Skip to main content
Log in

Extending tensors on polar manifolds

  • Published:
Mathematische Annalen Aims and scope Submit manuscript

Abstract

Given a polar action on a Riemannian manifold, we prove surjectivity of restriction to the section for general invariant tensors, and a sharper surjectivity result in the special case of metrics. These are related to the Chevalley Restriction Theorem and Michor’s Basic Forms Theorem. The proofs rely on results in the Invariant Theory of finite reflection groups and symmetric pairs, some of which may be of independent interest.

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

References

  1. Bourbaki, N.: Éléments de mathématique. Fasc. XXXIV. Groupes et algèbres de Lie. In: Chapitre IV: Groupes de Coxeter et systèmes de Tits. Chapitre V: Groupes engendrés par des réflexions. Chapitre VI: systèmes de racines. Actualités Scientifiques et Industrielles, vol. 1337. Hermann, Paris (1968)

  2. Bredon, G.E.: Introduction to compact Transformation Groups. Pure and Applied Mathematics, vol. 46. Academic Press, New York (1972)

  3. Chevalley, C.: Invariants of finite groups generated by reflections. Am. J. Math. 77, 778–782 (1955)

    Article  MathSciNet  MATH  Google Scholar 

  4. Coxeter, H.S.M.: The product of the generators of a finite group generated by reflections. Duke Math. J. 18, 765–782 (1951)

    Article  MathSciNet  MATH  Google Scholar 

  5. Dadok, J.: Polar coordinates induced by actions of compact Lie groups. Trans. Am. Math. Soc. 288(1), 125–137 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  6. Eisenbud, D.: Commutative Algebra. Graduate Texts in Mathematics, vol. 150. Springer, New York (1995). doi:10.1007/978-1-4612-5350-1. With a view toward algebraic geometry

  7. Eschenburg, J.H., Heintze, E.: On the classification of polar representations. Math. Z. 232(3), 391–398 (1999). doi:10.1007/PL00004763

    Article  MathSciNet  MATH  Google Scholar 

  8. Field, M.J.: Transversality in \(G\)-manifolds. Trans. Am. Math. Soc. 231(2), 429–450 (1977)

    MathSciNet  MATH  Google Scholar 

  9. Flatto, L.: Invariants of finite reflection groups and mean value problems. II. Am. J. Math. 92, 552–561 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  10. Flatto, L., Weiner, S.M.M.: Invariants of finite reflection groups and mean value problems. Am. J. Math. 91, 591–598 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  11. Fulton, W., Harris, J.: Representation Theory. Graduate Texts in Mathematics, vol. 129. Springer, New York (1991). A first course, Readings in Mathematics

  12. Geck, M., Hiss, G., Lübeck, F., Malle, G., Pfeiffer, G.: CHEVIE–a system for computing and processing generic character tables for finite groups of Lie type, Weyl groups and Hecke algebras. Appl. Algebra Eng. Commun. Comput. 7, 175–210 (1996)

    Article  MATH  Google Scholar 

  13. Grove, K., Ziller, W.: Polar manifolds and actions. J. Fixed Point Theory Appl. 11(2), 279–313 (2012). doi:10.1007/s11784-012-0087-y

    Article  MathSciNet  MATH  Google Scholar 

  14. Humphreys, J.E.: Reflection Groups and Coxeter Groups. Cambridge Studies in Advanced Mathematics, vol. 29. Cambridge University Press, Cambridge (1990)

  15. Hunziker, M.: Classical invariant theory for finite reflection groups. Transform Groups 2(2), 147–163 (1997). doi:10.1007/BF01235938

    Article  MathSciNet  MATH  Google Scholar 

  16. Iwasaki, K., Kenma, A., Matsumoto, K.: Polynomial invariants and harmonic functions related to exceptional regular polytopes. Exp. Math. 11(2), 313–319 (2002). http://projecteuclid.org/euclid.em/1062621224

  17. Joseph, A.: On a Harish-Chandra homomorphism. C. R. Acad. Sci. Paris Sér. I Math. 324(7), 759–764 (1997). doi:10.1016/S0764-4442(97)86940-6

  18. Kane, R.: Reflection Groups and Invariant Theory. CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC, vol. 5. Springer, New York (2001)

