This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Jackiw-Teitelboim and Kantowski-Sachs quantum cosmology

Georgios Fanaras and Alexander Vilenkin Institute of Cosmology, Department of Physics and Astronomy,
Tufts University, Medford, MA 02155, USA
Abstract

We study quantum cosmology of the 2D2D Jackiw-Teitelboim (JT) gravity with Λ>0\Lambda>0 and calculate the Hartle-Hawking (HH) wave function for this model in the minisuperspace framework. Our approach is guided by the observation that the JT dynamics can be mapped exactly onto that of the Kantowski-Sachs (KS) model describing a homogeneous universe with spatial sections of S1×S2S^{1}\times S^{2} topology. This allows us to establish a JT-KS correspondence between the wave functions of the models. We obtain the semiclassical Hartle-Hawking wave function by evaluating the path integral with appropriate boundary conditions and employing the methods of Picard-Lefschetz theory. The JT-KS connection formulas allow us to translate this result to JT gravity, define the HH wave function and obtain a probability distribution for the dilaton field.

1 Introduction

In quantum cosmology the entire universe is treated quantum mechanically and is described by a wave function, rather than by a classical spacetime. The wave function Ψ(g,ϕ)\Psi(g,\phi) is defined on the space of all 3-geometries (gg) and matter field configurations (ϕ\phi), called superspace. It can be found by solving the Wheeler-DeWitt (WDW) equation

Ψ=0,{\cal H}\Psi=0, (1.1)

where {\cal H} is the Hamiltonian operator. Alternatively, one can consider the transition amplitude from the initial state (g,ϕ)(g^{\prime},\phi^{\prime}) to the final state (g,ϕ)(g,\phi), which can be expressed as a path integral,

G(g,ϕ|g,ϕ)=(g,ϕ)(g,ϕ)eiS,G(g,\phi|g^{\prime},\phi^{\prime})=\int_{(g^{\prime},\phi^{\prime})}^{(g,\phi)}~{}e^{\textstyle{iS}}, (1.2)

where SS is the action and the integration is over the histories interpolating between the initial and final configurations. In general, GG is a Green’s function of the WDW equation Teitelboim . But if (g,ϕ)(g^{\prime},\phi^{\prime}) is at the boundary of superspace, or if the geometries that are being integrated over have a single boundary at gg, then GG is a solution of the WDW equation and the path integral (1.2) may be used to define a wave function of the universe.

The choice of the boundary conditions for the WDW equation and of the class of paths included in the path integral representation of Ψ\Psi has been a subject of ongoing debate. The most developed proposals in this regard are the Hartle-Hawking HH and the tunneling Vilenkin:1987kf ; Vilenkin:1994rn wave functions.111For early work closely related to HH and tunneling proposals, see Refs. Vilenkin:1982de and Linde:1983mx ; Rubakov:1984bh ; Vilenkin:1984wp ; Zeldovich:1984vk respectively. The intuition behind both of these proposals is that the universe originates ‘out of nothing’ in a non-singular way. But despite a large amount of work, a consensus on the precise definition of these wave functions has not yet been reached. In fact, the two wave functions are often confused with one another.

The Hartle-Hawking (HH) wave function is usually defined in terms of a Euclidean path integral,

ΨHH(g,ϕ)=(g,ϕ)eSE,\Psi_{HH}(g,\phi)=\int^{(g,\phi)}e^{\textstyle{-S_{E}}}, (1.3)

where SES_{E} is the Euclidean action and the integration is over regular compact geometries with a single boundary on which the boundary values (g,ϕ)(g,\phi) are specified. The tunneling wave function is defined either by an outgoing-wave boundary condition in superspace or by a path integral over Lorentzian histories interpolating between a vanishing 3-geometry and the configuration (g,ϕ)(g,\phi). Here we will focus on the HH wave function; the tunneling wave function will be discussed in a separate publication.

In the last few years there has been a renewed interest in quantum cosmology, inspired by recent work on the exactly soluble (1+1)(1+1)-dimensional quantum gravity model – the Jackiw-Teitelboim (JT) gravity Jackiw:1984je ; Teitelboim:1983ux . This model can be thought of as a quantum theory of a one-dimensional closed universe. Apart from the scale factor aa, it also includes an evolving scalar field ϕ\phi – the dilaton, which makes a comparison with higher-dimensional models somewhat less informative. On the positive side, one can hope that exact solubility of the model may provide new insights into the nature of the wave function of the universe.222 An exact quantization of JT model was first developed by Henneaux Henneaux . For recent discussions see Refs. Maldacena ; Iliesiu ; Trivedi ; Stanford and references therein.

The HH wave function for JT gravity has been recently discussed in the interesting paper by Maldacena, Turiaci and Yang (MTY) Maldacena . They calculated the wave function in the leading semiclassical order in the limit of large a and included the pre-exponential factor suggested by the Schwarzian analysis. In difference from the Hartle and Hawking approach, MTY focused on the outgoing branch of the wave function, describing expanding universes at large a. As a result the asymptotic behavior of Ψ\Psi is more consistent with the tunneling boundary conditions.

Another interesting recent work is the paper by Iliesiu, Kruthoff, Turiaci and Verlinde (IKTV) Iliesiu . They presented an exact solution to the WDW equation of JT gravity, which they interpreted as the HH wave function, but their choice of boundary conditions was different from the earlier literature. Hartle & Hawking and most of the subsequent authors required that geometries included in the path integral close off smoothly in the limit of small universes. Instead, IKTV imposed a boundary condition in the opposite limit, requiring that the wave function exhibits Schwarzian behavior when the universe is large. The assumption of regularity and closure is implicit in their discussion, but these conditions are not explicitly enforced. The resulting wave function agrees with the semiclassical analysis of MTY in the appropriate limit. However, IKTV note an unexpected feature: the wave function develops a strong singularity at a finite value of the scale factor.

In the present paper we take a different approach to JT quantum cosmology. It is based on the observation of MTY that JT model can be obtained from 4D4D gravity by dimensional reduction. We shall use this connection between 2D2D and 4D4D theories as a guide to defining the cosmological HH wave function in the JT model. In their paper MTY discussed a dimensional reduction from a nearly extremal Schwarzschild-de Sitter solution, with the extra two dimensions compactified on a sphere (see also Fabbri ; Bousso for earlier work). Since our emphasis is on the cosmological aspects of the theory, we find it more useful to consider a cosmological 4D4D model describing a homogeneous universe with spatial sections having S1×S2S^{1}\times S^{2} topology, known as the Kantowski-Sachs model. The main difference from the MTY and IKTV work is that we impose the boundary conditions in the small universe limit, requiring that the geometry closes off in a regular way.

We begin in the next section by reviewing JT gravity and its quantization, discussing in particular the semiclassical analysis of MTY and the exact solutions of IKTV. We argue that these solutions are not suitable to represent the HH wave function. We also discuss how JT model can be obtained by dimensional reduction from 4D4D gravity.

In Section 3 we review the quantum cosmology of the Kantowski-Sachs (KS) model, following largely the treatment of Halliwell and Louko (HL) HL . We establish an exact correspondence between the WDW equations for KS and JT models. Furthermore, we show that the transition amplitude between states with specified initial and final scale factors calculated by HL is closely related to the wave function found by IKTV. It follows from this analysis that their wave function satisfies an equation with a singular source and thus is not a solution of the WDW equation. This accounts for the divergence of the wave function pointed out by IKTV.

The semiclassical HH wave function for the KS model is discussed in Section 4. HL studied this wave function only for a vanishing cosmological constant, Λ=0\Lambda=0. Here we will need to extend their analysis to the case of Λ>0\Lambda>0, which is significantly more complicated. We impose the boundary conditions of smooth closure in the limit of small universes and follow standard methods to reduce the problem to evaluation of a lapse (NN) integral over some contour 𝒞{\cal C} in the complex NN plane. The choice of the contour 𝒞{\cal C} is restricted by the requirements that the HH wave function is expected to satisfy. We argue that there is only a single acceptable choice, with all other acceptable choices equivalent to it.

In the semiclassical limit the dominant contribution to the integral is given by saddle points of the action. We find these saddle points, as well as the steepest descent and ascent lines, and use the Picard-Lefschetz prescription to deform the contour so that the integral becomes absolutely convergent. The integral is then evaluated in the WKB approximation for the range of parameters most relevant for the connection to JT.

In Section 5 we use the HH wave function calculated in the preceding section to find the probability distribution for the radius of S2S^{2} at a given radius of S1S^{1} in our S1×S2S^{1}\times S^{2} model. In Section 6 we use the correspondence between JT and KS models to define the HH wave function in JT gravity. We use this wave function to determine the probability distribution for the dilaton field ϕ\phi. Our results are summarized and discussed in Section 7. Some technical details are relegated to the Appendix.

2 JT gravity

2.1 The action

The action of the JT model is Jackiw:1984je ; Teitelboim:1983ux

S=d2xgϕ(R2H2)2BϕbK,S=\int d^{2}x\sqrt{-g}\phi(R-2H^{2})-2\int_{B}\phi_{b}K, (2.1)

where RR is the 2D2D spacetime curvature, H=constH={\rm const}, ϕ\phi is the dilaton field, ϕb\phi_{b} is its value at the boundary, and KK is the extrinsic curvature of the boundary curve BB. Throughout the paper we shall assume that H>0H>0. Variation with respect to ϕ\phi yields R=2H2R=2H^{2}, telling us that the 2D2D spacetime is a de Sitter space with expansion rate HH.

With the metric represented as

ds2=N2dt2+a2dx2,ds^{2}=-N^{2}dt^{2}+a^{2}dx^{2}, (2.2)

where 0<x<2π0<x<2\pi and NN is the lapse function, the state vector is a functional

Ψ[a(x),ϕ(x)].\Psi[a(x),\phi(x)]. (2.3)

We can choose the gauge so that ϕb=const\phi_{b}={\rm const} at the boundary. Furthermore, we are going to adopt a minisuperspace picture, where a=a(t)a=a(t), independent of xx, and the boundary is a circle, t=constt={\rm const}. Then Ψ\Psi is an ordinary function Ψ(a,ϕ)\Psi(a,\phi). It has been shown in Iliesiu that due to the simplicity of the model, the wave functional (2.3) can be recovered from the minisuperspace wave function Ψ(a,ϕ)\Psi(a,\phi). Here, we shall restrict our analysis to the minisuperspace model with a=a(t),ϕ=ϕ(t)a=a(t),~{}\phi=\phi(t).

After integration by parts the action (2.1) can be represented as

S=4π𝑑t(a˙ϕ˙N+NH2aϕ),S=-4\pi\int dt\left(\frac{{\dot{a}}{\dot{\phi}}}{N}+NH^{2}a\phi\right), (2.4)

where dots stand for derivatives with respect to tt. We are going to use the gauge N=constN={\rm const}. The momenta conjugate to aa and ϕ\phi are

Πa=4πNϕ˙,Πϕ=4πNa˙.\Pi_{a}=-\frac{4\pi}{N}{\dot{\phi}},~{}~{}~{}~{}\Pi_{\phi}=-\frac{4\pi}{N}{\dot{a}}. (2.5)

The equations of motion obtained by varying the action with respect to aa and ϕ\phi are

a¨H2a=0,{\ddot{a}}-H^{2}a=0, (2.6)
ϕ¨H2ϕ=0,{\ddot{\phi}}-H^{2}\phi=0, (2.7)

where we have set N=1N=1. The Hamiltonian constraint is obtained by varying with respect to NN:

a˙ϕ˙=H2aϕ,{\dot{a}}{\dot{\phi}}=H^{2}a\phi, (2.8)

or

ΠaΠϕ=16π2H2aϕ.\Pi_{a}\Pi_{\phi}=16\pi^{2}H^{2}a\phi. (2.9)

The classical solution of these equations is

a=a0cosh(Ht),ϕ=ϕ0sinh(Ht)a=a_{0}\cosh(Ht),~{}~{}~{}~{}\phi=\phi_{0}\sinh(Ht) (2.10)

with a0,ϕ0=consta_{0},\phi_{0}={\rm const}. We shall set a0=H1a_{0}=H^{-1}, so that the metric covers the full de Sitter space in a nonsingular way.

2.2 Semiclassical wave function

To lowest order in the WKB approximation, the wave function is given by

ΨeiScl,\Psi\sim e^{iS_{cl}}, (2.11)

where SclS_{cl} is the classical action,

Scl=202π𝑑xaϕbK=4πaϕbK.S_{cl}=-2\int_{0}^{2\pi}dx~{}a\phi_{b}K=-4\pi a\phi_{b}K. (2.12)

Here, we used the fact that R=2H2R=2H^{2} in the classical solution, so only the surface terms make a contribution, and that ϕ\phi and KK are constant on the boundary. Following the no-boundary philosophy, we assume that the (Euclideanized) geometry closes off smoothly, so that there is no boundary contribution at a=0a=0.

