end{aligned}$$
(113)

In that case \(v_\infty (N) = 0\) for \(N \le N_*\), while \(v_\infty (N) = \infty \) for \(N > N_*\), viz. \(N_*\) is the maximum number of non-overlapping calottes with spherical radius \(\pi /6\) which can be placed on the unit sphere. By picking any ball, \(B\), from a hexagonal close packing (hcp) of \(\mathbb {R}^3\) with unit balls, then projecting its 12 nearest neighbors (unit balls) radially onto the surface of the central ball \(B\), one sees that \(N_*\ge 12\). And dividing the surface area of the unit sphere, \(4\pi \), by the area of a calotte, \(\pi (2-\sqrt{3})\), yields \(\approx 14.92820323\), giving the upper bound \(N_*< 15\). But how large is \(N_*\), exactly?

Interestingly, the sharp value for \(N_*\) is found by studying the related Tammes problem. For \(2 \le N \le 12\) one hasFootnote 41 \(\rho (N) > 1\). L. Fejes Tóth’s famous inequality ([79]),

$$\begin{aligned} \left[ \rho (N) \right] ^2 \le 4 - \Bigl [ \mathrm {cosec}\Big ( \frac{\pi }{6} \frac{N}{N-2} \Big ) \Bigr ]^2, \end{aligned}$$

where equality holds only for \(N = 3\), \(4\), \(6\) and \(12\), gives \(\rho (N) < 1\) for \(N \ge 14\). From [83] follows \(\rho (13) < 1\). Hence, \(N_* = 12\).

To our best knowledge, the following point sets are the only ones for which one can rigorously prove that they have minimal average standardized Riesz pair-energy for all \(s > -2\). One can easily characterize the minimizing configuration explicitly only when \(N = 2\) or \(3\) (as the antipodal and equilateral configuration, respectively). The minimizing configuration has been characterized explicitly also for \(N = 4\), \(6\), and \(12\) as the vertices of Platonic solidsFootnote 42 (tetrahedron, octahedron, and icosahedron), which are known to be universally optimal (see [84]); such configurations minimize the potential energy of completely monotonic pair-energy functions. The standardized Riesz pair-energies for \(s > -2\) (including the logarithmic pair-energy at \(s=0\)) fall into this category. The listed configurations for \(N = 2\), \(3\), \(4\), \(6\), and \(12\) exhaust the possibilities for universally optimal configurations on \(\mathbb {S}^2\); cf. [84, 85].

The surprisingly difficult task of finding a proof of minimality can, perhaps, be best illustrated with the only partly resolved five point problem on \(\mathbb {S}^2\). It is clear from [84, Prop. 14] that there is no universally optimal \(5\)-point configuration on \(\mathbb {S}^2\). Indeed, computational optimization reveals that the minimal-energy arrangement of five labeled points on \(\mathbb {S}^2\) changes many times as \(s\) varies over the real line.Footnote 43 Thus, for \(s \le -2.368335\ldots \) an antipodal arrangement with two labeled points in the South, and three in the North Pole (say) is the optimizer; at \(s=-2.368335\ldots \) a crossover takes place, and for \(-2.368335\ldots \le s \le -2\) the energy-minimizing arrangement of five labeled points is an isosceles triangle on a great circle, with one point in the North Pole and two labeled points each in the other two corners, with (numerically) optimized height. At \(s=-2\) the isosceles arrangement bifurcates off of a continuous family of rectangular pyramids with height \(h=5/4\), all of which have the same energy \(-3/4\) at \(s=-2\), and of which the isosceles arrangement is the degenerate limit. At \(s=-2\) also another crossover happens, and for \(-2\le s\le 15.048077392\dots \) the regular triangular bi-pyramid is the putative energy-minimizing configuration. At \(s = 15.048077392\dots \), yet another crossover happens, at which the triangular bi-pyramid and a square pyramid with height \(h\approx 1.1385\) have the same average (standardized) Riesz pair-energy. A square pyramid with (numerically) optimized heightFootnote 44 as function of \(s\) appears to have lower (standardized) Riesz pair-energy for \(s\in [15.04807\ldots ,\infty )\). Lastly, it is well-known that the triangular bi-pyramid and the square pyramid with height \(1\) both are particular best-packing configurations, with \(\varrho (\omega _5^\infty )=\sqrt{2}\), so that “at \(s=\infty \)” an “asymptotic crossover, or degenerate bifurcation,” happens.

