Skip to main content
Springer logoLink to Springer
. 2025 Aug 25;85(8):902. doi: 10.1140/epjc/s10052-025-14656-2

Reassessing the foundations of metric-affine gravity

J François 1,2,3, L Ravera 4,5,6,
PMCID: PMC12378725  PMID: 40874032

Abstract

We reassess foundational aspects of Metric-Affine Gravity (MAG) in light of the Dressing Field Method, a tool allowing to systematically build gauge-invariant field variables. To get MAG started, one has to deal with the problem of “gauge translations”. We first recall that Cartan geometry is the proper mathematical foundation for gauge theories of gravity, and that this problem never arises in that framework, which still allows to clarify the geometric status of gauge translations. Then, we show how the MAG kinematics is obtained via dressing in a technically streamlined way, which highlights that it reduces to a Cartan-geometric kinematics.

Introduction

In the mid 1970s was established [1] the now well-known fact that the mathematical foundation of Gauge Field Theory (GFT), à la Yang–Mills (YM), is the differential geometry of Ehresmann (or principal) connections on fiber bundles [2, 3]. General Relativity (GR) already being inherently a geometric theory, in the late 1970s and 1980s, treatments stressing the geometric structure of all fundamental interactions appeared, e.g. [46].

The field of gauge theoretic formulations of gravity has its roots in Einstein’s introduction, in 1925, of the local Lorentz group SO(1,3) and vielbein eaμ [7, 8], and in Weyl’s introduction of the local spin group SL(2,C) (acting on spinors) in the 1929 paper [9] in which he also introduces for the first time the Gauge Principle for U(1) [10]. But it started in earnest with Utiyama, who, in 1955, introduced (independently from the simple SU(2) model published in 1954 by Yang and Mills [11]) the general framework of GFT based on the general Gauge Principle for (i.e. of “gauging” of) an arbitrary Lie group G,  and thus applied it to the Lorentz group SO(1,3) to recover the structure of GR [12]. In the early 1960s, Kibble [13] and Sciama [14] (re-)introduced the Einstein–Cartan formulation, which allows for spacetime torsion sourced by spinor fields. Around that time, elucidation of the gauge structure of gravity was motivated, aside from deepening our understanding of it, by the expectations that it would facilitate its quantization – as quantization of other GFTs has been successful, notably in the Standard Model (SM) – and/or its unification with the other fundamental interactions. See e.g. [15]. The 1970s thus saw a considerable activity in model building based on the gauging of various groups, and supersymmetrization thereof leading to supergravity models – e.g. de Sitter groups by MacDowell and Mansouri [16], or the conformal group [17].

Gauge approaches to gravity have been dominated, understandably, by the heuristic habits of Yang–Mills theory: Meaning that whenever one wanted to “gauge” a Lie group G,  one postulated a gauge potential A with value in the Lie algebra g of G,  defined by its gauge transformations under the gauge group, supposed to be G; i.e. the group, under pointwise product, of maps g:UMG with M a (spacetime) manifold. Prominent examples of this are Poincaré gravity (PG), where G=SO(1,3)R4 [1822], and its Metric-Affine gravity (MAG) generalization to G=GL(n)Rn with n= dimM [2326]. See [27] for a bibliographic sample. In mathematical, bundle geometric terms, it means that the gravitational gauge potential is seen as (the local representative of, see after) an Ehresmann connection on a G-principal fiber bundle Q over M,  whose gauge group is G. However, such approaches feature systematically a group of “internal gauge translations”, conceptually redundant with the group of diffeomorphisms Diff(M), which gives rise to a set of issues that we shall review below. To even get started as theories of gravity, MAG and PG have to deal with these issues, which essentially implies to get rid of these gauge translations, or to try to identify them with Diff(M) (via the tetrad field/soldering form).

We have here two converging aims. First, we shall remind that a proper understanding of bundle geometry forbids such an identification (see our comment below Eq. (4) giving the short exact sequence characterizing the bundle Q), but most importantly we will highlight that the proper mathematical foundation of gauge theories of gravity is Cartan geometry [2831], i.e. the differential geometry of Cartan connection [32, 33] on fiber bundles, which captures the key Einsteinian insight about the nature of gravity and in which the issue of “gauge translation” simply does not arise. Second, we shall expose and use the Dressing Field Method (DFM) [3439], a systematic approach to building gauge-invariants in GFT, to eliminate gauge translations, thereby streamlining the construction of the basic MAG and PG kinematics. Doing so will stress that the latter are actually just the kinematics one would obtain by starting from Cartan geometry in the first place.

Considered together, these two goals result in our thesis that it is a priori misguided to attempt to build a gauge theory of gravity by “gauging” the affine or Poincaré groups – as this remains captive of the heuristic habits of YM theory, and overlooks the insight of Cartan geometry – but also that even doing so, MAG and PG actually cannot be understood as genuine gauge theories stemming from applications of the Gauge Principle to these groups. Indeed, we shall stress that the tinkering done to make it work – put on solid formal grounds via the DFM – leads back to Cartan geometry, where one should have started from the beginning.

The very same logic and conclusion apply e.g. to attempts to build theories of conformal gauge gravity by gauging the conformal group SO(2,4), as well as for supersymmetrization of all the aforementioned theories.1

The paper is thus structured as follows: In Sect. 2 we remind the bundle geometric structures underlying Yang–Mills type and gravitational gauge theories, emphasizing the distinct foundational role of Cartan geometry for the latter. Then, in Sect. 3, after briefly reviewing the gauge field-theoretic counterpart of the previously discussed global geometric structures, we present the key technical and conceptual aspects of the DFM. All this lays the groundwork for Sect. 4, where MAG (and PG) kinematics is obtained via the DFM, thereby streamlining its technical and conceptual foundations: in particular, we automatically reproduce the so-called “radius vector” usually introduced ad hoc to handle the issue of gauge translations. Finally, we further discuss the implications of the DFM approach to MAG and PG in Sect. 5, our main conclusion being that it only highlights that Cartan geometry is the sole sound foundation for gauge theories of gravity.

Cartan geometric foundation of gauge theories of gravity

The underlying geometry of GFTs à la YM is that of a principal bundle P,  a smooth manifold supporting the right action of a Lie group H (its structure group), P×HP, (p,h)ph=:Rhp, whose orbits are the fibers of P. The moduli space of fibers is itself a smooth manifold M:=P/H, so that there is a projection π:PM, pπ(p)=x, s.t. π(ph)=π(p). The bundle may be noted PM. Its tangent bundle TP contains thus the canonical vertical subbundle VPTP, defined as the kernel of the tangent application π:TPTM. The linearization of the right action of H defines the morphism of Lie algebras hVP, XXv, with h the Lie algebra of H and Xv a fundamental vertical vector induced by Xh.

