Skip to main content
Log in

T-duality simplifies bulk–boundary correspondence: the noncommutative case

  • Published:
Letters in Mathematical Physics Aims and scope Submit manuscript

Abstract

We state and prove a general result establishing that T-duality, or the Connes–Thom isomorphism, simplifies the bulk–boundary correspondence, given by a boundary map in K-theory, in the sense of converting it to a simple geometric restriction map. This settles in the affirmative several earlier conjectures of the authors and provides a clear geometric picture of the correspondence. In particular, our result holds in arbitrary spatial dimension, in both the real and complex cases, and also in the presence of disorder, magnetic fields, and H-flux. These special cases are relevant both to string theory and to the study of the quantum Hall effect and topological insulators with defects in condensed matter physics.

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

Source: Reproduced with permission from [25, p.  786]

Fig. 2

Similar content being viewed by others

Notes

  1. Quite often KR-theory groups, in the sense of Atiyah’s Real K-theory, are relevant, and these groups are defined for spaces with involutions. Such spaces are usually called “Real spaces”, which should not be confused with our usage of “real space” as synonymous with “position space”.

  2. Although we concentrate here on the case where the boundary has codimension 1, our notation and arguments have been written to suggest a way to include higher-codimensional boundaries. The main differences lie in the parity of the K-groups linked by Connes–Thom maps and the fact that there can be Mackey obstructions for the boundary.

  3. There may be no such symmetries even if the boundary occupies \(n\ge 1\) dimensions; for instance, it may not be completely straight. Generally speaking, the collection of boundary symmetries determine what topological invariants may be associated with it, and the more symmetries there are, the more boundary invariants are available.

  4. If \(\alpha \) defined a character of L, then the “periodicity” would give Bloch-type functions.

  5. \(A_\Theta \) can be defined as the universal \(C^*\)-algebra generated by d unitaries \(U_1,\ldots ,U_d\) subject to \(U_kU_j={\mathrm{exp}}(2\pi i\Theta _{jk})U_jU_k\), where we regard \(\Theta \) as a skew-symmetric \(d\times d\) matrix.

  6. We have switched convention for induced algebras in this Appendix compared to the main text. This is to make closer contact to Paschke’s original work, and the effect is that \(\alpha ^{-1}\) is replaced by \(\alpha \).

  7. There is no loss of generality in assuming that \({\mathcal {A}}\) is unital and that p is in \(\mathrm {Ind}({\mathcal {A}},\alpha )\) rather than in \(\mathrm {Ind}({\mathcal {A}},\alpha )\otimes M_n\), see [37].

References

  1. Atiyah, M.F., Donnelly, H., Singer, I.M.: Eta invariants, signature defects of cusps, and values of L-functions. Ann. Math. 118, 131–177 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bellissard, J., van Elst, A., Schulz-Baldes, H.: The noncommutative geometry of the quantum Hall effect. J. Math. Phys. 35(10), 5373–5451 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Bellissard, J.: \(K\)-theory of \(C^*\)-algebras in solid state physics. In: Statistical Mechanics and Field Theory: Mathematical Aspects (Groningen, 1985), Lecture Notes in Physics, vol. 257, pp. 99–156. Springer, Berlin (1986)

  4. Bouwknegt, P., Evslin, J., Mathai, V.: T-duality: topology change from \(H\)-flux. Commun. Math. Phys. 249, 383–415 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Bramwell, S.T., et al.: Measurement of the charge and current of magnetic monopoles in spin ice. Nature 461, 956–959 (2009)

    Article  ADS  Google Scholar 

  6. Chang, M.-C., Niu, Q.: Berry phase, hyperorbits, and the Hofstadter spectrum: semiclassical dynamics in magnetic Bloch bands. Phys. Rev. B 53(11), 7010–7023 (1996)

    Article  ADS  Google Scholar 

  7. Connes, A.: An analogue of the Thom isomorphism for crossed products of a \(C^*\)-algebra by an action of \({{\mathbb{R}}}\). Adv. Math. 39, 31–55 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  8. Cuntz, J.: \(K\)-theory and \(C^*\)-algebras. In: Algebraic \(K\)-Theory, Number Theory, Geometry and Analysis, Lecture Notes in Mathematics, vol. 1046, pp. 55–79, Springer, Berlin (1984)

  9. Cuntz, J., Meyer, R., Rosenberg, J.: Topological and bivariant \(K\)-theory. Birkhäuser, Basel (2007)

    MATH  Google Scholar 

  10. Echterhoff, S.: A categorical approach to imprimitivity theorems for \(C^*\)-dynamical systems. Mem. Am. Math. Soc. 805 (2006). arXiv:math/0205322

  11. Echterhoff, S., Nest, R., Oyono-Oyono, H.: Principal non-commutative torus bundles. Proc. Lond. Math. Soc. 99(3), 1–31 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  12. Echterhoff, S., Williams, D.P.: Locally inner actions on \(C_0(X)\)-algebras. J. Oper. Theory 45, 131–160 (2001)

    MathSciNet  MATH  Google Scholar 

  13. Fack, T., Skandalis, G.: Connes’ analogue of the Thom isomorphism for the Kasparov groups. Invent. Math. 64(1), 7–14 (1981)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Freed, D.S., Moore, G.W.: Twisted equivariant matter. Ann. Henri Poincaré 14(8), 1927–2023 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Gawedzki, K.: Bundle gerbes for topological insulators. Banach Center Publications. arXiv:1512.01028 (in press)

  16. Green, P.: The local structure of twisted covariance algebras. Acta Math. 140, 191–250 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  17. Hannabuss, K.C.: Representations of nilpotent locally compact groups. J. Funct. Anal. 34, 146–165 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  18. Hannabuss, K.C., Mathai, V.: Noncommutative principal torus bundles via parametrised strict deformation quantization. AMS Proc. Symp. Pure Math. 81, 133–148 (2010). arXiv:0911.1886

    Article  MathSciNet  MATH  Google Scholar 

  19. Hannabuss, K.C., Mathai, V.: Parametrised strict deformation quantization of \(C^*\)-bundles and Hilbert \(C^*\)-modules. J. Aust. Math. Soc. 90(1), 25–38 (2011). arXiv:1007.4696

