0% found this document useful (0 votes)
22 views18 pages

Date Key Words and Phrases

This document discusses applying results from canonical quantum gravity to cosmology. Specifically: 1. It quantizes gravity for spatially unbounded hyperbolic spacetimes and applies this to Friedmann universes. 2. It examines how this approach can provide insights into dark energy density, inflation, dark matter, and the missing antimatter problem. 3. The key equations derived are a wave equation and its equivalent temporal and spatial self-adjoint operator forms, which can model aspects of the early universe when applied to functions on a constant curvature Cauchy hypersurface.

Uploaded by

alex moreno
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
0% found this document useful (0 votes)
22 views18 pages

Date Key Words and Phrases

This document discusses applying results from canonical quantum gravity to cosmology. Specifically: 1. It quantizes gravity for spatially unbounded hyperbolic spacetimes and applies this to Friedmann universes. 2. It examines how this approach can provide insights into dark energy density, inflation, dark matter, and the missing antimatter problem. 3. The key equations derived are a wave equation and its equivalent temporal and spatial self-adjoint operator forms, which can model aspects of the early universe when applied to functions on a constant curvature Cauchy hypersurface.

Uploaded by

alex moreno
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
Download as pdf or txt
You are on page 1/ 18

APPLICATIONS OF CANONICAL QUANTUM GRAVITY

TO COSMOLOGY
arXiv:1908.02145v1 [gr-qc] 3 Aug 2019

CLAUS GERHARDT

Abstract. We apply quantum gravitational results to spatially un-


bounded Friedmann universes and try to answer some questions related
to dark energy, dark matter, inflation and the missing antimatter.

Contents
1. Introduction 1
2. The dark energy density 6
3. The inflationary period 9
4. The dark matter 10
5. The missing antimatter 15
6. Spherically symmetric eigenfunctions in hyperbolic space 16
References 17

1. Introduction
The quantization of gravity is one of the most challenging open problems
in physics. The Einstein equations are the Euler-Lagrange equations of the
Einstein-Hilbert functional and quantization of a Lagrangian theory requires
to switch from a Lagrangian view to a Hamiltonian view. In a ground break-
ing paper, Arnowitt, Deser and Misner [3] expressed the Einstein-Hilbert
Lagrangian in a form which allowed to derive a corresponding Hamilton func-
tion by applying the Legendre transformation. However, since the Einstein-
Hilbert Lagrangian is singular, the Hamiltonian description of gravity is only
correct if two additional constraints are satisfied, namely, the Hamilton con-
straint and the diffeomorphism constraint. Dirac [7] proved how to quantize
a constrained Hamiltonian system—at least in principle—and his method has
been applied to the Hamiltonian setting of gravity, cf. the paper by DeWitt
[6] and the monographs by Kiefer [16] and Thiemann [18]. In the general

Date: August 7, 2019.


2000 Mathematics Subject Classification. 83,83C,83C45.
Key words and phrases. quantization of gravity, quantum gravity, Friedmann universe,
dark energy density, dark matter, inflation, missing antimatter, cosmology, negative cos-
mological constant.
1
2 CLAUS GERHARDT

case, when arbitrary globally hyperbolic spacetime metrics are allowed, the
problem turned out to be extremely difficult and solutions could only be
found by assuming a high degree of symmetry, cf., e.g., [8].
However, in the papers [9, 10] we developed a model for the quantization of
gravity for general hyperbolic spacetimes. In these papers we eliminated the
diffeomorphism constraint by reducing the number of variables and proving
that the Euler-Lagrange equations for this special class of metrics were still
the full Einstein equations. The Hamiltonian description of the Einstein-
Hilbert functional then allowed a canonical quantization.
We quantized the action by looking at the Wheeler-DeWitt equation in
a fiber bundle E, where the base space is a Cauchy hypersurface of the
spacetime which has been quantized and the elements of the fibers are Rie-
mannian metrics. The fibers of E are equipped with a Lorentzian metric such
that they are globally hyperbolic and the transformed Hamiltonian, which
is now a hyperbolic operator, Ĥ, is a normally hyperbolic operator acting
only in the fibers. The Wheeler-DeWitt equation has the form Ĥu = 0
with u ∈ C ∞ (E, C) and we defined with the help of the Green’s operator a
symplectic vector space and a corresponding Weyl system.
The Wheeler-DeWitt equation seems to be the obvious quantization of
the Hamilton condition. However, Ĥ acts only in the fibers and not in the
base space which is due to the fact that the derivatives are only ordinary
covariant derivatives and not functional derivatives, though they are supposed
to be functional derivatives, but this property is not really invoked when a
functional derivative is applied to u, since the result is the same as applying
a partial derivative.
Therefore, we discarded the Wheeler-DeWitt equation in the paper [14]
and also in the monograph [15], and expressed the Hamilton condition dif-
ferently by looking at the evolution equation of the mean curvature of the
foliation hypersurfaces M (t) and implementing the Hamilton condition on
the right-hand side of this evolution equation. The left-hand side, a time de-
rivative, we replaced by the corresponding Poisson brackets. After canonical
quantization the modified evolution equation was transformed to an equation
satisfied by operators which acted on functions u ∈ C ∞ (E, C).
Since the Poisson brackets became a commutator we could now employ the
fact that the derivatives are functional derivatives, since we had to differen-
tiate the scalar curvature of a metric when we applied the operator equation
to a smooth function and tried to simplify the resulting equation. As a result
of the simplification of the commutator action we obtained an elliptic differ-
ential operator in the base space, the main part of which was the Laplacian
with respect to a fiber element. Here, we considered functions u depending
on the variables (x, gij ), where x is a point in the base space S0 , x ∈ S0 , and
gij is an element of the fibers. The fiber metrics have the form

