A Heavily Scattered Fast Radio Burst Is Viewed Through Multiple Galaxy Halos

Jakob T. Faber Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Vikram Ravi Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Stella Koch Ocker Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. The Observatories of the Carnegie Institution for Science, Pasadena, CA 91101, USA. Myles B. Sherman Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Kritti Sharma Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Liam Connor Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Casey Law Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Nikita Kosogorov Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Gregg Hallinan Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Charlie Harnach Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Greg Hellbourg Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Rick Hobbs Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. David Hodge Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Mark Hodges Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. James W. Lamb Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Paul Rasmussen Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA. Jean J. Somalwar Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. Sander Weinreb Cahill Center for Astronomy and Astrophysics, MC 249-17 California Institute of Technology, Pasadena CA 91125, USA. David P. Woody Owens Valley Radio Observatory, California Institute of Technology, Big Pine CA 93513, USA.
Abstract

We present a multi-wavelength study of the apparently non-repeating, heavily scattered fast radio burst, FRB 20221219A, detected by the Deep Synoptic Array 110 (DSA-110). The burst exhibits a moderate dispersion measure (DM) of 706.70.6+0.6subscriptsuperscript706.70.60.6706.7^{+0.6}_{-0.6}706.7 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT pccm3pcsuperscriptcm3\mathrm{pc}~{}\mathrm{cm}^{-3}roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and an unusually high scattering timescale of τobs=19.22.7+2.7subscript𝜏obssuperscriptsubscript19.22.72.7\tau_{\mathrm{obs}}=19.2_{-2.7}^{+2.7}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = 19.2 start_POSTSUBSCRIPT - 2.7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 2.7 end_POSTSUPERSCRIPT ms at 1.4 GHz. We associate the FRB with a Milky Way-like host galaxy at zhost=0.554subscript𝑧host0.554z_{\mathrm{host}}=0.554italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT = 0.554 of stellar mass log10(M,host)=10.200.03+0.04Msubscriptlog10subscript𝑀hostsubscriptsuperscript10.200.040.03subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{host}})=10.20^{+0.04}_{-0.03}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , roman_host end_POSTSUBSCRIPT ) = 10.20 start_POSTSUPERSCRIPT + 0.04 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.03 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We identify two intervening galaxy halos at redshifts zigh1=0.492subscript𝑧igh10.492z_{\mathrm{igh1}}=0.492italic_z start_POSTSUBSCRIPT igh1 end_POSTSUBSCRIPT = 0.492 and zigh2=0.438subscript𝑧igh20.438z_{\mathrm{igh2}}=0.438italic_z start_POSTSUBSCRIPT igh2 end_POSTSUBSCRIPT = 0.438, with low impact parameters, bigh1=43.011.3+11.3subscript𝑏igh1superscriptsubscript43.011.311.3b_{\mathrm{igh1}}=43.0_{-11.3}^{+11.3}italic_b start_POSTSUBSCRIPT igh1 end_POSTSUBSCRIPT = 43.0 start_POSTSUBSCRIPT - 11.3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 11.3 end_POSTSUPERSCRIPT kpc and bigh2=36.111.3+11.3subscript𝑏igh2superscriptsubscript36.111.311.3b_{\mathrm{igh2}}=36.1_{-11.3}^{+11.3}italic_b start_POSTSUBSCRIPT igh2 end_POSTSUBSCRIPT = 36.1 start_POSTSUBSCRIPT - 11.3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 11.3 end_POSTSUPERSCRIPT kpc, and intermediate stellar masses, log10(M,igh1)=10.010.02+0.02Msubscriptlog10subscript𝑀igh1subscriptsuperscript10.010.020.02subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{igh1}})=10.01^{+0.02}_{-0.02}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , igh1 end_POSTSUBSCRIPT ) = 10.01 start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and log10(M,igh2)=10.600.02+0.02Msubscriptlog10subscript𝑀igh2subscriptsuperscript10.600.020.02subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{igh2}})=10.60^{+0.02}_{-0.02}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , igh2 end_POSTSUBSCRIPT ) = 10.60 start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The presence of two such galaxies suggests that the sightline is significantly overcrowded compared to the median sightline to this redshift, as inferred from the halo mass function. We perform a detailed analysis of the sightline toward FRB 20221219A, constructing both DM and scattering budgets. Our results suggest that, unlike most well-localized sources, the host galaxy does not dominate the observed scattering. Instead, we posit that an intersection with a single partially ionized cloudlet in the circumgalactic medium of an intervening galaxy could account for the substantial scattering in FRB 20221219A and remain in agreement with typical electron densities inferred for extra-planar dense cloud-like structures in the Galactic and extragalactic halos (e.g., high-velocity clouds).

Radio bursts (1339), Radio transient sources (2008), Interstellar medium (847), Circumgalactic medium (1879), Intergalactic medium (813), Galaxy stellar halos (598), Interstellar scattering (854)
facilities: DSA-110, Keck-I/LRIS (Oke et al., 1995), Keck-II/DEIMOS (Faber et al., 2003)software: astropy (Price-Whelan et al., 2018); scipy (Virtanen et al., 2020), numpy (Harris et al., 2020), hmf (Murray, 2014), matplotlib (Hunter, 2007), LPipe (Perley, 2019), pPXF (Cappellari, 2017, 2022), PypeIt (Prochaska et al., 2020), prospector (Johnson et al., 2021), astropath (Aggarwal et al., 2021)

1 Introduction

Fast radio bursts (FRBs; Lorimer et al., 2007) are a class of highly luminous extragalactic radio transients with durations ranging from nanoseconds to seconds (Petroff et al., 2019; Cordes & Chatterjee, 2019; Majid et al., 2021; Nimmo et al., 2022; CHIME/FRB Collaboration et al., 2022). Their dispersion measures (DMs), which quantify the electron column density along the line of sight (LoS), exceed the values expected from Milky Way electron density models, suggesting extragalactic origins. This hypothesis has since been confirmed by an increasing number of well-localized FRBs that reside in host galaxies out to redshifts z similar-to\sim 1 (Ryder et al., 2022; Law et al., 2023; Gordon et al., 2023).

While the physical origins of FRBs remain unknown, many source models have been proposed, ranging from isolated neutron stars to compact object mergers. Recent observations have noted magnetars as viable sources, supported by detections of millisecond radio bursts from the galactic magnetar SGR 1935+++2154 by the STARE-2 and CHIME/FRB observatories, with measured isotropic-equivalent energies of 2.20.4+0.4×1035ergsubscriptsuperscript2.20.40.4superscript1035erg2.2^{+0.4}_{-0.4}\times 10^{35}~{}\mathrm{erg}2.2 start_POSTSUPERSCRIPT + 0.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.4 end_POSTSUBSCRIPT × 10 start_POSTSUPERSCRIPT 35 end_POSTSUPERSCRIPT roman_erg and 31.6+3×1035ergsuperscriptsubscript31.63superscript1035erg3_{-1.6}^{+3}\times 10^{35}~{}\mathrm{erg}3 start_POSTSUBSCRIPT - 1.6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 3 end_POSTSUPERSCRIPT × 10 start_POSTSUPERSCRIPT 35 end_POSTSUPERSCRIPT roman_erg, respectively, consistent with FRB source energetics (Bochenek et al., 2020; CHIME/FRB Collaboration et al., 2020).

Despite their unknown formation channels, sources, and emission mechanisms, if the source distance is known and intervening material is well-characterized, FRB DMs serve as stringent probes of baryon densities within galaxy halos, clusters, and filaments throughout the intergalactic medium (IGM; McQuinn, 2014; Macquart et al., 2020). FRB DMs can also facilitate the study of diffuse ionized gas in the Milky Way and neighboring galaxies (Connor & Ravi, 2022; Ravi et al., 2023a; Cook et al., 2023).

FRB observations are sensitive to inhomogeneities in intervening plasma through the effects of multipath propagation, or scattering. This propagation effect results in asymmetric, frequency-dependent pulse broadening and broadly exponential scattering tails (assuming a Gaussian scattered image). In the case of interstellar scattering, such tails are expected to follow a temporal delay τναproportional-to𝜏superscript𝜈𝛼\tau\propto\nu^{-\alpha}italic_τ ∝ italic_ν start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT, where α4greater-than-or-equivalent-to𝛼4\alpha\gtrsim 4italic_α ≳ 4 assuming an infinitely wide, thin scattering screen described by Kolmogorov turbulence (α=4.4𝛼4.4\alpha=4.4italic_α = 4.4 for scale-free power spectra, α=4𝛼4\alpha=4italic_α = 4 in the case of a diffractive scale below a finite inner scale, or where the density fluctuations follow a Gaussian power spectrum; Rickett et al., 2009). These assumptions are largely idealized, however, and may be challenged by flatter values (α4less-than-or-similar-to𝛼4\alpha\lesssim 4italic_α ≲ 4) inferred for Galactic pulsars (Löhmer et al., 2001; Deneva et al., 2009; Dexter et al., 2017). The diffractive scintillation bandwidth Δνd(2πτ)1similar-to-or-equalsΔsubscript𝜈dsuperscript2𝜋𝜏1\Delta\nu_{\mathrm{d}}\simeq(2\pi\tau)^{-1}roman_Δ italic_ν start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT ≃ ( 2 italic_π italic_τ ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT yields more information due to its preservation of phase information, though in practice, most of the convincing measurements have been made for scintillation caused by Galactic scattering (Masui et al., 2015; Gajjar et al., 2018; Hessels et al., 2019; Marcote et al., 2020; Bhandari et al., 2020; Schoen et al., 2021; Ocker et al., 2022a; Sammons et al., 2023), with the exception of FRB 150807 (Ravi et al., 2016). While some FRBs at lower Galactic latitudes (e.g., FRB 20121102A, FRB 20180916B) are predominantly scattered by the Milky Way (MW) (Ocker et al., 2021), the majority exhibit scattering timescales (τ(ν)𝜏𝜈\tau(\nu)italic_τ ( italic_ν )) that exceed those expected from purely Galactic contributions (Cordes & Lazio, 2002). The dominant scattering source for these FRBs is unconstrained in the absence of both scintillation bandwidth and scattering timescales, which can be used in tandem to constrain the presence of both Galactic and extragalactic screens (Cordes et al., 2016).

To date, sightline-independent correlations between DM and τ𝜏\tauitalic_τ have been explored for the broader FRB population (Ravi, 2018; CHIME/FRB Collaboration et al., 2019), and compared to the known τ𝜏\tauitalic_τ-DM correlation observed for MW radio pulsars (Cordes et al., 2016, 2022). Such comparisons have been difficult, however, as the canonical τ𝜏\tauitalic_τ-DM relation for pulsars relies on Galactic DMs, while the DMs typically assumed for FRBs in comparison are extragalactic (DMex; beyond the MW) and not specific to the host galaxy of the source. The relatively small sample of well-localized FRBs available has limited our ability to precisely constrain host DMs, and make what would be more apt comparisons between τ𝜏\tauitalic_τ and DMhost, rather than DMex. However, the recent advent of precise FRB localizations has begun alleviating this dilemma, which we discuss in §3.1.

Recent studies favor either the host galaxy or circumburst environment as the dominant scattering medium (Simha et al., 2020; Chittidi et al., 2021; Ocker et al., 2022c; Cordes et al., 2022), though it has been suggested that intervening halos may contribute meaningfully to scattering in select cases as well (Vedantham & Phinney, 2018; Connor et al., 2020; Chawla et al., 2022). Determining whether scattering is, in certain instances, dominated by the circumburst medium, host, or circumgalactic media (CGM) of intervening galaxies could further constrain source models and offer insights into these galactic halos, opening avenues for detailed studies of plasma turbulence in both the host and intervening galaxies (Cordes et al., 2016; Simard & Ravi, 2021; Ocker et al., 2022c).

For scattering to dominate in the circumburst medium, high electron densities ne102much-greater-thansubscript𝑛𝑒superscript102n_{e}\gg 10^{-2}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≫ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT cm-3 and high-amplitude density fluctuations (far exceeding a MW-like ISM; Ocker et al., 2020) would be expected, as well as strong magnetic fields B50μGgreater-than-or-equivalent-to𝐵50𝜇GB\gtrsim 50~{}\mu\mathrm{G}italic_B ≳ 50 italic_μ roman_G (similar to the magnetized filaments observed in the Crab nebula, for instance; Bietenholz & Kronberg, 1991) that induce extreme Faraday rotation. Highly magnetic environments do not, however, guarantee strong scattering (e.g., in the case of the first repeater FRB 20121102A; Michilli et al., 2018). To date, evidence for circumburst scattering has only been found in FRB 20190520B, which exhibits rapid variability in scattering between bursts (Ocker et al., 2022c).

Significant scattering in the interstellar medium (ISM) of FRB host galaxies appears to be more common (Masui et al., 2015; Ocker et al., 2022a). Such constraints require contemporaneous and discrepant scintillation and scattering measurements, such that a two-screen scattering model can be applied (Simard & Ravi, 2021). The two-screen model does, however, come with certain assumptions: e.g., that there are only two dominant scattering screens. Furthermore, placing constraints on the model necessitates measurable scintillation in the scattering tail of the burst. Scattering in one or more intervening galaxies, for instance, further complicates the two-screen model.

In this paper, we analyze the contributions to both DM and scattering in FRB 20221219A by intervening plasmas. In §2 we present the detection of the heavily scattered burst, including its localization with the DSA-110 and optical/IR follow-up observations of its host galaxy and two intervening galaxies. In §3.1 we constrain contributions to the DM along the LoS (the so-called “DM budget”) using HI-inferred HII column densities for intervening galaxies based on results from the COS-Halos Survey (Werk et al., 2014), and Illustris TNG simulation data (Zhang et al., 2021) for the intergalactic medium (IGM). In §3.2 we present a detailed investigation aimed at identifying the medium responsible for the uniquely high degree of scattering (the so-called “scattering budget”), with a focus on the circumgalactic media (CGMs) of two intervening galaxies. We conclude in §4. In this work, we adopt standard cosmological parameters from Planck Collaboration et al. (2016) for estimating DM contributions from the IGM (consistent with Illustris TNG; Pillepich et al., 2017) and from Planck Collaboration et al. (2020) for all other cosmological inferences.

2 DSA-110 Observation of FRB 20221219A

Table 1: Burst properties for FRB 20221219A (see Figure 1).
Parameter Value
Localization[R.A.J2000,Dec.J2000]\mathrm{Localization~{}[R.A.~{}J2000,~{}Dec.~{}J2000]}roman_Localization [ roman_R . roman_A . J2000 , roman_Dec . J2000 ] 17h10m31.15s,+713736.6superscript17superscript10𝑚superscript31.15𝑠superscript713736.617^{h}10^{m}31.15^{s},+71^{\circ}37\arcmin 36.6\arcsec17 start_POSTSUPERSCRIPT italic_h end_POSTSUPERSCRIPT 10 start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT 31.15 start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT , + 71 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 37 ′ 36.6 ″
LocalizationUncertainty[1σ,arcsec]LocalizationUncertainty1𝜎arcsec\mathrm{Localization~{}Uncertainty~{}[1\sigma,arcsec]}roman_Localization roman_Uncertainty [ 1 italic_σ , roman_arcsec ] 1.5, 0.9
Signal-to-NoiseSignal-to-Noise\mathrm{Signal}\text{-}\mathrm{to}\text{-}\mathrm{Noise}roman_Signal - roman_to - roman_Noise 8.91+1subscriptsuperscript8.9118.9^{+1}_{-1}8.9 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT
DispersionMeasure[pccm3]DispersionMeasuredelimited-[]pcsuperscriptcm3\mathrm{Dispersion~{}Measure~{}[pc~{}cm^{-3}]}roman_Dispersion roman_Measure [ roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ] 706.70.6+0.6subscriptsuperscript706.70.60.6706.7^{+0.6}_{-0.6}706.7 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT
IntrinsicFWHM[ms]IntrinsicFWHMdelimited-[]ms\mathrm{Intrinsic~{}FWHM~{}[ms]}roman_Intrinsic roman_FWHM [ roman_ms ] 0.120.08+0.65superscriptsubscript0.120.080.650.12_{-0.08}^{+0.65}0.12 start_POSTSUBSCRIPT - 0.08 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.65 end_POSTSUPERSCRIPT
ScatteringTimescale[ms]ScatteringTimescaledelimited-[]ms\mathrm{Scattering~{}Timescale~{}[ms]}roman_Scattering roman_Timescale [ roman_ms ] 19.22.7+2.7superscriptsubscript19.22.72.719.2_{-2.7}^{+2.7}19.2 start_POSTSUBSCRIPT - 2.7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 2.7 end_POSTSUPERSCRIPT
LinearPolarizationFraction[L/I]LinearPolarizationFractiondelimited-[]LI\mathrm{Linear~{}Polarization~{}Fraction~{}[L/I]}roman_Linear roman_Polarization roman_Fraction [ roman_L / roman_I ] 0.5020.08subscript0.5020.080.502_{0.08}0.502 start_POSTSUBSCRIPT 0.08 end_POSTSUBSCRIPT

2.1 DSA-110 Discovery and Localization

As described in Ravi et al. (2023b), the DSA-110 is a radio interferometer located at the Owens Valley Radio Observatory (OVRO) dedicated to the discovery and arcsecond-scale localization of FRBs. FRB 20221219A was detected during commissioning observations at a Modified Julian Date (MJD) of 59932.79297813 (arrival time at 1530 MHz at the observatory), with a real-time detection signal-to-noise ratio111This was calculated using a weighted sum matched filter. of 8.9. Antenna voltage measurements containing the burst were recorded in two linear polarizations with a time resolution of 32.768 μs𝜇𝑠\mu sitalic_μ italic_s across 6144 channels between 1311.25–1498.75 MHz. These data were used to localize the burst to a sky location of (R.A. J2000, Dec. J2000) = (+17h10m31.15s,+713736.6superscript17superscript10𝑚superscript31.15𝑠superscript713736.6+17^{h}10^{m}31.15^{s},+71^{\circ}37\arcmin 36.6\arcsec+ 17 start_POSTSUPERSCRIPT italic_h end_POSTSUPERSCRIPT 10 start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT 31.15 start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT , + 71 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 37 ′ 36.6 ″) with 1σ1𝜎1\sigma1 italic_σ uncertainties of 1.5″ in R.A. and 0.9″ in Dec. The data were also coherently combined to measure the various burst properties given in Table 1, and a dynamic spectrum of the burst is shown in Figure 1.

