Home Mathematics On Twisted Spacetimes: a new class of Galilean cosmological models
Article Open Access

On Twisted Spacetimes: a new class of Galilean cosmological models

  • Daniel de la Fuente EMAIL logo , Rafael M. Rubio and Jose Torrente-Teruel
Published/Copyright: October 1, 2025

Abstract

Within the generalized Newton-Cartan theory, Galilean Twisted spacetimes are introduced as dual models of the well-known relativistic twisted spacetimes. As a natural generalization, torqued vector fields in Galilean spacetimes are defined, showing that the local structure of a Galilean spacetime admitting a timelike torqued vector field is given by a Twisted spacetime. In addition, several results assuring the global splitting as Twisted spacetime are obtained. On the other hand, completeness of free falling observers is studied, as well as geodesic completeness.

MSC 2020: 53Z05; 53C80; 53B50

1 Introduction

The geometric formulation of Newton’s theory of gravity began with E. Cartan in the early twentieth century [1], [2]. However, it was not until the second half of the century that the so-called Newton-Cartan theory began to receive significant attention from various authors. In the 1960s, Malament studied the Lorentzian and Galilean groups, proving that Newtonian gravity constitutes a limit of General Relativity and how free particles are represented by geodesic curves of certain connections in both theories [3]. Additionally, Trautman showed the existence of a symmetric connection in both theories [4], arising due to the local nature of physical laws and simultaneously encoding the inertia principle and gravitational forces. Concretely, classical Newtonian gravitation is formulated as a covariant theory, showing that certain results previously considered characteristic or singular of the theory of Relativity are also shared by the (geometric) Newton-Cartan gravitational theory. In fact, Newtonian gravity is explained by the curvature of a connection in the spacetime, although in this case it no longer arises from any semi-Riemannian metric. In the geometric formulation of gravity, the spacetime structure is dynamic in the sense that it participates in the development of physics rather than being a fixed backdrop (see [3] and references therein).

Thereafter, several steps were taken towards the generalization of this theory, such as the introduction of new Newtonian models satisfying the cosmological principle [5]. Thus, obtaining the Galilean analogue of relativistic Robertson-Walker spacetimes. In [6], Newtonian theory achieves the status of a gauge theory and is shown to be a limit of the Einstein-Klein-Gordon theory. This view as a limit of General Relativity continued to develop during these years [7], concluding also that the results obtained from Newton-Cartan cosmology were similar on not too large scales to those derived from Einstein’s theory [8], [9].

In the last few decades, the simplicity of certain Newtonian models in Newton-Cartan theory, in contrast to the computational complexity of General Relativity, has led to new applications of this non-relativistic theory in Cosmology [10], Hydrodynamics [11], quantum collapse [12], AdS/CFT correspondence [13], condensed matter systems [14], quantum Hall effect [15] and other related phenomena. The many applications have motivated a review of the geometric structures associated with generalized Newton-Cartan theory [16] and new extensions such as Newton-Cartan Supergravity [17] and non-relativistic Strings [18], [19]. Furthermore, new models have been obtained, including Galilean Generalized Robertson-Walker spacetimes [20], the Newtonian Gödel spacetime [21], standard stationary Galilean spacetimes [22], as well as the embedding of these geometries in higher dimensional Lorentzian manifolds [23], [24], [25], [26], [27].

In this manuscript (Sect. 3), the cosmological models of Galilean Generalized Robertson-Walker spacetimes (GGRW) [20] are introduced in a wider family of Galilean model spacetimes. Definition 3.1 states

Given a real interval I R , a Riemannian manifold (F, h) and a positive smooth function fC (I × F), a Galilean Twisted (GT) spacetime is given by the tuple I × F , Ω = d t , g = f 2 π F * h , where t : = π I and ∇ is the only symmetric Galilean connection on I × F such that t t = 0 and Rot∂t = 0.

According to the cosmological principle, classical Robertson-Walker spacetimes (see, e.g., [28]) model a spatially homogeneous and isotropic universe composed of matter (galaxies) subject only to the action of its own gravity. Thus, the “absolute space at each moment in absolute time” must be locally Euclidean with constant sectional curvature. Furthermore, as large-scale experimental measurements (redshift) indicate, the expansion is assumed to be homogeneous, i.e., the expansion rate is the same at all points at a fixed (absolute) time. However, while the hypothesis of spatial homogeneity and isotropy may be a reasonable first approximation to the large scale structure of the universe, it could not be appropriate when we consider a more accurate scale. This led to the introduction of the family of GGRWs spacetimes, where the absolute space is modeled by an arbitrary Riemannian manifold, which does not necessarily have to be of constant curvature. Finally, the Twisted spacetimes presented here assume a non-homogeneous expansion rate, which means that it depends not only on the moment in time but also on the point in space. In other words, the “Hubble constant” is no longer constant.

Twisted products have been studied in the relativistic setting [29], [30], [31], [32]. They have been analyzed in relation to geometric models of fluids, resulting in a characterization in terms of the stress-energy tensor of an imperfect fluid [33] and obtaining the properties of static perfect fluids in Twisted Lorentzian spacetimes [34]. Similar results could be explored in the non-relativistic setting introduced here.

A Galilean Twisted (GT) spacetime possesses an infinitesimal symmetry given by the existence of a timelike irrotational and (spatially) conformally Leibnizian field of observers. Several geometrical properties, as well as physical interpretations of this family of spacetimes, are given in Section 4 and Section 5. We concretely focus on the completeness of its free falling observers. Our main result (Theorem 4.9) asserts

Let (F, h) be a complete Riemannian manifold and let f C ( R × F ) be a positive function. If inf  f > 0 and inf(f t /f) > − along finite times, then the GT spacetime M = R × F , Ω = d π R , g = f 2 π F * h , is geodesically complete.

From a physical perspective, this result is relevant since the completeness of the timelike trajectories is crucial in order to have a consistent theory (without observers that appear or disappear suddenly).

Section 6 is devoted to the study of Galilean spacetimes admitting a torqued vector field (see Definition 6.2), which is closely related to the previously introduced family of irrotational conformally Leibnizian vector fields [20]. The main result in this section is Theorem 6.5, which gives the following statement.

Let (M, Ω, g, ∇) be a Galilean spacetime. If it admits a Galilean torqued vector field K ∈ Γ(TM), then for each pM, there exists an open neighborhood of p, U , and a Galilean diffeomorphism Ψ : N U , where N is a GT spacetime.

Finally, Section 7 is devoted to face splitting-type problems, i.e., under which geometrical assumptions a Galilean spacetime decomposes globally into a Galilean Twisted spacetime. Theorem 7.5 states

Let (M, Ω, g, ∇) be a simply connected Galilean spacetime. It admits a global decomposition as a GT spacetime if and only if there exists a Galilean torqued vector field K ∈ Γ(TM) such that the associated field of observers, Z  : = K/Ω(K), is a uniform global generator.

Moreover, if the leaves of the foliation induced by Ω are compact, we can ensure again a global splitting result (Theorem 7.6).

2 Preliminaries

Let M be a (n + 1)-dimensional smooth manifold. As usual, it is assumed connected, Hausdorff and paracompact. A Leibnizian spacetime [16] is a triad (M, Ω, g) where (Ω, g) is a Leibnizian structure defined by a global non-vanishing one-form on M, Ω ∈ Λ1(M), and a positive definite metric g on the vector bundle defined by the kernel of Ω. Concretely, if AnΩ := {vTM   : Ω(v) = 0} is the n-distribution induced by Ω, then g is a smooth, bilinear, symmetric and positive definite 2-covariant tensor

g : Γ ( An Ω ) × Γ ( An Ω ) C ( M ) ,

where Γ(AnΩ) denote the subset of sections V ∈ Γ(TM) such that V p ∈ AnΩ, ∀pM.

Given a Leibnizian spacetime (M, Ω, g), the points pM are called events. If pM, Ω p is known as the absolute clock at p and An Ω p , g p is called the absolute space at p (see for example [35]). Adopting relativistic terminology, a tangent vector vT p M is said to be spacelike (resp. timelike) if Ω p (v) = 0 (resp. Ω p (v) ≠ 0). Moreover, a timelike vector v is called future-pointing (resp. past-pointing) if Ω p (v) > 0 (resp. Ω p (v) < 0).

On the other hand, an observer is defined as a smooth curve γ : I R M , being I an interval, such that its velocity is a normalized timelike future-pointing vector (standard timelike unit [35]), i.e., Ωγ(s)(γ′(s)) = 1, ∀sI. The parameter sI is the proper time of the observer γ. A field of observer Z ∈ Γ(TM) is a smooth vector field whose integral curves are observers, equivalently, Ω(Z) = 1.

In this non-relativistic setting, the synchronizability of observers is intrinsic to the Leibnizian structure. If Ω ∧dΩ = 0 (i.e., the smooth distribution AnΩ is integrable), the spacetime is called locally synchronizable. In this case, Frobenius Theorem (see [36]) states that there exists a foliation consisting of spacelike hypersurfaces. Moreover, locally Ω = λdT, for certain smooth functions λ > 0 and T, and the hypersurfaces of the foliation are exactly {T = constant}. Thus, every observer can locally synchronize its proper time with the so-called “compromise time” T. If the absolute clock is closed, dΩ = 0, the Leibnizian spacetime is said to be proper time locally synchronizable. Under this assumption, locally we can express Ω = dT, for certain smooth function T, and the observers can synchronize their proper times up to a constant. In the case Ω = dT, for TC (M), we will suppose every observer is parameterized by the absolute time T. Notice that, if M is simply-connected and Ω is closed, we can ensure the existence of an absolute time function making use of the Poincaré Lemma (see, for instance, [37]).