4
(1.1) gij = t n σij ,
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 3

where 0 < t < ∞ is a time-like fiber variable, which is referred to as time,


n ≥ 3, is the dimension of S0 and σij is a Riemannian metric, depending only
on x, subject to the requirement
(1.2) det σij = det χij ,
cf. [15, equs. (1.4.103) & (1.4.104), p. 29] and also [15, Remark 1.6.8]. The
arbitrary, but fixed, metric χij in S0 had been introduced to transform the
densities det gij to functions.
On the right-hand side of the evolution equation the interesting term was
H 2 , the square of the mean curvature. It transformed to a second time
derivative, the only remaining derivative with respect to a fiber variable,
since the differentiations with respect to the other variables canceled each
other. The resulting quantized equation is then a wave equation in a globally
hyperbolic spacetime
(1.3) Q = (0, ∞) × S0 ,
of the form
1 n2 4 n 4
(1.4) ü − (n − 1)t2− n ∆u − t2− n Ru + nt2 Λu = 0,
32 n − 1 2
where S0 is a Cauchy hypersurface of the original spacetime and the Laplacian
and the scalar curvature R are formed with respect to a metric σij satisfying
(1.2) and Λ is a cosmological constant. The function u depends on (x, t, σij ).
Since the metric χij is also a fiber metric we may choose σij = χij and
because it is also arbitrary we may set χij to be the original metric of the
Cauchy hypersurface S0 , cf. [15, Remark 1.6.8 on page 49]. The function u
then only depends on (t, x), u = u(t, x). For a detailed derivation of equation
(1.4) we refer to [15, Chapter 1.6] or [14, Section 6].
When S0 is a space of constant curvature then the wave equation, consid-
ered only for functions u which do not depend on x, is identical to the equa-
tion obtained by quantizing the Hamilton constraint in a Friedmann universe
without matter but including a cosmological constant, cf. [15, Remark 1.6.11
on page 50] or [14, Remark 6.11].
There exist temporal and spatial self-adjoint operators H0 resp. H1 such
that the hyperbolic equation is equivalent to
(1.5) H0 u − H1 u = 0,
where u = u(t, x). The operator H0 is defined by
1 n2
(1.6) H0 w = ϕ−1
0 {− ẅ − nt2 Λw},
32 n − 1
4
where w = w(t), w ∈ Cc∞ (R+ , C), and ϕ0 = t2− n , while the definition of H1
is given by
n
(1.7) H1 v = −(n − 1)∆v − Rv,
2
4 CLAUS GERHARDT

