Tachyacoustic Cosmology: An Alternative to Inflation
Dennis Bessada1,2 ,∗ William H. Kinney1,† , Dejan Stojkovic
1,‡
and John Wang1,3§
1
arXiv:0908.3898v2 [astro-ph.CO] 25 Sep 2009
2
Dept. of Physics, University at Buffalo, the State University of New York, Buffalo, NY 14260-1500, United States
INPE - Instituto Nacional de Pesquisas Espaciais - Divisão de Astrofı́sica, São José dos Campos, 12227-010 SP, Brazil
3
Dept. of Physics, Niagara University, NY 14109-2044, United States
We consider an alternative to inflation for the generation of superhorizon perturbations in the
universe in which the speed of sound is faster than the speed of light. We label such cosmologies, first
proposed by Armendariz-Picon, tachyacoustic, and explicitly construct examples of non-canonical
Lagrangians which have superluminal sound speed, but which are causally self-consistent. Such
models possess two horizons, a Hubble horizon and an acoustic horizon, which have independent
dynamics. Even in a decelerating (non-inflationary) background, a nearly scale-invariant spectrum
of perturbations can be generated by quantum perturbations redshifted outside of a shrinking acoustic horizon. The acoustic horizon can be large or even infinite at early times, solving the cosmological
horizon problem without inflation. These models do not, however, dynamically solve the cosmological flatness problem, which must be imposed as a boundary condition. Gravitational wave modes,
which are produced by quantum fluctuations exiting the Hubble horizon, are not produced.
PACS numbers: 98.80.Cq
I.
INTRODUCTION: WHY INFLATION WORKS
Inflationary cosmology [1, 2, 3] is the most successful
and widely accepted paradigm for understanding the very
early universe. By all measures inflation is a compelling
and scientifically useful theory, and makes quantitative
predictions which have been strongly supported by recent
data [4, 5, 6]. The two main hallmarks of inflationary cosmology are solutions to the flatness and horizon problems
of the standard Big Bang cosmology: why is the universe
so close to geometrically flat, and how did the apparent
acausal structure of the universe arise? “Acausal” more
specifically means that the universe is approximately homogeneous on scales larger than a Hubble length H −1 ,
and in addition exhibits a spectrum of density perturbations which is correlated on scales larger than a Hubble
length. Such superhorizon correlations are generated in
inflation by accelerated expansion, ä/a > 0, where a (t)
is the cosmological scale factor, which means that the
comoving Hubble length dH ≃ (aH)−1 shrinks with the
expansion of the universe,
d
−1
(aH) < 0.
(1)
d ln a
Therefore, quantum perturbations, which have constant
wavelength in comoving units, are smaller than the Hubble length at early times, and are redshifted to larger
than the Hubble length at late times, where they are
“frozen” as classical perturbations. Furthermore, as long
as the Hubble constant H is slowly varying with time,
the perturbations generated in inflation are nearly scaleinvariant, consistent with observation. Furthermore, the
∗
[email protected],
†
[email protected],
[email protected],
§
[email protected]
‡
solution to the horizon problem and the flatness problem
are linked in inflation via a conservation law,
d |Ω − 1|
= 0.
d ln a d2H
(2)
Through this conservation law, a universe with shrinking
comoving horizon size is identical to a universe which is
evolving toward flatness,
d
|Ω − 1| < 0.
d ln a
(3)
Inflation therefore solves the horizon and flatness problems of the standard Big Bang with a single mechanism:
accelerated expansion.
However, inflation is not the only way to accomplish
this goal, as can be seen from the fact that the acceleration ä appears nowhere in the conservation law (2). To
solve both the horizon and flatness problems, it is sufficient to have a shrinking comoving Hubble length. One
way to do this is accelerated expansion, but another is
to have a collapsing universe, H = (ȧ/a) < 0. A collapsing, matter- or radiation-dominated universe also has a
shrinking comoving Hubble length, which will generate
perturbations in a manner similar to inflation. This is
the mechanism used by the Ekpyrotic scenario [7] to construct a cosmology consistent with observations. It is also
possible to decouple the horizon and flatness problems,
for example in theories with a varying speed of light, so
that the causal horizon is much larger than the Hubble
length [8]. Such theories can in principle solve the horizon
problem, but not the flatness problem, since the conservation law (2) is violated. It is also possible to solve the
horizon problem by a universe which is much older than
a Hubble time as in string gas cosmology [11] or island
cosmology [9, 10], or by the inclusion of extra dimensions [12, 13]. However, it has been argued that inflation
and ekpyrosis are the only mechanisms for generating a
scale-invariant spectrum of perturbations [7, 14].
2
In this paper, we discuss a method of solving the cosmological horizon problem and seeding scale-invariant
primordial perturbations in a cosmology with decelerating expansion and a corresponding growing comoving
Hubble horizon. The key to implementing such a model
is the fact that curvature perturbations are not generated at the Hubble horizon, but at the acoustic horizon determined by the speed of sound of a scalar field.
For canonical field theories, the two are identical, but
for non-canonical field theories, they are not. If one
has a decaying, superluminal sound speed, curvature perturbations can be generated outside the Hubble horizon
without inflation. We propose the term tachyacoustic
for such cosmologies, which are closely related to varying
speed of light theories. This idea has some history: such
cosmologies were first proposed by Armendariz-Picon in
the context of modified dispersion relations [15], and the
generation of perturbations in such cosmologies was further considered by Piao [16]. The idea re-emerged in the
context of varying speed of light theories by Magueijo
[17], and non-canonical Lagrangians by Magueijo [18]
and Piao [19]. In this paper, we outline a general approach to such cosmologies based on the generalization of
the inflationary flow formalism [20] introduced by Bean,
et al. for the case of arbitrary Lagrangians [21]. We
find that there is a class of Lagrangians with the necessary properties for tachyacoustic cosmologies, and discuss
two interesting examples. We find that it is straightforward to generate nearly scale-invariant perturbations for
these Lagrangians, and show that they have the property of reducing to instanton-like solutions with infinite
sound speed on the initial-time boundary of the spacetime. We speculate that this property may allow a
self-consistent description of tachyacoustic cosmologies
within a Wheeler-deWitt description of quantum cosmology. Finally, we show that such models are causally selfconsistent, and argue that they form a viable class of
alternatives to inflation.
