Skip to main content
Log in

Spectrally Optimized Pointset Configurations

  • Published:
Constructive Approximation Aims and scope

Abstract

The search for optimal configurations of pointsets, the most notable examples being the problems of Kepler and Thompson, have an extremely rich history with diverse applications in physics, chemistry, communication theory, and scientific computing. In this paper, we introduce and study a new optimality criteria for pointset configurations. Namely, we consider a certain weighted graph associated with a pointset configuration and seek configurations that minimize certain spectral properties of the adjacency matrix or graph Laplacian defined on this graph, subject to geometric constraints on the pointset configuration. This problem can be motivated by solar cell design and swarming models, and we consider several spectral functions with interesting interpretations such as spectral radius, algebraic connectivity, effective resistance, and condition number. We prove that the regular simplex extremizes several spectral invariants on the sphere. We also consider pointset configurations on flat tori via (i) the analogous problem on lattices and (ii) through a variety of computational experiments. For many of the objectives considered (but not all), the triangular lattice is extremal.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. In some contexts, \(L^{{\mathbf {x}}}\) is referred to as the weighted unnormalized graph Laplacian [37], but we do not consider any of the other graph Laplacians here.

  2. The Voronoi cell is also sometimes referred to as the Dirichlet cell or Wigner–Seitz cell.

  3. A matrix \(A\in {\mathbb {Z}}^{n\times n}\) is unimodular if \(\mathrm {det} A =\pm 1\).

References

  1. Arenas, A., Díaz-Guilera, A., Kurths, J., Moreno, Y., Zhou, C.: Synchronization in complex networks. Phys. Rep. 469(3), 93–153 (2008). doi:10.1016/j.physrep.2008.09.002

    Article  MathSciNet  Google Scholar 

  2. Baernstein II, A.: A minimum problem for heat kernels of flat tori. Contemp. Math. 201, 227–243 (1997). doi:10.1090/conm/201/02604

    Article  MathSciNet  MATH  Google Scholar 

  3. Ballinger, B., Blekherman, G., Cohn, H., Giansiracusa, N., Kelly, E., Schürmann, A.: Experimental study of energy-minimizing point configurations on spheres. Exp. Math. 18(3), 257–283 (2009). doi:10.1080/10586458.2009.10129052

    Article  MathSciNet  MATH  Google Scholar 

  4. Belkin, M., Niyogi, P.: Laplacian eigenmaps for dimensionality reduction and data representation. Neural Comput. 15(6), 1373–1396 (2003). doi:10.1162/089976603321780317

    Article  MATH  Google Scholar 

  5. Belkin, M., Niyogi, P., Sindhwani, V.: Manifold regularization: a geometric framework for learning from labeled and unlabeled examples. J. Mach. Learn. Res. 7, 2399–2434 (2006)

    MathSciNet  MATH  Google Scholar 

  6. Berger, M.: Sur les premiéres valeurs propres des variétés Riemanniennes. Compos. Math. 26(2), 129–149 (1973)

    MATH  Google Scholar 

  7. Bétermin, L.: 2d theta functions and crystallization among Bravais lattices (2015). ArXiv: 1502.03839

  8. Bétermin, L., Zhang, P.: Minimization of energy per particle among Bravais lattices in \({R}^2\): Lennard–Jones and thomas-fermi cases. Commun. Contemp. Math. 17(6), 1450,049 (2015). doi:10.1142/S0219199714500497

    Article  MathSciNet  MATH  Google Scholar 

  9. Biyikoglu, T., Leydold, J., Stadler, P.F.: Laplacian Eigenvectors of Graphs. Springer, Berlin (2007). doi:10.1007/978-3-540-73510-6

    Book  MATH  Google Scholar 

  10. Björner, A., Lovász, L., Shor, P.W.: Chip-firing games on graphs. Eur. J. Comb. 12(4), 283–291 (1991). doi:10.1016/s0195-6698(13)80111-4

    Article  MathSciNet  MATH  Google Scholar 

  11. Borodachov, S.V., Hardin, D.P., Saff, E.B.: Low complexity methods for discretizing manifolds via Riesz energy minimization. Found. Comput. Math. 14(6), 1173–1208 (2014). doi:10.1007/s10208-014-9202-3

    Article  MathSciNet  MATH  Google Scholar 

  12. Borwein, J.M., Lewis, A.S.: Convex Analysis and Nonlinear Optimization. Springer, Berlin (2006). doi:10.1007/978-0-387-31256-9