The maximal group of automorphisms Aut(P) of P is the subgroup of its diffeomorphisms preserving its fiber structure – ΞDiff(P) s.t. Ξ(ph)=Ξ(p)h – i.e. mapping fibers to fibers. It thus induces by definition diffeomorphisms of M,  so that there is a surjective group morphism Aut(P)Diff(M). Its kernel is the normal subgroup Autv(P) of vertical automorphisms of P,  i.e. those automorphisms acting only along fibers and inducing the identity transformation on M,  i.e. idMDiff(M). These are thus generated by maps γ:PH with defining property γ(ph)=h-1γ(p)h, forming, under pointwise product, the gauge group H(P) of P: There is then a isomorphism Autv(P)H(P), given by Ξ(p)=pγ(p).2 The geometry of P is thus characterized by the short exact sequence (SES) of groups

idPAutv(P)H(P)Aut(P)π~Diff(M)idM. 1

The local structure of P is trivial, i.e. for UM it is the case that P|UU×H. Yet, in general PM×H.

As a manifold, P has a de Rham complex of forms (Ω(P),d), with d the exterior derivative. The equivariance of a form βΩ(P) is defined by the pullback action of the structure group, Rhβ. In particular, given a representation (ρ,V) of H,  one defines equivariant forms βΩeq(P,V) whose equivariance is controlled by the representation: Rhβ=ρ(h)-1β. The gauge transformation of a form β is defined by the pullback action by Autv(P), which, given the above isomorphism, is expressible in terms of the corresponding generating elements of H(P): βγ:=Ξβ. An important case is that of tensorial forms αΩtens(P,V), whose gauge transformations are homogeneous and controlled by the representation: αγ:=Ξα=ρ(γ)-1α. We remark that, in particular, gauge group elements are both equivariant 0-forms, Rhγ=h-1γh, and tensorial 0-forms acting on each other by ηγ:=Ξη=γ-1ηγ. In physics, tensorial 0-forms φΩtens0(P,V) represent matter fields.

Unfortunately, d does not preserve tensorial forms, i.e. dαΩtens(P,ρ). To get a first order differential operator on Ωtens(P,ρ), one needs to introduce an Ehresmann connection on P: that is ωΩeq1(P,h), with defining properties

(i)Rhω=Ad(h-1)ω=h-1ωh,and(ii)ω(Xv)=X. 2

These properties implies that ω gauge transforms inhomogeneously: ωγ=γ-1ωγ+γ-1dγ. It also implies that it induces a covariant derivative on tensorial forms, D_=d_+ρ(ω)_:Ωtens(P,V)Ωtens+1(P,V). This reflects the Gauge Principle, or gauge argument, of GFT. One shows that D2α=DDα=ρ(Ω)α, where Ω=dω+12[ω,ω]=dω+ω2Ωtens2(P,h) is the curvature of ω. It thus gauge transforms as Ωγ=γ-1Ωγ, and satisfies the Bianchi identity DΩ=dΩ+Ad(ω)Ω=dΩ+[ω,Ω]=0.

The above, when restricted to M,  or UM, provides the complete kinematics of a YM gauge field theory, as we shall review briefly in next section, the only missing ingredient being a Lagrangian providing a specific dynamics.

The mathematical underpinning of gauge theories of gravity is Cartan geometry; see e.g. [31] for a recent review, and [2830] for in depth treatments, while [33, 44] are of historical interest. A Cartan geometry (P,ω¯) is an H-principal bundle P endowed with a Cartan connection ω¯Ωeq1(P,g), with gh s.t. g/h=:p is a left H-module, satisfying the same two defining properties (2) of an Ehresmann connection, but also a third distinctive one:

(iii)ω¯:TPgis a linear isomorphism, i.e.ker(ω¯)=. 3

From this single key property stems the specificities of Cartan geometry, which essentially boil down to the fact that it ensures P encodes the geometry of M. Given the general-relativistic insight that gravity is the geometry of spacetime, Cartan geometry is thus perfectly adapted to describe the kinematics of gauge theories of gravitation, the Cartan connection ω¯ representing a generalized gravitational gauge potential.

We emphasize that since a Cartan connection ω¯ satisfies, like an Ehresmann connection ω, the properties (2), it transforms under the gauge group H(P) as ω¯γ=γ-1ω¯γ+γ-1dγ: in other words, there is no gauge transformation “associated to” p. Since in many cases pRn, there is no “internal gauge translations” in Cartan geometry.

Technically, ω¯ induces soldering on M,  i.e. the tangent bundle TM is isomorphic to the associated vector bundle to P with fiber p: we write TMP×Hp. In other words, vector fields on M are represented by Ωtens0(P,p), and all tensors on M are likewise represented by forms on P. Relatedly, given τ:gp, a Cartan connection induces a soldering form θ:=τ(ω¯)Ωtens1(P,p), which thus H(P)-transforms as θγ=γ-1θ. The curvature of ω¯ is Ω¯:=dω¯+12[ω¯,ω¯]=dω¯+ω¯2Ωtens2(P,g), and thus H(P)-transforms as Ω¯γ=γ-1Ω¯γ. It satisfies the Bianchi identity D¯Ω¯=dω¯+[ω¯,Ω¯]=0. The torsion of ω¯ is Θ:=τ(Ω¯)Ωtens2(P,p), and H(P)-transforms as Θγ=γ-1Θ. Manifestly, soldering and torsion are notions inexistent for Ehresmann connections, which on the upside accounts for the fact that Ehresmann geometry (P,ω) allows to describe an “enriched structure” over M,  unrelated to M’s intrinsic geometry, and is thus a perfect fit for YM type GFTs. See [45] for a discussion of this point.

A Cartan connection is normal if it is entirely expressed in term of its soldering part: ω¯N=ω¯N(θ). Typically this at least implies that Θ=0. This generalizes the notion of Levi-Civita connection. In reductive or parabolic Cartan geometries, one has a H-invariant decomposition g=hp, so ω¯=ω+θ where ωΩeq1(P,h) is an Ehresman connection on P. Correspondingly, Ω¯=Ω+Θ, but ΩΩtens2(P,h) in general is not ω’s curvature, yet contains it.

A flat Cartan geometry (P,ω¯) is isomorphic to a Klein geometry (G,ω¯MC), where G is a Lie group with Lie algebra g and closed subgroup H so that it is an H-bundle over the homogeneous space M:=G/H, i.e. GM. The Maurer–Cartan form ω¯MCΩeq1(G,g) satisfies (i)–(iii) and dω¯MC+12[ω¯MC,ω¯MC]=0; thus is a flat Cartan connection. In other words, Cartan flatness implies that the manifold M becomes (isomorphic to) a homogeneous space M, thus generalizing Riemann flatness which implies that M becomes (isomorphic to) the homogeneous space M=Rn.

On M,  or UM, the local representative of ω¯ and Ω¯ are A¯ and F¯, which describe respectively the gravitational gauge potential and its field strength. The local representative of θ is eΩ1(U,p), which is none other than a “vielbein”. Given an H-invariant non-degenerate bilinear form η:p×pR, the Cartan connection thus induces a metric on M by ge:Γ(TM)×Γ(TM)R, (X,Y)g(X,Y):=η(e(X),e(Y)).

