2. Appearance of Iterated Function Systems in Black Hole Quantum Physics and Bohr-Like Black Hole
Researchers in quantum gravity have the intuitive, common conviction that, in some respects, BHs are the fundamental bricks of quantum gravity in the same way that atoms are the fundamental bricks of quantum mechanics [
11]. This similarity suggests that the BH mass should have a discrete spectrum [
11]. On the other hand, the analogy generates an immediate and natural question: if the BH is the nucleus of the gravitational atom in quantum gravity, what are the electrons? One of us (Christian Corda) gave an intriguing answer to that question, showing that the BH quasi-normal modes (QNMs) triggered by the emission of Hawking quanta and by the potential absorptions of neighboring particles can be considered as the electrons of that gravitational atom [
1,
2,
3]. Thus, the intuitive picture is more than a picture as QNMs can be really interpreted in terms of BH quantum levels in a BH model somewhat similar to the semi-classical Bohr model of the structure of a hydrogen atom [
1,
2,
3]. This issue has important consequences on the BH information puzzle [
12]. In fact, showing BHs in terms of well defined quantum mechanical systems, having an ordered, discrete quantum spectrum, looks consistent with the unitarity of the underlying quantum gravity theory and with the idea that information should come out in BH evaporation [
1,
2,
3]. A fundamental feature of the Bohr-like BH model that we are going to resume is the discreteness of the BH horizon area as the function of the QNMs principal quantum number, which is consistent with various models of quantum gravity where the spacetime is fundamentally discrete [
13]. We also stress that, in our knowledge, the first who viewed BHs as similar to gravitational atoms was Bekenstein [
11]. In [
1,
2,
3], it has been indeed shown that the semi-classical evaporating Schwarzschild BH is somewhat similar to the historical semi-classical model of the structure of a hydrogen atom introduced by Bohr in 1913. The results in [
1,
2,
3] are founded on the non-thermal spectrum of Parikh and Wilczek [
14], which implies the countable character of subsequent emissions of Hawking quanta enabling a natural correspondence between Hawking radiation [
15] and the BH quasi-normal modes (QNMs) triggered by the emissions of Hawking quanta and by the potential absorptions of neighbouring particles. In such an approach, those QNMs represent the “electron” which jumps from a level to another one. The absolute values of the QNMs frequencies triggered by emissions (Hawking radiation) and absorption of particles represent, in turn, the energy “shells” of the gravitational hydrogen atom [
1,
2,
3]. Remarkably, the time evolution of BH evaporation is governed by a time-dependent Schrodinger equation and represents an independent approach to solve the BH information puzzle [
2,
3]. The results in [
1,
2,
3] are also in perfect agreement with previous existing results in the literature, starting from the famous result of Bekenstein on the area quantization [
16]. Using Planck units
for large values of the principal quantum number
n (i.e., for excited BHs), the energy levels of the Schwarzschild BH which is interpreted as gravitational hydrogen atom are given by [
1,
2,
3]
where
M is the initial BH mass and
is interpreted like the total energy emitted when the BH is excited at the level
n [
1,
2,
3]. During a quantum jump a discrete amount of energy is radiated and, for large values of
n, the analysis becomes independent of the other quantum numbers, in complete consistence with Bohr’s Correspondence Principle [
17], which states that transition frequencies at large quantum numbers should equal classical oscillation frequencies. In Bohr’s model electrons can only gain and lose energy by jumping from one allowed energy shell to another, absorbing or emitting radiation with an energy difference of the levels according to the Planck relation (in standard units)
where
is the Planck constant and
the transition frequency. In the analysis in [
1,
2,
3] QNMs can only gain and lose energy by jumping from one allowed energy shell to another, absorbing or emitting radiation (emitted radiation is given by Hawking quanta) with an energy difference of the levels according to [
1,
2,
3]
Equation (
2) represents the jump between the two levels
and
due to the emission of a particle having frequency
where
is the residual mass of the BH excited at the level
n, that is the original BH mass minus the total energy emitted when the BH is excited at the level
n [
1,
2,
3]. Thus,
[
1,
2,
3]. Then, the jump between the two levels depends only on the initial BH mass and on the correspondent values of the BH principal quantum number [
1,
2,
3]. In the case of an absorption one gets instead [
1,
2,
3]
The similarity with Bohr’s model is completed if one notes that the interpretation of Equation (
3) is of a particle, the electron, quantized on a circle of length [
1,
2,
3]
which is the analogous of the electron traveling in circular orbits around the hydrogen nucleus, similar in structure to the solar system, of Bohr model [
1,
2,
3].