    Book  MATH  Google Scholar 

  13. Boumal, N., Singer, A., Absil, P.A., Blondel, V.D.: Cramér–Rao bounds for synchronization of rotations. Inf. Inference 3(1), 1–39 (2013). doi:10.1093/imaiai/iat006

    MATH  Google Scholar 

  14. Boyd, S., Diaconis, P., Xiao, L.: Fastest mixing Markov chain on a graph. SIAM Rev. 46(4), 667–689 (2004). doi:10.1137/s0036144503423264

    Article  MathSciNet  MATH  Google Scholar 

  15. Cassels, J.W.S.: On a problem of Rankin about the Epstein zeta function. Proc. Glasg. Math. Assoc. 4, 73–80 (1959). doi:10.1017/s2040618500033906

    Article  MathSciNet  MATH  Google Scholar 

  16. Chung, F.R.K.: Spectral Graph Theory. AMS, Providence (1997). doi:10.1090/cbms/092

    MATH  Google Scholar 

  17. Cohen-Tannoudji, C., Dupont-Roc, J., Grynberg, G.: Atom–Photon Interactions. Wiley-Interscience, New York (1992). doi:10.1002/9783527617197

    Google Scholar 

  18. Cohn, H.: Order and disorder in energy minimization. In: Proceedings of the International Congress of Mathematicians, Hyderabad, India (2010). doi:10.1142/9789814324359_0152

  19. Cohn, H., Kumar, A.: Universally optimal distribution of points on spheres. J. Am. Math. Soc. 20(1), 99–148 (2007). doi:10.1090/S0894-0347-06-00546-7

    Article  MathSciNet  MATH  Google Scholar 

  20. Conway, J., Sloane, N.J.A.: Sphere Packings, Lattices and Groups, 3rd edn. Springer, Berlin (1999). doi:10.1007/978-1-4757-6568-7

    Book  MATH  Google Scholar 

  21. Cucker, F., Smale, S.: Emergent behavior in flocks. IEEE Trans. Autom. Control 52(5), 852–862 (2007). doi:10.1109/tac.2007.895842

    Article  MathSciNet  Google Scholar 

  22. Cucker, F., Smale, S.: On the mathematics of emergence. Jpn. J. Math. 2(1), 197–227 (2007). doi:10.1007/s11537-007-0647-x

    Article  MathSciNet  MATH  Google Scholar 

  23. Damelin, S.B., Grabner, P.J.: Energy functionals, numerical integration and asymptotic equidistribution on the sphere. J. Complex. 19(3), 231–246 (2003). doi:10.1016/s0885-064x(02)00006-7

    Article  MathSciNet  MATH  Google Scholar 

  24. Diananda, P.H.: Notes on two lemmas concerning the Epstein zeta function. Proc. Glasg. Math. Assoc. 6, 202–204 (1964). doi:10.1017/s2040618500035036

    Article  MathSciNet  MATH  Google Scholar 

  25. Economou, E.N.: Green’s Functions in Quantum Physics, vol. 7. Springer, Berlin (1983). doi:10.1007/978-3-662-02369-3

    Book  Google Scholar 

  26. Ennola, V.: A lemma about the Epstein zeta function. Proc. Glasg. Math. Assoc. 6, 198–201 (1964). doi:10.1017/s2040618500035024

    Article  MathSciNet  MATH  Google Scholar 

  27. Fiedler, M.: Algebraic connectivity of graphs. Czech. Math. J. 23, 298–305 (1973)

    MathSciNet  MATH  Google Scholar 

  28. Ganapati, V., Miller, O.D., Yablonovitch, E.: Light trapping textures designed by electromagnetic optimization for subwavelength thick solar cells. IEEE J. Photovol. 4(1), 175–182 (2014). doi:10.1109/jphotov.2013.2280340

    Article  Google Scholar 

  29. Ghosh, A., Boyd, S.: Growing well-connected graphs. In: Proceedings of the IEEE Conference Decision and Control, pp. 6605–6611 (2006). doi:10.1109/cdc.2006.377282

  30. Ghosh, A., Boyd, S.: Upper bounds on algebraic connectivity via convex optimization. Linear Algebra Appl. 418, 693–707 (2006). doi:10.1016/j.laa.2006.03.006

