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

Lyα\alpha Radiative Transfer: Monte-Carlo Simulation of the Wouthuysen-Field Effect

Kwang-il Seon11affiliation: Korea Astronomy and Space Science Institute, Daejeon 34055, Republic of Korea; [email protected] 22affiliation: Astronomy and Space Science Major, University of Science and Technology, Daejeon 34113, Republic of Korea and Chang-goo Kim33affiliation: Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA 44affiliation: Center for Computational Astrophysics, Flatiron Institute, New York, NY 10010, USA
Abstract

A three-dimensional Monte Carlo Lyα\alpha radiative transfer (RT) code, named LaRT, is developed to study the Lyα\alpha RT and the Wouthuysen-Field (WF) effect. Using the code, we calculate the line profile of Lyα\alpha radiation within the multiphase interstellar medium (ISM), with a particular emphasis on gas at low densities. We show that the WF effect is in action: the central portion of the line profile tends to approach a small slice of the Planck function with a color temperature equal to the kinetic temperature of the gas, even in a system with an optical thickness as low as τ0100500\tau_{0}\approx 100-500. We also investigate the effects of the turbulent motion of the ISM on the emergent Lyα\alpha spectrum and color temperature. The turbulent motion broadens, as generally expected, the emergent spectrum, but the color temperature is not affected by the turbulent motion in typical astrophysical environments. We utilize two multiphase ISM models, appropriate for the vicinity of the Sun, to calculate the 21-cm spin temperature of neutral hydrogen, including excitation via the Lyα\alpha resonant scattering. The first ISM model is a simple clumpy model, while the second is a self-consistent magnetohydrodynamics simulation model using the TIGRESS framework. Lyα\alpha photons originating from both H II regions and the collisionally cooling gas are taken into account. We find that the Lyα\alpha radiation field is, in general, likely to be strong enough to bring the 21-cm spin temperature of the warm neutral medium close to the kinetic temperature. The escape fraction of Lyα\alpha in our ISM models is estimated to be 720%\approx 7-20\%.

Subject headings:
line: formation – line: profiles – radiative transfer – scattering – dust, extinction — ISM: structure

1. INTRODUCTION

Temperature, density, and ionization degree of the interstellar medium (ISM) are essential parameters to understand heating, cooling, and feedback processes in star and galaxy formation. The 21-cm line originating from the hyperfine energy splitting of the ground state (n=1)(n=1) of neutral hydrogen is one of the most critical tracers to measure gas temperature and density in the neutral atomic phase of the diffuse ISM, circumgalactic medium (CGM), and intergalactic medium (IGM) (e.g., Schneider et al., 1983; Kulkarni & Heiles, 1987; Schneider, 1989; Swaters et al., 1997; Heiles & Troland, 2003a, b; Murray et al., 2015, 2017). It is often assumed that collisions with electrons, ions, and other H I atoms govern the population of the hyperfine levels of a hydrogen atom. Under a high-density condition in which collisions are frequent, the 21-cm spin temperature (or excitation temperature), which is defined by the relative population of the hyperfine levels, is equal to the gas kinetic temperature. In a low-density medium, however, collisions are not frequent enough to make the spin temperature a good tracer of the gas kinetic temperature; much of the hydrogen would then be “invisible” in observations via the 21-cm line against the cosmic microwave background (CMB) unless other mechanisms establish the spin temperature different from that of the CMB (Bahcall & Ekers, 1969; Deguchi & Watson, 1985).

The Wouthuysen-Field (WF) effect is a radiative mechanism where the resonance scattering (absorption and subsequent re-emission at the same frequency) of Lyα\alpha photons indirectly excites the 21-cm line via transitions involving the n=2n=2 state of hydrogen as an intermediate state (Wouthuysen, 1952; Field, 1958, 1959b). Wouthuysen (1952) suggested, based on a thermodynamical argument, that the central portion of the Lyα\alpha line profile should approach a small slice of the Planck function appropriate to the gas kinetic temperature as neutral hydrogen atoms repeatedly scatter Lyα\alpha photons. Field (1959b) explicitly found an asymptotic solution of the time-dependent radiative transfer (RT) equation for isotropic Doppler scattering and showed that the recoil of the scattering hydrogen atoms with a kinetic temperature of TKT_{{\rm K}} gives rise to a Lyα\alpha spectral profile proportional to the exponential function exp(hν/kBTK)\exp(-h\nu/k_{{\rm B}}T_{{\rm K}}) at the line center. The energy exchange between the Lyα\alpha photons and the atoms due to recoil causes the exponential slope (color temperature) of the Lyα\alpha line profile at the line center to be equal to that of the kinetic temperature. The Planck (or equivalently exponential) function shape at the line center of the Lyα\alpha spectrum yields a redistribution over the two hyperfine levels of the ground state of hydrogen.

The indirect pumping effect is found to be indeed important in exciting the 21-cm line in dilute IGM (Field, 1959a). Urbaniak & Wolfe (1981) found that the mean intensity of Lyα\alpha is strong enough to govern the spin temperature across an entire H I region that adjoins the H II regions photoionized by Lyman continuum radiation from a quasar. The Lyα\alpha recombination followed by ionization via intergalactic UV and X-ray radiation is expected to keep the spin temperature TsT_{{\rm s}} of the very low-density H I at outer parts of disk galaxies well above the CMB temperature of 2.7 K (Watson & Deguchi, 1984; Corbelli & Salpeter, 1993; Irwin, 1995). Keeney et al. (2005) found that the spin temperature and the kinetic temperature of the halo cloud in a nearby spiral galaxy NGC 3067 probed by the sightline to 3C 232 are approximately equal. They showed that the density of the NGC 3067 clouds is smaller than the density required for particle collisions to thermalize the spin temperature. They thereupon argued that the Lyα\alpha photons generated inside the halo cloud drive the spin temperature to the kinetic temperature.

The WF pumping effect is also crucial in understanding the epoch of cosmic reionization. Madau et al. (1997) showed that the non-ionizing UV continuum emitted by early generation stars prior to the epoch of full reionization and then redshifted by the Hubble expansion to the Lyα\alpha wavelength in the absorbing IGM rest frame is capable of heating the IGM to a temperature above that of cosmic background radiation (CBR) via the WF coupling. As a consequence, they concluded that most of the neutral IGM would be available for detection at 21 cm in emission. Kuhlen et al. (2006) used numerical hydrodynamic simulations of early structure formation in the Universe and demonstrated that overdense filaments could shine in emission against the CMB, possibly allowing future radio arrays to probe the distribution of neutral hydrogen before cosmic reionization.

In studying the phases of the ISM, the WF effect is also critical. Classical models of heating and cooling in the ISM predict the existence of two neutral phases, the high-density cold neutral medium (CNM) and the low-density warm neutral medium (WNM), which are in thermal and pressure equilibrium (Field et al., 1969; McKee & Ostriker, 1977; Wolfire et al., 1995, 2003). In the CNM, the 21-cm hyperfine transition of hydrogen is thermalized by collisions. On the other hand, in the WNM, collisions with particles cannot thermalize the 21-cm transition. Hence, if only collisions with particles were taken into account, the spin temperature is expected to be much less than the gas kinetic temperature (Field, 1958; Davies & Cummings, 1975; Deguchi & Watson, 1985; Liszt, 2001).

Liszt (2001) investigated how the Lyα\alpha scattering would be significant in studying the spin temperature of the WNM using analytic approximations for a uniform slab, derived by Neufeld (1990). The spin temperature was found to be thermalized to the kinetic temperature through the WF coupling when the Lyα\alpha radiation field is strong enough in the WNM. He also argued that the density and velocity distribution of the ISM is critical in calculating the Lyα\alpha intensity in the WNM. Kim et al. (2014) used three-dimensional (3D) numerical hydrodynamic simulations of the turbulent, multiphase atomic ISM (Kim et al., 2013) to construct synthetic H I 21-cm emission and absorption lines. They considered the WF effect in the synthetic H I 21-cm line observations, adopting a fixed value of the Lyα\alpha number density. From a stacking analysis of “residual” 21-cm emission line profiles, Murray et al. (2014) detected a widespread WNM component with excitation temperature Ts7200T_{{\rm s}}\sim 7200 K in the Milky Way, concluding that the Lyα\alpha pumping effect is more decisive in the ISM than previously assumed.

Quantifying the requirements for the WF effect is a crucial prerequisite in using the 21-cm line to probe the physical properties of the ISM, CGM, and IGM. Field (1959b) showed that the Lyα\alpha spectral profile would have an exponential slope corresponding to the gas kinetic temperature at the limit of an infinite number of scattering. However, surprisingly, very few studies have assessed how a large number of scatterings is required to thermalize the Lyα\alpha radiation by recoil. Adams (1971) was the first to quantify how many scatterings are needed for the recoil effect to affect the Lyα\alpha RT. He concluded that recoil is negligible unless Lyα\alpha undergoes more than 6×1010\sim 6\times 10^{10} scatterings. Using the large-velocity gradient (LVG) approximation, Deguchi & Watson (1985) found that a Sobolev optical depth of 106\sim 10^{6}, corresponding to a much smaller number of scatterings of 106\sim 10^{6}, is sufficient to bring the color temperature of Lyα\alpha to the kinetic temperature of hydrogen gas. Roy et al. (2009) numerically solved the time-dependent RT equation and showed that the color temperature approaches the kinetic temperature after 104105\sim 10^{4}-10^{5} or more scatterings. Recently, Shaw et al. (2017) claimed that the Lyα\alpha color temperature would rarely trace the kinetic temperature. These studies were based either on the approximate solution of the RT equation (Field, 1959b; Adams, 1971; Deguchi & Watson, 1985; Roy et al., 2009) or the escape probability method (Shaw et al., 2017).

The Monte-Carlo RT simulation would be the most reliable method to quantify the thermalization condition of the Lyα\alpha spectrum due to recoil. Many Monte-Carlo RT codes have been developed to simulate the Lyα\alpha RT (Avery & House, 1968; Ahn et al., 2000; Zheng & Miralda-Escudé, 2002a; Cantalupo et al., 2005; Tasitsiomi, 2006; Dijkstra et al., 2006; Verhamme et al., 2006; Semelin et al., 2007; Laursen & Sommer-Larsen, 2007; Laursen et al., 2009; Pierleoni et al., 2009; Forero-Romero et al., 2011; Yajima et al., 2012; Orsi et al., 2012; Gronke & Dijkstra, 2014; Smith et al., 2015). However, they have mostly focused on calculating the Lyα\alpha spectrum emergent from astrophysical systems; no RT simulation has been performed to investigate the thermalization condition of Lyα\alpha inside the neutral hydrogen medium where Lyα\alpha photons propagate.

The above situation motivated the present study. The primary purpose of this work is, therefore, to gain fundamental insights into the properties of the Lyα\alpha RT in connection with the WF effect, in particular, the conditions where the central portion of the Lyα\alpha spectrum is thermalized to the Planck function with a color temperature corresponding to the gas kinetic temperature. The Monte-Carlo RT method adopted in this paper is described in Section 2. Section 3 presents the test results of our code. The primary results on the WF effect are provided in Section 4. The WF effect in turbulent media is also investigated in the section. We apply the code to two ISM models that describe the ISM properties in the solar neighborhood to examine the significance of the WF effect in Section 5. A discussion is presented in Section 6, and Section 7 summarizes our main findings. In Appendixes, we present: (a) a novel algorithm to find a random velocity component of hydrogen atoms scattering Lyα\alpha photons, (b) newly-derived analytic solutions in slab and spherical geometries, (c) a brief discussion on the Lyα\alpha spectrum absorbed by dust grains, and (d) the physical principle of the WF effect.

2. Monte-Carlo Radiative Transfer Methods

Basic algorithms of Lyα\alpha transport in our code “LaRT” are mostly based on the previous Monte-Carlo RT algorithms. However, many new algorithms were developed for LaRT. The present version of LaRT uses a 3D cartesian grid. The physical quantities to describe a 3D medium are the density and temperature of neutral hydrogen gas, the dust density (or gas-to-dust ratio), and the velocity field. LaRT is also capable of dealing with the fine structure of the hydrogen atom.

LaRT is written in modern Fortran 90/95/2003 and uses the Message Passing Interface for parallel computation. We found that assigning an equal number of photon packages to each processor yields severe load imbalance because each photon packet experiences a significantly different number of scattering. Hence, the master-slave (manager-worker) algorithm was used to implement dynamic load balancing (Gropp et al., 2014). In our previous studies for the dust RT (e.g., Seon, 2015; Seon & Draine, 2016), the Keep It Simple and Stupid (KISS) random number generator has been used (Marsaglia & Zaman, 1993). In LaRT, the Mersenne Twister Random Number Generator (MT19937), which is much faster and has a more extended period, is used (Matsumoto & Nishimura, 1998).

Section 2.1 briefly outlines the general methodology, ignoring the dust effect and the fine structure of the hydrogen atom, implemented in LaRT. The methods to take into account the dust effect and the fine structure are described in Sections 2.2 and 2.3, respectively. We clarify the necessary conditions for the WF effect in Section 2.4. In Sections 2.5 and 2.6, we describe the methods to calculate the Lyα\alpha spectrum inside a medium and the number of Lyα\alpha scatterings that a hydrogen atom undergoes, which are necessary to investigate the WF effect.

2.1. Lyα\alpha Radiative Transfer

In the code, we generate photon packets and track each of them in physical and frequency spaces. The Monte-Carlo method for the scattering of Lyα\alpha photons by neutral hydrogen atoms proceeds through the following steps. First, we generate a position 𝐫\mathbf{r}, a frequency ν\nu, and a propagation direction 𝐤i\mathbf{k}_{i} of the photon packet. The initial location and frequency of the photon are drawn from the source’s spatial distribution function and an intrinsic frequency distribution under consideration, respectively. The initial propagation direction is generated to be isotropic. Second, the optical depth τ\tau that the photon will travel through before it interacts with a hydrogen atom is randomly drawn from the probability density function P(τ)=eτP(\tau)=e^{-\tau} as follows:

τ=lnξ,\tau=-\ln\xi, (1)

where ξ\xi is drawn from a uniform random distribution. The physical distance is then calculated by inverting the integral of optical depth over the photon pathlength, defined in Equation (2). Third, we generate the velocity vector of the atom that scatters the photon, as described later. Fourth, we draw the scattering angle of the photon from an appropriate scattering phase function and calculate the new propagation direction 𝐤f\mathbf{k}_{f}. The velocity of the scattering atom and the old and new direction vectors of the photon determine a new frequency following the energy and momentum conservation laws. The process of propagation and scattering is repeated until the photon escapes the system. For the present study, we used a number of photon packets of 10510^{5} up to 101010^{10}.

The optical depth τν(s)\tau_{\nu}(s) of a photon with frequency ν\nu along a path length ss is given by

τν(s)=0sn(v)σν𝑑v𝑑l,\tau_{\nu}(s)=\int_{0}^{s}\int_{-\infty}^{\infty}n(v_{\parallel})\sigma_{\nu}dv_{\parallel}dl, (2)

where n(v)n(v_{\parallel}) is the number density of neutral hydrogen atom with velocity component vv_{\parallel} parallel to the photon’s traveling path and σν\sigma_{\nu} is the scattering cross-section. It is convenient to express frequency in terms of the relative frequency, x(ννα)/ΔνDx\equiv(\nu-\nu_{\alpha})/\Delta\nu_{{\rm D}}, where να=2.466×1015\nu_{\alpha}=2.466\times 10^{15} Hz is the central frequency of Lyα\alpha, ΔνD=να(vth/c)=1.057×1011(TK/104K)1/2\Delta\nu_{{\rm D}}=\nu_{\alpha}(v_{{\rm th}}/c)=1.057\times 10^{11}(T_{{\rm K}}/10^{4}{\rm K})^{1/2} Hz is the thermal, Doppler frequency width, and vth=(2kBTK/mH)1/2=12.85(TK/104K)1/2v_{{\rm th}}=(2k_{{\rm B}}T_{{\rm K}}/m_{{\rm H}})^{1/2}=12.85(T_{{\rm K}}/10^{4}{\rm K})^{1/2} km s-1 is the thermal velocity dispersion, multiplied by 2\sqrt{2}, of atomic hydrogen gas with a kinetic temperature TKT_{{\rm K}}. Here, kBk_{{\rm B}} is the Boltzmann constant and mHm_{{\rm H}} is the hydrogen atomic mass. For the ray tracing of photon packets in evaluating the optical depth integral of Equation (2), we use the fast voxel traversal algorithm (Amanatides & Woo, 1987), which is used in our dust radiative transfer code MoCafe (Seon & Witt, 2012; Seon, 2015; Seon & Draine, 2016).

The scattering cross-section of a Lyα\alpha photon in the rest frame of a hydrogen atom is given by

σνrest=fαπe2mecϕνrest=fαπe2mecΓ/4π2(ννα)2+(Γ/4π)2,\sigma_{\nu}^{{\rm rest}}=f_{\alpha}\frac{\pi e^{2}}{m_{e}c}\phi_{\nu}^{{\rm rest}}=f_{\alpha}\frac{\pi e^{2}}{m_{e}c}\frac{\Gamma/4\pi^{2}}{(\nu-\nu_{\alpha})^{2}+(\Gamma/4\pi)^{2}}, (3)

where fα=0.4162f_{\alpha}=0.4162 is the oscillator strength of Lyα\alpha, and Γ(=Aα)\Gamma(=A_{\alpha}) is the damping constant. The resonance profile111The normalized profile of the resonance cross-section as a function of frequency is often referred to as the line profile. In this paper, we call it “the resonance profile” to distinguish it from the line profile arising as a result of the RT process. ϕνrest\phi_{\nu}^{{\rm rest}} is normalized such that 0ϕνrest𝑑ν=1\int_{0}^{\infty}\phi_{\nu}^{{\rm rest}}d\nu=1. The Einstein A coefficient of the Lyα\alpha transition between the n=2n=2 and n=1n=1 levels of a hydrogen atom is Aα=6.265×108A_{\alpha}=6.265\times 10^{8} s-1. Integrating over the Maxwellian velocity distribution of hydrogen atoms, the scattering cross-section in the laboratory frame is found to be

σν=χ0ΔνDH(x,a)π=χ0ΔνDϕx,\sigma_{\nu}=\frac{\chi_{0}}{\Delta\nu_{{\rm D}}}\frac{H(x,a)}{\sqrt{\pi}}=\frac{\chi_{0}}{\Delta\nu_{{\rm D}}}\phi_{x}, (4)

where χ0fαπe2/mec=(3c2/8πνα2)Aα=1.105×102\chi_{0}\equiv f_{\alpha}\pi e^{2}/m_{e}c=\left(3c^{2}/8\pi\nu_{\alpha}^{2}\right)A_{\alpha}=1.105\times 10^{-2} cm2 Hz, ϕxH(x,a)/π\phi_{x}\equiv H(x,a)/\sqrt{\pi}, and

H(x,a)=aπeu2(xu)2+a2𝑑uH(x,a)=\frac{a}{\pi}\int_{-\infty}^{\infty}\frac{e^{-u^{2}}}{(x-u)^{2}+a^{2}}du (5)

is the Voigt-Hjerting function. Here, a=Γ/(4πΔνD)=4.717×104(TK/104K)1/2a=\Gamma/(4\pi\Delta\nu_{{\rm D}})=4.717\times 10^{-4}\ (T_{{\rm K}}/10^{4}{\rm K})^{-1/2} is the natural linewidth relative to the thermal frequency width. The atomic velocity v is normalized by the thermal speed, i.e., 𝐮=v/vth\mathbf{u}=\textbf{v}/v_{{\rm th}}. We also note that the resonance profile ϕx\phi_{x} is normalized, e.g., ϕx𝑑x=1\int_{-\infty}^{\infty}\phi_{x}dx=1. In this paper, the monochromatic optical thickness τ0(χ0/ΔνD)NHIϕx(0)(χ0/πΔνD)NHI\tau_{0}\equiv(\chi_{0}/\Delta\nu_{{\rm D}})N_{{\rm HI}}\phi_{x}(0)\simeq(\chi_{0}/\sqrt{\pi}\Delta\nu_{{\rm D}})N_{{\rm HI}} measured at the line center is used to represent the total optical thickness of media. Here, NHIN_{{\rm HI}} is the column density of neutral hydrogen atoms. The optical depth has a temperature dependence of

τ0\displaystyle\tau_{0} \displaystyle\simeq 5.90×106(NHI/1020cm2)(TK/104K)1/2.\displaystyle 5.90\times 10^{6}\left(N_{{\rm HI}}/10^{20}\,{\rm cm}^{-2}\right)\left(T_{{\rm K}}/10^{4}\,K\right)^{-1/2}. (6)

The monochromatic optical thickness (τ0\tau_{0}) is related to the optical thickness integrated over the relative frequency (ττx𝑑x\tau_{*}\equiv\int_{-\infty}^{\infty}\tau_{x}dx), which is used in Neufeld (1990), by

τ0τπ(12aπ+a2).\tau_{0}\simeq\frac{\tau_{*}}{\sqrt{\pi}}\left(1-\frac{2a}{\sqrt{\pi}}+a^{2}-\cdots\right). (7)

We need an accurate yet fast algorithm to calculate the Voigt function because the function is repeatedly estimated in simulations. An approximate formula suggested by N. Gnedin, as described in Tasitsiomi (2006) and Laursen et al. (2009), is simple and thus might be appropriate in calculating the Lyα\alpha spectrum emerging from the hydrogen gas. However, we found that the approximation yields a systematic, spurious wavy feature at the line center of the Lyα\alpha spectrum calculated inside the medium. We, instead, adopt a slightly modified version of the function “voigt_king” in the VPFIT program (Carswell & Webb, 2014). This routine was found to be accurate and as fast as or even quicker than that given in Tasitsiomi (2006).

The velocity of the scattering atom is decomposed of the parallel (uu_{\parallel}) and perpendicular (𝐮\mathbf{u}_{\perp}) components with respect to the photon’s propagation direction 𝐤\mathbf{k}. The two (mutually orthogonal) perpendicular components 𝐮\mathbf{u}_{\perp} are independent of the photon frequency and thus are drawn from the following two-dimensional (2D) Gaussian distribution with zero mean and standard deviation 1/21/\sqrt{2}:

f(𝐮)=1πe|𝐮|2.f(\mathbf{u}_{\perp})=\frac{1}{\sqrt{\pi}}e^{-|\mathbf{u}_{\perp}|^{2}}. (8)

The two random velocity components are then generated as

u,1\displaystyle u_{\perp,1} =\displaystyle= (lnξ1)1/2cos(2πξ2)\displaystyle(-\ln\xi_{1})^{1/2}\cos(2\pi\xi_{2})
u,2\displaystyle u_{\perp,2} =\displaystyle= (lnξ1)1/2sin(2πξ2),\displaystyle(-\ln\xi_{1})^{1/2}\sin(2\pi\xi_{2}), (9)

where ξ1\xi_{1} and ξ2\xi_{2} are two independent random variates between 0 and 1. When a photon is in the core of the Voigt profile, it is trapped by resonance scattering in a small region, where it started, and thus cannot travel long distance until it is scattered off to the wing of the profile. Therefore, to speed up the calculation, one can use an acceleration scheme, known as “core-skipping,” which was first suggested by Ahn et al. (2002). The scheme artificially pushes the photon into the wing by using a truncated Gaussian distribution. This is achieved by generating the perpendicular components as

u,1\displaystyle u_{\perp,1} =\displaystyle= (xcrit2lnξ1)1/2cos(2πξ2)\displaystyle(x_{{\rm crit}}^{2}-\ln\xi_{1})^{1/2}\cos(2\pi\xi_{2})
u,2\displaystyle u_{\perp,2} =\displaystyle= (xcrit2lnξ1)1/2sin(2πξ2),\displaystyle(x_{{\rm crit}}^{2}-\ln\xi_{1})^{1/2}\sin(2\pi\xi_{2}), (10)

in which the critical frequency xcritx_{{\rm crit}} can be chosen as described in Dijkstra et al. (2006), Laursen et al. (2009), or Smith et al. (2017). This acceleration scheme is also implemented in LaRT. However, the scheme is not used for the present study because it gives rise to a serious underestimation of the number of scatterings. As we shall show later, measuring the number of scatterings is the most crucial point to study the WF effect. We also attempted to apply various bias techniques, such as the composite bias method (Baes et al., 2016) or others (Juvela, 2005), which have been developed for dust and nuclear physics, to improve the computation speed. However, none of the methods was found to be appropriate for the Lyα\alpha RT. This failure is because the number of scatterings in the Lyα\alpha RT is exceptionally high compared to that for dust.

The parallel component uu_{\parallel} of the scattering atom depends on the incident photon frequency and thus should be drawn from the following distribution function:

f(u|x,a)=aπH(x,a)eu2a2+(xu)2.f(u_{\parallel}|x,a)=\frac{a}{\pi H(x,a)}\frac{e^{-u_{\parallel}^{2}}}{a^{2}+(x-u_{\parallel})^{2}}. (11)

Note that this probability distribution function depends on not only the photon frequency but also the gas temperature through aa. The most commonly adopted algorithm is based on the acceptance/rejection method of Zheng & Miralda-Escudé (2002a). Detailed descriptions are given in Semelin et al. (2007) and Laursen et al. (2009). The algorithm is elaborated by Smith et al. (2017). For LaRT, we developed a novel method to generate random uu_{\parallel} using the ratio-of-uniforms method (Kinderman & Monahan, 1977), as described in Appendix 28. Our algorithm is efficient in the sense that fewer random numbers are rejected, especially for high xx values and high temperatures TK103T_{{\rm K}}\gtrsim 10^{3} K.

After the velocity components of the scattering atom is selected, the scattering direction is generated assuming the Rayleigh phase function P(θ)1+cos2θP(\theta)\propto 1+\cos^{2}\theta, where θ\theta is the angle between the incident direction 𝐤i\mathbf{k}_{i} and the outgoing direction 𝐤f\mathbf{k}_{f}. The random, scattering angle θ\theta is determined as described in Seon (2006):

μ\displaystyle\mu =\displaystyle= [2(2ξ1)+4(2ξ1)2+1]1/3\displaystyle\left[2(2\xi-1)+\sqrt{4(2\xi-1)^{2}+1}\right]^{1/3} (12)
[2(2ξ1)+4(2ξ1)2+1]1/3,\displaystyle-\left[2(2\xi-1)+\sqrt{4(2\xi-1)^{2}+1}\right]^{-1/3},

where μ=cosθ\mu=\cos\theta and ξ\xi is a uniform random variate. The azimuthal angle ϕ\phi is generated by ϕ=2πξ\phi=2\pi\xi^{\prime} using another random variate ξ\xi^{\prime}. More generally, the scattering phase function is known to have a form of P(θ)1+(R/Q)cos2θP(\theta)\propto 1+(R/Q)\cos^{2}\theta, in which R/QR/Q is the degree of polarization for 90 scattering and depends on the incident frequency (Stenflo, 1980). We will ignore the dependence of the scattering angle distribution on the photon frequency, which is not significant for the present purpose. The algorithm to draw the scattering angle θ\theta for this phase function will be described in a forthcoming paper, in which the Lyα\alpha polarization is investigated.

