On the reduction of general relativity to Newtonian gravitation

https://doi.org/10.1016/j.shpsb.2019.04.005Get rights and content

Highlights

  • Clarifies the physical interpretation of the geometrical relationship between general relativity and Newtonian gravitation.

  • Expounds this relationship as one of intertheoretic reduction.

  • Introduces topological structure on spacetime models to encode similarity of spacetime observables for this reduction.

  • Provides several concrete examples of Newtonian limits parameterized by distinct observables.

Abstract

Intertheoretic reduction in physics aspires to be both to be explanatory and perfectly general: it endeavors to explain why an older, simpler theory continues to be as successful as it is in terms of a newer, more sophisticated theory, and it aims to relate or otherwise account for as many features of the two theories as possible. Despite often being introduced as straightforward cases of intertheoretic reduction, candidate accounts of the reduction of general relativity to Newtonian gravitation have either been insufficiently general or rigorous, or have not clearly been able to explain the empirical success of Newtonian gravitation. Building on work by Ehlers and others, I propose a different account of the reduction relation that is perfectly general and meets the explanatory demand one would make of it. In doing so, I highlight the role that a topology on the collection of all spacetimes plays in defining the relation, and how the selection of the topology corresponds with broader or narrower classes of observables that one demands be well-approximated in the limit.

Introduction

In physics the concept of reduction is often used to describe how features of one theory can be approximated by those of another under specific circumstances. In such circumstances physicists say the former theory reduces to the latter, and often the reduction will induce a simplification of the features in question. (By contrast, the standard terminology in philosophy is to say that the less encompassing, approximating theory reduces the more encompassing theory being approximated.) Accounts of reductive relationships aspire to generality, as broader accounts provide a more systematic understanding of the relationships between theories and which of their features are relevant under which circumstances.

Thus reduction is naturally taken to be physically explanatory. Reduction can be more explanatory in other ways as well: sometimes the simpler theory is an older, predecessor of the theory being reduced. These latter “theories emeriti” are retained for their simplicity—especially regarding prediction, and sometimes understanding and explanation—and other pragmatic virtues despite being acknowledged as incorrect.1 Under the right circumstances, one incurs sufficiently little error in using the older theory; it is empirically adequate within desired bounds and domain of application. Reduction explains why these older, admittedly false (!) theories can still be enormously successful and, indeed, explanatory themselves.

The sort of explanation that reduction involves is similar to that which Weatherall (2011) identifies in the answer to the question of why inertial and gravitational mass are the same in Newtonian gravitation. In that case, the explanandum is a successfully applied fact about Newtonian systems. The explanans is a deductive argument invoking and comparing Newtonian theory and general relativity, showing that this fact is to be expected if the latter is true. In the case of reduction, the explanandum is rather the relation of non-identity between the predictions and explanations of two successful theories. The explanans is also a deductive argument invoking both theories, but one showing that if the reducing theory is adequate, then the theory it reduces to is adequate as well, within desired bounds and domain of application.

Despite the philosophical and scientific importance of reduction, it is astounding that so few reductive relationship in physics are understood with any detail. It is often stated that relativity theory reduces to Newtonian theory, but supposed demonstrations of this fact are almost always narrowly focused, not coming close to the level of generality to which an account of reduction aspires, which is to describe the relationship between the collections of all possibilities each theories allows. Indeed, much discussion of the Newtonian limit of relativity theory (e.g., Batterman (1995)) has focused on the “low velocity limit.” For example, the relativistic formula for the magnitude of the three-momentum, p, of a particle of mass m becomes the classical formula in the limit as its speed v (as measured in some fixed frame) becomes small compared with the speed of light c:p=mv1(v/c)2=mv(1+12(v/c)2+38(v/c)4+516(v/c)6+)v/c1mv.Similar formulas may be produced for other point quantities.

Another class of narrow demonstrations, so-called “linearized gravity,” concerns gravitation sufficiently far from an isolated source. In a static, spherically symmetric (and asymptotically flat) relativistic spacetime (M,gab), one can writegab=ηab+γab,where ηab is the Minkowski metric. If the components of γab are “small” relative to a fixed global reference frame, then in terms of Minkowski coordinates (t,x,y,z),gab1+2ϕdatdbt12ϕdaxdbx+daydby+dazdbz,Tabρdatdbt,where Tab is the stress-energy tensor, and the field ϕ satisfies Poisson's equation,2ϕ=4πρ.(Here 2 is the spatial Laplacian.) Accordingly, the trajectories of massive test particles far from the source will in this reference frame approximately follow trajectories expected from taking ϕ as a real Newtonian gravitational potential. In practice one often has a rough and ready handle on how to apply these approximations, but it is difficult to make all the mathematical details precise.2

It is, however, not often explicitly recognized that even the collection of all point quantity formulas, such as eq. (1), or the linearized expression for the metric (eq. (3)) along with trajectories of test particles far from an isolated massive body, together constitute only a small fragment of relativity theory. In particular, they say nothing about the nature of gravitation in other circumstances, or exactly how the connection between matter, energy, and spacetime geometry differs between relativistic and classical spacetimes.3 Insofar as one is interested in the general account of the reduction of one theory to another, these particular limit relations and series expansions cannot be understood to be a reduction of relativity theory to classical physics in any strict sense. Even the operationally minded would be interested in an account of how arbitrary relativistic observables can be approximated by their Newtonian counterparts.

