1 Introduction

Predictions of General Relativity (GR) have been tested with high accuracy ranging from the solar system to the cosmological scales. Despite this success, GR is an incomplete theory at short distance and time scales (for example, near black holes and cosmological singularities), and perhaps at large distances as well, where dark components and/or modifications of GR are invoked to explain the accelerated phase of the present Universe. It is expected that the inconsistencies at small scales can be resolved within the framework of quantum gravity (QG), which incorporates the principles of GR and quantum theory, and provides a description of the microstructure of space-time at the Planck scale.

Among the various attempts towards formulating a theory of QG, String/M-theory and Loop Quantum Gravity (LQG) remain as important candidates. A consequence of these theories is that space-time appears non-commutative (NC) at the fundamental level [1,2,3,4,5], and in some situations, may also exhibit a varying speed of light [6, 7]. This gives rise to non-local field theories and a modification of the dispersion relation of the quantum fields in a NC space-time. For example, one of the consequences of String Theory (as well as of M-Theory) is space-time non-commutativity [1], with the latter leading to modified dispersion relations [8]. Related to this is the fact that owing to quantum fluctuations, the usual canonical commutation relation also gets modified \([x,p]=i\hbar (1+\beta p^{2})\) [9,10,11,12,13,14] (see also Refs. [15,16,17,18]). It must be pointed out however, that there are several other approaches to QG that also predict the existence of a minimum measurable length, which in turn represents a natural cutoff and induces a departure from the relativistic dispersion relation. These approaches include space-time foam models [19,20,21], spin-network in LQG [22], space-time discreteness [23], spontaneous symmetry breaking of Lorentz invariance in string field theory [24] or in NC geometry [25], Horava’s approach [26, 27], and Doubly Special Relativity (DSR) [21, 28, 29]. In Ref. [30], the authors proposed an extension of DSR to include curvature, also known as Doubly General Relativity, in such a way that the geometry of space-time does depend on the energy E of the particle used to probe it (gravity’s rainbow) [21]. The general form of the modified dispersion relation (MDR) reads [31]

$$\begin{aligned} {{E}^{2}}f{{\left( E/{{E}_{P}} \right) }^{2}}-{{p}^{2}}\,c^2\,g{{\left( E/{{E}_{P}} \right) }^{2}}=m^{2}c^4, \end{aligned}$$
(1)

where the (rainbow) functions \(f(E/E_P)\) and \(g(E/E_P)\) depend on the Planck energy \(E_P=1.221\times 10^{19}\,\text {GeV}\) (for details see for instance Refs. [32,33,34,35,36,37,38,39]). Now, whenever \(f,g \ne 1\), i.e., one deviates from the standard relativistic dispersion relation, as we shall show below, the Planck radiation spectrum changes as well. This in turn may be able to explain the anomaly, which the Experiment to Detect the Global Epoch of Reionisation Signature (EDGES) collaboration has recently reported [40]. In the range \(z=15-20\), the EDGES collaboration found an anomalous absorption profile, with a brightness temperature minimum at \(z_{\mathrm{E}} \simeq 17.2\), which has a magnitude of about a factor of two greater than predicted by the \(\Lambda \)CDM model. It is this anomaly that we propose to explain using MDRs in this work. It turns out that the standard MDRs do not adequately explain the EDGES anomaly. However, by imposing redshift dependent MDR parameters, or by imposing a non-trivial power dependence for the MDRs, we are able to provide a viable explanation for the EDGES anomaly. A non-trivial power dependence of a MDR is also discussed in Ref. [41].

The rest of the paper is organized as follows. In the next section, we briefly review some of the important special cases of the above MDR. Following this, in Sect. 3, we estimate the parameters in the models that we consider from the results of the EDGES experiment. Finally, we summarize our results and conclude in Sect. 4.

