Dust–Gas Scaling Relations and OH Abundance in the Galactic ISM

, , , , , , , , , , , , and

Published 2018 July 20 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Hiep Nguyen et al 2018 ApJ 862 49 DOI 10.3847/1538-4357/aac82b

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/862/1/49

Abstract

Observations of interstellar dust are often used as a proxy for total gas column density NH. By comparing Planck thermal dust data (Release 1.2) and new dust reddening maps from Pan-STARRS 1 and 2MASS, with accurate (opacity-corrected) H i column densities and newly published OH data from the Arecibo Millennium survey and 21-SPONGE, we confirm linear correlations between dust optical depth τ353, reddening E(B − V), and the total proton column density NH in the range (1–30) × 1020 cm−2, along sightlines with no molecular gas detections in emission. We derive an NH/E(B − V) ratio of (9.4 ± 1.6) × 1021 cm−2 mag−1 for purely atomic sightlines at $| b| \gt 5^\circ $, which is 60% higher than the canonical value of Bohlin et al. We report a ∼40% increase in opacity σ353 = τ353/NH, when moving from the low column density (NH < 5 × 1020 cm−2) to the moderate column density (NH > 5 × 1020 cm−2) regime, and suggest that this rise is due to the evolution of dust grains in the atomic interstellar medium. Failure to account for H i opacity can cause an additional apparent rise in σ353 of the order of a further ∼20%. We estimate molecular hydrogen column densities ${N}_{{{\rm{H}}}_{2}}$ from our derived linear relations, and hence derive the OH/H2 abundance ratio of XOH ∼ 1 × 10−7 for all molecular sightlines. Our results show no evidence of systematic trends in OH abundance with ${N}_{{{\rm{H}}}_{2}}$ in the range ${N}_{{{\rm{H}}}_{2}}$ ∼ (0.1−10) × 1021 cm−2. This suggests that OH may be used as a reliable proxy for H2 in this range, which includes sightlines with both CO-dark and CO-bright gas.

Export citation and abstract BibTeX RIS

1. Introduction

Observations of neutral hydrogen in the interstellar medium (ISM) have historically been dominated by two radio spectral lines: the 21 cm line of atomic hydrogen (H i) and the microwave emission from carbon monoxide (CO), particularly the CO(J = 1–0) line. The former provides direct measurements of the warm neutral medium (WNM), and the cold neutral medium (CNM), which is the precursor to molecular clouds. The latter is widely used as a proxy for molecular hydrogen (H2), often via the use of an empirical "X-factor," (e.g., Bolatto et al. 2013). The processes by which CNM and molecular clouds form from warm atomic gas sows the seeds of structure into clouds, laying the foundations for star formation. Being able to observationally track the ISM through this transition is of key importance.

However, there is strong evidence for gas not seen in either H i or CO. This undetected material is often called "dark gas," following Grenier et al. (2005). These authors found an excess of diffuse gamma-ray emission from the Local ISM, with respect to the expected flux due to cosmic-ray interactions with the gas mass estimated from H i and CO. Similar conclusions have been reached using many different tracers, including γ-rays (e.g., Abdo et al. 2010; Ackermann et al. 2012, 2011), infrared emission from dust (e.g., Blitz et al. 1990; Reach et al. 1994; Douglas & Taylor 2007; Planck Collaboration et al. 2011, 2014a), dust extinction (e.g., Paradis et al. 2012; Lee et al. 2015), C ii emission (Pineda et al. 2013; Langer et al. 2014; Tang et al. 2016), and OH 18 cm emission and absorption (e.g., Wannier et al. 1993; Liszt & Lucas 1996; Barriault et al. 2010; Allen et al. 2012, 2015; Engelke & Allen 2018).

While a minority of studies have suggested that cold, optically thick H i could account for almost all the missing gas mass (Fukui et al. 2015), CO-dark H2 is generally expected to be a major constituent, particularly in the envelopes of molecular clouds (e.g., Lee et al. 2015). In diffuse molecular regions, H2 is effectively self-shielded, but CO is typically photodissociated (Tielens & Hollenbach 1985a, 1985b; van Dishoeck & Black 1988; Wolfire et al. 2010; Glover & Mac Low 2011; Lee et al. 2015; Glover & Smith 2016), meaning that CO lines are a poor tracer of H2 in such environments. Indeed Herschel observations of C ii suggest that between 20%–75% of the H2 in the Galactic plane may be CO-dark (Pineda et al. 2013).

For the atomic medium, the mass of warm H i can be computed directly from measured line intensities under the optically thin assumption. However, cold H i with spin temperature Ts ≲ 100 K suffers from significant optical depth effects, leading to an underestimation of the total column density. This difficulty is generally addressed by combining H i absorption and emission profiles observed toward (and immediately adjacent to) bright, compact continuum background sources. Such studies find that the optically thin assumption underestimates the true H i column by no more than a few tens of percent along most Milky Way sightlines (e.g., Dickey et al. 1983, 2000, 2003; Heiles & Troland 2003a, 2003b; Liszt 2014a; Lee et al. 2015); though, the fraction missed in some localized regions may be much higher (Bihr et al. 2015).

Since dust and gas are generally well mixed, absorption due to dust grains has been widely used as a proxy for total gas column density. Early work (e.g., Savage & Jenkins 1972, Bohlin et al. 1978) observed Lyα and H2 absorption in stellar spectra to calibrate the relationship between total hydrogen column density NH, and the color excess E(B − V). Similar work was carried out by comparing X-ray absorption with optical extinction, AV (Reina & Tarenghi 1973, Gorenstein 1975). Bohlin et al's value of NH/E(B − V) =5.8 × 1021 cm−2 mag−1 has become a widely accepted standard.

Dust emission is also a powerful tool and requires no background source population. The dust emission spectrum in the bulk of the ISM peaks in the FIR-to-millimeter range, and arises mostly from large grains in thermal equilibrium with the ambient local radiation field (Draine 2003, Draine & Li 2007). It has long been recognized that FIR dust emission could potentially be a better tracer of NH than H i and CO (de Vries et al. 1987, Heiles et al. 1988, Blitz et al. 1990; Reach et al. 1994). An excess of dust intensity and/or optical depth above a linear correlation with NH (as measured by H i and CO) is typically found in the range AV = 0.3–2.7 mag (Planck Collaboration et al. 2011, 2014a, 2014b; Martin et al. 2012), consistent with the range where CO-dark H2 can exist. Alternative explanations cannot be definitively ruled out, however. These include (1) the evolution of dust grains across the gas phases, (2) underestimation of the total gas column due to significant cold H i opacity, and (3) insufficient sensitivity for CO detection. It has also been impossible to rule out remaining systematic effects in the Planck data or bias in the estimate of τ353 introduced by the choice of the modified blackbody model.

In this study, we examine the correlations between accurately derived H i column densities and dust-based proxies for NH. We make use of opacity-corrected H i column densities derived from two surveys: the Arecibo Millennium Survey (MS, Heiles & Troland 2003b, hereafter HT03), and 21-SPONGE (Murray et al. 2015), both of which used on-/off-source measurements toward extragalactic radio continuum sources to derive accurate physical properties for the atomic ISM. We also make use of archival OH data from the Millennium Survey, recently published for the first time in a companion paper, Li et al. (2018). OH is an effective tracer of diffuse molecular regions (Wannier et al. 1993; Liszt & Lucas 1996; Barriault et al. 2010; Allen et al. 2012, 2015; Xu et al. 2016; Li et al. 2018), and has recently been surveyed at high sensitivity in parts of the Galactic plane (Dawson et al. 2014; Bihr et al. 2015). There exists both theoretical and observational evidence for the close coexistence of interstellar OH and H2. Observationally, they appear to reside in the same environments, as evidenced by tight relations between their column densities (Weselak & Krełowski 2014). Theoretically, the synthesis of OH is driven by the ions O+ and ${{\rm{H}}}_{3}^{+}$ but requires H2 as the precursor; once H2 becomes available, OH can be formed efficiently through the charge-exchange chemical reactions initiated by cosmic-ray ionization (van Dishoeck & Black 1986). Here we combine H i, OH, and dust data sets to obtain new measurements of the abundance ratio, XOH = NOH/${N}_{{{\rm{H}}}_{2}}$—a key quantity for the interpretation of OH data sets.

The structure of this article is as follows. In Section 2, the observations, the data processing techniques, and corrections on H i are briefly summarized. In Section 3, the results from OH observations are discussed. Section 4 discusses the relationship between τ353, E(B − V) and NH in the atomic ISM. We finally estimate the OH/H2 abundance ratio in Section 5 before concluding in Section 6.

2. Data Sets

In this study, we use the all-sky optical depth (τ353) map of the dust model data measured by Planck/IRAS (Planck Collaboration et al. 2014a—hereafter PLC2014a), the reddening E(B − V) all-sky map from Green et al. (2018), H i data from both the 21-SPONGE Survey (Murray et al. 2015) and the Millennium Survey (Heiles & Troland 2003a, HT03), OH data from the Millennium Survey (Li et al. 2018), and CO data from the Delingha 14 m Telescope, the Caltech Submillimeter Observatory (CSO), and the IRAM 30 m telescope (Li et al. 2018).

