Skip to main content
Log in

The non-abelian self-dual string

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

Abstract

We argue that the relevant higher gauge group for the non-abelian generalization of the self-dual string equation is the string 2-group. We then derive the corresponding equations of motion and discuss their properties. The underlying geometric picture is a string structure, i.e., a categorified principal bundle with connection whose structure 2-group is the string 2-group. We readily write down the explicit elementary solution to our equations, which is the categorified analogue of the ’t Hooft–Polyakov monopole. Our solution passes all the relevant consistency checks; in particular, it is globally defined on \(\mathbb {R}^4\) and approaches the abelian self-dual string of charge one at infinity. We note that our equations also arise as the BPS equations in a recently proposed six-dimensional superconformal field theory and we show that with our choice of higher gauge structure, the action of this theory can be reduced to four-dimensional supersymmetric Yang–Mills theory.

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. cf. Eq. (3.7).

  2. Considering (1, 0)-theories is indeed very natural: first, (2, 0)-theories are special cases of (1, 0)-theories. Second, recall that general M2-brane models were expected to posses \(\mathcal {N}=8\) supersymmetry in three dimensions. The ABJM model, however, only features \(\mathcal {N}=6\) supersymmetry, and full supersymmetry is restored by non-local “monopole operators.” The six-dimensional analogue would be a (1, 0)-theory whose supersymmetry is enhanced to that of the (2, 0)-theory by non-local “self-dual string operators.”

  3. A closely related model was derived in [25].

  4. Inversely since \(\mathbb {R}^3\backslash \{x\}\) is homotopy equivalent to \(S^2\), any principal bundle on \(\mathbb {R}^3\backslash \{x\}\) originates, up to isomorphism, from a pullback along the embedding \(S^2{\hookrightarrow }\mathbb {R}^3\backslash \{x\}\). This extends in an obvious way to \(\mathbb {R}^{n+1}\backslash \{x\}\) and \(S^n\) as well as to higher principal bundles.

  5. A slight generalization of a Lie groupoid: A 2-space is a category whose objects and morphisms form smooth manifolds and whose source, target, identity and composition maps are smooth. A simple notion of a map between 2-spaces is given by a functor consisting of smooth maps.

  6. A cover of \({\textsf {Spin}}(3)\) that extends to a simplicial cover of the nerve of \(\textsf {B}{\textsf {Spin}}(3)=({\textsf {Spin}}(3)\rightrightarrows *)\).

  7. This also informs the name skeletal, as an \(L_\infty \)-algebra with trivial \(\mu _1\) corresponds to a category where source and target maps agree, cf. [39].

  8. See “Appendix D” for definitions and notation.

  9. We use both terms interchangeably, cf. “Appendix B”.

  10. Unfortunately, one always has to employ slightly awkward grading conventions when working simultaneously with \(L_\infty \)-algebras, differential forms and NQ-manifolds.

  11. Interchanging two odd elements requires inserting a minus sign.

  12. See again “Appendix D” for definitions and notation.

  13. This condition is trivially satisfied in the skeletal case.

  14. As remarked earlier, one can also write down a gauge-invariant self-duality equation in six dimensions.

  15. Our construction works for arbitrary metric Lie algebra \(\mathfrak {g}\).

  16. Note that \({\textsf {Mfd}}^\infty \) does not contain all pullbacks which leads to some technicalities which we can ignore.

  17. To be precise, one would immediately specialize from 2-spaces to Lie groupoids and regard these as differentiable stacks or objects in a 2-category with the much more general bibundles as morphisms, see e.g., the discussion in [40]. This point can safely be ignored.

References

  1. Baez, J.C.: Higher Yang–Mills theory. arXiv:hep-th/0206130

  2. Baez, J.C., Schreiber, U.: Higher gauge theory: 2-connections on 2-bundles. arXiv:hep-th/0412325

  3. Baez, J.C., Schreiber, U.: Higher gauge theory. Contemp. Math. 431, 7 (2007). arXiv:math.DG/0511710

    MathSciNet  MATH  Google Scholar 

  4. Sati, H., Schreiber, U., Stasheff, J.: \(L_\infty \)-algebra connections and applications to String- and Chern-Simons \(n\)-transport. In: Fauser, B., Tolksdorf, J., Zeidler, E. (eds.) Quantum Field Theory, p. 303. Birkhäuser (2009). arXiv:0801.3480 [math.DG]

  5. Baez, J.C., Huerta, J.: An invitation to higher gauge theory. Gen. Relativ. Gravit. 43, 2335 (2011). arXiv:1003.4485 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  6. Saemann, C.: Lectures on higher structures in M-theory. In: Maeda, Y., Moriyoshi, H., Kotani, M., Watamura, S. (eds.) Noncommutative Geometry and Physics 4. Proceedings of the “Workshop on Strings, Membranes and Topological Field Theory,” Tohoku University, Sendai, March 2015, p. 171. arXiv:1609.09815 [hep-th]

  7. Howe, P.S., Lambert, N.D., West, P.C.: The self-dual string soliton. Nucl. Phys. B 515, 203 (1998). arXiv:hep-th/9709014

