6. Photon Physics

Photons, being neutral particles, behave much in the same manner as neutrons, traveling in straight lines and experiencing occasional collisions that change their energy and direction. Photons undergo four basic interactions as they pass through matter: coherent (Rayleigh) scattering, incoherent (Compton) scattering, photoelectric effect, and pair/triplet production. Photons with energy in the MeV range may also undergo photonuclear reactions with an atomic nucleus. In addition to these primary interaction mechanisms, all processes other than coherent scattering can result in the excitation/ionization of atoms. The de-excitation of these atoms can result in the emission of electrons and photons. Electrons themselves also can produce photons by means of bremsstrahlung radiation.

6.1. Photon Interactions

6.1.1. Coherent (Rayleigh) Scattering

The elastic scattering of a photon off a free charged particle is known as Thomson scattering. The differential cross section is independent of the energy of the incident photon. For scattering off a free electron, the differential cross section is

(1)\[\frac{d\sigma}{d\mu} = \pi r_e^2 ( 1 + \mu^2 )\]

where \(\mu\) is the cosine of the scattering angle and \(r_e\) is the classical electron radius. Thomson scattering can generally occur when the photon energy is much less than the rest mass energy of the particle.

In practice, most elastic scattering of photons off electrons happens not with free electrons but those bound in atoms. This process is known as Rayleigh scattering. The radiation scattered off of individual bound electrons combines coherently, and thus Rayleigh scattering is also known as coherent scattering. Even though conceptually we think of the photon interacting with a single electron, because the wave functions combine constructively it is really as though the photon is interacting with the entire atom.

The differential cross section for Rayleigh scattering is given by

(2)\[\begin{aligned} \frac{d\sigma(E,E',\mu)}{d\mu} &= \pi r_e^2 ( 1 + \mu^2 )~\left| F(x,Z) + F' + iF'' \right|^2 \\ &= \pi r_e^2 ( 1 + \mu^2 ) \left [ ( F(x,Z) + F'(E) )^2 + F''(E)^2 \right ] \end{aligned}\]

where \(F(x,Z)\) is a form factor as a function of the momentum transfer \(x\) and the atomic number \(Z\) and the term \(F' + iF''\) accounts for anomalous scattering which can occur near absorption edges. In a Monte Carlo simulation, when coherent scattering occurs, we only need to sample the scattering angle using the differential cross section in (2) since the energy of the photon does not change. In OpenMC, anomalous scattering is ignored such that the differential cross section becomes

(3)\[\frac{d\sigma(E,E',\mu)}{d\mu} = \pi r_e^2 ( 1 + \mu^2 ) F(x, Z)^2\]

To construct a proper probability density, we need to normalize the differential cross section in (3) by the integrated coherent scattering cross section:

(4)\[p(\mu) d\mu = \frac{\pi r_e^2}{\sigma(E)} ( 1 + \mu^2 ) F(x, Z)^2 d\mu.\]

Since the form factor is given in terms of the momentum transfer, it is more convenient to change variables of the probability density to \(x^2\). The momentum transfer is traditionally expressed as

(5)\[x = a k \sqrt{1 - \mu}\]

where \(k\) is the ratio of the photon energy to the electron rest mass, and the coefficient \(a\) can be shown to be

(6)\[a = \frac{m_e c^2}{\sqrt{2}hc} \approx 2.914329\times10^{-9}~\text{m}\]

where \(m_e\) is the mass of the electron, \(c\) is the speed of light in a vacuum, and \(h\) is Planck’s constant. Using (5), we have \(\mu = 1 - [x/(ak)]^2\) and \(d\mu/dx^2 = -1/(ak)^2\). The probability density in \(x^2\) is

(7)\[p(x^2) dx^2 = p(\mu) \left | \frac{d\mu}{dx^2} \right | dx^2 = \frac{2\pi r_e^2 A(\bar{x}^2,Z)}{(ak)^2 \sigma(E)} \left ( \frac{1 + \mu^2}{2} \right ) \left ( \frac{F(x, Z)^2}{A(\bar{x}^2, Z)} \right ) dx^2\]

where \(\bar{x}\) is the maximum value of \(x\) that occurs for \(\mu=-1\),

(8)\[\bar{x} = a k \sqrt{2} = \frac{m_e c^2}{hc} k,\]

and \(A(x^2, Z)\) is the integral of the square of the form factor:

(9)\[A(x^2, Z) = \int_0^{x^2} F(x,Z)^2 dx^2.\]

As you see, we have multiplied and divided the probability density by the integral of the squared form factor so that the density in (7) is expressed as the product of two separate densities in parentheses. In OpenMC, a table of \(A(x^2, Z)\) versus \(x^2\) is pre-generated and used at run-time to do a table search on the cumulative distribution function:

(10)\[\frac{\int_0^{x^2} F(x,Z)^2 dx^2}{\int_0^{\bar{x}^2} F(x,Z)^2 dx^2}\]

Once a trial \(x^2\) value has been selected, we can calculate \(\mu\) and perform rejection sampling using the Thomson scattering differential cross section. The complete algorithm is as follows:

  1. Determine \(\bar{x}^2\) using (8).
  2. Determine \(A_{max} = A(\bar{x}^2, Z)\) using the pre-generated tabulated data.
  3. Sample the cumulative density by calculating \(A' = \xi_1 A_{max}\) where \(\xi_1\) is a uniformly distributed random number.
  4. Perform a binary search to determine the value of \(x^2\) which satisfies \(A(x^2, Z) = A'\).
  5. By combining (5) and (8), calculate \(\mu = 1 - 2x^2/\bar{x}^2\).
  6. If \(\xi_2 < (1 + \mu^2)/2\), accept \(\mu\). Otherwise, repeat the sampling at step 3.

