Skip to article frontmatterSkip to article content

1General considerations

Radiative processes are crucial to the study of cosmic-ray acceleration for two reasons.

  1. As energy losses, they limit the maximum energy reached by astrophysical accelerators (cf. diffusion-loss equation (14));
  2. As a mechanism for producing neutral secondary particles (neutrinos, gamma-rays, neutrons) that can more easily escape the magnetized regions where cosmic-ray electrons, protons and nuclei are accelerated.

Stable neutral secondary particles (neutrinos, gamma-rays) are not deflected by intervening magnetic fields (in the interstellar medium, the galactic halo, the cosmic web) and thus trace the direction of the astrophysical accelerators more robustly than the primary charged particles.

Different loss mechanisms should be considered depending on whether the cosmic ray interacts with ambient protons (and helium nuclei), photons, or magnetic fields. For a given cosmic-ray species (e.g. electrons), the comparison of loss time scales, [1pdpdt]1\left[ \frac{1}{p} \frac{\dd p}{\dd t}\right]^{-1}, or loss lengths, [1pdpdx]1\left[ \frac{1}{p} \frac{\dd p}{\dd x}\right]^{-1}, allows us to identify the main channels of secondary-radiation production and to model them in detail as a function of energy.

As the losses accumulate following p˙loss=iprocessesp˙i\dot p_\text{loss} = \sum_{i \in \text{processes}} \dot p_i, the overall loss timescale is

tloss=pp˙loss=piprocessesp˙i=1iprocessesti1.\begin{align} t_\text{loss} &= \frac{p}{\dot p_\text{loss}} \\ & = \frac{p}{\sum_{i \in \text{processes}} \dot p_i} \\ & = \frac{1}{\sum_{i \in \text{processes}} t_i^{-1}}. \end{align}

As expected, the fastest losses (i.e. the ones with the shortest tit_i) dominate the fate of the primary particle.

When multiple cosmic-ray species are present (e.g. co-acceleration of electrons and protons), we have to identify the main losses for each species and account for the relative energy throughput per particle as well as for the number of particles per primary species. Radiative processes are divided in two main categories, hadronic and leptonic, depending on the nature of the primary (pp or ZAX^{A}_{Z}X versus ee^- or e+e^+).

2Hadronic processes

2.1Spallation and pp-pp processes

The baryonic composition of the interstellar medium and of the cosmic web is dominated by hydrogen (about half molecular, H2, and half atomic, as neutral HI or ionized HII in the Milky Way) and helium (about 10%10\,\% in number, that is 40%40\,\% in mass in the Milky Way, see Brinks (1990)).

In dense environments, such as molecular clouds near supernova remnants and the circum nuclear medium of starburst galaxies, proton-proton (pp-pp) and spallation processes can be dominant (see e.g. timescales in Condorelli et al. (2023)). Nearly all intermediate-mass nuclei such as Li-Be-B, are produced by spallation. Leaving such nuclei aside, the throughput of pp-pp and spallation processes is dominated by neutrons and protons, which in turns mainly interact through:

n+pp+p+πp+pp+p+π0p+pp+n+π+\begin{align} n+p &\rightarrow p+p+\pi^-\\ p+p &\rightarrow p+p+\pi^0\\ p+p &\rightarrow p+n+\pi^+ \end{align}

The three types of pions are produced in a ``democratic’’ fashion by pp-pp interactions, so that the average number of charged pion per neutral pion produced in these interactions is

KπNπ±Nπ0with Kπ(pp)2\begin{align} &K_\pi \equiv \frac{N_{\pi^\pm}}{N_{\pi^0}}\\ \text{with } &K_\pi(p-p) \approx 2 \end{align}

The subsequent decay of pions leads to the production of neutral secondaries:

π0γγπ+μ++νμe++νe+νˉμ+νμπμ+νˉμe+νˉe+νμ+νˉμ\begin{align} \pi^0 &\rightarrow \gamma \gamma\\ \pi^+ &\rightarrow \mu^+ + \nu_\mu \rightarrow e^+ + \nu_e + \bar\nu_\mu + \nu_\mu\\ \pi^- &\rightarrow \mu^- + \bar\nu_\mu \rightarrow e^- + \bar\nu_e + \nu_\mu + \bar\nu_\mu\\ \end{align}

Although neutrinos are produced in proportion (νe:νμ:ντ)source=(1:2:0)(\nu_e : \nu_\mu : \nu_\tau)|_\text{source} = (1 : 2 : 0), detection is expected in proportion (νe:νμ:ντ)=(1:1:1)(\nu_e : \nu_\mu : \nu_\tau)|_\oplus = (1 : 1 : 1) due to neutrino oscillations.

Solution to Exercise 1

The interaction of a proton leads to the production of charged and neutral pions with branching ratios of Kπ/(Kπ+1)K_\pi/ (K_\pi+1) and 1/(Kπ+1)1/(K_\pi+1), respectively. This yields a number of neutrinos Nν=dEνdNνdEν=3Nπ±3NpKπN_\nu = \int\dd E_\nu \frac{\dd N_\nu} {\dd E_\nu} = 3 N_{\pi^\pm} \propto 3 N_p K_\pi at energy Eν=Eπ/4E_\nu = E_\pi /4 and of gamma rays Nγ=dEγdNγdEγ=2Nπ02NpN_\gamma = \int\dd E_\gamma \frac{\dd N_\gamma}{\dd E_\gamma}= 2 N_{\pi^0} \propto 2 N_p at energy Eγ=Eπ/2E_\gamma = E_\pi /2 .

The gamma-ray and neutrino number flux thus follow the relation:

dEγdNγdEγ=23KπdEνdNνdEν \int\dd E_\gamma \frac{\dd N_\gamma}{\dd E_\gamma} = \frac{2}{3K_\pi} \int\dd E_\nu \frac{\dd N_\nu}{\dd E_\nu}

The gamma-ray energy flux is dEγdNγdEγEγ2NpEπ2\int\dd E_\gamma \frac{\dd N_\gamma}{\dd E_\gamma}E_\gamma \propto 2N_p\frac{E_\pi}{2} and the neutrino energy flux is dEνdNνdEνEν3KπNpEπ4\int\dd E_\nu \frac{\dd N_\nu}{\dd E_\nu}E_\nu \propto 3K_\pi N_p\frac{E_\pi}{4}, from which we get

dEγdNγdEγEγ=43KπdEνdNνdEνEν \int\dd E_\gamma \frac{\dd N_\gamma}{\dd E_\gamma}E_\gamma = \frac{4}{3K_\pi}\int\dd E_\nu \frac{\dd N_\nu}{\dd E_\nu}E_\nu

For pp-pp interactions, Kπ2K_\pi \approx 2 so that neutrinos are three times more numerous than gamma rays. Although individually the neutrinos are twice less energetic than gamma rays, they carry as a population a power 1.51. 5 times larger than the gamma rays.