  19. Kumar, S.: A refinement of the PRV conjecture. Invent. Math. 97(2), 305–311 (1989). doi:10.1007/BF01389044

    Article  MathSciNet  MATH  Google Scholar 

  20. Mathieu, O.: Construction d’un groupe de Kac–Moody et applications. Composit. Math. 69(1), 37–60 (1989). http://www.numdam.org/item?id=CM_1989__69_1_37_0

  21. Mehta, M.L.: Basic sets of invariant polynomials for finite reflection groups. Commun. Algebra 16(5), 1083–1098 (1988). doi:10.1080/00927878808823619

    Article  MathSciNet  MATH  Google Scholar 

  22. Mendes, R.A.E.: Equivariant tensors on polar manifolds. Ph.D. thesis, University of Pennsylvania (2011). http://repository.upenn.edu/dissertations/AAI3463026

  23. Michor, P.W.: Basic differential forms for actions of Lie groups. Proc. Am. Math. Soc. 124(5), 1633–1642 (1996). doi:10.1090/S0002-9939-96-03195-4

    Article  MathSciNet  MATH  Google Scholar 

  24. Michor, P.W.: Basic differential forms for actions of Lie groups. II. Proc. Am. Math. Soc. 125(7), 2175–2177 (1997). doi:10.1090/S0002-9939-97-03929-4

    Article  MathSciNet  MATH  Google Scholar 

  25. Palais, R.S., Terng, C.L.: A general theory of canonical forms. Trans. Am. Math. Soc. 300(2), 771–789 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  26. Schönert, M., et al.: GAP-Groups, Algorithms, and Programming—version 3 release 4 patchlevel 4. Lehrstuhl D für Mathematik, Rheinisch Westfälische Technische Hochschule, Aachen, Germany (1997)

  27. Schwarz, G.W.: When polarizations generate. Transform Groups 12(4), 761–767 (2007). doi:10.1007/s00031-006-0044-1

    Article  MathSciNet  MATH  Google Scholar 

  28. Smith, L.: Polynomial Invariants of Finite Groups. Research Notes in Mathematics, vol. 6. A K Peters Ltd., Wellesley (1995)

  29. Solomon, L.: Invariants of finite reflection groups. Nagoya Math. J. 22, 57–64 (1963)

    MathSciNet  MATH  Google Scholar 

  30. Springer, T.A.: Invariant Theory. Lecture Notes in Mathematics, vol. 585. Springer, Berlin (1977)

  31. Tevelev, E.A.: On the Chevalley restriction theorem. J. Lie Theory 10(2), 323–330 (2000)

    MathSciNet  MATH  Google Scholar 

  32. Wallach, N.R.: Invariant differential operators on a reductive Lie algebra and Weyl group representations. J. Am. Math. Soc. 6(4), 779–816 (1993). doi:10.2307/2152740

    Article  MathSciNet  MATH  Google Scholar 

  33. Warner G (1972) Harmonic Analysis on Semi-Simple Lie Groups. I. die Grundlehren der mathematischen Wissenschaften, Band 188. Springer, New York

  34. Weyl, H.: The Classical Groups: Their Invariants and Representations. Princeton University Press, Princeton (1939)

    MATH  Google Scholar 

Download references

Acknowledgments

Part of this work was completed during my Ph.D., and I would like to thank my advisor W. Ziller for the long-term support. I would also like to thank A. Lytchak and J. Tevelev for useful communication.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ricardo A. E. Mendes.

Appendix: Hessian Theorem for finite reflection groups

Appendix: Hessian Theorem for finite reflection groups

In this section we provide proofs of Theorem 4 and Observation 2. The latter relies on Calculation 1, which can be checked with a computer. See [22] for the source code for a script written in GAP [26] using the package CHEVIE [12], which performs these calculations.

Note that as far as the proofs of Theorem 2 and Observation 1 are concerned, one only needs to consider crystallographic reflection groups (see [14] for a definition). Our proofs include the non-crystallographic cases for the sake of completeness.

Recall some facts about finite reflection groups: First, the algebra of invariants, \(A=\mathbb {R}[\varSigma ]^W\), is a free polynomial algebra with n generators, where \(n=\dim \varSigma \). This is known as Chevalley’s Theorem—see [1, Chapter V]. Such a set \(\{\rho _i\}\) of homogeneous generators is called a set of basic invariants, and \(d_i=\deg \rho _i\) are called the degrees of W.

