<<

A unified view of curvature and torsion in metric-affine of gravity through affine-vector bundles

Bo-Hung Chen1, 2, ∗ and Dah-Wei Chiou3, 4, † 1Department of Physics, National Taiwan University, Taipei 10617, Taiwan 2Center for Theoretical Physics, National Taiwan University, Taipei 10617, Taiwan 3Department of Physics, National Sun Yat-sen University, Kaohsiung 80424, Taiwan 4Center for Condensed Matter Sciences, National Taiwan University, Taipei 10617, Taiwan Abstract One of the most appealing results of metric-affine gauge theory of gravity is a close parallel between the Riemann curvature two-form and the Cartan torsion two-form: While the former is the field strength of the Lorentz-group connection one-form, the latter can be understood as the field strength of the coframe one-form. This parallel, unfortunately, is not fully established until one adopts Trautman’s idea of introducing an affine-vector-valued zero-from, the meaning of which has not been satisfactorily clarified. This paper aims to derive this parallel from first principles without any ad hoc prescriptions. We propose a new mathematical framework of an associated affine-vector bundle as a more suitable arena for the affine group than a conventional vector bundle, and rigorously derive the covariant derivative of a local section on the affine-vector bundle in the formal Ehresmann-connection approach. The parallel between the Riemann curvature and the Cartan torsion arises naturally on the affine-vector bundle, and their geometric and physical meanings become transparent. The clear picture also leads to a conjecture about a kinematical

arXiv:2008.07750v2 [gr-qc] 14 Jul 2021 effect of the Cartan torsion that in principle can be measured `ala the Aharonov-Bohm effect.

∗Electronic address: [email protected] †Electronic address: [email protected]

1 I. INTRODUCTION

Gauge theories of gravity are the effort to cast gravitation in the language of Yang- Mills theory. Various approaches of gauge theories of gravity, differing from one another by considering different local gauge groups, have been intensively developed in the past 40 years, and they have revealed various profound geometric structures of spacetime and achieved many inspiring results. (See [1, 2] for comprehensive accounts; also see [3] for a brief review and [4] for a historical account.) Even though a satisfactory gauge theory of gravity remains elusive, undertaking of the research has enormously enhanced our knowledge about many fundamental issues — inertial effects of a spin particle [5, 6], relationship between spin and torsion [7], nonlinear effects of gravity, topological aspects of gravity, to name a few [1–3]. Among various gauge approaches of gravity, metric-affine gauge theory (MAG) of gravity is a well-developed formulation that has many intriguing features (see [3, 8] for reviews). In MAG, the gauge group is given by the affine group A(n, R) := Rn o GL(n, R) — i.e. the semidirect product of the vector space Rn and general linear group GL(n, R), and corre- spondingly the a(n, R)-valued affine connection A = Γ(L) + Γ(T ) plays the role of a gauge (T ) n potential, where Γ is associated with the translational subgroup T (n, R) ∼= R of the affine group and Γ(L) with the linear subgroup GL(n, R). Particularly, if the general linear group GL(n, R) is replaced by the Lorentz group SO(1, n 1), MAG is reduced to the Poincar´e − gauge theory of gravity. The result of MAG profoundly suggests that the Cartan torsion two-form T and the Riemann curvature two-form R can be understood on an equal footing. More precisely, we have

T = D(L)θ dθ + Γ(L) θ, (1.1a) ≡ ∧ R = D(L)Γ(L) dΓ(L) + Γ(L) Γ(L), (1.1b) ≡ ∧ where θ is the coframe one-form, Γ(L) is the gauge potential one-form of the group GL(n, R), and D(L) is the exterior covariant derivative associated with Γ(L). These draw a close parallel between T and R in the sense that the former is the field strength of θ while the latter the field strength of Γ(L).1 It is tantalizing to equate θ with Γ(T ) and furthermore to identify

1 This parallel can also be extended to the nonmetricity field Q, which can be viewed as the field strength

2 T and R as the translational part R(T ) and the linear part R(L), respectively, of the affine gauge curvature as defined in (2.28). Unfortunately, while R is to be identified as R(L), the fact that θ and Γ(L) transform differently under the A(n, R) gauge transformation spoils the beauty of identifying θ as Γ(L) and T as R(T ). To resolve this glitch, Trautman [9] proposed a solution that introduces an affine-vector-valued zero-form ξi and defines the new one-form as θ := Γ(T ) + D(L)ξ. (1.2)

The new one-form θ then can be identified as the coframe field, and correspondingly T and R(T ) are related via i (T )i (L)i j T = R + Rj ξ . (1.3)

Tremendous effort has been devoted to interpreting ξ and deriving it from different per- spectives.2 The field ξ also appears in the context of gauged nonlinear realizations of the translation group [12–16]. By requiring Γ(T ) = 0, ξ can be interpreted as Cartan’s “gener- (L) i i alized radius vector” [17, 18]. By requiring Dµ ξ = δµ, which is used to connect the coset n space A(n, R)/GL(n, R) ∼= R to the cotangent space of the spacetime manifold, ξ can be kinematically understood as the “Poincar´ecoordinates” [19–24]. The field ξ could also be explained in view of the theory of dislocations [25] or in terms of jet bundles [26]. The fact that Γ(T ) is not directly identical to θ might give rise to a gravitationally induced geometric phase [27]. Despite unceasing endeavors to understand ξ and its relation to translational symmetry, “the story of ξ has not yet come to an end [8]”, and “the geometric and physical meaning of the relation [(1.2)], especially the role of the field ξ, is perhaps not completely satisfactorily clarified, yet [3].” Furthermore, it should be noted that the parallel between T and R has not been com- pletely spelled out. It is well known that the Riemann curvature yields two geometric con- sequences: geodesic deviation and holonomy around a closed curve. Correspondingly, the

of the metric tensor. See Table 5 in [3] and Table 1 in [8]. For our purpose, we focus on the parallel between T and R only. 2 Meanwhile, there have been some approaches that apparently do not get involved with ξ but nevertheless give rise to the coframe one-form θ. For example, in [10, 11], the approach of a “reductive principle bundle” was proposed, whereby Γ(T ) is directly identified as θ without referring to ξ. However, this approach does not seem to manifest the complete symmetry of the affine group A(n, R), obscuring the close parallel between the Riemann curvature and the Cartan torsion.

3 Cartan torsion is expected to have the two analogous consequences as well. Geodesic devia- tion describes the tendency of bending towards or away from each other of two neighboring geodesics that are initially parallel to each other (see Ch. 11 of [28] for a detailed account). Analogously, the Cartan torsion indicates the tendency of closure failure of a parallelogram — i.e., it gives rise to a displacement vector, which describes how much the two endpoints of an infinitesimal parallelogram fail to coincide (see [3], especially Fig. 4 therein, for a detailed account). The Riemann curvature and the Cartan torsion are analogous to each other in the sense that both of them measure how much the curved spacetime is deviated from the flat one in regard to parallelogram distortion, albeit from different considerations. On the other hand, R(X,Y ) yields a value of the general linear algebra gl(n, R), which corresponds to the holonomy around an infinitesimal closed curve spanned by the two vec- tors X and Y as depicted in Fig. 2. The holonomy of R is well understood as a linear transformation that tells the difference between the initial and final states for a given vector that is parallel transported around the closed curve until it is back to the starting point (see Ch. 11 of [28] for a detailed account). Likewise, T (X,Y ) yields a value of the trans- lation algebra t(n, R). It is suggestive that the holonomy of T indicates the translational displacement of something that is parallel transported around the closed curve. However, it is unclear what precisely that something is and how it is parallel transported. It would be self-contradictory if the holonomy of T is naively interpreted as a displacement of location experienced by a round trip in some sense of parallel transport, because any round trip, by definition, identifies the final location with the initial location and thus makes no displace- ment of location. Unlike the case of R, the geometric meaning of T in terms of holonomy remains rather obscure. The goal of this paper is to rigorously derive the parallel between the Riemann curvature and the Cartan torsion from first principles, making precise sense of ξ without any ad hoc prescriptions and obtaining a clear geometric picture of the holonomy of T .3 Motivated

3 The phrase “from first principles” here is used in the ordinary sense, meaning that we make only a few fundamental postulations naturally motivated from physical considerations, once and for all, and everything else (regarding the interplay of R and T through ξ) follows deductively. It should be remarked that our work is focused solely on the kinematical aspects of MAG, and we leave the dynamical aspects for future research. The phrase “first principles” used here does not intend to connotate an ab initio approach to formulating a dynamical theory of MAG from the action principle. For various ab initio approaches starting from the action principle, see e.g. [8, 29]. Note that all the approaches studied in the

4 by careful consideration of the Einstein equivalence principle in regard to the Poincar´e symmetry, we propose a new mathematical framework called the associated affine-vector bundle, which provides a more suitable arena for the affine group A(n, R) than a conventional associated vector bundle. Our strategy is to first consider an Ehresmann connection endowed on the of A(n, R), and then correspondingly define the parallel transport and the covariant derivative on the associated affine-vector bundle in the same spirit of defining the covariant derivative on an associated vector bundle. The Ehresmann-connection approach is rather formal and less familiar to physicists, but it is advantageous for our purpose of generalizing the notion of parallel transport to the associated affine-vector bundle, as it gives a clear geometric picture of connection independent of local gauge choice. We obtain the desired results: The field ξ arises naturally in the associated affine-vector bundle as an arbitrary gauge choice of a reference point, and the Cartan torsion T and the Reimann curvature R exactly correspond to the holonomies appearing in the “affine” part and the “vector” part, respectively, of the associated affine-vector bundle. This paper is organized as follows. In Sec. II, we give a brief overview of the mathematical foundations of MAG. In Sec. III, we consider the local Poincar´esymmetry in depth from the perspective of the Einstein equivalence principle, which motivates us to construct the associated affine-vector bundle. In Sec. IV, the associated affine-vector bundle is rigorously formulated. In Sec. V and Sec. VI, we then rigorously derive the covariant derivative and the corresponding curvature appearing on the associated affine-vector bundle. In Sec. VII, we consider possible observational consequences of the holonomy of the Cartan torsion. Finally, the results are summarised and discussed in Sec. VIII. For the theory of MAG, we follow closely the line of [8]. For the formulation of connections on fiber bundles, we follow closely the line of [30]. Readers who are unfamiliar with the formal quotient-space construction of an associated vector bundle or the formal definition of parallel transport via an Ehresmann connection are advised to read Chapters 9 and 10 of [30] first. Throughout this paper, Latin letters i, j, . . . are used as internal indices for algebras or vectors on fibers, while Greek letters µ, ν, . . . are used as external indices for spacetime.4

literature so far are essentially based on a conventional vector bundle, instead of an affine-vector bundle as proposed in this paper. 4 This convention is the same as that adopted in [1] but opposite to that in [2, 3, 8].

