Skip to main content
Log in

The Capelli eigenvalue problem for Lie superalgebras

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

For a finite dimensional unital complex simple Jordan superalgebra J, the Tits–Kantor–Koecher construction yields a 3-graded Lie superalgebra \(\mathfrak {g}^\flat \cong \mathfrak {g}^\flat (-1)\oplus \mathfrak {g}^\flat (0)\oplus \mathfrak {g}^\flat (1)\), such that \(\mathfrak {g}^\flat (-1)\cong J\). Set \(V:=\mathfrak {g}^\flat (-1)^*\) and \(\mathfrak {g}:=\mathfrak {g}^\flat (0)\). In most cases, the space \(\mathscr {P}(V)\) of superpolynomials on V is a completely reducible and multiplicity-free representation of \(\mathfrak {g}\), and there exists a direct sum decomposition \(\mathscr {P}(V):=\bigoplus _{\lambda \in \Omega }V_\lambda \), where \(\left( V_\lambda \right) _{\lambda \in \Omega }\) is a family of irreducible \(\mathfrak {g}\)-modules parametrized by a set of partitions \(\Omega \). In these cases, one can define a natural basis \(\left( D_\lambda \right) _{\lambda \in \Omega }\) of “Capelli operators” for the algebra \(\mathscr {PD}(V)^{\mathfrak {g}}\) of \(\mathfrak {g}\)-invariant superpolynomial differential operators on V. In this paper we complete the solution to the Capelli eigenvalue problem, which asks for the determination of the scalar \(c_\mu (\lambda )\) by which \(D_\mu \) acts on \(V_\lambda \). We associate a restricted root system \(\varSigma \) to the symmetric pair \((\mathfrak {g},\mathfrak {k})\) that corresponds to J, which is either a deformed root system of type \(\mathsf {A}(m,n)\) or a root system of type \(\mathsf {Q}(n)\). We prove a necessary and sufficient condition on the structure of \(\varSigma \) for \(\mathscr {P}(V)\) to be completely reducible and multiplicity-free. When \(\varSigma \) satisfies the latter condition we obtain an explicit formula for the eigenvalue \(c_\mu (\lambda )\), in terms of Sergeev–Veselov’s shifted super Jack polynomials when \(\varSigma \) is of type \(\mathsf {A}(m,n)\), and Okounkov-Ivanov’s factorial Schur Q-polynomials when \(\varSigma \) is of type \(\mathsf {Q}(n)\). Along the way, we prove that the natural map from the centre of the enveloping algebra of \(\mathfrak {g}\) into \(\mathscr {PD}(V)^{\mathfrak {g}}\) is surjective in all cases except when \(J\cong F \), where \( F \) is the 10-dimensional exceptional Jordan superalgebra.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. We remark that \(\mathfrak {g}^\flat \) is a slight modification of the simple Lie superalgebra that is constructed from J by the Kantor functor (see Remark A.3).

  2. The classical Capelli operator appears as a special case of the operators \(D_\lambda \). For this reason, we call the \(D_\lambda \) the Capelli operators.

References

  1. Alldridge, A., Sahi, S., Salmasian, H.: Schur Q-functions and the Capelli eigenvalue problem for the Lie superalgebra \({\mathfrak{q}}(n)\). In: Christensen, J.G., Dann, S., Dawson, M. (eds.) Proceedings of the Conference on Representation Theory and Harmonic Analysis on Symmetric Spaces, Contemporary Mathematics, vol. 714, pp. 1–22. American Mathematical Society, Providence, RI (2018)

  2. Brini, A., Huang, R.Q., Teolis, A.G.B.: The umbral symbolic method for supersymmetric tensors. Adv. Math. 96(2), 123–193 (1992)

    MathSciNet  MATH  Google Scholar 

  3. Cantarini, N., Kac, V.G.: Classification of linearly compact simple Jordan and generalized Poisson superalgebras. J. Algebra 313(1), 100–124 (2007)

    MathSciNet  MATH  Google Scholar 

  4. Cheng, S.-J., Wang, W.: Remarks on the Schur–Howe–Sergeev duality. Lett. Math. Phys. 52(2), 143–153 (2000)

    MathSciNet  MATH  Google Scholar 

  5. Cheng, S.-J., Wang, W.: Howe duality for Lie superalgebras. Compos. Math. 128(1), 55–94 (2001)

    MathSciNet  MATH  Google Scholar 

  6. Cheng, S.J., Wang, W.: Dualities and Representations of Lie Superalgebras. Graduate Studies in Mathematics, vol. 144. American Mathematical Society, Providence (2012)

    MATH  Google Scholar 

  7. Coulembier, K.: The orthosymplectic superalgebra in harmonic analysis. J. Lie Theory 23(1), 55–83 (2013)

    MathSciNet  MATH  Google Scholar 

  8. Debiard, A.: Polynômes de Tchébychev et de Jacobi dans un espace euclidien de dimension \(p\). C. R. Acad. Sci. Paris Sér. I Math. 296(13), 529–532 (1983)

    MathSciNet  MATH  Google Scholar 

  9. Goodman, R., Wallach, N.R.: Symmetry, representations, and invariants, vol. 255 of Graduate Texts in Mathematics. Springer, Dordrecht (2009)

    Google Scholar 

  10. Helgason, S.: Some results on invariant differential operators on symmetric spaces. Am. J. Math. 114(4), 789–811 (1992)

    MathSciNet  MATH  Google Scholar 

  11. Howe, R.: Perspectives on invariant theory: Schur duality, multiplicity-free actions and beyond. In: The Schur lectures (1992) (Tel Aviv), vol. 8 of Israel Math. Conf. Proc., pp. 1–182. Bar-Ilan Univ., Ramat Gan (1995)

  12. Howe, R., Umeda, T.: The Capelli identity, the double commutant theorem, and multiplicity-free actions. Math. Ann. 290(3), 565–619 (1991)

    MathSciNet  MATH  Google Scholar 

  13. Humphreys, J.E.: Representations of semisimple Lie algebras in the BGG category \(\mathscr {O}\). Graduate Studies in Mathematics, vol. 94. American Mathematical Society, Providence, RI (2008)

  14. Ivanov, V.N.: A combinatorial formula for factorial Schur \(Q\)-functions. In: Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 256(Teor. Predst. Din. Sist. Komb. i Algoritm. Metody. 3), no. 265, pp. 73–94 (1999)

  15. Joseph, A.: A preparation theorem for the prime spectrum of a semisimple Lie algebra. J. Algebra 48(2), 241–289 (1977)

    MathSciNet  MATH  Google Scholar 

  16. Kac, V.G.: Classification of simple \(Z\)-graded Lie superalgebras and simple Jordan superalgebras. Commun. Algebra 5(13), 1375–1400 (1977)

