Skip to main content

Advertisement

Log in

Positivity certificates and polynomial optimization on non-compact semialgebraic sets

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

Abstract

In a first contribution, we revisit two certificates of positivity on (possibly non-compact) basic semialgebraic sets due to Putinar and Vasilescu (C R Acad Sci Ser I Math 328(6):495–499, 1999). We use Jacobi’s technique from (Math Z 237(2):259–273, 2001) to provide an alternative proof with an effective degree bound on the sums of squares weights in such certificates. As a consequence, it allows one to define a hierarchy of semidefinite relaxations for a general polynomial optimization problem. Convergence of this hierarchy to a neighborhood of the optimal value as well as strong duality and analysis are guaranteed. In a second contribution, we introduce a new numerical method for solving systems of polynomial inequalities and equalities with possibly uncountably many solutions. As a bonus, one can apply this method to obtain approximate global optimizers in polynomial optimization.

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

References

  1. Artin, E.: Über die zerlegung definiter funktionen in quadrate. Abhandlungen aus dem mathematischen Seminar der Universität Hamburg 5(1), 100–115 (1927)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bajbar, T., Stein, O.: Coercive polynomials and their Newton polytopes. SIAM J. Optim. 25(3), 1542–1570 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  3. Berg, C.: The multidimensional moment problem and semigroups. Proc. Symp. Appl. Math. 37, 110–124 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  4. Blekherman, G.: There are significantly more nonegative polynomials than sums of squares. Isr. J. Math. 153(1), 355–380 (2006)

    Article  MATH  Google Scholar 

  5. Blekherman, G., Gouveia, J., Pfeiffer, J.: Sums of squares on the hypercube. Math. Z. 284(1–2), 41–54 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  6. Blekherman, G., Smith, G.G., Velasco, M.: Sharp degree bounds for sum-of-squares certificates on projective curves. J. Math. Pures Appl. 129, 61–86 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  7. Boyd, S., Vandenberghe, L.: Convex Optimization. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  8. Curto, R.E., Fialkow, L.A.: Truncated K-moment problems in several variables. J. Oper. Theory 54, 189–226 (2005)

    MathSciNet  MATH  Google Scholar 

  9. Delzell, C.N.: Continous, piecewise-polynomial functions which solve Hilbert’s 17th problem. J. Reine Angew. Math. 1993(440), 157–174 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  10. Demmel, J., Nie, J., Powers, V.: Representations of positive polynomials on noncompact semialgebraic sets via KKT ideals. J. Pure Appl. Algebra 209(1), 189–200 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  11. Dickinson, P.J., Povh, J.: On an extension of Pólya’s Positivstellensatz. J. Global Optim. 61(4), 615–625 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  12. Djukić, D., Janković, V., Matić, I., Petrović, N.: The IMO Compendium: A Collection of Problems Suggested for the International Mathematical Olympiads: 1959–2009, 2nd edn. Springer, Berlin (2011)

    MATH  Google Scholar 

  13. Gershwin, S.B.: KKT examples. min J. 10, 2 (2010)

    Google Scholar 

  14. Greuet, A., Safey El Din, M.: Probabilistic algorithm for polynomial optimization over a real algebraic set. SIAM J. Optim. 24(3), 1313–1343 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Henrion, D., Lasserre, J.B.: Detecting global optimality and extracting solutions in GloptiPoly. In: Henrion, D., Garulli, A. (eds.) Positive Polynomials in Control, pp. 293–310. Springer, Berlin (2005)

    Chapter  Google Scholar 

  16. Hilbert, D.: Über die Darstellung Definiter Formen als Summe von Formenquadraten. Math. Ann. 32(3), 342–350 (1888)

    Article  MathSciNet  MATH  Google Scholar 

  17. Hu, S., Li, G., Qi, L.: A tensor analogy of Yuan’s theorem of the alternative and polynomial optimization with sign structure. J. Optim. Theory Appl. 168(2), 446–474 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  18. Jacobi, T.: A representation theorem for certain partially ordered commutative rings. Math. Z. 237(2), 259–273 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  19. Jeyakumar, V., Kim, S., Lee, G.M., Li, G.: Solving global optimization problems with sparse polynomials and unbounded semialgebraic feasible sets. J. Glob. Optim 65(2), 175–190 (2016)

    Article  MATH  Google Scholar 

  20. Jeyakumar, V., Lasserre, J.B., Li, G.: On polynomial optimization over non-compact semi-algebraic sets. J. Optim. Theory Appl. 163(3), 707–718 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Jibetean, D., Laurent, M.: Semidefinite approximations for global unconstrained polynomial optimization. SIAM J. Optim. 16(2), 490–514 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  22. Josz, C., Henrion, D.: Strong duality in Lasserre’s hierarchy for polynomial optimization. Optim. Lett. 10(1), 3–10 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  23. Krivine, J.L.: Anneaux préordonnés. J. Anal. Mathé. 12(1), 307–326 (1964)

    Article  MATH  Google Scholar 

  24. Lasserre, J.B.: Global optimization with polynomials and the problem of moments. SIAM J. Optim. 11(3), 796–817 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  25. Lasserre, J.B.: A sum of squares approximation of nonnegative polynomials. SIAM Rev. 49(4), 651–669 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  26. Lasserre, J.B.: An Introduction to Polynomial and Semi-algebraic Optimization, vol. 52. Cambridge University Press, Cambridge (2015)

    Book  MATH  Google Scholar 

  27. Lasserre, J.B., Laurent, M., Rostalski, P.: Semidefinite characterization and computation of zero-dimensional real radical ideals. Found. Comput. Math. 8(5), 607–647 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  28. Lasserre, J.B., Magron, V.: In SDP relaxations, inaccurate solvers do robust optimization. SIAM J. Optim. 29(3), 2128–2145 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  29. Lasserre, J.B., Netzer, T.: SOS approximations of nonnegative polynomials via simple high degree perturbations. Math. Z. 256(1), 99–112 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  30. Laurent, M.: Revisiting two theorems of Curto and Fialkow on moment matrices. Proc. Am. Math. Soc. 133(10), 2965–2976 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  31. Laurent, M.: Sums of squares, moment matrices and optimization over polynomials. In: Putinar, M., Sullivant, S. (eds.) Emerging Applications of Algebraic Geometry, pp. 157–270. Springer, Berlin (2009)

    Chapter  Google Scholar 

  32. Leep, D., Starr, C.: Polynomials in R[X, Y] that are sums of squares in R(X, Y). Proc Am Math Soc 129(11), 3133–3141 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  33. Lombardi, H., Perrucci, D., Roy, M.F.: An elementary recursive bound for effective Positivstellensatz and Hilbert 17-th problem. arXiv preprint arXiv:1404.2338 (2014)