5 II. MATHEMATICAL FOUNDATIONS OF METRIC-AFFINE GAUGE THEORY

This section gives a brief overview of the mathematical foundations of MAG, following the line of [8]. The main purpose is to introduce basic ideas and define notations for later use. As a secondary goal, we also endeavor to present these materials with the full rigor so that various confusions resulting from subtleties can be avoided. In the approach of MAG for an n-dimensional spacetime, the gauge group is taken to be the affine group A(n, R) := Rn o GL(n, R), i.e. the semidirect product of the vector space Rn and the degree-n general linear group GL(n, R). The Lie algebra a(n, R) associated with i A(n, R) is given by the generators Pi of n-dimensional translations and the generators L j of n-dimensional linear transformations, which satisfy the Lie brackets:

[Pi,Pj] = 0, (2.1a)

 i  i L j,Pk = δkPj, (2.1b)  i m  i m m i L j,L n = δ L j δ L n, (2.1c) n − j

for i, j, = 1, . . . , n, or i, j, = 0, 1, . . . , n 1, depending on the index convention. The ··· ··· − i subalgebra spanned by L j is gl(n, R), the Lie algebra associated with GL(n, R). On the { } other hand, the subalgebra spanned by P i is t(n, R), the Lie algebra associated with the { } n-dimensional translation group T (n, R).5

The Lie algebra a(n, R) admits the M¨obiusrepresentation ρM in the (n + 1) (n + 1) × matrix form, which reads as     j i k j ¯i k j i k ai ρn×n(L j) b ρn(Pk) ai L j b e¯k ρM (ai L j + b Pk) =     , (2.2) 0 0 ≡ 0 0

i where ρn×n : gl(n, R) a(n, R) Mn×n(R) is the “identity map”, which maps L j to the ⊂ → n same n n matrix, and ρn : t(n, R) a(n, R) R is the “position map”, which maps Pk × ⊂ →

5 In the literature, T (n, R) is often denoted as Rn. In this paper, we rigorously distinguish between the translation group T (n, R) and the defining module Rn of T (n, R) or GL(n, R).

6 n 6 to the n-dimensional basis vectore ¯k R ; more precisely, ∈

i l i l i l ρn×n(L j) (L¯ j) = δ δ , (2.3a) m ≡ m m j m m m (ρn(Pk)) (¯ek) = δ . (2.3b) ≡ k

It is straightforward to prove from (2.1) that ρM is indeed a linear representation of a(n, R), i.e.,

ρM ([α, β]) = [ρM (α), ρM (β)], (2.4) for any α, β a(n, R). ∈ Because Rn is the defining module of GL(n, R), the Lie algebra gl(n, R) naturally acts on Rn in the matrix form:

i l i l i l m l i i l ρn×n(L j)e ¯k (L¯ j e¯k) = (L¯ j) (¯ek) = δ δ = δ e¯j , (2.5) ≡ m j k k where (2.3) has been used. That is, Rn is the carrier space of the fundamental representation of gl(n, R): i i i ρn×n(L j)e ¯k L¯ j e¯k = δ e¯j. (2.6) ≡ k Meanwhile, the Lie algebra t(n, R) also provides a carrier space of the following representa- tion of gl(n, R): i i i ρtn (L j)Pk := [L j,Pk] = δkPj (2.7) according to (2.1b). This draws a parallel between Rn and tn t(n, R) that they are ≡ isomorphic to and thus interchangeable with each other as far as their transformations under gl(n, R) are concerned.7 More precisely, we have the following commutative diagrams:

ρn tn t(n, R) Rn ≡ ¯ ρtn (λ)=[λ, · ] ρn×n(λ)≡λ , (2.8)

ρn tn t(n, R) Rn ≡

6 n In this paper, we use barred notations to denote Mn×n(R)-valued or R -valued objects and unbarred notations for the corresponding gl(n, R)-valued or t(n, R)-valued objects. More precisely, for η = ηiP i p j i p p ∈ t(n, R) Ω (M) and κ = κi L j gl(n, R) Ω (M), where Ω (M) is the space of p-forms over M, we ⊗ ∈ ⊗ i n p j ¯i p denoteη ¯ := ρn(η) = η e¯i R Ω (M) andκ ¯ := ρn×n(κ) = κi L j Mn×n(R) Ω (M). ∈ ⊗ ∈ ⊗ 7 This point is rarely emphasized and clarified in the literature of MAG. For example, what is intended to i i i i i be ρn×n(L j)ρn(Pk) L¯ j e¯k or ρtn (L j)Pk [L j,Pk] is often denoted in shorthand as L jPk by abuse ≡ ≡ of notation. The shorthand may look confusing, because for α, β g we have [α, β] g but, rigorously ∈ ∈ speaking, αβ is not even well defined.

7 and correspondingly ρn tn t(n, R) Rn ≡ −1 ¯ Rtn (Λ)=Λ( · )Λ ρn×n(Λ)≡Λ , (2.9)

ρn tn t(n, R) Rn ≡ j i λ (λi L j ) where Λ e = e GL(n, R) and the representation Rtn of GL(n, R) acting on ≡ ∈ t(n, R) is given by the exponential of ρtn , i.e.,

−1 1 Rtn (Λ)Pi := ΛPiΛ Pi + [λ, Pi] + [λ, [λ, Pi]] + ... (2.10) ≡ 2!

Any element g A(n, R) can be specified by the two variables Λ GL(n, R) and τ = ∈ ∈ i 8 τ Pi t(n, R) as ∈

i g(Λ, τ) = eτ Λ = eτ Pi Λ, (2.11a)

i g(Λ, τ)−1 = Λ−1e−τ = Λ−1e−τ Pi , (2.11b)

τ i n λ where e T (n, R) A(n, R) is a translation by the vectorτ ¯ = τ e¯i R and Λ e ∈ ⊂ ∈ ≡ ∈ GL(n, R) A(n, R) is a linear transformation exponentiated by λ gl(n, R). The M¨obius ⊂ ∈ representation ρM of g(τ, Λ) then takes the form:   Λ¯ τ ¯ τ λ ρM (τ) ρM (λ) ρM (g) = ρM (e )ρM (e ) = e e =   , (2.12) 0 1 where n ∞     0τ ¯ 1n×n τ¯ ρM (τ) X 1 e :=   =   1 + ρM (τ), (2.13a) n! n=1 0 0 0 1 ≡  n     ∞ ¯ λ¯ ¯ ρ (λ) X 1 λ 0 e 0 Λ 0 e M :=   =     . (2.13b) n! n=1 0 0 0 1 ≡ 0 1

Similarly, we have   ¯ −1 ¯ −1 −1 Λ Λ τ¯ ρM g(Λ, τ) =  −  , (2.14) 0 1

8 j i i It seems more elegant to use a pair of algebra-valued variables, λ = λi L j gl(n, R) and τ = τ Pi ∈ ∈ t(n, R), to specify an element g A(n, R) as g = eτ eλ. But it turns out more convenient to use Λ ∈ ∈ GL(n, R) and τ t(n, R) instead. ∈

8 which implies

−1 −1 −1 −1 −1 g(Λ, τ) = g(Λ , Rtn (Λ) τ) = g(Λ , Λ τΛ), (2.15) − −

and   ¯ τ −1 1n×n Λ¯τ ρM (Λ e Λ ) =   , (2.16) 0 1 which implies −1 Λ eτ Λ−1 = e(Rtn τ) = e(ΛτΛ ). (2.17)

i n The affine group A(n, R) naturally acts on affine vectors x¯ = x e¯i R in the following ∈ affine transformation: 0 x¯ x¯ = ρM (g)¯x = Λ¯¯x +τ, ¯ (2.18) → which can be cast in the M¨obiusmatrix form as       x¯0 Λ¯ τ ¯ x¯   =     . (2.19) 1 0 1 1

To study the local affine symmetry on a manifold M in the Yang-Mills gauge approach, we locally introduce the affine connection A a(n, R) Ω(M) as ∈ ⊗

(L) (T ) i (L)j (T )i i (L)j µ (T )i µ A = Γ + Γ = L j Γi + Pi Γ = L j Γiµ dx + Pi Γµ dx , (2.20)

which in the M¨obiusrepresentation reads as     ¯(L) ¯(T ) ¯i (L)j (T )i Γ Γ L j Γi e¯i Γ A¯ := ρM (A) =   =   . (2.21) 0 0 0 0

(L)j (T )i (L) i Here, Γi and Γ are the coefficients of the gl(n, R)-valued one-form Γ in the basis L j (T ) and of the t(n, R)-valued one-form Γ in the basis Pi, respectively. These coefficients are by (L)j (L)j µ (T )i (T )i µ themselves one-forms and thus can be recast as Γi = Γiµ dx and Γ = Γµ dx . The connection one-form A transforms inhomogeneously under the affine gauge transformation g−1(x) as: A0(x) = g−1(x)A(x)g(x) + g−1(x)dg(x), (2.22)

The affine connection A can be viewed to have two parts: the “linear” part Γ(L) gl(n, R) ∈ ⊗ Ω(M) and the “translational” part Γ(T ) t(n, R) Ω(M). The gauge transformation rules ∈ ⊗

9 of them can be easily obtained in the M¨obiusrepresentation as

Γ¯0(L) = Λ¯ −1Γ¯(L)Λ¯ + Λ¯ −1dΛ¯, (2.23a) Γ¯0(T ) = Λ¯ −1Γ¯(T ) + Λ¯ −1 dτ¯ + Γ¯(L)τ¯ Λ¯ −1Γ¯(T ) + Λ¯ −1D¯ (L)τ.¯ (2.23b) ≡ Alternatively, substituting (2.11) and A = Γ(L) + Γ(T ) into (2.22), by virtue of (2.1), we also arrive at the equivalent form:9

Γ0(L) = Λ−1Γ(L)Λ + Λ−1dΛ, (2.24a)

Γ0(T ) = Λ−1Γ(T )Λ + Λ−1dτ + [Γ(L), τ]Λ Λ−1Γ(T )Λ + Λ−1D(L)τΛ ≡ −1 (T ) −1 (L) Rtn (Λ) Γ + Rtn (Λ) D n τ. (2.24b) ≡ t Here, the exterior covariant derivative with respect to Γ(L) is defined as

(L) (L) (L) (L) D D n η := dη + [Γ , η] dη + ρtn Γ η, (2.25) ≡ t ≡ ∧

which maps η t(n, R) Ωp(M) to D(L)η t(n, R) Ωp+1(M).10 For any linear representa- ∈ ⊗ ∈ ⊗ tion ρV of gl(nR) with V being the carrier space, the exterior covariant derivative associated with Γ(L) can also be generalized for a V -valued form v V Ωp(M), defined as ∈ ⊗

(L) (L) D v := dv + ρV Γ v. (2.26) V ∧ Particularly, we define the shorthand

¯ (L) (L) D := DRn . (2.27)

9 Particularly, we have

( 1) ( 1)2 e−τ Γ(L)eτ = Γ(L) + − [τ, Γ(L)] + − [τ, [τ, Γ(L)]] + = Γ(L) + [Γ(L), τ], 1! 2! ··· and Λ−1e−τ d (eτ Λ) = Λ−1dτ Λ + Λ−1dΛ.