2.1. H i and OH

H i data from the Millennium Arecibo 21 cm Absorption-Line Survey (hereafter MS) was taken toward 79 strong radio sources (typically S ≳ 2 Jy) using the Arecibo L-wide receiver. The two main lines of ground state OH at 1665.402 and 1667.359 MHz were observed simultaneously toward 72 positions, and OH absorption was detected along 19 of these sightlines (see also Li et al. 2018). The observations are described in detail by HT03. Briefly, their so-called Z16 observation pattern consists of one on-source absorption spectrum toward the background radio source and 16 off-source emission spectra with the innermost positions at 1.0 HPBW and the outermost positions at $\sqrt{2}$ HPBW from the central source. The off-source "expected" emission spectrum, the emission profile we would observe in the absence of the continuum source, is then estimated by modeling the 17-point measurements. In this work, we use the published values of HT03 for the total H i column density, NH i (scaled as described below), and use the off-source (expected) MS emission profiles to compute the H i column density under the optically thin assumption, ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$, where required. We compute OH column densities ourselves, as described in Section 3. All OH emission and absorption spectra are scaled to a main-beam temperature scale using a beam efficiency of ηb = 0.5 (Heiles et al. 2001), appropriate if the OH is not extended compared to the Arecibo beam size of 3'.

In order to increase the source sample, we also use H i data from the Very Large Array (VLA) 21-SPONGE Survey, which observed 30 continuum sources, including 16 in common with the Millennium Survey sample (Murray et al. 2015). 21-SPONGE used on-source absorption data from the VLA, combining them with off-source emission profiles observed with Arecibo. Murray et al. (2015) report an excellent agreement between the optical depths measured by the two surveys, demonstrating that the single dish Arecibo absorption profiles are not significantly contaminated with resolved 21 cm emission. Note that in this work we have used an updated scaling of the 21-SPONGE emission profiles, which applies a beam efficiency factor of 0.94 to the Arecibo spectra. The total number of unique sightlines presented in this work is therefore 93. The locations of all observed sources in Galactic coordinates are presented in Figure 1. Where sources were observed in both the MS and 21-SPONGE, we use the MS data.

Figure 1.

Figure 1. Locations of all 93 sightlines considered in this study, overlaid on the map of dust optical depth τ353. Squares show H i absorption detections (93/93); red circles show OH absorption detections (19/72); black circles show nondetections (51/72); red triangles show CO detections (19/44); and black triangles show nondetections (25/44). For purely atomic sightlines (those with no molecular detection at the threshold discussed in Section 4), the squares are colored red. Note that the absence of a symbol indicates that the sightline was not observed in that particular tracer. The labeled sightline toward 3C132 (far left) shows the single position detected in H i and OH but not detected in CO emissions. The "X" marker labels the center of the Milky Way. Note that the symbols for a small number of sightlines entirely overlap due to their proximity on the sky.

Standard image High-resolution image

2.1.1. H i Intensity Scale Corrections

We check our ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ against the Leiden–Argentine–Bonn survey (LAB, Hartmann & Burton 1997; Kalberla et al. 2005) and the HI4PI survey (HI4PI Collaboration et al. 2016). Both are widely regarded as a gold standard in the absolute calibration of Galactic H i. We find that the optically thin column densities derived from 21-SPONGE are consistent with LAB and HI4PI. However, the MS values are systematically lower than both LAB and HI4PI by a factor of ∼1.14. A possible explanation for this difference lies in the fact that (in contrast to 21-SPONGE) the MS did not apply a main-beam efficiency.

To bring the MS data set in-line with LAB, HI4PI, and 21-SPONGE, one might assume that both the on-source and off-source spectra must be rescaled, and the opacity-corrected column densities recomputed according to the method of HT03 (or equivalent). However, NH i may in fact be obtained from the tabulated values of HT03, with no need to perform a full reanalysis of the data. For warm components, the tabulated values of NH i are simply scaled by 1.14—appropriate since these were originally computed directly from the the integrated off-source (expected) profiles under the optically thin assumption. For cold components, we recall that the radiative transfer equations for the on-source and off-source (expected) spectra in the MS data set are given by:

Equation (1)

Equation (2)

where ${T}_{{\rm{B}}}^{\mathrm{OFF}}(v)$ and ${T}_{{\rm{B}}}^{\mathrm{ON}}(v)$ are the main-beam temperatures of the off-source spectrum and on-source spectrum, respectively. Ts is the spin temperature, τv is the optical depth, Trx is the receiver temperature (∼25 K), and Tc is the main-beam temperature of the continuum source, obtained from the line-free portions of the on-source spectrum. Tbg is the continuum background brightness temperature including the 2.7 K isotropic radiation from CMB and the Galactic synchrotron background at the source position. Equations (1) and (2) may be solved for τv and Ts:

Equation (3)

Equation (4)

From Equation (3), it is clear that optical depth is unchanged by any rescaling, which will affect both the numerator and denominator of the expression identically. Only Ts must be recomputed. This is done on a component-by-component basis from the tabulated Gaussian fit parameters for peak optical depth, τ0, peak brightness temperature (scaled by 1.14), and the linewidth Δv. The corrected NH i is obtained from

Equation (5)

where the factor 1.94 includes the usual constant 1.8224 and the 1.065 arising from the integration over the Gaussian line profile.

2.1.2. NH i versus ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$

We show in Figure 2 the correlation between NH i and ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ toward all 93 positions. While optically thin H i column density is comparable with the true column density in diffuse/low-density regions with NH i ≲ 5 × 1020 cm−2, opacity effects start to become apparent above ∼5 × 1020 cm−2.

Figure 2.

Figure 2. Ratio f = NH i/${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ as a function of opacity-corrected NH i along 93 sightlines from the MS and 21-SPONGE surveys. Circles show accurate NH i obtained via on- and off-source observations (HT03; scaled as described in the text), with the 34 atomic sightlines (selection criteria described in Section 4) filled gray and all other points filled black. Red triangles show NH i obtained from ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ assuming a single isothermal component of Ts ∼ 144 K. The vertical dashed line is plotted at NH i = 5 × 1020 cm−2; the horizontal dashed line marks where NH i = ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$.

Standard image High-resolution image

If a linear fit is performed to the data, the ratio f = NH i/${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ may be described as a function of log(${N}_{{\rm{H}}\,{\rm{I}}}^{* }$/1020) with a slope of (0.19 ± 0.02) and an intercept of (0.89 ± 0.02) (see also Lee et al. 2015). Alternatively, a simple isothermal correction to the optically thin ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ data with Ts ∼ 144 K also yields a good agreement with our data points, as illustrated in Figure 2 (see also Liszt 2014b). This approach also better fits the low NH i plateau, NH i < 5 × 1020 cm−2, below which ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ ≈ NH i. While a single component with a constant spin temperature is a poor physical description of interstellar H i, it can provide a reasonable (if crude) correction for opacity.

2.2. CO

As described in Li et al. (2018), a CO follow-up survey was conducted toward 44 of the sightlines considered in this work. The J = 1–0 transitions of 12CO, 13CO, and C18O were observed with the Delingha 13.7 m telescope in China. 12CO(J = 2–1) data for 45 sources and J = 3–2 data for 8 sources with strong 12CO emission were taken with the 10.4 m CSO on Maunakea, with further supplementary data obtained by the IRAM 30 m telescope. In this work, we use CO data solely to identify and exclude from some parts of the analysis positions with detected CO-bright molecular gas—comprising 19 of the 44 observed positions. These positions are identified in Figure 1.

2.3. Dust

To trace the total gas column density NH, we use publicly available all-sky maps of the 353 GHz dust optical depth (τ353) from the Planck satellite. The τ353 map was obtained by a modified blackbody (MBB) fit to the first 15 months of 353, 545, and 857 GHz data, together with IRAS 100 micron data (for details, see PLC2014a). The angular resolution of this data set is 5 arcmin. In this work, we use the R1.20 data release in Healpix15 format (Górski et al. 2005). For dust reddening, we employ the newly released all-sky 3D dust map of Green et al. (2018) at an angular resolution of 3farcm4–13farcm7, which was derived from 2MASS and the latest Pan-STARRS 1 data photometry. In contrast to emission-based dust maps that depend on the modeling of the temperature, optical depth, and the shape of the emission spectrum, in maps based on stellar photometry reddening values are more directly measured and not contaminated from zodiacal light or large-scale structure. Here we convert the Green et al. (2018) Bayestar17 dust map to E(B − V) by applying a scaling factor of 0.884, as described in the documentation accompanying the data release.16

3. OH Data Analysis

The Millennium Survey OH data consists of on-source and off-source "expected" spectra for each of the OH lines. In our companion paper (Li et al. 2018), we use the method of HT03 to derive OH optical depths, excitation temperatures and column densities. Namely, we obtain solutions for the excitation temperature, Tex, and τ via Gaussian fitting (to both the on-source and off-source spectra) that explicitly includes the appropriate treatment of the radiative transfer. In the present work, we use a simpler channel-by-channel method for the derivation of Tex.

