We obtain explicit formulas for the solution of the wave equation in certain Friedmann–Lemaître–Robertson–Walker (FLRW) spacetimes. Our method, pioneered by Klainerman and Sarnak, consists in finding differential operators that map solutions of the wave equation in these FLRW spacetimes to solutions of the conformally invariant wave equation in simpler, ultra-static spacetimes, for which spherical mean formulas are available. In addition to recovering the formulas for the dust-filled flat and hyperbolic FLRW spacetimes originally derived by Klainerman and Sarnak and generalizing them to the spherical case, we obtain new formulas for the radiation-filled FLRW spacetimes and also for the de Sitter, anti-de Sitter, and Milne universes. We use these formulas to study the solutions with respect to the Huygens principle and the decay rates and to formulate conjectures about the general decay rates in flat and hyperbolic FLRW spacetimes. The positive resolution of the conjecture in the flat case is seen to follow from known results in the literature.

The aim of this article is to obtain explicit formulas and exact decay rates for the solution of the wave equation in certain Friedmann–Lemaître–Robertson–Walker (FLRW) spacetimes, taken as fixed backgrounds. Besides its intrinsic interest in modeling electromagnetic or gravitational waves, the linear wave equation may be considered as a first step to understand the qualitative behavior of solutions of the Einstein equations, which usually requires a good quantitative grasp of the linearized Einstein equations.

Explicit formulas, albeit applying only to particular FLRW models, are very useful in that they provide detailed information about the solutions and may indeed inspire conjectures about the behavior of solutions of the wave equation in other spacetimes (see Sec. VI). Moreover, they highlight special properties of the solutions of the wave equation in the particular FLRW models where they do apply, such as (some form of) the Huygens principle.

The wave equation in cosmological spacetimes has been amply studied in the literature (see, for example, Refs. 114 and references therein). Our own contribution is described in more detail below.

The Klainerman–Sarnak method consists of finding an operator Ô that maps solutions ϕ of the wave equation in a given FLRW background to solutions Ôϕ of the conformally invariant wave equation on a simpler, ultra-static spacetime (characterized by the same spatial curvature), where a spherical means formula for Ôϕ is available; by inverting the operator Ô, one then obtains an explicit expression for ϕ. It should be noted, however, that such operators can only be found for very specific FLRW spacetimes.

Klainerman and Sarnak introduced this method in Ref. 8 for the dust-filled FLRW universes with flat and hyperbolic spatial sections. Abbasi and Craig extended their analysis in Ref. 1, where they elaborated on the decay rates of the solutions of the wave equation and also on the validity of the Huygens principle. Physical applications were later discussed in Refs. 15 and 16.

In this paper, we generalize the Klainerman–Sarnak method to the spherical dust-filled FLRW universe, as well as to a number of other FLRW spacetimes. More precisely, using the parameter K to denote the curvature of the spatial sections (so that K = 0, −1, 1 corresponds to flat, hyperbolic, or spherical geometries, respectively), we obtain new explicit formulas for the solution of the wave equation in radiation-filled (K = 0, −1, 1), de Sitter (K = 0, −1, 1), anti-de Sitter (K = −1), and Milne (K = −1) universes. In each case, we are able to find an operator Ô such that Ôϕ satisfies the conformally invariant wave equation in the following:

  • Flat case: Minkowski’s spacetime.

  • Hyperbolic case: The ultra-static universe with hyperbolic spatial sections (that is the hyperbolic analog of Einstein’s universe).

  • Spherical case: The ultra-static universe with spherical spatial sections (that is, Einstein’s universe).

The well-known fact that Minkowski’s spacetime is conformal to a globally hyperbolic region of Einstein’s universe makes it plausible that a spherical means formula for the conformally invariant wave equation should exist in this universe. A similar argument applies to its hyperbolic analog, which can be seen to be conformal to a globally hyperbolic region of Minkowski’s spacetime.

All solutions for which we obtain explicit formulas have the property that Ôϕ satisfies the strong Huygens principle: the values of Ôϕ at some spacetime point depend only on the values of the initial data at points on its light cone (that is information propagates along null geodesics). However, as noted by Abbasi and Craig in Ref. 1, this is not necessarily the case for ϕ itself. In this paper, we show that the strong Huygens principle does hold for the radiation-filled (K = 0, −1, 1) and the Milne universes. Moreover, a weaker form of the Huygens principle, dubbed the incomplete Huygens principle in Ref. 17, is satisfied by both the de Sitter (K = 0, −1, 1) and the anti-de Sitter universes.

The explicit formulas derived in this paper give very detailed information about the solutions, such as decay rates as t → +∞, where t is the physical time coordinate of the FLRW metric (see Sec. I E).

In the flat case, it is interesting to notice that the decay rates in both the dust-filled and the radiation-filled universes coincide with the decay rate t−1 of waves propagating in Minkowski’s spacetime, despite the different behavior of their scale factors (a(t)=t23 and a(t)=t12, respectively). These decay rates differ from those obtained in Ref. 2 for the nonzero Fourier modes in the case of toroidal spatial sections, which were, respectively, t23 and t12. The explanation for this discrepancy is dispersion: whereas in the toroidal models, the decay is solely due to the cosmological redshift resulting from the expansion, for noncompact spatial sections, the solutions can disperse across an unbounded region. The faster the expansion, the larger the decay due to the cosmological redshift, but the smaller the decay due to dispersion (as faraway regions are moving away faster). By adapting the analysis of the damped wave equation in Ref. 18, we show that indeed there exists a regime (corresponding to a slow enough expansion of the spatial sections) where the two effects exactly balance each other to give the same decay rate as in Minkowski’s spacetime. For faster expansion, the cosmological redshift does not fully compensate for the lack of dispersion, and the decay becomes slower, until no decay occurs at all.

