variational principles for nonsmooth metrics

10
PHYSICAL REVIEW D VOLUME 42, NUMBER 12 15 DECEMBER 1990 Variational principles for nonsmooth metrics Geoff Hayward* and Jorma ~ ouko~ Institute of Theoretical Physics, University of Alberta, Edmonton, Alberta, Canada T6G2JI (Received 16 May 1990) We analyze the variational principle of general relativity for two classes of metrics that are not necessarily smooth but for which the Einstein action is still well defined. The allowed singularities in the metric are (1)a jump discontinuity in the extrinsic curvature at a three-dimensional hypersur- face and (2) a conical singularity in the Euclidean Kantowski-Sachs metric on the manifold ~xs~. In agreement with general expectations, we demonstrate that in both cases the extremizing vacuum metrics are the usual smooth solutions to the Einstein equations, the smoothness conditions coming themselves out of the variational principle as part of the equations of motion. In the presence of a singular matter distribution on a three-dimensional hypersurface, we also demonstrate that the usu- al junction conditions for the metric are directly obtained from the variational principle. We argue that variational principles with nonsmooth Kantowski-Sachs metrics are of interest in view of con- structing minisuperspace path integrals on the manifold 5 XS2, in the context of both black-hole thermodynamics and quantum cosmology. The relation of nonsmooth Kantowski-Sachs variational principles to the issue of Lorentzian versus Euclidean path integrals in quantum gravity is briefly discussed. I. INTRODUCTION In the Euclidean path-integral approach to quantum gravity, the object of interest is a path integral of the form Here gp,, is a Euclidean metric defined on a four- dimensional manifold A, and I(g,,,) is the Einstein ac- tion I = - --- I I d4x(g +boundary terms . (1.2) 16r I? The boundary terms in (1.2) depend on M and on the boundary conditions imposed on the metric g,,,. ',* One application of this path-integral formalism is in black- hole thermodynamics, where Z is interpreted as a ther- modynamic partition function.2p4 Another application is in quantum cosmology, where Z is interpreted as the wave function of the ~ n i v e r s e . " ~ Much of the work with the path integral (1.1) has been done in the context of models where the four-metric is constrained to take a "spatially" homogeneous, 3 + 1 split form where hijdx 'dxJ is the metric on a compact homogeneous three-manifold 2, completely determined by the finitely many functions qa(t). With this ansatz, the action takes the form ~=~~(cj~,q~;~)dt+boundar~ terms . ( 1.4) Provided the ansatz is such that the Euler-Lagrange equations obtained from the Lagrangian in (1.4) are equivalent to the full Einstein equations for the ansatz, the issue of evaluating Z reduces to performing a quantum-mechanical path integral of the form In which sense the minisuperspace integral (1.5) could be hoped to reflect the properties of the full path integral is largely an open question. For example, the paths qa(t) contributing to (1.5) are expected to be continuous but nowhere differentiable in t, where the four-metrics g,,, contributing to the full integral (1.1) would not be expect- ed to be even continuous.' Nevertheless, there are prob- lems in the formalism that are shared by the minisuper- space integral and the full integral, and it is often as- sumed that the minisuperspace integral gives a simple arena for studying such problems. A well-known exam- ple is the issue of the contour of integration (see Ref. 8 and the references therein). As the minisuperspace ansatz (1.3) is defined in terms of a 3+ 1 decomposition of the metric, the minisuper- space path integral is directly applicable only for four- manifolds of the form I X Z, where the interval I may be either open, semiopen, or closed. For example, if the in- terval is closed, the integral can be understood as a prop- agation amplitude between an "initial" and a "final" three-surface, on which the "initial" and "final" data could be chosen to be the intrinsic three-metrics, or the extrinsic curvature tensors, or some combinations thereof. There are, however, physical situations which have prompted minisuperspace path-integral constructions for manifolds which are not globally of the form I X Z. The case of particular interest, both from the viewpoints of black-hole thermodynamics and quantum cosmology, is to take to be compact with a single connected bound- 4032 8 1990 The American Physical Society

Upload: jorma

Post on 03-Apr-2017

214 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Variational principles for nonsmooth metrics

PHYSICAL REVIEW D VOLUME 42, NUMBER 12 15 DECEMBER 1990

Variational principles for nonsmooth metrics

Geoff Hayward* and Jorma ~ o u k o ~ Institute of Theoretical Physics, University of Alberta, Edmonton, Alberta, Canada T6G2JI

(Received 16 May 1990)

We analyze the variational principle of general relativity for two classes of metrics that are not necessarily smooth but for which the Einstein action is still well defined. The allowed singularities in the metric are (1) a jump discontinuity in the extrinsic curvature at a three-dimensional hypersur- face and ( 2 ) a conical singularity in the Euclidean Kantowski-Sachs metric on the manifold ~ x s ~ . In agreement with general expectations, we demonstrate that in both cases the extremizing vacuum metrics are the usual smooth solutions to the Einstein equations, the smoothness conditions coming themselves out of the variational principle as part of the equations of motion. In the presence of a singular matter distribution on a three-dimensional hypersurface, we also demonstrate that the usu- al junction conditions for the metric are directly obtained from the variational principle. We argue that variational principles with nonsmooth Kantowski-Sachs metrics are of interest in view of con- structing minisuperspace path integrals on the manifold 5 XS2, in the context of both black-hole thermodynamics and quantum cosmology. The relation of nonsmooth Kantowski-Sachs variational principles to the issue of Lorentzian versus Euclidean path integrals in quantum gravity is briefly discussed.

I. INTRODUCTION

In the Euclidean path-integral approach to quantum gravity, the object of interest is a path integral of the form

Here gp,, is a Euclidean metric defined on a four- dimensional manifold A, and I(g,,,) is the Einstein ac- tion

I = - --- I I d 4 x ( g +boundary terms . (1.2) 1 6 r I?