In the classically allowed region (a>H1)(a>H^{-1}), the extrinsic curvature KK is given by

K=a˙a=Htanh(Ht)=a1H2a21,K=\frac{\dot{a}}{a}=H\tanh(Ht)=a^{-1}\sqrt{H^{2}a^{2}-1}, (2.13)

where we have used Eq. (2.10) with a0=H1a_{0}=H^{-1}. Substituting this in Eqs. (2.12) and (2.11), we obtain

Ψexp(4πiϕbH2a21)\Psi\propto\exp\left(-4\pi i\phi_{b}\sqrt{H^{2}a^{2}-1}\right) (2.14)

A linearly independent WKB wave function is a complex conjugate of (2.14). A general WKB solution is a linear combination of the two. The semiclassical approximation applies when the action is large, ϕbH2a211\phi_{b}\sqrt{H^{2}a^{2}-1}\gg 1.

The momentum operator Πϕ\Pi_{\phi} acting on Ψ\Psi gives

ΠϕΨ=iϕΨ=4πH2a21Ψ.\Pi_{\phi}\Psi=-i\partial_{\phi}\Psi=-4\pi\sqrt{H^{2}a^{2}-1}\Psi. (2.15)

The classical momentum is given by Eq. (2.5), so we get

a˙=H2a21.{\dot{a}}=\sqrt{H^{2}a^{2}-1}. (2.16)

This agrees with the expanding branch of the classical solution (2.10). The complex conjugate wave function describes a contracting universe.

2.3 Exact solutions of the WDW equation

The WDW equation corresponding to the Hamiltonian constraint (2.8) is

(aϕ+16π2H2aϕ)Ψ~=0.(\partial_{a}\partial_{\phi}+16\pi^{2}H^{2}a\phi){\tilde{\Psi}}=0. (2.17)

Here,

Ψ~=Ψ/a{\tilde{\Psi}}=\Psi/a (2.18)

and the factor 1/a1/a comes from the factor ordering indicated by the exact quantization of the JT model by Henneaux Henneaux (see IKTV Iliesiu for a detailed explanation). Following MTY Maldacena , we introduce new variables

u=ϕ2,v=(2π)2(H2a21).u=\phi^{2},~{}~{}~{}v=(2\pi)^{2}(H^{2}a^{2}-1). (2.19)

Then the WDW equation becomes

(uv+1)Ψ~=0.(\partial_{u}\partial_{v}+1){\tilde{\Psi}}=0. (2.20)

Yet another change of variables

T=uv,ξ=12lnvuT=\sqrt{uv},~{}~{}~{}\xi=\frac{1}{2}\ln\frac{v}{u} (2.21)

brings the equation to a separable form

1TT(TTΨ~)+1T2ξ2Ψ~4Ψ~=0.-\frac{1}{T}\partial_{T}(T\partial_{T}{\tilde{\Psi}})+\frac{1}{T^{2}}\partial_{\xi}^{2}{\tilde{\Psi}}-4{\tilde{\Psi}}=0. (2.22)

With the ansatz

Ψ~m=emξfm(T){\tilde{\Psi}}_{m}=e^{m\xi}f_{m}(T) (2.23)

we obtain an equation for fm(T)f_{m}(T):

fm′′+1Tfmm2T2fm+4fm=0.{f_{m}}^{\prime\prime}+\frac{1}{T}{f_{m}}^{\prime}-\frac{m^{2}}{T^{2}}f_{m}+4f_{m}=0. (2.24)

The solution is

fm(T)=Zm(2T),f_{m}(T)=Z_{m}(2T), (2.25)

where ZmZ_{m} is a Bessel function.

Following MTY, IKTV required that the wave function should describe an expanding universe in the limit of large aa. Then the appropriate choice of Bessel functions is Hm(2)(2T)H_{m}^{(2)}(2T). The general solution of the WDW equation is then a linear combination of functions of the form (2.23):

Ψ~m=(vu)m/2Hm(2)(2T).{\tilde{\Psi}}_{m}=\left(\frac{v}{u}\right)^{m/2}H_{m}^{(2)}(2T). (2.26)

In terms of the variables aa and ϕ\phi, the argument of the Bessel functions is

2T=4πϕH2a21.2T=4\pi\phi\sqrt{H^{2}a^{2}-1}. (2.27)

The asymptotic form of the Bessel functions at large TT is Hm(2)(2T)T1/2e2iTH_{m}^{(2)}(2T)\propto T^{-1/2}e^{-2iT}; hence

Ψm(Haϕ1)(aϕ)m+1/2exp(4πiϕH2a21),{\Psi}_{m}(Ha\phi\gg 1)\propto\left(\frac{a}{\phi}\right)^{m+1/2}\exp\left(-4\pi i\phi\sqrt{H^{2}a^{2}-1}\right), (2.28)

where we have accounted for the factor 1/a1/a relating Ψm\Psi_{m} and Ψ~m{\tilde{\Psi}}_{m}. All these functions have the same asymptotic exponential factor as the WKB wave function (2.14). So in order to choose between them one has to determine the pre-exponential factor.

In the path integral formulation, the semiclassical pre-factor is determined by quantum fluctuations about the classical solution. In the JT model these are fluctuations in the shape of the boundary curve, which are described by the Schwarzian theory and yield a one-loop pre-factor (ϕ/a)3/2(\phi/a)^{3/2} at large aa StanfordWitten . It is shown in StanfordWitten that this result is one-loop exact, so there are no further corrections. This pre-factor is obtained by setting m=2m=-2 in (2.28). Then the exact wave function takes the form

Ψ(a,ϕ)=aϕ2H2a21H2(2)(4πϕH2a21)(Ha>1).\Psi(a,\phi)=\frac{a\phi^{2}}{H^{2}a^{2}-1}H_{2}^{(2)}\left(4\pi\phi\sqrt{H^{2}a^{2}-1}\right)~{}~{}~{}~{}(Ha>1). (2.29)

Analytic continuation of this wave function to Ha<1Ha<1 is not unique because of the singularity at Ha=1Ha=1. IKTV specify the wave function in the entire range of aa by replacing the Hankel function in (2.29) with K2(iϕ2(H2a21iϵ))K_{2}\left(i\sqrt{\phi^{2}(H^{2}a^{2}-1-i\epsilon)}\right), which gives333The inclusion of the term iϵi\epsilon with ϵ+0\epsilon\to+0 is needed to make the solution well-defined on the branch cut.

Ψ(a,ϕ)=2iπaϕ2H2a21K2(4πϕH2a21)(Ha<1).\Psi(a,\phi)=\frac{2i}{\pi}\frac{a\phi^{2}}{H^{2}a^{2}-1}K_{2}\left(4\pi\phi\sqrt{H^{2}a^{2}-1}\right)~{}~{}~{}~{}(Ha<1). (2.30)

IKTV identify the wave function (2.29),(2.30) with the HH wave function for JT gravity. There are however some problems with this identification. We first note that one of the defining properties of the HH wave function is that it is real. This can be interpreted as reflecting the CPT invariance of the HH state HHH . On the other hand, the tunneling wave function is specified by the outgoing wave boundary condition, which in the present context requires that the large aa asymptotic of Ψ\Psi should only include terms corresponding to expanding universes. This seems to suggest that the wave function described by Eqs.(2.29),(2.30) is more appropriately interpreted as the tunneling wave function.

More importantly, the wave function (2.29) has a singularity at a=H1a=H^{-1}. It is actually not a solution of the WDW equation. We will show in Sec.III.C that it satisfies

Ψδ(ϕ)δ′′(aH1).{\cal H}\Psi\propto\delta(\phi)\delta^{\prime\prime}(a-H^{-1}). (2.31)

Hence it is not suitable for the role of HH or tunneling wave function.

IKTV have also proposed another candidate for the HH wave function:

Ψ(a,ϕ)=aϕ2H2a21J2(4πϕH2a21).\Psi(a,\phi)=\frac{a\phi^{2}}{H^{2}a^{2}-1}J_{2}\left(4\pi\phi\sqrt{H^{2}a^{2}-1}\right). (2.32)

This wave function is real and non-singular. However, its behavior in the classically forbidden range a<H1a<H^{-1} is very different from what is expected for the semiclassical HH wave function. We have

Ψ(a<H1)=aϕ21H2a2I2(4πϕ2(1H2a2))aϕ3/24π(1H2a2)5/4exp(4π|ϕ|1H2a2),\Psi(a<H^{-1})=\frac{a\phi^{2}}{1-H^{2}a^{2}}I_{2}\left(4\pi\sqrt{\phi^{2}(1-H^{2}a^{2})}\right)\sim\frac{a\phi^{3/2}}{\sqrt{4\pi}(1-H^{2}a^{2})^{5/4}}\exp\left(4\pi|\phi|\sqrt{1-H^{2}a^{2}}\right), (2.33)

where the last expression is the asymptotic form of Ψ\Psi assuming that the argument of I2I_{2} is large. As aa varies from a=0a=0 to aH1a\sim H^{-1}, the exponential factor in Ψ\Psi decreases, which is opposite to the expected behavior of the HH wave function.

2.4 Dimensional reduction

MTY discussed the relation between JT and Einstein 4D4D gravity using dimensional reduction from a nearly extremal Schwarzschild-de Sitter solution. Since our emphasis is on the cosmological aspects of the theory, we find it more useful to consider a cosmological 4D4D model describing a universe with spatial sections having S1×S2S^{1}\times S^{2} topology and the metric

ds2=dt2+a2(x,t)dx2+b2(x,t)dΩ2.ds^{2}=-dt^{2}+a^{2}(x,t)dx^{2}+b^{2}(x,t)d\Omega^{2}. (2.34)

Here, 0<x<2π0<x<2\pi and dΩ2d\Omega^{2} is the metric on a unit sphere. Substituting this in the Einstein-Hilbert action (in Planck units)

S=116πd4xg(4)(R(4)2Λ),S=\frac{1}{16\pi}\int d^{4}x\sqrt{-g^{(4)}}\left(R^{(4)}-2\Lambda\right), (2.35)

where Λ\Lambda is the 4D4D cosmological constant, and integrating over the angular variables we obtain

S=d2xg[b24R+12(b)2+1212Λb2].S=\int d^{2}x\sqrt{-g}\left[\frac{b^{2}}{4}R+\frac{1}{2}(\nabla b)^{2}+\frac{1}{2}-\frac{1}{2}\Lambda b^{2}\right]. (2.36)

Here, RR and gg are respectively the 2D2D curvature scalar and the metric determinant and we have omitted the boundary term.

Following Louis-Martinez , we can remove the gradient term in the action by a conformal rescaling

g¯μν=Ω2(b)gμν{\bar{g}}_{\mu\nu}=\Omega^{2}(b)g_{\mu\nu} (2.37)

with

dlnΩdlnb=12.\frac{d\ln\Omega}{d\ln b}=\frac{1}{2}. (2.38)

This has the solution

Ω=(b/2)1/2,\Omega=(b/2)^{1/2}, (2.39)

where we have chosen the normalization factor for future convenience. The action then reduces to

S=d2xg¯[ϕ¯R¯V(ϕ¯)],S=\int d^{2}x\sqrt{-{\bar{g}}}\left[{\bar{\phi}}{\bar{R}}-V({\bar{\phi}})\right], (2.40)

where ϕ¯=b2/4{\bar{\phi}}=b^{2}/4 and

V(ϕ¯)=2Λϕ¯12ϕ¯.V({\bar{\phi}})=2\Lambda\sqrt{\bar{\phi}}-\frac{1}{2\sqrt{\bar{\phi}}}. (2.41)

We define

ϕ¯=ϕ0+ϕ,{\bar{\phi}}=\phi_{0}+{\phi}, (2.42)

where ϕ0=1/4Λ\phi_{0}=1/4\Lambda, so that V(ϕ0)=0V(\phi_{0})=0. We shall assume that Λ1\Lambda\ll 1, so ϕ01\phi_{0}\gg 1. Then, for |ϕ|ϕ0|\phi|\ll\phi_{0} we can expand the potential (2.41) around ϕ=0\phi=0. Neglecting quadratic and higher order terms in the expansion, we obtain an approximate JT action

Sϕ0d2xg¯R¯+d2xg¯ϕ(R¯2Λ¯),S\approx\phi_{0}\int d^{2}x\sqrt{-{\bar{g}}}{\bar{R}}+\int d^{2}x\sqrt{-{\bar{g}}}{\phi}\left({\bar{R}}-2{\bar{\Lambda}}\right), (2.43)

where Λ¯=2Λ3/2{\bar{\Lambda}}=2\Lambda^{3/2}.

Since the second term in (2.43) already includes a factor of ϕ\phi, we can use

g¯μν12Λgμν,R¯2ΛR.{\bar{g}}_{\mu\nu}\approx\frac{1}{2\sqrt{\Lambda}}g_{\mu\nu},~{}~{}~{}{\bar{R}}\approx 2\sqrt{\Lambda}R. (2.44)

Hence, in the same approximation we can rewrite the action in terms of the original metric gμνg_{\mu\nu} and the cosmological constant Λ\Lambda as

Sϕ0d2xgR+d2xgϕ(R2Λ),S\approx\phi_{0}\int d^{2}x\sqrt{-{g}}{R}+\int d^{2}x\sqrt{-{g}}{\phi}\left({R}-2{\Lambda}\right), (2.45)

The above analysis suggests that in the appropriate limit the dynamics of the 4D4D cosmological model (2.34) is well approximated by that of the JT gravity (2.1) with Λ=H2\Lambda=H^{2}. The radius of the sphere S2S^{2} in this limit is bH1b\approx H^{-1}. We will focus on this regime in most of the paper, but in the next section we will see that in the minisuperspace setting the two models are even more closely related and can be mapped onto one another for arbitrary values of the scale factors aa and bb.

