A Window on the Earliest Star Formation: Extreme Photoionization Conditions of a High-ionization, Low-metallicity Lensed Galaxy at z ∼ 2*

, , , , and

Published 2018 June 4 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Danielle A. Berg et al 2018 ApJ 859 164 DOI 10.3847/1538-4357/aab7fa

Download Article PDF
DownloadArticle ePub

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

0004-637X/859/2/164

Abstract

We report new observations of SL2S J021737–051329, a lens system consisting of a bright arc at z = 1.84435, magnified ∼17× by a massive galaxy at z = 0.65. SL2S0217 is a low-mass (M < 109 M), low-metallicity (Z ∼ 1/20 Z) galaxy, with extreme star-forming conditions that produce strong nebular UV emission lines in the absence of any apparent outflows. Here we present several notable features from rest-frame UV Keck/LRIS spectroscopy: (1) Very strong narrow emission lines are measured for C iv λλ1548, 1550, He ii λ1640, O iiiλλ1661, 1666, Si iiiλλ1883, 1892, and C iiiλλ1907, 1909. (2) Double-peaked Lyα emission is observed with a dominant blue peak and centered near the systemic velocity. (3) The low- and high-ionization absorption features indicate very little or no outflowing gas along the sight line to the lensed galaxy. The relative emission-line strengths can be reproduced with a very high ionization, low-metallicity starburst with binaries, with the exception of He ii, which indicates that an additional ionization source is needed. We rule out large contributions from active galactic nuclei and shocks to the photoionization budget, suggesting that the emission features requiring the hardest radiation field likely result from extreme stellar populations that are beyond the capabilities of current models. Therefore, SL2S0217 serves as a template for the extreme conditions that are important for reionization and thought to be more common in the early universe.

Export citation and abstract BibTeX RIS

1. Introduction

Determining the precise epoch and source of reionization is currently a major focus of observational cosmology. Yet, the relative contributions to ionizing radiation from stellar and nuclear activity are still uncertain (e.g., Fontanot et al. 2014). However, there is a general consensus that metal-poor, low-mass galaxies host a substantial fraction of the star formation in the high-redshift universe and are likely the key contributors to reionization (e.g., Wise et al. 2014; Madau & Haardt 2015). Depending on the redshift, luminosity function, intergalactic medium (IGM) clumping factor, and ionizing photon production efficiency assumed, the estimated escape fraction of ionizing radiation needed to sustain cosmic reionization ranges from <13% up to 50% (Finkelstein et al. 2012). Therefore, detailed studies of these chemically unevolved galaxies at early epochs are necessary to probe the initial stages of galaxy evolution and assess whether the physical conditions allow sufficient Ly continuum (LyC) radiation to escape.

Surveys of galaxies at 2 ≲ z ≲ 4 reveal a rich diversity of galaxy properties and mark a critical stage of galaxy evolution at which the peak of both star formation and black hole accretion activity occur (see Shapley 2011, for a review). These surveys have provided us with a broad understanding of how galaxies assemble and evolve, but the spatial and spectral limitations inherent in observing faint, distant objects mean that many of the physical processes regulating this dynamic evolution are poorly constrained. Clearly, a more detailed understanding of this key epoch is needed.

Until recently, very few high-redshift galaxies had been observed spectroscopically in detail, as they are generally too faint to obtain high signal-to-noise ratio (S/N) observations with even 8−10 m class telescopes. Previously, this challenge has been overcome by building composite spectra of high-redshift galaxies (e.g., Steidel et al. 2001; Shapley et al. 2003; Rigby et al. 2018), the properties of which generally indicate that the sources are young (∼108 yr), metal-poor (Zneb < 0.5 Z) galaxies with high specific star formation rates and strong outflows of ionized gas. Currently, our most detailed information on the physical conditions in galaxies at moderate redshift comes from careful, high-S/N spectroscopic studies of gravitationally lensed objects (e.g., Pettini et al. 2000, 2002b; Fosbury et al. 2003; Hainline et al. 2009; Quider et al. 2009; Yuan & Kewley 2009; Dessauges-Zavadsky et al. 2010; Quider et al. 2010; Rigby et al. 2011; Jones et al. 2013; Amorín et al. 2014; Vanzella et al. 2016). However, observations of such galaxies are necessarily obtained one at a time, and so the number remains small and does not yet adequately sample the diversity among high-redshift galaxy populations. In particular, there are few detailed studies of low-mass, low-metallicity (Zneb < 0.1 Z) sources and only a handful of targets with direct determinations of their physical conditions such as electron temperature, density, and gas-phase oxygen abundance (e.g., Hainline et al. 2009; Bayliss et al. 2014; Christensen et al. 2014).

In order to understand how the physical conditions in galaxies have evolved over cosmic time, we need detailed studies of galaxies at early epochs that have undergone little chemical evolution in comparison with nearby counterparts. The Strong Lensing Legacy Survey (Tu et al. 2009) has discovered several such targets, including the z ∼ 1.85 gravitationally lensed arc SL2S J0217–0513 (hereafter SL2S0217). Low-resolution rest-frame UV and optical spectra of this arc display very strong, high-ionization, nebular emission lines, suggesting that SL2S0217 is both very metal poor and highly ionized (Tu et al. 2009; Brammer et al. 2012).

In preparation for the next frontier of high-redshift studies made possible by the forthcoming generation of space-based (i.e., JWST) and 30 m ground-based observatories (i.e., GMT, ELT, and TMT), we seek to understand the conditions producing the extreme UV emission lines seen in SL2S0217 and even higher-redshift sources. Because the intergalactic medium becomes increasingly neutral at z > 6, Lyα emission, which is typically the strongest emission feature in UV spectra of high-redshift Lyα-emitting galaxies,5 is increasingly scattered and therefore suppressed (Stark 2016). Instead, the next strongest UV emission lines, such as C iii, must be used to spectroscopically confirm redshifts. Additionally, these lines will provide useful constraints on the physical properties of the emitting source and may be used as important diagnostics for characterizing high-redshift galaxies. Recently, these strong UV nebular emission lines have been observed in high-redshift (z > 6) galaxies (e.g., Stark et al. 2015; Stark 2016), indicative of low metallicity and high ionization (e.g., Erb et al. 2010; Stark et al. 2014). Because high-ionization targets may also have large escape fractions of H-ionizing photons (e.g., Brinchmann et al. 2008; Jaskot & Oey 2013; Stark et al. 2014), SL2S0217 is a good candidate for studying the chemically unevolved, extreme physical conditions that are expected in the early galaxies responsible for reionizing the universe.

Here we present the analysis of new Keck/Low Resolution Imaging Spectrograph (LRIS) spectroscopy of SL2S0217. Section 2 describes the previous observations and derived properties of SL2S0217 presented in Brammer et al. (2012). In Section 3, we present our analysis of the existing photometry and an improved lensing model for SL2S0217. The spectral observing plan and data reduction for the new LRIS spectrum are laid out in Section 4. We examine the remarkable features of the resulting rest-frame UV spectrum in Section 5, paying particular attention to the nebular emission lines (Section 5.2), the double-peaked Lyα profile (Section 5.3), and the interstellar absorption line profiles (Section 5.4). The physical conditions of the gas and chemical abundances of oxygen, carbon, nitrogen, and silicon are estimated from the nebular emission lines in Section 7. Finally, in Section 8, we attempt to reproduce our spectra and constrain the ionization source by inspecting a large grid of photoionization models, considering contributions from stars, shocks (Section 8.3.1), and active galactic nuclei (AGNs; Section 8.3.2). The photoionization budget is discussed further in the context of the strong C iv (Section 8.4.1) and He ii (Section 8.4.2) emission. Throughout this paper, we adopt a flat FRW metric with Ωm = 0.3, ΩΛ = 0.7, and H0 = 70 km s−1 Mpc−1 and the solar metallicity scale of Asplund et al. (2009), where 12 + log(O/H) = 8.69.

2. Galaxy Properties and Past Observations

SL2S0217 is a gravitationally lensed galaxy at redshift z ∼ 1.85 that is lensed by a massive galaxy at z = 0.6459 as discovered by the Strong Lensing Legacy Survey (Tu et al. 2009). The general properties of SL2S0217 are well characterized owing to its location within both the Hubble Space Telescope (HST) CANDELS imaging survey (ACS F606W and F814W and WFC3 F125W and F160W; Grogin et al. 2011; Koekemoer et al. 2011) and the 3D-HST spectroscopic survey (WFC3 G141 grism spectroscopy; Brammer et al. 2012, hereafter B12). Additionally, a low-resolution Keck/LRIS spectrum (Tu et al. 2009) exists for the arc. All existing imaging and spectral observations for SL2S0217 are summarized in Table 1.

Table 1.  Observations of SL2S J0217–0513

Observing Program Band/Wavelength
Photometry:  
HST CANDELS ACS F606W (V), F814W (i)
  WFC3 F125W (J), F160W (H)
3D-HST F140W (Hwide)
Spitzer MIPS 24 μm (mJy)
HST Lyα WFC3 F343N, F390M
Spectroscopy:  
3D-HST Optical WFC3 G141; 1.10–1.65 μm
Keck/LRIS UV 600/4000; ∼3100–9000 Å

Note. Existing imaging and spectral observations of SL2S0217, including the new Keck/LRIS spectra presented here.

Download table as:  ASCIITypeset image

B12 report that the arc is very blue, with a UV slope of β = −1.7 ± 0.2 (fλ ∝ λβ) for λrest = 2100–2800 Å, but is rather red in the F160W band owing to strong [O iii] λλ4959, 5007 emission. The low-resolution WFC3 G141 grism spectrum of the lens system presented in B12 shows extremely high equivalent width rest-frame optical emission lines that are atypical of z ∼ 0 star-forming galaxies (namely, rest-frame EW(Hβ) = 1470 Å and EW([O iii] λλ4959, 5007) = 5690 Å). The emission lines are all spatially extended, indicating that strong nebular emission is coming from multiple clumps along the arc. From their analysis of this grism spectrum and modeling of the broadband spectral energy distribution, B12 concluded that the arc is a young, low-mass, low-metallicity (12+log(O/H) ∼ 7.5), starbursting (specific star formation rate [sSFR] ∼ 100 Gyr−1) galaxy, with similar characteristics to local blue compact dwarf galaxies. SL2S0217 is thus one of the lowest-metallicity star-forming galaxies yet identified at z > 1. Derived quantities from B12 are given in the top of Table 2.

Table 2.  Properties of SL2S J0217–0513

Parameter Value
R.A. 02:17:37.237
Decl. −05:13:29.78
z 1.84435 ± 0.00066
From Brammer et al. (2012)
log (age yr−1) 7.2 ± 0.2
μ 25 ± 1a
μ' ∼1.4a
AV (continuum) 0.09 ± 0.15
log (μ · M/M) 9.5 ± 0.1
β (2000–2800 Å) −1.7 ± 0.2
μ · SFRHβ (M yr−1) 390 ± 9
$\sqrt{\mu ^{\prime} }\cdot {r}_{e}$ 350 pc
Using Updated Lens Model
μtot 17.3 ± 1.2
μeff 19 ± 1.5
log(M/M) 8.26 ± 0.10
SFR (M yr−1)b 22.5 ± 1.6

Notes. Observed and derived quantities for SL2S0217. The HST CANDELS (Koekemoer et al. 2011) and MIPS (Labbé et al. 2006) photometry, as well as properties derived from the 3D-HST grism spectrum and SED fit, are taken from Brammer et al. (2012). Updated parameters incorporating the lensing model from this work are also listed. R.A. and decl. specify the optical center in units of hours, minutes, and seconds and degrees, arcminutes, and arcseconds, respectively.

aμ is the relative magnification between the integrated arc and the counterimage. μ' is the brightness magnification of the counterimage. The total lens magnification is μ = μ · μ'. bDetermined from the Hβ line luminosity.

Download table as:  ASCIITypeset image

3. Improved Photometry and Lensing Model

Using the available five-band HST imaging, we quantify the resolved photometric properties of SL2S0217. These data include optical F606W and F814W images observed with ACS/WFC (4856 and 11,300 s, respectively) and near-infrared F125W, F140W, and F160W images observed with WFC3/IR (1912, 812, and 3512 s, respectively). All of the data were reduced with the AstroDrizzle package (Fruchter et al. 2010) and were resampled to a pixel scale of 0farcs05. The absolute astrometry from the telescope differs between the different bands, but we ensured that each band was well registered by matching the apparent coordinates of the central lensing galaxy in the output resampled images. We used an unsaturated bright star 21'' from the lens as a model for the point-spread function (PSF) in our subsequent surface brightness and lens modeling.

The surface brightness distribution of the foreground lensing galaxy overlaps with the lensed arc features, and we therefore modeled and removed this light from each image using the following process. First, we manually masked several interloping objects that lie both within and around the lens system. Next, we constructed a mask for the arc and counterimage and simultaneously fit the foreground lensing galaxy light distribution in each band with a model that allowed three concentric Sérsic profiles. As a test of this procedure, we have also fit the foreground light in the F606W band while simultaneously modeling the light of the background source and found consistent results. Finally, the foreground surface brightness model was subtracted from the data to give an uncontaminated view of the background source structure.

The structure of the bright, lensed arc indicates that the background galaxy has a complex morphology. We therefore used the adaptive pixelated source modeling technique described in Vegetti & Koopmans (2009) to delens and "reconstruct" the lensed source. The intrinsic source surface brightness distribution was described on an irregular grid of pixels that approximately follows the magnification of the lens, with a PSF-deconvolved intensity at each pixel determined from the lens mass model and observed data. We modeled the lensing potential as an elliptical power-law mass distribution with an empirical external shear and constrained it with the F606W data, which have excellent spatial resolution and S/N. The reconstructed source surface brightness distribution is shown in the right panel of Figure 1.

Figure 1.

Figure 1. Left: HST F606W, F814W, and F125W image of the lens. Middle: same as the left panel, but with the resolution degraded to the resolution of the F160W image to perform aperture photometry within the shown apertures. Right: reconstructed source in the F606W filter, showing the intrinsic three regions corresponding to the apertures in the middle panel.

Standard image High-resolution image

The total observed (magnified) magnitudes are given in Table 3, with systematics-dominated uncertainties of ∼0.03 mag. While the increased freedom in our lens model generally increases the uncertainties on the magnification, our flexible source model allows a better fit to the data, decreasing the systematic uncertainty. Therefore, the lens modeling yields a total magnification of μ = 17.3 ± 1.2, significantly lower than the value of ∼35 determined by Cooray et al. (2011) and reported in B12. This discrepancy in magnification is largely driven by the choice of source model. In our analysis, we find that there is a significant amount of low-surface-brightness (and low-magnification) flux that is not encapsulated by the Cooray et al. (2011) sources, and for that reason the magnification (observed light)/(modeled source light) is larger for their model.6

Table 3.  Lens Model Photometry

Parameter Region
  Total
F606W 22.15
F814W 22.03
F125W 21.80
F140W 21.14
F160W 20.98
  Magenta Green White
F606W−F160W 1.03 ± 0.10 1.35 ± 0.19 1.79 ± 0.27
F814W−F160W 1.01 ± 0.10 1.24 ± 0.18 1.53 ± 0.25
F125W−F160W 0.79 ± 0.08 0.93 ± 0.14 1.01 ± 0.21
F140W−F160W 0.25 ± 0.08 0.28 ± 0.13 0.26 ± 0.20

Note. Top: the total observed (magnified) magnitudes, with systematics-dominated uncertainties of ∼0.03 mag. Bottom: the colors measured for the magenta, green, and white apertures show significant broadband color differences across the galaxy.