where v = v(x), v ∈ Cc∞ (S0 , C). More precisely, the operators Hi , i = 0, 1,


are the corresponding unique self-adjoint extensions of the operators defined
above in the appropriate function spaces.
Assuming Λ < 0 we proved that H0 has a pure point spectrum with
positive eigenvalues λi , cf. [15, Chapter 6.2], especially [15, Theorem 6.2.5 on
page 144], while, for H1 , it is possible to find corresponding eigendistributions
for each of the eigenvalues λi , if S0 is asymptotically Euclidean or if the
quantized spacetime is a black hole with a negative cosmological constant,
cf. [12, 11, 13] or [15, Chapters 3–5], and also if S0 is the hyperbolic space
S0 = Hn , n ≥ 3, cf. Section 6 on page 16.
Let wi , i ∈ N, be an orthonormal basis for the temporal eigenvalue prob-
lems
(1.8) H0 wi = λi wi
and vi be corresponding smooth eigendistributions for the spatial eigenvalue
problems
(1.9) H1 vi = λi vi ,
then
(1.10) ui = wi vi
are special solutions of the wave equation (1.4).
The temporal eigenvalues λi all have multiplicity 1, the spatial eigenval-
ues are the same eigenvalues, but they may have higher multiplicities. In
case of black holes this is caused by very compelling intrinsic mathematical
reasons, cf. [15, Chapter 6.4], but unless there are either convincing intrin-
sic or extrinsic reasons, like data, we choose the spatial eigenspaces to be
one-dimensional, because the spatial eigenvalues belong in general to the
continuous spectrum of the spatial Hamiltonian H1 . If S0 is the Cauchy hy-
persurface of a Friedmann universe we only considered smooth spherically
symmetric spatial eigenfunctions, which also leads to one-dimensional spa-
tial eigenspaces, cf. [15, Chapter 6.6] for the Euclidean case and Section 6 on
page 16 for the hyperbolic case.
One can then define an abstract Hilbert space H spanned by the ui and a
self-adjoint operator H, unitarily equivalent to H0 , such that
(1.11) Hui = λi ui .
e−βH is then of trace class in H for all β > 0 and the canonical extension
of H to the corresponding symmetric Fock space F , which is still called H,
shares this property. Hence, we can define the partition function Z,
(1.12) Z = tr e−βH ,
the operator density
(1.13) ρ̂ = Z −1 e−βH , ∀ β > 0,
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 5

the average energy and the von Neumann entropy in F . The eigenvectors ui
can also be viewed as to be elements of F and they are then also eigenvectors
of ρ̂.
In the present paper we want to apply these quantum gravitational results
to cosmology by looking at a Friedmann universe
(1.14) N = I × S0 ,
where S0 is a n-dimensional simply connected space of constant curvature κ̃,
(1.15) κ̃ ∈ {0, −1},
i.e., S0 is either R or the hyperbolic space Hn , n ≥ 3. We tried to an-
n

swer some questions related to dark energy, dark matter, inflation and the
missing antimatter. In doing so we shall also show that assuming a negative
cosmological constant is not a contradiction to the observational result of an
expanding universe. Usually a positive cosmological constant is supposed to
be responsible for the dark energy and dark matter is sometimes explained
by assuming so-called extended theories of gravity, confer, e.g., the papers
[5] and [4]. In this paper we rely on general relativity combined with some
quantum gravitational ingredients.
Let us summarize the main result as a theorem, where ρdm resp. ρde refer to
the dark matter resp. dark energy densities, which we defined as eigenvalues
of the operator density ρ̂ in F , and ρ3 is a conventional density. Z is the
partition function, T > 0 the absolute temperature and λ0 > 0 the smallest
eigenvalue of the Hamiltonian H.

Theorem 1.1. Let the cosmological constant Λ,


(1.16) − 1 < Λ < 0,
be given and consider the perfect fluid defined by the density
(1.17) ρ = ρdm + ρde + ρ3
satisfying the assumptions (4.23), (4.24), (4.36) and (4.37). Moreover, we
suppose that β = T −1 and the scale factor a are functions depending on t.
The initial value problems
ä κ2 2
(1.18) =− {(n − 2)ρ + np} + Λ
a n(n − 1) n(n − 1)
and
ρdm
(1.19) β̇ = −n ∂
a−1 ȧ.
∂β (ρdm + ρde )
with initial values (β0 , a0 , ȧ0 ) are then solvable in I = [t0 , ∞) provided β0 > 0
is so large that (2.12) on page 7 as well as
2κ2 1 2
(1.20) Z −1 {1 − (n − 2)α0 e−βλ0 } + Λ>0
n(n − 1) 2 n(n − 1)
6 CLAUS GERHARDT

are valid at β = β0 and a0 > 0 has to be chosen such that after adding
κ2 −n(1+ω3 )
(1.21) − (n(1 + ω3 ) − 2)γ3 a0
n(n − 1)
to the left-hand side of (1.20) the inequality still remains valid at β = β0 . The
initial value ȧ0 is supposed to be positive. The solution (β, a) then satisfies
(1.22) β̇ > 0,

(1.23) ȧ > 0,

(1.24) ä > 0
and
2 2
(1.25) κ2 ρ + Λ − κ̃a−2 > 0.
n(n − 1) n(n − 1)
In order that (β, a) also satisfies the first Friedmann equation ȧ0 has to be
chosen appropriately, namely, such that the first Friedmann equation is valid
for t = t0 , which is possible, in view of (1.25).

Remark 1.2. Let us also mention that we use (modified) Planck units in
this paper, i.e.,
(1.26) c = κ2 = ~ = KB = 1,
where κ2 is the coupling constant connecting the Einstein tensor with the
stress-energy tensor
(1.27) Gαβ + Λḡαβ = κ2 Tαβ .

