Skip to main content
Log in

PP-waves with torsion: a metric-affine model for the massless neutrino

  • Research Article
  • Published:
General Relativity and Gravitation Aims and scope Submit manuscript

Abstract

In this paper we deal with quadratic metric-affine gravity, which we briefly introduce, explain and give historical and physical reasons for using this particular theory of gravity. We then introduce a generalisation of well known spacetimes, namely pp-waves. A classical pp-wave is a 4-dimensional Lorentzian spacetime which admits a nonvanishing parallel spinor field; here the connection is assumed to be Levi-Civita. This definition was generalised in our previous work to metric compatible spacetimes with torsion and used to construct new explicit vacuum solutions of quadratic metric-affine gravity, namely generalised pp-waves of parallel Ricci curvature. The physical interpretation of these solutions we propose in this article is that they represent a conformally invariant metric-affine model for a massless elementary particle. We give a comparison with the classical model describing the interaction of gravitational and massless neutrino fields, namely Einstein–Weyl theory and construct pp-wave type solutions of this theory. We point out that generalised pp-waves of parallel Ricci curvature are very similar to pp-wave type solutions of the Einstein–Weyl model and therefore propose that our generalised pp-waves of parallel Ricci curvature represent a metric-affine model for the massless neutrino.

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. See “Appendix A” for our general spinor formalism and Sect. 3.3 for spinor formalism for pp-waves.

  2. See in particular [37] as well as [51, 53, 6871].

  3. Only the \(T^{(1)}\), or ‘tensor torsion’ irreducible piece of torsion is non-zero, see Eqs. (9), (10).

  4. A classical model describing the interaction of gravitational and electromagnetic fields.

  5. A classical model describing the interaction of gravitational and massless neutrino fields.

References

  1. Adamowicz, W.: Plane waves in gauge theories of gravitation. Gen. Relativ. Gravit. 12, 677–691 (1980)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  2. Alekseevsky, D.V.: Holonomy groups and recurrent tensor fields in Lorentzian spaces. In: Stanjukovich, K.P. (ed.) Problems of the Theory of Gravitation and Elementary Particles issue 5, pp. 5–17. Atomizdat, Moscow (in Russian) (1974)

  3. Audretsch, J.: Asymptotic behaviour of neutrino fields in curved space-time. Commun. Math. Phys. 21, 303–313 (1971)

    Article  MathSciNet  ADS  Google Scholar 

  4. Audretsch, J., Graf, W.: Neutrino radiation in gravitational fields. Commun. Math. Phys. 19, 315–326 (1970)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  5. Babourova, O.V., Frolov, B.N., Klimova, E.A.: Plane torsion waves in quadratic gravitational theories in RiemannCartan space. Class. Quantum Grav. 16, 1149–1162 (1999). gr-qc/9805005

  6. Berestetskii, V.B., Lifshitz, E.M., Pitaevskii, L.P.: Quantum Electrodynamics (Course of Theoretical Physics vol. 4) 2nd edn. Pergamon Press, Oxford (1982)

  7. Blagojevic, M.: Gravitation and Gauge Symmetries. Institute of Physics Publishing, Bristol (2002)

    Book  MATH  Google Scholar 

  8. Blagojević, M., Hehl, F.W.: Gauge Theories of Gravitation. A Reader with Commentaries. Imperial College Press, London (2013)

    Book  MATH  Google Scholar 

  9. Brill, D.R., Wheeler, J.A.: Interaction of neutrinos and gravitational fields. Rev. Mod. Phys. 29, 465–479 (1957)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  10. Brinkmann, M.W.: On Riemann spaces conformal to Euclidean space. Proc. Natl. Acad. Sci. USA 9, 1–3 (1923)

    Article  ADS  Google Scholar 

  11. Bryant, R.L.: Pseudo-Riemannian metrics with parallel spinor fields and vanishing Ricci tensor. Global Analysis and Harmonic Analysis (Marseille-Luminy, 1999) Sémin. Congr. 4 (Paris: Soc. Math. France), 53–94 (2000). math/0004073

  12. Buchbinder, I.L., Kuzenko, S.M.: Ideas and Methods of Supersymmetry and Supergravity. Institute of Physics Publishing, Bristol (1998)

    MATH  Google Scholar 

  13. Buchdahl, H.A.: Math. Rev. 20, 1238 (1959)

    Google Scholar 

  14. Collinson, C.D., Morris, P.B.: Space-time admitting neutrino fields with zero energy and momentum. J. Phys. A6, 915–916 (1972)

    ADS  Google Scholar 

  15. Cotton, É.: Sur les varits a trois dimensions. Annales de la Facult des Sciences de Toulouse II 14(14), 385–438 (1899)

    Article  MathSciNet  Google Scholar 

  16. Davis, T.M., Ray, J.R.: Ghost neutrinos in general relativity. Phys. Rev. D 9, 331–333 (1974)

    Article  ADS  Google Scholar 

  17. Eddington, A.S.: The Mathematical Theory of Relativity, 2nd edn. The University Press, Cambridge (1952)

    Google Scholar 

  18. Esser, W.: Exact Solutions of the Metric-Affine Gauge Theory of Gravity. Diploma Thesis, University of Cologne, Cologne (1996)

  19. Fairchild Jr, E.E.: Gauge theory of gravitation. Phys. Rev. D 14, 384–391 (1976)

    Article  MathSciNet  ADS  Google Scholar 

  20. Fairchild Jr, E.E.: Erratum: gauge theory of gravitation. Phys. Rev. D 14, 2833 (1976)

    Article  MathSciNet  ADS  Google Scholar 

  21. García, A., Macías, A., Puetzfeld, D., Socorro, J.: Plane fronted waves in metric affine gravity. Phys. Rev. D 62, 044021 (2000). gr-qc/0005038

