<<

2

II. INDUCED GRAVITATIONAL WAVES

We start with deriving the expression for the energy density of induced GWs. The evolution equation for the induced GWs hij(η, ~x) is of the form

00 0 2 hij + 2Hhij − ∂ hij = Λij,klSkl, (1) where a dash denotes differentiation with respect to the conformal time η and Λij,kl is the projection which extracts the transverse-traceless part of the source term Skl(η, ~x). A more explicit expression for Λij,klSkl will be given shortly below. We consider a universe filled with a perfect fluid whose equation of state parameter w is a constant. The scale factor is then given by a ∝ η2/(1+3w) and hence H := a0/a = [2/(1 + 3w)]/η. The Fourier components of hij are defined by

Z 3 X d k (A) ~ h (η, ~x) = h (η,~k)e (~k)eik·~x (A = +, ×). (2) ij (2π)3/2 A ij A (A) ~ The two eij (k) are defined as

(+) ~ 1 h ~ ~ ~ ~ i e (k) = √ ei(k)ej(k) − ei(k)ej(k) , (3) ij 2 (×) ~ 1 h ~ ~ ~ ~ i e (k) = √ ei(k)ej(k) + ei(k)ej(k) , (4) ij 2 ~ ~ ~ where ei(k) and ej(k) are unit vectors orthogonal to each other and to k. It follows from the definition of the 0 i (A) (A) ~ (A ) ~ polarization tensors that k eij = 0 and eij (k)eij (k) = δAA0 . Using the polarization tensors one can write the right-hand side of Eq. (1) as

Z 3 X d k (A) (A) Λ S (η, ~x) = e (~k)e (~k)S (η,~k), (5) ij,kl kl (2π)3/2 ij lm lm A where 3 Z d x ~ S (η,~k) := S (η, ~x)e−ik·~x (6) ij (2π)3/2 ij

~ (A) ~ ~ is the Fourier transform of Sij(η, ~x). We write SA(η, k) := eij (k)Sij(η, k). The formal solution to Eq. (1) in the Fourier domain is given by Z η ~ 1 ~ hA(η, k) = Gk(η, η¯)a(¯η)SA(¯η, k)d¯η, (7) a(η) 0 where the Green’s function is expressed in terms of the Bessel functions as π G (η, η¯) = η1/2η¯1/2 [Y (kη)J (kη¯) − J (kη)Y (kη¯)] , (8) k 2 ν ν ν ν with 3(1 − w) ν := . (9) 2(1 + 3w) The energy density of GWs is given by

M 2 X 1 ρ (η) = Pl hh0 (η, ~x)h0 (η, ~x)i, (10) GW 2 a2 A A A where h· · · i denotes a spatial average and

3 Z d k ~ h (η, ~x) := h (η,~k)eik·~x. (11) A (2π)3/2 A 3

Since 3 3 0 Z d k d k ~ ~ 0 hh0 (η, ~x)h0 (η, ~x)i = h0 (η,~k)h∗ 0(η,~k0)hei(k−k )·~xi, (12) A A (2π)3/2 (2π)3/2 A A and 3 ~ ~ 0 1 Z ~ ~ 0 (2π) hei(k−k )·~xi = ei(k−k )·~xd3x = δ(3)(~k − ~k0), (13) V V where V is a volume whose size is much larger than the wavelengths of interest, we have

M 2 1 X Z ρ (η) = Pl d3k|h0 (η,~k)|2. (14) GW 2 V a2 A A

In the subhorizon regime, kη  1, the time derivative of hA is approximately given by Z η 0 ~ 1 ~ hA(η, k) ' ∂ηGk(η, η¯)a(¯η)SA(¯η, k)d¯η. (15) a(η) 0 We find the following approximate expression for the time derivative of the Green’s function, π ∂ G (η, η¯) 'G (η, η¯) := (kη)1/2(kη¯)1/2 [Y (kη)J (kη¯) − J (kη)Y (kη¯)] . (16) η k k 2 ν−1 ν ν−1 ν More explicitly, for ν = 1/2 (w = 1/3) we have

Gk(η, η¯) = cos[k(η − η¯)], (17) and for ν = 3/2 (w = 0) we have sin[k(η − η¯)] G (η, η¯) = cos[k(η − η¯)] + . (18) k kη¯ It then follows that M 2 1 X Z Z η Z η ρ (η) = Pl d3k dη0 dη00a(η0)a(η00)G (η, η0)G (η, η00)S (η0,~k)S∗ (η00,~k). (19) GW 2 V a4(η) k k A A A 0 0

