Ingoing Eddington-Finkelstein Metric of an Evaporating Black Hole



Shohreh Abdolrahimi,† Don N. Page,‡ and Christos Tzounis§ Theoretical Physics Institute, University of Alberta, Edmonton, AB, Canada, T6G 2G7 (Dated: 2016 July 15) We present an approximate time-dependent metric in ingoing Eddington-Finkelstein coordinates for an evaporating black hole as a first-order perturbation of the Schwarzschild metric, using the linearized back reaction from a realistic approximation to the stress-energy tensor for the Hawking radiation in the Unruh quantum state.

arXiv:1607.05280v1 [hep-th] 18 Jul 2016

I.

INTRODUCTION

The physics of black holes is an abundant field in which the convergence of gravitation, quantum theory, and thermodynamics takes place. The original derivation of Hawking radiation from black holes is based on semi-classical effective field theory. Normally, quantum fields are considered test fields in the curved spacetime of a classical background geometry. A quantum field theory constructed on a curved background spacetime experiences gravitationally induced vacuum polarization and/or particle creation. These effects induce a nonzero expectation value for the stressenergy tensor. The calculations of the renormalized expectation value of the complete quantum stress-energy tensor outside the classical event horizon were performed using a framework established by Christensen and Fulling [1]. In this framework, the assumptions are that the stress-energy tensor is time-independent, satisfies local stress-energy conservation, and has a trace determined solely by the conformal anomaly, both for the fields that are classically conformally invariant, such as a massless scalar field and the electromagnetic field, and for the gravitational field. The quantum state considered is usually either the Hartle-Hawking state or the Unruh state. In the Hartle-Hawking state, one has thermal equilibrium, and zero net energy flux, with the outgoing Hawking radiation balanced by incoming radiation from an external heat bath at the Hawking temperature. In [2] a fairly good closed-form approximation for the energy density and stresses of a conformal scalar field in the Hartle-Hawking state everywhere outside a static black hole can be found. In the Unruh state, there is the absence of incoming radiation at both past null infinity and the past horizon, plus regularity of the stress-energy tensor on the future event horizon in the frame of a freely falling observer, representing a black hole formed from gravitational collapse, with nothing falling into the black hole thereafter. Matt Visser and others have constructed the quantum stress-energy tensor corresponding to a conformally massless scalar field in the Unruh state [3] in the Schwarzschild metric [4–6]. One of the important questions that one wants to answer concerns the effect of quantized matter on the geometry of black holes. Such effects in the Hartle-Hawking state have been studied [7] (for similar work see [8–10]), using the approximation found in [2] for the expectation value of the renormalized thermal equilibrium stress-energy tensor of a free conformal scalar field in a Schwarzschild black hole background as the source in the semiclassical Einstein equation. The back-reaction and new equilibrium metric are found perturbatively to first order in ~. The new metric is not asymptotically flat unless the system is enclosed by a reflecting wall. The nature of the modified black hole spacetime was explored in subsequent work [11–13]. In this paper, we construct the first-order back-reaction on the metric, using the expectation value of the quantum stress-energy tensor in the Unruh state as the source in the spherically symmetric Einstein equations. This metric represents to the first-order approximation the metric of an evaporating black hole.

II.

METRIC ANSATZ

To construct a metric using the expectation value of the quantum stress-energy tensor in the Unruh state as a source in the spherically symmetric Einstein equations, we first need to find an appropriate metric ansatz. To do so, we begin from the outgoing Vaidya metric. The outgoing Vaidya metric describes a spherically symmetric spacetime



Alberta Thy 26-16, arXiv: YYMM.NNNNN[hep-th] address: [email protected] ‡ Electronic address: [email protected] § Electronic address: [email protected] † Electronic

2

FIG. 1:

with radially outgoing null radiation. Here, we consider a black hole evaporating by Hawking emission. The outgoing Vaidya metric can be written in outgoing Eddington-Finkelstein coordinates as   2µ(u) ds2 = − 1 − du2 − 2dudr + r2 dΩ2 . (1) r Here u is the retarded or outgoing null time coordinate, and dΩ2 = dθ2 + sin2 θdφ2 . We use Planck units, ~ = c = G = 1. When the black hole mass µ is much larger than the reciprocal of the masses of all massive particles, the Hawking emission is almost entirely into massless particles (e.g., photons and gravitons for astrophysical mass black holes), and the Hawking emission rate is given by µ0 ≡

dµ α =− 2, du µ

(2)

where α is a constant coefficient that has been numerically evaluated to be about α ≈ 3.7474 × 10−5 [18–22]. Then we have 1

µ(u) = (−3αu) 3 .