3 Kantowski-Sachs model

3.1 Classical dynamics

As already mentioned, our focus in this paper will be on homogeneous minisuperspace models. Hence we will use a homogeneous version of the model (2.34), with the scale factors aa and bb independent of xx, for dimensional reduction. This is the Kantowski-Sachs (KS) model KS describing a homogeneous universe with spatial sections of S1×S2S^{1}\times S^{2} topology.

Following Halliwell and Louko HL , we represent the metric of the KS model as

ds2=N2a2dτ2+a2dx2+b2dΩ2,ds^{2}=-\frac{N^{2}}{a^{2}}d\tau^{2}+a^{2}dx^{2}+b^{2}d\Omega^{2}, (3.1)

where NN, aa and bb are functions of time τ\tau, which we can choose to vary in the range 0<τ<10<\tau<1. After substituting this in the Lorentzian Einstein-Hilbert action with a cosmological constant ΛH2\Lambda\equiv H^{2} and integrating over xx and over the angular variables, the action reduces to

S=π01𝑑τ[b˙c˙N+N(H2b21)],S=-\pi\int_{0}^{1}d\tau\left[\frac{{\dot{b}}{\dot{c}}}{N}+N(H^{2}b^{2}-1)\right], (3.2)

where we have introduced a new variable c=a2bc=a^{2}b.

The factor 1/a21/a^{2} is added in the first term of (3.1) in order to simplify the equations of motion, which take the form

b¨=0,{\ddot{b}}=0, (3.3)
c¨N22H2b=0,\frac{\ddot{c}}{N^{2}}-2H^{2}b=0, (3.4)

where overdots stand for derivatives with respect to τ\tau. The constraint equation is obtained by varying the action with respect to NN:

b˙c˙NN(H2b21)=0.\frac{{\dot{b}}{\dot{c}}}{N}-N(H^{2}b^{2}-1)=0. (3.5)

An important solution of these equations is obtained by setting b˙=0{\dot{b}}=0. Then Eq.(3.5) tells us that b=H1b=H^{-1} and Eq.(3.4), expressed in terms of the proper time variable t=𝑑τ/a(τ)t=\int d\tau/a(\tau), becomes

d2adt2=H2a,\frac{d^{2}a}{dt^{2}}=H^{2}a, (3.6)

which has a solution

a(t)=H1cosh(Ht),a(t)=H^{-1}\cosh(Ht), (3.7)

where we have set N=1N=1. This is the Nariai solution, which is a product of a 2D2D de Sitter space and a 2-sphere of radius H1H^{-1} Nariai .

It follows from Eq.(3.3) that b˙{\dot{b}} cannot change sign, indicating that the Nariai solution is unstable. If we perturb it by giving the radius of the sphere bb a slight velocity, the sphere will collapse if b˙<0{\dot{b}}<0 and will expand to infinite size if b˙>0{\dot{b}}>0.

The Euclidean continuation of the Nariai solution is a product of two spheres of radius H1H^{-1}. It is often referred to as the Nariai instanton and describes nucleation of extremal black holes in de Sitter space Ginsparg:1982rs ; Bousso:1996au .

3.2 WDW equation

The quantum cosmology of the KS model has been studied by a number of authors Laflamme ; HL ; Conradi ; Anninos:2012ft ; Hertog and Conti . Some exact solutions of the WDW equation have been found and semiclassical methods have been used to study more general solutions. Here we will follow the method of Halliwell and Louko (HL) HL which allows one not only to find the saddle points of the action, but also helps to determine which saddle points contribute to the semiclassical wave function. This method will also be useful for interpreting the solution (2.29) found by IKTV.

The momenta conjugate to the variables bb and cc are

pb=πc˙/N,pc=πb˙/N.p_{b}=-\pi{\dot{c}}/N,~{}~{}~{}~{}p_{c}=-\pi{\dot{b}}/N. (3.8)

Using this in the constraint equation (3.5) and replacing pbi/bp_{b}\to-i\partial/\partial b, pci/cp_{c}\to-i\partial/\partial c, we obtain the WDW equation

πΨ=[bc+π2(H2b21)]Ψ=0.\pi{\cal H}\Psi=\left[\partial_{b}\partial_{c}+\pi^{2}(H^{2}b^{2}-1)\right]\Psi=0. (3.9)

This equation can be simplified by introducing a new variable ξ\xi, which is related to bb as dξ=π2(H2b21)dbd\xi=\pi^{2}(H^{2}b^{2}-1)db. Choosing the integration constant so that ξ(Hb=1)=0\xi(Hb=1)=0, we have

ξ=π23H(H3b33Hb+2)=π23H(Hb1)2(Hb+2).\xi=\frac{\pi^{2}}{3H}(H^{3}b^{3}-3Hb+2)=\frac{\pi^{2}}{3H}(Hb-1)^{2}(Hb+2). (3.10)

We also introduce the variable ρ=cH3\rho=c-H^{-3}; then the WDW equation becomes

(ξρ+1)Ψ=0.(\partial_{\xi}\partial_{\rho}+1)\Psi=0. (3.11)

We note that this equation has the same form as Eq.(2.20) for the JT model. The difference is that Eq.(2.20) is for the function Ψ~=Ψ/a{\tilde{\Psi}}=\Psi/a, where the factor 1/a1/a appeared due to a particular choice of the factor ordering. Following the same steps as in Sec.2.3, we find that Eq.(3.11) has exact solutions

Ψm=(ρξ)m/2Hm(2)(2ξρ).\Psi_{m}=\left(\frac{\rho}{\xi}\right)^{m/2}H_{m}^{(2)}(2\sqrt{\xi\rho}). (3.12)

The solution with m=2m=-2 corresponds to the IKTV solution (2.29). We expect this solution to agree with (2.29) when bH1b\approx H^{-1}. The argument of the Hankel function in (3.12) is

2ξρ=2πH2Hb+23(Hb1)H3a2b12πH2(Hb1)H2a21.2\sqrt{\xi\rho}=\frac{2\pi}{H^{2}}\sqrt{\frac{Hb+2}{3}}(Hb-1)\sqrt{H^{3}a^{2}b-1}\approx\frac{2\pi}{H^{2}}(Hb-1)\sqrt{H^{2}a^{2}-1}. (3.13)

Comparing this with the argument of the Hankel function in (2.29), we can identify

Hb1H22ϕ.\frac{Hb-1}{H^{2}}\approx 2\phi. (3.14)

It is interesting to note that the correspondence between the two wave function extends beyond this approximation. If we define

a~=(Hb)1/2a,ϕ=(Hb1)2H2Hb+23,{\tilde{a}}=(Hb)^{1/2}a,~{}~{}~{}~{}{\phi}=\frac{(Hb-1)}{2H^{2}}\sqrt{\frac{Hb+2}{3}}, (3.15)

then a~Ψ(a~,ϕ){\tilde{a}}\Psi({\tilde{a}},{\phi}) exactly reproduces the wave function (2.29). More generally, the transformation (3.15) can be used to obtain a solution to the WDW equation of the JT model from that of the KS model and vice versa. Note also that a~{\tilde{a}} and ϕ{\phi} are simply related to the variables ξ\xi and ρ\rho:

ξ=4π2H3ϕ2,ρ=H3(H2a~21).\xi=4\pi^{2}H^{3}{\phi}^{2},~{}~{}~{}~{}\rho=H^{-3}(H^{2}{\tilde{a}}^{2}-1). (3.16)

We thus see that JT and KS minisuperspace models are formally equivalent to one another. This equivalence, however, does not extend beyond minisuperspace. In the JT case, the minisuperspace wave function can be extended to a wave function in full superspace, but in the KS model the number of variables in the wave function and the pre-exponential factor depend on which perturbation modes are included in the minisuperspace truncation. Equivalence between the two models at the minisuperspace level will nevertheless be sufficient for our purposes here.

3.3 Transition amplitude

We now consider the transition amplitude from the initial state {b,c}\{b^{\prime},c^{\prime}\} to the final state {b,c}\{b,c\}. We will be particularly interested in the initial state

{b,c}={H1,H3}\{b^{\prime},c^{\prime}\}=\{H^{-1},H^{-3}\} (3.17)

corresponding to the bounce point of the Nariai solution. We shall refer to it as ‘Nariai initial conditions’.

The general framework for calculating transition amplitudes has been discussed by HL HL . For a general minisuperspace model described by the Hamiltonian

=12fαβ(q)pαpβ+V(q),{\cal H}=\frac{1}{2}f^{\alpha\beta}(q)p_{\alpha}p_{\beta}+V(q), (3.18)

where qαq^{\alpha} are the generalized coordinates and pαp_{\alpha} their conjugate momenta, the transition amplitude between qq^{\prime} and qq is

G(q|q)=0𝑑N𝒟p𝒟qeiS=0𝑑Nq,N|q,0.G(q|q^{\prime})=\int_{0}^{\infty}dN\int\mathcal{D}p\mathcal{D}qe^{iS}=\int_{0}^{\infty}dN\langle q,N|q^{\prime},0\rangle. (3.19)

Here NN is the lapse parameter, the action is

S=01𝑑τ(pαq˙αN)S=\int_{0}^{1}d\tau\left(p_{\alpha}{\dot{q}}^{\alpha}-N{\cal H}\right) (3.20)

and the path integral is over histories interpolating between qq^{\prime} and qq.

HL show that the amplitude (3.19) satisfies

G(q|q)=iq,0|q,0=iδ(q,q),{\cal H}G(q|q^{\prime})=-i\langle q,0|q^{\prime},0\rangle=-i\delta(q,q^{\prime}), (3.21)

where

=122+ζR+V{\cal H}=-\frac{1}{2}\nabla^{2}+\zeta R+V (3.22)

is the Hamiltonian operator, 2\nabla^{2} and RR are the Laplacian and the curvature scalar in the metric fαβf_{\alpha\beta}, and ζ\zeta is the conformal coupling. The magnitude of ζ\zeta depends on the dimension of superspace and vanishes in the case of 2D2D, which is of interest to us here. We see from Eq.(3.21) that GG is a Green’s function of the WDW equation.

For the KS model the Hamiltonian is quadratic and the path integral in (3.19) can be performed exactly. Then, up to an overall multiplicative constant, HL found that the amplitude reduces to

G(b,c|b,c)=0dNNexp[i2(αNβN)],G(b,c|b^{\prime},c^{\prime})=\int_{0}^{\infty}\frac{dN}{N}\exp\left[\frac{i}{2}\left(\alpha N-\frac{\beta}{N}\right)\right], (3.23)

where

α=1H23(b2+bb+b2),β=(2π)2(cc)(bb).\alpha=1-\frac{H^{2}}{3}(b^{2}+bb^{\prime}+{b^{\prime}}^{2}),~{}~{}~{}\beta=(2\pi)^{2}(c-c^{\prime})(b-b^{\prime}). (3.24)

This amplitude should satisfy

G(b,c|b,c)iδ(bb)δ(cc).{\cal H}G(b,c|b^{\prime},c^{\prime})\propto-i\delta(b-b^{\prime})\delta(c-c^{\prime}). (3.25)

The integral over NN in (3.23) can be expressed in terms of Bessel functions HL . With Nariai initial conditions (3.17) we have

α=13(Hb1)(Hb+2),β=(2πH2)2(H3c1)(Hb1)\alpha=-\frac{1}{3}(Hb-1)(Hb+2),~{}~{}~{}\beta=\left(\frac{2\pi}{H^{2}}\right)^{2}(H^{3}c-1)(Hb-1) (3.26)

and

G=iπH0(2)(αβ)(H3c>1)G=-i\pi H_{0}^{(2)}\left(\sqrt{-\alpha\beta}\right)~{}~{}~{}~{}~{}(H^{3}c>1) (3.27)
G=2K0(αβ)(H3c<1)G=2K_{0}\left(\sqrt{\alpha\beta}\right)~{}~{}~{}~{}~{}(H^{3}c<1) (3.28)

The amplitude GG in Eqs.(3.27),(3.28) is similar to the IKTV wave function (2.29),(2.30). The difference is in the prefactor and in the index of the Bessel functions. The two objects are closely related, as we will now show.

