This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Cosmic Radio Background from Primordial Black Holes at Cosmic Dawn

Zhihe Zhang Department of Physics, College of Science, Tibet University, Lhasa 850000, P. R. China Key Laboratory of Cosmic Rays (Tibet University), Ministry of Education, Lhasa 850000, P. R. China National Astronomical Observatories, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100101, P. R. China School of Astronomy and Space Science, University of Chinese Academy of Sciences, Beijing 100049, P. R. China    Bin Yue [email protected] National Astronomical Observatories, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100101, P. R. China    Yidong Xu National Astronomical Observatories, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100101, P. R. China    Yin-Zhe Ma School of Chemistry and Physics, University of KwaZulu-Natal, Westville Campus, Private Bag X54001, Durban 4000, South Africa NAOC-UKZN Computational Astrophysics Centre (NUCAC), University of KwaZulu-Natal, Durban, 4000, South Africa National Institute for Theoretical and Computational Sciences (NITheCS), South Africa    Xuelei Chen National Astronomical Observatories, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100101, P. R. China School of Astronomy and Space Science, University of Chinese Academy of Sciences, Beijing 100049, P. R. China Department of Physics, College of Sciences, Northeastern University, Shenyang, 110819, China Center of High Energy Physics, Peking University, Beijing 100871, China    Maoyuan Liu [email protected] Department of Physics, College of Science, Tibet University, Lhasa 850000, P. R. China Key Laboratory of Cosmic Rays (Tibet University), Ministry of Education, Lhasa 850000, P. R. China
Abstract

The presence of an extra radio background besides the cosmic microwave background has important implications for the observation of the 21-cm signal during the cosmic Dark Ages, Cosmic Dawn, and epoch of Reionization. The strong absorption trough found in the 21-cm global spectrum measured by the EDGES experiment, which has a much greater depth than the standard model prediction, has drawn great interest to this scenario, but more generally it is still of great interest to consider such a cosmic radio background (CRB) in the early Universe. To be effective in affecting the 21-cm signal at early time, such a radio background must be produced by sources which can emit strong radio signals but modest amount of X-rays, so that the gas is not heated up too early. We investigate the scenario that such a radio background is produced by the primordial black holes (PBHs). For PBH with a single mass, we find that if the PBHs’ abundance log(fPBH)\log(f_{\rm PBH}) (ratio of total PBH mass density to total matter density) and mass satisfy the relation log(fPBH)1.8log(M/M)3.5\log(f_{\rm PBH})\sim-1.8\log(M_{\bullet}/{\rm M}_{\odot})-3.5 for 1MM300M1\,{\rm M}_{\odot}\lesssim M_{\bullet}\lesssim 300{\rm M}_{\odot}, and have jet emission, they can generate a CRB required for reproducing the 21-cm absorption signal seen by the EDGES. The accretion rate can be boosted if the PBHs are surrounded by dark matter halos, which permits lower fPBHf_{\rm PBH} value to satisfy the EDGES observation. In the latter scenario, since the accretion rate can evolve rapidly during the Cosmic Dawn, the frequency (redshift) and depth of the absorption trough can determine the mass and abundance of the PBHs simultaneously. For absorption trough redshift \sim 17 and depth 500\sim-500 mK, it corresponds to M1.05MM_{\bullet}\sim 1.05\,{\rm M}_{\odot} and fPBH1.5×104f_{\rm PBH}\sim 1.5\times 10^{-4}.

preprint: APS/123-QED

I Introduction

The 21-cm global spectrum is the result of the thermal evolution of the intergalactic medium (IGM), and cosmic star formation history during the Dark Ages, Cosmic Dawn, and Reionization Furlanetto et al. (2006); Pritchard and Loeb (2012). Standard cosmology models typically predicts an absorption trough at Cosmic Dawn, with depth up to 200\sim 200 mK Chen and Miralda-Escudé (2004, 2008). The location, depth, and detailed shape of the absorption trough reflect complicated interactions between the IGM hydrogen atoms, the X-ray and Lyα\alpha photons from the first stars and galaxies, and the cosmic microwave background (CMB) photons from the early Universe.

There have been several ground-based single-antenna experiments to measure the 21-cm global spectrum, such as EDGES Bowman et al. (2018), SARAS Singh et al. (2017), SCI-21 Voytek et al. (2014), LEDA Price et al. (2018), BIGHORNS Sokolowski et al. (2015), REACH de Lera Acedo et al. (2022), etc. There are also proposals to carry out the measurements in space, such as the DARE/DAPPER Burns (2021), DSL Chen et al. (2021), and FARSIDE Burns et al. (2021) projects. The EDGES experiment detected an absorption trough on 21-cm global spectrum Bowman et al. (2018). The frequency is in agreement with the 21-cm signal from the era of Cosmic Dawn (z17z\sim 17Bowman et al. (2018), though the depth is 500\sim 500 mK, much deeper than the maximum limit in the standard cosmology model Xu et al. (2018, 2021). The subsequent SARAS-3 experiment obtained a result contradictory with that of the EDGES Singh et al. (2022), but it may be still too early to reach a final conclusion  Chen (2022), as the experiment systematics are far from understood. In this paper we will take the view that the EDGES result represents a possible 21-cm global spectrum which requires a theoretical explanation.

Several theoretical models have been proposed to explain the surprising result of the EDGES experiment. These include the exotic dark matter particles which couple to and cool the baryonic gas Barkana (2018), and the extra cosmic radio background (CRB) in addition to the CMB Feng and Holder (2018); Ewall-Wice et al. (2018). Although the existence of this deep absorption signal is still debated, the scenario that at Cosmic Dawn there is an extra radio background and its impact on the 21-cm global spectrum is nevertheless an interesting and novel perspective. Many objects may produce an extra CRB at Cosmic Dawn, for example the first galaxies in small halos with high efficiency of star formation and radio production Mirocha and Furlanetto (2019), dark matter annihilations Yang (2018), the first black holes Ewall-Wice et al. (2018); Mirabel and Rodríguez (2022), and the primordial black holes (PBHs).

PBHs may arise from small-scale primordial curvature fluctuations, and form in the early Universe, much earlier than the formation of the first generation stars (Pop III stars) Cang et al. (2022a, 2021, b). In addition, they can be much more abundant than the black hole seeds formed after the death of Pop III stars. Therefore PBHs have the potential to generate a significant amount of the CRB through its accretion and jet mechanisms. In most models, the initial PBH mass ranges from the Planck mass to 105M\sim 10^{5}{\rm M}_{\odot} Carr (2005); Harada and Carr (2005). PBHs with mass M<1015M_{\bullet}<10^{15} g would have evaporated through Hawking radiation Carr and Rees (1979); Carr et al. (2010). The more massive ones could however survive and grow by accreting the ambient medium Harada and Carr (2005). They may form accretion disks with a significant angular momentum Poulin et al. (2017), or even launch relativistic jets like astrophysical black holes Takhistov et al. (2022). So far no PBH has been detected, but the abundance has been constrained by several observations, such as the gravitational wave Bird et al. (2016), gravitational lensing Barnacka et al. (2012), the impact of accretion on the cosmic thermal evolution Poulin et al. (2017), CMB spectra distortion Serpico et al. (2020); Green and Kavanagh (2021) and CMB anisotropy power spectrum Cang et al. (2021, 2022b).

The CRB from PBHs has been investigated by many authors. Most of these works focus on the Hawking radiation of smaller PBHs Mittal et al. (2022); Cang et al. (2022a); Saha and Laha (2022); Mittal and Kulkarni (2022), or the free-free emission from accretion disks of stellar-mass PBHs Tashiro et al. (2021); Ali-Haïmoud and Kamionkowski (2017). In these scenarios, it is not easy for the PBHs to produce sufficient CRB to boost the 21-cm absorption significantly, otherwise the IGM would be ionized/heated too early, in conflict with the observational constraints on the reionization history Acharya et al. (2022). It is generally unquestionable that the PBH can produce some X-rays Mena et al. (2019). If there is also radio emission associated with the X-ray radiation, then PBHs can produce a sufficient amount of CRB and boost the 21 cm absorption to the EDGES level Hasinger (2020). Since these PBHs also contribute to building up the Cosmic X-ray Background (CXB), the observed CXB also put tight limits on the contribution to CRB from PBHs Ziparo et al. (2022).

Astrophysical black holes can usually launch powerful jet which produces strong synchrotron radiation Paliya et al. (2017); Narayan et al. (2022). The same effect may also exist for PBH. PBH usually sits in the gaseous environment of the interstellar medium, so its accretion disk can naturally launch outflows such as winds and jets. Ref. Takhistov et al. (2022) showed that PBHs with significant spin can sustain powerful relativistic jets and generate associated cocoons. Ref. Hasinger (2020) pointed out that the PBH can play important role in amplifying the initial magnetic field. If the jet has a strong magnetic field, then its synchrotron could dominate its radio radiation. As a result, the synchrotron from PBHs may generate a much stronger CRB than the free-free emission, and influence the 21-cm absorption significantly. This is the basic motivation of our work. In this paper, we investigate if in this case PBHs can produce a strong CRB that boosts the 21-cm absorption signal at the Cosmic Dawn, while at the same time only heats the IGM slightly.

The paper is organized as follows. In Sec. II we consider the accretion rate evolution of the PBHs. We estimate the contribution to the CRB, the thermal history of the IGM, and the influence on the 21-cm signal at Cosmic Dawn, if an empirical radio luminosity-X-ray luminosity relation is applicable to the PBHs. In Sec. III we calculate the synchrotron radiation of PBHs using a jet model and evaluate the final influence on the 21-cm signal. Finally, we summarize the results and discuss some issues of this problem in Sec. IV.

II Methods

II.1 The accretion rate and radiative efficiency of PBHs

II.1.1 Naked PBHs

First we consider a naked PBH, by which we mean one without a surrounding dark matter halo. The accretion process of a massive PBH has been studied comprehensively in the past years (e.g. (Ricotti, 2007; Ricotti et al., 2008; Ali-Haïmoud and Kamionkowski, 2017; Tashiro et al., 2021)). Ignoring the angular momentum of the accreted gas, the accretion is described well by the spherical symmetric model and the accretion rate is Bondi (1952)

M˙B=4πλρ¯bveffrB2,\dot{M}_{\rm B}=4\pi\lambda\bar{\rho}_{\rm b}v_{\rm eff}r^{2}_{\rm B}, (1)

where the veff=vrel2+cs2v_{\rm eff}=\sqrt{v_{\rm rel}^{2}+c_{\rm s}^{2}}, vrelv_{\rm rel} is the velocity of the relative motion between the black hole and the ambient gas, and csc_{\rm s} is the sound speed. λ\lambda is a dimensionless parameter describing the net accretion; and ρ¯b=Ωbρcr(1+z)3\bar{\rho}_{\rm b}=\Omega_{\rm b}\rho_{\rm cr}(1+z)^{3} is the cosmic mean baryon density. ρcr=2.775h2×1011MMpc3\rho_{\rm cr}=2.775\,h^{2}\times 10^{11}\,{\rm M}_{\odot}\,{\rm Mpc}^{-3} is the critical density of the present Universe. The Bondi radius of a PBH with mass MM_{\bullet} is defined as the radius for which the escape velocity equals veffv_{\rm eff}:

rB=GMveff2.r_{\rm B}=\frac{GM_{\bullet}}{v_{\rm eff}^{2}}. (2)

The dark matter may have a relative motion with respect to the gas, once the two are both decoupled from each other and from the photons Ricotti et al. (2008). A PBH is therefore expected to have a proper motion relative to its ambient gas medium. In linear regime, the dispersion of the relative velocity vL21/2min[1,z/103]×30km/s\langle v_{\rm L}^{2}\rangle^{1/2}\approx{\rm min}[1,z/10^{3}]\times 30~{}{\rm km/s}, and the effective velocity is Ali-Haïmoud and Kamionkowski (2017),

veff{csvL21/2vL21/2cscsvL21/2cs.v_{\rm eff}\approx\begin{cases}\sqrt{c_{\rm s}\langle v_{\rm L}^{2}\rangle^{1/2}}&\langle v_{\rm L}^{2}\rangle^{1/2}\gg c_{\rm s}\\ c_{\rm s}&\langle v_{\rm L}^{2}\rangle^{1/2}\ll c_{\rm s}.\end{cases} (3)

The relative motion reduces the accretion rate, since M˙Bveff3\dot{M}_{\rm B}\propto v_{\rm eff}^{-3} .

