THE STRUCTURE OF THE CIRCUMGALACTIC MEDIUM OF GALAXIES: COOL ACCRETION INFLOW AROUND NGC 1097*

, , , , , , and

Published 2016 July 20 © 2016. The American Astronomical Society. All rights reserved.
, , Citation David V. Bowen et al 2016 ApJ 826 50 DOI 10.3847/0004-637X/826/1/50

0004-637X/826/1/50

ABSTRACT

We present Hubble Space Telescope far-UV spectra of four QSOs whose sightlines pass through the halo of NGC 1097 at impact parameters of ρ = 48–165 kpc. NGC 1097 is a nearby spiral galaxy that has undergone at least two minor merger events, but no apparent major mergers, and is relatively isolated with respect to other nearby bright galaxies. This makes NGC 1097 a good case study for exploring baryons in a paradigmatic bright-galaxy halo. Lyα absorption is detected along all sightlines and Si iiiλ1206 is found along the three sightlines with the smallest ρ; metal lines of C ii, Si ii, and Si iv are only found with certainty toward the innermost sightline. The kinematics of the absorption lines are best replicated by a model with a disk-like distribution of gas approximately planar to the observed 21 cm H i disk, which is rotating more slowly than the inner disk, and into which gas is infalling from the intergalactic medium. Some part of the absorption toward the innermost sightline may arise either from a small-scale outflow or from tidal debris associated with the minor merger that gives rise to the well known "dog-leg" stellar stream that projects from NGC 1097. When compared to other studies, NGC 1097 appears to be a "typical" absorber, although the large dispersion in absorption line column density and equivalent width in a single halo goes perhaps some way toward explaining the wide range of these values seen in higher-z studies.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

It is now thought that the circumgalactic medium (CGM) of a galaxy represents a genuine transition zone for the recycling of baryons. The term CGM refers to the gas that exists in some sense "outside" the interstellar medium (ISM) of a galaxy, yet "inside" its virial radius, distinct from the surrounding intergalactic medium (IGM; see, e.g., Shull 2014). The idea of the CGM has replaced that of the classic "gaseous halo" or "galactic corona," in large part because of a recognition of the many physical processes that probably compete to sustain it: baryons may be flowing out of galaxies as part of stellar winds or wind from an active galactic nucleus (AGN; Bordoloi et al. 2014a; Rubin et al. 2014, and references therein), or from inflows from the IGM (Martin et al. 2012; Rubin et al. 2012; Ford et al. 2014), or from the interplay between merging galaxies, stripped via tidal interactions (e.g., Bessiere et al. 2012; Rupke & Veilleux 2013); or they may simply be part of in situ cloud formations that arise from thermal instabilities in a halo (Mo & Miralda-Escude 1996; Hobbs et al. 2013, and references therein). These processes drive galactic evolution, so characterizing the CGM around galaxies is essential for understanding how galaxies evolve.

One way to study galaxy CGMs is through the absorption lines they cause in the spectra of background objects. The method of detecting QSO absorption line systems (QSOALSs) in background QSO spectra, and then searching for galaxies at the same redshift as the absorption, has been in practice for three decades. Since the launch of the Hubble Space Telescope (HST), QSOALSs have been found in the ultraviolet (UV) at very low redshifts; in particular, the installation of the Cosmic Origins Spectrograph (COS, Green et al. 2012) aboard HST has enabled the production of several extensive surveys that have targeted the CGMs of specific samples of relatively low-z galaxies, including, for example, "COS-Halos" (Tumlinson et al. 2013; Werk et al. 2013), "COS-Dwarfs" (Bordoloi et al. 2014b), "COS-GASS" (Borthakur et al. 2015), as well as several unnamed surveys (e.g., Borthakur et al. 2013; Stocke et al. 2013; Burchett et al. 2015, 2016). If the goal of a study is to understand the nature of the CGM around galaxies as a function of their properties, then surveys that look for absorption lines from selected galaxies are perhaps the most valuable, because selecting galaxies prior to searching for any absorption reduces the bias in only finding CGMs around particular types of galaxies or in particular environments.

All these surveys study the statistical properties of absorbing clouds that arise from single sightlines through single galaxies. Conclusions are drawn about the CGMs of galaxies assuming that galaxies are sufficiently similar that probing many galaxies with single sightlines is the same as probing a single galaxy with many sightlines. Or at least—as larger samples of galaxy–absorber pairs have become available, and a more nuanced approach has been possible—that galaxies with similar characteristics (the same luminosity, halo mass, type, star formation rates, etc.) will have similar enough CGMs that the properties of the absorbing clouds—if really linked to these properties—will become apparent.

Is it valid to assume that similar galaxies can have the same CGMs? Given two identical galaxies, can their properties dominate the state of the baryons within Rv so directly that they end up having highly similar CGMs? Or are the processes that manufacture the CGM inherently too chaotic to lead to reproducible structures? Indeed, can QSOALSs provide any unique signature of a CGM's origin at all?

One way to test these assumptions is to probe the CGM of individual galaxies along multiple sightlines. Absorption by the CGM of one galaxy should depend on far fewer variables in such experiments because, obviously, galaxy properties are fixed for a single galaxy. Such studies ought to be able to more easily expose which property of a galaxy a CGM most depends on.

There are several ways to probe galaxies along more than a single sightline. Multiple sightlines have been observed from either closely separated QSO sightlines (Bechtold et al. 1994; Fang et al. 1996; Dinshaw et al. 1997, 1998; D'Odorico et al. 1998) or lines of sight toward the multiple images of lensed QSOs (Smette et al. 1992; Monier et al. 1998; Petry et al. 1998; Rauch et al. 2001). In many cases, such investigations have focused on determining the sizes of absorbing clouds when lines were detected (or not) in common along the sightlines, although the use of these background sources also provided examples of rarer damped Lyα (DLA) systems (Smette et al. 1995; Zuo et al. 1997; Lopez et al. 1999; Kobayashi et al. 2002; Churchill et al. 2003; Ellison et al. 2004; Cooke et al. 2010), the detection of which might be expected for lensed QSOs when the lensing object is predicted to be relatively close to the sightlines. For these latter absorbers, multiple sightlines have now been used to provide evidence for changes in gas-phase metallicities on scales of a few kiloparsecs (Lopez et al. 2005).

For lensed QSOs, lensing galaxies have been at z ≳ 1, and the absorbing galaxy has been undetected (even if a lens redshift is inferred from the configuration of the lensed images), and the galaxy's properties remain unknown. There are only a small number of examples where both multiple background sightlines have been observed at a high enough resolution to show absorption systems and a galaxy has been confirmed spectroscopically to be at the same redshift as an absorber (Rauch et al. 2002; Rogerson & Hall 2012; Chen et al. 2014; Muzahid 2014; Zahedy et al. 2016). Explanations for the observed absorption have covered the entire gamut of CGM origins, from ancient, as well as recent, outflows, to tidal debris, inflowing streams and stripped gas. Even with these studies then, there is no clear agreement on a common origin for absorption from lensing galaxies.

An alternative method for studying multiple sightlines through a galaxy (and the one adopted in this paper) is to select nearby galaxies for study. Galaxies at very low redshifts have large angular footprints on the sky, which make it easier to identify multiple QSO sightlines that pass through the galaxy at interesting impact parameters. The main disadvantage of working at low redshifts is that most of the absorption lines of interest lie in the UV, and can only be reached using HST. Not only is access to the satellite limited due to high demand, but the telescope's modest mirror size restricts spectroscopic observations to only the brightest objects. Nevertheless, it was always expected prior to its launch that HST would probe galaxy halos along more than single sightlines (e.g., Monk et al. 1986), and indeed some attempts were made with the first generation of HST spectrographs (e.g., Norman et al. 1996; Bowen et al. 1997). Significant progress has been made in the last few years, however, as COS has enabled observations of the fainter QSOs that constitute a higher surface density on the sky.

Keeney et al. (2013) first made use of the enhanced throughput of COS by observing three QSO sightlines 74–172 kpc from the edge-on spiral ESO 157–49 at cz = 1673 km s−1; they detected absorption from Lyα, and both low- and high-ion species, and suggested that some of the absorbing clouds are the remnants of recycled galactic fountain gas. Perhaps a more obvious set of galaxies to probe with multiple QSOs, however, might be M31 and M33. Being the closest bright galaxies to the Milky Way, they have very large angular diameters on the sky, and finding a high number of background probes at interesting impact parameters is straightforward. The problem with both these galaxies is that their velocities are low. Maps of 21 cm emission show that the gas with high H i column density in the disk of M33 extends from ∼−60 to −300 km s−1 (e.g., Putman et al. 2009), and from ∼−50 to −600 km s−1 for M31 (e.g., Corbelli et al. 2010) (the latter having an angular diameter on the sky more than twice that of M33). The complexity of the H i on the largest scales between M31, M33, and the Magellanic Stream7 , and the difficulties in separating out Galactic, Local Group, and extragalactic H i can be seen in the maps of Braun & Thilker (2004). These low velocities make it difficult to separate any absorption from the galaxies with absorption from high-velocity clouds (HVCs) in the Milky Way (MW), particularly Complexes H & G (e.g., Figure 1 of Tripp & Song 2012); they also make it impossible to observe any extragalactic Lyα absorption, which is lost in the saturated Lyα profile of the Galaxy. Nevertheless, attempts to utilize the many QSOs behind both galaxies have been made recently by Rao et al. (2013) and Lehner et al. (2014). We return to these studies later (see Section 8.1).

In this paper, we investigate absorption—including that of Lyα—by multiple QSO sightlines that pass through the halo of NGC 1097, a galaxy that lies well away from the Local Group. Our paper is divided into several sections. In Section 2 we summarize the properties of NGC 1097, its local environment, and the large-scale structure (LSS) in which it is found. In Section 3 we briefly discuss some of the work that led to the selection of NGC 1097 and the background QSOs for study. In Section 4 we describe the HST observations and the reduction of the data, and in Section 5 we analyze in detail the characteristics of the absorption lines detected. Section 6 outlines the results of photoionization modeling of the results, although full details are presented in Appendix C. In the remaining sections we discuss the interpretation of our results. We consider possible configurations of the CGM of NGC 1097 based on the kinematics of the absorbing gas in Section 7, examine the statistical properties of the detected absorption in Section 8, and compare our results to other studies of absorbing galaxies in order to determine whether NGC 1097 is a "typical" absorbing galaxy in comparison to the growing set of higher-z absorbing galaxies (Section 8.1). We discuss our results in Section 9, where we suggest the most likely origin for the absorption (Section 9.1), and briefly compare our data with the results from numerical simulations (Section 9.2). We summarize our results in Section 10.

Throughout this paper we adopt the cosmology H0 = 73 km s−1 Mpc−1, Ωm = 0.27, and ΩΛ = 0.73, when required.

2. NGC 1097 AND ITS ENVIRONMENT

2.1. Properties of NGC 1097

NGC 1097 is an SB(s)b galaxy (Figure 1) and hosts a Type 1 AGN that is surrounded by a ∼700 pc ring of ongoing star formation (e.g., Barth et al. 1995; Prieto et al. 2005). A global systemic velocity for the galaxy can be taken from 21 cm emission line measurements, the two most pertinent H i maps having been made by Ondrechen et al. (1989) and Higdon & Wallin (2003). Ondrechen et al. cited a heliocentric systemic velocity of v = 1280 ± 10 km s−1, but Higdon & Wallin did not derive v from their data. Since these maps were produced, the Bright Galaxy Catalog of the HI Parkes All-Sky Survey (HIPASS) single-dish survey has listed v = 1271 ± 3 km s−1 (Koribalski et al. 2004).

Figure 1.

Figure 1. The field of NGC 1097 (cz = 1270 km s−1), showing its stellar tidal streams. Left: part of a multicolor (22 hr total) composite image of the galaxy taken with a Ceravolo 30 cm telescope and Apogee U16M camera by Rick Stevenson in Queensland, Australia in 2013. The two most prominent streams, toward the northeast, are easily seen: the well-known "dog-leg" stream is to the left, with its right-angle kink, while the stream to the right appears to end in a bulbous cloud of stars. To the southwest, the leftmost stream is fainter and shorter than the northern streams, but is clearly visible. The southern stream to the right is very faint (∼27.6 mag arcsec−2) but may be the longest of the four, and perhaps wider and more amorphous than the other three streams. The streams themselves are likely the stellar remains of a low-mass galaxy that has passed through the disk of NGC 1097 (Higdon & Wallin 2003; Amorisco et al. 2015). NGC 1097's companion galaxies are also marked—NGC 1097A (cz = 1368 km s−1) and NGC 1097B (see Appendix B). The position of one of the QSOs discussed in this paper, Q0244−303, is marked, but the remaining sightlines are outside the field of view (see instead Figure 3). Right: portion of a 400 s V-band FORS2 image extracted from the ESO Archive (originally published by Galianni et al. 2010). The dog-leg stellar trail is shown more clearly in relation to the position of Q0244−303, and the only part of the stream with a spectroscopic redshift, Knot-A, is indicated.

Standard image High-resolution image

Converting the heliocentric systemic velocity to a Hubble flow distance is not necessarily correct, since corrections for local group motions and motions of NGC 1097 toward the Fornax cluster (see Section 2.3) are not accounted for. The difference between v and the cosmic microwave background is 166 km s−1, giving a corrected velocity of 1105 ± 12 km s−1 if the HIPASS velocity is used, and a corresponding distance of 15.1 ± 1.1 Mpc, which we adopt in this paper. This gives a scale of 4.4 kpc arcmin−1 on the sky. This distance agrees well with a distance of 14.2 ± 2.4 Mpc given by Tully et al. (2008) based on a distance modulus derived from the relationship of Tully & Fisher (1977) between 21 cm linewidth and galaxy luminosity. The apparent magnitude of NGC 1097 is bJ = 10.1 (Doyle et al. 2005), which would translate to an absolute magnitude of M*(bJ) = −20.8 for D = 15.1 Mpc. If an L* galaxy has an absolute magnitude of M*(bJ) = −20.4 (Norberg et al. 2002), then L/L* = 1.4 for NGC 1097.

We calculate the virial radius Rv for NGC 1097 as follows. With the stellar mass of the galaxy known (Table 1), the dark matter (DM) mass can be calculated using the correlations between stellar mass and halo mass given by Behroozi et al. (2013): for log M* = 10.5 ± 0.1, log Mh(M) = 12.1. The error in converting between stellar mass and halo mass is of the same order as the error in the stellar mass itself, ∼0.2 dex. This halo mass is similar to log Mh(M) = 11.8 found by Higdon & Wallin (2003) when fitting a simple disk+halo model to the observed 21 cm rotation curve. Using the approximations given by Maller & Bullock (2004), the virial halo radius is Rv ≃ 280 kpc. Deviations of order 0.2 dex in stellar mass would change this value by ±40 kpc, so the value of Rv is not known precisely. The value of Rv = 280 kpc agrees with the conversion between galaxy luminosity and Rv, Mh given by Stocke et al. (2013, their Figure 1), providing their halo abundance matching curves are used; their "adopted minimum mass-to-light ratio" gives an Rv that is 30% smaller. For a halo with this mass, the corresponding virial velocity is 140 km s−1 and the virial temperature is 6.9 × 105 K.

Table 1.  Parameters for NGC 1097

    References
R.A., decl. 02:46:18.95, −30:16:28.8 1
Heliocentric velocity, v 1271 ± 10 km s−1 2
Hubble flow velocity 1105 ± 12 km s−1 3
Inclination, position angle of major axis 40° ± 5°, 130° 4
Mag bJ, M(bJ), L 10.1, −20.8, 1.4 L* 5
Diameter D(0)25(B) 9farcm55 ≡ 42 kpc 6
H i massa 7.7 × 109 M 2
Star formation ratea 5 ± 1 M yr−1 7
Stellar mass [log(M)]a 10.5 ± 0.1 8
Total mass [log(M)] ≃12.0 ± 0.2 9
Adopted distance, D 15.1 ± 1.1 Mpc 10
Virial radius, Rvir 280 kpc 10

Note.

aPublished values corrected for D = 15.1 Mpc.

References. (1) Position of nucleus in HST WFPC2 F218W image from data in the Hubble Legacy Archive. (2) From 21 cm emission line data of Koribalski et al. (2004). (3) Heliocentric velocity converted to velocity of the cosmic microwave background using the conversion given by the NASA/IPAC Extragalactic Database. (4) Higdon & Wallin (2003). (5) Doyle et al. (2005). (6) de Vaucouleurs et al. (1991, p. 1) (RC3). (7) Calzetti et al. (2010). (8) Skibba et al. (2011). (9) Converted from stellar mass—see Section 2.1. (10) See Section 2.1.

Download table as:  ASCIITypeset image

These, and other relevant physical parameters of NGC 1097 (some of which are discussed later in this paper) are listed in Table 1.

2.2. Local Environment

At the outskirts of the galaxy, NGC 1097 shows clear signs of minor merger events—one ongoing, the other occurring in the more distant past. The most obvious feature is the presence of NGC 1097A (see Figure 1), an elliptical galaxy that has a velocity of 1365 ± 45 km s−1, and which appears to be interacting with (and disrupting) the outer northwest spiral arm of the parent galaxy. The interaction produces two additional H i arms to the south of the galaxy, well beyond the optical disk (see Figure 4(b) of Higdon & Wallin 2003, and the outer contour of the 21 cm emission outlined in Figure 3). The galaxy has a magnitude of mB(0) = 13.6 or MB = −17.3 at a distance of 15.1 Mpc, largely equivalent to the Large Magellanic Cloud (LMC) near our own Galaxy.