Second, \(\mathbb {R}[\varSigma ]\) is a free \(A=\mathbb {R}[\varSigma ]^W\)-module, more precisely \(\mathbb {R}[\varSigma ]=A\otimes H\), where H is isomorphic to the regular representation of W (see Theorem B in [3]). In particular, \({\mathcal {M}}=\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^W\) is a free A-module of rank \((n^2+n)/2\).

Third, \(\varSigma \) is reducible as a W-representation if and only if \(\varSigma =\varSigma _1\times \varSigma _2\) and \(W=W_1\times W_2\) for two reflection groups \(W_k\subset O(\varSigma _k)\)—see section 2.2 in [14]. Thus the following proposition reduces the proofs of Theorem 4 and Observation 2 to the irreducible case.

Lemma 5

Let \(W_k\subseteq O(\varSigma _k),\) \(k=1,2\) be finite reflection groups in the Euclidean vector spaces \(\varSigma _k,\) and let \(W=W_1\times W_2\subset O(\varSigma )=O(\varSigma _1\times \varSigma _2)\). Then there are W-invariant polynomials on \(\varSigma \) whose Hessians generate \(\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^W\) if and only if the same holds for \(W_k\subseteq O(\varSigma _k),\) \(k=1,2\).

Proof

Assume there are \(Q_j\in \mathbb {R}[\varSigma ]^W\) whose Hessians generate \(\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^W\). Then the restrictions \(Q_j|_{\varSigma _1}\) generate \(\mathbb {R}[\varSigma _1,{\mathrm {Sym}}^2(\varSigma _1^*)]^{W_1}\) as an \(\mathbb {R}[\varSigma _1]^{W_1}\)-module.

Indeed, every \(\sigma \in \mathbb {R}[\varSigma _1,{\mathrm {Sym}}^2(\varSigma _1^*)]^{W_1}\) can be naturally extended to \({\tilde{\sigma }}\in \mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^W\) that is constant on each copy of \(\varSigma _2\). Then there are \(a_j\in \mathbb {R}[\varSigma ]^W\) such that \({\tilde{\sigma }}=\sum _j a_j{\mathrm {Hess}}(Q_j)\). Therefore

$$\begin{aligned} \sigma =\sum _j (a_j|_{\varSigma _1}){\mathrm {Hess}}(Q_j|_{\varSigma _1}) \end{aligned}$$

and similarly for \(\mathbb {R}[\varSigma _2,{\mathrm {Sym}}^2(\varSigma _2^*)]^{W_2}\).

For the converse, let \(\rho _j\in \mathbb {R}[\varSigma _1]^{W_1}\), \(j=1,\ldots , n_1\) and \(\psi _j\in \mathbb {R}[\varSigma _2]^{W_2}\), \(j=1,\ldots , n_2\) be basic invariants on \(\varSigma _1\) and \(\varSigma _2\) respectively; and \(Q_j\in \mathbb {R}[\varSigma _1]^{W_1}\), \(j=1,\ldots , (n_1^2+n_1)/2\), \(R_j\in \mathbb {R}[\varSigma _2]^{W_2}\), \(j=1,\ldots , (n_2^2+n_2)/2\) be homogeneous invariants whose Hessians form a basis for the corresponding spaces of equivariant symmetric 2-tensors.

We claim that the Hessians of the following set of \(W=W_1\times W_2\)-invariant polynomials on \(\varSigma =\varSigma _1\times \varSigma _2\) form a basis for the space of equivariant symmetric 2-tensors on \(\varSigma \):

$$\begin{aligned} \{Q_j\}\cup \{R_j\}\cup \{\rho _i\psi _j\} \end{aligned}$$

Indeed, \(\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^W\) decomposes as

$$\begin{aligned} \mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma _1^*)]^W\oplus \mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma _2^*)]^W\oplus \mathbb {R}[\varSigma ,\varSigma _1^*\otimes \varSigma _2^*]^W \end{aligned}$$