Download table as:  ASCIITypeset image

This low-surface-brightness flux is readily seen in the bluer filters, where the light from the lensing galaxy is less prominent, as extensions of the bright arc immediately north and south of the lensing galaxy. Our flexible source model allows magnification to vary across the source and results in the intrinsically higher surface brightness regions (highlighted in the right panel of Figure 1) having somewhat higher magnifications of ∼25. In other words, there is clearly differential magnification across the broadband image of the source. We simulated slit losses of ∼10% by convolving the light profile from the HST F606W image with the seeing disk and integrating through the 1'' slit of the LRIS spectrum. Accounting for the lensing magnification and applying the slit loss correction, we find that the flux within the spectroscopic aperture has an effective mean amplification of 19 ± 1.5.

As in B12, we find significant broadband color differences across the galaxy. To quantify this, we convolved each of the four bluer images with a Gaussian filter to match the resolution of the F160W image and determined aperture colors within the three regions shown in Figure 1. The colors measured for the magenta, green, and white apertures are listed in Table 3. The white region, which is clearly redder, is also much lower surface brightness and lower magnification than the other regions, and so it contributes less to the observed LRIS spectrum. Thus, the UV spectrum of SL2S0217 is largely produced by the neighboring magenta and green star-forming regions.

4. New Keck/LRIS Spectroscopic Observations and Data Reduction

4.1. Observations

We obtained spectra of SL2S0217 using LRIS (Oke et al. 1995) on the Keck I telescope on the UT date of 2015 October 12. The 600/4000 grism and 600/7500 grating were used with the blue and red detectors, respectively, with the dichroic at 560 nm, resulting in an observed-wavelength coverage of approximately 3100–9000 Å and an average resolution at FWHM of roughly 4.0 Å (275 km s−1) in the blue and 4.7 Å (190 km s−1) in the red. This corresponds to a rest-wavelength coverage of approximately 1100–3000 Å and resolutions of 1.4 (1.7) Å in the blue (red).

Internal and twilight calibration flats were obtained at the beginning of the night to account for differences between the chip illumination patterns on the sky and the internal lamps. Two standard stars, Feige 24 and Feige 110, with spectral energy distributions peaking in the blue, were observed over a range of hour angles throughout the night, allowing the flux calibration to be determined as a function of airmass. SL2S0217 was observed for 13 × 1800 s, or 6.5 hr, in clear, dark conditions with 0.7–0farcs8 seeing, over an airmass range of 1.0–1.7. A slit of 1farcs× 168'' at an angle of 3° was chosen in order to encompass the maximum flux from the arc, as shown in Figure 2. HgNeArCdZn arc lamps were observed after every other exposure to mitigate the effects of instrument flexure in the wavelength calibration.

Figure 2.

Figure 2. CANDELS HST ACS F606W image of SL2S0217. The 1farcs0 LRIS slit is overlaid in red, demonstrating that the majority of the arc light was captured in the slit. A second lensed source is visible in the top part of the slit, but it is spatially distinct (both along the slit and in redshift) and so does not affect the spectrum of SL2S0217.

Standard image High-resolution image

4.2. Spectra Reduction

The LRIS spectra were processed using ISPEC2D (Moustakas & Kennicutt 2006), a long-slit spectroscopy data reduction package written in IDL. Master sky and internal flats were constructed by taking the median at each pixel after normalizing the counts in the individual images. These calibration files were then used to bias-subtract, flat-field, and illumination-correct the raw science data frames. Misalignment between the trace of the light in the dispersion direction and the orientation of the CCD detector was rectified via the mean trace of the standard stars, providing alignment to within a pixel across the detector. A median sky subtraction was performed at each column along the dispersion, followed by a wavelength calibration applied from the HgNeArCdZn comparison lamps taken at the nearest airmass. The one-dimensional spectra were produced using a boxcar aperture that encompassed roughly 99% of the light in the Lyα feature. We note that while a second lensed source is within the top part of the slit (see Figure 2), it is distinct spatially, as well as in redshift (z ∼ 2.32; Cooray et al. 2011), and so does not affect the extracted spectrum of SL2S0217. Individual exposures were then flux calibrated using the sensitivity curve derived from the standard star observations taken throughout the night. Finally, the 13 subexposures were median combined, eliminating cosmic rays in the process.

The resulting 1D spectra were then compared as a check on the flux calibration: the overlapping regions of flux-calibrated blue and red spectra were found to be in good agreement within the dispersion of their continua. We then checked the absolute flux calibration of the combined blue+red LRIS spectrum using newly obtained HST images in the WFC3 F390M (covering rest frame ∼1300–1400 Å) and F343N (covering the Lyα emission line) filters; these images will be presented in a future paper. We used the PYSYNPHOT package in Python to predict the F390M AB magnitude from the spectrum and then scaled the spectrum to match the observed F390M AB magnitude (22.68 ± 0.07; D. K. Erb et al. 2018, in preparation). Note that we used the bluest possible continuum band, so that contamination of the lens is negligible. The resulting 1D spectrum and error spectrum are shown in Figure 3. Regions of strong sky line emission are shaded gray. Potential emission and absorption features are labeled accordingly.

Figure 3.

Figure 3. Absorption and emission features observed in the LRIS spectra (flux vs. wavelength) of SL2S0217, with the error spectrum shown in red. Regions of strong sky line emission are shaded gray. Note that the sky features in the red spectrum are stronger than in the blue and so are shaded darker gray accordingly. The locations of nebular emission lines are identified with dashed blue lines. Absorption features due to the interstellar medium (ISM) are identified by dashed purple lines, and corresponding fine-structure emission lines are labeled in black. Not all labels correspond to detections, but they identify the location of features discussed in the text.

Standard image High-resolution image

5. The Rest-frame UV Spectrum

The blue portion of the rest-frame UV LRIS spectrum of SL2S0217 is shown in Figure 4, where the middle and bottom panels are zoomed in on the vertical scale to highlight the numerous absorption and emission features observed. In comparison, we plot the composite spectrum of ∼1000 z ∼ 2 galaxies from Erb et al. (2010) and note two defining properties of the SL2S0217 spectrum: (1) the interstellar absorption features typical of z ∼ 2 galaxies appear to have velocity profiles that are roughly three times narrower in SL2S0217 and (2) significant, high-ionization, nebular emission is present in SL2S0217 that is atypical of z ∼ 2 galaxies. We further discuss some of the most striking features of the LRIS spectrum below.

Figure 4.

Figure 4. Blue arm of the Keck/LRIS spectrum of SL2S0217, where regions of significant sky contamination are designated by the gray shading. The top panel shows the full extent of the spectrum, whereas the bottom two panels are zoomed in to depict the numerous ISM absorption and nebular emission features. The z ∼ 2 composite spectrum from Erb et al. (2010) is overplotted in orange, demonstrating the staggering difference in the strength of the emission features and widths of the absorption features between the two spectra.

Standard image High-resolution image

5.1. Source Redshift

In order to analyze the velocity structure of the absorption and emission features seen in the SL2S0217 UV spectrum, it is necessary to first establish a reliable determination of the systemic redshift as the reference velocity zero point. Tu et al. (2009) used all of the strong emission lines present in their low-resolution Keck/LRIS spectrum (Lyα, C iv, He ii, O iii], and C iii]) to estimate zarc = 1.84691 ± 0.0024. However, our LRIS spectrum shows that the Lyα emission for SL2S0217 is double peaked (see Figure 6 below). Additionally, both the He ii and C iv emission can have complex profiles complicated by combinations of stellar and nebular contributions. Therefore, we choose not to use these lines in our redshift measurement. With the higher resolution of our LRIS spectrum, we are able to use the O iii] λλ1660, 1666, Si iii] λλ1883, 1892, and C iii] λλ1907, 1909 emission lines for our systemic redshift determination.

The line centers used to calculate the redshift are given in Table 4, resulting in a systemic redshift of zsys = 1.84435 ± 0.00066 for SL2S0217. Since this value is within the uncertainty of the redshift measured by Tu et al. (2009), the line centers of the observed Lyα, He ii, and C iv profiles must lie close to the systemic velocity.

Table 4.  Emission-line Redshift Determinations

Ion λlab (Å)a λobs (Å) z
iii] 1660.81 4723.97 1.84438
iii] 1666.15 4739.03 1.84430
Si iii] 1882.71 5355.23 1.84443
Si iii] 1892.03 5381.48 1.84429
[C iii] 1906.68 5423.25 1.84434
iii] 1908.73 5429.13 1.84437
Average: 1.84435 ± 0.00066

Note. Systemic redshift of SL2S0217 calculated from the strong nebular emission lines in the LRIS spectrum.

aVacuum wavelengths.

Download table as:  ASCIITypeset image

5.2. Nebular Emission Lines

The rest-frame UV spectrum of SL2S0217 contains numerous emission-line features that are sensitive to the ionizing stellar population, physical properties of the emitting gas, and kinematics of outflows. In particular, SL2S0217 shows double-peaked Lyα emission and exceptionally strong, nebular-like emission from high-ionization species such as C iv λλ1548, 1550, He ii λ1640, O iii] λλ1661, 1666, Si iii] λλ1883, 1892, and C iii] λλ1907, 1909, with additional weak detections of the low-ionization [O iii] λ2322 and C ii] λλλ2325, 2327, 2328 lines. While UV emission lines have rarely been observed at strengths comparable to SL2S0217 in star-forming galaxies at any epoch, these features are especially uncommon for the typically older, more massive star-forming galaxies studied at z ∼ 2 (see, e.g., Jones et al. 2012; Reddy et al. 2012; Whitaker et al. 2012). Figures 3 and 4 highlight these atypical emission features of SL2S0217. However, extreme emission-line features may be more common at higher redshifts, where we expect to find harder radiation fields associated with less evolved, more metal-poor galaxies. Therefore, the magnified emission-line spectrum of SL2S0217 provides a unique window to examine the conditions driving powerful photoionization in distant galaxies.

To characterize the emission-line spectrum, all emission-line strengths for our LRIS spectrum were measured using the SPLOT routine within IRAF.7 With the exceptions of Lyα, C iv, and He ii, which have multiple components, groups of nearby lines were constrained to a single Gaussian FWHM and a single shift in wavelength from vacuum when possible.8 Note that the nebular emission lines are unresolved, and so subsequent discussions of line widths do not account for the instrumental resolution. Following B12, who found that the Hγ/Hβ ratio from the grism spectrum was consistent with no reddening, we ignored the reddening from dust and simply corrected the line fluxes for the galactic extinction along the line of sight to SL2S0217 (E(BV) = 0.0194; Schlafly & Finkbeiner 2011)9 using the Cardelli et al. (1989) reddening law. The adopted emission-line fits are depicted in Figure 5, with the resulting line strengths and intensities given in Table 5. We examine a few of the interesting spectral features of SL2S0217 individually in the next sections and in the discussion (Section 8).

Figure 5.

Figure 5. Unusually strong nebular emission lines of SL2S0217. Lines are fit by Gaussian profiles (blue), with FWHMs and wavelength shifts tied together for nearby lines. When multiple components are fit to a single blended structure, the components are shown by dashed light-blue lines. In the case of He ii, we show both a single Gaussian fit and a two-component fit (inset window) allowing for narrow (∼280 km s−1) nebular and broad (∼710 km s−1) stellar contributions. Two features are particularly interestingly: (1) the C iv emission is surprisingly strong, and (2) the He ii emission is both strong and narrow, appearing mostly nebular in origin. The ±1σ error spectrum (shaded gray) is shown for comparison. Residuals to the fit are plotted below the corresponding spectral windows on a scale of −1 to 1.

Standard image High-resolution image

Table 5.  Nebular Emission Lines

Ion λlab (Å)a Fλb W (Å)c Iλa Iλ/Iλ1909
Lyαb 1215.67 58.7 ± 0.6 66.0 71.3 ± 1.2 8.94
Lyαr 1215.67 39.0 ± 0.4 43.2 47.3 ± 0.8 5.94
Lyαtot 1215.67 100.8 ± 1.0 113.0 122.4 ± 2.1 15.3
iv 1548.19 5.54 ± 0.07 3.1 6.40 ± 0.12 0.80
  1550.77 4.47 ± 0.06 2.5 5.17 ± 0.10 0.65
He ii 1640.42 4.62 ± 0.07 2.8 5.32 ± 0.10 0.67
He iin 1640.42 3.98 ± 0.06 2.4 4.58 ± 0.09 0.57
He iiw 1640.42 1.33 ± 0.05 0.8 1.53 ± 0.07 0.19
iii] 1660.81 3.50 ± 0.06 2.1 4.03 ± 0.09 0.51
  1666.15 7.60 ± 0.09 4.5 8.74 ± 0.15 1.10
iii]d 1749 0.52 ± 0.10 0.4 0.60 ± 0.12 0.08
Si iii] 1882.71 3.17 ± 0.06 2.1 3.66 ± 0.08 0.46
  1892.03 2.35 ± 0.05 1.6 2.72 ± 0.07 0.34
[C iii] 1906.68 10.4 ± 0.12 7.0 12.1 ± 0.19 1.51
iii] 1908.73 6.90 ± 0.08 4.7 7.98 ± 0.13 1.00
[O iii] 2321.66 0.64 ± 0.03 0.6 0.74 ± 0.08 0.09
ii] 2325.40 0.36 ± 0.03 0.4 0.42 ± 0.08 0.05
  2326.93 1.11 ± 0.03 0.9 1.29 ± 0.09 0.16
  2328.12 0.27 ± 0.03 0.4 0.32 ± 0.08 0.04
[Ne iii] 3870.16 30 ± 4 32.6 ± 4.4
Hδ 4101.00 19 ± 2 21.1 ± 2.2
Hγe 4341.69 35 ± 2 37.8 ± 2.2
[O iii]e 4364.44 5 ± 3 5.3 ± 3.2
Hβ 4862.69 74 ± 2 517 78.9 ± 2.3
[O iii] 4960.30 79 ± 2 84.6 ± 3.3
  5008.24 230 ± 2 2095f 244.2 ± 4.1

Notes. Rest-frame UV emission lines from this work and optical emission lines from the 3D-HST spectrum of B12. Emission features were fit with Gaussian profiles, where nearby lines were constrained to the same FWHMs and a single wavelength offset. The exceptions to this rule are the Lyα, C iv, and He ii profile fits, which have additional contributions from resonant scattering (Lyα and C iv) and stars (C iv and He ii). The two fits listed for Lyαb and Lyαr correspond to blue and red Gaussian components, whereas Lyαtot was determined by integrating the profile as a whole. For He ii, a single Gaussian fit is reported, as well as a two-component fit with narrow and (potential) wide parts. Following B12, who found negligible reddening in SL2S0217, we simply corrected for the galactic extinction (E(BV) = 0.0194; Schlafly & Finkbeiner 2011). Line fluxes (observed frame) and equivalent widths (rest frame) are listed in Columns (3) and (4), followed by the extinction-corrected intensities in Column (5).

aVacuum wavelengths. bUnits are 10−17 erg s−1 cm−2, uncorrected for lens magnification. cEquivalent widths are measured in the rest frame. diii] is the blend of N iii] λλ1748, 1749. eLines were completely blended in the 3D-HST grism spectrum. fTotal EW of the blended [O iii] λλ4959, 5007 doublet.