    MathSciNet  MATH  Google Scholar 

  17. Knop, F., Sahi, Siddhartha: Difference equations and symmetric polynomials defined by their zeros. Int. Math. Res. Not. 10, 473–486 (1996)

    MathSciNet  MATH  Google Scholar 

  18. Kostant, B., Sahi, S.: The Capelli identity, tube domains, and the generalized Laplace transform. Adv. Math. 87(1), 71–92 (1991)

    MathSciNet  MATH  Google Scholar 

  19. Macdonald, I.G.: Symmetric Functions and Hall Polynomials. Oxford Classic Texts in the Physical Sciences, 2nd edn. The Clarendon Press, Oxford University Press, New York (2015). (With contribution by A. V. Zelevinsky and a foreword by Richard Stanley, Reprint of the 2008 paperback edition (2015))

    Google Scholar 

  20. Meinrenken, E.: Clifford algebras and Lie theory, volume 58 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer, Heidelberg (2013)

  21. Nazarov, M.: Capelli identities for Lie superalgebras. Ann. Sci. École Norm. Sup. (4) 30(6), 847–872 (1997)

    MathSciNet  MATH  Google Scholar 

  22. Pragacz, P.: Algebro-geometric applications of Schur \(S\)- and \(Q\)-polynomials. In: Malliavin, M.-P. (ed.) Topics in invariant theory (Paris, 1989/1990), vol. 1478 of Lecture Notes in Math., pp. 130–191. Springer, Berlin (1991)

  23. Sahi, S.: The spectrum of certain invariant differential operators associated to a Hermitian symmetric space. In: Brylinski, J.-L., Brylinski, R., Guillemin, V., Kac, V. (eds.) Lie theory and geometry, vol. 123 of Progr. Math., pp. 569–576. Birkhäuser Boston, Boston (1994)

    Google Scholar 

  24. Sahi, S., Salmasian, H.: The Capelli problem for \(\mathfrak{gl}(m|n)\) and the spectrum of invariant differential operators. Adv. Math. 303, 1–38 (2016)

    MathSciNet  MATH  Google Scholar 

  25. Scheunert, M.: The Theory of Lie Superalgebras. Lecture Notes in Mathematics An introduction, vol. 716. Springer, Berlin (1979)

    MATH  Google Scholar 

  26. Schmid, W.: Die Randwerte holomorpher Funktionen auf hermitesch symmetrischen Räumen. Invent. Math. 9, 61–80 (1969/1970)

  27. Sekiguchi, J.: Zonal spherical functions on some symmetric spaces. Publ. Res. Inst. Math. Sci. 12, 455–459 (1976/77) (supplement)

    MathSciNet  MATH  Google Scholar 

  28. Sergeev, A.N., Veselov, A.P.: Deformed quantum Calogero–Moser problems and Lie superalgebras. Commun. Math. Phys. 245(2), 249–278 (2004)

    MathSciNet  MATH  Google Scholar 

  29. Sergeev, A.N., Veselov, A.P.: Generalised discriminants, deformed Calogero–Moser–Sutherland operators and super-Jack polynomials. Adv. Math. 192(2), 341–375 (2005)

    MathSciNet  MATH  Google Scholar 

  30. Weingart, M.: Spectral functions of invariant operators on skew multiplicity free spaces. ProQuest LLC, Ann Arbor, MI, 2007. Thesis (Ph.D.)—Rutgers The State University of New Jersey—New Brunswick

  31. Zou, Y.M.: Categories of finite-dimensional weight modules over type I classical Lie superalgebras. J. Algebra 180(2), 459–482 (1996)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The research of Siddhartha Sahi was partially supported by a Simons Foundation grant (509766), of Hadi Salmasian by NSERC Discovery Grants (RGPIN-2013-355464 and RGPIN-2018-04044), and of Vera Serganova by an NSF Grant (1701532). This work was initiated during the Workshop on Hecke Algebras and Lie Theory, which was held at the University of Ottawa. The first and the second named authors thank the National Science Foundation (DMS-162350), the Fields Institute, and the University of Ottawa for funding this workshop.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hadi Salmasian.

Additional information

Publisher's Note

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

Appendices

Appendix A. The TKK construction

Recall that a vector superspace \(J:=J_{\overline{\mathbf {0}}}\oplus J_{\overline{\mathbf {1}}}\) is called a Jordan superalgebra if it is equipped with a supercommutative bilinear product \(J\times J\rightarrow J\) which satisfies the Jordan identity

$$\begin{aligned}&(-1)^{|x||z|}[L_x,L_{yz}]+(-1)^{|y||x|}[L_y,L_{zx}]+(-1)^{|z||y|}[L_z,L_{xy}]=0 \ \text { for homogeneous }\\&\quad x,y,z\in J, \end{aligned}$$

where we define \(L_a:J\rightarrow J\) for \(a\in J\) to be the left multiplication map \(x\mapsto ax\), and denote the parity of a homogeneous element \(a\in J\) by |a|.

Following [3], by a short grading of a Lie superalgebra \(\mathfrak {l}\) we mean a \({\mathbb {Z}}\)-grading of \(\mathfrak {l}\) of the form \(\mathfrak {l}:=\bigoplus _{t\in {\mathbb {Z}}}\mathfrak {l}(t)\), such that \(\mathfrak {l}(t)=\{0\}\) for \(t\not \in \{0,\pm 1\}\). Using the Kantor functor, in [16] Kac associates to J a simple Lie superalgebra \(\mathfrak {g}_J^{}\) (the TKK Lie superalgebra) with a short grading

$$\begin{aligned} \mathfrak {g}_J^{}:= \mathfrak {g}_J^{}(-1)\oplus \mathfrak {g}_J^{}(0)\oplus \mathfrak {g}_J^{}(1). \end{aligned}$$

We recall the definition of \(\mathfrak {g}_J^{}\). Set \(\mathfrak {g}_J^{}(-1):=J\), \(\mathfrak {g}_J^{}(0):={\mathrm{Span}}_{\mathbb {C}}\{L_a,[L_a,L_b]\,:\,a,b\in J\}\subseteq \mathrm {End}_{\mathbb {C}}(J)\), and \( \mathfrak {g}_J^{}(1):={\mathrm{Span}}_{\mathbb {C}}\{ P,[L_a,P]\,:\,a\in J\}\subseteq \mathrm {Hom}_{\mathbb {C}}(\mathcal {S}^2(J),J), \) where \(P:\mathcal {S}^2(J)\rightarrow J\) is the map \(P(x,y):=xy\), and \( [L_a,P](x,y):=a(xy)-(ax)y-(-1)^{|x||y|}(ay)x \). The Lie superbracket of \(\mathfrak {g}_J^{}\) is defined by the following relations.