    Article  MathSciNet  MATH  Google Scholar 

  31. Ghosh, A., Boyd, S., Saberi, A.: Minimizing effective resistance of a graph. SIAM Rev. 50(1), 37–66 (2008). doi:10.1137/050645452

    Article  MathSciNet  MATH  Google Scholar 

  32. Godsil, C.D., Mohar, B.: Walk generating functions and spectral measures of infinite graphs. Linear Algebra Appl. 107, 191–206 (1988). doi:10.1016/0024-3795(88)90245-5

    Article  MathSciNet  MATH  Google Scholar 

  33. Kao, C.Y., Lai, R., Osting, B.: Maximization of Laplace–Beltrami eigenvalues on closed Riemannian surfaces. ESAIM Control Optim. Calc. Var. (2016). doi:10.1051/cocv/2016008

  34. Kirr, E., Weinstein, M.I.: Parametrically excited Hamiltonian partial differential equations. SIAM J. Appl. Math. 33(1), 16–52 (2001). doi:10.1137/s0036141099363456

    Article  MATH  Google Scholar 

  35. Kirr, E., Weinstein, M.I.: Metastable states in parametrically excited multimode Hamiltonian systems. Commun. Math. Phys. 236(2), 335–372 (2003). doi:10.1007/s00220-003-0820-x

    Article  MathSciNet  MATH  Google Scholar 

  36. Lin, L., Saad, Y., Yang, C.: Approximating spectral densities of large matrices. SIAM Rev. 58(1), 34–65 (2016). doi:10.1137/130934283

    Article  MathSciNet  MATH  Google Scholar 

  37. Luxburg, U.V.: A tutorial on spectral clustering. Stat. Comput. 17(4), 395–416 (2007). doi:10.1007/s11222-007-9033-z

    Article  MathSciNet  Google Scholar 

  38. Marcotte, E., Stillinger, F.H., Torquato, S.: Unusual ground states via monotonic convex pair potentials. J. Chem. Phys. 134(16), 164,105 (2011). doi:10.1063/1.3576141

    Article  Google Scholar 

  39. Matoušek, J.: Geometric Discrepancy: An Illustrated Guide. Springer, Berlin (1999). doi:10.1007/978-3-642-03942-3

    Book  MATH  Google Scholar 

  40. Miller, O.D.: Photonic design: from fundamental solar cell physics to computational inverse design. arXiv: 1308.0212 (2013)

  41. Miller, O.D., Yablonovitch, E.: Photon extraction: the key physics for approaching solar cell efficiency limits. In: Proceedings of the SPIE 8808, Active Photonic Materials, p. 880807 (2013). doi:10.1117/12.2024592

  42. Mohar, B.: The Laplacian spectrum of graphs. In: Alavi, Y., Chartrand, G., Oellermann, O.R., Schwenk, A.J. (eds.) Graph Theory, Combinatorics, and Applications, vol. 2, pp. 871–898. Wiley (1991)

  43. Mohar, B., Woess, W.: A survey on spectra of infinite graphs. Bull. Lond. Math. Soc. 21(3), 209–234 (1989). doi:10.1112/blms/21.3.209

    Article  MathSciNet  MATH  Google Scholar 

  44. Montgomery, H.L.: Minimal theta functions. Glasg. Math. J. 30(1), 75–85 (1988). doi:10.1017/s0017089500007047

    Article  MathSciNet  MATH  Google Scholar 

  45. Motsch, S., Tadmor, E.: A new model for self-organized dynamics and its flocking behavior. J. Stat. Phys. 144(5), 923–947 (2011). doi:10.1007/s10955-011-0285-9

    Article  MathSciNet  MATH  Google Scholar 

  46. Olfati-Saber, R., Fax, A., Murray, R.M.: Consensus and cooperation in networked multi-agent systems. Proc. IEEE 95(1), 215–233 (2007). doi:10.1109/jproc.2006.887293

    Article  Google Scholar 

  47. Osting, B., Brune, C., Osher, S.: Enhanced statistical rankings via targeted data collection. JMLR W&CP 28(1), 489–497 (2013)

    Google Scholar 

  48. Osting, B., Brune, C., Osher, S.: Optimal data collection for informative rankings expose well-connected graphs. J. Mach. Learn. Res. 15, 2981–3012 (2014)

    MathSciNet  MATH  Google Scholar 

  49. Osting, B., Marzuola, J.L., Cherkaev, E.: An isoperimetric inequality for integral operators on flat tori. Ill. J. Math. 59(3), 773–793 (2015). http://projecteuclid.org/euclid.ijm/1475266407

