1. Introduction
The quantization of gravity possesses a fundamental place in the realm of theoretical physics. As was argued by Claus Kiefer [
1], there are mainly three arguments that motivate such a quest: the first motivation is based more on philosophical reasons and the idea that all fundamental interactions should be unified at some energy scale. Hence, if a coherent quantum theory exists that describes those fundamental interactions, gravity should be included as well. The second motivation is related to the appearance of singularities in the classical theory of gravity, at is described by general relativity. The hope is that a quantum theory of gravity will be free of such pathological situations. Finally, the third motivation is related to the so-called “problem of time” as can be found in [
2,
3,
4,
5,
6,
7]. The main reason that such a problem arises is that in the conventional quantum theory, time is an external parameter, while in the general theory of relativity, there is no preferred or fundamental notion of time. This is due to the diffeomorphism invariance of the theory. Furthermore, this invariance results in a singular Lagrangian and Hamiltonian density respectively, which implies constraint equations in addition to the dynamics. The relation between constraint equations and diffeomorphisms can be found in [
8], as well as in [
9].
The appearance of constraints results in two main approaches in the quantization procedure: the first is to solve the constraints before quantization, while the second is to impose the constraints at the quantum level, as restrictions upon the states. One of the main paths to deal with the above approaches is the so-called “canonical quantization”, where the canonical variables that describe the gravitational field are promoted to operators and the Poisson brackets are replaced by operator commutators. Some of the original work in the direction of the Hamiltonian formalism of gravity was given in [
10,
11,
12,
13,
14,
15,
16,
17,
18,
19].
Related to the second procedure, the most famous result is the Wheeler–DeWitt equation [
20,
21,
22,
23,
24,
25], which results from the quadratic constraint, or in other words, from the variation of the gravitational action with respect to the lapse function, in the canonical formalism. This equation contains no external parameter in which the term “time” could be assigned. In fact, due to its hyperbolic nature, the “time” parameter is encoded in the degrees of freedom that appear, hence the name “intrinsic” time. Thus, additional hypothesis are needed in order to define a “time” and hence arrive at some sort of a Schrödinger equation. One of the problems arising due to the hyperbolic nature is the difficulty in defining a proper Hilbert space as in the conventional quantum theory.
When it comes to the first procedure, the usual approach is to introduce quantities
, which will be functions of the degrees of freedom and impose a suitable gauge choice. To this end, a “true” Hamiltonian density
can be constructed, which depends only on the true gravitational degrees of freedom. The constraint equation has the form:
where
are the conjugate momenta of the introduced quantities
. A Schrödinger-type equation can then be realized from the above equation, with the identification
:
An extensive analysis of this approach can be found in [
5,
26,
27]. Now, since in the constraint, the momenta appear quadratically, in many cases, it is very difficult to obtain a closed form density
. This is one of the reasons that many authors have assumed the existence of matter in order to define a standard of time [
28,
29,
30,
31]. In this case, a Hilbert space can be defined, and alongside this, a probabilistic interpretation could be provided [
5]. As was argued by Barvinsky [
6], choosing a gauge breaks the gauge invariance of the theory and, hence, for each different gauge (classically equivalent systems), one has different quantum theories, which must be proven to be equivalent. Another problem is that the Hamiltonian density
is the square root of the true degrees of freedom and the “time” variables, which implies that in many cases, it will remain real only for a specific domain of the “time” variable. Furthermore, since it is a square root operator, it is known that it can be handled only if the quantity whose root is being taken is self-adjoint and positive [
5]. Note also that
are “time” dependent, which implies that we cannot reproduce the Wheeler–DeWitt equation from (
2). This is not something to worry about; it simply reveals the different quantization approaches. An interesting new approach to the problem of time can be found in [
32].