How much of this has been proved rigorously? By traditional methods (see [86]) it can be shown that the triangular bi-pyramid consisting of two antipodal points at, say, the North and the South Pole, and three equally spaced points on the Equator, is the unique (up to orthogonal transformation) minimizer of the logarithmic average pair-energy. The proof that the same configuration maximizes the sum of distances (that is: assumes \(v_{-1}(5)\)) is computer-aided, exploiting interval methods and related techniques (see [87]). In [88] a computer-aided approach is proposed to show optimality of the triangular bi-pyramid for \(s = 1\) and \(s = 2\). The optimality of both the triangular bi-pyramid and the family of rectangular pyramids with height \(5/4\) at \(s = -2\) can be shown with elementary techniques. The rest of the \(s\)-parameter regime still awaits its rigorous treatment.Footnote 45

Numerical results in [36], carried out with varying \(s\in [1,400]\) for \(N\in \{2,\ldots ,16\}\) fixed, suggest that also for other values of \(N\not \in \{2,3,4,6,12\}\) the minimizing configuration \(\omega _N^s\) may generally change as \(s\) passes through critical values, and their number seem to depend on \(N\). In particular, for \(N=7\) there seem to be three(!) critical \(s\)-values in \([1,6]\) at which the minimizing configuration changes, and presumably a few more when \(s<1\), cf. [37] for \(s=-1\). The general dependence on \(s\) of the optimal \(N\)-point configurations \(\omega _N^s\) is one of the intriguing features of this minimization problem. All the same, it makes it plain why the rigorous determination of the optimizers is a highly nontrivial task even for moderate \(N\)-values other than the special ones \(2,3,4,6,12\), becoming hopelessly complicated when \(N\) increases.

Yet, the large-\(N\) asymptotics of the minimal average standardized Riesz pair-energy \(v_s (N)\) can be determined without seeking the exact Fekete points, see [6]. In particular, \(\lim _{N\rightarrow \infty }v_s (N)\) is for all \(s\) determined by the variational principleFootnote 46 (see [26, 38, 89, 90])

$$\begin{aligned} \lim _{N\rightarrow \infty }v_s (N) = \inf _{\mu \in \mathfrak {P}(\mathbb {S}^2)} \int \int _{\mathbb {S}^2\times \mathbb {S}^2} V_s(|{\varvec{{ p}} }-{\varvec{{ q}} }|)\mu (d{\varvec{{ p}} })\mu (d{\varvec{{ q}} }); \end{aligned}$$
(114)

here, \(\mathfrak {P}(\mathbb {S}^2)\) is the set of all Borel probability measures supported on \(\mathbb {S}^2\). For \(s\le -2\) the minimizer is not unique, but all minimizers are known; in particular, for \(s<-2\) the minimizer, after factoring out \(SO(3)\), is a symmetric measure which is concentrated on two antipodal points, see [38]. From classical potential theory (cf. [26] for \(s\in [0,2)\) and [38] for \(s\in (-2,0)\))Footnote 47 it is well-known that the uniform normalized (Lebesgue) surface area measure on \(\mathbb {S}^2\), denoted by \(\sigma \), uniquely minimizes the right-hand side in (114) for \(-2 < s < 2\). For \(s\ge 2\) the l.h.s. and r.h.s. of (114) are both \(\infty \); in this case the rate of divergence of \(v_s (N)\) can be determined. It “suffices” to know that for large \(N\) the Voronoi cells around the charges are mostly hexagons of a certain size; see [91] for an enlightening discussion.Footnote 48 The picture one should have in mind, when \(N\) is large, is a vast sea of hexagonal Voronoi cells around most of the points. Thus, the dual structure of the Voronoi cell decomposition, the Delaunay triangulation, is a network of mostly six-fold coordinated sites. The reason for the qualification “mostly” lies in the topology of the sphere, which gives rise to geometric frustration (see [92] for a thorough exposition of this notion). Certain points “pick up” a topological charge that measures the departure from the ideal coordination number, \(6\), of the planar triangular lattice. The celebrated Euler theorem of topology yields that the total topological charge on \(\mathbb {S}^2\) is always 12. This accounts, for example, for the appearance of 12 (isolated) pentagons in the common soccer ball design. For large \(N\) one observes “scars” emerging from these isolated centers that attract pentagon-heptagon pairs (having total topological charge \(0\)). Scars and other topology-induced defects of the hexagonal lattice become important when pushing the asymptotic analysis to higher order, and are not well understood. For instance, to the best of our knowledge it is an unresolved question if there are \(n\)-gon Voronoi cells with \(n \ge 8\) in a minimizing configuration. See [93] for an approach using elastic continuum formalism.