In the hyperbolic case, we obtain faster decay rates than in the flat case for both the dust-filled and the radiation-filled universes, as could be expected from the fact that dispersion is more effective for hyperbolic spatial geometry. Based on these two examples and also on the numerical results obtained in Ref. 19, we formulate a conjecture for the decay rates in general FLRW spacetimes with hyperbolic sections. We also correct a claim in Ref. 1 by showing that the decay rate in the dust-filled universe is t32 rather than t−2.

Generically, solutions of the wave equation blow up at the Big Bang (and also at the Big Crunch, when there is one). Nonetheless, some solutions (e.g., constant functions) have a well-defined limit at the Big Bang. In the cases of the dust-filled (K = 0, −1, 1) and radiation-filled (K = 0, −1, 1) universes, we show how to obtain explicit expressions for these solutions. In fact, by taking limits of solutions with appropriate initial data, we give explicit expressions for solutions with any given function as its limit at the Big Bang (see also Refs. 7 and 9).

Recall that a (3 + 1)-dimensional FLRW spacetime (R×Σ3,g) has metric
(1)
where dΣ32 is the standard Riemannian metric for Σ3=R3,H3, or S3 and a(t) is the scale factor. The conformal time coordinate is defined as
(2)
and we refer to t as the physical time coordinate. In what follows, we give a precise quantitative summary of the main results and conjectures derived from the exact formulas.

Flat case: The results for FLRW universes with flat spatial sections are summarized in Table I. The decay rates for the dust-filled and the de Sitter universes had already been obtained in Refs. 1 and 20, respectively. The fact that the de Sitter universe satisfies the incomplete Huygens principle was first noticed in Ref. 17.

TABLE I.

Properties of the solutions of the wave equation □gϕ = 0 in FLRW models with flat space sections.

SpacetimeScale factort(τ)Large timesHuygens
Minkowski a(τ) = 1 tτ |ϕ| ≲ t−1 Yes (strong) 
Dust a(τ) ∝ τ2 tτ3 |ϕ| ≲ t−1 No 
Radiation a(τ) ∝ τ tτ2 |ϕ| ≲ t−1 Yes (strong) 
de Sitter a(τ) ∝ τ−1 t ∝ − log τ |tϕ| ≲ e−2t Yes (incomplete) 
SpacetimeScale factort(τ)Large timesHuygens
Minkowski a(τ) = 1 tτ |ϕ| ≲ t−1 Yes (strong) 
Dust a(τ) ∝ τ2 tτ3 |ϕ| ≲ t−1 No 
Radiation a(τ) ∝ τ tτ2 |ϕ| ≲ t−1 Yes (strong) 
de Sitter a(τ) ∝ τ−1 t ∝ − log τ |tϕ| ≲ e−2t Yes (incomplete) 
We now consider the family of flat FLRW metrics whose scale factor satisfies a(t) = tp, with p ≥ 0. These correspond to universes filled with a perfect fluid with linear equation of state pm = m (and zero cosmological constant), where
(see  Appendix A). We summarize our (proved) conjecture for the decay of solutions of the wave equations in these spacetimes in Fig. 1. The precise statement can be found in Sec. VI.
FIG. 1.

Conjectured (and proved) decay rates in the flat case. The rates for w13 follow from Refs. 18 and 21 and the absence of decay for w<13 follows from Ref. 2. The Klainerman–Sarnak method gives the precise decay for the dust-filled (w = 0) and radiation-filled (w=13) universes.

FIG. 1.

Conjectured (and proved) decay rates in the flat case. The rates for w13 follow from Refs. 18 and 21 and the absence of decay for w<13 follows from Ref. 2. The Klainerman–Sarnak method gives the precise decay for the dust-filled (w = 0) and radiation-filled (w=13) universes.

Close modal

Hyperbolic case: The results for FLRW universes with hyperbolic spatial sections are summarized in Table II. The decay rate in the dust-filled universe corrects the claim in Ref. 1.

TABLE II.

Properties of the solutions of the wave equation □gϕ = 0 in FLRW models with hyperbolic space sections.

SpacetimeScale factort(τ)Large timesHuygens
Dust a(τ) = cosh τ − 1 t = sinh ττ |ϕ|t32 No 
Radiation a(τ) = sinh τ t = cosh τ |ϕ| ≲ t−2 Yes (strong) 
AdS a(τ) = sech τ t=2arctantanhτ2 ϕc Yes (incomplete) 
de Sitter a(τ) = −cosech τ t=logtanhτ2 |tϕ| ≲ e−2t Yes (incomplete) 
Milne a(τ) = eτ t = eτ |ϕ| ≲ t−2 Yes (strong) 
SpacetimeScale factort(τ)Large timesHuygens
Dust a(τ) = cosh τ − 1 t = sinh ττ |ϕ|t32 No 
Radiation a(τ) = sinh τ t = cosh τ |ϕ| ≲ t−2 Yes (strong) 
AdS a(τ) = sech τ t=2arctantanhτ2 ϕc Yes (incomplete) 
de Sitter a(τ) = −cosech τ t=logtanhτ2 |tϕ| ≲ e−2t Yes (incomplete) 
Milne a(τ) = eτ t = eτ |ϕ| ≲ t−2 Yes (strong) 

We now consider the family of hyperbolic FLRW universes filled with a perfect fluid with linear equation of state pm = m (and zero cosmological constant—see  Appendix A). We summarize our conjecture for the decay of solutions of the wave equations in these spacetimes in Fig. 2. The precise statement can be found in Sec. VI.

FIG. 2.

Conjectured decay rates in the hyperbolic case. The Klainerman–Sarnak method gives the precise decay for the dust-filled (w = 0) and radiation-filled (w=13) universes.

FIG. 2.