(3)

We are setting u = 0 at the final evaporation of the black hole, so that u is negative for the part of the spacetime being considered. Moreover, we are assuming that the black hole mass is infinite at negative infinite u, so going back in retarded time, the mass grows indefinitely, rather than having a black hole that forms at some particular time.

3 However, for a black hole that forms of some initial mass M0 , the metric we find should be good for values of u when µ(u) < M0 . For r  2µ, the retarded/outgoing time u can be written in terms of the advanced/ingoing time v and radius r as approximately u ≈ v − 2r − 4µ ln

r . 2µ

(4)

Then µ(u) can be written in terms of v and r as well, since µ3 ≈ −3αv + 6αr + 12αµ ln

r = −3αv + 12αµ 2µ



1 − ln z z

 ,

(5)

where 2µ . r

z≡

(6)

In terms of the zeroth-order solution 1

µ0 (v) ≡ (−3αv) 3 ,

(7)

the solution of the cubic equation (5) for µ with v and z given has the following approximate form for −3αv  1:   1 4α 1 µ(v, z) ≈ (−3αv) 3 + − ln z . (8) 1 (−3αv) 3 z Assuming also that z  1, in the second term we can take z ≈ 2µ0 /r, so µ(r, v) ≈ µ0 +

2αr 4α r ln + µ20 µ0 2µ0 1

− 32

≡ (−3αv) 3 + 2αr (−3αv)

+ 4α (−3αv)

− 31

r

ln

1

.

(9)

2 (−3αv) 3

Next, we transform the outgoing Vaidya metric to ingoing Eddington-Finkelstein coordinates with advanced/ingoing time v and radius r. Let us consider u as a function of v and r again as in equation (4). Then we have



r 4µdr du ≈ dv − 2dr − 4µ0 ln du − + 4µ0 du 2µ r   2µ ψ(v,r) ψ(v,r) du ≈ e dv − 2 1 + e dr , r

(10)

where e−ψ(v,r) ≡ 1 − 4µ0 + 4µ0 ln

r . 2µ

(11) (12)

For µ0  1, or −3αv  1   r 4α eψ(v,r) − 1 ≈ ψ(v, r) ≈ −4µ0 ln − 1 ≈ − 2 (ln z + 1) . 2µ µ Plugging back Eq. (10) into the metric (1), we get        2µ 2ψ(v,r) 2 2µ 2µ ψ(v,r) ds2 ≈ − 1 − e dv + 2eψ(v,r) 2 1 + 1− e − 1 dvdr r r r        2µ ψ(v,r) 2µ 2µ ψ(v,r) + 4 1+ e 1− 1+ 1− e dr2 + r2 dΩ2 . r r r

(13)

(14)

4 For r → ∞, and astrophysical black holes µ0  1, or −3αv  1, we have   2µ 2ψ(v,r) 2 2 e dv + 2eψ(v,r) drdv + r2 dΩ2 . ds ≈ − 1 − r

(15)

Now we modify this metric with corrections going as 1/µ2 with coefficients going mainly as functions of z in order to match the stress-energy tensor of the Hawking radiation. In particular, we make the following ansatz for the metric of the evaporating black hole:   2m(v, z) ˜ ˜ ds2 = −e2ψ(v,z) 1 − dv 2 + 2eψ(v,z) dvdr + r2 dΩ2 , (16) r with 1 ˜ eψ(v,r) − 1 ≈ ψ˜ ≈ 2 [g(z) − 4α ln z] , µ   h(z) m≈µ 1+ 2 , µ

(17) (18)

where µ(v, z) is given by Eq. (8), and r ≡ 2µ/z. For z → 0, we have g(z) and h(z) approaching the constants g0 and h0 respectively, so in this limit of infinitely large r/µ, m → µ + h0 /µ and ψ˜ → ψ + (g0 + 4α)/µ2 . It is only an approximation that ψ˜ and m are functions just of µ and z of this form, but for a large and hence very slowly evaporating black hole, it seems to be a very good approximation. Taking v and z to be the independent coordinates (at fixed angles θ and φ), we can rewrite the metric (16) as      4µ2 h(z) ˜ ˜ ˜ ψ(v,z) 2 2ψ(v,z) r,v dv 2 + 2eψ(v,z) r,z dzdv + 2 dΩ2 ds ≈ − e 1−z 1+ 2 − 2e µ z = −F dv 2 + 2Hdvdz + W dΩ2 ,