2 MDR and modification of thermal spectrum

As stated in the Introduction, MDR is predicted by various theories of QG, and has the most general form of Eq. (1). The rainbow functions can in the most general case be expressed in a power series expansion (MacLaurin series) as \(f\left( E/{{E}_{P}} \right) =\sum _{n=0}^\infty \frac{f^{(n)}(0)}{n!}\left( E/E_P\right) ^n\) and \(g\left( E/{{E}_P} \right) =\sum _{n=0}^\infty \frac{g^{(n)}(0)}{n!}\left( E/E_P\right) ^n\), where constraints \(f(0)=1\) and \(g(0)=1\) must be imposed to obtain the standard relativistic dispersion relation at low energies. Here we consider some of the interesting special cases.

  • Case 1: \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}} \right) =\sqrt{1-\eta \, (E/E_P)^\omega }\), which is one of the most studied in literature. Here \(\eta \) is a parameter which signifies the effective scale of the modification, and \(\omega \) is the order of the modification. A complete theory of QG should fix both of them. However, in this work we study the modifications for different values of \(\eta \) and \(\omega \), and in particular, we consider three special cases. The first case is compatible with LQG and NC space-time [42, 43], while the next two are compatible with the linear and quadratic Generalized Uncertainty Principle (GUP) respectively [44, 45]:

    1. i)

      \(\omega =1\) and \(\eta >0\) \(\Longrightarrow \) \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}} \right) =\sqrt{1-\eta \, E/E_P}~,\)

    2. ii)

      \(\omega =1\) and \(\eta =\mp 2\alpha _0\) \(\Longrightarrow \) \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}} \right) =\sqrt{1\pm 2\alpha _0\, E/E_P}~,\)

    3. iii)

      \(\omega =2\) and \(\eta =2\beta _0\) \(\Longrightarrow \) \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}} \right) =\sqrt{1-2\beta _0\, (E/E_P)^2}~,\)

    where restrictions on \(\alpha _0\) from Ref. [44] have been relaxed to include both positive and negative values. In general, \(f\left( E/{{E}_{P}} \right) \ne 1\) and, specifically, in the presence of a strong gravitational field \(f\left( E/{{E}_{P}} \right) =1/\sqrt{-g_{00}}\), where \(g_{00}\) is the 00 component of the metric [44, 45]. However, in the dark ages, most of the hydrogen gas was in a very weak field, and, therefore, we can set \(f\left( E/{{E}_{P}} \right) =1\), as far as space-time curvature corrections to the MDR are concerned.

  • Case 2: \(f\left( E/{{E}_{P}} \right) =\frac{{{e}^{\alpha E/E_P}}-1}{\alpha E/{{E}_{P}}},\) \(g\left( E/{{E}_{P}} \right) =1\), proposed for explaining the spectra from GRBs at cosmological distances [19].

  • Case 3: \(f\left( E/{{E}_{P}}\right) =1\), \(g\left( E/{{E}_{P}}\right) =[1+(\lambda E)^\gamma ]^\delta =[1+\lambda '\, (E/E_P)^\gamma ]^\delta \), with \(\lambda '=(\lambda E_P)^\gamma \). The case \(\delta =1\) has been proposed for models in which a varying speed of light occurs [7]. The case \(\gamma =\delta =1\) has been proposed in Refs. [21, 46]. For \(\gamma =1\), \(\delta =1/2\), and \(\lambda =-\eta \), we recover Case 1. GUP provides another case for this form of \(g\left( E/{{E}_{P}}\right) \) [45]:

    1. i)

      \(\delta =1\), \(\gamma =1\) and \(\lambda '=\pm \alpha _0\) \(\Longrightarrow \) \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}}\right) =1\pm \alpha _0\, E/E_P\) .

Here \(\eta , \alpha _0, \beta _0, \alpha , \lambda '\) are dimensionless parameters, with \(\alpha _0\) and \(\beta _0\) to be the linear and quadratic GUP parameters, respectively. It is often assumed that \(\eta , \alpha _0, \beta _0, \alpha ,\lambda '\sim \mathcal{O}(1)\), so that the modifications of the dispersion relations are non-negligible at Planck scales. However, one may relax such a restriction and investigate signals of new physics at a new intermediate scale \(\lambda _{new}\sim {\eta }{\ell _{P}} \sim {\alpha _0}{\ell _{P}} \sim {\sqrt{\beta _0}}{\ell _{P}} \sim {\alpha }{\ell _{P}} \sim {\lambda '}{\ell _{P}}\). Such a length (energy) scale \(\lambda _{new}\) cannot exceed the well-tested electroweak length scale, \(\lambda _{EW}\sim {10^{17}}{\ell _{P}}\), so the consequent upper bound \(\eta \sim \alpha _0 \sim \sqrt{\beta _0}\sim \alpha \sim \lambda ' \le 10^{17}\).