The radiative transfer equations for the on-source and off-source (expected) spectra are identical to those for H i, given above in Equations (1) and (2). All terms and their meanings are identical, with the exception that the spin temperature, Ts is replaced by Tex. Tbg is the continuum background brightness temperature including the 2.7 K isotropic radiation from CMB and the Galactic synchrotron background at the source position. For consistency with HT03 and Li et al. (2018), we estimate the synchrotron contribution at 1665.402 and 1667.359 MHz from the 408 MHz continuum map of Haslam et al. (1982), by adopting a temperature spectral index of 2.8, such that

Equation (6)

resulting in typical values of around 3.5 K. The background continuum contribution from Galactic H II regions may be safely ignored, since the continuum sources we observed are either at high Galactic latitudes or Galactic anti-center longitudes. Thus, in line-free portions of the off-source spectra:

Equation (7)

In the absence of information about the true gas distribution, we assume that OH clouds cover fully both the continuum source and the main beam of the telescope. We may therefore solve Equations (1) and (2) to derive Tex and τv for each of the OH lines, as shown in Equations (3) and (4) for the case of H i.

We fit each OH opacity spectrum (cf. Equation (3)) with a set of Gaussian profiles to obtain the peak optical depth (τ0,n), central velocity (v0,n), and FWHM (Δvn) of each component, n. Equation (4) is then used to calculate excitation temperature spectra. Examples of ${e}^{-{\tau }_{v}}$, ${T}_{{\rm{B}}}^{\mathrm{OFF}}$, and Tex spectra are shown in Figure 3, together with their associated Gaussian fits. It can be seen that the Tex spectra are approximately flat within the FWHM of each Gaussian component. We therefore compute an excitation temperature for each component from the mean Tex in the range v0,n ± Δv/2.

Figure 3.

Figure 3. Example of OH 1667 MHz ${e}^{-{\tau }_{v}}$ (top), expected ${T}_{{\rm{B}}}^{\mathrm{OFF}}$ (middle), and Tex (bottom) spectra for the source P0428+20. The FWHM of the Gaussian fits to the absorption profile are used to define the range over which Tex is computed for each component, shown as white regions in the bottom panel.

Standard image High-resolution image

Figure 4 compares the τ0 and Tex values obtained from our method with those of Li et al. (2018), demonstrating that the two methods generally return consistent results. Minor differences arise only for the most complex sightlines through the Galactic plane (G197.0+1.1, T0629+10), where the spectra are not simple to analyze; however, even these points are mostly consistent to within the errors.

Figure 4.

Figure 4. Comparison between derived values of the peak optical depth τ (left panel), and Tex (right panel) for both OH main lines, 1667 MHz (black) and 1665 MHz (gray), as obtained from our companion paper by Li et al. (2018) and the present work. The dashed lines mark where the two values are equal.

Standard image High-resolution image

We compute total OH column densities, NOH, independently from both the 1667 and 1665 MHz lines via:

Equation (8)

Equation (9)

where the constants include Einstein A-coefficients of A1667 = 7.778 × 10−11 s−1 and A1665 = 7.177 × 10−11 s−1 for the OH main lines (Destombes et al. 1977). All values of τ0, Tex, and NOH are tabulated in Table 1.