10 i i Let ζ = Tiζ be a g-valued p-form and η = Tiη be a g-valued q-form. The commutator of them is defined as pq i j pq j i i j [ζ, η] ζ η ( 1) η ζ := TiTj ζ η ( 1) TjTi η ζ = [Ti,Tj] ζ η , ≡ ∧ − − ∧ ∧ − − ∧ ∧ which is a g-valued (p + q)-form. In case ζ = η and p = q is odd, we have

[ζ, ζ] = 2 ζ ζ. ∧ In other cases, ζ η is not necessarily g-valued in general. See Section 10.3.2 in [30]. ∧ 10 The corresponding affine gauge curvature is defined as 1 R := dA + [A, A] dA + A A = R(L) + R(T ), (2.28) 2 ≡ ∧ which is again separated into the linear part R(L) gl(n, R) Ω2(M) and the translational ∈ ⊗ part R(T ) t(n, R) Ω2(M): ∈ ⊗ (L) i (L)j (L) 1 (L) (L) (L) (L) (L) R L jR = dΓ + [Γ , Γ ] dΓ + Γ Γ ≡ i 2 ≡ ∧ i (L)j 1  i m  (L)j (L)n i  (L)j (L)j (L)k = L j dΓ + L j,L n Γ Γ = L j dΓ + Γ Γ , (2.29a) i 2 i ∧ m i k ∧ i (T ) (T )i (T ) (L) (T ) (L) (T ) R PiR = dΓ + [Γ , Γ ] D Γ ≡ ≡ (T )i  i  (L)j (T )m  (T )i (L)i (T )j = Pi dΓ + L j,Pm Γ Γ = Pi dΓ + Γ Γ . (2.29b) i ∧ j ∧ Compared with (1.1b), the gl(n, R)-valued two-form R(L) in (2.29a) is the familiar Rie- mann curvature tensor R = D(L)Γ(L). Correspondingly, compared with (1.1a), it is at- tempting to identity the t(n, R)-valued two-form R(T ) in (2.29b) as the familiar Cartan torsion tensor T = D(L)θ, if we identity the t(n, R) connection Γ(T ) as the coframe one-form i ¯ −1 ¯ (L) θ = θ Pi t(n, R) Ω(M). Unfortunately, because of the inhomogeneous term Λ D τ¯, ∈ ⊗ or equivalently Λ−1(D(L)τ)Λ, appearing in (2.23b) and (2.24b), Γ(T ) cannot be identified as θ, which under the gauge transformation g−1(x) transforms as a vector, i.e.11

g−1(x) θ¯ θ¯0 = Λ¯ −1θ.¯ (2.30) −→ Following the suggestion by Trautman [9], one can introduce an affine-vector-valued zero- ¯ i n 0 from ξ = ξ e¯i R Ω (M), which transforms as ∈ ⊗ g−1(x) ξ¯ ξ¯0 = Λ¯ −1(ξ¯ τ¯) (2.31) −→ − in accordance with (2.18), and define the new one-form as (1.2), i.e.,

θ := Γ(T ) + D(L)ξ. (2.32)

It can be easily verified that, the transformation law of the one-form θ defined above just takes the form of (2.30) and therefore θ can be identified as the coframe one-form. Corre- spondingly, the Cartan torsion T is related to R(T ) via (1.3), i.e.,

T¯ = D¯ (L)θ¯ = R¯(T ) + R¯(L)ξ,¯ (2.33)

11 Do not confuse a vector with an affine vector. The former is insensitive to translation, while the latter transforms as (2.18) under the gauge transformation g(x).

11 or equivalently T = D(L)θ = R(T ) + [R(L), ξ]. (2.34)

However, as commented in Sec. I, the geometric and physical meaning of ξ remains obscure. In the rest of this paper, we first study the Poincar´esymmetry in more depth and then formulate the new framework of an affine-vector bundle to resolve the problem regarding ξ.

III. POINCARE´ SYMMETRY IN VIEW OF THE EINSTEIN EQUIVALENCE PRINCIPLE

The Einstein equivalence principle states that (Ch. 16 of [28]): “In any and every local Lorentz frame, anywhere and anytime in the universe, all the (nongravitational) laws of physics must take on their familiar special-relativistic forms.” However, on the passage from the special-relativistic laws to general-relativistic counterparts by applying the standard “comma-goes-to-semicolon” rule [28], the resulting covariant laws respect the Lorentz gauge symmetry as well as the symmetry of diffeomorphism, but seem to render irrelevant the translational part of the Poincar´esymmetry of special relativity.12 In this section, we address this issue in depth by considering the Dirac equation and the Dirac Lagrangian density as representative examples. The Poincar´egroup acting on an n-dimensional Minkowski (flat) spacetime is given by P (1, n 1) := Rn o SO(1, n 1), where SO(1, n 1) as a proper subgroup of GL(n, R) − − − is the Lorentz group. The Lie algebra p(1, n 1) associated with P (1, n 1) is given by − − ij the generators Pi of n-dimensional translations and the generators M of n-dimensional Lorentz transformations. The generators M ij so(1, n 1) are given by particular linear ∈ − i superpositions of the generators L j gl(n, R) as ∈ ij [ij] ik j jk i M M := η L k η L k, (3.1) ≡ −

12 The “comma-goes-to-semicolon” rule is also known as the minimal coupling prescription in Riemannian spacetimes. This prescription has also been studied for the cases that the covariant derivatives are not Riemannian ones. Particularly, it was recently argued that in general one cannot make sense of the notion of minimal coupling that is well defined for arbitrary cases [31, 32]. This problem does not concerns us here, as our purpose is to take the standard minimal coupling prescription as a typical example to address the issue that the translational part of the Poincar´esymmetry is not manifested. Other prescriptions beyond minimal coupling that have been studied in the literature so far generally bear the same issue.

12 where ηij is the (inverse of the) metric tensor of the n-dimensional Minkowski spacetime. Consequently, by (2.1), the Lie algebra of p(1, n 1) satisfies the Lie brackets: − P i,P j = 0, (3.2a) M ij,P k = ηikP j ηjkP i, (3.2b) − M ij,M kl = ηikM jl ηilM jk ηjkM il + ηjlM ik, (3.2c) − − i ij where P := η Pj. The Dirac equation in the Minkowski spacetime is given by

i  iγ ∂i m ψ(¯x) = 0, (3.3) − i n i i i wherex ¯ x e¯i R , x are the Minkowski coordinates, ∂i := ∂/∂x , and γ are the N N ≡ ∈ × gamma matrices (where N := 2bn/2c), which satisfy

γi, γj γiγj + γjγi = 2ηij. (3.4) { } ≡ In the Dirac spinor representation, the Lorentz generators are represented as i i σij σ[ij] := [γi, γj] γiγj γjγi . (3.5) ≡ 4 ≡ 4 − One can easily verify that

i jk ij k ik j  jl k i kl j i  m jk i m [γ , σ ] = i η γ η γ = i η (L¯ l) η (L¯ l) γ i(M¯ ) γ , (3.6) − m − m ≡ m which is just the infinitesimal form of the finite Lorentz transformation law of γi:

−1 i ¯ i j ΛD γ ΛD = (Λ) jγ , (3.7) where 1 ij Λ = exp ωijM SO(1, n 1) (3.8) 2 ∈ − is any Lorentz transformation and

 i ij ΛD := exp ωijσ (3.9) − 2 is the Dirac spinor representation of Λ. Under a Poincar´etransformation g(Λ, τ) P (1, n 1) A(n, R), according to (2.18), ∈ − ⊂ we have

x¯ x¯0 = Λ¯¯x +τ, ¯ (3.10a) → 0 −1 j ∂i ∂ = (Λ¯ ) ∂j. (3.10b) → i i 13 Correspondingly, the Dirac spinor field is transformed as ψ ψ0 via13 →

0 0 0 −1 0 −1 ψ(¯x) ψ (¯x ) = ΛDψ(¯x(¯x )) = ΛDψ(Λ¯ x¯ Λ¯ τ¯). (3.11) → −

The Dirac equation is invariant under the Poincar´etransformation in the sense that, if ψ(¯x) satisfies (3.3), then it implies

i 0  0 0 i 0  −1 0 −1 iγ ∂ m ψ (¯x ) = iγ ∂ m ΛDψ(Λ¯ x¯ Λ¯ τ¯) i − i − − −1 i −1 j  = ΛD iΛ γ ΛD(Λ¯ ) ∂j m ψ(¯x) D i − i k −1 j  = ΛD i(Λ)¯ γ (Λ¯ ) ∂j m ψ(¯x) k i − j  = ΛD iγ ∂j m ψ(¯x) = 0, (3.12) − ¯ i ¯ −1 j j ¯ −1 ¯ where (3.7), (3.10), and (Λ) k(Λ ) i = δk (i.e. Λ Λ = 1n×n) have been used. Furthermore, defining ψ¯ = ψ†γ0, (3.13) we have the transformation law:

