Skip to main content
Log in

General-Relativistic Covariance

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

This is an essay about general covariance, and what it says (or doesn’t say) about spacetime structure. After outlining a version of the dynamical approach to spacetime theories, and how it struggles to deal with generally covariant theories, I argue that we should think about the symmetry structure of spacetime rather differently in generally-covariant theories compared to non-generally-covariant theories: namely, as a form of internal rather than external symmetry structure.

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. For details of the pushforward operation on tensor fields, see e.g. [28, §1.5].

  2. Note that \(\eta _{ab}\) is not acted upon by the diffeomorphism, since it is considered part of the Minkowski spacetime rather than as a field (we will soon see what happens if it is instead treated as a field). Compare [39]’s “final version” of diffeomorphism invariance.

  3. Introductions to fibre bundles for philosophers may be found in [51, Appendix 1] or [21, Appendix B].

  4. Cf. Maudlin [29].

  5. That is, with the points whose associated base-space points are infinitesimally close to the original fibre’s base-space point. Note that this doesn’t typically yield an identification of arbitrary points of fibres with one another: unless the connection is “flat”, the identification between finitely separated fibres will be path-dependent.

  6. See Kolář et al. [25], and especially chapter 12, for an exposition of the theory of natural bundles.

  7. Cf. the notion of an induction in Curiel [15].

  8. This theory is discussed in [17, §3.5] and [19, § III.5].

  9. Note that this simpler definition of Newtonian spacetime (which I learned from David Wallace) still captures the relevant structure: the persistence of points of absolute space is represented by the projection \(\pi _X : T \times X \rightarrow X\) (i.e. two points (tx) and \((t', x')\) of Newtonian spacetime correspond to the same persisting points of absolute space just in case \(x = x'\)).

  10. Let \(\delta ^T_{ab}\) be the Euclidean metric on T, \(\delta ^X_{ab}\) be the Euclidean metric on X, and any vector \(\xi ^a\) over \(T \times X\) decompose as \((\xi ^a_T, \xi ^a_X)\). Then we may define \(\eta _{ab}\) by \(\eta _{ab}\xi ^a \zeta ^b := \delta ^T_{ab}\xi ^a_T \zeta ^b_T - \delta ^X_{ab}\xi ^a_X \zeta ^b_X\).

  11. See Barrett [4].

  12. For readers who are interested in these controversies, see Friedman [19], Earman [17], Saunders [44], Caulton [14], Dewar [16], as well as the papers in Brading and Castellani [8] and references therein.

  13. The primary text for the dynamical approach is Brown [10]; but see also Brown [9], Brown and Pooley [11], Read et al. [40], and Brown and Read [12].

  14. Cf. Norton [35].

  15. Norton [34] gives a detailed account of how these strategies feature in the work of Klein and Riemann.

  16. For a broader discussion of symmetries in the context of coordinate-based theories, see Olver [36].

  17. This corresponds to the older way of thinking about geometric objects, prior to the introduction of natural bundles in Nijenhuis [32]. For more on this tradition, see Schouten and Haantjes [46] and Nijenhuis [31]; for modern discussion of this tradition, see [2, chap. 1] and Pitts [38].

  18. Of course, these are just the expected transformation rules for a rank-(0, 2) tensor and a vector. But the rules are not to be thought of as manifestations of the fact that we are representing (independently defined) geometric objects in coordinates; rather, the rules are constitutive of those objects. This is why I specify the rules explicitly.

  19. See [10, chap. 4] for some of the history of how the symmetry group of Maxwell’s equations was determined.

  20. See also Myrvold [30] and Acuña [1] for further discussion and defence of the idea that the claim “spacetime symmetries are dynamical symmetries” is analytic. As Myrvold (§5) discusses, it is not clear to what extent Brown [10] should be interpreted as supporting this way of reading the dynamical approach (although Brown and Read [12] declare themselves sympathetic).

