<<

CP3-19-12

Measuring the Inflaton Coupling in the CMB

Marco Drewes

Centre for Cosmology, Physics and Phenomenology, Universit´ecatholique de Louvain, Louvain-la-Neuve B-1348, Belgium

Abstract We study the conditions under which simple relations between the inflaton couplings and CMB observables can be established. The crucial criterion is to avoid feedback effects during reheating, which tend to introduce a complicated dependence of the CMB observables on a large number of microphysical parameters that prohibits the derivation of meaningful constraints on any individual one of them. We find that the inflaton coupling can be “measured” with cosmological data when the effective potential during reheating can be approximated by a parabola, and when the coupling constants are smaller than an upper bound that it determined by the ratios between the inflaton mass and the Planck mass or the scale of inflation. The power at which these ratios appear is determined by the power at which the inflaton appears in a given interaction term, and the value of the upper bound is largely independent of the type of produced particle. Our results show that next generation CMB observatories may be able to constrain the inflaton couplings for various types of interactions, providing an important clue to understand how a given model of inflation may be embedded into a more fundamental microphysical theory of nature.

Contents

1 Reheating and the CMB 2

2 Conditions to avoid feedback during reheating 5

3 The range of measurable inflaton couplings 7 3.1 Self-interactions and the effective potential ...... 7 arXiv:1903.09599v2 [astro-ph.CO] 29 Jan 2021 3.2 Interactions with other scalars ...... 9 3.3 Two body decays into ...... 11 3.4 Gauge production from -like coupling...... 11

4 Discussion and conclusion 12 1 Reheating and the CMB

Inflation and the CMB. Cosmic inflation [1–3] is currently the most popular expla- nation for the overall homogeneity and isotropy of the observable universe, as well as the small deviations from it that are visible in the Cosmic Microwave Background (CMB) and formed the seeds for galaxy formation. While the general idea that he universe underwent a phase of accelerated expansion before it entered the radiation dominated epoch is in good agreement with all cosmological data sets, it is unclear what mechanism drove this expansion1 and how it should be embedded into a more fundamental theory of nature. In the simplest scenarios the energy density during inflation is dominated by the po- tential energy density V(ϕ) of the condensate ϕ = hφi of a single scalar field φ, which acts like a if ϕ changes only very slowly. In this case the evolution of ϕ is governed by the equation of motion

ϕ¨ + (3H + Γ)ϕ ˙ + ∂ϕV(ϕ) = 0, (1) see e.g. Refs. [5–8] and references therein for a derivations in quantum field theory. Here a is the scale factor, H =a/a ˙ is the Hubble rate, V(ϕ) is the effective potential for ϕ and Γ takes into account dissipative effects due to the interaction of ϕ with other fields.2 Dissipation is usually thought to be negligible during inflation, but plays a key role for the conversion of the inflaton energy into relativistic when inflation ends. This conversion, which is known as cosmic reheating [15–22], sets the initial conditions for the hot , i.e., the radiation dominated epoch of cosmic history. Given the high energy scales that most inflationary models invoke, it seems unlikely that the mechanism of inflation and its embedding in an underlying theory of nature can be directly probed in experiments in the foreseeable future.3 An indirect probe of the interactions between the inflaton field and other particles is provided by the imprint that the reheating process leaves in CMB observables. In the present work we study under which circumstances observational constraints on the duration of the reheating epoch in a given inflationary model can be translated into an indirect “measurement” of the inflaton’s coupling to lighter particles. More precisely, we estimate the regime in which the reheating process is sufficiently simple that CMB observables can be related to a small number of microphysical parameters in a semi-analytic way.

The effect of reheating. Reheating takes a finite amount of time that should be re- garded as a separate reheating epoch in cosmic history. This epoch begins when the universe stops accelerating (the equation of state exceeds w > −1/3) and ends when the energy 2 density of radiation ρR exceeds the energy density ρϕ ' ϕ˙ /2 + V(ϕ) stored in ϕ, i.e., the universe becomes radiation dominated (w = 1/3). The modified equation of state during

1A partial overview of scenarios can e.g. be found in ref. [4]. 2There is no fluctuation-term on the RHS of (1), as one might expect from a generalised fluctuation- dissipation theorem, because our definition of ϕ includes an average over thermal fluctuations. Near the potential minimum, where reheating occurs, these average to zero. The equation of motion for φ itself would contain a noise term, see e.g. [9–14]. 3A noteworthy exception from this rule is described in ref. [23].

2 reheating [24] affects the redshifting of cosmological perturbations and thereby leaves an imprint in the CMB [25]. Since we can only observe the time-integrated effect, the quan- tities that the CMB is directly sensitive to are the duration of the reheating era in terms of e-folds Nre and the averaged equation of statew ¯re during reheating, with 1 Z Nre w¯re = w(N)dN. (2) Nre 0 Within a given model of inflation, knowledge of V(ϕ) fixes almost all relevant param- eters. It permits to compute the initial spectrum of cosmological perturbations, fixes the moment when inflation ends (∂ϕV(ϕ) > 3Hϕ˙) and determines the equation of state during 4 reheating (and therefore the expansion rate H). However, the duration of reheating Nre is strongly affected by Γ: reheating ends and ρϕ is rapidly depleted once Γ exceeds H. As far as the effect on the expansion history is concerned this occurs almost instantaneously [25]. At first sight the dependence on Γ is quite inconvenient, as it poses a fundamental uncertainty in the interpretation of CMB data in the context of inflationary models that cannot be resolved without specifying the interactions of φ with other degrees of freedom [26]. It is, however, possible to turn the tables and use this dependence to learn something about the reheating era from the CMB for a given model V(ϕ) [25, 27]. This realisation has motivated a considerable number of studies of reheating, cf. e.g. refs. [25, 27–55]. The constraint on Nre is often expressed in terms of the energy density at the end of reheating ρre or, equivalently, the reheating temperature, which is defined by the relation π2g  90 1/4 ∗ 4 p Tre ≡ ρR Γ=H = ρre . as Tre = ΓMpl 2 . (3) 30 π g∗ Γ=H