  1. (i)

    \([A,a]:=A(a)\) for \(A\in \mathfrak {g}_J^{}(0)\) and \(a\in \mathfrak {g}_J^{}(-1)\).

  2. (ii)

    \([A,a](x):=A(a,x)\) for \(A\in \mathfrak {g}_J^{}(1)\), \(a\in \mathfrak {g}_J^{}(-1)\), and \(x\in J\).

  3. (iii)

    \([A,B](x,y):=A(B(x,y))-(-1)^{|A||B|}B(A(x),y)-(-1)^{|A||B|+|x||y|}B(A(y),x)\) for \(A\in \mathfrak {g}_J^{}(0)\), \(B\in \mathfrak {g}_J^{}(1)\), and \(x,y\in J\).

For the classification of finite dimensional complex simple Jordan superalgebras and their corresponding TKK Lie superalgebras, see the articles by Kac [16] and Cantarini and Kac [3].

If J has a unit \(1_J\in J\), then the elements \(e:=1_J\), \(f:=-2P\), and \(h:=2L_{1_J}\) of \(\mathfrak {g}_J^{}\) satisfy (1). It follows that \(\mathfrak {s}:={\mathrm{Span}}_{\mathbb {C}}\{e,f,h\}\) is a subalgebra of \(\mathfrak {g}_J^{}\) isomorphic to \(\mathfrak {sl}_2({\mathbb {C}})\). Indeed \(\mathfrak {s}\) is a short subalgebra of \(\mathfrak {g}_J^{}\) (see [3]). We recall the definition of a short subalgebra.

Definition A.1

Let \(\mathfrak {l}\) be a complex Lie superalgebra. A short subalgebra of \(\mathfrak {l}\) is a Lie subalgebra \(\mathfrak {a}\subseteq \mathfrak {l}_{\overline{\mathbf {0}}}\) that is isomorphic to \(\mathfrak {sl}_2({\mathbb {C}})\), with a basis efh that satisfies the relations (1), such that the eigenspace decomposition of \(\mathrm {ad}\left( -\frac{1}{2}h\right) \) defines a short grading of \(\mathfrak {l}\).

Remark A.2

Let \(\mathfrak {l}\) be a complex Lie superalgebra and let \(\mathfrak {a}\subseteq \mathfrak {l}_{\overline{\mathbf {0}}}\) be a short subalgebra of \(\mathfrak {l}\).

  1. (a)

    Assume that \(\mathfrak {l}\) is a subalgebra of another Lie superalgebra \(\tilde{\mathfrak {l}}\) such that \(\dim \tilde{\mathfrak {l}}=\dim \mathfrak {l}+1\). Since every finite dimensional \(\mathfrak {sl}_2({\mathbb {C}})\)-module is completely reducible, it follows that \(\tilde{\mathfrak {l}}\cong \mathfrak {l}\oplus {\mathbb {C}}\) as \(\mathfrak {a}\)-modules, so that \(\mathfrak {a}\) is a short subalgebra of \(\tilde{\mathfrak {l}}\) as well.

  2. (b)

    Every central extension \(0\rightarrow {\mathbb {C}}\rightarrow \hat{\mathfrak {l}}\rightarrow \mathfrak {l}\rightarrow 0\) splits on \(\mathfrak {a}\). An argument similar to part (a) implies that the image of \(\mathfrak {a}\) under the splitting section is a short subalgebra of \(\hat{\mathfrak {l}}\).

When J is isomorphic to \( gl (m,n)_+\), \( p (n)_+\), or \( q (n)_+\), it will be more convenient for us to replace \(\mathfrak {g}_J^{}\) by a non-simple Lie superalgebra which has a more natural matrix realization (see also Remark A.3). To this end, we define the Lie superalgebra \(\mathfrak {g}^\flat \) by

$$\begin{aligned} \mathfrak {g}^\flat := {\left\{ \begin{array}{ll} \mathfrak {gl}(2m|2n)&{}\text { if } J\cong gl (m,n)_+,\\ \mathfrak {p}(2n)&{}\text { if }J\cong p (n)_+,\\ \mathfrak {q}(2n)&{}\text { if }J\cong q (n)_+,\\ \mathfrak {g}_J^{}&{}\text { otherwise.} \end{array}\right. } \end{aligned}$$
(54)

For a precise description of \(\mathfrak {p}(2n)\) and \(\mathfrak {q}(2n)\) see Appendix B. From Remark A.2 it follows that the short subalgebra \(\mathfrak {s}\) of \(\mathfrak {g}_J^{}\) corresponds to a unique short subalgebra of \(\mathfrak {g}^\flat \). We use the same symbols \(\mathfrak {s}\), e, f, and h for denoting the short subalgebra of \(\mathfrak {g}^\flat \) and its corresponding basis.

By restriction of the adjoint representation, \(\mathfrak {g}^\flat \) is equipped with an \(\mathfrak {s}\)-module structure. This \(\mathfrak {s}\)-module is a direct sum of trivial and adjoint representations of \(\mathfrak {s}\), hence it integrates to a representation of the adjoint group \(\mathrm {PSL}_2({\mathbb {C}})\). Furthermore,

$$\begin{aligned} w:=\exp (\mathrm {ad}({f})) \exp (-\mathrm {ad}({e})) \exp (\mathrm {ad}({f})) \end{aligned}$$
(55)

represents the nontrivial element of the Weyl group of \(\mathrm {PSL}_2({\mathbb {C}})\).

Set \(\mathfrak {g}^\flat (t):=\{x\in \mathfrak {g}^\flat \, :\, [h,x]=-2tx\}\) for \(t\in \{0,\pm 1\}\). The Lie superalgebra \(\mathfrak {g}^\flat (0)\) naturally acts on \(\mathfrak {g}^\flat (-1)\cong J\). Set

$$\begin{aligned} \mathfrak {g}:=\mathfrak {g}^\flat (0) \text { and } V:=\mathfrak {g}^\flat (-1)^*:=\mathrm {Hom}_{\mathbb {C}}(V,{\mathbb {C}}^{1|0}). \end{aligned}$$

Thus the \(\mathfrak {g}\)-module V is the dual of the \(\mathfrak {g}\)-module J.

