<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by CERN Document Server

AEI-2002-011 The Real Wick rotations in quantum gravity

Arundhati Dasgupta1, Max-Planck-Institut f¨ur Gravitationsphysik, Am M¨uhlenberg 1, D-14476 Golm, Germany.

We discuss Wick rotations in the context of gravity, with emphasis on a non-perturbative Wick rotation proposed in hep-th/0103186 mapping real Lorentzian metrics to real Euclidean metrics in proper-time coordinates. As an application, we demonstrate how this Wick rotation leads to a correct answer for a two dimensional non-perturbative path-integral.

1email: [email protected]

1 1 Introduction

The evaluation of the path-integral for any field theory, essentially involves a rotation to Euclidean signature space-times. This renders the complex measures and oscillatory weight factors geıS real and the functional inte- gral computable. One then Wick rotatesD back in the final answer to get Lorentzian quantities, or invoke the Osterwalder-Schraeder theorems. The usual methods used for field theories in flat space-times are difficult to ex- tend to gravity as there is no distinguished notion of time. There have been attempts to define a Euclidean gravitational path-integral and relate them to Lorentzian quantities. Initially complex weights were replaced by real Boltz- mann weights and the sum performed over all possible Riemannian metrics [1]. This prescription is ad hoc as there is no apriori relation between the Lorentzian and Riemannian sums. However, if one believes that Lorentzian quantities are the relevant physical quantities, then a prescription must exist for identifying the corresponding Euclidean metrics to be summed over. The usual Wick rotations pertain to static metric solutions of Einstein’s equa- tions. Given the time coordinate, t, Wick rotation amounts to going to the space of complex metrics, and integrating over the axis. As evident, this prescription cannot be directly extended to a non-perturbative configuration space path integral where there is aprioiri no distinguished no- tion of time. The solution to this problem was proposed in [2], where one gauge fixes the metric and performs a Wick rotation on the space of gauge fixed metrics. This ‘non-perturbative’ Wick rotation is a first step in mapping Lorentzian path-integrals to Euclidean path-integrals, in configuration space. The map- ping crucially does not involve a complexification of space-time and maps the real Lorentzian metrics to corresponding real Riemannian metrics in the path-integral. To verify that our proposal works, we discuss some examples where an explicit answer can be obtained. In particular we show how in 2 dimension, the only case where the ‘non-perturbative’ sector can be explicitly evaluated, a correct 2-loop answer is obtained. We show how the rotation relates the Lorentzian and Euclidean sectors of the theory. A completely independent evaluation of the Euclidean path integral in 2 dimensions would include ‘baby universes’ and hence would not satisfy the stringent causality requirements which the Lorentzian histories have to satisfy. We first define the path-integral in Lorentzian space, gauge fix the path-integral, Wick ro- tate to Euclidean space to perform the computations. We also restrict the

2 topology of the manifold to be fixed. The Faddeev-popov determinant which arises due to gauge fixing gives the effective in proper-time coordi- nates with the only degree of freedom being g11 component of the metric. Using a hueristic arguement by coupling 2 dimensional gravity to matter, we show that this component of the metric has the product form g11 = γ(t)γ(x). This immediately reduces the problem to that of , and the zeta function regularisation of the Faddeev-Popov determinant gives the effective action as a quantum mechanical action. The path-integral can be explicitly evaluated once this is determined. We then rotate the final re- sult to obtain the physical Lorentzian answer. The answer obtained in the end is same as that obtained in a discretized evaluation of the path-integral [3]. A continuum calculation of the path-integral in proper-time coordinates had been previously obtained in [4]. Our approach is along the formalism of [2], where we derive the effective action from the Faddeev-Popov deter- minant arising in the measure unlike [4]. Note the choice of gauge which is Proper-time gauge plays a crucial role in all our arguements. In this gauge, causality is manifest, which restricts the path-integral. We also explain how our Wick rotation helps in identifying the divergence in the 3 and 4 dimen- sional Euclidean path-integral and briefly results in [2]. The second section is a detailed discussion of the Wick rotation proposed in [2], and the reason for it being non-perturbative is pointed out. Much of this appears in [5]. The third section enumerates the applications. An explicit calculation of the two dimensional path-integral is given, and there is a short review of the conformal mode cancellation of [2] to complete our argument. We end with a discussion.

