Stellar Black Holes and the Origin of Cosmic Acceleration
Chanda Prescod-Weinstein,1, 2, ∗ Niayesh Afshordi,1, † and Michael L. Balogh2, ‡
1
arXiv:0905.3551v2 [astro-ph.CO] 25 Jul 2009
2
Perimeter Institute for Theoretical Physics, 31 Caroline St. N., Waterloo, ON, N2L 2Y5, Canada
Department of Physics and Astronomy, University of Waterloo, Waterloo, ON, N2L 3G1, Canada
(Dated: July 25, 2009)
The discovery of cosmic acceleration has presented a unique challenge for cosmologists. As observational cosmology forges ahead, theorists have struggled to make sense of a standard model
that requires extreme fine tuning. This challenge is known as the cosmological constant problem.
The theory of gravitational aether is an alternative to general relativity that does not suffer from
this fine-tuning problem, as it decouples the quantum field theory vacuum from geometry, while
remaining consistent with other tests of gravity. In this paper, we study static black hole solutions
in this theory and show that it manifests a UV-IR coupling: Aether couples the spacetime metric
close to the black hole horizon, to metric at infinity. We then show that using the Trans-Planckian
ansatz (as a quantum gravity effect) close to the black hole horizon, leads to an accelerating cosmological solution, far from the horizon. Interestingly, this acceleration matches current observations
for stellar mass black holes. Based on our current understanding of the black hole accretion history
in the Universe, we then make a prediction for how the effective dark energy density should evolve
with redshift, which can be tested with future dark energy probes.
I.
INTRODUCTION
The discovery of recent acceleration of cosmic expansion was one of the most surprising findings in modern
cosmology [1, 2]. The standard cosmological model (also
known as the concordance model) drives this expansion
with a cosmological constant (CC). While the CC is consistent with (nearly) all current cosmological observations
[28], it requires an extreme fine-tuning of more than 60
orders of magnitude, known as the cosmological constant
problem [3].
In the context of the concordance cosmological model,
there are (at least) three different aspects of the CC problem. For decades, physicists worried about why the value
of the cosmological constant/vacuum energy seemed to
be nearly zero by particle physics standards (known as
the old CC problem)[4], and the conventional wisdom was
that it should vanish exactly, as a result of a yet unknown
symmetry of nature. The accelerated cosmic expansion
has thus challenged us to address this question on two
new fronts. First is the new CC problem: why is the vacuum energy density so close to zero, but non-vanishing?
Second is the coincidence problem: Why did the dark energy dominance and structure formation happen at approximately coincident times?
The race is on to simultaneously address these three
questions. A popular alternative approach to the cosmological constant is a model that modifies Einstein’s
theory of gravity. Traditionally, this involves adding
higher order curvature terms to the geometric side of Einstein’s equation. However, in [5], one of us proposed a
novel approach to modified gravity. This model intro-
∗ Electronic
address:
[email protected]
address:
[email protected]
‡ Electronic address:
[email protected]
† Electronic
duces gravitational aether, as a necessary ingredient to
decouple the quantum field theory vacuum from gravity
while simultaneously satisfying other tests of gravity. Unlike many models of modified gravity, the gravitational
aether model modifies the energy-momentum content of
the spacetime, instead of adding higher order curvature
terms.
In this model, the right hand side of the Einstein field
equation is modified as:
1
(8πG′ )−1 Gµν = Tµν − Tαα gµν + p′ (u′µ u′ν + gµν ),
4
(1)
where G′ is 4/3 times the Newton’s constant, and p′ and
u′µ are aether pressure and four-velocity that are fixed by
requiring the conservation of the energy-momentum tensor, Tµν , and the Bianchi identity. As argued in [5], while
consistent with precision tests of gravity, this theory is
preferred to the standard model by the combination of
cosmological observations (with the notable exception of
4
He primordial abundance).
In this paper, we pursue a detailed understanding of
static spherical black hole solutions in the gravitational
aether theory. The solution we find is, at first glance, a
perturbed Schwarzschild metric. However, upon closer
inspection we find that this perturbation is divergent
both near to and far away from the horizon (where we
refer to an infinite redshift surface as a horizon). Thus
the static solution in the presence of gravitational aether
is fundamentally different from Schwarzschild, which can
be characterized as a UV-IR connection: the metric near
and far from the horizon is set by the same integration
constant. Here, we will explore possible meanings of this
property, and whether the cosmological behavior is set by
a Trans-Planckian ansatz close to the black hole horizon.
We note that the static black hole solution found here
also applies to the cuscuton models [6, 7] which have
the same energy-momentum tensor as the gravitational
aether in the limit of vanishing cuscuton potential.
2
II.
BLACK HOLE IN GRAVITATIONAL
AETHER
We find a solution for the static black hole in the Gravitational Aether model using assumptions similar to those
that lead to the Schwarzschild solution. Namely, we assume a spacetime with no matter content, and we assume
spherical symmetry. Given that the aether takes fluid
form, the metric in this model is the same as the general
static, spherically symmetric metric that describes the
interior of a star, as modeled by a perfect fluid. The only
notable divergence from the star model is the absence of
a matter density, leaving an energy-momentum tensor of
the following form:
Tµν = p(uµ uν + gµν ).
