An emergent universe from a loop
David J. Mulryne1 , Reza Tavakol1, James E. Lidsey1 and George F. R. Ellis2
arXiv:astro-ph/0502589v1 28 Feb 2005
1
Astronomy Unit, School of Mathematical Sciences,
Queen Mary, University of London, London E1 4NS, U.K. and
2
Department of Applied Mathematics, University of Cape Town,Cape Town, South Africa
Closed, singularity-free, inflationary cosmological models have recently been studied in the context
of general relativity. Despite their appeal, these so called emergent models suffer from a number of
limitations. These include the fact that they rely on an initial Einstein static state to describe the
past eternal phase of the universe. Given the instability of such a state within the context of general
relativity, this amounts to a very severe fine tuning. Also in order to be able to study the dynamics
of the universe within the context of general relativity, they set the initial conditions for the universe
in the classical phase. Here we study the existence and stability of such models in the context of
Loop Quantum Cosmology and show that both these limitations can be partially remedied, once
semi-classical effects are taken into account. An important consequence of these effects is to give
rise to a static solution (not present in GR), which dynamically is a centre equilibrium point and
located in the more natural semi-classical regime. This allows the construction of emergent models
in which the universe oscillates indefinitely about such an initial static state. We construct an
explicit emergent model of this type, in which a non-singular past eternal oscillating universe enters
a phase where the symmetry of the oscillations is broken, leading to an emergent inflationary epoch,
while satisfying all observational and semi-classical constraints. We also discuss emergent models in
which the universe possesses both early- and late-time accelerating phases.
I.
INTRODUCTION
One of the fundamental questions of modern cosmology is whether the universe had a definite origin or
whether it is past eternal. The central paradigm for
structure formation in the universe is the inflationary
scenario. (For a review, see, e.g., Ref. [1]). Under very
general conditions, inflation is future eternal, in the sense
that once inflation has started, most of the volume of the
universe will remain in an inflating state [2]. Given certain assumptions, however, it has been argued [3] that
inflation can not be past null complete – and therefore
past eternal – if it is future eternal.
Leading alternatives to inflation that are motivated by
recent advances in string/M–theory are the pre–big bang
[4, 5] and ekpyrotic/cyclic [6] scenarios, respectively. The
fundamental postulate of the pre–big bang model is that
the set of initial data for the universe lies in the infinite
past in the perturbative regime of small string coupling
and spacetime curvature [4]. The universe then evolves
into a strongly coupled and highly curved regime before
exiting into the standard, post–big bang phase.
On the other hand, the big bang singularity is interpreted in the cyclic scenario in terms of the collisions of
two co–dimension one branes propagating in a higher–
dimensional spacetime [6]. In this picture, the branes
undergo an infinite sequence of oscillations where they
move towards and subsequently away from each other.
From the perspective of a four–dimensional observer, the
collision of the branes is interpreted as the bounce of a
collapsing universe into a decelerating, post–big bang expansion.
Despite these attractive properties, however, the process that leads to a non–singular transition between the
pre– and post–big bang phases is unclear. The search for
singularity-free inflationary models within the context of
classical General Relativity (GR) has recently led to the
development of the emergent universe scenario [7, 8]. In
this model, the universe is positively curved and initially
in a past eternal Einstein static (ES) state that eventually evolves into a subsequent inflationary phase. Such
models are appealing since they provide specific examples
of non–singular (geodesically complete) inflationary universes. Furthermore, it has been proposed that entropy
considerations favour the ES state as the initial state for
our universe [9].
Classical emergent models, however, suffer from a
number of important shortcomings. In particular, the
instability of the ES state (represented by a saddle equilibrium point in the phase space of the system) makes
it extremely difficult to maintain such a state for an infinitely long time in the presence of fluctuations, such as
quantum fluctuations, that will inevitably arise. Moreover, the initial ES state must be set in the classical domain in order to study the dynamics within the context
of GR. A more natural choice for the initial state of the
universe would, however, be in the semi-classical or quantum gravity regimes.
A leading framework for a non–perturbative theory of
quantum gravity is loop quantum gravity (LQG). (For reviews, see, e.g., Ref. [10, 11]). Recently, there has been
considerable interest in loop quantum cosmology (LQC),
which is the application of LQG to symmetric states [12].
Within the framework of LQC, there can exist a ‘semi–
classical’ regime, where spacetime is approximated by a
continuous manifold, but where non–perturbative quantum corrections modify the form of the classical Einstein
field equations [13].
Motivated by the above discussion, the present paper
studies the existence and nature of static solutions within
2
the framework of semi–classical LQC in the presence of a
self–interacting scalar field. Significant progress has recently been made in understanding the dynamics of this
system, and its relevance to inflation and non-singular
behaviour [14, 15, 16, 17, 18, 19, 20, 21]. In addition
to the ES universe of classical cosmology, LQC effects
result in a second equilibrium point in the phase space
that we refer to as the ‘loop static’ (LS) solution. Crucially, this solution corresponds to a centre equilibrium
point in the phase space and, consequently, the universe
may undergo a series of (possibly) infinite, non–singular
oscillations about this point. During these oscillations,
the scalar field can be driven up its potential [17]. This
leads us to consider a new picture for the origin of the
universe, where the universe is initially oscillating about
the LS static solution in the infinite past and eventually emerges into a classical inflationary era. The model
shares some of the attractive features of the pre–big bang
and cyclic scenarios, in the sense that it is past eternal,
although it exhibits a significant difference in that the cycles are broken by inflationary expansion. Additionally,
unlike the pre–big bang and cyclic scenarios, the emergent model considered here is genuinely non-singular.
The paper is organised as follows. The existence and
nature of static universes in semi-classical LQC is studied in Section 2. Section 3 discusses the dynamics that
leads to the emergence of an inflationary universe and
a particular inflationary model that is consistent with
present–day cosmological observations is outlined in Section 4. In Section 5 we discuss a class of models where
the field that drives the pre–inflationary oscillations may
also act as the source of dark energy in the present–day
universe. We conclude with a discussion in Section 6.