II.
TACHYACOUSTIC COSMOLOGY
In this paper, we consider a way of generating scaleinvariant superhorizon cosmological perturbations based
on non-canonical scalar field Lagrangians with a speed of
sound faster than the speed of light, cS > 1. If the universe is dominated by a scalar field with speed of sound
cS , the relevant horizon for the generation of density perturbations is not the Hubble horizon dH ≃ (aH)−1 but
the acoustic horizon,
cS
DH ≃
.
(4)
aH
Mode freezing at the acoustic horizon is well-known in
non-canonical inflation models, for example k-Inflation
[22] and DBI inflation [23]. In non-canonical inflation
models, the Hubble horizon and the acoustic horizon are
both shrinking in comoving units, resulting in the generation of density perturbations at the acoustic horizon
and gravitational wave perturbations at the Hubble horizon [24]. However, the comoving Hubble horizon need
not be shrinking to generate curvature perturbations: all
that is required is that the acoustic horizon be shrinking, dDH /d ln a < 0. In this case, if curvature perturbations are to be generated on scales larger than the Hubble horizon, it is necessary that the acoustic horizon be
larger than the Hubble horizon, which requires a speed
of sound greater than the speed of light. Such theories
were studied recently by Babichev et al. [25, 26], who
showed that k-essence theories with cS > 1 are causally
self-consistent (see Appendix A), and can be mapped to
bimetric theories with two “light cones”, one given by
the Hubble horizon, and the other given by the acoustic horizon, which can be larger than the Hubble horizon without the presence of closed timelike loops. This
opens the possibility that one can construct a decelerating cosmology which nonetheless generates perturbations on super-Hubble scales via a superluminal, shrinking acoustic cone.
To explicitly construct such a model, consider a DiracBorn-Infeld (DBI) Lagrangian,
q
L (φ, X) = −f −1 (φ) 1 − f (φ) φ̇2 + f −1 (φ) − V (φ) ,
(5)
where we take the metric to have negative signature
ηµν = diag.(1, −1, −1, −1), so that the kinetic term for
the field is positive,
φ̇2 ≡ g µν ∂µ φ∂ν φ > 0.
(6)
Such Lagrangians often arise in string theory, for example
in the case of DBI inflation. For the moment, we will not
attempt to make a connection with string theory, but will
take the Lagrangian (5) as a phenomenological ansatz.
Take the background spacetime to be a flat, FriedmannRobertson-Walker (FRW) space, gµν = a2 ηµν , so that
the scale factor evolves as
Z
a ∝ exp
Hdt ∝ e−N ,
(7)
where we define the number of e-folds N as1
Z
N ≡ − Hdt.
(8)
There is a class of exact solutions [27, 28] to the equation
of motion for the field φ characterized by two dimensionless flow parameters ǫ and s, where
ǫ≡
1 dH
= const.,
H dN
(9)
and
s≡−
1
1 dcS
= const.
cS dN
(10)
We use the usual convention that N → ∞ corresponds to early
time, and N → −∞ corresponds to late time.
3
The parameter ǫ has its usual interpretation in terms of
the equation of state of the scalar field,
2
p=ρ
ǫ−1 .
(11)
3
For ǫ = const., the scale factor evolves as a power-law,
a ∝ t1/ǫ , so that the expansion is accelerating (i.e. inflation) for ǫ < 1. The speed of sound evolves as
q
cS = 1 − f (φ) φ̇2 ∝ e−sN ,
(12)
III.
FLOW HIERARCHY FOR GENERAL
K-ESSENCE MODELS
We now generalize the discussion in the last section
to an arbitrary k-essence model. Consider a general
Lagrangian of the form L = L [X, φ], where 2X =
g µν ∂µ φ∂ν φ is the canonical kinetic term (X > 0 according to our choice of the metric signature). The energy
density ρ and pressure p are given by
p = L (X, φ) ,
ρ = 2XLX − L.
and the Hubble parameter evolves as
ȧ
H = ∝ eǫN .
a
(13)
The parameter ǫ is a positive-definite quantity for p ≥
−ρ, so that the Hubble constant always decreases with
expansion. In contrast, the parameter s can take either
sign, with s > 0 corresponding to a sound speed which
increases with expansion, and s < 0 corresponding to an
decreasing sound speed. (See Ref. [28] for a detailed
derivation of this solution.) The important dynamics for
the generation of perturbations is the time evolution of
the corresponding horizons in comoving units. The comoving Hubble horizon evolves as
−1
dH ∝ (aH)
(1−ǫ)N
∝e
∝ τ,
The speed of sound is given by
−1
LXX
pX
,
= 1 + 2X
c2S ≡
ρX
LX
Therefore the condition for a shrinking acoustic horizon,
1 − ǫ − s > 0, is not identical to accelerated expansion.
For ǫ > 1 and s < 1−ǫ, the expansion is non-inflationary,
the Hubble horizon is growing in comoving units, and the
acoustic horizon is shrinking. The initial singularity is at
τ = 0, and we see immediately that for the tachyacoustic
solution, the speed of sound in the scalar field is infinite at the initial singularity, and the acoustic horizon
is likewise infinite in size. Therefore, such a cosmology
presents no “horizon problem” in the usual sense, since
even a spatially infinite spacetime is causally connected
on the initial-time boundary. Furthermore, unlike in the
case of inflation, there is no period of reheating necessary, since the cosmological evolution can be radiationdominated throughout and the cosmic temperature is not
driven exponentially to zero.
In the next section, we use the the generalized flow
function approach of Bean, et al. [21] to construct a
class of Lagrangians with solutions of the type outlined
above, with constant flow parameters. In these solutions,
the scale-factor evolves as a power-law in time and the
equation of motion for curvature perturbations can be
solved exactly, which we discuss in Sec. VI.
(18)
where the subscript “X” indicates a derivative with respect to the kinetic term. Throughout this section, unless
otherwise stated, we will follow closely Bean et al. [21].
We define the first three flow parameters as derivatives
with respect to the number of e-folds, dN = −Hdt:2 .
