Skip to main content
Log in

Quantization of Hamiltonian loop group spaces

  • Published:
Mathematische Annalen Aims and scope Submit manuscript

Abstract

We prove a Fredholm property for spin-c Dirac operators \(\mathsf {D}\) on non-compact manifolds satisfying a certain condition with respect to the action of a semi-direct product group \(K\ltimes \Gamma \), with K compact and \(\Gamma \) discrete. We apply this result to an example coming from the theory of Hamiltonian loop group spaces. In this context we prove that a certain index pairing \([{\mathcal {X}}] \cap [\mathsf {D}]\) yields an element of the formal completion \(R^{-\infty }(T)\) of the representation ring of a maximal torus \(T \subset H\); the resulting element has an additional antisymmetry property under the action of the affine Weyl group, indicating \([{\mathcal {X}}] \cap [\mathsf {D}]\) corresponds to an element of the ring of projective positive energy representations of the loop group.

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. This is true if A, B are separable, as we are assuming.

  2. Under some conditions on A (in particular if \(A={\mathbb {C}}\)), \(\text{ KK }(A,-)\) is countably additive, cf. [9] Sect.  23.15.

  3. Or indeed any of the summands.

  4. In [28] we actually worked with a slightly larger space \({\mathcal {P}}H\) (a principal H-bundle over \(L{\mathfrak {h}}^*\)), which was desirable for certain purposes, although we have avoided it here for simplicity. The embedding referred to here can be constructed by the same method.

  5. Compare to [16] for example, where this condition is called ‘topological regularity’.

  6. Here and wherever possible below we have written E instead of \(\pi ^*E\).

References

  1. Alekseev, A., Malkin, A., Meinrenken, E.: Lie group valued moment maps. J. Differen. Geom. 48, 445–495 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  2. Alekseev, A., Meinrenken, E.: Dirac structures and Dixmier–Douady bundles. Int. Math. Res. Not. 2012, 904–956 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  3. Alekseev, A., Meinrenken, E., Woodward, C.: The Verlinde formulas as fixed point formulas. J. Symplectic Geom. 1, 1–46 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  4. Anghel, N.: An abstract index theorem on non-compact Riemannian manifolds. Houston J. of Math. 19, 223–237 (1993)

    MathSciNet  MATH  Google Scholar 

  5. Atiyah, M.F.: Elliptic Operators and Compact Groups, vol. 401. Springer, Berlin, Heildelberg, New York (2006)

    Google Scholar 

  6. Baaj, S., Julg, P.: Théorie bivariante de Kasparov et opérateurs non bornés dans les C\(^\ast \)-modules hilbertiens. CR Acad. Sci. Paris Sér. I Math 296(21), 875–878 (1983)

    MATH  Google Scholar 

  7. Bär, C., Ballmann, W.: Guide to elliptic boundary value problems for dirac-type operators. In: Ballmann, W., Blohmann, C., Faltings, G., Teichner, P., Zagier, D. (eds) Arbeitstagung Bonn 2013. Progress in Mathematics, vol 319. Birkhuser, Cham (2016)

  8. Baum, P., Guentner, E., Willett R.: Exactness and the Kadison-Kaplansky Conjecture. In: Doran, R., Park, E. (eds.) Operator algebras and their applications: a tribute to Richard V. Kadison. Contemporary Mathematics, vol 671. American Mathematical Society, Providence, Rhode Island (2016)

  9. Blackadar, B.: K-Theory for Operator Algebras, 2nd edn. Cambridge University Press, Cambridge (1998)

    MATH  Google Scholar 

  10. Blackadar, B.: Operator algebras. Theory of \(C^\ast \)-algebras and von Neumann algebras. Encyclopedia of Mathematical Sciences, vol. 122. Springer-Verlag, Berlin Heidelberg (2006)

  11. Bott, R., Tolman, S., Weitsman, J.: Surjectivity for Hamiltonian loop group spaces. Invent. Math. 155, 225–251 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  12. Braverman, M.: Index theorem for equivariant Dirac operators on noncompact manifolds. K-theory 27(1), 61–101 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Bunke, U.: A K-theoretic relative index theorem and Callias-type Dirac operators. Math. Annalen 303(1), 241–279 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  14. Chernoff, P.R.: Essential self-adjointness of powers of generators of hyperbolic equations. J. Funct. Anal. 12(4), 401–414 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  15. Chernoff, P.R.: Schrödinger and Dirac operators with singular potentials and hyperbolic equations. Pac. J. Math. 72(2), 361–382 (1977)

