1. Introduction
General Relativity (GR) is the current best theory for the gravitational interaction [
1,
2]. It passes all observational tests at the solar system scales (e.g., [
3]), and provides the basis for technological applications, such as GPS [
4]. GR predicts gravitational waves of coalescing massive stars that have been directly detected by the LVK Collaboration [
5]. We should not forget about black holes, which were directly imaged twice by the EHT Collaboration [
6,
7]. The standard cosmological model stems from GR; it predicts the expansion and large-scale structure of the universe [
8], along with the nucleosynthesis and relative abundances of light atomic nuclei (see [
9] and references therein) and the existence of the cosmic microwave background radiation [
10,
11], all of which were measured with great precision [
12,
13,
14,
15,
16,
17].
For all its astonishing successes, GR still has a few shortcomings. The description of the present-day universe hinges on the existence of the dark sector, composed of dark matter and dark energy. The latter is responsible for the observed accelerated expansion of the recent cosmos [
18]. On the other hand, dark matter is necessary to explain galaxy rotation curves [
19], the cohesion of galaxy clusters [
20], and even the large-scale structure of the universe [
21], seeded by CMB anisotropies. On top of that, GR has to deal with singularities where its equations break down [
22,
23]. This happens (again) in cosmology when we consider the very early universe, where the Big Bang took place and the time-like geodesics terminate (according to the Penrose–Hawking theorem). The ultraviolet regime problem of GR also includes the interior singularities at the centers of black holes. In any case, a quantum version of GR is missing, which makes it difficult to hope for descriptions of spacetime points with extremely high curvature [
24].
The limitations of GR motivate extensions (or modifications) of the theory [
25,
26,
27]. One of the possible modifications is to explore the class of scalar-tensor theories of gravity [
26,
28], where a scalar degree of freedom is added to the already present metric field of GR. Two great examples of such theories may be found in the Kaluza–Klein model [
29,
30] and in the Jordan–Brans–Dicke theory [
31].
Early in the 1950s, Lyra introduced a scalar field to Riemann spacetime based solely on geometrical motivations [
32]. In effect, Lyra added a scale function
to the very definition of the reference frame. For this reason, the Lyra manifold generalizes the Riemannian manifold of GR in many ways. The transformations between spacetime points will involve both diffeomorphisms and scale transformations. This extends the notion of connection, curvature, and torsion, allowing the construction of a fresh gravitational theory with novel consequences.
Early works on Lyra gravity are those by Sen [
33,
34] and collaborators [
35,
36]; see also Ref. [
37]. This first approach rendered a theory where a gauge potential
is introduced, resembling Weyl’s attempt to unify electromagnetism and gravity [
38]. However, the resulting theory suffered from formal and physical inconsistencies. First, field equations are not covariant under Lyra’s transformations and do not come from a well-defined variational problem. Second, Weyl’s theory had to be abandoned due to the non-integrability of vector lengths [
39], a feature that would produce modification of astrophysical spectral lines, an effect contrary to observations [
40,
41]. Sen tried to solve the problem using the gauge fixing condition
. However, this would render
null, and the resulting equations would become just Einstein field equations of GR [
42]. After recognizing the second problem of the original model, Sen and Dunn [
35] proposed a second model with field equations depending on the ordinary Einstein tensor
and the scalar field
as an additional degree of freedom, i.e., not only a gauge parameter. The drawback with the Sen–Dunn model goes beyond the lack of novelty (in effect, it was proven to be equal to Brans–Dicke theory for a specific value of the Brans–Dicke coupling
). The first problem was still an important issue.
In the face of these challenges, and inspired by Sen’s ideas, Ref. [
42] proposes a new scalar-tensor theory on Lyra’s manifold. The remedy built upon the theory from the fundamental symmetries of the Lyra manifold: diffeomorphisms and scale transformations. The field equations do not present the problems of the previous models. Instead, they exhibit the suitable properties for a consistent gravity theory involving the metric tensor
and the scale
as the fundamental dynamical fields, including the correct Newtonian limit in the weak field regime and low-velocity motion of test particles.
In the same paper, the simplest solution for the LyST field equations is built. An ansatz was proposed for the line element
, endowed with static and spherically symmetric character. Substitution in the field equations leads to differential equations involving the scale
and the function
, appearing as coefficients of the temporal and radial parts of
. The solution to these differential equations gives the Lyra–Schwarzschild line element, a generalization of the Schwarzschild metric. A subsequent study [
43] devoted attention to particle and light motion associated with the Lyra–Schwarzschild line element. To do so, the geodesic equations were built, and the study of possible trajectories was carried out.
The results described in the previous two paragraphs are reviewed by more fundamental methods in the present work.