ψ¯(¯x) ψ¯0(¯x0) = ψ¯(Λ¯ −1x¯0 Λ¯ −1τ¯)Λ−1. (3.14) → − D

It follows from (3.12) and (3.14) that the Dirac lagrangian density defined as

i  LDirac := ψ¯ iγ ∂i m ψ (3.15) − is invariant under the Poincar´etransformation as well. The Dirac equation and the Dirac Lagrangian can be generalized from the context of special relativity to that of general relativity. In general relativity, we have an n-dimensional manifold M as the curved spacetime, which is locally coordinated by xµ (not to be confused i with the Minkowski coordinates x ). At each point p M, we have a tangent space Tp(M), ∈ which is isomorphic to Rn and is to be identified as the local Lorentz frame of a freely falling laboratory (see Fig. 1). Consider a Dirac spinor field ϕ on M, i.e., ϕ : p M ∈ 7→ ϕ(xµ(p)) CN . In the immediate vicinity of a given point p, the field ϕ locally gives rise ∈ n N to a Dirac spinor field ψ :x ¯ R = Tp(M) ψ(¯x) C in the locality of the origin ∈ ∼ 7→ ∈

13 More precisely, the Dirac spinor field ψ :x ¯ Rn ψ(¯x) CN is to be regarded as a section of the fiber ∈ 7→ ∈ bundle with Rn being the base space and CN being the fiber. The Poincar´etransformation moves a given section ψ to a new section ψ0 :x ¯0 Rn ψ0(¯x0) CN . ∈ 7→ ∈ 14 e¯ n Tp(M) j R ∂ν θ¯ e¯ ei dxµ ≡ i µ

∂µ e¯i p x¯ = 0

? e¯j

M

p e¯i ξ¯

FIG. 1: A spacetime manifold M (lower left) naturally gives rise to a tangent space Tp(M) (upper

n left) at any point p M. Tp(M) is to be identified as the local Lorentz frame R of a freely falling ∈ i µ laboratory via the “soldering form” θ¯ e¯ie µdx (see upper right). Within the local frame in the ≡ locality of p, a local Lorentz transformation rotates or boosts the orthonormal coordinates xi to x0i while leaving the origin unmoved. On the other hand, a local Poincar´etranslation displaces the coordinate origin from the point p, and consequently the translated local frame can no longer be identified as Tp(M) (see lower right). x¯ = 0 as measured by a local freely falling laboratory. The local field ψ in the locality of x¯ = 0 is to be understood as the “best” linear approximation of ϕ in the vicinity of p; its exact formulation is to be properly prescribed in accordance with the Einstein equivalence principle as follows. Firstly, the value of ψ at the originx ¯ = 0 of Rn is prescribed to be identical to that of ϕ at p, i.e.,

µ ψ(¯x) x¯=0 = ϕ(x ) xµ=xµ(p). (3.16)

Secondly, in order to reflect the local features of ψ around the locality ofx ¯ = 0 in the local Lorentz frame, we need also to specify the derivatives of ψ with respect tox ¯. As the Minkowski coordinates xi of a freely falling frame are supposed to be locally identified as Riemann normal coordinates, it is most natural to take the prescription:

µ (D) µ ν µ (D) µ (∂iψ)(¯x) x¯=0 = ei(x )y D ϕ(x ) xµ=xµ(p) = ei (x )Dν ϕ(x ) xµ=xµ(p), (3.17) where D(D) is the exterior covariant derivative associated with the SO(1, n 1) spin con- − 15 nection, and µ ei ei ∂µ (3.18) ≡ are the frame field (also called vielbein), i.e. a set of n orthonormal vector fields.14 The frame field is related to the vector-valued coframe one-form field

i µ θ¯ e¯i e µdx (3.19) ≡ via µ j j ei e µ = δi . (3.20)

The coframe field θ¯ provides a mathematical facility that naturally “solders” a tangent vector µ ¯ ¯ ¯ i µ n X = X ∂µ Tp(M) to X = θ(X) X θ = e µX e¯i R in the local frame (see Fig. 1). ∈ ≡ y ∈ ¯ ¯ n Particularly, it maps ei Tp(M) to θ(ei) ei θ =e ¯i R . Furthermore, the exterior ∈ ≡ y ∈ covariant derivative D(D) is defined as

(D) (SO) (SO)  ij (SO) µ D := d + ρD Γ d + Γ ∂µ iσ Γ dx , (3.21) ≡ D ≡ − ijµ where Γ(SO) M ijΓ(SO) is the so(1, n 1)-valued connection one-form, which is the same as ≡ ij − (L) i (L)j Γ L j Γ except that its algebra value is now restricted to the subalgebra so(1, n 1) ≡ i − ij ij of gl(n, R), and ρD(M ) = iσ is the Dirac spinor representation of so(1, n 1), which − − maps an element of so(1, n 1) to an N N matrix. − × Under a local Lorentz gauge transformation Λ(xµ) SO(1, n 1), we have ∈ −

µ 0 µ µ µ ϕ(x ) ϕ (x ) = ΛD(x )ϕ(x ), (3.22a) → µ 0 µ µ −1 j µ ei(x ) e (x ) = (Λ(¯ x ) ) ej(x ), (3.22b) → i i (SO) µ 0(SO) µ µ (SO) µ −1 µ µ −1 µ Γ (x ) Γ (x ) = ΛD(x )Γ (x )Λ (x ) + ΛD(x )dΛ (x ), (3.22c) D → D D D D where (3.22c) is given in accordance with (2.24a).15 Consequently, these imply

(D) µ (D) 0 µ 0(SO) µ  0 µ µ (D) µ D ϕ(x ) (D ϕ) (x ) d + Γ (x ) ϕ (x ) = ΛD(x )D ϕ(x ). (3.23) → ≡ D

According to (3.16) and (3.17), from the viewpoint of a freely falling laboratory whose frame i n coordinates are given byx ¯ = x e¯i R , the local SO(1, n 1) gauge transformation upon ∈ −

14 The frame field is also known as a tetrad or vierbein in the case of n = 4. 15 Note that (2.22) is expressed for the gauge transformation g−1(x), instead of g(x). Thus, we have to interchange Λ and Λ−1 when applying (2.24a) here.

16 the local field reads as

0 0 ψ(¯x) ψ (¯x ) 0 = ΛD(p)ψ(¯x) , (3.24a) x¯=0 → x¯ =0 x¯=0 0 0 ¯ −1 j (∂iψ)(¯x) (∂iψ) (¯x ) 0 = (Λ(p) ) ΛD(p)(∂jψ)(¯x) . (3.24b) x¯=0 → x¯ =0 i x¯=0 Consequently, by the similar tricks used for deriving (3.12), it can be shown that

i 0 0 0 0  i  iγ (∂iψ) (¯x ) mψ (¯x ) 0 = ΛD(p) iγ (∂iψ)(¯x) mψ(¯x) . (3.25) − x¯ =0 − x¯=0 That is, in the local Lorentz frame of a freely falling laboratory, if ψ(¯x) in the locality of an event p satisfies the Dirac equation with respect to the local Minkowski coordinatesx ¯, then the SO(1, n 1)-transformed version ψ0(¯x0) in the locality of p satisfies the Dirac equation − as well. We have just affirmed the Einstein equivalence principle with regard to the Lorentz, i.e. SO(1, n 1), symmetry of the Dirac equation. Therefore, the covariant generalization − of the Dirac equation is prescribed as

i µ (D)  µ iγ ei D m ϕ(x ) = 0, (3.26) µ − which has been shown to respect the local Lorentz gauge symmetry. Similarly, the covariant generalization of the Dirac Lagrangian density is prescribed as

i µ (D)  LDirac :=ϕ ¯ iγ ei D m ϕ, (3.27) µ − which respects the local Lorentz gauge symmetry as well. Unfortunately, while the covariant Dirac equation and the covariant Dirac Lagrangian density are invariant under Lorentz gauge transformations as discussed above as well as under diffeomorphisms of xµ x0µ = x0µ(xµ) (as they are in covariant forms), they are → oblivious of the translational part of the Poincar´esymmetry. That is, we can make sense of a local Lorentz transformation corresponding tox ¯ x¯0 = Λ(¯ p)¯x within the local Lorentz → frame, but cannot prescribe a local translationx ¯ x¯0 =x ¯ +τ ¯(p) within the Lorentz frame → as any attempt to do so is meant to spoil the fact that the local Lorentz frame in the locality of p is identified as Tp(M). This obstacle is not only for the Dirac spinor field but for any generic field that is described as a local section of a vector bundle over M associated with the Lorentz group.16 Lacking an adequate mathematical language to address a local translation

16 Particularly, the Dirac field ϕ(xµ) can be viewed as a local section of the CN vector bundle over M associated with SO(1, n 1), and a scalar field φ(xµ) a local section of the R or C vector bundle. − 17 suggests that the Einstein equivalence principle has not been completely fulfilled in the standard comma-goes-to-semicolon approach, because special-relativistic laws are supposed to respect the full symmetry of the Poincar´egroup, not just its Lorentz subgroup. In order to make sense of the local Poincar´etransformation, in the next section, we propose a new mathematical framework called the associated affine-vector bundle, which extends the structure of an associated vector bundle by rendering the Poincar´etranslation as an affine transformation on the fiber. Let v V represents a local value of a generic field ∈ in the vector fiber V . The idea is to augment v with an affine vector ξ¯ Rn serving as a ∈ reference point, so that f = (ξ,¯ v) F = Rn V represents the local field with reference ∈ × to ξ¯, which indicates the displacement between the designated origin of the local frame and the point p. See Fig. 1. One might argue that designating the reference point is purely a matter of arbitrary gauge choice, so it makes no difference to take its involvement into account. This argument is true only if one considers the local physics that is completely independent of the neighboring local frame. If one considers the covariant derivative of the reference point in the sense of “parallel transport”, it will yield physical significance in relation to the Cartan torsion. This is the topic we will investigate closely. To keep our analysis as generic as possible, in the following, we will study the local translational transformation in the broader context of the affine group A(n, R) := Rn o GL(n, R), instead of the Poincar´egroup proper.

