Skip to main content
Log in

Measuring centrality and dispersion in directional datasets: the ellipsoidal cone covering approach

  • Published:
Journal of Global Optimization Aims and scope Submit manuscript

Abstract

Consider a finite collection \(\{\xi _k\}_{k=1}^p\) of vectors in the space \(\mathbb {R}^n\). The \(\xi _k\)’s are not to be seen as position points but as directions. This work addresses the problem of computing the ellipsoidal cone of minimal volume that contains all the \(\xi _k\)’s. The volume of an ellipsoidal cone is defined as the usual n-dimensional volume of a certain truncation of the cone. The central axis of the ellipsoidal cone of minimal volume serves to define the central direction of the datapoints, whereas its volume can be used as measure of dispersion of the datapoints.

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

Similar content being viewed by others

References

  1. Ahipasaoglu, S.D., Sun, P., Todd, M.J.: Linear convergence of a modified Frank–Wolfe algorithm for computing minimum-volume enclosing ellipsoids. Optim. Methods Softw. 23, 5–19 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  2. Astorino, A., Gaudioso, M., Seeger, A.: An illumination problem: optimal apex and optimal orientation for a cone of light. J. Glob. Optim. 58, 729–750 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  3. Astorino, A., Gaudioso, M., Seeger, A.: Central axes and peripheral points in high dimensional directional datasets. Comput. Optim. Appl. (2015). doi:10.1007/s10589-014-9724-2

  4. Ball, K.: An elementary introduction to modern convex geometry. In: Levy, S. (ed.) Flavors of Geometry, vol. 31, pp. 1–58. Math. Sci. Res. Inst. Publ., Cambridge Univ. Press, Cambridge (1997)

  5. Berman, A.: Cones, matrices and mathematical programming. In: Lecture Notes in Economics and Mathematical Systems, vol. 79. Springer, Berlin (1973)

  6. Danskin, J.M.: The theory of max–min, with applications. SIAM J. Appl. Math. 14, 641–664 (1966)

    Article  MathSciNet  MATH  Google Scholar 

  7. Goffin, J.-L.: The relaxation method for solving systems of linear inequalities. Math. Oper. Res. 5, 388–414 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  8. Gourion, D., Seeger, A.: Solidity indices for convex cones. Positivity 16, 685–705 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  9. Gruber, P.M.: John and Loewner ellipsoids. Discrete Comput. Geom. 46, 776–788 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  10. Henrion, R., Seeger, A.: On properties of different notions of centers for convex cones. Set-Valued Var. Anal. 18, 205–231 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  11. Henrion, R., Seeger, A.: Inradius and circumradius of various convex cones arising in applications. Set-Valued Var. Anal. 18, 483–511 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  12. Iusem, A., Seeger, A.: Distances between closed convex cones: old and new results. J. Convex Anal. 17, 1033–1055 (2010)

    MathSciNet  MATH  Google Scholar 

  13. Jerónimo-Castro, J., McAllister, T.B.: Two characterizations of ellipsoidal cones. J. Convex Anal. 20, 1181–1187 (2013)

    MathSciNet  MATH  Google Scholar 

  14. Loewy, R., Schneider, H.: Positive operators on the n-dimensional ice cream cone. J. Math. Anal. Appl. 49, 375–392 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  15. Rockafellar, R.T.: Convex Analysis. Princeton University Press, Princeton (1970)

    Book  MATH  Google Scholar 

  16. Rockafellar, R.T., Wets, R .J.-B.: Variational Analysis. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  17. Seeger, A., Torki, M.: Centers and partial volumes of convex cones. I. Basic theory. Beitr. Algebra Geom. 56, 227–248 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  18. Seeger, A., Torki, M.: Conic version of Loewner–John ellipsoid theorem. Math. Program. 155, 403–433 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Stern, R., Wolkowicz, H.: Invariant ellipsoidal cones. Linear Algebra Appl. 150, 81–106 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  20. Sun, P., Freund, R.M.: Computation of minimum-volume covering ellipsoids. Oper. Res. 52, 690–706 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  21. Todd, M.J., Yildirim, E.A.: On Khachiyan’s algorithm for the computation of minimum-volume enclosing ellipsoids. Discrete Appl. Math. 155, 1731–1744 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  22. Watson, G.S.: Statistics on Spheres. Wiley, New York (1983)

    MATH  Google Scholar 

Download references

Acknowledgments

We would like to thank both referees for many helpful and constructive comments. The final form of this paper owes much to their recommendations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alberto Seeger.

Additional information

José Vidal -Nuñez is supported by Consellería de la Educación de la Generalitat Valenciana, Spain, Pre-Doc Program Vali+d, DOCV 6791/07.06.2012, Grant ACIF-2013-156.

Appendix

Appendix

As mentioned already on a number of occasions, two different matrices \( B, C\in \mathbb {GL}(n)\) may produce the same ellipsoidal cone. The statement (d) in the next proposition provides an easily verifiable necessary condition for having the equality

$$\begin{aligned} B^{-1}(\mathbb {L}_n) = C^{-1}(\mathbb {L}_n). \end{aligned}$$
(46)

A necessary and sufficient condition for (46) is given in Corollary 6.4. If P and D are convex cones in \(\mathbb {R}^n\), then \(Q \circeq P\) indicates that either \(Q=P\) or \(Q=-P\).