The sound speed csc_{\rm s} depends on the kinetic temperature of the IGM gas, TkT_{\rm k}, while the gas is also heated by the X-ray radiation from PBHs, so TkT_{\rm k} must be solved self-consistently from the cosmic thermal evolution equations with X-ray heating from PBHs.

The parameter λ\lambda is determined by the PBH surrounding environment as well as various physical processes such as Hubble expansion, Compton drag and cooling of the CMB photons, and negative feedback. We calculate it following the Refs. Ricotti (2007); Ali-Haïmoud and Kamionkowski (2017). For a naked PBH,

λ(γ,β)\displaystyle\lambda(\gamma,\beta) =\displaystyle= 1λiso[λad+(λisoλad)(γc288+γc2)0.22]\displaystyle\frac{1}{\lambda_{\rm iso}}\left[\lambda_{\rm ad}+(\lambda_{\rm iso}-\lambda_{\rm ad})\left(\frac{\gamma^{2}_{\rm c}}{88+\gamma^{2}_{\rm c}}\right)^{0.22}\right] (4)
×\displaystyle\times [1(1+βc+1)2exp(9/23+βc3/4)],\displaystyle\left[\frac{1}{(\sqrt{1+\beta_{\rm c}}+1)^{2}}\exp\left(\frac{9/2}{3+\beta_{\rm c}^{3/4}}\right)\right],

where λad=(3/5)3/2/40.12,λiso=e3/2/41.12\lambda_{\rm ad}=(3/5)^{3/2}/4\approx 0.12,\lambda_{\rm iso}=e^{3/2}/4\approx 1.12; βc\beta_{\rm c} and γc\gamma_{\rm c} are the dimensionless Compton drag and cooling rates from the surrounding CMB photons respectively Ali-Haïmoud and Kamionkowski (2017):

βc\displaystyle\beta_{\rm c} =\displaystyle= 43xeσTργmpcrBveff\displaystyle\frac{4}{3}\frac{x_{\rm e}\sigma_{\rm T}\rho_{\gamma}}{m_{\rm p}c}\frac{r_{\rm B}}{v_{\rm eff}} (5)
γc\displaystyle\gamma_{\rm c} =\displaystyle= 83xeσTργmec(1+xe)rBveff,\displaystyle\frac{8}{3}\frac{x_{\rm e}\sigma_{\rm T}\rho_{\gamma}}{m_{\rm e}c(1+x_{\rm e})}\frac{r_{\rm B}}{v_{\rm eff}}, (6)

where xex_{\rm e} is the electron fraction in the IGM, cc is the speed of light, σT=6.65×1025\sigma_{\rm T}=6.65\times 10^{-25} cm2 is the Thomson scattering cross-section, and ργ=π2TCMB4/15\rho_{\gamma}=\pi^{2}T^{4}_{\rm CMB}/15 is the energy density of the CMB.

In Fig. 1 we plot the evolution of dimensionless accretion rate m˙M˙B/(LEdd/c2)\dot{m}\equiv\dot{M}_{\rm B}/(L_{\rm Edd}/c^{2}) for naked PBHs of masses 100,101,102,103M10^{0},10^{1},10^{2},10^{3}{\rm M}_{\odot} by thin curves, where

LEdd=1.26×1038Mergs1L_{\rm Edd}=1.26\times 10^{38}M_{\bullet}~{}{\rm erg~{}s}^{-1}

is the Eddington luminosity. At z17z\sim 17, m˙\dot{m} ranges from 105\sim 10^{-5} for M=1MM_{\bullet}=1\,{\rm M}_{\odot} to 102\sim 10^{-2} for M=103MM_{\bullet}=10^{3}\,{\rm M}_{\odot}. Given these small ratios, the accretion is not affected by the Eddington limit in this case.

Refer to caption
Figure 1: The evolution of the dimensionless accretion rate with redshift. The PBH masses corresponding to the curves from bottom to top are: 100M10^{0}{\rm M}_{\odot} (red solid), 101M10^{1}{\rm M}_{\odot} (green dashed), 102M10^{2}{\rm M}_{\odot} (blue dash-dotted), 103M10^{3}{\rm M}_{\odot} (magenta dotted). Of the two lines corresponding to each mass, the upper (thick) one is the case if PBHs are surrounded by dark matter halos and the lower (thin) one is the case of naked PBHs.

II.1.2 PBHs inside dark matter halos

If PBHs account for only a small fraction of the dark matter, they can attract the surrounding dark matter particles. A PBH with mass MM_{\bullet} will form an extended dark matter halo with mass Ricotti (2007)

Mh=M(30001+z)M_{\rm h}=M_{\bullet}\left(\frac{3000}{1+z}\right) (7)

and radius

rh\displaystyle r_{\rm h} =\displaystyle= 0.339rturn\displaystyle 0.339r_{\rm turn} (8)
=\displaystyle= 0.339581+z(MhM)13pc,\displaystyle 0.339\frac{58}{1+z}\left(\frac{M_{\rm h}}{{\rm M}_{\odot}}\right)^{\frac{1}{3}}\text{pc},

where rturnr_{\rm turn} is the turnaround radius.

We assume the dark matter halo has spherical symmetry. Because there is a black hole at the center of the dark matter halo, the interior of the density profile becomes steeper Adamek et al. (2019) than the predictions from the NFW model Navarro et al. (1996) or the Einasto model Einasto (1965). Therefore, we adopt the power-law density profile derived from numerical simulations Bertschinger (1985) to describe the dark matter distribution. The enclosed dark matter mass is then

Mh(<r)={Mh×(rrh)pr<rhMhr>rh,M_{\rm h}(<r)=\begin{cases}M_{\rm h}\times\left(\frac{r}{r_{\rm h}}\right)^{p}&\text{$r<r_{\rm h}$}\\ M_{\rm h}&\text{$r>r_{\rm h}$},\end{cases} (9)

where p=0.75p=0.75.

Obviously, at the redshifts of interest, the dark matter halo mass is much higher than the mass of PBH, so the velocity dispersion of dark matter is much larger than the escape velocity. Therefore, the accretion of dark matter to PBHs is negligible Ricotti et al. (2008). For a PBH inside a dark matter halo, basically the accretion rate is solved from conservation equations of mass and momentum, and gas thermodynamics. In what follows we give a brief summary of the formulae for calculating such an accretion rate. More details can be found in the original references.

Let the halo Bondi radius to halo radius ratio be χrB,h/rh\chi\equiv r_{\rm B,{\rm h}}/r_{\rm h}, where rB,h=GMh/veff2r_{\rm B,{\rm h}}=GM_{\rm h}/v_{\rm eff}^{2} is the halo Bondi radius. When χ>1\chi>1, we simply replace MM_{\bullet} with Mh+MM_{\rm h}+M_{\bullet} to obtain the new accretion rate. However, when χ<1\chi<1, only a fraction of the halo mass plays the role of sustaining Bondi accretion. In this case one should replace the Bondi radius in Eqs. (1, 5, 6) with the effective radius solved from the following equation:

rB,eff\displaystyle r_{\rm B,eff} =G(M+Mh(<rB,eff))veff2\displaystyle=\frac{G(M_{\bullet}+M_{\rm h}(<r_{\rm B,eff}))}{v_{\rm eff}^{2}}
rB,h(rB,effrh)p.\displaystyle\approx r_{\rm B,h}\left(\frac{r_{\rm B,eff}}{r_{\rm h}}\right)^{p}. (10)

λ\lambda is solved from the hydrodynamic equations Ricotti et al. (2008); Ali-Haïmoud and Kamionkowski (2017); Tashiro et al. (2021):

ρ~x2u\displaystyle-\widetilde{\rho}x^{2}u =\displaystyle= λ\displaystyle\lambda (11)
Mrx2dT~dxT~ρ~dρ~dxβcu\displaystyle-\frac{M_{r}}{x^{2}}-\frac{{\rm d}\widetilde{T}}{{\rm d}x}-\frac{\widetilde{T}}{\widetilde{\rho}}\frac{{\rm d}\widetilde{\rho}}{{\rm d}x}-\beta_{\rm c}u =\displaystyle= ududx\displaystyle u\frac{{\rm d}u}{{\rm d}x} (12)
udT~dx+uT~(23)1ρ~dρ~dx\displaystyle u\frac{{\rm d}\tilde{T}}{{\rm d}x}+u\widetilde{T}(-\frac{2}{3})\frac{1}{\widetilde{\rho}}\frac{{\rm d}\widetilde{\rho}}{{\rm d}x} =\displaystyle= γc(1T~),\displaystyle\gamma_{\rm c}(1-\widetilde{T}), (13)

where x=r/rB,effx=r/r_{\rm B,eff} is the dimensionless radial coordinate; u=vr(x)/veff<0u=v_{r}(x)/v_{\rm eff}<0 is the dimensionless radial velocity of accreting gas; ρ~=ρb(x)/ρb,\widetilde{\rho}=\rho_{b}(x)/\rho_{b,\infty} is the dimensionless gas density; T~=T(x)/TCMB\widetilde{T}=T(x)/T_{\rm CMB} is the dimensionless temperature; and MrM_{r} represents a mass proportion

Mr=M+Mh(<r)M+Mh(<rB,eff).M_{r}=\frac{M_{\bullet}+M_{\rm h}(<r)}{M_{\bullet}+M_{\rm h}(<r_{\rm B,eff})}. (14)

Defining v=uv=-u and replacing Eq. (11) with the derivative form:

dlnv+dlnρ~=2dlnx,d\ln v+d\ln\widetilde{\rho}=-2d\ln x,

Eqs. (11)()-(13) can be transformed to the matrix form:

(v2T~T~02v/3v110)(dlnvdlnρ~dlnT~)\displaystyle\begin{pmatrix}v^{2}&\widetilde{T}&\widetilde{T}\\ 0&2v/3&-v\\ 1&1&0\end{pmatrix}\begin{pmatrix}{\rm d}\ln v\\ {\rm d}\ln\widetilde{\rho}\\ {\rm d}\ln\widetilde{T}\end{pmatrix}
=(Mr/xβcvxγc(1/T~1)x2)dlnx.\displaystyle=-\begin{pmatrix}M_{r}/x-\beta_{\rm c}vx\\ -\gamma_{\rm c}(1/\widetilde{T}-1)x\\ 2\end{pmatrix}{\rm d}\ln x. (15)

The boundary conditions are T~(x)=1\widetilde{T}(x)=1 and v(x)=λ/x2v(x)=\lambda/x^{2} at large xx. There are two further requirements. The infalling velocity uu must increase monotonously towards the center; and there must exist a sonic point, where the infalling velocity is continuous. In such case a unique λ\lambda value can be derived from Eq. (15). Details can be found in Ref. Ali-Haïmoud and Kamionkowski (2017).

When Mr=1M_{r}=1, the hydrodynamic equation (15) reduces to the naked PBH model. We have checked that the solutions are indeed consistent with Eq. (4).

In Fig. 1 the dimensionless accretion rate m˙\dot{m} for PBHs inside dark matter halos is shown by the thick curves. When the PBH is surrounded by a dark matter halo, the larger MM_{\bullet}, the earlier the dark matter halo plays a significant role in enhancing the accretion rate. The accretion rate is boosted by 2\gtrsim 2 orders of magnitude at z17z\sim 17. For PBHs of M102MM_{\bullet}\gtrsim 10^{2}\,{\rm M}_{\odot}, the accretion rate can reach the Eddington limit of the central PBH before z10z\sim 10.

II.1.3 The bolometric luminosity

For a PBH with an accretion rate m˙\dot{m}, the bolometric luminosity can be written as

Lbol=ηm˙LEdd,L_{\rm bol}=\eta\dot{m}L_{\rm Edd}, (16)

where η\eta is the radiative efficiency and it depends on m˙\dot{m} and the geometry of the accretion flow. The geometry also depends on the accretion rate. If m˙1\dot{m}\gtrsim 1, then it forms the standard thin accretion disc111In some literatures m˙\dot{m} is defined as M˙B/(10LEdd/c2)\dot{M}_{\rm B}/(10L_{\rm Edd}/c^{2}), using this definition the critical accretion rate for accretion mode transition is 0.1\sim 0.1.; if m˙1\dot{m}\lesssim 1 then an advection dominated accretion flow (ADAF) forms Yuan and Narayan (2014); Narayan et al. (2022); Hektor et al. (2018). For the ADAF accretion mode, when m˙0.1\dot{m}\lesssim 0.1, the radiative efficiency is low and proportional to the accretion rate; when m˙0.1\dot{m}\gtrsim 0.1, the radiative efficiency approaches the thin disc which has a constant radiative efficiency. Therefore Narayan and McClintock (2008)

η=0.1×{m˙0.1m˙0.11m˙>0.1.\eta=0.1\times\begin{cases}\frac{\dot{m}}{0.1}~{}~{}&\dot{m}\leq 0.1\\ 1~{}~{}&\dot{m}>0.1.\end{cases} (17)

Since LbolLEddL_{\rm bol}\leq L_{\rm Edd}, we set a hard limit on m˙\dot{m} so that ηm˙1\eta\dot{m}\leq 1 always holds.