The first two pieces are freely generated over \(\mathbb {R}[\varSigma ]^W\) by Hess\(Q_j\) and Hess\(R_j\). The third piece can be rewritten as \(\mathbb {R}[\varSigma ,\varSigma _1^*\otimes \varSigma _2^*]^W=\mathbb {R}[\varSigma _1,\varSigma _1^*]^{W_1}\otimes \mathbb {R}[\varSigma _2,\varSigma _2^*]^{W_2}\). By Solomon’s Theorem [29], \(\mathbb {R}[\varSigma _k,\varSigma _k^*]^{W_k}\), \(k=1,2\), are freely generated by \(d\rho _j\) and \(d\psi _j\), so that \(\mathbb {R}[\varSigma ,\varSigma _1^*\otimes \varSigma _2^*]^W\) is freely generated by \((d\rho _j\otimes d\psi _j + d\psi _j\otimes d\rho _j)\). To finish the proof of the claim one uses the product rule

$$\begin{aligned} {\mathrm {Hess}}(\rho _i\psi _j)= d\rho _i\otimes d\psi _j + d\psi _j\otimes d\rho _j+ \rho _i{\mathrm {Hess}}(\psi _j)+\psi _j{\mathrm {Hess}}(\rho _i) \end{aligned}$$

Proof of Theorem 4

By Lemma 5, we may assume W is irreducible of classical type.

If W is of type A, type B, or dihedral, then all multi-variable invariants are generated by polarizations, by [15, 34]. Hence it is enough to show that \({\mathcal {P}}^3\cap {\mathcal {M}}\) is generated, as an A-module, by Hessians of invariants. This follows from the product rule

$$\begin{aligned} {\mathrm {Hess}}(\rho _i\psi _j)= d\rho _i\otimes d\psi _j + d\psi _j\otimes d\rho _j+ \rho _i{\mathrm {Hess}}(\psi _j)+\psi _j{\mathrm {Hess}}(\rho _i) \end{aligned}$$

If, on the other hand, W is of type D, then the multi-variable invariants are generated by generalized polarizations—see Theorems 3.1 and 3.4 in [15], or Proposition 2 in Appendix 2 of [32]. But degree considerations imply that \({\mathcal {M}}\) is in fact generated by (classical) polarizations, hence also by Hessians by the product rule.

Calculation 1

Let W of type \(H_3\), \(H_4\), \(F_4\), \(E_6\), \(E_7\) or \(E_8\). Then there is a choice of basic invariants \(\rho _1, \ldots \rho _n \text { with }\deg (\rho _1)<\cdots <\deg (\rho _n)\), and of a regular vector \(v\in \varSigma \), such that \( \{ {\mathrm {Hess}}(\rho ^*Q)(v)\ |\ Q\in T\}\) is linearly independent, where \(\rho =(\rho _1, \ldots , \rho _n):\varSigma \rightarrow \mathbb {R}^n\), and T is the set of polynomials on \(\mathbb {R}^n\) given in Table 1.

Table 1 Monomials with generating Hessians

We remark that the computation above is independent of the choice of basic invariants and regular vector—see Lemmas 6 and 7. Moreover, one may construct a set of basic invariants consisting of “orbit Chern classes” from some linear functional \(\lambda _0:\varSigma \rightarrow \mathbb {R}\) and the degrees \(d_i\). Namely, \(\rho _i=\sum _{\lambda \in W\lambda _0}\lambda ^{d_i}\)—see [9, 10, 22, 28]. See also [16, 21] for (other) explicit sets of basic invariants of exceptional groups. The degrees \(d_1,\ldots ,d_n\) of the exceptional groups are listed in Table 2 for the convenience of the reader (see [4]).

Table 2 Degrees of exceptional finite reflection groups

The following Lemma is analogous to Proposition 3.13 in [14]. We will use the special case \(U=\text {Sym}^2\varSigma ^*\) in proving both Observation 2 from Calculation 1, and independence of the choice of regular vector v in Calculation 1.

Lemma 6

Let \(W\subset O(\varSigma )\) be an irreducible finite reflection group,  and \(\eta : W\rightarrow O(U)\) an orthogonal representation,  with character \(\chi \). Choose a basis \(\{e_1,\ldots , e_l\}\) for U,  and let \(f_1, \ldots , f_l\in \mathbb {R}[\varSigma , U]^W\) be homogeneous elements given by \(f_i=\sum _j a_{ij} e_j\). Let \(D=\det (a_{ij})\in \mathbb {R}[\varSigma ]\). Then : 

  1. (a)

    D is divisible by the following product over all reflections \(r\in W{:}\)

    $$\begin{aligned} J_\eta =\prod _{r}(\lambda _r)^\frac{l-\chi (r)}{2} \end{aligned}$$

    with \(\lambda _r\) a linear functional whose kernel equals the hyperplane fixed by r.