We are interested in the case where SA is of the form Z d3q S (η,~k) = e(A)(~k) qiqjA(~q)A(~k − ~q)F (~k, ~q, η), (20) A ij (2π)3/2 where A(~q) is a Gaussian random field and F (~k, ~q, η) is some function. In the actual calculation, A(~q) will be the primordial amplitude of scalar perturbations and F contains the information of their time evolution. The power spectrum P(k) is defined by

2π2 hA(~k)A∗(~q)i = δ(3)(~k − ~q)P(k). (21) k3 (Now h· · · i denotes an average over the whole distribution.) Using Wick’s theorem, the two-point correlator of the source term SA can be written as Z 3 4 0 ~ ∗ 00 ~ 0 d q (A) ~ i j 2 8π (3) ~ ~ 0 ~ ~ 0 ~ 00 hSA(η , k)SA(η , k )i = (e (k)q q ) δ (k − k )P(q)P(|k − ~q|)F (k, ~q, η )F (k, ~q, η ). (22) (2π)3 ij q3|~k − ~q|3

Using this and [δ(3)(~k − ~k0)]2 = [V/(2π)3]δ(3)(~k − ~k0), one can write the energy density of GWs (19) as

2 Z 3 Z 3 Z η Z η 2MPl d k d q 0 00 0 00 0 00 ρGW(η) = 4 3 3 dη dη a(η )a(η )Gk(η, η )Gk(η, η ) a (η) (2π) (2π) 0 0 π4 × q4 sin4 θ P(q)P(|~k − ~q|)F (~k, ~q, η0)F (~k, ~q, η00), (23) q3|~k − ~q|3 4 where θ is the angle between ~q and ~k. The final step is to extract from the above expression the energy density parameter of GWs, ΩGW(η, k), defined by Z 2 2 ρGW(η) = 3MPlH ΩGW(η, k)d ln k. (24)

In practice, F depends on ~k, ~q, and η through qη and |~k − ~q|η. Therefore, it is convenient to introduce dimensionless variables u := |~k − ~q|/k and v := q/k. Using these variables, the energy density parameter is expressed as

Z ∞ Z 1+v Z η Z η 1 1 0 00 0 00 0 00 ΩGW(η, k) = 2 4 dv du dη dη a(η )a(η )Gk(η, η )Gk(η, η ) 3H a (η) 0 |1−v| 0 0 " #2 k4 v2 1 + v2 − u2 2 × 1 − P(ku)P(kv)F (u, v, kη0)F (u, v, kη00) 4 u2 2v 2 2 Z ∞ Z 1+v 2 "  2 2 2# k v 1 + v − u 2 = 2 dv du 2 1 − P(ku)P(kv)I (u, v, kη), (25) 12H 0 |1−v| u 2v where Z η k 0 0 0 0 I(u, v, kη) = dη a(η )Gk(η, η )F (u, v, kη ). (26) a(η) 0 The information of the initial conditions for scalar perturbations is encoded in the power spectrum P, while the time evolution of the perturbations determines the form of the integral I. The two distinct effects are thus separated. The gauge difference is essentially imprinted in I. This can be evaluated analytically for w = 0 and w = 1/3 [25] and numerically for the other values of w.

III. ACTION APPROACH TO INDUCED GRAVITATIONAL WAVES

To derive the gauge-ready form of the source term for induced GWs in a universe filled with an irrotational barotropic perfect fluid, it is convenient to employ the action-based approach, describing the fluid in terms of a shift-symmetric k-essence field. Our action is given by

Z √ M 2  S = d4x −g Pl R + P (X) , (27) 2

µν where X := −g ∂µφ∂ν φ/2. The k-essence field is equivalent to a cosmological perfect fluid whose energy density and pressure are given respectively by ρ = 2XPX − P and p = P . (Here and hereafter we write ∂P/∂X = PX .) Therefore, a w = p/ρ = const fluid can be mimicked by [26, 27]

P ∝ X(1+w)/2w. (28)