If the integrability condition Ω ∧dΩ = 0 does not hold, then the causal structure is not well-defined. In fact, if Ω(p) ∧dΩ(p) ≠ 0 at some pM, Caratheodory’s Theorem ensures that there exists an open neighborhood of p in M where all events are connected by some spacelike curve (see, e.g., [38]). Thus, even locally the notion of absolute space cannot be considered.

Remark 2.1.

Remarkably, the notion of Galilean synchronizability depends only on the Leibnizian structure, in contrast to relativistic synchronizability which is relative to a fixed field of observers. In the latter case, the (local) synchronizability of a field of observers Z is related to the spacetime connection. In particular, if Z is synchronizable, its vorticity must vanish identically.

Let (M, Ω, g) and ( M ̃ , Ω ̃ , g ̃ ) be two Leibnizian spacetimes. A diffeomorphism λ : M M ̃ is called Leibnizian if it preserves the absolute clock and space, that is, λ * Ω ̃ = Ω and λ * g ̃ = g .

To give complete physical meaning to a Leibnizian spacetime, the principle of inertia has to be included via an affine connection ∇. It is necessary to consider ∇ compatible with the absolute clock Ω and the absolute space (AnΩ, g), i.e.,

  1. ∇Ω = 0, equivalently, X(Ω(Y)) = Ω(∇ X Y), for all X, Y ∈ Γ(TM),

  2. g = 0, that is, Z(g(V, W)) = g(∇ Z V, W) + g(V, ∇ Z W), for all Z ∈ Γ(TM) and V, W ∈ Γ(AnΩ).

Note that, thanks to condition (1), the right-hand side of item (2) is well defined. Any such ∇ is called a Galilean connection. Consequently, the Leibnizian spacetime does not determine a canonical Galilean connection. The tuple (M, Ω, g, ∇) is known as a Galilean spacetime. Recall that if its torsion tensor vanishes, a connection is said to be symmetric, i.e.,

Tor ( X , Y ) = X Y Y X [ X , Y ] = 0 , X , Y Γ ( T M ) .

From a physical viewpoint, it is often convenient to choose a symmetric connection because it is determined by the associated geodesics [[39], Add. 1, Ch. 6]. Observe that the closeness of the absolute clock and the torsion tensor are related [[16], Lemma 13]

(1) Ω Tor = d Ω .

From now on, we will only consider symmetric Galilean connections, then using equation (1), dΩ = 0.

Given two Galilean spacetimes (M, Ω, g, ∇) and ( M ̃ , Ω ̃ , g ̃ , ̃ ) , a diffeomorphism λ : M M ̃ is called Galilean if it is Leibnizian and additionally λ * ̃ = , i.e., ̃ d λ ( X ) d λ ( Y ) = X Y , for all X, Y ∈ Γ(TM).

On the other hand, given a Galilean spacetime (M, Ω, g, ∇) and a field of observers Z ∈ Γ(TM), it is well-known that the Galilean connection ∇ can be characterized using Z. The gravitational field induced by ∇ in Z is the spacelike vector field G = Z Z and similarly, the vorticity or Coriolis field is the 2-form ω = 1 2 Rot ( Z ) . The space of symmetric Galilean connections is mapped bijectively onto Γ(AnΩ) ×Λ2(AnΩ), giving ( G Z , ω Z ) (see [[16], Cor. 28]). Moreover, the converse is obtained through a formula ’à la Koszul’. From the tangent space splitting TM = ⟨Z⟩⊕AnΩ, we can consider

X Y = P Z X Y + X ( Ω ( Y ) ) Z , X , Y Γ ( T M )

where P Z X = X − Ω(X)Z is the natural spacelike projection for Z, then for each V ∈ Γ(AnΩ) [[16], eq. (13)],

(2) 2 g P Z X Y , V = X g P Z Y , V + Y g P Z X , V V g P Z X , P Z Y + 2 Ω ( X ) Ω ( Y ) g G Z , V + 2 Ω ( X ) ω Z ( P Z Y , V ) + 2 Ω ( Y ) ω Z ( P Z X , V ) + Ω ( X ) g Z , P Z Y , V g [ Z , V ] , P Z Y Ω ( Y ) g Z , P Z X , V + g [ Z , V ] , P Z X + g P Z X , P Z Y , V g P Z Y , V , P Z X g P Z X , V , P Z Y .

2.1 Spatially conformally Leibnizian spacetimes

Let us recall the notion of spatially conformally Leibnizian vector field, which appears in several classes of cosmological models in the context of the generalized Newton-Cartan theory (see, e.g., [20]) and will be key to obtaining some of our results.

Definition 2.2.

Let (M, Ω, g) be a Leibnizian spacetime and C a vector field satisfying

(3) Ω ( [ C , V ] ) = 0 for all V Γ ( An Ω ) .

The vector field C is called spatially conformally Leibnizian if the Lie derivative of the absolute space metric satisfies

(4) L C g = 2 ρ g ,

for some smooth function ρC (M). The function ρ is called conformal factor of g associated to C.

Notice that condition (3) ensures that the previous notion is well defined. As a direct consequence, we have

(5) C ( g ( V , W ) ) = 2 ρ g ( V , W ) + g ( [ C , V ] , W ) + g ( [ C , W ] , V ) ,

for all V, W ∈ Γ(AnΩ).

The next result provides another condition to guarantee that (3) holds for a vector field.

Proposition 2.3.

Let (M, Ω, g, ∇) be a Galilean spacetime with symmetric connection ∇. Then, a vector field K satisfies equation (3) if and only if the function Ω(C) is spatially invariant, i.e., V(Ω(C)) = 0, V Γ A n Ω .

The vector field C ∈ Γ(TM) is called spatially Leibnizian if the conformal factor ρ vanishes identically, or equivalently, if its local flows Φ s preserve the absolute space, i.e., Φ s * g = g . If, in addition, they also preserve the absolute clock, i.e., Φ s * Ω = Ω , then C is called Leibnizian.

2.2 Spacelike differential map and spacelike gradient

Given a smooth function φC (M) on a Leibnizian spacetime (M, Ω, g), define the spacelike gradient of φ to be the section grad φ ∈ Γ(AnΩ), determined by

g ( grad φ , V ) = d φ ( V ) , V Γ ( An Ω ) .

Implicitly, we consider Ω(grad φ) = 0. This notion is a generalization of the usual gradient in semi-Riemannian manifolds.

When a smooth manifold M splits as a product M = I × F, for I R an open interval and F certain smooth manifold, the differential of some φC (I × F) can be canonically split. We can identify T ( t , p ) ( I × F ) T t I × T p F R × T p F . Then, define

φ t C ( I × F ) , φ t ( t , p ) : = d φ ( t , p ) ( t ) , d F φ ( t , p ) : T p F R , d F φ ( t , p ) ( v ) : = d φ ( t , p ) ( v ̄ ) ,

where t is the canonical vector field associated to the projection onto the first component t = π I and v ̄ R × T p F is the horizontal lift of a vector vT p F. If (F, h) is a Riemannian manifold, the usual Riemannian gradient grad h applies. Thus, we will denote d h = d F to unify the notation although the differential map only depends on the differential structure. From now on, the horizontal lift of a vector field W ∈ Γ(TF) to a product I × F will be denoted by W ̄ Γ ( T ( I × F ) ) . Moreover, the induced differential on F can be lifted as

d h ̄ φ ( t , p ) : R × T p F R , d h ̄ φ ( t , p ) ( a t + v ̄ ) : = d h φ ( t , p ) ( v ) ,

for all a R and vT p F.

On the other hand, we say that φC (I × F) is bounded from below (resp. from above) along finite times if, [s1, s2] ⊂ I, there exists a constant C s 1 , s 2 R such that φ > C s 1 , s 2 (resp. φ < C s 1 , s 2 ) in [s1, s2] × F. A smooth function φ is bounded along finite times if it is bounded from below and from above along finite times. Obviously, inf  φ > A, A R , along finite times is equivalent to φ being bounded from below along finite times for constants greater than A.

3 Galilean Twisted Spacetimes

We introduce a new class of Galilean spacetimes, which extends the family of cosmological models known as Galilean Generalized Robertson-Walker spacetimes (see [5], [20]).

Definition 3.1.

Let I R , 0 ∈ I, be an open real interval, (F, h) be an n-dimensional connected Riemannian manifold, and fC (I × F) be a positive smooth function. A Galilean spacetime (M, Ω, g, ∇) is called Galilean Twisted (GT) spacetime if M = I × F, Ω = dπ I , g is the restriction to the vector bundle AnΩ of the following (degenerate) metric on M:

(6) g ̄ = f 2 π F * h ,

where π I , π F are the canonical projections onto the open interval I and the manifold F, respectively, and ∇ is the only symmetric Galilean connection on M such that

(7) t t = 0  and  Rot t = 0 ,

where t = /∂t is the global coordinate vector field associated with t : = π I .

Abusing the notation, we will denote g = f 2 π F * h . Analogously to the Riemannian twisted product, the Riemannian manifold (F, h) is called fiber.