  2. (b)

    If \(\{f_i\}\) is a basis of \(\mathbb {R}[\varSigma , U]^W\) over \(\mathbb {R}[\varSigma ]^W,\) then D and \(J_\eta \) have the same degree.

  3. (c)

    \(\{f_i\}\) forms a basis if and only if \(D=cJ_\eta \) for some \(c\in \mathbb {R}-\{0\}\).

Proof

  1. (a)

    Let \(r\in W\) be a reflection. The transformation \(\eta (r):U\rightarrow U\) is diagonalizable with eigenvalues \(1,-1\). In particular the multiplicity of \(-1\) equals \(k=(l-\chi (r))/2\). Assume, without loss of generality, that \(e_1, \ldots , e_k\) is a basis for the eigenspace of \(\eta (r)\) associated to the eigenvalue \(-1\). Then the first k columns of \((a_{ij})\) are odd with respect to r. By multi-linearity, D vanishes to order k on the hyperplane fixed by r. In other words, D is divisible by \((\lambda _r)^k\). Since this is true for every reflection r, D is divisible by \(J_\eta \).

  2. (b)

    For any graded vector space \(E=\oplus _i E_i\), denote its Poincaré series by \(P_t(E)=\sum _i \dim (E_i)t^i\). Let

    $$\begin{aligned} P(t)=\frac{P_t(\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^{W})}{P_t(\mathbb {R}[\varSigma ]^W)} \end{aligned}$$

    Since \(P(t)=\sum _{i=1}^lt^{\deg (f_i)}\), we have \(\deg (D)=P'(1)\). On the other hand, \(P_t(\mathbb {R}[\varSigma ]^W)=\varPi _{i=1}^n(1-t^{d_i})^{-1}\), while the numerator can be computed using Molien’s formula [18, page 249]:

    $$\begin{aligned} P_t(\mathbb {R}[\varSigma ,{\mathrm {Sym}}^2(\varSigma ^*)]^{W})=\frac{1}{|W|}\sum _{g\in W}\frac{\chi (g)}{\det (1-tg)} \end{aligned}$$

    Note that for \(g=1\), \(\det (1-tg)=(1-t)^n\); for \(g=r\) a reflection, \(\det (1-tg)=(1-t)^{n-1}(1+t)\); and for all other g, \((1-t)^{n-1}\) does not divide \(\det (1-tg)\). Therefore, when computing \(P'(1)\), terms of the latter type vanish:

    $$\begin{aligned}&|W|P'(1)=\chi (1)\left. \frac{d}{dt}\right| _{t=1} \frac{\varPi _{i=1}^n(1-t^{d_i})}{(1-t)^n} + \sum _{r \text { refl.}}\chi (r)\left. \frac{d}{dt}\right| _{t=1} \frac{\varPi _{i=1}^n(1-t^{d_i})}{(1-t)^{n-1}(1+t)}\\&P'(1)=\frac{1}{|W|}\left( \frac{lN|W|}{2} - \sum _{r \text { refl.}}\chi (r)\frac{|W|}{2}\right) =\sum _{r \text { reflection}}\frac{l-\chi (r)}{2} \end{aligned}$$

    where N is the number of reflections and we have used the identities \(d_1\cdots d_n=|W|\) and \((d_1-1)+\cdots +(d_n-1)=N\). Thus \(P'(1)\) equals the degree of \(J_\eta \).