Hereafter we will assume that w ≥ 0. Note that the w = 0 case appears to be singular, but a careful inspection shows that the limit in fact makes sense [23]. The metric in the 3 + 1 Arnowitt-Deser-Misner (ADM) form is given by

2 2 2 i i  j j  ds = −N dt + gij dx + N dt dx + N dt . (29) We only fix the spatial coordinate, while leaving the temporal gauge degree of freedom unfixed, as we are interested only in the latter gauge difference. We therefore write the ADM variables in terms of the scalar and tensor perturbations as

2 −2ψ h N = 1 + δN, Ni = ∂iχ, gij = a e (e )ij, (30)

h where (e )ij = δij + hij + hikhkj/2 + ··· . The perturbed scalar field is written as φ = φ¯(t) + δφ. (31) 5

We will omit the bar from the background value if unnecessary. Using the temporal gauge degree of freedom one can eliminate one of χ, ψ, and δφ. The background equations are given by

2 2 3MPlH = 2XPX − P, (32) 2  2 ˙  MPl 3H + 2H = −P, (33) d   a3φP˙ = 0. (34) dt X

In order to derive the equations of motion for the perturbations, we substitute the metric (30) and the scalar field (31) to the action (27) and expand it to third order. At third order we only need the terms containing one tensor and two scalars, because the variation of such terms with respect to hij leads to the source terms for GWs induced by scalar perturbations. Thus, the action that suffices for our purpose is Z 3 h (2) (2) (3) i S = dtd x Ls + Lh + Lssh , (35) where

 ψ∂2ψ ∂2χ ∂2χ ∂2ψ  L(2) = a3M 2 −3ψ˙ 2 − − 2HδN − 2ψ˙ − 6HδNψ˙ + 2δN s Pl a2 a2 a2 a2 3 a  2  + a3 XP + 2X2P − 3M 2 H2 δN 2 + (P + 2XP ) δφ˙ − 2φδN˙ δφ˙ X XX Pl 2 X XX aP   − 3a3φP˙ ψδφ˙ + X δφ∂2δφ + 2φδφ∂˙ 2χ , (36) X 2 a3M 2  (∂ h )2  L(2) = Pl h˙ 2 − k ij , (37) h 8 ij a2 and

 1 1  L(3) = aM 2 2(HδN + ψ˙)χ h + (δN + 3ψ)χ h˙ − ψ ψ h + 2δN ψ h + ∂2 (χ χ ) h ssh Pl ,ij ij 2 ,ij ij ,i ,j ij ,i ,j ij 4a2 ,i ,j ij aP   + X h ∂ δφ∂ δφ + 2φ∂˙ χ∂ δφ . (38) 2 ij i j i j

(2) The quadratic Lagrangian Ls yields the linearized equations of motion for the scalar perturbations. The variation of the above action with respect to hij gives the equation of motion for hij sourced by the scalar perturbations, which takes the form of Eq. (1). Now it is straightforward to obtain

1 8 2 d S (η, ~x) = ∂2(χ χ ) + 8δN ψ − 4ψ ψ + (HδN + ψ0)χ − [a(δN + 3ψ)χ ] ij a2 ,i ,j (,i ,j) ,i ,j a ,ij a2 dη ,ij   4XPX 2H + 2 2 Q,iQ,j + χ(,iQ,j) , (39) MPlH a where we defined Q(t, ~x) := Hδφ/φ˙. Note that one can simplify the expression further by using the background 2 2 equation and write 4XPX /MPlH = 6(1 + w). This is the gauge-ready form of the source term for induced GWs. Moving to the Fourier domain, we have

Z d3q  k2  S (η,~k) = e(A)(~k) qiqj − χ(η, ~q)χ(η,~k − ~q) + 4δN(η, ~q)ψ(η,~k − ~q) + 4ψ(η, ~q)δN(η,~k − ~q) + ··· . A ij (2π)3/2 a2 (40)

This can be recast in the form of Eq. (20) by separating the scalar perturbations into the transfer functions and the primordial amplitudes. 6

IV. THE GAUGE DEPENDENCE

Now let us investigate the gauge dependence of induced GWs. Specifically, we consider the Newtonian gauge (χ = 0), the comoving gauge (δφ = 0), and the uniform curvature gauge (ζ = 0). The gauge dependence can be seen clearly by evaluating the integral I(u, v, kη). In radiation-dominated (RD) and matter-dominated (MD) universes this can be done analytically. For the other values of w, one needs to perform numerical integration to evaluate I precisely.