II.
A.
STATIC UNIVERSES
Einstein static universe in classical gravity
field. Equations (1) and (2) together admit the first integral (the Friedmann equation):
H2 =
i
2 h
ä
8πlPl
φ̇2 − V (φ) ,
=−
a
3
(1)
φ̈ + 3H φ̇ + V ′ (φ) = 0 ,
(2)
and local conservation of energy–momentum implies the
scalar field satisfies
where a prime denotes differentiation with respect to the
(3)
where as usual K = 0, ±1 parametrises the spatial curvature. Combining Eq. (3) with Eq. (1) yields
2 2
φ̇ +
Ḣ = −4πlPl
K
.
a2
(4)
The Einstein static universe is characterised by the
conditions K = 1 and ȧ = 0 = ä. In the presence of
a scalar field with a constant potential V , it is straightforward to verify that there exists a unique solution satisfying these conditions, where the scale factor is given
2 2 1/2
by a = a0 with a0 = (4πlPl
φ̇ ) . There are two important points to note regarding this solution. Firstly, it
represents a saddle equilibrium point in the phase space,
which is unstable to linear perturbations. Secondly, in
order to enable the analysis to be performed within the
context of GR, it is necessary to assume that a0 lies in
the classical domain.
In the emergent universe scenario, it is assumed that
the potential becomes asymptotically flat in the limit
φ → −∞ and that the initial conditions are specified
such that the ES configuration represents the past eternal state of the universe, out of which the universe slowly
evolves into an inflationary phase. However, the instability of the ES universe ensures that any perturbations
– no matter how small – rapidly force the universe away
from the ES state, thereby aborting the scenario.
In the next subsection we shall see that employing a
more general LQC setting can partially resolve both these
issues.
B.
Before considering static solutions within the context
of LQC, it is instructive to review the corresponding results for classical Einstein gravity.
Throughout this paper, we consider an isotropic and
homogeneous Friedmann–Robertson–Walker (FRW) universe sourced by a scalar field with energy density and
pressure given by ρ = 21 φ̇2 + V (φ) and p = 12 φ̇2 − V (φ),
respectively, where V (φ) represents the self–interaction
potential of the field. The Raychaudhuri equation for
such a universe is given by
i K
2 h
8πlPl
φ̇2 + V (φ) − 2 ,
3
a
Static solutions in semi-classical LQC
It has recently been shown that in the semi-classical
regime the cosmological evolution equations become
modified [13]. When restricted to FRW backgrounds, this
regime is characterised by the scale factor of the universe
√
lying in the range ai < a < a∗ , where ai ≡ γlPl determines the scale for the discrete quantum nature of spacetime to become important and γ ≈ p
0.274 is the Barbero–
Immirzi parameter [22], and a∗ ≡ γj/3lPl , where j is
a parameter that arises due to ambiguities in the quantization procedure [23]. It must take positive, half–integer
values but is otherwise arbitrary [38]. The standard classical cosmology is recovered above the scale a∗ and the
parameter j therefore sets the effective quantum gravity scale. The dynamics in this regime is determined
by replacing the inverse volume term a−3 that arises in
the classical matter Hamiltonian Hφ = 21 a−3 p2φ + a3 V (φ)
with a continuous function dj (a) that approximates the
eigenvalues of the geometrical density operator in LQC.
This ‘quantum correction’ function is given by dj (a) ≡
3
D(q)a−3 , where
6 n h
i
8
3
11
11
q 2 7 (q + 1) 4 − |q − 1| 4
D(q) =
77
h
io6
7
7
− 11q (q + 1) 4 − sgn(q − 1)|q − 1| 4
, (5)
and q ≡ (a/a∗ )2 . As the universe evolves through the
semi–classical phase, this function varies as D ∝ a15 for
a ≪ a∗ , has a global maximum at a ≈ a∗ , and falls
monotonically to D = 1 for a > a∗ .
The effective field equations then follow from the
Hamiltonian [13]:
3
1
Ĥ = −
ȧ2 + K a + dj p2φ + a3 V = 0 ,
8πℓ2Pl
2
The first integral of Eqs. (7) and (8) is given by the
modified Friedmann equation
"
#
2
φ̇2
K
8πlPl
2
(9)
+ V (φ) − 2 ,
H =
3
2D
a
where in the LQC context K can only take values 0, +1.
Combining Eq. (7) with Eq. (9) then implies that
1−
1 d ln D
6 d ln a
+
K
.
a2
(10)
Equation (7)–(10) can be rewritten in the form of a
three–dimensional dynamical system in terms of variables
{a, H, V } [20]. (The present study extends the results of
[20] and, more importantly, provides an analytical account of the dynamics). Employing the Friedmann equation (9) to eliminate the scalar field’s kinetic term, and
assuming that dV /dφ can be expressed as a function of
the potential, allows the complete dynamical system to
be described by equations:
A
2
V − 3H 2 1 −
Ḣ = 8πlPl
6
K A
−2 ,
(11)
+ 2
a
2
ȧ = Ha ,
(12)
1
2
6D
6DH 2
− 2DV +
, (13)
V̇ = V ′ (V )
2
2
2
8πlPl
8πlPl a
a = aeq ,
(14)
where aeq is given by solutions to the constraint equation
(6)
and the modified scalar field equation takes the form
1 d ln D
φ̈ = −3H 1 −
φ̇ − DV ′ .
(8)
3 d ln a
2 2
4πlPl
φ̇
D
Heq = 0,
A(aeq ) = B(aeq ),
where pφ = d−1
j φ̇ is the momentum canonically conjugate
to the scalar field. The Raycharduri equation becomes
2
2
ä
8πlPl
1 d ln D
8πlPl
+
φ̇2 1 −
=−
V (φ) ,
(7)
a
3D
4 d ln a
3
Ḣ = −
where the function A(a) ≡ d ln D/d ln a. In this Section,
with the aim of considering emergent universes we shall
take K = +1 and initially consider constant potentials.