We assumed, unless otherwise stated, that photons are injected at a monochromatic frequency of x=0x=0. If the initial photon packets are supposed to have the Voigt profile222The emission line profile is often regarded to be a Gaussian function. However, the ‘intrinsic’ profile of an emission line is a convolution of Lorentzian and Gaussian functions, representing the natural line profile from an undriven harmonically bound atom and the effect due to the gas thermal motion, respectively (Rybicki & Lightman, 1985; Hubeny & Mihalas, 2014). The ‘intrinsic’ absorption profile has the same shape as the ‘intrinsic’ emission profile by the principle of detailed balance (Hubeny & Mihalas, 2014)., its frequency is obtained by adding a random number for a Lorentzian distribution and a random number for a Gaussian distribution:

x=atan[π(ξ12)]+ξG2,x=a\tan\left[\pi\left(\xi-\frac{1}{2}\right)\right]+\frac{\xi_{{\rm G}}}{\sqrt{2}}, (13)

where ξ\xi is a uniform random number between 0 and 1, and ξG\xi_{{\rm G}} is a random number drawn from a Gaussian distribution with zero mean and unit variance. Here, we note that ξG/2\xi_{{\rm G}}/\sqrt{2} can be obtained, for instance, using Equation (9).

In the code, each cell has a bulk velocity (𝐮bulk=vbulk/vth\mathbf{u}^{{\rm bulk}}=\textbf{v}^{{\rm bulk}}/v_{{\rm th}}), which is expressed in units of the thermal speed that depends on the gas temperature. When a photon packet traverses from cell to cell (for instance, ii to jj), the gas temperature TKT_{{\rm K}} and the bulk velocity 𝐮bulk\mathbf{u}^{{\rm bulk}} may vary. The relative frequency should then be changed to

xj=(xi+𝐤𝐮ibulk)(ΔνDi/ΔνDj)1/2𝐤𝐮jbulk,x_{j}=\left(x_{i}+\mathbf{k}\cdot\mathbf{u}_{i}^{{\rm bulk}}\right)\left(\Delta\nu_{{\rm D}}^{i}/\Delta\nu_{{\rm D}}^{j}\right)^{1/2}-\mathbf{k}\cdot\mathbf{u}_{j}^{{\rm bulk}}, (14)

as the photon crosses the cell ii to jj, where 𝐤\mathbf{k} is the photon’s propagation vector with unit length.

After a scattering event, the incident photon frequency xix_{i} is changed to the outgoing frequency xfx_{f} given by

xf=xiu+𝐤f𝐮atomg(1𝐤f𝐤i),x_{f}=x_{i}-u_{\parallel}+\mathbf{k}_{f}\cdot\mathbf{u}_{{\rm atom}}-g_{*}\left(1-\mathbf{k}_{f}\cdot\mathbf{k}_{i}\right), (15)

where g=(hνα2/mHc2)/ΔνD=2.536×104(TK/104K)1/20.54ag_{*}=(h\nu_{\alpha}^{2}/m_{{\rm H}}c^{2})/\Delta\nu_{{\rm D}}=2.536\times 10^{-4}(T_{{\rm K}}/10^{4}{\rm K})^{-1/2}\simeq 0.54a is the recoil parameter. Here, hh is the Planck constant. The recoil parameter might be negligible for applications in which only the spectra emergent from galaxies are of interest. However, as we will show, it is this small effect that causes the WF effect.

2.2. Dust Scattering and Absorption

In addition to being scattered by neutral hydrogen atoms, Lyα\alpha photons also interact with dust grains; they can either be scattered or destroyed by dust. When the interaction with dust is included, the total optical depth along a pathlength ss is the sum of optical depths due to neutral hydrogen and dust, τ=τH+τdust\tau=\tau_{{\rm H}}+\tau_{{\rm dust}}. The absorption optical depth τabs\tau_{{\rm abs}} by dust is defined to be τabs=(1η)τdust\tau_{{\rm abs}}=(1-\eta)\tau_{{\rm dust}} for the dust extinction optical depth τdust\tau_{{\rm dust}} and the scattering albedo η\eta.

The probability of being interacted with dust is given by

𝒫dust=ndustσdustnHσH+ndustσdust.\mathcal{P}_{{\rm dust}}=\frac{n_{{\rm dust}}\sigma_{{\rm dust}}}{n_{{\rm H}}\sigma_{{\rm H}}+n_{{\rm dust}}\sigma_{{\rm dust}}}. (16)

If a random number ξ\xi is generated to be less than 𝒫dust\mathcal{P}_{{\rm dust}}, then the photon interacts with a dust grain; otherwise, it is scattered by a hydrogen atom.

To take into account the interaction with dust grains, we implemented two different methods in LaRT. The first method is to use the weight of photon packets, where the weight is initially set to unity (w=1w=1) and reduced by a factor of albedo (w=ηww^{\prime}=\eta w) as the photon packet undergoes dust scattering. The second method is to compare a uniform random number with the dust albedo and then assume the photon packet will be absorbed and destroyed by a dust grain if the random number is larger than the albedo.

When the photon is scattered by dust, the scattering angle θ\theta is found using the inversion method to follow the Henyey-Greenstein phase function for a given asymmetry factor gg (e.g., Witt, 1977):

μ=1+g22g12g(1g21g+2gξ)2,\mu=\frac{1+g^{2}}{2g}-\frac{1}{2g}\left(\frac{1-g^{2}}{1-g+2g\xi}\right)^{2}, (17)

where ξ\xi is drawn from a uniform random distribution.

We adopt the Milky-Way dust model of Weingartner & Draine (2001) and Draine (2003). The dust extinction cross-section per hydrogen atom is ndustσdust/nH=1.61×1021n_{{\rm dust}}\sigma_{{\rm dust}}/n_{{\rm H}}=1.61\times 10^{-21} cm2/H at Lyα\alpha. The asymmetry factor and albedo at Lyα\alpha are given by g=0.676g=0.676 and η=0.325\eta=0.325, respectively. The dust parameters (ndustσdust/nH,η,g)(n_{{\rm dust}}\sigma_{{\rm dust}}/n_{{\rm H}},\ \eta,\ g) for the other dust types are (5.32×10225.32\times 10^{-22} cm2/H, 0.255, 0.648) for the Large Magellanic Cloud, and (3.76×10223.76\times 10^{-22} cm2/H, 0.330, 0.591) for the Small Magellanic Cloud.

2.3. Fine Structure

The hyperfine structure of the n=2n=2 state should be considered in order to derive the formula for the spin temperature due to the indirect Lyα\alpha pumping of the hyperfine levels in the ground state (the lower level “0” and the upper level “1” in Figure 33). However, one can ignore the hyperfine splitting when calculating only the Lyα\alpha spectrum through the RT process because the Lyα\alpha line profile is much broader than the splitting. The fine structure of the n=2n=2 state also can be ignored in most of the cases presented in this paper. At low gas temperatures and low optical thicknesses, however, the fine-structure splitting of the n=2n=2 state must be taken into account; in these circumstances, the line width is comparable to or smaller than the frequency gap between the fine-structure levels of the excited state. The level splitting of 10.8 GHz in the excited state is equivalent to a Doppler shift of 1.341.34 km s-1; the Doppler width (ΔνD1.28\Delta\nu_{{\rm D}}\lesssim 1.28 km s-1) for the gas of T102T\lesssim 10^{2} K is thus smaller than the fine-structure splitting.

We use the term 1/2\nicefrac{{1}}{{2}} to denote the transition 1S1/222P1/2o21\,{}^{2}{\rm S}_{1/2}\leftrightarrow 2\,{}^{2}{\rm P}_{1/2}^{{\rm o}} and 3/2\nicefrac{{3}}{{2}} the transition 1S1/222P3/2o21\,{}^{2}{\rm S}_{1/2}\leftrightarrow 2\,{}^{2}{\rm P}_{3/2}^{{\rm o}}. Then, the cross-section is given by

σνrest\displaystyle\sigma_{\nu}^{{\rm rest}} =\displaystyle= χ0[(1/3)Γ/4π2(νν1/2)2+(Γ/4π)2+(2/3)Γ/4π2(νν3/2)2+(Γ/4π)2]\displaystyle\chi_{0}\left[\frac{(1/3)\Gamma/4\pi^{2}}{(\nu-\nu_{\nicefrac{{1}}{{2}}})^{2}+(\Gamma/4\pi)^{2}}+\frac{(2/3)\Gamma/4\pi^{2}}{(\nu-\nu_{\nicefrac{{3}}{{2}}})^{2}+(\Gamma/4\pi)^{2}}\right] (18)

in the rest frame of a hydrogen atom (e.g., Ahn & Lee, 2015). The cross-section is a combination of two Lorentzian functions with weights of 1:2. In a gas medium having a Maxwellian velocity distribution at temperature TT, the scattering cross-section is the sum of two Voigt functions:

σν=χ0ΔνD[13ϕ(x1/2)+23ϕ(x3/2)],\sigma_{\nu}=\frac{\chi_{0}}{\Delta\nu_{{\rm D}}}\left[\frac{1}{3}\phi(x_{\nicefrac{{1}}{{2}}})+\frac{2}{3}\phi(x_{\nicefrac{{3}}{{2}}})\right], (19)

where x1/2(νν1/2)/ΔνDx_{\nicefrac{{1}}{{2}}}\equiv(\nu-\nu_{\nicefrac{{1}}{{2}}})/\Delta\nu_{{\rm D}} and x3/2(νν3/2)/ΔνDx_{\nicefrac{{3}}{{2}}}\equiv(\nu-\nu_{\nicefrac{{3}}{{2}}})/\Delta\nu_{{\rm D}}. We also define x[ν(ν1/2+ν3/2)/2]/ΔνD=(x1/2+x3/2)/2x\equiv\left[\nu-(\nu_{\nicefrac{{1}}{{2}}}+\nu_{\nicefrac{{3}}{{2}}})/2\right]/\Delta\nu_{{\rm D}}=(x_{\nicefrac{{1}}{{2}}}+x_{\nicefrac{{3}}{{2}}})/2. The parallel component uu_{\parallel} of scattering atom is drawn from the following distribution function:

fFS(u|x)\displaystyle f_{{\rm FS}}(u_{\parallel}|x) =\displaystyle= 𝒫1/2f(u|x1/2)+(1𝒫1/2)f(u|x3/2),\displaystyle\mathcal{P}_{\nicefrac{{1}}{{2}}}f(u_{\parallel}|x_{\nicefrac{{1}}{{2}}})+(1-\mathcal{P}_{\nicefrac{{1}}{{2}}})f(u_{\parallel}|x_{\nicefrac{{3}}{{2}}}), (20)

where

𝒫1/2H(x1/2,a)H(x1/2,a)+2H(x3/2,a).\mathcal{P}_{\nicefrac{{1}}{{2}}}\equiv\frac{H(x_{\nicefrac{{1}}{{2}}},a)}{H(x_{\nicefrac{{1}}{{2}}},a)+2H(x_{\nicefrac{{3}}{{2}}},a)}. (21)

The random variates for this distribution are obtained using the composition method. If a uniform random number ξ\xi is selected to be smaller than the probability 𝒫1/2\mathcal{P}_{\nicefrac{{1}}{{2}}}, the photon is scattered by the 1/2\nicefrac{{1}}{{2}} transition; otherwise, by the 3/2\nicefrac{{3}}{{2}} transition. Then, either f(u|x1/2)f(u_{\parallel}|x_{\nicefrac{{1}}{{2}}}) or f(u|x3/2)f(u_{\parallel}|x_{\nicefrac{{3}}{{2}}}), depending on the transition, determines the parallel velocity component. The initial photon frequency is chosen to be either ν=ν1/2\nu=\nu_{\nicefrac{{1}}{{2}}} or ν=ν3/2\nu=\nu_{\nicefrac{{3}}{{2}}} with a probability ratio of 1:2. We note, in an optically thick medium, that the line ratio becomes equal to 1:1 as a result of the resonance-line scattering (e.g., Long et al., 1992).

In this paper, we present the results for TK=10T_{{\rm K}}=10 K and 10410^{4} K, unless otherwise stated. We consider the fine-structure splitting only when calculating the Lyα\alpha spectrum inside the medium of TK=10T_{{\rm K}}=10 K and occasionally of TK=102T_{{\rm K}}=10^{2} K in Section 4; the fine-structure splitting is negligible at TK103T_{{\rm K}}\gtrsim 10^{3} K. In Sections 3 and 5, the splitting effect is not taken into account. A more complete description of how to deal with the fine-structure splitting in Lyα\alpha RT will be given in a separate paper.

2.4. Requirements for the WF effect

Field (1958) found that the 21-cm spin temperature is a weighted mean of the three temperatures TRT_{{\rm R}}, TKT_{{\rm K}}, and TαT_{\alpha}, as follows:

Ts\displaystyle T_{{\rm s}} =\displaystyle= TR+ycTK+yαTα1+yc+yα,\displaystyle\frac{T_{{\rm R}}+y_{{\rm c}}T_{{\rm K}}+y_{\alpha}T_{\alpha}}{1+y_{{\rm c}}+y_{\alpha}}, (22)

where TRT_{{\rm R}} is the brightness temperature of the background radio radiation at 2121 cm (e.g., TR2.7T_{{\rm R}}\sim 2.7 K for the CMB), and TKT_{{\rm K}} is the kinetic temperature of hydrogen gas. The color temperature TαT_{\alpha} of Lyα\alpha is defined by the exponential slope of the mean intensity spectrum at the line center, as follows:

Jν=Jν(να)exp[h(ννα)kBTα].J_{\nu}=J_{\nu}(\nu_{\alpha})\exp\left[-\frac{h\left(\nu-\nu_{\alpha}\right)}{k_{{\rm B}}T_{\alpha}}\right]. (23)

The weighting factors in Equation (22) are

yc\displaystyle y_{{\rm c}} =\displaystyle= TTKP10cA10andyα=427TTαPαA10,\displaystyle\frac{T_{*}}{T_{{\rm K}}}\frac{P_{10}^{{\rm c}}}{A_{10}}\ \ {\rm and}\ \ y_{\alpha}=\frac{4}{27}\frac{T_{*}}{T_{\alpha}}\frac{P_{\alpha}}{A_{10}}, (24)

where Thν10/kB=0.0681T_{*}\equiv h\nu_{10}/k_{{\rm B}}=0.0681 K, P10cP_{10}^{{\rm c}} is the downward transition rate (per sec) from the hyperfine level “1” to “0” due to particle collisions (see Figure 33 for the level definition), and PαP_{\alpha} is the scattering rate (the number of Lyα\alpha scatterings per unit time for a single hydrogen atom). Here, A10A_{10} is the Einstein coefficient for the spontaneous emission of 21-cm line. In Appendix 33, we describe the above equations in more detail.

From the above Equations (22)-(24), we can identify two necessary conditions for the WF effect. First, the central part of the Lyα\alpha spectrum in the medium should be the exponential shape of Equation (23) with a color temperature that is equal to the kinetic temperature (e.g., Tα=TKT_{\alpha}=T_{{\rm K}}). Second, the weighting factor for the Lyα\alpha pumping should be much higher than that of the background radiation (e.g., yα1y_{\alpha}\gg 1). We note that the first condition requires individual Lyα\alpha photons to experience a sufficient number of resonance-line scatterings to thermalize the spectrum via atomic recoil. On the other hand, the second condition is primarily a matter of the amount of Lyα\alpha photons that are supplied to the medium. The amount of dust grains and the kinematics of gas are also critical factors that affect both requirements.

2.5. Spectrum of the Lyα\alpha Radiation Field

To study the condition for the first requirement of the WF effect, we first need to examine the spectral shape of Lyα\alpha in the medium. To calculate the spectrum of the Lyα\alpha radiation field in a volume element of the medium, we use the method that is described in Lucy (1999) and has been extensively adopted to estimate the absorbed energy of stellar radiation by dust grains and the re-emitted infrared emission in dust RT simulations (e.g., Misselt et al., 2001; Baes et al., 2011).

Let δi\delta\ell_{i} denote the pathlength between successive events and Δt\Delta t the total duration of the “sequential” simulation. Here, events indicate not only scatterings but also crossings across boundaries between volume elements. The energy carried by a single photon packet is denoted by ϵ=LΔt/Npacket\epsilon_{*}=L_{*}\Delta t/N_{{\rm packet}}, where LL_{*} and NpacketN_{{\rm packet}} are the total luminosity and the total number of photon packets used in the simulation, respectively. The segment of the packet’s trajectory between two successive events contributes ϵδti/Δt\epsilon_{*}\delta t_{i}/\Delta t to the time-averaged energy content of a volume element, where δti=δi/c\delta t_{i}=\delta\ell_{i}/c is the traveling time between events. Here, we note that Δt=k=1Npacketiδti\Delta t=\sum_{k=1}^{N_{{\rm packet}}}\sum_{i}\delta t_{i}. The mean intensity JνJ_{\nu} of the radiation field, expressed in the unit of photon number, in the frequency interval (ν,ν+Δν)(\nu,\ \nu+\Delta\nu) can be expressed in terms of the energy density uνu_{\nu} by hναJν=(c/4π)uνh\nu_{\alpha}J_{\nu}=(c/4\pi)u_{\nu}.333In this paper, JνJ_{\nu} is used to denote the mean intensity in terms of photon number. The mean intensity expressed in terms of energy is denoted by IνI_{\nu}. They are related by Iν=hναJνI_{\nu}=h\nu_{\alpha}J_{\nu}. The mean intensity is then given by summing all energy segments that contribute to the volume element:

hναJνΔν\displaystyle h\nu_{\alpha}J_{\nu}\Delta\nu =\displaystyle= c4π1ΔViϵδtiΔt,\displaystyle\frac{c}{4\pi}\frac{1}{\Delta V}\sum_{i}\epsilon_{*}\frac{\delta t_{i}}{\Delta t},
Jν\displaystyle J_{\nu} =\displaystyle= 14πΔVΔνL/hναNpacketiδi.\displaystyle\frac{1}{4\pi\Delta V\Delta\nu}\frac{L_{*}/h\nu_{\alpha}}{N_{{\rm packet}}}\sum_{i}\delta\ell_{i}. (25)

Thus, the spectrum of the radiation field JνJ_{\nu} can be obtained by adding the pathlengths of the packet’s trajectory between events. In the case of an infinite, plane-parallel slab, the luminosity should be replaced by L=ΔAL_{*}=\mathcal{L}_{*}\Delta A, where \mathcal{L_{*}} is the photon energy per unit area per unit time, and ΔA\Delta A is the area of the simulation box in the XYXY plane. The volume element is ΔV=HΔA\Delta V=H\Delta A for the height HH of the slab. We also note that all intensity spectra JxJνΔνDJ_{x}\equiv J_{\nu}\Delta\nu_{{\rm D}} in this paper (including the emergent spectra) were angle-averaged and normalized for a unit production-rate of photons; thus, they should be multiplied by the Lyα\alpha production rate (L/hναL_{*}/h\nu_{\alpha} or /hνα\mathcal{L}_{*}/h\nu_{\alpha}).

In Appendixes B and C, we derive the analytic formulae (Equations (B5) and (C3)) for the mean intensity at an arbitrary position in a static, infinite slab and sphere, respectively. The formulae are used to validate the Monte-Carlo simulation. Neufeld (1990) and Dijkstra et al. (2006) provided series solutions for the mean intensity in the slab and sphere, respectively. Our formulae are based on the series solutions, but we offer the solutions in more compact, analytic forms.

2.6. Scattering Rate

We also calculate the scattering rate PαP_{\alpha}, which is defined as the number of scatterings per unit time that a single atom undergoes. By the definition of the cross-section, the scattering rate can be calculated by integrating the scattering cross-section multiplied by the mean intensity:

Pα=4πJνσν𝑑ν.P_{\alpha}=4\pi\int J_{\nu}\sigma_{\nu}d\nu. (26)

We note that, at the central part of the Lyα\alpha line, the spectrum can be approximated to be constant as the cross-section peaks sharply. We then obtain

Pα\displaystyle P_{\alpha} \displaystyle\simeq 4πJν(να)σν𝑑ν=4πχ0Jν(να)\displaystyle 4\pi J_{\nu}(\nu_{\alpha})\int\sigma_{\nu}d\nu=4\pi\chi_{0}J_{\nu}(\nu_{\alpha}) (27)
=\displaystyle= 4πχ0ΔνDJx(0)\displaystyle 4\pi\frac{\chi_{0}}{\Delta\nu_{{\rm D}}}J_{x}(0)
=\displaystyle= 1.315×1012(TK/104K)1/2Jx(0)\displaystyle 1.315\times 10^{-12}\left(T_{{\rm K}}/10^{4}\,{\rm K}\right)^{-1/2}J_{x}(0)

In Section 4, we show that the radiation field spectrum is indeed constant if the atomic recoil effect is ignored or exponential if the recoil effect is considered, over a wide range of frequency. The exponential function can be approximated to a linear function at the central portion of the spectrum and thus the integral in Equation (27) should give rise to precisely the same result as for a constant spectrum. In Appendixes B and C, we provide useful formulae (Equations (B) and (C)) for the mean intensity at the line center Jx(0)J_{x}(0) in an any position of a slab and sphere to calculate the scattering rate PαP_{\alpha}.

In the Monte-Carlo RT, it is more natural to calculate the scattering rate by directly counting the number of scattering events in each volume element and normalize it by the number of atoms. The total number of photons injected during Δt\Delta t is LΔt/hναL_{*}\Delta t/h\nu_{\alpha}. The scattering rate can then be obtained by

Pα\displaystyle P_{\alpha} =\displaystyle= 1nHΔVL/hναNpacketNscattcell,\displaystyle\frac{1}{n_{{\rm H}}\Delta V}\frac{L_{*}/h\nu_{\alpha}}{N_{{\rm packet}}}N_{{\rm scatt}}^{{\rm cell}}, (28)

where NscattcellN_{{\rm scatt}}^{{\rm cell}} is the number of scattering events that occur in the volume element ΔV\Delta V. Again, for the infinite slab geometry, the luminosity and volume element should be replaced by L=ΔAL_{*}=\mathcal{L}_{*}\Delta A and ΔV=HΔA\Delta V=H\Delta A, respectively.

The third estimator of PαP_{\alpha} can be obtained by combing Equations (25) and (26), as follows:

Pα=1nHΔVL/hναNpacketi,νδτi,ν,P_{\alpha}=\frac{1}{n_{{\rm H}}\Delta V}\frac{L_{*}/h\nu_{\alpha}}{N_{{\rm packet}}}\sum_{i,\,\nu}\delta\tau_{i,\nu}, (29)

where δτi,ν=nHσνδli\delta\tau_{i,\nu}=n_{{\rm H}}\sigma_{\nu}\delta l_{i} is the optical depth segments between events. Here, the optical depths include only those estimated for hydrogen atoms, but not due to dust extinction.

Directly counting the number of scattering events is much simpler than measuring the radiation field spectrum at the line center. However, the method of using the radiation field spectrum is advantageous to derive approximate formulae for PαP_{\alpha} from analytic solutions of the RT equation and to check the self-consistency of the Monte-Carlo simulation. We also note that Equation (29) is the same as Equation (28), except that the number of scattering events is replaced with the summation of optical depths. The probability of scattering events in an optical depth interval of δτ\delta\tau is 𝒫=1eδτδτ\mathcal{P}=1-e^{-\delta\tau}\simeq\delta\tau; therefore, Equations (28) and (29) are in fact equivalent. An advantage of using Equation (29) is that the estimator gives a smooth, non-zero PαP_{\alpha} even in the cells in which the density of neutral hydrogen is very low and thus no scattering occurs (Nscattcell=0N_{{\rm scatt}}^{{\rm cell}}=0), in a practical sense, because of a limited number of photon packets. On the other hand, the estimator in Equation (28) yielded Nscattcell=0N_{{\rm scatt}}^{{\rm cell}}=0 in more than 50% of cells, corresponding to hot and fully ionized gas, when it was applied to the simulation snapshots in Section 5.4. The results in Section 5 were thus obtained using Equation (29). However, the above three methods of calculating PαP_{\alpha} gave no significant differences in our simulations.

The average number of scatterings per a photon before escape can be obtained by integrating the scattering rate for a single photon over the whole volume, after multiplying by the neutral hydrogen density:

Nscatt=PαnH𝑑V.\left\langle N_{{\rm scatt}}\right\rangle=\int P_{\alpha}n_{{\rm H}}dV. (30)

Using this equation and Equation (27), we can also derive the analytic formulae for the mean number of scatterings in the cases of a static slab and sphere, as described in Appendixes B and C, respectively. The resulting formulae for the mean number of scatterings are

Nscattslab\displaystyle\left\langle N_{{\rm scatt}}^{{\rm slab}}\right\rangle =\displaystyle= 1.612τ0,\displaystyle 1.612\tau_{0}, (31)
Nscattsphere\displaystyle\left\langle N_{{\rm scatt}}^{{\rm sphere}}\right\rangle =\displaystyle= 0.9579τ0\displaystyle 0.9579\tau_{0} (32)

for a slab and sphere, respectively. The equation for a slab is the same as the first derived by Harrington (1973); the equation for a sphere is newly-derived in Appendix C.

Refer to caption
Figure 1.— Emergent Lyα\alpha spectra from a static, homogeneous slab at (a) TK=10T_{{\rm K}}=10 K and (b) 10410^{4} K, with different optical thicknesses (τ0=105108\tau_{0}=10^{5}-10^{8}). The black curves are line profiles calculated with LaRT. The red curves denote the approximate, analytic solution given by Neufeld (1990). The upper abscissas show the wavelength shift in Å from the Lyα\alpha line center.
Refer to caption
Figure 2.— Emergent Lyα\alpha spectra from a static, homogeneous sphere at (a) TK=10T_{{\rm K}}=10 K and (b) 10410^{4} K, with different optical depths (τ0=105108\tau_{0}=10^{5}-10^{8}). The black curves are line profiles calculated with LaRT. The red curves denote the new analytic solution given in Appendix C, which was derived from the series solution of Dijkstra et al. (2006).
Refer to caption
Figure 3.— Emergent Lyα\alpha for the dynamic motion test cases, in which the gas expands isotropically, and has a temperature of TK=104T_{{\rm K}}=10^{4} K and a column density of NHI=2×1020N_{{\rm HI}}=2\times 10^{20} cm-2. The maximum velocity vmaxv_{{\rm max}} of the Hubble-like outflow is denoted in units of km s-1. The ordinate is the mean intensity integrated over the solid angle outgoing from the spherical surface.
Refer to caption
Refer to caption
Figure 4.— Number of scatterings for (a) slab and (b) sphere as a function of the central optical depth τ0\tau_{0}. The blue and red circles denote the cases of TK=10T_{{\rm K}}=10 K and TK=104T_{{\rm K}}=10^{4} K, respectively. The black dashed lines denote the theoretical number of scatterings (a) Nscatt=1.612τ0N_{{\rm scatt}}=1.612\tau_{0} and (b) Nscatt=0.958τ0N_{{\rm scatt}}=0.958\tau_{0}, which were derived at the limit of τ0\tau_{0}\rightarrow\infty.