  34. Magron, V., Safey El Din, M.: On exact polya and Putinar’s representations. In: Proceedings of the 2018 ACM International Symposium on Symbolic and Algebraic Computation, pp. 279–286 (2018)

  35. Mai, N.H.A., Magron, V., Lasserre, J.B.: A sparse version of Reznick’s Positivstellensatz. arXiv preprint arXiv:2002.05101 (2020)

  36. Marshall, M.: Approximating positive polynomials using sums of squares. Can. Math. Bull. 46(3), 400–418 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  37. Navascués, M., García-Sáez, A., Acín, A., Pironio, S., Plenio, M.B.: A paradox in bosonic energy computations via semidefinite programming relaxations. New J. Phys. 15(2), 023026 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  38. Nie, J.: Discriminants and nonnegative polynomials. J. Symb. Comput. 47(2), 167–191 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  39. Nie, J.: Optimality conditions and finite convergence of Lasserre’s hierarchy. Math. Program. 146(1–2), 97–121 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  40. Nie, J.: Tight relaxations for polynomial optimization and Lagrange multiplier expressions. Math. Program. 178(1–2), 1–37 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  41. Nie, J., Demmel, J., Sturmfels, B.: Minimizing polynomials via sum of squares over the gradient ideal. Math. Program. 106(3), 587–606 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  42. Nie, J., Schweighofer, M.: On the complexity of Putinar’s positivstellensatz. J. Complex. 23(1), 135–150 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  43. Parrilo, P.A.: Structured semidefinite programs and semialgebraic geometry methods in robustness and optimization. Ph.D. thesis, California Institute of Technology (2000)

  44. Pólya, G.: Über Positive Darstellung von Polynomen. Vierteljschr. Naturforsch. Ges. Zürich 73, 141–145 (1928)

    MATH  Google Scholar 

  45. Putinar, M.: Positive polynomials on compact semi-algebraic sets. Indiana Univ. Math. J. 42(3), 969–984 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  46. Putinar, M., Vasilescu, F.H.: Positive polynomials on semi-algebraic sets. C. R. Acad. Sci. Ser. I Math. 328(7), 585–589 (1999)

    MathSciNet  MATH  Google Scholar 

  47. Putinar, M., Vasilescu, F.H.: Solving moment problems by dimensional extension. C. R. Acad. Sci. Ser. I Math. 328(6), 495–499 (1999)

    MathSciNet  MATH  Google Scholar 

  48. Quijorna, M.L.: An experimental approach for global polynomial optimization based on moments and semidefinite programming. arXiv preprint arXiv:1809.09043 (2018)

  49. Reznick, B.: Uniform denominators in Hilbert’s seventeenth problem. Math. Z. 220(1), 75–97 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  50. Reznick, B.: Some concrete aspects of Hilbert’s 17th problem. Contemp. Math. 253, 251–272 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  51. Reznick, B.: On the absence of uniform denominators in Hilbert’s 17th problem. Proc. Am. Math. Soc. 133(10), 2829–2834 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  52. Schabert, R.: Uniform denominators in Hilbert’s 17th problem theorems by Pólya and Reznick. Manuscript (2019). http://www.math.uni-konstanz.de/~infusino/FachseminarMA-SS19/Schabert-Handout.pdf

  53. Scheiderer, C.: A Positivstellensatz for projective real varieties. Manuscr. Math. 138(1–2), 73–88 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  54. Schmüdgen, K.: The K-moment problem for compact semi-algebraic sets. Math. Ann. 289(1), 203–206 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  55. Schweighofer, M.: Iterated rings of bounded elements and generalizations of Schmüdgen’s Positivstellensatz. J. Reine Angew. Math. 554, 19–45 (2003)

    MathSciNet  MATH  Google Scholar 

  56. Schweighofer, M.: Global optimization of polynomials using gradient tentacles and sums of squares. SIAM J. Optim. 17(3), 920–942 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  57. Slinko, A.M.: USSR Mathematical Olympiads, 1989–1992, vol. 11. Australian Mathematics Trust, Bruce (1997)

    Google Scholar 

  58. Stengle, G.: A Nullstellensatz and a Positivstellensatz in semialgebraic geometry. Math. Ann. 207(2), 87–97 (1974)

    Article  MathSciNet  MATH  Google Scholar 

  59. Trnovska, M.: Strong duality conditions in semidefinite programming. J. Electr. Eng. 56(12), 1–5 (2005)

    MATH  Google Scholar 

  60. Vui, H.H., Son, P.T.: Representations of positive polynomials and optimization on noncompact semialgebraic sets. SIAM J. Optim. 20(6), 3082–3103 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  61. Waki, H., Kim, S., Kojima, M., Muramatsu, M., Sugimoto, H.: Algorithm 883: SparsePOP—a sparse semidefinite programming relaxation of polynomial optimization problems. ACM Trans. Math. Softw. 35(2), 15 (2008)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The second author was supported by the Tremplin ERC Stg Grant ANR-18-ERC2-0004-01 (T-COPS project) and by the FMJH Program PGMO (EPICS project) (Grant number P-2019-0038) and EDF, Thales, Orange et Criteo. This work has benefited from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie Actions, grant agreement 813211 (POEMA) as well as from the AI Interdisciplinary Institute ANITI funding, through the French “Investing for the Future PIA3” program under the Grant agreement \(\hbox {n}^{\circ }\)ANR-19-PI3A-0004.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Victor Magron.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Proof of Theorem 5

Jacobi [18, Theorem 7] proves another result of Putinar and Vasilescu [47, Theorem 4.2], which states that if \(f,g_1,\dots ,g_m\) are polynomials of even degree and \(\tilde{f}>0\) on \(S(\{\tilde{g}_1,\dots ,\tilde{g}_m\})\backslash \{0\}\), then \(\theta ^kf\in Q(g)\) for k large enough, where \(\tilde{h}\) is the homogenization of given polynomial h. The idea of Jacobi is to apply Putinar’s Positivstellensatz for the fact that \(\tilde{f}>0\) on the intersection of \(S(\{\tilde{g}_1,\dots ,\tilde{g}_m\})\) with the unit sphere. Our proof for Theorem 5 replaces \(\tilde{f}\) here by the perturbation of \(\tilde{f}\) and replaces the unit sphere by a sphere with an arbitrary radius \(L^{1/2}\). This will have a direct implication on the complexity analysis derived in Proposition 1.