The MDR given by Eq. (1), for the case of photons reads

$$\begin{aligned} E^2 - p^2 c^2 F^2 =0\,, \quad \text {with} \,\, F=\frac{g}{f}\,, \end{aligned}$$
(2)

so that, following Refs. [7, 46], one may derive the modified thermal spectrum \(\rho _{MDR}\).

The density of states per volume for photons (which have 2 polarization states) is written as

$$\begin{aligned} \Omega (p)=\frac{p^{2}}{\pi ^{2}\hbar ^{3}}. \end{aligned}$$
(3)

By considering the MDR in Eq. (2) and using \(\Omega (E)\,\mathrm {d}E=\Omega (p)\,\mathrm {d}p\), we obtain the following density of states

$$\begin{aligned} \Omega (E)=\frac{E^{2}}{\pi ^{2}{\hbar ^{3}}{\hat{c}}^{2}\tilde{c}}, \end{aligned}$$
(4)

where the two ‘speeds’ in the above equation turn out to be

$$\begin{aligned} {\hat{c}}=\frac{E}{p}=cF\,\,\,\,\,\,\mathrm {and}\,\,\,\,\,\,\tilde{c} =\frac{\mathrm {d}E}{\mathrm {d}p}=\frac{cF}{1-\frac{F'E}{F}}~, \end{aligned}$$
(5)

where \(F'=\mathrm {d}F/\mathrm {d}E\). Therefore, we can write the modified density of states as

$$\begin{aligned} \Omega (E)=\frac{E^2}{\pi ^2\hbar ^3c^3}\frac{1}{F^3}\left| 1-\frac{F'E}{F}\right| ~. \end{aligned}$$
(6)

The modified thermal spectrum is then obtained usingFootnote 1\(\rho _{MDR}(T,E)=2\pi \hbar E\,n(E)\,\Omega (E)\), where \(n(E)=\frac{1}{e^{\beta _T E}-1}\) is the Bose-Einstein distribution, \(\beta _T=\frac{1}{k_BT}\) is the inverse temperature and \(k_B\) is the Boltzmann constant. The modified thermal spectrum then reads as

$$\begin{aligned} \rho _{MDR}(E,T) = \rho (E,T) \frac{1}{F^3}{\left| 1-\frac{F'E}{F}\right| }\equiv \rho (E,T)R~, \end{aligned}$$
(7)

where

$$\begin{aligned} \rho (E,T)=\frac{2}{\pi \hbar ^2 c^3}\frac{E^3}{e^{\beta _T E}-1} \end{aligned}$$
(8)

is the standard thermal distribution of photons and R is the correction factor, formally defined in the following section. Note that the standard result from Eq. (8) is obtained from Eq. (7) when the MDR parameters vanish, i.e., \(\eta , \alpha _0, \beta _0, \alpha , \lambda '\longrightarrow 0\).

3 Experimental bounds

In this section we study the effects of the modified thermal spectrum given by Eq. (7), induced by the MDR given in Eq. (2), on the 21-cm cosmology. Details of 21-cm cosmology are given in Appendix A. This is related to the history of the universe, and represents a new framework for probing fundamental physics [49] (for other models see Refs. [50,51,52,53,54,55,56,57,58,59]). In particular, we focus on the recent release of the EDGES collaboration [40] (see also Ref. [60]).

