As the FRWq dynamics may be described completely in terms of a Lagrangian similar to that of a relativistic particle, we can proceed further in our scheme to quantize this model described by FRWq Lagrangian. The quantization of cosmological Universes has been a prosperous field for decades. This field, called quantum cosmology, attempts to construct a quantum theory for the entire Universe. However, there is not a unique form to achieve this. Several possibilities have been carried out (for a comprehensive review, we refer the reader to Reference [
5]). The most famous procedure corresponds to do a canonical quantization of the classical dynamical equations for the FRW Universe. This is done by replacing the momentum by a derivative operator on the scale factor variable. The final equation is known as the Wheeler–DeWitt equation [
11,
12]. This equation will depend on the main features we want to study of the Universe. Thus, quantum cosmologies have been studied for Universes with a dynamical vacuum in de Sitter cosmologies [
13], in anti-de Sitter spacetimes [
14], in radiation-dominated Universe [
15], in Universes with cosmological constant [
16,
17], in
gravity [
18,
19], in conformal theory [
20], and with a massive scalar field [
21], among plenty of other works.
In order to avoid problems with the procedure of canonical quantization of the FRWq system, we restrict ourselves to the cases where
, and consider static manifolds only [
22], where there exists a family of spacelike surfaces which are always orthogonal to a timelike Killing vector. This implies that
, or:
which means that the original potential
is a constant. Thereby, for the current quantization process,
is essentially equal to a cosmological constant. We have just restricted ourselves to the cases in which the FRWq Lagrangian is
independent. The associated Noether conservation law forces that
cannot change sign (see Equation (
9) for a
independent potential). Thereby,
may be used as the evolution (time) variable, in the form that the variational principle and quantization procedure suggest. Notice that the quintessence field has now the functionality of a super-time in this new description, where the particle evolves in an effective two-dimensional conformally flat spacetime.
Classically, it can be rigorously shown [
23] that the Hamiltonian for the system described by Lagrangian (
23) is:
where
is the canonical momentum. We used this Hamiltonian to construct the quantum theory for the FRWq system. In order to avoid factor ordering issues, the quantum Hamiltonian operator
may be constructed from its classical analogue (
28) as:
where
is the momentum operator defined as
because of
. In this way, the quantum equation that describe the quantization of the FRWq system is:
where
is the wavefunction of the FRWq system.
In principle, one may ask whether there are ways to construct other Hamiltonian operators that differ from Equation (
29), giving rise to quantum theories which are not equivalent to the one described by Equation (
31) (see, for instance [
24]). This point is subtle, and the answer is affirmative; however, the operator (
29) has the advantage in that it reproduces results from quantum field theory in curved spacetimes, as we will see below. In what follows, we proceed to quantize the FRWq theory in two majorly different ways. In the first case, we quantize the system using a procedure devised for spinless particles. This approach produces a Klein–Gordon equation for the wavefunction of the FRWq Universe. The second case corresponds to the quantization of the FRWq system as a Dirac particle.
4.1. Quantization of the FRWq System as a Klein–Gordon Particle
One way to canonically quantize the relativistic spinless particle can be obtained following the method developed by Gavrilov and Gitman [
25]. This procedure is a consistent way to construct the quantum theory along Dirac’s theory for gauge and constrained systems [
26,
27]. We will not reproduce the calculations for this quantization scheme here, but we limit ourselves to exhibit the results of applying this method. The quantization for the FRWq system produces the quantum equation [
25]:
(taking
for convenience) where
is the spinor:
and
is a matrix Hamiltonian:
Let us notice that these are not two dynamical equations for a spinor, but Equation (
32) will produce the constraint
. Therefore, Equation (
32) gives rise to a Klein–Gordon equation:
that it also can be written as:
with the notation from previous section, Equations (
24) and (
25),
and
. The wavefunction
represents the probability amplitude obtained by using the quantization of the FRWq Universe system as a Klein–Gordon particle. It obeys the same equation of the Klein–Gordon field in Minkowski space but now with an effective mass term whose origin is the conformal metric. Equation (
35)—or (
36)—is consistent with the principle of manifest covariance, as this quantum theory emerges from general relativity. A similar result is obtained when a scalar field is quantized in an expanding curved spacetime background [
28], obtaining a mass-corrected term due to a conformal time. However, we must emphasize that our treatment is different because, in our approach, it is the spacetime itself that is quantized.