These interesting questions are very difficult to tackle in the full theory of gravity. In order to acquire some insight, people have worked on simplified models such as minisuperspace Lagrangians, where the isometries of spacetime are enough, in order to render the system from infinite to finite dimensional. In cosmology, such examples are the FLRW geometries, the Bianchi types, and the Kantowski–Sachs metric [
33]. Additionally, there are the point-like sources such as Schwarzschild (S) [
34] and Reissner–Nordström (RN) [
35] spacetimes. For instance, in [
36], the author argued that the classical properties of an intrinsic time variable, as the scale factor in FLRW geometry, would emerge in some sense, through the interaction with fermionic degrees of freedom. The semiclassical quantum cosmology was investigated [
37] for the FLRW minisuperspace model with electromagnetic field perturbations around the background. Wave packets were constructed for the coupling of gravity and a massless, as well as a massive scalar field in the FLRW geometry [
38]. Another work with the same underlying geometry and in the presence of scalar field can be found here [
39]. For the case of general, spatial cosmologies in the presence of a scalar field, exact solutions have been found [
40]. In the case that the Hamilton–Jacobi equation is separable for the cosmological models under consideration, it has been proven [
41] that these models can be quantized as ordinary gauge systems. The solutions to the Wheeler–DeWitt equation for the Reissner–Nordström–de Sitter black hole were interpreted via the de Broglie–Bohm theory [
42]. The use of conditional symmetries was employed for the quantum description of the general Bianchi Type I, with and without the presence of a cosmological constant [
43]. In the same line of thought, the canonical quantization of the Reissner–Nordström black hole can be found in [
44], where the interpretation of the solutions to the Wheeler–DeWitt (WDW) equation is based on the semiclassical corrected geometry via Bohm’s analysis. Another work that shares some common ground with what we intend to do in this paper is [
45]. The authors managed to decouple the reparametrization invariance for some minisuperspace models and hence reduced the systems to the “true” degrees of freedom. Furthermore, they provided a generalized definition of probability. For an interpretation of the wavefunctions through the definition of homothetic time, we provide this work [
46]. A different perspective on the quantum mechanical corrections to the Schwarzschild black hole can be found in [
47]. The authors studied the contributions to the Bekenstein black hole entropy due to quantum corrections. Yet another different approach can be found in [
48,
49], where a Schrödinger-type equation was found for the Schwarzschild and Reissner–Nordström black holes correspondingly. The quantum black holes were treated as a gravitational hydrogen atom, and their energy spectrum was found. The authors in [
50] described a quantum Schwarzschild black hole by a non-singular wave packet composed of plane wave eigenstates of the momentum Dirac-conjugate to the mass operator. Furthermore, in [
51], they provided an interpretation to an emergent spacetime as the quantum mechanical Schwarzschild vacuum, which appears to be massless, but exhibits non-zero mass uncertainty. These are only a few examples of the full list of quantum minisuperspace models.
The purpose of this work is to provide some sort of a Schrödinger-type equation for the minisuperspace Reissner–Nordström black hole, without imposing any gauge conditions and without introducing additional functional for the purpose of choosing “time” variables. To do so, we recognize that one of the sources to the various problems is the existence of constraint equations and hence the singular nature of the Lagrangian density. Our intention is to satisfy the constraint equation without imposing any gauge condition and then construct a reduced Lagrangian density, which will be regular for some of the degrees of freedom and, furthermore, will not have a square root form. Since the gauge is not broken, we expect non-true degrees of freedom to be included in the Lagrangian density. Those will help us to identify the “time” parameter and hence to define a Schrödinger-type equation. For whatever wavefunctions are to be found, we intend to use the Bohm analysis in order to present an interpretation based on geometrical tools. Furthermore, some sort of criterion is provided with which we adjudicate which, between two spacetimes, is “more” singular in comparison. Once the Hilbert space is constructed, a probability density can be defined, to which we apply DeWitt’s idea on whether singularities are avoided (the probability density should be zero at the configuration points where the classical singularity appears).