6.1.2. Incoherent (Compton) Scattering

Before we noted that the Thomson cross section gives the behavior for photons scattering off of free electrons valid at low energies. The formula for photon scattering off of free electrons that is valid for all energies can be found using quantum electrodynamics and is known as the Klein-Nishina formula after the two authors who discovered it:

(11)\[\frac{d\sigma_{KN}}{d\mu} = \pi r_e^2 \left ( \frac{k'}{k} \right)^2 \left [ \frac{k'}{k} + \frac{k}{k'} + \mu^2 - 1 \right ]\]

where \(k\) and \(k'\) are the ratios of the incoming and exiting photon energies to the electron rest mass energy equivalent (0.511 MeV), respectively. Although it appears that the outgoing energy and angle are separate, there is actually a one-to-one relationship between them such that only one needs to be sampled:

(12)\[k' = \frac{k}{1 + k(1 - \mu)}.\]

Note that when \(k'/k\) goes to one, i.e., scattering is elastic, the Klein-Nishina cross section becomes identical to the Thomson cross section. In general though, the scattering is inelastic and is known as Compton scattering. When a photon interacts with a bound electron in an atom, the Klein-Nishina formula must be modified to account for the binding effects. As in the case of coherent scattering, this is done by means of a form factor. The differential cross section for incoherent scattering is given by

(13)\[\frac{d\sigma}{d\mu} = \frac{d\sigma_{KN}}{d\mu} S(x,Z) = \pi r_e^2 \left ( \frac{k'}{k} \right )^2 \left [ \frac{k'}{k} + \frac{k}{k'} + \mu^2 - 1 \right ] S(x,Z)\]

where \(S(x,Z)\) is the form factor. The approach in OpenMC is to first sample the Klein-Nishina cross section and then perform rejection sampling on the form factor. As in other codes, Kahn’s rejection method is used for \(k < 3\) and a direct method by Koblinger is used for \(k \ge 3\). The complete algorithm is as follows:

  1. If \(k < 3\), sample \(\mu\) from the Klein-Nishina cross section using Kahn’s rejection method. Otherwise, use Koblinger’s direct method.
  2. Calculate \(x\) and \(\bar{x}\) using (5) and (8), respectively.
  3. If \(\xi < S(x, Z)/S(\bar{x}, Z)\), accept \(\mu\). Otherwise repeat from step 1.

6.1.2.1. Doppler Energy Broadening

Bound electrons are not at rest but have a momentum distribution that will cause the energy of the scattered photon to be Doppler broadened. More tightly bound electrons have a wider momentum distribution, so the energy spectrum of photons scattering off inner shell electrons will be broadened the most. In addition, scattering from bound electrons places a limit on the maximum scattered photon energy:

(14)\[E'_{\text{max}} = E - E_{b,i},\]

where \(E_{b,i}\) is the binding energy of the \(i\)-th subshell.

Compton profiles \(J_i(p_z)\) are used to account for the binding effects. The quantity \(p_z = {\bf p} \cdot {\bf q}/q\) is the projection of the initial electron momentum on \({\bf q}\), where the scattering vector \({\bf q} = {\bf p} - {\bf p'}\) is the momentum gained by the photon, \({\bf p}\) is the initial momentum of the electron, and \({\bf p'}\) is the momentum of the scattered electron. Applying the conservation of energy and momentum, \(p_z\) can be written in terms of the photon energy and scattering angle:

(15)\[p_z = \frac{E - E' - EE'(1 - \mu)/(m_e c^2)}{-\alpha \sqrt{E^2 + E'^2 - 2EE'\mu}},\]

where \(\alpha\) is the fine structure constant. The maximum momentum transferred, \(p_{z,\text{max}}\), can be calculated from (15) using \(E' = E'_{\text{max}}\). The Compton profile of the \(i\)-th electron subshell is defined as

(16)\[J_i(p_z) = \int \int \rho_i({\bf p}) dp_x dp_y,\]

where \(\rho_i({\bf p})\) is the initial electron momentum distribution. \(J_i(p_z)\) can be interpreted as the probability density function of \(p_z\).

The Doppler broadened energy of the Compton-scattered photon can be sampled by selecting an electron shell, sampling a value of \(p_z\) using the Compton profile, and calculating the scattered photon energy. The theory and methods used to do this are described in detail in LA-UR-04-0487 and LA-UR-04-0488. The sampling algorithm is summarized below:

  1. Sample \(\mu\) from (13) using the algorithm described in Incoherent (Compton) Scattering.
  2. Sample the electron subshell \(i\) using the number of electrons per shell as the probability mass function.
  3. Sample \(p_z\) using \(J_i(p_z)\) as the PDF.
  4. Calculate \(E'\) by solving (15) for \(E'\) using the sampled value of \(p_z\).
  5. If \(p_z < p_{z,\text{max}}\) for shell \(i\), accept \(E'\). Otherwise repeat from step 2.