2 The Non-perturbative ‘Wick Rotation’

Since the perturbative path integral does not define a fundamental theory of quantum gravity, we have to consider, non-perturbative evaluations (see [6] for recent examples). Since there is no distinguished notion of time in gravity, definition of Euclidean path-integral is not well defined. A non-perturbative Wick rotation however can be implemented on the space of gauge fixed met- rics as shown in [2]. Here, after fixing the gauge in the Lorentzian path integral, the appropriate gauge-invariant metrics are written in coordinates with a distinguished time . We then ‘Wick rotate’ in the time-like component of the gauge invariant metric. To implement this, we need to identify a suit-

3 able gauge, and the proper-time gauge precisely fits in. Here the notion of time is distinguished, and causality is manifest. Our construction, including the ‘Wick rotation’, is non-perturbative and background-independent. Our gauge is defined by the use of Gaussian normal coordinates, where the d-dimensional metric takes the form

2 2 i j ds = dt + gijdx dx ,= 1,i=1, 2, ..., d 1, (1) ± − singling out a proper time t.Thevalue = 1 corresponds to Lorentzian- signature metrics. Our generalized Wick rotation− consists in mapping  , (2) 7→ − while keeping gij unchanged. It has the following properties:

ε − ε 2 2 i j d s = ε d t + gij dx dx 2 ε 2 j i ds = − dt + gij dx dx

EUCLIDEAN METRICS LORENTZIAN METRICS

Real Lorentzian metrics are mapped to real Euclidean metrics, thus • making the DeWitt metric and associated scalar products well-defined. Note that this is unlike the prescriptions used in [7] where complex- ification of the metric occurs. This is necessary for us as we need not only the action to be real but also the De-Witt metric Gµνρσ in the configuration space of metrics, and the associated scalar products 4 µνρσ = d x√ghµν G hρσ to be real too. A simple illustrative example is thatR of the de Sitter metric (foliated by flat hypersurfaces), ds2 = dt2 + e2Λt d~x2, (3) which is also a solution to the Einstein equations. The standard rota- tion t it takes this to a complex-Euclidean metric. Not only that, → 4 the scalar product defined above is clearly complex for d=4. The Gaus- sian normalisation conditions used to define the correct measure in the form h exp− =1 (4) Z D will no longer have a real exponent, and hence the ‘Euclideanised’ in- tegrals will also be ill defined. The Wick rotation to real Riemannian metrics which is achieved in (2) is a necessity here

The initial motivation for Wick rotation is of course the replacement • of complex amplitudes in the path-integral by real weights. This is very well achieved in our prescription. The Complex amplitudes which i√ S appears in the path integral can be written in this gauge as e − = 1, with the action S as: | −

 d (d 1)  ijkl S[g]= d x det gij − R 2Λ G( 2)(∂0gij)(∂0gkl) , −16πG − − 4 − Z q  (5) where Gijkl =(gikgjl + giljk + Cgijgkl)/2 is the DeWitt metric defined on the functional space of metrics and Λ the cosmological constant. (The one-parameter ambiguity C gets fixed by the form of the kinetic term in the action to C = 2.) Now clearly, the Wick rotation pre- scription given as   helps− in mapping the complex weights to real →−Seu Boltzmann weights e− The map is one-to-one and onto the space of gauge-fixed Euclidean • metrics. We assume that metrics of the form (1) are in a suitable sense dense in the quotient of metrics modulo diffeomorphisms, see [2] for a detailed discussion.

The non-perturbative nature of our Wick rotation becomes manifest • when we look at solutions to Einstein’s equations. In general, Lorentzian solutions of the Einstein equations are not mapped to solutions of the Einstein equations in Euclidean space by (2). In proper-time coordi- nates, as written in (1), the entire  dependence can be absorbed into a rescaling of the time coordinate by √ (See Appendix 1). Thus L E Lorentzian and Euclidean solutions of gij(τ,r)=gij (τ,r)havethe same functional dependence on the scaled parameter τ = √t,how- ever,   would implement the change t ıt. Note that in our →− → proposal for the Wick rotation, we assume that gij is independent of

5 . Thus we are essentially not looking at Classical metrics, or Saddle points of the path-integral. The usual ‘common-sense’ associated with Einstein’s equation solutions cannot be implemented in our space of metrics.