    ADS  MathSciNet  MATH  Google Scholar 

  8. Saemann, C., Wolf, M.: Non-abelian tensor multiplet equations from twistor space. Commun. Math. Phys. 328, 527 (2014). arXiv:1205.3108 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  9. Saemann, C., Wolf, M.: Six-dimensional superconformal field theories from principal 3-bundles over twistor space. Lett. Math. Phys. 104, 1147 (2014). arXiv:1305.4870 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  10. Jurco, B., Saemann, C., Wolf, M.: Semistrict higher gauge theory. JHEP 1504, 087 (2015). arXiv:1403.7185 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  11. Jurco, B., Saemann, C., Wolf, M.: Higher groupoid bundles, higher spaces, and self-dual tensor field equations. Fortschr. Phys. 64, 674 (2016). arXiv:1604.01639 [hep-th]

    MathSciNet  MATH  Google Scholar 

  12. Baez, J.C., Stevenson, D., Crans, A.S., Schreiber, U.: From loop groups to 2-groups. Homol. Homot. Appl. 9, 101 (2007). arXiv:math.QA/0504123

    MathSciNet  MATH  Google Scholar 

  13. Schommer-Pries, C.: Central extensions of smooth 2-groups and a finite-dimensional string 2-group. Geom. Top. 15, 609 (2011). arXiv:0911.2483 [math.AT]

    MathSciNet  MATH  Google Scholar 

  14. Sati, H., Schreiber, U., Stasheff, J.: Differential twisted String and Fivebrane structures. Commun. Math. Phys. 315, 169 (2012). arXiv:0910.4001 [math.AT]

    ADS  MATH  Google Scholar 

  15. Samtleben, H., Sezgin, E., Wimmer, R.: (1,0) superconformal models in six dimensions. JHEP 1112, 062 (2011). arXiv:1108.4060 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  16. Akyol, M., Papadopoulos, G.: (1,0) superconformal theories in six dimensions and Killing spinor equations. JHEP 1207, 070 (2012). arXiv:1204.2167 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  17. Singh, H.: Instantonic string solitons on M5 branes. arXiv:1704.07651 [hep-th]

  18. Bagger, J., Lambert, N.D.: Gauge symmetry and supersymmetry of multiple M2-branes. Phys. Rev. D 77, 065008 (2008). arXiv:0711.0955 [hep-th]

    ADS  MathSciNet  Google Scholar 

  19. Gustavsson, A.: Algebraic structures on parallel M2-branes. Nucl. Phys. B 811, 66 (2009). arXiv:0709.1260 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  20. Aharony, O., Bergman, O., Jafferis, D.L., Maldacena, J.M.: \({\cal{N}}=6\) superconformal Chern–Simons-matter theories, M2-branes and their gravity duals. JHEP 0810, 091 (2008). arXiv:0806.1218 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  21. Killingback, T.P.: World sheet anomalies and loop geometry. Nucl. Phys. B 288, 578 (1987)

    ADS  MathSciNet  Google Scholar 

  22. Stolz, S., Teichner, P.: What is an elliptic object? Topology, Geometry and Quantum Field Theory, volume 308 of London Mathematical Society Lecture Note series, pp. 247–343. Cambridge University Press, Cambridge (2004)

    Google Scholar 

  23. Redden, D.C.: Canonical metric connections associated to string structures, Ph.D. thesis. Notre Dame, Indiana (2006)

  24. Waldorf, K.: String connections and Chern–Simons theory. Trans. Am. Math. Soc. 365, 4393 (2013). arXiv:0906.0117 [math.DG]

    MathSciNet  MATH  Google Scholar 

  25. Chu, C.-S.: A theory of non-abelian tensor gauge field with non-abelian gauge symmetry \(G\times G\). Nucl. Phys. B 866, 43 (2013). arXiv:1108.5131 [hep-th]

    ADS  MATH  Google Scholar 

  26. Palmer, S., Saemann, C.: Six-dimensional (1,0) superconformal models and higher gauge theory. J. Math. Phys. 54, 113509 (2013). arXiv:1308.2622 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  27. Lavau, S., Samtleben, H., Strobl, T.: Hidden Q-structure and Lie 3-algebra for non-abelian superconformal models in six dimensions. J. Geom. Phys. 86, 497 (2014). arXiv:1403.7114 [math-ph]

    ADS  MathSciNet  MATH  Google Scholar 

  28. Mukhi, S., Papageorgakis, C.: M2 to D2. JHEP 0805, 085 (2008). arXiv:0803.3218 [hep-th]

    ADS  MathSciNet  Google Scholar 

  29. Palmer, S., Saemann, C.: The ABJM model is a higher gauge theory. Int. J. Geom. Methods Mod. Phys. 11, 1450075 (2014). arXiv:1311.1997 [hep-th]

    MathSciNet  MATH  Google Scholar 

  30. Chu, C.-S., Smith, D.J.: Multiple self-dual strings on M5-branes. JHEP 1001, 001 (2010). arXiv:0909.2333 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  31. Bunk, S., Saemann, C., Szabo, R.J.: The 2-Hilbert space of a prequantum bundle gerbe. Rev. Math. Phys. 30, 1850001 (2018). arXiv:1608.08455 [math-ph]

