Skip to main content

Generic Theory of Geometrodynamics from Noether’s Theorem for the \(\mathrm {Diff}(M)\) Symmetry Group

  • Chapter
  • First Online:
Discoveries at the Frontiers of Science

Abstract

We work out the most general theory for the interaction of spacetime geometry and matter fields—commonly referred to as geometrodynamics—for spin-0 and spin-1 particles. Actually, we present a Hamilton–Lagrange–Noether formulation of the gauge theory of gravitation. It is based on the minimum set of postulates to be introduced, namely (i) the action principle and (ii) the form-invariance of the action under the (local) diffeomorphism group. The second postulate thus implements the Principle of General Relativity, also referred to as the Principle of General Covariance. According to Noether’s theorem, this physical symmetry gives rise to a conserved Noether current, from which the complete set of theories compatible with both postulates can be deduced. This finally results in a new generic Einstein-type equation, which can be interpreted as an energy-momentum balance equation emerging from the Lagrangian \(\mathscr {L}_{R}\) for the source-free dynamics of gravitation and the energy-momentum tensor of the source system \(\mathscr {L}_{0}\). Provided that the system has no other symmetries—such as SU(N)—the canonical energy-momentum tensor turns out to be the correct source term of gravitation. For the case of massive spin-1 particles, this entails an increased weighting of the kinetic energy over the mass as the source of gravity, compared to the metric energy momentum tensor, which constitutes the source of gravity in Einstein’s General Relativity. We furthermore confirm that a massive vector field necessarily acts as a source for torsion of spacetime. From the viewpoint of our generic Einstein-type equation, Einstein’s General Relativity constitutes the particular case for scalar and massless vector particle fields, and the Hilbert Lagrangian \(\mathscr {L}_{R,\mathrm {H}}\) as the model for the source-free dynamics of gravitation.

Dedicated to the memory of Prof. Dr. Walter Greiner, our teacher, mentor, and friend.

In this contribution, we present the canonical transformation formalism in the realm of classical field theory, where spacetime is treated as a dynamical quantity, and apply it to formulate the gauge theory of gravitation. In this respect, it generalizes the Extended Hamilton–Lagrange–Jacobi formalism of relativistic point dynamics. Walter was very much interested in this formalism and therefore added several chapters on the matter to the second edition of his textbook “Classical Mechanics” [1]. To quote Walter from the Preface to the Second Edition: “It may come as a surprise that even for the time-honored subject such as Classical Mechanics in the formulation of Lagrange and Hamilton, new aspects may emerge.” We are sure, Walter would have loved the following elaboration.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 139.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 179.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 179.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. W. Greiner, Classical Mechanics, 2nd edn. (Springer, Berlin, 2010)

    Book  Google Scholar 

  2. J. Struckmeier, J. Muench, D. Vasak, J. Kirsch, M. Hanauske, H. Stoecker, Phys. Rev. D 95, 124048 (2017). https://doi.org/10.1103/PhysRevD.95.124048

    Article  ADS  MathSciNet  Google Scholar 

  3. A. Einstein, Private letter to Hermann Weyl (ETH Zürich Library, Archives and Estates, 1918)

    Google Scholar 

  4. D. Kehm, J. Kirsch, J. Struckmeier, D. Vasak, M. Hanauske, Astron. Nachr./AN 338(9–10), 1015 (2017). https://doi.org/10.1002/asna.201713421

  5. F.W. Hehl, P. von der Heyde, G.D. Kerlick, J.M. Nester, Rev. Mod. Phys. 48(3), 393 (1976). https://doi.org/10.1103/revmodphys.48.393

    Article  ADS  Google Scholar 

  6. E. Noether, Nachrichten der Königlichen Gesesellschaft der Wissenschaften Göttingen. Mathematisch-Physikalische Klasse 57, 235 (1918)

    Google Scholar 

  7. J. Struckmeier, A. Redelbach, Int. J. Mod. Phys. E 17, 435 (2008). https://doi.org/10.1142/s0218301308009458

    Article  ADS  Google Scholar 

  8. M. Gaul, C. Rovelli, Lect. Notes Phys. 541, 277 (2000)

    Article  ADS  Google Scholar 

  9. P. Matteucci, Rep. Math. Phys. 52, 115 (2003). https://doi.org/10.1016/S0034-4877(03)90007-3, arXiv:gr-qc/0201079

  10. M. Godina, P. Matteucci, Int. J. Geom. Methods Mod. Phys. 2, 159 (2005). https://doi.org/10.1142/S0219887805000624, arXiv:math/0504366

  11. J. Schouten, Ricci-Calculus (Springer. Berlin (1954). https://doi.org/10.1007/978-3-662-12927-2

    Article  Google Scholar 

  12. F.W. Hehl, Rep. Math. Phys. 9(3), 55 (1976)

    Article  ADS  Google Scholar 

  13. F.W. Hehl, in Proceedings, 15th Workshop on High Energy Spin Physics (DSPIN-13), Dubna, Russia, 8–12 October 2013 (2014)

