You are on page 1of 5

week ending

PRL 117, 021301 (2016) PHYSICAL REVIEW LETTERS 8 JULY 2016

Perfect Quantum Cosmological Bounce


Steffen Gielen
Theoretical Physics, Blackett Laboratory, Imperial College London, London SW7 2AZ, United Kingdom

Neil Turok
Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada
(Received 14 October 2015; revised manuscript received 14 March 2016; published 6 July 2016)
We study quantum cosmology with conformal matter comprising a perfect radiation fluid and a number
of conformally coupled scalar fields. Focusing initially on the collective coordinates (minisuperspace)
associated with homogeneous, isotropic backgrounds, we are able to perform the quantum gravity path
integral exactly. The evolution describes a “perfect bounce”, in which the Universe passes smoothly
through the singularity. We extend the analysis to spatially flat, anisotropic universes, treated exactly, and to
generic inhomogeneous, anisotropic perturbations treated at linear and nonlinear order. This picture
provides a natural, unitary description of quantum mechanical evolution across a cosmological bounce. We
provide evidence for a semiclassical description in which all fields pass “around” the cosmological
singularity along complex classical paths.

DOI: 10.1103/PhysRevLett.117.021301

Recent observations have revealed extraordinary sim- continuation could be performed around [the singularity]
plicity in the large-scale structure of the Universe: a but whether this would have any physical meaning is unclear”.
spatially flat geometry with nearly scale-invariant, We shall perform just such a continuation and interpret it.
Gaussian fluctuations. As yet, there are no indications of We study the quantum dynamics of a universe
tensor (gravitational wave) modes which would signal with conformal matter consisting of perfect radiation and
primordial inflation, nor complications such as curvature, conformally coupled scalar fields. We first analyze homo-
non-Gaussianity, or isocurvature modes. The simplicity of geneous, isotropic backgrounds, showing that the semi-
these findings seems at odds with expectations based on classical approximation to the path integral is exact. We
inflationary models predicting a “multiverse” with random then generalize to anisotropic spatially flat metrics, com-
and unpredictable behavior on large scales. We are there- puting the Feynman propagator and clarifying its analytic
fore encouraged to seek more economical explanations for properties. Finally, we tackle generic inhomogeneous
the state of the Universe. One of the oldest and simplest cosmologies, order by order in a perturbative expansion.
ideas [1] is that the big bang was a bounce. Such a bounce Although perturbation theory fails as we approach η ¼ 0,
is generally forbidden in classical general relativity, but we can maintain its validity by deforming the η-contour
might be allowed in quantum gravity [2]. into the complex η-plane and bypassing the singularity.
At the big bang singularity, the density and temperature This continuation respects all the symmetries of general
of matter diverges. General arguments indicate that the only relativity and yields an unambiguous result. For conformal
complete quantum field theories are those with a UV fixed matter, we show there is no quantum creation of scalar
point, i.e., which are conformally invariant at high energies density perturbations or gravitational waves across the
[3]. Therefore, it is of particular interest to study cosmol- bounce, indicating a well-defined vacuum. For theories
ogies with conformally invariant matter. A simple example with nontrivial running and/or soft breaking terms, one
is a spatially flat, homogeneous, and isotropic universe would find finite particle production.
filled with a perfect radiation fluid, with line element Weyl-invariant formulation.—We consider cosmology
with perfect radiation and M conformally coupled scalars
ds2 ¼ a2 ðηÞð−dη2 þ d~x2 Þ; ð1Þ χ~ ¼ ðχ 1 ; …; χ M Þ. It is convenient to “lift” Einstein gravity
to a classically Weyl-invariant action [5],
with aðηÞ ∝ η, where η is the conformal time. This is a good Z  
metric for all η ≠ 0, and furthermore possesses a unique pffiffiffiffiffiffi 1
S¼ d x D
−g ðð∂ϕÞ2 − ð∂~χ Þ2 Þ
analytic continuation around the singularity in the complex 2
η-plane. As we shall see, generic perturbations about this   
ðD − 2Þ 2 2 jJj μ
metric share these nice properties; hence, we term this þ ðϕ − χ~ ÞR − ρ pffiffiffiffiffiffi − J ∂ μ φ~ : ð2Þ
8ðD − 1Þ −g
cosmology a “perfect bounce”. Indeed, in a pioneering paper
[4] [see discussion following Eq. (5.19)], DeWitt anticipated The fluid is characterized by a densitized particle number
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffi
this idea, stating: “One might hope that an analytic flux J with jJj ≡ −gμν Jμ J ν ¼ −gn, where n is the
μ