Download table as:  ASCIITypeset image

5.2.1. Blended Emission Features

Neither the components of the double Lyα emission feature nor the individual lines of the C iii] λλ1907, 1909 or C iv λλ1548, 1550 doublets are resolved, but they are still well fit by two blended Gaussians. C iv and He ii are further complicated by the potential for their profiles to be modified by multiple components. In the case of He ii, which can have both nebular and stellar emission, we used a two-component fit with narrow and wide profiles, respectively. Both single- and multicomponent fits to the He ii profile are demonstrated in Figure 5. While the two-component model may offer a more visually appeasing fit to the He ii profile, it does not significantly reduce the fit residuals, and so the significance of the wide (∼700 km s−1) stellar component is difficult to assess with the current S/N and resolution. We therefore suggest that the reported fits are upper/lower limits for the stellar/nebular components. Regardless of the correct fit, the He ii emission in SL2S0217 is clearly dominated by a narrow component whose width is consistent with nebular emission (∼280 km s−1).

In the case of C iv, which is typically an ISM absorption feature or a P Cygni profile from the stellar winds of massive stars, it is clearly seen in SL2S0217 as an emission doublet. Because the C iv doublet, like Lyα, can also be resonantly scattered, its profile widths were not constrained to the pure nebular line widths. The resulting best fit had an observed-frame Gaussian FWHM 6.0 Å or about 410 km s−1. Compared to the nearby profiles of the O iii] nebular lines (285 km s−1), this 45% increase in profile width indicates that C iv is likely dominated by nebular emission but is broadened by resonant scattering. Additionally, strong emission may be masking an absorption component, and therefore we report the C iv emission-line strengths as lower limits.

5.3. Lyα

Lyα emission is predominantly produced by recombination in the ionized gas surrounding star formation, but also by collisionally excited neutral H gas (cooling radiation; Dijkstra 2014). However, the emission feature is increasingly obscured and scattered with increasing H i column density and dust attenuation. Therefore, the shape of the Lyα profile depends on the kinematics and distribution of the outflowing gas and dust in star-forming galaxies (e.g., Verhamme et al. 2006; Kornei et al. 2010; Steidel et al. 2010).

The Lyα velocity profile from the LRIS spectrum of SL2S0217 is shown in Figure 6, where we observe strong, double-peaked emission with a dominant blue peak and nearly systemic central velocity. The total, integrated Lyα emission has a large equivalent width of WLyα = 113 Å, consistent with values for other young, metal-poor galaxies (e.g., Cowie et al. 2011; Trainor et al. 2016). Fitting the emission with two independent Gaussian profiles, we measure the relative blue to red flux to be ∼1.5. The peak separation of this profile is ${{\rm{\Delta }}}_{{\rm{peaks}}}=371$ km s−1. From Figure 6 we can see that this model fits the profile fairly well; however, some of the flux at the peaks is missed, and a residual blue tail extends to ∼−1300 km s−1.

Figure 6.

Figure 6. Velocity profile of the double-peaked Lyα emission in SL2S0217. The overall profile is reasonably well fit using two Gaussians (blue lines), but a residual blue tail remains (shown in the bottom panel), indicating that a more complex fit is needed to fully describe the kinematic properties of the Lyα-emitting gas. Interestingly, the blue component is stronger than the red one and has a slightly larger velocity width. The Lyα peak separation is relatively small, shown here to be narrower than the velocity range of the average absorption feature (red dashed lines), with a nearly systemic central velocity.

Standard image High-resolution image

Multipeaked Lyα emission (multiplicity) is commonly associated with star-forming galaxies. In fact, recent studies have found multiplicity rates of 30%–50% among large samples of z ∼ 2−3 star-forming galaxies (e.g., Kulas et al. 2012) and Lyα emitters (LAEs; e.g., Yamada et al. 2012; Trainor et al. 2015). Targets with Lyα multiplicity are typically grouped according to whether they are blue- or red-peak dominant. While targets with red-dominant peaks are much more common than those with blue-dominant peaks, Trainor et al. (2015) find that 26% of their 129 multipeak LAEs are blue dominant, with an average peak separation of $\langle {\rm{\Delta }}{v}_{{\rm{peaks}}}\rangle \,=660\pm 300$ km s−1. Additionally, for blue-dominant profiles in star-forming galaxies, Kulas et al. (2012) found an average peak separation of $\langle {\rm{\Delta }}{v}_{{\rm{peaks}}}\rangle \sim 800$ km s−1, where the blueshifted peak is located at ${v}_{{\rm{peak}}}\sim -200$ km s−1. Note, however, that these results are based on low-resolution spectra (Δv ∼ 200–600 km s−1), which are more sensitive to larger peak separations, and so may be biased to a higher average value. Still, the peak separation we observe for SL2S0217 (${\rm{\Delta }}{v}_{{\rm{peaks}}}\sim 371$ km s−1) is small relative to these samples of star-forming galaxies at similar redshifts.

The simplest explanation for a double-peaked Lyα emission profile is scattering by a static, homogeneous, spherical or shell gas cloud. Given such a medium, the Lyα peak separation is set by the total H i column density. For the peak separation and Doppler parameter (thermal velocity dispersion) of $b\,=\sqrt{2{k}_{B}T/{m}_{{\rm{H}}}}=16.3$ km s−1 (for T = 16,100 K) measured in SL2S0217, the radiative transfer models of Verhamme et al. (2015) predict a column density of neutral gas of ${N}_{{\rm{H}}{\rm{I}}}\,\lesssim {10}^{20}$ cm−2. Observationally, the trend in Lyα peak separation with H i column density found for 12 Lyα emitters at z ∼ 2 by Hashimoto et al. (2015) also suggests ${N}_{{\rm{H}}{\rm{I}}}\approx {10}^{20}$ cm−2. However, static models produce red and blue emission components of equal strengths and widths, and so a more complex picture is needed to explain the larger intensity and wing of the blue Lyα peak in SL2S0217.

Additional explanations of the Lyα emission profile exist, including fluorescence, where ionizing photons that escape from galaxies can produce double-peaked fluorescent Lyα emission in cold clouds (e.g., Mas-Ribas & Dijkstra 2016). Like the Lyα profile of SL2S0217, fluorescence tends to produce a small peak separation (e.g., Gould & Weinberg 1996). Alternatively, Vanzella et al. (2017) used MUSE integral field spectroscopy of two extended Lyα systems to measure spatial-dependent emission and varying substructures, suggesting radiative transfer through clumpy/multiphase media as the source of blue-dominant Lyα emission (e.g., Gronke et al. 2016). Unfortunately, our observations of SL2S0217 lack sufficient resolution to assess the spatial extent of or variations in emission and are further complicated by lensing.

Even in a simple static model, viewing Lyα emission from different inclinations/orientations affects the relative blue and red Lyα emission strengths (Behrens & Braun 2014), especially in a differentially lensed system such as SL2S0217. Alternatively, in radiation transfer models, blue- and red-peak-dominant Lyα morphologies are associated with inflowing and outflowing gas, respectively (e.g., Dijkstra et al. 2006; Verhamme et al. 2006). In the case of infalling gas toward a central source, the red side of a double-peaked profile is depressed as the redshifted Lyα photons within the galaxy see higher optical depth owing to the line-of-sight infalling gas, resulting in dominant blue-peak Lyα emission (see, e.g., Yang et al. 2012). The Lyα profile of SL2S0217 is consistent with this infalling halo model, which predicts a peak separation of $\langle {\rm{\Delta }}{v}_{{\rm{peaks}}}\rangle $ ∼ 400 km s−1 (Verhamme et al. 2006, see their Figure 5). However, such a model does not eliminate the possibility of outflows, as we could have a restricted sight line along which the gas is static or cold pristine gas is flowing into the galaxy, and such that outflows are not visible to our viewing angle.

If we are in fact observing the effects of gas inflowing toward the star-forming regions in SL2S0217, then we may be witnessing an early episode of star formation, during which we might have expected relatively low ${N}_{{\rm{H}}{\rm{I}}}$ and negligible dust attenuation (inferred from the Balmer decrement of the grism spectrum) to favor a leakage of ionizing radiation. The fraction of escaping Lyα can be estimated by comparing the intrinsic and observed Lyα luminosities. We determine the intrinsic Lyα luminosity by multiplying the theoretical Lyα/Hβ ratio of 23.3 (assuming Te = 1.5 × 104 K and ne = 102 cm−3; Osterbrock 1989) by the Hβ intensity (only corrected for galactic extinction; see Section 5.2). Note, however, that the theoretical Lyα/Hβ ratio is density sensitive, and so we also consider the ne = 103 cm−3 case, in which Lyα/Hβ = 25.7. Finally, we use the total magnification (μtot = 17.3) to correct the observed Hβ intensity and the effective mean magnification (μeff = 19) to correct the observed Lyα intensity. Interestingly, the Lyα escape fraction, ${f}_{\mathrm{esc}}^{\mathrm{Ly}\alpha }$, is ${L}_{\mathrm{Ly}\alpha }^{\mathrm{obs}}$/${L}_{\mathrm{Ly}\alpha }^{\mathrm{int}}=0.061(0.055)$ for ne = 102 (103) cm−3, or only 6% of Lyα emission escapes along the line of sight.

For a sample of local LyC emitters, Verhamme et al. (2017) found that both Lyα and LyC escape fractions increased with increasing Lyα EWs, while Lyα peak separations decreased. Along these lines, the small Lyα peak separation (371 km s−1) and large EW in SL2S0217 (113 Å) argue for significant Lyα leakage (∼40%–60%; Verhamme et al. 2017), discordant with the observed estimated escape fraction. Note, however, that the Lyα escape fraction may be underestimated owing to slit losses. A preliminary analysis of the continuum-subtracted HST WFC3 F343N image suggests significant Lyα slit losses of roughly 30%–40%. Correcting the Lyα flux for this factor increases the Lyα escape fraction to 10%, still surprisingly small given the large observed Lyα EW. A complete analysis will be discussed in a future paper (D. K. Erb et al. 2018, in preparation).

While SL2S0217 is a clear outlier to the trends found for the sample of LyC emitters in Verhamme et al. (2017), Jaskot et al. (2017) recently reported an extreme Green Pea galaxy, J1608, with properties similar to SL2S0217. In particular, these authors found J1608 to be a very high ionization (from [O iii]/[O ii]) galaxy, with no evidence of outflows, and measured a Lyα escape fraction of 0.16, despite its strong Lyα EW of 163 Å and narrow peak separation of ${{\rm{\Delta }}}_{{\rm{peaks}}}=214$ km s−1. They suggest that multiple mechanisms for LyC escape exist, where such extreme targets may have suppressed superwinds, with radiation-dominated feedback driving Lyα escape, and likely escaping LyC emission. Similarly, despite the low Lyα escape fraction measured for SL2S0217, it has several characteristics, namely, strong [O iii]/[O ii], large Lyα EW, and small Lyα peak separation, that are indicative of LyC leakage (Henry et al. 2015; Izotov et al. 2016; Verhamme et al. 2017). SL2S0217 is, therefore, a unique template to probe the conditions of extreme galaxies that may have played a critical role in the reionization of the universe.

5.4. Interstellar Absorption Lines

Outflowing gas, if present, can be directly probed by examining the interstellar absorption line profiles. Given sufficient spectral resolution and S/N in these features, a map of the covering fraction of the absorbing gas as a function of velocity can be inferred for both high- and low-ionization states. To better understand the gas kinematics in SL2S0217, the UV spectrum was normalized using the Rix et al. (2004) continuum windows as a guide, with further continuum designated by eye.

The normalized profiles of the significant absorption-line detections are depicted in Figure 7, arranged by ion. Many of the strong, low-ionization absorption lines that are characteristic of z ∼ 2–3 galaxies (e.g., Pettini et al. 2002b; Shapley et al. 2003) are also present in the SL2S0217 spectrum, but with (∼3×) narrower velocity ranges. We can place constraints on the absorbing gas by measuring the properties of these profiles for various transitions. Using a bootstrap Monte Carlo simulation in which the 1σ uncertainty was used to perturb a given absorption profile and generate 1000 artificial spectra, we measured the flux-weighted centroid, velocity range, and integrated equivalent width for each line. These values, along with the oscillator strengths for each line (Morton 2003), are given in Table 6. The integrated profiles are depicted by the filled absorption features in Figure 7.

Figure 7.

Figure 7. Absorption features of both low- and high-ionization species in SL2S0217. The blue line represents the average absorption profile, composed of the portions of the spectra with thick, solid black lines. The average profile is centered close to 0 km s−1 and is nearly symmetric, with velocity limits of [−392, +319] km s−1 (red dashed vertical lines). In comparison to the f-values (listed in the lower right corner for each line), the strongest lines must be saturated.

Standard image High-resolution image

Table 6.  Interstellar Absorption Lines

Ion λlab f -value va Δvb W0c
  (Å)   (km s−1) (km s−1) (Å)
ii 1334.5323d 0.128 93 −350:+525 1.279 ± 0.119
Si iv 1402.7729 0.254 14 −314 + 282 0.511 ± 0.063
Si ii 1526.7070 0.133 −37 −432:+445 1.136 ± 0.102
Fe ii 1608.4511 0.0577 4 −332:+292 0.643 ± 0.072
Fe ii 1611.2005 0.00138 6 −197:+201 0.237 ± 0.040
Al ii 1670.7886 1.74 19 −317:+342 0.489 ± 0.088
Si ii 1808.0129 0.0021 39 −154:+310 0.271 ± 0.058
Fe ii 2249.8768 0.0018 −95 −299:+61 0.315 ± 0.079
Fe ii 2344.2189 0.114 −83 −323:+149 0.567 ± 0.074
Fe ii 2374.4612 0.0313 −27 −303:+282 0.927 ± 0.107
Fe ii 2382.7652 0.320 −63 −247:+124 0.395 ± 0.087
Ave.     −8 −392:+319 0.805 ± 0.050

Notes. Measurements of the cleanest absorption profiles in SL2S0217. Vacuum wavelengths and f-values are taken from Morton (2003). Columns (4) and (5) list the flux-weighted velocity centroids and velocity ranges over which the profiles were integrated in order to determine the equivalent widths given in Column (6). Characteristics of the error-weighted average absorption profile are given in the final row for comparison.

aVelocity relative to zsys = 1.844 (see Table 2). bVelocity range used for equivalent width measurements relative to line center (Column (2)) at zsys = 1.844. cRest-frame equivalent width. dThe unresolved blend of C ii λ1334.5323 + C ii* λ1335.6627.