Table 1.  Parameters for OH Main Lines

Source l/b OH(1665) OH(1667)
    τ Vlsr ΔV Tex N(OH) τ Vlsr ΔV Tex N(OH)
(Name) (°)   (km s−1) (km s−1) (K) (1014 cm−2)   (km s−1) (km s−1) (K) (1014 cm−2)
3C105 187.6/−33.6 0.0156 ± 0.0003 8.14 ± 0.01 0.95 ± 0.03 4.65 ± 1.86 0.29 ± 0.12 0.0265 ± 0.0004 8.17 ± 0.01 0.94 ± 0.02 3.95 ± 0.95 0.23 ± 0.06
3C105 187.6/−33.6 0.0062 ± 0.0003 10.22 ± 0.02 0.93 ± 0.06 8.5 ± 4.89 0.21 ± 0.12 0.0104 ± 0.0004 10.25 ± 0.02 0.96 ± 0.04 7.66 ± 3.41 0.18 ± 0.08
3C109 181.8/−27.8 0.0023 ± 0.0003 9.15 ± 0.11 1.03 ± 0.26 18.28 ± 27.06 0.18 ± 0.27 0.0036 ± 0.0004 9.24 ± 0.05 0.75 ± 0.12 24.58 ± 8.7 0.16 ± 0.06
3C109 181.8/−27.8 0.0036 ± 0.0003 10.45 ± 0.07 0.98 ± 0.15 13.97 ± 5.41 0.21 ± 0.09 0.0053 ± 0.0004 10.55 ± 0.04 1.02 ± 0.1 13.63 ± 4.48 0.18 ± 0.06
3C123 170.6/−11.7 0.0191 ± 0.0007 3.65 ± 0.06 1.19 ± 0.11 10.92 ± 3.26 1.05 ± 0.33 0.0348 ± 0.0009 3.71 ± 0.04 1.22 ± 0.07 10.92 ± 2.69 1.1 ± 0.28
3C123 170.6/−11.7 0.0431 ± 0.0023 4.43 ± 0.01 0.53 ± 0.03 8.06 ± 0.78 0.78 ± 0.09 0.0919 ± 0.0029 4.46 ± 0.0 0.53 ± 0.01 7.7 ± 0.65 0.89 ± 0.08
3C123 170.6/−11.7 0.0337 ± 0.0008 5.37 ± 0.01 0.91 ± 0.03 11.59 ± 4.3 1.53 ± 0.57 0.0784 ± 0.0009 5.47 ± 0.01 0.92 ± 0.01 8.79 ± 2.57 1.5 ± 0.44
3C131 171.4/−7.8 0.0065 ± 0.0005 4.55 ± 0.02 0.56 ± 0.05 12.52 ± 3.59 0.19 ± 0.06 0.0117 ± 0.0004 4.64 ± 0.01 0.78 ± 0.04 6.96 ± 1.98 0.15 ± 0.04
3C131 171.4/−7.8 0.0073 ± 0.0006 6.81 ± 0.06 2.91 ± 0.23 8.94 ± 4.54 0.82 ± 0.42 0.0089 ± 0.0005 5.84 ± 0.03 0.67 ± 0.08 11.04 ± 2.07 0.16 ± 0.04
3C131 171.4/−7.8 0.0166 ± 0.0007 6.59 ± 0.01 0.42 ± 0.02 5.69 ± 0.85 0.17 ± 0.03 0.0319 ± 0.0007 6.55 ± 0.01 0.45 ± 0.02 5.99 ± 0.84 0.2 ± 0.03
3C131 171.4/−7.8 0.0521 ± 0.0007 7.23 ± 0.0 0.55 ± 0.01 5.91 ± 0.64 0.72 ± 0.08 0.0927 ± 0.0005 7.22 ± 0.0 0.65 ± 0.01 5.98 ± 0.34 0.85 ± 0.05
3C132 178.9/−12.5 0.0033 ± 0.0003 7.82 ± 0.04 0.9 ± 0.1 15.55 ± 6.23 0.19 ± 0.08 0.0056 ± 0.0003 7.79 ± 0.02 0.79 ± 0.06 23.56 ± 2.17 0.25 ± 0.03
3C133 177.7/−9.9 0.1008 ± 0.001 7.66 ± 0.0 0.53 ± 0.0 4.47 ± 0.44 1.01 ± 0.1 0.2132 ± 0.0014 7.68 ± 0.0 0.52 ± 0.0 3.25 ± 0.27 0.85 ± 0.07
3C133 177.7/−9.9 0.0149 ± 0.001 7.94 ± 0.02 1.22 ± 0.04 7.08 ± 3.08 0.55 ± 0.24 0.0333 ± 0.0013 7.96 ± 0.01 1.23 ± 0.02 4.17 ± 0.99 0.4 ± 0.1
3C154 185.6/4.0 0.0266 ± 0.0006 −2.32 ± 0.02 0.74 ± 0.03 2.69 ± 1.93 0.23 ± 0.16 0.0429 ± 0.0006 −2.34 ± 0.01 0.71 ± 0.02 2.57 ± 0.75 0.19 ± 0.05
3C154 185.6/4.0 0.01 ± 0.0006 −1.39 ± 0.04 0.83 ± 0.09 5.2 ± 5.28 0.18 ± 0.19 0.0181 ± 0.0005 −1.34 ± 0.02 0.94 ± 0.05 4.46 ± 1.79 0.18 ± 0.07
3C154 185.6/4.0 0.0038 ± 0.0005 2.23 ± 0.07 1.14 ± 0.17 5.83 ± 6.56 0.11 ± 0.12 0.0054 ± 0.0004 2.19 ± 0.05 1.57 ± 0.13 0.54 ± 8.69 0.01 ± 0.17
3C167 207.3/1.2 0.0106 ± 0.0019 18.46 ± 0.12 1.49 ± 0.35 4.75 ± 17.95 0.32 ± 1.22 0.009 ± 0.0018 17.77 ± 0.15 1.76 ± 0.49 4.59 ± 9.57 0.17 ± 0.36
3C18 118.6/−52.7 0.0031 ± 0.0003 −8.52 ± 0.11 2.64 ± 0.27 10.92 ± 14.88 0.38 ± 0.52 0.006 ± 0.0003 −8.34 ± 0.05 2.61 ± 0.14 9.2 ± 6.04 0.34 ± 0.23
3C18 118.6/−52.7 0.0056 ± 0.0004 −7.82 ± 0.02 0.67 ± 0.07 6.45 ± 3.7 0.1 ± 0.06 0.0079 ± 0.0004 −7.85 ± 0.01 0.6 ± 0.04 4.83 ± 1.6 0.05 ± 0.02
3C207 213.0/30.1 0.015 ± 0.0002 4.55 ± 0.01 0.76 ± 0.01 2.94 ± 1.55 0.14 ± 0.08 0.0266 ± 0.0002 4.55 ± 0.0 0.77 ± 0.01 2.48 ± 0.46 0.12 ± 0.02
3C409 63.4/−6.1 0.0058 ± 0.0011 14.59 ± 0.27 1.68 ± 0.35 11.31 ± 8.53 0.47 ± 0.38 0.0055 ± 0.0015 14.68 ± 0.33 1.52 ± 0.4 7.83 ± 11.73 0.16 ± 0.24
3C409 63.4/−6.1 0.0204 ± 0.0025 15.4 ± 0.01 0.89 ± 0.05 3.18 ± 2.31 0.25 ± 0.18 0.0275 ± 0.0032 15.42 ± 0.01 0.86 ± 0.04 0.62 ± 1.0 0.03 ± 0.06
3C410 69.2/−3.8 0.0044 ± 0.0006 6.32 ± 0.04 1.89 ± 0.15 13.41 ± 11.22 0.47 ± 0.4 0.0079 ± 0.0005 6.38 ± 0.02 2.32 ± 0.09 6.4 ± 5.25 0.28 ± 0.23
3C410 69.2/−3.8 0.0089 ± 0.0006 6.21 ± 0.01 0.65 ± 0.04 8.46 ± 2.4 0.21 ± 0.06 0.0193 ± 0.0005 6.26 ± 0.01 0.81 ± 0.02 3.81 ± 1.28 0.14 ± 0.05
3C410 69.2/−3.8 0.0044 ± 0.0003 10.7 ± 0.03 0.71 ± 0.07 10.06 ± 5.89 0.13 ± 0.08 0.0085 ± 0.0002 10.71 ± 0.02 0.81 ± 0.04 4.15 ± 3.09 0.07 ± 0.05
3C410 69.2/−3.8 0.0054 ± 0.0002 11.67 ± 0.03 0.84 ± 0.07 4.83 ± 4.66 0.09 ± 0.09 0.0115 ± 0.0002 11.68 ± 0.02 0.82 ± 0.03 2.93 ± 3.0 0.07 ± 0.07
3C454.3 86.1/−38.2 0.0023 ± 0.0001 −9.67 ± 0.03 1.6 ± 0.09 4.63 ± 12.36 0.07 ± 0.19 0.0044 ± 0.0001 −9.54 ± 0.01 1.25 ± 0.04 8.13 ± 6.06 0.1 ± 0.08
3C75 170.3/−44.9 0.0071 ± 0.0005 −10.36 ± 0.04 1.3 ± 0.12 3.45 ± 5.41 0.14 ± 0.21 0.014 ± 0.0008 −10.36 ± 0.03 1.22 ± 0.09 3.51 ± 1.56 0.14 ± 0.06
4C13.67 43.5/9.2 0.0464 ± 0.0043 4.85 ± 0.05 1.1 ± 0.12 10.43 ± 2.77 2.28 ± 0.69 0.0567 ± 0.0057 4.89 ± 0.05 1.12 ± 0.14 10.34 ± 2.13 1.55 ± 0.4
4C22.12 188.1/0.0 0.0058 ± 0.001 −2.84 ± 0.07 0.79 ± 0.19 6.54 ± 7.03 0.13 ± 0.14 0.0102 ± 0.0011 −2.73 ± 0.04 0.78 ± 0.12 6.72 ± 2.32 0.13 ± 0.05
4C22.12 188.1/0.0 0.0172 ± 0.0012 −1.78 ± 0.02 0.56 ± 0.05 5.07 ± 2.12 0.21 ± 0.09 0.0354 ± 0.0013 −1.78 ± 0.01 0.54 ± 0.03 3.78 ± 0.74 0.17 ± 0.03
G196.6+0.2 196.6/0.2 0.0044 ± 0.0005 3.26 ± 0.11 1.94 ± 0.27 10.82 ± 12.22 0.4 ± 0.46 0.0062 ± 0.0005 3.4 ± 0.09 2.38 ± 0.22 8.85 ± 8.6 0.31 ± 0.3
G197.0+1.1 197.0/1.1 0.0126 ± 0.0005 4.83 ± 0.04 1.88 ± 0.09 6.94 ± 3.82 0.7 ± 0.39 0.0191 ± 0.001 4.73 ± 0.04 1.65 ± 0.1 4.8 ± 2.17 0.36 ± 0.16
G197.0+1.1 197.0/1.1 0.0059 ± 0.0009 7.46 ± 0.05 0.65 ± 0.11 0.31 ± 10.58 0.0 ± 0.17 0.0078 ± 0.0015 7.34 ± 0.06 0.65 ± 0.14 1.61 ± 5.87 0.02 ± 0.07
G197.0+1.1 197.0/1.1 0.0049 ± 0.0005 17.01 ± 0.12 2.47 ± 0.28 10.99 ± 9.9 0.57 ± 0.52 0.0081 ± 0.0034 16.26 ± 0.17 0.91 ± 0.3 6.45 ± 2.95 0.11 ± 0.08
G197.0+1.1 197.0/1.1 0.0052 ± 0.0015 17.59 ± 0.03 0.25 ± 0.08 9.87 ± 1.81 0.05 ± 0.03 0.0127 ± 0.0013 17.38 ± 0.2 1.46 ± 0.35 5.46 ± 4.35 0.24 ± 0.2
G197.0+1.1 197.0/1.1 0.0237 ± 0.001 32.01 ± 0.01 0.57 ± 0.03 4.75 ± 1.92 0.27 ± 0.11 0.043 ± 0.0017 32.01 ± 0.01 0.54 ± 0.02 3.96 ± 1.04 0.22 ± 0.06
P0428+20 176.8/−18.6 0.0014 ± 0.0002 3.6 ± 0.08 1.01 ± 0.19 13.48 ± 7.8 0.08 ± 0.05 0.0029 ± 0.0003 3.54 ± 0.03 0.69 ± 0.08 4.45 ± 9.98 0.02 ± 0.05
P0428+20 176.8/−18.6 0.0075 ± 0.0002 10.7 ± 0.02 1.09 ± 0.04 13.49 ± 3.45 0.47 ± 0.12 0.0136 ± 0.0002 10.7 ± 0.01 1.1 ± 0.02 12.72 ± 1.62 0.45 ± 0.06
T0526+24 181.4/−5.2 0.0172 ± 0.0073 7.55 ± 0.29 1.9 ± 1.13 13.7 ± 15.65 1.91 ± 2.59 0.043 ± 0.0102 7.56 ± 0.14 2.43 ± 0.75 10.19 ± 7.5 2.52 ± 2.1
T0629+10 201.5/0.5 0.0043 ± 0.0022 0.16 ± 0.0 0.65 ± 0.4 4.16 ± 2.97 0.05 ± 0.05 0.0103 ± 0.0035 0.35 ± 0.0 1.18 ± 0.41 3.25 ± 1.93 0.09 ± 0.07
T0629+10 201.5/0.5 0.0387 ± 0.0074 3.14 ± 0.13 1.1 ± 0.0 3.85 ± 0.47 0.7 ± 0.16 0.0577 ± 0.0113 3.07 ± 0.14 1.1 ± 0.0 4.54 ± 0.55 0.68 ± 0.16
T0629+10 201.5/0.5 0.0169 ± 0.0015 1.46 ± 0.07 1.39 ± 0.26 2.19 ± 2.11 0.22 ± 0.22 0.0281 ± 0.0029 1.51 ± 0.08 1.23 ± 0.25 2.9 ± 0.91 0.24 ± 0.09
T0629+10 201.5/0.5 0.1607 ± 0.0104 3.6 ± 0.01 0.61 ± 0.02 3.83 ± 0.35 1.59 ± 0.19 0.2536 ± 0.0154 3.6 ± 0.01 0.65 ± 0.03 4.72 ± 0.64 1.84 ± 0.29
T0629+10 201.5/0.5 0.0811 ± 0.002 4.62 ± 0.01 0.76 ± 0.03 6.43 ± 0.97 1.68 ± 0.27 0.1553 ± 0.0037 4.61 ± 0.01 0.67 ± 0.03 6.62 ± 1.01 1.63 ± 0.26
T0629+10 201.5/0.5 0.0747 ± 0.0018 6.09 ± 0.02 1.06 ± 0.05 5.44 ± 1.58 1.84 ± 0.54 0.1165 ± 0.003 6.06 ± 0.02 1.17 ± 0.07 6.52 ± 1.67 2.1 ± 0.55
T0629+10 201.5/0.5 0.0367 ± 0.0031 7.0 ± 0.02 0.49 ± 0.06 4.36 ± 0.68 0.33 ± 0.07 0.0631 ± 0.0056 7.0 ± 0.02 0.5 ± 0.06 4.2 ± 0.3 0.31 ± 0.05
T0629+10 201.5/0.5 0.0174 ± 0.0018 7.9 ± 0.05 0.83 ± 0.13 3.44 ± 1.67 0.21 ± 0.11 0.0307 ± 0.003 7.91 ± 0.05 0.82 ± 0.13 3.65 ± 1.21 0.22 ± 0.08

Download table as:  ASCIITypeset images: 1 2

4. Dust-based Proxies for Total Neutral Gas Column Density

In this section, we will investigate the correlations between dust properties and the total gas column density NH. Specifically, we consider dust optical depth at 353 GHz, τ353, and reddening, E(B − V), with data sets sourced as described in Section 2.3. When these quantities are used as proxies for NH, a single linear relationship between the measured quantity and NH is typically assumed. In this work, our H i data set provides accurate (opacity-corrected) atomic column densities, while complementary OH and CO data allow us to identify and exclude sightlines with molecular gas (dark or not). We are therefore able to measure τ353/NH and E(B − V)/NH along a sample of purely atomic sightlines for which NH is very well constrained.