2.2 Inference of Burst Properties

We optimize the DM for temporal structure in the burst by searching for the DM at which the forward-derivative of the dedispersed timeseries is maximized, following methods outlined in Gajjar et al. (2018), Hessels et al. (2019), and Josephy et al. (2019). We begin by calculating a 2D DM transform across the burst in time for 0.075 pccm3pcsuperscriptcm3\mathrm{pc}~{}\mathrm{cm}^{-3}roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT intervals, ranging from 692-722692-722692~{}\text{-}~{}722692 - 722 pccm3pcsuperscriptcm3\mathrm{pc}~{}\mathrm{cm}^{-3}roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. We then calculate the forward-derivative and convolve with a 3×5(2.7ms×0.5pccm3)352.7ms0.5pcsuperscriptcm33\times 5\left(2.7~{}\mathrm{ms}\times 0.5~{}\mathrm{pc}~{}\mathrm{cm}^{-3}\right)3 × 5 ( 2.7 roman_ms × 0.5 roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) smoothing kernel. Finally, we integrate the modulus of the forward-derivative in time raised to n(1,2,4)𝑛124n\in(1,2,4)italic_n ∈ ( 1 , 2 , 4 ). Gajjar et al. (2018) and Hessels et al. (2019) elect to use n=1𝑛1n=1italic_n = 1 and n=2𝑛2n=2italic_n = 2, respectively. Josephy et al. (2019), however, show that n>2𝑛2n>2italic_n > 2 is optimal for identifying one uniquely bright peak in the pulse-profile, while n2less-than-or-similar-to𝑛2n\lesssim 2italic_n ≲ 2 typically prefer a series of lower-amplitude peaks. For the curves corresponding to each value of n𝑛nitalic_n, we fit a complex polynomial and take its peak as the structure-maximizing DM. To estimate uncertainties, we fit a simple Gaussian function to the most prominent peak in the polynomial fit and take the 1σ𝜎\sigmaitalic_σ offset. For both n=2𝑛2n=2italic_n = 2 and n=4𝑛4n=4italic_n = 4, we find a structure-optimizing DM of 706.70.6+0.6pccm3subscriptsuperscript706.70.60.6pcsuperscriptcm3706.7^{+0.6}_{-0.6}~{}\mathrm{pc~{}cm}^{-3}706.7 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, as both contain an unambiguous peak.

Due to the low S/N of the burst, we confirm the presence of scattering by dividing the observing band into three 83similar-to-or-equalsabsent83\simeq 83≃ 83 MHz sub-bands and fit exponentially-modified Gaussian functions 𝔾(t)𝔾𝑡\mathbb{G}(t)blackboard_G ( italic_t ) to the timeseries of each respective sub-band, defined as

𝔾(tA,Δt,σ,τ)=A×[exp((tΔt)22σ2)]𝔾conditional𝑡AΔ𝑡𝜎𝜏Adelimited-[]superscript𝑡Δ𝑡22superscript𝜎2\displaystyle\mathbb{G}\left(t\mid\mathrm{A},\Delta t,\sigma,\tau\right)=% \mathrm{A}\times\left[\exp\left(-\frac{\left(t-\Delta t\right)^{2}}{2\sigma^{2% }}\right)\right]blackboard_G ( italic_t ∣ roman_A , roman_Δ italic_t , italic_σ , italic_τ ) = roman_A × [ roman_exp ( - divide start_ARG ( italic_t - roman_Δ italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ] (1)
[(tΔt)exp(tΔtτ)],absentdelimited-[]𝑡Δ𝑡𝑡Δ𝑡𝜏\displaystyle\ast\left[\mathbb{H}\left(t-\Delta t\right)\exp\left(-\frac{t-% \Delta t}{\tau}\right)\right],∗ [ blackboard_H ( italic_t - roman_Δ italic_t ) roman_exp ( - divide start_ARG italic_t - roman_Δ italic_t end_ARG start_ARG italic_τ end_ARG ) ] ,

where t𝑡titalic_t represents time, τ𝜏\tauitalic_τ indicates the scattering timescale at the central frequency of the band (in MHz), (t)𝑡\mathbb{H}(t)blackboard_H ( italic_t ) is the Heaviside unit step function, ΔtΔ𝑡\Delta troman_Δ italic_t is the time shift, σ𝜎\sigmaitalic_σ is Gaussian standard deviation, and \ast symbolizes a convolution. We find that the pulse broadening timescale evolves in accordance with α=2.6±1.8𝛼plus-or-minus2.61.8\alpha=2.6\pm 1.8italic_α = 2.6 ± 1.8, consistent with α4greater-than-or-equivalent-to𝛼4\alpha\gtrsim 4italic_α ≳ 4 scatter-broadening to within uncertainties. Fitting 𝔾(t)𝔾𝑡\mathbb{G}(t)blackboard_G ( italic_t ) to the timeseries obtained by integrating across the full observing band, we infer a scattering timescale τ=19.22.7+2.7𝜏superscriptsubscript19.22.72.7\tau=19.2_{-2.7}^{+2.7}italic_τ = 19.2 start_POSTSUBSCRIPT - 2.7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 2.7 end_POSTSUPERSCRIPT ms. The uncertainties in τ𝜏\tauitalic_τ are derived from the covariance matrix of the least-squares fit, which is re-calculated for a set of pulse profiles within the DM uncertainties to produce a total error that includes uncertainty in DM. These fits are shown in Figure 1.

Refer to caption
Figure 1: The dynamic (time-frequency) spectrum showing FRB 20221219A, downsampled in time by a factor of 18 and in frequency by a factor of 96 to for clarity. To better visualize the burst in the dynamic spectrum, we further smoothed the data using a Savitzky-Golay smoothing filter with a window of 13 time bins and a polynomial order of 1. Above the dynamic spectrum, we show the normalized, unsmoothed total-intensity (Inormsubscript𝐼normI_{\rm norm}italic_I start_POSTSUBSCRIPT roman_norm end_POSTSUBSCRIPT) and linearly polarized intensity (Lnormsubscript𝐿normL_{\rm norm}italic_L start_POSTSUBSCRIPT roman_norm end_POSTSUBSCRIPT) burst timeseries, as well as a non-linear least-squares fit for 𝔾(t)𝔾𝑡\mathbb{G}(t)blackboard_G ( italic_t ) (Eq. 1) to the pulse profile, which indicates a τ=19.22.7+2.7𝜏superscriptsubscript19.22.72.7\tau=19.2_{-2.7}^{+2.7}italic_τ = 19.2 start_POSTSUBSCRIPT - 2.7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 2.7 end_POSTSUPERSCRIPT ms. In the right-most panels, we fit timeseries to three 83similar-to-or-equalsabsent83\simeq 83≃ 83 MHz sub-bands, and the respective scattering timescales measured for each sub-band with errors with a best-fit power-law of τν2.6±1.8proportional-to𝜏superscript𝜈plus-or-minus2.61.8\tau\propto\nu^{-2.6\pm 1.8}italic_τ ∝ italic_ν start_POSTSUPERSCRIPT - 2.6 ± 1.8 end_POSTSUPERSCRIPT power law. Uncertainties for all fits of 𝔾(t)𝔾𝑡\mathbb{G}(t)blackboard_G ( italic_t ) are shown as shaded regions surrounding the fitted profiles.

2.3 Optical/IR Follow-Up of Host & Intervening Galaxies

Using the 10th data release of the Legacy Survey222https://www.legacysurvey.org/dr10/ (DR10 R-band image; Dey et al., 2019) and the Pan-STARRS1 (PS1; Chambers et al., 2016) optical survey, we identify a plausible host galaxy for FRB 20221219A (henceforth HG 20221219A) consistent with the radio localization ellipse (see R-band field in Figure 3). We also identify two closely neighboring galaxies (henceforth IGH1 and IGH2).

We use astropath (Aggarwal et al., 2021) for host galaxy validation, which implements a Bayesian approach to estimate association probabilities for neighboring galaxies with tunable priors based on the sky position of the FRB, as well as their R-band magnitude, position, and R-band radius (see SED fit in Figure 2; Sharma et al. (2023) and Law et al. (2023) provide details regarding host association methods). We were able to successfully associate the host galaxy to FRB 20221219A at a confidence level of 99.6%percent99.699.6\%99.6 % using Legacy Survey DR10 R-band imaging data from the Bei**g-Arizona Sky Survey (BASS). We discuss IGH1 and IGH2, also visible in the field (see Figure 3), in Sections 2.4, 3.2.4, and 3.2.5.

2.3.1 Optical/IR Photometry

We make photometric measurements using images from PS1 (g,r,i,z,y𝑔𝑟𝑖𝑧𝑦g,r,i,z,yitalic_g , italic_r , italic_i , italic_z , italic_y), DECam (g,r,z𝑔𝑟𝑧g,r,zitalic_g , italic_r , italic_z; Valdes et al., 2014), WISE (w1,w2𝑤1𝑤2w1,w2italic_w 1 , italic_w 2; Wright et al., 2010), as well as IR photometric data obtained with the Wide Field Infrared Camera (WIRC, Ks; Wilson et al., 2003) on August 16, 2022. HG 20221219A, IGH1, and IGH2 are characterized following the methods outlined in Sharma et al. (2023). We measure isophotes for each using PS1 i-band images, setting a best-fit coverage of 95%greater-than-or-equivalent-toabsentpercent95\gtrsim 95\%≳ 95 %. The isophote is then scaled in accordance with the point spread functions of various photometric bands to execute aperture photometry.

2.3.2 Optical/IR Spectroscopy

We observed HG 20221219A with the Deep Extragalactic Imaging Multi-Object Spectrograph (Keck-II/DEIMOS; Faber et al., 2003) on April 20, 2023, and both IGH1 and IGH2 with the Low-Resolution Imaging Spectrometer on the Keck I telescope (Keck-I/LRIS; Oke et al., 1995) on June 14, 2023. We created a mask of roughly 50 slits for our DEIMOS observation of the FRB 20221219A field. Slits were placed on the positions of candidate foreground galaxies from the Legacy Survey DR9 catalog, including on massive central members of the galaxy cluster J171039.6+713427.

We reduce our spectra using PypeIt (Prochaska et al., 2020) and the standard LPipe software (Perley, 2019) and calibrate using observations of the BD+28 4211 standard star. To account for slit losses, we scale the spectra to match the PS1 photometry.

To measure the spectroscopic redshifts and line fluxes, we use the penalized PiXel-Fitting software (pPXF; Cappellari, 2017, 2022). This jointly fits for the stellar continuum and nebular emission lines based on the MILES stellar library (Sanchez-Blazquez et al., 2006). This places the host galaxy at a redshift of zhost=0.554subscript𝑧host0.554z_{\mathrm{host}}=0.554italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT = 0.554, and find the two neighboring galaxies in the foreground at zigh1=0.492subscript𝑧igh10.492z_{\mathrm{igh1}}=0.492italic_z start_POSTSUBSCRIPT igh1 end_POSTSUBSCRIPT = 0.492 and zigh2=0.438subscript𝑧igh20.438z_{\mathrm{igh2}}=0.438italic_z start_POSTSUBSCRIPT igh2 end_POSTSUBSCRIPT = 0.438 (see Table 2).

2.3.3 SED Modeling

Using the spectral energy distribution (SED) fitting software Prospector (Johnson et al., 2021), we perform a non-parametric fit for the stellar properties of the HG 20221219A, IGH1 and IGH2. While we were unable to obtain star formation rates for IGH1 and IGH2 due to low S/N in the FIR region of the spectrum, we were able to identify HG 20221219A as a modestly star-forming galaxy, M˙=1.780.23+0.24Myr1subscript˙𝑀superscriptsubscript1.780.230.24subscript𝑀direct-productsuperscriptyr1\dot{M}_{\star}=1.78_{-0.23}^{+0.24}~{}M_{\odot}\mathrm{yr}^{-1}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1.78 start_POSTSUBSCRIPT - 0.23 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.24 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, of stellar mass log10(M,host)=10.200.04+0.03Msubscriptlog10subscript𝑀hostsubscriptsuperscript10.200.030.04subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{host}})=10.20^{+0.03}_{-0.04}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , roman_host end_POSTSUBSCRIPT ) = 10.20 start_POSTSUPERSCRIPT + 0.03 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.04 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (see Figure 2), and IGH1, IGH2 as galaxies with stellar masses straddling that of HG 20221219A, log10(M,igh1)=10.600.02+0.02Msubscriptlog10subscript𝑀igh1subscriptsuperscript10.600.020.02subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{igh1}})=10.60^{+0.02}_{-0.02}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , igh1 end_POSTSUBSCRIPT ) = 10.60 start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, log10(M,igh2)=10.010.02+0.02Msubscriptlog10subscript𝑀igh2subscriptsuperscript10.010.020.02subscript𝑀direct-product\mathrm{log}_{10}(M_{\star,\mathrm{igh2}})=10.01^{+0.02}_{-0.02}~{}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ , igh2 end_POSTSUBSCRIPT ) = 10.01 start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (see Table 2).

Refer to caption
Figure 2: A non-parametric SED fit to observed spectroscopic and photometric data (red) for the host galaxy (PSO J257.6335+71.625257.633571.625257.6335+71.625257.6335 + 71.625) of FRB 20221219A, obtained with Keck-II/DEIMOS. The SED fit is performed using Prospector with residuals for the best posterior sample (green).

2.4 Sightline Crowding

The extreme degree of scattering in FRB 20221219A and apparently unusual presence of closely neighboring intervening galaxies suggests that the galaxy population along the FRB 20221219A LoS may be overcrowded. To evaluate whether this is true, we first calculate the number density of galaxies, n50(z)subscript𝑛50𝑧n_{50}(z)italic_n start_POSTSUBSCRIPT 50 end_POSTSUBSCRIPT ( italic_z ), along the full sightline within a 50 kpc comoving radius using data taken from Legacy Survey DR10. A 50 kpc radius is chosen as the smallest radius that fully encompasses the two interveners, accounting for uncertainty in the localization, as shown in Figure 3; our conclusions are not sensitive to this choice.

We then compare this count to the expected number densities using a Monte Carlo-based approach. The predicted number density of galaxies, n50(z)subscript𝑛50𝑧n_{50}(z)italic_n start_POSTSUBSCRIPT 50 end_POSTSUBSCRIPT ( italic_z ), can be calculated using the halo mass function (HMF), which determines the number density n𝑛nitalic_n for a specific halo mass Mhsubscript𝑀M_{h}italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT as

dndln(Mh)=f(σm)ρm,0Mhdln(σm1)dln(Mh)𝑑𝑛𝑑subscript𝑀𝑓subscript𝜎𝑚subscript𝜌𝑚0subscript𝑀𝑑superscriptsubscript𝜎𝑚1𝑑subscript𝑀\frac{dn}{d\ln(M_{h})}=f\left(\sigma_{m}\right)\frac{\rho_{m,0}}{M_{h}}\frac{d% \ln\left(\sigma_{m}^{-1}\right)}{d\ln(M_{h})}divide start_ARG italic_d italic_n end_ARG start_ARG italic_d roman_ln ( italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) end_ARG = italic_f ( italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_m , 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG divide start_ARG italic_d roman_ln ( italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_d roman_ln ( italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) end_ARG (2)

where ρm,0subscript𝜌𝑚0\rho_{m,0}italic_ρ start_POSTSUBSCRIPT italic_m , 0 end_POSTSUBSCRIPT is the matter density at z=0𝑧0z=0italic_z = 0, σmsubscript𝜎𝑚\sigma_{m}italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is the root-mean-square variance of the linear density field, which varies with the linear matter power spectrum, and f(σm)𝑓subscript𝜎𝑚f\left(\sigma_{m}\right)italic_f ( italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) represents a function of σmsubscript𝜎𝑚\sigma_{m}italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT that is redshift-independent. We utilize the Tinker et al. (2008) HMF as implemented in the hmf Python package (Murray et al., 2013). The Tinker HMF is precisely calibrated to redshifts z2less-than-or-similar-to𝑧2z\lesssim 2italic_z ≲ 2 and evolves with redshift in accordance with a predefined overdensity threshold.

To simulate expected values for n50(z)subscript𝑛50𝑧n_{50}(z)italic_n start_POSTSUBSCRIPT 50 end_POSTSUBSCRIPT ( italic_z ) within the relevant spatial domain, we set a target redshift to that of the host, z=0.554𝑧0.554z=0.554italic_z = 0.554, and populate the comoving volume with dark matter halos following a Poisson distribution, based on the mean number of halos set by the HMF. Each halo is randomly assigned a mass in accordance with the HMF distribution, bounded by a mass range of 1011superscript101110^{11}~{}10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT-1015Msuperscript1015subscript𝑀direct-product~{}10^{15}M_{\odot}10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. These limits are informed by the halo mass estimates for each intervener, log(Mh,igh1)=11.960.02+0.02Msubscript𝑀igh1subscriptsuperscript11.960.020.02subscript𝑀direct-product\log\left(M_{h,\mathrm{igh1}}\right)=11.96^{+0.02}_{-0.02}M_{\odot}roman_log ( italic_M start_POSTSUBSCRIPT italic_h , igh1 end_POSTSUBSCRIPT ) = 11.96 start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, log(Mh,igh2)=11.520.01+0.01Msubscript𝑀igh2subscriptsuperscript11.520.010.01subscript𝑀direct-product\log\left(M_{h,\mathrm{igh2}}\right)=11.52^{+0.01}_{-0.01}M_{\odot}roman_log ( italic_M start_POSTSUBSCRIPT italic_h , igh2 end_POSTSUBSCRIPT ) = 11.52 start_POSTSUPERSCRIPT + 0.01 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.01 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which we obtain assuming the Moster et al. (2010) stellar-to-halo mass relation

MMh(z)=2A(z)[(MhMA(z))β(z)+(MhMA(z))γ(z)]1,subscript𝑀subscript𝑀𝑧2𝐴𝑧superscriptdelimited-[]superscriptsubscript𝑀subscript𝑀𝐴𝑧𝛽𝑧superscriptsubscript𝑀subscript𝑀𝐴𝑧𝛾𝑧1\frac{M_{\star}}{M_{h}}(z)=2A(z)\left[\left(\frac{M_{h}}{M_{A}(z)}\right)^{-% \beta(z)}+\left(\frac{M_{h}}{M_{A}(z)}\right)^{\gamma(z)}\right]^{-1},divide start_ARG italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG ( italic_z ) = 2 italic_A ( italic_z ) [ ( divide start_ARG italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_z ) end_ARG ) start_POSTSUPERSCRIPT - italic_β ( italic_z ) end_POSTSUPERSCRIPT + ( divide start_ARG italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_z ) end_ARG ) start_POSTSUPERSCRIPT italic_γ ( italic_z ) end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (3)

parameterized by Girelli et al. (2020) using the ΛΛ\Lambdaroman_ΛCDM DUSTGRAIN-pathfinder simulation by A=0.04290.0026+0.0026,MA=11.870.06+0.06,β=0.990.07+0.08,γ=0.6690.015+0.016formulae-sequence𝐴superscriptsubscript0.04290.00260.0026formulae-sequencesubscript𝑀𝐴superscriptsubscript11.870.060.06formulae-sequence𝛽superscriptsubscript0.990.070.08𝛾superscriptsubscript0.6690.0150.016A=0.0429_{-0.0026}^{+0.0026},\ M_{A}=11.87_{-0.06}^{+0.06},\ \beta=0.99_{-0.07% }^{+0.08},\ \gamma=0.669_{-0.015}^{+0.016}italic_A = 0.0429 start_POSTSUBSCRIPT - 0.0026 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.0026 end_POSTSUPERSCRIPT , italic_M start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 11.87 start_POSTSUBSCRIPT - 0.06 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.06 end_POSTSUPERSCRIPT , italic_β = 0.99 start_POSTSUBSCRIPT - 0.07 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.08 end_POSTSUPERSCRIPT , italic_γ = 0.669 start_POSTSUBSCRIPT - 0.015 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.016 end_POSTSUPERSCRIPT for a redshift range 0.2z0.50.2𝑧0.50.2\leq z\leq 0.50.2 ≤ italic_z ≤ 0.5 (see Table 2 in Girelli et al., 2020).

