Skip to main content
Log in

M-theory from the superpoint

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

Abstract

The “brane scan” classifies consistent Green–Schwarz strings and membranes in terms of the invariant cocycles on super Minkowski spacetimes. The “brane bouquet” generalizes this by consecutively forming the invariant higher central extensions induced by these cocycles, which yields the complete fundamental brane content of string/M-theory, including the D-branes and the M5-brane, as well as the various duality relations between these. This raises the question whether the super Minkowski spacetimes themselves arise as maximal invariant central extensions. Here, we prove that they do. Starting from the simplest possible super Minkowski spacetime, the superpoint, which has no Lorentz structure and no spinorial structure, we give a systematic process of consecutive “maximal invariant central extensions” and show that it discovers the super Minkowski spacetimes that contain superstrings, culminating in the 10- and 11-dimensional super Minkowski spacetimes of string/M-theory and leading directly to the brane bouquet.

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. The classification of these cocycles is also discussed by Movshev et al. [43] and Brandt [9,10,11]. A unified derivation of the cocycle conditions is given by Baez and Huerta [4, 5]. See also Foot and Joshi [31].

  2. A partial completion of the old brane scan can be achieved by classifying superconformal structures that may appear in the near horizon geometry of “solitonic” or “black” p-branes [8, 20].

  3. Notice that these are Lie n-algebras in the sense of Stasheff [38, 39, 47] as originally found in string field theory by Zwiebach [54, Section 4] not “n-Lie algebras” in the sense of Filippov. However, the two notions are not unrelated. At least the Filippov 3-Lie algebras that appear in the Bagger–Lambert model for coincident solitonic M2-branes may naturally be understood as Stasheff Lie 2-algebras equipped with a metric form [44, Section 2].

  4. Unfortunately, the “free differential algebras” of D’Auria and Fré are not free. In the parlance of modern mathematics, they are differential graded commutative algebras, where the underlying graded commutative algebra is free, but the differential is not. We will thus refer to them as “FDA”s, with quotes.

  5. The double arrows stand for the two different canonical inclusions of \({\mathbb {R}}^{d-1,1\vert N}\) into \({\mathbb {R}}^{d-1,1\vert N + N}\), being the identity on \({\mathbb {R}}^{d-1,1}\) and sending N identically either to the first or to the second copy in the direct sum \(N + N\).

References

  1. Achúcarro, A., Evans, J., Townsend, P., Wiltshire, D.: Super \(p\)-branes. Phys. Lett. B 198, 441–446 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  2. Baez, J.: The octonions. Bull. Am. Math. Soc. 39, 145–205 (2002). Errata in Bull. Am. Math. Soc. 42, 213 (2005). http://math.ucr.edu/home/baez/octonions/octonions.html

    Article  MathSciNet  MATH  Google Scholar 

  3. Baez, J., Crans, A.: Higher-dimensional algebra VI: Lie 2-algebras. Theory Appl. Categ. 12, 492–528 (2004). arXiv:hep-th/0307263

  4. Baez, J., Huerta, J.: Division algebras and supersymmetry I. In Doran, R., Friedman, G., Rosenberg, J. (eds.) Superstrings, Geometry, Topology, and \(C^\ast \)-algebras, Proc. Symp. Pure Math., vol. 81. AMS, Providence (2010), arXiv:0909.0551

  5. Baez, J., Huerta, J.: Division algebras and supersymmetry II. Adv. Theor. Math. Phys. 15, 1373–1410 (2011). arXiv:1003.3436

    Article  MathSciNet  MATH  Google Scholar 

  6. Bandos, I., Lechner, K., Nurmagambetov, A., Pasti, P., Sorokin, D., Tonin, M.: Covariant action for the super-fivebrane of M-theory. Phys. Rev. Lett. 78, 4332 (1997). arXiv:hep-th/9701149

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Bergshoeff, E., Sezgin, E., Townsend, P.: Supermembranes and eleven dimensional supergravity. Phys. Lett. B 189, 75–78 (1987)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Blencowe, M., Duff, M.: Supersingletons. Phys. Lett. B 203, 229–236 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  9. Brandt, F.: Supersymmetry algebra cohomology II: primitive elements in 2 and 3 dimensions. J. Math. Phys. 51, 112303 (2010). arXiv:1004.2978

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Brandt, F.: Supersymmetry algebra cohomology III: primitive elements in four and five dimensions. J. Math. Phys. 52, 052301 (2011). arXiv:1005.2102

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Brandt, F.: Supersymmetry algebra cohomology IV: primitive elements in all dimensions from D \(=\) 4 to D \(=\) 11. J. Math. Phys. 54, 052302 (2013). arXiv:1303.6211

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Candiello, A., Lechner, K.: Duality in supergravity theories. Nucl. Phys. B 412, 479–501 (1994). arXiv:hep-th/9309143

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Chryssomalakos, C., de Azcárraga, J., Izquierdo, J., Pérez, C.: The geometry of branes and extended superspaces. Nucl. Phys. B 567, 293–330 (2000). arXiv:hep-th/9904137

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Claus, P., Kallosh, R., Van Proeyen, A.: M5-brane and superconformal \((0,2)\) tensor multiplet in 6 dimensions. Nucl. Phys. B 518, 117–150 (1998). arXiv:hep-th/9711161

    Article  ADS  MATH  Google Scholar 

  15. Claus, P., Kallosh, R., Kumar, J., Townsend, P., Van Proeyen, A.: Conformal theory of M2, D3, M5 and D1 \(+\) D5 branes. J. High Energy Phys. 9806, 004 (1998). arXiv:hep-th/9801206