The typical value of λ\lambda calculated from Eqs. (4) or (15) is 𝒪(0.1)\mathcal{O}(0.1). Note however that when the accretion is non-spherical, λ\lambda can be an order of magnitude smaller Xie and Yuan (2012). If λ\lambda is smaller, we will need a larger PBH mass to produce the same amount of radiation.

II.2 An empirical LRLXL_{\rm R}-L_{\rm X} relation

Astrophysical black holes usually have strong X-ray signal that is either from the hot accretion flow (ADAF), or from the standard thin disc and the hot corona above the disc. If the black hole has a jet, then it also produces strong radio signal by the synchrotron emission of relativistic electrons inside the jet. Jet can also produce strong X-ray signal and it may dominate the X-ray radiation of the whole black hole when the accretion rate is extremely low Yuan and Cui (2005).

In the observed cases, when the accretion rate is high (m˙1\dot{m}\gtrsim 1), the thin disc forms, and only a small fraction of black holes have jet Rafter et al. (2009). However, when the accretion rate is lower (m˙1\dot{m}\lesssim 1), the ADAF forms, and the fraction of black holes with jet increases dramatically Yang et al. (2020). For most of the cases, our PBH has the accretion rate m˙\dot{m} well below 1\sim 1, therefore it is reasonable to assume that a sufficient large fraction of them have jets.

Ref. Merloni et al. (2003) analyzed the selected samples including supermassive black holes and stellar-mass black holes with both X-ray and radio signals. They found a tight correlation between the radio luminosity LRL_{R} (LR=ν5.0Lν5.0L_{R}=\nu_{5.0}L_{\nu_{5.0}}, ν5.0=5.0\nu_{5.0}=5.0 GHz), hard X-ray (2-10 keV) luminosity LXL_{X} and black hole mass MM_{\bullet},

logLR=0.60logLX+0.78logM+7.33,\log L_{R}=0.60\log L_{X}+0.78\log M_{\bullet}+7.33, (18)

where LRL_{R} and LXL_{X} are in the unit of ergs1{\rm erg}\,{\rm s}^{-1}, and MM_{\bullet} is in units of M{\rm M}_{\odot}. They pointed out that for models where the X-ray signal is mainly from the standard thin disc and corona, or mainly from the jet, the predictions do not agree with the observed scaling relation. However, if the radio signal is mainly from the jet and the X-ray signal is mainly from the ADAF accretion, such scaling relation is well interpreted Heinz and Sunyaev (2003).

We assume the PBH has radio from the jet and X-ray from the ADAF accretion. The radio luminosity and hard X-ray luminosity of the PBH also follow the scaling relation Eq. (18). To derive the radio luminosity of the PBH from Eq. (18), we use the following formula to calculate the hard X-ray luminosity

log(LX/LEdd)={2.3logm˙+1.1m˙0.020.25logm˙2.4m˙>0.02.\log(L_{X}/L_{\rm Edd})=\begin{cases}2.3\log\dot{m}+1.1~{}~{}&\dot{m}\leq 0.02\\ 0.25\log\dot{m}-2.4~{}~{}&\dot{m}>0.02.\end{cases} (19)

The first line fits the outputs of theoretical calculations of the ADAF accretion Merloni et al. (2003); the second line is derived from the bolometric correction of observed AGNs with accretion rate m˙0.02\dot{m}\gtrsim 0.02, assuming a radiative efficiency 0.1 Lusso et al. (2012).

We assume the radio spectrum has a power-law form with a constant spectrum index

Lν=LRν5.0(νν5.0)α,L_{\nu}=\frac{L_{R}}{\nu_{5.0}}\left(\frac{\nu}{\nu_{5.0}}\right)^{-\alpha}, (20)

where α=0.75\alpha=0.75 Ghisellini et al. (2017).

In Fig. 2, we show the radio luminosity as a function of redshift for different PBH masses. For naked PBHs, LRL_{R} is less evolved except at the highest redshifts. For PBHs inside dark matter halos, at high redshifts LRL_{R} is close to the naked PBHs. However at lower redshifts it increases rapidly until the accretion rate reaching the Eddington limit. At z20z\sim 20, when M=1103MM_{\bullet}=1-10^{3}\,{\rm M}_{\odot}, LR5×10151×109LEddL_{R}\sim 5\times 10^{-15}-1\times 10^{-9}L_{\rm Edd} for naked PBHs; and LR8×10149×109LEddL_{R}\sim 8\times 10^{-14}-9\times 10^{-9}L_{\rm Edd} for PBHs inside dark matter halos.

Refer to caption
Figure 2: The evolution of the radio luminosity at 5.0 GHz with redshift, for different PBH masses. Thin lines refer to the naked PBHs, while thick lines refer to PBHs surrounded by dark matter halos.

II.3 The cosmic radio background

To calculate the contribution of PBHs to the CRB, we need to know their mass distribution and number density. In this paper, for simplicity we assume that all PBHs have the same mass MM_{\bullet}, then the comoving number density of PBHs is

n=fPBHΩmρcrM,n_{\bullet}=\frac{f_{\rm PBH}\Omega_{\rm m}\rho_{\rm cr}}{M_{\bullet}}, (21)

where fPBHf_{\rm PBH} is the ratio of total PBH mass density to total matter density in the Universe, and Ωm\Omega_{\rm m} is the fractional matter density.

The CRB from PBHs is then

JCRB(ν,z)\displaystyle J_{\rm CRB}(\nu,z) =(1+z)34πzϵ(ν,z)cH(z)(1+z)dz\displaystyle=\frac{(1+z)^{3}}{4\pi}\int_{z}^{\infty}\epsilon(\nu^{\prime},z^{\prime})\frac{c}{H(z^{\prime})(1+z^{\prime})}{\rm d}z^{\prime} (22)
=(1+z)34πzfDCnLν(M)cH(z)(1+z)dz,\displaystyle=\frac{(1+z)^{3}}{4\pi}\int_{z}^{\infty}\frac{f_{\rm DC}n_{\bullet}L_{\nu^{\prime}}(M_{\bullet})c}{H(z^{\prime})(1+z^{\prime})}{\rm d}z^{\prime},

where

ν=ν(1+z1+z),\nu^{\prime}=\nu\left(\frac{1+z^{\prime}}{1+z}\right), (23)

is the frequency in the emitter’s rest frame. fDC1f_{\rm DC}\leq 1 is the duty-cycle for radio emission, which refers to the fraction of PBHs that is active in the radio band. Since the radio emission for these PBHs is produced by the jet, it would be equal to the ratio of the jet lifetime compared to the Hubble timescale.

In Fig. 3 we show the contribution to the 1.4 GHz CRB at z17z\sim 17 from PBHs as a function of fPBHf_{\rm PBH}, for naked PBHs and for PBHs inside dark matter halos respectively. However, even if we set fDC=1f_{\rm DC}=1, this contribution is still negligible compared with the CMB. In principle, if MM_{\bullet} is higher and/or fPBHf_{\rm PBH} is higher, one may obtain higher TCRBT_{\rm CRB}, but the X-ray heating (see Sec. II.4) can reduce the accretion rate, so TCRBT_{\rm CRB} increases more gently for higher MM_{\bullet} at high fPBHf_{\rm PBH} region. More importantly, the X-ray heating works more efficiently in reducing the 21-cm absorption than the extra CRB in boosting the 21-cm absorption. We have checked that using a higher PBH mass and a higher abundance than in Fig. 3 will not help boosting the 21-cm absorption signal.

Refer to caption
Refer to caption
Figure 3: The contribution to the CRB at z=17z=17 for various PBH masses, as a function of fPBHf_{\rm PBH}. Left: Naked PBHs; Right: PBHs inside dark matter halos. Since the X-ray heating can reduce the accretion rate, TCRBT_{\rm CRB} is not necessarily proportional to fPBHf_{\rm PBH}. In the left panel, the thick (thin) lines refer to the model without (with) PBH binaries.
Refer to caption
Refer to caption
Figure 4: The kinetic temperature of the IGM at z=17z=17 heated by PBHs with various masses, as a function of fPBHf_{\rm PBH}. Left: Naked PBHs; Right: PBHs inside dark matter halos. In the left panel, the thick (thin) lines refer to the model without (with) PBH binaries.

II.4 The X-ray heating and ionization of the IGM

PBHs emit not only radio photons but also X-ray photons. They can heat and ionize the IGM, which would reduce the 21-cm absorption signal.

The physical luminosity density of X-ray emission that interacts with the IGM is

lX=n(1+z)3EminLX(E)[1eτ(E)]dE,l_{X}=n_{\bullet}(1+z)^{3}\int_{E_{\rm min}}^{\infty}L_{X}(E)[1-e^{-\tau(E)}]{\rm d}E, (24)

where the optical depth τ(ν)=[σH(E)nH(1xHII)+σHe(E)nHe(1xHeII)](c/H)\tau(\nu)=[\sigma_{\rm H}(E)n_{\rm H}(1-x_{\rm HII})+\sigma_{\rm He}(E)n_{\rm He}(1-x_{\rm HeII})](c/H), and σH\sigma_{\rm H} and σHe\sigma_{\rm He} are photon-ionization cross-sections for hydrogen and helium respectively Maselli et al. (2003). Here we assume that the X-ray spectrum follows the form

LX(E)=L0(EEmin)0.5exp(E/E1)L_{X}(E)=L_{0}\left(\frac{E}{E_{\rm min}}\right)^{-0.5}\exp(-E/E_{1}) (25)

and adopt Emin=2E_{\rm min}=2 keV and E1=100E_{1}=100 keV, with the normalization

LX=2keV10keVLX(E)dE.L_{X}=\int_{\rm 2\,keV}^{\rm 10\,keV}L_{X}(E){\rm d}E. (26)

Ignoring the contribution from UV photons, the time evolution of hydrogen ionized fraction xHIIx_{\rm HII}, helium ionized fraction xHeIIx_{\rm HeII}, and temperature of the IGM TkT_{\rm k} follows

dxHIIdt\displaystyle\frac{{\rm d}x_{\rm HII}}{{\rm d}t} =\displaystyle= IH+χion,HlXnHEH\displaystyle I_{\rm H}+\chi_{\rm ion,H}\frac{l_{X}}{n_{\rm H}E_{\rm H}}
dxHeIIdt\displaystyle\frac{{\rm d}x_{\rm HeII}}{{\rm d}t} =\displaystyle= IHe+χion,HelXnHeEHe\displaystyle I_{\rm He}+\chi_{\rm ion,He}\frac{l_{X}}{n_{\rm He}E_{\rm He}}
dTkdt\displaystyle\frac{{\rm d}T_{\rm k}}{{\rm d}t} =\displaystyle= HT+23kBnχheatlX,\displaystyle H_{T}+\frac{2}{3k_{\rm B}n}\chi_{\rm heat}l_{X}, (27)

where nH=(1YHe)Ωbρcr(1+z)3/mHn_{\rm H}=(1-Y_{\rm He})\Omega_{\rm b}\rho_{\rm cr}(1+z)^{3}/m_{\rm H} and nHe=YHeΩbρcr(1+z)3/mHen_{\rm He}=Y_{\rm He}\Omega_{\rm b}\rho_{\rm cr}(1+z)^{3}/m_{\rm He} are physical number density of hydrogen and helium elements respectively, and YHe=0.24Y_{\rm He}=0.24; xe=xHII+nHe/nHxHeIIx_{\rm e}=x_{\rm HII}+n_{\rm He}/n_{\rm H}x_{\rm HeII}; the total number density of particles n=nH(1+xe)+nHen=n_{\rm H}(1+x_{\rm e})+n_{\rm He}; EH=13.6E_{\rm H}=13.6 eV and EHe=24.6E_{\rm He}=24.6 eV are ionization energy for hydrogen and helium; kBk_{\rm B} is the Boltzmann constant; χion,H\chi_{\rm ion,H}, χion,He\chi_{\rm ion,He} and χheat\chi_{\rm heat} are fractions of X-ray energy deposited into hydrogen ionization, helium ionization and heating Valdés and Ferrara (2008); the net ionization (IHI_{\rm H} and IHeI_{\rm He}) and heating (HTH_{T}) rates in the absence of PBH are:

IH\displaystyle I_{\rm H} =\displaystyle= [γH(Tk)(1xHII)αH(Tk)xHII]xenH\displaystyle[\gamma_{\rm H}(T_{\rm k})(1-x_{\rm HII})-\alpha_{\rm H}(T_{\rm k})x_{\rm HII}]x_{\rm e}n_{\rm H}
IHe\displaystyle I_{\rm He} =\displaystyle= [γHe(Tk)(1xHeII)αHe(Tk)xHeII]xenH\displaystyle[\gamma_{\rm He}(T_{\rm k})(1-x_{\rm HeII})-\alpha_{\rm He}(T_{\rm k})x_{\rm HeII}]x_{\rm e}n_{\rm H}
HT\displaystyle H_{T} =\displaystyle= 2H(z)Tk1n[nHdxHIIdt+nHedxHeIIdt]Tk\displaystyle-2H(z)T_{\rm k}-\frac{1}{n}\left[n_{\rm H}\frac{{\rm d}x_{\rm HII}}{{\rm d}t}+n_{\rm He}\frac{{\rm d}x_{\rm HeII}}{{\rm d}t}\right]T_{\rm k}
23kBnΛ(Tk),\displaystyle-\frac{2}{3k_{\rm B}n}\Lambda(T_{\rm k}),

where γH,He\gamma_{\rm H,He} and αH,He\alpha_{\rm H,He} are rates of collision ionization and recombination, for hydrogen and helium respectively Maselli et al. (2003); Λ(T)\Lambda(T) are the physical volume cooling rate (including the collisional ionization cooling, collisional excitation, Bremsstrahlung cooling and Compton heating/cooling) Maselli et al. (2003). Since m˙\dot{m} depends on TkT_{\rm k} and xex_{\rm e}, we solve the thermal evolution of the IGM and m˙\dot{m} self-consistently.

In Fig. 4, we show the kinetic temperature of the IGM at z17z\sim 17 as a function of fPBHf_{\rm PBH}, for various PBH masses same as in Fig. 3, for naked PBHs and PBHs inside dark matter halos respectively. In both cases, it is easier to heat the IGM temperature above the CMB than to generate the CRB larger than the CMB.

Refer to caption
Refer to caption
Figure 5: The 21-cm brightness temperature at z17z\sim 17 under the influence of PBHs with different masses, as a function of PBH abundance. Left: Naked PBHs; Right: PBHs surrounded by dark matter halos. The black solid line represents the maximum 21-cm absorption, i.e. 229-229 mK, at z17z\sim 17 in the standard cosmology. However, please note that the inhomogeneity of the IGM and shock heating effect can reduce this maximum absorption by 15%\sim 15\% Xu et al. (2018, 2021). In the left panel, the thick (thin) lines refer to the model without (with) PBH binaries.

II.5 The global spectrum of 21-cm signal

The brightness temperature of 21-cm global spectrum is Pritchard and Loeb (2012):

δTb\displaystyle\delta T_{\rm b} =\displaystyle= 27xHI(Ωbh20.023)(0.15Ωmh21+z10)1/2\displaystyle 27x_{\rm HI}\left(\frac{\Omega_{\rm b}h^{2}}{0.023}\right)\left(\frac{0.15}{\Omega_{\rm m}h^{2}}\frac{1+z}{10}\right)^{1/2} (29)
×\displaystyle\times (TSTRTS)[mK],\displaystyle\left(\frac{T_{\rm S}-T_{\rm R}}{T_{\rm S}}\right)~{}[\rm mK],

where the temperature of the radiation field TR=TCMB+TCRBT_{\rm R}=T_{\rm CMB}+T_{\rm CRB}. The spin temperature is

TS1=TR1+xαTk1+xcTk11+xα+xc,T_{\rm S}^{-1}=\frac{T_{\rm R}^{-1}+x_{\alpha}T_{\rm k}^{-1}+x_{\rm c}T_{\rm k}^{-1}}{1+x_{\alpha}+x_{\rm c}}, (30)

where xαx_{\alpha} and xcx_{\rm c} are coupling coefficients due to Lyα\alpha scattering and collision respectively. Calculations of them can be found in Ref. Mesinger et al. (2011) and references therein.

The PBH does not only generate radio background and X-ray which heats up the gas, but also contribute to the Lyα\alpha background. A fraction χα\chi_{\alpha}Valdés and Ferrara (2008) of the X-ray energy is deposited into Lyα\alpha photons, so the specific intensity of photon number is given by

Jα(z)=1hPνα14πχαlXναcH(z).J_{\alpha}(z)=\frac{1}{h_{\rm P}\nu_{\alpha}}\frac{1}{4\pi}\frac{\chi_{\alpha}l_{X}}{\nu_{\alpha}}\frac{c}{H(z)}. (31)

In Fig. 5, we show the 21-cm brightness temperature at z=17z=17 as a function of fPBHf_{\rm PBH}, for various PBH masses. To highlight the role played by the PBH, we ignore the radiation from any normal astrophysical objects, e.g. Pop III stars, first galaxies. It is interesting to note that, although the CRB from the PBH is unimportant in this scenario, for the 21-cm absorption there is an optimal fPBHf_{\rm PBH} where the absorption reaches the maximum. If fPBHf_{\rm PBH} is lower, it produces less Lyα\alpha background, which is not strong enough to couple TST_{\rm S} with TkT_{\rm k}, therefore the 21-cm absorption is weaker. If fPBHf_{\rm PBH} is higher however, the IGM is heated significantly, which still results in weaker 21-cm absorption. Even in the absence normal astrophysical objects, the PBHs can provide the necessary Lyα\alpha photons to make the 21-cm signal be close to the maximum absorption in the standard cosmology. This is not easy if only the Pop III stars are the first ionization sources Yajima and Khochfar (2015); Furlanetto (2006); Schauer et al. (2019).

II.6 The PBH binaries

In addition to the isolated PBHs, there are also PBH binaries. Ref. De Luca et al. (2020a) found that: for two PBHs with equal mass, if they form a binary system, their total accretion rate is the Bondi accretion rate of a single black hole with mass equal to the sum of their mass. Since Bondi accretion rate is generally proportional to M2\sim M^{2}_{\bullet}, if the indirect dependence of λ\lambda on MM_{\bullet} is ignored, the accretion rate of each PBH is then boosted by a factor \sim2, because in the binary system, the ambient medium density of each PBH is boosted. The boost factor of radiation (X-ray, radio) is even larger than 2 since the bolometric luminosity is proportional to m˙2\dot{m}^{2}. Therefore, the PBH binaries can help to increase the influence on the 21 cm signal. Ref. De Luca et al. (2020a) also pointed out that, the binary system will finally reach a stationary state where the two members have equal mass, even if at the beginning their mass is different, because the accretion evolution tends to balance each other.

On the other hand, the fraction of PBHs that are in binary systems is Raidal et al. (2017)

fbinary1eδdcfPBH,f_{\rm binary}\sim 1-e^{-\delta_{\rm dc}f_{\rm PBH}}, (32)

where δdc\delta_{\rm dc} describes the clustering of PBHs. For uniform distribution, δdc=1\delta_{\rm dc}=1. In such case, fbinaryf_{\rm binary} is small when fPBH0.1f_{\rm PBH}\lesssim 0.1, and the contribution to X-ray heating and CRB from PBH binaries is small. However, if for some reason the PBHs can be highly clustered, say δdc10\delta_{\rm dc}\gtrsim 10, then the PBH binaries may have a significant influence on the 21 cm signal, depending on fPBHf_{\rm PBH}. In the left panels of Fig. 3, 4 & 5 we show the TCRBT_{\rm CRB}, TkT_{\rm k} and δTb\delta T_{\rm b} of 21 cm signal if the contribution from PBH binaries is also taken into account by thin lines, assuming δdc=1000\delta_{\rm dc}=1000. We see that when fPBH104103f_{\rm PBH}\gtrsim 10^{-4}-10^{-3}, the influence of PBH binaries becomes more obvious. The gravitational wave observations put constraints on the joint parameter δdc16/37fPBH53/37\delta_{\rm dc}^{16/37}f_{\rm PBH}^{53/37}. For δdc=1000\delta_{\rm dc}=1000, fPBH104103f_{\rm PBH}\sim 10^{-4}-10^{-3} is still within the limit set by LIGO observations Raidal et al. (2017).

If the PBH binaries are located inside dark matter halos, absolutely they will also boost the influence on the 21 cm absorption signal. We do not show the results in the right panels because we suppose the accretion rate, in this case, can be much more complicated than the binaries of naked PBHs.

III The CRB and 21-cm signal from the jet model

So far we have calculated the radio and X-ray radiation from PBHs with the empirical relation Eq. (18). However, it is also possible that the PBHs do not necessarily follow this scaling relation, since they were born in a different environment. Moreover, it is pointed out that in Ref. Merloni et al. (2003) different sample objects may have different X-ray origins. For the low accretion rate objects, both the radio and X-ray signals can arise predominantly from the jet Mezcua et al. (2018). In this section, we tentatively calculate the radio emission using a jet model that can produce a stronger radio signal, then re-evaluate the CRB and 21-cm signal at Cosmic Dawn under reasonable assumptions about the model parameters.

III.1 The jet model

When accretion is weak, the radiative efficiency of the accretion is low, but the relativistic electrons in the jet can produce strong synchrotron radiation Yuan and Cui (2005); Paliya et al. (2017). In this case we can derive the radio luminosity from a jet model where all jet emission is considered as coming from a blob of electrons with uniform properties (Ghisellini et al., 2010).

The total synchrotron power of the blob is (Ghisellini, 2013)

Psyn=γminγmax43σTcγ2UBdNe(γ)dγdγ,P_{\rm syn}=\int_{\gamma_{\rm min}}^{\gamma_{\rm max}}\frac{4}{3}\sigma_{\rm T}c\gamma^{2}U_{B}\frac{dN_{\rm e}(\gamma)}{{\rm d}\gamma}{\rm d}\gamma, (33)

where dNe/dγ{\rm d}N_{\rm e}/{\rm d}\gamma is the energy distribution of relativistic electrons in the blob; UB=B2/(8π)U_{B}=B^{2}/(8\pi) is the energy density of magnetic field with strength BB; γmin\gamma_{\rm min} and γmax\gamma_{\rm max} are the minimum and maximum Lorentz factors respectively.

On the other hand, since the jet blob is far from the central black hole, in redshift ranges of our interests, the seed photons for inverse Compton scattering (ICS) are mainly the synchrotron photons (the synchrotron self-Compton, SSC) and the CMB photons. The total ICS power is

PICS=γminγmax43σTcγ2URdNe(γ)dγdγ,\displaystyle P_{\rm ICS}=\int_{\gamma_{\rm min}}^{\gamma_{\rm max}}\frac{4}{3}\sigma_{\rm T}c\gamma^{2}U_{\rm R}\frac{{\rm d}N_{\rm e}(\gamma)}{{\rm d}\gamma}{\rm d}\gamma, (34)

where UR=ργ+UsynU_{\rm R}=\rho_{\gamma}+U_{\rm syn} is the energy density of radiation field, consisting of CMB energy density (ργ\rho_{\gamma}) and synchrotron radiation field. The latter is

UsynPsynRblob/c4πRblob3/3,\displaystyle U_{\rm syn}\sim\frac{P_{\rm syn}R_{\rm blob}/c}{4\pi R_{\rm blob}^{3}/3}, (35)

where RblobR_{\rm blob} is the physical size of the blob.

Refer to caption
Figure 6: The different components of the SED of a PBH with M=10MM_{\bullet}=10\,{\rm M}_{\odot} at z=17z=17.

It is generally assumed that the injected relativistic electrons follows a power-law distribution with an exponential cutoff (e.g. Pepe et al. (2015a)),

Q(γ)=Q0γsexp(γγmax),Q(\gamma)=Q_{0}\gamma^{-s}\exp\left(-\frac{\gamma}{\gamma_{\rm max}}\right), (36)

where the normalization Q0Q_{0} is set by the relation

Pe=mec2γQ(γ)dγ.\displaystyle P_{\rm e}=\int m_{\rm e}c^{2}\gamma Q(\gamma){\rm d}\gamma. (37)

Pe=ηeηjetm˙LEddP_{\rm e}=\eta_{\rm e}\eta_{\rm jet}\dot{m}L_{\rm Edd} is the injected power of the relativistic electrons, where ηjet\eta_{\rm jet} is the jet efficiency and ηe\eta_{\rm e} is the fraction of jet power that is finally converted into the relativistic electrons.