For each simulation, we randomly select a sightline and calculate the number of halos within a 50 kpc comoving radius. Sightlines with a halo number density equal to or greater than the observed number density of two interveners are tallied, from which we construct a representative cumulative distribution function (CDF), thereby assessing the rarity of the sightline toward FRB 20221219A. We find that for FRB 20221219A CDF(n50(z)=2)0.99similar-tosubscript𝑛50𝑧20.99\left(n_{50}(z)=2\right)\sim 0.99( italic_n start_POSTSUBSCRIPT 50 end_POSTSUBSCRIPT ( italic_z ) = 2 ) ∼ 0.99, indicating that the field is significantly overcrowded. The number of intersected halos that would be expected along this sightline amounts to only Nhalos=0.18subscript𝑁halos0.18N_{\mathrm{halos}}=0.18italic_N start_POSTSUBSCRIPT roman_halos end_POSTSUBSCRIPT = 0.18.

Table 2: Summary of observed and derived parameters for the host galaxy HG 20221219A (see Figure 2), IGH1, and IGH2 (see Figure 3). We report spectroscopic redshifts (z𝑧zitalic_z), impact parameters (b𝑏bitalic_b), stellar masses (log10(M)subscript10subscript𝑀\log_{10}(M_{\star})roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT )), extinction-correct r-band magnitudes (r𝑟ritalic_r), and the star formation rate (SFR; M˙subscript˙𝑀\dot{M}_{\star}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT) of the host. The M˙subscript˙𝑀\dot{M}_{\star}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT values of the interveners were not measured due to the low S/N of their respective SEDs.
Object z𝑧zitalic_z b[kpc]𝑏delimited-[]kpcb~{}\left[\mathrm{kpc}\right]italic_b [ roman_kpc ] log10(M)[M]subscript10subscript𝑀delimited-[]subscript𝑀direct-product\log_{10}\left(M_{\star}\right)~{}\left[M_{\odot}\right]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) [ italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ] r[mag]𝑟delimited-[]magr~{}\left[\mathrm{mag}\right]italic_r [ roman_mag ] M˙[Myr1]subscript˙𝑀delimited-[]subscript𝑀direct-productsuperscriptyr1\dot{M}_{\star}~{}\left[M_{\odot}~{}\mathrm{yr}^{-1}\right]over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT [ italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ]
HG 20221219A 0.5540.5540.5540.554 10.200.04+0.03superscriptsubscript10.200.040.0310.20_{-0.04}^{+0.03}10.20 start_POSTSUBSCRIPT - 0.04 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.03 end_POSTSUPERSCRIPT 22.60.2+0.2superscriptsubscript22.60.20.222.6_{-0.2}^{+0.2}22.6 start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.2 end_POSTSUPERSCRIPT 1.780.23+0.24superscriptsubscript1.780.230.241.78_{-0.23}^{+0.24}1.78 start_POSTSUBSCRIPT - 0.23 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.24 end_POSTSUPERSCRIPT
IGH1 0.4920.4920.4920.492 4311.3+11.3superscriptsubscript4311.311.343_{-11.3}^{+11.3}43 start_POSTSUBSCRIPT - 11.3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 11.3 end_POSTSUPERSCRIPT 10.600.02+0.02superscriptsubscript10.600.020.0210.60_{-0.02}^{+0.02}10.60 start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT 21.60.2+0.2superscriptsubscript21.60.20.221.6_{-0.2}^{+0.2}21.6 start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.2 end_POSTSUPERSCRIPT
IGH2 0.4380.4380.4380.438 36.111.3+11.3superscriptsubscript36.111.311.336.1_{-11.3}^{+11.3}36.1 start_POSTSUBSCRIPT - 11.3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 11.3 end_POSTSUPERSCRIPT 10.010.02+0.02superscriptsubscript10.010.020.0210.01_{-0.02}^{+0.02}10.01 start_POSTSUBSCRIPT - 0.02 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.02 end_POSTSUPERSCRIPT 21.70.2+0.2superscriptsubscript21.70.20.221.7_{-0.2}^{+0.2}21.7 start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.2 end_POSTSUPERSCRIPT
Refer to captionRefer to caption
Figure 3: Legacy Survey (BASS) R-band cutout of the field surrounding FRB 20221219A. The white dashed-line circle indicates a comoving radius of 50 kpc. The host and intervening galaxies are outlined by white, purple, and magenta dashed circles (labels consistent with Table 2). The virial radii (R200subscriptR200\mathrm{R}_{200}roman_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT) are over-plotted as larger solid circles with consistent coloring. The 3σ𝜎\sigmaitalic_σ radio localization ellipse is over-plotted in violet. The upper panel shows the virial radius (R500subscriptR500\mathrm{R}_{500}roman_R start_POSTSUBSCRIPT 500 end_POSTSUBSCRIPT) of the intervening galaxy cluster J171039.6+713427 identified in the WISE survey (Wen et al., 2017).
Refer to caption
Figure 4: A schematic showing the FRB 20221219A sightline, including the two intervening galaxies (IGH1, IGH2), their respective impact parameters (b𝑏bitalic_b), and virial radii (R200subscriptR200\mathrm{R}_{200}roman_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT). We also show galaxy cluster J171039.6+713427 identified in the WISE survey (Wen et al., 2017), including its R500subscriptR500\mathrm{R}_{500}roman_R start_POSTSUBSCRIPT 500 end_POSTSUBSCRIPT. The presence of this cluster is meaningful when considering the DM contribution of the ICM, discussed in Section 3.1.

3 Analysis & Discussion

3.1 The Dispersion Measure Budget

The dispersion measure (DM) of the FRB is defined as the electron column density along the LoS DM=ne(z)𝑑lDMsubscript𝑛𝑒𝑧differential-d𝑙\mathrm{DM}=\int n_{e}(z)dlroman_DM = ∫ italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_z ) italic_d italic_l, where nesubscript𝑛𝑒n_{e}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the electron density. The DM, expressed in standard units of pccm3pcsuperscriptcm3\mathrm{pc}\ \mathrm{cm}^{-3}roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, is measured from the frequency-dependent delay in arrival times present in radio signals passing through cold, sparse plasmas along the LoS. As is typically done in the literature, we can express the measured DM for a source at redshift zhostsubscript𝑧hostz_{\mathrm{host}}italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT as the sum

DMobssubscriptDMobs\displaystyle\mathrm{DM}_{\mathrm{obs}}roman_DM start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT =DMmw+DMigm(zhost)absentsubscriptDMmwsubscriptDMigmsubscript𝑧host\displaystyle=\mathrm{DM}_{\mathrm{mw}}+\mathrm{DM}_{\mathrm{igm}}\left(z_{% \mathrm{host}}\right)= roman_DM start_POSTSUBSCRIPT roman_mw end_POSTSUBSCRIPT + roman_DM start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ( italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT ) (4)
+DMicm1+zcluster+DMigh1+zigh+DMhost1+zhostsubscriptDMicm1subscript𝑧clustersubscriptDMigh1subscript𝑧ighsubscriptDMhost1subscript𝑧host\displaystyle+\frac{\mathrm{DM}_{\mathrm{icm}}}{1+z_{\mathrm{cluster}}}+\frac{% \mathrm{DM}_{\mathrm{igh}}}{1+z_{\mathrm{igh}}}+\frac{\mathrm{DM}_{\mathrm{% host}}}{1+z_{\mathrm{host}}}+ divide start_ARG roman_DM start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_z start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT end_ARG + divide start_ARG roman_DM start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_z start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT end_ARG + divide start_ARG roman_DM start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT end_ARG

that includes components from the Milky Way (mw), intergalactic medium (igm), intersecting intracluster medium (icm), intervening galaxy halos (igh), and host galaxy (host). Due to cosmic variance, a statistical dependence of DMigmsubscriptDMigm\mathrm{DM}_{\mathrm{igm}}roman_DM start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT on redshift, z𝑧zitalic_z, is needed. The DMicmsubscriptDMicm\mathrm{DM}_{\mathrm{icm}}roman_DM start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT, DMighsubscriptDMigh\mathrm{DM}_{\mathrm{igh}}roman_DM start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT, and DMhostsubscriptDMhost\mathrm{DM}_{\mathrm{host}}roman_DM start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT terms require a further reduction of the rest-frame dispersion measures by factors of 1/(1+z)11𝑧1/(1+z)1 / ( 1 + italic_z ) for the respective redshifts of each. For simplicity, the host galaxy disk and halo, as well as the circumburst medium surrounding the FRB source, are treated as a single term. In the following sections, we will evaluate the DM contributions we expect from the Milky Way, IGM, ICM, and the intervening galaxies along the LoS. A schematic of the full sightline is shown in Figure 4.

Due to the presence of a closely neighboring galaxy cluster and two intervening galaxy halos, the DM budget is tight, leaving many of the diagnostics for DM outside the MW biased towards underestimates, as we describe later in §3.1.2 and §3.1.3. To alleviate this bias and remain conservative in our DM attributions to local media, we omit the inclusion of a DM contribution from the MW halo, as this is also a largely uncertain quantity. The main purpose of this is to maximize the host galaxy DM allotment, which, if determined to be smaller, would aid in our suggestion that the observed scattering does not occur in the host. Upper limits on the DM of the Milky Way halo have, however, been constrained to within DMhalo38less-than-or-similar-tosubscriptDMhalo38\mathrm{DM}_{\mathrm{halo}}\lesssim 38roman_DM start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT ≲ 38 pc cm-3 by Ravi et al. (2023a) using the non-repeating FRB 20220319D, and DMhalo52-111less-than-or-similar-tosubscriptDMhalo52-111\mathrm{DM}_{\mathrm{halo}}\lesssim 52~{}\text{-}~{}111roman_DM start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT ≲ 52 - 111 pc cm-3 by Cook et al. (2023) using the first FRB catalog from CHIME/FRB (CHIME/FRB Collaboration et al., 2021).

3.1.1 Milky Way and Halo

The DM contribution from the MW is characterized by the Galactic electron density model NE2001 (Cordes & Lazio, 2002), which estimates DMmw444+4pccm3similar-tosubscriptDMmwsubscriptsuperscript4444pcsuperscriptcm3\mathrm{DM}_{\mathrm{mw}}\sim 44^{+4}_{-4}~{}\mathrm{pc}\ \mathrm{cm}^{-3}roman_DM start_POSTSUBSCRIPT roman_mw end_POSTSUBSCRIPT ∼ 44 start_POSTSUPERSCRIPT + 4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 4 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (assuming 10%similar-toabsentpercent10\sim 10\%∼ 10 % uncertainty, in accordance with Ocker et al., 2020) along the LoS toward FRB 20221219A (positioned at 102.93superscript102.93102.93^{\circ}102.93 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, 33.53superscript33.5333.53^{\circ}33.53 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT), through the Galaxy.

3.1.2 Intergalactic Medium

While the IGM is a dominant contributor to the DM budgets of FRB signals (Zhang et al., 2021; Walker et al., 2023), it does not appear to noticeably scatter them. We assume this based on both modeling of gas densities in the IGM (Macquart & Koay, 2013) and the absence of a universally applicable correlation between observed scattering timescales and extragalactic DMs (CHIME/FRB Collaboration et al., 2021; Chawla et al., 2022). The mean DM contribution for a constant co-moving density in the IGM is given by (e.g., McQuinn, 2014)

DMigm(z)=ne0DH0z𝑑z(1+z)E(z)delimited-⟨⟩subscriptDMigm𝑧subscript𝑛subscript𝑒0subscript𝐷Hsuperscriptsubscript0𝑧differential-dsuperscript𝑧1superscript𝑧𝐸superscript𝑧\langle\mathrm{DM}_{\mathrm{igm}}(z)\rangle=n_{e_{0}}D_{\mathrm{H}}\int_{0}^{z% }dz^{\prime}\frac{\left(1+z^{\prime}\right)}{E\left(z^{\prime}\right)}⟨ roman_DM start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ( italic_z ) ⟩ = italic_n start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT italic_d italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT divide start_ARG ( 1 + italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_E ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG (5)

where ne0=2.2×107×fIGMsubscript𝑛subscript𝑒02.2superscript107subscript𝑓IGMn_{e_{0}}=2.2\times 10^{-7}\times f_{\mathrm{IGM}}italic_n start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 2.2 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT × italic_f start_POSTSUBSCRIPT roman_IGM end_POSTSUBSCRIPT is the IGM electron density at z=0𝑧0z=0italic_z = 0, fIGMsubscript𝑓IGMf_{\mathrm{IGM}}italic_f start_POSTSUBSCRIPT roman_IGM end_POSTSUBSCRIPT being the IGM baryon fraction based on Planck 18 cosmology (Planck Collaboration et al., 2020), DH=c/H0subscript𝐷H𝑐subscript𝐻0D_{\mathrm{H}}=c/H_{0}italic_D start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = italic_c / italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the Hubble distance and E(z)=𝐸𝑧absentE(z)=italic_E ( italic_z ) = [Ωm(1+z)3+1Ωm]1/2superscriptdelimited-[]subscriptΩmsuperscript1𝑧31subscriptΩm12\left[\Omega_{\mathrm{m}}(1+z)^{3}+1-\Omega_{\mathrm{m}}\right]^{1/2}[ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ( 1 + italic_z ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 1 - roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT for a flat ΛΛ\Lambdaroman_ΛCDM universe with a matter density ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT. With this, we generate and fit a DM distribution using the probability density function (PDF) (Macquart et al., 2020; Zhang et al., 2021)

pigm(Δ)=AΔbexp[(ΔaC0)22a2σDM2],Δ>0formulae-sequencesubscript𝑝igmΔ𝐴superscriptΔ𝑏superscriptsuperscriptΔ𝑎subscript𝐶022superscript𝑎2superscriptsubscript𝜎DM2Δ0p_{\mathrm{igm}}(\Delta)=A\Delta^{-b}\exp\left[-\frac{\left(\Delta^{-a}-C_{0}% \right)^{2}}{2a^{2}\sigma_{\mathrm{DM}}^{2}}\right],\quad\Delta>0italic_p start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ( roman_Δ ) = italic_A roman_Δ start_POSTSUPERSCRIPT - italic_b end_POSTSUPERSCRIPT roman_exp [ - divide start_ARG ( roman_Δ start_POSTSUPERSCRIPT - italic_a end_POSTSUPERSCRIPT - italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] , roman_Δ > 0 (6)

where Δ=DMigm/DMigmΔsubscriptDMigmdelimited-⟨⟩subscriptDMigm\Delta=\mathrm{DM}_{\mathrm{igm}}/\left\langle\mathrm{DM}_{\mathrm{igm}}\right\rangleroman_Δ = roman_DM start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT / ⟨ roman_DM start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ⟩, b𝑏bitalic_b depends on the halo gas density profile. We assume a=b=3𝑎𝑏3a=b=3italic_a = italic_b = 3 (Macquart et al., 2020). The effective standard deviation is σDMsubscript𝜎DM\sigma_{\mathrm{DM}}italic_σ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT, and C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT impacts the transverse position, which is fitted for. We also assume values for A𝐴Aitalic_A, σDMsubscript𝜎DM\sigma_{\mathrm{DM}}italic_σ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT, and C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, derived from Illustris TNG simulation data by Zhang et al. (2021).

The IGM sightline for FRB 20221219A is, however, slightly complicated by the presence of a proximate foreground galaxy cluster J171039.6+713427 cataloged in the WISE survey and DR9 group catalog, positioned at a redshift of zcluster=0.16subscript𝑧cluster0.16z_{\mathrm{cluster}}=0.16italic_z start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT = 0.16 with mass log10(Mcluster)14.10.2+0.2Msimilar-to-or-equalssubscript10subscript𝑀clustersubscriptsuperscript14.10.20.2subscript𝑀direct-product\log_{10}(M_{\mathrm{cluster}})\simeq 14.1^{+0.2}_{-0.2}M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT ) ≃ 14.1 start_POSTSUPERSCRIPT + 0.2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (the quoted 0.2 dex uncertainty for cluster masses log10(Mcluster)14Mgreater-than-or-equivalent-tosubscriptlog10subscript𝑀cluster14subscript𝑀direct-product\mathrm{log}_{10}\left(M_{\mathrm{cluster}}\right)\gtrsim 14M_{\odot}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT ) ≳ 14 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) and a LoS offset of 3.2similar-to-or-equalsabsentsuperscript3.2\simeq 3.2^{\prime}≃ 3.2 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, corresponding impact parameter b55311+11similar-to𝑏subscriptsuperscript5531111b\sim 553^{+11}_{-11}italic_b ∼ 553 start_POSTSUPERSCRIPT + 11 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 11 end_POSTSUBSCRIPT kpc (Wen et al., 2017). We evaluate the DM contribution of the intracluster medium (ICM) surrounding the cluster at its measured impacted parameter. The ne,icmsubscript𝑛𝑒icmn_{e,\mathrm{icm}}italic_n start_POSTSUBSCRIPT italic_e , roman_icm end_POSTSUBSCRIPT at b𝑏bitalic_b is estimated using a Navarro-Frenk-White profile as implemented by Vikhlinin et al. (2006) and Prochaska et al. (2019), contributing DMcluster14212+12similar-to-or-equalssubscriptDMclustersubscriptsuperscript1421212\mathrm{DM}_{\mathrm{cluster}}\simeq 142^{+12}_{-12}roman_DM start_POSTSUBSCRIPT roman_cluster end_POSTSUBSCRIPT ≃ 142 start_POSTSUPERSCRIPT + 12 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 12 end_POSTSUBSCRIPT pc cm-3. This uncertainty does not, however, account for spatial asymmetry in the baryon halos, which may be significant (Connor et al., 2023).