Here g∗ is the effective number of relativistic degrees of freedom. Tre can be interpreted as a physical temperature under the assumption that the plasma formed by the decay prod- ucts, commonly referred to as radiation,5 reaches thermal equilibrium instantaneously, cf. refs. [56–59] for a recent discussion. Alternatively the constraint on ρre can, via the Fried- mann equation and the relation Γ = H, be converted into a constraint on the dissipation coefficient. In terms of Nre, ρend andw ¯re it reads 1 ρ 1/2 Γ = end e−3(1+w ¯re)Nre/2 (4) Mpl 3 Within a given model of inflation, all parameters on the right hand side of eq. (4) can be constrained with CMB data.6 The details of this connection are not crucial for the 4The CMB anisotropies are generated during the slow roll phase and are not directly sensitive to the shape of the potential near its minimum. Within a specific theory the shape of V(ϕ) at its minimum can be only be inferred indirectly if all free parameters in the full potential can be fixed from observation.This leads to the practical limitation that one may not have enough cosmological observables to determine all parameters in V(ϕ). 5We collectively refer to all fields other than φ as “radiation”, irrespectively of their spin and charges. 6Throughout this work we assume a standard thermal history, i.e., radiation domination between the end of reheating and BBN. A critical discussion of this assumption can be found in [60].

3 present work, an explicit derivation of eq. (4) and expressions for Nre, ρend andw ¯re in terms of the spectral index ns, the scalar-to-tensor ratio r and the amplitude As of the CMB anisotropies are e.g. given in section 2.2 of ref. [46], following the approach taken in [36].

Reheating as a key to fundamental physics. Γ is determined by the inflaton’s in- teraction with other fields and contains information about the underlying particle physics model. In [32] it was pointed out that a determination of Γ could be used to constrain mi- crophysical parameters, i.e., parameter that are considered “fundamental” in an underlying particle physics model. This is not trivial because reheating can be a highly complicated far from equilibrium process [61] that can only be described with numerical simulations that usually do not allow to extract any simple relations between the parameters. We shall in the following distinguish between the inflaton’s self-interactions {λi}, the inflaton couplings to other fields {gi} and the interactions of those other fields with each other {αi}. Here {λi}, {gi} and {αi} symbolically refer to sets of microphysical parameters (the set {αi} e.g. includes the SM gauge interactions). The relation (4) can, together with the known relations between (Nre, ρend, w¯re) and (As, ns, r), be used to impose constraints on individual microphysical parameter if Γ in good approximation only depends on a small number of those parameters. In the following we assume for simplicity that one single interaction dominates the transfer of energy from ϕ during reheating and symbolically label the relevant by g.This is, for instance, the case if reheating is driven by elementary decays of inflaton quanta in vacuum. The main complication arises from feedback effects of the produced particles on the dissipation rate. The feedback not only makes the reheating process highly non-linear and difficult to simulate, but it also implies that Γ in general depends on a large number of parameters {αi} in the underlying particle physics model, cf. e.g. [6, 62–66] for specific examples. Hence, the regime where simple relations between CMB observables and the inflaton coupling g to other fields can be established can be identified with the absence of feedback. Feedback typically becomes relevant when the occupation numbers in the plasma reach unity. This can be avoided if g is sufficiently small, so that the transfer of energy from ϕ to radiation is rather slow and the occupation numbers are reduced due to redshift before they can reach unity.7 On the other hand, g cannot be made arbitrarily small because the reheating temperature (3) is typically proportional to g for perturbative reheating. If g is too small, the universe fails to reheat, i.e., the reheating temperature is smaller than the temperature needed to explain the observed abundances of light elements with big bang nucleosynthesis (BBN). These considerations have been used to constrain the inflaton coupling in α-attractor models in refs. [36, 46]. It was found that the constraint (4) on Γ can, roughly speaking, be converted into a constraint on g if g < 10−5. The reheating temperature is high enough to be consistent with the observational constraint from BBN if g > 10−15. This leaves

7Since the relation (4) constrains Γ at the moment when reheating ends, it is strictly speaking not necessary to demand that the occupation numbers remain below unity at al times as long as the feedback for not enhance Γ so much that it exceeds the rate of Hubble expansion H. More precisely, our results still apply as long as Γ is dominated by elementary processes in the moment when Γ = H.

4 about ten orders of magnitude for which the inflaton coupling can be “measured” from the CMB. These results reported in ref. [46] are rather independent of the specific interaction that couples the inflaton to radiation (i.e., whether g is a coupling to or fermions). In the present work we use simple well-known relations to investigate the origin of this general behaviour, which we find is not restricted to α-attractor models.

2 Conditions to avoid feedback during reheating

In the following we restrict the discussion to scenarios where the cosmos is reheated dur- ing oscillations of ϕ around its potential minimum. Other forms of reheating, such as tachyonic instabilities, are generally driven by non-perturbative processes that introduce a complicated dependence of Γ on time and the {αi}. In this situation, one can broadly distinguish the following mechanisms to reheat the universe,

(i) elementary decays of inflaton quanta in vacuum,

(ii) annihilations of inflaton quanta,

(iii) scatterings of inflaton quanta with other particles.

(iv) non-perturbative production of particles from the time-dependent background pro- vided by ϕ,

(v) gravitational particle production.

We shall neglect (v) in the following, as it tends to be sub-dominant when other channels exist. The conditions under which these mechanisms lead to a resonant production of particles have been studied in detail in ref. [20]. Here we use the results of this work for parametric estimates. We have explicitly checked their validity for some of specific interactions we consider in the following in refs. [32, 46]. The non-perturbative particle production (iv) tends to be very efficient, leading to O[1] occupation numbers, strong feedback, and resonant particle production within a few oscillations [16]. The resonance can be avoided in two ways: Either the produced particles decay into other particles within one oscillation, or the expansion of the universe redshifts the particles’ momenta out of the resonance bands. Since the decay rates of the produced particles necessarily depend on the {αi}, Γ in such scenarios necessarily depends on the {αi}. The redshifting, on the other hand, depends on the equation of state, which is dominated by ϕ, and therefore in good approximation independent of the {αi}. One can qualitatively distinguish between a broad resonance that produces particles over a broad range of momenta, and a narrow resonance that only affects relatively narrow momentum bands. A broad resonance cannot be avoided by redshifting, but only by decays. Hence, scenarios where the non-perturbative mechanism (iv) produces particles over a broad range of momenta, Γ always depends on the {αi}. If the production only affects a narrow range of momenta, on the other hand, the resonance can potentially be avoided by redshifting alone, and Γ is a function of g only.

