Previous Article in Journal
Reinterpretation of Fermi Acceleration of Cosmic Rays in Terms of Ballistic Surfing Acceleration in Supernova Shocks
Previous Article in Special Issue
New Conformally Invariant Born–Infeld Models and Geometrical Currents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Canonical Quantization of Metric Tensor for General Relativity in Pseudo-Riemannian Geometry

1
Basic Science Department, Faculty of Engineering, Future University in Egypt (FUE), New Cairo 11835, Egypt
2
Department of Physics, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
3
Department of Mathematics, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
4
Basic Science Department, Modern Academy for Engineering and Technology, Al Hadaba Al Wusta, Cairo 11571, Egypt
5
Physics Department, Faculty of Science, Benha University, Benha 13518, Egypt
*
Author to whom correspondence should be addressed.
Physics 2025, 7(4), 52; https://doi.org/10.3390/physics7040052
Submission received: 23 April 2025 / Revised: 17 August 2025 / Accepted: 1 September 2025 / Published: 20 October 2025
(This article belongs to the Special Issue Beyond the Standard Models of Physics and Cosmology: 2nd Edition)

Abstract

By extending the four-dimensional semi-Riemann geometry to higher-dimensional Finsler/Hamilton geometry, the canonical quantization of the fundamental metric tensor of general relativity, i.e., an approach that tackles a geometric quantity, is derived. With this quantization, the smooth continuous Finsler structure is transformed into a quantized Hamilton structure through the kinematics of a free-falling quantum particle with a positive mass, along with the introduction of the relativistic generalized uncertainty principle (RGUP) that generalizes quantum mechanics by integrating gravity. This transformation ensures the preservation of the positive one-homogeneity of both Finsler and Hamilton structures, while the RGUP dictates modifications in the noncommutative relations due to integrating consequences of relativistic gravitational fields in quantum mechanics. The anisotropic conformal transformation of the resulting metric tensor and its inverse in higher-dimensional spaces has been determined, particularly highlighting their translations to the four-dimensional fundamental metric tensor and its inverse. It is essential to recognize the complexity involved in computing the fundamental inverse metric tensor during a conformal transformation, as it is influenced by variables like spatial coordinates and directional orientation, making it a challenging task, especially in tensorial terms. We conclude that the derivations in this study are not limited to the structure in tangent and cotangent bundles, which might include both spacetime and momentum space, but are also applicable to higher-dimensional contexts. The theoretical framework of quantization of general relativity based on quantizing its metric tensor is primarily grounded in the four-dimensional metric tensor and its inverse in pseudo-Riemannian geometry.

1. Introduction

The fundamental principles of general relativity (GR) and quantum mechanics (QM) are believed to be incompatible [1,2]. Despite the extensive investigations, a solution to this long-standing problem has yet to be found. A novel approach has been proposed, which involves a canonical quantization approach to quantize the metric tensor of GR in pseudo-Riemann geometry. The main assumption of this novel method is the requirement for a proper generalization of the current formulations of GR and QM. The integration of relativistic gravitational fields in the uncertainty principle [3,4,5], gravitational self-energy in the spontaneous collapse of superposition principle [6,7], and topology change or interaction in the principle of general covariance are essential ingredients for the generalization for QM [8]. The generalization for pseudo-Riemann geometry is discussed in the present paper; for more information, see Refs. [9,10,11,12,13,14,15,16].
Before the pseudo-Finsler geometry and its particular mathematical properties are addressed, it is worth recalling what was first considered for the possible generalization of pseudo-Riemann geometry [17]. In pseudo-Riemann geometry, the metric g and the manifold M with dim ( M ) = n , that is ( M , g ) , there is a natural metric g ¯ on the tangent bundle T M , so ( T M , g ¯ ) represents a generalization of the Riemann manifold. This manifold is known as the Sasaki metric [18,19,20]. Eduardo Caianiello introduced a similar procedure for quantization of GR [21,22]. The corresponding line element can be expressed in the local coordinates on T M , namely ( x , x ˙ ) , as
d s ˜ 2 = g μ ν d x μ d x ν + g μ ν D x ˙ μ D x ˙ ν ,
where g μ ν is the fundamental metric tensor, the dot on top denotes the time derivative, and D x ˙ μ = d x ˙ μ + Γ ν δ μ x ˙ ν d x δ is the covariant derivative. Γ ν δ μ is the affine connection. That is, the generalized metric tensor can be expressed as [23]
g ˜ μ ν = g μ ν + g α β Γ ρ μ α Γ γ ν β ,
which leads to
(3) g ˜ ( n + μ ) ( n + ν ) = g μ ν , (4) g α β Γ ρ μ α Γ γ ν β = A μ ν , (5) g ˜ μ ( n + ν ) = g ν σ Γ ϵ μ σ x ˙ ϵ = B μ ν .
The Latin letter indexes denote the space coordinates.
The additional component that, when is added to the conventional metric tensor, produces the generalized metric (2), is represented by A μ ν . Now the Sasaki metric can be defined in line with Equation (2):
g ˜ μ ν = [ g μ ν + A μ ν B μ ν B μ ν T g μ ν ] .
By exploring the relationship between metric geometry, pseudo-Riemann geometry ( M , g ) , and the tangent bundle T M geometry, which is supplied with the Sasaki metric g ˜ , various types of rigidity have been suggested [24,25,26]; more details on the correct connections between ( M , g ) and ( T M , g ˜ ) can be found in the literature, such as Refs. [27,28]. The interconnection between difference space is contingent upon a specific parametrization. The extension to Finsler/Hamilton space, however, is there to introduce quantum mechanical components. Here, we convert the metric tensor on Finsler/Hamilton into a metric tensor on the pseudo-Riemann manifold.
As mentioned above, another type of generalization for the metric tensor which characterizes the Riemann geometry ( M , g ) , the canonical quantization approach proposed here, can be directly derived from the Finsler structure F, as discussed in Section 2. Consider an n-dimensional pseudo-Finslerian manifold ( M , F ) , where F : T M R . Here, we define the coordinated ( x , x ˙ ) as ( x μ , x ˙ μ ) the local coordinates on T M . Furthermore, π : T M { 0 } M represents the natural projection. Similar to other pseudo-Finslerian quantities, the Finsler metric is a function on T M [29] and can be related to essential quantities including the non-degenerate fundamental tensor [30]
g μ ν = 1 2 2 x ˙ μ x ˙ ν F 2 ( x , x ˙ ) ,
and the Cartan tensor [31]
C μ ν σ = 1 4 3 x ˙ μ x ˙ ν x ˙ σ F 2 ( x , x ˙ ) .
In Section 2, we introduce a discretized version (likely quantized) of Finsler structure, where the direction defined by x ˙ is replaced by that of the momentum space p. In this regard, a kinematic theory of a quantum particle that lives and moves in the discretized Finsler manifold to be utilized. Also, let us emphasize that the translation x ˙ m x ˙ p is not exactly similar to the one proposed for Hamilton geometry [32]. The Finsler geometry characterized by x and x ˙ is dual to the Hamilton geometry which is characterized by x and p. Both types of geometries are mappings connecting the spaces of velocity and of momentum, respectively [33]. As soon as duality of both the velocity and the momentum spaces is nonlinear, the Legendre transformation does not linearly translate x ˙ in Finsler to p in Hamilton geometry. The rigorous mathematical procedure we used to transform F ( x , x ˙ ) to F ( x , m x ˙ ) H ( x , p ) is essentially based on the mathematical properties of the Finsler structure. We employ this homogeneous characteristic to introduce at least discreteness, if not quantization, into the tangent bundle.
In order to determine the mathematical equivalence between F ( x , p ) and H ( x , p ) , let us analyze the procedure that strictly upholds the homogeneous property of F ( x , m x ˙ ) and compare it with the Legendre transformation of F ( x , x ˙ ) to H ( x , p ) . The Hamilton space is identified by the manifold M and the Hamilton structure H ( x , p ) . This space is described by the cotangent space T M ( x , p ) H ( x , p ) R , where the Hamilton structure H ( x , p ) is differentiable on the cotangent bundle T M and continuous on the null section π : T M M . The function T M ( x , x ˙ ) L ( x , p ) R defines a Lagrangian space. It is quite straightforward that the Lagrangian space is characterized by the manifold M and L ( x , x ˙ ) . The Legendre transformation leads to [34]
H ( x , p ) = p a x ˙ a L ( x , x ˙ ) ,
where x ˙ = { x ˙ a } are solutions of x ˙ b = L ( x ^ , x ˙ ^ ) / x ˙ b . In this regard, let us mention that the solution is obtained by identifying L ( x , x ˙ ) with F ( x a , m x ˙ b ) so that
x ˙ b = F ( x a , m x ˙ b ) x ˙ b = m F ( x a , x ˙ b ) x ˙ b + x b x a , x ˙ b 1 | x a | 2 ,
where indexes a , b { 0 , 1 , , 7 } . Also, it can be observed that Equation (9) is satisfied by this particular solution, wherein the indexes are lowered or raised with the corresponding metric tensor. One concludes that when L ( x , x ˙ ) is equivalent to F ( x a , m x ˙ b ) , then F ( x , p ) H ( x , p ) . This study focuses on a Finsler metric that can be derived from the Hessian of F 2 ( x , m x ˙ ) . The derivation of the Hamilton metric to be addressed elsewhere.
The introduction of a discrete Finsler structure and its transformation into a Hamiltonian framework are examples of the proposed generalization of spacetime geometry. The four-dimensional limitation is a respectable approximation that appears to hold true at all levels. At relatively low (quantum) scale, this approximation has to be either decreased or removed. The discretized Finsler structure is transformed into a Hamiltonian framework in order to generalize the spacetime geometry and thereby attenuate the conventional approximation. We make a further step by introducing generalization to quantum mechanics by making further adjustments based on a relativistic generalized uncertainty principle (RGUP). Using RGUP, the quantum mechanical framework becomes compatible with finite gravitational fields. In this paper, we provide a first-principal derivation of the quantum mechanically corrected metric tensor. Various studies already cover physical grounding, predictive ability, and technical clarity; see, e.g., [9,10,12,13,14,15,16].
The paper is organized as follows. The discretized Finsler structure is discussed in Section 2. This Section also describes a discretized Hamilton geometry in which four-dimensional spacetime is coupled to four-dimensional momentum space. The anisotropic conformal transformation of arbitrary high-dimensional metric is introduced in Section 3. The utilization of the Klein metric is considered for pedagogical reasons, as it belongs to the simplest Finsler metrics. In Section 4, a detailed derivation of the inverse for this metric is outlined. The translation from the quantized metric and its inverse on a higher-dimensional Hamilton manifold to the fundamental metric tensors on the four-dimensional Riemann manifold are discussed in Section 5 and Section 6, respectively. The conclusions are given in Section 7.