More strikingly, NGC 1097 also shows four optical filaments or streams (Wolstencroft & Zealey 1975; Arp 1976; Lorre 1978). The structures show no optical, X-ray, or radio emission (Wehrle et al. 1997, and references therein). All four streams can be seen in the left panel of Figure 1. The filament oriented to the northeast (NE) is remarkable for showing an acute right-angle bend, producing an L-shaped structure that has become known as the "dog-leg."

In discussing the possible origin of the streams, Higdon & Wallin (2003) suggested that they were caused by the minor merger with a low-mass dwarf disk galaxy, with the lack of any detectable H i and H ii emission due to ram-pressure stripping of the dwarf's gas by the disk of NGC 1097. Higdon & Wallin were able to show that X-shaped tidal streams—including features resembling the dog-leg seen toward NGC 1097—could be reproduced in N-body simulations of a dwarf disk galaxy with a mass of only 0.1% of the host galaxy mass (so in this case, M* ∼ 6 × 108 M) passing through a disk-dominated potential. The streams are, therefore, most likely the stellar remains of a disrupted galaxy, and are not caused by outflowing gas from the center of NGC 1097. More recently, Amorisco et al. (2015) modeled the streams using the infall of a disky dwarf galaxy that has passed its pericenter three times, and which has lost two dex in mass. They too were able to reproduce the right-angle kink seen in the dog-leg stream.

The interpretation that the streams are composed mainly of stars is consistent with spectra of two regions of the dog-leg obtained by Galianni et al. (2010). Within the dog-leg stream itself, there are several "knots" of stars. The only feature with a confirmed redshift is Knot-A, which Galianni et al. measured to be cz = 1259 ± 30 km s−1, and which likely gives a redshift for the stream at that position (Figure 1, right panel). Galianni et al. also observed a second source near the right angle of the trail (their "Knot-B," see Figure 1), but were unable to derive a redshift; for both sources, no emission lines were detected, as might be expected if the streams were actually hot outflows from the center of the galaxy. Moreover, the spectral energy distributions (SEDs) of the knots are inconsistent with thermal or synchrotron emission (Higdon & Wallin 2003).

2.3. Global Environment

In order to examine how the CGM of NGC 1097 might be related to (and influenced by) the wider IGM, we would like to know the galactic structures in which the galaxy is found. To examine the relationship of NGC 1097 to its environment, we extracted all galaxies with known redshifts within ±20° of the galaxy from the NASA/IPAC Extragalactic Database (NED). If we naively assume that the difference in velocity between the surrounding galaxies and NGC 1097 in the line-of-sight direction gives a distance of Δv/H0, then the nearest galaxies that have luminosities ≥0.8 times that of NGC 1097 lie ∼4–5 Mpc away (NGC 908, 1371, 1399, and 1398, and 6dF J0342193−352334).

NED's listing is not a magnitude-limited sample (there is currently no magnitude-limited survey covering this wide a field around NGC 1097) but the collation is useful in showing the proximity of NGC 1097 to the Fornax Cluster, which lies 12° away. The galaxy surface density of Fornax reaches the background at only ∼6° (Ferguson & Sandage 1989) from its center, which almost certainly places NGC 1097 outside the cluster. The left-hand panel of Figure 2 shows a roughly linear distribution of galaxies running North–South on the sky, defining the filamentary structure containing Fornax cluster galaxies, and the color coding in the figure highlights the filament's velocity gradient (Waugh et al. 2002). NGC 1097 appears to lie close to, but not obviously within, the filament; not only does its distance from the central concentration of galaxies suggest that it lies away from the filament, but its velocity is several hundred km s−1 lower than that of the majority of the background galaxies in that direction. If the distance from us to the center of Fornax is 19–20 Mpc (e.g., Madore et al. 1999; Dunn & Jerjen 2006), then NGC 1097 is ∼5–6 Mpc from its center.

Figure 2.

Figure 2. Distribution of galaxies within ±400 km s−1 of NGC 1097. The position of NGC 1097 is at the center of each plot (dotted lines). Left: all galaxies with redshifts cataloged by the NED are plotted. Galaxies that have redshifts greater than NGC 1097 are shown as small, red circles; those with redshifts less than NGC 1097 have larger, bluer circles. The magnitudes of the velocity differences, as a function of the color and size of the circles, are shown in the legend on the extreme right of the figure. The largest, bluest circles are for galaxies with the highest blueshifts relative to NGC 1097, while the smallest, reddest circles have the highest redshift relative to NGC 1097. Right: same as left panel, but only galaxies with luminosities ≳0.1 L* are plotted. The five red circles filled with green are the five closest galaxies with luminosities greater than 0.8 times that of NGC 1097, assuming that their radial redshifts really represent Hubble flow distances. The labelled (unfilled) green circle in each panel shows a 4° radius from the center of the Fornax cluster, which corresponds to ≈1.4 Mpc at the distance of Fornax.

Standard image High-resolution image

The right-hand panel of Figure 2 shows the same distribution of galaxies, but only after selecting galaxies with L > 0.1 L*. The magnitudes of galaxies listed by the NED are inhomogeneous in their filter selection, and the conversion of those magnitudes to a luminosity is approximate. The five closest galaxies mentioned above that have comparable luminosities to that of NGC 1097 are highlighted as circles with green centers and are all over 4–5 Mpc away. The panel demonstrates that there are far fewer bright galaxies near NGC 1097 than the left-hand panel suggests, and that there are no companion galaxies to NGC 1097 with comparable luminosity. Figure 2 therefore suggests that NGC 1097 is a relatively isolated galaxy, probably lying at the edge of a LSS filament.

While it is possible that NGC 1097 may be moving toward the cluster, there is no evidence that it has experienced any recent tidal stripping from any intracluster medium. In this sense, the CGM of the galaxy has most recently been defined primarily by the local interactions discussed in the previous section.

3. QSO TARGET SELECTION

Over the last few years we have been searching the fields of nearby galaxies to find those with multiple QSOs behind them. We began by compiling a list of the top 250 low-redshift galaxies with the largest angular diameters on the sky, and with velocities ≥700 km s−1 to ensure that Lyα absorption would be well away from the geocoronal Lyα emission line (see Section 4.1.1) in COS data. For each galaxy, we used available Galaxy Evolution Explorer (GALEX) satellite catalogs to collate all sources within 200 kpc that had a measured far-UV (FUV) flux greater than 40 μJy (Fλ = 5 × 10−16 erg cm−2 s−1 Å−1, or m(FUV) = 19.9). An additional cut was made to remove objects with GALEX FUV widths greater than 15 arcsec FWHM, in order to avoid extended sources that would pass only a small fraction of their cataloged flux through the small COS aperture.

To corroborate which of the UV sources were extragalactic, and not Milky Way stars, the GALEX-generated lists of FUV-bright objects were cross-correlated with the latest release of the SDSS redshift catalogs, the NED, and the SIMBAD database. The galaxies were then ranked simply by the surface density of the background probes. NGC 1097 was one of a small number of galaxies that had more than five sources satisfying the criteria described above and became one of the potential targets for HST follow-up.

In addition, NGC 1097 was observed as part of a GALEX Cycle 1 program that we designed to obtain FUV grism spectra of UV-bright objects in nearby galaxy fields. Unfortunately, we did not discover any new QSOs near NGC 1097, but did recover many of the objects already known. Further details on the original discoveries of the QSOs selected for our HST study are given in Appendix A along with notes on our GALEX grism spectra.

Table 2 lists the QSOs observed in this paper; in order to provide a more succinct nomenclature for the objects, we abbreviate the QSO names to the simpler forms of Q0244, S393Z082, Q0246, and HE0241 throughout the text. A fifth object, 2dFGRS S394Z150, believed to be an AGN at z = 0.131, was observed as part of our program. Its identification proved to be incorrect, however—the object is likely a white dwarf (WD) star instead. Additional details are given in Appendix A, but this object is listed at the end of Table 2 because it was used in the analysis of the other spectra, discussed below.

Table 2.  Journal of Observationsa

                  Impact Parameters
QSO R.A., decl. Galactic B(J)b FUVc   Observation Time Observed ρ ρ  
Name (J2000.0) l, b (mag) (μJy) zem Date (minutes) Fluxd (arcmin) (kpc)e ρ/Rvf
Q0244−303 02:46:49.87, 226.573, 18.4 126 0.53 2013−08−09 204 1.14 11.0 48.2 0.17
  −30:07:41.3 −64.569                  
2dFGRS 02:45:00.77, 226.556, 19.1 57 0.34 2013−05−27 294 0.70 19.2 84.2 0.30
S393Z082 −30:07:22.4 −64.962                  
Q0246−308 02:48:22.00, 227.747, 17.3 97 1.09 2013−08−06 245 1.28 34.2 149.9 0.54
  −30:38:06.8 −64.238                  
HE0241−3043 02:43:37.66, 227.488, 17.2 155 0.67 2013−06−21 116 3.90 37.6 164.8 0.59
  −30:30:48.1 −65.259                  
2dFGRS 02:46:13.34, 227.431 19.2 63 0.0g 2012−10−17/20 308 1.0h 13.3 ... ...
S394Z150 −30:29:42.2 −64.700                  

Notes.

aAll data were taken using COS and the G130M grating centered at 1291 Å and 1327 Å, and the PSA aperture. bOptical magnitudes are B(J) taken from the superCOSMOS Sky Survey web server or 2dFGRS database. cGALEX FUV flux. dFlux at 1240 Å in units of 10−15 erg cm−2 s−1 Å−1. eAssuming a distance of 15.1 Mpc to NGC 1097. fAssuming a virial radius Rv of 280 kpc. gThe object was expected to have zem = 0.131, but turned out to be a WD star—see Appendix A. hFlux at 1240 Å is zero due to strong Lyα absorption; the stated value is the flux in the continuum at 1350 Å.

Download table as:  ASCIITypeset image

4. COS OBSERVATIONS AND DATA REDUCTION

4.1. Co-addition of Sub-exposures

Our GO program (12988) observations were made with COS using the G130M grating and the Primary Science Aperture (PSA). A journal of the observations is given in Table 2. Data were taken in TIME-TAG mode. Two different grating positions were used, centered at 1291 and 1327 Å, to provide (after co-adding all the exposures) some data in the gap between the two segments of the photon-counting microchannel plate detector (Holland et al. 2012). Each sightline was observed over three to five orbits per visit, with all of the four available FP-POS offsets used to reduce fixed-pattern noise.

While the sub-exposures were wavelength-calibrated by the pipeline, their wavelengths disagreed with each other in pixel space, because of shifts in the dispersed light across the detector, caused by the difficulty of returning the grating to exactly the same position after any movement. In order to co-add spectra for a given sightline, we interpolated the flux arrays onto a common wavelength grid at a dispersion of 0.01 Å pix−1, close to the initial value from the pipeline. To account for wavelength shifts in individual exposures that were not corrected for by the pipeline calibration, we compared positions of the strongest lines visible in each exposure, and measured the fractional pixel shifts needed to align the features. Any offsets found were added as zero-point shifts to the wavelength arrays.

To co-add data at wavelengths obtained by only some of the exposures, and to account for exposures with unequal exposure times, we calculated a weighted average from the COUNT-RATE arrays in the x1d files, with weights formed from the inverse of a spectrum's variance. We propagated an error array σi at each pixel i from the gross counts n of a spectrum using the lower Poisson confidence levels given by Gehrels (1986). Each error array was smoothed by a 10 pixel boxcar average to eliminate effects from fluctuations in individual pixels. The average count-rate of all the exposures was formed by weighting each sub-exposure by the inverse of the variance array.

The long and short segments were co-added using the same weighting scheme described above. The final spectrum for each sightline was rebinned by a factor of 3 to 0.03 Å pix−1 (≡7 km s−1 pix−1 at 1220 Å) to better sample the data given the resolution of COS. The final spectra had signal-to-noise ratios (S/N) of 6–8 per rebinned pixel at 1220 Å, and 8–10 pix−1 at 1340 Å.

4.1.1. Background Contribution from Lyα Airglow

Searching for Lyα absorption lines at the redshifts of nearby galaxies is hampered by two problems. First, at redshifts close to zero, Lyα lines are unobservable because the QSO flux is completely extinguished by the DLA absorption from our own Milky Way. Second, geocoronal Lyα emission filling the 2.5 arcsec diameter circular COS PSA will contribute additional flux at the wavelengths of the expected extragalactic absorption. This would make lines seem weaker than they really are, and lead to an underestimate of H i column densities. The Lyα airglow is distributed over a wide spectral range, and when observing objects with low count rates, the emission can be a significant additional background source near the rest wavelength of Lyα.

Figure 3.

Figure 3. Positions of QSO sightlines observed in this paper relative to NGC 1097. The image is constructed from portions of the STScI Digitized Sky Survey data, aligned with each other to reproduce the correct angular distances on the sky. QSOs observed in our COS program are labelled with circles; we also show the position of S394Z150 even though our COS data found it to be a white dwarf. The green dotted line around NGC 1097 marks its optical diameter D(0)25, while the thick blue line shows the maximum extent of the galaxy detected by Higdon & Wallin (2003) at 21 cm, at a level of ≃1.3 × 1019 cm−2. If the virial radius Rv of NGC 1097 is 280 kpc then the furthest sightline from the galaxy, toward HE 0241–3043, is at an impact parameter of ≃0.6 Rv in this figure.

Standard image High-resolution image

Portions of the co-added spectra near rest-frame Lyα are shown in Figure 4. At the lowest velocities, geocoronal Lyα emission dominates. The spectrum of Q0244 shows the widest Lyα absorption from NGC 1097 (the extent of which is indicated by orange lines), but there is no flux at the center of the absorption line profile, suggesting that, at least by a redshift of cz = 1270 km s−1, Lyα airglow does not contribute to the spectrum's flux. Additional information comes from the observation of the WD star, S394Z150, where Lyα absorption in the star's atmosphere is so strong that there is no flux over a wavelength range much wider than that which is usually extinguished from interstellar Lyα in the Milky Way. The spectrum of the WD is shown in the bottom panel of Figure 4, where only geocoronal Lyα emission is left at the bottom of the DLA trough (black line). In our sample, the strongest geocoronal Lyα was observed toward Q0244, and Figure 4 shows the extent of the emission toward the WD after the peak of the Lyα emission has been scaled to that of Q0244 (red line). The figure shows that by ≃800 km s−1, any contribution from airglow is negligible and any Lyα absorption lines at those velocities would be unaffected.

Figure 4.

Figure 4. The top four panels show COS spectra for the QSOs whose sightlines pass through the halo of NGC 1097. The orange dashed regions shows the approximate extent of Lyα absorption from NGC 1097. The bottom panel shows the spectrum of the WD star S394Z150 that was also observed in our program. The black line shows the measured normalized flux, while the red line shows the same flux (binned by a factor of 2 for clarity) scaled (by a factor of 5.6) such that the geocoronal Lyα emission line would have the same flux as the line seen in the spectrum of Q0244, which had the highest Lyα emission line flux of the four QSOs. This panel shows that at velocities similar to that of NGC 1097, emission from the wings of Lyα geocoronal emission adds no additional background counts to the observed Lyα absorption lines seen from NGC 1097.

Standard image High-resolution image

4.1.2. Final Wavelength Accuracy

To provide a precise zero-point to the wavelength calibration, we tied the wavelengths of features in the COS data to the velocity of 21 cm emission measured toward the QSOs from H i profiles extracted from the Parkes Galactic All-Sky Survey (GASS, McClure-Griffiths et al. 2009). The median heliocentric velocities of the 21 cm H i emission along all four QSO sightlines were nearly identical, +9 km s−1. We matched this velocity to those of the strongest low-ionization ISM lines seen in the COS spectra, which are expected to arise in the same gas detected at 21 cm, namely Si iiλλ1190, 1193, the S iiλλ1250, 1253, 1259 triplet, Si iiλ1260, and C iiλ1334. In fact, the velocities of these lines did not all agree; in collating line centroids for all four sightlines, differences of 0–12 km s−1 (∼0–2 rebinned pixels) were found between the different ions. There are several possible reasons for this: line centers may be measured incorrectly because of low S/N in the line profile; weaker components of the strongest lines may have different profiles compared to weaker lines where such components are not detected; or errors may arise in the wavelength solution applied to the detectors, or to x-walk along the detector in regions where the gain-sag is high. Without knowing which of these effects dominate, we decided to match the 21 cm emission velocity of +9 km s−1 to the average velocity of all seven MW lines; as a consequence, between 1190 and 1334 Å, the error in measuring the wavelength of any narrow feature ought to be ∼±6 km s−1 or about one rebinned pixel. This wavelength range covers all the lines we expected to see from NGC 1097, except the Si ivλλ1393, 1402 doublet.

5. RESULTS

5.1. Absorption Line Measurements

In the final co-added COS spectra, we searched for absorption from NGC 1097 from a variety of species, which are listed in Tables 36. In this section, we describe the analysis of detected absorption lines, while in Section 5.2 we discuss how limits to column densities were derived when no absorption was detected. Much of this analysis follows the methods described by Bowen et al. (2008). In particular, the way in which the data were normalized, the procedures used to measure velocities v, column densities N, and Doppler parameters b, of the absorbing gas by generating theoretical Voigt line profiles and fitting these to the data, and the use of Monte Carlo (MC) simulations to account for errors in the fits arising from Poisson statistics, were identical to those described by Bowen et al. (2008) and are not repeated here.