In this case, Eq. (13) is trivially satisfied and the system reduces to the two–dimensional autonomous system
(11)–(12).
The static solutions (ȧ = 0 = ä) then correspond to
the equilibrium points of this system, which are given by
B(a) ≡
2
6(8πlPl
V a2 − 2)
.
2
(8πlPl V a2 − 3)
(15)
It follows from Eq. (10) that the field’s kinetic energy at
the equilibrium points is given by
φ̇2eq =
Deq
2
2
4πaeq lPl (1 −
Aeq /6)
.
(16)
Equation (15) implies that a necessary and sufficient
condition for the existence of the equilibrium points is
that the functions A(a) and B(a) should intersect.
Here we are interested in how the properties of the
equilibrium points alter as the potential is varied. This
is best seen by considering how the functions A(a) and
B(a) (and hence their points of intersection) change. The
function A has the form shown in the top panel of Fig. 1.
It asymptotes to the constant value A = 15 at a = 0,
decreases to a minimum value of Amin = −5/2 at a =
a∗ , and then asymptotes to zero from below as a → ∞.
Increasing the parameter j increases the value of a∗ and
this results in moving the turning point of the function
A to larger values of the scale factor. The important
point to note, however, is that the qualitative form of
the function A remains unaltered for all values of j, as
can be seen from Fig. 1.
The qualitative nature of the function B, on the other
hand, is sensitive to the sign of the potential, as can be
seen from the bottom panel of Fig. 1. We briefly discuss
the behaviour of B for each case in turn. For V = 0,
it is given by the (solid) horizontal line B = 4. For
V > 0 it represents a hyperbola (dot-dashed curve) with
a single (since the scale factor a is non-negative) vertical
asymptote given by
s
3
a=
(17)
2 V .
8πlPl
The region to the left of this asymptote defines the region in the {A/B, a} plane where the reality condition,
φ̇2 > 0, is satisfied. Thus, the upper–right branch of the
hyperbola plays no role in determining the existence of
the equilibrium points. In the limit of a → 0, the function B → 4 which coincides with the value of the function
in the case V = 0. As a is increased, the function takes
progressively smaller values as the vertical asymptote is
4
approached. As V tends to zero from above, it causes
the asymptote to move to progressively larger values of
a, and we may therefore formally view the V = 0 case as
the limit where the asymptote moves to infinity. Finally,
for negative potentials the qualitative behaviour of the
function B is changed to the dashed line in the bottom
panel of Fig. 1 and there is no vertical asymptote. As in
the case of positive potentials, B → 4 as a → 0, but the
function B now increases monotonically as a increases,
ultimately tending to the value 6 as a → ∞. As the
potential tends to zero from below, the function B still
tends to the asymptotic value B = 6, but at larger a.
An important point to note is that for all values of V ,
the function B satisfies the condition B < 6 (for physically relevant regions of the {A/B, a} plane), and hence
equilibrium points can only occur when A = B < 6.
Once the positions of the equilibrium points have been
determined, their nature can be found by linearising
about these points. The eigenvalues are given by
dA
1
4−A
+
λ =
2
a
6a(1 − A/6) da
2
.
(18)
15
10
5
0
−5
0
5
10
a
15
20
25
15
10
aeq
When λ2 < 0, this leads to imaginary eigenvalues and
implies that the equilibrium point is a centre, whereas
the point is a saddle when λ2 > 0. Although this expression is rather complicated, it can be verified that for
A < 6 (which is a necessary condition for the existence of
equilibrium points), λ2 is negative for a < a∗ and positive for a > a∗ . Hence we have the important result that,
equilibrium points occurring in the semi-classical regime
are centres and those occurring in the classical regime are
saddles.
To determine the existence of equilibrium points, we
again consider the different signs of the potential separately.
V = 0: In the case of a massless scalar field with V = 0,
the condition (15) has a single solution given by A = 4,
implying a single equilibrium point. The position of this
equilibrium point can be seen from the intersection point
in the top panel of Fig. 2. Since this point occurs at
a < a∗ , it is a centre, with its phase portrait consisting
of a continuum of concentric orbits (see the bottom panel
of Fig. 2).
V > 0: As described above, for small positive values
of the potential, the vertical asymptote of the function
B is far to the right of the origin, which results in two
points of intersections between the functions A and B
(see the top panel of Fig. 3). There are therefore two
equilibrium points in this case: the first occurs at a < a∗
and hence is a centre, while the second occurs at a > a∗
and is therefore a saddle. As the potential is increased,
the asymptote moves towards the origin causing the equilibrium points to eventually coalesce when V = V∗ . This
occurs at the point of tangency of the functions A and
B, i.e. at the minimum (which is a cusp) of the function
A located at a = a∗ . Thus, using B and the fact that
5
0
−5
0
5
a
10
15
FIG. 1: The top panel shows the plots for the function A
against a for j = 100 (solid line), j = 1000 (dashed line)
and j = 3000 (dot-dashed line). The bottom panel shows
plots of the function B against a for V = 0 (solid line), V > 0
(dot-dashed line) and V < 0 (dashed line). Note the dramatic
change in the shape of the function as the sign of V is altered.
A|a∗ = −5/2, we obtain
V∗ =
39
117
4
2 a2 = 136πγj mPl .
136πlPl
∗
(19)
For values of V > V∗ , the asymptote moves further
towards the origin. Consequently, the functions A and
B will no longer intersect in the part of the plane which
satisfies the reality condition, and therefore there will be
no equilibrium points in this case. It is also easy to understand the effect of changing the parameter j on V∗ .
Increasing j increases a∗ , which means that the point of
tangency occurs for smaller values of the potential. This
implies that larger values of the quantization parameter,
and hence a∗ , lead to smaller values of the potential beyond which no equilibrium points can exist.