2. Discretized Finsler Structure

For a quantum particle with positive mass m, the Finsler structure can be reformulated as
F x μ , m x ˙ ν = F ( x μ , p ν ) , m R + ,
where x μ and p ν are quantum mechanical operators whose auxiliary four-vectors are x μ x 0 μ = ( x 0 0 , x 0 i ) = ( c t , x 0 i ) and p μ p 0 μ = i / x 0 μ = ( p 0 0 , p 0 i ) = ( E / c , p 0 i ) , where, E represents energy, t stands for time, c denotes the speed of light, and is the reduced Planck constant. For the sake of simplicity, we assume that the operators x μ and p μ , as defined in QM without gravitational field, are considered to remain unchanged in generalized QM. QM without a gravitational field refers to the QM that relies on the Heisenberg uncertainty principle (HUP), which was developed in vanishing gravitational field [3,35,36,37,38]. For further details on recent studies in order to replace HUP by the generalized uncertainty principle (GUP), in which gravitational fields are included, see [3,37,38].
With respect to time operator x 0 0 , we intentionally refer to Ref. [39] by Eric Galapon where the time operator is considered to behave like energy, is bounded and self-adjoint. To be more specific, we assume that the time operator is canonically conjugated to the Hamiltonian H. The Hamiltonian is considered to have discrete spectra. This choice is discussed shortly. This gives us reason to believe that this pair of operators follows the time–energy canonical commutation relation [ T , H ] = i . In addition, this enables us to represent the known Heisenberg equation of motion d T / d t = [ T , H ] / i [39]. The latter indicates that time operators are always progressing in steps of parametric time t, which means that T meets the covariance requirement T ( t ) = T ( 0 ) + t for all times [39,40]. Given the semibounded Hamiltonian of quantum mechanics, we are aware of the historical rejection of the existence of any such kind of time operator [41,42]. As a non-self-adjoint, covariant, positive operator-valued measures (POVM) observable, time can now be reintroduced into quantum mechanics thanks to the recent critical finding that quantum observables need not be self-adjoint POVM [43]. In this regard, let us recall that Pauli’s claim [44] has been refuted by several counterexamples, leading to a growing focus on the study of self-adjoint time operators. This highlights that the assumption of a self-adjoint operator canonically conjugate to a semibounded or discrete Hamiltonian does not contradict [45]. The statement that there are just non-self-adjoint time operators in single Hilbert quantum mechanics appears to be non-universal. The present study is an illustration of that type.
In the various attempts to overcome Paulis’ highly-established no-go theorem, we consider bounded time operators, particularly for Hamiltonians that exhibit discrete spectra. The averaging involved in the proposed quantization, along with the lack of any other alternatives, suggests that Hamiltonians with discrete spectra are quite appropriate [9,10,12,13,14,15,16]. Other recent studies, such as Refs. [46,47], which have extended the bounded time operators to include the unbounded time operators and alternative formulations of Hilbert space, are discussed in another context, especially given the rigorous mathematical analysis presented in Ref. [48].
It is understandable that the Relativistic Generalized Uncertainty Principle (RGUP) [3] introduces modifications to the HUP where higher-order momentum terms are multiplied by the same order of the RGUP parameter. Therefore, the general Hamiltonian can be expanded in a Taylor series with respect to the RGUP parameter β . The corrections to the Hamiltonian H 0 = p 0 2 / ( 2 m ) + v ( t ) , where v ( t ) represents the correction part, are solely adjustments to the kinetic portion of the Hamiltonian, specifically the first term. This indicates that the corrections introduced are independent of the potential, represented by the second term. That is, we implement perturbations to H, which necessitates that the proposed RGUP corrections must be smaller than the uncertainties of the perturbations introduced to H. It is understandable that the present iteration of RGUP corresponds quite well to Hamiltonians with discrete spectra.
It was noted that GUP exhibits characteristics of being three-dimensional and non-relativistic, in addition to other shortcomings including breaking Lorentz covariance [5,49,50,51,52], violating linearity of the dispersion relations [53], and not coupling the uncertainty principle to the spacetime metric [54,55]. The relativistic generalization of the HUP was suggested to resolve the mentioned shortcomings that arise from HUP and GUP. The RGUP [3] is briefly reviewed in Appendix A. The RGUP impacts the operators and their auxiliary four-vectors so that x μ = x 0 μ and p μ = ( 1 + β p 0 ρ p 0 ρ ) p 0 μ , where ρ is a dummy index and β ’s upper bound can be empirically and/or observationally determined [56].
For the auxiliary four-vectors x 0 μ , x ˙ 0 μ , and p 0 ν , let us start with the conventional Finsler metric F ( x 0 μ , x ˙ 0 ν ) , which appears to preserve the entire mathematical properties of the Finsler structure. To use the RGUP approach (Appendix A), consider the kinematics of the free-falling quantum particle, by which determining the uncertainty principle becomes possible through the utilization of energy and momentum. The positive mass preserves the homogeneous properties of F, resulting in F ( x , m x ˙ ) = m F ( x , x ˙ ) . Then, RGUP can be directly applied on F ( x , m x ˙ ) . The RGUP also preserves the homogeneous properties of F ( x , p ) . In F ( x , p ) , the auxiliary four-moment becomes a subject of RGUP, p 0 = ϕ ( p 0 ) p 0 , where ϕ ( p 0 ) = 1 + β p 0 ρ p 0 ρ and the generalization of p 0 is provided by the left-hand site. The function ϕ ( p 0 ) assures that all quantities defining ϕ ( p 0 ) obey the same dimension as the generalized p 0 . The coefficients refer to the phase space coordinates ( x 0 , p 0 ) . Besides the positive homogeneity, i.e., F is positively 1-homogeneous in p 0 , the relativistic four-momentum so that F ( x , ϕ ( p 0 ) p ) = ϕ ( p 0 ) F ( x , p ) , ϕ ( p 0 ) R + ; the resulting discretized F ( x , ϕ ( p 0 ) p ) presumably preserves the following properties:
  • non-degeneracy, i.e., the metric tensor is non-degenerate on T M { 0 } ;
  • smoothness, i.e., F is smooth on T M { 0 } , on the complement of the zero section.
The function ϕ ( p 0 ) has a few key roles. So far, we consider the extension of the HUP to RGUP, and then a successful incorporation gravitational effects into QM, avoiding the shortcomings that have been discussed earlier [3]. Other roles such as (i) generalizing the fundamental tensor and thereby quantizing GR [57], and (ii) retaining the special curvature of the Randers metric (Appendix B), are elaborated Section 3.
For the discretized Finsler–Hamilton structure H ( x 0 , ϕ ( p 0 ) p 0 ) , to be 1-homogeneous in p 0 , the functions ϕ ( p 0 ) must be 0-homogeneous in p 0 . It is understandable that the RGUP parameter β is 2 -homogeneous, because both β and p 0 do not depend on x 0 . β = β 0 G / ( c 3 ) = β 0 ( l p / ) 2 where G is the gravitational constant and l p is the Planck length [37,38]. As soon as the function ϕ ( p 0 ) is positive and 0-homogeneous in p 0 , H ( x 0 , ϕ ( p 0 ) p 0 ) = ϕ ( p 0 ) H ( x 0 , p 0 ) is then 1-homogeneous in p 0 .
In the discretized Hamilton space of the spacetime cotangent bundle, it is conjectured that the fiber manifold constitutes the 4-momentum forming momentum space, while the base manifold is represented by the spacetime [58]. This constructs the eight-dimensional geometry composed of four-dimensional spacetime and four-dimensional momentum space. By using a parameterization ζ , a generic point can be characterized by the coordinates x a ( ζ ) ( x 0 μ ( ζ ) , p 0 ν ( ζ ) ) , which are phase space coordinates. Both position and momentum variables, x 0 and p 0 , respectively, are appropriately parameterized. The application of the RGUP approach suggests that p 0 ( 1 + β p 0 ρ p 0 ρ ) p 0 [37,38]. As so, then the quantized fundamental tensor can be derived as
g ˜ μ ν = 1 2 2 p 0 μ p 0 ν H 2 ( x 0 μ , ϕ ( p 0 ) p 0 ν ) .
The metric (12) is a Finsler-type metric because of the Hessian of the squared discretized structure. The coordinates in Finsler–Hamilton space are given as x 0 a = ( x 0 μ , p 0 ν ) .
By utilizing positive homogeneity, F ( x 0 μ , x ˙ 0 ν ) (Finsler) is transformed into H ( x 0 μ , p 0 ν ) (Hamilton). This transformation allows us to suggest that the resulting expression of H 2 ( x 0 μ , ϕ ( p 0 ) p 0 ν ) may be equivalently expressed by the corresponding Finsler structure F 2 ( x 0 μ , ϕ ( p 0 ) p 0 ν ) . For the sake of simplicity, let us suggest that the Klein metric represents F ( x 0 μ , ϕ ( p 0 ) p 0 ν ) and thereby H ( x 0 μ , ϕ ( p 0 ) p 0 ν ) . The Klein metric on B n ( 1 ) (Section 3), in which RGUP is integrated, can be given as
F 2 ( x 0 μ , ϕ ( p 0 ) p 0 ν ) = ϕ 2 ( p 0 ) | p 0 ν | 2 | x 0 μ | 2 | p 0 ν | 2 + x 0 μ , p 0 ν 2 ( 1 | x 0 μ | 2 ) 2 ,
where | · | and · are, respectively, the standard Euclidean norm and the inner product on R n . Accordingly, Equation (12) can be reexpressed as
g ˜ μ ν Finsler = 1 2 2 p 0 μ p 0 ν F 2 ( x 0 μ , ϕ ( p 0 ) p 0 ν ) ,
This represents an additional basis for asserting that the resulting metric is of the Finsler type.
It must be noted that the derivative of g ˜ μ ν (14) is valid for arbitrary higher-dimensional structures. Section 3 provides a outline of all the relevant particularities. Henceforth, we denote an arbitrary high-dimensional structure using the structure F.