Remark 3.2.

We are only considering symmetric Galilean connections ∇. However, a more general definition of Galilean Twisted spacetimes with torsion could be given. Once a field of observers Z is fixed, formula “a la Koszul” (2) gives a unique Galilean connection, ∇Tor, such that P Z ◦Tor = 0 and for which the coordinate field t is free falling and irrotational. The only extra elements of this extended family of Galilean models are locally synchronizable but no proper time locally synchronizable. Indeed, if dΩ = 0, then Ω◦Tor = 0 and therefore P Z ◦Tor ≡Tor; that is, symmetric and Z-symmetric are the same in the case of closed absolute clock.

Note that in the definition above, we continue to impose the condition that the connection ∇ is symmetric (due to its relevance in Physics, since such connections are completely determined by free-fall trajectories). However, a more general definition of Galilean Twisted spacetimes with torsion could be given: By Koszul type formula in [16], once an observer field Z is fixed, there exists a unique connection, ∇tor, for which the observer field t is free falling and irrotational in such a way that it is Z-symmetric, i.e., P Z ◦Tor = 0. However, as discussed in Section 2, only locally synchronizable spacetimes have physical meaning, then one of this type of spacetimes with nonzero torsion should be locally synchronizable but not proper time locally synchronizable, because in that case dΩ = 0, then, Ω◦Tor = 0 and, therefore, P Z ◦Tor ≡Tor; that is, symmetric and Z-symmetric are the same.

The coordinate vector field t is a field of observers in M, and the observers in t are called commovil observers in analogy with both the Galilean and relativistic Generalized Robertson-Walker spacetimes [20], [40]. Notice that the conditions (7) determine the Galilean connection ∇ (see equation (2)).

Example 3.3.

Consider the GT spacetime R × R n , d π 1 , g = f 2 π 2 * , , , where π 1 : R × R n R and π 2 : R × R n R n are the canonical projections onto the first and second factors, respectively, ⟨, ⟩ is the usual Euclidean product on R n and f C ( R × R n ) . If f does not depend on the fiber F = R n , i.e., f = f(t), t R , then we have a GGRW spacetime [20]. In the simplest case where f = constant, the Galilean connection is just the standard flat connection of R n + 1 .

4 Geodesic completeness of GT spacetimes

In this section we are going to study the completeness of the geodesics in a GT spacetime. In particular, this will allow us to know if the free falling observers live forever.

Remark 4.1.

From the compatibility condition of a Galilean connection ∇ with the absolute clock Ω in a Galilean spacetime, it is clear that the geodesics γ maintain their causal character. That is, Ω(γ′) is constant along the trajectory of γ. Then, the two relevant cases to consider will be spacelike geodesics (Ω(γ′) = 0) and free falling observers (Ω(γ′) = 1).

Given a geodesic γ : JM in a GT spacetime (M = I × F, Ω, g, ∇), where J R denotes an interval in the real line, we will be able to derive the corresponding equation of its projection π F (γ) on the fiber F. The following lemma gives us the necessary computations.

Lemma 4.2.

Let M = I × F , Ω = d π I , g = f 2 π F * h , be a GT spacetime. If W1, W2 ∈ Γ(TF), then

(8) t W ̄ = W ̄ t = u t W ̄ , W 1 ̄ W 2 ̄ = W 1 h W 2 ̄ + d h u ( W 1 ) W 2 ̄ + d h u ( W 2 ) W 1 ̄ h ( W 1 , W 2 ) grad h u ̄ ,

where u  : = log  f and grad h u denotes the usual gradient on the Riemannian manifold (F, h).

Proof.

It is enough to apply equation (2) for V = U ̄ , U ∈ Γ(TF), since a local reference frame { E i } i = 1 n in TF lift to a local reference frame { t } { E i ̄ } i = 1 n in TM. □

The previous formulae (8) allow us to obtain the equations of the projections of geodesics onto the fiber of a GT spacetime.

Remark 4.3.

Let γ be a geodesic γ  : JM in a GT spacetime (M = I × F, Ω, g, ∇), where J R denotes an interval in the real line. If γ is a free falling observer, we have

γ ( s ) = ( s , σ ( s ) ) and γ ( s ) = ( t γ ) ( s ) + σ ̄ ( s ) ,

where σ is a smooth curve in F. Consider X a smooth vector field, which extends σ′ on a neighbourhood of σ(s) ∈ F. Then the lift X ̄ extends σ ̄ and satisfies L t X ̄ = [ t , X ̄ ] = 0 . Thus, the geodesic equation turns out to be

0 = γ γ = γ ( t + X ̄ ) | γ = t t + σ ̄ t + γ X ̄ | γ = 2 t X ̄ + σ ̄ X ̄ | γ ,

where we have t t = 0 . Using (8), the free falling observers γ(s) = (s, σ(s)) are characterized by the following equation on (F, h),

(9) 0 = D h σ d s + 2 u t + d h u ( σ ) σ | σ | h 2 grad h u .

In the case γ(s) = (c, σ(s)), cI, is a spacelike geodesic, then γ = σ ̄ . The corresponding equation on (F, h) is given by

(10) 0 = D h σ d s + 2 d h u ( σ ) σ | σ | h 2 grad h u .

Note that this is now an autonomous equation in the Riemannian manifold (F, h). ⋄

For the rest of the section we suppose I = R since it is necessary to ensure the completeness of commovil observers.

Let (M, Ω, g, ∇) be a GT spacetime with M = R × F and g = f 2 π F * h . We can give a unique vector field G on T ( R × F ) such that, the timelike geodesic in M are represented by integral curves of G in T ( R × F ) . Indeed, let γ(s) = (s, σ(s)) be a free falling observer, s0J and (U, x1, …, x n ) a coordinate system in F, such that σ(s0) ∈ U. If we identify x(σ(s)) ≡ x(s) = (x1(s), …, x n (s)), equation (9) in R × U is given by

(11) x ̈ k ( s ) + i , j x ̇ i ( s ) x ̇ j ( s ) Γ i j k ( x ( s ) ) + 2 u t ( s , x ( s ) ) + l u x l ( s , x ( s ) ) x ̇ l ( s ) x ̇ k ( s ) α , β h α , β ( x ( s ) ) x ̇ α ( s ) x ̇ β ( s ) α h α k ( x ( s ) ) u x α ( s , x ( s ) ) = 0 ,

where k = 1, …, n and the Γ’s denote the Christoffel symbols of the Riemannian metric h. Making use of the standard procedure, we can introduce auxiliary variables v i = x ̇ i to convert the previous second-order system, in the following equivalent first-order system in twice the number of variables,

(12) x ̇ k ( s ) = v k ( s ) , v ̇ k ( s ) = i , j v i ( s ) v j ( s ) Γ i j k ( x ( s ) ) 2 u t ( t , x ( s ) ) l u x l ( s , x ( s ) ) v l ( s ) v k ( s ) + α , β h α , β ( x ( s ) ) v α ( s ) v β ( s ) α h α k ( x ( s ) ) u x α ( s , x ( s ) ) ,

for k = 0, …, n.

Treating (x1, …, x n , v1, …, v n ) as coordinates on T ( R × U ) , we recognize (12) as the equations for the flow of the vector field G X ( R × U ) given by

(13) G ( s , x , v ) = t ( s , x , v ) + k v k x k ( s , x , v ) i , j v i v j Γ i j k ( x ) + 2 u t ( s , x ) + l u x l ( s , x ) v l v k α , β h α , β ( x ( s ) ) v α v β α h α k ( x ) u x α ( s , x ) v k ( t , x , v ) .

To show that G define a global well-defined vector field on T ( R × F ) , it is enough to see that G acts on a smooth function φ by

G ( φ ) ( p , v ) = d d s s = 0 φ ( γ v ( s ) , γ ̇ v ( s ) )

where γ v (s) denote a geodesic satisfying γ(0) = p and γ ̇ v ( 0 ) = v , p R × F and v T p ( R × F ) . This last statement is clear from the previous coordinates expressions.

Making use of [[28], Lem. 1.56], we can assert that an integral curve ξ of G on an interval [ 0 , b [ R , for b < + , can be extended to b (as an integral curve) if and only if there exists a sequence in [0, b[, {s n }↗ b, such that {ξ(s n )} converges. Therefore, we obtain the following technical lemma.

Lemma 4.4.

A free falling observer γ  : [0, b[ → M, 0 < b < + , in a GT spacetime (M, Ω, g, ∇) can be extended to b as a geodesic of ∇ if and only if there exists a sequence in [0, b[, {s n }↗ b, such that γ ( s n ) converges in TM.

The following result gives natural conditions to ensure completeness of the free falling observers in a GT spacetime.

Theorem 4.5.

Let (F, h) be a complete Riemannian manifold and let f C ( R × F ) be a positive function. If inf  f > 0 and inf(f t /f) > − along finite times, then the trajectories of free falling observers in the GT spacetime M = R × F , Ω = d π R , g = f 2 π F * h , are complete.

Proof.

Given 0 < b < + , consider a free falling observer

γ : [ 0 , b [ R × F , γ ( s ) = ( s , σ ( s ) ) ,

where σ: [0, b[ → F is a smooth curve in F. The conditions on the function f admit to applying the following argument for past extensibility analogously. Hence, if we show γ is extensible towards b, the proof will be complete.

From equation (9) we have

d d s h ( σ , σ ) = 2 2 u t + d h u ( σ ) h ( σ , σ ) .