5
15
15
10
10
5
5
0
0
−5
0
a*
5
a
10
15
+∞
−5
0
a*
5
10
a
15
+∞
a
a
a
a
i
i
0
−∞
1
H
1
+∞
0
−∞
1
H
1
+∞
FIG. 2: The top panel shows the plots of functions A (solid
line) and B (dashed line) against a for V = 0. The vertical
dotted line denotes the position of a = a∗ . The bottom panel
is the corresponding phase portrait demonstrating the centre
equilibrium point that occurs in this case. The plot is compactified using x = arctan(H) and y = arctan(ln a), in order
to present the entire phase space. The vertical dotted lines
in the bottom panel demarcate the region in which the Hubble parameter is less than the Planck scale. The horizontal
dotted line marks ai , where the quantum regime begins. All
axes are in Planck units, and j = 100.
FIG. 3: The top panel shows the plots of functions A (solid
line) and B (dashed line) against a for V = 0.005. For this
value of V , the equilibrium points are close to coalescing. The
vertical dotted line denotes the position of a = a∗ . The bottom panel is the corresponding phase portrait demonstrating
the centre equilibrium point and saddle point which occurs in
this case with the dashed line indicating the separatrix. The
shaded area is the unphysical region where φ̇2 would be negative. The plot is compactified as in Fig. 2. The dotted lines
in the bottom panel are as described in Fig. 2. All axes are
in Planck units, and j = 100.
V < 0: Negative potentials are known to arise in
string/M-theory compactifications (e.g., [24]) and are of
interest in connection with the ekpyriotic/cyclic models considered recently [6]. In such cases the qualitative
change in the form of B results in only one equilibrium
point, which occurs inside the semi-classical region and
hence is a centre (see the bottom panel of Fig. 4). Furthermore, the centre equilibrium point will persist for any
negative value of V .
To summarise, the important consequence of considering LQC effects in this context is that they admit two
possible static solutions, rather than the single solution
in the case of GR. The first static solution corresponds
to a saddle point (as in GR) and is referred to as the Einstein Static (ES) solution. The second static solution is a
direct consequence of LQC effects. It is a centre and we
shall refer to it as the loop static (LS) solution. We have
shown analytically that the LS solution arises for a wide
range of values for the potential given by V ∈ (−Vlb , V ∗ ),
where the lower bound −Vlb is determined by the need
to satisfy the Planck bounds. The importance of the LS
solution is that, in contrast to the ES solution present in
GR, slight perturbations do not result in an exponential
divergence from the static universe, but lead instead to
6
15
III.
EMERGENT INFLATIONARY UNIVERSE
IN LQC
A.
The Dynamics of Emergence
10
5
0
−5
0
a*
5
a
10
15
+∞
a
a
i
0
−∞
1
H
1
+∞
FIG. 4: The top panel shows the plots of functions A (solid
line) and B (dashed line) against a for V = −0.013. The vertical dotted line denotes the position of a = a∗ . The bottom
panel is the corresponding phase portrait demonstrating the
centre equilibrium point that occurs in this case. The plot
is compactified as in Fig. 2. The dotted lines in the bottom
panel are as described in Fig. 2. All axes are in Planck units
and j = 100.
oscillations about it. Moreover, since the LS solution always lies within the semi-classical region a < a∗ , it is ideally suited to act as the past asymptotic initial state for
an emerging universe. Indeed, it is the only equilibrium
point in models with a vanishing or negative potential.
In the following Section, we employ these novel features
of LQC dynamics within the context of the emergent inflationary universe.
The phase plane analysis of the previous Section determined the qualitative LQC dynamics for the case of
a constant potential. However, any realistic inflationary
model clearly requires the potential to vary as the scalar
field evolves. Nevertheless, the constant potential dynamics (see Figs. 2–4) provides a good approximation to
more general dynamics if variations in the potential are
negligible over a few oscillations. This implies that cosmic dynamics with a variable potential may be studied by
treating the potential as a sequence of separate, constant
potentials. Here, with the emergent inflationary models
in mind (see below), we shall consider a general class of
potentials that asymptote to a constant V−∞ < V∗ as
φ → −∞ and rise monotonically once the value of the
scalar field exceeds a certain value.
As we saw in the previous Section, for any constant potential V < V∗ , there exists a region of parameter space
where the universe undergoes non–singular oscillations
about the point LS. For V > V∗ , the equilibrium points
LS and ES merge and the phase plane then resembles that
of classical GR: a collapsing universe bounces and asymptotically evolves into a de Sitter (exponential) expansion
in the infinite future. This leads us to propose the following picture for the origin of the universe. The universe is
initially at, or in the neighbourhood of, the static point
LS, with the field located on the plateau region of the
potential with a positive kinetic energy, φ̇init > 0. The
universe undergoes a series of non–singular oscillations
in a (possibly) past–eternal phase with the field evolving monotonically along the potential. As the magnitude
of the potential increases, the cycles are eventually broken by the emergence of the universe into an inflationary
epoch.
We proceed to discus the dynamics of this scenario in
more detail. The universe will exhibit cyclic behaviour
around the LS point for a very wide range of initial values φ̇init when V−∞ < 0. If 0 < V−∞ < V∗ , the range
becomes more limited as V−∞ is increased. An important feature to note is that in all cases the field’s kinetic
energy never vanishes during a given cycle. In the case
of a positive potential, the unphysical region of phase
space where the field violates the null energy condition
lies outside the cyclic region that is bounded by the separatrix. Indeed, for a constant potential, the evolution
of the scalar field is determined by integrating Eq. (8):
a 3 D
init
φ̇ = φ̇init
.
(20)
a
Dinit
Thus, the field will vary monotonically along the potential and eventually reach the region where the potential
begins to rise.