2. The dark energy density


In [15, Remark 6.5.5] we proposed to use the eigenvalue of the density
operator ρ̂ with respect to the vacuum vector η, which is Z −1 ,
(2.1) ρ̂η = Z −1 η,
as the source of dark energy density, and though this eigenvalue is the vac-
uum, or zero-point, energy and many authors have proposed the vacuum
energy to be responsible for the dark energy, these proposals all assumed the
cosmological constant to be positive, while we assume Λ < 0 because of the
spectral resolution of the wave equation, otherwise the temporal Hamiltonian
does not have a pure point spectrum. However, if Λ < 0 then we have to
assure that Z −1 dominates Λ which will only be the case if
(2.2) T < T0 = T0 (|Λ|).
Note that Z depends on the eigenvalues λi and on
(2.3) β = T −1 .
First, we emphasize that we shall treat
(2.4) ρde = Z −1
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 7

as a constant, i.e., we shall define the perfect fluid stress-energy tensor by


(2.5) Tαβ = −ρde ḡαβ .
Let λi > 0, i ∈ N, be the eigenvalues of the temporal Hamiltonian H0 for
a given Λ < 0 and let λ̄i be the eigenvalues for
(2.6) Λ = −1,
then
n−1
(2.7) λi = λ̄i |Λ| n ,
cf. [15, Lemma 6.4.9, p. 172], and define the parameter τ by
n−1
(2.8) τ = |Λ| n ,
where we now assume
(2.9) |Λ| < 1,
throughout the rest of the paper. We proved in [15, Theorem 6.5.6, p. 180]
that
(2.10) lim Z = ∞,
τ →0

or equivalently, that
(2.11) lim ρde = 0.
τ →0

However, we shall now derive a more precise estimate of ρde = Z −1 involving


β and Λ.

Lemma 2.1. For any Λ satisfying −1 < Λ < 0, there exists exactly one
T0 > 0 such that
(2.12) Z −1 (β) > |Λ| ∀ β > β0 = T0−1 ,
where we recall that
(2.13) β = T −1 .

Proof. In view of (2.7) we deduce that


(2.14) Z(β) ≡ Z(β, λi ) = Z̄(γ, λ̄i ) ≡ Z̄(γ),
where
n−1
(2.15) γ = β|Λ| n .
From the relations
∂ log Z ∂ log Z −1
(2.16) 0<E=− = ,
∂β ∂β
cf. [15, equations (6.5.30) and (6.5.32), p. 176],
(2.17) lim Z(β) = 1,
β→∞
8 CLAUS GERHARDT

and
(2.18) lim Z(β) = ∞,
β→0

cf. [15, Theorem 6.5.8, p. 181], we then conclude that there exists exactly
one γ0 such that
(2.19) Z̄ −1 (γ0 ) = |Λ|
and, furthermore, that
(2.20) Z̄ −1 (γ) > Z̄ −1 (γ0 ) ∀ γ > γ0 ,
completing the proof of the lemma. 
Thus, defining the dark energy density by (2.4) and (2.5), we immediately
deduce:

Theorem 2.2. Let T0 be the temperature defined in Lemma 2.1 and as-
sume that the temperature T satisfies T < T0 , then the dark energy density
guarantees that the Friedmann universe with negative cosmological constant
Λ,
(2.21) − 1 < Λ < 0,
is expanding such that
(2.22) ȧ > 0
as well as
(2.23) ä > 0.

Proof. The Friedmann equations for a perfect fluid with energy ρ and pressure
p are
ȧ2 2 2
(2.24) = κ2 ρ + Λ − κ̃a−2
a2 n(n − 1) n(n − 1)
and
ä κ2 2
(2.25) =− {(n − 2)ρ + np} + Λ.
a n(n − 1) n(n − 1)
Choosing ρ = ρde we also specified
(2.26) p = −ρde
yielding
ä 2κ2 2
(2.27) = ρde + Λ.
a n(n − 1) n(n − 1)
Moreover, in our units,
(2.28) κ2 = 1
and we also only consider space forms satisfying
(2.29) κ̃ ≤ 0,
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 9

hence the theorem is proved in view of Lemma 2.1. 

