In previous sections, we have shown that the Hamiltonian constraint (
33) describes the deep interior dynamics of scalar excitations without angular momentum falling into a macroscopic Schwarzschild black hole. The approximations used are based on assumptions that are satisfied by the (spherically symmetric) Hawking excitations produced during black hole evaporation. Thus, the simplified dynamics in the deep interior region, which will turn out to be analytically solvable (both at the quantum and classical level), offers a toy scenario to analyze key questions of black hole evaporation in a controlled scenario. We will show in this section that there is natural polymer quantization of the deep interior dynamics with remarkable simple properties such as: (like in the full theory of LQG) the discreteness of the area of constant area surfaces, a well-defined quantum dynamics across the singularity, an effective classical description, and a direct (to our knowledge novel) link with the continuum representation. The polymer quantization we propose does not suffer from the usual ambiguities associated with the so-called holonomy corrections [
35], as the Hamiltonian constraint evolution has (in our simple model) a clear-cut geometric interpretation that allows for a unique polymerization that is compatible with the continuum limit (defined by the Schroedinger representation). This special geometric property arises from the fact that the Hamiltonian constraint is linear in the momentum conjugate to the area of the spheres whose spectrum is discrete in our polymer representation. Ambiguities remain in the form of the so-called inverse volume corrections, which are necessary if one defines the quantum dynamics across the singularity.
3.1. Sketch of the Schrodinger quantization
In the standard Schrodinger representation, one would quantize the phase space of
Section 2.1 by promoting the variables
to self adjoint operators:
in the kinematical Hilbert space is
, equipped with the usual inner product
where we have chosen the momentum representation for the scalar field for convenience (as
is one of the constants of motion of the system). Eigenstates of the
operator are interpreted as distributions (they are not in the Hilbert space), and one usually writes
with
and form an orthonormal basis
The dynamics is imposed by solving the Hamiltonian constraint (
33), which, in the present representation, takes the precise form of a Schrodinger equation in the area variable
a, namely,
As usual, solutions of the constraint are certainly not square-integrable in the
a-direction; thus, physical states are outside of the kinematical Hilbert. The physical Hilbert space is defined as the space of square integrable functions of
m and
at fixed time
a—
—with inner product
which is preserved; i.e., it is independent of
a, by the Schrodinger Equation (evolution is unitary in
a).
Two important remarks are in order: First, note that we are formulating in detail the dynamics of the system in the near singularity approximation. The physical reason for this is that (as argued previously) it is only in this approximation that the system can be compared with a (spherically symmetric) black hole with spherically symmetric excitations falling inside. A side gain is also the simplification of the dynamics which will allow us a simpler quantization and the analysis of the possibility of a well-defined dynamics across the singularity when we undergo the LQG inspired quantization. One could, however, consider the quantization of the minisuperspace system without the near-singularity approximation. In that case, one would need to write a Schroedinger equation using the Hamiltonian (
26), now genuinely time-dependent (
r-dependent), for which unitary evolution would involve path ordered exponentials (as the Hamiltonian does not commute with itself at different
r values). In addition, one would need to work with either
variables or
variables without the luxury of the simplifications introduced by the use of the near singularity variables (
30).
3.2. The Polymer Quantization
We define now a representation of the phase-space variables that incorporates a key feature of the full theory of LQG: the area quantization. Such representation closely mimics the structure of the quantum theory in the fundamental theory in such a way that the area variable
a acquires a discrete spectrum. Mathematically, this is achieved by replacing the
structure of the inner product in the variable
by the inner product of the Bohr compactification of the
phase-space dimension. More precisely, one substitutes the kinematical inner product in the Schroedinger representation (
47) by
With this inner product, periodic functions of
with an arbitrary period are normalizable, and the conjugate
a-representation acquires the property of discreteness in a way that closely mimics the structure of the fundamental theory of loop quantum gravity [
11]. In particular, eigenstates of
exist
with
. These states form an orthonormal basis with inner product
in contrast with (
49). Discreteness of the spectrum of
comes at the prize of changing the kinematical Hilbert space structure in a way that precludes the infinitesimal translation operator
from existing. Instead, only finite translations (quasi periodic functions of
) can be represented as unitary operators in the polymer Hilbert space. Their action on the
a-basis is given by
Eigenstates of the finite translations (or shift operators) exist and are given by wave functions supported on discrete
a-lattices. Namely,
where the parameter
. The discrete lattices denoted
are the analogs of the spin-network graphs in LQG with the values of
a on lattice sites being the analogs of the corresponding spin labels. With all this, one has (using (
55)) that
Note that, unlike the Schroedinger representation where the eigen-space of the momentum operator is one dimensional, the eigen-spaces of the translation operator (labeled by the eigenvalue
) are infinite dimensional and nonseparable. This is explicit from the independence of the eigen-values of the continuous parameter
labeling eigenstates. Such huge added degeneracy in the spectrum of the shift operators is a general feature of the polymer representation. We will show that this degeneracy can show up in Dirac observables of central physical importance such as the mass operator in
Section 3.7.