Section 2 brings an overview of the defining features of the Lyra manifold. It shows how the basis vectors depend on the scale
; this propagates to the tensor transformation rules, the line element, and the volume element. The scale is also present in the expression of auto-parallel curves and geodesics (which are not automatically the same in the context of Lyra geometry). Connection, curvature, and torsion are also defined and compared to their Riemannian counterparts. The novelty of the procedure employed here rests on the use of the Cartan formalism [
44].
Section 3 builds Lyra scalar-tensor gravity (LyST) from a properly defined action
S towards the field equations. The Newtonian limit shows that these field equations recover the Poisson equation for a well-defined potential.
In
Section 4, we build the Lyra–Schwarzschild spacetime by invoking spherical symmetry. However, as can be seen from Equation (
50), the
coefficient and the
coefficient in
are not presumed to be inversely equal, in contrast with Ref. [
43]. As we shall see, solving the geodesic equations demonstrates not only the once assumed line element, but also demonstrates the spherically symmetric solution to be static; this is equivalent to Birkhoff’s theorem [
45]. Another difference between the approach of Ref. [
43] and the one in the present paper concerns the technique employed to solve the field equations. Herein, we used tetrads fields
instead of the standard technique. Cartan’s formalism is used to write the curvature in terms of the spin connection
. Imposing a metric-compatible torsion-free spacetime,
values are uniquely related to
by Cartan’s structure equation, which is then solved to provide the curvature 2-form and the curvature scalar.
Massive and massless particle motion is discussed in
Section 5. Here, as another distinct feature compared to previous approaches, we apply the Hamilton–Jacobi formalism in the spirit of Hagihara’s work [
46]. The effective potential
is defined for both massive and massless test particles. The different regions in the plots of
correspond to different types of motion. We discuss the motion of massive particles, including the determination of the innermost stable circular orbit (ISCO) and the periastron shift. In the case of massless particle motion, the last photon’s stable orbit is discussed. Gravitational redshift and causal structure are also discussed.
Section 6 brings our final remarks, physical interpretations, and perspectives.
4. Spherically Symmetric Solution
The task of finding spherically symmetric solutions of the field equations is usually carried out with Killing vector analysis. In rectangular coordinates, a spatial rotation is generated by the matrices
, where
denotes the
i-th Cartesian spatial axis. The corresponding Killing vectors have the coefficients
. In spherical coordinates
, those vectors take the form
Rotational invariance constrains the scalar field
and the metric tensor
. Applying the Lie derivative concerning the Killing vectors to the scalar field
, we see that it must be independent of the angular coordinates, i.e.,
. In the case of the metric coefficients, we have
that results in
This expression can be simplified by suitable coordinate transformations. To this end, we introduce the quantities
Next, we define a new radial coordinate through the relation
, where
is an arbitrary function. With this change, the radial coordinate
r can be expressed in terms of
through the relation
thus preserving the general structure of the line element. We then enforce a rescaling by extracting an overall factor
, which yields
A further redefinition of the temporal coordinate
allows the coefficient of the mixed term
to be eliminated, bringing the metric into a diagonal form. Locally, this is achieved through
where
is an appropriate integrating factor. The resulting expression, after suppressing the prime symbols for notational simplicity, can be written as
where
,
A, and
B are functions different from the former ones. In this case, the scalar field is independent of the transformed time coordinate, and
is diagonal in the transformed time and radial coordinates. Equation (
50) is the most general form of a spherically symmetric line element.
Seeking to employ Cartan’s formalism, it is useful to define the orthonormal 1-form tetrad basis:
where the undetermined functions have been written in exponential form:
By applying the first Cartan structure Equation (
20) to the Lorentz frame (
51), one can find the non-zero components of the connection 1-form:
The second Cartan structure Equation (
19) yields the following curvature coefficients:
The components of the Ricci tensor are calculated as
Now, Ricci’s scalar is given by
Therefore, the non-zero components of the Einstein tensor in LyST theory are given by
4.1. Vacuum Solution
Here, the energy-momentum tensor vanishes. If
is time independent, then
, and from Equation (53b), we have
The solution
is incompatible with Equation (53c); therefore, we must have
. In this case, Equation (53c) implies
, then
. The only time-dependent coefficient in (
50) is
A, and we can absorb
into the time coordinate by redefining
Since
is assumed to be smooth within our framework, it is also continuous and therefore integrable. This guarantees that the transformation above is well defined. This redefinition eliminates the explicit time dependence of the metric coefficients while preserving the general structure of the line element and introducing no mixed terms. Moreover, if we choose not to absorb
, we will find that the field equations impose no restriction on it, so we may fix it arbitrarily, for instance by setting
. Therefore, Birkhoff’s theorem
3 holds when the scale function is time-independent.