    Article  ADS  Google Scholar 

  16. D’Auria, R., Fré, P.: Geometric supergravity in D \(=\) 11 and its hidden supergroup. Nucl. Phys. B 201, 101–140 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  17. de Azcárraga, J., Townsend, P.: Superspace geometry and the classification of supersymmetric extended objects. Phys. Rev. Lett. 62(22), 2579–2582 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  18. de Azcárraga, J., Izquierdo, J.M.: Lie Groups, Lie Algebras, Cohomology and Some Applications in Physics, Cambridge Monographs of Mathematical Physics. Cambridge University Press, Cambridge (1995)

    Book  MATH  Google Scholar 

  19. Duff, M.: Supermembranes: the first fifteen weeks. Class. Quantum Gravity 5, 189 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  20. Duff, M.: Near-horizon brane-scan revived. Nucl. Phys. B 810, 193–209 (2009). arXiv:0804.3675

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Duff, M.: The World in Eleven Dimensions: Supergravity. Supermembranes and M-Theory. IOP, Bristol (1999)

    Book  MATH  Google Scholar 

  22. Duff, M., Howe, P., Inami, T., Stelle, K.: Superstrings in D \(=\) 10 from supermembranes in D \(=\) 11. Phys. Lett. B 191, 70 (1987)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Duff, M., Sutton, C.: The membrane at the end of the Universe. New Sci. 118, 67–71 (1988)

    Google Scholar 

  24. Evans, J.M.: Supersymmetric Yang–Mills theories and division algebras. Nucl. Phys. B 298, 92–108 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  25. Fiorenza, D., Rogers, C., Schreiber, U.: \(L_\infty \)-algebras of local observables. Homol. Homotopy Appl. 16, 107–142 (2014). arXiv:1304.6292

    Article  MATH  Google Scholar 

  26. Fiorenza, D., Sati, H., Schreiber, U.: Multiple M5-branes, string 2-connections, and 7d nonabelian Chern–Simons theory. Adv. Theor. Math. Phys. 18, 229–321 (2014). arXiv:1201.5277

    Article  MathSciNet  MATH  Google Scholar 

  27. Fiorenza, D., Sati, H., Schreiber, U.: Super lie \(n\)-algebra extensions, higher WZW models, and super \(p\)-branes with tensor multiplet fields. Int. J. Geom. Methods Mod. Phys. 12, 1550018 (2015). arXiv:1308.5264

    Article  MathSciNet  MATH  Google Scholar 

  28. Fiorenza, D., Sati, H., Schreiber, U.: Rational sphere valued supercocycles in M-theory and type IIA string theory. J. Geom. Phys. 114, 91–108 (2017). arXiv:1606.03206

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Fiorenza, D., Sati, H., Schreiber, U.: T-Duality from super Lie \(n\)-algebra cocycles for super p-branes arXiv:1611.06536

  30. Freed, D.: Five Lectures on Supersymmetry. American Mathematical Society, Providence (1999)

    MATH  Google Scholar 

  31. Foot, R., Joshi, G.C.: On a certain supersymmetric identity and the division algebras. Mod. Phys. Lett. A 3, 999–1004 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  32. Green, M., Schwarz, J.: Covariant description of superstrings. Phys. Lett. B 136, 367–370 (1984)

    Article  ADS  Google Scholar 

  33. Henneaux, M., Mezincescu, L.: A \(\sigma \)-model interpretation of Green–Schwarz covariant superstring action. Phys. Lett. B 152, 340–342 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  34. Horowitz, G., Susskind, L.: Bosonic M theory. J. Math. Phys. 42, 3152–3160. arXiv:hep-th/0012037

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Howe, P.: Weyl superspace. Phys. Lett. B 415, 149–155 (1997). arXiv:hep-th/9707184

    Article  ADS  MathSciNet  Google Scholar 

  36. Huerta, J.: Division algebras, supersymmetry and higher gauge theory. Ph.D. Thesis, Department of Mathematics, UC Riverside (2011). arxiv:1106.3385

  37. Huerta, J., Sati, H., Schreiber, U.: Real ADE-equivariant (co)homotopy and super M-branes. arXiv:1805.05987

  38. Lada, T., Stasheff, J.: Introduction to SH Lie algebras for physicists. Int. J. Theor. Phys. 32, 1087–1103 (1993). arXiv:hep-th/9209099

    Article  MathSciNet  MATH  Google Scholar 

  39. Lada, T., Markl, M.: Strongly homotopy Lie algebras. Commun. Algebra 23, 2147–2161 (1995). arXiv:hep-th/9406095

    Article  MathSciNet  MATH  Google Scholar 

  40. Lawson, H.B., Michelsohn, M.-L.: Spin Geometry. Princeton University Press, Princeton (1989)

    MATH  Google Scholar 

  41. Lott, J.: The geometry of supergravity torsion constraints. Commun. Math. Phys. 133, 563–615 (1990). arXiv:math/0108125

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. Moore, G.: Physical mathematics and the future, talk at Strings (2014). http://www.physics.rutgers.edu/~gmoore/PhysicalMathematicsAndFuture.pdf

  43. Movshev, M.V., Schwarz, A., Xu, R.: Homology of Lie algebra of supersymmetries and of super Poincaré Lie algebra. Nucl. Phys. B 854, 483–503 (2012). arXiv:1106.0335

    Article  ADS  MATH  Google Scholar 

  44. Palmer, S., Sämann, C.: M-brane models from non-abelian gerbes. J. High Energy Phys. 1207, 010 (2012). arXiv:1203.5757

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Polchinksi, J.: M-theory and the light cone. Prog. Theor. Phys. Suppl. 134, 158–170 (1999). arXiv:hep-th/9903165

