Skip to main content
Log in

Harmonic Properties of the Logarithmic Potential and the Computability of Elliptic Fekete Points

  • Published:
Constructive Approximation Aims and scope

Abstract

We investigate the properties of the function sending each N-tuple of points to minus the logarithm of the product of their mutual distances. We prove that, as a function defined on the product of N spheres, this function is subharmonic, and indeed its (Riemannian) Laplacian is constant. We also prove a mean value equality and an upper bound on the derivative of the function. We use these results to get sharp upper bounds for the precision needed to describe an approximation to elliptic Fekete points (in the sense demanded by Smale’s 7th problem). We also conclude that Smale’s 7th problem has solutions given by rational spherical points of bounded (small) bit length, proving that there exists an exponential running time algorithm which solves it on the Turing machine model.

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.

Similar content being viewed by others

Notes

  1. Smale refers to the unit sphere in ℝ3, but the two problems are equivalent by sending points \((a,b,c)\in \mathbb{S}\) to (2a,2b,2c−1).

  2. More precisely, the running time of our procedure is polynomial(N)⋅(20N)36N.

  3. As a linear operator in \(C^{2}(\mathcal{M})\), Δ is called the Laplace–Beltrami operator. Note that for some authors \(\Delta {f}=-\operatorname{div}(\operatorname{grad}(f))\) and hence has a minus sign, see for example [14, p. 88].

References

  1. Armentano, D., Beltrán, C., Shub, M.: Minimizing the discrete logarithmic energy on the sphere: the role of random polynomials. Trans. Am. Math. Soc. 363(6), 2955–2965 (2011)

    Article  MATH  Google Scholar 

  2. Bendito, E., Carmona, A., Encinas, A.M., Gesto, J.M., Gómez, A., Mouriño, C., Sánchez, M.T.: Computational cost of the Fekete problem. I. The forces method on the 2-sphere. J. Comput. Phys. 228(9), 3288–3306 (2009)

    Article  MathSciNet  Google Scholar 

  3. Bendito, E., Carmona, A., Encinas, A.M., Gesto, J.M.: Computational cost of the Fekete Problem II: on Smale’s 7th problem. To appear. Available at http://www-ma3.upc.es/users/bencar/papers.html

  4. Bergersen, B., Boal, D., Palffy-Muhoray, P.: Equilibrium configurations of particles on a sphere: the case of logarithmic interactions. J. Phys. A, Math. Gen. 27, 2579–2586 (1994)

    Article  MATH  Google Scholar 

  5. Besse, A.L.: Manifolds all of Whose Geodesics are Closed. Ergebnisse der Mathematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas], vol. 93. Springer, Berlin (1978). With appendices by D.B.A. Epstein, J.-P. Bourguignon, L. Bérard-Bergery, M. Berger and J.L. Kazdan

    Book  MATH  Google Scholar 

  6. Blum, L., Cucker, F., Shub, M., Smale, S.: Complexity and Real Computation. Springer, New York (1998)

    Book  Google Scholar 

  7. Blum, L., Shub, M., Smale, S.: On a theory of computation and complexity over the real numbers: NP-completeness, recursive functions and universal machines. Bull., New Ser., Am. Math. Soc. 21(1), 1–46 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  8. do Carmo, M.P.: Riemannian Geometry. Mathematics: Theory & Applications. Birkhäuser, Boston (1992). Translated from the second Portuguese edition by Francis Flaherty

    MATH  Google Scholar 

  9. Dragnev, P.D.: On the separation of logarithmic points on the sphere. In: Approximation Theory, X, St. Louis, MO, 2001. Innov. Appl. Math., pp. 137–144. Vanderbilt Univ. Press, Nashville (2002)

    Google Scholar 

  10. Dubickas, A.: On the maximal product of distances between points on a sphere. Liet. Mat. Rink. 36(3), 303–312 (1996)

    MathSciNet  Google Scholar 

  11. Evans, L.C.: Partial Differential Equations. Graduate Studies in Mathematics, vol. 19. American Mathematical Society, Providence (1998)