The Bessel functions appearing in Eqs.(3.27),(3.28) can be expressed as Z0(X)Z_{0}(X), where

X=H3c1f(b)=4πϕH2a~21X=\sqrt{H^{3}c-1}f(b)=4\pi{\phi}\sqrt{H^{2}{\tilde{a}}^{2}-1} (3.29)

with

f(b)=2πH2(Hb1)Hb+23=4πϕ,f(b)=\frac{2\pi}{H^{2}}(Hb-1)\sqrt{\frac{Hb+2}{3}}=4\pi{\phi}, (3.30)

where we have used the notation ZνZ_{\nu} for a Bessel function (of any kind) with index ν\nu and Eqs.(3.15) relating aa and bb to a~{\tilde{a}} and ϕ{\phi}. Differentiating twice with respect to cc, we obtain

2Z0c2=H6f2(b)4(H3c1)[Z1+1f(b)H3c1Z1]=H6f2(b)4(H3c1)Z2=(2πH3)2ϕ2a~21Z2.\frac{\partial^{2}Z_{0}}{\partial c^{2}}=\frac{H^{6}f^{2}(b)}{4(H^{3}c-1)}\left[-Z_{1}^{\prime}+\frac{1}{f(b)\sqrt{H^{3}c-1}}Z_{1}\right]=\frac{H^{6}f^{2}(b)}{4(H^{3}c-1)}Z_{2}=(2\pi H^{3})^{2}\frac{{\phi}^{2}}{{\tilde{a}}^{2}-1}Z_{2}. (3.31)

Here prime stands for a derivative with respect to the argument and we used an iteration formula for Bessel functions in the second step.

Now, the expression on the right-hand side of Eq.(3.31) has the same form as the wave function (2.29) of IKTV. We conclude that

ΨIKTV=C2c2G(b,c|H1,H3)\Psi_{IKTV}=C\frac{\partial^{2}}{\partial c^{2}}G(b,c|H^{-1},H^{-3}) (3.32)

with C=constC={\rm const}. Furthermore, it follows from Eq.(3.25) that444Note that /c\partial/\partial c commutes with {\cal H}.

ΨIKTViδ(bH1)δ′′(cH3).{\cal H}\Psi_{IKTV}\propto-i\delta(b-H^{-1})\delta^{\prime\prime}(c-H^{-3}). (3.33)

Hence ΨIKTV\Psi_{IKTV} is not a solution of the WDW equation. It has a distributional source at a=b=H1a=b=H^{-1}, which is more singular than that of a Green’s function.

4 Hartle-Hawking wave function

In the original Hartle & Hawling paper HH the HH wave function is defined as

Ψ(gb)=gb𝒟geSE(g),\Psi(g_{b})=\int^{g_{b}}{\cal D}ge^{-S_{E}(g)}, (4.1)

where the integration is over ”regular” 4D4D Euclidean geometries gg, having a single boundary {\cal B} with a 3-metric gbg_{b}. For simplicity we specialize to models without any matter fields. As it stands, this definition is rather problematic. The Euclidean action SES_{E} is unbounded from below, so the integral in (4.1) is divergent. This can often be dealt with by a suitable analytic continuation of the integration variables. Another problem is that the metrics contributing to the path integral are generally rather irregular, even non-differentiable. So the notion of integrating over regular geometries needs to be defined. The same problem arises in JT gravity.

IKTV attempted to get around this issue by focusing on the upper limit of integration in (4.1), with the hope that the regularity condition would somehow take care of itself. They allowed the boundary curve of the 2D2D geometry to fluctuate and required that the asymptotic form of the wave function agrees with the semiclassical pre-exponential factor resulting from these fluctuations. Another approach that they used was to calculate the path integral (4.1) for JT gravity with Λ<0\Lambda<0, where it is better defined, and then analytically continue to Λ>0\Lambda>0. IKTV find that the two approaches agree and yield the wave function (2.29). Our analysis shows, however, that this wave function is unsatisfactory, as it is not a solution of the WDW equation. Instead, it is related to the transition amplitude from a Nariai initial state with a=H1a=H^{-1} and ϕ=0\phi=0. The path integral (3.19) for this amplitude is over geometries with two boundaries – one with the specified values of aa and ϕ\phi and the other with the ‘Nariai’ values. This does not square well with the intuitive idea of quantum creation of the universe from nothing.

We therefore need to revisit the question of how the path integral over regular Euclidean geometries has to be defined. In the context of minisuperspace KS model, this issue has been discussed in detail by HL HL , whose approach we largely follow.

4.1 Boundary conditions

To discuss the boundary conditions for the no-boundary path integral, it is more convenient to represent the Euclideanized KS metric as

dsE2=N2dt2+a2(t)dx2+b2(t)dΩ2.ds_{E}^{2}=N^{2}dt^{2}+a^{2}(t)dx^{2}+b^{2}(t)d\Omega^{2}. (4.2)

The Euclidean action is then HL

SE=π01𝑑t[1N(ab˙2+2ba˙b˙)+Na(H2b21)]+Sb,S_{E}=\pi\int_{0}^{1}dt\left[-\frac{1}{N}\left(a{\dot{b}}^{2}+2b{\dot{a}}{\dot{b}}\right)+Na(H^{2}b^{2}-1)\right]+S_{b}, (4.3)

where SbS_{b} is the boundary term555 The Gibbons-Hawking boundary terms in the gravitational action cancel out after integration by parts if the geometry has two boundaries – at t=0t=0 and t=1t=1. But for a compact geometry with a single boundary at t=1t=1, the boundary term at t=0t=0 has to be kept HL .

Sb=[πNddt(ab2)]t=0.S_{b}=-\left[\frac{\pi}{N}\frac{d}{dt}(ab^{2})\right]_{t=0}. (4.4)

The time variable tt is defined so that 0<t<10<t<1 with t=1t=1 corresponding to the boundary {\cal B} and t=0t=0 corresponding to the ”bottom” 0{\cal B}_{0} of the 4-geometry gg.

The boundary conditions at t=1t=1 fix the values of {a,b}\{a,b\}, while the boundary conditions at t=0t=0 should be chosen so that the geometry closes smoothly at 0{\cal B}_{0}. HL show that for a classical 4-geometry(4.2) there are two choices:

a(0)=0,1Na˙(0)=±1,1Nb˙(0)=0a(0)=0,~{}~{}\frac{1}{N}{\dot{a}}(0)=\pm 1,~{}~{}\frac{1}{N}{\dot{b}}(0)=0 (4.5)

and

b(0)=0,1Nb˙(0)=±1,1Na˙(0)=0,b(0)=0,~{}~{}\frac{1}{N}{\dot{b}}(0)=\pm 1,~{}~{}\frac{1}{N}{\dot{a}}(0)=0, (4.6)

where overdots now stand for derivatives with respect to tt.

The time derivatives a˙{\dot{a}} and b˙{\dot{b}} are related to the (Euclidean) momenta conjugate to aa and bb:

pa=2πNbb˙,pb=2πN(ab˙+ba˙),p_{a}=-\frac{2\pi}{N}b{\dot{b}},~{}~{}p_{b}=-\frac{2\pi}{N}(a{\dot{b}}+b{\dot{a}}), (4.7)

so these boundary conditions correspond to fixing {a,pa,pb}\{a,p_{a},p_{b}\} or {b,pa,pb}\{b,p_{a},p_{b}\} at 0{\cal B}_{0}. It is however inconsistent in quantum theory to fix a coordinate and its conjugate momentum. Hence the best one can do is to impose two out of the three conditions, for example

a(0)=0,pb(0)=2πb(0)a(0)=0,~{}~{}p_{b}(0)=\mp 2\pi b(0) (4.8)

or

b(0)=0,pa(0)=0.b(0)=0,~{}~{}p_{a}(0)=0. (4.9)

HL note that if classical field equations hold, then with either of these choices all three conditions in (4.5) or (4.6) are satisfied. One can therefore expect that in the semiclassical regime the path integral will be dominated by regular geometries. Since we are interested in dimensional reduction of KS model with the S2S^{2} part integrated out, we will focus on the boundary conditions (4.8), which correspond to fixing aa and a˙{\dot{a}} on 0{\cal B}_{0}.

HL suggest that a better choice of variables, suitable for the boundary conditions (4.8), would be

A=2πb2,B=2πabA=2\pi b^{2},~{}~{}B=2\pi ab (4.10)

with the conjugate momenta

PA=a˙2N,PB=b˙N.P_{A}=-\frac{\dot{a}}{2N},~{}~{}P_{B}=-\frac{\dot{b}}{N}. (4.11)

The boundary conditions (4.8) then take the form

B=0,PA=12,B^{\prime}=0,~{}~{}{P_{A}}^{\prime}=\mp\frac{1}{2}, (4.12)

where primes indicate the values at t=0t=0.

The HH wave function can now be expressed as HL

ΨNB(A,B)=G(A,B|PA,B)=𝑑N𝒟Qα𝒟Pαexp(SE),\Psi_{NB}(A,B)=G(A,B|{P_{A}}^{\prime},B^{\prime})=\int dN\int{\cal D}Q^{\alpha}{\cal D}P_{\alpha}\exp(-S_{E}), (4.13)

where the path integral is taken over histories with fixed {A,B}\{A,B\} and {PA,B}\{{P_{A}}^{\prime},B^{\prime}\}. Unlike the case of fixed initial values aa^{\prime} and bb^{\prime}, this path integral cannot be evaluated exactly. We therefore use the WKB method to express ΨHH\Psi_{HH} approximately as

ΨHH(Qα)=Cμ(Qα,Zβ,N)exp[SE(Qα;N|Zβ)]𝑑N,\Psi_{HH}(Q^{\alpha})=\int_{C}\mu\left(Q^{\alpha},Z^{\beta},N\right)\exp\left[-S_{E}(Q^{\alpha};N|Z^{\beta})\right]dN, (4.14)

where Qα={A,B},Zβ={PA,B}Q^{\alpha}=\{A,B\},~{}Z^{\beta}=\{{P_{A}}^{\prime},B^{\prime}\}, μ\mu is the semiclassical prefactor of the propagator and

SE=π[H23N(b2+bb+b2)N1N(bb)(a2bB2b)+2b2PA+2BPB]S_{E}=\pi\left[\frac{H^{2}}{3}N(b^{2}+bb^{\prime}+{b^{\prime}}^{2})-N-\frac{1}{N}(b-b^{\prime})\left(a^{2}b-\frac{{B^{\prime}}^{2}}{b^{\prime}}\right)+2b^{\prime 2}{P_{A}}^{\prime}+2B^{\prime}{P_{B}}^{\prime}\right] (4.15)

is the Euclidean action evaluated on a history satisfying the boundary conditions and the second order equations (3.3),(3.4) for aa and bb, but not the constraint equation (3.5). Note that the last term in (4.15) can be neglected since it does not contribute to the path integral and the semiclassical prefactor. The integration contour CC in (4.14) is generally complex; we shall discuss the choice of this contour in Sec.4.4. The calculation of the prefactor is discussed in Sec. 4.3.

HL discussed the calculation of the HH wave functions for KS model only for the case of a vanishing cosmological constant, H=0H=0. Eqs.(4.14),(4.15) apply for arbitrary HH, but from this point on we cannot directly use the results of HL and will have to extend their analysis to H>0H>0.

The initial value bb^{\prime} in Eq.(4.15) has to be expressed in terms of the boundary values A,BA,B (or a,ba,b), B,PAB^{\prime},{P_{A}}^{\prime}, and the lapse NN. This can be done using the solutions a¯(τ),b¯(τ){\bar{a}}(\tau),{\bar{b}}(\tau) of the second order field equations (3.3),(3.4) (but not of the constraint equation). HL give these solutions in terms of the time variable τ{\tau}, which is related to tt as dτ=a(t)dtd{\tau}=a(t)dt:

b¯(τ)=(bb)τ+b{\bar{b}}({\tau})=(b-b^{\prime}){\tau}+b^{\prime} (4.16)
a¯2(τ)b¯(τ)=H2N23(bb)τ3H2N2bτ2+[a2ba2b+H2N23(b+2b)]τ+a2b.{\bar{a}}^{2}({\tau}){\bar{b}}({\tau})=-\frac{H^{2}N^{2}}{3}(b-b^{\prime}){\tau}^{3}-H^{2}N^{2}b^{\prime}{\tau}^{2}+\left[a^{2}b-{a^{\prime}}^{2}b^{\prime}+\frac{H^{2}N^{2}}{3}(b+2b^{\prime})\right]{\tau}+{a^{\prime}}^{2}b^{\prime}. (4.17)

Expressed in terms of τ\tau, the boundary condition (1/N)(da/dt)=±1(1/N)(da/dt)=\pm 1 takes the form

a¯Nda¯dτ(τ=0)=2PA=±1.\frac{\bar{a}}{N}\frac{d{\bar{a}}}{d{\tau}}({\tau}=0)=\mp 2{P_{A}}^{\prime}=\pm 1. (4.18)

To implement this boundary condition, we differentiate Eq.(4.17) with respect to τ{\tau} and take the limit τ0{\tau}\to 0. This gives (after dividing by b2{b^{\prime}}^{2})

4NPAbb=a2+B2b2H2N23b(b+2b).\frac{4N{P_{A}}^{\prime}}{b}{b^{\prime}}=-a^{2}+\frac{{B^{\prime}}^{2}}{{b^{\prime}}^{2}}-\frac{H^{2}N^{2}}{3b}(b+2b^{\prime}). (4.19)

where we have used a=B/ba^{\prime}=B^{\prime}/b^{\prime}.