The relative energy loss of the primary nucleon is called the inelasticity of the reaction, κ. The average inelasticity per pion is thus defined as Eπ=κπEp\langle E_\pi \rangle = \kappa_\pi E_p, with κπ20%\kappa_\pi \approx 20\,\% for pp-pp interactions (also valid for pp-γ interactions discussed in the following, see Ahlers & Halzen (2018)). This means that a population of 10PeV10\,\text{PeV} protons will produce on average 2PeV2\,\text{PeV} pions, which will decay in 1PeV1\,\text{PeV} gamma rays and 500TeV500\,\text{TeV} neutrinos. Note that while the neutrinos are likely to escape interactions in their environment and on their way to Earth, the 1PeV1\,\text{PeV} gamma rays could easily interact along their path with lower-energy photons to produce pairs of electrons and positrons (see section dedicated to leptonic processes). While neutrinos and gamma-rays are expected to be produced in about equal proportion by hadronic processes, they are not necessarily expected to arrive at Earth in such proportions!

2.2Photo-erosion and pp-γ processes

Photo-erosion

The photo-erosion process, also called photo-disintegration, describes the interaction of a nucleus with a photon:

ZAX+γZbAabY+an+bp, ^{A}_{Z}X + \gamma \rightarrow ^{A-a-b}_{Z-b}Y + an + bp,

with (a,b)N2(a,b) \in \mathbb{N}^2.

The Lorentz boost per nucleon is roughly conserved in such interactions. This is particularly true at the highest energies when the nuclei can be seen as a superposition of independent nucleons.

In the nucleus rest frame, R\mathcal {R}', different nuclear/hadronic processes should be considered depending on the target photon energy (see Allard et al. (2006)):

  • Giant dipole resonance for ϵγ1020MeV\epsilon_\gamma' \approx 10-20\,\text{MeV}, i.e. a collective nuclear mode in which the neutrons and protons oscillate in antiphase. The giant dipole resonance can lead, in the final state, to the release of nn, pp, 2n2n, 2p2p, npnp, α;
  • Quasi-deuteron process for ϵγ30MeV\epsilon_\gamma' \approx 30\,\text{MeV}, i.e. the interaction of the photon with a pair of nucleons that can be released by the parent nucleus;
  • Baryon resonance for ϵγ150MeV\epsilon_\gamma' \gtrsim 150\,\text{MeV}, i.e. the interaction of the photon with single nucleons. This process is similar to the pp-γ and nn-γ processes discussed in the following subsection.

Photo-production of pions and pairs

The main interactions of a nucleon in a photon field are:

  • the photo-production of pairs:
    • p+γp+e++ep+\gamma \rightarrow p + e^+ + e^-
    • n+γn+e++en+\gamma \rightarrow n + e^+ + e^-
  • the photo-production of pions:
    • p+γp+π0p + \gamma \rightarrow p + \pi^0 and p+γn+π+p + \gamma \rightarrow n + \pi^+
    • n+γn+π0n + \gamma \rightarrow n + \pi^0 and n+γp+πn + \gamma \rightarrow p + \pi^-

Note that the three pions are not produced in a ``democratic’’ fashion in pp-γ processes, rather:

KπNπ±Nπ0with Kπ(pγ)1\begin{align} &K_\pi \equiv \frac{N_{\pi^\pm}}{N_{\pi^0}}\\ \text{with } &K_\pi(p-\gamma) \approx 1 \end{align}

In the nucleon’s rest frame, the threshold for pair and pion production are about 1MeV1\,\text{MeV} and 140MeV140\,\text{MeV}, respectively. The energy loss-length, λ=(1EddEdX)1\lambda = \langle (\frac{1}{E}\frac{\dd dE}{\dd X})^{-1} \rangle, depends on the product of the cross section and of the inelasticity (energy-loss fraction) of the reaction. This product, which is nearly constant at energies larger than a few times the threshold energy, is two orders of magnitude larger for pion production than for pair production, so that the former dominates over the latter at the highest energies.

The various loss processes relevant for the propagation of ultra-high energy cosmic rays on cosmological scales are illustrated in Figure 1.

Energy loss length of ultra-high-energy cosmic rays as a function of energy. Different nuclei are illustrated in
different colors. The relevant processes are shown with different line styles.

Figure 1:Energy loss length of ultra-high-energy cosmic rays as a function of energy. Different nuclei are illustrated in different colors. The relevant processes are shown with different line styles.

At the lowest energies, below 1EeV1\,\text{EeV}, the losses are dominated by the so-called adiabatic cooling, which is driven by the expansion of the universe, indeed:

tadiab=1EdEdt1=1EdEdzdzdt1=(1+z)H(z)(1+z)1=1/H(z)\begin{align} t_\text{adiab} &= \left|\frac{1}{E}\frac{\dd E}{\dd t} \right|^{-1}\\ &= \left|\frac{1}{E}\frac{\dd E}{\dd z}\frac{\dd z}{\dd t} \right|^{-1}\\ &= \left|-\frac{(1+z)H(z)}{(1+z)}\right|^{-1}\\ &= 1/H(z) \end{align}
Solution to Exercise 2

As a refresher, two properties of the invariant mass are used when computing a threshold energy. First, the norm of the four-momentum is conserved in the reaction, i.e. it is the same in the initial and final states. Second, this norm is Lorentz invariant, which allows us to compute it in the rest frame for the final state rather than in the lab frame. The threshold energy is by definition such as the particles in the final state are produced at rest in the center-of-mass frame. In the following calculation, we adopt the convention c=1c=1 for simplicity.