3. The inflationary period


Immediately after the big bang the development of the universe will have to
be governed by quantum gravitational forces, i.e., by the eigenfunctions resp.
eigendistributions of the corresponding temporal and spatial Hamiltonians,
which we have combined to a single Hamiltonian H acting in an abstract
separable Hilbert space H spanned by the eigenvectors ui
(3.1) Hui = λi ui ,
where the eigenvalues all have multiplicity 1, are ordered
(3.2) 0 < λ0 < λ1 < · · ·
and converge to infinity
(3.3) lim λi = ∞.
i→∞
The dominant energies near the big bang will therefore be the eigenvalues
(3.4) λi = hHui , ui i
for large i and we shall assume, when considering the development of a Fried-
mann universe, that this development is driven by a perfect fluid
(3.5) Tαβ = −ρi ḡαβ ,
where
(3.6) ρi = λi .
Looking at the Friedmann equations
ȧ2 2κ2 2
(3.7) 2
= ρi + Λ − κ̃a−2
a n(n − 1) n(n − 1)
and
ä 2κ2 2
(3.8) = ρi + Λ
a n(n − 1) n(n − 1)
we conclude that the universe is expanding rapidly depending on the eigen-
value ρi = λi . The corresponding eigenvector, or particle, ui will decay after
some time and produce lower order eigenvectors or maybe particles that can
be looked at as matter or radiation satisfying the corresponding equations of
state.
After some time the inflationary period will have ended and only the stable
ground state u0 ,
(3.9) Hu0 = λ0 u0 ,
together with conventional matter and radiation will be responsible for the
further development of the Friedmann universe.
n−1
The eigenvalue λ0 is of the order |Λ| n in view of (2.7) on page 7, hence
it will dominate Λ for small values of |Λ|.
10 CLAUS GERHARDT

4. The dark matter


Let ρ̂ be the density operator acting in the Fock space F ,
(4.1) ρ̂ = Z −1 e−βH ,
where we use the same symbol H to denote the self-adjoint operator H in the
separable Hilbert space H as well its canonical extension to the corresponding
symmetric Fock space F+ (H) ≡ F . In Section 2 we defined the dark energy
density ρde by
(4.2) ρde = hρ̂η, ηi = Z −1
and we propose to define the dark matter density by
(4.3) ρdm = α0 hρ̂u0 , u0 i = α0 e−βλ0 Z −1 ,
where u0 is a unit eigenvector of H satisfying
(4.4) Hu0 = λ0 u0
and
(4.5) α0 > 1
an otherwise arbitrary constant. Its presence should guarantee that there
exists β0 > 0 such that

(4.6) (ρdm + ρde ) < 0 ∀ β ≥ β0 ,
∂β
as we shall now prove:

Lemma 4.1. Let α0 satisfy (4.5) and Λ


(4.7) − 1 < Λ ≤ Λ0 < 0,
then there exists β0 = β0 (α0 , |Λ0 |) such that the inequality (4.6) is valid.

Proof. In view (2.16) on page 7 we have



(4.8) (ρdm + ρde ) = −α0 λ0 e−βλ0 Z −1 + α0 e−βλ0 Z −1 E + Z −1 E,
∂β
where
X

λi λ0 X λi

(4.9) E= βλ
= βλ0 + ,
i=0
e i −1 e − 1 i=1 eβλi − 1
cf. [15, equ. (6.5.32), p. 176] or simply differentiate. Hence, we obtain
λ0 eβλ0 X ∞
λi
Eeβλ0 = +
e βλ0 − 1 i=1 eβ(λi −λ0 ) − e−βλ0
(4.10)
λ0 eβλ0 X ∞
λi
≤ βλ0 +
e − 1 i=1 eβ(λi −λ0 ) − 1
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 11

and we conclude
(4.11) lim Eeβλ0 = λ0 ,
β→∞

since
X

λi X

λi − λ0 X ∞
λ0
= +
i=1
eβ(λi −λ0 ) − 1 i=1
e β(λi −λ0 ) − 1 i=1 e β(λi −λ0) − 1

(4.12)
X

µi X ∞
µi
≤ βµ
+ λ0 (λ1 − λ0 )−1 βµ
,
i=1
e i −1 i=1
e i −1
where µi is defined by
(4.13) µi = λi − λ0 ≥ λ1 − λ0 > 0 ∀ i ≥ 1.
Thus the right-hand side of (4.12) is estimated from above by
(4.14) (1 + λ0 (λ1 − λ0 )−1 )E(β, µi )
and
(4.15) lim E(β, µi ) = 0,
β→∞

cf. [15, equ. (6.5.71), p. 181]. Furthermore, we know


n−1
(4.16) λ0 = λ̄0 |Λ| n ,
cf. (2.7). Combining these estimates we conclude that there exists
(4.17) β0 = β0 (α0 , |Λ0 |)
such that
∂ α0 − 1
(4.18) (ρdm + ρde ) ≤ − λ0 e−βλ0 Z −1 ∀ β ≥ β0 .
∂β 2
The limits in (4.11) and (4.15) are also uniform in |Λ| because of (4.7). 
Dark matter is supposed to be dust, i.e., its pressure vanishes, and hence,
ρdm cannot be constant which is tantamount to
(4.19) β 6≡ const,
since we assume that Λ is constant. Thus, ρde is also not constant, though
we still assume that its stress-energy tensor is defined by
(4.20) Tαβ = −ρde ḡαβ .
Therefore, we can only establish the continuity equation for
(4.21) ρdm + ρde
and not for each density separately. Let a dot or a prime indicate differenti-
ation with respect to time t, then the continuity equation has the form
(4.22) (ρdm + ρde )′ = −nρdma−1 ȧ,
because
(4.23) pdm = 0
12 CLAUS GERHARDT

and
(4.24) pde = −ρde .
The left-hand side of (4.22) is equal to

(4.25) (ρdm + ρde )β̇
∂β
and we see that the continuity equation can only be satisfied if
ρdm
(4.26) β̇ = −n ∂ a−1 ȧ.
∂β (ρ dm + ρ de )
From Lemma 4.1 we immediately derive

