3.2. Dimensional Analysis
To satisfy the requirements of dimensional analysis, only those functions of physical quantities will be admissible that transform under a scale transformation of ratio
according to the following rule: if
,
,…,
have dimensions
,
,…,
, then any function
,
,…,
with dimensions
that can be formed from them must satisfy
This is the only imposed restriction. It prohibits the presence of dimensionful constants in the theory.
Aldersley demonstrates the following two theorems in his work [
23]:
Let
be a class
tensor concomitant of the metric and its derivatives up to any order,
such that its dimensions are
, and assuming that condition (2) is satisfied—which is, in fact, an axiom of the theory—then, for a four-dimensional manifold, it holds that
is the curvature scalar and
is the electromagnetic field tensor, such that
where
is the Lagrangian.
Let
be a scalar density of class
, concomitant of the metric and its derivatives up to the arbitrary order
such that its dimensions are
, and assuming that axiom (2) is satisfied. Then, it follows that the manifold is four-dimensional, since
For both proofs, Aldersley employs arguments similar to those described so far, together with Lovelock’s theorem [
24]: If
is a tensorial concomitant satisfying
, then, in a four-dimensional manifold,
If dimensional constants had been permitted in —for example, with —the absence of the differentiability properties of with respect to (and the same holds for any tensor, not just density ) would have rendered the entire analysis presented above impossible.
If, instead of choosing two universal constants with dimensions that relate lengths, times, and charges as per Aldersley (
and
), it is considered that three universal constants exist, for example
,
, and
with
,
, this would lead to the time scale
being proportional to itself, satisfying the relation,
where
is a dimensionless constant.
The restriction of excluding dimensionful constants from the theory—aside from the aforementioned two universal constants relating length, time, and mass—does not affect the derivation of the Lagrangian densities for the Einstein–Maxwell fields. In particular, no dimensionful constants are introduced in the Lagrangian, and the cosmological constant Λ is treated as a variable, which may become constant only in specific cases. The remaining parameters appearing in the Lagrangian (, ,…, ) are regarded as dimensionless constants.
3.3. Einstein–Maxwell Lagrangian Formulation
In the formulation of Lagrangian densities, it is natural to employ the classical theory of concomitants, which studies the general form of mathematical entities depending on the field variables and their derivatives, combined with dimensional analysis. This approach is justified because the equations of motion for the various fields involved in the Lagrangian formulation of a theory can be derived by applying a variational principle to the action constructed from the corresponding Lagrangian density, which is a function of the fields and their derivatives. The outcome of applying the variational principle is the Euler–Lagrange equations, and it has been demonstrated that these expressions are concomitant operators [
25].
Weyl’s theorem states that the most general Lagrangian density that can be constructed from the metric and its derivatives up to the second order and that is linear in the second derivatives is the Einstein Lagrangian. Rigorous studies have been conducted on the Lagrangians commonly employed in semiclassical theories [
26], as well as on more general possible Lagrangians [
27]. The resulting field equations derived from the variation of the Lagrangian density must be gauge-invariant under the appropriate symmetry group.
Hypothesis:
The precise formulation of this hypothesis is summarized as follows:
- (i)
The Lagrangian density to be constructed must be a function of the metric , the electromagnetic vector potential , and a scalar field . The latter may be either a conventional field or a constant.
- (ii)
No dimensional constants are allowed in the Lagrangian, as their presence generally leads to non-renormalizable theories.
- (iii)
We permit first derivatives of the fields and because we require their field equations to be of the second order. For the metric, both first and second derivatives are allowed. No upper bound is imposed on the order of derivatives in the Lagrangian to keep the theory as general as possible.
- (iv)
Units are chosen such that
and
denotes the action. Hence,
- (v)
Finally, we require gauge invariance for the field equations.