    Article  MATH  Google Scholar 

  16. Freed, D.S., Hopkins, M.J., Teleman, C.: Loop groups and twisted K-theory III. Ann. Math. 4(4), 947–1007 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  17. Freed, D.S., Hopkins, M.J., Teleman, C.: Loop groups and twisted K-theory II. J. Am. Math. Soc. 26, 595–644 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  18. Guillemin, V., Sternberg, S.: Geometric quantization and multiplicities of group representations. Invent. Math. 67(3), 515–538 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  19. Higson, N.: A primer on KK-theory. Operator theory: operator algebras and applications 1, 239–283 (1990)

    MathSciNet  Google Scholar 

  20. Higson, N., Roe, J.: Analytic K-Homology. Oxford University Press, Oxford (2000)

    MATH  Google Scholar 

  21. Hochs, P., Song, Y.: Equivariant indices of Spin-c Dirac operators for proper moment maps. Duke Math. J. 166, 1125–1178 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  22. Kac, V.: Infinite-Dimensional Lie Algebras. Cambridge University Press, Cambridge (1994)

    Google Scholar 

  23. Kasparov, G.: Equivariant KK-theory and the Novikov conjecture. Invent. Math. 91, 147–201 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  24. Kucerovsky, D.: The KK-product of unbounded modules. K-theory 11(1), 17–34 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  25. Kucerovsky, D.: A short proof of an index theorem. In: Proceedings of the Am. Math. Soc. 3729–3736 (2001)

  26. Loizides, Y.: Geometric K-homology and the Freed-Hopkins-Teleman theorem. arXiv:1804.05213

  27. Loizides, Y.: Norm-square localization for Hamiltonian LG-spaces. J. Geom. Phys. 114, 420–449 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  28. Loizides, Y., Meinrenken, E., Song, Y.: Spinor modules for Hamiltonian loop group spaces. arXiv:1706.07493

  29. Loizides, Y., Song, Y.: Witten deformation for Hamiltonian loop group spaces. arXiv:1810.02347

  30. Meinrenken, E.: Symplectic surgery and the Spin-c Dirac operator. Adv. Math. 134, 240–277 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  31. Meinrenken, E.: Twisted K-homology and group-valued moment maps. Int. Math. Res. Not. 2012, 4563–4618 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  32. Meinrenken, E., Woodward, C.: Hamiltonian loop group actions and Verlinde factorization. J. Diff. Geom. 50, 417–469 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  33. Mislin, G., Valette, A.: Proper Group Actions and the Baum–Connes Conjecture. Birkhäuser, Basel (2012)

    MATH  Google Scholar 

  34. Paradan, P.-E.: Localization of the Riemann–Roch character. J. Fun. Anal. 187, 442–509 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  35. Paradan, P.-E., Vergne, M.: Equivariant Dirac operators and differentiable geometric invariant theory. arXiv:1411.7772

  36. Paradan, P.-E., Vergne, Michele: Witten non abelian localization for equivariant K-theory, and the \([Q, R]= 0\) theorem. arXiv:1504.07502

  37. Phillips, N.C.: An introduction to crossed product \(C^\ast \)-algebras and minimal dynamics. Lecture notes. http://pages.uoregon.edu/ncp/

  38. Pressley, A., Segal, G.B.: Loop groups. Clarendon Press, Oxford (1986)

    MATH  Google Scholar 

  39. Reed, M., Simon, B.: Methods of modern mathematical physics. vol. 4: Analysis of operators. Academic Press, New York-London (1978)

  40. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. Functional Analysis, vol. 1. Academic Press, New York (1980)

    MATH  Google Scholar 

  41. Shubin, M.A.: Spectral Theory of the Schrödinger Operators on Non-compact Manifolds: Qualitative Results. Spectral Theory and Geometry, pp. 226–283. Cambridge University Press, Cambridge (1999). (Edinburgh, 1998)