  1. The conservation of the invariant mass for the reaction p+γp+π0p+\gamma \rightarrow p+\pi^0 gives:
(Ep+Eγ)2(pp+pγ)2=(mp+mπ)2mp2+2EpEγ(1cosθ)=mp2+2mpmπ+mπ2Ep2mpmπ2Eγ(1cosθ)Ep2mpmπ4Eγ\begin{align} (E_p+E_\gamma)^2 - (\vec{p_p}+\vec{p_\gamma})^2 &= (m_p + m_\pi)^2\\ m_p^2 + 2E_p E_\gamma(1-\cos\theta) &= m_p^2 + 2m_p m_\pi + m_\pi^2\\ E_p &\approx \frac{2m_p m_\pi}{2E_\gamma(1-\cos\theta)}\\ E_p &\gtrsim \frac{2m_p m_\pi}{4E_\gamma} \end{align}

That is:

Ep, thmpmπ2×λγhc0.94GeV×140×106eV×1032×1.2eV×(λγ1mm)55EeV×(λγ1mm)\begin{align} E_{p,\text{ th}} &\approx \frac{m_p m_\pi}{2}\times \frac{\lambda_\gamma}{hc}\\ &\approx \frac{0.94\,\text{GeV}\times 140 \times 10^6\,\text{eV}\times 10^{3}}{2\times 1.2\,\text{eV} } \times \left (\frac {\lambda_\gamma}{1\,\text{mm}}\right)\\ &\approx 55\,\text{EeV} \times \left (\frac {\lambda_\gamma}{1\,\text{mm}}\right) \end{align}
  1. Similarly, for γ+γe++e\gamma+\gamma \rightarrow e^+ + e^-
(Eγ+ϵ)2(pγ+pϵ)2=(2me)22Eγϵ(1cosθ)=4me2Eγ=4me22ϵ(1cosθ)Eγ4me24ϵ\begin{align} (E_\gamma+\epsilon)^2 - (\vec{p_\gamma}+\vec{p_\epsilon})^2 &= (2m_e)^2\\ 2E_\gamma\epsilon(1-\cos\theta) &= 4m_e^2\\ E_\gamma &= \frac{4m_e^2}{2\epsilon(1-\cos\theta)}\\ E_\gamma &\geq \frac{4m_e^2}{4\epsilon} \end{align}

That is:

Eγ, th=me2×λγhc0.5112MeV21.2eV×(λγ1μm)220GeV×(λγ1μm)\begin{align} E_{\gamma,\text{ th}} &= m_e^2\times \frac{\lambda_\gamma}{hc}\\ &\approx \frac{0.511^2\,\text{MeV}^2}{1.2\,\text{eV} } \times \left(\frac {\lambda_\gamma}{1\,\mu\text{m}}\right)\\ &\approx 220\,\text{GeV} \times \left(\frac {\lambda_\gamma}{1\,\mu\text{m}}\right) \end{align}

2.3Proton synchrotron

A particle of energy EE, mass mm and charge ZeZe emits synchrotron radiation in a magnetic field BB. Averaged over the pitch angle θ, such as cosθ=βBβB\cos\theta = \frac{\vec{\beta}\cdot\vec{B}}{\beta B}, the synchrotron energy loss reads (see e.g. chapter 14 of Jackson (1999)):

dEdtsync=43σcuBβ2γ2,\left.-\frac{\dd E}{\dd t}\right|_\text{sync} = \frac{4}{3} \sigma c u_B \beta^2 \gamma^2,

with (γ,γβ)(\gamma, \gamma\vec{\beta}) the four-momentum of the particle and

σ=8π3[(Ze)24πϵ0mc2]2=σT×Z4×(mme)2.\begin{align} \sigma &= \frac{8\pi}{3} \left[ \frac{(Ze)^2}{ 4\pi\epsilon_0 mc^2} \right]^2\\ &= \sigma_\text{T} \times Z^4 \times \left( \frac{m}{ m_e} \right)^{-2}. \end{align}

In the ultra-relativistic limit, β1\beta \approx 1 adopted hereafter, the synchrotron loss rate is thus

tsync1=1EdEdtsync=43σcuBmc2×Emc2.\begin{align} t_\text{sync}^{-1} &= \left.-\frac{1}{E}\frac{\dd E}{\dd t}\right|_\text{sync}\\ &= \frac{4}{3} \frac{\sigma c u_B}{mc^2} \times \frac{E}{mc^2}. \end{align}

Note that the synchrotron loss rate increases with energy and that, for a fixed Lorentz boost γ=E/mc2\gamma = E/mc^2, the rate goes as 1/m31/m^3. If the loss rate becomes dominant with respect to the acceleration rate, the maximum energy is limited by the losses (so-called synchrotron burn-off limit).

The synchrotron radiation is emitted as photons over a range of frequencies, ν. The power emitted by a single proton per frequency unit reads

Pν(ν,γ)sync=43σcuBβ2γ2×fsync(ννc),\left.P_\nu(\nu,\gamma)\right|_\text{sync} = \frac{4}{3} \sigma c u_B \beta^2 \gamma^2 \times f_\text{sync}\left( \frac{\nu}{\nu_\text{c}} \right),

where dνfsync(ννc)=1\int \dd \nu\,f_\text{sync}\left( \frac{\nu}{\nu_\text{c}}\right)=1 so that P(γ)syncdνPν(ν,γ)sync=dEdtsync\left.P(\gamma)\right|_\text{sync} \equiv \int \dd \nu\, \left.P_\nu(\nu,\gamma)\right|_\text{sync} = \left.-\frac{\dd E}{\dd t}\right|_\text{sync}, and where the critical frequency, averaged over the pitch angle, is νc=316ZeBmγ2\nu_\text{c} = \frac{3}{16} \frac{ZeB}{m} \gamma^2.