Table 3.  Equivalent Widths and Column Densities Toward v20, Q0244

  v Δv σ(v) Wb bb σ(b)T    
IDa (km s−1) (km s−1) (km s−1) (Å) (km s−1) (km s−1) log(N) σ(N)T
(1) (2) (3) (4) (5) (6) (7) (8) (9)
        H i Lyα
 
A1 1227 −44 11 $\left.\begin{array}{c}{}^{-}| \\ 1.316\\ {}_{-}| \end{array}\right\}$ 62.7 −10.2, +8.7 15.57 −0.23, +0.55
A2 1261 −10 5 No unique fit
A3 1384 +113 2 18.9 ±2.0 15.07 −0.32, +0.24
A4 1456 +185 2 0.035 ≤9 ... ≥13.10 ...
A5 1491 +220 3 0.045 ≤9 ... ≥13.10 ...
Total [p]       1.385 + 0.035 – 0.058     ≥15.69c
Total [N = 90]       1.365 + 0.095 – 0.109      
        C iiλ1334
       
A1 1227 −44 11 ... (52) ... <13.9d ...
A2 1261 −10 5 0.062 10.9 −4.7, +5.3 13.62 −0.08,+0.12
A3 1384 +113 2 0.200 24.4 ±3.0 14.22 ±0.03
A4 1456 +185 2 <0.044 (9) ... <13.46 ...
A5 1491 +220 3 <0.046 (9) ... <13.48 ...
Total [p]       0.261 ± 0.027     14.31 ±0.03
Total [N = 60]       0.266 ± 0.064        
        Si ii e
 
A1 1227 −44 11 <0.028 (52) ... <12.24 ...
A2 1261 −10 5 <0.028 (11) ... <12.29 ...
A3 1384 +113 2 0.124 16.2 ±2.4 13.09 ±0.05
A4 1456 +185 2 <0.028 (9) ... <12.30 ...
A5 1491 +220 3 <0.028 (9) ... <12.30 ...
Total [p]       0.124 ± 0.013     13.09 ±0.05
Total [N = 30]       0.118 ± 0.020        
        Si iiiλ1206
 
A1 1227 −44 11 0.204 51.9 −10.3, +11.4 13.08 ±0.14
A2 1261 −10 5 ... (11) ... <12.2f ...
A3 1384 +113 2 0.183 24.9 ±5.0 13.16 ±0.07
A4 1456 +185 2 <0.090 (9) ... <13.01 ...
A5 1491 +220 3 <0.090 (9) ... <13.01 ...
Total [p]       0.387 ± 0.084     13.42 ±0.03
Total [N = 60]       0.380 + 0.134 – 0.162        
        Si ivλ1393g
 
A1 1227 −44 11 <0.034 (52) ... <12.60 ...
A2 1261 −10 5 <0.032 (11) ... <12.62 ...
A3 1384 +113 2 0.063 30.2 −9.2, +19.2 12.89 ±0.13
A4 1456 +185 2 <0.032 (9) ... <12.64 ...
A5 1491 +220 3 <0.032 (9) ... <12.64 ...
Total [p]       0.063 ± 0.020     12.89 ±0.13
Total [N = 30]       0.058 ± 0.024        
        N vλ1238
 
A1 1227 −44 11 <0.028 (52) ... <13.13 ...
A2 1261 −10 5 <0.028 (11) ... <13.18 ...
A3 1384 +113 2 <0.030 (30) ... <13.17 ...
A4 1456 +185 2 <0.030 (9) ... <13.23 ...
A5 1491 +220 3 <0.030 (9) ... <13.23 ...

Notes.

a"Total [p]" in this column refers to the total equivalent width (EW) and total column densities calculated from the blended theoretical line profiles used to derive N and b values for individual components; "total [N = ]" refers to the total EW measured over N pixels (see Section 5.1). bEW limits are 2σ(W)T values measured over 14 pixels. Doppler parameters in parentheses are the values used to convert these EW limits to upper limits on column density. cGiven the lack of information on the N(H i) from component A2, this total H i column density is likely only a lower limit. dThis is a conservative estimate based on blending a plausible line profile with b = 52 km s−1 with the fitted A2 component, and is not derived from an EW upper limit. See Section 5.3.1. eEquivalent width and column density limits are derived from the Si iiλ1260 line, but fitted b and N values are measured using all of the λλ1190, 1193, 1260, and 1304 lines. fThis is a conservative estimate based on blending plausible line profiles with the fitted A1 component and is not derived from an EW upper limit; column densities higher than this would alter the shape of the A1 component. See Section 5.3.1. gMeasurements from fitted lines based only on the Si ivλ1393 line because an O viλ1037 line at z = 0.35818 is at the expected wavelength of Si ivλ1403.

Download table as:  ASCIITypeset image

Table 4.  Equivalent Widths and Column Densities Toward v40, S393Z082

  v Δv σ(v) Wb bb σ(b)T    
IDa (km s−1) (km s−1) (km s−1) (Å) (km s−1) (km s−1) log(N) σ(N)T
(1) (2) (3) (4) (5) (6) (7) (8) (9)
        H i Lyα
       
B1 1240 −31 8 0.594 ± 0.036 19.1 ≤ b ≤ 46.2 17.60 ≥ log N ≥ 14.80
Total [N = 60]       0.603 + 0.101 – 0.119        
        C iiλ1334c
       
B1 1240 −31 8 <0.092 (22) ... <13.8 ...
        Si iiλ1260
       
B1 1240 −31 8 <0.036 (22) ... <12.37 ...
        Si iiiλ1206
       
B1 1240 −31 8 0.148±0.023 22.2 ±5.0 13.04 ±0.08
Total [N = 30]       0.140 + 0.054 – 0.060        
        Si ivλ1393
       
B1 1240 −31 8 <0.040 (22) ... <12.69 ...
        N vλ1238
       
B1 1240 −31 8 <0.048 (22) ... <13.41 ...

Notes.

aIn this column, "total [N = ]" refers to the total EW measured over N pixels (see Section 5.1). bEW limits are 2σ(W)T values measured over 14 pixels. Doppler parameters in parentheses are the values used to convert these EW limits to upper limits on column density. cC iiλ1334 lies at the same wavelength as a line likely to be O viλ1032 at z = 0.2986. The EW listed here is a 4σ(W) limit, and the column density is derived from that limit, assuming b = 22 km s−1. Such a line would be well distinguished from the narrow O vi interloper.

Download table as:  ASCIITypeset image

Table 5.  Equivalent Widths and Column Densities Toward v60, Q0246

  v Δv σ(v) Wb bb σ(b)T    
IDa (km s−1) (km s−1) (km s−1) (Å) (km s−1) (km s−1) log(N) σ(N)T
(1) (2) (3) (4) (5) (6) (7) (8) (9)
        H i Lyα
       
C1 1269 −20 11 0.269 34.7 −2.7, +3.1 13.93 −0.02, +0.03
C2 1342 +71 3 0.436 16.8 −1.6, +4.7 16.89 −1.39, +0.50
C3 1488 +217 14 0.070 28.7 −8.6, +11.7 13.16 −0.10, +0.13
Total [p]       0.707 + 0.056 – 0.047     16.89 −1.14, +0.35
Total [N = 60]       0.693 ± 0.055        
        C iiλ1334
       
C1 1269 −20 11 Blended
C2 1342 +71 3 0.063 14.5 −5.7,+6.6 13.59 −0.07, +0.09
C3 1488 +217 14 <0.044 (29) ... <13.37 ...
Total [p]       0.063 ± 0.019     13.59 −0.07, +0.09
        Si iiλ1260
       
C1 1269 −20 11 <0.036 (16) ... <12.36 ...
C2 1342 +71 3 <0.034 (17) ... <12.36 ...
C3 1488 +217 14 <0.040 (29) ... <12.41 ...
        Si iiiλ1206
       
C1 1269 −20 11 0.040 15.5 −12.6, +13.5 12.33 ±0.24
C2 1342 +71 3 0.085 9.3 −3.8, +5.1 12.92 −0.10, +0.56
C3 1488 +217 14 <0.058 (29) ... <12.49 ...
Total [p]       0.125 ± 0.029     13.02 −0.07, +0.43
Total[N = 35]       0.125 + 0.055 – 0.060        
        Si ivλ1393
       
C1 1269 −20 11 <0.040 (16) ... <12.68 ...
C2 1342 +71 3 <0.040 (17) ... <12.70 ...
C3 1488 +217 14 <0.044 (29) ... <12.73 ...
        N v
       
C1 1269 −20 11 <0.030c (16) ... <13.47 ...
C2 1342 +71 3 <0.030c (17) ... <13.49 ...
C3 1488 +217 14 <0.038 (29) ... <13.28 ...

Notes.

a"Total [p]" in this column refers to the total EW and total column densities calculated from the blended theoretical line profiles used to derive N and b values for individual components; "total [N =]" refers to the total EW measured over N pixels (see Section 5.1). bEW limits are 2σ(W)T values measured over 14 pixels. Doppler parameters in parentheses are the values used to convert these EW limits to upper limits on column density. cThese are the limits from the N vλ1242 line, because the λ1238 is blended with a higher-redshift line at the velocities of C1 and C2.

Download table as:  ASCIITypeset image

Table 6.  Equivalent Widths and Column Densities Toward v70, HE0241

  v Δv σ(v) Wb bb σ(b)T    
IDa (km s−1) (km s−1) (km s−1) (Å) (km s−1) (km s−1) log(N) σ(N)T
(1) (2) (3) (4) (5) (6) (7) (8) (9)
        H i Lyα
       
D1 1219 −52 4 0.087 16.0 −6.6, +11.5 13.34 −0.13, +0.22
D2 1309 +38 2 0.179 19.7 ±4.3 13.82 ±0.04
Total [p]       0.266 ± 0.050     13.95 ±0.04
        C iiλ1334
       
D1 1219 −52 4 <0.064 (16) ... <13.59 ...
D2 1309 +38 2 <0.060 (20) ... <13.54 ...
        Si iiλ1260
       
D1 1219 −52 4 <0.024 (16) ... <12.19 ...
D2 1309 +38 2 <0.024 (20) ... <12.19 ...
        Si iiiλ1206
       
D1 1219 −52 4 <0.044 (16) ... <12.39 ...
D2 1309 +38 2 <0.046 (20) ... <12.40 ...
        Si ivλ1402c
       
D1 1219 −52 4 <0.048 (16) ... <13.10 ...
D2 1309 +38 2 <0.048 (20) ... <13.08 ...
        N vλ1238
       
D1 1219 −52 4 <0.032 (16) ... <13.22 ...
D2 1309 +38 2 <0.032 (20) ... <13.22 ...

Notes.

a"Total [p]" in this column refers to the total EW and total column densities of Lyα calculated from the blended theoretical line profiles used to derive N and b values for the two individual components; no measurement from the actual spectrum is given because the components are slightly blended with Lyβ at z = 0.189. bEW limits are 2σ(W)T values measured over 14 pixels. Doppler parameters in parentheses are the values used to convert these EW limits to upper limits on column density. cThese are the limits from the Si ivλ1402 line, because the λ1392 is blended with a higher-redshift line.

Download table as:  ASCIITypeset image

The results of the line profile fits are given in Tables 36. For all the fits performed, b and N were allowed to vary as free parameters. In all cases, however, v was fixed to the values shown in column 2. The alignment of components between metal lines and Lyα is discussed in detail below for each sightline. Values of σ(b)T and σ(N)T are given in these tables (columns 7 and 9): these represent the errors in b and N resulting from a combination of continuum fitting errors and theoretical line profile fitting errors (again, see Bowen et al. 2008). The errors in velocity (column 4) are simply those found from the MC simulations, and do not include possible systematic errors discussed in Section 4.1.2.

For multicomponent fits, the total (summed) column density over all components is given at the bottom of column 8 for each species. The error on this total, given at the bottom of column 9 for each species, was calculated from the distribution of summed column densities derived for each MC run (and not from any combination of the errors from individual components, because these are not independent for multicomponent fits).

Tables 36 also include the equivalent widths (EWs) W of detected lines (or upper limits when no lines were detected—see below). Several of the absorption lines from NGC 1097 were blended with other higher-redshift lines, in which case their EWs were derived directly from their theoretical line profiles and not measured from the data. The EWs of the individual components are given in column 5 of the tables, but when more than one component exists, the total EW in column 5 (referred to in column 1 as "total [p]") is given for the EW of the blended components, which need not equal the sum of the individual component EWs.

Unfortunately, the resulting errors in W for EWs that rely on the profile fits cannot be obtained by using the errors in N and b because they are correlated in such a way that N and b can vary while W stays the same. Instead, total EWs were recorded for each of the 300 MC profile fits, and used to form a distribution of EWs. The values of W were always normally distributed, yielding an error σ(W). Again, W was measured for three possible continuum fits, yielding errors that were added in quadrature with σ(W) to give a total error. These are the errors that accompany the EW totals in column 5.

Tables 36 also list total EWs measured in the traditional way, directly from the spectra, calculated over a specific number of pixels (referred to as "total [N =]" in column 1). The values of N here are chosen simply to be large enough to cover all the detected components. These EWs are independent of the assumed line-spread function (LSF) or any other assumptions about how the absorption line profile should be modeled. The errors listed include errors from Poisson noise and deviations from the continuum fitting, analogous to the column density errors described above.

5.2. EW Limits and Column Density Limits for Undetected Metal Lines

In many cases, no metal lines were detected toward the QSO sightlines, and column density limits had to be derived from EW limits. Traditionally, the EW limit at some wavelength is taken to be ${\sigma }^{2}(W)\,=\,{\displaystyle \sum }_{i}^{N}\ {({\sigma }_{i}\delta \lambda )}^{2}$, where δλ is the wavelength dispersion in Å pix−1 and N is the number of pixels that the sum is made over, centered at the wavelength of interest. Or more simply, if σi is close to being the same value σ over N pixels (because the errors change slowly with wavelength) then $\sigma (W)\,=\,\sigma \delta \lambda \sqrt{N}$. The COS LSF consists of a Gaussian core superimposed on much broader wings, as a result of mid-frequency wavelength errors in the HST mirrors (Kriss 2011; Holland et al. 2012). If the LSFs generated for the profile fitting described above are integrated, then 95.5% of the area of a line is enclosed in 14 of our rebinned pixels. We adopted N = 14 as the value over which to measure EW limits, knowing that the most a line EW might be underestimated would be by ≃5%.

For undetected lines in the QSO spectra, we list upper limits of 2σ(W) in Tables 36. To derive column density limits we used b values from other lines where possible, and the values adopted are listed in parentheses in column 6 of Tables 36.

In the following sections, we look in detail at the absorption lines detected along each sightline, and discuss the measurements made from the line profiles.

5.3. Absorption Lines Toward Q0244–303

5.3.1. Metal Lines

The absorption lines seen toward Q0244 are shown in Figure 5. Metal lines are detected toward Q0244 at at least two velocities; in the first case, C iiλ1334, Si iiλ1193, Si iiλ1260, Si iiiλ1206, and Si ivλ1393 are found at a velocity of 1384 km s−1. This component is labelled A3 in Table 3, and categorizes the strongest metal-line absorbers along the sightline. There is also a marginal detection of Si iiλ1190 at this velocity, although its reality is based largely on the presence of the other Si ii lines since its strength is below our detection threshold. As expected therefore, the weakest line of the four Si ii lines, Si iiλ1304, is absent. Si ivλ1403 at 1384 km s−1 is contaminated by an O viλ1032 line at z = 0.35818, from a system that also shows corresponding Lyβ, Lyγ, Lyδ, and O viλ1037. Hence the λ1403 line is not used in the derivation of the Si iv column densities.

Figure 5.

Figure 5. Absorption lines found toward Q0244. The Lyα line is contaminated by two other lines, O viλ1031 at z = 0.18272 and H iλ930 at z = 0.31229; the expected profiles of these two lines are shown as dotted lines, although the data shown are already corrected for them (see text). As discussed in the text, only two components (A1 and A3) are fitted to the body of the Lyα line, but the position of a possible component seen in C iiλ1334 is labelled (A2).

Standard image High-resolution image

A second component is detected from Si iiiλ1206 at 1227 km s−1; although there are no other metal lines to confirm the reality of this component, there is clear evidence of Lyα at a similar velocity (see below), and so we consider this component to be real. This component is labelled A1.

Finally, a weak feature is detected near the C iiλ1334 A3 component, which, if a lower-velocity C ii absorption feature, would be at v = 1261 km s−1. This velocity is clearly different from that of A1 and A3. We label this possible system A2 in Table 3 and discuss its reality below.

In order to establish the velocity of components A1 and A3, all the metal lines were fit simultaneously. Fits were first made without including the Lyα line. (The Lyα components are clearly blended, so there is little information in the line profile to help establish the components' velocities.) Values of b and N were allowed to vary independently for each component of each species, and while v was also allowed to vary, it was constrained to be the same for each component of each ion.

Although A2 is defined only by the possible existence of a weak C ii line, Table 3 lists EW limits and column density limits for the other species at the same velocity. For Si iii, a limit is hard to define, since any absorption would be blended with the A1 component. The Si iii A1 component is broad, and the low S/N of the data makes the profile appear coarse, but there is no obvious need to add an additional component to represent possible absorption at the A2 velocity. We estimate a plausible column density for Si iii based on adding a fictitious A2 (unresolved) component and seeing how the blended profile changes with N(Si iii).