1 dH
,
H dN
1 dcS
s ≡ −
,
cS dN
1 dLX
.
s̃ ≡
LX dN
ǫ ≡
(14)
where τ is the conformal time. The Hubble horizon is
shrinking in comoving units for ǫ < 1, which is identical
to accelerated expansion, and is the usual condition for
inflation. The acoustic horizon behaves as
cS
DH ∝
∝ e(1−ǫ−s)N ∝ τ (1−ǫ−s)/(1−ǫ) .
(15)
aH
(16)
(17)
(19)
(20)
(21)
The Friedmann equation can be
√ written in terms of the
reduced Planck mass MP = 1/ 8πG
H2 =
1
1
ρ=
(2XLX − L) ,
3MP2
3MP2
(22)
and the continuity equation is
ρ̇ = 2H Ḣ = −3H (ρ + p) = −6HXLX .
(23)
For monotonic field evolution, the field value φ can be
used as a “clock”, and all other quantities expressed as
functions of φ, for example X = X (φ), L = L [X (φ) , φ],
and √
so on. We consider the homogeneous case, so that
φ̇ = 2X. Using
√
d
d
d
= φ̇
= 2X ,
dt
dφ
dφ
(24)
we can re-write the Friedmann and continuity equations
as the Hamilton Jacobi equations,
φ̇ =
√
2X = −
2MP2 ′
H (φ),
LX
(25)
2
3MP2 H 2 (φ) =
2
4MP4 H ′ (φ)
− L.
LX
(26)
The parameters s and s̃ correspond to the parameters κ and κ̃
in Bean, et al. [21]
4
where a prime denotes a derivative with respect to the
field φ. The number of e-folds dN can similarly be rewritten in terms of dφ by:
dN ≡ −Hdt
H
= −√
dφ
2X
LX
H (φ)
=
dφ.
2MP2 H ′ (φ)
(27)
(28)
The flow parameters ǫ, s, and s̃ (19) can therefore be
written as derivatives with respect to the field φ as:
2
2MP2 H ′ (φ)
,
LX
H (φ)
2MP2 H ′ (φ) c′S (φ)
s (φ) = −
,
LX H (φ) cS (φ)
2MP2 H ′ (φ) L′ X
.
s̃ (φ) =
LX H (φ) LX
(30)
(31)
Taking successive derivatives d/dN with respect to the
number of e-folds yields an infinite hierarchy of flow equations [20, 21],
= −ǫ (2ǫ − 2η̃ + s̃) ,
The simplest way to solve the flow equations derived in
the preceding section is to take all of the flow parameters
to be constant,
ds
ds̃
dℓ λ
dℓ α
dℓ β
dǫ
=
=
=
=
=
= 0.
dN
dN
dN
dN
dN
dN
Then, from (19-21) we easily find the following relations:
H ∝ eǫN ,
cS ∝ e−sN ,
LX ∝ es̃N ,
L′ X
3/2
LX
=
s̃
√ = const.,
MP 2ǫ
(37)
with solution
= −s̃ (ǫ − η̃ + 2s̃) + ǫ1 β,
= − λ [ℓ (s̃ + ǫ) − (ℓ − 1) η̃] +
ℓ+1
8ǫ
LX (φ) = 2
s̃
λ,
= −ℓ α [(ℓ − 1) (ǫ − η̃) + ℓs̃ + s] + ℓ+1 α,
= −ℓ β [(ℓ − 1) (ǫ − η̃) + (ℓ + 1) s̃] + ℓ+1 β, (32)
2MP2 H ′′ (φ)
λ=
,
LX H (φ)
ℓ−1
ℓ
2MP2
H ′ (φ)
ℓ
λ (φ) =
LX
H (φ)
ℓ−1
′
2 ℓ
H
(φ)
2M
ℓ
P
α (φ) =
LX
H (φ)
ℓ−1
′
2 ℓ
2MP
H (φ)
ℓ
β (φ) =
LX
H (φ)
(35)
The first two are identical to the DBI case, Eqs. (12)
and (13), but in the fully general case LX evolves independently of cS . It is straightforward to verify that the
full flow hierarchy (33) reduces to an exactly solvable set
of algebraic equations, with the higher-order parameters
expressed in terms of ǫ, s, and s̃. We can use the relations (29,30,31) to solve for H (φ), cS (φ), and LX (φ) as
follows: from Eqs. (29,31), we have
√ L′
2MP2 H ′ L′X
s̃ =
= Mp 2ǫ X = const.
(36)
LX
H LX
LX
= −s (ǫ − η̃ + s̃ + s) + ǫρ,
ℓ
(34)
We then have a differential equation for LX ,
= −η̃ (ǫ + s̃) + 2 λ,
where the higher-order flow parameters are defined as
follows, where ℓ = 1, . . . ∞ is an integer parameter:
η̃ (φ) =
COSMOLOGICAL SOLUTIONS FOR
CONSTANT FLOW PARAMETERS
(29)
ǫ (φ) =
dǫ
dN
dη̃
dN
ds
dN
ds̃
dN
dℓ λ
dN
dℓ α
dN
dℓ β
dN
IV.
Solutions to this infinite hierarchy of flow equations are
equivalent to solutions of the scalar field equation of motion. In the next section, we specialize to the case where
the flow parameters are constant, which results in an exactly solvable system.
MP
φ
2
,
(38)
where the integration constant has been absorbed into a
field redefinition. From Eq. (35), the field then evolves
as
φ2 ∝ e−s̃N ,
(39)
so that the direction of the field evolution depends on the
sign of s̃,
1
1
dℓ+1
H (φ) ,
H (φ) dφℓ+1
1
dℓ+1 −1
c (φ),
dφℓ+1 S
c−1
S (φ)
1 dℓ+1
LX . (33)
LX dφℓ+1
dφ
s̃
= − dN.
φ
2
Equation (29) then reduces to
′ 2
H
4ǫ2
= 2 2,
H
s̃ φ
(40)
(41)
with solution
H ∝ φ±2ǫ/s̃ .
(42)
The sign ambiguity can be resolved by requiring that the
universe be expanding, dH/dN > 0, so that
H ∝ φ−2ǫ/s̃ ∝ eǫN .
(43)
5
Finally, we solve for the speed of sound using Eq. (30),
which reduces to
c′S
2s
= const.,
(44)
=
cS
s̃
with solution
cS ∝ φ2s/s̃ .