In the subsequent development, the powers of the arbitrary degree of the fields and their derivatives will be admitted. We shall allow only derivatives of and so that the resulting field equations for them are of the second order and the second derivatives of , since it is not possible to construct a concomitant other than a constant using only the metric and its first partial derivatives. In any case, the Hilbert action is degenerate, and consequently, the field equations for the Einstein–Maxwell system derived from it are of the second order, even though the Lagrangian density itself is of the same order.
The correct units of
and
are obtained by inspecting the units of their kinetic terms in the Lagrangian,
with
y
The issue of the dimensionful constant is related to the fact that two tensors, rather than one, are involved in the field action. The perturbative treatment of the field requires its decomposition into a flat background metric
plus a perturbation
, which must be multiplied by the constant
so that the sum carries the correct units. Thus, the decomposition is
The inverse
and the determinant
expand into an infinite power series in
,
where the position of the indices is irrelevant, since raising and lowering is performed with
. In the coordinate gauge for
, one has
with
. Thus, gauge invariance reduces the degree of divergence in the Einstein–Maxwell case, making it sufficient to set several coefficients in the Lagrangian expression to zero.
Lemma 1. Let
denote the concomitant Lagrangian of the metric tensor, the electromagnetic potential (a covector), a scalar field, and their derivatives up to the order specified in the following equation: Based on condition (iii) and using the field dimensions expressed in (iv), under a scaling transformation applied to , we obtain Replacement Theorem. If satisfies Equation (3), as well as assumptions (i)–(iv), then for real manifold dimension
, the general Lagrangian density is [28]
where denote numerical constants. Lemma 2. By differentiating four times with respect to
, taking the limit
, and applying the replacement theorem, we obtainwhere are tensor densities, and the semicolon denotes covariant differentiation with respect to the Christoffel symbols . In a physical problem formulated within the framework of Differential Geometry, the Christoffel symbols correspond directly to the force field governing the physical system under consideration. The underlying differentiable manifold is typically Lorentzian, and particles are assumed to follow geodesic paths, that is, trajectories of stationary action. Lemma 3. The tensor densities
have been determined for the arbitrary order [29]. Based on these results, and assuming the spacetime dimension , it is shown that - (i)
- (ii)
- (iii)
- (iv)
is a linear combination of and , where the brackets [ ] denote antisymmetrization over the indices enclosed, and denote the concomitants of the metric tensor, which are known in the generic case. For example, - (v)
- (vi)
- (vii)
is a linear combination of y
- (viii)
is a linear combination of y
Observation 1. In this way, we obtain the following expression for the action:
We now return to the issue of gauge invariance. We require that all field equations, which must have physical significance, remain invariant under local transformations of the
group. Thus, suppose that
is gauge invariant. Then,
and consequently,
By differentiating Equation (4) with respect to
, it immediately follows that
. Then, by contracting Equation (4) with
and differentiating with respect to
, one finds that:
Multiplying by and contracting with , we obtain .
Let us now suppose that
is gauge invariant. Then,
which implies that
By differentiating Equation (5) with respect to , and evaluating at a point where the metric takes the form , setting and , and we obtain . Next, taking , , and , it follows that . Finally, with and , we find .
Differentiating Equation (5) with respect to
, and taking into account that
, we obtain
Now, by setting and, , we find that .
Finally, let us assume that
is gauge invariant. Then,
Since is linear in , fourth-order derivatives do not appear in . Finally, by differentiating with respect to and subsequently contracting with , it follows that . In this way, we have established the following,
Theorem 1. If is a Lagrangian density of the form,which satisfies hypotheses (i) through (v), then the gauge invariance of the Euler–Lagrange equations imply the gauge invariance of the Lagrangian, which takes the form The terms , , and can be related in four dimensions () through the Gauss–Bonnet theorem.