  22. García, A., Hehl, F.W., Heinicke, C., Macías, A.: The cotton tensor in Riemannian spacetimes. Class. Quantum Grav. 21, 10991118 (2004). gr-qc/0309008

  23. Griffiths, J.B.: Colliding Plane Waves in General Relativity. Oxford University Press, Oxford (1991)

    MATH  Google Scholar 

  24. Griffiths, J.B., Newing, R.A.: The two-component neutrino field in general relativity. J. Phys. A 3, 136–149 (1970)

    Article  MathSciNet  ADS  Google Scholar 

  25. Griffiths, J.B., Newing, R.A.: Tetrad equations for the two-component neutrino field in general relativity. J. Phys. A 3, 269–273 (1970)

    Article  MathSciNet  ADS  Google Scholar 

  26. Griffiths, J.B.: Some physical properties of neutrino-gravitational fields. Int. J. Theor. Phys. 5, 141–150 (1972)

    Article  Google Scholar 

  27. Griffiths, J.B.: Gravitational radiation and neutrinos. Commun. Math. Phys. 28, 295–299 (1972)

    Article  ADS  Google Scholar 

  28. Griffiths, J.B.: Ghost neutrinos in Einstein–Cartan theory. Phys. Lett. A 75, 441–442 (1980)

    Article  ADS  Google Scholar 

  29. Griffiths, J.B.: Neutrino fields in Einstein–Cartan theory. Gen. Relativ. Gravit. 13, 227–237 (1981)

    Article  MATH  ADS  Google Scholar 

  30. Hehl, F.W.: Spin und Torsion in der Allgemeinen Relativitätstheorie oder die Riemann-Cartansche Geometrie der Welt. Technischen Universität Clausthal, Habilitationsschrift (1970)

  31. Hehl, F.W.: Spin and torsion in general relativity I: foundations. Gen. Relativ. Gravit. 4, 333–349 (1973)

    Article  MathSciNet  ADS  Google Scholar 

  32. Hehl, F.W.: Spin and torsion in general relativity II: geometry and field equations. Gen. Relativ. Gravit. 5, 491–516 (1974)

    Article  MathSciNet  ADS  Google Scholar 

  33. Hehl, F.W., McCrea, J.D., Mielke, E.W., Ne’eman, Y.: Metric-affine gauge theory of gravity: field equations, Noether identities, world spinors, and breaking of dilation invariance. Phys. Rep. 258, 1–171 (1995). gr-qc/9402012

  34. Hehl, F.W., Macías, A.: Metric-affine gauge theory of gravity II. Exact solutions. Int. J. Mod. Phys. D 8, 399–416 (1999). gr-qc/9902076

  35. Hehl, F.W., von der Heyde, P., Kerlick, G.D., Nester, J.M.: General relativity with spin and torsion: foundations and prospects. Rev. Mod. Phys. 48, 393–416 (1976)

    Article  ADS  Google Scholar 

  36. Higgs, P.W.: Quadratic Lagrangians and general relativity. Nuovo Cimento 11, 816–820 (1959)

    Article  MathSciNet  MATH  Google Scholar 

  37. King, A.D., Vassiliev, D.: Torsion waves in metric-affine field theory. Class. Quantum Grav. 18, 2317–2329 (2001). gr-qc/0012046

  38. Kramer, D., Stephani, H., Herlt, E., MacCallum, M.: Exact Solutions of Einstein’s Field Equations. Cambridge University Press, Cambridge (1980)

    MATH  Google Scholar 

  39. Kröner, E.: Continuum theory of defects. In: Balian R., et al. (eds.) Physics of Defects, Les Houches, Session XXXV, 1980. North-Holland, Amsterdam (1980)

  40. Kröner, E.: The continuized crystal—a bridge between micro- and macromechanics. Gesellschaft angewandte Mathematik und Mechanik Jahrestagung Goettingen West Germany Zeitschrift Flugwissenschaften, vol. 66 (1986)

  41. Kuchowicz, C., Żebrowski, J.: The presence of torsion enables a metric to allow a gravitational field. Phys. Lett. A 67, 16–18 (1978)

    Article  MathSciNet  ADS  Google Scholar 

  42. Lanczos, C.: A remarkable property of the Riemann–Christoffel tensor in four dimensions. Ann. Math. 39, 842–850 (1938)

    Article  MathSciNet  Google Scholar 

  43. Lanczos, C.: Lagrangian multiplier and Riemannian spaces. Rev. Mod. Phys. 21, 497–502 (1949)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  44. Lanczos, C.: Electricity and general relativity. Rev. Mod. Phys. 29, 337–350 (1957)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  45. Landau, L.D., Lifshitz, E.M.: The Classical Theory of Fields (Course of Theoretical Physics vol. 2) 4nd edn. Pergamon Press, Oxford (1975)

  46. Mielke, E.W.: On pseudoparticle solutions in Yang’s theory of gravity. Gen. Relativ. Gravit. 13, 175–187 (1981)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  47. Nakahara, M.: Geometry, Topology and Physics. Institute of Physics Publishing, Bristol (1998)

    Google Scholar 

  48. Obukhov, Y.N.: Generalized plane fronted gravitational waves in any dimension. Phys. Rev. D 69, 024013 (2004). gr-qc/0310121

