Skip to main content

Advertisement

Log in

A generalization of Löwner-John’s ellipsoid theorem

  • Full Length Paper
  • Series A
  • Published:
Mathematical Programming Submit manuscript

Abstract

We address the following generalization \(\mathbf {P}\) of the Löwner-John ellipsoid problem. Given a (non necessarily convex) compact set \(\mathbf {K}\subset \mathbb {R}^n\) and an even integer \(d\in \mathbb {N}\), find an homogeneous polynomial \(g\) of degree \(d\) such that \(\mathbf {K}\subset \mathbf {G}:=\{\mathbf {x}:g(\mathbf {x})\le 1\}\) and \(\mathbf {G}\) has minimum volume among all such sets. We show that \(\mathbf {P}\) is a convex optimization problem even if neither \(\mathbf {K}\) nor \(\mathbf {G}\) are convex! We next show that \(\mathbf {P}\) has a unique optimal solution and a characterization with at most \({n+d-1\atopwithdelims ()d}\) contacts points in \(\mathbf {K}\cap \mathbf {G}\) is also provided. This is the analogue for \(d>2\) of Löwner-John’s theorem in the quadratic case \(d=2\), but importantly, we neither require the set \(\mathbf {K}\) nor the sublevel set \(\mathbf {G}\) to be convex. More generally, there is also an homogeneous polynomial \(g\) of even degree \(d\) and a point \(\mathbf {a}\in \mathbb {R}^n\) such that \(\mathbf {K}\subset \mathbf {G}_\mathbf {a}:=\{\mathbf {x}:g(\mathbf {x}-\mathbf {a})\le 1\}\) and \(\mathbf {G}_\mathbf {a}\) has minimum volume among all such sets (but uniqueness is not guaranteed). Finally, we also outline a numerical scheme to approximate as closely as desired the optimal value and an optimal solution. It consists of solving a hierarchy of convex optimization problems with strictly convex objective function and Linear Matrix Inequality constraints.

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

Similar content being viewed by others

Notes

  1. For instance some well-known NP-hard 0/1 optimization problems reduce to conic LP optimization problems over the convex cone of copositive matrices (and/or its dual) for which the associated membership problem is hard.

  2. We thank Pham Tien Son for providing these two examples.

  3. We have used the GloptiPoly software [20] dedicated to solving the generalized problem of moments.

  4. A semidefinite program is a finite-dimensional convex optimization problem which in canonical form reads: \(\min _\mathbf {x}\{\mathbf {c}^T\mathbf {x}:\mathbf {A}_0+\sum _{k=1}^t\mathbf {A}_kx_k\succeq 0\}\), where \(\mathbf {c}\in \mathbb {R}^t\) and the \(\mathbf {A}_k\)’s are real symmetric matrices. Importantly, up to arbitrary fixed precision it can be solved in time polynomial in the input size of the problem.

  5. A Linear Matrix Inequality (LMI) is a constraint of the form \(\mathbf {A}(\mathbf {x}):=\mathbf {A}_0+\sum _{\ell =1}^t\mathbf {A}_\ell x_\ell \succeq 0\) where each \(\mathbf {A}_\ell \), \(\ell =0,\ldots ,t\), is a real symmetric matrix; so each entry of the real symmetric matrix \(\mathbf {A}(\mathbf {x})\) is affine in \(\mathbf {x}\in \mathbb {R}^t\). An LMI always define a convex set, i.e., the set \(\{\mathbf {x}\in \mathbb {R}^t:\mathbf {A}(\mathbf {x})\succeq 0\}\) is convex.

References

  1. Anastassiou, G.A.: Moments in Probability Approximation Theory. Longman Scientific & Technical, UK (1993)

    Google Scholar 

  2. d’Aspremont, A.: Smooth optimization with approximate gradient. SIAM J. Optim. 19, 1171–1183 (2008)

    Article  MathSciNet  Google Scholar 

  3. Ash, R.B.: Real Analysis and Probability. Academic Press Inc., Boston (1972)

    Google Scholar 

  4. Ball, K.: Ellipsoids of maximal volume in convex bodies. Geom. Dedicata 41, 241–250 (1992)

    Article  MathSciNet  Google Scholar 

  5. Ball, K.: Convex geometry and functional analysis. In: Johnson, W.B., Lindenstrauss, J. (eds.) Handbook of the Geometry of Banach Spaces I, pp. 161–194. North Holland, Amsterdam (2001)

    Chapter  Google Scholar 

  6. Barvinok, A.I.: Computing the volume, counting integral points, and exponential sums. Discrete Comput. Geom. 10, 123–141 (1993)

    Article  MathSciNet  Google Scholar 

  7. Bastero, J., Romance, M.: John’s decomposition of the identity in the non-convex case. Positivity 6, 1–16 (2002)

    Article  MathSciNet  Google Scholar 

  8. Bayer, C., Teichmann, J.: The proof of Tchakaloff’s theorem. Proc. Am. Math. Soc. 134, 303–3040 (2006)

    Article  MathSciNet  Google Scholar 

  9. Bookstein, F.L.: Fitting conic sections to scattered data. Comp. Graph. Image. Process. 9, 56–71 (1979)

    Article  Google Scholar 

  10. Calafiore, G.: Approximation of \(n\)-dimnsional data using spherical and ellipsoidal primitives. IEEE Trans. Syst. Man. Cyb. 32, 269–276 (2002)