5 The mechanism (iii) can only be operational when there is already a plasma of particles. The efficiency of the scatterings depends on the composition and the occupation numbers in the plasma, which depend on time and the {αi} [67]. Therefore we demand that such processes are sub-dominant to ensure that Γ is independent of the {αi}. The elementary processes (i) and (ii) can occur in vacuum because they only require inflaton quanta in the initial state. The rates for both can be enhanced if the occupation numbers in the final states are large. Particularly dangerous are 1 → 2 decays, which produce monochromatic momentum distributions and can exhibit a narrow resonance due to induced transitions. A dependence of Γ on the {αi} can only be avoided if the produced particles’ momenta are efficiently redshifted out of the resonance bands. In summary, use the following conditions to ensure that Γ is a function of g only,

I) The momenta of the particles are concentrated in narrow bands.

II) Hubble expansion shifts the momenta out of these bands before the occupation num- bers reach unity.

We further demand that

III) the reheating temperature must be high enough to ensure primordial nucleosynthesis, i.e., Tre > TBBN. Note that III) is a hard requirement imposed by observation, while I) and II) only restrict the range of validity of the assumption that Γ is independent of the {αi}. In the following we shall assume that the potential is approximately parabolic dur- ing the inflaton oscillations. The reason is that terms of higher power in the potential generically lead to violations of adiabaticity and therefore non-perturbative particle pro- duction, cf. e.g. [24]. To illustrate this, let us consider a symmetric potential with a 1 2 2 1 4 6 at vanishing field value, which we expand as V = 2 mφφ + 4! λφ + O[φ ]. 2 2 2 2 The effective mass for the condensate reads Mϕ = mφ + λϕ + λT /4!, with T the temperature. For ϕ  (m2 /λ + T 2/4!)1/2 one can neglect the quadratic terms, and √ φ p 8 ϕ(t) ' ϕendcn(1/ 2, λ/6ϕendt), where ϕend is the field value at the end of inflation. 9 p Expanding the elliptic cosine one finds ϕ(t) ' ϕend cos(ωt) with ω ' λϕ/6 ϕend. Plug- 2 2 2 ging this back into the frequency of the field mode k,Ωk = Mϕ + k , it is straightforward 2 to see that it is impossible to fulfil the adiabaticity condition (∂tΩk)/Ωk for low momentum modes, and one necessarily encounters non-perturbative particle production that typically leads to feedback effects. To be conservative, we exclude this possibility here. This trans- 2 lates into the rather strong condition λ < (mφ/ϕend) on λ, which, however, is fulfilled for most of the parameter points considered in [46].

8This approximation is conservative because the oscillations usually do not immediately start after the end of inflation, but there is a brief period of “fast roll”. Also neglecting this is conservative in the present context because ϕ would decrease during this period. √ 9 P∞ e−π(n+1/2) To be explicit, cn(1/ 2, x) = A n=0 cn cos(bnx) with A = 4.79257, cn = 1+e−2π(n+1/2) , bn = π 3.70815 (2n + 1).

6 During oscillations in a parabolic potential, the equations for the mode functions Xk of fields that couple to ϕ can be brought into the form of the usual Mathieu equation, 00   Xk (z) + Ak − 2q cos(2z) Xk(z) = 0. (5) with z ∼ tω (ignoring prefactors and phase shifts). The Mathieu equation shows insta- bilities that indicate resonant particle production. The conditions I) and II) now read [20] I) The width of the resonance bands of (5) scales as ∼ q, so the condition to avoid a broad resonance translates into q  1.10 II) In the regime q < 1 the rate of resonant particle production can be estimated as ∼ qω. The effective width of the first band is roughly ∼ qω/2, and the time that particles spend within this band can be estimated as qH−1. The particle numbers then grow q2ω 2 as exp( 2H ). Hence, we require q ω  H. For simplicity and to be conservative we impose the condition at the beginning of reheating, which is sufficient to ensure that the condition holds during reheating in the scenarios that we consider here.

3 The range of measurable inflaton couplings

For the following discussion of the conditions under which feedback during reheating can be avoided for specific interactions. It is convenient to introduce the small parameters

pl ≡ mϕ/Mpl , ϕ ≡ mϕ/ϕend. (6)

3.1 Self-interactions and the effective potential Splitting φ = ϕ + η into the expectation value ϕ and fluctuations η, we shall first consider the production of inflaton particles η due to self-interactions. For later use, we consider general couplings of the form n m 4−m−n L ⊃ c(n,m)φ χ Λ (7) between φ and another scalar field χ, with dimensionless couplings c(n,m) and Λ some n l n−l m 4−m−n l heavy mass scale. These generate terms l c(n,m)ϕ η χ Λ with l ≤ n. ϕ(t) = l l [ϕend cos(mφt)] , can be decomposed into a series of terms ∝ ϕ cos(ωt), with ω given by integer multiples of mφ up to lmφ. These operators generate time dependent masses for η and χ, leading to particle production. At tree level, only the terms with (m, l) = (0, n − 2) and (m, l) = (2, n) contribute.11

10 This is a conservative criterion for which we assumed that Ak can be small for small momenta. If the produced particles are heavier than ω, then the range of Ak may be restricted to be larger than 1, in which case the width of the bands is smaller. 11A systematic way to include quantum corrections has e.g. been developed in [68]. Instead of the + mode functions the authors study the time evolution of the so-called statistical propagator ∆χ (x, y) ≡ 1 2 h{φ(x), φ(y)}i − ϕ(x)ϕ(y), which provides a measure for the energy density stored in χ-fluctuations. The + equation of motion for ∆χ in Wigner space is of the form (16), but contains integral kernels involving self-energies on the right hand side.