We find the following metric
−1
2m
ds2 = −e2φ dt2 + 1 −
dr2 + r2 dΩ2 .
r
(2)
80
60
40
20
0
-20
0.01
0.1
2m
(5)
10
100
-1
FIG. 1: Function f (r) [Eq. (10)] as a function of the distance
from the Schwarzschild radius (= 2m). The deviation from
the Schwarzschild metric is proportional to p0 f (r), where p0 is
the integration constant. If p0 is small, as we argue in Section
III, the corrections only become important at the horizon, and
on cosmological scales.
We see immediately that exp(φ) and p are inversely related:
p = p0 e−φ ,
(6)
where p0 is an integration constant. Notice that Eq. (6)
is equivalent to the condition of hydrostatic equilibrium
for aether, and is valid independent of the assumption of
spherical symmetry, for any static spacetime [29]. Now,
we may rewrite Eqn. 4:
m + 4πr3 p0 e−φ
dφ
=
dr
r(r − 2m)
(7)
We can solve this equation by noting that it is a firstorder inhomogeneous linear differential equation in eφ ,
with the standard solution:
#
1 "Z
−1/2 2
r
(1 − 2m
2m 2
φ(r)
r )
e
= 4πp0 1 −
dr + const. .
r
r − 2m
(8)
To put this into more familiar terms, we can set the
constant, so that we recover the Schwarzschild solution
as p0 → 0:
e
(4)
1
r
(3)
With components obeying the following differential equation, known as the Tolman-Oppenheimer-Volkoff equations [8]:
m + 4πr3 p
dφ
=
,
dr
r(r − 2m)
−p(m + 4πr3 p)
dp
=
.
dr
r(r − 2m)
100
fHrL m2
In Sec. II, we introduce our gravitational aether black
hole solution. We describe the properties of the solution,
including a preferred coordinate system and the location
of the event horizon. We also establish asymptotic properties of the black hole, which are characterized by the
same integration constant both close in and far away from
the horizon of the black hole.
Sec. III explores the Trans-Planckian ansatz, as a way
to fix the aforementioned integration constant, through
quantum gravity effects close to the horizon. We suggest a way to connect the presence of black holes to the
existence of a pervasive pressure that behaves like dark
energy on cosmological scales.
In Sec. IV, we present a study of the contribution that many such black holes would make to the
global/cosmological structure of space-time, while Sec.
V provides a census of average black hole mass through
cosmic history, which translates into a prediction for the
history of cosmic acceleration.
Finally, we will discuss open questions and future
prospects in Sec. VI.
Throughout the paper, we use the natural Planck
units: ~ = c = GN = kB = 1. Moreover, we will replace pressure p′ by 3p/4 in Eq. (1), so that the vacuum
field equations for the aether theory resembles general
relativity sourced by a perfect fluid with pressure p and
zero density.
φ(r)
1
2m 2
= 1−
[4πp0 f (r) + 1] ,
r
(9)
where f (r) is given by:
−1/2
2m
1−
−30m2 + 5mr + r2
r
"
1/2 #
2m
r
r
15 2
1−
, (10)
−1+
+ m ln
2
m
m
r
f (r) =
1
2
3
and is shown in Fig. (1). In the limit where r is large
(r ≫ m):
f (r) =
r2
+ 3mr + O[m2 ].
2
(11)
while close to the “Schwarzschild horizon” we find:
√ 5/2
2m
+ O[m3/2 (r − 2m)1/2 ].
(12)
f (r) = −8 √
−2m + r
Thus the correction to the Schwarzschild metric dominates in both UV and IR regimes (corresponding to close
to and far from the BH horizon). This a nice tie, even
for arbitrarily small values of the integration constant p0 .
Therefore, a very suggestive conclusion is that, unlike in
general relativity, the gravitational aether ties the formation of black hole horizons to cosmological dynamics.
But then, is there really an event horizon for this spacetime? Looking at the trace of the Einstein’s equation, we
find that the Ricci scalar is proportional to the pressure
of aether, p, which is in turn inversely proportional to
the 00 component of the metric, eφ . We define the surface where eφ → 0 as the black hole horizon. Therefore
the pressure at the horizon, and thus the Ricci scalar,
goes to infinity (p ∝ R → ∞) implying that this surface
coincides with a real metric singularity (as opposed to a
coordinate singularity).
It appears that we have established that any static
event horizon in a theory of gravitational aether (like the
one we have modeled) coincides with a real metric singularity. In a traditional formulation of general relativity,
such a scenario may be given to ambiguous physical interpretation. Cognizant of the fact that a modified gravity
will display properties divergent from traditional relativity, we expect that such a picture is best contextualized
by a more comprehensive theory of quantum gravity.