Proof

1. Let us prove the conclusion under condition (i). For every \(h \in {\mathbb {R}}[ x ]\), we define by \(\hat{h}\) the degree \(2d_1 ( h )\) homogenization of h, i.e.,

$$\begin{aligned} \hat{h}\left( {x,{x_{n + 1}}} \right) = x_{n + 1}^{2d_1 ( h )}h\left( {\displaystyle {x}/{{{x_{n + 1}}}}} \right) . \end{aligned}$$
(1.34)

where \({d_1}( h ): = {1 + \lfloor {{{\deg ( h )}}/{2}} \rfloor }\). Let \(\varepsilon > 0\) be fixed. Let \(\tilde{f}\) be the degree 2d homogenization of f. Then \(\tilde{f} + \varepsilon {\Vert {( {x,{x_{n + 1}}} )} \Vert _2^{2d}}\) is a homogeneous polynomial of degree 2d. Let \(\hat{g} := \{ {{\hat{g}_1},\dots ,{\hat{g}_m}} \}\) and \(L>0\). We will show that

$$\begin{aligned} \tilde{f} + \varepsilon {\Vert {\left( {x,{x_{n + 1}}} \right) }\Vert _2^{2d}} \ge \varepsilon L^{d}\text { on }S\left( \hat{g}, \left\{ L-\Vert {\left( {x,{x_{n + 1}}} \right) }\Vert _2^2\right\} \right) . \end{aligned}$$
(1.35)

Let \(( {y,{y_{n + 1}}}) \in S( \hat{g}, \{L-\Vert ( x,x_{n + 1} \Vert _2^2\})\) be fixed. By (1.34), one has

$$\begin{aligned} y_{n + 1}^{2d_1 \left( g_j \right) }{g_j}\left( {\displaystyle {y}/{{{y_{n + 1}}}}} \right) ={{\hat{g}}_j}\left( {y,{y_{n + 1}}} \right) \ge 0,\,j = 1,\dots ,m. \end{aligned}$$
(1.36)

We consider the following two cases:

  • Case 1: \({y_{n + 1}} \ne 0\). For \(j=1,\dots ,m\), by (1.36) and since \(y_{n + 1}^{2d_1 \left( g_j \right) }>0\), \({g_j}( {\displaystyle {y}/{{{y_{n + 1}}}}} ) \ge 0\). It implies that \(\displaystyle {y}/{{{y_{n + 1}}}} \in S\left( g\right) \). By assumption, \(f( \displaystyle {y}/{{{y_{n + 1}}}}) \ge 0\). From this, \(\tilde{f}( {y,{y_{n + 1}}}) = y_{n + 1}^{2d}f( {{y}/{{{y_{n + 1}}}}} ) \ge 0\). From this and since \(\Vert ( {y,{y_{n + 1}}} )\Vert _2^2=L\), \(\tilde{f}( {y,{y_{n + 1}}}) + \varepsilon {\Vert {( {y,{y_{n + 1}}} )} \Vert _2^{2d}} \ge \varepsilon L^{d}\).

  • Case 2: \({y_{n + 1}} = 0\). By definition of \(d_1(f)\), \(x_{n+1}\) divides \(\tilde{f}\) so \(\tilde{f}( {y,{y_{n + 1}}}) = 0\). From this and since \(\Vert ( {y,{y_{n + 1}}})\Vert _2^2=L\),

    $$\begin{aligned} \tilde{f}\left( {y,{y_{n + 1}}}\right) + \varepsilon {\Vert {\left( {y,{y_{n + 1}}} \right) } \Vert _2^{2d}} = \varepsilon L^{d}. \end{aligned}$$

Thus, (1.35) holds. It is not hard to show that \(S( \hat{g} \cup \{L-\Vert {( {x,{x_{n + 1}}} )}\Vert _2^2\})\) satisfies the Archimedean condition. From this and by applying Theorem 2 (i),

$$\begin{aligned} \tilde{f} + \varepsilon {\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert ^{2d}} \in Q\left( \hat{g},{\left\{ { L - {{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert }^2_2}} \right\} } \right) . \end{aligned}$$

Then there exist \(\psi _{j} \in \varSigma [ {x,{x_{n + 1}}} ],\ j = 0,1,\dots ,m\) and \(\varphi \in {\mathbb {R}}[ {x,{x_{n + 1}}} ]\) such that

$$\begin{aligned} \tilde{f} + \varepsilon {\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert _2^{2d}} = \psi _0 + \sum \limits _{j = 1}^m {\psi _j{{\hat{g}}_j}} + \left( {L - {{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert }_2^2}} \right) \varphi . \end{aligned}$$

Let \(( {z,{z_{n + 1}}}) \in {{\mathbb {R}}^{n + 1}}\backslash \{ 0\}\). By replacing \({( {x,{x_{n + 1}}} )}\) in the last equality by \(L^{1/2}\frac{{( {z,{z_{n + 1}}} )}}{{\Vert {( {z,{z_{n + 1}}} )} \Vert _2}}\) and the fact that \({\tilde{f} + \varepsilon {\Vert {( {x,{x_{n + 1}}})} \Vert _2^{2d}}}, \ \hat{g}_1,\dots ,\hat{g}_m\) are homogeneous polynomials of degree \(2d,\ 2d_1 \left( g_1 \right) ,\dots ,2d_1 \left( g_m\right) \) respectively, one has

$$\begin{aligned}&\left( {\tilde{f}\left( {z,{z_{n + 1}}} \right) + \varepsilon {{\Vert {\left( {z,{z_{n+ 1}}} \right) } \Vert }_2^{2d}}} \right) L^d{\Vert {\left( {z,{z_{n + 1}}} \right) } \Vert _2^{ -2d}}\\&\quad = \psi _0 \left( L^{1/2}\frac{\left( z,z_{n + 1}\right) }{\Vert \left( z,z_{n + 1} \right) \Vert _2} \right) \\&\qquad + \sum \limits _{j = 1}^m {\psi _j \left( L^{1/2}\frac{\left( z,z_{n + 1} \right) }{\Vert \left( z,z_{n + 1}\right) \Vert } \right) \hat{g}_j\left( z,z_{n + 1} \right) L^{d_1(g_j)}\Vert \left( z,z_{n + 1} \right) \Vert _2^{ - 2d_1 ( g_j )}} . \end{aligned}$$

Set

$$\begin{aligned} K: = \max \left\{ {2d,\deg (\psi _0 ),2d_1\left( g_1\right) + \deg \left( \psi _1\right) ,\dots ,2d_1\left( g_m\right) + \deg \left( \psi _m\right) ,2+2\lceil \deg (\varphi )/2\rceil }\right\} . \end{aligned}$$