Another positive point of this Wick rotation, is the fact that since gij(t) • remains unchanged, whatever properties the Lorentzian metric has as a function of Lorentzian time are retained in the Euclidean case. As an example, the metric for desitter space (foliated by spheres) is:

ds2 = dt2 +cosh2 t(dχ2 + dΩ2)(6) − The standard t it would give, → ds2 = dt2 +cos2 t(dχ2 + dΩ2)(7) − which is pathological at t = π/2. ± However, our Wick rotation gives

ds2 = dt2 +cosh2 t(dχ2 + dΩ2)(8)

with the result that the Euclidean metric is not pathological at t = π/2. Hence, once we have fixed the configuration space of metrics for the Lorentzian case, we can retain the mapped Euclidean set of metrics in the sum

The prescription (2) and the t it agree only in a few cases, for ex- ample, the Minkowski and Schwarzschild→ metrics (which are static). If we were primarily interested in a perturbative treatment around given classical solutions, we might worry about the last property. However, we need a Wick rotation that works on the space of all gauge fixed metrics, since this is the space we wish to sum over in the non-perturbative path integral, for given initial and final boundaries. We conclude that in a non-perturbative context the prescription   is perfectly well-defined, and clearly preferred to a rotation to imaginary7→ − (proper) time.

6 3 Applications

The Wick rotation discussed in the previous section has to be implemented in a actual path integral to illustrate that it actually works. 2 dimensions is simplest to start with, and a full non-perturbative path-integral can be evaluated.

3.1 PI in 2 dim In two dimensions there are many problems absent like the conformal factor problem in higher dimensions, and the theory is renormalisable. The exact integral helps in illustrating our Wick rotation, which is specific to the proper- time gauge we have chosen. We explicitly define the path-integral in two dimensional Lorentzian geometry before ‘Wick rotating’ to Euclidean space- times. As illustrated in sum over discretized geometries, the Lorentzian sum excludes higly branched polymer-like geometries found in conformal gauge calculations [3, 8]. Also, the Hausdorf dimension of space-time appears as 2 in [3]. We explicitly show below how a calculation in proper-time gauge with Lorentzian signature rules out any baby universe formation, and reduces the model to a quantum mechanical model, also discussed in [4]. This model is solved, and we reproduce the answer obtained in [3]. The starting point is the path-integral, with the action:

P Z = gµν J exp( ıS)(9) Z D − where J is the Faddeev-Popov determinant (FP) which arises due to the P gauge fixing to a metric gµν . After determining the dependence of the FP on the metric variable (which in two dimensions has only one free component), one proceeds to evaluate the entire path-integral. Here we have removed the diffeomorphism degrees of freedom. The form of the effective action Seff constitutes the reduced space. We start by giving a heuristic argument explaining how the imposition of restriction to cylinder topology implies a reduction to quantum mechanics. As seen in dicscrete calculations [3], one essentially studies correlations of loops √γdx = l(t) in time. To implement this in the path-integral, one has to go toR reduced space before evaluating the sum over histories.

7 Cylinder Topology For that we couple 2-dim gravity to matter fields and examine this re- striction which is independent of the gauge fixing of the metric. As explained in [9] the matter integral can be done exactly, and the effective energy momentum tensors determined. In other words: = δF δgab ,with

M 1 µν Z = x exp( √gg ∂µx∂ν x)=exp(F ) (10) Z D −2 Z − The conformal anamoly (in Euclidean space) is given as [9] D gµν =( )[R(ξ)+constant] (11) µν 24π The right hand side of the above equations is coordinate invariant, and hence the LHS is also. Since gµν is a classical contravariant tensor, the expectation value of the energy momentum tensor also transforms covariantly under a diffeomorphism transformation. Using the above knowledge, one sees how the above look in the conformal set of coordinates and in the proper-time coordinates. The two set of coordinates are: 2 2φ(ξ) 2 2 ds = e dξ0 + dξ1 (12) = dt2 +γ (t, x) dx2 (13)

8 Where the metric in (12) is in conformal gauge with φ denoting the con- formal mode and the one in (13) in proper-time gauge with γ denoting the g11 component of the metric. The Wick rotation to Euclidean coordinates involves  . The coordinate transformations are very simple, and gives a relation→− between the derivatives of the coordinates as (see Appendix B):

 t˙2 + t02 = e2φ (14) exp(2φ) x˙ 2 + x02 = (15) γ with t˙ = √γx0,t0 = √γx˙ (16) ± ±

(Where t˙ denotes differentiation of t with respect to ξ0 and x0 denotes differentiation of x with respect to ξ1) This coordinate transformation can be implemented on the energy momentum tensors in (11). Using this, and expression for the Conformal anamoly we find that the metric in proper-time gauge gets severely restricted. To see the constraint T01 = 0 in the gauges discussed above, one needs to change the gauge and introduce

g00 =1+h00,g01 = h01 (17) such that δZM h0µ=0 = T0µ = 0 (18) δh0µ |

We impose the constraint T01 = 0 in both the gauges initially to see the consequences. CF Thus imposing simultaneously T01 = T01 =0leadsto