Conjectured decay rates in the hyperbolic case. The Klainerman–Sarnak method gives the precise decay for the dust-filled (w = 0) and radiation-filled (w=13) universes.

Close modal

Spherical case: The results for FLRW universes with spherical spatial sections are summarized in Table III. The solutions in the dust-filled and radiation-filled universes blow up for generic initial data due to the Big Crunch singularity (see Sec. I D).

TABLE III.

Properties of the solutions of the wave equation □gϕ = 0 in FLRW models with spherical space sections.

SpacetimeScale factort(τ)Large timesHuygens
Dust a(τ) = 1 − cos τ t = τ − sin τ |ϕ| blows upa No 
Radiation a(τ) = sin τ t = 1 − cos τ |ϕ| blows upa Yes (strong) 
de Sitter a(t) = sec τ t=2arctanhtanτ2 |tϕ| ≲ e−2t Yes (incomplete) 
SpacetimeScale factort(τ)Large timesHuygens
Dust a(τ) = 1 − cos τ t = τ − sin τ |ϕ| blows upa No 
Radiation a(τ) = sin τ t = 1 − cos τ |ϕ| blows upa Yes (strong) 
de Sitter a(t) = sec τ t=2arctanhtanτ2 |tϕ| ≲ e−2t Yes (incomplete) 
a

Generically (it is possible to choose initial data such that |ϕ| does not blow up).

In this paper, we will assume that the initial data for the wave equation are smooth so that the solutions are also smooth. Moreover, we will assume that the initial data belong to any required Sobolev space or even that it has compact support. All these assumptions can be easily relaxed, but we shall refrain from doing so as this would distract us from the main objectives of this paper.