Observation 2. The brain’s spacetime, considered a manifold, constitutes a central theme of the present work. Building on Weyl’s program for the unification of long-range interactions (gravity and electromagnetism), we associate the purely geometric concept of curvature of a Weyl spacetime with the presence of the local cerebral electromagnetic field. General Relativity arises as a particular case of Equation (6) if we allow it to be expressed without dimensional constants in the Brans–Dicke fashion. The Lagrangian of the latter theory can be derived from Equation (6) by choosing , , and setting all other coefficients to zero. General Relativity is then recovered by identifying .
Corollary 1. Equation (6) enables us to derive the Lagrangian for the Einstein–Maxwell field equations, which we adapt to the Minkowskiian spacetime of the grid cells toroidal manifold by an appropriate choice of constants. The Einstein–Maxwell equations are obtained by replacing the partial derivatives in the Minkowski space formulation with covariant derivatives from Equation (6) by setting , , all other , and identifying .
Observation 3. However, the application of this Lagrangian entails several difficulties. Among them is the question of what guarantees that the manifold possesses metric properties sufficiently similar to those of Minkowski spacetime to allow for a decomposition such as (1). Furthermore, when treated perturbatively, the powers of the cosmological constant Λ that appear in the expansion lead to a loss of predictability in the theory. In addition, for such a decomposition to be carried out in practice, the manifold must exhibit topological properties akin to those of flat spacetime. Yet, topology remains completely undetermined in General Relativity and in the Einstein–Maxwell equations. In what follows, we shall develop this Lagrangian for a toroidal manifold associated with cerebral synaptic activity and examine its properties.
3.4. Synaptic Stability of Helical Geodesics in Einstein–Maxwell Brain Field Equations
Our previous work [
5] aimed to better understand how brain spiral waves arise by studying the geometric properties on regular surfaces of a given topological space where a metric has been incorporated, and specifically, we demonstrated that helices are the only non-trivial geodesics on Minkowskiian tori (
Figure 4).
The problem of particle motion can be simplified by considering a central potential and, specifically, by analysing the Lagrangian corresponding to the Lorentz force experienced by a particle in electromagnetic fields. A persistent difficulty in the analysis of cylindrical spacetimes arises from the fact that they do not asymptotically approach a Minkowski background at large distances [
30]. This behaviour is analogous to the Newtonian gravitational potential of a cylinder mass distribution, which exhibits a logarithmic divergence at large radii
. In General Relativity, a similar divergence appears in the metric component
, further complicating the interpretation of such geometries. As a consequence, the physical meaning of cylindrical spacetimes becomes increasingly ambiguous at large distances, posing limitations on their use in global models. A common resolution to this issue involves restricting attention to configurations that are locally cylindrical but spatially confined within a finite region. This problem can be overcome in systems like thin rings (tori) or needles (finite cylinders), since they are locally cylindrical and limited to a finite region of space.
The solution is constructed within the framework of Weyl’s formalism, which provides a systematic approach for obtaining axially symmetric solutions to the coupled Maxwell–Einstein field equations. The background spacetime is described by two scalar potentials,
and
, both exhibiting axial symmetry. These functions are determined by the following system of field equations:
Partial derivatives are indicated by commas. For convenience, these field equations can be reformulated using the gradient operator
and the Laplacian
defined with respect to the background metric:
If the gradient of
, denoted
, forms an angle
with the radial direction
, then the gradient of
,
, forms an angle of
relative to the same direction. For each solution of the field Equations (7)–(11) within the background space defined by (6), there corresponds an axially symmetric solution to the Einstein–Maxwell field equations characterized by the metric (6):
Different types of singular sources—points, lines, or surfaces—chosen for
in the background space yield different solutions. When these sources are confined to a bounded region, the functions
and
approach constant values at spatial infinity (i.e., as
). These constants can be taken as zero without changing the physics, resulting in an asymptotically Minkowskiian spacetime (Equation (12)) [
31].