With the energy loss due to synchrotron and ICS, the steady state electron distribution is Ghisellini and Tavecchio (2009); Gardner and Done (2014, 2018)

dNe(γ)dγ={KQ(γ)γ<γcool1γ˙γγmaxQ(γ)dγγ>γcool,,\displaystyle\frac{{\rm d}N_{\rm e}(\gamma)}{{\rm d}\gamma}=\begin{cases}KQ(\gamma)&\gamma<\gamma_{\rm cool}\\ -\frac{1}{\dot{\gamma}}\int_{\gamma}^{\gamma_{\rm max}}Q(\gamma^{\prime}){\rm d}\gamma^{\prime}&\gamma>\gamma_{\rm cool},\end{cases}, (38)

where

γ˙=1mec243σTcγ2(UR+UB).\displaystyle\dot{\gamma}=-\frac{1}{m_{\rm e}c^{2}}\frac{4}{3}\sigma_{\rm T}c\gamma^{2}(U_{\rm R}+U_{B}). (39)

γcool\gamma_{\rm cool} is the point where the cooling time equals the crossing timescale of the blob, with the cooling time given by

tcool=γmec243σTcγ2(UR+UB)t_{\rm cool}=\frac{\gamma m_{\rm e}c^{2}}{\frac{4}{3}\sigma_{\rm T}c\gamma^{2}(U_{\rm R}+U_{B})} (40)

and tcross=Rblob/ct_{\rm cross}=R_{\rm blob}/c. KK factor in Eq. (38) is

K=1γ˙Q(γ)γγmaxQ(γ)dγ|γ=γcool.\displaystyle K=-\frac{1}{\dot{\gamma}Q(\gamma)}\int_{\gamma}^{\gamma_{\rm max}}Q(\gamma^{\prime}){\rm d}\gamma^{\prime}\Big{|}_{\gamma=\gamma_{\rm cool}}. (41)
Refer to caption
Refer to caption
Refer to caption
Figure 7: The CRB temperature from PBHs (left), the IGM temperature (middle); and the 21-cm signal (right), at z=17z=17 as a function of PBH abundance. We show the results for three different fDCf_{\rm DC}. We assume all PBHs have the same mass of 10M10\,{\rm M}_{\odot}, and fiducial jet parameters are adopted. For comparison, in the CRB panel we plot the CMB temperature; in the kinetic temperature panel we plot the IGM temperature in the absence of extra heating; in the 21-cm signal panel we plot the maximal absorption signal in the absence of extra CRB and heating. All quantities are calculated for naked PBHs.

In practice, γcool\gamma_{\rm cool} is found by numerical iterations. Finally the synchrotron luminosity is

Lsyn(ν)=γminγmaxlsyn(ν,γ,B)dNe(γ)dγdγ,L_{\rm syn}(\nu)=\int_{\gamma_{\rm min}}^{\gamma_{\rm max}}l_{\rm syn}(\nu,\gamma,B)\frac{{\rm d}N_{\rm e}(\gamma)}{{\rm d}\gamma}{\rm d}\gamma, (42)

where lsynl_{\rm syn} is the synchrotron emission of a single electron Ghisellini (2013). We then obtain the SSC luminosity from Lsyn(ν)L_{\rm syn}(\nu) and dNe(γ)/dγ{\rm d}N_{\rm e}(\gamma)/{\rm d}\gamma, and the luminosity of the ICS between electrons and CMB photons Lefa (2012) from dNe(γ)/dγ{\rm d}N_{\rm e}(\gamma)/{\rm d}\gamma.

So the jet model has following parameters {Rbolb,B,ηjet,γmin,γmax,s}\{R_{\rm bolb},B,\eta_{\rm jet},\gamma_{\rm min},\gamma_{\rm max},s\} that are directly relevant to the radio signal. We assume that the PBH jet is analogous to the jet of typical stellar-mass black hole Cygnus X-1 Pepe et al. (2015a); Gandhi et al. (2017), in spite of its lower accretion rate. We set the jet parameters according to the following considerations:

  • For Cygnus X-1, according to Ref. Zdziarski (2012); Heinz (2006), the bulk of the low frequency radio emission is generated at the height 𝒪(1)×107RS\mathcal{O}(1)\times 10^{7}~{}R_{\rm S} along the jet axis, where RS=2GM/c2R_{\rm S}=2GM_{\bullet}/c^{2} is the Schwarzschild radius. We assume half of the opening angle of the blob is 0.1 radian, therefore take Rblob=106RSR_{\rm blob}=10^{6}~{}R_{\rm S}.

  • According to Ref. Pepe et al. (2015b), at the jet base which is 1.1×1081.1\times 10^{8} cm above the back hole, B=5×107B=5\times 10^{7} G. BB can be inversely proportional to the height along jet axis, therefore at the end of acceleration region which is 𝒪(1)×1012\sim\mathcal{O}(1)\times 10^{12} cm above the black hole, BB is serval thousand G. According to Ref. Zdziarski et al. (2014), at the onset of the energy dissipation B104106B\sim 10^{4}-10^{6} G. If most of the radio emission is produced at the height 10310^{3} times the onset, and BB is again inversely proportional to the height, then BB can be several tens to several thousand. We take B=2000B=2000 G as the fiducial value. We will also show the results for much lower BB.

  • The observed total jet power can be comparable or even larger than the total accretion power m˙LEdd\dot{m}L_{\rm Edd} Ghisellini et al. (2014), because the jet power is not directly from the accretion Ghisellini (2019), but supplied by the black hole spin. This is supported by the GRMHD simulations which found that the steady-state jet efficiency ηjet\eta_{\rm jet} can be \sim 140% Tchekhovskoy et al. (2011) for the ADAF-like accretion mode. The maximum jet efficiency can reach 300\sim 300% McKinney et al. (2012). Here we assume that a large fraction of the jet power is finally converted into the kinetic energy of the relativistic electrons, and take ηeηjet=0.5\eta_{e}\eta_{\rm jet}=0.5 as the optimistic fiducial value. We will also show the results for a much smaller ηeηjet\eta_{e}\eta_{\rm jet}.

  • Regarding the parameters of energy distribution of relativistic electrons, we take γmin=1\gamma_{\rm min}=1 and γmax106\gamma_{\rm max}\sim 10^{6} as in Ref. Zhang et al. (2014). However we take a steeper slope s=3.2s=3.2 as fiducial value, so that the low frequency signal is larger in the jet radiation. In some blazars the distribution of relativistic electrons with γ100\gamma\gtrsim 100 can have such steep slope Ghisellini et al. (2010). We will also show the results for a shallower slope.

In Fig. 6 we plot the different components of the SED for a 10M10\,{\rm M}_{\odot} PBH with such fiducial parameters at z=17z=17. We see that the SSC and ICS luminosities are much smaller than the synchrotron. The heating and ionization of the IGM will be dominated by accretion X-ray radiation, if the UV radiation is neglected.

III.2 The CRB and 21-cm signal at Cosmic Dawn

We solve the IGM thermal history and m˙\dot{m} evolution self-consistently, then compute the CRB and the 21-cm signal. The contributions from accretion X-ray radiation, synchrotron, SSC and the ICS are all taken into account.

For naked PBHs with M=10MM_{\bullet}=10\,{\rm M}_{\odot}, we plot TCRBT_{\rm CRB}, TkT_{\rm k} and δTb\delta T_{\rm b} at z=17z=17 as a function of fPBHf_{\rm PBH} for various fDCf_{\rm DC} in Fig. 7. For fDC0.01f_{\rm DC}\lesssim 0.01, the X-ray heating would be the dominant effect, the 21-cm signal is hardly enhanced.

In contrast, for fDC0.01f_{\rm DC}\gtrsim 0.01, when fPBHfDC105f_{\rm PBH}f_{\rm DC}\gtrsim 10^{-5}, they can significantly contribute to the CRB and boost the 21-cm absorption signal. Although the IGM is also heated, TkT_{\rm k} is still smaller than TCMB+TCRBT_{\rm CMB}+T_{\rm CRB}, so the 21-cm signal is still in absorption. For more massive PBHs and a higher abundance, the heating is more efficient, TkT_{\rm k} can be larger than TCMB+TCRBT_{\rm CMB}+T_{\rm CRB}, so it turns 21-cm signal to an emission. However, the resultant high ionization fraction from efficient heating makes the optical depth too large to be compatible with current Planck constraints. To see this, we calculate the augumented optical depth of the Thomson scattering of CMB photons as

Δτe(>z)=zσTnH(z)Δxe(z)cdzH(z)(1+z),\Delta\tau_{\rm e}(>z)=\int^{\infty}_{z}\sigma_{\rm T}n_{\rm H}(z^{\prime})\Delta x_{\rm e}(z^{\prime})\frac{c{\rm d}z^{\prime}}{H(z^{\prime})(1+z^{\prime})}, (43)

where Δxe\Delta x_{\rm e} is the fraction of extra IGM electrons generated by X-rays from PBHs. To avoid distorting the CMB signal, we require that Δτe(z>10)\Delta\tau_{\rm e}(z>10) is smaller than the 2σ2\sigma uncertainty level of the Planck observations, i.e. Δτe0.014\Delta\tau_{\rm e}\lesssim 0.014 Planck Collaboration et al. (2020).

Refer to caption
Figure 8: The contour map of δTb\delta T_{\rm b} as a function of PBH mass and abundance, for fDC=1f_{\rm DC}=1. We mark the region where extra CMB optical depth from PBHs at z>10z>10 is larger than the Planck observational 2σ2\sigma upper limit (above the solid curve), and the region where the 21-cm absorption can be boosted by the PBHs (above the dashed line).

In Fig. 8 we plot the contour map of the 21-cm signal with the PBH mass and abundance when fDC=1f_{\rm DC}=1. We see that, even in the presence of X-ray heating, there is still a large parameter space for which the 21-cm absorption can be boosted, For example, if M102MM_{\bullet}\gtrsim 10^{2}\,{\rm M}_{\odot}, the 21-cm signal is obviously boosted even fPBHf_{\rm PBH} is as low as 106\sim 10^{-6}. Even after excluding the parameters for which Δτe0.014\Delta\tau_{\rm e}\geq 0.014 (above the solid curve), it is still possible to boost the 21-cm absorption signal deeper than 500\sim-500 mK. In Fig. 8 the δTb=500\delta T_{b}=-500 mK curve corresponds to log(fPBH)=1.8log(M/M)3.5\log(f_{\rm PBH})=-1.8\log(M_{\bullet}/{\rm M}_{\odot})-3.5 for 1MM300M1\,{\rm M}_{\odot}\lesssim M_{\bullet}\lesssim 300\,{\rm M}_{\odot}. For a PBH distribution with a single mass, if M1MM_{\bullet}\lesssim 1\,{\rm M}_{\odot} and fPBH104f_{\rm PBH}\lesssim 10^{-4}, it will not significantly affect the 21-cm signal.

We also investigate the alternative duty-cycle values, and find that if fDC0.04f_{\rm DC}\lesssim 0.04, the boosted 21-cm absorption signal would be 500\leq 500 mK if we are to keep Δτe0.014\Delta\tau_{\rm e}\lesssim 0.014.

In Fig. 9 we show the ratio of the 21 cm signal with PBH binaries to that without PBH binaries, for the fiducial jet model parameters. When fPBH104f_{\rm PBH}\lesssim 10^{-4}, the PBH binaries have negligible influence on the 21 cm signal. When fPBH104f_{\rm PBH}\gtrsim 10^{-4}, the PBH binaries boost the 21 cm absorption signal for M10MM_{\bullet}\lesssim 10~{}M_{\odot}, while suppress it for M10MM_{\bullet}\gtrsim 10~{}M_{\odot}. This is because when the PBHs are heavier, the X-ray heating can somewhat reduce the absorption depth and/or the accretion rate, by increasing TkT_{k}, although such parameter region should be ruled out when the reionization constraints are taken into account.

Refer to caption
Figure 9: The ratio between the 21 cm signal with and without PBH binaries, jet model parameters are the same as in Fig. 8.