There is a relation between Cartan and Ehresmann geometry. Suppose the H-bundle P of the Cartan geometry (P,ω¯) can be embedded as a subbundle of a bundle QM with structure group GH, ι:PG , and with SES

idQAutv(Q)G(Q)Aut(Q)π~Diff(M)idM, 4

where G(Q) is the gauge group of Q,  whose elements are maps γ¯:QG with defining property γ¯(pg)=g-1γ¯(p)g. Observe that it contains gauge transformations “associated to” p, that is “gauge translations” in case p=Rn. It should be clear from the SES (4) that these can in no way be identified with Diff(M), contrary to what is often stated in the literature,3 as the whole gauge group G(Q) maps to idM.

An Ehresmann connection on Q is ϖΩeq1(Q,g) satisfying, mutatis mutandis, the two defining properties (2). It therefore transforms under the gauge group G(Q) as ϖγ¯=γ¯-1ϖγ¯+γ¯-1dγ¯, thus in particular under gauge transformations corresponding to p (“gauge translations”). An Ehresmann connection ϖ on Q induces by restriction a Cartan connection ω¯:=ιϖ on Pprovided the condition ker(ϖ)ι(TP)= is met – so that (iii) holds on P. Reciprocally, a Cartan connection ω¯ on P induces an Ehresmann connection ϖ on Q satisfying this condition. See [29], appendix A.3. The space of Ehresmann connections on Q thus contains the space of Cartan connections on P.

Locally, on UM, the local representatives of both ϖ and ω¯ are forms A¯Ω1(U,g), and the only way to distinguish them is by how they transform: A¯ is the local representative of ϖ on Q if it transforms under the local version G of G(Q), i.e. with maps γ¯:UG, and it is the local representative of ω¯ on P if it transforms under the local version H of H(P), i.e. with maps γ:UH (see Eq. (5) below for a more precise statement).

Gauge theories with gauge group G, as MAG and PG purport to be, are actually concerned with the geometry of the bundle (Q,ϖ) characterized by the SES (4), which is not a Cartan geometry and therefore not the proper framework for a gauge theory of gravity. Yet, as we shall demonstrate, MAG and PG are actually no such theories, as their kinematics involves the elimination of gauge translations, which we shall perform systematically via the Dressing Field Method presented next, thereby ending up with a kinematics with gauge group H stemming from the Cartan geometry (P,ω¯), as befitting gauge theories of gravity.

The Dressing Field Method of gauge symmetry reduction

The DFM [3439] is a systematic tool to produce gauge-invariant variables out of the field space Φ of a theory with gauge group H whose action on Φ defines gauge transformations. Let us briefly review its key aspects, to better appreciate the content of what comes next. We start with the local, field theoretic, version of the geometric structures discussed above.

Consider a gauge theory over an n-dimensional manifold M,  or region UM, based on a (finite-dimensional) Lie group H,  the structure group of the underlying principal bundle PM, with Lie algebra h. Its elementary variables are: A YM gauge potential 1-form A=AμdxμΩ1(U,h), with field strength 2-form F=dA+12[A,A]=dA+A2Ω2(U,h). They are respectively the local representatives of an Ehresmann connection ω and its curvature Ω on P. Alternatively (or in addition) if one is considering a gauge theory of gravity, a gravitational gauge potential 1-form is A¯=A¯μdxμΩ1(U,g), with curvature 2-form F¯=dA¯+12[A¯,A¯]=dA¯+A¯2Ω2(U,g). They are respectively the local representatives of a Cartan connection ω¯ and its curvature Ω¯ on P. Often the potential splits as A¯=A+e, where e=eaμdxμΩ1(U,p) is a soldering form. Matter fields are ϕΩ0(U,V), with V a representation space for H,  i.e. ρ:HGL(V), and ρ:hgl(V). Their minimal coupling to gauge potentials is given by the covariant derivative, Dϕ:=dϕ+ρ(A)ϕΩ1(U,V). One shows that D2ϕ=ρ(F)ϕ.

These fields are acted upon by the (infinite-dimensional) gauge group of the theory: the set of H-valued functions γ:UH, xγ(x), with point-wise group multiplication (γγ)(x)=γ(x)γ(x), defined by

H:=γ,η:UH|ηγ:=γ-1ηγ. 5

This action defines the gauge transformations of the fields,

Aγ:=γ-1Aγ+γ-1dγ,αγ:=ρ(γ)-1α, 6

where we designate collectively α={F,F¯,e,ϕ,Dϕ}, which are all “gauge-tensorial” and H-transform respectively via the adjoint (field strength), left (soldering), and ρ (matter fields and their covariant derivatives) representations, that we denote collectively ρ={Ad,,ρ}.

The Lagrangian of a theory is a top form L(A,e,ϕ)Ωn(U,R) required to be quasi-invariant under the action of H, i.e. L(Aγ,eγ,ϕγ)=L(A,e,ϕ)+db(γ;A,e,ϕ), so that the field equations E(A,e,ϕ)=0 are H-covariant: E(Aγ,eγ,ϕγ)=ρ(γ)-1E(A,e,ϕ)=0.

Now, consider a subgroup K of the structure group HKH, to which corresponds the gauge subgroup KH. A K-dressing field is a map u:UMK, i.e. a K-valued field, defined by its K-gauge transformation:

uκ:=κ-1u,forκK. 7

Given the existence of a K-dressing field, one defines the K-invariant dressed fields

Au:=u-1Au+u-1du,αu:=ρ(u)-1α. 8

In particular, a dressed field strength is Fu=u-1Fu=dAu+12[Au,Au], i.e. it is the field strength of the dressed potential. A dressed matter field is ϕu=ρ(u)-1ϕ, and (Dϕ)u=ρ(u)-1Dϕ=dϕu+ρ(Au)ϕu=:Duϕu, i.e. the dressed covariant derivative is the minimal coupling of the dressed matter field with the dressed gauge potential.

Noticing the formal similarity with the gauge-transformations (6), one sees the simplest case of the DFM “rule of thumb”: To obtain the dressing of an object (field or functional thereof), first compute its gauge transformation, then substitute in the resulting expression the gauge parameter γ with the dressing field u. The dressed object is K-invariant by construction.

Note, however, that the dressing field is not an element of the gauge group, uK, as it can be immediately seen by comparing (5) and (7). This is a crucial fact of the DFM: Despite the formal analogy with (6), the dressed fields (8) are not gauge transformations. Hence, {Au,αu} must not be confused with a gauge-fixing of the bare variables {A,α}. Contrary to a gauge-fixing, the dressing operation is not a map from Φ to itself, but from Φ to the space of dressed fields Φu, only isomorphic to a subspace (a subbundle) of Φ – cf. [35, 46] for details. Clearly, when u is an H-dressing field, s.t. uγ=γ-1u, the dressed fields are H-invariant.