The motion of a charged particle can be described by a variety of orbital paths in spacetime, which are solutions to the equations of motion—effectively the geodesics of the background geometry. Charged particles trace out geodesics along timelike curves, thereby revealing the properties of spacetime itself. Equatorial plane orbits, confined to , are characterized by the coordinates and , which obey a system of first-order differential equations representing the geodesic equations in this plane.
The motion is mathematically governed by the three Equations (7)–(9), which encode the conservation of angular momentum per unit mass and restrict the dynamics to the equatorial plane. These equations admit a classical interpretation in terms of radial kinetic energy, rotational energy, and potential energy. Equation (10), obtained from the line element (6), represents an equivalent formulation of the system. With three equations and three variables, the system is well-determined and suitable for numerical integration.
It is essential to find a different function for modelling the electromagnetic potential, ensuring that it is a solution to Laplace’s Equation (10), which plays a significant role in describing stationary processes and constructing vector fields from the potential.
With this formulation of the Maxwell–Einstein Lagrangian, the structure of the toroidal metric and the curved spacetime reduces to Minkowski spacetime in the asymptotic limit, indicating that the geometry is asymptotically flat. Equations (7)–(9) are also symmetric under reflection about the torus, namely [
32],
The resulting physical spacetime, as described by Equation (12), corresponds to a pure line monopole solution, for which
The source in the background space corresponding to a locally cylindrical yet globally toroidal metric exhibits distinctive characteristics. It is most appropriately modelled by considering the background as filled with an incompressible fluid undergoing steady-state potential flow with potential and momentum density , where denotes the fluid’s mass density, distinct from the cylindrical coordinate . This fluid expands over a finite height , and is emitted at a constant rate The fluid flow is confined by two rigid disks attached at the endpoints , each having radius , thereby restricting the fluid’s expansion within the background space.
When the fluid has moved sufficiently far from the confining disks,
, its flow is almost spherical, with a mass flow rate,
and potentials,
Thus, the physical spacetime metric (10) takes the asymptotic form
from which the total mass-energy
of the system can be determined in terms of the mass flow rate
in the (fictitious) background metric:
Near the source, the component of the flow is in the
direction, with
The solutions for the potential can be succinctly summarized as follows:
- (i)
behaves asymptotically, as
where
denotes the total mass-energy of the system.
- (ii)
The boundary conditions require
and
satisfies the Laplace equation,
, throughout the background space.
- (iii)
At spatial infinity,
vanishes according to
The corresponding potential
satisfies Equations (8) and (9) everywhere and tends to zero at infinity.
This set of conditions fully determines the metric coefficients of the physical spacetime, as given in Equation (12).
Field Equation (13), along with the boundary condition (14), impose that, in the vicinity of (
, the potential
assumes the form
Here,
and
represent polar coordinates defined with their origin at the edge of the disks:
The behaviour of
near the edge of the disks, as constrained by Equations (11) and (15), is characterized by
The constants
, and
are uniquely defined functions of the parameters
, and
, which can be determined by solving Equations (7)–(14) entirely. Inserting expression (16) into the physical metric (12) yields
By applying the coordinate transformation
results in the subsequent form
The constants of motion corresponding to the energy
and the angular momentum per unit mass
are introduced, allowing the behaviour of the field source to be determined. Focusing on geodesic trajectories, the Lagrangian reduces to
where a condition is imposed to distinguish the type of particle under consideration:
with
for null geodesics, and
for timelike geodesics. Imposing circular helical geodesics yields the following expression:
where the so-called effective potential is defined as
which is a function of
and allows the energy to be reinterpreted by reducing the problem to one dimension. Next, we analysed the motion of particles along timelike geodesics, though the analysis can be extended to null geodesics.
The model under study was solved using the fifth-order Runge–Kutta–Fehlberg method. Subsequently, we employed the developments by Gonzalez et al. [
33], based on the first three terms of the relativistic Kuzmin–Toomre disk solution family, which were derived from the generalized potential with an infinite number of terms. For each potential model, the dynamics of helical geodesics were analysed, the expression for the effective potential was determined, and under a specific set of initial conditions and parameters, it was concluded that the helical geodesics are always stable, as the effective potential exhibits a minimum regardless of the choice of angular momentum. Furthermore, the set of helical solutions is free from singularities, with the exception of the locally cylindrical ring source located at
, for
.