The analysis in [
1,
2,
3] permits to show that the famous formula of Bekenstein–Hawking entropy [
15,
18] is a function of the QNMs principal quantum number, i.e., of the BH quantum level [
3]
before the emission and
after the emission respectively.
On the other hand, it is a general belief that there is no reason to expect that Bekenstein–Hawking entropy will be the whole answer for a correct quantum gravity theory [
3]. For a better understanding of BH entropy we need to go beyond Bekenstein–Hawking entropy and identify the sub-leading corrections [
3]. Using the quantum tunneling approach one obtains the sub-leading corrections to the third order approximation [
19]. In this approach BH entropy contains four parts: the usual Bekenstein–Hawking entropy, the logarithmic term, the inverse area term and the inverse squared area term [
19]
In this way, the formulas of the total entropy that takes into account the sub-leading corrections to Bekenstein–Hawking entropy become
before the emission, and
after the emission, respectively. Thus, also the total BH entropy results a function of the BH excited state
Here we improve the result in [
3] where only the second order approximation has been taken into account. We stress that the present results are in perfect agreement with existing results in the literature. In fact, as we consider large
it is
, see [
3,
20] and references within. Thus, if one neglects the difference between the original BH mass and the residual mass
i.e.,
the Bekenstein–Hawking entropy reads (
and
)
which is consistent with the standard result, see [
3,
20] and references within,
Again, the consistence with well known and accepted results
cannot be a coincidence, but it is a confirmation of the correctness of the current analysis instead. Then, the total entropy reads
which is well approximated by
Also Equations (
12) and (
13) improve the results in [
3] where only the second order approximation has been taken into account. Now, let us explain the way in which the Bohr-like BH model works following [
3]. Let us consider a BH’s original mass
After an emission from the ground state to a state with large
or, alternatively, after a certain number of emissions (and potential absorptions as the BH can capture neighboring particles), the BH is at an excited level
and its mass is
where
is the absolute value of the frequency of the QNM associated to the excited level
. We recall again that
is interpreted as the total energy emitted at that time [
1,
2,
3]. The BH can further emit an energy to jump to the subsequent level:
. Now, the BH is at an excited level
n and the BH mass is
The BH can, in principle, return to the level
by absorbing an energy
. In [
1,
2,
3]. it has been also shown that the quantum of area is
the same for both absorption and emission and it is given by
which is exactly the original result of Bekenstein [
16]. Again, we stress that the Bohr-like BH model has important implications for the BH information paradox see [
12]. In fact, the results in [
1,
2,
3] show that BH QNMs are really the BH quantum levels in our Bohr-like semi-classical approximation. The time evolution of the Bohr-like BH obeys a
time dependent Schrodinger equation for the system composed by Hawking radiation and BH QNMs see [
2,
3]. Such a time evolution enables
pure quantum states to evolve in
pure quantum states, while subsequent emissions of Hawking quanta are entangled with BH QNMs [
2,
3]. On the other hand, consistence between the Bohr-like BH model and a recent approach to solve the BH information paradox [
21] has been recently highlighted in [
22]. Thus, the general conviction that BHs result in highly excited states representing both the “hydrogen atom” and the “quasi-thermal emission” in an unitary theory of quantum gravity is in perfect agreement with the Bohr-like BH model which seems to approach the final theory of quantum gravity in the same way the Bohr model of hydrogen atom approached the final theory of quantum mechanics.
Appearance of the logarithmic term in Equation (8) implies to the congruence of BH entropy with the metric entropy of a function
from a state space
X to itself for a fixed
n with
. The metric entropy can not work when we want to consider different states as a whole. More precisely, the BH entropy depends on
n. The problem is finding a suitable mathematical model to consider all the
n-states with
as a system. Our mathematical suggestion for considering this situation is an iterated function system (
)
where
X is a compact metric space. In the classical case we work with autonomous systems, i.e.,
, but our suggested model is a non-autonomous system.