To estimate the collective PDF, Pigm+icm(DM), for the DM contributions from the IGM and ICM, we assume a Gaussian PDF for picm(Δ)subscript𝑝icmΔp_{\mathrm{icm}}(\Delta)italic_p start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT ( roman_Δ ), given by

picm(Δ)=12πσDM2exp((Δμ)22σDM2)subscript𝑝icmΔ12𝜋superscriptsubscript𝜎DM2superscriptΔ𝜇22superscriptsubscript𝜎DM2p_{\mathrm{icm}}(\Delta)=\frac{1}{\sqrt{2\pi\sigma_{\mathrm{DM}}^{2}}}\exp% \left(-\frac{(\Delta-\mu)^{2}}{2\sigma_{\mathrm{DM}}^{2}}\right)italic_p start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT ( roman_Δ ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π italic_σ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG roman_exp ( - divide start_ARG ( roman_Δ - italic_μ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (7)

where Δ=DMicm/DMicmΔsubscriptDMicmdelimited-⟨⟩subscriptDMicm\Delta=\mathrm{DM}_{\mathrm{icm}}/\left\langle\mathrm{DM}_{\mathrm{icm}}\right\rangleroman_Δ = roman_DM start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT / ⟨ roman_DM start_POSTSUBSCRIPT roman_icm end_POSTSUBSCRIPT ⟩ and convolve with pigm(Δ)subscript𝑝igmΔp_{\mathrm{igm}}(\Delta)italic_p start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ( roman_Δ ) (see Eq. 6) to obtain Pigm+icm(DM). Taking the 50th percentile of the convolution (uncertainties given by the 15th and 85th percentiles), we estimate DMigm+icm,50=60859+60pccm3subscriptDMigmicm50superscriptsubscript6085960pcsuperscriptcm3\mathrm{DM}_{\mathrm{igm+icm},50}=608_{-59}^{+60}\mathrm{pc~{}cm}^{-3}roman_DM start_POSTSUBSCRIPT roman_igm + roman_icm , 50 end_POSTSUBSCRIPT = 608 start_POSTSUBSCRIPT - 59 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 60 end_POSTSUPERSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT in the host frame, as shown in Figure 5 and Table 3.

Refer to caption
Figure 5: PDFs for DMigm (black), DMigm+cluster (magenta), DMHost (purple), where P(DM |z=0.554|~{}z=0.554| italic_z = 0.554) is in units of [pc cm-3]-1. The 50th percentile values of each PDF, as well as 15th and 85th percentile uncertainties are indicated by black dotted lines and shaded regions, respectively. Due to the narrowness of the Milky Way and intervening galaxy PDFs (see §3.1.1 and §3.1.3), we omit them above for clarity.

3.1.3 Intervening Galaxies

We estimate DMigh for the CGM of both intervening galaxies based on HII column densities (NHII) inferred from direct NHI measurements made for similar nearby galaxies in the COS-Halos survey (for 44 gaseous halos with T104similar-to𝑇superscript104T\sim 10^{4}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K photoionized CGMs at impact parameters of b160less-than-or-similar-to𝑏160b\lesssim 160italic_b ≲ 160 kpc near z0.2similar-to𝑧0.2z\sim 0.2italic_z ∼ 0.2; Werk et al., 2014). Inferences of NHII rely on photoionization modeling relating NHI to NH, which shows that the CGMs are highly ionized, such that nHII/nH99%greater-than-or-equivalent-tosubscriptnHIIsubscriptnHpercent99\mathrm{n}_{\mathrm{HII}}/\mathrm{n}_{\mathrm{H}}\gtrsim 99\%roman_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT / roman_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ≳ 99 %. Interpolating measurements from the COS-Halos survey across the full (NHII, b𝑏bitalic_b, Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT) parameter space, we converge on the regions applicable to b𝑏bitalic_b and Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT for both interveners and estimate log10(NHII,igh1)19.50.1+0.1similar-to-or-equalssubscriptlog10subscriptNHIIigh1subscriptsuperscript19.50.10.1\mathrm{log}_{10}(\mathrm{N}_{\mathrm{HII},\mathrm{igh1}})\simeq 19.5^{+0.1}_{% -0.1}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( roman_N start_POSTSUBSCRIPT roman_HII , igh1 end_POSTSUBSCRIPT ) ≃ 19.5 start_POSTSUPERSCRIPT + 0.1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.1 end_POSTSUBSCRIPT cm-2 and log10(NHII,igh2)19.60.1+0.1similar-to-or-equalssubscriptlog10subscriptNHIIigh2subscriptsuperscript19.60.10.1\mathrm{log}_{10}(\mathrm{N}_{\mathrm{HII},\mathrm{igh2}})\simeq 19.6^{+0.1}_{% -0.1}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( roman_N start_POSTSUBSCRIPT roman_HII , igh2 end_POSTSUBSCRIPT ) ≃ 19.6 start_POSTSUPERSCRIPT + 0.1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.1 end_POSTSUBSCRIPT cm-2, which correspond to DMigh1(z=0.492)105+5pccm3similar-to-or-equalsdelimited-⟨⟩subscriptDMigh1𝑧0.492subscriptsuperscript1055pcsuperscriptcm3\langle\mathrm{DM}_{\mathrm{igh1}}(z=0.492)\rangle\simeq 10^{+5}_{-5}\mathrm{% pc}\mathrm{~{}cm}^{-3}⟨ roman_DM start_POSTSUBSCRIPT igh1 end_POSTSUBSCRIPT ( italic_z = 0.492 ) ⟩ ≃ 10 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and DMigh2(z=0.438)145+5pccm3similar-to-or-equalsdelimited-⟨⟩subscriptDMigh2𝑧0.438subscriptsuperscript1455pcsuperscriptcm3\langle\mathrm{DM}_{\mathrm{igh2}}(z=0.438)\rangle\simeq 14^{+5}_{-5}\mathrm{% pc}\mathrm{~{}cm}^{-3}⟨ roman_DM start_POSTSUBSCRIPT igh2 end_POSTSUBSCRIPT ( italic_z = 0.438 ) ⟩ ≃ 14 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Similar to the ICM, we assume a Gaussian PDF for pigh(Δ)subscript𝑝ighΔp_{\mathrm{igh}}(\Delta)italic_p start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT ( roman_Δ ). In estimating DMHost, we conservatively neglect the presence of any hot gas in the CGM of the interveners, which is again challenging to include in the DM budget due to tight constraints placed by the large contributions of the IGM and neighboring galaxy cluster.

3.1.4 Host Galaxy

To estimate DMhost, we convolve pigh+icmsubscript𝑝ighicmp_{\mathrm{igh+icm}}italic_p start_POSTSUBSCRIPT roman_igh + roman_icm end_POSTSUBSCRIPT with pMWsubscript𝑝MWp_{\mathrm{MW}}italic_p start_POSTSUBSCRIPT roman_MW end_POSTSUBSCRIPT, pigh1subscript𝑝igh1p_{\mathrm{igh1}}italic_p start_POSTSUBSCRIPT igh1 end_POSTSUBSCRIPT and pigh2subscript𝑝igh2p_{\mathrm{igh2}}italic_p start_POSTSUBSCRIPT igh2 end_POSTSUBSCRIPT to obtain PhostsubscriptPhost\mathrm{P}_{\mathrm{host}}roman_P start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT(DM). We assume Gaussian PDFs (similar to Eq. 7) for the MW and both interveners. Taking the 50th percentile of PhostsubscriptPhost\mathrm{P}_{\mathrm{host}}roman_P start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT(DM), we estimate DMhost,50=6243+48subscriptDMhost50superscriptsubscript624348\mathrm{DM}_{\mathrm{host,50}}=62_{-43}^{+48}roman_DM start_POSTSUBSCRIPT roman_host , 50 end_POSTSUBSCRIPT = 62 start_POSTSUBSCRIPT - 43 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 48 end_POSTSUPERSCRIPT pc cm-3 in the host-galaxy frame (see Figure 5 and Table 3). Given the large uncertainties in modeling the DMs contributed by the intervening galaxy and cluster halos, we also estimated DMhost,50subscriptDMhost50\mathrm{DM}_{\mathrm{host,50}}roman_DM start_POSTSUBSCRIPT roman_host , 50 end_POSTSUBSCRIPT in the absence of their contributions. In that case, we find DMhost,50=16593+81subscriptDMhost50superscriptsubscript1659381\mathrm{DM}_{\mathrm{host,50}}=165_{-93}^{+81}roman_DM start_POSTSUBSCRIPT roman_host , 50 end_POSTSUBSCRIPT = 165 start_POSTSUBSCRIPT - 93 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 81 end_POSTSUPERSCRIPT pc cm-3.

Table 3: Estimated DM contributions (in pc cm-3) for each term in Eq. 4 derived from the PDFs in Figure 5, quoted as the 50th percentile with 15th and 85th percentile uncertainties.
P(DM)PDM\mathrm{P(DM)}roman_P ( roman_DM ) DM50subscriptDM50\mathrm{DM}_{50}roman_DM start_POSTSUBSCRIPT 50 end_POSTSUBSCRIPT [pccm3]delimited-[]pcsuperscriptcm3\left[\mathrm{pc~{}cm}^{-3}\right][ roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ]
MWMW\mathrm{MW}roman_MW 4410+10subscriptsuperscript44101044^{+10}_{-10}44 start_POSTSUPERSCRIPT + 10 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 10 end_POSTSUBSCRIPT
IGM+ICMIGMICM\mathrm{IGM+ICM}roman_IGM + roman_ICM 60859+60superscriptsubscript6085960608_{-59}^{+60}608 start_POSTSUBSCRIPT - 59 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 60 end_POSTSUPERSCRIPT
IGH1IGH1\mathrm{IGH1}IGH1 105+5subscriptsuperscript105510^{+5}_{-5}10 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT
IGH2IGH2\mathrm{IGH2}IGH2 145+5subscriptsuperscript145514^{+5}_{-5}14 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT
HG 20221219AHG20221219A\mathrm{HG\,20221219A}roman_HG 20221219 roman_A 6243+48superscriptsubscript62434862_{-43}^{+48}62 start_POSTSUBSCRIPT - 43 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 48 end_POSTSUPERSCRIPT

3.2 The Scattering Budget

We observe a high scattering timescale of τobs=19.22.7+2.7subscript𝜏obssuperscriptsubscript19.22.72.7\tau_{\mathrm{obs}}=19.2_{-2.7}^{+2.7}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = 19.2 start_POSTSUBSCRIPT - 2.7 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 2.7 end_POSTSUPERSCRIPT ms for FRB 20221219A at 1.4 GHz. The pulse broadening, as measured for three individual sub-bands, is best described by τναproportional-to𝜏superscript𝜈𝛼\tau\propto\nu^{-\alpha}italic_τ ∝ italic_ν start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT, α=2.6±1.8𝛼plus-or-minus2.61.8\alpha=2.6\pm 1.8italic_α = 2.6 ± 1.8. To assess the origins of the scattering seen in FRB 20221219A, we can divide τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT for a source at redshift zhostsubscript𝑧hostz_{\mathrm{host}}italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT into a sum of terms corresponding to the various scattering media along the LoS as

τ(ν)obs=τmw(ν)+τigm(ν,z)+τigh(ν)(1+zigh)3+τhost(ν)(1+zhost)3+τcbm(ν)(1+zhost)3,𝜏subscript𝜈obsabsentsubscript𝜏mw𝜈limit-fromsubscript𝜏igm𝜈𝑧missing-subexpressionsubscript𝜏igh𝜈superscript1subscript𝑧igh3subscript𝜏host𝜈superscript1subscript𝑧host3subscript𝜏cbm𝜈superscript1subscript𝑧host3\begin{aligned} \tau(\nu)_{\mathrm{obs}}=&\tau_{\mathrm{mw}}(\nu)+\tau_{% \mathrm{igm}}(\nu,z)~{}+\\ &\frac{\tau_{\mathrm{igh}}(\nu)}{\left(1+z_{\mathrm{igh}}\right)^{3}}+\frac{% \tau_{\mathrm{host}}(\nu)}{\left(1+z_{\mathrm{host}}\right)^{3}}~{}+\frac{\tau% _{\mathrm{cbm}}(\nu)}{\left(1+z_{\mathrm{host}}\right)^{3}}\end{aligned},start_ROW start_CELL italic_τ ( italic_ν ) start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = end_CELL start_CELL italic_τ start_POSTSUBSCRIPT roman_mw end_POSTSUBSCRIPT ( italic_ν ) + italic_τ start_POSTSUBSCRIPT roman_igm end_POSTSUBSCRIPT ( italic_ν , italic_z ) + end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL divide start_ARG italic_τ start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT ( italic_ν ) end_ARG start_ARG ( 1 + italic_z start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_τ start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT ( italic_ν ) end_ARG start_ARG ( 1 + italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_τ start_POSTSUBSCRIPT roman_cbm end_POSTSUBSCRIPT ( italic_ν ) end_ARG start_ARG ( 1 + italic_z start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW , (8)

that, similarly to the DM budget, includes lens rest-frame components from the Milky Way (mw), the intergalactic medium (igm), intervening galaxies (igh), the host galaxy (host), and now the circumburst medium (cbm). We adopt a nominal power-law frequency scaling, τ(ν)ν4proportional-to𝜏𝜈superscript𝜈4\tau(\nu)\propto\nu^{-4}italic_τ ( italic_ν ) ∝ italic_ν start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. The last three terms of Eq. 8 scale with redshift to account for the fact that, for an observation frequency ν𝜈\nuitalic_ν, scattering goes as ν=ν(1+z)superscript𝜈𝜈1𝑧\nu^{\prime}=\nu(1+z)italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_ν ( 1 + italic_z ) in the galaxy’s rest frame and time dilation goes as (1+z)1𝑧(1+z)( 1 + italic_z ) (Macquart & Koay, 2013; Cordes et al., 2016). In the following sections, we will consider each scattering medium along the line of sight individually.

3.2.1 Electron Density Fluctuations

Density fluctuations are modeled within a volume of ionized cloudlets, parameterized by the combined quantity F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG, the density fluctuation parameter, defined as

F~=ζε2fv(lo2li)1/3(pc2km)1/3,~𝐹𝜁superscript𝜀2subscript𝑓vsuperscriptsuperscriptsubscript𝑙o2subscript𝑙i13superscriptsuperscriptpc2km13\widetilde{F}=\frac{\zeta\varepsilon^{2}}{f_{\mathrm{v}}\left(l_{\mathrm{o}}^{% 2}l_{\mathrm{i}}\right)^{1/3}}~{}~{}~{}\left(\mathrm{pc}^{2}\mathrm{~{}km}% \right)^{-1/3},over~ start_ARG italic_F end_ARG = divide start_ARG italic_ζ italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_f start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT ( italic_l start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_l start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG ( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT , (9)

where losubscript𝑙ol_{\mathrm{o}}italic_l start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT is the outer scale of the turbulence power spectrum (in pc), lisubscript𝑙il_{\mathrm{i}}italic_l start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT is the inner scale (in km), fvsubscript𝑓vf_{\mathrm{v}}italic_f start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT is the volume filling factor, ϵ2superscriptitalic-ϵ2\epsilon^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the intra-cloudlet variance in the electron density fluctuations, and ζ𝜁\zetaitalic_ζ is the inter-cloudlet variation with respect to the mean electron density within the cloudlets. The fluctuation parameter F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG, in combination with the geometric factor G𝐺Gitalic_G, strongly affects τ𝜏\tauitalic_τ.

We assess the physical validity of two scattering scenarios: scattering by a uniform, extended CGM (§3.2.4), and scattering by a single ionized cloudlet (§3.2.5).

3.2.2 Line of Sight Geometry

In addition to the electron density fluctuations within intervening media, the position of those media with respect to the source and observer serves as an impactful factor in determining their scattering power. Treatments in Cordes et al. (2016), Ocker et al. (2021), and Cordes et al. (2022) define a dimensionless geometric factor G𝐺Gitalic_G to describe this. In Euclidean space, it is defined as

G=layers(1s/d)𝑑shosts(1s/d)𝑑s,𝐺subscriptlayer𝑠1𝑠𝑑differential-d𝑠subscripthost𝑠1𝑠𝑑differential-d𝑠G=\frac{\int_{\mathrm{layer}}s(1-s/d)ds}{\int_{\mathrm{host}}s(1-s/d)ds},italic_G = divide start_ARG ∫ start_POSTSUBSCRIPT roman_layer end_POSTSUBSCRIPT italic_s ( 1 - italic_s / italic_d ) italic_d italic_s end_ARG start_ARG ∫ start_POSTSUBSCRIPT roman_host end_POSTSUBSCRIPT italic_s ( 1 - italic_s / italic_d ) italic_d italic_s end_ARG , (10)

where s𝑠sitalic_s represents a path element of the full path length d𝑑ditalic_d, and G𝐺Gitalic_G is of order unity for a source contained within in the scattering medium (e.g., the host galaxy), such that the distance to the source far exceeds the LoS extent of the scattering medium. Typically, however, G𝐺Gitalic_G strongly depends on the distribution of free electrons along the LoS and can increase by many orders of magnitude for intervening galaxies. In the case of FRB 20221219A, the intervening galaxies we consider are at non-negligible redshifts, so the expression for G𝐺Gitalic_G can be rewritten (see Cordes et al. (2022) for the complete derivation) as

G(z,zs)=2dsldloLdso,𝐺subscript𝑧subscript𝑧s2subscript𝑑slsubscript𝑑lo𝐿subscript𝑑soG\left(z_{\ell},z_{\mathrm{s}}\right)=\frac{2d_{\mathrm{sl}}d_{\mathrm{lo}}}{% Ld_{\mathrm{so}}},italic_G ( italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ) = divide start_ARG 2 italic_d start_POSTSUBSCRIPT roman_sl end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT roman_lo end_POSTSUBSCRIPT end_ARG start_ARG italic_L italic_d start_POSTSUBSCRIPT roman_so end_POSTSUBSCRIPT end_ARG , (11)

where dslsubscript𝑑sld_{\mathrm{sl}}italic_d start_POSTSUBSCRIPT roman_sl end_POSTSUBSCRIPT, dlosubscript𝑑lod_{\mathrm{lo}}italic_d start_POSTSUBSCRIPT roman_lo end_POSTSUBSCRIPT, dsosubscript𝑑sod_{\mathrm{so}}italic_d start_POSTSUBSCRIPT roman_so end_POSTSUBSCRIPT are, respectively, the source-to-intervener, intervener-to-observer, and source-to-observer angular diameter distances (see schematic in Figure 4), and L𝐿Litalic_L is the thickness of the scattering layer.