  3. (c)

    Assume \(\{f_i\}\) is a basis. By parts 1 and 2, \(D=cJ_\eta \) for some \(c\in \mathbb {R}\). Assume for a contradiction that \(c=0\). This means that \(\{f_i(v)\}\) is linearly dependent for every \(v\in \varSigma \). Take a regular v, and let B be a small open W-invariant neighborhood of the orbit Wv. Since W acts freely on B, on can construct \(\sigma \in C^\infty (\varSigma , U)^W\) with supp\((\sigma )\subset B\), and \(\sigma (v)\notin \text {span}\{f_i(v)\}\). This contradicts the fact that \(\mathbb {R}[\varSigma , U]^W\) is dense in \( C^\infty (\varSigma , U)^W\). Now assume \(D=cJ_\eta \) for some \(c\in \mathbb {R}-\{0\}\). Choose any homogeneous basis \(\{f_i'\}\) of \({\mathcal {M}}\), and define \(D'\) analogously to D. Writing \(f_i=\sum _j b_{ij} f_j'\), we see that \(\det (b_{ij})\in \mathbb {R}-\{0\}\), because \(D=D'\det (b_{ij})\). This implies that \((b_{ij})\) is invertible in the algebra of matrices with coefficients in A, so that \(\{f_i\}\) is a basis of \({\mathcal {M}}\) over A, too.

Proof of Observation 2

Assume Calculation 1. We claim that

$$\begin{aligned} \{f_i\}=\{\text {Hess}Q\ ,\ Q\in T\} \end{aligned}$$

forms a basis for \({\mathcal {M}}\) over A. Indeed, by inspection, the sum of the degrees of \(f_i\) equals \(\deg (J_\eta )\), which in this case is \(N(n-1)\). Using Lemma 6, we see that \(D=cJ_\eta \) for some \(c\ne 0\), so that \(\{f_i\}\) forms a basis.

Note that independence of the choice of regular vector follows from Lemma 6, because the zero set of \(J_\eta \) is contained in the singular set.

Lemma 7

Calculation 1 is independent of the choice of basic invariants \(\rho _i\).

Proof

Assume \(\{\rho _i\}\) and \(\{\psi _i\}\) are two sets of basic invariants, and that

$$\begin{aligned} s\{ {\mathrm {Hess}}(\rho ^*Q)(v)\ |\ Q\in T\} \end{aligned}$$

is linearly independent at some (hence all) regular vector \(v\in \varSigma \). By Lemma 6, \( \{ {\mathrm {Hess}}(\rho ^*Q)\ |\ Q\in T\}\) forms a basis of \({\mathcal {M}}\) as a free A-module. We claim that \( \{ {\mathrm {Hess}}(\psi ^*Q)\ |\ Q\in T\}\) is also a basis, so that in particular \( \{ {\mathrm {Hess}}(\psi ^*Q)(v)\ |\ Q\in T\}\) is linearly independent for every regular vector v.

Recall the graded version of Nakayama’s Lemma (see Exercise 4.6a in [6]): a set of homogeneous elements \(f_i\in {\mathcal {M}}\) generates \({\mathcal {M}}\) as an A-module if and only if their images in \({\mathcal {M}}/I{\mathcal {M}}\) span it as real vector space, where I is the ideal of A generated by the elements of positive degree.

Since the degrees \(d_i\) are all distinct, we may assume without loss of generality that \( \psi _i=\rho _i + R_i(\rho _1, \ldots , \rho _{i-1})\). Note that the Hessian of a product of three or more (not necessarily distinct) basic invariants belongs to \(I{\mathcal {M}}\), so that modulo \(I{\mathcal {M}}\) we have \(\text {Hess}(\psi _i \psi _j) \equiv \text {Hess}(\rho _i\rho _j)\), and

$$\begin{aligned} \text {Hess}(\psi _i)\equiv \text {Hess}(\rho _i)+ \sum _{j,k}c_{j,k}\text {Hess}(\rho _j \rho _k) \end{aligned}$$

where \(c_{j,k}\in \mathbb {R}\) vanishes unless \(d_j+d_k=d_i\).

By inspection, \(y_jy_k\in T\) whenever \(d_j+d_k=d_i\) for some i (see Table 2). Therefore \(\{\text {Hess}(\psi ^*Q)\ ,\ Q\in T\}\) is written in terms of \(\{\text {Hess}(\rho ^*Q)\ ,\ Q\in T\}\) (modulo \(I{\mathcal {M}}\)) using a triangular matrix with 1’s in the diagonal, showing that \( \{{\mathrm {Hess}}(\psi ^*Q)\ |\ Q\in T\}\) is a basis of \({\mathcal {M}}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mendes, R.A.E. Extending tensors on polar manifolds. Math. Ann. 365, 1409–1424 (2016). https://doi.org/10.1007/s00208-015-1319-4

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00208-015-1319-4

Mathematics Subject Classification

Navigation