EDGES High and Low band antennas probe the frequency ranges 90–200 MHz and 50–100 MHz, respectively, overall measuring the 21-cm signal within the redshift range \(z\in 6 - 27\), corresponding to an age of the Universe \(t_U\in (100 \mathrm{Myr} - 1 \mathrm{Gyr})\), i.e., the dark ages. This includes the epochs of reionization and cosmic dawn, in which the first astrophysical sources form. At \(z_{\mathrm{E}} \simeq 17.2\), the observed magnitude of the absorption lineFootnote 2 is about a factor of two greater than the one predicted by the \(\Lambda \)CDM model. At the redshift of the minimum of the 21-cm line, i.e., \(z_{\mathrm{E}} \simeq 17.2\), and frequency of CMB radiation, i.e., \(\nu _{21}(z_{\mathrm{E}})\simeq 78\,\mathrm{MHz}\), one has a 21-cm brightness temperature \(T_{21}(z_{\mathrm{E}})=-0.5_{-0.5}^{+0.2}\,\mathrm{K}\) (\(99\%\,\mathrm{C.L.}\), including estimates of systematic uncertainties). Since at \(z=z_E\) one has \((1+\delta _{\mathrm{b}})\,x_{H_I}(z_{\mathrm{E}}) \simeq 1\), Eq. (A5) implies \(T_{\gamma }(z_{\mathrm{E}})/T_S(z_{\mathrm{E}}) =15^{+15}_{-5.5}\) [49, 60]. Moreover, in the context of the \(\Lambda \)CDM model, one also gets

$$\begin{aligned} T_{\gamma }(z_{\mathrm{E}}) = T_{CMB}(z_{\mathrm{E}}) = T_{CMB,0}\,(1+z_\mathrm{E}) \simeq 50\,\mathrm{K} \end{aligned}$$
(9)

and

$$\begin{aligned} T_{\mathrm{gas}}(z_{\mathrm{E}}) \simeq T_{CMB}(z_{\mathrm{dec}}^\mathrm{gas})\,\left( \frac{1+z_{\mathrm{E}}}{1+z_{\mathrm{dec}}^{\mathrm{gas}}}\right) ^2 \simeq 6\,\mathrm{K}~, \end{aligned}$$
(10)

where \(z_{\mathrm{dec}}^{\mathrm{gas}}\simeq 150\) and \(T_{\mathrm{dec}}^{\mathrm{gas}} \simeq 410\,\mathrm{K}\) are the redshift and the temperature at the time when the gas and radiation decouple. Using (A5), one infers \(T_{21}(z_{\mathrm{E}}) \gtrsim -0.2\,\mathrm{K}\). Notice that the minimum is saturated for \(T_{S}(z_{\mathrm{E}}) = T_{\mathrm{gas}}(z_\mathrm{E})\), which corresponds to \(T_{\gamma }(z_{\mathrm{E}})/T_{\mathrm{gas}}(z_\mathrm{E}) \simeq 8\). As a consequence of these results, one finds that the best fit value for \(T_{21}(z_E)\) is \(\sim 2.5\) times lower than expected within the \(\Lambda \)CDM.

The 21-cm CMB photons absorbed at \(z_E\) fall clearly in the Rayleigh–Jeans tail since \(E_{21} \ll k_BT(z_{\mathrm{E}})\), where \(E_{21}\) is the hyperfine transition energy of the hydrogen atom. The energy density of the photons, i.e., Eq. (8), evaluated at \(T=T_{CMB}\), reads

$$\begin{aligned} \rho _{CMB}(E,z) = \frac{2}{\pi \hbar ^2 c^3}\,\frac{E^3}{e^{\beta _{T_{CMB}}\!(z)\,{E}}-1}~, \end{aligned}$$
(11)

where \(\beta _{T_{CMB}}\!(z)=\frac{1}{k_BT_{CMB}(z)}\). Only photons with energy \(E_{21}\) at \(z\simeq z_{\mathrm{E}}\) could be absorbed by the neutral hydrogen producing a 21-cm absorption global signal. For explaining the EDGES results, we consider the \(\rho _{MDR}\) given by Eq. (7). Therefore, we define the parameter R to study the discrepancy from the \(\Lambda \)CDM model as