A. Newtonian Gauge

We start with reproducing the standard Newtonian gauge result. The Newtonian gauge is defined by

χ = 0. (41)

Following the conventional notation we write δN = Φ and ψ = Ψ. From the Lagrangian (36) we obtain

k2 3 3 Q0 3  Ψ00 + 2HΨ0 + (Ψ − Φ) + HΦ0 + (1 − w)H2Φ − (1 + w)H2 + (1 − w)Q = 0, (42) 3 2 2 H 2 3(1 − w) 3(1 + w) Q0 3  − 3HΨ0 − k2Ψ + H2Φ − H2 + (1 − w)Q = 0, (43) 2w 2w H 2 Q00 + H (2Q0 − Φ0 − 3wΨ0) + wk2Q = 0, (44) where we moved to the Fourier domain and used the conformal time. The solution to the above set of equations is given by

~ Φ = Ψ = AΦ(k)fΦ(η, k), (45) A (~k) Q = Φ [2f + (1 + 3w)ηf 0 ] , (46) 3(1 + w) Φ Φ where √  wkη −ν−1 √ f (η, k) := Γ(ν + 2) J ( wkη), (47) Φ 2 ν+1

~ AΦ(k) is the amplitude of Φ at η = 0, and we discarded another independent solution that diverges at η = 0. We thus have Φ = Ψ = AΦ and Q = 2AΦ/[3(1 + w)] at η = 0. Substituting the above result to the source term (40), one can compute F (u, v, kη). In a RD universe, we find

1 18 ukη  vkη  FRD,χ(u, v, kη) = cos √ cos √ 54 u2v2k4η4 3 3 √ 2 3 ukη  vkη  + u2k2η2 − 9 sin √ cos √ u3v2k5η5 3 3 √ 2 3 ukη  vkη  + v2k2η2 − 9 cos √ sin √ u2v3k5η5 3 3 1 ukη  vkη  + 54 − 6(u2 + v2)k2η2 + u2v2k4η4 sin √ sin √ . (48) u3v3k6η6 3 3

Following Ref. [25] one can evaluate analytically the integral I(u, v, kη) for kη  1:

1 I (u, v, kη) = [I cos(kη) + I sin(kη)] , (49) RD,χ 4kη 1 2 7 where 2 2  2  27 u + v − 3 2 2 3 − (u + v) I1 := −4uv + (u + v − 3) ln , (50) 2 u3v3 3 − (u − v)2 27 (u2 + v2 − 3)2 √ I := π Θ(u + v − 3), (51) 2 2 u3v3 with Θ being the step function. Its oscillation average is therefore

I2 + I2 I2 = 1 2 . (52) RD,χ 32(kη)2

In a MD universe, we take the w → 0 limit in Eqs. (45) and (46) and obtain Φ = Ψ = AΦ, Q = (2/3)AΦ. Thus, it is easy to see 20 F = , (53) MD,χ 3 and thus 20[kη − sin(kη)] I = . (54) MD,χ (kη)2

B. Comoving Gauge v.s. Newtonian Gauge

The comoving gauge is defined by

δφ = 0. (55)

We write the comoving curvature perturbation as ζ = −ψ. The linear equations of motion in the comoving gauge are

2 3(1 − w) 6Hζ0 + 2k2ζ + Hk2χ + H2δN = 0, (56) a w ζ0 δN = , (57) H k2 k2 k2 3 ζ00 + 2Hζ0 + ζ + (χ0 + Hχ) − HδN 0 + δN − (1 − w)H2δN = 0. (58) 3 3a 3 2 The solution regular at η = 0 is given by

1 + 3w k A (~k) 3(1 + w) f 0  ζ = A (~k)f (k, η), δN = A (~k)ηf 0 , χ = − ζ ζ + (1 + 3w)kηf , (59) ζ ζ 2 ζ ζ a 2 w k ζ where √  wkη −ν √ f (k, η) := Γ(ν + 1) J ( wkη). (60) ζ 2 ν

It is well-known that the primordial amplitude in the comoving gauge is related to that in the Newtonian gauge by 5 + 3w A (~k) = − A (~k). (61) ζ 3(1 + w) Φ