    Article  ADS  MathSciNet  Google Scholar 

  46. Sakaguchi, M.: IIB-branes and new spacetime superalgebras. J. High Energy Phys. 0004, 019 (2000). arXiv:hep-th/9909143

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. 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, Competitive models, pp. 303–424. Birkhäuser, Basel (2009). arXiv:0801.3480

  48. Sati, H., Schreiber, U., Stasheff, J.: Twisted differential string- and fivebrane structures. Commun. Math. Phys. 315(1), 169–213 (2012). arXiv:0910.4001

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Schray, J.: The general classical solution of the superparticle. Class. Quantum Gravity 13, 27–38 (1996). arXiv:hep-th/9407045

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. Sorokin, D.: Superbranes and superembeddings. Phys. Rep. 329, 1–101 (2000). arXiv:hep-th/9906142

    Article  ADS  MathSciNet  MATH  Google Scholar 

  51. Sudbery, A.: Division algebras, (pseudo)orthogonal groups and spinors. J. Phys. A 17, 939–955 (1984)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. Varadarajan, V.S.: Supersymmetry for Mathematicians: An Introduction, Courant Lecture Notes in Mathematics, vol. 11. American Mathematical Society, Providence (2004)

    Google Scholar 

  53. Witten, E.: Magic, mystery, and matrix. Notices Am. Math. Soc. 45, 1124–1129 (1998). www.ams.org/notices/199809/witten.pdf

  54. Zwiebach, B.: Closed string field theory: quantum action and the Batalin–Vilkovisky master equation. Nucl. Phys. B 390, 33–152 (1993). arXiv:hep-th/9206084

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

We thank David Corfield for comments on an earlier version of this article. We also thank John Baez and an anonymous referee for comments and corrections. We thank the Max Planck Institute for Mathematics in Bonn for kind hospitality while the result reported here was conceived. U.S. thanks Roger Picken for an invitation to Instituto Superior Técnico, Lisbon, where parts of this article were written. U.S. was supported by RVO:67985840. J.H. was supported by the Portuguese science foundation Grant SFRH/BPD/92915/2013.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John Huerta.

A Background

A Background

For reference, we briefly recall some definitions and facts that we use in the main text.

1.1 A.1 Super Lie algebra cohomology

We recall the definition of super Lie algebras and their cohomology. All our vector spaces and algebras are over \({\mathbb {R}}\), and they are all finite dimensional. We write even for \(0 \in {\mathbb Z}_2\) and odd for \(1 \in {\mathbb Z}_2\).

Definition 15

The tensor category of supervector spaces is the category of \({\mathbb Z}_2\)-graded vector spaces and grade-preserving linear maps, equipped with the unique non-trivial braiding, \(\tau ^{\mathrm{super}}\). For any two supervector space V and W, \(\tau ^{\mathrm{super}}\) is the isomorphism

$$\begin{aligned} \begin{array}{rcl} \tau ^{\mathrm{super}} :V \otimes W &{} \rightarrow &{} W \otimes V \\ v_1 \otimes v_2 &{} \mapsto &{} (-1)^{\sigma _1 \sigma _2} \, v_2 \otimes v_1 , \end{array} \end{aligned}$$

for elements \(v_1 \in V\), \(v_2 \in W\) of homogeneous degree \(\sigma _i \in {\mathbb Z}_2\).

A super Lie algebra is a Lie algebra internal to supervector spaces. That is, it is a supervector space

$$\begin{aligned} \mathfrak {g} = \mathfrak {g}_{\mathrm {even}} \oplus \mathfrak {g}_{\mathrm {odd}} \end{aligned}$$

equipped with a bilinear map, called the Lie bracket:

$$\begin{aligned}{}[-,-] :\mathfrak {g} \otimes \mathfrak {g} \longrightarrow \mathfrak {g} \end{aligned}$$

which is graded skew symmetric:

$$\begin{aligned}{}[v_1, v_2] = -(-1)^{\sigma _1 \sigma _2} [v_2, v_1] \end{aligned}$$

and which satisfies the graded Jacobi identity:

$$\begin{aligned}{}[v_1, [v_2,v_3]] = [[v_1, v_2], v_3] + (-1)^{ \sigma _1 \sigma _2 } [v_2, [v_1, v_3]]. \end{aligned}$$

A homomorphism of super Lie algebras \(\mathfrak {g}_1 \longrightarrow \mathfrak {g}_2\) is a linear map preserving the \({\mathbb Z}_2\)-grading and the bracket.

Definition 16

(super Lie algebra cohomology) Let V be a finite-dimensional supervector space. Then, the super Grassmann algebra \(\Lambda ^\bullet V^*\) is the \({\mathbb {Z}} \times {\mathbb Z}_2\)-bigraded-commutative associative algebra freely generated by \(V^*\) in degree \(1 \in {\mathbb Z}\). That is to say, it is generated by the elements in \(V^*_{\mathrm {even}}\) regarded as being in bidegree \((1,\mathrm {even})\), and the elements in \(V^*_{\mathrm {odd}}\) regarded as being in bidegree \((1,\mathrm {odd})\), subject to the relation that for \(\alpha _i\) two elements of homogeneous bidegree \((n_i, \sigma _i)\), then

$$\begin{aligned} \alpha _1 \wedge \alpha _2 \;=\; (-1)^{n_1 n_2} (-1)^{\sigma _1 \sigma _2} \, \alpha _2 \wedge \alpha _1. \end{aligned}$$

In particular, this relation says that elements of bidegree \((1,\mathrm{even})\) anticommute with each other, those of bidegree \((1,\mathrm{odd})\) commute with each other, while an element of bidegree \((1,\mathrm{even})\) anticommutes with an element of bidegree \((1,\mathrm{odd})\).