Download table as:  ASCIITypeset image

Measurements of two different lines from the same ion can be used to assess the optical depth quantitatively. On the linear part of the curve of growth, W ∝ Nfλ2, where W is the equivalent width, N is the column density, f is the oscillator strength, and λ is the wavelength of a given transition. Then, the theoretical ratio of the Si ii lines is Wλ1527/Wλ1808 = 45.2, such that Si ii λ1527 is expected to be about 45 times stronger than Si ii λ1808 in the optically thin case. Instead, the observed ratio of Wλ1527/Wλ1808 = 1.136 Å/0.271 Å = 4.2 indicates line saturation and optically thick gas. Similar results are found for the low-ionization Fe ii λλ1608, 1611 lines (theoretical Wλ1608/Wλ1611 = 41.7 vs. 2.7 observed), suggesting that the Fe ii λ1608 line is strongly saturated. Deviations are also found for the high-ionization Si iv λλ1394, 1403 lines, but to a smaller degree, as the observed ratio is about half of the theoretical ratio. Therefore, all but perhaps the weakest absorption lines in the UV SL2S0217 spectrum are saturated to varying degrees. In this case, the equivalent width is dependent on both the velocity range and covering fraction of the absorbing gas, but high resolution is needed to model the covering fraction as a function of velocity. Given the low resolution of the UV SL2S0217 spectrum (FWHM ∼ 280 km s−1 at 1500 Å), a more detailed kinematic analysis of individual lines is unfeasible at this time.

5.5. Average Absorption Profiles

Differences between the average low- and high-ionization line profiles can inform gas properties. We determined error-weighted average velocity profiles from the (i) low-ionization, (ii) high-ionization, and (iii) combined absorption features for comparison. Given that the near-UV spectrum is significantly more affected by night-sky contamination and that the Fe ii profiles may be affected by emission filling (see Section 5.6), absorption features from the red side were not used in the average profile. From the blue spectrum, the S ii + Si ii λ1260 feature was excluded from the average owing to its blended nature. Far-UV regions affected by night-sky contamination were also omitted. The combined average profile, therefore, incorporated the appropriate portions of O ı λ1302, Si ii λ1304, Si iv λ1394, Si iv λ1403, Si ii λ1527, Fe ii λ1608, Fe ii λ1611, and Al ii λ1670, indicated by the black profiles in Figure 7.

The resulting profiles are nearly identical. In Figure 7, the combined average profile (blue) is overplotted on the individual absorption features, where the portions of each line used in the average are designated by a solid black line. Apparently the low- and high-ionization gas has similar kinematic properties, as all the observed species seem to be nominally consistent with the average profile. Using a bootstrap Monte Carlo simulation in which the 1σ uncertainty was used to perturb the average profile and generate 1000 artificial spectra, we measured a flux-weighted centroid of −7.8 km s−1, with limits of [−392, +319] km s−1 (designated by red vertical lines in Figure 7), and an equivalent width of 0.80 Å. Despite our reticence to interpret the gas kinematics (due to line saturation and low resolution), the average absorption profile of SL2S0217 is clearly uncharacteristic in terms of both velocity centroid and range relative to interstellar absorption observed in other z ∼ 2 galaxies (e.g., Pettini et al. 2000; Erb et al. 2010; Quider et al. 2010). Rather, the average profile is characterized by nearly symmetric, narrow absorption around the systemic velocity, a signature of little to no outflowing gas.

In Figure 8 we compare the average velocity profiles of the low- and high-ionization absorption lines to the O iii] emission features present in the UV spectrum of SL2S0217. Evidently, the absorption velocity profiles are remarkably similar to those of the emission lines, whose widths represent the instrumental profile, providing a vivid demonstration of the lack of outflows in SL2S0217.

Figure 8.

Figure 8. Comparison of the emission and absorption velocity profiles in SL2S0217. The strengths of the emission lines are arbitrarily scaled for visualization. An average profile was determined for the low-ionization species by combining the individual profiles (see Figure 7), whereas Si iv λ1403 alone was used to represent the high-ionization profile. Remarkably, the emission and absorption profiles are well matched in terms of line width and velocity centroid.

Standard image High-resolution image

5.6. Fine-structure and Resonant Emission Lines

Fine-structure lines, such as Si ii* λλ1265, 1309, 1533, C ii* λ1335, and Fe ii* λλ2365, 2396, are clearly visible and redshifted in other z ∼ 2 emission-line galaxies such as BX418 (emission; Erb et al. 2010) and cB58 (absorption; Pettini et al. 2000), but they are not detected in the UV spectra of local starbursts (e.g., Schwartz et al. 2006; Leitherer et al. 2011). Additionally, previous studies have found correlations between strong Fe ii* features and/or Mg ii λλ2796, 2803 resonant emission and galaxies with extreme properties similar to SL2S0217, namely, high sSFRs, low stellar masses, blue UV slopes, and large Lyα equivalent widths (e.g., Erb et al. 2012; Jones et al. 2012; Kornei et al. 2013). Based on these properties alone, we might have expected to see Fe fine-structure or Mg resonant emission in our spectrum of SL2S0217; however, we find no evidence of these features. On the other hand, from a sample of 184 z ∼ 1 star-forming galaxies, Du et al. (2017) find that Fe ii* strength does not vary significantly with stellar mass or color. While the significance of these trends is still a matter of debate, the fact that we do not significantly detect any of the Fe ii* or Mg ii features is interesting.

Previous attempts to reproduce the fine-structure emission lines using cloudy photoionization models have been unsuccessful, suggesting that these lines are not likely produced by nebular photoionization (Shapley et al. 2003; Erb et al. 2010). Erb et al. (2010) measure a consistent redshift of the fine-structure lines in the z = 2.3 galaxy Q2343-BX418 by ∼200 km s−1 and suggest an origin in the outflowing gas, potentially through resonant scattering, but they caution that this theory is incomplete, as it does not explain the narrowness of the observed lines. Rubin et al. (2011) proposed that Fe ii* emission, observed at or near the systemic velocity, arises from photon scattering in outflowing gas and so may trace the spatial extent of the outflows. This scenario is supported by the nearly systemic mean velocity centroid for the sample of 96 z ∼ 1–2 star-forming galaxies analyzed by Erb et al. (2012; Δv2626 = −20 km s−1).

The absence of Fe ii fine-structure features in SL2S0217 could then be due, in part, to the lack of outflowing gas observed along our line of sight (see discussion of the velocity structure of the absorption lines in Section 5.4). Nevertheless, if scattering in outflows is in fact the production mechanism responsible for Fe ii* emission, then extended emission in transverse outflows and subsequent slit losses are possible. Note, however, that these weak features may simply not be detectable given the low resolution and S/N of the LRIS spectrum.

Galactic outflow models have shown that the overall profiles of the Fe ii and Mg ii lines can be significantly altered by the effects of photon scattering and re-emitted photons (Scarlata & Panagia 2015). Emission filling, where the re-scattered emission "fills in" a portion of the absorption profile, may then be responsible for the bluer velocity centroids and reduced equivalent widths of many of the near-UV Fe ii profiles depicted in Figure 7 and listed in Table 6 (see also Erb et al. 2012). The effects of re-scattered emission may be even greater for Fe ii λ2383, whose only allowed transition after absorption is back to the ground state.

Mg ii absorption profiles may be filled by both resonance re-emission and nebular emission (e.g., Henry et al. 2018). From their composite spectra, Du et al. (2017) measure decreasing Mg ii absorption (−7 Å < W0,Mg ii < −1 Å) with increasing C iii] emission, attributing the effect to nebular emission filling. From the scatter in the LRIS spectrum, we estimate the upper limit of the Mg ii features to be W0,Mg ii ∼ 0.5 Å; only a small residual feature is possible, as expected for nebular and, possibly, significant re-scattered emission-line filling in a target with little to no outflows.

6. Absorption-line Abundances

While most of the absorption lines in the UV spectrum of SL2S0217 are likely saturated, the weakest lines offer the best opportunity to constrain relative abundances. Using the integrated equivalent widths from Table 6, we estimated the column density for the weakest lines observed, assuming optically thin conditions (apparent optical depth method). From Spitzer (1978), the linear part of the curve of growth can be characterized as

Equation (1)

where N is the column density, λ is the transition wavelength, and f is the oscillator strength (see Table 6). The resulting N values are listed in Table 7; however, if any of the lines are saturated, the column densities determined here should be considered lower limits.

Table 7.  Interstellar Abundances

Ion λ(Å) log $({N}_{{\rm{H}}{\rm{I}}}$ cm−2) log(X/H) log(X/H)b [X/H]c
H ı 1215 20.0a
Fe ii 1611 15.874 −4.126 −4.50 0.374
Si ii 1808 15.649 −4.351 −4.49 0.139

Notes. Interstellar abundances estimated from the weakest absorption features, assuming the optically thin case.

aValue of log(NH i cm s−1) predicted by homogeneous shell models with no expansion velocity, b = 10 km s−1, and Δv = 371 km s−1 (Verhamme et al. 2015). bSolar (meteoric) abundances from the compilation by Asplund et al. (2009). c[X/H] = log(X/H) - log(X/H).

Download table as:  ASCIITypeset image

Because the Lyα profile is seen purely in emission (see discussion in Section 5.3), we do not have a measure of the total column density of neutral hydrogen. As discussed in Section 5.3, the static, homogeneous radiative transfer shell models of Verhamme et al. (2015) predict a column density of neutral gas of ${N}_{{\rm{H}}{\rm{I}}}={10}^{20}$ cm−2 for the Lyα peak separation in SL2S0217. Adopting this value as indicative, we subsequently estimated the element abundances for Si ii and Fe ii, as reported in Table 7.

Since Si ii and Fe ii both have lines with very weak and similar oscillator strengths, assuming that these lines are unsaturated, we can estimate the Si ii/Fe ii ratio (independent of ${N}_{{\rm{H}}{\rm{I}}}$). Further, Si ii and Fe ii have similar ionization potentials and are the only ionization states that we observe in absorption for Si and Fe, so the ionization correction factor (ICF) will be small and, to first order, can be ignored. Using our measured values for SL2S0217 in Table 7, we find log(Si/Fe) = −0.225 ± 0.118, or 44%–77% solar. However, Si is expected, both empirically and theoretically, to be overabundant relative to Fe in metal-poor dwarf galaxies (e.g., Tolstoy et al. 2009). This is because Si is predominantly an alpha element (α), produced on relatively short timescales during Type II supernova (SN II; massive stars) explosions, whereas Fe is synthesized by SNe Ia (intermediate binaries with mass transfer) and returned much later. Since SNe Ia have longer timescales than SNe II, Si and Fe trace different stellar populations and timescales. Enhanced [α/Fe] occurs until SNe Ia begin to contribute to the chemical evolution 108–109 yr after the first episode of star formation.

Given the derived properties of SL2S0217, one may have expected a pronounced α-enhancement in a galaxy experiencing a young burst of star formation. Instead, the abnormal Si/Fe ratio measured for SL2S0217 could be evidence of an old stellar population that has had sufficient time to enrich the gas in Fe. Alternatively, the [α/Fe]-poor gas could be the result of larger Si than Fe depletion onto dust grains, although the authors of this work are not aware of any empirical evidence to support this scenario. The UV emission-line ratios indicate that the Si depletion may be as great as 65% in SL2S0217 (see Section 7.2.3). Unfortunately, we cannot estimate the Fe depletion from emission lines with existing data. The Si abundance and dust depletion are further discussed in Section 7.2.3.

7. Nebular Emission-line Abundances

7.1. Physical Parameters of the ISM

Directly measuring abundances from nebular emission lines requires knowledge of the physical properties of the emitting gas. While detailed observations of nearby star-forming regions reveal complex nebular structures (e.g., Pellegrini et al. 2012), simplified H ii region models leverage a spherical geometry with three separate ionization volumes. The advantage of such models is that they account for the fact that the low-, intermediate-, and high-ionization zones are governed by different physical properties and therefore require reliable electron temperature and densities for each volume to determine accurate abundances. Although the SL2S0217 arc is evidently composed of multiple star-forming regions, they are not easily disentangled from the UV spectrum, and so we employ a single H ii region model to estimate the nebular gas properties. The electron temperature and electron density of SL2S0217 were determined using the PyNeb package in python (Luridiana et al. 2012, 2015), assuming a five-level atom model (De Robertis et al. 1987).

7.1.1. Electron Temperature

Electron temperature is typically determined by observing a temperature-sensitive auroral-to-strong-line ratio. Historically, the [O iii] λ4363/[O iii] λ5007 emission-line ratio has been the ideal measure of electron temperature in the high-ionization zone of nebulae. However, the temperature-sensitive [O iii] λ4363 auroral line is not resolved in the HST grism spectrum.

Fortunately, the electron temperature can also be determined from the O iii] λ1666/[O iii] λ5007 ratio, as is commonly done in high-redshift targets where the intrinsically faint optical auroral line is often undetected. In the case of SL2S0217, this diagnostic combines space- and ground-based observations, potentially introducing flux and aperture mismatching issues, and so this calculation must be done carefully. Given the excellent flux calibration of the HST grism spectrum and the match of the flux-corrected LRIS continuum to the HST ACS F606W AB magnitude, we used the O iii] λ1666/[O iii] λ5007 ratio to measure Te,λ1666 = 15,400 ± 200 K. Adopting this Te,λ1666 measurement for the high-ionization zone electron temperature, the intermediate- and low-ionization zone temperatures (Te [S iii] and Te [O ii], respectively) were then determined from the theoretical temperature relationships of Garnett (1992). The temperatures used for each ionization zone are listed in Table 8.

