Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-28T13:21:21.982Z Has data issue: false hasContentIssue false

Spacetime Conventionalism Revisited

Published online by Cambridge University Press:  11 August 2023

Ufuk I. Tasdan*
Affiliation:
Department of Philosophy, University of Bristol, Bristol, BS8 1TH, United Kingdom
Karim P. Y. Thébault
Affiliation:
Department of Philosophy, University of Bristol, Bristol, BS8 1TH, United Kingdom
*
Corresponding author: Ufuk I. Tasdan; Email: iu.tasdan@bristol.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

We provide five rearticulations of the thesis that the structure of spacetime is conventional, rather than empirically determined, based upon variation of the structures that are empirically underdetermined and modal contexts in which this underdetermination occurs. Three of the five formulations of conventionalism will be found to fail. Two are found to open up new interesting problems for researchers in the foundations of general relativity. In all five cases, our analysis explores the interplay between geometric identities, symmetry, conformal structure, and the dynamical content of physical theories with the conventionalism dialectic deployed as a tool of explication, clarification, and exploration.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of the Philosophy of Science Association

1. Conventionalism about what?

1.1. Geometric conventionalism and universal forces

Geometric conventionalism is a form of underdetermination thesis regarding the space or spacetime structure of the physical theory. Footnote 1 One may demarcate varieties of geometric conventionalism on the basis of the answers given to the following two questions:

  1. (1) Which geometric statements regarding a model of space (or spacetime) are not empirically determined truths?

  2. (2) Which physical differences arise in the comparison of different conventions about such statements?

Reichenbach (Reference Reichenbach1928) famously responded to the first question that statements regarding the spatial or spacetime metric tensor of any given model are not empirically determined; and, to the second question that the physical difference between two metrics (and thus two models) would manifest as an apparent universal force. In particular, Reichenbach (Reference Reichenbach1928), like Poincaré (Reference Poincaré1902) in the disk experiment, compares the measuring rod of observers living in two different worlds having flat spacetime. In this thought experiment, inhabitants of one of the worlds have rulers under the effect of heat and perceive the spacetime as curved. However, “(h)eat as a force can thus be demonstrated directly” (13) because heat affects different materials differently. As a result, the geometry of spacetime is underdetermined if the heat is taken as a universal force: “Thus the only distinguishing characteristic of a field of heat is the fact that it causes different effects on different materials. But we could very well imagine that the coefficients of heat expansion of all materials might be equal then no difference would exist between a field of heat and the geometry of space” (Reichenbach, Reference Reichenbach1928, 26).

What forms of conventionalism might we consider beyond the specific form articulated in terms of universal forces? As just noted, in his original argument Reichenbach took metrical relations of spacetime to be conventional and made the specific assumption that the choice in convention could be underdetermined by introducing universal forces. Significantly, Dieks (Reference Dieks1987) argues that it is in the Reichenbachian spirit to interpret gravitation as a universal “force,” since any effect of the curvature of spacetime can be explained by a suitable gravitational effect. Indeed, in Reichenbach (Reference Reichenbach1928) we find the following text: “[The] universal effect of gravitation on all kinds of measuring instruments defines therefore a single geometry. In this respect we may say that gravitation is geometrized” (Reichenbach, Reference Reichenbach1928, 256; italics in original).

An explicit defence of this more liberal “universal effect” reading can be found in the later work of Carnap (Reference Carnap1966, 169): “Reichenbach called them ‘forces’, but it is preferable here to speak […] in a more general way, as two kinds of ‘effects’. (Forces can be introduced later to explain the effects.)” Further discussion of this point in the context of the interpretation of Reichenbach can be found in Sexl (Reference Sexl1970) and Duerr and Ben-Menahem (Reference Duerr and Ben-Menahem2022, section 2.4.2).

In what follows we will draw inspiration from these ideas for reframing of the debate via generalized forms of the conventionalist thesis that do not rely on force-based representations. In particular, we will consider different formulations of the spacetime conventionalist thesis in terms of physical differences that are taken to result from different proposals for the relevant universal effect.

1.2 Rearticulating spacetime conventionalism

We have just seen that one is able to historically situate a mildly generalized form of geometric conventionalism based upon universal effects rather than universal forces. For a foundational, as opposed to primarily historical, perspective the next step is then to consider the salience of wider forms of conventionalist thesis. Footnote 2 To do this productively, however, one must place some constraints upon the relevant argument pattern. With this in mind, we will consider the following generalized forms of our two questions:

  1. (1) Which statements regarding the structure of a spacetime theory are empirically underdetermined?

  2. (2) Which physical differences arise in the comparison of different conventions about such structure?

Setting the two questions up in this way allows us to confine ourselves within a generalized spacetime conventionalist argument pattern whilst allowing for rearticulation along various dimensions.

The first dimension of rearticulation we will pursue involves considering various different ways of formulating the relevant physical differences in (2). The second dimension of rearticulation is to vary the structure in question in (1). Such structure should, of course, still be physically well-motivated as the basic structure of a spacetime theory, but need not necessarily be a geometric structure. Third, and finally, we can identify a more subtle dimension of rearticulation in terms of the modal context of the empirical underdetermination thesis. The idea is that the context in which one is answering questions (1) and (2) is itself underdetermined to the degree that it might be specified as obtaining between: i) two observers within a single model of a theory; ii) two models within the same theory; iii) two models of different theories; or iv) two ways of interpreting the sets of models within a single theory. What is essential within our family of generalized notions of spacetime conventionalism is that in each and every case it is required that a basic structure of a spacetime theory is empirically underdetermined, and this underdetermination leads to the possibility for physical differences to arise between conventions regarding how to break the underdetermination.

In what follows we will tour a selection of sights of interest in the conceptual space defined along these three dimensions of rearticulation. Some of the pathways towards spacetime conventionalism that we will explore will take us rather far from the original debates, and others will ultimately be found to lead us into a cul-de-sac. We take this journey to be a valuable and productive one for two principal reasons.

First, in what follows, we will deploy our generalized framework for varieties of spacetime conventionalism as a tool of foundational analysis towards both clarification of old conceptual problems, or indeed pseudo-problems, and the identification of new problems. Articulation of new conceptual problems can, in turn, be understood as a constitutive element of scientific progress according to the framework articulated by Laudan (Reference Laudan1977), cf. Feyerabend (Reference Feyerabend1981). In this sense, our ultimate aim is to apply conventionalism as a lens to advance the scientific understanding of spacetime theory.

Second, we take our project to provide important clarificatory insights into the formal interplay between geometric identities, symmetry, conformal structure, and dynamical equations within the foundations of general relativity. The five forms of conventionalism articulated in what follows are woven together by twin formal threads of conformal structure and the Bianchi identities. Each of the forms of conventionalism we will consider, to at least some degree, trades on a package of rereading the interconnection between parts of the formalism of general relativity based on one or both of these aspects. We will consider putative conventionalisms relating to non-conformal structure, local conformal structure, and to the decomposition of the Einstein equation into conformally invariant and non-conformally invariant parts. Further, we will consider putative conventionalisms based upon understanding the role of the Bianchi identities as formal geometric identities, conservation equations, and dynamical field equations.

We start, in section 2, by considering a formulation of geometric conventionalism due to Weatherall and Manchak (Reference Weatherall and Manchak2014). We consider the implications of the no-go result proved by the authors in the context of a general framework for the analysis of the methodological implications of no-go theorems due to Dardashti (Reference Dardashti2021). We will find that this analysis of the Weatherall–Manchak no-go result further motivates the program for rearticulation along a different dimension that we have just set out.

In section 3 we present our first attempt at rearticulation of conventionalism in which we focus on inertial structure and putative universal effects that result from differing representations of local stress-energy 4-current by coincident observers following different world-lines. Ultimately, the view is found to be unconvincing, since there are good reasons to reject the physical salience of the energy 4-current and thus this prima facie plausible form of conventionalism can be dissolved. In section 4 we then consider the conformal structure and the universal effect of tidal deformation. We argue that we can dissolve a further potential source of conventionalism with regard to the deforming and non-deforming aspects of tidal effects by considering the invariant decomposition of the Riemann tensor into Weyl and Ricci parts. In each case, although the new forms of conventionalism are found to fail, our discussion serves to elucidate new perspectives on the representational roles of different objects within the foundations of general relativity.

In section 5, we pull our focus back to a wider form of interpretational question regarding the dynamical understanding of the Weyl tensor fields and the role of the Bianchi identities. In particular, we consider a putative underdetermination with regard to the nomic structure of general relativity. This analysis serves to open up the possibility for a very different form of conventionalism in terms of how to isolate the dynamical content of general relativity. In particular, we argue that underdetermination with regard to the structure that specifies which relativistic spacetimes are dynamically possible models leads to conventionalism with regard to which fields are understood as dynamical and which non-dynamical. Furthermore, we find that consideration of the initial value problem of general relativity serves to further establish, rather than dissolve, the putative underdetermination with regard to nomic structure that is at the heart of Spacetime Conventionalism 4.