IV. ASSOCIATED AFFINE-VECTOR BUNDLE

Given a principal bundle P (M,G) whose fiber is identical to the structure group G, one can formulate an associated vector bundle P ρ V with V being a vector space and ρ being × the presentation of G over V via the quotient-space construction (see Sec. 9.4.2 of [30]). The principal bundle P (M,G) can be endowed with an Ehresmann connection one-form ω and its local form as a pullback to the base space M is the well-known gauge potential (see Sec. 10.1 of [30]). The connection defined on P (M,G) naturally gives rise to the covariant derivative on the associated vector bundle P ρ V and the resulting covariant derivative × involves the same gauge potential (see Sec. 10.4 of [30]).17

17 Therefore, if we consider parallel transport on a vector bundle associated with A(n, R), we will simply obtain a result identical (or, more precisely, homomorphic) to (2.29). Without going beyond the framework

18 In the approach of MAG, the group G is given by the affine group A(n, R) := Rn o GL(n, R). To reflect the translational symmetry more faithfully, we develop a slightly dif- ferent framework by constructing an associated affine-vector bundle in the same spirit of defining an associated vector bundle. The associated affine-vector bundle is defined below n and denoted as E = P ρA F , where F = R V is the affine-vector space and ρA is the × × affine (nonlinear) representation of A(n, R) acting on F defined below. Let g = g(Λ, τ) G, where Λ GL(n, R) and τ t(n, R), act on (u, f) P F as ∈ ∈ ∈ ∈ × follows:

−1 g(u, f) g(u, (ξ,¯ v)) = (ug, ρA(g) f) ≡ −1 −1  ug, ρM (g) ξ,¯ ρV (Λ) v ≡ −1 −1  ug, (Λ¯ (ξ¯ τ¯), ρV (Λ) v) , (4.1) ≡ −

¯ n where f = (ξ, v) F = R V , ρV (Λ) is the linear representation of GL(n, R) acting ∈ × on v V , and ρM (g) is the M¨obiusrepresentation of A(n, R) acting on affine vectors as ∈ defined in (2.18). That is, the “vector” part V of F responses only to the GL(n, R) part of G, whereas the “affine” part Rn of F serves as a “reference point”, which transforms as 18 an affine vector. The associated affine-vector bundle E = P ρ F is then defined as the × A quotient space (P F/ ) via the following equivalence relation × ∼

−1 (u, f) (ug, ρA(g) f). (4.2) ∼

The affine-vector bundle E is a fiber bundle over M with F being the fiber in the sense that its fiber bundle structure is given by the projection, trivialization, and transition functions as defined in the following.

The projection πE : E M is defined as πE(u, f) = π(u), where π : P M is the → → projection of P (M,G). The projection πE is well defined under the equivalence relation −1 (4.2), since we have π(ug) = π(u) and it follows πE(ug, ρA(g) f) = πE(u, f). The local trivialization ψ : U F π−1(U ) is given by ψ (p, f) = (φ (p, e), f), where U M is i i × → E i i i i ⊂

of an associated vector bundle, we are unable to solve the problem of ξ. 18 Note that ρA(g): F F is a representation of g A(n, R) in the sense ρA(g1g1) = ρA(g1)ρA(g2) and → ∈ 1 0 ρA(e)f = f, where e = g(Λ = , τ = 0). However, it is a nonlinear representation. If ρA(g)f1 = f1 and 0 0 0 ρA(g)f2 = f2, it does not imply ρA(g)(αf1 + βf2) = αf1 + βf2. By contrast, the vector part ρV (Λ) is linear.

19 an open set of M, p U is a point in M, e G is the identity, and φ : U G π−1(U ) ∈ i ∈ i i × → i is the trivialization of P (M,G). It is easy to show that πE ψ (p, f) = p. ◦ i Let φ−1(u) = (p, g ), where u π−1(U ) and p = π(u). On U U = , the two i i ∈ i i ∩ j 6 ∅ trivializations φ and φ are related by a smooth map t : U U G via i j ij i ∩ j →

φj(p, gj) = φi(p, tij(p)gj), (4.3)

where the maps tij are the transition functions of P . As P is a principal bundle, the right action of g G on π−1(U ) is defined as φ−1(ug) = (p, g g), or equivalently ∈ i i i

ug = φi(p, gig). (4.4)

This leads to

ug = φj(p, gjg) = φi(p, tij(p)gjg) = φi(p, gig), (4.5) where gi = tij(p)gj. That is, the right action is well defined and independent of local trivializations. Now, regarding the trivializations of E, we have

ψ (p, f ) (φ (p, e), f ) = (φ (p, t (p)), f ) = (φ (p, e)t (p), f ) j j ≡ j j i ij j i ij j −1 −1 (φ (p, e), ρA(t (p)) f ) ψ (p, ρA(t (p)) f ), (4.6) ∼ i ij j ≡ i ij j where we have used (4.2) and (4.4). That is, the transition functions of E are given by −1 ρA(tij(p)) , where tij(p) are the transition functions of P .

V. COVARIANT DERIVATIVE ON THE AFFINE-VECTOR BUNDLE

If a principle bundle P is endowed with an Ehresmann connection one-form ω g T ∗P , ∈ ⊗ it naturally gives rise to the notion of parallel transport as one can define the horizontal lift of a given curve γ : [0, 1] M. A curveγ ˜ : [0, 1] P is said to be a horizontal lift of γ if → → π γ˜ = γ and ω(X˜) = 0, where X˜ =γ ˜∗(X) with X(t) Tγ(t)M being a tangent vector to ◦ ∈ γ(t). Let σ Γ(U ,P ): U P be an arbitrary local section of P on U . The well known i ∈ i i → i gauge potential A g Ω(U ) is given as the pullback of ω via σ , i.e. i ∈ ⊗ i i

∗ Ai = σi ω. (5.1)

20 As two sections are related by transition functions, i.e., σj(p) = σi(p)tij(p), it turns out Ai follows the well known gauge transformation rule:

−1 −1 Aj = tij Aitij + tij dtij. (5.2)

Conversely, if gauge potentials that are locally defined for an open covering of M satisfy (5.2), we can uniquely construct an Ehresmann connection one-form ω for P from these local gauge potentials. Given an Ehresmann connection on P , we can define the covariant derivative on an associated affine-vector bundle E = P ρ F just in the same spirit of defining the covariant × A derivative on an associated vector bundle P ρ V (see Sec. 10.4 of [30]). × First, we consider a local section sv Γ(U ,E) of E on U , i.e. sv : U E, as a ∈ i i i → representative of the equivalence class associated with the relation (4.2):

  sv(p) = σi(p), ξ¯i(p), v(p) , (5.3)

¯ n where σi : Ui P is a local section of P on Ui, ξi : Ui R , and v : Ui V . The → → → arbitrary choice of σi amounts to different gauge fixing as indicated in (5.1). In the affine- vector bundle associated with G = A(n, R), we have the additional arbitrariness of choosing different “reference points”, which amounts to specifying ξ¯i(p).

For a given curve γ(t), a section sv(γ(t)) along γ(t) is said to be parallel transported if,   in the representation sv(γ(t)) = (˜γ(t), (ξ¯i(γ(t)), v(γ(t)))) , ξ¯i and v remain constant (i.e., independent of t). The notion of parallel transport is well defined, since ifγ ˜0(t) is another horizontal lift of γ(t), we haveγ ˜0(t) =γ ˜(t)g for a constant g G and consequently ∈

0 −1 0 −1 −1 0 0 0 [˜γ, (ξ¯ , v)] = [˜γ g , (ξ¯ , v)] = [˜γ , (ρM (g) ξ¯ , ρV (g) v)] [˜γ , (ξ¯ , v )], (5.4) i i i ≡ i ¯0 0 which follows that ξi and v are constant as well. The notion of parallel transport enables us to define the covariant derivative of a sec- tion sv(p). Let p = γ(t = 0) M and X TpM be a tangent vector to γ at p, i.e., ∈ ∈ µ µ dx (γ(t))/dt t=0 = X . The covariant derivative of s(p) with respect to X is defined as   d ¯  X sv := γ˜(0), ξi(γ(t)), v(γ(t)) . (5.5) ∇ dt t=0 Again, it is easy to show that this definition is well defined, regardless of choosing a different 0 horizontal liftγ ˜ (t). Once we have defined X s, we can also define the exterior covariant ∇ 21 derivative : Γ(U, E) Γ(U, E) Ω(U), which maps a section sv Γ(U, E) to a section- ∇ → ⊗ ∈ valued one-form, by

sv(X) := X sv, (5.6) ∇ ∇ where X X (U) is a vector field over an open set U. ∈ So far, the ideas of parallel transport and covariant derivative on the associated affine- vector bundle are the same as those on the associated vector bundle. However, because the

section sv(p) has to specify ξ¯i(p) as an extra “gauge fixing”, the local expression for the covariant derivative leads to a crucial difference. Given a local section σ Γ(U ,P ), a horizontal liftγ ˜(t) of γ(t) can be expressed as i ∈ i γ˜(t) = σi(t)gi(t), where gi(t) g(Λi(t), τ(t)) := gi(γ(t)) A(n, R). A local section along ≡ ∈ γ(t) expressed in the form (5.3) then leads to

 −1   −1  sv(t) := sv(γ(t)) = γ˜(t)gi(t) , (ξ¯i(t), v(t)) = γ˜(t), ρA(gi(t)) (ξ¯i(t), v(t)) , (5.7) where f(t) ξ¯ (t), v(t) := ξ¯ (γ(t)), v(t). By the identity dg(t)−1/dt = ≡ i i g(t)−1(dg(t)/dt)g(t)−1, we then have −     d −1 ¯ X sv = γ˜(0), ρA(gi(t)) (ξi(t), v(t)) ∇ dt t=0   −1 dρA(t) −1 −1 df(t) = γ˜(0), ρA(t) ρA(t) f(t) + ρA(t) − dt dt t=0   −1 dρA(t) −1 df(t) = γ˜(0)gi(0) , ρA(t) f(t) + − dt dt t=0   ¯   dρM (t) −1 ¯ dξi(t) dρV (t) −1 dv(t) = σi(0), ρM (t) ξi(t) + , ρV (t) v(t) + , − dt dt − dt dt t=0 (5.8)

where ρA(t) := ρA(gi(t)), ρM (t) := ρM (gi(t)), and ρV (t) := ρV (Λi(t)). As σi is a section

of P (M,G) andγ ˜(t) = σi(γ(t))gi(t), it can be shown that (see Eq. 10.13 in [30] for more details) dgi(t) = ω(σ ∗X)g (t) = A (X)g (t), (5.9) dt − i i − i i where (5.1) has been used. This leads to   µ µ Γ¯(L) Γ¯(T ) dρM (t) −1 ¯ dx ¯ dx iµ iµ ρM (t) = Ai(X) Aiµ   (5.10) − dt ≡ dt ≡ dt 0 0