Given this geometric factor, a source at redshift zssubscript𝑧sz_{\mathrm{s}}italic_z start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, and a scattering region in an intervening galaxy at zsubscript𝑧z_{\ell}italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT, the scattering timescale can be expressed as

τ(DM,ν,z,zs)𝜏subscriptDM𝜈subscript𝑧subscript𝑧s\displaystyle\tau\left(\mathrm{DM}_{\ell},\nu,z_{\ell},z_{\mathrm{s}}\right)italic_τ ( roman_DM start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT , italic_ν , italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ) (12)
0.48ms×AτF~G(z,zs)DM,1002ν4(1+z)3,similar-to-or-equalsabsent0.48mssubscript𝐴𝜏~𝐹𝐺subscript𝑧subscript𝑧ssuperscriptsubscriptDM1002superscript𝜈4superscript1subscript𝑧3\displaystyle\quad\simeq 0.48\mathrm{~{}ms}\times\frac{A_{\tau}\widetilde{F}G% \left(z_{\ell},z_{\mathrm{s}}\right)\mathrm{DM}_{\ell,100}^{2}}{\nu^{4}\left(1% +z_{\ell}\right)^{3}},≃ 0.48 roman_ms × divide start_ARG italic_A start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT over~ start_ARG italic_F end_ARG italic_G ( italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ) roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( 1 + italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ,

where DMsubscriptDM\mathrm{DM}_{\ell}roman_DM start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT is in the rest frame of the intervening galaxy, which contributes to the measured DM as DM/(1+z)subscriptDM1subscript𝑧\mathrm{DM}_{\ell}/\left(1+z_{\ell}\right)roman_DM start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT / ( 1 + italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ), with ν𝜈\nuitalic_ν in GHzGHz\mathrm{GHz}roman_GHz and DM,100DM/(100pccm3)subscriptDM100subscriptDM100pcsuperscriptcm3\mathrm{DM}_{\ell,100}\equiv\mathrm{DM}_{\ell}/\left(100~{}\mathrm{pc}~{}% \mathrm{cm}^{-3}\right)roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT ≡ roman_DM start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT / ( 100 roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) (Cordes et al., 2022). The quantity Aτsubscript𝐴𝜏A_{\tau}italic_A start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, in turn, accounts for the shape of the pulse broadening function (typically falling between 1/6-116-11/6~{}\text{-}~{}11 / 6 - 1, though we henceforth assume Aτ=1subscript𝐴𝜏1A_{\tau}=1italic_A start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 1), and depends on both the spectral index β𝛽\betaitalic_β and inner scale lisubscript𝑙il_{\mathrm{i}}italic_l start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT of the scattering medium.

3.2.3 Host Galaxy & Circumburst Environment

It has been shown by Cordes et al. (2022) that the majority of FRBs detected by CHIME/FRB (repeating and non-repeating), assuming host galaxy-dominated scattering, emerge from an ISM with F~1less-than-or-similar-to~𝐹1\widetilde{F}\lesssim 1over~ start_ARG italic_F end_ARG ≲ 1 (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT, with the exception of the heavily scattered FRB 20191221A, which still suggests a host F~10less-than-or-similar-to~𝐹10\widetilde{F}\lesssim 10over~ start_ARG italic_F end_ARG ≲ 10 (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT (see their Fig. 5; Cordes et al., 2022).

We find HG 20221219A to be MW-like in its star formation rate (SFR; M˙subscript˙𝑀\dot{M}_{\star}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT). While the M,hostsubscript𝑀hostM_{\star,\mathrm{host}}italic_M start_POSTSUBSCRIPT ⋆ , roman_host end_POSTSUBSCRIPT is a factor of 4similar-toabsent4\sim 4∼ 4 smaller than that of the MW, and its SFR of M˙=1.780.23+0.24Myr1subscript˙𝑀superscriptsubscript1.780.230.24subscript𝑀direct-productsuperscriptyr1\dot{M}_{\star}=1.78_{-0.23}^{+0.24}~{}M_{\odot}\mathrm{yr}^{-1}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1.78 start_POSTSUBSCRIPT - 0.23 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.24 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, is comparable to the MW’s M˙=1.650.19+0.19Myr1subscript˙𝑀subscriptsuperscript1.650.190.19subscript𝑀direct-productsuperscriptyr1\dot{M}_{\star}=1.65^{+0.19}_{-0.19}~{}M_{\odot}\mathrm{yr}^{-1}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 1.65 start_POSTSUPERSCRIPT + 0.19 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.19 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Licquia & Newman, 2015). Under the assumption that the two galaxies are comparable in their properties, we can apply the established relation between τ𝜏\tauitalic_τ and DM values of Galactic pulsars (Cordes et al., 2016) to the host galaxy, and infer its scattering contribution based on DMhost estimated in Section 3.1.4. First discussed by Sutton (1971), Rickett (1977) and Cordes et al. (1991), and most recently measured by Cordes et al. (2022), this relation was fitted against scattering timescales and DMs for 568 Galactic pulsars. The canonical fitting function used for the pulsar data is τ^(DM)=A×DMa(1+B×DMb)^𝜏DM𝐴superscriptDM𝑎1𝐵superscriptDM𝑏\widehat{\tau}(\mathrm{DM})=A\times\mathrm{DM}^{a}\left(1+B\times\mathrm{DM}^{% b}\right)over^ start_ARG italic_τ end_ARG ( roman_DM ) = italic_A × roman_DM start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT ( 1 + italic_B × roman_DM start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT ) (Ramachandran et al., 1997), for which Cordes et al. (2016) found the empirical τ𝜏\tauitalic_τ-DM relation

[τ^(DM,ν)]mw,psr=1.90×107ms×ναDM1.5subscriptdelimited-[]^𝜏DM𝜈mwpsr1.90superscript107mssuperscript𝜈𝛼superscriptDM1.5\displaystyle\left[\widehat{\tau}\left(\mathrm{DM},\nu\right)\right]_{\mathrm{% mw},\mathrm{psr}}=1.90\times 10^{-7}\mathrm{~{}ms}\times\nu^{-\alpha}\mathrm{% DM}^{1.5}[ over^ start_ARG italic_τ end_ARG ( roman_DM , italic_ν ) ] start_POSTSUBSCRIPT roman_mw , roman_psr end_POSTSUBSCRIPT = 1.90 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT roman_ms × italic_ν start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT roman_DM start_POSTSUPERSCRIPT 1.5 end_POSTSUPERSCRIPT (13)
×(1+3.55×105DM3.0),absent13.55superscript105superscriptDM3.0\displaystyle\times\left(1+3.55\times 10^{-5}\mathrm{DM}^{3.0}\right),× ( 1 + 3.55 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT roman_DM start_POSTSUPERSCRIPT 3.0 end_POSTSUPERSCRIPT ) ,

at frequencies ν𝜈\nuitalic_ν in GHzGHz\mathrm{GHz}roman_GHz, with scatter σlogτ=0.76subscript𝜎𝜏0.76\sigma_{\log\tau}=0.76italic_σ start_POSTSUBSCRIPT roman_log italic_τ end_POSTSUBSCRIPT = 0.76 dex (Bhat et al., 2004; Cordes & Chatterjee, 2019).

Figure 6 shows the fitted τ𝜏\tauitalic_τ-DM relation from Cordes et al. (2016) for Galactic pulsars. The steepening at large DMs occurs due to the cloudy nature of the inner Galaxy, where high-DM pulsars are found, as opposed to those in the outer regions of the galaxy or near the solar neighborhood (Cordes et al., 1991; Cordes & Chatterjee, 2019). We over-plot a sample of well-localized FRBs in comparison whose host DMs are inferred probabilistically using similar methods as those outlined in §3.1. FRB 20221219A emerges from these populations as a clearly over-scattered outlier333FRB scattering timescales have scaled up by a factor 3, under the assumption that the scattered waves are planar and not spherical (Cordes et al., 2016). (to 6.37σlog10τsubscript𝜎subscriptlog10𝜏\sigma_{\mathrm{log}_{10}\tau}italic_σ start_POSTSUBSCRIPT roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT for [τ^(DM,ν)]MW,psrsubscriptdelimited-[]^𝜏DM𝜈MWpsr[\widehat{\tau}(\mathrm{DM},\nu)]_{\mathrm{MW},\mathrm{psr}}[ over^ start_ARG italic_τ end_ARG ( roman_DM , italic_ν ) ] start_POSTSUBSCRIPT roman_MW , roman_psr end_POSTSUBSCRIPT) given its inferred DMhost, reflecting just how unique this event is in the context of most FRBs that have been detected to date. We, therefore, tentatively suggest that the host is unlikely to contribute significantly to the observed scattering timescale. Still, there remains the possibility that a compact scattering structure (e.g., a plasma lens) exists within HG 20221219A and intersects the LoS. Such a structure could lead to strong deviations from the τ𝜏\tauitalic_τ-DM relation and weaken the inference made on the scattering prospects within HG 20221219A. As this scenario is difficult to constrain without scintillation information, we do not expand on it in this work.

Refer to caption
Figure 6: τ𝜏\tauitalic_τ-DM relation for Galactic pulsars. The fitted line (solid black) and ±1σplus-or-minus1𝜎\pm 1\sigma± 1 italic_σ variations (dashed black) are based on measurements and upper limits on τ𝜏\tauitalic_τ for Galactic pulsars (568 in total; data provided by courtesy of J. Cordes and also published in Cordes et al., 2016). The scattering timescales and inferred host galaxy DMs for a subset of well-localized FRBs are over-plotted as multi-colored diamonds (Cordes et al., 2022). All scattering timescales have been scaled to their nominal values at 1 GHz. FRB 20221219A (denoted by a red diamond) dramatically exceeds the expected scattering timescale from a Milky Way-like galaxy by 6.37σlog10τsubscript𝜎subscriptlog10𝜏\sigma_{\mathrm{log}_{10}\tau}italic_σ start_POSTSUBSCRIPT roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT for its DMhost (see §3.1.4).

One possible indicator of a dynamic circumburst medium is a high Faraday rotation measure (RM), defined as RM=0.81B(l)ne(l)dlRM0.81subscript𝐵𝑙subscript𝑛e𝑙differential-d𝑙\mathrm{RM}=0.81\int B_{\|}(l)n_{\mathrm{e}}(l)\mathrm{d}lroman_RM = 0.81 ∫ italic_B start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ( italic_l ) italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ( italic_l ) roman_d italic_l, where Bsubscript𝐵B_{\|}italic_B start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT is the magnetic field component along the LoS (in μG𝜇G\mu\mathrm{G}italic_μ roman_G), and nesubscript𝑛en_{\mathrm{e}}italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT is the electron number density. The low S/N of FRB 20221219A precludes any robust constraints on the RM by RM synthesis or QU𝑄𝑈QUitalic_Q italic_U-fitting techniques. We were, however, able to estimate a linear polarization fraction (L/I𝐿𝐼L/Iitalic_L / italic_I) in the non-derotated spectrum, which we take to be the lower limit. With this, we set an upper limit on its absolute |RM|RM|\mathrm{RM}|| roman_RM | via a Monte Carlo-based approach to estimate the degree to which a 100% linearly polarized burst could be Faraday depolarized.

We calculate the likelihood of a “true” (intrinsic) L/I𝐿𝐼L/Iitalic_L / italic_I exceeding the measured polarization fraction of FRB 20221219A (L/I0.502similar-to𝐿𝐼0.502L/I\sim 0.502italic_L / italic_I ∼ 0.502) by simulating a linearly polarized signal completely in Stokes Q𝑄Qitalic_Q, a priori. Using off-pulse noise statistics, we proceed to simulate Q𝑄Qitalic_Q and U𝑈Uitalic_U for a range of hypothetical RM values |RM|isubscriptRM𝑖|\mathrm{RM}|_{i}| roman_RM | start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT at the observing wavelength λ𝜆\lambdaitalic_λ as

Q𝑄\displaystyle Qitalic_Q =cos(|RM|i×λ2)absentsubscriptRM𝑖superscript𝜆2\displaystyle=\cos\left(|\mathrm{RM}|_{i}\times\lambda^{2}\right)= roman_cos ( | roman_RM | start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT × italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (14)
U𝑈\displaystyle Uitalic_U =sin(|RM|i×λ2),absentsubscriptRM𝑖superscript𝜆2\displaystyle=-\sin\left(|\mathrm{RM}|_{i}\times\lambda^{2}\right),= - roman_sin ( | roman_RM | start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT × italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ,

which we then sum in quadrature to obtain a linear polarization fraction L/I=Q2+U2𝐿𝐼superscriptdelimited-⟨⟩𝑄2superscriptdelimited-⟨⟩𝑈2L/I=\sqrt{\langle Q\rangle^{2}+\langle U\rangle^{2}}italic_L / italic_I = square-root start_ARG ⟨ italic_Q ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⟨ italic_U ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, where Qdelimited-⟨⟩𝑄\langle Q\rangle⟨ italic_Q ⟩ and Udelimited-⟨⟩𝑈\langle U\rangle⟨ italic_U ⟩ represent the mean of the simulated values over the signal data. From this distribution of simulated polarization fractions exceeding the measured value, we extract a lower limit L/I0.42similar-to𝐿𝐼0.42L/I\sim 0.42italic_L / italic_I ∼ 0.42 (estimated at similar-to\sim 10%) for FRB 20221219A. Thus, we can set an upper limit on |RM|RM|\mathrm{RM}|| roman_RM | of 345less-than-or-similar-toabsent345\lesssim 345≲ 345 rad m-2, a relatively low value, and conclude that the circumburst medium for FRB 20221219A is not highly magnetic.

There is some evidence that a high RM is accompanied by strong local scattering (such as in FRB 20190520B; Ocker et al., 2022c; Anna-Thomas et al., 2023). Since RM is a tracer of both Bsubscript𝐵B_{\|}italic_B start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT and nesubscript𝑛en_{\mathrm{e}}italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT, it is possible for a low RM to imply a less extreme local plasma environment and thus a lack of scattering power, as further evidenced by the high RMs of Galactic pulsars in the inner disk of the Milky Way, which exhibit high degrees of scattering (Lazio & Cordes, 1998; Wharton et al., 2012; Eatough et al., 2013; Cordes et al., 2022). We note, however, that there is not necessarily correspondence between the Bsubscript𝐵B_{\|}italic_B start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT traced by RM and the possibly turbulent electron density fluctuations that lead to scattering. FRB 20121102A, for instance, exhibits RM values 105greater-than-or-equivalent-toabsentsuperscript105\gtrsim 10^{5}≳ 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT rad m-2, indicating the presence of an extreme magneto-ionic local environment, but exhibits no significant scattering in its host or circumburst environment (Michilli et al., 2018).

To further assess the likelihood that scattering would originate from the circumburst environment, we can take a more empirical approach using our combined parameters outlined in §3.2.1 and §3.2.2 to compute the measurable parameters, namely G𝐺Gitalic_G, τ𝜏\tauitalic_τ, ν𝜈\nuitalic_ν, and zsubscript𝑧z_{\ell}italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT, and solve for F~×DM,1002~𝐹superscriptsubscriptDM1002\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to better constrain the electron column density and fluctuation statistics. If we consider the scenario in which all the observed scattering originates from the circumburst environment (where G=1𝐺1G=1italic_G = 1), we find that F~×DM,1002580pc4/3km1/3cm1/3similar-to-or-equals~𝐹superscriptsubscriptDM1002580superscriptpc43superscriptkm13superscriptcm13\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}\simeq 580~{}\mathrm{pc}^{4/3}% \mathrm{~{}km}^{-1/3}\mathrm{~{}cm}^{-1/3}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≃ 580 roman_pc start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT roman_km start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT. If we remain agnostic to the exact physics in the circumburst environment, and simply consider the lower limit on F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG imposed by DMhost (see Table 3), we find that F~1500greater-than-or-equivalent-to~𝐹1500\widetilde{F}\gtrsim 1500over~ start_ARG italic_F end_ARG ≳ 1500 (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT. While the exact range of F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG is poorly understood for local environments of FRB sources, we see that the lower limit far exceeds the maximal expected F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG for the thin disk ISM of a spiral galaxy by 103similar-toabsentsuperscript103\sim 10^{3}∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (104similar-toabsentsuperscript104\sim 10^{4}∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT for dwarf galaxies, 106similar-toabsentsuperscript106\sim 10^{6}∼ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT for elliptical galaxies; Ocker et al., 2022b). This assumption is purely illustrative, however, as it of course ignores the presence of an ISM in the host galaxy. Nonetheless, it highlights the requirement that the local environment be extremely turbulent in order for it to substantially contribute to τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT.

Next, we’ll evaluate the scattering contributions of each intervening galaxy, again utilizing the scattering formalism developed by Cordes et al. (2016), Ocker et al. (2021), and Cordes et al. (2022) to characterize the scattering power of intervening plasma based on the LoS geometry (§3.2.2), and electron density-fluctuation (i.e., turbulence) statistics (§3.2.1).