Remark A.3

The reason for replacing \(\mathfrak {g}_J^{}\) by \(\mathfrak {g}^\flat \) is to obtain a convenient way of associating partitions to the irreducible \(\mathfrak {g}\)-modules that occur in \(\mathscr {P}(V)\). For example, assume that \(J\cong gl (m,n)_+\), where \( gl (m,n)_+\) denotes the Jordan superalgebra of \((m+n)\times (m+n)\) matrices in (mn)-block form. Then \(\mathfrak {g}_J^{}\cong \mathfrak {sl}(2m|2n)\) if \(m\ne n\), and \(\mathfrak {g}_J^{}\cong \mathfrak {psl}(2m|2n)\) if \(m=n\). In both cases, \(\mathfrak {g}_J^{}(0)\) is closely related to \(\mathfrak {gl}(m|n)\oplus \mathfrak {gl}(m|n)\), but it is not isomorphic to it. However, \(\mathfrak {g}:= \mathfrak {g}^\flat (0)\cong \mathfrak {gl}(m|n)\oplus \mathfrak {gl}(m|n)\), and the irreducible summands of \(\mathscr {P}(V)\cong \mathscr {P}(({\mathbb {C}}^{m|n})^*\otimes {\mathbb {C}}^{m|n})\) are naturally parametrized by (mn)-hook partitions.

Appendix B. Classical Lie superalgebras

In this Appendix we give explicit realizations of classical Lie superalgebras \(\mathfrak {gl}(m|n)\), \(\mathfrak {gosp}(m|2n)\), \(\mathfrak {p}(n)\), and \(\mathfrak {q}(n)\). We describe root systems of \(\mathfrak {gl}(m|n)\), \(\mathfrak {gosp}(m|2n)\), and \(\mathfrak {q}(n)\), and choose Borel subalgebras in these Lie superalgebras.

1.1 B.1. The Lie superalgebra \(\mathfrak {gl}(m|n)\)

Let \(m,n\ge 1\) be integers. We use the usual realization of \(\mathfrak {gl}(m|n)\) as \((m+n)\times (m+n)\) matrices in (mn)-block form

$$\begin{aligned} \begin{bmatrix} A&B\\ C&D \end{bmatrix} \end{aligned}$$
(56)

where A is \(m\times m\) and D is \(n\times n\). The diagonal Cartan subalgebra of \(\mathfrak {gl}(m|n)\) is

$$\begin{aligned} \mathfrak {h}_{m|n}:=\left\{ \mathrm {diag}({\mathbf {s}},{\mathbf {t}})\,:\,\mathbf {s}:=(s_1,\ldots , s_m)\in {\mathbb {C}}^m\text { and }\mathbf {t}:=(t_1,\ldots ,t_n)\in {\mathbb {C}}^n\right\} . \end{aligned}$$
(57)

The standard characters \(\varepsilon _i,\delta _j:\mathfrak {h}_{m|n}\rightarrow {\mathbb {C}}\) are defined by

$$\begin{aligned} \varepsilon _i(\mathrm {diag}({\mathbf {s}},{\mathbf {t}})):=s_i\text { for }1\le i\le m\ \text { and }\ \delta _j(\mathrm {diag}({\mathbf {s}},{\mathbf {t}})):=t_j\text { for }1\le j\le n. \end{aligned}$$

We define \(\mathfrak {b}^{\mathrm {st}}_{m|n}\) (respectively, \(\mathfrak {b}^{\mathrm {op}}_{m|n}\)) to be the Borel subalgebras of \(\mathfrak {gl}(m|n)\) corresponding to the fundamental systems \({\Pi }^{\mathrm {st}}\) (respectively, \({\Pi }^{\mathrm {op}}\)), where

$$\begin{aligned} {\Pi }^{\mathrm {st}}:=\left\{ \varepsilon _i-\varepsilon _{i+1}\right\} _{i=1}^{m-1}\cup \left\{ \varepsilon _m-\delta _1\right\} \cup \left\{ \delta _j-\delta _{j+1}\right\} _{j=1}^{n-1} \text { and } {\Pi }^{\mathrm {op}}:=-{\Pi }^{\mathrm {st}} \end{aligned}$$

For every partition \(\lambda \in \mathcal {H}(m,n)\), we set

$$\begin{aligned} \lambda ^{\mathrm {st}}_{m|n}:= \sum _{i=1}^m\lambda _i\varepsilon _i+ \sum _{j=1}^n\langle \lambda _j'-m\rangle \delta _j. \end{aligned}$$
(58)

In the spacial case \(m=n\), we define \(\mathfrak {b}_{n|n}^{\mathrm {mx}} \) to be the Borel subalgebra corresponding to the fundamental system

$$\begin{aligned} {\Pi }^{\mathrm {mx}} :=\left\{ \delta _i-\varepsilon _i \right\} _{i=1}^n\cup \left\{ \varepsilon _{j}-\delta _{j-1} \right\} _{j=2}^{n}. \end{aligned}$$

1.2 B.2. The Lie superalgebra \(\mathfrak {gosp}(m|2n)\)

Let \(m,n\ge 1\) be integers. We begin with an explicit realization of \(\mathfrak {osp}(m|2n)\). Set \(r:=\lfloor \frac{m}{2}\rfloor \). Let \(J^+\) be the \(m\times m\) matrix defined by

$$\begin{aligned} J^+:=\begin{bmatrix} 1&0_{1\times r}&0_{1\times r}\\ 0_{r\times 1}&0_{r\times r}&I_{r\times r}\\ 0_{r\times 1}&I_{r\times r}&0_{r\times r} \end{bmatrix} \text { if } m=2r+1, \text { and } J^+:=\begin{bmatrix} 0_{r\times r}&I_{r\times r}\\ I_{r\times 1}&0_{r\times r} \end{bmatrix} \text { if } m=2r, \end{aligned}$$

Also, let \(J^-\) be the \(2n\times 2n\) matrix defined by

$$\begin{aligned} J^-:=\begin{bmatrix} 0_{n\times n}&I_{n\times n}\\ -I_{n\times n}&0_{n\times n} \end{bmatrix}. \end{aligned}$$

Let \(\{\mathsf {e}_i\}_{i=1}^m\cup \{\mathsf {e}'_j\}_{j=1}^{2n}\) be the standard homogeneous basis of \({\mathbb {C}}^{m|2n}\), and let \( \mathsf {B}:{\mathbb {C}}^{m|2n}\times {\mathbb {C}}^{m|2n}\rightarrow {\mathbb {C}}\) be the even supersymmetric bilinear form defined by

$$\begin{aligned} \mathsf {B}(\mathsf {e}_i,\mathsf {e}_j)=J^+_{i,j},\ \mathsf {B}(\mathsf {e}_i',\mathsf {e}_j')=J^-_{i,j}, \text { and } \mathsf {B}(\mathsf {e}_i,\mathsf {e}_j')=0. \end{aligned}$$

