You are on page 1of 9

Back-reaction of the Hawking radiation flux on a gravitationally collapsing star II:

Fireworks instead of firewalls


Laura Mersini-Houghton1, 2 and Harald P. Pfeiffer3, 4
1
DAMTP,University of Cambridge, Wilberforce Rd., Cambridge, CB3 0WA, England
2
Department of Physics and Astronomy, UNC-Chapel Hill, NC 27599, USA
3
Canadian Institute for Theoretical Astrophysics, 60 St. George Street, University of Toronto, Toronto, ON M5S 3H8, Canada
4
Canadian Institute for Advanced Research, 180 Dundas St. West, Toronto, ON M5G 1Z8, Canada
(Dated: September 8, 2014)
A star collapsing gravitationally into a black hole emits a flux of radiation, knowns as Hawking
radiation. When the initial state of a quantum field on the background of the star, is placed
in the Unruh vacuum in the far past, then Hawking radiation corresponds to a flux of positive
energy radiation travelling outwards to future infinity. The evaporation of the collapsing star can
be equivalently described as a negative energy flux of radiation travelling radially inwards towards
the center of the star. Here, we are interested in the evolution of the star during its collapse. Thus
arXiv:1409.1837v1 [hep-th] 5 Sep 2014

we include the backreaction of the negative energy Hawking flux in the interior geometry of the
collapsing star and solve the full 4-dimensional Einstein and hydrodynamical equations numerically.
We find that Hawking radiation emitted just before the star passes through its Schwarzschild radius
slows down the collapse of the star and substantially reduces its mass thus the star bounces before
reaching the horizon. The area radius starts increasing after the bounce. Beyond this point our
program breaks down due to shell crossing. We find that the star stops collapsing at a finite radius
larger than its horizon, turns around and its core explodes. This study provides a more realistic
investigation of the backreaction of Hawking radiation on the collapsing star, that was first presented
in [1].

PACS numbers: 04.25.D-, 04.25.dg, 04.30.-w, 04.30.Db

I. INTRODUCTION ergy Hawking radiation in its interior. Yet, the evolution


of the star could not be followed beyond the bounce with
the approximate analytic methods of [1].
The backreaction of Hawking radiation on the evolu-
tion of the collapsing star is the most important problem Given the fundamental importance of this problem and
in the quantum physics of black holes. This problem the intriguing results of [1], we here aim to study the
provides an arena for the interplay of quantum and grav- backreaction of Hawking radiation on the collapsing star
itational effects on black holes and their respective im- by considering a more realistic setting, namely: we allow
plications for the singularity theorem. A key feature of the star to be inhomogeneous and consider an Hawking
Hawking radiation, which was well established in seminal radiation flux of negative energy which propagates in the
works by [3, 4, 9, 11, 17–20], is that the radiation is pro- interior of the star, with its counterpart of positive en-
duced during the collapse stage of the star prior to black ergy flux travelling outwards to infinity. Hawking radia-
hole formation. The very last photon making it to future tion flux arises when the initial conditions imposed on the
infinity and thus contributing to Hawking radiation, is quantum field on the background of the star, are chosen
produced just before an horizon forms. However its effect to be in the Unruh vacuum state in the far past [9, 18]. In
on the collapse evolution of the star was considered for contrast to the Hartle-Hawking initial state which leads
the first time only recently [1]. As was shown in [1] once to an idealized time symmetric thermal bath of radia-
the backreaction of Hawking radiation is included in the tion present before and after the collapse, the choice of
interior dynamics of the star, then the collapse stops and the Unruh vacuum describes a flux of thermal radiation
the star bounces. Solving analytically for the combined which is zero before the collapse and switches on after
system of a collapsing star with the Hawking radiation the collapse. We solve numerically the full 4 dimensional
included, is quite a challenge. The system studied in [1] set of Einstein and of total energy conservation equations
was idealized in order to obtain an approximate analyti- leading to a complete set of hydrodynamic equations for
cal solution: there the star was taken to be homogeneous; this model. Numerical solutions allows us to follow the
the star’s fluid considered was dust; the star was placed evolution of the collapsing star beyond its bounce.
in a thermal bath of Hawking radiation which arises from The paper is organized as follows: Section II describes
the time-symmetric Hartle-Hawking initial conditions on the metric in the star’s interior, the stress energy tensor
the quantum field in the far past. Within these approx- for the star and for the Hawking radiation flux, which
imations, the main finding of [1] was that a singularity comprise the model we wish to study. We then set up
and an horizon do not form after the star’s collapse be- the system of the evolution equations that need to be
cause the star reverses its collapse and bounces at a finite solve for the combined system. In Sec. III we describe
radius due to the balancing pressure of the negative en- the numerical implementation and the two codes written,
2