(19)

where 8α2 1 − z ln z , z 2 (−3αv) 34 z (−3αv) 1 8α 2 + z − z ln z 2 r,z ≈ − 2 (−3αv) 3 − 3 . 1 z z (−3αv) 3 r,v ≈ −



2 3

+

(20) (21)

In summary, we started with the outgoing Vaidya metric as a first approximation for a spherically symmetric black hole metric evaporating by the emission of massless Hawking radiation, and then we switched to ingoing EddingtonFinkelstein coordinates and introduced the functions g(z) and h(z) to allow us to get a better approximation for the metric of an evaporating black hole with the stress-energy tensor of massless Hawking radiation. 1 For r  −v = µ30 /(3α), which implies that µ ≈ µ0 ≡ (−3αv) 3 , Eq. (18) above implies that at fixed z the mass of the black hole changes as dm α 1 = − 2 + O( 4 ). dv µ µ

(22)

The stress-energy tensor for the spherically symmetric Unruh quantum state |ψi on the spherically symmetric curved background of the Schwarzschild space-time with 2µ ≡ rz constant, ds2 = −(1 − z)dt2 +

dr2 + r2 dΩ2 , 1−z

can be written in the following form in an orthonormal frame:  ρ˜ f˜  f˜ −˜ τ hψ| T µˆνˆ |ψi =  0 0 0 0

0 0 p˜ 0

 0 0 . 0 p˜

(23)

(24)

5 Dimensional analysis shows that for massless fields at fixed z, the dependence of the orthonormal components of the stress-energy tensor on the mass µ of Schwarzschild metric at fixed z goes as µ−4 , so for the slowly evolving metric (19), we shall assume that the stress-energy tensor (24) has approximately the following form:  ρ(z) f (z) 0 0 1 f (z) −τ (z) 0 0  . hψ| T µˆνˆ |ψi = 4  0 0 p(z) 0  µ 0 0 0 p(z) 

(25)

According to Christensen and Fulling [1], the stress-energy tensor in the case of the Schwarzschild spacetime can be decomposed into four separately conserved quantities as hψ| T µˆνˆ |ψi = [Ttrace ]µˆνˆ + [Tpressure ]µˆνˆ + [T+ ]µˆνˆ + [T− ]µˆνˆ ,

(26)

where 2

z −T (z) + 1−z H(z)  0 ≡  0 0



µ4 [Ttrace ]µˆνˆ

1 ≡ 2

H(z)

µ4 [Tpressure ]µˆνˆ

1

Z z

0 1





G(z)

z

1 z 1 ≡ f+  1−z 0 0 

1 z −1 ≡ f− 0 1−z 0 

2

µ4 [T− ]µˆνˆ

 0 0 , 0 0

 0 0 0 z2 0 0  , 1−z G(z) 0 p(z) 0  0 0 p(z)

1 1 0 0

0 0 0 0

−1 1 0 0

 0 0 , 0 0 0 0 0 0

(27)

(28)

 2 3 − 2 p(¯ z )d¯ z, z¯3 z¯

2

µ4 [T+ ]µˆνˆ

0 0 0 0

T (¯ z) d¯ z, z¯2

 z2 2p(z) + 1−z G(z)  0 ≡  0

Z

0 z2 1−z H(z) 0 0

 0 0 , 0 0

(29)

(30)

(31)

(32)

The decompositions (27) and (29) make sense if the integrals G(z) and H(z) converge. Imposing mild integrability constraints on T (z) and p(z) at the horizon, which are satisfied for the Unruh state where T |z=1 and p|z=1 are actually finite, we have 1 T |z=1 (1 − z) + O[(1 − z)2 ], 2 G(z) = −pz=1 (1 − z) + O[(1 − z)2 ].