    MATH  Google Scholar 

  42. Song, Y.: Dirac operators on quasi-Hamiltonian G-spaces. J. Geom. Phys. 106, 70–86 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  43. Takata, D.: An analytic LT-equivariant index and noncommutative geometry. arXiv:1701.06055

  44. Takata, D.: LT-equivariant index from the viewpoint of KK-theory. arXiv:1709.06205

  45. Tian, Y., Zhang, W.: An analytic proof of the geometric quantization conjecture of Guillemin-Sternberg. Invent. Math. 132, 229–259 (1998)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We especially thank Eckhard Meinrenken for many helpful discussions and encouragement, and for providing feedback on an earlier draft. We also thank Nigel Higson for helpful discussions. Y. Song is supported by NSF grant 1800667.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yiannis Loizides.

Additional information

Communicated by Thomas Schick.

Appendices

Appendix A. Group automorphisms and the descent map

Throughout this section G will be a locally compact group equipped with a left-invariant Haar measure, \(\sigma \in \text{ Aut }(G)\) will be a group automorphism preserving the Haar measure, and \(\langle \sigma \rangle \ltimes G\) the semi-direct product group. We write \(g \mapsto g^\sigma \) for the action of \(\sigma \) on \(g \in G\).

Let A be a \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebra; A is also a G-\(C^*\) algebra by restricting the group action. Define the G-\(C^*\) algebra \(A^\sigma \) to be the \(C^*\) algebra A equipped with the new G-action \(\pi ^\sigma (g)=\pi (g^\sigma )\), where \(\pi \) is the G-action on A. Since A is a \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebra, there is an action map

$$\begin{aligned} \alpha ^A_\sigma :A \rightarrow A. \end{aligned}$$

Viewed as a map \(A \rightarrow A^\sigma \), \(\alpha ^A_\sigma \) is an isomorphism of G-\(C^*\) algebras. In particular, it induces maps

$$\begin{aligned} (\alpha ^A_\sigma )_*:\text{ K }_0^G(A) \rightarrow \text{ K }_0^G(A^\sigma ), \qquad (\alpha ^A_\sigma )^*:\text{ K }^0_G(A^\sigma ) \rightarrow \text{ K }^0_G(A). \end{aligned}$$

Below we usually omit the \(*\) from the notation.

For any group homomorphism \(\sigma :G_1 \rightarrow G_2\) one has a restriction homomorphism (cf. [23])

$$\begin{aligned} \sigma :\text{ KK }_{G_2}(A,B) \rightarrow \text{ KK }_{G_1}(A,B), \end{aligned}$$

where A, B become \(G_1\)-\(C^*\) algebras by pre-composing the \(G_2\) action with \(\sigma \). As a special case, if \(\sigma \in \text{ Aut }(G)\) is a group automorphism, one obtains a map

$$\begin{aligned} \sigma :\text{ KK }_G(A,B) \rightarrow \text{ KK }_G(A^\sigma ,B^\sigma ). \end{aligned}$$

Definition A.1

Let A, B be \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebras, with \(\alpha _\sigma ^A\), \(\alpha _\sigma ^B\) the corresponding action maps. Define

$$\begin{aligned} \tau _\sigma :\text{ KK }_G(A,B) \rightarrow \text{ KK }_G(A,B) \end{aligned}$$

to be the composition

$$\begin{aligned} \text{ KK }_G(A,B) \xrightarrow {\alpha _\sigma ^B} \text{ KK }_G(A,B^\sigma ) \xrightarrow {(\alpha _\sigma ^A)^{-1}} \text{ KK }_G(A^\sigma ,B^\sigma ) \xrightarrow {\sigma ^{-1}} \text{ KK }_G(A,B). \end{aligned}$$

In general \(\tau _\sigma \) is not the identity map (we saw an example in Sect.  4.4), but it does act as the identity on the image of the restriction map \(\text{ KK }_{\langle \sigma \rangle \ltimes G}(A,B) \rightarrow \text{ KK }_G(A,B)\).