    MATH  Google Scholar 

  12. Forrester, P.J.: Log-Gases and Random Matrices. Princeton University Press, Princeton (2010)

    MATH  Google Scholar 

  13. Jost, J.: Postmodern Analysis, 3rd edn. Universitext. Springer, Berlin (2005)

    MATH  Google Scholar 

  14. Jost, J.: Riemannian Geometry and Geometric Analysis, 5th edn. Universitext. Springer, Berlin (2008)

    MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  17. Rakhmanov, E.A., Saff, E.B., Zhou, Y.M.: Electrons on the sphere. In: Computational Methods and Function Theory, Penang, 1994. Ser. Approx. Decompos., vol. 5, pp. 293–309. World Scientific, River Edge (1995)

    Google Scholar 

  18. Schmutz, E.: Rational points on the unit sphere. Cent. Eur. J. Math. 6(3), 482–487 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  19. Shub, M., Smale, S.: Complexity of Bezout’s theorem. III. Condition number and packing. J. Complex. 9(1), 4–14 (1993). Festschrift for Joseph F. Traub, Part I

    Article  MathSciNet  MATH  Google Scholar 

  20. Smale, S.: Mathematical problems for the next century. Mathematics: Frontiers and Perspectives, pp. 271–294. Am. Math. Soc., Providence (2000)

    Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  22. Willmore, T.J.: Mean value theorems in harmonic Riemannian spaces. J. Lond. Math. Soc. 25, 54–57 (1950)

    Article  MathSciNet  MATH  Google Scholar 

  23. Zhong, Q.: Energies of zeros of random sections on Riemann surfaces. Indiana Univ. Math. J. 57(4), 1753–1780 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  24. Zhou, Y.: Arrangements of points on the sphere. Ph.D. Thesis. Math. Department, University of South Florida (1995)

  25. Hartman, P.: Ordinary Differential Equations. Wiley, New York (1964)

    MATH  Google Scholar 

Download references

Acknowledgements

Thanks to Cecilia Pola for many conversations, and to Jean Pierre Dedieu, Luis Miguel Pardo, Mike Shub, and two anonymous referees for comments and suggestions. Thanks also to Joaquim Ortega Cerdà for his comments on harmonic manifolds.

Partially supported by MTM2010-16051 (Spanish Ministry of Science and Innovation MICINN).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Carlos Beltrán.

Additional information

Communicated by Edward B. Saff.

Appendix: Topics from Riemannian Geometry and Harmonic Analysis in Manifolds

Appendix: Topics from Riemannian Geometry and Harmonic Analysis in Manifolds

To facilitate the reading of this manuscript, we include two short sections with some results from Riemannian geometry and harmonic analysis that have been used in the paper. The contents of this appendix are mainly taken from [8, 14].

1.1 A.1 Riemannian Geometry

By a “Riemannian manifold” \(\mathcal{M}\), we mean here a smooth (C ) differentiable manifold with a smooth Riemannian structure, that is, a smooth section \(\langle\cdot,\cdot\rangle:T\mathcal{M}\times T\mathcal{M}\,{\rightarrow}\,\mathbb{R}\), where \(T\mathcal{M}\) is the tangent bundle of \(\mathcal{M}\) and for each \(p\in \mathcal{M}\), \(\langle \cdot,\cdot\rangle_{p}:T_{p}\mathcal{M}\times T_{p}\mathcal{M}\,{\rightarrow}\,\mathbb{R}\) is a definite positive, symmetric bilinear map. We denote by n the dimension of \(\mathcal{M}\). Associated with 〈⋅,⋅〉 p , we consider the norm \(\|v\|_{p}=\langle v,v\rangle_{p}^{1/2}\) for \(v\in T_{p}\mathcal{M}\). We denote by C 2(U) the set of C 2 functions defined in some open set \(U\subseteq \mathcal{M}\).

Given a collection \(\mathcal{M}^{(1)},\ldots,\mathcal{M}^{(r)}\) of Riemannian manifolds, we can define in \(\mathcal{M}=\mathcal{M}^{(1)}\times\cdots\times \mathcal{M}^{(r)}\) a product Riemannian structure as follows: Let \(p=(p^{(1)},\ldots,p^{(r)})\in \mathcal{M}\), and let \(v^{(i)},w^{(i)}\in T_{p^{(i)}}\mathcal{M}^{(i)}\), 1≤ir, be a collection of tangent vectors. Then define

$$\bigl\langle\bigl(v^{(1)},\ldots,v^{(r)}\bigr),\bigl(w^{(1)},\ldots,w^{(r)}\bigr)\bigr\rangle_p=\bigl\langle v^{(1)},w^{(1)}\bigr\rangle_{p^{(1)}}+\cdots+\bigl\langle v^{(r)},w^{(r)}\bigr\rangle_{p^{(r)}}.$$