The peak of the photon spectrum is at Epeak=αhνcE_\text{peak} = \alpha h\nu_\text{c}, where α0.3\alpha \approx 0.3 corresponds to the maximum of fsync(x)f_\text{sync}(x).[1] Using the value of the nuclear magneton, μN=e2mpπ×1014MeVT1\mu_N = \frac{e\hbar}{2m_p} \approx \pi \times 10^ {-14}\,\text{MeV\,T}^{-1}, one finds:

Epeak=3πα4μN×Bγ2×(mmp)1(Z1)3π2×0.34×1014×104×1018MeV×(B1G)(γ109)2(mmp)1(Z1)2.2MeV×(B1G)(γ109)2(mmp)1(Z1)\begin{align} E_\text{peak} & = \frac{3\pi\alpha}{4} \mu_N \times B\gamma^2 \times \left(\frac{m}{m_p}\right)^{-1} \left(\frac{Z}{1} \right)\\ & \approx \frac{3\pi^2\times0.3}{4}\times 10^{-14} \times 10^{-4} \times 10^{18} \text{\,MeV} \times \left(\frac{B}{1\, \text{G}}\right) \left(\frac{\gamma}{10^9}\right)^2 \left(\frac{m}{m_p}\right)^{-1} \left(\frac{Z}{1}\right)\\ & \approx 2.2\text{\,MeV} \times \left(\frac{B}{1\,\text{G}}\right) \left(\frac{\gamma}{10^9}\right)^2 \left(\frac{m}{m_p}\right)^{-1} \left(\frac{Z}{1}\right) \end{align}

Synchrotron emission from ultra-high energy protons at 101810^{18}\,eV in a magnetic field of 1G1\,\text{G} can thus end up as photons in the MeV range. This process has sometimes been invoked to explain the gamma-ray emission from the jetted active galactic nuclei. In the next section, we will look at alternative leptonic mechanisms.

3Leptonic processes

3.1Electron synchrotron

For a given Lorentz boost, the synchrotron losses dEdtsync\left.-\frac{\dd E}{\dd t}\right|_\text{sync} go as m2m^{-2} so that the emission from electrons can be sizeable with respect to that of protons. In what follows, we mostly discuss electrons but synchrotron could also be expected e.g. from positrons if there are no electrons around. Otherwise, the positrons and electrons annihilate resulting, for particles nearly at rest, in a 511keV511\,\text{keV} emission that is observed e.g. along the Galactic ridge.

Equation (20) can be used to determine the peak energy of the emission from e.g. 500GeV500\,\text{GeV} electrons in a 1mG1\,\text{mG} field:

Epeaksync2.2MeV××103××106×1836×(B1mG)(γ106)24eV×(B1mG)(γ106)2\begin{align} \left.E_\text{peak}\right|_\text{sync} & \approx 2.2\text{\,MeV} \times \times 10^{-3} \times \times 10^{-6} \times 1836 \times \left (\frac{B} {1\,\text{mG}}\right) \left(\frac {\gamma}{10^6}\right)^2\\ & \approx 4\text{\,eV} \times \left(\frac{B}{1\,\text{mG}}\right) \left(\frac {\gamma}{10^6}\right)^2 \end{align}

So, in a mG field, 0.5TeV0.5\,\text{TeV} electrons radiate in the UV, while 5TeV5\,\text{TeV} electrons radiate in X-rays at 0.4keV0.4\,\text{keV}. Synchrotron radiation from electrons thus explains the emission from radio to optical wavelengths (sometimes up to X-rays) of most non-thermal sources.

We can take the calculation one step further and determine the expected spectrum for the synchrotron emission (a similar derivation can of course be made for protons). Consider a number density of electrons, nen_e, following a power law of energy or Lorentz boost γ, i.e.

ne(γ)=n0γs,n_e(\gamma) = n_0 \gamma^{-s},

with e.g. s2s \approx 2 for diffusive shock acceleration.

The synchrotron luminosity of such electrons filling a spherical region of radius RR that is optically thin (transparent) to its own emission reads:

Lνsync=43πR3dγne(γ)Pν(ν,γ)sync, \left.L_\nu\right|_\text{sync} = \frac{4}{3}\pi R^3 \int \dd \gamma\, n_e(\gamma) \left.P_\nu(\nu, \gamma)\right|_\text{sync},

where PνP_\nu is the power emitted by a single electron, as defined in Equation (19).

To simplify the problem, we assume that the distribution function of emission as a function of photon frequency is sharply peaked around ανc=γ2νref\alpha\nu_c = \gamma^2 \nu_\text{ref}, where νref=3α16eBme\nu_\text{ref} = \frac{3\alpha}{16} \frac {eB}{m_e} does not depend on the electron Lorentz boost. The assumption of a sharply peaked function is called the delta approximation:

fsync(ν/νc)δ(νγ2νref). f_\text{sync}(\nu/\nu_c) \approx \delta(\nu - \gamma^2 \nu_\text{ref}).

Then, considering only the ultra-relativistic electrons (β1\beta \approx 1):

Lνsync=43πR3×43σTcuB×dγne(γ)γ2δ(νγ2νref)=43πR3×43σTcuB×dγγ2νrefne(γ)×2γνrefδ(νγ2νref)=43πR3×43σTcuB×dγγ2νrefne(γ)×δ(γννref),as δ(f(x))f(x)=δ(x)=43πR3×23σTc×uBνref×ννrefne(ννref)=43πR3n0×23σTc×uBνref×(ννref)(s1)/2.\begin{align} \left.L_\nu\right|_\text{sync} &= \frac{4}{3}\pi R^3 \times \frac{4}{3}\sigma_\text{T}c u_B \times \int \dd \gamma\, n_e (\gamma) \gamma^2 \delta(\nu - \gamma^2 \nu_\text{ref})\\ &= \frac{4}{3}\pi R^3 \times \frac{4}{3}\sigma_\text{T}c u_B \times \int \dd \gamma\, \frac{\gamma}{2\nu_\text{ref}} n_e(\gamma) \times 2\gamma \nu_\text{ref}\, \delta(\nu - \gamma^2 \nu_\text{ref})\\ &= \frac{4}{3}\pi R^3 \times \frac{4}{3}\sigma_\text{T}c u_B \times \int \dd \gamma\, \frac{\gamma}{2\nu_\text{ref}} n_e(\gamma) \times \delta\left(\gamma- \sqrt{\frac{\nu}{\nu_\text{ref}}}\right), \text{as } \delta(f(x))f'(x) = \delta(x) \\ &= \frac{4}{3}\pi R^3 \times \frac{2}{3}\sigma_\text{T}c \times \frac{u_B}{\nu_\text{ref}} \times \sqrt{\frac{\nu} {\nu_\text{ref}}} n_e\left( \sqrt{\frac{\nu}{\nu_\text{ref}}} \right)\\ &= \frac{4}{3}\pi R^3n_0 \times \frac{2}{3}\sigma_\text{T}c \times \frac{u_B}{\nu_\text{ref}} \times \left(\frac{\nu} {\nu_\text{ref}}\right)^{-(s-1)/2}. \end{align}