Obviously, K is even. After multiplying the two sides of the last equality with \(\Vert ( z,z_{n + 1}) \Vert ^K\), one has

$$\begin{aligned}&\left( {\tilde{f}\left( {z,{z_{n + 1}}} \right) + \varepsilon {{\Vert {\left( {z,{z_{n + 1}}} \right) } \Vert }_2^{2d}}} \right) L^d{\Vert {\left( {z,{z_{n + 1}}} \right) } \Vert _2^{K -2d}}\\&\qquad = \bar{\psi }_0\left( z,z_{n + 1} \right) + \sum \limits _{j = 1}^m {L^{d_1\left( g_j\right) } \bar{\psi }_j \left( z,z_{n + 1} \right) \hat{g}_j\left( z,z_{n + 1} \right) } , \end{aligned}$$

where

$$\begin{aligned} \bar{\psi }_0&:=\psi _0 \left( L^{1/2}\frac{( x,x_{n + 1})}{\Vert ( x,x_{n + 1} ) \Vert _2} \right) \Vert ( x,x_{n + 1} ) \Vert _2^{ K}\, ,\\ \bar{\psi }_j&:=\psi _j \left( L^{1/2}\frac{( x,x_{n + 1} )}{\Vert ( x,x_{n + 1}) \Vert _2} \right) \Vert ( x,x_{n + 1} ) \Vert _2^{ K- 2d_1 ( g_j )},\,j=1,\dots ,m. \end{aligned}$$

Since \(( {z,{z_{n + 1}}}) \in {{\mathbb {R}}^{n + 1}}\backslash \{ 0\}\) is arbitrary,

$$\begin{aligned} \left( {\tilde{f} + \varepsilon {{\Vert {( {x,{x_{n + 1}}} )} \Vert }_2^{2d}}} \right) L^d{\Vert {( {x,{x_{n + 1}}} )} \Vert _2^{K - 2d}}= \bar{\psi }_0 + \sum \limits _{j = 1}^m {L^{d_1(g_j)}\bar{\psi }_j \hat{g}_j} . \end{aligned}$$
(1.37)

Let \(j\in \{0,\dots ,m\}\) be fixed and set \(g_0:=1\) and \(d_1(g_0):=0\). We will show that

$$\begin{aligned} \bar{\psi }_j=s_j+\xi _j \Vert (x,x_{n+1})\Vert _2, \end{aligned}$$
(1.38)

for some \(s_j\in \varSigma [x,x_{n+1}]_{K/2-d_1(g_j)}\) and \(\xi _j\in {\mathbb {R}}[x,x_{n+1}]\). By definition of K (\(K- 2d_1 ( g_j )\ge \deg (\psi _j)\), \(j=1,\dots ,m\)) and since \(\psi _j\) is SOS, \(\bar{\psi }_j=\sum \nolimits _{t = 1}^{l} {\phi _{t}^2}\) with \(\phi _{t}:=h_t+p_t\Vert (x,x_{n+1})\Vert _2\) for some \(h_t\in {\mathbb {R}}[x,x_n]_{K/2-d_1(g_j)}\) and \(p_t\in {\mathbb {R}}[x,x_n]_{K/2-d_1(g_j)-1}\). Then

$$\begin{aligned} \bar{\psi }_j=\sum \limits _{t = 1}^{l} {\left( h_t^2+p_t^2\Vert \left( x,x_{n+1}\right) \Vert _2^2\right) }+2\Vert \left( x,x_{n+1}\right) \Vert _2\sum \limits _{t = 1}^{l} {h_tp_t}. \end{aligned}$$

By setting \(s_j:=\sum \nolimits _{t = 1}^{l} {(h_t^2+p_t^2\Vert (x,x_{n+1})\Vert _2^2)}\) and \(\xi _j:=2\sum \nolimits _{t = 1}^{l} {h_tp_t}\), (1.38) follows. By (1.37) and (1.38),

$$\begin{aligned}&\left( {\tilde{f} + \varepsilon {{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert }_2^{2d}}} \right) L^d{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert _2^{K - 2d}}\\&\quad = s_0 + \sum \limits _{j = 1}^m {L^{d_1\left( g_j\right) }s_j \hat{g}_j}+\Vert \left( x,x_{n + 1} \right) \Vert _2\sum \limits _{j = 1}^m {L^{d_1\left( g_j\right) }\xi _j \hat{g}_j} . \end{aligned}$$

Note that \(\Vert {( {x,{x_{n + 1}}} )} \Vert _2\) is not a polynomial. From the last equality and since the right hand side is polynomial, the left hand side must be a polynomial, so \(\sum \nolimits _{j = 1}^m {L^{d_1(g_j)}\xi _j \hat{g}_j}=0\). Thus,

$$\begin{aligned} \left( {\tilde{f} + \varepsilon {{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert }_2^{2d}}} \right) L^d{\Vert {\left( {x,{x_{n + 1}}} \right) } \Vert _2^{K - 2d}}= s_0 + \sum \limits _{j = 1}^m {L^{d_1(g_j)}s_j \hat{g}_j} . \end{aligned}$$

By setting \(k_\varepsilon :=K/2-d\), \({s_0} \in \varSigma {[ x,x_{n+1} ]_{{k_\varepsilon } + d}}\), \({s _j} \in \varSigma {[ x,x_{n+1} ]_{{k_\varepsilon } + d - d_1( {{g_j}})}},\,j = 1,\dots ,m\) and \(L^d{\Vert {( {x,{x_{n + 1}}})} \Vert _2^{2{k_\varepsilon }}}( {\tilde{f} + \varepsilon {{\Vert {( {x,{x_{n + 1}}} )} \Vert }_2^{2d}}} ) = {s _0} + \sum \nolimits _{j = 1}^m {L^{d_1(g_j)}{s_j}{{\hat{g}}_j}} \). By replacing \(x_{n+1}\) by 1, \( L^d{\theta ^{{k_\varepsilon }}}( {f + \varepsilon {\theta ^{d}}} ) = {s _0}( {x,1} ) + \sum \nolimits _{j = 1}^m {L^{d_1(g_j)}{s_j}( {x,1} ){g_j}} \). Since \({s_0} \in \varSigma {[ x,x_{n+1}]_{{k_\varepsilon } + d}}\), \({s_j} \in \varSigma {[ x,x_{n+1} ]_{{k_\varepsilon } + d - d_1( {{g_j}} )}},\,j = 1,\dots ,m\), \({s _0}( {x,1} ) \in \varSigma {[ x ]_{{k_\varepsilon } + d}}\), \({s _j}( {x,1}) \in \varSigma {[ x]_{{k_\varepsilon } + d - d_1( {{g_j}})}},\,j = 1,\dots ,m\). Note that \(d_1(g_j)\ge d_2(g_j)=u_j\), \(j=1,\dots ,m\). Hence, \({\theta ^{{k_\varepsilon }}}( {f + \varepsilon {\theta ^{d}}} ) \in {Q_{{k_\varepsilon } + d}}( g )\) by the definition of truncated quadratic module.