3.3. Quantum Dynamics
The dynamics is dictated by the quantization and imposition of the constraint (
33). As the operator corresponding to
does not exist in our kinematical Hilbert space, we introduce a polymerized version. Traditionally, this is achieved by replacing the infinitesimal translation operator
The rule consists of making some ’minimal’ substitution of
by a periodic regularization satisfying that in the limit
the functional choice will approximate the original function. Such rule is intrinsically ambiguous, and it opens in general the door for an infinite set of possibilities. Such choices are to be interpreted as quantization ambiguities of the Hamiltonian constraint with potential quantitative dynamical consequences (for a general discussion see [
36]). Dynamical implications of these ambiguities can be analyzed in detail in simple models of quantum cosmology [
35] and black holes [
37].
In the full theory, a new perspective on the regularization issue has been introduced, motivated by the novel mathematical notion of generalized gauge covariant Lie derivatives [
38] and their geometric interpretation, allowing for the introduction of a natural regularization and (subsequent) anomaly-free quantization of the Hamiltonian constraint [
39]. Even when the procedure does not eliminate all ambiguities of quantization (choices are available in the part of the quantum constraint responsible for propagation [
40]), the new technique reduces drastically some of them in the part of the Hamiltonian that is more stringently constrained by the quantum algebra of surface deformations.
What we want to emphasize here is that the analogous procedure in the case of our symmetry-reduced Hamiltoinian has a similar effect. Indeed, because our classical Hamiltonian constraint is linear in the variable
(whose associated Hamiltonian vector field has a crystal-clear geometric interpretation of infinitesimal translations in
a), one has an unambiguous choice of quantization: the obvious choice to replace infinitesimal translations (which do not exist in the polymer representation) by finite translations or shifts. From this geometric perspective, the right polymerization is the regularization that makes the replacement
In other words, the differential time evolution in the Schrodinger equation must be represented in the polymer Hilbert space by a finite translation with a polymerization scale
. However, as such an action is associated with a clear geometric meaning, the geometric compatibility with the Schrodinger equation can be preserved if the second term in the classical Hamiltonian (
33) is exponentiated too in order to produce the well-known unitary evolution operator that produces finite-area evolution. Disregarding for the moment quantum corrections that will have to be included near the
region (see
Section 3.5), the quantum constraint is taken to be
whose action is well-defined in the polymer representation and whose solutions are easily found (by acting on the left with
) to be wave functions satisfying the discrete dynamics given by
The physical Hilbert space is defined via the usual inner product at fixed (discrete) time
a via
which is independent of the lattice sites as required (a property that we could identify with the intrinsic unitarity of the quantum constraint kernel). More precisely, the physical inner product is a constant of the quantum motion associated with the full history represented by the lattice
, as implied by unitarity. Explicitly, one has
Ambiguities of regularization that are usually associated with the polymerization procedure are thus completely absent in this model. The reason is the linear dependence of the Hamiltonian constraint in the polymerized variable which allows for a regularization fixed by the geometric interpretation of the classical Hamiltonian vector field associated with the corresponding variable. However, ambiguities remain when one studies the evolution across the would-be-singularity of the Kantowski–Sachs model at . We will study this in the next section.
Now we would like to concentrate on the evolution when we are away from the
. In such regime, the one step evolution (
61) can be composed to produce the arbitrary initial to final-area evolution
for arbitrary integers
. Evolving across the
point will be discussed later.
3.4. The Continuum Versus the Polymer Dynamics
The polymer dynamics that arises from the geometric action of the quantum constraint (
60) enjoys the appealing feature of being closely related to the dynamics that one would obtain in the continuum Schroedinger representation. This statement can be made precise as follows: any solution of the Schroedinger equation (
50) induces on any given lattice
a solution of (
60). Conversely, physical states of the polymer theory represent a discrete sampling of the continuum solutions of (
50). However, the Schroedinger evolution is ill-defined at the singularity
due to the divergence of the
factor in front of the second term of (
50). The polymer representation allows for a well-defined evolution across the singularity thanks to the deviations from the
behavior introduced by the analog of the ’inverse-volume’ corrections (see next section). Nevertheless, with the appropriate modification of the
factor in the Schroedinger equation, the correspondence between the discrete and continuum solutions continues to hold.