The development of Post-Newtonian (PN) theory, which is very general method for applying the “slow-motion” limit to a variety of formulas like equation (1) ameliorates this concern somewhat (Poisson and Will, 2014). Its goal is to facilitate complex general relativistic calculations in terms of quantities that are (in principle) measurable by experiments, although it typically only applies as a good approximation to a small region of a relativistic spacetime. Indeed, while it seems in many circumstances to yield helpful predictions, there is no guarantee that an arbitrary PN expansion actually converges. Part of the reason for this is that the theory typically relies heavily on particular coordinate systems for these small regions that may not always have felicitous properties. In a word, while enormously useful, the PN theory by itself does not constitute a general method for explaining the success of Newtonian theory.

One way to do so—to organize relatively succinctly the relationships between arbitrary relativistic observables and their classical counterparts—is to provide a sense in which the relativistic spacetimes themselves, and fields defined on them, reduce to classical spacetimes (and their counterpart fields), as all spacetime observables are defined in these terms. This geometrical way of understanding the Newtonian limit of relativity theory has been recognized virtually since the former's beginning (Minkowski, 1952, pp. 73–91), where it was observed that as the light cones of spacetime flatten out in Minkowski spacetime, hyperboloids of constant coordinate time become hyperplanes in the limit. This geometric account has since been developed further by Ehlers, 1981, Ehlers, 1988, Ehlers, 1991, Ehlers, 1997, Ehlers, 1998, Ehlers et al., 1986, Minkowski et al., 1952 and others (e.g., Künzle (1976); Malament (1986a, b)) for general, curved spacetimes.4

But the nature and interpretation of this limit has sat uneasily with many. The image of the widening light cones seems to suggest an interpretation in which the speed of light c. In one of the first discussions of this limiting type of reduction relation in the philosophical literature, Nickles (1973) considered this interpretation and pointed out some of its conceptual problems: what is the significance of letting a constant vary, and how is such variation to be interpreted physically? Rohrlich (1989) suggested that “c” can only be interpreted counterfactually—really, counterlegally, since it corresponds literally with a sequence of relativistic worlds whose speeds of light grow without bound. Thus interpreted, the limit only serves to connect the mathematical structure of relativistic spacetime with that of classical spacetime (Rosaler, 2019). It cannot explain the success of Newtonian physics, since such an explanation “specifies what quantity is to be neglected relative to what other quantity” (Rohrlich, 1989, p. 1165) in worlds with the same laws to determine the relative accuracy of classical formulas as approximations to the outcomes of observations.

So, we are left with a conundrum: existing approaches to the reduction of general relativity to Newtonian gravitation are explanatory, or they are general; moreover, there may be some concerns about the mathematical rigor of some approaches. The received view about these three approaches is summarized in Table 1.

The purpose of this essay is to develop and provide an interpretation of the geometrical limiting account that could in fact meet the explanatory demand required of it. It is thus both perfectly general and explanatory.5 To explicate how it works, I first present a unified framework for classical and relativistic spacetimes in §2. The models of each are instantiations of a more general “frame theory” that makes explicit the conceptual and technical continuity between the two. To define the limit of a sequence of spacetimes, however, one needs more structure than just the collection of all spacetime models. As I point out in §3, one way to obtain this structure is by putting a topology on this collection. The key interpretative move, as explicated in §4, exploits the freedom of each spacetime to represent many different physical situations through the representation of different physical magnitudes.6

To illustrate this move I describe three classes of examples, involving Minkowski, Schwarzschild, and cosmological (FLRW) spacetimes, that I also show have Newtonian limits in the sense defined here, with respect to a certain topology. (It still remains an open question—though I conjecture it to be true—whether every Newtonian spacetime is an appropriate limit of relativistic spacetimes.)

The topology, though, can make a significant difference to the evaluation of a potential reduction. In particular, whether the Newtonian limit of a particular sequence of spacetimes exists at all can depend on it, and it is not determined automatically from the spacetime theories themselves. I indicate in §5 that certain topologies correspond with a certain the classes of observables that one demands must converge in order for a sequence of spacetimes to converge. Requiring that more observables converge in the limit leads in general to more stringent convergence criteria, thus a finer topology (i.e., with more open sets). In light of this, I argue for a slightly more stringent criterion than has been used by other authors so that observables depending on compact sets of spacetime, such as the proper times along (bounded segments of) timelike curves, also converge. Finally, in §6, I summarize what topical and methodological conclusions I think can be drawn about these results for gravitation (including PN theory) and philosophy of science, and indicate how they might be applied more generally to other reduction relations of a similar limiting type.7