Indeed, any process (for example, quantum gravity)
that alleviates/regulates metric singularities, will inevitably remove event horizons from the theory of gravitational aether. In other words, static event horizons
cannot exist in a UV completion of gravitational aether.
This is independent of the assumption of spherical symmetry, and only relies on the aether hydrostatic equilibrium condition (6). However, we note that, as the singularity is a null surface, the spacetime does not violate
the weak cosmic censorship principle.
Back to the spherical aether black hole spacetime (9),
we now notice that the static metric solution is only welldefined for r ≥ 2m, as the solution becomes complex
inside the Schwarzschild radius, r < 2m. More surprisingly, for negative values of p0 , unlike a Schwarzschild
black hole, a free-falling observer can reach this boundary within a finite coordinate time. The reason is that
the redshift of a static source at the Schwarzschild radius
is now finite as seen by distant observers [30]:
1 + z = e−φ
#−1
"
1/2
2m
− 32πp0 m2
≃
1−
r
< 1 + zmax = −
1
.
32πp0 m2
(13)
As to what happens inside r = 2m, it is clear that our
current choice of coordinates do not give us a physical
metric for r < 2m. However, is it possible that with
an appropriate choice of coordinate, we can analytically
continue the static solution beyond the Schwarzschild radius? Indeed, we can define a new radial coordinate:
−1/2
2m
′
1
−
dr
dr grr =
λ ≡
r′
2m
2m
i
h
1/2
= 2 [2m(r − 2m)] + O (r − 2m)3/2 m−1/2 ,(14)
Z
r
′√
Z
r
which is equivalent to the constant-time proper radial
distance. In terms of λ, the metric takes the form:
ds2 = −e2φ dt2 + dλ2 + r(λ)2 dΩ2 ,
(15)
where
λ
+ O[p0 λ2 , λ3 m−3/2 ],
4m
λ2
+ O[λ4 /m2 ].
r(λ) = 2m +
8m
eφ = −32πp0 m2 +
(16)
(17)
In other words, the metric is analytic and real in terms
of the new radial coordinate, λ, at and beyond the
Schwarzschild radius, which now corresponds to λ = 0.
Moreover, a static event horizon, which as we argued corresponds to a real curvature singularity, now exists for all
(small) values of p0 , as eφ = 0 at:
λH ≃ 128πp0 m3 .
(18)
In the next section, we study the implications of this
solution for cosmology. However, we shall postpone a full
investigation of the causal structure of this spacetime,
as well as its possible analytic continuations, to future
studies.
III.
TRANS-PLANCKIAN ANSATZ AND
COSMIC ACCELERATION
In the last section, we saw that within spherical spacetimes in the gravitational aether theory, the integration
constant p0 ties the geometry close to the horizon to the
geometry at infinity. While, in the classical theory, p0 is
an arbitrary integration constant, here we speculate that
its value is fixed by quantum gravity effects, especially
since the horizon is now a curvature singularity, where
quantum gravity effects should become important.
We first note that the temperature of sources that fall
through the Schwarzschild horizon, as seen by distant
observers, approaches the Hawking temperature [9]:
TH =
1
,
8πm
(19)
4
black hole. Comparing the scale of p0 with the density
(≃ − pressure) of the cosmological dark energy, ρΛ :
1
ÈDΦÈaether
0.01
p0
2
= − θP−1
ρΛ
3
1 Msol
10-4
10 Msol
10-6
10-8
100 Msol
10-10
1
10
100
1000
104
DistanceHMpcL
FIG. 2: Predicted large distance deviation from the vacuum
Schwarzschild solution for 1, 10, and 100 M⊙ black holes,
based on the Trans-Planckian ansatz. Here, we assumed θP =
100 in Eq. (22) for non-rotating black holes to find p0 , which
is then plugged into Eq. (9) to find the metric. As pointed
out in the text, the corrections become important on today’s
cosmological horizon scale for solar/stellar-mass black holes.
Furthermore, we assume that the maximum rest-frame
temperature of sources is comparable to the Planck temperature (or one in Planck units):
Tmax = θP = O[1].
(20)
Here, θP is a dimensionless constant that measures Tmax
in units of Planck temperature, which we shall call the
Trans-Planckian parameter. We then adopt the TransPlanckian ansatz, which is the idea that the maximum
redshift at Schwarzschild radius (Eq. 13) is roughly set
by the ratio of the Planck to Hawking temperatures:
1 + zmax
1
Tmax
=−
=
= 8πθP m,
2
32πp0 m
TH
(21)
or
p0 = −
1
256π 2 θ
Pm
3
.
(22)
With this ansatz, we further see that:
λH = −
1
= O[1],
2πθP
(23)
i.e. the event horizon is roughly a Planck length away
from the Schwarzschild radius. Equivalently, the shortdistance aether corrections to the Schwarzschild metric
only become important at about a planck distance from
the horizon/singularity, which is a reasonable expectation from a possible quantum gravitational mechanism.
