You are on page 1of 5

Vorticity generation at second order in cosmological perturbation theory

Adam J. Christopherson1 , Karim A. Malik1 , and David R. Matravers2


1
Astronomy Unit, School of Mathematical Sciences, Queen Mary University of London,
Mile End Road, London, E1 4NS, United Kingdom
2
Institute of Cosmology and Gravitation, University of Portsmouth,
Dennis Sciama Building, Portsmouth, PO1 3FX, United Kingdom
(Dated: October 30, 2018)
We show that at second order in cosmological perturbation theory vorticity generation is sourced
by entropy gradients. This is an extension of Crocco’s theorem to a cosmological setting.
arXiv:0904.0940v3 [astro-ph.CO] 26 Jun 2009

PACS numbers: 98.80.Cq, 98.80.Jk, 47.10.-g Phys. Rev. D 79, 123523 (2009), arXiv:0904.0940v3

I. INTRODUCTION pressure perturbation for a fluid with this equation of


state as δP = ∂P ∂P
∂S δS + ∂ρ δρ, which can be written as
2
It is well known in fluid dynamics that vorticity gen- δP = δPnad + cs δρ , where cs is the adiabatic sound
eration is sourced by entropy gradients. This was first speed, defined as c2s ≡ P ′ /ρ′ , and we have defined
the
pointed out by Crocco in 1937 [1]. Despite vorticity be- non-adiabatic pressure perturbation δPnad ≡ ∂P ∂S ρ δS,

ing ubiquitous in nature, studies of vorticity in the early in addition to the pressure and energy density pertur-
universe and in cosmology have so far been rare. bations. For a detailed discussion of the non-adiabatic
At linear order in perturbation theory scalar and vec- pressure perturbation, see Ref. [18].
tor perturbations, classified according to their transfor-
mation behaviour on spatial 3-hypersurfaces, decouple Our main result, derived in detail below and given in
from each other. Scalar perturbations are much easier Eq. (3.6), can be written concisely as
to treat mathematically and play the dominant role in

structure formation on super-horizon scales. At linear ω2ij ∝ δρ1,[j δPnad1,i] , (1.1)
order vorticity, intrinsically of vectorial nature (in fluid
dynamics simply the curl of the fluid velocity) cannot that is, the gradients in the non-adiabatic pressure per-
be constructed from scalar quantities, and the vorticity turbation coupled to gradients in the density act as a
tensor constructed from vector perturbations is, in gen- source for vorticity at second order.
eral, sourced only by anisotropic stress and decays in its We use cosmological perturbation theory throughout,
absence [2, 3, 4, 5, 6]. 1 which will make the numerical implementation of the re-
Things are different though at second order in the sults straightforward and, compared to other approaches,
perturbations. Only recently, with second order cos- has the added benefit of actually being physically trans-
mological perturbation theory becoming mature (see parent. The Paper is organised as follows. In the
e.g. Refs. [8, 9, 10, 11, 12, 13, 14]), has the issue of next section we define and introduce the key variables.
vorticity generation beyond the standard linear order In Section III we define the vorticity tensor and give
begun to be addressed [15, 16, 17]. However, these its evolution. We discuss our results and conclude in
previous studies have been restricted to barotropic fluids Section IV. All the governing equations necessary to
with an equation of state P = P (ρ). We show that derive the results in Section III are given in the appendix.
allowing for a non-adiabatic pressure perturbation gives
qualitatively different results. In this Paper, we use conformal time, η, throughout,
denoting derivatives with respect to conformal time with
a prime. The scale factor is a, and the Hubble parameter
We focus here on the case of vorticity generation
is H, where H = a′ /a. Greek indices, µ, ν, λ run from
at second order in the perturbations, sourced only by
0, . . . , 3, while lower case Latin indices, i, j, k, take the
scalar and vector perturbations. We consider a perfect
value 1, 2, or 3. Covariant derivatives are denoted by a
fluid, namely a fluid whose energy-momentum tensor is
semi-colon, partial derivatives by a comma. The order
diagonal (the inclusion of anisotropic stress will source
of the perturbations is denoted with a subscript immedi-
vorticity already at first order in the perturbations).
ately after a perturbed quantity. We work in the uniform
Such a fluid has an equation of state P = P (ρ, S),
curvature gauge throughout.
where P is the pressure, ρ is the energy density and
S is the entropy of the fluid. We can expand the
II. DEFINITIONS