22 according to (2.21), and

µ dρV (t) −1  (L)  dx  (L) ρV (t) = ρV Γ (X) ρV Γ . (5.11) − dt i ≡ dt iµ

Acting (5.10) on ξ¯i via (2.19) yields

µ dρM (t) −1 dx  (L) (T ) ρM (t) ξ¯ (t) = Γ¯ ξ¯ + Γ¯ . (5.12) − dt i dt iµ i iµ

Substituting (5.11) and (5.12) into (5.8), we have  µ    dx ¯ (L) ¯ ¯(T ) (L) X sv = σi(0), Dµ ξi + Γiµ ,DV iµv , (5.13) ∇ dt t=0 where dxµ(t)/dt := dxµ(γ(t))/dt = Xµ(γ(t)), or equivalently

h  (L) (T ) (L) i sv = σi, D¯ ξ¯ + Γ¯ ,D v , (5.14) ∇ i i V i (L) ¯ (L) where DV and D are defined in (2.26) and (2.27). Remarkably, the new one-form θ as defined in (2.32) arises naturally. In terms of θ, the

covariant derivative of the section sv(p) Γ(U, E) can be recast as ∈

h  (L) i sv = σ , θ¯ ,D v , (5.15) ∇ i i V i

or equivalently

h  (L) i h  (L)  i X sv = σ , θ¯ (X),D (X)v σ , θ¯ (X), dv(X) + ρV Γ (X) v . (5.16) ∇ i i V i ≡ i i i

The one-forms appearing in a local expression for the covariant derivative are to be identified as the local gauge potentials. Therefore, the corresponding local gauge potentials are given (L) by θ for the “affine” part and Γ (or, more precisely, the ρV -representation thereof) for the “vector” part. Since the affine vector ξ¯ plays the role of a reference point of a local Lorentz frame as discussed in Sec. III, the corresponding potential θ dictates how a reference point is different from its neighbored value via parallel transport. The weak equivalence principle furthermore requires that the parallel transport of a reference point be universal regardless of the matter content. In other words, θ shall be independent of the vector space V , which is chosen to represent the matter field under consideration. Since θ¯ transforms as a vector-valued one- form under a local A(n, R) transformation as indicated in (2.30) and is independent of the

23 matter content, we must identify θ¯ as the coframe one-form (up to an arbitary universal factor). On the other hand, since Γ(L) is insensitive of the translational part of a local A(n, R) transformation as shown in (2.23a) or (2.24a), it is identified as the familiar gauge potential associated with GL(n, R). Having rigorously derived θ and Γ(L) as the local gauge potentials with regard to the covariant derivative of a section on the associated affine-vector bundle, we will study the corresponding curvatures (i.e., field strengths) of the gauge potentials in the next section.

VI. CURVATURE ON THE AFFINE-VECTOR BUNDLE

In the previous section, we define the covariant derivative X sv with respect to a tangent ∇ vector X TpM. This leads us to define the corresponding curvature. To begin with, we ∈ compute Y X sv for any two vector fields X,Y X (U). ∇ ∇ ∈ Let χ : [0, 1] U be a curve on U whose tangent vectors are given by Y , i.e., dxµ(t)/dt := → µ µ dx (χ(t))/dt = Y (χ(t)), andχ ˜(t) = σi(t)gi(t) be a horizontal lift of χ(t). Starting from (5.16) and following the same procedures used in (5.8), we have      d −1 ¯ (L)  Y X sv = χ˜(0), ρA(gi(t)) θi(X(t)), ρV Γi (X(t)) v(t) + dv(X(t)) ∇ ∇ dt t=0   µ ¯ µ ν dx dθiµ(t) µ ∂X dx = σ (0), ρ¯ (A (t)) θ¯ (X(t)) + X (t) + θ¯ (t) , i dt θ i i dt iµ ∂xν dt dxµ  d  ρ Γ(L)(t)ρ Γ(L)(X(t))v(t) + ρ Γ(L)(t) Xµ(t) v(t) dt V iµ V i dt V iµ  µ ν  µ ν   (L) ∂X dx d(∂µv(t)) µ ∂X dx + ρV Γiµ (t) ν v(t) + X (t) + ∂µv(t) ν ∂x dt dt ∂x dt t=0   1 = σ (0), Γ¯(L)(Y )θ¯ (X) + (Y dθ¯ )(X) + θ¯ Y [Xµ], i i i 2 y i iµ  1    ρ Γ(L)(Y )Γ(L)(X) + Y dΓ(L) (X) + Γ(L)Y [Xµ] v V i i 2 y i iµ   ν µ µ + (∂ν∂µv)Y X + (dv)µY [X ] , (6.1) t=0 where dxµ(t)/dt := dxµ(χ(t))/dt = Y µ(χ(t)), X(t) := X(χ(t)), Y (t) := Y (χ(t)), A(t) := ¯ ¯ (L) (L) A(χ(t)), θ(t) := θ(χ(t)), Γ (t) := Γ (χ(t)), and ρθ¯(A) is the representation of A acting on θ¯, which takes the form ¯ ¯(L) ¯ ρθ¯(A) θ = Γ θ (6.2)

24 in accordance with (2.30). Meanwhile, applying (5.16) again, we have

h  (L) i [X,Y ]sv = σ , θ¯ ([X,Y ]),D ([X,Y ])v . (6.3) ∇ i i V i

µ ν µ ν µ Note that [X,Y ] = X ∂νY Y ∂νX implies −

µ µ µ η([X,Y ]) ηµ[X,Y ] = ηµX[Y ] ηµY [X ] (6.4) ≡ − for any one-form η. Putting (6.1), (6.3), and (6.4) together, we have

(E) R (X,Y )sv := [ X , Y ]sv [X,Y ]sv ∇ ∇ − ∇ h  (L)   (L) (L) (L)  i = σ , dθ¯ + Γ¯ θ¯ (X,Y ), ρV dΓ + Γ Γ (X,Y ) v i i i ∧ i i i ∧ i   = σi, T¯i(X,Y ), ρV (Ri(X,Y )) v , (6.5) where T t(n, R) Ω2(M) is defined as ∈ ⊗ T := D(L)θ dθ + [Γ(L), θ], (6.6) ≡ or equivalently (L) (L) T¯ ρn(T ) := D¯ θ¯ dθ¯ + Γ¯ θ,¯ (6.7) ≡ ≡ ∧ and R gl(n, R) Ω2(M) is defined as ∈ ⊗ 1 R := dΓ(L) + [Γ(L), Γ(L)] dΓ(L) + Γ(L) Γ(L). (6.8) 2 ≡ ∧ Note that (6.6) is identical to the Cartan torsion two-form as defined in (1.1a) and (6.8) identical to the Riemann curvature two-form as defined in (1.1b). If X, Y, v are replaced by X0(p) = f(p)X(p), Y 0(p) = g(p)Y (p), and v0(p) = h(p)v(p), where f, g, h Ω0(M) are arbitrary scalar fields, it follows from (6.5) that ∈

(E) 0 0  0 0 0 0 0 R (X ,Y )sv0 = σi, T¯i(X ,Y ), ρV (Ri(X ,Y )) v   = σi, fg T¯i(X,Y ), fgh ρV (Ri(X,Y )) v

(E) fg R (X,Y )shv, (6.9) ≡ since T and R are two-forms, which by definition are linear when acting on tangent vectors. (E) R (X,Y )sv is said to be linear in X, Y , and v in the above sense. Although we define (E) R (X,Y )sv by considering two vector fields X,Y X (U) and a V -value field v V ∈ ∈ ⊗ 25 δaX − [δaX, δbY ] −

δbY − δbY

p δaX

(E) FIG. 2: A closed curve used to understand the geometric meaning of R (δaX, δbY )sv. Given by the two vector fields X,Y X (U), the curve is edged with the vectors δaX, δbY , and the ∈ ± ± negative of the “closer of the quadrilateral” [δaX, δbY ] = δaδb [X,Y ].

0 (E) Ω (U), it turns out that, because of the linearity, R (X,Y )sv is well defined for any two 19 tangent vectors X,Y TpM and any V -value v V . ∈ ∈ Following the same reasoning explained in Ch. 11 of [28], the geometric meaning of the (E) “curvature” operator R (X,Y ) acting on a section sv can be understood in terms of holonomy as follows. Consider an infinitesimal closed curve spanned by δaX and δbY as depicted in Fig. 2. Starting at the initial point p with an initial value of sv(p) as given by

(5.3), if we parallel transport sv(p) around the closed curve until we come back to the point 0 0 p, we end up with a new value sv0 (p). The difference between sv0 (p) and sv(p) is give by

0  0 0  s 0 (p) sv(p) σ (p), (ξ¯ (p) ξ¯ (p), v v) v − ≡ i i − i − (E) 2 = R (δaX, δbY )sv + O(δ )

   2 = δaδb σi, (Ti(X,Y ), ρV Ri(X,Y ) v) + O(δ ), (6.10)

0 where, as we are comparing s 0 and sv at the same point p U, we choose the same local v ∈ 0 section σ Γ(U, P ) to represent both s 0 and sv. i ∈ v Under the parallel transport around an infinitesimal closed curve, R gives a value of gl(n, R) that dictates how a given vector v V is linearly changed to v0. The geometric ∈ picture of R in terms of holonomy in the “vector” part of the affine-vector bundle is exactly the same as that of the standard GL(n, R) Yang-Mills theory on a conventional vector bundle. Analogously, on the other hand, T gives a value of t(n, R) that dictates how a given reference point ξ¯ is displaced to ξ¯0 under the parallel transport around the closed curve.

19 (E) It would not be the case if we define R (X,Y )sv as [ X , Y ]sv without taking into account the term ∇ ∇ sv. See Ch. 11 and especially Exercise 11.2 of [28] for more discussions related to this point. −∇[X,Y ]

26 This provides a clear geometric picture of T in terms of holonomy in the “affine” part of the affine-vector bundle.

Also note that whereas the local section sv Γ(U ,E) in (5.3) involves two gauge choices ∈ i ¯ n — σi Γ(Ui,P ) and ξi : Ui R , the honolomy of sv around an infinitesimal closed curve ∈ → is given by T and R, both of which are covariant under a local GL(n, R) transformation and invariant under a local translation. (Notably, the honolomy of sv around a closed curve is completely independent of the gauge ambiguity of ξ¯.) This suggests that both the holonomies in the “affine” and “vector” parts are physical and yield observational consequences, which will be discussed in the next section.