6.1.2.2. Compton Electrons

Because the Compton-scattered photons can transfer a large fraction of their energy to the kinetic energy of the recoil electron, which may in turn go on to lose its energy as bremsstrahlung radiation, it is necessary to accurately model the angular and energy distributions of Compton electrons. The energy of the recoil electron ejected from the \(i\)-th subshell is given by

(17)\[E_{-} = E - E' - E_{b,i}.\]

The direction of the electron is assumed to be in the direction of the momentum transfer, with the cosine of the polar angle given by

(18)\[\mu_{-} = \frac{E - E'\mu}{\sqrt{E^2 +E'^2 - 2EE'\mu}}\]

and the azimuthal angle \(\phi_{-} = \phi + \pi\), where \(\phi\) is the azimuthal angle of the photon. The vacancy left by the ejected electron is filled through atomic relaxation.

6.1.3. Photoelectric Effect

In the photoelectric effect, the incident photon is absorbed by an atomic electron, which is then emitted from the \(i\)-th shell with kinetic energy

(19)\[E_{-} = E - E_{b,i}.\]

Photoelectric emission is only possible when the photon energy exceeds the binding energy of the shell. These binding energies are often referred to as edge energies because the otherwise continuously decreasing cross section has discontinuities at these points, creating the characteristic sawtooth shape. The photoelectric effect dominates at low energies and is more important for heavier elements.

When simulating the photoelectric effect, the first step is to sample the electron shell. The shell \(i\) where the ionization occurs can be considered a discrete random variable with probability mass function

(20)\[p_i = \frac{\sigma_{\text{pe},i}}{\sigma_{\text{pe}}},\]

where \(\sigma_{\text{pe},i}\) is the cross section of the \(i\)-th shell, and the total photoelectric cross section of the atom, \(\sigma_{\text{pe}}\), is the sum over the shell cross sections. Once the shell has been sampled, the energy of the photoelectron is calculated using (19).

To determine the direction of the photoelectron, we implement the method described in Kaltiaisenaho, which models the angular distribution of the photoelectrons using the K-shell cross section derived by Sauter (K-shell electrons are the most tightly bound, and they contribute the most to \(\sigma_{\text{pe}}\)). The non-relativistic Sauter distribution for unpolarized photons can be approximated as

(21)\[\frac{d\sigma_{\text{pe}}}{d\mu_{-}} \propto \frac{1 - \mu_{-}^2}{(1 - \beta_{-} \mu_{-})^4},\]

where \(\beta_{-}\) is the ratio of the velocity of the electron to the speed of light,

(22)\[\beta_{-} = \frac{\sqrt{(E_{-}(E_{-} + 2m_e c^2)}}{E_{-} + m_e c^2}.\]

To sample \(\mu_{-}\) from the Sauter distribution, we first express (21) in the form:

(23)\[f(\mu_{-}) = \frac{3}{2} \psi(\mu_{-}) g(\mu_{-}),\]

where

(24)\[\begin{aligned} \psi(\mu_{-}) &= \frac{(1 - \beta_{-}^2)(1 - \mu_{-}^2)}{(1 - \beta_{-}\mu_{-})^2}, \\ g(\mu_{-}) &= \frac{1 - \beta_{-}^2}{2 (1 - \beta_{-}\mu_{-})^2}. \end{aligned}\]

In the interval \([-1, 1]\), \(g(\mu_{-})\) is a normalized PDF and \(\psi(\mu_{-})\) satisfies the condition \(0 < \psi(\mu_{-}) < 1\). The following algorithm can now be used to sample \(\mu_{-}\):

  1. Using the inverse transform method, sample \(\mu_{-}\) from \(g(\mu_{-})\) using the sampling formula

    \[\mu_{-} = \frac{2\xi_1 + \beta_{-} - 1}{2\beta_{-}\xi_1 - \beta_{-} + 1}.\]
  2. If \(\xi_2 \le \psi(\mu_{-})\), accept \(\mu_{-}\). Otherwise, repeat the sampling from step 1.

The azimuthal angle is sampled uniformly on \([0, 2\pi)\).

The atom is left in an excited state with a vacancy in the \(i\)-th shell and decays to its ground state through a cascade of transitions that produce fluorescent photons and Auger electrons.

6.1.4. Pair Production

In electron-positron pair production, a photon is absorbed in the vicinity of an atomic nucleus or an electron and an electron and positron are created. Pair production is the dominant interaction with matter at high photon energies and is more important for high-Z elements. When it takes place in the field of a nucleus, energy is essentially conserved among the incident photon and the resulting charged particles. Therefore, in order for pair production to occur, the photon energy must be greater than the sum of the rest mass energies of the electron and positron, i.e., \(E_{\text{threshold,pp}} = 2 m_e c^2 = 1.022\) MeV.