3. Anisotopic Conformal Transformation of Metric in Higher-Dimensional Geometry

Let us start with the components of the metric tensor associated with the arbitrary higher-dimensional structure F, i.e., that of g μ ν where
g μ ν = δ μ ν 1 | x 0 | 2 + x 0 σ x 0 γ δ ν σ δ μ γ ( 1 | x 0 | 2 ) 2 ,
and δ μ ν represents the Euclidean metric. Considering the anisotropic confromal transformation of F, one finds that
F ¯ = ϕ ( p 0 ) F .
As discussed in Section 2, ϕ ( p 0 ) is a 0-homogeneous function in p 0 and therefore can be expressed as
ϕ ( p 0 ) = 1 + κ ( p 0 0 ) 2 F 2 ,
where κ = β / ( p 0 0 ) 2 .
To determine the components of the metric tensor associated with F ¯ , namely g ¯ μ ν , we introduce the following lemma.
Lemma 1.
The first and second derivatives of the function ϕ ( p 0 ) , with respect to p 0 , are given separately by [59,60]
(18) ϕ μ : = ˙ μ ϕ ( p 0 ) = 2 κ F ( p 0 0 ) 3 p 0 0 l μ F δ 0 μ , (19) ϕ μ ν : = ˙ ν ˙ μ ϕ ( p 0 ) = 2 κ ( p 0 0 ) 2 g μ ν 4 κ F ( p 0 0 ) 3 l ν δ 0 μ + l μ δ 0 ν + 6 κ F 2 ( p 0 0 ) 4 δ 0 ν δ 0 μ ,
where ˙ μ / x ˙ μ and l μ : = F / p 0 μ .
Proof. 
If ϕ μ ν : = ˙ ν ˙ μ ϕ ( p 0 ) , Equation (19) reads
ϕ μ ν = 2 κ l ν ( p 0 0 ) 3 p 0 0 l μ F δ 0 μ 6 κ F ( p 0 0 ) 4 δ 0 ν p 0 0 l μ F δ 0 μ + 2 κ F ( p 0 0 ) 3 δ 0 ν l μ + p 0 0 l μ ν l ν δ 0 μ = 4 κ F ( p 0 0 ) 3 l ν δ 0 μ + l μ δ 0 ν + 2 κ F ( p 0 0 ) 2 l μ l ν + 6 κ F 2 ( p 0 0 ) 4 δ 0 ν δ 0 μ + 2 κ F ( p 0 0 ) 2 l μ ν ,
where
l μ ν = ˙ μ l ν ,
g μ ν = F l μ ν + l μ l ν .
Proposition 1.
Equation (14) expresses an anisotropic transformation of the metric tensor
g ˜ μ ν = ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) F 2 ( p 0 0 ) 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 l μ l ν 4 κ F 3 ( p 0 0 ) 3 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 ( l μ δ 0 ν + l ν δ 0 μ ) + 2 κ F 4 ( p 0 0 ) 4 3 ϕ ( p 0 ) + 2 κ F 2 ( p 0 0 ) 2 δ 0 μ δ 0 ν .
Proof. 
g ˜ μ ν = ϕ 2 ( p 0 ) g μ ν + F 2 ( ϕ μ ϕ ν + ϕ ( p 0 ) ϕ μ ν ) + 2 ϕ ( p 0 ) F ( l μ ϕ ν + l ν ϕ μ ) = ϕ 2 ( p 0 ) g μ ν + F 2 2 κ F ( p 0 0 ) 3 2 p 0 0 l μ F δ 0 μ p 0 0 l ν F δ 0 ν + ϕ ( p 0 ) F 2 2 κ ( p 0 0 ) 2 g μ ν Finsler 4 κ F ( p 0 0 ) 3 ( l ν δ 0 μ + l μ δ 0 ν ) + 6 κ F 2 ( p 0 0 ) 4 δ 0 ν δ 0 μ + 2 ϕ ( p 0 ) F 2 κ F ( p 0 0 ) 3 l ν p 0 0 l μ F δ 0 μ + l μ p 0 0 l ν F δ 0 ν = ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) F 2 ( p 0 0 ) 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 l μ l ν 4 κ F 3 ( p 0 0 ) 3 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 l μ δ 0 ν + l ν δ 0 μ + 2 κ F 4 ( p 0 0 ) 4 3 ϕ ( p 0 ) + 2 κ F 2 ( p 0 0 ) 2 δ 0 μ δ 0 ν .
Based on the function ϕ ( p 0 ) = 1 + κ ( p 0 0 ) 2 F 2 , Equation (23) can be rewritten as
(25) g ˜ μ ν = ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) ( p 0 0 ) 2 F 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 ( p 0 0 ) 2 3 ϕ ( p 0 ) 1 l μ l σ δ μ ν δ μ σ F p 0 0 3 ϕ ( p 0 ) 1 l σ δ 0 ν δ μ ν δ μ σ + δ 0 μ δ ν μ δ ν σ + 2 F 2 5 ϕ ( p 0 ) 2 δ 0 μ δ 0 σ δ ν μ δ ν σ g μ ν , (26) = ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) ( p 0 0 ) 2 F 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 ( p 0 0 ) 2 3 ϕ ( p 0 ) 1 l μ l σ g σ μ F p 0 0 3 ϕ ( p 0 ) 1 l σ δ 0 ν g σ μ + δ 0 μ g σ ν + 2 F 2 5 ϕ ( p 0 ) 2 δ 0 μ δ 0 σ g σ ν .
This is the resulting generalized metric on the phase space manifold. According to the properties of ϕ ( p 0 ) , especially that ϕ ( p 0 ) does not depend of x 0 , the second term in Equation (26) is non-vanishing.
The anisotopic conformal transformation of the inverse metric reveals further properties of the metric, which are readily apparent upon derivation. Section 4 just below complete elaboration on the derivation of the inverse metric.

4. Anisotopic Conformal Transformation of Inverse Metric in Higher-Dimensional Geometry

Determining the transformation of the inverse metric under an anisotropic conformal transformation in higher dimensions, where the conformal factor is dependent on both position and direction, is quite challenging task [61]. Nevertheless, we have managed to derive this transformation in a tensorial form, incorporating the specific conformal factor. The anisotropic conformal transformation of the inverse higher-dimensional metric tensor can be obtained by considering the following Lemma:
Lemma 2.
Let g μ ν and m μ ν be two n × n symmetric matrices and C = C μ be an n-dimensional vector, which satisfy the relation [62,63]
g μ ν = m μ ν + λ C μ C ν ,
where λ is a constant. Then
det g μ ν = 1 + λ C 2 det m μ ν .
Assuming that m μ ν is non-degenerate with m μ ν 1 = m μ ν and 1 + λ C 2 0 , g μ ν is invertible and g μ ν = g μ ν 1 is given by
g μ ν = m μ ν λ 1 + λ C 2 C μ C ν ,
where C μ = m μ ν C ν and C = m μ ν C μ C ν .
Proof. 
Under proposition 11.2.1, Ref. [62] (p. 287) provides the proof of Lemma 2. The proof is also given with Lemma 1.1 in Ref. [63] (p. 3). □
Let ( M , F ) be an n-dimensional manifold with the metric tensor g μ ν . Consider the anisotropic transformation F ˜ = ϕ ( p 0 ) F , where ϕ ( p 0 ) = 1 + κ F 2 / ( p 0 0 ) 2 . Then, the metric g ˜ μ ν corresponding to the higher-dimensional F is given by
g ˜ μ ν = ϕ 2 ( p 0 ) + 2 ϕ ( p 0 ) κ F 2 ( p 0 0 ) 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 l μ l ν 4 κ F 3 ( p 0 0 ) 3 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 ( l μ δ 0 μ + l ν δ 0 ν ) + 2 κ F 4 ( p 0 0 ) 4 3 ϕ ( p 0 ) + 2 κ F 2 ( p 0 0 ) 2 δ 0 μ δ 0 ν .
Let us rewrite g ˜ μ ν as
g ˜ μ ν = ϕ 2 ( p 0 ) + 2 ϕ ( p 0 ) κ F 2 ( p 0 0 ) 2 g μ ν + 4 κ F 2 ( p 0 0 ) 2 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 l μ F p 0 0 δ 0 μ l ν F p 0 0 δ 0 ν 2 κ ϕ ( p 0 ) F 4 ( p 0 0 ) 4 δ 0 μ δ 0 ν .
To find the inverse metric g ˜ μ ν Finsler one can now apply Lemma 2. This allows to rewrite g ˜ μ ν as
g ˜ μ ν = A g μ ν + 1 A C μ C ν 2 κ ϕ ( p 0 ) F 4 ( p 0 0 ) 4 A δ 0 μ δ 0 ν ,
where
A = ϕ 2 ( p 0 ) + 2 ϕ ( p 0 ) κ F 2 ( p 0 0 ) 2 , B = 2 ϕ ( p 0 ) + κ F 2 ( p 0 0 ) 2 , C μ = 4 κ B F 2 ( p 0 0 ) 2 l μ F p 0 0 δ 0 μ .
Assuming that
m μ ν = g μ ν + λ C μ C ν with λ : = 1 / A ,
m μ ν = g μ ν λ 1 + λ C 2 C μ C ν ,
where C μ : = g μ ν C ν . C 2 can be calculated as C 2 = C μ C μ ,
C 2 = 4 κ B F 2 ( p 0 0 ) 2 l μ F p 0 0 g 0 μ l μ F p 0 0 δ 0 μ = 4 κ B F 2 ( p 0 0 ) 2 1 F p 0 0 l 0 F p 0 0 l 0 + F 2 ( p 0 0 ) 2 g 00 .
Since l 0 = p 0 0 / F , then C 2 reads
C 2 = 4 κ B F 2 ( p 0 0 ) 2 F 2 g 00 ( p 0 0 ) 2 ( p 0 0 ) 2 .
From the Lemma 2, one gets
m μ ν = g μ ν 4 κ B F 2 ( p 0 0 ) 4 A ( p 0 0 ) 4 + 4 κ B F 2 F 2 g 00 ( p 0 0 ) 2 l μ F p 0 0 g 0 μ l ν F p 0 0 g 0 ν .
Now m μ ν can be simplified as
m μ ν = g μ ν Θ · l μ F p 0 0 g 0 μ l ν F p 0 0 g 0 ν ,
where
Θ : = 4 κ B ( p 0 0 ) 2 F 2 A ( p 0 0 ) 4 + 4 κ B F 2 F 2 g 00 ( p 0 0 ) 2 .
When applying the Lemma 2, the derivation of the inverse metric g ˜ μ ν leads to
m ¯ μ ν = m μ ν + λ ¯ C ¯ μ C ¯ ν ,
where λ ¯ = 2 κ ϕ ( p 0 ) F 4 ( p 0 0 ) 4 A . For C ¯ μ = δ 0 μ , one derives
C ¯ μ = m μ ν = g μ ν Θ · l μ F p 0 0 g 0 μ l ν F p 0 0 g 0 ν δ 0 ν ,
which can be simplified as
C ¯ μ = g 0 μ Θ · ( p 0 0 ) 2 F 2 g 00 p 0 0 F l μ F p 0 0 g 0 μ .
Now, let us calculate C ¯ 2 = C ¯ μ C ¯ μ :
C ¯ 2 = g 0 μ Θ · p 0 2 F 2 g 00 F p 0 0 l μ F p 0 0 g 0 μ δ 0 μ = g 00 Θ · ( p 0 0 ) 2 F 2 g 00 2 F 2 ( p 0 0 ) 2 .
Then, we construct g ˜ μ ν :
g ˜ μ ν = 1 A g μ ν Θ · l μ F p 0 0 g 0 μ l ν F p 0 0 g 0 ν λ ¯ 1 + λ ¯ c ¯ 2 g 0 μ Θ · ( ( p 0 0 ) 2 F 2 g 00 ) p 0 0 F l μ F p 0 0 g 0 μ g 0 ν Θ · ( ( p 0 0 ) 2 F 2 g 00 ) p 0 0 F l ν F p 0 0 g 0 ν = 1 A g μ ν Θ A 1 + λ ¯ Θ · ( ( p 0 0 ) 2 F 2 g 00 ) 2 ( p 0 0 ) 2 F 2 ( 1 + λ ¯ c ¯ 2 ) l μ l ν + Θ A p 0 0 F F 2 + λ ¯ ( p 0 0 ) 2 F 2 g 00 1 + λ ¯ c ¯ 2 1 + Θ · ( p 0 0 ) 2 F 2 g 00 ( p 0 0 ) 2 l μ g 0 ν + l ν g 0 μ Θ F 2 ( p 0 0 ) 2 + λ ¯ 1 + λ ¯ c ¯ 2 1 + Θ · ( p 0 0 ) 2 F 2 g 00 ( p 0 0 ) 2 2 + Θ · ( p 0 0 ) 2 F 2 g 00 ( p 0 0 ) 2 g 0 μ g 0 ν = 1 A g μ ν + 4 κ F 2 F 2 κ + 2 ( p 0 0 ) 2 ϕ ( p 0 ) 2 F 4 g 00 κ 2 F 2 κ ( p 0 0 ) 2 ( p 0 0 ) 4 ϕ ( p 0 ) ( p 0 0 ) 2 ϕ 3 ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × l μ l ν + 4 κ p 0 0 F 3 F 2 κ + 2 ( p 0 0 ) 2 ϕ ( p 0 ) ϕ 2 ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × l μ g 0 ν + l ν g 0 μ + 6 F 4 κ ( p 0 0 ) 2 ϕ ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × g 0 μ g 0 ν .
It is essential to resolve the distinct mathematical challenge that arises when trying to express the last three lines in terms of g μ ν .
Both g ˜ μ ν and its inverse tensor g ˜ μ ν are metrics in arbitrary higher-dimensions. The transformation of these tensors into fundamental metric tensors is discussed in Section 5 just below.