    MathSciNet  MATH  Google Scholar 

  32. Hooft, G.: Magnetic monopoles in unified gauge theories. Nucl. Phys. B 79, 276 (1974)

    ADS  MathSciNet  Google Scholar 

  33. Polyakov, A.M.: Particle spectrum in the quantum field theory. JETP Lett. 20, 194 (1974)

    ADS  Google Scholar 

  34. Prasad, M., Sommerfield, C.M.: Exact classical solution for the ’t Hooft monopole and the Julia-Zee dyon. Phys. Rev. Lett. 35, 760 (1975)

    ADS  Google Scholar 

  35. Wockel, C.: Principal 2-bundles and their gauge 2-groups. Forum Math. 23, 566 (2011). arXiv:0803.3692 [math.DG]

    MathSciNet  MATH  Google Scholar 

  36. Waldorf, K.: A global perspective to connections on principal 2-bundles. arXiv:1608.00401 [math.DG]

  37. Stolz, S.: A conjecture concerning positive Ricci curvature and the Witten genus. Math. Ann. 304, 785 (1996)

    MathSciNet  MATH  Google Scholar 

  38. Severa, P.: \(L_\infty \)-algebras as 1-jets of simplicial manifolds (and a bit beyond), arXiv:math.DG/0612349

  39. Baez, J., Crans, A.S.: Higher-dimensional algebra VI: Lie 2-algebras. Theory Appl. Categ. 12, 492 (2004). arXiv:math.QA/0307263

    MathSciNet  MATH  Google Scholar 

  40. Demessie, G.A., Saemann, C.: Higher gauge theory with string 2-groups. Adv. Theor. Math. Phys. 21, 1895 (2017). arXiv:1602.03441 [math-ph]

    MathSciNet  MATH  Google Scholar 

  41. Henriques, A.: Integrating \(L_\infty \)-algebras. Comput. Math. 144, 1017 (2008). arXiv:math.CT/0603563

    MathSciNet  MATH  Google Scholar 

  42. Myers, R.C.: Dielectric-branes. JHEP 9912, 022 (1999). arXiv:hep-th/9910053

    ADS  MathSciNet  MATH  Google Scholar 

  43. Saemann, C., Szabo, R.J.: Groupoid quantization of loop spaces, PoS CORFU2011, 46 (2012) arXiv:1203.5921 [hep-th]

  44. Fiorenza, D., Sati, H., Schreiber, U.: Multiple M5-branes, string 2-connections, and 7d nonabelian Chern–Simons theory. Adv. Theor. Math. Phys. 18, 229 (2014). arXiv:1201.5277 [hep-th]

    MathSciNet  MATH  Google Scholar 

  45. Aschieri, P., Jurco, B.: Gerbes, M5-brane anomalies and \(E_8\) gauge theory. JHEP 0410, 068 (2004). arXiv:hep-th/0409200

    ADS  Google Scholar 

  46. Palmer, S., Saemann, C.: M-brane models from non-abelian gerbes. JHEP 1207, 010 (2012). arXiv:1203.5757 [hep-th]

    ADS  MathSciNet  Google Scholar 

  47. Atiyah, M.F.: Complex analytic connections in fibre bundles. Trans. Am. Math. Soc. 85, 181 (1957)

    MathSciNet  MATH  Google Scholar 

  48. Bojowald, M., Kotov, A., Strobl, T.: Lie algebroid morphisms, Poisson sigma models, and off-shell closed gauge symmetries. J. Geom. Phys. 54, 400 (2005). arXiv:math.DG/0406445

    ADS  MathSciNet  MATH  Google Scholar 

  49. Kotov, A., Strobl, T.: Generalizing geometry—Algebroids and sigma models. In: Cortes, V. (ed.) Handbook on Pseudo-Riemannian Geometry and Supersymmetry. arXiv:1004.0632 [hep-th]

  50. Gruetzmann, M., Strobl, T.: General Yang–Mills type gauge theories for p-form gauge fields: from physics-based ideas to a mathematical framework OR from Bianchi identities to twisted Courant algebroids. Int. J. Geom. Methods Mod. Phys. 12, 1550009 (2014). arXiv:1407.6759 [hep-th]

    MathSciNet  MATH  Google Scholar 

  51. Saemann, C., Wolf, M.: On twistors and conformal field theories from six dimensions. J. Math. Phys. 54, 013507 (2013). arXiv:1111.2539 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  52. Ho, P.-M., Matsuo, Y.: Note on non-Abelian two-form gauge fields. JHEP 1209, 075 (2012). arXiv:1206.5643 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  53. Palmer, S., Saemann, C.: Self-dual string and higher instanton solutions. Phys. Rev. D 89, 065036 (2014). arXiv:1312.5644 [hep-th]

    ADS  Google Scholar 

  54. Fiorenza, D., Schreiber, U., Stasheff, J.: Čech cocycles for differential characteristic classes—an infinity-Lie theoretic construction. Adv. Theor. Math. Phys. 16, 149 (2012). arXiv:1011.4735 [math.AT]