We realize the Lie superalgebra \(\mathfrak {osp}(m|2n)\) as the subalgebra of \(\mathfrak {gl}(m|2n)\) that leaves the bilinear form \({\mathsf {B}}:{\mathbb {C}}^{m|2n}\times {\mathbb {C}}^{m|2n}\rightarrow {\mathbb {C}}\) invariant. For \({\mathbf {s}}\in {\mathbb {C}}^r\) and \({\mathbf {t}}\in {\mathbb {C}}^n\), set

$$\begin{aligned} {\mathsf {d}}({\mathbf {s}},{\mathbf {t}}):= {\left\{ \begin{array}{ll} \mathrm {diag}({\mathbf {s}},-{\mathbf {s}},{\mathbf {t}},-{\mathbf {t}})&{}\text { if }m=2r,\\ \mathrm {diag}(0,{\mathbf {s}},-{\mathbf {s}},{\mathbf {t}},-{\mathbf {t}})&{}\text { if }m=2r+1. \end{array}\right. } \end{aligned}$$

Recall from Appendix B.1 that we denote the standard Cartan subalgebra of \(\mathfrak {gl}(m|2n)\) by \(\mathfrak {h}_{m|2n}\). Then \(\overline{\mathfrak {h}}_{m|2n}:=\mathfrak {h}_{m|2n}\cap \mathfrak {osp}(m|2n)\) is a Cartan subalgebra of \(\mathfrak {osp}(m|2n)\). We have

$$\begin{aligned} \overline{\mathfrak {h}}_{m|2n}=\left\{ {\mathsf {d}}({\mathbf {s}},{\mathbf {t}})\,:\,{\mathbf {s}}\in {\mathbb {C}}^r\text { and }{\mathbf {t}}\in {\mathbb {C}}^n\right\} , \end{aligned}$$

and the standard characters of \(\overline{\mathfrak {h}}_{m|2n}\) are given by

$$\begin{aligned} \varepsilon _i({\mathsf {d}}({\mathbf {s}},{\mathbf {t}})):=s_i \text { for }1\le i\le r \text { and } \delta _j({\mathsf {d}}({\mathbf {s}},{\mathbf {t}})):=t_j \text { for }1\le j\le n. \end{aligned}$$

Let \(\breve{\mathfrak {b}}_{m|2n}\subseteq \mathfrak {osp}(m|2n)\) be the Borel subalgebra corresponding to the fundamental system \({\Pi }\), where

$$\begin{aligned} {\Pi } := {\left\{ \begin{array}{ll} \left\{ \varepsilon _i-\varepsilon _{i+1} \right\} _{i=1}^{r-1} \cup \left\{ \varepsilon _r-\delta _1\right\} \cup \left\{ \delta _j-\delta _{j+1} \right\} _{j=1}^{n-1} \cup \left\{ \delta _n \right\} &{}\text { if }m=2r+1,\\ \left\{ \varepsilon _i-\varepsilon _{i+1} \right\} _{i=1}^{r-1} \cup \left\{ \varepsilon _r-\delta _1\right\} \cup \left\{ \delta _j-\delta _{j+1} \right\} _{j=1}^{n-1} \cup \left\{ 2\delta _n \right\} &{}\text { if }m=2r. \end{array}\right. } \end{aligned}$$

Finally, we set \(\mathfrak {gosp}(m|2n):=\mathfrak {osp}(m|2n)\oplus {\mathbb {C}}I\subseteq \mathfrak {gl}(m|2n)\), where \(I:=I_{(m+2n)\times (m+2n)}\). We also set

$$\begin{aligned} \mathfrak {b}_{m|2n}:=\breve{\mathfrak {b}}_{m|2n}\oplus {\mathbb {C}}I. \end{aligned}$$
(59)

We extend the standard characters \(\varepsilon _i,\delta _j\) of \(\overline{\mathfrak {h}}_{m|2n}\) to the subalgebra \(\tilde{\mathfrak {h}}_{m|2n}:=\mathfrak {h}_{m|2n}\cap \mathfrak {gosp}(m|2n)\) of diagonal matrices in \(\mathfrak {gosp}(m|2n)\), by setting \(\varepsilon _i(I)=\delta _j(I)=0\). Let \(\zeta :\tilde{\mathfrak {h}}_{m|2n}\rightarrow {\mathbb {C}}\) be the linear functional defined uniquely by

$$\begin{aligned} \zeta \big |_{\overline{\mathfrak {h}}_{m|2n}}=0\text { and }\zeta (I)=1. \end{aligned}$$

The set \(\{\varepsilon _i\}_{i=1}^r\cup \{\delta _j\}_{j=1}^n\cup \{\zeta \}\) is a basis for the dual of \(\tilde{\mathfrak {h}}_{m|2n}\). We remark that when \(\mathfrak {g}\) is of type \(\mathfrak {gosp}\) (i.e., in Cases III and V), we have \(\zeta (h)=2\) where \(h\in \mathfrak {g}^\flat \) is defined as in (1).

In Case V, where \(\mathfrak {g}=\mathfrak {gosp}(2|4)\), we need to consider an exceptional Borel subalgebra

$$\begin{aligned} \mathfrak {b}_{2|4}^\mathrm {ex}:= \hat{\mathfrak {b}}^{}_{2|4}\oplus {\mathbb {C}}I, \end{aligned}$$
(60)

where \(\hat{\mathfrak {b}}^{}_{2|4}\) is the Borel subalgebra of \(\mathfrak {osp}(2|4)\) corresponding to the fundamental system

$$\begin{aligned} {\Pi }^{\mathrm {ex}}:= \{-\varepsilon _1-\delta _1,\delta _1-\delta _2,2\delta _2\}. \end{aligned}$$

1.3 B.3. The anisotropic embedding of \(\mathfrak {osp}(m|2n)\) in \(\mathfrak {gl}(m|2n)\)

We will need another realization of \(\mathfrak {osp}(m|2n)\) inside \(\mathfrak {gl}(m|2n)\) which will be used in the description of the spherical subalgebra \(\mathfrak {k}\). Set \(r:=\lfloor \frac{m}{2}\rfloor \), and let \(\tilde{J}^-\) be the \(2n\times 2n\) matrix defined by

$$\begin{aligned} \tilde{J}^-:= \mathrm {diag} (\underbrace{\tilde{J},\ldots ,\tilde{J}}_{n\text { times}}) \text { where } \tilde{J}:=\begin{bmatrix} 0&1\\ -1&0\end{bmatrix} . \end{aligned}$$

Let \(\tilde{\mathsf {B}}:{\mathbb {C}}^{m|2n}\times {\mathbb {C}}^{m|2n}\rightarrow {\mathbb {C}}\) be the even supersymmetric bilinear form which is given in the standard basis \(\{\mathsf {e}_i\}_{i=1}^m\cup \{\mathsf {e}_j\}_{j=1}^{2n}\) of \({\mathbb {C}}^{m|2n}\) by

$$\begin{aligned} \tilde{\mathsf {B}}(\mathsf {e}_i,\mathsf {e}_j):=\delta _{i,j},\ \tilde{\mathsf {B}}(\mathsf {e}'_i,\mathsf {e}'_j):=\tilde{J}^-_{i,j}, \text { and } \tilde{\mathsf {B}}(\mathsf {e}_i,\mathsf {e}_j'):=0. \end{aligned}$$

Thus, the matrix of \(\tilde{\mathsf {B}}(\cdot ,\cdot )\) in the standard basis of \({\mathbb {C}}^{m|2n}\) is

$$\begin{aligned} \begin{bmatrix}I_{m\times m}&0_{m\times 2n}\\0_{2n\times m}&\tilde{J}^-\end{bmatrix}. \end{aligned}$$

The subalgebra of \(\mathfrak {gl}(m|2n)\) that leaves the bilinear form \(\tilde{\mathsf {B}}\) invariant is isomorphic to \(\mathfrak {osp}(m|2n)\).