(45)
Since our choice of N = 0 corresponds to an arbitrary
renormalization of the scale factor a ∝ e−N , we can without loss of generality define cS = 1 at N = 0, so that the
general solution for the background evolution is given by
2
8ǫ MP
,
(46)
LX = 2
s̃
φ
H (φ) = H0
cS (φ) =
φ
φ0
φ
φ0
−2ǫ/s̃
2s/s̃
,
,
where the field evolves as
φ
= e−s̃N/2 .
φ0
(47)
(48)
(49)
We can derive the time dependence of the scale factor
using the Hamilton-Jacobi equation (25),
√
s̃
(50)
φ̇ = H (φ) φ = 2X,
2
so that the kinetic term can be written as
X (φ) =
s̃2
2
8 H
(φ) φ2 .
(51)
Integrating expression (50) gives
1
,
(52)
ǫt
so that the scale factor evolves as a power-law in time,
consistent with the relation (9) between ǫ and the equation of state w = p/ρ,
H (t) =
a (t) ∝ t1/ǫ = t2/3(1+w) .
(53)
Radiation-dominated evolution therefore corresponds to
ǫ = 2, and matter-dominated evolution corresponds to
ǫ = 3/2. Inflation corresponds to ǫ < 1. The comoving
Hubble horizon evolves proportional the to the conformal
time,
dH ∝ (aH)−1 ∝ e(1−ǫ)N ∝ τ,
(54)
and the acoustic horizon evolves as
cS
DH ∝
∝ e(1−ǫ−s)N ∝ τ (1−ǫ−s)/(1−ǫ) ,
(55)
aH
identically to the DBI case discussed in Sec. II. For ǫ >
1, the acoustic horizon is shrinking in comoving units
for s < 1 − ǫ. Note that this behavior is independent
of the parameter s̃, which determines the form of the
Lagrangian, as we discuss in the next section.
V.
RECONSTRUCTING THE ACTION
In the past two sections we have solved the flow hierarchy for a model characterized by constant flow parameters, which allowed us to solve for H (φ), cS (φ), and
LX (φ); only the derivative of the Lagrangian with respect to the kinetic term X is determined. Therefore
this solution corresponds not to a single action but a
class of actions. In this section we derive a general equation for Lagrangians in this class, and discuss two specific
examples.
From Eqs. (38) and (48), we see that the speed of
sound cS can be written in terms of LX
−2s/s̃
c2S = C −1 LX
−1
LXX
= 1 + 2X
,
LX
(56)
where we have used Eq. (18), and defined
C≡
s̃2 φ20
8MP2 ǫ
2s/s̃
.
(57)
The result is a differential equation for the function
L (X, φ):
2XLXX + LX − CLnX = 0,
(58)
where we have defined
n≡1+
2s
.
s̃
(59)
Therefore, by specifying a relationship between the parameters s and s̃, we can construct a Lagrangian as the
solution to the differential equation (58). For example, a
canonical Lagrangian with speed of sound cS = const. =
1 is just the case s = 0, so that n = 1 and C = 1, and
Eq. (58) becomes
LXX = 0,
(60)
L = f (φ) X − V (φ) .
(61)
with general solution
Here f (φ) and V (φ) are free functions which arise from
integration of the second-order equation (58). The function f (φ) can be eliminated by a field redefinition dϕ =
p
f (φ)dφ, resulting in a manifestly canonical Lagrangian
for ϕ, as we would expect from setting cS = 1. A canonical Lagrangian can support inflationary solutions, but
not tachyacoustic solutions, and is therefore not of interest here. However, other choices of n do yield tachyacoustic solutions, and we focus on two such choices:
1. n = 0: A Cuscuton-like model.
2. n = 3: A DBI model.
We discuss each case separately below.
6
A.
The Hamilton-Jacobi Equation (25) becomes:
n = 0: A Cuscuton-like model
The case n = 0 corresponds to s̃ = −2s in (59), with
solution
√
(62)
L (X, φ) = 2f (φ) X + CX − V (φ) .
This Lagrangian is similar to a “cuscuton” Lagrangian
[36], with the addition of a term proportional to X.
Unlike the original Cuscuton model, which represents a
causal field with infinite speed of sound, the solution obtained here is valid for the general case, in which the
speed of sound can be finite. A similar cuscuton-like Lagrangian was considered in Ref. [19].
As in the canonical case, the functions f (φ) and V (φ)
are free functions resulting from integrating Eq. (58).
Unlike the canonical case, however, neither can be removed by a field redefinition. However, both functions
are fully determined by our choice of solution with ǫ, s,
and s̃ constant. Differentiating Eq. (62) with respect to
X gives
2
f (φ)
2ǫ MP
LX = √ + C = 2
,
(63)
s
φ
X
where the right hand side is the solution (46). Then
#
2 " 2
√
φ0
MP
2ǫ
−1
f (φ) = X
s2
φ0
φ
2
√
2
2ǫ
MP
cS (φ) − 1
(64)
= X
2
s
φ0
where for 2s̃ = −s, the expression (48) for the speed of
sound becomes
φ0
cS (φ) =
.
(65)
φ
The Lagrangian (62) can then be written as
L=X
2ǫ
s2
MP
φ0
2
2
2cS (φ) − 1 − V (φ) .
(66)
The Hubble parameter (47) is given by
H (φ) = H0
φ
φ0
ǫ/s
,
(67)
and we can then express the kinetic term as a function
of φ using Eq. (51):
s2 φ20 H 2 (φ)
s2
,
X (φ) = H 2 φ2 =
2
2 c2s (φ)
3MP2 H 2 = 2MP2 ǫH 2 − L
M 2 ǫH 2
= V (φ) + P 2 ,
cS
and we have an expression for the potential V (φ),
ǫ
2
2
V (φ) = MP H (φ) 3 − 2
.