The boundary terms in (1.2) depend on M and on the boundary conditions imposed on the metric g,,,. ',* One application of this path-integral formalism is in black- hole thermodynamics, where Z is interpreted as a ther- modynamic partition function.2p4 Another application is in quantum cosmology, where Z is interpreted as the wave function of the ~ n i v e r s e . " ~

Much of the work with the path integral (1.1) has been done in the context of models where the four-metric is constrained to take a "spatially" homogeneous, 3 + 1 split form

where hijdx 'dxJ is the metric on a compact homogeneous three-manifold 2, completely determined by the finitely many functions qa ( t ) . With this ansatz, the action takes the form

~ = ~ ~ ( c j ~ , q ~ ; ~ ) d t + b o u n d a r ~ terms . ( 1.4)

Provided the ansatz is such that the Euler-Lagrange

equations obtained from the Lagrangian in (1.4) are equivalent to the full Einstein equations for the ansatz, the issue of evaluating Z reduces to performing a quantum-mechanical path integral of the form

In which sense the minisuperspace integral (1.5) could be hoped to reflect the properties of the full path integral is largely an open question. For example, the paths q a ( t ) contributing to (1.5) are expected to be continuous but nowhere differentiable in t , where the four-metrics g,,, contributing to the full integral (1.1) would not be expect- ed to be even continuous.' Nevertheless, there are prob- lems in the formalism that are shared by the minisuper- space integral and the full integral, and it is often as- sumed that the minisuperspace integral gives a simple arena for studying such problems. A well-known exam- ple is the issue of the contour of integration (see Ref. 8 and the references therein).

As the minisuperspace ansatz (1.3) is defined in terms of a 3 + 1 decomposition of the metric, the minisuper- space path integral is directly applicable only for four- manifolds of the form I X Z, where the interval I may be either open, semiopen, or closed. For example, if the in- terval is closed, the integral can be understood as a prop- agation amplitude between an "initial" and a "final" three-surface, on which the "initial" and "final" data could be chosen to be the intrinsic three-metrics, or the extrinsic curvature tensors, or some combinations thereof.

There are, however, physical situations which have prompted minisuperspace path-integral constructions for manifolds which are not globally of the form I X Z. The case of particular interest, both from the viewpoints of black-hole thermodynamics and quantum cosmology, is to take to be compact with a single connected bound-

4032 8 1990 The American Physical Society

Page 2: Variational principles for nonsmooth metrics

42 VARIATIONAL PRINCIPLES FOR NONSMOOTH METRICS 4033

ary.1,4-6 In terms of the ansatz (1.3), one would then take the boundary to be the three-surface 2, at (say) the upper limit of the coordinate time, t = t , , and understand the lower limit t =to as occurring at a coordinate singu- larity at the "bottom" of At. To construct the minisuper- space action in this case, one starts from the general in- stein action (1.2) defined on all of &, imposes the symme- try dictated by the ansatz (1.3) in the region that is covered by the coordinate system of the gnsatz, and derives the form of the minisuperspace action (1.4) pay- ing careful attention to the coordinate singularity at t =to . One can then analyze the variational principle as- sociated with the minisuperspace action and use general arguments of consistency to promote the minisuperspace variational principle into a minisuperspace path integral. Analyses of this kind have been given in Refs. 4, 10, and 11.

An important ambiguity remains on the issue of what conditions to impose at the coordinate singularity at t = to . In the analyses of Refs. 4, 10, and 11, one starts by considering metrics that are smooth on all of M , in par- ticular at t = t o , and then passes to the minisuperspace action (1.4). At the level of the minisuperspace path in- tegral (1.51, however, one can no longer expect to main- tain smoothness in the initial conditions. This is because in a 3+ 1 formulated path integral one expects to fix only a limited number of initial data at t = t o , and these data should further be a quantum-mechanically consistent set, in the sense of, for example, not attempting to fix simul- taneously both a coordinate and its conjugate momen- tum. Some possibilities of consistently relaxing the initial conditions at the level of the minisuperspace action have been discussed in Ref. 1 1.

The purpose of this paper is to demonstrate, in a par- ticular model, that one can relax the smoothness condi- tion at t =to to allow conical or perhaps worse singulari- ties already before passing to the minisuperspace action, and still recover a well-defined variational principle. In particular, the variational principle yields smoothness of the extremizing metrics as an equation of motion. The classical solutions emerging from such a variational prin- ciple are thus smooth solutions to the Einstein equations on all of &.

The question of how one might arrive at a well-defined minisuperspace path integral from this kind of a varia- tional principle, without explicitly appealing to smooth- ness in the initial conditions, will be left a subject of fu- ture work. We shall argue in Sec. V, however, that the potential problems to be confronted in this approach are not obviously more severe than in the path-integral con- structions of Refs. 4 and 1 1.

A further interesting consequence of our work relates to the treatment of two-dimensional singularities in gen- eral relativity, such as conical singularities associated with idealized cosmic strings. Motivated by the possibili- ty of distributional matter sources in general relativity, Geroch and ~ r a s c h e n " have raised the question of how singular a metric can be and still have a ~ i e m a n n tensor which is well defined in a distributional sense. They in- troduce a class of metrics, christened "regular" metrics, which are less smooth than C' but satisfy conditions

guaranteeing the existence of a distributionally well- defined Riemann tensor. They go on to prove that no metric in this "regularity" class can have a source con- centrated on a two-dimensional hypersurface. However, the question of whether one might-be able to arrive at a well-defined action and variational principle associated with conically singular metrics was not resolved by their analysis since these metrics do not belong to their "regu- larity" class. Here we show that, in fact, it is possible to define an action and a variational principle for at least some conically singular metrics.