and provide a consistency test for the codes by applying


them to the well known Oppenheimer Snyder model. In
Sect. IV we provide the results of the numerical solutions Gab = Ttotal
ab
, (5)
for the evolution of the interior geometry of the star and
conclude in Sect. V. the total energy conservation equations,

ab
∇b Ttotal = 0, ∇b T ab = −∇b τ ab , (6)
II. MODEL

and the baryon number conservation


Let us start with a spherically symmetric and inhomo-
geneous dust star, described by the following metric
∇a (nua ) = 0. (7)

ds2 = −e2Φ(r,t) dt2 + eλ dr2 + R2 dΩ2 (1) The form of the renormalized stress energy tensor for
the Hawking radiation flux in the Unruh vacua, at fu-
where R(r, t) is the areal radius and dΩ2 = dθ2 + ture infinity and near the surface of the star were cal-
sin(θ)2 dφ2 . This form of the metric is convenient for culated in Candelas [18]. Based on energy conservation,
describing a radiating star. The set of Einstein and en- the net inward negative energy flux in the star’s inte-
ergy conservation equations, were originally derived in rior is equal and opposite to its value at infinity, shown
Misner [12]. We will use the set of equations [12] and in [3, 18, 20, 23] who also showed that the Unruh vac-
[10], jointly known as the hydrodynamic equations, by uum best approximates the state that follows gravita-
modifying them to reflect our particular system. At the tional collapse since it such that it is nearly empty in the
onset of collapse the star gradually starts producing a far past before the star collapses, but then produces a
flux of Hawking radiation. The positive energy flux trav- flux of radiation once the collapse starts. On these ba-
els outwards to future infinity and an equivalent but neg- sis the quantum mechanical stress energy tensor of the
ative energy flux travels radially inwards in the interior Hawking radiation flux in the interior is as follows
of the star towards the center. Most of the outgoing
null radiation is produced on the last stages of collapse 1 1 0 0
 
[9, 11, 19, 20]. We wish to include the backreaction of the L −1 −1 0 0
τa b = 2 (8)
negative energy Hawking radiation flux on the interior of R 0 0 0 0
the star and solve for its evolution. 0 0 0 0
The stress energy tensor of the Hawking radiation flux
is given by The total energy conservation is best written as a set of
ab a b two equations which contain the radiative heat transfer
τ = qH k k (2)
from the fluid to the Hawking flux, its t-component and
where k a is an outgoing null vector k a ka = 0 and qH the r-component that are obtained by contracting T ab
is the Hawking radiation energy density related to the in 6 with ua and the unit vector na orthogonal to the
luminosity of radiation by L = 4πR2 qH . 3-hypersurface Σt at t = constant as follows
The fluid of the inhomogenenous dust star with a 4-
velocity ua , normalized such that ua ua = −1, has a stress Qa = ∇b τ ab = (e−Φ nC, e−λ/2 nCr , 0, 0) (9)
energy tensor
From Eqn. 9 and the Hawking flux stress energy tensor in
T ab = ua ub , (3) the interior Eqn. 8, it is immediately clear that Cr = −C.
The latter simply reflect the fact that the star is receiving
where the energy density  is expressed in terms of a spe- negative energy transfer travelling inwards in its interior.
cific inetrnal energy density per baryon e and the number We can now write explicitly the full set of hydrody-
density of baryons n by  = n(1 + e) = nh with h the namic equations, given above. First let us define the
enthalpy. We normalize ka by the condition ka ua = 1, so proper derivates Dr , Dt by the following relation in terms
that qH denotes the magnitude of Hawking energy flux of the metric
density of the inward moving radiation in the rest frame
of the fluid. Therefore the total stress energy tensor of
Dr = e−λ/2 ∂r Dt = e−Φ ∂t . (10)
the combined system of the star’s fluid (Tab )and of the
radiation (τab ) is
The evolutionary equations for baryon number density
ab
Ttotal = T ab + τ ab (4) n(r, t) from 7, specific internal energy e from 6, the en-
ergy of null-radiation with luminosity L given from the
The equations that describe the dynamics of the in- t-component of 9 and 6, the proper velocity of the co-
terior of the star with the backreaction of the Hawking moving fluid U , areal radius of comoving fluid elements
radiation flux included are: the Einstein equations, R(r, t), and the ’acceleration’ equation obtained from the
3

r-component of 9 and the Einstein equations 5 We being the motivation of our choice C(r, t) by not-
ing that for p = 0, Eq. (11b) simplifies to ∂t e = −C eΦ .
1 1 ∂ Therefore, we can achieve a spatially constant e (i.e. spa-
Dt n + 2 (R2 U ) = 4πRqH αf −1/2 , (11a)
n R ∂R   tially uniform mass evaporation) with the choice
1
Dt e = −C − pDt , (11b)
n C(r, t) = e−Φ(r,t) C̄(t), (19)
 