$$\begin{aligned} R \equiv \frac{\rho _{MDR}(E_{21},z_E)}{\rho _{CMB}(E_{21},z_E)}=\frac{1}{F^3}{\left| 1-\frac{F'E_{21}}{F}\right| }~, \end{aligned}$$
(12)

with \(\rho _{MDR}\) and \(\rho _{CMB}\) defined in Eqs. (7) and (11), respectively. It may appear that such a modification may affect the optical depth \(\tau _\nu \) (introduced in Appendix A) and, therefore, the intensity and shape of the 21-cm line profile. However, as shown in Appendix B, such a modification does not affect \(\tau _\nu \) in any way. The experimental values from the EDGES experiment can then be explained by imposing (see Ref. [60] for details)

$$\begin{aligned} R = 2.15^{+2.15}_{-0.8} \,. \end{aligned}$$
(13)

Parameter R in Eq. (12) is then only a function of F, \(F'\) and E, since everything else except the relevant correction cancels out. Since we can in general write the rainbow functions f and g as a power series in \(E/E_P\), we can also write the function \(F=g/f\) as a power series expansion

$$\begin{aligned} F\left( E/{{E}_{P}} \right) =\frac{g\left( E/{{E}_{P}} \right) }{f\left( E/{{E}_{P}} \right) }=\sum _{n=0}^\infty \frac{F^{(n)}(0)}{n!}\, \left( E/E_P\right) ^n~. \end{aligned}$$
(14)

Note that \(F(0)=1\), which corresponds to the standard \(\Lambda \)CDM result. The parameter R from Eq. (12) for such a general expression reads

$$\begin{aligned} R=\frac{\left| 1-\sum _{n=1}^\infty \frac{F^{(n)}(n-1)}{n!}\left( {E}/{E_P}\right) ^{n}\right| }{\left[ \sum _{m=0}^\infty \frac{F^{(m)}}{m!}\left( {E}/{E_P}\right) ^{m}\right] ^4}~. \end{aligned}$$
(15)

Either Eq. (12) or Eq. (15) above can be used to estimate R for the cases studied here, compare with experimentally measured values and obtain bounds on the various parameters.

  • Case 1: \(f\left( E/{{E}_{P}} \right) =1\), \(g\left( E/{{E}_{P}} \right) =\sqrt{1-\eta \, (E/E_P)^\omega }~\). The ratio R reads

    $$\begin{aligned} R=\frac{|1-\left( 1-\frac{\omega }{2}\right) \eta \, (E/E_P)^\omega |}{(1-\eta \, (E/E_P)^\omega )^{5/2}}~ \end{aligned}$$
    (16)

    for arbitrary parameters \(\eta \) and \(\omega \). We take a look at the special cases:

    1. i)

      For \(\omega =1\) and \(\eta >0\), we have

      $$\begin{aligned} R=\frac{|1-\eta \, E/2E_{P}|}{(1-\eta \, E/E_P)^{5/2}}~. \end{aligned}$$
      (17)

      The ratio R is plotted as a function of \(\eta \) in Fig. 1. To fit the EDGES experimental bounds, the parameter is fixed at \(\eta = 6.5_{-3.6}^{+4.0}\times 10^{32}\).

    2. ii)

      For \(\omega =1\) and \(\eta =\mp 2\alpha _0\) we have

      $$\begin{aligned} R =\frac{|1\pm \alpha _0\, E/E_P|}{[1\pm 2\alpha _0\, E/E_P]^{5/2}}~. \end{aligned}$$
      (18)

      The ratio R is plotted as a function of \(\alpha _0\) for both branches in Fig. 2. However, only the branch with \(\eta =+2\alpha _0\) can fix \(\alpha _0\). To fit the EDGES experimental bounds, the parameter is fixed at \(\alpha _0=3.2_{-1.8}^{+2.0}\times 10^{32}\).