We next describe the relationship between \(\tau _\sigma \) and the descent map \(j_G\). The G-equivariant \(^*\)-homomorphism \(\alpha ^A_\sigma :A \rightarrow A^\sigma \) induces a \(^*\)-homomorphism

$$\begin{aligned} \alpha ^A_\sigma :G \ltimes A \rightarrow G \ltimes A^\sigma \end{aligned}$$
(27)

(on \(C_c(G,A)\) it is given by applying \(\alpha ^A_\sigma \) point-wise). Recall that a \(^*\)-homomorphism \(\alpha :A \rightarrow B\) determines an element \(\alpha \in \text{ KK }(A,B)\), such that push-forward (resp. pull-back) by \(\alpha \) in K-theory (resp. K-homology) are given by an appropriate \(\text{ KK }\)-product. Thought of in this way, the corresponding element in \(\text{ KK }(G\ltimes A,G\ltimes A^\sigma )\) is the image of \(\alpha ^A_\sigma \in \text{ KK }_G(A,A^\sigma )\) under the descent map \(j_G\).

The group automorphism \(\sigma :G \rightarrow G\) also induces a \(^*\)-homomorphism

$$\begin{aligned} \sigma ^A :G \ltimes A \rightarrow G \ltimes A^\sigma , \end{aligned}$$
(28)

given on \(C_c(G,A)\) by \(a \mapsto a^\sigma \), where \(a^\sigma (g):=a(g^\sigma )\).

Proposition A.2

Let A,B be \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebras. For any \(\mathsf {x} \in \text{ KK }_G(A^\sigma ,B^\sigma )\) one has the following equality in \(\text{ KK }(G\ltimes A,G\ltimes B)\):

$$\begin{aligned} j_G(\sigma ^{-1}(\mathsf {x}))=\sigma ^A \otimes j_G(\mathsf {x})\otimes (\sigma ^B)^{-1}. \end{aligned}$$

Proof

Let \(({\mathcal {E}},\rho ,F)\) be a cycle representing \(\mathsf {x}\), thus in particular \({\mathcal {E}}\) is an \((A^\sigma ,B^\sigma )\)-bimodule with \(B^\sigma \)-valued inner product \((\cdot ,\cdot )\). To avoid confusion between the G-\(C^*\) algebras B and \(B^\sigma \), we will write all formulas in terms of the action map \((g^\prime ,b)\mapsto g^\prime .b\) for B, not\(B^\sigma \). Thus, for example, the \(G\ltimes B^\sigma \)-valued inner product for \(j_G(\mathsf {x})\) is expressed as

$$\begin{aligned} (e_1,e_2)_{G\ltimes B^\sigma }(g)=\int _G (g_1^\sigma )^{-1}.\big (e_1(g_1),e_2(g_1g)\big )\,\,dg_1, \end{aligned}$$

i.e. \((g_1^\sigma )^{-1}\) appears in the formula instead of \(g_1^{-1}\). The \((G\ltimes A,G\ltimes B)\)-bimodule structure for \(\sigma ^A \otimes j_G(\mathsf {x})\otimes (\sigma ^B)^{-1}\) is given by the formulas

$$\begin{aligned} (a\cdot e)(g)=\int _G a(g_1^\sigma )g_1.e(g_1^{-1}g)\,\,dg_1, \qquad (e\cdot b)(g)=\int _G e(g_1)g_1^\sigma .b\big ((g_1^\sigma )^{-1}g^\sigma \big )\,\,dg_1, \end{aligned}$$

and the \(G\ltimes B\)-valued inner product is

$$\begin{aligned} (e_1,e_2)_{G\ltimes B}(g)=\int _G (g_1^\sigma )^{-1}.\big (e_1(g_1),e_2(g_1g^{\sigma ^{-1}})\big )\,\, dg_1. \end{aligned}$$

On the other hand, the \((G\ltimes A,G\ltimes B)\)-bimodule structure for \(j_G(\sigma ^{-1}(\mathsf {x}))\) is given by the formulas

$$\begin{aligned} (a \star e)(g)=\int _G a(g_1)g_1^{\sigma ^{-1}}.e(g_1^{-1}g)\,\,dg_1, \qquad (e \star b)(g)=\int _G e(g_1)g_1.b(g_1^{-1}g)\,\,dg_1, \end{aligned}$$

and the \(G\ltimes B\)-valued inner product is

$$\begin{aligned} (e_1,e_2)^{\star }_{G\ltimes B}(g)=\int _G g_1^{-1}.\big (e_1(g_1),e_2(g_1g)\big )\,\,dg_1. \end{aligned}$$