7 The terms with m = 0 and n−l = 2 in (7) lead to tree level contributions to the effective 2 2 1/2 η masses and can therefore induce self-resonances. The criterion ϕend  (mφ/λ+T /4!) used at the beginning of this section is a necessary, but not a sufficient one to avoid a self-resonance. The mode equation for η then receives contributions from all terms in (7) with n − l = 2 and m = 0, " # 2 2 X n−2 4−n η¨k(t) + k + mφ + n(n − 1)c(n,0)ϕ Λ ηk(t) = 0 (8) n Here we consider only renormalisable interactions, as terms with n + m > 4 are suppressed by negative powers of Λ. This leaves us with the terms for n = 3 and n = 4. Assuming that only one or the other is important at a given time, we can use the following relations

(2ω )2 Λϕ A = k , q = 12c end , z = (tω + π)/2 [for the c Λφ3 interaction] (9) k ω2 (3,0) ω2 (3,0) and, identifying c(4,0) = λ/4! to connect to the previous section,

ω2 λ ϕ2 λ A = k + 2q , q = end , z = tω + π/2 [for the φ4 interaction],. (10) k ω2 8 ω2 4!

2 2 Using the standard relations H = ρ/(3Mpl) and ρend ' 4/3 × Vend with Vend = V(ϕend), this can be translated into

2 r 1/4 r 2 r 1/4 r ω 2 Vend ω ω ω 2 Vend ω ω λ  8 2 , λ  8 , c(3,0)  , c(3,0)  . (11) ϕend 3 ϕend ϕend Mpl Λϕend 3 ϕend Λ Mpl

1 2 2 Using Vend ' 2 mφϕend, ω = mφ and the notation (6), this simplifies to m m λ  82 , λ  5.5 3/21/2 , c   φ , c  5.5 1/21/2 φ . (12) ϕ ϕ pl (3,0) ϕ Λ (3,0) ϕ pl Λ Before moving on, we shall question our restriction to values n ≤ 4, which we justified by the suppression of higher order terms in (8) by powers of ϕ/Λ. This is justified if one interprets (7) as an effective field theory with cutoff Λ, assuming that this cutoff must be larger than ϕ. In the framework of perturbation theory one would typically identify Λ with the mass of some heavy degree of freedom that has been integrated out, and there is no reason why ϕ should remain smaller than the cutoff, as ϕ does not correspond to the mass of any physical particle. However, the non-perturbative mechanism (iv) can produce particle with masses much larger than ω. This is evident from the fact that the Mathieu 12 equation (5) has resonance bands for Ak  1. We therefore require that Λ in terms with

12In this context it is instructive to compare the requirement q  1 from condition I) to the adiabaticity 2 3 condition (∂tΩk)/Ωk that has to be violated to efficiently produce heavy particles. For the c(3,0)φ inter- 3/2 4 action the adiabaticity condition translates into 2q  Ak , for for quartic interaction λφ /4! one finds 2 2 3/2 1/2 3/2 q  (6Ak + 9Ak − 3 ((1 + 6Ak) − 1)) /6. For Ak  1 this coincides with the condition 2q  Ak , and for Ak  1 is reproduces the commonly used condition 2q  Ak.

8 n + m > 4 is larger than all other physical scales. In terms with n + m < 4, on the other hand, one should not interpret Λ as the cutoff in an effective field theory, but rather as a conveniently chosen scale to make the coupling dimensionless. We may therefore think of the ratio mφ/Λ in (12) as a parameter of order one. Ignoring all numerical factors, one can therefore roughly summarise (12) as

2 3/2 1/2 1/2 1/2 c(4,0)  min(ϕ, ϕ pl ) , c(3,0)  min(ϕ, ϕ pl ) (13) In summary, the previous considerations show that the upper bound on the inflaton self- couplings from the requirement that Γ is independent of the {αi} is given by the ratios (6) of mφ and other physical scales, and that the power at which these ratios appear is determined by the power n at which φ appears in the respective term. More precisely, the 2 power of the small parameters is given by the power at which ϕ appears in Ωk.

3.2 Interactions with other scalars The mode equation for χ then receives contributions from all terms in (7) with n − l = 0 and m = 2, " # 2 2 X n 2−n χ¨k(t) + k + mχ + 2c(n,2)ϕ Λ χk(t) = 0. (14) n Comparing to (5), we can inter that the parameter q quantifies the amplitude of the time- dependent mass and has a contribution that parametrically scales as

n 2−n 2 q ∼ c(n,2)ϕ Λ /ω (15) for the tree level mass term. Both conditions I) and II) impose upper bounds on q and therefore c(n,2). From (15) it is clear that the power n at which φ appears in (7) is decisive, and that the upper bound on the coupling c(n,2) is stronger for larger n. We therefore classify the different interactions according to n, which turns out to be more relevant than e.g. the spin of the produced particles.

Scalar two body decays. Consider coupling of the form gφχ2 to another scalar χ, where g = c(1,2)Λ in the notation of (7). The equation for χ-modes in Minkowski space reads h 2 2 i χ¨k(t) + k + mχ + 2gϕ(t) χk(t) = 0 (16) We can bring the mode equation (16) into the form (5) with 2 (2ωk) gϕend Ak = 2 , q = 4 2 , z = (tω + π)/2 (17) ω mφ

2 g p 2 Using the well-known expression Γϕ→χχ = 1 − (mϕ/(2mχ)) θ(mφ/(2mχ) − 1) ' 8πmφ 2 g . for the perturbative decay rate the conditions I)-III) translate into 8πmφ r  1/4 g mφ g mφ 1/4 g TBBN 32 ϕend < , ϕend < Vend , > p π g∗ . mφ 4 mφ 24Mpl mφ mφMpl 45

9 we can simplify this to  1/4 g 1 g √ g √ g∗  ϕ ,  0.2 ϕpl ,  3π pl , (18) mφ 4 mφ TBBN 106.75 where 106.75 is the value of g∗ in the symmetric phase of the SM and we have rounded all 13 numerical prefactors to ratios of integers. If we set TBBN ' 10 MeV, ϕend ∼ Mpl and ignore factors of order one, we get the very simple condition s  g 1/4 GeV g m 10−10 ∗ < < 10−19 φ . (19) 106.75 mφ mφ GeV This very simple condition marks the range of inflaton couplings for which the expression 2 Γϕ→χχ = g /(8πmφ) can be used to translate the relation (4) into a constraint on the inflaton coupling. One can also use the requirement that the lower bound on the coupling 6 remains below the upper bound to derive a condition on mφ, e.g. mφ > 10 GeV for large field models in which (19) can be applied.