3.5. Quantum Evolution across the Classical Singularity
Quantum evolution in
a for all values of
a, including the singularity, is dictated by the quantum corrected version of the constraint (
60) given by
where
denotes the quantum corrected expression for the operator
(the analog of inverse volume correction in cosmology) that can be implemented in various ways due to inherent ambiguities associated with the polymer quantization. In general, this will give deviations of the
behavior in the region
. This modifies the integral of
in a way characterized by the (to a large extend arbitrary [
35]) function
introduced in the second line. One among the many possibilities is the one that follows from the so-called Thiemann’s trick whose most elementary form is (see [
11] for its application in cosmology; see [
35] for a discussion of the multiplicity of variants)
Integration leads to the following
function in (
65)
whose graph is presented in
Figure 3.
The classical theory does not fix the evolution uniquely in this high curvature regime where quantum geometry effects cannot be neglected. Quantum geometry effects regularize the dynamics near the
singularity; one way of seeing this is that the factor
in the quantum evolution away from the singularity in (
60) receives inverse volume quantum geometry corrections. As discussed in [
35], these corrections are ambiguous (a fact that should not be surprising, given the expectation that the classical theory cannot guide us all the way to the deep UV in QFT). Instead of proposing one particular UV extension, as in the example shown where Thiemann regularization was used, one might simply keep all possibilities open and assume that the corresponding operator is regularized in the relevant region by some arbitrary function
. In regions where
, the quantum evolution leads to semiclassical equations that match exactly Einstein’s equations in the KS sector (more details in
Section 3.6).
3.6. Dynamics of Semiclassical States and Tunneling across the Singularity
Let us first consider the vacuum
case. This case is important because it should correspond to the Schrwarzschild interior in the region
(or
). The dynamical evolution (
65) becomes
In the
representation, the previous equation simply reads
i.e., a simple translation in momentum space. Such dynamical evolution in
a implies in an obvious manner that semiclassical states peaked at classical values
at area
with some given fluctuations will be simply evolved into the translated state with the very same fluctuation properties peaked at
at area
a. This is a remarkable property of the quantum system: independently of the quantum gravity effects encoded in the precise form of
, expectation values satisfy well-defined effective dynamical equations which are exact (not an approximation), and semiclassical states are not spread by the dynamical evolution (as in the simple case of the harmonic oscillator). Notice that the validity of effective dynamical equations in the context of black hole models have remained a conjecture in other formulations [
41].
In regions where
, the previous evolution gives
From the definitions (
30), we have that
and
. The constraint (
25) gives us
and the metric becomes
which corresponds to the Schrwarzschild solution with the two Dirac observables
M and
given in terms of the initial conditions by
When quantum inverse volume corrections are taken into account, then the quantum evolution is perfectly well-defined across the classical singularity. The evolution of the mean values of a semiclassical state is also well-defined and given by
Such solutions can also be obtained from the Hamiltonian constraint, given that one replaces the operator
with its quantum regularization (effective equations are in this sense exact). The metric for all values of
but otherwise arbitrary becomes
Since the Thiemann regularization produces a near , we see that the previous metric is just given by a two dimensional Minkowski metric fibrated with two dimensional sphere with time-dependent radius r. The shrinking of the spheres leads to a singularity at , where the spheres collapse and the spacetime geometry (as described by the effective line element) becomes a two-dimensional flat one at the singularity in the a-t plane. Despite the singular nature of the effective metric, the fundamental quantum evolution is well-defined across the singularity.