Table 8.  Ionic and Relative Nebular Abundances

ne used 100 cm−3 4500 cm−3
Te [O iii]opta 15,700 ± 3200 K
Te [O iii]combb used 15,400 ± 200 K
Te [S iii] used 14,500 ± 200 K
Te [O ii] used 13,800 ± 200 K
neiii] measured ${300}_{-300}^{+1300}$ cm−3
ne Si iii] measured ${4500}_{-1,400}^{+1,500}$ cm−3
O+/H+a,c 5.21 × 10−7 8.77 × 10−7
O+2/H+a 3.14 ± 0.12 × 10−5 3.13 ± 0.12 × 10−5
ICFd 1.055
12 + log(O/H)a ≥7.50e ≥7.51e
C+3/C+2 0.86 0.86
C+2/O+2 0.122 ± 0.005 0.124 ± 0.005
log U −1.50
ICF 1.27 ± 0.27
log(C/O) −0.81 ± 0.09 −0.80 ± 0.09
N+2/O+2 0.033 ± 0.062 0.029 ± 0.054
log(N/O) −1.48 ± 0.46 −1.53 ± 0.45
Si+2/C+2 0.072 ± 0.002 0.073 ± 0.002
ICF 2.31 ± 0.32
log(Si/C) −0.76 ± 0.09 −0.77 ± 0.09
log(Si/O) −1.59 ± 0.17 −1.58 ± 0.17

Notes. Physical conditions and abundances derived for the nebular gas in SL2S0217 assuming either ne = 100 cm−3 (consistent with the ne from C iii]) or ne = 4500 cm−3 (from Si iii]). Ionic abundances are reported for both the optical and UV species when available.

aDetermined from the HST optical grism spectrum values. bDetermined from the combined UV and optical λ1666/λ5007 line ratio. cThe [O ii] λ3727 emission-line intensity was inferred from the observed [O iii] λ5007 flux and the [O iii] λ5007/[O ii] λ3727 ratio from cloudy photoionization modeling. dThe oxygen ICF accounts for unseen O+3, estimated from the 3σ flux upper limit of O iv] λλ1401, 1404. eBecause we do not detect [O ii], the estimated oxygen abundance is reported as a lower limit. However, the O+ contribution is expected to be small at such high ionization.

Download table as:  ASCIITypeset image

Interestingly, Brammer et al. (2012) reported excess emission from the Hγ + [O iii] λ4363 blend in the grism spectrum, which can be attributed to [O iii] λ4363 and used to constrain the optical electron temperature diagnostic. We determined the Hγ flux contribution to the blend using the measured Hβ grism flux ((74 ± 2) × 10−17 erg s−1 cm−2) and the Case B Hγ/Hβ value from Hummer & Storey (1987) assuming Te = 15,000 K and ne = 100 cm−3 (conditions representative of SL2S0217; see Section 7.1.2 for ne). Subtracting this Hγ value from the blend ((40 ± 3) × 10−17 erg s−1 cm−2; B12), the flux corresponding to [O iii] λ4363 is Fλ4363 = 4.9 × 10−17 erg s−1 cm−2. We use the standard [O iii] λ4363/[O iii] λ5007 temperature diagnostic to estimate the high-ionization zone electron temperature to be Te,λ4363 = 15,700 ± 3200 K, in agreement with the combined UV/optical temperature diagnostic. A higher-S/N and higher-resolution rest-frame optical spectrum would be useful to more securely measure [O iii] λ4363 and compare the two temperature diagnostics.

7.1.2. Electron Density

In nearby objects, the gas-phase electron density is most commonly determined from the [S ii] and [O ii] optical collisionally excited doublets. For SL2S0217, the [O ii] lines were not detected and the [S ii] lines were outside the wavelength range covered by the HST grism spectrum. Instead, we observe two sets of density-sensitive UV emission-line doublets in our LRIS spectrum: Si iii] λλ1883, 1892 and C iii] λλ1907, 1909. We calculate densities of ${4500}_{-1,400}^{+1,500}$ cm−3 and ${300}_{-300}^{+1300}$ cm−3 for the Si iii] and C iii] ions, respectively.10 Since the relevant lines are all high S/N and should result purely from nebular emission, we find no obvious explanation for this discrepancy.

The critical densities of C iii] and Si iii] are both of the order of 104 cm−3, and so they are generally not useful density diagnostics for ne values below ∼103 cm−3. This is evident in Figure 9, where we plot the emissivity ratios of common rest-frame UV (Si iii] and C iii]) and optical ([O ii] and [S ii]) electron density diagnostic ratios, assuming a general electron temperature of 15,000 K. When measured, the optical line ratios provide constraints below ne ∼103 cm−3; however, they probe a different (low-)ionization zone and indicate significantly lower densities relative to the highly ionized gas in high-redshift star-forming galaxies (e.g., Christensen et al. 2012; Bayliss et al. 2014; James et al. 2014). Traditionally, the high-ionization zone (∼35–55 eV) is represented by O++, whereas the low-ionization zone (∼15–35 eV) is represented by the O+ or N+ ions, with the intermediate zone, based on S++, partially overlapping (∼23–35 eV). By these definitions, the UV Si iii] and C iii] density diagnostics span the low- to intermediate-ionization and intermediate- to high-ionization zones, respectively. Perhaps not surprisingly, the mixed-zone electron densities measured for high-redshift targets, including SL2S0217, are significantly larger than the pure low-ionization zone densities typical of H ii regions in nearby galaxies. Further, Sanders et al. (2016) find that z ∼ 2 galaxies have mean electron densities that are an order of magnitude higher relative to local galaxies at fixed stellar mass.

Figure 9.

Figure 9. Comparison of common rest-frame UV (Si iii] and C iii]) and optical ([O ii] and [S ii]) electron density diagnostic ratios. Emissivities were calculated using the PyNeb package in Python, assuming an electron temperature of 15,000 K. The dashed (dotted-dashed) black line and pink (orange) box represent the Si iii] (C iii]) ratio and uncertainty for SL2S0217. Unfortunately, neither C iii] nor Si iii] are sensitive to low densities, so we simply assume a typical nebular density of ne = 100 cm−3.

Standard image High-resolution image

While the insensitivity of the UV density diagnostics to low densities inhibits the density determination for SL2S0217, the C iii] ratio is at least in agreement with the low-density regime, and so we assumed a standard value of ne = 100 cm−3. Fortunately, the gas temperature and line emissivities are practically independent of the gas density in the low-density regime, and so this choice of density makes little difference. In fact, subsequently derived abundance ratios for both ne = 100 cm−3 (consistent with the ne from C iii]) and ne = 4500 cm−3 (from Si iii]), given in Table 8, show differences that are significantly smaller than the respective uncertainties. However, detailed studies of these diagnostics must be prioritized to understand whether targets such as SL2S0217 have abnormally high electron densities/density inhomogeneities, the UV diagnostics probe a higher-density regime than their optical counterparts, or some other effect is at play.

7.2. Ionic and Total Abundances

Due to the number of strong rest-frame UV emission lines observed in the spectrum of SL2S0217, we were able to determine relative C, O, N, and Si abundances. Using the rest-frame optical grism spectrum, ionic abundances were also determined relative to hydrogen:

Equation (2)

The emissivity coefficients, jλ(i), which are functions of both temperature and density, were determined using the five-level atom approximation and updated atomic data reported in Berg et al. (2015).

Typically, the total oxygen abundance (O/H) in an H ii region is calculated by simply summing the O+/H+ and O+2/H+ ionic abundances, as contributions from O+3/H+ (requiring an ionization energy of 54.9 eV) are thought to be negligible. Although no emission is seen for the strongest UV O iv transitions (λλ1401, 1404) in the LRIS spectrum, the presence of He ii from recombination (requiring an ionization energy of 54.4 eV) makes this argument less secure. To account for possible contributions from O+3, we measured the 3σ upper limit on the flux at λλ1401, 1404 to be ≤3.3 × 10−18 erg s−1 cm−2. Comparing this value to the O iii] λλ1661, 1666 line strengths, we find a marginal oxygen ICF of 1.055.

For SL2S0217 we can measure O+2/H+ from the [O iii] λλ4959, 5007/Hβ optical emission-line ratio; however, the corresponding lines for O+/H+ are not detected in the optical grism spectrum. In order to estimate 12+log(O/H) from the optical spectrum, we used the observed [O iii] λ5007 emission and results from photoionization models (see Section 8.1) to predict the [O ii] λ3727 emission-line flux. Based on these models, O+/H+ only contributes 2% to the total oxygen abundance (and O0 is truly negligible), but such a small contribution is expected at the extreme ionization level of SL2S0217. Therefore, we find SL2S0217 to be an extremely metal-poor (EMP; 12+log(O/H) < 7.7) galaxy, with an oxygen abundance, corrected for the unseen O+/H+ and O+3/H+ contributions, of 12 + log(O/H) = 7.50. The ionic and total oxygen abundances are listed in Table 8.

7.2.1. C/O Abundance

Abundance determinations for other elements require ICFs to account for their unobserved ionic species. The prominence of the high-ionization-potential He ii and C iv emission lines observed in SL2S0217 indicate that ICFs are particularly important for C. Although the relatively narrow emission profiles of the C iv doublet indicate the dominance of nebular emission, it is complicated by potential contributions from a range of stellar features, from pure absorption to P Cygni profiles. To avoid this complication, we used the ICF presented in Berg et al. (2016) to estimate the C+3 contribution at log U = −1.50 (our best model value; see Section 8.1). Despite the strong C iv emission present in the UV spectrum of SL2S0217, we determine a modest C ICF = 1.27 ± 0.27. The C/O abundance is then determined using

Equation (3)

to be log(C/O) = −0.81 ± 0.03, which is typical of star-forming dwarf galaxies with similar oxygen abundance (Garnett et al. 1995b; Berg et al. 2016).

Carbon and oxygen are produced on different timescales, by different mass populations of stars, and so the C/O trend with metallicity has been studied by many authors with observations for a variety of astrophysical objects, and using a variety of chemical evolution models and stellar yields. Oxygen is synthesized by SNe II (massive stars) and returned to the ISM quickly, whereas carbon is primarily produced by He burning through the triple-α process, a reaction that can occur in both massive (M > 8 M) and low- to intermediate-mass (1 M < M < 8 M) stars.

Historically, C/O has been difficult to measure, but recently C/O has been measured in dozens of nearby, metal-poor, dwarf galaxies using the UV lines (e.g., Berg et al. 2016; Senchyna et al. 2017). Similar to the early work of Garnett et al. (1995b), these studies find a general trend of increasing C/O with oxygen abundance, but with a significant unexplained scatter. These data can also be interpreted as a constant trend in C/O at low metallicities. We plot SL2S0217 (open star) on this C/O versus O/H trend in the left panel of Figure 10, where nearby dwarf galaxies are plotted as filled circles for comparison. The trend is extended to higher oxygen abundances by incorporating C/O determinations from optical recombination lines (filled squares: Esteban et al. 2002, 2009, 2014; Pilyugin & Thuan 2005; García-Rojas & Esteban 2007; López-Sánchez et al. 2007).

Figure 10.

Figure 10. Left: carbon-to-oxygen ratio vs. oxygen abundance of SL2S0217 (black star) in comparison to other star-forming galaxies. Comparison data of nearby galaxies are divided into two groups. Red circles are metal-poor dwarf galaxies as reported in Berg et al. (2016) and D. A. Berg et al. (2018b, in preparation), and green filled squares represent star-forming galaxies with larger oxygen abundances determined from recombination lines (Esteban et al. 2002, 2009, 2014; Pilyugin & Thuan 2005; García-Rojas & Esteban 2007; López-Sánchez et al. 2007). The dashed line is the least-squares fit from Garnett et al. (1995b, G95), and the dotted line is the weighted mean of the significant CEL C/O detections (filled red circles). More distant (z ∼ 2−3) galaxies are represented as gold triangles (Pettini et al. 2000; Fosbury et al. 2003; Erb et al. 2010; Rigby et al. 2011; Christensen et al. 2012; Bayliss et al. 2014; James et al. 2014; Stark et al. 2014; Steidel et al. 2016; Vanzella et al. 2016; Amorín et al. 2017; Rigby et al. 2018), including composite spectra from Steidel et al. 2016, Amorín et al. (2017), and Rigby et al. (2018). The background shaded region demonstrates the value for the z ∼ 3 Lyman break galaxy composite spectrum determined by Shapley et al. (2003). Right: trends in relative nitrogen abundance vs. oxygen abundance for galaxies spanning a range in redshift. Additional direct abundances are plotted for nearby dwarf galaxies (van Zee & Haynes 2006; Berg et al. 2012) and for nearby spiral galaxies (blue points: the CHAOS project; Berg et al. 2015; Croxall et al. 2015, 2016; D. A. Berg et al. 2018a, in preparation). SL2S0217 appears to have a normal N abundance with respect to both nearby, metal-poor dwarf galaxies and the few other z ∼ 2−3 galaxies with C/N observations.

Standard image High-resolution image

Significant UV C and O have also been measured for a handful of z ∼ 2−3 galaxies (gold triangles in Figure 10: Pettini et al. 2000; Fosbury et al. 2003; Erb et al. 2010; Christensen et al. 2012; Bayliss et al. 2014; James et al. 2014; Stark et al. 2014; Steidel et al. 2016; Vanzella et al. 2016; Amorín et al. 2017; Rigby et al. 2018), among which the C/O ratio of SL2S0217 appears to be average. Interestingly, C/O values for the z ∼ 2−3 galaxies are consistent with the nearby dwarf galaxies of similar metallicity, but potentially with a lower average value. The distinction between these populations of different ages may support the notion put forth by Berg et al. (2016) that higher C/O values may be due to a delayed pseudo-secondary carbon contribution from low- and intermediate-mass stars.

7.2.2. Relative N Abundance

We estimate the N/O ratio by combining the UV O iii] λλ1661, 1666 and C iii] λλ1907, 1909 emission lines and the N iii] λ1750 multiplet. Since the N iii] λ1750 feature is only weakly detected in the SL2S0217 LRIS spectrum, it is appropriate to think of the following N/O determination as an upper limit. Utilizing the fact that the ionization potentials of N++ and C++ are nearly the same (47.448 and 47.887 eV, respectively), we used the N++/C++ ion ratio to measure a relative N/C enrichment of log(N/C) = −0.68 ± 0.02. Then, combining this ratio with the ionization-corrected C/O ratio (see Table 8), we inferred a total N/O of log(N/O) = −1.48 ± 0.46, which is consistent with the range of log(N/O) seen for local dwarf galaxies with similar oxygen abundance (e.g., van Zee & Haynes 2006; Berg et al. 2012).

Relative N/O abundances are well studied in nearby dwarf and spiral galaxies, providing a robust sample to compare measurements from more distant targets. In the top right panel of Figure 10 we compare the N/O of SL2S0217 to nearby dwarf (red points: van Zee & Haynes 2006; Berg et al. 2012, 2016; D. A. Berg et al. 2018b, in preparation) and spiral galaxies (blue points: the CHAOS project; Berg et al. 2015; Croxall et al. 2015, 2016; D. A. Berg et al. 2018a, in preparation), as well as nearby galaxies with oxygen abundance measurements from recombination lines (green squares). SL2S0217 lies among the N/O range of both nearby dwarf galaxies and the handful of z ∼ 2−3 galaxies with N/O and O/H measurements (gold triangles). Note that the N/O value for SL2S0217 falls on the plateau expected for primary (metallicity-independent) N production, as expected for a low oxygen abundance of 12+log(O/H) ∼ 7.5 (see, e.g., the discussion in Pettini et al. 2002a; Pettini & Cooke 2012).