Finally, in section 6, we consider a proposal based upon the decomposition of the Einstein tensor into conformally invariant and non-conformally invariant parts. The potential for underdetermination of this decomposition is then taken to be the hallmark of a novel form of conventionalism whose validity is expressible in terms of (the failure of) a precise mathematical “Bach conjecture,” whose truth is as yet unknown.

2. Non-conformal structure and universal forces

Spacetime Conventionalism 1

  1. (1) The non-conformal structure of relativistic spacetime is not an empirically determined truth.

  2. (2) Physical differences between conformally equivalent models of relativistic spacetime with regard to the existence of universal forces arise in the comparison of different conventions about affine structure.

Any pair of conformal equivalent relativistic spacetimes are, by definition, such that $\left( {M,{g_{\mu \nu }}} \right)$ and $\left( {M,{{\tilde g}_{\mu \nu }}} \right)$ where ${\tilde g_{\mu \nu }} = {{\rm{\Omega }}^2}{g_{\mu \nu }}$ . Each of these spacetimes is accompanied by a metric-compatible, torsion-free, Levi–Civita derivative operator, $\nabla $ and $\tilde \nabla $ respectively. Relativistic spacetime conventionalism can then be formulated as the thesis that the same curve, $\gamma $ , can be a geodesic relative to $\nabla $ but be associated with an acceleration relative to $\tilde \nabla $ given by a “universal force” induced by a tensor field ${G_{\mu \nu }}$ via the expression $G_{{\rm{\;}}\mu }^\nu {\tilde \xi ^\mu }$ where $\tilde \xi $ is the tangent field to $\gamma $ with unit length relative to ${\tilde g_{\mu \nu }}$ . So formulated, a relativistic conventionalist thesis, equivalent to Spacetime Conventionalism 1, has a contradiction as a deductive consequence (Weatherall and Manchak, Reference Weatherall and Manchak2014, Proposition 2).

The result of Weatherall and Manchak constitutes a “no-go” theorem for a specific form of spacetime conventionalism and, so understood, the theorem might be taken to close off one of the principal avenues for the articulation of such a view within the context of spacetime theory. It is worth noting in this context that recent analysis due to Duerr and Ben-Menahem (Reference Duerr and Ben-Menahem2022) has strongly contested whether the characterization of conventionalism within the antecedent conditions of the Weatherall and Manchak no-go result is historically or conceptually adequate. Whilst our perspective on spacetime conventionalism intersects in a number of respects with Duer and Ben-Menahem, our attitude to the work of Weatherall and Manchak is rather different. This difference can be understood by drawing upon the hugely insightful work on no-go theorems in physics due to Dardashti (Reference Dardashti2021).

Within Dardashti’s framework, a no-go result has been established if and only if an inconsistency arises between a derived consequence of a set of physical assumptions, $P$ , represented by a mathematical structure, $M$ , and a goal, $G$ , within a framework, $F$ . By contraposition, the implications of the no-go result can then be understood in terms of different “methodological pathways” based upon a disjunct of the form $\neg P \vee \neg M \vee \neg F \vee \neg G$ . Presuming that we do not want to give up the goal, we may then reinterpret the implication as a “go theorem” with the implication $\neg P \vee \neg M \vee \neg F$ . That is, if we would like to achieve a certain goal, then the theorem provides us with a formal basis to articulate the specific mathematical, physical, or framework assumptions one must reject to achieve it.

Returning to the Weatherall–Manchak result. Since the theorem contained in their Proposition 2 is both valid and non-trivial, we take there to be good cause to explore its implications as a “go theorem” in the context of the negation of the various physical, mathematical, and framework assumptions. Footnote 3 That is, we take the result to motivate “methodological pathways” in which we seek to articulate different forms of the generalized spacetime conventionalist goal based upon different physical, mathematical, and framework assumptions. Our specific approach is based upon articulating a set of such pathways in terms of the different dimensions of rearticulation of conventionalism as per the analysis of the introduction. Thus, our project is further motivated by a “go theorem” interpretation of the Weatherall–Manchak result. We start by considering a form of spacetime conventionalism founded upon inertial structure and an intra-model modal context.

3. Inertial structure and local stress–energy flux

Spacetime Conventionalism 2

  1. (1) The inertial structure of a particular relativistic spacetime is not an empirically determined truth.

  2. (2) Physical differences between coincident observers with regard to local stress–energy flux arise in the comparison of different conventions about which time-like vector field is chosen as an observer’s inertial frame.

The second form of the conventionality of spacetime conventionality we will consider relates to the interpretation of different observers of the same model, and thus the same geometric facts, in terms of different local dynamical facts. The idea is that within a spacetime there can be observers that are coincident and yet will disagree about the dynamical interpretation of the same local geometric facts. The formal property that we will isolate in the privileged class of observers who have the analogue of a vanishing universal force is that they are Killing observers and thus the answer to the first of our two conventionalism questions is inertial structure. The physical differences that feature in the second question are then with regard to local stress–energy flux.

We can better understand the motivation for Conventionalism 2 by briefly considering the problem of the local definition of conservation of energy in general relativity. Footnote 4 The problem can be straightforwardly stated as follows. For a vanishing cosmological constant, the Einstein equations take the familiar form

(1) $${G_{\mu \nu }} = {R_{\mu \nu }} - {1 \over 2}R{g_{\mu \nu }} = 8\pi {T_{\mu \nu }},$$

where ${T_{\mu \nu }}$ is the energy–momentum tensor associated with all matter fields and their interactions, and ${G_{\mu \nu }}$ is the Einstein tensor which is determined by the metric g ${_{\mu \nu }}$ via the Ricci curvature ${R_{\mu \nu }}$ and Ricci scalar $R$ .

The Levi–Civita derivative, ${\nabla _\mu }$ , can be explicitly constructed from the metric and an arbitrarily chosen derivative operator. Given such a derivative operator on a manifold $M$ , it can be proved that there exists a unique smooth tensor field on $M$ such that, for all smooth fields ${\xi ^\sigma }$ ,

(2) $$R_{\sigma \mu \nu }^\rho {\xi ^\sigma } = - 2{\nabla _{[\mu }}{\nabla _{\nu ]}}{\xi ^\rho }$$

(Malament, Reference Malament2012, Lemma 1.8.1). $R_{\sigma \mu \nu }^\rho $ is then the unique Riemann curvature tensor field associated with ${\nabla _\mu }$ . The Ricci curvature tensor is defined via the contraction of the Riemann curvature as

(3) $${R_{\sigma \mu }} = R_{\sigma \mu \rho }^\rho .$$

An important identity that holds for any Riemann tensor is the second Bianchi identity. For the torsion-free case, this can be expressed in terms of the Levi–Civita derivative as

(4) $${\nabla _{[\alpha }}R_{\left| \sigma \right|\mu \nu ]}^\rho = 0.$$

This equation follows from the basic mathematical properties of the Riemann tensor and is thus naturally understood in this context as a formal identity without physical content. Footnote 5 Interestingly, however, when combined with the definition of the Einstein tensor, equation (4) can be shown to imply Footnote 6

(5) $${\nabla _\mu }{G^{\mu \nu }} = {\nabla _\mu }{T^{\mu \nu }} = 0.$$

At this point, it is very tempting to read the equation ${\nabla _\mu }{T^{\mu \nu }} = 0$ as an expression of the local conservation of energy (or stress–energy) and it seems like we have derived a dynamical conservation equation from a formal identity. However, such a naïve way of thinking about energy conservation in general relativity immediately runs into trouble.

Consider an observer moving along the integral curve of an arbitrary time-like vector field ${\xi ^\mu }$ . Let us define the stress–energy 4-current along the direction defined by this vector as $j\left[ {{\xi ^\mu }} \right] = T_{{\rm{\;}}\nu }^\mu {\xi ^\nu }$ . This 4-current does not in general vanish, so we cannot be guaranteed that ${\nabla _\mu }j\left[ {{\xi ^\mu }} \right] = 0$ by the fact that ${\nabla _\mu }{T^{\mu \nu }} = 0$ . This might be interpreted to mean that we no longer have “strict” conservation of energy according to the curved spacetime version of Gauss’ law (Wald, Reference Wald1984, 69–70). Moreover, we might plausibly interpret $j\left[ {{\xi ^\mu }} \right]$ as the conserved current associated with the observers’ time translations and thus understand ${\nabla _\mu }j\left[ {{\xi ^\mu }} \right] = 0$ to imply a failure of conservation of locally measured energy. Next, recall that the class of Killing vector fields is given by the satisfaction of the Killing equation:

(6) $${\nabla _\mu }{\xi _\nu } + {\nabla _\nu }{\xi _\mu } = 0.$$

This is equivalent to the statement that the metric is invariant along the integral curves of ${\xi _\nu }$ or, in terms of the Lie derivative, that we have ${\mathfrak{L}_\xi }{g_{\mu \nu }} = 0$ . The existence of a time-like Killing vector field is then (in this sense) equivalent to the symmetry of time translation invariance holding along the relevant time-like curves. The existence of a time-like Killing vector field is then necessary and sufficient for the vanishing of the stress–energy 4-current. This, in turn, implies that, for the class of Killing observers whose worldlines are integral curves of the vector fields solving (6), the vanishing of the stress–energy 4-current and thus “strict” conservation of energy is guaranteed.