In Fig. 10 we show the cases for which either ηeηjet\eta_{e}\eta_{\rm jet} is lower (top), or BB is lower (middle), or the energy distribution of injected relativistic electron is harder (bottom), than our fiducial values respectively. In these cases, we find that if ηeηjet0.02\eta_{e}\eta_{\rm jet}\lesssim 0.02, or B200B\lesssim 200 G, or s2.5s\lesssim 2.5, then it would be unable to boost the 21-cm absorption signal to 500\sim-500 mK, while keeping the reionization history almost unchanged at the same time. For the electron distribution with s=2.5s=2.5, the heating by X-rays from the synchrotron is even more important than that from the accretion radiation.

Refer to caption
Refer to caption
Refer to caption
Figure 10: Same to Fig. 8, but for ηeηjet=0.02\eta_{e}\eta_{\rm jet}=0.02 (top), or B=200GB=200\,{\rm G} (middle), or s=2.5s=2.5 (bottom). For each panel, except the specified parameter, other parameters are set to fiducial values.

If PBHs are inside dark matter halos, where the gravity of the dark matter halo is still negligible, the accretion rate decreases as the Univese expands, like the naked PBHs. However, the rate can increase at a critical redshift because the gravity of the dark matter halo dominates the accretion. This critical redshift depends on the PBH mass, as seen from Fig. 1. The 21-cm signal is then more complicated in this case.

In such a model, at very high redshift, when the effects of dark matter halo is still negligible, the 21-cm absorption decreases with decreasing redshift, just as the case for naked PBHs. However, the dark matter halo enhances the PBH gas accretion, so at some point the 21-cm absorption signal starts to increases as redshift decreases, because of the enhancement of CRB. The absorption reaches a maximum at a redshift ztroughz_{\rm trough}, then starts to decrease because of the X-ray heating and the expansion of the Universe. Therefore in the presence of dark matter halos, an absorption trough is produced. The depth and redshift of the trough depend on the PBH mass and abundance.

In Fig. 11 we plot the δTb\delta T_{b} at ztroughz_{\rm trough} as a function of fPBHf_{\rm PBH} and MM_{\bullet}, assuming fDC=1f_{\rm DC}=1. The δTb=500\delta T_{b}=-500 mK curve corresponds to log(fPBH)=max[1.7log(M/M)3.7,0.2log(M/M)5.0]\log(f_{\rm PBH})={\rm max}[-1.7\log(M_{\bullet}/{\rm M}_{\odot})-3.7,-0.2\log(M_{\bullet}/{\rm M}_{\odot})-5.0] for 0.3MM30M0.3\,{\rm M}_{\odot}\lesssim M_{\bullet}\lesssim 30\,{\rm M}_{\odot}. Since ztroughz_{\rm trough} is also a function of fPBHf_{\rm PBH} and MM_{\bullet}, the cross point of given δTb\delta T_{\rm b} and given ztroughz_{\rm trough} curves uniquely determines the mass and abundance of PBHs.

For example, to produce the maximal 21-cm absorption of δTb=500\delta T_{\rm b}=-500 mK at ztrough=17z_{\rm trough}=17, it requires M1.05MM_{\bullet}\sim 1.05\,{\rm M}_{\odot} and fPBH1.5×104f_{\rm PBH}\sim 1.5\times 10^{-4}. In Fig. 12 we show δTb\delta T_{\rm b} as a function of redshift for this PBH mass and abundance. For comparison we also plot the curves from some other PBH masses and abundances.

Similar to the case of naked PBHs, we find that if fDC0.04f_{\rm DC}\lesssim 0.04 then it would be hard to boost the 21-cm signal significantly.

A dark matter halo does not always boost the accretion rate of its embedded PBH. If a PBH falls into the gravitational well of a previously-formed dark matter halo, then the gravitational potential energy is converted into kinetic energy. In this case, the PBH and ambient medium can have a higher relative velocity than vLv_{L}. As a result, the accretion is suppressed and the dark matter halo is not helpful for producing the 21 cm absorption trough. Such a scenario is not involved in our model.

Refer to caption
Figure 11: The 21-cm brightness temperature at the absorption trough as a function of MM_{\bullet} and fPBHf_{\rm PBH}, for PBHs inside dark matter halos. We mark the location of δTb(ztrough)=500\delta T_{\rm b}(z_{\rm trough})=-500 mK, and the location of the trough redshift ztrough=17z_{\rm trough}=17. The intersection is M1.05MM_{\bullet}\sim 1.05\,{\rm M}_{\odot} and fPBH1.5×104f_{\rm PBH}\sim 1.5\times 10^{-4}.
Refer to caption
Figure 12: The 21-cm signal as a function of redshift (or ν\nu) for PBHs inside dark matter halos.

III.3 The Contribution to Present-day Cosmic Electromagnetic Background

The redshifted radiation from PBHs at >z>z forms part of the Cosmic Electromagnetic Background (CEB) in present-day Universe. The intensity can be written as

ICEB(ν,>z)=14πznL(ν)cdzH(z)(1+z),I_{\rm CEB}(\nu,>z)=\frac{1}{4\pi}\int^{\infty}_{z}n_{\bullet}L_{\bullet}(\nu)\frac{c{\rm d}z^{\prime}}{H(z^{\prime})(1+z^{\prime})}, (44)

where LL_{\bullet} is the total PBH luminosity, including the contribution from synchrotron, SSC, ICS of the CMB photons, and the accretion X-ray radiation. For the model to be consistent with observations, we require that the produced background radiation should not exceed what have been observed at different frequency bands.

In Fig. 13 we plot the CEB from radio to gamma-ray, contributed by PBHs inside dark matter halos at z>10z>10, for M=1.05MM_{\bullet}=1.05\,{\rm M}_{\odot} and fPBH=1.5×104f_{\rm PBH}=1.5\times 10^{-4}, fDC=1f_{\rm DC}=1. Compared with the observed CRB excess Fixsen et al. (2011), the cosmic X-ray background excess Cappelluti et al. (2017) and the isotropic gamma-ray background (IRGB) Ackermann et al. (2015), the contribution from PBHs in this model is much smaller, so this model is entirely consistent with the existing observations. Of course, it also means that to detect PBHs by using the CEB is a challenge. Such a background must be very isotropic compared with the background built up by astrophysical sources since the dark matter halos hosting the PBHs are much less clustered than galaxies.

Refer to caption
Figure 13: The Cosmic Electromagnetic Background contributed by our model of PBHs inside dark matter halos, assuming M=1.05MM_{\bullet}=1.05\,{\rm M}_{\odot}, fPBH=1.5×104f_{\rm PBH}=1.5\times 10^{-4} and z>10z>10. For comparison we plot the observed cosmic radio background excess Fixsen et al. (2011), the observed cosmic X-ray background (CXB) excess Cappelluti et al. (2017), and the observed isotropic gamma-ray background (IGRB) Ackermann et al. (2015). Note that the IGRB should be dominated by the unresolved star-forming galaxies Roth et al. (2021). Our PBH contribution is safely smaller than all these observational limits.

IV Summary & Discussion

In this paper we investigate the contribution to the CRB at Cosmic Dawn from PBHs, and the influence on the 21-cm global spectrum. We assume that a PBH can accrete the ambient IGM, form accretion flow and launch a jet. The radio and X-ray radiation of a PBH is calculated either from an empirical LRLXL_{R}-L_{X} relation derived from observed stellar-mass black holes and supermassive black holes that have jets, or directly from a jet model.

We found the following:

  • If the empirical LRLXL_{R}-L_{X} relation is applicable to PBHs, generally PBHs can only produce negligible CRB (compared with the CMB) at the Cosmic Dawn, no matter whether they are naked PBHs or inside dark matter halos. If PBHs produce a stronger CRB (with a larger MM_{\bullet} and/or a higher fPBHf_{\rm PBH}), then they will inevitably heat and ionize the IGM significantly, not only break the observational constraints on the IGM thermal history, but also reduce or even fully erase the 21-cm absorption signal.

    However, even when fPBHf_{\rm PBH} is small so the heating and ionization of the IGM are negligible, PBHs can still produce a strong Lyα\alpha background to couple TST_{\rm S} tightly with TkT_{\rm k} at Cosmic Dawn. So at least PBHs have the potential to make the 21-cm absorption reach the theoretical maximum in the absence of an extra radiation background, say 200\sim-200 mK.

  • If the empirical LRLXL_{R}-L_{X} relation is NOT applicable to PBHs, i.e. the PBHs have strong radio synchrotron radiation from the jets, without producing as much X-ray as inferred from the observed LRLXL_{R}-L_{X} relation, they can produce a CRB that is strong enough to boost the 21-cm absorption signal to 500\sim-500 mK at the Cosmic Dawn, while at the same time only cause negligible or modest heating/ionization on the IGM, as long as log(fPBH)=1.8log(M)3.5\log(f_{\rm PBH})=-1.8\log(M_{\bullet})-3.5 for 1MM300M1\,{\rm M}_{\odot}\lesssim M_{\bullet}\lesssim 300\,{\rm M}_{\odot}.

    In this case, the required minimum fraction of power finally converted into relativistic electrons compared with the accretion power is ηeηjet0.02\eta_{e}\eta_{\rm jet}\sim 0.02; or minimum magnetic field is 200\sim 200 G, or the minimum spectrum index of injected relativistic electron distribution is 2.5\sim 2.5. These parameter values are quite plausible, at least for the observed stellar-mass black holes with jets.

  • If PBHs are inside the dark matter halos, at first the accretion rate decreases with decreasing redshift, but then starts to increase since the dark matter halos help attracting the gas. As a result, the 21-cm absorption reaches a maximum at some redshift ztroughz_{\rm trough}. Before this time both the CRB and heating are weak, so the 21-cm absorption is weak; after this time, the heating is more efficient so that the 21-cm absorption is reduced even though the CRB keeps increasing, producing a trough in the 21-cm global spectrum δTb(ν)\delta T_{\rm b}(\nu). The redshift and depth of the the trough depends on the PBH mass and abundance. For M=1.05MM_{\bullet}=1.05\,{\rm M}_{\odot} and fPBH=1.5×104f_{\rm PBH}=1.5\times 10^{-4}, it gives ztrough17z_{\rm trough}\sim 17 and δTb(ztrough)500\delta T_{\rm b}(z_{\rm trough})\sim-500 mK.

In our model, for PBHs inside dark matter halos, if M=30MM_{\bullet}=30\,{\rm M}_{\odot} and z=10z=10, the radio flux of one such PBH at 1.41.4 GHz is less than 10510^{-5} mJy. This is far below the SKA sensitivity (Braun et al., 2019) limit, so individual emission of such objects are still beyond the reach of the upcoming large radio telescope. The influence on the 21-cm signal provides some potential for revealing its properties.

Launching a jet requires the angular momentum supply. Suppose that the angular momentum of a PBH gradually grows via accretion, then the PBH can only launch a jet after a sufficiently long time of accretion which builds up its angular momentum. Ref. De Luca et al. (2020b) provided an intensive study about the growth of spin for PBHs with various masses. They found that: for the PBHs smaller than 𝒪(30)M\mathcal{O}(30)~{}M_{\odot}, the spin is likely always negligible. However, for heavier PBHs, the spin increases fast via super-Eddington accretion after z30z\lesssim 30, and finally reaches the maximum in timescale \lesssim Hubble time. It provides not only the necessary mechanism for a PBH to launch the jet, but also a natural solution to the problem of PBHs generate too strong 21-cm absorption signal at stage much earlier than Cosmic Dawn in the naked PBH models. However, a quantitative investigation is beyond the scope of this paper, we leave it to future work.

For a regular black hole, the jet is supplied by its spin, and the spindown timescale is 0.2\sim 0.2 Gyr Narayan et al. (2022), which is comparable to the Hubble time at z=17z=17. Therefore the assumption for a fDC𝒪(1)f_{\rm DC}\sim\mathcal{O}(1) is not unreasonable. However, it is possible that some PBHs may not have enough time to collect sufficient angular momentum to launch the jet, so fDCf_{\rm DC} can also be significantly smaller than 1. In this case, our predicted 21-cm absorption should be reduced by a factor fDC\sim f_{\rm DC}. In other words, to produce the same amount of the signal, the required PBH abundance should be boosted by a factor 1/fDC\sim 1/f_{\rm DC}. But we do not expect fDC0.01f_{\rm DC}\lesssim 0.01, otherwise either the required fPBHf_{\rm PBH} is too high, or the X-ray heating effect will erase the 21-cm absorption.