1 Note, however, that there are certain situations in which vorticity The Friedmann-Robertson-Walker (FRW) metric ten-
is not sourced by anisotropic stress at linear order. See Ref. [7] sor, up to and including second order perturbations, has
for an example of first order vorticity sourced by heat flux. in the uniform curvature gauge the components (see,
2

e.g. Ref. [14]) which gives the well known result that at first order,
in the absence of an anisotropic pressure source term,
g00 = −a2 (1 + 2φ1 + φ2 ) , |ω1ij ω1ij | ∝ a−2 during radiation domination, where
c2s = 1/3 [4]. Hence the vorticity remains zero in this
 
2 1
g0i = a B1i + B2i , case, if it is initially zero.
2
However, at second order we get a non-zero source
gij = a2 δij , (2.1) term for the vorticity evolution equation, assuming zero
anisotropic pressure.2 Even assuming zero first order vor-
where we have assumed a flat (K = 0) background with-
ticity, ω1ij = 0, the second order vorticity evolves accord-
out loss of generality. We neglect tensor perturbations,
ing to
which will add another source term to the momentum
conservation equation, but are beyond the scope of this ′
ω2ij − 3Hc2s ω2ij (3.6)
work [19]. Here a = a(η) is the scale factor, φ1 and φ2 are
δρ δP
 
the lapse functions at first and second order, respectively, 2a 1,[j nad1,i]
= 3HV1[i δPnad1,j] + .
and B1i and B2i , represent the shear in this gauge. All ρ0 + P0 ρ0 + P0
perturbed quantities are function of xµ . Note B1i and
B2i can be further split into scalar and divergence-free Thus, we see that the non-adiabatic pressure perturba-
vector parts [14], though this step is unnecessary here. tion gradients coupled to density perturbation gradients
The fluid four velocity, uµ , obeying the normalisation act as a source for second order vorticity. In the case
condition uµ uµ = −1, has components of a vanishing entropy perturbation, as is the case for
barotropic fluids, we recover the result of Ref. [17]. For
completeness, we give the full second order vorticity evo-
 
1 1 2 1 k
u0 = −a 1 + φ1 + φ2 − φ1 + v1k v1 , (2.2) lution equation without assuming ω1ij = 0 in the Ap-
2 2 2
  pendix as Eq. (A8).
1
ui = a V1i + V2i − φ1 B1i , (2.3)
2
IV. DISCUSSION AND CONCLUSIONS
up to second order, where v i is the spatial three velocity,
and V i is the covariant spatial velocity, defined as V i =
vi + B i . In this work we have shown that at second order in
the perturbations vorticity is generated from first order
scalar and vector perturbations for a perfect fluid. This
III. VORTICITY is an extension of Crocco’s Theorem to an expanding,
dynamical background, namely a FRW universe.
Whereas in previous works barotropic flow was as-
The vorticity tensor is defined as the projected anti-
sumed, allowing for entropy gives a qualitatively novel re-
symmetrised covariant derivative of the fluid four veloc-
sult. This implies that the description of the cosmic fluid
ity, that is [4]
as a potential flow, which works exceptionally well at first
order in the perturbations, will break down at second
ωµν = Pµα Pνβ u[α;β] , (3.1)
order for non-barotropic flows. Similarly in barotropic
flow Kelvin’s theorem guarantees conservation of vortic-
where u[α;β] ≡ 21 (uα;β − uβ;α ), and the projection tensor
ity. This is no longer the case if the flow non-barotropic.
Pµν into the instantaneous fluid rest space is given by
Entropy perturbations will arise in many settings,
Pµν = gµν + uµ uν . (3.2) here we only considered the case of a single fluid
with non-zero intrinsic entropy. Even in simple single
The vorticity can then be decomposed, up to second field inflation models, after the end of the slow-roll
order in perturbation theory, as ωij ≡ ω1ij + 12 ω2ij , where regime, there is a non-zero entropy perturbation (or
the first order part is simply non-adiabatic pressure; see, e.g., Ref. [18]). Further-
more, in the case of multiple fluids we get in addition
ω1ij = aV1[i,j] , (3.3) to the intrinsic entropies of the individual fluids also
an entropy component stemming from the mixing of
and the second order part is the fluids (the relative entropy perturbation in the
terminology of Ref. [14]), even if there is no energy and
. momentum transfer between the fluids. This relative
h i