We begin in Sec. 11, after a brief general discussion of singular configurations in a variational principle, by con- sidering the more familiar variational problem for metrics which may have a jump discontinuity in the ex- trinsic curvature across a three-dimensional hypersur- face. We exhibit the Einstein action for such metrics and demonstrate, in the absence of matter singularities, that the variational principle gives the absence of jump discontinuities as "generalized" Einstein equations. W; also show how, in the presence of a singular matter distri- bution on a three-dimensional hypersurface, the standard junction conditions on this hypersurface'3 are directly ob- tained from the variational principle. Although these re- sults have been anticipated in the earlier literature, espe- cially in Refs. 2 and 14, they have not to our knowledge been previously explicitly stated.

In Sec. I11 we turn to the spatially homogeneous min- isuperspace ansatz ( 1.3). We consider metrics defined on the manifold 0 4 1 = D ~ ~ 2 , where is the closed two- dimensional disc, and we take these metrics to satisfy the ansatz (1.3) with S ' X S ~ spatial surfaces (known as the Kantowski-Sachs ansatzI5). This is a situation of interest both in quantum cosmology and in black-hole thermo- dynamics.3t4,11*16 We assume that the metric is smooth everywhere on At except possibly at the "center" s2, which is not covered by the coordinate system of the an- satz. We then derive the minisuperspace action for these metrics, in the special case where the singularity at the center is at worst conical. We demonstrate, in analogy with the extrinsic curvature discontinuities discussed in Sec. 11, that the minisuperspace variational principle gives the absence of a conical singularity as a generalized Einstein equation. In Sec. IV we argue how a similar minisuperspace variational principle ;odd be developed to incorporate a broader class of singularities than just conical ones.

The results are summarized and discussed in Sec. V. We exhibit some of the as vet unresolved issues that would need to be confronted if one wishes to promote our nonsmooth minisuperspace variational principles into genuine minisuperspace path integrals. We also com- ment on the relation of our minisuperspace variational principles to the question of Lorentzian versus Euclidean path integrals in quantum gravity.

11. VARIATIONAL PRINCIPLE FOR DISCONTINUITIES IN THE EXTRINSIC CURVATURE

A variational principle consists of an action functional whose stationary configurations subject to given bound-

Page 3: Variational principles for nonsmooth metrics

4034 GEOFF HAYWARD AND JORMA LOUKO 42

ary conditions are the solutions to the equations of motion subject to the same boundary conditions." In classical mechanics and classical field theory, the equa- tions of motion in general imply that the classical solu- tions will need to have certain regularity properties. When constructing an action principle, it is usually sufficient to define an action functional on configurations that belong to the same regularity class as the classical solutions. For example, to derive Newton's equation for a particle in a one-dimensional smooth (say, c') potential V(x ), it is sufficient to consider the action

on paths x ( t ) belonging to the same smoothness class as the classical solutions.

However, it can occur that an action functional ob- tained in this way can be meaningfully extended to configurations that are considerably less smooth than the classical solutions. The question then arises as to wheth- er the stationary configurations of the extended action are still the same as those of the original action.

For physically reasonable systems the answer would be expected to be positive. As a simple example, consider the action (2.1) with V ( x ) (say) c'. Suppose now that x ( t ) is continuous for O i t i T , and C' for O < t F T / 2 and T/2 5 t i T. The action can then be expressed as the sum

We wish to find the stationary configurations of this ac- tion subject to the boundary data x ( 0 ) = x o , x ( T ) = x I . For simplicity, vary the action first with respect to x ( t ) on 0 < t < T/2 and T /2 < t < T, with fixed x ( T / 2 ) = y , to obtain the ordinary equations of motion in the two re- gions. Then solve these equations and substitute back to the action to obtain

where Sc (xo ,y ;0 ,T /2 ) and S c ( y , x l ; T / 2 , T ) are the ac- tions of the classical solutions with the indicated bound- ary data. The generic path in (2.3) will not solve the clas- sical equations of motion at t =T/2 , and it will indeed not even be C' there. However, extremizing (2.3) with respect to y yields the classical solution for the whole in- terval 0 <_ t I T, and in particular yields the smoothness of the extremizing path at t = T/2 as an equation of motion. The reason is simply that by standard Hamilton-Jacobi theory dSc(xo,y ;O, T/2) /dy is the momentum of the first half of the path at y and dS,(y,z; T/2, T)/dy is the nega- tive of the momentum of the second half at y. This exam- ple is a special case of a discretization procedure which can be used to define path integrals for quantum- mechanical systems with curved configuration spaces (see Ref. 18 and the references therein).

In general relativity, one similarly anticipates that an action functional appropriately extended to nonsmooth

metrics would still have its stationary configurations at the usual solutions to Einstein's equations. Here the situ- ation is, however, considerably more complicated than in the example given above. Since the Ricci scalar is non- linear in the first derivatives of the metric, it becomes a very subtle question as to how singular a metric can be and still have a distributionally well-defined Riemann tensor and a well-defined action.12 In this section we shall consider metrics that have at most a jump discon- tinuity in the extrinsic curvature at a three-dimensional hypersurface. For such metrics the Riemann tensor is well defined as a distribution,12 but the jump discontinui- ties will give a nonvanishing contribution to the gravita- tional action. We shall demonstrate that when these con- tributions are properly taken into account, the situation is very similar to the example considered above.

Consider a general Euclidean line element in the Arnowitt-Deser-Misner (ADM) 3 + 1 formalism:

where N is the lapse function and N' is the shift vector (latin indices extend from 1 to 3). Confine attention to spatially compact manifolds and take the Euclidean time coordinate, t, to extend from a three-surface 2, at t = O to a three-surface Z1 at t =1. We begin assuming that the metric is smooth (for concreteness, C"), and we shall re- lax this assumption later.

With the above representation, the four-curvature sca- lar density takes the form19

RV'; = - N V ' ~ ( K , , K ' ~ - K ~ - ( ~ ' R ) - 2 ( d z ~ ) , ~

Here, Kjj is the extrinsic curvature tensor of a surface of constant 1:

where a vertical bar indicates covariant differentiation with respect to hij. The gravitational scalar density (2.5) integrates to yield the Einstein-Hilbert action

Here, L is a Lagrangian density first order in time deriva- tives:

In the Hamiltonian formulation, one defines a momen- tum conjugate to the "field variable" hij.