Increasing the magnitude of the potential over a series
of cycles has the effect of moving the location of the LS
7
point to progressively higher values of the scale factor,
although the shift is moderate. In the case of a positive potential, for example, a necessary condition for the
existence of LS is that 4 < A(aeq ) < −2.5, where the
function A is defined in Eq. (15), and this corresponds
to the range 0.94a∗ < a < a∗ . Eq. (16) then implies
that the field’s kinetic energy does not alter significantly,
since the universe remains in the vicinity of LS. On the
other hand, the saddle point ES occurs at progressively
smaller values of the scale factor as the magnitude of
the potential increases. As discussed in Section 2.2, this
equilibrium point occurs in the range a∗ ≤ a ≤ ∞, where
the limits are approached as V → V∗ and V → 0, respectively.
The overall effect of increasing the potential, therefore,
is to distort the separatrix in the phase plane, making it
narrower in the vertical direction but introducing little
change to the position of the LS point. This implies that
the dynamics varies only slightly from cycle to cycle for
orbits that are close to the LS point. If the magnitude
of the potential continues to increase as the field evolves,
however, a cycle is eventually reached where the trajectory that represents the universe’s evolution now lies outside the finite region bounded by the separatrix and this
effectively breaks the oscillatory cycles. From a physical point of view, the magnitude of the field’s potential
energy, relative to its kinetic energy, is now sufficiently
large that a recollapse of the universe is prevented, i.e.,
the strong energy condition of GR is violated, thereby
leading to accelerated expansion. The field decelerates as
it moves further up the potential, subsequently reaching
a point of maximal displacement and then rolling back
down. If the potential has a suitable form in this region,
slow–roll inflation will occur.
B.
The Energy Scale of Inflation
An important question to address is the energy scale at
the onset of inflation, Vinf . In general, inflation begins (in
the classical regime) when the strong energy condition is
violated:
V (φinf ) ≈ φ̇2inf ,
(21)
and, moreover, the structure of the phase space indicates
that the potential energy of the field remains dynamically negligible until the onset of inflation [39]. In this
case, the evolution of the scalar field prior to inflation is
determined by Eq. (20).
The inflationary energy scale may then be estimated
by considering the penultimate cycle before the onset of
inflation. The turnaround in the expansion occurs when
2 2
the field’s energy density satisfies ρ = 3/(8πlPl
a ). Since
the energy density of the field will not have changed significantly at the equivalent stage of the following cycle,
the scale factor at the onset of inflation is determined
1
2
approximately by ainf ≈ (4πlPl
Vinf )− 2 . Substituting this
condition into Eq. (20) then yields an estimate for the
magnitude of the potential in terms of initial conditions:
Vinf ≈
1
2 )
(4πlPl
3
2
Dinit
3
ainit φ̇init
.
(22)
For a universe located near to the equilibrium point
LS, the scale factor is given by ainit = f a∗ , where 0.94 ≤
f ≤ 1 and it may be further verified that Dinit = O(1)
in this range. The field’s initial kinetic energy is then
2 2
determined by Eq. (16): φ̇2init ≈ 3/(4πlPl
aeq ). It follows
from Eq. (22), therefore, that a universe ‘emerging’ from
the semi–classical LQC phase near to the LS point will
begin to inflate when
Vinf ≈
1
m4 .
2jf 2 Pl
(23)
As expected, this is in good agreement with the necessary condition (19) for the coalescence of the equilibrium
points LS and ES, since the scale factor can not evolve
until V > V∗ if the universe is located on or near the LS
point.
A precise measure on the set of initial data is presently
unknown, and we should therefore consider other possible
initial conditions. Having investigated the regime ainit ≈
a∗ , a second possibility is to consider initial conditions
where ai ≈ ainit ≪ a∗ . In this regime, the quantum
correction factor (5) asymptotes to a power–law, D ≈
(12/7)6 (a/a∗ )15 , and Eq. (22) is then equivalent to the
condition
−1
β 12 Vinf
φ̇init ≈ 20 3
m2Pl ,
(24)
j 2 m4Pl
where we have defined the ratio β ≡ ainit /a∗ . This ratio
may be constrained by imposing two necessary conditions
for the semi–classical framework to be valid. Firstly, the
initial conditions should be set in the regime where spacetime is approximated by a smooth manifold, ai /ainit < 1,
and this leads to the lower limit:
r
3
β>
.
(25)
j
Secondly, the scalar field’s kinetic energy must not exceed
the Planck scale at the onset of the classical regime. Since
the anti–frictional effects in the modified field equation
(8) accelerate the field when a < a∗ , the field’s kinetic
energy must initially be sub-Planckian, |φ̇init |/m2Pl ≤ 1.
Eq. (24) then leads to an upper limit on β:
β<
Vinf
m4Pl
1
12
1
j8 ,
(26)
and combining the constraints (25) and (26) results in a
necessary condition for the onset of inflation when ainit ≪
a∗ :
15
2 2
Vinf
>
.
(27)
4
mPl
j
8
oscillations of the inflaton. Since inflation will end after
the field has rolled back down the potential, this should
occur for φ ≤ φgrow . Finally, the region of the potential
that drove the last 60 e–foldings of inflationary expansion is then constrained by cosmological observations, as
in the standard scenario.
Examples of potentials that exhibit these generic properties are illustrated in Fig. 5. It is interesting that
potentials of this form have been considered previously
in a number of different settings, including cases where
higher–order curvature invariants of the form
V
LN
φ
FIG. 5: The figure depicts two possible forms of emergent
potentials that allow for conventional re-heating. The solid
line is the form of the potential motivated by the inclusion of
a R2 term in the Einstein–Hilbert action, and the dotted line
the form motivated by the inclusion of higher–order terms.
In other words, for a given value of the parameter j,
15
inflation must begin above the scale (2/j) 2 m4Pl .
Conditions (23) and (27) both imply that the inflationary energy scale is higher for lower values of the parameter j. Indeed, it is comparable to the Planck scale
for j ≤ O(10) and this generic behaviour is only weakly
dependent on the initial conditions. This has implications for the asymptotic form of the potential as the field
reaches progressively higher values. Unless the parameter j is sufficiently large, it is unlikely that the oscillatory dynamics will end if the potential asymptotes to
a constant value, or reaches a local maximum, that is
significantly below the Planck scale. In this sense, therefore, the scenario outlined above favours potentials that
increase monotonically once the value of the scalar field
has exceeded some critical value.