A similar problem exists for defining a C ii column density limit for A1, since any line at that velocity would be blended with the defining A2 C ii component. We again estimate an upper limit to C ii at A1 by blending a hypothetical component (this time with the b value of the broad Si iii component) with the existing C ii component.

5.3.2. Lyα

The Lyα absorption lines detected from NGC 1097 are complicated by contamination by lines from two higher-redshift systems. One contaminant is Lyλ930 arising from a system at z = 0.31229; the H i from this system is very well constrained, however, since Lyβ, Lyγ, Lyδ, Ly$\epsilon $, Lyλ923, and Lyλ920 lines are all present. Absorption can be fit with a single component to give log N(H i) = 15.88 and b = 38.9 km s−1. With these values we can calculate a profile for Lyλ930 (shown as a dashed line in Figure 5) and remove it from the data. There also likely exists a second contaminant to the Lyα complex, an O viλ1031 line at z = 0.18272. This system is defined primarily by Lyα, Lyβ, and Lyγ lines, and so appears secure. There is a feature at 1227.2 Å that is probably the O viλ1037 line from this system. The O vi absorption is weak and simple, and a fit to the O viλ1037 line gives log N(O vi) = 14.19, b = 20.6 km s−1. We again model the corresponding O viλ1031 line using these values and remove it from the Lyα lines from NGC 1097. Both these interloping Lyλ930 and O viλ1031 lines are relatively weak compared to the strong Lyα lines from NGC 1097 and contribute little to the total EW.

The Lyα profile clearly argues for at least two components, but determining the column density for A3 remains problematic; without including any more information from the metal lines (apart from their velocity) values of log N(H i) = 15.07 and b = 18.9 are obtained (Table 3), but the solution is not unique, and the same profile can be obtained for higher N(H i) and smaller b values. An upper limit for these can be set by the appearance of damping wings, which is not seen in the data, and which occurs when log N(H i) ≳ 17.1 and b ≲ 11 km s−1.

While a two-component fit can adequately represent the observed Lyα absorption, we noted in the previous section the possible presence of an additional C ii component—A2—which could have a corresponding Lyα component. There is no information in the Lyα profile to constrain the physical parameters of a component matching A2, but we can explore how badly the Lyα b-values and column densities of A1 and A3 might be affected by including A2. An upper limit to N(H i) for A2 can be set at log N(H i) ≃ 18.1, when the damping wings of A2 are wider than the entire profile. The maximum b-value of A2 is given by the b-value of the C ii component, which has b = 10.9 km s−1. If b(C ii) was defined purely by turbulent processes, then b(Lyα) would be the same as b(C ii), or 10.9 km s−1. If b(C ii) was entirely a thermal width, then the corresponding temperature would be ∼8.6 × 104 K. The width of a Lyα line at that temperature would be b(Lyα) = 38 km s−1. We can therefore use b = 10.9 km s−1 and b = 38 km s−1 as lower and upper limits to b for A2. We can then fix the velocities of all three components to those of the metal lines, vary N(H i) for A2 in fixed intervals, and then refit, to see how N(H i) and b change for A1 and A3.

In fact, for either value of b for A2, the values of b and N(H i) change little for A1 and A3 until the N(H i) of A2 reaches its upper limit of log N(H i) ≃ 18.1. For N(H i) less than this, most of the optical depth at the velocity of A2 is always taken up by A1, so increasing N(H i) for A2 makes little difference to the optical depth at the velocity of A2. This suggests that N(H i) for A1 and A3 may not be too badly affected by the presence of even a moderately strong Lyα component at the velocity of A2.

Finally, we note the presence of additional weak, higher velocity components, which we label A4 and A5 in Table 3 and Figure 5. The figure shows the Lyα profile after the contaminating O viλ1032 and H iλ 930 lines have been divided out of the data. We find values of N and b from the profile fitting, but the lines are unresolved. The limit of the COS resolution is ∼15 km s−1 FWHM (ignoring the non-Gaussian shape of the LSF) or b ≲ 9 km s−1. The EWs of both Lyα lines lie on the linear part of the Curve of Growth so long as b ≳ 7 km s−1 (i.e., T ≃ 3000 K) with log N(H i) ≃ 13.1. If b is really less than 7 km s−1, however, then N(H i) is a lower limit. As b is unconstrained, we give N(H i) as a lower limit in Table 3.

It is useful to summarize the above discussion in terms of the total N(H i) along the sightline, a value we will use later in this paper. The lower limit to N(H i) is set by A1, and is log N(H i) ≃ 15.6 (Table 3). The upper limit is set by a lack of a damping wing on the red side of A3, log N(H i) ≲ 17.1; however, the possible existence of H i corresponding to the C ii component A2 increases this limit to log N(H i) ≲ 18.1, at which point damping wings would be seen on both sides of the complex.

5.4. Absorption Lines Toward 2dFGRS S393Z082

5.4.1. Metal Lines

As shown in Figure 6, Si iiiλ1206 is detected unambiguously at 1240 km s−1 toward S393Z082, a system we label B1. A weak feature is also detected at the wavelength where C iiλ1334 is expected at the same velocity as B1, but the line could be O viλ1032 at z = 0.29863, because there is also a very weak feature that could be Lyβ at that redshift. The corresponding O viλ1037 line is not detected at this redshift, but given the weakness of the λ1032 line, this is not unexpected for unsaturated lines. Fitting Voigt profiles to the Lyβ, O viλ1032, and (to provide an upper limit) O viλ1037, a solution of b = 13.8 km s−1, log N(H i) = 13.7, and log N(O vi) = 13.5 can be found (Figure 7). Although an O vi/H i absorber with such a narrow linewidth is relatively rare, Tripp et al. (2008) have found such systems. Further, the ratio of the column densities in our fit, log[N(H i)/N(O vi)] = −0.17 is consistent with the values found by Tripp et al. for the value of log N(H i) = 13.7 that we find for the system. Given the precise alignment of the two lines, and the plausibility of physical parameters obtained from the profile fits when compared to other O vi systems, we cannot be sure that the line at 1340.1 Å is C ii from NGC 1097 and refrain from using it in our analysis. We note that if the identification is wrong, and the line is indeed C iiλ1334, then the profile fit to the line requires log N(C ii) ≃ 14.0.

Figure 6.

Figure 6. Absorption lines found toward S393Z082. The values of N(H i) and b are ill constrained for Lyα absorption from NGC 1097: two theoretical line profiles are shown in the top panel, each representing nearly identical best fits to the data: one has a solution with high N(H i) and low b (green), the other has low N(H i) and high b-values (red). The two profiles are nearly indistinguishable from each other. There exists a possible detection of C iiλ1334 (second panel), but the line is likely to be O viλ1032 at z = 0.2986. Si iiiλ1206 is the only metal line unambiguously detected from NGC 1097 (fourth panel).

Standard image High-resolution image
Figure 7.

Figure 7. While the line at 1340.1 Å toward S393Z082 could be C iiλ1334 absorption from NGC 1097 (see second panel of Figure 6) another possible interpretation of this line, along with another at 1332.0 Å, is that they represent O viλ1032 and Lyβ absorption at z = 0.29863 instead. This figure shows these lines (as well as the expected absorption from O vi λ1037) in that rest frame, as well as theoretical profile fits for b = 13.8 km s−1, log N(H i) = 13.7, and log N(O vi) = 13.5.

Standard image High-resolution image

5.4.2. Lyα

This is the simplest of the Lyα absorption lines associated with NGC 1097, but the physical parameters are not well constrained. The data have the lowest S/N of the four sightlines, and there is some ambiguity over the number of components to fit. A single component, which we label B1 in Figure 6, appears to be adequate; there is a slightly lower "shelf" at 1100 km s−1, and adding a component to match this feature changes N(H i) for B1 by a factor of 4. The MC simulations used to calculate the errors in N(H i) and b-values, however, suggest that the small amount of extra depth may well just be from noise, since in many cases the feature is not replicated in the synthetic spectra. We therefore assume that no additional absorption occurs at this lower velocity.

The Lyα line provides an example of the classic problem of measuring N(H i) and b when its EW lies on the flat part of the Curve of Growth. The line-profile fitting settles on two possible solutions, one with b = 46 km s−1 and log N(H i) = 14.8, and a second one with b = 19 km s−1 and log N(H i) = 17.6. The problem is exacerbated by the low S/N of the data, providing insufficient constraints on the shape of the wings of the line. An analysis in which the width of the line is determined by btherm and bturb components (see Appendix C) shows that the change in log N(H i) from 14.8 to 17.6 occurs when T > 1 × 104 K. Figure 6 better shows the problem, where theoretical lines profiles for both the high-N fit (green) and the low-N fit (red) are drawn; the profiles for the two solutions are almost identical.

5.5. Absorption Lines Toward Q0246–308

5.5.1. Metal Lines

Along this sightline, Si iii appears to be detected at the redshift of NGC 1097, and being shortward of Lyα, the reliability of its identification is strong. This component is labelled C2 in Table 5 and Figure 8. There is an additional weak Si iii component to the blue side of C2, which is designated as C1.

Figure 8.

Figure 8. Metal absorption lines found toward Q0246. The defining component of the system is labelled C2, seen in both Si iii and Lyα, and possibly in C ii; weaker Si iii and Lyα components exist at a more negative velocity, labelled C1. No Si ii or Si iv is detected from either C1 or C2. C3 is identified only by Lyα absorption. The absorption at the expected velocities of the C ii components is complex, and the strongest feature is probably a Lyα line at z = 0.10224. Although we identify C ii from system C2, it remains possible that this feature is part of that higher-z complex at z = 0.10272.

Standard image High-resolution image

Any C iiλ1334 from system C1 is lost in a line that, given its strength and width, is probably Lyα at z = 0.10224 (Figure 8, second panel) although there are no other corroborating lines. C ii from system C2 is expected at 1340.5 Å, and a weak feature is seen very close to that wavelength, although it is nestled next to the high-zLyα line. N(C ii) for the line is given in Table 5, and while we consider the detection to be real, it remains possible that the feature is simply an additional higher-velocity (z = 0.10272) component that is part of the strong high-z Lyα line.

A feature is also seen at 1244.32 Å that could be N vλ1238 from NGC 1097, but there is no N vλ1242 line to confirm this. Fitting the N vλ1238 line gives log N(N v) = 13.50 and b = 20 km s−1; with these values, the N vλ1242 line should be detected, but is not. We therefore list only upper limits in Table 5.

5.5.2. Lyα

The Lyα absorption from NGC 1097 toward Q0246 comprises three components. Two of these well match the systems C1 and C2 identified in the metals; a third, broader line is evident at a higher velocity, which we label C3 in Figure 8, but which has no corresponding metal-line absorption. The Lyα component in system C2 is another example of a saturated Lyα, and constraining N(H i) is challenging. For this spectrum, the S/N is higher than for S393Z082, which helps constrain the fits of theoretical Voigt profiles; nevertheless, there is still a large uncertainty in N(H i), which is reflected in the errors listed in Table 5.

5.6. Absorption Lines Toward HE0241–3043

5.6.1. Metal Lines

No metal-line absorption is detected from NGC 1097 along this sightline. The lack of detections is shown in Figure 9, and upper limits are listed in Table 6. Absorption by Si ivλ1393 from NGC 1097 would be blended with a pair of Ly$\epsilon $ lines at z ≃ 0.493, so column density limits are measured for Si ivλ1403 instead.

Figure 9.

Figure 9. Absorption lines found toward HE0241. Only Lyα is detected from NGC 1097.

Standard image High-resolution image

5.6.2. Lyα

Though no metal lines are detected, two Lyα components are seen from NGC 1097. These are labelled D1 and D2 in Figure 9 and Table 6, and lie close to Lyβ lines from a two-component complex at z = 0.18868 and z = 0.18945. D1 Lyα from NGC 1097 is slightly blended with the higher-z of these two Lyβ lines, as can be seen in the figure. Unfortunately, any Lyα components at velocities less than that of D1 will be lost in the Lyβ lines. Neither D1 or D2 can, instead, be an additional Lyβ component at z ≃ 0.189 because the corresponding Lyα lines are not seen at longer wavelengths. We can find no other systems for which D1 and D2 might be metal lines.

6. PHOTOIONIZATION MODELS OF ABSORBING CLOUDS

In order to better understand the physical characteristics of the absorbing clouds, we have attempted to reproduce the measured values of, or limits to, the column densities for all ion species along the QSO sightlines using photoionization models. The results depend on various assumptions, and as the list is rather extensive, we present the details of our analysis in Appendix C. Unsurprisingly, only components with detected metal lines, those labelled above as A1, A3, B1, and C2, give useful constraints. In summary, the metallicities are probably sub-solar, Z < 0.3 Z for A1, B1, and C2. For A3, however, we can find no single-phase solution, meaning that either the gas is not simply photoionized (or indeed collisionally ionized, which was also considered) or that the component-fitting model used to model N(H i) is inadequate.

7. CONSTRAINTS ON HALO GAS KINEMATICS

In this section we use the kinematics of the absorption lines toward NGC 1097—the profiles of the lines and velocity centroids—to model the gas in the halo of the galaxy. We focus specifically on the detected Lyα lines because the H i absorbing gas is the most sensitive tracer of cool gas available from our observations, and is detected along all sight lines.

7.1. Disk Models

Simulations of galaxy structures in a ΛCDM universe predict that the halos of galaxies should accrete gas from the IGM along cool gas streams that possess their own angular momentum (Stewart et al. 2013; Cen 2014; Danovich et al. 2015). This leads us to question whether the absorption seen toward NGC 1097 might trace similar structures. We consider two variations of gas rotation in NGC 1097: (a) a rotating disk model, and (b) a rotating-inflow disk model, where gas is flowing into the disk (akin to an accretion disk). For reasons that we explain, we will also consider a third option, that of the same rotating–inflowing disk model, but with a conic outflow launched from the center of the galaxy, perpendicular to the disk. For all these models, we use the galaxy parameters listed in Table 1.

7.1.1. Rotating Disk Models

In this model, we assume an inclined thin disk rotating at a fixed azimuthal speed vrot. A positive value of vrot indicates clockwise rotation of the disk shown in Figure 3—the analysis by Quillen et al. (1995) suggests that the southwest (SW) side of the galaxy lies nearer to us than the NE side, and therefore that the galaxy rotates with its spiral arms trailing the rotation. The column density of the disk is assumed to follow an exponential decline with the deprojected (e.g., Williams & Hodge 2001) radius $\widetilde{\rho }\propto \exp (-\tilde{\rho }/{\tilde{\rho }}_{d})$, where the disk scale is ${\tilde{\rho }}_{d}=11\,{\rm{kpc}}$. To model the absorption we assume that microturbulence in the gas produces lines with widths of vturb = 20 km s−1, which result in the observed line profiles after convolution with the COS LSF.

Our aim is to use the parameters described to predict the line profiles that would be expected along the four QSO sightlines. As there are many parameters that can be varied to produce the predicted profiles, we refrain from, e.g., attempting to minimize a χ2 statistic between the theoretical profiles and the data—there is no unique solution for the number of parameters available. Instead, we look for more qualitative agreements that might guide us toward finding a correct scenario for the origin of the absorption.

So, for example, by adopting of value of vrot ≃ +70 km s−1, we find that our predicted profiles show some agreement with the observed data. (This velocity is smaller than the circular speed of the inner disk observed from measurements of 21 cm emission, which is ∼200–300 km s−1, a point we return to in the next section.) Figure 10 shows the predicted profiles in comparison to the observed profiles. While the impact parameters of the two outer sightlines differ by only 15 kpc, the Lyα line profiles of each are markedly different, with the sightline toward HE0241 showing none of the saturated absorption seen toward Q0246. As the former sightline lies closer to the minor axis of the disk, a lower opacity would indeed be expected. Further, the absorption velocities toward Q0244 and HE0241 are centered around the systemic velocity of NGC 1097, as would be expected for sightlines at opposite ends of the galaxy's minor axis.

Figure 10.

Figure 10. Predictions of Lyα line profiles (red lines) for QSO sightlines passing through a rotating, geometrically thin disk model, compared to our COS data. The central figure is a simple schematic of the galaxy–QSO configurations, with a disk that has approximately the same inclination and position angle as NGC 1097 shown in Figure 3; the SW side of the disk is assumed to be closer to us, with the spiral arms trailing the rotation. The positions of the QSOs are indicated by red circles; they are arranged to roughly match their true configuration (Figure 3), and indicate which structures of the model are intercepted along a given sightline. (The inset image of the galaxy is not to scale, but is included to show the orientation of the spiral arms relative to the rotation.) The predicted line profiles are built from a model that follows the galaxy parameters in Table 1, as well as the positions of the QSOs on the sky, precisely. The disk has rotation speed of vrot = + 70 km s−1 and a turbulent velocity of 20 km s−1. The model is qualitatively consistent with the data toward Q0244 and Q0246, although only one broad component is predicted for Q0244 when there are clearly two in the data; the model also overpredicts the strength of the absorption toward S393Z082, and cannot reproduce the double line profile toward HE0241. Also shown are fmin curves (blue dashed-line histograms) that are used to estimate the allowed rotation speed interval for each sightline (as shown by the blue shaded regions). For comparison, profiles are also plotted (red dashed lines) for an absorbing disk counter-rotating with respect to NGC 1097. Such counter-rotation would produce absorption at velocities quite different from that observed toward S393Z082 and Q0246.