For a flat energy spectrum of electrons with s=2s=2, neglecting the redshift of the photons, the energy spectrum of synchrotron photons goes as:

Eγ2dNdEγ=νLν4πDL2n0R3×uBνref×ν1(s1)/2n0R3×B×ν0.5n0R3×B×Eγ0.5,\begin{align} E_\gamma^2 \frac{\dd N}{\dd E_\gamma} &= \frac{\nu L_{\nu}}{4\pi D_L^2}\\ &\propto n_0 R^3 \times \frac{u_B}{\nu_\text{ref}} \times \nu^{1-(s-1)/2}\\ &\propto n_0 R^3 \times B \times \nu^{0.5}\\ &\propto n_0 R^3 \times B \times E_\gamma^{0.5}, \end{align}

i.e. a differential photon spectrum following a power-law, dNdEγEγ1.5\frac{\dd N}{\dd E_\gamma} \propto E_\gamma^{-1.5}, from γmin2νrefγmin2B\gamma_\text{min} ^2 \nu_\text{ref} \propto \gamma_\text{min}^2 B to γmax2νrefγmax2B\gamma_\text{max}^2 \nu_\text{ref} \propto \gamma_\text{max}^2 B.

The observed range of frequencies covered by the photons can provide constraints on the range of Lorentz boosts covered by the electrons, provided the magnetic field, and the observed flux normalisation constrains the product of the magnetic field with the number of electrons in the emitting region.

3.2Inverse Compton

The inverse Compton process, which is symmetrical to the Compton process, consists of the scattering of an energetic electron onto a photon, resulting in an energy gain for the photon:

e1+γ1e2+γ2, e^-_1 + \gamma_1 \rightarrow e^-_2 + \gamma_2,

where the initial photon and electron have energies Eγ1=hν1E_{\gamma_1} = h\nu_1 and Ee1=γmec2E_{e_1} = \gamma m_e c^2 in the observer’s frame.

There are two regimes to be considered for this scattering in the rest frame R\mathcal{R}' of the electron (see Chapter 9.3 in Longair (2011)):

  • Thomson regime: hν1γhνmec2h\nu_1' \approx \gamma h\nu \ll m_ec^2. In the Thomson regime, the cross section is independent of the photon energy, σσT\sigma \approx \sigma_\text{T}, and the outgoing photon has an energy hν2=hν1h\nu_2' =h\nu_1' with an opposite momentum. Back in the observer frame, the energy of the outgoing photon is hν2γhν2γ2hν2h\nu_2 \approx \gamma h\nu_2 \approx \gamma^2 h\nu_2, which result in a net gain of energy by a factor γ2\propto \gamma^2.[2]
  • Klein-Nishina regime: hν1mec2h\nu_1' \gtrsim m_ec^2. In the Klein-Nishina regime, the recoil of the electron cannot be neglected and the cross section is suppressed as σ(hν1)1\sigma \propto (h\nu_1')^{-1}. The conservation of energy in the observer’s frame tells us that the energy gain of the photon is at most hν2hν1=(γ1)mec2h\nu_2 - h\nu_1 = (\gamma-1)m_e c^2, that is a gain γ\propto \gamma for an ultrarelativistic electron.

The inverse-Compton energy loss of an electron in an isotropic radiation field of energy density uradu_\text{rad} reads:

P(γ)IC=dEdtIC=43σcuradβ2γ2, \left.P(\gamma)\right|_\text{IC} = \left.-\frac{\dd E}{\dd t}\right|_\text{IC} = \frac{4}{3} \sigma c u_\text{rad} \beta^2 \gamma^2,

Note that synchrotron radiation can be viewed as the inverse-Compton scattering of virtual photons associated to the magnetic field (so-called method of virtual quanta). This is the fundamental reason behind the similarity of the inverse-Compton and synchrotron energy loss in Equation (16). We can then write

P(γ)IC   P(γ)sync=uraduB  . \boxed{\frac{\left.P(\gamma)\right|_\text{IC\ \ \ }}{\left.P(\gamma)\right|_\text{sync}} = \frac{u_\text{rad}} {u_{B\ \ }}}.

Similarly to synchrotron radiation, the power emitted by a single electron per unit photon frequency can be written:

Pν(ν,γ)IC=43σcβ2γ2dν1duraddν1×fIC(νγ2ν1),\left.P_\nu(\nu,\gamma)\right|_\text{IC} = \frac{4}{3} \sigma c \beta^2 \gamma^2 \int \dd \nu_1 \frac{\dd u_\text {rad}}{\dd \nu_1} \times f_\text{IC} \left(\frac{\nu}{\gamma^2\nu_1}\right),

where dνfIC(νγ2ν1)=1\int \dd \nu f_\text{IC} \left(\frac{\nu}{\gamma^2\nu_1}\right) = 1.

The inverse-Compton luminosity of ultrarelativistic (β1\beta \approx 1) electrons filling an optically-thin spherical region of radius RR is:

LνIC=43πR3dγne(γ)Pν(ν,γ)IC=43πR3×dγ43σcne(γ)γ2dν1duraddν1×fIC(νγ2ν1)\begin{align} \left.L_\nu\right|_\text{IC} &= \frac{4}{3}\pi R^3 \int \dd \gamma\, n_e(\gamma) \left.P_\nu(\nu, \gamma) \right|_\text{IC}\\ &= \frac{4}{3}\pi R^3 \times \int \dd \gamma\, \frac{4}{3} \sigma c n_e(\gamma) \gamma^2 \int \dd \nu_1 \frac{\dd u_\text {rad}}{\dd \nu_1} \times f_\text{IC} \left(\frac{\nu}{\gamma^2\nu_1}\right) \end{align}

In the Thomson regime, the function fIC(x)f_\text{IC}(x) peaks around x2x\approx2. The delta approximation can be roughly adopted so that

fIC(ν/γ2ν1)δ(ν2γ2ν1). f_\text{IC}(\nu/\gamma^2\nu_1) \approx \delta(\nu - 2\gamma^2 \nu_1).