We now have to solve Eq.(4.19) for bb^{\prime}, substitute the result into the action (4.15), and use it to evaluate SES_{E} and the pre-factor in (4.14) with the values of PAP_{A}^{\prime} and BB^{\prime} specified by the boundary conditions (4.12). This calculation is significantly simplified if we note that the solution of Eq.(4.19) minimizes the action (with our boundary conditions), and thus SE/b=0\partial S_{E}/\partial b^{\prime}=0. It follows that when calculating the derivative of SES_{E} with respect to BB^{\prime} in the determinantal prefactor in (4.14), we only need to account for an explicit dependence of SES_{E} on this variable and can disregard the dependence through bb^{\prime}. This means that we can substitute the boundary value B=0B^{\prime}=0 directly in Eq.(4.19). Then, instead of a cubic equation we get a linear equation for bb^{\prime}, with the solution

b=b2Na2+H2N2/32PA+H2N/3.b^{\prime}=-\frac{b}{2N}\frac{a^{2}+H^{2}N^{2}/3}{{2P_{A}}^{\prime}+H^{2}N/3}. (4.20)

Substituting this in the action (4.15) we find

SE=πN3(H2b23)πNa2b2πb24N2(a2+H2N23)22PA+H2N3πB2N(a2+H2N2+4NPA)a2+H2N2/3.S_{E}=\frac{\pi N}{3}(H^{2}b^{2}-3)-\frac{\pi}{N}a^{2}b^{2}-\frac{\pi b^{2}}{4N^{2}}\frac{\left(a^{2}+\frac{H^{2}N^{2}}{3}\right)^{2}}{2P_{A}^{\prime}+\frac{H^{2}N}{3}}-\frac{\pi{B^{\prime}}^{2}}{N}\frac{\left(a^{2}+H^{2}N^{2}+4NP_{A}^{\prime}\right)}{a^{2}+H^{2}N^{2}/3}. (4.21)

where we have not substituted the boundary values PA{P_{A}}^{\prime}, BB^{\prime} yet, in order to calculate the prefactor.

4.2 Saddle points

Without making any approximations for the action, the saddle points cannot be found in closed form. Since we will be mostly interested in the regime where bH1b\approx H^{-1}, we will first find the saddles for b=H1b=H^{-1} and then treat deviations from these points as small perturbations.

We also have to decide on the choice of sign in the boundary condition (4.12) for PAP_{A}^{\prime}. Here we will follow HL and pick PA=1/2P_{A}^{\prime}=-1/2. Their justification is that for this choice the final boundary \mathcal{B} is to the ”future” of the initial boundary 0\mathcal{B}_{0}. In fact, there is a stronger argument: it can be shown that choosing the opposite sign in (4.12) does not yield convergent contours for the HH wave function. Furthermore, the characteristic exponential factor exp(π/H2)\exp(\pi/H^{2}) can only be retrieved with the choice PA=1/2P_{A}^{\prime}=-1/2.

Thus, setting b=H1b=H^{-1}, B=0B^{\prime}=0 and PA=1/2P_{A}^{\prime}=-1/2 we have

SE0=2πN3πa2H2Nπ(3a2+H2N2)212H2N2(H2N3),S_{E0}=-\frac{2\pi N}{3}-\frac{\pi a^{2}}{H^{2}N}-\frac{\pi\left(3a^{2}+H^{2}N^{2}\right)^{2}}{12H^{2}N^{2}\left(H^{2}N-3\right)}, (4.22)

where the extra subscript ”0” indicates that the action is evaluated at Hb=1Hb=1. Note the singularities at N=0N=0 and N=3/H2N=3/H^{2}.

For the following analysis it will be convenient to introduce the rescaled variables

u=H2a2,N~=H2N,SE~=H2SEπu=H^{2}a^{2}\ \ ,\ \ \tilde{N}=H^{2}N\ \ ,\ \ \tilde{S_{E}}=\frac{H^{2}S_{E}}{\pi} (4.23)

The rescaled action (4.22) is given by

S~E0=2N~3uN~(3u+N~2)212N~2(N~3){\tilde{S}}_{E0}=-\frac{2\tilde{N}}{3}-\frac{u}{\tilde{N}}-\frac{\left(3u+\tilde{N}^{2}\right)^{2}}{12\tilde{N}^{2}\left(\tilde{N}-3\right)} (4.24)

In order to evaluate the integral (4.14) using the method of steepest descent, we first determine the extrema of the action S~E0\tilde{S}_{E0}. These are given by

SE0~N~=0.\frac{\partial\tilde{S_{E0}}}{\partial\tilde{N}}=0. (4.25)

The resulting equation is quintic in N~\tilde{N}:

(N~22N~+u)(N~34N~23N~u+6u)=0\left(\tilde{N}^{2}-2\tilde{N}+u\right)\left(\tilde{N}^{3}-4\tilde{N}^{2}-3\tilde{N}u+6u\right)=0 (4.26)

Its solutions are

N~1,2=1±1u\tilde{N}_{1,2}=1\pm\sqrt{1-u} (4.27)
N~3=43+16+9u3A+A3\tilde{N}_{3}=\frac{4}{3}+\frac{16+9u}{3A}+\frac{A}{3} (4.28)
N~4,5=43(1i3)(16+9u)6A(1±i3)A6\tilde{N}_{4,5}=\frac{4}{3}-\frac{\left(1\mp i\sqrt{3}\right)(16+9u)}{6A}-\frac{\left(1\pm i\sqrt{3}\right)A}{6} (4.29)

where the quantity AA is given by

A=(6427u+9128u39u29u3)1/3A=(64-27u+9\sqrt{-128u-39u^{2}-9u^{3}})^{1/3} (4.30)

The solutions 3,4,53,4,5 are expressed here in a rather complicated form. Taking a closer look at the quantity AA we can express it in a more convenient way with the Euler representation of complex numbers. After some straightforward calculations we arrive at

A=16+9ueiθ3,A=\sqrt{16+9u}\ e^{\textstyle i\frac{\theta}{3}}, (4.31)

where θ\theta is given by

θ=arctan[99u3+39u2+128u6427u]\theta=\arctan\left[\frac{9\sqrt{9u^{3}+39u^{2}+128u}}{64-27u}\right] (4.32)

The saddles N~3,4,5\tilde{N}_{3,4,5} are then simplified to:

N~3=43+216+9u3cosθ3\tilde{N}_{3}=\frac{4}{3}+\frac{2\sqrt{16+9u}}{3}\cos{\frac{\theta}{3}} (4.33)
N~4,5=43216+9u3cos[θ±π3]\tilde{N}_{4,5}=\frac{4}{3}-\frac{2\sqrt{16+9u}}{3}\cos\left[{\frac{\theta\pm\pi}{3}}\right] (4.34)

It is clear that the saddles 3,4,53,4,5 are always real. N~1{\tilde{N}}_{1} and N~2{\tilde{N}}_{2} are also real for u1u\leq 1, while for u>1u>1 they form a complex conjugate pair.

4.3 Prefactor

The semiclassical prefactor for the propagator is given by Schulman

μ=f1/4Df1/4\mu=f^{-1/4}\sqrt{D}f^{\prime-1/4} (4.35)

where ff^{\prime} and ff are the determinants of the minisuperspace metric fμνf_{\mu\nu} evaluated at t=0t=0 and t=1t=1 respectively and DD is the Van Vleck-Morette determinant VV Morette . In the representation A=2πb2A=2\pi b^{2} and B=2πabB=2\pi ab the Hamiltonian for the KS model takes the form HL

=PB24APAPBB+1H22πA=12fμνPμPν+V\mathcal{H}=-P_{B}^{2}-\frac{4AP_{A}P_{B}}{B}+1-\frac{H^{2}}{2\pi}A=\frac{1}{2}f^{\mu\nu}P_{\mu}P_{\nu}+V (4.36)

where V=1AH2/(2π)V=1-AH^{2}/(2\pi)\ . Thus the minisuperspace metric and its determinant are

fμν=(B28A2B4AB4A0),fB2A2f_{\mu\nu}=\begin{pmatrix}\frac{B^{2}}{8A^{2}}&-\frac{B}{4A}\\ -\frac{B}{4A}&0\end{pmatrix},~{}~{}~{}~{}f\propto\frac{B^{2}}{A^{2}} (4.37)

The action (4.21) expressed in variables {A,B}\{A,B\} is given by

SE=πN3(AH22π3)B24NπA8N2(B22πA+H2N23)22PA+H2N3πB2N(B22πA+H2N2+4NPAB22πA+H2N23)S_{E}=\frac{\pi N}{3}\left(\frac{AH^{2}}{2\pi}-3\right)-\frac{B^{2}}{4N\pi}-\frac{A}{8N^{2}}\frac{\left(\frac{B^{2}}{2\pi A}+\frac{H^{2}N^{2}}{3}\right)^{2}}{2P^{\prime}_{A}+\frac{H^{2}N}{3}}-\frac{{\pi B^{\prime}}^{2}}{N}\left(\frac{\frac{B^{2}}{2\pi A}+H^{2}N^{2}+4NP^{\prime}_{A}}{\frac{B^{2}}{2\pi A}+\frac{H^{2}N^{2}}{3}}\right) (4.38)

and the Van Vleck-Morette determinant DD can be calculated as

Ddet[2SEQαZβ]=3BBAN2(H2N+6PA),D\equiv\det\left[-\frac{\partial^{2}S_{E}}{\partial Q^{\alpha}\partial Z^{\beta}}\right]=\frac{3BB^{\prime}}{AN^{2}\left(H^{2}N+6P^{\prime}_{A}\right)}, (4.39)

where, as before, Qα={A,B},Zβ={PA,B}Q^{\alpha}=\{A,B\},~{}Z^{\beta}=\{{P_{A}}^{\prime},B^{\prime}\}. Inserting the above relations in Eq.(4.35) for the prefactor and switching back to variables {a,b}\{a,b\}, we obtain666We note that there is an error in the expression (6.5) for the prefactor in Ref.HL . We are grateful to Jorma Louko for a discussion of this point.

μ(a,b,N)bNH2N3.\mu\left(a,b,N\right)\propto\frac{b^{\prime}}{N\sqrt{H^{2}N-3}}. (4.40)

where bb^{\prime} is a function of a,b,Na,b,N and is given by Eq.(4.20) and we have now inserted the boundary value of PA=1/2P^{\prime}_{A}=-1/2.

The prefactor in Eq.(4.40) introduces a branch cut, which we can choose to lie at N>3/H2N>3/H^{2} along the real axis. From the analysis that follows, we will see that the choice of a suitable contour will not be affected significantly by this branch cut.

4.4 Integration contours

One of the key issues that remains to be addressed is the choice of the integration contour over NN in Eq.(4.14). No general prescription for this choice has yet been given. Integration over real or pure imaginary values of NN yields divergent integrals, so one has to look for a non-trivial contour in the complex plane that would make the integral convergent. The consensus view appears to be that the contour CC should satisfy the following three criteria (see, e.g., Halliwell:1989dy ). (1) CC should not have ends: it must be either infinite or closed. This guarantees that ΨHH\Psi_{HH} is a solution of the WDW equation (rather than a Green’s function). (2) The HH wave function should be real. This can be achieved by choosing a contour which is symmetric with respect to the real NN axis. This requirement can be thought of as an expression of the CPTCPT invariance of the HH state HHH . (3) The wave function should predict a classical spacetime when the universe is large. This means that in the appropriate limit ΨHH\Psi_{HH} should be a superposition of rapidly oscillating terms of the form eiSe^{iS}, where SS is the classical action. We shall adopt these criteria as the defining properties of ΨHH\Psi_{HH}.

Refer to caption
Figure 1: Examples of convergent infinite contours in the complex NN plane. The singularities N={0,3}N=\{0,3\} are shown in circles.

Let us first consider infinite contours. From Eq.(4.22) we find that SE3N/8S_{E}\sim-3N/8 for |N||N|\to\infty. It follows that the integral over NN can be convergent only if the asymptotic directions of the contour are at |argN|>π/2|{\rm arg}~{}N|>\pi/2. Some inequivalent choices are illustrated in Fig.1. These contours are symmetric with respect to the real axis, so they define a real wave function. We will first consider the contour BB which crosses the real axis at 0<N<30<N<3 and will comment on the other choices at the end of this section. Note that the contour BB may almost coincide with the imaginary axis. It could cross the real axis at N=+ϵN=+\epsilon and asymptote to argN=±(π/2+ϵ){\rm arg}~{}N=\pm(\pi/2+\epsilon^{\prime}), where ϵ,ϵ+0\epsilon,\epsilon^{\prime}\to+0.

The integration contour BB can be turned into a closed contour by adding to it an infinite arc |N|=const|N|={\rm const}\to\infty. The integral over the arc vanishes in the limit, so the original integral remains unchanged. The resulting contour can be distorted into a finite closed contour which encircles the singularity at N=0N=0 but does not encircle the singularity at N=3N=3. Thus the infinite contours of type BB are equivalent to this kind of closed contours.

Refer to caption
Figure 2: The steepest descent contours for u>1u>1 and Hb=1Hb=1. The arrowheads point to the direction where Re(S~E)Re(-\tilde{S}_{E}) decreases. The saddles N~i\tilde{N}_{i} are marked with solid dots and the singularities with circles. Note the branch cut at N~(3+,+)\tilde{N}\in(3^{+},+\infty). The HH contour corresponds to the solid curve encircling the singularity N=0N=0; it is dominated by saddles N1N_{1}, N2N_{2}.