    Article  MathSciNet  MATH  Google Scholar 

  20. Hannabuss, K.C., Mathai, V., Thiang, G.C.: T-duality simplifies bulk-boundary correspondence: the parametrised case. Adv. Theor. Math. Phys. 20(5), 1193–1226 (2016). arXiv:1510.04785

    Article  MathSciNet  MATH  Google Scholar 

  21. Kane, C.L., Mele, E.J.: Quantum spin Hall effect in graphene. Phys. Rev. Lett. 95(22), 226801 (2005)

    Article  ADS  Google Scholar 

  22. Kane, C.L., Mele, E.J.: \({\mathbb{Z}}_2\) topological order and the quantum spin Hall effect. Phys. Rev. Lett. 95(14), 146802 (2005)

    Article  ADS  Google Scholar 

  23. Kellendonk, J., Richter, T., Schulz-Baldes, H.: Edge current channels and Chern numbers in the integer quantum Hall effect. Rev. Math. Phys. 14(1), 87–119 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  24. Kitaev, A.: Periodic table for topological insulators and superconductors. AIP Conf. Proc. 1134(1), 22–30 (2009)

    Article  ADS  MATH  Google Scholar 

  25. Kleinert, H.: Gauge Fields in Condensed Matter, vol. 2. World Scientific, Singapore (1989)

    Book  MATH  Google Scholar 

  26. Kleppner, A.: Multipliers on abelian groups. Math. Ann. 158, 11–34 (1965)

    Article  MathSciNet  MATH  Google Scholar 

  27. Lawson, H., Michelsohn, M.-L.: Spin Geometry, Princeton Mathematical Series, vol. 38. Princeton University Press, Princeton (1989)

    MATH  Google Scholar 

  28. Lee, S.T., Packer, J.: Twisted group algebras for two-step nilpotent and generalized discrete Heisenberg groups. J. Oper. Theory 33, 91–124 (1995)

    MathSciNet  MATH  Google Scholar 

  29. Marcolli, M.: Solvmanifolds and noncommutative tori with real multiplication. Commun. Number Theory Phys. 2(2), 421–476 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  30. Mathai, V., Rosenberg, J.: T-duality for torus bundles via noncommutative topology. Commun. Math. Phys. 253, 705–721 (2005). arXiv:hep-th/0401168

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Mathai, V., Rosenberg, J.: T-duality for torus bundles with H-fluxes via noncommutative topology, II. The high-dimensional case and the T-duality group. Adv. Theor. Math. Phys. 10, 123–158 (2006). arXiv:hep-th/0508084

    Article  MathSciNet  MATH  Google Scholar 

  32. Mathai, V., Thiang, G.C.: T-duality of topological insulators. J. Phys. A Math. Theor. (Fast Track Commun.) 48(42), 42FT02 (2015). arXiv:1503.01206

    Article  MathSciNet  MATH  Google Scholar 

  33. Mathai, V., Thiang, G.C.: T-duality simplifies bulk-boundary correspondence. Commun. Math. Phys. 345(2), 675–701 (2016). arXiv:1505.05250

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Mathai, V., Thiang, G.C.: T-duality simplifies bulk-boundary correspondence: some higher dimensional cases. Ann. Henri Poincaré 17(12), 3399–3424 (2016). arXiv:1506.04492

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Mathai, V., Thiang, G.C.: Differential topology of semimetals. Commun. Math. Phys. 355(2), 561–602 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Packer, J., Raeburn, I.: Twisted crossed products of \(C^*\)-algebras. Math. Proc. Cambridge Philos. Soc. 106, 293–311 (1989)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Paschke, W.: On the mapping torus of an automorphism. Proc. Am. Math. Soc. 88, 481–485 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  38. Pimsner, M., Voiculescu, D.: Exact sequences for \(K\)-groups and \(EXT\)-groups of certain cross-product \(C^*\)-algebras. J. Oper. Theory 4, 93–118 (1980)

    MathSciNet  MATH  Google Scholar 

  39. Prodan, E., Schulz-Baldes, H.: Bulk and Boundary Invariants for Complex Topological Insulators: From \(K\)-Theory to Physics. Mathematical Physics Studies. Springer, Cham (2016)

    Book  MATH  Google Scholar 

  40. Raeburn, I., Rosenberg, J.: Crossed products of continuous-trace \(C^*\) algebras by smooth actions. Trans. Am. Math. Soc. 305, 1–45 (1988)

    MathSciNet  MATH  Google Scholar 

  41. Raeburn, I., Williams, D.: Morita Equivalence and Continuous-Trace \(C^*\)-Algebras. Mathematical Surveys and Monographs. American Mathematical Society, Providence (1998)

    MATH  Google Scholar 

  42. Ran, Y., Zhang, Y., Vishwanath, A.: One-dimensional topologically protected modes in topological insulators with lattice dislocations. Nat. Phys. 5, 298–303 (2009)

    Article  Google Scholar 

  43. Ray, M.W., Ruokokoski, E., Kandel, S., Möttönen, M., Hall, D.S.: Observation of Dirac monopoles in a synthetic magnetic field. Nature 505, 657–660 (2014)

    Article  ADS  Google Scholar 

  44. Rieffel, M.A.: Connes’ analogue for crossed products of the Thom isomorphism. Contemp. Math. 10, 143–154 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  45. Rieffel, M.A.: Strong Morita equivalence of certain transformation group \(C^*\)-algebras. Math. Ann. 222(1), 7–22 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  46. Rørdam, M., Larsen, M., Laustsen, M.: An Introduction to \(K\)-theory for \(C^*\)-algebras. London. Math. Soc. Student Texts 19. Cambridge Univ. Press, Cambridge (2000)

  47. Rosenberg, J.: Some results on cohomology with Borel cochains, with applications to group actions on operator algebras. Oper. Theory Adv. Appl. 17, 301–330 (1986)

    MathSciNet  MATH  Google Scholar 

  48. Rosenberg, J.: \(C^*\)-algebras, positive scalar curvature, and the Novikov conjecture–III. Topology 25(3), 319–336 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  49. Schröder, H.: \(K\)-Theory for Real \(C^*\)-Algebra and Applications. Pitman Research Notes in Mathemathical Series. Longman, Harlow (1993)