Table 2.  34 Atomic Sightlines

Sources l/b NH i ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ στ(OH1667) ${N}_{{{\rm{H}}}_{2}}$(upper limit)a τ353 E(B − V)
(Name) (°) (1020 cm−2) (1020 cm−2) (10−4) (1020 cm−2) (10−6) (10−2 mag)
3C33 129.4/−49.3 3.25 ± 0.0 3.2 ± 0.1 12.14 0.6 2.16 ± 0.07 3.54 ± 0.42
3C142.1 197.6/−14.5 25.11 ± 2.6 19.6 ± 0.8 10.55 0.52 21.53 ± 0.72 21.71 ± 0.81
3C138 187.4/−11.3 22.9 ± 1.1 19.9 ± 0.3 5.02 0.25 21.63 ± 0.81 17.47 ± 0.59
3C79 164.1/−34.5 10.86 ± 1.2 9.8 ± 0.8 37.03 1.84 9.23 ± 0.37 12.67 ± 0.78
3C78 174.9/−44.5 11.69 ± 0.5 10.3 ± 0.2 13.25 0.66 12.45 ± 0.63 14.64 ± 1.07
3C310 38.5/60.2 4.29 ± 0.1 4.0 ± 0.1 16.19 0.8 3.48 ± 0.15 2.75 ± 0.53
3C315 39.4/58.3 5.48 ± 0.4 4.7 ± 0.0 12.96 0.64 3.98 ± 0.09 5.63 ± 0.26
3C234 200.2/52.7 1.84 ± 0.0 1.9 ± 1.1 12.66 0.63 0.78 ± 0.03 1.64 ± 0.56
3C236 190.1/54.0 1.38 ± 0.0 1.3 ± 1.2 10.72 0.53 0.7 ± 0.03 2.04 ± 0.49
3C64 157.8/−48.2 7.29 ± 0.2 6.9 ± 0.8 33.12 1.65 6.55 ± 0.35 9.01 ± 0.36
P0531+19 186.8/−7.1 27.33 ± 0.7 25.4 ± 0.3 6.37 0.32 20.75 ± 0.62 20.54 ± 1.3
P0820+22 201.4/29.7 4.82 ± 0.2 4.8 ± 1.1 7.09 0.35 3.73 ± 0.11 2.56 ± 0.4
3C192 197.9/26.4 4.56 ± 0.1 4.5 ± 0.1 20.66 1.03 3.38 ± 0.08 4.05 ± 0.53
3C98 179.8/−31.0 12.7 ± 0.5 11.3 ± 1.3 12.26 0.61 13.72 ± 0.41 17.73 ± 1.02
3C273 289.9/64.4 2.35 ± 0.0 2.3 ± 0.1 21.0 1.04 1.3 ± 0.09 1.98 ± 0.41
DW0742+10 209.8/16.6 2.77 ± 0.0 2.8 ± 0.9 8.01 0.4 1.6 ± 0.03 1.99 ± 0.25
3C172.0 191.2/13.4 8.89 ± 0.2 8.6 ± 1.1 13.02 0.65 5.66 ± 0.08 4.7 ± 0.52
3C293 54.6/76.1 1.46 ± 0.1 1.5 ± 1.1 6.24 0.31 1.31 ± 0.09 2.83 ± 0.75
3C120 190.4/−27.4 18.17 ± 2.1 10.7 ± 0.1 28.29 1.41 29.26 ± 1.08 22.74 ± 1.03
CTA21 166.6/−33.6 10.97 ± 0.4 10.0 ± 0.8 27.35 1.36 10.39 ± 0.44 13.3 ± 0.96
P1117+14 240.4/65.8 1.79 ± 0.0 1.8 ± 0.3 15.0 0.75 1.5 ± 0.04 2.73 ± 0.54
3C264.0 237.0/73.6 1.97 ± 0.0 2.0 ± 0.4 6.29 0.31 1.64 ± 0.08 2.88 ± 0.34
3C208.1 213.6/33.6 3.15 ± 0.1 3.2 ± 0.2 18.67 0.93 3.12 ± 0.04 2.93 ± 0.43
3C208.0 213.7/33.2 3.41 ± 0.1 3.5 ± 0.2 19.69 0.98 3.38 ± 0.07 4.37 ± 0.47
4C32.44 67.2/81.0 1.23 ± 0.0 1.3 ± 0.6 9.88 0.49 0.81 ± 0.02 1.94 ± 0.29
3C272.1 280.6/74.7 2.82 ± 0.0 2.8 ± 0.3 10.24 0.51 1.73 ± 0.28 2.43 ± 0.31
4C07.32 322.2/68.8 2.43 ± 0.0 2.4 ± 0.3 30.7 1.53 2.32 ± 0.06 4.3 ± 0.41
3C245 233.1/56.3 2.39 ± 0.0 2.4 ± 0.2 11.36 0.56 2.22 ± 0.06 2.71 ± 0.36
3C348 23.0/29.2 6.56 ± 0.2 6.0 ± 0.1 16.51 0.82 5.7 ± 0.15 9.8 ± 0.33
3C286 56.5/80.7 2.33 ± 0.1 2.4 ± 2.7 7.13 0.35 0.81 ± 0.05 2.75 ± 0.74
4C13.65 39.3/17.7 10.56 ± 0.2 9.9 ± 0.1 20.01 0.99 11.8 ± 0.39 15.63 ± 0.6
3C190.0 207.6/21.8 3.21 ± 0.0 3.4 ± 0.9 17.57 0.87 2.41 ± 0.03 1.96 ± 0.42
3C274.1 269.9/83.2 2.74 ± 0.0 2.6 ± 0.1 13.13 0.65 2.34 ± 0.02 2.64 ± 0.42
3C298 352.2/60.7 2.39 ± 0.4 2.6 ± 0.1 10.69 0.53 1.3 ± 0.07 1.97 ± 0.39

Note.

aEstimated from OH(1667) 3σ detection limits using Tex = 3.5 K, FWHM = 1 km s−1 and NOH/${N}_{{{\rm{H}}}_{2}}$ = 10−7 (see Section 5).

Download table as:  ASCIITypeset image

In the following, we consider 34/93 sightlines to be "purely atomic." These are defined as either (a) sightlines where CO and OH were observed and not detected in emission (16/93), or (b) sightlines where CO was not observed but OH was observed but not detected (18/93 sightlines). In both cases, we require that OH be undetected in the 1667 MHz line to a detection limit of NOH < 1 × 1013 cm−2 (see Li et al. 2018), which excludes some positions with weaker continuum background sources. We may confidently assume that these sightlines contain very little or no H2 and note that all but one of them lie outside the Galactic plane ($| b| \gt 10^\circ $). Figure 5 shows maps of the immediate vicinity of these sightlines in τ353 and E(B − V). Identical maps for the 19 sightlines with OH detections (see also Section 5), are shown in Figure 6.

Figure 5.
Standard image High-resolution image
Figure 5.
Standard image High-resolution image
Figure 5.

Figure 5. Maps of the immediate vicinity of the 34 "purely atomic" sightlines toward background radio sources. Dust maps (3° × 3° in Galactic coordinates) are adopted from Planck Collaboration et al. (2014a; τ353, Nside = 2048) and Green et al. (2018; E(B − V), Nside = 1024). The "X" markers show the locations of the radio sources. The contours represent the integrated intensity WCO(1−0) from the all-sky extension to the maps of Dame et al. (2001; T. Dame 2018, private communication). The base level is at 0.25 K km s−1, the typical sensitivity of the CfA CO survey, and the other contour levels are evenly spaced from the base to the maximum in each map area.

Standard image High-resolution image

In all of the following subsections, NH is taken to be equal to NH i, the opacity-corrected H i column density, as derived along sightlines with no molecular gas detected in emission.

4.1. NH from Dust Optical Depth τ353

We adopt the all-sky map of dust optical depth τ353 computed by PLC2014a. This was derived from an MBB empirical fit to IRAS and Planck maps at 3000, 857, 545, and 353 GHz, described by the expression:

Equation (10)

Here, τ353, dust temperature, Tdust, and spectral index, βdust, are the three free parameters, and Bν(Tdust) is the Planck function for dust at temperature Tdust which is, in this model, considered to be uniform along each sightline (see PLC2014a for more details). The relation between dust optical depth and total gas column density can then be written as:

Equation (11)

where σ353 is the dust opacity, κ353 is the dust emissivity cross-section per unit mass (cm2 g−1), r is the dust-to-gas mass ratio, μ is the mean molecular weight, and mH is the mass of a hydrogen atom.