3. Tests of the Code

3.1. Static Media

As a test of our code, we performed the RT calculation for the cases of a static, plane-parallel slab, and sphere, for which Neufeld (1990) and Dijkstra et al. (2006), respectively, derived analytic approximations of the emergent mean intensity at the limit of large optical thickness. We also present new analytic formulae, which are slightly different from theirs, in Appendixes B and C. For these tests, we ignored the recoil effect unless otherwise stated; in other words, the recoil parameter gg_{*} in Equation (15) was assumed to be zero.

We first consider the infinite slab case in which the Lyα\alpha source is located at the mid-plane and injects monochromatic photons at frequency x=0x=0 (ν=να\nu=\nu_{\alpha}). Figure 1 shows our simulation results for the slab with a temperature of (a) TK=10T_{{\rm K}}=10 K and (b) TK=104T_{{\rm K}}=10^{4} K together with the approximate, analytic solution of Neufeld (1990). The optical thickness at line center (τ0\tau_{0}) varies from 10510^{5} to 10810^{8}, as indicated in the figure. Our results agree with the analytic solution, in general. However, as the optical thickness decreases and the temperature increases, the analytic solution begins to deviate from the simulation results. This discrepancy is because the analytic solution was derived in the limit of optically thick media (2aτ0=9.4(TK/104K)1/2(τ0/104)1032a\tau_{0}=9.4(T_{{\rm K}}/10^{4}{\rm K})^{-1/2}(\tau_{0}/10^{4})\gtrsim 10^{3}).

As the second test, Figure 2 shows the simulation results for the case of a sphere in which photons are injected at frequency x=0x=0 from the center of the sphere. The temperature of the medium is assumed to be (a) TK=10T_{{\rm K}}=10 K and (b) TK=104T_{{\rm K}}=10^{4} K, and the optical depth at line center varies from 10510^{5} to 10810^{8}, as for the slab geometry. In the figure, Equation (C3) given in Appendix C is compared. The analytic curves are in good agreement with the simulation results, except for TK=104T_{{\rm K}}=10^{4} K and τ0106\tau_{0}\lesssim 10^{6}.444The analytic solutions (Equations (B5) and (C3)) for slab and sphere geometries were derived by assuming the line wing approximation for the Voigt function, i.e., H(x,a)a/(π1/2x2)H(x,a)\approx a/(\pi^{1/2}x^{2}). They, thus, do not well reproduce the broad and deep U-shape feature at x0x\approx 0, which is caused by the core scattering, as shown for TK=104T_{{\rm K}}=10^{4} K and τ0=105\tau_{0}=10^{5} in Figure 1. If we use the exact Voigt function in the equations, the U-shape is well reproduced. However, in that case, the resulting line profile gives rise to an underestimation of the total flux, for the models with τ0106\tau_{0}\lesssim 10^{6}, integrated over frequency.

Refer to caption
Figure 5.— Emergent and dust-absorbed Lyα\alpha spectra for a sphere. The gas temperature was assumed to be TK=10T_{{\rm K}}=10 K for the top panels (a) and (b), and TK=104T_{{\rm K}}=10^{4} K for the bottom panels (c) and (d). The H I optical depth is τ0=105\tau_{0}=10^{5} for the left panels (a) and (c), and τ0=106\tau_{0}=10^{6} for the right panels (b) and (d). The black and red lines represent the emergent (JescJ_{{\rm esc}}) and absorbed (JabsJ_{{\rm abs}}) spectra, respectively. The absorbed spectra were scaled up by multiplying a denoted number for clarity. The dust properties and gas-to-dust ratio of the Milky Way were adopted in the models. The absorption optical depths (τabs\tau_{{\rm abs}}) are (a) 6.0×1056.0\times 10^{-5}, (b) 6.0×1046.0\times 10^{-4}, (c) 1.9×1031.9\times 10^{-3}, and (d) 1.9×1021.9\times 10^{-2}.

3.2. The Hubble-like flow

To test the code for the dynamic case, we examine the Hubble-like flow model. In the model, we consider an isothermal, homogeneous sphere, which is isotropically expanding or contracting. The bulk velocity of a fluid element at a distance rr from the center is assumed to be

vbulk(𝐫)=(vmaxR)𝐫,\textbf{v}_{{\rm bulk}}(\mathbf{r})=\left(\frac{v_{{\rm max}}}{R}\right)\mathbf{r}, (33)

where RR is the radius of the medium and vmaxv_{{\rm max}} is the maximum velocity at the edge of the sphere (r=Rr=R). There is no analytical solution of the emergent spectrum for a moving medium with non-zero temperature.555The analytic solution presented in Loeb & Rybicki (1999) is for the zero temperature. In other words, no thermal broadening was taken into account. We thus use the same parameters as the models of Laursen et al. (2009), Yajima et al. (2012), and Smith et al. (2017) and compare our simulation results with theirs. In the models, the gas medium is assumed to have a temperature of TK=104T_{{\rm K}}=10^{4} K and a column density of NHI=2×1020N_{{\rm HI}}=2\times 10^{20} cm-2 (corresponding to τ01.2×107\tau_{0}\simeq 1.2\times 10^{7}). The sphere of gas expands isotropically with vmax=20v_{{\rm max}}=20, 200, or 2000 km s-1. Figure 3 shows excellent agreement of our results with those of Laursen et al. (2009), Yajima et al. (2012), and Smith et al. (2017). We note that, as vmaxv_{{\rm max}} increases, the blue peak is suppressed and disappears entirely at vmax200v_{{\rm max}}\sim 200 km s-1. The peak location is also pushed further from the line center. These trends are because photons are Doppler shifted out of the line center due to the velocity gradient. However, at the extreme velocity of vmax=2000v_{{\rm max}}=2000 km s-1, the peak approaches the line center again because the extreme velocity gradient allows photons to escape more easily. If a collapsing sphere (vmax<0v_{{\rm max}}<0) is considered, then the red wing is suppressed, and the blue wing is enhanced, as opposed to the outflowing case.

3.3. Number of Scatterings

The average number of scatterings of Lyα\alpha experienced before escaping from the gas cloud is also an interesting quantity to examine. Figures 4(a) and (b) show the average number of scatterings for the infinite slab and sphere, respectively, as a function of the central optical thickness τ0\tau_{0}. We calculated the number of scatterings in the cases of the gas temperature TK=10T_{{\rm K}}=10 K (blue circles) and TK=104T_{{\rm K}}=10^{4} K (red dots) for a wide range of optical thickness τ0=1109\tau_{0}=1-10^{9}. The results cover the widest range reported to date. Overplotted as black dashed lines are the theoretical prediction (a) for the slab geometry (Equation (31)) and (b) for the spherical geometry (Equation (32)).

It is evident that the mean number of scatterings in the slab is higher than that in the sphere, because the slab geometry is infinitely extended along the XY plane and thus photons can escape only in the direction perpendicular to the XY plane.We also note that, in the optical thickness range of 102τ010610^{2}\lesssim\tau_{0}\lesssim 10^{6}, Lyα\alpha photons in the lower-temperature medium undergo less scattering than in the higher-temperature medium. This is not caused by poor statistics; we increased the number of photon packets from 10510^{5} up to 10910^{9} and found no significant differences in the results.

Note that the mean numbers of scatterings for τ0=104\tau_{0}=10^{4} and 10510^{5} in the bottom left panel of Figure 1 in Dijkstra et al. (2006) are lower than our results as well as the analytic approximation. This may be due to Dijkstra et al. (2006) having used the weighting scheme developed by Avery & House (1968), as described in Appendix B1 of Dijkstra et al. (2006). The distribution of the number of scatterings has a long tail to a large number of scatterings. The weighting scheme gives less weight to the photons that experience a large number of scatterings and thus underestimates the average number of scatterings.

Refer to caption
Figure 6.— Escape fraction fescf_{{\rm esc}} of Lyα\alpha photons from a dusty, infinite slab with a temperature (a) TK=10T_{{\rm K}}=10 K and (b) TK=104T_{{\rm K}}=10^{4} K as a function of (aτ0)1/3τabs(a\tau_{0})^{1/3}\tau_{{\rm abs}}. The escape fraction was calculated for the H I optical depths of τ0=104108\tau_{0}=10^{4}-10^{8}, and by varying the dust absorption optical depth τabs\tau_{{\rm abs}}. The escaped fraction is fairly well represented by Equation (34) with different ζ\zeta values for different optical depths and temperatures, as denoted. The function with ζ=0.525\zeta=0.525, as given in Neufeld (1990), is shown for the highest optical depth of τ0=108\tau_{0}=10^{8}.

3.4. Dust Effect and Escape Fraction

So far, dust is neglected in the tests. The presence of dust grains in the medium suppresses the Lyα\alpha intensity escaping from the medium. Figure 5 shows the emergent spectra (JescJ_{{\rm esc}}) from a dusty, homogeneous sphere of which the gas-to-dust ratio is that of the Milky Way (100\sim 100 by mass). The spectra were obtained for four combinations of the gas temperature (TK=10T_{{\rm K}}=10 and 10410^{4} K) and the optical thickness (τ0=105\tau_{0}=10^{5} and 10610^{6}). As shown in the figure, the overall shape of the emergent Lyα\alpha spectrum is not significantly altered by dust. In the figure, we also show the spectra of photons that are absorbed by dust (Jabs)J_{{\rm abs}}). The absorbed spectra are not observable but may be useful when calculating the heating of dust grains in the vicinity of H II regions by Lyα\alpha photons.

It is worthwhile to note that there is a spiky dip feature at the line center of the dust-absorption spectra for the gas of TK=10T_{{\rm K}}=10 K. The feature indicates that Lyα\alpha photons are less absorbed at the very line core than at the outside of the core. Verhamme et al. (2006) also found a similar dip feature, as shown in Figure 3 in their paper, but it is much broader and deeper than ours. We found that the dip is noticeable only when the optical depth and temperature are rather low. As the optical depth and temperature increase, the dip feature is found to be smeared out and eventually disappear in the absorbed spectra. The dip feature is discussed in more detail in Appendix 32.

We also estimated the escape fraction of Lyα\alpha photons from a dusty, infinite slab. Neufeld (1990) derived an analytic equation for the escape fraction at a limit of large optical depths, as follows:

fesc=1/cosh[3π5/2ζ{(aτ0)1/3τabs}1/2].f_{{\rm esc}}=\left.1\right/\cosh\left[\frac{\sqrt{3}}{\pi^{5/2}\zeta}\left\{\left(a\tau_{0}\right)^{1/3}\tau_{{\rm abs}}\right\}^{1/2}\right]. (34)

He estimated ζ0.525\zeta\approx 0.525 by comparing this equation with numerical results of Hummer & Kunasz (1980), obtained for τ0=5×107\tau_{0}=5\times 10^{7} and a=4.7×102a=4.7\times 10^{-2} (or equivalently TK=1T_{{\rm K}}=1 K). Figure 6 shows the escape fraction as a function of (aτ0)1/3τabs(a\tau_{0})^{1/3}\tau_{{\rm abs}} calculated for τ0=104108\tau_{0}=10^{4}-10^{8} and TK=10T_{{\rm K}}=10, 10410^{4} K. The dust optical thickness was varied so that (aτ0)1/3τabs(a\tau_{0})^{1/3}\tau_{{\rm abs}} ranges from 10210^{-2} to 10210^{2}. To compare our results with those of others, we assumed the dust scattering albedo of 0.5 and isotropic scattering (g=0g=0). It appears that Equation (34) with ζ=0.525\zeta=0.525 is indeed a good approximation for the cases of τ0>107\tau_{0}>10^{7}. However, ζ=0.525\zeta=0.525 overpredicts the escape fraction when the optical depth is lower than 10710^{7}. From the figure, it is also clear that ζ0.525\zeta\approx 0.525 is more adequate for lower temperature. This is because the equation was derived under the condition of (aτ0)1/3>τabs(a\tau_{0})^{1/3}>\tau_{{\rm abs}} and the width (aa) of the Voigt profile is proportional to TK1/2T_{{\rm K}}^{-1/2}. However, adopting a different value of ζ\zeta for different temperature TT and optical depth τ0\tau_{0}, we found that the equation still provides a reasonably good representation of the escape fraction even when the condition required for the equation is not satisfied. The ζ\zeta values suitable for different temperatures TKT_{{\rm K}} and optical depths τ0\tau_{0} are shown in Figure 6.

Refer to caption
Figure 7.— Scattering rate PαP_{\alpha} for a slab at the temperature of (a) TK=10T_{{\rm K}}=10 K and (b) TK=104T_{{\rm K}}=10^{4} K. The scattering rates for τ0=103\tau_{0}=10^{3} and 10710^{7} are shown in each panel. The black lines represent PαP_{\alpha} calculated from Monte-Carlo simulations by directly counting the number of scatterings. The red dots are those calculated using the spectrum Jx(0)J_{x}(0) obtained from the simulations. The green dashed lines denote the analytical solution, which overlaps with the simulation results for τ0=107\tau_{0}=10^{7}. We also note that the simulation results for τ0105\tau_{0}\gtrsim 10^{5} in (a) and τ0106\tau_{0}\gtrsim 10^{6} in (b), but not shown here, almost coincide with the results for τ0=107\tau_{0}=10^{7}.
Refer to caption
Figure 8.— Scattering rate PαP_{\alpha} for a sphere at the temperature of (a) TK=10T_{{\rm K}}=10 K and (b) TK=104T_{{\rm K}}=10^{4} K. The scattering rates for τ0=103\tau_{0}=10^{3} and 10710^{7} are shown in each panel. The black lines represent PαP_{\alpha} calculated from Monte-Carlo simulations by directly counting the number of scatterings. The red dots are those calculated using the spectrum Jx(0)J_{x}(0) obtained from the simulations. The green dashed lines denote the analytical solution, which overlap with the simulation results for τ0=107\tau_{0}=10^{7}. The simulation results for τ0105\tau_{0}\gtrsim 10^{5} in (a) and τ0106\tau_{0}\gtrsim 10^{6} in (b) almost coincide with the results for τ0=107\tau_{0}=10^{7}.

4. THE WF EFFECT IN SIMPLE GEOMETRIES

In this section, we present our main results that are relevant to the WF effect but calculated in simple geometries such as a uniform, infinite slab, and a sphere. We also compare the simulation results with the new analytic formulae derived in Appendixes B and C.

4.1. Scattering Rate

We present the scattering rate obtained for an unit generation rate of Lyα\alpha photons. Figure 7 compares the scattering rate calculated through a Monte-Carlo RT simulation and that obtained using an analytic approximation for the slab geometry. The results for the sphere geometry are shown in Figure 8. In the figures, the solid black lines represent the scattering rate calculated by directly counting the scattering events in the simulation. The red dots were estimated using the Lyα\alpha spectrum at the line center Jx(x=0)J_{x}(x=0) and Equation (27). The green dashed lines denote the analytical, approximate solutions at the limit of a large optical thickness. The approximate formulae for the scattering rate in slab and sphere geometries are, respectively, obtained by combining Equations (B) and (C) for Jx(0)J_{x}(0) with Equation (27). Note that the analytic solutions overlap with the simulation results for τ0=107\tau_{0}=10^{7} and are barely distinguishable in the figures.

The functional shape of PαP_{\alpha} as a function of τ/τ0\tau/\tau_{0} is more or less independent of the gas temperature and total optical thickness τ0\tau_{0}. However, its absolute strength is proportional to TK1/2T_{{\rm K}}^{-1/2}, as shown in Equation (27). Therefore, in Figures 7 and 8, the scattering rate PαP_{\alpha} at TK=10T_{{\rm K}}=10 K is greater than that at TK=104T_{{\rm K}}=10^{4} K by a factor of 31\sim 31.6 for a given optical depth τ0\tau_{0}. We also note that the functional form of PαP_{\alpha} in a sphere is much steeper than that in a slab. This tendency is due to photons in an infinite slab being able to escape the medium only through one direction that is perpendicular to the slab.

As τ0\tau_{0} increases, the scattering rate Pα(τ/τ0)P_{\alpha}(\tau/\tau_{0}) gradually decreases and converges to an asymptotic form, which is well represented by the analytic formulae denoted by the green dashed lines in Figures 7 and 8. The convergence is achieved at a lower optical thickness when the medium has a lower temperature. As denoted in the figures, good agreement between the formula and simulation was obtained at τ0105\tau_{0}\gtrsim 10^{5} for TK=10T_{{\rm K}}=10 K, but at a slightly higher optical depth of τ0106\tau_{0}\gtrsim 10^{6} for TK=104T_{{\rm K}}=10^{4} K. A similar trend can be also found in Figure 4, where the average number of scatterings begins to approach the asymptotic solution at a lower optical depth when the medium is in a lower temperature (TK=10T_{{\rm K}}=10 K) than in a higher temperature (TK=104T_{{\rm K}}=10^{4} K).

The scattering rate PαP_{\alpha} is affected by the presence of dust grains in the medium as well as by the systematic, bulk motion of the gas. Figure 9 shows the dependence of PαP_{\alpha} on the dust-absorption optical depth in a slab. In the figure, the gas temperature and optical depth are TK=104T_{{\rm K}}=10^{4} K and τ0=107\tau_{0}=10^{7}, respectively, and the dust-absorption optical depth varies from τabs=0\tau_{{\rm abs}}=0.0 (no dust) to τabs=0.24\tau_{{\rm abs}}=0.24. These results were adopted from the models calculated for Figure 6. As expected, when τabs\tau_{{\rm abs}} gradually increases, PαP_{\alpha} begins to drop. We note that, if the gas-to-dust ratio of the Milky Way is assumed, the absorption optical depth of the medium with TK=104T_{{\rm K}}=10^{4} K and τ0=107\tau_{0}=10^{7} is τabs=0.14\tau_{{\rm abs}}=0.14. This value is similar to the second-highest τabs\tau_{{\rm abs}} in Figure 9, indicating that dust grains effectively absorb Lyα\alpha photons and reduce the number of Lyα\alpha scatterings at a location of τ/τ0=0.5\tau/\tau_{0}=0.5 by a factor of 10\sim 10.

Figure 10 shows the effect of outflow on the scattering rate. In the figure, we assumed a Hubble-like, expanding sphere with a temperature of TK=100T_{{\rm K}}=100 K and an optical depth of τ0=103\tau_{0}=10^{3}. The maximum velocity is increased from vmax=0v_{{\rm max}}=0 to 10 km s-1. As expected, the scattering rate rapidly decreases as the expansion velocity rises. Figures 9 and 10 illustrate how dramatically the number of scatterings can be reduced by the presence of dust and the bulk motion of the gas.

Refer to caption
Figure 9.— Dependence of the scattering rate on the dust absorption optical depth of a slab. The gas kinetic temperature and H I optical depth are assumed to be TK=104T_{{\rm K}}=10^{4} K and τ0=107\tau_{0}=10^{7}, respectively, which corresponds to (aτ0)1/3=16.77(a\tau_{0})^{1/3}=16.77. The dust-absorption optical depth varies from 0.0 (the uppermost) to 0.24 (the lowest).
Refer to caption
Figure 10.— Dependence of the scattering rate on the maximum velocity in a Hubble-like expanding, spherical medium with TK=100T_{{\rm K}}=100 K and τ0=103\tau_{0}=10^{3}. The maximum expanding velocity varies from 0 km s-1 (the uppermost) to 10 km s-1 (the lowest).

4.2. Lyα\alpha Line Profile inside the Medium

We first examine the line profile of Lyα\alpha formed within a spherical medium when the recoil effect is ignored, i.e., g=0g_{*}=0 in Equation (15). Figures 11 and 12 show the line profiles at various radial locations of a spherical medium with a temperature of TK=10T_{{\rm K}}=10 and 10410^{4} K, respectively. In each figure, the results for two different optical depths (τ0=105\tau_{0}=10^{5} and 10710^{7} in Figure 11, and τ0=106\tau_{0}=10^{6} and 10710^{7} in Figure 12) are shown. The black lines in the figures denote the simulation results, while the red lines show the analytic solutions given in Equation (C3). It is clear that the approximate formula reproduces the simulation results reasonably well, in particular, when the optical depth is large enough, i.e., τ0107\tau_{0}\gtrsim 10^{7} and the gas temperature is low (TK=10T_{{\rm K}}=10 K). For TK=104T_{{\rm K}}=10^{4} K and τ0<106\tau_{0}<10^{6}, however, the analytic solution significantly underestimates the intensity. We also note that the analytic solution tends to underestimate the Lyα\alpha intensity in the central portion and overestimate in outer parts. The same trend was found by Higgins & Meiksin (2012) who solved the moment equations numerically. We also found a similar trend for a slab, but not shown in this paper. It should also be mentioned that the Lyα\alpha spectrum evolves progressively to a typical double-peak shape as upon approaching the system boundary (τ/τ00.95\tau/\tau_{0}\gtrsim 0.95), but not clearly shown in the figures.

Refer to caption
Figure 11.— Spectra of the Lyα\alpha radiation field at various radial locations (corresponding to τ/τ0=\tau/\tau_{0}= 0.50, 0.55, \cdots, 0.95) in a spherical medium with a temperature of TK=10T_{{\rm K}}=10 K. No recoil effect is taken into account so that the line center is flat. The optical depth at line center is (a) τ0=105\tau_{0}=10^{5} and (b) τ0=107\tau_{0}=10^{7}. Black lines show our simulation results and red lines are from the analytical approximation of Equation (C3).
Refer to caption
Figure 12.— Spectra of the Lyα\alpha radiation field at various radial locations (corresponding to τ/τ0=\tau/\tau_{0}= 0.50, 0.55, \cdots, 0.95) in a spherical medium with a temperature of TK=104T_{{\rm K}}=10^{4} K. No recoil effect is taken into account so that the line center is flat. The optical depth at line center is (a) τ0=106\tau_{0}=10^{6} and (b) τ0=107\tau_{0}=10^{7}. Black lines show our simulation results and red lines are from the analytical approximation of Equation (C3).
Refer to caption
Refer to caption
Figure 13.— Recoil effect on the Lyα\alpha radiation field within a slab. The gas has a temperature of (a)(b) TK=10T_{{\rm K}}=10 K, (c)(d) TK=102T_{{\rm K}}=10^{2} K, and (e)(f) TK=104T_{{\rm K}}=10^{4} K. The optical depth at line center is (a)(c)(e) τ0=102\tau_{0}=10^{2} and (b)(d)(f) τ0=5×102\tau_{0}=5\times 10^{2}. The figures show the spectra in linear scale. The black solid lines represent the spectra obtained when the fine structure was ignored, while the blue solid lines were obtained after taking it into account. The blue dotted lines denote the frequency interval |x|x|x|\leq x_{*}. Here, x=0.84+|νKνH|/2ΔνDx_{*}=0.84+\left|\nu_{{\rm K}}-\nu_{{\rm H}}\right|/2\Delta\nu_{{\rm D}} for (a) to (d), and x=0.84x_{*}=0.84 for (e) and (f).
Refer to caption
Refer to caption
Figure 14.— Recoil effect on the Lyα\alpha radiation field within a slab. The same as Figure 13 except that the spectra are shown in logarithmic scale. The gas has a temperature of (a)(b) TK=10T_{{\rm K}}=10 K, (c)(d) TK=102T_{{\rm K}}=10^{2} K, and (e)(f) TK=104T_{{\rm K}}=10^{4} K. The optical depth at line center is (a)(c)(e) τ0=102\tau_{0}=10^{2} and (b)(d)(f) τ0=5×102\tau_{0}=5\times 10^{2}.
Refer to caption
Figure 15.— Lyα\alpha line profiles in an expanding medium with vmax=5v_{{\rm max}}=5 km s-1, TK=100T_{{\rm K}}=100 K, and τ0=103\tau_{0}=10^{3}. The spectra are shown at optical depths of τ/τ0=0.0,\tau/\tau_{0}=0.0, 0.2, 0.4, 0.6, 0.8, and 1.0. They were normalized by the maximum intensities for convenience.
Refer to caption
Figure 16.— Thermalization of the Lyα\alpha line profiles inside an expanding medium with vmax=5v_{{\rm max}}=5 km s-1. The model is the same as in Figure 15, but the line profiles are shown in logarithmic scale for four optical depth bins.

We next consider the case in which the atomic recoil effect is taken into account in the simulation. Figure 13 shows the Lyα\alpha spectra in a static, infinite slab for six different sets of physical parameters. The gas temperature is assumed to be TK=10T_{{\rm K}}=10 K in Figures 13(a) and (b), while it is TK=102T_{{\rm K}}=10^{2} K in Figures 13(c) and (d), and TK=104T_{{\rm K}}=10^{4} K in Figures 13(e) and (f). The optical thickness for the system is τ0=102\tau_{0}=10^{2} for Figures 13(a), (c), and (e), and τ0=5×102\tau_{0}=5\times 10^{2} for Figures 13(b), (d), and (f). In each figure, the left panels show the Lyα\alpha spectra measured in an optical depth bin of 0.2<τ/τ0<0.50.2<\tau/\tau_{0}<0.5; the right panels show the spectra in 0.5<τ/τ0<0.80.5<\tau/\tau_{0}<0.8.

As noted in Section 2.3, the frequency gap between the fine-structure levels of the n=2n=2 state can be comparable to the line width formed in the medium with a relatively low optical depth and low temperature (i.e., TK102T_{{\rm K}}\lesssim 10^{2} K). In Figures 13(a) to (d), we, therefore, show also the Lyα\alpha spectra obtained when the fine structure is considered. In the figures, the black and blue solid lines denote the Lyα\alpha spectra calculated when the fine structure is ignored and taken into consideration, respectively. In Figure 13(a) for τ0=102\tau_{0}=10^{2}, we see that the spectra, denoted in blue, show two steps due to the fine structure. However, as the optical thickness increases higher than τ05×102\tau_{0}\approx 5\times 10^{2}, the step-like shape disappears, as shown in Figure 13(b) for τ0=5×102\tau_{0}=5\times 10^{2}. The intensity ratio of the doublet line becomes 1:1 at high optical depths as a result of a considerable number of resonance scatterings, and thus the step signature of the doublet disappears. The step shape disappeared at a lower optical depth as temperature increases. For instance, in the case of TK=102T_{{\rm K}}=10^{2} K (Figure 13(c)), the step shape is not found already at τ0=102\tau_{0}=10^{2}.

The most notable thing in Figures 13(a) to (d) is that the Lyα\alpha spectra are tilted in the central part, unlike those shown in Figures 11 and 12 in which no recoil effect was considered. To compare with the expectation in the WF effect theory, we overlay an exponential function exp[h(ννα)/kBTK]\exp\left[-h(\nu-\nu_{\alpha})/k_{{\rm B}}T_{{\rm K}}\right] for the gas temperature TKT_{{\rm K}} with a red dashed line. As shown in the figures, the spectra clearly appears to be consistent with the function expected for the WF effect. However, in Figures 13(e) and (f) for the highest temperature of TK=104T_{{\rm K}}=10^{4} K, the spectral shape appears to be nearly constant at the central part. In fact, the exponential function for a high TKT_{{\rm K}} is almost constant. Hence, at first glance, it seems that it is not possible to confirm whether the Lyα\alpha spectra within the medium at a high temperature follow an expected functional form or not.