While this may imply that tests of strong gravity close
to the horizon of a black hole may have a hard time
testing the influence of aether on the spacetime metric,
the Trans-Planckian ansatz has a curious prediction for
the numerical value of p0 , i.e. aether pressure far from the
m
85 M⊙
−3
,
(24)
where we assumed ΩΛ = 0.7 and H0 = 70 km/s/Mpc.
The resulting deviation from the Schwarzschild metric is
shown in Fig. (2) for stellar mass black holes.
This leads us to a very interesting possibility, which
was first conjectured in [5]: that the formation of stellarmass black holes could trigger the onset of cosmic acceleration, especially since aether and dark energy have
similar pressures, assuming that the aether pressure is
set by the Trans-Planckian ansatz for stellar mass black
holes. To see this, we can explicitly compare the black
hole spacetime (Eqs. 3 and 11) far from the black hole
(r ≫ m):
ds2 = −(1 + 2πp0 r2 )2 dt2 + dr2 + r2 dΩ2 ,
(25)
with the de-Sitter spacetime:
ds2 = −(1−8πρΛr2 /3)dt2 +(1−8πρΛr2 /3)−1 dr2 +r2 dΩ2 .
(26)
We thus notice that non-relativistic particles close to
the origin, but far from the black hole horizon (2m ≪
r ≪ |p0 |−1/2 ) see the same Newtonian potential (or gtt )
in both spacetimes, if p0 = −2ρΛ /3. In other words,
close-by non-relativistic test particles (such as galaxies,
stars, or other black holes) accelerate away from the origin/black hole, similar to a de-Sitter space. Moreover,
this acceleration will correspond to the current cosmological observations, if the mass of the black holes is roughly:
−1/3
m ≃ 85 θP
M⊙ .
(27)
So far, our solution has neglected the effects of black
hole spin. Indeed, spin is expected in realistic black holes,
which are fed by astrophysical accretion disks. For example, the dimensionless spin parameter, a∗ = a/m was
recently measured for two stellar-mass black holes, to be
within the range 0.65 − 0.85 [10]. In order to include
3
this effect, we conjecture that p0 scales as TH
(as suggested in [5]), for general black hole spins. This is justified, as the Trans-Planckian ansatz is controlled by the
Hawking temperature, TH , while f (r) also depends on
the surface gravity close to the black hole horizon, which
is also proportional to TH . With this assumption, the
scale-dependence should go as:
h
−1/2 i−3
p0 (m, a∗ )
= 8 1 + 1 − a2∗
,
p0 (m, 0)
(28)
which is in the range 0.2 − 0.6 for a∗ = 0.65 − 0.85.
While this paper only deals with static vacuum solutions, it was shown in [5] that for non-relativistic fluids
(e.g. stars, planets): p′ ≈ −Tαα /4+const., i.e. the local
matter density sets the aether pressure up to a constant.
5
One expects that the constant term would be set by the
boundary conditions at infinity, or by cosmology. Alternatively, what we suggest in this section is that the
boundary condition can be set at the horizons of the
black holes. The fact that this can naturally explain the
onset of cosmic acceleration is certainly very suggestive,
but the best way to test this hypothesis is to see how/if
this boundary condition can emerge from the process of
(classical or quantum) gravitational collapse into a black
hole. We leave this question to future studies.
A further implication of this hypothesis is that solar/stellar mass is the minimum mass of black holes allowed in the model. A discovery of significantly subsolar mass black holes (e.g. primordial black holes
with M ≪ M⊙ ) could potentially rule out the TransPlanckian ansatz, as it would imply much larger than
observed cosmic acceleration rates for θP ∼ 1.
Of course, we also need to patch together and coarsegrain individual black hole spacetimes into a de-Sitter
space, in order to rigorously prove this correspondence.
However, the above argument is already very suggestive, as long as there are many black holes within the
cosmological/de-Sitter horizon, so that one can trust the
above Newtonian argument. In the next section, we provide an approximation to the cosmological spacetime of
multiple black holes.
IV.