H(z) =

(33) (34)

This is enough to imply that the two tensors (27) and (29) are individually regular at both the past and future horizon. In Kruskal null coordinates, [T+ ] is singular on the future horizon H + and regular on the past horizon H− . On the other hand, [T− ] is singular on the past horizon H − and regular on the future horizon H+ . The two tensors (31) and (32) correspond to outgoing and ingoing null fluxes, respectively. The constants f+ and f− determine the overall flux. The Unruh state must be regular on the future horizon, so we need f+ = 0. However, such a condition

6 na¨ıvely seems to exclude any outgoing radiation. Nevertheless, we can get outgoing radiation by making f− negative. It is convenient to set f− = −f0 β, where f0 is a positive quantity, and β ≡ 1/(213 32 5π 2 ) ≡ 1/(368 640π 2 ). In what follows, we define f (z) = βf0

z2 . 1−z

(35)

Thus,

µ4 [T+ ]µˆνˆ + µ4 [T− ]µˆνˆ

  −f (z) f (z) 0 0  f (z) −f (z) 0 0 . ≡ 0 0 0 0 0 0 0 0

(36)

Therefore, we have ˆˆ

µ4 T 00 = ρ(z) = 2p(z) + ˆˆ

z2 (H(z) + G(z)) − T (z) − f (z) , 1−z

ˆˆ

µ4 T 10 = T 01 = f (z) , z2 ˆˆ µ4 T 11 = −τ (z) = (H(z) + G(z)) − f (z) , 1−z ˆˆ

(37) (38) (39)

ˆˆ

µ4 T 22 = µ4 T 33 = p(z) .

(40)

Since the stress-energy tensor is given in the orthonormal frame, we rewrite Einstein equations in the orthonormal frame, i.e., Gµˆνˆ = 8πTµˆνˆ .

(41)

To find the approximate time-dependent metric for an evaporating black hole as a first-order perturbation of the Schwarzschild metric, using the linearized back reaction from stress-energy tensor (37)-(40) of the Hawking radiation in the Unruh quantum state in the Schwarzschild spacetime, we solve (41) up to relative corrections of the order of 1/µ2 in the Planck units that we are using. To bring Gµν to the orthonormal frame, note that we have − (ω 0 )2 + (ω 1 )2 + (ω 2 )2 + (ω 3 )2 = −F dv 2 + 2Hdvdz + W dΩ2 ,

(42)

where √

H F dv − √ dz, F H ˆ 1 ω = √ dz, F √ √ ˆ ˆ 2 ω = W dθ, ω 3 = W sin θdφ. ˆ

ω0 =

(43) (44) (45)

We have ˆ

ˆ

ˆ

ˆ

ˆ

ˆ

ˆ

ˆ

ˆ

ˆ

G = Gˆ0ˆ0 ω 0 ω 0 + 2Gˆ0ˆ1 ω 0 ω 1 + Gˆ1ˆ1 ω 1 ω 1 + Gˆ2ˆ2 ω 2 ω 2 + Gˆ3ˆ3 ω 3 ω 3 = Gvv dv 2 + 2Gvz dvdz + Gzz dz 2 + Gθθ dθ2 + Gφφ dφ2 .

(46)

Therefore, we derive Gvv , F Gvz Gvv = + , H F Gzz F Gvz Gvv = +2 + , 2 H H F Gθθ . = W

Gˆ0ˆ0 =

(47)

Gˆ1ˆ0

(48)

Gˆ1ˆ1 Gˆ2ˆ2

(49) (50)

Deriving the Einstein tensor components Gvv , Gvz , Gzz , and Gθθ for the metric (19) and using (47-50), we obtain the Einstein tensor components in the orthonormal frame.

7 III.

SOLUTION

We now solve the Einstein equation (41) to first order in the perturbation of the metric from the Schwarzschild metric, using the stress-energy tensor whose components are proportional to 1/µ4 . From the zero-one component of the Einstein equation, we get f (z) =

αz 2 . 16 π (1 − z)

(51)

Comparing this to Eq.(35), we have f0 = α/(16πβ). Therefore, the Hawking radiation luminosity of the black hole is not only L = −dm/dv ≈ α/µ2 but also 16πf0 β/µ2 . From the zero-zero component, we find ρ(z) = −

   z 4 h,z z2 2α 2z 2 − 1 + z 2 (1 − z)h,z = −f (z) 2z 2 − 1 − . 32 π (1 − z) 32π

(52)

From the one-one component of Einstein equation (41), we get   z2 −2α 1 − 8z + 6z 2 + 2z(1 − z)2 g,z − z 2 (1 − z)h,z 32 π (1 − z)  z 4 h,z 1 3 z (1 − z)g,z − . = f (z) −1 + 8z − 6z 2 + 16π 32π