In the following Section, we consider a concrete example of a potential that exhibits the appropriate asymptotic behaviour.
IV.
A SPECIFIC MODEL OF AN EMERGING
UNIVERSE
From a dynamical point of view, the emergent universe scenario can be realised within the context of semi–
classical LQC if the potential satisfies a number of rather
weak constraints. Asymptotically, it should have a horizontal branch as φ → −∞ such that dV /dφ → 0 and
increase monotonically in the region φ > φgrow , where
without loss of generality we may choose φgrow = 0. The
reheating of the universe imposes a further constraint
that there should be a global minimum in the potential
at Vmin = 0 if reheating is to proceed through coherent
N
1 X
ǫi R i
=
2
16πlPl
i=1
(28)
are introduced into the Einstein–Hilbert action, where R
is the Ricci scalar, ǫi are coupling constants and ǫ1 = 1.
Such corrections are required when attempting to renormalize theories of quantum gravity [25] and also arise
in low-energy limits of superstring theories [26]. In general, such theories are conformally equivalent to Einstein
gravity sourced by a minimally coupled, self–interacting
scalar field. For example, potentials with a nonzero
asymptote at φ → −∞ (as shown by the solid line in
Fig. 5) can be obtained from theories that include a
quadratic term in the action, whereas those with a zero
asymptotic value and a local maximum (shown by a dotted line) arise when cubic and higher–order terms in the
Ricci scalar are introduced [27, 28]. In general, all these
potentials possess a global minimum at V = 0. Finally,
potentials of this form have also been considered within
the context of the classical emergent universe [8].
Motivated by the above discussion, we consider, as an
example, the potential
h
i2
√
V = α (exp(βφ/ 3) − 1 ,
(29)
where α and β are constants. This potential is qualitatively similar to that illustrated by the solid line in
Fig. 5. The parameters in the potential are constrained
primarily by the Cosmic Microwave Background (CMB)
anisotropy power spectrum. Assuming inflation proceeds
in the region φ ≫ 0, the parameter β determines the
spectral index of the scalar perturbation spectrum together with the relative amplitude of the gravitational
wave perturbations. The parameter α is then constrained
by the COBE normalization of the power spectrum on
large–scales [29].
We chose α = 10−12 and β = 0.1 as representing typical values satisfying the constraints imposed recently by
the WMAP satellite [30] and numerically integrated the
field equations (7)–(9) for a universe starting from an initial state close to the LS static state. The results of the
integration are illustrated in Figs. 6 and 7. The field
starts in the asymptotically flat region of the potential
and gradually increases in value, as the scale factor oscillates about the LS point. The field moves past the
minimum and climbs up the potential. The scale factor continues to oscillate until the field reaches the point
9
29
20
a
x 10
10
0
0
200
400
200
400
600
800
1000
600
800
1000
100
φ
50
0
0
time
FIG. 6: Time evolution of the scale factor (top panel) and
scalar field (bottom panel) with the field initially on the
asymptotically flat region of the potential (29) with α = 10−12
and β = 0.1. The field increases in value from initial conditions close to the LS (centre) solution defined by Eq. (14).
5
4
a
3
2
400
450
500
550
600
650
700
450
500
550
time
600
650
700
120
100
φ
80
60
400
FIG. 7: Plots illustrating the magnification of Fig. 6 around
the region where emergence commences, the oscillations cease
and inflation begins.
where it slows down significantly, thereby bringing the
oscillations to an end and initiating the inflationary expansion. The behaviour at this stage is qualitatively similar to that discussed previously for a quadratic potential
[17].
V.
EMERGING QUINTESSENTIAL INFLATION
One drawback of reheating through inflaton decay in
the emergent scenario is that the coupling of the scalar
field to the standard model degrees of freedom must be
strongly suppressed if the field is to survive for a (possibly) infinite time as it emerges from the oscillatory semi–
classical phase. It is more natural, therefore, to invoke a
‘sterile’ inflaton that is not coupled directly to standard
model fields, and where reheating proceeds through an
alternative mechanism such as gravitational particle production [31]. In this case, the potential need not exhibit
a minimum and the field could continue to roll back down
the potential towards φ → −∞ at late times.
It is notable that the general requirements for a sterile inflaton with a potential exhibiting a decaying tail as
φ → −∞ are precisely those features that are characteristic of the quintessential inflationary scenario, where the
field that drove early universe inflation is also identified
as the source of dark energy today [32]. In the present
context, this suggests that the scalar field could play a
three–fold role in the history of the universe, acting as
the mechanism that enables the universe to emerge into
the classical domain, and then subsequently driving both
the early– and late–time accelerated expansion.
The specific constraints that must be satisfied by the
potential in standard quintessential inflation have been
considered in detail in Ref. [33]. In particular, one
of the simplest asymptotic forms for the low–energy
tail that simultaneously leads to tracking behaviour and
satisfies primordial nucleosynthesis bounds is given by
V ∝ (m/φ)k exp(λφ/mPl ), where m, k and λ are constants. Cyclic behaviour in LQC will arise for such a
potential.
A further requirement is that the potential must be
sufficiently steep immediately after the end of inflation if
the field’s energy density is to redshift more rapidly than
the sub–dominant radiation component. This requires a
second point of inflexion in the potential, as illustrated
qualitatively in Fig. 8. Beyond this region, the potential must continue to rise in order for the oscillatory dynamics to come to end and, as discussed above, this is
expected to occur near to the Planck scale. From a dynamical point of view, there are no further constraints
on how rapidly the potential energy need increase in this
region. The only remaining consideration is that a phase
of successful slow–roll inflation should arise as the field
rolls back down the potential. Given the ease with which
the inflaton is able to move up the potential due to LQC
effects, we anticipate that any potential satisfying the
existing constraints for successful quintessential inflation
will also lead to a successful emergence of the classical
universe.