We can then formulate Spacetime Conventionalism 2 by interpreting the stress–energy flux for the non-Killing observers as playing the functional role of a Carnapian universal effect. Killing and non-Killing observers can be coincident within the same spacetime patch but will attribute different interpretations to the same geometric facts in terms of the presence or absence of this additional associated flux. Putative physical differences with regard to local stress–energy flux thus can be understood to arise in the comparison of different conventions about which timelike vector field is chosen as an observer’s rest frame. The existence or not of such effects depends upon a convention as to the choice of timelike vector field as a rest frame.

Three clear lines of response to this argument for Spacetime Conventionalism 2 are available. The first is simply to note that in the general class of Einstein spacetimes Killing observers are not guaranteed to exist at all. The existence or not of Killing vector fields depends upon there being non-trivial continuous isometries of the metric, and there are good reasons to believe that such transformations do not exist in general for physically realistic spacetimes. Footnote 7 In a generic spacetime we can therefore expect that there will not be any set of “freely falling” observers who measure locally vanishing stress–energy 4-current. Spacetime Conventionalism 2 would then be best understood as a viable but highly limited conventionalist position. Footnote 8

The second response relies upon the introduction of further structure into the theory. The first step is to note that the magnitude of the stress–energy 4-current depends upon the affine properties of the world-line of the observer. This means that the quantity given the flux into a region is still conventional in the sense that it will be different for different observers moving within the same spacetime patch irrespective of the existence or not of Killing vector fields. In this context, one could consider introduction of a gravitational-energy–momentum tensor ${t^{\mu \nu }}\left[ {{g_{\alpha \beta }},{\nabla _{\alpha \beta }}} \right]$ which is such that a total energy–momentum complex is conserved with respect to the flat metric’s torsion-free derivative operator (Pitts, Reference Pitts2010, Reference Pitts2016, Reference Pitts2021). This would eliminate any potential for conventionalism with regard to total energy conservation but at the cost of introducing extra background or auxiliary structures into the theory.

This brings us to our third ultimately more satisfactory response to Killing conventionalism: the rejection of the physical salience of the energy 4-current. The key observation is that it is far from clear that focusing on localized energy fluxes is germane to the context of relativistic physics; arguably the ambiguities we have encountered are a product of focusing on concepts and quantities that are a relic of the bygone, pre-relativistic era. Consider in particular the fact that the failure of the vanishing of the stress–energy 4-current can occur for non-Killing observers even in flat spacetimes. This surely indicates that we should not think of it as encoding the results of additional universal effect resulting in the failure of energy conservation. Rather, it motivates us to look for a completely general characterization of the causal basis of genuine universal gravitational effects due to curvature without the need for background structure: this is to reframe the discussion to focus on tidal forces and their explicit geometric representation, rather than energy conservation. To do this it will prove essential to consider the physical significance of conformal structure within general relativity.

4. Conformal structure and tidal deformation

Spacetime Conventionalism 3

  1. (1) The local conformal structure of a particular relativistic spacetime is not an empirically determined truth.

  2. (2) Physical differences between coincident observers with regard to mechanical stress due to tidal forces arise in the comparison of different conventions about the decomposition of the Riemann tensor into deforming and non-deforming parts.

Consider a smooth one-parameter family of geodesics. Define two vector fields: a timelike vector field ${\xi ^\mu }$ tangent to the family of geodesics and a second vector field ${\chi ^\nu }$ that represents the infinitesimal displacement to an infinitesimally nearby geodesic. For a given Riemann curvature tensor, the acceleration due to geodesic deviation is given by

(7) $${a^\mu } = - R_{\beta \nu \alpha }^\mu {\xi ^\alpha }{\xi ^\beta }{\chi ^\nu }.$$

In general terms, tidal “forces” should be understood as the effects of geodesic deviation as induced by Riemann curvature. There are two physically distinct senses in which a geodesic can undergo deviation. The first is through conformal rescaling, which would change local spatial scales but preserve local shape. The second is through deformation, which would preserve local spatial scales but change local shape.

Each effect can have non-trivial physical consequences. However, there are compelling physical reasons to expect the manifestations of the effects for local observers to be rather different. Since geodesic deviation through conformal rescaling does not change the shape of a material body it will not induce mechanical stress upon that body. Rather, physical manifestations will be indirect and contingent on the material content and relevant scales and coupling constants, for example, redshifts or thermal effects due to expansion. By contrast, geodesic deviation through deformation will change local shape and thus manifest as a “universal effect” which induces mechanical stress on a material body. We have direct empirical evidence that such stresses are able to overwhelm internal forces in realistic physical situations through the observed phenomena of spaghettification in which the deforming geodesic deviation induces an extreme form of tidal shear in the region outside a black hole event horizon that rips stars into thin strips.

Spacetime Conventionalism 3 is then the view that seeks to exploit underdetermination in the representation of tidal forces. The proposal would be that different coincident observers may decompose the consequences of the geodesic deviation given by equation (7) differently, and so adopt different conventions with regard to the degree of tidal deformation and associated mechanical stress. This would not only underdetermine the tidal “forces” in a physically significant sense but also underdetermine the local conformal structure of spacetime.

The proposed new formulation of spacetime conventionalism has, however, been set up to fail. It is precisely in the relevant respect that the Riemann curvature tensor has a unique decomposition in terms of its Ricci and Weyl parts. This decomposition equates to exactly the decomposition of geodesic deviation into deforming and non-deforming parts. We can see this as follows. The Weyl curvature tensor, ${C_{\rho \sigma \mu \nu }}$ , can be defined via the Riemann and Ricci curvature tensors as the traceless tensor

(8) $${C_{\rho \sigma \mu \nu }} = {R_{\rho \sigma \mu \nu }} - {1 \over 2}\left( {{g_{\rho [\nu }}{R_{\mu ]\sigma }} + {g_{\sigma [\mu }}{R_{\nu ]\rho }}} \right) - {R \over 6}\left( {{g_{\rho [\mu }}{g_{\nu ]\sigma }}} \right)$$

The tensorial nature of this expression of course means that the decomposition of Riemann curvature into Ricci and Weyl parts is invariant under the push-forward of diffeomorphisms $\phi $ of the spacetime manifold $M$ . That is, for any three tensors which are related such that $A = B + C$ we will have ${\phi ^*}\left( A \right) = {\phi ^*}\left( {B + C} \right) = {\phi ^*}B + {\phi ^*}C$ . Observers using different coordinate systems in a given spacetime will necessarily agree on the decomposition. Moreover, the decomposition of the Riemann curvature tensor into Ricci and Weyl components has its origin in the symmetry transformation properties of the Riemann tensor. In particular, we can explicitly derive the decomposition by considering the symmetric and antisymmetric parts of a decomposition of the Riemann tensor in terms of SO(3,1) irreducible tensors. Footnote 9 We thus find that there is a solid mathematical foundation for taking the distinction as observer-independent and intrinsic to the structure of a given spacetime.

A further decomposition of the Weyl tensor then allows an explicit and insightful connection to tidal deformation. Footnote 10 Let us first interpret the Weyl tensor as the free gravitational field, and the metric tensor as its (second-order) potential field, and consider a timelike unit vector field ${\xi ^\nu }{\xi _\nu } = - 1$ representing a family of observers. Next, parallel to the way one can split the electromagnetic field into electric and magnetic parts in the rest frame of ${\xi ^\nu }$ , we can split the Weyl curvature tensor, ${C_{\rho \sigma \mu \nu }}$ , into electric and magnetic parts constructed as symmetric traceless tensors orthogonal to ${\xi ^\nu }$ . The explicit expressions are

(9) $${E_{\rho \sigma }} = {C_{\rho \sigma \mu \nu }}{\xi ^\mu }{\xi ^\nu },\;\;\;\;\;\;\;\;{H_{\rho \sigma }} = {1 \over 2}{\varepsilon _{\rho \nu \alpha }}C_{\sigma \mu }^{\nu \alpha }{\xi ^\mu },$$

where ${\varepsilon _{\rho \sigma \mu }}$ is the effective volume element in the rest space of the comoving observer and is explicitly given by

(10) $${\varepsilon _{\rho \sigma \mu }} = \sqrt {\left| {{\rm{det\ g}}} \right|} \delta _{[\rho }^0\delta _\sigma ^1\delta _\mu ^2\delta _{\nu ]}^3{\xi ^\nu }$$

(Goswami and Ellis, Reference Goswami and Ellis2021, appendix A). The equation for geodesic deviation due to the Weyl curvature is then simply

(11) $${a^\mu } = - C_{\beta \nu \alpha }^\mu {\xi ^\alpha }{\xi ^\beta }{\chi ^\nu } = E_\nu ^\mu {\chi ^\nu }.$$

We can thus understand the electric part of the Weyl tensor as uniquely responsible for the mechanical stress associated with the tidal deformation effect—that is, the geodesic deviation that changes the shape of bodies in geodesic motion. Footnote 11 Tidal deformation is a non-conventional effect which can be associated with a specific part of the Weyl curvature tensor of any given spacetime. Spacetime Conventionalism 3 is a resounding failure.