Now let \((\mathfrak {g}, [-,-])\) be a finite-dimensional super Lie algebra. Then, its Chevalley–Eilenberg algebra \(\mathrm {CE}(\mathfrak {g})\) is the super Grassmann algebra \(\Lambda ^\bullet \mathfrak {g}^*\) equipped with the differential \(d_{\mathrm{CE}}\) defined as follows. On the generators \(\mathfrak {g}^*\), \(d_{\mathrm{CE}}\) acts as the linear dual of the Lie bracket:

$$\begin{aligned}{}[-,-]^* :\mathfrak {g}^*\rightarrow \Lambda ^2 \mathfrak {g}^*. \end{aligned}$$

The action of \(d_{\mathrm{CE}}\) on generators is then extended to all of \(\Lambda ^\bullet \mathfrak {g}^*\) as a derivation, bigraded of bidegree \((1,\mathrm{even})\). This makes \(\mathrm{CE}(\mathfrak {g})\) into a differential graded algebra. A calculation shows \(d_{\mathrm{CE}}^2 = 0\), so \(\mathrm{CE}(\mathfrak {g})\) is also a cochain complex.

For \(p \in {\mathbb {N}}\), we say that a \((p+2)\)-cocycle on \(\mathfrak {g}\) with coefficients in \({\mathbb {R}}\) is a \(d_{\mathrm {CE}}\)-closed element in \(\Lambda ^{p+2} \mathfrak {g}^*\). We say that cocycle is even if its degree in \({\mathbb Z}_2\) is even, and odd if it is odd. The super Lie algebra cohomology of \(\mathfrak {g}\) with coefficients in \({\mathbb {R}}\) is the cohomology of its Chevalley–Eilenberg algebra, regarded as a cochain complex:

$$\begin{aligned} H^\bullet (\mathfrak {g}, {\mathbb {R}}) := H^\bullet (\mathrm {CE}(\mathfrak {g})). \end{aligned}$$

The \({\mathbb Z}_2\)-grading on \(\mathrm{CE}(\mathfrak {g})\) makes \(H^p({\mathfrak {g}},{\mathbb {R}})\) into a supervector space for each p. We will be interested in its even part, \(H_{\mathrm {even}}^p({\mathfrak {g}},{\mathbb {R}})\)

Example 17

Let \(\mathfrak {g}\) be a finite-dimensional super Lie algebra, and let \(\omega \in \Lambda ^2 \mathfrak {g}^*\) be an even 2-cocycle as in Definition 16. Then, there is a new super Lie algebra \(\widehat{\mathfrak {g}}\) whose underlying supervector space is

$$\begin{aligned} \widehat{\mathfrak {g}} := \underset{\mathrm {even}}{\underbrace{\mathfrak {g}_{\mathrm {even}} \oplus {\mathbb {R}}}} \oplus \underset{\mathrm {odd}}{\underbrace{\mathfrak {g}_{\mathrm {odd}}}} \end{aligned}$$

and with super Lie bracket given by

$$\begin{aligned}{}[ (x_1,c_1), (x_2,c_2) ] \;=\; ([x_1,x_2], \omega (x_1, x_2)). \end{aligned}$$

We have a short exact sequence giving \({\widehat{\mathfrak {g}}}\) as a central extension of \(\mathfrak {g}\):

$$\begin{aligned} 0 \rightarrow {\mathbb R}\rightarrow {\widehat{\mathfrak {g}}} \rightarrow \mathfrak {g}\rightarrow 0 . \end{aligned}$$

In the paper, we will often write this short exact sequence as follows; in the style, an algebraic topologist might use to write down a fibration:

Just as for ordinary Lie algebras, this construction establishes a natural equivalence between central extensions of \(\mathfrak {g}\) by \({\mathbb {R}}\) (in even degree) and even super Lie algebra 2-cocycles on \(\mathfrak {g}\).

More generally, a central extension in even degree is by some vector space \(V \simeq {\mathbb {R}}^n\)

which is equivalently the result of extending by n even 2-cocycles, one after the other, in any order.

We will be interested not in the full super Lie algebra cohomology, but in the invariant cohomology with respect to some action:

Definition 18

For \(\mathfrak {g}\) a super Lie algebra (Definition 15), its automorphism group is the Lie subgroup

$$\begin{aligned} \mathrm {Aut}(\mathfrak {g}) \hookrightarrow \mathrm {GL}(\mathfrak {g}_{\mathrm {even}}) \times \mathrm {GL}(\mathfrak {g}_{\mathrm {odd}}) \end{aligned}$$

of the group of degree-preserving linear isomorphisms on the underlying vector space, consisting of those elements which are super Lie algebra isomorphisms.

Proposition 19

For \(\mathfrak {g}\) a super Lie algebra, the Lie algebra of its automorphism Lie group (Definition 18)

$$\begin{aligned} \mathfrak {aut}(\mathfrak {g}) \end{aligned}$$

is called the automorphism Lie algebra of \(\mathfrak {g}\). It is the Lie algebra of those linear maps \(\Delta :\mathfrak {g} \rightarrow \mathfrak {g}\) which preserve the degree and satisfy the derivation property:

$$\begin{aligned} \Delta [X,Y] = [\Delta X, Y] + [X, \Delta Y] \end{aligned}$$

for all \(X, Y\in \mathfrak {g}\). The Lie bracket on \(\mathfrak {aut}(\mathfrak {g})\) is the commutator:

$$\begin{aligned}{}[\Delta _1, \Delta _2] := \Delta _1 \Delta _2 - \Delta _2 \Delta _1. \end{aligned}$$

We caution the reader that, even though \(\mathfrak {g}\) is a super Lie algebra, its automorphism algebra \(\mathfrak {aut}(\mathfrak {g})\) is merely a Lie algebra. This is because we want elements of \(\mathfrak {aut}(\mathfrak {g})\) to preserve the degree on \(\mathfrak {g}\).