(71)
cS (φ)
The Hubble parameter H (φ) and the speed of sound
cS (φ) are given by Eqs. (67) and (65), respectively. For
φ/φ0 ≪ 1, the speed of sound is much greater than the
speed of light, cS ≫ 1, and the potential is approximately
V (φ) ≃ 3MP2 H 2 (φ) = 3MP2 H02
φ
φ0
2ǫ/s
,
(72)
which can be recognized as a slow-roll-like solution dominated by the potential H 2 ≃ V 2 /3MP2 . For s < 0, the
field is rolling away from the origin, and for s < 1 − ǫ the
comoving acoustic horizon is shrinking and the solution
is tachyacoustic.
B.
n = 3: The DBI model
The case n = 3, corresponds to s̃ = s; then, from (30)
and (31), we find LX = c−1
S . Eq. (59) is then
c2S =
1
,
CL2X
(73)
so that we can take C = 1 without loss of generality.
Therefore, the Lagrangian assumes the well-known DBI
form,
p
L (X, φ) = −f −1 (φ) 1 − f (φ) X + f −1 (φ) − V (φ) .
(74)
The DBI model with constant flow parameters is extensively discussed in Ref. [28], and the reader is referred
this paper for further details. For ǫ and s constant, the
functions V and f are fully determined and are given by
2ǫ
1
V (φ) = 3MP2 H 2 (φ) 1 −
,
3 1 + cS (φ)
1
1 − c2S (φ)
f (φ) =
.
(75)
2
2MP ǫ H 2 (φ) cS (φ)
The Hubble parameter and speed of sound are given by:
(68)
The Lagrangian (62) can then be written entirely as a
function of the field φ,
1
L = MP2 ǫH 2 (φ) 2 − 2
− V (φ) .
(69)
cS (φ)
(70)
H (φ) = H0
cS (φ) =
φ
φ0
−2ǫ/s
,
(76)
and
φ
φ0
2
.
(77)
7
DBI Lagrangians allow for either inflationary or tachyacoustic evolution [18], depending on the values of ǫ and
s. Note that for cS > 1, the function f is negative, which
has consequences for embedding such a model in string
theory, which we discuss in Sec. VII.
In this section, we have explicitly constructed Lagrangians, including fully determined potentials, for
which the flow parameters are constant and the background evolution can be solved exactly. For suitable
choices of the flow parameters, the evolution is tachyacoustic, i.e. with a growing comoving Hubble horizon and
a shrinking comoving acoustic horizon. In the next section, we discuss the generation of curvature perturbations
at the acoustic horizon and show that such perturbations
are nearly scale-invariant, consistent with observation.
VI.
COSMOLOGICAL PERTURBATIONS FOR
CONSTANT FLOW PARAMETERS
We can deal with cosmological perturbations in this
general k-essence model with constant flow parameters
in the same way as performed in [28]. Following the approach of Garriga and Mukhanov [24] we start with the
perturbed Einstein equations,
d δφ
2M 2 c2
= Φ + 2 P S ∇2 Φ
dt φ̇
a (ρ + p)
a(ρ + p) δφ
d
,
(78)
(aΦ) =
dt
2MP2
φ̇
where Φ is the Bardeen potential and δφ is the perturbation of the field φ. Equations (78) can be cast into a
more convenient form by changing the perturbations Φ
and δφ to the new variables ζ and ξ defined by
ξ =
2MP2 Φa
H
δφ
+ Φ,
ζ = H
φ̇
(79)
so that the perturbed Einstein equations (78) become
a(ρ + p)
ξ˙ =
ζ,
H2
2
2
c H
∇2 ξ.
ζ̇ = 3 S
a (ρ + p)
(80)
As usual, we introduce a new variable z and the gaugeinvariant Mukhanov potential u as
z=
a(ρ + p)1/2
,
cS H
u = zζ;
(81)
then, from (80) we derive the mode equation for u(τ ) ∝
uk (τ ) exp(ik · x), given by
z ′′
2
′′
uk = 0,
(82)
uk − (cS k) +
z
where a prime denotes a derivative with
respect to con
formal time, ds2 = a2 (τ ) dτ 2 − dx2 . It is easy to show
that the variable z, defined by (81) can be cast into the
following form,
√
aMP 2ǫ
z=−
;
(83)
cS
then, using
d
d
= −aH
,
dτ
dN
(84)
we can evaluate the ratio z ′′ /z in (82) in terms of the
flow parameters (29-33); the result is
where
z ′′
= a2 H 2 F̄ ǫ, η̃, s, s̃, 2 λ, 1 α, 1 β ,
z
(85)
3
5
F̄ ≡ 2 + 2ǫ − 3η̃ − 3s + s̃ + 2ǫ2 + s̃2 − 2ss̃
2
4
5
+ η̃ 2 + 2ǫ(s̃ − s) + 3η̃s − η̃s̃ − 4η̃ǫ + 2 λ
2
1 1
− ǫ α + ǫ 1β .
(86)
2
Next, it is convenient to change the conformal time, τ ,
to the ratio of wavenumber to the sound horizon,
y≡
cS k
;
aH
(87)
then, conformal time derivatives switch to
d
d
= −aH (1 − ǫ − s) y ,
dτ
dy
(88)
and
2
d
d2
2 2 d
2 2
1
,
= a H (1 − ǫ − s) y
+ Ḡ ǫ, η̃, s, s̃, α y
dτ 2
dy 2
dy
(89)
where
Ḡ ≡ −s + ǫ(2s + s̃) + s(2s + s̃) + 2ǫ2 − 2ǫη̃ − sη̃
− ǫ 1α .
(90)
It is important to stress that the functions F and G
derived above hold in general; they reduce to the well
known expressions in the DBI limit [28], which, in this
case, s = s̃ and 1 α = 1 β = ρ. Substituting (85) and (89)
into the mode equation (82), we find
2
(1 − ǫ − s) y 2
d 2 uk
duk 2
+ Ḡy
+ y − F̄ uk = 0, (91)
2
dy
dy
which is an exact equation, without any assumption of
slow-roll.
In the case where the flow parameters are constant,
we can use the differential equations (32) to reduce the
number of independent parameters. We have
η̃ =
1
(2ǫ + s̃) ,
2
8
1
(2ǫ + s̃) (ǫ + s̃) ,
2
s
1
α =
(2s + s̃) ,
2ǫ
2
3s̃
1
β =
;
2ǫ
2
whereas in the late-time behavior y → 0 the mode function behaves as
λ =
(92)
then, substituting these values into expressions (86) and
(90), we find, respectively,
3
1
9
F̄ = 2 − ǫ − 3s + s̃2 − ss̃ + ǫs − s2 ,
4
4
2
(93)
Ḡ = s(−1 + ǫ + s).