    Article  Google Scholar 

  11. Chernousko, F.L.: Guaranteed estimates of undetermined quantities by means of ellipsoids. Sov. Math. Dodkl. 21, 396–399 (1980)

    Google Scholar 

  12. Croux, C., Haesbroeck, G., Rousseeuw, P.J.: Location adjustment for the minimum volume ellipsoid estimator. Stat. Comput. 12, 191–200 (2002)

    Article  MathSciNet  Google Scholar 

  13. Giannopoulos, A., Perissinaki, I., Tsolomitis, A.: A John’s theorem for an arbitrary pair of convex bodies. Geom. Dedicata 84, 63–79 (2001)

    Article  MathSciNet  Google Scholar 

  14. Dyer, M.E., Frieze, A.M., Kannan, R.: A random polynomial-time algorithm for approximating the volume of convex bodies. J. ACM 38, 1–17 (1991)

    Article  MathSciNet  Google Scholar 

  15. Faraut, J., Korányi, A.: Analysis on Symmetric Cones. Clarendon Press, Oxford (1994)

    Google Scholar 

  16. Freitag, E., Busam, R.: Complex Analysis, 2nd edn. Springer, Berlin (2009)

    Google Scholar 

  17. Gander, W., Golub, G.H., Strebel, R.: Least-squares fitting of circles and ellipses. BIT 34, 558–578 (1994)

    Article  MathSciNet  Google Scholar 

  18. Helton, J.W., Nie, J.: A semidefinite approach for truncated K-moment problems. Fond. Comput. Math. 12, 851–881 (2012)

    Article  MathSciNet  Google Scholar 

  19. Henk, M.: Löwner-John ellipsoids. Documenta Math. 95–106 (2012) (extra volume: optimization stories)

  20. Henrion, D., Lasserre, J.B., Lofberg, J.: Gloptipoly 3: moments, optimization and semidefinite programming. Optim. Method. Softw. 24, 761–779 (2009)

    Article  MathSciNet  Google Scholar 

  21. Henrion, D., Lasserre, J.B., Savorgnan, C.: Approximate volume and integration of basic semi-algebraic sets. SIAM Rev. 51, 722–743 (2009)

    Article  MathSciNet  Google Scholar 

  22. Henrion, D., Peaucelle, D., Arzelier, D., Sebek, M.: Ellipsoidal approximation of the stability domain of a polynomial. IEEE Trans. Autom. Control 48, 2255–2259 (2003)

    Article  MathSciNet  Google Scholar 

  23. Henrion, D., Lasserre, J.B.: Inner approximations for polynomial matrix inequalities and robust stability regions. IEEE Trans. Autom. Control 57, 1456–1467 (2012)

    Article  MathSciNet  Google Scholar 

  24. Henrion, D., Sebek, M., Kucera, V.: Positive polynomials and and robust stabilization with fixed-order controllers. IEEE Trans. Autom. Control 48, 1178–1186 (2003)

    Article  MathSciNet  Google Scholar 

  25. Hiriart-Urruty, J.B., Lemarechal, C.: Convex Analysis and Minimization Algorithms I. Springer, Berlin (1993)

    Google Scholar 

  26. Hiriart-Urruty, J.B., Lemarechal, C.: Convex Analysis and Minimization Algorithms II. Springer, Berlin (1993)

    Google Scholar 

  27. Karimi, A., Khatibi, H., Longchamp, R.: Robust control of polytopic systems by convex optimization. Automatica 43, 1395–1402 (2007)

    Article  MathSciNet  Google Scholar 

  28. Kemperman, J.H.B.: Geometry of the moment problem. In: H.J. Landau (Ed.) Moments in Mathematics. Proceedings of Symposia in Applied Mathematics, vol. 37 pp. 16–53 (1987)

  29. Kemperman, J.H.B.: The general moment problem, a geometric approach. Annals Math. Stat. 39, 93–122 (1968)

    Article  MathSciNet  Google Scholar 

  30. Lasserre, J.B.: Global optimization with polynomials and the problem of moments. SIAM J. Optim. 11, 796–817 (2001)

    Article  MathSciNet  Google Scholar 

  31. Lasserre, J.B.: Moments, Positive Polynomials and Their Applications. Imperial College, London (2009)

    Book  Google Scholar 

  32. Lasserre, J.B.: Recovering an homogeneous polynomial from moments of its level set. Discrete Comput. Geom. 50, 673–678 (2013)