5. Nomic structure and dynamical fields

Spacetime Conventionalism 4

  1. (1) The nomic structure that specifies which relativistic spacetimes are dynamically possible models is not an empirically determined truth.

  2. (2) Physical differences with regard to dynamical and non-dynamical interpretations of field equations arise in the comparison of different conventions about whether certain fields are understood to couple or interact within equations.

Whilst the dynamical role of the Weyl tensor forecloses the possibility for a putative conventionalism about tidal deformation, it opens up an opportunity for a more interesting, and ultimately more tenable, form of conventionalism with regard to interpretations of the equations that enforce the connection between Weyl and Ricci curvature. In this section, we will consider a novel form of conventionalism that is built upon underdetermination with regard to the nomic or law-like structure that specifies which relativistic spacetimes are dynamically possible models. In particular, an underdetermination with regard to which equations are taken to encode nomic structure that directly restricts dynamically possible models and which equations are taken to be gauge identities or constraints that restrict kinematically possible models. This underdetermination then, in turn, leads to conventionalism with regard to which fields are understood to couple or interact, which in turn implies potential physical differences. We can articulate the line of thought behind this form of conventionalism as follows.

First, recall that solving the Einstein equation (1) does not uniquely determine the Riemann curvature. This can be seen most obviously in the example of vacuum Einstein spacetimes, which are defined as the class of Ricci-flat Lorentzian manifolds, i.e., Einstein spacetimes in which the Ricci curvature tensor ${R_{\mu \nu }}$ is zero. The Riemann curvature of a Ricci-flat Lorentzian manifold is entirely determined by the Weyl curvature tensor ${C_{\mu \nu \alpha \beta }}$ . In contrast, we might also consider Weyl-flat spacetimes, in which the Ricci curvature encodes all geometric degrees of freedom. Weyl-flat spacetime cannot admit any local deformations of shape and are illustrated most vividly by the Friedmann–Lemaître–Robertson–Walker spacetimes that approximately describe our universe. The intersection of the Ricci-flat and Weyl-flat cases is then the unique Riemann-flat spacetime: Minkowski spacetime. In general, a spacetime will have non-zero Ricci and Weyl curvature and the obvious question is then how these two forms of curvature can be related if the Einstein equation only relates to the Ricci part.

The answer is found in the Bianchi identities. Equation (4) can be rewritten as

(12) $${\nabla _{[\alpha }}{R_{\mu \nu ]\rho \sigma }} = 0.$$

This can be expanded via the Weyl–Ricci decomposition (8) as

(13) $${\nabla _\nu }{C^{\rho \sigma \mu \nu }} = {\nabla ^{[\sigma }}{R^{\rho ]\mu }} + {1 \over 2}{g^{\mu [\sigma }}{\nabla ^{\rho ]}}R$$

In the context of the further decomposition of the Weyl tensor as per equation (9), the contracted Bianchi identities then lead directly to the temporal and spatial derivatives of the electric and magnetic parts of the Weyl tensor, and these equations then lead in vacuo to the propagation equations for gravitational waves (Goswami and Ellis, Reference Goswami and Ellis2021). Footnote 12

The next crucial step towards Spacetime Conventionalism 4 is to consider the form of Maxwell’s equations,

(14) $${\nabla _\sigma }{F^{\rho \sigma }} = {J^\rho },$$

where ${F^{\rho \sigma }}$ is the electromagnetic field tensor and ${J^\rho }$ is the source current, and note that we can rewrite the equations (13) such that they take an analogous form:

(15) $${\nabla _\nu }{C^{\rho \sigma \mu \nu }} = {J^{\rho \sigma \mu }},$$

where ${J^{\rho \sigma \mu }} = {\nabla ^{[\sigma }}{R^{\rho ]\mu }} + {1 \over 2}{g^{\mu [\sigma }}{\nabla ^{\rho ]}}R$ (Hawking and Ellis, Reference Hawking and Ellis1973, 85). This striking analogy suggests we might think of equation (13) as field equations for the Weyl curvature, just as the Einstein equations provide field equations for the Ricci curvature. Furthermore, we can use the Einstein equation to rewrite the source current directly in terms of the stress–energy tensor such that we have an expression of the form

(16) $${\nabla _\nu }{C^{\rho \sigma \mu \nu }} = {\nabla ^{[\sigma }}{T^{\rho ]\mu }} + {1 \over 2}{g^{\mu [\sigma }}{\nabla ^{\rho ]}}T_\nu ^\nu $$

We would then seem to have a dynamical equation for the Weyl curvature as sourced by the stress–energy tensor (Danehkar, Reference Danehkar2009). In such circumstances, the difference between Weyl and Ricci curvature appears to have disappeared in the sense that both can be seen as dynamical fields encoding geometric degrees of freedom that are sourced by the stress–energy tensor.

Notwithstanding the argument just given, we can also arrive at a non-dynamical interpretation of the Weyl tensor by starting with a different reading of the modal status of the Bianchi identities. That is, we understand the identities as encoding restrictions on kinematical rather than dynamical possibilities. A Bianchi identity is, in the general case, defined as the consequence of the “gauge” symmetry properties of a theory via Noether’s second theorem. Footnote 13 The Noether derivation of Bianchi identities does not require stationarity of the relevant action. Then, in the context of general relativity, as was discussed in section 3, the second Bianchi identity follows from the basic mathematical properties of the Riemann tensor and is thus naturally understood as a formal identity without physical content. As such, equation (13) is derivable completely independently of the action and would hold for any spacetime theory formulated on Riemannian geometries with a torsion-free connection.

These considerations lead us to a second non-dynamical interpretational option in which we treat the Bianchi identities as encoding kinematic rather than nomic structure. In the context of general relativity, this means that the space of kinematically possible models of the theory is preselected such that the identities are obeyed. This, in turn, means that the Ricci and Weyl curvatures are kinematically constrained to be coordinated such that the Riemann curvature obeys the relevant identity. Dynamics is then encoded solely within the Einstein equation which then fixes the dynamically allowed Ricci curvature. Residual freedom within the Weyl curvature can be understood as fixed via the choice of initial conditions. Under such an interpretation the relation between the Ricci and Weyl curvature is a product of a pre-established kinematical harmony, not a substantive dynamical relationship. The Weyl tensor is a non-dynamical field.

A useful way of understanding the connection between the interpretation of equations and the interpretation of fields has been suggested by Lehmkuhl (Reference Lehmkuhl2011, 469):

[A]n interaction demands that all fields present are dynamical fields […] it seems sensible to make a distinction between speaking of two fields interacting and two fields coupling. For a dynamical field can couple to a non-dynamical field […] but we would not speak of an interaction if only one of the two fields was dynamical: a non-dynamical field acts without being acted upon if it couples to a dynamical field. Hence, two fields interacting should be seen as sufficient but not necessary for the fields to couple, whereas two fields coupling is necessary but not sufficient for the two fields to interact.

Following this formulation, we would then have that if the Bianchi identities (13) describe an interaction then the Weyl tensor is a dynamical field of the same status as the Ricci tensor; they each act whilst simultaneously being acted upon. However, if the Bianchi identities (13) describe a coupling, then we should offer a non-dynamical interpretation of the Weyl curvature tensor: it merely couples to the Ricci tensor, the two fields do not interact; the Ricci tensor acts upon the Weyl tensor without being acted upon. Footnote 14

What would appear to be the best strategy for breaking this interpretational underdetermination is to consider the initial value problem. Footnote 15 In particular, if the initial value problem of general relativity were found to be fully specifiable independently of the Bianchi identities, one could argue that only the Einstein equations encode nomic structure, and the non-dynamical interpretation is supported. Conversely, if the Bianchi identities play an explicit role in encoding interactions between degrees of freedom within the initial value problem then they would encode nomic structure and the dynamical interpretation would be supported. Footnote 16

Approaches to the explicit solution of the initial value problem in general relativity proceed via hyperbolic reductions based upon a particular gauge choice. The most famous of these is the original “harmonic” or “wave” gauge treatment due to Choqouet-Bruhat. Footnote 17 In this approach, although the Bianchi identities do play a substantive role, they do not form part of the basic system of hyperbolic dynamical equations that are obtained as a reduction of the Einstein equations. This would seem to support the non-dynamical interpretation. However, at least in vacuum, there exists an alternative approach to the hyperbolic reduction of general relativity due to Friedrich (Reference Friedrich1996) in which the Bianchi identities are explicitly understood as hyperbolic propagation equations for the Weyl tensor. Explicitly, and considering the vacuum case, within this approach the Weyl tensor is treated as one of the fundamental dynamical variables in a system of equations given by

(17) $$R_{\nu \lambda \rho }^\mu = C_{\nu \lambda \rho }^\mu, {\rm{\;\;\;\;\;\;\;\;}}{\nabla _\mu }C_{\nu \lambda \rho }^\mu = 0,$$

which are equivalent to the vacuum forms of the Einstein equation and Bianchi identity respectively. This is precisely to understand the Einstein equations and Bianchi identities as interaction equations of equal status and, moreover, to explicitly treat the Weyl tensor as a dynamical field. Footnote 18

Study of the initial value problem of general relativity thus serves to further establish, rather than dissolve, the putative underdetermination with regard to nomic structure that is at the heart of Spacetime Conventionalism 4. If we appeal to the initial value problem to identify the nomic structure of general relativity then we find that such a specification depends upon a choice as to which approach to hyperbolic reduction we apply. At least in the vacuum case, this then explicitly allows for different options with regard to the attribution of a dynamical or non-dynamical role to the Weyl tensor.