  21. The sense in which Minkowski spacetime is a substructure of \(\mathbb {R}^4\) is just that all the structure of the former can be defined in terms of the structure of the latter: that is, one can define the Minkowski metric in \(\mathbb {R}^4\) by \(\eta _{\mu \nu } \xi ^\mu \zeta ^\nu := \xi ^0 \zeta ^0 - \xi ^1 \zeta ^1 \xi ^2 \zeta ^2 - \xi ^3 \zeta ^3\).

  22. Contra Friedman [19], Earman [17].

  23. Barrett [4]

  24. For more discussion of the relationship between symmetries and relativity principles, see Brown and Sypel [13].

  25. By “manifold structure”, I mean the structure of a differentiable manifold (i.e., a set equipped with an atlas of compatible charts).

  26. The stipulation that the coordinate system go to the identity as we approach the boundary is necessary because we are, after all, only treating some subregion of spacetime. For more discussion of the subtleties here, see Belot [6].

  27. See Norton [33] for further discussion.

  28. This need not be a criticism of the dynamical approach per se, insofar as the dynamical approach could perhaps be adapted to generally covariant theories (by, for instance, dropping the claim of ontological reduction): see [10, chap. 5] and [24, §3]. (I thank an anonymous referee and James Read for pressing this point.)

  29. Cf. Norton [34].

  30. This observation, of course, goes back to Kretschmann [26].

  31. Alternatively, one could argue that the theory’s spatiotemporal commitments should be the intersection of the commitments of its various formulations. But it’s not clear to me what the motivation for this move would be, beyond a vague appeal to supervaluationist semantics; and in any event, the effect of such a prescription will be to claim that any theory with a generally covariant formulation is—at most—committed to spacetime being a manifold.

  32. How to pick out the appropriate subset will be addressed in Sect. 6.

  33. Trivially, of course, whatever \(g_{\alpha \beta }\) were this claim would be true for the group \(\{\delta ^\alpha _\beta \}\). But the existence of a nontrivial group of transformation matrices, under which \(g_{\alpha \beta }\) is invariant, reflects important facts about \(g_{\alpha \beta }\).

  34. I say “as a group” because the matrices themselves will not be the “Lorentz matrices” one sometimes meets in introductions to special relativity: those are the real-valued matrices which preserve the coefficients \(\eta _{\mu \nu }\), i.e., which preserve the diagonalised form of the metric. But as discussed below, there are transformations which transform between \(\eta _{\mu \nu }\) and \(g_{\mu \nu }\); these same transformations will transform between and , constituting a group isomorphism between the two sets of matrices.

  35. Without a proviso of this sort, we would end up counting arbitrary permutations of the space of models as dynamical symmetries (cf. Belot [5]’s “Fruitless Definition” of symmetry).

  36. It is a “confined object” in the sense of Pitts [37].

  37. My use of uppercase indices for internal vector spaces follows Weatherall [51].

  38. For more on the tetrad formalism, see Rovelli [43] and Wallace [48]. Read et al. [40] offer an alternative analysis of how to get at the Lorentz symmetry of a generally-covariant theory, by seeking to show that minimally coupled dynamical equations will be invariant under (only) Poincaré transformations. However, they are using a somewhat different sense of “invariant”: see Appendix 1.

  39. For discussion of and references on the tetrad formalism applied to Newton-Cartan theory, see Read and Teh [41].

  40. Cf. Weatherall [51].

  41. Knox [24]. Note that as Robertson [42] observes, almost any theoretical quantity could be taken to be functionally defined by its total role in that specific theory; the hard (and interesting) project is finding smaller functional roles that are present in distinct theories, and so let us identify physical quantities across theories.

  42. [10, p. 169]

  43. The existence of such coordinates is related to the considerations raised in Sect. 5: if \(\xi ^\mu \) are Riemann normal coordinates for p, and \(x^\mu \) are our original coordinates, then the matrix given by the value at p of \(\partial \xi ^\mu /\partial x^\alpha \) will diagonalise the metric in the manner of equation (13). But the presence of inertial coordinates is stronger than the possibility of such a diagonalisation, for two reasons: (a) diagonalising amounts to a choice of basis at each point, with no requirement that it be a coordinate (or “holonomic”) basis; (b) inertial coordinates have the further property that the connection coefficients vanish.