Lemma 4.2. Let the assumptions of Lemma 4.1 be satisfied and suppose
that ȧ > 0, then, for any solution β = β(t) of (4.26) in the interval
(4.27) I = [t0 , b), t0 < b ≤ ∞,
with initial value
(4.28) β(t0 ) ≥ β0
the inequality
(4.29) β̇ > 0
is valid and hence
(4.30) β(t) ≥ β0 ∀ t ∈ I.
Furthermore, β̇ can be expressed in the form
(4.31) β̇ = nδ(α0 − 1)−1 α0 a−1 ȧ,
where δ = δ(t, β0 ) satisfies
(4.32) 1≤δ≤2
and
(4.33) lim δ = 1,
β0 →∞

i.e.,
(4.34) β(t) − β(t0 ) ≈ nδα0 (α0 − 1)−1 (log a(t) − log a(t0 )).

Proof. (4.29)“ Follows from (4.6) and (4.26).



(4.31)“ To prove the claim we combine (4.8), (4.26) and (4.11).

(4.32)“ and (4.33)“ Same argument as before.
” ”
(4.34)“ Obvious in view of (4.31) and (4.33). 

APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 13

Now, we are prepared to solve the Friedmann equations (2.24) and (2.25)
on page 8 for
(4.35) ρ = ρdm + ρde + ρ3 ,
where ρ3 is a conventional density satisfying the equation of state
(4.36) p 3 = ω 3 ρ3
assuming
(4.37) ω3 > −1.
ρ3 is only added for good measure and we are allowed to assume
(4.38) ρ3 = 0,
since its presence is not essential.
We also emphasize that we have to solve an additional third equation,
namely, equation (4.26). We shall solve the Friedmann equations and (4.26)
in the interval
(4.39) I = [t0 , ∞), t0 > 0,
for the unknown functions (a, β) with prescribed positive initial values
(a0 , ȧ0 , β0 ). β0 can be arbitrary but large enough such that the assump-
tions in Lemma 4.1 and Lemma 4.2 are satisfied. If ρ3 vanishes then a0 > 0
can be arbitrary, otherwise it has to be large enough. The last initial value
ȧ0 > 0 cannot be arbitrary, instead it has to be chosen such that the first
Friedmann equation is initially valid at t = t0 .
If these assumptions are satisfied then we shall solve the equations (2.25)
on page 8 and (4.26). The first Friedmann equation will then be valid auto-
matically. For simplicity we shall only consider the case
(4.40) ρ3 > 0
to avoid case distinctions. Then we deduce, from the continuity equation,
(4.41) ρ3 = γ3 a−n(1+ω3 ) ,
where γ3 > 0 is a given constant.
Let us now prove:

Theorem 4.3. Let the cosmological constant Λ,


(4.42) − 1 < Λ < 0,
be given and consider the perfect fluid defined by the density
(4.43) ρ = ρdm + ρde + ρ3
satisfying the assumptions (4.23), (4.24), (4.36) and (4.37). Moreover, we
suppose that β = T −1 and the scale factor a are functions depending on t.
The initial value problems
ä κ2 2
(4.44) =− {(n − 2)ρ + np} + Λ
a n(n − 1) n(n − 1)
14 CLAUS GERHARDT

and
ρdm
(4.45) β̇ = −n ∂
a−1 ȧ.
∂β (ρ dm + ρde )
with initial values (β0 , a0 , ȧ0 ) are then solvable in I = [t0 , ∞) provided β0 > 0
is so large that (2.12) on page 7 as well as
2κ2 1 2
(4.46) Z −1 {1 − (n − 2)α0 e−βλ0 } + Λ>0
n(n − 1) 2 n(n − 1)
are valid at β = β0 and a0 > 0 has to be chosen such that after adding
κ2 −n(1+ω3 )
(4.47) − (n(1 + ω3 ) − 2)γ3 a0
n(n − 1)
to the left hand side of (4.46) the inequality still remains valid at β = β0 . The
initial value ȧ0 is supposed to be positive. The solutions (β, a) then satisfy
(4.48) β̇ > 0,

(4.49) ȧ > 0,

(4.50) ä > 0
and
2 2
(4.51) κ2 ρ + Λ − κ̃a−2 > 0.
n(n − 1) n(n − 1)
In order that (β, a) also satisfy the first Friedmann equation ȧ0 has to be
chosen appropriately, namely, such that the first Friedmann equation is valid
for t = t0 , which is possible, in view of (4.51).