Following the Picard-Lefschetz prescription,777 For a simple review of Picard-Lefschetz theory see, e.g., Ref. Feldbrugge:2017kzv . the closed contour can now be distorted so that it passes through saddle points following the steepest descent and ascent lines, making the integral absolutely convergent. Let us first consider the case of a>H1a>H^{-1}. The saddle points for this case are shown in Fig.2. The steepest descent and ascent lines are defined by ImSE=const{\rm Im}~{}S_{E}={\rm const}. The lines passing through the saddle points are also shown in the figure, with arrows indicating the direction in which ReSE-{\rm Re}~{}S_{E} is decreasing. The contour encircling the singularity at N=0N=0 can be distorted into the contour passing through the saddle points N1N_{1}, N4N_{4}, N2N_{2} and N5N_{5}. This contour is dominated by the saddles at N1N_{1} and N2N_{2}.

We now consider the case of a<H1a<H^{-1}, when all saddle points lie on the real axis. The steepest descent and ascent lines in this case are illustrated in Fig.3. The contour encircling N=0N=0 can now be deformed into the contour passing through N4N_{4} and N5N_{5}. It is dominated by the saddle at N4N_{4}.

Refer to caption
Figure 3: The steepest descent contours for u<1u<1 and Hb=1Hb=1. In this case, all the saddles are real. The HH contour corresponds to the solid curve encircling the singularity N=0N=0 and dominated by saddle N4N_{4}.

We finally comment on other possible choices of the integration contour. A contour of type AA in Fig.1 crosses the real axis at N<0N<0. After it is closed by adding an infinite arc, this contour does not encircle any singularities, so it can be continuously shrunk to a point. The wave function defined by this contour is therefore identically zero.

Another possibility is to choose the branch cut to lie at N~<3{\tilde{N}}<3 on the real axis and choose the contour that crosses the real axis at N~>3{\tilde{N}}>3 (see contour CC in Fig.1). For a>H1a>H^{-1} such a contour can be deformed into a contour that runs along the steepest descent/ascent lines from N3N_{3} to N1N_{1}, then takes a turn and goes to N5N_{5}, and from there runs above the branch cut along the real axis to NN\to-\infty. This has to be supplemented by another half of the contour that runs symmetrically from N3N_{3} through N2N_{2} and N5N_{5} to -\infty in the lower half-plane. The resulting wave function is then non-oscillating, with the main contribution given by the real saddle point N3N_{3}. This is in conflict with the defining property (3) of the HH wave function. We thus conclude that the only acceptable choice of integration contour is an infinite contour of type B or equivalently a closed contour encircling the singularity at N=0N=0.

4.5 Perturbing the Saddle Points

To make a connection with JT gravity, we need to know the KS wave function for bb very close but not equal to H1H^{-1}. The saddle points and the steepest descent/ascent lines will then be slightly different from those we found in the subsections B and D. For a>H1a>H^{-1} we are interested in the complex saddle points N1,2N_{1,2} which dominate the integral. Let us define the shift xx of the saddle point N~i\tilde{N}_{i} as

N~=N~i+x,\tilde{N}=\tilde{N}_{i}+x, (4.41)

where i=1,2i=1,2, N~i\tilde{N}_{i} are given by (4.27) and |x|<<1|x|<<1. We insert this into the action and expand to second order in xx. This gives :

S~E12i(1Hb)u12(1Hb)x+f(u)x2+O(x3)\tilde{S}_{E}\approx-1\mp 2i(1-Hb)\sqrt{u-1}-2(1-Hb)x+f(u)x^{2}+O(x^{3}) (4.42)

where the upper and lower signs are for N1N_{1} and N2N_{2} respectively and the function f(u)f(u) is given in the Appendix.

The action is extremized with

x=1Hbf(u).x=\frac{1-Hb}{f(u)}. (4.43)

Since xx depends linearly on (1Hb)(1-Hb) the contribution of the x-dependent terms to the action is 𝒪[(1Hb)2]{\cal O}[(1-Hb)^{2}]. Thus the action of the dominant saddle points is

SEπH22iπH2(1Hb)H2a21+𝒪[(1Hb)2].S_{E}\approx-\frac{\pi}{H^{2}}\mp\frac{2i\pi}{H^{2}}(1-Hb)\sqrt{H^{2}a^{2}-1}+{\cal O}[(1-Hb)^{2}]. (4.44)

The steepest descent/ascent lines can be calculated numerically for any values of aa and bb. For Ha>1Ha>1 the character of these lines changes when HbHb is moved away from 1, even for an arbitrarily small amount. For Hb>1Hb>1, the steepest descent contour passing through N1N_{1} follows nearly the same path as for Hb=1Hb=1, but short of reaching N5N_{5} it makes a turn and runs along the real axis towards N=0N=0. Then it runs back, repeats the same path symmetrically in the lower half-plane and arrives at N2N_{2}. From there it runs towards N4N_{4} following nearly the same path as for Hb=1Hb=1, but short of reaching N4N_{4} it makes a turn and runs towards the singularity at N=3N=3. Finally it runs back symmetrically and returns to N1N_{1}. This contour is illustrated in Fig.4 . The only change compared to the original Hb=1Hb=1 contour is that small segments near N4N_{4} and N5N_{5} are now replaced by sharp spikes running towards the singularities and back. The integrals over the upper and lower halves of these spikes nearly cancel one another, so their combined contribution to the wave function is very small.

Refer to caption
Figure 4: The perturbed steepest descent contours for u>1u>1 and Hb>1Hb>1. The HH contour does not pass through the saddles N4N_{4} and N5N_{5}.
Refer to caption
Figure 5: The perturbed steepest descent contours for u>1u>1 and Hb<1Hb<1. The HH contour does not pass through the saddles N4N_{4} and N5N_{5}.

.

For Hb<1Hb<1 the situation is very similar, except now instead of shooting to the right the spikes shoot to the left. The spike originating near N5N_{5} runs along the negative real axis to -\infty and back, and the spike originating near N4N_{4} runs to N=0N=0 and back (see Fig.5 ). As before, the contour is dominated by the saddles N1N_{1} and N2N_{2}, with the spikes making a very small contribution.

For Ha<1Ha<1, small deviations of HbHb from 1 do not change the character of the steepest descent lines. The contour is still dominated by the real saddle N4N_{4}, and the steepest descent line passing through this saddle also passes through N5N_{5}. At some critical value of Hb>1Hb>1 the saddles N1N_{1} and N4N_{4} merge, and at greater values they become a complex conjugate pair . A similar situation occurs for saddles N2N_{2} and N4N_{4} when Hb<1Hb<1 . Generally, the saddle N4N_{4} remains real in a range of (Hb1)(Hb-1) that depends on uu. It can be shown that as uu increases from 0 to 11, the range of (Hb1)(Hb-1) for which N4N_{4} remain real shrinks from 1\sim 1 until it reaches zero at u=1u=1, where the saddles N1,2,4N_{1,2,4} merge. For example, when u1u\ll 1 the contour behavior does not change qualitatively for 0<Hb<30<Hb<\sqrt{3} and for u=0.9u=0.9 the range is |1Hb|0.008|1-Hb|\sim 0.008. It would be interesting to map the behavior of the contours and saddles for the full range of aa and bb, but we will not attempt this here.

4.6 The Hartle-Hawking wave function

We are now ready to calculate the semiclassical Hartle-Hawking wave function.

4.6.1 Ha>1Ha>1

For Ha>1Ha>1 we need to expand the action (4.22) up to quadratic order in (NNi)(N-N_{i}) at saddle points NiN_{i} (i=1,2i=1,2) and then do the Gaussian integrals. Up to an overall numerical factor, the corresponding contributions to the wave function are

b(Ni)Ni3H2Ni1SNN(Ni)eSE(Ni).\frac{b^{\prime}(N_{i})}{N_{i}\sqrt{3-H^{2}N_{i}}}\sqrt{\frac{1}{S_{NN}(N_{i})}}e^{-S_{E}(N_{i})}. (4.45)

Here the factor b(Ni)/(Ni3H2Ni)b^{\prime}(N_{i})/\left(N_{i}\sqrt{3-H^{2}N_{i}}\right) comes from Eq.(4.40), SE(Ni)S_{E}(N_{i}) from Eq.(4.44) and

SNN=2SE(N)2.S_{NN}=\frac{\partial^{2}S_{E}}{(\partial N)^{2}}. (4.46)

At N=NiN=N_{i} we have

SNN(N1,2)=6π(H2a21)a4H2(H2a2+3)[2±2iH2a21±iH2a2H2a21]S_{NN}(N_{1,2})=\frac{6\pi\left(H^{2}a^{2}-1\right)}{a^{4}H^{2}\left(H^{2}a^{2}+3\right)}\left[-2\pm 2i\sqrt{H^{2}a^{2}-1}\pm iH^{2}a^{2}\sqrt{H^{2}a^{2}-1}\right] (4.47)

and

b(Ni)[Ni2(3H2Ni)SNN(Ni)]1/2=H6π(H2a21),b^{\prime}(N_{i})\left[N_{i}^{2}\left(3-H^{2}N_{i}\right)S_{NN}(N_{i})\right]^{-1/2}=\frac{H}{\sqrt{6\pi\left(H^{2}a^{2}-1\right)}}, (4.48)

Combining the contributions of the two saddle points, we obtain an approximate semiclassical HH wave function. Up to an overall constant factor it is given by

ΨHH(Ha>1)𝒜exp(πH2)cos(2πH2(1Hb)H2a21),\Psi_{HH}(Ha>1)\propto{\cal A}\exp\left(\frac{\pi}{H^{2}}\right)\cos\left(\frac{2\pi}{H^{2}}(1-Hb)\sqrt{H^{2}a^{2}-1}\right), (4.49)

where

𝒜=1H2a21.{\cal A}=\frac{1}{\sqrt{H^{2}a^{2}-1}}. (4.50)

Note that we neglected corrections 𝒪(1Hb){\cal O}(1-Hb) in the prefactor, but kept them in the exponent, which includes a large factor H2H^{-2}. The WKB approximation is essentially an expansion in powers of HH. It is easily verified that the wave function (4.49) satisfies the WDW equation to the leading order in HH and (Hb1)(Hb-1).

As one might expect, ΨHH\Psi_{HH} exhibits the characteristic WKB divergence at the turning point a=1/Ha=1/H. This divergence is much milder than that in the IKTV solution. The WKB approximation breaks down near the turning point, and we expect the exact wave function to remain finite there.

4.6.2 Ha<1Ha<1

For Ha<1Ha<1 the integral over NN is dominated by the saddle point N4N_{4}. This case is difficult to handle analytically, so we will only consider the limiting case Ha1Ha\ll 1. In this regime, we are able to go beyond the approximation Hb1Hb\approx 1, as long as the qualitative behavior of the contours and the respective saddles does not change (see section 4.5).

Let us first note that for Ha1Ha\ll 1 and Hb1Hb\approx 1, Eq.(4.34) gives

H2N432Ha.H^{2}N_{4}\approx\sqrt{\frac{3}{2}}Ha. (4.51)

Let us further assume (to be verified shortly) that H2N4=𝒪(Ha)H^{2}{N}_{4}={\cal O}(Ha) in a wide range of Hb1Hb\lesssim 1. Then the action can be approximated by888 This can be seen from Eq.(A.1) for SES_{E} in the Appendix by noticing that the first two terms in both of the big parentheses in that equation are 𝒪(Ha){\cal O}(Ha) while the last terms are 𝒪(H2a2){\cal O}(H^{2}a^{2}).

SEπN3(H2b23)πNa2b2S_{E}\approx\frac{\pi N}{3}(H^{2}b^{2}-3)-\frac{\pi}{N}a^{2}b^{2} (4.52)

and the corresponding saddles are

N4,5=±ab1H2b2/3.N_{4,5}=\pm\frac{ab}{\sqrt{1-H^{2}b^{2}/3}}. (4.53)

We note that Eq.(4.53) agrees with our assumption that H2N4=𝒪(Ha)H^{2}{N}_{4}={\cal O}(Ha) for Hb1Hb\lesssim 1, verifying that this assumption is consistent. Substituting (4.53) into the action we obtain

SE2πab33H2b2.S_{E}\approx-\frac{2\pi ab}{\sqrt{3}}\sqrt{3-H^{2}b^{2}}. (4.54)

The WKB prefactor in the regime Ha1Ha\ll 1 can be found along the same lines as for Ha>1Ha>1. To lowest order in HaHa, it is proportional to a\sqrt{a}. Thus, the wave function is given by

ΨHHaexp(2πab33H2b2).\Psi_{HH}\propto\sqrt{a}\exp\left(\frac{2\pi ab}{\sqrt{3}}\sqrt{3-H^{2}b^{2}}\right). (4.55)

This approximation applies for HHa1H\ll Ha\ll 1, which is a wide range in the sub-Planckian regime H1H\ll 1. Also, we have to constrain the values of bb to Hb<3Hb<\sqrt{3} in order for the saddles to remain real and the wavefunction to be non-oscillatory. (Note that our approximation breaks down at Hb3Hb\approx\sqrt{3}, where the lapse parameter NN in Eq.(4.53) becomes large.)