However, we note that the spectrum can be expressed by

kBTKhΔνDln(Jx/J0)=x.\frac{k_{{\rm B}}T_{{\rm K}}}{h\Delta\nu_{{\rm D}}}\ln\left(J_{x}/J_{0}\right)=-x. (35)

Thus, if the spectrum is drawn in logarithmic scale, as in the left side of the above equation, and compared with x-x, we can readily illustrate that the spectrum follows the expected function. Figure 14 compares the Lyα\alpha spectra represented in logarithmic scale with x-x, which is indicated by the dashed red line, for all cases shown in Figure 13. Consequently, we conclude that the Lyα\alpha spectrum is well described by an exponential function with a slope of the gas temperature, even at the highest temperature (TK=104T_{{\rm K}}=10^{4} K) presented here. Interestingly, Figures 13(a) and 14(a) show that the double steps (blue lines) have the same exponential slope corresponding to the gas temperature, although they are disconnected.

Here, we note that the exponential shape should remain valid across the frequency interval corresponding to the width of the resonance profile ϕx\phi_{x}, i.e., |x|x=0.84(=ln2)|x|\lesssim x_{*}=0.84\ (=\ln 2). In this paper, we take this width as the full-width at half maximum (FWHM) of the Voigt function for a1a\ll 1, or equivalently the FWHM of the Gaussian function with a standard deviation of 1/21/\sqrt{2}. If we additionally consider the line splitting due to the fine structure, the exponential shape should be kept over the whole range of the frequency gap of the doublet plus the width of the resonance profile, i.e., |x|x=0.84+|ν3/2ν1/2|/(2ΔνD)|x|\lesssim x_{*}=0.84+|\nu_{\nicefrac{{3}}{{2}}}-\nu_{\nicefrac{{1}}{{2}}}|/(2\Delta\nu_{{\rm D}}). Otherwise, the 21-cm spin temperature could not approach the gas temperature. Note that |ν3/2ν1/2|/(2ΔνD)=5.83×102(TK/104K)1/2|\nu_{\nicefrac{{3}}{{2}}}-\nu_{\nicefrac{{1}}{{2}}}|/(2\Delta\nu_{{\rm D}})=5.83\times 10^{-2}\left(T_{{\rm K}}/10^{4}\,\text{K}\right)^{-1/2}. In order to check if this is the case, the vertical blue dashed lines are overlaid in Figures 13 and 14 (and in other figures as well where necessary) to denote the frequency interval |x|x|x|\leq x_{*}.

In the case of TK=10T_{{\rm K}}=10 K and τ0=102\tau_{0}=10^{2} shown in Figure 14(a), it appears that the spectral shape does not satisfy the requirement because of the discontinuous step feature. On the other hand, we found that the spectra obtained for τ03×102\tau_{0}\gtrsim 3\times 10^{2} and TK=10T_{{\rm K}}=10 K follows the required exponential shape over the frequency interval |x|x|x|\leq x_{*}, for instance, as shown in Figure 14(b). In the case of TK=102T_{{\rm K}}=10^{2} K (τ0=102\tau_{0}=10^{2} and τ0=5×102\tau_{0}=5\times 10^{2}), the desired profile is more clearly seen in Figures 14(c) and (d). Figures 14(e) and (f) also show that the line profiles follow the expected shape very well, over the frequency interval |x|x|x|\leq x_{*}, even for the highest temperature TK=104T_{{\rm K}}=10^{4} K. We examined many different cases and found that the necessary condition for the Lyα\alpha spectral profile is fulfilled at an optical depth as low as τ0(15)×102\tau_{0}\simeq(1-5)\times 10^{2}. In general, as the temperature is low, it appeared that the requirement is satisfied at a lower optical depth. If the optical depth is lower than τ0102\tau_{0}\simeq 10^{2}, the Lyα\alpha spectrum was, in general, not fully thermalized in the temperature range of 10KTK104K10\,\text{K}\leq T_{{\rm K}}\leq 10^{4}\,\text{K}, examined in this paper.

The above results were obtained in static media. It would, therefore, be worthwhile to examine whether the Lyα\alpha spectrum will have a shape of the required exponential function in a medium in motion. For instance, we show the spectra within a Hubble-like, expanding medium in Figures 15 and 16. Both figures were obtained for an expanding medium with τ0=103\tau_{0}=10^{3}, TK=102T_{{\rm K}}=10^{2} K. Figure 15 shows the Lyα\alpha spectra at several locations, expressed in terms of optical depth, in a spherical medium with a maximum expanding velocity of vmax=5v_{{\rm max}}=5 km s-1. In Figure 15, the spectra near the line center (x=0x=0) seem flat. However, more detailed views in Figure 16 clearly illustrate that the spectra calculated at various optical depth bins follow the exponential function for TK=102T_{{\rm K}}=10^{2} K. We, therefore, can conclude that the Lyα\alpha line profile follows the desired exponential function with a slope corresponding to the gas temperature even in an expanding medium, if the expanding velocity is not too fast. If a medium rapidly expands or, more precisely, has a velocity gradient, across a region with an optical thickness of τ0100500\tau_{0}\approx 100-500, larger than the thermal velocity dispersion, Lyα\alpha photons will escape the region before they fully develop the spectral shape required for the WF effect. The relevant issue in moving media is further addressed in the next section.

The initial frequency of Lyα\alpha was fixed at ν=να\nu=\nu_{\alpha} (x=0x=0) for the above discussion. The same conclusion was obtained even when we assumed a Voigt profile or a flat continuum. The Lyα\alpha line is broadened or adjusted very quickly in a medium with an optical thickness of τ0102\tau_{0}\gtrsim 10^{2} so that the initial frequency distribution did not significantly alter the resulting spectra; only the spectra in the vicinity of the source showed a weak dependence on the initial frequency distribution.

Refer to caption
Figure 17.— Turbulence effect on the Lyα\alpha line profile emerging from a sphere with the central optical depth (a) τ0=10\tau_{0}=10 and (b) τ0=105\tau_{0}=10^{5}. The black lines are the spectra emerging from a sphere with TK=102T_{{\rm K}}=10^{2} K when random turbulent motion with vturb=1.28v_{{\rm turb}}=1.28 km s-1 is applied. The red lines were obtained when no random turbulent motion was taken into consideration, but the temperature of the medium was assumed to be TK=2×102T_{{\rm K}}=2\times 10^{2} K, corresponding to the Doppler temperature TDT_{{\rm D}}, defined in Equation (36). All spectra are shown as functions of the same relative frequency xTx_{T}, which is defined for TK=200T_{{\rm K}}=200 K, for ease of comparison.
Refer to caption
Figure 18.— Turbulence effect, when the coherence scale of optical depth varies, on the Lyα\alpha line profile in a spherical medium with the central optical depth τ0=103\tau_{0}=10^{3} and temperature TK=102T_{{\rm K}}=10^{2} K. From the top to bottom panels, the coherence scale decreases from τcoh=50\tau_{{\rm coh}}=50 to 2. The red dashed lines represent the exponential function for TK=102T_{{\rm K}}=10^{2} K, while the green dashed lines are for TK=200T_{{\rm K}}=200 K.
Refer to caption
Figure 19.— Lyα\alpha line profiles in a multiphase medium in which three kinetic temperatures (TK=T_{{\rm K}}= 100, 500, and 1000 K) are equally mixed. The central optical depth was assumed to be τ0=103\tau_{0}=10^{3}, 5×1035\times 10^{3}, 10410^{4}, and 5×1045\times 10^{4} from left to right. The coherence optical depth is τcoh=10\tau_{{\rm coh}}=10, 5050, 10210^{2}, and 5×1025\times 10^{2}, respectively, from the first to fourth column. The red dashed lines denote the exponential functions for (a) TK=100T_{{\rm K}}=100 K, (b) 500 K, and (c) 1000 K. The black lines show the Lyα\alpha line profiles when there is no random motion was applied while the red lines represent the models that include a random, turbulent motion with a velocity dispersion of vturb=4.06v_{{\rm turb}}=4.06 km s-1.

4.3. Turbulence Effect on Line Profile

In the above calculation of the Lyα\alpha RT, only the thermal motion and systematic bulk flow of gas were taken into account. Aside from the thermal broadening, the line width of Lyα\alpha may also be broadened by the presence of the turbulent velocity field. It is convenient to define the Doppler temperature TDT_{{\rm D}} as in Liszt (2001), to incorporate the turbulence effect (see also Draine, 2011):

2kBTD/mH=2kBTK/mH+vturb2,2k_{{\rm B}}T_{{\rm D}}/m_{{\rm H}}=2k_{{\rm B}}T_{{\rm K}}/m_{{\rm H}}+v_{{\rm turb}}^{2}, (36)

where vturbv_{{\rm turb}} is the velocity dispersion of turbulent motion. The Doppler parameter bb defined by

b=vth2+vturb2b=\sqrt{v_{{\rm th}}^{2}+v_{{\rm turb}}^{2}} (37)

is also useful to describe effects caused by the turbulent motions, as in Verhamme et al. (2006). The Doppler temperature TDT_{{\rm D}} (or Doppler parameter bb) has been used in place of the thermal kinetic temperature (or thermal velocity) to describe the Lyα\alpha spectral shape. However, it has not been verified whether this prescription of incorporating the turbulent fluid motion is appropriate in the context of Lyα\alpha RT, especially with regards to the WF effect. Hydrodynamic simulations for galaxies, where the turbulence effects are included, have been adopted to calculate the emergent Lyα\alpha spectra. However, in these studies, the turbulence effect could not be separated from the pure thermal motion effect.

Before we go forward, we need to classify random or turbulent motions into two categories, microturbulence and macroturbulence, as described in Leung & Liszt (1976). Microturbulence is used to refer to the case in which the correlation length of the velocity field (or the size of a typical turbulent element) is small compared with the photon mean free path. Macroturbulence refers to the case in which the scale length of the turbulent fluctuations is large enough compared with the mean free path of photons. In the context of the WF effect, we need to compare the correlation length with a length corresponding to τ0100500\tau_{0}\approx 100-500, instead of the mean free path, which is required to make the color temperature equal to the kinetic temperature.

In this paper, to quantify the degree of spatial coherence (or correlation) of the medium, we define the coherence length as the size over which the physical quantities, such as density, temperature, and bulk motion, remain uniform. We also define the coherence optical depth as an optical depth corresponding to the coherence length. In our simulation, each cell has a constant density, temperature, and bulk motion velocity and thus the coherence length (or optical depth) is the size (optical depth) of individual cells. In the following models, we utilize a 3D Cartesian box to implement spherical geometry by setting the density zero outside of a certain radius.

First, we examine how the random, turbulence motion affects the emergent Lyα\alpha spectrum by adopting a simple, random motion model to mimic turbulence. We employ a random, 3D velocity field in which three velocity components in each cell are independently produced to follow a Gaussian distribution with a standard deviation of vturb/2v_{{\rm turb}}/\sqrt{2}. According to the above prescription to describe the turbulence effect, the emergent spectrum from a medium with a kinetic temperature TKT_{{\rm K}} and having a Gaussian random motion with a velocity dispersion vturbv_{{\rm turb}} is supposed to be equivalent to that obtained from a medium with a gas temperature of TD=(TK+mHvturb2/2kB)1/2T_{{\rm D}}=\left(T_{{\rm K}}+m_{{\rm H}}v_{{\rm turb}}^{2}/2k_{{\rm B}}\right)^{1/2}, but with no turbulence motion.

In Figure 17, we show the results for TK=102T_{{\rm K}}=10^{2} K, vturb=1.28v_{{\rm turb}}=1.28 km s-1, and TD=2×102T_{{\rm D}}=2\times 10^{2} K. In the figure, the spectrum emerging out of a medium with the turbulence motion is denoted in black, and the emergent spectrum from a medium with only the pure thermal motion corresponding to the Doppler temperature TDT_{{\rm D}} in red. The medium is assumed to have an optical depth of τ0=102\tau_{0}=10^{2} and τ0=105\tau_{0}=10^{5}, respectively, in Figures 17(a) and 17(b). The coherence optical depth is taken to be τcoh=0.5\tau_{{\rm coh}}=0.5 for the case of optical depth τ0=102\tau_{0}=10^{2} and τcoh=103\tau_{{\rm coh}}=10^{3} for τ0=105\tau_{0}=10^{5}; in other words, the number of cells in the simulation box is 4003400^{3} for Figure 17(a) and 2003200^{3} for Figure 17(b). Thus, Figure 17(a) illustrates the case of microturbulence and Figure 17(b) macroturbulence. When dealing with several models in which turbulent motion or many different velocities are concerned, we need to use a common abscissa in the spectra. We use xTKx_{T_{{\rm K}}} to denote the relative frequency defined for the temperature TKT_{{\rm K}}; for instance, x200x_{200} in Figure 17 represents the relative frequency defined when TK=200T_{{\rm K}}=200 K. In the figures, we find that the two results obtained using two different methods agree reasonably well with each other; therefore, the commonly adopted prescription provides an excellent method to predict the emergent Lyα\alpha spectrum from a turbulent medium regardless of the coherence length.

Second, we investigate the turbulence effect on the Lyα\alpha spectrum measured inside the medium. We calculated the Lyα\alpha radiation field spectra in a spherical medium with TK=102T_{{\rm K}}=10^{2} K and τ0=103\tau_{0}=10^{3}. The turbulent velocity is again vturb=1.28v_{{\rm turb}}=1.28 km s-1, and the coherence optical depth was adjusted to be τcoh=50\tau_{{\rm coh}}=50, 10, and 2 by varying the cell size (or the number of cells of the system). Figure 18 shows the spectra at two radial distance intervals corresponding to 0.2<τ/τ0<0.50.2<\tau/\tau_{0}<0.5 and 0.5<τ/τ0<0.80.5<\tau/\tau_{0}<0.8 for each model. In the figure, the solid black lines represent the calculated Lyα\alpha spectra, and the red and green dashed lines denote exponential functions for TKT_{{\rm K}} and TDT_{{\rm D}}, respectively. From Figure 18(a) (τcoh=50\tau_{{\rm coh}}=50) to 18(c) (τcoh=2\tau_{{\rm coh}}=2), the spectra gets shallower from x100-x_{100} (for TKT_{{\rm K}}) to x100/2-x_{100}/2 (for TDT_{{\rm D}}). These results indicate that in the case of large coherence optical depths (macroturbulence), the color temperature is solely determined by the gas kinetic temperature; in other words, the prescription to incorporate turbulent motions is only valid for the microturbulence case (τcohτWF100\tau_{{\rm coh}}\ll\tau_{{\rm WF}}\approx 100).

The ISM and IGM will not have a constant temperature, but rather a continuously varying temperature structure; this is called the multiphase medium. Thus, there would be a possibility that Lyα\alpha photons that were thermalized to one temperature in a volume element could then be perturbed or disturbed while they pass through an adjacent parcel with a different temperature. It may also be possible that the Lyα\alpha photons can be no longer characterized by the gas temperature of the volume element that the photons traverse. It is, therefore, worthwhile to examine whether thermalization can immediately occur after photons enter the next parcel with a different temperature.

To investigate this possibility in a multiphase medium, we used a uniform density medium and randomly assigned one of three temperatures of TK=100T_{{\rm K}}=100, 500, and 1000 K to every cell with an equal probability of 1/31/3 and estimated the average spectra taken from the cells with each temperature. In Figure 19, we show the results (black lines) for τ0=103\tau_{0}=10^{3}, 5×1035\times 10^{3}, 10410^{4}, and 5×1045\times 10^{4} from the first to fourth column. For each optical depth, the top, middle, and bottom panels show the spectra of the phases with TK=100T_{{\rm K}}=100, 500, and 1000 K, respectively. Each cell has a size of optical depth of τcoh=τ0/100\tau_{{\rm coh}}=\tau_{0}/100, and thus, for instance, τcoh=10\tau_{{\rm coh}}=10 for the case of τ0=103\tau_{0}=10^{3} and τcoh=5×102\tau_{{\rm coh}}=5\times 10^{2} for τ0=5×104\tau_{0}=5\times 10^{4}. In the first and second columns (τcoh=10\tau_{{\rm coh}}=10 and 5050), we find that the Lyα\alpha spectrum is not thermalized to the gas temperature (TKT_{{\rm K}}) at all in the cells with the higher temperatures TK=5×102T_{{\rm K}}=5\times 10^{2} and 10310^{3} K; the spectrum is steeper than that expected, because of the influence of spectra having lower temperatures in adjacent cells. However, in the cells with the lowest temperature (TK=100T_{{\rm K}}=100 K), the spectrum is fully thermalized even at a coherence optical depth as low as τcoh=50\tau_{{\rm coh}}=50 (second column). For the intermediate temperature (TK=500T_{{\rm K}}=500 K) and τcoh=50\tau_{{\rm coh}}=50, the spectrum appears to have a color temperature that is slightly different from the gas temperature. These results are because the thermalization is achieved more quickly in a lower temperature medium than in a higher temperature medium. On the other hand, in the case of the higher optical depth of τ0=5×104\tau_{0}=5\times 10^{4}, as shown in the rightmost panels, in which the optical depth of individual cells is 5×1025\times 10^{2}, the Lyα\alpha spectra are found to be fully thermalized to the gas temperatures in all cells regardless of the gas temperature.

In Figure 19, we also show the spectra (red solid lines) taken from the models in which a random velocity field with vturb=4.06v_{{\rm turb}}=4.06 km s-1 (corresponding to a temperature of 103 K) was additionally applied to the same media. For low optical depths (first and second columns), the resulting Lyα\alpha line profiles show shallower slopes compared to those (black lines) obtained when no random, turbulent motion is employed. The shallow slopes are attributable to the influence of turbulent motion (TD>TKT_{{\rm D}}>T_{{\rm K}}). However, even in this case, we find the same conclusion as before that thermalization is more readily achieved in high optical depths and low temperatures.

In summary, we conclude that the emergent spectrum from a turbulent medium becomes wider according to the commonly adopted prescription regardless of the coherence length. However, the color temperature of the Lyα\alpha radiation field within the turbulent medium approaches the gas temperature, as opposed to the common assumption, unless the coherence scale is too small. In most astrophysical environments, the Lyα\alpha optical depth would be very large for a turbulent length scale, as discussed in Section 6.2, and thus, we conclude that the color temperature in the WF effect theory should include only the gas temperature, but not the turbulence effect.

Refer to caption
Figure 20.— Spin temperature for the Ferriere model at the solar vicinity (Ferriere, 1998). (a) Spin temperature distribution of the WNM as a function of height from the Galactic plane. (b) Spin temperature distribution of the CNM. Lyα\alpha photons were produced either from H II regions or cooling hot gas. The label “tot” indicates the model that includes both sources. The gas temperature assumed for each phase is also denoted in each panel. The hatch areas and error bars indicate the standard deviation of temperature at a given height. The abscissa values were slightly moved to clarify the data points. The green lines and dots denote the spin temperature when no Lyα\alpha pumping is taken into account.
Refer to caption
Figure 21.— The emergent Lyα\alpha spectra obtained from the Ferriere model. The Lyα\alpha production rate (ΨLyα\Psi_{{\rm Ly}\alpha}) and escape fraction (fescf_{{\rm esc}}) for each source (H II regions and cooling gas) are shown together with those for the combined model (“tot”).

5. THE WF EFFECTS IN ISM MODELS

In the previous section, we present basic properties of the Lyα\alpha RT with regard to the WF effect. We now apply our methods to more realistic ISM models that may be appropriate to the solar neighborhood in the Milky Way galaxy. We first describe the Lyα\alpha production mechanisms (Section 5.1) and the collisional transition rates for the H I hyperfine levels (5.2), which are required to examine the WF effect in the ISM models. We use two multiphase ISM models, (1) a rather idealized theoretical model for the vertical profiles of density and volume filling factors (Ferriere, 1998; Section 5.3) and (2) snapshots of a full 3D MHD simulation result (Kim & Ostriker, 2017; Section 5.4). The present results can be readily applied to other hydrodynamic models for whole galaxies.

5.1. Lyα\alpha Sources: Photoionization and Collisional Cooling

There are two primary mechanisms to produce Lyα\alpha photons. First, photoionization of hydrogen by the UV radiation field is followed by the recombination, leaving the atom in an excited state. The subsequent radiative cascades to the ground state can then produce a Lyα\alpha photon, which is often called the ‘fluorescent’ radiation in the Lyα\alpha astronomy community. H II regions that are photoionized by bright OB associations are the most prominent sources of Lyα\alpha photons in the ISM. Second, the collision between an electron and a hydrogen atom can excite or ionize the atom, and then the hydrogen produces a Lyα\alpha photon as the atom decays down to the ground state or recombines. This process is referred to as Lyα\alpha production via ‘cooling.’

Both Lyα\alpha sources are included in the following RT simulations. For the Lyα\alpha recombination photons, we inject initial Lyα\alpha photons near the Galactic plane smoothly following the vertical distribution of H II regions in the vicinity of the Sun. We adopt a Gaussian vertical distribution with a scale height of 81 pc, while the horizontal distribution is uniform. Lyα\alpha photons are injected with frequencies according to a Voigt profile for the H II region temperature of TK=104T_{{\rm K}}=10^{4} K. Note that the initial photon frequency did not significantly alter our RT simulation results.

Vacca et al. (1996) estimated the production rate of ionizing UV photons (Lyman continuum; Lyc) from 429 O- and early B stars within 2.5 kpc of the Sun to be ΨLyc=3.7×107\Psi_{{\rm Lyc}}=3.7\times 10^{7} photons cm-2 s-1. Each ionizing photon should produce 0.68\sim 0.68 Lyα\alpha photons in the case B condition. Hence, the production rate of Lyα\alpha photons by H II regions in the vicinity of the Sun is ΨLyα=2.52×107\Psi_{{\rm Ly}\alpha}=2.52\times 10^{7} photons cm-2 s-1. The production rate is expressed by ΨLyα=/hνα\Psi_{{\rm Ly}\alpha}=\mathcal{L}_{*}/h\nu_{\alpha} in terms introduced in Section 2.5.

To calculate the Lyα\alpha production rate due to the collisional cooling, we assumed that the gas is in the collisional ionization equilibrium (CIE), in which the ionization fraction of hydrogen is calculated by balancing the collisional ionization rate with the case A recombination rate of hydrogen. The ionization and recombination rate coefficients of hydrogen are given in Equations (13.11) and (14.5), respectively, of Draine (2011). To calculate the electron density, we also included electrons produced by collisional ionization of helium. The collisional ionization fraction of helium was calculated using the collisional ionization and recombination rate coefficients provided in Cen (1992).

The total Lyα\alpha emissivity caused by the collisional excitation and ionization is given by

4πjLyαhνα=nenHCLyα(TK)hνα+nenpαBPB(Lyα),\frac{4\pi j_{{\rm Ly}\alpha}}{h\nu_{\alpha}}=n_{e}n_{{\rm H}}\frac{C_{{\rm Ly}\alpha}(T_{{\rm K}})}{h\nu_{\alpha}}+n_{e}n_{p}\alpha_{{\rm B}}P_{{\rm B}}({\rm Ly}\alpha), (38)

where nen_{e}, nHn_{{\rm H}}, and npn_{p} are the densities of electron, neutral hydrogen, and proton, respectively. We calculated the cooling rate coefficient CLyα(TK)C_{{\rm Ly}\alpha}(T_{{\rm K}}) via Lyα\alpha emission using version 9 of the CHIANTI666www.chiantidatabase.org spectral code (Dere et al., 1997, 2019), which we found to be well represented by

CLyα(TK)=4.13×1019exp(117744TK)ergcm3s1,C_{{\rm Ly}\alpha}(T_{{\rm K}})=4.13\times 10^{-19}\exp\left(-\frac{117744}{T_{{\rm K}}}\right)\ \text{erg}\ {\rm cm^{3}}\ {\rm s}^{-1}, (39)

within 10% error in most cases and 20% error at the very most.777We compared Equation (39) with those of Cen (1992) and Giovanardi et al. (1987). The polynomial equation given in Giovanardi et al. (1987) was found to underestimate the cooling rate systematically by \sim30–35%. We also found that adopting that of Cen (1992) did not significantly alter the present results obtained using Equation (39). Note that the total cooling rate from hydrogen is about 20% higher than that estimated using the equation. The case B recombination rate coefficient αB\alpha_{{\rm B}} is given in Equation (14.6) of Draine (2011). The probability PB(Lyα)P_{{\rm B}}({\rm Ly}\alpha), as a function of gas temperature, that a recombination event leads to the production of a Lyα\alpha photon in the case B condition is given by Cantalupo et al. (2008).

5.2. Collisional Transition Rate of the Hyperfine Levels

The downward transition rate from the hyperfine level “1” to “0” due to particle collisions P10cP_{10}^{{\rm c}} in Equation (24) is the sum of the downward rates by all collision partners (hydrogen atoms, protons, and electrons). The downward rate associated with interactions with electrons and protons were taken from Liszt (2001). We adopted the formula given by Shaw (2005) for the downward transition rate arising from collisions with other hydrogen atoms.

5.3. The Ferriere Model

Ferriere (1998) describes a global model of the ISM in our Galaxy. The ISM comprises of two neutral media and two ionized media; the WNM and CNM, and the warm ionized medium (WIM) and hot ionized medium (HIM). The CNM has a temperature of 80 K, the WNM and WIM a temperature of 8000 K, and the HIM a temperature of 10610^{6} K. Partial ionization is not considered in this model so that no free electrons coexist with the neutral hydrogen. Therefore, the recombination process that follows collisional ionization (WIM) or photoionization (H II regions) is the only Lyα\alpha source.

The space-averaged densities in the CNM, WNM, WIM, and HIM near the Sun are approximated, as a function of height zz perpendicular to the Galactic plane, by

ρCNM(z)ρCNM0\displaystyle\frac{\left\langle\rho_{{\rm CNM}}(z)\right\rangle}{\rho_{{\rm CNM}}^{0}} =\displaystyle= 0.859e(z/H1)2+0.047e(z/H2)2+0.094e|z|/H3\displaystyle 0.859e^{-\left(z/H_{1}\right)^{2}}+0.047e^{-\left(z/H_{2}\right)^{2}}+0.094e^{-\left|z\right|/H_{3}}
ρWNM(z)ρWNM0\displaystyle\frac{\left\langle\rho_{{\rm WNM}}(z)\right\rangle}{\rho_{{\rm WNM}}^{0}} =\displaystyle= 0.456e(z/H1)2+0.403e(z/H2)2+0.141e|z|/H3\displaystyle 0.456e^{-\left(z/H_{1}\right)^{2}}+0.403e^{-\left(z/H_{2}\right)^{2}}+0.141e^{-\left|z\right|/H_{3}}
ρWIM(z)ρWIM0\displaystyle\frac{\left\langle\rho_{{\rm WIM}}(z)\right\rangle}{\rho_{{\rm WIM}}^{0}} =\displaystyle= 0.948e|z|/1kpc+0.052e|z|/150pc\displaystyle 0.948e^{-\left|z\right|/1{\rm kpc}}+0.052e^{-\left|z\right|/150{\rm pc}}
ρHIM(z)ρHIM0\displaystyle\frac{\left\langle\rho_{{\rm HIM}}(z)\right\rangle}{\rho_{{\rm HIM}}^{0}} =\displaystyle= 0.12+0.88e|z|/1.5kpc,\displaystyle 0.12+0.88e^{-\left|z\right|/1.5{\rm kpc}},

where the densities at the Galactic plane are ρCNM0=0.340\rho_{{\rm CNM}}^{0}=0.340 cm-3, ρWNM0=0.226\rho_{{\rm WNM}}^{0}=0.226 cm-3, ρWIM0=0.025\rho_{{\rm WIM}}^{0}=0.025 cm-3, and ρHIM0=4.8×104\rho_{{\rm HIM}}^{0}=4.8\times 10^{-4} cm-3, and the three scaleheights are H1=0.127H_{1}=0.127 kpc, H2=0.318H_{2}=0.318 kpc, and H3=0.403H_{3}=0.403 kpc.