CF 1 =˙xx0  = 0 (19) −γ ! Due to the presence of the conformal anamoly we do not impose the con- straints T00 = T11 = 0 but one is left with the following options from the above: 1 x˙ =0, orx =0 = (20) 0 00 γ 11

The latter condition is disallowed as from simple considerations, T00 is the Hamiltonian density in this gauge, and T11 is proportional to the Lagrangian density. These two are not equal but differ from each other as shall be

9 justified later on when we have determined the form of the effective action. The form of the T00 and T11 in [4] also confirm the above. Also at this stage we examine the consequences of only the imposition of T01 =0. It also happens that when we continue to Lorentzian signature by (14),the condition x0 = 0 is not possible. What we have left isx ˙ = 0 and hence by (14), t0 = 0, which implies that γ = γ(x)γ(t)andφ(x, t)=φ(t). realised in the proper-time coordinates. No Topology change in the form of baby universe formation or wormholes are allowed. This is reminiscent of the work done in [10] where Lorentzian cobordism severely restricts topology, albeit the analyses is at the classical level.

Branching Topology In [4] this additonal restriction was derived as a solution to the constraints of matter coupled gravity. It is interesting to note that this precise restriction when translated to conformal gauge makes the conformal mode depend only on the time coordinate and reduces it to a quantum mechanical model. The implementation of T01 = 0 or the equation of motion for the g01 component, is tied up with the factorisation of correlations along the spatial sections and time like sections. This is confirmed by the discrete model calculations, where correlations are studied between time-slices. Using the above now, we implement the path-integral in configuration space, using our Wick rotation.

10 Implement our gauge the path integral takes the following form:

Z = γJ(, γ)ei√Λ d2x√γ (21) D Z R The Wick rotation which takes   is very simple now, and we com- pute the entire path integral calculation→− in Euclidean space. The main task is the evaluation of the Faddeev-Popov determinant J and it’s dependence on γ. To evaluate the determinant explicitly, we refer to [2] and use the Gaussian normalisation condition to fix the exact value. Expanding the metric around the diffeomorphism orbits:

PT PT δgµν = δg + µξν + νξµ = δg + Lξ (22) µν ∇ ∇ µν For our purposes we set C=0, a value predetermined by the action we have chosen. (Note that this is very specific to 2-dimensions) Using the explicit expression in (22), we have:

δgµν exp( <δgµν ,δgµν >)=1 Z D − Using the coordinate transformation given in (22), and denoting the Jacobian of the transformation as J, the induced measure on the tangent space is determined. Note that since the exponent is diffeomorphism invariant, we determine the scalar product at the base point on the gauge slice. As given in Appendix C, the Faddeev-Popov determinant can be written as a product of three scalar determinants. After a few redefinitions as in Appendix C, the entire exponent can be written as a sum of complete squares.

2 δg11 ξ0 ξ1 exp( d xδg11( 11)δg11 + ξ0 ( 0)ξ0 + ξ10 ( 1)ξ10 ) (23) Z D D D − Z ∇ ∇ ∇ Where the vectors have been re-defined in order to complete the squares as given in Appendix C. The Gaussian integrals can now be done to yield the appropriate Jacobians in tangent space. The Jacobians evaluated here also determine the measure in base functional space. The Faddeev-Popov determinant is thus obtained as a product of three scalar determinants.

J = det ( )det ( )det ( ) (24) S ∇0 S ∇1 S ∇11 where the explicit expressions for the 0, 1, 11 are given as in Appendix C. Being the simplest, we concentrate∇ on ∇ first.∇ ∇0 11 1 (lnγ˙ )2 detS( 0)=detS 4(∂0†∂0 + ∂1†∂1 + ) (25) ∇ " 2γ 4 # The operator in the above is thus:

1 (lnγ˙ )2 P = 4(∂0†∂0 + ∂1†∂1 + ) √γδ(t t0)δ(x x0) (26) " 2γ 4 # − − The determinant of this operator then satisfies the following eigenvalue equation: Pψ = n√γψ (27) where the eigenfunctions have to satisfy appropriate boundary conditions on t =0andt = T surfaces. We impose Dirichlet boundary conditions, and as shown below, assume that the operator equation can be solved by a separation of variables. Once the separable solutions are found, any other solution can be expressed as a linear combination of those [12]. To achieve separability of the solution, we write the operator equation in the following form