We also estimated the contributions from PBH binaries to X-ray heating and CRB. If the PBHs are highly clustered, for example δdc=1000\delta_{\rm dc}=1000, then their binaries may have important contributions when fPBH104103f_{\rm PBH}\gtrsim 10^{-4}-10^{-3}. Such an estimation is preliminary, but could be considered as a useful references, though the following ignored effects should be noted:

a) For dense star/PBH clusters, if the accretion radii of the members overlap significantly, then the clusters will be in the coherent accretion mode, for which the accretion rate of each member is boosted by a factor of NcN_{\rm c}, where NcN_{\rm c} is the number of cluster members Lin and Murray (2007); Kaaz et al. (2019); Hütsi et al. (2019). Ref. Hütsi et al. (2019) pointed out that for PBHs the coherent boost effect is sizable when M10MM_{\bullet}\gtrsim 10~{}{\rm M}_{\odot} and fPBH0.1f_{\rm PBH}\gtrsim 0.1.

b) PBH binaries formed much earlier than the Cosmic Dawn. Those with small semi-major axes have short coalescence time, and can therefore merge into single black holes before the Cosmic Dawn Peters (1964). Moreover, the merger rate is generally enhanced if PBH binaries are located in clusters Raidal et al. (2017); Young and Byrnes (2020); De Luca et al. (2020a). Mergers may modify the role played by PBH binaries. However, PBH binaries can be either disrupted or hardened (the binding energy is increased) when they interact with the surrounding PBHs Raidal et al. (2017, 2019); Young and Byrnes (2020); De Luca et al. (2020a); Jedamzik (2020); Vaskonen and Veermäe (2020), and the interaction may happen frequently in clusters. Therefore the merger rate and the role played by the PBH binaries have complicated dependence on the clustering and fPBHf_{\rm PBH}. For example, Ref. Atal et al. (2020) found that when fPBH103f_{\rm PBH}\gtrsim 10^{-3}, in clusters the merger rate decreases with increasing fPBHf_{\rm PBH}.

There are observational constraints on the abundance of PBHs based on its non-radiative characteristics. For example, CMB anisotropy implies fPBH101106f_{\rm PBH}\lesssim 10^{-1}-10^{-6} for PBH masses from M=1MM_{\bullet}=1\,{\rm M}_{\odot} to M=102MM_{\bullet}=10^{2}\,{\rm M}_{\odot} Serpico et al. (2020); the constraints from stochastic background of gravitational wave possibly produced by PBH mergers gives fPBH102103f_{\rm PBH}\lesssim 10^{-2}-10^{-3} for M=1MM_{\bullet}=1\,{\rm M}_{\odot} to 102M10^{2}\,{\rm M}_{\odot} Chen and Huang (2020); and gravitational lensing imposes a bound of fPBH101f_{\rm PBH}\lesssim 10^{-1} in stellar mass range Oguri et al. (2018). The PBH abundance and mass for producing the strong 21-cm absorption signal as considered in this work are consistent with these constraints.

Ref. Hasinger (2020) found that the PBHs can easily produce sufficient CRB to explain the 21 cm absorption depth as observed by EDGES, provided that \sim5% of them have radio luminosity 1000 times higher than the value predicted by the empirical LRLXL_{R}-L_{X} relation in Ref. Wang et al. (2006). It implies that, even ignoring the X-ray heating, at z17z\sim 17 the TCRBT_{\rm CRB} from the PBHs must be 2TCMB\gtrsim 2T_{\rm CMB}. Our empirical model adopted the LRLXL_{R}-L_{X} relation for black holes with jet activities in Ref. Merloni et al. (2003), and incorporated the X-ray heating/ionization and their feedback self-consistently. The results of our empirical model are not as optimistic as Ref. Hasinger (2020), probably because Ref. Hasinger (2020) adopted a complicated PBH mass spectrum and allowed the cumulative fPBH1f_{\rm PBH}\sim 1, while we investigated the case with monochromatic mass in the limit of fPBH1f_{\rm PBH}\ll 1. We further investigated the PBH radio signal in greater detail by using a jet model, and found the jet parameters required for the PBHs to produce the amount of CRB that can explain the EDGES results. We found that it requires a strong magnetic field and a high fraction of energy deposited into relativistic electrons, which is still physically feasible.

Acknowledgements.
We thank Dr. Erlin Qiao for the very helpful discussions. This work is supported by National SKA Program of China no. 2020SKA0110402 and no. 2020SKA0110401, the Inter-Governmental Cooperation China-South Africa Program grant 2018YFE0120800, the Chinese Academy of Science (CAS) Scientific Research Instrument and Equipment Development Project ZDKYYQ20200008, the National Natural Science Foundation of China grant No. 11973047, and CAS frontier research grant QYZDJ-SSW-SLH017. Y.Z.M. is supported by the National Research Foundation of South Africa under grant No. 120385 and No. 120378, NITheCS program “New Insights into Astrophysics and Cosmology with Theoretical Models confronting Observational Data”, and National Natural Science Foundation of China with project 12047503. This work used computing resources of the Astronomical Big Data Joint Research Center, co-founded by National Astronomical Observatories, Chinese Academy of Sciences and Alibaba Cloud.