ω2ij = aV2[i,j] +2a V1[i V1j] + φ1,[i (V1 + B1 )j] − φ1 B1[i,j]
(3.4)
Using the evolution equations given in Appendix A,
the first order part evolves as 2 The calculation of the vorticity evolution equation uses first
order evolution and field equations which we give in Appendix

ω1ij − 3Hc2s ω1ij = 0 , (3.5) A.
3

entropy perturbation can be another source of vorticity linear order in perturbation theory, only tensor pertur-
in the multi-fluid case. However in the cosmic plasma it bations (or gravitational waves) will produce B-mode
is likely that the relative entropy perturbation between polarisation since scalar perturbations only produce
fluids is subdominant compared to those mechanisms E-modes [30], and vector perturbations will become
discussed below. subdominant during inflation, and will decay with the
expansion of the universe. However, at second order the
Our result, that the vorticity is non-zero at second or- former is no longer true and, as we have shown above,
der in the presence of non-adiabatic or entropy pertur- density perturbations can generate vorticity. This could
bations, has immediate implications for the generation then produce B-mode polarisation large enough to be ob-
of magnetic fields in the early universe, since Biermann servable by future CMB experiments such as Planck [31].
showed that the generation of magnetic fields is related
to vorticity [20] (see also Harrison, Ref. [21]). Acknowledgments
Previous works either used momentum exchange
between multiple fluids to generate vorticity, as in The authors are grateful to Kishore Ananda, Chris
Refs. [15, 16, 22, 23, 24, 25, 26], or used intermediate Clarkson and Roy Maartens for useful discussions and
steps to first generate vorticity for example by using comments. AJC is supported by the Science and Tech-
shock fronts as in Ref. [27]. The vorticity generated in nology Facilities Council (STFC). We used the computer
the latter case was then used to source turbulence. As algebra package Cadabra [32] to obtain some of the con-
we have shown above, we do not require such extra steps. servation equations, and thank Kasper Peeters for useful
We will study the evolution of magnetic fields in the case discussions on using the package.
of non-zero entropy perturbations, and including tensor
perturbations, numerically in a forthcoming paper [19].
APPENDIX A
At first glance, one might expect that because the vor-
ticity is generated as a second order effect by products
Here we give the evolution and constraint equations
of first order density and entropy perturbations, the vor-
necessary to derive the results of Section III. For more
ticity will be of order 10−10 . However, the source term
details see Ref. [14].
is in fact constructed from gradients of the density and
entropy perturbations and, because of this, the effect can
readily be larger than O(10−10 ). In Fourier space, this
1. Energy-momentum conservation
source term will be ‘boosted’ by the wavenumbers on
small scales. Since our mechanism will manifest itself on
sub-horizon scales and hence the wavenumbers are large, In the background the energy conservation equation is
this has the potential to increase the effect. For example,
the authors of Ref. [28] recently showed, in the context ρ′0 = −3H (ρ0 + P0 ) . (A1)
of second order tensor perturbations, that on sub-horizon
At first order we have for the energy conservation equa-
scales this effect can ‘boost’ second order quantities to be
tion
of comparable amplitude as first order quantities. In fact,
the WMAP five year results results [29] give only a non- k
δρ′1 + 3H (δρ1 + δP1 ) + (ρ0 + P0 ) v1,k = 0 , (A2)
zero upper bound on entropy perturbations on horizon
scales. Since our effect is on sub-horizon scales, following and for the momentum conservation equation
the argument of WMAP5 and assuming a blue spectrum
for the entropy perturbation, it is likely that this effect h δP
1
i
V1i′ + H 1 − 3c2s V1i +