We are interested in solving the wave equation,
(3)
in a (3 + 1)-dimensional FLRW background (R×Σ3,g), with metric
(4)
where dΣ32 is the standard Riemannian metric for Σ3=R3,H3, or S3. Introducing the conformal time
(5)
we can write the wave equation in the form
(6)
where ΔΣ3 is the Laplacian operator for the metric dΣ32 and
(7)
Now, let Ô be an operator acting on the space of smooth functions that commutes with ΔΣ3. If the assumption
(8)
holds, where K ∈ {−1, 0, 1} represents the curvature of the spatial sections we are considering, then (6) implies that Ôϕ satisfies the conformally invariant wave equation,
(9)
We will search a suitable operator Ô by taking
(10)
where f and g are functions to be determined. Plugging this expression into (8), we obtain the condition
(11)
We will now try to satisfy this condition for each of the three different types of spatial sections.
Let K = 0. Equation (11) holds if and only if
(12)
with α,β,κR. Let us consider the case
(13)
with τ > 0. It can be easily checked that there are non-trivial solutions f and g of the constraint equations (12) only for j ∈ {−1, 0, 1, 2}, provided that α, β, κ are chosen appropriately. We will analyze each of these four possibilities next.
If j = 0, the constraint equations (12) force us to choose α = 0 so that the functions f and g become
(14)
In this case, a(τ) ≡ 1, and so our FLRW universe is simply Minkowski’s spacetime. Indeed, if ϕ is a solution of the wave equation in Minkowski’s spacetime, then
(15)
is a solution as well.
If j = 2, conditions (12) force the functions f and g to be given by
(16)
and therefore,
(17)
In this case, a(τ) = τ2 and
(18)
so that
(19)
This scale factor a(t) is a solution of the Friedmann equations when K = 0, Λ = 0, and w = 0 (see  Appendix A), corresponding to a dust-filled flat FLRW model (also known as the Einstein–de Sitter universe).
Here and in what follows, we will denote the initial data for the wave equation at t = t0 by
(20)
Therefore, the initial data for
(21)
at the initial conformal time τ = τ0 are
(22)
Here, we used Eq. (6) to write
Using Kirchhoff’s formula for Ôϕ, we have
(23)
An explicit expression for ϕ can be obtained by noticing that Ôϕ=τ1τ(τ3ϕ) and then integrating Kirchhoff’s formula,
(24)
This is the formula obtained by Klainerman and Sarnak in Ref. 8. Abbasi and Craig used this formula in Ref. 1 to obtain decay estimates by noticing that
(25)
[and similarly for the other integrals in (24)], since
(26)
and
(27)
Therefore,
(28)
and so, provided that the relevant norms are finite, we have
(29)
From (24), we expect the solutions of the wave equation to diverge as t → 0, which can be traced back to the fact that the FLRW metric is singular at t = 0. However, as noted in Ref. 1 (see also Refs. 7 and 9), it is possible to construct special solutions, which have a well-defined limit as t → 0. Indeed, if we fix ϕ0, ϕ1 and τ and take the limit τ0 → 0 in Eqs. (22) and (24), we obtain
(30)
It is clear that this expression determines a solution of the wave equation satisfying
(31)
If j = 1, conditions (12) tell us that the functions f and g must be of the form
(32)
In this case, a(τ) = τ and
(33)
which implies
(34)
This scale factor is a solution of the Friedmann equations for K = 0, Λ = 0 and w=13 (see  Appendix A). Therefore, this model corresponds to a flat, radiation-filled universe. We note that in this case, Ô can be taken as the multiplicative operator,
(35)
The reason for this is that the energy–momentum tensor of a radiation fluid is traceless, and thus, the scalar curvature R of the four-dimensional spacetime vanishes. Consequently, the wave equation coincides with the conformally invariant wave equation, and so the scalar field = τϕ satisfies the wave equation in Minkowski spacetime.
The initial data for Ôϕ=τϕ at the initial conformal time τ = τ0 are given by
(36)
Therefore, if the appropriate Sobolev norms of the initial data are finite, by the standard decay of the wave equation in Minkowski spacetime (see, e.g., Ref. 22), we have, as τ → +∞,
(37)
the same decay rate as in Minkowski spacetime.
The explicit expression for ϕ is again given by Kirchhoff’s formula (23),
(38)
As one would expect, ϕ satisfies the strong Huygens principle; a related result was obtained in Refs. 23 and 24 for gravitational waves in the Regge–Wheeler gauge.
Again, it is possible to construct solutions with a well-defined limit as t → 0 by fixing ϕ0, ϕ1, and τ and taking the limit τ0 → 0 in Eqs. (36) and (38),
(39)
It is easy to check that this expression determines a solution of the wave equation satisfying
(40)
If j = −1, we can choose α = β = 0 so that the functions f and g become
(41)
These lead to the operator Ô=τ1τ. In this case, a(τ) = τ−1, and so
(42)
This yields
(43)
corresponding to the contracting flat de Sitter spacetime (which is a globally hyperbolic region of the full de Sitter universe); the expanding case can be trivially recovered by changing the time coordinate from t to −t.
The initial data for
(44)
at the corresponding conformal time τ = τ0 is given by
(45)
where we used (6) to write
(46)
If the appropriate Sobolev norms of the initial data are finite, then we have, by boundedness of the solution of the wave equation in the region 0 < ττ0 of Minkowski’s spacetime,
(47)
as t → −∞ (that is, as τ → 0). This translates to |tϕ| ≲ e−2t as t → +∞ in the expanding case, confirming the result in Ref. 20.
An explicit expression for ϕ can be obtained by integrating Kirchhoff’s formula (23),
(48)
In the particular case, when ϕ1 ≡ 0, we obtain
(49)
Using the divergence theorem and Green’s identities, this expression can be rewritten as
(50)
In particular, ϕ(τ, x) depends only on the values of ϕ0 and ∇ϕ0 on Bττ0(x), that is, ϕ satisfies the strong Huygens principle in the special case when ϕ1 ≡ 0. This fact was previously noticed by Yagdjian in Ref. 17, who dubbed this property the incomplete Huygens principle.
Next, we want to find suitable operators Ô in the case K = −1. We start by noting that the constraints (11) require the functions f and g in operators of the form Ô=f(τ)τ+g(τ) to satisfy
(51)
These conditions can be met for certain choices of scale factors a(τ) satisfying the Friedmann equations (see  Appendix A). In these cases, it is possible to obtain an explicit formula for ϕ from the expression of Ôϕ in terms of spherical means, analogous to Kirchhoff’s formula in R3, which we now discuss.
We will now obtain a spherical means formula for the solution of the conformally invariant wave equation (9) in the spacetime given by the metric
(52)
where dΣ32 is the line element for the hyperbolic space H3. What follows is a simplified version of the more general derivation in Ref. 8.
Consider the Cauchy problem
(53)