GLOBAL CONTRIBUTION OF MULTIPLE
BLACK HOLES
In this section, we will seek an approach to approximately find the spacetime of multiple black holes with
gravitational aether, which can be used to describe an
approximate FRW cosmology. Here, for simplicity, we
focus on the quasi-static Newtonian regime, where we
could assume hydrostatic equilibrium for aether in the
vacuum (6). For simplicity, we ignore the matter inbetween black holes [31], and assume that black holes
are much farther apart than their horizon sizes, but are
much closer than the cosmological horizon. In this limit,
using Eq. (6) we have:
∇2 ln p = −∇2 φ = 0,
Green’s function, which can be found using the method
of images (e.g. [11]):
GD (x, x′ ) =
(30)
For n spheres (black holes) at positions xi and with radii
ai (= 2mi ), we may expand this Green’s function, up to
first image, as
GD (x, x′ ) =
−
n
X
i=1
1
|x − x′ |
a2
,
+O
a2i
|∆x|3
′
|x′ − xi ||x − xi − |x′ −x
2 (x − xi )|
|
i
ai
(31)
which is a good approximation, as long as the distance
between the spheres/black holes is much larger than their
sizes. Now, using Green’s theorem, we can find aether
pressure in-between the black holes, in terms of the pressure on the surfaces of the spheres, pi ’s:
n I
1 X
∂GD
ln p(x) − ln p̄ = −
[ln pi (x′ ) − ln p̄],
ds′ ·
4π i=1 Si
∂x′
H (32)
where ln p̄ is the log of pressure at infinity, and Si ds′
are surface integrals over the horizons of the black holes
(assuming a flat geometry), while pi ∝ m−3
are fixed by
i
the masses of the blacks holes, using the Trans-Planckian
ansatz (22). Since the Green’s function (31) is analogous to superposition of electrostatic potentials of point
charges, we can use Gauss’s theorem to evaluate the surface integrals:
ln p(x) − ln p̄ =
n
X
ai [ln pi (x′ ) − ln p̄]
i=1
|x − xi |
.
(33)
Now, using the assumption of statistical homogeneity,
we expect the spatial/ensemble average of ln p to be the
same as ln p̄. If we take ensemble averages of both sides
of Eq. (33), this yields:
ln p̄ =
(29)
where the assumption of ∇2 φ = 0 is the equivalent of the
Poisson equation in Newtonian gravity, for zero matter
density (which also applies to aether). We thus see that
fixing the aether pressure in the vicinity of black holes,
through the Trans-Planckian ansatz (22), is equivalent to
solving the Laplace equation (29) with Dirichlet boundary conditions at (or close) to the horizon of the black
holes [32].
This problem is analogous to finding the electrostatic
potential between multiple conducting spheres, which
can be solved using the Green’s function for the appropriate geometry. For a single sphere of radius at the origin
(and in a flat space), there is an exact expression for the
1
a
−
.
|x − x′ | x′ |x − xa′22 x′ |
hai ln pi i
,
hai i
(34)
or alternatively:
2
1
hmi ln mi i
p̄ = − ρDE,eff = −
, ln m∗ ≡
,
3
256π 2θP m3∗
hmi i
(35)
where we used pi ∝ m−3
and
a
=
2m
,
as
well
as
Eq.
i
i
i
(22). In other words, in the presence of multiple black
holes, the mean aether pressure, and thus FRW cosmology, is set by m∗ , which is the mass-weighted geometric mean of black hole masses. Subsequently, the correspondence of this mean aether pressure with an effective Dark Energy or cosmological constant density was
demonstrated in the last section.
6
Furthermore, taking the Laplacian of Eq.(33), we can
find an equation for the perturbation of effective Dark
Energy, for sub-horizon perturbations (but on scales
larger than the size of the blacks holes):
∇2 δDE,eff = −8πρBH ln(pi /p̄),
(36)
where δDE,eff is the overdensity of the effective dark energy, while ρBH is the black hole density. Eq. (36) can be,
in principle, used to track cosmological structure formation and the impact on CMB anisotropies (through the
Integrated Sachs-Wolfe effect), but we postpone a study
of these effects to future work.
In the next section, we will provide a quantitative picture of how the cosmic history of accretion into stellar
and super-massive black holes (or active galactic nuclei)
leads to an estimate of m∗ as a function of redshift, and
its implications for the effective dark energy scenarios.
V.
COSMIC HISTORY OF BLACK HOLES AND
COSMIC ACCELERATION
An up-to-date inventory of cosmic energy at the
present day, including the contribution from stellar-mass
and supermassive black holes, is provided by [12]. In order to measure m∗ (z) we need to take this a step farther
and understand the mass distribution of such black holes,
and their redshift evolution.
The mass distribution of stellar–mass black holes is
not well-determined observationally, but estimates are
that it is fairly broad, with a mean of around ∼ 7M⊙
[13, 14]. We base our calculations on the theoretical predictions of [15], which show that the distribution can be
approximately represented by a power-law such that the
number density of black holes decreases by a factor 5
between M = 3M⊙ and M = 15M⊙. Assuming this
distribution, the average black hole mass is 8.2M⊙ . We
will treat the uncertainty in this distribution by varying
the slope sufficiently to alter this mean mass by ±1M⊙.
To determine the redshift evolution in black hole abundance, we use observations of the cosmic star formation
history, from [16]. There is significant uncertainty in the
shape of this history; however it must obey the integral
constraint that the total stellar mass density today be
ρ∗ = 0.0027ρcrit = 3.67 × 108 M⊙ Mpc−3 , which is known
to a precision of ∼ 30 per cent [12]. We will therefore
normalize the black hole number density at z = 0 to
1.46 × 106 Mpc−3 at z = 0 [12]. We assume that changes
to the initial mass function do not significantly alter the
shape of the star formation rate density evolution, but
primarily affect the number of black holes formed. By
default we assume a Kroupa IMF [17], which is the “second model” considered in [12]. For this choice, 0.19 per
cent of stars formed end up as black holes; a more useful number is that for every solar mass of stars formed
0.0025 black holes are created. These numbers change by
less than 5 per cent if we assume a Chabrier IMF [18];
we expect therefore the uncertainty on the normalization
of the black hole mass function to be dominated by the
30 per cent uncertainty in the present day stellar mass
function. Note, however, that a pure Salpeter IMF [19]
would produce significantly fewer black holes, only 0.0013
for every solar mass formed.