The formulas are not the same, but there is a linear map

$$\begin{aligned} e \in C_c(G,{\mathcal {E}}) \mapsto e^\sigma \in C_c(G,{\mathcal {E}}), \qquad e^\sigma (g):=e(g^\sigma ) \end{aligned}$$

which intertwines the bimodule structures and \(C_c(G,B)\)-valued inner products, i.e. \((a \star e)^\sigma =a\cdot e^\sigma \), \((e \star b)^\sigma =e^\sigma \cdot b\) and \((e_1^\sigma ,e_2^\sigma )_{G\ltimes B}=(e_1,e_2)^{\star }_{G\ltimes B}\) (one uses the \(\sigma \)-invariance of the Haar measure). Thus, the map extends to an isometric isomorphism between the completions, intertwining the bimodule structures. \(\square \)

Definition A.3

Let A be a \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebra. Let

$$\begin{aligned} \tau ^A_\sigma =(\alpha ^A_\sigma )^{-1} \circ \sigma ^A :G\ltimes A \rightarrow G\ltimes A, \end{aligned}$$

where \(\alpha ^A_\sigma \) (resp. \(\sigma ^A\)) is as in (27) (resp. (28)). Then \(\tau ^A_\sigma \) is an automorphism of \(G\ltimes A\). On \(C_c(G,A)\) it is given by the formula

$$\begin{aligned} \tau ^A_\sigma (f)(g)=(\alpha ^A_\sigma )^{-1}(f(g^\sigma )). \end{aligned}$$

The corresponding element of \(\text{ KK }(G\ltimes A,G\ltimes A)\) is the \(\text{ KK }\)-product

$$\begin{aligned} \tau ^A_\sigma =\sigma ^A \otimes (\alpha ^A_\sigma )^{-1} \in \text{ KK }(G\ltimes A,G\ltimes A). \end{aligned}$$

The following is a corollary of Proposition A.2.

Corollary A.4

Let A, B be \(\langle \sigma \rangle \ltimes G\)-\(C^*\) algebras. For any \(\mathsf {x} \in \text{ KK }_G(A,B)\),

$$\begin{aligned} j_G(\tau _\sigma (\mathsf {x}))=\tau ^A_\sigma \otimes j_G(\mathsf {x}) \otimes (\tau ^B_\sigma )^{-1}. \end{aligned}$$

Similar to \(\tau _\sigma \), \(\tau ^A_\sigma \) acts trivially on elements which are \(\langle \sigma \rangle \ltimes G\)-equivariant:

Proposition A.5

Let \(\widetilde{G}=\langle \sigma \rangle \ltimes G\). Let A be a \(\widetilde{G}\)-\(C^*\) algebra. Let

$$\begin{aligned} \iota _G :G\ltimes A \rightarrow \widetilde{G} \ltimes A \end{aligned}$$

denote the \(^*\)-homomorphism induced by the open inclusion \(G \hookrightarrow \widetilde{G}\). Then

$$\begin{aligned} \tau ^A_\sigma \otimes \iota _G = \iota _G \in \text{ KK }(G\ltimes A,\widetilde{G} \ltimes A). \end{aligned}$$

Proof

The automorphism \(\tau ^A_\sigma \in \text{ Aut }(G\ltimes A)\) extends to an automorphism of \(\widetilde{G} \ltimes A\), given by the same formula with \(\widetilde{g}^\sigma :=\text{ Ad }_\sigma (\widetilde{g}) \in \widetilde{G}\). Thus

$$\begin{aligned} \iota _G \circ \tau ^A_\sigma =\widetilde{\tau }^A_\sigma \circ \iota _G, \end{aligned}$$

where \(\widetilde{\tau }^A_\sigma \) denotes the extended automorphism. In fact \(\widetilde{\tau }^A_\sigma \) is an inner automorphism, i.e. there exists a unitary \(u_\sigma \) in the multiplier algebra of \(\widetilde{G} \ltimes A\) such that \(\widetilde{\tau }^A_\sigma =\text{ Ad }_{u_\sigma }\), cf. [10, II.10.3.10]. The result follows from the fact that any inner automorphism of a \(C^*\) algebra D induces the identity element in \(\text{ KK }(D,D)\). \(\square \)