2 2 h 2 2 Inflaton annihilations. We now consider the operator c(2,2)φ χ ≡ 4 φ χ . Comparing (14) and (16) it is straightforward to obtain the expressions for Ak and q by replacing λ → h in (10), ω2 h ϕ2 λ A = k + 2q , q = end , z = tω + π/2 [for the φ4 interaction],, (20) k ω2 8 ω2 4! and obtain from (13)

2 3/2 1/2 h  min(ϕ, ϕ pl ). (21) In order to evaluate condition III) we need the anihilation rate during adiabatic harmonic 2 2 2 2 h ϕ p 2 h ϕ oscillations [6, 69] Γϕϕ→χχ = 1 − (mϕ/mχ) θ(mφ/mχ − 1) ' . This rate 256πmφ 256πmϕ can intuitively be interpreted as an annihilation of two φ-quanta into two χ-quanta. Deter- 2 mining TR is a bit more tricky because Γ ∝ ϕ , while H ∝ |ϕ|. As a result,√ the ratio Γ/H 2 2 (averaged over a few oscillations) decreases with time. For ϕend > 256 2πmϕ/(3h Mpl) this ratio is larger than one already when the oscillations commence, for smaller ϕend it remains below one at all times, and the universe is not reheated. The requirement to reheat the universe through inflaton annihilations can be translated into the condition on 1/4p √ h > 16 × 2 π/3 ϕpl. Comparing this to (21) shows that it is not possible to reheat the universe with elementary annihilations (as it was e.g. assumed in [69]): one either encounters a resonance, or the universe is never reheated. Before moving on, we should add that several details have been wiped under the carpet here. We implicitly assumed an averaging over a few oscillations when comparing Γ to H without specifying how exactly this averaging should be done. This may be important because the ϕ-dependence of Γ introduces a non-linearity in (1), so that the motion of ϕ may not be that of a simple 1 2 2 damped harmonic oscillator even if we set V = 2 mφϕ . 13This is conservative because in reality ϕ is typically sub-Planckian when the oscillations commence.

10 3.3 Two body decays into fermions. It is usually assumed reheating primarily produces bosons because quantum statistical ef- fects lead to an enhancement of the rate, while they suppress the production of fermions due to Pauli’s principle. However, fermions can also be produced, and they can also experi- ence resonant enhancement [22]. This can have interesting implications even if the production is sub-dominant compared to that of bosons, e.g. for leptogenesis or production [70, 71]. For the present discussion the fermion production is only relevant if it dominates the reheating process and sets Nre. This requires that the coupling to fermions is considerably larger than that to light bosons. We here consider the production of a Dirac ¯ fermion ψ of mass mψ through a Yukawa coupling yφψψ. A resonance parameter can be 2 ϕ2 y2ϕ2 defined as q = y 2 , which yields q ∼ 2 if the potential is approximately quadratic ω mφ [22, 71]. The underlying mechanism for the resonant production of fermions is quite differ- ent from bosons. In particular, redshifting does not reduce the efficiency of the resonance, but rather helps to avoid Pauli blocking. Hence, we cannot use condition II) here. In spite of the differences in the underlying microphysics, it turns out that the condition q  1 still provides a good criterion to determine the largest coupling where reheating is perturbative [22].14 The lower bound on y can again be obtained from condition III). Comparing the y2 perturbative decay rate Γϕ→ψψ = 8π mφ to Γϕ→χχ reveals that we can literally apply the lower bound in conditions (18)-(19) with the replacement g → y. We finally arrive at the mφ range  1/4 TBBN √ g∗ 1 3π pl  y  ϕ (22) mφ 106.75 4

3.4 production from axion-like coupling. −1 ˜µν Let us now consider axion-like couplings αφΛ FµνF to the field strength tensor Fµν of vector bosons.15 We assume that the gauge bosons are massless for ϕ = 0, which is justified in the symmetric phase of gauge theories, and consider an Abelian gauge theory. This situation is qualitatively different from the scalar interaction because the vertex is momentum dependent, and the time dependent mass term of the gauge fields have a non- trial Lorentz structure. The Mathieu equation for the circular polarisations of the field reads [74]  k m  ∂2 + k2 ∓ 4α φ ϕ a = 0 (23) t Λ k±

2 k ϕ mφ Comparison to (16) gives Ak = (2k/ω) with k = |k| and q = 8α ω Λ ω . It is immediately clear that condition I) is always violated for large momenta. On the other hand this does not necessarily imply non-perturbative particle production in this case because for

14It should be noted that the treatment used in [22] underestimates the fermion production in the presence of large bosonic occupation numbers [72]. This is, however, not relevant in the present discussion because in that case the coupling to bosons would set Nre. 15Though may not be needed to solve the strong CP problem [73], axion-like couplings appear in many theories beyond the .