In the bottom right panel of Figure 10 we examine the C/N ratio, noting the remarkably flat trend for both nearby and distant galaxies. Berg et al. (2016) found that the constant ratio of C/N seen over a large range in O/H for nearby galaxies may indicate that carbon is predominantly produced by similar nucleosynthetic mechanisms to nitrogen (see also Steidel et al. 2016). However, a larger sample of C/N measurements in distant galaxies is needed to determine whether the hypothesis of Berg et al. (2016) also applies at z ∼ 2−3, or whether different nucleosynthetic processes dominate C and N.

Recently, observations of the standard rest-frame optical emission-line [O iii]/Hβ versus [N ii]/Hα diagnostic plot (the BPT diagram; Baldwin et al. 1981) have revealed a pronounced offset of ∼2–3 galaxies from the main sequence of local star-forming galaxies (e.g., Masters et al. 2014; Steidel et al. 2014; Shapley et al. 2015). Several studies have investigated the physical origin of this offset by analyzing the BPT parameter space in terms of other physical properties for large surveys of local galaxies. Using ∼100,000 galaxies from the Sloan Digital Sky Survey Data Release 12, Masters et al. (2016) found that the region of the BPT occupied by ∼2–3 galaxies corresponds to galaxies with elevated SFR surface densities, stellar masses, and N/O ratios, concluding that the higher N/O ratios at fixed [O iii]/Hβ are the proximate cause of the offset. In contrast, the sample of z ∼ 2−3 galaxies plotted in Figure 10 occupies the same range in N/O as the local comparison sample. SL2S0217, in particular, has an average N/O abundance relative to the local dwarf galaxies; however, this value is an upper limit estimated from the weak detection of the N iii] λ1750 line. Further, despite the unusual strength of its UV and optical emission lines, the [O iii]/Hβ and [N ii]/Hα (inferred from photoionization modeling; Section 8.1.1) ratios of SL2S0217 indicate that it may not coincide with the offset region of the BPT. Clearly there is still much to understand about the z ∼ 2−3 BPT diagram.

7.2.3. Relative Si Abundance

Silicon abundances relative to carbon were determined following the logic presented in Garnett et al. (1995a): the Si iii] and C iii] doublets arise from comparable energy levels (∼6.57 and 6.50 eV, respectively), and therefore the Si/C abundance ratio is relatively insensitive to uncertainties in the electron temperature. Additionally, Si/C depends little on the assumed density, as both Si iii] and C iii] have very high critical densities, meaning that their volume emissivities have a similar dependence on density. Finally, the proximity of the Si iii] and C iii] emission lines in wavelength eliminates strong effects of differential extinction.

The Si iii] and C iii] emission-line features are extremely prominent in SL2S0217; we measured their strengths with S/N > 20 (see Figure 5). As described in Section 7.1.2, we assume that the ionized gas of SL2S0217 is in the low-density limit (<103 cm−3), where collisional de-excitation does not significantly impact the Si+2/C+2 ratio derived from the UV lines. Because C+2 has a larger ionization potential than Si+2 (47.9 and 33.5 eV, respectively), we expect a larger fraction of Si to be in a higher, unobserved ionization state than C, and so an ICF is needed to determine the Si/C abundance. We determine the Si/C element abundance using the equation

Equation (4)

where X(Si+2) and X(C+2) are the Si+2 and C+2 volume fractions, respectively. The ICF is modeled as a function of the ionization parameter using the photoionization code cloudy (Ferland et al. 2013), with BPASS models as the input ionizing radiation field (see Section 8.1). We sought simple models that reproduced the conditions in SL2S0217 as closely as possible; see Section 8.1 for a detailed discussion.

The ionization fractions of Si and C species as a function of ionization parameter are shown in Figure 11. As we will show in Section 8.1.1, we can reproduce many aspects of the observed spectrum using cloudy models of a stellar population with binaries and an ionization parameter of log U = −1.5; this value is reasonable given the very hard ionizing radiation field predicted by the SED modeling of B12. We estimate the uncertainty in the ICF as the scatter among the different models considered at a given C+2 volume fraction. The resulting Si ICF = 2.31 ± 0.32 for 12+log(O/H) = 7.5 and t = 107.0 yr is significant, correcting the relative Si abundance from log(Si/C) = −1.14 to log(Si/C) = −0.76 ± 0.09. When combined with the C/O abundance derived for SL2S0217, we estimate log(Si/O) = −1.59 ± 0.17. The relevant Si abundance values are reported in Table 8.

Figure 11.

Figure 11. Ionization fraction of C and Si species as a function of ionization parameter from cloudy models of Zneb = 0.05 Z (Z = 0.001) BPASS stellar models. Two different ages of instantaneous bursts are shown, demonstrating the limits of the parameter space explored: t = 106.4 and 107.0 yr. For the older model, symbols are color coded by the gas-phase oxygen abundance, whereas the younger model is depicted by gray symbols. In the top two panels the different ionic species are designated by triangles, circles, squares, and stars in order of increasing ionization. The bottom panel plots the ICF vs. ionization parameter, where the dashed and dotted-dashed curves indicate the ICF for the 12+log(O/H) = 7.5 model at an age of 106.4 and 107.0 yr, respectively, and the gray shading highlights the range spanned by metallicity. The gray vertical dashed line marks the predicted ionization parameter for SL2S0217.

Standard image High-resolution image

Due to the need of far-UV spectroscopic observations of nearby galaxies, there exists a general paucity of nebular measurements of Si/O to compare to. Perhaps the best-known study of Si/O comes from Garnett et al. (1995a), who measured the UV Si iii] emission lines for seven extragalactic H ii regions using the Faint Object Spectrograph on HST. These authors found Si/O to be constant over the observed range in oxygen abundance, with a weighted mean value of log(Si/O) = −1.59 ± 0.07. Our log(Si/O) = −1.59 ± 0.17 measurement for SL2S0217 is remarkably consistent with this average from the Garnett et al. (1995a) study but indicates significant Si/O enrichment or a reduced Si depletion onto dust relative to the z ∼ 2.4 galaxies of Steidel et al. (2016; log(Si/O) = −1.81 ± 0.10).

Note that our Si/O value does not account for the relative depletion of Si and O onto dust grains. While one might expect this effect to be small given the low predicted dust reddening in SL2S0217 and the probability of grain destruction or erosion in the presence of a hard ionizing radiation field, our Si/O value is significantly lower than the solar ratio of log(Si/O) = −1.15 (Jenkins 2009). Many studies argue that both Si and O are predominantly products of massive star nucleosynthesis, and so their ratio is expected to vary little with time (e.g., Timmes et al. 1995). Note, however, that some studies find that Si production can have a substantial contribution from SN Ia outbursts (see Matteucci & Greggio 1986; Kobayashi et al. 2006). Since oxygen does not deplete as strongly, the observed underabundance can likely be attributed to the depletion of Si onto dust (Steidel et al. 2016).

We estimated the logarithmic depletion of Si in the ISM following the prescription laid out in Jenkins (2009):11

Equation (5)

where, assuming that the observed subsolar Si/O ratio is solely due to Si depletion, we use Si/O in place of Si/H and use the solar values from Asplund et al. (2009) for our reference. This results in a dust depletion of [Si/H]gas = −0.43 for SL2S0217, a value typical with respect to Garnett et al. (1995a), but smaller than the value derived by Steidel et al. (2016). From observations of nearby dwarf galaxies we have come to expect [Si/Fe] enrichment at low [Fe/H] due to early contamination of Si by SNe II. However, this exercise indicates that a significant fraction of the Si produced in SL2S0217 is missing, perhaps resulting in the subsolar Si/Fe ratio measured from the absorption lines (Section 6).

Jenkins (2009) interpreted the abundances of 17 different elements from more than 100 studies of sight lines to 243 different stars to construct a unified representation of depletions that is linearly dependent on the severity of depletion along a given sight line, F*. Using this model and the inferred depletion value of Si, we estimated the depletion of the other significant elements in the SL2S0217 UV spectrum—C, N, and O. Using the fit from Jenkins (2009) for Si, [X/H]gas = BX + AX(F* − zX) (coefficients reported in their Table 4), we found a depletion severity of F* = 0.182. Then, we determine the following dust depletion factors: δ(Si) = 0.628, δ(C) = 0.259, δ(N) = 0.222, and δ(O) = 0.11112 ([C/H]gas = −0.130, [N/H]gas = −0.109, and [O/H]gas = −0.051). While these depletion values imply a systemically higher relative Si/O abundance (Δ ∼ 0.4 dex), the O/H, C/O, and N/O ratios would remain relatively unchanged (Δ ∼ 0.05 dex).

8. Discussion

8.1. Photoionization Models

While the number of objects with detections of extreme UV emission-line features continues to grow, many questions remain regarding their ionizing sources. To gain insight, we investigated the potential to produce strong UV emission lines from an ionizing radiation field powered by stars. Previous studies have also used photoionization models to examine the factors that regulate UV emission lines, focusing on C iii] in particular. Stark et al. (2014) found that metal-poor (Zneb < 0.4 Z), young (<50 Myr) galaxies with subsolar C/O ratios and large ionization parameters were capable of reproducing the large C iii] EWs observed for high-redshift galaxies. More recently, Jaskot & Ravindranath (2016) created a large grid of photoionization models of star-forming galaxies to examine C iii] EWs by varying starburst age, metallicity, ionization parameter, C/O abundance, dust content, gas density, optical depth, and nebular geometries. These authors found that low metallicities and high-ionization parameters enhance C iii] emission. Further, the largest C iii] EWs were produced by stellar populations that incorporated binary interactions among massive stars. Note, however, that models that include stellar rotation have also been shown to extend the lifetime (up to 5–7 Myr) and increase the strength of nebular emission (e.g., Byler et al. 2017; Choi et al. 2017). We focus here on models using binaries alone, as they can produce nebular emission up to ∼10 Myr, but note that likely both rotation and binaries play a role in the observed line strengths.

Our primary goal was to reproduce the strong UV emission- line ratios observed for SL2S0217. Building on the lessons of Jaskot & Ravindranath (2016), we ran a grid of photoionization models using cloudy17.00 in which ionization parameter, gas-phase metallicity, and relative C, N, and Si abundances were varied. For the input ionizing radiation field, we used a suite of models from the new release of "Binary Population and Spectral Synthesis" (BPASSv2.14; Eldridge & Stanway 2016; Stanway et al. 2016), with an initial mass function (IMF) index of −1.30 for 0.1–0.5 M and −2.35 for M > 0.5 M.

We ran six sets of models depending on three variables: whether (1) an IMF upper mass cutoff of 100 or 300 M was used for (2) continuous or a burst of star formation, and (3) whether or not binary evolution was included. Limited parameter ranges were informed by the SED modeling of Brammer et al. (2012) for the stellar population properties and by the UV emission-line ratios measured from the LRIS spectrum for the nebular gas properties. In particular, Jaskot & Ravindranath (2016) found that C iii] EWs as strong as those measured for SL2S0217 (>10 Å) only result from a hard radiation field (log U > −2). Each model set considered an age range of 106.4–107.0 yr, a range in ionization parameter of −2.0 < log U < −1.0, with stellar metallicity of Z = 0.001 (Z = 0.05 Z) to match the estimated gas-phase abundance (12+log(O/H) = 7.5 or Zneb = 0.065 Z). The GASS10 solar abundance ratios within cloudy were used to initialize the relative gas-phase abundances. These abundances were then scaled to cover a range in nebular metallicity (7.25 < 12 + log(O/H) < 8.25) and relative C, N, and Si abundances (0.1 < (X/O)/(X/O) < 0.5). The range in relative N/O, C/O, and Si/O abundances was motivated by our measured values for SL2S0217 and the observed values for nearby metal-poor dwarf galaxies (e.g., Garnett et al. 1999; Berg et al. 2012, 2016). The resulting X/O = [0.1, 0.3, 0.5] X/O model parameters (written as CNSi = [0.1, 0.3, 0.5] for shorthand) were log(C/O) = [−1.26, −0.78, −0.56], log(N/O) = [−1.86, −1.38, −1.16], and log(Si/O) = [−2.18, −1.70, −1.48].

8.1.1. Best Model

To determine the best model from our grid, we first eliminated sets of models that were obviously poor matches. The single-star model sets do not reproduce the UV nebular emission-line feature strengths (see also Jaskot & Ravindranath 2016) and so were not investigated further. For the young ages considered here (t < 107.2 yr), the continuous star formation models are consistent with the burst models, and so only the burst models were examined as representative of both types of star formation. Note, however, that continuous star formation models do produce very different results from bursts for older systems. Finally, the models with an IMF cutoff of 300 M do produce more He ii ionizing photons than the 100 M models, yet they still fall short of producing the He ii flux observed for SL2S0217 and significantly overpredict the Si iii]/C iii] ratio. Therefore, we employed an error-weighted χ2-minimization routine of the grid of binary burst models with an IMF cutoff of 100 M. We find we can reproduce most of the features in the UV emission-line spectrum of SL2S0217 with a model similar to the SED modeling of Brammer et al. (2012): the best model is a high-ionization (log U = −1.5), metal-poor (12+log(O/H) = 7.75; CNSi = 0.3) gas cloud, ionized by Z = 0.001 star formation with binaries at an age of t = 107.0 yr.

In Figure 12 we compare the normalized LRIS spectrum of SL2S0217 to the output model spectrum at the spectral resolution of the observed spectrum. The C iv, O iii], N iii], Si iii], and C iii] emission lines are all relatively well matched, while the He ii emission feature is clearly discrepant. Notably, Fosbury et al. (2003) also observed strong, narrow nebular He ii and C iv emission from the z ∼ 3.4 Lynx Arc, reportedly due to a young, low-metallicity stellar cluster with an extreme ionization parameter of log U = −1.0. However, to date, no pure stellar photoionization studies have reported the levels of nebular He ii emission observed in SL2S0217 or the Lynx Arc. Interestingly, we do not see the N iv] emission line in our SL2S0217 spectrum despite the fact that we see strong C iv emission and C++ and N++ have similar ionization potentials (47.888 and 47.445 eV, respectively). The N iv] emission line is also absent from the best model, indicating that N emission is mostly sensitive to metallicity and, therefore, not expected in metal-poor galaxies like SL2S0217, whereas C iv is more sensitive to temperature/cooling. This fact has motivated the use of the strong-line [N ii]/Hα metallicity indicator in nearby galaxies; however, the line ratio is not available from the ground at high redshifts (z ≳ 2.5) and is highly sensitive to ionization parameter at larger metallicities (e.g., Kewley & Dopita 2002). Since exceptionally low nitrogen abundances seem to be common among distant galaxies with similar extreme emission-line features (e.g., Bayliss et al. 2014; James et al. 2014; Smit et al. 2017), the lack of UV N iii] and N iv] could help identify additional metal-poor galaxies at high redshifts.

Figure 12.