    Article  MathSciNet  Google Scholar 

  33. Morosov, A., Shakirov, S.: New and old results in resultant theory. Theor. Math. Phys. 163, 587–617 (2010)

    Article  Google Scholar 

  34. Morosov, A., Shakirov, S.: Introduction to integral discriminants, J. High Energy Phys. 12 (2009). arXiv:0911.5278v1

  35. Nurges, U.: Robust pole assignment via reflection coefficientsof polynomials. Automatica 42, 1223–1230 (2006)

    Article  MathSciNet  Google Scholar 

  36. O’Rourke, J., Badler, N.I.: Decomposition of three-dimensional objets into spheres. IEEE Trans. Pattern Anal. Machine Intell. 1, 295–305 (1979)

    Article  Google Scholar 

  37. Pratt, V.: Direct least squares fitting of algebraic surfaces. ACM J. Comp. Graph. 21, 145–152 (1987)

  38. Putinar, M.: Positive polynomials on compact semi-algebraic sets. Indiana Univ. Math. J. 42, 969–984 (1993)

    Article  MathSciNet  Google Scholar 

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

    Google Scholar 

  40. Rosen, J.B.: Pattern separation by convex programming techniques. J. Math. Anal. Appl. 10, 123–134 (1965)

    Article  MathSciNet  Google Scholar 

  41. Rosin, P.L.: A note on the least squares fitting of ellipses. Pattern Recog. Lett. 14, 799–808 (1993)

    Article  Google Scholar 

  42. Rosin, P.L., West, G.A.: Nonparametric segmentation of curves into various representations. IEEE Trans. Pattern Anal. Machine Intell. 17, 1140–1153 (1995)

    Article  Google Scholar 

  43. Rousseeuw, P.J., Leroy, A.M.: Robust Regression and Outlier Detection. John Wiley, New York (1987)

    Book  Google Scholar 

  44. Royden, H.L.: Real Analysis. Macmillan, New York (1968)

    Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  46. Taubin, G.: Estimation of planar curves, surfaces and nonplanar space curves defined by implicit equations, with applications to to edge and range image segmentation. IEEE Trans. Pattern Anal. Machine Intell. 13, 1115–1138 (1991)

    Article  Google Scholar 

  47. Vandenberghe, L., Boyd, S.: Semidefinite programming. SIAM Rev. 38, 49–95 (1996)

    Article  MathSciNet  Google Scholar 

  48. Widder, D.V.: The Laplace Transform. Princeton University Press, Princeton (1946)

    Google Scholar 

Download references

Acknowledgments

This work was partially supported by a grant from the Gaspar Monge Program for Optimization and Operations Research (PGMO) of the Fondation Mathématique Jacques Hadamard (France).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean B. Lasserre.

Appendix

Appendix

1.1 First-order KKT-optimality conditions

Consider the finite dimensional optimization problem:

$$\begin{aligned} \inf \,\{\,f(\mathbf {x}):\,\mathbf {A}\mathbf {x}\,=\,\mathbf {b};\,\mathbf {x}\in \,C\,\}, \end{aligned}$$

for some real matrix \(\mathbf {A}\in \mathbb {R}^{m\times n}\), vector \(\mathbf {b}\in \mathbb {R}^m\), some closed convex cone \(C\subset \mathbb {R}^n\) (with dual cone \(C^*=\{\,\mathbf {y}:\mathbf {y}^T\mathbf {x}\ge 0,\,\forall \mathbf {x}\,\in C\,\}\)) and some convex and differentiable function \(f\) with domain \(D\). Suppose that \(C\) has a nonempty interior \(\mathrm{int}(C)\) and Slater’s condition holds, that is, there exists \(\mathbf {x}_0\in D\cap \mathrm{int}(C)\) such that \(\mathbf {A}\mathbf {x}_0=\mathbf {b}\). The normal cone at a point \(0\ne \mathbf {x}\in C\) is the set \(N_C(\mathbf {x})=\{\mathbf {y}\in C^*:\,\langle \mathbf {y},\mathbf {x}\rangle =0\}\) (see e.g. [25, p. 189]).

Then by Theorem 5.3.3, p. 188 in [26], \(\mathbf {x}^*\in C\) is an optimal solution if and only if there exists \((\lambda ,\mathbf {y})\in \mathbb {R}^m\times N_C(\mathbf {x}^*)\) such that:

$$\begin{aligned} \mathbf {A}\mathbf {x}^*=\mathbf {b};\quad \nabla f(\mathbf {x}^*)+\mathbf {A}^T\lambda \,=\,\mathbf {y}\end{aligned}$$
(7.1)

and \(\langle \mathbf {x}^*,\mathbf {y}\rangle =0\) follows because \(\mathbf {y}\in N_C(\mathbf {x}^*)\).

1.2 Measures with finite support

We restate the following important result stated in [29, Theorem 1] and [1, Theorem 2.1.1, p. 39].

Theorem 7.1

([1, 29]) Let \(f_1,\ldots ,f_N\) be real-valued Borel measurable functions on a measurable space \(\Omega \) and let \(\mu \) be a probability measure on \(\Omega \) such that each \(f_i\) is integrable with respect to \(\mu \). Then there exists a probability \(\nu \) with finite support in \(\Omega \) and such that:

$$\begin{aligned} \int _\Omega f_i(\mathbf {x})\,\mu (d\mathbf {x})\,=\,\int _\Omega f_i(\mathbf {x})\,\nu (d\mathbf {x}),\quad i=1\ldots ,N. \end{aligned}$$

One can even attain that the support of \(\nu \) has at most \(N+1\) points.