5. Quantized Metric Tensor in Pseudo-Riemannian Geometry

When examining a relativistic particle with mass m moving through a background spacetime, the following action functional represents the dynamics of the corresponding quantum field:
S [ p ˙ 0 μ ( τ ) ] = m 0 1 g μ ν p 0 μ p 0 ν d τ ,
where τ is a parametrization in a single coordinate for the worldline of a point particle on a one-dimensional manifold. The proper time τ along the world line is determined by the equation d τ 2 = g μ ν p 0 μ p 0 ν = d s 2 . The effects of quantum gravity result in the identification of a zero-point length in spacetime [64,65]. Under the duality transformation d s l P 2 / d s , the invariance of the path integral amplitude describes the way in which quantum gravity transforms the spacetime interval from ( x ( i ) y ( f ) ) 2 to ( x ( i ) y ( f ) ) 2 + l P 2 , where x ( i ) = ( x 0 ( i ) μ , p 0 ( i ) ν ) and y ( f ) = ( x 0 ( f ) μ , p 0 ( f ) ν ) are initial and final locations [64]. The application of the proposed quantization approach to predict the zero-point length can be carried out elsewhere. Actually, GUP and the Duality Principle of the zero-point length of spacetime are what Equation (46) shows, along with the zero-point length of spacetime and duality.
Let us now focus on the line element and metric tensor on the higher-dimensional manifold. As elaborated in the above Sections, the higher-dimensional metric, such as a Finster-type metric, can be derived from the Hessian of the corresponding structure (14),
g ˜ μ ν Finsler = 1 2 2 p 0 μ p 0 ν ϕ 2 ( p 0 ) F 2 ( x 0 μ , p 0 ν ) ,
where F 2 ( x 0 μ , p 0 ν ) can express any higher-dimensional metric, for instance, Klein metric [17]. The Euler theorem, as is briefly reviewed in Appendix C, describes a process for connecting the resulting metric to the structure F 2 ( x 0 μ , p 0 ν ) , which—for instance—can be given by Equation (13). Equation (47) can be reexpressed as
ϕ 2 ( p 0 ) F 2 ( x 0 μ , p 0 ν ) = g ˜ μ ν Finsler p 0 μ p 0 ν .
The Hessian matrix considers the derivative with regard to a particular variable as if the function is exclusively dependent on that variable. The proper mathematical derivative is defined by these two considerations. It is apparent that Equation (48) is related to the quadratic condition governing the line element as described in GR and therefore any attempts to correlate the corresponding line element with the one utilized in spacetime geometry neglect the significant advantages that phase space geometry provides.
Now, let us focus to the derivation associated with the proposed parametrization ζ , which is characterized by a set of embedded four-coordinates. It is assumed that these coordinates effectively represent both geometries along with their corresponding metric tensor. The modified line element related to the higher-dimensional manifold
d s ˜ 2 Finsler = g ˜ μ ν Finsler d x μ ( ζ ) d x ν ( ζ ) .
The line element obtained by the quantized fundamental metric tensor, g ˜ α β , is
d s ˜ 2 Riemann = g ˜ μ ν Riemann d ζ μ d ζ ν .
The introducing of higher-dimensional coordinates x μ = ( x 0 μ , ϕ ( p 0 ) p 0 μ ) extends the conventional spacetime framework incorporating the momentum space. On the other hand, the higher-dimensional coordinates permit the use of partial derivatives and quantum mechanical revisions such as RGUP. With the proposed parametrization, the modified line element associated with the T M metric is formulated in the corresponding bases as d x μ ( ζ ) and d x ν ( ζ ) in Equation (49). The line element is accordingly modified by the quantized fundamental tensor g ˜ μ ν and the bases d ζ μ and d ζ ν in Equation (50), so that
g ˜ μ ν Riemann = g ˜ μ ν Finsler x μ ( ζ ) ζ μ x ν ( ζ ) ζ ν .
At the proposed generic point, which is characterized by four-dimensional spacetime x 0 and four-dimensional momentum space p 0 , at which the proposed parameterization is valid, we replace
g ˜ μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ , ϕ ( p 0 ) p 0 μ ζ μ x 0 ν ζ ν , ϕ ( p 0 ) p 0 ν ζ ν = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν ,
where the normalization factor F = m A represents the maximal proper force, where A stands for the maximum proper acceleration given by A = c 7 / G . In this context, F 2 represents the normalization applied to the square of the resultant proper force. In the derivation of Equation (52), it was assumed that the products of mixed derivatives entirely vanish.
Several observations can be made regarding the line elements (49) and (50). First, the measurements associated with these line elements are no longer characterized by precision or continuity. Second, the proposed quantization introduces corresponding modifications to both line elements. Last, equating these two line elements allows for the assessment of their modifications, thereby promoting the relationship between manifolds of four dimensions and those of higher dimensions.
The quantum geometry predicts that the world lines correspond to physical particles with a maximum acceleration [21]. Caianiello proposed that the simplest theoretical framework with the maximal proper acceleration must satisfy physical invariance [66,67,68]. This is not the line element of classical four-dimensional spacetime, but rather an eight-dimensional phase space, for instance. The quantized four-dimensional first-order fundamental tensor can be derived from Equation (52).
Lemma 3.
By substituting Equation (26) into Equation (52), one obtains
g ˜ μ ν Riemann = 1 + ϕ 4 ( p 0 ) + ϕ 3 ( p 0 ) 2 κ ( p 0 0 ) 2 F 2 p ˙ 0 μ p ˙ 0 ν F 2 g μ ν Riemann + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν d μ ν Finsler ,
where ϕ 2 ( p 0 ) = 1 + 2 β p 0 ρ p 0 ρ , ϕ 3 ( p 0 ) = 1 + 3 β p 0 ρ p 0 ρ , ϕ 4 ( p 0 ) = 1 + 4 β p 0 ρ p 0 ρ , and the symmetric tensor d μ ν Finsler is given as
d μ ν Finsler = 4 κ F 2 ( p 0 0 ) 2 ( p 0 0 ) 2 3 ϕ ( p 0 ) 1 l μ l σ g σ μ F p 0 0 3 ϕ ( p 0 ) 1 l σ δ 0 ν g σ μ + δ 0 μ g σ ν + 2 F 2 5 ϕ ( p 0 ) 2 δ 0 μ δ 0 σ g σ ν .
Proof. 
Equation (26) represents the tensor g ˜ μ ν Finsler along with its associated metrics, g μ ν Finsler and d μ ν Finsler . The former metric is characterized by four dimensions, whereas the latter is defined in eight dimensions. By substituting Equation (26) into Equation (52), one gets
g ˜ μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 g ˜ μ ν Finsler p 0 μ ζ μ p 0 ν ζ ν , = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) ( p 0 0 ) 2 F 2 g μ ν Riemann + d μ ν Finsler p 0 μ ζ μ p 0 ν ζ ν .
The first term can be substituted by the line element in corresponding geometry:
g μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν .
At relatively small β , the function ϕ ( p 0 ) can be redefined
g ˜ μ ν Riemann = 1 + ϕ 4 ( p 0 ) + ϕ 3 ( p 0 ) 2 κ ( p 0 0 ) 2 F 2 p ˙ 0 μ p ˙ 0 ν F 2 g μ ν Riemann + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν d μ ν Finsler ,
where p ˙ 0 μ = d p 0 μ / ζ μ and p ˙ 0 ν = d p 0 ν / ζ ν . The proper force squared, p ˙ 0 μ p ˙ 0 ν , is ad hoc normalized to the maximal proper force squared, F 2 , so that 0 p ˙ 0 μ p ˙ 0 ν / F 2 1 . The lower limit removes entirely the quantization contribution from the first line of Equation (57), while the upper limit maximizes it. Due to the complexity of the higher-dimensional metric d μ ν Finsler , the second line of Equation (57) is not assessed; the assessment can be undertaken elsewhere. □
By examining only the first line of Equation (57), it becomes apparent that the quantized fundamental tensor can be expressed as a conformal transformation of its unquantized form. Given that all values within the square brackets are certainly positive, one concludes that g ˜ μ ν Riemann > g μ ν Riemann . The conformal coefficient, represented by the square brackets in Equation (57), indicates that the proposed quantization perspective entirely retains the unquantized fundamental tensor, thereby affirming the foundations of conventional GR. An essential aspect of the proposed quantization approach is its proposed quantization, which broadens the applicability of GR. The quantized GR, as it is predicted by its fundamental tensor, is no longer restricted to its traditional large-scale framework; it also finds relevance at lower (quantum) scales, where the second term in the conformal coefficient in the square brackets in Equation (57) becomes highly significant.
Section 6 just below outlines the derivatives of quantized fundamental inverse tensor in pseudo-Riemann geometry.