  44. In the context of Newtonian gravitation, for instance, Knox [23] argues that the local inertial frames are the free-fall frames, and hence that the spacetime structure is encoded in the curved Newton-Cartan connection rather than the flat Galilean connection: in other words, that the connection coefficients are not the Galilean coefficients \(\Gamma ^\rho _{\mu \nu }\), but rather the Newton-Cartan coefficients \(\Gamma ^\rho _{\mu \nu } + t_{\mu \nu } h^{\rho \sigma } \nabla _\sigma \phi \) (where \(\phi \) is the gravitational field).

  45. [22, p. 348]

  46. That said, there are ways in which one can see the role of spacetime structure in the field equations as being what gives rise to such equations of motion [52]: one can use a variational analysis to ground a certain kind of conservation condition, and then employ that condition to prove an appropriate equation of motion. The best-known example of this kind of construction is the geodesic theorem in GR, but one can similarly prove a geodesic theorem in Newtonian theories [50], and the Lorentz force law in electromagnetism [20].

  47. von Helmholtz [47], Weyl [53]; for contemporary discussion, see Bernard [7], Scholz [45], or Eisenthal [18].

  48. Lam and Wüthrich [27], Baker [3]

  49. Assuming that this is how the left-hand equality is meant to be justified.

  50. The remarks below draw heavily on correspondence with James Read.

  51. This paraphrases an argument put to me by Read (in correspondence).

References

  1. Acuña, P.: Minkowski spacetime and Lorentz invariance: the cart and the horse or two sides of a single coin? Stud. Hist. Philos. Sci. Part B 55, 1–12 (2016)

    Article  MathSciNet  Google Scholar 

  2. Anderson, J.L.: Principles of Relativity Physics. Academic Press, New York (1967)

    Book  Google Scholar 

  3. Baker, D.J. On spacetime functionalism [preprint]. (2019). Retrieved 24 April 2019, from http://philsci-archive.pitt.edu/15860/

  4. Barrett, T.W.: Spacetime Structure. Stud. Hist. Philos. Sci. Part B 51, 37–43 (2015)

    Article  MathSciNet  Google Scholar 

  5. Belot, G.: Symmetry and equivalence. In: Batterman, R.W. (ed.) The Oxford Handbook of Philosophy of Physics. Oxford University Press, New York (2013)

    Google Scholar 

  6. Belot, G.: Fifty million elvis fans can’t be wrong. Noûs 52(4), 946–981 (2018). https://doi.org/10.1111/nous.12200

    Article  MathSciNet  Google Scholar 

  7. Bernard, J.: Riemann’s and Helmholtz-Lie’s problems of space from Weyl’s relativistic perspective. Stud. Hist. Philos. Sci. Part B 61, 41–56 (2018)

    Article  MathSciNet  Google Scholar 

  8. Brading, K., Castellani, E. (eds.): Symmetries in Physics: Philosophical Reflections. Cambridge University Press, Cambridge (2003)

    MATH  Google Scholar 

  9. Brown, H.R.: The origins of length contraction: I. The FitzGerald–Lorentz deformation hypothesis. Am. J. Phys. 69(10), 1044–1054 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  10. Brown, H.R.: Physical Relativity: Space-Time Structure from a Dynamical Perspective. Oxford University Press, Oxford (2005)

    Book  Google Scholar 

  11. Brown, H.R., Pooley, O.: Minkowski Space-Time: A Glorious Non-Entity. In: Dieks, D. (ed.) The Ontology of Spacetime, vol. 1, pp. 67–89. Elsevier, Amsterdam (2006)

    Chapter  Google Scholar 

  12. Brown, H.R., Read, J.: The dynamical approach to spacetime theories. In: Knox, E., Wilson, A. (eds) The Routledge Companion to Philosophy of Physics, page 25. Routledge, London. Forthcoming (philsci-archive: 14592) (2018)

  13. Brown, H.R., Sypel, R.: On the meaning of the relativity principle and other symmetries. Int. Stud. Philos. Sci. 9(3), 235–253 (1995)