The photon can also interact in the field of an atomic electron. This process is referred to as “triplet production” because the target electron is ejected from the atom and three charged particles emerge from the interaction. In this case, the recoiling electron also absorbs some energy, so the energy threshold for triplet production is greater than that of pair production from atomic nuclei, with \(E_{\text{threshold,tp}} = 4 m_e c^2 = 2.044\) MeV. The ratio of the triplet production cross section to the pair production cross section is approximately 1/Z, so triplet production becomes increasingly unimportant for high-Z elements. Though it can be significant in lighter elements, the momentum of the recoil electron becomes negligible in the energy regime where pair production dominates. For our purposes, it is a good approximation to treat triplet production as pair production and only simulate the electron-positron pair.

Accurately modeling the creation of electron-positron pair is important because the charged particles can go on to lose much of their energy as bremsstrahlung radiation, and the subsequent annihilation of the positron with an electron produces two additional photons. We sample the energy and direction of the charged particles using a semiempirical model described in Salvat. The Bethe-Heitler differential cross section, given by

(25)\[\frac{d\sigma_{\text{pp}}}{d\epsilon} = \alpha r_e^2 Z^2 \left[ (\epsilon^2 + (1-\epsilon)^2) (\Phi_1 - 4f_C) + \frac{2}{3}\epsilon(1-\epsilon)(\Phi_2 - 4f_C) \right],\]

is used as a starting point, where \(\alpha\) is the fine structure constant, \(f_C\) is the Coulomb correction function, \(\Phi_1\) and \(\Phi_2\) are screening functions, and \(\epsilon = (E_{-} + m_e c^2)/E\) is the electron reduced energy (i.e., the fraction of the photon energy given to the electron). \(\epsilon\) can take values between \(\epsilon_{\text{min}} = k^{-1}\) (when the kinetic energy of the electron is zero) and \(\epsilon_{\text{max}} = 1 - k^{-1}\) (when the kinetic energy of the positron is zero).

The Coulomb correction, given by

(26)\[\begin{aligned} f_C = \alpha^{2}Z^{2} \big[&(1 + \alpha^{2}Z^{2})^{-1} + 0.202059 - 0.03693\alpha^{2}Z^{2} + 0.00835\alpha^{4}Z^{4} \\ &- 0.00201\alpha^{6}Z^{6} + 0.00049\alpha^{8}Z^{8} - 0.00012\alpha^{10}Z^{10} + 0.00003\alpha^{12}Z^{12}\big] \end{aligned}\]

is introduced to correct for the fact that the Bethe-Heitler differential cross section was derived using the Born approximation, which treats the Coulomb interaction as a small perturbation.

The screening functions \(\Phi_1\) and \(\Phi_2\) account for the screening of the Coulomb field of the atomic nucleus by outer electrons. Since they are given by integrals which include the atomic form factor, they must be computed numerically for a realistic form factor. However, by assuming exponential screening and using a simplified form factor, analytical approximations of the screening functions can be derived:

(27)\[\begin{aligned} \Phi_1 &= 2 - 2\ln(1 + b^2) - 4b\arctan(b^{-1}) + 4\ln(Rm_{e}c/\hbar) \\ \Phi_2 &= \frac{4}{3} - 2\ln(1 + b^2) + 2b^2 \left[ 4 - 4b\arctan(b^{-1}) - 3\ln(1 + b^{-2}) \right] + 4\ln(Rm_{e}c/\hbar) \end{aligned}\]

where

(28)\[b = \frac{Rm_{e}c}{2k\epsilon(1 - \epsilon)\hbar}.\]

and \(R\) is the screening radius.

The differential cross section in (25) with the approximations described above will not be accurate at low energies: the lower boundary of \(\epsilon\) will be shifted above \(\epsilon_{\text{min}}\) and the upper boundary of \(\epsilon\) will be shifted below \(\epsilon_{\text{max}}\). To offset this behavior, a correcting factor \(F_0(k, Z)\) is used:

(29)\[\begin{aligned} F_0(k, Z) =~& (0.1774 + 12.10\alpha Z - 11.18\alpha^{2}Z^{2})(2/k)^{1/2} \\ &+ (8.523 + 73.26\alpha Z - 44.41\alpha^{2}Z^{2})(2/k) \\ &- (13.52 + 121.1\alpha Z - 96.41\alpha^{2}Z^{2})(2/k)^{3/2} \\ &+ (8.946 + 62.05\alpha Z - 63.41\alpha^{2}Z^{2})(2/k)^{2}. \end{aligned}\]

To aid sampling, the differential cross section used to sample \(\epsilon\) (minus the normalization constant) can now be expressed in the form

(30)\[\frac{d\sigma_{\text{pp}}}{d\epsilon} = u_1 \frac{\phi_1(\epsilon)}{\phi_1(1/2)} \pi_1(\epsilon) + u_2 \frac{\phi_2(\epsilon)}{\phi_2(1/2)} \pi_2(\epsilon)\]

where

(31)\[\begin{aligned} u_1 &= \frac{2}{3} \left(\frac{1}{2} - \frac{1}{k}\right)^2 \phi_1(1/2), \\ u_2 &= \phi_2(1/2), \end{aligned}\]
(32)\[\begin{aligned} \phi_1(\epsilon) &= \frac{1}{2}(3\Phi_1 - \Phi_2) - 4f_{C}(Z) + F_0(k, Z), \\ \phi_2(\epsilon) &= \frac{1}{4}(3\Phi_1 + \Phi_2) - 4f_{C}(Z) + F_0(k, Z), \end{aligned}\]

and

(33)\[\begin{aligned} \pi_1(\epsilon) &= \frac{3}{2} \left(\frac{1}{2} - \frac{1}{k}\right)^{-3} \left(\frac{1}{2} - \epsilon\right)^2, \\ \pi_2(\epsilon) &= \frac{1}{2} \left(\frac{1}{2} - \frac{1}{k}\right)^{-1}. \end{aligned}\]