where Δ is the Laplacian operator in H3. Using geodesic polar coordinates about some point xH3, we know that
(54)
where 2 is the line element for the unit sphere S2, and also that
(55)
where ΔS2 is the Laplacian on S2. The geodesic sphere S(x, r) about the point x is defined as
(56)
where expx:TxH3H3 is the geodesic exponential map and Br(0) is the ball of radius r about 0 in TxH3. We define the spherical mean of a function ϕ(t, x) on the geodesic sphere of radius r about x as
(57)
where S2 is the unit sphere in TxH3. Note that this last formula allows us to extend the spherical mean to negative values of the second argument, yielding
(58)
Now, if ϕ is a solution of problem (53), we have
(59)
where we used (57), (53), and the divergence theorem on the sphere. From (53), we have the initial data
(60)
Using (59), it can be checked that ω(t, r, x) ≔ sinh(r)Mϕ(t, r, x) satisfies
(61)
Given the initial data
(62)
the solution of Eq. (61) is given by the d’Alembert formula,
(63)
Since Mϕ(t, r, x) → ϕ(t, x) as r → 0, we obtain
(64)
where we used L’Hôpital’s rule and the fact that γ and ψ are odd functions. From (57), the solution of the Cauchy problem (53) is then given by
(65)
with cz(t) = expx(tz).
Let us consider the case
(66)
with τ > 0. This corresponds to the dust-filled hyperbolic model, since a(τ) solves Friedmann’s equations for K = −1, Λ = 0 and w = 0 (see  Appendix A). With this choice of the scale factor, the wave equation (6) becomes
(67)
In this case, the constraint equations (51) are satisfied if we choose κ = 1, α = 0, and β=32, that is, if we choose
(68)
We note that we can rewrite this as
(69)
which is the expression that can be found in Ref. 8. The initial data for Ôϕ at the initial conformal time τ = τ0 are then
(70)
where we used (67) to write
(71)
Then, (65) and (69) give the following expression for the solution:
(72)
Let us analyze the decay of the solution for τ → +∞. The first integral, for instance, can be estimated by noting that
(73)
and that, rewriting the integral on S2 as an integral over the geodesic sphere,
(74)
Similar computations apply to the other integrals in expression (72). After dividing both sides by sinh3τ2, we can write
(75)
and so, provided that the relevant norms are finite, we have
(76)
We note that, in general, a faster decay rate is not possible (see  Appendix B). To obtain this estimate in terms of the coordinate time t, we see from
(77)
that, as τ → +∞,
(78)
whence
(79)
As in the flat case, there exist solutions that have a well-defined limit at the Big Bang: if we take the limit τ0 → 0 while keeping ϕ0, ϕ1, and τ fixed, we obtain from (70) and (72) the limit solution
(80)
Using L’Hôpital’s rule, one can easily check that
(81)
Let us now consider the case
(82)
with τ > 0. This corresponds to the radiation-dominated hyperbolic universe, since a(τ) = sinh(τ) solves the Friedmann equations for K = −1, Λ = 0, and w=13 (see  Appendix A). With this choice of the scale factor, our wave equation(6) becomes
(83)
In this case, the constraint equations(51) are satisfied if we choose α = κ = 0 and β = 1, that is, if we choose
(84)
Again, this operator can be taken as multiplicative because in this universe, the wave equation coincides with the conformally invariant wave equation.
The initial data for Ôϕ at the initial conformal time τ = τ0 are
(85)
The solution in terms of spherical means is then given by (65),
(86)
As in the flat case, ϕ satisfies the strong Huygens principle.
Let us now analyze the decay of the solution as τ → +∞. We note that after dividing the formula above by sinh(τ), the first term of the solution can be estimated as
(87)
In fact, assuming for simplicity that the initial data have compact support, we have
(88)
where cz(r) = expx(rz), whence
(89)
for all R > 0. The remaining two terms have the same decay rate, which can be found after similar computations. Therefore, if the relevant Sobolev norms of the initial data are finite, we have
(90)
To obtain the decay rate in terms of the time variable t, we use
(91)
so that as t → +∞,
(92)
Again there exist solutions with a well-defined limit at the Big Bang: if we take the limit τ0 → 0 while keeping ϕ0, ϕ1, and τ fixed, we obtain from (85) and (86) the limit solution
(93)
which clearly converges to ϕ0(x) as τ → 0.
Let us now analyze the case
(94)
This corresponds to a globally hyperbolic region of the anti-de Sitter universe, since a(τ) = sech(τ) is a solution for the Friedmann equations for K = −1, Λ = −3, and ρ0 = 0 (see  Appendix A). We can obtain the scale factor as a function of the physical time t by noting that
(95)
and so
(96)
with tπ2,π2.
With this choice of the scale factor, (6) becomes
(97)
We can satisfy the constraint equations (51) by choosing α = β = 0 and κ = 1 so that we have
(98)
As before, we can express the solution in terms of spherical means. We have the following initial data for Ôϕ at the initial conformal time τ = τ0,
(99)
where we used (97) to write
(100)
The spherical means solution is then
(101)
In the particular case when ϕ1 ≡ 0, we obtain
(102)
where we defined r(y)=distH3(y,x). Using the divergence theorem and Green’s identities, together with Δ cotanh(r) = 0, this expression can be rewritten as
(103)
Therefore, ϕ(τ, x) depends only on the values of ϕ0 and ∇ϕ0 on Bττ0(x), that is, ϕ satisfies Yagdjian’s incomplete Huygens principle. In particular, if ϕ0 is compactly supported, then we have
(104)
Interestingly, in the case when ϕ0 ≡ 0, we have, for compactly supported ϕ1,
(105)
(in particular, this limit does not depend on x). By the superposition principle, this result holds, in general, for compactly supported initial data; it can be interpreted by noting that for fixed x, the curve τ ↦ (τ, x) approaches a single point in the full anti-de Sitter spacetime as τ → +∞ (see Ref. 25).
To check (105), we note that for ϕ0 ≡ 0, Eq. (101) can be written in the form
(106)
Discarding the boundary term, whose limit will vanish for compactly supported initial data, and using (99), we obtain
(107)
Let us now consider the case
(108)
with τ < 0. This corresponds to a globally hyperbolic region of the de Sitter universe, since a(τ) = −cosech(τ) is a solution for the Friedmann equations for K = −1, Λ = −3, and ρ0 = 0 (see  Appendix A). We can obtain the scale factor as a function of the physical time t by noticing that
(109)
and so
(110)
with t ∈ (0, +∞).
This choice of the scale factor in (6) leads to
(111)
We can satisfy the constraint equations (51) by choosing α = β = 0 and κ = −1 so that
(112)
Again we can repeat the same steps as in the previous cases to obtain the solution in terms of spherical means. The initial data for Ôϕ at the initial conformal time τ = τ0 are
(113)
where we used (111) to write
(114)
The spherical means solution is then
(115)