To understand the gauge dependence in a RD universe, it will be helpful to see the behavior of the scalar pertur- bations for kη  1. For w = 1/3, we have √ sin(kη/ 3) √ k √ √ ζ = Aζ √ , δN ≈ Aζ cos(kη/ 3), χ ≈ − 3Aζ sin(kη/ 3). (62) kη/ 3 a 8

This should be contrasted with the behavior of the Newtonian gauge perturbations in a RD universe for kη  1: √ √ cos(kη/ 3) 3 sin(kη/ 3) Φ = Ψ ≈ −9AΦ ,Q ≈ AΦ √ . (63) k2η2 2 kη/ 3 We see that in the comoving gauge the source term contains the terms that do not decay at late times. Accordingly, we have       12  2 2 ukη vkη FRD,δφ(u, v, kη) = −2 − 3 − 2 u + v cos √ cos √ u2v2k2η2 3 3 √ 2 3 ukη  vkη  + 6 3 − 2(u2 + v2) + u2 −3 + u2 + 2v2 k2η2 sin √ cos √ u3v2k3η3 3 3 √ 2 3 ukη  vkη  + 6 3 − 2(u2 + v2) + v2 −3 + 2u2 + v2 k2η2 cos √ sin √ u2v3k3η3 3 3 1  + −36 3 − 2(u2 + v2) u3v3k4η4  ukη  vkη  + k2η2 6(u2 + v2) − k2η2u2v2 3 − (u2 + v2) sin √ sin √ , (64) 3 3 which does not decay at late times, in contrast to the Newtonian gauge result (48). It then follows that  2      2 3 2 2 kη 2 2 kη IRD,δφ = IRD,χ − 3(2u − 3uv + 2v ) cos (u − v)√ + 3(2u + 3uv + 2v ) cos (u + v)√ 3 2u2v2kη 3 3 √ 3   kη   kη  + (u − v) sin (u − v)√ − (u + v) sin (u + v)√ , (65) 2uv 3 3 where we took the limit kη  1. If we had only the first term in Eq. (65), the energy density of induced GWs would always be gauge-invariant, as the factor (2/3)2 is canceled in the final result due to the relation (61). The first line decays as ∼ η−1, while the second line just oscillates, and hence the latter in fact dominates at late times, resulting in a large gauge dependence. This is essentially due to the first term in the source (39). The large gauge dependence we have observed is in fact generic to the other values of w (> 0). In the Newtonian gauge, we have, for kη  1, Φ = Ψ ∼ η−ν−3/2,Q ∼ η−ν−1/2. (66) However, in the comoving gauge we have k ζ ∼ η−ν−1/2, δN ∼ χ ∼ η−ν+1/2, (67) a which shows that the first term in the source (39) always overwhelms the other contributions and causes a large gauge dependence. Let us then consider a MD universe. Since one has w in the denominator in Eq. (59), the w → 0 limit in the k- essence description of a fluid seems particularly subtle in the comoving gauge. However, for w  1, fζ is approximated by w f = 1 − k2η2, (68) ζ 10 and using this one finds k 1 ζ = A , δN = 0, χ = − A kη (69) ζ a 5 ζ in the w → 0 limit [23]. Thus, one can safely take the w → 0 limit. Note that (k/a)χ grows in time. This is again different from the behavior of the Newtonian gauge variables in a MD universe: Φ, Ψ, and Q remain constant. This difference gives rise to a growing contribution in F and I: k2η2 F = 2 − , (70) MD,δφ 25 32 kη I = I − . (71) MD,δφ 5 MD,χ 5 9

If we had only the first term in Eq. (71), there would be no gauge dependence in induced GWs, given that Eq. (61) accounts for the factor (3/5)2. However, this term decays as ∼ η−1, and so the second term dominates at late times. Therefore, there is a large gauge dependence also in this case. Again, this is caused by the first term in the source (39).