(94)
It is important to notice that F̄ is different from the corresponding expression found in the DBI case [28], since
the gauge-dependent s̃ comes into play. However, Ḡ is
identical to its DBI analog, and it is expected since basically it comes from the change of variables τ → y, which
depends solely on the parameters cS and H, and not on
LX . For constant flow parameters we can solve equation
(91) exactly, and the solutions are given by
1−ǫ
y
uk (y) = y 2(1−ǫ−s) c1 Hν(1)
1−ǫ−s
y
+ c2 Hν(2)
,
(95)
1−ǫ−s
(1)
(2)
where c1 and c2 are constants, and Hν , Hν are Hankel
functions of first and second kind, respectively. The order
ν of the Hankel function is given by
9 − 6ǫ − 12s + 9s̃2 − 3ss̃ + 4ǫs − 2s2 + ǫ2
ν2 =
; (96)
4(1 − ǫ − s)2
next, using (35), (84) and (88) we find that
cS ∝ y s/(ǫ+s−1) ;
(97)
then, imposing the Bunch-Davies vacuum c2 = 0, and
normalizing the mode amplitudes by means of the canonical quantization condition
u∗k
duk
du∗
i
− uk k =
,
dy
dy
cs k(1 − ǫ − s)
r
π
cs k
r
y
Hν
1−ǫ−s
y
1−ǫ−s
1
eiy/(1−ǫ−s) ,
2cs k
From (101) we can derive the expression for the scalar
spectral index ns . Using the definition of the power spectrum of curvature perturbations
PR (k) =
k 3 uk
2π 2 z
2
,
(102)
and substituting expressions (83) and (101) into (102),
we find
2
PR =
|f (ν)| H 2
8π 2 MP2 cS ǫ
(103)
at horizon crossing, where f (ν) is a constant given by
f (ν) = 2ν−3/2
Γ(ν)
(1 − ǫ − s)ν−1/2 ;
Γ(3/2)
(104)
then, from the definition of the scalar spectral index
ns − 1 ≡
d ln PR
,
d ln k
(105)
and using
d
=−
d ln k
1
1−ǫ−s
d
,
dN
(106)
we see that the spectral index ns assumes the form
ns = 1 −
2ǫ + s
,
1−ǫ−s
(107)
which does not depend on the gauge-dependent parameter s̃, and is identical to its DBI analog. This is expected
since the power spectrum evaluated at the horizon crossing, equation (103), depends solely on H and cS , whose
derivatives with respect to N are related to the gaugeinvariant flow parameters ǫ and s. The scale-invariant
limit is s = −2ǫ.
VII.
,
(99)
which differs from the DBI case only in the order of the
Hankel function (96). In the small wavelength limit y →
∞ the early-time behavior of uk will be identical to DBI
[28] for constant flow parameters
uk = √
Γ(ν)
y 1/2−ν
(1 − ǫ − s)ν−1/2 √
. (101)
Γ(3/2)
2cs k
(98)
we find
1
uk (y) =
2
|uk (y)| → 2ν−3/2
(100)
CONCLUSIONS
In this paper we have demonstrated that accelerated
expansion or a collapsing universe are not the only ways
to dynamically generate a scale-invariant spectrum of
superhorizon curvature perturbations. There is a third
way: a superluminal acoustic cone which is shrinking in
comoving coordinates. Curvature perturbations generated at the acoustic horizon are familiar from inflationary scenarios based on non-canonical Lagrangians such
as k-inflation and DBI inflation. Such non-canonical Lagrangians arise naturally in string theory. However, in
these scenarios, both the Hubble horizon and the acoustic
9
horizon are shrinking in comoving units, and the acoustic
horizon is typically smaller than the Hubble horizon, i.e.
cS < 1. It is natural to ask whether tachyacoustic models have a similar, natural stringy embedding, especially
since the DBI action (5) naturally admits tachyacoustic solutions. Such an embedding is nontrivial, however,
since the frequently considered case of a 3+1 dimension
d-brane evolving in a higher-dimensional throat is illdefined in the cS > 1 limit. To see this, consider the
full ten-dimensional metric of throat plus brane [29],
ds210 = h2 (r) ds24 + h−2 (r) dr2 + r2 ds2X5 .
(108)
The field φ is simply
related to the coordinate in the
√
throat r as φ = T3 r, where the brane tension T3 depends on the string scale ms and the string coupling gs
as [39]
T3 =
m4s
3
(2π) gs
.
(109)
The Lagrangian for the field φ can be shown to be of the
DBI form (5), where the inverse brane tension f (φ) is
given in terms of the warp factor h (φ) by
f (φ) =
1
.
T3 h4 (φ)
(110)
The problem is immediately evident: superluminal propagation cS > 1 requires f < 0, so that the factor h2 (φ)
appearing in the metric (108) is imaginary, and the metric is ill-defined. Therefore, although the DBI action itself admits tachyacoustic solutions, this limit does not
correspond to a well-defined string solution. It is not
clear whether or not string manifolds exist which selfconsistently admit solutions with cS > 1.
We calculate the scalar spectral index of perturbations
for tachyacoustic solutions, and find
ns = 1 −
2ǫ + s
.
1−ǫ−s
(111)
Unlike inflationary models, radiation-dominated tachyacoustic models do not require a period of explosive entropy production to transition to a “hot” Big Bang cosmology. The early universe must be scalar-field dominated, but the temperature of the universe is not driven
exponentially to zero, since the scalar has a radiation
equation of state at all times, and entropy density is conserved (for any radiation component with density ργ , the
ratio ρφ /ργ = const). The scalar field φ must eventually
decay to Standard Model degrees of freedom, but as long
as this happens before primordial nucleosynthesis, the
model will match observations. A slow or late decay of
φ into other degrees of freedom would also suppress the
production of unwanted relics such as monopoles or gravitinos. For radiation-dominated tachyacoustic expansion
with ǫ = 2, the spectral index is
n=1+
4+s
,
1+s
where we have s < −3 for a shrinking comoving acoustic
cone. For s < −4, the spectral index is blue, n > 1,
which is ruled out by observation. The WMAP 2σ limit
n = 0.96 ± 0.026 [5] corresponds to s = [−3.814, −3.959].