The length of a piecewise C 1 curve \(\gamma:[a,b]\rightarrow \mathcal{M}\) with tangent vector \(\dot{\gamma}\) (i.e., \(\dot{\gamma}(t)\in T_{\gamma (t)}\mathcal{M},t\in[a,b]\)) is defined as

$$L(\gamma)=\int_a^b\|\dot{\gamma}\|\,dt.$$

The distance between two points \(p,q\in \mathcal{M}\) is then defined as the infimum of the lengths of piecewise C 1 curves with extremes p,q. This gives \(\mathcal{M}\) a structure of metric space and allows us to define open and closed balls as usual.

Given a coordinate chart \(\phi:U\subseteq \mathcal{M}\,{\rightarrow}\, V\subseteq \mathbb{R}^{n}\), the Riemannian product is represented by a positive definite, symmetric matrix

$$G(x)=\bigl(g_{ij}(x)\bigr),\quad x\in V,$$

such that given \(v,w\in T_{p} \mathcal{M}\) with pU, v=(v 1,…,v n ) and w=(w 1,…,w n ) in coordinates, we have

$$\langle v,w\rangle_x=w^TG(x)v,$$

where w T is the transpose of w. A function f:U → ℝ is called (Lebesgue-)measurable or integrable if f(ϕ −1(x))|det(DG(x))|1/2 is (Lebesgue-)measurable or integrable as a function of xV. In that case, the integral of f in U is defined as

$$\int_Uf(p)\,dp=\int_Vf\bigl(\phi^{-1}(x)\bigr)\bigl|\det\bigl(DG(x)\bigr)\bigr|^{1/2}\,dx.$$

Given \(f:\mathcal{M}\,{\rightarrow}\,\mathbb{R}\), we define

$$\int_\mathcal{M}f(p)\,dp=\sum_{\alpha}\int_{U_\alpha}f(p)\rho_\alpha(p)\,dp,$$

where {ρ α } is a partition of the unity subordinate to some open cover of \(\mathcal{M}\) by coordinate charts {(U α ,ϕ α )} (f is called measurable or integrable if it is so in each U α ). The volume of some measurable subset \(U\subseteq \mathcal{M}\) is

$$\operatorname {Vol}{(U)}=\int_\mathcal{M}\chi_U(p)\,dp,\quad \chi_U(p)=\begin{cases}1&p\in U,\\0&\mathrm{otherwise}.\end{cases}$$

If \(\operatorname {Vol}{(U)}<\infty\), we define the expected value of f:U → ℝ in U as

$$\mathchoice {{\setbox0=\hbox{$\displaystyle{\textstyle-}{\int}$}\vcenter{\hbox{$\textstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\textstyle{\scriptstyle-}{\int}$}\vcenter{\hbox{$\scriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptscriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}\!\int _U f(p)\,dp=\frac{1}{\operatorname {Vol}{(U)}}\int_Uf(p)\,dp.$$

We denote by g ij the components of the inverse matrix G(x)−1. The Chirstoffel symbols associated with ϕ are then

$$\varGamma _{jk}^i=\frac{1}{2}\sum _{l=1}^ng^{il} \biggl(\frac{\partial g_{jl}}{\partial x_k}+\frac{\partial g_{kl}}{\partial x_j}-\frac{\partial g_{jk}}{\partial x_l} \biggr).$$

A smooth curve \(\gamma:[a,b]\rightarrow \mathcal{M}\) is a geodesic if it is a critical point of the energy functional \(\int_{a}^{b}\|\dot{\gamma}\|^{2}\,dt\). In coordinates, denoting x(t)=(x 1(t),…,x n (t))=ϕ(γ(t)), for i∈{1,…n},

$$\ddot{x}_i(t)=-\dot{x}(t)^T\varGamma ^i\bigl(x(t)\bigr)\dot{x}(t),\quad \varGamma ^i=\bigl(\varGamma ^i_{jk}\bigr)_{j,k=1\ldots n}.$$