The conserved Klein–Gordon probability density is not positive definitive. For the case at hand, one can calculate, from Equation (
35), the probability density
for the Klein–Gordon field as:
where
is the complex conjugated of the wavefunction (
41), and
is some value of the scale factor that can be equal or larger than its present value. We notice that the probability density depends on
. In the same way, for our case, the expected value of the scale factor is [
29]:
which depends on
, as well.
We can solve Equation (
35) exactly for
and any constant
V. Assuming the dependence
, for a constant parameter
E, the previous equation becomes:
For a spatially flat Universe
, where
—being
a constant related to the cosmological constant—, the general solution is found to be:
where
is the Euler gamma function, and
is the modified Bessel function of the first kind of order
n. By appropiate choice of constants
and
, the wavefunction
can be real.
To find solutions for
for the wavefunction
, we proceed assuming that the wavefunction can be decomposed as:
where the reason to go back to the
a-representation in the variables will be clear in the following. Here,
is a parameter that, as we will see below, can be associated with the energy of the
n-state of the Klein–Gordon field
. Using this decomposition, we can find a suitable equation to be solved. First, let us define the field
. This new field follows the differential equation:
where we have introduced the effective potential:
that depends on the value of
k. Equation (
42) can be converted in the following Ricatti equation:
where:
The above Ricatti equation can be generally solved if a particular solution can be found for any k.
This is not a trivial task. An approximated solution of Equation (
42) for the field
can be obtained using the Spectral Method (SM) [
30,
31] in the expansion (
41). This method is usually utilized in similar quantum theories for the Universe [
17]. As the wavefunction of the Universe must vanish at the origin, as well as in infinity, the SM uses the approximation that the wavefunction vanishes at some length
L (as large as we require). Thus, the SM allows us to expand the wavefunction
in a Fourier series:
where
are constant coefficients that depends on the
n-state. Notice that this expansion implies, from Equation (
41), that the wavefunction
, which is the desired behavior. Here,
N is a number that can be chosen arbitrarily. The approximation improves as
N increases. According to the SM, we can expand also the following functions:
where again
and
are coefficients depending on the
n-state. It is straightforward to find the relation between
and
with
. Using Equation (
46) in Equation (
47), we get that
, and
, where:
Now, with the expansion in Equations (
46) and (
47) in Equation (
42), we can finally obtain the eigenvector equation:
for the
vector (with
N components
) and where
corresponds to the eigenvalues. Here,
is the
inverse matrix of
(with components
), and the
matrix
has the components
. The dimensions of the matrices
and
will be fixed once a cut on the series Equation (
46) in Equation (
47) will be done, while the better the approximation, the larger will be the matrices. Thus, the system is completed solved. The values of
correspond to the energies of the different possible states of the Universe. This also allows us to identify the variable
as a super-time, as it was previously discussed.
If one is being less restrictive with the assumption that the potential
is constant, Equation (
35) can be written in the form of the Wheeler–DeWitt Super-Hamiltonian formalism [
12,
21]. For example, for a closed universe
—and the case quinteessence
—, it is possible to rewrite Equation (
35) as:
where
, and
. To obtain this equation, we chose
and
, where
is the mass of the field. Here,
is usually called the Wheeler–DeWitt Super-Hamiltonian [
12,
21]. Thus, the quantization of the FRWq system as a Klein–Gordon particle proposed here can reproduce known results of quantization using the Super-Hamiltonian formalism.
On the other hand, another possible solution of Equation (
42) could be achieved using the Frobenius (polynomial) method, which is different from the SM. This solution corresponds to a polynomial expansion in
a, where all the coefficients can be found from recurrence relations. A polynomial solution for
has the form:
where
and
are constants. For the sake of simplicity, we choose
. Using the previous expansion in Equation (
42) we find:
where
V is defined after Equation (
9). Equating every term to zero, we can readily find the solutions:
with also
,
, and the recurrence relation for
is:
where now the problem is completely solved. The weakness of this method is that it does not give any physical meaning to the constant
in the ansatz (
41).