Keeping the scale factor aa fixed, the solution has a maximum at Hb=3/2Hb=\sqrt{3/2}. This peak is not in the range Hb1Hb\approx 1 which is of most interest for dimensional reduction to JT. Note, however, that the maximum of the wave function (4.55) at a fixed a~=aHb{\tilde{a}}=a\sqrt{Hb} is at Hb=1Hb=1. This is more relevant, since a~{\tilde{a}} plays the role of the scale factor after dimensional reduction. A numerical WKB solution for ΨHH\Psi_{HH} in the full range a<1/Ha<1/H with Hb=1Hb=1 is shown in Fig.6.

Finally, it can be verified that the wave function (4.55) satisfies the WDW equation to the leading order in HH and that it grows exponentially with aa, as expected.

Refer to caption
Figure 6: A graph of (logΨHH,a)(\log\Psi_{HH},a) for bH=1bH=1. A numerical WKB solution for the HH wavefunction is shown by the solid curve. It diverges abruptly at aH=1aH=1, due to the WKB prefactor. The blue dashed line corresponds to the aH<<1aH<<1 approximation. Note that both curves diverge to -\infty at a=0a=0 due to the pre-exponential factor a\sqrt{a} in (4.55).

5 Probability distribution

Probability distributions in minisuperspace quantum cosmology can be expressed in terms of the conserved current density

jα=iffαβ(ΨβΨΨβΨ),j^{\alpha}=i\sqrt{-f}f^{\alpha\beta}\left(\Psi^{*}\partial_{\beta}\Psi-\Psi\partial_{\beta}\Psi^{*}\right), (5.1)
αjα=0,\partial_{\alpha}j^{\alpha}=0, (5.2)

where fαβf^{\alpha\beta} is the minisuperspace metric and f=det(fαβ)f=\det(f_{\alpha\beta}). In a dd-dimensional minisuperspace one of the coordinates (or a combination of coordinates), call it TT, can be designated as a ”clock”. Then the probability distribution for the other coordinates on surfaces of constant TT is given by

dPjαdΣα,dP\propto j^{\alpha}d\Sigma_{\alpha}, (5.3)

where dΣαd\Sigma_{\alpha} is the (d1)(d-1)-dimensional surface element. If the clock variable TT exhibits semiclassical behavior and its classical evolution is monotonic (which are the properties one can reasonably require from a good clock), then it can be shown that the probability defined by Eq.(5.3) is positive definite Vilenkin:1988yd .

An obvious problem with the HH wave function is that it is real, so the current is identically zero. In the classically allowed range Ha>1Ha>1 this can be circumvented if we calculate the current using only the branch of the wave function describing expanding universes. Eq.(4.49) for ΨHH\Psi_{HH} would then be replaced by

Ψ+(Ha>1)1H2a21exp[2πiH2(Hb1)H2a21+F(a)(Hb1)2+𝒪((Hb1)3)],\Psi_{+}(Ha>1)\propto\frac{1}{\sqrt{H^{2}a^{2}-1}}\exp\left[\frac{2\pi i}{H^{2}}(Hb-1)\sqrt{H^{2}a^{2}-1}+F(a)(Hb-1)^{2}+{\cal O}\left((Hb-1)^{3}\right)\right], (5.4)

where we have dropped the constant factor exp[π/H2]\exp\left[\pi/H^{2}\right]. We have also included a quadratic in (Hb1)(Hb-1) correction in the exponential; we shall see that it plays an important role in the probability distribution. The coefficient function F(a)F(a) is given in the Appendix.

In the KS model, a natural choice for the clock variable is the scale factor aa (on the expanding branch of the wave function). Alternatively, we can use c=a2bc=a^{2}b, since we are working in the regime where bH1=constb\approx H^{-1}={\rm const}. We shall adopt the latter choice, which is more convenient. It is also more appropriate for the dimensional reduction, since ca~2c\propto{\tilde{a}}^{2}, where a~{\tilde{a}} is the scale factor of the JT model. Then the probability distribution for bb, Eq.(5.3), takes the form999Note that with bb and cc used as minisuperspace coordinates, the metric is fbc=constf^{bc}={\rm const}, fbb=fcc=0f^{bb}=f^{cc}=0, and its determinant is f=(fbc)2=constf=-(f^{bc})^{-2}={\rm const}

dPjcdb,dP\propto j^{c}db, (5.5)

where

jci(ΨbΨΨbΨ).j^{c}\propto-i\left(\Psi^{*}\partial_{b}\Psi-\Psi\partial_{b}\Psi^{*}\right). (5.6)

Substituting the wave function Ψ+\Psi_{+} in jcj^{c} and setting Hb=1Hb=1 in the pre-exponential factor, we obtain

dPdbH2a21exp[2(Hb1)2ReF(a)].dP\propto\frac{db}{\sqrt{H^{2}a^{2}-1}}\exp\left[2(Hb-1)^{2}{\rm Re}F(a)\right]. (5.7)

The real part of F(a)F(a) is calculated in the Appendix:

ReF(a)=2π3H2(H2a21).{\rm Re}F(a)=-\frac{2\pi}{3H^{2}(H^{2}a^{2}-1)}. (5.8)

Thus the probability distribution is

dPdbH2a21exp[4π(Hb1)23H2(H2a21)].dP\propto\frac{db}{\sqrt{H^{2}a^{2}-1}}\exp\left[-\frac{4\pi(Hb-1)^{2}}{3H^{2}(H^{2}a^{2}-1)}\right]. (5.9)

There are a few interesting things to note about this distribution. It is peaked at b=H1b=H^{-1} with variance δbHa\delta b\sim Ha. Our approximations are accurate for δbH1\delta b\ll H^{-1}, that is for

H1aH2.H^{-1}\lesssim a\ll H^{-2}. (5.10)

The distribution is obviously normalizable. Moreover, we note that

dPdb𝑑b=const,\int_{-\infty}^{\infty}\frac{dP}{db}db={\rm const}, (5.11)

independent of aa, so we can normalize the distribution to one. This is a direct consequence of conservation of jαj^{\alpha}.

6 Back to JT

Once we have found the HH wave function ΨHH(KS)(a,b)\Psi_{HH}^{(KS)}(a,b) for the KS model, we can define the HH wave function for JT gravity as the wave function obtained from ΨHH(KS)\Psi_{HH}^{(KS)} by dimensional reduction. This amounts to expressing the scale factors a,ba,b in terms of the JT variables a~,ϕ{\tilde{a}},\phi using the connection formulas (3.15) and adding an extra factor of a~{\tilde{a}} to account for the difference between Ψ\Psi and Ψ~\tilde{\Psi} in the JT model.

An issue that needs to be addressed is that of the boundary conditions (4.12) that we used in the path integral for ΨHH(KS)\Psi_{HH}^{(KS)}. These boundary conditions were imposed to ensure a smooth closure of the geometry at t=0t=0; they are equivalent to a(0)=0,a˙(0)/N=1a(0)=0,~{}{\dot{a}}(0)/N=1. However, after dimensional reduction the new scale factor is given by a~=aHb{\tilde{a}}=a\sqrt{Hb}, so the appropriate boundary conditions would now be

a~(0)=0,a~˙(0)/N=1.{\tilde{a}}(0)=0,~{}{\dot{\tilde{a}}}(0)/N=1. (6.1)

This is equivalent to (4.12) only if Hb(0)=1Hb(0)=1. Requiring in addition that the dilaton is smooth at t=0t=0, we should add the condition

ϕ˙(0)=0.{\dot{\phi}}(0)=0. (6.2)

As we discussed in Sec.4.1, a smooth closure of a classical S1×S2S^{1}\times S^{2} geometry requires three boundary conditions, while only two conditions can be consistently imposed in quantum theory. Classically, the three conditions are not independent and any two of them imply the third. In the present case the situation is similar: it follows from the boundary conditions (4.12) that Hb(0)=1Hb(0)=1 if the constraint equation (3.5) is satisfied. Thus the boundary conditions that we used are equivalent to smooth closure conditions for JT model at the classical level. Quantum mechanically, different choices of two conditions out of three may be inequivalent, yielding wave functions satisfying WDW equations with different factor orderings. One can expect, however, that in the semiclassical regime these wave functions will be close to one another, differing perhaps only in the pre-exponential factor.101010 The semiclassical wave function for an FRW universe with a uniform scalar field ϕ\phi was studied in Ref.Vilenkin:1987kf for different factor orderings in the WDW equation. A change of factor ordering had an effect on pre-exponential factor, but it did not affect the semiclassical probability distribution for ϕ\phi. One can expect a similar situation to occur in the JT model. Since no particular choice appears to be preferred, we shall proceed to use our ΨHH(KS)\Psi_{HH}^{(KS)} for dimensional reduction.

For Hb1Hb\approx 1 the relations (3.15) become

Hb1+2ϕH2,aa~(1ϕH2).Hb\approx 1+2\phi H^{2}\ \ ,\ \ a\approx\tilde{a}(1-\phi H^{2}). (6.3)

Then in the classically allowed range Ha>1Ha>1 the wave function (4.49) becomes

ΨHHa~H2a~21exp(πH2)cos(4πϕH2a~21),\Psi_{HH}\sim\frac{\tilde{a}}{\sqrt{H^{2}\tilde{a}^{2}-1}}\exp\left(\frac{\pi}{H^{2}}\right)\cos\left(4\pi\phi\sqrt{H^{2}\tilde{a}^{2}-1}\right), (6.4)

were we have neglected H2H^{2} corrections to the prefactor.

In the classically forbidden region we can use the small-HaHa solution (4.55) to obtain the JT wave function for Ha~1H{\tilde{a}}\ll 1. We find111111For this calculation, it is helpful to rearrange the relation of ϕ\phi and bb in the form:2Hb16ϕ2H4=3H2b2\sqrt{\frac{2}{Hb}}\sqrt{1-6\phi^{2}H^{4}}=\sqrt{3-H^{2}b^{2}}.

ΨHHa~3/2exp[4a~π6H16ϕ2H4],\Psi_{HH}\sim\tilde{a}^{3/2}\exp\left[\frac{4\tilde{a}\pi}{\sqrt{6}H}\sqrt{1-6\phi^{2}H^{4}}\right], (6.5)

where we have neglected corrections to the prefactor. This expression is valid for a~H1\tilde{a}H\ll 1 and ϕ2H4<1/6\phi^{2}H^{4}<1/6. The wave function peaks at ϕ=0\phi=0 at a fixed a~\tilde{a}. It can be shown that it satisfies the WDW equation of JT gravity (2.17) to the leading order.

Similarly to the KS model, the expanding branch of the wave function (6.4) can be used to find the probability distribution for the dilaton field ϕ\phi. The difference here is that now the conserved current is given by Eq.(5.1) with Ψ\Psi replaced by Ψ~=Ψ/a{\tilde{\Psi}}=\Psi/a. The reason is that the differential operator in Eq.(2.17) is not the covariant Laplacian, because of nonstandard factor ordering in Henneaux’s quantization of the JT model. As a result the probability distribution is obtained by simply using the connection formulas (3.15) in Eq.(5.9) without any extra factors of a~{\tilde{a}}:

dPdϕH2a~21exp[16πH2ϕ23(H2a~21)].dP\propto\frac{d\phi}{\sqrt{H^{2}{\tilde{a}}^{2}-1}}\exp\left[-\frac{16\pi H^{2}\phi^{2}}{3(H^{2}{\tilde{a}}^{2}-1)}\right]. (6.6)

We expect this expression to be accurate for ϕH1δbH2\phi\lesssim H^{-1}\delta b\ll H^{-2} and a~{\tilde{a}} satisfying the conditions (5.10).

We note that the classical solutions (2.10) of the JT model

a=H1cosh(Ht),ϕ=ϕ0sinh(Ht)a=H^{-1}\cosh(Ht),~{}~{}~{}~{}\phi=\phi_{0}\sinh(Ht) (6.7)

satisfy

ϕ2H2a21=ϕ02=const.\frac{\phi^{2}}{H^{2}a^{2}-1}=\phi_{0}^{2}={\rm const}. (6.8)

These solutions are parametrized by ϕ0\phi_{0} and Eq.(6.6) gives a probability distribution for this parameter:

dPdϕ0exp(16πH2ϕ023).dP\propto d\phi_{0}\exp\left(-\frac{16\pi H^{2}\phi_{0}^{2}}{3}\right). (6.9)

This distribution can be interpreted as describing an ensemble of (1+1)D(1+1)D universes that nucleate with a~H1{\tilde{a}}\approx H^{-1}, ϕ0\phi\approx 0 and ϕ˙Hϕ0\dot{\phi}\approx H\phi_{0} and then evolve according to Eqs.(6.7). Even though the approximations we used to derive the wave function break down at a~H2{\tilde{a}}\gtrsim H^{-2}, the classical solutions become increasingly accurate at large a~{\tilde{a}} and we expect the distribution (6.9) to remain accurate as well.