The functions in (32) are non-negative and maximum at \(\epsilon = 1/2\). In the interval \((\epsilon_{\text{min}}, \epsilon_{\text{max}})\), the functions in (33) are normalized PDFs and \(\phi_i(\epsilon)/\phi_i(1/2)\) satisfies the condition \(0 < \phi_i(\epsilon)/\phi_i(1/2) < 1\). The following algorithm can now be used to sample the reduced electron energy \(\epsilon\):

  1. Sample \(i\) according to the point probabilities \(p(i=1) = u_1/(u_1 + u_2)\) and \(p(i=2) = u_2/(u_1 + u_2)\).

  2. Using the inverse transform method, sample \(\epsilon\) from \(\pi_i(\epsilon)\) using the sampling formula

    \[\begin{aligned} \epsilon &= \frac{1}{2} + \left(\frac{1}{2} - \frac{1}{k}\right) (2\xi_1 - 1)^{1/3} ~~~~&\text{if}~~ i = 1 \\ \epsilon &= \frac{1}{k} + \left(\frac{1}{2} - \frac{1}{k}\right) 2\xi_1 ~~~~&\text{if}~~ i = 2. \end{aligned}\]
  3. If \(\xi_2 \le \phi_i(\epsilon)/\phi_i(1/2)\), accept \(\epsilon\). Otherwise, repeat the sampling from step 1.

Because charged particles have a much smaller range than the mean free path of photons and because they immediately undergo multiple scattering events which randomize their direction, it is sufficient to use a simplified model to sample the direction of the electron and positron. The cosines of the polar angles are sampled using the leading order term of the Sauter–Gluckstern–Hull distribution,

(34)\[p(\mu_{\pm}) = C(1 - \beta_{\pm}\mu_{\pm})^{-2},\]

where \(C\) is a normalization constant and \(\beta_{\pm}\) is the ratio of the velocity of the charged particle to the speed of light given in (22).

The inverse transform method is used to sample \(\mu_{-}\) and \(\mu_{+}\) from (34), using the sampling formula

(35)\[\mu_{\pm} = \frac{2\xi - 1 + \beta_{\pm}}{(2\xi - 1)\beta_{\pm} + 1}.\]

The azimuthal angles for the electron and positron are sampled independently and uniformly on \([0, 2\pi)\).

6.2. Secondary Processes

New photons may be produced in secondary processes related to the main photon interactions discussed above. A Compton-scattered photon transfers a portion of its energy to the kinetic energy of the recoil electron, which in turn may lose the energy as bremsstrahlung radiation. The vacancy left in the shell by the ejected electron is filled through atomic relaxation, creating a shower of electrons and fluorescence photons. Similarly, the vacancy left by the electron emitted in the photoelectric effect is filled through atomic relaxation. Pair production generates an electron and a positron, both of which can emit bremsstrahlung radiation before the positron eventually collides with an electron, resulting in annihilation of the pair and the creation of two additional photons.

6.2.1. Atomic Relaxation

When an electron is ejected from an atom and a vacancy is left in an inner shell, an electron from a higher energy level will fill the vacancy. This results in either a radiative transition, in which a photon with a characteristic energy (fluorescence photon) is emitted, or non-radiative transition, in which an electron from a shell that is farther out (Auger electron) is emitted. If a non-radiative transition occurs, the new vacancy is filled in the same manner, and as the process repeats a shower of photons and electrons can be produced.

The energy of a fluorescence photon is the equal to the energy difference between the transition states, i.e.,

(36)\[E = E_{b,v} - E_{b,i},\]

where \(E_{b,v}\) is the binding energy of the vacancy shell and \(E_{b,i}\) is the binding energy of the shell from which the electron transitioned. The energy of an Auger electron is given by

(37)\[E_{-} = E_{b,v} - E_{b,i} - E_{b,a},\]

where \(E_{b,a}\) is the binding energy of the shell from which the Auger electron is emitted. While Auger electrons are low-energy so their range and bremsstrahlung yield is small, fluorescence photons can travel far before depositing their energy, so the relaxation process should be modeled in detail.

Transition energies and probabilities are needed for each subshell to simulate atomic relaxation. Starting with the initial shell vacancy, the following recursive algorithm is used to fill vacancies and create fluorescence photons and Auger electrons:

  1. If there are no transitions for the vacancy shell, create a fluorescence photon assuming it is from a captured free electron and terminate.
  2. Sample a transition using the transition probabilities for the vacancy shell as the probability mass function.
  3. Create either a fluorescence photon or Auger electron, sampling the direction of the particle isotropically.
  4. If a non-radiative transition occurred, repeat from step 1 for the vacancy left by the emitted Auger electron.
  5. Repeat from step 1 for vacancy left by the transition electron.