  49. Obukhov, Y.N.: Plane waves in metric-affine gravity. Phys. Rev. D 73, 024025 (2006). gr-qc/0601074

  50. Olesen, P.: A relation between the Einstein and the Yang–Mills field equations. Phys. Lett. B 71, 189–190 (1977)

    Article  MathSciNet  ADS  Google Scholar 

  51. Pasic, V.: New vacuum solutions for quadratic metric-affine gravity—a metric affine model for the massless neutrino? Math. Balk. New Ser. 24, Fasc 3–4, 329 (2010)

  52. Pasic, V., Barakovic, E., Okicic, N.: A new representation of the field equations of quadratic metric-affine gravity. Adv. Math. Sci. J. 3(1), 33–46 (2014)

  53. Pasic, V., Vassiliev, D.: PP-waves with torsion and metric-affine gravity. Class. Quantum Grav. 22, 3961–3975 (2005). gr-qc/0505157

  54. Pauli, W.: Zur Theorie der Gravitation und der Elektrizität von Hermann Weyl. Physik. Zaitschr. 20, 457–467 (1919)

    MATH  Google Scholar 

  55. Pavelle, R.: Unphysical solutions of Yang’s gravitational-field equations. Phys. Rev. Lett. 34, 1114 (1975)

    Article  ADS  Google Scholar 

  56. Penrose, O., Rindler, W.: Propagating modes in gauge field theories of gravity, vol. 2. Cambridge University Press, Oxford (1984, 1986)

  57. Peres, A.: Some gravitational waves. Phys. Rev. Lett. 3, 571–572 (1959)

    Article  MATH  ADS  Google Scholar 

  58. Peres, A.: PP—WAVES preprint (reprinting of [57]) (2002). hep-th/0205040

  59. Pirani, F.A.E.: Introduction to Gravitational Radiation Theory. Lectures on General Relativity. Prentice-Hall, Inc. Englewood Cliffs, New Jersey (1964)

  60. Singh, P.: On axial vector torsion in vacuum quadratic Poincaré gauge field theory. Phys. Lett. A 145, 7–10 (1990)

    Article  MathSciNet  ADS  Google Scholar 

  61. Singh, P.: On null tratorial torsion in vacuum quadratic Poincaré gauge field theory. Class. Quantum Grav. 7, 2125–2130 (1990)

    Article  MATH  ADS  Google Scholar 

  62. Singh, P., Griffiths, J.B.: On neutrino fields in Einstein–Cartan theory. Phys. Lett. A 132, 88–90 (1988)

    Article  MathSciNet  ADS  Google Scholar 

  63. Singh, P., Griffiths, J.B.: A new class of exact solutions of the vacuum quadratic Poincaré gauge field theory. Gen. Relativ. Gravit. 22, 947–956 (1990)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  64. Stephenson, G.: Quadratic Lagrangians and general relativity. Nuovo Cimento 9, 263–269 (1958)

    Article  MathSciNet  MATH  Google Scholar 

  65. Streater, R.F., Wightman, A.S.: PCT, spin and statistics, and all that. In: Princeton Landmarks in Physics, Princeton University Press, Princeton. ISBN:0-691-07062-8, corrected third printing of the 1978 edition (2000)

  66. Thompson, A.H.: Yang’s gravitational field equations. Phys. Rev. Lett. 34, 507–508 (1975)

    Article  ADS  Google Scholar 

  67. Thompson, A.H.: Geometrically degenerate solutions of the Kilmister–Yang equations. Phys. Rev. Lett. 35, 320–322 (1975)

    Article  MathSciNet  ADS  Google Scholar 

  68. Vassiliev, D.: Pseudoinstantons in metric-affine field theory. Gen. Relativ. Gravit. 34, 1239–1265 (2002). gr-qc/0108028

  69. Vassiliev, D.: Pseudoinstantons in metric-affine field theory. In: Brambilla, N., Prosperi, G.M. (eds.) Quark Confinement and the Hadron Spectrum V, pp. 273–275. World Scientific, Singapore (2003)

    Chapter  Google Scholar 

  70. Vassiliev, D.: Quadratic non-Riemannian gravity. J. Nonlinear Math. Phys. 11(Supplement), 204–216 (2004)

    Article  MathSciNet  ADS  Google Scholar 

  71. Vassiliev, D.: Quadratic metric-affine gravity. Ann. Phys. (Lpz.) 14, 231–252 (2005). gr-qc/0304028

  72. Weyl, H.: Eine neue Erweiterung der Relativitätstheorie. Ann. Phys. (Lpz.) 59, 101–133 (1919)

    Article  MATH  ADS  Google Scholar 

  73. Wilczek, F.: Geometry and interaction of instantons. In: Stump, D.R., Weingarten, D.H. (eds.) Quark Confinement and Field theory, pp. 211–219. Wiley-Interscience, New York (1977)

    Google Scholar 

  74. Yang, C.N.: Integral formalism for gauge fields. Phys. Rev. Lett. 33, 445–447 (1974)

    Article  MathSciNet  ADS  Google Scholar 

Download references

Acknowledgments

The authors are very grateful to D Vassiliev, J B Griffiths and F W Hehl for helpful advice and to the Ministry of Education and Science of the Federation of Bosnia and Herzegovina, which supported our research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vedad Pasic.