4.2. Quantization of the FRWq System à la Dirac-Pauli
Basically, the quantization process proposed here consists in finding the square root of the Hamiltonian operator (
29). In principle, one can use matrices to find the square root, but its use implies the notion that the cosmological model behaves as a Dirac particle. At a first glance, it may appear strange to quantize a model for a relativistic pointlike particle with a quantum spin theory. However, in 1928, Breit [
7] showed that there exists a correspondence between the Dirac and the relativistic pointlike particle Hamiltonians. In that work, it is shown that one can obtain the Dirac equation via a prescription of replacement of the particle velocity and the Dirac matrices, as well as the prescription in Schrödinger or Klein–Gordon theories where the energy and momentum and can be replaced by the time and space derivatives. Thus, Breit’s prescription implies a classical and geometrical interpretation of the spin. We leave the calculations and the deep discussion of this idea to
Section 5. For now, in this section, we restrict ourselves to follow Breit’s interepretation and perform the quantization of the FRWq Universe using Dirac matrices.
We propose that the Hamiltonian (
29) can be written using Dirac matrices (
and
). This will give us the Hamiltonian operator:
In
Section 5, we justify this choice. This Hamiltonian allows us to quantize a FRWq Universe as if it were a relativistic spin particle. Using the operator (
55), the quantum mechanical Equation (
31) now reads:
where
now is a bi-spinor. Here,
stands for any of the Dirac matrices
. Choosing the Dirac representation
and
, we can operate Equation (
56) by
by the left to find that:
This equation describes the quantum theory for the FRWq Universe modeled as a Dirac particle. Also, it can be obtained directly from the theory of the Dirac equation in curved spacetimes, thus giving validity to our quantization scheme (shown in
Appendix A).
Dirac matrices are
, and as we are describing a two-dimensional conformal system, we may anticipate that the above equation is reducible. This means that Dirac Equation (
57) couples the wavefunction in pairs, implying that the two pairs of wavefunctions satisfy the same equation. This gives us a hint that a completely equivalent quantization formalism to the previous one can be achieved using Pauli matrices. Solving the Hamiltonian (
29) using Pauli matrices—notice that there is no restriction to this ansatz—, the Hamiltonian operator (
29) can be written as:
where
and
can be any two different Pauli matrices. The quantum mechanical Equation (
31) describing the FRWq system becomes (setting
):
where
is the unit matrix, and
represents a spinor field. We can use the Pauli matrices properties to put the previous equation in the following form:
where
. It can be proven that choosing
and
gives the same dynamical equation as choosing
in the Dirac equation.
On the other hand, from Equation (
57), we notice a that defining the bi-spinor
, we can obtain the equation:
which is a flat 1+1 spacetime massless Dirac equation with a scalar potential. The principle of manifest covariance of this quantum cosmology model can be explicitly seen here, as we could have re-written the previous equation as:
These kinds of equations have been extensively studied and approximated solutions have been found [
32,
33,
34,
35].
Finally, using the wavefunction
(given by solving either the Dirac or Pauli equations), we can calculate the probability density of the Dirac field as:
where
is the transpose conjugated of the wavefunction
. In the previous expression,
and
are written in terms of
a. In a similar fashion, the expected value of the scale factor for the Universe under the Dirac quantization is obtained:
depending again on the values of super-time
.
To further study the system, let us do the bi-spinor descomposition:
where
are constants. Using Equation (
65) in Equation (
57), gives:
where we have made the particular choice of the Dirac matrix:
The same mathematical equations can be obtained for the fields
and
, reflecting that the system can also be studied using Pauli matrices. Now the fields
and
appear coupled. From those equations, it is impossible to recover the Klein–Gordon Equation (
39). The reason is that wavefunctions are coupled to the spacetime metric through the potential due to the quintessence field.
Let us notice that a simple exact solution of Equation (
66) can be found when the fields do not depend on quintessence, i.e., when
. In this case, the solutions are:
where we define:
However, these solutions are not well-behaved at () as it diverges. Therefore, we will seek solutions with .
A more general solution can be obtained in the following way. Let us define
, and
. Thus, the Equation (
66) can be re-written as:
where
, such that
. The previous equations are coupled, but we can find the following second-order equation, which holds for each of the fields:
The above equation can be reduced to familiar expressions doing the change
. The equation for
is reduced to a Ricatti equation:
with:
A general solution of the Ricatti Equation (
72) can be found if we are able to find a particular solution.