We base our estimate of the supermassive black hole
mass distribution on observations of the quasar luminosity function. This requires assumptions about the
lifetime and obscuring column density of quasars; for
this we adopt the model of [20] who describe a mergerdriven scenario of black hole growth. Using this model,
the z = 0 mass density of supermassive black holes is
5
−3
2.9+2.3
. This is somewhat smaller than
−1.2 × 10 M⊙ Mpc
5
the value of 5.4×10 M⊙ Mpc−3 determined from the correlation between black hole mass and bulge luminosity
[21, 22], as computed by [12]. However, the uncertainty
on the latter is a factor of two, and a lower value of
3.4 × 105 M⊙ Mpc−3 is obtained [12] if one uses the correlation with velocity dispersion for early type galaxies
[23, 24] rather than luminosity.
With this in hand we are able to compute the expected
m∗ (z), and this is shown in the bottom panel of Figure 3.
Our best estimate of the local, mass-weighted geometric
mean of black hole masses is m∗ (0) = 12.7M⊙. The
dashed lines represent the range of uncertainty on this
z = 0 normalization. A larger value of m∗ is obtained
by reducing the contribution of stellar-mass black holes
(assuming the local density is 30% lower than our fiducial model, and assuming the mass distribution is more
steeply weighted to lower masses, so the average mass is
7.2M⊙), and increasing the contribution of supermassive
black holes (by increasing the z = 0 space density within
the 1σ uncertainty, to 5.2 × 105 M⊙ Mpc−3 ). This yields
m∗ (0) = 24.7M⊙ . Pushing the numbers in the opposite
direction, we obtain m∗ (0) = 10.5M⊙. Using Eq. (27)
for the current effective density of dark energy, and ignoring the spin of black holes, this range in m∗ (0) translates
to a range for the Trans-Planckian parameter θP :
θP = (0.4 − 5) × 102 .
(37)
We can consider spinning black holes, using our scaling
argument from Section III and taking a nominal value
of a∗ = 0.75. This implies a lower range for the TransPlanckian parameter, θP = 20 − 300, in order to match
the current rate of cosmic acceleration. The fact that
θP ∼ 1, further justifies a Trans-Planckian, or quantum
gravitational origin for the observed “dark energy phenomenon”.
The evolution of the stellar-mass black hole mass density is dependent upon the shape of the star-formationrate density plot from [16]. To consider the effect of this,
we construct two star formation histories that are consistent with those data within the 1σ error bars, but which
produce as many stars as possible at either high redshift
(z > 1) or at low redshift (z < 0). We still renormalize this to match the local stellar mass density. These
extremes are shown in Figure 3 as dashed lines. The
evolution of the supermassive black hole distribution is
7
We can define a mean equation of state as:
1 + w̄(< z) ≡
FIG. 3: Bottom panel: The mass-weighted geometric mean
of black hole masses, m∗ , in units of M⊙ as a function of redshift. Our fiducial model (solid, black line) assumes our best
estimates of the mass distribution evolution of the black hole
mass distribution. Dashed lines indicate the range of uncertainty expected due to the unknown relative contribution of
supermassive and stellar-mass black holes at z = 0, while
the dotted lines represent the uncertainty in the shape of the
star formation density evolution from [16]. Top panel: The
prediction of the equation of state parameter w̄(< z) from
Equation 39, for the same models. The dashed area shows
the region excluded at 68% confidence level for this parameter, as measured from independent observations [25].
very model dependent, and not well constrained. We note
that the two different predictions shown by [20], which
make different assumptions about the quasar space density evolution at z > 2, have a subdominant effect on the
predictions shown here, relative to the other uncertainties considered.
Within an effective dark energy description of FRW
cosmology, a fixed dark energy equation of state, w, implies that dark energy density evolves as (1 + z)3(1+w), as
a function of redshift, z. The effective equation of state
(which is simply a way to parameterize cosmic expansion
history) is observationally constrained to
w(z) = −1.06 ± 0.14 + (0.36 ± 0.62)
z
,
1+z
(38)
at 68% confidence level, based on cosmic microwave background, baryonic acoustic oscillations, and supernovae
Ia observations, assuming a spatially flat cosmology [25],
and a linear dependence of w(z) on the cosmological scale
factor = (1 + z)−1 .