Appendices

Appendix A: Spinor formalism

This appendix provides the spinor formalism used throughout our work. Unless otherwise stated, we work in a general metric compatible spacetime with torsion. When introducing our spinor formalism, we were faced with the problem that there doesn’t seem to exist a uniform convention in the existing literature on how to treat spinors. Optimally, we would have wanted to achieve the following:

  1. (i)

    consecutive raising and lowering of a spinor index does not change the sign of a rank 1 spinor;

  2. (ii)

    the metric spinor \(\epsilon ^{ab}\) is the raised version of \(\epsilon _{ab}\) and vice versa;

  3. (iii)

    the spinor inner product is invariant under raising and lowering of indices, i.e. \(\xi _a \eta ^a = \xi ^a \eta _a \).

Unfortunately, it becomes clear that it is not possible to satisfy all three desired properties, as shown in [59]. This inconsistency is related to the well known fact (see for example Section 19 in [6] or Section 3–5 in [65]), that a spinor does not have a particular sign—for example, a spatial rotation of the coordinate system by \(2\pi \) leads to a change of sign. Also see [56] for more helpful insight about the problem of choice of the spinor formalism, as well as e.g. [6, 7, 12, 29, 59, 63] for insight to various approaches to spinor formalism.

We decided to define our spinor formalism in the following way. We define the ‘metric spinor’ as

$$\begin{aligned} \epsilon _{ab}=\epsilon _{\dot{a}\dot{b}}= \epsilon ^{ab}=\epsilon ^{\dot{a}\dot{b}}= \left( \begin{array}{cc} 0&{}\quad 1\\ -1&{}\quad 0 \end{array} \right) \end{aligned}$$
(57)

with the first index enumerating rows and the second enumerating columns. We raise and lower spinor indices according to the formulae

$$\begin{aligned} \xi ^a=\epsilon ^{ab}\xi _b, \qquad \xi _a=\epsilon _{ab}\xi ^b, \qquad \eta ^{\dot{a}}=\epsilon ^{\dot{a}\dot{b}}\eta _{\dot{b}}, \qquad \eta _{\dot{a}}=\epsilon _{\dot{a}\dot{b}}\eta ^{\dot{b}}. \end{aligned}$$
(58)

Our definition (57), (58) has the following advantages:

  • The spinor inner product is invariant under the operation of raising and lowering of indices, i.e. \((\epsilon _{ac}\xi ^c)(\epsilon ^{ad}\eta _d)=\xi ^a\eta _a\).

  • The ‘contravariant’ and ‘covariant’ metric spinors are ‘raised’ and ‘lowered’ versions of each other, i.e. \(\epsilon ^{ab}=\epsilon ^{ac}\epsilon _{cd}\epsilon ^{bd}\) and \(\epsilon _{ab}=\epsilon _{ac}\epsilon ^{cd}\epsilon _{bd}\).

The disadvantage of our definition (57), (58) is that the consecutive raising and lowering of a single spinor index leads to a change of sign, i.e. \(\epsilon _{ab}\epsilon ^{bc}\xi _c=-\xi _a\). In formulae where the sign is important we will be careful in specifying our choice of sign; see, for example, (59), (63). We in a sense intentionally ‘sacrificed’ this property in order to guarantee that the other two properties, which in our view have greater physical significance, are satisfied.

Let \(\mathfrak {v}\) be the real vector space of Hermitian \(2\times 2\) matrices \(\sigma _{a\dot{b}}\). Pauli matrices \(\sigma ^\alpha {}_{a\dot{b}}\), \(\alpha =0,1,2,3\), are a basis in \(\mathfrak {v}\) satisfying \(\sigma ^\alpha {}_{a\dot{b}}\sigma ^{\beta c\dot{b}} +\sigma ^\beta {}_{a\dot{b}}\sigma ^{\alpha c\dot{b}} =2g^{\alpha \beta }\delta _a{}^c\) where

$$\begin{aligned} \sigma ^{\alpha a\dot{b}}:= \epsilon ^{ac}\sigma ^\alpha {}_{c\dot{d}}\epsilon ^{\dot{b}\dot{d}}. \end{aligned}$$
(59)

At every point of the manifold \(M\) Pauli matrices are defined uniquely up to a Lorentz transformation. Define

$$\begin{aligned} \sigma _{\alpha \beta ac}:=\frac{1}{2} \bigl ( \sigma _{\alpha a\dot{b}}\epsilon ^{\dot{b}\dot{d}}\sigma _{\beta c\dot{d}} - \sigma _{\beta a\dot{b}}\epsilon ^{\dot{b}\dot{d}}\sigma _{\alpha c\dot{d}} \bigr ). \end{aligned}$$
(60)

These ‘second order Pauli matrices’ are polarized, i.e.

$$\begin{aligned} *\sigma =\pm i\sigma \end{aligned}$$
(61)

depending on the orientation of ‘basic’ Pauli matrices \(\sigma ^\alpha {}_{a\dot{b}}\), \(\alpha =0,1,2,3\).

We define the covariant derivatives of spinor fields as

$$\begin{aligned} \nabla _\mu \xi ^a&= \partial _\mu \xi ^a+{\varGamma }^a{}_{\mu b}\xi ^b, \qquad \nabla _\mu \xi _a=\partial _\mu \xi _a-{\varGamma }^b{}_{\mu a}\xi _b,\\ \nabla _\mu \eta ^{\dot{a}}&= \partial _\mu \eta ^{\dot{a}} +\bar{\varGamma }^{\dot{a}}{}_{\mu \dot{b}}\eta ^{\dot{b}}, \qquad \nabla _\mu \eta _{\dot{a}}=\partial _\mu \eta _{\dot{a}} -\bar{\varGamma }^{\dot{b}}{}_{\mu \dot{a}}\eta _{\dot{b}}, \end{aligned}$$