2 −2Φ 2Φ
 ∂U Lα
Dt L = 4πR nC − e Dr Le − 2L − , where C̄(t) is a not yet determined function of time.
∂R Rf 1/2
(11c) To determine C̄(t), we consider Eq. (11c), which relates
the luminosity L to the energy-conversion rate C. If we
Dt R = U, Dr R = f, (11d) disregard all gradient terms and any other term indepen-
∂Φ m L dent of C, then Eq. (11c) becomes ∂t L = 4πR2 eΦ nC.
Dt U = f − + . (11e) This represents the radiation creation rate in a spherical
∂R R2 R
shell at radius R. Integrating over radius with volume
Two auxiliary quantities are given by radial ODEs, element eλ/2 dr, we find a total radiation creation rate of
∂m    Z rs
= 4πR2 ε − qH 1 + αU f −1/2 , (11f)
∂R Ctotal = dr 4πR2 eΦ nC eλ/2 (20)
∂Φ ∂p 0
+ nCf −1/2 ,
Z rS
(ε + p) =− (11g) αR0
∂R ∂R = C̄(t) dr 4πR2 n √ . (21)
0 f
with boundary conditions
If we disregard radiation propagation effects, i.e. assume
m = 0, L = 0, at r = 0, for all t, (12) that the star is unchanged during a light-crossing time,
Φ=0 at r = rstar , for all t. (13) then we expect that LS ≈ Ctotal . Combining this with
Eq. (18), we arrive at
Finally, ε, f , α and R0 are short-cuts for
rS
 p 2 Z αR0
2m LH = C̄(t) US + fS dr 4πR2 n √ , (22)
f = 1 + U2 − , (14) 0 f
R
ε = n(1 + e), (15) from which it follows that
R0
α = 0 = signR0 , (16) LH
Z rS 0 −1

|R | 2 αR
C̄(t) = √ 2 dr 4πR n √ . (23)
∂R US + fS 0 f
R0 = . (17)
∂r

III. NUMERICAL IMPLEMENTATION


A. Choice of energy conversion rate C
The numerical implementation will use coordinates t
The remaining quantities pressure p and the heating and r. t represents the proper time of the fluid-element
transfer rate C are free data, that must be specified on the surface of the star (this identification is enforced
through their functional dependence on the evolved vari- through the boundary condition (13). r is the radial
ables. For the present work, we assume pressure-less coordinate comoving with the fluid. Using r as radial
dust, i.e. p = 0 throughout. In the future we will ex- coordinate in the numerical implementation ensures that
tend the analysis to the case p = w with w 6= 0 the the star always covers the same range r ∈ [0, rstar ]. We
equation of state for the fluid. rewrite Eqs. (11) with partial derivatives according to
Having to specify C is somewhat inconvenient, as our Eq. (10).
goal is a solution of Eqs. (11a)–(11e) for the scenario
where the luminosity at infinity is equal to the expected
Hawking radiation:
A. Discretization
p
L∞ = LH,S (t)(US + fS )2 = LH . (18)
We implement Eqs. (11) in a finite-difference code, us-
where LH,S (t) is the luminosity of the outgoing null radi- ing the Crank-Nicholson method. A uniform grid in co-
ation at the moving surface of the star and LH the Hawk- moving radius r is employed. Spatial derivatives are dis-
ing flux at future infinity. Unfortunately, Eqs. (11a)– cretized with second order accuracy using centered finite-
(11e) do not allow to specify L freely. Rather, L follows difference stencils in the interior and one-sided stencils at
from the choice of the energy-conversion rate C. There- the boundary. Denoting the vector of evolved variables
fore, our task is to choose a suitable function C(r, t). with u = {n, e, L, R, U }, then discretized equations have
4

the form
0.4
mS
u(ri , t + ∆t) − u(ri , t) 1 0.3
= (R[u](ri , t) + R[u](ri , t + ∆t)) . RS/10
∆t 2 0.2
(24)
Here, R represents the right-hand-sides of Eq. (11). R 0.1
involves second order finite-difference stencils (centered 0
-4
in the interior, one-sided at the boundary). When eval- 10 |dtk - dt2|
uating R at the origin, r = R = 0, the rule of L’Hospital 10
-5