Example 20

The super Minkowski super Lie algebras \({\mathbb {R}}^{d-1,1\vert N}\) from Definition 22 all carry an automorphism action of the abelian Lie algebra \({\mathbb {R}}\) which is spanned by the scaling derivation that acts on vectors \(v \in {\mathbb {R}}^{d-1,1}\) by

$$\begin{aligned} v \mapsto 2v \end{aligned}$$

and on spinors \(\psi \in N\) by

$$\begin{aligned} \psi \mapsto \psi . \end{aligned}$$

Definition 21

Let \(\mathfrak {g}\) be a super Lie algebra (Definition 15). Clearly, every automorphism of \(\mathfrak {g}\) will induce an automorphism of the Chevalley–Eilenberg algebra \(\mathrm{CE}(\mathfrak {g})\) (Definition 16). Explicitly, this works as follows. Let \(\Delta \in \mathfrak {aut}(\mathfrak {g})\) be an infinitesimal automorphism (Proposition 19). The induced automorphism \(\Delta _\mathrm{CE}:\mathrm{CE}(\mathfrak {g}) \rightarrow \mathrm{CE}(\mathfrak {g})\) acts on the generators \(\mathfrak {g}^*\) of \(\mathrm{CE}(\mathfrak {g})\) as the linear dual \(\Delta ^*\):

$$\begin{aligned} \Delta _{\mathrm {CE}} :\mathfrak {g}^*\overset{\Delta ^*}{\longrightarrow } \mathfrak {g}^*. \end{aligned}$$

This is then extended to all of \(\mathrm{CE}(\mathfrak {g})\) as a derivation of bidegree \((0,\mathrm{even})\). The fact that \(\Delta _\mathrm{CE}\) commutes with \(d_\mathrm{CE}\) is equivalent to the fact that \(\Delta \) is a derivation of \(\mathfrak {g}\).

Now let \(\mathfrak {h} \hookrightarrow \mathfrak {aut}(\mathfrak {g})\) be a Lie subalgebra of its automorphism Lie algebra. The elements of \(\mathrm {CE}(\mathfrak {g})\) which are annihilated by \(\Delta _{\mathrm {CE}}\) for all \(\Delta \in \mathfrak {h}\) form a differential graded subalgebra of \(\mathrm{CE}(\mathfrak {g})\):

$$\begin{aligned} \mathrm {CE}(\mathfrak {g})^{\mathfrak {h}} \hookrightarrow \mathrm {CE}(\mathfrak {g}). \end{aligned}$$

We say an \(\mathfrak {h}\)-invariant \((p+2)\)-cocycle on \(\mathfrak {g}\) is an element in \(\mathrm {CE}(\mathfrak {g})^{\mathfrak {h}}\) which is \(d_{\mathrm {CE}}\)-closed and the \(\mathfrak {h}\)-invariant cohomology of \(\mathfrak {g}\) with coefficients in \({\mathbb {R}}\) is the cochain cohomology of this subcomplex:

$$\begin{aligned} H^\bullet (\mathfrak {g}, {\mathbb {R}})^{\mathfrak {h}} := H^\bullet (\mathrm {CE}(\mathfrak {g})^{\mathfrak {h}}). \end{aligned}$$

We define even and odd invariant cocycles as before. The vector space \(H^p(\mathfrak {g}, {\mathbb R})^\mathfrak {h}\) is \({\mathbb Z}_2\)-graded for each p, and our focus will be on its even part, \(H^p_\mathrm{even}(\mathfrak {g},{\mathbb R})^\mathfrak {h}\).

1.2 A.2 Super Minkowski spacetimes

We recall the definition of “super Minkowski super Lie algebras” (Definition 22) as well as their construction, on the one hand via Majorana or symplectic Majorana spinors (Example 25) and on the other hand via the four normed division algebras (Proposition 32). We freely use basic facts about spinors, as may be found in the book of Lawson and Michelsohn [40].

Definition 22

(super Minkowski Lie algebras) Let \(d \in {\mathbb {N}}\) (spacetime dimension) and let N be a real spinor representation of \(\mathrm{Spin}(d-1,1)\), the double cover of the connected Lorentz group \(\mathrm{SO}_0(d-1,1)\). Then, d-dimensional N-supersymmetric super Minkowski spacetime \({\mathbb {R}}^{d-1,1\vert N}\) is the super Lie algebra (Definition 15) whose underlying supervector space is

$$\begin{aligned} {\mathbb {R}}^{d-1,1\vert N} := \underset{\mathrm {even}}{\underbrace{{\mathbb {R}}^{d-1,1}}} \oplus \underset{\mathrm {odd}}{\underbrace{N}}. \end{aligned}$$

The Lie bracket is nonzero only on N and is a choice of symmetric, bilinear, \(\mathrm{Spin}(d-1,1)\)-equivariant map:

$$\begin{aligned}{}[-,-] :N \otimes N \longrightarrow {\mathbb {R}}^{d-1,1} . \end{aligned}$$

Such a map is always available in spacetime signature \((d-1,1)\) [30], though there may be more than one choice [52].