    3. iii)

      For \(\omega =2\) and \(\eta =2\beta _0\) we have

      $$\begin{aligned} R=\frac{1}{(1-2\beta _0\, (E/E_P)^2)^{5/2}}~. \end{aligned}$$
      (19)

      The ratio R is plotted as a function of \(\beta _0\) in Fig. 3. To fit the EDGES experimental bounds, the parameter is fixed at \(\beta _0=5.7_{-3.3}^{+3.9}\times 10^{65}\).

  • Case 2: \(f\left( E/{{E}_{P}} \right) =\frac{{{e}^{\alpha E/Ep}}-1}{\alpha E/{{E}_{P}}}\), \(g\left( E/{{E}_{P}} \right) =1~\). The ratio R reads

    $$\begin{aligned} R=\frac{e^{\alpha E/E_{P}} (e^{\alpha E/E_{P}} -1)^2}{(\alpha E/E_P)^2 }\,. \end{aligned}$$
    (20)

    The ratio R is plotted as a function of \(\alpha \) in Fig. 4. To fit the EDGES experimental bounds, the parameter is fixed at \(\alpha =7.8_{-4.7}^{+6.9}\times 10^{32}\).

  • Case 3: \(f=1\), \(g=\left[ 1+\lambda '\, (E/E_P)^\gamma \right] ^\delta \). The ratio R reads

    $$\begin{aligned} R=\frac{|1+(1-\delta \gamma )\lambda '\, (E/E_P)^\gamma |}{[1+\lambda '\, (E/E_P)^\gamma ]^{3\delta +1}}~, \end{aligned}$$
    (21)

    for arbitrary parameters \(\lambda '\), \(\gamma \) and \(\delta \). We take a look at the special case:

    1. i)

      For \(\delta =1\), \(\gamma =1\) and \(\lambda '=\pm \alpha _0~\), we have

      $$\begin{aligned} R=\frac{1}{(1\pm \alpha _0E/E_P)^4}~. \end{aligned}$$
      (22)

      The ratio R is plotted as a function of \(\alpha _0\) for both branches in Fig. 5. However, only the branch with \(\lambda '=-\alpha _0\) can fix \(\alpha _0\). To fit the EDGES experimental bounds, the parameter is fixed at \(\alpha _0=3.6_{-2.1}^{+2.7}\times 10^{32}\).

Fig. 1
figure 1

R vs \(\eta \) for fixed energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV

Fig. 2
figure 2

R vs \(\alpha _0\) for fixed energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV. The \(\eta =-2\alpha _0\) branch is presented in dash-dot blue and the \(\eta =+2\alpha _0\) branch is presented in solid black

Fig. 3
figure 3

R vs \(\beta _0\) for fixed energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV

At this point it should be stressed that the above plots indicate that the MDRs provided by cases 1, 2 and 3, give \((\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda ')\left| _{z=z_E} \approx 10^{32}\right. \) at redshift \(z=z_E\). These values are much larger than the bound set by the electroweak scale \(\lambda _{EW}/\ell _P\lesssim 10^{17}\). To verify the compatibility with known observations and obtain the bounds on the above parameters in the current epoch (\(z=0\)), we compare the experimental precision of the CMB temperature \(\left( \frac{\delta T}{T}\right) _{exp}=2\times 10^{-4}\) [61] (see also Refs. [62, 63]) of a perfect black body to the theoretical deviation due to MDRs in the current epoch

$$\begin{aligned} \frac{\delta T}{T}(z=0)= & {} (R(E)-1)\nonumber \\&\frac{\cosh {\left( \beta _{T_{CMB}}\!(0)\,E\right) } -1}{e^{\beta _{T_{CMB}}\!(0)\,E}-1}\,\frac{2}{\beta _{T_{CMB}}\!(0)\,E}~. \end{aligned}$$
(23)

In the above, R(E) is given by Eq. (12) and \(\beta _{T_{CMB}}\!(0)\) is given in terms of the CMB temperature in the current epoch. We obtain Eq. (23) by expressing \(\frac{\delta T}{T}\) from \(\rho _{MDR}(E,T)=\rho (E,T) R\approx \rho (E,T)+\frac{\mathrm {d}\rho }{\mathrm {d}T}(E,T)\,\delta T\). The parameters in the current epoch then must satisfy an upper bound of \((\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda ')\left| _{z=0} < 10^{28}\right. \) to be consistent with the observed CMB spectrum in the current epoch. The bound obtained from the electroweak experiments is stronger than that, so it should be used as the relevant MDR parameter bound in the current epoch. The EDGES anomaly at \(z=z_E\) combined with the above bound at \(z=0\) suggest that the above parameters should be increasing functions of the redshift z. Therefore, we also expect R to increase with z for a given energy E and have a value of \(R\approx 1\) at \(z=0\).