is used to evaluate terms that would lead to a division -6


10
by zero, e.g L/R → 0, U/R → ∂r U/∂r R. -7
We solve Eq. (24) separately for each of the variables 10
n, e, L, R, U , fixing all other variables to guesses of their 10
-4
|Nrk - Nr512|
values at the new time. Once all variables have been -5
10
solved for, the auxiliary variables f , m, and Φ are up- -6
dated. Now the updated variables are copied into the 10
-7
guesses, and if necessary the procedure is repeated. We 10
iterate until the maximal norm of the correction is less 0 4 8 12
than 10−10 . This takes typically 3-6 iterations. t
When integrating Eq. (11f), care must be taken to cor-
FIG. 1: Convergence test for the numerical code. The top
rectly represent the m ∝ r3 behavior close to the origin.
panel shows the evolution of areal radius RS and total mass
A second order finite-difference method cannot resolve mS for the run discussed in detail in Sec. IV. This evolution
this cubic behavior of m(r), and would therefore lead to was performed at five different time-steps, dt = 2, 4, 8, 16, 32
a loss of accuracy. Instead, we define the auxiliary quan- (in units of τcollapse /10000), and the middle panel shows the
tity difference between runs at the larger for time-steps and the
smallest time-step. The evolution was also performed at five
3m different radial resolutions (N r = 32, 64, 128, 256, 512), and
η≡ , (25)
4πR3 the lower panel shows the difference between the finest res-
olution and the coarser ones. The thick black lines in the
which is approximately constant close to the origin. We lower two panels are representative of the errors in the runs
rewrite Eq. (11f) as discussed in this paper.

∂η 3 h   i
= ε − q 1 + αU f −1/2 − η , (26)
∂R R i.e. the lines Λ(r, t)=const are null. To achieve this, Λ
must satisfy
and discretize the partial derivative directly in terms of
the areal radius R. After solving for η, we recover m = ∂Λ ∂Λ
4 3 0 = k(Λ) = e−φ + e−λ/2 . (28)
3 πR η. ∂t ∂r
Figure 1 demonstrates the expected second order con-
vergence of this finite-difference code. Solving for ∂Λ/∂t results in an evolution equation for Λ,

∂Λ φ α f ∂Λ
= −e . (29)
B. Diagnostics ∂t R0 ∂r
This is an advection equation with positive advection
Besides monitoring the evolved variables and the aux- speed, i.e. the characteristics of Λ (the null rays) al-
iliary variables, we utilize two additional diagnostics. ways move toward larger values of r. This is an expected
First, we compute the expansion θ(k) of r =const sur- result for a comoving coordinate r. Because Λ is always
faces along the outgoing null-normal k µ = Dt + Dr = advected outward, we must supply a boundary condition
e−Φ ∂t + e−λ/2 ∂r . This quantity is computed as at r = 0. This boundary condition sets the value of Λ for
the null ray starting at (t, 0), and we choose
2  p 
θ(k) = α f +U . (27)
R Λ(t, 0) = t. (30)
When θ(k) = 0, an apparent horizon has formed. Specif- We also need initial conditions at t = 0, where we set
ically, we will plot Θ(k) R/2, a quantity which is unity in
flat space. Λ(0, r) = −r. (31)
Second, we track during the evolution outgoing null
geodesics, to gain a deepened understanding of the causal These initial- and boundary conditions ensure that Λ ≤ 0
structure of the space-time. The null-geodesics are rep- represents null rays that intersect the t = 0 surface,
resented by the level-sets of an auxiliary variable Λ(t, r), whereas Λ ≥ 0 represent null rays that emenate from
5

1.2 RS/R0
5 R(ri, t) 0.8
0.6 θS RS/2
mS
4 0.4
0 nS
US
-0.6 3 0
13.5 13.8 14.1
0.4
1000 LS RS Oppenh.-Sny.
1000 L∞
0.2 1000 LH 2
2mS(t)
0
eS
1
-0.2

-0.4
0
0 5 10 15 0 5 10 15
t t
FIG. 2: Evolution of pressure-free collapse with Hawking FIG. 3: Evolution of pressure-free collapse with Hawking
radiation (R0 = 4, M0 = 0.4). The thick lines indicate the radiation (R0 = 4, M0 = 0.4). Shown are the area radii
evolution of quantities evaluated at the stellar surface: areal at sixteen equally-spaced comoving coordinate radii ri . The
radius RS , radial velocity US , total mass mS , number density thick red line indicates 2mS , which is scaled such that RS =
nS , expansion θS . The lower panel shows the internal energy 2mS would indicate the formation of a black hole. The thin
eS , luminosity at the stellar surface LS , luminosity at infinity dashed blue line indicates the areal radius of a Oppenheimer-
L∞ , and our target luminosity LH . For easy of plotting, the Snyder collapse.
luminosities are scaled by a factor of 1000. The thin dashed
lines in the upper panel represent the respective quantities
in a pure Oppenheimer-Snyder collapse of the same initial 0.4 U(ri, t)
conditions.

the origin at coordinate time t = Λ. The boundary 0


conditions (30) and (31) are consistent with the evolu-
tion equation (29), because at the origin φ(0, 0) = 0,
f (0, 0) = α(0, 0) = R0 (0, 0) = 1. Equation (29) is added
-0.4
as an extra evolution equation to Eqs. (11).