In this formulation, the Einstein-Hilbert action becomes

Page 4: Variational principles for nonsmooth metrics

VARIATIONAL PRINCIPLES FOR NONSMOOTH METRICS 4035

As first noticed by ~ o r k , ' the Einstein-Hilbert action (2.10) is appropriate for a variational principle in which the values of n-'i, but not those of hi,, are fixed on 2, and 2,. To obtain an action appropriate for fixing the canoni- cal coordinates hi, on 2, and 2,, one must subtract boundary terms which cancel the boundary terms in (2.10). The resulting action is the York action'^^

Note, however, that these statements about the variation- al principle remain largely formal. In general, little is known about the existence of solutions to the classical boundary-value problem with data fixed on 2, and 2 , . In the context of minisuperspace models, a discussion of this boundary-value problem is given in Ref. 8.

We shall now relax the smoothness assumptions about the metric. Consider metrics which are regular every-

where on our manifold .M except at a surface 2, of con- stant time, t =T. We assume the intrinsic three-metric h,, to be continuous, but we allow the extrinsic curvature K,, to have a finite discontinuity (jump discontinuity) at the surface 2,. We need to evaluate the action for such metrics.

Break the time integral into three components:

+boundary terms

= I , +I2 +I3 +boundary terms , (2.12)

where T* ,lime +,T~E. Since the metric is regular every- where except at T , actions I, and I3 are well defined. For instance,

Meanwhile for I2 one has

In the first and third terms above, the worst irregularities are only jump discontinuities. So only the second term con- tributes:

This is the contribution of the jump discontinuity to the action integral. For definiteness, assume the York action (2.1 1). Substituting the expressions for I,, I,, and Z3, into Eq. (2.12) gives

I = ' - ~ d 'x d t + I L d 'x dr +boundary terms . (2.16) 0

This is the final form of our action, appropriate for metrics with a jump discontinuity in the extrinsic curvature on the surface 2,. We shall now take the action (2.16) as the starting point of a variational principle and derive the corre- sponding equations of motion.

Under variation with respect to the metric components and their derivatives, the action functional (2.16) yields

where greek indices range from 0 to 3 and latin range from 1 to 3. After integration by parts, this becomes

It is vital to note that if the action is to be extremized un- fore wish also here to consider variations for which h,, is der arbitrary variations 6g,, for 0 < t < . r and T+ < t < 1 fixed at these two boundaries but otherwise arbitrary. as well as arbitrary variations 6hi, at t =T, each term in Requiring 61, (2.18) to vanish under such variations (2.18) mus t vanish independently. yields the usual Einstein equations

Recall that the York action (2.11) for smooth metrics is appropriate for keeping hi, fixed on Z, and 2,. We there- GP"=O (2.19)

Page 5: Variational principles for nonsmooth metrics

4036 GEOFF HAYWARD AND JORMA LOUKO - 42

for O < t < r a n d r < t < l , whereasat t = r w e r e c o v e r the equations

Equations (2.20) are the well-known surface Einstein equations for the "vacuum" case.I3

The above treatment easily generalizes to account for the presence of matter. One starts from an action func- tional

By hypothesis, we take the matter to have a regular dis- tribution everywhere except at t =T, where there is an infinitely dense, infinitely thin matter shell. Hence, the matter Lagrangian density is assumed to include a term of the form

Define the intrinsic stress energy tensor of the surface by

. . 1 a S'I= --,- d h ah,

(L$urfv'K 1 .

Then an extremization procedure, analogous to the one performed above, yields the ordinary Einstein equations plus the surface Einstein equations in the presence of a matter shell:13

The standard way to derive the surface Einstein equa- tions is to assume that the ordinary Einstein equations are satisfied across a matter shell and then take the limit- ing case that the shell is infinitely thin. We have shown that they may be derived directly from the variational principle once the contribution of jump discontinuities in the extrinsic curvature have been included in the action.

Before leaving this discussion, it is appropriate to have a closer look at the boundary variations in (2.18). At both 2, and XI extremization of the York action yields the condition

Clearly, setting 6hi, = O at the boundaries (i.e., fixing the intrinsic three-metric there) is sufficient to guarantee that the boundary variations are zero. However, one can im- agine physical situations where the intrinsic three-metric is not fixed on the boundary. This would be the gravita- tional version of "natural boundary conditions" in the variational problem. A familiar example of such natural boundary conditions is a flag whose end is allowed to fly freely in the breeze;20 another example is provided by the open relativistic string.21 If now arbitrary variations of hi, are allowed at a boundary then a condition for ex- tremizing the gravitational action is

at the boundary where hij is not fixed. This means that

not fixing hi, at a boundary forces the extrinsic curvature to vanish there. It might be of interest to study whether such "natural boundary conditions" would be relevant for path-integral constructions in the context of quantum cosmology and black-hole thermodynamics.

111. KANTOWSKI-SACHS ANSATZ ON X S WITH CONICAL SINGULARITIES

We shall now consider metrics defined on the manifold & = B X s 2 , where B is the closed two-dimensional disc. We take the metrics to obey the Kantowski-Sachs an- satzI5

where t is periodic with period 27; and d f l : is the line ele- ment for the unit two-sphere. The Euclidean time coor- dinate t is interpreted as the radial coordinate on the disc. We assume the metrics to be smooth (for concreteness, C m ) everywhere on J22 except possibly at the "center" s 2 , which is not covered by the coordinate system of the an- satz. Further, we assume the possible singularities at the center to be so mild that S u M d 4 x ( g ) 1 ' 2 ~ , and hence the Einstein action, is still well defined.