  50. Osting, B., Weinstein, M.I.: Emergence of periodic structure from maximizing the lifetime of a bound state coupled to radiation. SIAM J. Multiscale Model. Simul. 9(2), 654–685 (2011). doi:10.1137/100813221

    Article  MathSciNet  MATH  Google Scholar 

  51. Purcell, E.M.: Spontaneous emission probabilities at radio frequencies. Phys. Rev. 69, 681 (1946)

    Article  Google Scholar 

  52. Purcell, E.M.: Research in nuclear magnetism. Science 118(3068), 431–436 (1953)

  53. Rankin, R.A.: A minimum problem for the Epstein zeta-function. In: Proceedings of the Glasgow Mathematical Association, vol. 1, pp. 149–158 (1953). doi:10.1017/s2040618500035668

  54. Saff, E.B., Kuijlaars, A.B.J.: Distributing many points on a sphere. Math. Intell. 19(1), 5–11 (1997). doi:10.1007/bf03024331

    Article  MathSciNet  MATH  Google Scholar 

  55. Shannon, C.E.: A mathematical theory of communication. ACM SIGMOBILE Mob. Comput. Commun. Rev. 5(1), 3–55 (2001)

    Article  MathSciNet  Google Scholar 

  56. Shi, J., Malik, J.: Normalized cuts and image segmentation. IEEE Trans. Pattern Anal. Mach. Intell. 22(8), 888–905 (2000). doi:10.1109/34.868688

    Article  Google Scholar 

  57. Singer, A.: From graph to manifold Laplacian: the convergence rate. Appl. Comput. Harmonic Anal. 21(1), 128–134 (2006). doi:10.1016/j.acha.2006.03.004

    Article  MathSciNet  MATH  Google Scholar 

  58. Smale, S.: Mathematical problems for the next century. Math. Intell. 20(2), 7–15 (1998). doi:10.1007/bf03025291

    Article  MathSciNet  MATH  Google Scholar 

  59. Sun, J., Boyd, S., Xiao, L., Diaconis, P.: The fastest mixing Markov process on a graph and a connection to a maximum variance unfolding problem. SIAM Rev. 48(4), 681–699 (2006). doi:10.1137/s0036144504443821

    Article  MathSciNet  MATH  Google Scholar 

  60. 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. Ser. 6 7(39), 237–265 (1904). doi:10.1080/14786440409463107

    Article  MATH  Google Scholar 

  61. Torquato, S., Stillinger, F.H.: Jammed hard-particle packings: from Kepler to Bernal and beyond. Rev. Mod. Phys. 82(3), 2633–2672 (2010). doi:10.1103/revmodphys.82.2633

    Article  Google Scholar 

  62. Tóth, L.F.: Regular Figures. Pergamon Press, Oxford (1964)

    MATH  Google Scholar 

  63. Van Hove, L.: The occurrence of singularities in the elastic frequency distribution of a crystal. Phys. Rev. 89(6), 1189–1193 (1953). doi:10.1103/physrev.89.1189

    Article  MathSciNet  MATH  Google Scholar 

  64. Wendland, H.: Scattered Data Approximation. Cambridge University Press, Cambridge (2004). doi:10.1017/cbo9780511617539

    Book  MATH  Google Scholar 

  65. Whyte, L.L.: Unique arrangements of points on a sphere. Am. Math. Mon. 59(9), 606–611 (1952). doi:10.2307/2306764

    Article  MathSciNet  MATH  Google Scholar 

  66. Yablonovitch, E.: Inhibited spontaneous emission in solid-state physics and electronics. Phys. Rev. Lett. 58(20), 2059–2062 (1987). doi:10.1103/physrevlett.58.2059

    Article  Google Scholar 

  67. Yu, Z., Raman, A., Fan, S.: Fundamental limit of nanophotonic light trapping in solar cells. Proc. Natl. Acad. Sci. 107(41), 17491–17496 (2010). doi:10.1073/pnas.1008296107

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Mikael Rechtsman for starting them down the path of this work, as well as Laurent Betermin, Elena Cherkaev, Owen Miller, and Peter Mucha for valuable discussions along the way to completing it.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Braxton Osting.

Additional information

Communicated by Edward B. Saff.

BO gratefully acknowledges support from NSF DMS-1461138 and NSF DMS-1619755. JLM is supported by NSF Applied Math Grant DMS-1312874 and NSF CAREER Grant DMS-1352353.