IV. RESULTS -0.8


0 4 8 12
We will investigate stars with different initial radius t
R0 and with different initial masses M0 . Figure 2
shows a rather typical outcome, a star with initial com- FIG. 4: Radial velocity U for the case R0 = 4, M0 = 0.4.
pactness R0 /M0 = 10. The evolution proceeds first Shown are the area radii at sixteen equally-spaced comoving
through a phase which is almost identical to standard coordinate radii ri , with the blue thick lines indicating the
Oppenheimer-Snyder collapse: The star collapses with stellar surface, r = rS and the radius r = rS /2.
increasing inward velocity. The luminosity is very small
in this early phase, resulting in a negligible change in
internal energy e and mass ms . The upper panel of verges, resulting in substantial reduction of internal en-
Fig. 2 also shows the various quantities for a standard ergy e and the total mass mS . The internal energy drops
Oppenheimer-Snyder collapse (dashed lines), which es- to e ∼ −0.4, and the total mass drops to about half its
sentially overlap the one for the Hawking-radiation col- initial value.
lapse case. As can be seen, the velocity of the fluid changes be-
The similarities between Oppenheimer-Snyder collapse haviour from negative to positive at the bounce radius,
and the Hawking-radiation collapse end, however, as the describing a collapsing phase switching to an expanding
star approaches its Schwarzschild-radius, which would phase of the fluid. The inner layers of the star pick up a
happen when θ(k) crosses zero. Just before the star larger positive velocity than the outer layers, resulting in
crosses its Schwarzschild-radius, the luminosity LS di- shell crossing. At that point, due to shell crossing which
6

Areal radius R Potential φ Potential λ


14.0 14.0 0.0 14.0 0.7
13.9 3.6 13.9 0.1 13.9 0.1
3.0 0.2 0.9
13.8 2.4 13.8 0.3 13.8 1.7
2.5
13.7 1.8 13.7 0.4 13.7 3.3
t

t
13.6 1.2 13.6 0.5 13.6 4.1
0.6 4.9
13.5 0.6 13.5 13.5 5.7
0.0 0.7 6.5
13.4 Areal radius R
140.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
13.4 Potential φ
140.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
13.4 Potential λ
140.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
12 r 12 r 0.0 12 r 0.7
3.6 0.1 0.1
10 3.0 10 0.2 10 0.9
2.4 1.7
8 8 0.3 8 2.5
1.8 0.4 3.3
t

t
6 6 6 4.1
1.2 0.5 4.9
4 4 0.6 4
0.6 5.7
2 0.0 2 0.7 2 6.5
0 0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
r r r

FIG. 5: Contour plots of metric quantities for the simulation with R0 = 4, M0 0 = 0.4. The lower panel show the entire
time-range on the y-axis, the upper panels zoom into the late-time phase with strong Hawking radiation.

is an artifact of p = 0 choice in this model, the density Φ drops quickly to values around −0.5. The potential λ
n(r, t) at the surface of the star diverges ∝ (tcrit − t)−1/2 , shows also two distinct phases: During the Oppenheimer-
and the simulation crashes. The divergence in n(r, t) Snyder phase, λ remains approximately spatially con-
at finite stellar radius and mass indicating shell-crossing stant, but slowly decreases. During the Hawking radi-
of the outer layers of the star, can not be handled by ation phase, λ quickly decreases close to the surface of
our code. A more sophisticated numerical method, in the star, presumably in accord with the diverging number
combination with a fluid-description with pressure would density n close to the stellar surface.
be needed to proceed further and will be done in fu- Figure 6 explores the parameter space of different R0
ture work. Nevertheless, we consider the behavior before and M0 . Generically, we find the behavior shown in
shell-crossing as generic. Figs. 2 and 3: The stars first collapse at nearly con-
The behavior of this collapse with Hawking radiation stant mass until they reach a size close to the horizon (in
is explored further in Fig. 3. Shown are 16 areal radii Fig. 6, this represents a diagonal line at mS = 2RS ). As
(at constant comoving coordinates), as well as 2mS (t). the stars approach their horizons, luminosity increases
The main panel shows clearly the homologous collapse and the mass drops significantly, with the star remain-
at early times t . 13.5. The inset shows an enlargement ing just outside RS = 2mS . All trajectories approach the
of the late-time behavior, showing clearly that the mass same limiting behavior, thus we find the expected generic
mS is reduced such that the outer surface of the star behavior of Hawking radiation.
remains just outside its horizon. We have shown results here for very compact stars as
Figure 4 shows radial velocity U at 16 comoving radii. those are the more likely ones to collapse into a black
When Hawking radiation becomes dominant during the hole. We conducted the analysis for stars with low mass,
last ∼ 1 of the evolution, the inward motion is signifi- see the M = 0.1 data in Fig. 6. As expected these stars
cantly slowed down in the outer layers, and even reversed behave differently, and substantially evaporate and ex-
in the inner layers of the star. The radial gradient in the plode before they approach their horizon.
acceleration (with larger outward velocities toward the
center of the star) is responsible for the diverging num-
ber density in the star. To proceed further will either V. CONCLUSIONS
require a treatment of shocks, or a modification of our
model to avoid shocks. Einstein equations tell us that the final destiny of a
Figure (5), finally presents the metric quantities for gravitationally collapsing massive star is a black hole [2].
this evolution. The areal radius R is very similar to the This system satisfies all the conditions of the Penrose
standard Oppenheimer-Snyder collapse; the small devia- Hawking singularity theorem [25]. However a collaps-
tions during the Hawking radiation phase are not easily ing star has a spacetime dependent gravitational field
visible. The lapse-potential Φ clearly shows two distinct which by the theory of quantum fields on curved space-
phases: During the quasi Oppenheimer-Snyder collapse time should give rise to a flux of particles created [11].
Φ ≈ 0. When the Hawking radiation becomes significant, Hawking discovered in the early 0 700 s that this is in-
7