VNS, as a low-dimensional manifold with toroidal topology, determines the influence of ongoing neural activity through helicoidal geodesics on synaptic strength stability. So, the above results show that VNS is geodesically complete in a Lorentzian manifold with a Minkowskiian spacetime, but what is the physical interpretation of this geodesic completeness?
How, then, can we characterize a singularity? By far, the most satisfactory approach is to consider the presence of ‘holes’ left after the removal of singularities as a criterion for their existence [
34]. These ‘holes’ may be detected by the fact that geodesics of finite length exist; in other words, there should be geodesics that are inextensible in at least one direction, either future or past, but which have only a finite range of affine parameters [
35]. Such geodesics are said to be incomplete.
Accordingly, a spacetime is said to be singular, i.e., it contains singularities, if it possesses at least one incomplete geodesic (in the case of a Lorentzian manifold with a Minkowskiian spacetime, geodesic completeness is equivalent to Cauchy completeness) [
36]. Many examples arise from the failure of geodesic completeness, which corresponds to the intuitive notion of removing singular ‘holes’. In a compact spacetime, every sequence of points has an accumulation point; thus, in a strong intuitive sense, it cannot contain ‘holes’.
From Hawking’s perspective, spacetime is considered singularity-free if it is timelike complete—that is, all timelike geodesics can be extended indefinitely—and if the metric is a well-defined tensor field [
37]. Timelike geodesic incompleteness has immediate physical significance, as it suggests the existence of freely falling observers or particles whose worldlines terminate after a finite amount of proper time. Misner refined this concept by analogy with the Riemannian case, proposing geodesic incompleteness as a necessary condition for spacetime singularities [
38]. He further argued that it suffices for some scalar polynomial constructed from the curvature tensor and its covariant derivatives to become unbounded on an open geodesic segment of finite length, since this implies the geodesic cannot be extended in any extension of the spacetime.
Penrose introduced a new definition of singularities characterized by two novel features: (1) it is perspectival, meaning that the singularity’s significance depends on its causal relationship with points it can influence, and (2) incomplete causal geodesics serve as the hallmark of singularities, replacing the older notion of non-extendable timelike curves [
39]. Earlier, Penrose had proven that under appropriate curvature conditions, a spacetime containing a non-compact Cauchy surface and a trapped surface cannot be future null geodesically complete. Importantly, he focused on causal geodesics rather than arbitrary curves, choosing the former on physical grounds [
40].
3.5. Geodesic Completeness
Now, our interest is to study the behaviour of the geodesics of the torus for this metric
. Geodesics on a Riemannian manifold are obtained as the curves that solve the system of differential equations:
where
is the dimension of the manifold.
is a coordinate system, and the constants
are given in terms of the derivatives of the coefficients
of the metric
:
denotes the coefficients of the inverse of
:
In the case of the torus,
; hence,
Then,
and system (17) is reduced to system (18):
To simplify, let us take
and
, reducing system (18) to
Our objective at this stage is to demonstrate that every non-spacelike geodesic can be extended to arbitrary values of its affine parameter. For this purpose, we shall examine the geodesic equations in metric (12), which, after standard and straightforward calculations, take the form given in (19).
The strategy of the proof is to establish finite bounds for the first derivatives, which implies [
41] that the field is nonsingular and the geodesics are complete. We also consider the second derivatives of the coordinates to demonstrate that these cannot diverge. Our discussion focuses solely on geodesics propagating towards the future; those propagating towards the past can be treated in an analogous manner. To proceed, we divided our analysis into steps, beginning with the simpler geodesics and then addressing the general case.