    Google Scholar 

  50. Scott, P.: The geometries of 3-manifolds. Bull. Lond. Math. Soc. 15(5), 401–487 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  51. Takai, H.: On a duality for crossed product algebras. J. Funct. Anal. 19, 25–39 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  52. Thiang, G.C.: On the \(K\)-theoretic classification of topological phases of matter. Ann. Henri Poincaré 17(4), 757–794 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Wegge-Olsen, N.E.: \(K\)-Theory and \(C^*\)-Algebras. Oxford University Press, Oxford (1993)

    MATH  Google Scholar 

  54. Wu, Y.-S., Zee, A.: Cocycles and magnetic monopoles. Phys. Lett. B 152, 98–102 (1985)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

Varghese Mathai and Guo Chuan Thiang were supported by the Australian Research Council via ARC Discovery Project Grants DP150100008, FL170100020, and DE170100149, respectively. The authors thank the Erwin Schrödinger Institute (ESI), Vienna, for its hospitality during the ESI Program on Higher Structures in String Theory and Quantum Field Theory, when part of this research was completed.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Varghese Mathai.

Appendices

Appendix A: Paschke’s map and the Connes–Thom isomorphism

This appendix explains why Paschke’s map [37] is, up to a sign convention, the Connes–Thom isomorphism [7] composed with an isomorphism from Green’s imprimitivity theorem. This result belongs to a family of ideas which can be found in [7,8,9, 37] and may be known to experts, but we provide a detailed argument for the reader’s reference. The Paschke map is explicitly defined for each \({\mathbb {Z}}\)-algebra \(({\mathcal {A}},\alpha )\), whereas the Connes–Thom map is more abstractly defined. As we will see, there is an analogous abstract characterization of the Paschke map, which determines its formula uniquely.

1.1 Generalities on Connes–Thom isomorphism

Recall from [7] that the Connes–Thom isomorphism is a natural transformation between the functors \(K_\bullet (\cdot )\) and \(K_{\bullet +1}((\cdot )\rtimes {\mathbb {R}})\), \(\bullet \in {\mathbb {Z}}_2\), i.e. it assigns to every \({\mathbb {R}}\)-\(C^*\)-algebra \(({\mathscr {A}},\alpha )\) an isomorphism \(\phi ^\bullet _\alpha : K_\bullet ({\mathscr {A}})\rightarrow K_{\bullet +1}({\mathscr {A}}\rtimes _\alpha {\mathbb {R}})\), such that for any morphism \(\rho :({\mathscr {A}},\alpha )\rightarrow (\mathscr {B},\beta )\), the following diagram commutes:

(28)

Here, \(\rho \rtimes {\mathbb {R}}\) is the morphism of crossed products induced by the equivariant map \(\rho \). Furthermore, Connes shows that \(\phi _\alpha ^\bullet \) is an isomorphism, which is determined uniquely by this naturality property together with a normalization condition (for \(({\mathcal {A}},\alpha )=({{\mathbb {C}}},{\mathrm {id}})\)) and compatibility with suspensions.

1.2 The analogous axioms for Paschke’s map

If \(\rho :({\mathcal {A}},\alpha )\rightarrow ({\mathcal {B}},\beta )\) is a morphism of \({\mathbb {Z}}\)-\(C^*\)-algebras (not necessarily unital), the induction functor \(\mathrm {Ind}:=\mathrm {Ind}_{\mathbb {Z}}^{\mathbb {R}}(\cdot )\) gives a morphism of induced algebras \(\tilde{\rho }:\mathrm {Ind}({\mathcal {A}},\alpha )\rightarrow \mathrm {Ind}({\mathcal {B}},\beta )\) which is equivariant for the respective translation actions \(\tau ^\alpha , \tau ^\beta \) of \({\mathbb {R}}\). Elements of \(\mathrm {Ind}({\mathcal {A}},\alpha )\) are viewed either as bounded continuous functions \(f:{\mathbb {R}}\rightarrow {\mathcal {A}}\) satisfying an equivariance conditionFootnote 6, \(f(x+1)=\alpha (f(x)), x\in {\mathbb {R}}\), or alternatively, continuous \(f:[0,1]\rightarrow {\mathcal {A}}\) such that \(f(1)=\alpha (f(0))\) (the mapping torus). The translation action \(\tau ^\alpha \) of \({\mathbb {R}}\) on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) is \((\tau ^\alpha _t f)(s)=f(s+t), s,t\in {\mathbb {R}}\).

Let \(S\alpha \) be the suspended \({\mathbb {Z}}\)-action on \(S{\mathcal {A}}\) (acting by \(\alpha \) on \({\mathcal {A}}\) and trivially on the suspension variable). Note that \(S(\mathrm {Ind}({\mathcal {A}},\alpha ))\cong \mathrm {Ind}(S{\mathcal {A}},S\alpha )\) and \(S({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\cong S{\mathcal {A}}\rtimes _{S\alpha }{\mathbb {Z}}\), and write \(S^\bullet _{(\cdot )}\) for the natural suspension isomorphisms \(K_\bullet (\cdot )\rightarrow K_{\bullet -1}(S(\cdot ))\).

For \(\bullet \in {\mathbb {Z}}_2\), let \(\gamma ^\bullet \) be a natural transformation of the functors \(({\mathcal {A}},\alpha )\mapsto K_\bullet (\mathrm {Ind}({\mathcal {A}},\alpha ))\) and \(({\mathcal {A}},\alpha )\mapsto K_{\bullet +1}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\) from \({\mathbb {Z}}\)-algebras to abelian groups, satisfying the following three axioms:

Axiom A.1

(Normalization) If \(({\mathcal {A}},\alpha )=({{\mathbb {C}}},{\mathrm {id}})\), then \(\gamma ^0_{{\mathrm {id}}}:K_0(C({\mathbb {R}}/{\mathbb {Z}}))=K_0(C({\mathbb {T}}))\rightarrow K_1(C^*({\mathbb {Z}}))\) takes \([1_{C({\mathbb {T}})}]\) to the (Bott) generator [b] of \(K_1(C^*({\mathbb {Z}}))\) corresponding to \(1\in {\mathbb {Z}}\) regarded as an element of \(C^*({\mathbb {Z}})\).