The paper is structured as follows: There are four basic sections, the classical description in
Section 2, the quantum description in
Section 3, the Bohm analysis in
Section 4, and the Discussion in
Section 5. Related to
Section 2: the starting point is
Section 2.1 is the introduction to the line element and the electromagnetic potential for static spherically-symmetric spacetimes, prior to the solution of the equations. In
Section 2.2, the well-known singular Lagrangian density is provided. The reduced, “time”-dependent, regular Lagrangian density is presented in
Section 2.3. The last subsection,
Section 2.4, is dedicated to some variable transformations in the minisuperspace geometry and obtaining the solution in an arbitrary gauge. When it comes to the quantum description: Some variable redefinition are presented in
Section 3.1, and the canonical Hamiltonian density is constructed in
Section 3.2. In
Section 3.3, the “time”-covariant Schrödinger equation is defined. Two subsections follow,
Section 3.3.1 and
Section 3.3.2, where wavefunctions to the above equation are found for specific initial states. Next is
Section 4, where the Bohm analysis is performed in
Section 4.1 and
Section 4.2, for the two wavefunctions obtained in the previous section. The singularity criterion is presented in
Section 4.3 for the comparison of the classical and semi-classical trajectories. Finally,
Section 4.4 is related to the existence of event horizons for the semi-classical solutions. Lastly, in
Section 5, a discussion can be found for the overall results.
4. Bohm Analysis
Even though we have a well-defined inner product and the solutions at hand and we could calculate expectation values and so fourth, we find it more insightful to interpret the wavefunctions by their impact on the geometry. That is to say, what are the quantum corrections to the classical geometry?
On of the ways to do such a thing is based on Bohm’s analysis [
53,
54,
55]. A de Broglie–Bohm interpretation for the full quantum-gravitational system can be found in [
56]. For a recent review, check [
57]. Let us briefly recall the main idea properly adjusted in our example: Suppose some Hamiltonian density and a Schrödinger equation of the form:
Furthermore, assume that the states can be cast into the form:
where
are real functions respectively called the amplitude and phase of the state. By use of (
106) in (
105) and equating to zero the real and imaginary parts of the resulting expression, two equations come up:
where:
the so-called “quantum potential”. Equation (
107) has the form of a continuity equation with
the “density” and
the “velocity” of the “fluid”. This is basically the local expression for the probability conservation. Additionally, (
108) is a modified version of the Hamilton–Jacobi equation, by the quantum potential term. Thus, in the absence of it, the classical Hamilton–Jacobi equation is retrieved, hence the justification of the term “quantum potential”.
Following Bohm, the connection to the classical regime is to identify the “fluid momenta”
to the classical one, given by the Lagrangian density describing the system
), that is:
Equation (
110) is a system of first order differential equations for the “positions”
and will provide us with what is called the “semi-classical trajectories”.
For later purposes, the easiest way to compare the semi-classical trajectories to the classical ones is to make the usual gauge choice:
which implies that at
, the line element acquires the form:
4.1. Delta Function Initial State
In this case, the quantum potential is calculated to be equal to zero. This was expected since as we can see from (
93), the state’s amplitude is independent of the coordinates
. The continuity equation is satisfied with a “current probability density”:
As is expected, Equation (
110) is of the form:
which yields the classical solutions:
As we can see, the constants that were introduced in the initial quantum state appear as integration constants for the semi-classical trajectory. The most important outcome is that, for the above quantum description, when the initial state is localized, or the wavefunction has infinite “uncertainty” in the “momenta”, the semi-classical solution coincides with the classical one.
4.2. Gaussian Initial State
The introduction of a Gaussian initial state is expected to give us one of the possible ways to deviate from the classical trajectories, since it approaches the above delta state only at the limit
, as we have already pointed out. For the solutions to be unburdened from unnecessary, arbitrary constants, which were introduced for a matter of unit conventions, we introduce the parameter
, which will be defined instead of
via the relation:
or in terms of Planck’s length,
Hence, the delta initial state is expected to be retrieved at the limit and so the classical trajectories. Note also that has units of length.