τ (z) =

(53)

From the two-two component, we get  z3  16α + (2 − 5z) g,z − 2zh,z + 2 z (1 − z)g,zz − z 2 h,zz 64 π  z3  = 4f (z) z (1 − z) + (2 − 5z) g,z − 2zh,z + 2 z (1 − z)g,zz − z 2 h,zz . 64 π

p(z) =

(54)

Now let us suppose that in the stress-energy tensor (37)-(40), the functions p(z) and T (z) are explicitly given. We then solve for the metric functions h(z) and g(z) in terms of p(z) and T (z). From Eq. (51), we have f (z). From Gˆ0ˆ0 = 8π Tˆ0ˆ0 and Gˆ1ˆ1 = 8π Tˆ1ˆ1 , we get  Z  2α(1 − 2z 2 ) 32π ρ(z) h(z) = − dz + C1 z 2 (1 − z) z4 Z Z Z T (z) − 2p(z) (1 + z) H(z) + G(z) dz + 32π dz + 4α dz + C2 , (55) = −32π z 2 (1 − z) z4 z2 Z Z 2α(1 − 4z + 2z 2 )dz τ (z) − ρ(z) g(z) = + 16π dz + C2 z(1 − z)2 z 3 (1 − z) Z Z Z T (z) − 2p(z) dz H(z) + G(z) dz + 16π dz + 4α + C2 . (56) = −32π z(1 − z)2 z 3 (1 − z) z Note that the functions H(z) and G(z) are given in terms of T (z) and p(z) by Eqs. (28) and (30). From (33) and (34), near z = 1, we can write Z Z (2 − z 2 ) (1 + z) h(z)|z→1 = 16π (−2p(z) + T (z)) dz + 4α dz + C1 , (57) z4 z2 Z Z (z + 1) dz g(z)|z→1 = 16π (−2p(z) + T (z)) dz + 4α + C2 , (58) z3 z For any conformally invariant quantum field, the trace of the stress tensor is known exactly and is given by the conformal anomaly. In the Schwarzschild space-time, the trace is T (z), where T (z) = γz 6 = ξβz 6

(59)

with β ≡ 1/(213 32 5π 2 ). Here ξ is a dimensionless coefficient of the trace. For spins 0, 1/2, 1, 3/2, and 2, ξ is respectively 96, 168, −1248, −5592, and 20352 [14]. Eq. (28) then gives H(z) =

γ (1 − z 5 ). 10

(60)

8 Therefore, we can derive h(z) and g(z) using these special forms for T (z) and H(z):     Z  z2 8 16 1 1 2 3 2p(z) + G(z) dz − πγ 2z + z + (4α − πγ) ln z − + C1 , h(z) = −32π − 6z z4 1−z 5 5 z  Z   p(z) 8 G(z) 16 + g(z) = −32π dz − πγ 12z + 9z 2 + 8z 3 + (4α − πγ) ln(z) + C2 2 3 z(1 − z) z (1 − z) 15 5

(61) (62)

A 4-term polynomial is believed to be a good approximation for the function p(z). There is evidence [15] that p(z) starts off at order z 3 , and the anomalous trace introduces a term of z 6 . Therefore, we consider p(z) to be a polynomial of the form p(z) = β(k3 z 3 + k4 z 4 + k5 z 5 + k6 z 6 ).

(63)