There is a canonical action of \(\mathrm {Spin}(d-1,1)\) on \({\mathbb {R}}^{d-1,1\vert N}\) by Lie algebra automorphisms, and the corresponding semidirect product Lie algebra is the super Poincaré super Lie algebra

$$\begin{aligned} \mathfrak {iso}({\mathbb {R}}^{d-1,1\vert N}) = {\mathbb {R}}^{d-1,1\vert N} \rtimes \mathfrak {so}(d-1,1). \end{aligned}$$

It is also called the supersymmetry algebra.

Remark 23

(number of supersymmetries) In the physics literature, the choice of real spin representation in Definition 22 is often referred to as the “number of supersymmetries.” While this is imprecise, it fits well with the convention of labeling irreducible representations by their dimension in boldface. For example, when \(d = 10\) there are two irreducible real spinor representations, both of real dimension 16, but of opposite chirality, and hence traditionally denoted \({\mathbf {16}}\) and \(\overline{{\mathbf {16}}}\). Hence, we may speak of \(N = {\mathbf {16}}\) supersymmetry (also called \(N = 1\), type I or heterotic) and \(N = {\mathbf {16}} \oplus \overline{{\mathbf {16}}}\) supersymmetry (also called \(N = (1,1)\) or type IIA) and \(N = {\mathbf {16}} \oplus {\mathbf {16}}\) supersymmetry (also called \(N = (2,0)\) or type IIB).

In Sect. 3, the generalization of the last of these cases plays a central role, where for any given real spin representation N we pass to the doubled supersymmetry \(N \oplus N\). Observe that the two canonical linear injections \(N \rightarrow N \oplus N\) into the direct sum induce two super Lie algebra homomorphisms

The following degenerate variation of super Minkowski spacetime will play a key role:

Definition 24

(Superpoint) A superpoint is the super Lie algebra

$$\begin{aligned} {\mathbb {R}}^{0|N}, \end{aligned}$$

which has zero Lie bracket, and whose underlying supervector space is concentrated in odd degree, where it is of dimension N.

We will use two different ways of constructing real spinor representation, and hence super Minkowski spacetimes: via “Majorana” or “symplectic Majorana” spinors (Example 25) and via real normed division algebras (Proposition 32).

Example 25

(Majorana representations) For \(d = 2\nu \) or \(2 \nu +1\), there exists a complex representation of the Clifford algebra \(\mathrm {Cl}({\mathbb {R}}^{d-1,1})\otimes {\mathbb {C}}\), hence of the spin group \(\mathrm {Spin}(d-1,1)\) on \({\mathbb {C}}^{2^\nu }\) such that

  1. 1.

    all skew-symmetrized products of \(p \ge 1\) Clifford elements \(\Gamma _{a_1 \cdots a_p}\) are traceless;

  2. 2.

    \(\Gamma _0^\dagger = \Gamma _0\) and \(\Gamma _i^\dagger = - \Gamma _i\), for \(1 \le i \le d - 1\).

This is the Dirac representation, a complex representation of \(\mathrm{Spin}(d-1,1)\). For \(d = 2\nu \), this is the direct sum of two subrepresentations on \({\mathbb {C}}^{2^{\nu }-1}\), the Weyl representations.

For \(d \in \{1,2,3,4,8,9,10,11\}\), there exists a real structure J on the complex Dirac representation, restricting to the Weyl representations for \(d = 2\) or \(d = 10\). This is a \(\mathrm {Spin}(d-1,1)\)-equivariant antilinear endomorphism \(J :S \rightarrow S\) which squares to the identity: \(J^2 = +1\). It carves out a real representation called the Majorana representation \(N := \mathrm {Eig}(J,+1)\), the eigenspace of J of eigenvalue +1, whose elements are called the Majorana spinors. In this case, the Dirac conjugation \(\psi \mapsto \psi ^\dagger \Gamma _0\) on elements \(\psi \in {\mathbb {C}}^{2^\nu }\) restricts to N and is called the Majorana conjugation. We write it as simply \(\psi \mapsto {\overline{\psi }}\). In terms of this matrix representation, then the spinor bilinear pairing that appears in Definition 22 is given by the following matrix product expression:

$$\begin{aligned}{}[\psi ,\phi ] \;=\; \left( {\overline{\psi }} \Gamma ^a \phi \right) _{a = 0}^{d-1}. \end{aligned}$$

Similarly, for \(d \in \{5,6,7\}\) there exists a quaternionic structure on the Dirac representation. This is a \(\mathrm{Spin}(d-1,1)\)-equivariant antilinear endomorphism \({\tilde{J}}\) which squares to minus the identity, \({\tilde{J}}^2 = -1\). It follows that

$$\begin{aligned} J := \left( \begin{array}{c@{\quad }c} 0 &{} -\,{\tilde{J}} \\ {\tilde{J}} &{} 0 \end{array} \right) \end{aligned}$$

is a real structure on the direct sum of the Dirac representation with itself. Hence as before \(N := \mathrm {Eig}(J,+1)\) is a real subrepresentation, called the symplectic Majorana representation. The spinor-to-vector bilinear pairing for symplectic Majorana spinors is similar to the case of Majorana spinors.