is non-negligible. + φ1 = 0 . (A3)
ρ0 + P0 ,i
One possible observational signature of vorticity in
the early universe is that of B-mode polarisation of the At second order we only need the momentum conserva-
cosmic microwave background (CMB) radiation. At tion equation

h i′ h i h i′
+ 4H (ρ0 + P0 ) V2i + δP2 + (ρ0 + P0 ) φ2 + 2 (δρ1 + δP1 ) φ1,i + 2 (δρ1 + δP1 ) V1i
(ρ0 + P0 ) V2i (A4)
,i
h i h   i
+8H (δρ1 + δP1 ) V1i − 2 (ρ0 + P0 ) φ′1 B1i + φ21 ,i − 2φ1 (ρ0 + P0 ) V1i′ + ρ′0 + P0′ + 4H (ρ0 + P0 ) V1i


k k k
+2 (ρ0 + P0 ) v1, k V1i + 2 (ρ0 + P0 ) v1 V1i,k − 2 (ρ0 + P0 ) v1k B1, i = 0 .

2. Field equations We only need the field equations up to first order in


the perturbations, a considerable simplification, since the
At zeroth order the evolution of the scale factor is gov-
erned by the Friedmann equation, given by
8πG 2
H2 = a ρ0 . (A5)
4

second order lapse function in Eq. (A4) above cancels 3. Vorticity


when calculating the vorticity evolution. The Einstein
constraint equations at first order are
k
2HB1,k + 6H2 φ1 = −8πGa2 δρ1 , (A6)
Using the definition of the vorticity tensor, Eq. (3.4)
and above, and the equations given in the previous subsec-
tion, we get the evolution equation for the vorticity ten-
∇2 B1i − B1,ki
k
− 4Hφ1,i = 16πGa2 (ρ0 + P0 )V1i . (A7) sor at second order

" ′ #
δP 1 + δρ1

ω2ij − 3Hc2s ω2ij +2 + k
V1,k − k
X1,k ω1ij (A8)
ρ0 + P0
k
+2 V1k − X1k ω1ij,k − 2 X1,j k k k
  
− V1,j ω1ik + 2 X1,i − V1,i ω1jk
 
a  1 
= 3H V1i δPnad1,j − V1j δPnad1,i + δρ1,j δPnad1,i − δρ1,i δPnad1,j ,
ρ0 + P0 ρ0 + P0

where X1i is given entirely in terms of matter perturba-


tions as
 
4πG 2 
X1i = ∇−2 a 3H(ρ0 + P0 )V1i − δρ1,i . (A9)
H

[1] L. Crocco, ZAMM 17, 1 (1937). [arXiv:astro-ph/0410687].