Then, the inverse-Compton luminosity of the source reads:

LνIC=43πR3×43σTc×dγne(γ)γ2dν1duraddν1(ν1)×δ(ν2γ2ν1)=43πR3×43σTc×dγ12ne(γ)dν1duraddν1(ν1)×2γ2δ(ν2γ2ν1)=43πR3×43σTc×dγ12ne(γ)dν1duraddν1(ν1)×δ(ν1ν2γ2)=43πR3×23σTc×dγne(γ)×duraddν1(ν2γ2).\begin{align} \left.L_\nu\right|_\text{IC} &= \frac{4}{3}\pi R^3 \times \frac{4}{3} \sigma_\text{T} c \times \int \dd \gamma\, n_e (\gamma) \gamma^2 \int \dd \nu_1 \frac{\dd u_\text {rad}}{\dd \nu_1}(\nu_1) \times \delta(\nu - 2\gamma^2 \nu_1)\\ &= \frac{4}{3}\pi R^3 \times \frac{4}{3} \sigma_\text{T} c \times \int \dd \gamma\, \frac{1}{2} n_e (\gamma) \int \dd \nu_1 \frac{\dd u_\text{rad}}{\dd \nu_1}(\nu_1) \times 2\gamma^2 \delta(\nu - 2\gamma^2 \nu_1)\\ &= \frac{4}{3}\pi R^3 \times \frac{4}{3} \sigma_\text{T} c \times \int \dd \gamma\, \frac{1}{2} n_e (\gamma) \int \dd \nu_1 \frac{\dd u_\text{rad}}{\dd \nu_1}(\nu_1) \times \delta(\nu_1 - \frac{\nu}{2\gamma^2})\\ &= \frac{4}{3}\pi R^3 \times \frac{2}{3} \sigma_\text{T} c \times \int \dd \gamma\, n_e (\gamma) \times \frac{\dd u_\text{rad}}{\dd \nu_1}\left(\frac{\nu}{2\gamma^2}\right). \end{align}

The radiation field can have different origins depending on the environment. We speak of external inverse Compton when the photon field does not come directly from the electron population. For example, it can be an optical/infrared photon field from the environment of the compact object (e.g. accretion disk) or a diffuse photon field such as the CMB for an extended emission region.

A special case, found in many non-thermal sources, is when the scattered photon field comes from the electrons themselves, namely their synchrotron emission. This is known as the synchrotron self-Compton (SSC) model. The energy density in the spherical region of radius RR is related to the synchrotron luminosity of this region by the relation (see e.g. Section 6 in Ghisellini (2013)):

43πR3duraddν1dν1=tcross×Lν1syncdν1,\frac{4}{3}\pi R^3 \frac{\dd u_\text{rad}}{\dd \nu_1} \dd \nu_1 = t_\text{cross} \times \left.L_{\nu_1}\right|_\text {sync} \dd \nu_1,

where tcross=34Rct_\text{cross} = \frac{3}{4}\frac{R}{c} is the average crossing time of the photons through the spherical region.

Using Equations (33) and (34), the SSC losses can then be written as:

LνSSC=σTR2×dγne(γ)×Lν1sync(ν2γ2).\begin{align} \left.L_\nu\right|_\text{SSC} &= \frac{\sigma_\text{T} R}{2} \times \int \dd \gamma\, n_e (\gamma) \times \left.L_{\nu_1}\right|_\text{sync}\left(\frac{\nu}{2\gamma^2}\right). \end{align}

The SSC luminosity of an electron population of density n0n_0 and Lorentz factor γˉ\bar \gamma is simply its synchrotron luminosity scaled by a factor n0σTR/2n_0 \sigma_\text{T} R/2 and shifted to frequencies larger by a factor of 2γˉ22 {\bar \gamma}^2. The same conclusion can be reached for a power-law distribution of the Lorentz boost of the electrons as in Equation (22):

LνSSC=σTR2×43πR3n0×23σTc×uBνref×dγn0γs×(ν2γ2νref)(s1)/2=σT2cR3×43πR3n0×uBνref×(ν2νref)(s1)/2×dγn0γ1=43πR3×n02ln(γmaxγmin)×σT2cR3×uBνref×(ν2νref)(s1)/2.\begin{align} \left.L_\nu\right|_\text{SSC} &= \frac{\sigma_\text{T} R}{2} \times \frac{4}{3}\pi R^3 n_0\times \frac{2}{3} \sigma_\text{T}c \times \frac{u_B}{\nu_\text{ref}} \times \int \dd \gamma\, n_0 \gamma^{-s} \times \left(\frac{\nu}{2\gamma^2\nu_\text{ref}}\right)^{-(s-1)/2}\\ &= \frac{\sigma_\text{T}^2 c R}{3} \times \frac{4}{3}\pi R^3n_0 \times \frac{u_B}{\nu_\text{ref}} \times \left(\frac {\nu}{2\nu_\text{ref}}\right)^{-(s-1)/2} \times \int \dd \gamma\, n_0 \gamma^{-1}\\ &=\frac{4}{3}\pi R^3 \times n_0^2 \ln\left( \frac{\gamma_\text{max}}{\gamma_\text{min}}\right) \times \frac{\sigma_\text {T}^2 c R}{3} \times \frac{u_B}{\nu_\text{ref}} \times \left(\frac{\nu}{2\nu_\text{ref}}\right)^{-(s-1)/2}. \end{align}

The SSC spectrum has the same photon index as the synchrotron spectrum. As we have shown in Equation (26), the frequency range covered by the synchrotron photons goes from γmin2νref\gamma_\text{min}^2 \nu_\text{ref} to γmax2νref\gamma_\text{max}^2 \nu_\text{ref}, where νrefB\nu_\text{ref} \propto B. The frequency range covered by the SSC photons should then go from 2γmin4νref2\gamma_\text{min}^4 \nu_\text{ref} to 2γmax4νref2\gamma_\text {max}^4 \nu_\text{ref}, when we neglect the Klein-Nishina effect. Note though that for γmaxhν1mec2\gamma_\text{max} h\nu_1 \gtrsim m_e c^2, which is verified e.g. for γmax=106\gamma_\text{max}=10^6 and hν1=Epeaksync4eV×(B1mG)(γ106)2h\nu_1 = \left.E_\text{peak}\right|_\text {sync} \approx 4\text{\,eV} \times \left(\frac{B}{1\,\text{mG}}\right) \left(\frac {\gamma}{10^6}\right)^2, the scattering occurs in the Klein-Nishina regime so that the SSC peak emission can reach an energy of EpeakSSCγmaxmec20.5TeV\left.E_\text{peak}\right|_\text{SSC} \approx \gamma_\text{max} m_e c^2 \approx 0.5 \text{\,TeV}.