Matt Visser [6] has performed a least-squares fit on the transverse pressure data of Ottewill, McLaughlin and Jensen [16] for the case of spin zero for the Unruh state, giving the constants k4 = 26.565, k5 = −59.021, k6 = 38.207 and f0 = 5.349, when k3 = 0. For the case with k3 6= 0 Bardeen has kindly provided us the values of the constant k3 = 0.264, k4 = 25.438, k5 = −57.460, k6 = 37.503 and f0 = 5.319 in private communications. For spin 1 particles we are aware of only one calculation [17], which gives k3 = 0, k4 = −239.06, k5 = −166.54, k6 = −1158 and f0 = 2.439. Using Eq. (63) for p(z) in Eq. (30) gives    2k5 − 3k4 2k6 − 3k5 3k6 k3 2 2 3 4 5 1 − 4z + 3z + k4 (1 − z ) + (1 − z ) + (1 − z ) − (1 − z ) G(z) = β 2 3 4 5  k3 1 − 4z + 3z 2 2k5 − 3k4 = β(1 − z) + k4 (1 + z) + (1 + z + z 2 ) 2 1−z 3  3k6 2k6 − 3k5 2 3 2 3 4 (1 + z + z + z ) − (1 + z + z + z + z ) . + 4 5 (64) Finally, we can write f (z), ρ(z) and τ (z) in the following form, z2 , 1−z −z 2 β  60f0 + 60z 3 k4 − 60z 2 k4 + 75z 4 k5 − 80z 3 k5 + 84k6 z 5 60(1 − z)  −90z 4 k6 + 30z 2 k3 − 6ξ − 54ξz 5 + 6k6 + 5k5 −30k3 + 60ξz 4      3 7 z2β 4 5 − z + k6 z 4 − z −f0 + k4 z 2 (1 − z) + z 3 k5 1−z 3 4 2 5    9 ξ k6 k5 k3 −ξz 4 1 − z + − − + (1 − z 2 ) , 10 10 10 12 2  z2β −6ξ + 6ξz 5 − 36k6 z 5 + 30z 4 k6 − 45z 4 k5 + 40z 3 k5 60(1 − z)  −60z 3 k4 + 60z 2 k4 −90z 2 k3 + 120zk3 + 6k6 + 5k5 −30k3 + 60f0      ξ 5 1 3 2 3 z2β z + − z z 4 k6 + − z z 3 k5 1 − z 10 2 5 3 4   ξ k6 k5 3 1 + (1 − z) z 2 k4 − + + + f0 +k3 2z − z 2 − . 10 10 12 2 2

f (z) = f0 β ρ(z) =

=

τ (z) =

=

(65)

(66)

(67)

We want at asymptotic infinity the stress energy tensor to be that of an outgoing flux of positive radiation, requiring ρ(z) → f (z) asymptotically as z → 0 [1]. Picking up the dominant terms [O(z 2 )] in (65) and (66), we see that f0 =

ξ k3 k5 k6 + − − . 20 4 24 20

(68)

In the case that k3 = 0, this constraint is the same as Eq. (29) in [6]. If now we apply the constraint (68), we can replace one of the constants k3 , k4 , k5 , or k6 . For example, we can write   ξ k6 α k3 k5 = 24 − − + , (69) 20 20 16πβ 4

9 Then equations for ρ(z) and τ (z) can be written as follows:  z2β 10f0 + 9ξz 5 + 30z 4 k6 − 25z 4 ξ + 300z 4 f0 − 75z 4 k3 + 16z 3 ξ − 320z 3 f0 + 80z 3 k3 10(1 − z)  −16z 3 k6 − 10z 3 k4 + 10z 2 k4 − 5z 2 k3 − 14k6 z 5      z2β ξ 3 8 7 2 3 4 2 3 = f0 1 − 32z + 30z − z 16 − 9z + 25z + z k6 3z − − z 1−z 10 5 5   1 13 +z 2 k4 (1 − z) + z 2 k3 8z − − z 2 , 2 2  5 z2β τ (z) = ξz − 6k6 z 5 + 14z 4 k6 − 9z 4 ξ + 180z 4 f0 − 45z 4 k3 + 8z 3 ξ − 160z 3 f0 10(1 − z)  +40z 3 k3 − 8z 3 k6 − 10z 3 k4 + 10z 2 k4 −15z 2 k3 + 20zk3 − 10f0       ξ 4 7 3 4 z2β z 2 − 9z + z 3 + k6 z 3 z − z2 − + f0 16z 3 + 18z 2 − 1 = 1 − z 10 5 5 5 5   9 3 . +k4 z 2 (1 − z) + zk3 2 − z + 4z 2 − z 3 2 2 ρ(z) =

(70)

(71) (72)

Now solving the Einstein equation (41), we obtain 8πβ z [A1 (z)k6 + A2 (z)ξ + A3 (z)f0 − 60k4 +A4 (z)k3 ] + C2 , 15  8πβ  2 z 8z (ξ − k6 ) + D1 (z)f0 −D2 (z)k3 + C1 , g(z) = − 15

h(z) =

(73) (74)

where A1 (z) ≡ −28z 2 + 48z, A2 (z) ≡ 18z 2 − 48z, A3 (z) ≡ 900z − 120, A4 (z) ≡ −225z + 30, D1 (z) ≡ 180z + 240, D2 (z) ≡ 45z + 60.