Appendices

Appendix A: Sensitivity Analysis of Spectral Quantities

In Sects. 2.1 and 2.2, we introduced the weighted adjacency matrix and graph Laplacian associated to a pointset configuration, discussed some spectral properties of the matrices, and recalled a variety of spectral quantities that are of interest in various applications. In this appendix, we discuss the sensitivity of these spectral quantities with respect to changes in the pointset configuration.

We say that a function \(J :{\mathbb {R}}^n \rightarrow {\mathbb {R}}\) is a symmetric if \(J(x) = J([x])\), where \([\cdot ]:{\mathbb {R}}^n \rightarrow {\mathbb {R}}^n\) rearranges the components of a vector in nondecreasing order. In other words, the value of J is invariant to permuting the components of the argument. Recall that we have defined \(\mathbf{S}^n\) to be the set of real symmetric \(n \times n\) matrices, and let \(\lambda :\mathbf{S}^n \rightarrow {\mathbb {R}}^n\) be the vector containing the ordered eigenvalues of a symmetric matrix. We say that the composition \(J\circ \lambda \) is a spectral function if J is a symmetric function. Roughly speaking, \(J\circ \lambda \) is convex and differentiable when J is convex and differentiable [12].

Recall that the weighted adjacency matrix associated with a Euclidean pointset \(\{{{\mathbf {x}}}_i \}_{i\in [n]}\) is defined as in (2). For the spectral function \(J\circ \lambda \), we consider the composition

$$\begin{aligned} J \circ \lambda \circ W^{{\mathbf {x}}}. \end{aligned}$$

The following proposition shows how the composite function, \( J \circ \lambda \circ W^{{\mathbf {x}}}\), changes as we move a single point in the configuration, say \({{\mathbf {x}}}_i \in {\mathbb {R}}^d\).

Proposition A.1

Let J be differentiable at \(\lambda (W^{{\mathbf {x}}})\). The gradient of the objective function, \( J \circ \lambda \circ W^{{\mathbf {x}}}:{\mathbb {R}}^{N\times 2} \rightarrow {\mathbb {R}}\), with respect to \({{\mathbf {x}}}_i\) is given by

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J \circ \lambda \circ W^{{\mathbf {x}}}) = 4 \sum _k \left( U \ \mathrm {diag} \nabla J (\lambda ) \ U^t \right) _{ik} f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_k) \end{aligned}$$

for any diagonalizing matrix \(U\in O^n\) satisfying \(W^{{\mathbf {x}}}= U \ \mathrm {diag}\ \lambda (W^{{\mathbf {x}}}) \ U^t\).

Proof

The proof is an exercise in the chain rule. When J is differentiable at \(\lambda (A)\), the composition \(J\circ \lambda :\mathbf{S}^N \rightarrow {\mathbb {R}}\) is Fréchet differentiable with derivative

$$\begin{aligned} \nabla (J\circ \lambda ) (A) = U \ \mathrm {diag} \nabla J (\lambda ) \ U^t \end{aligned}$$

for any diagonalizing matrix \(U\in O^n\) satisfying \(A = U \ \mathrm {diag}\ \lambda (A) \ U^t\). See, for example, [12, Corollary 5.2.5]. For simplicity, we write \(V = U \ \mathrm {diag} \nabla J (\lambda ) \ U^t\). Thus, we compute

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i}( J \circ \lambda \circ W^{{\mathbf {x}}})&= \langle V , \nabla _{{{\mathbf {x}}}_i} W^{{\mathbf {x}}}\rangle _F \end{aligned}$$
(30a)
$$\begin{aligned}&= \sum _{jk} V_{jk} \nabla _{{{\mathbf {x}}}_i} W_{jk}^{{\mathbf {x}}}\end{aligned}$$
(30b)
$$\begin{aligned}&= \sum _{jk} V_{jk} \left( \delta _{ij} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}+ \delta _{ik} \nabla _{{{\mathbf {x}}}_i} W_{ji}^{{\mathbf {x}}}\right) \end{aligned}$$
(30c)
$$\begin{aligned}&= \sum _{k} V_{ik} \ \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}+ \sum _{j} V_{ji} \ \nabla _{{{\mathbf {x}}}_i} W_{ji}^{{\mathbf {x}}}\end{aligned}$$
(30d)
$$\begin{aligned}&= 2 \sum _{k} V_{ik} \ \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}. \end{aligned}$$
(30e)