Proof. By introducing a new variable


(4.52) ϕ = ȧ
we may consider a flow equation for (β, a, ϕ), where ϕ̇ replaces ä and
(4.53) ȧ = ϕ
is an additional equation.
Choosing then β0 , a0 as above and ϕ0 > 0 arbitrary the flow has a solution
on an maximal time interval
(4.54) I = [t0 , t1 ), t1 > t0 ,
because of Lemma 4.1 and Lemma 4.2. It is also obvious that the relations
(4.48)–(4.51) are valid, in view of these lemmata.
Furthermore, if the interval I was bounded, then the flow would have a
singularity at t = t1 which is not possible, in view of the relation (4.34),
which would imply that β, β̇ as well as a and ȧ would tend to infinity by
approaching t1 which, however, contradicts the second Friedmann equation
(4.44) from which we then would infer
(4.55) 0 < ä ≤ ca ∀t ∈ I
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 15

an apparent contradiction. Hence we deduce


(4.56) I = [t0 , ∞).
It remains to prove that the first Friedmann equation is satisfied if ȧ0 is
chosen appropriately, Define
2 2
(4.57) Φ = ȧ2 − { κ2 ρ + Λ}a2 + κ̃,
n(n − 1) n(n − 1)
then we obtain
(4.58) Φ̇ = 0,
in view of the continuity equations and (4.44), yielding
(4.59) Φ(t) = Φ(t0 ) = 0 ∀ t ∈ I.


5. The missing antimatter


In [15, Theorem 4.3.1, p. 110] we proved that a temporal eigenfunction
w = w(t) defined in R+ can be naturally extended past the big bang singu-
larity {t = 0} by defining
(5.1) w(−t) = −w(t), ∀ t > 0.
2,α
The extended function is then of class C ,
2,α
(5.2) w∈C (R),
for some 0 < α < 1 and its restriction to {t < 0} is also a solution of the
variational eigenvalue problem. Hence we have two quantum spacetimes
(5.3) Q− = R− × S0
and
(5.4) Q+ = R+ × S0
2,α
and a C transition between them. If we assume that the common time
function t is future directed in both quantum spacetimes, then the singularity
in {t = 0} would be a big crunch for Q− and a big bang for Q+ and simi-
larly for the corresponding Friedmann universes N∓ governed by the Einstein
equations. No further singularities will be present, i.e., the spacetime N− will
have no beginning but will end in in a big crunch and will be recreated with
a big bang as the spacetime N+ .
This scenario would be acceptable if it would describe a cyclical universe.
However, there are no further cycles, there would only be one transition from
a big crunch to a big bang. Therefore, the mathematical alternative, namely,
that at the big bang two universes with opposite light cones will be created, is
more convincing, especially, if the CPT theorem is taken into account which
would require that the matter content in the universe with opposite time
direction would be antimatter. This second scenario would explain what
happened to the missing antimatter.
16 CLAUS GERHARDT

6. Spherically symmetric eigenfunctions in hyperbolic space


The spatial Hamiltonian H1 is a linear elliptic operator
n
(6.1) H1 v = −(n − 1)∆v − Rv,
2
where the Laplacian is the Laplacian in S0 and R the corresponding scalar
curvature. We are then looking for eigenfunctions or, more precisely, eigendis-
tributions v,
(6.2) H1 v = λv,
such that, for each temporal eigenfunction (λi , wi ) there exists a matching
spatial pair (λi , vi ). The product
(6.3) ui = wi vi
would then be a solution of the wave equation (1.4) on page 3.
If S0 is the hyperbolic space Hn , n ≥ 3, we have
(6.4) R = −n(n − 1)
and, given any temporal eigenvalue λi , we would have to find functions vi
satisfying
n2
(6.5) − (n − 1)∆vi = (λi − (n − 1))vi .
2
We are also looking for spherically symmetric eigenfunctions vi . In hyperbolic
space the radial eigenfunctions, known as spherical functions, are well-known:
For each µ ∈ C there exists exactly one radial eigenfunctions ϕµ of the
Laplacian satisfying
(6.6) − ∆ϕµ = (µ2 + ρ2 )ϕµ
and
(6.7) ϕµ (0) = 1,
where
n−1
(6.8) ρ= ,
2
see e.g., [1, Section 2] and the references therein. Here, we introduced geodesic
polar coordinates (r, ξ) in Hn and the ϕµ only depend on r. The ϕµ have the
integral representation
Z r
n−3
(6.9) ϕµ (r) = cn (sinh r)2−n (cosh r − cosh t) 2 e−iµt dt,
−r

cf. [2, equation (6), p. 4].