Section snippets

Ehlers's frame theory

The conceptual unification of relativistic and classical spacetimes that the frame theory affords requires some preliminaries. In particular, one must recast the empirical content of Newtonian gravitation, a theory of flat spacetime with a gravitational potential, in the language of Newton-Cartan theory, which describes gravitation through a curved connection.8

The Newtonian Limit

Before developing how limits can be interpreted in this approach, one must first determine what, mathematically, a limit of a family of spacetimes is supposed to be in the first place. One way to formalize this is to put a topology on the models of the frame theory, or at least the subclass of those models that represents relativistic and classical spacetimes. In particular, if O0 and {Oλ}λ(0,1] are models of the frame theory, then limλ0Oλ=O0 if and only if for any open neighborhood N of O0,

Interpretation of the Newtonian Limit

As mathematical objects, the models of the frame theory are completely well-defined, and it is a mathematical matter whether a particular family, parameterized by λ, has a Newtonian limit. According to Proposition 3.1, the existence of such a limit as λ0 means that, given ε>0 any and any finite set of spacetime points, the values in any basis of the components of the temporal and spatial metrics, connection, and stress-energy (and their partial derivatives to second order) of members of the

Topology and observables

We can now return to the question I posed after the definition of the Newtonian limit: why use the point-open topology? Since there is no canonical topology for the spacetime metrics (Fletcher, 2014, 2016), it must be justified relative to the nature of the investigation. In light of the foregoing discussion of the legal interpretation of the limit, it is clear that a topology is equivalent with a set of relevant observables that one requires be well-approximated by the Newtonian limit

Conclusions

One can draw a number of topical and methodological conclusions from the above discussion. In §2, I described how one can give a unified description of both general relativity and Newton-Cartan theory under the banner of Ehlers's frame theory. It is only superficially an example of a unification in the traditional sense in philosophy of science, for the theories thereby “unified” do not have different domains but in fact concern the same phenomena: one (general relativity) is a successor to the

Acknowledgements

Thanks to Jim Weatherall and the anonymous referees for detailed comments, and to the audience members in Munich, London, Oxford, Budapest, Cologne, Rotterdam, Belgrade, and Boulder for illuminating discussion, especially Harvey Brown, Jeremy Butterfield, Roman Frigg, Eleanor Knox, Oliver Pooley, Miklós Rédei, David Wallace, and Charlotte Werndl. David Malament has my gratitude for providing invaluable copies of the most difficult-to-find of Ehlers's papers. Part of the research leading to the

References (42)

  • J. Ehlers

    On limit relations between, and approximative explanations of, physical theories

  • D.B. Malament

    Gravity and spatial geometry

  • J. Rosaler

    Local reduction in physics

    Studies in History and Philosophy of Modern Physics

    (2015)
  • R.W. Batterman

    Theories between theories: Asymptotic limiting intertheoretic relations

    Synthese

    (1995)
  • G. Belot

    Whose devil? Which details?

    Philosophy of Science

    (2005)
  • É. Cartan

    Sur les variétés a connexion affine et la théorie de la relativité généralisée

    Annales Scientifiques de l’École Normale Superieur

    (1923)
  • É. Cartan

    Sur les variétés a connexion affine et la théorie de la relativité généralisée

    Annales Scientifiques de l’École Normale Superieur

    (1924)
  • J. Ehlers

    Über den Newtonschen Grenzwert der Einsteinschen Gravitationstheorie

  • J. Ehlers

    The Newtonian limit of general relativity

    (1988)
  • J. Ehlers

    The Newtonian limit of general relativity

  • J. Ehlers

    Examples of Newtonian limits of relativistic spacetimes

    Classical and Quantum Gravity

    (1997)
  • J. Ehlers

    The Newtonian limit of general relativity

  • S.C. Fletcher

    Similarity and spacetime: Studies in intertheoretic reduction and physical significance

    (2014)
  • S.C. Fletcher

    Similarity, topology, and physical significance in relativity theory

    The British Journal for the Philosophy of Science

    (2016)
  • S.C. Fletcher

    Global spacetime similarity

    Journal of Mathematical Physics

    (2018)
  • S.C. Fletcher

    On representational capacities, with an application to general relativity

    Foundations of Physics

    (2019)
  • K. Friedrichs

    Eine invariante Formulierung des Newtonschen Gravitationsgesetzes und des Grenzübergangs vom Einsteinschen zum Newtonschen Gesetz

    Mathematische Annalen

    (1927)
  • M. Golubitsky et al.

    Stable mappings and their singularities

    (1973)
  • H.-P. Künzle

    Covariant Newtonian limit of Lorentz space-times

    General Relativity and Gravitation

    (1976)
  • D. Lehmkuhl

    Introduction: Towards a theory of spacetime theories

  • D.B. Malament

    Newtonian gravity, limits, and the geometry of space

  • View full text