Irrespective of the scattering regime, the ratio of the amplitude of the SSC and synchrotron peaks provides constraints on the product of the density of the region of its size, while the peak energies (or better their ratio in the Thomson regime) constrain the maximum Lorentz factor of the electrons.

3.3Bremsstrahlung

The Bremsstrahlung emission, i.e. the radiation of an electron in the electromagnetic field of a nucleus, is relevant in an astrophysical context when the density of matter (neutral or ionised hydrogen/helium) is sufficiently high, as for the ppp-p process. Through Bremsstrahlung emission, the electron can lose a substantial fraction of its kinetic energy T=(γ1)mec2T=(\gamma-1)m_e c^2. For a Maxwellian electron velocity distribution with β1\beta\ll1 and T12me(βc)2T \approx \frac{1}{2}m_e(\beta c)^2, we speak of thermal Bremsstrahlung or free-free emission, as the electron is not bound to the nucleus with which it interacts in either the initial or final states.[3] Thermal Bremsstrahlung emission is found, for example, in the warm-hot plasma of galaxy clusters and explains their X-ray continuum radiation up to 12me(βc)2=2.5keV×(β0.1)2\frac{1}{2}m_e(\beta c)^2 = 2.5\,\text{keV}\times \left(\frac{\beta}{0.1}\right)^2.

Non-thermal Bremsstrahlung is found in the interstellar medium, particularly in molecular clouds close to supernova remnants. This non-thermal radiation explains some of the gamma-ray emission from molecular clouds up to energies of (γ1)mec20.5TeV×(γ106)(\gamma-1)m_e c^2 \approx 0.5\,\text{TeV}\times \left(\frac{\gamma}{10^6}\right), in addition to the emission processes discussed above.

The energy loss of an electron through Bremsstrahlung depends on the radiation length of the electrons in the medium, X0X_0 in gcm2\text{g}\,\text{cm}^{-2}, so that:

dEdtBrem=nmpcX0×γmec2,\left.-\frac{\dd E}{\dd t}\right|_\text{Brem} = \frac{n m_p c}{X_0} \times \gamma m_e c^2,

where nmpn m_p is the mass density of protons.

The Bremsstrahlung loss rate, 1EdEdtBrem-\left.\frac{1}{E}\frac{\dd E}{\dd t}\right|_\text{Brem}, is independent of the electron energy, contrarily to synchrotron and inverse Compton loss rates which go as E\propto E. As all electrons are affected by Bremsstrahlung in the same manner, the photon index of the Bremsstrahlung spectrum is the same as the electron index.

A comparison of the various emission mechanisms invoked in a lepto-hadronic model for the emission of a supernova remnant interacting with a molecular cloud are shown in Figure 2. The energy flux of the synchrotron and external inverse-Compton components goes as Eγ2dNγdEγEγ0.5E_\gamma^2 \frac{\dd N_\gamma}{\dd E_\gamma} \propto E_\gamma^{0.5}. We can thus infer that electrons are injected in the model with a slope of s2s \approx 2. This confirmed by the Bremsstrahlung component, which is flat (Eγ2dNγdEγEγ0E_\gamma^2 \frac{\dd N_\gamma}{\dd E_\gamma} \propto E_\gamma^0 i.e. dNγdEγEγ2\frac{\dd N_\gamma}{\dd E_\gamma} \propto E_\gamma^{-2}) between 1MeV{\sim}\,1\,\text{MeV} and 1TeV{\sim}\,1\,\text{TeV}. Finally, while the leptonic emission fully explains the synchrotron peak from radio to X rays, hadronic emission is included to model the high-energy component. This hadronic emission from the ppp-p process results in π0\pi^0 decay, with a flat spectrum from 100MeV{\sim}\,100\,\text{MeV} to 1TeV{\sim}\,1\,\text{TeV} corresponding to the proton index s2s \approx 2.[4] Note the characteristic feature of the pion bump, with a sharp break close to half the pion mass at 70MeV{\sim}\,70\,\text{MeV}. This break has now been observed in the several bright supernova remnants interacting with molecular clouds in the energy range covered by the Fermi-LAT satellite, demonstrating the co-acceleration of protons and electrons in these environments.

Spectral energy distribution, i.e. energy flux as a function of photon energy from
radio wavelengths to TeV gamma rays, of the supernova remnant SN 1006 interacting with a molecular cloud. The
radiation processes are indicated with lines of different colors and styles, as indicated in the figure. Extracted
from .

Figure 2:Spectral energy distribution, i.e. energy flux as a function of photon energy from radio wavelengths to TeV gamma rays, of the supernova remnant SN 1006 interacting with a molecular cloud. The radiation processes are indicated with lines of different colors and styles, as indicated in the figure. Extracted from H.E.S.S. Collaboration (2010).