γ(t) (lnγ˙ )2 1 4 ∂0†∂0 + ψ(t) nγ(t)= ∂1†∂1 ψ(x) (28) ψ(t) 4 ! − 2ψ(x)γ(x)   What is evident from the above is the fact that the above equation is eas- ily separable in the variables but with the consequence that the eigenvalue n depends on the time dependent operators only. This is a clear indication of the reduction of the system to a quantum mechanical system with the reduced action depending only on the time coordinate. (The effective action is derived from J =Πn. Also the splitting we found in γ = γ(t)γ(x)con- tributes crucially to our conclusion. Note that by assuming separability of the solutions, we have assumed that the node of the eigenfunction  agrees with constant t surfaces, in other words, whatever the boundary condition, the  = 0 surface coincides with a t = constant surface. The eigenvalue equation for the t dependent part becomes now

1 ˙ ˙ 2 4γ(t) (∂0 + ln γ)∂0 ψ(t)+γ(t)(ln γ) (t) nγ(t)ψ(t)= αψ(t) (29) − 2  − − where α is a positive constant independent of t, x.

12 Finding the effective action hence reduces to the problem of finding the determinant of the operator

2 α 4(∂†∂ )+(ln˙γ) + . (30) 0 0 γ In order to determine the eigenvalue, we perform a zeta function regular- isation. We follow [11] to perform this. Since it is difficult to evaluate the exact determinant, we try to guess the dependence on the γ(t), by assuming that the effective action has the form

2 Γ= dt f(γ)(∂0γ) + V (γ) (31) Z   ˙ Thus, for configurations which have ln γ = 0, we recover V (γ0), where V (γ0) is the potential depending on the constant value γ0. In other words the potential is simply given by the zeta function regularisation of the operator in (30) with the value of γ0:

dtV (γ0)= ζ0(0) (32) Z − Note that the metric in (13) is still non-constant due to the dependence on γ(x). Thus, we consider the operator for γ0 which reduces to the following form

2 α 4∂0 + (33) − γ0 The Heat Kernel for this operator satisfies the following equation:

2 α (4∂0 )G(t, t0,τ)= ∂τ G (34) − − γ0 − with the boundary condition that 1 G(t, t0,τ =0)= δ(t t0) (35) γ √ 0 −

(the metric in this direction is flat) The heat Kernel G(t, t0,τ)canbesolved to give: 2 (t t ) α π − 0 τ G(t, t0,τ)= e− 16τ − γ0 (36) s4τγ0

13 The ζ function is thus:

α 1 ∞ s 1 π τ ζ(s)= dττ − dt e− γ0 (37) Γ(s) Z0 Z r4τ The dt is not infinite, and cancels with the volume element in the LHS of (32).R The task thus reduces to the determination of

s 1/2 Γ(s 1/2) γ0 − γ0 ζ0(0) = √π − ln +Ψ(s 1/2) Ψ(s) Γ(s) α − α − −       s=0

(38) At this juncture, we keep only finite terms as s 0, and hence only the second term of the above integral. It is easily justifiable→ as the first term, ζ (0) due to the pole of 1/Γ(s), at s 0 contributes vanishingly to e− 0 . Thus the potential has a dependence→ of the form: 1 dtV (γ0) dt√γ0 (39) Z ≈ Z γ0

Next we look at the operator of the form 1, which has the form as given in the Appendix: ∇ 0 4 0 1 1 1 1 = 1 1+(∇ )− 1 (40) ∇ γ2 ∇ γ ∇ ! with 2 0 γ γ ˙ (ln γ0) 1 = ∂1†∂1 + ∂0†∂0 + (ln γ)∂0 + ∇ 2 2 4 !

The above operator is very similar to the operator 0 except that now the measure is of 1/ γ . However, this does not modify∇ the final answer, and √ 0 one gets a very similar analyses as above, except for a change of factor 2. And the contribution to the effective potential is of the form dt1/√γ0 To determine the derivative term in (31), we resort to the demandR of locality of the action. To see what such a term in the action should look like we make ascalingofjustthet coordinate to go to conformal gauge with the metric

2 2 2 ds = γ( dt0 + dx ). − 1 2 The local term in these set of coordinates would be dt0 4 (∂t0 ln √γ) .Trans- lating this back to original coordinates we concludeR that the effective action can be written as:

14 2 1 d√γ(t) β I = dt + +Λ√γ(t) (41)   Z 4 γ(t) dt ! √γ(t) The constant β is finite. q Remarkably, the above coincides with the effective action found in [4]. To extract the finite part of our effective action, the observation of separability proved crucial. In principle we have not argued for the absence of higher derivative terms, however conditions of locality are stringent. The fact that the potential has the form expected, is a very remarkable result. Once the action has been determined, one needs to complete the evalua- tion of the propagator

l2 l G(l1,l2,T)= D exp( I(l)) (42) Zl1 l − where we are now using the variable γ(t)=l. Though the action is quan- tum mechanical, it is not clear thatq the evaluation of (42) is simple. An exact evaluation is left for a future publication. Instead, one resorts to the following representation of the propagator.