Repeating the same steps as was done for the anti-de Sitter solution, one can show that when ϕ1 ≡ 0, we have
(116)
Therefore, again ϕ satisfies Yagdjian’s incomplete Huygens principle.
Similarly to what happens in the flat de Sitter universe, the time derivative of the solution decays exponentially as t → +∞,
(117)
Indeed, if we take the partial derivative with respect to τ of both sides of (115), we have, as τ → 0,
(118)
where we are assuming that the L norms of ψ0, 0, and ψ1 are finite.
We will now work with the scale factor
(119)
This model represents the Milne universe, as we can see by writing the scale factor as a function of the physical variable t,
(120)
whence
(121)
for t ∈ (0, +∞).
With this choice, our wave equation (6) becomes
(122)
To satisfy the constraint equations (51), we can pick α = β = 1 and κ = 0. In this case, the operator is
(123)
To get an explicit expression for solutions, we consider the following initial data for Ôϕ at the initial conformal time τ = τ0:
(124)
The solution in terms of spherical means is given by (65),
(125)
As could be expected from the fact that the Milne universe is a globally hyperbolic region of Minkowski’s spacetime, ϕ satisfies the strong Huygens principle. Note that the decay of the solution as τ → +∞ is the same as that of the solution (86) if we assume that the appropriate norms are finite. Thus, as t → +∞, we have
(126)
This decay rate can be understood by viewing the Milne universe as the region t > |x| of Minkowski’s spacetime and recalling the well-known estimate
(127)
for solutions of the wave equation with compactly supported initial data (see, for instance, Ref. 26).
Finally, we consider the case K = 1. The constraint equations (11) now give
(128)
Again these conditions are satisfied for certain choices of a(τ), for which it is possible to obtain an explicit expression for ϕ from the expression of Ôϕ in terms of spherical means.
Let us find the solution of the conformally invariant wave equation (9) in the Einstein universe, that is, in the spacetime given by
(129)
where dΣ32 is the line element for the three-sphere S3. We therefore consider the Cauchy problem
(130)
where Δ is the Laplacian operator in S3. Following the exact same steps as in the hyperbolic case but replacing the hyperbolic functions by their trigonometric counterparts, we arrive at the Kirchhoff-like formula,
(131)
where again cz(t) = expx(tz). It is interesting to note that ϕ is periodic in t with period 2π.
Let us consider the case
(132)
with τ ∈ (0, 2π). This corresponds to the dust-filled spherical model, since a(τ) solves Friedmann’s equations for K = 1, Λ = 0, and w = 0 (see  Appendix A). With this choice of the scale factor, the wave equation (6) becomes
(133)
The constraint equations(128) are satisfied if we choose κ = 1, α = 0, and β=32, that is, if we choose
(134)
which can also be written as
(135)
as remarked in Ref. 8. The initial data for Ôϕ at the initial conformal time τ = τ0 are given by
(136)
where we used (133) to write
(137)
Finally, the spherical means formulas (131) and (135) give
(138)
In general, these solutions will diverge as τ → 0 and as τ → 2π. As in the flat case, there exist solutions that have a well-defined limit at the Big Bang: if we take the limit τ0 → 0 while keeping ϕ0, ϕ1, and τ fixed, we obtain from (136) and (138) the limit solution
(139)
Using L’Hôpital’s rule, one can easily check that
(140)
Similarly, we can find solutions with a well-defined limit at the Big Crunch by taking the limit τ0 → 2π while keeping ϕ0, ϕ1, and τ fixed. In this case, we obtain the limit solution
(141)
which, again using L’Hôpital’s rule, satisfies
(142)
Interestingly, if ϕ0 satisfies
(143)
for all xS3, then the two limit solutions coincide, yielding a solution with the same limit at the Big Bang and at the Big Crunch (this includes, of course, the constant solutions).
Next, let us take the scale factor
(144)
with τ ∈ (0, π). This corresponds to a radiation-dominated spherical universe, since a(τ) = sin(τ) solves the Friedmann equations for K = 1, Λ = 0, and w=13 (see  Appendix A). With this choice of the scale factor, our wave equation (6) becomes
(145)
In this case, the constraint equation (51) are satisfied if we choose α = κ = 0 and β = 1, that is, if we choose
(146)
Again, this operator can be taken as multiplicative because in this universe, the wave equation coincides with the conformally invariant wave equation.
The initial data for Ôϕ at the initial conformal time τ = τ0 is given by
(147)
The spherical means solution is then given by (131),
(148)
As in the flat and hyperbolic cases, ϕ satisfies the strong Huygens principle.
Again, these solutions will, in general, diverge as τ → 0 and as τπ, but there exist solutions that have a well-defined limit at the Big Bang: if we take the limit τ0 → 0 while keeping ϕ0, ϕ1, and τ fixed, we obtain from (147) and (148) the limit solution
(149)
which clearly converges to ϕ0(x) as τ → 0. Similarly, we can find solutions with a well-defined limit at the Big Crunch by taking the limit τ0π while keeping ϕ0, ϕ1, and τ fixed. In this case, we obtain the limit solution
(150)
which converges to ϕ0(x) as τπ. Interestingly, if ϕ0 is an even function on the sphere S3, then the two limit solutions coincide, yielding a solution with the same limit at the Big Bang and at the Big Crunch (this includes, of course, the constant solutions).
Finally, we consider the case
(151)
with τπ2,π2. This corresponds to the spherical (full) de Sitter universe, since a(τ) = sec(τ) is a solution for the Friedmann equations for K = 1, Λ = 3, and ρ0 = 0 (see  Appendix A). We can obtain the scale factor as a function of the physical time t by noting that
(152)
and so
(153)
With this choice of the scale factor, (6) becomes
(154)
We can satisfy the constraint equation (128) by choosing α = β = 0 and κ = 1 so that we have
(155)
As before, we can express the solution in terms of spherical means. We have the following initial data for Ôϕ at the initial conformal time τ = τ0:
(156)
where we used (154) to write
(157)
The spherical means solution is then
(158)
In the particular case when ϕ1 ≡ 0, we obtain
(159)
where we defined r(y)=distS3(y,x). Using the divergence theorem and Green’s identities, together with Δcotan(r) = 0, this expression can be rewritten as
(160)
Therefore, ϕ(τ, x) depends only on the values of ϕ0 and ∇ϕ0 on Bττ0(x), that is, ϕ satisfies Yagdjian’s incomplete Huygens principle.
As expected by analogy with the flat and hyperbolic de Sitter universes, the time derivative of the solution decays exponentially as t → +∞,
(161)
Indeed, if we take the partial derivative with respect to τ of both sides of (158), we have, as τπ2,
(162)
Similarly, one can show that the time derivative of the solution decays exponentially as t → −∞,
(163)