Note that

d d s ( u γ ) = d u ( γ ) = d u t + σ ̄ = u t + d h u ( σ ) .

Therefore, the square h − norm of σ′ is given by

(14) h ( σ , σ ) = A exp 2 0 s u t ( γ ( r ) ) d r exp 2 u ( γ ( s ) ) = A f 2 ( γ ( s ) ) exp 2 0 s u t ( γ ( r ) ) d r ,

for some A R determined by the initial values γ(0) and γ′(0). Hence, given T > 0 such that [0, b[ ⊂ [0, T], the hypothesis on f imply h(σ′, σ′) is bounded. Calling Lemma 4.4, we get γ is extensible. □

Remark 4.6.

In this finite-dimension case, it is no necessary to endow the tangent fiber bundle with a Riemannian metric. Indeed, under the assumptions of Lemma 4.4, it is not difficult to see that each integral curve of the vector field G can be included in a compact subset of TM.

Remark 4.7.

Note that this result extends the one obtained for Generalized Galilean Robertson–Walker (GGRW) spacetimes [[20], Th. 5], where the twisting function f > 0 depends only on t, and therefore the hypotheses of the previous theorem are automatically satisfied.

In the case of spacelike geodesics, we obtain an analogous result.

Theorem 4.8.

Let (F, h) be a complete Riemannian manifold and let f C ( R × F ) be a positive function. If inf  f > 0 along finite times, then the spacelike geodesics in the GT spacetime M = R × F , Ω = d π R , g = f 2 π F * h , are complete.

Proof.

Given a spacelike geodesic γ(s) = (c, σ(s)), we can reason analogously by defining a suitable vector field on TF, whose integral curves are determined by equation (10). Now,

d d s h ( σ , σ ) = 2 d h u ( σ ) h ( σ , σ ) .

and d d s ( u γ ) = d h u ( σ ) . Therefore,

(15) h ( σ , σ ) = A f 2 ( γ ( s ) ) , A R .

Using now inf  f > 0 on [0, T] ⊃ [0, b[, we have h(σ′, σ′) is bounded. An easy adaptation of Lemma 4.4 imply γ is extensible. □

Note that, if the fiber (F, h) is complete, Theorem 4.8 assures the geodesic completeness of any Riemannian manifold (F, g t ), for g t = f 2 ( t , ) π F * h , under a natural assumption on f.

We can ensure the geodesic completeness of a GT spacetime by unifying the results on the completeness of free falling observers and spacelike geodesics.

Theorem 4.9.

Let (F, h) be a complete Riemannian manifold and let f C ( R × F ) be a positive function. If inf  f > 0 and inf(f t /f) > − along finite times, then the GT spacetime M = R × F , Ω = d π R , g = f 2 π F * h , is geodesically complete.

If F is compact, the hypothesis for f C ( R × F ) in Theorem 4.9 are fulfilled trivially.

Corolary 4.10.

Let (F, h) be a compact Riemannian manifold and f C ( R × F ) a positive function. Then, the GT spacetime M = R × F , Ω = d π R , g = f 2 π F * h , is geodesically complete.

The following example prove that the hypotheses in Theorem 4.9 are optimal.

Example 4.11.

Consider the Riemannian manifold ( F , h ) = ( R , ) , to be the one-dimensional Euclidean space and f C ( R 2 ) a smooth positive function. They define the GT spacetime

( R × R , Ω = d t , g = f 2 π 2 * ( ) f 2 d x 2 , ) ,

where t   : = π1 and x   : = π2. Then, free falling observers γ(s) = (s, σ(s)) satisfy the following second-order nonlinear differential equation (see (9)),

0 = d π 2 D γ d s = σ + 2 f t f ( s , σ ( s ) ) σ + f x f ( s , σ ( s ) ) ( σ ) 2 .

From (14) (which can be integrated explicitly in this case), we know

(16) σ ( s ) = C f ( s , σ ( s ) ) exp 0 s u t ( r , σ ( r ) ) d r , C R ,

for a constant C = σ ̄ { 0 } × R = f ( 0 , σ ( 0 ) ) σ ( 0 ) (see equation (18)).

If we consider an incomplete geodesic γ = ( , σ ( ) ) : [ 0 , b [ R × R , for b > 0, then lim s b | σ ( s ) | = . For example, if inf  f > 0 along finite times, then it is necessary lim s b u t ( s , σ ( s ) ) = . Thus,

lim s b σ ( s ) = ± and lim x ± u t ( b , x ) = .

On the other hand, if inf  u t > − along finite times, then lim s b f ( s , σ ( s ) ) = 0 . Hence,

lim s b σ ( s ) = ± and lim x ± f ( b , x ) = 0 .

Note that the conditions of boundedness from below on f and f t /f arise naturally when there are incomplete geodesics.

In the case f : F = R R is time-independent, free falling observers γ(s) = (s, σ(s)) can be integrated as

(17) σ ( s ) = G 1 ( C s + D ) , C , D R ,

where G is a primitive function of f. Note that G is a diffeomorphism on R onto its image since G′ = f > 0.

For instance, if f ( x ) = ( 1 + x 2 ) 1 , free falling observers are given by γ(s) = (s, σ(s)) = (s, tan(Cs + D)), where C , D R are constants fixed by σ(0) and σ′(0). These are incomplete observers.

5 Physical interpretation of geodesics

Several interesting physical observations can be obtained from the expressions in the previous section.

Remark 5.1.

In the conditions of previous theorems, from equation (14) a free falling observer γ satisfies

(18) P t γ ( s ) F s = P t γ ( 0 ) F 0 e 0 s u t ( γ ) ,

where F s is the norm induced by g on the spacelike slice F s = { s } × F , for all s R . For a spacelike geodesic γ = (c, σ(⋅)), equation (15) implies

(19) P t γ ( s ) F c = P t γ ( 0 ) F c .

This last property is obvious from the fact that σ = P t γ is a geodesic in the Riemannian manifold ( F c , g c ) .

Now, equation (18) allows us to heuristically deduce the behaviour of the complete timelike geodesics in a GT spacetime in relation to certain properties of the twisting function. Consider two simultaneous events (s0, p), (s0, q) ∈ M. Their spacelike distance (in the absolute space) is given by

(20) dist F s 0 ( ( s 0 , p ) , ( s 0 , q ) ) = inf ξ C F ( [ 0,1 ] , p , q ) 0 1 f s 0 , ξ ( r ) ξ ( r ) h d r ,

where C F ([0, 1], p, q) is the set of admissible curves in F from p to q (see, for instance, [41]). Note that the function dist F s 0 ( ( s 0 , p ) , ( s 0 , q ) ) determine the distance between the commovil observer θ p (s0) = (s0, p) and θ q (s0) = (s0, q) in the absolute space F s 0 . Moreover, if β(r) = (s0, ξ(r)), r ∈ [0, 1], is a minimizing geodesic in F s 0 from p to q, then we have

(21) dist F s 0 ( ( s 0 , p ) , ( s 0 , q ) ) = 0 1 β d r = β ,

where ‖ ⋅‖ is the norm induced by g in F s 0 .

Obviously, if f t > 0 (resp. f t < 0) the absolute space is expanding (resp. contracting) over time. If for instance f t > 0, then the spacelike component of the velocity γ′(s) of a free falling observer (no commovil) γ, relative to the corresponding instant commovil observer, decreases in norm. Taking this into account, if we assume that the twisting function f t tends to + for t ↗ + , each free falling observer γ evolves asymptotically to commovil observer directions (see Figure 1(a)). If otherwise f t < 0 and we assume f t tends to − for t ↗ + and inf  f > 0, then every free falling observer γ evolves asymptotically to some spacelike direction (see Figure 1(b)).

Figure 1: 
Evolution of free falling observers in a Galilean Twisted spacetime: (a) f
t
(s0, p) > 0: Every free falling observer γ evolves asymptotically to a commovil observer. (b) f
t
(s0, p) < 0: Each free falling observer γ evolves to some spacelike direction.
Figure 1:

Evolution of free falling observers in a Galilean Twisted spacetime: (a) f t (s0, p) > 0: Every free falling observer γ evolves asymptotically to a commovil observer. (b) f t (s0, p) < 0: Each free falling observer γ evolves to some spacelike direction.

Now, it is natural to ask how free falling observers measure the speed of each other. If two points in the fiber p, qF are closely enough, free falling observers passing through one of them can measure the velocity of the others. If the leaves of the foliation induced by Ω are affine Euclidean spaces, the classical notion of relative velocity of an observer with respect to another is determined by the existence of a position vector between them. As a generalization, we can give the following.

Definition 5.2.

Let γ, σ be two observers in a (symmetric) Galilean spacetime and F s 0 the spacelike leaf in the absolute time t = s 0 R . If there exists a spacelike geodesic β : [ 0,1 ] F s 0 from γ(s0) to σ(s0) in F s 0 , the relative spacelike velocity of γ with respect to σ at t = s0, V γ ( s 0 ) , σ ( s 0 ) , is defined as the spacelike projection (with respect to σ′) of the parallel transport along β of γ′(s0), i.e.,

V γ ( s 0 ) , σ ( s 0 ) : = P σ P γ ( s 0 ) , σ ( s 0 ) β γ ( s 0 ) ,

where P β denotes the parallel transport along β.