Therefore, the nonzero components of the Einstein tensor can be written as follows:
The solution of the corresponding field equations is given by
where
,
a is an integration constant, and
is still arbitrary.
By defining a new coordinate
, one finds that the line element has the same functional form as the one in the Schwarzschild case, with
being the Schwarzschild radius:
This is an important result, as it shows that the Schwarzschild line element is not just a particular solution. In fact, it is the only solution, equivalent to any other solution with a non-trivial scale function, but expressed in a specific Lyra reference system. This solution can be expressed in another reference system with a non-trivial scale function, but keeping a Schwarzschild-type form with respect to the metric components
Here, the scale function is
and
where
is the Schwarzschild radius in the new Lyra reference system and
is a positive integration constant called the Lyra radius.
If the line element (
57) is interpreted as an exterior solution, the function
must remain positive, which holds only within the range
. This suggests the possibility of a spatially limited spacetime. However, cosmological observations indicate the existence of structures on extremely large scales, implying that spacetime is either infinite or bounded by a value of
far beyond the currently observed scales of the universe. The latter scenario is also mathematically plausible, provided that cosmological observations occur within the range
, in a limit where the coordinates
r and
are practically indistinguishable. Given its significant implications, attention will be payed to the expression (
57) of the line element, and it will be referred to as the Lyra–Schwarzschild solution.
4.2. Correction to Newtonian Gravity
In a spherically symmetric spacetime, the
r equation of motion in the Newtonian limit, as given in Equation (
48), simplifies to
where
. The weak field perturbations
and
are found by Taylor expanding the functions (
58) and (
59) within the range
:
and
In (
62) and (
63), we made explicit the corrections to the Newtonian limit:
Thus, the
r equation can be written as follows:
where
is the geometric mass defined by
Equation (
66) reveals that the geometric mass, which acts as the source of the gravitational field, depends not only on
but also on
. Meanwhile, second-order effects in Equation (
65) include a constant repulsive acceleration and a term linear in
r analogous to an anti-de Sitter contribution.
The existence of a Newtonian limit allows the identification of the Newtonian mass, which is a real and positive quantity. Accordingly, solution (
59) can be rewritten by expressing the constant
in terms of
and
:
Substituting (
67) into (
59), one obtains
while
retains the form (
58). With this parametrization, the line element becomes
The limit
, which is equivalent to
, defines the Schwarzschild limit in GR. In this case,
in this limit. This confirms that the solution admits a Schwarzschild limit and, consequently, that the parameter
may be regarded as the generalization, within Lyra geometry, of the Schwarzschild radius.
4.3. The Cosmological Constant
The gravitational field experienced by test particles at spatial infinity can be obtained by taking the limit
in (
59). In this case, the dominant contribution is quadratic:
For comparison, the Schwarzschild–de Sitter solution at large distances reads
From this, it follows that the parameter
is directly related to both the source mass and the cosmological constant:
The sign of
determines whether
is positive or negative. The solution is asymptotically de Sitter (
) when
. However, in this case, the function
is negative for all positive values of
r. This also implies
and
, leading to a metric with a signature of the form
. On the other hand, if
or
, the spacetime at large distances from the gravitational source is of the AdS type. Therefore, the large distance behavior may be grouped into three distinct classes:
Class 1: , asymptotically AdS solutions with positive ;
Class 2: , asymptotically AdS solutions with negative ;
Class 3: , asymptotically de Sitter solutions.
In these,
exhibits two positive real roots, one at
and the other at
. The point
corresponds to a minimum of
. In Class 2 solutions,
increases monotonically with the radial coordinate, approaching an AdS regime as
. Class 3 solutions, by contrast, approach the de Sitter regime for large
r. Nevertheless, in this case, the function
remains negative for all positive values of
r. From this point onward, our analysis will focus on the Class 1 case, defined by
, since this case is manifestly the most physically relevant.
6. Final Remarks
Lyra scalar-tensor theory is formulated on a Lyra manifold with a torsion-free, metric-compatible connection. The fundamental fields of this theory are the metric tensor and the scale function . The presence of the scalar field modifies the canonical basis of the tensor bundles, leading to a transformation law for Lyra reference systems that differs from that of General Relativity. The metric-compatible and torsion-free conditions are imposed to make LyST theory as similar to General Relativity as possible. In this sense, LyST theory can be viewed as an extension of General Relativity on a Lyra manifold with a nontrivial scale function.
Field equations in LyST theory are derived from the standard variational principle applied to a Lyra-invariant action. This action generalizes the Einstein–Hilbert action via the minimal coupling principle, incorporating a Lyra-invariant volume element. Equation (
48) shows that LyST theory is compatible with Newtonian gravity in the linearized and static approximation. The resulting Newtonian gravitational potential is given by
, where
denotes the perturbation of the Lyra scale function. Thus, gravity in LyST theory is also governed by this perturbation, as it represents a quantity independent of the metric.