    Article  MathSciNet  Google Scholar 

  14. Caulton, A.: The role of symmetry in the interpretation of physical theories. Stud. Hist. Philos. Sci. Part B 52, 153–162 (2015)

    Article  MathSciNet  Google Scholar 

  15. Curiel, E.: On Geometric Objects, the Non-Existence of a Gravitational Stress-Energy Tensor, and the Uniqueness of the Einstein Field Equation. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. Forthcoming. References are to draft of February 2nd, 2017 (2017)

  16. Dewar, N.: Sophistication about symmetries. Br. J. Philos. Sci. Forthcoming (2017)

  17. Earman, J.: World Enough and Space-Time: Absolute versus Relational Theories of Space and Time. MIT Press, Cambridge, MA (1989)

    Google Scholar 

  18. Dewar, N., Eisenthal J.: A raum with a view: Hermann Weyl and the problem of space. In: Beisbart, C., Sauer, T., Wüthrich, C. (eds.) Thinking About Space and Time. Springer. Forthcoming

  19. Friedman, M.: Foundations of Space-Time Theories: Relativistic Physics and Philosophy of Science. Princeton University Press, Princeton, NJ (1983)

    Google Scholar 

  20. Geroch, R., Weatherall, J.O.: The motion of small bodies in space-time. Commun. Math. Phys. 364(2), 607–634 (2018). https://doi.org/10.1007/s00220-018-3268-8

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Healey, R.: Gauging What’s Real. Oxford University Press, Oxford, UK (2007)

    Book  Google Scholar 

  22. Knox, E.: Effective spacetime geometry. Stud. Hist. Philos. Sci. Part B 44(3), 346–356 (2013)

    Article  MathSciNet  Google Scholar 

  23. Knox, E.: Newtonian spacetime structure in light of the equivalence principle. Br. J. Philos. Sci. 65(4), 863–880 (2014)

    Article  MathSciNet  Google Scholar 

  24. Knox, E.: Physical relativity from a functionalist perspective. Stud. Hist. Philos. Sci. Part B (2017)

  25. Kolář, I., Michor, P.W., Slovák, J.: Natural Operations in Differential Geometry, Electronic edn. Springer-Verlag, Berlin, Heidelberg (1993)

    Book  Google Scholar 

  26. Kretschmann, E.: über den physikalischen Sinn der Relativitätspostulate, A. Einsteins neue und seine ursprüngliche Relativitätstheorie. Ann. Phys. 53(16), 575–614 (1917)

    Article  Google Scholar 

  27. Lam, V., Wüthrich, C.: Spacetime is as spacetime does. Stud. Hist. Philos. Sci. Part B 64, 39–51 (2018). https://doi.org/10.1016/j.shpsb.2018.04.003

    Article  MathSciNet  MATH  Google Scholar 

  28. Malament, D.B.: Topics in the Foundations of General Relativity and Newtonian Gravitation Theory. University of Chicago Press, Chicago, IL (2012)

    Book  Google Scholar 

  29. Maudlin, T.: Suggestions from Physics for Deep Metaphysics. In: Maudlin, T. (ed.) The Metaphysics Within Physics, pp. 78–103. Oxford University Press, Oxford (2007)

    Chapter  Google Scholar 

  30. Myrvold, W.C.: How could relativity be anything other than physical? Stud. Hist. Philos. Sci. Part B (2017). https://doi.org/10.1016/j.shpsb.2017.05.007

    Article  MATH  Google Scholar 

  31. Nijenhuis, A.: Theory of the Geometric Object. PhD thesis, Universiteit van Amsterdam (1952)

  32. Nijenhuis, A.: Natural bundles and their general properties. Differ. Geom. Honor of K. Yano, Kinokuniya, Tokyo 317, 334 (1972)