Standard image High-resolution image

In contrast, the sightlines toward S393Z082 and Q0246 lie close to the opposite ends of the major axis and the absorption ought to be shifted away from the galaxy's systemic velocity, and with the velocities in opposite directions. The implied sense of rotation is consistent with that assumed for the inner disk; for comparison, Figure 10 also shows the results from a disk model where the gas counter-rotates with the galaxy (vrot = −70 km s−1): here the predicted absorption clearly fails to match the data. As expected, the discrepancies are largest toward the sightlines along the major axis where differences in vrot would be most noticeable.

To estimate the uncertainties associated with vrot, we define a minimum fraction of the observed line profile that lies on either side of the centroid velocity of the predicted absorption vc, i.e., fmin(vrot) = min[EW(v > vc(vrot)), EW(v < vc(vrot))]/EW. Hence a model whose predicted velocity is centered around the observed absorption would have fmin ≃ 0.5. For predicted line profiles whose vc significantly deviates from the observed centroid velocity, fmin ≪ 1. To estimate the uncertainty of vrot, we calculate fmin for a range of vrot values, and take the profile to be an acceptable replica of the data if fmin > 0.2. This variation of fmin as a function of vrot is shown as a dashed line at the bottom of each panel in Figure 10. As noted above, some sightlines are more sensitive to the disk dynamics than others, depending on where they probe the disk. For example, the sightline toward Q0244 is largely insensitive to disk rotation (top left panel of Figure 10), because its sightline passes close to the minor axis of NGC 1097, and hence fmin changes little for any value of vrot (fmin ≃ 0.5 for −500 km s−1 < vrot < 500 km s−1). The other sightlines provide more stringent constraints on vrot, as Figure 10 shows, and the similarity in the velocities at which fmin peaks in the three other panels, ∼70 km s−1, suggests that this is a good estimate of vrot.

Although this model reproduces some of the absorption line features, there are notable discrepancies between the rotating disk model and the data. For example, the model fails to predict the two-component absorption seen toward HE0241. This might be the easiest discrepancy to understand. One possibility is that a thin disk model might no longer apply at such large radii. In fact, at a radius of ∼200 kpc, the dynamical time $\sim 2\pi \tilde{\rho }/{v}_{{\rm{rot}}}$ is comparable to the Hubble time. Alternatively, the two-component absorption may mean that at large radii, H i clouds no longer form a homogeneous thin disk. Additionally, discrepancies associated with under- and overprediction of the opacity along particular sightlines can be resolved if either vturb or the column density profile deviates with radius (see, e.g., the variations in the rotation curves around a mean value in Ondrechen et al. 1989; Hsieh et al. 2011). More worryingly, the model predicts only one absorption component toward Q0244, whereas at least two are observed; and the absorption toward S393Z082 is somewhat narrower than the model predicts.

7.1.2. Rotating–Inflowing Disk Models with a Conic Outflow

As noted above, the rotation speed of an extended disk with vrot ≃ 70 km s−1 is significantly smaller than the 200–300 km s−1 rotation speed of the inner disk (Ondrechen et al. 1989, see also Section 9.1). This fact, combined with results from ΛCDM models that imply the existence of cold-flow accretion, leads us to consider a model in which the absorbing gas has an inward radial velocity vr.

The results of such a model are shown in Figure 11; data toward sightlines that lie in the direction of the minor axis are more sensitive to vr, and imply vr < 100 km s−1. We find that a model with vrotvr ≃ 70 km s−1 improves the match between data and model toward some of the sightlines. Toward Q0244, the blue component, A1, is better reproduced by the model, but at the expense of A3, which is not predicted at all. The predicted absorption toward S393Z082 is improved, while the predicted absorption toward Q0246 is unchanged. One of the components toward HE0241, D2, now matches the predictions, but the other component, D1, is still not reproduced.

Figure 11.

Figure 11. Predicted Lyα line profiles (red lines) for QSO sightlines passing through gas accreting onto NGC 1097 via a rotating, geometrically thin disk model, as compared to our COS data. The disk has vturb = 20 km s−1, a rotation speed of vrot = 70 km s−1, and an inflow speed along the disk's (inward) radial direction of vr = 70 km s−1. For comparison, the predicted line profiles for gas flowing away from the disk at a speed of vr = 70 km s−1 are shown as blue lines. Compared to the thin-disk model shown in Figure 10, this model better reproduces the absorption seen toward S393Z082 and Q0246, component A1 toward Q0244, and D2 toward HE0241. However, it fails to account for components A3 toward Q0244 and D1 from HE0241. One way to explain component A3 toward Q0244 is to add a conic outflow (purple cone in the middle schematic) with an outflow velocity of 300 km s−1 subtending an angular range of 65° < θ < 70° and extending out to at least 50 kpc above the disk. Such a model leads to the green line in the top left panel, with the blend of the profiles from the two models shown in magenta. The inset diagram (middle, framed bottom panel) represents a cut through the model at the position of Q0244, which shows further how positive velocity absorption (i.e., component A3) can be obtained from a wide conic outflow (purple regions) as it emerges from the (green) disk. So long as θ ≥ 90 − i, the edge of the outflow will have a velocity greater than the systemic velocity. In this particular scenario, outflowing gas occurs on only one side of the galaxy.

Standard image High-resolution image

The values of vrot and vr are not entirely unique, and similar (but not superior) fits can be obtained for other combinations of vrot and vr. This model offers an improvement over the simple rotating disk presented above, but still fails to match all the observed features.

NGC 1097 shows moderate amounts of star formation at its center (Calzetti et al. 2010; Hsieh et al. 2011) so it is reasonable to consider the possible effects of an outflow in our models. In several studies where outflows have been traced by gas and dust emission (most notably toward AGNs) a hollow-cone geometry has been inferred whose axis lies along the disk's angular momentum vector (Veilleux et al. 2001, 2005) and where the wind may be detected in emission out to several kiloparsecs above the disk, and potentially out to even larger scales in absorption (Murray et al. 2011).

We consider the effect of adding a hollow-cone model to our rotating–inflowing model by including gas that flows out of the galaxy along radial trajectories at a speed vrad. This addition is also shown in Figure 11. The most notable result from adding an outflow component is that we can replicate the additional A3 absorption component (Table 3) toward Q0244, using a model that has a cone with vrad ≃ 300 km s−1 and that is so wide—65° < θ < 70°—that part of the cone wall is still receding from us, providing an absorption component that has a positive velocity (redshift) with respect to the rest of the galaxy. This scheme is highlighted in the inset diagram in Figure 11. Wide-angle outflow cones have been seen in the inner regions of some galaxies, e.g., toward ESO 097-G013 (Greenhill et al. 2003), so a weak outflow that contributes to the absorption is plausible. A wide, amorphous outflow more like the superwind seen toward M82 might be the best analog (e.g., Leroy et al. 2015, and references therein).

7.1.3. Contributions from Warped Disks

Galactic disks are known to exhibit warps, especially in systems undergoing interactions with other galaxies. If the inclination angle of the disk changes with radius, for example, and/or the position angle of the major axis changes with radius, absorption line profiles could be replicated to better match the data with these added variables. For example, the simplest case is where the inclination angle varies with radius, which would allow for a wider range of vrot at different radii. Among the various possibilities for geometrically thin disks we mention only two: good agreement between data and line profiles can be found when the outer disk has vrot = 250 km s−1 and i = 10° (i.e., nearly face-on), or vrot = 60 km s−1, i = 75°, and a very large disk scale of ∼30 kpc. So, given the inclination of NGC 1097 (40° ± 5°), substantial warping would be required to account for the observed absorption. There is some indication from the 21 cm H i maps of Higdon & Wallin (2003) that the outer spiral arms of NGC 1097 are disturbed by tidal interactions (which we discuss later), so distortions in the structure of any extended disk are a possibility. However, far more additional data on the nature of the warps in NGC 1097's disk would be needed before they could be reliably added to our model.

7.2. Outflow Models

While a conic outflow can be added to the disk model to explain the additional absorption (component A3) seen toward Q0244, its inclusion is somewhat ad hoc, and leads us to examine whether, in fact, outflow models alone—with no contribution from an absorbing disk of gas—could adequately reproduce the observations. We assume bi-conic outflows whose symmetry axis is parallel to the angular momentum vector of the inner disk. Geometrically, the outflow is defined by an inner and an outer polar angle, as well as an inner and outer radius. This parameterization allows us to construct both spherical and polar outflows. By specifying an inner and outer radius to the outflow, intermittent flows or shells—rather than continuous winds—can also be modeled. We set the outer radius to arbitrarily large values; whether outflowing gas ejected from the center of a galaxy can travel well into, and survive within, the outer regions of a galactic halo is a separate issue, and beyond the scope of this paper. For the sake of brevity, we do not reproduce a figure similar to Figures 10 and 11, but simply describe the results from our analysis.

We assume that the outflows are coasting at a constant radial outflow velocity, vrad. If mass is conserved along outflow lines (e.g., no mass loading), then the gas density is nr−2, where r is the radial coordinate. For continuous spherical winds, this also implies that the absorbing column density declines as 1/ρ. In contrast, for thin-shell geometries whose size is larger than the impact parameter range in question, the column density is independent of ρ. Clearly, significant deviations from the above scalings could occur if vrad varies with r, or if there are multiple ionization phases within the outflow (Rupke & Veilleux 2013; Scannapieco & Brüggen 2015). Such models cannot be constrained by the current data set.

7.2.1. Spherical Outflow Models

By virtue of isotropy, the absorption line profiles predicted from a spherical outflow model are centered symmetrically around the systemic velocity. For example, expanding shells would generally produce two-component line profiles as the sightlines toward the background QSOs intercept the front and back sides of the outflow. This well matches the double component profile (D1 and D2) toward HE0241, and we used these components to calibrate a thin-shell model with δrr (a similar approach was adopted by Tripp et al. 2011). Specifically, by matching the model to the D1 and D2 components, we find that the mass in the outflowing shell is >3 × 105η M, where 1/η is the mass fraction of the gas in the form of H i, a potentially small number related to the uncertainties in the ionization corrections and/or the multiphase nature of the outflow. The mass-loss rate is then of order 4πr2ηN(H i)mp(vrad/r) ≃ 10−4η(r/200 kpc)(N(H i)/1014 cm−2)(vrad/40 km s−1) M yr−1, which is probably much smaller than the current rate of star formation. Irrespective of this argument, thin-shell models fail to explain the decreasing gas opacity with impact parameter—the predicted column densities are always too low to produce the strong absorption seen along all the sightlines. Moreover, the asymmetric absorption around the systemic velocity along sightlines at opposite ends of the disk's major axis (i.e., toward S393Z082 and Q0246) cannot be reproduced.

We find that the simplest models of spherical outflows seem to be inadequate for reproducing the observed absorption, and given the limited set of observational constraints, a more detailed analysis seems unwarranted.

7.2.2. Polar Outflow Models

In Section 7.1.2, we showed how adding a polar outflow to the rotating disk models could account for the A3 component seen toward Q0244. In general, however, models that include only polar outflows and no disk component run into difficulties explaining the absorption along the sightlines that lie in the direction of the major axis and are at large impact parameters, unless the opening angle is large or the flow extends far enough above the disk. For example, assuming an outflow that extends out to Rv, an opening angle of >50° is needed to produce absorption toward Q0246.

The model shown in Figure 11 assumes outflow from only one side of the galaxy, but a bipolar outflow would produce additional positive velocity absorption. Outflowing gas from the far side of the galaxy would need to extend further than the front-side flow to intercept the sightline. We do indeed see weak components—A4 and A5—at higher velocities, which might fit such a scenario. Whether the extent and covering factor of the outer regions of an outflow could account for such weak absorption is unclear.

We should note that gas outflowing at the observed velocities would need to travel over timescales that are comparable to the dynamical time of the halo, which would most likely invalidate the simple conic geometry. For example, gas may be dragged perpendicular to the outflow if the halo is rotating, or minor mergers may disrupt the geometry of the escaping gas. Again, modeling the absorbing properties of such disrupted outflows is beyond the scope of this paper.

8. GLOBAL PROPERTIES OF THE ABSORPTION

We have observed four QSOs with impact parameters of ρ = 48–165 kpc from NGC 1097 and searched for absorption from the galaxy along each sightline. The lowest quality spectrum has a 2σ EW limit of 35 mÅ at short wavelengths, defining the limiting EW of our survey for Lyα and Si iiiλ1206. These limits correspond to column densities of log N(H i) =12.9 and log N(Si iii) = 12.3 (for Doppler parameters of b ≥ 7.1 km s−1, or thermal temperatures ≥3000 K). We have seen that while the EWs of all the detected lines are measured precisely, N(H i) derived from the Lyα lines for three of the sightlines is less well defined, because the lines have EWs where b and N(H i) are degenerate. For the four sightlines, our most basic statistic suggests that the covering fractions of H i and Si iii are 100% and 75%, respectively, at a limit of ρ = 165 kpc.

The values of Lyα and Si iiiEWs, and H i and Si iii column densities, plotted against their impact parameters normalized by the virial radius Rv, are shown in the left-hand column of Figure 12. The gas in the halo of the galaxy does not appear to be smoothly distributed. While a rank correlation test suggests that ρ/Rv and W(Lyα) are significantly anticorrelated, a Pearson's product-moment analysis shows that the correlation is not significant (the probability of the null hypothesis is p = 0.18, well above the 0.05 significance level often used to define a significant correlation). The same test for W(Si iii) with ρ/Rv actually finds more significance for anticorrelation.

Figure 12.

Figure 12. Left: distribution of total N(H i) summed over all components (top panel), Lyα EW (second panel), total N(Si iii) (third panel), and Si iiiλ1206 EW (bottom panel) toward NGC 1097 as functions of impact parameter normalized by virial radius. Our data are shown as red circles for detections, and pink circles for upper limits. For the N(H i) values in the top panel, our measurements are uncertain for the sightlines with the two smallest ρ, and we represent the range in possible N(H i) as two red/orange vertical bars. For Q0244, there are two possible upper limits to N(H i), one set by the maximum possible from component A3, which is shown in red, and one set by the maximum toward A2, extended in orange (Section 5.3). For S393Z082, the range is set by the two possible values of N(H i) that arise from two line profiles that are nearly identical (Section 5.4). For the three lower panels, power-law fits to the four points are shown as black lines, and the beige regions indicate errors obtained from χ2 confidence intervals (see text). Right: contour plots of χ2 centered on the minimum values (shown with crosses) found for power-law fits to Lyα EWs (top panel), Si iii EWs (middle panel), and N(Si iii) (bottom panel). The best-fit values are listed in the top right corner of each panel. Sets of α and W* or N* at the extreme deviations of the 1σ confidence intervals (black circles) are also indicated, and these two sets of values are the ones used to draw the shaded beige regions in the left-hand side of the plot.

Standard image High-resolution image

Irrespective of these rank correlation tests, however, we would like to measure how ρ/Rv and W(Lyα) vary if the underlying relationship is a power law, in part because such a relationship is often suggested for higher-z collations between EW and impact parameter in systems comprising a single galaxy and single QSO sightline (see Section 8.1). Even though we have only four sightlines through NGC 1097, the sampling of multiple points in a single galactic halo is unique, and we therefore examine the results of fitting a power law to the NGC 1097 data.

Both EWs and column densities are assumed to be related as

Equation (1)

We fit the Lyα and Si iii EWs, and the distribution of Si iii column densities; we do not fit the H i column densities because we only have a range of values for two sightlines. We also take the Si iii EW and upper limit on column density toward HE0241 to be a detected value.

If a power-law fit to the data is justified, then the distributions do not vary smoothly with ρ/Rv. We can fit a power law by varying W* or N*, and α, and minimizing the χ2 statistic between the NGC 1097 data and values expected from Equation (1), but the reduced χ2 is ${\chi }_{\nu }^{2}\gg 1$ given the small errors relative to the overall dispersion in N and W. To account for this, we introduce an additional "error" that accounts for the clumpiness in the distribution. This approach is the same as that used to account for patchiness in absorption from O vi in the Milky Way described by Bowen et al. (2008) (after the work of Savage et al. 1990). For each value of W measured toward the QSOs behind NGC 1097 we introduce an additional relative error of σCL beyond that of the EW error σ(W) described in Section 5.2:

Equation (2)

For the given points, σCL can be varied until ${\chi }_{\nu }^{2}\,=\,1$. The values of σCL that we find are listed in the right-hand panels of Figure 12; for both the Lyα and Si iii EW, σCL is similar, around 0.2 dex (top two panels), which might be expected if the Lyα and Si iii absorbing regions are the same. For N(Si iii), σCL is a little higher, ≃0.3 dex. Of course, σCL need not be the same for EW and column density measurements, for while the two parameters are obviously related, the EW of a line represents both the column density of absorbing clouds and the dispersion between components comprising the line. In all three cases, σCLσ(W), as expected. With σCL fixed to the value that gives ${\chi }_{\nu }^{2}\,=\,1$, the corresponding values of W*, N*, and α are considered to be the best-fit values. These are again listed in the right-hand panels of Figure 12, and the resulting curves are shown in the bottom three panels of the left-hand column of Figure 12 as black solid lines. The values of α for N(Si iii) are identical, α = −1.4, which may reflect a more direct relationship between the simple, weak Si iii EWs and their column densities.