The most general spherically symmetric solution in LyST theory depends on a single parameter and a freely selectable scale function, as shown in Equation (
55). This freedom suggests the existence of an infinite family of solutions satisfying the field equations. However, by defining a new radial coordinate
, one obtains the exact expression of the standard Schwarzschild solution, where the aforementioned parameter is the Schwarzschild radius
. This leads us to the analogue of Birkhoff’s theorem in the case of Lyra gravity: the Schwarzschild solution is the only physical solution, and the general expression is simply a result of a coordinate transformation involving the scale function. Requiring that the general expression preserve the structure of the Schwarzschild metric leads to a specific choice of the scale function:
, where
is a new parameter, known as the Lyra radius, with
. The resulting metric is referred to as the Lyra–Schwarzschild solution. It has a coordinate singularity at
and
, yet it is equivalent to the Schwarzschild metric under the above transformation.
In the limit , one recovers the standard Schwarzschild metric, with the radial coordinates becoming identical, . This suggests that the Lyra–Schwarzschild metric is merely an alternative representation of the Schwarzschild solution, adapted to the case of a spatially finite universe bounded by . Could this apparent finiteness be an actual feature of the spatial extent of our universe, or is it merely a consequence of adopting a particular coordinate system? According to the standard ΛCDM model, the universe is generally considered spatially flat, which implies an infinite spatial extension. Thus, the finiteness in the Lyra–Schwarzschild expression seems to be just a matter of the coordinate system. However, the converse is also plausible: the universe might be spatially finite, characterized by an extremely large Lyra radius well beyond the bounds of any cosmological scale currently accessible to observation. In this case, the Lyra–Schwarzschild description would be nearly indistinguishable from that of the Schwarzschild solution, since the relevant scales satisfy . This raises the possibility that we may be inadvertently confusing coordinate systems, interpreting physical results within an unintended frame.
In the Newtonian approximation of the spherically symmetric solution, the equation of motion for a massive particle, formulated in the Lyra–Schwarzschild description, is given by Equation (
65). The geometric mass
, expressed in both coordinate systems, is provided by Equation (
66). In the first order in
the equation of motion reduces to the standard expression in Newtonian gravity, with the geometric mass still influenced by the Lyra radius. Second-order effects introduce a constant repulsive acceleration and an additional term analogous to an anti-de Sitter contribution. It should be noted that particle motion remains affected by the Lyra radius even in the absence of matter distributions, since the anti-de Sitter term persists when
. The contribution of this term to the equation of motion increases as
r approaches
. Therefore, while uncertainties in measurements at solar-system-size scales can be used to set lower bounds to
, Equation (
65) could be employed to estimate its value by observing particle trajectories at large scales, where
.
The Hamilton–Jacobi method offers a formal and straightforward approach to derive the equations of motion for the spherically symmetric solution. More than merely serving as a consistency check with respect to the classical approach of integrating the equations of motion, this formalism provides a robust and effective framework for identifying the dynamical invariants that underlie the system’s integrability and its solutions. This is particularly valuable for more complex scenarios that may be addressed in future work, such as Lyra–Reissner–Nordstrom (charged black holes) and Lyra–Kerr–Newmann (rotating black hole) solutions. In the Lyra–Schwarzschild description, these equations are given by Equation (79). They are equivalent under the transformation
to those of the standard Schwarzschild case, as provided by Equation (80). An analysis of the effective potential in both coordinate systems suggests that
appears as a new critical point introduced by the scale function. However, this value lies outside the physical domain, since
is forbidden. As a result, the overall behavior of the effective potential, whether expressed in terms of
r or
, remains nearly identical, i.e., yielding the same types of motion in both cases. Consequently, the particle trajectories in the Lyra–Schwarzschild description can be directly obtained from the well-known trajectories in the standard Schwarzschild case via the aforementioned coordinate transformation. In fact, the solutions expressed in terms of the inverse radial coordinate differ only by the constant
. In the case of photons, the redshift is given by Equation (
120), which reduces to the standard Schwarzschild expression when both the source and the observer are located far from the event horizon at
(for a fixed
).
On the other hand, the causal structure presented by the Lyra–Schwarzschild description is very similar to that of the Schwarzschild case in the regime
. The only physical singularity remains at the coordinate origin
, as indicated by Equation (
103). The main difference in the causal structure is that the lightcones become significantly narrower in the regime
, eventually collapsing at
. These results encourage the study of other spherically symmetric solutions, such as an analog of the Reissner–Nordstrom metric in GR, as well as other cases of LyST gravity applied to more general solutions with different symmetries, gravitational waves, and even more general constructions like the Einstein–Cartan theory.