    MATH  Google Scholar 

  33. Norton, J.D.: Coordinates and covariance: Einstein’s view of space-time and the modern view. Found. Phys. 19(10), 1215–1263 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  34. Norton, J.D.: Geometries in Collision: Einstein, Klein and Riemann. In: Gray, J.J. (ed.) The Symbolic Universe: Geometry and Physics 1890–1930, pp. 128–144. Oxford University Press, Oxford; New York (1999)

    Google Scholar 

  35. Norton, J.D.: Why constructive relativity fails. Br. J. Philos. Sci. 59(4), 821–834 (2008)

    Article  MathSciNet  Google Scholar 

  36. Olver, P.J.: Applications of Lie Groups to Differential Equations. Springer-Verlag, New York (1986)

    Book  Google Scholar 

  37. Pitts, J.B.: Absolute objects and counterexamples: Jones-Geroch dust, Torretti constant curvature, tetrad-spinor, and scalar density. Stud. Hist. Philos. Sci. Part B 37(2), 347–371 (2006)

    Article  MathSciNet  Google Scholar 

  38. Pitts, J.B.: The nontriviality of trivial general covariance: How electrons restrict ‘time’ coordinates, spinors (almost) fit into tensor calculus, and 716 of a tetrad is surplus structure. Stud. Hist. Philos. Sci. Part B 43(1), 1–24 (2012)

    Article  MathSciNet  Google Scholar 

  39. Pooley, O.: Background independence, diffeomorphism invariance, and the meaning of coordinates. In: Lehmkuhl, D., Schiemann, G., Scholz, E. (eds.) Towards a Theory of Spaceme Theories, number 13 in Einstein Studies. Birkhäuser, Basel (2017)

    Google Scholar 

  40. Read, J., Brown, H.R., Lehmkuhl, D.: Two miracles of general relativity. Stud. Hist. Philos. Sci. Part B 64, 14–25 (2018). https://doi.org/10.1016/j.shpsb.2018.03.001

    Article  MathSciNet  MATH  Google Scholar 

  41. Read, J., Teh, N.J.: The teleparallel equivalent of Newton–Cartan gravity. Class. Quantum Gravity 35(18), 18LT01 (2018)

    Article  MathSciNet  Google Scholar 

  42. Robertson, K.: Functionalism fit for physics. Unpublished (2018)

  43. Rovelli, C.: Quantum Gravity. Cambridge University Press, Cambridge, New York (2004)

    Book  Google Scholar 

  44. Saunders, S.: To what physics corresponds. In: French, S., Kamminga, H. (eds.) Correspondence, Invariance and Heuristics, number 148 in Boston Studies in the Philosophy of Science, pp. 295–325. Kluwer, Dordrecht (1993)

    Chapter  Google Scholar 

  45. Scholz, E.: The problem of space in the light of relativity: The views of Hermann Weyl and Elie Cartan. In: Bioesmat-Martagon, L. (ed.) Eléments d’une Biographie de l’espace Géométrique, pp. 255–312. Presses Universitaires de Nancy, Nancy (2016)

    Google Scholar 

  46. Schouten, J.A., Haantjes, J.: On the theory of the geometric object. Proc. Lond. Math. Soc. s2–42(1), 356–376 (1937)

    Article  MathSciNet  Google Scholar 

  47. von Helmholtz, H.: über den Ursprung und die Bedeutung der geometrischen Axiome. In: Vorträge Und Reden, volume 2, pp. 1–31. Vieweg, Braunschweig, 4th edition (1896)

  48. Wallace, D.: Fields as Bodies: A unified presentation of spacetime and internal gauge symmetry. (2015). arXiv:1502.06539

  49. Wallace, D.: Who’s afraid of coordinate systems? An essay on representation of spacetime structure. Stud. Hist. Philos. Sci. Part B. Forthcoming (2017)

  50. Weatherall, J.O.: The motion of a body in Newtonian theories. J. Math. Phys. 52(3), 032502 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  51. Weatherall, J.O.: Fiber bundles, Yang-Mills theory, and general relativity. Synthese 193(8), 2389–2425 (2016)

    Article  MathSciNet  Google Scholar 

  52. Weatherall, J.O.: Conservation, inertia, and spacetime geometry. Stud. Hist. Philos. Sci. Part B. Forthcoming (2017)

  53. Weyl, H.: Mathematische Analyse des Raumproblems. Verlag von Julius Springer, Berlin (1923)

    Book  Google Scholar 

Download references

Acknowledgements

I’m very grateful to the participants in the conference “Thinking About Space and Time” (University of Bern) for their comments and questions, and to James Read and two anonymous referees for comments on earlier drafts.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Neil Dewar.