Since the Hubble horizon is growing in comoving units,
no gravitational wave modes are produced.
Tachyacoustic models are not a fully convincing alternative to inflation, since they solve only the horizon problem and not the flatness problem, and inflation solves
both at once. However, inflation has initial conditions
problems of its own, in particular the fact that the initial
inflationary “patch” must be larger than a horizon size
for inflation to start [37]. Furthermore it has been shown
that inflationary spacetimes are in general geodesically
past-incomplete [38]. The initial conditions for tachyacoustic cosmology are quite different than those for inflation due to the presence of a true “Big Bang” singularity
at zero time. However, in this limit, the sound speed is
infinite and the tachyacoustic solution approaches an instanton. To see this, examine the form of the DBI field
Lagrangian (5) near the τ = 0 boundary of a tachyacoustic spacetime. From Eq. (12), the cS → ∞ limit
corresponds to φ → ∞ and f (φ) φ̇2 → −∞, so that
(112)
where
φ̇
L → p − V (φ) ,
|f |
φ̇ ≡
p
g µν ∂µ φ∂ν φ.
(113)
(114)
From Eq. (75), the asymptotic behavior of V (φ) and
f (φ) are
V (φ) → 3MP2 H 2 ∝ φ−4ǫ/s ,
1 cS
f (φ) → −
∝ φ2(1+2ǫ/s) .
2MP2 ǫ H
(115)
The scale-invariant limit s = −2ǫ is especially interesting,
since
1
p
→ µ2 = const.,
|f |
(116)
and the Lagrangian takes the form
L → µ2 φ̇ − V (φ) ,
(117)
where V (φ) ∝ φ2 . This can be identified as exactly the
“cuscuton” Lagrangian, suggested by Afshordi, et al. as a
candidate for Dark Energy [36, 40, 41]. Similarly, the n =
0 solution considered in Sec. V approaches a cuscuton
on the initial boundary surface. The cuscuton is a nondynamical, instanton-like solution with infinite speed of
sound. Consider the action for the Lagrangian (117),
Z
i
√ h
Sφ =
d4 x −g µ2 φ̇ − V (φ)
Z
Z
Z
√
2
3 √
= µ
dtφ̇ d x −g − d4 x −gV (φ)
10
= µ2
Z
dφΣ (φ) −
Z
√
d4 x −gV (φ),
(118)
where Σ (φ) is the volume of a constant-φ hypersurface
in the spacetime. The classical solutions to the cuscuton
action are constant mean curvature hypersurfaces, analogous to soap bubbles [36]. It is interesting to speculate
that this property of the cuscuton action may provide a
self-consistent cosmological boundary condition, or (even
more speculatively) be useful as a solution to the cosmological flatness problem. A full analysis, however, would
require inclusion of the gravitational action and solution
in a Wheeler-De Witt framework, or perhaps an embedding of the model in string theory or an alternate gravity
theory such as Horava-Lifshitz [42, 43]. This is the subject of future work.
APPENDIX A: STABLE CAUSALITY
Since tachyacoustic cosmology deals with superluminal
propagation of perturbations, it is important to address
the issue of causality in this model. Babichev et. al. [26]
have discussed the conditions that must be fulfilled by a
general k-essence model with superluminal propagation
in order to avoid causal paradoxes (i.e., the presence of
closed causal curves - CCC). In this appendix we outline
the main ideas of this work and apply to our tachyacoustic model.
To begin with let us introduce some key definitions
[44]. Let gµν be a metric with Lorentzian signature defined on a given manifold M. Given a point p ∈ M,
let tµ be a timelike vector at p; then, from this timelike
vector we construct a second metric, g̃µν , related to the
background metric gµν by
g̃µν ≡ gµν − tµ tν ,
(A1)
which clearly has a Lorentzian signature. Then, the
spacetime (M, gµν ) is said to be stably causal if there
is a continuous timelike vector field tµ such that the
spacetime (M, g̃µν ) possesses no closed timelike curves.
The following theorem (8.2.2. in [44]) establishes the
necessary and sufficient conditions for a spacetime to be
stably causal:
A spacetime (M, gµν ) stably causal if and only if
there exists a differentiable function f on M such that
∇µ f is a past directed timelike vector field.
We can apply this theorem to k-essence models as
follows [26]. First, we must find the analog of the
induced metric (A1) for the case of k-essence models,
which can be obtained by means of the equation of
motion for a scalar field described by a Lagrangian
[1] A. H. Guth, Phys. Rev. D 23, 347 (1981).
L (X, φ),
G̃µν ∇µ ∇ν φ + 2XLXφ − Lφ = 0,
(A2)
where G̃ , called “effective” or “acoustic” metric is given
by
µν
G̃µν (φ, ∇φ) = LX g µν + LXX ∇µ φ∇ν φ.
(A3)
It is convenient to use the metric [26]
Gµν ≡
cS µν
G̃
L2X
(A4)
which is conformally equivalent to G̃µν , and hence, defines the same causal structure. The inverse metric G−1
µν
is given by
LX
2 LXX
G−1
≡
g
−
c
∇
φ∇
φ
;
(A5)
µν
µ
ν
µν
S
cS
LX
notice that it has the same form of (A1), since ∇µ φ is
timelike. Using this definition, we can now apply the
theorem stated above and check the stable causality of
k-essence models. Let t be time coordinate with respect
to the background metric (which is everywhere future
directed), which we take to be FRW. Since ∇µ t∇ν t = 1,
we have, using (A3) and (A4),
cS
LXX 2
µν
G ∇µ t∇ν t =
1+
(A6)
φ̇ ;
LX
LX
then, since for a homogeneous scalar field holds φ̇2 = 2X,
we have, from (18) and (A6) that
Gµν ∇µ t∇ν t =
1
> 0,
cS LX
(A7)
provided the Null Energy Condition (NEC) is satisfied,
that is, LX > 0. Therefore, t plays a role ofglobal time
for both spacetimes (M, gµν ) and M, G−1
µν , and then
the conditions of the theorem are fulfilled. Then, there
is no CCC in superluminal k-essence models built from
homogeneous scalar fields on a FRW background. Since
this is exactly the case of the models introduced in this
paper, we conclude that there are no causal paradoxes in
tachyacoustic cosmology.