In fact if \(\mathcal {M}(\Omega )_+\) denotes the space of probability measures on \(\Omega \), then the moment space

$$\begin{aligned} Y_N:=\left\{ \mathbf {y}\,=\,\left( \int _\Omega f_k(\mathbf {x})d\mu (\mathbf {x})\right) ,\,k=1,\ldots ,N,\quad \text{ for } \text{ some } \mu \in \mathcal {M}(\Omega )_+\right\} \end{aligned}$$

is the convex hull of the set \(f(\Omega )=\{(f_1(\mathbf {x}),\ldots ,f_N(\mathbf {x})):\,\mathbf {x}\in \Omega \}\) and each point \(\mathbf {y}\in Y_N\) can be represented as the convex hull of at most \(N+1\) point \(f(\mathbf {x}_i)\), \(i=1,\ldots ,N+1\). (See e.g. Sect. 3, p. 29 in Kemperman [28].)

In the proof of Theorem 3.2 one uses Theorem 7.1 with the \(f_i\)’s being all monomials \((\mathbf {x}^\alpha )\) of degree equal to \(d\) (and so \(N={n+d-1\atopwithdelims ()d}\)). We could also use Tchakaloff’s Theorem [8] but then we would potentially need \({n+d\atopwithdelims ()d}\) points. An alternative would be to use Tchakaloff’s Theorem after “de-homogenizing” the measure \(\mu \) so that \(n\)-dimensional moments of order \(\vert \alpha \vert =d\) become \((n-1)\)-dimensional moments of order \(\vert \alpha \vert \le d\), and one retrieves the bound \({n-1+d\atopwithdelims ()d}\).

1.3 Proof of Theorem 3.2

Proof

  1. (a)

    As \(\mathcal {P}\) is a minimization problem, its feasible set \(\{\,g\in \mathbf {H}[\mathbf {x}]_{d}:1-g\in C_{d}(\mathbf {K})\,\}\) can be replaced by the smaller set

    $$\begin{aligned} F\,:=\,\left\{ g\in \mathbf {H}[\mathbf {x}]_{d}\,:\,\begin{array}{l}\displaystyle \int _{\mathbb {R}^n}\exp (-g(\mathbf {x}))\,d\mathbf {x}\,\le \,\int _{\mathbb {R}^n}\exp (-g_0(\mathbf {x}))\,d\mathbf {x}\\ 1-g\in \,C_{d}(\mathbf {K}) \end{array}\right\} , \end{aligned}$$

for some \(g_0\in \mathbf {P}[\mathbf {x}]_{d}\). Notice that \(F\subset \mathbf {P}[\mathbf {x}]_d\) and \(F\) is a closed convex set since the convex function \(g\mapsto \int _{\mathbb {R}^n}\exp (-g)d\mathbf {x}\) is continuous on the interior of its domain.

Next, let \(\mathbf {z}=(z_\alpha ), \alpha \in \mathbb {N}^n_{d}\), be a (fixed) element of \(\mathrm{int}(C_{d}(\mathbf {K})^*)\) (hence \(z_0>0\)). By Lemma 2.6 such an element exists and \(\langle \mathbf {z},\mathbf {g}\rangle \ge 0\) (as \(g\in \mathbf {P}[\mathbf {x}]_d\) is nonnegative). Next as \(\langle \mathbf{z}, 1-\mathbf{g} \rangle \ge 0\) one has \(z_0 \ge \langle \mathbf{z, g}\rangle \) and so by Corollary I.1.6 in Faraut and Korányi [15], g is bounded. Therefore the set \(F\) is a compact convex set. Finally, since \(g\mapsto \int _{\mathbb {R}^n}\exp (-g(\mathbf {x}))d\mathbf {x}\) is strictly convex, it is continuous on the interior of its domain and so it is continuous on \(F\). Hence problem \(\mathcal {P}\) has a unique optimal solution \(g^*\in \mathbf {P}[\mathbf {x}]_d\).

  1. (b)

    We may and will consider any homogeneous polynomial \(g\) as an element of \(\mathbb {R}[\mathbf {x}]_{d}\) whose coefficient vector \(\mathbf {g}=(g_\alpha )\) is such that \(g^*_\alpha =0\) whenever \(\vert \alpha \vert <d\). And so Problem \(\mathcal {P}\) is equivalent to the problem

    $$\begin{aligned} \mathcal {P}':\quad \left\{ \begin{array}{ll}\rho =\displaystyle \inf _{g\in \mathbb {R}[\mathbf {x}]_{d}}&{}\displaystyle \int _{\mathbb {R}^n} \exp (-g(\mathbf {x}))\,d\mathbf {x}\\ \text{ s.t. }&{}g_\alpha =0,\quad \forall \,\alpha \in \mathbb {N}^n_{d};\,\vert \alpha \vert <d\\ &{}1-g\,\in \,C_{d}(\mathbf {K}),\end{array}\right. \end{aligned}$$
    (7.2)