Senovilla [
42] proposed a singularity-free solution to Einstein’s equations for a perfect-fluid energy-momentum tensor, which also fulfils the stricter causality conditions. Both pressure and energy density are positive throughout the spacetime. Subsequently, Senovilla et al. [
43] showed that the solution is geodesically complete and examined how this result aligns with the general conclusions of the highly powerful singularity theorems.
Using the developments of Senovilla et al., the geodesic completeness of metric (12) for Equation (19) can be demonstrated in the following cases: (i) geodesics in the fluid congruence, (ii) geodesics along the axis, (iii) radial null geodesics, (iv) radial timelike geodesics, (v) null geodesics with no angular velocity, (vi) null geodesics on the hypersurfaces , and (vii) general nonspacelike geodesics. So, we can conclude that in our case of study all the geodesics are complete. Given that each spacelike hypersurface constitutes a global Cauchy surface and that functions as a time coordinate in the solution, it follows that every non-spacelike curve (geodesic or otherwise) can be extended to arbitrary values of its generalized affine parameter. This implies that the solution is free of singularities.
3.7. Geodesic Completeness and Curvature Singularities on the Lorentzian Torus in a Curved Spacetime
Both the space of synaptic activity of grid cells and the associated manifold with toroidal topology exhibit event horizons. In terms of analytical mappings, one can say that the real singularities of the map define the asymptotic regions of spacetime [
45]. Each real singularity corresponds to a specific pattern of synaptic activity, and in particular, the critical points determine the horizons. In the standard interpretation, these singularities connect causally disconnected events; however, the cost of such connectivity is the global loss of distinction between past and future within the region bounded by the singularity’s horizon [
46]. As a consequence, at spatial-type singularities, the time orientation is inverted (
Figure 5) [
47]. The choice of topology corresponds to the choice of boundary conditions for the fields on the horizon, and, therefore, to the definition of the global structure of the manifold.
So, timelike geodesic completeness has an immediate physical significance in that it excludes the possibility that there could be freely moving observers or particles whose histories did not exist after (or before) a finite interval of proper time. A spacetime singularity is a breakdown in spacetime, either in its geometry, in some other basic physical structure, or in its causal chain of events. We investigated the existence and stability of tori equipped with Lorentzian metrics and found that Lorentzian tori –with a Minkowski spacetime– exhibit maximal stability. This result indicates that, in contrast to the Riemannian Hopf theorem, the absence of singularities in the Minkowski spacetime context is neither exceptional nor rigid. The above results build a brain model around the central idea that the dynamics associated with the synaptic spacetime of grid cells associated with a Minkowski spacetime is free of spacetime singularities.
In their original work, Gardner et al. [
1] employed a Lorentzian torus with a Minkowski spacetime in their topological approach to the synaptic space of individual grid cell modules in the VNS. As the simplest example of a four-dimensional Lorentzian manifold, Minkowski space is topologically trivial and globally asymptotically flat, making it the most elementary model of spacetime within the framework of General Relativity. Trivial topology ensures that local flatness extends globally, yielding a Minkowskiian geometry. In contrast, non-trivial topologies restrict flatness to a local property, potentially allowing for curvature and spacetime singularities.
For Minkowskiian manifolds, the compactness of the manifold implies completeness. In contrast, there are Lorentzian metrics on the torus that are not complete. This striking fact motivated the search for sufficient assumptions under which compactness implies geodesical completeness of such a manifold or, more generally, of a compact indefinite Lorentzian manifold [
48]. Consequently, we may ask: What is the relationship between curvature and completeness? The question is quite general, and its answer is somewhat surprising: complete and incomplete Lorentzian metrics with the same curvature on a torus exist.