    MathSciNet  MATH  Google Scholar 

  55. Horava, P., Witten, E.: Eleven-dimensional supergravity on a manifold with boundary. Nucl. Phys. B 475, 94 (1996). arXiv:hep-th/9603142

    ADS  MathSciNet  MATH  Google Scholar 

  56. Witten, E.: Five-brane effective action in M-theory. J. Geom. Phys. 22, 103 (1997). arXiv:hep-th/9610234

    ADS  MathSciNet  MATH  Google Scholar 

  57. Schreiber, U.: Differential cohomology in a cohesive infinity-topos. Habilitation Thesis (2011) arXiv:1310.7930 [math-ph]

  58. Adawi, T., Cederwall, M., Gran, U., Nilsson, B.E.W., Razaznejad, B.: Goldstone tensor modes. JHEP 9902, 001 (1999). arXiv:hep-th/9811145

    ADS  MathSciNet  MATH  Google Scholar 

  59. Strominger, A.: Open p-branes. Phys. Lett. B 383, 44 (1996). arXiv:hep-th/9512059

    ADS  MathSciNet  MATH  Google Scholar 

  60. Townsend, P.K.: Brane surgery. Nucl. Phys. Proc. Suppl. 58, 163 (1997). arXiv:hep-th/9609217

    ADS  MathSciNet  MATH  Google Scholar 

  61. Nepomechie, R.I.: Magnetic monopoles from antisymmetric tensor gauge fields. Phys. Rev. D 31, 1921 (1985)

    ADS  MathSciNet  Google Scholar 

  62. Akyol, M., Papadopoulos, G.: Brane solitons of (1,0) superconformal theories in six dimensions with hypermultiplets. Class. Quantum Gravity 31, 065012 (2014). arXiv:1307.1041 [hep-th]

    ADS  MATH  Google Scholar 

  63. Brunner, I., Karch, A.: Branes and six dimensional fixed points. Phys. Lett. B 409, 109 (1997). arXiv:hep-th/9705022

    ADS  MathSciNet  Google Scholar 

  64. Hanany, A., Zaffaroni, A.: Branes and six dimensional supersymmetric theories. Nucl. Phys. B 529, 180 (1998). arXiv:hep-th/9712145

    ADS  MathSciNet  MATH  Google Scholar 

  65. Baez, J.C., Lauda, A.D.: Higher-dimensional algebra V: 2-groups. Theory Appl. Categ. 12, 423 (2004). arXiv:math.QA/0307200

    MathSciNet  MATH  Google Scholar 

  66. Chatterjee, D.S.: On gerbs, Ph.D. thesis. Trinity College Cambridge (1998)

  67. Murray, M.K.: Bundle gerbes. J. Lond. Math. Soc. 54, 403 (1996). arXiv:dg-ga/9407015

    MathSciNet  MATH  Google Scholar 

  68. Murray, M.K.: An introduction to bundle gerbes. In: Garcia-Prada, O., Bourguignon, J.-P., Salamon, S. (eds.) The Many Facets of Geometry: A Tribute to Nigel Hitchin. Oxford University Press, Oxford (2010). arXiv:0712.1651 [math.DG]

    Google Scholar 

  69. Zwiebach, B.: Closed string field theory: quantum action and the B-V master equation. Nucl. Phys. B 390, 33 (1993). arXiv:hep-th/9206084

    ADS  MathSciNet  Google Scholar 

  70. Lada, T., Stasheff, J.: Introduction to SH Lie algebras for physicists. Int. J. Theor. Phys. 32, 1087 (1993). arXiv:hep-th/9209099

    MathSciNet  MATH  Google Scholar 

  71. Lada, T., Markl, M.: Strongly homotopy Lie algebras. Commun. Algebra 23, 2147 (1995). arXiv:hep-th/9406095

    MathSciNet  MATH  Google Scholar 

  72. Fregier, Y., Rogers, C.L., Zambon, M.: Homotopy moment maps. Adv. Math. 303, 954 (2016). arXiv:1304.2051 [math.DG]

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank Branislav Jurčo, Michael Murray, Urs Schreiber, Konrad Waldorf and Martin Wolf for useful discussions. C.S. was supported in part by the STFC Consolidated Grant ST/L000334/1 Particle Theory at the Higgs Centre. L.S. was supported by an STFC PhD studentship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Sämann.

Additional information

Publisher's Note

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

Appendix

Appendix

Below, we collect some definitions, results and hopefully helpful references. While “Appendix A” is a rough overview over the categorification of the mathematical notions involved in the definition of principal bundles, “Appendices B, C and D” explain in detail various aspects of categorified gauge algebras. Potentially useful and more detailed introductions into some of the following material are found in [5, 6].