Recall that in a Riemannian manifold, the existence and uniqueness of geodesics between two given points is locally ensured. Globally, the existence result holds in the complete case. Moreover, the non-existence of conjugate points globally guarantees uniqueness.

Let us deduce some interesting geometric properties of this definition in a GT spacetime. The fixed reference observer will be taken as a commovil observer. Indeed, given a free falling observer γ in a GT spacetime, denote γ ( s 0 ) = θ p ( s 0 ) + v , for v ( A n d t ) ( s 0 , p ) . The parallel transport P ( s 0 , p ) , ( s 0 , q ) β ( γ ( s 0 ) ) = V ( 1 ) is defined by a parallel vector field V along β with V(0) = γ′(s0). This can be split into

V ( r ) = a ( r ) t | ( s 0 , β ( r ) ) + U ( r ) ,

where a is a smooth function on [0,1] and U is a spacelike vector field along β. The initial value condition V(0) = γ′(s0) turns out

a ( 0 ) = 1 , U ( 0 ) = v .

From the fact that V is parallel and β can be lifted from a curve in F, it is easy to obtain that a′ = 0 and D U d r = u t β . Therefore, a ≡ 1 and U satisfy the initial value problem

D U d r = u t β , U ( 0 ) = v .

The modulus of the relative spacelike velocity of γ′(s0) measured by θ q (s0), V γ ( s 0 ) , θ q ( s 0 ) = U ( 1 ) = g ( U ( 1 ) , U ( 1 ) ) , can be explicitly obtained in this case. It is enough to compute

d d r U 2 = 2 u t g ( U , β ) , d d r g ( U , β ) = u t β 2 ,

using β is a geodesic. Obviously, ‖β′‖ is a constant. Taking into account the initial values, we can integrate both two differential equations, obtaining

(22) V γ ( s 0 ) , θ q ( s 0 ) = 0 1 u t ( s 0 , β ( r ) ) d r w v F s 0 ,

where v = P t γ ( 0 ) and the vector w: = β′(0). This expression, takes into account the geometric structure of the slice F s 0 , as well as, the distance between the points (s0, p) and (s0, q) (see equations (20) and (21)). In particular, if γ is a commovil observer, the velocity of θ p (s0) measured by θ q (s0) is

(23) V θ p ( s 0 ) , θ q ( s 0 ) = 0 1 u t ( s 0 , β ( r ) ) d r β = 0 1 u t ( s 0 , β ( r ) ) d r dist F s 0 ( s 0 , p ) , ( s 0 , q ) ,

where |⋅| is the absolute value function. We see that it is proportional to the spacelike distance between the points (s0, p) and (s0, q).

Finally, if the GT spacetime can be reduced to a GGRW spacetime, f = f(t), then expression (23) turns out V θ p ( s 0 ) , θ q ( s 0 ) = | u t ( s 0 ) | dist F s 0 ( s 0 , p ) , ( s 0 , q ) .

6 Galilean torqued vector fields

Along this section, we will see that there exists a natural generalization of GT spacetimes in a rich geometrical structure characterized by the existence of a distinguished vector field on the Galilean spacetime. In fact, the main theorem will state that these Galilean spacetimes are locally GT spacetimes.

Proposition 6.1.

Let M = I × F , Ω = d t , g = f 2 π F * h , be a GT spacetime. Then, the vector field K = f∂t satisfies

(24) X K = f t X + α ( X ) K and α ( K ) = 0 , X Γ ( T M ) ,

where α = d h ̄ ( log f ) 1 ( M ) (see subsection 2.2).

Proof.

Let X ∈ Γ(TM) be a vector field. Consider { E i } i = 1 n a local reference frame in TF that lifts to a local reference frame { t } { E i ̄ } i = 1 n in TM. Then, there exist local smooth functions g0, g1, …, g n such that

X = g 0 t + i g i E i ̄ .

Then, if u: = log  f, we have

X K = X ( f ) t + f X t = X ( u ) K + f i g i u t E i ̄ = X ( u ) K + f u t ( X g 0 t ) = f t X + d h ̄ u ( X ) K .

As we announced, a generalization of GT spacetimes is defined by means of a vector field satisfying an analogous expression to equation (24).

Definition 6.2.

Let (M, Ω, g, ∇) a Galilean spacetime. A future-pointing timelike vector field K ∈ Γ(TM) satisfying

(25) X K = ρ X + α ( X ) K and α ( K ) = 0 , X Γ ( T M ) ,

for some ρC (M) and α ∈⋀1(M), is called a Galilean torqued vector field. The function ρ is called the torqued function and α the torqued form.

The torqued function and form are fully determined by the function Ω(K), as we prove in the following lemma.

Lemma 6.3.

Let (M, Ω, g, ∇) be a Galilean spacetime and suppose it admits a Galilean torqued vector field K ∈ Γ(TM) with torqued function ρC (M) and torqued form α ∈⋀1(M). Then,

(26) α ( V ) = 1 Ω ( K ) V Ω ( K ) , V Γ ( An Ω ) .

and

(27) ρ = 1 Ω ( K ) K Ω ( K ) .

Proof.

Consider the field of observers Z = 1 Ω ( K ) K . Then, for any V ∈ Γ(AnΩ),

(28) V Z = V 1 Ω ( K ) K + 1 Ω ( K ) V K = ρ Ω ( K ) V + α ( V ) Ω ( K ) + V 1 Ω ( K ) K .

Now, making use of the compatibility of the connection with Ω we have Ω(∇ V Z) = V(Ω(Z)) = 0. Then, from (28) we get (26).

On the other hand, from equation (25) we have ∇ K K = ρK. Consequently,

K ( Ω ( K ) ) = Ω ( K K ) = ρ Ω ( K ) .

Thus, it follows expression (27). □

Remark 6.4.

Let K be a Galilean torqued vector field in a Galilean spacetime (M, Ω, g, ∇) with torqued function ρC (M) and torqued form α ∈⋀1(M). Consider the associated field of observers Z : = 1 Ω ( K ) K . Note that we have obtained in the proof of the lemma the following identity,

(29) V Z = ρ Ω ( K ) V , V Γ ( An Ω ) .

Since V(Ω(Z)) = 0 for all V ∈ Γ(AnΩ), from equation (29) we can compute

L Z g ( V , W ) = Z g ( V , W ) g ( [ Z , V ] , W ) g ( V , [ Z , W ] ) = g ( V Z , W ) + g ( V , W Z ) = 2 ρ Ω ( K ) g ( V , W ) ,

for all V, W ∈ Γ(AnΩ) (see subsection 2.1). Therefore, Z preserves conformally the absolute metric.⋄

Finally, we obtain the main theorem of this section, which determines the local structure of Galilean spacetimes admitting a Galilean torqued vector field.

Theorem 6.5.

Let (M, Ω, g, ∇) be a Galilean spacetime. If it admits a Galilean torqued vector field K ∈ Γ(TM), then for each pM, there exists an open neighborhood of p, U , and a Galilean diffeomorphism Ψ : N U , where N is a GT spacetime.

Proof.

This is in fact a generalization of the corresponding proof of the local structure Theorem for ICL spacetimes [[20], Th. 12]. In order to make the paper self-contained we give the details.

Suppose there exists a torqued vector field K. Let ρC (M) and α ∈⋀1(M) be the associated torqued function and one-form, and Z : = 1 Ω ( K ) K be the associated field of observers. Given an event pM, take a neighborhood U p of p in the leaf F p of the foliation induced by Ω, and I R , 0 ∈ I, a suitable interval such that the local flow of Z,

Ψ : I × U p M ( s , q ) ϕ s ( q ) ,

is well-defined and one-to-one. It satisfies

d Ψ ( s , q ) ( 1,0 ) = Z ϕ s ( q ) , d Ψ ( s , q ) ( 0 , v ) = d ϕ s q ( v ) ,

for all (s, q) ∈ I × U p and vT p U p . Now, if we identify tπ I and t ≡ (1, 0), it follows by construction

Ω ( d Ψ ( t ) ) = Ω ( Z ) = 1 = d t ( t ) .

Moreover, the flow ϕ s of Z preserves Ω on spacelike vector fields, L Z Ω(V) = 0, ∀V ∈ Γ(AnΩ). Then,

Ω d Ψ ( s , q ) ( 0 , v ) = Ω ( d ϕ s ) q ( v ) = ϕ s * Ω q ( v ) = Ω q ( v ) = 0 .

Consequently, Ψ*Ω = dt, and each level set of t corresponds with certain open set U F a of some leaf of the foliation F induced by Ω.

From Remark 6.4, Z is spatially conformally Leibnizian:

(30) L Z g ( V , W ) = 2 ρ Ω ( K ) g ( V , W ) , V , W Γ ( An ( Ω ) ) .

Consider v, wT q U p . If we define

η q ( s ) : = Ψ * g ( s , q ) ( ( 0 , v ) , ( 0 , w ) ) = g ϕ s ( q ) d ϕ s q ( v ) , d ϕ s q ( w ) ,

it follows from equation (30),

η q ( s ) = 2 ρ Ω ( K ) ϕ s ( q ) η q ( s ) .

Then, taking

(31) f ( s , q ) : = exp 0 s ρ Ω ( K ) ϕ l ( q ) d l , ( s , q ) I × U p ,

we obtain η q (s) = f2(s, q)g q (v, w). Therefore, Ψ*g(s, q) = f2(s, q)g q .