Given \(p\in \mathcal{M}\) and \(v\in T_{p}\mathcal{M}\), there exists ε>0 and a unique geodesic \(\gamma:[0,\varepsilon]\,{\rightarrow}\,\mathcal{M}\) such that γ(0)=p and \(\dot{\gamma}(0)=v\). From the geodesic equation above, we can easily see that if \(\mathcal{M}^{(1)}\times\cdots\times \mathcal{M}^{(r)}\) has the product structure, then a curve \(\gamma=(\gamma^{(1)},\ldots,\gamma ^{(r)}):[a,b]\,{\rightarrow}\,\mathcal{M}\) is a geodesic in \(\mathcal{M}\) if and only if \(\gamma ^{(i)}:[a,b]\,{\rightarrow}\,\mathcal{M}^{(i)}\) is a geodesic in \(\mathcal{M}^{(i)}\) for i∈{1,…,r}. If \(\mathcal{M}\subseteq \mathbb{R}^{k}\) is a smooth submanifold of ℝk with the Riemannian structure inherited from ℝk (that is, 〈v,w p =〈v,w〉 the usual inner product in ℝk), the geodesic equation reads

$$\bigl\|\dot{\gamma}(t)\bigr\|=\text{constant},\qquad\ddot{\gamma}(t)\perp T_{\gamma(t)}\mathcal{M}\subseteq \mathbb{R}^k,\quad\text{for all }t.$$

We are most interested in the cases \(\mathcal{M}=\mathbb{S}\) with the Riemannian structure inherited from ℝ3 and \(\mathcal{M}=\mathbb{S}^{N}=\mathbb{S}\times\cdots \times \mathbb{S}\) (N times) with the product Riemannian structure. Of course, this product structure is also the Riemannian structure of \(\mathbb{S}^{N}\) as a submanifold of ℝ3N. Note that \(\mathbb{S}\) has dimension 2 and \(\mathbb{S}^{N}\) has dimension 2N. Geodesics in \(\mathbb{S}\) are great circles parametrized with constant speed. More exactly, the geodesic γ(t) such that \(\gamma(0)=p\in \mathbb{S}\) and \(\dot{\gamma}(0)=v\in T_{p}\mathbb{S}\), where ∥v∥=1, is given by

$$ \gamma_{p,v}(t)=p\cos(2t)+\frac{v}{2}\sin(2t)+\frac{1}{2}\bigl(0,0,1-\cos(2t)\bigr)^T.$$
(A.1)

Indeed, the reader may check that \(\gamma(0)=p,\dot{\gamma}(0)=v,\dot{\gamma}(t)\perp T_{\gamma(t)}\mathbb{S},\|\dot{\gamma}(t)\|=1\) for all t.

1.2 A.2 Harmonic Analysis in Manifolds

The Hessian of a function fC 2(U) is a bilinear form defined as

$$\operatorname{Hess}({\mathcal{E}})(p) (v,w)=X\bigl(Y(f)\bigr)-(\nabla_XY)(f),\quad p\in U,v,w\in T_p\mathcal{M},$$

where ∇ is the Levi-Civita connection and X,Y are vector fields such that X(p)=v, Y(p)=w. In coordinates,

$$\operatorname{Hess}({f})\bigl(p(x)\bigr) (v,w)=w^t\bigl (h_{ij}(x)\bigr)v,\quad h_{ij}(x)=\frac{\partial^2f}{\partial x_i\partial x_j}-\sum_{k=1}^n\frac{\partial f}{\partial x_k}\varGamma _{ij}^k.$$

If γ(t) is the unique geodesic such that γ(0)=p and \(\dot{\gamma}(0)=v\), then

$$\operatorname{Hess}({f})(p) (v,v)=\frac{d^2}{dt^2}\bigg {|}_{t=0}f\bigl(\gamma (t)\bigr).$$

The LaplacianFootnote 3 \(\Delta {f}=\operatorname {div}(\operatorname{grad}(\mathcal{E}))\) of f is the trace of \(\operatorname{Hess}({f})\) as a bilinear operator, that is, the trace of (h ij ), or in coordinates,

$$\Delta {f}=\sum_{i=1}^nh_{ii}(x)=\sum_{i=1}^n \biggl(\frac{\partial^2f}{\partial x_i\partial x_i}\biggr)-\sum_{k=1}^n\frac{\partial f}{\partial x_k}\Biggl(\sum_{i=1}^n\varGamma _{ii}^k\Biggr).$$