0031-9007=16=117(2)=021301(5) 021301-1 © 2016 American Physical Society


week ending
PRL 117, 021301 (2016) PHYSICAL REVIEW LETTERS 8 JULY 2016
Z    
particle number density and ρðnÞ is the energy density [6]. m 1 α α
The Lagrange multiplier φ~ enforces particle number con- S¼ dt x_ x_ α − Nðκx xα þ 1Þ − φm
_ ; ð4Þ
2 N
servation: for a homogeneous, isentropic fluid there are no
additional constraints. The action (2) is invariant under
local Weyl transformations. While ϕ has the “wrong sign”
kinetic term, there is no physical ghost: one can go to with φ ≔ φV ~ 0 ðdn=dmÞ. Equation (4) is the action for a
“Einstein gauge” where ϕ2 − χ~2 is a constant, obtaining relativistic oscillator (κ > 0), free particle (κ ¼ 0), or
Einstein gravity coupled to scalars with positive kinetic “upside-down” oscillator (κ < 0), with mass m, extending
energy. However, other Weyl gauges may be more earlier work (see, e.g., Ref. [7]) for the case ρ ¼ 0.
convenient. From Eq. (2), the effective Newton’s constant is fixed by
Quantum mechanics of a cosmological bounce.—For ðϕ2 − χ~2 Þ ¼ −2ρx2 so that, for positive radiation density,
homogeneous, isotropic cosmologies, one can choose a there are two timelike regions of superspace, with x2 < 0
Weyl gauge in which the metric is static and the scalars and x0 > 0 or x0 < 0, respectively, describing “gravity,”
ðϕ; χ~Þ encode all of the dynamics. While the metric is and a spacelike region, x2 > 0, describing “antigravity.”
nonsingular in this gauge, the theory is still problematic The presence of radiation allows real solutions which pass
because the effective Planck mass, given by the coefficient smoothly from a “gravity” region through an “antigravity”
of R, can vanish, so that gravity becomes strongly coupled. region and back to a “gravity” region, i.e., classical bounces
Our strategy is to first identify this singularity in the [5]. Once anisotropies and inhomogeneities are included,
quantum propagator, and then understand how to analyti- generically there are no regular, real “bounce” solutions;
cally continue around it. Our key assumption, which we but there are regular, complex solutions which are defor-
shall test in various calculations, is that there are no mations of the classical bounces. We claim these are
singularities obstructing such a continuation. We set legitimate saddle points of the path integral and provide
D ¼ 4 unless otherwise stated. a consistent semiclassical description of a quantum bounce.
Fixing the metric to ds2 ¼ −N 2 ðtÞdt2 þ hij dxi dxj , To describe these complex solutions, it is convenient to
where hij is a metric of constant three-curvature define the Einstein-frame scale factor a by xα ¼ avα with v
Rð3Þ ¼ 6κ, Eq. (2) reduces to an (M þ 1)-vector satisfying v2 ¼ −1, so that real negative
Z _2 _ 2    and positive values of a represent the two “gravity” regions,
χ~ − ϕ κ 2 2 while imaginary values of a represent the “antigravity”
S ¼ V 0 dt þ N ðϕ − χ~ Þ − ρ − φ~ n_ ; ð3Þ
2N 2 region. We shall study generic complex solutions by
R pffiffiffi analytically continuing a into the complex plane and
where V 0 ¼ d3 x h is the comoving spatial volume around the singularity at a ¼ 0.
3
(which we take to be finite) and n ∝ ρ4 . Define Let us begin with the propagator for homogeneous,
xα ≔ ð2ρÞ−1=2 ðϕ; χ~Þ, with α ¼ 0; …; M. Then, with ηαβ ¼ isotropic universes.
diagð−1; 1; 1; …Þ and m ¼ 2V 0 ρ, which is a coordinate Feynman propagator from path integral.—The propa-
scalar, Eq. (3) becomes gator is given by a phase-space path integral [8],

Z  Z   
0 0 α
1=2
α Pα Pα m α
Gðx; mjx ; m Þ ≔ Dx DPα DmDpm DN exp i dt x_ Pα þ mp
_ m−N þ ðκx xα þ 1Þ : ð5Þ
−1=2 2m 2