2. Let us show the conclusion under condition (ii). We do a similar process as part 1 (under condition (i)). The difference is that \(\hat{h}\) here is defined as the degree \(2d_2 ( h )\) (instead of \(2d_1 ( h )\)) homogenization of \(h \in {\mathbb {R}}[ x ]\) and the proof for (1.35). To show (1.35) in Case 2: \(y_{n+1}=0\) here, we rely on the constraint \(g_m=f+\lambda \) for some \(\lambda \ge 0\) (instead of \(d \ge d_1(f)\) and that \(x_{n+1}\) divides \(\tilde{f}\)). More explicitly, we assume by contradiction that

$$\begin{aligned} \tilde{f}({y,{y_{n + 1}}}) + \varepsilon {\Vert {( {y,{y_{n + 1}}} )}\Vert _2^{2d}} <\varepsilon L^{d}. \end{aligned}$$

From this and since \(0 \le {\hat{g}_m}(y,{y_{n + 1}}) = \hat{f}(y,0)\),

$$\begin{aligned} \begin{array}{rl} \varepsilon \Vert ({y,{y_{n + 1}}})\Vert _2^{2d} &{}\le 0^{2d-2d_2(f)}\hat{f}(y,0) + \varepsilon \Vert ({y,{y_{n + 1}}})\Vert _2^{2d}\\ &{}= y_{n+1}^{2d-2d_2(f)}\hat{f}(y,y_{n+1}) + \varepsilon \Vert ({y,{y_{n + 1}}})\Vert _2^{2d} \\ &{}= \tilde{f}({y,{y_{n + 1}}}) + \varepsilon {\Vert {( {y,{y_{n + 1}}} )}\Vert _2^{2d}} <\varepsilon L^{d}. \end{array} \end{aligned}$$

It follows that \(\Vert ({y,{y_{n + 1}}})\Vert _2^2<L=\Vert ({y,{y_{n + 1}}})\Vert _2^2\). It is impossible. Thus, \(\tilde{f}({y,{y_{n + 1}}}) + \varepsilon {\Vert {( {y,{y_{n + 1}}} )}\Vert _2^{2d}} \ge \varepsilon L^{d}\). \(\square \)

1.2 Proof of Proposition 1

Proof

We keep all the notation from the proof of Theorem 5. Without loss of generality, let us assume that \(L=1/4\). By (1.35), \(S( \hat{g},{\{ { L - {{\Vert {( {x,{x_{n + 1}}} )} \Vert }^2_2}} \}} )\subset (-1,1)^{n+1}\) and

$$\begin{aligned} \min \{\tilde{f} + \varepsilon {\Vert {( {x,{x_{n + 1}}} )} \Vert _2^{2d}}\,:( {x,{x_{n + 1}}} )\in S( \hat{g},{\{ { L - {{\Vert {( {x,{x_{n + 1}}} )} \Vert }^2_2}} \}} )\}\ge 2^{-2d}\varepsilon .\qquad \end{aligned}$$
(1.39)

By definition of \(\hat{g}\) and since \(0_{{\mathbb {R}}^n}\in S(g)\), \((0_{{\mathbb {R}}^n},L^{1/2})\in S( \hat{g},{\{ { L - {{\Vert {( {x,{x_{n + 1}}} )} \Vert }^2_2}} \}} )\). Thus, \(S( \hat{g},{\{ { L - {{\Vert {( {x,{x_{n + 1}}} )} \Vert }^2_2}} \}} )\) is nonempty. By definition of K,

$$\begin{aligned} \tilde{f} + \varepsilon {\Vert {( {x,{x_{n + 1}}} )} \Vert _2^{2d}} \in Q_{K/2}( \hat{g},{\{ { L - {{\Vert {( {x,{x_{n + 1}}} )} \Vert }^2_2}} \}} ). \end{aligned}$$

Since \(\Vert {(x,{x_{n + 1}})} \Vert _2^{2d} = \sum \limits _{\bar{\alpha }\in {\mathbb {N}}_d^{n + 1}} {c_{n+1}(\bar{\alpha })(x,{x_{n + 1}})^{2\bar{\alpha }}} \), one has

$$\begin{aligned} \begin{array}{rl} \Vert {\tilde{f} + \varepsilon \Vert {(x,{x_{n + 1}})} \Vert _2^{2d}} \Vert _{\max }&{}\le \Vert {\tilde{f} } \Vert _{\max } +\varepsilon \Vert {\Vert {(x,{x_{n + 1}})} \Vert _2^{2d}} \Vert _{\max }\\ &{}=\Vert {\tilde{f} } \Vert _{\max } +\varepsilon \max \{ \displaystyle {c_{n+1}(\bar{\alpha })}/{c_{n+1}(2\bar{\alpha })}\,:{\bar{\alpha }\in {\mathbb {N}}^{n+1}_d}\}. \end{array} \end{aligned}$$

Note that \(\tilde{f} = \sum \nolimits _\alpha {f_\alpha x^\alpha x_{n+1}^{2d-|\alpha |}}\), so \(\Vert \tilde{f} \Vert _{\max }=\Vert f \Vert _{\max ,d}\). Thus,

$$\begin{aligned} \Vert {\tilde{f} + \varepsilon \Vert {(x,{x_{n + 1}})} \Vert _2^{2d}} \Vert _{\max }\le \Vert { f } \Vert _{\max , d} +\varepsilon \max \{ \displaystyle {c_{n+1}(\bar{\alpha })}/{c_{n+1}(2\bar{\alpha })}\,:{\bar{\alpha }\in {\mathbb {N}}^{n+1}_d}\}. \end{aligned}$$
(1.40)

From these and using Theorem 3, one can choose

$$\begin{aligned} \frac{K}{2}\ge C\exp \left( \left( 4^{d+1}d^2(n + 1)^{2d}\left( \varepsilon ^{-1} \Vert f \Vert _{\max , d}+ \max \left\{ \displaystyle \frac{c_{n+1}(\bar{\alpha })}{c_{n+1}(2\bar{\alpha })}\,:{\bar{\alpha }\in {\mathbb {N}}^{n+1}_d}\right\} \right) \right) ^C \right) , \end{aligned}$$

for some \(C>0\). The right hand side of this inequality comes from (1.39), (1.40) and the fact that the function \(t \mapsto c\exp \left( {b{t^c}} \right) \) with positive constants b and c is increasing on \([0,\infty )\). By setting \(k_\varepsilon =K/2-d\), the conclusion follows. \(\square \)