Proposition 6.3

For a pair of matrices \( B, C\in \mathbb {GL}(n)\), the following conditions are equivalent:

  1. (a)

    \(B^{-1}(\mathbb {L}_n) \circeq C^{-1}(\mathbb {L}_n)\).

  2. (b)

    \((BC^{-1})(\mathbb {L}_n) \circeq \mathbb {L}_n\).

  3. (c)

    \((BC^{-1})^\sharp \) is a positive multiple of \(J_n\).

  4. (d)

    \(B^\sharp \) is a positive multiple of \(C^\sharp \).

Proof

Clearly, (a) \(\Leftrightarrow \) (b) and (c) \(\Leftrightarrow \) (d). To prove the equivalence (b) \(\Leftrightarrow \) (c), there is no loss of generality in working with the particular case \(C=I_n\). So, we must check that

$$\begin{aligned} B(\mathbb {L}_n) \circeq \mathbb {L}_n\;\; \Leftrightarrow \;\; B^\sharp =\mu J_n \text{ for } \text{ some } \mu >0. \end{aligned}$$
(47)

In fact, the above equivalence is a known result by Loewy and Schneider [14, Theorem 2.4]. We give below an alternative proof of (47) that is based on Proposition 3.3 and the spectral decomposition

$$\begin{aligned} B^\sharp \;= \;U \mathrm{Diag}(\lambda _1,\ldots ,\lambda _n)U^T \,=\; \sum _{i=1}^{n-1}\lambda _i u_iu_i^T+ \lambda _nu_n u_n^T. \end{aligned}$$

Let \(B(\mathbb {L}_n) \circeq \mathbb {L}_n\). Hence, the ellipsoidal cone \(Q=B^{-1}(\mathbb {L}_n)\) is equal to \( \mathbb {L}_n\) or equal to \(-\mathbb {L}_n\). In particular, Q admits either \(u_n=e_n\) or \(u_n=-e_n\) as axial symmetry center. In addition, all the semiaxes lenghts of Q are equal to 1. From Proposition 3.3 we get \(\lambda _1= \ldots = \lambda _{n-1} =-\lambda _n\) and, a posteriori,

$$\begin{aligned} B^\sharp = \left[ \begin{array}{cc} -\lambda _n I_{n-1}&{}\quad 0 \\ 0 &{}\quad \lambda _n \\ \end{array} \right] = \lambda _n J_n. \end{aligned}$$

This shows that \(B^\sharp \) is a positive multiple of \(J_n\). Conversely, suppose that \(B^\sharp =\mu J_n \) for some \(\mu >0\). In such a case, \(u_n\in \{ e_n,-e_n\}\) and

$$\begin{aligned} \lambda _1= \ldots = \lambda _{n-1} =-\mu \,, \quad \lambda _n= \mu . \end{aligned}$$
(48)

Hence, the ellipsoidal cone \(Q=B^{-1}(\mathbb {L}_n)\) is symmetric with respect to the line generated by \(e_n\) and, thanks to (48) and Proposition 3.3, all the semi-axes lengths of Q are equal to 1. In conclusion, Q is either the Lorentz cone or the opposite of the Lorentz cone. \(\square \)

Corollary 6.4

Consider a pair \( B, C\in \mathbb {M}_n\) of invertible matrices and write

$$\begin{aligned} BC^{-1} = \left[ \begin{array}{cc} M &{}\quad z \\ q^T &{}\quad t \\ \end{array} \right] , \end{aligned}$$

with \(t\in \mathbb {R}, M\in \mathbb {M}_{n-1}\), and \(z,q\in \mathbb {R}^{n-1}\). Then

$$\begin{aligned} B^{-1}(\mathbb {L}_n) = C^{-1}(\mathbb {L}_n)\quad \Leftrightarrow \quad \left\{ \begin{array}{lll} t>\Vert z\Vert \,,\; M^Tz=t\, q,\; \text{ and } \\ M^TM-qq^T= (t^2- \Vert z\Vert ^2)I_{n-1}. \end{array} \right. \end{aligned}$$

Proof

By using Loewy and Schneider [14, Theorem 2.4] or the equivalence (b) \(\Leftrightarrow \) (c) in Proposition 6.3, we get

$$\begin{aligned} BC^{-1}(\mathbb {L}_n)\circeq \mathbb {L}_n\quad \Leftrightarrow \quad \left\{ \begin{array}{lll} t^2- \Vert z\Vert ^2 &{}>&{}0\\ M^Tz-t\, q&{}=&{}0\\ M^TM-qq^T&{}=&{} (t^2- \Vert z\Vert ^2)I_{n-1}. \end{array} \right. \end{aligned}$$

The case \(BC^{-1}(\mathbb {L}_n)= \mathbb {L}_n\) occurs when the last entry of \( BC^{-1}e_n \) is positive, i.e., when \(t>0\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Seeger, A., Vidal -Nuñez, J. Measuring centrality and dispersion in directional datasets: the ellipsoidal cone covering approach. J Glob Optim 68, 279–306 (2017). https://doi.org/10.1007/s10898-016-0464-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10898-016-0464-y

Keywords

Mathematics Subject Classification

Navigation