6.2.2. Electron-Positron Annihilation

When a positron collides with an electron, both particles are annihilated and generally two photons with equal energy are created. If the kinetic energy of the positron is high enough, the two photons can have different energies, and the higher-energy photon is emitted preferentially in the direction of flight of the positron. It is also possible to produce a single photon if the interaction occurs with a bound electron, and in some cases three (or, rarely, even more) photons can be emitted. However, the annihilation cross section is largest for low-energy positrons, and as the positron energy decreases, the angular distribution of the emitted photons becomes isotropic.

In OpenMC, we assume the most likely case in which a low-energy positron (which has already lost most of its energy to bremsstrahlung radiation) interacts with an electron which is free and at rest. Two photons with energy equal to the electron rest mass energy \(m_e c^2 = 0.511\) MeV are emitted isotropically in opposite directions.

6.2.3. Bremsstrahlung

When a charged particle is decelerated in the field of an atom, some of its kinetic energy is converted into electromagnetic radiation known as bremsstrahlung, or ‘braking radiation’. In each event, an electron or positron with kinetic energy \(T\) generates a photon with an energy \(E\) between \(0\) and \(T\). Bremsstrahlung is described by a cross section that is differential in photon energy, in the direction of the emitted photon, and in the final direction of the charged particle. However, in Monte Carlo simulations it is typical to integrate over the angular variables to obtain a single differential cross section with respect to photon energy, which is often expressed in the form

(38)\[\frac{d\sigma_{\text{br}}}{dE} = \frac{Z^2}{\beta^2} \frac{1}{E} \chi(Z, T, \kappa),\]

where \(\kappa = E/T\) is the reduced photon energy and \(\chi(Z, T, \kappa)\) is the scaled bremsstrahlung cross section, which is experimentally measured.

Because electrons are attracted to atomic nuclei whereas positrons are repulsed, the cross section for positrons is smaller, though it approaches that of electrons in the high energy limit. To obtain the positron cross section, we multiply (38) by the \(\kappa\)-independent factor used in Salvat,

(39)\[\begin{aligned} F_{\text{p}}(Z,T) = & 1 - \text{exp}(-1.2359\times 10^{-1}t + 6.1274\times 10^{-2}t^2 - 3.1516\times 10^{-2}t^3 \\ & + 7.7446\times 10^{-3}t^4 - 1.0595\times 10^{-3}t^5 + 7.0568\times 10^{-5}t^6 \\ & - 1.8080\times 10^{-6}t^7), \end{aligned}\]

where

(40)\[t = \ln\left(1 + \frac{10^6}{Z^2}\frac{T}{\text{m}_\text{e}c^2} \right).\]

\(F_{\text{p}}(Z,T)\) is the ratio of the radiative stopping powers for positrons and electrons. Stopping power describes the average energy loss per unit path length of a charged particle as it passes through matter:

(41)\[-\frac{dT}{ds} = n \int E \frac{d\sigma}{dE} dE \equiv S(T),\]

where \(n\) is the number density of the material and \(d\sigma/dE\) is the cross section differential in energy loss. The total stopping power \(S(T)\) can be separated into two components: the radiative stopping power \(S_{\text{rad}}(T)\), which refers to energy loss due to bremsstrahlung, and the collision stopping power \(S_{\text{col}}(T)\), which refers to the energy loss due to inelastic collisions with bound electrons in the material that result in ionization and excitation. The radiative stopping power for electrons is given by

(42)\[S_{\text{rad}}(T) = n \frac{Z^2}{\beta^2} T \int_0^1 \chi(Z,T,\kappa) d\kappa.\]

To obtain the radiative stopping power for positrons, (42) is multiplied by (39).

While the models for photon interactions with matter described above can safely assume interactions occur with free atoms, sampling the target atom based on the macroscopic cross sections, molecular effects cannot necessarily be disregarded for charged particle treatment. For compounds and mixtures, the bremsstrahlung cross section is calculated using Bragg’s additivity rule as

(43)\[\frac{d\sigma_{\text{br}}}{dE} = \frac{1}{\beta^2 E} \sum_i \gamma_i Z^2_i \chi(Z_i, T, \kappa),\]

where the sum is over the constituent elements and \(\gamma_i\) is the atomic fraction of the \(i\)-th element. Similarly, the radiative stopping power is calculated using Bragg’s additivity rule as

(44)\[S_{\text{rad}}(T) = \sum_i w_i S_{\text{rad},i}(T),\]

where \(w_i\) is the mass fraction of the \(i\)-th element and \(S_{\text{rad},i}(T)\) is found for element \(i\) using (42). The collision stopping power, however, is a function of certain quantities such as the mean excitation energy \(I\) and the density effect correction \(\delta_F\) that depend on molecular properties. These quantities cannot simply be summed over constituent elements in a compound, but should instead be calculated for the material. The Bethe formula can be used to find the collision stopping power of the material:

(45)\[S_{\text{col}}(T) = \frac{2 \pi r_e^2 m_e c^2}{\beta^2} N_A \frac{Z}{A_M} [\ln(T^2/I^2) + \ln(1 + \tau/2) + F(\tau) - \delta_F(T)],\]

where \(N_A\) is Avogadro’s number, \(A_M\) is the molar mass, \(\tau = T/m_e\), and \(F(\tau)\) depends on the particle type. For electrons,

(46)\[F_{-}(\tau) = (1 - \beta^2)[1 + \tau^2/8 - (2\tau + 1) \ln2],\]

while for positrons

(47)\[F_{+}(\tau) = 2\ln2 - (\beta^2/12)[23 + 14/(\tau + 2) + 10/(\tau + 2)^2 + 4/(\tau + 2)^3].\]

The density effect correction \(\delta_F\) takes into account the reduction of the collision stopping power due to the polarization of the material the charged particle is passing through by the electric field of the particle. It can be evaluated using the method described by Sternheimer, where the equation for \(\delta_F\) is

(48)\[\delta_F(\beta) = \sum_{i=1}^n f_i \ln[(l_i^2 + l^2)/l_i^2] - l^2(1-\beta^2).\]

Here, \(f_i\) is the oscillator strength of the \(i\)-th transition, given by \(f_i = n_i/Z\), where \(n_i\) is the number of electrons in the \(i\)-th subshell. The frequency \(l\) is the solution of the equation

(49)\[\frac{1}{\beta^2} - 1 = \sum_{i=1}^{n} \frac{f_i}{\bar{\nu}_i^2 + l^2},\]

where \(\bar{v}_i\) is defined as

(50)\[\bar{\nu}_i = h\nu_i \rho / h\nu_p.\]

The plasma energy \(h\nu_p\) of the medium is given by

(51)\[h\nu_p = \sqrt{\frac{(hc)^2 r_e \rho_m N_A Z}{\pi A}},\]

where \(A\) is the atomic weight and \(\rho_m\) is the density of the material. In (50), \(h\nu_i\) is the oscillator energy, and \(\rho\) is an adjustment factor introduced to give agreement between the experimental values of the oscillator energies and the mean excitation energy. The \(l_i\) in (48) are defined as

(52)\[\begin{aligned} l_i &= (\bar{\nu}_i^2 + 2/3f_i)^{1/2} ~~~~&\text{for}~~ \bar{\nu}_i > 0 \\ l_n &= f_n^{1/2} ~~~~&\text{for}~~ \bar{\nu}_n = 0, \end{aligned}\]

where the second case applies to conduction electrons. For a conductor, \(f_n\) is given by \(n_c/Z\), where \(n_c\) is the effective number of conduction electrons, and \(v_n = 0\). The adjustment factor \(\rho\) is determined using the equation for the mean excitation energy:

(53)\[\ln I = \sum_{i=1}^{n-1} f_i \ln[(h\nu_i\rho)^2 + 2/3f_i(h\nu_p)^2]^{1/2} + f_n \ln (h\nu_pf_n^{1/2}).\]

6.2.3.1. Thick-Target Bremsstrahlung Approximation

Since charged particles lose their energy on a much shorter distance scale than neutral particles, not much error should be introduced by neglecting to transport electrons. However, the bremsstrahlung emitted from high energy electrons and positrons can travel far from the interaction site. Thus, even without a full electron transport mode it is necessary to model bremsstrahlung. We use a thick-target bremsstrahlung (TTB) approximation based on the models in Salvat and Kaltiaisenaho for generating bremsstrahlung photons, which assumes the charged particle loses all its energy in a single homogeneous material region.

To model bremsstrahlung using the TTB approximation, we need to know the number of photons emitted by the charged particle and the energy distribution of the photons. These quantities can be calculated using the continuous slowing down approximation (CSDA). The CSDA assumes charged particles lose energy continuously along their trajectory with a rate of energy loss equal to the total stopping power, ignoring fluctuations in the energy loss. The approximation is useful for expressing average quantities that describe how charged particles slow down in matter. For example, the CSDA range approximates the average path length a charged particle travels as it slows to rest:

(54)\[R(T) = \int^T_0 \frac{dT'}{S(T')}.\]

Actual path lengths will fluctuate around \(R(T)\). The average number of photons emitted per unit path length is given by the inverse bremsstrahlung mean free path:

(55)\[\lambda_{\text{br}}^{-1}(T,E_{\text{cut}}) = n\int_{E_{\text{cut}}}^T\frac{d\sigma_{\text{br}}}{dE}dE = n\frac{Z^2}{\beta^2}\int_{\kappa_{\text{cut}}}^1\frac{1}{\kappa} \chi(Z,T,\kappa)d\kappa.\]

The lower limit of the integral in (55) is non-zero because the bremsstrahlung differential cross section diverges for small photon energies but is finite for photon energies above some cutoff energy \(E_{\text{cut}}\). The mean free path \(\lambda_{\text{br}}^{-1}(T,E_{\text{cut}})\) is used to calculate the photon number yield, defined as the average number of photons emitted with energy greater than \(E_{\text{cut}}\) as the charged particle slows down from energy \(T\) to \(E_{\text{cut}}\). The photon number yield is given by

(56)\[Y(T,E_{\text{cut}}) = \int^{R(T)}_{R(E_{\text{cut}})} \lambda_{\text{br}}^{-1}(T',E_{\text{cut}})ds = \int_{E_{\text{cut}}}^T \frac{\lambda_{\text{br}}^{-1}(T',E_{\text{cut}})}{S(T')}dT'.\]

\(Y(T,E_{\text{cut}})\) can be used to construct the energy spectrum of bremsstrahlung photons: the number of photons created with energy between \(E_1\) and \(E_2\) by a charged particle with initial kinetic energy \(T\) as it comes to rest is given by \(Y(T,E_1) - Y(T,E_2)\).