The errors in W* or N*, and α, can be estimated by calculating χ2 as these values move away from a minimum χ2. Full details are given in Bowen et al. (2008), which followed an identical procedure. To summarize: as W* or N*, and α, vary, contours in χ2 can be selected at particular confidence levels; in the right-hand column of Figure 12 we show the confidence levels at 0.38, 0.68, 0.95, and 0.997, which resemble the confidence intervals applied to a normal distribution. In that figure, we also list the values of W* or N*, and α, that correspond to the extreme deviations of the 1σ confidence interval for both parameters (the end point of the major axis of the contour); and for these pairings, we plot envelopes to the deviations in the left-hand panels of Figure 12 (shown in beige). Unsurprisingly, given only four data points, the errors are large. In fact, for one of the 1σ confidence intervals, the deviations are consistent with no correlation at all for the Lyα EW and N(Si iii).

8.1. Comparison to Other Datasets

In this section, we compare the properties of the absorption lines seen toward NGC 1097 with the properties of absorption lines seen toward galaxies in three other low-redshift studies. We first review the samples and galaxy types in Sections 8.1.18.1.4, and draw comparisons between the published results and NGC 1097 in Section 8.1.5. Our goal is to determine whether NGC 1097 is a typical QSOALS.

8.1.1. Comparison to COS-Halos

In Figure 12 of Section 8, we showed how Lyα and Si iii column densities and EWs changed with impact parameter. We also introduced a comparison set of data, taken from the COS-Halos program, and in this section we discuss these results further. The properties of Lyα, and of metal lines, from the absorbers were studied by Tumlinson et al. (2013) and Werk et al. (2013), respectively. To compare with our NGC 1097 data, we need to derive galactic virial radii in the same way as we have for NGC 1097. Tumlinson et al. list stellar masses for all their galaxies, derived from SDSS colors, from which we can estimate halo masses, and hence Rv, again using the relationships between the two described by Behroozi et al. (2013).

In Figure 13, the COS-Halos data are shown as light pink squares, while the Si iii values are purple squares. The H i column densities and EWs show a large dispersion, and include a set of sensitive non-detections at very small impact parameters, which Tumlinson et al. suggested was due to the lack of a cool halo around early type galaxies. These non-absorbing galaxies are only a small fraction of the total sample, and the bulk of the absorbers show unity covering fractions out to the ∼0.6 Rv limit of the survey. Our range of N(H i) values, and the inferred covering fraction, appears consistent with the COS-Halos sample.

Figure 13.

Figure 13. Same as Figure 12, now with the addition of values from published data sets, and with the x-axis extended beyond 1Rv. Data from COS-Halos are shown as light pink squares for H i (Tumlinson et al. 2013) and purple squares for Si iii (Werk et al. 2013). The black lines in the bottom three panels show the power-law fits to the NGC 1097 data alone from Figure 12; the dashed purple lines for Si iii show the fits from Werk et al. (2013) (where the slopes were found to be −0.84 ± 0.24 and −1.11 ± 0.29 for W(Si iii) and N(Si iii), respectively). The green line in the second panel represents a least-squares fit to the COS-Halos and LC14 data, ignoring their upper limits.

Standard image High-resolution image

For Si iii lines, Werk et al. fitted a power law to the distribution of N(Si iii) and W(Si iii) with ρ/Rv (ignoring both upper and lower limits). These fits are reproduced in the bottom two panels of Figure 13; their fits to N(Si iii) and W(Si iii) are close to what we find for the four sightlines toward NGC 1097. They also calculated a covering fraction for Si iii of ${73}_{-14}^{+9}$% over a radius of 0–160 kpc for galaxies with stellar masses log M*(M) ≥ 10.5. This again is consistent with the three or four detections in the halo of NGC 1097 at similar radii and at the same stellar mass limit.

In Figure 14 we introduce the detections and limits for C ii (orange points), Si ii (green points), and Si iv (blue points) for the COS-Halos galaxies. With C ii, Si ii, and Si iv detections toward only Q0244, and a possible C ii detection toward Q0246 (which we include) comparisons with NGC 1097 are tenuous. We note simply that the strengths of the metals toward NGC 1097 are similar to the COS-Halos sample. They lie in regions where the covering fractions of the metals in the COS-Halos sample are less than unity, and our non-detections are compatible with those low values.

Figure 14.

Figure 14. Distribution of total column densities summed over all components, and equivalents widths, for C ii, Si ii, and Si iv, as functions of impact parameter normalized by virial radius. Our data toward NGC 1097 are shown as red circles for detections, and pink circles for upper limits. Several other data sets are included for comparison, and their origin is shown in the legend in the first panel. Unfilled symbols represent upper limits, while lower limits (usually from mildly saturated lines) are indicated as filled symbols with the appropriate upward pointing arrow. The data of Liang & Chen (2014) contain a large number of EW upper limits beyond 1Rv, and to avoid confusion, we draw the median of those upper limits with a single solid gray line and indicate the range of their limits with two dashed gray lines. Individual upper limits at <1 Rv are still plotted. Errors in ρ/Rv are shown for the COS-Halos data, but to avoid confusion, they are omitted for the rest of the data sets. In most cases, uncertainties in Rv are similar.

Standard image High-resolution image

8.1.2. Comparison to ESO 157–49

Three background probes were used by Keeney et al. (2013) to probe the edge-on galaxy ESO 157−49, a much less luminous (L/L* = 0.12) and weaker star-forming galaxy (0.2–1.1 M yr−1) than NGC 1097. We assume a viral mass of log M(M) = 10.6 for the galaxy (the lower limit set by Keeney et al.), which corresponds to Rv = 88 kpc. This puts their sightlines at impact parameters of ρ/Rv ≃ 0.9, 1.1, and 2.0, distances larger than those probed for NGC 1097. We adopt the sum of the three N(H i) components toward the outermost sightline, HE 0435–5304, to be log N(H i) = 14.22 ± 0.11 and calculate the total EW by reconstructing the line profile from their Voigt profile fits, W(Lyα) = 0.63 Å. The Lyα lines toward the inner two sightlines are of similar strength to the inner two sightlines toward NGC 1097, and Keeney et al. had the same problem in determining N(H i) as we have had, with moderately strong W(Lyα) yielding ambiguous column densities. Figure 13 shows these points compared to the NGC 1097 values; although the errors are large, N(H i) (filled cyan squares) appears to continue to decline with ρ/Rv beyond 0.9Rv. Si iii is detected toward the inner two ESO 157–49 sightlines, beyond the last limit from NGC 1097. No C ii or Si ii is detected toward any of the sightlines (open cyan squares) but Si iv is seen at distances beyond where the COS-Halos or the NGC 1097 data probe.

With the absorbers at either side of the major axis of the edge-on galaxy both having negative velocities relative to the galaxy, Keeney et al. concluded that the absorption (at least for one of the sightlines) was unlikely to be associated with any galactic rotation, and that it was more likely to have arisen in material once ejected from the galaxy. This interpretation is clearly quite different from the ones discussed in Section 7 that were designed to account for the absorption toward NGC 1097.

8.1.3. Comparison to M31

M31 offers an interesting comparison with NGC 1097 in that M31 has a luminosity and halo mass close to that of NGC 1097 (van der Marel et al. 2012a, 2012b). The one important difference is that M31 is interacting with another galaxy of comparable mass only 0.8 Mpc away, with the same halo mass and luminosity—namely the Milky Way; as discussed in Section 2.3, there are no galaxies as bright as NGC 1097 within 4–5 Mpc. Although there are a large number of QSOs that can be found at interesting impact parameters behind M31, the low velocity of the galaxy makes it difficult to study in absorption (Section 1). Nevertheless, Rao et al. (2013) found no absorption beyond 0.2 Rv along the major axis of the disk (assuming Rv = 300 kpc for M31), but low-ionization absorption from four of six sightlines below 0.1 Rv. These data are hard to incorporate in this study because (i) the low resolution of the COS G140L and G230L observations makes it difficult to fully separate out absorption from the strong MW ISM absorption lines, (ii) understanding the relationship between the velocities of the detected lines and the velocity of MS+HVC gas (if any) at the positions probed is beyond the scope of this paper, and (iii) both the Lyα and Si iiiλ1206 lines, which are integral to our study, are not available in their low-resolution data. For these reasons, these data are not plotted in Figures 13 and 14. Nevertheless, Rao et al. drew an important distinction in claiming that M31 is not a typical Mg ii absorbing galaxy compared to higher-redshift samples.

On the other hand, Lehner et al. (2015) used COS G130M and G160M archival data of 18 sightlines passing within 2Rv of M31. They distinguished a specific velocity range over which to measure absorption that they considered sufficient to avoid contamination from MW ISM absorption and the MS. These data are included in Figures 13 and 14 (shown as triangles). Lehner et al. did not provide EW measurements from their data, and so we estimate EWs directly from the column densities, assuming the lines are optically thin. These represent lower limits to the true EWs because not only may some of the lines be mildly saturated, but also because the wings of the absorption profile are often lost in the blended MW ISM and the MS lines either side of the absorption from M31.

8.1.4. Comparison to Intermediate-redshift Sample

Like the COS-Halos sample, the collation by Liang & Chen (2014) is of absorption along single sightlines through many galaxies. Nevertheless, the sample is useful in that the sightlines probe beyond the impact parameters of the COS-Halos galaxies, and measurements of many species of metal lines, as well as Lyα, are provided. This sample of galaxies (hereafter designated as LC14) has a median redshift of 0.02 and samples galaxies with stellar masses of 107–1011 M. Their derivation of Rv is the same as described above, so we use their tabulated values. Only EWs are given by Liang & Chen, and since these cannot be converted to column densities with any reliability, column densities are not plotted. In Figure 13 all the data for Lyα and Si iiiλ1206 are included, but in Figure 14, most points at ρ/Rv > 1.0 are only upper limits, and not sensitive enough to show any significant decline in EW above those radii. To avoid confusion with points from the other data sets, we show only the median EW limit at ρ/Rv > 1.0 as a gray horizontal bar.

8.1.5. Is NGC 1097 a Typical Absorber?

Our fitting of a power law to the data obtained for NGC 1097 shows that the evidence for a decline in H i and Si iii column densities and EWs with ρ/Rv—based only on the four sightlines—is weak. Tumlinson et al. reported no decline in N(H i) or W(Lyα) with ρ/Rv in the COS-Halos sample, which we see recreated in the top panel of Figure 12. However, Figure 13 shows that when we combine the COS-Halos results with the LC14 data it seems more plausible that Lyα and Si iiidecline with impact parameter, and indeed that NGC 1097 follows the same decline. If we exclude the upper limits in the COS-Halos and LC14 data sets, then a linear regression analysis of the two larger samples confirms a strong correlation, with a high probability of being able to exclude the null hypothesis that there is no correlation. A least-squares fit to the W(Lyα) data (shown as a green line in the second panel of Figure 13) has logW* = −0.37 ± 0.02 Å, α = −0.45 ± 0.04; but more importantly, the standard error in the regression, s, is 0.27 dex, while for NGC 1097, s = 0.21 dex. A similar result is found for W(Si iii). This standard error is the average distance that the observed values fall from the fitted line, and conventionally, 95% of all points should fall within 2s. The fact that the values of s for NGC 1097 and the two samples of COS-Halos and LC14 are similar suggests that the overall variability or clumpiness within the halo of NGC 1097 is largely the same as the dispersion seen from the galaxies in the larger single-probe data sets.

Of course, while the Lyα and Si iii absorption in the halo of NGC 1097 may have much the same characteristics as the other data sets, we only know this to be true for impact parameters <0.6Rv. Beyond 1Rv, LC14 have shown that the number of non-detections of Lyα increases rapidly, i.e., the covering fraction declines to ∼60% at an EW threshold of 0.05 Å (their Figure 11). Keeney et al.'s observations of ESO 157–49 do not show such a decrease (and remain as flat as the values for NGC 1097 at smaller ρ/Rv). Whether this represents the way H i is distributed around all galaxies, or whether the detection/non-detection of Lyα points to some other property of a galaxy that influences the extent of H i around its CGM, is not clear. Observations of additional probes beyond 0.6 Rv for NGC 1097 would help to explain how the covering fraction really declines in a single halo at large radii.

We conclude that based on the absorbing column densities and EWs alone, NGC 1097—and probably M31 and ESO 157–49—are "typical" absorbers compared to the COS-Halos and LC14 samples. This is curious, because the structures that have been suggested as origins of the absorbing gas (an isolated galaxy for NGC 1097, an interacting group for the MW–M31–MS system, and post-ejection products from ESO 157–49) are quite different. It may be that while galaxies do indeed possess definable CGMs that decline with radius, some absorption line properties—in this case N and EWs—are relatively insensitive to the structures that contain the baryons. For NGC 1097 we have also used kinematics to attempt to better understand the origin of absorbing gas, leveraging the use of multiple probes of a single galaxy to constrain a model (an approach that is far more ambiguous for single sightlines through single galaxies). Again, though, a similar consideration of the kinematics of the three sightlines toward ESO 157–49 suggests an origin quite different than that for NGC 1097, and it remains to be determined if either set of results is universal for galaxies.

9. DISCUSSION

9.1. Likely Origin of the Absorbing Gas

We have shown in Section 7.1 that most of the absorption line profiles can be reproduced by a model in which gas is distributed in an extended rotating disk whose geometry matches that of the inner H i disk of NGC 1097. This implies that we are able to trace the rotation curve of the galaxy out to distances of 0.6 Rv. Figure 15 shows the 21 cm rotation curves obtained by Ondrechen et al. (1989) and Higdon & Wallin (2003) as functions of $\tilde{\rho }/{R}_{v}$, in comparison to the Lyα lines toward the background QSOs. Figure 15 shows how velocities would decline for a Keplerian rotation; an extended disk model is only marginally consistent with being rotationally supported if there were no mass exterior to the inner H i disk (i.e., beyond ∼40 kpc). Obviously, this is at odds with ΛCDM theory in general, and specifically, with the NGC 1097 system, where a large DM halo is required to produce the stellar streams (Higdon & Wallin 2003; Amorisco et al. 2015). Instead, we are likely seeing absorption from an extended disk with a cool gaseous component, probably of low metallicity, falling into the inner disk of NGC 1097. Geometrically thick quasi-disk gas distributions that are accreting onto a central object from large scales have been shown to occur naturally in simulations of structure formation (Kaufmann et al. 2006; Stewart et al. 2013; Danovich et al. 2015, e.g., their Figure 13). There is evidence that such cold accretion inflows are now being detected at high redshift (Martin et al. 2015).

Figure 15.

Figure 15. Top: rotation curve for NGC 1097's inner disk from 21 cm data (pink and blue circles), compared to the absorbers in the CGM (gray squares) assuming that the absorbing gas is in the form of a disk and has the same inclination and position angle as the inner disk. All values are corrected for inclination effects. The sightline toward Q0244 leads to constraints on vrot that are larger than the scale plotted on the y-axis, (because the sightline lies along the minor axis so has only a very small Doppler shift, and because the absorption is very broad) and so the sightline is excluded. The dashed line shows the expected vrot if there were no dark matter beyond 25 kpc, and the black solid curve is a rescaled version of the galaxy–halo rotation curve of Seljak (2002). The vertical red line marks the expected virial speed for NGC 1097. If the absorbing gas really is in the configuration of a disk, the fact that vrot is below the DM model expectations may imply either that the gas is infalling or that the disk is warped at large $\tilde{\rho }$. Bottom: specific angular momentum of the gas. The dotted line shows a polynomial fit to the data.

Standard image High-resolution image

Figure 15 also shows that the specific angular momentum of the absorbing gas along the axis perpendicular to the inclined disk ($\tilde{\rho }\,{v}_{{\rm{rot}}}$) is consistent with that deduced from the 21 cm observations at the outer edges of NGC 1097's disk. This implies that the accreting gas is infalling and does not lose angular momentum until it joins the inner halo at ∼0.3 Rv. This agrees with initial semi-analytic works on galaxy formation (e.g., Fall & Efstathiou 1980) as well as the more recent simulations of Danovich et al. (2015) (see their Figure 20) that explain the origin of angular momentum of galaxies by the deposition of gas from cold inflowing streams.

Our interpretation suggests that the spin parameter of the extended inclined disk (Dutton & van den Bosch 2012, and references therein) is ${\lambda }_{\mathrm{extended}\mbox{--}\mathrm{disk}}\simeq 0.15$ $(\tilde{\rho }/100\,{\rm{kpc}})$ $({v}_{{\rm{rot}}}/70$ km s−1). This is about four times larger than the typical spins values of DM halos (Dutton & van den Bosch 2012; Genel et al. 2015) and is consistent with the theoretical expectations for the spin of cold dense gas that originates from larger scales and accretes onto a halo.

Conversely, there is little obvious evidence to suggest that the absorption seen along the four QSO sightlines arises in outflowing gas from NGC 1097. Our attempts to replicate the kinematics of the absorption do a poor job when only outflowing gas is included, compared to the inflowing/rotating thick disk model. Nevertheless, adding an outflow to the thick disk model helps to account for the A3 component toward Q0244. The primary constraint on such an outflow is that the opening angle must be large.