Axiom A.2

(Naturality) If \(\rho :({\mathcal {A}},\alpha )\rightarrow ({\mathcal {B}},\beta )\) is a morphism of \({\mathbb {Z}}\)-algebras, then the following diagram commutes:

(29)

Axiom A.3

(Suspension) \(\gamma \) is compatible with S in the sense that

$$\begin{aligned} S^{(\bullet +1)}_{\mathcal {A\rtimes _\alpha {\mathbb {Z}}}}\circ \gamma ^\bullet _\alpha =\gamma ^{(\bullet -1)}_{S\alpha }\circ S^\bullet _{\mathrm {Ind}({\mathcal {A}},\alpha )}. \end{aligned}$$
(30)

Proposition A.4

There is a unique \(\gamma \) satisfying Axioms A.1A.3.

Proof

Outline: We use a modification of Connes’ argument for the uniqueness of \(\phi \). Axiom A.3 and Bott periodicity ensure that we only need to look at \(\gamma ^0\) (which we will simply denote by \(\gamma \) subsequently). We need to show that \(\gamma _\alpha [p]\in K_1({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\) is uniquely determined for any projection \(p\equiv \{p_\theta \}_{\theta \in [0,1]}\) in \(\mathrm {Ind}({\mathcal {A}},\alpha )\). The basic idea is to construct a \({\mathbb {Z}}\)-action \(\alpha '\), exterior equivalent to \(\alpha \), which fixes \(p_0\). Furthermore, the new induced algebra \(\mathrm {Ind}({\mathcal {A}},\alpha ')\), with its translation action \(\tau ^{\alpha '}\), is \({\mathbb {R}}\)-equivariantly isomorphic to the original \(\mathrm {Ind}_{\alpha }{\mathcal {A}}\) with a modified action \(\tau '\), where \(\tau '\) is exterior equivalent to the original \(\tau ^\alpha \) and fixes p. Thus, the projection \(p'\in \mathrm {Ind}({\mathcal {A}},\alpha ')\) corresponding to \(p\in \mathrm {Ind}({\mathcal {A}},\alpha )\) under this isomorphism is itself translation invariant under \(\tau ^{\alpha '}\). Then, the Axioms determine what \(\gamma _{\alpha '}[p']\) has to be. Finally, we need to argue that the modification \(\alpha \mapsto \alpha '\) can be assumed because the above exterior equivalences determine unique maps, consistent with the Axioms, which yield the desired \(\gamma _\alpha [p]\) from \(\gamma _{\alpha '}[p']\).

Details: As in [7], we may assume without loss that \({\mathcal {A}}\) is unital and \(p\in \mathrm {Ind}({\mathcal {A}},\alpha )\). By [7] Proposition 4 (c.f. [9] Lemma 10.16, [44]), we may also assume that on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) there is an action \({\mathbb {R}}\)-action \(\tau '\), exterior equivalent to \(\tau ^\alpha \), which fixes p. Let \(U\equiv \{U_t\}_{t\in {\mathbb {R}}}\) be the 1-cocycle on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) which relates \(\tau ^\alpha \) and \(\tau '\) via \(\tau '_t=\mathrm {Ad}(U_t)\circ \tau ^\alpha _t\). Thus, \(\{U_t\}_{t\in {\mathbb {R}}}\) satisfies the cocycle condition and equivariance condition,

$$\begin{aligned} U_{t_1+t_2}=U_{t_1}\tau ^\alpha _{t_2}(U_{t_2}),\quad U_t(1)=\alpha (U_t(0)), \quad t,t_1,t_2\in {\mathbb {R}}. \end{aligned}$$
(31)

Define the modified \({\mathbb {Z}}\)-action \(\alpha '\) on \({\mathcal {A}}\) via

$$\begin{aligned} \alpha '_n=\mathrm {Ad}(U_n(0))\circ \alpha _n, \quad n\in {\mathbb {Z}}. \end{aligned}$$
(32)

We can verify, from (31), that the assignment \(u:n\mapsto U_n(0)=:u_n,\, n\in {\mathbb {Z}}\) defines a unitary 1-cocycle \(\{u_n\}_{n\in {\mathbb {Z}}}\) on \(({\mathcal {A}},\alpha )\), i.e. \(u_{m+n}=u_m\alpha ^m(u_n)\). Thus, \({\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}\cong {\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}\) (a canonical isomorphism \(\varphi _u\) is given in Lemma A.5 later).

Since \(\tau '_1=\mathrm {Ad}(U_1)\circ \tau ^\alpha _1\) fixes p, it follows that \(p_0:=p(0)\in {\mathcal {A}}\) fixed by \(\alpha '\). Define the isomorphism \(\Psi _U:\mathrm {Ind}({\mathcal {A}},\alpha )\rightarrow \mathrm {Ind}({\mathcal {A}},\alpha ')\) by

$$\begin{aligned} \Psi _Uf(s)=\mathrm {Ad}(U_s(0))[f(s)],\quad s\in {\mathbb {R}}, f\in \mathrm {Ind}({\mathcal {A}},\alpha ). \end{aligned}$$
(33)

We can verify that \(\Psi _Uf\) does satisfy \(\alpha '\)-equivariance and intertwines \(\tau '\) on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) with the new translation action \(\tau ^{\alpha '}\) on \(\mathrm {Ind}({\mathcal {A}},\alpha ')\), i.e. \((\tau ^{\alpha '}_t(\Psi _U f))= (\Psi _U(\tau '_t f))\).