Now, the quantum potential is not zero in this case:
and the continuity equation is again satisfied with the probability current reading:
Equation (
110) is of the form:
which is not the same as in the classical case. The semi-classical solutions read:
where the integration constants
are redefined in order to coincide with the corresponding integration constants of the classical solution. Hence, the radii
can be re-introduced. Furthermore, the correspondence between the integration constants of the classical and semi-classical trajectories can be found from the classical limit
to be
. That being taken into account, the semi-classical metric and electromagnetic potential have the form:
where the function
reads:
In addition to the different form of the metric and electromagnetic potential, (
124) and (
125) do not satisfy the free Einstein–Maxwell equations. An additional energy-momentum and a four-current tensor have to be taken into account,
Both
are considered quantum mechanical in origin. For completeness, their non-zero components are provided:
By evaluating carefully the limit of the above expressions at , we notice that all of them tend to zero, as expected. Thus, the realization of the quantization effect is localized in the alteration of the forms of the original fields and the appearance of sources, in both sets of equations.
4.3. Singularity
Let us now discuss the singularities. For the classical solution, the Ricci scalar is equal to zero, due to the special behavior of the electromagnetic field in four dimensions, that is the trace of the energy momentum tensor is equal to zero. This is a generic feature and does not depend on whether or not the solution is classical or semi-classical. As we have seen, for the semi-classical solution, an additional energy momentum tensor has to be taken into account. This leads to a non-zero Ricci scalar, which vanishes at the limit
, sharing the same property of
. The behavior of this scalar for small values of
r (we present only the divergent terms) is:
Thus, we can immediately conclude that this spacetime is not singularity free. Since in the classical case,
, the Ricci scalar is not a good measure of whether this spacetime is “more” or “less” singular than the classical Reissner–Nordström one. Let us explain what we mean by “more” or “less” singular: Consider the cases of Schwarzschild and Reissner–Nordström black holes. Since in both cases,
, we are going to use the Kretschmann scalar
for our purposes, which reads respectively:
In order to work with dimensionless objects, we multiply both expressions with
and provide
in terms of
as well, specifically
,
, with (
, so that no naked singularity appears):
Now, the difference between the two is defined:
and we provide the contour plot in the space of the two parameters
.
As can be inferred from
Figure 1, the function
X acquires negative values in the accepted parameter space. This means that
; hence, we can “move” towards smaller values of
x (or
r) in the case of Reissner–Nordström rather than that of Schwarzschild, before the curvature “blows up” to infinity. In some sense, the addition of the electromagnetic field (a point charge) leads to a more localized singularity around the value
, and hence, by this criteria, a “less” singular spacetime.
To this end, we are going to compare the classical Reissner–Nordström and the quantum corrected one (QRN) based on the above described method. To proceed, the constant
is redefined in terms of
as follows
, where for
, the classical Reissner–Nordström is retrieved. The quantum RN has many divergent terms, but we keep only the higher ones, which coincide in powers of
r with those of the classical solution. For completeness,
Now, the space of parameters is three-dimensional; thus, we could give a density plot. However, in trying, we found out that it is not very presentable. Thus, we decide to construct some contour plots in the space of variables
for some specific values of
x. The results are presented in
Figure 2 and
Figure 3.
Some remarks are at hand: The first thing that we observe is that there are values of the parameters that imply , meaning that the quantum corrected spacetime is “more” singular than the classical one based on the above criterion. The second thing is that in all four cases, there exists the deep blue region, which corresponds to and, hence, a “less” singular spacetime. Finally, as we move closer to , the desired region becomes smaller. Therefore, we can always achieve “less” singular spacetimes for some values of , but the characteristics of those black holes (encoded in ) are bounded to specific values. If we now focus on a specific distance x, then we observe that as grows, meaning stronger quantum effects, the region of accepted values of is closer to , which corresponds to a Schwarzschild black hole.