Z  
Equation (5) is obtained from a canonical analysis of 0 0 0 m
Eq. (3) using Dirac’s algorithm, after “solving” two second- Gðx; mjx ; m Þ ¼ iδðm − m Þ dτ exp −i τ
2
class constraints to eliminate φ and its momentum [9].  pffiffiffi ðMþ1Þ=2
m κ
Fixing the gauge N_ ¼ 0 allows us to replace the path × pffiffiffi
integral over N with an ordinary integral over proper time 2iπ sinð κ τÞ
 pffiffiffi 
τ ¼ N in this gauge [10]. Integrating over m and pm yields pffiffiffi ðx2 þ x02 Þ cosð κτÞ − 2x⋅x0
a delta function for m conservation. The remaining path × exp im κ pffiffiffi :
2 sinð κτÞ
integrals are Gaussian and are computed exactly using the
classical solution ð7Þ
pffiffiffi pffiffiffi
x sin ½ κ τðt þ 1=2Þ þ x0 sin ½ κτð1=2 − tÞ For the Feynman propagator, τ runs from 0 to ∞. We
xðtÞ ¼ pffiffiffi : ð6Þ
sinð κτÞ insert a convergence factor −ϵτ=ð2mÞ in the exponent or,
alternatively, define the τ-contour by steepest descent from
The final result is the appropriate saddle point. The propagator has the usual

021301-2
week ending
PRL 117, 021301 (2016) PHYSICAL REVIEW LETTERS 8 JULY 2016

short-distance singularities but is otherwise regular. It is Ôa Gða; mja0 ; m0 Þ ¼ −2ima−ðMþD−2Þ δðm − m0 Þδða − a0 Þ;
defined by analytic continuation in x0 (or a) around these
ð10Þ
singularities through the half-plane in which it converges.
For example, in extending the amplitude from values for where
which x0 −x00 <j~x −~x0 j to values for which x0 −x00 >
j~x −~x0 j, we pass around the singularity in the lower-half d2 MþD−2 d C
x0 -plane. For κ ¼ 0, we obtain the massive free-particle Ôa ≡ þ − þ m2 ; ð11Þ
propagator on flat spacetime, for which these analyticity da2 a da a2
properties are well known [9,11]. with C ≡ − 14 ðM − 1Þ2 − ζ 2 − k2A þ ξðD − 2Þð2M þ D − 3Þ.
For homogenous, isotropic quantum cosmology with Here, ζ is the momentum on HM , kA that on the space
conformal matter, these calculations explicitly demonstrate of anisotropies, and ξ is a parameter governing the ordering
that the Feynman propagator is perfectly regular at the “big ambiguity identified by DeWitt [13] and clarified (in the
bang singularity” x2 ¼ 0 and that, since the path integrals phase-space path integral) by Kuchař [14]. Halliwell has
are Gaussian, the semiclassical approximation is exact. For made a strong case that ξ should be taken to be the
a large universe, the background is “heavy”: there is little conformal coupling on superspace, and further shown that
quantum spreading or backreaction from perturbations this is consistent with the DeWitt inner product [10].
[12]. More explicitly, provided radiation dominates, the Formally, setting D ¼ 2 and kA ¼ 0 in Eq. (11) removes
action associated with its density [the first exponent in the anisotropy degrees of freedom and reduces the equation
Eq. (7)] may be expressed as ð3=8πGÞHE V E , where H E to that for the isotropic, flat case.
and V E are the Hubble constant and three-volume in The Feynman propagator is obtained by the Wronskian