1 ln[ρDE,eff (z)/ρDE,eff (0)]
3
ln(1 + z)
ln[m∗ (z)/m∗ (0)]
,
=−
ln(1 + z)
(39)
since, ρDE,eff (z) ∝ m−3
∗ (z), as we saw in the last section. We show this estimate of w for the models described
above, in the top panel of Figure 3. Our fiducial model
predicts a value of w that deviates from −1 by less than
5 per cent out to z ∼ 2, but predicts it should reach
w = −0.8 by z = 3. There is considerable uncertainty
on this, however, due both the unknown distribution of
black hole masses at z = 0 (dashed lines) and the unknown shape of the star formation rate density evolution
(dotted lines).
While most these models are consistent with the current bounds on the effective dark energy equation of state
(using Eq. 38):
z
w̄(< z) = −1.06±0.14+(0.36±0.62) 1 −
,
(1 + z) ln(1 + z)
(40)
stage IV dark energy missions, as quantified by the dark
energy task force report [26], are expected to have percent
level sensitivity to w̄(< 1 − 3), and thus should be able to
distinguish the aether model with these m∗ (z) histories
from a cosmological constant.
VI.
CONCLUSIONS AND FUTURE
PROSPECTS
We have shown that static black hole solutions exist in
the gravitational aether model of [5]. The model is an attractive alternative to the cosmological constant, which
does not suffer from the tremendous fine-tuning problem
of vacuum energy in standard model. We find that in
the presence of a gravitational aether, the Schwarzschild
black hole is sufficiently perturbed so as to result in
a Trans-Planckian connection between physics near the
black hole horizon and cosmology. This could be a phenomenological product of quantum gravity, and it naturally explains the present-day acceleration of cosmic
expansion as a result of formation of stellar/solar-mass
black holes.
Indeed, the recent discovery of cosmic acceleration, or
dark energy [1, 2] might be the first concrete evidence for
quantum gravity and/or Trans-Planckian physics. Future work may include an exploration of quantum properties of this black hole solution. In particular, a natural
next step would be to understand how quantum gravity
can resolve the null singularity at the event horizon.
As discussed in Sec. III, another important question
yet to be addressed is whether dynamical evolution could
lead to the static solutions found in this work. While
prior to formation of black holes, the integration constant
8
logical probes over the next decade [5].
p0 is set by large-scale conditions, as black hole horizons
form, we speculate that the constant is instead set by
conditions at the event horizon. In order to understand
the causal transition between these two boundaries, and
how fast the effect will propagate away from the black
hole, a more complete dynamical picture is necessary.
Furthermore, in the presence of multiple black holes
with relative motion, the aether is expected to be locally
dragged by different black hole horizons. However, for
black holes at large separations compared to their horizon sizes and non-relativistic velocities (as expected in
astrophysical situations), the perturbations to the static
solution is expected to be small.
To conclude, we would like to entertain the exciting
possibility that the gravitational aether [5] might provide
a complete solution to the three aspects of the cosmological constant (CC) problem, as discussed in the Introduction:
3. Coincidence problem: As we showed in Sec. (V),
the stellar mass black holes expected in standard
star formation, can naturally lead to the observed
present-day acceleration of the Universe. The competition between the contribution of stellar mass
black holes, and super-massive black holes leads
to an evolution of the effective dark energy density, which can be tested with NASA’s future Joint
Dark Energy Mission (JDEM) [33] or its European
counterpart Euclid [34].
1. Old CC problem: Gravitational aether theory decouples quantum vacuum from geometry, which
allows a nearly flat spacetime even in the presence of large vacuum energy densities expected
from the standard model of particle physics. The
model makes specific predictions for physics at big
bang nucleosynthesis and radiation-matter transition era, which will be tested with precision cosmo-
We would like to acknowledge helpful discussions with
Simon DeDeo, Ghazal Geshnizjani, Jurjen Koksma, Rob
Myers, Lee Smolin, and Mark Wyman. CP and NA are
supported by Perimeter Institute for Theoretical Physics.
Research at Perimeter Institute is supported by the Government of Canada through Industry Canada and by the
Province of Ontario through the Ministry of Research &
Innovation.
[1] A. G. Riess et al., , Astron. J. 116 (1998) 1009–1038,
arXiv:astro-ph/9805201.
[2] S. Perlmutter et al., , Astrophys. J. 517 (1999) 565–586,
arXiv:astro-ph/9812133.
[3] S. Weinberg, Reviews of Modern Physics 61 (Jan.,
1989) 1–23.
[4] N. Straumann, arXiv:gr-qc/0208027.
[5] N. Afshordi, ArXiv e-prints (July, 2008) ,
arXiv:0807.2639.
[6] N. Afshordi, D. J. H. Chung, and G. Geshnizjani, ,
Phys. Rev. D75 (2007) 083513, arXiv:hep-th/0609150.
[7] N. Afshordi, D. J. H. Chung, M. Doran, and
G. Geshnizjani, , Phys. Rev. D75 (2007) 123509,
arXiv:astro-ph/0702002.
[8] S. M. Carroll, Spacetime and geometry. An introduction
to general relativity. 2004.