Since the ϕµ are distributions they are smooth in Hn , cf. [17, Theorem
3.2, p. 125]. Furthermore, for each i ∈ N we can choose µi ∈ C such that
n2
(6.10) (n − 1)(µ2i + ρ2 ) = λi − (n − 1).
2
APPLICATIONS OF CANONICAL QUANTUM GRAVITY TO COSMOLOGY 17

Obviously, there are two solutions µi and −µi , but the corresponding eigen-
functions are identical as can be easily checked.

References
[1] Jean-Philippe Anker and Vittoria Pierfelice, Wave and Klein-Gordon equations on
hyperbolic spaces, Anal. PDE 7 (2014), 953–995, arXiv:1104.0177.
[2] Jean-Philippe Anker, Vittoria Pierfelice, and Maria Vallarino, The wave equation on
hyperbolic spaces, (2010), arXiv:1010.2372.
[3] R. Arnowitt, S. Deser, and C. W. Misner, The dynamics of general relativity, Grav-
itation: an introduction to current research (Louis Witten, ed.), John Wiley, New
York, 1962, pp. 227–265.
[4] Salvatore Capozziello, Francisco S. N. Lobo, and José P. Mimoso, Generalized energy
conditions in Extended Theories of Gravity, Phys. Rev. D91 (2015), no. 12, 124019,
1407.7293, doi:10.1103/PhysRevD.91.124019.
[5] Sayantan Choudhury, Manibrata Sen, and Soumya Sadhukhan, Can dark matter be
an artifact of extended theories of gravity?, The European Physical Journal C 76
(2016), no. 9, 494, doi:10.1140/epjc/s10052-016-4323-2.
[6] Bryce S. DeWitt, Quantum Theory of Gravity. I. The Canonical Theory, Phys. Rev.
160 (1967), 1113–1148, doi:10.1103/PhysRev.160.1113.
[7] Paul A. M. Dirac, Lectures on quantum mechanics, Belfer Graduate School of Science
Monographs Series, vol. 2, Belfer Graduate School of Science, New York, 1967, Second
printing of the 1964 original.
[8] Claus Gerhardt, Quantum cosmological Friedman models with an initial sin-
gularity, Class. Quantum Grav. 26 (2009), no. 1, 015001, arXiv:0806.1769,
doi:10.1088/0264-9381/26/1/015001.
[9] , The quantization of gravity in globally hyperbolic spacetimes,
Adv. Theor. Math. Phys. 17 (2013), no. 6, 1357–1391, arXiv:1205.1427,
doi:10.4310/ATMP.2013.v17.n6.a5.
[10] , A unified quantum theory I: gravity interacting with a Yang-Mills
field, Adv. Theor. Math. Phys. 18 (2014), no. 5, 1043–1062, arXiv:1207.0491,
doi:10.4310/ATMP.2014.v18.n5.a2.
[11] , The quantization of a black hole, (2016), arXiv:1608.08209.
[12] , The quantum development of an asymptotically Euclidean Cauchy hypersur-
face, (2016), arXiv:1612.03469.
[13] , The quantization of a Kerr-AdS black hole, Advances in Mathemati-
cal Physics vol. 2018 (2018), Article ID 4328312, 10 pages, arXiv:1708.04611,
doi:10.1155/2018/4328312.
[14] , The quantization of gravity, Adv. Theor. Math. Phys. 22 (2018), no. 3, 709–
757, arXiv:1501.01205, doi:10.4310/ATMP.2018.v22.n3.a4.
[15] , The Quantization of Gravity, 1st ed., Fundamental Theories of Physics, vol.
194, Springer, Cham, 2018, doi:10.1007/978-3-319-77371-1.
[16] Claus Kiefer, Quantum Gravity, 2nd ed., International Series of Monographs on
Physics, Oxford University Press, 2007.
[17] J.-L. Lions and E. Magenes, Non-homogeneous boundary value problems and appli-
cations. Vol. I, Springer-Verlag, New York, 1972, Translated from the French by P.
Kenneth, Die Grundlehren der mathematischen Wissenschaften, Band 181.
[18] Thomas Thiemann, Modern canonical quantum general relativity, Cambridge Mono-
graphs on Mathematical Physics, Cambridge University Press, Cambridge, 2007, With
a foreword by Chris Isham.
18 CLAUS GERHARDT

Ruprecht-Karls-Universität, Institut für Angewandte Mathematik, Im Neuen-


heimer Feld 205, 69120 Heidelberg, Germany
E-mail address: [email protected]
URL: http://www.math.uni-heidelberg.de/studinfo/gerhardt/

You might also like