Figure 7 shows the correlation between NH and τ353. A tight linear trend can be seen with a Pearson coefficient of 0.95. The value of σ353 from the orthogonal distance regression (Boggs & Rogers 1990) linear fit is (7.9 ± 0.6) × 10−27 cm2 H−1 (the intercept is set to 0), where the quoted uncertainties are the 95% confidence limits estimated from pair bootstrap resampling. This is consistent to within the uncertainties with that obtained by PLC2014a based on all-sky H i data from LAB, (6.6 ± 1.7) × 10−27 cm2 H−1. Note that here we have quoted the PLC2014a measurement made toward low NH i positions, because the lack of any H i opacity correction in that work makes this value the most reliable. However, our fit is consistent with all of the σ353 values presented in that work (which was based on the Planck R1.20 data release), to within the quoted uncertainties.

Figure 6.
Standard image High-resolution image
Figure 6.

Figure 6. Maps of the immediate vicinity of the 19 OH-detected sightlines toward background radio sources. Dust maps (3° × 3° in Galactic coordinates) are adopted from Planck Collaboration et al. (2014a; τ353, Nside = 2048) and Green et al. (2018; E(BV), Nside = 1024). The "X" markers show the locations of the radio sources. The contours represent the integrated intensity WCO(1−0) from the all-sky extension to the maps of Dame et al. (2001; T. Dame 2018, private communication). The base level is at 0.25 K km s−1, the typical sensitivity of the CfA CO survey, and the other contour levels are evenly spaced from the base to the maximum in each map area.

Standard image High-resolution image

Small systematic deviations from the linear fit, evident at the high and low column density ends of the plot, are discussed further in Section 4.3.

In order to examine the possible contribution of molecular gas to NH along the 34 atomic sightlines, we estimate upper limits on ${N}_{{{\rm{H}}}_{2}}$ from the 3σ OH detection limits using an abundance ratio of XOH = 10−7 (see Section 5). These values are tabulated in Table 2, and the resulting upper limits on NH are shown as gray triangles in Figure 7. As expected, the σ353 obtained from the fit to these upper limits is lower, at (6.4 ± 0.3) × 10−27 cm2 H−1. However, while some molecular gas may indeed be present at low levels, these limits should be considered as extreme upper bounds on the true molecular column density. This is particularly true for the most diffuse sightlines with the lowest column density (NH i < 5 × 1020 cm−2), where the observational upper limits may appear to raise NH by up to ∼50%. Molecules are not expected to be well-shielded at such low columns (and indeed even CNM is largely absent along these sightlines in our data). Even for higher column density data points, it can be readily seen from Figures 5 that all sightlines considered in this analysis lie well away from even the faintest outskirts of CO-bright molecular gas complexes. We also note that the deviations from the linear fit that will be discussed in more detail below could not be removed by any selective addition of molecular gas at levels up to these limits.

We next compare our results with the dust opacity σ353 derived by Fukui et al. (2015) (plotted on Figure 7 as a dashed line). These authors derived a smaller value than in the present work (by a factor of ∼1.5), by restricting their fit to only the warmest dust temperatures, under the assumption that these most reliably select for genuinely optically thin H i. They then applied this factor to the Planck τ353 map (excluding $| b| \lt 15^\circ $ and CO-bright sightlines) to estimate NH i, assuming that the contribution from CO-dark H2 was negligible. This resulted in NH i values ∼2–2.5 times higher than under the optically thin assumption, and motivated their hypothesis that significantly more optically thick H i exists than is usually assumed. However, we find that while the σ353 of Fukui et al. (2015) may be a good fit to some sightlines in the very low NH i regime (≲3 × 1020 cm−2), it overestimates NH i at larger column densities by ∼50%. Indeed, as will be discussed below, σ353 is not expected to remain constant as dust evolves. This (combined with some contribution from CO-dark H2) may reconcile the apparent discrepancy between their findings and absorption/emission-based measurements of the opacity-corrected H i column.

4.2. NH from Dust Reddening E(B − V)

Reddening caused by the absorption and scattering of light by dust grains is defined as:

Equation (12)

where AV is the dust extinction, RV is an empirical coefficient correlated with the average grain size, and all other symbols are defined as before. In the Milky Way, RV is typically assumed to be 3.1 (Schultz & Wiemer 1975), but it may vary between 2.5 and 6.0 along different sightlines (Goodman et al. 1995; Draine 2003).

The ratio $\langle {N}_{{\rm{H}}}/E(B-V)\rangle =5.8\times {10}^{21}\,{\mathrm{cm}}^{-2}\,{\mathrm{mag}}^{-1}$ (Bohlin et al. 1978) is a widely accepted standard, used in many fields of astrophysics to connect reddening measurements to gas column density. This value was derived from Lyα and H2 line absorption measurements toward 100 stars (see also Savage et al. 1977), and has been replicated over the years via similar methodology (e.g., Shull & van Steenberg 1985; Diplas & Savage 1994; Rachford et al. 2009). However, a number of recent works using H i 21 cm data have found significantly higher values (PLC2014a; Liszt 2014a; Lenz et al. 2017).

Here we use the all-sky map of E(B − V) from Green et al. (2018) to estimate the ratio NH/E(B − V) for our sample of purely atomic sightlines, at $| b| \gt 5^\circ $. The results are shown in Figure 8. It can be seen that E(B − V) and NH are strongly linearly correlated, with a Pearson coefficient of 0.93. The ratio obtained from the linear fit is NH/E(B − V) = (9.4 ± 1.6) × 1021 cm−2 mag−1 (the intercept is also set to be 0), where the quoted uncertainties are the 95% confidence limits estimated from pair bootstrap resampling. This value is a factor of 1.6 higher than that in Bohlin et al. (1978).

Figure 7.

Figure 7. τ353 vs. NH along the 34 purely atomic sightlines described in the text. Gray triangles indicate the upper limits for NH along these 34 atomic sightlines with ${N}_{{{\rm{H}}}_{2}}$ estimated from the 3σ OH detection limits using an abundance ratio NOH/${N}_{{{\rm{H}}}_{2}}$ = 10−7. The thick solid line shows the linear fit to the data in this work, the dotted line shows the conversion factor derived by PLC2014a, and the dashed line shows the conversion factor derived by Fukui et al. (2015). (Note that all these works use the same τ353 map). τ353 error bars are from the uncertainty map of PLC2014a; the shaded region represents the 95% confidence intervals for the linear fit, estimated from pair bootstrap resampling.

Standard image High-resolution image

The value obtained here is consistent with the estimate of Lenz et al. (2017): NH/E(B − V) = 8.8 × 1021 cm−2 mag−1 (no uncertainty is given in that work). These authors compared optically thin H i column density from HI4PI (Collaboration et al. 2016) with various estimates of E(B − V) from Schlegel et al. (1998), Peek & Graves (2010), Schlafly et al. (2014), PLC2014a, and Meisner & Finkbeiner (2015). We note that the estimate of Lenz et al. (2017) is only valid for NH < 4 × 1020 cm−2, where it seems safe to assume that the 21 cm emission is optically thin. Our value is also close to that of Liszt (2014a), who find NH i/E(B − V) = 8.3 × 1021 cm−2 mag−1 (also given without uncertainty) for $| b| \geqslant 20^\circ $ and 0.015 ≲ E(B − V) ≲ 0.075, by comparing H i data from LAB and E(B − V) from Schlegel et al. (1998). The methodology used by these two studies differs in a number of details. For instance, Liszt (2014a) did not apply a gain correction to the Schlegel et al. (1998) map (whereas Lenz et al. 2017 scaled it down by 12%), and did not smooth it to the LAB angular resolution (30'). However, Liszt (2014a) did apply an empirical correction factor to account for H i opacity (albeit one whose effects on high-latitude sightlines was small). These details may account for the difference between the values obtained by these two otherwise similar studies.

We also note that, like the present work, these studies did not take into account the potential contribution of dust associated with the diffuse warm ionized gas (WIM). This would tend to produce a flattening of the E(B − V) versus NH i relation at low NH i and therefore increase the value of NH i/E(B − V) artificially. Because we are able to accurately probe a large column density range (up to 3 × 1021 cm−2), we would naively expect our estimate of NH/E(B − V) to be less affected by WIM bias than either Liszt (2014a) or Lenz et al. (2017; which would tend to have a greater effect on lower column data points). While more work is needed to quantify the contribution of the WIM on dust emission/absorption measurements at low E(B − V), we consider it unlikely to account for the difference between our work and historically lower measurements of the NH/E(B − V) ratio.

Despite minor differences between these three studies, it is clear that they point to a NH i/E(B − V) value of (∼8–9) × 1021 cm−2 mag−1. This is 40%–60% higher than the traditional value of Bohlin et al. (1978), which has been used by most models of interstellar dust as a reference point to set the dust-to-gas ratio (e.g., Draine & Fraisse 2009; Jones et al. 2013). We note that if NH is replaced with upper limits (as discussed in Section 4.1), NH/E(B − V) climbs yet higher, leaving this key conclusion unaffected.