Definition 26

(Cayley–Dickson double [2, Section 2.2]) Let \({\mathbb {K}}\) be a real \(*\)-algebra. This is a real, not necessarily associative algebra \({\mathbb K}\) equipped with a conjugation \(\overline{(-)} :{\mathbb {K}} \rightarrow {\mathbb {K}}\), satisfying:

$$\begin{aligned} \overline{a + b} = {\overline{a}} + {\overline{b}}, \quad {\overline{ab}} = {\overline{b}} \, {\overline{a}}, \quad \overline{{\overline{a}}} = a , \end{aligned}$$

for any \(a, b \in {\mathbb K}\). Then, the Cayley–Dickson double \({\mathbb {K}}_{\mathrm{dbl}}\) of \({\mathbb {K}}\) is the real \(*\)-algebra obtained from \({\mathbb {K}}\) by adjoining one element \(\ell \) such that \(\ell ^2 = -1\) and such that the following relations hold, for all \(a, b \in {\mathbb {K}}\):

$$\begin{aligned} a(\ell b) = \ell ({\overline{a}} b),\;\;\;\; (a \ell ) b = (a {\overline{b}}) \ell ,\;\;\;\; (\ell a)(b \ell ) = -\, \overline{(a b)}. \end{aligned}$$

Finally, the conjugation \(\overline{(-)}\) on \({\mathbb {K}}_{\mathrm {dbl}}\) acts on elements of \({\mathbb K}\) by the conjugation on \({\mathbb K}\), and sends the new generator \(\ell \) to \(-\ell \).

Example 27

Consider \({\mathbb {R}}\) the real numbers regarded as a \(*\)-algebra with trivial conjugation \({\overline{a}} = a\). Then, its Cayley–Dickson double (Definition 26) is the complex numbers \({\mathbb {C}}\) with the usual conjugation, the Cayley–Dickson double of \({\mathbb C}\) is the quaternions \({\mathbb {H}}\), and the Cayley–Dickson double of \({\mathbb H}\) is the octonions \({\mathbb {O}}\).

By a classical result of Hurwitz, these four algebras are the only normed division algebras over the real numbers, as reviewed by Baez [2].

In the next proposition and elsewhere in the text, we will use \(n \times n\) matrices over \({\mathbb K}\) to describe real linear operators on \({\mathbb K}^n\). We will write \({\mathbb K}[n]\) for the set of all \(n \times n\) matrices with entries in \({\mathbb K}\). For any such matrix, there are two natural ways for it to induce a linear operator, one using left multiplication in \({\mathbb K}\) and the other right multiplication.

Definition 28

(Matrices over\({\mathbb K}\) as linear operators) Let \({\mathbb K}\) be a normed division algebra. Any element of \(a \in {\mathbb K}\) induces a linear endomorphism on \({\mathbb K}\) by left or right multiplication, which we will denote by \(a_L\) or \(a_R\), respectively:

$$\begin{aligned} \begin{array}{rcl} a_L :{\mathbb K}&{} \rightarrow &{} {\mathbb K}\\ x &{} \mapsto &{} ax \end{array} , \quad \begin{array}{rcl} a_R :{\mathbb K}&{} \rightarrow &{} {\mathbb K}\\ x &{} \mapsto &{} xa . \end{array} \end{aligned}$$

More generally, any \(n \times n\) matrix \(A \in {\mathbb K}[n]\) induces a linear endomorphism on \({\mathbb K}^n\) via either left multiplication or right multiplication:

$$\begin{aligned} \begin{array}{rcl} A_L :{\mathbb K}^n &{} \rightarrow &{} {\mathbb K}^n \\ x &{} \mapsto &{} \sum a_{ij} x_j \end{array} , \quad \begin{array}{rcl} A_R :{\mathbb K}^n &{} \rightarrow &{} {\mathbb K}^n \\ x &{} \mapsto &{} \sum x_j a_{ij} \end{array} \end{aligned}$$

where we are using the subscript \(x_j\) to denote the jth coordinate of \(x \in {\mathbb K}^n\), and \(A = (a_{ij})\). In other words, \(A_L\) and \(A_R\) are the linear maps on \({\mathbb K}^n\) with components \(\left( (a_{ij})_L \right) \) and \(\left( (a_{ij})_R \right) \), respectively. We say that \(A_L\) is the left action of A and \(A_R\) is the right action of A. We caution that because \({\mathbb K}\) is non-associative, \((AB)_L \ne A_L B_L\) in general, and because \({\mathbb K}\) is non-associative and non-commutative, \((AB)_R \ne A_R B_R\) in general.

Remark 29

The left action of a matrix A by \(A_L\) is just the usual matrix multiplication, so we will sometimes write:

$$\begin{aligned} A_L x = Ax . \end{aligned}$$

The utility of defining the linear transformation \(A_L\) is that the composition of linear transformations is associative, so we do not need to worry about the non-associativity of \({\mathbb K}\) when we compose them. For example:

$$\begin{aligned} A_L B_L C_L x = A(B(Cx)) . \end{aligned}$$

Since \({\mathbb K}\) comes with a conjugation, we can define the conjugate of any matrix in \({\mathbb K}[n]\) by taking the conjugate of each entry, and the conjugate of any element of \({\mathbb K}^n\) by taking the conjugate of each coordinate. It is then an elementary calculation to show that the action of a matrix A by \(A_L\) and \(A_R\) is related by conjugation:

Proposition 30

Let \(A \in {\mathbb K}[n]\) be an \(n \times n\) matrix over the normed division algebra \({\mathbb K}\) (as defined in Example 27). Then,

$$\begin{aligned} \overline{A_L x} = ({\overline{A}})_R \, {\overline{x}} \text{ and } \overline{A_R x} = ({\overline{A}})_L \, {\overline{x}} \end{aligned}$$

for all \(x \in {\mathbb K}^n\).

The next definition is straightforward, but is central to realizing spin representations via normed division algebras.