The volume filling factors, defined as the ratios of their space-averaged densities to their true densities, of the three phases (CNM, WNM, and WIM) are calculated using Equations (38) to (40) of Ferriere (1998), as functions of the filling factor of the HIM phase and the thermal pressures. The filling factor of the HIM phase at the solar position (R=RR=R_{\odot}) is given as a function of height from the Galactic plane in Figure 12 of Ferriere (1998); we digitized the figure for the present work.

We used these relations to calculate the volume filling factors and true densities of the four phases at each vertical height bin. We then randomly assigned one of four phases to every volume element by comparing a random number with the given filling factors at that vertical height. The temperatures and true densities are then determined for each volume element. For this model, the simulation box was assumed to have a physical size of (±3\pm 3 kpc)3, and the number of volume elements was 1003100^{3}. The periodic boundary condition along the XYXY plane was implemented. The Lyα\alpha RT simulations were performed for H II regions and Lyα\alpha cooling sources, separately, to calculate the scattering rate PαP_{\alpha} in every volume element. The spin temperature was then calculated using Equation (22).

Figure 20 shows the resulting average vertical profiles of spin temperature for the WNM and CNM phases, respectively, in the top and bottom panels. In the figure, we show the results when Lyα\alpha photons originating either from H II regions (red) or cooling hot gas (blue) are separately taken into account. The profiles obtained when both sources are included (black) and no Lyα\alpha source is considered (green) are also shown. First, we find that the Lyα\alpha pumping (black) is strong enough to make the spin temperature of the WNM approach very close to the kinetic temperature in a region (|z|1kpc|z|\lesssim 1\,{\rm kpc} 2H3\approx 2H_{3}) that contains most of the WNM. However, this is not the case when no Lyα\alpha is considered (green). Second, H II regions (red) are, in general, the most significant Lyα\alpha source for the WF effect. At high altitudes (|z|1kpc|z|\gtrsim 1\,{\rm kpc}), however, cooling Lyα\alpha photons (blue) originating from the WIM can also play a vital role in raising the spin temperature. Third, as expected, the spin temperature of the CNM is virtually the same as the kinetic temperature, even without Lyα\alpha, up to a height (|z|0.5kpc|z|\approx 0.5\,{\rm kpc}) where the CNM resides.

However, we note that the spin temperature in the WNM still differs from the kinetic temperature, by tens of percents as one moves away from the midplane. To examine whether this deviation matters in the context of 21-cm observations, we calculated the “observed” spin temperature, at a velocity channel corresponding to the 21-cm line center, along the vertical direction. Using the synthesis method of Kim et al. (2014), the observed spin temperature was estimated to be Ts,obs=7705±715T_{{\rm s,obs}}=7705\pm 715 K. We, therefore, conclude that the deviation at high altitudes does not give a significant observational effect. This is because low-density elements at high altitudes have a negligible optical depth at 21 cm, and thus their contribution to the observed spin temperature is minor (see Equation (9) of Kim et al. 2014).

Figure 21 shows the angle-averaged emergent Lyα\alpha spectra when Lyα\alpha photons originate from H II regions, cooling, and both, as for Figure 20. The emergent spectra were obtained by averaging the spectra of Lyα\alpha photons escaping from both boundaries at z=±3z=\pm 3 kpc. The Lyα\alpha production rates and the escape fractions of Lyα\alpha photons are also shown in the figure. Because the ISM in this model is static, the spectrum originating from H II regions is doubly-peaked. However, the Lyα\alpha line center is not entirely suppressed since the WIM occupies most of the volume at high altitudes (|z|0.4|z|\gtrsim 0.4 kpc).888Gronke et al. (2016, 2017) derived a formula for a minimum number of clumps along a random sightline, required to suppress the flux at the line center. In their model, all clumps have the same density, while, in our ISM model, the clump’s density depends on the distance from the midplane; thus, the criterion is not, strictly speaking, appropriate to our model. However, the criterion appears to help understand the non-zero flux at the line center in Figure 21. For our model, the mean number of clumps, measured outside the FWHM of the Lyα\alpha source along the vertical direction, is found to be fc7f_{{\rm c}}\approx 7; this is lower than the threshold fc,crit2aτ0,cl/(3π1/4)12f_{{\rm c,\,{\rm crit}}}\approx 2\sqrt{a\tau_{0,{\rm cl}}}/(3\pi^{1/4})\approx 12, calculated using the “mean” optical thickness τ0,cl\left\langle\tau_{0,{\rm cl}}\right\rangle of clumps. Thus, the flux at the line center is not entirely removed. A substantial portion of cooling Lyα\alpha photons can escape relatively freely out of the medium, since they were produced in high altitudes, before suffering enough scatterings to change their frequencies significantly; therefore, the emergent spectrum from cooling Lyα\alpha shows a single peak at the line center. It is noteworthy that the cooling Lyα\alpha radiation is an efficient source to thermalize the hydrogen atoms in high altitudes, even though its luminosity is less than 10% of that originating from H II regions. We also note that the escape fraction of Lyα\alpha photons is about 8.2%, as denoted in the figure, and most of the escaped Lyα\alpha photons originate from H II regions.

Refer to caption
Refer to caption
Figure 22.— Density and temperature slices for a snapshot of a fiducial TIGRESS model at t=393t=393 Myr. From the first to the fourth panel are shown the total hydrogen density (ρ\rho), neutral hydrogen density (ρneu\rho_{{\rm neu}}), kinetic temperature (TKT_{{\rm K}}), and spin temperature (TsT_{{\rm s}}), respectively. The velocity field of fluid is also displayed overlaid on the second panel. The velocity scale is shown at the top of the second panel.
Refer to caption
Refer to caption
Figure 23.— Density and temperature slices for a snapshot of a fiducial TIGRESS model at t=464t=464 Myr. From the first to the fourth panels are shown the total hydrogen density, neutral hydrogen density, kinetic temperature, and spin temperature, respectively. The velocity field of fluid is also displayed overlaid on the second panel. The velocity scale is shown at the top of the second panel.
Refer to caption
Figure 24.— Spin temperature calculated for a snapshot of a fiducial TIGRESS model at t=393t=393 Myr. The top panels show the spin temperature as a function of height and the bottom panels the spin temperature vs the kinetic temperature. From left to right, the panels show the spin temperature (a) when no Lyα\alpha pumping is considered, (b) when only Lyα\alpha photons originating from H II regions are taken into consideration, (c) when cooling Lyα\alpha photons are the only pumping source, and (d) when both Lyα\alpha sources are taken into account. The white dots and lines represent the mode and mean of the spin temperature, weighted with the neutral hydrogen density, at each abscissa bin, respectively. The colors represent the 2D cumulative distribution of occurrence of (z,Ts)(z,T_{{\rm s}}) or (TK,Ts)(T_{{\rm K}},T_{{\rm s}}). The contours denote the locations that discriminate cumulative occurrences of Pc=P_{c}= 0.05, 0.3, 0.6, and 0.9; the outermost contour corresponds to Pc=0.05P_{c}=0.05.
Refer to caption
Figure 25.— Spin temperature calculated for a snapshot of a fiducial TIGRESS model at t=464t=464 Myr. The top panels show the spin temperature as a function of height and the bottom panels the spin temperature vs the kinetic temperature. From left to right, the panels show the spin temperature (a) when no Lyα\alpha pumping is considered, (b) when only Lyα\alpha photons originating from H II regions are taken into consideration, (c) when cooling Lyα\alpha photons are the only pumping source, and (d) when both Lyα\alpha sources are taken into account. The white dots and lines represent the mode and mean of the spin temperature, weighted with the neutral hydrogen density, at each abscissa bin, respectively. The colors and contours show the 2D cumulative frequency of occurrence, as in Figure (24). The contours denote the locations that discriminate cumulative occurrences of Pc=P_{c}= 0.05, 0.3, 0.6, and 0.9.
Refer to caption
Figure 26.— The emergent spectra obtained from the same snapshot (at t=363t=363 Myr) as that used in Figure 24. The red and blue lines show the Lyα\alpha spectra originating from H II regions and collisional processes, respectively. The black line represents the total Lyα\alpha spectrum. The escape fraction and production rate for each source are also shown.
Refer to caption
Figure 27.— The emergent spectra calculated for the same snapshot (at t=464t=464 Myr) as that used in Figure 25. The red and blue lines show the Lyα\alpha spectra originating from H II regions and collisional processes, respectively. The black line represents the total Lyα\alpha spectrum. The escape fraction and production rate for each source are also shown.

5.4. The TIGRESS Model

We now use a multiphase ISM model simulated by the Three-phase Interstellar Medium in Galaxies Resolving Evolution with Star Formation and Supernova Feedback (TIGRESS) framework (Kim & Ostriker, 2017) to investigate the spin temperature of the WNM. In the TIGRESS framework, the ideal magnetohydrodynamics equations are solved in a local, shearing box, representing a small patch of a differentially rotating galactic disk. The framework includes self-gravity, star formation (followed by sink particles), and stellar feedback in the forms of FUV photoelectric heating and supernova and yields realistic, fully self-consistent 3D ISM models with self-regulated star formation. The resulting simulation shows realistic gas properties, including mass/volume fractions, turbulence velocities, and scale heights of cold, warm, and hot phases. Furthermore, self-consistently modeled supernova feedback enables to produce inflow/outflow cycles in the extraplanar region via warm fountains and hot winds (Kim & Ostriker, 2018; Vijayan et al., 2019). Both mean and turbulent magnetic fields are generated and saturated through galactic dynamos, reproducing realistic polarized dust emission maps (Kim et al., 2019).

In the present work, we utilize a fiducial ISM model suitable for the solar neighborhood with a uniform spatial resolution of 8 pc999The fiducial resolution of the solar neighborhood model is 4 pc, which is utilized in other work. Due to computationally demanding RT calculations performed here, we use a lower resolution run, in which all global properties are well converged (see Kim & Ostriker, 2017 for convergence study). Even with 4 pc resolution, the CNM structure is not well resolved (e.g., Gong et al., 2018; Kim et al., 2019; Mao et al., 2019). However, the WF effect on the WNM is well converged.. The simulation box size is Lx=Ly=1024L_{x}=L_{y}=1024 pc and Lz=7168L_{z}=7168 pc. The shearing-periodic boundary conditions were applied when photon packets cross the simulation domain horizontally. We subtract a velocity difference due to the galactic differential rotation, vshear=qΩLx𝐲^\textbf{v}_{{\rm shear}}=-q\Omega L_{x}\hat{\mathbf{y}}, when crossing the horizontal boundaries. Here, we adopt a flat rotation curve with the shear parameter q|dlnΩ/dlnR|=1q\equiv\left|d\ln\Omega/d\ln R\right|=1 and galactic rotational speed Ω=28\Omega=28 km s-1 kpc-1. We apply the same methods to inject Lyα\alpha photons as explained in Section 5.1. We also assumed that Lyα\alpha photons comove with the fluid elements from which they originate.101010The results, for the two snapshots analyzed in this paper, were significantly altered when Lyα\alpha photons were injected from a lab frame with a zero velocity. In that case, the emergent spectra were found to have a single peak at the line center. Such single-peak spectra were obtained for the snapshots that have a mean fluid velocity of v90\left\langle v\right\rangle\gtrsim 90 km s-1 in the galactic plane. Non-comoving Lyα\alpha photons emitted from fluid elements with a large velocity escape the system relatively quickly without experiencing significant frequency shifts and consequently yield a single-peak spectrum. It is physical to assume photons comoving with their sources, and indeed found to be essential; otherwise, we see single peaked spectra that are inconsistent with observations of the star-forming galaxies. In principle, we can utilize star cluster particles formed and followed in the simulation to inject Lyα\alpha photons more self-consistently. Although the current TIGRESS framework does not follow ionizing radiation dynamically, the EUV/FUV radiation field is post-processed using the adaptive ray tracing (Kim et al., 2018) to study ionized gas distribution in the solar neighborhood model (Kado-Fong et al., 2020). We defer more comprehensive investigation using those snapshots in the subsequent papers.

Figures 22 and 23 show the slices of the total hydrogen density, neutral hydrogen density, kinetic temperature, and spin temperature at X=0X=0 kpc for two snapshots at t=393t=393 Myr and 464 Myr, respectively. The fluid velocity fields are also shown. The spin temperature was obtained by taking into account the Lyα\alpha-pumping effect due to H II regions and cooling gas. We have analyzed 8 snapshots over one orbit time of evolution (\sim 224 Myr). These two snapshots were chosen to present typical results from the Lyα\alpha RT for two distinct conditions realized in the simulation. The snapshot at t=393t=393 Myr (Figure 22) shows a breakout of hot-bubble driven by recent supernova explosions, resulting in a large volume filling factor of hot gas near the midplane (fV,hot0.4f_{V,\,{\rm hot}}\sim 0.4) and vertically extended neutral medium with a scale height of H500H\sim 500 pc. In the simulation, star formation activities have been low since then. The neutral medium has fallen back (H200H\sim 200 pc) and filled up most of the volume near the midplane (fV,hot0.1f_{V,\,{\rm hot}}\sim 0.1) at t=464t=464 Myr (Figure 23). Consequently, the midplane density of the neutral medium is lower at t=393t=393 Myr than at t=464t=464 Myr. The snapshot at t=393t=393 Myr also has a high-velocity tail in the velocity distribution measured near the midplane and in general higher fluid velocities.

Figures 24 and 25 show the spin temperature calculated for the two snapshots at t=393t=393 Myr and 464 Myr, respectively. In the figures, the top panels show the spin temperature as a function of the Galactic altitude, and the bottom panels compare the spin temperature with the gas kinetic temperature. The first to the fourth columns, respectively, show the spin temperature obtained (1) when no Lyα\alpha pumping is applied at all, (2) when H II regions are the only Lyα\alpha source, (3) when Lyα\alpha photons originate only from collisional cooling, and (4) when H II regions and collisional processes are both taken into account.

In the figures, the white dots and curves show the mode and mean of the spin temperature weighted with the density of neutral hydrogen at each abscissa bin, respectively. The colors represent the 2D cumulative distribution of occurrence of a pair (x,y)=(z,Ts)(x,y)=(z,T_{{\rm s}}) or (TK,Ts)(T_{{\rm K}},T_{{\rm s}}). To build the 2D cumulative distribution, we first constructed a 2D histogram Hx,yH_{x,y} of the number of volume elements in the 2D space (x,y)(x,y). The 2D histogram was made to have the number of bins 100×100100\times 100 and then smoothed with a Gaussian function with σ=1.2\sigma=1.2 bins. The 2D cumulative distribution was then obtained by calculating Pc(x,y)=i,jHi,jP_{c}(x,y)=\sum_{i,j}H_{i,j}, where the summation was performed over the 2D bins (i,j)(i,j) satisfying Hi,jHx,yH_{i,j}\leq H_{x,y}, and by normalizing it with the maximum of Pc(x,y)P_{c}(x,y). Figures 24 and 25 show the contours that correspond to Pc=P_{c}= 0.05, 0.3, 0.6, and 0.9.

As gas density drops at high altitudes, the spin temperature appears to be very low at high altitudes when no Lyα\alpha pumping is considered. As shown in the first bottom panels, the spin temperature traces well the kinetic temperature in the low-temperature regime of TK30004000T_{{\rm K}}\lesssim 3000-4000 K, even without Lyα\alpha pumping, but traces very poorly at higher temperatures. The spin temperature decreases down to a few hundred Kelvin (even lower than 100 K at some locations) as the kinetic temperature increases (TK5000T_{{\rm K}}\gtrsim 5000 K). This trend is because the gas density decreases to a too low level to bring the spin temperature to the kinetic temperature with collisional pumping alone as the height increases above 200\sim 200 pc. The result would lead to difficulty in discriminating the thermally stable WNM from the thermally unstable H I gas.

When we include the indirect pumping effect by Lyα\alpha photons originating from H II regions, the majority of the WNM is found to have spin temperatures that are very close to the kinetic temperature, as seen in the second bottom panels of Figures 24 and 25. However, a minor fraction of the WNM with TK6000T_{{\rm K}}\gtrsim 6000 K at high altitudes (|z|500|z|\gtrsim 500 pc) can still have very low spin temperatures (Ts1000T_{{\rm s}}\lesssim 1000 K) as shown in the snapshot at t=t= 464 Myr (Figure 25). This result indicates that a substantial portion of Lyα\alpha photons is absorbed by dust grains in the galactic plane, and thus not all Lyα\alpha photons can reach high altitudes.

As shown in the third columns of Figures 24 and 25, cooling Lyα\alpha photons can also pump the spin temperature of the WNM up to a few thousand Kelvin, although not high enough. A substantial fraction of cooling Lyα\alpha photons are produced at high altitudes and relatively free from the destruction by dust grains, and thus are fairly efficient in thermalizing the spin temperature of the WNM, although the Lyα\alpha production rate is as low as 24\sim 2-4% of that by H II regions.

The last columns of Figures 24 and 25 show the final spin temperature as Lyα\alpha photons from both H II regions and collisional cooling gas play roles in raising the spin temperature. The degree of thermalization in Figure 25 is less than perfect compared to the case in Figure 24, which is mainly due to the enhanced dust absorption of the particular snapshot at t=464t=464 Myr. Still, in the snapshot at t=464t=464 Myr, only a minor fraction of the WNM appears to have more or less low spin temperatures, but not as low as Ts1000T_{{\rm s}}\lesssim 1000 K. In the additional six snapshots we analyzed (not presented here), the thermalization by the WF effect is efficient enough to make the spin temperature as close as the kinetic temperature for the WNM (mostly due to Lyα\alpha photons from H II regions, but cooling Lyα\alpha photons can maintain Ts>1000T_{s}>1000 K at high altitudes).

Figures 26 and 27 present the average Lyα\alpha spectra emerging from the top and bottom of the simulation domain for the same snapshots as those for Figures 24 and 25, respectively. In the figures, the emergent Lyα\alpha spectra originating from H II regions and cooling gas are shown in red and blue, respectively. The total spectra, including both sources, are shown in black. The escape fractions of Lyα\alpha photons are also denoted in the figures. As noted in the case of the Ferriere-ISM model of Section 5.3, the escape fraction of cooling Lyα\alpha photons is, in general, higher than that from H II regions. The spectra obtained from both snapshots are asymmetric in the sense that the red part is stronger than the blue part. However, the blue part did not wholly disappear, and it is still strong.

There are three differences between Figures 26 and 27. First, the spectrum emerging from the snapshot of t=393t=393 Myr is found to be in general narrower, but have more extended wings, than that from the snapshot at t=464t=464 Myr. Second, the escape fraction (fescLyα=f_{{\rm esc}}^{{\rm Ly}\alpha}= 23.6%) from the first snapshot is as more than twofold higher than that (fescLyα=f_{{\rm esc}}^{{\rm Ly}\alpha}= 11.5%) in the second snapshot. Third, the average number of scatterings (Nscatt=2.97×106N_{{\rm scatt}}=2.97\times 10^{6}) in the first snapshot is slightly lower than that (Nscatt=3.29×106N_{{\rm scatt}}=3.29\times 10^{6}) in the second. These differences are mainly due to the difference in the scale height. The snapshots at t=393t=393 and 464 Myr is peak and dip in the scale height evolution; therefore, the snapshot at t=464t=464 Myr has slightly higher densities, lower temperatures, and slower fluid velocities in the midplane, where most scattering events occur, compared to the snapshot at t=393t=393 Myr. These differences are shown in Figures 22 and 23. The effects gave rise to a broader spectrum and stronger absorption of Lyα\alpha photons by dust grains for the snapshot at t=464t=464 Myr. The extended wings in the spectrum emerging from the snapshot at t=393t=393 Myr (Figure 26) are mainly due to a high-velocity tail in the velocity distribution that happened in the midplane of that snapshot. These two snapshots bracket the characteristics shown in the additional six snapshots analyzed (but not shown). The escape fraction was estimated to be in general fescLyα720%f_{{\rm esc}}^{{\rm Ly}\alpha}\approx 7-20\%.

One other noteworthy aspect is that we had to use at least more than (27)×107(2-7)\times 10^{7} photon packets, depending on a snapshot, to calculate the scattering rate for the TIGRESS model with a spatial resolution of 8 pc. With only 10710^{7} or fewer photon packets, the calculated scattering rate was not fully converged, in a statistical sense, at high altitudes, and the resulting spin temperature was found to be underestimated, compared to those obtained with a larger number of photon packets, in a considerable portion of the simulation box. The emergent spectrum and escape fraction converged within \sim2% with only 10710^{7} photons. We used 10810^{8} photon packets to obtain Figures 24 to 27.

6. DISCUSSION

The WF effect has been extensively studied to use the 21-cm signal to investigate the epoch of reionization (e.g., Furlanetto et al., 2006). In this context, a Monte-Carlo RT method was used in Semelin et al. (2007) and Baek et al. (2009). However, they only counted the number of scatterings of Lyα\alpha photons to calculate the 21-cm spin temperature. No Monte-Carlo Lyα\alpha RT simulation has been performed, as far as we know, to study the effect of recoil on the slope of the Lyα\alpha spectrum or to understand the WF effect in the context of the ISM. The present study provides the most unambiguous results on the necessary condition for the WF effect. In the following, we discuss some topics that are relevant to our results.

6.1. The critical optical depth for the WF effect

Deguchi & Watson (1985) found that a Sobolev optical depth of τLVG105106\tau_{{\rm LVG}}\approx 10^{5}-10^{6} will bring the color temperature to the kinetic temperature. In order to solve the RT equation, they assumed the LVG condition, the so-called Sobolev approximation. The Sobolev optical depth is a measure of the optical thickness of the resonance zone over which the velocity gradient in a moving medium induces a velocity difference (ΔV\Delta V) equal to the thermal velocity (vthv_{{\rm th}}), along a given ray, as follows:

τLVGnHχ0(να/c)|dV/ds|.\tau_{{\rm LVG}}\equiv\frac{n_{{\rm H}}\chi_{0}}{(\nu_{\alpha}/c)\left|dV/ds\right|}. (44)

The resonance zone can be regarded as a stationary patch of the medium in the comoving frame. Therefore, the result of Deguchi & Watson (1985) indicates that the minimum optical depth for the WF effect is τ0105106\tau_{0}\approx 10^{5}-10^{6}, which is too large compared to ours. In the LVG method, a resonance line at any location (i.e., resonance zone) in the cloud is assumed to be completely decoupled from resonance at all other locations in the cloud due to point-to-point velocity differences. In other words, photons should not be diffused out more than one Doppler width ΔνD\Delta\nu_{{\rm D}} in frequency in a local resonance zone. Otherwise, the photons would have opportunities to be absorbed at other locations due to the velocity gradient. The frequency shift due to radiative diffusion in a local resonance zone with an optical depth τ0\tau_{0} is given by |x|(aτ0)1/3|x|\approx\left(a\tau_{0}\right)^{1/3} (Neufeld, 1990). Thus, the condition for the LVG method should be |x|1|x|\lesssim 1 or τLVG1/a=2.12×103(TK/104K)1/2\tau_{{\rm LVG}}\lesssim 1/a=2.12\times 10^{3}(T_{{\rm K}}/10^{4}\,\text{K})^{1/2}. This condition is not self-consistent with τLVG105106\tau_{{\rm LVG}}\approx 10^{5}-10^{6} and indicates that the LVG method is not adequate for the problem to find the minimum optical depth for the WF effect.

Meiksin (2006) numerically solved the RT equation for resonant photons in various approximate forms to find the time-scale over which the color temperature relaxes to the kinetic temperature of the gas, and obtained results, in some solutions, that the color temperature approaches the kinetic temperature. However, his solutions showed weird oscillatory behaviors near the line center, most likely due to adopted approximations and unstable numerical methods. Roy et al. (2009) made use of a reliable numerical technique to solve the time-dependent RT equation for a homogeneous and isotropically expanding infinite medium. They found that the color temperature of Lyα\alpha becomes close to the kinetic temperature after photons have undergone 104\sim 10^{4} scatterings for TK=10T_{{\rm K}}=10 K and 105\sim 10^{5} scatterings for TK=100T_{{\rm K}}=100 K (see their Figure 4). These critical optical depths are 102103\sim 10^{2}-10^{3} times higher than ours. It is not clear what caused this discrepancy.

Using a Monte-Carlo method, Higgins & Meiksin (2012) calculated the Lyα\alpha spectral shape inside the scattering medium to verify their numerical method. However, no Monte-Carlo simulation was performed to investigate the effect of recoil. Instead, they numerically solved the RT equation, including the recoil effect, for an expanding medium with TK=10T_{{\rm K}}=10 K and mentioned, with regard to their Figure 10, that the expected spectral gradient is recovered across the line center; no detailed analysis, however, was made about the basic requirements for the WF effect.

6.2. Turbulence Effect

In Section 4, we investigated the effect of turbulent motion on the Lyα\alpha spectrum within the medium of atomic hydrogen. Spectroscopic observations of many molecular clouds have revealed that the internal velocity dispersion (vturbv_{{\rm turb}}, deduced from the line width) of each region is well correlated with its size (LL) and mass, and these correlations are approximately of power-law form (Larson, 1981; Heyer & Brunt, 2004), as follows:

vturb=1.10(L/pc)0.38kms1.v_{{\rm turb}}=1.10\left(L/\text{pc}\right)^{0.38}\ \text{km}\,\text{s}^{-1}. (45)

The above size-line width dispersion gives the ratio of the internal velocity dispersion to the thermal dispersion of clouds in terms of the optical thickness τ0=(χ0/ΔνDπ)nHL\tau_{0}=\left(\chi_{0}/\Delta\nu_{{\rm D}}\sqrt{\pi}\right)n_{{\rm H}}L for Lyα\alpha photons:

vturbvth=9×103(TK104K)0.31(τ0/500nH/cm3)0.38.\frac{v_{{\rm turb}}}{v_{{\rm th}}}=9\times 10^{-3}\left(\frac{T_{{\rm K}}}{10^{4}\,\text{K}}\right)^{-0.31}\left(\frac{\tau_{0}/500}{n_{{\rm H}}/\text{cm}^{-3}}\right)^{0.38}. (46)

Therefore, in most astrophysical cases, the turbulent velocity dispersion of a cloud with an optical depth of τ0500\tau_{0}\approx 500 appears to be much smaller than the thermal velocity dispersion. This relation implies that a region corresponding to τ0500\tau_{0}\approx 500 can be almost always considered to be coherent in velocity. In other words, a coherent zone in which the point-to-point velocity fluctuation is less than the thermal velocity dispersion would have an optical thickness much larger than τ0500\tau_{0}\approx 500. Thus, turbulent motions in most, if not all, astrophysical circumstances can be regarded as the macroturbulence case concerning the WF effect. This implies that the turbulence motion should not be taken into account in calculating the color temperature of Lyα\alpha radiation.