6. Quantized Inverse Tensor in Pseudo-Riemann Geometry

In Section 5, it was highlighted that the use of higher-dimensional coordinates x a = ( x 0 μ , ϕ ( p 0 ) p 0 ν ) establishes a link between four-dimensional momentum space and the conventional four-dimensional spacetime:
g ˜ μ ν Riemann = g ˜ μ ν Finsler x μ ζ μ x ν ζ ν .
In phase space geometry, the modification proposed by the RGUP leads to
g ˜ μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν .
Lemma 4.
By substituting Equation (45) into Equation (59), we obtain
g ˜ μ ν Riemann = 1 + 1 1 + ϕ 1 ( p 0 ) ϕ 2 ( p 0 ) 2 κ ( p 0 0 ) 2 F 2 p ˙ 0 μ p ˙ 0 ν F 2 g μ ν Riemann + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν d μ ν Finsler ,
where ϕ 1 ( p 0 ) = 1 + β p 0 ρ p 0 ρ and the symmetric tensor d μ ν Finsler is given as
d μ ν Finsler = 4 κ F 2 F 2 κ + 2 ( p 0 0 ) 2 ϕ ( p 0 ) 2 F 4 g 00 κ 2 F 2 κ ( p 0 0 ) 2 ( p 0 0 ) 4 ϕ ( p 0 ) ( p 0 0 ) 2 ϕ 3 ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × l μ l ν + 4 κ p 0 0 F 3 F 2 κ + 2 ( p 0 0 ) 2 ϕ ( p 0 ) ϕ 2 ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × l μ g 0 ν + l ν g 0 μ + 6 F 4 κ ( p 0 0 ) 2 ϕ ( p 0 ) 6 F 4 g 00 κ 4 F 2 κ ( p 0 0 ) 2 + ( p 0 0 ) 4 ϕ ( p 0 ) 2 F 2 κ + ( p 0 0 ) 2 ϕ ( p 0 ) × g 0 μ g 0 ν .
Proof. 
The Equation (45) expresses the tensor g ˜ μ ν Finsler along with its associated metrics, g μ ν Finsler and d μ ν Finsler . Substituting Equation (45) into Equation (59) results in
g ˜ μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 g ˜ μ ν Finsler p 0 μ ζ μ p 0 ν ζ ν , = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν + ϕ 2 ( p 0 ) F 2 1 ϕ 2 ( p 0 ) + 2 κ ϕ ( p 0 ) ( p 0 0 ) 2 F 2 g μ ν Riemann + d μ ν Finsler p 0 μ ζ μ p 0 ν ζ ν .
The first term can be substituted by the line element in corresponding geometry:
g μ ν Riemann = g ˜ μ ν Finsler x 0 μ ζ μ x 0 ν ζ ν
At comparably small β , the function ϕ ( p 0 ) can be approximated. Hence, the metric g μ ν Riemann can be redefined as
g ˜ μ ν Riemann = 1 + 1 1 + ϕ 1 ( p 0 ) ϕ 2 ( p 0 ) 2 κ ( p 0 0 ) 2 F 2 p ˙ 0 μ p ˙ 0 ν F 2 g μ ν Riemann + ϕ 2 ( p 0 ) F 2 p 0 μ ζ μ p 0 ν ζ ν d μ ν Finsler .
The complexity of the higher-dimensional metric d μ ν Finsler (61) results in the second line of Equation (64) not being evaluated. This evaluation can be performed elsewhere. □

7. Discussion and Conclusions

In this paper, we discussed the mathematical methodology for quantizing the fundamental metric tensor. This strategy is coherent from a mathematical point of view. The concept of geometric quantization presents a physical alternative to the numerous theories introduced in the last century that aimed to integrate quantum mechanics with GR and formulate a consistent theory of quantum gravity.
In this context, let us refer to Edward Witten’s paper [69], which thoroughly discusses the challenges related to the quantization of gravity. In particular, the perspective on matters concerning differential geometry is limited to GR at the “trivial” (2+1) dimensions, which refers to flat spacetime. Witten’s quantization process involves “constructing the classical phase space, defining Poisson brackets on this space, and then interpreting the functions on phase space as quantum mechanical operators” [69]. Furthermore [69], “the vierbein formalism of GR makes GR remarkably similar to a gauge theory.” This specific formalism makes GR appear similar to a gauge theory. In this regard, it was concluded that achieving a “quantum gravity theory” is a significant challenge [69]. The vierbein formalism could be studied elsewhere. The approach proposed in the current paper is based on the quite straightforward consideration that the metric tensor defines the geometry. This is actually the main characteristic of conventional GR. The results obtained by us based on this are highly promising [9,10,12,13,14,15,16].
The foundation of GR is established on specific postulates and assumptions that are integral to the theory. When attempting to quantize GR, it becomes crucial to keep all postulates and assumptions preserved, as the unquantized version of GR has reliably passed all assessments conducted over the past century. Pseudo-Riemann geometry belongs to the assumptions and assumptions of the conventional GR. The quantization proposal made here focuses on the fundamental metric tensor, the quantity that encodes the entire geometric properties. The mathematical framework of GR certainly relies on this metric tensor. The construction of the affine connection is based on the fundamental tensor. Utilizing the affine connection allows for the derivation of the Riemann curvature tensor. Subsequently, the Riemann curvature tensor can be contracted to yield the Ricci curvature tensor, which further leads to the Ricci tensor. This sequence of derivations ultimately results in the formulation of the Einstein tensor. In the context of the Einstein field equations, Appendix E, the right-hand side of Equation (A16) is represented by the stress–energy tensor, which can be obtained from the Einstein–Hilbert action. It is worth noting that the Einstein–Hilbert action is also influenced by the fundamental tensor.
In this paper, we demonstrated the derivations of the quantized fundamental tensor, starting with the extension of the four-dimensional pseudo-Riemann geometry to the pseudo-Finsler geometry. We exclusively considered the homogeneity property of the pseudo-Finsler structure, presuming the kinematics of a free-falling quantum particle with a positive mass. The Finsler structure is subsequently translated into a Hamilton structure, where momentum replaces velocity. We also introduced a secondary physical component inspired by QM. It is crucial to generalize QM to accommodate the influences of relativistic gravitational fields. Thus, the momentum parameter is modified through the ϕ ( p 0 ) function. Integration of momentum modification into the Hamilton structure allows for the straightforward deduction of the corresponding metric once the Hamilton structure is prepared including quantum mechanical ingredients.
The derivatives are valid in arbitrary higher dimensions. Particularly, the derivatives are not limited to the Hamiltonian space, which typically encompasses spacetime and momentum space. This principle is also applied for the inverse metric, as both can be represented as conformal transformations of their corresponding metrics. Given the context of GR within pseudo-Riemann geometry, the quantized metric in higher dimensions is shown needed to be translated to represent the fundamental tensor in four dimensions.
The main point of this paper was to present the “derivation of the quantized metric tensor”. We have already considered the “geometric quantization” such as in our earlier studies [9,10,12,13,14,15,16]. Geometric quantization, as we impose it there, is a mathematical method for defining a quantum theory that corresponds to a specific classical theory, GR. The method we applied also relates to symplectic geometry, which is closely related to the Weyl quantization, one of the first attempts for natural quantization of GR. The Weyl’s approach connects quantum mechanical observables, which are self-adjoint operators on a Hilbert space, with a real-valued function in classical phase space. We focused on position and momentum and their mapping to the Heisenberg group generators in particular. Groenewold’s product of such a pair of observables serves as the foundation for our method [70,71]. In this regard, we emphasize that the approach we employed, being closely related to Weyl quantization, differs from Kirillov’s “geometric quantization,” known as Kirillov–Kostant–Souriau two-form which considers coadjoint orbit and calls for prequantization, polarization, and then metaplectic correction [72].
In summary, we have successfully derived the higher-dimensional quantized metric. Additionally, despite encountering significant mathematical challenges, we found a way to obtain the inverse metric. Furthermore, we have derived the fundamental tensor in four dimensions. Finally, it is now possible to construct Einstein field equations incorporating quantum revisions so that GR can be applied in both relatively large and relatively short quantum scales. The quantum mechanically revisited Einstein field equations are described in further detail in Appendix E.
As an outlook, we believe that a comprehensive study on how and why vierbein formalism appears in the proposed quantization to be conducted elsewhere.

Author Contributions

Conceptualization, methodology, writing—original draft prparation, visualization, validation, supervision, A.N.T.; formal analysis, investigation, A.N.T. and S.G.E.; writing—review and editing, S.S. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors declare that this research received no specific grants from any funding agency in the public, commercial, or not-for-profit sectors.