Moreover, galactic outflows are expected to have metallicities of the order of the solar value, while our investigations suggest that the absorbers might have metallicities less than this (Appendix C). The A3 component toward Q0244, again, is a possible exception, since we are unable to constrain its metallicity from photoionization considerations.

Finally, there is the issue of tidal debris around NGC 1097. As discussed in Section 2.2, NGC 1097 is interacting with an LMC-like galaxy (NGC 1097A), and has seen the passage of another, now destroyed, satellite galaxy that has given rise to the dog-leg stellar trail shown in Figure 1. It has been suggested that galaxy mergers and interactions are an effective way of distributing gas out to large distances and relative velocities from galaxies, and that complex, multicomponent absorption line systems found in QSO spectra could be probing such debris (Bowen et al. 1994; Keeney et al. 2011; Fox et al. 2014; Rosenberg et al. 2014). The three outer QSO sightlines toward NGC 1097 show no obvious evidence for complex multicomponent structures spanning many hundreds of km s−1, which we might use to argue for the presence of tidal debris. The sightline toward Q0244, however, lies only ≃15 kpc from Knot-A (at least, in projection), and only ≈10 kpc from the limit of the trail discernable on the image. Components A1–A3 span only 157 km s−1, but if the very weak positive-velocity Lyα components (A4 and A5) are included, the absorption covers at least 260 km s−1. Indeed, the spread in velocity from components arising in tidal debris will depend strongly on the orientation of the debris structure. Structures that are roughly confined to a plane that is parallel to our sightline will show components that have velocities determined only by internal random motions rather than the bulk flow. Hence complex absorption lines may not be a necessary signature for tidal debris, particularly for minor mergers, where the disk of the more massive galaxy might be less perturbed.

Higdon & Wallin (2003) found no H i associated with the streams to a limit of 0.06 M pc−2 or 8 × 1018 cm−2; at the position of the right angle of the dog-leg trail, they set a limit of 0.021 M pc−2 or 3 × 1018 cm−2 in a 4 arcsec aperture. (This is consistent with the value of N(H i) found toward Q0244.) Higdon & Wallin's N-body simulations suggest that X-shaped tidal tails develop only after a second or third pass of the accreting galaxy through the host (1.7–3.8 Gyr from the first collision), and that the dwarf's ISM will have been destroyed by ram pressure from the disk. In that case, there may be no remnants of cool gas associated with the dwarf or its interaction with the host galaxy. Our absorption line detections toward all four QSO sightlines, however, are sensitive to the much lower column densities that may arise in more highly ionized gas, and it remains possible that the absorption we see is related to the fate of the gas from the cannibalism of the dwarf galaxy.

We have seen in Section 7.1.2 that while a rotating, infalling disk can explain most of the kinematics of the absorption lines from NGC 1097, an additional model component of a conic outflow is needed to explain the A3 component to the absorption. It is possible that, instead, this component—and the weak A4 and A5 Lyα lines—come from tidal debris associated with the stellar trails. This may help to explain the difficulty in finding suitable solutions for photoionized gas in Appendix C.1.2 for A3. The models of the debris constructed by Higdon & Wallin (their Figure 13) show that stellar material ought to be distributed out to the sightlines of Q0244 and S393Z082, but, as expected, the covering fraction would probably depend quite strongly on the galaxy's orientation to the line of sight. Modeling the evolution of gas associated with tidal mergers is beyond the scope of this paper. While the kinematics of the gas and the arguments made above show consistency with the models described, it is worth emphasizing that with only four sightlines the correspondence could be fortuitous.

9.2. Comparisons with Simulations

As argued above, there are good reasons to believe that most of the absorbing gas we see in the CGM of NGC 1097 comes from material that is being accreted onto the galaxy from the outskirts of the halo. We conclude by comparing our data with one further set of simulations that evolve galactic structures down to z = 0, those of van de Voort & Schaye (2012). Their calculations include radiative cooling, sub-grid star formation, and feedback prescriptions. Approximately 500 simulated halos are used to quantify the statistical properties of the halo gas. Conveniently, their results include a halo mass that is consistent with that of NGC 1097. Two "modes" of gaseous structures are identified, a "cold-mode" component in the form of cool <105.5 K gas filaments, and a "hot-mode" component that contains a volume-filling gas component. Our data imply gas temperatures of <105 K, and so would be related to the cold-mode structures in the halo.

Figure 16 compares the results from the simulations with our analyses of the absorption systems. We find good agreement with many of the variables recorded in the simulations as they vary with ρ/Rv: the surface density profile is consistent with the model within the (non-negligible) uncertainties; the deduced gas temperatures coincide with the simulated values; and the absorbing gas metallicities also show good correspondence with the model.

Figure 16.

Figure 16. Comparison between the results for the absorbers toward NGC 1097 and those predicted by the simulations of cold-mode gas accretion at z = 0 made by van de Voort & Schaye (2012). The dashed blue lines mark the median of the property from their simulations, while the blue shaded regions mark the 16th to 84th percentiles around the median. Triangles mark upper or lower limits, while vertical bars mark the range of values, found from our analysis of the absorbers. The simulations calculate distances in real (3D) space, while we only probe projected impact parameters ρ; as shown in previous plots, however, ρ and the deprojected values $\tilde{\rho }$ are similar for all but the outermost sightline, and we plot our data in terms of ρ/Rv. Panels (a) to (e) show the gas density, temperature, metallicity, inflow velocity, and mass accretion rate assuming azimuthal symmetry in the plane of the disk. In panel (c) we plot inner disk metallicities of NGC 1097 (gray circles) taken from Moustakas et al. (2010) (using the normalization from Pilyugin & Thuan 2005). For the mass accretion rates in panel (e), van de Voort & Schaye do not quote estimates in the scatter of $\dot{M}$, so no blue shaded region is drawn. In panel (d) two sets of limits exist, either for inflowing or outflowing gas, while the vertical bars indicate the allowed inflow velocity range based on the rotating–inflowing disk model (Section 7.1.2). Overall, there is good agreement between the absorber properties deduced in this paper and these particular simulations.

Standard image High-resolution image

The comparison between gas kinematics is shown on the bottom panel of the first column of Figure 16. It is impossible to distinguish an outflow that is launched away from us from the more distant side of a disk, from an infall of gas toward the side of a disk nearest to us. If we assume that the absorbers are falling toward the center of NGC 1097, then the absorption velocities are consistent with values in the simulation. We note, however, that we only measure line-of-sight velocities, with a radial velocity component being potentially larger.

We can crudely estimate the mass infall rate from our data as a function of galactocentric radius, by assuming azimuthal symmetry: $\dot{M}(\tilde{\rho })\,\simeq \,\pi {\tilde{\rho }}^{2}N$(H)$(\tilde{\rho })({v}_{{\rm{rad}}}/\tilde{\rho })$. If vrad is independent of $\tilde{\rho }$, we find $\dot{M}(\tilde{\rho })$ rises monotonically with $\tilde{\rho }$, as predicted in the simulations. Specifically, if we take a fiducial value of vrad = 70 km s−1, there is qualitative agreement between a mass infall rate of ≃1 M yr−1 at small $\tilde{\rho }/{R}_{v}$ and the current star formation rate, despite the ≳1 Gyr dynamical timescales implied.

Figure 16 suggests good agreement between theoretical simulations and the absorption line data for NGC 1097. Intriguingly, van de Voort & Schaye noted that in their simulations, the probability of detecting these cold streams was low. Similarly, Goerdt et al. (2012, their Figure 4) calculated a covering fraction of only ∼20% for cool gas within 1Rv at z = 1.4, with the value dropping with time. In the case of NGC 1097, and despite the small numbers of sightlines used, the covering fraction at log N(H i) ≳ 14 is of the order of unity out to ≃0.6 Rv. While it is possible that NGC 1097 may be unique in some way, there are no clear indications from the absorption line data that this is so. In fact, in Section 8.1.5, we argued that NGC 1097 was a "typical" absorber, in comparison to the ensemble of higher-redshift absorbers. If our results are not entirely due to a fortuitous combination of sightlines, then the apparent discrepancy could be related to problems with the simulations, a topic that is beyond the scope of this paper.

10. SUMMARY

In this paper, we have observed four QSO whose sightlines pass through the halo of NGC 1097: Q0244−303 (referred to as Q0244 in the text), 2dFGRS S393Z082 (S393Z082), Q0246−308 (Q0246), and HE 0241−3043 (HE0241). These sightlines lie at impact parameters of 48, 84, 150, and 165 kpc, or ≈0.2, 0.3, 0.5, and 0.6 times the virial radius of the galaxy, which we take to be Rv = 280 kpc. We can summarize our results as follows.

  • 1.  
    We detect Lyα absorption toward all four sightlines. The total EWs range from ≃0.3–1.4 Å, but the column densities toward the inner two sightlines are difficult to determine due to saturation of the lines. Nevertheless, they all have N(H i) ≥ 14.0, which implies that the cross section of NGC 1097 is unity out to 165 kpc (0.6 Rv) at that limit. Si iiiλ1206 is detected toward the three inner sightlines, with W = 0.13–0.39 Å, all lines having log N(Si iii) > 13.0, again suggesting a covering fraction of unity at 150 kpc (0.5 Rv) at this column density.
  • 2.  
    The strongest Lyα absorption, and the one that shows the most metal lines (including C ii, Si ii, Si iii, and Si iv lines) occurs along the innermost sightline, Q0244−303 at 48 kpc (0.2 Rv), and comprises a minimum of two components. A lower limit for the integrated column density of log N(H i) ≥ 15.7 is robust, but the upper limit depends on whether a third component also exists, which is defined by the presence of a single C ii component and no other metal lines. For a two-component model, we establish log N(H i) ≤ 17.1, but if a third component exists, we cannot rule out a column density as high as log N(H i) ≲ 18.1. It is therefore possible that the absorption along this sightline is equivalent to a higher-redshift Lyman limit system. This may also be true for the second innermost sightline, 2dFGRS S393Z082 (ρ ≃ 0.3 Rv), because although the Lyα line appears to consist of a single component, a fit with a high N(H i) and low b profile is indistinguishable from a solution with lowN(H i) and high b.
  • 3.  
    As noted above, Si iv is detected only along the innermost sightline, with log N(Si iv) = 12.9. Beyond this, Si iv is absent at a level of log N(Si iv) ≃ 12.7–13.1, implying a small covering fraction at these limits beyond 48 kpc (0.2 Rv).
  • 4.  
    There is only marginal evidence that N(H i) and N(Si iii) decline with galactocentric radius, and the distribution of gas is clearly inhomogeneous within the CGM. We have fitted a power law to the decline in column density with impact parameter, and for W(Lyα) and log N(Si iii), the errors in the fit show that there could be no anticorrelation at all. An anticorrelation in W(Si iii) with ρ appears more plausible. In all cases, the CGM appears clumpy with respect to any possible smooth decline with ρ, and we characterize the clumpiness to be more than ∼0.2 dex for an assumed power-law decline of W and N with ρ for both Lyα and Si iii.
  • 5.  
    We have modeled the observed column densities and column density limits using CLOUDY, assuming that the gas is photoionized. There are considerable uncertainties in the results, either because N(H i) is not well known and the SED of NGC 1097 is difficult to account for—as in the case toward Q0244−303—or because few metals are actually detected (as toward the outer sightlines). The simplest solutions suggest that the gas has metallicities that are sub-solar, if the gas is dust-free.
  • 6.  
    We have attempted to replicate the observed absorption lines using a suite of models for the CGM around NGC 1097, including gas that is rotating in a huge extended disk, inflowing from the IGM, or outflowing from the galaxy. None of the models replicates the observations exactly, but one that has a disk of gas rotating roughly planar to, but more slowly than, the inner disk observed from 21 cm data, and which also includes gas infalling from the IGM, recreates the data well. The models that provide the poorest predictions of the data are those dominated by conical or spherical outflows. The exception to this statement arises in modeling the absorption toward Q0244−303, where the most positive velocity component requires additional absorption either from a broad outflow or from tidal debris.
  • 7.  
    While these models provide a plausible insight into how the CGM may be distributed around NGC 1097, it is also possible that some of the absorption could arise from ionized gas associated with the stellar streams that exist around the galaxy as a result of a multiple-bypass minor merger. Gas associated with a cannabalized dwarf galaxy may have been stripped during its encounter, but its remains may contribute to the observed absorption.
  • 8.  
    In comparison only to the column densities and EWs observed from the CGMs of other galaxies obtained from recent COS surveys, the absorption from NGC 1097 seems typical. Given the variations in these quantities in the single halo of NGC 1097, however, it may be that column density and EW measurements are not a particularly sensitive way to disentangle how the origin of CGMs depends on the properties of the host galaxies. Our data demonstrate the necessity to observe absorption lines from the nearest local galaxies where the CGM can be defined a priori from multiple techniques, and that multiple probes of single-galaxy CGMs is now a priority.

Big thanks to Gero Rupprecht for help with the calibration of ESO Archive data, to Katy Oldfield for her expertise with the 3D CAD software package SolidWorks, and to Rick Stevenson for permission to use his image of NGC 1097. For the work required to identify QSOs behind galaxies, D.V.B. was supported by a NASA Long Term Space Astrophysics (LTSA) grant NNG05GE26G, and by NNG05GE31G for the analysis of GALEX Cycle 1 data; funding for the reduction of the COS spectra obtained from the GO program was provided by NASA grant number 12988 from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. The research by D.C. is partially supported by FP7/IRG PIRG-GA-2009-256434. Computations presented in this paper have made use of the Hive computer cluster at the University of Haifa, which is partly funded by ISF grant 2155/15.

Extensive use has been made of archival databases in this research: the SuperCOSMOS Sky Survey material is based on photographic data originating from the UK, Palomar, and European Southern Observatory (ESO) Schmidt telescopes and is provided by the Wide-Field Astronomy Unit, Institute for Astronomy, University of Edinburgh; the 2dF QSO Redshift Survey (2QZ) and the 2dF Galaxy Redshift Survey (2dFGRS) were compiled by the 2QZ and sdFGRS survey teams from observations made with the 2-degree Field on the Anglo-Australian Telescope (AAT); the NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA; and the SIMBAD database is operated at the Centre de Données astronomiques de Strasbourg (CDS) in Strasbourg, France.

APPENDIX A: NOTES ON INDIVIDUAL QSOS AND GALEX GRISM OBSERVATIONS

As noted in Section 3, we observed NGC 1097 as part of a GALEX Cycle 1 Guest Investigator program (Program 085) designed to obtain FUV grism spectra of UV-bright objects in nearby galaxy fields with the aim of discovering new extragalactic sources. In this appendix we summarize these observations, as well as remark on the original discoveries of QSOs in the field of NGC 1097 that made the observations discussed in this paper possible.

Many of the galaxies with the largest angular diameter in our original list of target fields were to be observed as part of the GALEX Nearby Galaxy Survey (NGS) (Gil de Paz et al. 2007). This made it possible to add grism observations as part of a GALEX SNAP program without needing to acquire long UV images beforehand. We observed NGC 1097 for 1786 s on 2004 October 26 as part of this program, and a total of 117 grism spectra were extracted with good quality data in both FUV and NUV channels. Of these, 80% showed spectra with very irregular continua, decreasing to zero flux below 2000 Å, indicative of the SED of stars. Of the remaining objects with discernable FUV flux, nine already had published identifications, leaving three unidentified. These three unfortunately showed no discernable emission line features and could not be added to our list of potential targets. Hence for this field, we were not able to discover any new UV-bright AGNs or QSOs near NGC 1097.

Even though our selection of QSOs behind nearby galaxies was constructed by first selecting UV-bright objects from GALEX data (Section 3), our eventual choice of targets relied on them having extant redshifts. While NED catalogs the majority of redshifts from all published redshift surveys, it turns out that for NGC 1097, much of the redshift information comes from surveys conducted three decades ago by Wolstencroft et al. (1983) and Arp et al. (1984). The apparent overdensity of X-bright QSOs in the field, and a hypothesized alignment of some QSOs with the galaxy's jets, made this field of particular interest to the latter author (Arp 1998, p. 43; Arp & Carosati 2007). Several redshifts were measured for objects in the field of NGC 1097 as part of that research.

Q0244 was originally discovered as a z = 0.528 X-ray source by Wolstencroft et al. (1983) (their "Q1097.2"); Arp et al. (1984) labelled this as "#27" in a follow-up paper. It was later recovered by the 2dF QSO Redshift Survey (Croom et al. 2004) as 2QZ J024649.8−300742, which confirmed the original detection of Mg ii in emission, as well as several Balmer emission lines. We detected C iv emission at the correct wavelength in our GALEX grism data.

S393Z082 was detected as part of the 2dF Galaxy Redshift Survey (2dFGRS) (Colless et al. 2001), where broad, weak Hβ, Hγ, and Mg ii were clearly detected at z = 0.339. Equally narrow and weak Lyα emission was also detected in our GALEX grism data at the expected wavelength.

Q0246 was discovered as an X-ray bright object by Arp et al. (1984), their "#35." Their redshift of z = 1.09 was determined unambiguously from optical data, and we detected strong and broad Lyα emission in our GALEX grism data.

HE0241 was discovered by Wisotzki et al. (2000) as part of the Hamburg/ESO survey for bright QSOs. We obtained no GALEX grism spectra for the QSO, since it lies just outside the GALEX field centered on NGC 1097.