C. Uniform Curvature Gauge v.s. Newtonian Gauge

Finally, let us consider the uniform curvature gauge defined by

ψ = 0. (72)

The evolution of the scalar perturbations in the uniform curvature gauge is governed by

3 2w H 1 − w HQ0 + (1 − w)H2Q − k2χ − H2δN = 0, (73) 2 3(1 + w) a 1 + w 3(1 + w) δN = Q, (74) 2  H  Q00 + H(2Q0 − δN 0) + wk2 Q + χ = 0. (75) a

The non-decaying solution is given by

3(1 + w) k 3(1 + w) f 0 Q = A (~k)f (η, k), δN = A (~k)f , χ = A (~k) Q , (76) Q Q 2 Q Q a 2w Q k where √  wkη −ν √ f (η, k) = Γ(ν + 1) J ( wkη). (77) Q 2 ν

This function is the same as Eq. (60). The primordial amplitude AQ is related to Aζ (and AΦ) by

5 + 3w A (~k) = −A (~k) = A (~k). (78) Q ζ 3(1 + w) Φ

In a RD universe, we have √ √ δN sin(kη/ 3) k cos(kη/ 3) Q = = AQ √ , χ ≈ 6AQ (79) 2 kη/ 3 a kη for kη  1. Therefore, unlike in the comoving gauge, the perturbations in the uniform curvature gauge decay as ∼ η−1. It is then straightforward to compute √ 2 2     2 2     ~ 12(u + v − 3) ukη vkη 12 3(u + v − 3) ukη vkη FRD,ζ (k, ~q, η) = cos √ cos √ − sin √ cos √ u2v2k2η2 3 3 u3v2k3η3 3 3 √ 12 3(u2 + v2 − 3) ukη  vkη  36(u2 + v2 − 3) ukη  vkη  − cos √ sin √ + sin √ sin √ . (80) u2v3k3η3 3 3 u3v3k4η4 3 3

This expression is clearly different from the Newtonian gauge result (48). However, integrating this to get I we find

22 I = I . (81) RD,ψ 3 RD,χ

Taking into account the relation (78), we see from Eq. (81) that the comoving gauge and the uniform curvature gauge give the identical result on the energy density of induced GWs. To see whether this is accidental or not, we evaluate I numerically for the other values of w (. 1). Examples of our numerical investigation are presented in Figs. 1–3. In Figs. 1 and 2 we present the comparison of I in the 10

0.00 ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ -0.02 ×◦ ×◦ ×◦ -0.04 ×◦ ×◦ -0.06 ×◦ ×◦ -0.08 3(1+w) 2 ◦   ℐ,χ ×◦ 5+3w -0.10 ×◦ × ℐ,ψ -0.12 ×◦ ◦ -0.14 × 0.0 0.2 0.4 0.6 0.8 1.0 w

FIG. 1. Comparison of I(2, 2, 250) (as a function of w) computed in the Newtonian and uniform curvature gauges.

0.00 ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ -0.01 ×◦ ×◦ ×◦ -0.02 ×◦

3(1+w) 2 ×◦ 0.03 ◦   ℐ,χ - 5+3w ×◦

× ℐ,ψ -0.04 ×◦ ×◦ 0.0 0.2 0.4 0.6 0.8 1.0 w

FIG. 2. Comparison of I(5, 5.5, 250) (as a function of w) computed in the Newtonian and uniform curvature gauges.

Newtonian and uniform curvature gauges for different values of w with u, v, kη being fixed. We also show in Fig. 3 the comparison of I as a function of kη for w = 2/3. These results imply that the following relation holds:

3(1 + w)2 I = I . (82) ,ψ 5 + 3w ,χ

We thus conclude that the Newtonian gauge and the uniform curvature gauge give the identical result on ΩGW for w > 0. This is, however, not true in the case of w = 0. Again, there is a subtlety regarding w in the denominator, but this 2 2 can be circumvented in the same way as in the comoving gauge. Since fQ = 1 − wk η /10 for w  1, we have k 3 χ = − A kη. (83) a 10 Q 11

0.06 ×◦×◦◦ ×◦◦ ×◦ ×◦ ×◦ ××◦ ×◦ × 25 ×◦ ◦ ×◦ ◦ ℐ,χ ×◦ × 0.04 ×◦ 49 ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ × ℐ,ψ ×◦ ×◦ 0.02 ×◦ ×◦ ×◦ ×◦ ×◦ 0.00 ×◦ ×◦ ×◦ ×◦ ×◦ ×◦ -0.02 ×◦ ×◦ ×◦ × ◦ ×◦ ×◦ × -0.04 ◦ ×◦ ×◦ ×◦ ×◦ × × ◦ ◦ ×◦×◦×◦×◦ -0.06 200 202 204 206 208 210 kη

FIG. 3. Comparison of I(2, 2, kη) computed in the Newtonian and uniform curvature gauges for w = 2/3.