6.3. Lyα\alpha Sources and Escape Fraction

In Section 5, we considered two types of Lyα\alpha sources: (1) Lyα\alpha photons originating from H II regions near the galactic plane and (2) cooling radiation from the ionized gas. In order to calculate the emissivity for the ionized gas, we assumed the CIE condition. Concerning the spatial distribution of H II regions, we assumed a Gaussian distribution in the vertical direction. However, in reality, H II regions would be mostly clustered around the OB associations. The H II regions surrounded by dense neutral gas clouds would then produce Lα\alpha photons that are eventually absorbed locally by the adjacent clouds. Lallement et al. (2011) used the Voyager measurements to detect Lyα\alpha emission from our Galaxy. They estimated the escape fraction of the Lyα\alpha radiation from brightest H II regions to be on the order of 3%, but it was highly spatially variable. This effect might strongly suppress the strength of the Lyα\alpha radiation from H II regions, and consequently the coupling of Lyα\alpha with the 21-cm spin temperature would be less effective than we found in this paper.

However, the WIM, which is known to be produced by the leakage of Lyc photons originating from H II regions to the diffuse, low-density regions (Haffner et al., 2009), could be another important source of Lyα\alpha photons. In Figures 24 and 25, it is shown that Lyα\alpha radiation from cooling gas under the CIE condition could be an efficient, though not perfect, source to raise the spin temperature of the WNM in the regions where Lyα\alpha photons from H II regions could not reach. The luminosity of diffuse Hα\alpha radiation is known to occupy 2050\sim 20-50% of the total Hα\alpha luminosity in star-forming galaxies. If the diffuse Hα\alpha radiation originates from the WIM, the same WIM should also produce Lyα\alpha photons that are comparable, in amount, with the Lyα\alpha photons from H II regions (for an alternative explanation on the origin of the diffuse Hα\alpha, see Witt et al., 2010; Seon & Witt, 2012). This amount is much higher than that estimated to be produced by the cooling gas in the CIE condition. Therefore, Lyα\alpha photons originating from the WIM could play a significant role in thermalizing the spin temperature of the WNM. In addition to these sources, cosmic rays can also produce Lyα\alpha, both by direct excitation by suprathermal secondary electrons and by collisional ionization followed by recombination. As a consequence of all these Lyα\alpha sources, we expect that the WF effect would be efficient, at least in the Milky Way-like galaxies.

The escape fraction (fescLyαf_{{\rm esc}}^{{\rm Ly}\alpha}) and spectral shape of the emergent Lyα\alpha radiation is highly sensitive to the density structure and kinematic properties of the neutral ISM, including dust. It is, therefore, worthwhile to compare the escape fraction predicted from our models with the observations of external galaxies. Hayes et al. (2011) compiled the observational results from the literature on the escape fraction and derived an average relationship between fescLyαf_{{\rm esc}}^{{\rm Ly}\alpha} and the color excess E(BV)E(B-V), i.e., fescLyα=0.445×105.52E(BV)f_{{\rm esc}}^{{\rm Ly}\alpha}=0.445\times 10^{-5.52E(B-V)}. However, observational data show a rather large variation in fescLyαf_{{\rm esc}}^{{\rm Ly}\alpha} for a given E(BV)E(B-V). E(BV)E(B-V) for our TIGRESS snapshots is 0.15\approx 0.15 and it gives fescLyα7%f_{{\rm esc}}^{{\rm Ly}\alpha}\sim 7\%. Our results (fescLyα720%f_{{\rm esc}}^{{\rm Ly}\alpha}\approx 7-20\%) are slightly higher than this value, but yet within the variation range observed from external galaxies.

6.4. The Escape Probability Method

Shaw et al. (2017) claimed that the common assumptions that the “excitation” temperature (TexcT_{{\rm exc}}) for Lyα\alpha transition in hydrogen atoms traces the kinetic temperature (TKT_{{\rm K}}) of gas, and the Lyα\alpha radiation drives the 21-cm spin temperature to the kinetic temperature are not valid. It is, therefore, worthwhile to clarify what made their results differ from ours. We first stress that TexcT_{{\rm exc}} is defined by the relative population between the 2p2p and 1s1s states of hydrogen atoms; on the other hand, the “color” temperature (TαT_{\alpha}) of Lyα\alpha radiation refers to the slope of Lyα\alpha spectrum (or the occupation number density). The excitation temperature becomes the same as the color temperature if the 2p2p state is mainly populated by the Lyα\alpha resonance scattering, rather than by the balance between recombination into 2p2p and the radiative (and collisional) decay into 1s1s.

In Shaw et al. (2017), Lyα\alpha photons were assumed to be produced by recombination following photoionization and cosmic ray ionization or by collisional de-excitation. They adopted “the first-order escape probability” method to derive the level populations of 1S1/221\,{}^{2}S_{1/2} and 2P1/2,3/2o22\,{}^{2}P_{1/2,3/2}^{{\rm o}} states by balancing recombination and de-excitation. The Lyα\alpha excitation temperature was then determined by the level population between the two states. In the escape probability method, any transfer of photons in space and frequency is ignored (Rybicki, 1984; Hubeny, 2001). Lyα\alpha photons either travel zero distance and are immediately re-absorbed (eventually destroyed by dust grains) at the same position or escape the medium altogether in a single flight. In this sense, the escape probability method is also called “the dichotomous model” or “the normalized on-the-spot approximation.” In other words, only the locally-created Lyα\alpha photons are taken into account in estimating the level populations and no diffusion in space and frequency due to the RT effect is taken into consideration. They also appear to investigate the effect of Lyα\alpha pumping by the attenuated external, stellar radiation field, but no RT of the external continuum photons was considered. The discrepancy between our results and those of Shaw et al. (2017) is primarily because they did not take into account the RT effect.

A (frequency) redistribution function, which gives the probability that an incoming photon of frequency ν\nu^{\prime} is scattered as an outgoing photon of frequency ν\nu, is, in general, a linear combination of the distribution functions for coherent scattering and complete redistribution in the atom’s frame (Hubeny & Mihalas, 2014). The redistribution functions for coherent scattering and complete redistribution are designated as RIIR_{{\rm II}} and RIIIR_{{\rm III}}, respectively (Hummer, 1962). The escape probability method is based on the complete redistribution function, which is valid for describing the complete decorrelation (or reshuffling) between the absorbed and re-emitted frequencies in a collision-dominant, high-density medium. The branching ratio between them is given by a ratio of the elastic collision rate to the spontaneous emission rate. The density in the ISM is much lower than the critical density at which the collisional de-excitation rate is equal to the rate of spontaneous radiative decay, as discussed in Dijkstra et al. (2006). The rate at which excited hydrogen atom is perturbed is at most 0.10.1 s-1, even at the electron density of 102\sim 10^{2} cm-3, which is nine orders of magnitude below the spontaneous decay rate 108\sim 10^{8} s-1. Therefore, the Lyα\alpha scattering must be coherent in the atom’s rest frame.

It is known that even for a pure RIIR_{{\rm II}} redistribution function without a contribution of the complete redistribution part, the source function is constant with frequency in the line core for line center optical depths in Lyα\alpha larger than about 10310^{3}, as first demonstrated by Hummer (1969). Note that the recoil effect was not included in this RIIR_{{\rm II}} redistribution function. Field (1959b) showed that the Lyα\alpha line profile progressively widens and flattens as scattering occurs repeatedly, if no recoil effect of hydrogen atoms were taken into account. The analytical solution of the RT in Neufeld (1990) for a slab also shows flat spectra. Higgins & Meiksin (2012) numerically solved the RT equation in a static sphere using moment equations and showed that the line profile within the medium is indeed flat at the central portion. We also showed that the line profile is indeed a constant when no recoil effect is considered and the optical depth is larger than 102\sim 10^{2}.

The photon loses energy due to recoil in the rest frame of the atom. This effect yields a slope of h/kBTK-h/k_{{\rm B}}T_{{\rm K}} at the central portion of the Lyα\alpha profile. Adams (1971) speculated that recoil might not play an important role in the escape of resonance radiation. He claimed that, in the observer’s frame, the small recoil redshift might be almost entirely swamped by the Doppler shifts due to the thermal motion of atoms. He also estimated how many scatterings are needed for recoil to affect the escape of Lyα\alpha from a very thick medium. Photons would receive a frequency shift of one Doppler width in each scattering. The recoil shift is about 10410^{4} times smaller than the Doppler shift at TK=104T_{{\rm K}}=10^{4} K, and thus a very large number of scatterings is needed for the recoil effect to significantly shift the photon frequency to the red wing and then boost the escape of Lyα\alpha from the medium. He concluded that recoil can be neglected when Lyα\alpha undergoes less than 5.6×10105.6\times 10^{10} scatterings. However, the effect of recoil is systematic while the Doppler shift is random. This systematic effect, though it would be tiny, is what shapes the Lyα\alpha line profile near the line center, as we demonstrated in Section 4. Shaw et al. (2017) argued that the results of Adams (1971) on the effect of recoil on the Lyα\alpha RT support that the recoil effect is not strong enough to bring the spin temperature to the gas kinetic temperature. However, it should be stressed that the study of Adams (1971) is not about the effect of recoil on the Lyα\alpha spectral shape within the medium, but mainly on the escape of Lyα\alpha from a medium. What matters in the WF effect is a small slope in Lyα\alpha profile near the line center, rather than a significantly large redshift of the Lyα\alpha frequency to the red wing.

7. SUMMARY

In this paper, we have studied the Lyα\alpha RT and WF effect in detail using a Monte-Carlo RT code, named LaRT. LaRT is capable of dealing with not only the WF effect but also the effects due to the fine-structure splitting. LaRT has been successfully used to model the Lyα\alpha spectra and surface brightness profiles of the Lyα\alpha emitting galaxies at z=36z=3-6 (Song et al. 2019, submitted to ApJ). The main aim of the present study was to investigate the WF effect on the spin temperature of the WNM. However, most of our results are applicable in the context of the CGM and IGM. The numerical results of the models will be made available to the community via the authors’ WWW site 111111http://seoncafe.github.io. Our calculations were performed using an unprecedentedly large number of photon packets and thus will be useful for benchmark tests. The principal conclusions of this paper are as follows:

  1. 1.

    As a result of the recoil effect, the color temperature of the Lyα\alpha line profile approaches the gas kinetic temperature even in a medium with as low as τ0100500\tau_{0}\approx 100-500, which is much lower than those predicted in previous studies.

  2. 2.

    Even in an expanding or turbulent medium, the Lyα\alpha spectrum within the medium is fully thermalized to have a color temperature equal to the kinetic temperature due to recoil of hydrogen atom, unless the optical thickness of a coherent region, over which the flow velocity and kinetic temperature can be regarded to be nearly constant, is less than 100500\approx 100-500.

  3. 3.

    We demonstrated that the turbulent motion widens the emergent spectra. However, we showed that the turbulent motion in most astrophysical systems does not change the color temperature of the Lyα\alpha spectrum. In other words, the color temperature of Lyα\alpha is equal to the gas kinetic temperature even in a turbulent medium.

  4. 4.

    To bring the spin temperature to the kinetic temperature, the Lyα\alpha radiation field should also be strong enough. Dust grains may suppress the strength of Lyα\alpha radiation by absorbing them. How Lyα\alpha photons are effectively destroyed or survive in dusty clouds would strongly depend on the density distribution and kinematic properties of the clouds.

  5. 5.

    Utilizing the ISM model of Ferriere (1998) and a MHD simulation using the TIGRESS framework of Kim & Ostriker (2018), we calculated the spin temperature of the WNM in the vicinity of the Sun and found that the resulting spin temperature is reasonably close to the kinetic temperature in the majority of the ISM. In the Lyα\alpha simulations, we included both H II regions and the collisionally cooling gas as Lyα\alpha sources.

  6. 6.

    We found that Lyα\alpha photons originating from sources in relatively dust-free, high altitude regions, such as the collisionally cooling gas, can play a significant role in regions where Lyα photons from H II regions cannot reach due to suppression by dust grains.

  7. 7.

    In addition to the above main results, we developed a new, efficient algorithm to produce the velocity component of scattering atoms that is parallel to the photon propagation direction, as presented in Appendix 28.

  8. 8.

    We calculated the average number of scatterings in an infinite slab and a sphere for a wide range of optical depths from τ0=1\tau_{0}=1 to 10910^{9} and two temperatures of TK=10T_{{\rm K}}=10 K and 10410^{4} K. We also derived a new analytic formula for the average number of scatterings in a spherical medium. New analytic equations for the scattering rate PαP_{\alpha} and Lyα\alpha line profile in a sphere and slab are also provided121212We note that some of these solutions were recently independently derived by Lao & Smith (2020).. These formulae, as shown in Appendixes B and C, well reproduce the Monte-Carlo simulation results at the limit of large optical depth.

This work was supported by a National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) (No. 2017R1A2B4008291 and 2020R1A2C1005788). This work was also supported by the National Institute of Supercomputing and Network/Korea Institute of Science and Technology Information with supercomputing resources including technical support (KSC-2018-S1-0005). The work of CGK was supported in part by the Simons Foundation (CCA 528307, ECO) and in part by NASA (NNX17AG26G). Numerical simulations were partially performed by using a high performance computing cluster at the Korea Astronomy and Space Science Institute. CHIANTI is a collaborative project involving George Mason University, the University of Michigan (USA), University of Cambridge (UK), and NASA Goddard Space Flight Center (USA). We thank the anonymous referee for valuable comments, which helped clarify the manuscript.

Appendix A The Parallel Velocity of Scattering Atom

Refer to caption
Figure 28.— Two-dimensional space 𝒮\mathcal{S} from which the distribution for the parallel velocity component uu of scattering atoms is required to be generated: (a) F(β,γ)dβdγ=f0(γ/β)dβdγF(\beta,\gamma)d\beta d\gamma=f_{0}(\gamma/\beta)d\beta d\gamma or (b) F¯(β,u)dβdu=f0(u)βdβdu\bar{F}(\beta,u)d\beta du=f_{0}(u)\beta d\beta du

The probability distribution function for the velocity component uu of a scattering atom parallel to the incoming photon is

f(u)=aπH(x,a)eu2(ux)2+a2,f(u)=\frac{a}{\pi H(x,a)}\frac{e^{-u^{2}}}{(u-x)^{2}+a^{2}}, (A1)

which is not analytically integrable. In order to generate random numbers that follow the above distribution, the acceptance/rejection method of Zheng & Miralda-Escudé (2002a) usually has been used. However, we found that the algorithm is fairly inefficient, especially when xx is large, and TKT_{{\rm K}} is high (aa is small).

Here, we introduce a new algorithm inspired by the ratio-of-uniforms method (Kinderman & Monahan, 1977). Since the distribution function f(u)f(u) is symmetric under the transformation of (x,u)(x,u)(x,u)\leftrightarrow(-x,-u), the following discussion is limited to x>0x>0.

Suppose that a bivariate random variable (β,γ)(\beta,\gamma) is uniformly distributed while satisfying the following inequality:

0βf(γ/β),0\leq\beta\leq\sqrt{f(\gamma/\beta)}, (A2)

for any nonnegative function f(u)f(u). It can then be proven that u=γ/βu=\gamma/\beta follows a probability density function proportional to f(u)f(u). This theorem can be generalized for a product of two distribution functions. If f(u)=f0(u)f1(u)f(u)=f_{0}(u)f_{1}(u) and (β,γ)(\beta,\gamma) has a probability density proportional to f0(γ/β)f_{0}(\gamma/\beta) over the region of the plane defined by

𝒮={(β,γ):0βf1(γ/β)},\mathcal{S}=\{(\beta,\gamma):0\leq\beta\leq\sqrt{f_{1}(\gamma/\beta)}\}, (A3)

then u=γ/βu=\gamma/\beta follows a density distribution proportional to f(u)f(u).

Refer to caption
Figure 29.— Schematic shape of Q(β)Q(\beta) for the case of (a) x<xc(=1+2)x<x_{c}\ (=1+\sqrt{2}) and (b) x>xcx>x_{c}. The three domains are denoted by S0S_{0}, S1S_{1} and S2S_{2}, respectively. Note that the figure is not in scale and the height h2h_{2} is not always higher than h0h_{0}. In (b), the two areas S0S_{0} and S1S_{1} become negligible and β0,β11\beta_{0},\,\beta_{1}\ll 1 when h2>2h0h_{2}>2h_{0}. The solid line represents the function Q(β)Q(\beta) defined in Equation (A8) and the shaded areas are the piecewise comparison function in Equation (A20).

Let us define two functions to apply the above theorem to our problem:

f0(u)=1π1(ux)2+a2andf1(u)=eu2.f_{0}(u)=\frac{1}{\pi}\frac{1}{(u-x)^{2}+a^{2}}\ \ {\rm and}\ \ f_{1}(u)=e^{-u^{2}}. (A4)

The region 𝒮\mathcal{S} where f0(u)f_{0}(u) is defined is equivalent to the following condition:

|γ|β2lnβor|u|2lnβ,\left|\gamma\right|\leq\beta\sqrt{-2\ln\beta}\ \ {\rm or}\ \ \left|u\right|\leq\sqrt{-2\ln\beta}, (A5)

which is illustrated in Figure 28. In other words, the following 2D distribution of (β,γ)(\beta,\gamma) or (β,u)(\beta,u) should be generated over the area defined by Equation (A5):

F(β,γ)dβdγ\displaystyle F(\beta,\gamma)d\beta d\gamma \displaystyle\equiv f0(γ/β)dβdγ,\displaystyle f_{0}(\gamma/\beta)d\beta d\gamma, (A6)

or

F¯(β,u)dβdu\displaystyle\bar{F}(\beta,u)d\beta du \displaystyle\equiv f0(u)βdβdu.\displaystyle f_{0}(u)\beta d\beta du. (A7)

For our purpose, it is more convenient to deal with two variables (β,u)(\beta,u) instead of (β,γ)(\beta,\gamma).

To make bivariate random variables (β,u)(\beta,u) according to F¯(β,γ)\bar{F}(\beta,\gamma), defined over the light-blue region defined in Figure 28(b), we consider the following one-dimensional marginal distribution of β\beta:

Q(β)\displaystyle Q(\beta) \displaystyle\equiv p(β)p(β)F¯(β,u)𝑑u\displaystyle\int_{-p(\beta)}^{p(\beta)}\bar{F}(\beta,u)du (A8)
=\displaystyle= βaπ[tan1(p(β)xa)+tan1(p(β)+xa)],\displaystyle\frac{\beta}{a\pi}\left[\tan^{-1}\left(\frac{p(\beta)-x}{a}\right)+\tan^{-1}\left(\frac{p(\beta)+x}{a}\right)\right],

where

p(β)2lnβ.p(\beta)\equiv\sqrt{-2\ln\beta}. (A9)

A bivariate random variable (β,u)(\beta,u) can be obtained by selecting a random number β\beta from the distribution function Q(β)Q(\beta) and then generating uu from f0(u)f_{0}(u) over the range of |u|p(β)\left|u\right|\leq p(\beta). The distribution function f0(u)f_{0}(u) is analytically invertible, and thus a random number uu can be easily obtained for the given random variate β\beta. The question that is then raised is how to generate random numbers that follow the density distribution Q(β)Q(\beta).

We generate the random numbers for Q(β)Q(\beta) using an acceptance/rejection method. To use the technique, we need to construct a comparison function C(β)C(\beta) that is similar to, but always higher than, Q(β)Q(\beta). The overall shape of Q(β)Q(\beta) is shown in Figure 29. We first decompose Q(β)Q(\beta) into two domains: p(β)>xp(\beta)>x and p(β)<xp(\beta)<x. Let us define β0\beta_{0}, which divides the domains, as follows:

β0ex2/2.\beta_{0}\equiv e^{-x^{2}/2}. (A10)

For the first domain (S0S_{0}; β<β0\beta<\beta_{0}, p(β)>xp(\beta)>x in Figure 29), we observe that

Q(β)<F¯(β,u)𝑑u=βa.Q(\beta)<\int_{-\infty}^{\infty}\bar{F}(\beta,u)du=\frac{\beta}{a}. (A11)

Here, the last term is a linear function with a slope 1/a1/a. For the second domain (S1+S2S_{1}+S_{2}; β>β0\beta>\beta_{0}, p(β)<xp(\beta)<x in Figure 29), we find the following inequality, since 1/[(ux)2+a2]<1/(ux)21/\left[(u-x)^{2}+a^{2}\right]<1/(u-x)^{2}:

Q(β)\displaystyle Q(\beta) <\displaystyle< Q(β)βπp(β)p(β)du(ux)2\displaystyle Q^{*}(\beta)\equiv\frac{\beta}{\pi}\int_{-p(\beta)}^{p(\beta)}\frac{du}{(u-x)^{2}} (A12)
=2βπp(β)x2p2(β).\displaystyle\ \ \ \ \ \ \ \ \ \ \ =\frac{2\beta}{\pi}\frac{p(\beta)}{x^{2}-p^{2}(\beta)}.

We note that Q(β)Q^{*}(\beta) is very close to Q(β)Q(\beta) in the range of β>β1\beta>\beta_{1} (minimum point in Figure 29(b)), but rapidly diverges to infinity as ββ0\beta\rightarrow\beta_{0}. These inequalities are summarized as follows:

Q(β)\displaystyle Q(\beta) <\displaystyle< {β/aforβ<β0,p(β)>x,Q(β)forβ>β0,p(β)<x.\displaystyle\begin{cases}\beta/a&{\rm for}\ \beta<\beta_{0},\ \ \ p(\beta)>x,\\ \\ Q^{*}(\beta)&{\rm for}\ \beta>\beta_{0},\ \ \ p(\beta)<x.\end{cases} (A13)

A notable point is that Q(β)Q(\beta) has a right triangular shape in the first domain (β<β0\beta<\beta_{0}), as shown in Figure 29. In the second domain (β>β0\beta>\beta_{0}), Q(β)Q(\beta) appears to be more complex. If x<xc1+2x<x_{c}\equiv 1+\sqrt{2}, Q(β)Q(\beta) is a monotonically decreasing function of β>β0\beta>\beta_{0}, as shown in Figure 29(a). On the other hand, if x>xcx>x_{c}, Q(β)Q(\beta) shows a minimum and maximum in the domain S1+S2S_{1}+S_{2}. The existence condition of the minimum (β=β1\beta=\beta_{1} in Figure 29(b)) and maximum (β=β2\beta=\beta_{2}) points in the second domain can be easily derived from dQ/dβ=0dQ^{*}/d\beta=0; the resulting condition is that x>xc=1+2x>x_{c}=1+\sqrt{2}.

For the efficiency in applying the rejection/acceptance method, we further divide the second domain into two subdomains, as denoted by S1S_{1} and S2S_{2} in Figure 29. The basic idea for our comparison function is shown in Figure 29: a triangle for S0S_{0} and two rectangles for S1S_{1} and S2S_{2}. In the case of x<xcx<x_{c}, the location β=β1\beta=\beta_{1}, where S1S_{1} and S2S_{2} are divided, is obtained by minimizing the area S1+S2=(β1β0)h0+(1β1)h1S_{1}+S_{2}=(\beta_{1}-\beta_{0})h_{0}+(1-\beta_{1})h_{1}. Here, h0h_{0}, the half of the height of the right triangle in S0S_{0}, and h1h_{1} are given by, respectively,

h0\displaystyle h_{0} \displaystyle\equiv β02aQ(β0)=β0aπtan1(2xa),\displaystyle\frac{\beta_{0}}{2a}\geq Q(\beta_{0})=\frac{\beta_{0}}{a\pi}\tan^{-1}\left(\frac{2x}{a}\right), (A14)
h1\displaystyle h_{1} \displaystyle\equiv Q(β1).\displaystyle Q^{*}(\beta_{1}). (A15)

After expanding d(S1+S2)/dβd(S_{1}+S_{2})/d\beta in a power series of ββ0\beta-\beta_{0} and aa, and setting the derivative to be zero, we obtain

β1β0+[2aπ(1β0)β0x]1/2.\beta_{1}\equiv\beta_{0}+\left[\frac{2a}{\pi}\left(1-\beta_{0}\right)\beta_{0}x\right]^{1/2}. (A16)

We found that this is also a good choice even for x>xcx>x_{c}; however, this β1\beta_{1} is not the minimum point in S1+S2S_{1}+S_{2}.

In Figure 29(b) (x>xcx>x_{c}), the height h2h_{2} that tightly constrains the true height of S2S_{2} is obtained to be

h2\displaystyle h_{2} \displaystyle\equiv 2e1/2p(e1/2)π1x21.373=0.3861x21.373.\displaystyle\frac{2e^{-1/2}p(e^{-1/2})}{\pi}\frac{1}{x^{2}-1.373}=\frac{0.3861}{x^{2}-1.373}. (A17)

This was found by noticing that β=e1/2\beta=e^{-1/2} yields the maximum of Q(β)Q^{*}(\beta) at the limit of xx\rightarrow\infty (Q(β)=Q(β)2βp(β)/πx2Q(\beta)=Q^{*}(\beta)\rightarrow 2\beta p(\beta)/\pi x^{2}). However, Q(β=e1/2)Q^{*}(\beta=e^{-1/2}) was slightly lower than the true maximum of Q(β)Q(\beta). We, therefore, attempted to find a suitable height h2h_{2} that tightly constrains the maximum of Q(β)Q(\beta). In the end, h2h_{2} in Equation (A17) was obtained by adjusting p(β=e1/2)=1p(\beta=e^{-1/2})=1 in the denominator of Q(β=e1/2)Q^{*}(\beta=e^{-1/2}) to 1.373. This h2h_{2} provides a very tight constraint on the maximum of Q(β)Q(\beta) for any xx.

Refer to caption
Figure 30.— The regions in the (TK,x)(T_{{\rm K}},x) space where we define the three different comparison functions in Equations (A18)-(A20).
Refer to caption
Figure 31.— The distribution of parallel velocities f(u)f(u_{\parallel}) for different values of incoming frequency xx. The black curves denote the histograms for the random numbers generated using the present algorithm. The colored curves show the theoretical distribution functions, as calculated from Equation (A1).

Let us now define our comparison function. It is more convenient to combine two or three domains and define a simpler comparison function depending on h2h_{2}. If h2>2h0h_{2}>2h_{0}, the areas S0S_{0} and S1S_{1} are negligible and we thus combine the three domains as a single domain S0+S1+S2S_{0}+S_{1}+S_{2}. In this case, we choose a single comparison function for the whole domain, as follows:

C(β)h2.C(\beta)\equiv h_{2}. (A18)

If h0<h2<2h0h_{0}<h_{2}<2h_{0}, we combine S1S_{1} and S2S_{2} and consider two domains S0S_{0} and S1+S2S_{1}+S_{2}. This leads us to define the following comparison function:

C(β){β/aifββ0,h2otherwise,C(\beta)\equiv\begin{cases}\beta/a&{\rm if}\ \ \beta\leq\beta_{0},\\ \\ h_{2}&\text{otherwise},\end{cases} (A19)

The areas of the first and second domains are S0=β0h0S_{0}=\beta_{0}h_{0} and S1=(1β0)h2S_{1}=(1-\beta_{0})h_{2}, respectively. The total area of the comparison function is Stot=S0+S1S_{{\rm tot}}=S_{0}+S_{1}.