Observe that \(p'(\theta )=p'_0=p_0\) since \(p'=\Psi _U(p)\) is fixed by \(\tau ^{\alpha '}\). The homomorphism

$$\begin{aligned} \omega :({{\mathbb {C}}},{\mathrm {id}})\rightarrow ({\mathcal {A}},\alpha '),\;\;\;\;\;\omega (\lambda )=\lambda p'_0 \end{aligned}$$

is \({\mathbb {Z}}\)-equivariant, and we write \(\tilde{\omega }\) for the corresponding inflated map \(\tilde{\omega }:C({\mathbb {T}})\cong \mathrm {Ind}({{\mathbb {C}}},{\mathrm {id}})\rightarrow \mathrm {Ind}({\mathcal {A}},\alpha ')\). Note that

$$\begin{aligned} (\tilde{\omega } (1_{C({\mathbb {T}})}))(\theta )\equiv \omega (1_{C({\mathbb {T}})}(\theta ))=p'_0=p'(\theta ),\quad \theta \in [0,1], \end{aligned}$$

so \(p'=\tilde{\omega }(1_{C({\mathbb {T}})})\). Then naturality and normalization mean that \(\gamma _{\alpha '}[p']\) must be determined by the equation

$$\begin{aligned} \gamma _{\alpha '}[p']=\gamma _{\alpha '}\circ \tilde{\omega }_*[1_{C({\mathbb {T}})}]=(\omega \rtimes {\mathbb {Z}})_*\circ \gamma _{{\mathrm {id}}}[1_{C({\mathbb {T}})}]=(\omega \rtimes {\mathbb {Z}})_*[b]. \end{aligned}$$
(34)

Corollary A.7 gives the details of how \(\gamma [p]\) is obtained from \(\gamma _{\alpha '}[p']\). \(\square \)

1.2.1 Connes’ \(2\times 2\) matrix trick

Lemma A.5

(c.f. [7] Lemma 2) Let \(\alpha , \alpha '\) be exterior equivalent \({\mathbb {Z}}\)-actions on \({\mathcal {A}}\) related by a unitary 1-cocycle \({\mathbb {Z}}\ni n \mapsto u_n \in U{\mathcal {M}}{\mathcal {A}}\). Then

$$\begin{aligned} \kappa \begin{pmatrix}a_{11} &{} a_{12} \\ a_{21} &{} a_{22} \end{pmatrix}= \begin{pmatrix} \alpha (x_{11}) &{}\quad \alpha (x_{12})u_1^* \\ u_1\alpha (x_{21}) &{}\quad \alpha '(x_{22}) \end{pmatrix} \end{aligned}$$
(35)

is a \({\mathbb {Z}}\)-action on \(M_2({\mathcal {A}})\) which restricts to \(\alpha \) in the top-left corner and to \(\alpha '\) in the bottom-right corner. Let \(\iota , \iota '\) be the equivariant inclusions of \({\mathcal {A}}\) into their respective corners in \(M_2({\mathcal {A}})\). There is a unique isomorphism \(\varphi _u:{\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}\rightarrow {\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}\) such that \((\iota '\rtimes {\mathbb {Z}})(\varphi _u(y))=\mathrm {Ad}(\begin{pmatrix} 0 &{}\quad 1 \\ 1 &{}\quad 0\end{pmatrix})[(\iota \rtimes {\mathbb {Z}})(y)]\) for all \(y\in {\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}\).

Similarly, there are inclusions \(\tilde{\iota }, \tilde{\iota '}\) of the induced algebras into the respective corners of \(\mathrm {Ind}_\kappa M_2({\mathcal {A}})\). Let \(\tau ^\alpha , \tau '\) be the exterior equivalent \({\mathbb {R}}\)-actions on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) as in the proof of Proposition A.4, related by the 1-cocycle \(\{U_t\}_{t\in {\mathbb {R}}}\). A straightforward computation shows:

Lemma A.6

Define \(X\in \mathrm {Ind}_{\kappa } M_2({\mathcal {A}})\) by

$$\begin{aligned} X(s)=\begin{pmatrix} 0 &{}\quad U_s^*(0)\\ U_s(0) &{}\quad 0 \end{pmatrix}. \end{aligned}$$

Then the isomorphism \(\Psi _U\) satisfies \(\tilde{\iota '}(\Psi _U(f))=\mathrm {Ad}(X)[\tilde{\iota }f], f\in \mathrm {Ind}({\mathcal {A}},\alpha )\).

Corollary A.7

Suppose \(\gamma \) satisfies Axioms A.1A.3, and \(\alpha , \alpha ', U\) are as above. Then \(\gamma ^\bullet _{\alpha '}=(\varphi _u)_*\circ \gamma ^\bullet _\alpha \circ (\Psi _U^{-1})_*\).

Proof

Recall that inner automorphisms induce the identity map in K-theory. Thus, Lemmas A.5 and A.6 show that \((\iota \rtimes {\mathbb {Z}})_*=(\iota '\rtimes {\mathbb {Z}})_*\circ (\varphi _u)_*\) and \(\tilde{\iota }_*=\tilde{\iota '}_*\circ (\Psi _U)_*\). Using the naturality axiom, and dropping the superscript \({}^\bullet \) for now,

$$\begin{aligned}&(\iota '\rtimes {\mathbb {Z}})_*\circ \gamma _{\alpha '}= \gamma _\kappa \circ \tilde{\iota '}_*= \gamma _\kappa \circ \tilde{\iota }_*\circ (\Psi _U^{-1})_*\\&\quad = (\iota \rtimes {\mathbb {Z}})_*\circ \gamma _\alpha \circ (\Psi _U^{-1})_*= (\iota '\rtimes {\mathbb {Z}})_*\circ (\varphi _u)_*\circ \gamma _\alpha \circ (\Psi _U^{-1})_*, \end{aligned}$$

so injectivity of \((\iota '\rtimes {\mathbb {Z}})_*\) (see [7] Proposition 2–3) gives the required result. \(\square \)

1.3 Paschke’s isomorphisms

Paschke’s isomorphisms [37]

$$\begin{aligned} \gamma ^{\bullet ,{{{\mathrm{Paschke}}}}}_\alpha : K_\bullet (\mathrm {Ind}({\mathcal {A}},\alpha ))\rightarrow K_{\bullet +1}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}) \end{aligned}$$

are first defined for unital \({\mathcal {A}}\) and \(\bullet =0\) on [p], by exhibiting the existence of a path of unitaries \(\theta \mapsto w_\theta \in {\mathcal {A}}\) such that

$$\begin{aligned} p(\theta )=\mathrm {Ad}(w_\theta )(p_0),\quad \theta \in [0,1]. \end{aligned}$$
(36)