1.3 Proof of Proposition 3

Proof

We denote by P and \(P^\star \) feasible set and optimal solutions set for the moment SDP (4.22), respectively. We claim that P is nonempty. Indeed, with z being the moment sequence of the 1-atomic measure \(\theta (\bar{x})^{-1}\delta _{\bar{x}}\) for some \(\bar{x}\in S(g)\), it is not hard to check that the truncation of z is a feasible solution of (4.22). By setting \(C: = {L_z}({\theta ^k}(f + \varepsilon {\theta ^d}))\), \(P^\star \) is also the set of optimal solutions for

$$\begin{aligned} \begin{array}{rl} {\tau _k^2}(\varepsilon ) = \inf &{}{L_y}( {{\theta ^k}( {f + \varepsilon {\theta ^d}} )} )\\ \text {s.t.}&{}y = {(y_\alpha )_{\alpha \in {\mathbb {N}}^n_{2( {d + k})}}} \subset {\mathbb {R}},\\ &{}{L_y}( {{\theta ^k}( {f + \varepsilon {\theta ^d}} )} )\le C,\\ &{}{M_{k + d - {u_j}}}( {{g_j}y}) \succeq 0 ,\;j = 0,\dots ,m,\\ &{}{L_y}( {{\theta ^k}}) = 1, \end{array} \end{aligned}$$
(1.41)

with \(g_0=1\). Denote by \(\bar{P}\) the feasible set of (1.41). Note that \(\bar{P}\) is nonempty since the truncation of z is also a feasible solution of (1.41). We will prove that \(\bar{P}\) is bounded. By definition of \(\theta \), \({\theta ^r} = \sum \nolimits _{\alpha \in {\mathbb {N}}_r^n} {{C_\alpha ^r }{x^{2\alpha }}}\) for all \(r\in {\mathbb {N}}\), where \(C_\alpha ^r\ge 1\) for all \(\alpha \in {\mathbb {N}}_r^n\). Since \(f-f^\star \ge 0\) on S(g), by Theorem 5, there exists \(K\in {\mathbb {N}}\) such that

$$\begin{aligned} {\theta ^K}\left( {f - f^\star + \displaystyle \frac{\varepsilon }{2} {\theta ^d}} \right) \in Q_{K + d}{( g )}. \end{aligned}$$

Note that K depends only on f, g and \(\varepsilon \). Assume that \(k\ge K\). Since \(\theta ^{k-K} \in \varSigma {[ x ]_{k-K}}\), one has

$$\begin{aligned} {\theta ^k}\left( {f - {f^\star }+ \displaystyle \frac{\varepsilon }{2} {\theta ^d} }\right) = \theta ^{k - K} {\theta ^{K}} \left( {f - {f^\star }+ \displaystyle \frac{\varepsilon }{2} {\theta ^d}} \right) \in Q _{k + d}{( g)}. \end{aligned}$$

Then there exist \(G_j\succeq 0\), \(j=1,\dots ,m\) such that

$$\begin{aligned} {\theta ^k}\left( f - {f^ \star } + \frac{\varepsilon }{2}{\theta ^d}\right) = \sum \limits _{j = 0}^m {v_{k + d - {u_j}}^T{G_j}{v_{k + d - {u_j}}}{g_j}} = \sum \limits _{j = 0}^m {{\hbox {trace}}({G_j}{v_{d + k - {u_j}}}v_{d + k - {u_j}}^Tg_j)} . \end{aligned}$$

Let \(y\in \bar{P}\). From these and since \({M_{k + d - {u_j}}}( {{g_j}y})\succeq 0\),

$$\begin{aligned} {L_y}\left( {{\theta ^k}\left( {f - {f^ \star } + \frac{\varepsilon }{2}{\theta ^d}} \right) } \right) = \sum \limits _{j = 0}^m {{\hbox {trace}}({G_j}{M_{d + k - {u_j}}}({g_j}y))} \ge 0. \end{aligned}$$

Thus, for every \(\beta \in {\mathbb {N}}_{k + d}^n\),

$$\begin{aligned} \begin{array}{rl} \displaystyle \frac{\varepsilon }{2} {y_{2\beta }} &{}\le \displaystyle \frac{\varepsilon }{2} C_\beta ^{k + d}{y_{2\beta }} \le \displaystyle \frac{\varepsilon }{2} \sum \limits _{\alpha \in {\mathbb {N}}_{k + d}^n} {C_\alpha ^{k + d}{y_{2\alpha }}} = \displaystyle \frac{\varepsilon }{2} \sum \limits _{\alpha \in {\mathbb {N}}_{k + d}^n} {C_\alpha ^{k + d}{L_y}({x^{2\alpha }})} \\ &{} = \displaystyle \frac{\varepsilon }{2} {L_y}({\theta ^{k + d}}) \le \displaystyle \frac{\varepsilon }{2} {L_y}({\theta ^{k + d}}) + {L_y}\left( {\theta ^k}\left( f - f^\star +\displaystyle \frac{\varepsilon }{2} \theta ^d\right) \right) \\ &{}= {L_y}({\theta ^k}(f + \varepsilon {\theta ^d})) - f^\star {L_y}({\theta ^k})\le C - f^\star , \end{array}\ \end{aligned}$$

since every element \(y_{2\alpha }\) on the diagonal of the positive semidefinite matrix \({M_{k + d}}( {y})\) is nonnegative. Thus, \({y_{2\beta }}\le 2(C - f^\star )\varepsilon ^{-1}\) for every \(\beta \in {\mathbb {N}}_{k + d}^n\). Since \({M_{k + d }}( {y})\succeq 0\), \(| {{y_{\alpha + \beta }}} | \le {y_{2\beta }} \le 2(C - \underline{f})\varepsilon ^{-1}\) for all \(\alpha ,\beta \in {\mathbb {N}}_{k + d}^n\). This implies that \(\Vert y\Vert _2\) is bounded by \(2(C - \underline{f})\varepsilon ^{-1}\sqrt{s(2(d + k))}\). Since the objective function of (1.41) is linear and the feasible set of (1.41) is closed and bounded, the set \(P^\star \) of optimal solutions of (1.41) is nonempty and bounded. By using Trnovska’s result [59, Corrollary 1], \(\rho _k^2(\varepsilon )=\tau _k^2(\varepsilon )\), yielding the desired conclusion. \(\square \)