Finally, from equation (29), and any arbitrary tensor field V ∈ Γ(AnΩ), we have

Rot ( Z ) ( V , W ) = g V Z , W g W Z , V = 0 ,

and

Z Z = ρ Ω ( K ) Z + Z 1 Ω ( K ) K = 0 .

Thus, Z is an irrotational field of observers. Since dΨ(t) = Z, from [[16], Cor. 28], the Galilean connection ∇ is Ψ − related to the corresponding GT connection ∇ GT in I × U p . Therefore, Ψ is a Galilean diffeomorphism on

I × U p , d π I , f 2 π U p * g , G T .

Note that the conclusion is similar when we consider ICL spacetimes [20]. In this case, the distinction lies in the fact that the torqued factor is not necessarily spatially uniform.

Remark 6.6.

In the context of Lorentzian geometry, Chen [42] proved an analogous result. Indeed, Definition 6.2 applies to the notion of Lorentzian torqued vector fields with respect to the Levi-Civita connection, and the existence of such a vector field guarantees a local splitting of a Lorentzian spacetime as a Twisted spacetime. The local diffeomorphism is equivalent to the one presented in the proof of Theorem 6.5, but in the relativistic case it must be shown that it is a local isometry rather than a Galilean local diffeomorphism.

7 Global splitting

This section is devoted to give natural geometric conditions to characterize a Galilean spacetime as a global GT spacetime. We will use strongly the geometric properties satisfied by a Galilean torqued vector field.

The following argument will include the construction of an auxiliary pseudo-Riemannian metric, in contrast to the proof in [20], where the bijection is explicitly constructed and verified. In the lemma, we define this auxiliary semi-Riemannian metric in terms of the Galilean structure of a Galilean spacetime (see [43]). The spacelike gradient defined in subsection 2.2 allows us to write the associated Levi-Civita connection in terms of the Galilean connection.

Lemma 7.1.

Let (M, Ω, g, ∇) be a Galilean spacetime and let Z ∈ Γ(TM) be an irrotational spatially conformally Leibnizian field of observers. Given a smooth function φC (M) and ϵ ∈ {−1, + 1}, define on M the semi-Riemannian metric

g ̄ ϵ , φ : = ϵ Ω Ω + e 2 φ g ( P Z , P Z ) .

Then, the associated Levi-Civita connection ̄ of g ̄ ϵ , φ is given by

(32) ̄ X Y = X Y + d φ ( X ) P Z Y + d φ ( Y ) P Z X Ω ( X ) Ω ( Y ) G Z g ̄ ϵ , φ ( P Z X , P Z Y ) grad φ ϵ Z ( φ ) + ρ ̃ g ̄ ϵ , φ P Z X , P Z Y Z ,

for all X, Y ∈ Γ(TM), where G Z is the gravitational field of Z, ρ ̃ is the conformal factor associated to Z and grad φ is the spacelike gradient of φ in the Galilean structure.

Proof.

Consider X, Y ∈ Γ(TM) and V ∈ Γ(AnΩ). Koszul formula [[28], Th. 11] gives

(33) 2 g ̄ ϵ , φ ̄ X Y , V = X g ̄ ϵ , φ ( Y , V ) + Y g ̄ ϵ , φ ( X , V ) V g ̄ ϵ , φ ( X , Y ) + g ̄ ϵ , φ ( [ X , Y ] , V ) g ̄ ϵ , φ ( [ X , V ] , Y ) g ̄ ϵ , φ ( [ Y , V ] , X ) = e 2 φ 2 X ( φ ) g P Z Y , V + X g P Z Y , V + 2 Y ( φ ) g P Z X , V + Y g P Z X , V 2 V ( φ ) g P Z X , P Z Y V g P Z X , P Z Y + g P Z X , P Z Y , V Ω ( X ) g P Z Y , Z , V + Ω ( Y ) g P Z X , Z , V g P Z X , V , P Z Y + Ω ( X ) g [ V , Z ] , P Z Y g P Z Y , V , P Z X + Ω ( Y ) g [ V , Z ] , P Z X

Considering ω(Z) = 0, we can derive from equation (2) the following expression,

(34) 2 g ̄ ϵ , φ X Y , V = 2 e 2 φ g P Z X Y , V = e 2 φ X g P Z ( Y ) , V + Y g P Z ( X ) , V V g P Z ( X ) , P Z ( Y ) + 2 Ω ( X ) Ω ( Y ) g G Z , V + Ω ( X ) g Z , P Z ( Y ) , V g [ Z , V ] , P Z ( Y ) Ω ( Y ) g Z , P Z ( X ) , V + g [ Z , V ] , P Z ( X ) + g P Z ( X ) , P Z ( Y ) , V g P Z ( Y ) , V , P Z ( X ) g P Z ( X ) , V , P Z ( Y ) .

Substituting (34) in (33), we obtain

g ̄ ϵ , φ ̄ X Y , V = g ̄ ϵ , φ X Y , V + d φ ( X ) g ̄ ϵ , φ ( Y , V ) + d φ ( Y ) g ̄ ϵ , φ ( X , V ) e 2 φ V ( φ ) g P Z X , P Z Y Ω ( X ) Ω ( Y ) g ̄ ϵ , φ G Z , V .

On the other hand, applying (2) again and the relation L Z g = 2 ρ ̃ g , we have

(35) 2 g ̄ ϵ , φ ̄ X Y , Z = X g ̄ ϵ , φ ( Y , Z ) + Y g ̄ ϵ , φ ( X , Z ) Z g ̄ ϵ , φ ( X , Y ) + g ̄ ϵ , φ ( [ X , Y ] , Z ) g ̄ ϵ , φ ( [ X , Z ] , Y ) g ̄ ϵ , φ ( [ Y , Z ] , X ) = ϵ X ( Ω ( Y ) ) + ϵ Y ( Ω ( X ) ) ϵ Z ( Ω ( X ) ) Ω ( Y ) ϵ Ω ( X ) Z ( Ω ( Y ) ) 2 e 2 φ Z ( φ ) g P Z X , P Z Y e 2 φ Z g P Z X , P Z Y + ϵ Ω ( [ X , Y ] ) ϵ Ω ( [ X , Z ] ) Ω ( Y ) e 2 φ g P Z [ X , Z ] , P Z Y ϵ Ω ( [ Y , Z ] ) Ω ( X ) e 2 φ g P Z [ Y , Z ] , P Z X = 2 ϵ X ( Ω ( Y ) ) 2 e 2 φ Z ( φ ) g P Z X , P Z Y 2 e 2 φ ρ ̃ g P Z X , P Z Y .

Using g ̄ ϵ , φ X Y , Z = ϵ Ω X Y = ϵ X ( Ω ( Y ) ) , equation (35) gives

g ̄ ϵ , φ ̄ X Y , Z = g ̄ ϵ , φ X Y , Z e 2 φ Z ( φ ) g P Z X , P Z Y e 2 φ ρ ̃ g P Z X , P Z Y .

We will describe how the induced semi-Riemannian structure on a Galilean spacetime can be split into a semi-Riemannian twisted product I × f F , for an interval I, a Riemannian manifold F and certain positive smooth function f C ( I × F ) . This requires a slight adaptation of [[44], Th. 1], in which the geodesic completeness of the leaves L is replaced in the following special case.

Consider a smooth r-dimensional manifold N and suppose N admits two smooth transverse foliations L and K , 1-dimensional the first one and (r − 1)-dimensional the second one. We can assume that L is defined by a smooth vector field X L .

Definition 7.2.

The foliation L admits a uniform global generator X L if there exists a transverse leaf K 0 K and an open real interval I, such that the maximal flow ϕ of X L is well-defined and onto on I × K 0 .

The proof of [[44], Th. 1] remains valid in the following case.

Lemma 7.3.

Let (M, g) be a simply connected semi-Riemannian manifold endowed with two orthogonal foliations L and K . Suppose L is one-dimensional and admits a uniform global generator X ∈ Γ(TM) with associated interval I R and transverse leaf K 0 K . If L is totally geodesic and K is totally umbilic, then (M, g) is isometric to a (semi-Riemannian) twisted product I × f K 0 .

The above lemma allows us to prove

Theorem 7.4.

Let (M, Ω, g, ∇) be a simply connected Galilean spacetime and let K ∈ Γ(TM) be a Galilean torqued vector field. Given a smooth function φC (M) and ϵ ∈ {−1, + 1}, if the field of observers Z:= K/Ω(K) is a uniform global generator with associated interval I R and transverse leaf F in the foliation induced by Ω, then the semi-Riemannian manifold

( M , g ̄ ϵ , φ ) , with g ̄ ϵ , φ = ϵ Ω Ω + e 2 φ g ( P Z , P Z ) ,

is isometric to the (semi-Riemannian) twisted product

( I × F , g ̄ ) , with g ̄ = ϵ π I * d t 2 + f 2 π F * g | F ,

where g | F is the restriction of g to F and f is a positive smooth function on I × F .

Moreover, Z is a spatially conformal vector field of g ̄ ϵ , φ .

Proof.

Note first that we are in the hypothesis of Lemma 7.1 with a conformal factor ρ ̃ = ρ / Ω ( K ) (cf. equation (29)).

The induced foliation by Z is parameterized by its integral curves ϕ p (t), for tI and p F . From equation (32) and using ∇ Z Z = 0, we obtain ̄ Z Z = 0 . Therefore, the integral curves ϕ p (t) are geodesics for both ∇ and ̄ , that is, the associated foliation is totally geodesic.