    where we replaced \(g\in \mathbf {P}[\mathbf {x}]_d\) with the equivalent constraints \(g\in \mathbb {R}[\mathbf {x}]_{d}\) and \(g_\alpha :=0\) for all \(\alpha \in \mathbb {N}^n_{d}\) with \(\vert \alpha \vert <d\). Next, doing the change of variable \(h=1-g\), \(\mathcal {P}\)’ reads:

    $$\begin{aligned} \mathcal {P}':\quad \left\{ \begin{array}{ll}\rho =\displaystyle \inf _{h\in \mathbb {R}[\mathbf {x}]_{d}}&{}\displaystyle \int _{\mathbb {R}^n} \exp (h(\mathbf {x})-1)\,d\mathbf {x}\\ \text{ s.t. }&{}h_\alpha =0,\quad \forall \,\alpha \in \mathbb {N}^n_{d};\,0<\vert \alpha \vert <d\\ &{}h_0=1\\ &{}h\,\in \,C_{d}(\mathbf {K}),\end{array}\right. \end{aligned}$$
    (7.3)

As \(\mathbf {K}\) is compact, there exists \(\theta \in \mathbf {P}[\mathbf {x}]_{d}\) such that \(1-\theta \in \mathrm{int}(C_{d}(\mathbf {K}))\), i.e., Slater’s condition holds for the convex optimization problem \(\mathcal {P}'\). Indeed, choose \(\mathbf {x}\mapsto \theta (\mathbf {x}):=M^{-1}\Vert \mathbf {x}\Vert ^{d}\) for \(M>0\) sufficiently large so that \(1-\theta >0\) on \(\mathbf {K}\). Hence with \(\Vert g\Vert _1\) denoting the \(\ell _1\)-norm of the coefficient vector of \(g\) (in \(\mathbb {R}[\mathbf {x}]_{d}\)), there exists \(\epsilon >0\) such that for every \(h\in B(\theta ,\epsilon )(:=\{h\in \mathbb {R}[\mathbf {x}]_{d}:\Vert \theta -h\Vert _1<\epsilon \}\)), the polynomial \(1-h\) is (strictly) positive on \(\mathbf {K}\).

Therefore, the unique optimal solution \((1-g^*)=:h^*\in \mathbb {R}[\mathbf {x}]_{d}\) of \(\mathcal {P}\)’ in (7.3) satisfies the Karush-Kuhn-Tucker (KKT) optimality conditions (7.1) which for problem (7.3) read:

$$\begin{aligned} \int _{\mathbb {R}^n}\mathbf {x}^\alpha \,\exp (h^*(\mathbf {x})-1)\,d\mathbf {x}&= y^*_\alpha , \quad \forall \vert \alpha \vert =d\end{aligned}$$
(7.4)
$$\begin{aligned} \int _{\mathbb {R}^n}\mathbf {x}^\alpha \,\exp (h^*(\mathbf {x})-1)\,d\mathbf {x}+ \gamma _\alpha&= y^*_\alpha ,\quad \forall \,\vert \alpha \vert <d\end{aligned}$$
(7.5)
$$\begin{aligned} \langle h^*,\mathbf {y}^*\rangle \,=\,0;\quad h^*_0\,=\,1;\,h^*_\alpha&= 0,\quad \forall \,0<\vert \alpha \vert <d \end{aligned}$$
(7.6)

for some \(\mathbf {y}^*=(y^*_\alpha )\), \(\alpha \in \mathbb {N}^n_{d}\), in the dual cone \(C_{d}(\mathbf {K})^*\subset \mathbb {R}^{s(d)}\) of \(C_{d}(\mathbf {K})\), and some vector \(\gamma =(\gamma _\alpha )\), \(0\le \vert \alpha \vert <d\). By Lemma 2.5,

$$\begin{aligned} C_{d}(\mathbf {K})^*\,=\,\left\{ \mathbf {y}\in \mathbb {R}^{s(d)}\,{:}\, \exists \mu \in \mathcal {M}(\mathbf {K})_+ \text{ s.t. } y_\alpha =\int _\mathbf {K}\mathbf {x}^\alpha \,d\mu ,\,\alpha \in \mathbb {N}^n_{d}\,\right\} , \end{aligned}$$

and so (3.2) is just (7.4) restated in terms of \(\mu ^*\).

Next, the condition \(\langle h^*,\mathbf {y}^*\rangle =0\) (or equivalently, \(\langle 1-g^*,\mathbf {y}^*\rangle =0\)), reads:

$$\begin{aligned} \int _{\mathbf {K}} (1-g^*)\,d\mu ^*\,=\,0, \end{aligned}$$

which combined with \(1-g^*\in C_{d}(\mathbf {K})\) and \(\mu ^*\in \mathcal {M}(\mathbf {K})_+\), implies that \(\mu ^*\) is supported on \(\mathbf {K}\cap \{\mathbf {x}:g^*(\mathbf {x})=1\}=\mathbf {K}\cap \mathbf {G}^*_1\).