7 Discussion

Our main goal in this paper was to define and calculate the Hartle-Hawking wave function ΨHH\Psi_{HH} in a (1+1)(1+1)-dimensional minisuperspace JT model with a cosmological constant Λ=H2>0\Lambda=H^{2}>0. This model is closely related to that of a homogeneous 4D4D universe with the same cosmological constant and having spatial topology S1×S2S^{1}\times S^{2} (the KS model). Our approach was first to find ΨHH\Psi_{HH} for the KS model using its definition in terms of a Euclidean path integral and then to use the exact correspondence between the two models to define ΨHH\Psi_{HH} for JT gravity. In our analysis of KS quantum cosmology we followed the work of Halliwell and Louko HL . However, this work was mostly limited to the case of a vanishing Λ\Lambda, so to implement our program we had to tackle the nontrivial task of extending it to Λ>0\Lambda>0.

The wave function that we found is normalizable, so we could obtain a normalized probability distribution for the dilaton in the JT model. Note, on the other hand, that the leading-order semiclassical wave function found by Maldacena et al Maldacena is not normalizable, even after including a Schwarzian prefactor.

Our wave function is different from the exact WDW solution obtained earlier by Iliesiu et al Iliesiu . This difference is due to a different choice of the boundary conditions. The HH wave function was originally defined as a path integral over smooth Euclidean geometries with a single boundary. We adopted this definition here and imposed the condition of smooth closure at a=0a=0, where aa is the radius of S1S^{1}. On the other hand, Ref.Iliesiu imposed boundary conditions at large aa requiring that ΨHH\Psi_{HH} has the asymptotic form suggested by Schwarzian theory, which accounts for quantum fluctuations of the boundary curve. Both boundary conditions seem to be reasonable, but it appears that they are not compatible with one another.

The wave function obtained in Iliesiu using the Schwarzian boundary conditions has a strong singularity at a=H1a=H^{-1}.121212An alternative approach, suggested in Ref.Maldacena , was to derive ΨHH\Psi_{HH} by analytic continuation from JT gravity with a negative cosmological constant. A Euclidean dS metric (with H=1H=1) is ds2=(1r2)dθ2+(1r2)1dr2ds^{2}=(1-r^{2})d\theta^{2}+(1-r^{2})^{-1}dr^{2}. With r=coshρr=\cosh\rho, this becomes ds2=dρ2sinh2ρdθ2ds^{2}=-d\rho^{2}-\sinh^{2}\rho d\theta^{2}, which is minus Euclidean AdS metric. This method gives the same wave function as the Schwarzian boundary condition Iliesiu . We note that the origin ρ=0\rho=0 in AdS analytically continues to the horizon (r=1r=1) in dS. This may explain why the transition amplitude from ”nothing” (ρ=0\rho=0) to some ρ>0\rho>0 in AdS could be related to the transition amplitude from the horizon (a=H1a=H^{-1}) to some a>H1a>H^{-1} in dS. We found that this wave function is not a solution of the WDW equation. Instead, it satisfies an equation with a singular source at a=H1a=H^{-1}. Furthermore, we showed that this wave function is closely related to the transition amplitude in the KS model from a=b=H1a=b=H^{-1} to specified values of aa and bb at the boundary, where bb is the radius of S2S^{2}. Since a state with a=b=H1a=b=H^{-1} (which corresponds to a=H1,ϕ=0a=H^{-1},~{}~{}\phi=0 in the JT model) can hardly be interpreted as ”nothing”, we believe that the wave functions discussed in Iliesiu cannot be interpreted as the HH wave function. On the other hand, the wave function we found in the present paper satisfies the WDW equation and has only a mild singularity at a=H1a=H^{-1}, which one always expects in a WKB wave function at a classical turning point. It seems therefore that our choice of boundary conditions yields a more reasonable result for the HH wave function.

A possible reason why the Schwarzian boundary condition at large aa fails to yield a suitable candidate for the HH wave function is that it leaves the geometry at small aa completely unconstrained. It is assumed that the geometry closes off at a=0a=0 in a nonsingular way. However, this condition is not explicitly enforced, so one should not be surprised if geometries included in the path integral include conical singularities and even gaps.

It is perhaps not surprising that the wave function we found using the boundary condition of smooth closure does not exhibit Schwarzian asymptotic behavior. An obvious reason is that our analysis was restricted to minisuperspace, so the Schwarzian degrees of freedom were not included. On the other hand, Iliesiu et al Iliesiu argued that the minisuperspace wave function is simply related to the wave functional of full JT gravity. This issue needs to be better understood. Another possibility is that the condition of smoothness (absence of a conical singularity) at a=0a=0 is too restrictive. We know that the metrics contributing to the path integral are generally rather irregular, so the Hartle-Hawking proposal of integrating over smooth metrics should not be taken too literally. Finally, the Schwarzian boundary condition was imposed in Ref.Iliesiu assuming that the dynamics of the boundary curve at large aa are completely decoupled from the geometry at small aa. It is conceivable, however, that the decoupling is incomplete, so the conditions of closure and maybe smoothness modify the asymptotic behavior of ΨHH\Psi_{HH}.

In this paper we utilized the familiar 4D4D minisuperspace framework in order to explore the closely related JT quantum cosmology and to define the corresponding HH wave function. Due to the simplicity of JT theory, one can hope that with better understanding the relation between the two models will be reversed and JT cosmology will provide important insights for the 4D4D case. Towards this goal, it would be interesting to do a path integral calculation of ΨHH\Psi_{HH} directly in JT model, without a reference to KS and without using the minisuperspace truncation. This would help to understand the Schwarzian issue that we referred to above. It would also be interesting to define the tunneling wave function in both JT and KS models. We hope to return to these problems in future work.

Acknowledgements

We are grateful to Jose Blanco-Pillado, Jonathan Halliwell, Oliver Janssen, Jorma Louko, Juan Maldacena and Sandip Trivedi for very useful discussions and to Raymond Laflamme for sending us his 1986 thesis.

Appendix: Higher order corrections

The rescaled action of Eq.(4.21) can be split into two components. The first is S~E0=S~E(bH=1)\tilde{S}_{E0}=\tilde{S}_{E}(bH=1) and the second depends linearly on (1H2b2)(1-H^{2}b^{2}). Specifically, using the definitions for S~E,N~,u\tilde{S}_{E},\tilde{N},u and defining v=H2b2v=H^{2}b^{2} we can decompose the action in the following way:

S~E=(2N~3uN~(N~2+3u)212N~2(N~3))+(N~3uN~(N~2+3u)212N~2(N~3))(v1)\tilde{S}_{E}=\left(-\frac{2\tilde{N}}{3}-\frac{u}{\tilde{N}}-\frac{(\tilde{N}^{2}+3u)^{2}}{12\tilde{N}^{2}(\tilde{N}-3)}\right)+\left(\frac{\tilde{N}}{3}-\frac{u}{\tilde{N}}-\frac{(\tilde{N}^{2}+3u)^{2}}{12\tilde{N}^{2}(\tilde{N}-3)}\right)\left(v-1\right) (A.1)

Note that this expression is exact and not an expansion to first order in (1H2b2)(1-H^{2}b^{2}).

Setting v=1v=1 in the action (A.1) and taking a derivative with respect to N~\tilde{N}, we obtain Eq.(4.26) for the saddles. These are the saddles of the zeroth order action S~E0\tilde{S}_{E0}. We will refer to the 5 solutions of this equation as N~i\tilde{N}_{i} with i={1,2,3,4,5}i=\{1,2,3,4,5\}.

It can be shown that if we introduce a perturbation xx as in Eq.(4.41), N~=N~i+x{\tilde{N}}={\tilde{N}}_{i}+x, the action is extremized for

x=(1v2v)1f(N~i,u)+1vh(N~i,u)x=\left(\frac{1-v}{2v}\right)\frac{1}{f(\tilde{N}_{i},u)}+\frac{1}{v}h(\tilde{N}_{i},u) (A.2)

where the function hh vanishes at all saddles N~i\tilde{N}_{i}. Thus the perturbed saddles will be given by

N~i=N~i+(1v2v)1f(N~i,u)\tilde{N}_{i}^{\star}=\tilde{N}_{i}+\left(\frac{1-v}{2v}\right)\frac{1}{f(\tilde{N}_{i},u)} (A.3)

Setting N~=N~i\tilde{N}=\tilde{N}_{i}^{\star} in the action (A.1) and expanding to 2nd order in (1v)(1-v) we notice the following. The 0-th order term does not depend on the function ff, as expected. The 1st order term takes the form

[(Ni~3uN~i(N~i2+3u)212N~i2(N~i3))+38f(N~i,u)(N~i22N~i+u)(N~i34N~i23N~iu+6u)(N~i3)2N~i3](v1)\left[\left(\frac{\tilde{N_{i}}}{3}-\frac{u}{\tilde{N}_{i}}-\frac{(\tilde{N}_{i}^{2}+3u)^{2}}{12\tilde{N}_{i}^{2}(\tilde{N}_{i}-3)}\right)+\frac{3}{8f(\tilde{N}_{i},u)}\frac{\left(\tilde{N}_{i}^{2}-2\tilde{N}_{i}+u\right)\left(\tilde{N}_{i}^{3}-4\tilde{N}_{i}^{2}-3\tilde{N}_{i}u+6u\right)}{(\tilde{N}_{i}-3)^{2}\tilde{N}_{i}^{3}}\right](v-1) (A.4)

From (4.26) we see that the term depending on ff vanishes. Thus, the action up to first order corrections is

S~E=(2Ni~3uN~i(N~i2+3u)212N~i2(N~i3))+(N~i3uN~i(N~i2+3u)212N~i2(N~i3))(v1)+𝒪[(1v)2]\tilde{S}_{E}=\left(-\frac{2\tilde{N_{i}}}{3}-\frac{u}{\tilde{N}_{i}}-\frac{(\tilde{N}_{i}^{2}+3u)^{2}}{12\tilde{N}_{i}^{2}(\tilde{N}_{i}-3)}\right)+\left(\frac{\tilde{N}_{i}}{3}-\frac{u}{\tilde{N}_{i}}-\frac{(\tilde{N}_{i}^{2}+3u)^{2}}{12\tilde{N}_{i}^{2}(\tilde{N}_{i}-3)}\right)\left(v-1\right)+\mathcal{O}\left[(1-v)^{2}\right] (A.5)

This equation is the same as Eq.(A.1) with N~=N~i\tilde{N}=\tilde{N}_{i}. This means that the 1st order corrections to the action are obtained by finding the saddles for v=1v=1 and inserting them into the action for v1v\neq 1. The perturbed saddles contribute only to 2nd order and higher corrections to the action. Note that we did not make any specification for which saddle we are referring to. This analysis is true for all 5 saddles.

In the classically allowed region the contributing saddles for the Hartle-Hawking solution are N1,N2N_{1},N_{2}. In this regime the function f(N1,2,a)f(N_{1,2},a) is given by

f(N1,2)=3(1H2a2)2±2iH2a21±iH2a2H2a21f(N_{1,2})=\frac{3(1-H^{2}a^{2})}{2\pm 2i\sqrt{H^{2}a^{2}-1}\pm iH^{2}a^{2}\sqrt{H^{2}a^{2}-1}} (A.6)

Thus, the first order correction to the saddles with respect to (1Hb)\left(1-Hb\right) is given by

H2N1,2=1±iH2a21(1Hb)(2±2iH2a21±iH2a2H2a213(H2a21))H^{2}N_{1,2}=1\pm i\sqrt{H^{2}a^{2}-1}-\left(1-Hb\right)\left(\frac{2\pm 2i\sqrt{H^{2}a^{2}-1}\pm iH^{2}a^{2}\sqrt{H^{2}a^{2}-1}}{3(H^{2}a^{2}-1)}\right) (A.7)

and the action evaluated at N1,2N_{1,2} up to second order corrections is

SE(N1,2)SE1(N1,2)π(1Hb)23H2(2±2iH2a21±iH2a2H2a211H2a2)S_{E}(N_{1,2})\approx S_{E1}(N_{1,2})-\frac{\pi(1-Hb)^{2}}{3H^{2}}\left(\frac{2\pm 2i\sqrt{H^{2}a^{2}-1}\pm iH^{2}a^{2}\sqrt{H^{2}a^{2}-1}}{1-H^{2}a^{2}}\right) (A.8)

where SE1(N1,2)S_{E1}(N_{1,2}) is the action evaluated at N1,2N_{1,2} up to first order corrections in (1Hb)(1-Hb). From the above we can specify the coefficient function F(a)F(a) in Eq. (5.4) as

F(a)=π3H2(2±2iH2a21±iH2a2H2a211H2a2)F(a)=\frac{\pi}{3H^{2}}\left(\frac{2\pm 2i\sqrt{H^{2}a^{2}-1}\pm iH^{2}a^{2}\sqrt{H^{2}a^{2}-1}}{1-H^{2}a^{2}}\right) (A.9)

Its real part is

ReF(a)=2π3H2(1H2a2)\textrm{Re}F(a)=\frac{2\pi}{3H^{2}(1-H^{2}a^{2})} (A.10)

References

References