Moreover, for all vector fields V, W ∈ Γ(AnΩ), we have

2 g ̄ ϵ , φ ( ̄ V W , Z ) = 2 g ̄ ϵ , φ ( W , ̄ V Z ) = Z g ̄ ϵ , φ ( V , W ) g ̄ ϵ , φ ( V , [ W , Z ] ) + g ̄ ϵ , φ ( W , [ Z , V ] ) = e 2 φ 2 Z ( φ ) g ( V , W ) g ( V , W Z ) g ( V Z , W ) = 2 Z ( φ ) + ρ Ω ( K ) e 2 φ g ( V , W ) = 2 Z ( φ ) + ρ Ω ( K ) g ̄ ϵ , φ ( V , W ) .

Then, the second fundamental form of each spacelike leaf of the foliation in ( M , g ̄ ϵ , φ ) is given by I I ̄ ( V , W ) = K ( φ ) + ρ Ω ( K ) g ̄ ϵ , φ ( V , W ) Z ; hence, the leaves of the foliation induced by AnΩ are totally umbilic for the semi-Riemannian metric g ̄ ϵ , φ . From Lemma 7.3, ( M , g ̄ ϵ , φ ) is isometric to a (semi-Riemmanian) twisted product I × f F .

Furthermore, from equation (32) we have

̄ V Z = 1 Ω ( K ) ( ρ + K ( φ ) ) V ,

for all spacelike vector fields V. Then,

L Z g ̄ ϵ , φ ( V , W ) = g ̄ ϵ , φ ( ̄ V Z , W ) + g ̄ ϵ , φ ( V , ̄ W Z ) = 2 Ω ( K ) ( ρ + K ( φ ) ) g ̄ ϵ , φ ( V , W ) ,

for all spacelike vector fields V, W, that is, Z is spatially conformal for g ̄ ϵ , φ . □

Finally, we are in the position to obtain a converse for Proposition 6.1.

Theorem 7.5.

Let (M, Ω, g, ∇) be a simply connected Galilean spacetime. It admits a global decomposition as a GT spacetime if and only if there exists a Galilean torqued vector field K ∈ Γ(TM) such that the associated field of observers, Z:= K/Ω(K), is a uniform global generator relative to the foliation defined by AnΩ.

Proof.

From Proposition 6.1, the necessary condition is obvious. Now, if the flow of Z is well-defined and onto on I × F , for I R an open interval and F a leaf of the foliation induced by Ω, Theorem 7.4 states that there exists a semi-Riemannian isometry map

Ψ : M , g ̄ 1,0 I × F , d t 2 + f 2 g | F ,

where we take ϵ = −1 and φ = 0. Consider Ψ as a map between Galilean spacetimes, where ∇0 is the corresponding GT connection. It suffices to prove that Ψ is a Galilean diffeomorphism.

First, Ψ maps the foliation induced by Z to the canonical foliation induced by t in I × F , with t = π I : I × F I . Consequently, if ϕ q (s) is the flow of Z for q F , we have

Ψ ( ϕ q ( s ) ) = ( β ( s ) , q ) , s I ,

for certain q F and a smooth function βC (I). Since Ψ preserves the causal character of vector fields for the semi-Riemannian structures, d Ψ ( Z ) = d d s ( Ψ ϕ q ) must be timelike unitary. Thus, (β′)2 = 1.

On the one hand, if β′ = −1, we may redefine the isometry map as follows,

Ψ ̃ : M I × F , p Ψ ̃ ( p ) = ψ 1 ( p ) , ψ 2 ( p ) ,

for all pM, where ψ i are the component functions of Ψ, Ψ ( p ) = ψ 1 ( p ) , ψ 2 ( p ) with ψ1(p) ∈ I and ψ 2 ( p ) F . Hence, the orientation of the isometry Ψ is reversed and the previous argument gives d Ψ ̃ ( Z ) = β ̃ = 1 .

Now, if β′ = 1, we have

( Ψ ϕ q ) ( s ) = ( s + c , q ) , s I ,

with cI, and dΨ(Z) = t. The manifold M is simply connected and dΩ = 0, then there exists a function TC (M), determined except for a constant, such that Ω = dT. It turns out (t◦Ψ)(ϕ q (s)) = s + c and T ( ϕ q ( s ) ) = s + c ̃ , with c ̃ R . If necessary, redefine T via a translation to get t◦Ψ = T. As a consequence, Ψ*dt = dT. Using the fact that Ψ preserves the semi-Riemannian structures, from Theorem 7.4 we obtain Ψ * f 2 g | F = g .

Note that dΨ(Z) = t and taking into account

d Ψ ( P Z X ) = d Ψ ( X ) Ω ( X ) d Ψ ( Z ) = P t d Ψ ( X ) ,

it is straightforward that formula ’à la Koszul’ (2) gives

2 g | F ( P t d Ψ ( X ) 0 d Ψ ( Y ) , d Ψ ( V ) ) = 2 g P Z X Y , V ,

for all X, Y ∈ Γ(TM) and all V ∈ Γ(AnΩ). Therefore ∇ and ∇0 are Ψ − related. □

When the spacelike leaves are compact, we can assure the existence of a global splitting.

Theorem 7.6.

Let (M, Ω, g, ∇) be a simply connected Galilean spacetime. If there exists a Galilean torqued vector field K ∈ Γ(TM) and the leaves of the distribution induced by AnΩ are compact, then M is a GT spacetime.

Proof.

It is enough to show that the vector field Z = K Ω ( K ) is a uniform global generator. Indeed, let Φ : D M be the maximal local flow of Z.

For any p F , we have I p × U p D , with I p R an open interval and U p an open set in F . Since F is compact, F = i = 1 k U i , k N . Then, Φ is well-defined on some domain I × F , for I R maximal interval.

In the case I = R , the flow is complete and the proof is ended using Theorem 7.5. Suppose now I = ]a, b[, with b < + . Given q F and ϵ > 0, if Φ(⋅, q) is defined on ]a, b + ϵ[, then there exists δ > 0 such that ] δ , δ [ × F Φ ( b , q ) D . Therefore, the flow can be extended,

Φ ̄ ( t , q ) = Φ ( t , q )  if  t ( a , b δ ) , Φ t b + δ 2 , Φ ( b δ 2 , q )  if  t ( b δ , b + δ ) .

This is a contradiction.

Finally, we prove that Φ : I × F M is onto. Suppose again by contradiction, that there exists q M \ Φ ( I × F ) and consider the maximal interval J R where Φ : J × F q M is defined. If Φ ( I × F ) Φ ( J × F q ) is not empty, the flow could be extended towards I. Therefore, since I is maximal, Φ ( I × F ) is open and closed in a connected manifold M. Hence, Φ ( I × F ) = M , and the proof is ended using Theorem 7.5. □


Corresponding author: Daniel de la Fuente, Departamento de Matemáticas, Universidad de Oviedo, 33003, Gijón, Spain, E-mail: 

Acknowledgments

We sincerely thank the referees for their helpful comments, which have contributed to improving the final presentation of the manuscript.

  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The author states no conflict of interest.

  6. Research funding: The three authors are partially supported by Spanish MICINN project PID2021-126217NB-I00. In addition, this research was supported by an FPU grant from the Spanish MICIU to the third author.

  7. Data availability: Not applicable.

References

[1] E. Cartan, “Les variètés a conexion affine,” Ann. Ec. Norm. Sup., vol. 40, no. 1, pp. 1–25, 1923.Search in Google Scholar

[2] E. Cartan, “Les variètés a conexion affine (suite),” Ann. Ec. Norm. Sup., vol. 41, no. 1, pp. 325–412, 1924.10.24033/asens.753Search in Google Scholar

[3] D. B. Malament, Topics in the Formulations of General Relativity and Newtonian Gravitation Theory, Chicago, Chicago lectures in Physics, University of Chicago Press, University of Chicago Press, 2012.Search in Google Scholar

[4] A. Trautman, Comparison of Newtonian and Relativistic Theories of space-time, Bloomington, Indiana, Perspectives in Geometry, Indiana Univ. Press, 1966.Search in Google Scholar

[5] F. Muler-Hoissen, “The cosmological principle and a generalization of Newton’s theory of gravitation,” Gen. Relativ. Gravit., vol. 15, no. 1, pp. 1051–1066, 1983. https://doi.org/10.1007/bf00760059.Search in Google Scholar

[6] C. Duval and H. P. Künzle, “Minimal gravitational coupling in the Newtonian theory and the covariant Schrödinger equation,” Gen. Relat. Gravit., vol. 16, no. 1, pp. 333–347, 1984. https://doi.org/10.1007/bf00762191.Search in Google Scholar

[7] H. Leihkauf, “On Newton-Cartan-theory,” Ann. Physik, vol. 46, no. 4, pp. 312–314, 1989. https://doi.org/10.1002/andp.19895010411.Search in Google Scholar

[8] W. A. Rodrigues, Q. A. de Souza, and Y. Bozhkov, “The mathematical structure of Newtonian spacetime: classical dynamics and gravitation,” Found. Phys., vol. 25, no. 1, pp. 871–924, 1995. https://doi.org/10.1007/bf02080568.Search in Google Scholar