Next, let \(s:=\sum _{\vert \alpha \vert =d}g^*_\alpha y^*_\alpha \,(=y^*_0)\). From \(\langle 1-g^*,\mu ^*\rangle =0\), the measure \(s^{-1}\mu ^*=:\psi \) is a probability measure supported on \(\mathbf {K}\cap \mathbf {G}^*_1\), and satisfies \(\int \mathbf {x}^\alpha d\psi =s^{-1}y^*_\alpha \) for all \(\vert \alpha \vert =d\) (and \(\langle 1-g^*,\psi \rangle =0\)).

Hence by Theorem 7.1 there exists an atomic probability measure \(\nu ^*\in \mathcal {M}(\mathbf {K}\cap \mathbf {G}^*_1)_+\) such that

$$\begin{aligned} \int _{\mathbf {K}\cap \mathbf {G}^*_1}\mathbf {x}^\alpha d\nu ^*(\mathbf {x})\,=\,\int _{\mathbf {K}\cap \mathbf {G}^*_1}\mathbf {x}^\alpha d\psi (\mathbf {x})\,=\, s^{-1}\,y^*_\alpha ,\qquad \forall \,\vert \alpha \vert =d. \end{aligned}$$

In addition \(\nu ^*\) may be chosen to be supported on at most \(N={n+d-1\atopwithdelims ()d}\) points in \(\mathbf {K}\cap \mathbf {G}^*_1\) and not \(N+1\) points as predicted by Theorem 7.1. This is because one among the \(N\) conditions

$$\begin{aligned} \int _{\mathbf {K}\cap \mathbf {G}_1^*}\mathbf {x}^\alpha \,d\nu ^*\,=\,s^{-1}\, y^*_\alpha ,\quad \vert \alpha \vert \,=\,d, \end{aligned}$$

is redundant as \(\langle g^*,\mathbf {y}^*\rangle =y^*_0\) and \(\nu ^*\) is supported on \(\mathbf {K}\cap \mathbf {G}_1^*\). In other words, \(\mathbf {y}^*\) is not in the interior of the moment space \(Y_N\). Hence in (3.2) the measure \(\mu ^*\) can be substituted with the atomic measure \(s\,\nu ^*\) supported on at most \(n+d-1\atopwithdelims ()d\) contact points in \(\mathbf {K}\cap \mathbf {G}^*_1\).

To obtain \(\mu ^*(\mathbf {K})=\frac{n}{d}\int _{\mathbb {R}^n}\exp (-g^*)\), multiply both sides of (7.4)-(7.5) by \(h^*_\alpha \) for every \(\alpha \ne 0\), sum up and use \(\langle h^*,\mathbf {y}^*\rangle =0\) to obtain

$$\begin{aligned} -y^*_0=\sum _{\alpha \ne 0} h^*_\alpha \,y^*_\alpha&= \int _{\mathbb {R}^n}(h^*(\mathbf {x})-1)\exp (h^*(\mathbf {x})-1)\,d\mathbf {x}\\&= -\int _{\mathbb {R}^n}g^*(\mathbf {x})\exp (-g^*(\mathbf {x}))\,d\mathbf {x}\\&= -\frac{n}{d}\int \exp (-g^*(\mathbf {x}))\,d\mathbf {x}, \end{aligned}$$

where we have also used (2.5).

  1. (c)

    Let \(\mu ^*:=\sum _{i=1}^s \lambda _i\delta _{\mathbf {x}_i}\) where \(\delta _{\mathbf {x}_i}\) is the Dirac measure at the point \(\mathbf {x}_i\in \mathbf {K}\), \(i=1,\ldots ,s\). Next, let \(y^*_\alpha :=\int \mathbf {x}^\alpha d\mu ^*\) for all \(\alpha \in \mathbb {N}^n_{d}\), so that \(\mathbf {y}^*\in C_d(\mathbf {K})^*\). In particular \(\mathbf {y}^*\) and \(g^*\) satisfy

    $$\begin{aligned} \langle 1-g^*,\mathbf {y}^*\rangle \,=\,\int _\mathbf {K}( 1-g^*)d\mu ^*\,=\,0, \end{aligned}$$

    because \(g^*(\mathbf {x}_i)=1\) for all \(i=1,\ldots ,s\). In other words, the pair \((g^*,\mathbf {y}^*)\) satisfies the KKT-optimality conditions associated with the convex problem \(\mathcal {P}\). But since Slater’s condition holds for \(\mathcal {P}\), those conditions are also sufficient for \(g^*\) to be an optimal solution of \(\mathcal {P}\), the desired result

\(\square \)