A key aspect of the DFM is that a dressing field should be extracted/built from the (bare) field content ϕ={A,α} of the theory, i.e. u=u[ϕ], so that u[ϕ]κ:=u[ϕκ]=κ-1u[ϕ]. In such a case the dressed field ϕu[ϕ] have a natural interpretation as relational variables: they encodes the gauge-invariant relations among the physical degrees of freedom (d.o.f.) embedded redundantly in the bare fields (among pure gauge modes). The relational aspect of the theory is encoded in the gauge symmetry of the bare theory, which is then said substantive, see [35, 47]. If a dressing is introduced by fiat, as additional d.o.f., one is dealing with a new, distinct theory: The dressing field is ad hoc, not built from the original d.o.f.; the dressed fields thus cannot be interpreted as representing the physical, relational content of the original theory. The gauge symmetry of the new theory is said to be artificial [47].4

Residual transformations

Being K-invariant, the dressed fields (8) are expected to display residual transformations under what remains of the gauge group. For these to be well-defined, it must be that K is a normal subgroup of HKH, so that H/K=:J is again a Lie group. Correspondingly then, KH and J=H/K is a gauge subgroup of H. In this case, the dressed fields may exhibit well-defined residual J-gauge transformations, which are named residual transformations of the 1st kind. Now, since the J-transformations of the bare fields are known, as a special case of their H-transformations given by (6), to find the residual J-transformations of the K-invariant dressed fields (8) one only needs to determine that of the dressing field u. An interesting case is given by the following

Proposition 1

If a K-dressing field transforms as uη=η-1uη for ηJ, then the dressed fields (8) are standard J-gauge fields with J-gauge transformations

graphic file with name 10052_2025_14656_Equ9_HTML.gif 9

In particular, the dressed curvature transforms as (Fu)η=η-1Fuη, and a dressed matter field and its dressed covariant derivative as (ϕu)η=ρ(η-1)ϕu and (Dϕu)η=ρ(η-1)Dϕu.

Dressed variables may also be subject to residual transformations stemming from a possible “ambiguity” in the choice of the dressing field: Two dressing fields uu are a priori related by u=uζ, for ζ:UK a map s.t. ζκ=ζ. Under pointwise product, these maps form a group we denote G and call the group of residual transformations of the 2nd kind. Its action on a dressing we may denote uζ:=uζ, so that, while it does not act on bare fields ϕζ:=ϕ, it does act naturally on dressed ones as (ϕu)ζ:=ϕuζ. Explicitly, we have

Proposition 2

If the K-dressing field transforms as uζ=uζ for ζG, then the dressed fields (8) G-transform as

graphic file with name 10052_2025_14656_Equ10_HTML.gif 10

With in particular,  (Fu)ζ=ζ-1Fuζ, while (ϕu)ζ=ρ(ζ-1)ϕu and (Dϕu)ζ=ρ(ζ-1)Dϕu.

However, when u=u[ϕ] is built from the fields, the constructive procedure may typically reduce G to a discrete group, reflecting the finite choices among the d.o.f. available. This happens e.g. in the context of classical and quantum mechanics, where residual transformations of the 2nd kind (10) are shown to encode physical reference frame covariance [48]. No such restriction naturally exists for ad hoc dressing fields: in that case, the action of G on ϕu essentially just replicates the action of K on ϕ, the two situation are isomorphic.

Dressed dynamics

As is clear from what precedes, a dressing is an operation performed at the kinematical level, turning the bare kinematics into a dressed one. Regarding dynamics, there are two possibilities to consider.

First, suppose that, following a Gauge Principle for the gauge group H, one had built an H-quasi-invariant Lagrangian for the bare variables L(A,e,ϕ), with bare field equations E(A,e,ϕ)=0, defining an H-gauge theory. Then, given a dressing field u,  one defines the dressed Lagrangian by L(Au,eu,ϕu):=L(A,e,ϕ)+db(u;A,e,ϕ) – again seen to be a case of the DFM rule of thumb – from which derives dressed field equations E(Au,eu,ϕu)=0, i.e. equations for the dressed fields, which have just the same functional expression as the bare field equations. In case Proposition 1 obtains, L(Au,eu,ϕu) is J-quasi-invariant, and the dressed field equations E(Au,eu,ϕu)=0 are J-covariant. The H-theory is thus seen to be reduced to a J-gauge theory.

Alternatively, one may start from the dressed kinematics, considered as the only physically relevant one, and take advantage of the greater freedom afforded by following a Gauge Principle for the residual gauge subgroup J. That is, one may define a J-gauge theory by building a J-quasi-invariant Lagrangian L(Au,eu,ϕu) from which derive J-covariant field equations for the dressed fields E(Au,eu,ϕu)=0. Such a theory, in general not H-quasi-invariant, obviously cannot be considered to follow from a standard application of the Gauge Principle to H.

As it turns out, this second strategy is the one followed by model building in MAG and PG, as we stress below. So, contrary to heuristic claims often encountered in the literature, neither MAG nor PG follow straightforwardly from gauging of the affine and Poincaré groups.

Reduction of gauge translations in MAG via the DFM

We shall now apply the DFM to MAG. What follows essentially applies mutatis mutandis to PG, i.e. by simply substituting the general linear group GL(n) by the Lorentz group SO(1,3) – or its spin cover Spin(1,3)SL(2,C). Let us start with reminding its a priori kinematical setup, introducing a compact matrix notation.

MAG starts with the gauging of the affine group GL(n)Tn, whose elements are pairs (G,t) with composition law (G,t)·(G,t)=(GG,t+Gt). The neutral element is (1,0), so the inverses are (G,t)-1=(G-1,-G-1t). By definition of a semidirect product group, the subgroup Tn is normal in GL(n)Tn, i.e. (G,t)-1·(1,t)·(G,t)Tn. This group can be embedded in GL(n+1): we have the injective group morphism

GL(n)TnGL(n+1),(G,t)g=Gt01,(G,t)-1g-1=G-1-G-1t.01, 11

such that indeed the semidirect group law of the affine group is reproduced by simple matrix multiplication,

gg=Gt01·Gt01=GGGt+t.01. 12

The associated gauge group is GL(n)Tn:={γ:UGL(n)Tn|}, similarly to the general case (5). Abusing notations slightly, we shall write gauge elements as γ=Gt01. So, the gauge transformation of gauge elements is

ηγ=γ-1ηγ=Gt01-1·Gt01·Gt01=G-1GGG-1t01, 13

which indicates in particular that the gauge elements t are GL(n)-tensorial and Tn-invariant, Tn-valued function. Correspondingly, an element of the gauge subgroup GL(n) is

G=G001,with inverseG-1=G-1001, 14

while an element of the additive Abelian normal gauge subgroup Tn is given by the upper triangular matrix

T=1t01,with inverseT-1=1-t.01. 15

Using this matrix embedding, the gauge potential of MAG and its field strength are

A¯=AV00,andF¯=dA¯+A¯2=FT00=dA+A2dV+AV00, 16

where V is the gauge potential of translations, A that of general linear rotations, and T is the “torsion” 2-form of A. As the potential A¯ is being seen as the local, field-theoretical representatives of an Ehresmann connection on the G=(GL(n)Tn)-bundle Q over M – rather than of a Cartan connection on the H=GL(n)-bundle P over M – the fields (16) transform under the gauge group GL(n)Tn as

A¯γ=γ-1A¯γ+γ-1dγ=Gt01-1AV00Gt01+Gt01-1dGdt01=G-1AG+G-1dGG-1(V+Dt).00,F¯γ=γ-1F¯γ=Gt01-1FT00Gt00=G-1FGG-1(T+Ft).00, 17

where Dt:=dt+At is the covariant derivative of the GL(n)-tensorial and Tn-invariant Tn-valued function t. In particular, this specializes to give the transformation of the MAG potential and field strength under “internal gauge translations”,