The truly hard regime is the vast intermediate range of \(N\) which are generically too large to allow for an explicit determination of the minimizing configuration, but not large enough for the asymptotic formulas to yield sufficiently accurate results.

This concludes our brief survey of this fascinating field. Further information can be found in the survey articles [5, 41],[91], and on the websites [18] and [94]. See also the delightful article [95] where, based on numerical evidence, the first dozen minimizers are discussed mostly for Thomson’s problem (\(s=1\)).

Appendix 2

In this appendix we supply the proofs of relations (111) and (112) which control the limit \(s\rightarrow \infty \).

1.1 Proof of Relation (111)

Suppose \(\omega _N = \{ {\varvec{{ q}} }_1, \dots , {\varvec{{ q}} }_N \} \subset \mathbb {S}^2\) is a fixed \(N\)-point set, with separation distance \(\varrho ( \omega _N ) = \min _{1 \le i < j \le N} | {\varvec{{ q}} }_i - {\varvec{{ q}} }_j |\). Then using that the function \(f(x) \equiv x^{-1/s}\) is strictly decreasing for \(s > 0\), we find that

$$\begin{aligned} \left[ \langle V_s\rangle (\omega _N) +{\frac{1}{s}}\right] ^{-1/s}&= \left[ \frac{1}{s} \frac{2}{N(N-1)} \mathop {\sum \sum }_{1 \le i < j \le N} \frac{1}{\left| {\varvec{{ q}} }_i - {\varvec{{ q}} }_j \right| ^{s}} \right] ^{-1/s} \\&= \varrho ( \omega _N ) \left[ \frac{1}{s} \frac{2}{N(N-1)}\mathop {\sum \sum }_{1 \le i < j \le N} \left( \frac{\varrho (\omega _N)}{\left| {\varvec{{ q}} }_i - {\varvec{{ q}} }_j\right| }\right) ^{s}\right] ^{-1/s}\\&\ge \varrho ( \omega _N ) \left( \frac{1}{s} \right) ^{-1/s} \\&\ge \varrho ( \omega _N ). \end{aligned}$$

On the other hand, retaining only one of the least distance pairs in the double sum yields

$$\begin{aligned} \left[ \langle V_s\rangle (\omega _N) +{\frac{1}{s}}\right] ^{-1/s}&\le \left( \frac{1}{s} \frac{2}{N(N-2)} \frac{1}{\varrho (\omega _N)^s}\right) ^{-1/s}\\&= \varrho ( \omega _N ) \left( \frac{1}{s} \right) ^{-1/s} \left( \frac{2}{N(N-2)} \right) ^{-1/s} \\&\rightarrow \varrho ( \omega _N )\qquad \text{ as }\quad s\rightarrow \infty . \end{aligned}$$

This completes the proof of (111).

1.2 Proof of Relation (112)

Let \(\omega _N^s = \{ {\varvec{{ q}} }_1^s, \dots , {\varvec{{ q}} }_N^s \} \subset \mathbb {S}^2\) denote a minimizing \(N\)-point set, and suppose \(\omega _N^\infty \) is a best-packing configuration with \(\varrho ( \omega _N^\infty ) = \rho (N)\).