2 Hawking flux produced only during the collapse. Here


we solved the full set of Einstein and energy conserva-
tion equations numerically therefore the exact solutions
1 we obtain are robust. We find for a range of initial
masses and radii of the collapsing star that the evolution
proceeds through two phases: First, a collapsing phase
mS(t)

0.4 where Hawking radiation is unimportant, and the star


follows very closely standard Oppenheimer-Snyder col-
1.2
lapse. When the star approaches formation of an horizon,
0.2 then Hawking radiation sets in. This slows down the col-
0.8 lapse (e.g. Fig. 4) while significantly reducing the mass
of the star. Both effects (slowdown and mass-loss) bal-
0.1 ance such that the evaporating star remains very slightly
0.4
outside its event horizon, cf. Fig. 6.
1 2 The key idea that enables this program is the fact that
0.04 radiation is produced during the collapse stage of the
star, prior to black hole formation. Once the star be-
0.3 1 3 10 30 100
comes a black hole nothing can escape it, not even light.
RS(t) Since Hawking radiation is produced prior to black hole
formation then its backreaction on the star’s evolution
FIG. 6: Collapsing stars with Hawking radiation for several can be included in the set of hydrodynamic equations
choices of initial mass and initial radius. Shown are trajec- for the coupled system of the quantum field and the
tories in the (RS (t), mS (t))-plane, with the circles indicating star. Thus we solve numerically the set of hydrodynamics
the initial conditions of each run. The filled blue circle marks equations describing the evolutio of collapse for the inho-
the evolution discussed in detail in Figs. 2–5. The dashed mogeneous dust star absorbing negative energy Hawking
orange line indicates the Schwarzschild horizon, RS = 2mS .
flux during its collapse. We discover that the stars ex-
The inset shows an enlargement, to emphasize the universal
nature of the near-horizon behavior. plode instead of collapsing to a black hole, as they get
close to their last stage of collapse, just when Hawking
radiation reaches a maximum, but they never cross what
deed the case for stars collapsing to a black hole, they would have been the apparent horizon. Just like Hawk-
produce Hawking radiation [3]. The conclusions derived ing radiation which is universal, we discover that this be-
from both theories, the existence of black holes from Ein- haviour of the collapsing stars bouncing and exploding
stein’s theory of gravity and the existence of Hawking before the horizon and singularity would have formed,
radiation from the theory of quantum fields in curved is universal, i.e independent of their characteristics such
spacetime, were soon found to be in high friction with one as mass and size. Physically the backreaction of ingoing
another, (see [26] for an interesting treatment) . They negative energy Hawking radiation reduces the gravita-
led to a series of paradoxes, most notably the informa- tional binding energy in the star with the maximum loss
tion loss [5] and firewalls [7]. Being forced to give up on near the last stages of collapse, while taking momentum
either unitarity or causality is at the heart of the prob- away from the star. This is the reason for the universal
lem. But the quantum theory is violated if unitarity is feature of the explosion in the star instead of its collapse
broken. Einstein’s theory is violated if causality is bro- to a black hole singularity. Independent of what size and
ken. Violations of the quantum theory imply Hawking mass the star starts from, most if its radiation will be pro-
radiation may not exist. Violations of Einstein’s theory duced as the star nears its future horizon. At that stage
of gravity, on which the singularity theorem is based, im- the drop in mass and internal energy is maximum and
ply black holes may not exist. Thus black hole physics the star goes through a bounce from collapse to explo-
provides the best arena for understanding the friction be- sion. The negative energy Hawking radiation absorbed
tween quantum and gravitational physics. In this light, by the star, violates the energy condition of the singular-
an investigation of the evolution of the star’s interior as ity theorem [25] thus it is not surprising that a singularity
it is approaching its singularity with the backreaction of and an horizon do not form, features traditionally associ-
the Hawking radiation produced by quantum effects, is ated with the definition of black holes. Stated simply our
of fundamental importance. findings indicate that singularities and horizon do not ex-
This problem was first investigated with semianalyt- ist due to quantum effect and that universally stars blow
ical methods and a series of approximations, such as a up on their last stage of collapse.
homogeneous star placed in the Hartle Hawking thermal More specifically, we find that collapsing stars slow
bath in [1]. down their collapse right outside their horizon, while
In this work we extend our investigation to study a substantially reducing their mass through Hawking ra-
more realistic system of a collapsing star with the quan- diation. In the cases studied, the mass of the black hole
tum field in the Unruh vacuum, which gives rise to a is reduced by roughly a factor of 2, with the radius of the
8