We shall consider the York action (2.1 11, which is ap- propriate for fixing the intrinsic three-metric on the boundary a d H . This is directly the action relevant for the no-boundary proposal in quantum cosmology.5~6 In black-hole thermodynamics, one would usually consider an action with the additional boundary term

which normalizes the action and the black-hole energy to agree with the conventional results in asymptotically flat space.2f3314 This boundary term is, however, a function of the intrinsic boundary three-metric only, and its presence will therefore not affect our discussion of the variational principle with fixed boundary three-metric.

We need to evaluate the action (2.1 1) for our metrics and to examine the resulting variational principle. Without loss of generality, we can take t = O to corre- spond to the s2 "center" of &, and t = 1 to correspond to the three-surface 2 , at a h . For 0 < t I I , the coordinate system of the ansatz is regular, and we can insert the an- satz into (2.1 I ) , integrate over the three-surfaces, and per- form a standard integration by parts to eliminate the second time d e r i ~ a t i v e s . ~ ~ The action can thus be written as

where the Lagrangian L is given by

Page 6: Variational principles for nonsmooth metrics

42 - VARIATIONAL PRINCIPLES FOR NONSMOOTH METRICS 4037

Let us first recall what happens if we take the metrics to be smooth on all of A.4810311,16 In this case the func- tions a (t), b (t), and N ( t ) must at t -0 satisfy the condi- tions

as well as further conditions on the higher t derivatives. Also, the last term in (3.3) will vanish. Using (3.3, the action can be written as

where the quantities at t = O are understood as limits as t -+O.

When interpreting this action, it is important to bear in mind that the coordinate singularity at t = O is not a boundary of A. So while the term --ab2Ir=, in (3.6) may resemble a "boundary term," it, in fact, is not. We now coin the expression "limit term" which will refer to terms in the 3+ 1 decomposed action and its variation which are evaluated either at t = O or at the boundary at t = l .

Varying the action (3.6) yields

where SL /6a and 6L /Sb are the usual Euler-Lagrange variational derivatives:

The limit terms at t = O in (3.7) vanish by virtue of the as- sumption of smoothness, as is seen from (3.5). If the boundary three-metric is taken to be fixed, the limit terms at t = 1 vanish as well. Extremization of the action in this class of metrics therefore gives the standard Einstein equations

for t > 0, and our regularity assumptions guarantee that the solutions to (3.9) can be extended to smooth solutions on all of A . We therefore see that the minisuperspace ac- tion (3.6), together with the regularity assumptions at t+O, gives a minisuperspace variational principle ap- propriate for the manifold JN with the intrinsic three- metric fixed on a&. The classical solutions are well known to be part of the Euclidean Schwarzschild solu- tion, with mass determined by the values of the boundary

scale factor^.^,^','^ Two points here should be emphasized. First, al-

though the action (3.6) was written in a 3+ l form, the boundary conditions for the metrics have been specified in an intrinsically four-dimensional way. In particular, the assumption of four-dimensional smoothness at the coordinate singularity at t = O implies all the conditions (3.51, as well as further conditions for the higher t deriva- tives. No question of counting "independent" pieces of initial data has arisen. This question would only arise at the next step, when attempting to give a 3 + 1 formulated variational principle with initial data that could be used in a minisuperspace path integral of the form (1.5).10"'

Second, the limit term at t = O in the action (3.6) could have been written in a number of different forms which are all equivalent under the assumption of smoothness. For example, the term could have been written as

Depending what form is chosen for this term, the limit terms at t = O in SI can take a number of superficially different forms, which nevertheless are all equivalent and vanishing under the smoothness assumptions. Again, the question of choosing between the different forms of the limit term in (3.6) would only arise at the level of a 3+ 1 formulated variational principle.']

We now turn to metrics which are not necessarily smooth at the center of M . For such metrics the condi- tions (3.5) need not necessarily hold, nor need the last term in the action (3.3).

necessarily vanish. We wish to evaluate (3.1 1 ) for such metrics and find the corresponding minisuperspace varia- tional principle.

In the rest of this section we shall assume that the singularity at t-0 is at worst conical. More precisely, we assume that the metric can be written as

where

such that f and b are positive-valued C" functions of x and y, f (0,O) = 1, and - w < a < 1. Here the coordinates ( x , y ) cover all of the disc, the center being at x = y =O. The value of the deficit angle at the conical singularity is 2 a a .

Computation of (3.11) is now a standard result. Inter- preting ( g ) ' ' 2 ~ as a distribution-valued density on ,,212, we have explicitly

Page 7: Variational principles for nonsmooth metrics

GEOFF HAYWARD AND JORMA LOUKO

since

and all the other terms under the integral in (3.14) have a vanishing limit as q-+0+. The result (3.14) is just the product of 4 r a , coming from the conical singularity on the disc, and 4 a b 2 ( 0 ) , coming from the volume of the s2.

For x ~ + ~ ~ > O , it is straightforward to construct the transformation from ( x , y ) to the coordinates of the min- isuperspace ansatz. For the functions a ( t ) , b ( t ) , and N ( t ) , our conical singularity assumptions (3.12) and (3.13) imply

and

as t -0, as well as further conditions on the higher time derivatives. Using these conditions and the result (3.141, we now see that the minisuperspace action (3.3) takes a form identical to (3.6). The quantities at t =O are again understood as limits as t -0.

Let us now vary the minisuperspace action (3.61, keep- ing fixed the values of a and b at the boundary t = 1 , re- quiring the variables to obey the conical singularity con- ditions as t -0, but not fixing the actual value of the deficit angle at the conical singularity. The variation is again given by (3.7). Recall that if the action is to be ex- tremized under arbitrary variations, each term in (3.7) must vanish independently. The limit terms at t = l again vanish by the assumption of fixing the boundary values of a and b. Similarly, the second limit term at t =O vanishes by the equations (3.16) which are implied by our conical singularity assumptions. However, the first limit term at t = O is not identically vanishing, since the conical singularity conditions (with unspecified deficit angle) do not fix the limiting values of b and a / N as t -0.