Then, first of all,

$$\begin{aligned} \langle V_s \rangle (\omega _N^s ) \le \langle V_s \rangle ( \omega _N^\infty ); \end{aligned}$$

but \(\langle V_s \rangle (\omega _N^s ) = v_s (N)\), and so, by (111), we have

$$\begin{aligned} \liminf _{s \rightarrow \infty } \left[ v_s (N) +{\textstyle \frac{1}{s}}\right] ^{-1/s} \ge \varrho ( \omega _N^\infty ) \equiv \rho (N). \end{aligned}$$
(115)

On the other hand, using that \(| {\varvec{{ q}} }_i^s - {\varvec{{ q}} }_j^s | = \varrho ( \omega _N^s )\) for at least one pair \((i,j)\), and furthermore that \(\varrho ( \omega _N^s ) \le \rho (N)\), we obtain

$$\begin{aligned} \langle V_s \rangle (\omega _N^s ) + {\frac{1}{s}}&= {\frac{1}{s} \frac{2}{N(N-1)}} \mathop {\sum \sum }_{1 \le i < j \le N} \left| {\varvec{{ q}} }_i^s -{\varvec{{ q}} }_j^s\right| ^{-s} \\&= \left[ \rho (N) \right] ^{-s} \frac{1}{s} \frac{2}{N(N-1)} \mathop {\sum \sum }_{1 \le i < j \le N} \left[ \frac{\rho (N)}{\left| {\varvec{{ q}} }_i^s - {\varvec{{ q}} }_j^s \right| }\right] ^{s} \\&\ge \left[ \rho (N) \right] ^{-s} \frac{1}{s} \frac{2}{N(N-1)} > 0. \end{aligned}$$

Hence,

$$\begin{aligned} \left[ \langle V_s \rangle ( \omega _N^s ) + {\frac{1}{s}}\right] ^{-1/s}&\le \rho (N) s^{1/s} \left( N(N-1)/2 \right) ^{1/s} \\&\rightarrow \rho (N)\qquad \text{ as }\quad s\rightarrow \infty ; \end{aligned}$$

but again, \(\langle V_s \rangle (\omega _N^s ) = v_s (N)\), and so we get

$$\begin{aligned} \limsup _{s \rightarrow \infty } \left[ v_s (N) +{\textstyle \frac{1}{s}}\right] ^{-1/s} \le \rho (N). \end{aligned}$$
(116)

By (115) and (116)

$$\begin{aligned} \lim _{s \rightarrow \infty } \left[ v_s (N) +{\textstyle \frac{1}{s}}\right] ^{-1/s} = \rho (N). \end{aligned}$$

This completes the proof of (112).

Appendix 3

We now prove the strict monotonic increase of \(s\mapsto v_s (N)\).

1.1 Proof of Relation (14)

We begin with the observation that the map \(s\mapsto V_s (r)\) is monotone increasing for all \(r\in (0,2]\), in fact strictly so except when \(r = 1\). To see this, take the first partial derivative of \(V_s (r)\) w.r.t. \(s\) to get

$$\begin{aligned} \partial _s V_s (r) = -s^{-2}\left( r^{-s}-1\right) -s^{-1} r^{-s} \ln r. \end{aligned}$$
(117)

Clearly, r.h.s.(117) \(=0\) if \(r=1\). We now show that r.h.s.(117) \(>0\) if \(r\ne 1\).