In particular, \(p_1=\mathrm {Ad}(w_1)(p_0)\). Then \(\gamma ^{0,\mathrm {Paschke}}_\alpha [p]\) is defined to be

$$\begin{aligned} \gamma ^{0,{{{\mathrm{Paschke}}}}}_\alpha [p]=[L^*w_1p_0+1-p_0], \end{aligned}$$
(37)

where \(L\in {\mathcal {M}}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\) is the unitary implementing \(\alpha \) (i.e. \(\alpha (a)=LaL^*, a\in {\mathcal {A}}\)), with this map shown to be well-defined in K-theory.Footnote 7

The \(\bullet =1\) case is defined by compatibility with suspensions. Since \(\gamma ^{0,\mathrm {Paschke}}_\alpha \) may be expressed using representative projections and unitaries (rather than their K-theory classes), it is easy to see that the naturality Axiom A.2 is satisfied. It is also clear that normalization Axiom A.1 is satisfied up to a minus sign.

On the other hand, there is a commutative diagram due to the Connes–Thom natural isomorphisms,

(38)

Composing \(\phi _{\tau ^\alpha }^\bullet \) with the natural K-theory isomorphisms

$$\begin{aligned} M_\alpha ^{\bullet +1,{\mathrm{Green}}}:K_{\bullet +1}(\mathrm {Ind}({\mathcal {A}},\alpha )\rtimes _{\tau ^\alpha }{\mathbb {R}})\cong K_{\bullet +1}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}) \end{aligned}$$

given by the Morita equivalence \(\mathrm {Ind}({\mathcal {A}},\alpha )\rtimes _{\tau ^\alpha }{\mathbb {R}}\sim _{\mathrm{M.E.}}{\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}\) (this is a special case of Green’s imprimitivity theorem [16]), we obtain another \(\gamma \) satisfying Axioms A.1A.3: The normalization axiom follows from the fact that \(\phi _{\tau ^{{\mathrm{id}}}}\) is an isomorphism, so there is only one possibility (apart from a sign choice) for \(M_{\mathrm{id}}^{0,{\mathrm{Green}}}\circ \phi _{\tau ^{\mathrm{id}}}:{\mathbb {Z}}[1_{C({\mathbb {T}})}]\rightarrow {\mathbb {Z}}[b]\). Compatibility with suspensions is inherited from the Connes–Thom map, while naturality follows from naturality of implementing the Morita equivalence (see [10] Chapter 4). By Proposition A.4, we have

Theorem A.8

The map \(\gamma ^{\bullet ,\,{{{\mathrm{Paschke}}}}}_\alpha :K_\bullet (\mathrm {Ind}({\mathcal {A}},\alpha ))\rightarrow K_{\bullet +1}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\) is, up to a sign, the composition \(M_\alpha ^{\bullet +1,\,{\mathrm{Green}}}\circ \phi ^\bullet _{\tau ^\alpha }\).

1.3.1 Paschke’s unitaries and Connes’ cocycle

Given a \(p\in \mathrm {Ind}({\mathcal {A}},\alpha )\), we may start from Connes’ cocycle \(\{U_t\}_{t\in {\mathbb {R}}}\) in (31) above and obtain a path of unitaries \(w_\theta :=U^*_\theta (0), \theta \in [0,1]\) in \({\mathcal {A}}\). Then \(w_0=1\) and

$$\begin{aligned} p_\theta =p(\theta )\,{=}\,\tau ^\alpha _\theta (p(0))\,{=}\,(\mathrm {Ad}(U^*_\theta )\circ \tau '(p))(0)=\mathrm {Ad}(U^*_\theta (0))(p(0))=\mathrm {Ad}(w_\theta )(p_0), \end{aligned}$$

thus \(\{w_\theta \}_{\theta \in [0,1]}\) is precisely a Paschke path whose existence was proved in [37].

Conversely, given a Paschke path \(\{w_\theta \}_{\theta \in [0,1]}\) such that \(w_0=1\) and \(\mathrm {Ad}(w_\theta )(p_0)=p_\theta \), we can reconstruct Connes’ cocycle for the action \(\tau ^\alpha \) on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) as follows. Regard \(p\in \mathrm {Ind}({\mathcal {A}},\alpha )\) as a projection-valued function \({\mathbb {R}}\rightarrow {\mathcal {A}}\) through the equivariance condition \(p(s)\equiv p_s=\alpha (p_{s-1}), s\in {\mathbb {R}}\). Similarly, extend Paschke’s \(w:[1,0]\mapsto {\mathcal {A}}\) to a continuous unitary function \(w:{\mathbb {R}}\ni s\mapsto w_s\in {\mathcal {A}}\) by the recursion relation \(w_s=\alpha (w_{s-1})\cdot w_1\). It is easy to check that \(p_s=\mathrm {Ad}(w_s)(p_0)\) for all \(s\in {\mathbb {R}}\) and that \(n\mapsto w_n\) defines a 1-cocycle for \(({\mathcal {A}},\alpha )\). The projection \(p_0\) is then fixed by \(\alpha '_n=\mathrm {Ad}(w^*_n)\circ \alpha , n\in {\mathbb {Z}}\), since

$$\begin{aligned} \alpha '(p_0)=\mathrm {Ad}(w^*_1)(\alpha (p_0))=\mathrm {Ad}(w^*_1)\circ \mathrm {Ad}(w_1)(p_0)=p_0. \end{aligned}$$

The assignment \(t\mapsto U_t\in \mathrm {Ind}({\mathcal {A}},\alpha )\) defined by

$$\begin{aligned} U_t(s)=w_sw_{s+t}^*, \quad s,t\in {\mathbb {R}} \end{aligned}$$
(39)

is the desired Connes 1-cocycle for \((\mathrm {Ind}({\mathcal {A}},\alpha ),\tau ^\alpha )\); we can verify that each \(U_t\) is \(\alpha \)-equivariant and satisfies the cocycle condition. If we define the \({\mathbb {R}}\)-action \(\tau '\) on \(\mathrm {Ind}({\mathcal {A}},\alpha )\) by \(\tau '_t=\mathrm {Ad}(U_t)\circ (\tau ^\alpha _t)\), we find that the projection p is now fixed by \(\tau '\).