[9] M. Nouri-Zonoz and T. Padmanabhan,
arXiv:gr-qc/9812088.
[10] R. Shafee, J. E. McClintock, R. Narayan, S. W. Davis,
L.-X. Li, and R. A. Remillard, ApJ 636 (Jan., 2006)
L113–L116, arXiv:astro-ph/0508302.
[11] J. D. Jackson, Classical Electrodynamics. John Wiley &
Sons, New York, 3 ed., 1998.
[12] M. Fukugita and P. J. E. Peebles, ApJ 616 (Dec., 2004)
643–668, arXiv:astro-ph/0406095.
[13] C. D. Bailyn, R. K. Jain, P. Coppi, and J. A. Orosz,
ApJ 499 (May, 1998) 367–+, arXiv:astro-ph/9708032.
[14] K. A. Postnov and A. M. Cherepashchuk, Astronomy
Reports 47 (Dec., 2003) 989–999.
[15] C. L. Fryer and V. Kalogera, ApJ 554 (June, 2001)
548–560, arXiv:astro-ph/9911312.
[16] A. M. Hopkins and J. F. Beacom, ApJ 651 (Nov., 2006)
142–154, arXiv:astro-ph/0601463.
[17] P. Kroupa, MNRAS 322 (Apr., 2001) 231–246,
arXiv:astro-ph/0009005.
[18] G. Chabrier, PASP 115 (July, 2003) 763–795,
arXiv:astro-ph/0304382.
[19] E. E. Salpeter, ApJ 121 (Jan., 1955) 161–+.
[20] P. F. Hopkins, L. Hernquist, T. J. Cox, T. Di Matteo,
B. Robertson, and V. Springel, ApJS 163 (Mar., 2006)
1–49, arXiv:astro-ph/0506398.
[21] L. Ferrarese, ApJ 578 (Oct., 2002) 90–97,
arXiv:astro-ph/0203469.
[22] K. Gebhardt, J. Kormendy, L. C. Ho, R. Bender,
G. Bower, A. Dressler, S. M. Faber, A. V. Filippenko,
R. Green, C. Grillmair, T. R. Lauer, J. Magorrian,
J. Pinkney, D. Richstone, and S. Tremaine, ApJ 543
(Nov., 2000) L5–L8, arXiv:astro-ph/0007123.
[23] D. Merritt and L. Ferrarese, ApJ 547 (Jan., 2001)
140–145, arXiv:astro-ph/0008310.
[24] S. Tremaine, K. Gebhardt, R. Bender, G. Bower,
A. Dressler, S. M. Faber, A. V. Filippenko, R. Green,
C. Grillmair, L. C. Ho, J. Kormendy, T. R. Lauer,
J. Magorrian, J. Pinkney, and D. Richstone, ApJ 574
(Aug., 2002) 740–753, arXiv:astro-ph/0203468.
[25] E. Komatsu, J. Dunkley, M. R. Nolta, C. L. Bennett,
B. Gold, G. Hinshaw, N. Jarosik, D. Larson, M. Limon,
L. Page, D. N. Spergel, M. Halpern, R. S. Hill,
A. Kogut, S. S. Meyer, G. S. Tucker, J. L. Weiland,
E. Wollack, and E. L. Wright, ApJS 180 (Feb., 2009)
2. New CC problem: Formation of black holes leads
to a UV-IR coupling, which connects near-horizon
Planck-scale physics to cosmology, and can naturally lead to cosmic acceleration, even in the absence of a real dark energy component.
9
[26]
[27]
[28]
[29]
[30]
330–376, arXiv:0803.0547.
A. Albrecht, G. Bernstein, R. Cahn, W. L. Freedman,
J. Hewitt, W. Hu, J. Huth, M. Kamionkowski, E. W.
Kolb, L. Knox, J. C. Mather, S. Staggs, and N. B.
Suntzeff, ArXiv Astrophysics e-prints (Sept., 2006) ,
arXiv:astro-ph/0609591.
N. Afshordi, G. Geshnizjani, and J. Khoury, ArXiv
e-prints (Dec., 2008) , arXiv:0812.2244.
See [27] for an account of observational anomalies in the
standard cosmological model.
This follows from the relativistic Euler
equation:(ρ + p)u · ∇u = −∇⊥ p, assuming a static
spacetime and zero density, ρ = 0.
Here, distant observers are located at
[31]
[32]
[33]
[34]
2m ≪ r ≪ (−p0 )−1/2 .
This is not a bad approximation since, as we argued in
the last section, the effect of matter on the aether
pressure is localized and does not extend into vacuum
in the non-relativistic regime.
Since pressure approaches p0 at several BH horizon
radii for individual black holes, as long as the distance
in-between black holes is much larger than their horizon
radii, the exact radius at which the boundary condition
is set is not important.
http://jdem.gsfc.nasa.gov/
http://sci.esa.int/sciencee/www/object/index.cfm?fobjectid=42266