In (30a), \(\langle \cdot , \cdot \rangle _F \) denotes the Frobenius inner product. In (30c), \(\delta _{ij}\) denotes the Kronecker delta function. In (30e), we used the symmetry of V and W. We compute

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}= f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) ) \ \nabla _{{{\mathbf {x}}}_i} d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) = 2 f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_k), \end{aligned}$$
(31)

from which the result follows. \(\square \)

Example

Consider the spectral function \(J(\lambda ) = \sum _i \lambda _i^2\). In this case, \(U \ \mathrm {diag} \nabla J (\lambda ) \ U^t = 2 W^{{\mathbf {x}}}\). Thus from Proposition A.1 we can check that

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J \circ \lambda \circ W^{{\mathbf {x}}}) = 2 \langle W^{{\mathbf {x}}}, \nabla _{{{\mathbf {x}}}_i} W^{{\mathbf {x}}}\rangle _F = \nabla _{{{\mathbf {x}}}_i} \Vert W^{{\mathbf {x}}}\Vert _F^2 . \end{aligned}$$

This is a trivial identity since \( \Vert W^{{\mathbf {x}}}\Vert _F^2 = \mathrm {tr} [ (W^{{\mathbf {x}}})^2 ] = \sum _i \lambda _i(W^{{\mathbf {x}}})^2\).

As in (3), the graph Laplacian associated with a pointset \(\{{{\mathbf {x}}}_i \}_{i\in [n]}\) is given by \(L^{{\mathbf {x}}}= D^{{\mathbf {x}}}- W^{{\mathbf {x}}}\). For the spectral function \(J\circ \lambda \), we consider the composition

$$\begin{aligned} J \circ \lambda \circ L^{{\mathbf {x}}}. \end{aligned}$$

The following proposition shows how this composite function, \( J \circ \lambda \circ L^{{\mathbf {x}}}\), changes as we move a single point in the configuration, say \({{\mathbf {x}}}_i \in {\mathbb {R}}^n\).

Proposition A.2

Let J be differentiable at \(\lambda (L^{{\mathbf {x}}})\). The gradient of the objective function, \( J \circ \lambda \circ L^{{\mathbf {x}}}:{\mathbb {R}}^{N\times 2} \rightarrow {\mathbb {R}}\), with respect to \({{\mathbf {x}}}_i\) is given by

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J \circ \lambda \circ L^{{\mathbf {x}}}) = 2 \sum _k(V_{kk} - 2 V_{ik} + V_{ii}) \ f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_k), \end{aligned}$$
(32)

where \(V = \left( U \ \mathrm {diag} \nabla J (\lambda ) \ U^t \right) _{ik}\) for any diagonalizing matrix \(U\in O^n\) satisfying \(L^{{\mathbf {x}}}= U \ \mathrm {diag}\ \lambda (L^{{\mathbf {x}}}) \ U^t\). The gradient can alternatively be expressed using the arc-vertex incidence matrix decomposition (7),

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J \circ \lambda \circ L^{{\mathbf {x}}}) = 2 \sum _{k = (i,j)} g_k \ f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_j) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_j), \end{aligned}$$
(33)

where \(g = \sum _{j=1}^n (Bu_j)^2 \big ( \nabla J(\lambda ) \big )_j\).

Proof

As in the proof of Proposition A.1, when J be differentiable at \(\lambda (L^{{\mathbf {x}}})\), the composition \(J\circ \lambda :\mathbf{S}^n \rightarrow {\mathbb {R}}\) is Fréchet differentiable with derivative

$$\begin{aligned} (J\circ \lambda )' (L^{{\mathbf {x}}}) = U \ \mathrm {diag} \nabla J (\lambda ) \ U^t \end{aligned}$$

for any diagonalizing matrix \(U\in O^n\) satisfying \(L^{{\mathbf {x}}}= U \ \mathrm {diag} \lambda (L^{{\mathbf {x}}}) \ U^t\). Writing \(V = U \ \mathrm {diag} \nabla J (\lambda ) \ U^t\), it follows that