Fig. 4
figure 4

R vs \(\alpha \) for fixed energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV

Fig. 5
figure 5

R vs \(\alpha _0\) for fixed energy \(E=E_{12}\simeq 5.9 \times 10^{-6}\)eV. The \(\lambda '=+\alpha _0\) branch is presented in dash-dot blue and the \(\lambda =-\alpha _0\) branch is presented in solid black

The compatibility of such MDRs with epochs earlier than \(z_E\) should be taken into consideration as well. For example, in the epoch of the Big Bang Nucleosynthesis (BBN), at \(z\approx 3\times 10^8\) [64], a bound of \(\beta _0\lesssim 10^{87}\) was obtained in [65] for the quadratic GUP parameter \(\beta _0\), which corresponds to an upper bound \(\lesssim 10^{44}\) for the MDR parameters. Therefore, the values of the MDR parameters, measured by the EDGES anomaly are consistent with the BBN measurements, even if they increase to \(\sim 10^{44}\) at \(z\approx 3\times 10^{8}\). This supports the increasing trend of the redshift dependence of the MDR parameters and may in fact provide a clue in determining the exact form of this dependence. Estimations of the MDR parameters from the modified CMB spectrum would not be relevant in the BBN epoch, since it has not been created until the epoch of recombination at \(z=1090\) [64].

The standard MDRs used in this work can be found in Refs. [21, 42,43,44,45,46] as mentioned in Sect. 2, but they consider the MDR parameters as constants. The assumption that the MDR parameters are functions of another parameter, such as redshift, is fairly new. However, such an assumption is indirectly supported by Ref. [66], where the author finds a mass/radius dependent GUP parameter. This is also supported by the difference between estimations of the quadratic GUP parameter in tabletop experiments, where \(\beta _0>0\) [67,68,69,70,71], and astrophysical/cosmological observations, where \(\beta _0<0\) [66, 72,73,74,75,76]. This shows that the MDR parameters can in fact be dependent on scale or redshift.

Since the usual models of modified dispersion relations can not explain the EDGES anomaly, it is also legitimate to investigate if it can be explained by considering the cases analyzed here in which \(\eta , \alpha _0, \lambda '=10^{17}\), namely they are fixed to the electroweak scale, while \(\omega \) and \(\delta \) are treated as free parameters. We only consider cases 1 and 3, since case 2 has no other parameters to tweak. Also, we did not separately consider the special case 1, iii), because it is automatically considered as \(\omega \longrightarrow 2\).

Fig. 6
figure 6

R vs \(\gamma \) for fixed energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV. The \(\lambda '=+10^{17}\) branch is presented in dash-dot blue and the \(\lambda '=-10^{17}\) branch is presented in solid black

Fig. 7
figure 7

R vs \(\omega \) for fixed \(\eta ,\alpha _0=10^{17}\) and energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV. The solid black, the dash-dot blue and dotted red lines represent cases 1) i) and ii) (positive and negative branch) respectively

Fig. 8
figure 8

R vs \(\delta \) for fixed \(\lambda '=10^{17}\) and energy \(E=E_{12}\simeq 5.9\times 10^{-6}\)eV. The dash-dot blue, solid black and dotted red lines represent \(\gamma =0.49,\,0.50,\,0,51\) respectively

In Fig. 6, we plot R from Eq. (21) vs \(\gamma \) for fixed \(\lambda ' =\pm 10^{17}\) and for fixed \(\delta =1\). The values of R for \(\lambda '=10^{17}\) fall outside the EDGES bounds and cannot provide an explanation for the EDGES anomaly. However, the values of R for \(\lambda '=-10^{17}\) fall inside the EDGES bounds twice in a narrow range of \(\gamma \) around \(\gamma \approx 0.5\) and can therefore provide an explanation for the EDGES anomaly. Changing the \(\delta \) parameter only moves the peak to a different location.