11 sufficiently small α almost all modes remain adiabatic. Here we restrict ourselves to a very simple estimate. Both, the perturbative decay and a narrow resonance produce particles with energy mφ/2, so that we can fix the momentum to this value and use (9) with the 2 replacements ωk → mφ/2 and g → αmφ/Λ in Ak and q. We obtain the upper bounds on α from conditions I) and II) by making this replacement in (18)-(19). To obtain the lower 3 α2 mφ bound from BBN, we consider the elementary decay rate into vector bosons Γϕ→VV = 4π Λ2 . g2 Comparison with Γϕ→χχ = reveals that the lower bound can be obtained by making 8πmφ 2 the same replacement g → αmφ/Λ in Γ and multiplying the total rate with a factor 2 to account for the two photon polarisations in the final state, leading to

 1/4 mφ 1 mφ √ mφ TBBN √ g∗ α  ϕ , α  0.2 ϕpl , α  2 pl . (24) Λ 4 Λ Λ mφ 106.75 4 Discussion and conclusion

We studied the conditions under which a simple relation between the inflaton coupling g and the CMB parameters (As, ns, r) can be established, so that g can be “measured” in the CMB. A crucial requirement for this is to avoid feedback effects during reheating, which in general introduce a complicated dependence of (As, ns, r) on many parameters, and in particular on the masses and couplings of the particles produced during reheating. At the classical level, to which re restrict ourselves here, feedback can be studied in terms of the instability bands of the Mathieu equation (5). While quantum corrections will certainly change the details of the reheating process, the classical analysis should be sufficient for the parametric estimates that we are interested in here. The conditions I) and II) outlined in section 2 under which feedback can be avoided can be translated into upper bounds on the parameter q in the Mathieu equation (5). If we identify the inflaton coupling g with the coefficients c(n,2), c(n+2,0) in 7, the simple scaling relation (15) suggests that the upper bounds on g from conditions I) and II) can be expressed in terms of n and the ratios (6), ϕ = mφ/ϕend and pl = mφ/Mpl,

m 2−n √ m 2−n g  n φ , g   n−1/2 φ (25) ϕ Λ pl ϕ Λ

The explicit results (21), (18) and (13) confirm this scaling. Following the comments after (12), we should interpret Λ as a cutoff in an effective field theory when considering operators with mass dimension larger than four, while it Λ should be viewed as a conveniently chosen mass scale to make the coupling dimensionless in terms with smaller mass dimension. In √ the latter case the constrain (25) is equal or weaker than requiring g < min(ϕ, plϕ) (depending on whether Λ is e.g. identified with mφ or mχ). In the former case, for n > 2, the factors Λ/mφ  1 on the right hand sides of the inequalities in (25) help to alleviate the upper bound on g. The strongest bound therefore holds for n = 2, i.e., for the coupling φ2χ2 and the self-interaction φ4. Allowing for derivative couplings softens these bounds, as it leads to positive powers of (Λ/k) on the right hand sides of the inequalities in (25),

12 which are larger than unity for the most relevant modes k = 0 and k = mφ/2. The explicit result (24) confirms this conclusion. Finally, similar scalings apply to fermions, with the main difference being that the time-dependent contribution to the squared effective fermion n 2 n 1−n mass scales as ∼ (gϕ ) , so that we can expect constraints of the form g  ϕ(mφ/Λ) . This scaling is consistent with (22). In summary, we studied the possibility to “measure” the inflaton coupling to other fields in scenarios where reheating occurs during coherent oscillations of the inflaton field ϕ around its potential minimum. We find that a simple relation between the inflaton coupling g and the CMB parameters (As, ns, r) can be established when the effective potential V(ϕ) can be approximated by a parabola during the oscillations, and when g is not too large. The upper bound on g from this requirement is more or less independent of the type of produced particle (scalar, fermion, vector); it is primarily determined by the ratios between the inflaton mass and other physical scales in the problem (such as Mpl and ϕend). The power at which these ratios appear is determined by the power n at which φ appears in p a given operator. All renormalisable couplings should be smaller than mφ/Mpl, which puts considerable limitations on the perspectives to “measure” the inflaton couplings in n small field models. In addition, the upper bound scales as ∝ (mφ/ϕend) , implying that the range of “measurable” inflaton couplings is the largest when mφ and ϕend are of the same order, and both near the Planck scale. The strongest constraints exist for the coefficients of renormalisable Higgs-portal type couplings φ2χ2, and for the inflaton self-coupling φ4. The aforementioned limitations can be alleviated by factors Λ/mφ for non-renormalisable interactions. The perspective to probe the inflaton couplings to other fields with CMB data is ex- tremely interesting from both, the viewpoint of cosmology and particle physics. It provides an indirect probe of fundamental microphysical parameters that most likely can never be measured directly in the laboratory, but have an immense impact on the evolution of the cosmos by setting the stage for the hot big bang. Further, it can provide one of the very few observational hints on how a given model of inflation can be embedded in a more fundamental theory of nature. The precision at which ns and r are constrained by current data is insufficient to extract any meaningful bound on the inflaton coupling. However, reducing the error bar on ns by an order of magnitude would be sufficient to pin down the order of magnitude of the inflaton coupling [46].

Acknowledgments.

I would like to thank Gilles Buldgen, Drazen Glavan, Jin U Kang, Lei Ming, and Vincent Vennin for their helpful comments on the draft.

References

[1] A. A. Starobinsky, A New Type of Isotropic Cosmological Models Without Singularity, Adv. Ser. Astrophys. Cosmol. 3 (1987) 130–133.

13 [2] A. H. Guth, The Inflationary Universe: A Possible Solution to the Horizon and Flatness Problems, Phys. Rev. D 23 (1981) 347–356. [3] A. D. Linde, A New Inflationary Universe Scenario: A Possible Solution of the Horizon, Flatness, Homogeneity, Isotropy and Primordial Monopole Problems, Phys. Lett. B 108 (1982) 389–393. [4] J. Martin, C. Ringeval and V. Vennin, Encyclopædia Inflationaris, Phys. Dark Univ. 5-6 (2014) 75–235, [1303.3787]. [5] A. Berera, I. G. Moss and R. O. Ramos, Warm Inflation and its Microphysical Basis, Rept. Prog. Phys. 72 (2009) 026901, [0808.1855].