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i}( J \circ \lambda \circ L^{{\mathbf {x}}})&= \langle V , \nabla _{{{\mathbf {x}}}_i} (D^{{\mathbf {x}}}- W^{{\mathbf {x}}}) \rangle _F \nonumber \\&= \sum _k V_{kk} \nabla _{{{\mathbf {x}}}_i} D_{kk}^{{\mathbf {x}}}- 2 \sum _k V_{ik} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}, \end{aligned}$$
(34)

where we used (30e). From

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} D_{kk}^{{\mathbf {x}}}= \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}+ \delta _{ik} \sum _j \nabla _{{{\mathbf {x}}}_i} W_{ij}^{{\mathbf {x}}}, \end{aligned}$$

we then have that

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i}( J \circ \lambda \circ L^{{\mathbf {x}}})&= \sum _k V_{kk} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}+ \sum _k V_{kk} \delta _{ik} \sum _j \nabla _{{{\mathbf {x}}}_i} W_{ij}^{{\mathbf {x}}}- 2 \sum _k V_{ik} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}\\&= \sum _k V_{kk} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}+ V_{ii} \sum _j \nabla _{{{\mathbf {x}}}_i} W_{ij}^{{\mathbf {x}}}- 2 \sum _k V_{ik} \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}\\&= \sum _k (V_{kk} - 2 V_{ik} + V_{ii}) \nabla _{{{\mathbf {x}}}_i} W_{ik}^{{\mathbf {x}}}. \end{aligned}$$

Equation (32) now follows from (31).

From (34), we could also write

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J\circ \lambda \circ L)&= \langle V, \nabla _{{{\mathbf {x}}}_i} (B^t \text {diag}(w^{{\mathbf {x}}}) B) \rangle _F \nonumber \\&= \langle V, B^t \text {diag}( \nabla _{{{\mathbf {x}}}_i} w) B \rangle _F \nonumber \\&= \langle (BU) \ \mathrm {diag} \nabla J (\lambda ) \ (BU)^t, \text {diag}( \nabla _{{{\mathbf {x}}}_i} w^{{\mathbf {x}}}) \rangle _F \nonumber \\&= \Big \langle g , \nabla _{{{\mathbf {x}}}_i} w^{{\mathbf {x}}}\Big \rangle , \end{aligned}$$
(35)

where

$$\begin{aligned} g = \text {diag} \left( (BU) \ \mathrm {diag} \nabla J (\lambda ) \ (BU)^t \right) . \end{aligned}$$

Thus, \(g\in {\mathbb {R}}^{\left( {\begin{array}{c}n\\ 2\end{array}}\right) }\) has entries given by

$$\begin{aligned} g_k = \sum _{j=1}^n (BU)_{k,j}^2 \big ( \nabla J(\lambda ) \big )_j. \end{aligned}$$

Finally, we compute

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} w_k = {\left\{ \begin{array}{ll} 2 f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_j) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_j), &{} k = (i,j), \\ 0, &{} \text {otherwise}, \end{array}\right. } \end{aligned}$$

which gives (33). \(\square \)

Example

Consider the spectral function \(J(\lambda ) = \sum _i \lambda _i\) so that the objective function is simply \(\text {tr} L^{{\mathbf {x}}}\). In this case, \(V= \text {Id}\), and

$$\begin{aligned} \nabla _{{{\mathbf {x}}}_i} (J \circ \lambda \circ L^{{\mathbf {x}}}) = 4 \sum _k f'( d^2({{\mathbf {x}}}_i, {{\mathbf {x}}}_k) ) \ ({{\mathbf {x}}}_ i - {{\mathbf {x}}}_k) = \nabla _{{{\mathbf {x}}}_i} \sum _{j,k} f( d^2({{\mathbf {x}}}_j, {{\mathbf {x}}}_k) ). \end{aligned}$$

Remark A.3

Equation (35) in the proof of Proposition A.2 implies that the variation of a spectral function of the graph Laplacian with respect to the edge weights is generally given by

$$\begin{aligned} \nabla _w [J \circ \lambda \circ (B^t \ \text {diag}(w) \ B) ] = \text {diag}[ (BU) \ \text {diag}\big (\nabla J(\lambda ) \big ) \ (BU)^t]. \end{aligned}$$
(36)

For example, the gradient of the trace, \( \text {tr} L(w)\), is given by

$$\begin{aligned} \frac{\delta \text {tr} L}{\delta w} = \text {diag} (B B^t) = 2 \cdot 1_{\left( {\begin{array}{c}n\\ 2\end{array}}\right) }. \end{aligned}$$