3.2.4 IGH Model A: Extended CGM

To evaluate the scattering power of a uniform, extended CGM, we assume DMigh and a viral radius, R200subscript𝑅200R_{200}italic_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT, based on the halo mass of each intervener. With these assumptions, we take a similar approach to §3.2.3 and solve for F~×DM,1002~𝐹superscriptsubscriptDM1002\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. First, we take the path length L𝐿Litalic_L through the CGM (scattering layer) to be the geometric chord length through a spherically symmetric dark matter halo of R200subscript𝑅200R_{200}italic_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT, corresponding to IGH1 or IGH2. With this, we calculate the geometrical factor G𝐺Gitalic_G using Eq. 11 for the angular diameter distances shown in Figure 4. Substituting the DM values inferred for IGH1 and IGH2 in §3.1.3 based on the COS-Halos survey data (Werk et al., 2014), we estimate the value of F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG that would be required to reach τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT using Eq. 12 (see Table 4). We find the fluctuation parameters for each respective intervener, F~[1100610+3300,410190+580]similar-to~𝐹subscriptsuperscript11003300610subscriptsuperscript410580190\widetilde{F}\sim\left[1100^{+3300}_{-610},410^{+580}_{-190}\right]over~ start_ARG italic_F end_ARG ∼ [ 1100 start_POSTSUPERSCRIPT + 3300 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 610 end_POSTSUBSCRIPT , 410 start_POSTSUPERSCRIPT + 580 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 190 end_POSTSUBSCRIPT ] (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT, that would be required to reach τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT far exceed the physically valid range expected for the CGM based on prior studies (F~103less-than-or-similar-to~𝐹superscript103\tilde{F}\lesssim 10^{-3}over~ start_ARG italic_F end_ARG ≲ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT; Ocker et al., 2021), requiring values more appropriate for the Galactic thin disk (F~1greater-than-or-equivalent-to~𝐹1\widetilde{F}\gtrsim 1over~ start_ARG italic_F end_ARG ≳ 1 (pc2km)1/3superscriptsuperscriptpc2km13\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT; Ocker et al., 2021). The specific range of predicted F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG is largely informed by the volume filling fraction of the gas, estimated to be fv104similar-tosubscript𝑓vsuperscript104f_{\mathrm{v}}\sim 10^{-4}italic_f start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT (Vedantham & Phinney, 2018), inferred from areal covering fractions measured using quasar (QSO) absorption spectroscopy and photoionization modeling (Prochaska & Hennawi, 2008; Stocke et al., 2013; Hennawi et al., 2015; Lau et al., 2016; McCourt et al., 2017). The volume filling fraction relies on the assumption that the CGM can be described as a two-phase medium: a warm T106greater-than-or-equivalent-to𝑇superscript106T\gtrsim 10^{6}italic_T ≳ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K medium of tenuous coronal gas that hosts within it a cooler T106less-than-or-similar-to𝑇superscript106T\lesssim 10^{6}italic_T ≲ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K medium, which becomes unstable around T104similar-to𝑇superscript104T\sim 10^{4}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K and fragments or shatters into discrete photoionized cloudlets that populate the tenuous gas with a high covering fraction, but low filling factor (McCourt et al., 2017). We conclude that scattering by an extended CGM is physically unreasonable.

This conclusion is, of course, highly sensitive to our estimate of G𝐺Gitalic_G, and therefore Lhalosubscript𝐿haloL_{\mathrm{halo}}italic_L start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT. Minimizing Lhalosubscript𝐿haloL_{\mathrm{halo}}italic_L start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT to by a factor of 102similar-toabsentsuperscript102\sim 10^{2}∼ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to of order 1kpcsimilar-toabsent1kpc\sim 1~{}\mathrm{kpc}∼ 1 roman_kpc, while kee** DMhalosubscriptDMhalo\mathrm{DM}_{\mathrm{halo}}roman_DM start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT fixed, while an unphysical assumption, would better accommodate the scenario in which τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT originates from an extended CGM. This illustrates that in order for scattering to occur in an IGH at all, it would need to originate from a highly confined scattering layer in the halo. We consider this possibility as follows in §3.2.5.

Table 4: For each IGH, we estimate halo masses (log10(Mh)subscript10subscript𝑀\log_{10}\left(M_{h}\right)roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT )), virial radii (R200subscript𝑅200R_{200}italic_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT), impact parameters (b𝑏bitalic_b), geometrical chord lengths through the spherical halos (Lhalosubscript𝐿haloL_{\mathrm{halo}}italic_L start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT). We then calculate geometrical factors (G𝐺Gitalic_G; Eq. 11) and solve for F~×DM,1002~𝐹superscriptsubscriptDM1002\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (with units pc4/3km1/3cm1/3superscriptpc43superscriptkm13superscriptcm13\mathrm{pc}^{4/3}\mathrm{~{}km}^{-1/3}\mathrm{~{}cm}^{-1/3}roman_pc start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT roman_km start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT, using Eq. 12). For IGH model A, we infer DMighsubscriptDMigh\mathrm{DM}_{\mathrm{igh}}roman_DM start_POSTSUBSCRIPT roman_igh end_POSTSUBSCRIPT using the COS-Halos survey data (described in §3.1.3; Werk et al., 2014), and consequent values for F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG. For IGH model B, we assume a cloudlet width (Lcloudsubscript𝐿cloudL_{\mathrm{cloud}}italic_L start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT, which we take to be the outer scale losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT), electron density fluctuation statistics (ζε2𝜁superscript𝜀2\zeta\varepsilon^{2}italic_ζ italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), the Fresnel scale (rFsubscript𝑟𝐹r_{F}italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, which we take to be 10×10\times10 × the inner scale lisubscript𝑙𝑖l_{i}italic_l start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT), and a volume-filling factor (fvsubscript𝑓𝑣f_{v}italic_f start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT), to estimate the fluctuation parameter (F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG), and infer a DMcloudsubscriptDMcloud\mathrm{DM}_{\mathrm{cloud}}roman_DM start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT required to achieve τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT. These assumptions are described and motivated in §3.2.5. All assumed or inferred parameters are indicated in bold font to distinguish them from data-driven measurements.
Parameter IGH1 IGH2
log10(Mh)[M]subscript10subscript𝑀delimited-[]subscript𝑀direct-product\log_{10}\left(M_{h}\right)\left[M_{\odot}\right]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) [ italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ] 11.520.05+0.05subscriptsuperscript11.520.050.0511.52^{+0.05}_{-0.05}11.52 start_POSTSUPERSCRIPT + 0.05 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.05 end_POSTSUBSCRIPT 11.960.01+0.01subscriptsuperscript11.960.010.0111.96^{+0.01}_{-0.01}11.96 start_POSTSUPERSCRIPT + 0.01 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.01 end_POSTSUBSCRIPT
R200subscriptR200\mathrm{R}_{200}roman_R start_POSTSUBSCRIPT 200 end_POSTSUBSCRIPT [kpc] 1225+5subscriptsuperscript12255122^{+5}_{-5}122 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT 1742+2subscriptsuperscript17422174^{+2}_{-2}174 start_POSTSUPERSCRIPT + 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT
b𝑏bitalic_b [kpc] 4311+11subscriptsuperscript43111143^{+11}_{-11}43 start_POSTSUPERSCRIPT + 11 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 11 end_POSTSUBSCRIPT 3611+11subscriptsuperscript36111136^{+11}_{-11}36 start_POSTSUPERSCRIPT + 11 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 11 end_POSTSUBSCRIPT
                IGH  Model  A:  Extended  CGM
Lhalosubscript𝐿haloL_{\mathrm{halo}}italic_L start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT [kpc] 22514+14subscriptsuperscript2251414225^{+14}_{-14}225 start_POSTSUPERSCRIPT + 14 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 14 end_POSTSUBSCRIPT 3405+5subscriptsuperscript34055340^{+5}_{-5}340 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT
G𝐺Gitalic_G 110070+70subscriptsuperscript110070701100^{+70}_{-70}1100 start_POSTSUPERSCRIPT + 70 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 70 end_POSTSUBSCRIPT 130019+19subscriptsuperscript130019191300^{+19}_{-19}1300 start_POSTSUPERSCRIPT + 19 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 19 end_POSTSUBSCRIPT
F~×DM,1002~𝐹superscriptsubscriptDM1002\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 112+2subscriptsuperscript112211^{+2}_{-2}11 start_POSTSUPERSCRIPT + 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 81+1subscriptsuperscript8118^{+1}_{-1}8 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT
DMigh[pccm3]subscriptDMighdelimited-[]pcsuperscriptcm3\mathrm{\textbf{DM}}_{\mathrm{\textbf{igh}}}\left[\mathrm{pc~{}cm}^{-3}\right]DM start_POSTSUBSCRIPT igh end_POSTSUBSCRIPT [ roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ] 105+5subscriptsuperscript105510^{+5}_{-5}10 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT 145+5subscriptsuperscript145514^{+5}_{-5}14 start_POSTSUPERSCRIPT + 5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT
𝑭~[(pc2km)1/3]~𝑭delimited-[]superscriptsuperscriptpc2km13\widetilde{\bm{F}}\left[\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}\right]over~ start_ARG bold_italic_F end_ARG [ ( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT ] 1100610+3300subscriptsuperscript110033006101100^{+3300}_{-610}1100 start_POSTSUPERSCRIPT + 3300 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 610 end_POSTSUBSCRIPT 410190+580subscriptsuperscript410580190410^{+580}_{-190}410 start_POSTSUPERSCRIPT + 580 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 190 end_POSTSUBSCRIPT
     IGH  Model  B:  Partially  Ionized  CGM  Cloudlet
𝑳𝐜𝐥𝐨𝐮𝐝subscript𝑳𝐜𝐥𝐨𝐮𝐝\bm{L_{\mathrm{cloud}}}bold_italic_L start_POSTSUBSCRIPT bold_cloud end_POSTSUBSCRIPT [losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT; pc] 10 10
𝑮𝑮\bm{G}bold_italic_G 2.4×1072.4superscript1072.4\times 10^{7}2.4 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT 4.4×1074.4superscript1074.4\times 10^{7}4.4 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT
𝑭~×𝐃𝐌,𝟏𝟎𝟎𝟐~𝑭superscriptsubscript𝐃𝐌bold-ℓ1002\widetilde{\bm{F}}\times\bm{\mathrm{DM}_{\ell,100}^{2}}over~ start_ARG bold_italic_F end_ARG × bold_DM start_POSTSUBSCRIPT bold_ℓ bold_, bold_100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_2 end_POSTSUPERSCRIPT 50.7+0.7×104subscriptsuperscript50.70.7superscript1045^{+0.7}_{-0.7}\times 10^{-4}5 start_POSTSUPERSCRIPT + 0.7 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.7 end_POSTSUBSCRIPT × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 20.3+0.3×104subscriptsuperscript20.30.3superscript1042^{+0.3}_{-0.3}\times 10^{-4}2 start_POSTSUPERSCRIPT + 0.3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.3 end_POSTSUBSCRIPT × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
𝒓𝐅subscript𝒓𝐅\bm{r_{\mathrm{F}}}bold_italic_r start_POSTSUBSCRIPT bold_F end_POSTSUBSCRIPT [10×li10subscript𝑙𝑖10\times l_{i}10 × italic_l start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT; km] 3.6×1083.6superscript1083.6\times 10^{8}3.6 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 4.8×1084.8superscript1084.8\times 10^{8}4.8 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT
𝜻𝜺𝟐𝜻superscript𝜺2\bm{\zeta\varepsilon^{2}}bold_italic_ζ bold_italic_ε start_POSTSUPERSCRIPT bold_2 end_POSTSUPERSCRIPT 1 1
𝒇𝐯subscript𝒇𝐯\bm{f_{\mathrm{v}}}bold_italic_f start_POSTSUBSCRIPT bold_v end_POSTSUBSCRIPT 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
𝑭~[(pc2km)1/3]~𝑭delimited-[]superscriptsuperscriptpc2km13\widetilde{\bm{F}}\left[\left(\mathrm{pc}^{2}\mathrm{~{}km}\right)^{-1/3}\right]over~ start_ARG bold_italic_F end_ARG [ ( roman_pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_km ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT ] 0.070.070.070.07 0.060.060.060.06
𝐃𝐌𝐜𝐥𝐨𝐮𝐝[pccm3]subscript𝐃𝐌𝐜𝐥𝐨𝐮𝐝delimited-[]pcsuperscriptcm3\bm{\mathrm{DM}_{\mathrm{cloud}}}\left[\mathrm{pc~{}cm}^{-3}\right]bold_DM start_POSTSUBSCRIPT bold_cloud end_POSTSUBSCRIPT [ roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ] 8.50.5+0.5subscriptsuperscript8.50.50.58.5^{+0.5}_{-0.5}8.5 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT 5.80.5+0.5subscriptsuperscript5.80.50.55.8^{+0.5}_{-0.5}5.8 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT

3.2.5 IGH Model B: Partially Ionized CGM Cloudlet

The likelihood of intersecting a partially ionized cloudlet can be understood based on the areal covering fraction of the cool (T104similar-to𝑇superscript104T\sim 10^{4}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K) gas. Multiple studies investigating the prevalence of absorption features in QSO spectra and lensing of those features (Churchill et al., 2003; Xavier Prochaska et al., 2012; Lau et al., 2016), as well as fluorescent Ly-α𝛼\alphaitalic_α emission in halos (Hennawi et al., 2015), support the conclusion that the areal covering fraction exceeds unity. Hence, the scenario in which a single sightline intersects a cool cloudlet is highly probable.

To evaluate the scattering power of a single, partially ionized cloudlet, we again estimate F~×DM,1002~𝐹superscriptsubscriptDM1002\widetilde{F}\times\mathrm{DM}_{\ell,100}^{2}over~ start_ARG italic_F end_ARG × roman_DM start_POSTSUBSCRIPT roman_ℓ , 100 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (Eq. 9) based on a fiducial width of the cloudlet residing in an intervening CGM. The width is taken to be a typical upper limit on the size of a stable over-density of ionized gas in a cool (T104similar-to𝑇superscript104T\sim 10^{4}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K), fragmented CGM, Lcloud10similar-tosubscript𝐿cloud10L_{\mathrm{cloud}}\sim 10italic_L start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 10 pc (McCourt et al., 2017). We then estimate F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG, based on fiducial values for the metric that describes local and ensemble electron density variations ζε2𝜁superscript𝜀2\zeta\varepsilon^{2}italic_ζ italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the volume-filling factor fvsubscript𝑓vf_{\mathrm{v}}italic_f start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT, as well as the inner (lisubscript𝑙𝑖l_{i}italic_l start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT) and outer (losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT) scales of the turbulence power spectrum within the cloudlet. Under the well-justified assumption of homogeneous and isotropic turbulence, we take ζε21similar-to𝜁superscript𝜀21\zeta\varepsilon^{2}\sim 1italic_ζ italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ 1 (Spangler & Gwinn, 1990; Armstrong et al., 1995; Rickett et al., 2009). We also assume fv102similar-tosubscript𝑓vsuperscript102f_{\mathrm{v}}\sim 10^{-2}italic_f start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (typical for T104Ksimilar-to𝑇superscript104KT\sim 10^{4}~{}\mathrm{K}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_K gas; Ocker et al., 2021). It ought to be noted, however, that this factor is largely uncertain for individual clouds due to the inability to resolve and characterize this gas at sub-pc scales.

We set the inner scale lisubscript𝑙𝑖l_{i}italic_l start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT to be a fraction444This is a somewhat arbitrary choice and can be adjusted. We set this inner scale as 0.1×rF0.1subscript𝑟F0.1\times r_{\mathrm{F}}0.1 × italic_r start_POSTSUBSCRIPT roman_F end_POSTSUBSCRIPT as it concretely satisfies the condition that, in order for scattering to occur, the inner scale of the medium must lie below rFsubscript𝑟Fr_{\mathrm{F}}italic_r start_POSTSUBSCRIPT roman_F end_POSTSUBSCRIPT. of the Fresnel scale rFsubscript𝑟Fr_{\mathrm{F}}italic_r start_POSTSUBSCRIPT roman_F end_POSTSUBSCRIPT, in this case li=0.1×rFsubscript𝑙𝑖0.1subscript𝑟Fl_{i}=0.1\times r_{\mathrm{F}}italic_l start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.1 × italic_r start_POSTSUBSCRIPT roman_F end_POSTSUBSCRIPT, at the distance of the intervener, defined as (Cordes et al., 2017)

rF=λdsldlo2πdsosubscript𝑟F𝜆subscript𝑑slsubscript𝑑lo2𝜋subscript𝑑sor_{\mathrm{F}}=\sqrt{\frac{\lambda d_{\mathrm{sl}}d_{\mathrm{lo}}}{2\pi d_{% \mathrm{so}}}}italic_r start_POSTSUBSCRIPT roman_F end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_λ italic_d start_POSTSUBSCRIPT roman_sl end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT roman_lo end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π italic_d start_POSTSUBSCRIPT roman_so end_POSTSUBSCRIPT end_ARG end_ARG (15)

and an outer scale losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT consistent with the width of the cloudlet (here Lcloud10pcsimilar-tosubscript𝐿cloud10pcL_{\mathrm{cloud}}\sim 10~{}\mathrm{pc}italic_L start_POSTSUBSCRIPT roman_cloud end_POSTSUBSCRIPT ∼ 10 roman_pc. With these assumptions, we estimate F~~𝐹\widetilde{F}over~ start_ARG italic_F end_ARG using Eq. 9, and the DM required to produce τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT in Table 4. Under the assumptions outlined in Table 4, we find that for IGH2, a DMcloud of only 5.80.5+0.5pccm3subscriptsuperscript5.80.50.5pcsuperscriptcm35.8^{+0.5}_{-0.5}~{}\mathrm{pc~{}cm}^{-3}5.8 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT is required to achieve τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT.