where \(\bar{\varGamma }^{\dot{a}}{}_{\mu \dot{b}}=\overline{{\varGamma }^a{}_{\mu b}}\). The explicit formula for the spinor connection coefficients \({\varGamma }^a{}_{\mu b}\) can be derived from the following two conditions:

$$\begin{aligned} \nabla _\mu \epsilon ^{ab}=0, \quad \nabla _\mu j^\alpha =\sigma ^\alpha {}_{a\dot{b}}\nabla _\mu \zeta ^{a\dot{b}}, \end{aligned}$$
(62)

where \(\zeta \) is an arbitrary rank 2 mixed spinor field and \(j^\alpha :=\sigma ^\alpha {}_{a\dot{b}}\zeta ^{a\dot{b}}\) is the corresponding vector field (current). Conditions (62) give a system of linear algebraic equations for \(\mathrm{Re}\,{\varGamma }^a{}_{\mu b}\), \(\mathrm{Im}\,{\varGamma }^a{}_{\mu b}\) the unique solution of which is

$$\begin{aligned} {\varGamma }^a{}_{\mu b}=\frac{1}{4} \sigma _\alpha {}^{a\dot{c}} \left( \partial _\mu \sigma ^\alpha {}_{b\dot{c}} +{\varGamma }^\alpha {}_{\mu \beta }\sigma ^{\beta }{}_{b\dot{c}} \right) . \end{aligned}$$
(63)

See section 3 of [24] for more background on covariant differentiation of spinors.

Appendix B: Massless Dirac equation

The generally accepted point of view [2932, 35] is that a massless neutrino field is a metric compatible spacetime with or without torsion, described by the action

$$\begin{aligned} S_\mathrm{neutrino}:=2i\int \Bigl ( \xi ^a\,\sigma ^\mu {}_{a\dot{b}}\,(\nabla _\mu \bar{\xi }^{\dot{b}}) \ -\ (\nabla _\mu \xi ^a)\,\sigma ^\mu {}_{a\dot{b}}\,\bar{\xi }^{\dot{b}} \Bigr ), \end{aligned}$$
(64)

see formula (11) of [29]. We first vary the action (64) with respect to the spinor \(\xi \), while keeping torsion and the metric fixed. A straightforward calculation produces the massless Dirac (or Weyl’s) equation

$$\begin{aligned} \sigma ^\mu {}_{a\dot{b}}\nabla _\mu \,\xi ^a -\frac{1}{2}T^\eta {}_{\eta \mu }\sigma ^\mu {}_{a\dot{b}}\,\xi ^a=0, \end{aligned}$$
(65)

which can be equivalently rewritten as

$$\begin{aligned} \sigma ^\mu {}_{a\dot{b}}\{\!\nabla \!\}_\mu \,\xi ^a \pm \frac{\mathrm{i}}{4} \varepsilon _{\alpha \beta \gamma \delta }T^{\alpha \beta \gamma } \sigma ^\delta {}_{a\dot{b}}\,\xi ^a=0. \end{aligned}$$
(66)

1.1 Energy momentum tensor

In this subsection we give the derivation of the energy momentum tensor of the action \(S_\mathrm{neutrino}\), where we vary the metric keeping the spinor fixed. The covariant and contravariant metric change in the following way

$$\begin{aligned} g_{\alpha \beta }\mapsto g_{\alpha \beta } + \delta g_{\alpha \beta }, \qquad g^{\alpha \beta }\mapsto g^{\alpha \beta } -g^{\alpha \alpha '}(\delta g_{\alpha '\beta '})g^{\beta \beta '}, \end{aligned}$$
(67)

while the Pauli matrices transform in the following way

$$\begin{aligned} \sigma _{\alpha }\mapsto \sigma _{\alpha } + \frac{1}{2}\delta g_{\alpha \beta } g^{\beta \gamma }\sigma _{\gamma }, \ \quad \ \sigma ^\alpha \mapsto \sigma ^\alpha - \frac{1}{2}g^{\alpha \beta }(\delta g_{\beta \gamma }) \sigma ^{\gamma }. \end{aligned}$$
(68)

Formulae describe a ‘symmetric’ variation of the Pauli matrices caused by the (symmetric) variation of the (symmetric) metric.

Remark 10

We do most of the following calculations under the assumption that the metric is the Minkowski metric \(g_{\mu \nu }=\mathrm diag (1,-1,-1,-1)\) and that the connection is Levi-Civita.

Now we need to look at the \(\delta {\varGamma }^{\alpha }{}_{\beta \gamma }\). Using the definition of the Levi-Civita connection, Eq. (67) and metric compatibility (\(\nabla g \equiv 0\)), we get that the connection transforms as

$$\begin{aligned} \delta {\varGamma }^{\kappa }{}_{\mu \nu } = \frac{1}{2}g^{\kappa \lambda }(\nabla _\mu \delta g_{\lambda \nu } +\nabla _\nu \delta g_{\lambda \mu }-\nabla _\lambda \delta g_{\mu \nu }). \end{aligned}$$
(69)

Lemma 3

The variation of the covariant derivative of \(\xi \) with respect to the metric is

$$\begin{aligned} \delta \nabla _\mu \xi ^a =\frac{1}{8}\left( \sigma _\alpha {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma _{\alpha }{}_{c\dot{d}}\right) \xi ^c\delta {\varGamma }^\alpha {}_{\mu \beta }. \end{aligned}$$
(70)