The factor of two simply indicates that each edge weight effects the degree of two vertices. When \(\lambda _2 \big ( L(w)\big )\) is a simple eigenvalue, the gradient is given by

$$\begin{aligned} \frac{\delta \lambda _2}{\delta w} = (B u_2)^2, \end{aligned}$$

which agrees with the formulas in [29, 30]. Finally, the gradient of the total effective resistance, \(R_{tot} = \sum _j \lambda _j^{-1}\), is computed

$$\begin{aligned} \frac{\delta R_{tot}}{\delta w} = - \text {diag} \big ( B U \text {diag}(\lambda ^{-2} ) (BU)^t \big ) = - \text {diag}( B (L^\dag )^2 B^t), \end{aligned}$$

which can also be found in [31].

For a general spectral function J, the gradients computed in Propositions A.1 and A.2 can be used together with a quasi-Newton optimization method to efficiently search for a locally optimal pointset configuration; see Sect. 5.

Appendix B: Parameterization of Lattices

Let \(B = [b_1,\ldots ,b_n]\in {\mathbb {R}}^{n\times n}\) have linearly independent columns. The lattice generated by the basis B is the set of integer linear combinations of the columns of B,

$$\begin{aligned} {\mathcal {L}}(B)=\{Bx:x \in {\mathbb {Z}}^n \}. \end{aligned}$$

Let B and C be two lattice bases. We recall that \({\mathcal {L}}(B)={\mathcal {L}}(C)\) if and only if there is a unimodularFootnote 3 matrix U such that \(B = CU\). Thus, there is a one-to-one correspondence between the unimodular \(2\times 2\) matrices and the bases of a two-dimensional lattice.

We say that two lattices are isometric if there is a rigid transformation that maps one to the other. The following proposition parameterizes the space of two-dimensional, unit-volume lattices modulo isometry.

Proposition B.1

Every two-dimensional lattice with volume one is isometric to a lattice parameterized by the basis

$$\begin{aligned} \begin{pmatrix} \frac{1}{\sqrt{b}}&{}\frac{a}{\sqrt{b}} \\ 0 &{} \sqrt{b} \end{pmatrix}, \end{aligned}$$

where the parameters a and b are constrained to the set

$$\begin{aligned} U := \left\{ (a,b) \in {\mathbb {R}}^2 :b>0, \ a \in [ 0,1/2 ],\ \text {and} \ a^2 + b^2 \ge 1 \right\} . \end{aligned}$$

The set U defined in Proposition B.1 is illustrated in Fig. 8.

Fig. 8
figure 8

The set U in Proposition B.1. Parameters (ab) corresponding to square, triangular, rectangular, rhombic, and oblique lattices are also indicated

Proof

Consider an arbitrary lattice with unit volume. We first choose the basis vectors so that the angle between them is acute. After a suitable rotation and reflection, we can let the shorter basis vector (with length \(\frac{1}{\sqrt{b}}\)) be parallel to the x axis and the longer basis vector (with length \( \sqrt{ \frac{a^2}{b} + b} = \sqrt{\frac{1}{b} \left( a^2 + b^2 \right) } \ge \sqrt{ \frac{1}{b}}\)) lie in the first quadrant. Multiplying on the right by a unimodular matrix, \( \begin{pmatrix} 1 &{} 1 \\ 0 &{} 1 \end{pmatrix}\), we compute

$$\begin{aligned} \begin{pmatrix} \frac{1}{\sqrt{b}} &{} \frac{a}{\sqrt{b}} \\ 0 &{} \sqrt{b} \end{pmatrix} \begin{pmatrix} 1 &{} 1 \\ 0 &{} 1 \end{pmatrix} = \begin{pmatrix} \frac{1}{\sqrt{b}} &{} \frac{a+1}{\sqrt{b}} \\ 0 &{} \sqrt{b} \end{pmatrix} . \end{aligned}$$

Since this is equivalent to taking \(a\mapsto a+1\), it follows that we can identify the lattices associated with the points (ab) and \((a+1,b)\). Thus, we can restrict the parameter a to the interval \(\left[ 0, 1/2 \right] \) by symmetry. For a complete picture of this restriction and how the symmetry naturally arises, see [33, Proposition 3.2 Figure 3]. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Osting, B., Marzuola, J. Spectrally Optimized Pointset Configurations. Constr Approx 46, 1–35 (2017). https://doi.org/10.1007/s00365-017-9365-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00365-017-9365-7

Keywords

Mathematics Subject Classification

Navigation