VI.
DISCUSSION
In this paper, we have investigated the occurrence
of static solutions in loop quantum cosmology settings
sourced by a scalar field with a constant potential. We
have shown that there are in principle two such solutions,
depending upon the sign and magnitude of the potential.
The point ES is always in the classical domain and cor-
10
V
φ
FIG. 8: Plots illustrating the generic form of the potential
that leads to early– and late–time accelerating phases. The
potential exhibits a decaying tail as φ → −∞. As the field
moves up this tail and increases in value, the universe can
oscillate about the LS point. The inflationary regime rises
(possibly) towards the Planck scale for large values of φ. As
the field turns round, it can drive a phase of inflation and,
if the potential exhibits a sufficiently steep middle section
around φ ≈ 0, reheating may proceed through gravitational
particle production. Consequently, the field may survive until
the present epoch, where it can act as the source of dark
energy by slowly rolling along the tail towards φ → −∞.
responds to the unstable saddle point that is also present
in GR. The important characteristic of this solution is
that any perturbations, no matter how small, necessarily force the universe to deviate exponentially from the
static state. The second solution LS is always in the
semi-classical domain and corresponds to a centre. This
is a solution made possible by LQC effects. We have
shown that it exists for a wide range of values of the potential, including positive, zero and negative values. Its
importance lies in the fact that it allows a universe that
is slightly displaced from the static state to oscillate in
the neighbourhood of the static solution for an arbitrarily
long time.
We have exploited these characteristics to develop a
working scenario of the emerging inflationary universe,
in which a past–eternal, cyclic cosmology eventually enters a phase where the symmetry of the oscillations is
broken by the scalar field potential, thereby leading in
principle to a phase of successful slow–roll inflation. The
mechanism that enables the universe to emerge depends
very weakly on the form of the potential, and requires
only that it asymptotes to a constant value at φ → −∞
and grows in magnitude at larger φ in order to break
the cycles. The asymptotic value of the potential can be
either positive, zero or negative.
The above emergent scenario has a further advantage
that the initial state of the universe is set in the more
natural semi-classical regime, rather than the classical
arena of GR. Nevertheless, an important question that
arises is the likelihood of these initial conditions within a
more general framework. In particular, there is the issue
of why the scalar field should initially be located in the
asymptotic low-energy region of the potential. Although
a detailed study of such questions is beyond the scope of
the present paper, it has been argued [34] that for the
case of a constant potential, the wavefunction in LQC
most closely resembles the Hartle–Hawking no–boundary
wavefunction [35]. More specifically, the difference equation in LQC requires the wavefunction to tend to zero
near to the classical singularity [36], and in this sense resembles DeWitt’s initial condition [37]. However, within
the context of solutions to the Wheeler–DeWitt equation, requiring the wavefunction to be bounded as a → 0
selects the exponentially increasing WKB mode [34] and
this corresponds to the no–boundary wavefunction.
This is of interest since the square of the wavefunction in quantum cosmology is interpreted as the probability distribution for initial values of the scalar field in
an ensemble of universes. For the no–boundary boundary condition, the probability, P ∝ exp(3/[8πlP2 V (φ)]),
is peaked at V (φ) = 0, thereby suggesting that the no–
boundary proposal disfavours inflation. In the present
context, however, this implies that the most natural initial condition for the field is to be located either at the
minimum of the potential, or in the case where the potential has no minimum, at φ = −∞. It would be interesting
to explore this possibility further. In particular, previous
analyses have so far neglected kinetic terms in the matter
Hamiltonian and these may be important in the emerging universe scenario. There is also the related question
of whether the field moves from left to right initially.
In general, we have found that the onset of slow–roll inflation will occur at a relatively high energy scale, unless
the quantization parameter j is extremely large. This indicates that a large amount of slow–roll inflation should
arise, at least for a wide class of smoothly varying potentials, and it is expected, therefore, that the density of
the present–day universe should be exponentially close
to the critical density, Ω0 = 1 + ǫ, where ǫ ≪ 1. In principle, therefore, the emergent scenario we have proposed
could be ruled out if a significant detection of spatial
curvature is ultimately reported by future cosmological
observations.
It is also worth remarking that if the field is pushed
sufficiently far up the potential, perhaps as high as the
Planck scale, the inflating universe may enter a phase
of eternal self–reproduction, where quantum fluctuations
in the inflaton become more important than its classical
dynamics [2].
Finally, there is an interesting symmetry in the emerging quintessential scenario between the initial and final
states of the universe. Although the size of the universe
differs by many orders of magnitude, the field evolves
along the tail of the potential at both early and late times,
φ(t → −∞) = φ(t → +∞), but with its kinetic energy
11
having changed sign. This implies that a reconstruction
of the dark energy equation of state at the present epoch
could yield direct observational insight into the nature of
the pre–inflationary potential in this scenario.
[1] A. R. Liddle and D. H. Lyth, Cosmological Inflation
and Large-Scale Structure (Cambridge University Press,
Cambridge, 2000).
[2] A. D. Linde, Phys. Lett. 175B, 395 (1986).
[3] A. Borde and A Vilenkin, Phys. Rev. Lett. 72,
3305 (1994); Phys. Rev. D 56, 717 (1997); A.
Borde, A.H. Guth and A Vilenkin, Phys. Rev.
Lett. 90,
151301 (2003) [arXiv:gr-qc/0110012];
A.H. Guth, arXiv:astro-ph/0101507; A Vilenkin,
arXiv:gr-qc/0204061.
[4] M. Gasperini and G. Veneziano, Phys. Rep. 373, 1 (2003)
[arXiv:hep-th/0207130].
[5] J. E. Lidsey, D. Wands and E. J. Copeland, Phys. Rep.
337, 343 (2000) [arXiv:hep-th/9909061].