A¯T=T-1A¯T+T-1dT=AV+Dt.00andFT=T-1FT=FT+Ft.00. 18

We may also observe that the fundamental representation of the affine group is RnTn, the corresponding right action of the gauge group on XΩ0(U,Rn) is X(G,t)-1X=G-1(X-t), or using the matrix embedding,

X¯:=X1X¯γ=γ-1X¯=G-1(X-t)1. 19

The covariant derivative induced by A¯ is thus,

D¯X¯=dX¯+A¯X¯=DX+V1,and s.t.(D¯X¯)γ=D¯γX¯γ=γ-1D¯X¯, 20

with DX:=dX+AX. One furthermore shows that D¯2X¯=F¯X¯=FX+T1. The object X¯Ω0(U,Rn+1) is the local representative of a tensorial 0-form on the (GL(n)Tn)-bundle QM, and can equally well be understood as a the section of an associated bundle P×GL(n)TnRn+1. Typically in gauge theory, these represent matter fields, or “Higgs” fields if the Lagrangian of the theory features a potential term V(X¯). A priori, one may attempt to see X¯, i.e. X,  as the spacetime velocity of a material point particle on M. However, there is obstruction to such an interpretation.

The very existence of “internal” gauge translations is a problem. First, and most notably, they are redundant conceptually with diffeomorphisms Diff(M). And yet, contrary to what is often claimed in the MAG and Poincaré gravity literature,5 they can bear no relation to them because, as we stressed at the end of Sect. 2 and as the SES (4) makes clear, Tn as a gauge subgroup acts trivially on M,  i.e. induces the identity of Diff(M).

Then, Tn makes impossible to identify the most basic objects in the fundamental representation XΩ0(U,Rn) with (components of) vector fields XΓ(TM) of M;  it is indeed clear by (19) that while the former are GL(n)-tensorial, as expected from vector field components, they are not Tn-invariant. Consequently, the covariant derivative DX in D¯X¯ cannot be understood as the covariant derivative of a vector field on M – and D¯X¯=0 is not a geodesic equation in M.6 Idem for objects in the dual of the fundamental representation XΩ0(U,Rn), which cannot be identified with (components of) covectors, 1-forms, on M. So that, in general, tensorial objects built from tensoring X and X are not related to tensors of M.

Finally, gauge translations make impossible the identification of the translation potential VΩ1(U,Tn) with a soldering form inducing a metric on M (and consequently make unclear the relation between T and a true torsion tensor on M): Indeed, given a GL(n)-invariant non-degenerate bilinear form η:Tn×TnR, if one tries to define a metric as g:=ηV:Γ(TM)×Γ(TM)R, X,Yg(X,Y):=η(V(X),V(Y)), then this metric is not Tn-invariant.

Thus, to even get started with MAG as a gravity theory, one must somehow get rid of the gauge subgroup Tn.7 The DFM provides just the systematic framework that allows to do so naturally.

Dressed connection and curvature

A dressing field for the gauge subgroup Tn of “internal translations” is a map u:UMTn defined by uT=T-1u, for TTn. Using again the matrix embedding, we write

u:=1ξ01,withξΩ0(U,Tn),s.t.uT=T-1u=1-t.011ξ01=1ξ-t.01. 21

We stress that, without the matrix embedding, one might have defined the dressing for Tn directly by ξ:UTn s.t. ξt=ξ-t, as an additive Abelian version of the general definition of a dressing field.

Given a dressing field as above, applying (8), we easily build the Tn-invariant dressed potential

A¯u:=u-1A¯u+u-1du=1-ξ01AV001ξ01+1-ξ010dξ00=AV+Dξ.00=:Ae.00, 22

where Dξ:=dξ+Aξ. Correspondingly, the Tn-invariant dressed field strength, i.e. the field strength of A¯u, is

F¯u:=u-1F¯u=1-ξ01FT001ξ01=FT+Fξ00=:FΘ00. 23

Comparison of (22)–(23) with (18) illustrates the DFM rule of thumb. We observe that the Tn-invariant dressed field e:=V+DξΩ1(U,Tn) in (22) is called the “key relation” of MAG in [23].8 Finally, one can build the dressed 0-form and its dressed covariant derivative

X¯u=u-1X¯=X-ξ1=:Xξ1andD¯uX¯u=dX¯u+AuX¯u=DXξ+e1, 24

with DXξ=dXξ+AXξ. Now, Xξ=X-ξΩ0(U,Rn), being Tn-invariant, is potentially identifiable as a vector field on M if it retains the correct GL(n)-tensorial transformation of its bare counterpart X. Similarly, the Tn-invariant form e:=V+DξΩ1(U,Tn) is a soldering form, and Θ=De=de+Ae is a true torsion 2-form on M,  only if both are GL(n)-tensorial. To ascertain these questions, we must assess the residual transformations of the 1st kind of the above dressed fields.

Residual GL(n) transformations

After reducing the normal gauge subgroup Tn via dressing, we expect residual transformations of the 1st kind under GL(n). As per the general explanations of Sect. 3, as we already know the GL(n)-transformations of the bare variables by (17) and (19), we need only find that of the dressing field u.

By assumption, the dressing field ξ is in the fundamental representation of GL(n), so that ξξG:=G-1ξ. Using the matrix embedding, and (14), this is

uG=G-1uG=G-10011ξ01G001=1G-1ξ01. 25

This is a special case of Proposition 1, which allows us to immediately conclude that the dressed fields are standard GL(n)-gauge fields, so that by (9) we have:

(A¯u)G=G-1A¯uG+G-1dG=G-1AG+G-1dGG-1e.00,(F¯u)G=G-1F¯uG=G-1FGG-1Θ.00,(X¯u)G=G-1X¯u=G-1Xξ1,and(DuX¯u)G=G-1D¯uX¯u=G-1(DXξ+e)1. 26

The full group of local transformations of the dressed MAG kinematics is thus Diff(M)GL(n). From (26) it is now clear that Xξ is indeed identifiable with a vector field X of M,  and can now represent the spacetime velocity of a point particle on M. So, DXξ in D¯uX¯u is the covariant derivative of vector fields, and DXξ=0 describes a geodesic on M. Tensors obtained from tensoring Xξ and Xξ are true tensors of M.

Furthermore, it is clear that A¯u is but the (local representative of a) Cartan connection associated to a Cartan-affine geometry, with curvature F¯u, inducing via eΩ1(U,Tn), a true soldering form on M,  a gauge-invariant metric on M by g:=ηe:Γ(TM)×Γ(TM)R, X,Yg(X,Y):=η(e(X),e(Y)). We have now a good kinematics for a gauge theory of gravity: But it is just the local version of the Cartan-affine geometry (P,ω¯), with PM a H=GL(n)-principal bundle (i.e. the frame bundle of M), we would have started with had we heeded the insight of Cartan geometry.

Discussion

Several observations and comments are in order. First, we did not yet say how the Tn-dressing field u (21) is to be found: in the DFM, the physical picture changes significantly depending if the dressing is field-dependent or not.