VII. OBSERVATIONAL CONSEQUENCES

The observational consequences of the Reimann curvature R are well understood. Typ- ically, R = 0 imposes the geodetic effect and frame-dragging on the dynamics of a matter 6 field (see [33] for the experimental verification). Additionally, even if the trajectories of a matter field are confined to a region where R = 0, a two-path interference experiment will still yield different interference patterns `ala the Aharonov-Bohm effect, in response to the holonomy of Γ(L) along the closed curve composed of the two worldline paths. The holonomy H Γ(L) of Γ(L) along a finite closed curve γ is defined as the pathwise integral P e γ , which is R equal to evγ — i.e. the exponential of the flux integral of R over the surface enclosed by γ — according to (6.5). It is possible to have R = 0 everywhere along γ but still have nontrivial holonomy, thus yielding an observable result of a nontrivial interference pattern [34–36]. On the other hand, the observational consequences of the Cartan torsion T are much less understood. In the theories of first-order Lagrangians, it has been shown that torsion is inextricably bound to spinning matter and vanishes wherever there is no spinning matter [7]. Consequently, it does not propagate through vacuum and only gives rise to a “direct interaction” in contact with spin source for the dynamics of a matter field. Even in the theories of second-order Lagrangians including torsion-torsion coupling, the dynamics of torsion does not change in an essential way (see [37] for more discussions). Therefore, the Cartan torsion has no “at-a-distance” effects analogous to the geodetic effect and frame- dragging of the Riemann curvature.

27 Nevertheless, the geometric meaning of T in terms of holonomy leads us to anticipate a kinematical effect of torsion that in principle can be measured `ala the Aharonov-Bohm effect, analogous to the case of R. If the surface enclosed by the two paths contains spinning H H θ θ T 20 matter, we may have nontrivial holonomy P e γ e γ = evγ according to (6.5). The ≡ logarithm of the holonomy, H θ¯ = T¯, is to be understood as the displacement vector γ vγ between the initial reference point and the final reference point under the parallel transport around γ. Although the gauge choice of the reference point ξ¯ is arbitrary, the displacement vector associated with the holonomy is non-arbitrary and physical. Consider that a local Lorentz frame starting at a given point is parallel transported along two worldline paths separately. If the two paths join again at a destination point, the two transported Lorentz frames overlap but their coordinate origins might not coincide, and the difference between the frame origins is described by the displacement vector of the holonomy along the closed curve composed of the two paths. Since the interference experiment measures the two wavefunctions along the two paths superimposed at the destination point, the difference between the origins of the two overlapped frames and consequently the displacement vector of the honolomy should be manifested in the interference pattern. Therefore, despite the fact that torsion only gives a direct-contact interaction, its flux can still be measured `ala the Aharonov-Bohm effect without being probed directly. It should be emphasized that the conjecture given above is based purely on kinemati- cal considerations and remains speculative. To ensure that this effect is indeed physically measurable, one has to take into account dynamical considerations to see whether the dy- namics dictates that the Lorentz frames are parallel transported in accordance with the A(n, R)-connection. More precisely, the Lorentz frames should be operationally defined with reference to some physical objects, namely, matter fields, and in principle it is the dynamics of matter fields that dictates how the Lorentz frames are transported. For the dynamics of matter fields in the gauge theory of a general G-connection (with the gauge group G being different from SO(1, 3)), it has been shown that the equations of motion of ordinary matter fields do not follow the parallel transport law in accordance with the G-connection, but instead they follow the parallel transport law either of the Levi-Civita connection (i.e., the Riemannian connection of the metric) alone or of a hybrid

20 Because [P ,P ] = 0, the pathwise symbol P here is superfluous and consequently we have H θ = T . i j γ vγ

28 connection resulting from both the Levi-Civita connection and the general G-connection [38–42]. Furthermore, it has been argued that, by starting from the action principle, there is no clear way to formulate the dynamics of matter fields whereby the matter fields follow the parallel transport law of the G-connection (see Section VI.C of [43]). The “no-go” results from dynamical considerations of matter fields seem to suggest that the conjectured Aharonov-Bohm-like effect is not measurable after all. However, we should bear in mind that, for the matter dynamics studied so far in the literature, a matter field µ ϕ(x ) is essentially treated as a local section of an associated vector bundle P ρ V , where × V is the vector space in which ϕ(xµ) resides and ρ is the linear presentation of G over V (recall Footnote 16 for the case of G = SO(1, n 1)). This treatment does not faithfully − reflect the affine structure of A(n, R) as argued in Sec. III, and cannot make sense of an affine vector ξ¯ as a reference point (recall Footnote 17). The results of our work suggest that a more suitable arena for the gauge theory of A(n, R) is given by an associated affine-vector n bundle P ρA F , where F = R V is the affine-vector space and ρA is the affine (nonlinear) × × representation of A(n, R). If we attempt to formulate the dynamical theory of matter fields in the framework of an affine-vector bundle, a matter field should be treated as a local section of P ρ F , instead of that of P ρ V . That is, a matter field is locally described by × A × the pair (ξ¯(xµ), ϕ(xµ)), rather than ϕ(xµ) alone. With the inclusion of the accompanying field ξ¯(xµ), the resulting dynamics of matter fields could be qualitatively different from the results that have been studied so far. Therefore, it would be premature to jump to the conclusion that the Aharonov-Bohm-like effect is not physically measurable. Of course, it remains a challenging open question how the dynamical theory of matter fields should be formulated through affine-vector bundles. Especially, it is unclear how the reference variable ξ¯(xµ) should be included in the action. Should it be treated as a dynamical variable or merely an algebraic one? Depending on the details of formulation, the dynamics might result in a complicated form of transport upon the Lorentz frames. However complicated, nevertheless, it is still likely to give rise to a measurable Aharonov- Bohm-like effect, although the detailed form of the effect could be quantitatively different from what we conjectured solely from kinematical considerations. As regards the dynamics of matter fields, one important issue in relation to the weak equivalence principle is whether the variable ξ¯(xµ) should be treated as universal for different matter species or different matter species should be associated with different ξ¯(xµ) variables. If the former is the case,

29 the displacement vector inferred from the interference pattern of the Aharonov-Bohm-like effect is expected to be identical regardless of the matter species used for the experiment. If the latter is the case, the weak equivalence principle is not fully satisfied in the sense that we cannot directly identify θ¯ appearing in (5.15) as the coframe one-form, or alternatively different matter species are said to see intricately different spacetime structures of different coframe one-forms. In any case, studying the dynamics in the framework of an affine-vector bundle will shed new light on and open new possibilities for various important issues that have been studied in the literature.

VIII. SUMMARY AND DISCUSSION

Investigating the Poincar´esymmetry more carefully in view of the Einstein equivalence principle motivates us to propose the framework of an associated affine-vector bundle, which contains an affine vector ξ¯ on the fiber serving as a reference point and thus provides a more suitable arena for the affine group A(n, R). The associated affine-vector bundle is rigorously defined in the same spirit of the formal quotient-space construction for an associated vector

bundle. Choosing a local section sv Γ(U, E) on the affine-vector bundle E amounts to the ∈ familiar gauge fixing of choosing a local section σ Γ(U, P ) on the principle bundle P plus i ∈ the extra gauge fixing of choosing an arbitrary reference point ξ¯i, The formal definitions of the parallel transport and the covariant derivative on an asso- ciated vector bundle in the Ehresmann-connection approach can be naturally generalized to the associated affine-vector bundle. Rigorously deriving the covariant derivative of a local

section sv on the affine-vector bundle, we obtain the result in (5.15) and (5.16). Remarkably, θ and Γ(L) naturally appear as the gauge potentials for the “affine” part and the “vector” part, respectively. Because θ¯ transforms as a vector under a local A(n, R) transformation and the weak equivalence principle suggests that θ¯, being a measure of the parallel transport of a reference point, has to be universal regardless of the matter content, the gauge potential θ is to be identified as the coframe one-form field.

Applying the covariant derivative twice on the local section sv, we rigorously derive the (E) curvature operator R (X,Y ) acting on sv as given in (6.5). As expected, the Cartan torsion two-form T and the Riemann curvature two-form R naturally appear as the gauge field strengths for the “affine” part and the “vector” part, respectively. The geometric

30 meanings of T and R in terms of holonomy become clear. Our approach rigorously derives the appealing parallel between R and T from first prin- ciples without any ad hoc prescriptions, and provides a clear geometric and physical picture of them. Believing that the affine-vector bundle is more fundamental than a conventional vector bundle on the grounds that any local Lorentz frame shall manifest the full symme- try of the Poincar´egroup including its translational part, we arrive at a conjecture about a kinematical effect of the Cartan torsion that in principle can be measured `ala the Aharonov- Bohm effect. The exact quantitative prediction of this effect, however, might depend on the dynamical theory of matter fields. It should be remarked that, as noted in the last sentence of Sec. III, our analysis is carried out in the broader context of the affine group A(n, R), but it has not addressed the issue of how the gauge group A(n, R) is reduced to the Poincar´egroup. In other words, we view the coframe field θ¯ defined in (3.19) simply as a universal vector-valued one-form, but so far have not takeen into account the important fact that θ gives rise to the spacetime metric tensor via i j gµν = ηije µe ν. (8.1)

It is curious why the coframe field has two very different geometric meanings: as the gauge potential associated with translation and as the “square root” of the metric tensor. Some theories that include the nonmetricity field as an additional dynamical variable [1–3] might eventually provide a dynamical explanation for the dual role of θ. Although this paper does not consider this question at all, the clear geometric and physical picture we have obtained in the framework of an associated affine-vector bundle might offer valuable new insight about the intriguing relation between the local translational symmetry and the spacetime metric. Finally, we comment that the attention of this paper is focused on the kinematical aspects of MAG, and we leave the dynamical aspects of formulating the Lagrangian of MAG for future research. Akin to what we have observed for the covariant Dirac equation and the covariant Dirac Lagrangian in Sec. III, various Lagrangian actions of gravity, particularly the Einstein-Cartan action, formulated in the context of MAG are invariant under local Lorentz transformations as well as under diffeomorphisms but not under local translations (see Sec. 3.2.3 of [1]). In this sense, they are not really cast as “true” gauge theories of the affine group. According to Trautman’s idea that the affine-vector-valued field ξ¯ can be viewed as a “generalized Higgs field” [44], perhaps it is possible to formulate an explicit

31 spontaneous symmetry breaking mechanism responsible for “hiding” the local translational symmetry in the Lagrangian action. There have been various approaches in this direction (see e.g. [45] and the recent work [46]), but a satisfactory theory is still missing. The formalism of the associated affine-vector bundle we have devised might provide a better mathematical framework for implementing such a theory. As has been remarked in the end of Sec. VII, investigation of the dynamical theories in the framework of an affine-vector bundle will offer new insight into various important issues in the literature.