Let us consider the family of flat FLRW universes described by the scale factor a(t) = tp, with p ≥ 0. In this context, we obtained the Minkowski spacetime decay rate of t−1 for the wave equation when p=23 [dust-filled universe, see (29)] and when p=12 [radiation-filled universe, see (37)]. The independence of these results on p (as well as the numerical results in Ref. 19) suggests a wider range of validity for this decay.

Conjecture 6.1
(decay in the flat case). Let ϕ be a solution of the wave equation in a flat FLRW universe with scale factor a(t) = tp, where p ≥ 0. Assume that the initial data ϕ0(x) ≔ ϕ(t0, x) and ϕ1(x) ≔ tϕ(t0, x) are sufficiently regular and belong to appropriate Sobolev spaces. Then,
(164)
or, equivalently,
(165)
Moreover, we have
and there is no decay for p > 1 (i.e., 1<w<13).

The absence of decay for p > 1 (but not for p = 1) follows from the results in Ref. 2. For 0 ≤ p ≤ 1, we have the following result:

Theorem 6.2.

Conjecture 6.1 holds for 0 ≤ p ≤ 1 (i.e., w13).

Proof.
The case p = 0 corresponds to the wave equation in Minkowski’s spacetime. For 0 < p < 1, the original wave equation (6) (written in conformal time) can be seen as a damped wave equation in Minkowski’s spacetime,
(166)
where =τ2+Δ and
(167)
This equation was studied in Ref. 18, Theorem 3.5 (which can be extended to the L1L case, as stressed in Remark 3.3), where it was proved that
When p = 1, the respective damped equation has constant coefficients. By a simple rescaling, ϕ̃(τ,x):=ϕτ2,x2, we have
The diffusive structure of the above partial differential equation (PDE) was explicitly described in Ref. 21, where the decay
was established.□

A related conjecture for the hyperbolic case is motivated by the decay rates obtained in Secs. IV B and IV C and by the numerical results in Ref. 19.

Conjecture 6.3
(Decay in the hyperbolic case). Let ϕ be a solution of the wave equation in a hyperbolic FLRW universe, with the scale factor satisfying the Friedmann equation (A1) with zero cosmological constant and equation of state (A2). Assume that the initial data ϕ0(x) ≔ ϕ(t0, x) and ϕ1(x) ≔ tϕ(t0, x) are sufficiently regular and belongs to appropriate Sobolev spaces. Then,
(168)
Moreover, the decay is slower than (1+t)3(w+1)2 for 13<w<0, and there is no decay for 1w<13.

We thank João Costa and Alex Vañó-Viñuales for many useful discussions. This work was partially supported by FCT/Portugal through CAMGSD, IST-ID (Project Nos. UIDB/04459/2020 and UIDP/04459/2020). Flavio Rossetti was supported by FCT/Portugal through the Ph.D. scholarship (Grant No. UI/BD/152068/2021).

The authors have no conflicts to disclose.

José Natário: Conceptualization (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Flavio Rossetti: Conceptualization (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Meaningful choices of the scale factor a(t) in the FLRW metric (4) are obtained by solving the Friedmann equations, which result from the Einstein equations with an ideal fluid source. In three spatial dimensions, these are given by (see, for instance, Ref. 27),
(A1)
where K is the curvature of the spatial sections, ρm is the fluid’s energy density, pm is the fluid’s pressure, and Λ is the cosmological constant. If we assume that the fluid satisfies the linear equation of state
(A2)
where w is a constant (the square of the fluid’s speed of sound), then (A1) can be rewritten as
(A3)
where ρ0 ≥ 0 is an integration constant and the derivatives are now taken with respect to the conformal time. Interesting values of w are given in Table IV.
TABLE IV.

Fluid types for different values of w.