star shrinking in proportion, to preserve RS /mS > 2, that universally collapsing stars bounce into an expand-
cf. Figs. 3 and 6. The velocity of the collapsing mat- ing phase and probably blow up, instead of collapsing to
ter then reverses toward an expanding phase, cf. Fig. 4 a black hole. Thus ’fireworks’ should replace ’firewalls’.
with the inner layers expanding faster than the outer
layers. Unfortunately the latter behaviour, i.e. the ’un-
folding’ of the star from inside out leads numerically to
shell-crossing which our code can not handle. Once shell Acknowledgments
crossing occurs we can not follow the evolution further
to determine whether in the expanding phase, the stellar LMH is grateful to DAMTP Cambridge University
remnant explode, or simply evaporate away, despite that for their hospitality when this work was done. LMH
from the low mass cases and from the velocity results the would like to thank M.J. Perry, L. Parker, J. Bekenstein,
explosion seems to be the case. P. Spindel, and G. Ellis for useful discussions. LMH ac-
The star never crosses its horizon, so neither unitarity knowledges support from the Bahnson trust fund. HPP
nor causality are violated, thereby solving the longstand- gratefully acknowledges support from NSERC of Canada
ing information loss paradox. This investigation shows and the Canada Research Chairs Program.

[1] L. Mersini-Houghton, arXiv:1406.1525 [hep-th]. [8] S. W. Hawking, arXiv:1401.5761 [hep-th].