At this point a number of comments are in order. First, the power dependencies on \(\omega \) and \(\delta \) of these cases are shown in Figs. 7 and 8, respectively. It is easily seen that the EDGES anomaly can be explained by powers \(\omega _{max}< 0.544\) and \(\delta _{max}< -0.05\). We also notice that we can only set an upper bound to the powers \(\omega \) and \(\delta \), since the electroweak length scale \(\lambda _{EW}\) is an upper bound for the new length scale. Second, the stringent values for \(\omega \) and \(\delta \), to resolve the EDGES anomaly, with their respective errors will be available in the future, when the true new length scale will be estimated and known with higher energy accelerators and astrophysical observations. Third, we also point out that power \(\delta <0\), which means that the correction decreases with increasing E as also seen in case 1. It may be noted that negative \(\delta \) and positive \(\lambda '\) is equivalent to positive \(\delta \) and negative \(\lambda '\) to leading order.

4 Conclusion

In this work, we have studied the possibility that MDRs can account for the recent results of the EDGES collaboration, which has discovered an anomalous absorption signal in the CMB radiation spectrum. This signal is larger by about a factor of 2 with respect to the expected value (assuming that the background is described by the \(\Lambda \)CDM model), i.e., the EDGES anomaly. In particular, we have shown that the most commonly considered MDRs, namely cases 1-3, lead to a modified thermal spectrum and to the subsequent estimation of the parameters \((\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda ')\left| _{z=z_E} \approx 10^{32}\right. \). Unfortunately, the parameter values at redshift \(z=z_E\) are outside the bounds allowed by, e.g., the electroweak experiments, since \((\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda ')\left| _{z=z_E} \gg \lambda _{EW}/\ell _P=10^{17}\right. \). However, given the precision of the CMB temperature in the current epoch, \(z=0\), we were able to constrain these parameters to an upper bound \((\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda ')\left| _{z=0}\right. <10^{28}\) to be consistent with the observed CMB black body spectrum. The estimation of the MDR parameters from the EDGES anomaly at \(z=z_E\), the bound obtained from electroweak experiments at \(z=0\) and the BBN bound at \(z\approx 3\times 10^8\) suggest that the MDR parameters should be functions of redshift z and as such could explain the EDGES anomaly. We can assume that the evolution of MDR parameters with time in the current epoch is slow or nearly constant, since we observe the same physics in all observable astrophysical objects such as distant galaxies. However, the time evolution of MDR parameters could have been faster in the early stages of the Universe as the EDGES anomaly suggests.

There is also another way out! As seen in Figs. 6, 7 and 8 and explained there, letting the powers \(\omega \), \(\gamma \) and \(\delta \) vary does also explain the anomaly for finite ranges of those powers. To precisely fit the EDGES experiment, and set \(\eta , \alpha _0, \sqrt{\beta _0}, \alpha , \lambda '=10^{17}\), bounded by the electroweak scale, we have studied the possibility of treating the powers \(\omega \), \(\gamma \) and \(\delta \) of the MDRs as free parameters and estimating upper bounds to their values. Similar results were found in Ref. [41]. However, MDRs with non-trivial power dependencies require further research to better understand their importance for QG theories.

The results in this work indicate that MDRs originating from existing theories and thought experiments with minimal measurable length can provide a mechanism which explains the EDGES anomaly only if the MDR parameters are increasing functions of redshift z. Also, if the true QG theory with minimum measurable length predicts non-trivial deformation parameters as obtained from Figs. 6, 7 and 8, then such a theory can also provide a viable mechanism to explain the EDGES anomaly as well. It will be interesting to study the consequences of such deformation parameters in various contexts, such as GRBs physics [19]. We hope to report on this in the future.