Requiring that the action be stationary under arbitrary variations ha, 66, and 6 N satisfying our boundary condi- tions, we now obtain the usual Einstein equations (3.9) for t > 0 , as well as the equation

Taken together with the conical singularity conditions we imposed at t =0, the condition (3.18) guarantees that the solutions to the Einstein equations (3.9) for t > 0 can be extended to sinooth solutions of the Einstein equations over all of JtZ. These solutions are the same Euclidean Schwarzschild solutions that are recovered from the vari- ational principle in which one initially restricts the varia- tions to smooth metrics on all of &. The difference be- tween the two variational principles is only in the class of the metrics included in the variations: with our conical singularity conditions the smoothness at t = O is not put in as an assumption, but it emerges from varying the ac- tion as an equation of motion.

In the above discussion we have understood the conical singularity conditions at t -0 to be those following from the metric (3.12) and (3.13). We now note that the condi- tions in the minisuperspace variational principle at t =O can in fact be relaxed to consist only of the set (3.16). For consider the action (3.6) as defined on functions of t , as- suming only that a, b, ci / N , and b / N have finite limits as t -0 and that the conditions (3.16) are satisfied. The variation of the action is given by (3.7). Requiring that the action be stationary under variations for which a and b are fixed at t = 1 gives again the Einstein equations for (3.9) for t > 0 and the condition (3.18) at t =O. Combined with (3.16), this is sufficient to guarantee that the classical solutions are extendible to smooth solutions of the Ein- stein equations on all of A.

We summarize. The action (3.61, defined on smooth functions of t E ( O , l ] such that a , b, a / N , and b / N have finite limits as t -0, subject to conditions

a, b fixed at t = 1 , (3.19)

gives a minisuperspace variational principle whose ex- tremizing metrics are smooth solutions to the Einstein equations on A. The conditions of (3.20), with the Ein- stein equations for t > 0 , do not by themselves guarantee smoothness of the solutions of all of ,4l. However, varia- tion of the action gives, in addition to the usual Einstein equations for t > 0, one more condition (3.18) as t -0, and this will be sufficient to guarantee smoothness of the solutions on all of A.

It is worth noting that this variational principle is to some extent analogous to the "natural boundary condi- tions" discussed at the end of Sec. 11. In both cases the variation of the action contains, in addition to the in- tegrals of the usual Euler-Lagrange terms, also a limit term. A variation of this kind is nevertheless well

Page 8: Variational principles for nonsmooth metrics

42 VARIATIONAL PRINCIPLES 1 :OR NONSMOOTH METRICS 4039

defined. The situation is similar to the more familiar variational principles for a flag whose end is allowed to fly freelyZ0 and for the open relativistic string.''

IV. KANTOWSKI-SACHS ANSATZ ON 5 xSZ: MORE GENERAL SINGULARITIES?

In the preceding section we restricted the singularities in the Kantowski-Sachs metric at the center of Jlit to be at worst conical. For discussing variational principles based on the Einstein action this restriction is unnecessarily strong, since one only needs the singularity to be so mild that , d 4 ~ ( g ) " 2 ~ is still well defined. For example, in (3.12) the smoothness conditions for b ( x , y ) at x = y = O could be to some extent relaxed.

It might be a problem of interest to give an exhaustive classification of the sufficiently mild singularities and the corresponding variational principles with the Kantowski-Sachs ansatz on DXS'. It might also be of interest to investigate whether such singularities could in some sense be understood as limiting cases of metrics that have a jump discontinuity in the extrinsic curvature at a three-surface of constant t .

With a conical singularity metric, a limiting procedure of this kind can be given as follows. For a given number E satisfying 0 < E < 1 , consider a continuous metric g , which coincides with the prescribed conical singularity metric for t 1 E , has a jump discontinuity in the extrinsic curvature at the surface t =E, and is smooth for t 5~ with b ( t ) = b ( ~ ) and a ( t ) = S ; N ( t ' ) d t f . At the limit E+O, the action of g , is easily seen to approach the ac- tion of the prescribed conical singularity metric. In par- ticular, the contribution (3.11) from the conical singulari- ty is obtained as the limit 6-0 of the contribution from the surface t = E ,

1 7r d -J _ ( K + - K - )V'Z d3x=- - ( 8~ t - t N dt 1 =c+

where K - and K + are the limiting forms of the extrinsic curvature scalar on the surface t = E when approaching this surface, respectively, from t < E and t > E . If an analogous limiting procedure can be justified for singular- ities worse than conical, comparison of (3.3) and (4.1) suggests that our action (3.6) may be the appropriate minisuperspace action also for more general singularities than just conical ones.

We shall not attempt to develop the above ideas to a more precise level. Recall that in Sec. I11 we first intro- duced the minisuperspace variational principle with coni- cal singularities in a formulation where the conical singu- larity conditions were given in terms of the four- dimensional metric (3.12). These conical singularity con- ditions included, but were not restricted to, the condi- tions (3.16). In the end, however, we were able to give a minisuperspace variational principle with weaker initial conditions consisting just of (3.16). For analyzing the consistency of the minisuperspace variational principle in

its own right, there was no need to establish a direct con- nection between the full action (2.1 1 ) and the minisuper- space action (3.6) under the weaker initial conditions (3.16): it was sufficient to notice that these actions coin- cided for the classical solutions that came out of the min- isuperspace variational principle. We shall therefore not attempt to give a four-dimensional analysis of singulari- ties more general than conical. Rather, motivated by the considerations in the preceding paragraph, we shall now just start from the minisuperspace action (3.61, and inves- tigate the variational principle given by this action when the initial conditions used in Sec. I11 for the functions a ( t ) , b ( t ) , and N ( t ) are relaxed.