That is why currently, a singularity represented by a geodesic incompleteness ca be classified according to one or more of the following types: (i) a singularity manifested by the curvature scalar: a scalar polynomially constructed from , and its covariant derivatives diverges along the geodesics; (ii) a singularity manifested by parallel propagation of curvature: no scalar diverges, but a component of the tensor or one of its covariant derivatives diverges along the geodesic; and (iii) a non-curvature singularity: neither curvature scalars nor curvature tensor components diverge. Therefore, the study of the geodesic completeness of compact Lorentzian manifolds requires a more complex analysis.
It is essential to (a) find a physical interpretation for these global singularities and (b) identify a suitable physical condition on spacetimes, such that any compact spacetime satisfying it is geodesically complete. In either case, the study of geodesic completeness in compact manifolds leads to a deeper understanding of the issues posed by singularities. In particular, examining the completeness of compact Lorentzian manifolds sheds light on the concept of singularity in any Lorentzian manifold.
Historically, compact curved spacetimes have been largely excluded from studies in General Relativity. This is primarily because they contain closed timelike curves, which carry troublesome implications from the standpoint of causality. However, since the study of wormholes has gained prominence in cosmology, there has been speculation that the laws of physics may allow—or even necessitate—the existence of closed timelike curves [
49,
50,
51]. As a result, the main objection to using curved compact spacetimes as models of our physical universe has diminished. In fact, the usual constraints imposed by causality theory (such as global hyperbolicity and strong causality) may be weakened and replaced by alternative conditions that still prevent paradoxes [
52,
53].
From a less radical standpoint, we can present the following argument, which is fully consistent with causality theory. The exceptional importance of field theory on compact curved Lorentzian manifolds is well known. However, one must bear in mind that the standard procedure for applying this framework to a flat Lorentzian manifold becomes problematic when the Lorentzian geometry is arbitrarily curved [
54].
Milnor established a certain connection between curvature and completeness by proving that every flat compact Lorentzian manifold is complete [
55]. A natural next step following this result is to address the question: Is every compact Lorentzian manifold complete?
If the manifold also admits a temporal Killing (or conformal Killing) vector field, then the answer is affirmative. A natural follow-up question is: Which compact flat Lorentzian manifolds admit a temporal Killing vector field? Furthermore, the lack of temporal orientability would obstruct the existence of such a vector field; however, this does not preclude completeness if the temporally orientable finite-sheeted Lorentzian cover admits a Killing (or conformal Killing) vector field. More generally, one may ask: Which compact Lorentzian manifolds admit a finite-sheeted Lorentzian cover that possesses a temporal Killing (or conformal Killing) vector field? It can be shown that every flat Lorentzian 2-torus admits such a cover. From the arguments in [
56], it follows that every flat Lorentzian
-torus also admits one. That is, every flat Lorentzian 2-torus admits a temporal Killing vector field. Consequently, since there exist Lorentzian tori that are not temporally orientable, none of these can be conformally flat [
57]. That is, only curved spacetimes can contain Lorentzian tori with temporal singularities.
The line element (12) admits an Abelian symmetry group , with the Killing vectors and being globally defined. Both are spacelike, mutually orthogonal, and orthogonally transitive. Since (12) is globally hyperbolic, no Cauchy horizon can exist. The fluid congruence is trivially complete, and through every point of the manifold, there passes a worldline of this congruence. Moreover, given the properties of the solution, any possible singularities would necessarily possess some extension and thus would be reflected in the curvature invariants. Accordingly, if a component of the tensor diverges, the singularity in the arbitrarily curved Lorentzian torus must correspond to type (ii) in the classification above.
The absence of a Hopf–Rinow-type theorem for indefinite compact manifolds opens the door to numerous new problems related to completeness. Despite some significant results, such as those by Marsden [
58] and Carrière [
59], the list of naturally arising open questions remains extensive. For instance, (a) until recently, the Clifton–Pohl torus was the only known example of an incomplete compact semi-Riemannian manifold in the literature, and (b) very little is known about the structure of the conformal curved moduli space of Lorentzian metrics on a torus.