1.4 Paschke’s formula from axioms

Recall that \(\gamma ^0_\alpha [p]\) can be computed by passing to \(\alpha '\), using \(\gamma ^0_\alpha [p]=(\varphi _u^{-1})_*\circ \gamma ^0_{\alpha '}\circ (\Psi _U)_*[p]\). In (34), we saw that \(\gamma ^0_{\alpha '}\circ (\Psi _U)_*[p]=\gamma ^0_{\alpha '}[p']=(\omega \rtimes {\mathbb {Z}})_*[b].\) Note that even if \({\mathcal {A}}\) is unital, the homomorphism \(\omega \) (and thus \(\omega \rtimes {\mathbb {Z}}\)) is nonunital in general, so the induced map \((\omega \rtimes {\mathbb {Z}})_*\) in \(K_1\) is slightly tricky to compute (c.f. Proposition 8.1.6 of [46]). Namely, we have to append a formal unit \(\mathbf{1}\) to \(C({\mathbb {T}})\) and \({\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}\), replace [b] by \([\mathbf{1}-1_{C({\mathbb {T}})}+b]\), then compute \((\omega \rtimes {\mathbb {Z}})_*\) using the unital extension \((\omega \rtimes {\mathbb {Z}})^+:C({\mathbb {T}})^+\rightarrow ({\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}})^+\). This gives

$$\begin{aligned} ((\omega \rtimes {\mathbb {Z}})^+)_*[\mathbf{1}-1_{C({\mathbb {T}})}+b]&=[\mathbf{1}+(\omega \rtimes {\mathbb {Z}})(-1_{C({\mathbb {T}})}+b)]\\&=[\mathbf{1}-L'p'_0+p'_0]\\&=[\mathbf{1}-1_{{\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}}+1_{{\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}}+L'p'_0-p'_0], \end{aligned}$$

and mapping back to unitaries in \(\mathrm {Ind}({\mathcal {A}},\alpha ')\), we obtain

$$\begin{aligned} (\omega \rtimes {\mathbb {Z}})_*[b]=[1+L'p'_0-p'_0]=[L'p_0+(1-p_0)]. \end{aligned}$$

Finally, recall that the isomorphism \(\varphi _u\) implements the isomorphism \({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}}\cong {\mathcal {A}}\rtimes _{\alpha '}{\mathbb {Z}}\) induced by the exterior equivalence \(\alpha '_n=\mathrm {Ad}(w^*_n)\circ \alpha _n\), so it effects \(w_1^*L\mapsto L'\). Thus,

$$\begin{aligned} \gamma ^0_\alpha [p]=(\varphi _u^{-1})_*\circ \gamma ^0_{\alpha '}\circ (\Psi _U)_*[p]=(\varphi _u^{-1})_*[L'p_0+(1-p_0)]=[w_1^*Lp_0+(1-p_0)], \end{aligned}$$

which, up to a sign, concides with \(\gamma ^{0,{{{\mathrm{Paschke}}}}}_\alpha [p]\).

1.5 The real \(C^*\)-algebra case

A general reference for the K-theory of real \(C^*\)-algebras is [49]. The Connes–Thom isomorphisms and Pimsner–Voiculescu exact sequence also hold for real \(C^*\)-algebras \({\mathcal {A}}\) and real crossed products, see [48, 49]. Similarly, Paschke’s map \(\gamma ^{0,{{{{{\mathrm{Paschke}}}}}}}_\alpha \) still makes sense on real \({\mathbb {Z}}\)-algebras \(({\mathcal {A}},\alpha )\). We can then repeat the arguments in this appendix to obtain the real version of Theorem A.8. The only significant change is in the normalization axiom, and in taking \(\bullet \in {\mathbb {Z}}_8\) rather than in \({\mathbb {Z}}_2\).

Recall that \({\mathbb {C}}\rtimes _{\mathrm{id}}{\mathbb {Z}}=C_{\mathbb {R}}^*({\mathbb {Z}})\cong C({\mathbb {T}};\varsigma )\) where \(\varsigma \) is the involution \(\theta \mapsto -k\theta \) on \({\mathbb {T}}=\widehat{{\mathbb {Z}}}\) inherited from complex conjugation of characters (parametrized by \(\theta \in [0,1]\)), and \(C({\mathbb {T}};\varsigma )\) is the real \(C^*\)-algebra of continuous functions \(f\in {\mathbb {T}}\rightarrow {{\mathbb {C}}}\) such that \(\overline{f(\theta )}=f(-\theta )\). These functions are precisely the ones with real Fourier coefficients. It is known (pp. 40 of [49]) that \(K_1(C({\mathbb {T}};\varsigma ))\cong {\mathbb {Z}}[z]\oplus {\mathbb {Z}}_2[-1_{C({\mathbb {T}},\varsigma )}]\), where z denotes the unitary function \(\theta \mapsto e^{2\pi i\theta }\) with winding number 1. Furthermore, [z] is the Bott element implementing (1,1) periodicity in KR-theory (c.f. Theorem 1.5.4 of [49], Theorem 10.3 of [27]), so we also write it as \([b_r]\). We can also identify it with the unitary L implementing the (trivial) automorphism in \({\mathbb {R}}\rtimes _{\mathrm{id}}{\mathbb {Z}}\).