wFluid type
w < 0 Unphysical (imaginary speed of sound) 
w = 0 Dust 
w=13 Radiation 
w = 1 Stiff fluid 
w > 1 Unphysical (superluminal speed of sound) 
wFluid type
w < 0 Unphysical (imaginary speed of sound) 
w = 0 Dust 
w=13 Radiation 
w = 1 Stiff fluid 
w > 1 Unphysical (superluminal speed of sound) 
If ρ0 > 0, Λ = 0, K = 0, and w13, the solutions of (A3) are given by
(A4)
for an appropriate choice of ρ0 (amounting to a choice of units). In this work, we consider the following scale factors:
(A5)
The de Sitter universe corresponds to ρ0 = 0 and Λ > 0 (here, we chose units such that Λ = 3).
If ρ0 > 0, Λ = 0, K = −1, and w13, the solutions of (A3) are given by
(A6)
for an appropriate choice of ρ0. In particular, we consider the following scale factors:
(A7)
The anti-de Sitter, de Sitter, and Milne universes correspond to the cases where ρ0 = 0 and Λ = −3, Λ = 3, and Λ = 0, respectively.
Finally, if ρ0 > 0, Λ = 0, K = 1, and w13, the solutions of (A3) are given by
(A8)
for an appropriate choice of ρ0. In particular, we consider
(A9)
Again, the de Sitter universe corresponds to ρ0 = 0 and Λ = 3.
Let us show that the decay (76) is optimal by fixing a point x̄H3 and choosing initial data such that |ϕ(τ,x̄)|e32τ. This can be done as follows: given τ0 > 0, we choose the functions ϕ0 and ϕ1 such that ψ0 ≡ 0. From (70), this implies that
(B1)
with A=32cosh(τ0)12sinh2(τ0) and B = cosh(τ0) − 1. Next, we choose the radially symmetric function ϕ0(x)=φ(distH3(x,x̄)), where
(B2)
and N > 0 is a free parameter. Note that φ is smooth in (0, +∞) and moreover that its behavior at r = 0 implies that ϕ0 is of class C2. Plugging this choice of ϕ0 into the right-hand side of the spherical means solution (72) evaluated at x̄, we obtain
(B3)
where we used (55) and the substitution r = sτ0. By (B2), the above integral is constant for τ > τ0 + 1. We can choose τ0 such that this constant is non-zero (for τ0 = 1, τ = 2 and N = 50, the evaluation of the integral in (B3) using Mathematica gives 0.151503, and this value grows linearly with N), and so, after this choice of ϕ0 and ϕ1, (72) implies
(B4)
Since the C2 functions on a manifold can be approximated by smooth functions (Whitney approximation theorem, see, e.g., Ref. 28), we can choose smooth initial data that approximates the chosen ϕ0 and extend this construction to smooth initial data. The above reasoning disproves the decay rate of e−2τ in the conformal time variable stated by Abbasi and Craig in Ref. 1. Indeed, the inequality proved in Theorem 3.1 of their article seems to hold specifically for the term they choose, but not for the last terms in their solutions (3.9) and (3.10).
1.
B.
Abbasi
and
W.
Craig
, “
On the initial value problem for the wave equation in Friedmann-Robertson-Walker spacetimes
,”
Proc. R. Soc. A
470
,
20140361
(
2014
).
2.
J. L.
Costa
,
J.
Natário
, and
P. F. C.
Oliveira
, “
Decay of solutions of the wave equation in expanding cosmological spacetimes
,”
J. Hyperbolic Differ. Equ.
16
,
35
58
(
2019
).
3.
M.
Dafermos
and
I.
Rodnianski
, “
The wave equation on Schwarzschild-de Sitter spacetimes
,” arXiv:0709.2766 (
2007
).
4.
D.
Gajic
, “
Linear waves on constant radius limits of cosmological black hole spacetimes
,”
Adv. Theor. Math. Phys.
22
,
919
1005
(
2018
).
5.
A.
Galstian
,
T.
Kinoshita
and
K.
Yagdjian
, “
A note on wave equation in Einstein & de Sitter spacetime
,”
J. Math. Phys.
51
,
052501
(
2010
).
6.
A.
Galstian
and
K.
Yagdjian
, “
Microlocal analysis for waves propagating in Einstein & de Sitter spacetime
,”
Math. Phys. Anal. Geom.
17
,
223
246
(
2014
).
7.
P. M.
Girão
,
J.
Natário
, and
J. D.
Silva
, “
Solutions of the wave equation bounded at the Big Bang
,”
Classical Quantum Gravity
36
,
075016
(
2019
).
8.
S.
Klainerman
and
P.
Sarnak
, “
Explicit solutions of □u = 0 on the Friedmann-Robertson-Walker space-times
,”
Ann. Inst. Henri Poincare, Sect. A
35
,
253
257
(
1981
).
9.
H.
Ringström
, “
A unified approach to the Klein-Gordon equation on Bianchi backgrounds
,”
Commun. Math. Phys.
372
,
599
656
(
2019
).
10.
H.
Ringström
, “
Linear systems of wave equations on cosmological backgrounds with convergent asymptotics
,”
Astérisque
420
,
1
526
(
2020
).
11.
H.
Ringström
, “
Wave equations on silent big bang backgrounds
,” arXiv:2101.04939 (
2021
).
12.
H.
Ringström
, “
On the geometry of silent and anisotropic big bang singularities
,” arXiv:2101.04955 (
2021
).
13.
V.
Schlue
, “
Global results for linear waves on expanding Kerr and Schwarzschild de Sitter cosmologies
,”
Commun. Math. Phys.
334
,
977
1023
(
2015
).
14.
K.
Yagdjian
and
A.
Galstian
, “
Fundamental solutions of the wave equation in Robertson-Walker spaces
,”
J. Math. Anal. Appl.
346
,
501
520
(
2008
).
15.
W.
Craig
and
M.
Reddy
, “
On the initial value Problem for the electromagnetic wave Equation in Friedmann-Robertson-Walker space-times
,”
C. R. Math. Rep. Acad. Sci. Canada
41
,
45
56
(
2019
).
16.
D.
Starko
and
W.
Craig
, “
The wave equation in Friedmann-Robertson-Walker space-times and asymptotics of the intensity and distance relationship of a localised source
,”
J. Math. Phys.
59
,
042502
(
2018
).
17.
K.
Yagdjian
, “
Huygens’ principle for the Klein-Gordon equation in the de Sitter spacetime
,”
J. Math. Phys.
54
,
091503
(
2013
).
18.
J.
Wirth
, “
Solution representations for a wave equation with weak dissipation
,”
Math. Methods Appl. Sci.
27
,
101
124
(
2004
).
19.
F.
Rossetti
and
A.
Vañó-Viñuales
, “
Decay of solutions of the wave equation in cosmological spacetimes–A numerical analysis
,” arXiv:2302.06666 (
2023
).
20.
J.
Natário
and
A.
Sasane
, “
Decay of solutions to the Klein–Gordon equation on some expanding cosmological spacetimes
,”
Ann. Henri Poincaré
23
,
2345
2389
(
2022
).
21.
K.
Nishihara
, “
LpLq estimates of solutions to the damped wave equation in 3-dimensional space and their application
,”
Math. Z.
244
,
631
649
(
2003
).
22.
L.
Hörmander
,
Lectures on Nonlinear Hyperbolic Differential Equations
(
Springer
,
1997
).
23.
W.
Kulczycki
and
E.
Malec
, “
Gravitational waves in Friedman-Lemaitre-Robertson-Walker cosmology, material perturbations and cosmological rotation, and the Huygens principle
,”
Classical Quantum Gravity
34
,
135014
(
2017
).
24.
E.
Malec
and
G.
Wyleżek
, “
The Huygens principle and cosmological gravitational waves in the Regge-Wheeler gauge
,”
Classical Quantum Gravity
22
,
3549
(
2005
).
25.
J.
Natário
,
An Introduction to Mathematical Relativity
(
Springer
,
2021
).
26.
C.
Sogge
,
Lectures on Non-linear Wave Equations
(
International Press
,
2008
).
27.
S.
Chen
,
G. W.
Gibbons
,
Y.
Li
, and
Y.
Yang
, “
Friedmann’s equations in all dimensions and Chebyshev’s theorem
,”
J. Cosmol. Astropart. Phys.
2014
,
035
.
28.
J.
Lee
,
Introduction to Riemannian Manifolds
(
Springer
,
2018
).
Published open access through an agreement with Universidade de Lisboa Instituto Superior Tecnico