In the presence of matter, the situation is a bit more involved due to the factor
appearing in matter’s contribution to the Hamiltonian constraint (
60). However, in the spirit of applying this analysis to macroscopic black holes and modeling the dynamics of a weak scalar excitation (a Hawking particle) falling into the singularity, it is natural to focus on semiclassical states (Gaussian) peaked on values such that
with fluctuations
. As in the vacuum case
, i.e.,
, is a constant of motion and its spread,
. The dynamics of the conjugate variable (the mean value
of the variable
) can be evaluated using stationary phase methods, and the result is
which can be seen to correspond to the classical solutions found in
Section 2.3. Notice that, as expected from the form of the matter coupling, there are here quantum corrections characterized by terms proportional to
. The spread in the variable
is not time-independent if we take into account higher order corrections, namely,
The previous equations are derived assuming that the scalar field is in an eigenstate of the momentum
. This is an idealization that simplifies the analysis of the dynamical evolution of the geometry. Similarly, if we assume that the geometry state was in an eigenstate of
m, then we can easily analyze the dynamics of the scalar field assuming that it is initially in a Gaussian semiclassical state picked about
and
. In accordance with the classical solutions, we get
One way to quickly derive these equations by inspection is to realize that the Hamiltonian constraint (
33) is that of a nonrelativistic point particle with mass proportional to our geometric variable
m evolving in
. Note that (
77) implies that the back-reaction of the scalar field enters only through a simple modification of the exponential conformal factor in front of the 2-metric in the
a-
t ‘plane’ in Equation (
76).
Unlike the geometry degrees of freedom, the fluctuations in the scalar field grow as one approaches the would-be singularity: for a given geometry semiclassical state picked around the mass
M, the spread of the scalar field
in an initial eigenstate of
at area
a grows to a maximum value close to the would-be singularity such that
which are small in the interior if we take
, as expected from the appearance of
in the fundamental commutation relations [
42]. Note that, if we take into account inverse volume corrections of the type suggested by LQG (see
Figure 3), the scalar field reaches a critical point at
where its area velocity vanishes.
3.7. The Mass Operator (in the Vacuum Case)
In this section, we concentrate on the vacuum case for simplicity, as the mass can be directly read off the form of the metric, in this case via a simple comparison with the classical Schwarzschild solution. In this case, the result is
where we used the vacuum solution (
16) (in its
approximation) and (
30). It is easy to verify that the previous is indeed a Dirac observable by showing that it commutes with Hamiltonian constraint (
33). Its nonlinear dependence on the basic variables anticipates factor-ordering ambiguities when it comes to promoting the mass to a quantum operator. Here, we focus on the choice
where
. The eigenstates equation
turns into the differential equation
If we expand the eigenstate in the
basis,
where the sum runs over the discrete lattice
as defined in (
56) when introducing the dynamical constraint (
60). This differential eigenvalue equation is solved by
where
is a Bessel function. One can explicitly verify that the quantum dynamics (
60) preserves the eigenstates by explicitly showing that the evolution between arbitrary lattice points
sends the wave function of the eigenstate at the
lattice point to the
lattice point (as expected for a Dirac observable); or equivalently, the eigenstates of the mass are physical states solving (
60). Explicitly,
Now, the evolution across
requires inverse volume corrections, which modifies the previous dynamical law by replacing
. The mass Dirac observable still exists once inverse volume corrections are turned on. It corresponds to the modification of (
82) via the substitution
. Eigenstates are also obtained by the same substitution in (
86) and satisfy the expected Dirac observable condition (which now holds for lattice points at different sides across the singularity)
When supported on the same lattice, one can show that they satisfy the orthogonality relation
where the inner product is computed with the physical inner product (
62). Thus, the spectrum of the mass operator is continuous. It was argued in the context of the full LQG theory in [
28,
43,
44] that the eigenspaces of the mass should be infinitely degenerate due to the underlying discrete structure of the fundamental theory and the existence of defects that would not be registered in the ADM mass operator. Interestingly, the conjectured property is illustrated explicitly in our simple toy model, as for the eigenvectors (
85) for a given eigenvalue,
M, there are infinitely many and labeled by a continuum parameter. More precisely, they are associated with wave functions of the form (
86) supported on lattices with different values of
. Thus, eigenstates of the mass should then be denoted
with orthogonality relation
where
is the Kronecker delta symbol. The existence of such a large degeneracy is a generic feature of the polymer representation. Even when this is a toy model of quantum gravity, this feature is likely to reflect a basic property of the representation of the algebra of observables in the full LQG context. Here, we are showing that the mass operator is hugely degenerate, suggesting that the usual assumption of the uniqueness of the vacuum in background-dependent treatments of quantum field theory might fail in a full loop quantum gravity context.
Alternative factor orderings of the quantum operator M could be treated similarly (some simple choices lead to slightly different eigenvectors written also in terms of Bessel functions). Such an ambiguity is not relevant for our purposes (and it does not change the key fact that the spectrum of M is infinitely degenerate), as the aim of the model is not to construct any quantitative physical prediction but rather to use it as a toy model to investigate possibly sufficiently generic features that could actually survive in the full theory. The large degeneracy of the mass spectrum is, in our view, an interesting example of one such feature.