4.3. Disentangling the Effects of Grain Evolution and Dark Gas on σ353

A number of studies have used the correlation between τ353 and NH, particularly with regards to the search for dark gas (e.g., Planck Collaboration et al. 2011; Fukui et al. 2014, 2015; Reach et al. 2015). It is clear that τ353 and NH are in general linearly correlated only if σ353 is a constant. However, it is recognized that σ353 is sensitive to grain evolution, and significant variations in the ratio NH/τ353 have been observed, particularly when transitioning to the high-density, molecular regime (e.g., Planck Collaboration et al. 2014a, 2015; Okamoto et al. 2017; Remy et al. 2017). The origin of observed variations in σ353 may relate to a change in dust properties via κ353, and/or a variation in the dust-to-gas ratio r, but may also include a contribution due to the presence of dark gas, if this is unaccounted for in the estimated NH.

PLC2014a presented the variation in σ353 with NH at 30' resolution over the entire sky. In that work, NH was derived from (${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ + XCO WCO), thus dark gas (both optically thick H i and CO-dark H2) was unaccounted for. We reproduce their data in Figure 9. It can be seen that σ353 is roughly flat and at a minimum in a narrow, low column density range NH =(1−3) × 1020 cm−2, then increases linearly until NH = 15 ×1020 cm−2, by which point it is almost a factor of 2 higher. It then remains approximately constant for the canonical value of XCO = 2.0 × 1020 cm−2 K−1 km−1 s. A key issue for dark gas studies is disentangling how much of the initial rise in σ353 is due to changing grain properties and how much is due to the contribution of unseen material, whether it be opaque H i or diffuse H2. (Note also the upturn in σ353 seen at the lowest NH, which may be due to the presence of unaccounted-for protons in the warm ionized medium.)

Figure 8.

Figure 8. Correlation between NH and dust reddening E(B − V) from Green et al. (2018) along 34 atomic sightlines. Gray triangles indicate the upper limits for NH along these 34 atomic sightlines, with ${N}_{{{\rm{H}}}_{2}}$ estimated from the 3σ OH detection limits using an abundance ratio NOH/${N}_{{{\rm{H}}}_{2}}$ = 10−7. The errorbar on E(B − V) along each sightline is the standard deviation of the 20 Markov Chain realizations of E(B − V) at infinite distance; the shaded region represents the 95% confidence intervals for the linear fit, estimated from pair bootstrap resampling.

Standard image High-resolution image

The column density range probed by our purely atomic sightlines, NH = (1 ∼ 30) × 1020 cm−2 well samples the range where σ353 undergoes its first linear increase. Dark gas is also fully accounted for in our data, since H i is opacity-corrected, and no molecular gas is detected in emission along these sightlines. To quantify the effect of ignoring H i opacity on σ353, we compare σ353 deduced from the true, opacity-corrected NH i with that deduced under the optically thin assumption. The results are shown in Figure 10. In low column density regions (NH < 5 × 1020 cm−2), each σ353 pair from NH i and ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ are comparable. However, at higher column densities (NH > 5 × 1020 cm−2) σ353 from true NH i is systematically lower than that measured from ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$. On average, σ353 obtained from optically thin H i column density increases by ∼1.6 when going from low to high column density regions; whereas σ353 from true NH i increases by ∼1.4. This suggests that if H i opacity is not explicitly corrected for, it can account for around one-third (1/3) of the increase of σ353 observed during the transition from diffuse to dense atomic regimes. The remaining of two-thirds (2/3) must arise due to changes in dust properties.

Figure 9.

Figure 9. Dust opacity σ353 vs. total column density NH along the 34 purely atomic sightlines presented in this work (red points), overlaid on σ353 derived for the whole sky at 30' resolution from PLC2014a. Here, blue points assume an X-factor of XCO = 1.0 × 1020, black assume XCO = 2.0 × 1020, and violet assume XCO = 3.0 × 1020. The gray envelope is the standard deviation of these all-sky measurements for XCO = 2.0 × 1020. The red and black dashed lines show, respectively, the constant σ353 derived from the linear fit in Section 4.1 and that obtained from PLC2014a for the low column density regime.

Standard image High-resolution image

From Equation (11), we see that σ353 is a function of the dust-to-gas mass ratio, r, and the dust emissivity cross-section, κ353, which depends on the composition and structure of dust grains. Given the uncertainties on the efficiency of the physical processes involved in the evolution of interstellar dust grains, it is difficult at this point to conclude if the variations of σ353 observed here are due to an increase of the dust mass (i.e., r) or to a change in the dust emission properties (i.e., κ353). Using the dust model of Jones et al. (2013), Ysard et al. (2015) suggest that most of the variations in the dust emission observed by Planck in the diffuse ISM could be explained by relatively small variations in the dust properties. That interpretation would favor a scenario in which the increase of σ353 from diffuse to denser gas is caused by the growth of thin mantles via the accretion of atoms and molecules from the gas phase. Even though this process would increase the mass of grains (and therefore increase r), the change of the structure of the grain surface would lead to a larger increase in κ353. Alternatively, it is possible that this systematic variation of τ353/NH could be due to residual large-scale systematic effects in the Planck data, or to the fact that the modified blackbody model introduces a bias in the estimate of τ353. Neither of these explanations can be ruled out.

Figure 9 shows σ353 as a function of NH superimposed on the results from PLC2014a. It can be seen that we observe a similar rise in σ353 in the column density range (∼5–30) × 1020 cm−2, but less extreme. In particular, most of our data points in the higher column density range (NH > 5 × 1020 cm−2) are found below the PLC2014a trend, which is derived from the mean values of σ353 over the whole sky in NH bins. This is true even if we use ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ rather than NH i to derive σ353, indicating that optically thick H i alone cannot shift our data points high enough for a perfect match. This is consistent with the fact that we are examining purely atomic sightlines, and likely happens because we are sampling comparatively low number densities (nH ≲ 10–100 cm−3; a mixture of WNM and CNM), whereas the sample in PLC2014a includes molecular gas in the NH bins, presumably with a higher κ353. However, in diffuse regions with NH < 5 ×1020 cm−2, the mean value of σ353 from our sample is comparable with that from PLC2014a.

4.4. E(B − V) as the More Reliable Proxy for NH?

We have seen that along 34 atomic sightlines E(B − V) shows a tight linear correlation with NH in the column density range NH = (1 ∼ 30) × 1020 cm−2. τ353 also shows a good linear relation with NH but with systematic deviations as described above.

Figure 11 replicates Figure 10 but for E(B − V) rather than τ353. Although the sample used here is small, these figures demonstrate clearly that the ratio E(B − V)/NH is more stable than τ353/NH over the range of column densities and sightlines covered by our analysis. In fact, with NH corrected for optical depth effects, our data are compatible with a constant value for E(B − V)/NH, up to NH = 30 × 1020 cm−2. On the other hand, we have observed an increase of τ353/NH with NH, which we suggest may be due to an increase of the dust emissivity (an increase of r and/or κ353 without significantly affecting the dust absorption cross-section). While we are unfortunately unable to follow how these relations evolve at higher AV and in molecular gas, our results nevertheless suggest that the E(B − V) maps of Green et al. (2018) are a more reliable proxy for NH than the current release of Planck τ353 in low-to-moderate column density regimes.

Figure 10.

Figure 10. Dust opacity σ353 vs. total column density NH along the 34 purely atomic sightlines presented in this work using true NH i (red) and ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ (black) as the total gas column density NH. The large data points are the average values for the low-density (NH < 5 × 1020 cm−2) and high-density (NH > 5 ×1020 cm−2) regions (error bars on these points are the standard error of the mean). Note that two data points, one black (σ353 = 27.2 × 10−27 cm2 H−1) and one red (σ353 = 16.1 × 10−27 cm2 H−1), at NH = 18.2 × 1020 cm−2 are not shown, but are included in the averages.

Standard image High-resolution image

5. OH Abundance Ratio XOH

The rotational lines of CO are widely used to probe the physical properties of H2 clouds, but in diffuse molecular regimes where CO is not detectable in emission other species and transitions must be considered as alternative tracers of H2. Among these, the ground-state main lines of OH are a promising dark gas tracer; they are readily detectable in translucent/diffuse molecular clouds (e.g., Magnani & Siskind 1990; Barriault et al. 2010), and since OH is considered to be a precursor molecule necessary for the formation of CO in diffuse regions (Black & Dalgarno 1977; Barriault et al. 2010), it is expected to be abundant in low-CO density/abundance regimes.