Acknowledgments

The authors would like to thank two anonymous reviewers for their valuable suggestions, which have helped to improve the manuscript greatly. This work was supported in part by the Ministry of Science and Technology, Taiwan under the Grants MOST 107-2119-M-002- 031-MY3 and MOST 109-2112-M-110-021.

[1] M. Blagojevic, Gravitation and Gauge Symmetries. Bristol: IOP Publishing, 2002. [2] E. W. Mielke, Geometrodynamics of Gauge Fields: On the Geometry of Yang-Mills and Grav- itational Gauge theories. Mathematical Physics Studies, Berlin: Springer, 2017. [3] F. Gronwald and F. W. Hehl, “On the gauge aspects of gravity,” in International School of Cosmology and Gravitation: 14th Course: Quantum Gravity, pp. 148–198, 5 1995, [arXiv:gr- qc/9602013 [gr-qc]]. [4] M. Blagojevi´cand F. W. Hehl, Gauge Theories of Gravitation: A Reader with Commentaries. London: Imperial College Press, 2013, [arXiv:1210.3775 [gr-qc]]. [5] F. W. Hehl and W.-T. Ni, “Inertial effects of a Dirac particle,” Physical Review D, vol. 42, no. 6, p. 2045, 1990. [6] Y. N. Obukhov, A. J. Silenko, and O. V. Teryaev, “Spin in an arbitrary gravitational field,” Physical Review D, vol. 88, no. 8, p. 084014, 2013, [arXiv:1308.4552 [gr-qc]]. [7] F. W. Hehl, P. Von der Heyde, G. D. Kerlick, and J. M. Nester, “General relativity with spin and torsion: Foundations and prospects,” Reviews of Modern Physics, vol. 48, no. 3, p. 393, 1976.

32 [8] F. W. Hehl, J. D. McCrea, E. W. Mielke, and Y. Ne’eman, “Metric-affine gauge theory of grav- ity: Field equations, Noether identities, world spinors, and breaking of dilation invariance,” Physics Reports, vol. 258, no. 1-2, pp. 1–171, 1995, [arXiv:gr-qc/9402012 [gr-qc]]. [9] A. Trautman, “On the structure of the Einstein-Cartan equations,” in Differential Geometry, Symposia Mathematica, vol. 12, pp. 139–162, London: Academic Press, 1973. [10] R. W. Sharpe, Differential geometry: Cartan’s generalization of Klein’s Erlangen program, vol. 166. Berlin: Springer, 2000. [11] S. Sternberg, Curvature in mathematics and physics, New York: Dover, 2013. [12] T. Kawai, “A Poincar´egauge theory of gravity,” General Relativity and Gravitation, vol. 18, no. 10, pp. 995–1018, 1986. erratum 19 (1987) 1285. [13] T. Kawai and N. Toma, “An extended new general relativity as a reduction of Poincar´e gauge theory of gravity: Generators of internal and coordinate transformations,” Progress of Theoretical Physics, vol. 85, no. 4, pp. 901–926, 1991. [14] E. A. Lord, “A unified approach to the gauging of space-time and internal symmetries,” General Relativity and Gravitation, vol. 19, no. 10, pp. 983–1002, 1987. [15] A. L´opez-Pinto, A. Tiemblo, and R. Tresguerres, “Ordinary matter in non-linear affine gauge theories of gravitation,” Classical and Quantum Gravity, vol. 12, no. 6, p. 1503, 1995, [arXiv:gr- qc/9412045 [gr-qc]]. [16] J. Julve, A. L´opez-Pinto, A. Tiemblo, and R. Tresguerres, “Nonlinear gauge realization of spacetime symmetries including translations,” General Relativity and Gravitation, vol. 28, no. 6, pp. 759–768, 1996, [arXiv:gr-qc/9412043 [gr-qc]]. [17] E. J. Cartan, On manifolds with an affine connection and the theory of general relativity. Napoli: Bibliopolis, 1986. English translation from the original in French. [18] E. W. Mielke, J. D. McCrea, Y. Ne’eman, and F. W. Hehl, “Avoiding degenerate coframes in an affine gauge approach to quantum gravity,” Physical Review D, vol. 48, no. 2, p. 673, 1993, [arXiv:gr-qc/9301023 [gr-qc]]. [19] K. Hayashi and T. Nakano, “Extended translation invariance and associated gauge fields,” Progress of Theoretical Physics, vol. 38, no. 2, pp. 491–507, 1967. [20] K. Hayashi and T. Shirafuji, “Gravity from Poincar´egauge theory of the fundamental particles. i: General formulation,” Progress of Theoretical Physics, vol. 64, no. 3, pp. 866–882, 1980. [21] T. Shirafuji and M. Suzuki, “Gauge theory of gravitation: A unified formulation of Poincar´e

33 and (anti-) de Sitter gauge theories,” Progress of theoretical physics, vol. 80, no. 4, pp. 711–730, 1988. [22] F. M¨uller-Hoissen,“A gauge theoretical approach to space-time structures,” in Annales de l’IHP Physique th´eorique, vol. 40, pp. 21–34, 1984. [23] E. Ivanov and J. Niederle, “On gauge formulations of gravitation theories,” in Group Theo- retical Methods in Physics, pp. 545–551, Berlin: Springer, 1980. [24] G. Grignani and G. Nardelli, “Gravity and the Poincar´egroup,” Physical Review D, vol. 45, no. 8, p. 2719, 1992. [25] G. Sardanashvily and M. Gogberashvily, “The dislocation treatment of gauge fields of space- time translations,” Modern Phys. Lett. A, vol. 2, p. 609, 1987. [26] J. Hennig and J. Nitsch, “Gravity as an internal Yang-Mills gauge field theory of the Poincar´e group,” General Relativity and Gravitation, vol. 13, no. 10, pp. 947–962, 1981. [27] H. Morales, A. M. T´ecotl, and E. Mielke, “Geometric phases and translations,” in 1st Mexican School on Gravitation and Mathematical Physics (A. Macıas et al., eds.), Singapore: World Scientific, 1995. [28] C. W. Misner, K. Thorne, and J. Wheeler, Gravitation. San Francisco: W. H. Freeman, 1973. [29] T. Koivisto, M. Hohmann, and T. Zlo´snik,“The General Linear Cartan Khronon,” Universe, vol. 5, no. 6, p. 168, 2019, [arXiv:1905.02967 [gr-qc]]. [30] M. Nakahara, Geometry, Topology and Physics. Bristol: IOP Publishing, 2003. [31] A. Delhom, “Minimal coupling in presence of non-metricity and torsion,” Eur. Phys. J. C, vol. 80, no. 8, p. 728, 2020, [arXiv:2002.02404 [gr-qc]]. [32] J. B. Jim´enez,L. Heisenberg, and T. Koivisto, “The coupling of matter and spacetime ge- ometry,” Classical and Quantum Gravity, vol. 37, no. 19, p. 195013, 2020, [arXiv:2004.04606 [hep-th]]. [33] C. Everitt et al., “Gravity Probe B: Final Results of a Space Experiment to Test General Relativity,” Phys. Rev. Lett., vol. 106, p. 221101, 2011, [arXiv:1105.3456 [gr-qc]]. [34] J. S. Dowker, “A gravitational Aharonov-Bohm effect,” Il Nuovo Cimento B (1965-1970), vol. 52, no. 1, pp. 129–135, 1967. [35] V. B. Bezerra, “Gravitational analogue of the Aharonov-Bohm effect in four and three dimen- sions,” Phys. Rev. D, vol. 35, pp. 2031–2033, Mar 1987. [36] V. B. Bezerra, “Some remarks on loop variables, holonomy transformation, and gravitational

34 Aharonov-Bohm effect,” Annals of Physics, vol. 203, no. 2, pp. 392–409, 1990.

[37] V. D. Heyde, “The equivalence principle in the U4 theory of gravitation,” Lettere al Nuovo Cimento (1971-1985), vol. 14, no. 7, pp. 250–252, 1975. [38] J. Audretsch, “Dirac Electron in Space-times With Torsion: Spinor Propagation, Spin Pre- cession, and Nongeodesic Orbits,” Phys. Rev. D, vol. 24, pp. 1470–1477, 1981. [39] J. A. R. Cembranos, J. G. Valcarcel, and F. J. Maldonado Torralba, “Fermion dynamics in torsion theories,” JCAP, vol. 04, p. 039, 2019, [arXiv:1805.09577 [gr-qc]]. [40] K. Hayashi, K. Nomura, and T. Shirafuji, “Spin precession in space-time with torsion,” Prog. Theor. Phys., vol. 84, pp. 1085–1099, 1990. [41] H. Rumpf, “Quasiclassical Limit of the Dirac Equation and the Equivalence Principle in the Riemann-Cartan Geometry,” in Cosmology and Gravitation. Spin, Torsion, Rotation, and Supergravity, Proceedings, NATO Advanced Study Institute, Erice, Italy, May 6–18, 1979, NATO Science Series Ser. B, vol. 58, pp. 93–104, New York: Plenum, 1980. [42] F. W. Hehl, “Spin and torsion in general relativity: I. foundations,” General Relativity and Gravitation, vol. 4, no. 4, pp. 333–349, 1973. [43] J. B. Jim´enezand A. Delhom, “Instabilities in metric-affine theories of gravity with higher or- der curvature terms,” The European Physical Journal C, vol. 80, Jun 2020, [arXiv:2004.11357 [gr-qc]]. [44] A. Trautman, “The geometry of gauge fields,,” Czech. J. Phys. B, vol. 29, pp. 107–116, 1979. [45] D. K. Wise, “MacDowell-Mansouri gravity and Cartan geometry,” Class. Quant. Grav., vol. 27, p. 155010, 2010, [arXiv:gr-qc/0611154 [gr-qc]]. [46] Y. N. Obukhov and F. W. Hehl, “General relativity as a special case of Poincar´egauge gravity,” Phys. Rev. D, vol. 102, no. 4, p. 044058, 2020, [arXiv:2007.00043 [gr-qc]].

35