Consider thus the minisuperspace action (3.6) defined on functions o f t such that a , b , a / N , and b / N have finite limits as t-+0, but the values of these limits are not specified. The variation of this action is again given by (3.7). Requiring the action to be stationary under varia- tions which keep a and b fixed at t = 1 , we recover now the standard Einstein equations (3.9) for t >O, as well as the conditions

I t is straightforward to verify that the conditions (4.21, combined to the Einstein equations for t > 0, are indeed sufficient to guarantee that the classical solutions can be extended to smooth solutions on all of ,42.

We have thus shown that the action (3.6) gives a min- isuperspace variational principle appropriate for solu- tions defined on all of A , even when the boundary condi- tions in the variational principle only consists of specify- ing the final values of a and b at t = 1 but "nothing" at t =O. The conditions at t = O that are necessary to make the solutions extendible to all of Jlit will themselves come out of the variational principle as equations of motion. One might regard this minisuperspace variational princi- ple as being the one most closely analogous to the varia- tional principle appropriate for the manifold JM, in the full theory.

V. DISCUSSION

In this paper we have analyzed the variational princi- ple of general relativity for two classes of metrics that are not necessarily smooth but for which the Einstein action can still be defined in an unambiguous way. These were metrics with ( 1 ) a jump discontinuity in the extrinsic cur- vature at a three-dimensional hypersurface, and (2) a con- ical singularity occurring in Euclidean Kantowski-Sachs metrics on the manifold X S ~ . In both cases we demon- strated that the vacuum variational principle gives, in ad- dition to the usual Einstein equations, the smoothness of the extremizing metrics as part of the equations of motion. This means that the classical solutions are the same smooth metrics that would be obtained when the variational principle is initially defined only for smooth metrics. With the former class of metrics, we also

Page 9: Variational principles for nonsmooth metrics

4040 GEOFF HAYWARD AND JORMA LOUKO 42

showed that in the presence of singular matter distribu- tion on a three-dimensional hypersurface the usual junc- tion conditions on this hypersurface are directly recovered from our variational principle as equations of motion.

At the purely classical level, our variational principles d o not contain anything surprising. A variational princi- ple in general relativity consists of an action functional whose stationary configurations subject to given bound- ary conditions are the solutions to the Einstein equations subject to the same boundary conditions. In the vacuum theory the classical solutions of interest are smooth metrics, and in the classical variational principle it is then sufficient to take also the neighboring, nonclassical metrics to be smooth. If the action functional obtained in this way can be extended also to nonsmooth metrics, such that neither the four-manifold on which the metrics live nor the boundary conditions for these metrics are changed, it is then expected that the stationary configurations of the extended action be still those of the old action: the extended variational principle should give the smoothness of the extremizing metrics as "general- ized" Einstein equations. Similarly, if we introduce matter with a singular Lagrangian but a well-defined ac- tion, a carefully defined total action would be expected to lead to the appropriate "generalized" Einstein equations at the singular matter source, provided these equations exist in some suitable distributional sense.I2 The varia- tional principles presented in this paper are just special cases of this construction, under specific choices for the potential singularities.

Our minisuperspace variational principles become more interesting when viewed as a starting point for con- structing a minisuperspace path integral. A path integral must in general be defined by a careful regularization pro- cedure, and the contribufing paths will usually not be smooth even when the classical variational principle is in- itially defined for smooth paths. For example, the paths q "( t ) contributing to a quantum-mechanical path integral of the type (1.5) are expected to be continuous but no- where differentiable in t. When we now wish to under- stand a minisuperspace path integral of the type (1.5) as a sum over metrics on a manifold which does not admit a global 3 + 1 decomposition, the smoothness of the metrics becomes an issue already one step before the actual regu- larization of the integral in (1.51, namely, at the stage of choosing the end-point conditions in this integral at the upper and lower limits of t. As the quantum-mechanical end-point conditions should be consistent with the boundary conditions of the classical variational principle, it is thus of importance to choose the classical minisuper- space variational principle to correspond to an appropri- ate smoothness class of metrics. The choice of this ap- propriate smoothness class would, in the end, have to be justified by establishing a connection between the minisu- perspace path integral and the path integral of the full theory.

With Kantowski-Sachs metrics on the manifold M = 5 x S 2 , we showed explicitly how the choice of the smoothness assumptions at the "center" of .M is reflected in the "initial" conditions of the minisuperspace varia-

tional principle. Our analysis thus complements those given in Refs. 4, 10, and 11, where the metrics in the vari- ational principle were first taken to be smooth on all of M and the smoothness conditions were relaxed only at the level of the minisuperspace action.

To follow Refs. 4 and 11 and to promote our minisu- perspace variational principles into genuine path in- tegrals, one would need to give a detailed definition of the path measure, including a specification of a convergent contour of integration. We shall not attempt to embark on this task here, but we would nevertheless like to end by briefly discussing some issues that could arise in this context.

For definiteness, consider the variational principle presented in Sec. IV: the action is given by (3.6), and the fixed quantities in the variational principle are the values of a and b at t = 1 but "nothing" at t =O. The first ques- tion in the path integral would then be how to implement the initial "nothing." One way to proceed is to notice from the limit term in (3.7) that the action (3.6) gives a generically well-defined variational principle for metrics on the manifold [0, 1 ] XS' x S 2 , provided the fixed quan- tities both at t = O and t = 1 are the values of a and b. It should therefore be possible to construct with this action a path integral between fixed initial a,, b, and final a , , b , . The path integral for D x S 2 , fixing "nothing" at the center, could then be obtained by integrating over all choices of the initial data.

where p could be chosen to depend on a, and 6,. To proceed from here, there are at least two different routes. One could start by defining first the B N B a B b path in- tegral on the right-hand side of (5.1), and only after that address the two ordinary integrals over a. and b,. The ordinary integrals would then formally amount to a (gen- eralized) Laplace-Fourier transform.23 Alternatively, one could first take one or both of the ordinary integrals un- der the path integral, for example, by giving an explicit discretization, and address the definition of the path in- tegral only after that. One would perhaps hope that ei- ther method should in the end lead to the same answer; however, to substantiate this hope one would need to give precise definitions of each of the respective path integrals, including the contours of integration. One would in par- ticular need to specify the contours for the Laplace- Fourier transformations in a way compatible with the contours for the path integrals.