A function fC 2(U), \(U\subseteq \mathcal{M}\) open, is called harmonic (subharmonic) if Δf=0 (Δf≥0). Note that Δ is a uniformly elliptic operator and satisfies the hypotheses of [13, Sect. 24]. Thus, the classical strong maximum principle applies. Namely:

Theorem 6.1

(Strong maximum principle)

Let fC 2(U) be a nonconstant subharmonic function. Then for any open set ΩU such that \(\overline{\varOmega }\subseteq U\), we have

$$x\in \varOmega \quad\Rightarrow\quad f(x)<\sup_{x\in \varOmega }f.$$

In particular, the supremum is reached only in the boundary of Ω.

The trace of (h ij ) is equal to the trace of Q (h ij )Q for any orthogonal matrix Q. Hence, we also have

$$\Delta {f}=\sum_{{v^{(1)},\ldots,v^{(2N)}}\text{\ an orthonormal basis of}\ T_x\mathbb{S}^N}\operatorname{Hess}({f})(x) \bigl(v^{(i)},v^{(i)} \bigr).$$

The classical mean value theorem for harmonic functions in ℝn (see for example [11, Sect. 2.2]) claims that, if f:U⊆ℝn → ℝ is harmonic (U an open set), then \(f(x)= \mathchoice {{\setbox0=\hbox{$\displaystyle{\textstyle-}{\int}$}\vcenter{\hbox{$\textstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\textstyle{\scriptstyle-}{\int}$}\vcenter{\hbox{$\scriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptscriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}\!\int _{B(x,\varepsilon)}f(y)\,dy\) for each ball B(x,ε)⊆U. Unfortunately, no similar equality is valid in general for harmonic functions on Riemannian manifolds, but there is a class of manifolds for which it holds. These are called locally harmonic manifolds, and the corresponding mean value equality was first proved by Willmore [22] (see also [5, Chap. 6]). Following the proof of [5, Prop. 6.21], if \(\mathcal{M}\) is a locally harmonic manifold and \(f:\mathcal{M}\,{\rightarrow}\,\mathbb{R}\) is such that Δf=C is constant, then for small enough t>0, we have

$$ \frac{d}{dt} \mathchoice {{\setbox0=\hbox{$\displaystyle{\textstyle-}{\int}$}\vcenter{\hbox{$\textstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\textstyle{\scriptstyle-}{\int}$}\vcenter{\hbox{$\scriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptscriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}\!\int _{S(x,t)}f=C \frac{\mathrm {Vol}(B(x,t))}{\mathrm {Vol}(S(x,t))}.$$
(A.2)

The sphere \(\mathbb{S}\subseteq \mathbb{R}^{3}\) is locally harmonic, and moreover, (A.2) can be stated for every t>0 such that f is defined in B(x,t). Thus, the classical mean value theorem is valid in the case of \(\mathbb{S}=\mathcal{M}\):

Theorem 6.2

Let f:U → ℝ (\(U\subseteq \mathbb{S}\) an open set) be such that Δf=C. Let pU. Then, for every ε>0 such that B p (ε)⊆U, we have

$$\mathchoice {{\setbox0=\hbox{$\displaystyle{\textstyle-}{\int}$}\vcenter{\hbox{$\textstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\textstyle{\scriptstyle-}{\int}$}\vcenter{\hbox{$\scriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}{{\setbox0=\hbox{$\scriptscriptstyle{\scriptscriptstyle-}{\int}$}\vcenter{\hbox{$\scriptscriptstyle-$}}\kern-.5\wd0}}\!\int _{S(p,\varepsilon)}f=f(p)+C\int_0^\varepsilon \frac{\pi\sin^2t}{\pi\sin(2t)}=f(p)-\frac{C}{2}\log(\cos \varepsilon).$$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Beltrán, C. Harmonic Properties of the Logarithmic Potential and the Computability of Elliptic Fekete Points. Constr Approx 37, 135–165 (2013). https://doi.org/10.1007/s00365-012-9158-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00365-012-9158-y

Keywords

Mathematics Subject Classification

Navigation