It is clear that, an iterated function system creates a multifunction with finite range [
23].
The orbit of corresponding to a sequence with is the sequence , where and .
Let
be an open cover for a compact topological space
X. Then we define:
where
and
for
and
. It is clear that for each
,
is an open cover for the space
X.
If is the number of sets in with the smallest cardinality (the number of the members) which covers the space X, then .
We use of the following lemma.
Lemma 1. For a given open cover α we have . Moreover if is an onto map for some , then .
Proof. Let be a subcover of for X. Then is a subcover of . So . Now, let be an onto map, and be a subcover of . Then is a subcover of . Hence . ☐
For two open covers
and
, we define
and
If and are two open covers for the space X, then the open cover is called a refinement of if each member of is a subset of a member of . In this case we write .
Let
be a member of
. Then
. Thus
for each two covers
. So we have
It is not necessary that
(see Example 1). Similarly we have
, for each
, and
Thus
for every finite covers
.
The following example shows that the converse of the Inequality (
17) is not always true.
Example 1. Consider the IFS where and are defined by and . If and on X, then we havebut Lemma 2. exists.
Proof. Consider the sequence
which
for all
. Then for each
we have
So
Thus
is a subadditive sequence [
7]. Hence we have
. ☐
Now we define the topological entropy of , based on the open covers of X.
Definition 1. We define the topological entropy of relative to α by:and the topological entropy by This is well known that topological entropy is an invariant of topological conjugacy. Now we define topological conjugacy for iterated function systems and in Theorem 1 we prove the same result for iterated function systems.
Let and be two compact topological spaces and be a finite set. If and are two , then we say that is topologically conjugate to if there is a homeomorphism such that , for all .
Remark 1. Let α be an open cover for X and let be an onto continuous map. Then (Remark 5, Chapter 5, [7]). Theorem 1. With the above assumptions, if and are topologically conjugate then
Proof. Since
, for all
, then by Remark 1 we have
Hence Similarly, by replacing with we have . So ☐
Theorem 2. Let be an , and let be homeomorphisms. Then , where the is defined by: Proof. So . Similarly we have . Thus . ☐
3. Metric Entropy
Let
be an
with continuous maps
. For a given
, we define a metric
on
X by:
where
,
and
.
A neighborhood of
x with the radius
with respect to
is:
where
,
and
is a neighborhood of
x with the radius
with respect to
d.
Let K be a compact subset of X. A subset E of K is called -separated if for each we have or . denotes the largest cardinality of -separated sets of K.
A subset W of X is called -spanning set for a compact subset K, if for every there is a with . denotes the smallest cardinality of -spanning sets of K.
Now we present the notion of metric entropy for .
Definition 2. The metric entropy of an is: Next theorem shows that the metric entropy and the topological entropy of an are equal.
Theorem 3. If is an on the compact metric X, then .
Proof. Suppose that is a finite open cover for X and , where . Let E be an -separated set with the cardinality and let be two distinct members of E. Since then can not lay in the one member of , so . Hence .
Now we prove
. Let
be an open cover of
X with the Lebesgue number
. For an
-spanning set
W with the cardinality
we have
where
,
. Since for
there exists a member
such that
then
Hence This implies that . ☐
We write .
It is well known that for every continuous map , the power rule for its entropy holds, i.e., for any positive integer m. By Theorem 4 we prove a similar result for .
Definition 3. If is an [24]. Then we define the by:where for all and . Lemma 3. Every -spanning set of an , is a -spanning set for the .
Proof. Let
W be an
-spanning set for an
. Then for every
,
we have
where
,
and
.
Hence
where
,
,
and
. So
W is an
-spanning set for the
. ☐
Theorem 4. Suppose that is an , where are continuous maps, and , then .