References

  • Furlanetto et al. (2006) S. R. Furlanetto, S. P. Oh, and F. H. Briggs, Phys. Rep. 433, 181 (2006), eprint astro-ph/0608032.
  • Pritchard and Loeb (2012) J. R. Pritchard and A. Loeb, Reports on Progress in Physics 75, 086901 (2012), eprint 1109.6012.
  • Chen and Miralda-Escudé (2004) X. Chen and J. Miralda-Escudé, ApJ 602, 1 (2004), eprint astro-ph/0303395.
  • Chen and Miralda-Escudé (2008) X. Chen and J. Miralda-Escudé, ApJ 684, 18 (2008), eprint astro-ph/0605439.
  • Bowman et al. (2018) J. D. Bowman, A. E. E. Rogers, R. A. Monsalve, T. J. Mozdzen, and N. Mahesh, Nature 555, 67 (2018), eprint 1810.05912.
  • Singh et al. (2017) S. Singh, R. Subrahmanyan, N. Udaya Shankar, M. Sathyanarayana Rao, A. Fialkov, A. Cohen, R. Barkana, B. S. Girish, A. Raghunathan, R. Somashekar, et al., ApJ 845, L12 (2017), eprint 1703.06647.
  • Voytek et al. (2014) T. C. Voytek, A. Natarajan, J. M. Jáuregui García, J. B. Peterson, and O. López-Cruz, ApJ 782, L9 (2014), eprint 1311.0014.
  • Price et al. (2018) D. C. Price, L. J. Greenhill, A. Fialkov, G. Bernardi, H. Garsden, B. R. Barsdell, J. Kocz, M. M. Anderson, S. A. Bourke, J. Craig, et al., MNRAS 478, 4193 (2018), eprint 1709.09313.
  • Sokolowski et al. (2015) M. Sokolowski, R. B. Wayth, S. E. Tremblay, S. J. Tingay, M. Waterson, J. Tickner, D. Emrich, F. Schlagenhaufer, D. Kenney, and S. Padhi, ApJ 813, 18 (2015), eprint 1509.06125.
  • de Lera Acedo et al. (2022) E. de Lera Acedo, D. I. L. de Villiers, N. Razavi-Ghods, W. Handley, A. Fialkov, A. Magro, D. Anstey, H. T. J. Bevins, R. Chiello, J. Cumner, et al., Nature Astronomy 6, 984 (2022).
  • Burns (2021) J. O. Burns, Philosophical Transactions of the Royal Society of London Series A 379, 20190564 (2021), eprint 2003.06881.
  • Chen et al. (2021) X. Chen, J. Yan, L. Deng, F. Wu, L. Wu, Y. Xu, and L. Zhou, Philosophical Transactions of the Royal Society of London Series A 379, 20190566 (2021), eprint 2007.15794.
  • Burns et al. (2021) J. Burns, G. Hallinan, T.-C. Chang, M. Anderson, J. Bowman, R. Bradley, S. Furlanetto, A. Hegedus, J. Kasper, J. Kocz, et al., arXiv e-prints arXiv:2103.08623 (2021), eprint 2103.08623.
  • Xu et al. (2018) Y. Xu, B. Yue, and X. Chen, ApJ 869, 42 (2018), eprint 1806.06080.
  • Xu et al. (2021) Y. Xu, B. Yue, and X. Chen, ApJ 923, 98 (2021), eprint 2102.12865.
  • Singh et al. (2022) S. Singh, N. T. Jishnu, R. Subrahmanyan, N. Udaya Shankar, B. S. Girish, A. Raghunathan, R. Somashekar, K. S. Srivani, and M. Sathyanarayana Rao, Nature Astronomy 6, 607 (2022), eprint 2112.06778.
  • Chen (2022) X. Chen, Nature Astronomy 6, 887 (2022).
  • Barkana (2018) R. Barkana, Nature 555, 71 (2018), eprint 1803.06698.
  • Feng and Holder (2018) C. Feng and G. Holder, ApJ 858, L17 (2018), eprint 1802.07432.
  • Ewall-Wice et al. (2018) A. Ewall-Wice, T. C. Chang, J. Lazio, O. Doré, M. Seiffert, and R. A. Monsalve, ApJ 868, 63 (2018), eprint 1803.01815.
  • Mirocha and Furlanetto (2019) J. Mirocha and S. R. Furlanetto, MNRAS 483, 1980 (2019), eprint 1803.03272.
  • Yang (2018) Y. Yang, Physical Review D 98 (2018), URL https://doi.org/10.1103%2Fphysrevd.98.103503.
  • Mirabel and Rodríguez (2022) I. F. Mirabel and L. F. Rodríguez, New A Rev. 94, 101642 (2022), eprint 2203.12741.
  • Cang et al. (2022a) J. Cang, Y. Gao, and Y.-Z. Ma, J. Cosmology Astropart. Phys. 2022, 012 (2022a), eprint 2108.13256.
  • Cang et al. (2021) J. Cang, Y. Gao, and Y.-Z. Ma, J. Cosmology Astropart. Phys. 2021, 051 (2021), eprint 2011.12244.
  • Cang et al. (2022b) J. Cang, Y.-Z. Ma, and Y. Gao, arXiv e-prints arXiv:2210.03476 (2022b), eprint 2210.03476.
  • Carr (2005) B. J. Carr, arXiv e-prints astro-ph/0511743 (2005), eprint astro-ph/0511743.
  • Harada and Carr (2005) T. Harada and B. J. Carr, Phys. Rev. D 71, 104009 (2005), eprint astro-ph/0412134.
  • Carr and Rees (1979) B. J. Carr and M. J. Rees, Nature 278, 605 (1979).
  • Carr et al. (2010) B. J. Carr, K. Kohri, Y. Sendouda, and J. Yokoyama, Phys. Rev. D 81, 104019 (2010), eprint 0912.5297.
  • Poulin et al. (2017) V. Poulin, P. D. Serpico, F. Calore, S. Clesse, and K. Kohri, Phys. Rev. D 96, 083524 (2017), eprint 1707.04206.
  • Takhistov et al. (2022) V. Takhistov, P. Lu, K. Murase, Y. Inoue, and G. B. Gelmini, MNRAS (2022), eprint 2111.08699.
  • Bird et al. (2016) S. Bird, I. Cholis, J. B. Muñoz, Y. Ali-Haïmoud, M. Kamionkowski, E. D. Kovetz, A. Raccanelli, and A. G. Riess, Phys. Rev. Lett. 116, 201301 (2016), eprint 1603.00464.
  • Barnacka et al. (2012) A. Barnacka, J. F. Glicenstein, and R. Moderski, Phys. Rev. D 86, 043001 (2012), eprint 1204.2056.
  • Serpico et al. (2020) P. D. Serpico, V. Poulin, D. Inman, and K. Kohri, Physical Review Research 2, 023204 (2020), eprint 2002.10771.
  • Green and Kavanagh (2021) A. M. Green and B. J. Kavanagh, Journal of Physics G Nuclear Physics 48, 043001 (2021), eprint 2007.10722.
  • Mittal et al. (2022) S. Mittal, A. Ray, G. Kulkarni, and B. Dasgupta, J. Cosmology Astropart. Phys. 2022, 030 (2022), eprint 2107.02190.
  • Saha and Laha (2022) A. K. Saha and R. Laha, Phys. Rev. D 105, 103026 (2022), eprint 2112.10794.
  • Mittal and Kulkarni (2022) S. Mittal and G. Kulkarni, MNRAS 510, 4992 (2022), eprint 2110.11975.
  • Tashiro et al. (2021) H. Tashiro, K. T. Abe, and T. Minoda, arXiv e-prints arXiv:2108.01916 (2021), eprint 2108.01916.
  • Ali-Haïmoud and Kamionkowski (2017) Y. Ali-Haïmoud and M. Kamionkowski, Phys. Rev. D 95, 043534 (2017), eprint 1612.05644.
  • Acharya et al. (2022) S. K. Acharya, J. Dhandha, and J. Chluba, MNRAS (2022), eprint 2208.03816.
  • Mena et al. (2019) O. Mena, S. Palomares-Ruiz, P. Villanueva-Domingo, and S. J. Witte, Phys. Rev. D 100, 043540 (2019), eprint 1906.07735.
  • Hasinger (2020) G. Hasinger, J. Cosmology Astropart. Phys. 2020, 022 (2020), eprint 2003.05150.
  • Ziparo et al. (2022) F. Ziparo, S. Gallerani, A. Ferrara, and F. Vito, MNRAS 517, 1086 (2022), eprint 2209.09907.
  • Paliya et al. (2017) V. S. Paliya, L. Marcotulli, M. Ajello, M. Joshi, S. Sahayanathan, A. R. Rao, and D. Hartmann, ApJ 851, 33 (2017), eprint 1711.01292.
  • Narayan et al. (2022) R. Narayan, A. Chael, K. Chatterjee, A. Ricarte, and B. Curd, MNRAS 511, 3795 (2022), eprint 2108.12380.
  • Ricotti (2007) M. Ricotti, ApJ 662, 53 (2007), eprint 0706.0864.
  • Ricotti et al. (2008) M. Ricotti, J. P. Ostriker, and K. J. Mack, ApJ 680, 829 (2008), eprint 0709.0524.
  • Bondi (1952) H. Bondi, MNRAS 112, 195 (1952).
  • Adamek et al. (2019) J. Adamek, C. T. Byrnes, M. Gosenca, and S. Hotchkiss, Phys. Rev. D 100, 023506 (2019), eprint 1901.08528.
  • Navarro et al. (1996) J. F. Navarro, C. S. Frenk, and S. D. M. White, ApJ 462, 563 (1996), eprint astro-ph/9508025.
  • Einasto (1965) J. Einasto, Trudy Astrofizicheskogo Instituta Alma-Ata 5, 87 (1965).
  • Bertschinger (1985) E. Bertschinger, ApJS 58, 39 (1985).
  • Yuan and Narayan (2014) F. Yuan and R. Narayan, ARA&A 52, 529 (2014), eprint 1401.0586.
  • Hektor et al. (2018) A. Hektor, G. Hütsi, L. Marzola, M. Raidal, V. Vaskonen, and H. Veermäe, Phys. Rev. D 98, 023503 (2018), eprint 1803.09697.
  • Narayan and McClintock (2008) R. Narayan and J. E. McClintock, New A Rev. 51, 733 (2008), eprint 0803.0322.
  • Xie and Yuan (2012) F.-G. Xie and F. Yuan, MNRAS 427, 1580 (2012), eprint 1207.3113.
  • Yuan and Cui (2005) F. Yuan and W. Cui, ApJ 629, 408 (2005), eprint astro-ph/0411770.
  • Rafter et al. (2009) S. E. Rafter, D. M. Crenshaw, and P. J. Wiita, AJ 137, 42 (2009), eprint 0809.3977.
  • Yang et al. (2020) X. Yang, S. Yao, J. Yang, L. C. Ho, T. An, R. Wang, W. A. Baan, M. Gu, X. Liu, X. Yang, et al., ApJ 904, 200 (2020), eprint 2007.11327.
  • Merloni et al. (2003) A. Merloni, S. Heinz, and T. di Matteo, MNRAS 345, 1057 (2003), eprint astro-ph/0305261.
  • Heinz and Sunyaev (2003) S. Heinz and R. A. Sunyaev, MNRAS 343, L59 (2003), eprint astro-ph/0305252.
  • Lusso et al. (2012) E. Lusso, A. Comastri, B. D. Simmons, M. Mignoli, G. Zamorani, C. Vignali, M. Brusa, F. Shankar, D. Lutz, J. R. Trump, et al., MNRAS 425, 623 (2012), eprint 1206.2642.
  • Ghisellini et al. (2017) G. Ghisellini, C. Righi, L. Costamante, and F. Tavecchio, MNRAS 469, 255 (2017), eprint 1702.02571.
  • Maselli et al. (2003) A. Maselli, A. Ferrara, and B. Ciardi, MNRAS 345, 379 (2003), eprint astro-ph/0307117.
  • Valdés and Ferrara (2008) M. Valdés and A. Ferrara, MNRAS 387, L8 (2008), eprint 0803.0370.
  • Mesinger et al. (2011) A. Mesinger, S. Furlanetto, and R. Cen, MNRAS 411, 955 (2011), eprint 1003.3878.
  • Yajima and Khochfar (2015) H. Yajima and S. Khochfar, MNRAS 448, 654 (2015), eprint 1405.7385.
  • Furlanetto (2006) S. R. Furlanetto, MNRAS 371, 867 (2006), eprint astro-ph/0604040.
  • Schauer et al. (2019) A. T. P. Schauer, B. Liu, and V. Bromm, ApJ 877, L5 (2019), eprint 1901.03344.
  • De Luca et al. (2020a) V. De Luca, G. Franciolini, P. Pani, and A. Riotto, J. Cosmology Astropart. Phys. 2020, 044 (2020a), eprint 2005.05641.
  • Raidal et al. (2017) M. Raidal, V. Vaskonen, and H. Veermäe, J. Cosmology Astropart. Phys. 2017, 037 (2017), eprint 1707.01480.
  • Mezcua et al. (2018) M. Mezcua, J. Hlavacek-Larrondo, J. R. Lucey, M. T. Hogan, A. C. Edge, and B. R. McNamara, MNRAS 474, 1342 (2018), eprint 1710.10268.
  • Ghisellini et al. (2010) G. Ghisellini, F. Tavecchio, L. Foschini, G. Ghirlanda, L. Maraschi, and A. Celotti, MNRAS 402, 497 (2010), eprint 0909.0932.
  • Ghisellini (2013) G. Ghisellini, Radiative Processes in High Energy Astrophysics, vol. 873 (2013).
  • Pepe et al. (2015a) C. Pepe, G. S. Vila, and G. E. Romero, A&A 584, A95 (2015a), eprint 1509.08514.
  • Ghisellini and Tavecchio (2009) G. Ghisellini and F. Tavecchio, MNRAS 397, 985 (2009), eprint 0902.0793.
  • Gardner and Done (2014) E. Gardner and C. Done, MNRAS 438, 779 (2014), eprint 1311.4320.
  • Gardner and Done (2018) E. Gardner and C. Done, MNRAS 473, 2639 (2018), eprint 1710.00316.
  • Lefa (2012) E. Lefa, Ph.D. thesis, Ruprecht-Karls University of Heidelberg, Germany (2012).
  • Gandhi et al. (2017) P. Gandhi, M. Bachetti, V. S. Dhillon, R. P. Fender, L. K. Hardy, F. A. Harrison, S. P. Littlefair, J. Malzac, S. Markoff, T. R. Marsh, et al., Nature Astronomy 1, 859 (2017), eprint 1710.09838.
  • Zdziarski (2012) A. A. Zdziarski, MNRAS 422, 1750 (2012), eprint 1105.4291.
  • Heinz (2006) S. Heinz, ApJ 636, 316 (2006), eprint astro-ph/0509777.
  • Pepe et al. (2015b) C. Pepe, G. S. Vila, and G. E. Romero, A&A 584, A95 (2015b), eprint 1509.08514.
  • Zdziarski et al. (2014) A. A. Zdziarski, P. Pjanka, M. Sikora, and Ł. Stawarz, MNRAS 442, 3243 (2014), eprint 1403.4768.
  • Ghisellini et al. (2014) G. Ghisellini, F. Tavecchio, L. Maraschi, A. Celotti, and T. Sbarrato, Nature 515, 376 (2014), eprint 1411.5368.
  • Ghisellini (2019) G. Ghisellini, Rendiconti Lincei. Scienze Fisiche e Naturali 30, 137 (2019), eprint 1809.07323.
  • Tchekhovskoy et al. (2011) A. Tchekhovskoy, R. Narayan, and J. C. McKinney, MNRAS 418, L79 (2011), eprint 1108.0412.
  • McKinney et al. (2012) J. C. McKinney, A. Tchekhovskoy, and R. D. Blandford, MNRAS 423, 3083 (2012), eprint 1201.4163.
  • Zhang et al. (2014) J. Zhang, B. Xu, and J. Lu, ApJ 788, 143 (2014), eprint 1405.1138.
  • Planck Collaboration et al. (2020) Planck Collaboration, N. Aghanim, Y. Akrami, M. Ashdown, J. Aumont, C. Baccigalupi, M. Ballardini, A. J. Banday, R. B. Barreiro, N. Bartolo, et al., A&A 641, A6 (2020), eprint 1807.06209.
  • Fixsen et al. (2011) D. J. Fixsen, A. Kogut, S. Levin, M. Limon, P. Lubin, P. Mirel, M. Seiffert, J. Singal, E. Wollack, T. Villela, et al., ApJ 734, 5 (2011), eprint 0901.0555.
  • Cappelluti et al. (2017) N. Cappelluti, Y. Li, A. Ricarte, B. Agarwal, V. Allevato, T. Tasnim Ananna, M. Ajello, F. Civano, A. Comastri, M. Elvis, et al., ApJ 837, 19 (2017), eprint 1702.01660.
  • Ackermann et al. (2015) M. Ackermann, M. Ajello, A. Albert, W. B. Atwood, L. Baldini, J. Ballet, G. Barbiellini, D. Bastieri, K. Bechtol, R. Bellazzini, et al., ApJ 799, 86 (2015), eprint 1410.3696.
  • Roth et al. (2021) M. A. Roth, M. R. Krumholz, R. M. Crocker, and S. Celli, Nature 597, 341 (2021), eprint 2109.07598.
  • Braun et al. (2019) R. Braun, A. Bonaldi, T. Bourke, E. Keane, and J. Wagg, arXiv e-prints arXiv:1912.12699 (2019), eprint 1912.12699.
  • De Luca et al. (2020b) V. De Luca, G. Franciolini, P. Pani, and A. Riotto, J. Cosmology Astropart. Phys. 2020, 052 (2020b), eprint 2003.02778.
  • Lin and Murray (2007) D. N. C. Lin and S. D. Murray, ApJ 661, 779 (2007), eprint astro-ph/0703807.
  • Kaaz et al. (2019) N. Kaaz, A. Antoni, and E. Ramirez-Ruiz, ApJ 876, 142 (2019), eprint 1901.03649.
  • Hütsi et al. (2019) G. Hütsi, M. Raidal, and H. Veermäe, Phys. Rev. D 100, 083016 (2019), eprint 1907.06533.
  • Peters (1964) P. C. Peters, Physical Review 136, 1224 (1964).
  • Young and Byrnes (2020) S. Young and C. T. Byrnes, J. Cosmology Astropart. Phys. 2020, 004 (2020), eprint 1910.06077.
  • Raidal et al. (2019) M. Raidal, C. Spethmann, V. Vaskonen, and H. Veermäe, J. Cosmology Astropart. Phys. 2019, 018 (2019), eprint 1812.01930.
  • Jedamzik (2020) K. Jedamzik, J. Cosmology Astropart. Phys. 2020, 022 (2020), eprint 2006.11172.
  • Vaskonen and Veermäe (2020) V. Vaskonen and H. Veermäe, Phys. Rev. D 101, 043015 (2020), eprint 1908.09752.
  • Atal et al. (2020) V. Atal, A. Sanglas, and N. Triantafyllou, J. Cosmology Astropart. Phys. 2020, 036 (2020), eprint 2007.07212.
  • Chen and Huang (2020) Z.-C. Chen and Q.-G. Huang, J. Cosmology Astropart. Phys. 2020, 039 (2020), eprint 1904.02396.
  • Oguri et al. (2018) M. Oguri, J. M. Diego, N. Kaiser, P. L. Kelly, and T. Broadhurst, Phys. Rev. D 97, 023518 (2018), eprint 1710.00148.
  • Wang et al. (2006) R. Wang, X.-B. Wu, and M.-Z. Kong, ApJ 645, 890 (2006), eprint astro-ph/0603514.