ACKNOWLEDGMENTS
This research is supported in part by the National Science Foundation under grant NSF-PHY-0757693. DB
thanks the Brazilian agency CAPES for financial support.
[2] A. D. Linde, Phys. Lett. B 108, 389 (1982).
11
[3] A. Albrecht and P. J. Steinhardt, Phys. Rev. Lett. 48,
1220 (1982).
[4] J. Dunkley et al. [WMAP Collaboration], Astrophys. J.
Suppl. 180, 306 (2009) [arXiv:0803.0586 [astro-ph]].
[5] E. Komatsu et al. [WMAP Collaboration], Astrophys. J.
Suppl. 180, 330 (2009) [arXiv:0803.0547 [astro-ph]].
[6] W. H. Kinney, E. W. Kolb, A. Melchiorri and A. Riotto,
Phys. Rev. D 78, 087302 (2008) [arXiv:0805.2966 [astroph]].
[7] S. Gratton, J. Khoury, P. J. Steinhardt and N. Turok,
Phys. Rev. D 69, 103505 (2004) [arXiv:astroph/0301395].
[8] A. J. Albrecht and J. Magueijo, Phys. Rev. D 59, 043516
(1999) [arXiv:astro-ph/9811018].
[9] S. Dutta and T. Vachaspati, Phys. Rev. D 71, 083507
(2005) [arXiv:astro-ph/0501396].
[10] S. Dutta, Phys. Rev. D 73, 063524 (2006) [arXiv:astroph/0511120].
[11] R. H. Brandenberger, arXiv:0902.4731 [hep-th].
[12] G. D. Starkman, D. Stojkovic and M. Trodden, Phys.
Rev. D 63, 103511 (2001) [arXiv:hep-th/0012226].
[13] G. D. Starkman, D. Stojkovic and M. Trodden, Phys.
Rev. Lett. 87, 231303 (2001) [arXiv:hep-th/0106143].
[14] J. Khoury and F. Piazza, arXiv:0811.3633 [hep-th].
[15] C. Armendariz-Picon, JCAP 0610, 010 (2006)
[arXiv:astro-ph/0606168].
[16] Y. S. Piao, Phys. Rev. D 75, 063517 (2007) [arXiv:grqc/0609071].
[17] J. Magueijo, Phys. Rev. Lett. 100, 231302 (2008)
[arXiv:0803.0859 [astro-ph]].
[18] J. Magueijo, Phys. Rev. D 79, 043525 (2009)
[arXiv:0807.1689 [gr-qc]].
[19] Y. S. Piao, arXiv:0807.3226 [gr-qc].
[20] W. H. Kinney, Phys. Rev. D 66, 083508 (2002)
[arXiv:astro-ph/0206032].
[21] R. Bean, D. J. H. Chung and G. Geshnizjani, Phys. Rev.
D 78, 023517 (2008) [arXiv:0801.0742 [astro-ph]].
[22] C. Armendariz-Picon, T. Damour and V. F. Mukhanov,
Phys. Lett. B 458, 209 (1999) [arXiv:hep-th/9904075].
[23] E. Silverstein and D. Tong, Phys. Rev. D 70, 103505
(2004) [arXiv:hep-th/0310221].
[24] J. Garriga and V. F. Mukhanov, Phys. Lett. B 458, 219
(1999) [arXiv:hep-th/9904176].
[25] E. Babichev, V. F. Mukhanov and A. Vikman, JHEP
0609, 061 (2006) [arXiv:hep-th/0604075].
[26] E. Babichev, V. Mukhanov and A. Vikman, JHEP 0802,
101 (2008) [arXiv:0708.0561 [hep-th]].
[27] L. P. Chimento and R. Lazkoz, Gen. Rel. Grav. 40, 2543
(2008) [arXiv:0711.0712 [hep-th]].
[28] W. H. Kinney and K. Tzirakis, Phys. Rev. D 77, 103517
(2008) [arXiv:0712.2043 [astro-ph]].
[29] I. R. Klebanov and M. J. Strassler, JHEP 0008, 052
(2000) [arXiv:hep-th/0007191].
[30] A. G. Muslimov, Class. Quant. Grav. 7, 231 (1990).
[31] D. S. Salopek and J. R. Bond, Phys. Rev. D 42, 3936
(1990).
[32] J. E. Lidsey, A. R. Liddle, E. W. Kolb, E. J. Copeland,
T. Barreiro and M. Abney, Rev. Mod. Phys. 69, 373
(1997) [arXiv:astro-ph/9508078].
[33] W. H. Kinney, Phys. Rev. D 56, 2002 (1997) [arXiv:hepph/9702427].
[34] M. Spalinski, JCAP 0704, 018 (2007) [arXiv:hepth/0702118].
[35] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions, New York: Dover, 1972.
[36] N. Afshordi, D. J. H. Chung and G. Geshnizjani, Phys.
Rev. D 75, 083513 (2007) [arXiv:hep-th/0609150].
[37] T. Vachaspati and M. Trodden, Phys. Rev. D 61, 023502
(1999) [arXiv:gr-qc/9811037].
[38] A. Borde, A. H. Guth and A. Vilenkin, Phys. Rev. Lett.
90, 151301 (2003) [arXiv:gr-qc/0110012].
[39] J. E. Lidsey and I. Huston, JCAP 0707, 002 (2007)
[arXiv:0705.0240 [hep-th]].
[40] N. Afshordi, D. J. H. Chung, M. Doran and G. Geshnizjani, Phys. Rev. D 75, 123509 (2007) [arXiv:astroph/0702002].
[41] V. F. Mukhanov and A. Vikman, JCAP 0602, 004 (2006)
[arXiv:astro-ph/0512066].
[42] Y. S. Piao, arXiv:0904.4117 [hep-th].
[43] N. Afshordi, arXiv:0907.5201 [hep-th].
[44] R. M. Wald, “General Relativity,” Chicago, USA: Univ.
Pr. ( 1984)