On the other hand, when the solutions depend on the quintessence field, we can use the SM to completely solve the Dirac Equation (
66), as well as in the previous section. As it was shown before, this method allows us to reduce the complicated Equation (
66) to an eigenvalue problem for any
k. First, we define the wavefunctions
and
, where now
and
should be written in terms of
a; similar changes can be done for the fields
and
. These two new functions satisfy the equations:
We can now use the SM for the functions
and
. Notice that every term in Equation (
74)’s approach to zero as
a goes to zero. With the SM, we can assume the following dependence for the different functions:
where again
and
are constant coefficients that depend on the
n-state. Anew, the wavefunction (
65) behaves as
. Similarly, we get:
where the relations between the coefficients are
,
,
and
, with the matrix elements:
Using Equation (
75) in Equation (
76), and the previous relations on Equation (
74), we find, after some algebra, the eigenvector equation:
where
corresponds to the eigenvalues (associated to the energy), and the eigenvector
is formed by the
components
and
as:
and the
matrix
is such that:
where the
matrices
and
are constructed by the components given in Equation (
77);
is the identity matrix. The evolution of the system is completely determined by solving the eigenvector and eigenvalue Equation (
78). The approximated solution improves by increasing
N.
4.3. Quantization of the FRWq System à la Majorana
Strictly speaking, the Dirac description of the FRQW system implies that the Universe can interact with self-electromagnetic fields, as the particle modeled can have charge. One way to avoid this issue is to use Majorana matrices instead of Dirac matrices. In this way, the quantization scheme models a Universe with quintessence as a neutral relativistic quantum particle. The quantum mechanical equation is:
where now
are the Majorana matrices, and
represents the wavefunction of the FRWq system in the Majorana scheme of quantization.
Similar to the previous case, the expected value of the scale factor for the Universe under the Majorana quantization is:
where
is the transpose conjugated of the wavefunction
, and the probability density is
. As well as all the previous cases, the expected value of the scale factor depends on
.
Analytical representation of the solutions can be obtained by performing the descomposition:
and choosing:
then Equation (
81) may be rewritten as:
Again the wavefunction is coupled to . Similarly, and are coupled by the same mathematical equations.
Notice that unlike the Dirac case, when
, the equations in the Majorana scheme decouple. In this case, the simple solutions can be obtained as:
where we use the definition (
69) for
. As in Dirac scheme, these solutions again diverge for
(
). For
, we can find a modified Klein–Gordon equation for
:
which contains an effective mass term depending on curvature variations, which can be compared with Equation (
39). For the special case of
, it can be shown that this equation reduces to the equation for a diatomic molecule decribed by the Morse potential [
6]. Performing the change of variables
, Equation (
87) can be simplified to:
As discussed in the begining of
Section 4, the general potential
must be positive. Therefore, we can choose a representation of
, where
is a constant. Making another change of variables
and defining
, Equation (
88) can be put in the form:
which is the quantum equation for a diatomic molecule described by the Morse potential. The wavefunctions and energy spectrum of this problem are already known.
Similar to previous sections, analytical approximated solutions for any
k can be found using the SM approach. As before, we can perform this task by defining the variables
and
that satisfy the following equations:
Again, notice that the SM allows us to have a well-defined behavior of the wavefunction (
83) as
. Applying the SM means we have to use similar decompositions, as in Equations (
75)–(
77), to the Majorana case. For simplicity, we use the same notation as before. Equation (
90) can be finally written as:
where again
are the eigenvalues and
is the vector (
79). The
matrix
is now:
where again
and
are
matrices constructed by Equation (
77), and
is the
zero matrix.
Another important feature of Equation (
85) deserves to be highlighted. Defining the new wavefunctions
and
, then Equation (
85) can be re-expressed as:
where the operators are defined as [
6]
Notice that Equation (
93) represents the equations for supersymmetric quantum mechanics [
35,
36,
37]. Each wavefunction satisfies:
where the Hamiltonians:
can be used to define the Super-Hamiltonian:
used to write the above Equation (
95) as:
The operators (
94) can be used to define the supercharges:
which are operators that can change bosonic (fermionic) states into fermionic (bosonic) ones. The above supersymmetric system exhibits the same features of any other supersymmetric quantum theory [
37].