We can draw an analogy between this situation and the role of the cosmological constant. Footnote 19 Consider a “ ${\rm{\Lambda }}$ -vacuum” spacetime of the form

(18) $${G_{\mu \nu }} + {g_{\mu \nu }}{\rm{\Lambda }} = 0.$$

Writing the equation in this way makes it natural to think of ${\rm{\Lambda }}$ as a non-dynamical constant of nature. We might, of course, alternatively rewrite the same equation as

(19) $${G_{\mu \nu }} = - {g_{\mu \nu }}{\rm{\Lambda }}.$$

In this context, it is natural to think of ${\rm{\Lambda }}$ as a dynamical source for the Ricci degrees of freedom. These are completely trivial rewritings of the same equation, so it is not clear what could hang on the dynamical vs. non-dynamical interpretation of ${\rm{\Lambda }}$ . However, there are good reasons coming from outside the theory to treat the cosmological constant vs. dark energy interpretational question as a substantive physical issue (Peebles and Ratra, Reference Peebles, James and Ratra2003; Huterer, Reference Huterer and Goodstein2011).

This requirement for reference to external physical theories or physics principles could also be taken to be required for the case of our two interpretations of the Ricci–Weyl relationship. Within general relativity, the equations themselves do not give priority to the differing dynamical and non-dynamical interpretations. However, in a wider theoretical context, and considering, in particular, the relevance to quantization, the distinction between the interpretation may in fact ground genuine physical differences. For example, whereas it is generally the case that kinematical restrictions are converted to super-selection rules in quantization, dynamical restrictions, such as conservation laws, are applied as quantum nomological restrictions, which allows for their violation subject to the uncertainty principle. Footnote 20 Spacetime Conventionalism 4 might thus offer insight into alternative heuristic strategies for theory extension.

Finally, an alternative strategy for breaking the underdetermination is to consider the deeper mathematical structure behind the tensor fields in question and appeal to the primacy of such structure as the basis upon which to designate which fields are dynamical and which non-dynamical. In particular, one can understand the Weyl and Ricci tensors as composite objects built out of the metric tensor, spinors, or connection coefficients, and then argue that since the Bianchi identities, but not the Einstein equations, follow from the mathematical definition of these objects, only the latter can be understood as nomic structure. Footnote 21 In the context of such a view, Spacetime Conventionalism 4 would, in its failure, again provide a valuable heuristic, in this case towards a particular, non-trivial, mathematical ontology of relativity theory. Such an avenue of investigation warrants further consideration.

6. The Bach Conjecture

Spacetime Conventionalism 5

  1. (1) The geometric structure of spacetime is not an empirically determined truth.

  2. (2) Physical differences with regard to which tensor fields play the role of geometric structure arise through different conventions in the decomposition of the Einstein tensor into conformally invariant and non-conformally invariant parts.

The final form of conventionalism we will consider is, in a sense, a marriage of the formulation of geometric conventionalism in the spirit of Weatherall and Manchak (Reference Weatherall and Manchak2014) with the idea discussed above of understanding conformally invariant tensors as representing the geometric structure of spacetime. The view is, however, to our knowledge entirely novel. Furthermore, it leads to a precise mathematical conjecture the truth of which is as yet unproven.

The basic starting point for this particular strategy is to look to re-express the Einstein equation in a conformally invariant form. That is, seek to show that Einstein’s equation is true if and only if

(20) $${B_{\mu \nu }} = 8\pi {T_{\mu \nu }} + {A_{\mu \nu }},$$

where ${T_{\mu \nu }}$ is the matter–energy tensor appearing in Einstein’s equation, ${B_{\mu \nu }}$ is a conformally invariant tensor, and ${A_{\mu \nu }}$ is a further tensor that is not conformally invariant. So far this is purely a formal exercise in analysis. The interpretation move is then to take ${B_{\mu \nu }}$ to characterize facts about geometric spacetime structure, which is, by assumption, understood to be conformally invariant. By contrast, the non-conformally invariant objects ${T_{\mu \nu }}$ and ${A_{\mu \nu }}$ are taken to characterize a matter–energy field and a further “universal effect,” restrictively. Footnote 22 Clearly, by the theorem of Weatherall and Manchak (Reference Weatherall and Manchak2014), ${A_{\mu \nu }}$ will not be expressible as a Newtonian force.

The reformulation (20) we are looking for is indeed possible, by defining ${B_{\mu \nu }}$ to be the Bach tensor Footnote 23 which can be expressed without introducing further structure in terms of the Weyl and Ricci tensors as

(21) $${B_{\mu \nu }}: = {\nabla ^\sigma }{\nabla ^\rho }{C_{\rho \mu \nu \sigma }} + {1 \over 2}{C_{\rho \mu \nu \sigma }}{R^{\rho \sigma }}.$$

The Bach tensor is conformally invariant in $D = 4$ (Bach, Reference Bach1921). Thus, defining ${A_{\mu \nu }}: = {B_\mu } - {G_{\mu \nu }}$ , we have that for $D = 4$ equation (20) holds if and only if the Einstein equation does, as desired. Footnote 24

The idea of a “pure geometry” interpretation of the Bach tensor is noted in the mathematical physics literature. In particular, (Bergman, Reference Bergman2004, 18) offers the observation that: “The Bach tensor is a tensor built up from pure geometry, and thereby captures necessary features of a space being conformally Einstein in an intrinsic way.”

The pivotal issue in the context of Spacetime Conventionalism 5 is then whether the Bach tensor ${B_{\mu \nu }}$ is the unique conformally invariant two-place tensor, at least up to a multiplicative constant. Adopting the interpretation of geometric structure and matter–energy above, uniqueness would imply that this decomposition allows one to distinguish uniquely between geometric structure ( ${B_{\mu \nu }}$ ) and matter–energy ( ${T_{\mu \nu }}$ ) and the “universal effect” ( ${A_{\mu \nu }}$ ); and non-uniqueness would amount to a sense in which the distinction between geometric structure and matter–energy is conventional.

Fascinatingly, although this question is both mathematically precise and foundationally important, its status appears, as yet, unsettled. The issue at hand is one which can be more generally associated with finding the conformal invariants in the theory of Lorentzian manifolds: “Classically known conformally invariant tensors include the Weyl conformal curvature tensor, which plays the role of the Riemann curvature tensor, its three-dimensional analogue the Cotton tensor, and the Bach tensor in dimension four. Further examples are not so easy to come by” (Fefferman and Graham, Reference Fefferman and Robin Graham2012, 1). Thus, although the review of the literature does not reveal any other known conformal invariants in four dimensions, there is no proof of the absence of such alternatives either.

One might thus propose that the debate over conventionality and non-conventionality, as described above, is not just a matter of philosophical debate, but of settling a precise mathematical conjecture whose truth is not yet known:

Bach Conjecture. The Bach tensor is the unique (up to a multiplicative constant) conformally invariant rank-two tensor field on a four-dimensional Lorentzian manifold.

If the Bach Conjecture is false, then Spacetime Conventionalism 5 is true; if the Bach Conjecture is true, then Spacetime Conventionalism 5 is false. We thus take the proof or refutation of the Bach Conjecture to be an important problem in the foundations of general relativity that warrants considered attention.

7. Summary

One might reasonably judge proposals for the conventionality of spacetime structure in terms of their respective novelty, interestingness, and plausibility. Our discussion has characterized and critically evaluated five formulations of the conventionality thesis and we hope the reader will agree that all five have proved worthy of consideration. The first three options considered have been argued to be implausible on grounds that vary from direct refutation (Spacetime Conventionalism 3) to provable falsity under certain assumptions (Spacetime Conventionalism 1) to rejection on the grounds of the lack of physical salience of the central undetermined object (Spacetime Conventionalism 2). Nevertheless, each of the three views illustrates important morals regarding the interplay between empirical and mathematical structures within the foundations of spacetime theory. The final two forms of conventionalism we considered to be novel, interesting, and plausible. Moreover, in terms of analysis of the dynamical role of the Bianchi identities (Spacetime Conventionalism 4) and proof or refutation of the Bach Conjecture (Spacetime Conventionalism 5), they each open up a new direction of enquiry that we may hope to be profitably pursued in future research.