To this end, now take the first partial derivative of \(\partial _s V_s (r)\) w.r.t. \(r\) to get

$$\begin{aligned} \partial _{r} \partial _{s} V_s (r) = r^{-s-1} \ln r. \end{aligned}$$
(118)

Clearly, r.h.s.(118) \(=0\) iff \(r=1\); thus, \(r\mapsto \partial _s V_s (r)\) has a critical point at \(r=1\), and only this one. Finally, take the second partial derivative of \(\partial _s V_s (r)\) w.r.t. \(r\) to get

$$\begin{aligned} \partial _{r}^{2} \partial _{s} V_s (r) = r^{-s-2}\left( 1- (1+s) \ln r \right) . \end{aligned}$$
(119)

Evaluating r.h.s.(119) at \(r=1\) yields \(\partial _{r}^{2} \partial _{s} V_s (1)=1\); thus, for each \(s\) the map \(r\mapsto \partial _s V_s (r)\) has a non-degenerate minimum at \(r=1\) with value \(0\), and no other critical point. Therefore, \(\partial _s V_s (r)\ge 0\), with “\(=0\)” iff \(r=1\).

With the help of this calculus result we now conclude that whenever \(s>t\), then

$$\begin{aligned} v_s (N) = \langle V_s\rangle (\omega _N^s) > \langle V_t\rangle (\omega _N^s) \ge \langle V_t\rangle (\omega _N^t) = v_t (N); \end{aligned}$$
(120)

here, the strict inequality holds because there is no optimizing \(N\)-point configuration on \(\mathbb {S}^2\) with all pairs having distance 1.

This completes the proof.

Appendix 4: Spherical Digital Nets

Maximal sum-of-distance points (i.e. optimal configurations \(\omega _N^s\) for \(s = -1\)) provide optimal integration nodes for equal-weight numerical integration rules on the sphere; see [49] and [50, 96]. In general, such configurations are obtained by solving a highly non-linear optimization problem which makes them impractical for large number of points.

Digital nets and sequences introduced in [97] are efficiently computable so-called low-discrepancy point systems in the unit square that define effective Quasi-Monte Carlo rules for integrating functions on the unit square

$$\begin{aligned} \frac{1}{N} \sum _{j=1}^N f( {\varvec{{ x}} }_j ) \approx \int \int _{[0,1]^2} f( {\varvec{{ x}} }) \, d {\varvec{{ x}} }. \end{aligned}$$

Informally speaking, the points of a digital net in \([0,1]^2\) are distributed in such a way that a large number of elementary rectangles contain precisely the fraction of all points that corresponds to their area; cf. [98] and Fig. 15.

Fig. 15
figure 15

\(2048\) Sobol’ points in the unit square generated using a method by [99]

This distribution property remains unchanged when lifting a digital net and elementary rectangles (now called spherical rectangles) to \(\mathbb {S}^2\) using the area-preserving Lambert transformation of the map makers. These spherical digital nets are studied in [100] and [101]. Of particular interest is that numerically they are comparable with maximal sum-of-distance points. It is conjecturedFootnote 49 that \(\langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}} )\) will approach the same limit as \(v_{-1}(N)\) with the same rate of convergence as \(N\rightarrow \infty \).

We tested for local concavity a sequence of \(N\)-point spherical digital nets formed by the first \(N\) points of a Sobol’ sequence lifted to the sphere. For the implementation of the Sobol’ points we used [99]; cf. Fig. 15. The graph of \(N \mapsto \langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}} )\) (Fig. 16) shows an overall concave shape.

Fig. 16
figure 16

\(\langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}} )\) based on the \(2048\) Sobol’ points given in Fig. 15

Yet, some irregularities are clearly discernible in Fig. 16; in fact, the discrete second derivative of \(N\mapsto \langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}})\) reveals that \(N\mapsto \langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}})\) for this sequence of spherical digital nets is locally highly non-concave; see Fig. 17.

Fig. 17
figure 17

Discrete second derivative of \(\langle V_{-1} \rangle ( \omega _N^{\mathrm {sphDN}} )\) based on the \(2048\) Sobol’ points given in Fig. 15. The discrete points are joint by lines to guide the eye

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nerattini, R., Brauchart, J.S. & Kiessling, M.KH. Optimal \(N\)-Point Configurations on the Sphere: “Magic” Numbers and Smale’s 7th Problem. J Stat Phys 157, 1138–1206 (2014). https://doi.org/10.1007/s10955-014-1107-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-1107-7

Keywords

Navigation