Einstein gauge, and G is Newton’s constant. We shall method from the positive and negative frequency modes,
perform our analytic continuations along complex paths for defined to be the solutions to Ôa ΨðaÞ ¼ 0 which
which this quantity increases at a rate sufficient to maintain are regular in the lower- and upper-half complex a-plane,
the Wentzel-Kramers-Brillouin (WKB) expansion. respectively. Up to normalization, they are given by
We now turn to anisotropic, spatially flat cosmologies.
Again, the Feynman propagator will be defined by analytic ð2;1Þ
Ψþ;− ðaÞ ¼ a−ðMþD−3Þ=2 H ν ðmaÞ; ð12Þ
continuation around its singularities.
Anisotropies.—Consider a spatially flat, anisotropic pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð1;2Þ
where ν ¼ 12 4C þ ðM þ D − 3Þ2 , and H ν are the
metric in a conformal gauge where the determinant of
Hankel functions of the first and second kind. The
the spatial metric is static,
propagator is
D−1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X X
D−1
ds2 ¼ −N 2 ðtÞdt2 þ e4 ðD−1Þ=ðD−2Þλi ðtÞ dx2i ; λi ¼ 0 π
Gða;mja0 ;m0 Þ ¼ mδðm − m0 Þðaa0 Þ−ðMþD−3Þ=2
i¼1 i¼1 2
ð8Þ ð1Þ ð2Þ
× ðHν ðma0 ÞHν ðmaÞθða − a0 Þ þ ða↔a0 ÞÞ:
ð13Þ
(restoring the dimension D). The action (4) becomes
Z    X    For the flat, isotropic case, we obtain the free-particle
m 1
S¼ dt x_ 2 − x2 λ_ 2i − N − φm
_ : ð9Þ propagator in the (M þ 1)-dimensional Minkowski space
2 N i
0
of fields ðϕ; χ~Þ; for M ¼ 0, G ¼ δðm − m0 Þe−imja−a j .
The modes appearing in the propagator have remarkable
This is again the action of a massive free particle,P now properties: an incoming positive-frequency mode continues
moving in a curved “superspace” metric dx2 − x2 i dλ2i , to an outgoing positive-frequency mode, without even a
with a Lorentzian signature for a2 ¼ −x2 > 0. Recall, phase shift. This surprising behavior follows from these
xα ¼ avα , with v2 ¼ −1, so the vector v parameterizes a facts: (i) the positive- (negative-)frequency modes are real
unit hyperboloid HM . Taking into account the constraint
P and decay exponentially as a runs to negative- (positive-)
i λi ¼ 0, the λi parameterize R
D−2
. Although the path imaginary values and (ii) the WKB approximation holds at
integration is hard to perform directly, there is sufficient large jaj, throughout the lower-half (upper-half) complex
symmetry present for the differential equation that the a-plane. Thus, the modes continue to the positive or
propagator satisfies, the Wheeler-DeWitt equation, to negative real a-axis where, by Schwarz reflection about
determine it completely. Fourier transforming to the con- the imaginary a-axis, their form is identical. The behavior
served momenta on HM × RD−2 , only the dependence on a of the DeWitt inner product is also of interest. In our
remains to be determined. This, however, is fixed uniquely example, where only the a dependence ↔ is nontrivial, this
by the Wheeler-DeWitt equation and analyticity properties. reduces to h1j2i ≡ aMþD−2 Ψ1 ðaÞi∂ a Ψ2 ðaÞ. If we normal-
The Wheeler-DeWitt equation reads ize the positive- and negative-frequency modes at negative