HT G(l ,l ,T)= (43) 1 2 1| | 2 This is similar to the approach in [4]. Being a quantum mechanical action, the spectrum of the Hamiltonian can be determined. The derivation of the Hamiltonian is done by identifying the momentum conjugate to the variable l(t), and then subsequently replacing that by ∂/∂l. The eigenvalue equation for the determination of the spectrum becomes: d2 β l Λl Ψ(l)=EΨ(l) (44) dl2 − l − ! By scaling the variable l by 2√Λ, one obtaines a differential equation, which has as its solution the Whittaker functions [13]. d 1 β E ψ(z) = 0 (45) dz2 − 4 − z2 − 2√Λz ! with the following form

√β+1/4+1/2 z/2 1 E 1 1 Ψ(x)=z e− Φ( β + + ;2 β + +1;z) (46) s 4 − 2√Λ 2 s 4

15 One then uses the above in (43). To find this, we have to define E to take discrete values in units of the 1/√Λ. They can be related to Laguarre polynomials as in [4], with β +1/4 E + 1 = n where n is an integer. − 2√Λ 2 − Using this quantised E, andq appropriate normalisation constant, the wave function now has the form:

1 z/2 a/2 a Ψ(x)= e− z Ln(z) (47) (n+a) Γ(1 + a)Cn q with a =2 β +1/4 + 1. Using this, and substituting this in (43), one evaluates the sumq to get

1 2√Λl l 2√Λ(l1+l2)coth(2√ΛT ) 1 2 G(l1,l2,T)=e− Ia (48) sinh(√ΛT ) sinh(√ΛT )!

This amplitude is the same obtained in different approaches in [3, 4], with a = 1 coincides with the result found in [3]. However, the interesting part is the rotation back in the final result to get a physical Lorentzian answer. The Wick rotation initially involved setting   = 1 to get the Euclidean set of metrics to be summed over. In the final→− answer, this rotation is hidden in the definition of proper-time T , or the Euclidean time is √T .(Thiscanbe traced back to equation (21)). Substituting T iT , a physical Lorentzian answer is obtained. → In two dimensions, our Wick rotation gives an exact and correct answer for the non-perturbative path-integral. Hence it is interesting to see how far we can implement this in higher dimensions.

3.2 In higher dimensions Our formalism crucially rests on the proper-time gauge. In higher dimen- sions, as discussed in [2], the action is as given in (5). The Euclidean and Lorentzian actions are related by  . We will set  = 1 in what follows. 7→ − As discussed in [2], we showed that the above helps in the identifica- tion of the negative definite part of the action, which is real unlike what a complexification of the Lorentzian metric would achieve.

ND (d 1)(d 2) d (d 1)λ˜ ˜ 2 S = − − d x e − (∂0λ) (49) − 16πG Z 16 with logg ˜ λ˜ = λ + . (50) 2(d 1) − (λ is the conformal mode). The full action can be written as S = SND(λ˜)+ rest ˜ rest S (˜gij, λ), where we refrain from giving the remainder S explicitly [2]. The main task is to now deal with the presence of the eλ dependence in the measure for λ. In other words we need a transformation which is similar to that used in 2-dimensional gravity [14]: ˜ ˜ g˜ λ λ λ J(λ, g¯) g˜g λg¯ (51) D e g¯D e g¯ → D ¯D We assume that this is achieved with the Jacobian J which is a local function S of the fields and hence by symmetries has the form e− L which for obvious symmetry reasons has to be of the following form

d ˜ 2 ijkl 2 SL = d x√g¯ A ∂ λ + B R¯ + C (R¯ijklR¯ )+D ( R)+... (52) 0 ∇ Z     where A, B, C, D.. are constants whose value will depend on the regularistion chosen. What is however clear that the kinetic term for the conformal mode appears in the same form as in the original action and can be absorbed back into the action with just a renormalisation of the coupling constant Inspired by a conjecture lying at the heart of 2d Liouville quantum gravity [14], we will now assume that an elimination of the conformal factors from the space-time measures in the exponent and the determinants (such that (d 1)λ √ge˜ − √g˜) is equivalent to a renormalization of the action. We can now pull all7→ Jacobians and other λ˜-independent terms out of the λ˜-integral. The leading divergence of what is left under the integral is of the form