[9] C. Rüede and N. Straumann, “On Newton-Cartan cosmology,” Helv. Phys. Acta, vol. 70, no. 1, pp. 318–335, 1997.Search in Google Scholar

[10] U. Brauer, A. Rendall, and O. Reula, “The cosmic no hair theorem and the nonlinear stability of homogeneous Newtonian cosmological models,” Classical Quant. Grav., vol. 11, no. 1, p. 2283, 1994. https://doi.org/10.1088/0264-9381/11/9/010.Search in Google Scholar

[11] M. Geracie, K. Prabhu, and M. M. Roberts, “Fields and fluids on curved non-relativistic spacetimes,” J. High Energy Phys., vol. 8, no. 42, pp. 1–39, 2015. https://doi.org/10.1007/jhep08(2015)042.Search in Google Scholar

[12] R. Penrose, “On gravity’s role in quantum state reduction,” Gen. Relat. Gravit., vol. 8, no. 5, pp. 581–600, 1996. https://doi.org/10.1007/bf02105068.Search in Google Scholar

[13] F. L. Lin and S. Y. Wu, “Non-relativistic holography and singular black hole,” Phys. Lett. B, vol. 679, no. 1, pp. 65–72, 2009. https://doi.org/10.1016/j.physletb.2009.07.002.Search in Google Scholar

[14] T. Brauner, S. Endlich, A. Monin, and R. Penco, “General coordinate invariance in quantum many-body systems,” Phys. Rev. D, vol. 90, no. 1, pp. 1–14, 2014. https://doi.org/10.1103/physrevd.90.105016.Search in Google Scholar

[15] M. Geracie, D. T. Son, C. Wu, and S. F. Wu, “Spacetime symmetries of the quantum Hall effect,” Phys. Rev. D, vol. 91, no. 1, pp. 1–18, 2015. https://doi.org/10.1103/physrevd.91.045030.Search in Google Scholar

[16] A. N. Bernal and M. Sánchez, “Leibnizian, Galilean and Newtonian structures of space-time,” J. Math. Phys., vol. 32, no. 3, pp. 77–108, 2003.10.1063/1.1541120Search in Google Scholar

[17] R. Andringa, E. Bergshoeff, J. Rosseel, and E. Sezgin, “3D Newton-Cartan supergravity,” Classical Quant. Grav., vol. 30, no. 20, pp. 1–19, 2013. https://doi.org/10.1088/0264-9381/30/20/205005.Search in Google Scholar

[18] R. Andringa, E. Bergshoeff, J. Gomis, and M. de Roo, “Stringy’ Newton–Cartan gravity,” Classical Quant. Grav., vol. 29, no. 23, pp. 1–44, 2012. https://doi.org/10.1088/0264-9381/29/23/235020.Search in Google Scholar

[19] E. A. Bergshoeff, J. Gomis, J. Rosseel, C. Simsek, and Z. Yan, “String theory and string Newton-Cartan geometry,” J. Phys. A, vol. 53, no. 1, pp. 1–38, 2020. https://doi.org/10.1088/1751-8121/ab56e9.Search in Google Scholar

[20] D. de la Fuente and R. M. Rubio, “Galilean Generalized Robertson-Walker spacetimes: a new family of Galilean geometrical models,” J. Math. Phys., vol. 59, no. 2, p. 022903, 2018, https://doi.org/10.1063/1.4997115.Search in Google Scholar

[21] L. F. Costa and J. Natário, “The Coriolis field,” Am. J. Phys., vol. 84, no. 5, pp. 1–9, 2016.10.1119/1.4938056Search in Google Scholar

[22] D. de la Fuente, J. A. S. Pelegrín, and R. M. Rubio, “On the geometry of stationary Galilean spacetimes,” Gen. Relat. Gravit., vol. 53, no. 8, pp. 1–15, 2021.10.1007/s10714-020-02772-1Search in Google Scholar

[23] X. Bekaert and K. Morand, “Embedding nonrelativistic physics inside a gravitational wave,” Phys. Rev. D, vol. 88, no. 6, pp. 1–61, 2013. https://doi.org/10.1103/physrevd.88.063008.Search in Google Scholar

[24] X. Bekaert and K. Morand, “Connections and dynamical trajectories in generalised Newton-Cartan gravity. II. An ambient perspective,” J. Math. Phys., vol. 59, no. 7, p. 072503, 2018, https://doi.org/10.1063/1.5030328.Search in Google Scholar

[25] E. Minguzzi, “Classical aspects of lightlike dimensional reduction,” Class. Quant. Grav., vol. 23, no. 23, pp. 7085–7110, 2006. https://doi.org/10.1088/0264-9381/23/23/029.Search in Google Scholar

[26] E. Minguzzi, “Eisenhart’s theorem and the causal simplicity of Eisenhart’s spacetime,” Class. Quant. Grav., vol. 24, no. 11, pp. 2781–2807, 2007. https://doi.org/10.1088/0264-9381/24/11/002.Search in Google Scholar

[27] E. Minguzzi, “A connection between Lorentzian distance and mechanical least action. Talk at ’Noncommutative structures and nonrelativistic (super)symmetries’ (Tours),” 2010. Slides available online at: http://www.lmpt.univ-tours.fr/bekaert/Conference_2010/Minguzzi.pdf.Search in Google Scholar

[28] B. O’Neill, Semi-Riemannian Geometry with Applications to Relativity, ISSN. Los Angeles, Academic Press, 1983.Search in Google Scholar

[29] B.-Y. Chen, “Totally umbilical submanifolds,” Soochow J. Math., vol. 5, pp. 9–37, 1979.Search in Google Scholar

[30] B.-Y. Chen, Differential Geometry of Warped Product Manifolds and Submanifolds, Singapore, World Scientific, 2017.10.1142/10419Search in Google Scholar

[31] C. A. Mantica and L. G. Molinari, “Generalized Robertson–Walker spacetimes – a survey,” Int. J. Geomet. Methods Mod. Phys., vol. 14, no. 03, p. 1730001, 2017, https://doi.org/10.1142/s021988781730001x.Search in Google Scholar

[32] A. Soria, “Spacelike hypersurfaces in twisted product spacetimes with complete fiber and Calabi-Bernstein-type problems,” J. Geom., vol. 114, no. 16, 2023, https://doi.org/10.1007/s00022-023-00677-3.Search in Google Scholar

[33] C. A. Mantica and L. G. Molinari, “Twisted Lorentzian manifolds: A characterization with torse-forming time-like unit vectors,” Gen. Relativ. Gravit., vol. 49, no. 51, 2017, https://doi.org/10.1007/s10714-017-2211-1.Search in Google Scholar

[34] S. Güler, U. C. De, and B. Ünal, “Geometry of twisted products and applications on static perfect fluid spacetimes,” Int. Electron. J. Geom., vol. 16, no. 2, pp. 598–607, 2023, https://doi.org/10.36890/iejg.1286525.Search in Google Scholar

[35] A. N. Bernal, M. López, and M. Sánchez, “Fundamental units of length and time,” Found. Phys., vol. 32, no. 1, pp. 77–108, 2002.10.1023/A:1013800914617Search in Google Scholar

[36] F. W. Warner, Foundations of Differentiable Manifolds and Lie Groups, New York, Graduate Texts in Mathematics. Springer-Verlag, 1983.10.1007/978-1-4757-1799-0Search in Google Scholar

[37] M. Spivak, A Comprehensive Introduction to Differential Geometry, vol. 1, Houston, Texas, Publish or Perish, 1979.Search in Google Scholar

[38] Theodore Frankel, The Geometry of Physics: An Introduction, 3 ed. Cambridge, Cambridge University Press, 2011.Search in Google Scholar

[39] M. Spivak, A Comprehensive Introduction to Differential Geometry, vol. 2, Houston, Texas, Publish or Perish, 1979.Search in Google Scholar

[40] L. J. Alías, A. Romero, and M. Sánchez, “Uniqueness of complete spacelike hypersurfaces of constant mean curvature in generalized Robertson-Walker spacetimes,” Gen. Relat. Gravit., vol. 27, no. 1, pp. 71–84, 1995. https://doi.org/10.1007/bf02105675.Search in Google Scholar

[41] M. P. do Carmo, Riemannian Geometry. Mathematics: Theory & Applications, 1 ed. Boston, MA, Birkhäuser, 1992.10.1007/978-1-4757-2201-7Search in Google Scholar

[42] B.-Y. Chen, “Rectifying submanifolds of Riemannian manifolds and torqued vector fields,” Kragujevac J. Math., vol. 41, no. 1, pp. 93–103, 2017, https://doi.org/10.5937/kgjmath1701093c.Search in Google Scholar

[43] D. de la Fuente, J. A. S. Pelegrín, and R. M. Rubio, “Semi-Riemannian structures for symmetric Galilean spacetimes,” Rev. Real Acad. Cienc. Exactas Fis. Nat. Ser. A-Mat., vol. 118, no. 164, 2024, https://doi.org/10.1007/s13398-024-01664-2.Search in Google Scholar

[44] R. Ponge and H. Reckziegel, “Twisted products in pseudo-Riemannian geometry,” Geom. Dedicata., vol. 48, no. 1, pp. 15–25, 1993. https://doi.org/10.1007/bf01265674.Search in Google Scholar

Received: 2024-10-21
Accepted: 2025-07-30
Published Online: 2025-10-01

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 22.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2023-0197/html?lang=en
Scroll to top button