Proof

Using Eq. (63), the fact that \(\xi \) does not contribute to the variation and the assumptions in Remark 10, we obtain

$$\begin{aligned} 4\delta \nabla _\mu \xi ^a = \sigma _\alpha {}^{a\dot{d}}\left( \partial _\mu (\delta \sigma ^\alpha {}_{c\dot{d}}) +(\delta {\varGamma }^\alpha {}_{\mu \beta })\sigma ^{\beta }{}_{c\dot{d}} \right) \xi ^c. \end{aligned}$$

Using Eq. (67) and metric compatibility we get that

$$\begin{aligned} \partial _\mu (\delta \sigma ^\alpha {}_{c\dot{d}})= -\frac{1}{2} g^{\alpha \eta }\sigma _{\zeta }{}_{c\dot{d}}\ \delta {\varGamma }^{\zeta }{}_{\mu \eta } -\frac{1}{2} \delta ^\alpha {}_{\zeta }\ \sigma ^{\xi }{}_{c\dot{d}}\ \delta {\varGamma }^{\zeta }{}_{\mu \xi }. \end{aligned}$$

Combining this with the formula for the variation of \(\nabla \xi \), we get the equivalent to Eq. (70)

$$\begin{aligned} 4\delta \nabla _\mu \xi ^a = -\frac{1}{2}\sigma ^\beta {}^{a\dot{d}} \sigma _{\alpha }{}_{c\dot{d}}\ \xi ^c \delta {\varGamma }^{\alpha }{}_{\mu \beta } -\frac{1}{2}\sigma _\alpha {}^{a\dot{d}} \ \sigma ^{\beta }{}_{c\dot{d}}\ \xi ^c \delta {\varGamma }^{\alpha }{}_{\mu \beta } +\sigma ^{\beta }{}_{c\dot{d}}\sigma _\alpha {}^{a\dot{d}}\xi ^c \delta {\varGamma }^\alpha {}_{\mu \beta }. \end{aligned}$$

\(\square \)

We now combine Eqs. (69) and (70) to get

$$\begin{aligned} \delta \nabla _\mu \xi ^a =\frac{1}{16}\left( \sigma ^\lambda {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma ^{\lambda }{}_{c\dot{d}}\right) \xi ^c\ \left( \partial _\mu \delta g_{\lambda \beta } +\partial _\beta \delta g_{\lambda \mu } -\partial _\lambda \delta g_{\mu \beta }\right) . \end{aligned}$$

As the first derivative is symmetric over \(\lambda ,\beta \) and the Pauli matrices are antisymmetric over these indices, we get

$$\begin{aligned} \left( \sigma ^\lambda {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma ^{\lambda }{}_{c\dot{d}}\right) \partial _\mu \delta g_{\lambda \beta }= -\left( \sigma ^\lambda {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma ^{\lambda }{}_{c\dot{d}}\right) \partial _\mu \delta g_{\beta \lambda } =0. \end{aligned}$$

Hence,

$$\begin{aligned} \delta \nabla _\mu \xi ^a&= \frac{1}{16}\left( \sigma ^\lambda {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma ^{\lambda }{}_{c\dot{d}}\right) \xi ^c \partial _\beta \delta g_{\lambda \mu }\\&-\frac{1}{16}\left( \sigma ^\beta {}^{a\dot{d}}\sigma ^{\lambda }{}_{c\dot{d}} -\sigma ^\lambda {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}}\right) \xi ^c\partial _\beta \delta g_{\mu \lambda }. \end{aligned}$$

So finally, we get the formula for the variation of the covariant derivative of \(\xi \):

$$\begin{aligned} \delta \{\! \nabla \! \}_\mu \xi ^a =\frac{1}{8}\xi ^c \left( \sigma ^\alpha {}^{a\dot{d}}\sigma ^{\beta }{}_{c\dot{d}} -\sigma ^\beta {}^{a\dot{d}}\sigma ^{\alpha }{}_{c\dot{d}}\right) \partial _\beta \delta g_{\mu \alpha }. \end{aligned}$$
(71)

Lemma 4

The energy momentum tensor of the action (64) is Eq. (48).

Proof

Varying the action (64) with respect to the metric, we get

$$\begin{aligned} \delta S&= 2i\delta \int \left( \xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}}) \ -\ (\{\! \nabla \! \}_\eta \xi ^a)\,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}\right) \sqrt{|\det g|} \\&= 2i\int \xi ^a\,(\delta \sigma ^\eta {}_{a\dot{b}})\,(\{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}}) +\xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\delta \{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}}) - (\delta \{\! \nabla \! \}_\eta \xi ^a)\,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}} \\&\quad -\,(\{\! \nabla \! \}_\eta \xi ^a)\,(\delta \sigma ^\eta {}_{a\dot{b}})\,\overline{\xi }^{\dot{b}} \\&\quad +\,\frac{1}{2}\left( \xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}}) \ -\ (\{\! \nabla \! \}_\eta \xi ^a)\,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}\right) g^{\mu \nu }\delta g_{\mu \nu } \end{aligned}$$

and using Eq. (68) we get

$$\begin{aligned} \delta S&= 2i\int \left( -\frac{1}{4}\xi ^a\,g^{\eta \mu }\sigma ^{\nu }{}_{a\dot{b}}\, \left( \{ \nabla \}_\eta \overline{\xi }^{\dot{b}}\right) -\frac{1}{4}\xi ^a\,g^{\eta \nu }\sigma ^{\mu }{}_{a\dot{b}}\, \left( \{ \nabla \}_\eta \overline{\xi }^{\dot{b}}\right) \right. \\&\quad +\,\left. \frac{1}{4}\left( \{\! \nabla \! \}_\eta \xi ^a\right) \,g^{\eta \mu }\sigma ^{\nu }{}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}+\frac{1}{4}\left( \{\! \nabla \! \}_\eta \xi ^a\right) \,g^{\eta \nu }\sigma ^{\mu }{}_{a\dot{b}}\,\overline{\xi }^{\dot{b}} +\frac{1}{2}\xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}})g^{\mu \nu } \right. \\&\quad -\,\left. \frac{1}{2}\left( \{\! \nabla \! \}_\eta \xi ^a\right) \,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}g^{\mu \nu }\right) \delta g_{\mu \nu } \\&\quad +\,\xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\delta \{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}}) -(\delta \{\! \nabla \! \}_\eta \xi ^a)\,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}. \end{aligned}$$