4.4. Event Horizons
This solution exhibits event horizons as the classical Reissner–Nordström solution. From (
124), the solution
leads to the divergence of the second diagonal element of the metric tensor. There are four solutions to the above equation, but only two of them correspond to a positive sign for
r. The result is:
where
are the usual Reissner–Nordström inner and outer horizons:
There are some remarks worth mentioning.
The existence of both inner and outer horizons is possible only if the following relations are true:
In the case
, there is only one horizon, which we are going to call
:
In the case
, there is no horizon at all. Better expressed, the value of the coordinate
r at which
is imaginary. For this particular value
, the parameter
becomes a function of
,
. In the absence of a horizon, we recognize the existence of a naked point singularity, since as can be inferred from
Figure 4, there are no values
for which
.
From the above, two remarks are that: As the quantum effects become stronger, meaning , the horizons are vanishing. For instance, a black hole with would consist of a region where no information could escape. In the case , this region is defined by . Since , information could escape from regions from which it could not before.
There is also the extremal horizon for which
, implying a degeneracy in this case as well:
This extremal case could not exist for the value .
In the accepted domain of
, in order for
to exist, we find that the following inequalities must hold:
Furthermore, an additional one holds for
:
Thus, the horizons of the QRN are larger than those of the classical solutions.
5. Discussion
The primary purpose of this work is to construct a Schrödinger-like equation, instead of a Wheeler–DeWitt one, for the Reissner–Nordström black hole, which holds the property of being covariant, under reparametrizations of the defined “time” variable. To do so, a different procedure was followed from those appearing in the literature, to the best of our knowledge. Specifically, the singular Lagrangian density (
23)–(
25) describing the dynamics of the system was a function of the degrees of freedom
and the lapse
n. The variation of the minisuperspace action integral with respect to
n yields a constraint Equation (
21), while dynamical equations appear from the variations with respect to the other degrees of freedom (
18)–(
20). Since many of the problems that appear in the quantization procedures are related to the existence of constraint equations, we thought that it is better if it did not exist, or better yet, if it were satisfied identically, once the solutions to the dynamical equations are obtained. One possible way to do so, without breaking the gauge freedom, is to solve the constraint equation with respect to the lapse function. To this end, from the reduced equations, only two of them are independent, (
27) and (
28), while there are three degrees of freedom left
. We argued that this implies that the gauge freedom still exists, and we could have arrived at these two independent equations from the beginning, if somehow we were able to guess the proper form of the lapse as a function of the other degrees of freedom (
26).
The reduced Lagrangian density has a square root form (
32) and is still singular if
are all considered degrees of freedom. However, since there are only two independent equations (the third is satisfied once the other two are solved), these equations of motion can be reproduced via the variation of the action from the reduced Lagrangian density, with respect to only two of the degrees of freedom. There is the freedom to choose any of the pairs
,
, or
. We chose
; hence,
b is now considered as a mere function of the coordinate
r and not a degree of freedom. To this end, the Lagrangian density is regular with respect to
; it is “time” dependent due to the appearance of
, and the gauge freedom still exists, encoded in the choice of
. Note that due to the explicit position that
appears in the spacetime metric, we cannot gauge fix it to a constant value.
As we have pointed out, the Lagrangian density has a square root form, and as is known, problems will arise in the quantization procedure when the square root Hamiltonian density will have to be turned into an operator. To avoid such kinds of problems, we searched for a quadratic in the “velocities” Lagrangian density, which reproduces the equations of motion, and we succeeded in finding one (
36)–(
38).