(75) (76) (77) (78) (79) (80)

Therefore, knowing the value of the trace anomaly coefficient ξ for the appropriate massless field, i.e., scalar, electromagnetic or gravitational, and the constants k3 , k4 , k5 , and k6 for the corresponding field, we have an approximate time-dependent metric for an evaporating black hole as a first-order perturbation of the Schwarzschild metric given by (73)-(80) and (17)-(21) in (v, z) coordinates with µ(v, z) from (8). Alternatively, using the functions (73)-(80), the metric is given by (16)-(18) in (v, r) coordinates with µ(v, r) from (9), where one needs to replace z from z = 2µ/r. We choose values of C1 and C2 such that for z = 1, h(0) = 0 and g(0) = 0, giving 8 πβ(8ξ + 420f0 − 105k3 − 8k6 ), 15 8 C2 = − πβ(−6ξ + 156f0 − 39k3 − 12k4 + 4k6 ). 3 C1 =

(81)

˜ for spin 0 and spin 1 particles. For functions h(z), In Figs. (2)-(4), we have plotted the functions g(z), h(z) and µ2 ψ, 2˜ and µ ψ, the lines for spin 0 with k3 = 0 and k3 6= 0 overlap. For the function g(z), there is a tiny difference between the two cases where k3 = 0 and k3 6= 0, which is shown in figure (2), by the difference between the solid line and the dashed line.

10

FIG. 2: Behaviour of the function g(z) for spin 0 in the case k3 = 0 (dashed line), for spin 0 in the case k3 6= 0 (solid line), and spin 1 (dotted line) fields.

FIG. 3: Behaviour of the function h(z) for spin 0 (dashed line) and spin 1 (solid line) fields.

IV.

COMPARISON OF METRICS

The metric (16) is good for −3αv  1. Next, we want to convert it back to outgoing Eddington-Finkelstein coordinates to see how our metric behaves in the coordinate system (u, r) when µ/r = (−3αu)1/3 /r  1. Consider the metric (16) with functions h and g known from Eqs. (73) and (74), using the relation " # r 1/3 v ≈ u + 2r + 4 (−3αu) ln , (82) 1/3 2 (−3αu)

11

FIG. 4: Behaviour of the function µ2 ψ˜ for spin 0 (dashed line) and spin 1 (solid line) fields.

we get dv = Ψ1 du + Ψ2 dr ,

(83)

where 4α ln Ψ1 = 1 − Ψ2 = 2 +

h

r 2(−3αu)1/3

i

(−3αu)2/3

+

4α , (−3αu)2/3

4(−3αu)1/3 . r

(84) (85)

The metric (16) with functions h and g known from Eqs. (73) and (74) becomes       2m 2m ˜ ˜ ˜ 2 2ψ 2 2 2ψ ψ ds = −e 1− Ψ1 du − 2 e 1− Ψ1 Ψ2 − e Ψ1 dudr r r     2m ˜ ˜ Ψ22 dr2 + r2 dΩ2 + 2eψ Ψ2 − e2ψ 1 − r = guu du2 + 2gur dudr + grr dr2 + r2 dΩ2 .

(86)

    2(−3αu)1/3 2 r 1 4α ln − 8α − h − 2g + O( − ), r (−3αu)1/3 r 2(−3αu)1/3 (−3αu)2/3

(87)

For −3αu  1 we have guu = −1 +

8(−3αu)2/3 1 + O( 2 ) , r2 r −96αu (−3αu)4/3 16(−3αu)2/3 + + O( ). = 2 3 r r r4

gur = −1 +

(88)

grr

(89)

Eqs. (86)-(89) give us the corrections to the Vaidya metric for an evaporating black hole, in outgoing EddingtonFinkelstein coordinates. When we consider only terms of the order of unity and of first order in the small quantity

12 µ/r = (−3αu)1/3 , we get guu = −1 +

2(−3αu)1/3 , r

gur = −1 , grr = 0 . Therefore, for µ/r = (−3αu)1/3  1 we get the outgoing Vaidya metric,   2(−3αu)1/3 2 du2 − 2dudr + r2 dΩ2 . ds = −1 + r V.