1.4 Proof of Proposition 4

Proof

We do a similar process as the proof of Proposition 3. The difference is the bound of y here obtained by the inequality constraint \(g_m=f-\underline{f}\ge 0\) in stead of Positivstellensatz. Thus, we do not need k sufficient large here. More explicitly, one has \(\Vert y\Vert _2 \le (C - \underline{f})\varepsilon ^{-1}\sqrt{s(2(d + k))}\), since for every \(\beta \in {\mathbb {N}}_{k + d}^n\),

$$\begin{aligned} \begin{array}{rl} \varepsilon {y_{2\beta }} &{}\le \varepsilon C_\beta ^{k + d}{y_{2\beta }} \le \varepsilon \sum \limits _{\alpha \in {\mathbb {N}}_{k + d}^n} {C_\alpha ^{k + d}{y_{2\alpha }}} = \varepsilon \sum \limits _{\alpha \in {\mathbb {N}}_{k + d}^n} {C_\alpha ^{k + d}{L_y}({x^{2\alpha }})} \\ &{}\le \varepsilon \sum \limits _{\alpha \in {\mathbb {N}}_{k + d}^n} {C_\alpha ^{k + d}{L_y}({x^{2\alpha }})} + \sum \limits _{\alpha \in {\mathbb {N}}_k^n} {C_\alpha ^k{L_y}({x^{2\alpha }}{g_m})} \\ &{} = \varepsilon {L_y}({\theta ^{k + d}}) + {L_y}({\theta ^k}(f - \underline{f} )) = {L_y}({\theta ^k}(f + \varepsilon {\theta ^d})) - \underline{f} {L_y}({\theta ^k})\\ &{}\le C - \underline{f}. \end{array} \end{aligned}$$

1.5 Proof of Lemma 4

Proof

Let us show that \((\xi _t)_{t=0,\dots ,n}\) is real sequence. Obviously, \(\xi _t\ge 0\), \(t=0,\dots ,n\). Set \(r_t=\xi _t^{1/2}\), \(t=0,\dots ,n\). Then

$$\begin{aligned} \left\{ \begin{array}{l} {r_0} = d( {{a_0},S( g,h )} )\ge 0,\\ {r_t} = d( {{a_t},S( g,h ) \cap \partial {B( {{a_0},{r_0}} )} \cap \dots \cap \partial {B( {{a_{t - 1}},{r_{t - 1}}} )} } )\ge 0,\, t = 1,\dots ,n, \end{array} \right. \end{aligned}$$
(1.42)

where \(d(a,A):=\inf \{\Vert x-a\Vert _2:\, x\in A\}\) for \(a\in {\mathbb {R}}^n\) and \(A\subset {\mathbb {R}}^n\). It is sufficient to prove that \(r_t\) is real, \(t=0,\dots ,n\). It is easy to see that S(gh) is closed and S(gh) is nonempty by assumption. From this, \({r_0}\) is nonnegative real and \(S( g,h ) \cap \partial {B( {{a_0},{r_0}} )} \) is also closed and nonempty. It implies that \(r_1\) is nonnegative real and

$$\begin{aligned} {S( g,h) \cap \partial {B( {{a_0},{r_0}} )} \cap \partial {B( {{a_{1}},{r_{1}}} )} } \end{aligned}$$

is also closed and nonempty. By induction, for \(t \in \{ {0,\dots ,n} \}\), \(r_t\) is nonnegative real and

$$\begin{aligned} {S( g,h ) \cap \partial {B( {{a_0},{r_0}} )} \cap \dots \cap \partial {B( {{a_{t}},{r_{t}}} )} } \end{aligned}$$

is closed and nonempty. Thus,

$$\begin{aligned}&S\left( {g,h \cup \{ {{\xi _t} - {{ \Vert {x - {a_t}} \Vert }_2^2}:\, t = 0,\dots ,n} \}} \right) \\&\qquad =S( g,h ) \cap \partial {B( {{a_0},{r_0}})} \cap \dots \cap \partial {B\left( {{a_n},{r_n}} \right) } \ne \emptyset . \end{aligned}$$

Let \({x^\star } \in S( {g,h \cup \{ {{\xi _t} - {{ \Vert {x - {a_t}} \Vert }_2^2}:\, t = 0,\dots ,n} \}} )\). Then \({x^\star } \in \partial B\left( {{a_0},{r_0}} \right) \cap \dots \cap \partial B( {{a_n},{r_n}} )\). It follows that \({\Vert {{x^\star } - {a_t}} \Vert _2^2} = r_t^2=\xi _t,\, t = 0,\dots ,n\). For \(t = 1,\dots ,n\),

$$\begin{aligned} \xi _t - \xi _0 = {\Vert {{x^\star } - {a_t}} \Vert _2^2} - {\Vert {{x^\star } - {a_0}} \Vert _2^2} = -2{( {{a_t} - {a_0}} )^T}{x^\star } + {\Vert {{a_t}} \Vert _2^2} - {\Vert {{a_0}} \Vert _2^2}. \end{aligned}$$

It implies (4.31). Denote

$$\begin{aligned} A = \left( {\begin{array}{*{20}{c}} {{{( {{a_1} - {a_0}} )}^T}}\\ {\dots }\\ {{{( {{a_n} - {a_0}} )}^T}} \end{array}} \right) \qquad \text { and }\qquad b = -\frac{1}{2}\left( {\begin{array}{*{20}{c}} {\xi _1 - \xi _0 - {{\Vert {{a_1}} \Vert }_2^2} + {{\Vert {{a_0}} \Vert }_2^2}}\\ {\dots }\\ {\xi _n - \xi _0 - {{\Vert {{a_n}} \Vert }_2^2} + {{\Vert {{a_0}} \Vert }_2^2}} \end{array}} \right) . \end{aligned}$$

The system (4.31) can be rewritten as \(Ax^\star =b\). Since \({a_j} - {a_0},\, j = 1,\dots ,n\) are linearly independent in \({\mathbb {R}}^n\), A is invertible. Hence, \(x^\star \) is determined uniquely by \(x^\star =A^{-1}b\). \(\square \)

1.6 Proof of Theorem 8

Proof

We will prove by induction that for \(t \in \left\{ {0,\dots ,n} \right\} \),

$$\begin{aligned} \exists {K_t} \in {\mathbb {N}}:\,\forall k \ge {K_t},\,\forall j \in \{ {0,\dots ,t} \},\, \eta _k^j = {\xi _j}. \end{aligned}$$
(1.43)