On the other hand, the induced algebra for \(({\mathbb {R}},{\mathrm{id}})\) is just the continuous real-valued functions on the circle \(C({\mathbb {T}};{\mathbb {R}})\). It is also known that

$$\begin{aligned} KO_0(C({\mathbb {T}};{\mathbb {R}}))={\mathbb {Z}}[1_{C({\mathbb {T}};{\mathbb {R}})}]\oplus {\mathbb {Z}}_2\left[ [P_{{\mathrm{Mob}}}]-[1_{C({\mathbb {T}};{\mathbb {R}})}]\right] , \end{aligned}$$

where \(P_{\mathrm{Mob}}\) is the Möbius projection in \(M_2(C({\mathbb {T}};{\mathbb {R}}))\) given by

$$\begin{aligned} P_{\mathrm{Mob}}(\theta )=\mathrm {Ad}(R(\theta )) \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 0 \end{pmatrix}=\mathrm {Ad}\begin{pmatrix}\cos \,\pi \theta &{} -\sin \,\pi \theta \\ \sin \,\pi \theta &{} \cos \,\pi \theta \end{pmatrix}\begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 0 \end{pmatrix}. \end{aligned}$$

It is straightforward to compute the real Paschke map on \(KO_0(C({\mathbb {T}};{\mathbb {R}}))\). For the constant projection 1, the Paschke path of unitaries is trivial, so we have \(\gamma ^{0,{{{\mathrm{Paschke}}}}}_{\mathrm{id}}[1]=[L+1-1]=[L]=[b_r]\). For the Möbius projection, the Paschke path of unitaries is \(w_\theta =R(\theta )\), so \(w_1\) is minus the identity. Then

$$\begin{aligned}&\gamma ^{0,{\mathrm{Paschke}}}_{\mathrm{id}} [P_{\mathrm{Mob}}]=[-LP_{\mathrm{Mob}}(0)+1_2-P_{\mathrm{Mob}}(0)]\\&\quad =\left[ \begin{pmatrix} -z &{}\quad 0 \\ 0 &{}\quad 1_{C({\mathbb {T}},\varsigma )}\\ \end{pmatrix}\right] =[-z]=[-1_{C({\mathbb {T}},\varsigma )}]+[z], \end{aligned}$$

so \(\gamma ^{0,{{{\mathrm{Paschke}}}}}_{\mathrm{id}}([P_{\mathrm{Mob}}]-[1_{C({\mathbb {T}};{\mathbb {R}})}])=[-1_{C({\mathbb {T}},\varsigma )}]\), i.e. the torsion generators are correctly mapped to each other.

The normalization axiom in the real \(C^*\)-algebra case needs to be replaced by

Axiom A.9

(Normalization) If \(({\mathcal {A}},\alpha )=({\mathbb {R}},{\mathrm {id}})\), then \(\gamma ^0_{{\mathrm {id}}}:KO_0(C({\mathbb {T}};{\mathbb {R}}))\cong KO^0({\mathbb {T}})\rightarrow KO_1(C_{\mathbb {R}}^*({\mathbb {Z}}))\cong KO_1(C({\mathbb {T}};\varsigma ))\) takes \([1_{C({\mathbb {T}};{\mathbb {R}})}]\) to the (Bott) generator \([b_r]\) and \(([P_{\mathrm{Mob}}]-[1_{C({\mathbb {T}};{\mathbb {R}})}])\) to \([-1_{C({\mathbb {T}},\varsigma )}]\).

Theorem A.10

There is a unique natural transformation \(\gamma \) of the functors \(({\mathcal {A}},\alpha )\mapsto KO_\bullet (\mathrm {Ind}({\mathcal {A}},\alpha ))\) and \(({\mathcal {A}},\alpha )\mapsto KO_{\bullet +1}({\mathcal {A}}\rtimes _\alpha {\mathbb {Z}})\), \(\bullet \in {\mathbb {Z}}_8\) , from the category of real \({\mathbb {Z}}\)-algebras to abelian groups, which satisfy Axiom A.9 and Axioms A.2A.3 with K replaced by KO. Thus, the real version of Corollary A.8 holds: \(\gamma ^{\bullet ,\,{{{{{\mathrm{Paschke}}}}}}}_\alpha =M_\alpha ^{\bullet +1,\,{\mathrm{Green}}}\circ \phi ^\bullet _{\tau ^\alpha }\) up to a sign convention.

Appendix B: Multipliers for lattice groups

Let \(N={\mathbb {Z}}^d\) and for a multiplier \(\sigma \), let \(\widetilde{\sigma }(n,n^\prime ) = \sigma (n,n^\prime )/ \sigma (n^\prime ,n)\) be the antisymmetric bicharacter labelling its class in \(H^2(N,{\mathbb {T}})\). We reconstruct a canonical representative multiplier in this class starting from \(\widetilde{\sigma }\). Choose generators \(\{e_j\}_{j\in J}\), where J is a totally ordered set, and, expanding \(n = \sum _j n_je_j \in N\), and \(n^\prime \) similarly, we define the multiplier

$$\begin{aligned} \sigma _J(n,n^\prime ) = \prod _{j<k}\widetilde{\sigma }(e_j,e_k)^{n_jn_k^\prime }, \end{aligned}$$

which is also a bicharacter (but not antisymmetric). Then we see that

$$\begin{aligned} \widetilde{\sigma }_J(n,n^\prime )= & {} \frac{\sigma _J(n,n^\prime )}{\sigma _J(n^\prime ,n)}\\= & {} \prod _{j<k}\widetilde{\sigma }(e_j,e_k)^{n_jn_k^\prime }\prod _{k<j}\widetilde{\sigma }(e_k,e_j)^{-n_jn_k^\prime }\\= & {} \prod _{j\ne k}\widetilde{\sigma }(e_j,e_k)^{n_jn_k^\prime }\\= & {} \widetilde{\sigma }(n,n^\prime ). \end{aligned}$$

Thus, we see that \(\widetilde{\sigma }_J = \widetilde{\sigma }\), and consequently, \(\sigma _J\) is cohomologous to \(\sigma \), and so also to any \(\sigma _{J^\prime }\) for any other totally ordered set \(J^\prime \). We note that when \(N = {\mathbb Z}^{2}\) and \(J = \{1,2\}\) indexes the usual generators with the natural ordering, the antisymmetric bicharacter \(\widetilde{\sigma }(n,n^\prime ) = \exp (i(n_1n_2^\prime - n_2n_1^\prime ))\) produces \(\sigma _J(n,n^\prime ) = \exp (i(n_1n_2^\prime ))\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hannabuss, K.C., Mathai, V. & Thiang, G.C. T-duality simplifies bulk–boundary correspondence: the noncommutative case. Lett Math Phys 108, 1163–1201 (2018). https://doi.org/10.1007/s11005-017-1028-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-017-1028-x

Keywords

Mathematics Subject Classification

Navigation