From the perspective of completeness, connections arising from an indefinite metric on a compact manifold lie somewhere between affine connections (there exist simple examples of incomplete affine connections on
) [
60] and Lorentzian connections (where no incomplete Lorentzian connections exist on compact manifolds). In particular, for the construction of Lorentzian tori, one first builds a class of incomplete metrics on a cylinder featuring closed, incomplete geodesics. This provides an intuitive framework to understand how incompleteness can arise in geodesics whose images are contained within a compact set.
Thus, the technical need to induce Lorentzian tori within a compact setting requires the use of coordinate systems in which incompleteness is concealed. As a result, when examining the final expression of the metric on the torus, it is not immediately intuitive that the metric is incomplete. The fact that the Clifton–Pohl torus is not geodesically connected was already noted in [
61], where three families of incomplete metrics on
(
are presented. When these are induced on a torus, the resulting space is geodesically connected for the first two metrics and geodesically disconnected for the third [
62].
While the study of the conformal moduli of Riemannian metrics on the torus
is fully completed, the analogous problem in the Lorentzian setting appears far more complex and remains largely unsolved [
63]. This situation is not new in Lorentzian geometry; consider, for example, the related problem: How many conformal classes of Lorentzian metrics exist on simply connected surfaces? The Riemannian counterpart of this problem, thanks to the uniformization theorem, is completely resolved. However, the Lorentzian case introduces new elements—such as the assignment of a conformally invariant boundary and the emergence of characteristic points, twins, corners, barriers, and so forth—that significantly increase the complexity of the problem and, to date, only allow for very particular results [
64].
In [
65], Tipler demonstrated that all geodesics are complete both to the future and to the past in a spacetime containing a compact maximal Cauchy surface, provided the following condition holds: there exist fixed positive constants
and
, such that
for every timelike geodesic (with tangent vector
) intersecting orthogonally at
, where
denotes the affine parameter. This condition, however, is not satisfied in metric (10), since for any fixed pair of constants
and
, the above integral is always positive but not bounded away from zero. Indeed, by selecting geodesics with sufficiently large initial
, the integral can be made arbitrarily small and thus less than any preassigned constant
.
From the preceding result, it follows that a Lorentzian torus embedded in a curved spacetime cannot be assumed to be free of global singularities. Specifically, when analysing the geodesic connection, there is the possibility that, unlike the Minkowskiian case, complete geodesics exist on compact indefinite manifolds with velocities that are not contained within the compact region (
Figure 6) [
66].
A spacetime singularity denotes a failure in the manifold’s structure. All causal geodesics are terminated by generalized curvature singularities, which are categorized through the interplay between causal structure and curvature intensity. We consider the emergence of both local and global singularities in the brain as a curved Lorentzian manifold, originating from the collapse of a toroidal manifold dust cloud governed by Einstein–Maxwell dynamics with a non-zero velocity profile. In this framework, future-directed null geodesics exit the boundary of the synaptic activity cloud, whereas in the past, they converge to the singularity. Déjà vu may serve as an observable manifestation of such an effect. To preserve physical plausibility, the singularity must satisfy the conditions of a strong curvature singularity.
So, déjà vu may be reinterpreted from a geometric standpoint, specifically as a curvature spacetime singularity embedded in the global dynamics of the brain toroidal manifold. What occurs in the brain during such experiences? Neuroscientific research suggests that déjà vu is not indicative of a pathological condition, nor is it a failure of memory [
67,
68]. Despite this, no unified explanatory model has gained consensus. Nevertheless, a prevailing hypothesis posits that déjà vu arises when specific neural circuits interpret a current experience as resembling a past event, even in the absence of explicit recall. This suggests that the phenomenon may be rooted in memory processes, wherein the brain detects partial overlaps between present and previously encoded experiences. Einstein–Maxwell’s brain field equations lead to the formation of inherent curvature spacetime singularities; so, déjà vu could be an inherent process in synaptic dynamics.