Data Availability Statement

All data generated or analyzed during this study are included in this published article. The data used to support the findings of this study are included within the published article and properly cited! All of the material is owned by the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Relativistic Generalized Uncertainty Principle (RGUP)

According to Refs. [4,14,15,16], the four-dimensional physical position and momentum coordinates of a test particle in curved spacetime are not canonically conjugate variables. To relax the energy regime of the possible implications of the GUP approach on relativistic quantum mechanics, quantum field theory, and quantum gravity, the physical position and momentum coordinates of a test particle are conjectured to be given in terms of their auxiliary four-vectors x 0 μ and p 0 ν , i.e., x μ x 0 μ , p 0 μ and p μ x 0 μ , p 0 μ . Under the general static metric g μ ν , x ^ 0 μ , p ^ 0 ν = i g μ ν and p ^ 0 μ = i / x ^ 0 μ [4,73,74]. The GUP corrections allow us to distinguish between the temporal and spatial components:
(A1) x μ = ( x 0 , x i ) , (A2) p μ = ( p 0 , f ( p j ) ) .
With Snyder’s algebra [75], we assume that [74]
(A3) x μ = ( c t , x 0 i ) , (A4) p μ = E / c , p 0 i 1 + β p 0 ρ p 0 ρ .
The stress–energy tensor describing the test particle in curved spacetime is used to calculate x 0 0 = c t and p 0 0 = E / c .
The relativistic generalized noncommutation relation [76], which is Lorentz covariant and preserves the linear law of addition of moments, is proposed as [4,74,77]
x ^ μ , p ^ ν = i 1 + β 1 p ρ p ρ g μ ν + 2 β 2 p μ p ν ,
where β 1 and β 2 are distinct RGUP parameters. Assuming Snyder’s algebra [75] and an isotropic property of spacetime, we can revisit the physical space and momentum coordinates as
(A6) x μ = x 0 μ , (A7) p μ = 1 + β p 0 ρ p 0 ρ p 0 μ .
The position and momentum noncommutation relation, Equation (A5), determines the uncertainties of their measurements based on the Robertson uncertainty relation [78] and the Schrödinger uncertainty relation [79]
Δ x μ Δ p ν 1 2 x ^ μ , p ^ ν .
As a result, the uncertainties on the position and momentum measurements in curved spacetime could be estimated as
Δ x μ Δ p ν i 2 g μ ν + β 1 p 2 β 2 Δ p ν 2 + β 2 p ν 2 .
Here, the Einstein summation is used to address uncertainty in the momentum operator Δ p .

Appendix B. Function ϕ(p0) and the Curvature Properties of Randers Metric

Appendix A discussed how the function ϕ ( p 0 ) affects the gravitation of the Heisenberg uncertainty principle. In this Appendix, we introduce how ϕ ( p 0 ) = 1 + β p 0 ρ p 0 ρ is capable of quantizing GR while retaining the special curvature properties of Randers’ metric [80]
ϕ ( p 0 ) = 1 + β p 0 ρ p 0 ρ = 1 + β F 2 ( x 0 ρ , p 0 ρ ) ,
where β is the RGUP constant and p 0 is the auxiliary four-momentum with either a covariant or contravariant dummy index ρ .
For simplicity, we denote F ( x 0 ρ , p 0 ρ ) as F. Multiplying both sides of Equation (A10) by F and assuming F ¯ = ϕ ( p 0 ) F and β ¯ = β F 3 , one finds:
F ¯ = ( 1 + β F 2 ) F = F + β ¯ .
The Randers metric tensor can be obtained using Equation (A11). This is subject to the conditions on the right-hand side of Equation (A11). The first term, F, should be the Riemann metric: F 2 = g μ ν y μ y ν , where y T p M . At p, the second term, β ¯ , must take the 1-form. β ¯ = b i ( x ) y i , where | | β ¯ | | p : = sup y T p M ( β ¯ ( y ) / F ( y ) ) is the expression to use. We discovered that β ¯ is 1-homogeneous in p ^ 0 . To fulfill Randers’ metric and thus the unification of gravity and electromagnetism, β ¯ has to be linear in y i . With these two conditions, one draws the conclusion that the function ϕ ( p 0 ) preserves the special curvature properties of Randers metric and plays the role of unifying gravity and electromagnetism [80], in addition to its crucial roles in gravitating QM and quantizing GR.

Appendix C. Euler Theorem and Homogeneous Finsler– Hamilton Function

A function F : V R is positively homogeneous of degree n R and F : T M { 0 } is a smooth function if F ( λ v ) = λ n F ( v ) for all v = v μ · e μ V and λ R + . The Finsler–Hamilton function F is positively homogeneous of degree one, because F ( λ v ) = λ F ( v ) for all v V and λ R + . Then, for 1-homogeneous F in p 0 ν , i.e., n = 1 , and then, F / p 0 ν becomes homogeneous of degree zero in p 0 ν with
F p 0 ν v v ν = F .
To realize that Equation (A12) represents Euler theorem, let us assume a smooth curve c : ( a , b ) M is stationary for c a b F c ^ ( t ) d t , where ∘ stands for group product and t is a proper parametrization, if and only if at each value of t exist local coordinates around c ^ ( t ) such that
F x 0 μ c ^ d d t F p 0 ν c ^ = 0 .
For the ( x 0 μ , p 0 ν ) , ( x ˜ 0 μ , p ˜ 0 ν ) local coordinates around p 0 ν T x M , where x ˜ 0 μ = x ˜ 0 μ ( x ) , p ˜ 0 ν = p ˜ 0 ν ( x , y ) , and p 0 ν are coordinates in the / x 0 μ basis, then / x 0 μ | y , / p 0 ν | y , / x ˜ 0 μ | y , / p ˜ 0 ν | y T ( T M { 0 } ) satisfy the transformation rules
(A14) x 0 μ y = x ˜ 0 γ x 0 μ x ˜ 0 γ y + 2 x ˜ 0 γ x 0 μ ( x 0 μ ) 1 ( p 0 ν ) 1 p ˜ 0 γ y , (A15) p 0 ν | y = x ˜ 0 γ x 0 μ p ˜ 0 γ y ,
which can be proved from p ˜ 0 ν = x 0 ν x 0 γ p 0 γ . Then, by implementing the transformation rules (A14) and (A15) to Equation (13), one arrives at Equation (48).

Appendix D. Remarks on Proposed Quantization

The realization that the origin of quantization is intrinsically linked to the discretization of the energy spectrum implies that quantization must involve discretization. This realization motivated us to commence with the discretization of the Finsler (or Hamiltonian-type) structure. Thus, we have successfully quantized a free-falling particle, which originally necessitated a continuous energy spectrum.
As mentioned in Section 1, the primary objective is to reconcile the principles of QM and GR. The methodology outlined in this study involves several key steps. Initially, it is essential to incorporate gravitational effects into the fundamental framework of QM. Subsequently, the concept of quantum geometry must be introduced, which entails the addition of curvatures to replicate the proposed quantization of Riemann spacetime. To facilitate this, we have extended the four-dimensional Riemann manifold to a Finsler manifold. In order to revert the resulting quantized metric tensor to pseudo-Riemann geometry, we utilized the characteristic that the ratio of lengths between any two colinear vectors is independent of the underlying metric tensor. The fundamental aspects of the approach proposed are derived from this essential premise. It is crucial to note that our approach is predicated on the assumption that the measurements of both line elements are accurate and deterministic. Therefore, achieving a thorough quantization is reliant on the presence of a quantum line element and the consideration of uncertain measurements, as discussed in Appendix C.
To express the line element in a quantum terms, it is essential to integrate the probability distribution with the metric tensor and the 1-form d x μ . Alternatively, one might propose the introduction of noncommutation relations. It is worth noting that neither g μ ν nor d x μ , including their generalized forms, possess a straightforward noncommutation interpretation [81]. Conversely, the establishment of a noncommutative differential calculus [82,83] alongside a noncommutative metric tensor [84] facilitates the integration of these approaches into a coherent definition of a noncommutative measure for the line element [85]. Moreover, the revised approach to relativistic kinematics, characterized by pseudo-Finslerian geometry, provides insights into a potential vacuum state of quantum gravity at low energies [86]. The adoption of a Finslerian length element, along with RGUP discretization, leads to a quantization of spacetime. A thorough examination of this topic is warranted in future research.
Aside from any nonconservative manifestations of the limitations inherent in the line element measure, we have succeeded in deriving a quantized metric tensor, as shown in Equation (59). This derivation is characterized as a qualitative and approximate estimation. In this context, we nonconservatively underscore that the g a b term in Equation (59) is largely truncated. Therefore, some quantum mechanical aspects may not be sufficiently accounted for. This form of approximation is presently unavoidable, despite the inadequately justified physical and mathematical constraints it entails.

Appendix E. Einstein Field Equations

The Einstein field equations are expressed as a second-order tensor. In traditional pseudo-Riemannian geometry, this is defined by
G μ ν = R μ ν 1 2 g μ ν R ,
where R μ ν is the Ricci tensor and R = g μ ν R μ ν the (Ricci) scalar curvature. All curvature quantities derive from the metric g μ ν via the Levi-Civita connection Γ μ ν ρ . These quantities are all derived from the quantized Christoffel symbols Γ ˜ μ ν ρ = Γ μ ν ρ + δ Γ μ ν ρ . The latter enables the construction of the quantized Riemann curvature tensor
R ˜ σ μ ν ρ = μ Γ ˜ ν σ ρ ν Γ ˜ μ σ ρ + Γ ˜ μ λ ρ Γ ˜ ν σ λ Γ ˜ ν λ ρ Γ ˜ μ σ λ ,
which transforms as a ( 1 , 3 ) tensor under diffeomorphisms.
First, by using the perturbative expansion and only keeping first-order corrections in δ Γ , one obtains
R ˜ σ μ ν ρ = R σ μ ν ρ + μ δ Γ ν σ ρ ν δ Γ μ σ ρ + δ Γ μ λ ρ Γ ν σ λ + Γ μ λ ρ δ Γ ν σ λ δ Γ ν λ ρ Γ μ σ λ Γ ν λ ρ δ Γ μ σ λ + O ( δ Γ 2 ) .
Secondly, the quantized Ricci tensor is defined by contracting the first and third indices with the quantized inverse metric g ˜ μ ν .
R ˜ σ ν = R ˜ σ λ ν λ ,
so that
R ˜ σ ν = R σ ν + λ δ Γ ν σ λ ν δ Γ λ σ λ + δ Γ λ μ λ Γ ν σ μ + Γ λ μ λ δ Γ ν σ μ δ Γ ν μ λ Γ λ σ μ Γ ν μ λ δ Γ λ σ μ + O ( δ Γ 2 ) .
The quantized Ricci scalar is then derived through an additional contraction with g ˜ σ ν :
R ˜ = g ˜ σ ν R ˜ σ ν = R + g σ ν δ R σ ν Ξ ( p 0 ) p ˙ 0 σ p ˙ 0 ν F 2 R σ ν + O ( δ 2 ) ,
where
Ξ ( p 0 ) = 1 + 4 β p 0 ρ p 0 ρ + 1 + 3 β p 0 ρ p 0 ρ 2 κ ( p 0 0 ) 2 K 2 .
with ρ as a dummy index and K representing the Klein metric which belongs to the simplest Finsler metrics. Finally, the quantized Einstein tensor is defined by
G ˜ σ ν = R ˜ σ ν 1 2 g ˜ σ ν R ˜ ,
which, upon expanding to first order in δ g and δ R , becomes
G ˜ σ ν = G σ ν + δ R σ ν 1 2 g σ ν δ R 1 2 δ g σ ν R + O ( δ 2 ) , = G σ ν + λ δ Γ ν σ λ ν δ Γ λ σ λ 1 2 g σ ν g α β δ R α β + 1 2 Ξ ( p 0 ) p ˙ 0 α p ˙ 0 β F 2 g σ ν R α β 1 2 Ξ ( p 0 ) p ˙ 0 σ p ˙ 0 ν F 2 R .
Regarding the physical implications, one observes that the momentum derivatives p ˙ 0 μ and the maximal force scale F become dynamic during periods of high curvature, such as inflation, leading to a spacetime-dependent Ξ ( p 0 ) . Then, in the semi-classical Einstein equations, δ G σ ν serves as a quantum mechanically corrected source term
G ˜ σ ν = 8 π G T σ ν .
The effective equation of state in the early universe appears to be modified by this finding, as well as the dispersion of gravitational waves and the expansion of perturbations. Thorough dimensional and tensorial bookkeeping, as demonstrated in the examples in this Appendix, allows for the consistent integration of these modifications into phenomenological and numerical studies of quantum gravitational effects in cosmology.