[6] J. Khoury, B. A. Ovrut, P. J. Steinhardt and N. Turok,
Phys. Rev. D 64, 123522 (2001) [arXiv:hep-th/0103239];
P. J. Steinhardt and N. Turok, Science 296, 1436 (2002);
P. J. Steinhardt and N. Turok, Phys. Rev. D 65, 126003
(2002) [arXiv:hep-th/0111098]; J. Khoury, P. J. Steinhardt and N. Turok, Phys. Rev. Lett. 92, 031302 (2004)
[arXiv:hep-th/0307132].
[7] G.F.R. Ellis, R. Maartens, Class. Quant. Grav. 21, 223
(2004).
[8] G.F.R. Ellis, J. Murugun and C.G. Tsagas, Class. Quant.
Grav. 27, 233 (2004).
[9] G. W. Gibbons, Nucl. Phys. B 292, 784 (1987); B 310,
636 (1980).
[10] C.
Rovelli,
Liv. Rev. Rel.
1,
1 (1998)
[arXiv:gr-qc/9710008].
[11] T. Thiemann, Lect. Notes Phys. 631, 41 (2003)
[arXiv:gr-qc/0210094].
[12] M. Bojowald, Class. Quantum Grav. 19, 2717 (2002)
[arXiv:gr-qc/0202077].
[13] M. Bojowald, Phys. Rev. Lett. 89, 261301 (2002)
[arXiv:gr-qc/0206054].
[14] P. Singh and A. Toporensky, Phys. Rev. D 69, 104008
(2004) [arXiv:gr-qc/0312110].
[15] S. Tsujikawa, P. Singh and R. Maartens, Class. Quant.
Grav. 21, 5767 (2004) [arXiv:astro-ph/0311015].
[16] M. Bojowald, J.E. Lidsey, D.J. Mulryne, P. Singh
& R. Tavakol, Phys. Rev. D 70, 043530 (2004)
[arXiv:gr-qc/0403106].
[17] J.E. Lidsey, D.J. Mulryne, N.J. Nunes and R. Tavakol,
Phys. Rev. D70, 063521 (2004) [arXiv:gr-qc/0406042].
[18] D.J. Mulryne, N.J. Nunes R. Tavakol and J.E. Lidsey,
arXiv:gr-qc/0411125.
[19] J.E.
Lidsey,
JCAP,
0412,
007
(2004)
[arXiv:gr-qc/0411124].
[20] G.V. Vereshchaign, JCAP, 0407, 013 (2004).
[21] M. Bojowald, R. Maartens & P. Singh, Phys. Rev. D 70,
083517 (2004) [arXiv:hep-th/0407115].
[22] M. Domagala and J. Lewandowski, Class. Quant.
Acknowledgments
DJM is supported by PPARC.
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
Grav. 21, 5233 (2004) [arXiv:gr-qc/0407051]; K.
A. Meissner, Class. Quant. Grav. 21, 5245 (2004)
[arXiv:gr-qc/0407052].
M. Gaul and C. Rovelli, Class. Quant. Grav. 18,
1593 (2001) [arXiv:gr-qc/0011106]; M. Bojowald, Class.
Quant. Grav. 19, 5113 (2002) [arXiv:gr-qc/0206053].
P. K. Townsend and M. N. R. Wohlfarth, Phys.
Rev. Lett. 91, 061302 (2003) [arXiv:hep-th/0303097];
N. Ohta, Phys. Rev. Lett. 91, 061303 (2003)
[arXiv:hep-th/0303238]; R. Emparan and J. Garriga, JHEP 0305, 028 (2003) [arXiv:hep-th/0304124];
N. Ohta, Prog. Theor. Phys. 110, 269 (2003)
[arXiv:hep-th/0304172]; C.-M. Chen, P.-M. Ho, I. P. Neupane, N.Ohta, and J. E. Wang, JHEP 0310, 058 (2003)
[arXiv:hep-th/0306291].
I. Antoniadis, E. T. Tomboulis, Phys. Rev. D 33, 2756
(1986).
Candelas P, G T Horowitz, A Strominger, E Witten,
Nucl. Phys. B 258, 46 (1985).
K. Maeda, Phys. Rev. D 39, 3159 (1989).
J.D. Barrow and S. Cotsakis, Phys. Lett. 258B, 305
(1991).
E. F. Bunn, A. R. Liddle and M. White, Phys. Rev. D
54, R5917 (1996); E. F. Bunn and M. White, Astrophys.
J. 480, 6 (1997).
D. N. Spergel et al., Astrophys. J. Suppl. 148, 175 (2003)
[arXiv:astro-ph/0302209]; H. V. Peiris et al., Astrophys.
J. Suppl. 148, 213 (2003) [arXiv:astro-ph/0302225].
L. H. Ford, Phys. Rev. D 35, 2955 (1987); L. P. Grishchuk
and Y. V. Sidorov, Phys. Rev. D 42, 3413 (1990); B.
Spokoiny, Phys. Lett. B 315, 40 (1993).
P. J. E. Peebles and A. Vilenkin, Phys. Rev. D 59, 063505
(1999).
K. Dimopoulos, Astropart. Phys. 18, 287 (2002)
[arXiv:astro-ph/0111417].
M. Bojowald & K. Vandersloot, Phys. Rev. D 67, 124023
(2003) [arXiv:gr-qc/0303072].
J. B. Hartle and S. W. Hawking, Phys. Rev. D 28, 2960
(1983).
M. Bojowald, Phys. Rev. Lett. 87, 121301 (2001).
B. S. DeWitt, Phys. Rev. 160, 1113 (1967).
More specifically, any irreducible SU(2) representation
with spin j may be chosen in the quantization scheme
when rewriting the classical scale factor in terms of
holonomies. The fundamental representation corresponds
to j = 1/2.
For a quadratic potential, numerical integration of the
field equations indicates that this estimate yields a very
good measure of the energy scale at the onset of inflation when the universe undergoes a large number of pre–
inflationary oscillations [18].