While there are a variety of circumgalactic media that could provide this electron column density, the most well-understood in the context of the Milky Way are high-velocity clouds (HVCs). The Wisconsin Hα𝛼\alphaitalic_α Mapper (WHAM; Tufte et al., 1998a, b) has observed HVCs at high galactic latitudes. As the extinction at high latitudes is presumed to be minimal, the Hα𝛼\alphaitalic_α luminosity can be related directly to emission measure (EM=ne2𝑑lEMsuperscriptsubscript𝑛e2differential-d𝑙\mathrm{EM}=\int n_{\mathrm{e}}^{2}dlroman_EM = ∫ italic_n start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_l), for which WHAM report an EM=0.18cm6pcEM0.18superscriptcm6pc\mathrm{EM}=0.18\mathrm{~{}cm}^{-6}\mathrm{pc}roman_EM = 0.18 roman_cm start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT roman_pc. Assuming an HII temperature of T8000similar-to𝑇8000T\sim 8000italic_T ∼ 8000 K and a small ionization fraction, NHII/NHI0.08similar-tosubscriptNHIIsubscriptNHI0.08\mathrm{N}_{\mathrm{HII}}/\mathrm{N}_{\mathrm{HI}}\sim 0.08roman_N start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT / roman_N start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ∼ 0.08 (as HVCs are known to be primarily neutral, ionized in part by external radiation), the inferred HII column densities along an individual sightline reach an upper limit of NHII2×1019{}_{\mathrm{HII}}\sim 2\times 10^{19}start_FLOATSUBSCRIPT roman_HII end_FLOATSUBSCRIPT ∼ 2 × 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT cm-2 (Tufte et al., 1998b). This column density corresponds to a dispersion measure of DMHVC6.5pccm3{}_{\mathrm{HVC}}\simeq 6.5~{}\mathrm{pc}~{}\mathrm{cm}^{-3}start_FLOATSUBSCRIPT roman_HVC end_FLOATSUBSCRIPT ≃ 6.5 roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, which exceeds the DMcloud5.80.5+0.5pccm3{}_{\mathrm{cloud}}\simeq 5.8^{+0.5}_{-0.5}~{}\mathrm{pc}~{}\mathrm{cm}^{-3}start_FLOATSUBSCRIPT roman_cloud end_FLOATSUBSCRIPT ≃ 5.8 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT roman_pc roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT sufficient to achieve τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT.

While most Galactic HVCs lie at extra-planar distances 10less-than-or-similar-toabsent10\lesssim 10≲ 10 kpc (Wakker et al., 2008; Thom, 2006; Thom et al., 2008), there are notable exceptions, such as those present in the Magellanic Stream (a stream of HVCs traversing 100greater-than-or-equivalent-toabsentsuperscript100\gtrsim 100^{\circ}≳ 100 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT over the Galactic South Pole), which can extend out to 55-150similar-toabsent55-150\sim 55~{}\text{-}~{}150∼ 55 - 150 kpc (Besla et al., 2010). Streams of this kind are not necessarily unique to the Milky Way, as tidal interactions between more massive galaxies and their neighboring dwarfs are not uncommon, particularly at higher z𝑧zitalic_z. Hence, while the presence of HVCs extending to extraplanar distances of order b𝑏bitalic_b for IGH1 or IGH2 is challenging to constrain and largely uncertain, it is certainly possible. Galactic and Local Group HI surveys have also placed a subpopulation of so-called “compact” HVCs out to distances of 40-80similar-toabsent40-80\sim 40~{}\text{-}~{}80∼ 40 - 80 kpc from the Galactic disk, and 50similar-toabsent50\sim 50∼ 50 kpc around Andromeda (Putman et al., 2003; Pisano et al., 2007; Westmeier, 2007). Similar, even smaller “ultra-compact” HVCs have been observed within the Local Group, though their associations with individual galaxies remain uncertain (Adams et al., 2013). Still, they remain interesting candidate sources of FRB scattering.

Absorption line spectroscopy of QSOs has illuminated the intricate structures and dynamic processes characterizing the CGMs of galaxies beyond the Milky Way as well. High-velocity cloud-like structures, akin to those observed in the Milky Way’s halo, have been detected in the CGMs of other galaxies, showcasing a wide range of ionization states and physical scales. QSOs intersecting the CGMs of intervening galaxies reveal the presence of ionized cloudlets through the detection of species ranging from low ions like Si II and Mg II to highly ionized ions such as O VI and Ne VIII (Fox et al., 2014; Prochaska et al., 2017). These findings suggest that the CGM is a complex, multiphase medium where different regions exhibit varying degrees of ionization (Werk et al., 2013; Tumlinson et al., 2017). The column density distribution and the spatial coherence of QSO absorption features suggest that these ionized clouds range from tens to hundreds of parsecs across, similar to HVCs in the Milky Way (Stocke et al., 2013; Prochaska et al., 2017).

The distributions of these ionized structures within the galactic halos are substantial, extending from several tens to hundreds of kiloparsecs. This vast reach underscores the CGM’s role as a reservoir of baryonic material that can influence galaxy evolution (Werk et al., 2014; Burchett et al., 2016). Ionization models employed to interpret the absorption spectra indicate that some regions within these clouds are nearly fully ionized, revealing a complex CGM phase structure (Oppenheimer & Schaye, 2013; Stocke et al., 2013). They also indicate that these clouds occupy diverse thermal phases of the CGM, from cool (T104similar-to𝑇superscript104T\sim 10^{4}italic_T ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K), photoionized gas to hot (T106greater-than-or-equivalent-to𝑇superscript106T\gtrsim 10^{6}italic_T ≳ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K), collisionally ionized plasma (Werk et al., 2016; Tumlinson et al., 2017).

The evidence garnered from spectroscopic studies not only confirms the presence of multiphase, dynamic HVC-like structures but also reveals their substantial physical scales and diverse ionization fractions, making it all the more reasonable that an intersection with a single, marginally ionized cloudlet could occur for both IGH1 and IGH2, given their respective impact parameters, and therefore dominate the scattering budget of FRB 20221219A.

4 Conclusions

By constructing detailed DM and scattering budgets for FRB 20221219A, we find that the event is highly over-scattered in its host galaxy and lies along an unusually rich sightline. We also find that the IGM and ICM likely dominate the DM budget, leaving a modest allowance for DMhost.

We consider the possibility of scattering occurring in the CGMs of two closely intervening galaxies, and find it plausible. Specifically, we note that a fortuitous intersection with a single, partially ionized cloudlet in one of the CGMs could produce τobssubscript𝜏obs\tau_{\mathrm{obs}}italic_τ start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT. The geometric leverage offered by compact scattering structures at large distances from both the source and observer, as well as the potential presence of partially ionized structures at the impact parameters of both IGH1 and IGH2 along the LoS, make this scenario likely. We further find that nothing about the host galaxy or local environment is unusual with respect to other FRBs (and pulsars) with far lower scattering, lending credence to this suggestion.

The fact that all FRBs are not scattered by the CGMs of intervening galaxies can be explained by the far sparser distributions of such galaxies along most sightlines, as well as the known bias against observing highly scattered FRBs. We further note that scattering is highly stochastic for fixed DMs, as evidenced by MW pulsars, inflating this bias even more.

In future case studies, it would be of great use, where possible, to measure scintillation bandwidths in addition to pulse broadening to better understand FRB scattering mechanisms and disentangle distinct LoS scattering media. Comparisons between scintillation and pulse broadening of a single burst can assist in constraining whether the scattering effects in FRBs are due to one or multiple screens along the line of sight. Still, it is crucial to search for repeat bursts from heavily scattered FRBs to characterize additional time-variable effects.

5 Acknowledgements

The authors would like to thank Jim Cordes for insightful conversations, as well as the staff members of the Owens Valley Radio Observatory and the Caltech radio group whose efforts were vital to the success of the DSA-110, including Kristen Bernasconi, Stephanie Cha-Ramos, Sarah Harnach, Tom Klinefelter, Lori McGraw, Corey Posner, Andres Rizo, Michael Virgin, Scott White, and Thomas Zentmyer. The DSA-110 is supported by the National Science Foundation Mid-Scale Innovations Program in Astronomical Sciences (MSIP) under grant AST-1836018.

This research is based in part on data gathered with the W.M. Keck Observatory, which is operated, in scientific partnership, by the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck foundation.

The Legacy Surveys consist of three individual and complementary projects: the Dark Energy Camera Legacy Survey (DECaLS; Proposal ID #2014B-0404; PIs: David Schlegel and Arjun Dey), the Bei**g-Arizona Sky Survey (BASS; NOAO Prop. ID #2015A-0801; PIs: Zhou Xu and Xiaohui Fan), and the Mayall z-band Legacy Survey (MzLS; Prop. ID #2016A-0453; PI: Arjun Dey). DECaLS, BASS and MzLS together include data obtained, respectively, at the Blanco telescope, Cerro Tololo Inter-American Observatory, NSF’s NOIRLab; the Bok telescope, Steward Observatory, University of Arizona; and the Mayall telescope, Kitt Peak National Observatory, NOIRLab. Pipeline processing and analyses of the data were supported by NOIRLab and the Lawrence Berkeley National Laboratory (LBNL). The Legacy Surveys project is honored to be permitted to conduct astronomical research on Iolkam Du’ag (Kitt Peak), a mountain with particular significance to the Tohono O’odham Nation.

NOIRLab is operated by the Association of Universities for Research in Astronomy (AURA) under a cooperative agreement with the National Science Foundation. LBNL is managed by the Regents of the University of California under contract to the U.S. Department of Energy.

This project used data obtained with the Dark Energy Camera (DECam), which was constructed by the Dark Energy Survey (DES) collaboration. Funding for the DES Projects has been provided by the U.S. Department of Energy, the U.S. National Science Foundation, the Ministry of Science and Education of Spain, the Science and Technology Facilities Council of the United Kingdom, the Higher Education Funding Council for England, the National Center for Supercomputing Applications at the University of Illinois at Urbana-Champaign, the Kavli Institute of Cosmological Physics at the University of Chicago, Center for Cosmology and Astro-Particle Physics at the Ohio State University, the Mitchell Institute for Fundamental Physics and Astronomy at Texas A&M University, Financiadora de Estudos e Projetos, Fundacao Carlos Chagas Filho de Amparo, Financiadora de Estudos e Projetos, Fundacao Carlos Chagas Filho de Amparo a Pesquisa do Estado do Rio de Janeiro, Conselho Nacional de Desenvolvimento Cientifico e Tecnologico and the Ministerio da Ciencia, Tecnologia e Inovacao, the Deutsche Forschungsgemeinschaft and the Collaborating Institutions in the Dark Energy Survey. The Collaborating Institutions are Argonne National Laboratory, the University of California at Santa Cruz, the University of Cambridge, Centro de Investigaciones Energeticas, Medioambientales y Tecnologicas-Madrid, the University of Chicago, University College London, the DES-Brazil Consortium, the University of Edinburgh, the Eidgenossische Technische Hochschule (ETH) Zurich, Fermi National Accelerator Laboratory, the University of Illinois at Urbana-Champaign, the Institut de Ciencies de l’Espai (IEEC/CSIC), the Institut de Fisica d’Altes Energies, Lawrence Berkeley National Laboratory, the Ludwig Maximilians Universitat Munchen and the associated Excellence Cluster Universe, the University of Michigan, NSF’s NOIRLab, the University of Nottingham, the Ohio State University, the University of Pennsylvania, the University of Portsmouth, SLAC National Accelerator Laboratory, Stanford University, the University of Sussex, and Texas A&M University.

BASS is a key project of the Telescope Access Program (TAP), which has been funded by the National Astronomical Observatories of China, the Chinese Academy of Sciences (the Strategic Priority Research Program “The Emergence of Cosmological Structures” Grant #XDB09000000), and the Special Fund for Astronomy from the Ministry of Finance. The BASS is also supported by the External Cooperation Program of Chinese Academy of Sciences (Grant #114A11KYSB20160057), and Chinese National Natural Science Foundation (Grant #12120101003, #11433005).

The Legacy Survey team makes use of data products from the Near-Earth Object Wide-field Infrared Survey Explorer (NEOWISE), which is a project of the Jet Propulsion Laboratory/California Institute of Technology. NEOWISE is funded by the National Aeronautics and Space Administration.

The Legacy Surveys imaging of the DESI footprint is supported by the Director, Office of Science, Office of High Energy Physics of the U.S. Department of Energy under Contract No. DE-AC02-05CH1123, by the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility under the same contract; and by the U.S. National Science Foundation, Division of Astronomical Sciences under Contract No. AST-0950945 to NOAO.

The Pan-STARRS1 Surveys (PS1) and the PS1 public science archive have been made possible through contributions by the Institute for Astronomy, the University of Hawaii, the Pan-STARRS Project Office, the Max-Planck Society and its participating institutes, the Max Planck Institute for Astronomy, Heidelberg and the Max Planck Institute for Extraterrestrial Physics, Garching, The Johns Hopkins University, Durham University, the University of Edinburgh, the Queen’s University Belfast, the Harvard-Smithsonian Center for Astrophysics, the Las Cumbres Observatory Global Telescope Network Incorporated, the National Central University of Taiwan, the Space Telescope Science Institute, the National Aeronautics and Space Administration under Grant No. NNX08AR22G issued through the Planetary Science Division of the NASA Science Mission Directorate, the National Science Foundation Grant No. AST-1238877, the University of Maryland, Eotvos Lorand University (ELTE), the Los Alamos National Laboratory, and the Gordon and Betty Moore Foundation.