1.4 B.4. The exceptional embedding of \(\mathfrak {osp}(1|2)\oplus \mathfrak {osp}(1|2)\) in \(\mathfrak {gosp}(2|4)\)

We consider the realization of \(\mathfrak {osp}(2|4)\) given in Appendix  B.2. Set

$$\begin{aligned} g:=\begin{bmatrix} 0&\sqrt{-1}&0&0&0&0\\ -\sqrt{-1}&0&0&0&0&0\\ 0&0&0&0&0&-\sqrt{-1}\\ 0&0&0&0&\sqrt{-1}&0\\ 0&0&0&-\sqrt{-1}&0&0\\ 0&0&\sqrt{-1}&0&0&0 \end{bmatrix}. \end{aligned}$$

We set \(\mathfrak {k}^\mathrm {ex}\) to be the subalgebra of fixed points of the map \(\mathfrak {osp}(2|4)\rightarrow \mathfrak {osp}(2|4)\) given by \(x\mapsto \mathrm {Ad}_g(x)\). One can verify that \(\mathfrak {k}^\mathrm {ex}\cong \mathfrak {osp}(1|2)\oplus \mathfrak {osp}(1|2)\). We will consider \(\mathfrak {k}^\mathrm {ex}\) as a subalgebra of \(\mathfrak {gosp}(2|4)\).

1.5 B.5. The Lie superalgebra \(\mathfrak {p}(n)\)

Let \(n\ge 1\) be an integer, and let \(\check{\mathsf {B}}:{\mathbb {C}}^{n|n}\times {\mathbb {C}}^{n|n}\rightarrow {\mathbb {C}}\) be the odd supersymmetric bilinear form defined by

$$\begin{aligned} \check{\mathsf {B}}(\mathsf {e}_i,\mathsf {e}_j'):=\delta _{i,j},\ \check{\mathsf {B}}(\mathsf {e}_i,\mathsf {e}_j):=0, \text { and } \check{\mathsf {B}}(\mathsf {e}'_i,\mathsf {e}'_j):=0, \end{aligned}$$

where \(\{\mathsf {e}_i\}_{i=1}^n \cup \{\mathsf {e}_i'\}_{i=1}^n\) is the standard homogeneous basis of \({\mathbb {C}}^{n|n}\). The Lie superalgebra \(\mathfrak {p}(n)\) is the subalgebra of \(\mathfrak {gl}(n|n)\) that leaves \(\check{\mathsf {B}}(\cdot ,\cdot )\) invariant. It consists of matrices in (nn)-block form

$$\begin{aligned} \begin{bmatrix} A&B\\ C&-A^T \end{bmatrix},\text { where }B=B^T\text { and }C=-C^T. \end{aligned}$$

In this paper we will not need a description of the root system and highest weight modules of \(\mathfrak {p}(n)\).

1.6 B.6. The Lie superalgebra \(\mathfrak {q}(n)\)

Let \(n\ge 1\) be an integer. The Lie superalgebra \(\mathfrak {q}(n)\) is the subalgebra of \(\mathfrak {gl}(n|n)\) that consists of matrices in (nn)-block form

$$\begin{aligned} \begin{bmatrix} A&B\\ B&A \end{bmatrix}. \end{aligned}$$

Let \(\mathfrak {h}\) be the subalgebra of matrices of the latter form where A and B are diagonal. Then \(\mathfrak {h}\) is a Cartan subalgebra of \(\mathfrak {q}(n)\). The standard characters \(\left\{ \varepsilon _i\right\} _{i=1}^n\) of \(\mathfrak {h}_{\overline{\mathbf {0}}}\) are the restrictions of the corresponding standard characters of \(\mathfrak {gl}(n|n)\). Let \(\mathfrak {b}^{\mathrm {st}}_n\) (respectively, \(\mathfrak {b}^{\mathrm {op}}_n\)) be the Borel subalgebra of \(\mathfrak {q}(n)\) associated to the fundamental system \( {\Pi }^{\mathrm {st}}:= \{\varepsilon _i-\varepsilon _{i+1}\}_{i=1}^{n-1}\) (respectively, \( {\Pi }^{\mathrm {op}}:= \{\varepsilon _{i+1}-\varepsilon _{i}\}_{i=1}^{n-1}\)). For every partition \(\lambda \in \mathcal {DP}(n)\), we set \(\lambda ^{\mathrm {st}}_n:=\sum _{i=1}^n\lambda _i\varepsilon _i\).

Appendix C. Facts from supergeometry

All of the supermanifolds that are considered in this appendix are complex analytic. We denote the underlying complex manifold of a supermanifold \(\mathcal {X}\) by \(|\mathcal {X}|\), and the sheaf of superfunctions on \(\mathcal {X}\) by \(\mathcal {O}_{{\mathcal {X}}}\). Morphisms of supermanfolds are expressed as \((\mathsf {f},\mathsf {f}^\#):{\mathcal {X}}\rightarrow {\mathcal {Y}}\), where \(\mathsf {f}:|{\mathcal {X}}|\rightarrow |{\mathcal {Y}}|\) is the complex analytic map between the underlying spaces and \(\mathsf {f}^\#:\mathcal {O}_{{\mathcal {Y}}}\rightarrow \mathsf {f}_*\mathcal {O}_{{\mathcal {X}}}\) is the associated morphism of sheaves of superalgebras.