Solution to Exercise 3
  1. We have seen that the peak synchrotron energy depends linearly on the magnetic field and quadratically on the electron Lorentz boost. Inverting Equation (21) to find the electron Lorentz boost, one gets:
    γ106×(hν4eV)1/2(B1mG)1/2106×500×100×(hν2keV)1/2(B10μG)1/22×108×(hν2keV)1/2(B10μG)1/2\begin{align} \gamma &\approx 10^6 \times \left( \frac{h\nu}{4\,\text{eV}} \right)^{1/2} \left( \frac{B}{1\,\text{mG}} \right)^ {-1/2}\\ &\approx 10^6 \times \sqrt{500} \times \sqrt{100} \times \left( \frac{h\nu}{2\,\text{keV}} \right)^{1/2} \left( \frac{B} {10\,\mu\text{G}} \right)^{-1/2}\\ &\approx 2 \times 10^8 \times \left( \frac{h\nu}{2\,\text{keV}} \right)^{1/2} \left( \frac{B} {10\,\mu\text{G}} \right)^{-1/2} \end{align}
    This corresponds to an electron energy of Ee=γmec2100TeVE_e = \gamma m_e c^2 \approx 100\,\text{TeV}.
  2. The synchrotron cooling timescale of an electron is
    tsync=[1EdEdtsync]1=34mec2σTcuB×(Emec2)1\begin{align} t_\text{sync} &= \left[\left.-\frac{1}{E} \frac{\dd E}{\dd t}\right|_\text{sync} \right]^{-1} \\ &= \frac{3}{4} \frac{m_e c^2}{\sigma_\text{T} c u_B} \times \left(\frac{E}{m_e c^2}\right)^{-1} \end{align}
    Noting that uB=B22μ0=2.5MeVm3×(B10μG)2u_B = \frac{B^2}{2\mu_0} = 2.5\,\text{MeV\,m}^{-3} \times \left(\frac{B}{10\,\mu\text{G}} \right)^2, we get
    tsync=3×0.54×0.671028×3108×2.5×2×108s×(B10μG)2(γ2×108)1=1.2kyr×(B10μG)2(γ2×108)1\begin{align} t_\text{sync} &= \frac{3 \times 0.5}{4 \times 0.67 \cdot 10^{-28} \times 3 \cdot 10^8 \times 2.5 \times 2 \times 10^8}\,\text{s} \times \left (\frac{B}{10\,\mu\text{G}} \right)^{-2} \left(\frac{\gamma}{2 \times 10^8}\right)^{-1}\\ & = 1.2\,\text{kyr} \times \left (\frac{B}{10\,\mu\text{G}} \right)^{-2} \left(\frac{\gamma}{2 \times 10^8}\right)^{-1} \end{align}
    i.e.
    ctsync=0.31×1.2kpc×(B10μG)2(γ2×108)1=370pc×(B10μG)2(γ2×108)1\begin{align} c t_\text{sync} &= 0.31 \times 1.2\,\text{kpc} \times \left(\frac{B}{10\,\mu\text{G}} \right)^{-2} \left(\frac{\gamma}{2 \times 10^8}\right)^{-1}\\ &= 370\,\text{pc} \times \left(\frac{B}{10\,\mu\text{G}} \right)^{-2} \left(\frac{\gamma}{2 \times 10^8}\right)^{-1} \end{align}
  3. At time t=tsynct = t_\text{sync}, the diffusion length of the electrons goes as
    rdiff=4Dtsync=4pc3eBtsync2γ3BcμBctsync2×21083×109197×10155.81011×3.11016×370pc×(B10μG)3/26×3.7×1039.3×5.8×104pc×(B10μG)3/22pc×(B10μG)3/2\begin{align} r_\text{diff} &= \sqrt{4Dt_\text{sync}}\\ &= \sqrt{\frac{4pc}{3eB}t_\text{sync}}\\ &\approx \sqrt{\frac{2 \gamma}{3B} \frac{\hbar c}{\mu_\text{B}} ct_\text{sync}}\\ &\approx \sqrt{\frac{2\times 2 \cdot 10^8}{3 \times 10^{-9}} \frac{197 \times 10^{-15}}{5.8 \cdot 10^{-11} \times 3. 1 \cdot 10^{16}} \times 370}\,\text{pc} \times \left(\frac{B}{10\,\mu\text{G}} \right)^{-3/2}\\ &\approx \sqrt{\frac{6 \times 3.7 \times 10^{-3}}{9.3 \times 5.8 \times 10^{-4}}}\,\text{pc} \times \left(\frac{B} {10\,\mu\text{G}} \right)^{-3/2}\\ &\approx 2\,\text{pc} \times \left(\frac{B}{10\,\mu\text{G}} \right)^{-3/2}\\ \end{align}
  4. The observed thickness of the filaments is R=dtanθ2.4kpc×2.53600×π18030mpcR = d \tan \theta \approx 2.4\,\text{kpc} \times \frac{2.5}{3600} \times \frac{\pi}{180} \approx 30\,\text{mpc}. If the observed size of the filaments is limited by diffusion, rdiff=Rr_\text{diff} =R, we get a magnetic field
    B=10μG×(R2pc)2/3=10μG×(20.03)2/3×(R30mpc)2/3=160μG×(R30mpc)2/3\begin{align} B &= 10\,\mu\text{G} \times \left(\frac{R}{2\,\text{pc}}\right)^{-2/3}\\ &= 10\,\mu\text{G} \times \left(\frac{2}{0.03}\right)^{2/3} \times \left(\frac{R}{30\,\text{mpc}}\right)^ {-2/3}\\ &= 160\,\mu\text{G} \times \left(\frac{R}{30\,\text{mpc}}\right)^{-2/3}\\ \end{align}

This is indeed the order of magnitude of the magnetic field inferred in such environments.

Footnotes
  1. The exact form of the function ff, which depends on the survival function of the modified Bessel function of order 5/35/3, is not necessary for the development of the argument. For more details, see Rybicki & Lightman (1986).

  2. Averaging over the angular configurations, one finds hν2=43(γβ)2hν2h\nu_2 = \frac{4}{3} (\gamma\beta)^2 h\nu_2.

  3. The opposite term, bound-bound radiation, is used for the emission of spectral lines by partially ionised atoms.

  4. For an energy-independent inelasticity, the pions and the photons have the same index as the protons, following the same argument as for Bremsstrahlung emission.

References
  1. Brinks, E. (1990). The cool phase of the interstellar medium - Atomic gas. In H. A. Thronson Jr. & J. M. Shull (Eds.), The Interstellar Medium in Galaxies (Vol. 161, pp. 39–65). 10.1007/978-94-009-0595-5_3
  2. Condorelli, A., Boncioli, D., Peretti, E., & Petrera, S. (2023). Testing hadronic and photohadronic interactions as responsible for ultrahigh energy cosmic rays and neutrino fluxes from starburst galaxies. \prd, 107(8), 083009. 10.1103/PhysRevD.107.083009
  3. Ahlers, M., & Halzen, F. (2018). Opening a new window onto the universe with IceCube. Progress in Particle and Nuclear Physics, 102, 73–88. 10.1016/j.ppnp.2018.05.001
  4. Allard, D., Ave, M., Busca, N., Malkan, M. A., Olinto, A. V., Parizot, E., Stecker, F. W., & Yamamoto, T. (2006). Cosmogenic neutrinos from the propagation of ultrahigh energy nuclei. \jcap, 2006(9), 005. 10.1088/1475-7516/2006/09/005
  5. Jackson, J. D. (1999). Classical electrodynamics (3rd ed.). Wiley. http://cdsweb.cern.ch/record/490457