[6] K. Mukaida, K. Nakayama and M. Takimoto, Fate of Z2 Symmetric , JHEP 12 (2013) 053, [1308.4394]. [7] Y.-K. E. Cheung, M. Drewes, J. U. Kang and J. C. Kim, Effective Action for Cosmological Scalar Fields at Finite Temperature, JHEP 08 (2015) 059, [1504.04444]. [8] G. Buldgen, M. Drewes, J. U. Kang and U. R. Mun, General Markovian Equation for Scalar Fields in a Slowly Evolving Background, 1912.02772. [9] K.-c. Chou, Z.-b. Su, B.-l. Hao and L. Yu, Equilibrium and Nonequilibrium Formalisms Made Unified, Phys. Rept. 118 (1985) 1–131. [10] M. Morikawa, Classical Fluctuations in Dissipative Quantum Systems, Phys. Rev. D 33 (1986) 3607. [11] C. Greiner and S. Leupold, Stochastic interpretation of Kadanoff-Baym equations and their relation to Langevin processes, Annals Phys. 270 (1998) 328–390, [hep-ph/9802312]. [12] J. Yokoyama, Fate of oscillating scalar fields in the thermal bath and their cosmological implications, Phys. Rev. D 70 (2004) 103511, [hep-ph/0406072]. [13] D. Boyanovsky, K. Davey and C. M. Ho, Particle abundance in a thermal plasma: Quantum kinetics vs. Boltzmann equation, Phys. Rev. D 71 (2005) 023523, [hep-ph/0411042]. [14] A. Anisimov, W. Buchmuller, M. Drewes and S. Mendizabal, Nonequilibrium Dynamics of Scalar Fields in a Thermal Bath, Annals Phys. 324 (2009) 1234–1260, [0812.1934]. [15] A. D. Dolgov and D. P. Kirilova, ON PARTICLE CREATION BY A TIME DEPENDENT SCALAR FIELD, Sov. J. Nucl. Phys. 51 (1990) 172–177. [16] J. H. Traschen and R. H. Brandenberger, Particle Production During Out-of-equilibrium Phase Transitions, Phys. Rev. D 42 (1990) 2491–2504.

14 [17] L. Kofman, A. D. Linde and A. A. Starobinsky, Reheating after inflation, Phys. Rev. Lett. 73 (1994) 3195–3198, [hep-th/9405187]. [18] Y. Shtanov, J. H. Traschen and R. H. Brandenberger, Universe reheating after inflation, Phys. Rev. D 51 (1995) 5438–5455, [hep-ph/9407247]. [19] D. Boyanovsky, H. J. de Vega, R. Holman and J. F. J. Salgado, Analytic and numerical study of preheating dynamics, Phys. Rev. D 54 (1996) 7570–7598, [hep-ph/9608205]. [20] L. Kofman, A. D. Linde and A. A. Starobinsky, Towards the theory of reheating after inflation, Phys. Rev. D 56 (1997) 3258–3295, [hep-ph/9704452]. [21] P. B. Greene, L. Kofman, A. D. Linde and A. A. Starobinsky, Structure of resonance in preheating after inflation, Phys. Rev. D 56 (1997) 6175–6192, [hep-ph/9705347]. [22] P. B. Greene and L. Kofman, Preheating of fermions, Phys. Lett. B 448 (1999) 6–12, [hep-ph/9807339]. [23] F. Bezrukov and D. Gorbunov, Light inflaton Hunter’s Guide, JHEP 05 (2010) 010, [0912.0390]. [24] S. Antusch, D. G. Figueroa, K. Marschall and F. Torrenti, Energy distribution and equation of state of the early Universe: matching the end of inflation and the onset of radiation domination, Phys. Lett. B 811 (2020) 135888, [2005.07563]. [25] J. Martin and C. Ringeval, First CMB Constraints on the Inflationary Reheating Temperature, Phys. Rev. D 82 (2010) 023511, [1004.5525]. [26] J. Martin, C. Ringeval and V. Vennin, Shortcomings of New Parametrizations of Inflation, Phys. Rev. D 94 (2016) 123521, [1609.04739]. [27] P. Adshead, R. Easther, J. Pritchard and A. Loeb, Inflation and the Scale Dependent Spectral Index: Prospects and Strategies, JCAP 02 (2011) 021, [1007.3748]. [28] R. Easther and H. V. Peiris, Bayesian Analysis of Inflation II: Model Selection and Constraints on Reheating, Phys. Rev. D 85 (2012) 103533, [1112.0326]. [29] L. Dai, M. Kamionkowski and J. Wang, Reheating constraints to inflationary models, Phys. Rev. Lett. 113 (2014) 041302, [1404.6704]. [30] J. B. Munoz and M. Kamionkowski, Equation-of-State Parameter for Reheating, Phys. Rev. D 91 (2015) 043521, [1412.0656]. [31] J. Martin, C. Ringeval and V. Vennin, Observing Inflationary Reheating, Phys. Rev. Lett. 114 (2015) 081303, [1410.7958]. [32] M. Drewes, What can the CMB tell about the microphysics of cosmic reheating?, JCAP 03 (2016) 013, [1511.03280].

15 [33] J. L. Cook, E. Dimastrogiovanni, D. A. Easson and L. M. Krauss, Reheating predictions in single field inflation, JCAP 04 (2015) 047, [1502.04673].

[34] R.-G. Cai, Z.-K. Guo and S.-J. Wang, Reheating phase diagram for single-field slow-roll inflationary models, Phys. Rev. D 92 (2015) 063506, [1501.07743].

[35] V. Domcke and J. Heisig, Constraints on the reheating temperature from sizable tensor modes, Phys. Rev. D 92 (2015) 103515, [1504.00345].

[36] Y. Ueno and K. Yamamoto, Constraints on α-attractor inflation and reheating, Phys. Rev. D 93 (2016) 083524, [1602.07427].

[37] R. J. Hardwick, V. Vennin, K. Koyama and D. Wands, Constraining Curvatonic Reheating, JCAP 08 (2016) 042, [1606.01223].

[38] K. D. Lozanov and M. A. Amin, Equation of State and Duration to Radiation Domination after Inflation, Phys. Rev. Lett. 119 (2017) 061301, [1608.01213].

[39] I. Dalianis, G. Koutsoumbas, K. Ntrekis and E. Papantonopoulos, Reheating predictions in Theories with Derivative Coupling, JCAP 02 (2017) 027, [1608.04543].

[40] R. Kabir, A. Mukherjee and D. Lohiya, Reheating constraints on K¨ahlermoduli inflation, Mod. Phys. Lett. A 34 (2019) 1950114, [1609.09243].