Before moving to the quantum description, we noticed that the minisuperspace metric corresponds to a flat, Minkowski spacetime. This implies the existence of coordinates
so that the minisuperspace metric acquires the diagonal form with eigenvalues
. To this end, the reduced equations acquired a very simple form (
44), and the solution in an arbitrary gauge was obtained, by solving one algebraic and one linear, second order, differential equation. We find that this is one of the simplest ways to obtain the solution in an arbitrary gauge.
At that point, a remarkable equivalence appeared: these equations of motion have exactly the same form and are equal in number, with the equations of motion for a 3D electromagnetic pp-wave spacetime, which we studied in some previous work [
52]. This bizarre and interesting coincidence inspired us to assume that there might be a kind of classification of spacetimes+matter, based on the equivalence in the form of Einstein’s equations, with the underlying reason being the geometrical characteristics of the minisuperspace metric. This idea, however, will be pursued in some future work.
Obtaining the Hamiltonian density was an easy task. Note that since it is a “time”-dependent quantity, it is not conserved. However, there is the
(
66) and three Noether charges (
67) that are conserved modulo the equations of motion. For the quantum description, we followed the canonical quantization procedure, meaning the elevation of observables into self-adjoint operators and the replacement of the Poisson brackets with the commutator. Since the Hamiltonian density is regular, the construction of a Schrödinger-like equation is possible. Furthermore, the hyperbolic nature of the Hamiltonian density operator implies that we have not defined an “intrinsic time” variable (we have not broken the gauge invariance). The idea is that we use as the parameter of “time” the coordinate (r), which appears explicitly in the Hamiltonian density through the function
(or
equivalently). This Equation (
77) looks like a hybrid between the “intrinsic time” Schrödinger equation and the Wheeler–DeWitt equation that we described in the Introduction. The difference with the first lies in the “time”-covariance property; hence, the quantum descriptions for each different gauge choice are, by definition, equivalent. For the second, the difference appears in the evolution of states in our description, in contrast to the “frozen” picture related to the Wheeler–DeWitt equation. Note at this point that the word “time” was used nominally in the various places inside the text, since as is inferred from the line element, the variable
r is a spatial coordinate. Alongside this, any reference to evolution of states is with respect to this external parameter
r and should not be confused with the usual quantum mechanical sense of the evolution of states in time.
The solution to this equation was obtained for wavefunctions that are common to the operators
. The evolution of two initial states was studied, Dirac’s delta and Gaussian. These two approach each other at a certain limit of the parameter that appears in the Gaussian distribution and control its width. For both cases, the wavefunctions were obtained, and the probability densities were calculated (
96) and (
103), respectively. In order to acquire some sense of the quantum corrections to the spacetime and interpret the wavefunctions in geometrical terms, the Bohm analysis was employed. The results are the following: For Dirac’s delta initial state, the quantum potential is zero; hence, the semi-classical “trajectories” coincide with the classical ones. The Gaussian initial state becomes a delta function only at a certain limit; hence, the obtained result is a non-zero quantum potential, resulting in different semi-classical trajectories from the classical “trajectories” (
124)–(
126). The quantum corrections manifest themselves as the appearance of an additional energy-momentum tensor
and a four-current density
in the system of Einstein–Maxwell equations. Their explicit formulas can be found in (
129)–(
132). One important aspect of the semi-classical solutions is that they contain the classical solutions as a limit, specifically in the limit that
,
vanish.
As we pointed in the Introduction that one of the motivations to construct a quantum theory of gravity is for the purpose of “healing” the singularities from which the classical theory suffers. As we obtained in
Section 4.3, the quantum corrected RN is still singular. One way to observe this is due to the Ricci scalar, whose appearance is solely due to
. For small values of the variable
r, this object scales as
; hence, as
, the Ricci scalar tends to infinity. At that point, since for the classical RN solution, the Ricci scalar is zero, we found it more insightful to compare the Kretschmann scalars of these two spacetimes. We defined a criterion in order to adjudicate whether or not the situation gets better in the case of QRN regarding the divergence of the Kretschmann scalar. The criterion states the following:
Between two spacetimes that suffer from the same kind of singularity, in this specific case, a point-like, “less” singular one will characterize the one whose singularity is more localized around the point of interest.