References

  1. Einstein, A.; Podolsky, B.; Rosen, N. Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 1935, 47, 777–780. [Google Scholar] [CrossRef]
  2. Macias, A.; Camacho, A. On the incompatibility between quantum theory and general relativity. Phys. Lett. B 2008, 663, 99–102. [Google Scholar] [CrossRef]
  3. Tawfik, A.N.; Alshehri, A. Relativistic generalized uncertainty principle for a test particle in four-dimensional spacetime. Mod. Phys. Lett. A 2024, 39, 2450079. [Google Scholar] [CrossRef]
  4. Todorinov, V. Relativistic Generalized Uncertainty Principle and Its Implications. Ph.D. Thesis, University of Lethbridge, Lethbridge, AB, Canada, 2020. [Google Scholar] [CrossRef]
  5. Tkachuk, V.M. Galilean and Lorentz transformations in a space with generalized uncertainty principle. Found. Phys. 2016, 46, 1666–1679. [Google Scholar] [CrossRef]
  6. Diosi, L. A universal master equation for the gravitational violation of quantum mechanics. Phys. Lett. A 1987, 120, 377–381. [Google Scholar] [CrossRef]
  7. Penrose, R. On gravity’s role in quantum state reduction. Gen. Relat. Grav. 1996, 28, 581–600. [Google Scholar] [CrossRef]
  8. Berglund, P.; Hübsch, T.; Mattingly, D.; Minic, D. Gravitizing the quantum. Int. J. Mod. Phys. D 2022, 31, 2242024. [Google Scholar] [CrossRef]
  9. Tawfik, A.N.; Alshehri, A.A.; Pasqua, A. Expansion evolution of nonhomogeneous metric with quantum-mechanically revisited fundamental metric tensor. Nucl. Phys. B 2025, 1015, 116893. [Google Scholar] [CrossRef]
  10. Tawfik, A.N.; Dabash, T.F.; Amer, T.S.; Shaker, M.O. Einstein–Gilbert–Straus solution of Einstein field equations: Timelike geodesic congruence with conventional and quantized fundamental metric tensor. Nucl. Phys. B 2025, 1014, 116866. [Google Scholar] [CrossRef]
  11. Tawfik, A.N.; Farouk, F.T.; Tarabia, F.S.; Maher, M. Quantum-induced revisiting space–time curvature in relativistic regime. Int. J. Mod. Phys. A 2024, 39, 2443016. [Google Scholar] [CrossRef]
  12. Tawfik, A.N.; Pasqua, A.; Waqas, M.; Alshehri, A.A.; Haldar, P.K. Quantum geometric perspective on the origin of quantum-conditioned curvatures. Class. Quant. Grav. 2024, 41, 195018. [Google Scholar] [CrossRef]
  13. Tawfik, A.; Dabash, T.F. Timelike geodesic congruence in the simplest solutions of general relativity with quantum-improved metric tensor. Int. J. Mod. Phys. D 2023, 32, 2350097. [Google Scholar] [CrossRef]
  14. Farouk, F.T.; Tawfik, A.N.; Tarabia, F.S.; Maher, M. On possible minimal length deformation of metric tensor, Levi-Civita connection, and the Riemann curvature tensor. Physics 2023, 5, 983–1002. [Google Scholar] [CrossRef]
  15. Tawfik, A.N.; Dabash, T.F. Born reciprocity and discretized Finsler structure: An approach to quantize GR curvature tensors on three-sphere. Int. J. Mod. Phys. D 2023, 32, 2350068. [Google Scholar] [CrossRef]
  16. Tawfik, A.N.; Dabash, T.F. Born reciprocity and relativistic generalized uncertainty principle in Finsler structure: Fundamental tensor in discretized curved spacetime. Int. J. Mod. Phys. D 2023, 32, 2350060. [Google Scholar] [CrossRef]
  17. Mo, X. An Introduction to Finsler Geometry; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2006. [Google Scholar] [CrossRef]
  18. Sasaki, S. On the differential geometry of tangent bundles of Riemannian manifolds. Tohoku Math. J. 1958, 10, 338–354. [Google Scholar] [CrossRef]
  19. Dombrowski, P. On the geometry of the tangent bundle. J. Reine Angew. Math. 1962, 210, 73–88. [Google Scholar] [CrossRef]
  20. Gudmundsson, S.; Kappos, E. On the geometry of tangent bundles. Expo. Math. 2002, 20, 1–41. [Google Scholar] [CrossRef]
  21. Caianiello, E.R. Is there a maximal acceleration? Lett. Nuovo Cim. 1981, 32, 65–70. [Google Scholar] [CrossRef]
  22. Caianiello, E.R.; Feoli, A.; Gasperini, M.; Scarpetta, G. Quantum corrections to the space–time metric from geometric phase space quantization. Int. J. Theor. Phys. 1990, 29, 131–139. [Google Scholar] [CrossRef]
  23. Yampolski, A. The curvature of the Sasaki metric of tangent sphere bundles. J. Math. Sci. 1990, 48, 108–117. [Google Scholar] [CrossRef]
  24. Kowalski, O. Curvature of the induced Riemannian metric on the tangent bundle of a riemannian manifold. J. Reine Angew. Math. 1971, 250, 124–129. [Google Scholar] [CrossRef]
  25. Musso, E.; Tricerri, F. Riemannian metrics on tangent bundles. Ann. Mat. Pura Appl. 1988, 150, 1–19. [Google Scholar] [CrossRef]
  26. Raei, Z. Some properties of Sasaki metric on the tangent bundle of Finsler manifolds. J. Finsler Geom. Appl. 2021, 2, 23–42. [Google Scholar] [CrossRef]
  27. Aso, K. Notes on some properties of the sectional curvature of the tangent bundle. Yokohama Math. J. 1981, 29, 1–5. Available online: https://ynu.repo.nii.ac.jp/search?page=1&size=20&sort=-custom_sort&search_type=2&q=647 (accessed on 2 September 2025).
  28. Abbassi, M.; Sarih, M. On some hereditary properties of Riemannian g-natural metrics on tangent bundles of Riemannian manifolds. Differ. Geom. Appl. 2005, 22, 19–47. [Google Scholar] [CrossRef]
  29. Wu, B.Y. Some results on the geometry of tangent bundle of Finsler manifolds. Publ. Math. Debrecen 2007, 71, 185–193. [Google Scholar] [CrossRef]
  30. Asanov, G. Finsleroid-Finsler spaces of positive-definite and relativistic types. Rep. Math. Phys. 2006, 58, 275–300. [Google Scholar] [CrossRef]
  31. Mignani, R.; Scipioni, R. On the solutions of the Cartan equation in metric affine gravity. Gen. Relat. Grav. 2001, 33, 683–711. [Google Scholar] [CrossRef]
  32. Miron, R.; Hrimiuc, D.; Shimada, H.; Sabau, S. The Geometry of Hamilton and Lagrange Spaces; Springer: Dordrecht, The Netherlands, 2002. [Google Scholar] [CrossRef]
  33. Albuquerque, S.; Bezerra, V.B.; Lobo, I.P.; Macedo, G.; Morais, P.H.; Santos, L.C.N.; Varo, G. Quantum configuration and phase spaces: Finsler and Hamilton geometries. Physics 2023, 5, 90–115. [Google Scholar] [CrossRef]
  34. Miron, R.; Anastasiei, M. The Geometry of Lagrange Spaces: Theory and Applications; Springer: Dordrecht, The Netherlands, 1993. [Google Scholar] [CrossRef]
  35. Heisenberg, W. Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik. Z. Phys. 1927, 43, 172–198. [Google Scholar] [CrossRef]
  36. Heisenberg, W. The Actual Content of Quantum Theoretical Kinematics and Mechanics. NASA Technical Memorandum Report NASA-TM-77379; NASA: Washington, DC, USA, 1983. Available online: https://ntrs.nasa.gov/citations/19840008978 (accessed on 2 September 2025).
  37. Tawfik, A.N.; Diab, A.M. A review on the generalized uncertainty principle. Rep. Prog. Phys. 2015, 78, 126001. [Google Scholar] [CrossRef]
  38. Tawfik, A.; Diab, A. Generalized uncertainty principle: Approaches and applications. Int. J. Mod. Phys. 2014, 23, 1430025. [Google Scholar] [CrossRef]
  39. Galapon, E.A. Self-adjoint time operator is the rule for discrete semi-bounded Hamiltonians. Proc. R. Soc. Lond. A Math. Phys. Engin. Sci. 2002, 458, 2671–2689. [Google Scholar] [CrossRef]
  40. Farrales, R.A.E.; Galapon, E.A. Characteristic time operators as quantum clocks. Phys. Lett. A 2025, 532, 130192. [Google Scholar] [CrossRef]
  41. Pauli, W. General Principles of Quantum Mechanics; Springer: Berlin/Heidelberg, Germany, 1980. [Google Scholar] [CrossRef]
  42. Srinivas, M.D.; Vijayalakshmi, R. The ‘time of occurrence’ in quantum mechanics. Pramana 1981, 16, 173–199. [Google Scholar] [CrossRef]
  43. Busch, P.; Grabowski, M.; Lahti, P.J. Operational Quantum Physics; Springer: Berlin/Heidelberg, Germany, 1995. [Google Scholar]
  44. Kraus, K. Remark on the uncertainty between angle and angular momentum. Z. Phys. 1965, 188, 374–377. [Google Scholar] [CrossRef]