Appendix B. Schrödinger-type operators

If A is a self-adjoint operator on a Hilbert space H with domain \(\text{ dom }(A)\) and spectrum in \([1,\infty )\), then one defines an associated positive definite quadratic form

$$\begin{aligned} q_A(u_1,u_2)=(Au_1,u_2) \end{aligned}$$

for all \(u_1,u_2 \in \text{ dom }(A)\). The completion of \(\text{ dom }(A)\) using the inner product \(q_A\) is a Hilbert space \(\text{ dom }(q_A)\) which can be identified with \(\text{ dom }(A^{1/2})\), and is known as the form domain of A (cf. [40, VIII.6]). Given self-adjoint operators AB with spectrum in \([1,\infty )\) one writes

$$\begin{aligned} A \ge B \end{aligned}$$

if

$$\begin{aligned} \text{ dom }(q_A) \subset \text{ dom }(q_B) \quad \text {and} \quad q_A(v,v)\ge q_B(v,v) \quad \forall v \in \text{ dom }(q_A) \end{aligned}$$

(cf. [39, XIII.2, p. 85]). Equivalently, \(A \ge B\) if the inclusion mapping

$$\begin{aligned} (\text{ dom }(q_A),q_A) \hookrightarrow (\text{ dom }(q_B),q_B) \end{aligned}$$

is norm-decreasing. It is enough to check the inequality \(q_A(v,v)\ge q_B(v,v)\) on a core for A. More generally if AB are self-adjoint operators with spectrum in \([-c,\infty )\) for some \(c \ge 0\) then one writes \(A \ge B\) if \(A+c+1 \ge B+c+1\).

Proposition B.1

Let \({\mathcal {Y}}\) be a complete Riemannian manifold. Let \(\mathsf {D} \) be a symmetric \(1^{st}\) order elliptic differential operator with finite propagation speed acting on sections of a Hermitian vector bundle S, and let \(H=L^2({\mathcal {Y}},S)\). Let \(\rho \), V be continuous functions such that \(\rho \) is bounded, V is bounded below, and V is proper on the support of \(\rho \). Let A be a self-adjoint operator with spectrum in \((0,\infty )\), and suppose

$$\begin{aligned} A \ge \mathsf {D} ^2+V. \end{aligned}$$

Then the operator \(\rho A^{-1}\) is compact.

Remark B.2

Consider the special case when V is proper and bounded below. The operator \(\mathsf {D}^2+V\) is a Schrödinger-type operator with potential going to infinity at infinity. In this case, it is known that \(\mathsf {D}^2+V\) has discrete spectrum. Many proofs of this appear in the literature, cf. [41] (for \(\mathsf {D}^2=-\Delta \)), [25] (for a proof based on a method of Gromov-Lawson). It is also closely related to a Fredholm criterion of Anghel [4], and to the property of being ‘\(\kappa \)-coercive’ for all \(\kappa >0\) in [7, Corollary 5.6].

Proof

Note that if \(\rho (A+c)^{-1}\) is compact for some \(c>0\), then so is \(\rho A^{-1}\) since

$$\begin{aligned} \rho A^{-1}-\rho (A+c)^{-1}=c\rho (A+c)^{-1}A^{-1} \end{aligned}$$

and the right hand side is compact since \(\rho (A+c)^{-1}\) is compact and \(A^{-1}\) is bounded. Therefore we may as well assume that \(V \ge 1\) and the spectrum of A is contained in \([1,\infty )\).

The operator \(\mathsf {D}^2+V\) is self-adjoint (cf. [15, Theorem 4.6]) and positive. Let q denote the corresponding quadratic form, defined initially on the domain of \(\mathsf {D}^2+V\) by

$$\begin{aligned} q(u_1,u_2)=\big ((\mathsf {D}^2+V)u_1,u_2\big ) \end{aligned}$$

and then extended to the completion \(H(q)=\text{ dom }(q) \subset H\) of the domain of \(\mathsf {D}^2+V\) with respect to the inner product q. Since

$$\begin{aligned} q(u,u)=\Vert \mathsf {D}u\Vert ^2+\Vert \sqrt{V}u\Vert ^2 \end{aligned}$$