Regarding the QRN and RN black holes, there are admissible values of the parameters for which QRN is “less” singular, as well as values for which is “more” singular. In the figures that are presented, there is a tendency that states that: as (translated as stronger quantum effects, moving further away from Dirac’s delta distribution), the characteristics of these black holes that are encoded as values of for which QRN is still “less” singular are close to zero. The upper value of for a specific value of depends also on the proximity to the point .
Another interesting result also related to the singularities is the following: Since we obtained a well-defined Hilbert space, and we calculated the probability densities, we may follow DeWitt’s proposal [
23], which states that: if the wave function or, better, the probability density vanishes at the configuration points where the classical singularity appears, then it is avoided at the quantum level. The classical singularity appears for
, which translates into
. By carefully evaluating at this limit the probability densities for both cases of initial states, (
96) and (
103), we find that they are equal to zero. Hence, in light of the above proposal, the singularities disappear in the quantum regime. Therefore, a discrepancy appears between this result and the result obtained via Bohm’s analysis. One possible explanation might be that Bohm’s analysis is just an approximate scheme, with the approximation localized in the identification (
110) of the classical and quantum momenta. A recent article discussing the singularity resolution and its dependence on the choice of clock can be found in [
58].
Since we used the probability densities to comment on the existence of singularities, it is rather instructive to comment also on the possible interpretation of these probability densities. As we explained, the external parameter
r is considered as the “time” variable in the Schrödinger equation. To this end,
is to be understood as the probability density for the variables
at “time”
r. This kind of interpretation is not unknown; it appears also in the “intrinsic time” formalism of the full quantum theory, where a multi-“time” Schrödinger equation is constructed. For more information, see [
5]. In this particular work, we used symmetries to restrict ourselves to a minisuperspace model, which implies, due to the minisuperspace model at hand, that a spacelike parameter
r remains to be interpreted as “time”. If a cosmological minisuperspace model were to be used instead, then the usual timelike coordinate
t would appear. Regarding the collapse of the wave-function after some measurement we may say that the usual quantum mechanical interpretation could be used. Lastly, due to the lack of a complete physical theory of quantum gravity, it is far from obvious to know whether such a quantum minisuperspace model would appear as a limiting case of the full theory and, hence, render it physical. At this point in time, we would say that it is more of a mathematical model with a possible physical meaning through the use of the Bohmian interpretation; it might be possible to contrast the quantum corrected metric with observations.
Event horizons (inner and outer) also appear for QRN, under the assumption that the constant related to the quantum effects, , satisfies the condition , where is the outer RN horizon. In comparison with the event horizons of the classical solution, we found that they have a larger radius. The reason for that may be attributed as well to the presence of the additional energy-momentum tensor in the quantum corrected metric. There exists the case of extremal horizons as well. Furthermore, for different values of the parameter , which controls the localization of and , there are the cases where two horizons exist, only one degenerate and none, implying a naked singularity. We observed that with the above defined criteria for the singularity resolution, the naked singularity does not disappear. Of course, let us point out once again that based on DeWitt’s proposal, there are no singularities at all. Hence, this is another strand of the revealed discrepancy. All the results that were obtained via Bohm’s analysis can be reduced to the Schwarzschild black hole by the simple substitution . For future work, it may be interesting to study the geodesic equation and find the corrections that are imposed due to quantum effects. Furthermore, some experimental boundaries may be imposed on the values of the parameter .
Finally, we believe that this “hybrid” procedure on the construction of a Schrödinger equation, for the case of minisuperspace models, will provide some new perspective on the problem of time. It is our intention to study, in some future work, other minisuperspace models as well and, perhaps, even try to formulate this procedure in the full theory of gravity.