ddxλ∂˜ ∂ λ˜ λe˜ 0† 0 . (53) D Z R

This formally yields 1/ det ( ∂†∂ ), which is badly divergent, since ∂†∂ is S − 0 0 0 0 a positive operator. Nevertheless,q it is of exactly the same functional form as the inverse of scalar part of the Faddeev-Popov determinant as determined in [2]. One can then cancel the two divergent determinants, leaving behind the integral in gij⊥. This completes the argument for the cancellation of the conformal divergence.

17 4 Discussions

In this article we have argued for a novel type of Wick rotation which seems particularly suited for a non-perturbative treatment of the gravitational path integral. It is an example of how the need for going beyond perturbative treatments in quantum gravity forces us to abandon a well-loved, but essen- tially perturbative concept like imaginary time. As an application, we gave a 2dimensional non-perturbative path integral calculation and outlined how in higher dimensions, the conformal sickness of the Wick-rotated path integral is cured non-perturbatively in this new framework. The two dimensional path- integral revealed that the Wick rotation does work in a non-perturbative set up. And in higher dimensions the cancellation of the conformal mode di- vergence leads to the hope that the path-integral can be evaluated to yield exact results. Acknowledgements: This paper grew out of extended discussions with R. Loll. The author wishes to thank her for a careful reading of the manuscript, pointing out of errors, criticism of arguments, many helpful comments and encouragement to publish.

5 Appendix A

The non-zero Christoffels: 0  Γ = gij, (54) ij −2 0 1 Γi = gikg (55) 0j 2 kj,0 i 1 ik (2) i Γ = g (glk,j + gjk,l glj,k)=Γ (56) lj 2 − lj The Ricci Tensor components

1 ik 1 jm kn R00 = g gik,0 + g g gjn,0gkm,0 (57) 2 ,0 4   (2)j 1 jk (2)k 1 jm (2)j 1 kl Ri =Γ g gik, ,j+Γ g gkm, Γ g gil, (58) 0 ij,0 − 2 0 ij 2 0 − jk 2 0 (2)   lm km Rij = R +  gij, , + gij, g glm, gik, gjm, g (59) ij 0 0 0 0 − 0 0 Scaling the time coordinate by √, and putting the above in Einstein’s equation, the entire  dependence is removed from them (we also need to use 2  = 1). So, one obtains gij = gij(√t, x)

18 6 Appendix B

The coordinate transformations relating metric components in conformal gauge (12) and (13).

t˙2 + γ x˙ 2 = e2φ (60) − − tt˙ 0 + γ xx˙ 0 = 0 (61) 2 − 2 2φ t0 + γx0 = e (62) − Dividing (60) by (62) and using (61) one obtains:

t˙ = √γx0 (63) t0 = √γx˙ (64)

Putting the above two in (60) and (62) one obtains (14,15).

7 Appendix C

This is a summary of the Faddeev-Popov determinant calculations in our gauge. A ab initio derivation of the determinant is required as our gauge is non-covariant and two dimensions is the easiest place to demonstrate that. By ab-initio we mean that we start with the Gaussian-Normalisation con- dition: <δg,δg> gµν e− = 1 (65) Z D where δgµν is fluctuation in tangent space around a given base point in the functional space of metrics. We assume that our gauge fixed proper-time metric, is of the following form: 10 gµν = (66) 0 γ ! with the following non-zero Christoffel symbols.

0 1 1 1 ˙ 1 1 Γ = γ,˙ Γ = (ln γ), Γ = (ln γ)0 11 −2 01 2 11 2

γ˙ ∂t,γ0 ∂x denote the derivatives and γ = γ0(t)γ1(x). ≡Using the≡ above:

19 2 1 <δg,δg> = d2x g (2 ξ )2 + ( ξ + ξ )2 + (δg +2 ξ )2 √ 0 0 0 1 1 0 2 11 1 1 Z " ∇ γ ∇ ∇ γ ∇ # (δg )2 = d2x γ 11 +4ξ˙2+ √ 2 0 Z " γ 4 1 1 2 1 1 + (ξ0 )+ γξ˙ (ln γ)0 ξ + ξ0 + γξ˙ (ln g)0ξ δg 2 1 0 1 1 0 1 11 γ  2 − 2   2 − 2  ! 2 ˙ ˙ 2 + ξ1 + ξ0 (ln γ)ξ1 (67) γ − #   Using partial integrations, the above can be written in a more convenient form: ˙ 2 2 1 2 1 (ln γ) = d x√γ δg +4ξ ∂†∂ + ∂†∂ + ξ  2 11 0  0 0 1 1  0 Z γ 2γ 4 4 ˙  1  4 1 ˙  + ξ0(ln γ)(∂1 ln g0)ξ1 ξ1 ∂0 + (ln γ) ∂1ξ0 γ  − 2  − γ   2   2 4 γ γln˙γ (ln γ0) + ξ1 ∂1†∂1 + ∂0†∂0 + ∂0 + ξ1 γ2 ( 2 2 4 ! ) 4 2 ˙ + ξ1∂1†δg11 + (ln γ)ξ0δg11 (68) γ2 γ # ˙ (where ∂0† = ∂0 +1/2(ln γ), ∂1† = ∂1 1/2(ln γ)0) One can− then complete the squares− to do− the ξ integrations and obtain:

2 1 2 = d x δg + ξ0 ξ0 + ξ0 ξ0 W W (69) 2 11 0 0 0 1 1 1 0 1 Z γ ∇ ∇ − − where,

2 1 (ln˙γ) =4(∂†∂ + ∂†∂ + ) (70) ∇0 0 0 2γ 1 1 4 2 4 γ γ (ln γ0) = (∂†∂ + ∂†∂ + (ln˙γ)∂ + ) ∇1 γ2 1 1 2 0 0 2 0 4 1 1 1 + (∂† ln˙γ)∂ ( )− ∂†(∂ 2ln˙γ) (71) γ 0 − 1 ∇0 γ 1 0 −

20 ˙ 1 1 ln γ W = δg ln˙γ − δg (72) 0 γ 11 ∇0 γ 11

1 1 1 W = δg ∂ + ln˙γ∂†(∂ 2ln˙γ)∂ − − 1 11 γ 1 1 0 − 1∇0 ∇1   ˙ ˙ 1 ln γ (∂0† ln γ)∂1 0− + ∂1† δg11 (73) × − ∇ γ !

After this, one can just evaluate the gaussian integrals in δg11, ξ0,ξ1 in the (65) to get the Jacobian in terms of three scalar determinants. The determi- nants apparently look quite complicated, however are scalar determinants. However for non-zero γ it becomes a Herculian task to determine the exact form of the effective action. However, following the zeta function regulari- sation methods used in [11], the form of the effective action can be guessed by examining around a value of the metric for which ln˙γ =0.Thisgives precisely the same action found in [4] for the determinant 0 as explained in section 3.1. For the rest of the determinants of operators∇ (71) and (73), we also do a zeta function regularisation. But most give diverging results due to presence of inverse of operators, and we absorb these in an infinite renormalisation of the effective action.

References

[1] S. W. Hawking, ‘The Path-Integral Approach to Quantum Gravity’ Gen- eral Relativity: An Einstein Centenary Syurvey, eds. (Cambridge Uni- versity Press, 1979)

[2] A. Dasgupta and R. Loll, Nucl. Phys. B 606, 357 (2001).

[3] J. Ambjørn, R. Loll, Nucl. Phys. B 536 407 (1998).

[4] R. Nakayama, Phys. Lett B 325 347 (1994).

[5] A. Dasgupta, R. Loll, ‘ The Real and the Sick: Wick Rotations and conformal divergence in the gravitational path integral’, AEI-2001-070, unpublished essay.

[6] J. Ambjørn, J. Jurkiewicz, R. Loll, Phys. Rev. Lett. 85 (2000) 924.

21 [7] J. D. Brown, E. A. Martinez, J, W. York, Jr, Phys. Rev. Letters 66 2281 (1991)

[8] J. Ambjørn, J. Correia, C. Kristjansen, R. Loll, Phys. Lett. B475 (2000) 24.

[9] A. M. Polyakov, Phys. Lett. B 103 207 (1981)

[10] M. Visser, Phys. Rev. D41, 1116 (1990)

[11] P. Ramond, Field Theory a Modern Primer Addison Wesley Publishing inc (1990).

[12] P. M. Morse, H. Feshbach, Methods of Theoretical , Mcgraw Hill (1953).

[13] I. S. Gradshteyn, I. M. Ryzhik Tables of Integrals, Series and Products Academic Press (2000) (sixth edition).

[14] J. Distler and H. Kawai, Nucl. Phys. B 321 (1989)

22