To simulate the emission of bremsstrahlung photons, the total stopping power and bremsstrahlung differential cross section for positrons and electrons must be calculated for a given material using (43) and (44). These quantities are used to build the tabulated bremsstrahlung energy PDF and CDF for that material for each incident energy \(T_k\) on the energy grid. The following algorithm is then applied to sample the photon energies:

  1. For an incident charged particle with energy \(T\), sample the number of emitted photons as

    \[N = \lfloor Y(T,E_{\text{cut}}) + \xi_1 \rfloor.\]
  2. Rather than interpolate the PDF between indices \(k\) and \(k+1\) for which \(T_k < T < T_{k+1}\), which is computationally expensive, use the composition method and sample from the PDF at either \(k\) or \(k+1\). Using linear interpolation on a logarithmic scale, the PDF can be expressed as

    \[p_{\text{br}}(T,E) = \pi_k p_{\text{br}}(T_k,E) + \pi_{k+1} p_{\text{br}}(T_{k+1},E),\]

    where the interpolation weights are

    \[\pi_k = \frac{\ln T_{k+1} - \ln T}{\ln T_{k+1} - \ln T_k},~~~ \pi_{k+1} = \frac{\ln T - \ln T_k}{\ln T_{k+1} - \ln T_k}.\]

    Sample either the index \(i = k\) or \(i = k+1\) according to the point probabilities \(\pi_{k}\) and \(\pi_{k+1}\).

  3. Determine the maximum value of the CDF \(P_{\text{br,max}}\).

  1. Sample the photon energies using the inverse transform method with the tabulated CDF \(P_{\text{br}}(T_i, E)\) i.e.,

    \[E = E_j \left[ (1 + a_j) \frac{\xi_2 P_{\text{br,max}} - P_{\text{br}}(T_i, E_j)} {E_j p_{\text{br}}(T_i, E_j)} + 1 \right]^{\frac{1}{1 + a_j}}\]

    where the interpolation factor \(a_j\) is given by

    \[a_j = \frac{\ln p_{\text{br}}(T_i,E_{j+1}) - \ln p_{\text{br}}(T_i,E_j)} {\ln E_{j+1} - \ln E_j}\]

    and \(P_{\text{br}}(T_i, E_j) \le \xi_2 P_{\text{br,max}} \le P_{\text{br}}(T_i, E_{j+1})\).

We ignore the range of the electron or positron, i.e., the bremsstrahlung photons are produced in the same location that the charged particle was created. The direction of the photons is assumed to be the same as the direction of the incident charged particle, which is a reasonable approximation at higher energies when the bremsstrahlung radiation is emitted at small angles.

6.3. Photon Production

In coupled neutron-photon transport, a source neutron is tracked, and photons produced from neutron reactions are transported after the neutron’s history has terminated. Since these secondary photons form the photon source for the problem, it is important to correctly describe their energy and angular distributions as the accuracy of the calculation relies on the accuracy of this source. The photon production cross section for a particular reaction \(i\) and incident neutron energy \(E\) is defined as

(57)\[\sigma_{\gamma, i}(E) = y_i(E)\sigma_i(E),\]

where \(y_i(E)\) is the photon yield corresponding to an incident neutron reaction having cross section \(\sigma_i(E)\).

The yield of photons during neutron transport is determined as the sum of the photon yields from each individual reaction. In OpenMC, production of photons is treated in an average sense. That is, the total photon production cross section is used at a collision site to determine how many photons to produce rather than the photon production from the reaction that actually took place. This is partly done for convenience but also because the use of variance reduction techniques such as implicit capture make it difficult in practice to directly sample photon production from individual reactions.

In OpenMC, secondary photons are created after a nuclide has been sampled in a neutron collision. The expected number of photons produced is

(58)\[n = w\frac{\sigma_{\gamma}(E)}{\sigma_T(E)},\]

where \(w\) is the weight of the neutron, \(\sigma_{\gamma}\) is the photon production cross section for the sampled nuclide, and \(\sigma_T\) is the total cross section for the nuclide. \(\lfloor n \rfloor\) photons are created with an additional photon produced with probability \(n - \lfloor n \rfloor\). Next, a reaction is sampled for each secondary photon. The probability of sampling the \(i\)-th reaction is given by \(\sigma_{\gamma, i}(E)/\sum_j\sigma_{\gamma, j}(E)\), where \(\sum_j\sigma_{\gamma, j} = \sigma_{\gamma}\) is the total photon production cross section. The secondary angle and energy distributions associated with the reaction are used to sample the angle and energy of the emitted photon.