Acknowledgements

The form, scope, and content of this paper were transformed through hugely productive interactions with Bryan Roberts. In particular, both the framing of varieties of conventionalism via the two questions and the formulation of the Bach Conjecture are due directly to Bryan. We further greatly profited from detailed written comments and lengthy discussions with Bryan without which the present paper would be a shadow of what it is. The paper has also profited from a number of other detailed discussions that KT has had over the last years with (among others) Erik Curiel, Juliusz Doboszewsk, Sam Fletcher, Henrique Gomes, Sean Gryb, James Ladyman, Dennis Lehmkuhl, Oliver Pooley, James Read, Simon Saunders, Dave Sloan, and Jim Weatherall. We are also grateful to online audiences at the DPG conference Heidelberg and the Philosophy of Physics and Chemistry group in Buenos Aires, Argentina, and to Patrick Duerr and two anonymous referees for helpful comments. The authors take full responsibility for any errors, omissions, or inaccuracies.

Appendix

A. Conformal tensors and the Bianchi identities

The Bianchi identities can also be related to the Cotton tensor ${{\cal C}_{\rho \sigma \nu }}$ (Cotton, Reference Cotton1899) and Bach tensor ${B_{ab}}$ as follows. Following García et al. (Reference Garca, Hehl, Heinicke and Macías2004), we can decompose curvature into irreducible representations with respect to the pseudo-orthogonal group for certain dimensions, $n$ , as follows:

(A1) $$n = 1 \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \to {R_{\alpha \beta }} = 0,$$
(A2) $$n = 2 \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \to {R_{\alpha \beta }} = Scala{r_{\alpha \beta }},$$
(A3) $$n = 3 \ \ \ \ \ \ \ \ \ \ \ \ \to {R_{\alpha \beta }} = Scala{r_{\alpha \beta }} + {\tilde R_{\alpha \beta }},$$
(A4) $$n \ge 4 \ \ \ \ \to {R_{\alpha \beta }} = Scala{r_{\alpha \beta }} + {\tilde R_{\alpha \beta }} + {C_{\alpha \beta }},$$

where $\tilde R$ denotes the traceless Ricci part of the decomposition and every tensor is in its Weyl 2-form. For $n = 3$ , the Cotton tensor is given by $D{R_{\alpha \beta }} = D{C_{\alpha \beta }} + \left( {2/\left( {n - 2} \right)} \right){\upsilon _{[\alpha }} \wedge {{\cal C}_{\beta ]}} = 0$ , where $D$ denotes the exterior covariant derivative and $\upsilon $ denotes the coframe of a Riemannian space of $n$ dimensions,

(A5) $${{\cal C}_{\rho \sigma \nu }} = 2\left( {{\nabla _{[\alpha }}{R_{\beta ]\gamma }} - {1 \over {2\left( {n - 1} \right)}}{\nabla _{[\alpha }}R{g_{\beta ]\gamma }}} \right),$$

and, for $n = 4$ , the Bach tensor can be defined as

(A6) $${B_{\alpha \beta }}: = {\nabla ^\mu }{{\cal C}_{\alpha \mu \beta }} + {L^{\mu \nu }}{C_{\alpha \mu \beta \nu }},$$

where ${L^{\mu \nu }}$ is called Schouten tensor and is given by

(A7) $${L_{\mu \nu }} = {R_{\mu \nu }} - {1 \over {2\left( {n - 1} \right)}}R{g_{\mu \nu }}.$$

Although the Bach and Weyl tensors appear to be independent in tensorial form, they can be directly associated with one another via a variational principle. In particular, we can drive the Bach tensor from a Lagrangian composed of Weyl tensors that takes the form Footnote 25

(A8) $$L = {C_{\rho \mu \nu \sigma }}{C^{\rho \mu \nu \sigma }} + k{g^{\mu \nu }}{T_{\mu \nu }},$$

where ${T_{\mu \nu }}$ is the relevant matter field with a non-zero constant. The relevant Euler–Lagrange equations are then Bach’s equation of motion,

(A9) $${B_{\mu \nu }} = k{T_{\mu \nu }},$$

where ${T_{\mu \nu }}$ is traceless due to conformal invariance of ${B_{\mu \nu }}$ .

Additionally, considering the vacuum case, the Bach tensor exhibits parallel behavior to the Weyl tensor in relation to the Bianchi identities given in (17): Footnote 26

(A10) $${\nabla ^d}{C_{abcd}} = {K^d}{C_{abcd}},{\rm{\;\;\;\;\;\;\;\;}}{B_{\mu \nu }} = 0.$$

Footnotes

1 The thesis of conventionalism regarding the geometry of space was first proposed by Poincaré (Reference Poincaré1902), following in the tradition of Riemann (Reference Riemann1854), and developed in a variety of forms by Dingler (Reference Dingler1911), Carnap (Reference Carnap1922), and Reichenbach (Reference Reichenbach1928). See Ben-Menahem (Reference Ben-Menahem2006) and Duerr and Ben-Menahem (Reference Duerr and Ben-Menahem2022) for an overview. Ivanova (Reference Ivanova2015a,b) gives a helpful analysis of Poincaré’s conventionalism, and Torretti (Reference Torretti1978) of Dingler’s. A translation of the 1919–21 doctoral dissertation of Carnap (Reference Carnap1922) can be found in Carus et al. (Reference Carus, Kienzler, Richardson and Schlotter2019). Further analysis with varying degrees of historical grounding can be found in Grünbaum (Reference Grünbaum1968); Sklar (Reference Sklar1974, Reference Sklar1985); Glymour (Reference Glymour1977); Friedman (Reference Friedman1983); Malament (Reference Malament1985); Norton (Reference Norton, Majer and Schmidt1994); Weatherall and Manchak (Reference Weatherall and Manchak2014); Dewar et al. (Reference Dewar, Linnemann and Read2022).

2 Our project is not primarily a historical one and so we will not seek to defend the historical as opposed to foundational salience of our generalizations. That said, our framework for generalization overlaps in several core respects with the reconstruction of Poincaré’s conventionalism provided by (Duerr and Ben-Menahem, Reference Duerr and Ben-Menahem2022, 167).

3 See Duerr and Ben-Menahem (Reference Duerr and Ben-Menahem2022) for specific arguments that involve questioning the Weatherall–Manchak result on the basis of both its mathematical assumptions—(NORM) that $\tilde \xi $ is a unit norm, and (RIEM) that $\left( {M,{{\tilde g}_{\mu \nu }}} \right)$ is a Riemannian geometry—and physical assumptions—(CONF) that the old and new metrics are conformally related, ${\tilde g_{\mu \nu }} = {{\rm{\Omega }}^2}{g_{\mu \nu }}$ , and (FORCE) that the “universal force” is induced by a tensor field ${G_{\mu \nu }}$ via $G_{{\rm{\;}}\mu }^\nu {\tilde \xi ^\mu }$ .

4 See Pitts (Reference Pitts2010, Reference Pitts2016, Reference Pitts2021); Lam (Reference Lam2011); Read (Reference Read2020); Dewar and Weatherall (Reference Dewar and Weatherall2018); Duerr (Reference Duerr2019a,b) for discussions of the problem of defining gravitational energy. A further noteworthy paper, along slightly different lines, is Lehmkuhl (Reference Lehmkuhl2011). There it is argued that in the context of general relativity energy-momentum density as expressed by ${T_{\mu \nu }}$ , cannot, in fact, be regarded as an intrinsic property of matter, but rather should be understood as a relational property that matter possesses only in virtue of its relation to spacetime structure. For an outstanding general discussion of energy conditions in general relativity see Curiel (Reference Curiel2014).

5 See (2012, 69–79) and Landsman (Reference Landsman2021, 61) for proofs.

6 See Malament (Reference Malament2012, 322). Note that one also uses further symmetry properties of the Riemann tensor in this short derivation.

7 Following the results of Fischer (Reference Fischer, Carmeli, Fickler and Witten1970), what can be proved explicitly is that for globally hyperbolic Einstein spacetimes which admit an initial value formulation, the topology of the space of physically distinguishable initial states (i.e., the “superspace” composed of distinct Riemannian three-geometries) is such that the geometries admitting continuous isometries are singular points. Furthermore, initial data without non-trivial Killing vector fields is generic in the sense that “geometries of high symmetry are completely contained in the boundary of geometries of lower symmetry” (303). There is thus a good formal basis to expect that a generic (globally hyperbolic) Einstein spacetime will lack Killing vector fields. At a simpler, more intuitive, level it is not hard to convince oneself that any physically realistic spacetime will be too “messy” in terms of inhomogeneities and anisotropies to contain Killing vector fields.

8 It is true that, in general, local conservation of stress–energy will approximately hold for “small enough” patches of spacetime, with how small depending on the relevant curvature (Wald, Reference Wald1984, 70). Thus, there will, in general, exist “approximate Killing observers” for whom within some “small enough” region there are no tidal forces and gravitational stress–energy 4-current vanishes. Operationally, however, observers cannot be arbitrarily small, and thus making such a move would violate the spirit of the Reichenbachian argument.