[2] J. R. Oppenheimer and H. Snyder, Phys. Rev. 56, 455 [9] W. G. Unruh, Phys. Rev. D 14, 870 (1976).
(1939). [10] S. Yamada [astro-ph/9601042].
[3] S. W. Hawking Nature 248, 30-31, (1974); S. W. Hawk- [11] L. Parker, Phys. Rev. Lett. 21, 562 (1968); L. Parker,
ing, Commun. Math. Phys. 43, 199, (1975). Phys. Rev. 183, 1057 (1969); L. Parker, Phys. Rev. D 3,
[4] L. Parker, Phys. Rev. D 12, 1519 (1975); L. Parker and 346 (1971) [Erratum-ibid. D 3, 2546 (1971)]; L. Parker,
S. A. Fulling, Phys. Rev. D 7, 2357 (1973). Phys. Rev. Lett. 28, 705 (1972) [Erratum-ibid. 28, 1497
[5] R. Penrose, “The big bang, quantum gravity and (1972)]; L. Parker, In *Cincinnati 1976, Proceedings,
black-hole information loss,” Foundations of Space and Asymptotic Structure Of Space-time*, New York 1977,
Time: Reflections on Quantum Gravity 10-14 Aug 107-226; L. Parker and S. A. Fulling, Phys. Rev. D 9,
2009. Cape Town, South Africa, CNUM: C09-08-10.2; 341 (1974).
W. Israel and Z. Yun, Phys. Rev. D 82, 124036 [12] W. C. Hernandez and C. W. Misner, Astrophys. J. 143,
(2010) [arXiv:1009.0879 [hep-th]]; M. Varadarajan, J. 452 (1966).
Phys. Conf. Ser. 140, 012007 (2008); T. Vachaspati, [13] S. Bose, L. Parker and Y. Peleg, Phys. Rev. D 53, 7089
D. Stojkovic and L. M. Krauss, Phys. Rev. D 76, (1996) [gr-qc/9601035].
024005 (2007) [gr-qc/0609024]; S. W. Hawking, Phys. [14] Ingemar Bengtsson Spherical Symmetry and Black
Rev. D 72, 084013 (2005) [hep-th/0507171]; S. Hossen- Holes http://www.fysik.su.se/ ingemar/sfar.pdf; Lu-
felder, arXiv:1401.0288 [hep-th]; C. Kiefer, Lect. Notes ciano Rezzolla An introduction to gravitational col-
Phys. 633, 84 (2003) [gr-qc/0304102]; G. ’t Hooft, gr- lapse to black holes http://www.aei.mpg.de/ rez-
qc/9509050; C. G. Callan, Jr., In *Meribel les Al- zolla/lnotes/mondragone/collapse.pdf.
lues 1994, Electroweak interactions and unified theo- [15] J. B. Hartle and S. W. Hawking, Phys. Rev. D 13, 2188
ries* 311-327; T. Banks, Nucl. Phys. Proc. Suppl. 41, (1976).
21 (1995) [hep-th/9412131]; C. R. Stephens, G. ’t Hooft [16] K. W. Howard and P. Candelas, Phys. Rev. Lett. 53, 403
and B. F. Whiting, Class. Quant. Grav. 11, 621 (1994) (1984).
[gr-qc/9310006]; T. M. Fiola, J. Preskill, A. Strominger [17] S. M. Christensen and S. A. Fulling, Phys. Rev. D 15,
and S. P. Trivedi, Phys. Rev. D 50, 3987 (1994) [hep- 2088 (1977); P. C. W. Davies, S. A. Fulling, S. M. Chris-
th/9403137]; S. B. Giddings, Phys. Rev. D 49, 4078 tensen and T. S. Bunch, Annals Phys. 109, 108 (1977);
(1994) [hep-th/9310101]; L. Smolin, In *College Park N. D. Birrell and P. C. W. Davies, “Quantum Fields
1993, Directions in general relativity, vol. 2* 237-292 [gr- in Curved Space,”, Cambridge Monogr.Math.Phys;
qc/9301016]. N. D. Birrell and P. C. W. Davies, Nature 272, 35 (1978);
[6] R. Brout, S. Massar, R. Parentani, Ph. Spindel, 1995. N. D. Birrell and P. C. W. Davies, Phys. Rev. D 18, 4408
1 Phys. Rept. 260, (1995), 329-454 ; F. Englert, Ph. (1978); G. ’t Hooft, Nucl. Phys. Proc. Suppl. 203-204,
Spindel, JHEP 1012 (2010) 065. 155 (2010).
[7] S. L. Braunstein, H. -J. Sommers and K. Zy- [18] P. Candelas, Phys. Rev. D 21, 2185 (1980).
czkowski, arXiv:0907.0739 [quant-ph]; A. Almheiri, [19] P. C. W. Davies and S. A. Fulling, Proc. Roy. Soc. Lond.
D. Marolf, J. Polchinski and J. Sully, JHEP 1302, 062 A 348, 393 (1976); P. C. W. Davies and S. A. Fulling,
(2013) [arXiv:1207.3123 [hep-th]];A. Almheiri, D. Marolf, Proc. Roy. Soc. Lond. A 356, 237 (1977); P. Candelas and
J. Polchinski, D. Stanford and J. Sully, JHEP 1309, D. J. Raine, J. Math. Phys. 17, 2101 (1976);L. Parker
018 (2013) [arXiv:1304.6483 [hep-th]]; A. Almheiri, and D. J. Toms, Gen. Rel. Grav. 17, 167 (1985).
D. Marolf, J. Polchinski, D. Stanford and J. Sully, JHEP [20] P. Davies Roy. Soc. Pub. P. Davies, Fulling and Unruh
1309, 018 (2013) [arXiv:1304.6483 [hep-th]]; S. L. Braun- Phys. Rev. D 13 (1975);S. A. Fulling and W. G. Unruh,
stein and A. K. Pati, Phys. Rev. Lett. 98, 080502 (2007) Phys. Rev. D 70, 048701 (2004)
[gr-qc/0603046]. [21] G. Abreu and M. Visser, Phys. Rev. D 82, 044027 (2010)
9

[arXiv:1004.1456 [gr-qc]]. 314, 529 (1970).


[22] C. Barcelo, S. Liberati, S. Sonego and M. Visser, Phys. [26] J. T. Firouzjaee and G. F. R. Ellis, arXiv:1408.0778
Rev. Lett. 97, 171301 (2006) [gr-qc/0607008]. [gr-qc];G. F REllis, R. Goswami, A. I. M. Hamid and
[23] Hideo Kodama Prog.Theor.Phys. 63 (1980) 1217. S. D. Maharaj, arXiv:1407.3577 [gr-qc].
[24] W. Israel, Phys. Lett. A 57, 107 (1976); W. Israel, Phys. [27] K. z. Felde, arXiv:1407.6205 [physics.gen-ph];H. Culetu,
Rev. 153, 1388 (1967); W. Israel Aspects of Black Hole arXiv:1407.7119 [gr-qc];M. Visser, arXiv:1407.7295 [gr-
Entropy (1985). qc];J. M. Bardeen, arXiv:1406.4098 [gr-qc].
[25] S. W. Hawking and R. Penrose, Proc. Roy. Soc. Lond. A

You might also like