For \(t=0\), (4.33) is the SDP relaxation of order \(k+w\) of

$$\begin{aligned} {\xi _0} = \min \left\{ {{{\Vert {x - {a_0}} \Vert }_2^2}:\, x \in S\left( {g\cup \{ {L - {{\Vert {x } \Vert }_2^2}} \},h} \right) } \right\} . \end{aligned}$$

By assumption, \({( {\eta _k^0} )_{k \in {\mathbb {N}}}}\) finitely converges to \(\xi _0\), i.e. there exist \(K_0\in {\mathbb {N}}\) such that \(\eta _k^0 = {\xi _0}\) for all \(k\ge K_0\). It follows that (1.43) is true for \(t=0\). Assume that (1.43) is true for \(t=T\), i.e.

$$\begin{aligned} \exists {K_T} \in {\mathbb {N}}:\,\forall k \ge {K_T},\,\forall j \in \{ {0,\dots ,T} \},\, \eta _k^j = {\xi _j}. \end{aligned}$$
(1.44)

We will show that (1.43) is true for \(t=T+1\). By (4.33) and (1.44), for all \(k\ge K_T\),

$$\begin{aligned} \begin{array}{rl} \eta _k^{T+1} = \mathop {\inf }\limits _{y \in {{\mathbb {R}}^{s\left( {2(k+w)} \right) }}}&{} {L_y}\left( {{{\Vert {x - {a_{T+1}}} \Vert }_2^2}} \right) \\ \text {s.t. }&{}{M_{k+w - {u_j}}}\left( {{g_j}y} \right) \succeq 0,\\ &{}{M_{k+w - {w_q}}}\left( {h_qy} \right) = 0,\\ &{}{M_{k+w - 1}}\left( {\left( {{\xi _ j} - {{\Vert {x - {a_j}} \Vert }_2^2}}\right) y}\right) = 0,\, j = 0,\dots ,T,\\ &{}{y_0} = 1 \end{array} \end{aligned}$$

is the SDP relaxation of order \(k+w\) of problem

$$\begin{aligned} {\xi _{T+1}} = \min \left\{ {{{\Vert {x - {a_t}} \Vert }_2^2}:\, x \in S\left( {g,h \cup \left\{ {{\xi _j} - {{\Vert {x- {a_j}} \Vert }_2^2},\, j = 0,\dots ,T} \right\} } \right) } \right\} . \end{aligned}$$

By assumption, \({( {\eta _k^{T+1}})_{k \ge K_T}}\) finitely converges to \(\xi _{T+1}\), i.e. there exist \(K_{T+1}\ge K_T\) such that \(\eta _k^{T+1} = {\xi _{T+1}}\) for all \(k\ge K_{T+1}\). It follows that (1.44) is true for \(t=T+1\). Thus, (1.44) is true for \(t=0,\dots ,n\). For \(t=n\), there exists \(K=K_n\in {\mathbb {N}}\) such that for all \(k\ge K\), \(\eta _k^t = {\xi _t}\), \(t = 0,\dots ,n\). Let \(k\ge K\) be fixed. Let y be the solution of problem

$$\begin{aligned} \begin{array}{rl} \eta _k^n: = \mathop {\inf }\limits _{y \in {{\mathbb {R}}^{s\left( {2(k+w)}\right) }}}&{} {L_y}\left( {{{\Vert {x - {a_n}} \Vert }_2^2}} \right) \\ \text {s.t. }&{}{M_{k+w - {u_j}}}\left( {{g_j}y} \right) \succeq 0,\\ &{}{M_{k+w - {w_q}}}\left( {h_qy}\right) = 0,\\ &{}{M_{k+w - 1}}\left( {\left( {{\xi _ j} - {{\Vert {x - {a_j}} \Vert }_2^2}} \right) y} \right) \succeq 0,\, j = 0,\dots ,n-1,\\ &{}{y_0} = 1, \end{array} \end{aligned}$$

which is the SDP hierarchy relaxation of order \(k+w\) of problem

$$\begin{aligned} {\xi _{n}} = \min \left\{ {{{\Vert {x - {a_n}} \Vert }_2^2}:\, x \in S\left( {g,h \cup \left\{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\, j = 0,\dots ,n-1} \right\} } \right) } \right\} . \end{aligned}$$

We will prove that this latter problem has a unique minimizer \(x^\star \). Set \(\hat{h}:=h\cup \{ \xi _0-\Vert x-a_0\Vert _2^2\}\). Then

$$\begin{aligned} \left\{ \begin{array}{l} {\xi _0} = \min \{ {{{\Vert {x - {a_0}} \Vert }_2^2}:\, x \in S( g,\hat{h})} \},\\ {\xi _t} = \min \{ {{{\Vert {x - {a_t}} \Vert }_2^2}:\, x \in S( { g,\hat{h} \cup \{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\, j = 0,\dots ,t - 1} \}} )} \},\\ \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \ \ \ \ \ \ t = 1,\dots ,n. \end{array} \right. \ \end{aligned}$$

By Lemma 4 (with \(S(g,h):=S(g,\hat{h})\)), there exists \(x^\star \) such that

$$\begin{aligned}&S\left( {g,h \cup \left\{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\,j = 0,\dots ,n } \right\} } \right) \nonumber \\&\quad = S\left( { g, \hat{h} \cup \left\{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\,j = 0,\dots ,n }\right\} } \right) = \left\{ {{x^\star }} \right\} . \end{aligned}$$
(1.45)

Let a be a minimizer of the above POP with value \(\xi _n\). Then

$$\begin{aligned} a\in S( {g,h \cup \{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\, j = 0,\dots ,n-1} \}} ), \end{aligned}$$

and \({\Vert {a - {a_n}} \Vert }_2^2=\xi _n\). It follows that

$$\begin{aligned} a\in S( {g,h \cup \{ {{\xi _j} - {{\Vert {x - {a_j}} \Vert }_2^2}:\,j = 0,\dots ,n } \}} ). \end{aligned}$$

From this and by (1.45), \(a=x^\star \). Since the above POP with optimal value \(\xi _n\) has a unique minimizer \(x^\star \) and its Lasserre’s hierarchy has finite convergence, the solution y of the SDP with optimal value \(\eta _k^n\) must have a representing 1-atomic measure \(\mu \) supported on \(x^\star \). Then y satisfies the flat extension condition when k is large enough and \({\mathop \mathrm{supp}\nolimits }( \mu ) = \{ {{x^\star }}\}\subset S(g,h)\). The conclusion follows. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mai, N.H.A., Lasserre, JB. & Magron, V. Positivity certificates and polynomial optimization on non-compact semialgebraic sets. Math. Program. 194, 443–485 (2022). https://doi.org/10.1007/s10107-021-01634-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-021-01634-1

Keywords

Mathematics Subject Classification

Navigation