Definition 31

[49] For \(A \in \mathfrak {h}_2({\mathbb K})\), a hermitian matrix with coefficients in one of the four real normed division algebras from Example 27. Then, its trace reversal is

$$\begin{aligned} {\widetilde{A}} := A - \mathrm {tr}(A)\cdot {\mathbf {1}}. \end{aligned}$$

Proposition 32

[4] Let \({\mathbb K}\in \{{\mathbb {R}}, {\mathbb {C}}, {\mathbb {H}}, {\mathbb {O}}\}\) be one of the normed division algebras as in Example 27. Write \(\mathfrak {h}_2({\mathbb K})\) for the real vector space of \(2 \times 2\) hermitian matrices with coefficients in \({\mathbb K}\), and k for the dimension of \({\mathbb K}\).

Then:

  1. 1.

    There is an isomorphism of inner product spaces (“forming Pauli matrices over \({\mathbb {K}}\)”)

    $$\begin{aligned} ({\mathbb {R}}^{k+1,1}, \eta ) {\mathop {\longrightarrow }\limits ^{\simeq }} \left( \mathfrak {h}_2({\mathbb K}), -\mathrm {det} \right) \end{aligned}$$

    identifying \({\mathbb {R}}^{k+1,1}\) equipped with its Minkowski inner product

    $$\begin{aligned} \eta (A,B) := -A^0 B^0 + A^1 B^1 + \cdots + A^{k+1} B^{k+1}, \text{ for } A, B \in {\mathbb R}^{k+1,1} \end{aligned}$$

    with the space of hermitian matrices equipped with the negative of the determinant operation.

  2. 2.

    Let \(N_+\) and \(N_-\) both denote the vector space \({\mathbb K}^2\). Then, \(N_+ \oplus N_-\) is a module of the Clifford algebra \({\mathcal {C}\ell }(k+1,1)\), with the action of a vector in \(A \in {\mathbb R}^{k+1,1}\) given by

    $$\begin{aligned} \Gamma (A) (\psi , \phi ) = (\tilde{A}_L \phi , A_L \psi ) \end{aligned}$$

    for any element \((\psi , \phi ) \in N_+ \oplus N_-\), where we are using the identification of vectors with \(2 \times 2\) hermitian matrices. Here, \(\widetilde{(-)}\) is the trace reversal operation from Definition 31, and \((-)_L\) denotes the linear map given by left multiplication as in Definition 28

  3. 3.

    Realizing the spin group \(\mathrm{Spin}(k+1,1)\) inside the Clifford algebra \({\mathcal {C}\ell }(k+1,1)\) by the standard construction, this induces irreducible representations \(\rho _\pm \) of \(\mathrm{Spin}(k+1,1)\) on \(N_\pm \). Explicitly, recall that \(\mathrm{Spin}(k+1,1)\) is the subgroup of the Clifford algebra generated by products of pairs of unit vectors of the same sign:

    $$\begin{aligned} \mathrm{Spin}(k+1, 1) = \langle AB \in {\mathcal {C}\ell }(k+1,1) \, : \, A, B \in {\mathbb R}^{k+1,1}, \, \eta (A,A) = \eta (B,B) = \pm 1 \rangle . \end{aligned}$$

    Then, restricting the Clifford action to these elements, a generator AB of \(\mathrm{Spin}(k+1,1)\) acts as

    $$\begin{aligned} \rho _+(AB) = \tilde{A}_L B_L \text{ on } N_+ \end{aligned}$$

    and as

    $$\begin{aligned} \rho _-(AB) = A_L \tilde{B}_L \text{ on } N_-, \end{aligned}$$

    where again \(\widetilde{(-)}\) is the trace reversal operation from Definition 31 and where \((-)_L\) denotes the linear map given by left multiplication as in Definition 28. For \({\mathbb {K}} \in \{{\mathbb {R}}, {\mathbb {C}}\}\), then these two representations are in fact isomorphic and are the Majorana representation of \(\mathrm {Spin}(2,1)\) and \(\mathrm{Spin}(3,1)\), respectively, while for \({\mathbb {K}} \in \{ {\mathbb {H}}, {\mathbb {O}} \}\) they are the two non-isomorphic symplectic Majorana representations of \(\mathrm {Spin}(5,1)\) and Majorana–Weyl representations of \(\mathrm {Spin}(9,1)\), respectively.

  4. 4.

    Under the above identifications, the symmetric bilinear \(\mathrm{Spin}(k+1,1)\)-equivariant spinor-to-vector pairings are given by

    $$\begin{aligned} \begin{array}{rcl} [-,-] :N_+ \otimes N_+ &{} \rightarrow &{} {\mathbb R}^{k+1,1} \\ \psi \otimes \phi &{} \mapsto &{} \tfrac{1}{2}\left( \widetilde{ \psi \phi ^\dagger + \phi \psi ^\dagger } \right) \, \end{array} \end{aligned}$$

    and

    $$\begin{aligned} \begin{array}{rcl} [-,-] :N_- \otimes N_- &{} \rightarrow &{} {\mathbb R}^{k+1,1} \\ \psi \otimes \phi &{} \mapsto &{} \tfrac{1}{2}\left( \psi \phi ^\dagger + \phi \psi ^\dagger \right) \, \end{array} \end{aligned}$$
  5. 5.

    There is a bilinear symmetric, non-degenerate and \(\mathrm {Spin}(k+1,1)\)-invariant spinor-to-scalar pairing given by

    $$\begin{aligned} \begin{array}{rcl} \langle -, -\rangle :N_\pm \otimes N_{\mp } &{} \rightarrow &{} {\mathbb R}\\ \psi \otimes \phi &{} \mapsto &{} \mathrm {Re}(\psi ^\dagger \phi ). \end{array} \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huerta, J., Schreiber, U. M-theory from the superpoint. Lett Math Phys 108, 2695–2727 (2018). https://doi.org/10.1007/s11005-018-1110-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-018-1110-z

Keywords

Mathematics Subject Classification

Navigation