If h2<h0h_{2}<h_{0} or x<xcx<x_{c}, we define the comparison function to be

C(β){β/aifββ0,h0ifβ0<ββ1,h1ifβ>β1andxxc.max(h1,h2)ifβ>β1andx>xc.C(\beta)\equiv\begin{cases}\beta/a&{\rm if}\ \ \beta\leq\beta_{0},\\ \\ h_{0}&{\rm if}\ \ \beta_{0}<\beta\leq\beta_{1},\\ \\ h_{1}&{\rm if}\ \ \beta>\beta_{1}\ \ \ {\rm and}\ \ \ x\leq x_{c}.\\ \\ \max(h_{1},h_{2})&{\rm if}\ \ \beta>\beta_{1}\ \ \ {\rm and}\ \ \ x>x_{c}.\end{cases} (A20)

In this case, the areas of the first and second domains are S0=β0h0S_{0}=\beta_{0}h_{0} and S1=(β1β0)h0S_{1}=(\beta_{1}-\beta_{0})h_{0}, respectively. The area of the third domain is given by S2=(1β1)h1S_{2}=(1-\beta_{1})h_{1} if x<xcx<x_{c} or S2=(1β0)max(h1,h2)S_{2}=(1-\beta_{0})\max(h_{1},h_{2}) if x>xcx>x_{c}. The total area of the comparison function is Stot=S0+S1+S2S_{{\rm tot}}=S_{0}+S_{1}+S_{2}.

We note that the comparison function in S0S_{0} is a linear function. A random variate for the linear part of the comparison function is obtained using the inversion method; β=β0ξ1/2\beta=\beta_{0}\xi^{1/2} for a uniform random number ξ\xi.

Supposing that ξ1,ξ2,\xi_{1},\xi_{2},\cdots are mutually-independent, uniform random numbers between 0 and 1, our algorithm is given as follows:

  1. 1.

    If h22h0h_{2}\geq 2h_{0}, we choose β\beta as follows:

    β\displaystyle\beta =\displaystyle= ξ1.\displaystyle\xi_{1}. (A21)

    If h0h2<2h0h_{0}\leq h_{2}<2h_{0}, β\beta is obtained by:

    β={β0ξ2ifξ1S0/Stot,β0+(1β0)ξ3otherwise.\beta=\begin{cases}\beta_{0}\sqrt{\xi_{2}}&{\rm if}\ \ \xi_{1}\leq S_{0}/S_{{\rm tot}},\\ \\ \beta_{0}+\left(1-\beta_{0}\right)\xi_{3}&\text{otherwise}.\end{cases} (A22)

    If h2<h0h_{2}<h_{0} or x<xcx<x_{c}, β\beta is given by

    β={β0ξ2ifξ1S0/Stot,β0+(β1β0)ξ3ifS0/Stot<ξ11S2/Stot,β1+(1β1)ξ4otherwise.\beta=\begin{cases}\beta_{0}\sqrt{\xi_{2}}&{\rm if}\ \ \xi_{1}\leq S_{0}/S_{{\rm tot}},\\ \\ \beta_{0}+\left(\beta_{1}-\beta_{0}\right)\xi_{3}&{\rm if}\ \ S_{0}/S_{{\rm tot}}<\xi_{1}\leq 1-S_{2}/S_{{\rm tot}},\\ \\ \beta_{1}+\left(1-\beta_{1}\right)\xi_{4}&\text{otherwise}.\end{cases} (A23)
  2. 2.

    If ξ6Q(β)/C(β)\xi_{6}\leq Q(\beta)/C(\beta), we accept β\beta. Otherwise, go back to step 1 and repeat the procedure until β\beta is accepted.

  3. 3.

    For the accepted β\beta, the parallel velocity component uu is readily obtained by solving the following equation about uu:

    p(β)uF¯(β,u)𝑑u=Q(β)ξ5.\int_{-p(\beta)}^{u}\bar{F}(\beta,u^{\prime})du^{\prime}=Q(\beta)\xi_{5}. (A24)

    This gives the following result:

    u=x+atan[(aπβ)Q(β)ξ7tan1(p(β)+xa)].u=x+a\tan\left[\left(\frac{a\pi}{\beta}\right)Q(\beta)\xi_{7}-\tan^{-1}\left(\frac{p(\beta)+x}{a}\right)\right]. (A25)

If the input xx was negative, we take the negative of uu that is obtained after replacing xx with its absolute value.

We note that h2h_{2} is a function of xx, and h0h_{0} is a function of xx and aa (or TKT_{{\rm K}}). Figure 30 shows the regions in the (TK,x)(T_{{\rm K}},x) space where the three different comparison functions in Equations (A18)-(A20) are defined. Figure 31 compares the distribution of random numbers generated using the algorithm developed in this paper with the theoretical probability distribution function Equation (A1). As shown in the figure, our algorithm reproduces the probability distribution function for the parallel velocity component very well.

Our method seems not to be intuitive to understand at first glance. However, the present algorithm is uniformly efficient for any xx and aa. One may use our algorithm only in the cases of h2>2h0h_{2}>2h_{0} and h0<h2<2h0h_{0}<h_{2}<2h_{0} while adopting the method of Zheng & Miralda-Escudé (2002b) in the case of h2<h0h_{2}<h_{0}. We use the algorithm of Zheng & Miralda-Escudé (2002b) for x<1x<1 (after setting u0=0u_{0}=0), and the present algorithm otherwise.

Appendix B Analytic Solutions for a Static, Uniform Slab

Neufeld (1990) derived an approximate series expression for the mean intensity at any point in an infinite slab:

J(x,τ)\displaystyle J(x,\tau) =\displaystyle= F(ττs2τ0,|σσi|2τ0),\displaystyle F\left(\frac{\tau^{\prime}-\tau^{\prime}_{s}}{2\tau_{0}},\frac{\left|\sigma-\sigma_{i}\right|}{2\tau_{0}}\right), (B1)
F(2τ0ττs2τ0,|σσi|2τ0)\displaystyle-F\left(\frac{2\tau_{0}-\tau^{\prime}-\tau^{\prime}_{s}}{2\tau_{0}},\frac{\left|\sigma-\sigma_{i}\right|}{2\tau_{0}}\right)

where

F(w,y)\displaystyle F(w,y) \displaystyle\equiv 68πn=1[cos(nπw)enπynπ],\displaystyle\frac{\sqrt{6}}{8\pi}\sum_{n=1}^{\infty}\left[\frac{\cos(n\pi w)e^{-n\pi y}}{n\pi}\right], (B2)
τ\displaystyle\tau^{\prime} \displaystyle\equiv τ(123ϕ(x)πτ0)τ(12x23aτ0),\displaystyle\tau\left(1-\frac{2}{3\phi(x)\sqrt{\pi}\tau_{0}}\right)\simeq\tau\left(1-\frac{2x^{2}}{3a\tau_{0}}\right),
σ\displaystyle\sigma \displaystyle\equiv (2π3)1/2x33a,\displaystyle\left(\frac{2\pi}{3}\right)^{1/2}\frac{x^{3}}{3a},
ϕ(x)\displaystyle\phi(x) \displaystyle\equiv 1πH(x,a)aπx2.\displaystyle\frac{1}{\sqrt{\pi}}H(x,a)\simeq\frac{a}{\pi x^{2}}.

Here, the optical depth at the location of the point source is denoted by τs\tau_{s} and the initial frequency xix_{i} is represented by σi=(2π/3)1/2xi3/(3a)\sigma_{i}=(2\pi/3)^{1/2}x_{i}^{3}/(3a). Note that Neufeld (1990) defines the optical depth as the total optical depth integrated over frequency, i.e., τNeufeld=πτ\tau^{{\rm Neufeld}}=\sqrt{\pi}\tau and σNeufeld=πσ\sigma^{{\rm Neufeld}}=\sqrt{\pi}\sigma. We can further simplify F(w,y)F(w,y) as

F(w,y)=616π2ln(12eπycosπw+e2πy).F(w,y)=-\frac{\sqrt{6}}{16\pi^{2}}\ln\left(1-2e^{-\pi y}\cos\pi w+e^{-2\pi y}\right). (B3)

This is obtained by applying the identity cosx=(eix+eix)/2\cos x=\left(e^{ix}+e^{-ix}\right)/2, and then simplifying the summation with ln(1x)=n=1xn/n\ln\left(1-x\right)=-\sum_{n=1}^{\infty}x^{n}/n and coshx=(ex+ex)/2\cosh x=\left(e^{x}+e^{-x}\right)/2. The mean intensity in the slab can then be written as

J(x,τ)=616π2ln[cosh(π|σσi|2τ)+cos(πτ+τs2τ0)cosh(π|σσi|2τ0)cos(πττs2τ0)].J(x,\tau)=\frac{\sqrt{6}}{16\pi^{2}}\ln\left[\frac{\cosh\left(\pi\frac{\left|\sigma-\sigma_{i}\right|}{2\tau}\right)+\cos\left(\pi\frac{\tau^{\prime}+\tau_{s}^{\prime}}{2\tau_{0}}\right)}{\cosh\left(\pi\frac{\left|\sigma-\sigma_{i}\right|}{2\tau_{0}}\right)-\cos\left(\pi\frac{\tau^{\prime}-\tau_{s}^{\prime}}{2\tau_{0}}\right)}\right]. (B4)

Setting τs=0\tau_{s}=0 and xi=σi=0x_{i}=\sigma_{i}=0, the mean intensity becomes:

J(x,τ)=616π2ln[cosh(π|σ|/2τ0)+cos(πτ/2τ0)cosh(π|σ|/2τ0)cos(πτ/2τ0)].J(x,\tau)=\frac{\sqrt{6}}{16\pi^{2}}\ln\left[\frac{\cosh(\pi\left|\sigma\right|/2\tau_{0})+\cos(\pi\tau^{\prime}/2\tau_{0})}{\cosh(\pi\left|\sigma\right|/2\tau_{0})-\cos(\pi\tau^{\prime}/2\tau_{0})}\right]. (B5)

This equation evaluated at τ=τ0\tau=\tau_{0} seems to be different from Equation (2.24) of Neufeld (1990) but gives almost the same numerical results.

The mean intensity at x=0x=0 (i.e., y=σ=0y=\sigma=0 and τ=τ\tau^{\prime}=\tau) is then given by

Jx(0,τ)\displaystyle J_{x}(0,\tau) =\displaystyle= 68π2ln[tan(π4ττ0)]\displaystyle-\frac{\sqrt{6}}{8\pi^{2}}\ln\left[\tan\left(\frac{\pi}{4}\frac{\tau}{\tau_{0}}\right)\right]
\displaystyle\simeq {68π2ln(π4ττ0)forτ/τ01616π(1ττ0)forτ/τ01.\displaystyle\begin{cases}-\frac{\sqrt{6}}{8\pi^{2}}\ln\left(\frac{\pi}{4}\frac{\tau}{\tau_{0}}\right)&{\rm for}\ \ \tau/\tau_{0}\ll 1\\ \\ \frac{\sqrt{6}}{16\pi}\left(1-\frac{\tau}{\tau_{0}}\right)&{\rm for}\ \ \tau/\tau_{0}\approx 1.\end{cases}

As is explained in Section 4, we can use Jx(0,τ)J_{x}(0,\tau) to calculate the scattering rate PαP_{\alpha} (see Equation (27)).

The average number of scatterings that a photon undergoes before escaping the medium is an integral of the scattering rate over total number of atoms in the whole volume:

Nscatt\displaystyle\left\langle N_{{\rm scatt}}\right\rangle =\displaystyle= PαnH𝑑z\displaystyle\int P_{\alpha}n_{{\rm H}}dz
\displaystyle\simeq 4πLL0Jν(0,z)σνnH𝑑ν𝑑z,\displaystyle 4\pi\int_{-L}^{L}\int_{0}^{\infty}J_{\nu}(0,z)\sigma_{\nu}n_{{\rm H}}d\nu dz,

where nHn_{{\rm H}} is the number density of neutral hydrogen atom. Using Equation (B), we can derive the average number of scatterings to be

Nscatt=(46πCπ2)τ0=1.612τ0,\left\langle N_{{\rm scatt}}\right\rangle=\left(\frac{4\sqrt{6\pi}C}{\pi^{2}}\right)\tau_{0}=1.612\tau_{0}, (B8)

where Ck=0(1)k(2k+1)20.915966C\equiv\sum_{k=0}^{\infty}(-1)^{k}(2k+1)^{-2}\simeq 0.915966 is Catalan’s constant. Harrington (1973) also obtained the same result by integrating a series solution. The number of scatterings in a slab geometry was also discussed in Adams (1972).

Appendix C Analytic Solution for a Static, Uniform Sphere

Dijkstra et al. (2006) provide a series solution to the RT equation for a uniform sphere, in which photons are emitted at radius rsr_{s}, assuming the Eddington approximation:

J(x,r)=616π2R2n=1sin(λnrs)sin(λnr)(λnrs)(r/R)eλn|σ|/κ0,J(x,r)=\frac{\sqrt{6}}{16\pi^{2}R^{2}}\sum_{n=1}^{\infty}\frac{\sin\left(\lambda_{n}r_{s}\right)\sin\left(\lambda_{n}r\right)}{\left(\lambda_{n}r_{s}\right)\left(r/R\right)}e^{-\lambda_{n}\left|\sigma\right|/\kappa_{0}}, (C1)

where

λn\displaystyle\lambda_{n} \displaystyle\equiv nπR(111+(3/2)ϕ(x)πτ0),\displaystyle\frac{n\pi}{R}\left(1-\frac{1}{1+(3/2)\phi(x)\sqrt{\pi}\tau_{0}}\right),
σ\displaystyle\sigma \displaystyle\equiv (2π3)1/2x33a\displaystyle\left(\frac{2\pi}{3}\right)^{1/2}\frac{x^{3}}{3a}
ϕ(x)\displaystyle\phi(x) 1π\displaystyle\equiv\frac{1}{\sqrt{\pi}} H(x,a)aπx2.\displaystyle H(x,a)\simeq\frac{a}{\pi x^{2}}.

However, the series solution is not practical for an application. They, therefore, derived an analytic formula for the Lyα\alpha spectrum emerging from the sphere by setting r=Rr=R and simplifying the series solution. In the present study, we need a more general formula for the spectrum of radiation field within the medium at an arbitrary radius. We found that, using the identities sinx=(eixeix)/2i\sin x=\left(e^{ix}-e^{-ix}\right)/2i, ln(1x)=n=1xn/n\ln\left(1-x\right)=-\sum_{n=1}^{\infty}x^{n}/n, cosx=(eix+eix)/2\cos x=\left(e^{ix}+e^{-ix}\right)/2, and coshx=(ex+ex)/2\cosh x=\left(e^{x}+e^{-x}\right)/2, the series solution of J(x,r)J(x,r) can be rewritten as

J(x,r)\displaystyle J(x,r) =\displaystyle= 14πR2616π1q(rs/R)(r/R)\displaystyle\frac{1}{4\pi R^{2}}\frac{\sqrt{6}}{16\pi}\frac{1}{q(r_{s}/R)(r/R)} (C2)
×ln[cosh(q|σ|/τ0)cos(q(r+rs)/R)cosh(q|σ|/τ0)cos(q(rrs)/R)],\displaystyle\times\ln\left[\frac{\cosh\left(q\left|\sigma\right|/\tau_{0}\right)-\cos\left(q(r+r_{s})/R\right)}{\cosh\left(q\left|\sigma\right|/\tau_{0}\right)-\cos\left(q(r-r_{s})/R\right)}\right],

where

qλ1Rπ(12πx23aτ0).q\equiv\lambda_{1}R\simeq\pi\left(1-\frac{2\sqrt{\pi}x^{2}}{3a\tau_{0}}\right).

However, the above equation has a singularity at the first term as rs0r_{s}\rightarrow 0. For the case of rs=0r_{s}=0, we use the asymptotic property sin(x)/x1\sin(x)/x\rightarrow 1 as x0x\rightarrow 0 in the series solution (Equation (C1)) and find an analytic solution to be

J(x,r)\displaystyle J(x,r) =\displaystyle= 14πR268π\displaystyle\frac{1}{4\pi R^{2}}\frac{\sqrt{6}}{8\pi} (C3)
×(Rr)sin(qr/R)cosh(q|σ|/τ0)cos(qr/R).\displaystyle\times\left(\frac{R}{r}\right)\frac{\sin\left(qr/R\right)}{\cosh\left(q\left|\sigma\right|/\tau_{0}\right)-\cos\left(qr/R\right)}.

At the limit of r=Rr=R and τ0\tau_{0}\rightarrow\infty, this equation becomes the same as Equation (C17) of Dijkstra et al. (2006). From the equation, the mean intensity at x=0x=0 is obtained to be

Jx(0,r)\displaystyle J_{x}(0,r) =\displaystyle= 14πR268π(Rr)cot(π2rR)\displaystyle\frac{1}{4\pi R^{2}}\frac{\sqrt{6}}{8\pi}\left(\frac{R}{r}\right)\cot\left(\frac{\pi}{2}\frac{r}{R}\right)
\displaystyle\simeq {14πR264π2(Rr)2forr/R114πR2616(Rr)(1rR)forr/R1,\displaystyle\begin{cases}\frac{1}{4\pi R^{2}}\frac{\sqrt{6}}{4\pi^{2}}\left(\frac{R}{r}\right)^{2}&{\rm for}\ \ r/R\ll 1\\ \\ \frac{1}{4\pi R^{2}}\frac{\sqrt{6}}{16}\left(\frac{R}{r}\right)\left(1-\frac{r}{R}\right)&{\rm for}\ \ r/R\approx 1,\end{cases}

which can be used to estimate the scattering rate PαP_{\alpha} (see Equation (27)). Interestingly, the solution indicates the free streaming radiation in the vicinity of the source (r/R1r/R\ll 1). As in Appendix B, the number of scatterings is the number of scattering events per unit time over the whole medium divided by the generation rate. We derive the average number of scatterings that a photon undergoes before escaping the sphere to be

Nscatt\displaystyle\left\langle N_{{\rm scatt}}\right\rangle =\displaystyle= 4π0RPαnH4πr2𝑑r\displaystyle 4\pi\int_{0}^{R}P_{\alpha}n_{{\rm H}}4\pi r^{2}dr (C5)
\displaystyle\simeq 4π0R0Jν(0,r)nHσν4πr2𝑑ν𝑑r\displaystyle 4\pi\int_{0}^{R}\int_{0}^{\infty}J_{\nu}(0,r)n_{{\rm H}}\sigma_{\nu}4\pi r^{2}d\nu dr
=\displaystyle= (32πln4)τ0=0.9579τ0.\displaystyle\left(\sqrt{\frac{3}{2\pi}}\ln 4\right)\tau_{0}=0.9579\tau_{0}.

This is the solution at the limit of τ0\tau_{0}\rightarrow\infty. We also note that the analytic solutions for the mean intensity and scattering rate in an expanding medium with zero temperature were derived in Loeb & Rybicki (1999) and Higgins & Meiksin (2012).

Appendix D Dust-Absorbed Lyα\alpha Spectrum

Refer to caption
Figure 32.— Emergent and dust-absorbed Lyα\alpha spectra for an infinite slab geometry with a temperature of TK=104T_{{\rm K}}=10^{4} K and the H I optical depth of τ0=106\tau_{0}=10^{6}. The black dotted line represents the emergent Lyα\alpha spectrum when no dust is included. The red, green, and blue curves denote the cases for the dust absorption optical depths of τabs=1.0×102\tau_{{\rm abs}}=1.0\times 10^{-2}, 1.5×1021.5\times 10^{-2}, and 2.0×1022.0\times 10^{-2}, respectively. The solid and dotted lines indicate the emergent and absorbed spectra, respectively.

To understand the origin of the dip feature in the dust-absorbed Lyα\alpha spectra, as shown in Figure 5, we need to consider two competing factors, as pointed out in Verhamme et al. (2006): (1) a relatively, very low probability of being absorbed by dust grains compared to that of the strong H I scattering at the line center (x=0x=0), and (2) a large number of resonance scatterings that photons undergo in the line core, increasing the opportunity of interacting with dust in the end. The ratio of dust extinction to the H I scattering cross-section is proportional to TK1/2T_{{\rm K}}^{1/2} at the Lyα\alpha line center, and thus the relative likelihood of photons to be absorbed by dust grains decreases as temperature decreases. This implies that the first factor is more important at lower temperatures. It can also be readily understood that the first effect would be more important at lower optical depths. As the optical depth increases, the increased number of interactions compensates the inefficiency of dust absorption at the line center and makes the second factor important. Even a slight increase of τabs\tau_{{\rm abs}} could yield a much stronger absorption; the dip feature would then disappear as τabs\tau_{{\rm abs}} increases, as shown in Figure 32.

Figure 32 shows some spectra resulting from attempts to reproduce Figure 3 of Verhamme et al. (2006). The three models in Figure 32 were selected to be reasonably close to their model. We note that Figure 32 assumes much higher dust-absorption optical depths than in Figure 5. In Figure 32, we found no clear dip feature in the absorbed spectra, expect a minor sign of deviation from the perfect flatness at x=0x=0, as opposed to Figure 3 of Verhamme et al. (2006). The cause of the discrepancy is not clear. We also found that the dip feature slightly changes as a less accurate Voigt profile function, for instance, used in Tasitsiomi (2006) and Laursen et al. (2009), is adopted.

Appendix E Physics of the WF Effect

Refer to caption
Figure 33.— The hyperfine energy levels (n=1n=1 and n=2n=2) of hydrogen. The energy levels are designated by LJF{}_{F}L_{J} and labeled in parentheses. The solid red and dashed black lines denote permitted and forbidden transitions, respectively. The solid blue lines are also permitted transitions, but complementary in the sense that they are not relevant to exciting the 21-cm line. The relative line strengths (S51:S41:S40:S31:S30:S21=5:1:2:2:1:1S_{51}:S_{41}:S_{40}:S_{31}:S_{30}:S_{21}=5:1:2:2:1:1) of the six permitted transitions are also denoted. Frequencies between the hyperfine states are also shown.

In the context of the ISM research, the WF effect has not been fully described. It is, therefore, worthwhile to summarize the fundamental physics of the WF effect to clarify the necessary conditions to fulfill the WF effect. Figure 33 shows the energy level diagram for the hyperfine structure of hydrogen. There are three kinds of excitation (or de-excitation) mechanisms that determine the relative population between the two hyperfine levels of the 12S1/21^{2}S_{1/2} ground state of hydrogen: collisional transition, radiative transition, and Lyα\alpha pumping. In the stationary state, the rate equation for the population of the hyperfine states “0” and “1” can be written as follows:

n0(P01R+P01c+P01α)=n1(P10R+P10c+P10α),n_{0}\left(P_{01}^{{\rm R}}+P_{01}^{{\rm c}}+P_{01}^{\alpha}\right)=n_{1}\left(P_{10}^{{\rm R}}+P_{10}^{{\rm c}}+P_{10}^{\alpha}\right), (E1)

where n0n_{0} and n1n_{1} are populations of the sublevels “0” and “1,” and PRP^{{\rm R}}, PcP^{{\rm c}}, and PαP^{\alpha} are transition rates (per sec) caused by radiation (21 cm), collisions, and Lyα\alpha pumping, respectively. The subscript jiji denotes the transition from the sublevel jj to ii. According to the definition of spin temperature (TsT_{{\rm s}}), the relative population between the sublevels “0” and “1” is given by

n1n0=g1g0exp(hν10kBTs)3(1TTs),\frac{n_{1}}{n_{0}}=\frac{g_{1}}{g_{0}}\exp\left(-\frac{h\nu_{10}}{k_{{\rm B}}T_{{\rm s}}}\right)\simeq 3\left(1-\frac{T_{*}}{T_{{\rm s}}}\right), (E2)

where gig_{i} is the statistical weight of the sublevel ii, and Thν10/kB=0.0681T_{*}\equiv h\nu_{10}/k_{{\rm B}}=0.0681 K. In this study, we are interested in only the cases in which TsTT_{{\rm s}}\gg T_{*}.

The brightness temperature of radio radiation at ν10=21\nu_{10}=21 cm is defined by

TR=c22kBν102I(ν10).T_{{\rm R}}=\frac{c^{2}}{2k_{{\rm B}}\nu_{10}^{2}}I(\nu_{10}). (E3)

Here, I(ν)I(\nu) denotes the mean intensity of the radiation field expressed in energy at frequency ν\nu. It is also often convenient to express intensities in terms of the number of photons; in this paper, the mean intensity in photon number unit is referred to as J(ν)I(ν)/hνJ(\nu)\equiv I(\nu)/h\nu. The transition rates due to 21-cm radiation are then given by

P01R\displaystyle P_{01}^{{\rm R}} =\displaystyle= B01I(ν10)=3TRTA10,\displaystyle B_{01}I(\nu_{10})=3\frac{T_{{\rm R}}}{T_{*}}A_{10},
P10R\displaystyle P_{10}^{{\rm R}} =\displaystyle= A10+B10I(ν10)=(1+TRT)A10.\displaystyle A_{10}+B_{10}I(\nu_{10})=\left(1+\frac{T_{{\rm R}}}{T_{*}}\right)A_{10}. (E4)

Here, we used the following relation between three Einstein coefficients, B10=(g0/g1)B01=(c2/2hν103)A10B_{10}=(g_{0}/g_{1})B_{01}=(c^{2}/2h\nu_{10}^{3})A_{10}. The detailed-balance relation between the upward and downward rates for any colliding particle states:

P01cP10c=g1g0exp(hν10kBTK)3(1TTK).\frac{P_{01}^{{\rm c}}}{P_{10}^{{\rm c}}}=\frac{g_{1}}{g_{0}}\exp\left(-\frac{h\nu_{10}}{k_{{\rm B}}T_{{\rm K}}}\right)\simeq 3\left(1-\frac{T_{*}}{T_{{\rm K}}}\right). (E5)

Now, let us define the effective color temperature TαT_{\alpha} of UV radiation at Lyα\alpha by

P01αP10α3(1TTα).\frac{P_{01}^{\alpha}}{P_{10}^{\alpha}}\equiv 3\left(1-\frac{T_{*}}{T_{\alpha}}\right). (E6)

Substituting Equations (E4), (E5), and (E6) into Equation (E1), we then find

Ts\displaystyle T_{{\rm s}} =\displaystyle= T+TR+ycTK+yαTα1+yc+yα,\displaystyle\frac{T_{*}+T_{{\rm R}}+y_{{\rm c}}T_{{\rm K}}+y_{\alpha}T_{\alpha}}{1+y_{{\rm c}}+y_{\alpha}}, (E7)

where

yc\displaystyle y_{{\rm c}} =\displaystyle= TTKP10cA10andyα=TTαP10αA10.\displaystyle\frac{T_{*}}{T_{{\rm K}}}\frac{P_{10}^{{\rm c}}}{A_{10}}\ {\rm and}\ y_{\alpha}=\frac{T_{*}}{T_{\alpha}}\frac{P_{10}^{\alpha}}{A_{10}}. (E8)

Note that TT_{*} in the numerator of Equation (E7) is negligible and thus usually ignored. Therefore, TsT_{{\rm s}} is a weighted mean of the three temperatures TRT_{{\rm R}}, TcT_{{\rm c}}, and TαT_{\alpha}.