    Google Scholar 

  14. C.W. Misner, K.S. Thorne, J.A. Wheeler, Gravitation (W. H. Freeman and Company, New York, 1973)

    Google Scholar 

  15. P. Jordan, Ann. der Phys. 428(1), 64 (1939)

    Article  ADS  Google Scholar 

  16. D. Sciama, Mon. Not. R. Astron. Soc. 113, 34 (1953)

    Article  ADS  MathSciNet  Google Scholar 

  17. R. Feynman, W. Morinigo, W. Wagner, Feynman Lectures On Gravitation, Frontiers in Physics (Westview Press, Boulder, 2002)

    Google Scholar 

  18. S. Hawking, The Theory of Everything (New Millenium Press, 2003)

    Google Scholar 

  19. S. Carroll, Spacetime and Geometry (Prentice Hall, Englewood Cliffs, 2013)

    Google Scholar 

  20. T.W.B. Kibble, J. Math. Phys. 2, 212 (1961)

    Article  ADS  Google Scholar 

  21. D.W. Sciama, in Recent Developments in General Relativity (Pergamon Press, Oxford; PWN, Warsaw, 1962), pp. 415–439. Festschrift for Infeld

    Google Scholar 

  22. J. Plebanski, A. Krasinski, An Introduction to General Relativity and Cosmology (Cambridge University Press, Cambridge, 2006)

    Book  Google Scholar 

  23. J. Struckmeier, P. Liebrich, J. Muench, M. Hanauske, J. Kirsch, D. Vasak, L. Satarov, H. Stoecker, Int. J. Mod. Phys. E 28(1), 1950007 (2019). https://doi.org/10.1142/S0218301319500071, arXiv:1711.10333

  24. K. Hayashi, T. Shirafuji, Prog. Theor. Phys. 64(3), 866, 883, 1435, 2222 (1980). https://doi.org/10.1143/PTP.64.866

Download references

Acknowledgements

First of all, we want to remember our revered academic teacher Walter Greiner, whose charisma and passion for physics inspired us to stay engaged in physics for all of our lives.

The authors thank Patrick Liebrich and Julia Lienert (Goethe University Frankfurt am Main and FIAS), and Horst Stoecker (FIAS, GSI, and Goethe University Frankfurt am Main) for valuable discussions. D.V. and J.K. thank the Fueck Foundation for its support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jürgen Struckmeier .

Editor information

Editors and Affiliations

Appendices

Identity for a Scalar-Valued Function S of an (nm)-Tensor T and the Metric

Proposition 1

Let \(S=S(g,T)\in \mathbb {R}\) be a scalar-valued function constructed from the metric tensor \(g_{\mu \nu }\) and an (nm)-tensor , where \((m-n)/2\in \mathbb {Z}\). Then the following identity holds:

(89)

Proof

The induction hypothesis is immediately verified for scalars constructed from second rank tensors and, if necessary, the metric, hence, , \(S=T^{\alpha \beta }\,g_{\alpha \beta }\), and \(S=T_{\alpha \beta }\,g^{\alpha \beta }\). Let Eq. (89) hold for an (nm)-tensor . We first consider an \((n+1,m+1)\)-tensor with the last indices contracted in order to again make up a scalar. Setting up S according to (89) with the tensor \(\bar{T}\), one encounters the two additional terms

Equation (89) thus also holds for the scalar S formed from the \((n+1,m+1)\)-tensor \(\bar{T}\).

For the case that \(\bar{T}\) represents an \((n+2,m)\)-tensor , the scalar S must have one additional factor \(g_{\alpha \beta }\). One thus encounters four additional terms:

For the case that \(\bar{T}\) represents an \((n,m+2)\)-tensor , the scalar S must have one additional factor \(g^{\alpha \beta }\). Owing to

$$\begin{aligned} {\frac{\partial {S}}{\partial {g_{\mu \beta }}}}g_{\nu \beta }=-{\frac{\partial {S}}{\partial {g^{\alpha \nu }}}}g^{\alpha \mu },\qquad {\frac{\partial {S}}{\partial {g_{\alpha \mu }}}}g_{\alpha \nu }=-{\frac{\partial {S}}{\partial {g^{\nu \beta }}}}g^{\mu \beta }, \end{aligned}$$

one thus encounters the four additional terms:

The derivative of a Lagrangian \(\mathscr {L}\) with respect to the metric \(g_{\mu \nu }\) can thus always be replaced by the derivatives with respect to the appertaining tensors T that are made into a scalar by means of the metric. The identity thus provides the correlation of the metric and the canonical energy-momentum tensors of a given system.

Corollary 1

The contraction of Eq. (89) then yields a condition for the scalar S:

(90)

Proof

Contracting Eq. (89) directly yields Eq. (90).