Proof. By Lemma 3 each -spanning set of an is an -spanning set of . So . Thus
Now, we prove the other inequality. Since each
is continuous and
X is compact, then for
there is a
such that
for all
. If
E is a
-spanning set for
K with respect to the
, then for each
there is
such that
where
and
. So
for every
,
. Hence every
-spanning set of
K with respect to the
is a
-spanning set of
K with respect to the
. Therefore
, this implies that
. Thus
☐
This is well known that if are two continuous functions then . Now we consider the product of two and prove the similar property.
Definition 4. Suppose that and are two compact metric spaces, and , are two . Then the product of is defined by:where and . Additionally, is a compact metric space, where Theorem 5. Let and be two compact metric spaces, and , are two . Then Proof. Consider
and
as two
-spanning sets for
and
respectively. For each
there are
such that
and
where
,
and
,
and
. If we take
with
then
where
,
,
,
and
. Hence
is an
-spanning set for the
and consequently
. Thus
. If
and
are
-separated subsets of
X and
Y respectively, then
is an
-separated subset of
. Thus
, and we have
. ☐
In the following example we compute topological entropy for an .
Example 2. Suppose that . By using of Formula (18) we havewhere , and . Now suppose K is a compact subset of with and E is an -separated subset of K. Since , then , and , hence . Similar calculations imply that if , then , where .
4. Average Entropy
In this section we present another method to define the entropy of an .
Let X be a topological space, and let be an on X, where , are distinct continuous maps. A typical element of can be denoted by and we use the notation for . and we denote the set of by .
Definition 5. For an , , and we define the topological entropy of by:where , and is the topological entropy of the . Lemma 4. Let and be two sequences of positive numbers. Then Theorem 6. Suppose that is an then for every Proof. Since for all
we have
, then
☐
Example 3. Let and let Put and , .
The following theorem is the main result of this paper. It gives a new type of topological entropy for an , based on the usual topological entropy of .
Theorem 7. Let be an , then Proof. Let
be an
with distinct continuous maps
. Define the maps
by:
We define
by:
This means that for each
,
. So
If where then . Now we claim that
To prove this claim, take
, where
and
. For every
and
, there exist
such that:
If
, then
. Thus
where
Hence, . Since for each we have , then . Thus . Hence the claim is proved and ☐
The method of the proof of Theorem 7 yields that for any there is at least such that , where and .
This fact and the above theorem motivate us to present the following definition.
Definition 6. Let be an . Then we define the entropy of by Theorem 8. If is topologically conjugate to , then
Proof. Let
be a homeomorphism such that
, for all
. Then
☐
Theorem 9. If and are two compact metric spaces, and if , are two , then Theorem 10. Let be an on a compact topological space X then
- (a)
If are homeomorphisms, then
- (b)
.
Proof. (a)
| Theorem 7.3 of [7] |
(b) Suppose that
where
for
and
, then
| Theorem 7.10 of [7] |
☐
Example 4. Suppose X is the unit interval [0,1]. We consider the 2-ary expansion 0. for and let be the shift map, then (Theorem 7.12, [7]). In addition, if is the tent map, then (Example 13, [25]). Thus for the we have . Example 5. In Example 2 with we have . By a similar method one can show that and where and . Thus
4.1. Non Wandering Sets and Topological Entropy of IFS
In this section we restrict ourself to on a compact manifold M which all of it’s functions are Axiom A diffeomorphisms.
A diffeomporphism is an Axiom A diffeomorphism if
- (a)
is a hyperbolic set and
- (b)
the periodic points of f are dense in , where is the set of non-wandering points of f.
We recall that, a point is called a non-wandering point if for each neighborhood U of x there is an such that .
Theorem 11. Let satisfies Axiom A then there is a neighborhood of f such that for every have [26]. Now we extend this theorem for iterated function systems.
Theorem 12. Let be an such that for every , and it satisfies Axiom A. Then there is an such that for each , with for every , where , we have
Proof. In Theorem 11 there are
with
implies that
,
. We choose
. Then
, where
. So
☐
Theorem 13. Let be an such that for every , and satisfies Axiom A and let σ be an arbitrary sequence in . Then there is an such that for every , with for we have .
Proof. We assume that
and
where
. Consider the number
and the
,
as in the proof of Theorem 12. So
with
. Thus
where
, and
is the topological entropy of the
. Since
, for every
and
, then
☐