Let \({\mathcal {L}}\) be a connected Lie supergroup and let \({\mathcal {M}}\) be a Lie subsupergroup of \({\mathcal {L}}\). Set \(\mathfrak {l}:=\mathrm {Lie}({\mathcal {L}})\) and \(\mathfrak {m}:=\mathrm {Lie}({\mathcal {M}})\). The right action of \({\mathcal {L}}\) on \({\mathcal {L}}\) induces a canonical isomorphism of superalgebras from \(\mathbf {U}(\mathfrak {l})\) onto the algebra of left invariant holomorphic differential operators on \({\mathcal {L}}\). Under this isomorphism elements of \(\mathbf {U}(\mathfrak {l})^{\mathcal {M}}\), the subalgebra of \({\mathcal {M}}\)-invariants in \(\mathbf {U}(\mathfrak {l})\), are mapped to holomorphic differential operators which are left \({\mathcal {L}}\)-invariant and right \({\mathcal {M}}\)-invariant. The latter differential operators induce \({\mathcal {L}}\)-invariant differential operators on the homogeneous space \({\mathcal {L}}/{\mathcal {M}}\). Consequently, we obtain a homomorphism of superalgebras

$$\begin{aligned} \Psi _{{\mathcal {L}},{\mathcal {M}}}:\mathbf {U}(\mathfrak {l})^{{\mathcal {M}}}\rightarrow \mathscr {D}({\mathcal {L}}/{\mathcal {M}}), \end{aligned}$$
(61)

where \(\mathscr {D}({\mathcal {L}}/{\mathcal {M}})\) denotes the algebra of \({\mathcal {L}}\)-invariant differential operators on \({\mathcal {L}}/{\mathcal {M}}\). By a superization of the argument of [20, Prop. 9.1], we obtain the following statement.

Proposition C.1

Let \({\mathscr {D}}^{(d)}({\mathcal {L}}/{\mathcal {M}})\) denote the subspace of elements of \({\mathscr {D}}({\mathcal {L}}/{\mathcal {M}})\) of order at most d. Assume that there exists an \({\mathcal {M}}\)-invariant complement of \(\mathfrak {m}\) in \(\mathfrak {l}\). Then

$$\begin{aligned} \Psi _{{\mathcal {L}},{\mathcal {M}}}\left( \mathbf {U}^{(d)}(\mathfrak {l})^{\mathcal {M}}\right) = {\mathscr {D}}^{(d)}({\mathcal {L}}/{\mathcal {M}}) \text { for every } d\ge 0. \end{aligned}$$

In the rest of this appendix we will assume that J is a Jordan superalgebra of type \(\mathsf {A}\). Let \(\mathfrak {g}\), \(\mathfrak {k}\), and V be as in Sect. 1.

Lemma C.2

There is a vector \(v_{\mathfrak {k}}\in V_{\overline{\mathbf {0}}}\) such that \(\mathfrak {k}=\mathrm {stab}_{\mathfrak {g}}(v_{\mathfrak {k}})\).

Proof

In all of the cases where J is of type \(\mathsf {A}\) we have \(V\cong V^*\cong J\) as \(\mathfrak {k}\)-modules, hence we can set \(v_{\mathfrak {k}}\) equal to the element of \(V_{\overline{\mathbf {0}}}\) corresponding to \(1_J\in J\). \(\square \)

Lemma C.3

The map \(\mathfrak {g}\rightarrow V\), \(x\mapsto x\cdot v_{\mathfrak {k}}\) is surjective.

Proof

The kernel of the linear map \(x\mapsto x\cdot v_{\mathfrak {k}}\) is \(\mathfrak {k}\). The statement now follows in all of the cases by verifying that the graded dimension of the image of this map and of V are the same. \(\square \)

Let \(\tilde{\mathfrak {b}}:=\tilde{\mathfrak {h}}\oplus \tilde{\mathfrak {n}}\) be a Borel subalgebra of \(\mathfrak {g}\) such that \(\mathfrak {g}=\tilde{\mathfrak {b}}+\mathfrak {k}\). Let \({\mathcal {G}}\) be a complex Lie supergroup such that \(\mathrm {Lie}({\mathcal {G}})=\mathfrak {g}\), and let \({\mathcal {V}}\) be the complex affine superspace corresponding to V. We assume that \(|{\mathcal {G}}|\) is a connected Lie group, and that the action of \(\mathfrak {g}\) on V can be globalized to an action of \({\mathcal {G}}\) on \({\mathcal {V}}\). The stabilizer of \(v_{\mathfrak {k}}\in |{\mathcal {V}}|= V_{\overline{\mathbf {0}}}\) is a complex Lie supergroup \(({\mathcal {K}},\mathcal {O}_{\mathcal {K}})\) such that \(\mathrm {Lie}({\mathcal {K}})=\mathfrak {k}\).

Proposition C.4

The orbit map of \(v_{\mathfrak {k}}\) factors through an embedding \((\mathsf {p}_{v_{\mathfrak {k}}},\mathsf {p}_{v_{\mathfrak {k}}}^\#):{\mathcal {G}}/{\mathcal {K}}\hookrightarrow {\mathcal {V}}\) whose image is an open subsupermanifold of \({\mathcal {V}}\).

Proof

This follows from the fact that the differential of the orbit map \((\mathsf {p}_{v_{\mathfrak {k}}},\mathsf {p}_{v_{\mathfrak {k}}}^\#)\) is a bijection for all \(g\in {\mathcal {G}}\), which is a consequence of Lemma C.3 and \({\mathcal {G}}\)-equivariance of \((\mathsf {p}_{v_{\mathfrak {k}}},\mathsf {p}_{v_{\mathfrak {k}}}^\#)\). \(\square \)

Remark C.5

Using the embedding \({\mathcal {G}}/{\mathcal {K}}\hookrightarrow {\mathcal {V}}\) of Proposition C.4 and the natural injection \(\mathscr {P}(V)\hookrightarrow \mathcal {O}_{{\mathcal {V}}}(|{\mathcal {V}}|)\), we obtain a \({\mathcal {G}}\)-equivariant embedding

$$\begin{aligned} \mathsf {p}_{v_{\mathfrak {k}}}^\#(|{\mathcal {V}}|)\big |_{\mathscr {P}(V)}: \mathscr {P}(V)\hookrightarrow \mathcal {O}_{{\mathcal {G}}/{\mathcal {K}}}\left( |{\mathcal {G}}/{\mathcal {K}}|\right) . \end{aligned}$$

Furthermore, connectedness of \(|{\mathcal {G}}|\) implies \(\mathscr {PD}(V)^{\mathfrak {g}}=\mathscr {PD}(V)^{\mathcal {G}}\). Therefore we can restrict every \(D\in \mathscr {PD}(V)^{\mathfrak {g}}\) to the open subsupermanifold \({\mathcal {G}}/{\mathcal {K}}\) of \({\mathcal {V}}\), and indeed \(D\big |_{{\mathcal {G}}/{\mathcal {K}}}\in \mathscr {D}({\mathcal {G}}/{\mathcal {K}})\).