Corollary 2

Let \(\tilde{\mathscr {L}}=\tilde{\mathscr {L}}(g,T_k)\in \mathbb {R}\) be a scalar density (i.e. a relative scalar of weight one) valued function of the (symmetric) metric \(g^{\mu \nu }\) and a sum of k tensors of respective rank \((n_k,m_k)\), where \((m_k-n_k)/2\in \mathbb {Z}\). Then the following identity holds:

(91)

Proof

Combine Eq. (89) with (58).

Equation (91) is obviously a representation of Euler’s theorem on homogeneous functions in the realm of tensor calculus.

Examples for Identities (89) Involving the Riemann Tensor

1.1 Riemann Tensor Squared

As a scalar, any Lagrangian \(\mathscr {L}_{R}(R,g)\) built from the Riemann–Cartan tensor (41) and the metric satisfies the identity (89)

(92)

The factors “2” emerge from the symmetry of the metric and the skew-symmetry of the Riemann–Cartan tensor in its last index pair. The identity is easily verified for a Lagrangian linear and quadratic in the Riemann tensor:

The left-hand side of Eq. (92) evaluates to

which indeed agrees with the terms obtained from the right-hand side:

1.2 Ricci Scalar

The Ricci scalar R is defined as the following contraction of the Riemann tensor

(93)

With the scalar \(\mathscr {L}_{R}=R\) and the tensor T the Riemann tensor, the general Eq. (89) takes on the particular form

$$\begin{aligned} {\frac{\partial {R}}{\partial {g^{\nu \beta }}}}g^{\mu \beta }+{\frac{\partial {R}}{\partial {g^{\beta \nu }}}}g^{\beta \mu }- {\frac{\partial {R}}{\partial {R_{\mu \alpha \xi \lambda }}}}R_{\nu \alpha \xi \lambda }- {\frac{\partial {R}}{\partial {R_{\eta \mu \xi \lambda }}}}R_{\eta \nu \xi \lambda }- {\frac{\partial {R}}{\partial {R_{\eta \alpha \mu \lambda }}}}R_{\eta \alpha \nu \lambda }- {\frac{\partial {R}}{\partial {R_{\eta \alpha \xi \mu }}}}R_{\eta \alpha \xi \nu }\equiv 0. \end{aligned}$$
(94)

Without making use of the symmetries of the Riemann tensor and the metric, this identity is actually fulfilled as

Similarly

The derivative terms of the Riemann tensor are

and

which obviously cancel the four terms emerging from the derivatives with respect to the metric.

Making now use of the skew-symmetries of the Riemann tensor in its first and second index pair and of the symmetry of the metric, Eq. (94) simplifies to

$$\begin{aligned} {\frac{\partial {R}}{\partial {g^{\nu \beta }}}}g^{\mu \beta }\equiv {\frac{\partial {R}}{\partial {R_{\mu \alpha \xi \lambda }}}}R_{\nu \alpha \xi \lambda }+ {\frac{\partial {R}}{\partial {R_{\eta \alpha \xi \mu }}}}R_{\eta \alpha \xi \nu }. \end{aligned}$$
(95)

For zero torsion, the Riemann tensor has the additional symmetry on exchange of both index pairs. Then

(96)

1.3 Ricci Tensor Squared

The scalar made of the (not necessarily symmetric) Ricci tensor \(R_{\eta \alpha }\) is defined by the following contraction with the metric

(97)

With Eq. (97), the general Eq. (89) now takes on the particular form

$$\begin{aligned} {\frac{\partial {\mathscr {L}_{R}}}{\partial {g^{\nu \beta }}}}g^{\mu \beta }+{\frac{\partial {\mathscr {L}_{R}}}{\partial {g^{\beta \nu }}}}g^{\beta \mu }- {\frac{\partial {\mathscr {L}_{R}}}{\partial {R_{\mu \beta }}}}R_{\nu \beta }-{\frac{\partial {\mathscr {L}_{R}}}{\partial {R_{\beta \mu }}}}R_{\beta \nu }\equiv 0. \end{aligned}$$
(98)

Without making use of the symmetries of the Ricci tensor and the metric, this identity is actually fulfilled as

Similarly

The derivative terms of the Ricci tensor are

and

which obviously cancel the four terms emerging from the derivatives with respect to the metric.

For zero torsion, the Ricci tensor is symmetric. Then

(99)

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Struckmeier, J., Vasak, D., Kirsch, J. (2020). Generic Theory of Geometrodynamics from Noether’s Theorem for the \(\mathrm {Diff}(M)\) Symmetry Group. In: Kirsch, J., Schramm, S., Steinheimer-Froschauer, J., Stöcker, H. (eds) Discoveries at the Frontiers of Science. FIAS Interdisciplinary Science Series. Springer, Cham. https://doi.org/10.1007/978-3-030-34234-0_12

Download citation

Publish with us

Policies and ethics