Additional information

Publisher's Note

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

Covariance, Unvariance, and Transformations

Covariance, Unvariance, and Transformations

[40, Appendix A] seek to show that “minimally coupled dynamical equations in GR manifest local Poincaré symmetry, when written in normal coordinates at any \(p \in M\).” Here, I critically review their proof.

Read et al. begin by assuming that any minimally coupled dynamical equation in GR is of the form

$$\begin{aligned} O_1 + O_2 + \cdots + O_m = 0 \end{aligned}$$
(18)

where each \(O_i\) is either:

  • a tensor;

  • a partial derivative of a tensor; or

  • a partial derivative of a connection coefficient.

The reason to exclude (undifferentiated) connection coefficients—according to Read et al.—is that we are assuming the equation (18) is written in normal coordinates, in which (at the point p under consideration) connection coefficients vanish. However, this already risks introducing confusion. There is, I claim, a significant difference between an equation involving a partial derivative, and an equation involving a covariant derivative whose connection coefficients happen to be zero, even though the formal expression of these two equations will be the same. Specifically, two such equations will involve different transformation rules, and hence will have different invariance properties.

To see this, consider the two equations

$$\begin{aligned} \nabla _\mu J^\nu = 0 \end{aligned}$$
(19)

and

$$\begin{aligned} \partial _\mu J^\nu = 0 \end{aligned}$$
(20)

and suppose that we are in a flat space to which our coordinates are adapted, such that \(\nabla _\mu J^\nu = \partial _\mu J^\nu \) (since \(\Gamma ^\rho _{\mu \nu } = 0\)). It follows that the two equations pick out exactly the same class of solutions. However, if we apply a coordinate transformation

$$\begin{aligned} x^\mu \mapsto \widetilde{x}^\mu \end{aligned}$$
(21)

then we transform \(\nabla _\mu J^\nu \) as a rank-(1, 1) tensor, but transform \(\partial _\mu J^\nu \) as the partial derivative of a components of a vector. This means that our two equations are transformed into

$$\begin{aligned} \frac{\partial \widetilde{x}^\alpha }{\partial x^\mu } \frac{\partial x^\nu }{\partial \widetilde{x}^\beta } \widetilde{\nabla }_\alpha \widetilde{J}^\beta = 0 \end{aligned}$$
(22)

and

$$\begin{aligned} \frac{\partial \widetilde{x}^\alpha }{\partial x^\mu } \widetilde{\partial }_\alpha \left( \frac{\partial x^\nu }{\partial \widetilde{x}^\beta } \widetilde{J}^\beta \right) = 0 \end{aligned}$$
(23)

respectively. The first of these is equivalent to \(\widetilde{\nabla }_\alpha \widetilde{J}^\beta = 0\) (in the sense of having the same solutions), and so equation (19) is invariant under the coordinate transformation; but the second is not equivalent to \(\widetilde{\partial }_\alpha \widetilde{J}^\beta = 0\), and so equation (20) is not invariant under the coordinate transformation. Another way of seeing what’s going on here is to observe that equation (19) is more fully expressed as

$$\begin{aligned} \partial _\mu J^\nu + \Gamma ^\nu _{\mu \rho } = 0 \end{aligned}$$
(24)

and although \(\Gamma ^\nu _{\mu \rho }\) takes the value zero, its transformation rule means that the coordinate transformation (21) will (in general) transform it away from zero—in just such a way, of course, as to cancel out the extra terms arising from the transformation of \(\partial _\mu J^\nu \). Equation (20), on the other hand, has no connection coefficients figuring at all (whether zero-valued or not); so such coefficients cannot step out from the shadows to guarantee invariance, in the way they do for equation (19).