1.1 Categorified Lie groups, Lie algebras and principal bundles

Let \({\textsf {Mfd}}^\infty \) be the category of smooth spaces with smooth maps between them. A 2-space (e.g., [2]) is formally a category internal toFootnote 16\({\textsf {Mfd}}^\infty \). That is, a 2-space consists of two manifolds \(M_1, M_0\) with smooth source and target maps \(\textsf {s},\textsf {t}{:}\,M_1\rightrightarrows M_0\), a smooth embedding \(\mathrm {id}{:}\,M_0{\hookrightarrow }M_1\) and a smooth composition of morphisms \(\circ {:}\, M_1\times _{M_0} M_1\rightarrow M_1\) satisfying the usual axioms of the structure maps of a category. Morphisms between 2-spaces can be defined as functors internal to \({\textsf {Mfd}}^\infty \), which are smooth maps between the morphism- and object-manifolds of the 2-spaces.Footnote 17 Note that any manifold M gives trivially rise to a 2-space \(M\rightrightarrows M\) in which the only morphisms are the identity morphisms. Also, Lie groupoids are 2-spaces in which each morphism is invertible.

In its simplest, strict form, a Lie 2-group is a category internal to \({\textsf {Grp}}\), the category of groups. Such strict Lie 2-groups are most conveniently described as crossed modules of Lie groups [65]. More generally, a Lie 2-group is a Lie groupoid which is simultaneously a monoidal category and in which morphisms are strictly invertible and objects are weakly invertible with respect to the monoidal product [65].

There is a higher analogue of Lie differentiation, which takes a Lie 2-group to a Lie 2-algebra [39]. Strict Lie 2-groups differentiate to strict Lie 2-algebras, which are conveniently described by crossed modules of Lie algebras. More generally, a Lie 2-group differentiates to a 2-term \(L_\infty \)-algebra [38], see also [11] for a detailed discussion of an extremely general case. Since we use these extensively, we present some more details in “Appendices B, C and D”.

Let us now come to the categorification of principal bundles. As explained in Sect. 2.1, one can use any of the equivalent definitions of principal bundles and replace all notions by categorified ones. To describe transition functions of principal \(\textsf {G}\)-bundles, one can use the approach via functors from the Čech groupoid\(\check{\mathscr {C}}(Y)\) of a surjective submersion \(\sigma {:}\,Y\twoheadrightarrow M\) over some manifold M to the category \(\textsf {B}\textsf {G}=\textsf {G}\rightrightarrows *\). One may imagine Y to be an open cover \(Y=\sqcup U_i\), for concreteness sake. There are now two obvious projections from the fiber product \(Y^{[2]}=Y\times _M Y\) (which is the space of double overlaps in the case \(Y=\sqcup U_i\)) to Y as well as a diagonal embedding \(Y{\hookrightarrow }Y^{[2]}\), which, together with the composition \((y_1,y_2)\circ (y_2,y_3)=(y_1,y_3)\) we can use to form the 2-space

$$\begin{aligned} \check{\mathscr {C}}(Y)=(Y^{[2]}\rightrightarrows Y). \end{aligned}$$
(A.1)

Since morphisms are invertible with \((y_1,y_2)^{-1}=(y_2,y_1)\), this is in fact a Lie groupoid. The transition functions of a principal \(\textsf {G}\)-bundle for some Lie group \(\textsf {G}\) are given by functors \(g{:}\,\check{\mathscr {C}}(Y)\rightarrow \textsf {B}\textsf {G}\) and isomorphism or gauge transformations correspond to natural transformations between functors. This picture readily extends to categorified groups, where we consider higher functors between the Čech groupoid, trivially regarded as a higher Lie groupoid, and the delooping \(\textsf {B}\mathscr {G}\) of a higher Lie group \(\mathscr {G}\). For \(\mathscr {G}=\textsf {B}\textsf {U}(1)\), we recover abelian gerbes, or principal \(\textsf {B}\textsf {U}(1)\)-bundles in the form of Hitchin–Chatterjee gerbes [66]. These are stably isomorphic to Murray’s more general and more useful bundle gerbes [67, 68]. Abelian gerbes, or principal \(\textsf {B}\textsf {U}(1)\) bundles are described by a characteristic class, called the Dixmier–Douady class in \(H^3(M,\mathbb {Z})\), which is the analogue of the first Chern class of line bundles in \(H^2(M,\mathbb {Z})\).

To add connections, one can either glue together the local description as derived in Sect. 3.2 or use insights from Lie differentiation, as done in [10, 11].

1.2 \(L_\infty \)-algebras and Lie n-algebras

Strong homotopy Lie algebras or \(L_\infty \)-algebras comprise Lie algebras and are useful descriptions of all their categorifications. They play important roles in BV quantization, string field theory and higher geometry in general; the original references are [69,70,71].

An \(L_\infty \)-algebra is an \(\mathbb {Z}\)-graded vector space \(\textsf {L}=\oplus _{k\in \mathbb {Z}}\textsf {L}_k\) which is endowed with a set of totally antisymmetric, multilinear maps \(\mu _i{:}\,\wedge ^i\textsf {L}\rightarrow \textsf {L}\), \(i\in \mathbb {N}\), of degree \(i-2\), which satisfy the higher or homotopy Jacobi relations

$$\begin{aligned} \sum _{r+s=i}\sum _{\sigma }(-1)^{rs}\chi (\sigma ;\ell _1,\dots ,\ell _{r+s})\mu _{s+1}(\mu _r(\ell _{\sigma (1)},\dots ,\ell _{\sigma (r)}),\ell _{\sigma (r+1)},\dots ,\ell _{\sigma (r+s)})\ =\ 0 \end{aligned}$$
(B.1)