1.4 Proof of Theorem 4.1

Proof

First observe that (4.2) reads

$$\begin{aligned} \mathcal {P}:\quad \inf _{\mathbf {a}\in \mathbb {R}^n}\,\left\{ \inf _{g\in \mathbf {P}[\mathbf {x}]_d}\,\left\{ \mathrm{vol}(\mathbf {G}^\mathbf {a}_1)\,:\,1-g_\mathbf {a}\in C_d(\mathbf {K})\right\} \right\} , \end{aligned}$$
(7.7)

and notice that the constraint \(1-g_\mathbf {a}\in C(\mathbf {K})\) is the same as \(1-g\in C(\mathbf {K}-\mathbf {a})\). And so for every \(\mathbf {a}\in \mathbb {R}^n\), the inner minimization problem

$$\begin{aligned} \inf _{g\in \mathbf {P}[\mathbf {x}]_d}\,\left\{ \mathrm{vol}(\mathbf {G}^\mathbf {a}_1)\,:\,1-g_\mathbf {a}\in C_d(\mathbf {K})\right\} \end{aligned}$$

of (7.7) reads

$$\begin{aligned} \rho _\mathbf {a}\,=\,\inf _{g\in \mathbf {P}[\mathbf {x}]_d}\,\left\{ \mathrm{vol}(\mathbf {G}_1)\,:\,1-g\in C_d(\mathbf {K}-\mathbf {a})\right\} . \end{aligned}$$
(7.8)

From Theorem 3.2 (with \(\mathbf {K}-\mathbf {a}\) in lieu of \(\mathbf {K}\)), problem (7.8) has a unique minimizer \(g^\mathbf {a}\in \mathbf {P}[\mathbf {x}]_d\) with value \(\rho _\mathbf {a}=\int _{\mathbb {R}^n}\exp (-g^\mathbf {a})d\mathbf {x}=\int _{\mathbb {R}^n}\exp (-g^\mathbf {a}_\mathbf {a})d\mathbf {x}\).

Therefore, in a minimizing sequence \((\mathbf {a}_\ell ,g^{\mathbf {a}_\ell })\subset \mathbb {R}^n\times \mathbf {P}[\mathbf {x}]_d\), \(\ell \in \mathbb {N}\), for problem \(\mathcal {P}\) in (4.2) with

$$\begin{aligned} \rho =\lim _{\ell \rightarrow \infty }\,\int _{\mathbb {R}^n}\exp (-g^{\mathbf {a}_\ell })d\mathbf {x}, \end{aligned}$$

we may and will consider that for every \(\ell \), the homogeneous polynomial \(g^{\mathbf {a}_\ell }\in \mathbf {P}[\mathbf {x}]_d\)) solves the inner minimization problem (7.8) with \(\mathbf {a}_\ell \) fixed. For simplicity of notation rename \(g^{\mathbf {a}_\ell }\) as \(g^\ell \) and \(g^{\mathbf {a}_\ell }_{\mathbf {a}_\ell }\) (\(=g^{\mathbf {a}_\ell }(\mathbf {x}-\mathbf {a}_\ell )\)) as \(g^\ell _{\mathbf {a}_\ell }\).

As observed in the proof of Theorem 3.2, there is \(\mathbf {z}\in \mathrm{int}(C_d(\mathbf {K})^*)\) such that \(\langle 1-g^\ell _{\mathbf {a}_\ell },\mathbf {z}\rangle \ge 0\) and by Corollary I.1.6 in Faraut et Korányi [15], the set \(\{h\in C_{d}(\mathbf {K}):\langle \mathbf {z},h\rangle \le z_0\}\) is compact.

Also, \(\mathbf {a}_\ell \) can be chosen with \(\Vert \mathbf {a}_\ell \Vert \le M\) for all \(\ell \) (and some \(M\)), otherwise the constraint \(1-g_{\mathbf {a}_\ell }\in C_d(\mathbf {K})\) would impose a much too large volume \(\mathrm{vol}(\mathbf {G}^{\mathbf {a}_\ell }_1)\).

Therefore, there is a subsequence \((\ell _k)\), \(k\in \mathbb {N}\), and a point \((\mathbf {a}^*,\theta ^*)\in \mathbb {R}^n\times C_d(\mathbf {K})\) such that

$$\begin{aligned} \lim _{k\rightarrow \infty }\mathbf {a}_{\ell _k}\,=\,\mathbf {a}^*;\qquad \lim _{k\rightarrow \infty }(g^{\ell _k}_{\mathbf {a}_{\ell _k}})_\alpha \,=\,\theta ^*_\alpha ,\quad \forall \alpha \in \mathbb {N}^n_d. \end{aligned}$$

Recall the definition (4.1) of \(g^\ell _{\mathbf {a}_\ell }(\mathbf {x})=g^\ell (\mathbf {x}-\mathbf {a}_\ell )\) for the homogeneous polynomial \(g^\ell \in \mathbf {P}[\mathbf {x}]_d\) with coefficient vector \(\mathbf {g}^\ell \), i.e.,

$$\begin{aligned} \left( g^\ell _{\mathbf {a}_{\ell }}\right) _\alpha \,=\,p_\alpha (\mathbf {a}_\ell ,\mathbf {g}^\ell ), \qquad \forall \alpha \in \mathbb {N}^n_d, \end{aligned}$$

for some polynomials \((p_\alpha )\subset \mathbb {R}[\mathbf {x},\mathbf {g}]\), \(\alpha \in \mathbb {N}^n_d\). In particular, for every \(\alpha \in \mathbb {N}^n_d\) with \(\vert \alpha \vert =d\), \(p_\alpha (\mathbf {a}_\ell ,\mathbf {g}^\ell )=(g^\ell )_\alpha \). And so for every \(\alpha \in \mathbb {N}^n_d\) with \(\vert \alpha \vert =d\),

$$\begin{aligned} \theta ^*_\alpha \,=\,\lim _{k\rightarrow \infty }\,=\,(g^{\ell _k})_\alpha . \end{aligned}$$