What this means is that we should include connection coefficients on the list of possible ingredients for our equation: although such coefficients might be zero in the coordinate system we start with, if we are assessing which coordinate transformations preserve the form of the equations, we need to check that they will preserve the vanishing of those connection coefficients! Fortunately, adding them to the list of ingredients doesn’t make a significant difference to Read et al.’s next observation: that for affine coordinate transformations, all the ingredients transform tensorially. Recall that an affine coordinate transformation is a transformation of the form

(25)

where and \(a^\mu \) are constant. Note that

(26)

and

(27)

where is the inverse to (i.e. is the matrix such that ). As is well-known, connection coefficients transform tensorially under affine coordinate transformations (since the non-tensorial part of the transformation rule features a second partial derivative). In the interests of space, I do not reproduce Read et al.’s proof that partial derivatives of tensors or partial derivatives of connection coefficients transform tensorially under affine transformations.

However, they then proceed to say

We have found that each of the \(O_i\) featuring in any minimally coupled dynamical equation in GR, written in normal coordinates at a point \(p \in M\), is covariant—i.e., transforms tensorially—under affine coordinate transformations. However, we have yet to show that all such equations are invariant—i.e. take the same form—under affine coordinate transformations. In fact, this is in general not the case.

Prima facie, this claim is somewhat surprising. For consider again the expression (18). In order for the left-hand-side to be well-formed, each \(O_i\) must have the same index structure: i.e., they must have the same free covariant and contravariant indices (where a “free index” is one that has not been contracted with another index). But if two terms have the same index structure, then when they are transformed tensorially, they will pick up the same partial derivative terms; owing to the linearity of tensor calculus these terms can then be uniformly multiplied away, as we did in observing that equation (22) is equivalent to \(\widetilde{\nabla }_\alpha \widetilde{J}^\beta = 0\). And note that the presence of bound indices (those which have been contracted) doesn’t make any difference: if we have an expression of the form

(28)

then applying the tensorial transformation rule yields

(29)

and so such indices “cancel out”.

Read et al. argue that this doesn’t hold, in general, “due to the potential contraction of indices in some terms with respect to the metric” (p. 11). As an example, they give (in my notation)

$$\begin{aligned} \partial _\nu F^{\mu \nu } = J^\mu \end{aligned}$$
(30)

They argue that the affine transformation (25) transforms this (again, using my notation) into

(31)

and that this latter equation is only equivalent to (30) if ; i.e., if is a Lorentz transformation (and hence, (25) a Poincaré transformation).

Now, this last assertion is correct in the sense that the right-hand equality in (31) is, indeed, only equivalent to (30) if is a Lorentz transformation. But the left-hand equality also only holds if is a Lorentz transformation: that’s the only way to use \(\eta \) to raise or lower indices and convert into .Footnote 49 And if we look at the leftmost term in equation (31), we observe that—just as our general discussion of contracted indices would lead us to expect—we have a matrix term and its inverse ; cancelling these out, we see that (30) transforms into

(32)

and (32) is equivalent to (30). Thus, the supposed counterexample is invariant under arbitrary affine transformations (not just Poincaré transformations).

Indeed, it seems to me that we have good grounds to expect equations formed from minimal coupling to be invariant under arbitrary coordinate transformations (not just affine transformations)—i.e. to be generally covariant. The reason why the above proof was limited to affine transformations is that, in general, partial derivatives and connection coefficients will transform non-tensorially, and hence we will get “extra” terms showing up in the transformed equation—terms which will prevent the transformed equation from being equivalent to the original. But if partial derivatives and connection coefficients show up together, then it may be that the extra terms from the one cancel out the extra terms from the other, and we do get invariance under (non-affine) coordinate transformations.