S394Z150 was identified as an AGN in the 2dFGRS with z = 0.131, and we selected the target based on the cataloged redshift and measured FUV magnitude. No spectrum was supplied in the original grism catalog produced after the GALEX data were taken, but more recent GALEX Data Releases now provide a grism spectrum for the object. It shows the flux rising toward the FUV at levels predicted from the FUV magnitude, but is featureless. After obtaining the HST COS spectrum, it became apparent that the object is not extragalactic, and is likely to be a local white dwarf star, based on the extremely strong DLA absorption profile at z = 0 and the weakness of any metal lines that would normally be expected from a sightline passing through the Milky Way. A retrospective examination of the optical spectrum from the 2dFGRS database shows that no emission or absorption lines were detected, making the original redshift measurement erroneous.

One final note: Wolstencroft et al. (1983) found a relatively bright X-ray QSO, J024645.60−300051.9 (their "Q1097.3"), with z = 1.00. We recovered this QSO independently from our GALEX grism data, with Lyα emission clearly visible at the expected redshift. However, there is no FUV flux below a strong discontinuity at 1900 Å, which could be due to Lyman limit absorption at the redshift of the QSO itself. For this reason, this QSO was never a potential HST COS target.

APPENDIX B: NOTES ON NGC 1097B

Figures 1 and 3 also show the position of NGC 1097B. This faint galaxy was reported by Higdon & Wallin (2003) to have a 21 cm velocity of 1015 ± 11 km s−1. (The position of the galaxy appears to be mis-reported in their Table 1 but the position discussed in their text is only 13'' from the optical position of the 2dFGRS source.) It was also observed as part of the 2dFGRS, designated S393Z052 with bJ = 18.1, and assigned a redshift of z = 0.0645 (≃19,300 km s−1), but there is little evidence from the galaxy spectrum that this is correct. Indeed, although the spectrum is of low S/N, there is better evidence for a 4000 Å break and, perhaps, Hα emission, close to the 1015 km s−1 given by Higdon & Wallin (2003). The galaxy has an absolute magnitude M(bJ) = −12.5 if at the redshift measured by Higdon & Wallin, reminiscent of the ultracompact dwarf galaxies (UCDs) found by Phillipps et al. (2001) in the Fornax cluster, but NGC 1097B does not look as compact as those UCDs. It is possible that NGC 1097B constitutes a third satellite of NGC 1097, although more accurate redshift information is obviously needed. There is no evidence for any absorption lines at the velocity of 1015 km s−1 and there is no reason to suppose that the galaxy contributes directly to the detected absorption. It would, nevertheless, be of interest to determine whether this is indeed another dwarf galaxy in the halo of NGC 1097.

APPENDIX C: NOTES ON PHOTOIONIZATION CALCULATIONS FOR THE ABSORPTION SYSTEMS

All of the fits in Tables 36 are based on allowing the b-values of individual species to vary independently of one another. (The only parameter in common for all the lines is their fixed velocities.) An alternative approach might be to fit all the species, demanding that the same Doppler parameter be used, an approach that would be acceptable if all the linewidths were dominated by nonthermal processes. However, at certain temperatures, the widths of Lyα lines and the widths of metal lines can be dominated by different processes: it is well known that the width of a line is usually represented by summing, in quadrature, the width expected from thermal broadening by gas at a temperature T, btherm, with a width that describes a turbulent velocity, bturb, such that the total width in km s−1 is ${b}^{2}\,=\,{b}_{{\rm{therm}}}^{2}+{b}_{{\rm{turb}}}^{2}$, or for an ion of atomic mass A,

Equation (2)

assuming that the turbulent width can be adequately described by a Gaussian. (Here, as usual, kB is the Boltzmann constant and mp is the mass of the proton.) Hence if the kinetic temperature of H i is, e.g., 3000 K, the thermal width of a Lyα line is 7 km s−1, while for metals such as, e.g., silicon, btherm is only ∼1 km s−1. So while the thermal widths of the metal lines are largely unresolved at these temperatures in COS data, the thermal width of H i can be a significant fraction of the total linewidth as T increases.

While btherm and bturb cannot be determined independently, we can make a series of fits to the H i and metal lines together, assuming a set of temperatures that fixes btherm for each line, and which lets bturb vary, while always constructing a total width from Equation (2) (which is used for fitting the observed profile). This yields a grid of column densities and values of bturb for each T (or btherm). Uncertainties on each model can be calculated using the same Monte Carlo procedures described in Section 5.1. An example of the results from this multiple-T fitting is shown in Figure 17 for both components of A1 and A3 toward Q0244. This method enables us to better understand the range of column densities—particularly N(H i) in this complicated system—that are likely causing the absorption. The only drawback is that we need to assume that the temperatures of the two components (i.e., the btherm values) are the same. This assumption need not be true, but so long as btherm is not considerably different for the two components, the results remain largely valid.

Figure 17.

Figure 17. Voigt profile fits to the lines toward Q0244, assuming that the gas has a range of thermal temperatures. Left: simultaneous fits to both the Lyα and Si iiiλ1206 lines for a two-component model when the b-value is determined by a combination of turbulent and thermal broadening, with the thermal component btherm fixed by the four indicated temperatures T. (The thermal widths are by necessity assumed to be the same for both components.) Although only two species are shown here, the profiles are a result of simultaneous fits to all available lines. The dotted profiles indicate the individual components A1 and A3 that comprise the fit for btherm = 1 × 103 K and are shown only to highlight the assumed two-component nature of the line. Fits to the high-velocity weak Lyα lines (A4 and A5 in Figure 5) are not shown here. The profiles demonstrate that the absorbing H i cannot be at temperatures T ≳ 105 K. Right: variations in the column densities when btherm—and hence bturb—changes with T. Red points are for component A1 and blue points are for A3. The metal column densities are largely insensitive to different temperatures, but N(H i) may change by ∼1 dex for A3, making it difficult to reliably constrain the physical properties of the system. The purple-shaded regions mark the temperature range where the CLOUDY predictions discussed in Appendix C.1 match the observed values. Bottom right: the deduced turbulent velocities of the gas are shown for both A1 and A3, while the thermal speeds for H i are shown as black circles.

Standard image High-resolution image

Further, we can use these results to construct models for photoionized gas of density n and metallicity Z for each value of T. We do this using the CLOUDY software package (Ferland et al. 2013, version C08.01). The analysis assumes that all abundances are relative to the total hydrogen density and are scaled to each other in a solar abundance pattern; the solar values for carbon and silicon come from Allende Prieto et al. (2002) and Holweger (2001), respectively. Unless otherwise stated in the sections that follow, the models are dust-free. We adopt a UV background based on the metagalactic field of Haardt & Madau (1996, 2012, hereafter HM), but discuss the possible contribution from NGC 1097 itself when appropriate. The total hydrogen column density N(H) is obtained by scaling the solution to the measured N(H i) at that temperature, and the absorber scale length l = N(H)/n is deduced.

In the following sections we discuss the results of this analysis for the metal line components toward three of the sightlines. For HE0241, the two absorption components D1 and D2 (Table 6 and Figure 9) are only detected in H i, and constraints on their physical properties are of little use. Hence no results are reported for these.

C.1. Absorption System toward Q0244

This system is the most complicated of the four studied. Our analysis is restricted to the A1 and A3 systems listed in Table 3 and labelled in Figure 5. Component A2 has no unique H i column density (and may not be real) while components A4 and A5 show only weak Lyα and no metal lines. Figure 17 shows that N(H i) for component A1 remains largely constant for all values of btherm, log N(H i) ≃ 15.6. For component A3, log N(H i) = 14.8 for T ∼ 5 × 103 K but decreases at higher temperatures. The metal-line column densities remain constant for all values of btherm. We should emphasize, however, that this apparently well-behaved change in N(H i) with temperature is strongly dependent on a likely oversimplified two-component model: obviously, it is always possible that the entire Lyα complex could be made up of many more components with high N(H i) and low b that are blended together, and which we are unable to separate.

C.1.1. Photoionization Models for Component A1

A single-phase solution (i.e., a single cloud with a uniform density) for this component that satisfies all the observational constraints is obtained for n ≲ 7 × 10−4 cm−3 and Z ≲ 0.3 Z, which yields log N(H) ≳ 18 (Figure 18). Taking the extreme values of this solution, the resulting length scale is 0.5 kpc ≲ l ≲ 10 kpc, which implies that the gas occupies only a small fraction of the volume as l/ρ < 0.2.

Figure 18.

Figure 18. Photoionization models for component A1 toward Q0244. Left: variations in temperature for different densities and metallicities, T = T(n, Z), for photoionized, dust-free, optically thin gas exposed to the metagalactic background radiation field under nonlocal thermodynamic equilibrium (non-LTE) conditions. Middle: the phase space of n and Z allowed by the different column density ratios. Solid curves mark ratios for which measured column densities were available for both ions, plotted for ±1σ variations as a result of the errors in the column densities. Dashed curves mark ion ratios for which only upper limits are available for one of the ions: upper limits to n and Z are marked in red while lower limits are shown in blue. The colored shaded region near the center marks the phase space that satisfies all the constraints, with the colors representing the total H column density (shown by the inset color-bar). For simplicity, only ion ratios that provide useful constraints are plotted. Right: the upper panel shows models for which the contribution of flux from NGC 1097 is included (see text) while the lower panel shows the effects of including Milky Way dust-to-metals mixtures (the latter yielding no solution). The colors of the lines in these panels are the same as those in the middle panel.

Standard image High-resolution image

The contribution of NGC 1097 to the flux ionizing the absorbing gas along this sightline may not be negligible, given the relatively small impact parameter of Q0244. Assessing the galaxy's contribution is difficult because the UV and X-ray flux that the absorbers see may depend on the galaxy's inclination relative to where the clouds are, on their distance and orientation to star-forming regions in the disk, and on any contribution from the central AGN (all of which are highly uncertain). The SED of the UV background, which is dominated by galaxies at z = 0, will likely have, to first order, a similar shape to that of NGC 1097, so the contribution from the galaxy can be accounted for by simply scaling-up the background flux level. In this case, a solution is obtained for the same ratio of photon density to particle density, and the allowed density range scales with the flux level (by roughly +0.6 dex given the luminosity of NGC 1097).

Alternatively, we can at least qualitatively assess the effects of assuming that the SED below the He ii edge is similar to that of the Milky Way (Fox et al. 2005), by using a flux below the He ii edge that is twice as high as that given by the HM SED (and which includes a contribution from both quasars and galaxies); the results are shown in Figure 18, where a single-phase solution can only be found if the upper limit to the Si iv column density is underestimated by 0.1 dex. The resulting model then has n ∼ 10−3 cm−3, Z ∼ 0.5 Z, log N(H) = 18.0 (l ∼ 0.3 kpc), which implies slightly higher gas metallicities and slightly more compact absorbers.

We also consider the possibility that the elements might be depleted onto dust grains. Adopting any depletion patterns that are similar to the Milky Way is difficult, since these are known to differ widely for, e.g., disk and halo gas (e.g., Savage & Sembach 1996). CLOUDY offers a compromise of using an average of values from both warm and cold gas, and we use these to see how our calculations might change with dust included. Ignoring any contribution to the UV flux from NGC 1097, using all the available column density ratios and limits, there is no single-phase solution. Only if we raise the upper limits on column density by a factor of 2, and use ±2σ intervals around the measured values, do we find a solution that has Z ∼ 3 Z and n ≳ 1 × 10−4 cm−3. We consider it unlikely, however, that the errors and upper limits for the metal lines in A1 are so badly determined by our analysis, and we reject these values.

C.1.2. Photoionization Models for Component A3

The analysis of this component yields no single-phase solution that can simultaneously account for all the observational constraints. The phase spaces for many line ratios are shown in Figure 19, and we see that none of the regions intersect. Considering the N(C ii)/N(H i) ratio alone (for example), the phase space available is completely ruled out because the allowed area is completely outside the region plotted in Figure 19 (unless Z > 10 Z). As discussed in Section 5.3, N(H i) is poorly constrained for this component (although apparently less so than A1), which certainly adds to the difficulty in finding a suitable solution. Nevertheless, even allowing for a wide range of N(H i), there is no single-phase solution that accounts for all the metal line ratios alone (Figure 19).

Figure 19.

Figure 19. Allowed n, Z phase-space regions set by the different ion ratios for the A3 system toward Q0244. The inset legend shows how individual ion ratios are colored in the figure. No single-phase solution solution exists for the model because no region exists where all the colored regions intersect. The C ii/H i ratio requires Z > 10 Z and so is not shown at the scales used for this plot (i.e., 10−2 Z < Z < 10 Z). Possible reasons for the poor agreement between model and observations include (a) that N(H i) is poorly determined, (b) that the relative metal abundances are different from the solar values, and/or (c) that multiple thermal phases are responsible for the absorption. If we consider only the silicon lines, a model with slightly super-solar abundances and densities of ≳10−4 cm−3 is implied.

Standard image High-resolution image

This problem remains even if additional flux is added from NGC 1097, or if the absorbing gas is assumed to be depleted by dust like the Galactic ISM, and it persists even when pure collisional heating is considered instead of photoionization. The difficulties can be seen most acutely when considering low-ionization line ratios (in particular the N(Si ii)/N(C ii) ratio, which seems to be more shifted toward lower densities than the other line ratios), which suggests that the problem is not simply because we have searched for a single phase to explain all the line ratios simultaneously. For the particular case of N(Si ii)/N(C ii), if carbon is underabundant by a factor of ∼3 compared to the solar value, then the phase-space traced by this ratio is in better agreement with those of other metal-line column density ratios.8

We can further attempt to constrain the physical conditions of the absorber by narrowly focusing on only the silicon lines, for which three ionization levels are observed: the gas density is then restricted by N(Si iv)/N(Si ii) to be 10−4 cm−3n ≲ 10−3 cm−3 over nearly three decades in metallicity. Constraints for all silicon column density ratios yield a single-phase solution at 2 × 10−4 cm−3 < n < 3 × 10−4 cm−3 with 1.4 Z < Z < 3 Z. The total hydrogen column density depends on the uncertain N(H i), but is in the range 17.3 < log N(H) < 19.5, which gives a scale length l ∼ 0.2–50 kpc.

In conclusion, the physical parameters of the A3 absorbing cloud remain far from clear. It is possible, for example, that the gas is not in equilibrium. In that case, a more detailed treatment of this system would require additional data.

C.2. Absorption System toward S393Z082

This sightline shows a single system B1 (Table 4 and Figure 6), and the results of our analysis are shown in Figure 20 (left panel). A solution is obtained within a range of ±0.5 dex around values centered at n ∼ 3 × 10−4 cm−3, Z ∼ 0.1 Z, and log N(H) ≲ 19. This gives l ∼ 10 kpc (l/ρ ∼ 0.1) with deviations of up to ±1 dex of that value. At ρ = 84 kpc, the contribution of any flux from NGC 1097 is smaller than it is for Q0244 (see above), which itself led to only small changes in metallicity and density for plausible changes in the SED. Assuming a Galactic-like ISM gas-to-dust mixture results in a solution with similar densities and total column densities, but with a metallicity increased to ∼0.5 Z.

Figure 20.

Figure 20. Allowed phase space for the absorption system B1 toward S393Z082 (left) and C2 toward Q0246 (right). (See Figure 18 for details.) The hatched surfaces correspond to the allowed phase space when dust is included (see text).

Standard image High-resolution image

C.3. Absorption System toward Q0246

This sightline shows two closely blended components C1 and C2 (Table 5 and Figure 8), and a third more kinematically distinct component C3. Only C2 shows metal-line absorption, so we only discuss our results for that component.

The solutions we obtain for this system depends on whether the detection of the C ii line is real (Section 5.5). Without including this line, we obtain a solution with 5 × 10−5 cm−3 < n < 3 × 10−3 cm−3 and 0.02 Z < Z < 0.2 Z (Figure 20, right panel), and a total column density in the range 3 × 1018 cm−2 < N(H) < 3 × 1019 cm−2, implying 0.3 kpc < l < 300 kpc. When including the ISM-like gas-to-dust patterns used in CLOUDY, the gas is found to have a similar density range, but with 0.2 Z < Z < 0.6 Z and log N(H) ≳ 19.0. As discussed above, there is a wide variation in the depletion patterns that exist in the Galactic ISM: our analysis shows merely that for depletions that are at least similar to the Galaxy, the deduced metallicity still appears to be sub-solar.

If we assume that the C ii line is real, then a single-phase solution is found only when the confidence levels are increased to 3σ for C ii. The allowed region is shown in magenta in Figure 20. The deduced metallicity changes little, but the density is confined to a much smaller range.

Footnotes

  • Based on observations with the NASA/ESA Hubble Space Telescope (HST) obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555.

  • The Magellanic Clouds (MCs) also have very large angular extents, of course; background QSOs have been cataloged (e.g., Cioni et al. 2013, and references therein), but the MCs are at even lower velocities than M31 and M33, and separating absorption from the MW disk and HVCs, and from the MS stream, is complicated. The situation is just as difficult when stars in the MCs themselves are observed, and these stars probe only the very centers of the galaxies. Many HST spectra exist of stars in the MCs, but given that they are not ideal probes of galactic halos, we do not include results from these studies in this paper.

  • Note that our photoionization models are not strictly self-consistent in this case because we assume that the relative abundances of the ions are in solar proportions. Nevertheless, since such models cannot be reliably constrained by the current data we do not consider deviations of relative metal abundances from solar composition in this work.

Please wait… references are loading.
10.3847/0004-637X/826/1/50