9 Following Ramond (Reference Ramond1997), we can decompose $R_{\sigma \mu \nu }^\rho $ in terms of ${\rm{SU}}\left( 2 \right) \otimes {\rm{SU}}\left( 2 \right)$ as it is locally isomorphic to ${\rm{SO}}\left( {3,1} \right)$ . We decompose the Riemann curvature tensor into symmetric parts,

$$\left( {3 \otimes 3} \right) \oplus \left( {5 \otimes 1} \right) \oplus \left( {1 \otimes 5} \right) \oplus \left( {1 \otimes 1} \right) \oplus \left( {1 \otimes 1} \right),$$

and antisymmetric parts, $\left( {3 \otimes 3} \right) \oplus \left( {3 \otimes 1} \right) \oplus \left( {1 \otimes 3} \right)$ . We then find that the Weyl tensor ${C_{\rho \sigma \mu \nu }}$ transforms as $\left( {5 \otimes 1} \right) \oplus \left( {1 \otimes 5} \right)$ and is conformally symmetric. The object that transforms as $\left( {3 \otimes 3} \right)$ in the symmetric part corresponds to the traceless part of the Ricci tensor $R$ .

10 Here we are following Goswami and Ellis (Reference Goswami and Ellis2021). The origin of this decomposition is Matte (Reference Matte1953). Further discussion can be found in Bel (Reference Bel2000). Thanks to Juliusz Doboszewski for help with these historical sources.

11 There is also an important connection between the electric part of the Weyl tensor and gravitational waves—see Goswami and Ellis (Reference Goswami and Ellis2021) for more details on the formal relation. A recent insightful discussion of gravitational waves and gravitational energy in the context of the analogy with electromagnetic waves can be found in Gomes and Rovelli (Reference Gomes and Rovelli2023).

12 For the original treatments along related lines see Newman and Penrose (Reference Newman and Penrose1962); Newman and Unti (Reference Newman and Unti1962); Hawking (Reference Hawking1966).

13 For a detailed discussion of Noether’s second theorem in a historical context see Kosmann-Schwarzbach (Reference Kosmann-Schwarzbach2010). For a rigorous formal overview see Olver (Reference Olver1993). For discussion in the context of gauge theories and quantization see Henneaux and Teitelboim (Reference Henneaux and Teitelboim1992).

14 This idea parallels in certain key respects the discussion of the distinction between matter and spacetime found in Martens and Lehmkuhl (Reference Martens and Lehmkuhl2020).

15 Thanks to Henrique Gomes for suggesting this idea to us.

16 The idea that (well-posed) initial value problems could play a privileged role in picking out the ontic structure of physical theories is explored in a rather different context by Callender (Reference Callender2017). See also also Baron and Evans (Reference Baron and Evans2021); James (Reference James2022).

17 Here we are following Choquet-Bruhat (Reference Choquet-Bruhat2008, §VI) and Landsman (Reference Landsman2021, 7.5).

18 For a more general application of the idea of treating the Bianchi identities as alternative representations of the field equations in the context of matter systems see Friedrich and Rendall (Reference Friedrich, Rendall and Bernd2000).

19 See Martens and Lehmkuhl (Reference Martens and Lehmkuhl2020) for a related discussion regarding modified gravity and dark matter.

20 See Gryb and Thébault (Reference Gryb and Thébault2016) for more discussion in the context of the problem of time and the cosmological constant.

21 Thanks to Henrique Gomes for suggesting this argument to us.

22 The interpretation of ${T_{\mu \nu }}$ as matter is equivalent to matter criterion G of Martens and Lehmkuhl (Reference Martens and Lehmkuhl2020).

23 Named for Rudolf Bach, not the Baroque German composer. Although, the former was in fact a pseudonym for Rudolf Förster. See appendix A for further formal details.

24 For $D = 2$ , all relativistic spacetimes are trivial vacuum solutions to Einstein’s equation since it can be proved both that the Einstein tensor is identically zero and that the natural candidate for a two-dimensional Weyl tensor is identically zero (Fletcher et al., Reference Fletcher, Manchak, Schneider and Weatherall2018). For $D = 3$ , the Riemann curvature tensor is completely determined by the local matter distribution (Giddings et al., Reference Giddings, Abbott and Kucharr1984), so although there are interesting conformally invariant tensors in three dimensions, such as the Weyl tensor and Cotton tensor, the dynamics of three-dimensional gravity is such that these are constrained to be zero. Thus, Conventionalism 5 for $D \lt 4$ is uninteresting.