[2] A. R. Liddle and D. H. Lyth, Cosmological inflation and [16] T. Kobayashi, R. Maartens, T. Shiromizu and
large-scale structure, Cambridge, UK: Univ. Pr. (2000). K. Takahashi, Phys. Rev. D 75, 103501 (2007)
[3] J. M. Bardeen, Phys. Rev. D 22, 1882 (1980). [arXiv:astro-ph/0701596].
[4] H. Kodama and M. Sasaki, Prog. Theor. Phys. Suppl. [17] T. C. Lu, K. Ananda, C. Clarkson and R. Maartens,
78, 1 (1984). JCAP 0902, 023 (2009) [arXiv:0812.1349 [astro-ph]].
[5] A. Lewis, Phys. Rev. D 70 (2004) 043518 [18] A. J. Christopherson and K. A. Malik, Phys. Lett. B 675
[arXiv:astro-ph/0403583]. (2009) 159 [arXiv:0809.3518 [astro-ph]].
[6] L. Hollenstein, C. Caprini, R. Crittenden and [19] A. J. Christopherson, K. A. Malik, and D. R. Matravers,
R. Maartens, Phys. Rev. D 77 (2008) 063517 in preparation (2009)
[arXiv:0712.1667 [astro-ph]]. [20] L. Biermann, Z. Naturforsch. Teil A, 5, 65 (1950).
[7] R. Maartens and J. Triginer, Phys. Rev. D 56 (1997) [21] E. R. Harrison, Mon. Not. R. atr. Soc. 147 279 (1970).
4640 [arXiv:gr-qc/9707018]. [22] R. Gopal and S. Sethi, Mon. Not. Roy. Astron. Soc. 363,
[8] V. F. Mukhanov, L. R. W. Abramo and R. H. Bran- 521 (2005) [arXiv:astro-ph/0411170].
denberger, Phys. Rev. Lett. 78, 1624 (1997) [23] K. Takahashi, K. Ichiki, H. Ohno and
[arXiv:gr-qc/9609026]. H. Hanayama, Phys. Rev. Lett. 95, 121301 (2005)
[9] M. Bruni, S. Matarrese, S. Mollerach and [arXiv:astro-ph/0502283].
S. Sonego, Class. Quant. Grav. 14, 2585 (1997) [24] K. Ichiki, K. Takahashi, H. Ohno, H. Hanayama
[arXiv:gr-qc/9609040]. and N. Sugiyama, Science 311, 827–829, 2006,
[10] H. Noh and J. c. Hwang, Phys. Rev. D 69, 104011 (2004). [arXiv:astro-ph/0603631].
[11] K. A. Malik and D. Wands, Class. Quant. Grav. 21, L65 [25] E. R. Siegel and J. N. Fry, Astrophys. J. 651, 627 (2006)
(2004) [arXiv:astro-ph/0307055]. [arXiv:astro-ph/0604526].
[12] K. Nakamura, Prog. Theor. Phys. 113, 481 (2005) [26] S. Maeda, S. Kitagawa, T. Kobayashi and T. Shiromizu,
[arXiv:gr-qc/0410024]. arXiv:0805.0169 [astro-ph].
[13] K. A. Malik and D. R. Matravers, Class. Quant. Grav. [27] D. Ryu, H. Kang, J. Cho and S. Das, Science 320, 909–
25, 193001 (2008) [arXiv:0804.3276 [astro-ph]]. 912, 2008, arXiv:0805.2466 [astro-ph].
[14] K. A. Malik and D. Wands, Phys. Rept. 475 (2009) 1 [28] D. Baumann, P. J. Steinhardt, K. Takahashi
[arXiv:0809.4944 [astro-ph]]. and K. Ichiki, Phys. Rev. D 76 (2007) 084019
[15] S. Matarrese, S. Mollerach, A. Notari and [arXiv:hep-th/0703290].
A. Riotto, Phys. Rev. D 71, 043502 (2005) [29] E. Komatsu et al. [WMAP Collaboration], Astrophys. J.
5

Suppl. 180 (2009) 330 [arXiv:0803.0547 [astro-ph]]. [32] K. Peeters, Comput. Phys. Commun. 176 (2007) 550
[30] D. N. Spergel and M. Zaldarriaga, Phys. Rev. Lett. 79 [arXiv:cs/0608005], K. Peeters, arXiv:hep-th/0701238.
(1997) 2180 [arXiv:astro-ph/9705182].
[31] [Planck Collaboration], arXiv:astro-ph/0604069.

You might also like