the Hilbert space H(q) can be described more simply as

$$\begin{aligned} H(q)=\{u \in H^1|\sqrt{V}u \in H\}, \end{aligned}$$

where \(H^1:=\text{ dom }(\mathsf {D})\) (a Sobolev space). The operator \(A^{-1}\) defines a bounded linear map \(H \rightarrow \text{ dom }(A)\), where the domain \(\text{ dom }(A)\) of A is equipped with the A-norm \(\Vert u\Vert _A:=\Vert Au\Vert \) (cf. [40, VIII]). The Cauchy-Schwartz inequality and \(q_A(u,u)\ge q(u,u)\) imply that the inclusion maps

$$\begin{aligned} \text{ dom }(A) \subset \text{ dom }(q_A) \hookrightarrow H(q) \end{aligned}$$

are bounded. Thus \(\rho A^{-1}\) factors as a composition of bounded linear maps:

$$\begin{aligned} H \xrightarrow {A^{-1}}\text{ dom }(A) \hookrightarrow H(q) \xrightarrow {M_\rho } H \end{aligned}$$

where \(M_\rho \) denotes the operator given by multiplication by the function \(\rho \). It therefore suffices to show that the map

$$\begin{aligned} M_\rho :H(q) \rightarrow H \end{aligned}$$

is a compact mapping. Without loss of generality assume \(|\rho |\le 1\). For \(n \in {\mathbb {Z}}_{\ge 0}\) let \(W_n=V^{-1}((-\infty ,n])\); by assumption \(\text{ supp }(\rho )\cap W_n\) is compact. For each n, let \(g_n:{\mathcal {Y}}\rightarrow [0,1]\) be a bump function equal to 1 on \(W_n\) and supported in \(W_{n+1}\). Since \(g_n\rho \) has compact support,

$$\begin{aligned} M_{g_n\rho }:H(q) \rightarrow H \end{aligned}$$

is compact, by the Rellich lemma. We claim that the difference

$$\begin{aligned} M_\rho -M_{g_n\rho }=M_{(1-g_n)\rho } \end{aligned}$$

converges to zero in norm, hence \(M_\rho \) is also compact. Indeed, if \(q(u,u)\le 1\) then

$$\begin{aligned} 1 \ge q(u,u) \ge \int _{{\mathcal {Y}}\setminus W_n} n|u|^2 \qquad \Rightarrow \qquad \int _{{\mathcal {Y}}\setminus W_n}|u|^2 \le \tfrac{1}{n}. \end{aligned}$$

Thus

$$\begin{aligned} \Vert M_{(1-g_n)\rho }\Vert ^2=\sup _{q(u,u)\le 1} \, \, \int _{{\mathcal {Y}}} (1-g_n)^2|\rho u|^2 \le \sup _{q(u,u)\le 1} \, \, \int _{{\mathcal {Y}}\setminus W_{n}}|u|^2 \le \tfrac{1}{n}, \end{aligned}$$

which goes to 0 as \(n \rightarrow \infty \). \(\square \)

Remark B.3

More generally, suppose a compact group K acts isometrically on \({\mathcal {Y}}\), commuting with A, \(\mathsf {D}\) and that V is K-invariant. The Hilbert space H decomposes into isotypic components

$$\begin{aligned} H=\bigoplus _{\pi \in \text{ Irr }(K)} H_\pi , \end{aligned}$$

and the operator A decomposes accordingly into self-adjoint operators \(A_\pi \) with \(\text{ dom }(A_\pi )=\text{ dom }(A)\cap H_\pi \subset H_\pi \), and similarly for \(\mathsf {D}^2+V\). Suppose the conditions in Proposition B.1 are satisfied except that the inequality holds only on \(H_\pi \), that is

$$\begin{aligned} A_\pi \ge (\mathsf {D}^2+V)_\pi . \end{aligned}$$

Essentially the same argument, with \(\rho A_\pi ^{-1}\) factored as

$$\begin{aligned} H_\pi \xrightarrow {A_\pi ^{-1}} \text{ dom }(A_\pi ) \hookrightarrow H(q)\xrightarrow {M_\rho } H \end{aligned}$$

shows that \(\rho A_\pi ^{-1}:H_\pi \rightarrow H\) is a compact operator.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Loizides, Y., Song, Y. Quantization of Hamiltonian loop group spaces. Math. Ann. 374, 681–722 (2019). https://doi.org/10.1007/s00208-018-1771-z

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00208-018-1771-z

Navigation