Therefore, the evolution of the scalar perturbations in a MD universe is very similar to that in the comoving gauge. This yields the following result on F and I:

3 9 32 9 F = − k2η2, I = I − kη. (84) MD,ψ 2 100 MD,ψ 5 MD,χ 20 Therefore, in a MD universe, the uniform curvature gauge and the Newtonian gauge give very different results. Rather, the uniform curvature gauge is similar to the comoving gauge and their results are different only by a factor at late times: IMD,ψ ≈ (9/4)IMD,δφ. Here again, this is due to the first term in the source (39). The factor 9/4 comes from the coefficients of the solution of χ: (3/10)2 = (9/4) × (1/5)2. It is exactly this factor that explains the difference between the comoving and uniform curvature gauges in the numerical calculation in [24].

V. CONCLUDING REMARKS

In this paper, we have revisited the issue of the gauge dependence of gravitational waves (GWs) induced at second order from scalar perturbations. We have evaluated the energy density of induced GWs in different gauges in a universe dominated by a perfect fluid whose equation-of-state parameter w is constant, and arrived at the following conclusions: (i) the amplitude of induced GWs in the comoving gauge is significantly larger than that in the Newtonian gauge for any w (≥ 0), and this huge gauge dependence is a consequence of the presence of the shift vector; (ii) for w > 0 the Newtonian gauge result agrees with that of the uniform curvature gauge; (iii) for w = 0 the uniform curvature gauge result differs only by a factor from that of the comoving gauge, but deviates significantly from that of the Newtonian gauge. Our calculation has been done analytically for w = 0 and w = 1/3 using the method of Ref. [25]. The above conclusions are consistent with the previous numerical result [24]. The gauge dependence has been clarified based only on the evolution of the perturbations, and hence our result is robust against the input form of the primordial power spectrum of the scalar perturbations. For simplicity and clarity, we have focused on the ideal case with w = const rather than the realistic and conventional cosmological setup. Nevertheless, we believe that the present paper would be of help to gaining a deeper understanding of the gauge dependence of scalar-induced GWs. As was noted in Ref. [24], the appropriate gauge one should choose depends on what quantity one measures in each observation. Given that there is a large gauge dependence of second-order GWs, it would be important to address this issue and identify the true observables.

Note added: While we were in the final stage of this work, the paper by J. O. Gong [28] appeared in the arXiv, where the gauge dependence of induced gravitational waves was studied analytically by comparing the Newtonian 12 and comoving gauge results in a matter-dominated universe. Our conclusion agrees with his where we overlap.

ACKNOWLEDGMENTS

We are grateful to K. Inomata and R. Saito for fruitful discussions. The work of KT was supported by the Rikkyo University Special Fund for Research. The work of TK was supported by MEXT KAKENHI Grant Nos. JP15H05888, JP17H06359, JP16K17707, and JP18H04355.