If it is introduced as a separate object from the bare GL(n)Tn kinematics, i.e. as extra d.o.f., then it is what in Sect. 3 we called an ad hoc dressing field. In that form it reproduces what is known as the “radius vector”, e.g. mentioned early in [23] (and attributed to Trautman [56]), an object indeed introduced in MAG to get rid of gauge translations Tn. But according to the DFM, this means that in MAG (and PG), which is thus the bare GL(n)Tn kinematics supplemented by an ad hoc dressing field u/radius vector, Tn is an artificial gauge symmetry – also called “fake” gauge symmetry by [57] – with no physical signature.

Only the residual GL(n) (or SO(1,3) in PG) gauge group has physical significance and is thus substantive. As a matter of fact, and as observed already at the end of Sect. 3 (and footnote 7), model building in MAG (and PG) starts only after eliminations of gauge translations Tn and the Lagrangians are only required to be (quasi-) invariant under the residual GL(n)-transformations.9 Usually, no invariance property under Tn is required. It is thus incorrect to claim that MAG, or PG, are built from “gauging” the affine or Poincaré groups à la Yang–Mills.10 As we just showed above, at best MAG (and PG) kinematics is just the kinematics of Cartan-affine geometry.

Looking at it the other way around, we see that one may have started with a Cartan-affine kinematics, i.e. A¯ and F¯ supporting GL(n)-gauge transformations like (26). Then, we may enforce an artificial “gauge translations” group Tn, acting trivially on A¯ and F¯, by introducing a Stueckelberg field u-1:UTn (i.e. -ξ) s.t. (u-1)T=u-1T, and then defining the fields A¯:=uA¯u-1+udu-1 and F¯:=uF¯u-1, transforming under Tn as (18): i.e. we end-up with a GL(n)Tn kinematics where Tn is “fake”. Clearly, A¯=A¯u and F¯=F¯u; as stressed earlier, the DFM encompasses Stueckelberg tricks when dressing fields are ad hoc.

Furthermore, it could be argued that since in MAG/PG the dressing field is ad hoc, according to Proposition 2 and Eq. (10), the dressed fields (22)–(24) may a priori support G-transformations of the 2nd kind, which all but reproduce the action (18)–(19) of Tn on bare variables.

A way out is to notice that in the field content Inline graphic there is a natural candidate for a Tn-dressing field: we may indeed define

graphic file with name 10052_2025_14656_Equ27_HTML.gif 27
graphic file with name 10052_2025_14656_Equ28_HTML.gif 28

Said otherwise, this is the field-dependent dressing field ξ=ξ[X¯]=X. It allows to define the Tn-invariant fields

A¯u[X¯]=AV+DX.00=:Ae.00andF¯u[X¯]=FT+FX.00=:FΘ.00 29

by (22)–(23), as well as

X¯u[X¯]=u[X¯]-1X¯=01and(D¯X¯)u[X¯]=dX¯u[X¯]+Au[X¯]X¯u[X¯]=e1, 30

similarly to (24). Their residual GL(n)-transformations are given by (26). The fields (29)–(30) may be understood as relational variables encoding the Tn-invariant relations among the “internal translational” d.o.f. of A¯ and X¯. It is indeed consistent with the a priori postulate of a GL(n)Tn kinematics, based on the bundled QM, that fields would have internal translational d.o.f. – which are still entirely unrelated to Diff(M), as noted earlier. The object X¯ would then describe the “generalised spacetime velocity” of a point particle with such translational internal d.o.f. and X¯u[X¯]=(0,1) simply expresses that such a particle “sees” itself at rest in its own reference frame.11

Pushing the idea a step further, suppose we have a collection of N such point particles, so ϕ={A¯,F¯,X¯1,X¯N}. Then, clearly, we have N choices to define a dressing field ui=u[X¯i], i{1,N}, giving rise to dressed fields ϕui={A¯ui,F¯ui,X¯1uiX¯iuiX¯Nui}, which are the relational variables describing the Tn-invariant relations among internal d.o.f. within ϕ as seen from the frame of X¯i. The change of particle perspective – i.e. of “physical” reference frame – is encoded by residual transformations of the 2nd kind, where G is the discrete group of elements ζij s.t. uj=uiζij, so that ϕuj=ρ(ζij)-1ϕui. This is analogous to the application of the DFM in non-relativistic classical and quantum mechanics [48].

This view is not so bad, it could be interesting if one was in the business of writing theories for the affine gauge group GL(n)Tn, in which case one might expect empirical consequences to the presence of the symmetry Tn. That could also be the case e.g. if, instead of seeing X¯ as a type of matter field, one was trying to interpret it as a sort of “gravitational Higgs field”, by embedding it into an invariant potential V(X¯), possibly leading to Higgs-type mechanism.12 The issue, as stated earlier, is that this is not what MAG/PG model building is usually about, since it starts only after kinematical elimination of Tn via dressing, and is constrained only by (quasi-)invariance under residual GL(n)-transformations. These approaches have no observable consequences associated to gauge translations, so one is only really concerned by the physics underpinned by Cartan geometry.

So, even if superficially MAG and PG approaches seemed to be taking a road to gauge gravity distinct from Cartan geometry, by gauging the affine or Poincaré groups à la Yang–Mills – implying to consider A¯ as the local representative of an Ehresmann connection on the G-bundle QM – the actual practice to get them started, involving the reduction of the gauge translation group Tn via the DFM, circles back to Cartan geometry and only highlights it as the sole sound foundation of classical gauge theories of gravity.

Funding

This research work was supported by the Austrian Science Fund (FWF), grant [P36542], by the Czech Science Foundation (GAC̆R), grant GA24-10887S, and by the GrIFOS research project, funded by the Ministry of University and Research (MUR, Ministero dell’Università e della Ricerca, Italy), PNRR Young Researchers funding program, MSCA Seal of Excellence (SoE), CUP E13C24003600006, IDSOE2024_0000103.

Data Availability Statement

This manuscript has no associated data. [Author’s comment: Data sharing not applicable to this article as no datasets were generated or analysed during the current study.]

Code Availability Statement

This manuscript has no associated code/software. [Author’s comment: Code/Software sharing not applicable to this article as ne code/software was generated or analysed during the current study.]

Footnotes

1

Relatedly, but distinct from supergravity model building, conceptual confusion about which group to gauge and misunderstanding of Cartan geometry are the root causes leading to obstacles in applying the supersymmetric framework to obtain a general approach to a unified description of matter and interaction fields – beyond the special 3D “unconventional susy” (or AVZ) model, first proposing it [4042]. Attempts at building models of such a framework in 4D stumbled upon the issue of having to make sense of “internal gauge translations” and to identify the translation potential with a separately postulated soldering form, leading to uncompelling efforts to justify a “double metric structure”. These matters are addressed and solved via the DFM in [37, 43].

2

One may indeed check the defining automorphism property Ξ(ph)=(ph)γ(ph)=, showing the defining property of γ to be essential.

3

For example, early in the well-known review [23] one reads “On the face of it, local diffeomorphisms can be considered as locally gauged translations [...]”. And in footnote 12, it is claimed that “the group [of gauge translations] is locally isomorphic to the group of active diffeomorphisms”.