Figure 12. The photoionization model (orange line) best matched to the blue arm of the UV SL2S0217 spectrum. SL2S0217 can be described by a high-ionization (log U = −1.5), metal-poor (12+log(O/H) = 7.75; X/O = 0.3 X/O) gas cloud, ionized by a Z = 0.001 starburst BPASS binary model at an age of t = 107.0 yr. All of the emission features, except He ii, which is notoriously difficult to emulate with pure photoionization models, are reproduced with this model.

Standard image High-resolution image

The predicted nebular emission-line ratios from our photoionization models are plotted relative to the ratios measured from the LRIS spectrum of SL2S0217 in Figure 13. Different gas-phase metallicity models are indicated by color, where relative C/O, N/O, and Si/O abundances increase with point size. The statistically best model is indicated by the dashed yellow line (12+log(O/H) = 7.75, CNSi = 0.3, t = 107.0 yr), where the corresponding yellow-shaded region extends from t = 106.4 to 107.0 yr. Note that nebular He ii emission peaks before t = 107.0 yr (∼106.8 yr) in the models, and so the dashed yellow line in the He ii/C iii] plot lies within the yellow shaded region. Interestingly, although C iv also has a very high ionization potential, its nebular line strength continues to increase past t = 107.0. Measured line ratios and uncertainties for SL2S0217 are shown as solid black horizontal lines and gray shaded boxes. Good agreement between model and measurement is found near an ionization parameter of log U = −1.50 for all of the UV emission-line ratios except He ii/C iii].

Figure 13.

Figure 13. UV emission-line ratios from cloudy photoionization models using the BPASS Z = 0.001 Z binary models for the ionizing source, assuming an IMFup = 100 M over a range in age of 106.4–107.0 yr. The color-coded points represent a range of gas-phase metallicities assumed: 7.25 ≤ 12+log(O/H) ≤ 8.25. We also explored a range in C/O, N/O, and Si/O abundances relative to their respective solar ratios, as indicated by various point sizes (see description in Section 8.1). Observed line ratios and corresponding uncertainties for SL2S0217 are plotted as solid black lines and gray extensions, respectively. For the 12+log(O/H) = 7.75 models, the change in line ratio as a function of age is shown by the spread of the shaded yellow region, where the dashed yellow line marks the upper limit of 107.0 yr.

Standard image High-resolution image

Just as Stark et al. (2014), Berg et al. (2016), and Senchyna et al. (2017) observed strong C iii] emission resulting from subsolar C/O in metal-poor systems (Z ≤ 0.5 × Z), reproducing the observed emission lines in SL2S0217 also requires subsolar C, N, and Si abundance ratios relative to O. The low gas metallicity prevents efficient metal cooling by oxygen, resulting in a high electron temperature that secures a stable production of the collisionally excited O, Si, and C nebular emission lines. If a portion of the metals are also locked up in dust, C and Si may be underabundant as a result. However, no evidence of dust reddening exists, and the extreme ionizing radiation field likely destroyed or eroded the dust in the H ii regions (but also see discussion of dust depletion in Section 7.2.3).

8.2. Stellar Ionization

The rate of ionizing photons needed to power the arc can be estimated from the demagnified Hβ flux. At the redshift measured in this work, we calculate a luminosity distance of Dl = 1.444 × 104 Mpc (Wright 2006). The hydrogen ionizing photon luminosity is then given by

Equation (6)

where μ is the lensing magnification and αB and ${\alpha }_{{\rm{H}}\beta }^{\mathrm{eff}}$ are the total and effective Hβ Case B H recombination coefficients, respectively. We calculate the recombination coefficients for Te = 1.6 × 104 K using the equations of Pequignot et al. (1991). Given the lensing magnification determined here (μ = 17.3) and the Hβ flux from the grism spectrum, we estimate an ionizing photon luminosity of Qion. = 7.3 × 1053 photons s−1. The cluster IMF and metallicity are needed in order to then constrain the nature of the ionizing stellar population; however, this value is still indicative of a large number of massive stars powering the observed spectrum (>103 O stars), which can live to greater ages owing to rotation and binary evolution.

8.3. Multiple Ionization Sources

While many of the emission-line features observed in the rest-frame UV spectra of SL2S0217 can be reproduced by photoionization from stars alone, outliers such as He ii prompted an investigation into additional ionization sources. Indeed, the ratios of rest-frame far-UV emission lines provide a means to discern between nuclear activity and star formation as the ionizing sources (e.g., Feltre et al. 2016). However, Feltre et al. (2016) use the photoionization models of Gutkin et al. (2016), which are not appropriate for our young target, as they assume continuous star formation for 100 Myr and do not include the effects of binaries (see Section 8.1.1). Instead, we investigate the effects of incorporating shocks and AGNs into the photoionization models presented above.

8.3.1. Shocks

Even in star-forming galaxies, a substantial portion of the nebular emission may be generated by shocks (e.g., Krabbe et al. 2014; Rich et al. 2014). This fact motivated Jaskot & Ravindranath (2016) to incorporate contributions from shocks into their pure photoionization models. In their study, shock+precursor models were taken from Mappings III (Allen et al. 2008), assuming a metallicity of Z ∼ 0.003 (the Small Magellanic Cloud models), a velocity range of 125–1000 km s−1, and magnetic field strengths of 0.5–10 μG. The authors find that the addition of the shock models results in an increase of both C iii] and He ii emission relative to Hβ; therefore, they suggest that nebular He ii may serve as a useful diagnostic of shocks for UV spectra.

In light of these findings, we investigated emission from shock ionization as a source for the significant He ii emission seen in SL2S0217. Following the methodology of Jaskot & Ravindranath (2016), we scaled the shock emission and added it to the emission from our photoionization models such that the shocks contribute 10%, 30%, 50%, 70%, or 90% of the total predicted emission in Hα (see their Section 2.3 for a more detailed description).

The emission-line ratios for a varying shock contribution to our best photoionization model are depicted in Figure 14(a) relative to those measured for SL2S0217. These results demonstrate that the addition of shocks to the BPASS photoionization does not improve the general match of models to the SL2S0217 features. While the extreme He ii/C iii] ratio can be reproduced when the ionizing radiation is dominated by shocks, the other observed line ratios are generally overpredicted and cannot be matched with the same conditions. Further, it is physically unlikely that the gas in SL2S0217 is 90% shock ionized, and so we dismiss shocks as a main source of ionization in SL2S0217.

Figure 14.

Figure 14. Cumulative UV emission-line ratios from cloudy photoionization models plus a (a) shock or (b) AGN contribution, while each color, slightly offset for better visualization, indicates a different fractional contribution from shocks or AGNs to the total Hα emission. The input stellar photoionization model is the best model fit to the SL2S0217 spectrum (see Figure 12). Z ∼ 0.003 shock+precursor models from Mappings III (Allen et al. 2008) are used for the shock contribution, where the spread represents a range in velocity of 125–1000 km s−1 and magnetic field strengths of 0.5–10 μG. Z = 0.25 Z dust-free AGN models over a range of power-law exponents (symbol size) are taken from Groves et al. (2004). While the extreme He ii/C iii] ratio observed for SL2S0217 can be reproduced by models with a significant shock or AGN component, the other line ratios cannot be matched with the same conditions.

Standard image High-resolution image

8.3.2. Active Galactic Nuclei

The complex suite of spectral features observed in extreme galaxies such as SL2S0217 may be due to a complex combination of ionizing sources, such as AGN emission diluted by stellar radiation. For supermassive black holes (SMBHs) in massive galaxies, numerous studies have found that the SMBH mass is well correlated with the velocity dispersion of the host-galaxy bulge (e.g., Gebhardt et al. 2000; Gültekin et al. 2009). However, it is unclear whether this well-known scaling relation extends to the galaxies as small as SL2S0217 (see discussion in Kormendy & Ho 2013). Black hole masses in dwarf AGNs have been measured in several low-mass star-forming galaxies in the local universe (e.g., Reines et al. 2011, 2014), with the lowest reported mass being just ∼50,000 M (Baldassare et al. 2015). Recently, Baldassare et al. (2017) found evidence for dwarf AGNs in 11 nearby (z < 0.055) composite dwarf galaxies. These authors measured black hole masses ranging from 104.9 to 106.1 MBH/M for galaxies with stellar masses of 108.1–109.4 M/M, on the order of SL2S0217. While these recent studies of local dwarf galaxies indicate that a non-negligible fraction of low-mass galaxies might harbor AGNs, we caution that we do not know how these results extrapolate to galaxies at high redshifts.

Given the uncertainties surrounding AGNs in low-mass galaxies, it is worthwhile to explore the possibility of producing the extreme emission-line ratios observed in SL2S0217 with the addition of emission from AGNs. For this purpose, we used the dust-free, low-metallicity (Z = 0.25 Z), narrow-line AGN photoionization models from Groves et al. (2004), assuming simple power-law radiation fields (fν ∝ νβ) with a range of powers. As in Section 8.3.1, we scale the AGN emission such that the AGNs contribute 10%, 30%, 50%, 70%, or 90% of the total emission in Hα.

Emission-line ratios for the best BPASS + AGN models are plotted in Figure 14(b) relative to those measured for SL2S0217. The resulting shifts in line ratios due to the contribution of emission from AGNs look very similar to those seen for the BPASS + Shock models. In either case, the additional hard radiation increases the relative emission from the highest-ionization lines: He ii, O iii], and C iv. This allows the C iv/C iii] ratio to be reproduced with small AGN contributions. Typical narrow-line Type II AGN observations have stronger C iv than C iii] emission (e.g., Humphreys et al. 2008; Matsuoka et al. 2009; Hainline et al. 2011; Cassata et al. 2013), but our models suggest that the C iii] dominant emission in SL2S0217 (C iii]/C iv = 1.46) could be produced by a gas-enshrouded AGN.

On the contrary, Figure 14 shows that the Si iii]/C iii] ratio is significantly less affected by contributions from AGNs than from shocks. This is due, in part, to the fact that the full range of shock velocities is considered at each value of log U, such that the maximum contribution from shocks for any ion is always included (near v ∼ 175 km s−1 for Si iii]). On the other hand, the AGN contribution to Si iii] emission peaks at low-ionization parameter; however, it has little effect on the total Si iii]/Hα emission, which is dominated by photoionization.

An additional difference is seen in the He ii/C iii] ratio, where the value observed in SL2S0217 is only reached by AGN-dominated models (>90% contribution; purple points), yet such models significantly overestimate the C iv/C iii] and O iii]/C iii] ratios. Further, if the strong nebular emission lines in SL2S0217 were powered by a Type II AGN, we would also expect to detect several high-ionization emission lines that are missing from the rest-frame UV spectrum: N v λ1240, Si iv λ1403, N iv λ1486 (e.g., Vanden Berk et al. 2001; Hainline et al. 2011). Again, the suite of lines observed for SL2S0217 cannot be simultaneously matched by BPASS photoionization models with an additional source of hard radiation, in this case, from AGNs.

8.4. High-ionization Emission Lines

While we postulate that SL2S0217 is indeed a star-forming galaxy, its strong He ii and C iv emission is rare and interesting. In fact, only 1% of UV-selected galaxies at z ∼ 3 show strong nebular C iv emission (e.g., Reddy et al. 2008; Hainline et al. 2011; Mainali et al. 2017). In comparison, z ∼ 2−3 LBGs with large Lyα EWs have been reported to measure relatively weak C iv emission with strengths ≤1% of Lyα (e.g., Erb et al. 2010). Notably, significant narrow C iv and/or He ii emission is also detected in several nearby dwarf galaxies, suggesting that extremely hard ionizing radiation fields may be common in low-mass, low-metallicity galaxies (e.g., Kehrig et al. 2011; Berg et al. 2016). This idea is further supported by several recent studies that have found nebular C iv and/or He ii in distant metal-poor star-forming galaxies (Erb et al. 2010; Christensen et al. 2012; Stark et al. 2014; de Barros et al. 2016; Steidel et al. 2016; Vanzella et al. 2016). Of the few extreme UV emission-line galaxies that have been observed at z ≳ 2 (e.g., Erb et al. 2010; Christensen et al. 2012; Bayliss et al. 2014; Stark et al. 2014, 2015; Caminha et al. 2016; de Barros et al. 2016; Vanzella et al. 2016; Smit et al. 2017), none show a comparable combination of emission-line ratios.

8.4.1. Nebular C iv λλ1548, 1550

Typically galaxies show C iv in (i) absorption from the surrounding ISM or circumgalactic medium (CGM) gas or (ii) a P Cygni profile from the stellar winds of massive stars (e.g., Leitherer et al. 2001). To the contrary, AGNs can produce C iv emission, but line widths are typically broader than nebular emission lines (hundreds of kilometers per second). To date, narrow C iv has only been observed in a handful of strongly lensed high-redshift galaxies (e.g., Christensen et al. 2012; Stark et al. 2014; Smit et al. 2017); therefore, the significant C iv emission detected in SL2S0217 is another interesting piece of this peculiar puzzle.

Since the C iv λλ1548, 1550 doublet is expected to be produced as a combination of nebular, stellar, and ISM/CGM components, we can use a simple combination of absorption and emission profiles to offer a possible model for the C iv profile observed in SL2S0217. Despite the absence of apparent C iv absorption, if we assume that the absorption in both C iv and Si iv is due to high-ionization clouds in the surrounding ISM of the galaxy, we can use the average absorption profile (which is nearly identical to the Si iv λ1403 profile; see Figure 7) to model the C iv absorption. Although the oscillator strength of Si iv λ1394 is roughly two times larger than that of Si iv λ1403, their relative absorption profile depths are closer to a factor of 1.5, likely due to saturation. Similarly, C iv λ1548 absorption is expected to be twice as strong as that of C iv λ1550. Following the observed model of Si iv, we assume relative absorption profile strengths of 3:2 for both Si iv λλ1394, 1403 and C iv λλ1548, 1550. The total profile is then produced by co-adding the absorption profile with the emission predicted by the best photoionization model, n × Fλ, where n is an arbitrary scaling of the emission to best match the observed spectrum of SL2S0217. The resulting model profiles and their components are plotted against the SL2S0217 spectrum in Figure 15. These model fits offer a solution in which the Si iv and C iv observed profiles could be produced, but we caution that this interpretation cannot be confirmed.

Figure 15.

Figure 15. Model to produce the observed Si iv absorption and C iv emission features by combining scaled versions of the average absorption profile and the emission-line ratios predicted by the photoionization described in Section 8.1.1.

Standard image High-resolution image

8.4.2. Nebular He ii λ1640

Arguably the most aberrant feature of the SL2S0217 UV spectrum is the impressive nebular He ii emission strength. Nebular He ii emission is produced by recombination from the He+2 ionization state and so requires very hard ionizing radiation with photons at energies ≥54 eV in order to ionize He++. The interpretation of He ii emission is further complicated by the fact that it can have both nebular and stellar origins. Stellar He ii is produced by the stellar winds of massive stars, such as Wolf-Rayet (W-R) stars, and so is broadened in relation to the wind velocity (e.g., Brinchmann et al. 2008; Erb et al. 2010; Cassata et al. 2013).