[1] LIGO Scientific, Virgo collaboration, B. P. Abbott et al., Observation of Gravitational Waves from a Binary Black Hole Merger, Phys. Rev. Lett. 116 (2016) 061102 [1602.03837]. [2] LIGO Scientific, Virgo collaboration, B. P. Abbott et al., GW151226: Observation of Gravitational Waves from a 22-Solar-Mass Binary Black Hole Coalescence, Phys. Rev. Lett. 116 (2016) 241103 [1606.04855]. [3] LIGO Scientific, VIRGO collaboration, B. P. Abbott et al., GW170104: Observation of a 50-Solar-Mass Binary Black Hole Coalescence at Redshift 0.2, Phys. Rev. Lett. 118 (2017) 221101 [1706.01812]. [4] LIGO Scientific, Virgo collaboration, B. P. Abbott et al., GW170814: A Three-Detector Observation of Gravitational Waves from a Binary Black Hole Coalescence, Phys. Rev. Lett. 119 (2017) 141101 [1709.09660]. [5] LIGO Scientific, Virgo collaboration, B. P. Abbott et al., GW170817: Observation of Gravitational Waves from a Binary Neutron Star Inspiral, Phys. Rev. Lett. 119 (2017) 161101 [1710.05832]. [6] LIGO Scientific, Virgo collaboration, B. P. Abbott et al., GW170608: Observation of a 19-solar-mass Binary Black Hole Coalescence, Astrophys. J. 851 (2017) L35 [1711.05578]. [7] S. Mollerach, D. Harari and S. Matarrese, CMB polarization from secondary vector and tensor modes, Phys. Rev. D69 (2004) 063002 [astro-ph/0310711]. [8] K. N. Ananda, C. Clarkson and D. Wands, The Cosmological background from primordial density perturbations, Phys. Rev. D75 (2007) 123518 [gr-qc/0612013]. [9] D. Baumann, P. J. Steinhardt, K. Takahashi and K. Ichiki, Gravitational Wave Spectrum Induced by Primordial Scalar Perturbations, Phys. Rev. D76 (2007) 084019 [hep-th/0703290]. [10] R. Saito and J. Yokoyama, Gravitational wave background as a probe of the primordial black hole abundance, Phys. Rev. Lett. 102 (2009) 161101 [0812.4339]. [11] H. Assadullahi and D. Wands, Gravitational waves from an early matter era, Phys. Rev. D79 (2009) 083511 [0901.0989]. [12] R. Saito and J. Yokoyama, Gravitational-Wave Constraints on the Abundance of Primordial Black Holes, Prog. Theor. Phys. 123 (2010) 867 [0912.5317]. [13] E. Bugaev and P. Klimai, Induced gravitational wave background and primordial black holes, Phys. Rev. D81 (2010) 023517 [0908.0664]. [14] K. Jedamzik, M. Lemoine and J. Martin, Generation of gravitational waves during early between cosmic inflation and reheating, JCAP 1004 (2010) 021 [1002.3278]. [15] L. Alabidi, K. Kohri, M. Sasaki and Y. Sendouda, Observable Spectra of Induced Gravitational Waves from Inflation, JCAP 1209 (2012) 017 [1203.4663]. [16] L. Alabidi, K. Kohri, M. Sasaki and Y. Sendouda, Observable induced gravitational waves from an early matter phase, JCAP 1305 (2013) 033 [1303.4519]. [17] K. Inomata and T. Nakama, Gravitational waves induced by scalar perturbations as probes of the small-scale primordial spectrum, Phys. Rev. D99 (2019) 043511 [1812.00674]. [18] K. Inomata, K. Kohri, T. Nakama and T. Terada, Gravitational Waves Induced by Scalar Perturbations during a Gradual Transition from an Early Matter Era to the Radiation Era, 1904.12878. [19] K. Inomata, K. Kohri, T. Nakama and T. Terada, Enhancement of Gravitational Waves Induced by Scalar Perturbations due to a Sudden Transition from an Early Matter Era to the Radiation Era, Phys. Rev. D100 (2019) 043532 [1904.12879]. [20] R.-g. Cai, S. Pi and M. Sasaki, Gravitational Waves Induced by non-Gaussian Scalar Perturbations, Phys. Rev. Lett. 122 (2019) 201101 [1810.11000]. [21] R.-G. Cai, S. Pi and M. Sasaki, Universal infrared scaling of gravitational wave background spectra, 1909.13728. [22] F. Arroja, H. Assadullahi, K. Koyama and D. Wands, Cosmological matching conditions for gravitational waves at second order, Phys. Rev. D80 (2009) 123526 [0907.3618]. [23] L. Boubekeur, P. Creminelli, J. Norena and F. Vernizzi, Action approach to cosmological perturbations: the 2nd order metric in matter dominance, JCAP 0808 (2008) 028 [0806.1016]. [24] J.-C. Hwang, D. Jeong and H. Noh, Gauge dependence of gravitational waves generated from scalar perturbations, Astrophys. J. 842 (2017) 46 [1704.03500]. [25] K. Kohri and T. Terada, Semianalytic calculation of gravitational wave spectrum nonlinearly induced from primordial curvature perturbations, Phys. Rev. D97 (2018) 123532 [1804.08577]. [26] S. Matarrese, On the Classical and Quantum Irrotational Motions of a Relativistic Perfect Fluid. 1. Classical Theory, Proc. Roy. Soc. Lond. A401 (1985) 53. 13

[27] J. Garriga and V. F. Mukhanov, Perturbations in k-inflation, Phys. Lett. B458 (1999) 219 [hep-th/9904176]. [28] J. O. Gong, Analytic integral solutions for induced gravitational waves, 1909.12708.