The condition under which such cancellations happen is, of course, that the partial derivatives and connection coefficients in the equation are such as to form a covariant derivative—indeed, the whole point of covariant differentiation is that the result of applying a covariant derivative to a tensor is, itself, another tensor (as reflected in the transformation (22)). But now consider Read et al.’s definition of minimal coupling (pp. 2–3):

Minimally coupled dynamical equations for matter fields in GR are constructed from dynamical equations for matter fields featuring coupling to a fixed Minkowski metric field \(\eta _{ab}\) and no curvature terms, by replacing all instances of \(\eta _{ab}\) with a generic Lorentzian metric field \(g_{ab}\), and replacing all instances of the torsion-free derivative operator compatible with \(\eta _{ab}\) with the torsion-free derivative operator compatible with \(g_{ab}\).

Thus, on the face of it, one would expect a minimally coupled dynamical equation to consist of tensors and covariant derivatives of tensors (with respect to the torsion-free derivative operator compatible with \(g_{ab}\))—that is, tensors and tensors. And clearly, if all the terms \(O_i\) in (18) are tensors (with the same index structure), then (18) will be invariant under arbitrary coordinate transformations.

The above, I claim, is the standard argument for the claim that generally covariant equations—including those obtained through minimal coupling—to be invariant under arbitrary (smooth) coordinate transformations. However, there is a different analysis one can give.Footnote 50 Suppose that our dynamical equation features the metric; very schematically, we give it the form

$$\begin{aligned} \cdots g_{\mu \nu } \cdots = 0 \end{aligned}$$
(33)

Now, if we have written this equation in normal coordinates, then \(g_{\mu \nu } = \eta _{\mu \nu }\), where \(\eta _{\mu \nu }\) denotes the matrix of coefficients (6). What is it for this expression to “retain the same form” under a coordinate transformation, from \(x^\mu \) to \(\widetilde{x}^\mu \)? First answer: it is for it to be (or to be equivalent to) an equation with the same syntactic structure, albeit with tildes over everything; schematically, the form is preserved if (33) is transformed into something (equivalent to)

$$\begin{aligned} \widetilde{\cdots } \widetilde{g}_{\mu \nu } \widetilde{\cdots } = 0 \end{aligned}$$
(34)

(A non-schematic example is given by the comparison of (30) with (32).) This first answer is the answer that the argument above assumed, and so this is the sense in which its conclusion holds.

Second answer: it is for it to have the same syntactic structure, and for certain simplifying identities to continue to hold. In the case at hand, this will mean that the form is preserved if (33) is transformed into something equivalent to (34), and—in addition\(\widetilde{g}_{\mu \nu } = \eta _{\mu \nu }\). In defence of this answer, one can argue that part of what makes the normal-coordinate form the “simplest” form is that writing the equation out in components would be considerably simpler if the metric diagonalises than if it does not; and it is in this sense that a transformation away from normal coordinates makes the equation into one with a less simple form, and so ipso facto one with a different form.Footnote 51

Evidently, if we require this second (stronger) sense of invariance, then any equation featuring the metric will be invariant only under coordinate transformations which are Poincaré in form, i.e. where

$$\begin{aligned} \frac{\partial \widetilde{x}^\alpha }{\partial x^\mu } \frac{\partial \widetilde{x}^\beta }{\partial x^\nu } \eta _{\alpha \beta } = \eta _{\mu \nu } \,, \end{aligned}$$
(35)

since it is only these equations which preserve the metric’s being diagonal.

Thus, the question becomes which of these two senses is more appropriate. The problem with the latter sense is that it risks overgenerating, and will lead us to underestimate the size of a theory’s symmetry group. In the context of electromagnetic theory on curved spacetime, for example, there will exist at any point Riemann normal coordinates with respect to which \(J^\mu = (\rho , 0, 0, 0)\). Adopting those coordinates will simplify the dynamical equations, just as the adoption of Riemann normal coordinates does; and such equations will be preserved only under spatial translations and rotations. This suggests that on the stronger sense of invariance just canvassed, the equations of electromagnetism turn out to be invariant only under the Newton group of transformations. This seems to me a reason to prefer the former, weaker, notion of invariance.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dewar, N. General-Relativistic Covariance. Found Phys 50, 294–318 (2020). https://doi.org/10.1007/s10701-019-00256-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-019-00256-0

Keywords

Navigation