4

This case encompasses the so-called Stueckelberg trick, whereby one implements a gauge symmetry in a theory via the introduction of extra d.o.f., the Stueckelberg fields: It is clear that the latter are ad hoc dressing fields, and what we described above is the inverse procedure of a Stueckelberg trick (which is thus an “undressing” operation) [35]. More broadly, the DFM underlies the so-called “Massive Yang–Mills” models and, as a framework, it encompasses both the electroweak model and Stueckelberg-type models [35, 47].

5

See again footnote 3: one finds attempts to justify such claims by heuristic arguments, e.g. in [23] that “this view gains some additional justification from the fact that the gravitational field is coupled to the energy-momentum tensor density, i.e. to the translational current.

6

Had we dealt with PG, e.g. in n=4 dimension, instead of MAG, we would have found relatedly that the object ψΩ0(U,C2) in the spin cover of the fundamental representation cannot be understood as a spinor field on M,  for it lacks invariance under gauge translations: so fermionic matter is not naturally represented. Nor is its minimal coupling to gravity, which is not DψΩ1(U,C2).

7

That much is clear from [23] where there is indeed no mention of a Lagrangian or field equations for MAG of PG before the issue of gauge translations is dealt with.

8

Comparable results have been achieved in [49] via the notion of “nonlinear realisations/representations” (NR), as defined e.g. in [5052]. One key difference between the DFM and NR is that, while in the former a dressing field is group-valued, the “dressing factor” in NR is coset-valued. Furthermore, the DFM having a clear bundle geometric formulation, a dressing field (like all geometric objects) has a clean geometric equivariance under the gauge subgroup being eliminated (i.e. a “linear”, covariant, transformation), while the “non-linear” transformation of the coset-valued field of NR seems non-geometric. The closest bundle geometric result that seems to connect to NR (that [49] indeed appears to be hinting at) is the Bundle Reduction Theorem (BRT) – see e.g. [4, 5355]. As detailed in [34], the DFM and the BRT can coincide when the structure group of the bundle under consideration is a (semi-)direct product of two subgroups. Which is precisely the case here.

9

That much is hinted at early in [23] where we read e.g. that “Only after a certain reduction, the translational connection and curvature are converted into coframe and torsion, respectively”[...]”, in line with our comments after (24) and (26) above.

10

For example, [23] insists on considering “[Poincaré] gravitational theories from the point of view of a Yang–Mills like gauging of the Poincaré group.”, while its section 2.7 is entitled “Metric-affine gauge theories: gauging the [affine group] [...]”. The preface of [27] states that “If one applies the gauge-theoretical ideas to [the Poincaré group], one arrives at the Poincaré gauge theory of gravity (PG)”. In the forewords of [27], Kibble states “applying the gauge principle to [the] Poincaré-group symmetries leads most directly not precisely to Einstein’s general relativity, but to a variant, originally proposed by Élie Cartan, which [...] uses a spacetime with torsion.

11

Something exactly analogous to (27)–(30) can be done in conformal Cartan geometry, and conformal gravity, where a dressing field u[Y¯] may be built from the dilaton embedded in the tractor field Y¯Ω0(U,R6) in the fundamental representation of HG=SO(2,4), and used to reduce Weyl gauge rescalings; this allows in particular to produce Dirac spinors from twistors [58].

12

The review [23], borrowing from Trautman [4, 56], suggests to understand ξ as a “generalized Higgs field”. As we noted above, this terminology would be apt only if a potential term appears in the Lagrangian. We remark that [58] did implement such an idea in the context of conformal Cartan gravity mentioned in footnote 10, treating the tractor field as a Higgs field embedded in a potential implementing a Lorentz SO(1,3) symmetry breaking mechanism.