We now examine the necessary condition for Equation (E6) to be held. As shown in Figure 33, there are six permitted transitions between n=2n=2 and n=1n=1 hyperfine levels. The transitions between “5” and “1” and between “2” and “1” give no contribution to the excitation or de-excitation of the 21-cm transition. According to the sum rule, the line strength of the sum of all transitions from a given nFJnFJ to all nJn^{\prime}J^{\prime} levels (summed over FF^{\prime}) is proportional to 2F+12F+1. Therefore, the relative line strengths of the four downward transitions to n=1n^{\prime}=1, J=1/2J^{\prime}=1/2 are

S51:S41+S40:S31+S30:S21\displaystyle S_{51}:S_{41}+S_{40}:S_{31}+S_{30}:S_{21} =\displaystyle= 5:3:3:1,\displaystyle 5:3:3:1, (E9)

where SjiS_{ji} denotes the line strength from sublevel jj to sublevel ii in Figure 33. Similarly, the relative strengths for the upward transitions are given by

S04:S14+S15\displaystyle S_{04}:S_{14}+S_{15} =\displaystyle= 1:3,\displaystyle 1:3,
S03:S12+S13\displaystyle S_{03}:S_{12}+S_{13} =\displaystyle= 1:3.\displaystyle 1:3. (E10)

Here, we note that S50=S20=0S_{50}=S_{20}=0 and Sij=SjiS_{ij}=S_{ji}. From these proportionality expressions, we obtain the relative line strengths between the six transitions

S51:S41:S40:S31:S30:S21=5:1:2:2:1:1,S_{51}:S_{41}:S_{40}:S_{31}:S_{30}:S_{21}=5:1:2:2:1:1, (E11)

which is also denoted in Figure 33. Using the relation between the line strength and the Einstein spontaneous emission coefficient (AjiA_{ji}) for the downward transitions

Sji/Sα=(gj/gn=2)(Aji/Aα),S_{ji}/S_{\alpha}=(g_{j}/g_{n=2})(A_{ji}/A_{\alpha}), (E12)

where Sα=j=25i=01SjiS_{\alpha}=\sum_{j=2}^{5}\sum_{i=0}^{1}S_{ji}, Aα=i=01AjiA_{\alpha}=\sum_{i=0}^{1}A_{ji} (j=2,,5j=2,\cdots,5), gj=2Fj+1g_{j}=2F_{j}+1, and gn=2=j=25gj=12g_{n=2}=\sum_{j=2}^{5}g_{j}=12, we obtain the Einstein’s spontaneous emission coefficients:

A20\displaystyle A_{20} =\displaystyle= A50=0\displaystyle A_{50}=0
A21\displaystyle A_{21} =\displaystyle= A51=Aα\displaystyle A_{51}=A_{\alpha}
A30\displaystyle A_{30} =\displaystyle= A41=Aα/3\displaystyle A_{41}=A_{\alpha}/3
A31\displaystyle A_{31} =\displaystyle= A40=(2/3)Aα.\displaystyle A_{40}=(2/3)A_{\alpha}. (E13)

See also Tozzi et al. (2000) for the derivation of the above relations.

We can write, according to the definition, the transition rates P01αP_{01}^{\alpha} and P10αP_{10}^{\alpha} caused by Lyα\alpha pumping between the two hyperfine levels “1” and “0” as follows:

P01α\displaystyle P_{01}^{\alpha} =\displaystyle= j=25B0jI(ν0j)Aj1i=01Aji\displaystyle\sum_{j=2}^{5}B_{0j}I(\nu_{0j})\frac{A_{j1}}{\sum_{i=0}^{1}A_{ji}} (E14)
=\displaystyle= j=34gjg0c22hν0j3Aj0I(ν0j)Aj1i=01Aji\displaystyle\sum_{j=3}^{4}\frac{g_{j}}{g_{0}}\frac{c^{2}}{2h\nu_{0j}^{3}}A_{j0}I(\nu_{0j})\frac{A_{j1}}{\sum_{i=0}^{1}A_{ji}}
=\displaystyle= 23[𝒩(ν03)+𝒩(ν04)]Aα.\displaystyle\frac{2}{3}\left[\mathcal{N}(\nu_{03})+\mathcal{N}(\nu_{04})\right]A_{\alpha}.

Here, the relations between Einstein coefficients are used and 𝒩(ν)c2I(ν)/2hν3\mathcal{N}(\nu)\equiv c^{2}I(\nu)/2h\nu^{3} is the occupation number per state. Similarly, we find

P10α\displaystyle P_{10}^{\alpha} =\displaystyle= j=25B1jI(ν1j)Aj0i=01Aji\displaystyle\sum_{j=2}^{5}B_{1j}I(\nu_{1j})\frac{A_{j0}}{\sum_{i=0}^{1}A_{ji}} (E15)
=\displaystyle= 29[𝒩(ν13)+𝒩(ν14)]Aα.\displaystyle\frac{2}{9}\left[\mathcal{N}(\nu_{13})+\mathcal{N}(\nu_{14})\right]A_{\alpha}.

The ratio between the transition rates is, then, obtained to be

P01αP10α\displaystyle\frac{P_{01}^{\alpha}}{P_{10}^{\alpha}} =\displaystyle= 3𝒩(ν03)+𝒩(ν04)𝒩(ν13)+𝒩(ν14).\displaystyle 3\frac{\mathcal{N}(\nu_{03})+\mathcal{N}(\nu_{04})}{\mathcal{N}(\nu_{13})+\mathcal{N}(\nu_{14})}. (E16)

Assuming that the occupation number is proportional to a Wien’s law function with a color temperature TαT_{\alpha}, i.e., 𝒩(νji)exp(hνji/kBTα)\mathcal{N}(\nu_{ji})\propto\exp(-h\nu_{ji}/k_{{\rm B}}T_{\alpha}), we can write

P01αP10α=3exp(hν10kBTα)3(1hν10kBTα).\frac{P_{01}^{\alpha}}{P_{10}^{\alpha}}=3\exp\left(-\frac{h\nu_{10}}{k_{B}T_{\alpha}}\right)\simeq 3\left(1-\frac{h\nu_{10}}{k_{B}T_{\alpha}}\right). (E17)

Note that this equation holds approximately as well even when J(νji)=2νji2/c2𝒩(νji)exp(hνji/kBTα)J(\nu_{ji})=2\nu_{ji}^{2}/c^{2}\mathcal{N}(\nu_{ji})\propto\exp(-h\nu_{ji}/k_{{\rm B}}T_{\alpha}). Therefore, it is now clear that the first condition for the WF effect is that the spectrum of the mean intensity at the Lyα\alpha line center must be described as a Wiens’ law or an exponential function with the kinetic temperature (i.e., Tα=TKT_{\alpha}=T_{{\rm K}}).

Equation (E7) indicates that if either ycy_{{\rm c}} or yαy_{\alpha} is large, the spin temperature (TsT_{{\rm s}}) approaches the kinetic temperature (TKT_{{\rm K}}). Therefore, the second condition for the Lyα\alpha pumping to be dominant in determining the spin temperature is to have a strong Lyα\alpha radiation field J(να)J(\nu_{\alpha}) to increase yαy_{\alpha} in Equation (E7). The weighting factor yαy_{\alpha} can be expressed either in terms of the mean radiation field strength or in terms of the number of scatterings. For the first expression, we rewrite the mean radiation field as

I(να)=c4πhναnα=c4πhναNα2Δν~D,I(\nu_{\alpha})=\frac{c}{4\pi}h\nu_{\alpha}n_{\alpha}=\frac{c}{4\pi}h\nu_{\alpha}\frac{N_{\alpha}}{2\Delta\tilde{\nu}_{{\rm D}}}, (E18)

where nαn_{\alpha} is the number density and NαN_{\alpha} is the number of Lyα\alpha photons per cm3 within the frequency range of ±Δν~D=±να(νth/2)/c=±ΔνD/2\pm\Delta\tilde{\nu}_{{\rm D}}=\pm\nu_{\alpha}(\nu_{{\rm th}}/\sqrt{2})/c=\pm\Delta\nu_{{\rm D}}/\sqrt{2} from the line center. We then find

yα\displaystyle y_{\alpha} =\displaystyle= h(λαc)236πkBmHkBAαA10ν10ναNαTαTK1/2\displaystyle\frac{h(\lambda_{\alpha}c)^{2}}{36\pi k_{{\rm B}}}\sqrt{\frac{m_{{\rm H}}}{k_{{\rm B}}}}\frac{A_{\alpha}}{A_{10}}\frac{\nu_{10}}{\nu_{\alpha}}\frac{N_{\alpha}}{T_{\alpha}T_{{\rm K}}^{1/2}} (E19)
=\displaystyle= (7.9×1011cm3s1)Nα/(TαTK1/2).\displaystyle\left(7.9\times 10^{11}\ {\rm cm}^{3}\ {\rm s}^{-1}\right)N_{\alpha}/(T_{\alpha}T_{{\rm K}}^{1/2}).

Note that 5.9×10115.9\times 10^{11} in Equation (28) of Field (1958) is a typo. This method to estimate yαy_{\alpha} was initially proposed in Field (1958). To utilize this method, we need to measure spectra within the medium.

The second method, which is more convenient, is to count the number of Lyα\alpha resonance scatterings. We can readily imagine that the radiation field strength at a given volume is determined by how long photons will stay (or be confined) within the volume and thus the strength is closely associated with the number of scattering events. The transition rates due to Lyα\alpha pumping in Equations (E14) and (E15) can be represented as

P01α\displaystyle P_{01}^{\alpha} \displaystyle\simeq 43c22hνα3AαI(να)=49BαabsI(να),\displaystyle\frac{4}{3}\frac{c^{2}}{2h\nu_{\alpha}^{3}}A_{\alpha}I(\nu_{\alpha})=\frac{4}{9}B_{\alpha}^{{\rm abs}}I(\nu_{\alpha}),
P10α\displaystyle P_{10}^{\alpha} \displaystyle\simeq 49c22hνα3AαI(να)=427BαabsI(να),\displaystyle\frac{4}{9}\frac{c^{2}}{2h\nu_{\alpha}^{3}}A_{\alpha}I(\nu_{\alpha})=\frac{4}{27}B_{\alpha}^{{\rm abs}}I(\nu_{\alpha}), (E20)

where BαabsB_{\alpha}^{{\rm abs}} is the Einstein’s absorption coefficient. Here, the relation between the spontaneous emission coefficient and the absorption coefficient is used, i.e., Aα=(gn=1/gn=2)(2hνα3/c2)Bαabs=(1/3)(2hνα3/c2)Bαabs.A_{\alpha}=(g_{n=1}/g_{n=2})(2h\nu_{\alpha}^{3}/c^{2})B_{\alpha}^{{\rm abs}}=(1/3)(2h\nu_{\alpha}^{3}/c^{2})B_{\alpha}^{{\rm abs}}. By definition, the number of scatterings per unit time for an atom is given by

Pα\displaystyle P_{\alpha} =\displaystyle= 4πIνhνσν𝑑ν=BαabsI(να),\displaystyle 4\pi\int\frac{I_{\nu}}{h\nu}\sigma_{\nu}d\nu=B_{\alpha}^{{\rm abs}}I(\nu_{\alpha}), (E21)

in which the relation σν=(hν/4π)Bαabsϕν\sigma_{\nu}=(h\nu/4\pi)B_{\alpha}^{{\rm abs}}\phi_{\nu} is used. The transition rates due to Lyα\alpha pumping can then be expressed in terms of the scattering rate:

P10α=427PαandP01α=49Pα.P_{10}^{\alpha}=\frac{4}{27}P_{\alpha}\ \ {\rm and}\ \ P_{01}^{\alpha}=\frac{4}{9}P_{\alpha}. (E22)

Therefore, the weighting factor yαy_{\alpha} can be represented using PαP_{\alpha}:

yα\displaystyle y_{\alpha} =\displaystyle= 427hν10kBTαPαA10=3.52×1012PαTα.\displaystyle\frac{4}{27}\frac{h\nu_{10}}{k_{{\rm B}}T_{\alpha}}\frac{P_{\alpha}}{A_{10}}=3.52\times 10^{12}\frac{P_{\alpha}}{T_{\alpha}}. (E23)

In Section 4, we demonstrate that the mean intensity at the line center J(x=0)J(x=0) gives rise to results consistent with those obtained using PαP_{\alpha}.

References

  • Adams (1971) Adams, T. F. 1971, ApJ, 168, 575
  • Adams (1972) —. 1972, ApJ, 174, 439
  • Ahn & Lee (2015) Ahn, S.-H., & Lee, H.-W. 2015, Journal of The Korean Astronomical Society, 48, 195
  • Ahn et al. (2000) Ahn, S.-H., Lee, H.-W., & Lee, H. M. 2000, 33, 29
  • Ahn et al. (2002) —. 2002, ApJ, 567, 922
  • Amanatides & Woo (1987) Amanatides, J., & Woo, A. 1987, in Proceedings of Eurographics ’87, ed. G. Marechal (New York: Elsevier), 3–10
  • Avery & House (1968) Avery, L. W., & House, L. L. 1968, ApJ, 152, 493
  • Baek et al. (2009) Baek, S., Di Matteo, P., Semelin, B., Combes, F., & Revaz, Y. 2009, A&A, 495, 389
  • Baes et al. (2016) Baes, M., Gordon, K. D., Lunttila, T., et al. 2016, A&A, 590, A55
  • Baes et al. (2011) Baes, M., Verstappen, J., De Looze, I., et al. 2011, ApJS, 196, 22
  • Bahcall & Ekers (1969) Bahcall, J. N., & Ekers, R. D. 1969, ApJ, 157, 1055
  • Cantalupo et al. (2008) Cantalupo, S., Porciani, C., & Lilly, S. J. 2008, ApJ, 672, 48
  • Cantalupo et al. (2005) Cantalupo, S., Porciani, C., Lilly, S. J., & Miniati, F. 2005, ApJ, 628, 61
  • Carswell & Webb (2014) Carswell, R. F., & Webb, J. K. 2014, Astrophysics Source Code Library, ascl:1408.015
  • Cen (1992) Cen, R. 1992, ApJS, 78, 341
  • Corbelli & Salpeter (1993) Corbelli, E., & Salpeter, E. E. 1993, ApJ, 419, 94
  • Davies & Cummings (1975) Davies, R. D., & Cummings, E. R. 1975, MNRAS, 170, 95
  • Deguchi & Watson (1985) Deguchi, S., & Watson, W. D. 1985, ApJ, 290, 578
  • Dere et al. (2019) Dere, K. P., Del Zanna, G., Young, P. R., Landi, E., & Sutherland, R. S. 2019, ApJS, 241, 22
  • Dere et al. (1997) Dere, K. P., Landi, E., Mason, H. E., Monsignori Fossi, B. C., & Young, P. R. 1997, A & A Supplement series, 125, 149
  • Dijkstra et al. (2006) Dijkstra, M., Haiman, Z., & Spaans, M. 2006, ApJ, 649, 14
  • Draine (2003) Draine, B. T. 2003, ApJ, 598, 1017
  • Draine (2011) Draine, B. T. 2011, Physics of The Interstellar and Intergalactic Medium (Princeton, New Jersey: Princeton University Press)
  • Ferriere (1998) Ferriere, K. 1998, ApJ, 503, 700
  • Field (1958) Field, G. B. 1958, in Proceedings of the IRE, Harvard College Observatory, Cambridge, Mass., 240–250
  • Field (1959a) Field, G. B. 1959a, ApJ, 129, 536
  • Field (1959b) —. 1959b, ApJ, 129, 551
  • Field et al. (1969) Field, G. B., Goldsmith, D. W., & Habing, H. J. 1969, ApJ, 155, L149
  • Forero-Romero et al. (2011) Forero-Romero, J. E., Yepes, G., Gottlöber, S., et al. 2011, MNRAS, 415, 3666
  • Furlanetto et al. (2006) Furlanetto, S. R., Oh, S. P., & Briggs, F. H. 2006, Physics Reports, 433, 181
  • Giovanardi et al. (1987) Giovanardi, C., Natta, A., & Palla, F. 1987, A&AS, 70, 269
  • Gong et al. (2018) Gong, M., Ostriker, E. C., & Kim, C.-G. 2018, ApJ, 858, 16
  • Gronke & Dijkstra (2014) Gronke, M., & Dijkstra, M. 2014, MNRAS, 444, 1095
  • Gronke et al. (2016) Gronke, M., Dijkstra, M., McCourt, M., & Oh, S. P. 2016, ApJ, 833, L26
  • Gronke et al. (2017) Gronke, M., Dijkstra, M., McCourt, M., & Peng Oh, S. 2017, A&A, 607, A71
  • Gropp et al. (2014) Gropp, W., Lusk, E., & Skjellum, A. 2014, Using MPI: Portable Programming with the Message-Passing Interface, 3rd Ed. (Cambridge, Massachusetts: The MIT Press)
  • Haffner et al. (2009) Haffner, L. M., Dettmar, R.-J., Beckman, J. E., et al. 2009, RMP, 81, 969
  • Harrington (1973) Harrington, J. P. 1973, 162, 43
  • Hayes et al. (2011) Hayes, M., Schaerer, D., Östlin, G., et al. 2011, ApJ, 730, 8
  • Heiles & Troland (2003a) Heiles, C., & Troland, T. H. 2003a, ApJS, 145, 329
  • Heiles & Troland (2003b) —. 2003b, ApJ, 586, 1067
  • Heyer & Brunt (2004) Heyer, M. H., & Brunt, C. M. 2004, ApJ, 615, L45
  • Higgins & Meiksin (2012) Higgins, J., & Meiksin, A. 2012, MNRAS, 426, 2380
  • Hubeny (2001) Hubeny, I. 2001, Spectroscopic Challenges of Photoionized Plasmas, ed. G. Ferland & D. W. Savin, Vol. 247 (San Francisco, CA: ASP Conference Series)
  • Hubeny & Mihalas (2014) Hubeny, I., & Mihalas, D. 2014, Theory of Stellar Atmospheres (Princeton, New Jersey: Princeton University Press)
  • Hummer (1962) Hummer, D. G. 1962, 125, 21
  • Hummer (1969) —. 1969, MNRAS, 145, 95
  • Hummer & Kunasz (1980) Hummer, D. G., & Kunasz, P. B. 1980, ApJ, 236, 609
  • Irwin (1995) Irwin, J. A. 1995, PASP, 107, 715
  • Juvela (2005) Juvela, M. 2005, A&A, 440, 531
  • Kado-Fong et al. (2020) Kado-Fong, E., Kim, J.-G., Ostriker, E. C., & Kim, C.-G. 2020, arXiv, arXiv:2006.06697
  • Keeney et al. (2005) Keeney, B. A., Momjian, E., Stocke, J. T., Carilli, C. L., & Tumlinson, J. 2005, ApJ, 622, 267
  • Kim et al. (2019) Kim, C.-G., Choi, S. K., & Flauger, R. 2019, ApJ, 880, 106
  • Kim & Ostriker (2017) Kim, C.-G., & Ostriker, E. C. 2017, ApJ, 846, 133
  • Kim & Ostriker (2018) —. 2018, ApJ, 853, 173
  • Kim et al. (2013) Kim, C.-G., Ostriker, E. C., & Kim, W.-T. 2013, ApJ, 776, 1
  • Kim et al. (2014) —. 2014, ApJ, 786, 64
  • Kim et al. (2018) Kim, J.-G., Kim, W.-T., & Ostriker, E. C. 2018, ApJ, 859, 68
  • Kinderman & Monahan (1977) Kinderman, A. J., & Monahan, J. F. 1977, ACM Transactions on Mathematical Software (TOMS), 3, 257
  • Kuhlen et al. (2006) Kuhlen, M., Madau, P., & Montgomery, R. 2006, ApJ, 637, L1
  • Kulkarni & Heiles (1987) Kulkarni, S. R., & Heiles, C. 1987, in Interstellar Processes, ed. D. J. Hollenbach & J. Thronson, Harley A., Vol. 134, 87
  • Lallement et al. (2011) Lallement, R., Quémerais, E., Bertaux, J.-L., Sandel, B. R., & Izmodenov, V. 2011, Science, 334, 1665
  • Lao & Smith (2020) Lao, B.-X., & Smith, A. 2020, arXiv, arXiv:2005.09692
  • Larson (1981) Larson, R. B. 1981, MNRAS, 194, 809
  • Laursen et al. (2009) Laursen, P., Razoumov, A. O., & Sommer-Larsen, J. 2009, ApJ, 696, 853
  • Laursen & Sommer-Larsen (2007) Laursen, P., & Sommer-Larsen, J. 2007, ApJ, 657, L69
  • Leung & Liszt (1976) Leung, C. M., & Liszt, H. S. 1976, ApJ, 208, 732
  • Liszt (2001) Liszt, H. 2001, A&A, 371, 698
  • Loeb & Rybicki (1999) Loeb, A., & Rybicki, G. B. 1999, ApJ, 524, 527
  • Long et al. (1992) Long, K. S., Blair, W. P., Vancura, O., et al. 1992, ApJ, 400, 214
  • Lucy (1999) Lucy, L. B. 1999, A&A, 344, 282
  • Madau et al. (1997) Madau, P., Meiksin, A., & Rees, M. J. 1997, ApJ, 475, 429
  • Mao et al. (2019) Mao, S. A., Ostriker, E. C., & Kim, C.-G. 2019, arXiv, arXiv:1911.05078
  • Marsaglia & Zaman (1993) Marsaglia, G., & Zaman, A. 1993, The KISS generator, Technical report, Department of Statistics (Tallahassee, FL: Univ. of Florida)
  • Matsumoto & Nishimura (1998) Matsumoto, M., & Nishimura, T. 1998, ACM Transactions on Modeling and Computer Simulation, 8, 3
  • McKee & Ostriker (1977) McKee, C. F., & Ostriker, J. P. 1977, ApJ, 218, 148
  • Meiksin (2006) Meiksin, A. 2006, MNRAS, 370, 2025
  • Misselt et al. (2001) Misselt, K. A., Gordon, K. D., Clayton, G. C., & Wolff, M. J. 2001, ApJ, 551, 277
  • Murray et al. (2017) Murray, C. E., Stanimirovic, S., Kim, C.-G., et al. 2017, ApJ, 837, 55
  • Murray et al. (2014) Murray, C. E., Lindner, R. R., Stanimirovic, S., et al. 2014, ApJ, 781, L41
  • Murray et al. (2015) Murray, C. E., Stanimirovic, S., Goss, W. M., et al. 2015, ApJ, 804, 89
  • Neufeld (1990) Neufeld, D. A. 1990, ApJ, 350, 216
  • Orsi et al. (2012) Orsi, A., Lacey, C. G., & Baugh, C. M. 2012, MNRAS, 425, 87
  • Pierleoni et al. (2009) Pierleoni, M., Maselli, A., & Ciardi, B. 2009, MNRAS, 393, 872
  • Roy et al. (2009) Roy, I., Xu, W., Qiu, J.-M., Shu, C.-W., & Fang, L.-Z. 2009, ApJ, 694, 1121
  • Rybicki (1984) Rybicki, G. B. 1984, IN: Methods in radiative transfer (A85-16126 05-90). Cambridge and New York, 21
  • Rybicki & Lightman (1985) Rybicki, G. B., & Lightman, A. P. 1985, Radiative Processes in Astrophysics (New York: Wiley)
  • Schneider (1989) Schneider, S. E. 1989, ApJ, 343, 94
  • Schneider et al. (1983) Schneider, S. E., Helou, G., Salpeter, E. E., & Terzian, Y. 1983, ApJ, 273, L1
  • Semelin et al. (2007) Semelin, B., Combes, F., & Baek, S. 2007, A&A, 474, 365
  • Seon (2006) Seon, K.-I. 2006, PASJ, 58, 439
  • Seon (2015) —. 2015, JKAS, 48, 57
  • Seon & Draine (2016) Seon, K.-I., & Draine, B. T. 2016, ApJ, 833, 201
  • Seon & Witt (2012) Seon, K.-I., & Witt, A. N. 2012, ApJ, 758, 109
  • Shaw (2005) Shaw, G. 2005, PhD thesis, University of Kentucky
  • Shaw et al. (2017) Shaw, G., Ferland, G. J., & Hubeny, I. 2017, ApJ, 843, 0
  • Smith et al. (2017) Smith, A., Bromm, V., & Loeb, A. 2017, MNRAS, 464, 2963
  • Smith et al. (2015) Smith, A., Safranek-Shrader, C., Bromm, V., & Milosavljević, M. 2015, MNRAS, 449, 4336
  • Stenflo (1980) Stenflo, J. O. 1980, A&A, 84, 68
  • Swaters et al. (1997) Swaters, R. A., Sancisi, R., & van der Hulst, J. M. 1997, ApJ, 491, 140
  • Tasitsiomi (2006) Tasitsiomi, A. 2006, ApJ, 645, 792
  • Tozzi et al. (2000) Tozzi, P., Madau, P., Meiksin, A., & Rees, M. J. 2000, ApJ, 528, 597
  • Urbaniak & Wolfe (1981) Urbaniak, J. J., & Wolfe, A. M. 1981, ApJ, 244, 406
  • Vacca et al. (1996) Vacca, W. D., Garmany, C. D., & Shull, J. M. 1996, ApJ, 460, 914
  • Verhamme et al. (2006) Verhamme, A., Schaerer, D., & Maselli, A. 2006, A&A, 460, 397
  • Vijayan et al. (2019) Vijayan, A., Kim, C.-G., Armillotta, L., Ostriker, E. C., & Li, M. 2019, arXiv, arXiv:1911.07872
  • Watson & Deguchi (1984) Watson, W. D., & Deguchi, S. 1984, ApJ, 281, L5
  • Weingartner & Draine (2001) Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
  • Witt (1977) Witt, A. N. 1977, ApJS, 35, 1
  • Witt et al. (2010) Witt, A. N., Gold, B., Barnes, F. S. I., et al. 2010, ApJ, 724, 1551
  • Wolfire et al. (1995) Wolfire, M. G., Hollenbach, D., McKee, C. F., Tielens, A. G. G. M., & Bakes, E. L. O. 1995, ApJ, 443, 152
  • Wolfire et al. (2003) Wolfire, M. G., McKee, C. F., Hollenbach, D., & Tielens, A. G. G. M. 2003, ApJ, 587, 278
  • Wouthuysen (1952) Wouthuysen, S. A. 1952, AJ, 57, 31
  • Yajima et al. (2012) Yajima, H., Li, Y., Zhu, Q., & Abel, T. 2012, 424, 884
  • Zheng & Miralda-Escudé (2002a) Zheng, Z., & Miralda-Escudé, J. 2002a, ApJ, 578, 33
  • Zheng & Miralda-Escudé (2002b) —. 2002b, ApJ, 568, L71