References

  • Adams et al. (2013) Adams, E. A. K., Giovanelli, R., & Haynes, M. P. 2013, ApJ, 768, 77, doi: 10.1088/0004-637X/768/1/77
  • Aggarwal et al. (2021) Aggarwal, K., Budavári, T., Deller, A. T., et al. 2021, The Astrophysical Journal, 911, 95, doi: 10.3847/1538-4357/abe8d2
  • Anna-Thomas et al. (2023) Anna-Thomas, R., Connor, L., Dai, S., et al. 2023, Science, 380, 599–603, doi: 10.1126/science.abo6526
  • Armstrong et al. (1995) Armstrong, J. W., Rickett, B. J., & Spangler, S. R. 1995, The Astrophysical Journal, 443, 209, doi: 10.1086/175515
  • Besla et al. (2010) Besla, G., Kallivayalil, N., Hernquist, L., et al. 2010, The Astrophysical Journal, 721, L97–L101, doi: 10.1088/2041-8205/721/2/l97
  • Bhandari et al. (2020) Bhandari, S., Sadler, E. M., Prochaska, J. X., et al. 2020, The Astrophysical Journal, 895, L37, doi: 10.3847/2041-8213/ab672e
  • Bhat et al. (2004) Bhat, N. D. R., Cordes, J. M., Camilo, F., Nice, D. J., & Lorimer, D. R. 2004, The Astrophysical Journal, 605, 759, doi: 10.1086/382680
  • Bietenholz & Kronberg (1991) Bietenholz, M. F., & Kronberg, P. P. 1991, The Astrophysical Journal, 368, 231, doi: 10.1086/169686
  • Bochenek et al. (2020) Bochenek, C. D., Ravi, V., Belov, K. V., et al. 2020, Nature, 587, 59, doi: 10.1038/s41586-020-2872-x
  • Burchett et al. (2016) Burchett, J. N., Tripp, T. M., Bordoloi, R., et al. 2016, The Astrophysical Journal, 832, 124, doi: 10.3847/0004-637x/832/2/124
  • Cappellari (2017) Cappellari, M. 2017, MNRAS, 466, 798, doi: 10.1093/mnras/stw3020
  • Cappellari (2022) —. 2022, arXiv e-prints, arXiv:2208.14974. https://arxiv.longhoe.net/abs/2208.14974
  • Chambers et al. (2016) Chambers, K. C., Magnier, E. A., Metcalfe, N., et al. 2016, arXiv e-prints, arXiv:1612.05560. https://arxiv.longhoe.net/abs/1612.05560
  • Chawla et al. (2022) Chawla, P., Kaspi, V. M., Ransom, S. M., et al. 2022, The Astrophysical Journal, 927, 35, doi: 10.3847/1538-4357/ac49e1
  • CHIME/FRB Collaboration et al. (2019) CHIME/FRB Collaboration, Andersen, B. C., Bandura, K., et al. 2019, ApJ, 885, L24, doi: 10.3847/2041-8213/ab4a80
  • CHIME/FRB Collaboration et al. (2020) CHIME/FRB Collaboration, Andersen, B. C., Bandura, K. M., et al. 2020, Nature, 587, 54, doi: 10.1038/s41586-020-2863-y
  • CHIME/FRB Collaboration et al. (2021) CHIME/FRB Collaboration, Amiri, M., Andersen, B. C., et al. 2021, The Astrophysical Journal Supplement Series, 257, 59, doi: 10.3847/1538-4365/ac33ab
  • CHIME/FRB Collaboration et al. (2022) CHIME/FRB Collaboration, Andersen, B. C., Bandura, K., et al. 2022, Nature, 607, 256, doi: 10.1038/s41586-022-04841-8
  • Chittidi et al. (2021) Chittidi, J. S., Simha, S., Mannings, A., et al. 2021, The Astrophysical Journal, 922, 173, doi: 10.3847/1538-4357/ac2818
  • Churchill et al. (2003) Churchill, C. W., Mellon, R. R., Charlton, J. C., & Vogt, S. S. 2003, The Astrophysical Journal, 593, 203–214, doi: 10.1086/376444
  • Connor & Ravi (2022) Connor, L., & Ravi, V. 2022, Nature Astronomy, 6, 1035, doi: 10.1038/s41550-022-01719-7
  • Connor et al. (2020) Connor, L., van Leeuwen, J., Oostrum, L. C., et al. 2020, MNRAS, 499, 4716, doi: 10.1093/mnras/staa3009
  • Connor et al. (2023) Connor, L., Ravi, V., Catha, M., et al. 2023, The Astrophysical Journal Letters, 949, L26, doi: 10.3847/2041-8213/acd3ea
  • Cook et al. (2023) Cook, A. M., Bhardwaj, M., Gaensler, B. M., et al. 2023, The Astrophysical Journal, 946, 58, doi: 10.3847/1538-4357/acbbd0
  • Cordes & Chatterjee (2019) Cordes, J. M., & Chatterjee, S. 2019, Annual Review of Astronomy and Astrophysics, 57, 417, doi: 10.1146/annurev-astro-091918-104501
  • Cordes & Lazio (2002) Cordes, J. M., & Lazio, T. J. W. 2002, arXiv e-prints, astro. https://arxiv.longhoe.net/abs/astro-ph/0207156
  • Cordes et al. (2022) Cordes, J. M., Ocker, S. K., & Chatterjee, S. 2022, The Astrophysical Journal, 931, 88, doi: 10.3847/1538-4357/ac6873
  • Cordes et al. (2017) Cordes, J. M., Wasserman, I., Hessels, J. W. T., et al. 2017, The Astrophysical Journal, 842, 35, doi: 10.3847/1538-4357/aa74da
  • Cordes et al. (1991) Cordes, J. M., Weisberg, J. M., Frail, D. A., Spangler, S. R., & Ryan, M. 1991, Nature, 354, 121, doi: 10.1038/354121a0
  • Cordes et al. (2016) Cordes, J. M., Wharton, R. S., Spitler, L. G., Chatterjee, S., & Wasserman, I. 2016, Radio Wave Propagation and the Provenance of Fast Radio Bursts, arXiv, doi: 10.48550/ARXIV.1605.05890
  • Deneva et al. (2009) Deneva, J. S., Cordes, J. M., & Lazio, T. J. W. 2009, The Astrophysical Journal, 702, L177–L181, doi: 10.1088/0004-637x/702/2/l177
  • Dexter et al. (2017) Dexter, J., Deller, A., Bower, G. C., et al. 2017, Monthly Notices of the Royal Astronomical Society, 471, 3563–3576, doi: 10.1093/mnras/stx1777
  • Dey et al. (2019) Dey, A., Schlegel, D. J., Lang, D., et al. 2019, The Astronomical Journal, 157, 168, doi: 10.3847/1538-3881/ab089d
  • Eatough et al. (2013) Eatough, R. P., Falcke, H., Karuppusamy, R., et al. 2013, Nature, 501, 391–394, doi: 10.1038/nature12499
  • Faber et al. (2003) Faber, S. M., Phillips, A. C., Kibrick, R. I., et al. 2003, in SPIE Proceedings, ed. M. Iye & A. F. M. Moorwood (SPIE), doi: 10.1117/12.460346
  • Fox et al. (2014) Fox, A. J., Wakker, B. P., Barger, K. A., et al. 2014, The Astrophysical Journal, 787, 147, doi: 10.1088/0004-637x/787/2/147
  • Gajjar et al. (2018) Gajjar, V., Siemion, A. P. V., Price, D. C., et al. 2018, The Astrophysical Journal, 863, 2, doi: 10.3847/1538-4357/aad005
  • Girelli et al. (2020) Girelli, G., Pozzetti, L., Bolzonella, M., et al. 2020, Astronomy & Astrophysics, 634, A135, doi: 10.1051/0004-6361/201936329
  • Gordon et al. (2023) Gordon, A. C., fai Fong, W., Kilpatrick, C. D., et al. 2023, The Astrophysical Journal, 954, 80, doi: 10.3847/1538-4357/ace5aa
  • Harris et al. (2020) Harris, C. R., Millman, K. J., van der Walt, S. J., et al. 2020, Nature, 585, 357–362, doi: 10.1038/s41586-020-2649-2
  • Hennawi et al. (2015) Hennawi, J. F., Prochaska, J. X., Cantalupo, S., & Arrigoni-Battaia, F. 2015, Science, 348, 779–783, doi: 10.1126/science.aaa5397
  • Hessels et al. (2019) Hessels, J. W. T., Spitler, L. G., Seymour, A. D., et al. 2019, The Astrophysical Journal, 876, L23, doi: 10.3847/2041-8213/ab13ae
  • Hunter (2007) Hunter, J. D. 2007, Computing in Science & Engineering, 9, 90, doi: 10.1109/MCSE.2007.55
  • Johnson et al. (2021) Johnson, B. D., Leja, J., Conroy, C., & Speagle, J. S. 2021, The Astrophysical Journal Supplement Series, 254, 22, doi: 10.3847/1538-4365/abef67
  • Josephy et al. (2019) Josephy, A., Chawla, P., Fonseca, E., et al. 2019, The Astrophysical Journal Letters, 882, L18, doi: 10.3847/2041-8213/ab2c00
  • Lau et al. (2016) Lau, M. W., Prochaska, J. X., & Hennawi, J. F. 2016, The Astrophysical Journal Supplement Series, 226, 25, doi: 10.3847/0067-0049/226/2/25
  • Law et al. (2023) Law, C. J., Sharma, K., Ravi, V., et al. 2023, Deep Synoptic Array Science: First FRB and Host Galaxy Catalog, arXiv, doi: 10.48550/ARXIV.2307.03344
  • Lazio & Cordes (1998) Lazio, T. J. W., & Cordes, J. M. 1998, The Astrophysical Journal, 505, 715–731, doi: 10.1086/306174
  • Licquia & Newman (2015) Licquia, T. C., & Newman, J. A. 2015, The Astrophysical Journal, 806, 96, doi: 10.1088/0004-637x/806/1/96
  • Löhmer et al. (2001) Löhmer, O., Kramer, M., Mitra, D., Lorimer, D. R., & Lyne, A. G. 2001, ApJ, 562, L157, doi: 10.1086/338324
  • Lorimer et al. (2007) Lorimer, D. R., Bailes, M., McLaughlin, M. A., Narkevic, D. J., & Crawford, F. 2007, Science, 318, 777, doi: 10.1126/science.1147532
  • Macquart & Koay (2013) Macquart, J.-P., & Koay, J. Y. 2013, The Astrophysical Journal, 776, 125, doi: 10.1088/0004-637x/776/2/125
  • Macquart et al. (2020) Macquart, J.-P., Prochaska, J. X., McQuinn, M., et al. 2020, Nature, 581, 391, doi: 10.1038/s41586-020-2300-2
  • Majid et al. (2021) Majid, W. A., Pearlman, A. B., Prince, T. A., et al. 2021, The Astrophysical Journal Letters, 919, L6, doi: 10.3847/2041-8213/ac1921
  • Marcote et al. (2020) Marcote, B., Nimmo, K., Hessels, J. W. T., et al. 2020, Nature, 577, 190, doi: 10.1038/s41586-019-1866-z
  • Masui et al. (2015) Masui, K., Lin, H.-H., Sievers, J., et al. 2015, Nature, 528, 523, doi: 10.1038/nature15769
  • McCourt et al. (2017) McCourt, M., Oh, S. P., O’Leary, R., & Madigan, A.-M. 2017, Monthly Notices of the Royal Astronomical Society, 473, 5407–5431, doi: 10.1093/mnras/stx2687
  • McQuinn (2014) McQuinn, M. 2014, ApJ, 780, L33, doi: 10.1088/2041-8205/780/2/L33
  • Michilli et al. (2018) Michilli, D., Seymour, A., Hessels, J. W. T., et al. 2018, Nature, 553, 182, doi: 10.1038/nature25149
  • Moster et al. (2010) Moster, B. P., Somerville, R. S., Maulbetsch, C., et al. 2010, The Astrophysical Journal, 710, 903–923, doi: 10.1088/0004-637x/710/2/903
  • Murray (2014) Murray, S. 2014, HMF: Halo Mass Function calculator, Astrophysics Source Code Library, record ascl:1412.006
  • Murray et al. (2013) Murray, S., Power, C., & Robotham, A. 2013, Astronomy and Computing, 3–4, 23–34, doi: 10.1016/j.ascom.2013.11.001
  • Nimmo et al. (2022) Nimmo, K., Hessels, J. W. T., Kirsten, F., et al. 2022, Nature Astronomy, 6, 393, doi: 10.1038/s41550-021-01569-9
  • Ocker et al. (2022a) Ocker, S., Cordes, J. M., Chatterjee, S., et al. 2022a, The Astrophysical Journal, 931, 87, doi: 10.3847/1538-4357/ac6504
  • Ocker et al. (2020) Ocker, S. K., Cordes, J. M., & Chatterjee, S. 2020, The Astrophysical Journal, 897, 124, doi: 10.3847/1538-4357/ab98f9
  • Ocker et al. (2021) —. 2021, The Astrophysical Journal, 911, 102, doi: 10.3847/1538-4357/abeb6e
  • Ocker et al. (2022b) Ocker, S. K., Cordes, J. M., Chatterjee, S., & Gorsuch, M. R. 2022b, The Astrophysical Journal, 934, 71, doi: 10.3847/1538-4357/ac75ba
  • Ocker et al. (2022c) Ocker, S. K., Cordes, J. M., Chatterjee, S., et al. 2022c, Monthly Notices of the Royal Astronomical Society, 519, 821–830, doi: 10.1093/mnras/stac3547
  • Oke et al. (1995) Oke, J. B., Cohen, J. G., Carr, M., et al. 1995, Publications of the Astronomical Society of the Pacific, 107, 375, doi: 10.1086/133562
  • Oppenheimer & Schaye (2013) Oppenheimer, B. D., & Schaye, J. 2013, Monthly Notices of the Royal Astronomical Society, 434, 1043–1062, doi: 10.1093/mnras/stt1043
  • Perley (2019) Perley, D. A. 2019, Publications of the Astronomical Society of the Pacific, 131, 084503, doi: 10.1088/1538-3873/ab215d
  • Petroff et al. (2019) Petroff, E., Hessels, J. W. T., & Lorimer, D. R. 2019, A&A Rev., 27, 4, doi: 10.1007/s00159-019-0116-6
  • Pillepich et al. (2017) Pillepich, A., Nelson, D., Hernquist, L., et al. 2017, Monthly Notices of the Royal Astronomical Society, 475, 648–675, doi: 10.1093/mnras/stx3112
  • Pisano et al. (2007) Pisano, D. J., Barnes, D. G., Gibson, B. K., et al. 2007, The Astrophysical Journal, 662, 959–968, doi: 10.1086/517986
  • Planck Collaboration et al. (2016) Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. 2016, A&A, 594, A13, doi: 10.1051/0004-6361/201525830
  • Planck Collaboration et al. (2020) Planck Collaboration, Aghanim, N., Akrami, Y., et al. 2020, Astronomy & Astrophysics, 641, A6, doi: 10.1051/0004-6361/201833910
  • Price-Whelan et al. (2018) Price-Whelan, A. M., Sipőcz, B. M., Günther, H. M., et al. 2018, The Astronomical Journal, 156, 123, doi: 10.3847/1538-3881/aabc4f
  • Prochaska & Hennawi (2008) Prochaska, J. X., & Hennawi, J. F. 2008, The Astrophysical Journal, 690, 1558, doi: 10.1088/0004-637x/690/2/1558
  • Prochaska et al. (2020) Prochaska, J. X., Hennawi, J. F., Westfall, K. B., et al. 2020, arXiv e-prints, arXiv:2005.06505. https://arxiv.longhoe.net/abs/2005.06505
  • Prochaska et al. (2017) Prochaska, J. X., Werk, J. K., Worseck, G., et al. 2017, The Astrophysical Journal, 837, 169, doi: 10.3847/1538-4357/aa6007
  • Prochaska et al. (2019) Prochaska, J. X., Macquart, J.-P., McQuinn, M., et al. 2019, Science, 366, 231, doi: 10.1126/science.aay0073
  • Putman et al. (2003) Putman, M. E., Bland‐Hawthorn, J., Veilleux, S., et al. 2003, The Astrophysical Journal, 597, 948–956, doi: 10.1086/378555
  • Ramachandran et al. (1997) Ramachandran, R., Mitra, D., Deshpande, A. A., McConnell, D. M., & Abies, J. G. 1997, Monthly Notices of the Royal Astronomical Society, 290, 260, doi: 10.1093/mnras/290.2.260
  • Ravi (2018) Ravi, V. 2018, Monthly Notices of the Royal Astronomical Society, 482, 1966–1978, doi: 10.1093/mnras/sty1551
  • Ravi et al. (2016) Ravi, V., Shannon, R. M., Bailes, M., et al. 2016, Science, 354, 1249, doi: 10.1126/science.aaf6807
  • Ravi et al. (2023a) Ravi, V., Catha, M., Chen, G., et al. 2023a, Deep Synoptic Array science: a 50 Mpc fast radio burst constrains the mass of the Milky Way circumgalactic medium, arXiv, doi: 10.48550/ARXIV.2301.01000
  • Ravi et al. (2023b) —. 2023b, The Astrophysical Journal Letters, 949, L3, doi: 10.3847/2041-8213/acc4b6
  • Rickett et al. (2009) Rickett, B., Johnston, S., Tomlinson, T., & Reynolds, J. 2009, Monthly Notices of the Royal Astronomical Society, 395, 1391, doi: 10.1111/j.1365-2966.2009.14471.x
  • Rickett (1977) Rickett, B. J. 1977, Annual Review of Astronomy and Astrophysics, 15, 479–504, doi: 10.1146/annurev.aa.15.090177.002403
  • Ryder et al. (2022) Ryder, S. D., Bannister, K. W., Bhandari, S., et al. 2022
  • Sammons et al. (2023) Sammons, M. W., Deller, A. T., Glowacki, M., et al. 2023, Monthly Notices of the Royal Astronomical Society, 525, 5653–5668, doi: 10.1093/mnras/stad2631
  • Sanchez-Blazquez et al. (2006) Sanchez-Blazquez, P., Peletier, R. F., Jimenez-Vicente, J., et al. 2006, Monthly Notices of the Royal Astronomical Society, 371, 703–718, doi: 10.1111/j.1365-2966.2006.10699.x
  • Schoen et al. (2021) Schoen, E., Leung, C., Masui, K., et al. 2021, Research Notes of the AAS, 5, 271, doi: 10.3847/2515-5172/ac3af9
  • Sharma et al. (2023) Sharma, K., Somalwar, J., Law, C., et al. 2023, The Astrophysical Journal, 950, 175, doi: 10.3847/1538-4357/accf1d
  • Simard & Ravi (2021) Simard, D., & Ravi, V. 2021, Measuring interstellar turbulence in fast radio burst host galaxies, arXiv, doi: 10.48550/ARXIV.2107.11334
  • Simha et al. (2020) Simha, S., Burchett, J. N., Prochaska, J. X., et al. 2020, The Astrophysical Journal, 901, 134, doi: 10.3847/1538-4357/abafc3
  • Spangler & Gwinn (1990) Spangler, S. R., & Gwinn, C. R. 1990, The Astrophysical Journal, 353, L29, doi: 10.1086/185700
  • Stocke et al. (2013) Stocke, J. T., Keeney, B. A., Danforth, C. W., et al. 2013, The Astrophysical Journal, 763, 148, doi: 10.1088/0004-637x/763/2/148
  • Sutton (1971) Sutton, J. M. 1971, Monthly Notices of the Royal Astronomical Society, 155, 51–64, doi: 10.1093/mnras/155.1.51
  • Thom (2006) Thom, C. 2006, PhD thesis, Swinburne University of Technology, Australia
  • Thom et al. (2008) Thom, C., Peek, J. E. G., Putman, M. E., et al. 2008, The Astrophysical Journal, 684, 364–372, doi: 10.1086/589960
  • Tinker et al. (2008) Tinker, J., Kravtsov, A. V., Klypin, A., et al. 2008, The Astrophysical Journal, 688, 709, doi: 10.1086/591439
  • Tufte et al. (1998a) Tufte, S. L., Reynolds, R. J., & Haffner, L. M. 1998a, The Astrophysical Journal, 504, 773, doi: 10.1086/306103
  • Tufte et al. (1998b) —. 1998b, H-alpha from High Velocity Clouds, arXiv, doi: 10.48550/ARXIV.ASTRO-PH/9811270
  • Tumlinson et al. (2017) Tumlinson, J., Peeples, M. S., & Werk, J. K. 2017, Annual Review of Astronomy and Astrophysics, 55, 389–432, doi: 10.1146/annurev-astro-091916-055240
  • Valdes et al. (2014) Valdes, F., Gruendl, R., & DES Project. 2014, in Astronomical Society of the Pacific Conference Series, Vol. 485, Astronomical Data Analysis Software and Systems XXIII, ed. N. Manset & P. Forshay, 379
  • Vedantham & Phinney (2018) Vedantham, H. K., & Phinney, E. S. 2018, Monthly Notices of the Royal Astronomical Society, 483, 971, doi: 10.1093/mnras/sty2948
  • Vikhlinin et al. (2006) Vikhlinin, A., Kravtsov, A., Forman, W., et al. 2006, The Astrophysical Journal, 640, 691, doi: 10.1086/500288
  • Virtanen et al. (2020) Virtanen, P., Gommers, R., Oliphant, T. E., et al. 2020, Nature Methods, 17, 261, doi: 10.1038/s41592-019-0686-2
  • Wakker et al. (2008) Wakker, B. P., York, D. G., Wilhelm, R., et al. 2008, The Astrophysical Journal, 672, 298–319, doi: 10.1086/523845
  • Walker et al. (2023) Walker, C. R. H., Spitler, L. G., Ma, Y.-Z., et al. 2023, The Dispersion Measure Contributions of the Cosmic Web, arXiv, doi: 10.48550/ARXIV.2309.08268
  • Wen et al. (2017) Wen, Z. L., Han, J. L., & Yang, F. 2017, Monthly Notices of the Royal Astronomical Society, 475, 343, doi: 10.1093/mnras/stx3189
  • Werk et al. (2013) Werk, J. K., Prochaska, J. X., Thom, C., et al. 2013, Astrophys. J. Suppl. Ser., 204, 17
  • Werk et al. (2014) Werk, J. K., Prochaska, J. X., Tumlinson, J., et al. 2014, The Astrophysical Journal, 792, 8, doi: 10.1088/0004-637x/792/1/8
  • Werk et al. (2016) Werk, J. K., Prochaska, J. X., Cantalupo, S., et al. 2016, The Astrophysical Journal, 833, 54, doi: 10.3847/1538-4357/833/1/54
  • Westmeier (2007) Westmeier, T. 2007, PhD thesis, Rheinische Friedrich Wilhelms University of Bonn, Germany
  • Wharton et al. (2012) Wharton, R. S., Chatterjee, S., Cordes, J. M., Deneva, J. S., & Lazio, T. J. W. 2012, The Astrophysical Journal, 753, 108, doi: 10.1088/0004-637x/753/2/108
  • Wilson et al. (2003) Wilson, J. C., Eikenberry, S. S., Henderson, C. P., et al. 2003, in Instrument Design and Performance for Optical/Infrared Ground-based Telescopes, ed. M. Iye & A. F. M. Moorwood (SPIE), doi: 10.1117/12.460336
  • Wright et al. (2010) Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, The Astronomical Journal, 140, 1868–1881, doi: 10.1088/0004-6256/140/6/1868
  • Xavier Prochaska et al. (2012) Xavier Prochaska, J., Hennawi, J. F., & Simcoe, R. A. 2012, The Astrophysical Journal, 762, L19, doi: 10.1088/2041-8205/762/2/l19
  • Zhang et al. (2021) Zhang, Z. J., Yan, K., Li, C. M., Zhang, G. Q., & Wang, F. Y. 2021, The Astrophysical Journal, 906, 49, doi: 10.3847/1538-4357/abceb9