References

  • 1.T.T. Wu, C.N. Yang, Concept of nonintegrable phase factors and global formulation of gauge fields. Phys. Rev. D 12, 3845–3857 (1975) [Google Scholar]
  • 2.M. Hamilton, Mathematical Gauge Theory: With Applications to the Standard Model of Particle Physics. Universitext, 1 edn (Springer, Berlin, 2018)
  • 3.J. François, Differential geometry of gauge theory: an introduction. PoS, Modave 2020, 002 (2021)
  • 4.A. Trautman, Fiber Bundles, Gauge Field and Gravitation, in General Relativity and Gravitation, vol. 1 (Plenum Press, New York, 1979) [Google Scholar]
  • 5.T. Eguchi, P. Gilkey, A. Hanson, Gravitation, gauge theories and differential geometry. Phys. Rep. 66(6), 213–393 (1980) [Google Scholar]
  • 6.M. Göckeler, T. Schücker, Differential Geometry, Gauge Theory and Gravity. Cambridge Monographs on Mathematical Physics (Cambridge University Press, Cambridge, 1987)
  • 7.A. Unzicker, T. Case, Translation of Einstein’s attempt of a unified field theory with teleparallelism (2005)
  • 8.T. Sauer, Field equations in teleparallel space-time: Einstein’s fernparallelismus approach toward unified field theory. Hist. Math. 33(4), 399–439 (2006). Special Issue on Geometry and its Uses in Physics, 1900–1930
  • 9.H. Weyl, Gravitation and the electron. Proc. Natl. Acad. Sci. 15(4), 323–334 (1929) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.L. O’Raifeartaigh, The Dawning of Gauge Theory. Princeton Series in Physics (Princeton University Press, Princeton, 1997)
  • 11.C.N. Yang, R.L. Mills, Conservation of isotopic spin and isotopic gauge invariance. Phys. Rev. 96, 191–195 (1954) [Google Scholar]
  • 12.R. Utiyama, Invariant theoretical interpretation of interaction. Phys. Rev. 101, 1597–1607 (1956) [Google Scholar]
  • 13.T.W.B. Kibble, Lorentz invariance and the gravitational field. J. Math. Phys. 2(2), 212–221 (1961) [Google Scholar]
  • 14.D.W. Sciama, The physical structure of general relativity. Rev. Mod. Phys. 36, 463–469 (1964) [Google Scholar]
  • 15.Y. Ne’eman, Gauge theories of gravity. Acta Phys. Pol., B 29, 827–843 (1998) [Google Scholar]
  • 16.S.W. McDowell, F. Mansouri, Unified geometric theory of gravity and supergravity. Phys. Rev. Lett. 38, 739–742 (1977) [Google Scholar]
  • 17.M. Kaku, P.K. Townsend, P. Van Nieuwenhuizen, Gauge theory of the conformal and superconformal group. Phys. Lett. 69B, 304–308 (1977) [Google Scholar]
  • 18.F.W. Hehl, Y.N. Obukhov, Conservation of energy–momentum of matter as the basis for the gauge theory of gravitation. Fundam. Theor. Phys. 199, 217–252 (2020) [Google Scholar]
  • 19.Y.N. Obukhov, Poincare gauge gravity: selected topics. Int. J. Geom. Methods Mod. Phys. 3, 95–138 (2006) [Google Scholar]
  • 20.E.W. Mielke, Geometrodynamics of Gauge Fields. On the Geometry of Yang–Mills and Gravitational Gauge Theories. Mathematical Physics Studies (Springer, Berlin, 2017)
  • 21.Y.N. Obukhov, Poincaré gauge gravity: an overview. Int. J. Geom. Methods Mod. Phys. 15(supp01), 1840005 (2018) [Google Scholar]
  • 22.F.W. Hehl, Four lectures on Poincaré gauge field theory, in International School of Cosmology and Gravitation: Spin, Torsion, Rotation and Supergravity (2023)
  • 23.F.W. Hehl, J.D. McCrea, E.W. Mielke, Y. Ne’eman, Metric-affine gauge theory of gravity: field equations, Noether identities, world spinors, and breaking of dilation invariance. Phys. Rep. 258(1), 1–171 (1995) [Google Scholar]
  • 24.F.W. Hehl, A. Macias, Metric affine gauge theory of gravity. 2. Exact solutions. Int. J. Mod. Phys. D 8, 399–416 (1999) [Google Scholar]
  • 25.V. Vitagliano, T.P. Sotiriou, S. Liberati, The dynamics of metric-affine gravity. Ann. Phys. 326, 1259–1273 (2011). [Erratum: Annals Phys. 329, 186–187 (2013)]
  • 26.R. Percacci, Towards metric-affine quantum gravity. Int. J. Geom. Methods Mod. Phys. 17(supp01), 2040003 (2020) [Google Scholar]
  • 27.M. Blagojević, F.W. Hehl, T.W.B. Kibble, Gauge Theories of Gravitation (Imperial College Press, London, 2013) [Google Scholar]
  • 28.S. Kobayashi, Transformation Groups in Differential Geometry (Springer, Berlin, 1972) [Google Scholar]
  • 29.R.W. Sharpe, Differential Geometry: Cartan’s Generalization of Klein’s Erlangen Program. Graduate text in Mathematics, vol. 166 (Springer, Berlin, 1996)
  • 30.A. Cap, J. Slovák, Parabolic Geometries I: Background and General Theory. Mathematical Surveys and Monographs, vol. 1 (American Mathematical Society, Providence, 2009)
  • 31.J. François, L. Ravera, Cartan geometry, supergravity, and group manifold approach. Arch. Math. 60, 4 (2024) [Google Scholar]
  • 32.S. Kobayashi, On connections of Cartan. Can. J. Math. 8, 145–156 (1956) [Google Scholar]
  • 33.C.M. Marle, The works of Charles Ehresmann on connections: from Cartan connections to connections on fibre bundles, in Geometry and Topology of Manifolds, vol. 76 (Banach Center Publication, Warsaw, 2007)
  • 34.J. François, Reduction of gauge symmetries: a new geometrical approach. Thesis, Aix-Marseille Université, September 2014 (2014)
  • 35.J.T. François, L. Ravera, Geometric relational framework for general-relativistic gauge field theories. Fortschr. Phys. 73, 2400149 (2024) [Google Scholar]
  • 36.J. François, L. Ravera, Dressing fields for supersymmetry: the cases of the Rarita–Schwinger and gravitino fields. J. High Energy Phys. 2024(7), 41 (2024) [Google Scholar]
  • 37.J. François, L. Ravera, Relational supersymmetry via the dressing field method and matter-interaction supergeometric framework. Annalen Phys. e00121 (2025). arXiv:2503.19077 [hep-th]
  • 38.J. François, L. Ravera, Off-shell supersymmetry via manifest invariance. Phys. Lett. B 868, 139633 (2025). arXiv:2504.06392 [hep-th]
  • 39.J. François, L. Ravera, There is no boundary problem (2025). arXiv:2504.20945 [gr-qc]
  • 40.P.D. Alvarez, M. Valenzuela, J. Zanelli, Supersymmetry of a different kind. JHEP 04, 058 (2012) [Google Scholar]
  • 41.P.D. Alvarez, L. Delage, M. Valenzuela, J. Zanelli, Unconventional SUSY and conventional physics: a pedagogical review. Symmetry 13(4), 628 (2021) [Google Scholar]
  • 42.P.D. Alvarez, P. Pais, J. Zanelli, Unconventional supersymmetry and its breaking. Phys. Lett. B 735, 314–321 (2014) [Google Scholar]
  • 43.J. François, L. Ravera, Unconventional supersymmetry via the dressing field method. Phys. Rev. D 111(12), 12 (2025). arXiv:2412.01898 [hep-th]
  • 44.S. Kobayashi, Theory of connections. Annali di Matematica 43(1), 119–194 (1957) [Google Scholar]
  • 45.J. François, L. Ravera, On the meaning of local symmetries: epistemic-ontological dialectic. Accepted for publication in Found. Phys. 55(3), 38 (2025). arXiv:2404.17449 [physics.hist-ph] [DOI] [PMC free article] [PubMed]
  • 46.P. Berghofer, J. François, Dressing vs. fixing: on how to extract and interpret gauge-invariant content. Found. Phys. 54(6), 72 (2024) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.J. François, Artificial versus substantial gauge symmetries: a criterion and an application to the electroweak model. Philos. Sci. 86(3), 472–496 (2019) [Google Scholar]
  • 48.J. François, L. Ravera, Relational bundle geometric formulation of non-relativistic quantum mechanics (2025). arXiv:2501.02046 [quant-ph]
  • 49.A.A. Tseytlin, Poincaré and de sitter gauge theories of gravity with propagating torsion. Phys. Rev. D 26, 3327–3341 (1982) [Google Scholar]
  • 50.S. Coleman, J. Wess, B. Zumino, Structure of phenomenological Lagrangians. I. Phys. Rev. 177, 2239–2247 (1969) [Google Scholar]
  • 51.C.G. Callan, S. Coleman, J. Wess, B. Zumino, Structure of phenomenological Lagrangians. II. Phys. Rev. 177, 2247–2250 (1969) [Google Scholar]
  • 52.Y.M. Cho, Nonlinear realization and unified interaction. Phys. Rev. D 18, 2810–2812 (1978) [Google Scholar]
  • 53.S. Sternberg, Group Theory and Physics (Cambridge University Press, Cambridge, 1994) [Google Scholar]
  • 54.S. Kobayashi, K. Nomizu, Foundations of Differential Geometry, vol. I (Wiley, New York, 1963) [Google Scholar]
  • 55.S. Kobayashi, K. Nomizu, Foundations of Differential Geometry, vol. II (Wiley, New York, 1969) [Google Scholar]
  • 56.A. Trautman, On the structure of the Einstein–Cartan equations. Symp. Math. 12, 139–162 (1973) [Google Scholar]
  • 57.R. Jackiw, S.Y. Pi, Fake conformal symmetry in conformal cosmological models. Phys. Rev. D 91, 067501 (2015) [Google Scholar]
  • 58.J. François, Dilaton from tractor and matter field from twistor. J. High Energy Phys. 2019(6), 18 (2019) [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This manuscript has no associated data. [Author’s comment: Data sharing not applicable to this article as no datasets were generated or analysed during the current study.]

This manuscript has no associated code/software. [Author’s comment: Code/Software sharing not applicable to this article as ne code/software was generated or analysed during the current study.]


Articles from The European Physical Journal. C, Particles and Fields are provided here courtesy of Springer

RESOURCES