Now we look at the terms involving the variation of \(\{\! \nabla \! \} \xi \) on their own. Using Eq. (71) we get

$$\begin{aligned} I_1&= \frac{i}{4} \int \xi ^a\,\sigma ^\mu {}_{a\dot{b}}\, \overline{\xi }^{\dot{d}}(\sigma ^\nu {}^{c\dot{b}}\sigma ^{\eta }{}_{c\dot{d}} -\sigma ^\eta {}^{c\dot{b}}\sigma ^{\nu }{}_{c\dot{d}})\partial _\eta \delta g_{\mu \nu }\\&- \xi ^c (\sigma ^\nu {}^{a\dot{d}}\sigma ^{\eta }{}_{c\dot{d}} -\sigma ^\eta {}^{a\dot{d}}\sigma ^{\nu }{}_{c\dot{d}}) \partial _\eta \delta g_{\mu \nu } \,\sigma ^\mu {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}. \end{aligned}$$

Integrating by parts and using the simplifications from Remark 10, we get

$$\begin{aligned} I_1&= \frac{i}{4} \int \overline{\xi }^{\dot{b}}\{\! \nabla \! \}_\eta \xi ^a \, \left( -\sigma ^\mu {}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^{\eta }{}_{c\dot{b}} +\sigma ^\mu {}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^{\nu }{}_{c\dot{b}}+ \sigma ^{\eta }{}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}} \right. \\&\quad \left. -\sigma ^{\nu }{}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}}-\sigma ^\mu {}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^{\eta }{}_{c\dot{b}} +\sigma ^\mu {}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^{\nu }{}_{c\dot{b}} +\sigma ^{\eta }{}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}} \right. \\&\quad \left. -\sigma ^{\nu }{}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}} \right) \delta g_{\mu \nu }. \end{aligned}$$

Since we have \(\displaystyle (\sigma ^\mu {}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^{\nu }{}_{c\dot{b}} -\sigma ^{\nu }{}_{a\dot{d}}\sigma ^\eta {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}}) \delta g_{\mu \nu } =0, \) as it is a product of symmetric and antisymmetric tensors, as well as (after a lengthy but straightforward calculation)

$$\begin{aligned} \sigma ^{\eta }{}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^\mu {}_{c\dot{b}}+ \sigma ^{\eta }{}_{a\dot{d}}\sigma ^\mu {}^{c\dot{d}}\sigma ^\nu {}_{c\dot{b}} -\sigma ^\mu {}_{a\dot{d}}\sigma ^\nu {}^{c\dot{d}}\sigma ^{\eta }{}_{c\dot{b}} -\sigma ^\nu {}_{a\dot{d}}\sigma ^\mu {}^{c\dot{d}}\sigma ^{\eta }{}_{c\dot{b}} = 0, \end{aligned}$$

we have shown that the terms involving \(\delta \{\! \nabla \! \} \xi \) do not contribute to the variation, i.e. \(I_1 = 0.\) We now return to the variation of the whole action, which after some simplification becomes

$$\begin{aligned} \frac{\delta S}{\delta g}&= \frac{i}{2}\int \left( \sigma ^{\nu }{}_{a\dot{b}} ((\{\! \nabla \! \}^\mu \xi ^a)\overline{\xi }^{\dot{b}}-\xi ^a\{\! \nabla \! \}^\mu \overline{\xi }^{\dot{b}}) +\sigma ^{\mu }{}_{a\dot{b}} ((\{\! \nabla \! \}^\nu \xi ^a)\overline{\xi }^{\dot{b}}-\xi ^a\{\! \nabla \! \}^\nu \overline{\xi }^{\dot{b}})\right) \delta g_{\mu \nu }\\&\quad +\,i\int \left( \xi ^a\,\sigma ^\eta {}_{a\dot{b}}\,(\{\! \nabla \! \}_\eta \overline{\xi }^{\dot{b}})g^{\mu \nu } - (\{\! \nabla \! \}_\eta \xi ^a)\,\sigma ^\eta {}_{a\dot{b}}\,\overline{\xi }^{\dot{b}}g^{\mu \nu }\right) \delta g_{\mu \nu }. \end{aligned}$$

Finally we can conclude that the energy momentum tensor of the action (64) is exactly Eq. (48). \(\square \)

Appendix C: Correction of explicit form of the second field equation and future work