  45. Galapon, E.A. What could we have been missing while Pauli’s theorem was in force? In Time and Matter; Bigi, I.I., Faessler, M., Eds.; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2006; pp. 133–144. [Google Scholar] [CrossRef]
  46. Hiroshima, F.; Teranishi, N. Unbounded self-adjoint time operator for discrete semibounded Hamiltonians. arXiv 2024, arXiv:2405.02851. [Google Scholar] [CrossRef]
  47. Kokoulin, R. Self-adjoint time operator in a weighted energy space. arXiv 2025, arXiv:2504.17830. [Google Scholar] [CrossRef]
  48. Schmüdgen, K. On the Heisenberg commutation relation I. J. Funct. Anal. 1983, 50, 8–49. [Google Scholar] [CrossRef]
  49. Tawfik, A.N.; Magdy, H.; Ali, A.F. Lorentz invariance violation and generalized uncertainty principle. Phys. Part. Nucl. Lett. 2016, 13, 59–68. [Google Scholar] [CrossRef]
  50. Samar, M.I.; Tkachuk, V.M. Perturbation hydrogen-atom spectrum in a space with the Lorentz-covariant deformed algebra with minimal length. J. Phys. Stud. 2010, 14, 1001. [Google Scholar] [CrossRef]
  51. Quesne, C.; Tkachuk, V.M. Lorentz-covariant deformed algebra with minimal length. Czech. J. Phys. 2006, 56, 1269–1274. [Google Scholar] [CrossRef]
  52. Hussar, P.E.; Kim, Y.S.; Noz, M.E. Time energy uncertainty relation and Lorentz covariance. Am. J. Phys. 1985, 53, 142–147. [Google Scholar] [CrossRef]
  53. Camacho, A. Generalized uncertainty principle and deformed dispersion relation induced by nonconformal metric fluctuations. Gen. Rel. Grav. 2002, 34, 1839–1845. [Google Scholar] [CrossRef]
  54. Petruzziello, L.; Wagner, F. Gravitationally induced uncertainty relations in curved backgrounds. Phys. Rev. D 2021, 103, 104061. [Google Scholar] [CrossRef]
  55. Filho, R.N.C.; Braga, R.P.M.; Lira, J.H.S.; Andrade, J.S. Extended uncertainty from first principles. Phys. Lett. B 2016, 755, 367–370. [Google Scholar] [CrossRef]
  56. Diab, A.M.; Tawfik, A.N. A Possible Solution of the Cosmological Constant problem based on GW170817 and Planck Observations with minimal length uncertainty. Adv. High Energy Phys. 2022, 2022, 9351511. [Google Scholar] [CrossRef]
  57. Tawfik, A.N. On quantum-induced revisiting Einstein tensor in the relativistic regime. Astron. Notes/Astron. Nachr. 2023, 344, e220071. [Google Scholar] [CrossRef]
  58. Munteanu, G. Complex Spaces in Finsler, Lagrange and Hamilton Geometries; Springer Science+Business Media: Dordrecht, The Netherlands, 2004. [Google Scholar] [CrossRef]
  59. Elgendi, S.G. On the classification of Landsberg spherically symmetric finsler metrics. Int. J. Geom. Meth. Mod. Phys. 2021, 18, 2150232. [Google Scholar] [CrossRef]
  60. Elgendi, S. Solutions for the Landsberg unicorn problem in Finsler geometry. J. Geom. Phys. 2021, 159, 103918. [Google Scholar] [CrossRef]
  61. Youssef, N.L.; Elgendi, S.G.; Kotb, A.A.; Taha, E.H. Anisotropic conformal change of conic pseudo-Finsler surfaces, I. Class. Quant. Grav. 2024, 41, 17505. [Google Scholar] [CrossRef]
  62. Bao, D.; Chern, S.-S.; Shen, Z. An Introduction to Riemann–Finsler Geometry; Springer Science+Business Media: New York, NY, USA, 2000. [Google Scholar] [CrossRef]
  63. Cheng, X.; Shin, Z. Finster Geometry: An Approach via Randers Spaces; Science Press Ltd.: Beijing, China; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar] [CrossRef]
  64. Padmanabhan, T. Duality and zero point length of space–time. Phys. Rev. Lett. 1997, 78, 1854–1857. [Google Scholar] [CrossRef]
  65. Mondal, V. Duality principle of the zero-point length of spacetime and Generalized Uncertainty Principle. EPL (Europhys. Lett.) 2020, 132, 10005. [Google Scholar] [CrossRef]
  66. Caianiello, E.R. Some remarks on quantum mechanics and relativity. Lett. Nuovo Cim. 1980, 27, 89–96. [Google Scholar] [CrossRef]
  67. Caianiello, E.R.; Marmo, G.; Scarpetta, G. (Pre)quantum geometry. Nuovo Cim. A 1985, 86, 337–355. [Google Scholar] [CrossRef]
  68. Brandt, H.E. Finslerian fields in the spacetime tangent bundle. Chaos Solitons Fractals 1999, 10, 267–282. [Google Scholar] [CrossRef]
  69. Witten, E. 2+1 dimensional gravity as an exactly soluble system. Nucl. Phys. B 1988, 311, 46–78. [Google Scholar] [CrossRef]
  70. Groenewold, H.J. On the Principles of elementary quantum mechanics. Physica 1946, 12, 405–460. [Google Scholar] [CrossRef]
  71. Woodhouse, N.M.J. Geometric Quantization; Clarendon Press: New York, NY, USA; Oxford University Press: New York, NY, USA, 1992. [Google Scholar] [CrossRef]
  72. Kirillov, A.A. Elements of the Theory of Representations; Springer: Berlin/Heidelberg, Germany, 1976. [Google Scholar] [CrossRef]
  73. Brandt, H.E. Riemann curvature scalar of space–time tangent bundle. Found. Phys. Lett. 1992, 5, 43–55. [Google Scholar] [CrossRef]
  74. Xun, Y.C. Generalized Uncertainty Principle and Its Applications. Ph.D. Thesis, National University of Singapore, Singapore, 2014. [Google Scholar]
  75. Snyder, H.S. Quantized space–time. Phys. Rev. 1947, 71, 38–41. [Google Scholar] [CrossRef]
  76. Liang, S.-D. Dirac theory in noncommutative phase spaces. Physics 2024, 6, 945–963. [Google Scholar] [CrossRef]
  77. Benczik, S.; Chang, L.N.; Minic, D.; Okamura, N.; Rayyan, S.; Takeuchi, T. Short distance versus long distance physics: The classical limit of the minimal length uncertainty relation. Phys. Rev. D 2002, 66, 026003. [Google Scholar] [CrossRef]
  78. Robertson, H.P. The uncertainty principle. Phys. Rev. 1929, 34, 163–164. [Google Scholar] [CrossRef]
  79. Schrödinger, E. Zum Heisenbergschen Unschaerfeprinzip. Sitzungsber. Preuß. Akad. Wiss. Phys.-Math. Kl. 1930, 1930, 296–303. Available online: https://books.google.ch/books?id=QP3jAAAAMAAJ&pg=PA296 (accessed on 2 September 2025).
  80. Randers, G. On an asymmetric metric in the four-space of general relativity. Phys. Rev. 1941, 59, 195–199. [Google Scholar] [CrossRef]
  81. Martinetti, P. Line element in quantum gravity: The Examples of DSR and noncommutative geometry. Int. J. Mod. Phys. A 2009, 24, 2792–2801. [Google Scholar] [CrossRef]
  82. Dubois-Violette, M. Lectures on graded differential algebras and noncommutative geometry. Math. Phys. Stud. 2001, 23, 245–306. [Google Scholar] [CrossRef]
  83. Madore, J. An Introduction to Noncommutative Differential Geometry and Its Physical Applications; Cambridge University Press: Cambridge, UK, 1999. [Google Scholar] [CrossRef]
  84. Ulhoa, S.C.; Santos, A.F.; Amorim, R.G.G. On Non-commutative correction of the Gödel-type metric. Gen. Relativ. Gravit. 2015, 47, 99. [Google Scholar] [CrossRef]
  85. FitzGerald, P.L. The Superfield quantisation of a superparticle action with an extended line element. Int. J. Mod. Phys. A 2005, 20, 2639–2655. [Google Scholar] [CrossRef]
  86. Pfeifer, C.; Relancio, J.J. Deformed relativistic kinematics on curved spacetime: A geometric approach. Eur. Phys. J. C 2022, 82, 150. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tawfik, A.N.; Elgendi, S.G.; Shenawy, S.; Hanafy, M. Canonical Quantization of Metric Tensor for General Relativity in Pseudo-Riemannian Geometry. Physics 2025, 7, 52. https://doi.org/10.3390/physics7040052

AMA Style

Tawfik AN, Elgendi SG, Shenawy S, Hanafy M. Canonical Quantization of Metric Tensor for General Relativity in Pseudo-Riemannian Geometry. Physics. 2025; 7(4):52. https://doi.org/10.3390/physics7040052

Chicago/Turabian Style

Tawfik, Abdel Nasser, Salah G. Elgendi, Sameh Shenawy, and Mahmoud Hanafy. 2025. "Canonical Quantization of Metric Tensor for General Relativity in Pseudo-Riemannian Geometry" Physics 7, no. 4: 52. https://doi.org/10.3390/physics7040052

APA Style

Tawfik, A. N., Elgendi, S. G., Shenawy, S., & Hanafy, M. (2025). Canonical Quantization of Metric Tensor for General Relativity in Pseudo-Riemannian Geometry. Physics, 7(4), 52. https://doi.org/10.3390/physics7040052

Article Metrics

Back to TopTop