[41] S.-M. Choi and H. M. Lee, Inflection point inflation and reheating, Eur. Phys. J. C 76 (2016) 303, [1601.05979].

[42] P. Cabella, A. Di Marco and G. Pradisi, Fiber inflation and reheating, Phys. Rev. D 95 (2017) 123528, [1704.03209].

[43] K. Nozari and N. Rashidi, Perturbation, non-Gaussianity, and reheating in a Gauss-Bonnet α-attractor model, Phys. Rev. D 95 (2017) 123518, [1705.02617].

[44] A. Di Marco, P. Cabella and N. Vittorio, Constraining the general reheating phase in the α-attractor inflationary cosmology, Phys. Rev. D 95 (2017) 103502, [1705.04622].

[45] S. Bhattacharya, K. Dutta and A. Maharana, Constraints on K¨ahlermoduli inflation from reheating, Phys. Rev. D 96 (2017) 083522, [1707.07924].

[46] M. Drewes, J. U. Kang and U. R. Mun, CMB constraints on the inflaton couplings and reheating temperature in α-attractor inflation, JHEP 11 (2017) 072, [1708.01197].

[47] D. Maity, Constraints through decaying inflaton: maximum reheating temperature, 1709.00251.

16 [48] I. Dalianis and Y. Watanabe, Probing the BSM physics with CMB precision cosmology: an application to , JHEP 02 (2018) 118, [1801.05736].

[49] D. Maity and P. Saha, Connecting CMB anisotropy and cold dark phenomenology via reheating, Phys. Rev. D 98 (2018) 103525, [1801.03059].

[50] A. Di Marco, G. Pradisi and P. Cabella, Inflationary scale, reheating scale, and pre-BBN cosmology with scalar fields, Phys. Rev. D 98 (2018) 123511, [1807.05916].

[51] D. Maity and P. Saha, (P)reheating after minimal Plateau Inflation and constraints from CMB, JCAP 07 (2019) 018, [1811.11173].

[52] N. Zhang, Y.-B. Wu, J.-W. Lu, C.-W. Sun, L.-J. Shou and H.-Z. Xu, Constraints on the generalized natural inflation after Planck 2018, Chin. Phys. C 44 (2020) 095107, [1807.03596].

[53] N. Rashidi and K. Nozari, α-Attractor and reheating in a model with noncanonical scalar fields, Int. J. Mod. Phys. D 27 (2018) 1850076, [1802.09185].

[54] A. Nautiyal, Reheating constraints on Inflation, Phys. Rev. D 98 (2018) 103531, [1806.03081].

[55] F. Takahashi and W. Yin, ALP inflation and Big Bang on Earth, JHEP 07 (2019) 095, [1903.00462].

[56] A. Mazumdar and B. Zaldivar, Quantifying the reheating temperature of the universe, Nucl. Phys. B 886 (2014) 312–327, [1310.5143].

[57] K. Harigaya and K. Mukaida, Thermalization after/during Reheating, JHEP 05 (2014) 006, [1312.3097].

[58] K. Harigaya, M. Kawasaki, K. Mukaida and M. Yamada, Dark Matter Production in Late Time Reheating, Phys. Rev. D 89 (2014) 083532, [1402.2846].

[59] K. Mukaida and M. Yamada, Thermalization Process after Inflation and Effective Potential of Scalar Field, JCAP 02 (2016) 003, [1506.07661].

[60] R. Allahverdi et al., The First Three Seconds: a Review of Possible Expansion Histories of the Early Universe, 2006.16182.

[61] M. A. Amin, M. P. Hertzberg, D. I. Kaiser and J. Karouby, Nonperturbative Dynamics Of Reheating After Inflation: A Review, Int. J. Mod. Phys. D 24 (2014) 1530003, [1410.3808].

[62] J. Garcia-Bellido, D. G. Figueroa and J. Rubio, Preheating in the Standard Model with the Higgs-Inflaton coupled to gravity, Phys. Rev. D 79 (2009) 063531, [0812.4624].

17 [63] K. Mukaida and K. Nakayama, Dynamics of oscillating scalar field in thermal environment, JCAP 01 (2013) 017, [1208.3399].

[64] K. Mukaida and K. Nakayama, Dissipative Effects on Reheating after Inflation, JCAP 03 (2013) 002, [1212.4985].

[65] S. Enomoto, O. Fuksi´nska and Z. Lalak, Influence of interactions on particle production induced by time-varying mass terms, JHEP 03 (2015) 113, [1412.7442].

[66] K. Freese, E. I. Sfakianakis, P. Stengel and L. Visinelli, The can delay Reheating after Inflation, JCAP 05 (2018) 067, [1712.03791].

[67] M. Drewes and J. U. Kang, The Kinematics of Cosmic Reheating, Nucl. Phys. B 875 (2013) 315–350, [1305.0267].

[68] J. Berges and J. Serreau, Parametric resonance in quantum field theory, Phys. Rev. Lett. 91 (2003) 111601, [hep-ph/0208070].

[69] J. P. B. Almeida, N. Bernal, J. Rubio and T. Tenkanen, Hidden Inflaton Dark Matter, JCAP 03 (2019) 012, [1811.09640].

[70] G. F. Giudice, M. Peloso, A. Riotto and I. Tkachev, Production of massive fermions at preheating and leptogenesis, JHEP 08 (1999) 014, [hep-ph/9905242].

[71] P. B. Greene and L. Kofman, On the theory of fermionic preheating, Phys. Rev. D 62 (2000) 123516, [hep-ph/0003018].

[72] J. Berges, D. Gelfand and J. Pruschke, Quantum theory of fermion production after inflation, Phys. Rev. Lett. 107 (2011) 061301, [1012.4632].

[73] W.-Y. Ai, J. S. Cruz, B. Garbrecht and C. Tamarit, Absence of CP violation in the strong interactions, 2001.07152.

[74] P. Carenza, A. Mirizzi and G. Sigl, Dynamical evolution of axion condensates under stimulated decays into , Phys. Rev. D 101 (2020) 103016, [1911.07838].

18