Leitherer et al. (1999) showed that the number of W-R stars relative to O stars peaks around 106.6 yr, and so stellar winds may no longer be prominently contributing to the He ii emission at a starburst age of 107 yr, in line with the fact that we do not observe a significant broad component to the He ii emission in SL2S0217. Further, Eldridge & Stanway (2009) demonstrated that the W-R He ii λ1640 feature is only prominent in binary models with stellar metallicities of Z = 0.004 or 0.008, which do not pair well with the gas-phase oxygen abundance of SL2S0217 (Zneb ∼ 0.001). Instead, the He ii emission in SL2S0217 appears to be nebular in origin. In the photoionization models of Jaskot & Ravindranath (2016), the narrow, nebular He ii emission seen in SL2S0217 only becomes strong for more metal-rich systems during the W-R phase or when ionized by shocks. Other sources of high-energy photons may be required to produce strong nebular He ii, such as AGNs, exotic Pop III stars, or high-mass X-ray binaries (e.g., Shirazi & Brinchmann 2012; Kehrig et al. 2015).

As discussed in Section 8.3.1, we dismiss a significant contribution to the ionizing budget from shocks, as the models significantly overpredict the O iii]/C iii] budget. In addition to overshooting O iii]/C iii], the AGN models (Section 8.3.2) also produce an excess of C iv relative to C iii] emission, but this argument is not definitive, as some of the emission may be obscured by overlapping C iv absorption. However, AGNs typically have large He ii EWs (≳8 Å; Hainline et al. 2011; Cassata et al. 2013), which are generally greater than those measured for z ∼ 2 galaxies (Scarlata et al. 2009; Erb et al. 2010) and for SL2S0217 (EW(He iin) = 2.4 Å or EW(He iitot) = 2.8 Å, where the difference is due to the difficulty in deblending the narrow and wide components).

Photoionization by AGNs has also been investigated using diagnostic plots of the C iv/He ii and C iii]/He ii line ratios, both of which are sensitive to metallicity and ionization parameter (Villar-Martin et al. 1997). In Figure 16 we plot our stellar photoionization models described in Section 8.1.1 in comparison to photoionization from pure shocks and AGNs for the C iv/He ii versus C iii]/He ii ratios. The observations for SL2S0217 (black star) are extended to larger values (shaded region) in case C iv is underestimated (due to overlapping C iv absorption) and/or He ii is overestimated (due to a stellar contribution), clearly suggesting a stellar origin of the photoionization budget. Note, however, that the stellar photoionization models still fail to reproduce the absolute strength/EW of the He ii emission. Motivated by the uncertainties of these line ratios, we will present a preferable analysis in a forthcoming paper that allows us to distinguish between sources of ionizing radiation using line ratios that are not complicated by stellar and CGM components.

Figure 16.

Figure 16. C iv/He ii vs. C iii]/He ii emission-line diagnostic diagram. The stellar photoionization models described in Section 8.1 are shown as filled colored circles, where the metallicity increases with color, ionization parameter increases with point size, and age increases with color saturation. As an example, for the CNSi = 0.1 and 12 + log(O/H) = 7.25 model, dashed lines connect points of constant age, and black dotted lines connect points of constant ionization parameter. Pure shock and AGN models are also shown. The observed line ratios of SL2S0217 are depicted as the black star, where the gray shaded region indicates the possible values of SL2S0217 if C iv or He ii is not purely nebular emission.

Standard image High-resolution image

In observations of star-forming galaxies, He ii emission is commonly attributed to W-R stars, which typically show strong and broad (Δv > 1000 km s–1; Crowther 2007) emission lines produced by their fast stellar winds. For example, Chandar et al. (2004) measured an extreme He ii λ1640 EW in the local universe from W-R stars in the massive cluster NGC 3125-1, with a broad width of ∼1000 km s−1, and further supported by strong detections of the N iv λ1488 and N v λ1720 WN W-R features. We therefore rule out W-R stars as the source of the He ii emission on the account that we do not observe significant broad He ii emission in SL2S0217, or any other indicative feature, and that the W-R phase occurs in a very short and early window that is unlikely the period of our observations. In binary star models, where the effects of mixing and mass transfer extend the life of these massive stars, He ii emission peaks ∼50 Myr after a burst of Z = 0.004 star formation. However, the He ii emission from W-R stars depends on the metallicity, and the lower-metallicity stellar spectra representative of SL2S0217 are not capable of reproducing the observed He ii (see Figure 13).

Theoretically, gravitational cooling radiation from gas accretion is another mechanism that can produce He ii (as well as Lyα) emission profiles like those in SL2S0217. In these models, infalling gas onto dark matter can be heated up to temperatures of T ∼ 105 K and then is cooled by H and He line emission (Kereš et al. 2005; Yang et al. 2006). It has been suggested that Lyα halos (Lyα "blobs" that show extended Lyα emission on scales of tens of kiloparsecs or more; e.g., Caminha et al. 2016; Patrício et al. 2016) are the observational signature of this cooling radiation in forming galaxies (Haiman et al. 2000; Fardal et al. 2001). While inflowing gas would also help explain the strong blue peak in the Lyα profile, it is unlikely that SL2S0217 is a Lyα halo. Our observations and lensing model show a small galaxy with Lyα emission coming from three dense star-forming regions, where the Lyα emission extends beyond the continuum on much smaller spatial scales (as detected in narrowband HST imaging; D. K. Erb et al. 2018, in preparation).

The current arsenal of stellar population models in the literature are only able to reproduce strong He ii emission lines with pure photoionization for extreme stellar populations. Due to its high ionization potential, He ii emission has been used as a probe to find Pop III stars (e.g., Visbal et al. 2015). A top-heavy IMF in extremely low metallicity regions can produce large He ii line luminosities (Tumlinson & Shull 2000; Schaerer 2003; Yoon et al. 2012). However, the He ii EW is very sensitive to metallicity and so requires Z < 10−5 for this model. Steidel et al. (2016) predicted that H ii regions with normal oxygen abundances can exhibit extreme UV features if they are powered by stars that are O/Fe enhanced relative to solar. Such a scenario would be expected if the ISM was enriched by core-collapse SNe, and then the best-matching stellar model would actually have a metallicity several times lower than the gas-phase oxygen abundance. Since stellar spectra are primarily dependent on the total opacity, which is largely modulated by the Fe abundance (e.g., Rix et al. 2004), much harder ionizing radiation would be produced than is expected from the nebular metallicity. Alternatively, Tornatore et al. (2007) predicted that zero-metallicity regions can survive in low-density regions around large overdensities, allowing Pop III stars to form, until as recently as z ∼ 2.

Other studies do not require large modifications to relative abundances or metal-free stars to produce narrow He ii emission. Kudritzki (2002) showed that certain O stars are indeed hot enough to ionize He ii, and Brinchmann et al. (2008) suggested that at low metallicities nebular He ii is predominantly produced by O stars where optically thin winds can be penetrated by ionizing photons. The energy production of low-metallicity massive stars may also be increased if they are fast rotators (Meynet & Maeder 2007). Further support is provided by the study of Kehrig et al. (2015), who found that the He ii ionization budget of I Zw 18 can only be produced by models of either low-metallicity supermassive O stars or rotating metal-free stars. However, the details of very low metallicity O stars are still widely debated and further complicated by the fact that some He ii nebulae do not appear to have surviving O or W-R stars (Kehrig et al. 2011). Gräfener & Vink (2015) posed an alternative scenario for the origin of narrow He ii emission in which metal-poor, very massive stars form strong but slow W-R-type stellar winds owing to their proximity to the Eddington limit. Post-AGB stars from an older stellar population may also produce He ii emission. The combined radiation field of post-AGB stars will dominate the ionizing spectrum about 100 Myr after a burst of star formation and would be strong enough to ionize He ii (Binette et al. 1994). However, the He ii luminosities produced by post-AGB stars are insufficient to explain the emission-line strength seen for SL2S0217 (Cassata et al. 2013).

It seems likely that very massive stars play a significant role in the He ii line strength. Unfortunately, the nature of metal-poor massive stars remains poorly understood, and therefore very massive stars are often withheld from current population synthesis models. Theoretical work on rotation and binary evolution of massive stars (e.g., Eldridge & Stanway 2009; de Mink et al. 2014) has produced a vast range in the predicted lifetimes and energy output of low-metallicity massive stars. Further, these models make simplifying assumptions about the rotation velocities and distribution of binary separations that are poorly motivated. In the future, more sophisticated models of metal-poor massive stars may be able to explain the strength of high-ionization nebular lines observed for SL2S0217. However, since narrow He ii emission is more common at z ∼ 3 (Cassata et al. 2013) than at z ∼ 0 (Kehrig et al. 2011), it is possible that the stellar populations that produce He ii ionizing radiation are more common at higher redshifts and are fundamentally different from typical populations in the local universe.

9. Conclusions

We present the rest-frame UV Keck/LRIS spectrum of SL2S J021737–051329 (SL2S0217), a lensed galaxy magnified by a factor of roughly 17 by a massive galaxy at z = 0.65. SL2S0217 is particularly interesting given its very low mass (M < 109 M), low estimated metallicity (Z ∼ 1/20 Z), and extreme star-forming conditions that produce strong nebular UV emission lines in the absence of any apparent outflows. Because these characteristics are more common at higher redshifts (e.g., Brinchmann et al. 2008; Stark et al. 2014) and have been shown to be correlated with escaping ionizing radiation (e.g., Verhamme et al. 2017), we present SL2S0217 as a template to study the extreme conditions that may be responsible for the reionization era.

In our analysis of the UV spectrum of SL2S0217 we found the Lyα, interstellar absorption, and nebular emission features to be notable in the following ways:

  • 1.  
    The Lyα feature is observed purely in emission that is double peaked and centered near the systemic velocity. The dominant blue peak and small peak separation indicate little to no outflowing neutral gas along the line of sight. Given the large Lyα EW, we measure an unexpectedly low Lyα escape fraction from the ${L}_{\mathrm{Ly}\alpha }^{\mathrm{obs}}/{L}_{\mathrm{Ly}\alpha }^{\mathrm{int}}$ ratio of 7%.
  • 2.  
    Nearly all of the observed absorption features are saturated owing to optically thick ionized gas. However, we were able to use the weakest Si ii and Fe ii lines to estimate a relative abundance of log(Si/Fe) = −0.225 ± 0.118, or 44%–77% solar, where this underabundance can be accounted for by Si depletion onto dust grains. Similar to the Lyα emission, both the low- and high-ionization absorption features show profiles with nearly systemic velocity centers and small velocity ranges, indicating very little or no outflowing ionized gas along the sight line to the lensed galaxy.
  • 3.  
    Large equivalent widths are measured for the C iv λλ1548, 1550, He ii λ1640, O iii] λλ1661, 1666, Si iii] λλ1883, 1892, and C iii] λλ1907, 1909 nebular emission lines. In the case of C iii] and C iv, these line strengths are similar to those observed for the extreme star-forming galaxies at higher redshifts (z > 7; e.g., Stark et al. 2015; Stark 2016).

With the exception of He ii, the relative emission-line strengths can be reproduced by ionization from a very high ionization, low-metallicity starburst with binaries. We rule out large contributions from AGNs and shocks to the photoionization budget, suggesting that the emission features requiring the hardest radiation field likely result from extreme stellar populations that are beyond the capabilities of current models. We therefore argue that the combination of features observed for SL2S0217 is the result of an extreme episode of star formation in a young galaxy along a sight line that does not cross any outflowing gas.

We have studied the extreme spectral features of SL2S0217, finding both its absorption and emission features to be distinct from the general population of z ∼ 2−3 galaxies. In particular, our models show that the emission-line strengths of SL2S0217 require very hard ionizing radiation, and so it adds to the small population of galaxies with UV emission lines powered by the hard ionizing stellar spectra that may be responsible for the reionization era. Therefore, due to its optimal properties and leveraged by its lensed nature, SL2S0217 is an ideal template to study the extreme conditions that are important for reionization and thought to be more common in the early universe.

This work evolved over the course of the project thanks to many creative and insightful conversations with colleagues, especially John Chisholm, Anne Jaskot, and Claudia Scarlata. We extend our gratitude to the Lorentz Center for useful discussions during the "Characterizing Galaxies with Spectroscopy with a view for JWST" workshop. Additionally, we are grateful to the referee for valuable comments that improved and clarified the paper. We wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

D.A.B. and D.K.E. are supported by the U.S. National Science Foundation through the Faculty Early Career Development (CAREER) Program, grant AST-1255591. Models in this work were computed using the computer cluster, Nemo, at the Leonard E. Parker Center for Gravitation, Cosmology and Astrophysics supported by NASA, the National Science Foundation, UW-Milwaukee College of Letters and Science, and UW-Milwaukee Graduate School. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of these organizations. This work was supported by a NASA Keck PI Data Award, administered by the NASA Exoplanet Science Institute. Data presented herein were obtained at the W. M. Keck Observatory from telescope time allocated to the National Aeronautics and Space Administration through the agency's scientific partnership with the California Institute of Technology and the University of California. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation.

Footnotes

  • Typical, moderately massive z ∼ 2–3 galaxies actually have net Lyα absorption.

  • Cooray et al. (2011) used two exponential profiles for the source and a singular isothermal elliptical (SIE) lens, perturbed by a distant singular isothermal sphere (SIS) halo. This model fixes the external shear field and the radial distribution. The subsequent lack of structure in the source surface brightness model results in significant residuals (see Figure 2(h) of Cooray et al. 2011).

  • IRAF is distributed by the National Optical Astronomical Observatories.

  • For the current spectrum, the emission-line shape is largely determined by the instrument resolution, which is wavelength dependent, and so the FWHM changes slightly across the spectrum (∼1.5–1.7 Å in the rest frame).

  • B12 presented further evidence for minimal dust. They determined the extinction of the arc continuum to be AV = 0.09 from their 2D model fit to the observed 3D-HST grism spectrum using a Calzetti et al. (2000) reddening law. Assuming RV = 4.05, this corresponds to a very small reddening value of E(BV) = 0.022 and so was concluded to be negligible.

  • 10 

    We note that a preliminary Si iii] density determination using TEMDEN in IRAF produced unphysical results for SL2S0217. We subsequently found that the emissivity ratios for the Si iii] density diagnostic approach an unexpected value of 3 using the default atomic data in IRAF. Alternatively, we find that the PyNeb package in Python uses the correct Si iii] emissivities.

  • 11 

    More recently, Jenkins & Wallerstein (2017) updated the elemental depletion equations for a given sight line in the Large Magellanic Cloud, including updated parameters for the Si depletion. However, this work does not analyze the depletion relationships for C, N, or O. Therefore, we have used the Jenkins (2009) equations for consistency.

  • 12 

    For dust depletion of $D(x)=\left[\tfrac{X}{{\rm{H}}}\right]$, the dust depletion factor is $\delta (x)\,=1-{10}^{D(x)}$.

Please wait… references are loading.
10.3847/1538-4357/aab7fa