for all \(\ell _1,\dots ,\ell _{r+s}\in \textsf {L}\), where the second sum runs over all (rs) unshuffles, i.e., permutations \(\sigma \) of \(\{1,\dots ,r+s\}\) with the first r and the last s images of \(\sigma \) ordered: \(\sigma (1)<\dots <\sigma (r)\) and \(\sigma (r+1)<\cdot <\sigma (r+s)\). Also, \(\chi (\sigma ;\ell _1,\dots ,\ell _i)\) denotes the graded Koszul sign defined through the graded antisymmetrized products

$$\begin{aligned} \ell _1\wedge \dots \wedge \ell _i\ =\ \chi (\sigma ;\ell _1,\dots ,\ell _i)\,\ell _{\sigma (1)}\wedge \dots \wedge \ell _{\sigma (i)}. \end{aligned}$$
(B.2)

If an \(L_\infty \)-algebra is non-trivial in degrees \(0,\dots ,n-1\), we call it an n-term \(L_\infty \)-algebra. This is a useful and very general notion of a Lie n-algebra. We therefore often use the term Lie n-algebra, when we actually mean an n-term \(L_\infty \)-algebra.

A cyclic structure on an \(L_\infty \)-algebra over \(\mathbb {R}\) is a graded symmetric, non-degenerate bilinear form

$$\begin{aligned} \langle -,-\rangle {:}\,\textsf {L}\odot \textsf {L}\ \rightarrow \ \mathbb {R}, \end{aligned}$$
(B.3)

such that

$$\begin{aligned} \langle \ell _1,\mu _i(\ell _2,\dots ,\ell _{i+1})\rangle \ =\ (-1)^{i+|\ell _{i+1}|(|\ell _1|+\dots +|\ell _i|)}\langle \ell _{i+1},\mu _i(\ell _1,\dots ,\ell _{i})\rangle \end{aligned}$$
(B.4)

for \(\ell _1,\dots ,\ell _{i+1}\in \textsf {L}\).

We are particularly interested in Lie 2- and 3-algebras. The lowest homotopy Jacobi relations are equivalent to the following ones:

$$\begin{aligned} \mu _1(\mu _1(\ell _1))= & {} 0,\nonumber \\ \mu _1(\mu _2(\ell _1,\ell _2))= & {} \mu _2(\mu _1(\ell _1),\ell _2)+(-1)^{|\ell _1|}\mu _2(\ell _1,\mu _1(\ell _2)),\nonumber \\ \mu _1(\mu _3(\ell _1,\ell _2,\ell _3))= & {} -\mu _3(\mu _1(\ell _1),\ell _2,\ell _3)-(-1)^{|\ell _1|}\mu _3(\ell _1,\mu _1(\ell _2),\ell _3)\nonumber \\&-\,(-1)^{(|\ell _1|+|\ell _2|)}\mu _3(\ell _1,\ell _2,\mu _1(\ell _3))+\mu _2(\ell _1,\mu _2(\ell _2,\ell _3))\nonumber \\&-\,\mu _2(\mu _2(\ell _1,\ell _2),\ell _3)-(-1)^{|\ell _1|\,|\ell _2|}\mu _2(\ell _2,\mu _2(\ell _1,\ell _3)).\nonumber \\ \end{aligned}$$
(B.5)

The first two identities are the same as for a differential graded Lie algebra, but the third identity is a controlled lifting of the Jacobi identity.

A cyclic structure on a Lie 3-algebra satisfies

$$\begin{aligned} \begin{aligned} \langle \ell _1,\mu _1(\ell _2)\rangle&= (-1)^{1+|\ell _2|\,|\ell _1|}\langle \ell _2,\mu _1(\ell _1)\rangle ,\\ \langle \ell _1,\mu _2(\ell _2,\ell _3)\rangle&= (-1)^{1+|\ell _3|(|\ell _1|+|\ell _2|)+|\ell _1|\,|\ell _2|}\langle \ell _3,\mu _2(\ell _2,\ell _1)\rangle ,\\ \langle \ell _1,\mu _3(\ell _2,\ell _3,\ell _4)\rangle&= (-1)^{1+|\ell _4|(|\ell _1|+|\ell _2|+\ell _3)+|\ell _3|(|\ell _1|+|\ell _2|)}\langle \ell _4,\mu _3(\ell _2,\ell _3,\ell _1)\rangle ,\\ \langle \ell _1,\mu _4(\ell _2,\ell _3,\ell _4,\ell _5)\rangle&= (-1)^{1+|\ell _5|(|\ell _1|+|\ell _2|+\ell _3+\ell _4)+|\ell _4|(|\ell _1|+|\ell _2|+|\ell _3|)}\langle \ell _5,\mu _4(\ell _2,\ell _3,\ell _4,\ell _1)\rangle \end{aligned} \end{aligned}$$
(B.6)

for all \(\ell _1,\dots ,\ell _5\in \textsf {L}\).