In calculating the Bianchi identity for curvature in “Appendix B” from [53], there was an arithmetic error, hence equation (B.11) from that work should read

$$\begin{aligned} \nabla _\eta \textit{Ric}^\eta {}_\lambda = -\frac{1}{2} \textit{Ric}^\eta {}_\xi T_\eta {}^\xi {}_{\lambda } -\frac{1}{2} \fancyscript{W}^{\eta \zeta }{}_{\lambda \xi }(T_\eta {}^\xi {}_{\zeta }-T_\zeta {}^\xi {}_{\eta }), \end{aligned}$$

and from here equation (B.12) should read

$$\begin{aligned} \nabla _\eta \fancyscript{W}^\eta {}_{\mu \lambda \kappa }&= \fancyscript{W}^\eta {}_{\mu \kappa \xi }(T_\eta {}^\xi {}_{\lambda }-T_\lambda {}^\xi {}_{\eta } )+\fancyscript{W}^\eta {}_{\mu \lambda \xi }(T_\kappa {}^\xi {}_{\eta }-T_\eta {}^\xi {}_{\kappa })\\&\quad +\,\frac{1}{4}(T_\zeta {}^\xi {}_{\eta }-T_\eta {}^\xi {}_{\zeta })(g_{\mu \lambda }\fancyscript{W}^{\eta \zeta }{}_{\kappa \xi }-g_{\mu \kappa }\fancyscript{W}^{\eta \zeta }{}_{\lambda \xi })\\&+\,\frac{1}{4} \textit{Ric}^\eta {}_\xi (g_{\mu \lambda }T_\eta {}^\xi {}_{\kappa }-g_{\mu \kappa }T_\eta {}^\xi {}_{\lambda })\\&\quad +\,\frac{1}{2}\left[ \nabla _\lambda \textit{Ric}_{\mu \kappa }- \nabla _\kappa \textit{Ric}_{\mu \lambda } + \textit{Ric}^\eta {}_\kappa (T_{\lambda \eta \mu }-T_{\eta \lambda \mu }) \right. \\&\left. +\textit{Ric}^\eta {}_\lambda (T_{\eta \kappa \mu }-T_{\kappa \eta \mu })\right] \end{aligned}$$

Consequently, when these two are used in calculating the explicit form of the field equation (3) this produces a different result to the one presented in [53]. Namely, Eq. (3) in its explicit form (i.e. equation (3) from [53]) should read

$$\begin{aligned}&d_6 \nabla _\lambda \textit{Ric}_{\kappa \mu } - d_7 \nabla _\kappa \textit{Ric}_{\lambda \mu }\\&\quad +\,d_6\left( Ric^\eta {}_\kappa \left( T_{\eta \mu \lambda }-T_{\lambda \mu \eta }\right) +\frac{1}{2} g_{\mu \lambda } \fancyscript{W}^{\eta \zeta }{}_{\kappa \xi }(T_\eta {}^\xi {}_{\zeta }-T_\zeta {}^\xi {}_{\eta }) +\frac{1}{2} g_{\mu \lambda } \textit{Ric}^\eta {}_\xi T_\eta {}^\xi {}_{\kappa } \right) \\&\quad +\,d_7 \left( \textit{Ric}^\eta {}_\lambda \left( T_{\eta \mu \kappa } - T_{\kappa \mu \eta } \right) +\frac{1}{2} g_{\kappa \mu } \fancyscript{W}^{\eta \zeta }{}_{\lambda \xi }(T_\eta {}^\xi {}_{\zeta }-T_\zeta {}^\xi {}_{\eta }) +\frac{1}{2} g_{\kappa \mu }\textit{Ric}^\eta {}_\xi T_\eta {}^\xi {}_{\lambda } ) \right) \\&\quad +\,b_{10} \left( T_\eta {}^\xi {}_{\zeta }-T_\zeta {}^\xi {}_{\eta })(g_{\mu \kappa }\fancyscript{W}^{\eta \zeta }{}_{\lambda \xi }-g_{\mu \lambda }\fancyscript{W}^{\eta \zeta }{}_{\kappa \xi }\right) \\&\quad +\,2b_{10}\left( \fancyscript{W}^\eta {}_{\mu \kappa \xi }(T_\eta {}^\xi {}_{\lambda }-T_\lambda {}^\xi {}_{\eta } ) + \fancyscript{W}^\eta {}_{\mu \lambda \xi }(T_\kappa {}^\xi {}_{\eta }-T_\eta {}^\xi {}_{\kappa }) - \fancyscript{W}^{\xi \eta }{}_{\kappa \lambda } T_{\eta \mu \xi } \right) =0. \end{aligned}$$

This change in no way affects the proof of the main theorem, as this form of the field equation is even simpler and does not contain any additional terms, but we believed it was important to point out for purposes of future work. This mistake was noticed in the process of generalising the explicit form of the field equations (2), (3), i.e. the equations equation (26) and (27) from [53], see [52].

The two papers of Singh [60, 61] are of particular interest to us, where the author constructs solutions for the Yang–Mills case (4) with purely axial and purely trace torsion respectively, see (9), and unlike the solution of [63], \(\{Ric\}\) is not assumed to be zero. It is obvious that these solutions differ from the ones presented in our work, as the torsion of generalised pp-waves is assumed to be purely tensor. It would however be of interest to us to see whether this construction of Singh’s can be expanded to our most general \(\mathrm{O}(1,3)\)-invariant quadratic form \(q\) with 16 coupling constants.

We also hope to see whether it is possible to produce torsion waves which are purely axial or trace and combine them with the pp-wave metric, in a similar fashion as was done with purely tensor torsion waves, in order to produce new solutions of quadratic metric-affine gravity and also give their physical interpretation in the near future.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pasic, V., Barakovic, E. PP-waves with torsion: a metric-affine model for the massless neutrino. Gen Relativ Gravit 46, 1787 (2014). https://doi.org/10.1007/s10714-014-1787-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-014-1787-y

Keywords

Navigation