Finally, we note that our minisuperspace variational principles and path integrals, especially (5.1), bear a superficial similarity to path integrals that have been ad- vocated by a number of workers from purely Lorentzian c o n ~ i d e r a t i o n s . ~ ~ ~ ~ It should be emphasized that our starting point in the minisuperspace analysis was essen- tially topological: the manifold M =B x S 2 . Although we formulated the variational problem in the Euclidean signature, we can relax this by taking the functions a ( t ), b ( t ) , and N ( t ) to be complex valued. In fact, in the clas-

Page 10: Variational principles for nonsmooth metrics

VARIATIONAL PRINCIPLES FOR NONSMOOTH METRICS

sical boundary-value problem the signature of t h e solu- tions cannot be specified as part of the problem, but t h e signature will come ou t a s part of the solution. F o r cer- tain (real) values of t h e boundary scale factors the only solutions t o o u r boundary-value problem are indeed nei- ther Euclidean nor Lorentzian bu t genuinely complex- valued metric^.^,""^ T o what extent the resulting pa th integrals, such as (5.1), can be thought of as being Lorentzian, falls then under t h e question of specifying the contour of integration. Wi th pa th integrals constructed from a minisuperspace action initially defined for metrics

that a r e smooth on all of A, this question has been dis- cussed in Ref. 1 1.

ACKNOWLEDGMENTS

W e are grateful t o David Brown, Jonathan Halliwell, Werner Israel, Claus Kiefer, Bernard Whiting, and J im York for valuable discussions and comments. This work was supported in par t by T h e Natura l Sciences a n d En- gineering Research Council of Canada. G.H. is indebted t o the Alberta Heritage foundation for financial support .

'Electronic address: ghay@ualtamts. Address after January 1, 1991: Department of Physics, University of British Colum- bia, Vancouver, B.C., Canada V6T 2A6.

t~lectronic address: jlouko@ualtavm. 'J. W. York, Jr., Phys. Rev. Lett. 28, 1082 (1972). IG. W. Gibbons and S. W. Hawking, Phys. Rev. D 15, 2752

(1977). 3J. W. York, Jr., Phys. Rev. D 33, 2092 (1986). %. F. Whiting and J. W. York, Jr., Phys. Rev. Lett. 61, 1336

(1988). 5 ~ . W. Hawking, in Astrophysical Cosmology, proceedings of the

Study Week on Cosmology and Fundamental Physics, edited by H. A. Briick, G. V. Coyne, and M. S. Longair (Pontificiae Academiae Scientiarum Scripta Varia, Vatican City, 1982).

6 ~ . B. Hartle and S. W. Hawking, Phys. Rev. D 28,2960 (1983). 'L. S. Schulman, Techniques and Applications of Path Integra-

tion (Wiley, New York, 19811. 'J. J. Halliwell and J. B. Hartle, Phys. Rev. D 41, 1815 (1990). 9 ~ . Teitelboim, Phys. Rev. D 25, 3159 (1982); 28, 297 (1983). 'OJ. Louko, Phys. Lett. B 202, 201 (1988). "J. J. Halliwell and J. Louko, preceding paper, Phys. Rev. D

42, 3997 (1990). 12R. Geroch and J. Traschen, Phys. Rev. D 36, 1017 (1987). 13K. Lanczos, Ann. Phys. (Leipzig) 74, 518 (1924); W. Israel,

Nuovo Cimento 44B, 1 (1966). I4S. W. Hawking, in General Relativity: An Einstein Centenary

Survey, edited by S. W. Hawking and W. Israel (Cambridge University Press, Cambridge, England, 1979).

- - -

I5R. Kantowski and R. K. Sachs, J. Math. Phys. 7, 443 (1966); M. P. Ryan and L. C. Shepley, Homogeneous Relativistic Cosmologies (Princeton University Press, Princeton, 1975).

1 6 ~ . Laflamme, Ph.D. thesis, University of Cambridge, 1988. "C. Lanczos, The Variational Principles of Mechanics (Universi-

ty of Toronto Press, Toronto, 1966). IRK. V. Kuchai, J. Math. Phys. 24,2122 (1983). 19B. S. DeWitt, Phys. Rev. 160, 11 13 11967). 20C. Lanczos, The Variational Principles of Mechanics (Ref. 17),

p. 70. 2lM. B. Green, J. H . Schwarz, and E. Witten, Superstring

Theory (Cambridge University Press, Cambridge, England, 1987), Vol. I, Sec. 2.1.

22Note that on a manifold with topology D X S ~ subject to the above assumptions, t = O is not a boundary of the manifold so no "boundary term" is removed at the lower limit of the t in- tegral.

23H. W. Braden, B. F. Whiting, and J. W. York, Jr., Phys. Rev. D 36, 3614 11987).

24A. Linde, Zh. Eksp. Teor. Fiz. 87, 369 (1984) [Sov. Phys. JETP 60, 211 (1984)l; Nuovo Cimento 39, 401 (1984); A. Vilenkin, Phys. Rev. D 30, 509 (1984); J. M. Cline, Phys. Lett. B 224, 53 (1989); M. J. Duncan and L. G. Jensen, Nucl. Phys. B328, 171 (1989); E. Farhi, Phys. Lett. B 219, 403 (1989); A. Strominger, Nucl. Phys. B319, 722 (1988); W.-M. Suen and K. Young, Phys. Rev. D 39, 2201 (1989); J. D. Brown, ibid. 41, 1125 (1990); J. D. Brown and E. A. Martinez, ibid. 42, 1931 (1990).