1.3 Lie n-algebras as NQ-manifolds

A very useful and elegant definition of \(L_\infty \)-algebras can be given in terms of NQ-manifolds, which are known to physicists from BRST quantization and string field theory. In this picture, the cyclic structure arises from a symplectic form. Below, we briefly explain this point of view.

An NQ-manifold\((\mathcal {M},Q)\) is an \(\mathbb {N}_0\)-graded manifold \(\mathcal {M}\) endowed with a vector field Q which is of degree 1 and nilquadratic: \(Q^2=0\). Due to a similar argument as that for smooth (real) supermanifolds, NQ-manifolds can be regarded as \(\mathbb {N}\)-graded vector bundles over the body \(\mathcal {M}_0\) of the manifold \(\mathcal {M}\).

Archetypical examples of NQ-manifolds are \((T[1]M,\mathrm {d})\), the grade-shifted tangent bundle together with the de Rham differential as well as the grade-shifted Lie algebra \((\mathfrak {g}[1],Q)\) with \(Q=-\tfrac{1}{2} f_{\alpha \beta }^\gamma \xi ^\alpha \xi ^\beta \frac{\partial }{\partial \xi ^\gamma }\) a vector field of degree 1 in some coordinates \(\xi ^\alpha \) on \(\mathfrak {g}[1]\), which are necessarily of degree 1. Note that \(Q^2=0\) is equivalent to the Jacobi identity in the latter case.

The latter example indicates the relation of NQ-manifolds to Lie n-algebras: Given a Lie n-algebra \(\textsf {L}\), we should first grade-shift the underlying graded vector space:

$$\begin{aligned} \textsf {L}=(\textsf {L}_0\leftarrow \dots \leftarrow \textsf {L}_{n-1})\quad \rightarrow \quad \textsf {L}[1]=(*\leftarrow \textsf {L}_0[1]\leftarrow \dots \leftarrow \textsf {L}_{n-1}[1]). \end{aligned}$$
(C.1)

Correspondingly, the degree of the maps \(\mu _i\) changes from \(i-2\) to \(-1\). The sum of these maps forms a codifferential

$$\begin{aligned} D=\mu _1+\mu _2+\mu _3+\cdots , \end{aligned}$$
(C.2)

which acts on the coalgebra \(\wedge ^\bullet \textsf {L}[1]\). If we now dualize to the algebra of functions on \(\textsf {L}[1]\), the corresponding differential is a vector field of degree 1 on the NQ-manifold

$$\begin{aligned} *\leftarrow \textsf {L}_0[1]\leftarrow \cdots \leftarrow \textsf {L}_{n-1}[1]. \end{aligned}$$
(C.3)

The condition \(Q^2=0\) translates to (B.1).

Altogether, we arrived at a description of a Lie n-algebra \(\textsf {L}\) as a differential graded algebra\(\big (\mathcal {C}^\infty (\textsf {L}[1]), Q\big )\). This differential graded algebra is also called the Chevalley–Eilenberg algebra\(\mathrm{CE}(\textsf {L})\) of \(\textsf {L}\).

It is not hard to see that a cyclic structure on a Lie n-algebra \(\textsf {L}\) corresponds to a symplectic form on \(\textsf {L}[1]\) of degree \(n+1\), and we use this in Sect. 6.1. For example, the inner product \((x_1,x_2)=g_{\alpha \beta }\xi _1^\alpha \xi _2^\beta \) on a Lie algebra \(\mathfrak {g}\) is described by the symplectic form \(\omega =\tfrac{1}{2} g_{\alpha \beta }\mathrm {d}\xi ^\alpha \wedge \mathrm {d}\xi ^\beta \) of degree 2 on \(\mathfrak {g}[1]\).

1.4 Morphisms and equivalences of Lie n-algebras

Let us now come to categorical equivalence between Lie n-algebras. For simplicity, we first restrict ourselves to Lie 2-algebras \(\textsf {L}=W\leftarrow V\). A morphism of Lie n-algebras, being a morphism of graded spaces, is most readily derived in the Chevalley–Eilenberg picture. There, a morphism \(\Phi {:}\,\textsf {L}\rightarrow \tilde{\textsf {L}}\) between Lie 2-algebras \(\textsf {L}=(W\leftarrow V)\) and \(\tilde{\textsf {L}}=\tilde{W}\leftarrow \tilde{V})\) is given by a morphism of differential graded algebras \(\Phi ^*\). That is,

$$\begin{aligned} \Phi ^*{:}\,\mathcal {C}^\infty (\tilde{\textsf {L}}[1])\rightarrow \mathcal {C}^\infty (\textsf {L}[1]),\quad \tilde{Q}\circ \Phi ^*= \Phi ^*\circ Q. \end{aligned}$$
(D.1)

Since \(\textsf {L}[1]\) is a graded vector space, this morphism is determined by its image on coordinate functions. Because degrees have to be preserved, \(\Phi ^*\) (and thus \(\Phi \)) is encoded in maps

$$\begin{aligned} \phi _0{:}\,W\rightarrow \tilde{W},\quad \phi _0{:}\,V\rightarrow \tilde{V},\quad \phi _1{:}\,W\wedge W\rightarrow \tilde{V}. \end{aligned}$$
(D.2)