021301-3
week ending
PRL 117, 021301 (2016) PHYSICAL REVIEW LETTERS 8 JULY 2016

a, their norm is preserved as we pass to positive a, again by Eq. (1): since aðηÞ ∝ η, we can equivalently analytically
Schwarz reflection through ReðaÞ ¼ 0. continue in a or in conformal time η. Assuming planar
It may seem strange that the singular potential in Ôa symmetry, the perturbed metric is
has no effect on scattering states. However, the classical 
theory echoes this behavior. Consider the Hamiltonian
ds ∝ η ð−1 þ 2ϵϕÞdη2 þ ½1 þ 2ϵðψ þ γÞdx2
2 2
H ¼ N 12 ðp2a þ C=a2 − m2 Þ, with C a constant and N a
Lagrange multiplier imposing H ¼ 0. For C > 0 (which   
1 T
can only occur when anisotropies are present), the classical þ 1 þ ϵ 2ψ þ h dy2
solutions bounce and are nonsingular, 2
   
1
þ 1 þ ϵ 2ψ − hT dz2 ; ð15Þ
a2 ðτÞ ¼ ðτ − τ0 Þ2 þ C=m2 ; ð14Þ 2

with τ ¼ Nt and τ0 an arbitrary constant. We have aðτÞ ≈ where ϕ, ψ, γ, and hT are functions of η and x only. For
ðτ − τ0 Þ at large negative or positive τ. The potential simplicity, we set the second tensor mode in gyz to zero (its
introduces no net time advance or delay: it slows the dynamics are analogous to those of hT ).
trajectory but also causes a to bounce sooner. For C < 0, We adopt a coordinate system in which the radiation is
the classical solutions are singular but become well defined at rest everywhere (comoving gauge); the radiation
if one gives the solutions an infinitesimal imaginary part. density is ρðη; xÞ ¼ ρ0 ðηÞ½1 þ ϵδr ðη; xÞ. We expand the
As a runs in from −∞ and approachespthe ffiffiffiffiffiffiffi origin, it P and Einstein equations in powers of ϵ:
perturbations
turns down the imaginary a-axis to a ¼ −i −C=m before ϕðη; xÞ ¼ n≥1 ϵn−1 ϕðnÞ ðη; xÞ, etc. At order ϵn, Einstein’s
heading back to the origin and out to þ∞. The speed-up equation for the tensors hTðnÞ is
due to the potential and the delay from the excursion into
imaginary a cancel. This excursion represents the “anti- ∂ 2 hTðnÞ 2 ∂hTðnÞ ∂ 2 hTðnÞ
þ − ¼ Jn ðη; xÞ; ð16Þ
gravity” phase described in Ref. [5]. This suggests that the ∂η2 η ∂η ∂x2
“antigravity” regime is a consequence of trying to make the
quantum behavior look classical, and that it should not be where Jn is nonlinear in the lower-order perturbations.
taken literally as a new classical phase. There are four Einstein equations analogous to Eq. (16) for
Noting that, for C > 0, one of the two modes diverges the scalar perturbations δr , ϕ, ψ, and γ. In Ref. [9], we
at a ¼ 0, some authors advocate setting Ψð0Þ ¼ 0. derive the tensor and scalar Green’s functions and solve
Effectively this renders Ψ real, but (i) a real Ψ cannot Einstein’s equations order by order in ϵ.
describe a state of nonzero a-momentum, i.e., an expanding Consider the nonlinear extension of positive frequency
or a contracting universe, (ii) the DeWitt norm is zero, so solutions of the linearized equations, of wave number k0,
there is no meaningful notion of unitarity, and (iii) the pffiffi
correspondence principle is violated, due to a phase shift in ð1Þ e−ik0 η= 3
ψ ¼ A cosðk0 xÞ ;
the reflected mode which (as is easily seen) is independent k0 η
of m. Furthermore, as kA is increased, C goes negative. For e−ik0 η
small C < 0, both solutions vanish at a ¼ 0, so there would hTð1Þ ¼ B cosðk0 xÞ ; ð17Þ
be a jump in the allowed modes. At larger C < 0 their k0 η
prescription again gives a phase shift which violates the ð1Þ
with ϕð1Þ , γ ð1Þ , and δr determined in terms of ψ ð1Þ .
correspondence principle. None of these problems occur In Ref. [9], we give the positive frequency perturbations
with our modes. at second order, involving gamma functions and loga-
Inhomogeneities.—Finally, we tackle inhomogeneities rithms. After suitable definition of branch cuts, they are
in a perturbative expansion around a flat (κ ¼ 0) back- analytic in the lower-half η-plane, allowing us to avoid the
ground. The amplitude between in and out states is singularity at η ¼ 0 (see Fig. 1), just as for the homo-
calculated in the saddle point (semiclassical) expansion geneous background. To calculate the quantum production
around the appropriate complex classical solution. At of scalar or tensor perturbations across the bounce, we
lowest order, the initial state for the perturbations is a compute the leading terms of the nonlinear perturbations as
Gaussian wave functional for the incoming adiabatic η → ∞. At large jηj, hTð2Þ is
vacuum. The corresponding classical perturbation is the
pffiffiffi pffiffiffi
positive frequency mode: any mode mixing with negative pffiffi ð27 þ 16 3Þ cosð2k0 xÞ − 6 − 5 3
frequency modes across the bounce signals particle pro- ABe−ið1þ1= 3Þk0 η pffiffiffi þ ;
duction. Here, for simplicity, we set M ¼ 0 and study only ð21 þ 11 3Þik0 η
planar perturbations, taken to nonlinear order. ð18Þ
We solve the equations of motion following from Eq. (2)
in Einstein gauge. The background metric is given in where    are terms subleading in ðk0 ηÞ, and

021301-4
week ending
PRL 117, 021301 (2016) PHYSICAL REVIEW LETTERS 8 JULY 2016

cosmological bounce, paving the way for detailed inves-


tigations of new, simpler and more predictive bouncing
models.
The work of S. G. is supported by the People Programme
(Marie Curie Actions) of the European Union’s Seventh
Framework Programme (FP7/2007-2013) under REA
Grant No. 622339. Research at Perimeter Institute is
supported by the Government of Canada through
Industry Canada and by the Province of Ontario through
the Ministry of Research and Innovation.

[1] A. Friedmann, Über die Krümmung des Raumes, Z. Phys.