If we define the homogeneous polynomial \(g^*\) of degree \(d\) by \((g^*)_\alpha =\theta ^*_\alpha \) for every \(\alpha \in \mathbb {N}^n_d\) with \(\vert \alpha \vert =d\), then

$$\begin{aligned} \lim _{k\rightarrow \infty }\left( g^{\ell _k}_{\mathbf {a}_{\ell _k}}\right) _\alpha \,=\, \lim _{k\rightarrow \infty }p_\alpha \left( \mathbf {a}_{\ell _k},\mathbf {g}^{\ell _k}\right) , \,=\,p_\alpha (\mathbf {a}^*,\mathbf {g}^*),\quad \forall \alpha \in \mathbb {N}^n_d. \end{aligned}$$

This means that for every \(\alpha \in \mathbb {N}^n_d\),

$$\begin{aligned} \theta ^*(\mathbf {x})\,=\,g^*(\mathbf {x}-\mathbf {a}^*),\quad \mathbf {x}\in \mathbb {R}^n. \end{aligned}$$

In addition, as \(\mathbf {g}^{\ell _k}\rightarrow \mathbf {g}^*\) as \(k\rightarrow \infty \), one has the pointwise convergence \(g^{\ell _k}(\mathbf {x})\rightarrow g^*(\mathbf {x})\) for all \(\mathbf {x}\in \mathbb {R}^n\). Therefore, by Fatou’s Lemma (see e.g. Ash [3]),

$$\begin{aligned} \rho \,=\,\lim _{k\rightarrow \infty }\,\int _{\mathbb {R}^n}\exp (-g^{\ell _k})\,d\mathbf {x}\,\ge \, \int _{\mathbb {R}^n}\liminf _{k\rightarrow \infty }\exp (-g^{\ell _k})\,d\mathbf {x}\,=\,\int _{\mathbb {R}^n}\exp (-g^*)\,d\mathbf {x}, \end{aligned}$$

which proves that \((\mathbf {a}^*,g^*)\) is an optimal solution of (4.2).

In addition \(g^*\in \mathbf {P}[\mathbf {x}]_d\) is an optimal solution of the inner minimization problem in (7.8) with \(\mathbf {a}:=\mathbf {a}^*\). Otherwise an optimal solution \(h\in \mathbf {P}[\mathbf {x}]_d\) of (7.8) with \(\mathbf {a}=\mathbf {a}^*\) would yield a solution \((\mathbf {a}^*,h)\) with associated cost \(\int _{\mathbb {R}^n}\exp (-h)\) strictly smaller than \(\rho \), a contradiction.

Hence by Theorem 3.2 (applied to problem (7.8)) there is a finite Borel measure \(\mu ^*\in \mathcal {M}(\mathbf {K}-\mathbf {a}^*)_+\) such that

$$\begin{aligned} \int _{\mathbb {R}^n}\mathbf {x}^\alpha \exp (-g^*)d\mathbf {x}&= \int _{\mathbf {K}-\mathbf {a}^*} \mathbf {x}^\alpha \,d\mu ^*,\qquad \forall \vert \alpha \vert =d\\ \int _{\mathbf {K}-\mathbf {a}^*}(1-g^*)\,d\mu ^*&= 0;\quad \mu (\mathbf {K}-\mathbf {a}^*)=\frac{n}{d}\int _{\mathbb {R}^n}\exp (-g^*)\,d\mathbf {x}. \end{aligned}$$

And so \(\mu ^*\) is supported on the set

$$\begin{aligned} V=\{\,\mathbf {x}\in \mathbf {K}-\mathbf {a}^*:\,g^*(\mathbf {x})=1\}\,=\,\{\,\mathbf {x}\in \mathbf {K}: \,g^*(\mathbf {x}-\mathbf {a}^*)=1\,\}\,=\,\mathbf {K}\cap \mathbf {G}^{\mathbf {a}^*}_1. \end{aligned}$$

Invoking again [1, Theorem 2.1.1, p. 39], there exists an atomic measure \(\nu ^*\in \mathcal {M}(\mathbf {K}-\mathbf {a}^*)_+\) supported on at most \({n-1+d\atopwithdelims ()d}\) of \(\mathbf {K}-\mathbf {a}^*\) with same moments of order \(d\) as \(\mu ^*\).   \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lasserre, J.B. A generalization of Löwner-John’s ellipsoid theorem. Math. Program. 152, 559–591 (2015). https://doi.org/10.1007/s10107-014-0798-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-014-0798-5

Keywords

Mathematics Subject Classification

Navigation