The fact that \(\Phi ^*\) is a morphism of differential graded algebras implies that

$$\begin{aligned} \begin{aligned} \phi _0\left( \mu _2(w_1,w_2)\right)&=\tilde{\mu }_2\left( \phi _0(w_1),\phi _0(w_2)\right) +\tilde{\mu }_{1}(\phi _1(w_1,w_2)),\\ \phi _0\left( \mu _2(w,v)\right)&=\tilde{\mu }_2(\phi _0(w),\phi _0(v))+\phi _1 (w,\mu _{1}(v)),\\ \phi _0\left( \mu _3(w_1,w_2,w_3)\right)&=\tilde{\mu }_3(\phi _0(w_1),\phi _0(w_2), \phi _0(w_3))+\left[ \phi _1(w_1,\mu _2(w_2,w_3))\right. \\&\quad \left. +\,\tilde{\mu }_2\left( \phi _0(w_1),\phi _1(w_2,w_3)\right) +\text {cyclic}\,(w_1,w_2,w_3)\right] \end{aligned} \end{aligned}$$
(D.3)

for all \(w,w_{1,2}\in W\) and \(v\in V\). This reproduces the definition of a morphism of Lie 2-algebra from [39].

Two morphisms \(\Phi =(\phi _0,\phi _1)\) and \(\Psi =(\psi _0,\psi _1)\) compose as follows:

$$\begin{aligned} (\Psi \circ \Phi )_0(\ell )=\psi _0(\phi _0(\ell )),\quad (\Psi \circ \Phi )_{1}(w_1,w_2)=\psi _0(\phi _1(w_1,w_2))+\psi _1(\phi _0(w_1),\phi _0(w_2)) \end{aligned}$$
(D.4)

for all \(\ell \in \textsf {L}\) and \(w_{1,2}\in W\). The identity morphism reads as \(\mathrm {id}_\textsf {L}=(\mathrm {id}_\textsf {L},0)\). The above data of 2-term \(L_\infty \)-algebras and their morphisms, together with the identity morphism and composition combines to a category of Lie 2-algebras, \({\textsf {Lie2alg}}\).

Note that the inverse of a morphism \(\Phi =(\phi _0,\phi _1)\) is defined if and only if \(\phi _0\) is invertible:

$$\begin{aligned} (\Phi ^{-1})_0(\ell )=\phi _0^{-1}(\ell ),\quad (\Phi ^{-1})_{-1}(w_1,w_2) =-\phi _0^{-1}(\phi _1(\phi _0^{-1}(w_1),\phi _0^{-1}(w_2))), \end{aligned}$$
(D.5)

again for \(\ell \in \textsf {L}\) and \(w_{1,2}\in W\).

A 2-morphism between two morphisms \(\Phi ,\Psi {:}\,\textsf {L}\rightarrow \tilde{\textsf {L}}\) is a chain homotopy \(\chi {:}\,W\rightarrow \tilde{V}\) such that

$$\begin{aligned} \phi _1(w_1,w_2)-\psi _1(w_1,w_2)=\mu _2(w_1,\chi (w_2))+\mu _2(\chi (w_1), \psi _0(w_2))-\chi (\mu _2(w_1,w_2)) \end{aligned}$$
(D.6)

for all \(w_1,w_2\).

Finally, an equivalence or a quasi-isomorphism of Lie 2-algebras between Lie 2-algebras \(\textsf {L}=W\xleftarrow {~\mu _1~}V\) and \(\tilde{\textsf {L}}=\tilde{W}\xleftarrow {~\tilde{\mu }_1~}\tilde{V}\) is a morphism \(\Phi _1{:}\,\textsf {L}\rightarrow \tilde{\textsf {L}}\) and a morphism \(\Phi _2{:}\,\tilde{\textsf {L}}\rightarrow \textsf {L}\) such that \(\Phi _1\circ \Phi _2\cong \mathbb {1}_{\tilde{frg}}\) and \(\Phi _2\circ \Phi _1\cong \mathbb {1}_{\textsf {L}}\). An explicit example of an equivalence between Lie 2-algebras is found in Sect. 2.3.

The above discussion readily generalizes to morphisms \(\Phi \) between Lie n-algebras \(\textsf {L}\) and \(\tilde{\textsf {L}}\). These are given by totally antisymmetric maps \(\phi _i{:}\,\textsf {L}^{\wedge i+1}\rightarrow \textsf {L}\), \(i=0,\dots ,n-1\), of degree \(-i\) such that appropriate extensions of (D.3) hold, see [72] for details. An isomorphism of Lie n-algebras is a morphism of Lie n-algebras with \(\phi _0\) an isomorphisms. Equivalences between \(L_\infty \)-algebras are then captured by Lie n-algebra quasi-isomorphisms which are morphisms of Lie n-algebras which induce an isomorphism on the cohomology of the complex underlying the Lie n-algebras.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sämann, C., Schmidt, L. The non-abelian self-dual string. Lett Math Phys 110, 1001–1042 (2020). https://doi.org/10.1007/s11005-019-01250-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-019-01250-3

Keywords

Mathematics Subject Classification

Navigation