25 See [Fiedler and Schimming (Reference Fiedler and Schimming1980)] for the exact solutions.

26 See [Bergman (Reference Bergman2004), 39] for a discussion.

References

Bach, Rudolf. 1921. “Zur Weylschen Relativitätstheorie und der Weylschen Erweiterung des Krümmungstensorbegriffs”. Mathematische Zeitschrift 9 (1–2):110–35.CrossRefGoogle Scholar
Baron, Sam and Evans, Peter W.. 2021. “What’s So Spatial About Time Anyway?The British Journal for the Philosophy of Science 72 (1):159–83.CrossRefGoogle Scholar
Bel, Louis. 2000. “Radiation States and the Problem of Energy in General Relativity”. General Relativity and Gravitation 32 (10):2047–78.CrossRefGoogle Scholar
Ben-Menahem, Yemima. 2006. Conventionalism: From Poincaré to Quine. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Bergman, Jonas. 2004. “Conformal Einstein Spaces and Bach Tensor Generalization in $n$n Dimensions”. PhD diss., Linköpings universitet.Google Scholar
Callender, Craig. 2017. What Makes Time Special? Oxford: Oxford University Press.CrossRefGoogle Scholar
Carnap, Rudolf. 1922. Der Raum: Ein Beitrag zur Wissenschaftslehre. Berlin: Verlag von Reuther & Reichard.Google Scholar
Carnap, Rudolf. 1966. Philosophical Foundations of Physics: An Introduction to the Philosophy of Science. New York: Basic Books Inc. Edited by Martin Gardner.Google Scholar
Carus, A. W., Michael Friedman, Kienzler, Wolfgang, Richardson, Alan, and Schlotter, Sven, eds. 2019. Rudolf Carnap, Early Writings: The Collected Works of Rudolf Carnap, Volume 1. Oxford: Oxford University Press.Google Scholar
Choquet-Bruhat, Yvonne. 2008. General Relativity and the Einstein Equations. Oxford: Oxford University Press.CrossRefGoogle Scholar
Cotton, Émile. 1899. “Sur les variétés a trois dimensions”. Annales de la Faculté des Sciences de Toulouse: Mathématiques 1 (4):385–38. https://doi.org/10.5802/afst.160.Google Scholar
Curiel, Erik. 2014. “A Primer on Energy Conditions”. Preprint, arXiv:1405.0403.Google Scholar
Danehkar, Ashkbiz. 2009. “On the Significance of the Weyl Curvature in a Relativistic Cosmological Model”. Modern Physics Letters A 24 (38):3113–27.CrossRefGoogle Scholar
Dardashti, Radin. 2021. “No-Go Theorems: What Are They Good For?Studies in History and Philosophy of Science Part A 86:4755.CrossRefGoogle ScholarPubMed
Dewar, Neil, Linnemann, Niels, and Read, James. 2022. “The Epistemology of Spacetime”. Philosophy Compass 17 (4):e12821. https://doi.org/10.1111/phc3.12821.CrossRefGoogle Scholar
Dewar, Neil and Weatherall, James O.. 2018. “On Gravitational Energy in Newtonian Theories”. Foundations of Physics 48 (5):558–78.CrossRefGoogle Scholar
Dieks, Dennis. 1987. “Gravitation as a Universal Force”. Synthese 73 (2):381–97.CrossRefGoogle Scholar
Dingler, Hugo. 1911. Die Grundlagen Der Angewandten Geometrie: Eine Untersuchung Über Den Zusammenhang Zwischen Theorie Und Erfahrung in Den Exakten Wissenschaften. Leibzig: Akademische Verlagsgesellschaft.Google Scholar
Duerr, Patrick M. 2019a. “Against ‘Functional Gravitational Energy’: A Critical Note on Functionalism, Selective Realism, and Geometric Objects and Gravitational Energy”. Synthese 199:299333.CrossRefGoogle Scholar
Duerr, Patrick M. 2019b. “It Ain’t Necessarily So: Gravitational Waves and Energy Transport”. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 65:2540.CrossRefGoogle Scholar
Duerr, Patrick M and Ben-Menahem, Yemima. 2022. “Why Reichenbach Wasn’t Entirely Wrong, and Poincaré Was Almost Right, About Geometric Conventionalism”. Studies in History and Philosophy of Science 96:154–73.CrossRefGoogle ScholarPubMed
Fefferman, Charles and Robin Graham, C.. 2012. The Ambient Metric. Princeton, NJ: Princeton University Press.Google Scholar
Feyerabend, Paul K. 1981. “More Clothes from the Emperor’s Bargain Basement”. The British Journal for the Philosophy of Science 32:5794.CrossRefGoogle Scholar
Fiedler, Bernd and Schimming, R. 1980. “Exact Solutions of the Bach Field Equations of General Relativity”. Reports on Mathematical Physics 17 (1):1536.CrossRefGoogle Scholar
Fischer, Arthur E. 1970. “The Theory of Superspace”. In Relativity: Proceedings of the Relativity Conference in the Midwest, edited by Carmeli, Moshe, Fickler, Stuart I., and Witten, Louis, 303–57. New York: Plenum Press.CrossRefGoogle Scholar
Fletcher, Samuel C., Manchak, John B., Schneider, Mike D., and Weatherall, James O.. 2018. “Would Two Dimensions Be World Enough for Spacetime?Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 63:100113.CrossRefGoogle Scholar
Friedman, Michael. 1983. Foundations of Space-time Theories. Princeton, NJ: Princeton University Press.Google Scholar
Friedrich, Helmut. 1996. “Hyperbolic Reductions for Einstein’s Equations”. Classical and Quantum Gravity 13 (6):1451.CrossRefGoogle Scholar
Friedrich, Helmut and Rendall, Alan. 2000. “The Cauchy Problem for the Einstein Equations”. In Einstein’s Field Equations and their Physical Implications: Selected Essays in Honour of Jürgen Ehlers, edited by Bernd, G. Schmidt, 127223. New York: Springer.Google Scholar
Garca, Alberto A., Hehl, Friedrich W., Heinicke, Christian, and Macías, Alfredo. 2004. “The Cotton Tensor in Riemannian Spacetimes”. Classical and Quantum Gravity 21 (4):1099.CrossRefGoogle Scholar
Giddings, Steven, Abbott, James, and Kucharr, Karel. 1984. “Einstein’s Theory in a Three-Dimensional Space-time”. General Relativity and Gravitation 16:751–75.CrossRefGoogle Scholar
Glymour, Clark. 1977. “The Epistemology of Geometry”. Noûs 11 (3):227251.CrossRefGoogle Scholar
Gomes, Henrique and Rovelli, Carlo. 2023. “On the Analogies Between Gravitational and Electromagnetic Radiative Energy”. Preprint, arXiv:2303.14064.Google Scholar
Goswami, Rituparno and Ellis, George F. R.. 2021. “Tidal Forces are Gravitational Waves”. Classical and Quantum Gravity 38 (8):085023. https://doi.org/10.1088/1361-6382/abdaf3.CrossRefGoogle Scholar
Grünbaum, Adolf. 1968. Geometry and Chronometry in Philosophical Perspective. Minneapolis, MN: University of Minnesota Press.Google Scholar
Gryb, Sean and Thébault, Karim. 2016. “Schrödinger Evolution for the Universe: Reparametrization”. Classical and Quantum Gravity 33 (6):065004.CrossRefGoogle Scholar
Hawking, Stephen W. 1966. “Perturbations of an Expanding Universe”. The Astrophysical Journal 145:544–54.CrossRefGoogle Scholar
Hawking, Stephen W. and Ellis, George F. R.. 1973. The Large Scale Structure of Space-time. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Henneaux, Marc and Teitelboim, Claudio. 1992. Quantization of Gauge Systems. Princeton, NJ: Princeton University Press.CrossRefGoogle Scholar
Huterer, Dragan. 2011. “The Accelerating Universe”. In Adventures In Cosmology, edited by Goodstein, David, 321–53. Singapore: World Scientific.CrossRefGoogle Scholar
Ivanova, Milena. 2015a. “Conventionalism About What? Where Duhem and Poincaré Part Ways”. Studies in History and Philosophy of Science Part A 54:8089.CrossRefGoogle Scholar
Ivanova, Milena. 2015b. “Conventionalism, Structuralism and Neo-Kantianism in Poincaré’s Philosophy of Science”. Studies in History and Philosophy of Modern Physics 52:114–22.CrossRefGoogle Scholar
James, Lucy. 2022. “A New Perspective on Time and Physical Laws”. The British Journal for the Philosophy of Science 73 (4):849–77.CrossRefGoogle Scholar
Kosmann-Schwarzbach, Yvette. 2010. The Noether Theorems: Invariance and Conservation Laws in the Twentieth Century. New York: Springer.Google Scholar
Lam, Vincent. 2011. “Gravitational and Nongravitational Energy: The Need for Background Structures”. Philosophy of Science 78 (5):1012–23.CrossRefGoogle Scholar
Landsman, Klaas. 2021. Foundations of General Relativity: From Einstein to Black Holes. Nijmegen, Radboud University Press.CrossRefGoogle Scholar
Laudan, Larry. 1977. Progress and its Problems. Berkeley, CA: University of California Press.Google Scholar
Lehmkuhl, Dennis. 2011. “Mass–Energy–Momentum: Only There Because of Spacetime?The British Journal for the Philosophy of Science 62 (3):453–88.CrossRefGoogle Scholar
Malament, David. 1985. “A Modest Remark About Reichenbach, Rotation, and General Relativity”. Philosophy of Science 52 (4):615–20.CrossRefGoogle Scholar
Malament, David B. 2012. Topics in the Foundations of General Relativity and Newtonian Gravitation Theory. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Martens, Niels C. M. and Lehmkuhl, Dennis. 2020. “Dark Matter = Modified Gravity? Scrutinising the Spacetime–Matter Distinction Through the Modified Gravity/Dark Matter Lens”. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 72:237–50. https://doi.org/10.1016/j.shpsb.2020.08.003.CrossRefGoogle Scholar
Matte, Alphonse. 1953. “Sur De Nouvelles Solutions Oscillatoires Des Equations De La Gravitation”. Canadian Journal of Mathematics 5:116. https://doi.org/10.4153/CJM-1953-001-3.CrossRefGoogle Scholar
Newman, Ezra and Penrose, Roger. 1962. “An Approach to Gravitational Radiation by a Method of Spin Coefficients”. Journal of Mathematical Physics 3 (3):566–78.CrossRefGoogle Scholar
Newman, Ezra T. and Unti, Theodore W. J.. 1962. “Behavior of Asymptotically Flat Empty Spaces”. Journal of Mathematical Physics 3 (5):891901.CrossRefGoogle Scholar
Norton, John D. 1994. “Why Geometry Is Not Conventional: The Verdict of Covariance Principles”. In Semantical Aspects of Spacetime Theories, edited by Majer, Ulrich and Schmidt, Heinz-Jürgen, 159–67. Mannheim: B. I. Wissenschaftsverlag.Google Scholar
Olver, Peter J. 1993. Applications of Lie Groups to Differential Equations. New York: Springer.CrossRefGoogle Scholar
Peebles, P. James, E. and Ratra, Bharat. 2003. “The Cosmological Constant and Dark Energy”. Reviews of Modern Physics 75 (2):559.CrossRefGoogle Scholar
Pitts, J. Brian. 2010. “Gauge-Invariant Localization of Infinitely Many Gravitational Energies from All Possible Auxiliary Structures”. General Relativity and Gravitation 42 (3):601–22. https://doi.org/10.1007/s10714-009-0874-y.CrossRefGoogle Scholar
Pitts, J Brian. 2016. “Einstein’s Physical Strategy, Energy Conservation, Symmetries, and Stability: ‘But Grossmann & I believed that the conservation laws were not satisfied’.” Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 54:5272. https://doi.org/10.1016/j.shpsb.2016.04.003.CrossRefGoogle Scholar
Pitts, J Brian. 2021. “Conservation of Energy: Missing Features in its Nature and Justification and Why They Matter”. Foundations of Science 26 (3):559–84. https://doi.org/10.1007/s10699-020-09657-1.CrossRefGoogle ScholarPubMed
Poincaré, Henri. 1902. La Science et l’Hypothèse. Paris: E. Flamarion. Published in English in 1952 as Science and Hypothesis, New York: Dover Publications Inc. https://archive.org/details/lascienceetlhypo00poin.Google Scholar
Ramond, Pierre. 1997. Field Theory: A Modern Primer. Boulder, CO: Westview Press.Google Scholar
Read, James. 2020. “Functional Gravitational Energy”. The British Journal for the Philosophy of Science 71 (1):205–32.CrossRefGoogle Scholar
Reichenbach, Hans. 1928. Philosophie Der Raum-Zeit-Lehre. Berlin and Leipzig: Walter de Gruyter. Published in English as The Philosophy of Space and Time (1958), translated by Maria Reichenbach and John Freund, New York: Dover Publications, Inc.CrossRefGoogle Scholar
Riemann, Bernhard. 1854. “Ueber Die Hypothesen, Welche DerGeometrie Zu Grunde Liegen”. Abhandlungen der Königlichen Gesellschaft der Wissenschaften zu Göttingen 13.Google Scholar
Sexl, Roman U. 1970. “Universal Conventionalism and Space-time”. General Relativity and Gravitation 1:159–80.CrossRefGoogle Scholar
Sklar, Lawrence. 1974. Space, Time and Spacetime. Berkeley, CA: University of California Press.Google Scholar
Sklar, Lawrence. 1985. Philosophy and Spacetime Physics. Berkeley, CA: University of California Press.Google Scholar
Torretti, Roberto. 1978. “Hugo Dingler’s Philosophy of Geometry”. Diálogos. Revista de Filosofía de la Universidad de Puerto Rico 32:85128.Google Scholar
Wald, Robert M. 1984. General Relativity. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Weatherall, James O. and Manchak, John B.. 2014. “The Geometry of Conventionality”. Philosophy of Science 81 (2):233–47.CrossRefGoogle Scholar