(90) (91) (92)

(93)

SUMMARY

In this paper we have constructed an approximate time-dependent metric for an evaporating black hole as a firstorder perturbation of the Schwarzschild metric, using the linearized back reaction from the stress-energy tensor of the Hawking radiation in the Unruh quantum state in the unperturbed spacetime. We used a metric ansatz in ingoing Eddington-Finkelstein coordinates (v, r). Our ansatz is such that at infinity we get the Vaidya metric in the outgoing Eddington-Finkelstein coordinates (u, r). We have solved the corresponding Einstein equations for the metric functions to first order in the stress-energy tensor of the unperturbed Schwarzschild metric. Therefore, our metric is a very good approximation when the mass is large in Planck units.

[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22]

Steven M. Christensen and Stephen A. Fulling, “Trace Anomalies and the Hawking Effect,” Phys. Rev. D 15, 2088 (1977). Don N. Page,“Thermal Stress Tensors in Static Einstein Spaces,” Phys. Rev. D 25, 1499 (1982). William G. Unruh, “Origin of the Particles in Black Hole Evaporation,” Phys. Rev. D 15, 365 (1977). B. P. Jensen, J. McLaughlin and A. C. Ottewill, “Renormalized Electromagnetic Stress Tensor for an Evaporating Black Hole,” Phys. Rev. D 43, 4142 (1991). B. P. Jensen, J. G. McLaughlin, and A. C. Ottewill, “Stress-energy Tensor of an Electromagnetic Field in Schwarzschild Spacetime.” , Phys. Rev. D 55, 809 (1996). Matt Visser, “Gravitational Vacuum Polarization. 4: Energy Conditions in the Unruh Vacuum,” Phys. Rev. D 56, 936 (1997). James W. York, Jr., “Black Hole in Thermal Equilibrium With a Scalar Field: The Back Reaction,” Phys. Rev. D 31, 775 (1985). Paul R. Anderson, William A. Hiscock, Janet Whitesell and James W. York, Jr., “Semiclassical Black Hole in Thermal Equilibrium with a Nonconformal Scalar Field,” Phys. Rev. D 50, 6427 (1994). David Hochberg and Sergei V. Sushkov, “Black Hole in Thermal Equilibrium with a Spin-2 Quantum Field,” Phys. Rev. D 53, 7094 (1996). David Hochberg and Thomas W. Kephart, “Gauge field back reaction on a black hole,” Phys. Rev. D 47, 1465 (1993). David Hochberg, Thomas W. Kephart and James W. York, Jr., “Positivity of Entropy in the Semiclassical Theory of Black Holes and Radiation,” Phys. Rev. D 48, 479 (1993). David Hochberg, Thomas W. Kephart and James W. York, Jr., “Effective Potential of a Black hole in Thermal Equilibrium with Quantum Fields,” Phys. Rev. D 49, 5257 (1994) [gr-qc/9307037]. William A. Hiscock, Shane L. Larson and Paul R. Anderson, “Semiclassical Effects in Black Hole Interiors,” Phys. Rev. D 56, 3571 (1997) [gr-qc/9701004]. Nicholas D. Birrell, Paul C. W. Davies, “Quantum Fields in Curved Space”, Cambridge University Press, (1982). James M. Bardeen (private communication, Feb. 15 2016). Bruce P. Jensen, James McLaughlin and Adrian C. Ottewill, “Renormalized Electromagnetic Stress Tensor for an Evaporating Black Hole,” Phys. Rev. D 43, 4142 (1991). Jerzy Matyjasek, “Stress energy Tensor of an Electromagnetic Field in Schwarzschild Space-time,” Phys. Rev. D 55, 809 (1997). Don N. Page, “Particle Emission Rates from a Black Hole: Massless Particles from an Uncharged, Nonrotating Hole,” Phys. Rev. D 13, 198 (1976). Don N. Page, “Particle Emission Rates from a Black Hole. 2. Massless Particles from a Rotating Hole,” Phys. Rev. D 14, 3260 (1976). Don N. Page, “Particle Emission Rates from a Black Hole. 3. Charged Leptons from a Nonrotating Hole,” Phys. Rev. D 16, 2402 (1977). Don N. Page, “Comment on ‘Entropy Evaporated by a Black Hole’,” Phys. Rev. Lett. 50, 1013 (1983). Don N. Page, “Hawking Radiation and Black Hole Thermodynamics,” New J. Phys. 7, 203 (2005) [hep-th/0409024].