For the next proposition, recall that every linear functional \(\varphi :\tilde{\mathfrak {h}}\rightarrow {\mathbb {C}}\) induces a natural homomorphism of \({\mathbb {C}}\)-algebras \(\mathcal {S}\big (\tilde{\mathfrak {h}}\big )\rightarrow {\mathbb {C}}\) given by \(x_1\cdots x_k\mapsto \varphi (x_1)\cdots \varphi (x_k)\) for \(x_1,\ldots ,x_k\in \tilde{\mathfrak {h}}\). Set \(\mathcal {S}^{(d)}(\tilde{\mathfrak {h}}):=\bigoplus _{i=0}^d\mathcal {S}^i\big (\tilde{\mathfrak {h}}\big )\).

Proposition C.6

Assume that \(\mathscr {P}(V)\) is a completely reducible and multiplicity-free \(\mathfrak {g}\)-module, and let \(D\in \mathscr {PD}^{(d)}(V)^{\mathfrak {g}}\). Then there exists an element \(x_D\in \mathcal {S}^{(d)}\big (\tilde{\mathfrak {h}}\big )\) such that for every irreducible \(\mathfrak {g}\)-module \(W\subseteq \mathscr {P}(V)\), the action of D on W is by the scalar \({\tilde{\lambda }}(x_D)\), where \({\tilde{\lambda }}\) is the \(\tilde{\mathfrak {b}}\)-highest weight of W.

Proof

Let \({\mathcal {G}}\), \({\mathcal {V}}\), and \({\mathcal {K}}\) be defined as above, and let \((\mathsf {q},\mathsf {q}^\#):{\mathcal {G}}\rightarrow {\mathcal {G}}/{\mathcal {K}}\) be the canonical quotient map. Then \(\mathsf {q}^\#(|{\mathcal {G}}/{\mathcal {K}}|):\mathcal {O}_{{\mathcal {G}}/{\mathcal {K}}}(|{\mathcal {G}}/{\mathcal {K}}|)\rightarrow \mathcal {O}_{{\mathcal {G}}}(G)\) is an injection. For any \(\tilde{X}\in \mathbf {U}(\mathfrak {g})\), let \(\mathrm {L}_{{\tilde{X}}}\) (respectively, \(\mathrm {R}_{{\tilde{X}}}\)) denote the action of \({\tilde{X}}\) on \(\mathcal {O}_{\mathcal {G}}(G)\) by left invariant (respectively, right invariant) differential operators. By Proposition C.1, there exists \({\tilde{D}}\in \mathbf {U}^{(d)}(\mathfrak {g})^{{\mathcal {K}}}\) such that \(\mathsf {q}^\#(|{\mathcal {G}}/{\mathcal {K}}|)(Df)= \mathrm {L}_{{\tilde{D}}} \mathsf {q}^\#(|{\mathcal {G}}/{\mathcal {K}}|)(f)\) for every \(f\in \mathcal {O}_{{\mathcal {G}}/{\mathcal {K}}}(|{\mathcal {G}}/{\mathcal {K}}|)\).

Now set \(f:=\mathsf {p}_{v_{\mathfrak {k}}}^\#(|V|)(\phi _{{\tilde{\lambda }}})\) where \(\phi _{{\tilde{\lambda }}}\in \mathscr {P}(V)\) is a highest weight vector of W, and let \({\tilde{f}}:=\mathsf {q}^\#(|{\mathcal {G}}/{\mathcal {K}}|)(f)\). Let \({\mathcal {N}}\) be the connected Lie subsupergroup of \({\mathcal {G}}\) such that \(\mathrm {Lie}({\mathcal {N}})=\tilde{\mathfrak {n}}\). Then \({\tilde{f}}\) is left \({\mathcal {N}}\)-invariant and right \({\mathcal {K}}\)-invariant. We can express \({\tilde{D}}\) in the form \({\tilde{D}}=D_1+D_2+D_3\), where

$$\begin{aligned} D_1\in \tilde{\mathfrak {n}}\mathbf {U}^{(d-1)}(\mathfrak {g}), \ D_2\in \mathbf {U}^{(d)}\big (\tilde{\mathfrak {h}}\big ), \text { and } D_3\in \mathbf {U}^{(d-1)}(\mathfrak {g})\mathfrak {k}. \end{aligned}$$

From \({\mathcal {K}}\)-invariance of f it follows that \(\mathrm {L}_{D_3}{\tilde{f}}=0\). Furthermore, we can write \(D_1\) as a sum of elements of the form \(XD'\) where \(X\in \tilde{\mathfrak {n}}\) and \(D'\in \mathbf {U}(\mathfrak {g})\). Let \({\mathcal {H}}:=(H,\mathcal {O}_{{\mathcal {H}}})\) denote the connected Lie subsupergroup of \({\mathcal {G}}\) such that \(\mathrm {Lie}({\mathcal {H}})=\tilde{\mathfrak {h}}\). For \(h\in H\) we have

$$\begin{aligned} \mathrm {L}_{XD'}{\tilde{f}}(h) =\mathrm {L}_X\left( \mathrm {L}_{D'}{\tilde{f}} \right) (h)= \mathrm {R}_{-\mathrm {Ad}_hX} \left( \mathrm {L}_{D'}{\tilde{f}} \right) (h). \end{aligned}$$

Since \(\mathrm {L}_{D'}{\tilde{f}}\) is left \({\mathcal {N}}\)-invariant and \(\mathrm {Ad}_h\left( \tilde{\mathfrak {n}}\right) \subseteq \tilde{\mathfrak {n}}\), it follows that \(\mathrm {L}_{XD'}{\tilde{f}}(h)=0\). Consequently, we have shown that for \(x_D:=D_2\),

$$\begin{aligned} \mathrm {L}_{{\tilde{D}}}{\tilde{f}}(h)=\mathrm {L}_{D_2}{\tilde{f}}(h)=\tilde{\lambda }\big (D_2\big ){\tilde{f}}(h) ={\tilde{\lambda }}(x_D){\tilde{f}}(h) . \end{aligned}$$

It remains to prove that \({\tilde{f}}\ne 0\). To this end, note that the canonical multiplication morphism

$$\begin{aligned} {\mathcal {N}}\times {\mathcal {H}}\times {\mathcal {K}}\rightarrow {\mathcal {G}}\end{aligned}$$

is a local isomorphism at the identity element. Thus, from analyticity, left \({\mathcal {N}}\)-invariance, and right \({\mathcal {K}}\)-invariance of \({\tilde{f}}\), it follows that \({\tilde{f}}\big |_{H}\) is not identically zero. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sahi, S., Salmasian, H. & Serganova, V. The Capelli eigenvalue problem for Lie superalgebras. Math. Z. 294, 359–395 (2020). https://doi.org/10.1007/s00209-019-02289-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-019-02289-7

Navigation