FIG. 1. Following a contour in the complex η-plane inside the 10, 377 (1922).
annulus ϵ < k0 jηj < 1=ϵ. The dashed region indicates branch [2] M. Bojowald, Absence of Singularity in Loop Quantum
cuts in the upper half-plane. Cosmology, Phys. Rev. Lett. 86, 5227 (2001); A. Ashtekar
and P. Singh, Loop quantum cosmology: A status report,
Classical Quantum Gravity 28, 213001 (2011).
A2 −ð2=pffiffi3Þik0 η [3] A. Shomer, A pedagogical explanation for the non-
ψ ð2Þ ðη; xÞ ∼ e ½1 þ 2 cosð2k0 xÞ þ    : ð19Þ renormalizability of gravity, arXiv:0709.3555.
12
[4] B. S. DeWitt, Quantum theory of gravity. I. The canonical
There is no mixing of positive and negative frequencies, theory, Phys. Rev. 160, 1113 (1967).
and, hence, no particle production. The full functions hTð2Þ [5] I. Bars, S. H. Chen, P. J. Steinhardt, and N. Turok, Anti-
and ψ ð2Þ decay exponentially for negative imaginary η. gravity and the big crunch/big bang transition, Phys. Lett. B
715, 278 (2012); I. Bars, P. J. Steinhardt, and N. Turok,
These properties extend to all orders in the perturbation
Local conformal symmetry in physics and cosmology, Phys.
expansion, unambiguously defining nonlinear positive Rev. D 89, 043515 (2014).
frequency modes [9]. [6] D. Brown, Action functionals for relativistic perfect fluids,
On the real η-axis, ψ ð2Þ does not decay at large jηj, Classical Quantum Gravity 10, 1579 (1993).
indicating a breakdown of perturbation theory. Metric [7] A. Ashtekar and R. S. Tate, An algebraic extension of Dirac
perturbations are gauge dependent; the “gauge-invariant” quantization: Examples, J. Math. Phys. (N.Y.) 35, 6434
Newtonian potential Φ [15] is Φð2Þ ∼ Oð1=k0 ηÞ at large jηj (1994).
at second order in ϵ. Comparing with the first-order [8] M. Henneaux and C. Teitelboim, Quantization of Gauge
potential Φð1Þ ∼ Oð1=k20 η2 Þ still indicates a breakdown of Systems (Princeton University Press, Princeton, NJ, 1994).
the expansion when k0 jηj ∼ 1=ϵ. This phenomenon has a [9] S. Gielen and N. Turok, Quantum cosmology with con-
simple physical explanation. The radiation fluid is gov- formal matter (to be published).
[10] J. J. Halliwell, Derivation of the Wheeler-DeWitt equation
erned by nonlinear dynamics, eventually leading to shocks.
from a path integral for minisuperspace models, Phys. Rev.
A careful analysis [16] shows that perturbation theory D 38, 2468 (1988).
breaks down when k0 jηj ∼ 1=ϵ. Perturbation theory can [11] N. D. Birrell and P. C. W. Davies, Quantum Fields in
thus be trusted for ϵ < k0 jηj < 1=ϵ (Fig. 1), where we Curved Space (Cambridge University Press, Cambridge,
obtain a perturbation expansion in ϵ for nonsingular, England, 1982).
nonlinear solutions in the complex η-plane. [12] C. Rovelli and E. Wilson-Ewing, Why are the effective
Conclusions.—Our results indicate that a valid semi- equations of loop quantum cosmology so accurate?, Phys.
classical approximation to quantum cosmology with con- Rev. D 90, 023538 (2014).
formal matter can be obtained from complex classical paths [13] B. DeWitt, Dynamical theory in curved spaces. I. A review
which avoid the classical big bang singularity. For homo- of the classical and quantum action principles, Rev. Mod.
geneous, isotropic backgrounds, and for anisotropic flat Phys. 29, 377 (1957).
[14] K. Kuchař, Measure for measure: Covariant skeletoniza-
backgrounds, we computed the Feynman propagator
tions of phase space path integrals for systems moving on
exactly, showing its large jaj behavior to be insensitive Riemannian manifolds, J. Math. Phys. (N.Y.) 24, 2122
to the divergent potential introduced by anisotropies near (1983).
the singularity. Finally, including inhomogeneities, we [15] J. M. Bardeen, Gauge-invariant cosmological perturbations,
found global, positive frequency modes; perturbation Phys. Rev. D 22, 1882 (1980).
theory fails at large times, but this is physically well [16] U.-L. Pen and N. Turok, Shocks in the early Universe,
understood [16]. Our investigations suggest the existence arXiv:1510.02985.
of a consistent and complete semiclassical description of a

021301-5

You might also like