The utility of OH as a tracer of CO-dark H2 depends on our ability to constrain the OH/H2 abundance ratio, XOH =NOH/${N}_{{{\rm{H}}}_{2}}$. From an observational perspective, this requires good estimates of both the OH and H2 column densities, the latter of which often cannot be observed directly. Many efforts (both modeling and observational) have been devoted to deriving XOH in different environmental conditions, which we summarize below:

  • 1.  
    Astrochemical models by Black & Dalgarno (1977) found XOH ∼ 10−7 for the case of ζ Ophiuchi cloud.
  • 2.  
    Nineteen comprehensive models of diffuse interstellar clouds with nH from 250 to 1000 cm−3, Tk from 20 to 100 K and AV from 0.62 to 2.12 mag (van Dishoeck & Black 1986) found OH/H2 abundances from 1.6 × 10−8 to 2.9 × 10−7.
  • 3.  
    The OH abundance with respect to H2 from chemical models of diffuse clouds was found to vary from 7.8 × 10−9 to 8.3 × 10−8 with AV = (0.1–1) mag, TK = (50–100) K and n = (50–1000) cm−3 (Viala 1986).
  • 4.  
    Six model calculations (that differ in depletion factors of heavy elements and cosmic-ray ionization rate) by Nercessian et al. (1988) toward molecular gas in front of the star HD 29647 in Taurus found OH/H2 ratios between 5.3 × 10−8 and 2.5 × 10−6.
  • 5.  
    From OH observations toward high-latitude clouds using the 43 m NRAO telescope, Magnani et al. (1988) derived XOH values between 4.8 × 10−7 to 4 × 10−6 in the range of AV = (0.4–1.1) mag, assuming that ${N}_{{{\rm{H}}}_{2}}$ = 9.4 ×1020 AV. However, we note that the excitation temperatures of the OH main lines were assumed to be equal, Tex,1665 = Tex,1667, likely resulting in overestimation of NOH (see Crutcher 1979; Dawson et al. 2014).
  • 6.  
    Andersson & Wannier (1993) obtained an OH abundance of ∼10−7 from models of halos around dark molecular clouds.
  • 7.  
    Combining NOH data from Roueff (1996) and Felenbok & Roueff (1996) with measurements of ${N}_{{{\rm{H}}}_{2}}$ from Savage et al. (1977, using UV absorption), Rachford et al. (2002, using UV absorption) and Joseph et al. (1986, using CO emission), Liszt & Lucas (2002) find XOH = (1.0 ± 0.2) × 10−7 toward diffuse clouds.
  • 8.  
    Weselak et al. (2010) derived OH abundances of (1.05 ± 0.24) × 10−7 from absorption-line observations of five translucent sightlines, with molecular hydrogen column densities ${N}_{{{\rm{H}}}_{2}}$ measured through UV absorption by (Rachford et al. 2002, 2009).
  • 9.  
    Xu et al. (2016) report that XOH decreases from 8 × 10−7 to 1 × 10−7 across a boundary region of the Taurus molecular cloud, over the range AV = 0.4–2.7 mag. ${N}_{{{\rm{H}}}_{2}}$ was obtained from an integration of AV-based estimates of the H2 volume density (assuming ${N}_{{{\rm{H}}}_{2}}$ = 9.4 × 1020 AV).
  • 10.  
    Recently, Rugel et al. (2018) report a median XOH ∼ 1.3 × 10−7 from THOR Survey observations of OH absorption in the first Milky Way quadrant, with ${N}_{{{\rm{H}}}_{2}}$ estimated from 13CO(1-0).

Overall, while model calculations tend to produce some variation in the OH abundance ratio over different parts of parameter space (8 × 10−9–4 × 10−6), observationally determined measurements of XOH cluster fairly tightly around 10−7, with some suggestion that this may decrease for denser sightlines.

In this paper, we determine our own OH abundances, using the MS data set to provide NOH and NH i; then employing τ353 and E(B − V) (along with our own conversion factors) to compute molecular hydrogen column densities as ${N}_{{{\rm{H}}}_{2}}=\tfrac{1}{2}({N}_{{\rm{H}}}-{N}_{{\rm{H}}{\rm{I}}})$. We note that since this dust-based estimate of ${N}_{{{\rm{H}}}_{2}}$ cannot be decomposed in velocity space, the OH abundances are determined in an integrated fashion for each sightline, and not on a component-by-component basis. While CO was detected along all but one sightline, it was not detected toward all velocity components, meaning that our abundances are generally computed for a mixture of CO-dark and CO-bright H2 (for further details, see Li et al. 2018).

The OH column densities derived in Section 3 are derived from direct measurements of Tex and τ. This means that they should be accurate compared to methods that rely on assumptions about these variables (see, e.g., Crutcher 1979; Dawson et al. 2014). In computing NH, we assume that the linear correlations (deduced from τ353, E(B − V) and NH i toward 34 atomic sightlines) still hold in molecular regions. In this manner, estimates of the OH/H2 abundance ratio can be obtained within a range of visual extinction AV = (0.25–4.8) mag. We note that, of our 19 OH-bright sightlines, 5 produce ${N}_{{{\rm{H}}}_{2}}$ that is either negative or consistent with zero to within the measurement uncertainties; these are excluded from the analysis.

Figure 12 shows NOH and XOH as functions of ${N}_{{{\rm{H}}}_{2}}$. We find that NOH increases approximately linearly with ${N}_{{{\rm{H}}}_{2}}$, and the OH/H2 abundance ratio is approximately consistent for the two methods, with no evidence of and systematic trends with increasing column density. Differences arise due to the overestimation of NH derived from τ353 along dense sightlines compared to NH from E(B − V). As discussed in Section 4, σ353 varies by up to a factor of 2 in the range of NH = (1 ∼ 30) × 1020 cm−2, whereas the ratio $\langle {N}_{{\rm{H}}}/E(B-V)\rangle $ is quite constant. The mean and standard deviation of the XOH distribution deduced from E(B − V) is (0.9 ± 0.6) × 10−7, which is close to the canonical value of ∼1 × 10−7, and double the XOH from τ353, (0.5 ± 0.3) ×10 −7. We regard the higher value as more reliable.

Figure 11.

Figure 11. Ratio E(B − V)/NH as a function of NH along the 34 purely atomic sightlines presented in this work, using true NH i (red) and ${N}_{{\rm{H}}\,{\rm{I}}}^{* }$ (black) as the total gas column density NH. The large data points are the average values for the low-density (NH < 5 × 1020 cm−2) and high-density (NH > 5 × 1020 cm−2) regions (error bars for these points are the standard error on the mean).

Standard image High-resolution image
Figure 12.

Figure 12. Left: NOH as a function of ${N}_{{{\rm{H}}}_{2}}$ obtained from the two NH proxies, E(B − V) (blue) and τ353 (red). Right: XOH derived from the two proxies as a function of ${N}_{{{\rm{H}}}_{2}}$.

Standard image High-resolution image

6. Conclusions

We have combined accurate, opacity-corrected H i column densities from the Arecibo Millennium Survey and 21-SPONGE with thermal dust data from the Planck satellite and the new E(B − V) maps of Green et al. (2018). We have also made use of newly published Millennium Survey OH data and information on CO detections from Li et al. (2018). In combination, these data sets allow us to select reliable subsamples of purely atomic (or partially molecular) sightlines, and hence assess the impact of H i opacity on the scaling relations commonly used to convert dust data to total proton column density NH. They also allow us to make new measurements of the OH/H2 abundance ratio, which is essential in interpreting the next generation of OH data sets. Our key conclusions are as follows:

  • 1.  
    H i opacity effects become important above NH i > 5 ×1020 cm−2; below this value the optically thin assumption may usually be considered reliable.
  • 2.  
    Along purely atomic sightlines with NH = NH i = (1–30) × 1020 cm−2, the dust opacity, σ353 = τ353/NH, is ∼40% higher for moderate-to-high column densities than low (defined as above and below NH = 5 × 1020 cm−2). We have argued that this rise is likely due to the evolution of dust grains in the atomic ISM, although large-scale systematics in the Planck data cannot be definitively ruled out. Failure to account for H i opacity can cause an additional apparent rise of the order of ∼20%.
  • 3.  
    For purely atomic sightlines, we measure a NH/E(B − V) ratio of (9.4 ± 1.6) × 1021 cm−2 mag−1. This is consistent with Lenz et al. (2017) and Liszt (2014a), but 60% higher than the canonical value from Bohlin et al. (1978).
  • 4.  
    Our results suggest that NH derived from the E(B − V) map of Green et al. (2018) is more reliable than that obtained from the τ353 map of PLC2014a in low-to-moderate column density regimes.
  • 5.  
    We measure the OH/H2 abundance ratio, XOH, along a sample of 16 molecular sightlines. We find XOH ∼ 1 ×10−7, with no evidence of a systematic trend with column density. Since our sightlines include both CO-dark and CO-bright molecular gas components, this suggests that OH may be used as a reliable proxy for H2 over a broad range of molecular regimes.

J.R.D. is the recipient of an Australian Research Council (ARC) DECRA Fellowship (project number DE170101086). D.L. thanks the supports from the National Key R&D Program of China (2017YFA0402600) and the CAS International Partnership Program (No.114A11KYSB20160008). N.M.-G. acknowledges the support of the ARC through Future Fellowship FT150100024. L.B. acknowledges the support from CONICYT grant PFB06. We are indebted to Professor Mark Wardle for providing us with valuable advice and support. We gratefully acknowledge discussions with Dr. Cormac Purcell and Anita Petzler. Finally, we thank the anonymous referee for comments and criticisms that allowed us to improve the paper.

This research has made use of the NASA/IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.

Footnotes

Please wait… references are loading.
10.3847/1538-4357/aac82b